Next Article in Journal
The Effect of Solidification Rate on the Microstructure and Mechanical Properties of Pure Magnesium
Previous Article in Journal
Structure and Multistage Martensite Transformation in Nanocrystalline Ti-50.9Ni Alloy
Previous Article in Special Issue
Cooling under Applied Stress Rejuvenates Amorphous Alloys and Enhances Their Ductility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

High-Pressure Torsion for Synthesis of High-Entropy Alloys

1
WPI International Institute for Carbon-Neutral Energy Research (WPI-I2CNER), Kyushu University, Fukuoka 819-0395, Japan
2
Hefei General Machinery Research Institute, Hefei 230031, China
3
Scientific Center in Chernogolovka, Russian Academy of Sciences, 142432 Chernogolovka, Russia
4
School of Applied Sciences, University of Campinas (FCA-UNICAMP), Limeira 13484-350, Brazil
5
Institute for Materials Physics, University of Göttingen, 37077 Göttingen, Germany
*
Author to whom correspondence should be addressed.
Metals 2021, 11(8), 1263; https://doi.org/10.3390/met11081263
Submission received: 12 July 2021 / Revised: 9 August 2021 / Accepted: 10 August 2021 / Published: 11 August 2021
(This article belongs to the Special Issue Mechanical and Mechanochemical Synthesis of Alloys)

Abstract

:
High-pressure torsion (HPT) is widely used not only as a severe plastic deformation (SPD) method to produce ultrafine-grained metals but also as a mechanical alloying technique to synthesize different alloys. In recent years, there have been several attempts to synthesize functional high-entropy alloys using the HPT method. In this paper, the application of HPT to synthesize high-entropy materials including metallic alloys, hydrides, oxides and oxynitrides for enhanced mechanical and hydrogen storage properties, photocatalytic hydrogen production and high light absorbance is reviewed.

Graphical Abstract

1. Introduction

High-pressure torsion (HPT) is a severe plastic deformation (SPD) method [1], which is widely used to generate nanograins or ultrafine grains (UFG) in metallic materials [2]. In this method, as schematically shown in Figure 1, a small disc sample is compressed between two anvils under a pressure of several giga-Pascals and, concurrently, shear strain is introduced in the sample by rotating one of the anvils with respect to the other one (γ = 2πrN/h, γ: shear strain, r: distance from disc center, N: anvil turns, h: sample thickness) [1,2]. The method was originally introduced by Bridgman in 1935 [3] to examine polymorphic phase transformations under high pressure (see the historical development of the HPT method in [4,5]). In 1988, Valiev et al. [6] introduced the HPT method as an SPD technique to achieve nanograined microstructures in metallic materials. This study resulted in the development of the Nano-SPD field [7] and accelerated the research activities on the HPT method by many other research groups [8,9,10,11].
Within the past three decades, the technology of the HPT method has not changed significantly; however, the method has been used for a wider range of materials and applications. Besides the popular application of the method to produce UFG materials [8,9,10,11], it has also been used to control phase transformations [12,13,14], consolidate powders [15,16,17] and conduct mechanical alloying [18,19,20]. Moreover, the application of HPT is no longer limited to metallic materials—the method has been applied to nonmetallic materials such as carbon polymorphs [21,22], silicon polymorphs [21,23], oxides [24,25], oxynitrides [26] and glasses [27,28] as well. Compared to other SPD methods such as equal-channel angular pressing (ECAP) [29], accumulative roll-bonding (ARB) [30] and twist extrusion (TE) [31], HPT has the unique capability of introducing extremely large strains in almost any kind of materials, including hard and brittle ones, due to its high processing pressure [1,2,3,4].
It has been shown that the HPT method can introduce extremely large shear strains of over 1000 (up to 100,0000), so that the microscopic phases can be geometrically thinned to the subnanometer range [20,32]. The shear strain of 1000 is the nominal strain that is geometrically required to reduce the size of a phase with 1 μm to less than 1 nm; however, the required shear strain for mechanical alloying can be lower or much higher than this level depending on the size of phases, their miscibility, their co-deformation, and their recrystallization [32]. In general, the strains required for mechanical alloying are a few orders of magnitude higher than those required for only grain refinement by HPT processing [2,4]. The introduction of such large strains, which has received the terminology of ultra-SPD [20,32], is quite effective for the synthesis of new materials even from immiscible systems such as Mg-Ti and Mg-Zr with hydrogen storage capability [33,34] or Al-Zr and Al-La-Ce with high strength [35,36]. Although high pressure is a key parameter of HPT to deform any kind of materials, shear strain plays the most important role in mechanical alloying. Mechanical alloying by HPT is usually conducted at ambient temperature with a low deformation speed (usually 1 rpm), and the temperature rise during the process is not significant [37]. All these HPT processing conditions provide an opportunity for the simultaneous synthesis and cold consolidation of bulk nanostructured alloys from their elemental powders [20,32]. Compared to ball milling as a popular mechanical alloying technique [38], in the HPT method, the synthesis time is shorter, the rotation speed is lower, the contamination is minor, and the final product is in the bulk form, which does not require further consolidation processes [2,4]. However, the main disadvantage of HPT is the low quantity of the synthesized product; thus, scaling up the sample [39,40] and making the process continuous [41,42] are currently major challenges for the commercialization of HPT.
High-entropy alloys (HEAs), containing a mixture of at least five principal elements (usually 5–35 at% for each element) [43,44], are one of the latest groups of materials that have been processed by the SPD methods [45,46]. Although HEAs have generally high strength [47,48,49,50], it was shown that their strength can be further enhanced by SPD processing. For example, the HPT method was successfully used for the hardening of several HEAs, such as CoCrFeMnNi [51,52], Al0.3CoCrFeNi [53,54], Al0.5CoCrFeMnNi [55], AlCrFeCoNiNb [56], TiZrNbHfTa [57], CrFe2NiMnV0.25 [58], CoCrFeNi [59], TiZrCrMnFeNi [60], TiAlFeCoNi [61], TiAlFeCoNi-C [62] and CoCrFeMnNi-C [63]. More recently, the HPT method was used to impart ultra-SPD and synthesize CoCrFeMnNi with high hardness [64]. Since the concept of HEAs has been extended to non-metallic materials such as oxides [65,66], nitrides [67,68], carbides [69,70], borides [71,72] and ceramics [73,74], there have been some attempts to use the HPT method to synthesize non-metallic HEAs such as high-entropy hydride MgTiVCrFe-H for hydrogen storage [75], high-entropy oxide TiZrHfNbTaO11 for photocatalytic hydrogen production [76] and high-entropy oxynitride TiZrHfNbTaO6N3 as a photocatalyst with large light absorbance [77].
In this article, the main achievements in the synthesis of HEAs by ultra-SPD via the HPT method are reviewed and the significance of the process for attaining advanced functionalities is discussed.

