Next Article in Journal
The Impact of 45S5-Bioactive Glass on Synovial Cells in Knee Osteoarthritis—An In Vitro Study
Next Article in Special Issue
Improving Mechanical Properties of Co-Cr-Fe-Ni High Entropy Alloy via C and Mo Microalloying
Previous Article in Journal
Effect of Alloying Additives and Moulding Technology on Microstructure, Tightness, and Mechanical Properties of CuSn10 Bronze
Previous Article in Special Issue
The Effects of the Al and Zr Contents on the Microstructure Evolution of Light-Weight AlxNbTiVZry High Entropy Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Microstructures, Mechanical Properties, and Deformation Mechanism of B2-Hardened NbTiAlZr-Based Refractory High-Entropy Alloys

1
Shi-Changxu Innovation Center for Advanced Materials, Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, China
2
School of Metallurgy, Northeastern University, Shenyang 110819, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(24), 7592; https://doi.org/10.3390/ma16247592
Submission received: 27 October 2023 / Revised: 26 November 2023 / Accepted: 27 November 2023 / Published: 11 December 2023

Abstract

:
The NbTiAlZrHfTaMoW refractory high-entropy alloy (RHEA) system with the structure of the B2 matrix (antiphase domains) and antiphase domain boundaries was firstly developed. We conducted the mechanical properties of the RHEAs at 298 K, 1023 K, 1123 K, and 1223 K, as well as typical deformation characteristics. The RHEAs with low density (7.41~7.51 g/cm3) have excellent compressive-specific yield strength ( σ YS / ρ ) at 1023 K (~131 MPa·cm3/g) and 1123 K (~104.2 MPa·cm3/g), respectively, which are far superior to most typical RHEAs. And, they still keep appropriate plastic deformability at room temperature (ε > 0.35). The superior specific yield strengths are mainly attributed to the solid solution strengthening induced by the Zr element. The formation of the dislocation slip bands with [111](10 1 _ ) and [111](11 2 _ ) directions and their interaction provide considerable plastic deformation capability. Meanwhile, dynamic recrystallization and dislocation annihilation accelerate the continuous softening after yielding at 1123 K.

1. Introduction

With the development of high-temperature metallic materials, Ni-based superalloys have been unable to meet the requirements of aerospace engines. It is urgent to find a new high-performance alloy to replace the Ni-based superalloy [1]. Refractory high-entropy alloys (RHEAs) based on refractory elements with superior high-temperature mechanical properties have recently attracted intensive research attention, which are regarded as promising candidates for high-temperature structural materials [2,3,4]. However, their inferior plasticity along with high ductile–brittle transition temperature and high density seriously limits their aerospace applications, where the requirements are mainly focus on the deformability and high specific strength of alloys.
Research efforts have been attempting to overcoming these shortcomings via component and structure regulation [5,6]. In order to improve plasticity at room temperature (RT), a new refractory high-entropy alloy (RHEA), HfNbTaTiZr [7], with excellent RT compression plasticity (ε > 50%), has been developed. However, its specific yield strength (SYS) at high temperatures is undesirability (such as <60 MPa·cm3/g at 1073 K). For better SYS, Han et al. [5] and Wang et al. [8] added Ti or V to NbMoTaW and NbMoTa RHEAs, respectively. Unfortunately, their SYSs are still lower than 90 MPa·cm3/g at 1073 K. Hence, the harmonization of strength, plasticity, and density is an intractable problem in RHEAs. Meanwhile, the strain softening phenomenon in RHEAs at high temperatures is ubiquitous, but the quantitative analysis of the softening mechanism is still lacking [9,10,11,12].
For the two crucial issues mentioned above, in this work, we realized the structural transition from disorder to order of a plastic NbTi-based RHEA via Al-Zr regulation. Meanwhile, high-temperature solid solution strengthening and grain boundary (GB) strengthening were also introduced into the RHEA by adding appropriate amounts of Hf, Ta, Mo, and W. On this basis, the fresh Nb42Ti25Al15Zr5Hf5Ta5Mo2W1 (NbTiAl-1), Nb39Ti23Al15Zr10Hf5Ta5Mo2W1 (NbTiAl-2), and Nb37Ti20Al15Zr15Hf5Ta5Mo2W1 (NbTiAl-3) RHEAs with the B2 matrix and antiphase domain boundaries (APBs) are firstly developed. The RHEAs exhibit excellent SYS at high temperatures, maintaining the top level in the field of plastic RHEAs. In addition, the deformation mechanism and strain softening characteristic at RT and 1123 K are also discussed in detail.

