• Matter and Radiation at Extremes
  • Vol. 7, Issue 5, 058403 (2022)
Evgeny F. Talantseva)
Author Affiliations
  • M. N. Miheev Institute of Metal Physics, Ural Branch, Russian Academy of Sciences, 18, S. Kovalevskoy St., Ekaterinburg 620108, Russia and NANOTECH Centre, Ural Federal University, 19 Mira St., Ekaterinburg 620002, Russia
  • show less
    DOI: 10.1063/5.0091446 Cite this Article
    Evgeny F. Talantsev. Universal Fermi velocity in highly compressed hydride superconductors[J]. Matter and Radiation at Extremes, 2022, 7(5): 058403 Copy Citation Text show less

    Abstract

    The Fermi velocity vF is one of the primary characteristics of any conductor, including any superconductor. For conductors at ambient pressure, several experimental techniques have been developed to measure vF, and, for instance, Zhou et al. [Nature 423, 398 (2003)] reported that high-Tc cuprates exhibited a universal nodal Fermi velocity vF,univ=2.7±0.5×105 m/s. However, there have been no measurements of vF in highly compressed near-room-temperature superconductors (NRTS), owing to experimental challenges. Here, to answer the question of the existence of a universal Fermi velocity in NRTS materials, we analyze the full inventory of data on the ground-state upper critical field Bc2(0) for these materials and find that this class of superconductors exhibits a universal Fermi velocity vF,univ=1/1.3×2Δ0/kBTc×105 m/s, where Δ(0) is the ground-state amplitude of the energy gap. The ratio 2Δ0/kBTc varies within a narrow range 3.22Δ0/kBTc5, and so vF,univ in NRTS materials lies in the range 2.5 × 105 m/s ≤ vF,univ ≤ 3.8 × 105 m/s, which is similar to the range of values found for the high-Tc cuprate counterparts of these materials.

    I. INTRODUCTION

    Since pivotal experimental discovery of the first near-room-temperature superconductor (NRTS) H3S by Drozdov et al.,1 nearly two dozen highly compressed hydrogen-rich superconducting phases have been synthesized in binary and ternary systems.2–17 Experimental studies of NRTS are well supported by first-principles calculations,18–30 but experimental characterizations of NRTS phases are limited by the narrow range of techniques that are available for materials inside diamond anvil cells (DACs).25–27 These comprise x-ray diffraction (XRD) phase analysis, Raman spectroscopy, and magnetoresistance measurements,31–35 although, with the use of some advanced techniques, Hall effect measurements can also be performed.31 Because of this, only two characteristic values of the superconducting state of the NRTS phases are commonly extracted from experimental data, namely, the transition temperature Tc and the extrapolated value for the ground-state upper critical field Bc20 or the ground-state superconducting coherence length ξ0, which can be derived from the Ginzburg–Landau36 expressionξ0=ϕ02πBc20,where ϕ0 = h/2e is the superconducting flux quantum, with h being Planck’s constant and e the electric charge of the electron.

    Other important parameters of NRTS materials, among which we can mention the Fermi velocity vF, have not been measured to date, owing to the challenges of performing such measurements on samples inside DACs. However, considering that all NRTS superconductors are hydrides, there is an expectation that these materials will exhibit a universal Fermi velocity vF,univ, as has been discovered in cuprates, for which vF,univ=2.7±0.5×105 m/s, as reported by Zhou et al.37 (see Fig. 1).

    Universal nodal Fermi velocity vF,univ=2.7±0.5×105 m/s for cuprate superconductors. These are the raw data reported by Zhou et al.37 for (La2−xSrx)CuO4 (LSCO), (La2−x−yNdySrx)CuO4 (Nd-LSCO), Bi2Sr2CaCu2O8 (Bi-2212), Bi2Sr2CuO6 (Bi-2201), (Ca2−xNax)CuO2Cl2 (Na-CCOC), and Tl2Ba2CuO6 (Tl-2201).

    Figure 1.Universal nodal Fermi velocity vF,univ=2.7±0.5×105 m/s for cuprate superconductors. These are the raw data reported by Zhou et al.37 for (La2−xSrx)CuO4 (LSCO), (La2−xyNdySrx)CuO4 (Nd-LSCO), Bi2Sr2CaCu2O8 (Bi-2212), Bi2Sr2CuO6 (Bi-2201), (Ca2−xNax)CuO2Cl2 (Na-CCOC), and Tl2Ba2CuO6 (Tl-2201).

    The theoretical motivation for the quest for a universal Fermi velocity in NRTS comes, on the one hand, from the recent understanding38 that sulfur in H3S is analogous to the oxygen in cuprates, and, on the other hand, from the fact that highly compressed hydrides fit nicely with the main global scaling laws for superconductors.39–43 However, it should be noted that an analysis of whether these materials also comply with other scaling laws for superconductors44,45 requires more experimental data on normal-state resistivity ρT31,46–48 and ground-state London penetration depth λ0.1,49,50

    Here, we report the results of our search for a universal Fermi velocity in NRTS materials, based on an analysis of the full inventory of values for the ground-state upper critical field Bc20 in these materials. We find that a universal Fermi velocity, vF,univ, does indeed exist in NRTS materials and obeys the empirical lawvF,univ=11.3×2Δ0kBTc×105m/s,where kB is Boltzmann’s constant and Δ0 is the ground-state superconducting energy gap.

    II. DESCRIPTION OF APPROACH

    In the Bardeen–Cooper–Schieffer (BCS) theory of superconductivity,51 the ground-state coherence length ξ0 and the amplitude of the ground-state energy gap Δ0 are linked through the expressionξ0=vFπΔ0,where is the reduced Planck’s constant. BCS theory also involves a dimensionless ratioα=2Δ0kBTc.

    Substitution of Eqs. (3) and (4) into Eq. (1) gives the dependence of the ground-state upper critical field on the transition temperature:Bc20=πϕ0kB282α2vF2Tc2,where the multiplicative prefactor in square brackets is a constant:A=πϕ0kB282=1.38×107Tm2/(s2K2).

