Skip to main content
National Science Review logoLink to National Science Review
. 2019 Sep 2;7(1):21–26. doi: 10.1093/nsr/nwz129

Superconductivity in chromium nitrides Pr3Cr10-xN11 with strong electron correlations

Wei Wu 1,2, Kai Liu 3, Yanjie Li 1,2, Zhenhai Yu 4, Desheng Wu 1,2, Yuting Shao 1,2, Shihang Na 1,2, Gang Li 1,5, Ruizhen Huang 2,6, Tao Xiang 1,2,6, Jianlin Luo 1,2,5,
PMCID: PMC8288919  PMID: 34692013

Abstract

Exploration of superconductivity in Cr-based compounds has attracted considerable interest because only a few Cr-based superconductors (CrAs, A2Cr3As3 and ACr3As3 (A = K, Rb, Cs, Na)) have been discovered so far and they show an unconventional pairing mechanism. We report the discovery of bulk superconductivity at 5.25 K in chromium nitride in Pr3Cr10-xN11 with a cubic lattice structure. A relatively large upper critical field of Hc2(0) ∼ 12.6 T is determined, which is larger than the estimated Pauli-paramagnetic pair-breaking magnetic field. The material has a large electronic specific-heat coefficient of 170 mJ K−2 mol−1—about 10 times larger than that estimated by the electronic structure calculation, which suggests that correlations between 3d electrons are very strong in Pr3Cr10-xN11, and thus quantum fluctuations might be involved. Electronic structure calculations show that the density of states at the Fermi energy are contributed predominantly by Cr 3d electrons, implying that the superconductivity results mainly from the condensation of Cr 3d electrons. Pr3Cr10-xN11 represents a rare example of possible unconventional superconductivity emerging in a 3D system with strong electron correlations. Nevertheless, clarification of the specific pairing symmetry needs more investigation.

Keywords: Cr-based superconductors, nitrides, strong electron correlations, unconventional superconductivity

INTRODUCTION

The 3d transition-metal oxides or pnictides exhibit rich quantum phases with novel quantum states, such as long-range magnetic orders, charge or spin density waves, metal-insulator transitions, high-Tc superconductivity and colossal magnetoresistance. In particular, unconventional high-Tc superconductivity has been discovered in cuprates as well as in iron-based superconductors [1,2]. Many compounds with 3d transition-metal elements can become superconducting at low temperatures. However, it is relatively difficult to find a superconducting material in chromium-based compounds because most Cr-based compounds have strong magnetism, which generally is not in favor of superconductivity. In fact, CrAs, A2Cr3As3 and ACr3As3 (A = Na, K, Rb, Cs) are the only Cr-based superconductors so far [3–6].

Superconductivity in CrAs was discovered in 2014 [3,4]. CrAs undergoes a first-order antiferromagnetic transition with a double helical spin structure at TN ≈ 265 K [3,7,8]. A bulk superconductivity with Tc ≈ 2 K emerges above a pressure Pc ≈ 8 kbar, where the antiferromagnetic order is completely suppressed. Both the NMR and neutron-scattering measurements for CrAs under high pressure revealed that there are strong magnetic fluctuations and line nodes may exist in the superconducting gap function [9–13], as in Sr2RuO4 or some heavy fermion superconductors [14,15]. After the discovery of superconductivity in CrAs, superconductivity has also been found in the quasi-1D compounds A2Cr3As3 (A = K, Rb, Cs) [3,4]. The upper critical field Hc2 of K2Cr3As3 is about three times larger than the Pauli-paramagnetic limit.