2. Materials and Methods

The RHEAs (the purity of elements was higher than 99.9 wt%) were prepared by vacuum melting in a furnace with a Ti-gettered argon atmosphere. The Zr-Ti-Al and Nb-Hf-Ta-Mo-W intermediate alloys were prepared, respectively in order to avoid the volatility of Al and Ti. Then, they were blended for the final button ingots, which were remelted at least six times to ensure chemical homogeneity. Compression samples with sizes of Φ5 mm × 7.5 mm and Φ8 mm × 12 mm were obtained directly from the button ingots. The 298 K compression tests (Φ5 mm × 7.5 mm) at RT were conducted on an Instron 5582 machine with a strain rate 0.001/s. The 1123 K compression tests (Φ8 mm × 12 mm) were conducted on a GLEEBLE 3800 machine with the same rate. Microstructure and composition analyses were performed with a scanning electron microscope (SEM; XL30-FEG, FEI, Eindhoven, Holland). In this work, both SEM and electron backscatter diffraction (EBSD) samples were, firstly, mechanically ground with 1200 to 2000 grit SiC sandpapers, followed by mechanical polishing. The mechanically polished surfaces were ultrasonically cleaned with alcohol and then electrolytically polished with an electrolyte of 6% HClO4 + 35% CH3(CH2)3OH + 59% CH3OH at about −30 °C. Phase identification was conducted using an X-ray diffractometer (XRD; Rigaku D/max-2500PC, Tokyo, Japan) and transmission electron microscope (TEM; Talos F200X, ThermoFisher, Waltham, MA, USA). The as-cast TEM samples, slices cut from the bulk samples, were mechanically ground to a thickness of ~70 μm. Disks with a diameter of 3 mm punched out from the thin foils were further ground to ~40 μm and finally thinned using a Gatan Model 695 after dimpling [13]. For TEM samples after deformation, thin specimens cut from the core of the column sample were mechanically ground to a thickness of approximately ~70 μm. The subsequent thinning process was consistent with the as-cast TEM samples.

