Next Article in Journal
Effect of Low- and High-Si/Al Synthetic Zeolites on the Performance of Renovation Plasters
Previous Article in Journal
Mechanical and Environmental Properties of Cemented Paste Backfill Prepared with Bayer Red Mud as an Alkali-Activator Substitute
Previous Article in Special Issue
Microstructure and Properties of AlxCr1−xCoFeNi High-Entropy Alloys Prepared by Spark Plasma Sintering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Increased TaNbV on the Microstructure, Mechanical Properties, and Energy Release Characteristics of High-Entropy Alloy HfZrTi(TaNbV)x

Military Chemical Defense Institute, Beijing 102200, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(20), 4713; https://doi.org/10.3390/ma18204713
Submission received: 20 September 2025 / Revised: 8 October 2025 / Accepted: 9 October 2025 / Published: 14 October 2025
(This article belongs to the Special Issue Fabrication, Characterization, and Application of High Entropy Alloy)

Abstract

In this study, we propose a novel energetic structural material, HfZrTi(TaNbV)x (x = 0.1, 0.3, 0.5, 0.7, 0.9, Ta:Nb: V = 1:1:1), to improve the ductility and toughness of the HfZrTi high-entropy alloy (HEAs). The transformation of the single-phase Hexagonal Close-Packed (HCP) HfZrTi-based alloy into a Body-Centered Cubic (BCC) phase HfZrTiTaNbV alloy can be achieved by tuning the concentration of Group VB β-stabilizing elements. The proposed alloy combines the insensitivity and excellent mechanical strength of conventional inert alloys with the ability to react with air under high-velocity impact for energy release. The mechanical properties and energy release characteristics of HZTXx (H = Hf, Z = Zr, T = Ti, X = TaNbV) at various strain rates are systematically investigated, and comprehensive microstructural characterization is performed, establishing a clear structure–property relationship. Under high-rate loading, the rapid oxidation of reactive elements, such as Hf and Zr, with atmospheric oxygen releases substantial chemical energy, which can be further enhanced by an adiabatic temperature rise, inducing local thermal softening through adiabatic shear bands. This study elucidates the connection between the deformation response mechanism of HZTXx under dynamic loading and the microstructure, providing crucial insights for advancing the application of high-entropy alloys in energetic systems.

1. Introduction

High-Entropy Alloys (HEAs) represent a novel class of alloys composed of five or more elements in equimolar or near-equimolar ratios. The high mixing entropy and severe lattice distortion inherent to these alloys typically contribute to their favorable balance between strength and ductility [1,2,3,4,5,6,7,8]. In 2010, researchers from the U.S. Air Force Research Laboratory, Senkov et al. [9,10,11,12], pioneered developing novel high-strength equiatomic single-phase BCC high-entropy alloys WNbMoTa and WNbMoTaV, composed of Group VB and VIB refractory elements. Subsequently, Group IVB transition metals, other transition metals such as Cr and Al, and non-metallic elements (including C, B, O, N, and Si), were gradually incorporated into the BCC high-entropy alloy systems. This expansion significantly broadened the compositional and property design space for these alloys. The BCC high-entropy alloys developed to date primarily form single-phase BCC solid solutions or structures featuring precipitates such as B2 or Laves phases within a BCC matrix. These alloys typically exhibit exceptionally high strength at room and elevated temperatures, with high-temperature mechanical properties that rival or surpass conventional superalloys [1]. This has attracted extensive research interest globally, making it a significant focus in high-entropy alloy research [13,14]. Research on the properties of BCC high-entropy alloys encompasses their physical, chemical, and mechanical performance. These alloys typically exhibit outstanding high-temperature mechanical properties and resistance to high-temperature softening—often surpassing those of traditional Ni-based superalloys [13,15,16,17,18,19]. Moreover, under extreme conditions such as ultra-high temperatures and hypervelocity impact [20,21,22], high-entropy alloys (HEAs), particularly refractory high-entropy alloys (RHEAs), exhibit further enhanced properties owing to the diverse composition and unique microstructures, demonstrating great potential as advanced structural materials. Furthermore, the compositional flexibility of HEAs enables the incorporation of highly reactive elements (e.g., Ti, Zr), making them promising candidates for energetic structural materials [20]. In 2017, Zhang et al. [23] proposed the concept of high-entropy alloys (HEAs) as energetic structural materials (ESMs) and demonstrated their feasibility, which represents a novel class of materials capable of releasing chemical energy under shock loading [24,25]. Ren et al. [24] studied the compressive properties and energy release characteristics of TiZrNbV high-entropy alloy under shock loading. The quasi-static compressive yield strength of TiZrNbV is 786 MPa. Furthermore, Chen et al. [26] investigated the WFeNiMo high-entropy alloy; its quasi-static compressive yield strength is 965 MPa. During penetration of thin steel targets, fragments of the WFeNiMo high-entropy alloy exhibited more intense energy release than tungsten alloys, with significant chemical reactions triggered by adiabatic shear.
Research by Zhang et al. [23] demonstrated that HfZrTiTa0.53, the quasi-static yield strength, compressive strength, and fracture strain of HfZrTiTa0.53 alloys reach 786 MPa, 1314 MPa, and 13.5%, respectively. TiZrNbV [27] HEAs, and Ti-Zr-Ta medium-entropy alloys exhibit substantial energy release during high-velocity impact [2], showcasing remarkable impact-induced energy release characteristics. However, RHEAs generally exhibit limited strain hardening capacity and, consequently, poor uniform deformability [28,29,30], due to a single deformation mechanism such as planar slip of dislocations and insufficient dislocation interactions within their single BCC phase or BCC-based matrix [31]. The increasing strain rate exacerbates this situation. When deformed under adiabatic dynamic conditions, thermal effects cannot be neglected, and adiabatic shear banding tends to dominate, leading to rapid failure [32]. The inadequate ductility of RHEAs hinders their application as structural materials due to their limited load-bearing reliability and processability.
As mentioned above, most BCC high-entropy alloys exhibit poor ductility at room temperature. Investigations into their mechanical properties have predominantly relied on compression testing, with relatively few studies focusing on tensile properties. These studies have primarily concentrated on the HfZrTiTaNb alloy system. In this work, we fabricated a series of HZTXx (H = Hf, Z = Zr, T = Ti, X = TaNbV). High-entropy alloys by tuning the content of Ta, Nb, and V in an HfZrTi-based matrix alloy. This approach aims to enhance the room-temperature ductility and deformability of BCC-structured high-entropy alloys, thereby providing a strategy for their application in energetic structural materials such as casings.

