Next Article in Journal
A Novel Method for Growing α-Ga2O3 Films Using Mist-CVD Face-to-face Heating Plates
Previous Article in Journal
Electrochemical Detection of Olivetol Based on Poly(L-Serine) Film Layered Copper Oxide Modified Carbon Paste Electrode (p-L-Serine/CuO/CPE)
Previous Article in Special Issue
Superconducting Properties of YBa2Cu3O7−δ with a Multiferroic Addition Synthesized by a Capping Agent-Aided Thermal Treatment Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quantifying Nonadiabaticity in Major Families of Superconductors

by
Evgueni F. Talantsev
1,2
1
M. N. Miheev Institute of Metal Physics, Ural Branch, Russian Academy of Sciences, 18 S. Kovalevskoy Str., 620108 Ekaterinburg, Russia
2
NANOTECH Centre, Ural Federal University, 19 Mira Str., 620002 Ekaterinburg, Russia
Nanomaterials 2023, 13(1), 71; https://doi.org/10.3390/nano13010071
Submission received: 2 December 2022 / Revised: 17 December 2022 / Accepted: 19 December 2022 / Published: 23 December 2022
(This article belongs to the Special Issue Superconducting Nanostructures and Materials)

Abstract

:
The classical Bardeen–Cooper–Schrieffer and Eliashberg theories of the electron–phonon-mediated superconductivity are based on the Migdal theorem, which is an assumption that the energy of charge carriers, k B T F , significantly exceeds the phononic energy, ω D , of the crystalline lattice. This assumption, which is also known as adiabatic approximation, implies that the superconductor exhibits fast charge carriers and slow phonons. This picture is valid for pure metals and metallic alloys because these superconductors exhibit ω D k B T F < 0.01 . However, for n-type-doped semiconducting SrTiO3, this adiabatic approximation is not valid, because this material exhibits ω D k B T F 50 . There is a growing number of newly discovered superconductors which are also beyond the adiabatic approximation. Here, leaving aside pure theoretical aspects of nonadiabatic superconductors, we classified major classes of superconductors (including, elements, A-15 and Heusler alloys, Laves phases, intermetallics, noncentrosymmetric compounds, cuprates, pnictides, highly-compressed hydrides, and two-dimensional superconductors) by the strength of nonadiabaticity (which we defined by the ratio of the Debye temperature to the Fermi temperature, T θ T F ). We found that the majority of analyzed superconductors fall into the 0.025 T θ T F 0.4 band. Based on the analysis, we proposed the classification scheme for the strength of nonadiabatic effects in superconductors and discussed how this classification is linked with other known empirical taxonomies in superconductivity.

1. Introduction

The majority of experimental works in superconductivity utilize the classical Bardeen–Cooper–Schrieffer (BCS) [1] and Migdal–Eliashberg (ME) [2,3] theories as primary tools to analyze measured data. However, it should be clarified that these theories are valid for superconductors which satisfy the condition designated by the Born–Oppenheimer–Migdal approximation [4]:
ω D k B T F = T θ T F = 88   K 1.1 × 10 5   K | P b = 8 × 10 4 1
where is the reduced Planck constant, ω D is the Debye frequency, k B is the Boltzmann constant, T θ is the Debye temperature, T F is the Fermi temperature, and data for lead were reported by Poole [5]. The Born–Oppenheimer–Migdal approximation allows the separation of electronic and ionic motions in metals, because Equation (1) implies that the conductor exhibits fast charge carriers (for which characteristic energy scale is related to the Fermi temperature, T F ) and relatively slow phonons (for which characteristic energy scale is related to the Debye temperature, T θ ).
However, Equation (1) satisfies for many, but not for all superconductors, and the first discovered superconductor for which Equation (1) was found to be violated is n-type doped semiconducting SrTiO3 [6]:
ω D k B T F = T θ T F = 627   K 13   K | SrTiO 3 = 48 1
where data for SrTiO3 is taken from [7,8]. The theoretical description of the superconductivity in materials, in which the charge carriers and the lattice vibrations exhibit characteristic energy scales similar to Equation (2), is complicated, and the general designation of these superconductors are as nonadiabatic superconductors [9,10,11,12,13,14,15,16]. This theory [9,10,11,12,13,14,15,16] provides a general equation for the superconducting transition temperature, T c , in nonadiabatic superconductors [9]: T c = 1.134 × ε F k B × e 1 λ n a d , where ε F is the Fermi energy, and λ n a d is the coupling strength constant in nonadiabatic superconductors, which serves a similar role to the electron–phonon coupling strength, λ e p h , in the BCS [1] and ME [2,3] theories. In addition, one of the primary fundamental theoretical problems is calculating this constant with acceptable accuracy to describe the experiment [4,5,6,7,8,9,10,11,12,13,14,15,16].
For experimentalists, it is important to have a simple practical routine to establish the strength of nonantiadiabatic effects in newly discovered superconductors. The most obvious parameter, which serves as an experimentally measured value to quantify the strength of nonantiadiabaticity, is the T θ T F ratio. For practical use of this criterion, there is a need for the taxonomy of possible T θ T F values.
To establish the taxonomy, we performed the analysis for a broad a range as possible of superconductors; these range from two- to three-dimensional materials, from elements to compounds of up to five elements, from low- T c (with T c ~ 0.1   K ) to record high- T c (with T c = 240   K ) hydrides, and from materials that exhibit a high order of crystalline lattice symmetry to the materials with low symmetry. Namely, we tried to cover all superconductors for which primary characteristic parameters (apart T c , T θ , and T F ), such as the London penetration depth, λ ( 0 ) , the coherence length, ξ ( 0 ) , the amplitude of the superconducting energy gap, Δ ( 0 ) , and the electron–phonon coupling strength constant, λ e p h , were established. In the results, we presented the analysis of more than 40 superconductors within the families of main superconductors.
Based on our analysis, we proposed the following classification scheme:
{             T θ T F < 0.025 a d i a b a t i c   s u p e r c o n d u c t o r ;             0.025 T θ T F 0.4 m o d e r a t e l y   s t r o n g   n o n a d i a b a t i c i t y ;     0.4 < T θ T F   n o n a d i a b a t i c   s u p e r c o n d u c t o r .                    
One of our findings is that for weakly nonadiabatic superconductors (i.e., for materials exhibited 0.025 T θ T F 0.4 ), the predicting power of the BCS-ME theories (for instance, the prediction of the superconducting transition temperature) is reasonably accurate. However, all these superconductors are located outside of the BCS corner in the Uemura plot.
We also showed how the proposed classification scheme is linked to other known empirical scaling laws and taxonomies in superconductivity [13,17,18,19,20,21]; meanwhile, the search for the link of the proposed taxonomy with the recently reported big data [22,23] is under progress.

