Next Article in Journal
Optimization of Open Die Ironing Process through Artificial Neural Network for Rapid Process Simulation
Next Article in Special Issue
Open Porous α + β Titanium Alloy by Liquid Metal Dealloying for Biomedical Applications
Previous Article in Journal
Coupling Finite Element Analysis and the Theory of Critical Distances to Estimate Critical Loads in Al6060-T66 Tubular Beams Containing Notches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nanoporous High-Entropy Alloy by Liquid Metal Dealloying

by
Artem Vladimirovich Okulov
1,
Soo-Hyun Joo
2,3,
Hyoung Seop Kim
4,
Hidemi Kato
2 and
Ilya Vladimirovich Okulov
2,5,6,7,*
1
Division of Materials Mechanics, Institute of Materials Research, Helmholtz-Zentrum Geesthacht, 21502 Geesthacht, Germany
2
Institute for Materials Research, Tohoku University, Katahira 2-1-1, Sendai 980-8577, Japan
3
Department of Materials Science and Engineering, Dankook University, 119 Dandae-ro, Cheonan 31116, Korea
4
Department of Materials Science and Engineering Pohang, University of Science and Technology, 77 Cheongam-Ro, 37673 Pohang, Korea
5
Institute of Natural Sciences and Mathematics, Ural Federal University, 620000 Ekaterinburg, Russia
6
Faculty of Production Engineering, University of Bremen, Badgasteiner Str. 1, 28359 Bremen, Germany
7
Leibniz Institute for Materials Engineering-IWT, Badgasteiner Str. 3, 28359 Bremen, Germany
*
Author to whom correspondence should be addressed.
Metals 2020, 10(10), 1396; https://doi.org/10.3390/met10101396
Submission received: 23 September 2020 / Revised: 12 October 2020 / Accepted: 15 October 2020 / Published: 21 October 2020
(This article belongs to the Special Issue Nanoporous and Nanocomposite Materials by Dealloying)

Abstract

:
High-entropy nanomaterials possessing high accessible surface areas have demonstrated outstanding catalytic performance, beating that found for noble metals. In this communication, we report about the synthesis of a new, nanoporous, high-entropy alloy (HEA) possessing open porosity. The nanoporous, high-entropy Ta19.1Mo20.5Nb22.9V30Ni7.5 alloy (at%) was fabricated from a precursor (TaMoNbV)25Ni75 alloy (at%) by liquid metal dealloying using liquid magnesium (Mg). Directly after dealloying, the bicontinuous nanocomposite consisting of a Mg-rich phase and a phase with a bulk-centered cubic (bcc) structure was formed. The Mg-rich phase was removed with a 3M aqueous solution of nitric acid to obtain the open, porous, high-entropy Ta19.1Mo20.5Nb22.9V30Ni7.5 alloy (at%). The ligament size of this nanoporous HEA is about 69 ± 9 nm, indicating the high surface area in this material.