2. Materials and Methods

2.1. Alloy Preparation

The HZTXx high-entropy alloy was synthesized via suspension melting (LiDe Equipment Technology Co., Ltd., Hangzhou, China). Initially, raw metal materials (purity = 99.95%) were weighed and ultrasonically cleaned (KQ-500DB Yijin New Material Technology Co., Ltd., Beijing, China). The inner wall of the chamber and the crucible were polished with sandpaper, cleaned with ethanol, and dried. The prepared alloy was loaded into the levitation melting furnace crucible with low-melting-point elements placed at the bottom and high-melting-point elements on top. The furnace door was closed, and the four sealing knobs were tightened. The chamber was first evacuated to a low vacuum of 5 Pa using a mechanical pump, followed by evacuation to a high vacuum of 5 × 10−3 Pa using a molecular pump. The chamber was purged three times with argon before being backfilled with argon to a negative pressure of −0.05 MPa. The current regulator was adjusted to 70 A, and the melting process was maintained for 2 min. The current control valve was then shut off, and after the alloy cooled, the above steps were repeated for remelting. This melting cycle was repeated five times. The sample was removed after the furnace cooled to room temperature.

2.2. Microstructure Characterization

The crystal structure of the HZTXx high-entropy alloy was characterized by X-ray diffraction (XRD; D8 ADVANCE) (Bruker, Billerica, MA, USA) with Cu-Ka radiation, and the measured 2θ ranges from 0 to 90°. Microstructure and elemental distribution were analyzed using scanning electron microscopy (SEM; Zeiss Gemini 300, Zhongke Baice Testing Technology Co., Ltd., Beijing China) equipped with backscattered electron (BSE) imaging, energy-dispersive spectroscopy (EDS) and transmission electron microscopy (TEM FEI Talos F200X, Zhongke Baice Testing Technology Co., Ltd., Beijing China) Electron backscatter diffraction (EBSD, Zhongke Baice Testing Technology Co., Ltd., Beijing China) analysis was performed to characterize the crystal orientation, grain/phase boundaries, phase distribution, and strain in the high-entropy alloy.

2.3. Mechanical Properties

Figure 1a shows a dog-bone-shaped specimen subjected to quasi-static tensile testing, with standardized cross-sectional dimensions of 10 mm × 4 mm × 1.8 mm. (b) illustrates a cylindrical specimen with dimensions of Φ 4 mm × 6 mm (aspect ratio of 1.5:1) used for quasi-static compression testing. (c) depicts a cylindrical specimen machined to dimensions of Φ 5 mm × 5 mm for dynamic mechanical testing.
Figure 2 shows a split Hopkinson pressure bar (SHPB) system (Longke Measurement and Control Technology Co., Ltd. Qinhuangdao, China), coupled with a high-speed camera was set up to investigate the dynamic compressive mechanical properties and fracture behavior of the material. The SHPB setup consisted of a striker bar (Φ20 mm × 400 mm), an incident bar (Φ20 mm × 1200 mm), and a transmission bar (Φ20 mm × 1200 mm), all made of 18NiCr350 maraging steel. Before conducting dynamic compression testing, strain gauges were mounted at the midpoints of the incident and transmission bars to record strain variations. Upon projectile impact with the incident bar, a high-speed camera was used to record the deformation process and fracture behavior of the specimen during dynamic loading.

