Next Article in Journal
Permeability Coefficient of Concrete under Complex Stress States
Previous Article in Journal
Hybrid BO-XGBoost and BO-RF Models for the Strength Prediction of Self-Compacting Mortars with Parametric Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Intrinsic Coherence Length Anisotropy in Nickelates and Some Iron-Based Superconductors

by
Evgeny F. Talantsev
1,2
1
M. N. Miheev Institute of Metal Physics, Ural Branch, Russian Academy of Sciences, 18, S. Kovalevskoy St., 620108 Ekaterinburg, Russia
2
NANOTECH Centre, Ural Federal University, 19 Mira St., 620002 Ekaterinburg, Russia
Materials 2023, 16(12), 4367; https://doi.org/10.3390/ma16124367
Submission received: 11 May 2023 / Revised: 11 June 2023 / Accepted: 12 June 2023 / Published: 13 June 2023

Abstract

:
Nickelate superconductors, R1−xAxNiO2 (where R is a rare earth metal and A = Sr, Ca), experimentally discovered in 2019, exhibit many unexplained mysteries, such as the existence of a superconducting state with Tc (up to 18 K) in thin films and yet absent in bulk materials. Another unexplained mystery of nickelates is their temperature-dependent upper critical field, B c 2 ( T ) , which can be nicely fitted to two-dimensional (2D) models; however, the deduced film thickness, d s c , G L , exceeds the physical film thickness,   d s c , by a manifold. To address the latter, it should be noted that 2D models assume that d s c is less than the in-plane and out-of-plane ground-state coherence lengths, d s c < ξ a b ( 0 ) and d s c < ξ c ( 0 ) , respectively, and, in addition, that the inequality ξ c ( 0 ) < ξ a b ( 0 ) satisfies. Analysis of the reported experimental B c 2 ( T ) data showed that at least one of these conditions does not satisfy for R1-xAxNiO2 films. This implies that nickelate films are not 2D superconductors, despite the superconducting state being observed only in thin films. Based on this, here we propose an analytical three-dimensional (3D) model for a global data fit of in-plane and out-of-plane B c 2 ( T ) in nickelates. The model is based on a heuristic expression for temperature-dependent coherence length anisotropy: γ ξ ( T ) = γ ξ ( 0 ) 1 1 a × T T c , where a > 1 is a unitless free-fitting parameter. The proposed expression for γ ξ ( T ) , perhaps, has a much broader application because it has been successfully applied to bulk pnictide and chalcogenide superconductors.

1. Introduction

High-temperature superconductivity in Nd1-xSrxNiO2 was experimentally discovered by Li et al. [1] in 2019, while the first theoretical report in which the oxidation state of Ni+ was established to be a condition under which nickelates become similar to cuprates (created by Li et al. [1]) was published by Anisimov et al. [2] in 1999. This experimental discovery initiated further theoretical and experimental studies on R1-xAxNiO2 (where R is a rare earth and A = Sr, Ca) thin films [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50] and bulk [51]. It is a widely accepted point of view that the superconducting state in nickelates is exhibited only in thin films [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51], with a thickness of d s c 15   n m . This is one of the primary mysteries in nickelate superconductors.
Another unexplained mystery of nickelates is the temperature dependence of the upper critical field, B c 2 ( T ) . For instance, when this fundamental field is measured for an applied field oriented in a perpendicular direction to the (00L) planes of the film (which will be designated as B c 2 , p e r p ( T ) , herein), the dependence can be nicely fitted to the Ginzburg–Landau (GL) model [52]:
B c 2 , p e r p ( T ) = ϕ 0 2 π 1 ξ a b 2 ( T ) = ϕ 0 2 π ( 1 T T c ) ξ a b 2 ( 0 ) ,
where ϕ 0 is a superconducting flux quantum, and ξ a b ( 0 ) is ground state in-plane coherence length. When an external field is oriented in parallel to the (00L) planes of the film (which will be designated as B c 2 , p a r a ( T ) ), the dependence reported by many research groups can be nicely fitted to the two-dimensional Ginzburg–Landau (2D-GL) model [52,53]:
B c 2 , p a r a ( T ) = ϕ 0 2 π 12 d s c , G L 1 ξ a b ( T ) = ϕ 0 2 π 12 d s c , G L 1 T T c ξ a b ( 0 ) ,
where d s c , G L is the film thickness associated with the use of Equation (2) for data fit.
This result should be interpreted as direct evidence of 2D superconductivity in nickelates, supporting the experimental observation that the superconducting state is observed only in thin films of nickelates. However, the deduced film thickness d s c , G L (from the B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) data fit to Equations (1) and (2) [23]) by to 2–3 times exceeds the physical film thickness d s c [23]. Here we found that an identical problem, i.e., d s c d s c , G L , does exist for nearly all nickelate films for which experimental data were reported.
To resolve this issue, here we point out that the derivation of Equation (2) [52,53] is based on the assumption that physical film thickness d s c is much less than the ground-state coherence length ξ ( 0 ) of the superconductor, which means that a 2D superconductor is defined by two conditions:
d s c ξ a b ( 0 ) ,
d s c ξ c ( 0 ) ,
where ξ c ( 0 ) is the out-of-plane coherence length. In addition, there is another hidden assumption for the derivation of Equation (2), which is:
ξ c ( 0 ) < ξ a b ( 0 ) .
As shown below, at least one of these conditions (Equations (3)–(5)) is not satisfied for the studied nickelate films. From this, we conclude that there is an incident, that B c 2 , p a r a ( T ) data of nickelate films are nicely approximated by the square root of an independent variable of a two-fluid model:
B c 2 , p a r a ( T ) 1 T T c ,
and that the deeper physics behind this dependence should be determined.
In this paper, we propose to resolve this problem by accepting that superconductivity in nickelates is a three-dimensional (3D) phenomenon and, thus, the upper critical field should be described by standard 3D Ginzburg–Landau equations:
{ (7) B c 2 , p e r p ( T ) = ϕ 0 2 π 1 ξ a b 2 ( T ) = ϕ 0 2 π ( 1 T T c ) ξ a b 2 ( 0 )                                     (8) B c 2 , p a r a ( T ) = ϕ 0 2 π 1 ξ c ( T ) 1 ξ a b ( T ) = ϕ 0 2 π 1 ξ a b ( T ) γ ξ ( T ) 1 ξ a b ( T ) = ϕ 0 2 π γ ξ ( T ) ( 1 T T c ) ξ a b 2 ( 0 )            
where γ ξ ( T ) = ξ a b ( T ) ξ c ( T ) denotes the temperature-dependent coherence length anisotropy. By experimenting with many analytical functions, we found a remarkably simple and robust heuristic expression for γ ξ ( T ) , which surprisingly enough can also be applied to iron-based superconductors:
γ ξ ( T ) = ξ a b ( T ) ξ c ( T ) = ξ a b ( 0 ) ξ c ( 0 ) 1 1 1 a × T T c = γ ξ ( 0 ) 1 1 1 a × T T c = γ ξ ( 0 ) 1 1 T T γ ,
where a is a free-fitting parameter (varies within a narrow range of 1.2 < a < 2.3 for all studied superconductors), and T γ a × T c .