In this paper, we report a novel Cr-based nitride superconductor with a cubic structure, Pr3Cr10-xN11, which was first synthesized and characterized by Broil et al. in 1995 [16]. The compound crystallizes in space group Fm-3m (No. 225) with lattice constant a = 12.891 Å. It contains 192 atoms in a face-centered cubic (FCC) cell with three kinds of building blocks, as illustrated in Fig. 1a. The building blocks are Pr6N9, Cr8N13 and Cr6. Previous studies showed that there are Cr vacancies in the lattice structure. However, no low-temperature physical properties were reported [16,17]. Our key finding of this study is the observation of superconductivity in Pr3Cr10-xN11 with Tc ∼ 5.25 K. The high-quality samples have a shielding fraction of 85% at 2 K from zero-field-cooled (ZFC) magnetic susceptibility and a prominent superconducting peak in the specific-heat measurement. A relatively large upper critical field is found at the zero-temperature limit, Hc2(0) ∼ 12.6 T, which is larger than the Pauli-paramagnetic pair-breaking field. From electronic structure calculations, we find that the density of states (DOS) at the Fermi energy are predominately contributed by Cr 3d electrons. The present results demonstrate that Pr3Cr10-xN11 is the first Cr-based superconductor discovered in chromium nitrides and it represents a rare example that possibly unconventional superconductivity emerges in a 3D system with strong electron correlations.

Figure 1.

Figure 1.

(a) The face-centered cubic cell of Pr3Cr10-xN11 with three kinds of building blocks: Pr6N9, Cr8N13 and Cr6, respectively. The blocks are shown in a shrunken form. There are two Cr positions: Cr1 and Cr2. (b) Typical Rietveld refinement of Pr3Cr10-xN11 under ambient conditions. The vertical bars represent the calculated Bragg reflection positions of the diffraction peaks for Pr3Cr10-xN11. The difference between the observed (scatters) and the fitted patterns (line) is shown at the bottom of the diffraction peaks.

RESULTS AND DISCUSSION

Sample and characterization

Polycrystalline samples of Pr3Cr10-xN11 were prepared by direct reactions of the corresponding binary nitrides with solid reaction. Synthesized powder shows a dark-brown color and is air-sensitive, as it is easily oxidized to Pr2O3 within a few hours. Figure 1b shows the General Structure Analysis System (GSAS) refinement of Pr3Cr10-xN11 under ambient conditions, which indicates that Pr3Cr10-xN11 crystallized in an FCC structure with space group Fm-3m. All the Pr3Cr10-xN11 reflections can be well indexed based on a cubic cell with lattice parameter a = 12.8521 Å, which is consistent with those reported in the literature (a = 12.891 Å) [16], indicating their ideal composition. A small amount of raw reactant Cr2N was marked with a red vertical bar.

Temperature-dependent resistivity in Pr3Cr10-xN11

Figure 2a shows the temperature-dependent resistivity for Pr3Cr10-xN11 from 1.8 to 300 K at zero field. The normal-state resistivity is metallic, with no phase transition observed. At low temperatures, a sharp superconducting transition is observed with onset Tc of about 5.25 K, as shown in the inset of Fig. 2a. The Tc of Pr3Cr10-xN11 is higher than that of CrAs with Tmax ∼ 2 K under pressure and is close to K2Cr3As3 with Tc ∼ 6.1 K [12].

Figure 2.

Figure 2.

(a) Temperature dependence of the resistivity for Pr3Cr10-xN11 from 1.8 to 300 K at the zero field. The inset shows the resistivity near superconducting transition. (b) The temperature dependence of the upper critical magnetic field. The solid circle shows the upper critical field obtained from WHH fitting. The inset shows the temperature dependence of the resistivity at different fields.

The new superconductor Pr3Cr10-xN11 shows a relatively large upper critical field. Figure 2b shows resistivity data in magnetic fields up to 9 T. As the field increases, the transition temperature Tc shifts to a lower temperature and the transition width is gradually broadened, similar to the iron-based superconductors [18,19]. The upper critical field Hc2 (T) curve obtained from the field-dependent transition temperatures shows a remarkably high critical field of 12.6 T using the formula Hc2(T) = Hc2(0)(1 − t2), where t is the reduced temperature t = T/Tc and of 22 T using the Werthamer−Helfand−Hohenberg (WHH) theory [20]. On the other hand, the Pauli-paramagnetic limit for the upper critical field is HP = 1.84Tc ≈ 9.6 T in the case of an isotropic full superconducting gap without considering spin–orbit coupling [21,22]. The Hc2 (0) in Pr3Cr10-xN11 is 130% as large as HP. Usually, the high superconducting upper critical field can be originated from multi-band effects, the strong-coupling effect, the spin-triplet pairing and the strong spin–orbit coupling effect in a low-dimensional system [23–25]. The origin of large Hc2(0) in Pr3Cr10-xN11 needs to be further studied. The obtained Inline graphic allows us to estimate the Ginzburg–Landau coherence length Inline graphic= 51 Å according to the relationship: Inline graphic, where Inline graphic =2.067 × 10−15 Wb is the magnetic flux quantum.