3. Results and Discussion

3.1. Microstructure Characteristics

Figure 3a–e shows the quasi-static tensile fracture morphologies of the HZTX0, HZTX0.3, HZTX0.5, HZTX0.7, and HZTX0.9 alloy specimens. All fracture surfaces exhibit dimples that are either elongated or equiaxed, with the dimple size decreasing progressively as the content of X is increased. In contrast, the HZTX0.1 HEA demonstrates relatively poor ductility and is therefore represented solely by its quasi-static compressive fracture morphology. The fractographic features reveal cleavage planes in Figure 3f, thus indicating brittle fracture. These observations suggest that the content of X within the base alloy induces a transition from brittle to ductile fracture behavior.
The backscattered electron SEM image shown in Figure 4 reveals that the as-cast alloy exhibits equiaxed grains with a size of approximately 180 μm. The intragranular structure consists of a basket-weave morphology formed by interlaced elongated platelets with widths ranging from 0.1 to 1.5 μm. The basketweave structure is commonly observed in alpha titanium alloys. During the transformation from the high-temperature BCC phase to the low-temperature HCP phase, a single unit cell of BCC-Ti can give rise to 12 distinct HCP orientation variants. HCP lamellae nucleate and grow along these different orientations, resulting in the formation of an interlaced basketweave structure. As the concentration of Ta, Nb, and V increases, the HCP phase within the alloy completely disappears, leading to the formation of dark grain boundary segregation. The size of this segregation exhibits significant coarsening and a tendency to diffuse into the grain interior with increasing Ta, Nb, and V concentrations [33]. Meanwhile, Figure 5 shows EDS point scanning, performed on both the dendrites and interdendritic regions of the alloys with the addition of X to determine the distribution of each element. As the content of X is increased, the V element became progressively enriched in the interdendritic regions, while the Ta element was increasingly concentrated in the dendrite cores.
The phenomenon arises because the solidus temperature of the high-melting-point element Ta is first reached during the initial cooling of the molten alloy. As a result, the initially formed solid nuclei become preferentially incorporated into the developing dendritic cores, leading to significant Ta enrichment within these regions. As effective stabilizers of the BCC structure, increased concentrations of Ta, Nb, and V enhance the tendency toward single BCC phase formation. Specifically, higher Ta and Nb contents provide more high-melting-point solvent elements that solidify first to form the dendritic framework, while a greater amount of V, being a lower-melting-point solute, is rejected into the liquid phase. Consequently, the extent of compositional segregation intensifies with increasing total content of these elements, making the coexistence of V-enriched interdendritic regions and Ta-rich dendritic cores more pronounced.
Figure 6 and Figure 7 show the phase composition and structural evolution of as-cast HZTXx HEAs, which were characterized by XRD and EBSD analyses. The results reveal a targeted change in phase stability. Figure 6 and Figure 7 reveal a single hexagonal close-packed (HCP) phase structure in the HZTX0 high-entropy alloy, with its XRD pattern exhibiting exclusively HCP diffraction peaks. As the x content increases, the EBSD maps detect a predominant BCC phase along with a minor HCP phase. In contrast, the peak of the Laves phase in HZTX0.7 likely indicates the formation of HfV2. The limited volume fraction of the residual HCP phase, however, was not detected by XRD. It can be observed that with the increase in x content, the volume fraction of the BCC phase increases, indicating that the formation of the hcp phase is suppressed while the formation of the BCC phase is promoted. SAED further confirms that the crystal undergoes a phase transition from HCP to BCC with an increased content of X.
The EBSD analysis in Figure 8 shows that, as the X content increased from 0.1 to 0.7, the alloy grains progressively transitioned from a basket-weave morphology to irregular polygons with random orientations. However, HZTX0.9 exhibited a preferred ⟨100⟩ orientation. This is because the ⟨100⟩ direction in BCC-structured HEAs has the highest atomic linear density and lowest growth resistance, making it energetically favorable. Grains with ⟨100⟩ parallel to the heat flow thus grew fastest and outcompeted others [34]. The excessive addition of high-melting-point elements (Ta, Nb, V) elevated the liquidus temperature of the HEAs, thereby increasing the undercooling and the thermal energy required for dissipation during solidification. The resulting large undercooling and steep thermal gradients provided a stronger, sustained driving force for competitive ⟨100⟩ growth.
The formation of equiaxed grains requires a key mechanism: constitutional supercooling [35]. This means that the enrichment of solute at the solidification front depresses the local melting point, thereby forming a Constitutional Supercooling Zone (CSZ), which creates conditions for new heterogeneous nucleation. Elements like Hf, Zr, and Ti have a pronounced segregation tendency during solidification. In contrast, Ta, Nb, and V (particularly Ta and Nb) have relatively lower diffusion coefficients in the BCC structure and exhibit good compatibility with the principal elements. Figure 9 shows the pole figure, which indicates a predominance of grains with the <100> orientation in agreement with the above conclusion.
As evidenced by SAED (Selected Area Electron Diffraction) in Figure 10 and Figure 11, an increase in the content of X was found to promote a phase transition in the high-entropy alloy from an HCP to a BCC structure, which is consistent with the results shown in Figure 5 and Figure 6.
The ω phase was observed in alloys with low Ta, Nb, and V contents, such as HZTX0.1. This is as evidenced by the diffraction spots at positions 1 and 2 circled in Figure 11(a2) and further confirmed by the bright-field (BF) images in Figure 11(b1,b2). Additionally, the XRD pattern of HZTX0.1 in Figure 6 shows diffraction peaks corresponding to the ω phase. This is attributed to the fact that Ti, Zr, and Hf are typical ω-phase-forming elements, which tend to undergo a diffusionless, displacive transformation from the β-phase [36]. The ω-phase, which has a hexagonal structure, is hard and brittle, and generally detrimental to ductility. In contrast, elements such as Ta, Nb, and V are strong β-stabilizers. They effectively lower the free energy of the β-phase (BCC structure) and expand the temperature range of the β-phase field, thereby stabilizing the β-phase over a wider composition and temperature range and suppressing the precipitation of other phases, such as the α- and ω-phases. When the content of these strong β-stabilizers is reduced, the average electron concentration (e/a) of the alloy system decreases [37]. A lower e/a ratio is directly associated with instability of the β-phase, making metastable phases such as the ω-phase thermodynamically more favorable and prone to precipitation. Furthermore, the formation of the ω-phase is a diffusionless transformation whose nucleation strongly depends on local soft-mode vibrations and lattice instabilities in the parent BCC phase. With a reduction in Ta, Nb, and V content, the composition becomes dominated by Ti, Zr, and Hf. The atomic size mismatch among these elements exacerbates local strain along specific crystallographic directions in the BCC lattice. This strain provides an ideal condition for the coordinated atomic displacement (collapse) required for ω-phase formation, thereby facilitating its nucleation.
Also shown in Figure 12(a3,b3) and Figure 13(a3,b3), the IFFT of HZTXx reveals that the density of lattice distortion increases with increasing X content, which is attributed to the large atomic radius difference between Ta and V. The increased lattice distortion density significantly enhances the solid solution strengthening effect. The addition of Ta, Nb, and V alters the elastic modulus in local regions, requiring dislocations to expend additional energy to traverse these regions with varying elastic moduli.
A strong interaction occurs between the internal stress field generated by the distortion and the stress field of the dislocations themselves, resulting in the pinning of dislocations and the hindrance of their slip, ultimately leading to an improvement in the material’s strength. Severe lattice distortion effectively strengthens the alloy while considerably restricting the mobility of dislocations. The initiation and propagation of dislocations become difficult, necessitating high stress during the initial stages of deformation. Once deformation begins, dislocations rapidly entangle and accumulate, impeding coordinated plastic deformation.
The severe lattice distortion may cause a high work hardening rate initially, but dislocation multiplication and storage soon reach saturation, leading to a reduction in uniform plastic deformation capability. The elongation is typically reduced, which is consistent with the quasi-static tensile properties shown in Figure 14a.