2. The Upper Critical Field Definition

Before Equations (7)–(9) will be applied for B c 2 ( T ) data fit, we should clarify the definition of the B c 2 ( T ) , because different research groups define this fundamental field using different criteria.
In many reports on nickelates, the upper critical field, B c 2 ( T ) , and, as a direct consequence of it, the coherence length, ξ ( T ) , were defined/deduced from the magnetoresistance curves, R ( T , B ) , by utilizing 50% of the normal state resistance criterion, i.e., R ( T ) R n o r m ( T ) = 0.5 (it should be noted that some research groups [49] utilized the criterion of R ( T ) R n o r m ( T ) 1.0 , which returns the most overestimated T c and B c 2 ( T ) and the most underestimated ξ a b ( 0 ) and ξ c ( 0 ) values).
However, in direct experiments performed by Harvey et al. [35], it was shown that the diamagnetic response in nickelate films always appears at temperatures well below the zero-resistance temperature, T c , z e r o (see, for instance, Figure 2 in [35]). Because diamagnetism is an essential and unavoidable property of the superconducting state, the definition of the fundamental superconducting field (i.e., the upper critical field, B c 2 ( T ) ) at the condition at which the superconducting state does not exist (and thus neither the Abrikosov’s vortices, nor the phase coherence of the order parameter and the amplitude coherence of the order parameter exist) is incorrect. The definition by R ( T ) R n o r m ( T ) = 0.5 , or by any other similar ratios, except R ( T ) R n o r m ( T ) 0 , causes many confusions, and the most notable one is the claim that the Pauli limiting field is violated in practically all thin film superconductors [36,54,55]. However, the primary reason for claimed violation is the definition of the upper critical field, B c 2 ( T ) , by the criterion at which the superconducting state does not yet exist.
It should be reaffirmed that because the upper critical field, B c 2 ( T ) , is defined as the magnetic flux density at which the superconducting state collapses and the diamagnetism is an essential property of the superconducting state, the definition of the B c 2 ( T ) should be made based on the disappearance of the diamagnetic response or, if it is impossible to measure, by the R ( T ) R n o r m ( T ) 0 criterion. However, it should be mentioned that these definitions have been implemented in very few studies [56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74].
Based on remarkably overestimated B c 2 ( T ) values, defined by R ( T ) R n o r m ( T ) = 0.5 or R ( T ) R n o r m ( T ) 1 [49,54,55,75,76,77,78], and a very broad resistive transition width in some thin film superconductors, new effects/phenomena can be claimed (for instance, the Pauli limiting field violation [36,54,55,75,76,77,78]). However, these new effects/phenomena can be explained by the misinterpretation of the thermodynamic fluctuations of the phase and the amplitude of the order parameter in superconductors with low charge carrier density [79,80,81] as the superconducting state. A strongly fluctuating Fermi sea is not an ordered superconducting condensate, where amplitude and phase coherence have been established across the entire sample. Based on this, it is incorrect to apply basic interpretations developed for superconducting condensate (at which a non-zero Meissner response should exist or, at least, zero resistance can be measured experimentally) to a system with strong local fluctuations in space and time [79,80,81], which are manifested as a drop of several percent in the resistance.
Emery and Kivelson [80] proposed the temperature scale T φ for the phase fluctuations of the superconducting order parameter:
T φ = A ϕ 0 2 a 4 π μ 0 λ a b 2 ( 0 ) ,
where a is the mean spacing between superconducting sheets for 2D superconductors and a = π ξ ( 0 ) for 3D superconductors, A = 0.9 for 2D superconductors, and A = 2.2 for 3D superconductors.
Alternatively, Bulaevskii et al. [81] propose the following equation for the amplitude fluctuations of the superconducting order parameter:
T a m p = π 3 A T φ = ϕ 0 2 a 12 π μ 0 λ a b 2 ( 0 ) ,
If T a m p and/or T φ are close to mean-field T c , then thermal fluctuations are expected to break Cooper pairs, which causes the suppression of the observed superconducting transition temperature. For instance, in cuprates, strong phase fluctuations reduce the experimentally observed transition temperature, T c , to well below its mean-field value, by more than 30% [80].
Returning now to the definition of the upper critical field, we can note that in several studies performed on perfect single-phase superconductors, the diamagnetic response was detected at T c , d i a o n s e t which is below (and in many cases is well below) the temperature at which the resistance drops to zero; that is, T c , d i a o n s e t T c , z e r o [35,82,83,84,85,86,87,88,89,90]. These rare but high-quality experimental reports provide additional evidence for the need to define the B c 2 ( T ) by at least the R ( T ) R n o r m ( T ) 0 criterion, which is the most accurate experimental value for the superconducting state collapse/emergence if only resistive measurements have been performed for the given sample. This B c 2 ( T ) definition has been implemented, but, again, in very rare studies [82,87,91,92,93] in comparison with the majority of studies, where B c 2 ( T ) was defined by the 0.5 R ( T ) R n o r m ( T ) < 1 criterion (see, for instance, [94,95]).
It is interesting to note that Mandal et al. [87] defined B c 2 ( T ) as by T c , d i a o n s e t (and for this definition, the derived B c 2 ( 0 ) = 3.79   T ) as well as by the R ( T ) R n o r m ( T ) = 0.5 criterion (and for this definition, the derived B c 2 ( 0 ) = 5.44   T ) for bulk Zr2Ir single crystals. This result demonstrates that B c 2 ( 0 ) defined by the R ( T ) R n o r m ( T ) = 0.5 criterion can be overestimated by a factor of 1.4.
However, for Nd0.775Sr0.225NiO2 nickelate films, this difference is much larger, shown in Figure 4 in [23], where the difference between B c 2 , p e r p ( T ) defined by R ( T = 6   K ) R n o r m ( T = 6 K ) = 0.5 and by R ( T = 6   K ) R n o r m ( T = 6   K ) = 0.01 is approximately five times (the subscripts perp and para are for the direction of the applied magnetic field to the surface of the films). A much larger difference between B c 2 ( T ) defined by different R ( T ) R n o r m ( T ) criteria was reported by Xiang et al. [22] for Nd0.8Sr0.2NiO2 thin films. These observed differences for nickelate films are much larger in comparison with bulk Zr2Ir [87] and bulk FeSe [71], which both exhibit a similar T c 7   K value and where the difference in B c 2 ( T ) defined by different R ( T ) R n o r m ( T ) criteria is about 40%.
Recently, independent direct confirmation that T c , d i a o n s e t < T c , z e r o in the nickelate films has been reported by Zeng et al. [45], who observed the inequality for films with thickness 5.5   n m d s c 15.2   n m .
Outstanding differences between absolute B c 2 ( T ) values defined by different criteria were recently demonstrated in isotropic superconductor TaReSi by Shang et al. [96], and for anisotropic superconductor LaFeAsO by Zhigadlo et al. [97], while the vanishing of the diamagnetic response [96,97,98,99] remains the most accurate criterion to determine the upper critical field. In regard to the B c 2 ( T ) anisotropy, which is the topic of the study, different criteria of the B c 2 ( T ) definition lead to not only different γ ξ ( T ) values, but also to the change in the function trend, i.e., decrease/increase type [97].
Based on the above, in this report we define the B c 2 ( T ) at the lowest possible R ( T , B ) R n o r m ( T , B = 0 ) criterion which can be applied to given experimental R ( T , B ) datasets (which depend on the signal/noise ratio and other real-world experimental issues).

3. Results

3.1. La0.8Ca0.2NiO2 Film

Chow et al. [36] reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets for La0.8Ca0.2NiO2 film (which has a physical thickness d s c = 15   n m ) defined by R ( T , B ) R n o r m ( T , B = 0 ) = 0.10 , R ( T , B ) R n o r m ( T , B = 0 ) = 0.50 , and R ( T , B ) R n o r m ( T , B = 0 ) = 0.90 criteria. By following our discussion in the previous section, in Figure 1 we show reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets defined by the R ( T , B ) R n o r m ( T , B = 0 ) = 0.10 criterion and global data fit to the 2D-GL model (Figure 1a,b; Equations (1) and (2)) and to our model (Figure 1c–e; Equations (7)–(9)). The quality of both fits (for which we used R-squared value [100]) is high, and the fitting parameters have low mutual dependence. However, deduced film thickness, d s c , G L = 8.0 ± 0.1   n m , differs by nearly two times from physical film thickness d s c = 15   n m , which is a manifestation of the general problem associated with utilization of Equations (1) and (2) for nickelate films [23], as we discussed above.
It should be noted that the deduced ξ a b , G L ( 0 ) = 3.99 ± 0.02   n m is much smaller than any of the two film thicknesses (i.e., the physical thickness, d s c , and the deduced thickness, d s c , G L , from the fit):
ξ a b , G L ( 0 ) 4   n m < d s c , G L = 8   n m
ξ a b , G L ( 0 ) 4   n m   d s c = 15   n m
Considering that there is an expectation that ξ c ( 0 ) ξ a b ( 0 ) , Equations (12) and (13) imply that Equations (1) and (2) cannot be used to fit B c 2 ( T ) data for this film, because the film is not thin.
Deduced coherence lengths from our model (Figure 1c–e) are ξ a b ( 0 ) = 4.01 ± 0.03   n m and ξ c ( 0 ) = ξ a b ( 0 ) γ ξ ( 0 ) = 2.64   n m . These values agree with our assumption that the film exhibits three-dimensional (3D) superconductivity:
ξ c ( 0 ) 2.6   n m < ξ a b ( 0 ) 4   n m   d s c = 15   n m ,
The physical meaning of this part of the γ ξ ( T ) curve we discuss in the Discussion section; however, in short, it can be mentioned that the anisotropy should also exist for the phase and amplitude fluctuations of the order parameter above the transition temperature, T c . Because all nickelates exhibit reasonably wide resistive transitions, similar to other unconventional superconductors (such as cuprates [80] and pnictides [101,102]), we can propose that the a × T c value can be interpreted as the onset for the establishing of the anisotropy in superconducting fluctuations, T γ , in the given material.
Thus, our interpretation of the T γ = a × T c is based on the assumption that there is a universal temperature dependence for the anisotropy of the superconducting order parameter and of the fluctuations of this order above the superconducting transition, which, at least from the first glance, looks like a reasonable assumption.

3.2. La0.8Sr0.2NiO2 Film

Wang et al. [103] reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets for La0.8Sr0.2NiO2 film (which has a physical thickness d s c ~ 7   n m ) defined by the R ( T , B ) R n o r m ( T , B = 0 ) = 0.50 criterion. In Figure 2, we show reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets and global data fit to the 2D-GL model (Figure 2a,b; Equations (1) and (2)) and to our model (Figure 2c–e; Equations (7)–(9)). The quality of both fits is high, and the fitting parameters have low mutual dependence.
However, the deduced film thickness, d s c , G L = 9.5 ± 0.1   n m , exceeds physical film thickness d s c = 7   n m (Figure 2a,b). In addition, the inequality of:
ξ a b ( 0 ) 4   n m < d s c ~ 7   n m < d s c , G L = 9.5   n m
shows that the 2D-GL model cannot be used for the analysis (because the film is not sufficiently thin), despite a good fit quality.
Deduced coherence lengths from the fit to our model (Equations (7)–(9)) are ξ a b ( 0 ) = 4.13 ± 0.03   n m and ξ c ( 0 ) = ξ a b ( 0 ) γ ξ ( 0 ) = 2.74   n m . These values agree with the assumption of our model that the film exhibits 3D superconductivity:
ξ c ( 0 ) 2.7   n m < ξ a b ( 0 ) 4.1   n m < d s c ~ 7   n m ,
In Figure 2c, we show the calculated temperature-dependent anisotropy of the coherence length: γ ξ ( T ) = γ ξ ( 0 ) 1 1 1 a × T T c , where all γ ξ ( T ) values for temperatures in the range of T c T < a × T c T γ are shown by the dashed line.

3.3. Pr0.8Sr0.2NiO2 Film

Wang et al. [103] reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets for Pr0.8Sr0.2NiO2 film (which has a physical thickness d s c ~ 7   n m ) defined by the R ( T , B ) R n o r m ( T , B = 0 ) = 0.50 criterion. In Figure 3 we show reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets and global data fit to the 2D-GL model (Figure 3a,b; Equations (1) and (2)) and to our model (Figure 3c–e; Equations (7)–(9)). The quality of both fits is high, and the fitting parameters have low mutual dependence.
Overall, inequalities (similar to those obtained for the other nickelates (Equations (12)–(16)) were revealed for the Pr0.8Sr0.2NiO2 film:
ξ a b ( 0 ) 4   n m < d s c ~ 7   n m < d s c , G L = 9.5   n m ,

3.4. Nd0.825Sr0.175NiO2 Film

Wang et al. [103] reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets for Nd0.825Sr0.175NiO2 film (which has a physical thickness d s c ~ 7   n m ) defined by the R ( T , B ) R n o r m ( T , B = 0 ) = 0.50 criterion. In Figure 4, we show reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets and global data fit to the 2D-GL model (Figure 4a,b; Equations (1) and (2)) and to our model (Figure 4c–e; Equations (7)–(9)). The quality of both fits is high, and the fitting parameters have low mutual dependence.
Deduced coherence lengths from the fit to our model (Equations (7)–(9)) are:
ξ c ( 0 ) 3.4   n m < ξ a b ( 0 ) 4.0   n m < d s c ~ 7   n m ,
which confirmed the 3D superconductivity of the Nd0.825Sr0.175NiO2 film.