3. Results and Discussion

The XRD results of the as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3 RHEAs are shown in Figure 1. All three alloys have a BCC structure. The XRD pattern of the NbTiAl-3 shows the presence of weakly ordered B2 phase structural diffraction peaks at about 26°. As shown in the enlarged illustration in Figure 1, the (110) diffraction peaks are obviously shifted to the left with the increase in Zr elements, indicating a significant increase in the lattice constants, which are 0.3302 nm for the NbTiAl-1, 0.3310 nm for the NbTiAl-2, and 0.3317 nm for the NbTiAl-3, respectively. Combined with the XRD results, Zr is the element with the largest atomic size in this alloy system, which results in the largest solid solution effect and thus a significant increase in the lattice constant. Meanwhile, the density values of NbTiAl-1, NbTiAl-2, and NbTiAl-3 are 7.51 g/cm3, 7.44 g/cm3, and 7.41 g/cm3, respectively.
As shown in Figure 2a, the as-cast structure of NbTiAl-1 is equiaxial dendritic segregation. NbTiAl-2 and NbTiAl-3 are dendritic structures, as shown in Figure 2b,c, which are mainly due to the rapid solidification process of elemental segregation. The EDS results shown in Figure 2d reveal that, due to the differences in the melting points of the constituent elements, the high-melting-point elements Nb, Ta, Mo, and W firstly solidify in the dendrite region (DR), and then drive the lower-melting-point Zr and Al elements into the inter-dendrite region (ID), which promotes the formation of the dendritic structure, whereas the distribution of the Ti and Hf elements is relatively uniform. This dendritic structure induced by the difference in elemental melting point is highly consistent with the as-cast microstructure of previous RHEAs [14,15].
Figure 3 shows the dark-field TEM (DF-TEM) images with the superlattice spots of (100) of as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3, which indicates that the as-cast structure of the RHEAs is composed of an ordered B2 matrix and APBs [16,17,18]. The presence of an ordered B2 structure is again verified by the superlattice spots in the electron diffraction pattern (EDP) with the [001] zone axis. This phenomenon is consistent with the result that the structure factor ratio F 100 F 200 is small due to a counterbalancing of heavy and lighter elements on the different lattice sites according to the site occupation factors S O F i L S in MoCrTiAl [18]. From the DF-TEM results, the volume fraction of B2 phase keeps increasing with the increase in Zr element, i.e., the density of APBs decreases significantly because Al-Zr partitioning can promote the formation of the B2 structure during the rapid solidification process. This partitioning trend is consistent with previous studies [19,20]. It should be noted that the effect of the density of APBs on the yield strength of the RHEAs is limited because the APBs had no notable influence on the motion and spatial distribution of dislocations, as demonstrated in our previous work and that of others [21,22,23].
Figure 4 shows the compression true stress–strain curves of the NbTiAl-1, NbTiAl-2, and NbTiAl-3 at different temperatures. The yield strengths of the three RHEAs at 298 K are 1198.1 ± 7.5 MPa, 1333.1 ± 7.2 MPa, and 1418.1 ± 7.4 MPa, respectively. The yield strength of the RHEAs increases significantly with increasing Zr content. It indicates that the alloy has an obvious solid solution strengthening effect after the addition of Zr [17,24]. Similarly, the yield strengths of the RHEAs at 1023 K and 1123 K shown in Figure 4b,c maintain that trend, proving that the solid solution strengthening induced by the Zr element are still effective in that temperature range. The 1223 K compression results in Figure 4d show that the yield strengths of NbTiAl-2 and NbTiAl-3 are similar, indicating that the solid solution strengthening induced by the addition of Zr elements are weakening. The obvious serration behavior of NbTiAl-2 from dynamic strain aging (DSA), shown in Figure 4b, is caused by the tearing-off and re-pining of alternating dislocations [25,26].
In order to assess the effect of solid solution strengthening in the RHEAs, a power-law function was used to plot the relationship between the yield strength increment Δσ of NbTiAl-1, 2, and 3 RHEAs and the concentration of Zr in the B2 matrix (CZr, Table 1): Δσ~ C Z r n . A good linear fit between Δσ and CZr was obtained for NbTiAl-1, 2, and 3 RHEAs at 298 K, 1023 K, and 1123 K when n = 0.25, as shown in Figure 5. Therefore, the increase in strength of NbTiAl-1, 2, and 3 RHEAs with increasing Zr content at 298 K, 1023 K, and 1123 K is doubtlessly caused by the solid solution strengthening effect. There is no longer a linear relationship between Δσ~ C Z r 0.25 at 1223 K, which proves that the solid solution strengthening is significantly reduced at this temperature, which is also consistent with the trend that the solid solution strengthening effect decreases with increasing temperature [27]. More interestingly, the strain softening of the RHEAs becomes more and more pronounced with the increase in Zr content in the range of 1023 K–1223 K. The specific strain softening mechanism will be analyzed below. The compressive yield strength (σYS) and SYS (σYS/ρ) of the three RHEAs at different temperatures are shown in Table 2.
