Open access peer-reviewed chapter - ONLINE FIRST

High-Entropy Superconducting Materials

Written By

Pablo Cayado and Jens Hänisch

Submitted: 22 December 2023 Reviewed: 01 February 2024 Published: 16 July 2024

DOI: 10.5772/intechopen.1005291

High Entropy Alloys - Composition and Microstructure Design IntechOpen
High Entropy Alloys - Composition and Microstructure Design Edited by Yu Yin

From the Edited Volume

High Entropy Alloys - Composition and Microstructure Design [Working Title]

Dr. Yu Yin, Prof. Han Huang, Dr. Mingxing Zhang and Prof. Libo Zhou

Chapter metrics overview

23 Chapter Downloads

View Full Metrics

Abstract

High-entropy alloys are a rapidly evolving field of materials research and have gained increasing attention in recent years. Characterized by their multicomponent compositions, they were originally created by mixing a multitude of metallic elements and expanded the idea of traditional alloy design, opening new paths for materials research. In particular, the discoveries of superconductivity in some of these alloys since 2014 provided a new impetus for exploring novel superconducting phenomena and materials. In fact, the concept of increasing entropy in superconducting compounds, firstly restricted to alloys or intermetallics, was soon extended to other types of superconductors enriching drastically the research in the field. The high-entropy superconductors are nowadays a matter of intense study. This chapter provides a comprehensive review of the most recent and relevant research on the different types of high-entropy superconductors. The current state of research, synthesis methods, and characterization techniques are included. This information may serve as a reference for future research on this topic and inspire further exploration and innovation in the application of these materials.

Keywords

  • superconductor
  • high-entropy alloy
  • high-entropy ceramic
  • high-entropy oxide
  • REBCO

1. Introduction

High-entropy alloys (HEAs) are a category of materials that has attracted attention in recent times in different fields of materials research. This new paradigm for alloy design, proposed around two decades ago [1, 2], consists of multiple species of atoms randomly intermixed, resulting in a notable level of disorder, specifically in high configurational entropy. This severe disorder appears due to the mixing of several atom species in considerable proportions on specific lattice sites.

The definition of a HEA, or, in other words, the difference between HEA and medium-entropy alloys or even the usual low-entropy alloys, is based on the number of elements present in the alloy and their proportions. An alloy is usually called HEA if it contains a mixture of at least five elements, typically in proportions between 5% and 35% [1], although some authors claim that mixtures of four elements may also be considered HEAs [3]. The entropy of such mixtures can be evaluated as:

Smix=Ri=1nxilnxiE1

where xi is the molar fraction of component i, n is the total number of components, and R = 8.314 J/mol K is the gas constant. An equimolar system with five components has a ΔSmix value of 1.609R, which is also often used as an equivalent, stricter criterion for HEA [4]. In the following, we keep the definition of at least five components taking part in mixing while softening occasionally the criteria of composition and ΔSmix.

The change in entropy lowers the Gibbs energy according to:

Gmix=HmixTSmixE2

with Hmix being the mixing enthalpy, which allows the stabilization of disordered compounds into certain singular crystal structures with substantial chemical substitutional disorder, rather than leading to the creation of multiple crystalline phases.

Among all the novel properties that emerged in HEAs, the appearance of superconductivity in some of them is one of the most interesting ones. This was seen for the first time in 2014 in the system Ta34Nb33Hf8Zr14Ti11 [5] with a rather high critical temperature, Tc, of ∼7.3 K [5]. This finding was the starting point of intense research into the topic as the combination of superconductivity and superior mechanical properties that HEAs typically gives them some application potential to be evaluated and utilized. Due to the large degree of disorder and hence short electron mean free paths and therefore coherence lengths, all HEA superconductors are type II. Often, their Bc2 is measured resistively, rarely also their Bc1 magnetically. Critical current densities are usually not reported with the exception of the REBa2Cu3O7−x (REBCO, RE rare earth elements) compounds. Most HEAs are fabricated by arc melting. In recent years, further preparation methods were investigated, such as solid-state reaction (SSR), spark-plasma sintering, hot-press sintering, and thin-film growth (sputtering, pulsed laser deposition, and chemical solution deposition).

Over the years, the concept of “high-entropy,” that is, incorporating multiple components within the same matrix, has been expanded to various types of superconductors like BiS2-based, Van der Waals, rock-salt-type, A15-type, and transition-metal-zirconide (TrZr2)-type superconductors. Most recently, this approach has been further extended to intermetallics as well as compounds containing non-metallic constituents like oxides, carbides, borides, nitrides, and sulfides. Particularly interesting in the field of superconductivity are the high-entropy oxides (HEO), in which the well-known REBCO compounds, probably the superconductors with the largest potential from the applications point of view, are included. The research on HE REBCO has started only few years ago, but it is a topic with promising perspectives. There is considerable variety in stoichiometry notation in literature. We occasionally carefully modified it for better comparability. For the alloys, we adopted 100% notation and (near) equimolar mixes without further subscripts, that is, the number outside of the brackets represents the sum of the chemical compositions of the elements in parentheses, for instance, (TaNb)67 refers to Ta1/3Nb1/3. Parts in 6 or 7, we note as 1/6 and 1/7 to avoid odd numbers and false accuracy. For the HE compounds, we mostly used the stoichiometry notation of the parent compounds.

1.1 Categorizing

Considering the composition and the structural and physical characteristics of the different HEAs, Sun et al. proposed a first classification by dividing them into four different groups [4]. The group of type A includes HEA superconductors made of metals from the left side of the transition-metal region of the periodic table like the different stoichiometries of the Ta-Nb-Hf-Zr-Ti system and most of the related compounds coming from elemental substitutions such as Nb-Re-Hf-Zr-Ti or Hf-Nb-Ti-V-Zr. These type A HEAs crystallize in bcc lattices and show Tc values between 4.0 K and 9.2 K and a maximum Bc2(0) of ∼11.7 T. The type B HEA superconductors are mainly composed of early transition metals (groups 3 to 7 of the periodic table), but crystallize in rather complex lattices, which could be large bcc α-Mn-type lattices or smaller bcc and hcp lattices, depending on the stoichiometry. Compounds like (ZrNb)30(MoReRu)70, (HfTaWIr)70Re30, or (HfTaWPt)50Re50 belong to this group. The Tc in this class is usually lower than for type A HEAs with values between 1.9 K and 5.7 K, and the highest known Bc2(0) is ∼7.9 T. The third group, type C HEA superconductors, crystallizes in a CsCl-type structure and is composed of early transition metals combined with late transition metals, such as the Sc-Zr-Nb-Ta-Rh-Pd system. They show Tc values between 3.9 K and 9.3 K and Bc2(0) values up to ∼10.7 T. The fourth group, type D HEA superconductors, crystallizes in the hcp structure. Examples are Re56Nb11Ti11Zr11Hf11, (MoReRuRh)95Ti5, and Nb10+2xMo35−xRu35−xRh10Pd10. This last alloy shows the highest Tc of this class, with a value of 6.19 K for x = 2.5. With the appearance of more and more different types of superconducting high-entropy compounds, this classification was extended by naming them simply according to their crystal structure.

One parameter that may have a great influence on the stability of HEAs is the valence electron count (VEC), which is better expressed as the number of valence electrons per atom in the particular case of HEAs. The VEC reflects the total density of states at the Fermi level and plays a crucial role in determining the Tc of superconductors that follow the Bardeen-Cooper-Schriefer (BCS) theory. In fact, a robust correlation between VEC and Tc is evident in binary and ternary superconducting transition-metal alloys, commonly known as the Matthias rule [6]. In the case of HEAs, there is also a relationship between the VEC and Tc similar to the Matthias rule. In Figure 1, the dependence of Tc on the VEC is shown for different types of superconducting HEAs. The type A HEAs that crystallize in bcc structure appear in the VEC range 4.2–4.8, reaching a maximum Tc for a VEC of 4.7. In the case of type C HEAs, with CsCl-type structures, superconductivity emerges in the VEC range of 5.9–6.3, with a maximum Tc at a VEC of ∼5.9. These two types of HEAs have similar behavior to the simpler binary alloys with the characteristic two-peak character seen for the near VECs of 4.7 and 5.9. However, the type B HEAs with a more complex α-Mn structure do not follow this general trend. Although the bcc and CsCl crystal structures are closely related to the α-Mn structure, their atomic arrangements differ in a remarkable way, making their stabilities to be limited to narrow ranges of VEC. The type D HEAs, the ones that show hcp structure, show superconducting behavior for VEC values similar to the type C ones. Finally, it is possible to add to this analysis the more recent HEAs. The A15-type HEAs show similar behavior as the binary A15 superconductors that follow the Matthias rule, with VEC falling between 4.55 and 4.85. Also, the σ-phase HEAs resemble their relatives, the binary σ-phase superconductors whose VEC values extend from 6.07 to 6.84 with a Tc maximum at VEC ∼ 6.5 in accordance with the Matthias rule [6]. The VECs of the σ-phase HEAs are located at the electron-rich side of this region and are higher than those of most HEA superconductors. Note that the VEC ranges of the σ phase and the α-Mn type HEA superconductors are very close. These dependencies show that the VEC serves as a crucial tool for predicting potential HEA superconductors and can be useful for understanding the superconducting properties of the new systems.

Figure 1.

Dependence of the critical temperature Tc on the valence electron count (VEC) for different types of high-entropy superconductors. Data extracted from Refs. [4, 7, 8].

Below, the state of the art of each of these HE superconductor classes will be reported. Special attention will be paid to the superconducting properties of each compound and their prospects for further development.

Advertisement

2. Superconducting high-entropy alloys

2.1 The Ta-Nb-Hf-Zr-Ti system

Superconductivity in high-entropy alloys was first discovered by Koželj et al. in 2014 for a particular stoichiometry of the Ta-Nb-Hf-Zr-Ti system [5]. Investigations on this system had already started years prior due to its superior structural and mechanical properties [9]. Varying the original equimolar ratio of the elements led to the Ta34Nb33Hf8Zr14Ti11 compound that appeared to be superconducting below a critical temperature of ∼7.3 K. The original compound was a polycrystalline sample with grains in the range of 2–300 μm fabricated by arc melting. It crystallized in a highly distorted body-centered cubic (bcc) structure (a = 3.36 Å), Figure 2a, due to the difference in the atomic radii of the constituent elements. Apart from its ∼7.3 K Tc, it shows an upper critical field Bc2(0) ∼ 8.2 T, a lower critical field Bc1 ∼ 32 mT, and an energy gap opening of ∼2.2 meV. Moreover, first principle calculations showed that it could be classified as a strong electron-phonon coupling superconductor with a coupling constant λ = 1.16 [11].

Figure 2.

Crystal structures of several high-entropy superconductors. Reproduced with permission from [10].

