Fully-gapped superconductivity with preserved time-reversal symmetry in
NiBi3 single crystals
Abstract
We report a study of NiBi3 single crystals by means of electrical-resistivity-, magnetization-, and muon-spin rotation and relaxation (SR) measurements. As a single crystal, NiBi3 adopts a needle-like shape and exhibits bulk superconductivity with K. By applying magnetic fields parallel and perpendicular to the -axis of NiBi3, we establish that its lower- and upper critical fields, as well as the magnetic penetration depths show slightly different values, suggesting a weakly anisotropic superconductivity. In both cases, the zero-temperature upper critical fields are much smaller than the Pauli-limit value, indicating that the superconducting state is constrained by the orbital pair breaking. The temperature evolution of the superfluid density, obtained from transverse-field SR, reveals a fully-gapped superconductivity in NiBi3, with a shared superconducting gap = 2.1 and magnetic penetration depths = 223 and 210 nm for - and , respectively. The lack of spontaneous fields below indicates that time-reversal symmetry is preserved in NiBi3. The absence of a fast muon-spin relaxation and/or precession in the zero-field SR spectra definitely rules out any type of magnetic ordering in NiBi3 single crystals. Overall, our investigation suggests that NiBi3 behaves as a conventional -type superconductor.
I Introduction
Topological superconductors offer many attractive properties, which range from the possibility to host Majorana quasiparticles to enabling topological quantum computing [1, 2, 3]. This has spurred researchers to explore different routes and/or materials to realize them. For instance, superconductors with pairing have been predicted to support Majorana bound states at their vortices [4]. Besides relying on bulk superconductors with nontrivial electronic bands, another approach towards topological superconductivity (SC) consists in combining conventional -wave superconductors with topological insulators or ferromagnets, so as to form heterostructures. In both cases, proximity effects at the interface may lead to a two-dimensional superconducting state with an unconventional pairing [5, 6, 7].
Among the different types of heterostructures, epitaxial Bi/Ni bilayers have been proposed as a promising candidate for hosting chiral -wave SC [8]. In fact, tunneling experiments in such bilayers have shown the coexistence of ferromagnetism (FM) and SC below the superconducting transition, at K [9]. More recently, surface magneto-optic Kerr-effect measurements revealed spontaneous magnetic fields occurring below the onset of SC, which suggest the breaking of time-reversal symmetry (TRS) in the superconducting state of Bi/Ni bilayers [10]. Finally, time-domain terahertz spectroscopy experiments found a nodeless SC state in Bi/Ni bilayers [11], more consistent with a chiral -wave SC [12].
At the same time, there is mounting evidence that the occurrence of SC in Bi/Ni bilayers could be related to the formation of a superconducting NiBi3 phase (with 4 K [13, 14, 15]) during the thin-film growth [16, 17, 18, 19, 19, 20, 21, 22, 23]. This because NiBi3-free Bi/Ni bilayers show no signs of superconductivity [20, 21]. These observations have triggered the researchers’ interests to study the superconducting properties of pure NiBi3. Similar to the Bi/Ni bilayers [9], also bulk NiBi3 shows the coexistence of SC and FM [24, 25, 26, 27]. Thus, magnetization data in its superconducting state, show a clear loop of ferromagnetic hysteresis [25, 26, 15]. Generally, superconductivity and ferromagnetism are antagonistic ground states. However, akin to certain U-based materials [28, 29, 30], NiBi3, too, belongs to the rare class of ferromagnetic superconductors, potentially able to host spin-triplet pairing.
To date, the superconducting properties of NiBi3 have been mostly investigated via magnetic- and transport measurements [13, 14, 15, 24, 25, 26, 27]. Yet, the microscopic nature of its SC, in particular, the superconducting order parameter, has not been explored and awaits further investigation. In addition, it is not yet clear if the observed breaking of TRS and the unconventional SC of epitaxial Bi/Ni bilayers can be attributed to the formation of the NiBi3 phase or not [10].
To clarify the above issues, we synthesized NiBi3 single crystals using the flux method, and studied their superconducting properties by means of electrical-resistivity-, magnetization-, and muon-spin rotation and relaxation (SR) measurements. We found that NiBi3 exhibits a fully-gapped superconducting state with a preserved TRS. The absence of any magnetic order, on the surface or in the bulk of NiBi3 single crystals, was also confirmed. Our results suggest that NiBi3 behaves as a conventional -type superconductor and, thus, the unconventional superconducting properties observed in Bi/Ni bilayers cannot be attributed to the NiBi3 phase.