2. Utilized Models

Proposed taxonomy is based on the knowledge of three fundamental temperatures of the superconductor, which are T c , T θ , and T F . The superconducting transition temperature, T c , is directly measured in either temperature resistance or in magnetization experiments. It is also important to mention the primary experimental techniques and theoretical models utilized to deduce the Debye temperature, T θ , and the Fermi temperature, T F , in superconductors.
There are two primary techniques to determine the Debye temperature, T θ . One technique is to analyze the measured temperature-dependent normal-state specific heat, C p ( T ) , from which the electronic specific heat coefficient, γ n , and the Debye temperature, T θ , are deduced (see, for instance [24,25,26]):
C p ( T ) T = γ n + β T 2 + α T 4
where β is the Debye law lattice heat-capacity contribution, and α is from higher order lattice contributions. The Debye temperature can be calculated:
T θ = ( 12 π 4 R p 5 β ) 1 3
where R is the molar gas constant, and p is the number of atoms per formula unit.
Another technique is to fit normal-state temperature dependent resistance, R(T), to the Bloch–Grüneisen (BG) equation [24,25,26,27,28]:
R ( T ) = 1 1 R s a t + 1 R 0 + A × ( T T θ ) 5 × 0 T θ T x 5 ( e x 1 ) ( 1 e x ) d x
where, R s a t is the saturated resistance at high temperatures which is temperature independent, R 0 is the residual resistance at T 0   K , and A is free fitting parameter. Many research groups utilized both techniques (i.e., Equations (4)–(6)) to deduce T θ [24,25,26,27,29].
From the measured T c and the deduced T θ , one can derive the electron–phonon coupling constant, λ e p h , as a root of either the original McMillan equation [30], or its recently revisited form [27]:
T c = ( 1 1.45 ) × T θ × e ( 1.04 ( 1 + λ e p h ) λ e p h μ * ( 1 + 0.62 λ e p h ) ) × f 1 × f 2 *
f 1 = ( 1 + ( λ e p h 2.46 ( 1 + 3.8 μ * ) ) 3 / 2 ) 1 / 3
f 2 * = 1 + ( 0.0241 0.0735 × μ * ) × λ e p h 2
where μ * is the Coulomb pseudopotential, 0.10 μ * 0.15 [27,30].
There are several experimental techniques to derive the Fermi temperature, T F , from experimental data. One of these techniques is to measure the temperature dependent Seebeck coefficient, S ( T ) , and fit a measured dataset to the equation [8]:
| S ( T ) T | = π 2 3 k B e 1 T F
Another approach is to measure the magnetic quantum oscillations [31], from which the magnitude of charge carrier mass, m * = m e ( 1 + λ e p h ) (where m e is bare mass of electron), together with the size of the Fermi wave vector, k F , can be obtained and plugged into [31]:
T F = 2 2 k B k F 2 m *
An alternative approach is based on the extraction of the charge carriers mass, m * , and density, n , as two of four parameters from the simultaneous analysis of C p ( T ) , R ( T ) , the muon spin relaxation (μSR), the lower critical field data, B c 1 ( T ) , and the upper critical field data, B c 2 ( T ) [32], and plugging these parameters into the equation for an isotropic spherical Fermi surface [32]:
T F = 2 2 k B 1 m * ( 3 π 2 n s ) 2 3
where n s is bulk charge curriers density at T 0   K . For 3D superconductors, n s is given by the equation [33]:
n s ( 0 ) = m * μ 0 e 2 1 λ 2 ( 0 )
where μ 0 is the permeability of free space, l is the charge carrier mean free path, λ ( 0 ) is the ground state London penetration depth, and ξ ( 0 ) is the ground state coherence length.
It should be noted that λ ( 0 ) can be also deduced from the ground state lower critical field [28,34]:
B c 1 ( 0 ) = ϕ 0 4 π l n ( 1 + 2 κ ( 0 ) ) λ 2 ( 0 )
where κ ( 0 ) = λ ( 0 ) ξ ( 0 ) is the ground state Ginzburg–Landau parameter.
For two dimensional (2D) superconductors, T F can be determined from μSR measurements and crystallographic data [18]:
T F = π 2 k B 1 m * n s × c i n t
where c i n t is the average distance between superconducting planes.
If measuring techniques are limited to the magnetoresistance measurements, R ( T , B ) (which was the case in the field of highly-compressed near-room temperature superconductors (NRTS) [35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50], until recent experimental progress by Minkov et al. [51,52]), T F can be estimated by the equation [53]:
T F = π 2 m * 2 k B 2 × ξ 2 ( 0 ) × Δ 2 ( 0 )
where Δ ( 0 ) is the ground state amplitude of the superconducting energy gap, which is varying in a reasonably narrow range 3.2 2 Δ ( 0 ) k B T c 5.0 , so that the ballpark value for T F can be estimated. For instance, ξ ( 0 ) can be deduced from magnetoresistance measurements [54] and the electron–phonon coupling strength constant, λ e p h , can be assumed to be the average value of values calculated by first-principles calculations [55,56].

3. Results

In Table 1, we present data for major groups of superconductors, where data sources for T c , T θ , T F , and other parameters (for instance, λ e p h ) are given.
In Figure 1, we show the T c   vs . T F dataset in a log–log plot, which is the traditional data representation in the well-known Uemura plot [18].
In Figure 2, we represent the same superconducting materials, but here we display the λ e p h   vs . T θ T F dataset in a semi-log plot. To our best knowledge, the λ e p h   vs . T θ T F plot was first plotted by Pietronero et al. [13] in linear–linear scales. However, because the T θ T F ratio for main families of superconductors is varied within four orders of magnitude (Table 1), and 0.4 λ e p h 3.0 , it is more suitable to use the semi-log plot (Figure 2).
Finally, in Figure 3, we represented the same superconducting materials, but here we displayed the T c   vs . T θ T F dataset in a log–log plot. This type of plot was chosen because as T c , as T θ T F are varied within several orders of magnitude.

4. Discussion

The family of near-room temperature superconductors (NRTS) is represented in Table 1 and Figure 1 by H3S (P = 155 GPa), SnH12 (P = 190 Gpa), and La1-xNdxH10 (x = 0.09, P = 180 Gpa). Two independent approaches were used to perform calculations in H3S:
  • T F was calculated based on Equations (12) and (13). In these calculations λ ( 0 ) = 37   nm (extracted from the analysis of DC magnetization experiments reported by Minkov et al. [51,52]) was used.
  • T F was calculated based Equations (6)–(9) and (18), in which utilized ξ ( 0 ) values were extracted [53] from magnetoresistance measurements reported by Mozaffari et al. [54]).
It should be noted that, in both approaches, the electron–phonon coupling strength constant, λ e p h , was assumed to be λ e p h = 1.76 , which is the average value of values calculated by first-principles calculations [55,56], and values extracted from experimental R(T) data [27].
It can be seen in Table 1 and Figure 1 that the calculated T F values for H3S, by two alternative approaches, are in a very good agreement with each other. To demonstrate the acceptable level of variation in T F values for the same material, in Table 1 and Figure 1 we present the results of the calculations for pure metals, where T F was calculated by the two approaches mentioned above and the use of experimental data reported by different research groups.
T F in HTS cuprates were calculated by the Equations (13) and (15), which do not require the knowledge of the electron–phonon coupling constants, λ e p h . This is despite Ledbetter et al. [7] reporting the so-called effective electron–phonon coupling strength, λ e p h , e f f , from which the effective mass can be deduced, m * = ( 1 + λ e p h , e f f ) × m e .
In addition, it should be noted that for YBa2Cu3O7, Uemura [83] reported the relation [83]:
m * m e = 2.5
from which λ e p h = 1.5 can be derived. Calculated values are in a reasonable agreement with experimental m * m e values reported by several research groups [108,109,110] in YBa2Cu3O7-x.
However, because the phenomenology of the electron–phonon mediated superconductivity cannot describe the superconducting state in cuprates, and the T θ for cuprates were taken as experimental values (see, for instance, report by Ledbetter et al. [7,88]), all cuprate superconductors are shown in Figure 1 and Figure 3 and are not shown in Figure 2.
It should be mentioned that the result of the T F calculation in MATBG (Table 1), T F = 16.5   K , which was primarily based on the London penetration depth, λ ( 0 ) = 1860   nm , was deduced in Ref. [96] from the self-field critical current density, J c ( s f , T ) , by the approach proposed by us [84]:
J c ( s f , T ) = ϕ 0 4 π μ 0 l n ( 1 + 2 κ ( T ) ) λ 3 ( T )
The remarkable agreement of the deduced value, T F = 16.5   K , and the value reported in the original work on MATBG by Cao et al. [95], T F = 17   K , which was calculated based on normal state charge carriers density in MATBG, independently validates our primary idea [84] about the fundamental nature of the self-field critical current in weak-links samples [84,85,111]. This concept was recently proven by Paturi and Huhtinen [112], who utilized the fact that the London penetration depth, λ ( 0 ) , in real samples, depends on the mean free-path of charge carriers, l :
λ ( 0 ) = λ c l e a n   l i m i t ( 0 ) 1 + ξ ( 0 ) l
where λ ( 0 ) is the effective penetration depth, and λ c l e a n   l i m i t ( 0 ) is the penetration depth in samples, exhibiting a very long mean free-path, l ξ ( 0 ) . Paturi and Huhtinen [112] varied l in YBa2Cu3O7-x films and showed that the change in J c ( s t , T ) satisfies Equations (18) and (19).
Materials, in which λ ( 0 ) was deduced by the mean temperature dependent self-field critical current density, J c ( s f , T ) (Equation (18)), have designation “ J c ( s f , T ) ” in Figure 1, Figure 2 and Figure 3.
The MATBG does not show in Figure 2, because the derivation of λ e p h cannot be performed by the used phenomenology: m * = ( 1 + λ e p h ) × m e , because m * m e = 0.2 [96]; however, this material is shown in Figure 1 and Figure 3, because λ e p h is not required for these plots.
Returning back to hydrides, we need to note that Durajski [56] performed first-principles and studied the strength of the nonadiabatic effects in highly-compressed sulfur hydride and phosphorus hydride. Calculations show that the strength of the nonadiabatic effects can be quantified as moderately weak in comparison with the classical nonadiabatic superconductor SrTiO3. This is in a good agreement with our result (see Figure 3 and Table 1), that all deduced T θ T F values for NRTS are within the range of:
0.03 T θ T F 0.3
Moreover, the classical nonadiabatic superconductor SrTiO3 falls into the intermediate zone between unconventional and BCS superconductors; this is because this material exhibits T c T F = 0.0066 , and by this criterion, SrTiO3 is similar to the Laves phase materials, intermetallics, A-15 alloys, and Heusler alloys, which cannot be considered to be a correct manifestation of primary uniqueness for this nonadiabatic material.
More unexpectedly, a two dimensional LiC6 (which is a lithium-doped graphene) superconductor falls into the BCS metals zone in the Uemura plot (Figure 1), despite the fact that this material exhibits reasonable strength in the nonadiabatic effects, T c T F = 0.15 [99].
However, in Figure 2 and Figure 3, the outstanding separations of all nonadiabatic superconductors from their adiabatic and moderate nonadiabatic counterparts are clearly manifested.
By looking at the data in Figure 2 and Figure 3, it is easy to recognize that ¾ (32 of 42) of the analyzed superconductors fall into a reasonably narrow band:
0.025 T θ T F 0.4
Based on this, we proposed that the values in Equation (21) were used as empirical limits for the adiabatic superconductors ( T θ T F 0.025 ), moderate nonadiabatic superconductors ( 0.025 T θ T F 0.4 ), and strong nonadiabatic superconductors ( T θ T F 0.4 ).
It also follows from our analysis that all strong nonadiabatic superconductors exhibit low superconducting transition temperatures, T c 1.2   K (Figure 3).