This discovery soon caught the attention of several groups, who started to investigate this system, preparing the samples, as in the original work, by an arc melting method as summarized in Table 1. A stoichiometry series, studied a couple of years later by von Rohr et al., was [TaNb]100−x(HfZrTi)x, where x took 12 different values from 20 to 84 [13]. This study revealed Tc variations between 4.49 K and 7.92 K for the different stoichiometries, with the largest upper critical field of ∼11.67 T for the sample with x = 0.5. A particular stoichiometry of this compound, (TaNb)67(HfZrTi)33, was further studied by Guo et al. [14] at high pressures, revealing an exceptionally resilient zero-resistance superconductivity under pressures up to 190.6 GPa. In their study, the authors observed that initially, the Tc of this compound rose from 7.7 K at ambient pressure to 10 K at ∼60 GPa, followed by a slight, gradual decrease to 9 K at 190.6 GPa, which is close to the pressure in the outer core of the Earth. The authors proposed therefore that this compound is a potential candidate for new applications under extreme pressure conditions. Similar results confirming the robustness of the superconductivity in the Ta-Nb-Hf-Zr-Ti system were obtained by Vrtnik et al. [12], who prepared samples with different numbers of elements (4 and 5) as well as differing stoichiometry (equimolar and off-equimolar) by arc melting and subjected them to different thermal treatments. Their results show that the samples are superconducting independently of the thermal treatment and the resulting microstructure. In addition, Jasiewicz et al. [19] further investigated the pressure effects on the electronic structure, electron-phonon interaction, and superconductivity of (TaNb)67(HfZrTi)33 from 0 to 100 GPa. Their investigation confirmed the previously reported results, observing the same Tc trend with pressure. However, they could also prove that this evolution of Tc with pressure can be well explained in the framework of the classical electron-phonon mechanism. A certain variant also investigated was Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6 by Kim et al. [16]. With a Tc = 7.85 K, it shows a Bc1(0) ∼ 23 mT and a Bc2(0) ∼ 12.05 T, a higher value than the one found for the (TaNb)50(HfZrTi)50 commented before. This work was the first one to report about critical current density and pinning mechanisms in this type of systems. The authors reported a critical current density (Jc) of ∼10.7 kA/cm2 at 2 K in self-field and found that the pinning mechanism agreed with the double exponential model, which was first introduced for MgB2 by Horvat et al. [20], indicating that there were two types of pinning mechanism active. Kim et al. [17] investigated recently the effect of vacuum annealing at several temperatures and at different times on the superconducting and mechanical properties of this compound. Even though Tc and Bc2 slightly decreased upon annealing at medium temperatures (550°C), it had a positive effect on both Jc and Vickers hardness (HVIT). Record values of Jc(0) = 520 kA/cm2 and HVIT = 528 were reported, an increase of 1860% and 32%, respectively. The best samples of Refs. [16, 17], Table 1, investigated are (also) included in Table 2 for comparison of Jc. The already high value of Bc2(0) found in [16] was recently overcome in the Nb34Ti33Zr14Ta11Hf8 alloy with Bc2(0) ~ 12.2 T, although it has a lower Tc = 7.5 [18]. For a summary of the Ta-Nb-Hf-Zr-Ti HEAs, see Table 1.

Nominal compositionTc (K)Bc1(0) (mT)Bc2(0) (T)Reference
Ta34Nb33Hf8Zr14Ti117.3328.2[5]
Ta34Nb33Hf8Zr14Ti117.88.15[12]
[TaNb]70(ZrHfTi)308.036.67[13]
[TaNb]67(ZrHfTi)337.757.75[13]
[TaNb]67(ZrHfTi)337.7 (1 atm)[14]
[TaNb]67(ZrHfTi)3310 (60 GPa)[14]
[TaNb]67(ZrHfTi)339 (190.6 GPa)[14]
[TaNb]60(ZrHfTi)407.568.43[13]
[TaNb]50(ZrHfTi)506.4611.67[13]
[TaNb]16(ZrHfTi)844.529.02[13]
Ta20Nb21Hf20Zr20Ti196.910.45[12]
Ta20Nb20Hf20Zr20Ti207.12[15]
Ta22Nb24Hf21Zr23Ti106.2[12]
Ta1/6Nb2/6Hf1/6Zr1/6Ti1/67.852312.05[16]
Ta1/6Nb2/6Hf1/6Zr1/6Ti1/68.111.0[17]
Ta11Nb34Hf8Zr14Ti337.51512.2[18]

Table 1.

Selected properties of superconducting Ta-Nb-Hf-Zr-Ti HEAs, all prepared by arc melting.

Nominal compositionSynthesis methodTc (K)Bc2(0) (T)Jc (kA/cm2)Reference
Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6Arc melting7.8512.0510.7 (2 K, sf)[16]
Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6Arc melting8.111520 (4 K, sf)[17]
Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6SPS7.8010.5030.5 (2 K, 0.01 T)
73.2 (4 K, 0.01 T)
[21]
Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6Hot-press sintering7.90[22]
Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6PLD7.2812.101.8 (4.2 K, sf)[23]
(TaNb)12(HfZrTi)88Sputtering2.776.15[24]
(TaNb)24(HfZrTi)76Sputtering4.6110.44[24]
(TaNb)35(HfZrTi)65Sputtering5.6011.05[24]
(TaNb)46(HfZrTi)54Sputtering6.149.93[24]
(TaNb)57(HfZrTi)43Sputtering6.768.77[24]
(TaNb)67(HfZrTi)33Sputtering6.437.05[24]
(TaNb)79(HfZrTi)21Sputtering6.335.78[24]
(TaNb)87(HfZrTi)13Sputtering6.024.29[24]
(TaNb)96(HfZrTi)04Sputtering5.572.95[24]
(TaNb)70(HfZrTi)30 (600 nm)Sputtering5.315.80[25]
(TaNb)70(HfZrTi)30 (100 nm)Sputtering2.693.20[25]
(TaNbZrTi)90W10Sputtering6.20[26]
(TaNbZrTi)90V10Sputtering6.20[26]

Table 2.

Selected properties of superconducting Ta-Nb-Hf-Zr-Ti HEAs prepared by methods different from the conventional arc melting.

2.2 Substitutions in Ta-Nb-Hf-Zr-Ti

The natural evolution of the field led to the study of new compositions and mixtures of elements, first by only replacing one of the elements. These compounds, summarized in Table 3, crystallize in the bcc structure and belong to type A too. An alternative soon investigated by Marik et al. [27] was the equimolar Nb-Re-Hf-Zr-Ti HEA, the result of substituting Re for Ta in the original system. Prepared by arc melting too, it was among the first equimolar HEA superconductors. With Tc ∼ 5.3 K, it shows an upper critical field Bc2(0) of 8.88 T, a value close to the BCS weak coupling Pauli limit, in contrast to other compounds of the Ta-Nb-Hf-Zr-Ti system [17]. The same elements in different stoichiometry were employed by Motla et al. [28] in the synthesis of Nb60Re10Zr10Hf10Ti10. This variant shows a Tc = 5.7 K, and it was studied in detail by means of different electronic characterization techniques, including muon spin relaxation and rotation, finding electronic properties similar to conventional binary and ternary superconducting compounds, that is, s-wave symmetry of the order parameter and phonon-mediated pairing [18]. Ta was also substituted by Mo in the equimolar HfMoNbTiZr, resulting in Tc = 4.1 K and Bc2(0) = 5.55 T [30]. Re, on the other hand, was also used by Hattori et al. [29] to substitute not the Ta but the Zr, leading to the Ti-Hf-Nb-Ta-Re system. Different stoichiometries with VEC values ranging from 4.6 to 5.0 were investigated, finding Tc values between 3.25 K and 4.38 K and Bc2(0) values between 7.90 T and 5.91 T, with the maximum Tc occurring for a VEC of 4.7 and the maximum Bc2(0) for a VEC of 4.6 [27]. The Zr was also substituted, in this case by Mo by Zeng et al. [31], to prepare equimolar TiHfNbTaMo samples that showed a Tc = 3.42 K and Bc2(0) = 3.95 T [28]. The Zr was also replaced by V in the equimolar HfNbTaTiV HEA [34]. The detailed characterization determined Tc = 5.0 K, Bc1 = 19 mT and Bc2(0) = 6.63 T [30]. Another substitution, V for Ta, led to Hf21Nb25Ti15V15Zr24, which shows a Tc value of 5.3 K [33]. In the case of [TaNb]31(TiUHf)69, the substitution was U for Zr [35]. This was the first HEA containing an f-electron element, and it extended the HEA research to actinide elements, opening the possibility of leveraging HEAs as functional waste forms for a variety of radioisotopes. This system shows phonon-mediated superconductivity with Tc ∼ 3.2 K and Bc2(0) ∼ 6.4 T [31].

Nominal compositionTc (K)Bc1(0) (mT)Bc2(0) (T)Reference
Re20Nb20Hf20Zr20Ti205.3338.88[27]
Re10Nb60Hf10Zr10Ti105.77.956.31[28]
Ta5Nb25Hf25Re10Ti353.954.877.90[29]
Ta5Nb35Hf20Re10Ti304.386.606.94[29]
Ta15Nb35Hf15Re10Ti254.104.566.65[29]
Ta25Nb35Hf10Re10Ti203.627.155.85[29]
Ta35Nb35Hf5Re10Ti153.256.875.91[29]
Mo20Nb20Hf20Zr20Ti204.15.25.82[30]
Ta20Nb20Hf20Mo20Ti203.4222.83.95[31]
(TaNb)0.67(HfMoW)0.334.31.45[32]
V15Nb25Hf21Zr24Ti155.30[33]
Ta20Nb20Hf20V20Ti204.936.63[34]
[TaNb]31(HfUTi)693.206.40[35]
[{Sc33Cr67}Nb]67(HfZrTi)335.60[36]
[Ta{Sc33Cr67}]67(HfZrTi)334.40[36]
[TaNb]67({Sc67Cr33}ZrTi)337.50[36]
[TaNb]67(Hf{Sc67Cr33}Ti)337.40[36]
[TaNb]67(HfZr{Sc67Cr33})337.60[36]
[{Y33Mo67}Nb]67(HfZrTi)334.70[36]
[Ta{Y33Mo67}]67(HfZrTi)333.50[36]
[TaNb]67({Y67Mo33}ZrTi)337.60[36]
[TaNb]67(Hf{Y67Mo33}Ti)336.70[36]
[TaNb]67(HfZr{Y67Mo33})337.50[36]
[{Sc33Mo67}Nb]67(HfZrTi)334.40[36]
[Ta{Sc33Mo67}]67(HfZrTi)332.90[36]
[TaNb]67({Sc67Mo33}ZrTi)337.50[36]
[TaNb]67(Hf{Sc67Mo33}Ti)336.60[36]
[TaNb]67(HfZr{Sc67Mo33})337.50[36]
Ta20Nb20Fe20Zr20Ti206.87[15]
Ta20Nb20Ge20Zr20Ti208.46[15]
Ta1/6Nb1/6Hf1/6Zr1/6Ti1/6V1/65.09[15]
Ta1/6Nb1/6Ge1/6V1/6Zr1/6Ti1/68.40[15]
Ta1/6Nb1/6Si1/6V1/6Zr1/6Ti1/64.29[15]
Ta1/6Nb1/6Si1/6Ge1/6Zr1/6Ti1/67.40[15]
Ta20Nb20Hf20Sc20Ti206.6013.10[37]
Ta20Nb20Hf20Zr20Sc207.7012.40[37]
Ta20Nb20Sc20Zr20Ti207.9019.30[37]
Ta1/6Nb1/6Hf1/6Zr1/6Ti1/6Sc1/67.2014.10[37]
All5Nb24V5Zr26Ti405.36[38]
All5Nb14V5Zr41Ti354.73[38]
All5Nb24V5Zr31Ti355.45[38]
All5Nb34V5Zr21Ti356.47[38]
All5Nb44V5Zr11Ti356.75[38]
All5Nb14V5Zr31Ti454.96[38]
All5Nb34V5Zr31Ti256.18[38]
All5Nb44V5Zr31Ti156.98[38]

Table 3.

Selected properties of superconducting type A Ta-Nb-Hf-Zr-Ti-based HEAs with different substitutions (bold). All samples are prepared by arc melting.