II Experimental details
The NiBi3 single crystals were grown from molten bismuth flux. High-purity Bi pieces (Alfa Aesar, 99.99%) and Ni powders (Alfa Aesar, 99.9%) with a 10:1 ratio were loaded in an Al2O3 crucible, which was sealed in a quartz ampoule under high vacuum. The quartz ampoule was heated up to 1050 ∘C and then kept at this temperature for over 24 hours. Finally, it was slowly cooled down to 360 ∘C at a rate of 2 ∘C/h, with the remaining Bi flux being removed via centrifugation. The typical dimensions of the obtained needle-like single crystals were about 3 mm 0.3 mm 0.3 mm (see below). The crystals were checked by powder x-ray diffraction (XRD), recorded using a Bruker D8 diffractometer. Further characterization involved electrical-resistivity- and magnetization measurements, performed on a Quantum Design physical property measurement system (PPMS) and a magnetic property measurement system (MPMS), respectively. For the electrical-resistivity measurements, the dc current was applied along the -axis.
The bulk SR measurements were carried out at the multipurpose surface-muon spectrometer (Dolly) on the E1 beamline of the Swiss muon source at Paul Scherrer Institut, Villigen, Switzerland. The crystals were aligned and mounted on a 25-m thick copper foil, which ensured thermalization at low temperatures. The magnetic fields were applied both parallel and perpendicular to the -axis of the NiBi3 single crystal. The time-differential SR data were collected upon heating and then analyzed by means of the musrfit software package [31].

III Results and discussion
III.1 Crystal structure
The crystal structure and phase purity of NiBi3 single crystals were checked via XRD at room temperature. As shown in Fig. I, we performed Rietveld refinements of the XRD pattern by using the Fullprof suite [32]. Consistent with previous results, we confirm that NiBi3 crystallizes in a centrosymmetric orthorhombic structure with a space group (No. 62) [13, 14, 15, 24, 25, 26, 27]. In this structure, bismuth atoms form an octahedral array, while nickel atoms form part of linear chains [25]. The crystal structure of a unit cell is shown in the inset of Fig. I(b). The refined lattice parameters, = 8.8799(1) Å, = 4.09831(6) Å, and = 11.4853(1) Å, are in good agreement with the results reported in the literature [13, 14, 15, 24, 25, 26, 27]. According to the refinements in Fig. I, a tiny amount of elemental bismuth (1%) was also identified, here attributed to the remanent Bi-flux on the surface of single crystals. Considering its small amount and the fact that Bi is not superconducting in the studied temperature range [33], the bismuth presence has negligible effects on the superconducting properties of NiBi3. We also performed XRD measurement on a NiBi3 single crystal. As shown in Fig. I(b), only the (004) reflection [the strongest among the (00)-reflections] was detected, confirming the single-crystal nature of the NiBi3 sample.
III.2 Bulk superconductivity

The bulk superconductivity of NiBi3 single crystals was first characterized by magnetic-susceptibility measurements, using both field-cooling (FC) and zero-field-cooling (ZFC) protocols in an applied field of 1 mT. As shown in Fig. II(a), a clear diamagnetic response appears below the superconducting transition at = 4.05 K. Similar results were obtained when applying the magnetic field parallel to the -axis. After accounting for the demagnetization factor , which is estimated to be 0.5 for and 0 for [34], the superconducting shielding fraction of NiBi3 was found to be close to 100%, indicative of bulk SC.
The temperature-dependent electrical resistivity of NiBi3 single crystals was measured from 2 K up to room temperature. It reveals a metallic behavior, without any anomalies associated with phase transitions in the normal state [see inset in Fig. II(b)]. The electrical resistivity in the low-temperature region is plotted in the main panel of Fig. II(b), where the superconducting transition (with K, K and K) is clearly seen. As shown by the dashed-line in Fig. II, the value is consistent with the onset of the superconducting transition in the magnetic susceptibility. Therefore, the values were used to determine the upper critical field of NiBi3. The curve can be described by the Bloch-Grüneisen-Mott (BGM) formula [35, 36]. Here, the first term is the residual resistivity due to the scattering of conduction electrons on the static defects of the crystal lattice, while the second term describes the electron-phonon scattering, with being the characteristic (Debye) temperature and a coupling constant. The third term represents a contribution due to - interband scattering, being the Mott coefficient [37, 38]. The fit in Fig. II (solid black line) results in µcm, µcm, K, and 10-7 µcmK-3. The derived is comparable with the value determined from low- specific heat [13]. The fairly large residual resistivity ratio (RRR), i.e., (300 K)/ 18, and the sharp superconducting transition ( = 0.15 K) both indicate a good sample quality.
III.3 Lower and upper critical fields
To determine the lower critical field , the field-dependent magnetization of a NiBi3 single crystal was measured at various temperatures up to 3.75 K. Figure III.2(a)-(b) plots the for and , respectively. The estimated values at different temperatures (accounting for a demagnetization factor) are summarized in Fig. III.2(c). The solid lines are fits to and yield the lower critical fields = 7.4(1) and 6.2(1) mT for and , respectively. The zero-temperature / ratio (1.2) suggests a weakly anisotropic superconductivity in the NiBi3 single crystal.