3.5. La0.8Sr0.2NiO2 Film

Now we return to the La0.8Sr0.2NiO2 compound, for which Wei et al. [70] recently reported the record high-superconducting transition temperature for nickelates. Wei et al. [70] also reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets defined by the R ( T , B ) R n o r m ( T , B = 0 ) = 0.01 , R ( T , B ) R n o r m ( T , B = 0 ) = 0.50 , and R ( T , B ) R n o r m ( T , B = 0 ) = 0.90 criteria. Despite our understanding that B c 2 ( T ) should be defined by the lowest possible R ( T , B ) R n o r m ( T , B = 0 ) criterion, in Figure 5 we analyze the B c 2 ( T ) data defined by the R ( T , B ) R n o r m ( T , B = 0 ) = 0.50 criterion [70] to make it possible to make a comparison of deduced parameters for the La0.8Sr0.2NiO2 film in Section 3.2 (Figure 2). The film thickness is d s c ~ 6.8   n m [70], which is practically the same as the one in the report by Wang et al. [103]. In Figure 5a,b we show B c 2 ( T ) data and global data fit to the 2D-GL model (Equations (1) and (2)) and to our model (Figure 5c–e); Equations (7)–(9)).
Deduced coherence lengths from the fit to our model (Equations (7)–(9)) are:
ξ c ( 0 ) 3.4   n m < ξ a b ( 0 ) 4.0   n m < d s c ~ 7   n m ,
confirm the 3D superconductivity of the La0.8Sr0.2NiO2 film.
In the following sections, we demonstrate that the high-quality fit of B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets to Equations (1) and (2) cannot be considered as evidence for 2D superconductivity because we obtained high-quality fits to Equations (1) and (2) for B c 2 ( T ) data for bulk iron-based superconductors (IBS).
IBS were experimentally discovered by Hosono’s group [104,105] more than 15 years ago, and to the best of our knowledge, there has been no proposal for an analytical expression for the temperature-dependent coherence length anisotropy, γ ξ ( T ) , to this family of superconductors. Here, we show that our 3D model (Equations (7)–(9)) can be extended to IBS materials.

3.6. Bulk Tl0.58Rb0.42Fe1.72Se2

Jiao et al. [106] reported R ( T , B p e r p ) and R ( T , B p a r a ) datasets for bulk single crystals of Tl0.58Rb0.42Fe1.72Se2. These authors [106] derived extrapolative values for B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) defined by the R ( T , B ) R n o r m ( T , B = 0 ) 0.0 and R ( T , B ) R n o r m ( T , B = 0 ) 1.0 criteria. Because these datasets of extrapolated values do not represent values measured in the experiment, in Figure 6 we analyze datasets, deduced by the R ( T , B ) R n o r m ( T , B = 0 ) = 0.50 criterion, which represent the measured values.
In Figure 6a,b, we fit B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets to Equations (1) and (2) to prove that high-quality fits to the 2D model can be obtained (and, even, “the thickness” of the superconductor, d s c , G L , can be deduced) for B c 2 ( T ) data measured for bulk anisotropic superconductors.
This implies (Figure 6a,b) that the thickness, d s c , G L , of the “2D superconductor” can be deduced from B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets measured for bulk superconductors by utilizing the widely used [54,107] Equations (1) and (2) proposed by Ginzburg and Landau in the early 1960s [52,53].
Thus, we argue that Equations (1) and (2) are incorrect to use in data analysis because these equations represent reasonably flexible fitting functions, which can be used to smooth data for some superconductors. However, one parameter in these equations—that is, d s c , G L —which exhibits a unit of length, does not have any physical meaning for bulk superconductors.
This implies that the traditional interpretation (see, for instance, [23]) that the violation of the following equation:
d s c d s c , G L ,
in thin film superconductors (including atomically thin superconductors) should indicate that there is a deep underlying physical effect (for instance, spin-orbit scattering [108,109]), cannot be found to be valid.
In Figure 6c,d, we show the B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets for Tl0.58Rb0.42Fe1.72Se2 which were fitted to Equations (7)–(9). The fits are of high quality. The deduced parameters:
a = 1.29 ± 0.03 ,
γ ξ ( 0 ) = 1.24 ± 0.04 ,
are within the same ranges as those deduced for the nickelate films (Figure 1, Figure 2, Figure 3, Figure 4 and Figure 5). This is evidence that the nickelates exhibit 3D superconductivity.
However, Figure 6 demonstrates that the 3D model (Equations (7)–(9)) can be extended to a broader range of superconductors, in particular, on bulk chalcogenides. To demonstrate this, in the next section we apply Equations (7)–(9) to another bulk single-crystal chalcogenide superconductor, Fe1.11Te0.6Se0.4 [110].

3.7. Bulk Fe1.11Te0.6Se0.4

Fang et al. [110] reported B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets for bulk single crystals of Fe1.11Te0.6Se0.4 [110] defined by the R ( T , B ) R n o r m ( T , B = 0 ) = 0.05 criterion. In Figure 7a,b we fit B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets to Equations (1) and (2) to demonstrate that high-quality fits to the 2D-GL model can be obtained.
Figure 7a,b show that “the thickness of the 2D superconductor”, d s c , G L , of several nanometers (i.e., within a typical range usually deduced for thin film superconductors, including nickelates) can be deduced from the fit for this bulk anisotropic superconductor.
In Figure 7c,d, the same B c 2 , p e r p ( T ) and B c 2 , p a r a ( T ) datasets (as shown in Figure 7a,b, respectively), were fitted to Equations (7)–(9). The deduced parameters are in the expected ranges.
However, the ground-state anisotropy of the coherence length is less than unity:
γ ξ ( 0 ) = 0.88 ± 0.01 < 1.0
It should be noted that γ ξ ( 0 ) < 1.0 was reported for several iron-based superconductors, and this topic has been discussed (see, for instance Refs. [111,112]).

3.8. Bulk KFe2As2

Zocco et al. [113] reported B c 2 , p e r p ( T ) , B c 2 , p a r a ( T ) , and γ ξ ( T ) = B c 2 , p a r a ( T ) B c 2 , p e r p ( T ) datasets for a bulk single-crystal KFe2As2 superconductor. In Figure 8a, the reported γ ξ ( T ) is fitted to Equation (9) (for this fit, we fixed the transition temperature to the value observed in the experiment, T c = 3.4   ( f i x e d ) ).
The fit is of high quality and has low mutual parameters dependence. The deduced parameters (Figure 8) are within the ranges reported above for nickelates and iron-based superconductors.

3.9. Bulk LiFeAs

Khim et al. [114] and Zhang et al. [115] reported B c 2 , p e r p ( T ) , B c 2 , p a r a ( T ) , and γ ξ ( T ) = B c 2 , p a r a ( T ) B c 2 , p e r p ( T ) datasets for a bulk single-crystal pnictide LiFeAs superconductor. In Figure 8b,c we show the fits of the reported γ ξ ( T ) to Equation (9) (for this fit, we set the transition temperature to the value observed in the experiments).
The fits are of a high quality and have low parameters dependence. The deduced parameters for two datasets reported by independent research groups are close to each other, within acceptable levels of parameter differences.

4. Discussion

The physical origin of our model (which is primarily based on Equation (9)) can be understood based on an analogy with the temperature-dependent DC magnetic susceptibility, χ ( T ) , in antiferromagnetic materials [116,117]. The temperature-dependent χ ( T ) in any material obeys the Curie–Weiss law (Figure 9):
χ ( T ) = C T θ ,
where θ is Curie–Weiss temperature, and C is the Curie constant.
In the schematic representations of Equation (24) in Figure 9, there are three types of magnetic materials that primarily depend on the sign of the Curie–Weiss temperature:
  • θ > 0   K for ferromagnetic materials;
  • θ = 0   K for paramagnetic materials;
  • θ < 0   K for antiferromagnetic materials.
To be consistent with the form of Equation (24), we can rewrite Equation (9):
γ ( T ) = a T c γ ( 0 ) a T c T = a T c γ ( 0 ) T a T c = C T θ ,
where C = a T c γ ( 0 ) , and θ = a T c .
Despite the negative sign (in K units) of the Curie–Weiss temperature, θ , for antiferromagnetic materials, this value represents one of the fundamental constants of the antiferromagnet, which quantifies the strength of the antiferromagnetic interaction in the material.
In antiferromagnetic materials, the χ ( T ) would obey the Curie–Weiss law down to very low temperatures, T 0   K (Figure 9c). However, at the Neel temperature, T N > 0   K , a phase transition occurs, and the χ ( T ) does not obey the Curie–Weiss law at T < T N .
In our model, γ ξ ( T ) would obey the Equation (9) up to high temperatures, T a T c T γ ; however, at T = T c , a superconductor–normal state phase transition occurs, and γ ξ ( T ) becomes undefined at T > T c . However, the latter does not mean that a T c T ξ does not represent any physical value for the material, and our current interpretation of this value is that T = a T c T γ represents the threshold temperature for the appearance of the anisotropy in the fluctuations of the order parameter in a superconductor.
It should also be stressed that our model (Equations (7)–(9)) utilizes the simplistic GL expression for the temperature-dependent in-plane coherence length:
ξ a b ( T ) = ξ a b ( 0 ) 1 T T c ,  
Perhaps it would be more accurate to use the Werthamer–Helfand–Hohenberg (WHH) theory [118] or its advanced version developed for two-band superconductors by Gurevich [119]. This type of advanced analysis, in conjunction with high-field experimental studies, has been implemented in several studies on IBS [120,121] and nickelates [23].
However, the high flexibility of primary WHH functions (exhibited several parameters, especially for the two-band model), and the nonexistence of WHH functions for T > T c , makes it impossible to extract a simple analytical expression, similar to Equation (9), for the temperature dependence of the coherence length anisotropy, γ ( T ) , which we propose herein. That is, the fact that:
ξ a b ( T ) | T T c ,  
ξ c ( T ) | T T c ,  
does not exclude the following:
γ ξ ( T ) = ξ a b ( T ) ξ c ( T ) | T T c .
The rigorous mathematical expression for the primary message of this study is as follows:
T a T c < 1 ,     γ ξ ( T ) > 0 ,  
where the standard mathematical symbols are used.
In this regard, the simple heuristic expression for γ ξ ( T ) (proposed herein (that is, Equation (9))) might be modified to be more accurate (and unfortunately, more complicated); however, our primary message that the anisotropy of the thermodynamic fluctuations in anisotropic superconductors (exhibiting low charge-carrier density) is established at some temperature T γ a T c ( a > 1 ) above the entire superconducting transition temperature, T c , should remain unchanged.
The hypothetical analogy between temperature dependencies for some primary physical quantities for the antiferromagnetism and the superconductivity phenomena proposed herein (Figure 9) does not have any deeper meaning, except that the shape of the temperature dependencies can be similar. However, perhaps there is a much deeper physical ground between these two physical phenomena, because recently Fowlie et al. [37] showed that nickelates (regardless of the rare earth ion or doping) exhibit an intrinsic magnetic ground state arising from the short-range antiferromagnetic order of the nickel sublattice. Considering that the interplay/relation between the magnetism and the superconductivity in other unconventional superconductors has been discussed for decades [122,123,124], there is a chance that deeper relations can be revealed by further theoretical and experimental investigations.
All superconductors exhibit the second fundamental characteristic length, which is the London penetration depth, λ ( T ) . In anisotropic superconductors, the λ ( T ) also has two components, which are in-plane London penetration depth, λ a b ( T ) , and out-of-plane London penetration depth, λ c ( T ) . We expect that the same approach, to that described herein for γ ξ ( T ) , should be applied for the temperature-dependent London penetration depth anisotropy, γ λ ( T ) = λ a b ( T ) λ c ( T ) [72,125,126,127]. However, the discussion of this topic is far beyond the framework of the current study.

