Next Article in Journal
Analysis of Plastic Improvement and Interference Behavior in Current-Assisted Riveting of CFRP Laminates
Next Article in Special Issue
Metallization on Sapphire and Low-Temperature Joining with Metal Substrates
Previous Article in Journal
Effect of Curing Regime on the Mechanical Strength, Hydration, and Microstructure of Ecological Ultrahigh-Performance Concrete (EUHPC)
Previous Article in Special Issue
Effect of Radiofrequency Plasma Spheroidization Treatment on the Laser Directed Energy Deposited Properties of Low-Cost Hydrogenated-Dehydrogenated Titanium Powder
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Designing High Entropy Bulk Metallic Glass (HE-BMG) by Similar Element Substitution/Addition

1
Marine Equipment and Technology Institute, Jiangsu University of Science and Technology, Zhenjiang 212100, China
2
School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(5), 1669; https://doi.org/10.3390/ma15051669
Submission received: 31 December 2021 / Revised: 10 February 2022 / Accepted: 15 February 2022 / Published: 23 February 2022
(This article belongs to the Special Issue Physical Metallurgy of Metals and Alloys)

Abstract

:
In this paper, we report that two newly designed high entropy bulk metallic glasses (HE-BMGs), Ti20Hf20Cu20Ni20Be20 with a critical diameter of 2 mm, and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 with a critical diameter of 1.5 mm, can be fabricated by copper mold casting method. These newly developed HE-BMGs exhibited a high fracture strength over 2300 MPa. The glass forming ability and atomic size distribution characteristics of the HE-BMGs are discussed in detail. Moreover, a parameter δ′ was proposed to evaluate the atomic size distribution characteristics in different HEAs. It showed that this new parameter is closely related to the degree of lattice distortion and phase selection of high-entropy alloys. Adjusting the value of δ′ parameter by similar element substitution/addition would be beneficial for designing high entropy bulk metallic glasses.

1. Introduction

In the past few decades, bulk metallic glasses (BMGs) [1,2,3,4,5,6,7,8,9] and high entropy alloys (HEAs) [10,11,12,13,14,15,16,17,18] have attracted much attention, owing to their unique structure and properties, such as high strength/hardness, good corrosion/wear resistance, etc. Previously, BMGs and HEAs were developed separately in most cases, following different composition design and fabrication routes. While recent studies show that intersections exist between these two domains, namely some HEAs with meticulously designed composition could be made into BMGs, and hence the high entropy bulk metallic glasses (HE-BMGs) were developed [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41]. An investigation into HE-BMGs is beneficial for understanding the phase formation rules of HEAs and fundamental issues of BMGs, so it is very important to develop more HE-BMGs.
In our previous work, a Ti20Zr20Cu20Ni20Be20 HE-BMG with a critical diameter of 3 mm was successfully obtained by copper mold casting method [24]. By introducing Hf as the sixth constituent element, Ti16.7Zr16.7Hf16.7Cu16.7Ni16.7Be16.7 with a critical diameter of 15 mm [25] and a series of Ti20Zr20Hf20(Cu20xNix)Be20 HE-BMGs with critical diameters of larger than 12 mm was developed [26,27]. These results indicate that similar element substitution/addition is an effective way for developing new HE-BMGs, just the same as traditional BMGs. Since Hf is an element chemically similar to Zr while Nb and Zr are also very close in the periodic table of elements, it is reasonable to suppose that by substituting Zr with Hf, or by adding Nb in the Ti20Zr20Cu20Ni20Be20 quinary HEA system, new HE-BMG with good properties can be obtained. Accordingly, two new HEAs, namely Ti20Hf20Cu20Ni20Be20 and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7, were designed to verify this assumption, and their glass-forming ability, atomic size distribution characteristics, lattice distortion, and phase selection rules of HEAs are discussed in detail.

2. Experimental

The master alloy ingots with nominal compositions of Ti20Hf20Cu20Ni20Be20 and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 in equal atomic ratio were prepared by arc melting the mixtures of high purity Ti, Hf, Cu, Ni, Zr, Nb plates, and Be granules (purity higher than 99.99 wt.%) within a pure argon gas environment. Cylindrical rods with different diameters were prepared by copper mold injection or suction casting method. Arc melting and casting was conducted on multi-functional high vacuum arc-melting and melt-spinning system, which was produced by SKY Technology Development Corporation, Shenyang, China. The glassy nature of these as-prepared samples was examined by X-ray diffraction (XRD) technique using a Rigaku D/max-RB XRD spectrometry (Rigaku Corporation, Tokyo, Japan) with Cu Kα radiation (λ = 0.15406 nm). Thermal properties of the glassy alloys were examined by a Shimadzu DSC-60 differential scanning calorimeter (Shimadzu Corporation, Kyoto, Japan) instrument under the protection of N2 gas (flow rate: 50 mL/min). The applied heating rate was set as 20 K/min. The DSC instrument was calibrated with In and Zn standard specimens. The errors are within ±1 K. Compression tests with specimens of Ø2 × 4 mm and Ø1.5 × 3 mm in size were carried out on WDW-100 testing machine (Shanghai Precision Instrument Co., Ltd, Shanghai, China) under a stain rate of 4 × 10−4 s−1.