Magnetic-susceptibility measurements in Pr3Cr10-xN11

The bulk superconductivity in Pr3Cr10-xN11 was confirmed by magnetic-susceptibility measurements. Figure 3a shows susceptibility χ at low temperatures with ZFC and field cool (FC) under a magnetic field of 30 Oe. χ starts to drop below Tc and the diamagnetic signal tends to saturate at low temperatures. The shielding fraction estimated from the ZFC magnetic susceptibility at 2 K is 85%, confirming bulk superconductivity in the sample. The normal-state susceptibility χ increases with decreasing temperature in Pr3Cr10-xN11, showing a Curie–Weiss behavior as shown in the inset of Fig. 3a. Such behavior is different from that of isostructural material La3Cr10-xN11, which shows a Pauli paramagnetism with a nearly temperature-independent susceptibility [16]. Since La3+ ion has no occupied 4f electrons while each Pr3+ ion has two occupied 4f electrons, it is natural to attribute the Curie–Weiss behavior of χ(T) in Pr3Cr10-xN11 to the magnetic moments of Pr3+ 4f electrons. Using a Curie–Weiss fit with formula χ(T) = χ0 + C/(T-θ), we obtained the effective moment of each Pr ion at about 3.6 μB, which is very close to the calculated moment of 3.5 μB for Pr3+ by Hund’s rule. The negative value of θ indicates the correlations between Pr ions are antiferromagnetic.

Figure 3.

Figure 3.

(a) Temperature dependence of the dc magnetic susceptibility with zero field cool (ZFC) and field cool (FC) measured in a field of 30 Oe. The inset of Fig. 3a is the normal-state susceptibility measured in a field of 2 T. (b) The magnetic hysteresis of the sample measured at 2 K. The left inset shows the lower critical magnetic field.

Further confirmation of superconductivity is shown by the magnetic hysteresis of the sample measured at 2 K in Fig. 3b, which displays the typical magnetic hysteresis curve for a type-II superconductor. From the inset of Fig. 3b, the lower critical magnetic fields Hc1 of 170 Oe can be obtained.

Specific-heat measurements

Figure 4a shows the specific-heat coefficient C/T as a function of T2 from 2 to 10 K at the zero field. The bulk nature of superconductivity is confirmed by a pronounced anomaly around Tc = 5.25 K, consistently with the resistivity and susceptibility measurements. Extrapolating C/T to zero temperature gives a residual value of γ0 = 0.061 mJ g−1 K−2. As indicated previously, there is raw reactant phase Cr2N in the sample. We measured the specific heat of Cr2N and found that it can be well fitted by C = γT + βT3 below 10 K with γ = 22 mJ mol−1 K2 and β = 0.0373 mJ mol−1 K4. Given the residual specific-heat γ0 origin from the Cr2N phase, we can then obtain the specific heat of pure Pr3Cr10-xN11 by subtracting that of Cr2N from the total specific heat, as shown in the inset of Fig. 4a.

Figure 4.

Figure 4.

(a) The specific-heat coefficient C/T of Pr3Cr10-xN11 as a function of T2. The inset shows C/T versus T2 for a sample with Cr2N. (b) Temperature dependence of normalized electronic specific-heat Ce/ T.