    Thus, if hydrogen-rich superconductors exhibit a universal Fermi velocity vF,univ, then a fit of the full inventory of the Bc20 vs Tc dataset to the equationBc20=AfTcβ,where β and f=α2/vF2 are free fitting parameters, should reveal thatβ2,and, if this is the case, then the universal Fermi velocity vF,univ can be calculated from the deduced free-fitting parameter f asvF,univ=αf=1f2Δ0kBTc.

    It should be noted that α=2Δ0/kBTc in highly compressed hydrogen-rich superconductors varies within the range4,8,12,27,42,49,52–553.22Δ0kBTc5,where the lower limit is the value deduced from experiment42,50,52 and the upper limit is based on the many results obtained from first-principles calculations, which always predict 4.32Δ0/kBTc in NRTS materials,4,8,12,27,53–55 including very high values of 2Δ0/kBTc5.0 for some NRTS phases.4,8,12,27

    III. EXTRAPOLATION MODEL FOR THE GROUND-STATE UPPER CRITICAL FIELD

    Equation (7) has the ground-state upper critical field Bc20 as dependent variable. However, it is important to note that this value can be determined by the use of several extrapolative models56–59 that utilize experimental Bc2T data measured at high reduced temperatures T/Tc. The primary reason why there is a necessity for extrapolative models is that all highly compressed hydrogen-rich superconductors have Bc2T0K>20T, which cannot be measured by the conventional and widely used Physical Property Measurement System (manufactured by Quantum Design), for which the highest measurable magnetic field Bappl = 9–16 T (depending on the specific model). It should be also stressed that Bc2T0K for the NRTS compounds H3S, LaH10, YH6/YH9 and (La,Y)H10 are so high that even measurements at the best available quasi-DC magnetic field facilities worldwide31,48,60 cover only the range of reduced temperatures 12T/Tc.

    In this paper, from the several extrapolative Bc2T models that are available,56–59 we use the following analytical approximate expression from Werthamer–Helfand–Hohenberg (WHH) theory,61,62 which was proposed by Baumgartner et al.:59Bc2T=10.693ϕ02πξ201TTc0.1531TTc20.1521TTc4,where ξ(0) and TcTc(B = 0) are two free fitting parameters (this equation is referred to as the B-WHH model hereinafter). Equation (11) was originally proposed for the extrapolation of Bc2T data for neutron-irradiated Nb3Sn alloys,59 and recently several research groups have found that provides good approximations for a variety of superconducting materials.4,63–68 On this basis, in the present study, we used Eq. (11) as a good, robust and simple analytical tool to extrapolate the Bc2(T) curve to the low-temperature high-field region.

    It is a necessary to describe the criterion for extracting Bc2(T) datasets from experimentally measured R(T, Bappl) curves. Several criteria are available for the definitions of Tc, Bc2(T), and Tc(Bappl), which have been discussed recently for the case of NRTS in Ref. 69. We have found69,70 that the best match between the electron–phonon coupling constant λe–ph extracted from R(T, Bappl = 0) curves and the λe–ph computed by first-principles calculation is obtained when Tc is defined at a value of the ratio R(T)/Rnorm that is as low as practically possible (where Rnorm is the normal-state resistance just above the transition). By analyzing the full inventory of R(T, Bappl) data for NRTS materials herein, we have come to the conclusion that owing to noise and slope issues with real-world R(T, Bappl) curves and the fact that highly compressed superhydrides contain several superconducting phases, the appropriate criterion isRT,BapplRTconset,Bappl=0.05,and we use this henceforth in this study.

    IV. RESULTS

    A. Unannealed highly compressed sulfur hydride

    In the first paper on NRTS superconductors, Drozdov et al.1 reported R(T, Bappl) data for unannealed highly compressed sulfur hydride (P = 155 GPa) in their Fig. 3(a). By using the criterion of Eq. (12) [which is RT,Bapplcriterion=23mΩ for the R(T, Bappl) curves shown in bottom insert in Fig. 3(a) in Ref. 1], we extracted the Bc2(T) dataset for this sample, which is shown in Fig. 2. Because this Bc2(T) dataset covers a significant part of the full temperature range 0 K < TTc, there was no need to use an extrapolative fit, and instead we fitted this dataset to the model in Ref. 52, which allowed to deduce Δ0, 2Δ0/kBTc, and ΔC/γTc (the last of which is the relative jump in electronic specific heat at Tc, with γ being the so-called Sommerfeld constant):Bc2T=ϕ02πξ201.770.43TTc2+0.07TTc41.772×112kBT0dεcosh2ε2+Δ2T2kBT,where the temperature-dependent superconducting gap Δ(T) is given by71,72ΔT=Δ0tanhπkBTcΔ0ηΔCγTcTcT1,with η = 2/3 for s-wave superconductors.

    Upper critical field data Bc2(T) and data fit to Eqs. (13) and (14) for unannealed highly compressed sulfur hydride (P = 190 GPa). The raw R(T, Bappl) dataset was that reported by Drozdov et al.1 The deduced values of ξ0, Δ0, Tc, and ΔC/γTc are shown on the figure. The 95% confidence bands are shown by the pink shaded area. The fit quality is R = 0.9985.

    Figure 2.Upper critical field data Bc2(T) and data fit to Eqs. (13) and (14) for unannealed highly compressed sulfur hydride (P = 190 GPa). The raw R(T, Bappl) dataset was that reported by Drozdov et al.1 The deduced values of ξ0, Δ0, Tc, and ΔC/γTc are shown on the figure. The 95% confidence bands are shown by the pink shaded area. The fit quality is R = 0.9985.