1. Introduction

High-entropy alloys (HEAs) have attracted intensive research during the past decade due to their unique properties [1,2,3,4]. Unlike the conventional alloy systems based on one principal element, or occasionally two principal elements, HEAs contain at least five major elements at concentrations ranging between 5 and 35 at%. The implicit hypothesis in the term “high-entropy” is that the high mixing entropy of these modern types of alloys can exceed the enthalpies of compound formation. Therefore, single-phase solid solutions are formed. The outstanding chemical and mechanical properties of HEAs [1,4,5,6,7,8,9,10,11] are directly related to the four “core effects”, i.e., high entropy, sluggish diffusion, severe lattice distortion, and the cocktail effect [12]. Especially, the sluggish diffusion effect is important for the exceptional high-temperature strength [13,14] and high-temperature structural stability of HEAs [13]. Recently, a number of HEA nanomaterials possessing high intrinsic catalytic activity for the decomposition of ammonia, CO oxidation, and water splitting were discovered [15,16,17]. In this case, the excellent functional, e.g., catalytic, properties can only be revealed for HEAs possessing a very large accessible surface area.
Several methods have been used for the synthesis of high-entropy nanomaterials possessing large surface areas, including chemical dealloying [17], sputtering [18], and carbothermal shock synthesis [4]. Recently, Joo et al. demonstrated the synthesis of nanoporous HEAs by liquid metal dealloying (LMD) [19]. Liquid metal dealloying, invented by Wada and Kato et al. in 2011 [20], is a metallurgical method for the synthesis of porous and nanocomposite materials [21,22,23,24]. In contrast to conventional chemical dealloying, the LMD method utilizes a metallic melt instead of a chemical etchant. Thus, during the LMD process, oxidation is prevented and reactive materials such as Mg [25] can be synthesized. To date, a wide variety of porous materials obtained by LMD have been reported, including silicon (Si) [26], steels [24,27,28,29,30], graphene (C) [31,32,33,34], chromium (Cr) [27], niobium (Nb) [35,36], titanium (Ti) [37,38], and titanium alloys (TiZr, TiHf, TiNb, TiFe, and TiMo) [23,38,39,40,41]. Moreover, LMD was successfully applied for the surface functionalization of biomedical Ti–6Al–4V and Ti–6Al–7Nb alloys [42,43], design of low modulus composites such as Fe–Mg [39,44], and synthesis of high-coercivity NdFeB-based permanent magnets [45]. Since the LMD process is typically conducted at elevated temperatures, it demonstrates significantly higher dealloying rates as compared to chemical dealloying. On the other hand, the drawback of LMD is a significant and unavoidable microstructural coarsening due to the high temperatures. The coarsening destroys the functional properties of the nanoporous materials associated with their high surface area. In our recent study, we demonstrated a material design principle enabling avoiding thermal coarsening in nanoporous materials [19]. In this communication, we report about the synthesis of a new, open, porous, high-entropy Ta19.1Mo20.5Nb22.9V30Ni7.5 (at%) alloy by the liquid metal dealloying of the (TaMoNbV)25Ni75 precursor alloy (at%).

2. Materials and Methods

The design of a material system for liquid metal dealloying begins from the selection of elements. The elements for the liquid metal dealloying are selected based on the enthalpy of mixture ( Δ H ( M g e l e m e n t ) m i x ) between a corrosive medium such as magnesium melt and the considered element (Figure 1a). Elements exhibiting a positive value of Δ H ( M g e l e m e n t ) m i x such as Ti, V, Cr, Fe, Mn, Co, Zr, Nb, Mo, Hf, and Ta (Figure 1b) are immiscible in Mg and, therefore, self-organize into bicontinuous structures upon dealloying. At the same time, elements with a negative value of Δ H ( M g e l e m e n t ) m i x such as B, Al, Si, P, Ca, Ni, Cu, Zn, Sr, Pd, Ag, In, Sn, Pt, and Au (Figure 1b) dissolve in Mg upon the dealloying process. Thereby, a large number of element combinations can be selected for liquid/solid metal dealloying and, particularly, for obtaining multicomponent porous scaffolds, including porous high-entropy alloys. To demonstrate the effectiveness of the above-described approach, a precursor (TaMoNbV)25Ni75 alloy (at%) was designed and dealloyed in magnesium at 1123 K for 20 min.
The precursor (TaMoNbV)25Ni75 alloy (at%) in the shape of rods (1 mm in diameter) was fabricated and prepared from pure metals (99.99%) with an arc melting device coupled with a suction casting set-up under an argon (Ar) atmosphere (Mini Arc Melter MAM-1, Edmund Bühler, Germany). To carry out dealloying, rods 2 mm long together with a magnesium mesh were heated in a glassy carbon crucible under Ar flow using an infrared furnace (IRF 10, Behr, Switzerland). Upon heating (at a heating rate of about 40 K s−1), magnesium metal melts and diffuses into the precursor to selectively dissolve Ni out of the precursor alloy. The remaining elements diffuse along the metal/liquid interface [20,22] to form bicontinuous ligament structures. After dealloying, the crucible is cooled down to room temperature and the nanocomposite consisting of new phases, namely Mg-rich and high-entropy phases, is formed (Figure 2). To obtain porous HEA material, the Mg-rich phase was selectively etched out in 3 M HNO3 for 5 h (Figure 2). Structural investigation of the precursor alloys and porous samples was performed by X-ray diffraction in Bragg–Brentano geometry (D8 Advance, Bruker, Germany) with Cu-Kα radiation. The device was equipped with a position-sensitive detector (LynxEye, Bruker, Germany), enabling us to achieve acceptable signal-to-noise ratios within a few hours of measuring time despite the smallness of the samples. Scanning electron microscopy (Nova Nanolab 200, FEI, Hillsboro, OR, USA) coupled with energy-dispersive X-ray analysis (EDAX, Weiterstadt, Germany) was used to explore the microstructure and composition.

