Single crystal growth and superconductivity in RbNi2Se2
Abstract
We report the synthesis and characterization of RbNi2Se2, an analog of the iron chalcogenide superconductor RbxFe2Se2, via transport, angle resolved photoemission spectroscopy, and density functional theory calculations. A superconducting transition at = 1.20 K is identified. In normal state, RbNi2Se2 shows paramagnetic and Fermi liquid behaviors. A large Sommerfeld coefficient yields a heavy effective electron mass of . In the superconducting state, zero-field electronic specific-heat data can be described by a two-gap BCS model, indicating that RbNi2Se2 is a multi-gap superconductor. Our density functional theory calculations and angle resolved photoemission spectroscopy measurements demonstrate that RbNi2Se2 exhibits relatively weak correlations and multi-band characteristics, consistent with the multi-gap superconductivity.
I Introduction
Since the discovery of copper oxide superconductors, researchers have extensively searched for superconductivity in materials with transition metalsBednorz et al. (1987); Takada et al. (2003). Significant progress has been made in iron pnictide and chalcogenide compounds, where several structural systems have been identified with the highest of 55 K achieved in LaFeAsORen et al. (2008). Superconductivity was also observed in the compounds consisting of Cr and Mn under pressure, such as CrAs, KCrAs, and MnPWu et al. (2014); Mu et al. (2017); Cheng et al. (2015). Among all of them, superconductivity has been found to be in the vicinity of an antiferromagnetic (AF) order, suggesting that spin fluctuations may play an important role in the mechanism of superconductivity. Nickel oxide materials have analogous structures with copper oxide superconductors. Superconductivity with K has also been observed in films of nickel-based compoundsLi et al. (2019); Gu and Wen (2022).
Ni2Se2 ( K, Cs, and Tl) crystalizes in the ThCr2Si2 structure and shows metallic behavior and Pauli paramagnetism. At low temperatures, superconductivity emerges with the superconducting (SC) transition temperatures of 0.8 K for KNi2Se2 polycrystals, 2.7 K for CsNi2Se2, and 3.7 K for TlNi2Se2 single crystalsNeilson et al. (2012); Chen, Huimin and Yang, Jinhu and Cao, Chao and Li, Lin and Su, Qiping and Chen, Bin and Wang, Hangdong and Mao, Qianhui and Xu, Binjie and Du, Jianhua and Fang, Minghu (2016); Wang et al. (2013). While K0.95Ni1.86Se2 single crystals do not show superconductivity down to 0.3 K, yielding that the superconductivity is sensitive to the stoichiometry of the samplesLei et al. (2014). As a comparison, AxFe2Se2 system exhibits s ranging from 20 30 K. With different amount of iron vacancies, AxFe2-δSe2 exhibits a variety of AF orders and iron vacancy ordersGuo et al. (2010); Fang et al. (2013); Dai (2015); Wang et al. (2016). The replacement of Fe by Co suppresses the superconductivity and induces a ferromagnetic (FM) order in RbCo2Se2Yang et al. (2013); Huang et al. (2021). The Ni2Se2 superconductors with a formal valence of Ni1.5+ have been revealed to exhibit remarkable properties. In particular, they usually exhibit a large Sommerfeld coefficient , suggesting a large density of states and unconventional pairing at low temperatureWang et al. (2013); Chen, Huimin and Yang, Jinhu and Cao, Chao and Li, Lin and Su, Qiping and Chen, Bin and Wang, Hangdong and Mao, Qianhui and Xu, Binjie and Du, Jianhua and Fang, Minghu (2016). One possibility that was proposed was that the large Sommerfeld coefficient might be induced by local charge orderNeilson et al. (2012). However, angle resolved photoemission spectroscopy (ARPES) measurements yield weak electronic correlations in KNi2Se2 and the origin of the large Sommerfeld coefficient may be driven by the large density states and the Van Hove singularity in the vicinity of the Fermi energyFan et al. (2015).