5. Conclusions

In this study, we analyze the temperature dependence of the upper critical field anisotropy and the coherence length anisotropy, γ ξ ( T ) , in nickelate superconductors for which we propose a simple heuristic expression (Equation (9)) within the 3D Ginzburg-Landau model. The proposed expression for γ ξ ( T ) (Equation (9)) is also applicable to some chalcogenide and pnictide superconductors.
Overall, we show that 2D Ginzburg–Landau model, proposed by Tinkham and co-workers [52,107] (and which was the dominant model to describe the upper critical field in thin film superconductors for more than 50 years) is not a unique approach to describe experimental upper critical field data in thin film superconductors, and alternative 3D models can be used to analyze experimental data for thin film superconductors (which satisfy the condition ξ c d s c ).

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.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Li, D.; Lee, K.; Wang, B.Y.; Osada, M.; Crossley, S.; Lee, H.R.; Cui, Y.; Hikita, Y.; Hwang, H.Y. Superconductivity in an infinite-layer nickelate. Nature 2019, 572, 624–627. [Google Scholar] [CrossRef]
  2. Anisimov, V.I.; Bukhvalov, D.; Rice, T.M. Electronic structure of possible nickelate analogs to the cuprates. Phys. Rev. B 1999, 59, 7901. [Google Scholar] [CrossRef]
  3. Nomura, Y.; Hirayama, M.; Tadano, T.; Yoshimoto, Y.; Nakamura, K.; Arita, R. Formation of a two-dimensional single component correlated electron system and band engineering in the nickelate superconductor NdNiO2. Phys. Rev. B 2019, 100, 205138. [Google Scholar] [CrossRef]
  4. Zeng, S.; Tang, C.S.; Yin, X.; Li, C.; Li, M.; Huang, Z.; Hu, J.; Liu, W.; Omar, G.J.; Jani, H.; et al. Phase diagram and superconducting dome of infinite layer Nd1-xSrxNiO2 thin films. Phys. Rev. Lett. 2020, 125, 147003. [Google Scholar] [CrossRef]
  5. Osada, M.; Wang, B.Y.; Goodge, B.H.; Lee, K.; Yoon, H.; Sakuma, K.; Li, D.; Miura, M.; Kourkoutis, L.F.; Hwang, H.Y. A superconducting praseodymium nickelate with infinite layer structure. Nano Lett. 2020, 20, 5735. [Google Scholar] [CrossRef] [PubMed]
  6. Gu, Q.; Li, Y.; Wan, S.; Li, H.; Guo, W.; Yang, H.; Li, Q.; Zhu, X.; Pan, X.; Nie, Y.; et al. Single particle tunnelling spectrum of superconducting Nd1-xSrxNiO2 thin films. Nat. Commun. 2020, 11, 6027. [Google Scholar] [CrossRef]
  7. Talantsev, E.F. Classifying superconductivity in an infinite-layer nickelate Nd0.8Sr0.2NiO2. Results Phys. 2020, 17, 103118. [Google Scholar] [CrossRef]
  8. Wang, Y.; Kang, C.-J.; Miao, H.; Kotliar, G. Hund’s metal physics: From SrNiO2 to LaNiO2. Phys. Rev. B 2020, 102, 161118. [Google Scholar] [CrossRef]
  9. Li, D.; Wang, B.Y.; Lee, K.; Harvey, S.P.; Osada, M.; Goodge, B.H.; Kourkoutis, L.F.; Hwang, H.Y. Superconducting dome in Nd1-xSrxNiO2 infinite layer films. Phys. Rev. Lett. 2020, 125, 027001. [Google Scholar] [CrossRef]
  10. Jiang, M.; Berciu, M.; Sawatzky, G.A. Critical nature of the Ni spin state in doped NdNiO2. Phys. Rev. Lett. 2020, 124, 207004. [Google Scholar] [CrossRef] [PubMed]
  11. Lechermann, F. Late transition metal oxides with infinite-layer structure: Nickelates versus cuprates. Phys. Rev. B 2020, 101, 081110. [Google Scholar] [CrossRef]
  12. Wu, X.; Sante, D.D.; Schwemmer, T.; Hanke, W.; Hwang, H.Y.; Raghu, S.; Thomale, R. Robust dx2-y2-wave superconductivity of infinite-layer nickelates. Phys. Rev. B 2020, 101, 060504. [Google Scholar] [CrossRef]
  13. Botana, A.S.; Norman, M.R. Similarities and differences between LaNiO2 and CaCuO2 and implications for superconductivity. Phys. Rev. X 2020, 10, 011024. [Google Scholar] [CrossRef]
  14. Sakakibara, H.; Usui, H.; Suzuki, K.; Kotani, T.; Aoki, H.; Kuroki, K. Model construction and a possibility of cupratelike pairing in a new d9 nickelate superconductor (Nd,Sr)NiO2. Phys. Rev. Lett. 2020, 125, 077003. [Google Scholar] [CrossRef]
  15. Kitatani, M.; Si, L.; Janson, O.; Arita, R.; Zhong, Z.; Held, K. Nickelate superconductors—A renaissance of the one-band Hubbard model. npj Quantum Mater. 2020, 5, 59. [Google Scholar] [CrossRef]
  16. Adhikary, P.; Bandyopadhyay, S.; Das, T.; Dasgupta, I.; Saha-Dasgupta, T. Orbital-selective superconductivity in a two-band model of infinite-layer nickelates. Phys. Rev. B 2020, 102, 100501. [Google Scholar] [CrossRef]
  17. Wang, Z.; Zhang, G.-M.; Yang, Y.-f.; Zhang, F.-C. Distinct pairing symmetries of superconductivity in infinite-layer nickelates. Phys. Rev. B 2020, 102, 220501. [Google Scholar] [CrossRef]
  18. Werner, P.; Hoshino, S. Nickelate superconductors: Multiorbital nature and spin freezing. Phys. Rev. B 2020, 101, 041104. [Google Scholar] [CrossRef]
  19. Zhang, Y.-H.; Vishwanath, A. Type-II t—J model in superconducting nickelate Nd1−xSrxNiO2. Phys. Rev. Res. 2020, 2, 023112. [Google Scholar] [CrossRef]
  20. Lechermann, F. Multiorbital processes rule the Nd1-xSrxNiO2 normal state. Phys. Rev. X 2020, 10, 041002. [Google Scholar] [CrossRef]
  21. Leonov, I.; Skornyakov, S.L.; Savrasov, S.Y. Lifshitz transition and frustration of magnetic moments in infinite-layer NdNiO2 upon hole doping. Phys. Rev. B 2020, 101, 241108. [Google Scholar] [CrossRef]
  22. Xiang, Y.; Li, Q.; Li, Y.; Yang, H.; Nie, Y.; Wen, H.-H. Physical properties revealed by transport measurements for superconducting Nd0.8Sr0.2NiO2 thin films. Chin. Phys. Lett. 2021, 38, 047401. [Google Scholar] [CrossRef]
  23. Wang, B.Y.; Li, D.; Goodge, B.H.; Lee, K.; Osada, M.; Harvey, S.P.; Kourkoutis, L.F.; Beasley, M.R.; Hwang, H.Y. Isotropic Pauli-limited superconductivity in the infinite layer nickelate Nd0.775Sr0.225NiO2. Nat. Phys. 2021, 17, 473–477. [Google Scholar] [CrossRef]
  24. Goodge, B.H.; Li, D.; Lee, K.; Osada, M.; Wang, B.Y.; Sawatzky, G.A.; Hwang, H.Y.; Kourkoutis, L.F. Doping evolution of the Mott-Hubbard landscape in infinite-layer nickelates. Proc. Natl. Acad. Sci. USA 2021, 118, e2007683118. [Google Scholar] [CrossRef] [PubMed]
  25. Botana, A.S.; Bernardini, F.; Cano, A. Nickelate superconductors: An ongoing dialog between theory and experiments. J. Exp. Theor. Phys. 2021, 132, 618–627. [Google Scholar] [CrossRef]
  26. Lechermann, F. Doping-dependent character and possible magnetic ordering of NdNiO2. Phys. Rev. Mater. 2021, 5, 044803. [Google Scholar] [CrossRef]
  27. Wan, X.; Ivanov, V.; Resta, G.; Leonov, I.; Savrasov, S.Y. Exchange interactions and sensitivity of the Ni two-hole spin state to Hund’s coupling in doped NdNiO2. Phys. Rev. B 2021, 103, 075123. [Google Scholar] [CrossRef]
  28. Kang, C.-J.; Kotliar, G. Optical properties of the infinite-layer La1−xSrxNiO2 and hidden Hund’s physics. Phys. Rev. Lett. 2021, 126, 127401. [Google Scholar] [CrossRef]
  29. Lu, H.; Rossi, M.; Nag, A.; Osada, M.; Li, D.F.; Lee, K.; Wang, B.Y.; Garcia-Fernandez, M.; Agrestini, S.; Shen, Z.X.; et al. Magnetic excitations in infinite-layer nickelates. Science 2021, 373, 213–216. [Google Scholar] [CrossRef]
  30. Lin, J.Q.; Villar Arribi, P.; Fabbris, G.; Botana, A.S.; Meyers, D.; Miao, H.; Shen, Y.; Mazzone, D.G.; Feng, J.; Chiuzbaian, S.G.; et al. Strong super-exchange in a d9-δ nickelate revealed by resonant inelastic X-ray scattering. Phys. Rev. Lett. 2021, 126, 087001. [Google Scholar] [CrossRef] [PubMed]
  31. Leonov, I. Effect of lattice strain on the electronic structure and magnetic correlations in infinite-layer (Nd,Sr)NiO2. J. Alloys Compd. 2021, 883, 160888. [Google Scholar] [CrossRef]
  32. Choubey, P.; Eremin, I.M. Electronic theory for scanning tunneling microscopy spectra in infinite layer nickelate superconductors. Phys. Rev. B 2021, 104, 144504. [Google Scholar] [CrossRef]
  33. Zhou, X.-R.; Feng, Z.-X.; Qin, P.-X.; Yan, H.; Wang, X.-N.; Nie, P.; Wu, H.-J.; Zhang, X.; Chen, H.-Y.; Meng, Z.-A.; et al. Negligible oxygen vacancies, low critical current density, electric-field modulation, in-plane anisotropic and high-field transport of a superconducting Nd0.8Sr0.2NiO2/SrTiO3 heterostructure. Rare Met. 2021, 40, 2847–2854. [Google Scholar] [CrossRef]
  34. Chow, L.E.; Sudheesh, S.K.; Nandi, P.; Zeng, S.W.; Zhang, Z.T.; Du, X.M.; Lim, Z.S.; Chia, E.E.; Ariando, A. Pairing symmetry in infinite-layer nickelate superconductor. arXiv 2022, arXiv:2201.10038. [Google Scholar] [CrossRef]