3. Results

Figure 1 shows the XRD spectra of the as-cast Ti20Hf20Cu20Ni20Be20 and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 rods with different diameters. No sharp diffraction peak corresponding to the crystalline phase was observed in the Ø2 mm Ti20Hf20Cu20Ni20Be20 and Ø1.5 mm Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 samples, indicating that they both possess a fully amorphous structure.
The DSC curves of the Ti20Hf20Cu20Ni20Be20 and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 samples are shown in Figure 2. The highest test temperature reached 1273 K (1000 °C). However, since the endothermic peak is very high in the high temperature part, glass transition would be very ambiguous in the curve. In order to demonstrate the glass transition phenomenon (which is very important for glasses) clearly, we just cut out temperature less than 1000 K in Figure 2. The glass transition temperature Tg and initial crystallization temperature Tx were marked with arrows. Tg, Tx, Tm (melting temperature) and Tl (liquidus temperature) were measured as 717 K, 760 K, 1095 K, and 1220 K for the Ti20Hf20Cu20Ni20Be20 HE-BMG, and 684 K, 739 K, 1066 K, and 1218 K for the Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 HE-BMG, respectively. These data were listed in Table 1.
The stress strain curves of Ø2 × 4 mm Ti20Hf20Cu20Ni20Be20 and Ø1.5 × 3 mm Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 HE-BMG samples in uniaxial compression test were shown in Figure 3. The fracture strength σb was 2425 MPa for Ti20Hf20Cu20Ni20Be20 HE-BMG, the yield strength σ0.2, fracture strength σb and plasticity εp were 2330 MPa, 2450 MPa and 0.5% for Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 HE-BMG, respectively, which were also listed in Table 1. The specimens fractured in a shear mode. It is interesting to note that both Ti20Hf20Cu20Ni20Be20 and Ti20Zr20Cu20Ni20Be20 quinary HE-BMGs fractured without any plasticity [24], while Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 and Ti16.7Zr16.7Hf16.7Cu16.7Ni16.7Be16.7 senary HE-BMGs exhibited a compressive plasticity of about 0.5%, as well as serration behavior [25]. The reason of this difference remains unclear.

4. Discussion

4.1. Glass Forming Ability (GFA) of High Entropy Alloys by Element Addition/Substitution

The parameters of supercooled liquid region ΔT (= TxTg), reduced glass transition temperature Trg (= Tg/Tl), and γ parameter (= Tx/(Tg + Tl)) are calculated as 43 K, 0.588, and 0.392 for Ti20Hf20Cu20Ni20Be20, while 55 K, 0.562, and 0.388 for Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7, respectively. Compared with Ti20Zr20Cu20Ni20Be20 alloy (3 mm), the critical diameter of Ti20Hf20Cu20Ni20Be20, alloy (2 mm) and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 (1.5 mm) both decreased. It is noticed that by substitution Zr with Hf, although Trg remains the same, ΔT and γ decreased; by addition of Nb as the sixth element, both Trg and γ decreased, although ΔT increased [24]. It implies that the parameter γ is better than Trg and ΔT in judging the GFA in these high-entropy glassy alloys; meanwhile high entropy is not always beneficial to the GFA of the HEAs. The substitution of element Hf and the addition of Nb brings the liquidus temperature Tl higher than that of Ti20Zr20Cu20Ni20Be20 alloy [24]. As a result, the GFA of the HEA was slightly deteriorated. On the other hand, by the addition of Hf as the sixth element, liquidus temperature Tl was lowered down. Therefore, the GFA of the Ti16.7Zr16.7Hf16.7Cu16.7Ni16.7Be16.7 was greatly improved as compared with Ti20Zr20Cu20Ni20Be20 alloy [25]. These results indicate that lowering down liquidus temperature would be helpful for enhancing the GFA.

4.2. Atomic Radius Characteristics of HE-BMG

The atomic size distribution characteristics of existing HE-BMGs were shown in Table 2. Based on the atomic radius of constituent elements, they were divided into five categories, namely super large atom (r > 0.165 nm), large atom (r ≈ 0.16 nm), medium atom (r ≈ 0.14 nm), small atom (r ≈ 0.12 nm), and ultra-small atom (r < 0.12 nm). It is noticed that most HE-BMGs were comprised of 3 to 4 categories, except for those containing nonmetal element such as Si, P, B, C, etc [29,33,36]. In high entropy alloys, larger atomic radius difference leads to larger lattice distortion. In case that lattice distortion exceeds some degree, the lattice collapse and amorphous structure formed accordingly. This is in agreement with Zhang’s work [13].