In this work, we report the successful synthesis and characterization of RbNi2Se2 single crystals. The crystal structure, electronic band structure, and transport properties have been investigated. We find that RbNi2Se2 is a Pauli paramagnetism and exhibits a SC transition at = 1.20 K. Normal-state specific heat measurements suggest an effective electronic mass enhancement with . In SC state, a two-gap BCS model can match well with the zero-field electronic specific heat, indicating that RbNi2Se2 is a multi-gap superconductor. Comparison with the density functional theory (DFT) calculations and ARPES measurements reveals that RbNi2Se2 is a weakly correlated superconductor with multi bands crossing the Fermi level.
II Experimental and calculation details
Single crystals of RbNi2Se2 were grown by the self-flux method. First, the precursor NiSe was prepared by heating Ni powders and Se pellets at C. Then, NiSe powders and Rb were put into alumina crucibles according to stoichiometry and sealed in an evacuated silica tube. The mixture was kept at C for 5 h, then heated to C in 40 h and kept for 5 h, after that heated to C in 40 h and held for 5 h. Finally, the temperature was cooled down to C at a rate of C/h. To prevent the reaction of Rb with water and air, all of the processes were conducted in an argon-filled glove box. Shiny plate-like single crystals with a typical size of 4 5 1 mm3 were grown as shown in the inset of Fig. 1(b).
Formula weight | 223.14 |
---|---|
Crystal system | Tetragonal |
Space group | I4/mmm |
Unit-cell parameters | a = b = 3.9272 (3) Å |
c = 13.8650(5) Å | |
Atomic parameters | |
Rb | 2b(0,0,1/2) |
Ni | 4d(0,1/2,3/4) |
Se | 4e(1/2,1/2,0.6502(1)) |
Density | 3.466 g/cm3 |
F(000) | 198 |
Radiation | Mo ( = 0.7107Å) |
2 for data collection | to |
Index ranges | -4 h 5, -5 k 5, |
-11 l 18 | |
Reflections collected | 808 |
Independent reflections | 119 |
Data/restraints/parameters | 119/0/5 |
Goodness-of-fit on F2 | 1.204 |
Final R indexes [I 2(I)] | = 0.0378, =0.1001 |
Largest diff. peak/hole/e Å-3 | 3.06/-2.04 |
Single crystal x-ray diffraction (XRD) were conducted on a SuperNova (Rigaku) x-ray diffractometer. The sample was blowed by N2 during the data collection to avoid exposure to air. The elemental analysis was measured by using an energy-dispersive x-ray spectroscopy (EDS) (EVO, Zeiss). Electrical transport, magnetic and specific heat measurements were performed on a physical property measurement system (PPMS, Quantum Design). The in-plane resistivity was measured using the standard four-probe method on a rectangular sheet crystal to keep current flowing in the ab-plane. The Vienna Ab initio Simulation Package (VASP) was employed for the DFT calculationsKresse and Hafner (1993). ARPES measurements were performed on a helium-lamp based system with a DA30 electron analyzer. Single crystals were cleaved in-situ in ultra-high vacuum with a base pressure better than Torr at 30 K. Energy and angular resolutions were better than 20 meV and , respectively.
III Results and discussions
All peaks from single crystal XRD can be indexed with the ThCr2Si2-type structure (space group: I4/mmm), which is illustrated in the inset of Fig. 1(a). The determined lattice parameters are , and Å at 150 K with the volume of unit cell between that of KNi2Se2 and CsNi2Se2. Details of the atom coordinates and other key information are shown in Table 1. To show the quality of the samples, we present a 2 scan of a single crystal along the in Fig. 1(a) and an XRD pattern on a powder sample in Fig. 1(b), where are Miller indices in reciprocal lattice units. No peaks from impurity could be identified. The EDS results for several single crystals are rather homogenous and the determined average atomic ratios are Rb:Ni:Se = 1.16(4):2.04(3):2.00(6) when the content of Se is normalized to be 2, close to the stoichiometry of RbNi2Se2.