5. Conclusions

In this work, we proposed a new classification scheme to quantify the effects of nonadiabaticity in superconductors. By performing the analysis of experimental data for more than 40 superconductors, which represent the primary families of superconductors, we found that ¾ of all analyzed superconductors fall into a narrow 0.025 T θ T F 0.4 band. Based on this, we proposed the taxonomy for the strength of the nonadiabatic effects in superconductors.

Funding

This research was funded by the Ministry of Science and Higher Education of the Russian Federation, grant number No. AAAA-A18-118020190104-3 (theme “Pressure”). The research funding from the Ministry of Science and Higher Education of the Russian Federation (Ural Federal University Program of Development within the Priority-2030 Program) is gratefully acknowledged. The APC was funded by MDPI.

Data Availability Statement

Not applicable.

Acknowledgments

The author thanks Luciano Pietronero (Universita’ di Roma) for comments about the limitations of the applicability of Migdal–Eliashberg (ME) theory of the electron–phonon mediated superconductivity. The author also thanks Dmitry V. Semenok (Skolkovo Institute of Science and Technology) and Dominik Szczesniak (Jan Dlugosz University in Czestochowa) for reading and commenting on the paper.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Bardeen, J.; Cooper, L.N.; Schrieffer, J.R. Theory of superconductivity. Phys. Rev. 1957, 108, 1175–1204. [Google Scholar] [CrossRef] [Green Version]
  2. Migdal, A.B. Interaction between electrons and lattice vibrations in a normal metal. Sov. Phys. JETP 1958, 7, 996–1001. [Google Scholar]
  3. Eliashberg, G.M. Interactions between electrons and lattice vibrations in a superconductor. Sov. Phys. JETP 1960, 11, 696–702. [Google Scholar]
  4. Pietronero, L.; Strässler, S.; Grimaldi, C. Nonadiabatic superconductivity. I. Vertex corrections for the electron-phonon interactions. Phys. Rev. B 1995, 52, 10516–10529. [Google Scholar] [CrossRef]
  5. Poole, C.P. Handbook of Superconductivity, 1st ed.; Poole, C.P., Jr., Ed.; Academic Press: New York, NY, USA, 1999; pp. 31–32. [Google Scholar]
  6. Takada, Y. Theory of superconductivity in polar semiconductors and its application to n-type semiconducting SrTiO3. J. Phys. Soc. Jpn. 1980, 49, 1267–1275. [Google Scholar] [CrossRef]
  7. Ledbetter, H.; Lei, M.; Kim, S. Elastic constants, Debye temperatures, and electron-phonon parameters of superconducting cuprates and related oxides. Phase Transit. 1990, 23, 61–70. [Google Scholar] [CrossRef]
  8. Lin, X.; Zhu, Z.; Fauque, B.; Behnia, K. Fermi surface of the most dilute superconductor. Phys. Rev. X 2013, 3, 021002. [Google Scholar] [CrossRef] [Green Version]
  9. Takada, Y. Plasmon mechanism of superconductivity in two- and three-dimensional electron systems. J. Phys. Soc. Jpn. 1978, 45, 786–794. [Google Scholar] [CrossRef]
  10. Grimaldi, C.; Pietronero, L.; Strässler, S. Nonadiabatic superconductivity. II. Generalized Eliashberg equations beyond Migdal’s theorem. Phys. Rev. B 1995, 52, 10530–10546. [Google Scholar] [CrossRef]
  11. Grimaldi, C.; Cappelluti, E.; Pietronero, L. Isotope effect on m* in high Tc materials due to the breakdown of Migdal’s theorem. Europhys. Lett. 1998, 42, 667–672. [Google Scholar] [CrossRef] [Green Version]
  12. Cappelluti, E.; Ciuchi, S.; Grimaldi, C.; Pietronero, L.; Strässler, S. High Tc superconductivity in MgB2 by nonadiabatic pairing. Phys. Rev. Lett. 2002, 88, 117003. [Google Scholar] [CrossRef] [Green Version]
  13. Pietronero, L.; Boeri, L.; Cappelluti, E.; Ortenzi, L. Conventional/unconventional superconductivity in high-pressure hydrides and beyond: Insights from theory and perspectives. Quantum Stud. Math. Found. 2018, 5, 5–21. [Google Scholar] [CrossRef]
  14. Klimin, S.N.; Tempere, J.; Devreese, J.T.; He, J.; Franchini, C.; Kresse, G. Superconductivity in SrTiO3: Dielectric function method for non-parabolic bands. J. Supercond. Nov. Magn. 2019, 32, 2739–2744. [Google Scholar] [CrossRef] [Green Version]
  15. Gor’kov, L.P. Phonon mechanism in the most dilute superconductor n-type SrTiO3. Proc. Natl. Acad. Sci. USA 2016, 113, 4646–4651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Kresin, V.Z.; Wolf, S.A. Colloquium: Electron-lattice interaction and its impact on high Tc superconductivity. Rev. Mod. Phys. 2009, 81, 481–501. [Google Scholar] [CrossRef] [Green Version]
  17. Harshman, R.; Fiory, A.T. High-Tc superconductivity in hydrogen clathrates mediated by Coulomb interactions between hydrogen and central-atom electrons. J. Supercond. Nov. Magn. 2020, 33, 2945–2961. [Google Scholar] [CrossRef]
  18. Uemura, Y.J. Classifying superconductors in a plot of Tc versus Fermi temperature TF. Physical C 1991, 185–189, 733–734. [Google Scholar] [CrossRef]
  19. Talantsev, E.F.; Crump, W.P.; Tallon, J.L. Universal scaling of the self-field critical current in superconductors: From sub-nanometre to millimetre size. Sci. Rep. 2017, 7, 10010. [Google Scholar] [CrossRef]
  20. Homes, C.C.; Dordevic, S.V.; Strongin, M.; Bonn, D.A.; Liang, R.; Hardy, W.N.; Komiya, S.; Ando, Y.; Yu, G.; Kaneko, N.; et al. A universal scaling relation in high-temperature superconductors. Nature 2004, 430, 539–541. [Google Scholar] [CrossRef] [Green Version]
  21. Koblischka, R.; Koblischka-Veneva, A. Calculation of Tc of superconducting elements with the Roeser–Huber formalism. Metals 2022, 12, 337. [Google Scholar] [CrossRef]