4.3. Assessing Degree of Lattice Distortion in High Entropy Alloys by Parameter δ′

The phase formation rule in HEA is of great importance both scientifically and technologically. The formed phase(s) in HEAs (solid solution, intermetallics and amorphous phase) at certain conditions (alloy composition, preparation method, service environment, etc.) remain unknown for most HEAs [10,13,15,18,42]. Many researchers proposed various criteria to solve this problem, such as the δ-ΔHmix diagram proposed by Zhang et al. [13], VEC criteria proposed by Guo et al. [15], electronegativity mismatch Dc proposed by Toda-Caraballo et al. [43], etc. Lattice distortion is a crucial factor in HEAs, and it is also very important in determining phase formation. However, the relationship between lattice distortion and phase formation is still not clear. Much research has been devoted to characterizing the degree of lattice distortion, and to further illustrate its correlation with phase formation, such as the γ parameter proposed by Wang et al. [44], the α2 parameter proposed by Wang et al. [45], etc. However, it is far from clearly understanding. Further investigation is still required.
It is noticed from Table 2 that in most HE-BMGs, the atomic radius of the constituent elements atomic sizes distribute in a wide range; while for many solid solution forming HEAs, atomic sizes are more concentrated (especially for CuCoCrNiFe [10] and Cantor alloy [11], they both possess FCC structure, meanwhile atomic size difference of the constituent elements are very small). However, this is a qualitative description, and it is somehow ambiguous. As a result, a quantitative exemplification is needed.
Based on Table 2, here we propose a new parameter δ′ to assess the degree of lattice distortion in HEAs. Supposing that a HEA contains N elements, the atomic fractions are c1, c2 …… cN, respectively, and the atomic radii are r1, r2 …… rN (r1 < r2 < …… < rN), respectively (data from ref. [46]). Then, the average atomic size is defined as r ¯ :
r ¯ = 1 N c i r i
The lattice distortion parameter δ′ is defined as
δ = 100   1 N 1 c i + 1 + c i 2   r i + 1 r i r ¯
In particular, for equal atomic alloy, δ′ is given as
δ = 100 N r N r 1 r ¯
According to Formula (2), lattice distortion parameter δ′ for some typical HEAs were calculated and listed in Table 3. For clarity, the relationship between atomic size distribution, lattice distortion parameter δ′, and phase selection is demonstrated in Figure 4. It is noticed that δ‘ is closely related to phase selection in HEAs: when atomic size difference is relatively small, δ′ is also small (δ′ < 2.2), FCC solid solution would be formed; when the atomic size difference became larger, δ′ increased, FCC + BCC solid solution would tend to form as 2.2 < δ′ < 2.9; with even larger δ′ (2.9 < δ′ < 4.9), BCC solid solution would be formed; amorphous phase would be formed as δ′ exceeds 4.9.
The parameter δ′ can be understood from the point of view of dense atomic packing. Since it is correlated with the degree of lattice distortion, when δ′ was small, dense random packing FCC phase formed (its density is about 74% for monolic element); with the increase of lattice distortion, δ′ became larger, looser BCC phase (density of about 68% for monolic element) appeared, and its concentration increased accordingly; when lattice distortion became even serious, lattice collapse and amorphous phase would form eventually. Then, adjusting the lattice distortion or the parameter δ′ of HEAs, such as by similar element substitution or addition, could be helpful in designing high-entropy metallic glasses.
The new parameter δ′ we proposed here is somewhat similar with the δ parameter proposed by Zhang et al. [13]; it is also affected by the number, type, and concentration of the elements, while its value is smaller than δ, as is demonstrated in Table 3. As compared with δ, δ′ is more sensitive to addition/substitution of an ultra large/small atom. Taking alloy 11 and 12 in Table 3 for example, it can be seen that by substituting Hf element with much smaller Be element in the Ti-Zr-Hf-Cu-Ni HEA, δ increased from 10.324 to 12.514, the growth rate is 21%; while δ′ increased from 4.977 to 7.065, the growth rate is 42%, much larger than that of δ. It indicates that the new parameter δ′ is more sensitive than δ in certain circumstance.
Additionally, it is noticed from Formula (3), for equiatomic high entropy alloys, as the number of elements N increased, δ′ decreased and lattice distortion is mitigated accordingly. As a result, it is not beneficial for amorphous phase formation, especially for N > 10. This is in consistent with Cantor’s result that an alloy with 16 to 20 elements in equiatomic concentration does not form amorphous phase [11].

