Next Article in Journal
Instantaneous Ablation Behavior of Laminated CFRP by High-Power Continuous-Wave Laser Irradiation in Supersonic Wind Tunnel
Next Article in Special Issue
High-Strength Ductility Joining of Multicomponent Alloy to 304 Stainless Steel Using Laser Welding Technique
Previous Article in Journal
In Situ Synthesis of Hierarchical Flower-like Sn/SnO2 Heterogeneous Structure for Ethanol GAS Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Microstructure of High-Activity Spherical TaNbTiZr Refractory High-Entropy Alloy Powders

State Key Laboratory of Powder Metallurgy, Central South University, Changsha 410083, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(2), 791; https://doi.org/10.3390/ma16020791
Submission received: 15 December 2022 / Revised: 9 January 2023 / Accepted: 10 January 2023 / Published: 13 January 2023

Abstract

:
High-activity spherical TaNbTiZr refractory high-entropy alloy (REHA) powders were successfully prepared by electrode induction melting gas atomization (EIGA) and plasma rotating electrode process (PREP) methods. Both the EIGAed and PREPed TaNbTiZr RHEA powders have a single-phase body-centered cubic (BCC) structure and low oxygen content. Compared with the EIGAed powders, the PREPed powders exhibit higher sphericity and smoother surface, but larger particle size. The average particle sizes of the EIGAed and PREPed powders are 51.8 and 65.9 μm, respectively. In addition, both the coarse EIGAed and PREPed powders have dendritic structure, and the dendrite size of the EIGAed powders is larger than that of the PREPed powders. Theoretical calculation indicates that the cooling rate of the PREPed powders is one order of magnitude higher than that of the EIGAed powders during the solidification process, and the dendritic structure has more time to grow during EIGA, which is the main reason for the coarser dendrite size of the EIGAed powders.

1. Introduction

High-entropy alloys (HEAs), especially the refractory high entropy alloys (RHEAs), have been attracting tremendous attention due to their attractive mechanical properties [1,2,3,4,5]. RHEAs usually consist of four or more refractory metallic elements, such as W, Mo, Ta, Nb, Hf, Ti, Zr and V, etc., which endow the alloy system outstanding high-temperature properties, including high melting point [6,7,8,9], high strength and hardness [10,11,12], good oxidation resistance [13], and outstanding thermal stability [14]. For example, WNbMoTa RHEA with a single-phase body-centered cubic (BCC) structure has an ultra-high yield strength of 405 MPa at 1600 °C [9,10], which makes it possible to replace the widely used Ni-based superalloys and be used as the next-generation high-temperature materials. However, the RHEAs containing W and Mo elements generally show high room-temperature brittleness and densities (>13 g/cm3 [9]). The TaNbTiZr RHEA developed recently by replacing W and Mo with Ti and Zr can overcome the brittle and heavy bottlenecks and exhibits low density (8.9 g/cm3 [15]), as well as high yield strength (410 MPa at 1000 °C [16]) and fracture strain (>48% at room temperature [17]), which has attracted extensive interests.
However, the high melting temperature of RHEAs will inevitably bring challenges for traditional manufacturing technologies, e.g., casting, extrusion and forging to produce high-quality complex components. Additive manufacturing (AM) is an emerging direct forming technology capable of producing complex components that can effectively solve the troubles of poor formability and extend the applications of RHEAs. Spherical powders are the raw material for AM, and the sphericity is a key parameter affecting product quality [18,19]. Although extensive efforts have shown that the preparation of RHEAs can be achieved by AM of elemental powders, the AMed products have severe compositional segregation and unstable mechanical properties due to the large differences in the melting points of the employed elements [20,21]. The use of pre-alloyed spherical powders with a targeted composition can eliminate the above problems, thereby the preparation of the pre-alloyed powders is extremely important.
At present, inductively coupled thermal plasma spheroidization method is the main method for preparing spherical RHEA pre-alloyed powders. For example, Xia et al. [22] prepared the WTaMoNbZr RHEA powders by vacuum electron beam melting (VEBM), hydrogenation, and disk-milled and plasma spheroidization. The spheroidized powders have good sphericity and homogeneous elemental distribution, but a high oxygen content of 1678 ppm. Na et al. [23] also fabricated the spherical TaNbHfZrTi RHEA powders through vacuum arc melting, hydrogenation, jaw crushing, and ball-milling and plasma spheroidization, and the oxygen content still reaches 1650 ppm. The above studies show that although the plasma spheroidization method can achieve the preparation of spherical RHEA powders, however, the preparation process contains a complex hydrogenation-dehydrogenation reaction involving crucible melting and mechanical crushing, which can easily lead to the introduction of impurity elements, especially the oxygen.
Hence, how to realize the preparation of high-activity RHEA powders with high sphericity and purity is still a difficult problem to solve up to now. In this work, we chose the methods of cold crucible levitation melting (CCLM) combined with electrode induction melting gas atomization (EIGA) and plasma rotating electrode process (PREP), respectively, to prepare spherical TaNbTiZr RHEA powders. Since the preparation process is simple and all of the above preparation methods belong to crucible-free fabrication technology [24,25], the oxygen content of the achieved powders is low. More importantly, industrial-scale production of more than 100 kg of powder has been obtained by both the EIGA and PREP methods. The detailed characterization of the powders and the influence of the cooling rate during the solidification process on the microstructure of the powders were investigated.