    We used Eqs. (13) and (14) to extract ξ0, Δ0, Tc, and ΔC/γTc from Bc2(T) datasets for a variety of superconductors, including two highly compressed hydride phases of H3S,52 SnH12,42 V3Si,73 and several iron-based superconductors.73 However, it should be stressed that the approach using Eqs. (13) and (14) is only applicable for Bc2(T) datasets defined by Eq. (12) or by a stricter criterion.

    One of the most important deduced parameters, α=2Δ0/kBTc=3.2±0.3, is in remarkable agreement with the corresponding values deduced for highly compressed annealed H3S (P = 155–160 GPa), α = 3.20 ± 0.0249 and 3.55 ± 0.31,52 and for highly compressed annealed SnH12 (P = 190 GPa), α = 3.28 ± 0.18.42 The deduced ΔC/γTc = 0.7 ± 0.1 is also below the weak-coupling limit of BCS theory ΔC/γTc = 1.43, as is the corresponding value for the annealed H3S material, ΔC/γTc = 1.2 ± 0.3.52 It should be mentioned that to deduce ΔC/γTc with higher accuracy requires more Bc2(T) data points, especially at TTc. The deduced Bc20 and Tc are given in Table I.

    Phase and data sourceFiguresPressure (GPa)Tc (K)ΔTc (K)Bc2(0) (T)ΔBc2(0) (T)
    Unannealed sulfur hydride [Fig. 3(a) in Ref. 1]215513.90.36.30.4
    Annealed H3S (Fig. 3 in Ref. 31)S1(a)155185298.81.2
    Annealed H3S (Figs. S1 and S2 in Ref. 31)S1(b)155196.10.671.11.1
    Annealed H3S (Fig. 3 in Ref. 31)S1(c)160143.91.459.22.3
    Annealed CeH9 [Fig. S7(a) in Ref. 12], coolingS2(a)8838.80.416.51
    Annealed CeH9 [Fig. 1(c) in Ref. 12], warmingS2(b)13988.60.322.20.7
    Annealed CeH9 [Fig. 1(d) in Ref. 12], coolingS2(c)13781.90.718.40.7
    Annealed CeH9 [Fig. 1(d) in Ref. 12], warmingS2(d)13782.70.718.70.6
    Annealed LaH10 [Fig. 3(a) in Ref. 60]S3(a)120174.80.8903
    Annealed LaH10 [Fig. 3(b) in Ref. 60]S3(b)136206.20.81363
    Annealed YD6 [Fig. S13(a) in Ref. 4]S4(a)172157.70.2124.92.4
    Annealed YH6 [Fig. S16(c) in Ref. 4]S4(b)200206.20.297.21.4
    Annealed (La,Y)H10 [Fig. S27(b) in Ref. 8]S5(a)183203.50.2101.61.8
    Annealed (La,Y)H10 [Fig. S28(a) in Ref. 8]S5(b)1822340.1135.81.5
    Annealed (La,Y)H10 [Fig. S28(a) in Ref. 8]S5(c)186234.50.11341
    Annealed SnH12 [Fig. 4(a) in Ref. 11], coolingS6(a)19062.80.490.2
    Annealed SnH12 [Fig. 4(a) in Ref. 11], warmingS6(b)19064.10.58.90.2
    Annealed ThH9 [Fig. 4(a) in Ref. 16]S7(a)170151.21.5320.9
    Annealed ThH10 [Fig. 4(a) in Ref. 16]S7(b)170150.60.443.40.6
    Th4H15 (Ref. 74)Ambient8.20.152.750.25

    Table 1. Deduced Bc20 and Tc values for hydrogen-rich superconductors for which raw R(T, Bappl) data are available to date.

    B. Annealed highly compressed hydrides

    We processed reported R(T, Bappl) datasets for several annealed highly compressed hydrides by using Eq. (12) to extract Bc2(T) datasets. The obtained datasets were fitted to Eq. (11), and the deduced values are given in Table I. These materials are as follows:Sulfur superhydride H3S (P = 155 and 160 GPa), for which the raw data were reported by Mozaffari et al.31Cerium superhydride CeHn (P = 88, 137, and 139 GPa), for which the raw data were reported by Chen et al.12Lanthanum superhydride LaH10 (P = 120, 136 GPa), for which the raw data were reported by Sun et al.60Yttrium superhydride/superdeuteride YH6/YD6 (P = 172 and 200 GPa), for which the raw data were reported by Troyan et al.4Lanthanum–yttrium superhydride (La,Y)H10 (P = 182, 183, and 186 GPa), for which the raw data were reported by Semenok et al.8Tin superhydride SnH12 (P = 190 GPa), for which the raw data were reported by Hong et al.11Thorium superhydrides ThH9 and ThH10 (P = 170 GPa), for which the raw data were reported by Semenok et al.16

    The respective fits are shown in Figs. S1–S7 in the supplementary material.

    C. Analysis of Bc2(0) vs Tc for superhydride phases

    All deduced Bc20 and Tc values for superhydride phases are collected in Table I, where we have also added data for the Th4H15 phase reported by Satterthwaite and Toepke.74

    The full dataset from Table I is shown in Fig. 3, together with the fit to Eq. (7). Although this dataset has a large scatter, it can be seen in Fig. 3(a) that the free-fitting power-law exponent β = 2.07 ± 0.14 is practically undistinguishable from the expected value β ≡ 2 [Eq. (5)]. When β is the free-fitting parameter [Fig. 3(a)], the deduced f=1.19±0.90×1010 s2/m2 has a large uncertainty. However, when β is fixed to 2 [Fig. 3(b)], the free-fitting parameter f can be deduced with high accuracy asf=α2vF,univ2=1.68±0.08×1010s2/m2,from which we can obtainvF,univ=α1.30±0.03×105m/s11.3×2Δ0kBTc×105m/s.

    Total Bc2(0) vs Tc dataset for hydrogen-rich superconductors deduced in this work (Table I) and data fits to (a) Eq. (7) and (b) Eq. (5). (a) Free-fitting β = 2.07 ± 0.14 and f=1.19±0.90×10−10 s2/m2; the fit quality is R = 0.9361. (b) β = 2.0 (fixed) and free-fitting f=1.68±0.08×10−10 s2/m2; the fit quality is R = 0.9354.