2. Synthesis of High-Entropy Alloys with High Hardness

Nanostructured materials show not only enhanced functional properties but also high hardness due to the Hall–Petch strengthening mechanism [1,7]. In an attempt to produce nanostructured HEAs, Kilmametov et al. [64] synthesized a CoCrFeMnNi alloy with a face-centered cubic (FCC) structure starting from the elemental powders, applying HPT with 100 turns. They selected this alloy (also known as the Cantor alloy [38]) because it is one of the first designated HEAs and there have been numerous studies to synthesize and process the alloy by different methods. They used atom probe tomography analysis and reported a successful atomic-scale mixture of elements, as shown in Figure 2. They also showed that, besides mechanical alloying, good consolidation to a bulk form and nanograin formation occurred. These microstructural features resulted in a high hardness level of 700 HV in the alloy. They also reported that such high hardness is not only due to microstructural refinement but also due to the formation of small amounts of chromium oxide particles, also shown in Figure 2 [64].
In another attempt, González-Masís et al. [78] synthesized TiNb, TiNbZr, TiNbZrTa and TiNbZrTaHf alloys from the elemental powders using the HPT process. These alloys were selected because of their biocompatibility and their tendency to form single body-centered cubic (BCC) phases. All materials showed the formation of fully BCC structures with the presence of nanograins. In addition to the successful synthesis of materials by the HPT method, they reported that there are increases in hardness when increasing the configurational entropy or the number of elements, as shown in Figure 3. Detailed analyses confirmed that these hardness changes cannot be justified only by conventional strengthening mechanisms such as solution hardening, dislocation hardening, grain-boundary hardening and precipitation hardening, because the ternary, quaternary and quinary alloys in Figure 3 had similar nanograin sizes, similar dislocation densities and similar atomic-size-mismatch parameters [78]. Since the steady-state hardness of HPT-processed metals is their intrinsic property and depends on the atomic bonding energy [79], and since this energy can be reduced by increasing the entropy [43,44], it was hypothesized by González-Masís et al. that high configurational entropy and a decrease in the internal energy of the system can also influence the steady-state hardness of HEAs [78].

3. Synthesis of High-Entropy Hydrides for Hydrogen Storage

The safe and high-density storage of hydrogen as a clean fuel is still a challenging issue for the future development of a hydrogen-based society [80]. Hydrides provide the most compact technology for the storage of hydrogen, although there are still drawbacks regarding the kinetics and thermodynamics of hydrogenation and dehydrogenation in hydrides [81,82]. The synthesis of hydrides is a critical issue because the performance of hydrides can be significantly influenced by the processing route [81,82]. Mechanochemistry routes such as HPT are considered effective to produce hydrides with promising hydrogen storage properties [83,84]. It has been shown that the HPT method is not only effective to synthesize hydrogen storage materials but also to enhance their activity, kinetics, and air resistance [85]. The enhancement of the activity and kinetics of hydrogen storage materials by HPT processing is mainly attributed to the formation of lattice defects and grain boundaries, which can act as quick pathways for hydrogen transport from surface to bulk [84,85].
In one of the first attempts to synthesize high-entropy hydrogen storage materials, de Marco et al. [75] synthesized MgVCr-H and MgVTiCrFe-H from the elemental powders using ball milling, followed by the HPT method and hydrogenation. The alloys after HPT processing showed a mixture of BCC and amorphous phases, and energy dispersive X-ray spectroscopy confirmed good mechanical mixing of the elements. This first study on the synthesis of high-entropy hydrogen storage materials by the HPT method confirmed the potential of the method, although the reported hydrogen storage capacity was not high, as shown in Figure 4. Since recent studies suggested the potential of high-entropy hydrogen storage materials (HfNbTiVZr [86], TiZrNbHfTa [87], MgZrTiFe0.5Co0.5Ni0.5 [88], TiZrNbMoV [89], TiZrHfScMo [90], CoFeMnTiVZr [91], ZrTiVCrFeNi [92], (VFe)60(TiCrCo)40−xZrx [93], TiVZrNbTa [94], AlCrFeMnNiW [95]) and particularly those materials designated by inputs from the theoretical calculations (TiZrCrMnFeNi [59], TiZrNbFeNi [96], and TiZrNbCrFe [97]), a combination of theoretical design and mechanical synthesis by HPT processing should be an effective strategy to explore new hydrogen storage materials. Edalati et al. [59] and Floriano et al. [96] suggested three criteria to explore high-entropy materials for room-temperature hydrogen storage: (i) selection of an AB2 or AB system, where A represents the hydride-forming elements such as Mg, Ti, Zr, V, Nb, etc., and B represents elements with low chemical affinity with hydrogen, such as Cr, Mn, Fe, Co, Ni, etc.; (ii) valence electron concentration of 6.4–6.5; (iii) Laves phase stability, which should be examined by thermodynamic calculations using the CALPHAD (calculation of phase diagram) method. Although these suggested criteria were used successfully to design room-temperature hydrogen storage materials [59,96], there are some less successful attempts to design HEAs with higher A/B ratio with the Laves or BCC structures [75,97].

4. Synthesis of High-Entropy Oxides for Photocatalytic Hydrogen Production

The production of hydrogen without CO2 emission is one major challenge in realizing hydrogen as a clean fuel for the future [80]. The cleanest technology to produce hydrogen is photocatalytic water splitting under sunlight [98]. In photocatalysis, light absorption by a semiconductor photocatalyst leads to the excitation of electrons to the conduction band and the formation of holes in the valence band. These electrons and holes on the surface of the catalyst contribute to the oxidation and reduction of water to produce O2 and H2, respectively. Despite the significant scientific interest in photocatalysis, the efficiency of this technology for hydrogen production is still far from commercialization [99,100]. There are now significant efforts being made around the world to develop new photocatalysts that can produce large amounts of hydrogen from water. Large light absorption, long electron/hole recombination time and high chemical stability are some parameters that are important in designing new photocatalysts [99,100]. High-entropy photocatalysts are one of the newest family of photocatalysts developed recently for hydrogen production [76,77].
In one of the first attempts to produce high-entropy photocatalysts [76], powders of Ti, Zr, Hf, Nb and Ta were mixed by HPT processing with 200 turns to produce a BCC TiZrHfNbTa alloy. The Ti, Zr, Hf, Nb and Ta elements were selected because their cations generate a d0 electronic configuration and it is empirically known that oxides with the d0 electronic configuration can usually exhibit good photocatalytic properties for hydrogen production [99,100]. The BCC alloy was further oxidized at 1443 K for 24 h to produce a dual-phase high-entropy oxide with an orthorhombic and a monoclinic phase and a general composition of TiZrHfNbTaO11. The synthesized dual-phase oxide had an orange color (Figure 5a), which corresponds to the absorbance of blue light based on the Itten color ring (Figure 5b). It had large particle sizes (Figure 5c) compared to its grain sizes, which were at the nanometer level (Figure 5d). Examination of the photocatalytic performance of the material, as shown in Figure 6a, confirmed that the material could produce hydrogen from water splitting under UV light. Moreover, examination of the material by X-ray diffraction before and after three cycles of photocatalytic testing (Figure 6b) confirmed that the material is chemically stable. High stability is a general feature of many metallic [47,48,49,50] and non-metallic [65,66,67,68,69,70,71,72,73,74] high-entropy materials, which results from the high entropy of mixing [43,44]. Although the hydrogen production rate on TiZrHfNbTaO11 was low, similar to other available photocatalysts [99,100], the study introduced the HPT method as a potential pathway to produce new materials for energy-related applications such as photocatalysis [76]. It should be noted that the HPT method as a processing route can also enhance the photocatalytic hydrogen production through three main mechanisms [24,25,26]: (i) bandgap narrowing by the introduction of vacancies; (ii) the introduction of lattice defects as active sites for photocatalytic water splitting; (iii) the stabilization of high-pressure phases such as TiO2-II with higher photocatalytic activity.