5. Conclusions

In this paper, two new high entropy bulk metallic glasses (HE-BMGs) have been successfully fabricated using copper mold casting method, namely Ti20Hf20Cu20Ni20Be20 with a critical diameter of 2 mm and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 with a critical diameter of 1.5 mm. These two HE-BMGs exhibit high fracture strength over 2300 MPa. The glass forming ability and atomic size distribution characteristics of the HE-BMGs are discussed, and it is found that atomic radius spans over a wide range in HE-BMGs. Moreover, we propose a new parameter δ′ to assess the degree of lattice distortion in high entropy alloys (HEAs). It emphasizes the difference between atoms with adjacent atomic size, and it is closely related to phase selection in HEAs. When δ′ is relatively small (δ′ < 2.2), FCC solid solution formed; when 2.2 < δ′ < 2.9, FCC + BCC phases formed; when 2.9 < δ′ < 4.9, BCC phase formed; while δ′ > 4.9, amorphous phase would be formed. This new parameter δ′ is beneficial for understanding lattice distortion and phase selection in HEAs. The present work suggests that through adjusting the parameter δ′ by similar element substitution/addition, that is, adjusting the lattice distortion, is an effective way for designing high entropy bulk glassy alloy.

Author Contributions

Conceptualization, H.D.; Data curation, H.X.; Formal analysis, H.L.; Funding acquisition, H.D. and K.Y.; Investigation, H.B.; Methodology, H.B.; Project administration, K.Y.; Software, H.X.; Validation, H.X.; Visualization, H.L.; Writing–Original draft, H.D.; Writing–review & editing, H.L., H.B. and K.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Basic Research and Development Program (Grant No. 2016YFB00300500), National Natural Science Foundation of China (Grant Nos. 51571127 and 51871129). and Youth Fund of Jiangsu Natural Science Foundation (Grant No. BK20190979).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the plots within this paper and other findings of this study are available from the corresponding authors upon reasonable request.

Conflicts of Interest

The authors declare no competing interests.We declare that we have no financial and personal relationships with other people or organizations that can inappropriately influence our work.