We also performed temperature-dependent electrical-resistivity measurements at various applied magnetic fields, as well as studied the field-dependent magnetization at various temperatures by applying the magnetic field either parallel or perpendicular to the -axis of the NiBi3 single crystal. As shown in Fig. III.3(a), upon increasing the magnetic field, the superconducting transition in shifts to lower temperatures. Similarly, in the data, the diamagnetic signal vanishes once the applied magnetic field exceeds the upper critical field . For , no zero resistivity is observed down to K for 0.35 T. For , the zero resistivity already vanishes for 0.25 T, implying that, in NiBi3, the value along the -axis is larger than that perpendicular to it.
Figure III.3(e) and (f) summarizes the upper critical fields vs. the reduced superconducting transition temperature for both and , as identified from the and data. The temperature evolution of the upper critical field is well described by the Werthamer-Helfand-Hohenberg (WHH) model [39]. As shown by the dash-dotted lines in Fig. III.3(e)-(f), the WHH model agrees remarkably well with the experimental data and provides = 0.70(2) T and = 0.41(1), respectively. These values are comparable to previous results [13, 15, 27] and both are much smaller than the Pauli-limiting field (i.e., 1.86 7.4 T). Consequently, the orbital pair-breaking mechanism seems to be dominant in NiBi3. Similarly to the lower critical fields, also the / ratio ( 1.7) confirms the weakly anisotropic SC in NiBi3.
In the Ginzburg-Landau (GL) theory of superconductivity, the magnetic penetration depth is related to the coherence length , and the lower critical field via ln, where T nm2 is the quantum of magnetic flux, = / is the GL parameter [40]. By using and values [calculated from = /2], the resulting and 238(2) nm for and are both compatible with the experimental value determined from SR data. All the superconducting parameters are summarized in Table IV.

III.4 TF-SR and fully-gapped superconductivity

To investigate the single-crystal anisotropy of superconducting pairing in NiBi3, we carried out systematic transverse-field (TF-) SR measurements, where we applied the field along different orientations. Representative TF-SR spectra collected in a field of 14.5 mT in the superconducting- and normal states of NiBi3 are shown in Fig. III.4(a) for -axis and in Fig. III.4(b) for -axis. In the latter case, the TF-SR spectra were also collected in a field of 30 mT [see Fig. III.4(c)]. In case of a type-II superconductor (see, e.g., the 0.3-K data in Fig. III.4), the development of a flux-line lattice (FLL) causes an inhomogeneous field distribution and, thus, it gives rise to an additional damping in the TF-SR spectra [41]. These are generally modeled using [42]:
(1) |
Here (85%), (15%) and , are the initial asymmetries and local fields sensed by the implanted muons in the sample and sample holder (i.e., Cu) or in the residual bismuth, /2 = 135.53 MHz/T is the muon gyromagnetic ratio, is a shared initial phase, and is the Gaussian relaxation rate of the th component. Note that, in the studied temperature range, the effects of the residual bismuth are negligible due to its non-magnetic and non-superconducting nature. We find that two oscillations (i.e., ) are required to describe the TF-SR spectra for both - and -axis (see solid lines in Fig. III.4). Indeed, as shown in Fig. III.4(a)-(c), oscillations with two distinct frequencies can be clearly identified and, generally, the model with two oscillations provides a better fit than that with a single oscillation. For example, in Fig. III.4(c), the two-oscillation model yields a goodness-of-fit parameter 1.4, twice smaller than the single-oscillation fit ( 3.2). The fast Fourier transforms (FFT) of the TF-SR spectra at 0.3 K are shown by dash-dotted lines in Fig. III.4(d)-(f), which illustrate the two components ( and ) and the background signal (). The two-peak FFT in NiBi3 might be related to two different muon-stopping sites, a feature to be confirmed by future DFT calculations.
The derived muon-spin relaxation rates are small and temperature-independent in the normal state, but below they start to increase due to the onset of FLL and the increased superfluid density. At the same time, a diamagnetic field shift, , appears below , with , where and is the applied external field (see, e.g., the inset of Fig. III.4). The effective Gaussian relaxation rate can be calculated from [42]. Then, the superconducting Gaussian relaxation rate , can be extracted by subtracting the nuclear contribution according to . Here, is the nuclear relaxation rate, almost constant in the covered temperature range and relatively small in NiBi3, as confirmed also by the ZF-SR data (see Fig. III.5). Since the applied TF fields (14.5 and 30 mT) are not really small compared to the modest upper critical fields of NiBi3 [see Fig. III.3(e) and (f)], to calculate the magnetic penetration depth from we had to consider the overlap of the vortex cores. Consequently, in our case, was calculated by means of , where [43, 44].