5. Synthesis of High-Entropy Oxynitrides with Large Light Absorbance

Large light absorbance is a major requirement in some applications, such as photocatalysis [99,100]. Photocatalysts are mainly oxides, but most of the oxides can only absorb UV light (due to their large bandgap), and UV light accounts for only 6% of the sunlight on Earth [101]. Nitrides have a lower bandgap than oxides due to the higher energy level of the nitrogen 2p orbital as compared to the oxygen 2p orbital; however, nitrides suffer from poor stability [102]. Therefore, there are now significant attempts to develop oxynitrides that possess both large light absorbance and high stability.
To develop oxynitrides with low bandgap and high stability for photocatalytic application, Edalati et al. [77] synthesized a high-entropy oxynitride based on the mixture of Ti, Zr, Hf, Nb and Ta cations, which can generate a d0 electronic configuration that is appropriate for photocatalytic activity [101,102]. They developed a three-step synthesis technique: (i) mechanical alloying of elemental powders by HPT processing for 100 turns to produce a BCC TiZrHfNbTa alloy; (ii) oxidation of the BCC alloy at 1373 K for 24 h to produce TiZrHfNbTaO11 oxide; (iii) nitriding the oxide in ammonia at 1373 K for 7 h to produce an oxynitride. The final product was a dual-phase high-entropy oxynitride with monoclinic and FCC structures and a general composition of TiZrHfNbTaO6N3. Its grain sizes were small, in the range of nanometers, as shown in Figure 7. The material showed a narrow bandgap of 1.6 eV, which is one of smallest bandgaps reported for oxynitrides [101,102]. Moreover, its valence band and conduction band satisfied the energy requirements for photocatalytic water splitting, as shown in Figure 8a (conduction band bottom was more negative than 0 eV vs. NHE for H2O to H2 reduction and valence band top was more positive than 1.23 eV vs. NHE for H2O to O2 oxidation [98]). Furthermore, the light absorbance of this oxynitride was better than all binary oxides formed in the Ti-Zr-Hf-Nb-Ta system as well the corresponding high-entropy oxide, as shown in Figure 8. This high-entropy oxynitride with high chemical stability could produce hydrogen from photocatalytic water splitting, introducing high-entropy oxynitrides as new low-bandgap and stable photocatalysts [77].

6. Concluding Remarks

In summary, the HPT method can be considered not only a processing route for HEAs [51,52,53,54,55,56,57,58,59,60,61,62,63] but also a fast synthesis route to produce various kinds of high-entropy materials, including metallic alloys with high hardness [63], hydrides for hydrogen storage [75], oxides for photocatalytic hydrogen production [76] and oxynitrides with large light absorbance [77]. Compared to conventional synthesis routes for metallic [43,44,47,48,49,50] and non-metallic [65,66,67,68,69,70,71,72,73,74] HEAs, the HPT method provides a fast laboratory-scale route that is applicable to almost any kind of materials [21,22,23,24,25,26,27,28]. One requirement to synthesize the materials by HPT is imparting ultra-SPD with large shear strains, which can be realized by increasing the number of HPT turns [20,32,33,34,35,36]. The materials that are synthesized by the HPT method usually have small crystal size [18,19,20], which is beneficial for some properties such as mechanical strength and hydrogen storage properties [1,2,3,4,5,6,7,8,9,10,11]. The phenomenon of cold consolidation by HPT [15,16,17] results in the small surface area of the final product, which is not beneficial for some applications such as photocatalysis. Although research on the HPT method as a synthesis route for HEAs is in its early stages, the method has high potential for the exploration of new materials for future applications. The next stage in this research field is the application of HPT to a wider range of high-entropy materials with new functional applications, although scaling up the sample size remains a major challenge for the commercialization of this method [39,40,41,42].

Author Contributions

Writing—review and editing, K.E., H.-W.L., A.K., R.F. and C.B.; project administration, K.E. All authors have read and agreed to the published version of the manuscript.