Above Tc, the good linear T2 dependence of C/T indicates that the normal-state specific heat consists of two parts of contributions: the electronic part, which is proportional to T, and the phonon part, which is proportional to T3 at low temperatures. By fitting the normal-state specific heat C with the formula C = γnT + βT3, we find that γn = 173 mJ K−2 mol−1 and β = 1.186 mJ K−4 mol−1. The normal-state electronic coefficient γn is proportional to the DOS at the Fermi level. Assuming that most of the DOS comes from the Cr 3d electrons, the value of γn per mole Cr is γn = 17.3 mJ K−2 mol Cr−1 for Pr3Cr10-xN11, which is much larger than the corresponding value for CrAs (∼7 mJ K−2 mol−1) and MnP (∼8.3 mJ K−2 mol−1) at ambient pressure [11,26] and it is slightly less than that for K2Cr3As3 (∼23.3 mJ K−2 mol Cr−1) and KCr3As3(∼27.1 mJ K−2 mol Cr−1) [14,27,28]. The relatively large γn for Pr3Cr10-xN11 indicates strong correlations of Cr 3d electrons. The Debye temperature θD obtained from β is 339 K.

In Fig. 4b, the normalized specific-heat jump at Tc is found to be ΔCnTc = 1.51. This value is much smaller than those of K2Cr3As3 (∼2.5) and LaNiAsO (∼1.9), which are regarded as strong-coupling superconductivity [14,27,29]. The normalized ΔC/γnTc reflects the coupling strength between the conducting electrons and the pairing glue. We can then estimate the electron–phonon coupling constant λ = 0.6 from the modified McMillian formula [25,30,31]:

graphic file with name M5.gif

where μ is a Coulomb pseudopotential and ω is a logarithmic averaged phonon frequency. ω can be determined from the specific-heat jump at Tc using the formula: ΔCnTc = 1.43[1 + 53(Tc/ω)2ln(ω/3Tc)]. Taking μ = 0.10 and Tc = 5.25 K, we obtained ω = 320 K and λ = 0.6. For such a small electron–phonon coupling constant, the large Hc2(0) value is not likely due to the strong-coupling effect.

Theoretical calculations

In order to examine which atomic species contribute most around the Fermi level (EF), we have plotted the local density of states (LDOS) for Pr3Cr9.5N11 as shown in Fig. 5. Electronic structure calculations show that the DOS at the Fermi energy are contributed predominantly by Cr 3d electrons, implying that the superconductivity results mainly from the condensation of Cr 3d electrons similar to that in CrAs [32]. The primitive cell of Pr3Cr10-xN11 is shown in the inset of Fig. 5, in which there are two types of nonequivalent Cr atoms, labeled as Cr1 and Cr2, respectively. Both the Cr1 and Cr2 atoms have large contributions around EF, mainly originating from the 3d orbitals of Cr atoms. In contrast, the vast majority of states for Pr and N atoms are far from the Fermi level. A higher density of Cr2 vacancies, such as in Pr3Cr9N11, does not change the results very much. According to our calculations, the total DOS of Pr3Cr9.5N11 at the Fermi level N(EF) is 7.37 states/(eVInline graphicf.u.). As a result, the corresponding calculated electronic specific-heat coefficient γe is 17.4 mJ K−2 mol−1. Experimentally, the measured γe for Pr3Cr10-xN11 is about 173 mJ K−2 mol−1—about 10 times more than the band calculations, which indicates the strong mass-enhancement effect.

Figure 5.

Figure 5.

Local density of states (LDOS) for Pr, Cr1, Cr2 and N atoms in Pr3Cr9.5N11. Inset shows a primitive cell with the yellow, blue, maroon and gray balls representing Pr, Cr1, Cr2 and N atoms, respectively. The atomic vacancy at the Cr2 site is highlighted by a red dashed circle.

CONCLUSION

Superconductivity in Pr3Cr10-xN11 shows several novel characters. First, electronic structure calculations show that most of DOS at Fermi energy is contributed by Cr 3d electrons, suggesting that the superconductivity is originated from the condensation of Cr 3d electrons. So Pr3Cr10-xN11 is the first Cr-based superconductor discovered in nitrides. A few other known Cr-based superconductors (CrAs, A2Cr3As3 and ACr3As3 (A = K, Rb, Cs, Na)) are arsenide. Superconductivity in CrAs emerges in the vicinity of a quantum critical point and antiferromagnetic spin fluctuations associated with the quantum criticality could act as an important glue medium for Cooper pairing. Superconductivity in the quasi-1D compounds A2Cr3As3 (A = K, Rb, Cs) shows non-s-wave pairing. Both of them show an unconventional pairing mechanism. So the Cr d-electrons play an important role in electron correlations and possibly unconventional superconductivity in Pr3Cr10-xN11 with Cr d-electrons.