Figure III.4 summarizes the temperature-dependent inverse square of the magnetic penetration depth [proportional to the superfluid density, i.e., ] for both and . For , the data collected in a field of 30 mT are also presented. The was analyzed by applying different models, generally described by:
(2) |
Here, is the Fermi function [40]; is an angle-dependent gap function, where is the maximum gap value and is the angular dependence of the gap, equal to 1 and for an - and -wave model, respectively, with being the azimuthal angles. The temperature dependence of the gap is assumed to follow [40, 45], where is the gap value at 0 K. As can be clearly seen in Fig. III.4 (see dashed line), the -wave model exhibits a significant deviation from the experimental data. While the temperature-invariant superfluid density below /3 strongly suggests the absence of low-energy excitations and, hence, a fully-gapped superconducting state in NiBi3. Consequently, the is more consistent with the -wave model. As shown by the solid line in Fig. III.4, in both cases (i.e., - and -axis), the -wave model describes very well across the entire temperature range and yields a zero-temperature gap = 2.10(5) . Such isotropic gap value along different crystal directions is consistent with previous Andreev-reflection spectroscopy results [46]. For , the estimated is 223(2) nm, while for , we find = 216(2) and 210(2) nm for TF-14.5 mT and TF-30 mT, respectively. The small / ratio (here, 1.03) and the similar gap sizes along different crystal directions confirm once more the weakly anisotropic SC of NiBi3 single crystals. Note that, although the single-oscillation model also gives a temperature-independent superfluid density below 1/3 as the two-oscillation model, the estimated values by using the former model are significantly different from the values estimated from the magnetization data.
III.5 ZF-SR and preserved time-reversal symmetry

To search for possible ferromagnetism and breaking of the time-reversal symmetry in a NiBi3 single crystal, we performed zero-field (ZF-) SR measurements covering both the normal- and superconducting states. As shown in Fig. III.5, neither coherent oscillations nor fast decays could be identified in the spectra collected above- (4.5 K) and below (0.3 K), hence implying the lack of any magnetic order or fluctuations. As clearly demonstrated in the inset of Fig. III.5(a), the ZF-SR spectra are almost flat on a short time scale (i.e., ¡ 0.5 s), confirming again the absence of fast oscillations which would be caused by a possible ferromagnetic ordering, as instead reported in previous work [24, 25]. Thus, in the absence of external fields, the muon-spin relaxation in NiBi3 is mainly due to the randomly oriented nuclear moments, which can be modeled by a Gaussian Kubo-Toyabe relaxation function [47, 41]. Here, is the zero-field Gaussian relaxation rate. The ZF-SR spectra were fitted by considering also an additional electronic contribution. The solid lines in Fig. III.5(a) represent fits to , where is the zero-field exponential muon-spin relaxation rate, and are the same as in TF-SR [see Eq. (1)]. The derived and values as a function of temperature are shown in Fig. III.5(b) and (c), respectively. Neither nor show a systematic enhancement below . Here, the jump of at 3.5 K [Fig. III.5(c)] is not related to a TRS-breaking effect, but to the correlated decrease of [Fig. III.5(b)]. The lack of an additional relaxation below excludes a possible TRS-breaking effect in the superconducting state of the NiBi3 single crystal, here also reflected in the practically overlapping datasets shown in Fig. III.5(a).
IV Discussion

First, we discuss the absence of ferromagnetism in NiBi3. Previous works found evidence for either extrinsic or intrinsic ferromagnetism in NiBi3 single crystals. In the extrinsic case, the amorphous Ni impurities in NiBi3 lead to a clear drop in magnetization at temperatures close to their Curie temperature [15]. In NiBi3, such FM can even coexist with SC, as demonstrated by the ferromagnetic hysteresis loops observed in the magnetization data [24, 25, 26]. Besides bulk crystals with Ni impurities, also submicrometer-sized NiBi3 particles, or quasi-one-dimensional nanoscale strips of NiBi3 undergo a ferromagnetic order [26]. In addition, below 150 K, electron-spin-resonance data reveal ferromagnetic-like fluctuations on the surface of NiBi3 crystals [27]. The above evidence suggests that NiBi3 is a possible ferromagnetic superconductor, which might enable spin-triplet SC pairing, analogous to that found in other ferromagnetic superconductors [28, 29, 30]. Here, we applied the ZF-SR technique to detect a possible ferromagnetic order in NiBi3 single crystals. Due to their large magnetic moment, muons represent one of the most sensitive magnetic probes, able to sense very low internal fields ( mT), and thus, to detect local magnetic fields of either nuclear or electronic origin [41, 53]. In contrast to previous studies, our ZF-SR results do not evidence any ferromagnetic order or fluctuations in NiBi3, thus implying that the previously reported ferromagnetism is most likely of extrinsic nature. In addition, our ZF-SR results indicate that time-reversal symetry is preserved in the superconducting state of NiBi3. This conclusion is further supported by recent high-resolution surface magneto-optic Kerr effect measurements in NiBi3 single crystals, which also do not find any traces of a spontaneous Kerr signal in the superconducting state [54].