    Figure 3.Total Bc2(0) vs Tc dataset for hydrogen-rich superconductors deduced in this work (Table I) and data fits to (a) Eq. (7) and (b) Eq. (5). (a) Free-fitting β = 2.07 ± 0.14 and f=1.19±0.90×1010 s2/m2; the fit quality is R = 0.9361. (b) β = 2.0 (fixed) and free-fitting f=1.68±0.08×1010 s2/m2; the fit quality is R = 0.9354.

    By substituting the lower [2Δ0/kBTc=3.2] and upper [2Δ0/kBTc=5.0] limits of the ratio 2Δ0/kBTc [Eq. (10)] into Eq. (16), we can establish the lower and upper limits of vF,univ in superhydrides:2.5×105m/svF,univ3.8×105m/s.

    The deduced vF,univ for hydrogen-rich superconductors [Eq. (16)] is similar to the corresponding value for high-Tc cuprates, vF,univ=2.7±0.5×105 m/s,37 if Eq. (10) is taken into account.

    V. DISCUSSION

    The primary assumption of the Migdal–Eliashberg (ME) theory of electron–phonon-mediated superconductivity75–77 is that the ratio of characteristic phonon energy ℏωD (where ωD is the Debye frequency) to the Fermi energy EF is very small, ℏωD/EF ≪ 1. In normal metals ℏωD/EF ≲ 10−2,78–82 and this is why ME theory is quantitatively accurate. However, for many high-temperature superconductors, the application of ME theory cannot be justified. In fact, in our previous studies,69,70 we deduced the Debye temperature Tθ ≅ 1500 K, in highly compressed H3S from a fit of experimentally measured temperature-dependent resistance R(T) to the Bloch–Grüneisen equation.83,84 This temperature can be converted into the Debye energy kBTθ = ℏωD ≅ 0.13 eV, and, considering that the Fermi energy was deduced in our previous study52 as EF = 0.5–1.0 eV, we can conclude that 0.13 ≤ ℏωD/EF ≤ 0.26, and thus ME theory75,76 does not provide an exact description of highly compressed H3S.

    The first concern that nonadiabatic effects (i.e., effects beyond ME theory75,76) are important in highly compressed H3S was expressed by Pietronero et al.,82 who also pointed out that: “… The fingerprints of non-adiabatic effects are: – position of the material in the Uemura diagram;…” Although the traditional way to position a material in the Uemura plot requires knowledge of the ground-state London penetration depth λ0 (which has only recently been reported for H3S46), the present author utilized the ground-state coherence length ξ0 [deduced from the Bc2(T) data] and found52 that H3S falls into the unconventional superconductors band in the Uemura plot.41 In later work,42,85–88 it was established that LaH10, Th4H15, ThH9, ThH10, YH6, SH12, and H3(S,C) also fall into the unconventional superconductors band in the Uemura plot. This is direct evidence that the ratio ℏωD/EF in superhydrides has values well above those typical of conventional superconductors, ℏωD/EF ≲ 10−2.

    On the basis of the derived universal Fermi velocity in superhydrides, vF,univ [Eq. (16)], we can conclude that the strength of the nonadiabatic effects in a superhydride (as quantified by the ratio ℏωD/EF) can be revealed if the Debye temperature Tθ of the compound can be deduced from the normal part of the temperature-dependent resistance R(T).69,70,86 It should be noted that the Debye temperature in superhydrides varies from Tθ ≅ 870 K (D3S, P = 173 GPa69) up to Tθ ≅ 1700 K (R3m-phase of H3S, P = 133 GPa69).

    An analysis of the first experimental R(T) data89 measured for metallic hydrogen phase III (compressed at P = 402 GPa) revealed that Tθ ≅ 730 K.70 If we assume that metallic hydrogen phase III complies with the established vF,univ [Eq. (16)] and that it has 2Δ0/kBTc=3.53 and exhibits no effective mass enhancement, then the ratio ℏωD/EF can be estimated as ℏωD/EF = 0.3. This implies that metallic hydrogen should exhibit pronounced nonadiabatic effects,78–82 which could prevent the emergence of a superconducting state in this metal at high temperature.

    It should be also mentioned that first principles calculations (FPC) are an essential part of current NRTS phase searches.27 The accuracy and powerful capabilities of FPC became obvious after the pivotal prediction of the Im3̄mH3S phase,90,91 which was later discovered experimentally.1 Another achievement of the FPC approach was demonstrated recently when Li et al.92 and Ma et al.93 reported the discovery of a calcium superhydride phase with transition temperature Tc = 200–215 K (at a pressure P = 160–190 GPa), which was predicted by Wang et al.94 in 2012. However, it should also be mentioned that the superconductivity predicted by FPC in some binary systems (e.g., AlH395,96), has never been observed experimentally. This implies that further development of FPC techniques to take account of non-adiabatic effects is highly desirable.

    Overall, remarkable progress has been achieved in this field from the first theoretical predictions of high-temperature superconductivity in metallic hydrogen97,98 and hydrogen-dominated alloys74,99 five decades ago to the remarkable experimental and FPC results1,27 reported recently. It should be stressed that all the primary discoveries in the field (e.g., the direct searches for and syntheses of the H3S, LaH10, and YH6 phases) have come from a perfect conjunction of theory and experiment. An excellent example of this is the story of the discovery of near-room-temperature superconductivity in highly compressed sulfur hydride.1 In February 2014, Li et al.100 reported results of FPC calculations for highly compressed sulfur hydride. These calculations showed that at P = 160 GPa, the sulfur hydride retains the composition of H2S and that this compound exhibits a superconducting transition temperature of Tc ∼ 80 K. In November 2014, an alternative theoretical result was reported by Duan et al.,91 who performed thorough FPC using USPEX software101–103 and predicted that sulfur hydride would decompose into a mixture of elemental sulfur and an (H2S)2H2 phase at high pressure. This result was in a good accord with an earlier report by Strobel et al.,104 who showed experimentally that at P > 3.2 GPa, the H2S–H2 mixture exhibited structural ordering with the formation of the (H2S)2H2 phase (with four formula units per unit cell). The predicted transition temperature for the (H2S)2H2 phase was Tc = 191–204 K at P = 200 GPa.91 On December 1, 2014, Drozdov et al.105 reported a milestone experimental result on the observation of Tc ≈ 190 K in sulfur hydride compressed at P > 150 GPa.