2. Materials and Methods

The TaNbTiZr RHEA ingots were fabricated from pure metals (purity > 99%) by CCLM and subsequently cast into the dimensions of Φ50 × 300 mm. Then, these bars were atomized into spherical powders via the EIGA and PREP methods, respectively, under high-purity argon atmosphere. Figure 1 shows the schematic diagrams of the EIGA and PREP systems. The EIGA method uses a tightly coupled nozzle to atomize the molten alloy into fine droplets by a high-pressure airflow, while the PREP method relies on the centrifugal force of the electrode rotating at high speed to crush the molten alloy into fine droplets [26,27,28]. The droplets cool down rapidly under an argon atmosphere and eventually form pre-alloyed powders. The vacuum before the argon purge in the EIGA and PREP devices is lower than 5 Pa and 10 Pa, respectively. The atomization pressure in the EIGA device is 3.8 MPa, and the electrode rotation speed in the PREP device is 20,000 rpm. Subsequently, both the EIGAed and PREPed powders with the particle size less than 105 μm were sieved for further microstructural characterization.
Metallic element contents were measured by an inductively coupled plasma mass spectroscopy (ICP-MS). Oxygen/hydrogen/nitrogen and carbon contents were determined by an O/N/H analyzer (LECO, TCH-600) and a C/S analyzer (LECO, CS-600), respectively. Powder size distribution was analyzed by a laser diffraction particle size analyzer (Malvern, Micro-plus). Phase constitution was detected by an X-ray diffractometer (XRD, Bruker D8) using Cu Kα radiation. Microstructural characterization was investigated by a scanning electron microscopy (SEM, FEI Nova Nano230) equipped with an electron-backscattered diffraction (EBSD) analyzer. Chemical analysis at the macroscale was performed by an Electron Probe Micro-analyzer (EPMA, JXA-8530F).

3. Results

3.1. Characterization of the TaNbTiZr RHEA Bars

The TaNbTiZr RHEA bars with 50 mm in diameter and 300 mm in length were prepared by CCLM, as shown in Figure 2a. The XRD pattern of the TaNbTiZr RHEA is given in Figure 2b, showing a single-phase BCC structure. The microstructure of the TaNbTiZr RHEA is displayed in Figure 2c, and the inverse pole figure (IPF) map reveals that the TaNbTiZr RHEA consists of equiaxial grains with a random grain-orientation distribution. The grain size distribution is summarized in Figure 2d, and the average grain size is nearly 134.69 μm. The oxygen content of the TaNbTiZr RHEA bars is ~488 ppm. To further investigate the composition uniformity of the TaNbTiZr RHEA bars, Figure 3 demonstrates the back-scattered electron (BSE) image of the TaNbTiZr RHEA and the corresponding elemental distribution maps of Ta, Nb, Ti, and Zr. It can be clearly seen that the elemental segregation is very slight, and no detectable secondary phases exist, which indicates that the TaNbTiZr RHEA bars prepared by CCLM have good compositional homogeneity. This provides a strong guarantee for the subsequent preparation of powders with uniform composition and low oxygen content.