Second, we discuss the weakly anisotropic superconductivity in NiBi3. Although NiBi3 single crystals exhibit a quasi-one dimensional needle-like shape [see inset in Fig. II(a)], their crystal structure is definitely three dimensional, with =8.8799 Å, = 4.09831 Å, and = 11.4853 Å (see Fig. I). Such an intrinsically three-dimensional structure could naturally explain the almost isotropic- (or weakly anisotropic) electronic properties of NiBi3. As confirmed by our study, the lower critical field (Fig. III.2), the upper critical field (Fig. III.3), the magnetic penetration depth , and the superconducting gap (Fig. III.4), all show comparable values for - and -axis. These results clearly indicate a weakly anisotropic SC in NiBi3 single crystals, as also observed in NiBi3 thin films [18]. Moreover, electrical-resistivity measurements reveal almost identical upper critical fields when the magnetic field is rotated in the -plane of a NiBi3 single crystal [13]. Future electronic band-structure calculations could provide further hints about the observed weak anisotropy.
Finally, let us compare the superconducting parameters of NiBi3 with those of other superconductors. By using the SC parameters obtained from the measurements presented here, we calculate an effective Fermi temperature 1.5 103 K for NiBi3 (here, we consider the case). We recall that is proportional to /, where and are the carrier density and the effective mass. The estimated 8 , indicates a moderate degree of electronic correlations in NiBi3. Different families of superconductors can be classified according to their / ratios into a so-called Uemura plot [49]. In such plot, many conventional superconductors, as e.g., Al, Sn, and Zn, are located at 10-4. Conversely, several types of unconventional superconductors, including heavy fermions, organics, high- iron pnictides, and cuprates, all lie in the 10-2 10-1 band. Between these two categories lie a few different types of superconductors, including multigap- and noncentrosymmetric superconductors, etc. [48, 49, 50, 51, 52]. Although located clearly far off the conventional band, our TF- and ZF-SR results suggest that the NiBi3 superconductivity is more consistent with a conventional fully-gapped SC with preserved time-reversal symmetry. Such conclusion is also supported by the linear suppression of under applied pressure. For instance, the of NiBi3 decreases from 4 K to 3 K when the pressure increases from 0 to 2.2 GPa [55].
Property | Unit | ||
---|---|---|---|
K | 4.05(5) | 4.05(5) | |
K | 3.9(1) | 3.9(1) | |
mT | 7.4(1) | 6.2(1) | |
T | 0.70(1) | 0.41(1) | |
(-wave) | 2.10(5) | 2.10(5) | |
(-wave) | 3.00(5) | 3.00(5) | |
nm | 223(2) | 216(2) | |
nm | 21.7(2) | 28.3(3) | |
nm | 229(2) | 238(2) | |
– | 10.5(2) | 8.4(2) |
V Conclusion
To summarize, we investigated the superconducting properties of flux-grown NiBi3 single crystals by means of electrical resistivity-, magnetization-, and SR measurements. NiBi3 was shown to exhibit bulk SC with 4.1 K. By applying magnetic fields along different crystal directions, we obtained the lower critical field , the upper critical field , and the magnetic penetration depth for both the - and cases. Both and values are much smaller than the Pauli-limit field, implying that the orbital pair-breaking mechanism is dominant in NiBi3. Although the NiBi3 single crystals show needle-like shapes, their superconducting properties are only weakly anisotropic, as reflected in the small ratios of /, /, and /. The temperature dependence of the NiBi3 superfluid density reveals a nodeless SC, well described by an isotropic -wave model. The relatively large gap value, here, = 2.1 , suggests strong electron-phonon interactions in NiBi3. Further, the lack of spontaneous magnetic fields below indicates that the time-reversal symmetry is preserved in the NiBi3 superconductor. The absence of a fast relaxation and/or oscillations in the ZF-SR spectra excludes a magnetic order on the surface or in the bulk of NiBi3 single crystals, thus implying that the previously observed ferromagnetism is most likely of extrinsic origin. Overall, our systematic results suggest that, in contrast to the unconventional superconductivity observed in Bi/Ni bilayers, NiBi3 behaves as a conventional -type superconductor.
Acknowledgements.
This work was supported by the Natural Science Foundation of Shanghai (Grants No. 21ZR1420500 and 21JC1402300), Natural Science Foundation of Chongqing (Grant No. 2022NSCQ-MSX1468), and the Schweizerische Nationalfonds zur Förderung der Wissenschaftlichen Forschung (SNF) (Grants No. 200021_188706 and 206021_139082). Y.X. acknowledges support from the Shanghai Pujiang Program (Grant No. 21PJ1403100). We acknowledge the allocation of beam time at the Swiss muon source (Dolly SR spectrometer).References
- Qi and Zhang [2011] X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83, 1057 (2011).
- Kitaev [2001] A. Y. Kitaev, Unpaired Majorana fermions in quantum wires, Phys. Usp. 44, 131 (2001).