References

  1. Greer, A.L. Metallic glasses. Science 1995, 267, 1947–1953. [Google Scholar] [CrossRef] [PubMed]
  2. Johnson, W.L. Bulk glass-forming metallic alloys: Science and technology. MRS Bull. 1999, 24, 42–56. [Google Scholar] [CrossRef]
  3. Inoue, A. Stabilization of metallic supercooled liquid and bulk amorphous alloys. Acta Mater. 2000, 48, 279–306. [Google Scholar] [CrossRef]
  4. Wang, W.H.; Dong, C.; Shek, C.H. Bulk metallic glasses. Mater. Sci. Eng. R 2004, 44, 45–89. [Google Scholar] [CrossRef]
  5. Inoue, A.; Takeuchi, A. Recent development and applications of bulk glassy alloys. Int. J. Appl. Glass Sci. 2010, 1, 273–295. [Google Scholar] [CrossRef]
  6. Wang, W.H. The elastic properties, elastic models and elastic perspectives of metallic glasses. Prog. Mater. Sci. 2012, 57, 487–656. [Google Scholar] [CrossRef]
  7. Greer, A.L. Metallic glasses. Phys. Metall. 2014, 35, 305–385. [Google Scholar] [CrossRef]
  8. Li, H.X.; Lu, Z.C.; Wang, S.L.; Wu, Y.; Lu, Z.P. Fe-based bulk metallic glasses: Glass formation, fabrication, properties and applications. Prog. Mater. Sci. 2019, 103, 235–318. [Google Scholar] [CrossRef]
  9. Yang, C.; Zhang, C.; Chen, Z.J.; Li, Y.; Yan, W.Y.; Yu, H.B.; Liu, L. Three-dimensional hierarchical porous structures of metallic glass/copper composite catalysts by 3D printing for efficient wastewater treatments. ACS Appl. Mater. Inter. 2021, 13, 7227–7237. [Google Scholar] [CrossRef]
  10. 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 principle elements: Novel alloy design concepts and outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  11. Cantor, B.; Chang, I.; Knight, P.; Vincent, A. Microstructural development in equiatomic multicomponent alloys. Mater. Sci. Eng. A 2004, 375–377, 213–218. [Google Scholar] [CrossRef]
  12. Zhou, Y.J.; Zhang, Y.; Wang, Y.L.; Chen, G.L. Solid solution alloys of AlCoCrFeNiTix with excellent room-temperature mechanical properties. Appl. Phys. Lett. 2007, 90, 181904. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Zhou, Y.J.; Lin, J.P.; Chen, G.L.; Liaw, P.K. Solid-solution phase formation rules for multi-component alloys. Adv. Eng. Mater. 2008, 10, 534–538. [Google Scholar] [CrossRef]
  14. 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]
  15. Guo, S.; Ng, C.; Lu, J.; Liu, C.T. Effect of valence electron concentration on stability of fcc or bcc phase in high entropy alloys. J. Appl. Phys. 2011, 109, 103505. [Google Scholar] [CrossRef] [Green Version]
  16. 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 2014, 345, 1153–1158. [Google Scholar] [CrossRef] [Green Version]
  17. Lu, Y.P.; Dong, Y.; Guo, S.; Jiang, L.; Kang, H.J.; Wang, T.M.; Wen, B.; Wang, Z.J.; Jie, J.C.; Cao, Z.Q.; et al. A promising new class of high-temperature alloys: Eutectic high-entropy alloys. Sci. Rep. 2014, 4, 6200. [Google Scholar] [CrossRef]
  18. Ye, Y.F.; Wang, Q.; Lu, J.; Liu, C.T.; Yang, Y. Design of high entropy alloys: A single-parameter thermodynamic rule. Scr. Mater. 2015, 104, 53–55. [Google Scholar] [CrossRef]
  19. Ma, L.Q.; Wang, L.M.; Zhang, T.; Inoue, A. Bulk Glass Formation of Ti–Zr–Hf–Cu–M (M = Fe,Co,Ni) Alloys. Mater. Trans. 2002, 43, 277–280. [Google Scholar] [CrossRef] [Green Version]
  20. Zhao, K.; Xia, X.X.; Bai, H.Y.; Zhao, D.Q.; Wang, W.H. Room temperature homogeneous flow in a bulk metallic glass with low glass transition temperature. Appl. Phys. Lett. 2011, 98, 141913. [Google Scholar] [CrossRef]
  21. Gao, X.Q.; Zhao, K.; Ke, H.B.; Ding, D.W.; Wang, W.H.; Bai, H.Y. High mixing entropy bulk metallic glasses. J. Non-Cryst. Solids 2011, 357, 3557–3560. [Google Scholar] [CrossRef]