Other elemental substitutions of two or more elements in the Ta-Nb-Hf-Zr-Ti system were studied by von Rohr et al. [36] with systematic isoelectronic replacements, using Mo-Y, Mo-Sc, and Cr-Sc mixtures and also adding Al to the initial [TaNb]67(HfZrTi)33 compound ([TaNb]67(HfZrTi)33Alx). This study concluded that Tc strongly depends on the elemental composition of the alloy. This change is clearly visible when substituting Nb and also Ta, observing a 60% decay in Tc of [Ta{Sc33Mo67}]67(HfZrTi)33. The replacement of the other three metals does not change Tc so drastically, which remains almost unchanged for the replacement of Hf and Ti and being ∼1 K lower for the replacement of Zr. In the case of the Al addition, an improvement of the crystallinity of the compounds was realized until x = 40, when the stabilization of a simple bcc lattice breaks down and the material starts to crystallize in the β-uranium structure, resulting in a suppression of superconductivity below 1.8 K [34]. Similar conclusions can be extracted from the work of Wu et al. [15], who added Fe, Ge, Hf, Si, and/or V instead of Hf to the original NbTaTiZr-based matrix. The analysis of the physical properties of the prepared compounds, including up to seven elements, revealed that the superconducting behavior varies with the composition distribution of Nb+Ta and the degree of lattice distortion generated by the change in the total number of elements. As the number of elements increases from four to seven, the Nb+Ta content reduces, causing a decrease in Tc [20]. The addition of Sc to the NbTaTiZr-based matrix generates a two-phase structure, with one phase being bcc, very similar to the Hf-Nb-Ta-Ti-Zr parent system and the other one hcp, which absorbs practically all Sc. Superconductivity emerges in the bcc phase with Tc = 7.2 K and Bc2(0) = 14.1 T, very similar to the Hf-Nb-Ta-Ti-Zr parent alloys [37]. Another studied system was Al-Nb-Ti-V-Zr. Harayama et al. [39] prepared Al5NbxTi35V5Zr55−x and Al5NbxTi59−xV5Zr31 samples, including the particular composition Al5Nb24Ti40V5Zr26, which have an interesting gum-metal-like behavior characterized by being superelastic and having a low Young’s modulus [35]. The Tc values of these compounds vary from 5.36 K to 7.01 K, being the maximum for the Al5Nb44Ti15V5Zr31 sample, following the Tc dependence on the VEC typical for the bcc-type HEA superconductors of type A [38]. The (TaNb)67(HfMoW)33 HEA was also investigated in detail. This HEA exhibits a metallic behavior at high temperature, while at Tc ∼ 4.3 K, it becomes superconducting with an upper critical field Bc2(0) ∼ 1.45 T [32].

2.3 Preparation methods

Until recent times, superconducting HEAs were generally prepared by arc melting. In the search for new features, alternative preparation methods were started, summarized in Table 2. Some of them, like Spark-Plasma Sintering (SPS) employed to synthesize Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6 samples, serve to obtain bulk samples similar to those prepared by arc melting. The Tc = 7.8 K and Bc2(0) = 10.5 T of these samples are similar to the one obtained in the samples by arc melting. However, the difference in the preparation method led to a very different pinning landscape. While the behavior of arc-melted samples is described by the surface pinning model, the SPS sintered sample includes also point pinning defects that sum up to the contribution of the surface pinning. This leads to a Jc of 30.5 kA/cm2 at 2 K and 0.01 T and 73.2 kA/cm2 at 4 K and 0.01 T, values that are 286% and 687% larger, respectively, to those obtained in the arc-melted sample [21]. Samples with exactly the same composition were also prepared by hot-press sintering. Also these samples show a similar Tc = 7.9 K as samples prepared by arc melting [22].

However, the most remarkable and interesting advances in alternative preparation methods are observed in the HEA films. The first work about HEA films was published by Zhang et al. and was about the preparation of (TaNb)100−x(HfZrTi)x films on Si substrates by DC magnetron co-sputtering at room temperature [24]. These films, with different stoichiometries and thicknesses (between 600 and 950 nm), crystallize, like their bulk relative, into a bcc structure. All the films are superconducting with Tc values in the range of 2.77–6.76 K. The largest Tc is found for a x = 43, and it is around 1 K lower than the bulk sample with the same composition due to, probably, the confined-size effect. The largest Bc2(0) = 11.05 T is found, however, for x = 65, which even exceeds the Pauli paramagnetic limit [24]. Films of a different stoichiometry of the same system, the (TaNb)70(HfZrTi)30, were also deposited by magnetron sputtering to investigate how the thickness and the external hydrostatic pressure affect the superconducting properties. It was proved that both Tc and Bc2(0) decay with decreasing thickness. The Tc values decrease from the original 7.8 K for the bulk to 5.31 K for the 600 nm film and 2.69 K for the 100 nm film. Films with a thickness of 30 nm are not superconducting. The same trend is observed for Bc2(0), with values of 6.7, 5.8, and 3.2 T, for bulk, 600 nm and 100 nm, respectively. The application of hydrostatic pressure up to 3.3 GPa has different consequences depending on the thickness of the film. In the case of the 600 nm film, Tc decreases with the pressure, which is different from the behavior observed on the bulk samples. However, no significant Tc change was observed when pressure was applied to the 100 nm film. This could indicate, as expected, that thinner films increase the disorder due to amorphization [25]. The same deposition technique was used in the work of Shu et al. to deposit (TaNbZrTi)100−xWx and (TaNbZrTi)100−xVx superconducting films. In this work, the influence of alloying the TaNbZrTi-based system with W and V and the deposition temperature on the phase formation and superconducting behavior of the films was investigated. The results show that all the films with different values of x show a bcc structure similar to the original TaNbZrTi system. In terms of Tc values, the highest one, 8.0 K, was reached for the original TaNbZrTi system. Then, for both alloying elements, Tc dropped monotonically as a function of x, almost at the same rate. When increasing the deposition temperature from room temperature to 400°C, Tc increases while the compositions remain unchanged. Moreover, the experimental observations led to the conclusion that the Tc values may also relate to the atomic radii difference and electronegativity difference of involved elements beyond the valence electron number [26]. A very similar study was done by Zhang et al. with (TaNb)100–x(HfZrTi)xMoy films. Films with different stoichiometries of ∼300 nm were deposited on Si substrates by magnetron sputtering. The films show again a bcc structure for low x values. For near-equimolar composition, the crystalline HEA grains transform into amorphous aggregations with a size of a few nanometers, forming a crystal/glass nanocomposite. This particular microstructure causes a broadening of the normal-to-superconducting transition and suppresses the zero-resistivity critical temperature to a lower constant value of approximately 2.9 K [40]. A different technique, namely pulsed laser deposition (PLD), was employed by Jung et al. to deposit Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6 films. This HEA, studied in several works in bulk samples, was deposited on Al2O3 substrates at different temperatures ranging from 270 to 620°C. These films, which crystallize as expected in bcc structure, show Tc values between ∼5.4 and 7.3 K, being the maximum value for deposition temperature of 520°C. Also, for the same deposition temperature, the largest Bc2(0) = 12.1 T is observed. All films showed significantly larger Jc values than the bulk sample, while a self-field value above 1 MA/cm2 (∼ 1.8 MA/cm2) at 4.2 K was measured for the film deposited at 520°C. The Jc values were maintained above 0.1 MA/cm2 under magnetic fields of approximately 3.4 and 2.0 T at 2.0 and 4.2 K, respectively, which is approximately eight times larger than in the bulk sample at the same conditions. Another remarkable result of this work is the effect of ion irradiation on the properties of the films. It was demonstrated that the HEA films are over 1000 times more robust to ion irradiation than other superconducting materials, such as Nb3Sn, MgB2, Fe-based superconductors, or REBCO. This means that the Tc of this HEA remains stable for much larger ion doses [23].

2.4 Other alloys

Rather different from the Ta-Nb-Hf-Zr-Ti system are the pentanary (ScZrNb)1−x[RhPd]x and hexanary (ScZrNbTa)100−x[RhPd]x systems reported by Stolze et al. [41]. They have investigated several mixtures of five and six elements showing a CsCl-type lattice, which is an ordered bcc structure, but had not been found in the original Ta-Nb-Hf-Zr-Ti system. It was also observed that Tc increases gradually with decreasing valence electron count (VEC) within a solution series. The highest Tc and Bc2(0) are observed in the particular case of (ScZrNb)65[RhPd]35 with values of ∼9.3 K and 10.7 T, respectively. Another important parameter that was observed in this case was the chemical composition, where the presence of niobium has a positive impact on the superconductivity. In any case, also samples without any niobium were superconducting proving that superconductivity is an intrinsic feature of the bulk material and not just an average of the constituents’ elemental properties [26]. The elastic properties of such a hexanary HEA superconductor, (ScZrNbTa)68.5[RhPd]31.5, was investigated by Pan et al. [42]. Its Young’s modulus E = 121.0 GPa and shear modulus of G = 44.9 GPa are around 10% larger than for Nb-Ti of similar Tc and ductility, which is promising for applications.

The authors of Ref. [41] reported also on three other new pentanary systems: (ZrNb)1−x[MoReRu]x, (HfTaWIr)100−x[Re]x, and (HfTaWPt)100−x[Re]x, which crystallize in a bcc α-Mn-type structure firstly observed in superconducting HEAs here [43]. The last two systems reported in this work are the first purely 5d-metal-based superconducting HEAs. Several stoichiometries were tested, showing Tc values between 2.1 K and 6.3 K. Just as in the case of type A Re-Nb-Hf-Zr-Ti, the presence of Re is important to understand the behavior of these systems. Here, Re plays a key role in the stabilization of this structure type [23]. Liu et al. [44] observed the same α-Mn-type structure in Nb25Mo5+xRe35Ru25−xRh10 (0 ≤ x ≤ 10), of which Nb25Mo10Re35Ru20Rh10 shows the largest Tc = 5.1 K and Bc2(0) = 8.3 T [41].

The general tendency of the superconducting HEAs to crystallize in cubic structures was not followed by the discovery of the Re56Nb11Ti11Zr11Hf11 alloy by Marik et al. [45], the first hexagonal superconducting HEA. These HEAs with hexagonal-close-packed (hcp) structure are in general rather interesting from the practical point of view because of their high hardness. Re56Nb11Ti11Zr11Hf11 has a Tc = 4.4 K, Bc1(0) = 2.3 mT, and Bc2(0) = 3.6 T [42]. Other superconducting HEAs crystallizing in hcp structure are, for example, (MoReRuRh)95Ti5 and (MoReRuRh)90Ti10 with Tc = 3.6 K and 4.7 K, respectively [46], or Nb10+2xMo35−xRu35−xRh10Pd10 (0 ≤ x ≤ 5) with a maximum Tc = 6.19 K for x = 2.5 [47]. A variant of the archetypical hcp structure is found in transition-metal antimonides M1–xPtxSb (M = equimolar mix of Ru, Rh, Pd, and Ir) that crystallize in a pseudo-hexagonal NiAs-type crystal structure [48]. The synthesis for various Pt contents, x, and heat treatment conditions lead to entropy-stabilized transition-metal monoantimonides. Superconductivity was only found for a particular composition, x = 0.2, for which a superconducting transition at Tc = 2.15 was observed. This Tc is similar to the highest values previously reported for transition-metal monoantimonides [45].