- Nayak et al. [2008] C. Nayak, S. H. Simon, A. Stern, M. Freedman, and S. Das Sarma, Non-abelian anyons and topological quantum computation, Rev. Mod. Phys. 80, 1083 (2008).
- Fu and Kane [2008] L. Fu and C. L. Kane, Superconducting proximity effect and Majorana fermions at the surface of a topological insulator, Phys. Rev. Lett. 100, 096407 (2008).
- Xu et al. [2014] J.-P. Xu, C. Liu, M.-X. Wang, J. Ge, Z.-L. Liu, X. Yang, Y. Chen, Y. Liu, Z.-A. Xu, C.-L. Gao, D. Qian, F.-C. Zhang, and J.-F. Jia, Artificial topological superconductor by the proximity effect, Phys. Rev. Lett. 112, 217001 (2014).
- Bergeret et al. [2005] F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Odd triplet superconductivity and related phenomena in superconductor-ferromagnet structures, Rev. Mod. Phys. 77, 1321 (2005).
- Buzdin [2005] A. I. Buzdin, Proximity effects in superconductor-ferromagnet heterostructures, Rev. Mod. Phys. 77, 935 (2005).
- Wang et al. [2017] J. Wang, X. Gong, G. Yang, Z. Lyu, Y. Pang, G. Liu, Z. Ji, J. Fan, X. Jing, C. Yang, F. Qu, X. Jin, and L. Lu, Anomalous magnetic moments as evidence of chiral superconductivity in a Bi/Ni bilayer, Phys. Rev. B 96, 054519 (2017).
- LeClair et al. [2005] P. LeClair, J. S. Moodera, J. Philip, and D. Heiman, Coexistence of ferromagnetism and superconductivity in Ni/Bi bilayers, Phys. Rev. Lett. 94, 037006 (2005).
- Gong et al. [2017] X. Gong, M. Kargarian, A. Stern, D. Yue, H. Zhou, X. Jin, V. M. Galitski, V. M. Yakovenko, and J. Xia, Time-reversal symmetry-breaking superconductivity in epitaxial bismuth/nickel bilayers, Sci. Adv. 3, e1602579 (2017).
- Chauhan et al. [2019] P. Chauhan, F. Mahmood, D. Yue, P.-C. Xu, X. Jin, and N. P. Armitage, Nodeless bulk superconductivity in the time-reversal symmetry breaking Bi/Ni bilayer system, Phys. Rev. Lett. 122, 017002 (2019).
- Chao [2019] S.-P. Chao, Superconductivity in a Bi/Ni bilayer, Phys. Rev. B 99, 064504 (2019).
- Fujimori et al. [2000] Y. Fujimori, S.-i. Kan, B. Shinozaki, and T. Kawaguti, Superconducting and normal state properties of NiBi3, J. Phys. Soc. Jpn. 69, 3017 (2000).
- Kumar et al. [2011] J. Kumar, A. Kumar, A. Vajpayee, B. Gahtori, D. Sharma, P. K. Ahluwalia, S. Auluck, and V. P. S. Awana, Physical property and electronic structure characterization of bulk superconducting Bi3Ni, Supercond. Sci. Technol. 24, 085002 (2011).
- Silva et al. [2013] B. Silva, R. F. Luccas, N. M. Nemes, J. Hanko, M. R. Osorio, P. Kulkarni, F. Mompean, M. García-Hernández, M. A. Ramos, S. Vieira, and H. Suderow, Superconductivity and magnetism on flux-grown single crystals of NiBi3, Phys. Rev. B 88, 184508 (2013).
- Siva et al. [2015] V. Siva, K. Senapati, B. Satpati, S. Prusty, D. K. Avasthi, D. Kanjilal, and P. K. Sahoo, Spontaneous formation of superconducting NiBi3 phase in Ni-Bi bilayer films, J. Appl. Phys. 117, 083902 (2015).
- Siva et al. [2016] V. Siva, P. C. Pradhan, G. Santosh Babu, M. Nayak, P. K. Sahoo, and K. Senapati, Superconducting proximity effect in NiBi3-Ni-NiBi3 trilayer system with sharp superconductor-ferromagnet boundaries, J. Appl. Phys. 119, 063902 (2016).
- Wang et al. [2018] W.-L. Wang, Y.-M. Zhang, Y.-F. Lv, H. Ding, L. Wang, W. Li, K. He, C.-L. Song, X.-C. Ma, and Q.-K. Xue, Anisotropic superconductivity and elongated vortices with unusual bound states in quasi-one-dimensional nickel-bismuth compounds, Phys. Rev. B 97, 134524 (2018).
- Bhatia et al. [2018] E. Bhatia, A. Talapatra, J. R. Mohanty, and K. Senapati, Superconductivity, Kondo effect, and observation of self-organized pattern formation in intermetallic NiBi3 thin films, Intermetallics 94, 160 (2018).