3.2. Characterization of the TaNbTiZr RHEA Powders

SEM images of the EIAGed and PREPed TaNbTiZr RHEA powders are shown in Figure 4. It can be seen that the EIGAed powders (Figure 4a) are nearly spherical in shape, but there are also some irregular shaped particles and satellite powders. The reason for the formation of satellite powders is mainly due to the fact that during gas atomization, the small molten droplets have a faster cooling rate and flying speed, which can easily collide with large molten droplets and solidify around them to form satellite powders [29]. The particle size distribution of the EIGAed powders is plotted in Figure 5, showing a typical Gaussian distribution. The particle size of the EIGA powder is mainly in the range of 30 to 100 μm, with a mean particle size (D50) of 51.8 μm. The high-magnified SEM images represented in Figure 4a1,a2 show that the fine powders have a smooth surface, while the surface of the coarse powders is rough due to the appearance of the dendritic structure. Compared with the EIGAed powders, the PREPed powders exhibit higher sphericity and smoother surface (Figure 4b). The PREPed powders also have a narrower size distribution with a D50 of 65.9 μm. Figure 4b1,b2 represent the SEM images of the PREPed powders at higher magnification, where the surfaces of the PREPed powders are smoother than the EIGAed powder at similar particle size. In addition, XRD analysis performed on the EIGAed and PREPed powders are shown in Figure 6, and all the powders keep a single-phase BCC structure from the as-cast state, indicating that no secondary phase formed during the preparation of powders.
The chemical composition of the EIGAed and PREPed TaNbTiZr RHEA powders are listed in Table 1. It can be seen that the actual compositions of the EIGAed and PREPed powders are in good agreement with the nominal composition, indicating that the powders fabricated by these two methods have good composition uniformity. In addition, the contents of impurity elements, especially the oxygen, are low. The oxygen contents of the spherical TaNbTiZr RHEA powders prepared in this work and other similar spherical RHEA powders [22,23,30] are listed in Table 2. Obviously, the oxygen content in the EIGAed powders (845 ppm) and PREPed powders (777 ppm) is significantly lower than that of the spherical high-activity RHEA powders prepared by plasma spheroidization method (~1700 ppm).
The cross-sectional microstructure of the EIGAed and PREPed TaNbTiZr RHEA powders are shown in Figure 7. The fine powders and coarse powders exhibit different microstructure, i.e., the fine powders exhibit dendrite-free structure (Figure 7a,c), while the coarse powders have typical dendritic structure (Figure 7b,d). Moreover, it can also be seen that the dendritic structure of the EIGAed powders is coarser than that of the PREPed powders. The dendrite arm spacings in the EIGAed and PREPed powders are measured, and the corresponding values are 1.88 μm and 1.62 μm, respectively. The elemental line scannings in the coarse EIGAed and PREPed powders are shown in Figure 7b1,d1, respectively, and the curves of each element fluctuate obviously, indicating that there is elemental segregation in the dendritic structure. The composition fluctuation of the EIGAed powders is significantly larger than that of the PREPed powders, which means that the segregation in the EIGAed powders is more serious. The EMPA elemental mapping (Figure 8) was conducted to further investigate the elemental distribution of each element in the dendritic structure, and the EIGAed and PREPed powders show a similar elemental distribution tendency, that is, the Ta element with a higher melting point is enriched in the dendrite region, the Ti and Zr elements with lower melting point exhibit the opposite tendency, and the Nb element is evenly distributed in the dendritic and interdendritic regions.