Temperature dependence of the resistivity for RbNi2Se2 is shown in Fig. 2(a). The electric current is applied in the ab-plane. The value of is about 92.1 cm at 300 K and only about 1.5 cm at 1.8 K. The residual resistivity ratio (RRR) of 61.4 [] suggests remarkable metallicity and high quality of the single crystalsLei et al. (2014); Ehrlich and Rivier (1968); Böhmer et al. (2016). The resistivity measured at 0, 0.15, and 11 T shows a metallic behavior without any anomaly or strong magnetic field dependence. The resistivity below 40 K can be well described by the equation TT2 as shown in the inset of Fig. 2(a), where = 1.211 cm and A = 0.012 cm/K2, revealing a paramagnetic Fermi liquid behaviorAnalytis et al. (2014). The magnetic susceptibility is nearly independent of temperature, yielding a Pauli paramagnetic behaviour as shown in Fig. 2(b). A weak FM sign is revealed from the hysteresis loop shown in the inset of Fig. 2(b), which could be ascribed to a small amount of Ni impurityLi et al. (2020). The specific heat from 2 to 200 K shown in Fig. 2(c) also suggests that no phase transition occurs in this temperature range.
To explore the possible superconductivity, we show measurements down to 50 mK in Fig. 3. The magnetic susceptibility at low temperatures is shown in Fig. 3(a). A clear diamagnetic response appears below 1.20 K under zero field, indicating a SC transition. The transition shifts to lower temperatures with an increase of the bias field, consistent with the Meissner effect of superconductivity. With knowing 0.8 K for KNi2Se2Neilson et al. (2012) and 2.7 K for CsNi2Se2Chen, Huimin and Yang, Jinhu and Cao, Chao and Li, Lin and Su, Qiping and Chen, Bin and Wang, Hangdong and Mao, Qianhui and Xu, Binjie and Du, Jianhua and Fang, Minghu (2016), the diamagnetic response at 1.20 K should correspond to the SC transition of RbNi2Se2. We plot the onset SC transition temperatures at various magnetic fields and fit the upper critical field using the Ginzburg-Landau theory with the formula , where is the reduced temperature . As shown in Fig. 3(b), the resultant = 1.38 T is within the Pauli limit, = 2.37 T, indicating a weak coupling behaviorBao et al. (2015).
stop here
Figure 3(c) displays the specific heat measured at low temperatures. In normal state, a fit to the specific heat from 1.5 to 3.8 K using = + results in the Sommerfeld coefficient = 30.30 mJmol-1K-2 and = 3.67 mJmol-1K-4, as shown in the inset. The Debye temperature is estimated to be of 167 K from the equation = , where is the atomic number in each formula unit and is the ideal gas constant. The for RbNi2Se2 is comparable with that of KNi2Se2 (44 mJmol-1K-2)Neilson et al. (2012), CsNi2Se2 (77 mJmol-1K-2)Chen, Huimin and Yang, Jinhu and Cao, Chao and Li, Lin and Su, Qiping and Chen, Bin and Wang, Hangdong and Mao, Qianhui and Xu, Binjie and Du, Jianhua and Fang, Minghu (2016), and TlNi2Se2 (40 mJmol-1K-2)Wang et al. (2013).The effective mass of electrons can be estimated through Eq. 1DeLong et al. (1985):
(1) |
where is the Boltzmann constant and the carrier density is calculated by the number of electrons () per cell volume (). Using a spherical Fermi surface approximation, the Fermi wave vector can be estimated by = . Assuming that Ni contributes 1.5 electrons ( = 6), we obtain = m-1. The estimated effective mass of electrons = 6 for RbNi2Se2 is significantly enhanced compared with the bare electron mass . Combining the parameters from fitting the electronic specific heat and the quadratic temperature dependent regime of resistivity, the Kadowaki-Woods ratio, /, is calculated to be 0.94 cm(molK2mJ)2, close to 10cm(molK2mJ)2 of heavy fermion systemsKadowaki and Woods (1986). This scaling relation yields RbNi2Se2 with the heavy-fermion behavior.
In SC state, the specific heat data reveals a clear -anomaly under zero field with a maximum at 0.94 K [Fig. 3(c)], suggesting bulk superconductivity. Applying an external magnetic field, the SC transition moves quickly to lower temperatures. As shown in the inset of Fig. 3(c), an upturn appears on below K2. The small upturn could be described by the Schottky anomaly for paramagnetic impurity spinsNeilson et al. (2012), which can be well fitted by = with = 0.07 mJKmol-1.