    Another remarkable story should also be mentioned here, namely, the discovery of the Fm3̄mLaH10 phase, for which Tc ≅ 274–286 K at P = 210 GPa was theoretically predicted by Liu et al.106 in June 2017. This NRTS phase was experimentally discovered by Drozdov et al.107 on August 21, 2018, and, two days later, Somayazulu et al.108 confirmed this discovery.

    VI. CONCLUSIONS

    In this study, we have proposed that hydrogen-rich superconductors exhibit a universal Fermi velocity vF, which is given by empirical expression vF,univ=1/1.3×2Δ0/kBTc×105 m/s. Considering that the gap-to-transition temperature ratio 2Δ0/kBTc in hydrogen-rich superconductors varies within the range 3.22Δ0/kBTc5.0, we conclude that vF,univ varies within the range 2.5 × 105 m/s ≤ vF,univ ≤ 3.8 × 105 m/s.

    The Debye temperature Tθ can be deduced from the temperature-dependent resistance R(T) of the conductor,69,83,84 and so this universal Fermi velocity vF in superhydrides [Eq. (16)] can be used to calculate the ratio ℏωD/EF, which determines the strength of nonadiabatic effects in the superconductor. Calculations for metallic hydrogen phase III (compressed at P = 402 GPa) have shown that ℏωD/EF = 0.3, which implies strong nonadiabatic effects in this metal.

    SUPPLEMENTARY MATERIAL

    ACKNOWLEDGMENTS

    Acknowledgment. The author thanks Luciano Pietronero (Universita’ di Roma) for comments about the limitations of the applicability of the Migdal–Eliashberg (ME) theory of electron–phonon-mediated superconductivity to high-temperature superconductors.

    The author is grateful for financial support provided by the Ministry of Science and Higher Education of the Russian Federation through the theme “Pressure” Grant No. AAAA-A18-118020190104-3 and through a Ural Federal University project within the Priority-2030 Program.

    References

    [1] I. A.Troyan, A. P.Drozdov, M. I.Eremets, S. I.Shylin, V.Ksenofontov. Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system. Nature, 525, 73-76(2015).

    [2] A. P.Drozdov et al. Superconductivity at 250 K in lanthanum hydride under high pressures. Nature, 569, 528-531(2019).

    [3] M.Somayazulu et al. Evidence for superconductivity above 260 K in lanthanum superhydride at megabar pressures. Phys. Rev. Lett., 122, 027001(2019).

    [4] I. A.Troyan et al. Anomalous high-temperature superconductivity in YH6. Adv. Mater., 33, 2006832(2021).

    [5] P.Kong et al. Superconductivity up to 243 K in yttrium hydrides under high pressure. Nat. Commun., 12, 5075(2021).

    [6] L.Ma et al. High-temperature superconducting phase in clathrate calcium hydride CaH6 up to 215 K at a pressure of 172 GPa. Phys. Rev. Lett., 128, 167001(2022).

    [7] C. L.Zhang et al. Superconductivity above 80 K in polyhydrides of hafnium.

    [8] D. V.Semenok et al. Superconductivity at 253 K in lanthanum–yttrium ternary hydrides. Mater. Today, 48, 18-28(2021).

    [9] D.Zhou et al. Superconducting praseodymium superhydrides. Sci. Adv., 6, eaax6849(2020).

    [10] T.Matsuoka et al. Superconductivity of platinum hydride. Phys. Rev. B, 99, 144511(2019).

    [11] F.Hong et al. Possible superconductivity at ∼70 K in tin hydride SnHx under high pressure. Mater. Today Phys., 22, 100596(2022).

    [12] A. R.Oganov, X.Huang, D. V.Semenok, H.Shu, X.Li, D.Duan, W.Chen, T.Cui. High-temperature superconducting phases in cerium superhydride with a Tc up to 115 K below a pressure of 1 Megabar. Phys. Rev. Lett., 127, 117001(2021).

    [13] M.Sakata et al. Superconductivity of lanthanum hydride synthesized using AlH3 as a hydrogen source. Supercond. Sci. Technol., 33, 114004(2020).

    [14] W.Chen et al. Synthesis of molecular metallic barium superhydride: Pseudocubic BaH12. Nat. Commun., 12, 273(2021).

    [15] M. A.Kuzovnikov, M.Tkacz. High-pressure synthesis of novel polyhydrides of Zr and Hf with a Th4H15-type structure. J. Phys. Chem. C, 123, 30059-30066(2019).

    [16] D. V.Semenok et al. Superconductivity at 161 K in thorium hydride ThH10: Synthesis and properties. Mater. Today, 33, 36-44(2020).

    [17] N. N.Wang et al. A low-Tc superconducting modification of Th4H15 synthesized under high pressure. Supercond. Sci. Technol., 34, 034006(2021).

    [18] I. A.Troyan, A. P.Drozdov, M. I.Eremets. Superconductivity above 100 K in PH3 at high pressures.

    [19] M.Shao et al. Superconducting ScH3 and LuH3 at megabar pressures. Inorg. Chem., 60, 15330(2021).

    [20] J.Chen et al. Computational design of novel hydrogen-rich YS–H compounds. ACS Omega, 4, 14317-14323(2019).

    [21] J. A.Alarco, P. C.Talbot, I. D. R.Mackinnon. Identification of superconductivity mechanisms and prediction of new materials using density functional theory (DFT) calculations. J. Phys.: Conf. Ser., 1143, 012028(2018).