4. Discussion

The microstructure of the powders is usually closely related to the cooling rate [31]. To analyze the effect of the cooling rate on the microstructure during the rapid solidification process, the cooling rate of spherical powders with various particle sizes under different preparation processes are calculated by numerical simulation.
For the EIGAed TaNbTiZr RHEA powders, the cooling rate Vc1 of the powders can be calculated by the following formula [32,33]:
V c   1 = 6 h 1 C d ρ d D T d     T g
where D is the diameter of the TaNbTiZr RHEA droplets; Cd is the theoretical specific heat capacity of the TaNbTiZr RHEA; ρd is the theoretical density of the TaNbTiZr RHEA droplets, which is usually close to that of the solid state; h1 is the convection heat transfer coefficient; Td is the TaNbTiZr RHEA droplet temperature; Tg is the argon temperature; and h1 can be expressed by the Ranz-Marshall relation [34]:
h 1 = k g D 2 + 0.6 R e P r 3
where kg is the argon thermal conductivity; R e = ρ g D u d u g μ g is the droplet Reynolds; P r = C p g μ g k g is the argon Prandtl number; ρg is the argon density; u d     u g is the velocity difference between droplet and airflow; μg is the argon dynamic viscosity; and Cpg is the argon specific heat capacity per unit mass. When ignoring the speed difference between the high-speed argon and the atomized droplets flying during the atomization process [33], the Vc1 can be rewritten as:
V c 1   = 12 k g C d ρ d D 2 T d   T g
For the PREPed TaNbTiZr RHEA powders, the cooling rate Vc2 of the powders can be described as [31]:
V c 2 = 6 h 2 C d ρ d D T d   T g
where h2 is the heat transfer coefficient, and can be calculated by the following formula [31]:
h 2 = 2 k g D + 0.6 k g 4 ρ g 3 C g 2 y g 1 6 v D 1 2
where Cg is the argon specific heat capacity; yg is the viscosity of argon; and v is the linear velocity of the rotating electrode edge, which is calculated to be 5.24 × 104 mm s−1, since the electrode has a diameter d of 50 mm and rotation speed r is 20,000 rpm. Hence, the Vc2 can be rewritten as:
V c 2 = 12 k g C d ρ d D 2 T d   T g + 3.6 h 2 C d ρ d T d     T g k g 4 ρ g 3 C g 2 y g 1 6 v 1 2 1 D 3 2
Table 3 summarizes the thermophysical constants of the TaNbTiZr RHEA and argon, and the preparation parameters. By substituting the above constants/parameters into Equations (3) and (6), respectively, the cooling rate of the EIGAed and PRERed powders can be expressed as:
V c 1 = 4.68   ×   10 8 1 D 2
V c 2 = 4.68   ×   10 8 1 D 2 + 108.65   ×   10 6 1 D 3 2
According to the above formulas, the cooling rate of powders is negatively correlated with particle size during the solidification process. Meanwhile, for the powders with similar particle sizes, the cooling rate varies greatly under different preparation methods. For example, when the particle size is 18 μm, the cooling rates of the EIGAed and PREPed powders are 1.44 × 106 and 2.87 × 106 K/s, respectively; when the particle size is 80 μm, the cooling rates of the EIGAed and PREPed powders are 7.31 × 104 and 2.25 × 105 K/s, respectively. Obviously, for the same preparation method, the coarse powders have a longer solidification time than fine powders, so the dendritic structure is hard to form in the fine powders while the coarse powders have dendritic structure. For powders with the same particle size, the cooling rate of the PREPed powders is one order of magnitude higher than that of the EIGAed powders. The time for dendrite growth during PREP is significantly shorter than that in the EIGA process. Hence, the dendrite size in the PREPed powders is finer. Similar phenomena were also reported in the preparation of other pre-alloyed powders [35,36]. For example, He et al. [31] found that the microstructure of the PREPed high-Nb TiAl powders is closely related to the particle size, that is, coarse particles usually present dendritic structures, while finer particles exhibit featureless smooth structure.

5. Conclusions

In this study, novel high-activity spherical TaNbTiZr REHA powders were successfully prepared by EIGA and PREP methods. The phase composition and microstructure of the powders were characterized. The effect of cooling rate during the solidification process on microstructure evolution was investigated. The main conclusions are summarized as follows:
(1) Both the EIGAed and PREPed TaNbTiZr powders exhibit a single-phase BCC structure and have a low oxygen content (845 and 777 ppm for the EIGAed and PREPed powders, respectively).
(2) The PREPed powders show higher sphericity and smoother surface compared with the EIGAed powders. The average particle size (65.9 μm) of the PREPed powders is slightly larger than that of the EIGAed powders (51.8 μm).
(3) The dendritic structure appears in the coarse powders, and the dendrite size in the EIGAed powders is larger than that in the PREPed powders. The low cooling rate during EIGA is considered to be responsible for the larger dendrite size of the EIGAed powders.
(4) This work highlights the characters of TaNbTiZr RHEA powders prepared by EIGA and PREP method, but variable parameters is not studied. Future work focusing on the relation between processing parameters and powder properties is warranted.