The SYS values and strain of the RHEAs at RT are compared with other representative RHEAs, such as NbMoTa [28], AlNbTiV [24], ZrTiHfV0.5Nb0.5, Zr2.0TiHfVNb2.0 and ZrTiHfNb0.5Ta0.5 [29], Al1.5NbTa0.5Ti1.5Zr0.5 and AlMo0.5NbTa0.5TiZr [30], HfNbTaTiZr, HfMoTaTiZr, HfMoNbTiZr and HfMoNbTaTi [31], NbMoTaW and NbMoTaWV [4], CrNbTiVZr [32,33], and HfMo0.5NbTaTiZr [34]. The results show that Al1.5NbTa0.5Ti1.5Zr0.5, CrNbTiVZr, and AlMo0.5NbTa0.5TiZr RHEAs have high SYSs, but their plasticity is inferior, which seriously affects the application of these RHEAs as structural materials. Although the SYSs of NbTiAl-1, 2, and 3 RHEAs at RT are in the middle of the range of representative RHEAs, all of them have superior plasticity compared with most of the RHEAs. Namely, a good combination of the SYS and the compression plasticity is realized in these RHEAs. A comparison of the SYSs of NbTiAl-1, 2, and 3 RHEAs with other representative RHEAs at different temperatures is shown in Figure 6b, such as NbMoTaW and VNbMoTaW [4], TaNbHfZrTi [7], NbTaTiV [35], CrNbTiZr and CrNbTiVZr [32], Zr2.0TiHfVNb2.0, [29], and Ti35Al15V20Nb20Mo10 [36]. The results show that the SYSs of NbTiAl-1, 2, and 3 RHEAs at 298 K are only lower than that of two RHEAs, CrNbTiZr and CrNbTiVZr. However, the SYSs of the NbTiAl-1, 2, and 3 RHEAs are higher than that of other RHEA systems, and the SYS of the NbTiAl-3 is the highest in the temperature range of 973 K~1173 K. It is obvious that the NbTiAl-3 has a substantially higher SYS ( σ YS / ρ ) from 298 K (191.3 MPa·cm3/g) to 1123 K (112.7 MPa·cm3/g) compared with other typical RHEAs. The superior mechanical property is mainly attributed to the solid solution strengthening induced by the Zr element in the B2 matrix. Meanwhile, the critical shear stress for the slip of dislocations in ordered crystalline structures is much higher than that in their solid solution counterparts [17].
Serious plastic deformation occurs near the GB after the fracture of the NbTiAl-3, and a large number of slip bands appear. The slip bands have two orientations, represented as slip band 1 and slip band 2 in Figure 7a, respectively, with an orientation difference of about 26° between them. The two-beam condition BF-TEM image with g = (01 1 _ ) shows two typical dislocation slip systems ([111](10 1 _ ) and [111](11 2 _ )), which are consistent with the orientation pile-up of slip bands shown in Figure 7a. There is an obvious dislocation pile-up in the slip bands while the two types of slip bands interact continually, thus providing considerable strain hardening capacity, as shown in Figure 4a. In particular, the appearance of {112} glide plane under RT deformation further provides the NbTiAl-3 with superior plastic deformation capability.
The electron backscatter diffraction (EBSD) results of the NbTiAl-3 after compression at 1123 K are shown in Figure 8. The compression axis (CA) is parallel to the vertical direction of Figure 8a. The grains are elongated along the horizontal direction, and the GBs slip significantly during compression, as shown in Figure 8a. Moreover, the geometrically necessary dislocation (GND) map shown in Figure 8b indicates that the dislocations are accumulated around the GBs, while the dislocation density is much lower inside the grains. Figure 8c shows that the dislocation density is 5.0 × 1012/m2 at the GBs, which is similar to the as-cast annealing state [37,38], whereas the dislocation densities around the GBs and inside the grains are 3.8 × 1014/m2 and 1.6 × 1014/m2, as shown in Figure 8d,e, respectively. It is precisely because of this heterogeneous deformation structure that it provides the necessary nucleation sites and driving forces for necklace structural dynamic recrystallization (DRX) at the GBs. Finally, the NbTiAl-3 forms the deformed structure that fine DRX grains are surrounded by high-density dislocations at the GBs after deformation, and the dislocation density gradually decreases, extending into the inside of the grains. Moreover, the overall dislocation density (1012/m2~1014/m2) of the sample is relative low [39], which proves that the dislocation annihilation occurs obviously during compression at 1123 K, resulting in the reduction in the dislocation strengthening effect. Hence, the DRX and dislocation annihilation accelerate the continuous strain softening after yielding.
The TEM results are also consistent with the EBSD analysis. Figure 9a shows the fine DRX grains along the initial GBs. Meanwhile, sub-grains with low-angle GBs are also detected, shown in Figure 9a, which indicates that the dynamic recovery (DRV) also occurs during compression. Accordingly, the dislocation density at the GBs is significantly reduced due to the consumption of DRX and DRV. Obvious dislocation tangles and high-density dislocation debris are detected surrounding the DRX and DRV regions, as shown in Figure 9b. Despite the fact that the NbTiAl-3 RHEA has superior SYS conferred by the existence of the ordered B2 matrix, DRX and DRV effectively reduce the stored elastic energy by consuming dislocations along the GBs. Meanwhile, dislocation cross-slip/climb and annihilation accelerated by thermal actuation are apparent inside the grains at 1123 K, resulting in continuous strain softening [9,10]. Figure 9c shows a schematic diagram of the microstructure of recrystallization and dislocation after compression at 1123 K.