A transformation from the original hcp structure to a new face-centered cubic (fcc) one was observed in (MoReRu)(1−2x)∕3(PdPt)x and carbon-added (MoReRu)(1−2x)∕3(PdPt)xCy HEAs by Zhu et al. [49]. The (MoReRu)(1−2x)∕3(PdPt)x alloys were found to crystallize in an hcp structure for 0.042 ≤ x ≤ 0.167 and in an fcc structure for x = 0.333. Their VEC values are somewhat outside the expected ranges, between 7.0 and 8.0. The addition of a certain amount of C that depends on x causes also a transformation to the fcc structure. As in the C-free case, the structural transformation is likely driven by the combined effects of VEC, lattice distortion, and increase in entropy. The Tc values of the hcp phases range from 1.75 K to 8.17  K, while the ones of the fcc structures are always below 2.8 K [46].

A different crystal structure is observed in the Ta5[Mo35–xW5+x]Re35Ru20 and [Ta5+yMo35–y]W5Re35Ru20 HEAs [8]. In this case, the HEAs crystallize into tetragonal sigma (σ) phases. These systems exhibit fully gapped bulk superconductivity with Tc values between 5.27 K and 6.29 K, reaching the maximum at x = y = 0 [47]. The same type of structure was observed in Ta10Mo35−xCrxRe35Ru20 (5 ≤ x ≤ 13) [50], highlighted by the presence of Cr, which is not commonly found in superconducting materials. These HEAs exhibit Tc values up to 4.79 K and show a paramagnetic behavior in the normal state due to the magnetic moment of Cr [48].

A transformation from the σ-phase into a new structure happens in Ta10Mo5W30Re35Ru20 upon adding C [51]. The Ta10Mo5W30Re35Ru20Cx HEA crystallizes in the centrosymmetric σ-type structure for 0 ≤ x ≤ 2 but transforms into the noncentrosymmetric β-Mn structure for 16 ≤ x ≤ 20. The HEA with the β-Mn structure shows Tc = 5.34 K, which is larger than for the original σ-type structure. The Bc2(0) = 9.3 T is close to the Pauli paramagnetic limit. This phase transformation is likely caused by the decrease in VEC and the increase in lattice distortion and mixing entropy [49]. The same β-Mn structure is found in the Cr5+xMo35−xW12Re35Ru13C20 HEA 0 ≤ x ≤ 9 [52]. By increasing x, Tc was observed to decrease from 5.49 K to 3.35 K due to the magnetic pair breaking caused by Cr moments. For all the different stoichiometries, Bc2(0) is comparable to the Pauli paramagnetic limit BP(0) = 1.86Tc, having a maximum value of 9.7 T [50].

Based on the relationship between the VEC and Tc values for HEAs and other binary or ternary superconducting alloys, new combinations of elements and stoichiometries were tested approaching the VEC values where superconductivity appears. As a result, several new HEA systems have been discovered. One example is the new boron-based HEA superconductor Mo11W11V11Re34B33 (MWVRB) [53]. With a VEC = 5.24, this compound crystallized in the tetragonal CuAl2 crystal structure and showed a Tc = 4.0 K and a Bc2(0) = 7.3 T [51]. For a summary of these HEAs different from TaNbHfZrTi basis, see Table 4.

StructureNominal compositionTc (K)Bc2(0) (T)Reference
CsCl(ScZrNb)60[RhPd]405.202.10[41]
(ScZrNb)62[RhPd]389.208.90[41]
(ScZrNb)63[RhPd]379.309.60[41]
(ScZrNb)65[RhPd]359.7010.70[41]
(ScZrNbTa)67[RhPd]334.202.10[41]
(ScZrNbTa)684[RhPd]3166.408.80[41]
(ScZrNbTa)685[RhPd]3157.0[42]
α-Mn(ZrNb)10[MoReRu]905.307.86[43]
(HfTaWIr)20[Re]805.90[43]
(HfTaWIr)40[Re]604.004.64[43]
(HfTaWPt)20[Re]806.30[43]
(HfTaWPt)40[Re]604.405.90[43]
(HfTaWPt)50[Re]502.40[43]
Nb25Mo5Re35Ru25Rh104.667.50[44]
Nb25Mo10Re35Ru20Rh105.108.30[44]
Nb25Mo15Re35Ru15Rh105.107.90[44]
Nb5Mo35Re15Ru35Rh107.548.90[44]
Nb5Mo30Re20Ru35Rh106.697.50[44]
Nb5Mo25Re25Ru35Rh106.517.50[44]
Nb5Mo20Re30Ru35Rh105.466.10[44]
hcpRe56Nb11Ti11Zr11Hf114.403.60[45]
(MoReRuRh)95Ti53.60[46]
(MoReRuRh)90Ti104.70[46]
Nb10Mo35Ru35Rh10Pd105.586.90[47]
Nb15Mo32.5Ru32.5Rh10Pd106.198.10[47]
Nb20Mo30Ru30Rh10Pd106.108.30[47]
Mo30Re29Ru36Pd1Pt48.17[49]
Mo27Re25Ru34Pd5Pt94.91[49]
Mo24Re20Ru30Pd12Pt142.22[49]
Mo22Re22Ru28Pd10Pt181.64[49]
NiAs(RuRhPdIr)80Pt20Sb (SSR)2.15[48]
fccMo28Re28Ru35Pd4Pt5C302.46[49]
Mo29Re27Ru36Pd4Pt4C442.72[49]
Mo26Re25Ru33Pd8Pt9C272.28[49]
Mo27Re25Ru34Pd6Pt9C402.30[49]
Mo26Re22Ru31Pd9Pt13C221.86[49]
Mo23Re23Ru30Pd11Pt14C311.91[49]
Mo21Re21Ru27Pd13Pt18C251.87[49]
Mo21Re22Ru27Pd14Pt17C301.75[49]
σ-typeTa5Mo35W5Re35Ru206.29[8]
Ta5Mo30W10Re35Ru206.20[8]
Ta5Mo25W15Re35Ru206.10[8]
Ta5Mo20W20Re35Ru205.69[8]
Ta5Mo15W25Re35Ru205.46[8]
Ta5Mo10W30Re35Ru205.45[8]
Ta5Mo5W35Re35Ru204.78[8]
Ta7Mo33W5Re35Ru206.13[8]
Ta9Mo31W5Re35Ru205.70[8]
Ta11Mo29W5Re35Ru205.32[8]
Ta13Mo27W5Re35Ru205.27[8]
Ta10Mo30Cr5Re35Ru204.796.10[50]
Ta10Mo25Cr10Re35Ru204.415.80[50]
Ta10Mo22Cr13Re35Ru203.984.90[50]
Ta10Mo5W30Re35Ru204.876.70[51]
Ta10Mo5W30Re35Ru20C24.80[51]
β-MnTa10Mo5W30Re35Ru20C165.36[51]
Ta10Mo5W30Re35Ru20C185.32[51]
Ta10Mo5W30Re35Ru20C205.349.30[51]
Cr5Mo35W12Re35Ru13C205.499.70[52]
Cr8Mo32W12Re35Ru13C204.738.40[52]
Cr11Mo29W12Re35Ru13C203.837.30[52]
Cr14Mo26W12Re35Ru13C203.355.30[52]
CuAl2Mo11W11V11Re34B334.007.30[53]

Table 4.

Selected properties of non-bcc structure-type HEA superconductors. SSR solid-state reaction; all other samples prepared by arc melting.

Advertisement

3. Superconducting high-entropy compounds

3.1 Intermetallics

Other high-entropy superconductors with different compositions present alternative structures to the previous ones, Table 5. For example, (V0.5Nb0.5)3−xMoxAl0.5Ga0.5 (0.2 < x ≤ 1.4) is polymorphic with bcc and A15 structure (similar to α- and β-W actually) [7]. Furthermore, heating also induces a polymorphic transformation from original bcc structure to the A15 one, as shown for x = 0.2. The resulting compound with this A15 structure shows bulk superconductivity with Tc = 10.2 K and an estimated Bc2(0) = 20.1 T [52]. Here, mixing on both distinct lattice sites occurs, and a true alloy resembling β-W is forming. A similar case is the system V5+2xNb35−xMo35−xIr10Pt15 (0 ≤ x ≤ 10) with maximum Tc = 5.18 K and Bc2(0) = 6.4 T for x = 0 [55].

StructureNominal compositionTc (K)Bc2(0) (T)Reference
A15(V0.5Nb0.5)2.8Mo0.2Al0.5Ga0.510.2020.10[7]
(V0.5Nb0.5)2.6Mo0.4Al0.5Ga0.59.2017.70[7]
(V0.5Nb0.5)2.4Mo0.6Al0.5Ga0.58.9017.00[7]
(V0.5Nb0.5)2Mo1Al0.5Ga0.56.109.90[7]
(V0.5Nb0.5)1.8Mo1.2Al0.5Ga0.54.807.60[7]
(V0.5Nb0.5)1.6Mo1.4Al0.5Ga0.53.204.80[7]
V5Nb35Mo35Ir10Pt155.186.40[55]
V15Nb30Mo30Ir10Pt154.495.70[55]
V25Nb25Mo25Ir10Pt153.614.40[55]
Nb3Al0.2Sn0.2Ge0.2Ga0.2Si0.29.0010.40[56]
Nb3Al0.3Sn0.3Ge0.2Ga0.1Si0.111.0013.30[56]
V3Al0.07Si0.30Ga0.08Ge0.30Sn0.256.308.8[57]
CuAl2Co0.2Ni0.1Cu0.1Rh0.3Ir0.3Zr28.012[58]
Fe0.09Co0.19Ni0.11Rh0.27Ir0.33Zr27.88.5[54]
Fe0.11Co0.30Ni0.20Rh0.07Ir0.32Zr26.76.6[54]
Fe0.19Co0.19Ni0.20Rh0.21Ir0.21Zr25.45.5[54]
Fe0.29Co0.19Ni0.30Rh0.09Ir0.12Zr24.84.8[54]
Fe0.09Co0.20Ni0.20Cu0.17Rh0.19Ir0.15Zr25.7[59]

Table 5.

Selected properties of A15 and TrZr2 high-entropy superconductors. All samples have been prepared by arc melting. The compositions of Ref. [54] were rounded to two decimals for simplicity.

The A15 structure type can be also found in high-entropy compounds like Nb3Al0.2Sn0.2Ge0.2Ga0.2Si0.2 with a Tc = 9.0 K and Bc2(0) = 10.4 T [56], Nb3Al0.3Sn0.3Ge0.2Ga0.1Si0.1 with a Tc = 11.0 K and Bc2(0) = 13.3 T [56], and V3Al0.07Si0.30Ga0.08Ge0.30Sn0.25 with a Tc = 6.3 K and Bc2(0) = 8.8 T [57]. In all these cases, mixing occurs primarily on just one of the lattice sites (disregarding natural disorder), and a compound resembling Cr3Si (or Nb3Sn, for that matter) is formed. The V system is also interesting because it introduced the concept of compositionally complex alloys (CCAs) as a new category of multi-principal-element materials closely related to high-entropy alloys, but containing more than two phases in an alloy sample. In the particular case of V3Al0.07Si0.30Ga0.08Ge0.30Sn0.25, it crystallizes into five high-entropy A15 phases with different compositions. This concept of CCAs is thought to be a practical way to improve the properties of the superconducting HEAs.