  22. Takeuchi, A.; Chen, N.; Wada, T.; Yokoyama, Y.; Kato, H.; Inoue, A.; Yeh, J.W. Pd20Pt20Cu20Ni20P20 high-entropy alloy as a bulk metallic glass in the centimeter. Intermetallics 2011, 19, 1546–1554. [Google Scholar] [CrossRef]
  23. Li, H.F.; Xie, X.H.; Zhao, K.; Wang, Y.B.; Zheng, Y.F.; Wang, W.H.; Qin, L. In vitro and in vivo studies on biodegradable CaMgZnSrYb high-entropy bulk metallic glass. Acta Biomater. 2013, 9, 8561–8573. [Google Scholar] [CrossRef] [PubMed]
  24. Ding, H.Y.; Yao, K.F. High entropy Ti20Zr20Cu20Ni20Be20 bulk metallic glass. J. Non-Cryst. Solids 2013, 364, 9–12. [Google Scholar] [CrossRef]
  25. Ding, H.Y.; Shao, Y.; Gong, P.; Li, J.F.; Yao, K.F. A senary TiZrHfCuNiBe high entropy bulk metallic glass with large glass-forming ability. Mater. Lett. 2014, 125, 151–153. [Google Scholar] [CrossRef]
  26. Zhao, S.F.; Yang, G.N.; Ding, H.Y.; Yao, K.F. A quinary Ti-Zr-Hf-Be-Cu high entropy bulk metallic glass with a critical size of 12 mm. Intermetallics 2015, 61, 47–50. [Google Scholar] [CrossRef]
  27. Zhao, S.F.; Shao, Y.; Liu, X.; Chen, N.; Ding, H.Y.; Yao, K.F. Pseudo-quinary Ti20Zr20Hf20Be20(Cu20−xNix) high entropy bulk metallic glasses with large glass forming ability. Mater. Des. 2015, 87, 625–631. [Google Scholar] [CrossRef]
  28. Huo, J.T.; Huo, L.S.; Men, H.; Wang, X.M.; Inoue, A.; Wang, J.Q.; Chang, C.T.; Li, R.W. The magnetocaloric effect of Gd-Tb-Dy-Al-M (M= Fe, Co and Ni) high-entropy bulk metallic glasses. Intermetallics 2015, 58, 31–35. [Google Scholar] [CrossRef]
  29. Qi, T.L.; Li, Y.H.; Takeuchi, A.; Xie, G.Q.; Miao, H.T.; Zhang, W. Soft magnetic Fe25Co25Ni25(B, Si)25 high entropy bulk metallic glasses. Intermetallics 2015, 66, 8–12. [Google Scholar] [CrossRef]
  30. Chen, C.; Pang, S.J.; Cheng, Y.Y.; Zhang, T. A centimeter-size Zr40Hf10Ti4Y1Al10Cu25Ni7Co2Fe1 bulk metallic glass with high mixing entropy designed by multi-substitution. J. Non-Cryst. Solids 2015, 410, 39–42. [Google Scholar] [CrossRef]
  31. Kim, J.; Oh, H.S.; Kim, J.; Ryu, C.W.; Lee, G.W.; Chang, H.J.; Park, E.S. Utilization of high entropy alloy characteristics in Er-Gd-Y-Al-Co high entropy bulk metallic glass. Acta Mater. 2018, 155, 350–361. [Google Scholar] [CrossRef]
  32. Li, J.; Xue, L.; Yang, W.M.; Yuan, C.C.; Huo, J.T.; Shen, B.L. Distinct spin glass behavior and excellent magnetocaloric effect in Er20Dy20Co20Al20RE20 (RE = Gd, Tb and Tm) high-entropy bulk metallic glasses. Intermetallics 2018, 96, 90–93. [Google Scholar] [CrossRef]
  33. Xu, Y.Q.; Li, Y.H.; Zhu, Z.W.; Zhang, W. Formation and properties of Fe25Co25Ni25(P, C, B, Si)25 high-entropy bulk metallic glasses. J. Non-Cryst. Solids 2018, 487, 60–64. [Google Scholar] [CrossRef]
  34. Wu, L.; Zhao, Y.; Li, J.J.; Wu, J.L.; Zhang, B. Correlation between mechanical and thermodynamic properties for La–Ce–Ni–Cu–Al high-entropy metallic glasses. J. Iron Steel Res. Int. 2018, 25, 658–665. [Google Scholar] [CrossRef]
  35. Wu, K.N.; Liu, C.; Li, Q.; Huo, J.T.; Li, M.C.; Chang, C.T.; Sun, Y.F. Magnetocaloric effect of Fe25Co25Ni25Mo5P10B10 high-entropy bulk metallic glass. J. Magn. Magn. Mater. 2019, 489, 165404. [Google Scholar] [CrossRef]
  36. Li, C.Z.; Li, Q.; Li, M.C.; Chang, C.T.; Li, H.X.; Dong, Y.Q.; Sun, Y.F. New ferromagnetic (Fe1/3Co1/3Ni1/3)80(P1/2B1/2)20 high entropy bulk metallic glass with superior magnetic and mechanical properties. J. Alloys Compd. 2019, 791, 947–951. [Google Scholar] [CrossRef]
  37. Wada, T.; Jiang, J.; Yubuta, K.; Kato, H.; Takeuchi, A. Septenary Zr–Hf–Ti–Al–Co–Ni–Cu high-entropy bulk metallic glasses with centimeter-scale glass-forming ability. Materialia 2019, 7, 100372. [Google Scholar] [CrossRef]