In order to get information about the SC gap, we extract the electronic specific heat by subtracting the phonon contribution and Schottky anomaly from the total , . The electronic specific heat data against are shown in Fig. 3(d), where has been revealed in the normal state. The is determined to be 0.94 K using an equal-area entropy construction. The heat jump at transition is 0.63, smaller than that of the theoretical value of 1.43 in the BCS weak-coupling scenarioDaams and Carbotte (1981). We employ both one-gap and two-gap BCS models ) to fit the electronic specific heat, where is the size of the SC gap at 0 K. The one-gap model reveals = 0.30 that is smaller than the value of 1.76 in the BCS theoryBardeen et al. (1957). In the two-gap model, the total specific heat can be considered as the sum of electronic contributions from two bands. Our fitting yields the sizes of two gaps of = 2.51 and = 0.25, respectively. The ratio of the contributions from the two gaps is as presented by the dashed lines in Fig. 3(d). The better fitting using the two-gap model indicates that RbNi2Se2 may be a multi-gap superconductor.
The electron-phonon coupling strength is also calculated by employing the inverted McMillan formulaMcMillan (1968):
(2) |
where represents the Coulomb repulsion pseudopotential, which we adopt = 0.13 for this systemAmon et al. (2018). Generally, the for strongly coupled superconductors are close to 1, and 0.5 is viewed as weak coupled superconductorsMcMillan (1968). The for RbNi2Se2 is 0.49, suggesting that RbNi2Se2 is a weakly coupled superconductor, consistent with the specific heat analysis.
To check the multi-band character, we conducted DFT calculations and ARPES measurements on the electronic band structure. Figures 4(a) and 4(b) show the calculated electronic bands of Ni ions and the integrated density of states of Ni, Rb, and Se. The electronic states near the Fermi surface are governed by the orbitals of Ni ions. In Fig 4(c), the measured electronic structure of RbNi2Se2 is compared with the band structure obtained from the DFT calculations. The calculated curves (black lines) are scaled and overlaid onto the photoemission intensity along the direction. The theoretical and experimental data match qualitatively with a renormalization factor of about 1.8. This factor is relatively small compared to other iron-based superconductorsYi et al. (2017), indicating that the electronic correlations are moderate. The moderate correlations suggest that the superconductivity is most likely originated from the electron-phonon coupling, similar to the two-gap superconductor MgB2Bouquet et al. (2001a, b). Figure 4(d) shows the photoemission intensity map of the Fermi surface, overlaid with the DFT calculated Fermi surfaces. The multiple Fermi pockets observed are consistent with the multi-band behavior of this compound, reinforcing the condition for the two-gap feature in the SC state. The band structure is reminiscent of the closely-related compound RbCo2Se2Huang et al. (2021). However, the flat band that induces itinerant ferromagnetism in RbCo2Se2 is observed here to be well below the Fermi level, due to the electron doping resulted from the replacement of Co by Ni.
As Ni2Se2 ( K, Cs, and Tl) superconductors, RbNi2Se2 also exhibits an enhanced effective electron mass with a large Sommerfeld coefficient from specific heat. However, ARPES data reveal that the electronic correlation is not strong. To understand the reason of the large Sommerfeld coefficient, is estimated with and (0) at the Fermi level through the relationshipXiao et al. (2021):
(3) |
From the DOS in Fig. 4(b), is estimated to be 11.68 states/eV per formula, resulting in = 40.98 mJmol-1K-2 that is close to the experimental value of 30.30 mJmol-1K-2. Therefore, the result suggests that the large is related to the large DOS at the Fermi level as proposed in KNi2Se2Fan et al. (2015), instead of the heavy fermion state. For Ni2Se2 ( = K, Rb, and Cs), the increase of the atomic radiuses of alkali metals works as applying a pressure to the Ni-Se layers. The DOS of Ni orbitals at the Fermi surface, the Sommerfeld coefficient , and the electronic correlations are all enhanced. In the framework of the BCS theory, the SC transition temperature increases accordingly, as the experimental observations in KNi2Se2, RbNi2Se2, and CsNi2Se2Neilson et al. (2012); Chen, Huimin and Yang, Jinhu and Cao, Chao and Li, Lin and Su, Qiping and Chen, Bin and Wang, Hangdong and Mao, Qianhui and Xu, Binjie and Du, Jianhua and Fang, Minghu (2016); Wang et al. (2013).