  35. Harvey, S.P.; Wang, B.Y.; Fowlie, J.; Osada, M.; Lee, K.; Lee, Y.; Li, D.; Hwang, H.Y. Evidence for nodal superconductivity in infinite-layer nickelates. arXiv 2022, arXiv:2201.12971. [Google Scholar] [CrossRef]
  36. Chow, L.E.; Yip, K.Y.; Pierre, M.; Zeng, S.W.; Zhang, Z.T.; Heil, T.; Deuschle, J.; Nandi, P.; Sudheesh, S.K.; Lim, Z.S.; et al. Pauli-limit violation in lanthanide infinite-layer nickelate superconductors. arXiv 2022, arXiv:2204.12606. [Google Scholar] [CrossRef]
  37. Fowlie, J.; Hadjimichael, M.; Martins, M.M.; Li, D.; Osada, M.; Wang, B.Y.; Lee, K.; Lee, Y.; Salman, Z.; Prokscha, T.; et al. Intrinsic magnetism in superconducting infinite-layer nickelates. Nat. Phys. 2022, 18, 1043–1047. [Google Scholar] [CrossRef]
  38. Chen, H.; Hampel, A.; Karp, J.; Lechermann, F.; Millis, A. Dynamical mean field studies of infinite layer nickelates: Physics results and methodological implications. Front. Phys. 2022, 10, 835942. [Google Scholar] [CrossRef]
  39. Lu, C.; Hu, L.-H.; Wang, Y.; Yang, F.; Wu, C. Two-orbital model for possible superconductivity pairing mechanism in nickelates. Phys. Rev. B 2022, 105, 054516. [Google Scholar] [CrossRef]
  40. Xie, T.Y.; Liu, Z.; Cao, C.; Wang, Z.F.; Yang, J.L.; Zhu, W. Microscopic theory of superconducting phase diagram in infinite-layer nickelates. Phys. Rev. B 2022, 106, 035111. [Google Scholar] [CrossRef]
  41. Jiang, M. Characterizing the superconducting instability in a two-orbital d-s model: Insights to infinite-layer nickelate superconductors. arXiv 2022, arXiv:2201.12967. [Google Scholar] [CrossRef]
  42. Karp, J.; Hampel, A.; Millis, A.J. Superconductivity and antiferromagnetism in NdNiO2 and CaCuO2: A cluster DMFT study. Phys. Rev. B 2022, 105, 205131. [Google Scholar] [CrossRef]
  43. Kreisel, A.; Andersen, B.M.; Roemer, A.T.; Eremin, I.M.; Lechermann, F. Superconducting instabilities in strongly-correlated infinite-layer nickelates. Phys. Rev. Lett. 2022, 129, 077002. [Google Scholar] [CrossRef]
  44. Krieger, G.; Martinelli, L.; Zeng, S.; Chow, L.E.; Kummer, K.; Arpaia, R.; Moretti Sala, M.; Brookes, N.B.; Ariando, A.; Viart, N.; et al. Charge and spin order dichotomy in NdNiO2 driven by the SrTiO3 capping layer. Phys. Rev. Lett. 2022, 129, 027002. [Google Scholar] [CrossRef]
  45. Zeng, S.W.; Yin, Z.M.; Li, C.J.; Chow, L.E.; Tang, C.S.; Han, K.; Huang, Z.; Cao, Y.; Wan, D.Y.; Zhang, Z.T.; et al. Observation of perfect diamagnetism and interfacial effect on the electronic structures in infinite layer Nd0.8Sr0.2NiO2 superconductors. Nat. Commun. 2022, 13, 743. [Google Scholar] [CrossRef] [PubMed]
  46. Tam, C.C.; Choi, J.; Ding, X.; Agrestini, S.; Nag, A.; Wu, M.; Huang, B.; Luo, H.; Gao, P.; García-Fernández, M.; et al. Charge density waves in infinite-layer NdNiO2 nickelates. Nat. Mater. 2022, 21, 1116–1120. [Google Scholar] [CrossRef]
  47. Rossi, M.; Osada, M.; Choi, J.; Agrestini, S.; Jost, D.; Lee, Y.; Lu, H.; Wang, B.Y.; Lee, K.; Nag, A.; et al. A broken translational symmetry state in an infinite-layer nickelate. Nat. Phys. 2022, 18, 869–873. [Google Scholar] [CrossRef]
  48. Pan, G.A.; Segedin, D.F.; LaBollita, H.; Song, Q.; Nica, E.M.; Goodge, B.H.; Pierce, A.T.; Doyle, S.; Novakov, S.; Carrizales, D.C.; et al. Superconductivity in a quintuple-layer square-planar nickelate. Nat. Mater. 2022, 21, 160–164. [Google Scholar] [CrossRef] [PubMed]
  49. Wang, N.N.; Yang, M.W.; Yang, Z.; Chen, K.Y.; Zhang, H.; Zhang, Q.H.; Zhu, Z.H.; Uwatoko, Y.; Gu, L.; Dong, X.L.; et al. Pressure-induced monotonic enhancement of Tc to over 30 K in superconducting Pr0.82Sr0.18NiO2 thin films. Nat. Commun. 2022, 13, 4367. [Google Scholar] [CrossRef]
  50. Chow, L.E.; Rubi, K.; Yip, K.Y.; Pierre, M.; Leroux, M.; Liu, X.; Luo, Z.; Zeng, S.; Li, C.; Goiran, M.; et al. Dimensionality control and rotational symmetry breaking superconductivity in square-planar layered nickelates. arXiv 2023, arXiv:2301.07606. [Google Scholar] [CrossRef]
  51. Li, Q.; He, C.; Si, J.; Zhu, X.; Zhang, Y.; Wen, H.-H. Absence of superconductivity in bulk Nd1−xSrxNiO2. Commun. Mater. 2020, 1, 16. [Google Scholar] [CrossRef]
  52. Tinkham, M. Introduction to Superconductivity, 2nd ed.; McGraw Hill: New York, NY, USA, 1996; pp. 141–143. [Google Scholar]
  53. Harper, F.; Tinkham, M. The mixed state in superconducting thin films. Phys. Rev. 1968, 172, 441–450. [Google Scholar] [CrossRef]
  54. Cao, Y.; Park, J.M.; Watanabe, K.; Taniguchi, T.; Jarillo-Herrero, P. Pauli-limit violation and re-entrant superconductivity in moiré graphene. Nature 2021, 595, 526–531. [Google Scholar] [CrossRef] [PubMed]
  55. Saito, Y.; Nojima, T.; Iwasa, Y. Highly crystalline 2D superconductors. Nat. Rev. Mater. 2017, 2, 16094. [Google Scholar] [CrossRef]
  56. Buntar, V.; Sauerzopf, F.M.; Weber, H.W.; Fischer, J.E.; Kuzmany, H.; Halushka, M. Mixed-state parameters and vortex pinning in single-crystalline K3C60 fullerene superconductors. Phys. Rev. B 1997, 56, 14128–14137. [Google Scholar] [CrossRef]
  57. Perkins, G.K.; Moore, J.; Bugoslavsky, Y.; Cohen, L.F.; Jun, J.; Kazakov, S.M.; Karpinski, J.; Caplin, A.D. Superconducting critical fields and anisotropy of a MgB2 single crystal. Supercond. Sci. Technol. 2002, 15, 1156–1159. [Google Scholar] [CrossRef]
  58. Zehetmayer, M.; Eisterer, M.; Jun, J.; Kazakov, S.M.; Karpinski, J.; Wisniewski, A.; Weber, H.W. Mixed-state properties of superconducting MgB2 single crystals. Phys. Rev. B 2002, 66, 052505. [Google Scholar] [CrossRef]
  59. Honda, F.; Kobayashi, S.; Kawamura, N.; Kawaguchi, S.; Koizumi, T.; Sato, Y.J.; Homma, Y.; Ishimatsu, N.; Gouchi, J.; Uwatoko, Y.; et al. Pressure-induced structural transition and new superconducting phase in UTe2. J. Phys. Soc. Jpn. 2023, 92, 044702. [Google Scholar] [CrossRef]
  60. Kasahara, S.; Suzuki, H.; Machida, T.; Sato, Y.; Ukai, Y.; Murayama, H.; Suetsugu, S.; Kasahara, Y.; Shibauchi, T.; Hanaguri, T.; et al. Quasiparticle nodal plane in the Fulde-Ferrell-Larkin-Ovchinnikov state of FeSe. Phys. Rev. Lett. 2021, 127, 257001. [Google Scholar] [CrossRef]
  61. Zhao, C.C.; Wang, L.-S.; Xia, W.; Yin, Q.W.; Ni, J.M.; Huang, Y.Y.; Tu, C.P.; Tao, Z.C.; Tu, Z.J.; Gong, C.S.; et al. Nodal superconductivity and superconducting domes in the topological Kagome metal CsV3Sb5. arXiv 2021, arXiv:2102.08356. [Google Scholar] [CrossRef]
  62. Tayama, T.; Harita, A.; Sakakibara, T.; Haga, Y.; Shishido, H.; Settai, R.; Onuki, Y. Unconventional heavy-fermion superconductor CeCoIn5: Dc magnetization study at temperatures down to 50 mK. Phys. Rev. B 2002, 65, 180504. [Google Scholar] [CrossRef]
  63. Yan, D.Y.; Yang, M.; Wang, C.X.; Song, P.B.; Yi, C.J.; Shi, Y.G. Superconductivity in centrosymmetric topological superconductor candidate TaC. Supercond. Sci. Technol. 2021, 34, 035025. [Google Scholar] [CrossRef]
  64. Yan, D.; Geng, D.; Gao, Q.; Cui, Z.; Yi, C.; Feng, Y.; Song, C.; Luo, H.; Yang, M.; Arita, M.; et al. Superconductivity and Fermi-surface nesting in the candidate Dirac semimetal NbC. Phys. Rev. B 2020, 102, 205117. [Google Scholar] [CrossRef]
  65. Li, Y.; Garcia, J.; Franco, G.; Lu, J.; Lu, K.; Rong, B.; Shafiq, B.; Chen, N.; Liu, Y.; Liu, L.; et al. Critical magnetic fields of superconducting aluminum-substituted Ba8Si42Al4 clathrate. J. Appl. Phys. 2015, 117, 213912. [Google Scholar] [CrossRef]
  66. Buta, F.; Bonura, M.; Matera, D.; Bovone, G.; Ballarino, A.; Hopkins, S.C.; Bordini, B.; Chaud, X.; Senatore, C. Very high upper critical fields and enhanced critical current densities in Nb3Sn superconductors based on Nb–Ta–Zr alloys and internal oxidation. J. Phys. Mater. 2021, 4, 025003. [Google Scholar] [CrossRef]
  67. Skourski, Y.; Fuchs, G.; Kerschl, P.; Kozlov, N.; Eckert, D.; Nenkov, K.; Mueller, K.-H. Magnetization and magneto-resistance measurements of bulk YBa2Cu3O7-x in pulsed magnetic fields up to 50 T. Phys. B 2004, 346, 325–328. [Google Scholar] [CrossRef]
  68. Zhang, C.L.; He, X.; Li, Z.W.; Zhang, S.J.; Min, B.S.; Zhang, J.; Lu, K.; Zhao, J.K.; Shi, L.C.; Peng, Y.; et al. Superconductivity above 80 K in polyhydrides of hafnium. Mater. Today Phys. 2022, 27, 100826. [Google Scholar] [CrossRef]
  69. Vedeneev, S.I.; Piot, B.A.; Maude, D.K.; Sadakov, A.V. Temperature dependence of the upper critical field of FeSe single crystals. Phys. Rev. B 2013, 87, 134512. [Google Scholar] [CrossRef]