3. Results and Discussion

Figure 3 shows the X-ray diffraction pattern and microstructure of the final open nanoporous HEA. According to the X-ray analysis, the nanoporous HEA mainly consists of a phase with a bulk-centered cubic (bcc) structure and a minor amount of an unknown phase (Figure 3). According to the energy-dispersive X-ray analysis (EDX), the bcc phase is a solid solution of five elements with the following chemical composition: Ta19.1Mo20.5Nb22.9V30Ni7.5 (at%). The single-phase structure and multicomponent chemical composition (the concentration of each element is above 5 at%) of the current porous material indicate that this is a high-entropy alloy. The SEM analysis reveals the nanoporous structure of the designed HEA (Figure 3). The average size of the ligaments is about 69 ± 9 nm, which is 1–2 orders of magnitude lower than is typically observed for the porous materials obtained by LMD [23,38]. The ligament coarsening during the liquid metal dealloying is primarily associated with surface diffusion [37]. Thus, the activation energy for the surface diffusion controls the coarsening. It seems that the sluggish diffusion effect observed in the high-entropy alloys [47] prevents coarsening in the current nanoporous HEA. However, the controversial results regarding the sluggish diffusion phenomenon require a deeper understanding of the basic mechanisms suppressing coarsening in the nanoporous HEAs.
As was proposed by Chen and Sieradzki [48], the microstructural length scale (ligament size) of porous materials fabricated by chemical dealloying at room temperature is correlated with the inverse “homologous dealloying temperature” or 1/TH = Tmelting point/T298K. This universal correlation was expanded by McCue et al. [49] to the porous materials obtained by liquid metal dealloying. It was shown that the porous materials obtained by LMD follow a similar trend. The homologous dealloying temperature, in this case, was modified to TH = Tmelting point/Tdealloying temperature. Later, Joo and Okulov et al. demonstrated that porous HEAs synthesized by LMD possess a different correlation of ligament size with homologous temperature [19]. Specifically, the universal relationship for the nanoporous HEAs shifts down by one order of magnitude towards the smaller-size regime. Similar to the reported nanoporous HEAs [19], the currently designed nanoporous HEA also deviates from the universal relationship proposed by Chen and Sieradzki [48] and validates the recent high-entropy design strategy [19] for suppressing thermal coarsening in nanoporous materials obtained by dealloying (Figure 4).
In summary, we have successfully designed a new, open, nanoporous, high-entropy Ta–Mo–Nb–V–Ni alloy by liquid metal dealloying based on the design strategy proposed in our recent study [19]. The currently developed open, porous HEA predominantly consists of a body-centered cubic (bcc) solid solution phase. The ligament size of the nanoporous Ta–Mo–Nb–V–Ni alloy after 20 min of dealloying at 1123 K is 69 ± 9 nm, which indicates its high stability against thermal coarsening. The nanoporous, high-entropy materials are promising candidates for functional applications such as catalysis.