- Liu et al. [2018] L. Y. Liu, Y. T. Xing, I. L. C. Merino, H. Micklitz, D. F. Franceschini, E. Baggio-Saitovitch, D. C. Bell, and I. G. Solórzano, Superconductivity in Bi/Ni bilayer system: Clear role of superconducting phases found at Bi/Ni interface, Phys. Rev. Materials 2, 014601 (2018).
- Vaughan et al. [2020] M. Vaughan, N. Satchell, M. Ali, C. J. Kinane, G. B. G. Stenning, S. Langridge, and G. Burnell, Origin of superconductivity at nickel-bismuth interfaces, Phys. Rev. Res. 2, 013270 (2020).
- Das et al. [2022] B. Das, M. Sahoo, A. Patra, A. K. Yadav, S. N. Jha, P. Samal, K. Senapati, and P. K. Sahoo, Phase evolution in thermally annealed Ni/Bi multilayers studied by X-ray absorption spectroscopy, Phys. Chem. Chem. Phys. 24, 4415 (2022).
- Doria et al. [2022] M. M. Doria, L. Liu, Y. Xing, I. L. C. Merino, F. J. Litterst, and E. Baggio-Saitovitch, Shape resonances and the dependence on film thickness of Ni/Bi systems, Supercond. Sci. Technol. 35, 015012 (2022).
- Liu et al. [2020] L. Liu, Y. Xing, I. L. C. Merino, D. F. Franceschini, I. G. Solórzano, and E. Baggio-Saitovitch, Magnetic properties of superconducting phases NiBi and NiBi3 formed during pulsed laser deposition of Ni-Bi films, J. Magn. Magn. Mater. 514, 167275 (2020).
- Piñeiro et al. [2011] E. L. M. Piñeiro, B. L. R. Herrera, R. Escudero, and L. Bucio, Possible coexistence of superconductivity and magnetism in intermetallic NiBi3, Solid State Commun. 151, 425 (2011).
- Herrmannsdörfer et al. [2011] T. Herrmannsdörfer, R. Skrotzki, J. Wosnitza, D. Köhler, R. Boldt, and M. Ruck, Structure-induced coexistence of ferromagnetic and superconducting states of single-phase Bi3Ni seen via magnetization and resistance measurements, Phys. Rev. B 83, 140501(R) (2011).
- Zhu et al. [2012] X. Zhu, H. Lei, C. Petrovic, and Y. Zhang, Surface-induced magnetic fluctuations in a single-crystal NiBi3 superconductor, Phys. Rev. B 86, 024527 (2012).
- Saxena et al. [2000] S. S. Saxena, P. Agarwal, K. Ahilan, F. M. Grosche, R. K. W. Haselwimmer, M. J. Steiner, E. Pugh, I. R. Walker, S. R. Julian, P. Monthoux, G. G. Lonzarich, A. Huxley, I. Sheikin, D. Braithwaite, and J. Flouquet, Superconductivity on the border of itinerant-electron ferromagnetism in UGe2, Nature 406, 587 (2000).
- Aoki et al. [2001] D. Aoki, A. Huxley, E. Ressouche, D. Braithwaite, J. Flouquet, J.-P. Brison, E. Lhotel, and C. Paulsen, Coexistence of superconductivity and ferromagnetism in URhGe, Nature 413, 613 (2001).
- Huy et al. [2007] N. T. Huy, A. Gasparini, D. E. de Nijs, Y. Huang, J. C. P. Klaasse, T. Gortenmulder, A. de Visser, A. Hamann, T. Görlach, and H. v. Löhneysen, Superconductivity on the border of weak itinerant ferromagnetism in UCoGe, Phys. Rev. Lett. 99, 067006 (2007).
- Suter and Wojek [2012] A. Suter and B. M. Wojek, Musrfit: A free platform-independent framework for SR data analysis, Phys. Procedia 30, 69 (2012).
- Rodríguez-Carvajal [1993] J. Rodríguez-Carvajal, Recent advances in magnetic structure determination by neutron powder diffraction, Physica B: Condens. Matter 192, 55 (1993).
- Prakash et al. [2017] O. Prakash, A. Kumar, A. Thamizhavel, and S. Ramakrishnan, Evidence for bulk superconductivity in pure bismuth single crystals at ambient pressure, Science 355, 52 (2017).
- Prozorov and Kogan [2018] R. Prozorov and V. G. Kogan, Effective demagnetizing factors of diamagnetic samples of various shapes, Phys. Rev. Appl. 10, 014030 (2018).
- Bloch [1930] F. Bloch, Zum elektrischen Widerstandsgesetz bei tiefen Temperaturen, Z. Phys. 59, 208 (1930).
- Blatt [1968] F. J. Blatt, Physics of Electronic Conduction in Solids (McGraw-Hill, New York, 1968) p. 185–190.
- Mott and Jones [1958] N. F. Mott and H. Jones, The Theory of the Properties of Metals and Alloys (Oxford University Press, London, 1958).