Second, Pr3Cr10-xN11 has a relatively large upper critical field Hc2(0) ∼ 12.6 T, exceeding the Pauli limit of the paramagnetic pair-breaking field; this is rare in 3D-structure superconductors. The upper critical field provides very important information on the superconducting pairing. This behavior resembles that unconventional superconducting in K2Cr3As3 in which Hc2 is three times the Pauli-paramagnetic limit that is regarded as evidence of spin-triplet superconductivity.

Last but not least, the measured γe for Pr3Cr10-xN11 is about 173 mJ mol−1 K−2, equivalent to 17.3 mJ K−2 mol Cr−1. This γ value is close to that of K2Cr3As3 (mJ K−2 mol Cr−1), indicating enhanced electron correlations in Pr3Cr10-xN11. The experimental value of γe is about 10 times higher than the electronic calculations. This large renormalization factor cannot be explained by electron–phonon interactions and thus quantum fluctuations might be involved.

In conclusion, we report the experimental result for a novel Cr-based superconductor nitrides Pr3Cr10-xN11 with a cubic lattice structure. Bulk superconductivity with Tc ∼ 5.25 K is observed from the resistivity, susceptibility and specific-heat measurements. Further theoretical and experimental studies are needed to determine the pairing symmetry and the corresponding mechanism, especially the role of Cr 3d electrons, for the observed superconductivity in Pr3Cr10-xN11.

METHODS

Sample preparation

Polycrystalline samples of Pr3Cr10-xN11 were prepared by direct reactions of the corresponding binary nitrides, staring with PrN (99%) and a mixture of the chromium nitrides (CrN and Cr2N) in a mass ratio of 4:6. The operations were all performed in an Ar-filled glovebox. Cold-pressed pellets of the mixtures were sealed in an evacuated quartz tube (<10−4 Pa). The pellets were gradually heated in 1 day to 1000°C, held at that temperature for 50 hours, then cooled to the room temperature in the furnace. The products were again ground in a glovebox, pressed into pellets, wrapped in Ta foil and heated in an evacuated quartz tube at 1165°C for 120 hours. It is worth noting that, after this treatment, the samples were usually still contaminated by the binary nitrides. The treatment of these samples with diluted hydrochloric acid only dissolved the rare-earth nitrides. In between these treatments, the pellets were ground to a fine powder, the decomposed rare-earth nitride was dissolved in hydrochloric acid and fresh rare -earth nitride was added again to the mixture to maintain the proper composition. The reaction temperature was carefully selected to avoid decomposition of the products, meanwhile obtaining a good crystallization. In our study, we found that Pr3Cr10-xN11 partially decomposed above 1200°C. Synthesized powder shows a dark-brown color and is air-sensitive, as it is easily oxidized to Pr2O3 within a few hours.

Measurements

The electrical transport measurement was carried out on a physical-property measurement system (PPMS-9, Quantum Design). The resistivity was measured by a standard four-probe method, employing silver-paste contacts cured at room temperature, used for resistivity measurements, with the electric current applied in an arbitrary direction. The magnetic susceptibility was measured in a Quantum Design SQUID VSM. The specific-heat measurements were performed up to 9 T in a PPMS.

Theoretical modeling

We carried out first-principles electronic structure calculations on Pr3Cr10-xN11. The first-principles calculations were performed using the projector augmented-wave method [33], as implemented in the VASP package [34]. The generalized gradient approximation of the Perdew–Burke–Ernzerhof type [35] was adopted for the exchange-correlation function.

Supplementary Material

nwz129_Supplemental_File

Acknowledgements

The authors are grateful to Dr. Zhao Yu Liu and Prof. Shi Liang Li (Institute of Physics) for their help in experiments.

FUNDING

This work was supported by the National Basic Research Program of China (2017YFA0302901, 2014CB921500), the National Natural Science Foundation of China (11674375, 11634015) and the Strategic Priority Research Program and Key Research Program of Frontier Sciences of the Chinese Academy of Sciences (XDB07020200).