    [22] A. R.Oganov, A. G.Kvashnin, D. V.Semenok, I. A.Kruglov. Actinium hydrides AcH10, AcH12, and AcH16 as high-temperature conventional superconductors. J. Phys. Chem. Lett., 9, 1920-1926(2018).

    [23] I.Errea, M. I.Eremets, C. J.Pickard. Superconducting hydrides under pressure. Annu. Rev. Condens. Matter Phys., 11, 57-76(2020).

    [24] M.Eremets, J. A.Flores-Livas, G.Profeta, A.Sanna, R.Arita, L.Boeri. A perspective on conventional high-temperature superconductors at high pressure: Methods and materials. Phys. Rep., 856, 1-78(2020).

    [25] A.Goncharov. Phase diagram of hydrogen at extreme pressures and temperatures; updated through 2019 (Review article). Low Temp. Phys., 46, 97(2020).

    [26] C.Ji, H.-K.Mao, E.Gregoryanz, P.Dalladay-Simpson, B.Li, R. T.Howie. Everything you always wanted to know about metallic hydrogen but were afraid to ask. Matter Radiat. Extremes, 5, 038101(2020).

    [27] B.Lilia et al. The 2021 room-temperature superconductivity roadmap. J. Phys.: Condens. Matter, 34, 183002(2022).

    [28] G.Yang, Y.Zhao, X.Zhang. Superconducting ternary hydrides under high pressure. Wiley Interdiscip. Rev.: Comput. Mol. Sci., 12, e1582(2021).

    [29] M. L.Cohen, M.Dogan. Anomalous behaviour in high-pressure carbonaceous sulfur hydride. Physica C, 583, 1353851(2021).

    [30] T.Wang et al. Absence of conventional room temperature superconductivity at high pressure in carbon doped H3S. Phys. Rev. B, 104, 064510(2021).

    [31] S.Mozaffari et al. Superconducting phase diagram of H3S under high magnetic fields. Nat. Commun., 10, 2522(2019).

    [32] V. B.Prakapenka, V. S.Minkov, M. I.Eremets, E.Greenberg. A boosted critical temperature of 166 K in superconducting D3S synthesized from elemental sulfur and hydrogen. Angew. Chem., Int. Ed., 59, 18970-18974(2020).

    [33] R.Matsumoto et al. Electrical transport measurements for superconducting sulfur hydrides using boron-doped diamond electrodes on beveled diamond anvil. Supercond. Sci. Technol., 33, 124005(2020).

    [34] D.Laniel et al. Novel sulfur hydrides synthesized at extreme conditions. Phys. Rev. B, 102, 134109(2020).

    [35] X.Huang et al. High-temperature superconductivity in sulfur hydride evidenced by alternating-current magnetic susceptibility. Natl. Sci. Rev., 6, 713-718(2019).

    [36] L. D.Landau, V. L.Ginzburg. Z. Eksp. Teor. Fiz., 20, 1064(1950).

    [37] X. J.Zhou et al. High-temperature superconductors: Universal nodal Fermi velocity. Nature, 423, 398(2003).

    [38] D. K.Sunko. High-temperature superconductors as ionic metals. J. Supercond. Novel Magn., 33, 27-33(2020).

    [39] D. R.Harshman, A. T.Fiory. High-Tc superconductivity in hydrogen clathrates mediated by Coulomb interactions between hydrogen and central-atom electrons. J. Supercond. Novel Magn., 33, 2945-2961(2020).

    [40] A. T.Fiory, D. R.Harshman. The superconducting transition temperatures of C–S–H based on inter-sublattice S–H4-tetrahedron electronic interactions. J. Appl. Phys., 131, 015105(2022).

    [41] Y. J.Uemura. Bose-Einstein to BCS crossover picture for high-Tc cuprates. Physica C, 282-287, 194-197(1997).

    [42] E. F.Talantsev. Comparison of highly-compressed C2/m-SnH12 superhydride with conventional superconductors. J. Phys.: Condens. Matter, 33, 285601(2021).

    [43] J. L.Tallon, W. P.Crump, E. F.Talantsev. Universal scaling of the self-field critical current in superconductors: From sub-nanometre to millimetre size. Sci. Rep., 7, 10010(2017).

    [44] X.Zhao, D. N.Basov, R.Liang, G.Yu, N.Kaneko, S.Komiya, M.Strongin, T.Timusk, S. V.Dordevic, C. C.Homes, M.Greven, D. A.Bonn, W. N.Hardy, Y.Ando. A universal scaling relation in high-temperature superconductors. Nature, 430, 539-541(2004).

    [45] M. R.Koblischka, A.Koblischka-Veneva. Calculation of Tc of superconducting elements with the Roeser–Huber formalism. Metals, 12, 337(2022).

    [46] S.Mozaffari, D.Sun, P. P.Kong, S. I.Shylin, A. P.Drozdov, M. I.Eremets, V. S.Minkov, F. F.Balakirev, L.Balicas, S. L.Bud’ko, V.Ksenofontov, R.Prozorov. High-temperature superconductivity in hydrides: Experimental evidence and details. J. Supercond. Novel Magn., 35, 965-977(2022).

    [47] O.Moulding, S.Cross, O.Lord, S.Friedemann, J.Buhot, T.Muramatsu, A.Brooks, I.Osmond, T.Fedotenko. Clean-limit superconductivity in Im3̄m H3S synthesized from sulfur and hydrogen donor ammonia borane. Phys. Rev. B, 105, L220502(2022).

    [48] D. V.Semenok et al. Effect of magnetic impurities on superconductivity in LaH10. Adv. Mater.(2022).

    [49] J. L.Tallon, E. F.Talantsev, W. P.Crump, J. G.Storey. London penetration depth and thermal fluctuations in the sulphur hydride 203 K superconductor. Ann. Phys., 529, 1600390(2017).