Funding

The author K.E. acknowledges MEXT, Japan for the Grants-in-Aid for Scientific Research on Innovative Areas (No. 19H05176, 21H00150) and the author A.K. thanks Volkswagen Stiftung for a grant (No. 97751).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Valiev, R.Z.; Estrin, Y.; Horita, Z.; Langdon, T.G.; Zehetbauer, M.J.; Zhu, Y.T. Producing bulk ultrafine-grained materials by severe plastic deformation. JOM 2006, 58, 33–39. [Google Scholar] [CrossRef] [Green Version]
  2. Zhilyaev, A.P.; Langdon, T.G. Using high-pressure torsion for metal processing: Fundamentals and applications. Prog. Mater. Sci. 2008, 53, 893–979. [Google Scholar] [CrossRef]
  3. Bridgman, P.W. Effects of high shearing stress combined with high hydrostatic pressure. Phys. Rev. 1935, 48, 825–847. [Google Scholar] [CrossRef]
  4. Edalati, K.; Horita, Z. A review on high-pressure torsion (HPT) from 1935 to 1988. Mater. Sci. Eng. A 2016, 652, 325–352. [Google Scholar] [CrossRef]
  5. Bryła, K.; Edalati, K. Historical studies by polish scientist on ultrafine-grained materials by severe plastic deformation. Mater. Trans. 2019, 60, 1553–1560. [Google Scholar] [CrossRef] [Green Version]
  6. Valiev, R.Z.; Kaibyshev, O.A.; Kuznetsov, R.I.; Musalimov, R.S.; Tsenev, N.K. Low-temperature superplasticity of metallic materials. Dokl. Akad. Nauk. SSSR 1988, 301, 864–866. [Google Scholar]
  7. Valiev, R.Z.; Islamgaliev, R.K.; Alexandrov, I.V. Bulk nanostructured materials from severe plastic deformation. Prog. Mater. Sci. 2000, 45, 103–189. [Google Scholar] [CrossRef]
  8. Pippan, R.; Scheriau, S.; Taylor, A.; Hafok, M.; Hohenwarter, A.; Bachmaier, A. Saturation of fragmentation during severe plastic deformation. Annu. Rev. Mater. Res. 2010, 40, 319–343. [Google Scholar] [CrossRef]
  9. Figueiredo, R.B.; Kawasaki, M.; Langdon, T.G. An evaluation of homogeneity and heterogeneity in metals processed by high-pressure torsion. Acta Phys. Pol. A 2012, 122, 425–429. [Google Scholar] [CrossRef]
  10. Starink, M.J.; Cheng, X.C.; Yang, S. Hardening of pure metals by high-pressure torsion: A physically based model employing volume-averaged defect evolutions. Acta Mater. 2013, 61, 183–192. [Google Scholar] [CrossRef] [Green Version]
  11. Lee, D.J.; Yoon, E.Y.; Ahn, D.H.; Park, B.H.; Park, H.W.; Park, L.J.; Estrin, Y.; Kim, H.S. Dislocation density-based finite element analysis of large strain deformation behavior of copper under high-pressure torsion. Acta Mater. 2014, 76, 281–293. [Google Scholar] [CrossRef]
  12. Bachmaier, A.; Pippan, R. High-pressure torsion deformation induced phase transformations and formations: New material combinations and advanced properties. Mater. Trans. 2019, 60, 1256–1269. [Google Scholar] [CrossRef] [Green Version]
  13. Levitas, V.I. High-pressure phase transformations under severe plastic deformation by torsion in rotational anvils. Mater. Trans. 2019, 60, 1294–1301. [Google Scholar] [CrossRef] [Green Version]
  14. Mazilkin, A.; Straumal, B.; Kilmametov, A.; Straumal, P.; Baretzky, B. Phase transformations induced by severe plastic deformation. Mater. Trans. 2019, 60, 1489–1499. [Google Scholar] [CrossRef] [Green Version]
  15. Alexandrov, I.V.; Zhu, Y.T.; Lowe, T.C.; Islamgaliev, R.K.; Valiev, R.Z. Microstructures and properties of nanocomposites obtained through SPTS consolidation of powders. Metall. Mater. Trans. A 1998, 29, 2253–2260. [Google Scholar] [CrossRef]
  16. Lee, Z.; Zhou, F.; Valiev, R.Z.; Lavernia, E.J.; Nutt, S.R. Microstructure and microhardness of cryomilled bulk nanocrystalline Al-7.5%Mg alloy consolidated by high pressure torsion. Scr. Mater. 2004, 51, 209–214. [Google Scholar] [CrossRef]
  17. Borchers, C.; Garve, C.; Tiegel, M.; Deutges, M.; Herz, A.; Edalati, K.; Pippan, R.; Horita, Z.; Kirchheim, R. Nanocrystalline steel obtained by ball milling of iron and graphite subsequently compacted by high-pressure torsion. Acta Mater. 2015, 97, 207–215. [Google Scholar] [CrossRef]
  18. Ibrahim, N.; Peterlechner, M.; Emeis, F.; Wegner, M.; Divinski, S.V.; Wilde, G. Mechanical alloying via high-pressure torsion of the immiscible Cu50Ta50 system. Mater. Sci. Eng. A 2017, 685, 19–30. [Google Scholar] [CrossRef]
  19. Han, J.K.; Jang, J.I.; Langdon, T.G.; Kawasaki, M. Bulk-state reactions and improving the mechanical properties of metals through high-pressure torsion. Mater. Trans. 2019, 60, 1131–1138. [Google Scholar] [CrossRef] [Green Version]
  20. Edalati, K. Metallurgical alchemy by ultra-severe plastic deformation via high-pressure torsion process. Mater. Trans. 2019, 60, 1221–1229. [Google Scholar] [CrossRef] [Green Version]
  21. Blank, V.D.; Popov, M.Y.; Kulnitskiy, B.A. The effect of severe plastic deformations on phase transitions and structure of solids. Mater. Trans. 2019, 60, 1500–1505. [Google Scholar] [CrossRef] [Green Version]
  22. Gao, Y.; Ma, Y.; An, Q.; Levitas, V.; Zhang, Y.; Feng, B.; Chaudhuri, J.; Goddard, W.A., III. Shear driven formation of nano-diamonds at sub-gigapascals and 300 K. Carbon 2019, 146, 364–368. [Google Scholar] [CrossRef] [Green Version]
  23. Ikoma, Y.; Hayano, K.; Edalati, K.; Saito, K.; Guo, Q.; Horita, Z. Phase transformation and nanograin refinement of silicon by processing through high-pressure torsion. Appl. Phys. Lett. 2012, 101, 121908. [Google Scholar] [CrossRef]
  24. Razavi-Khosroshahi, H.; Edalati, K.; Hirayama, M.; Emami, H.; Arita, M.; Yamauchi, M.; Hagiwara, H.; Ida, S.; Ishihara, T.; Akiba, E.; et al. Visible-light-driven photocatalytic hydrogen generation on nanosized TiO2-II stabilized by high-pressure torsion. ACS Catal. 2016, 6, 5103–5107. [Google Scholar] [CrossRef]
  25. Edalati, K. Review on recent advancement in severe plastic deformation of oxides by high-pressure torsion (HPT). Adv. Eng. Mater. 2019, 21, 1800272. [Google Scholar] [CrossRef]