Conflict of interest statement. None declared.

REFERENCES

  • 1. Bednorz JG, Müller KZ. Possible high Tc superconductivity in the Ba−La−Cu−O system. Z Phys B-Condens Mat 1986; 64: 189–93. [Google Scholar]
  • 2. Kamihara Y, Watanabe T, Hirano Met al. Iron-based layered superconductor La [O1-xFx] FeAs (x = 0.05−0.12) with Tc = 26 K. J Am Chem Soc 2008; 130: 3296–7. [DOI] [PubMed] [Google Scholar]
  • 3. Wu W, Cheng JG, Matsubayashi Ket al. Superconductivity in the vicinity of antiferromagnetic order in CrAs. Nat Commun 2014; 5: 5508. [DOI] [PubMed] [Google Scholar]
  • 4. Kotegawa H, Nakahara S, Tou Het al. Superconductivity of 2.2 K under pressure in Helimagnet CrAs. J Physical Soc Japan 2014; 83: 238–41. [Google Scholar]
  • 5. Bao JK, Liu JY, Ma CWet al. Superconductivity in quasi-one-dimensional K2Cr3As3 with significant electron correlations. Phys Rev X 2015; 5: 5568–77. [Google Scholar]
  • 6. Tang ZT, Bao JK, Liu Yet al. Unconventional superconductivity in quasi-one-dimensional Rb2Cr3As3. Phys Rev B 2015; 91: 020506; Tang ZT, Bao JK and Wang Z et al. Superconductivity in quasi-one-dimensional Cs2Cr3As3 with large interchain distance. Sci China Mater 2015; 58: 16–20; Mu QG, Ruan BB and Pan BJ et al. Superconductivity at 5 K in quasi-one-dimensional Cr-based KCr3As3 single crystals. Phys Rev B 2017; 96: 140504; Liu T, Mu QG and Pan BJ et al. Superconductivity at 7.3 K in 133-type Cr-based RbCr3As3 single crystals. Europhys Lett 2017; 120: 27006. [Google Scholar]
  • 7. Cheng JG, Matsubayashi K, Wu Wet al. Pressure induced superconductivity on the border of magnetic order in MnP. Phys Rev Lett 2015; 114: 117001. [DOI] [PubMed] [Google Scholar]
  • 8. Cheng JG, Luo JL. Special issue on pressure-induced superconductivity in CrAs and MnP. J Phys Condens Matter 2017; 29: 440301. [DOI] [PubMed] [Google Scholar]
  • 9. Kotegawa H, Nakahara S, Akamatsu Ret al. Detection of an unconventional superconducting phase in the vicinity of the strong first-order magnetic transition in CrAs using 75As-nuclear Quadrupole resonance. Phys Rev Lett 2015; 114: 117002. [DOI] [PubMed] [Google Scholar]
  • 10. Keller L, White JS, Frontzek Met al. Pressure dependence of the magnetic order in CrAs: a neutron diffraction investigation. Phys Rev B 2015; 91: 020409. [Google Scholar]
  • 11. Shen Y, Wang Q, Hao YQet al. Structural and magnetic phase diagram of CrAs and its relationship with pressure-induced superconductivity. Phys Rev B 2016; 93: 060503. [Google Scholar]
  • 12. Matsuda M, Ye F, Dissanayake SEet al. Pressure dependence of the magnetic ground states in MnP. Phys Rev B 2016; 93: 100405. [Google Scholar]
  • 13. Matsuda M, Lin FK, Yu Ret al. Evolution of magnetic double helix and quantum criticality near a dome of superconductivity in CrAs. Phys Rev X 2018; 8: 031017. [Google Scholar]
  • 14. Mackenzie AP, Haselwimmer RKW, Tyler AWet al. Extremely strong dependence of superconductivity on disorder in Sr2RuO4. Phys Rev Lett 1998; 80: 161–4. [Google Scholar]
  • 15. Mathur ND, Grosche FM, Julian SRet al. Magnetically mediated superconductivity in heavy fermion compounds. Nature 1998; 394: 39–43. [Google Scholar]
  • 16. Broil S, Jeitschko W. The ternary rare earth chromium nitrides Ce2CrN3 and Ln3Cr10-XN11 with Ln = La, Ce, Pr. Z Naturforsch B 1995; 50b: 905–12. [Google Scholar]