Author Contributions

S.G.: Methodology, Formal analysis, Writing—original draft. A.F.: Methodology, Investigation, Formal analysis. Z.X.: Investigation, Formal analysis. T.L.: Investigation, Writing—review and editing. Y.C.: Writing—review and editing. B.L.: Funding acquisition, Project administration, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by the National Natural Science Foundation of China [No. 52020105013 and No. 52104365] and the Hubei Provincial Natural Science Foundation of China [No. 2022CFB894].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. George, E.P.; Raabe, D.; Ritchie, R.O. High-entropy alloys. Nat. Rev. Mater. 2019, 4, 515–534. [Google Scholar] [CrossRef]
  2. Baker, I. Interstitial strengthening in fcc metals and alloys. Adv. Powder Mater. 2022, 1, 100034. [Google Scholar] [CrossRef]
  3. Fu, A.; Liu, B.; Li, Z.; Wang, B.; Cao, Y.; Liu, Y. Dynamic deformation behavior of a FeCrNi medium entropy alloy. J. Mater. Sci. Technol. 2022, 100, 120–128. [Google Scholar] [CrossRef]
  4. Coury, F.G.; Kaufman, M.; Clarke, A.J. Solid-solution strengthening in refractory high entropy alloys. Acta Mater. 2019, 175, 66–81. [Google Scholar] [CrossRef]
  5. Fu, A.; Liu, B.; Lu, W.; Liu, B.; Li, J.; Fang, Q.; Li, Z.; Liu, Y. A novel supersaturated medium entropy alloy with superior tensile properties and corrosion resistance. Scr. Mater. 2020, 186, 381–386. [Google Scholar] [CrossRef]
  6. Lee, C.; Kim, G.; Chou, Y.; Musicó, B.L.; Gao, M.C.; An, K.; Song, G.; Chou, Y.C.; Keppens, V.; Chen, W.; et al. Temperature dependence of elastic and plastic deformation behavior of a refractory high-entropy alloy. Sci. Adv. 2020, 6, eaaz4748. [Google Scholar] [CrossRef]
  7. Xiong, W.; Guo, A.X.; Zhan, S.; Liu, C.T.; Yeh, J.W.; Cao, S.C. Refractory high-entropy alloys: A focused review of preparation methods and properties. J. Mater. Sci. Technol. 2022, 142, 196–215. [Google Scholar] [CrossRef]
  8. Ren, X.; Li, Y.; Qi, Y.; Wang, B. Review on Preparation Technology and Properties of Refractory High Entropy Alloys. Materials 2022, 15, 2931. [Google Scholar] [CrossRef]
  9. 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]
  10. Senkov, O.N.; Wilks, G.B.; Scott, J.M.; Miracle, D.B. Mechanical properties of Nb25Mo25Ta25W25 and V20Nb20Mo20Ta20W20 refractory high entropy alloys. Intermetallics 2011, 19, 698–706. [Google Scholar] [CrossRef]
  11. Senkov, O.N.; Senkova, S.V.; Woodward, C.J.A.M. Effect of aluminum on the microstructure and properties of two refractory high-entropy alloys. Acta Mater. 2014, 68, 214–228. [Google Scholar] [CrossRef]
  12. Senkov, O.N.; Woodward, C.F. Microstructure and properties of a refractory NbCrMo0.5Ta0.5TiZr alloy. Mater. Sci. Eng. A 2011, 529, 311–320. [Google Scholar] [CrossRef]
  13. Gorr, B.; Azim, M.; Christ, H.J.; Mueller, T.; Schliephake, D.; Heilmaier, M. Phase equilibria, microstructure, and high temperature oxidation resistance of novel refractory high-entropy alloys. J. Alloys Compd. 2015, 624, 270–278. [Google Scholar] [CrossRef]
  14. Sun, C.; Guo, Y.; Yang, Z.; Li, J.; Xi, S.; Jie, Z.; Xu, T. Microstructurally stable nanocomposite WTaMoNb/Cu prepared by mechanical alloying and hot pressing sintering. Mater. Lett. 2022, 306, 130894. [Google Scholar] [CrossRef]