Author Contributions

Conceptualization—H.K., H.S.K., and I.V.O.; formal analysis—A.V.O., S.-H.J., and I.V.O.; funding acquisition—H.K., H.S.K., and I.V.O.; investigation—A.V.O., S.-H.J., and I.V.O.; supervision—I.V.O., H.S.K., and H.K.; validation—A.V.O., I.V.O., and S.-H.J.; writing—original draft, A.V.O. and I.V.O.; writing—review and editing—all. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the International Collaboration Center, Institute for Materials Research (ICC-IMR), Tohoku University, Japan (no grant number) and the German Science Foundation under the Leibniz Program, grant number MA 3333/13-1.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yeh, J. Recent Progress in High-entropy Alloys. Ann. Chim. Sci. Mater. 2006, 31, 633–648. [Google Scholar] [CrossRef]
  2. 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]
  3. 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]
  4. Yao, Y.; Huang, Z.; Xie, P.; Lacey, S.D.; Jacob, R.J.; Xie, H.; Chen, F.; Nie, A.; Pu, T.; Rehwoldt, M.; et al. Carbothermal shock synthesis of high-entropy-alloy nanoparticles. Science (80-.) 2018, 359, 1489–1494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Gludovatz, B.; Hohenwarter, A.; Catoor, D.; Chang, E.H.; George, E.P.; Ritchie, R.O. A fracture-resistant high-entropy alloy for cryogenic applications. Science (80-.) 2014, 345, 1153–1158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. 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]
  7. Senkov, O.N.; Scott, J.M.; Senkova, S.V.; Miracle, D.B.; Woodward, C.F. Microstructure and room temperature properties of a high-entropy TaNbHfZrTi alloy. J. Alloys Compd. 2011, 509, 6043–6048. [Google Scholar] [CrossRef]
  8. Okulov, I.V.; Wendrock, H.; Volegov, A.S.; Attar, H.; Kühn, U.; Skrotzki, W.; Eckert, J. High strength beta titanium alloys: New design approach. Mater. Sci. Eng. A 2015, 628, 297–302. [Google Scholar] [CrossRef]
  9. Okulov, I.V.; Bönisch, M.; Okulov, A.V.; Volegov, A.S.; Attar, H.; Ehtemam-Haghighi, S.; Calin, M.; Wang, Z.; Hohenwarter, A.; Kaban, I.; et al. Phase formation, microstructure and deformation behavior of heavily alloyed TiNb- and TiV-based titanium alloys. Mater. Sci. Eng. A 2018, 733, 80–86. [Google Scholar] [CrossRef]
  10. Tsai, M.H.; Yeh, J.W.; Gan, J.Y. Diffusion barrier properties of AlMoNbSiTaTiVZr high-entropy alloy layer between copper and silicon. Thin Solid Films 2008, 516, 5527–5530. [Google Scholar] [CrossRef]
  11. Li, Z.; Pradeep, K.G.; Deng, Y.; Raabe, D.; Tasan, C.C. Metastable high-entropy dual-phase alloys overcome the strength–ductility trade-off. Nature 2016, 534, 227–230. [Google Scholar] [CrossRef] [PubMed]
  12. Yeh, J.W. Physical Metallurgy of High-Entropy Alloys. J. Miner. Met. Mater. Soc. 2015, 67, 2254–2261. [Google Scholar] [CrossRef]
  13. Senkov, O.N.; Wilks, G.B.; Scott, J.M.; Miracle, D.B. Mechanical properties of Nb25Mo25Ta25W25 and V20Nb20Mo20Ta20W20. Intermetallics 2011, 19, 698–706. [Google Scholar] [CrossRef]
  14. Senkov, O.N.; Senkova, S.V.; Woodward, C. Effect of aluminum on the microstructure and properties of two refractory high-entropy alloys. Acta Mater. 2014, 68, 214–228. [Google Scholar] [CrossRef]
  15. Xie, P.; Yao, Y.; Huang, Z.; Liu, Z.; Zhang, J.; Li, T.; Wang, G.; Shahbazian-Yassar, R.; Hu, L.; Wang, C. Highly efficient decomposition of ammonia using high-entropy alloy catalysts. Nat. Commun. 2019, 10, 1–12. [Google Scholar] [CrossRef] [Green Version]