  38. Pang, C.M.; Chen, L.; Xu, H.; Guo, W.; Lv, Z.W.; Huo, J.T.; Cai, M.J.; Shen, B.L.; Wang, X.L.; Yuan, C.C. Effect of Dy, Ho, and Er substitution on the magnetocaloric properties of Gd-Co-Al-Y high entropy bulk metallic glasses. J. Alloys Compd. 2020, 827, 154101. [Google Scholar] [CrossRef]
  39. Li, Y.H.; Wang, S.W.; Wang, X.W.; Yin, M.L.; Zhang, W. New FeNiCrMo(P, C, B) high-entropy bulk metallic glasses with unusual thermal stability and corrosion resistance. J. Mater. Sci. Technol. 2020, 43, 32–39. [Google Scholar] [CrossRef]
  40. Shao, L.L.; Wang, Q.Q.; Xue, L.; Zhu, M.Y.; Wang, A.D.; Luan, J.H.; Yin, K.B.; Luo, Q.; Zeng, Q.S.; Sun, L.T.; et al. Effects of minor Si addition on structural heterogeneity and glass formation of GdDyErCoAl high-entropy bulk metallic glass. J. Mater. Res. Technol. 2021, 11, 378–391. [Google Scholar] [CrossRef]
  41. Jalali, A.; Malekan, M.; Park, E.S.; Rashidi, R.; Bahmani, A.; Yoo, G.H. Thermal behavior of newly developed Zr33Hf8Ti6Cu32Ni10Co5Al6 high-entropy bulk metallic glass. J. Alloys Compd. 2021, 892, 162220. [Google Scholar] [CrossRef]
  42. Luan, H.W.; Shao, Y.; Li, J.F.; Mao, W.L.; Han, Z.D.; Shao, C.L.; Yao, K.F. Phase stabilities of high entropy alloys. Scr. Mater. 2020, 179, 40–44. [Google Scholar] [CrossRef]
  43. Toda-Caraballo, I.; Rivera-Diaz-del-Castillo, P.E.J. A criterion for the formation of high entropy alloys based on lattice distortion. Intermetallics 2016, 71, 76–87. [Google Scholar] [CrossRef]
  44. Wang, Z.J.; Huang, Y.H.; Yang, Y.; Wang, J.C.; Liu, C.T. Atomic-size effect and solid solubility of multicomponent alloys. Scr. Mater. 2015, 94, 28–31. [Google Scholar] [CrossRef]
  45. Wang, Z.J.; Qiu, W.F.; Yang, Y.; Liu, C.T. Atomic-size and lattice-distortion effects in newly developed high-entropy alloys with multiple principal elements. Intermetallics 2015, 64, 63–69. [Google Scholar] [CrossRef]
  46. Senkov, O.N.; Miracle, D.B. Effect of the atomic size distribution on glass forming ability of amorphous metallic alloys. Mater. Res. Bull. 2001, 36, 2183–2198. [Google Scholar] [CrossRef]
Figure 1. XRD spectra of the Ø2 mm Ti20Hf20Cu20Ni20Be20 rod sample and Ø1.5 mm Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 rod sample.
Figure 1. XRD spectra of the Ø2 mm Ti20Hf20Cu20Ni20Be20 rod sample and Ø1.5 mm Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 rod sample.
Materials 15 01669 g001
Figure 2. DSC curves of the Ti20Hf20Cu20Ni20Be20 and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 HE-BMGs.
Figure 2. DSC curves of the Ti20Hf20Cu20Ni20Be20 and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 HE-BMGs.
Materials 15 01669 g002
Figure 3. Stress strain curves of the Ti20Hf20Cu20Ni20Be20 and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 HE-BMGs.
Figure 3. Stress strain curves of the Ti20Hf20Cu20Ni20Be20 and Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 HE-BMGs.
Materials 15 01669 g003
Figure 4. Correlation between atomic size distribution, lattice distortion degree parameter δ‘ and phase selection in some typical HEAs [10,11,12].
Figure 4. Correlation between atomic size distribution, lattice distortion degree parameter δ‘ and phase selection in some typical HEAs [10,11,12].
Materials 15 01669 g004
Table 1. Thermal and mechanical properties of some HE-BMGs.
Table 1. Thermal and mechanical properties of some HE-BMGs.
CompositionTg
(K)
Tx
(K)
Tm
(K)
Tl
(K)
σ0.2
(MPa)
σb
(MPa)
εp
(%)
Year
Ti20Zr20Cu20Ni20Be2068372910761161-231502013 [24]
Ti16.7Zr16.7Hf16.7Cu16.7Ni16.7Be16.768175110191100194320640.62014 [25]
Ti20Hf20Cu20Ni20Be2071776010951220-24250This work
Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.768473910661218233024500.5This work
Table 2. Atomic size distribution characteristics of existing HE-BMGs.