  15. Yang, M.; Shao, L.; Duan, J.M.; Chen, X.T.; Tang, B.Y. Temperature dependence of mechanical and thermodynamic properties of Ti(25+x)Zr25Nb25Ta(25-x) (x ≤ 20) refractory high entropy alloys: Influences of substitution of Ti for Ta. Phys. B 2021, 606, 412851. [Google Scholar] [CrossRef]
  16. Nguyen, V.T.; Qian, M.; Shi, Z.; Tran, X.Q.; Fabijanic, D.M.; Joseph, J.; Qu, D.D.; Matsumura, S.; Zhang, C.; Zhang, F.; et al. Cuboid-like nanostructure strengthened equiatomic Ti-Zr-Nb-Ta medium entropy alloy. Mater. Sci. Eng. A 2020, 798, 140169. [Google Scholar] [CrossRef]
  17. Nguyen, V.T.; Qian, M.; Shi, Z.; Song, T.; Huang, L.; Zou, J. A novel quaternary equiatomic Ti-Zr-Nb-Ta medium entropy alloy (MEA). Intermetallics 2018, 101, 39–43. [Google Scholar] [CrossRef]
  18. Han, C.; Fang, Q.; Shi, Y.; Tor, S.B.; Chua, C.K.; Zhou, K. Recent advances on high-entropy alloys for 3D printing. Adv. Mater. 2020, 32, 1903855. [Google Scholar] [CrossRef]
  19. Liu, C.; Zhu, K.Y.; Ding, W.W.; Liu, Y.; Chen, G.; Qu, X.H. Additive Manufacturing of Wmotati Refractory High-Entropy Alloy by Employing Fluidised Powders. Powder Metall. 2022, 65, 413–425. [Google Scholar] [CrossRef]
  20. Zhang, H.; Zhao, Y.; Cai, J.; Ji, S.; Geng, J.; Sun, X.; Li, D. High-strength NbMoTaX refractory high-entropy alloy with low stacking fault energy eutectic phase via laser additive manufacturing. Mater. Des. 2021, 201, 109462. [Google Scholar] [CrossRef]
  21. Ron, T.; Leon, A.; Popov, V.; Strokin, E.; Eliezer, D.; Shirizly, A.; Aghion, E. Synthesis of Refractory High-Entropy Alloy WTaMoNbV by Powder Bed Fusion Process Using Mixed Elemental Alloying Powder. Materials 2022, 15, 4043. [Google Scholar] [CrossRef] [PubMed]
  22. Xia, M.; Chen, Y.; Chen, K.; Tong, Y.; Liang, X.; Shen, B. Synthesis of WTaMoNbZr refractory high-entropy alloy powder by plasma spheroidization process for additive manufacturing. J. Alloys Compd. 2022, 917, 165501. [Google Scholar] [CrossRef]
  23. Na, T.W.; Park, K.B.; Lee, S.Y.; Yang, S.M.; Kang, J.W.; Lee, T.W.; Park, J.M.; Park, K.; Park, H.K. Preparation of spherical TaNbHfZrTi high-entropy alloy powders by a hydrogenation-dehydrogenation reaction and thermal plasma treatment. J. Alloys Compd. 2020, 817, 152757. [Google Scholar] [CrossRef]
  24. Wei, M.; Chen, S.; Liang, J.; Liu, C. Effect of atomization pressure on the breakup of TA15 titanium alloy powder prepared by EIGA method for laser 3D printing. Vacuum 2017, 143, 185–194. [Google Scholar] [CrossRef]
  25. Tang, J.; Nie, Y.; Lei, Q.; Li, Y. Characteristics and atomization behavior of Ti-6Al-4V powder produced by plasma rotating electrode process. Adv. Powder Technol. 2019, 30, 2330–2337. [Google Scholar] [CrossRef]
  26. Ma, X.Z.; Shen, J.; Qi, X.M.; Jia, J. Cooling Rates of Prep Ti48al Alloy. J. Mater. Sci. Technol. 2001, 17, 91–92. [Google Scholar]
  27. Cui, Y.; Zhao, Y.F.; Numata, H.; Yamanaka, K.; Bian, H.K.; Aoyagi, K.; Chiba, A. Effects of Process Parameters and Cooling Gas on Powder Formation During the Plasma Rotating Electrode Process. Powder Technol. 2021, 393, 301–311. [Google Scholar] [CrossRef]