  16. Qiu, H.J.; Fang, G.; Wen, Y.; Liu, P.; Xie, G.; Liu, X.; Sun, S. Nanoporous high-entropy alloys for highly stable and efficient catalysts. J. Mater. Chem. A 2019, 7, 6499–6506. [Google Scholar] [CrossRef]
  17. Jin, Z.; Lv, J.; Jia, H.; Liu, W.; Li, H.; Chen, Z.; Lin, X.; Xie, G.; Liu, X.; Sun, S.; et al. Nanoporous Al-Ni-Co-Ir-Mo High-Entropy Alloy for Record-High Water Splitting Activity in Acidic Environments. Small 2019, 15, 1–7. [Google Scholar] [CrossRef]
  18. Löffler, T.; Meyer, H.; Savan, A.; Wilde, P.; Garzón Manjón, A.; Chen, Y.T.; Ventosa, E.; Scheu, C.; Ludwig, A.; Schuhmann, W. Discovery of a Multinary Noble Metal–Free Oxygen Reduction Catalyst. Adv. Energy Mater. 2018, 8, 1–7. [Google Scholar] [CrossRef]
  19. Joo, S.-H.; Bae, J.W.; Park, W.-Y.; Shimada, Y.; Wada, T.; Kim, H.S.; Takeuchi, A.; Konno, T.J.; Kato, H.; Okulov, I.V. Beating Thermal Coarsening in Nanoporous Materials via High-Entropy Design. Adv. Mater. 2020, 32, 1906160. [Google Scholar] [CrossRef]
  20. Wada, T.; Yubuta, K.; Inoue, A.; Kato, H. Dealloying by metallic melt. Mater. Lett. 2011, 65, 1076–1078. [Google Scholar] [CrossRef]
  21. McCue, I.; Gaskey, B.; Geslin, P.A.; Karma, A.; Erlebacher, J. Kinetics and morphological evolution of liquid metal dealloying. Acta Mater. 2016, 115, 10–23. [Google Scholar] [CrossRef] [Green Version]
  22. Geslin, P.; Mccue, I.; Erlebacher, J.; Karma, A. Topology-generating interfacial pattern formation during liquid metal dealloying. Nat. Commun. 2015, 6, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Okulov, I.V.; Okulov, A.V.; Soldatov, I.V.; Luthringer, B.; Willumeit-Römer, R.; Wada, T.; Kato, H.; Weissmüller, J.; Markmann, J. Open porous dealloying-based biomaterials as a novel biomaterial platform. Mater. Sci. Eng. C 2018, 83, 95–103. [Google Scholar] [CrossRef] [PubMed]
  24. Joo, S.-H.; Kato, H. Transformation mechanisms and governing orientation relationships through selective dissolution of Ni via liquid metal dealloying from (FeCo)xNi100−x precursors. Mater. Des. 2020, 185, 108271. [Google Scholar] [CrossRef]
  25. Okulov, I.V.; Lamaka, S.V.; Wada, T.; Yubuta, K.; Zheludkevich, M.L.; Weissmüller, J.; Markmann, J.; Kato, H. Nanoporous magnesium. Nano Res. 2018, 11, 6428–6435. [Google Scholar] [CrossRef]
  26. Wada, T.; Ichitsubo, T.; Yubuta, K.; Segawa, H.; Yoshida, H.; Kato, H. Bulk-nanoporous-silicon negative electrode with extremely high cyclability for lithium-ion batteries prepared using a top-down process. Nano Lett. 2014, 14, 4505–4510. [Google Scholar] [CrossRef]
  27. Wada, T.; Kato, H. Three-dimensional open-cell macroporous iron, chromium and ferritic stainless steel. Scr. Mater. 2013, 68, 723–726. [Google Scholar] [CrossRef]
  28. Joo, S.-H.; Wada, T.; Kato, H. Development of porous FeCo by liquid metal dealloying: Evolution of porous morphology and effect of interaction between ligaments and melt. Mater. Des. 2019, 180, 107908. [Google Scholar] [CrossRef]
  29. Mokhtari, M.; Wada, T.; Le Bourlot, C.; Mary, N.; Duchet-Rumeau, J.; Kato, H.; Maire, E. Low cost high specific surface architectured nanoporous metal with corrosion resistance produced by liquid metal dealloying from commercial nickel superalloy. Scr. Mater. 2019, 163, 5–8. [Google Scholar] [CrossRef]