    [50] R. J.Husband, V. S.Minkov, S. L.Bud’ko, S.Chariton, V. B.Prakapenka, H. P.Liermann, F. F.Balakirev, M. I.Eremets. Magnetic field screening in hydrogen-rich high-temperature superconductors. Nat. Commun., 13, 3194(2022).

    [51] J.Bardeen, J. R.Schrieffer, L. N.Cooper. Theory of superconductivity. Phys. Rev., 108, 1175-1204(1957).

    [52] E. F.Talantsev. Classifying superconductivity in compressed H3S. Mod. Phys. Lett. B, 33, 1950195(2019).

    [53] I.Errea et al. Quantum crystal structure in the 250-kelvin superconducting lanthanum hydride. Nature, 578, 66-69(2020).

    [54] L.Boeri, S.di Cataldo, C.Heil, G. B.Bachelet. Superconductivity in sodalite-like yttrium hydride clathrates. Phys. Rev. B, 99, 220502(R)(2019).

    [55] J. A.Camargo-Martínez et al. The higher superconducting transition temperature Tc and the functional derivative of Tc with α2F(ω) for electron–phonon superconductors. J. Phys.: Condens. Matter, 32, 505901(2020).

    [56] H.Casimir, C. J.Gorter. On supraconductivity I. Physica, 1, 306-320(1934).

    [57] C. K.Jones, B. S.Chandrasekhar, J. K.Hulm. Upper critical field of solid solution alloys of the transition elements. Rev. Mod. Phys., 36, 74(1964).

    [58] L. P.Gor’kov. The critical supercooling field in superconductivity theory. Sov. Phys. JETP, 10, 593-599(1960).

    [59] M.Eisterer, T.Baumgartner, C.Scheuerlein, L.Bottura, R.Flükiger, H. W.Weber. Effects of neutron irradiation on pinning force scaling in state-of-the-art Nb3Sn wires. Supercond. Sci. Technol., 27, 015005(2014).

    [60] D.Sun et al. High-temperature superconductivity on the verge of a structural instability in lanthanum superhydride. Nat. Commun., 12, 6863(2021).

    [61] N. R.Werthamer, E.Helfand. Temperature and purity dependence of the superconducting critical field, Hc2. II. Phys. Rev., 147, 288-294(1966).

    [62] P. C.Hohenberg, E.Helfand, N. R.Werthamer. Temperature and purity dependence of the superconducting critical field, Hc2. III. Electron spin and spin-orbit effects. Phys. Rev., 147, 295-302(1966).

    [63] H.Ninomiya et al. Superconductivity in a scandium borocarbide with a layered crystal structure. Inorg. Chem., 58, 15629-15636(2019).

    [64] H.Xie et al. Superconducting zirconium polyhydrides at moderate pressures. J. Phys. Chem. Lett., 11, 646-651(2020).

    [65] W.Zhang et al. A new superconducting 3R-WS2 phase at high pressure. J. Phys. Chem. Lett., 12, 3321-3327(2021).

    [66] M.Scuderi et al. Nanoscale analysis of superconducting Fe(Se,Te) epitaxial thin films and relationship with pinning properties. Sci. Rep., 11, 20100(2021).

    [67] K.Ma et al. Group-9 transition-metal suboxides adopting the filled-Ti2Ni structure: A class of superconductors exhibiting exceptionally high upper critical fields. Chem. Mater., 33, 8722-8732(2021).

    [68] M.Boubeche et al. Enhanced superconductivity with possible re-appearance of charge density wave states in polycrystalline Cu1-xAgxIr2Te4 alloys. J. Phys. Chem. Solids, 163, 110539(2022).

    [69] E. F.Talantsev. Advanced McMillan’s equation and its application for the analysis of highly-compressed superconductors. Supercond. Sci. Technol., 33, 094009(2020).

    [70] E. F.Talantsev. The electron–phonon coupling constant and the Debye temperature in polyhydrides of thorium, hexadeuteride of yttrium, and metallic hydrogen phase III. J. Appl. Phys., 130, 195901(2021).

    [71] F.Gross et al. Anomalous temperature dependence of the magnetic field penetration depth in superconducting UBe13. Z. Phys. B: Condens. Matter, 64, 175-188(1986).

    [72] K.Andres, P. J.Hirschfeld, B. S.Chandrasekhar, F.Gro?-Alltag, D.Einzel. London field penetration in heavy fermion superconductors. Z. Phys. B: Condens. Matter, 82, 243-255(1991).

    [73] E. F.Talantsev. In-plane p-wave coherence length in iron-based superconductors. Results Phys., 18, 103339(2020).

    [74] C. B.Satterthwaite, I. L.Toepke. Superconductivity of hydrides and deuterides of thorium. Phys. Rev. Lett., 25, 741-743(1970).

    [75] A. B.Migdal. Interaction between electrons and lattice vibrations in a normal metal. Sov. Phys. JETP, 7, 996-1001(1958).

    [76] G. M.Eliashberg. Interactions between electrons and lattice vibrations in a superconductor. Sov. Phys. JETP, 11, 696-702(1960).

    [77] F.Marsiglio. Eliashberg theory: A short review. Ann. Phys., 417, 168102(2020).

    [78] C.Grimaldi, S.Str?ssler, L.Pietronero. Nonadiabatic superconductivity. I. Vertex corrections for the electron-phonon interactions. Phys. Rev. B, 52, 10516-10529(1995).

    [79] C.Grimaldi, L.Pietronero, S.Str?ssler. Nonadiabatic superconductivity. II. Generalized Eliashberg equations beyond Migdal’s theorem. Phys. Rev. B, 52, 10530-10546(1995).

    [80] C.Grimaldi, E.Cappelluti, L.Pietronero. Isotope effect on m* in high Tc materials due to the breakdown of Migdal’s theorem. Europhys. Lett., 42, 667(1998).