  26. Edalati, K.; Uehiro, R.; Takechi, S.; Wang, Q.; Arita, M.; Watanabe, M.; Ishihara, T.; Horita, Z. Enhanced photocatalytic hydrogen production on GaN-ZnO oxynitride by introduction of strain-induced nitrogen vacancy complexes. Acta Mater. 2020, 185, 149–156. [Google Scholar] [CrossRef]
  27. Révész, Á.; Kovács, Z. Severe plastic deformation of amorphous alloys. Mater. Trans. 2019, 60, 1283–1293. [Google Scholar] [CrossRef] [Green Version]
  28. Edalati, K.; Fujita, I.; Sauvage, X.; Arita, M.; Horita, Z. Microstructure and phase transformations of silica glass and vanadium oxide by severe plastic deformation via high-pressure torsion straining. J. Alloys Compd. 2019, 779, 394–398. [Google Scholar] [CrossRef]
  29. Segal, V.M.; Reznikov, V.I.; Drobyshevskiy, A.E.; Kopylov, V.I. Plastic working of metals by simple shear. Russ. Metall. 1981, 1, 99–105. [Google Scholar]
  30. Saito, Y.; Utsunomiya, H.; Tsuji, N.; Sakai, T. Novel ultra-high straining process for bulk materials: Development of the accumulative roll-bonding (ARB) process. Acta Mater. 1999, 47, 579–583. [Google Scholar] [CrossRef]
  31. Beygelzimer, Y.; Varukhin, V.; Synkov, S.; Sapronov, A.; Synkov, V. New techniques for accumulating large plastic deformations using hydroextrusion. Fiz. i Tekhnika Vusokih Davlenii 1999, 9, 109–111. [Google Scholar]
  32. Edalati, K.; Uehiro, R.; Fujiwara, K.; Ikeda, Y.; Li, H.W.; Sauvage, X.; Valiev, R.Z.; Akiba, E.; Tanaka, I.; Horita, Z. Ultra-severe plastic deformation: Evolution of microstructure, phase transformation and hardness in immiscible magnesium-based systems. Mater. Sci. Eng. A 2017, 701, 158–166. [Google Scholar] [CrossRef]
  33. Edalati, K.; Emami, H.; Staykov, A.; Smith, D.J.; Akiba, E.; Horita, Z. Formation of metastable phases in magnesium-titanium system by high-pressure torsion and their hydrogen storage performance. Acta Mater. 2015, 99, 150–156. [Google Scholar] [CrossRef]
  34. Edalati, K.; Emami, H.; Ikeda, Y.; Iwaoka, H.; Tanaka, I.; Akiba, E.; Horita, Z. New nanostructured phases with reversible hydrogen storage capability in immiscible magnesium-zirconium system produced by high-pressure torsion. Acta Mater. 2016, 108, 293–303. [Google Scholar] [CrossRef]
  35. Mohammadi, A.; Enikeev, N.A.; Murashkin, M.Y.; Arita, M.; Edalati, K. Examination of inverse Hall-Petch relation in nanostructured aluminum alloys by ultra-severe plastic deformation. J. Mater. Sci. Technol. 2021, 91, 78–89. [Google Scholar] [CrossRef]
  36. Mohammadi, A.; Enikeev, N.A.; Murashkin, M.Y.; Arita, M.; Edalati, K. Developing age-hardenable Al-Zr alloy by ultra-severe plastic deformation: Significance of supersaturation, segregation and precipitation on hardening and electrical conductivity. Acta Mater. 2021, 203, 116503. [Google Scholar] [CrossRef]
  37. Edalati, K.; Hashiguchi, Y.; Pereira, P.H.R.; Horita, A.; Langdon, T.G. Effect of temperature rise on microstructural evolution during high-pressure torsion. Mater. Sci. Eng. A 2018, 714, 167–171. [Google Scholar] [CrossRef] [Green Version]
  38. Suryanarayana, C. Mechanical alloying and milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  39. Edalati, K.; Horita, Z. Scaling-up of high pressure torsion using ring shape. Mater. Trans. 2009, 50, 92–95. [Google Scholar] [CrossRef] [Green Version]
  40. Horia, Z.; Tang, Y.; Masuda, T.; Takizawa, Y. Severe plastic deformation under high pressure: Upsizing sample dimensions. Mater. Trans. 2020, 61, 1177–1190. [Google Scholar] [CrossRef]
  41. Edalati, K.; Lee, S.; Horita, Z. Continuous high-pressure torsion using wires. J. Mater. Sci. 2012, 47, 473–478. [Google Scholar] [CrossRef]
  42. Ivanisenko, Y.; Kulagin, R.; Fedorov, V.; Mazilkin, A.; Scherer, T.; Baretzky, B.; Hahn, H. High pressure torsion extrusion as a new severe plastic deformation process. Mater. Sci. Eng. A 2016, 664, 247–256. [Google Scholar] [CrossRef]
  43. Yeh, J.W.; Chen, S.K.; Lin, S.J.; Gan, J.Y.; Chin, T.S.; Shun, T.T.; Tsau, C.H.; Chang, S.Y. Nanostructured high-entropy alloys with multiple principal elements: Novel alloy design concepts and outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  44. Cantor, B.; Chang, I.; Knight, P.; Vincent, A. Microstructural development in equiatomic multicomponent alloys. Mater. Sci. Eng. A 2004, 375, 213–218. [Google Scholar] [CrossRef]
  45. Hammond, V.H.; Atwater, M.A.; Darling, K.A.; Nguyen, H.Q.; Kecskes, L.J. Equal-channel angular extrusion of a low-density high-entropy alloy produced by high-energy cryogenic mechanical alloying. JOM 2014, 66, 2021–2029. [Google Scholar] [CrossRef]
  46. Shahmir, H.; Mousavi, T.; He, J.; Lu, Z.; Kawasaki, M.; Langdon, T.G. Microstructure and properties of a CoCrFeNiMn high-entropy alloy processed by equal-channel angular pressing. Mater. Sci. Eng. A 2017, 705, 411–419. [Google Scholar] [CrossRef] [Green Version]
  47. Zhang, Y.; Zuo, T.T.; Tang, Z.; Gao, M.C.; Dahmen, K.A.; Liaw, P.K.; Lu, Z.P. Microstructures and properties of high-entropy alloys. Prog. Mater. Sci. 2014, 61, 1–93. [Google Scholar] [CrossRef]
  48. Miracle, D.B.; Senkov, O.N. A critical review of high entropy alloys and related concepts. Acta Mater. 2017, 122, 448–511. [Google Scholar] [CrossRef] [Green Version]
  49. Li, Z.; Zhao, S.; Ritchie, R.O.; Meyers, M.A. Mechanical properties of high-entropy alloys with emphasis on face-centered cubic alloys. Prog. Mater. Sci. 2019, 102, 296–345. [Google Scholar] [CrossRef]
  50. George, E.P.; Curtin, W.A.; Tasan, C.C. High entropy alloys: A focused review of mechanical properties and deformation mechanisms. Acta Mater. 2020, 188, 435–474. [Google Scholar] [CrossRef]
  51. Schuh, B.; Mendez-Martin, F.; Völker, B.; George, E.P.; Clemens, H.; Pippan, R.; Hohenwarter, A. Mechanical properties, microstructure and thermal stability of a nanocrystalline CoCrFeMnNi high-entropy alloy after severe plastic deformation. Acta Mater. 2015, 96, 258–268. [Google Scholar] [CrossRef] [Green Version]