3.2. Mechanical Properties of HZTXx

Figure 14 demonstrates the quasi-static tensile properties of X. The addition of TaNbV enhanced both the strength and ductility of the HEAs, except HZTX0.1, where the sample exhibited brittleness and poor tensile properties—consistent with the dynamic mechanical testing results. When the content of X exceeds 0.5, the strength of the high-entropy alloy increases while the elongation after fracture decreases. Notably, the elongation after fracture of HZTX0.9 shows an increase compared to that of HZTX0.7. The increase in plasticity may be attributed to grain growth resulting from the preferred orientation of grains. The yield strength, ultimate tensile strength, and elongation at fracture are summarized in Table 1.
The figure shows the engineering stress–strain curves of the HZTXx at a strain rate of 4000 s−1. Except for the HZTX0.1 specimen, which exhibited brittle fracture, all other samples demonstrated appreciable plastic deformation. The strength of the alloy increased with higher TaNbV content. Meanwhile, the HEAs exhibit positive strain rate sensitivity under high strain rate conditions.
Firstly, the yield strength of the HEAs increased with rising TaNbV content, albeit at the expense of reduced elongation in HZTX0.7 and HZTX0.9. This strength-ductility trade-off originated from the severe lattice distortion induced by the large atomic radius mismatch of V. The distortion created effective barriers to dislocation motion, significantly increasing the resistance to glide and thereby elevating the yield strength.
Secondly, solute atoms (such as Ta, Nb, and V) engage in elastic interactions with dislocations in the matrix. Larger Ta and Nb atoms tend to aggregate in the compressive stress regions beneath dislocation lines, while smaller V atoms may accumulate in tensile stress regions, thereby “pinning” the dislocations. Additionally, the elastic moduli of these elements differ. Dislocations tend to propagate preferentially through regions with lower modules. The introduction of Ta, Nb, and other elements introduces modulus mismatches, which further hinder dislocation motion. The increased content of Ta, Nb, and V directly enhances this solid solution strengthening effect, making plastic deformation more difficult, thereby improving strength but reducing ductility.
Furthermore, Ta and Nb are strong carbide-forming elements with high melting points and inherent stability. Their addition is likely to enhance the interatomic bonding strength. Stronger atomic bonds require higher energy to break for dislocation glide to occur, which also contributes to the increased strength and decreased plasticity.

3.3. SHPB Test of HZTXx Alloy

Based on the split Hopkinson pressure bar (SHPB) tests, the deformation and fracture process of the HEA-ESM specimen under dynamic loading in an air atmosphere was presented. The HEA-ESM specimen was sandwiched between the incident bar and the transmission bar. High-speed camera images from Hopkinson bar experiments at a strain rate of 4000 s−1 were selected for HZTX0, HZTX0.1, and HZTX0.9.
Figure 10 shows the adiabatic shear band formed in HZTX0 at t = 0.14 ms. Spark emission is observed at t=0.69 ms. This phenomenon is attributed to the following mechanism: upon application of a high-strain-rate load, homogeneous plastic deformation is initiated. The vast majority (typically over 90%) of the plastic work is converted into heat, resulting in adiabatic temperature rise. This effect is particularly pronounced in refractory high-entropy alloys such as HZT, which exhibit high strength and relatively low thermal conductivity. The local accumulation of heat causes a sharp temperature increase in specific regions. The rise in temperature leads to a decrease in flow stress, a phenomenon known as thermal softening. Consequently, the localized region becomes softer than the surrounding material. Spark formation is attributed to oxidation reactions between fragments and oxygen. During loading, heat is generated within the fragments due to adiabatic temperature rise caused by shear friction.
According to the principle of least resistance, further plastic deformation becomes concentrated preferentially within this softened zone. This concentration of deformation generates additional heat, leading to even higher temperatures and more significant thermal softening, which in turn further softens the region. As the material softens, deformation becomes increasingly localized, producing yet more heat. Instability is triggered when the rate of heat generation exceeds the capacity for heat dissipation, resulting in the formation of a highly localized shear band dominated by the thermal softening effect.
Due to the substantial precipitation of the ω-phase, the HZTX0.1 specimen exhibits low strength in static mechanical tests. In dynamic tests, this low strength leads to more complete fragmentation. To prevent severe oxidation and adhesion to the incident bar during Hopkinson bar tests resulting from extensive fragmentation, the specimen was clamped between spacers made of the same material as the incident and transmission bars, as shown in Figure 15b. Figure 15b displays intense sparking and fragment ejection observed in the HZTX0.1 specimen at t = 4.64 ms. Furthermore, active elements in the high-entropy alloy react with air under local high-temperature conditions, ultimately producing sparks.
The faster the time to maximum brightness upon specimen fragmentation and the greater the light intensity, the more efficient the energy release. To semi-quantify the energy release characteristics of the HEAs, the high-speed photographic images were binarized, with the threshold for optical radiation intensity set to 125–245. Pixels exhibiting grayscale values greater than 125 were identified as those corresponding to energy release reactions. Statistical analysis was performed on the luminous regions associated with energy release reactions in the high-speed images. As indicated in the Figure 16, HZTX0.1 exhibits the earliest ignition and the highest combustion luminosity. Its brittle characteristics facilitate more complete fragmentation, leading to a more violent oxidation reaction. HZTX0.9 ignites slightly earlier and shows higher luminosity than HZTX0. These results demonstrate that HZTX0.9, while satisfying the requirements of high strength and high plasticity, still maintains a specific energy release performance, making it a potential candidate for energetic liner materials.