  70. Wei, W.; Sun, W.; Sun, Y.; Pan, Y.; Jin, G.; Yang, F.; Li, Y.; Zhu, Z.; Nie, Y.; Shi, Z. Large upper critical fields and dimensionality crossover of superconductivity in infinite-layer nickelate La0.8Sr0.2NiO2. arXiv 2023, arXiv:2304.14196. [Google Scholar] [CrossRef]
  71. Zhang, H.; Zhong, W.; Meng, Y.; Yue, B.; Yu, X.; Wang, J.-T.; Hong, F. Superconductivity above 12 K with possible multiband features in CsCl-type PbS. Phys. Rev. B 2023, 107, 174502. [Google Scholar] [CrossRef]
  72. Gupta, R.; Das, D.; Mielke, C.H., III; Guguchia, Z.; Shiroka, T.; Baines, C.; Bartkowiak, M.; Luetkens, H.; Khasanov, R.; Yin, Q.; et al. Microscopic evidence for anisotropic multigap superconductivity in the CsV3Sb5 kagome superconductor. Npj Quantum Mater. 2022, 7, 49. [Google Scholar] [CrossRef]
  73. Ruan, B.-B.; Chen, L.-W.; Shi, Y.-Q.; Yi, J.-K.; Yang, Q.-S.; Zhou, M.-H.; Ma, M.-W.; Chen, G.-F.; Ren, Z.-A. Superconductivity in Mo4Ga20As with endohedral gallium clusters. J. Phys. Condens. Matter 2023, 35, 214002. [Google Scholar] [CrossRef]
  74. Shang, T.; Meng, J.; Zhu, X.Y.; Zhang, H.; Yu, B.C.; Zhen, Z.X.; Wang, Y.H.; Xu, Y.; Zhan, Q.F.; Gawryluk, D.J.; et al. Fully-gapped superconductivity with preserved time-reversal symmetry in NiBi3 single crystals. Phys. Rev. B 2023, 107, 174513. [Google Scholar] [CrossRef]
  75. Xing, Y.; Zhang, H.-M.; Fu, H.-L.; Liu, H.; Sun, Y.; Peng, J.-P.; Wang, F.; Lin, X.; Ma, X.-C.; Xue, Q.K.; et al. Quantum Griffiths singularity of superconductor-metal transition in Ga thin films. Science 2015, 350, 542–545. [Google Scholar] [CrossRef]
  76. De la Barrera, S.C.; Sinko, M.R.; Gopalan, D.P.; Sivadas, N.; Seyler, K.L.; Watanabe, K.; Taniguchi, T.; Tsen, A.W.; Xu, X.; Xiao, D.; et al. Tuning Ising superconductivity with layer and spin–orbit coupling in two-dimensional transition-metal dichalcogenides. Nat. Commun. 2018, 9, 1427. [Google Scholar] [CrossRef]
  77. Liu, Y.; Wang, Z.; Shan, P.; Tang, Y.; Liu, C.; Chen, C.; Xing, Y.; Wang, Q.; Liu, H.; Lin, X.; et al. Anomalous quantum Griffiths singularity in ultrathin crystalline lead films. Nat. Commun. 2019, 10, 3633. [Google Scholar] [CrossRef] [PubMed]
  78. Park, J.M.; Cao, Y.; Xia, L.-Q.; Sun, S.; Watanabe, K.; Taniguchi, T.; Jarillo-Herrero, P. Robust superconductivity in magic-angle multilayer graphene family. Nat. Mater. 2022, 21, 877–883. [Google Scholar] [CrossRef]
  79. Bulaevskii, L.N.; Ginzburg, V.L.; Sobyanin, A.A. Macroscopic theory of superconductors with small coherence length. Phys. C 1988, 152, 378–388. [Google Scholar] [CrossRef]
  80. Emery, V.J.; Kivelson, S.A. Importance of phase fluctuations in superconductors with small superfluid density. Nature 1995, 374, 434–437. [Google Scholar] [CrossRef]
  81. Larkin, A.; Varlamov, V. Theory of Fluctuations in Superconductors; Oxford University Press: Oxford, UK, 2005. [Google Scholar]
  82. He, T.; Huang, Q.; Ramirez, A.P.; Wang, Y.; Regan, K.A.; Rogado, N.; Hayward, M.A.; Haas, M.K.; Slusky, J.S.; Inumara, L.; et al. Superconductivity in the non-oxide perovskite MgCNi3. Nature 2001, 411, 54–56. [Google Scholar] [CrossRef]
  83. Chen, J.; Jiao, L.; Zhang, J.L.; Chen, Y.; Yang, L.; Nicklas, M.; Steglich, F.; Yuan, H.Q. BCS-like superconductivity in the noncentrosymmetric compounds NbxRe1−x (0.13 < x < 0.38). Phys. Rev. B 2013, 88, 144510. [Google Scholar] [CrossRef]
  84. Pan, J.; Jiao, W.H.; Hong, X.C.; Zhang, Z.; He, L.P.; Cai, P.L.; Zhang, J.; Cao, G.H.; Li, S.Y. Nodal superconductivity and superconducting dome in the layered superconductor Ta4Pd3Te16. Phys. Rev. B 2015, 92, 180505. [Google Scholar] [CrossRef]
  85. 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]
  86. Zhou, N.; Sun, Y.; Xi, C.Y.; Wang, Z.S.; Zhang, J.L.; Zhang, Y.; Zhang, Y.F.; Xu, C.Q.; Pan, Y.Q.; Feng, J.J.; et al. Disorder-robust high-field superconducting phase of FeSe single crystals. Phys. Rev. B 2021, 104, L140504. [Google Scholar] [CrossRef]
  87. Mandal, M.; Patra, C.; Kataria, A.; Singh, D.; Biswas, P.K.; Lord, J.S.; Hillier, A.D.; Singh, R.P. Superconducting ground state of nonsymmorphic superconducting compound Zr2Ir. Phys. Rev. B 2021, 104, 054509. [Google Scholar] [CrossRef]
  88. 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]
  89. Takenaka, T.; Ishihara, K.; Roppongi, M.; Miao, Y.; Mizukami, Y.; Makita, T.; Tsurumi, J.; Watanabe, S.; Takeya, J.; Yamashita, M.; et al. Strongly correlated superconductivity in a copper-based metal-organic framework with a perfect kagome lattice. Sci. Adv. 2021, 7, eabf3996. [Google Scholar] [CrossRef]
  90. Gui, X.; Cava, R.J. LaIr3Ga2: A superconductor based on a Kagome lattice of Ir. Chem. Mater. 2022, 34, 2824–2832. [Google Scholar] [CrossRef]
  91. Wang, J.; Ying, T.; Deng, J.; Pei, C.; Yu, T.; Chen, X.; Wan, Y.; Yang, M.; Dai, W.; Yang, D.; et al. Superconductivity in an orbital-reoriented SnAs square lattice: A case study of Li0.6Sn2As2 and NaSnAs. Angew. Chem. 2022, 62, e202216086. [Google Scholar] [CrossRef]
  92. Yu, F.H.; Ma, D.H.; Zhuo, W.Z.; Liu, S.Q.; Wen, X.K.; Lei, B.; Ying, J.J.; Chen, X.H. Unusual competition of superconductivity and charge-density-wave state in a compressed topological kagome metal. Nat. Commun. 2021, 12, 3645. [Google Scholar] [CrossRef]
  93. Sun, Y.; Shi, Z.; Tamegai, T. Review of annealing effects and superconductivity in Fe1+yTe1-xSex superconductors. Supercond. Sci. Technol. 2019, 32, 103001. [Google Scholar] [CrossRef]
  94. Yuan, J.; Chen, Q.; Jiang, K.; Feng, Z.; Lin, Z.; Yu, H.; He, G.; Zhang, J.; Jiang, X.; Zhang, X.; et al. Scaling of the strange-metal scattering in unconventional superconductors. Nature 2022, 602, 431–436. [Google Scholar] [CrossRef] [PubMed]
  95. Saito, Y.; Ge, J.; Watanabe, K.; Taniguchi, T.; Young, A.F. Independent superconductors and correlated insulators in twisted bilayer graphene. Nat. Phys. 2020, 16, 926–930. [Google Scholar] [CrossRef]
  96. Shang, T.; Zhao, J.Z.; Hu, L.H.; Gawryluk, D.J.; Zhu, X.Y.; Zhang, H.; Meng, J.; Zhen, Z.X.; Yu, B.C.; Zhou, Z.; et al. Fully-gapped superconductivity and topological aspects of the noncentrosymmetric TaReSi superconductor. arXiv 2023, arXiv:2305.17381. [Google Scholar] [CrossRef]
  97. Zhigadlo, N.D.; Puzniak, R.; Philip, J.W.; Moll, P.J.W.; Bernardini, F.; Shiroka, T. Emergence of superconductivity in single-crystalline LaFeAsO under simultaneous Sm and P substitution. J. Alloys Compd. 2023, 958, 170384. [Google Scholar] [CrossRef]
  98. Leung, C.K.W.; Zhang, X.; von Rohr, F.; Lortz, R.; Jäck, B. Evidence for isotropic s-wave superconductivity in high-entropy alloys. Sci. Rep. 2022, 12, 12773. [Google Scholar] [CrossRef] [PubMed]
  99. 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]
  100. Devore, J.L. Probability and Statistics for Engineering and the Sciences, 8th ed.; Cengage Learning: Boston, MA, USA, 2011; pp. 508–510. [Google Scholar]
  101. Iida, K.; Hänisch, J.; Hata, S.; Yamamoto, A. Recent progress on epitaxial growth of Fe-based superconducting thin films. Supercond. Sci. Technol. 2023, 36, 063001. [Google Scholar] [CrossRef]
  102. Li, D.; Shen, P.; Tian, J.; He, G.; Ni, S.; Wang, Z.; Xi, C.; Pi, L.; Zhang, H.; Yuan, J.; et al. A disorder-sensitive emergent vortex phase identified in high-Tc superconductor (Li,Fe)OHFeSe. Supercond. Sci. Technol. 2022, 35, 064007. [Google Scholar] [CrossRef]
  103. Wang, B.Y.; Wang, T.C.; Hsu, Y.-T.; Osada, M.; Lee, K.; Jia, C.; Duffy, C.; Li, D.; Fowlie, J.; Beasley, M.R.; et al. Rare-earth control of the superconducting upper critical field in infinite-layer nickelates. arXiv 2022, arXiv:2205.15355. [Google Scholar] [CrossRef]
  104. Kamihara, Y.; Hiramatsu, H.; Hirano, M.; Kawamura, R.; Yanagi, H.; Kamiya, T.; Hosono, H. Iron-based layered superconductor:  LaOFeP. J. Am. Chem. Soc. 2006, 128, 10012–10013. [Google Scholar] [CrossRef] [PubMed]
  105. Kamihara, Y.; Watanabe, T.; Hirano, M.; Hosono, H. Iron-based layered superconductor La[O1−xFx]FeAs (x = 0.05−0.12) with Tc = 26 K. J. Am. Chem. Soc. 2008, 130, 3296–3297. [Google Scholar] [CrossRef] [PubMed]