4. Conclusions

Overall, three novel RHEAs, Nb42Ti25Al15Zr5Hf5Ta5Mo2W1 (NbTiAl-1), Nb39Ti23Al15Zr10Hf5Ta5Mo2W1 (NbTiAl-2), and Nb37Ti20Al15Zr15Hf5Ta5Mo2W1 (NbTiAl-3), with the B2 matrix and APBs were developed. The microstructure, RT, and high-temperature mechanical properties of the RHEAs in the as-cast state were systematically analyzed, and the deformation mechanism was also studied by taking the NbTiAl-3 RHEA as an example, with the following specific conclusions:
(1)
The RHEAs exhibit an excellent combination of high yield strength, low density (7.41~7.51 g/cm3), and large plasticity (ε > 0.35) at RT. The SYSs of the NbTiAl-3 RHEA are ~131 MPa·cm3/g at 1023 K and ~104.2 MPa·cm3/g at 1123 K, respectively, which are far superior to most typical RHEAs. These excellent mechanical properties are mainly attributed to the solid solution strengthening induced by the Zr element.
(2)
The deformation of the NbTiAl-3 RHEA at 298 K is dominated by the planar slip of dislocations, resulting in two dislocation slip bands oriented in the [111](10 1 _ ) and [111](11 2 _ ) directions. Dislocation pile-up in the slip bands and the two types of slip bands interact continually, providing considerable plastic deformation capability.
(3)
The DRX formed at GBs and low dislocation density inside the grains induced via dislocation annihilation accelerate continuous strain softening after yielding during the 1123 K compression of the NbTiAl-3 RHEA.