Further high-entropy intermetallics are the so-called TrZr2 compounds (Tr (mix of) transition metal) with CuAl2 crystal structure. They have been discovered by Mizuguchi et al. in Co0.2Ni0.1Cu0.1Rh0.3Ir0.3Zr2 and can be regarded as mixing of single-Tr zirconides, of which many are superconductors themselves [58]. This high-entropy compound shows a Tc of 8.0 K and a Bc2(0) of 12 T. Shortly after, superconductivity was also found in (Fe,Co,Ni,Rh,Ir)Zr2 samples of different stoichiometries [54]. The same group found an unusual broadening of the specific heat jump at Tc for this system (developing with the number of Tr elements), which they attributed to an inhomogeneity of the Cooper pair formation and hence the superconducting gap [59]. Pugliese et al. [60] investigated possible local inhomogeneities in these systems by extended X-ray absorption fine structure (EXAFS) measurements and found a correlation between Tc and the Zr 4d-Tr nd hybridization. The Zr-Zr bonds seem to get stiffer with increasing mixing entropy, which suggests nanoscale texturing with possible local order.

3.2 Non-oxide ceramics

The concept of high-entropy alloys primarily focused on metal alloys, however, was quickly extended to other types of materials, Table 6. This approach, centered on maximizing the configurational entropy to stabilize equimolar or near-equimolar mixtures, was employed in a mixture of oxides to synthesize the first high-entropy oxide (HEO) in 2015 [71], giving rise to a new category known as high-entropy ceramics (HECs). HECs encompass all high-entropy materials with ceramic properties. Similar to HEAs, HECs consist of multicomponent elements in a single phase, where their substantial configurational entropy plays a crucial role in their formation. In contrast to metallic HEAs, HECs typically exhibit semiconductor or insulator characteristics, making them potentially useful as functional materials. These systems have proven to be versatile in various technologies, including thermal barrier coatings, thermoelectrics, catalysts, and batteries, as well as wear-resistant and corrosion-resistant coatings [72, 73, 74]. However, in this chapter, the focus will be on those materials that show superconducting properties.

StructureNominal compositionTc (K)Bc2(0) (T)Reference
NaCl(TiZrNbHfTa)C2.350.51[62]
(MoNbTaVW)C0.93.43.37[63]
(AgInSnPbBi)Te2.62.8[64]
({AgSnPbBi}(1−x)/4Inx)Te<2.8[65]
1 T-NiTe2(Co,Au)0.2(Rh,Ir,Pd,Pt)0.8Te24.5[66]
Co0.03Au0.06Rh0.23Ir0.24Pd0.16Pt0.28Te22.5[66]
W5Si3(Nb0.1Mo0.3W0.3Re0.2Ru0.1)5Si33.305.00[67]
(Nb0.2Mo0.3W0.3Re0.1Ru0.1)5Si33.205.10[67]
LaOBiS2La0.1Ce0.1Pr0.2Nd0.3Sm0.3O0.5F0.5BiS24.9[68]
La0.2Ce0.2Pr0.2Nd0.2Sm0.2O0.5F0.5BiS23.97[69]
La0.3Ce0.3Pr0.2Nd0.1Sm0.1OBiS23.40.65/11[61]
La0.1Ce0.3Pr0.3Nd0.2Sm0.1OBiS24.30.69/15[61]
La0.2Ce0.2Pr0.2Nd0.2Sm0.2OBiS23.30.24/4.9[61]
La0.1Ce0.3Pr0.1Nd0.2Sm0.3OBiS24.60.42/16[61]
La0.1Ce0.3Pr0.3Nd0.1Sm0.2OBiS23.7[70]
La0.2Ce0.3Pr0.3Sm0.1Gd0.1OBiS23.0[70]
La0.2Ce0.3Pr0.2Nd0.1Sm0.1Gd0.1OBiS22.9[70]

Table 6.

Selected properties of high-entropy superconducting ceramics. For Ref. [61], Bc2(0) is given for both major field directions, B||c and Bc.

High-entropy carbide ceramics (HECCs) are a subgroup of materials belonging to the more general one of the HEC. The high-entropy carbides are coming from the binary transition metal carbides (TMCs), composed of carbon and a transition metal, that are widely used as ultra-high-temperature ceramics in structural applications. They are made by combining different TMCs to synthesize a single-phase multi-principal elemental HECC [75]. Among them, (TiZrNbHfTa)C was the first one to show superconducting properties [62]. This HECC, prepared by SPS, crystallizes in NaCl structure and shows bulk superconductivity with Tc = 2.35 K, Bc1(0) ∼ 26.1 mT, and Bc2(0) ∼ 0.51 T. Those superconducting properties are robust and were kept almost invariable under pressure up to 80 GPa [74]. Very recently, Shu et al. discovered the coexistence of ferromagnetism and superconductivity in (MoNbTaVW)C0.9 and (TaTiNbZr)C [63].

Some superconducting telluride high-entropy compounds crystallize in the NaCl structure too (Figure 2b). Superconductivity was first seen in AgInSnPbBiTe5 by Mizuguchi with a Tc of ∼2.6 K and an extrapolated Bc2(0) of ∼2.8 T [64]. Upon applying pressure up to 35 GPa, the crystal structure changes via a Pnma structure (at around 12 GPa) to the CsCl type with increased Tc values of ∼5.3 K [76]. The stoichiometry on the M site was varied shortly after by adjusting the In content as (AgSnPbBi)(1−x)/4InxTe with x = 0, 0.1, 0.2, 0.3, and 0.4, where the sample without In is not superconducting, and Tc increases slightly with x [65]. A special case of superconducting high-entropy tellurides is (Co,Au)0.2(Rh,Ir,Pd,Pt)0.8Te2 and Co0.03Au0.06Rh0.23Ir0.24Pd0.16Pt0.28Te2 as layered compounds with 1 T-NiTe2 structure. They are called high-entropy van der Waals materials (HEX) and show superconductivity up to 4.5 K [66].

The case of the high-entropy silicides (HESs) is quite similar to the carbides. They are made by combining Si with other elements, mainly metals. The binary transition-metal silicides M5Si3 were studied extensively for high-temperature structural applications due to their high melting points. Again, combining these binary silicides, different HESs were synthesized. However, the first superconducting high-entropy silicides appeared only very recently [67]. (Nb0.1Mo0.3W0.3Re0.2Ru0.1)5Si3 and (Nb0.2Mo0.3W0.3Re0.1Ru0.1)5Si3 were prepared by arc melting and are weakly coupled bulk superconductors. They both crystallize in the W5Si3 structure and have Tc values of 3.3 K and 3.2 K, and Bc2(0) values of 5.0 T and 5.1 T, respectively [65].

Superconducting high-entropy sulfides are derived from the layered superconductors RE(O,F)BiS2 by mixing the rare earth element (RE) (Figure 2c)[77]. The first reported system was (La,Pr,Ce,Nd,Sm)O0.5F0.5BiS2, where Tc depends linearly on lattice parameter a, and La0.1Ce0.1Pr0.2Nd0.3Sm0.3O0.5F0.5BiS2 showed the highest value of ∼4.9 K [68]. The same group investigated shortly after the dependence of structural and superconducting properties on the mixing entropy by compounds with one to five RE elements, where Tc and superconducting volume fraction increase linearly with mixing entropy [69]. Shortly after, superconductivity was also confirmed in F-free HE REOBiS2 single crystals [61, 70], for which the anisotropy of the upper critical field could be determined.

3.3 Oxides

The first high-entropy oxides (HEOs) were reported by Rost et al. in 2015 [71]. In this work, the authors demonstrated that the entropy drives a reversible solid-state transformation between a multiphase and a single-phase state and was able to synthesize (Mg0.2Co0.2Ni0.2Cu0.2Zn0.2)O, the first HEO [61]. However, the term HEO was only introduced 1 year later by Bérardan et al., making the analogy to HEAs, to classify these multicationic, equiatomic oxide systems [78]. Over the years, the HEOs have gained significant interest due to their unique structural characteristics and related possibilities for tailoring functional properties [79, 80, 81, 82]. One particularity of the HEOs, in contrast to HEAs, is that here the elemental mixing is restricted to one or only some of the atomic positions, similar to the metallic and ceramic compounds mentioned above. Superconductivity in HEOs could only be found in compounds of the REBa2Cu3O7−δ (REBCO) family so far, probably the best-known and studied superconducting oxides. A recent search, for example, for superconducting high-entropy Ruddlesden-Popper cuprates has not been successful so far [83].

The concept of HEO was applied for the first time to REBCO compounds in 2020 by Shukunami et al. when the first high-entropy REBCO (HE REBCO) compound was reported [10]. The REBCO compounds exhibit a triple perovskite structure, whose center is the RE atom. This RE site is occupied with a mixture of rare earth elements, so that is where entropy is introduced (Figure 2d).

To the best of our knowledge, although the number has been increasing year by year since 2020, only 10 articles have been reported on HE REBCO to date. The different HE REBCO compounds presented in those articles can be classified as bulks or films according to the synthesis method followed (Table 7). Within the group of the films, one can further distinguish between films by pulsed laser deposition (PLD) and by chemical solution deposition (CSD).

Nominal composition on RE siteTc (K)Jc (MA/cm2)Reference
Y0.28Nd0.16Sm0.18Eu0.18Gd0.2093.00.02 (2 K, sf)
0.0117 (2 K, 1 T)
[10]
Y0.18La0.24Nd0.14Sm0.14Eu0.15Gd0.1593.00.081 (2 K, 1 T)[10]
Y0.2La0.2Sm0.2Dy0.2Er0.292.0[84]
Y0.2Ho0.2Er0.2Tm0.2Yb0.292.0[84]
Y0.5La0.125Sm0.125Dy0.125Er0.12592.00.02 (2 K, sf)
0.002 (77 K, sf)
[84]
Y0.5Sm0.125Eu0.125Gd0.125Dy0.12592.0[84]
Y0.2La0.2Sm0.2Eu0.2Gd0.293.40.0109 (JcGlobalat 2 K, 1 T)
6.5 (Jclocal at 2 K, sf)
[85]
Y0.2Nd0.2Sm0.2Eu0.2Gd0.293.40.0117 (2 K, 1 T)[85]
Y0.2Sm0.2Eu0.2Dy0.2Ho0.293.00.0346 (2 K, 1 T)[85]
Y1/6La1/6Nd1/6Sm1/6Eu1/6Gd1/693.10.0081 (2 K, 1 T)[85]
Y1/6Sm1/6Eu1/6Dy1/6Ho1/6Yb1/692.00.0499 (2 K, 1 T)[85]
Y1/7Sm1/7Eu1/7Dy1/7Ho1/7Yb1/7Lu1/791.40.0535 (JcGlobalat 2 K, 1 T)
5.5 (Jclocal at 2 K, sf)
[85]
Dy0.16Ho0.17Er0.20Tm0.22Yb0.2588.8[86]
Gd0.23Dy0.17Ho0.15Er0.15Tm0.19Yb0.2089.8[86]
Gd0.20Dy0.13Ho0.10Er0.15Tm0.09Yb0.20Lu0.1589.4[86]
Y0.12Sm0.08Eu0.26Dy0.18Ho0.36 (PLD film)90.56.5 (2 K, 1 T)
2.3 (4.2 K, 7 T)
[87]
Y0.7Gd0.2Dy0.2Sm0.2Eu0.2 (CSD film)93.010 (30 K, sf)
4 (65 K, sf)
2 (77 K, sf)
[88]
Gd0.2Sm0.2Nd0.2Eu0.2Y0.2 (CSD film)90.42.4 (77 K, sf)[89]
Gd0.2Dy0.2Y0.2Ho0.2Er0.2 (+12% BHO, CSD)91.93.5 (77 K, sf)[90]

Table 7.