  106. Jiao, L.; Kohama, Y.; Zhang, J.L.; Wang, H.D.; Maiorov, B.; Balakirev, F.F.; Chen, Y.; Wang, L.N.; Shang, T.; Fang, M.H.; et al. Upper critical field and thermally activated flux flow in single-crystalline Tl0.58Rb0.42Fe1.72Se2. Phys. Rev. B 2012, 85, 064513. [Google Scholar] [CrossRef]
  107. Matsuoka, H.; Nakano, M.; Shitaokoshi, T.; Ouchi, T.; Wang, Y.; Kashiwabara, Y.; Yoshida, S.; Ishizaka, K.; Kawasaki, M.; Kohama, Y.; et al. Angle dependence of Hc2 with a crossover between the orbital and paramagnetic limits. Phys. Rev. Res. 2020, 2, 012064. [Google Scholar] [CrossRef]
  108. Wu, X.S.; Adams, P.W.; Yang, Y.; McCarle, R.L. Spin proximity effect in ultrathin superconducting Be–Au bilayers. Phys. Rev. Lett. 2006, 96, 127002. [Google Scholar] [CrossRef]
  109. Shalom, M.B.; Sachs, M.; Rakhmilevitch, D.; Palevski, A.; Dagan, Y. Tuning spin–orbit coupling and superconductivity at the SrTiO3/LaAlO3 interface: A magnetotransport study. Phys. Rev. Lett. 2010, 104, 126802. [Google Scholar] [CrossRef]
  110. Fang, M.; Yang, J.; Balakirev, F.F.; Kohama, Y.; Singleton, J.; Qian, B.; Mao, Z.Q.; Wang, H.; Yuan, H.Q. Weak anisotropy of the superconducting upper critical field in Fe1.11Te0.6Se0.4 single crystals. Phys. Rev. B 2010, 81, 020509. [Google Scholar] [CrossRef]
  111. Bristow, M.; Knafo, W.; Reiss, P.; Meier, W.; Canfield, P.C.; Blundell, S.J.; Coldea, A.I. Competing pairing interactions responsible for the large upper critical field in a stoichiometric iron-based superconductor CaKFe4As4. Phys. Rev. B 2020, 101, 134502. [Google Scholar] [CrossRef]
  112. Zhang, J.L.; Jiao, L.; Chen, Y.; Yuan, H.Q. Universal behavior of the upper critical field in iron based superconductors. Front. Phys. 2011, 6, 463–473. [Google Scholar] [CrossRef]
  113. Zocco, D.A.; Grube, K.; Eilers, F.; Wolf, T.; Löhneysen, H.V. Pauli-limited multiband superconductivity in KFe2As2. Phys. Rev. Lett. 2013, 111, 057007. [Google Scholar] [CrossRef]
  114. Khim, S.; Lee, B.; Kim, J.W.; Choi, E.S.; Stewart, G.R.; Kim, K.H. Pauli-limiting effects in the upper critical fields of a clean LiFeAs single crystal. Phys. Rev. B 2011, 84, 104502. [Google Scholar] [CrossRef]
  115. Zhang, J.L.; Jiao, L.; Balakirev, F.F.; Wang, X.C.; Jin, C.Q.; Yuan, H.Q. Upper critical field and its anisotropy in LiFeAs. Phys. Rev. B 2011, 83, 174506. [Google Scholar] [CrossRef]
  116. Kittel, C. Introduction to Solid State Physics, 7th ed.; John Wiley and Sons, Inc.: New York, NY, USA, 1957. [Google Scholar]
  117. Mugiraneza, S.; Hallas, A.M. Tutorial: A beginner’s guide to interpreting magnetic susceptibility data with the Curie-Weiss law. Commun. Phys. 2022, 5, 95. [Google Scholar] [CrossRef]
  118. Werthamer, N.R.; Helfand, E.; Hohenberg, P.C. Temperature and purity dependence of the superconducting critical field, Hc2. III. Electron spin and spin–orbit effects. Phys. Rev. 1966, 147, 295–302. [Google Scholar] [CrossRef]
  119. Gurevich, A. Enhancement of the upper critical field by nonmagnetic impurities in dirty two-gap superconductors. Phys. Rev. B 2003, 67, 184515. [Google Scholar] [CrossRef]
  120. Hänisch, J.; Iida, K.; Kurth, F.; Reich, E.; Tarantini, C.; Jaroszynski, J.; Förster, T.; Fuchs, G.; Hühne, R.; Grinenko, V.; et al. High field superconducting properties of Ba(Fe1−xCox)2As2 thin films. Sci. Rep. 2015, 5, 17363. [Google Scholar] [CrossRef]
  121. Pan, Y.; Sun, Y.; Zhou, N.; Yi, X.; Wang, J.; Zhu, Z.; Mitamura, H.; Tokunaga, M.; Shi, Z. Novel anisotropy of upper critical fields in Fe1+yTe0.6Se0.4. arXiv 2023, arXiv:2305.04515. [Google Scholar] [CrossRef]
  122. Stewart, G.R. Superconductivity in iron compounds. Rev. Mod. Phys. 2011, 83, 1589–1652. [Google Scholar] [CrossRef]
  123. Caspary, R.; Hellmann, P.; Keller, M.; Sparn, G.; Wassilew, C.; Köhler, R.; Geibel, C.; Schank, C.; Steglich, F.; Phillips, N.E. Unusual ground-state properties of UPd2Al3: Implications for the coexistence of heavy-fermion superconductivity and local-moment antiferromagnetism. Phys. Rev. Lett. 1993, 71, 2146–2149. [Google Scholar] [CrossRef]
  124. Tallon, J.L.; Bernhard, C.; Niedermayer, C. Muon spin relaxation studies of superconducting cuprates. Supercond. Sci. Technol. 1997, 10, A38. [Google Scholar] [CrossRef]
  125. Prozorov, R.; Tanatar, M.A.; Gordon, R.T.; Martin, C.; Kim, H.; Kogan, V.G.; Ni, N.; Tillman, M.E.; Bud’ko, S.L.; Canfield, P.C. Anisotropic London penetration depth and superfluid density in single crystals of iron-based pnictide superconductors. Phys. C 2009, 469, 582–589. [Google Scholar] [CrossRef]
  126. Kogan, V.G.; Prozorov, R.; Koshelev, A.E. Temperature-dependent anisotropies of upper critical field and London penetration depth. Phys. Rev. B 2009, 100, 014518. [Google Scholar] [CrossRef]
  127. Torsello, D.; Piatti, E.; Ummarino, G.A.; Yi, X.; Xing, X.; Shi, Z.; Ghigo, G.; Daghero, D. Nodal multigap superconductivity in the anisotropic iron-based compound RbCa2Fe4As4F2. Npj Quantum Mater. 2022, 7, 10. [Google Scholar] [CrossRef]
Figure 1. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the La0.8Ca0.2NiO2 film with physical thickness d s c = 15   n m . The raw data were reported by Chow et al. [36]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 8.0 ± 0.1   n m , T c = 7.35 ± 0.03   K , ξ a b , G L ( 0 ) = 3.99 ± 0.02   n m . The goodness of fit is (a) 0.9975 and (b) 0.9907. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) data are: T c = 7.41 ± 0.05   K , ξ a b ( 0 ) = 4.01 ± 0.03   n m , γ ξ ( 0 ) = 1.52 ± 0.04 , a = 1.26 ± 0.03 . The goodness of fit is (c) 0.9963 and (d) 0.9832. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Figure 1. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the La0.8Ca0.2NiO2 film with physical thickness d s c = 15   n m . The raw data were reported by Chow et al. [36]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 8.0 ± 0.1   n m , T c = 7.35 ± 0.03   K , ξ a b , G L ( 0 ) = 3.99 ± 0.02   n m . The goodness of fit is (a) 0.9975 and (b) 0.9907. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) data are: T c = 7.41 ± 0.05   K , ξ a b ( 0 ) = 4.01 ± 0.03   n m , γ ξ ( 0 ) = 1.52 ± 0.04 , a = 1.26 ± 0.03 . The goodness of fit is (c) 0.9963 and (d) 0.9832. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Materials 16 04367 g001
Figure 2. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the La0.8Sr0.2NiO2 film with physical thickness d s c ~ 7   n m . The raw data were reported by Wang et al. [103]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 9.5 ± 0.1   n m , T c = 6.45 ± 0.01   K , ξ a b , G L ( 0 ) = 4.11 ± 0.03   n m . The goodness of fit is (a) 0.9964 and (b) 0.9837. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 6.61 ± 0.03   K , ξ a b ( 0 ) = 4.13 ± 0.03   n m , γ ξ ( 0 ) = 1.51 ± 0.03 , a = 1.54 ± 0.05 . The goodness of fit is (c) 0.9959 and (d) 0.9897. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Figure 2. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the La0.8Sr0.2NiO2 film with physical thickness d s c ~ 7   n m . The raw data were reported by Wang et al. [103]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 9.5 ± 0.1   n m , T c = 6.45 ± 0.01   K , ξ a b , G L ( 0 ) = 4.11 ± 0.03   n m . The goodness of fit is (a) 0.9964 and (b) 0.9837. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 6.61 ± 0.03   K , ξ a b ( 0 ) = 4.13 ± 0.03   n m , γ ξ ( 0 ) = 1.51 ± 0.03 , a = 1.54 ± 0.05 . The goodness of fit is (c) 0.9959 and (d) 0.9897. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Materials 16 04367 g002
Figure 3. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the Pr0.8Sr0.2NiO2 film with physical thickness d s c ~ 7   n m . The raw data were reported by Wang et al. [103]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 9.1 ± 0.1   n m , T c = 12.11 ± 0.01   K , ξ a b , G L ( 0 ) = 3.11 ± 0.03   n m . The goodness of fit is (a) 0.9982 and (b) 0.9976. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 12.15 ± 0.04   K , ξ a b ( 0 ) = 3.12 ± 0.01   n m , γ ξ ( 0 ) = 1.06 ± 0.02 , a = 1.31 ± 0.03 . The goodness of fit is (c) 0.9984 and (d) 0.9985. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Figure 3. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the Pr0.8Sr0.2NiO2 film with physical thickness d s c ~ 7   n m . The raw data were reported by Wang et al. [103]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 9.1 ± 0.1   n m , T c = 12.11 ± 0.01   K , ξ a b , G L ( 0 ) = 3.11 ± 0.03   n m . The goodness of fit is (a) 0.9982 and (b) 0.9976. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 12.15 ± 0.04   K , ξ a b ( 0 ) = 3.12 ± 0.01   n m , γ ξ ( 0 ) = 1.06 ± 0.02 , a = 1.31 ± 0.03 . The goodness of fit is (c) 0.9984 and (d) 0.9985. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Materials 16 04367 g003