4. Conclusions

In summary, this work developed an HfZrTiTaNbV(HEA) with excellent mechanical properties under both quasi-static and dynamic loading by tailoring the content of Ta, Nb, and V elements. The main findings are as follows:
(1)
As the content of Ta, Nb, and V was increased, the HZTXx high-entropy alloy underwent a transition in mechanical behavior from ductile to brittle and then back to ductile, accompanying by a phase evolution from a single hexagonal close-packed (HCP) structure to a body-centered cubic (BCC) phase with omega (ω)-phase precipitation, and ultimately to a single-phase BCC structure.
(2)
Under quasi-static tension, the profuse precipitation of the ω phase in HZTX0.1 resulted in brittle fracture and low strength, leading to poor tensile properties. As the content of Ta, Nb, and V increased, the grain size was progressively reduced, and the resultant grain refinement endowed the other alloys with improved tensile performance. The concomitant lattice distortion further contributed to an enhancement in yield strength. HZTX0.9 exhibited a yield strength of 1065 MPa, an ultimate tensile strength of 1119 MPa, and a fracture elongation of 16.4%
(3)
In the split-Hopkinson pressure bar (SHPB) tests, high-speed imaging revealed that the selected alloys exhibited energy-releasing oxidation reactions under dynamic high-strain-rate loading. The primary mechanism is attributed to the oxidation of active elements (e.g., Hf, Zr) with atmospheric oxygen during the fragmentation process, resulting in visible sparking. The brittle HZTX0.1 alloy fragments more thoroughly, producing finer fragments that facilitate more complete oxidation of the active elements. This results in the earliest observed ignition and the most intense sparking. Meanwhile, the HZTX0.9 alloy exhibits higher strength and superior ductility compared to the base alloy A, while also demonstrating a greater energy release effect. This combination of properties provides key material support for the future development of energetic structural materials.