IV Summary
In summary, we have successfully synthesized single crystals of RbNi2Se2 and characterized the physical properties. RbNi2Se2 is found to be a weakly coupled superconductor with K. In normal state, RbNi2Se2 exhibits Fermi liquid behavior and Pauli paramagnetism. In the SC state, the zero-field electronic specific heat can be well described with a two-gap BCS model, indicating that RbNi2Se2 possesses multi-gap feature. DFT calculations and ARPES measurements demonstrate that multi electronic bands of the orbitals of Ni ions cross the Fermi level. Our analyses reveal that the large Sommerfeld coefficient of RbNi2Se2 is originated from the large DOS at the Fermi surface.
V acknowledgments
We thank Shiliang Li for fruitful discussions. Work at Sun Yat-Sen University was supported by the National Natural Science Foundation of China (Grants No. 12174454, 11904414, 11904416, 11974432, U2130101), the Guangdong Basic and Applied Basic Research Foundation (No. 2021B1515120015), National Key Research and Development Program of China (No. 2019YFA0705702, 2018YFA0306001, 2017YFA0206203), and GBABRF-2019A1515011337. Work at SLAB was supported by NSF of China with Grant No. 12004270, and GBABR-2019A1515110517. ARPES work at Rice is supported by the Robert A. Welch Foundation Grant No. C-2024 (M. Y.). P. D. is also supported by US Department of Energy, BES under Grant No. DE-SC0012311. Work at Berkeley was funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Materials Sciences and Engineering Division under Contract No. DE-AC02-05-CH11231 (Quantum Materials program KC2202).
References
- Bednorz et al. (1987) J. G. Bednorz, M. Takashige, and K. A. Müller, Europhys. Lett. 3, 379 (1987), URL http://link.springer.com/10.1007/BF01303701.
- Takada et al. (2003) K. Takada, H. Sakurai, E. Takayama-Muromachi, F. Izumi, R. A. Dilanian, and T. Sasaki, Nature 422, 53 (2003), ISSN 0028-0836, URL http://www.nature.com/articles/nature01450.
- Ren et al. (2008) Z.-A. Ren, G.-C. Che, X.-L. Dong, J. Yang, W. Lu, W. Yi, X.-L. Shen, Z.-C. Li, L.-L. Sun, F. Zhou, et al., EPL (Europhysics Lett. 83, 17002 (2008), ISSN 0295-5075, URL https://iopscience.iop.org/article/10.1209/0295-5075/83/17002.
- Wu et al. (2014) W. Wu, J. Cheng, K. Matsubayashi, P. Kong, F. Lin, C. Jin, N. Wang, Y. Uwatoko, and J. Luo, Nat. Commun. 5, 5508 (2014), ISSN 2041-1723, URL http://dx.doi.org/10.1038/ncomms6508http://www.nature.com/articles/ncomms6508.
- Mu et al. (2017) Q.-G. Mu, B.-B. Ruan, B.-J. Pan, T. Liu, J. Yu, K. Zhao, G.-F. Chen, and Z.-A. Ren, Phys. Rev. B 96, 140504 (2017), ISSN 2469-9950, URL https://link.aps.org/doi/10.1103/PhysRevB.96.140504.
- Cheng et al. (2015) J.-G. Cheng, K. Matsubayashi, W. Wu, J. P. Sun, F. K. Lin, J. L. Luo, and Y. Uwatoko, Phys. Rev. Lett. 114, 117001 (2015), ISSN 0031-9007, URL https://link.aps.org/doi/10.1103/PhysRevLett.114.117001.
- Li et al. (2019) D. Li, K. Lee, B. Y. Wang, M. Osada, S. Crossley, H. R. Lee, Y. Cui, Y. Hikita, and H. Y. Hwang, Nature 572, 624 (2019), ISSN 0028-0836, URL http://dx.doi.org/10.1038/s41586-019-1496-5http://www.nature.com/articles/s41586-019-1496-5.
- Gu and Wen (2022) Q. Gu and H.-H. Wen, Innov. 3, 100202 (2022), ISSN 26666758.