Figure 4. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the Nd0.825Sr0.175NiO2 film with physical thickness d s c ~ 7   n m . The raw data were reported by Wang et al. [103]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 12.4 ± 0.2   n m , T c = 13.52 ± 0.04   K , ξ a b , G L ( 0 ) = 3.96 ± 0.04   n m . The goodness of fit is (a) 0.9982 and (b) 0.9976. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 14.2 ± 0.1   K , ξ a b ( 0 ) = 4.00 ± 0.03   n m , γ ξ ( 0 ) = 1.16 ± 0.03 , a = 2.3 ± 0.2 . The goodness of fit is (c) 0.9846 and (d) 0.9900. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Figure 4. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the Nd0.825Sr0.175NiO2 film with physical thickness d s c ~ 7   n m . The raw data were reported by Wang et al. [103]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 12.4 ± 0.2   n m , T c = 13.52 ± 0.04   K , ξ a b , G L ( 0 ) = 3.96 ± 0.04   n m . The goodness of fit is (a) 0.9982 and (b) 0.9976. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 14.2 ± 0.1   K , ξ a b ( 0 ) = 4.00 ± 0.03   n m , γ ξ ( 0 ) = 1.16 ± 0.03 , a = 2.3 ± 0.2 . The goodness of fit is (c) 0.9846 and (d) 0.9900. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Materials 16 04367 g004
Figure 5. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the La0.8Sr0.2NiO2 film with physical thickness d s c = 6.8   n m . The raw data were reported by Wei et al. [70]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 7.49 ± 0.07   n m , T c = 17.62 ± 0.02   K , ξ a b ( 0 ) = 2.94 ± 0.01   n m . The goodness of fit is (a) 0.9978 and (b) 0.9962. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 17.83 ± 0.05   K , ξ a b ( 0 ) = 2.96 ± 0.01   n m , γ ξ ( 0 ) = 1.26 ± 0.02 , a = 1.31 ± 0.02 . The goodness of fit is (c) 0.9965 and (d) 0.9976. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Figure 5. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for the La0.8Sr0.2NiO2 film with physical thickness d s c = 6.8   n m . The raw data were reported by Wei et al. [70]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 7.49 ± 0.07   n m , T c = 17.62 ± 0.02   K , ξ a b ( 0 ) = 2.94 ± 0.01   n m . The goodness of fit is (a) 0.9978 and (b) 0.9962. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 17.83 ± 0.05   K , ξ a b ( 0 ) = 2.96 ± 0.01   n m , γ ξ ( 0 ) = 1.26 ± 0.02 , a = 1.31 ± 0.02 . The goodness of fit is (c) 0.9965 and (d) 0.9976. (e) deduced ξ a b ( T ) and γ ξ ( T ) . The 95% confidence bands are indicated by pink-shaded areas.
Materials 16 04367 g005
Figure 6. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for bulk single crystals of Tl0.58Rb0.42Fe1.72Se2. The raw data reported by Jiao et al. [106]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 5.57 ± 0.06   n m , T c = 32.26 ± 0.01   K , ξ a b ( 0 ) = 2.38 ± 0.01   n m . The goodness of fit is (a) 0.9978 and (b) 0.9962. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 32.8 ± 0.1   K , ξ a b ( 0 ) = 2.40 ± 0.01   n m , γ ξ ( 0 ) = 1.24 ± 0.04 , a = 1.29 ± 0.03 . The goodness of fit is (c) 0.9984 and (d) 0.9994. The 95% confidence bands are indicated by pink-shaded areas.
Figure 6. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for bulk single crystals of Tl0.58Rb0.42Fe1.72Se2. The raw data reported by Jiao et al. [106]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 5.57 ± 0.06   n m , T c = 32.26 ± 0.01   K , ξ a b ( 0 ) = 2.38 ± 0.01   n m . The goodness of fit is (a) 0.9978 and (b) 0.9962. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 32.8 ± 0.1   K , ξ a b ( 0 ) = 2.40 ± 0.01   n m , γ ξ ( 0 ) = 1.24 ± 0.04 , a = 1.29 ± 0.03 . The goodness of fit is (c) 0.9984 and (d) 0.9994. The 95% confidence bands are indicated by pink-shaded areas.
Materials 16 04367 g006
Figure 7. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for bulk single crystals of Fe1.11Te0.6Te0.4. Raw data reported by Fang et al. [110]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 9.65 ± 0.01   n m , T c = 13.70 ± 0.01   K , ξ a b ( 0 ) = 2.65 ± 0.01   n m . The goodness of fit is (a) 0.9986 and (b) 0.9966. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 13.74 ± 0.04   K , ξ a b ( 0 ) = 2.60 ± 0.01   n m , γ ξ ( 0 ) = 0.88 ± 0.01 , a = 1.40 ± 0.03 . The goodness of fit is (a) 0.9988 and (b) 0.9989. The 95% confidence bands are indicated by pink-shaded areas.
Figure 7. Global data fits to the 2D-GL model (Equations (1) and (2); panels (a,b)) and to our model (Equations (7)–(9); panels (ce)) for bulk single crystals of Fe1.11Te0.6Te0.4. Raw data reported by Fang et al. [110]. Deduced parameters for (a) B c 2 , p e r p ( T ) and (b) B c 2 , p a r a ( T ) are: d s c , G L = 9.65 ± 0.01   n m , T c = 13.70 ± 0.01   K , ξ a b ( 0 ) = 2.65 ± 0.01   n m . The goodness of fit is (a) 0.9986 and (b) 0.9966. Deduced parameters for (c) B c 2 , p e r p ( T ) and (d) B c 2 , p a r a ( T ) are: T c = 13.74 ± 0.04   K , ξ a b ( 0 ) = 2.60 ± 0.01   n m , γ ξ ( 0 ) = 0.88 ± 0.01 , a = 1.40 ± 0.03 . The goodness of fit is (a) 0.9988 and (b) 0.9989. The 95% confidence bands are indicated by pink-shaded areas.
Materials 16 04367 g007
Figure 8. γ ξ ( T ) data and data fit to Equation (9) for single-crystal pnictides (a) KFe2As2 and (b,c) LiFeAs. Raw data reported by (a) Zocco et al. [113], (b) Khim et al. [114], and (c) Zhang et al. [115]. Deduced parameters are (a): T c = 3.4   ( f i x e d ) , γ ξ ( 0 ) = 3.15 ± 0.03 , a = 1.62 ± 0.02 . The goodness of fit is 0.9963. (b) T c = 17.4   ( f i x e d ) [114], γ ξ ( 0 ) = 1.22 ± 0.03 , a = 1.62 ± 0.05 . The goodness of fit is 0.9777. (c) T c = 17.5   ( f i x e d ) [115], γ ξ ( 0 ) = 1.39 ± 0.02 , a = 1.89 ± 0.06 . The goodness of fit is 0.9871. The 95% confidence bands are indicated by pink-shaded areas.
Figure 8. γ ξ ( T ) data and data fit to Equation (9) for single-crystal pnictides (a) KFe2As2 and (b,c) LiFeAs. Raw data reported by (a) Zocco et al. [113], (b) Khim et al. [114], and (c) Zhang et al. [115]. Deduced parameters are (a): T c = 3.4   ( f i x e d ) , γ ξ ( 0 ) = 3.15 ± 0.03 , a = 1.62 ± 0.02 . The goodness of fit is 0.9963. (b) T c = 17.4   ( f i x e d ) [114], γ ξ ( 0 ) = 1.22 ± 0.03 , a = 1.62 ± 0.05 . The goodness of fit is 0.9777. (c) T c = 17.5   ( f i x e d ) [115], γ ξ ( 0 ) = 1.39 ± 0.02 , a = 1.89 ± 0.06 . The goodness of fit is 0.9871. The 95% confidence bands are indicated by pink-shaded areas.
Materials 16 04367 g008
Figure 9. Schematic representation of the DC magnetic susceptibility χ ( T ) obeys the Curie–Weiss law for (a) ferromagnetic, (b) paramagnetic, and (c) antiferromagnetic materials. θ is Curie–Weiss temperature (the analogue value in our model is T γ a T c ), C is Curie constant (the analogue value in our model is ( a T c ) × γ ξ ( 0 ) ), T N is Neel temperature (the analogue value in our model is T c ).
Figure 9. Schematic representation of the DC magnetic susceptibility χ ( T ) obeys the Curie–Weiss law for (a) ferromagnetic, (b) paramagnetic, and (c) antiferromagnetic materials. θ is Curie–Weiss temperature (the analogue value in our model is T γ a T c ), C is Curie constant (the analogue value in our model is ( a T c ) × γ ξ ( 0 ) ), T N is Neel temperature (the analogue value in our model is T c ).
Materials 16 04367 g009
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. Intrinsic Coherence Length Anisotropy in Nickelates and Some Iron-Based Superconductors. Materials 2023, 16, 4367. https://doi.org/10.3390/ma16124367

AMA Style

Talantsev EF. Intrinsic Coherence Length Anisotropy in Nickelates and Some Iron-Based Superconductors. Materials. 2023; 16(12):4367. https://doi.org/10.3390/ma16124367

Chicago/Turabian Style

Talantsev, Evgeny F. 2023. "Intrinsic Coherence Length Anisotropy in Nickelates and Some Iron-Based Superconductors" Materials 16, no. 12: 4367. https://doi.org/10.3390/ma16124367

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