Author Contributions

Conceptualization, G.T., X.S., A.W., J.L. and H.Z. (Hongwei Zhang); Methodology, G.T., J.P., Y.J., A.W. and H.Z. (Hongwei Zhang); Validation, J.P., J.L. and H.Z. (Haifeng Zhang); Formal analysis, X.S. and Y.J.; Resources, J.P. and H.Z. (Haifeng Zhang); Data curation, X.S. and Y.J.; Writing—Original draft, G.T. and X.S.; Writing—Review & editing, J.P., J.L. and H.Z. (Hongwei Zhang); Visualization, J.P. and J.L.; Supervision, J.L., H.Z. (Haifeng Zhang) and H.Z. (Hongwei Zhang); Project administration, H.Z. (Hongwei Zhang). All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by Liaoning Key Research and Development Program (2022JH2/101300080), the Chinese Academy of Sciences (ZDBS-LY-JSC023), the Youth Innovation Promotion Association CAS (2021188), the IMR Innovation Fund (2023-PY16), and the Natural Science Foundation of Liaoning Province (2023-BS-012).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pollock, T.M.; Tin, S. Nickel-Based Superalloys for Advanced Turbine Engines: Chemistry, Microstructure and Properties. J. Propuls. Power 2006, 22, 361–374. [Google Scholar] [CrossRef]
  2. 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, 33, 3092–3128. [Google Scholar] [CrossRef]
  3. Hua, X.-J.; Hu, P.; Xing, H.-R.; Han, J.-Y.; Ge, S.-W.; Li, S.-L.; He, C.-J.; Wang, K.-S.; Cui, C.-J. Development and Property Tuning of Refractory High-Entropy Alloys: A Review. Acta Metall. Sin. 2022, 35, 1231–1265. [Google Scholar] [CrossRef]
  4. 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]
  5. Han, Z.D.; Chen, N.; Zhao, S.F.; Fan, L.W.; Yang, G.N.; Shao, Y.; Yao, K.F. Effect of Ti additions on mechanical properties of NbMoTaW and VNbMoTaW refractory high entropy alloys. Intermetallics 2017, 84, 153–157. [Google Scholar] [CrossRef]
  6. Yi, J.; Cao, F.; Xu, M.; Yang, L.; Wang, L.; Zeng, L. Phase, microstructure and compressive properties of refractory high-entropy alloys CrHfNbTaTi and CrHfMoTaTi. Int. J. Miner. Metall. Mater. 2022, 29, 1231–1236. [Google Scholar] [CrossRef]
  7. Senkov, O.N.; Scott, J.M.; Senkova, S.V.; Meisenkothen, F.; Miracle, D.B.; Woodward, C.F. Microstructure and elevated temperature properties of a refractory TaNbHfZrTi alloy. J. Mater. Sci. 2012, 47, 4062–4074. [Google Scholar] [CrossRef]
  8. Wang, M.; Ma, Z.L.; Xu, Z.Q.; Cheng, X.W. Effects of vanadium concentration on mechanical properties of VxNbMoTa refractory high-entropy alloys. Mater. Sci. Eng. A 2021, 808, 140848. [Google Scholar] [CrossRef]
  9. Pang, J.; Zhang, H.; Zhang, L.; Zhu, Z.; Fu, H.; Li, H.; Wang, A.; Li, Z.; Zhang, H. A ductile Nb40Ti25Al15V10Ta5Hf3W2 refractory high entropy alloy with high specific strength for high-temperature applications. Mater. Sci. Eng. A 2022, 831, 142290. [Google Scholar] [CrossRef]
  10. 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]
  11. Eleti, R.R.; Chokshi, A.H.; Shibata, A.; Tsuji, N. Unique high-temperature deformation dominated by grain boundary sliding in heterogeneous necklace structure formed by dynamic recrystallization in HfNbTaTiZr BCC refractory high entropy alloy. Acta Mater. 2020, 183, 64–77. [Google Scholar] [CrossRef]
  12. Jia, Y.; Wang, G.; Wu, S.; Mu, Y.; Yi, Y.; Jia, Y.; Liaw, P.K.; Zhang, T.; Liu, C.-T. A lightweight refractory complex concentrated alloy with high strength and uniform ductility. Appl. Mater. Today 2022, 27, 101429. [Google Scholar] [CrossRef]
  13. Pang, J.; Xiong, T.; Wei, X.; Zhu, Z.; Zhang, B.; Zhou, Y.; Shao, X.; Jin, Q.; Zheng, S.; Ma, X. Oxide MnCr2O4 induced pitting corrosion in high entropy alloy CrMnFeCoNi. Materialia 2019, 6, 100275. [Google Scholar] [CrossRef]
  14. An, Z.; Mao, S.; Liu, Y.; Wang, L.; Zhou, H.; Gan, B.; Zhang, Z.; Han, X. A novel HfNbTaTiV high-entropy alloy of superior mechanical properties designed on the principle of maximum lattice distortion. J. Mater. Sci. Technol. 2021, 79, 109–117. [Google Scholar] [CrossRef]
  15. 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]
  16. Feng, R.; Zhang, C.; Gao, M.C.; Pei, Z.; Zhang, F.; Chen, Y.; Ma, D.; An, K.; Poplawsky, J.D.; Ouyang, L.; et al. High-throughput design of high-performance lightweight high-entropy alloys. Nat. Commun. 2021, 12, 4329. [Google Scholar] [CrossRef]
  17. Qiao, D.; Liang, H.; Wu, S.; He, J.; Cao, Z.; Lu, Y.; Li, T. The mechanical and oxidation properties of novel B2-ordered Ti2ZrHf0.5VNb0.5Alx refractory high-entropy alloys. Mater. Charact. 2021, 178, 111287. [Google Scholar] [CrossRef]
  18. Chen, H.; Kauffmann, A.; Seils, S.; Boll, T.; Liebscher, C.H.; Harding, I.; Kumar, K.S.; Szabó, D.V.; Schlabach, S.; Kauffmann-Weiss, S.; et al. Crystallographic ordering in a series of Al-containing refractory high entropy alloys Ta–Nb–Mo–Cr–Ti–Al. Acta Mater. 2019, 176, 123–133. [Google Scholar] [CrossRef]
  19. Soni, V.; Gwalani, B.; Alam, T.; Dasari, S.; Zheng, Y.; Senkov, O.N.; Miracle, D.; Banerjee, R. Phase Inversion in a Two-phase, BCC+B2, Refractory High Entropy Alloy. Acta Mater. 2019, 185, 89–97. [Google Scholar] [CrossRef]