  52. Shahmir, H.; He, J.; Lu, Z.; Kawasaki, M.; Langdon, T.G. Effect of annealing on mechanical properties of a nanocrystalline CoCrFeNiMn high-entropy alloy processed by high-pressure torsion. Mater. Sci. Eng. A 2016, 676, 294–303. [Google Scholar] [CrossRef] [Green Version]
  53. Tang, Q.H.; Huang, Y.; Huang, Y.Y.; Liao, X.Z.; Langdon, T.G.; Dai, P.Q. Hardening of an Al0.3CoCrFeNi high entropy alloy via high-pressure torsion and thermal annealing. Mater. Lett. 2015, 151, 126–129. [Google Scholar] [CrossRef] [Green Version]
  54. Tang, Q.H.; Huang, Y.; Cheng, H.; Liao, X.Z.; Langdon, T.G.; Dai, P.Q. The effect of grain size on the annealing-induced phase transformation in an Al0.3CoCrFeNi high entropy alloy. Mater. Des. 2016, 105, 381–385. [Google Scholar] [CrossRef] [Green Version]
  55. Reddy, T.S.; Wani, I.S.; Bhattacharjee, T.; Reddy, S.R.; Saha, R.; Bhattacharjee, P.P. Severe plastic deformation driven nanostructure and phase evolution in a Al0.5CoCrFeMnNi dual phase high entropy alloy. Intermetallics 2017, 91, 150–157. [Google Scholar] [CrossRef]
  56. Edalati, P.; Mohammadi, A.; Ketabchi, M.; Edalati, K. Ultrahigh hardness in nanostructured dual-phase high-entropy alloy AlCrFeCoNiNb developed by high-pressure torsion. J. Alloys Compd. 2021, 884, 161101. [Google Scholar] [CrossRef]
  57. Schuh, B.; Völker, B.; Todt, J.; Schell, N.; Perrière, L.; Li, J.; Couzinié, J.P.; Hohenwarter, A. Thermodynamic instability of a nanocrystalline, single-phase TiZrNbHfTa alloy and its impact on the mechanical properties. Acta Mater. 2018, 142, 201–212. [Google Scholar] [CrossRef]
  58. Shahmir, H.; Tabachnikova, E.; Podolskiy, A.; Tikhonovsky, M.; Langdon, T.G. Effect of carbon content and annealing on structure and hardness of CrFe2NiMnV0.25 high-entropy alloys processed by high-pressure torsion. J. Mater. Sci. 2018, 53, 11813–11822. [Google Scholar] [CrossRef] [Green Version]
  59. Gubicza, J.; Hung, P.T.; Kawasaki, M.; Han, J.K.; Zhao, Y.; Xue, Y.; Lábár, J.L. Influence of severe plastic deformation on the microstructure and hardness of a CoCrFeNi high-entropy alloy: A comparison with CoCrFeNiMn. Mater. Charact. 2019, 154, 304–314. [Google Scholar] [CrossRef]
  60. Edalati, P.; Floriano, R.; Mohammadi, A.; Li, Y.; Zepon, G.; Li, H.W.; Edalati, K. Reversible room temperature hydrogen storage in high-entropy alloy TiZrCrMnFeNi. Scr. Mater. 2020, 178, 387–390. [Google Scholar] [CrossRef]
  61. Edalati, P.; Floriano, R.; Tang, Y.; Mohammadi, A.; Danielle Pereira, K.; Ducati Luchessi, A.; Edalati, K. Ultrahigh hardness and biocompatibility of high-entropy alloy TiAlFeCoNi processed by high-pressure torsion. Mater. Sci. Eng. C 2020, 112, 110908. [Google Scholar] [CrossRef]
  62. Edalati, P.; Mohammadi, A.; Tang, Y.; Floriano, R.; Fuji, M.; Edalati, K. Phase transformation and microstructure evolution in ultrahard carbon-doped AlTiFeCoNi high-entropy alloy by high-pressure torsion. Mater. Lett. 2021, 302, 130368. [Google Scholar] [CrossRef]
  63. Lu, Y.; Mazilkin, A.; Boll, T.; Stepanov, N.; Zherebtzov, S.; Salishchev, G.; Ódor, É.; Ungar, T.; Lavernia, E.; Hahn, H.; et al. Influence of carbon on the mechanical behavior and microstructure evolution of CoCrFeMnNi processed by high pressure torsion. Materialia 2021, 16, 101059. [Google Scholar] [CrossRef]
  64. Kilmametov, A.; Kulagin, R.; Mazilkin, A.; Seils, S.; Boll, T.; Heilmaier, M.; Hahn, H. High-pressure torsion driven mechanical alloying of CoCrFeMnNi high entropy alloy. Scr. Mater. 2019, 158, 29–33. [Google Scholar] [CrossRef]
  65. Rost, C.M.; Sachet, E.; Borman, T.; Moballegh, A.; Dickey, E.C.; Hou, D.; Jones, J.L.; Curtarolo, S.; Maria, J.P. Entropy-stabilized oxides. Nat. Commun. 2015, 6, 8485. [Google Scholar] [CrossRef] [Green Version]
  66. Sarkar, A.; Djenadic, R.; Wang, D.; Hein, C.; Kautenburger, R.; Clemens, O.; Hahn, H. Rare earth and transition metal based entropy stabilised perovskite type oxides. J. Eur. Ceram. Soc. 2018, 38, 2318–2327. [Google Scholar] [CrossRef]
  67. Chen, T.; Shun, T.; Yeh, J.; Wong, M. Nanostructured nitride films of multi-element high-entropy alloys by reactive DC sputtering. Surf. Coat. Technol. 2004, 188, 193–200. [Google Scholar] [CrossRef]
  68. Lewin, E. Multi-component and high-entropy nitride coatings–a promising field in need of a novel approach. J. Appl. Phys. 2020, 127, 160901. [Google Scholar] [CrossRef]
  69. Braic, M.; Braic, V.; Balaceanu, M.; Zoita, C.; Vladescu, A.; Grigore, E. Characteristics of (TiAlCrNbY)C films deposited by reactive magnetron sputtering. Surf. Coat. Technol. 2010, 204, 2010–2014. [Google Scholar] [CrossRef]
  70. Jiang, S.; Shao, L.; Fan, T.W.; Duan, J.M.; Chen, X.T.; Tang, B.Y. Elastic and thermodynamic properties of high entropy carbide (HfTaZrTi)C and (HfTaZrNb)C from ab initio investigation. Ceram. Int. 2020, 46, 15104–15112. [Google Scholar] [CrossRef]
  71. Gild, J.; Zhang, Y.; Harrington, T.; Jiang, S.; Hu, T.; Quinn, M.C.; Mellor, V.M.; Zhou, N.; Vecchio, K.; Luo, J. High-entropy metal diborides: A new class of high-entropy materials and a new type of ultrahigh temperature ceramics. Sci. Rep. 2016, 6, 37946. [Google Scholar] [CrossRef]
  72. Qin, M.; Yan, Q.; Liu, Y.; Luo, J. A new class of high-entropy M3B4 borides. J. Adv. Ceram. 2021, 10, 166–172. [Google Scholar] [CrossRef]
  73. Wright, A.J.; Luo, J. A step forward from high-entropy ceramics to compositionally complex ceramics: A new perspective. J. Mater. Sci. 2020, 55, 9812–9827. [Google Scholar] [CrossRef] [Green Version]
  74. Oses, C.; Toher, C.; Curtarolo, S. High-entropy ceramics. Nat. Rev. Mater. 2020, 5, 295–309. [Google Scholar] [CrossRef]
  75. De Marco, M.O.; Li, Y.; Li, H.W.; Edalati, K.; Floriano, R. Mechanical synthesis and hydrogen storage characterization of MgVCr and MgVTiCrFe high-entropy alloy. Adv. Eng. Mater. 2020, 22, 1901079. [Google Scholar] [CrossRef]
  76. Edalati, P.; Wang, Q.; Razavi-Khosroshahi, H.; Fuji, M.; Ishihara, T.; Edalati, K. Photocatalytic hydrogen evolution on a high-entropy oxide. J. Mater. Chem. A 2020, 8, 3814–3821. [Google Scholar] [CrossRef]