Critical temperatures and selected Jc data of high-entropy REBCO phases. The samples, except otherwise stated, are prepared by solid-state reaction. sf, self-field or remanent state.

Shukunami et al. manufactured polycrystalline REBCO bulk samples via solid-state reaction, combining up to six different rare earths in 2020 [10]. So far, all bulk HE REBCO samples have been prepared by SSR. In particular, they presented Y0.28Nd0.16Sm0.18Eu0.18Gd0.20Ba2Cu3O7−δ and Y0.18La0.24Nd0.14Sm0.14Eu0.15Gd0.15Ba2Cu3O7−δ as the first two superconducting HE REBCO compounds with Tc = 93 K in both cases. The Jc values of these compounds were calculated from magnetic measurements at 2 K and 1 T. The values were 11.7 kA/cm2 for Y0.28Nd0.16Sm0.18Eu0.18Gd0.20Ba2Cu3O7−δ (∼20 kA/cm2 at self-field) and 8.1 kA/cm2 for Y0.18La0.24Nd0.14Sm0.14Eu0.15Gd0.15Ba2Cu3O7−δ. The authors also identify a clear correlation between Tc and the orthorhombicity parameter (OP), which is defined as 2|ab|/(a+b) with lattice constants a and b. For OP > 0.012, the Tc of the samples is almost constant, but for OP < 0.012, Tc starts to decay. For OP < 0.007, superconductivity was not observed anymore. Interestingly and accordingly, two further HE REBCO compounds that they prepared were not superconducting, namely, Y0.23La0.25Nd0.15Sm0.19Eu0.18Ba2Cu3O7−δ and the Y0.23La0.23Nd0.17Sm0.18Gd0.19Ba2Cu3O7−δ. These results indicate the importance of wisely selecting the RE elements to include in the HE REBCO to maximize the superconducting properties. One year later, Wang et al. [84] prepared four new HE REBCO compounds with a mixture of five elements in each case: Y0.2La0.2Sm0.2Dy0.2Er0.2Ba2Cu3O7−δ, Y0.2Ho0.2Er0.2Tm0.2Yb0.2Ba2Cu3O7−δ, Y0.5La0.125Sm0.125Dy0.125Er0.125Ba2Cu3O7−δ, and Y0.5Sm0.125Eu0.125Gd0.125Dy0.125Ba2Cu3O7−δ. Magnetic measurements showed that all these samples exhibit a very similar Tc of ∼92 K. The self-field Jc values of Y0.5La0.125Sm0.125Dy0.125Er0.125Ba2Cu3O7−δ were 20 kA/cm2 at 1.8 K and 2 kA/cm2 at 77 K. These samples show smaller irreversible fields than traditional YBCO, indicating again that the selection of the RE elements has a huge influence on the properties of these compounds. In 2022, Yamashita et al. of the same group that reported the first HE REBCO went a step further [85]. They prepared six new compounds by solid-state reaction including one with seven different RE: Y0.2La0.2Sm0.2Eu0.2Gd0.2Ba2Cu3O7−δ, Y0.2Nd0.2Sm0.2Eu0.2Gd0.2Ba2Cu3O7−δ, Y0.2Sm0.2Eu0.2Dy0.2Ho0.2Ba2Cu3O7−δ, Y1/6 La1/6Nd1/6Sm1/6Eu1/6Gd1/6Ba2Cu3O7−δ, Y1/6Sm1/6Eu1/6Dy1/6Ho1/6Yb1/6Ba2Cu3O7−δ, and Y1/7Sm1/7Eu1/7Dy1/7Ho1/7Yb1/7Lu1/7Ba2Cu3O7−δ. Tc values of these compounds are between 93.4 K and 91.4 K, slightly decreasing with RE complexity. Similar to their first work on Tc, they correlated the OP here with Jc. Again, there is an obvious correlation between both parameters, being the largest Jc = 53.5 kA/cm2 (2 K, 1 T), for the largest OP that corresponds to the compound with seven RE, even though Tc was the smallest. This value of Jc, measured magnetically on cube-shaped samples, corresponds to JcGlobal. However, this Jc is affected by the typical weak links from grain to grain found in the REBCO compounds due to their short coherence lengths. Therefore, the authors estimate Jc within the grains (Jclocal) in fine powders of the same composition as the samples by magnetic measurements. At 2 K, Jclocal of ∼6.5 MA/cm2 was found for Y0.2Sm0.2Eu0.2Dy0.2Ho0.2Ba2Cu3O7−δ and ∼ 5.5 MA/cm2 for Y1/7Sm1/7Eu1/7Dy1/7Ho1/7Yb1/7Lu1/7Ba2Cu3O7−δ. At 20 K, the values for the different compounds are very similar, ∼2 MA/cm2 independent of the number of RE [84]. Also, in 2023, the use of solid-state reaction synthesis has been continued. Suzuki et al. [86] followed this method to prepare HE REBCO single-crystal whiskers with dimensions 0.5–1.0 mm in length, 10 μm in thickness, and 10–50 μm in width. The analytical HE REBCO compositions were Dy0.16Ho0.17Er0.20Tm0.22Yb0.25Ba2Cu3O7−δ, Gd0.23Dy0.17Ho0.15Er0.15Tm0.19Yb0.20Ba2Cu3O7−δ, and Gd0.20Dy0.13Ho0.10Er0.15Tm0.09Yb0.20Lu0.15Ba2Cu3O7−δ. The Tc zero values of these compounds were around 89 K, which indicates that Tc of HE REBCO compounds with a similar average RE ionic radius does not depend on the mixture entropy at the RE site. Therefore, the advantages of high-entropy alloys may be kept for a high degree of entropy without reducing the superconducting properties of the REBCO compounds. Finally, Pryanichnikov et al. [91] synthesized equimolar Y0.2Nd0.2Eu0.2Sm0.2Ho0.2Ba2Cu3Oy bulk samples and studied their superconducting properties. The purpose of this work was to change the oxygen content by annealing in appropriate conditions and to study the crystal structure features and superconducting properties. It was concluded that OP, Tc, and Jc of Y0.2Nd0.2Eu0.2Sm0.2Ho0.2Ba2Cu3Oy increase with oxygen content y, very similar to their relatives of “low-entropy” REBCO compounds.

Within the group of HE REBCO compounds prepared as films, the first work published was the one of Yamashita et al., who prepared Y0.12Sm0.08Eu0.26Dy0.18Ho0.36Ba2Cu3O7−δ films by PLD [87]. This sample containing five REs was compared with other low- and medium-entropy compounds. All samples grew with c-axis-preferential orientation. In addition, the c-axis of the samples increases with the average diameter of the REs mixture and so does the width of the (005) peak, indicating, as expected, that there is an increase in the disorder of the crystal lattice. The Tc of the Y0.12Sm0.08Eu0.26Dy0.18Ho0.36Ba2Cu3O7−δ film was 90.5 K, the same as for the medium-entropy films with three and four REs. It shows Jc = 2.3 MA/cm2 at 4.2 K and 7 T, and that stays above 1.0 MA/cm2 up to 20 K and 7 T. These values are very similar to the medium-entropy film with four RE and to the standard YBCO film in the high-field region. Few months later, Chen et al. reported the first HE REBCO film deposited by fluorine-free CSD [88], namely, an 800-nm thick nominal Y0.7Gd0.2Dy0.2Sm0.2Eu0.2Ba2Cu3O7−δ film on technical substrates. This compound presents the particularity that it was nominally off stoichiometry on the RE molarity because the authors argued this type of off-stoichiometric composition ratio provided better superconducting properties than standard 1:2:3, possibly mostly due to the formation of RE2O3 nanoparticles, which serve as pinning centers. The structural and superconducting properties of this HE REBCO were compared with two control samples prepared the same way, Y1Dy0.5Ba2Cu3O7−δ (also off stoichiometry on the RE molarity) and YBCO. They showed that high-quality c-axis-oriented Y0.7Gd0.2Dy0.2Sm0.2Eu0.2Ba2Cu3O7−δ films with Tc = 93 K can be obtained. The in-field Jc values of this HE REBCO film are significantly improved, especially at 77 K or below 4 T. Self-field Jc values of 2, 4, and 10 MA/cm2 were obtained at 77, 65, and 30 K, respectively. The maximum pinning force density of the HE REBCO film at 77 K was 1.7 times higher than that of the standard YBCO film. However, it is difficult to confirm if this enhancement of Jc originates from the effect of the HE RE mixing or from the contribution of the RE2O3 pinning centers that were also found in the films. Fluorine-free solutions were also used by Masuda et al. to deposit 300-nm Gd0.2Sm0.2Nd0.2Eu0.2Y0.2Ba2Cu3O7−δ films on LaAlO3 substrates [89]. The films show a Tc = 90.4 K and a self-field Jc = 2.4 MA/cm2 at 77 K, both exceeding a standard GdBCO film of the same study. Cayado et al. raised the level of complexity further by introducing perovskite nanoparticles into the HE REBCO film [90]. They prepared Gd0.2Dy0.2Y0.2Ho0.2Er0.2Ba2Cu3O7−δ nanocomposite films with 12 mol% BaHfO3 (BHO) nanoparticles on SrTiO3 substrates by CSD using full-TFA solutions. Both Tc and self-field Jc at 77 K of such a film, 91.9 K and 3.5 MA/cm2, respectively, are comparable with the ones of low- and medium-entropy REBCO films prepared in the same way. By an exhaustive TEM study, a homogeneous distribution of the BHO nanoparticles as well as a homogeneous distribution of the REs at the atomic level instead of RE clustering, were observed, proving a true HE REBCO. Grünewald et al. [92] of the same group continued these TEM analyses of such films, focusing on the mixing of REs into the BHO nanoparticles. This is important as the possible incorporation of REs in the nanoparticles could lead to off-stoichiometries in the REBCO phase, causing a deviation from the expected growth process and reduction of the superconducting properties. The degree of REs incorporation in BHO was found to depend on both the ionic radius and the content of the REs in the REBCO phase. The REs incorporation increases with decreasing ionic radii and shows a stronger tendency for smaller ionic radii to intermix in medium- and high-entropy REBCO than in the low-entropy ones.

Advertisement

4. Conclusion

In this comprehensive review, we have described the state of the art of superconducting high-entropy materials, focusing on fabrication methods and superconducting properties. The extension of high-entropy principles to the field of superconductivity has opened up new frontiers in the search for new superconductors and offers promising perspectives for innovative technological advancements. The potential applications of superconducting high-entropy materials cover a wide spectrum, from energy transmission to quantum computing. Their ability to maintain superconductivity under varying external conditions positions them as promising candidates for different applications. The discovery of new high-entropy superconductors has been continuous as the first one was found in 2014, while the interest in this topic is growing year by year. Due to their chemical similarities, transition metals and rare earth elements are predestined for mixing on certain lattice sites. The first superconducting HE alloys contained only low transition metals, which were later extended to higher transition metals and even main group metals. These alloys with simple bcc crystal structures were further developed toward more complex crystal structures and further to compounds. In two layered compound classes, REOBiS2 and REBCO, high-entropy mixing on the RE site has been shown. Similar investigations might be interesting in the iron-based superconducting LnFeAsO system, Ln lanthanoid, which is still missing. Combining transition metal mixing and rare earth mixing on appropriate lattice sites might be a further increase in complexity. There is certainly still plenty of room for new, exciting investigations in this field.