  20. Soni, V.; Senkov, O.N.; Gwalani, B.; Miracle, D.B.; Banerjee, R. Microstructural Design for Improving Ductility of An Initially Brittle Refractory High Entropy Alloy. Sci. Rep. 2018, 8, 8816. [Google Scholar] [CrossRef]
  21. Pang, J.; Zhang, H.; Ji, Y.; Zhu, Z.; Zhang, L.; Li, H.; Wang, A.; Zhang, H. High-temperature structural and mechanical stability of refractory high-entropy alloy Nb40Ti25Al15V10Ta5Hf3W2. Mater. Charact. 2023, 205, 113321. [Google Scholar] [CrossRef]
  22. Yurchenko, N.; Panina, E.; Tojibaev, A.; Novikov, V.; Salishchev, G.; Zherebtsov, S.; Stepanov, N. Effect of B2 ordering on the tensile mechanical properties of refractory AlxNb40Ti40V20−x medium-entropy alloys. J. Alloys Compd. 2023, 937, 168465. [Google Scholar] [CrossRef]
  23. Yurchenko, N.; Panina, E.; Tojibaev, A.; Novikov, V.; Salishchev, G.; Zherebtsov, S.; Stepanov, N. Tuning the grain and domain sizes to achieve superior room-temperature tensile ductility in a B2-ordered refractory Al15Nb40Ti40V5 medium-entropy alloy. Mater. Sci. Eng. A 2023, 874, 145073. [Google Scholar] [CrossRef]
  24. Yurchenko, N.Y.; Stepanov, N.D.; Zherebtsov, S.V.; Tikhonovsky, M.A.; Salishchev, G.A. Structure and mechanical properties of B2 ordered refractory AlNbTiVZrx (x = 0–1.5) high-entropy alloys. Mater. Sci. Eng. A 2017, 704, 82–90. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Liu, J.P.; Chen, S.Y.; Xie, X.; Liaw, P.K.; Dahmen, K.A.; Qiao, J.W.; Wang, Y.L. Serration and noise behaviors in materials. Prog. Mater. Sci. 2017, 90, 358–460. [Google Scholar] [CrossRef]
  26. Ge, S.; Fu, H.; Zhang, L.; Mao, H.; Li, H.; Wang, A.; Li, W.; Zhang, H. Effects of Al addition on the microstructures and properties of MoNbTaTiV refractory high entropy alloy. Mater. Sci. Eng. A 2020, 784, 139275. [Google Scholar] [CrossRef]
  27. Toda-Caraballo, I.; Rivera-Diaz-del-Castillo, P.E.J. Modelling solid solution hardening in high entropy alloys. Acta Mater. 2015, 85, 14–23. [Google Scholar] [CrossRef]
  28. Li, Q.Y.; Zhang, H.; Li, D.C.; Chen, Z.H.; Huang, S.; Lu, Z.L.; Yan, H.Q. WxNbMoTa Refractory High-Entropy Alloys Fabricated by Laser Cladding Deposition. Materials 2019, 12, 533. [Google Scholar] [CrossRef]
  29. Chen, Y.; Li, Y.; Cheng, X.; Wu, C.; Cheng, B.; Xu, Z. The Microstructure and Mechanical Properties of Refractory High-Entropy Alloys with High Plasticity. Materials 2018, 11, 208. [Google Scholar] [CrossRef]
  30. 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]
  31. Tseng, K.-K.; Juan, C.-C.; Tso, S.; Chen, H.-C.; Tsai, C.-W.; Yeh, J.-W. Effects of Mo, Nb, Ta, Ti, and Zr on Mechanical Properties of Equiatomic Hf-Mo-Nb-Ta-Ti-Zr Alloys. Entropy 2019, 21, 15. [Google Scholar] [CrossRef]
  32. 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]
  33. Senkov, O.N.; Senkova, S.V.; Woodward, C.; Miracle, D.B. Low-density, refractory multi-principal element alloys of the Cr–Nb–Ti–V–Zr system: Microstructure and phase analysis. Acta Mater. 2013, 61, 1545–1557. [Google Scholar] [CrossRef]
  34. Juan, C.-C.; Tseng, K.-K.; Hsu, W.-L.; Tsai, M.-H.; Tsai, C.-W.; Lin, C.-M.; Chen, S.-K.; Lin, S.-J.; Yeh, J.-W. Solution strengthening of ductile refractory HfMoxNbTaTiZr high-entropy alloys. Mater. Lett. 2016, 175, 284–287. [Google Scholar] [CrossRef]
  35. Guo, W.; Liu, B.; Liu, Y.; Li, T.; Fu, A.; Fang, Q.; Nie, Y. Microstructures and mechanical properties of ductile NbTaTiV refractory high entropy alloy prepared by powder metallurgy. J. Alloys Compd. 2019, 776, 428–436. [Google Scholar] [CrossRef]
  36. Xu, Z.Q.; Ma, Z.L.; Wang, M.; Chen, Y.W.; Tan, Y.D.; Cheng, X.W. Design of novel low-density refractory high entropy alloys for high-temperature applications. Mater. Sci. Eng. A 2019, 755, 318–322. [Google Scholar] [CrossRef]
  37. Thirathipviwat, P.; Song, G.; Jayaraj, J.; Bednarcik, J.; Wendrock, H.; Gemming, T.; Freudenberger, J.; Nielsch, K.; Han, J. A comparison study of dislocation density, recrystallization and grain growth among nickel, FeNiCo ternary alloy and FeNiCoCrMn high entropy alloy. J. Alloys Compd. 2019, 790, 266–273. [Google Scholar] [CrossRef]
  38. Williamson, G.K.; Smallman, R.E., III. Dislocation densities in some annealed and cold-worked metals from measurements on the X-ray debye-scherrer spectrum. Philos. Mag. A J. Theor. Exp. Appl. Phys. 1956, 1, 34–46. [Google Scholar] [CrossRef]
  39. Jiang, J.; Britton, T.B.; Wilkinson, A.J. Evolution of dislocation density distributions in copper during tensile deformation. Acta Mater. 2013, 61, 7227–7239. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3 RHEAs.
Figure 1. XRD pattern of as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3 RHEAs.
Materials 16 07592 g001
Figure 2. SEM images of as-cast (a) NbTiAl-1, (b) NbTiAl-2, (c) NbTiAl-3 RHEAs, (d) EDS maps of as-cast NbTiAl-3 RHEA.
Figure 2. SEM images of as-cast (a) NbTiAl-1, (b) NbTiAl-2, (c) NbTiAl-3 RHEAs, (d) EDS maps of as-cast NbTiAl-3 RHEA.
Materials 16 07592 g002
Figure 3. DF-TEM images with the superlattice spots of (100) of as-cast (a) NbTiAl-1, (b) NbTiAl-2, (c) NbTiAl-3 RHEAs. The insets in (ac) are selected-area electron diffraction (SAED) patterns with [001] zone axis.