  22. Zhang, T.; Jiang, Y.; Song, Z.; Huang, H.; He, Y.; Fang, Z.; Weng, H.; Fang, C. Catalogue of topological electronic materials. Nature 2019, 566, 475–479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Xu, Y.; Vergniory, M.G.; Ma, D.-S.; Mañes, J.L.; Song, Z.-D.; Bernevig, B.A.; Regnault, N.; Elcoro, L. Catalogue of topological phonon materials. arXiv 2022, arXiv:2211.11776. [Google Scholar] [CrossRef]
  24. Pyon, S.; Kudo, K.; Matsumura, J.-I.; Ishii, H.; Matsuo, G.; Nohara, M.; Hojo, H.; Oka, K.; Azuma, M.; Garlea, V.O.; et al. Superconductivity in noncentrosymmetric iridium silicide Li2IrSi3. J. Phys. Soc. Jpn. 2014, 83, 093706. [Google Scholar] [CrossRef] [Green Version]
  25. Kudo, K.; Hiiragi, H.; Honda, T.; Fujimura, K.; Idei, H.; Nohara, M. Superconductivity in Mg2Ir3Si: A fully ordered Laves phase. J. Phys. Soc. Jpn. 2020, 89, 013701. [Google Scholar] [CrossRef]
  26. Sobczak, Z.; Winiarski, M.J.; Xie, W.; Cava, R.J.; Klimczuk, T. Superconductivity in the intermetallic compound Zr5Al4. EPL 2019, 127, 37005. [Google Scholar] [CrossRef]
  27. Talantsev, E.F. Advanced McMillan’s equation and its application for the analysis of highly-compressed superconductors. Supercond. Sci. Technol. 2020, 33, 094009. [Google Scholar] [CrossRef]
  28. Talantsev, E.F. The electron–phonon coupling constant and the Debye temperature in polyhydrides of thorium, hexadeuteride of yttrium, and metallic hydrogen phase III. J. Appl. Phys. 2021, 130, 195901. [Google Scholar] [CrossRef]
  29. Mayoh, D.A.; Barker, J.A.T.; Singh, R.P.; Balakrishnan, G.; Paul, D.M.; Lees, M.R. Superconducting and normal-state properties of the noncentrosymmetric superconductor Re6Zr. Phys. Rev. B 2017, 96, 064521. [Google Scholar] [CrossRef] [Green Version]
  30. McMillan, W.L. Transition temperature of strong-coupled superconductors. Phys. Rev. 1968, 167, 331–344. [Google Scholar] [CrossRef]
  31. Vignolle, B.; Vignolles, D.; LeBoeuf, D.; Lepault, S.; Ramshaw, B.; Liang, R.; Bonn, D.A.; Hardy, W.N.; Doiron-Leyraud, N.; Carrington, A.; et al. Quantum oscillations and the Fermi surface of high-temperature cuprate superconductors. Comptes Rendus Phys. 2011, 12, 446–460. [Google Scholar] [CrossRef]
  32. Kleiner, R.; Buckel, W. Superconductivity: An Introduction, 3rd ed.; Wiley-VCH: Berlin, Germany, 2016. [Google Scholar]
  33. Tinkham, M. Introduction to Superconductivity., 2nd ed.; Dover Publications: New York, NY, USA, 2004. [Google Scholar]
  34. Talantsev, E.F. Critical de Broglie wavelength in superconductors. Mod. Phys. Lett. B 2018, 32, 1850114. [Google Scholar] [CrossRef]
  35. Drozdov, A.P.; Eremets, M.I.; Troyan, I.A.; Ksenofontov, V.; Shylin, S.I. Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system. Nature 2015, 525, 73–76. [Google Scholar] [CrossRef] [PubMed]
  36. Drozdov, A.P.; Kong, P.P.; Minkov, V.S.; Besedin, S.P.; Kuzovnikov, M.A.; Mozaffari, S.; Balicas, L.; Balakirev, F.F.; Graf, D.E.; Prakapenka, V.B.; et al. Superconductivity at 250 K in lanthanum hydride under high pressures. Nature 2019, 569, 528–531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Somayazulu, M.; Ahart, M.; Mishra, A.K.; Geballe, Z.M.; Baldini, M.; Meng, Y.; Struzhkin, V.V.; Hemley, R.J. Evidence for superconductivity above 260 K in lanthanum superhydride at megabar pressures. Phys. Rev. Lett. 2019, 122, 027001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Semenok, D.V.; Kvashnin, A.G.; Ivanova, A.G.; Svitlyk, V.; Fominski, V.Y.; Sadakov, A.V.; Sobolevskiy, O.A.; Pudalov, V.M.; Troyan, I.A.; Oganov, A.R. Superconductivity at 161 K in thorium hydride ThH10: Synthesis and properties. Mater. Today 2020, 33, 36–44. [Google Scholar] [CrossRef] [Green Version]
  39. Chen, W.; Semenok, D.V.; Kvashnin, A.G.; Huang, X.; Kruglov, I.A.; Galasso, M.; Song, H.; Duan, D.; Goncharov, A.F.; Prakapenka, V.B.; et al. Synthesis of molecular metallic barium superhydride: Pseudocubic BaH12. Nat. Commun. 2021, 12, 273. [Google Scholar] [CrossRef]
  40. Troyan, I.A.; Semenok, D.V.; Kvashnin, A.G.; Sadakov, A.V.; Sobolevskiy, O.A.; Pudalov, V.M.; Ivanova, A.G.; Prakapenka, V.B.; Greenberg, E.; Gavriliuk, A.G.; et al. Anomalous high-temperature superconductivity in YH6. Adv. Mater. 2021, 33, 2006832. [Google Scholar] [CrossRef]
  41. Kong, P.; Minkov, V.S.; Kuzovnikov, M.A.; Drozdov, A.P.; Besedin, S.P.; Mozaffari, S.; Balicas, L.; Balakirev, F.F.; Prakapenka, V.B.; Chariton, S.; et al. Superconductivity up to 243 K in yttrium hydrides under high pressure. Nat. Commun. 2021, 12, 5075. [Google Scholar] [CrossRef]
  42. Ma, L.; Wang, K.; Xie, Y.; Yang, X.; Wang, Y.; Zhou, M.; Liu, H.; Yu, X.; Zhao, Y.; Wang, H.; et al. High-temperature superconducting phase in clathrate calcium hydride CaH6 up to 215 K at a pressure of 172 GPa. Phys. Rev. Lett. 2022, 128, 167001. [Google Scholar] [CrossRef]
  43. Semenok, D.V.; Troyan, I.A.; Ivanova, A.G.; Kvashnin, A.G.; Kruglov, I.A.; Hanfland, M.; Sadakov, A.V.; Sobolevskiy, O.A.; Pervakov, K.S.; Lyubutin, I.S.; et al. Superconductivity at 253 K in lanthanum–yttrium ternary hydrides. Mater. Today 2021, 48, 18–28. [Google Scholar] [CrossRef]
  44. Zhou, D.; Semenok, D.V.; Duan, D.; Xie, H.; Chen, W.; Huang, X.; Li, X.; Liu, B.; Oganov, A.R.; Cui, T. Superconducting praseodymium superhydrides. Sci. Adv. 2020, 6, eaax6849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Hong, F.; Shan, P.F.; Yang, L.X.; Yue, B.B.; Yang, P.T.; Liua, Z.Y.; Sun, J.P.; Dai, J.H.; Yu, H.; Yin, Y.Y.; et al. Possible superconductivity at ∼70 K in tin hydride SnHx under high pressure. Mater. Today Phys. 2022, 22, 100596. [Google Scholar] [CrossRef]