  77. Edalati, P.; Shen, X.F.; Watanabe, M.; Ishihara, T.; Arita, M.; Fuji, M.; Edalati, K. High-entropy oxyitride as low-bandgap and stable photocatalyst for hydrogen production. J. Mater. Chem. A 2021, 9, 15076–15086. [Google Scholar] [CrossRef]
  78. González-Masís, J.; Cubero-Sesin, J.M.; Campos-Quirós, A.; Edalati, K. Synthesis of biocompatible high-entropy alloy TiNbZrTaHf by high-pressure torsion. Mater. Sci. Eng. A 2021, 825, 141869. [Google Scholar] [CrossRef]
  79. Edalati, K.; Horita, Z. Correlations between hardness and atomic bond parameters of pure metals and semi-metals after processing by high-pressure torsion. Scr. Mater. 2011, 64, 161–164. [Google Scholar] [CrossRef]
  80. Zuttel, A.; Borgschulte, A.; Schlapbach, L. Hydrogen as a Future Energy Carrier, 1st ed.; Wiley VCH Verlag GmbH & Co, KGaA: Weinheim, Germany, 2008. [Google Scholar]
  81. von Colbe, J.B.; Ares, J.R.; Barale, J.; Baricco, M.; Buckley, C.; Capurso, G.; Gallandat, N.; Grant, D.M.; Guzik, M.N.; Jacob, I.; et al. Application of hydrides in hydrogen storage and compression: Achievements, outlook and perspectives. Int. J. Hydrog. Energy 2019, 44, 7780–7808. [Google Scholar] [CrossRef]
  82. Hirscher, M.; Yartys, V.A.; Baricco, M.; von Colbe, J.B.; Blanchard, D.; Bowman, R.C., Jr.; Broom, D.P.; Buckley, C.E.; Chang, F.; Chen, P.; et al. Materials for hydrogen-based energy storage–past, recent progress and future outlook. J. Alloys Compd. 2020, 827, 153548. [Google Scholar] [CrossRef]
  83. Huot, J.; Cuevas, F.; Deledda, S.; Edalati, K.; Filinchuk, Y.; Grosdidier, T.; Hauback, B.C.; Heere, M.; Jensen, T.R.; Latroche, M.; et al. Mechanochemistry of metal hydrides: Recent advances. Materials 2019, 12, 2778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Révész, A.; Gajdics, M. High-pressure torsion of non-equilibrium hydrogen storage materials: A review. Energies 2021, 14, 819. [Google Scholar] [CrossRef]
  85. Edalati, K.; Akiba, E.; Horita, Z. High-pressure torsion for new hydrogen storage materials. Sci. Technol. Adv. Mater. 2018, 19, 185–193. [Google Scholar] [CrossRef] [PubMed]
  86. Karlsson, D.; Ek, G.; Cedervall, J.; Zlotea, C.; Møller, K.T.; Hansen, T.C.; Bednarčík, J.; Paskevicius, M.; Sørby, M.H.; Jensen, T.R.; et al. Structure and hydrogenation properties of a HfNbTiVZr high-entropy alloy. Inorg. Chem. 2018, 57, 2103–2110. [Google Scholar] [CrossRef]
  87. Zlotea, C.; Sow, M.A.; Ek, G.; Couzinié, J.P.; Perrière, L.; Guillot, I.; Bourgon, J.; Møller, K.T.; Jensen, T.R.; Akiba, E.; et al. Hydrogen sorption in TiZrNbHfTa high entropy alloy. J. Alloys Compd. 2019, 775, 667–674. [Google Scholar] [CrossRef]
  88. Zepon, G.; Leiva, D.R.; Strozi, R.B.; Bedoch, A.; Figueroa, S.J.A.; Ishikawa, T.T.; Botta, W.J. Hydrogen-induced phase transition of MgZrTiFe0.5Co0.5Ni0.5 high entropy alloy. Int. J. Hydrog. Energy 2018, 43, 1702–1708. [Google Scholar] [CrossRef]
  89. Kunce, I.; Polanski, M.; Bystrzycki, J. Microstructure and hydrogen storage properties of a TiZrNbMoV high entropy alloy synthesized using laser engineered net shaping (LENS). Int. J. Hydrog. Energy 2014, 39, 9904–9910. [Google Scholar] [CrossRef]
  90. Hu, J.; Shen, H.; Jiang, M.; Gong, H.; Xiao, H.; Liu, Z.; Sun, G.; Zu, X. A DFT Study of Hydrogen storage in high-entropy alloy TiZrHfScMo. Nanomaterials 2019, 9, 461. [Google Scholar] [CrossRef] [Green Version]
  91. Kao, Y.F.; Chen, S.K.; Sheu, J.H.; Lin, J.T.; Lin, W.E.; Yeh, J.W.; Len, S.J.; Liou, T.H.; Wang, C.W. Hydrogen storage properties of multi-principal-component CoFeMnTixVyZrz alloys. Int. J. Hydrog. Energy 2010, 35, 9046–9059. [Google Scholar] [CrossRef]
  92. Kunce, I.; Polanski, M.; Bystrzycki, J. Structure and hydrogen storage properties of a high entropy ZrTiVCrFeNi alloy synthesized using laser engineered net shaping (LENS). Int. J. Hydrog. Energy 2013, 38, 12180–12189. [Google Scholar] [CrossRef]
  93. Yang, S.; Yang, F.; Wu, C.; Chen, Y.; Mao, Y.; Luo, L. Hydrogen storage and cyclic properties of (VFe)60(TiCrCo)40−xZrx (0 ≤ x ≤ 2) alloys. J. Alloys Compd. 2016, 663, 460–465. [Google Scholar] [CrossRef]
  94. Nygård, M.M.; Ek, G.; Karlsson, D.; Sahlberg, M.; Sørby, M.H.; Hauback, B.C. Hydrogen storage in high-entropy alloys with varying degree of local lattice strain. Int. J. Hydrog. Energy 2019, 44, 29140–29149. [Google Scholar] [CrossRef] [Green Version]
  95. Dewangan, S.K.; Sharma, V.K.; Sahu, P.; Kumar, V. Synthesis and characterization of hydrogenated novel AlCrFeMnNiW high entropy alloy. Int. J. Hydrog. Energy 2020, 45, 16984–16991. [Google Scholar] [CrossRef]
  96. Floriano, R.; Zepon, G.; Edalati, K.; Fontana, G.L.B.G.; Mohammadi, A.; Ma, Z.; Li, H.W.; Contieri, R. Hydrogen storage in TiZrNbFeNi high entropy alloys, designed by thermodynamic calculations. Int. J. Hydrog. Energy 2020, 45, 33759–33770. [Google Scholar] [CrossRef]
  97. Floriano, R.; Zepon, G.; Edalati, K.; Fontana, G.L.B.G.; Mohammadi, A.; Ma, Z.; Li, H.W.; Contieri, R.J. Hydrogen storage properties of new A3B2-type TiZrNbCrFe high-entropy alloy. Int. J. Hydrog. Energy 2021, 46, 23757–23766. [Google Scholar] [CrossRef]
  98. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  99. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  100. Wang, W.; Tade, M.O.; Shao, Z. Research progress of perovskite materials in photocatalysis- and photovoltaics-related energy conversion and environmental treatment. Chem. Soc. Rev. 2015, 44, 5371–5408. [Google Scholar] [CrossRef] [PubMed]
  101. Maeda, K.; Domen, K. New non-oxide photocatalysts designed for overall water splitting under visible light. J. Phys. Chem. C 2007, 111, 7851–7861. [Google Scholar] [CrossRef]
  102. Takata, T.; Pan, C.; Domen, K. Recent progress in oxynitride photocatalysts for visible-light-driven water splitting. Sci. Technol. Adv. Mater. 2015, 16, 033506. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Principles of high-pressure torsion method, which includes a disc-shaped sample and two pressure-resistant anvils, made from tool steels or WC—11 wt% Co composites, with shallow and circular flat-bottomed holes on their surfaces.