Author Contributions

Conceptualization, Y.M. (Yusong Ma); Methodology, C.C.; Software, Y.M. (Yanqi Mei); Validation, C.C. and K.Z.; Formal analysis, Y.P.; Resources, Methodology, X.G.; Visualization, C.C. and X.L.; Supervision, M.W. and J.L. 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. 52506292).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflict 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. Tang, Y.; Wang, R.; Xiao, B.; Zhang, Z.; Li, S.; Qiao, J.; Bai, S.; Zhang, Y.; Liaw, P.K. A Review on the Dynamic-Mechanical Behaviors of High-Entropy Alloys. Prog. Mater. Sci. 2023, 135, 101090. [Google Scholar] [CrossRef]
  3. Tsai, M.-H.; Yeh, J.-W. High-Entropy Alloys: A Critical Review. Mater. Res. Lett. 2014, 2, 107–123. [Google Scholar] [CrossRef]
  4. 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. Available online: https://advanced.onlinelibrary.wiley.com/doi/abs/10.1002/adem.200300567 (accessed on 17 September 2025). [CrossRef]
  5. Wang, R.; Tang, Y.; Li, S.; Zhang, H.; Ye, Y.; Zhu, L.; Ai, Y.; Bai, S. Novel Metastable Engineering in Single-Phase High-Entropy Alloy. Mater. Des. 2019, 162, 256–262. [Google Scholar] [CrossRef]
  6. Peng, J.; Li, L.; Li, F.; Liu, B.; Zherebtsov, S.; Fang, Q.; Li, J.; Stepanov, N.; Liu, Y.; Liu, F.; et al. The Predicted Rate-Dependent Deformation Behaviour and Multistage Strain Hardening in a Model Heterostructured Body-Centered Cubic High Entropy Alloy. Int. J. Plast. 2021, 145, 103073. [Google Scholar] [CrossRef]
  7. Hilhorst, A.; Jacques, P.J.; Pardoen, T. Towards the Best Strength, Ductility, and Toughness Combination: High Entropy Alloys Are Excellent, Stainless Steels Are Exceptional. Acta Mater. 2023, 260, 119280. [Google Scholar] [CrossRef]
  8. Ding, Q.; Zhang, Y.; Chen, X.; Fu, X.; Chen, D.; Chen, S.; Gu, L.; Wei, F.; Bei, H.; Gao, Y.; et al. Tuning Element Distribution, Structure, and Properties by Composition in High-Entropy Alloys. Nature 2019, 574, 223–227. [Google Scholar] [CrossRef]
  9. Senkov, O.N.; Jensen, J.K.; Pilchak, A.L.; Miracle, D.B.; Fraser, H.L. Compositional Variation Effects on the Microstructure and Properties of a Refractory High-Entropy Superalloy AlMo0.5NbTa0.5TiZr. Mater. Des. 2018, 139, 498–511. [Google Scholar] [CrossRef]
  10. Guo, N.N.; Wang, L.; Luo, L.S.; Li, X.Z.; Chen, R.R.; Su, Y.Q.; Guo, J.J.; Fu, H.Z. Hot Deformation Characteristics and Dynamic Recrystallization of the MoNbHfZrTi Refractory High-Entropy Alloy. Mater. Sci. Eng. A 2016, 651, 698–707. [Google Scholar] [CrossRef]
  11. 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]
  12. 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]
  13. Miracle, D.B.; Tsai, M.-H.; Senkov, O.N.; Soni, V.; Banerjee, R. Refractory High Entropy Superalloys (RSAs). Scr. Mater. 2020, 187, 445–452. [Google Scholar] [CrossRef]
  14. Development of a Refractory High-Entropy Superalloy. Available online: https://www.mdpi.com/1099-4300/18/3/102 (accessed on 17 September 2025).
  15. Li, C.; Li, J.C.; Zhao, M.; Jiang, Q. Effect of Aluminum Contents on Microstructure and Properties of AlxCoCrFeNi Alloys. J. Alloys Compd. 2010, 504, S515–S518. [Google Scholar] [CrossRef]
  16. Zhou, Y.J.; Zhang, Y.; Wang, Y.L.; Chen, G.L. Solid Solution Alloys of AlCoCrFeNiTix with Excellent Room-Temperature Mechanical Properties. Appl. Phys. Lett. 2007, 90, 181904. Available online: https://pubs.aip.org/aip/apl/article-abstract/90/18/181904/332807/Solid-solution-alloys-of-AlCoCrFeNiTix-with?redirectedFrom=fulltext (accessed on 17 September 2025). [CrossRef]
  17. 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]
  18. Wang, M.; Ma, Z.L.; Xu, Z.Q.; Cheng, X.W. Designing VxNbMoTa Refractory High-Entropy Alloys with Improved Properties for High-Temperature Applications. Scr. Mater. 2021, 191, 131–136. [Google Scholar] [CrossRef]
  19. Lee, C.; Kim, G.; Musicó, B.L.; Gao, M.C.; An, K.; Song, G.; Chou, Y.-C.; Keppens, V.; Chen, W.; Liaw, P.K. Temperature Dependence of Elastic and Plastic Deformation Behavior of a Refractory High-Entropy Alloy. Sci. Adv. 2020, 6, eaaz4748. Available online: https://www.science.org/doi/full/10.1126/sciadv.aaz4748 (accessed on 17 September 2025). [CrossRef]
  20. Chakraborty, P.; Sarkar, A.; Ali, K.; Jha, J.; Jothilakshmi, N.; Arya, A.; Tewari, R. Design and Development of Low-Density, High Strength ZrNbAlVTi High Entropy Alloy for High Temperature Applications. Int. J. Refract. Met. Hard Mater. 2023, 113, 106222. [Google Scholar] [CrossRef]
  21. Li, T.; Liu, T.; Zhao, S.; Chen, Y.; Luan, J.; Jiao, Z.; Ritchie, R.O.; Dai, L. Ultra-Strong Tungsten Refractory High-Entropy Alloy via Stepwise Controllable Coherent Nanoprecipitations. Nat. Commun. 2023, 14, 3006. [Google Scholar] [CrossRef]
  22. Wu, S.; Qiao, D.; Zhang, H.; Miao, J.; Zhao, H.; Wang, J.; Lu, Y.; Wang, T.; Li, T. Microstructure and Mechanical Properties of CxHf0.25NbTaW0.5Refractory High-Entropy Alloys at Room and High Temperatures. J. Mater. Sci. Technol. 2022, 97, 229–238. [Google Scholar] [CrossRef]
  23. Zhang, Z.; Zhang, H.; Tang, Y.; Zhu, L.; Ye, Y.; Li, S.; Bai, S. Microstructure, Mechanical Properties and Energetic Characteristics of a Novel High-Entropy Alloy HfZrTiTa0.53. Mater. Des. 2017, 133, 435–443. [Google Scholar] [CrossRef]
  24. Ren, K.; Liu, H.; Chen, R.; Tang, Y.; Guo, B.; Li, S.; Wang, J.; Wang, R.; Lu, F. Compression Properties and Impact Energy Release Characteristics of TiZrNbV High-Entropy Alloy. Mater. Sci. Eng. A 2021, 827, 142074. [Google Scholar] [CrossRef]
  25. Wang, R.; Tang, Y.; Li, S.; Ai, Y.; Li, Y.; Xiao, B.; Zhu, L.; Liu, X.; Bai, S. Effect of Lattice Distortion on the Diffusion Behavior of High-Entropy Alloys. J. Alloys Compd. 2020, 825, 154099. [Google Scholar] [CrossRef]
  26. Dynamic Mechanical Behavior and Penetration Performance of Wfenimo High-Entropy Alloy. Available online: https://lxxb.cstam.org.cn/en/article/Y2020/I5/1443 (accessed on 17 September 2025).
  27. Meng, J.; Shen, B.; Wang, J.; Xue, R.; Chen, J.; Li, S.; Tang, Y. Energy-Release Behavior of TiZrNbV High-Entropy Alloy. Intermetallics 2023, 162, 108036. [Google Scholar] [CrossRef]
  28. Wu, Y.D.; Cai, Y.H.; Wang, T.; Si, J.J.; Zhu, J.; Wang, Y.D.; Hui, X.D. A Refractory Hf25Nb25Ti25Zr25 High-Entropy Alloy with Excellent Structural Stability and Tensile Properties. Mater. Lett. 2014, 130, 277–280. [Google Scholar] [CrossRef]
  29. Huang, H.; Sun, Y.; Cao, P.; Wu, Y.; Liu, X.; Jiang, S.; Wang, H.; Lu, Z. On Cooling Rates Dependence of Microstructure and Mechanical Properties of Refractory High-Entropy Alloys HfTaTiZr and HfNbTiZr. Scr. Mater. 2022, 211, 114506. [Google Scholar] [CrossRef]
  30. Yuan, Y.; Wu, Y.; Yang, Z.; Liang, X.; Lei, Z.; Huang, H.; Wang, H.; Liu, X.; An, K.; Wu, W.; et al. Formation, Structure, and Properties of Biocompatible TiZrHfNbTa High-Entropy Alloys. Mater. Res. Lett. 2019, 7, 225–231. [Google Scholar] [CrossRef]