  • 17. Chevire F, Ranjan C, Disalvo F. Synthesis, crystal and electronic structures of La3Cr2N6. Solid State Commun 2009; 149: 273–6. [Google Scholar]
  • 18. Chen GF, Li Z, Li Get al. Superconducting properties of the Fe-based layered superconductor LaFeAsO0.9F0.1−δ. Phys Rev Lett 2008; 101: 057007. [DOI] [PubMed] [Google Scholar]
  • 19. Chen GF, Li Z, Wu Det al. Superconductivity at 41 K and its competition with spin-density-wave instability in layered CeO1−xFxFeAs. Phys Rev Lett 2008; 100: 247002. [DOI] [PubMed] [Google Scholar]
  • 20. Werthamer NR, Helfand E. Temperature and purity dependence of the superconducting critical field, Hc2. Phys Rev Lett 1964; 13: 686–8. [Google Scholar]
  • 21. Clogston AM. Upper limit for the critical field in hard superconductors. Phys Rev Lett 1962; 9: 266–7. [Google Scholar]
  • 22. Chandrasekhar BS. A note on the maximum critical field of high-field superconductors. Appl Phys Lett 1962; 1: 7–8. [Google Scholar]
  • 23. Gurevich A. Enhancement of the upper critical field by nonmagnetic impurities in dirty two-gap superconductors. Phys Rev B 2003; 67: 184515. [Google Scholar]
  • 24. Carbotte JP. Properties of boson-exchange superconductors. Rev Mod Phys 1990; 62: 1027–158. [Google Scholar]
  • 25. Lee IJ, Chaikin PM, Naughton MJ. Exceeding the Pauli paramagnetic limit in the critical field of (TMTSF)2PF6. Phys Rev B 2000; 62: 14669–72. [Google Scholar]
  • 26. Zheng P, Xu YJ, Wu Wet al. Orbital-dependent charge dynamics in MnP revealed by optical study. Sci Rep 2017; 7: 14178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Shao YT, Wu XX, Wang Let al. Evidence of line nodes in superconducting gap function in K2Cr3As3 from specific-heat measurements. Europhys Lett 2018; 123: 57001. [Google Scholar]
  • 28. Mu QG, Ruan BB, Pan BJet al. Dynamic instabilities in strongly correlated VSe2 monolayers and bilayers. Phys Rev B 2017; 96: 140524. [Google Scholar]
  • 29. Li Z, Chen GF, Dong Jet al. Strong-coupling superconductivity in the nickel-based oxypnictide LaNiAsO1−xFx. Phys Rev B 2008; 78: 060504. [Google Scholar]
  • 30. Allen PB, Dynes RC. Transition temperature of strong-coupled superconductors reanalyzed. Phys Rev B 1975; 12: 905–22. [Google Scholar]
  • 31. Klimczuk T, Ronning F, Sidorov Vet al. Physical properties of the noncentrosymmetric superconductor Mg10Ir19B16. Phys Rev Lett 2007; 99: 257004. [DOI] [PubMed] [Google Scholar]
  • 32. Autieri C, Noce C. First principles study of structural, magnetic and electronic properties of CrAs. Philos Mag 2017; 97: 3276–95. [Google Scholar]
  • 33. Blöchl PE. Projector augmented-wave method. Phys Rev B 1994; 50: 17953–79; Kresse G and Joubert D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys Rev B 1999; 59: 1758–75. [DOI] [PubMed]
  • 34. Kresse G, Hafner J. Ab initio molecular dynamics for liquid metals. Phys Rev B 1993; 47: 558–61; Kresse G and Furthmüller J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comp Mater Sci 1996; 6: 15–50; Kresse G and Furthmüller J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys Rev B1996; 54: 11169–86. [DOI] [PubMed] [Google Scholar]
  • 35. Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made simple. Phys Rev Lett 1996; 77: 3865–8. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

nwz129_Supplemental_File

Articles from National Science Review are provided here courtesy of Oxford University Press

RESOURCES