- Neilson et al. (2012) J. R. Neilson, A. Llobet, A. V. Stier, L. Wu, J. Wen, J. Tao, Y. Zhu, Z. B. Tesanovic, N. P. Armitage, and T. M. McQueen, Phys. Rev. B 86, 054512 (2012), ISSN 1098-0121, URL https://link.aps.org/doi/10.1103/PhysRevB.86.054512.
- Chen, Huimin and Yang, Jinhu and Cao, Chao and Li, Lin and Su, Qiping and Chen, Bin and Wang, Hangdong and Mao, Qianhui and Xu, Binjie and Du, Jianhua and Fang, Minghu (2016) Chen, Huimin and Yang, Jinhu and Cao, Chao and Li, Lin and Su, Qiping and Chen, Bin and Wang, Hangdong and Mao, Qianhui and Xu, Binjie and Du, Jianhua and Fang, Minghu, Supercond. Sci. Technol. 29, 045008 (2016), URL https://iopscience.iop.org/article/10.1088/0953-2048/29/4/045008.
- Wang et al. (2013) H. Wang, C. Dong, Q. Mao, R. Khan, X. Zhou, C. Li, B. Chen, J. Yang, Q. Su, and M. Fang, Phys. Rev. Lett. 111, 207001 (2013), ISSN 0031-9007, URL https://link.aps.org/doi/10.1103/PhysRevLett.111.207001.
- Lei et al. (2014) H. Lei, M. Abeykoon, K. Wang, E. S. Bozin, H. Ryu, D. Graf, J. B. Warren, and C. Petrovic, J. Phys. Condens. Matter 26, 015701 (2014), ISSN 0953-8984, URL https://iopscience.iop.org/article/10.1088/0953-8984/26/1/015701.
- Guo et al. (2010) J. Guo, S. Jin, G. Wang, S. Wang, K. Zhu, T. Zhou, M. He, and X. Chen, Phys. Rev. B 82, 180520 (2010), ISSN 1098-0121, URL https://link.aps.org/doi/10.1103/PhysRevB.82.180520.
- Fang et al. (2013) M. Fang, H. Wang, C. Dong, and Q. Huang, J. Phys. Conf. Ser. 449, 012015 (2013), ISSN 1742-6596, URL https://iopscience.iop.org/article/10.1088/1742-6596/449/1/012015.
- Dai (2015) P. Dai, Rev. Mod. Phys. 87, 855 (2015), ISSN 0034-6861, eprint 1503.02340, URL https://link.aps.org/doi/10.1103/RevModPhys.87.855.
- Wang et al. (2016) M. Wang, M. Yi, W. Tian, E. Bourret-Courchesne, and R. J. Birgeneau, Phys. Rev. B 93, 075155 (2016), ISSN 2469-9950, eprint 1508.04482, URL https://link.aps.org/doi/10.1103/PhysRevB.93.075155.
- Yang et al. (2013) J. Yang, B. Chen, H. Wang, Q. Mao, M. Imai, K. Yoshimura, and M. Fang, Phys. Rev. B 88, 064406 (2013), ISSN 1098-0121, URL https://link.aps.org/doi/10.1103/PhysRevB.88.064406.
- Huang et al. (2021) J. Huang, Z. Wang, H. Pang, H. Wu, H. Cao, S.-K. Mo, A. Rustagi, A. F. Kemper, M. Wang, M. Yi, et al., Phys. Rev. B 103, 165105 (2021), ISSN 2469-9950, URL https://link.aps.org/doi/10.1103/PhysRevB.103.165105.
- Fan et al. (2015) Q. Fan, X. P. Shen, M. Y. Li, D. W. Shen, W. Li, X. M. Xie, Q. Q. Ge, Z. R. Ye, S. Y. Tan, X. H. Niu, et al., Phys. Rev. B 91, 125113 (2015), ISSN 1098-0121, eprint arXiv:1411.6455v1, URL https://link.aps.org/doi/10.1103/PhysRevB.91.125113.
- Kresse and Hafner (1993) G. Kresse and J. Hafner, Phys. Rev. B 47, 558 (1993), ISSN 0163-1829, URL https://link.aps.org/doi/10.1103/PhysRevB.47.558.
- Ehrlich and Rivier (1968) A. Ehrlich and D. Rivier, J. Phys. Chem. Solids 29, 1293 (1968), ISSN 00223697, URL https://linkinghub.elsevier.com/retrieve/pii/0022369768901819.