Figure 1. Principles of high-pressure torsion method, which includes a disc-shaped sample and two pressure-resistant anvils, made from tool steels or WC—11 wt% Co composites, with shallow and circular flat-bottomed holes on their surfaces.
Metals 11 01263 g001
Figure 2. Distribution of elements examined by atom probe tomography in high-entropy alloy, CoCrFeMnNi, synthesized by high-pressure torsion, reproduced from [64], with permission from ELSEVIER, 2019.
Figure 2. Distribution of elements examined by atom probe tomography in high-entropy alloy, CoCrFeMnNi, synthesized by high-pressure torsion, reproduced from [64], with permission from ELSEVIER, 2019.
Metals 11 01263 g002
Figure 3. Variation of hardness versus configurational entropy for TiNb, TiNbZr, TiNbZrTa and high-entropy alloy TiNbZrTaHf, synthesized by high-pressure torsion, reproduced from [78], with permission from ELSEVIER, 2021. Grain size, dislocation density and atomic-size-mismatch parameters were similar for ternary, quaternary, and quinary alloys.
Figure 3. Variation of hardness versus configurational entropy for TiNb, TiNbZr, TiNbZrTa and high-entropy alloy TiNbZrTaHf, synthesized by high-pressure torsion, reproduced from [78], with permission from ELSEVIER, 2021. Grain size, dislocation density and atomic-size-mismatch parameters were similar for ternary, quaternary, and quinary alloys.
Metals 11 01263 g003
Figure 4. Hydrogen–pressure–composition isotherms at 673 K for high-entropy hydride MgTiVFeNi-H, synthesized by ball milling and high-pressure torsion for hydrogen storage, reproduced from [75], with permission from Wiley, 2020.
Figure 4. Hydrogen–pressure–composition isotherms at 673 K for high-entropy hydride MgTiVFeNi-H, synthesized by ball milling and high-pressure torsion for hydrogen storage, reproduced from [75], with permission from Wiley, 2020.
Metals 11 01263 g004
Figure 5. (a) Powder image, (b) Itten color wheel corresponding to orange color of powder, and (c,d) micrographs of powders by scanning electron microscopy for dual-phase high-entropy oxide TiZrHfNbTaO11, synthesized by high-pressure torsion and high-temperature oxidation, reproduced from [76], with permission from ROYAL SOCIETY OF CHEMISTRY, 2020.
Figure 5. (a) Powder image, (b) Itten color wheel corresponding to orange color of powder, and (c,d) micrographs of powders by scanning electron microscopy for dual-phase high-entropy oxide TiZrHfNbTaO11, synthesized by high-pressure torsion and high-temperature oxidation, reproduced from [76], with permission from ROYAL SOCIETY OF CHEMISTRY, 2020.
Metals 11 01263 g005
Figure 6. (a) Photocatalytic hydrogen production for three cycles and (b) X-ray diffraction patterns before and after photocatalysis for dual-phase high-entropy oxide TiZrHfNbTaO11, synthesized by high-pressure torsion and high-temperature oxidation, reproduced from [76], with permission from ROYAL SOCIETY OF CHEMISTRY, 2020.
Figure 6. (a) Photocatalytic hydrogen production for three cycles and (b) X-ray diffraction patterns before and after photocatalysis for dual-phase high-entropy oxide TiZrHfNbTaO11, synthesized by high-pressure torsion and high-temperature oxidation, reproduced from [76], with permission from ROYAL SOCIETY OF CHEMISTRY, 2020.
Metals 11 01263 g006
Figure 7. (a) Bright-field image, (b) selected area electron diffraction pattern, (c) dark-field image, (d) high-resolution image, (e) lattice image of FCC phase and (f) lattice image of monoclinic phase taken by transmission electron microscopy for nanograined dual-phase high-entropy oxynitride TiZrHfNbTaO6N3, synthesized by high-pressure torsion followed by high-temperature oxidation and nitriding, reproduced from [77], with permission from ROYAL SOCIETY OF CHEMISTRY, 2021.
Figure 7. (a) Bright-field image, (b) selected area electron diffraction pattern, (c) dark-field image, (d) high-resolution image, (e) lattice image of FCC phase and (f) lattice image of monoclinic phase taken by transmission electron microscopy for nanograined dual-phase high-entropy oxynitride TiZrHfNbTaO6N3, synthesized by high-pressure torsion followed by high-temperature oxidation and nitriding, reproduced from [77], with permission from ROYAL SOCIETY OF CHEMISTRY, 2021.
Metals 11 01263 g007
Figure 8. (a) Band structure compared with required potentials for water splitting, and (b) light absorbance compared with binary and high-entropy oxides for nanograined dual-phase high-entropy oxynitride TiZrHfNbTaO6N3, synthesized by high-pressure torsion followed by high-temperature oxidation and nitriding, reproduced from [77], with permission from ROYAL SOCIETY OF CHEMISTRY, 2021.
Figure 8. (a) Band structure compared with required potentials for water splitting, and (b) light absorbance compared with binary and high-entropy oxides for nanograined dual-phase high-entropy oxynitride TiZrHfNbTaO6N3, synthesized by high-pressure torsion followed by high-temperature oxidation and nitriding, reproduced from [77], with permission from ROYAL SOCIETY OF CHEMISTRY, 2021.
Metals 11 01263 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Edalati, K.; Li, H.-W.; Kilmametov, A.; Floriano, R.; Borchers, C. High-Pressure Torsion for Synthesis of High-Entropy Alloys. Metals 2021, 11, 1263. https://doi.org/10.3390/met11081263

AMA Style

Edalati K, Li H-W, Kilmametov A, Floriano R, Borchers C. High-Pressure Torsion for Synthesis of High-Entropy Alloys. Metals. 2021; 11(8):1263. https://doi.org/10.3390/met11081263

Chicago/Turabian Style

Edalati, Kaveh, Hai-Wen Li, Askar Kilmametov, Ricardo Floriano, and Christine Borchers. 2021. "High-Pressure Torsion for Synthesis of High-Entropy Alloys" Metals 11, no. 8: 1263. https://doi.org/10.3390/met11081263

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