  46. Chen, W.; Semenok, D.V.; Huang, X.; Shu, H.; Li, X.; Duan, D.; Cui, T.; Oganov, A.R. High-temperature superconducting phases in cerium superhydride with a Tc up to 115 K below a pressure of 1 Megabar. Phys. Rev. Lett. 2021, 127, 117001. [Google Scholar] [CrossRef] [PubMed]
  47. Osmond, I.; Moulding, O.; Cross, S.; Muramatsu, T.; Brooks, A.; Lord, O.; Fedotenko, T.; Buhot, J.; Friedemann, S. Clean-limit superconductivity in Im3m H3S synthesized from sulfur and hydrogen donor ammonia borane. Phys. Rev. B 2022, 105, L220502. [Google Scholar] [CrossRef]
  48. Semenok, D.V.; Troyan, I.A.; Sadakov, A.V.; Zhou, D.; Galasso, M.; Kvashnin, A.G.; Ivanova, A.G.; Kruglov, I.A.; Bykov, A.A.; Terent’ev, K.Y.; et al. Effect of magnetic impurities on superconductivity in LaH10. Adv. Mater. 2022, 34, 2204038. [Google Scholar] [CrossRef]
  49. Bi, J.; Nakamoto, Y.; Zhang, P.; Shimizu, K.; Zou, B.; Liu, H.; Zhou, M.; Liu, G.; Wang, H.; Ma, Y. Giant enhancement of superconducting critical temperature in substitutional alloy (La,Ce)H9. Nat. Commun. 2022, 13, 5952. [Google Scholar] [CrossRef]
  50. Li, Z.; He, X.; Zhang, C.; Wang, X.; Zhang, S.; Jia, Y.; Feng, S.; Lu, K.; Zhao, J.; Zhang, J.; et al. Superconductivity above 200 K discovered in superhydrides of calcium. Nat. Commun. 2022, 13, 2863. [Google Scholar] [CrossRef]
  51. Minkov, V.S.; Bud’ko, S.L.; Balakirev, F.F.; Prakapenka, V.B.; Chariton, S.; Husband, R.J.; Liermann, H.P.; Eremets, M.I. Magnetic field screening in hydrogen-rich high-temperature superconductors. Nat. Commun. 2022, 13, 3194. [Google Scholar] [CrossRef]
  52. Minkov, V.S.; Ksenofontov, V.; Budko, S.L.; Talantsev, E.F.; Eremets, M.I. Trapped magnetic flux in hydrogen-rich high-temperature superconductors. arXiv 2022, arXiv:2206.14108. [Google Scholar] [CrossRef]
  53. Talantsev, E.F. Classifying superconductivity in compressed H3S. Mod. Phys. Lett. B 2019, 33, 1950195. [Google Scholar] [CrossRef] [Green Version]
  54. Mozaffari, S.; Sun, D.; Minkov, V.S.; Drozdov, A.P.; Knyazev, D.; Betts, J.B.; Einaga, M.; Shimizu, K.; Eremets, M.I.; Balicas, L. Superconducting phase-diagram of H3S under high magnetic fields. Nat. Commun. 2019, 10, 2522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Errea, I.; Calandra, M.; Pickard, C.J.; Nelson, J.; Needs, R.J.; Li, Y.; Liu, H.; Zhang, Y.; Ma, Y.; Mauri, F. High-pressure hydrogen sulfide from first principles: A strongly anharmonic phonon-mediated superconductor. Phys. Rev. Lett. 2015, 856, 157004. [Google Scholar] [CrossRef] [PubMed]
  56. Durajski, A.P. Quantitative analysis of nonadiabatic effects in dense H3S and PH3 superconductors. Sci. Rep. 2016, 6, 38570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Prozorov, R.; Giannetta, R.W.; Carrington, A.; Fournier, P.; Greene, R.L.; Guptasarma, P.; Hinks, D.G.; Banks, A.R. Measurements of the absolute value of the penetration depth in high-Tc superconductors using a low-Tc superconductive coating. Appl. Phys. Lett. 2000, 77, 4202–4204. [Google Scholar] [CrossRef] [Green Version]
  58. Poole, P.P.; Farach, H.A.; Creswick, R.J.; Prozorov, R. Superconductivity; Academic Press: London, UK, 2007; Chapter 1. [Google Scholar]
  59. Carbotte, J.P. Properties of boson-exchange superconductors. Rev. Mod. Phys. 1990, 62, 1027–1157. [Google Scholar] [CrossRef]
  60. Peabody, G.E.; Meservey, R. Magnetic flux penetration into superconducting thin films. Phys. Rev. B 1972, 6, 2579–2595. [Google Scholar] [CrossRef]
  61. Guritanu, V.; Goldacker, W.; Bouquet, F.; Wang, Y.; Lortz, R.; Goll, G.; Junod, A. Specific heat of Nb3Sn: The case for a second energy gap. Phys. Rev. B 2004, 70, 184526. [Google Scholar] [CrossRef] [Green Version]
  62. Orlando, T.P.; McNiff, E.J., Jr.; Foner, S.; Beasley, M.R. Critical fields, Pauli paramagnetic limiting, and material parameters of Nb3Sn and V3Si. Phys. Rev. B 1979, 19, 4545–4561. [Google Scholar] [CrossRef]
  63. Zhang, R.; Gao, P.; Wang, X.; Zhou, Y. First-principles study on elastic and superconducting properties of Nb3Sn and Nb3Al under hydrostatic pressure. AIP Adv. 2015, 5, 107233. [Google Scholar] [CrossRef] [Green Version]
  64. Junod, A.; Muller, J. Comment on specific heat of a transforming V3Si crystal. Solid State Commun. 1980, 36, 721–724. [Google Scholar] [CrossRef]
  65. Yamashita, A.; Matsuda, T.D.; Mizuguchi, Y. Synthesis of new high-entropy alloy-type Nb3(Al, Sn, Ge, Ga, Si) superconductors. J. Alloys Compd. 2021, 868, 159233. [Google Scholar] [CrossRef]
  66. Winterlik, J.; Fecher, G.H.; Felser, C.; Jourdan, M.; Grube, K.; Hardy, F.; von Löhneysen, H.; Holman, K.L.; Cava, R.J. Ni-based superconductor: Heusler compound ZrNi2Ga. Phys. Rev. B 2008, 78, 184506. [Google Scholar] [CrossRef]
  67. Klimczuk, T.; Wang, C.H.; Gofryk, K.; Ronning, F.; Winterlik, J.; Fecher, G.H.; Griveau, J.-C.; Colineau, E.; Felser, C.; Thompson, J.D.; et al. Superconductivity in the Heusler family of intermetallics. Phys. Rev. B 2012, 85, 174505. [Google Scholar] [CrossRef] [Green Version]
  68. Singh, D.; Barker, J.A.T.; Thamizhavel, A.; Hillier, A.D.; McK Paul, D.; Singh, R.P. Superconducting properties and μSR Study of the noncentrosymmetric superconductor Nb0.5Os0.5. J. Phys. Condens. Matter 2018, 30, 075601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Karki, A.B.; Xiong, Y.M.; Vekhter, I.; Browne, D.; Adams, P.W.; Young, D.P.; Thomas, K.R.; Chan, J.Y.; Kim, H.; Prozorov, R. Structure and physical properties of the noncentrosymmetric superconductor Mo3Al2C. Phys. Rev. B 2010, 82, 064512. [Google Scholar] [CrossRef] [Green Version]