  28. Ning, J.; Zhang, H.B.; Chen, S.M.; Zhang, L.J.; Na, S.J. Intensive Laser Repair through Additive Manufacturing of High-Strength Martensitic Stainless Steel Powders (I)-Powder Preparation, Laser Cladding and Microstructures and Properties of Laser-Cladded Metals. J. Mater. Res. Technol. 2021, 15, 5746–5761. [Google Scholar] [CrossRef]
  29. Xie, B.; Fan, Y.; Zhao, S. Characterization of Ti6Al4V powders produced by different methods for selective laser melting. Mater. Res. Express 2021, 8, 076510. [Google Scholar] [CrossRef]
  30. Park, K.B.; Park, J.Y.; Do Kim, Y.; Na, T.W.; Mo, C.B.; Choi, J.I.; Choi, J.; Kang, H.S.; Park, H.K. Spark plasma sintering behavior of TaNbHfZrTi high-entropy alloy powder synthesized by hydrogenation-dehydrogenation reaction. Intermetallics 2021, 130, 107077. [Google Scholar] [CrossRef]
  31. He, W.; Liu, Y.; Tang, H.; Li, Y.; Liu, B.; Liang, X.; Xi, Z. Microstructural characteristics and densification behavior of high-Nb TiAl powder produced by plasma rotating electrode process. Mater. Des. 2017, 132, 275–282. [Google Scholar] [CrossRef]
  32. Vedovato, G.; Zambon, A.; Ramous, E. A simplified model for gas atomization. Mater. Sci. Eng. A 2001, 304, 235–239. [Google Scholar] [CrossRef]
  33. Sang, L.; Xu, Y.; Fang, P.; Zhang, H.; Cai, Y.; Liu, X. The influence of cooling rate on the microstructure and phase fraction of gas atomized NiAl3 alloy powders during rapid solidification. Vacuum 2018, 157, 354–360. [Google Scholar] [CrossRef]
  34. Heringer, R.; Gandin, C.A.; Lesoult, G.; Henein, H. Atomized droplet solidification as an equiaxed growth model. Acta Mater. 2006, 54, 4427–4440. [Google Scholar] [CrossRef]
  35. Yi, X.; Meng, X.; Cai, W.; Zhao, L. Multi-Stage Martensitic Transformation Behaviors and Microstructural Characteristics of Ti-Ni-Hf High Temperature Shape Memory Alloy Powders. J. Alloys Compd. 2019, 781, 644–656. [Google Scholar] [CrossRef]
  36. Zhang, M.; Zhang, Z.M. Numerical Simulation Study on Cooling of Metal Droplet in Atomizing Gas. Mater. Today Commun. 2020, 25, 101423. [Google Scholar] [CrossRef]
Figure 1. Schematic diagrams of the (a) EIGA and (b) PREP systems.
Figure 1. Schematic diagrams of the (a) EIGA and (b) PREP systems.
Materials 16 00791 g001
Figure 2. (a) TaNbTiZr RHEA bars; (b) XRD pattern, (c) EBSD IPF map, and (d) grain size distribution.
Figure 2. (a) TaNbTiZr RHEA bars; (b) XRD pattern, (c) EBSD IPF map, and (d) grain size distribution.
Materials 16 00791 g002
Figure 3. (a) SEM image of the TaNbTiZr RHEA bars; EDS mappings of (b) Ta, (c) Nb, (d) Ti, and (e) Zr, respectively.
Figure 3. (a) SEM image of the TaNbTiZr RHEA bars; EDS mappings of (b) Ta, (c) Nb, (d) Ti, and (e) Zr, respectively.
Materials 16 00791 g003
Figure 4. SEM images of the spherical TaNbTiZr RHEA powders at different magnifications: (a,a1,a2) the EIGAed powders and (b,b1,b2) the PREPed powders.
Figure 4. SEM images of the spherical TaNbTiZr RHEA powders at different magnifications: (a,a1,a2) the EIGAed powders and (b,b1,b2) the PREPed powders.
Materials 16 00791 g004
Figure 5. Particle size distributions of the EIGAed and PREPed TaNbTiZr RHEA powders.
Figure 5. Particle size distributions of the EIGAed and PREPed TaNbTiZr RHEA powders.
Materials 16 00791 g005
Figure 6. XRD patterns of the EIGAed and PREPed TaNbTiZr RHEA powders.
Figure 6. XRD patterns of the EIGAed and PREPed TaNbTiZr RHEA powders.