  31. Xiong, W.; Guo, A.X.Y.; Zhan, S.; Liu, C.-T.; Cao, S.C. Refractory High-Entropy Alloys: A Focused Review of Preparation Methods and Properties. J. Mater. Sci. Technol. 2023, 142, 196–215. [Google Scholar] [CrossRef]
  32. Yan, N.; Li, Z.; Xu, Y.; Meyers, M.A. Shear Localization in Metallic Materials at High Strain Rates. Prog. Mater. Sci. 2021, 119, 100755. [Google Scholar] [CrossRef]
  33. Leyens, C.; Peters, M. Titanium and Titanium Alloy; Wiley Online Books: Weinheim, Germany, 2003; pp. 1–36. Available online: https://onlinelibrary.wiley.com/doi/book/10.1002/3527602119 (accessed on 17 September 2025).
  34. Li, Y.; Du, Z.; Zhou, F.; Zhang, Y.; Lv, Y.; Li, R.; Pan, K.; Fan, J.; Han, Y. Effect of Rolling Reduction on the Texture Evolution and Mechanical Properties Hot-Rolled WMoTaV Refractory High Entropy Alloy with Interfacial Segregation. Mater. Sci. Eng. A 2025, 927, 148010. [Google Scholar] [CrossRef]
  35. Tu, C.-H.; Lai, Y.-C.; Wu, S.-K.; Lin, Y.-H. The Effects of Annealing on Severely Cold-Rolled Equiatomic HfNbTiZr High Entropy Alloy. Mater. Lett. 2021, 303, 130526. [Google Scholar] [CrossRef]
  36. Ji, X.; Chong, Y.; Emura, S.; Tsuchiya, K. Heterogeneous Distribution of Isothermal ω Precipitates Prevents Brittle Fracture in Aged β-Ti Alloys. Scr. Mater. 2024, 241, 115879. [Google Scholar] [CrossRef]
  37. Ma, X.; Li, N.; Feng, T.; Xiao, L. Overcoming the Strength-Ductility Trade-Off in ω -Phase Strengthened Metastable β Titanium Alloys via Dual Heterogeneous Structure. Mater. Res. Lett. 2025, 13, 844–853. [Google Scholar] [CrossRef]
Figure 1. (a) Specimen subjected to quasi-static tensile testing, (b) Specimen subjected to quasi-static compressive testing, (c) Specimen tested via the split Hopkinson bar technique.
Figure 1. (a) Specimen subjected to quasi-static tensile testing, (b) Specimen subjected to quasi-static compressive testing, (c) Specimen tested via the split Hopkinson bar technique.
Materials 18 04713 g001
Figure 2. Split Hopkinson pressure bar apparatus.
Figure 2. Split Hopkinson pressure bar apparatus.
Materials 18 04713 g002
Figure 3. (ae) SEM of the fracture surfaces of HZTX0, HZTX0.3, HZTX0.5, HZTX0.7, and HZTX0.9 specimens subjected to quasi-static tensile testing. The dashed boxes in (d,e) represent close-up views of the dimples. (f) SEM of the fracture surfaces of HZTX0.1 specimens subjected to quasi-static compression testing.
Figure 3. (ae) SEM of the fracture surfaces of HZTX0, HZTX0.3, HZTX0.5, HZTX0.7, and HZTX0.9 specimens subjected to quasi-static tensile testing. The dashed boxes in (d,e) represent close-up views of the dimples. (f) SEM of the fracture surfaces of HZTX0.1 specimens subjected to quasi-static compression testing.
Materials 18 04713 g003
Figure 4. (a) BSE micrograph of the HZTX0, (b) BSE micrograph of the HZTX0.1, (c) BSE micrograph of the HZTX0.3, (d) BSE micrograph of the HZTX0.5, (e) BSE micrograph of the HZTX0.7, (f) BSE micrograph of the HZTX0.9. A1–A5: EDS point analysis locations on dendritic core; B1–B5: on interdendritic region.
Figure 4. (a) BSE micrograph of the HZTX0, (b) BSE micrograph of the HZTX0.1, (c) BSE micrograph of the HZTX0.3, (d) BSE micrograph of the HZTX0.5, (e) BSE micrograph of the HZTX0.7, (f) BSE micrograph of the HZTX0.9. A1–A5: EDS point analysis locations on dendritic core; B1–B5: on interdendritic region.
Materials 18 04713 g004
Figure 5. (a,b) Variations in the composition of A and B were determined by EDS.
Figure 5. (a,b) Variations in the composition of A and B were determined by EDS.
Materials 18 04713 g005
Figure 6. XRD patterns of HZTXx HEAs.
Figure 6. XRD patterns of HZTXx HEAs.
Materials 18 04713 g006
Figure 7. (a) EBSD phase maps of the HEAs of HZTX0, (b) EBSD phase maps of the HEAs of HZTX0.1, (c) EBSD phase maps of the HEAs of HZTX0.3, (d) EBSD phase maps of the HEAs of HZTX0.5, (e) EBSD phase maps of the HEAs of HZTX0.7, (f) EBSD phase maps of the HEAs of HZTX0.9. In the dashed boxes: close-up views of the secondary phase from EBSD analysis.
Figure 7. (a) EBSD phase maps of the HEAs of HZTX0, (b) EBSD phase maps of the HEAs of HZTX0.1, (c) EBSD phase maps of the HEAs of HZTX0.3, (d) EBSD phase maps of the HEAs of HZTX0.5, (e) EBSD phase maps of the HEAs of HZTX0.7, (f) EBSD phase maps of the HEAs of HZTX0.9. In the dashed boxes: close-up views of the secondary phase from EBSD analysis.
Materials 18 04713 g007
Figure 8. (af) presents the EBSD orientation maps for the HZTXx high-entropy alloy system with the X content increasing from 0 to 0.9.
Figure 8. (af) presents the EBSD orientation maps for the HZTXx high-entropy alloy system with the X content increasing from 0 to 0.9.
Materials 18 04713 g008
Figure 9. The pole figure of HZTX0.9 with <100>, <110>, <111> orientation.
Figure 9. The pole figure of HZTX0.9 with <100>, <110>, <111> orientation.
Materials 18 04713 g009
Figure 10. (a) HRTEM, (b) SAED, (c) IFFT of HZTX0.
Figure 10. (a) HRTEM, (b) SAED, (c) IFFT of HZTX0.
Materials 18 04713 g010
Figure 11. (a1) HRTEM, (a2,a3) SAED, (b1,b2) BF, (b3) IFFT of HZTX0.1.
Figure 11. (a1) HRTEM, (a2,a3) SAED, (b1,b2) BF, (b3) IFFT of HZTX0.1.
Materials 18 04713 g011
Figure 12. (a1) HRTEM, (a2) SAED, (a3) IFFT of HZTX0.3. (b1) HRTEM, (b2) SAED, (b3) IFFT of HZTX0.5.
Figure 12. (a1) HRTEM, (a2) SAED, (a3) IFFT of HZTX0.3. (b1) HRTEM, (b2) SAED, (b3) IFFT of HZTX0.5.
Materials 18 04713 g012aMaterials 18 04713 g012b
Figure 13. (a1) HRTEM, (a2) SAED, (a3) IFFT of HZTX0.7. (b1) HRTEM, (b2) SAED, (b3) IFFT of HZTX0.9.
Figure 13. (a1) HRTEM, (a2) SAED, (a3) IFFT of HZTX0.7. (b1) HRTEM, (b2) SAED, (b3) IFFT of HZTX0.9.
Materials 18 04713 g013
Figure 14. (a) Quasi-static tension stress–strain curve at a strain rate of 1 × 10−3s−1; (b) stress–strain curve at a strain rate of 4000 s−1.
Figure 14. (a) Quasi-static tension stress–strain curve at a strain rate of 1 × 10−3s−1; (b) stress–strain curve at a strain rate of 4000 s−1.
Materials 18 04713 g014
Figure 15. (a) The deformation and fracture process under Air atmospheres of HZTX0, (b) The deformation and fracture process under Air atmospheres of HZTX0.1, (c) The deformation and fracture process under Air atmospheres of HZTX0.9.
Figure 15. (a) The deformation and fracture process under Air atmospheres of HZTX0, (b) The deformation and fracture process under Air atmospheres of HZTX0.1, (c) The deformation and fracture process under Air atmospheres of HZTX0.9.
Materials 18 04713 g015
Figure 16. Time-Dependent Light Intensity Profiles of HZTX0, HZTX0.1, HZTX0.9.
Figure 16. Time-Dependent Light Intensity Profiles of HZTX0, HZTX0.1, HZTX0.9.
Materials 18 04713 g016
Table 1. The mechanical property parameters of the HZTXx HEAs.
Table 1. The mechanical property parameters of the HZTXx HEAs.
CategoryYield Strength (MPa)Ultimate Tensile Strength (MPa)Elongation at Fracture (%)
HZT64281113.5%
HZTX0.383086922.0%
HZTX0.589493427.3%
HZTX0.7963101112.3%
HZTX0.91065111916.4%
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