References

  1. 1. Yeh JW, Chen SK, Lin SJ, Gan JY, Chin TS, Shun TT, et al. Nanostructured high-entropy alloys with multiple principal elements: Novel alloy design concepts and outcomes. Advanced Engineering Materials. 2004;6(5):299-303
  2. 2. Cantor B, Chang ITH, Knight P, Vincent AJB. Microstructural development in equiatomic multicomponent alloys. Materials Science and Engineering: A. 2004;375–377(1-2 Spec. issue):213-218
  3. 3. Yuan Y, Wu Y, Luo H, Wang Z, Liang X, Yang Z, et al. Superconducting Ti15Zr15Nb35Ta35 high-entropy alloy with intermediate electron-phonon coupling. Frontiers in Materials. 2018;5
  4. 4. Sun L, Cava RJ. High-entropy alloy superconductors: Status, opportunities, and challenges. Physical Review Materials. 2019;3(9):090301
  5. 5. Koželj P, Vrtnik S, Jelen A, Jazbec S, Jagličić Z, Maiti S, et al. Discovery of a superconducting high-entropy alloy. Physical Review Letters. 2014;113(10):107001
  6. 6. Matthias BT. Empirical relation between superconductivity and the number of valence electrons per atom. Physical Review. 1955;97:74
  7. 7. Wu J, Liu B, Cui Y, Zhu Q, Xiao G, Wang H, et al. Polymorphism and superconductivity in the V-Nb-Mo-Al-Ga high-entropy alloys. Science China Materials. 2020;63(5):823-831
  8. 8. Liu B, Wu J, Cui Y, Zhu Q, Xiao G, Wang H, et al. Formation and superconductivity of single-phase high-entropy alloys with a tetragonal structure. ACS Applied Electronic Materials. 2020;2(4):1130-1137
  9. 9. Senkov ON, Scott JM, Senkova SV, Miracle DB, Woodward CF. Microstructure and room temperature properties of a high-entropy TaNbHfZrTi alloy. Journal of Alloys and Compounds. 2011;509(20):6043-6048
  10. 10. Shukunami Y, Yamashita A, Goto Y, Mizuguchi Y. Synthesis of RE123 high-Tc superconductors with a high-entropy-alloy-type RE site. Physica C: Superconductivity and Its Applications. 2020;572:1353623
  11. 11. Jasiewicz K, Wiendlocha B, Korbeń P, Kaprzyk S, Tobola J. Superconductivity of Ta34Nb33Hf8Zr14Ti11 high entropy alloy from first principles calculations. Physica Status Solidi (RRL)–Rapid Research Letters. 2016;10(5):415-419
  12. 12. Vrtnik S, Koželj P, Meden A, Maiti S, Steurer W, Feuerbacher M, et al. Superconductivity in thermally annealed Ta-Nb-Hf-Zr-Ti high-entropy alloys. Journal of Alloys and Compounds. 2017;695:3530-3540
  13. 13. Von Rohr F, Winiarski MJ, Tao J, Klimczuk T, Cava RJ. Effect of electron count and chemical complexity in the Ta-Nb-Hf-Zr-Ti high-entropy alloy superconductor. Proceedings of the National Academy of Sciences. 2016;113(46):E7144-E7150
  14. 14. Guo J, Wang H, Von Rohr F, Wang Z, Cai S, Zhou Y, et al. Robust zero resistance in a superconducting high-entropy alloy at pressures up to 190 GPa. Proceedings of the National Academy of Sciences. 2017;114(50):13144-13147
  15. 15. Wu KY, Chen SK, Wu JM. Superconducting in equal molar NbTaTiZr-based high-entropy alloys. Natural Science. 2018;10(03):110
  16. 16. Kim G, Lee MH, Yun JH, Rawat P, Jung SG, Choi W, et al. Strongly correlated and strongly coupled s-wave superconductivity of the high entropy alloy Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6 compound. Acta Materialia. 2020;186:250-256
  17. 17. Kim J, Jung SG, Han Y, Kim JH, Rhyee JS, Yeo S, et al. Thermal-driven gigantic enhancement in critical current density of high-entropy alloy superconductors. Journal of Materials Science and Technology. 2024;189:60-67
  18. 18. Idczak R, Nowak W, Rusin B, Topolnicki R, Ossowski T, Babij M, et al. Enhanced superconducting critical parameters in a new high-entropy alloy Nb0.34Ti0.33Zr0.14Ta0.11Hf0.08. Materials. 2023;16(17):5814
  19. 19. Jasiewicz K, Wiendlocha B, Górnicka K, Gofryk K, Gazda M, Klimczuk T, et al. Pressure effects on the electronic structure and superconductivity of (TaNb)0.67(HfZrTi)0.33 high entropy alloy. Physical Review B. 2019;100(18):184503
  20. 20. Horvat J, Soltanian S, Pan AV, Wang XL. Superconducting screening on different length scales in high-quality bulk MgB2 superconductor. Journal of Applied Physics. 2004;96(8):4342-4351
  21. 21. Kim JH, Hidayati R, Jung SG, Salawu YA, Kim HJ, Yun JH, et al. Enhancement of critical current density and strong vortex pinning in high entropy alloy superconductor Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6 synthesized by spark plasma sintering. Acta Materialia. 2022;232:117971
  22. 22. Hong VTA, Jang H, Jung SG, Han Y, Kim JH, Hidayati R, et al. Probing superconducting gap of the high-entropy alloy Ta1/6Nb2/6Hf1/6Zr1/6Ti1/6 via Andreev reflection spectroscopy. Physical Review B. 2022;106(2):024504
  23. 23. Jung SG, Han Y, Kim JH, Hidayati R, Rhyee JS, Lee JM, et al. High critical current density and high-tolerance superconductivity in high-entropy alloy thin films. Nature Communications. 2022;13(1):3373
  24. 24. Zhang X, Winter N, Witteveen C, Moehl T, Xiao Y, Krogh F, et al. Preparation and characterization of high-entropy alloy (TaNb)1−x(ZrHfTi)x superconducting films. Physical Review Research. 2020;2(1):013375
  25. 25. Pristáš G, Bačkai J, Orendáč M, Gabáni S, Košuth F, Kuzmiak M, et al. Superconductivity in medium-and high-entropy alloy thin films: Impact of thickness and external pressure. Physical Review B. 2023;107(2):024505
  26. 26. Shu R, Zhang X, Rao SG, Le Febvrier A, Eklund P. Effects of alloying and deposition temperature on phase formation and superconducting properties of TiZrTaNb-based high entropy-alloy films. Applied Physics Letters. 2022;120(15):151901
  27. 27. Marik S, Varghese M, Sajilesh KP, Singh D, Singh RP. Superconductivity in equimolar Nb-Re-Hf-Zr-Ti high entropy alloy. Journal of Alloys and Compounds. 2018;769:1059-1063
  28. 28. Motla K, Meena PK, Singh D, Biswas PK, Hillier AD, Singh RP. Superconducting and normal-state properties of the high-entropy alloy Nb-Re-Hf-Zr-Ti investigated by muon spin relaxation and rotation. Physical Review B. 2022;105(14):144501
  29. 29. Hattori T, Watanabe Y, Nishizaki T, Hiraoka K, Kakihara M, Hoshi K, et al. Metallurgy, superconductivity, and hardness of a new high-entropy alloy superconductor Ti-Hf-Nb-Ta-Re. Journal of Alloys and Metallurgical Systems. 2023;3:100020
  30. 30. Kitagawa J, Hoshi K, Kawasaki Y, Koga R, Mizuguchi Y, Nishizaki T. Superconductivity and hardness of the equiatomic high-entropy alloy HfMoNbTiZr. Journal of Alloys and Compounds. 2022;924:166473
  31. 31. Zeng L, Zhan J, Boubeche M, Li K, Li L, Yu P, et al. Superconductivity in the bcc-type high-entropy alloy TiHfNbTaMo. Advanced Quantum Technologies. 2023;6:2300213
  32. 32. Sobota P, Topolnicki R, Ossowski T, Pikula T, Pikul A, Idczak R. Superconductivity in the high-entropy alloy (NbTa)0.67(MoHfW)0.33. Physical Review B. 2022;106(18)
  33. 33. Ishizu N, Kitagawa J. New high-entropy alloy superconductor Hf21Nb25Ti15V15Zr24. Results in Physics. 2019;13:102275
  34. 34. Sarkar NK, Prajapat CL, Ghosh PS, Garg N, Babu PD, Wajhal S, et al. Investigations on superconductivity in an equi-atomic disordered Hf-Nb-Ta-Ti-V high entropy alloy. Intermetallics (Barking). 2022;144:107503
  35. 35. Nelson WL, Chemey AT, Hertz M, Choi E, Graf DE, Latturner S, et al. Superconductivity in a uranium containing high entropy alloy. Scientific Reports. 2020;10(1):4717
  36. 36. Von Rohr FO, Cava RJ. Isoelectronic substitutions and aluminium alloying in the Ta-Nb-Hf-Zr-Ti high-entropy alloy superconductor. Physical Review Materials. 2018;2(3):034801
  37. 37. Krnel M, Jelen A, Vrtnik S, Luzar J, Gačnik D, Koželj P, et al. The effect of scandium on the structure, microstructure and superconductivity of equimolar Sc-Hf-Nb-Ta-Ti-Zr refractory high-entropy alloys. Materials. 2022;15(3):1122
  38. 38. Harayama Y, Kitagawa J. Superconductivity in Al-Nb-Ti-V-Zr multicomponent alloy. Journal of Superconductivity and Novel Magnetism. 2021;34(11):2787-2794
  39. 39. Zherebtsov S, Yurchenko N, Panina E, Tikhonovsky M, Stepanov N. Gum-like mechanical behavior of a partially ordered Al5Nb24Ti40V5Zr26 high entropy alloy. Intermetallics (Barking). 2020;116:106652
  40. 40. Zhang X, Eklund P, Shu R. Superconductivity in (TaNb)1-x(ZrHfTi)xMoy high-entropy alloy films. Applied Physics Letters. 2023;123(5):051902
  41. 41. Stolze K, Tao J, Von Rohr FO, Kong T, Cava RJ. Sc–Zr–Nb–Rh–Pd and Sc–Zr–Nb–Ta–Rh–Pd high-entropy alloy superconductors on a CsCl-type lattice. Chemistry of Materials. 2018;30(3):906-914
  42. 42. Pan Y, He X, Zhou B, Strong D, Zhang J, Bin YH, et al. Elastic properties of a Sc–Zr–Nb–Ta–Rh–Pd high-entropy alloy superconductor. Materials Today Communications. 2022;33:104265
  43. 43. Stolze K, Cevallos FA, Kong T, Cava RJ. High-entropy alloy superconductors on an α-Mn lattice. Journal of Materials Chemistry C. 2018;6(39):10441-10449
  44. 44. Liu B, Wu J, Cui Y, Zhu Q, Xiao G, Wu S, et al. Structural evolution and superconductivity tuned by valence electron concentration in the Nb-Mo-Re-Ru-Rh high-entropy alloys. Journal of Materials Science and Technology. 2021;85:11-17
  45. 45. Marik S, Motla K, Varghese M, Sajilesh KP, Singh D, Breard Y, et al. Superconductivity in a new hexagonal high-entropy alloy. Physical Review Materials. 2019;3(6):060602
  46. 46. Lee YS, Cava RJ. Superconductivity in high and medium entropy alloys based on MoReRu. Physica C: Superconductivity and its Applications. 2019;566:1353520