  30. Mokhtari, M.; Wada, T.; Le Bourlot, C.; Duchet-Rumeau, J.; Kato, H.; Maire, E.; Mary, N. Corrosion resistance of porous ferritic stainless steel produced by liquid metal dealloying of Incoloy 800. Corros. Sci. 2020, 108468. [Google Scholar] [CrossRef]
  31. Yu, S.G.; Yubuta, K.; Wada, T.; Kato, H. Three-dimensional bicontinuous porous graphite generated in low temperature metallic liquid. Carbon N. Y. 2016, 96, 403–410. [Google Scholar] [CrossRef]
  32. Greenidge, G.; Erlebacher, J. Porous graphite fabricated by liquid metal dealloying of silicon carbide. Carbon N. Y. 2020, 165, 45–54. [Google Scholar] [CrossRef]
  33. Park, W.-Y.; Wada, T.; Joo, S.-H.; Han, J.; Kato, H. Novel hierarchical nanoporous graphene nanoplatelets with excellent rate capabilities produced via self-templating liquid metal dealloying. Mater. Today Commun. 2020, 24, 101120. [Google Scholar] [CrossRef]
  34. Shao, J.-C.; Jin, H.-J. From liquid metal dealloying to liquid metal expulsion. J. Mater. Sci. 2020, 55, 8337–8345. [Google Scholar] [CrossRef]
  35. Kim, J.W.; Tsuda, M.; Wada, T.; Yubuta, K.; Kim, S.G.; Kato, H. Optimizing niobium dealloying with metallic melt to fabricate porous structure for electrolytic capacitors. Acta Mater. 2015, 84, 497–505. [Google Scholar] [CrossRef]
  36. Kim, J.W.; Wada, T.; Kim, S.G.; Kato, H. Sub-micron porous niobium solid electrolytic capacitor prepared by dealloying in a metallic melt. Mater. Lett. 2014, 116, 223–226. [Google Scholar] [CrossRef]
  37. Tsuda, M.; Wada, T.; Kato, H. Kinetics of formation and coarsening of nanoporous α-titanium dealloyed with Mg melt. J. Appl. Phys. 2013, 114. [Google Scholar] [CrossRef]
  38. Okulov, I.V.; Weissmüller, J.; Markmann, J. Dealloying-based interpenetrating-phase nanocomposites matching the elastic behavior of human bone. Sci. Rep. 2017, 7, 20. [Google Scholar] [CrossRef]
  39. Okulov, A.V.; Volegov, A.S.; Weissmüller, J.; Markmann, J.; Okulov, I.V. Dealloying-based metal-polymer composites for biomedical applications. Scr. Mater. 2018, 146, 290–294. [Google Scholar] [CrossRef]
  40. Okulov, I.V.; Okulov, A.V.; Volegov, A.S.; Markmann, J. Tuning microstructure and mechanical properties of open porous TiNb and TiFe alloys by optimization of dealloying parameters. Scr. Mater. 2018, 154, 68–72. [Google Scholar] [CrossRef]
  41. Berger, S.A.; Okulov, I.V. Open porous α + β titanium alloy by liquid metal dealloying for biomedical applications. Metals 2020, in press. [Google Scholar]
  42. Okulov, I.V.; Joo, S.-H.; Okulov, A.V.; Volegov, A.S.; Luthringer, B.; Willumeit-Römer, R.; Zhang, L.; Mädler, L.; Eckert, J.; Kato, H. Surface Functionalization of Biomedical Ti-6Al-7Nb Alloy by Liquid Metal Dealloying. Nanomaterials 2020, 10. [Google Scholar] [CrossRef]
  43. Fukuzumi, Y.; Wada, T.; Kato, H. Surface Improvement for Biocompatibility of Ti-6Al-4V by Dealloying in Metallic Melt BT-Interface Oral Health Science 2014; Sasaki, K., Suzuki, O., Takahashi, N., Eds.; Springer: Tokyo, Japan, 2015; pp. 93–101. [Google Scholar]
  44. Okulov, I.V.; Geslin, P.-A.; Soldatov, I.V.; Ovri, H.; Joo, S.-H.; Kato, H. Anomalously low modulus of the interpenetrating-phase composite of Fe and Mg obtained by liquid metal dealloying. Scr. Mater. 2019, 163, 133–136. [Google Scholar] [CrossRef]
  45. Volegov, A.S.; Andreev, S.V.; Selezneva, N.V.; Ryzhikhin, I.A.; Kudrevatykh, N.V.; Mädler, L.; Okulov, I.V. Additive manufacturing of heavy rare earth free high-coercivity permanent magnets. Acta Mater. 2020, 188, 733–739. [Google Scholar] [CrossRef]