Chen, C.; Ma, Y.; Wei, M.; Gai, X.; Peng, Y.; Mei, Y.; Liu, X.; Zhang, K.; Li, J. Influence of Increased TaNbV on the Microstructure, Mechanical Properties, and Energy Release Characteristics of High-Entropy Alloy HfZrTi(TaNbV)x. Materials 2025, 18, 4713. https://doi.org/10.3390/ma18204713

AMA Style

Chen C, Ma Y, Wei M, Gai X, Peng Y, Mei Y, Liu X, Zhang K, Li J. Influence of Increased TaNbV on the Microstructure, Mechanical Properties, and Energy Release Characteristics of High-Entropy Alloy HfZrTi(TaNbV)x. Materials. 2025; 18(20):4713. https://doi.org/10.3390/ma18204713

Chicago/Turabian Style

Chen, Chong, Yusong Ma, Manhui Wei, Xiqiang Gai, Yue Peng, Yanqi Mei, Xinglong Liu, Kaichuang Zhang, and Jianbin Li. 2025. "Influence of Increased TaNbV on the Microstructure, Mechanical Properties, and Energy Release Characteristics of High-Entropy Alloy HfZrTi(TaNbV)x" Materials 18, no. 20: 4713. https://doi.org/10.3390/ma18204713

APA Style

Chen, C., Ma, Y., Wei, M., Gai, X., Peng, Y., Mei, Y., Liu, X., Zhang, K., & Li, J. (2025). Influence of Increased TaNbV on the Microstructure, Mechanical Properties, and Energy Release Characteristics of High-Entropy Alloy HfZrTi(TaNbV)x. Materials, 18(20), 4713. https://doi.org/10.3390/ma18204713

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