- Böhmer et al. (2016) A. E. Böhmer, V. Taufour, W. E. Straszheim, T. Wolf, and P. C. Canfield, Phys. Rev. B 94, 024526 (2016), ISSN 2469-9950, eprint 1606.00500, URL https://link.aps.org/doi/10.1103/PhysRevB.94.024526.
- Analytis et al. (2014) J. G. Analytis, H.-h. Kuo, R. D. McDonald, M. Wartenbe, P. M. C. Rourke, N. E. Hussey, and I. R. Fisher, Nat. Phys. 10, 194 (2014), ISSN 1745-2473, URL http://dx.doi.org/10.1038/nphys2869http://www.nature.com/articles/nphys2869.
- Li et al. (2020) Q. Li, C. He, J. Si, X. Zhu, Y. Zhang, and H.-H. Wen, Communications Materials 1, 16 (2020), ISSN 2662-4443, eprint 1911.02420.
- Bao et al. (2015) J.-K. Bao, J.-Y. Liu, C.-W. Ma, Z.-H. Meng, Z.-T. Tang, Y.-L. Sun, H.-F. Zhai, H. Jiang, H. Bai, C.-M. Feng, et al., Phys. Rev. X 5, 011013 (2015), ISSN 2160-3308, eprint 1412.0067, URL https://link.aps.org/doi/10.1103/PhysRevX.5.011013.
- DeLong et al. (1985) L. E. DeLong, R. P. Guertin, S. Hasanain, and T. Fariss, Phys. Rev. B 31, 7059 (1985), ISSN 0163-1829, URL https://link.aps.org/doi/10.1103/PhysRevB.31.7059.
- Kadowaki and Woods (1986) K. Kadowaki and S. Woods, Solid State Commun. 58, 507 (1986), ISSN 00381098, URL https://linkinghub.elsevier.com/retrieve/pii/0038109886907854.
- Daams and Carbotte (1981) J. M. Daams and J. P. Carbotte, J. Low Temp. Phys. 43, 263 (1981), ISSN 0022-2291, URL http://link.springer.com/10.1007/BF00116155.
- Bardeen et al. (1957) J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957), ISSN 0031-899X, URL https://www.taylorfrancis.com/books/9780429964251%****␣RbNi2Se2_draft.bbl␣Line␣375␣****https://link.aps.org/doi/10.1103/PhysRev.108.1175.
- McMillan (1968) W. L. McMillan, Phys. Rev. 167, 331 (1968), ISSN 0031-899X, URL https://link.aps.org/doi/10.1103/PhysRev.167.331.
- Amon et al. (2018) A. Amon, E. Svanidze, R. Cardoso-Gil, M. N. Wilson, H. Rosner, M. Bobnar, W. Schnelle, J. W. Lynn, R. Gumeniuk, C. Hennig, et al., Phys. Rev. B 97, 014501 (2018), ISSN 2469-9950, URL https://link.aps.org/doi/10.1103/PhysRevB.97.014501.
- Yi et al. (2017) M. Yi, Y. Zhang, Z. Shen, and D. Lu, npj Quantum Materials 2, 57 (2017).
- Bouquet et al. (2001a) F. Bouquet, Y. Wang, R. A. Fisher, D. G. Hinks, J. D. Jorgensen, A. Junod, and N. E. Phillips, Europhys. Lett. 56, 856 (2001a), ISSN 0295-5075, eprint 0107196, URL https://iopscience.iop.org/article/10.1209/epl/i2001-00598-7.
- Bouquet et al. (2001b) F. Bouquet, R. A. Fisher, N. E. Phillips, D. G. Hinks, and J. D. Jorgensen, Phys. Rev. Lett. 87, 047001 (2001b), ISSN 0031-9007, URL https://link.aps.org/doi/10.1103/PhysRevLett.87.047001.
- Xiao et al. (2021) G. Xiao, Q. Zhu, Y. Cui, W. Yang, B. Li, S. Wu, G.-H. Cao, and Z. Ren, Sci. China Physics, Mech. Astron. 64, 107411 (2021), ISSN 1674-7348, URL https://link.springer.com/10.1007/s11433-021-1731-3.