  46. Takeuchi, A.; Inoue, A. Metallic Glasses By Atomic Size Difference, Heat of Mixing and Period of Constituent Elements and Its Application To Characterization of the Main Alloying Element. Mater. Trans. 2005, 46, 2817–2829. [Google Scholar] [CrossRef] [Green Version]
  47. Tsai, K.Y.; Tsai, M.H.; Yeh, J.W. Sluggish diffusion in Co-Cr-Fe-Mn-Ni high-entropy alloys. Acta Mater. 2013, 61, 4887–4897. [Google Scholar] [CrossRef]
  48. Chen, Q.; Sieradzki, K. Spontaneous evolution of bicontinuous nanostructures in dealloyed Li-based systems. Nat. Mater. 2013, 12, 1102–1106. [Google Scholar] [CrossRef]
  49. McCue, I.; Karma, A.; Erlebacher, J. Pattern formation during electrochemical and liquid metal dealloying. MRS Bull. 2018, 43, 27–34. [Google Scholar] [CrossRef]
Figure 1. Selection of elements for liquid metal dealloying. (a) The values of Δ H ( A B ) m i x (kJ/mol) calculated by Miedema’s model for atomic pairs between Mg and elements indicated in the plot [46]; (b) Required relationship of the values of enthalpy of mixing between elements for the liquid metal dealloying of a master alloy AB in a liquid metal C.
Figure 1. Selection of elements for liquid metal dealloying. (a) The values of Δ H ( A B ) m i x (kJ/mol) calculated by Miedema’s model for atomic pairs between Mg and elements indicated in the plot [46]; (b) Required relationship of the values of enthalpy of mixing between elements for the liquid metal dealloying of a master alloy AB in a liquid metal C.
Metals 10 01396 g001
Figure 2. Schematic illustration of liquid metal dealloying process.
Figure 2. Schematic illustration of liquid metal dealloying process.
Metals 10 01396 g002
Figure 3. X-ray diffraction pattern and scanning electron micrograph of the nanoporous high-entropy TaMoNbVNi alloy.
Figure 3. X-ray diffraction pattern and scanning electron micrograph of the nanoporous high-entropy TaMoNbVNi alloy.
Metals 10 01396 g003
Figure 4. Ligament size versus homologous temperature in conventional nano- and microporous materials and the nanoporous high-entropy alloys (HEAs) [19] (note: non-LMD represents the nanoporous materials obtained by chemical dealloying [49]; the dealloying time is different from that for the LMD-based materials: 10 min—open symbols, 20 min—half-open symbols, and 60 min—closed symbols).
Figure 4. Ligament size versus homologous temperature in conventional nano- and microporous materials and the nanoporous high-entropy alloys (HEAs) [19] (note: non-LMD represents the nanoporous materials obtained by chemical dealloying [49]; the dealloying time is different from that for the LMD-based materials: 10 min—open symbols, 20 min—half-open symbols, and 60 min—closed symbols).
Metals 10 01396 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Okulov, A.V.; Joo, S.-H.; Kim, H.S.; Kato, H.; Okulov, I.V. Nanoporous High-Entropy Alloy by Liquid Metal Dealloying. Metals 2020, 10, 1396. https://doi.org/10.3390/met10101396

AMA Style

Okulov AV, Joo S-H, Kim HS, Kato H, Okulov IV. Nanoporous High-Entropy Alloy by Liquid Metal Dealloying. Metals. 2020; 10(10):1396. https://doi.org/10.3390/met10101396

Chicago/Turabian Style

Okulov, Artem Vladimirovich, Soo-Hyun Joo, Hyoung Seop Kim, Hidemi Kato, and Ilya Vladimirovich Okulov. 2020. "Nanoporous High-Entropy Alloy by Liquid Metal Dealloying" Metals 10, no. 10: 1396. https://doi.org/10.3390/met10101396

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