- Mott [1964] N. F. Mott, Electrons in transition metals, Adv. Phys. 13, 325 (1964), p. 403.
- Werthamer et al. [1966] N. R. Werthamer, E. Helfand, and P. C. Hohenberg, Temperature and purity dependence of the superconducting critical field, . III. Electron spin and spin-orbit effects, Phys. Rev. 147, 295 (1966).
- Tinkham [1996] M. Tinkham, Introduction to Superconductivity, 2nd ed. (Dover Publications, Mineola, NY, 1996).
- Yaouanc and de Réotier [2011] A. Yaouanc and P. D. de Réotier, Muon Spin Rotation, Relaxation, and Resonance: Applications to Condensed Matter (Oxford University Press, Oxford, 2011).
- Maisuradze et al. [2009] A. Maisuradze, R. Khasanov, A. Shengelaya, and H. Keller, Comparison of different methods for analyzing SR line shapes in the vortex state of type-II superconductors, J. Phys.: Condens. Matter 21, 075701 (2009), and references therein.
- Barford and Gunn [1988] W. Barford and J. M. F. Gunn, The theory of the measurement of the London penetration depth in uniaxial type-II superconductors by muon spin rotation, Physica C 156, 515 (1988).
- Brandt [2003] E. H. Brandt, Properties of the ideal Ginzburg-Landau vortex lattice, Phys. Rev. B 68, 054506 (2003).
- Carrington and Manzano [2003] A. Carrington and F. Manzano, Magnetic penetration depth of MgB2, Physica C 385, 205 (2003).
- Zhao et al. [2018] G. J. Zhao, X. X. Gong, P. C. Xu, B. C. Li, Z. Y. Huang, X. F. Jin, X. D. Zhu, and T. Y. Chen, Singlet superconductivity in a single-crystal NiBi3 superconductor, Supercond. Sci. Technol. 31, 125005 (2018).
- Kubo and Toyabe [1967] R. Kubo and T. Toyabe, A stochastic model for low-field resonance and relaxation, in Magnetic Resonance and Relaxation. Proceedings of the XIVth Colloque Ampère, edited by R. Blinc (North-Holland, Amsterdam, 1967) pp. 810–823.
- Shang et al. [2019a] T. Shang, A. Amon, D. Kasinathan, W. Xie, M. Bobnar, Y. Chen, A. Wang, M. Shi, M. Medarde, H. Q. Yuan, and T. Shiroka, Enhanced and multiband superconductivity in the fully-gapped ReBe22 superconductor, New J. Phys. 21, 073034 (2019a).
- Uemura et al. [1991] Y. J. Uemura, L. P. Le, G. M. Luke, B. J. Sternlieb, W. D. Wu, J. H. Brewer, T. M. Riseman, C. L. Seaman, M. B. Maple, M. Ishikawa, D. G. Hinks, J. D. Jorgensen, G. Saito, and H. Yamochi, Basic similarities among cuprate, bismuthate, organic, Chevrel-phase, and heavy-fermion superconductors shown by penetration-depth measurements, Phys. Rev. Lett. 66, 2665 (1991).
- Uemura [2006] Y. J. Uemura, Twin spin/charge roton mode and superfluid density: Primary determining factors of in high- superconductors observed by neutron, ARPES, and SR, Physica B 374, 1 (2006).
- Xu et al. [2020] Y. Xu, S. Jöhr, L. Das, J. Kitagawa, M. Medarde, T. Shiroka, J. Chang, and T. Shang, Crossover from multiple- to single-gap superconductivity in Nb5Ir3-xPtxO alloys, Phys. Rev. B 101, 134513 (2020).
- Shang et al. [2019b] T. Shang, J. Philippe, J. A. T. Verezhak, Z. Guguchia, J. Z. Zhao, L.-J. Chang, M. K. Lee, D. J. Gawryluk, E. Pomjakushina, M. Shi, M. Medarde, H.-R. Ott, and T. Shiroka, Nodeless superconductivity and preserved time-reversal symmetry in the noncentrosymmetric Mo3P superconductor, Phys. Rev. B 99, 184513 (2019b).
- Amato [1997] A. Amato, Heavy-fermion systems studied by SR technique, Rev. Mod. Phys. 69, 1119 (1997).
- Wang et al. [2022] J. Wang, C. Farhang, D. Yue, X. Jin, X. Zhu, and J. Xia, Absence of spontaneous time-reversal symmetry breaking and ferromagnetism in superconducting NiBi3 single crystal, arXiv:2208.10645 [cond-mat.supr-con] (2022).
- Gati et al. [2019] E. Gati, L. Xiang, L.-L. Wang, S. Manni, P. C. Canfield, and S. L. Bud’ko, Effect of pressure on the physical properties of the superconductor NiBi3, J. Phys.: Condens. Matter 31, 035701 (2019).