  70. Górnicka, K.; Gui, X.; Wiendlocha, B.; Nguyen, L.T.; Xie, W.; Cava, R.J.; Klimczuk, T. NbIr2B2 and TaIr2B2–New low symmetry noncentrosymmetric superconductors with strong spin–orbit coupling. Adv. Funct. Mater. 2020, 31, 2007960. [Google Scholar] [CrossRef]
  71. Barker, J.A.T.; Breen, B.D.; Hanson, R.; Hillier, A.D.; Lees, M.R.; Balakrishnan, G.; Paul, D.M.; Singh, R.P. Superconducting and normal-state properties of the noncentrosymmetric superconductor Re3Ta. Phys. Rev. B 2018, 98, 104506. [Google Scholar] [CrossRef] [Green Version]
  72. Gong, C.; Wang, Q.; Wang, S.; Lei, H. Superconducting properties of MgCu2-type Laves phase compounds SrRh2 and BaRh2. J. Phys. Condens. Matter 2020, 32, 295601. [Google Scholar] [CrossRef]
  73. Gutowska, S.; Górnicka, K.; Wójcik, P.; Klimczuk, T.; Wiendlocha, B. Strong-coupling superconductivity of SrIr2 and SrRh2: Phonon engineering of metallic Ir and Rh. Phys. Rev. B 2021, 104, 054505. [Google Scholar] [CrossRef]
  74. Horie, R.; Horigane, K.; Nishiyama, S.; Akimitsu, M.; Kobayashi, K.; Onari, S.; Kambe, T.; Kubozono, Y.; Akimitsu, J. Superconductivity in 5d transition metal Laves phase SrIr2. J. Phys. Condens. Matter 2020, 32, 175703. [Google Scholar] [CrossRef]
  75. Mao, Z.Q.; Rosario, M.M.; Nelson, K.D.; Wu, K.; Deac, I.G.; Schiffer, P.; Liu, Y.; He, T.; Regan, K.A.; Cava, R.J. Experimental determination of superconducting parameters for the intermetallic perovskite superconductor MgCNi3. Phys. Rev. B 2003, 67, 094502. [Google Scholar] [CrossRef] [Green Version]
  76. Ryżyńska, Z.; Chamorro, J.R.; McQueen, T.M.; Wiśniewski, P.; Kaczorowski, D.; Xie, W.; Cava, R.J.; Klimczuk, T.; Winiarski, M.J. RuAl6—An endohedral aluminide superconductor. Chem. Mater. 2020, 32, 3805–3812. [Google Scholar] [CrossRef]
  77. Langenberg, E.; Ferreiro-Vila, E.; Leborán, V.; Fumega, A.O.; Pardo, V.; Rivadulla, F. Analysis of the temperature dependence of the thermal conductivity of insulating single crystal oxides. APL Mater. 2016, 4, 104815. [Google Scholar] [CrossRef] [Green Version]
  78. Adroja, D.; Bhattacharyya, A.; Biswas, P.K.; Smidman, M.; Hillier, A.D.; Mao, H.; Luo, H.; Cao, G.-H.; Wang, Z.; Wang, C. Multigap superconductivity in ThAsFeN investigated using μSR measurements. Phys. Rev. B 2017, 96, 144502. [Google Scholar] [CrossRef] [Green Version]
  79. Albedah, M.A.; Nejadsattari, F.; Stadnik, Z.M.; Wang, Z.-C.; Wang, C.; Cao, G.H. Absence of the stripe antiferromagnetic order in the new 30 K superconductor ThFeAsN. J. Alloys Compd. 2017, 695, 1128–1136. [Google Scholar] [CrossRef]
  80. Bhattacharyya, A.; Adroja, D.T.; Smidman, M.; Anand, V.K. A brief review on μSR studies of unconventional Fe- and Cr-based superconductors. Sci. China Phys. Mech. Astron. 2018, 61, 127402. [Google Scholar] [CrossRef] [Green Version]
  81. Sonier, J.E.; Sabok-Sayr, S.A.; Callaghan, F.D.; Kaiser, C.V.; Pacradouni, V.; Brewer, J.H.; Stubbs, S.L.; Hardy, W.N.; Bonn, D.A.; Liang, R.; et al. Hole-doping dependence of the magnetic penetration depth and vortex core size in YBa2Cu3Oy: Evidence for stripe correlations near 1/8 hole doping. Phys. Rev. B 2007, 76, 134518. [Google Scholar] [CrossRef] [Green Version]
  82. Kiefl, R.F.; Hossain, M.D.; Wojek, B.M.; Dunsiger, S.R.; Morris, G.D.; Prokscha, T.; Salman, Z.; Baglo, J.; Bonn, D.A.; Liang, R.; et al. Direct measurement of the London penetration depth in YBa2Cu3O6.92 using low-energy µSR. Phys. Rev. B 2010, 81, 180502. [Google Scholar] [CrossRef] [Green Version]
  83. Uemura, Y.J. Muon spin relaxation studies on high- Tc organic, heavy-fermion, and Chevrel phase superconductors. Physical B 1991, 169, 99–106. [Google Scholar] [CrossRef]
  84. Talantsev, E.F.; Tallon, J.L. Universal self-field critical current for thin-film superconductors. Nat. Commun. 2015, 6, 7820. [Google Scholar] [CrossRef] [Green Version]
  85. Talantsev, E.; Crump, W.P.; Tallon, J.L. Thermodynamic parameters of single- or multi-band superconductors derived from self-field critical currents. Ann. Der Phys. 2017, 529, 1700197. [Google Scholar] [CrossRef] [Green Version]
  86. Wagner, P.; Hillmer, F.; Frey, U.; Adrian, H. Thermally activated flux movement and critical transport current density in epitaxial Bi2Sr2CaCu2O8+δ films. Phys. Rev. B 1994, 49, 13184. [Google Scholar] [CrossRef]
  87. Holstein, W.L.; Wilker, C.; Laubacher, D.B.; Face, D.W.; Pang, P.; Warrington, M.S.; Carter, C.F.; Parisi, L.A. Critical current density and resistivity measurements for long patterned lines in Tl2Ba2CaCu2O8 thin films. J. Appl. Phys. 1993, 74, 1426–1430. [Google Scholar] [CrossRef]
  88. Ledbetter, H. Dependence of on Debye temperature θD for various cuprates. Phys. C 1994, 235–240, 1325–1326. [Google Scholar] [CrossRef]
  89. Krusin-Elbaum, L.; Tsuei, C.C.; Gupta, A. High current densities above 100 K in the high-temperature superconductor HgBa2CaCu2O6+δ. Nature 1995, 373, 679–681. [Google Scholar] [CrossRef]
  90. Hänisch, J.; Attenberger, A.; Holzapfel, B.; Schultz, L. Electrical transport properties of Bi2Sr2Ca2Cu3O10+δ thin film [001] tilt grain boundaries. Phys. Rev. B 2002, 65, 052507. [Google Scholar] [CrossRef]
  91. Cava, R.J.; Batlogg, B.; Krajewski, J.J.; Farrow, R.; Rupp, L.W.; White, A.E.; Short, K.; Peck, W.F.; Kometani, T. Superconductivity near 30 K without copper: The Ba0.6K0.4BiO3 perovskite. Nature 1988, 332, 814–816. [Google Scholar] [CrossRef]