Materials 16 00791 g006
Figure 7. Cross-section microstructure of the (a,b) EIGAed and (c,d) PREPed TaNbTiZr RHEA powders; (b1) and (d1) are the elemental line scanning of the marked regions in (b) and (d), respectively.
Figure 7. Cross-section microstructure of the (a,b) EIGAed and (c,d) PREPed TaNbTiZr RHEA powders; (b1) and (d1) are the elemental line scanning of the marked regions in (b) and (d), respectively.
Materials 16 00791 g007
Figure 8. SEM images of the (a,b) EIGAed and (c,d) PREPed TaNbTiZr RHEA powders; (b1b4) and (d1d4) are the elemental mapping of the marked regions in (a) and (c), respectively.
Figure 8. SEM images of the (a,b) EIGAed and (c,d) PREPed TaNbTiZr RHEA powders; (b1b4) and (d1d4) are the elemental mapping of the marked regions in (a) and (c), respectively.
Materials 16 00791 g008
Table 1. Chemical composition of the EIGAed and PREPed TaNbTiZr RHEA powders.
Table 1. Chemical composition of the EIGAed and PREPed TaNbTiZr RHEA powders.
PowderElementTa (at%)Nb (at%)Ti (at%)Zr (at%)O (ppm)N (ppm)C (ppm)H (ppm)
Nominal25.0025.0025.0025.00
EIGAedActual25.1224.1425.4825.2684512415012
PREPedActual25.1924.4924.5525.77777175317
Table 2. Comparison of the oxygen content of the spherical TaNbTiZr RHEA powders with other similar RHEA powders prepared by different methods.
Table 2. Comparison of the oxygen content of the spherical TaNbTiZr RHEA powders with other similar RHEA powders prepared by different methods.
AlloysPreparation MethodsO (ppm)Ref
WTaMoNbZrMelting + Hydrogenation + Crushing + Spheroidization1678[22]
TaNbHfZrTiMelting + Hydrogenation + Dehydrogenation1770[30]
TaNbHfZrTiMelting + Hydrogenation + Dehydrogenation + Spheroidization1650[23]
TaNbTiZrCCLM + EIGA845This work
TaNbTiZrCCLM + PREP777This work
Table 3. The thermophysical parameters of the TaNbTiZr RHEA and argon, and the corresponding preparation parameters [15,17,31].
Table 3. The thermophysical parameters of the TaNbTiZr RHEA and argon, and the corresponding preparation parameters [15,17,31].
ParametersSymbolValue
Specific heat capacity (TaNbTiZr)Cd0.23 J g−1 K−1
Droplet density (TaNbTiZr)ρd8.91 g cm−3
Droplet temperatureTd2547.15 K
Argon temperatureTg298.15 K
Argon thermal conductivitykg3.55 × 10−2 W m−1 K−1
Argon densityρg9.7 × 10−4 g cm−3
Argon specific heat capacity per unit massCpg5.21 × 10−1 J g−1 K−1
Argon viscosityyg4.62 × 10−4 g cm−1 s−1
Electrode diameterd50 mm
Electrode rotation speedr20,000 rpm
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

Gao, S.; Fu, A.; Xie, Z.; Liao, T.; Cao, Y.; Liu, B. Preparation and Microstructure of High-Activity Spherical TaNbTiZr Refractory High-Entropy Alloy Powders. Materials 2023, 16, 791. https://doi.org/10.3390/ma16020791

AMA Style

Gao S, Fu A, Xie Z, Liao T, Cao Y, Liu B. Preparation and Microstructure of High-Activity Spherical TaNbTiZr Refractory High-Entropy Alloy Powders. Materials. 2023; 16(2):791. https://doi.org/10.3390/ma16020791

Chicago/Turabian Style

Gao, Shenghan, Ao Fu, Zhonghao Xie, Tao Liao, Yuankui Cao, and Bin Liu. 2023. "Preparation and Microstructure of High-Activity Spherical TaNbTiZr Refractory High-Entropy Alloy Powders" Materials 16, no. 2: 791. https://doi.org/10.3390/ma16020791

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