Figure 3. DF-TEM images with the superlattice spots of (100) of as-cast (a) NbTiAl-1, (b) NbTiAl-2, (c) NbTiAl-3 RHEAs. The insets in (ac) are selected-area electron diffraction (SAED) patterns with [001] zone axis.
Materials 16 07592 g003
Figure 4. Compressive true stress–strain curves of as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3 RHEAs at (a) 298 K, (b) 1023 K, (c) 1123 K, and (d) 1223 K.
Figure 4. Compressive true stress–strain curves of as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3 RHEAs at (a) 298 K, (b) 1023 K, (c) 1123 K, and (d) 1223 K.
Materials 16 07592 g004
Figure 5. The increment in the yield strength, Δσ, at T = 298 K, 1023 K, 1123 K, and 1223 K as a function of the Zr concentration, CZr, in the matrix phase of the RHEAs.
Figure 5. The increment in the yield strength, Δσ, at T = 298 K, 1023 K, 1123 K, and 1223 K as a function of the Zr concentration, CZr, in the matrix phase of the RHEAs.
Materials 16 07592 g005
Figure 6. (a) The SYS ( σ YS / ρ )-fracture strain of as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3 RHEAs and other representative RHEAs at RT, (b) the SYS ( σ YS / ρ )-temperature of as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3 RHEAs and other representative RHEAs [4,7,24,28,29,30,31,32,33,34,35,36].
Figure 6. (a) The SYS ( σ YS / ρ )-fracture strain of as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3 RHEAs and other representative RHEAs at RT, (b) the SYS ( σ YS / ρ )-temperature of as-cast NbTiAl-1, NbTiAl-2, and NbTiAl-3 RHEAs and other representative RHEAs [4,7,24,28,29,30,31,32,33,34,35,36].
Materials 16 07592 g006
Figure 7. (a) SEM image showing two independent slip traces and (b) two-beam condition BF-TEM image with g = (01 1 _ ) showing two typical dislocation slip systems ([111](10 1 _ ) and [111](11 2 _ )) of as-cast NbTiAl-3 RHEA after fracture at 298 K.
Figure 7. (a) SEM image showing two independent slip traces and (b) two-beam condition BF-TEM image with g = (01 1 _ ) showing two typical dislocation slip systems ([111](10 1 _ ) and [111](11 2 _ )) of as-cast NbTiAl-3 RHEA after fracture at 298 K.
Materials 16 07592 g007
Figure 8. EBSD results after compression true strain 50% at 1123 K. (a) Inverse pole figure (IPF) maps. (b) GND maps. (ce) GND values corresponding to the selected regions ①–③, at the GBs, nearby the GBs, and the grain interior, respectively, shown in (b).
Figure 8. EBSD results after compression true strain 50% at 1123 K. (a) Inverse pole figure (IPF) maps. (b) GND maps. (ce) GND values corresponding to the selected regions ①–③, at the GBs, nearby the GBs, and the grain interior, respectively, shown in (b).
Materials 16 07592 g008
Figure 9. (a) BF-TEM image, (b) two-beam condition BF-TEM image with g = (00 1 _ ), and (c) deformation schematic diagram of as-cast NbTiAl-3 RHEA after 50% compression at 1123 K.
Figure 9. (a) BF-TEM image, (b) two-beam condition BF-TEM image with g = (00 1 _ ), and (c) deformation schematic diagram of as-cast NbTiAl-3 RHEA after 50% compression at 1123 K.
Materials 16 07592 g009
Table 1. Chemical compositions of the as-cast NbTiAl-1, 2, and 3 RHEAs (at.%).
Table 1. Chemical compositions of the as-cast NbTiAl-1, 2, and 3 RHEAs (at.%).
Nominal CompositionNbTiAlZrHfTaMoW
Nb42Ti25Al15Zr5Hf5Ta5Mo2W141.225.714.34.65.26.01.41.6
Nb39Ti23Al15Zr10Hf5Ta5Mo2W140.423.713.68.65.15.81.11.7
Nb37Ti20Al15Zr15Hf5Ta5Mo2W137.720.314.113.55.36.21.81.0
Table 2. Values of compression yield strength and SYS of as-cast NbTiAl-1, 2, and 3 RHEAs at different temperatures (Unit: MPa and MPa·g−1·cm3).
Table 2. Values of compression yield strength and SYS of as-cast NbTiAl-1, 2, and 3 RHEAs at different temperatures (Unit: MPa and MPa·g−1·cm3).
Temperature (K)298102311231223
NbTiAl-1σYS1198.1 ± 7.5705.6 ± 12.8705.5 ± 12.5386.2 ± 9.2
σYS/ρ159.593.993.951.4
NbTiAl-2σYS1333.1 ± 7.2857.9 ± 13.1764.6 ± 12.7451.8 ± 11
σYS/ρ179.1115.3102.760.7
NbTiAl-3σYS1418.1 ± 7.4951.8 ± 11.5835.5 ± 13466.3 ± 11.4
σYS/ρ191.3128.4112.762.9
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

Tang, G.; Shao, X.; Pang, J.; Ji, Y.; Wang, A.; Li, J.; Zhang, H.; Zhang, H. The Microstructures, Mechanical Properties, and Deformation Mechanism of B2-Hardened NbTiAlZr-Based Refractory High-Entropy Alloys. Materials 2023, 16, 7592. https://doi.org/10.3390/ma16247592

AMA Style

Tang G, Shao X, Pang J, Ji Y, Wang A, Li J, Zhang H, Zhang H. The Microstructures, Mechanical Properties, and Deformation Mechanism of B2-Hardened NbTiAlZr-Based Refractory High-Entropy Alloys. Materials. 2023; 16(24):7592. https://doi.org/10.3390/ma16247592

Chicago/Turabian Style

Tang, Guangquan, Xu Shao, Jingyu Pang, Yu Ji, Aimin Wang, Jinguo Li, Haifeng Zhang, and Hongwei Zhang. 2023. "The Microstructures, Mechanical Properties, and Deformation Mechanism of B2-Hardened NbTiAlZr-Based Refractory High-Entropy Alloys" Materials 16, no. 24: 7592. https://doi.org/10.3390/ma16247592

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