Table 2. Atomic size distribution characteristics of existing HE-BMGs.
CompositionSuper Large Atom
r > 0.165 nm
Large Atom
r ≈ 0.16 nm
Medium Atom
r ≈ 0.14 nm
Small Atom
r ≈ 0.12 nm
Ultra Small Atom
r < 0.12 nm
Year
Ti20Zr20Hf20Cu20Ni20 Zr, HfTiCu, Ni 2002 [19]
Sr20Ca20Yb20Mg20Zn20Sr, Ca, YbMgZn 2011 [20,23]
Er20Tb20Dy20Ni20Al20Tb, Dy, Er AlNi 2011 [21]
Pd20Pt20Cu20Ni20P20 Pt, PdCu, NiP2011 [22]
Ti20Zr20Cu20Ni20Be20 ZrTiCu, NiBe2013 [24]
Ti16.7Zr16.7Hf16.7Cu16.7Ni16.7Be16.7 Zr, HfTiCu, NiBe2014 [25]
Ti20Zr20Hf20(Cu20−xNix)Be20 Zr, HfTiCu, NiBe2015 [26,27]
Ho20Er20Co20Al20Dy20Dy, Ho, Er AlCo 2015 [28]
Fe25Co25Ni25(B, Si)25 Co, Ni, FeSi, B2015 [29]
Zr40Hf10Ti4Y1Al10Cu25Ni7Co2Fe1YZr, HfTi, AlCu, Co, Ni, Fe 2015 [30]
Er18Gd18Y20Al24Co20Y, Gd, Er AlCo 2018 [31]
Er20Dy20Co20Al20RE20 (RE = Gd, Tb, Tm)Gd/Tb, Dy, ErTmAlCo 2018 [32]
Fe25Co25Ni25(P0.4C0.2B0.2Si0.2)25 Co, Ni, FeSi, P, B, C2018 [33]
La25–35Ce25–35Ni5–15Cu5–15Al20La, Ce AlCu, Ni 2018 [34]
Fe25Co25Ni25Mo5P10B10 MoCo, Ni, FeP, B2019 [35]
(Fe1/3Co1/3Ni1/3)80(P1/2B1/2)20 Co, Ni, FeP, B2019 [36]
Zr35Hf17.5Ti5.5Al12.5Co7.5Ni12Cu10 Zr, HfTi, AlCu, Co, Ni 2019 [37]
Gd25Co25Al25Y15RE10 (RE = Dy, Ho, Er)Y, Gd, (Dy, Ho, Er) AlCo 2020 [38]
Fe20–35Ni20Cr20–30Mo5–15(P0.6C0.2B0.2)20 MoCr, Ni, FeP, B, C2020 [39]
(Gd0.2Dy0.2Er0.2Co0.2Al0.2)99.5Si0.5Gd, Dy, Er AlCoSi2021 [40]
Zr33Hf8Ti6Cu32Ni10Co5Al6 Zr, HfTi, AlCu, Co, Ni 2021 [41]
Ti20Hf20Cu20Ni20Be20 HfTiCu, NiBeThis work
Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 ZrTi, NbCu, NiBeThis work
Table 3. Correlation between atomic size distribution, lattice distortion and phase selection in some typical HEAs.
Table 3. Correlation between atomic size distribution, lattice distortion and phase selection in some typical HEAs.
No.Compositionr > 0.165 nmr ≈ 0.16 nmr ≈ 0.14 nmr ≈ 0.12 nmr < 0.12 nm r ¯ δ [13]δVEC [15]Phase
1CrMnFeCoNi MnCo, Cr, Ni, Fe 1.267443.2671.7178FCC [11]
2CuCoCrNiFe Cu, Co, Cr, Ni, Fe 1.253041.0310.5878.8FCC [10]
3Al0.3CuCoCrFeNi AlCu, Co, Cr, Ni, Fe 1.263153.4162.0428.472FCC [10]
4A0.5lCuCoCrFeNi AlCu, Co, Cr, Ni, Fe 1.269284.1612.1788.273FCC [10]
5Al0.8CuCoCrFeNi AlCu, Co, Cr, Ni, Fe 1.277684.9122.3638FCC + BCC [10]
6AlCuCoCrFeNi AlCu, Co, Cr, Ni, Fe 1.282815.2712.4757.833FCC + BCC [10]
7Al2.5CuCoCrFeNi AlCu, Co, Cr, Ni, Fe 1.312596.4663.1066.867FCC + BCC [10]
8Al2.8CuCoCrFeNi AlCu, Co, Cr, Ni, Fe 1.317176.5543.2016.718BCC [10]
9Al3.0CuCoCrFeNi AlCu, Co, Cr, Ni, Fe 1.320046.5983.2596.625BCC [10]
10AlCoCrFeNi AlCo, Cr, Ni, Fe 1.283785.7672.9687.2BCC [12]
11Ti20Zr20Hf20Cu20Ni20 Zr, HfTiCu, Ni 1.4330810.3244.977-BMG [19]
12Ti20Zr20Cu20Ni20Be20 ZrTiCu, NiBe1.3431812.5147.065-BMG [24]
13Ti16.7Zr16.7Hf16.7Cu16.7Ni16.7Be16.7 Zr, HfTiCu, NiBe1.3822312.7735.721-BMG [25]
14Ti20Hf20Cu20Ni20Be20 HfTiCu, NiBe1.3381811.9936.718-BMG (this work)
15Ti16.7Zr16.7Nb16.7Cu16.7Ni16.7Be16.7 ZrTi, NbCu, NiBe1.3549511.5465.826-BMG (this work)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ding, H.; Luan, H.; Bu, H.; Xu, H.; Yao, K. Designing High Entropy Bulk Metallic Glass (HE-BMG) by Similar Element Substitution/Addition. Materials 2022, 15, 1669. https://doi.org/10.3390/ma15051669

AMA Style

Ding H, Luan H, Bu H, Xu H, Yao K. Designing High Entropy Bulk Metallic Glass (HE-BMG) by Similar Element Substitution/Addition. Materials. 2022; 15(5):1669. https://doi.org/10.3390/ma15051669

Chicago/Turabian Style

Ding, Hongyu, Hengwei Luan, Hengtong Bu, Hongjie Xu, and Kefu Yao. 2022. "Designing High Entropy Bulk Metallic Glass (HE-BMG) by Similar Element Substitution/Addition" Materials 15, no. 5: 1669. https://doi.org/10.3390/ma15051669

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