  47. 47. Liu B, Wu J, Cui Y, Zhu Q, Xiao G, Wu S, et al. Superconductivity in hexagonal Nb-Mo-Ru-Rh-Pd high-entropy alloys. Scripta Materialia. 2020;182:109-113
  48. 48. Hirai D, Uematsu N, Saitoh K, Katayama N, Takenaka K. Superconductivity in high-entropy antimonide M1–x Pt x Sb (M = equimolar Ru, Rh, Pd, and Ir). Inorganic Chemistry. 2023;62(35):14207-14215
  49. 49. Zhu Q, Xiao G, Cui Y, Yang W, Song S, Cao GH, et al. Structural transformation and superconductivity in carbon-added hexagonal high-entropy alloys. Journal of Alloys and Compounds. 2022;909:164700
  50. 50. Liu B, Wu J, Cui Y, Zhu Q, Xiao G, Wu S, et al. Superconductivity and paramagnetism in Cr-containing tetragonal high-entropy alloys. Journal of Alloys and Compounds. 2021;869:159293
  51. 51. Xiao G, Zhu Q, Yang W, Cui Y, Song S, Cao GH, et al. Centrosymmetric to noncentrosymmetric structural transformation in a superconducting high-entropy alloy due to carbon addition. Science China Materials. 2023;66(1):257-263
  52. 52. Xiao G, Yang W, Zhu Q, Song S, Cao GH, Ren Z. Superconductivity with large upper critical field in noncentrosymmetric Cr-bearing high-entropy alloys. Scripta Materialia. 2023;223:115099
  53. 53. Motla K, Soni V, Meena PK, Singh RP. Boron based new high entropy alloy superconductor Mo0.11W0.11V0.11Re0.34B0.33. Superconductor Science and Technology. 2022;35(7):074002
  54. 54. Kasem MR, Yamashita A, Goto Y, Matsuda TD, Mizuguchi Y. Synthesis of high-entropy-alloy-type superconductors (Fe,Co,Ni,Rh,Ir)Zr2 with tunable transition temperature. Journal of Materials Science. 2021;56(15):9499-9505
  55. 55. Liu B, Wu J, Cui Y, Zhu Q, Xiao G, Wu S, et al. Superconductivity in cubic A15-type V–Nb–Mo–Ir–Pt high-entropy alloys. Frontiers of Physics. 2021;9
  56. 56. Yamashita A, Matsuda TD, Mizuguchi Y. Synthesis of new high-entropy alloy-type Nb3 (Al, Sn, Ge, Ga, Si) superconductors. Journal of Alloys and Compounds. 2021;868:159233
  57. 57. Nakahira Y, Kiyama R, Yamashita A, Itou H, Miura A, Moriyoshi C, et al. Tuning of upper critical field in a vanadium-based A15 superconductor by the compositionally-complex-alloy concept. Journal of Materials Science. 2022;6:15990-15998
  58. 58. Mizuguchi Y, Kasem MR, Matsuda TD. Superconductivity in CuAl2-type Co0.2Ni0.1Cu0.1Rh0.3Ir0.3Zr2 with a high-entropy-alloy transition metal site. Materials Research Letters. 2021;9(3):141-147
  59. 59. Kasem MR, Yamashita A, Hatano T, Sakurai K, Oono-Hori N, Goto Y, et al. Anomalous broadening of specific heat jump at Tc in high-entropy-alloy-type superconductor TrZr2. Superconductor Science and Technology. 2021;34:125001
  60. 60. Pugliese GM, Tortora L, Tomassucci G, Kasem RM, Mizokawa T, Mizuguchi Y, et al. Possible local order in the high entropy TrZr2 superconductors. Journal of Physics and Chemistry of Solids. 2023;174:111154
  61. 61. Fujita Y, Kinami K, Hanada Y, Nagao M, Miura A, Hirai S, et al. Growth and characterization of ROBiS2 high-entropy superconducting single crystals. ACS Omega. 2020;5(27):16819-16825
  62. 62. Zeng L, Wang Z, Song J, Lin G, Guo R, Luo SC, et al. Discovery of the high-entropy carbide ceramic topological superconductor candidate (Ti0.2Zr0.2Nb0.2Hf0.2Ta0.2)C. Advanced Functional Materials. 2023;33:2301929
  63. 63. Shu H, Zhong W, Feng J, Zhao H, Yue B. Coexistence of superconductivity and ferromagnetism in high entropy carbide ceramics. arXiv:2307.16438
  64. 64. Mizuguchi Y. Superconductivity in high-entropy-alloy telluride AgInSnPbBiTe5. Journal of the Physical Society of Japan. 2019;88:124708
  65. 65. Kasem MR, Ishii R, Katase T, Miura O, Mizuguchi Y. Tuning of carrier concentration and superconductivity in high-entropy-alloy-type metal telluride (AgSnPbBi)(1-x)/4InxTe. Journal of Alloys and Compounds. 2022;920:166013
  66. 66. Ying T, Yu T, Shiah YS, Li C, Li J, Qi Y, et al. High-entropy van der Waals materials formed from mixed metal dichalcogenides, halides, and phosphorus trisulfides. Journal of the American Chemical Society. 2021;143(18):7042-7049
  67. 67. Liu B, Yang W, Xiao G, Zhu Q, Song S, Cao GH, et al. High-entropy silicide superconductors with W5Si3-type structure. Physical Review Materials. 2023;7:014805
  68. 68. Sogabe R, Goto Y, Mizuguchi Y. Superconductivity in REO0.5F0.5BiS2 with high-entropy-alloy-type blocking layers. Applied Physics Express. 2018;11:053102
  69. 69. Sogabe R, Goto Y, Abe T, Moriyoshi C, Kuroiwa Y, Miura A, et al. Improvement of superconducting properties by high mixing entropy at blocking layers in BiS2 -based superconductor REO0.5F0.5BiS2. Solid State Communications. 2019;295:43-49
  70. 70. Fujita Y, Nagao M, Miura A, Urushihara D, Mizuguchi Y, Maruyama Y, et al. Effects of equivalent composition on superconducting properties of high-entropy REOBiS2 (RE = La, Ce, Pr, Nd, Sm, Gd) single crystals. Physica C: Superconductivity and its Applications. 2023;608:1354254
  71. 71. Rost CM, Sachet E, Borman T, Moballegh A, Dickey EC, Hou D, et al. Entropy-stabilized oxides. Nature Communications. 2015;6(1):8485
  72. 72. Oses C, Toher C, Curtarolo S. High-entropy ceramics. Nature Reviews Materials. Nature Research. 2020;5:295-309
  73. 73. Akrami S, Edalati P, Fuji M, Edalati K. High-entropy ceramics: Review of principles, production and applications. Materials Science and Engineering R. 2021;146:100644
  74. 74. Zhang RZ, Reece MJ. Review of high entropy ceramics: Design, synthesis, structure and properties. Journal of Materials Chemistry A. Royal Society of Chemistry. 2019;7:22148-22162
  75. 75. Zhang J, Xu B, Xiong Y, Ma S, Wang Z, Wu Z, et al. Design high-entropy carbide ceramics from machine learning. npj Computational Materials. 2022;8(1):5
  76. 76. Kasem MR, Nakahira Y, Yamaoka H, Matsumoto R, Yamashita A, Ishii H, et al. Robustness of superconductivity to external pressure in high-entropy-alloy-type metal telluride AgInSnPbBiTe5. Scientific Reports. 2022;12:7789
  77. 77. Mizuguchi Y, Demura S, Deguchi K, Takano Y, Fujihisa H, Gotoh Y, et al. Superconductivity in novel BiS 2-based layered superconductor LaO1-xFxBiS2. Journal of the Physical Society of Japan. 2012;81:114725
  78. 78. Bérardan D, Franger S, Dragoe D, Meena AK, Dragoe N. Colossal dielectric constant in high entropy oxides. Physica Status Solidi - Rapid Research Letters. 2016;10(4):328-333
  79. 79. Sarkar A, Wang Q, Schiele A, Chellali MR, Bhattacharya SS, Wang D, et al. High-entropy oxides: Fundamental aspects and electrochemical properties. Advanced Materials. 2019;31:1806236
  80. 80. Li H, Zhou Y, Liang Z, Ning H, Fu X, Xu Z, et al. High-entropy oxides: Advanced research on electrical properties. Coatings. 2021;11:628
  81. 81. Brahlek M, Gazda M, Keppens V, Mazza AR, McCormack SJ, Mielewczyk-Gryń A, et al. What is in a name: Defining ‘high entropy’ oxides. APL Materials. 2022;10:110902
  82. 82. Aamlid SS, Oudah M, Rottler J, Hallas AM. Understanding the role of entropy in high entropy oxides. Journal of the American Chemical Society. American Chemical Society. 2023;145:5991-6006
  83. 83. Mazza AR, Gao X, Rossi DJ, Musico BL, Valentine TW, Kennedy Z, et al. Searching for superconductivity in high entropy oxide Ruddlesden–Popper cuprate films. Journal of Vacuum Science & Technology A. 2022;40:013404
  84. 84. Wang K, Hou Q, Pal A, Wu H, Si J, Chen J, et al. Structural and physical properties of high-entropy REBa2Cu3O7-δ oxide superconductors. Journal of Superconductivity and Novel Magnetism. 2021;34(5):1379-1385
  85. 85. Yamashita A, Shukunami Y, Mizuguchi Y. Improvement of critical current density of RE Ba2Cu3O7-δ by increase in configurational entropy of mixing. Royal Society Open Science. 2022;9(3):211874
  86. 86. Suzuki Y, Nagao M, Maruyama Y, Watauchi S, Tanaka I. Growth of REBa2Cu3Ox single-crystal whiskers utilizing the concept of high-entropy alloys. Japanese Journal of Applied Physics. 2023;62:033001
  87. 87. Yamashita A, Hashimoto K, Suzuki S, Nakanishi Y, Miyata Y, Maeda T, et al. Fabrication of high-entropy REBa2Cu3O7− δ thin films by pulsed laser deposition. Japanese Journal of Applied Physics. 2022;61(5):050905
  88. 88. Chen J, Huang R, Zhou X, Zhou D, Li M, Bai C, et al. Nucleation and epitaxy growth of high-entropy REBa2Cu3O7–δ (RE= Y, Dy, Gd, Sm, Eu) thin films by metal organic deposition. Journal of Rare Earths. 2023;41(7):1091-1098
  89. 89. Masuda H, Ishii R, Kita R, Miura O. Superconducting properties of high-entropy type RE123 thin films by fluorine-free MOD method. IEEE Transactions on Applied Superconductivity. 2023;33(5):7200203
  90. 90. Cayado P, Grünewald L, Erbe M, Hänisch J, Gerthsen D, Holzapfel B. Critical current density improvement in CSD-grown high-entropy REBa2Cu3O7−δ films. RSC Advances. 2022;12(44):28831-28842
  91. 91. Pryanichnikov S, Vidmid’ L, Titova S. High-entropy superconducting oxides (Y,Nd,Eu,Sm,Ho)Ba2Cu3Oy with different oxygen contents. Journal of Superconductivity and Novel Magnetism. 2023;36(3):871-875
  92. 92. Grünewald L, Cayado P, Erbe M, Hänisch J, Holzapfel B, Gerthsen D. Analytical electron microscopy study of the composition of BaHfO3 nanoparticles in REBCO films: The influence of rare-earth ionic radii and REBCO composition. Materials Advances. 2023;4:6507-6521

Written By

Pablo Cayado and Jens Hänisch

Submitted: 22 December 2023 Reviewed: 01 February 2024 Published: 16 July 2024