  92. Uemura, Y.J.; Luke, G.M.; Sternlieb, B.J.; Brewer, J.H.; Carolan, J.F.; Hardy, W.N.; Kadono, R.; Kempton, J.R.; Kiefl, R.F.; Kreitzman, S.R. Universal correlations between Tc and n S m * (carrier density over effective mass) in high-Tc cuprate. Phys. Rev. Lett. 1989, 62, 2317–2320. [Google Scholar] [CrossRef]
  93. Szczesniak, D.; Kaczmarek, A.Z.; Drzazga-Szczesniak, E.A.; Szczesniak, R. Phonon-mediated superconductivity in bismuthates by nonadiabatic pairing. Phys. Rev. B 2021, 104, 094501. [Google Scholar] [CrossRef]
  94. Hundley, M.F.; Thompson, J.D.; Kwei, G.H. Specific heat of the cubic high-Tc superconductor Ba0.6K0.4BiO3. Solid State Commun. 1989, 70, 1155–1158. [Google Scholar] [CrossRef]
  95. Cao, Y.; Fatemi, V.; Fang, S.; Watanabe, K.; Taniguchi, T.; Kaxiras, E.; Jarillo-Herrero, P. Unconventional superconductivity in magic-angle graphene superlattices. Nature 2018, 556, 43–50. [Google Scholar] [CrossRef] [Green Version]
  96. Talantsev, E.F.; Mataira, R.C.; Crump, W.P. Classifying superconductivity in Moiré graphene superlattices. Sci. Rep. 2020, 10, 212. [Google Scholar] [CrossRef]
  97. Cocemasov, A.I.; Nika, D.L.; Balandin, A.A. Engineering of the thermodynamic properties of bilayer graphene by atomic plane rotations: The role of the out-of-plane phonons. Nanoscale 2015, 7, 12851. [Google Scholar] [CrossRef] [Green Version]
  98. Ludbrook, B.M.; Levy, G.; Nigge, P.; Zonno, M.; Schneider, M.; Dvorak, D.J.; Veenstra, C.N.; Zhdanovich, S.; Wong, D.; Dosanjh, P.; et al. Evidence for superconductivity in Li-decorated monolayer graphene. Proc. Natl. Acad. Sci. USA 2015, 112, 11795–11799. [Google Scholar] [CrossRef] [Green Version]
  99. Szczesniak, D.; Szczesniak, R. Signatures of nonadiabatic superconductivity in lithium-decorated graphene. Phys. Rev. B 2019, 99, 224512. [Google Scholar] [CrossRef] [Green Version]
  100. Park, S.; Kim, S.Y.; Kim, H.K.; Kim, M.J.; Kim, T.; Kim, H.; Choi, G.S.; Won, C.J.; Kim, S.; Kim, K.; et al. Superconductivity emerging from a stripe charge order in IrTe2 nanoflakes. Nat. Commun. 2021, 12, 3157. [Google Scholar] [CrossRef]
  101. Eremets, M.I.; Shimizu, K.; Kobayashi, T.C.; Amaya, K. Metallic CsI at pressures of up to 220 gigapascals. Science 1998, 281, 1333–1335. [Google Scholar] [CrossRef]
  102. Talantsev, E.F. Fermi-liquid nonadiabatic highly compressed cesium iodide superconductor. Condens. Matter 2022, 7, 65. [Google Scholar] [CrossRef]
  103. Talantsev, E.F. The dominance of non-electron–phonon charge carrier interaction in highly-compressed superhydrides. Supercond. Sci. Technol. 2021, 34, 115001. [Google Scholar] [CrossRef]
  104. Talantsev, E.F. Universal Fermi velocity in highly compressed hydride superconductors. Mater. Radiat. Extrem. 2022, 7, 058403. [Google Scholar] [CrossRef]
  105. Talantsev, E.F. Electron–phonon coupling constant and BCS ratios in LaH10−y doped with magnetic rare-earth element. Supercond. Sci. Technol. 2022, 35, 095008. [Google Scholar] [CrossRef]
  106. Shimizu, K.; Suhara, K.; Ikumo, M.; Eremets, M.I.; Amaya, K. Superconductivity in oxygen. Nature 1998, 393, 767–769. [Google Scholar] [CrossRef]
  107. Talantsev, E.F. An approach to identifying unconventional superconductivity in highly compressed superconductors. Supercond. Sci. Technol. 2020, 33, 124001. [Google Scholar] [CrossRef]
  108. Sebastian, S.E.; Harrison, N.; Lonzarich, G.G. Towards resolution of the Fermi surface in underdoped high-Tc superconductors. Rep. Prog. Phys. 2012, 75, 102501. [Google Scholar] [CrossRef] [Green Version]
  109. Hsua, Y.-T.; Hartstein, M.; Davies, A.J.; Hickey, A.J.; Chan, M.K.; Porras, J.; Loew, T.; Taylor, S.V.; Liua, H.; Eaton, A.G.; et al. Unconventional quantum vortex matter state hosts quantum oscillations in the underdoped high-temperature cuprate superconductors. Proc. Natl. Acad. Sci. USA 2021, 118, e2021216118. [Google Scholar] [CrossRef]
  110. Padilla, W.J.; Lee, Y.S.; Dumm, M.; Blumberg, G.; Ono, S.; Segawa, K.; Komiya, S.; Ando, Y.; Basov, D.N. Constant effective mass across the phase diagram of high-Tc cuprates. Phys. Rev. B 2005, 72, 060511. [Google Scholar] [CrossRef] [Green Version]
  111. Talantsev, E.F.; Crump, W.P. Weak-links criterion for pnictide and cuprate superconductors. Supercond. Sci. Technol. 2018, 31, 124001. [Google Scholar] [CrossRef]
  112. Paturi, P.; and Huhtinen, H. Roles of electron mean free path and flux pinning in optimizing the critical current in YBCO superconductors. Supercond. Sci. Technol. 2022, 35, 065007. [Google Scholar] [CrossRef]
Figure 1. Uemura plot (Tc vs. TF) for primary superconducting families. References on original data (Tc and TF) can be found in Table 1.
Figure 1. Uemura plot (Tc vs. TF) for primary superconducting families. References on original data (Tc and TF) can be found in Table 1.
Nanomaterials 13 00071 g001
Figure 2. Plot of T θ T F   vs .   λ e p h for primary superconducting families. This type of plot proposed by Pietronero et al. [13]. References for original data (Tθ, λ e p h , TF) can be found in Table 1.
Figure 2. Plot of T θ T F   vs .   λ e p h for primary superconducting families. This type of plot proposed by Pietronero et al. [13]. References for original data (Tθ, λ e p h , TF) can be found in Table 1.
Nanomaterials 13 00071 g002
Figure 3. Plot T θ T F   vs .   T c for primary superconducting families. References on original data (Tθ, Tc, TF) can be found in Table 1.