    [81] S.Ciuchi, S.Str?ssler, E.Cappelluti, L.Pietronero, C.Grimaldi. High Tc superconductivity in MgB2 by nonadiabatic pairing. Phys. Rev. Lett., 88, 117003(2002).

    [82] L.Pietronero, L.Ortenzi, L.Boeri, E.Cappelluti. Conventional/unconventional superconductivity in high-pressure hydrides and beyond: Insights from theory and perspectives. Quantum Stud.: Math. Found., 5, 5-21(2018).

    [83] F.Bloch. Zum elektrischen widerstandsgesetz bei tiefen temperaturen. Z. Phys., 59, 208-214(1930).

    [84] F. J.Blatt. Physics of Electronic Conduction in Solids, 185-190(1968).

    [85] E. F.Talantsev. Classifying hydrogen-rich superconductors. Mater. Res. Express, 6, 106002(2019).

    [86] E. F.Talantsev. An approach to identifying unconventional superconductivity in highly-compressed superconductors. Supercond. Sci. Technol., 33, 124001(2020).

    [87] E. F.Talantsev, R. C.Mataira. Classifying superconductivity in ThH-ThD superhydrides/superdeuterides. Mater. Res. Express, 7, 016003(2020).

    [88] E. F.Talantsev. The electron-phonon coupling constant, Fermi temperature and unconventional superconductivity in the carbonaceous sulfur hydride 190 K superconductor. Supercond. Sci. Technol., 34, 034001(2021).

    [89] M. I.Eremets, P. P.Kong, A. P.Drozdov. Metallization of hydrogen(2021).

    [90] Y.Li, Y.Li, Y.Ma, H.Liu, J.Hao. The metallization and superconductivity of dense hydrogen sulfide. J. Chem. Phys., 140, 174712(2014).

    [91] F.Tian, T.Cui, D.Li, Y.Liu, X.Huang, D.Duan, W.Tian, H.Yu, B.Liu, Z.Zhao. Pressure-induced metallization of dense (H2S)2H2 with high-Tc superconductivity. Sci. Rep., 4, 6968(2014).

    [92] Z.Li et al. Superconductivity above 200 K discovered in superhydrides of calcium. Nat. Commun., 13, 2863(2022).

    [93] M.Zhou, X.Yang, K.Wang, G.Liu, L.Ma, X.Yu, Y.Wang, H.Wang, Y.Ma, Y.Zhao, Y.Xie, H.Liu. High-temperature superconducting phase in clathrate calcium hydride CaH6 up to 215 K at a pressure of 172 GPa. Phys. Rev. Lett., 128, 167001(2022).

    [94] J. S.Tse, Y.Ma, T.Iitaka, H.Wang, K.Tanaka. Superconductive sodalite-like clathrate calcium hydride at high pressures. Proc. Natl. Acad. Sci. U. S. A., 109, 6463-6466(2012).

    [95] I.Goncharenko, M. I.Eremets, I. A.Trojan, J. S.Tse, M.Hanfland, M.Amboage, Y.Yao. Pressure-induced hydrogen-dominant metallic state in aluminum hydride. Phys. Rev. Lett., 100, 045504(2008).

    [96] F.Belli, P.Hou, I.Errea, R.Bianco. Strong anharmonic and quantum effects in Pm3̄nAlH3 under high pressure: A first-principles study. Phys. Rev. B, 103, 134305(2021).

    [97] N. W.Ashcroft. Metallic hydrogen: A high-temperature superconductor?. Phys. Rev. Lett., 21, 1748-1749(1968).

    [98] V. L.Ginzburg. Superfluidity and superconductivity in the universe. J. Stat. Phys., 1, 3-24(1969).

    [99] R. H.Caton, C. B.Satterthwaite, J. F.Miller. Low-temperature heat capacity of normal and superconducting thorium hydride and thorium deuteride. Phys. Rev. B, 14, 2795(1976).

    [100] J.Hao, Y.Ma, Y.Li, Y.Li. The metallization and superconductivity of dense hydrogen sulphide(2014).

    [101] C. W.Glass, A. R.Oganov. Crystal structure prediction using ab initio evolutionary techniques: Principles and applications. J. Chem. Phys., 124, 244704-244715(2006).

    [102] A. R.Oganov, M.Valle, A. O.Lyakhov. How evolutionary crystal structure prediction works—And why. Acc. Chem. Res., 44, 227-237(2011).

    [103] Q.Zhu, H. T.Stokes, A. R.Oganov, A. O.Lyakhov. New developments in evolutionary structure prediction algorithm USPEX. Comput. Phys. Commun., 184, 1172-1182(2013).

    [104] T. A.Strobel, P.Ganesh, R. J.Hemley, P. R. C.Kent, M.Somayazulu. Novel cooperative interactions and structural ordering in H2S–H2. Phys. Rev. Lett., 107, 255503(2011).

    [105] M. I.Eremets, I. A.Troyan, A. P.Drozdov. Conventional superconductivity at 190 K at high pressures(2014).

    [106] R.Hoffmann, R. J.Hemley, N. W.Ashcroft, H.Liu, I. I.Naumov. Potential high-Tc superconducting lanthanum and yttrium hydrides at high pressure. Proc. Natl. Acad. Sci. U. S. A., 114, 6990-6995(2017).

    [107] D. A.Knyazev, P. P.Kong, V. S.Minkov, A. P.Drozdov, M. I.Eremets, M. A.Kuzovnikov, S. P.Besedin. Superconductivity at 215 K in lanthanum hydride at high pressures(2018).

    [108] M.Ahart, V. V.Struzhkin, Y.Meng, R. J.Hemley, M.Somayazulu, A. K.Mishra, Z. M.Geballe, M.Baldini. Evidence for superconductivity above 260 K in lanthanum superhydride at megabar pressures(2018).

    Evgeny F. Talantsev. Universal Fermi velocity in highly compressed hydride superconductors[J]. Matter and Radiation at Extremes, 2022, 7(5): 058403
    Download Citation