Figure 3. Plot T θ T F   vs .   T c for primary superconducting families. References on original data (Tθ, Tc, TF) can be found in Table 1.
Nanomaterials 13 00071 g003
Table 1. Superconductors and their parameters used in the work. In all calculations (except some original sources, μ * = 0.13 ) .
Table 1. Superconductors and their parameters used in the work. In all calculations (except some original sources, μ * = 0.13 ) .
Type/Chemical
Composition
λ(0)
(nm)
ξ(0)
(nm)
λe-phTc
(K)
Tθ
(K)
2 Δ ( 0 ) k B T c TF
(103 K)
Tθ/TF
Pure metals
Aluminium 1.18 [57,58]394 [5] 136 [5] 2.9 × 10 3
Aluminium50 [57]1550 [58]0.43 [59]1.18 [57,58]394 [5]3.535 [59]18.9 (Equation (12)) 2.1 × 10 2
Tin 3.72 [58]170 [5] 118 [5] 1.4 × 10 3
Tin77 [60]180 [58]0.72 [59]3.72 [58]170 [5]3.705 [59]10.0 (Equation (12)) 1.2 × 10 2
Lead 7.20 [58]88 [5] 110 [5] 8 × 10 4
Lead64 [60]87 [58]1.55 [59]7.20 [60]88 [5]4.497 [59]11.2 (Equation (12)) 7.8 × 10 3
Niobium 9.25 [58]265 [5] 61.8 [5] 4.3 × 10 3
Niobium52 [58]39 [58]0.98 [59]9.25 [58]265 [5]3.964 [59]16.1 (Equation (12)) 1.6 × 10 2
Gallium52 [58]39 [58]2.25 [59]1.09 [5,32,58,59]325 [5,32,58,59] 120 [5,32,58] 2.7 × 10 3
A15 Alloys
Nb3Sn124 [61]3.6 [61]1.8 [62]17.9 [61]234 [61]4.2 [62]4.5 (Equation (12)) 5.2 × 10 2
V3Si62 [62]3.3 [62]0.96 [62]16.4 [63]297 [64]3.7 [62]12.8 (Equation (12)) 2.3 × 10 2
Nb3Ge90 [58]3.0 [58]1.60 [59]23.2 [58]302 [65]4.364 [59]7.1 (Equation (12)) 4.3 × 10 2
Heusler alloys
ZrNi2Ga350 [66]15 [66]0.551 [66]2.85 [66]300 [66] 1.4 (Equation (12)) 2.2 × 10 1
YPd2Sn196 [67]19 [67]0.70 [67]4.7 [67]210 [67]4.1 [67]2.9 (Equation (12)) 7.2 × 10 2
HfPd2Al225 [67]13 [67]0.68 [67]3.66 [67]182 [67]3.74 [67]2.4 (Equation (12)) 7.5 × 10 2
Noncentrosymmetric
Nb0.5Os0.5654 [68]7.8 [68]0.53 [68]3.07 [68]367 [68]3.62 [68]0.60 (Equation (12)) 6.1 × 10 1
Re6Zr (mSR)356 [29]3.7 [29]0.67 [29]6.75 [29]338 [29]3.72 [29]1.3 2.6 × 10 1
Re6Zr (magnetization)247 [29]3.3 [29]0.67 [29]6.75 [29]237 [29]3.72 [29]2.1 1.1 × 10 1
Mo3Al2C376 [69]4.2 [69]0.74
(Equations (7)–(9))
9.2 [69]339 [69]4.03 [69]1.2 2.8 × 10 1
NbIr2B2 [70]2234.50.747.18274 2.4 1.1 × 10 1
TaIr2B2 [70]3424.70.705.1230 1.4 1.7 × 10 1
Re3Ta [71] 0.624.7321 0.64 5.0 × 10 1
Laves phases
BaRh2 [72]3408.40.805.6178 1.4 1.3 × 10 1
SrRh2 [72]2299.10.715.4237 2.3 1.0 × 10 1
SrRh2 [73]1218.60.935.4250 5.3 4.7 × 10 2
SrIr2 [74]2377.50.845.9180 2.3 8.2 × 10 2
Intermetallics
MgCNi3 [75]2484.60.74
(Equations (7)–(9))
7.6284 2.1 1.4 × 10 1
RuAl6 [76]26527.70.811.21458 1.9 2.4 × 10 1
Perovskite
SrTiO3 0.2 [7]0.086 [8]690 [77] 1.3 × 10 2 [8] 5.3 × 10 1
Pnictides
ThFeAsN375 [78] 1.48 [78]28.1 [78]332 [79] 0.47 (Equation (17)) c i n t = 8.5   [78] 7.0 × 10 1
KCa2Fe4As4F2230 [80] 1.59 [80]33.4 [80]366 [80] 1.3 (Equation (17)) c i n t = 8.5   [80] 2.9 × 10 1
RbCa2Fe4As4F2232 [80] 1.45 [80]29.2 [80]332 [80] 1.2 (Equation (17)) c i n t = 8.5   [80] 2.8 × 10 1
CsCa2Fe4As4F2244 [80] 1.44 [80]28.3 [80]344 [80] 1.1 (Equation (17)) c i n t = 8.5   [80] 3.1 × 10 1
Cuprates
YBa2Cu3O7 [81]115 [81,82]2.5 [81]1.5 [83]93.2 [81]437 [7] 3.4 (Equation (17)) c i n t = 5.8   [83] 1.2 × 10 1
(Y,Dy)Ba2Cu3O7 [84]128 [84,85]2.5 [81]1.5 [83]90.4 [84,85]437 [7]4.24 [84,85]2.9 (Equation (17)) c i n t = 5.8   [83] 1.5 × 10 1
Bi2Sr2CaCu2O8 [86]196 [85]1.2 [85]4.7 [7]82.7 [85]240 [7]3.9 [85]1.2 (Equation (17)) c i n t = 6   [83] 2.0 × 10 1
Tl2Ba2CaCu2O8 [87]179 [85]1.2 [85] 103 [85]425 [88]4.3 [85]1.5 (Equation (17)) c i n t = 6   [83] 2.9 × 10 1
HgBa2CaCu2O8 [89]188 [85]1.6 [85] 120 [85]525 [88]3.3 [85]1.3 (Equation (17)) c i n t = 6   [83] 3.9 × 10 1
Bi2Sr2Ca2Cu3O10 [90]175 [85]1.0 [85]4.585 [85]319 [7]4.5 [85]1.5 (Equation (17)) c i n t = 6   [83] 2.1 × 10 1
Bismuthates
Ba1-xKxBiO3
(x = 0.4) [91,92]
1.10 [93]23 [92]210 [94]3.8-4.1 [93]1.5 [92] 1.4 × 10 1
Ba1-xKxBiO3
(x = 0.5) [91,92]
1.10 [93]14 [92]210 [94]3.8-4.0 [93] 1.1 [92] 1.9 × 10 1
2D superconductors
MATBG [95]2180 [96] m * m e = 0.2 [96]1.2 [96]1864 [97]4.4 [96] 16.5 × 10 3 (Equation (15)) c i n t = 1   nm 1.1 × 10 2
MATBG [95] 61.4 [96] [96]1.2 [96]1864 [97]4.4 [96] 28.6 × 10 3 (Equation (16)) 6.5 × 10 1
Li-doped graphene, LiC6 [98] 0.61 [99]5.9 [98]2240 [99] 15.5 [99] 1.45 × 10 1
IrTe2 [100] (sample thickness is 21 nm)600 [100]75 [100] 1.6 [100] 5.46 [100] 0.118   ( Equation   ( 15 ) ) c i n t = 0.54   nm
Ionic Salt
CsI
(P = 206 GPa) [101]
0.445 [102]1.1 [101]339 [102] ( 20 ± 4 ) × 10 2 [102] 17 ± 4 [102]
NRTS hydrides
H3S
(P = 155 GPa) [35]
37 [52]1.9 [54]2.2 [103]197 [103]1427 [103] 21.6 (Equation (12)) and [104] 6.6 × 10 2
H3S
(P = 155 GPa) [35]
1.9 [54]1.76 [55,56]197 [103]1427 [103]3.55 [53] 10 ± 3
(Equation (16)) and [104]
( 1.4 ± 0.3 ) × 10 1
LaH10
(P = 150 GPa) [36]
30 [51]1.5 [51]2.77 [27]240 [27]1310 [27] 27.0 (Equation (12)) 2.7 × 10 2
La1-xNdxH10 (x = 0.15)
(P = 180 GPa) [48]
2.3 [105]1.65 [105]122 [105]1156 [105]4.0 [105] 4.4 [105]
(Equation (16))
2.6 × 10 1
Compressed oxygen
ζ-O2
(P = 115 GPa) [106]
42 [107]0.42 [107]0.64 [107]306 [107] 3.5 × 10 2 [107] (Equation (16))8.7
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Talantsev, E.F. Quantifying Nonadiabaticity in Major Families of Superconductors. Nanomaterials 2023, 13, 71. https://doi.org/10.3390/nano13010071

AMA Style

Talantsev EF. Quantifying Nonadiabaticity in Major Families of Superconductors. Nanomaterials. 2023; 13(1):71. https://doi.org/10.3390/nano13010071

Chicago/Turabian Style

Talantsev, Evgueni F. 2023. "Quantifying Nonadiabaticity in Major Families of Superconductors" Nanomaterials 13, no. 1: 71. https://doi.org/10.3390/nano13010071

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop