Josephson Junction of Nodal Superconductors with Rashba and Ising Spin-Orbit coupling
Abstract
We study the effect of a Rashba spin-orbit coupling on the nodal superconducting phase of an Ising superconductor. Such nodal phase was predicted to occur when applying an in-plane field beyond the Pauli limit to a superconducting monolayer transition metal dichalcogenides (TMD). Generically, Rashba spin-orbit is known to lift the chiral symmetry that protects the nodal points, resulting in a fully gapped phase. However, when the magnetic field is applied along the line, a residual vertical mirror symmetry protects a nodal crystalline phase. We study a single-band tight-binding model that captures the low energy physics around the pocket of monolayer TMD. We calculate the topological properties, the edge state structure, and the current phase relation in a Josephson junction geometry of the nodal crystalline phase. We show that while the nodal crystalline phase is characterized by localized edge modes on non-self-reflecting boundaries, the current phase relation exhibits a trivial periodicity in the presence of Rashba spin-orbit coupling.
I Introduction
Transition metal dichalcogenides (TMDs) such as and have been proposed and experimentally confirmed to be an ideal platform for in-depth explorations for unconventional superconductivity - both intrinsic and externally induced [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14].
More recently, cutting-edge advances in fabrication techniques have facilitated the engineering of layered systems from these TMDs where the constituent layers are held together by weak Van der Waals force[15, 16]. Here, some systems are found to retain their superconducting property even down to the monolayer limit [17, 18, 19, 20, 21, 22, 23, 14, 24, 25].
Unlike their bulk counterparts, many monolayer and few-layer TMD’s break inversion symmetry, thereby giving rise to a very strong Ising spin-orbit coupling (SOC) [15, 9, 17, 18, 20, 21, 19, 26, 14] which pins the electron spins perpendicular to the plane. The most remarkable consequence of this strong SOC is that superconductivity survives at high in-plane magnetic fields even beyond the Pauli critical limit [17, 19, 2, 20, 22, 23, 14, 27, 28, 29, 30, 31].
It was proposed that the presence of an in-plane field can induce a topological transition into a nodal superconducting phase [3, 7] protected by a combination of an effective time reversal and particle-hole symmetry. The nodal superconducting phase is expected to be accompanied by Majorana flat bands [32, 33], indication of which have been reported [34, 35], as well as distinct periodic Josephson current for the transverse momenta in-between the nodal points [36].
In this paper we study the effect of Rashba SOC on the nodal superconducting phase, focusing on the boundary modes and the Josephson current phase relation. Rashba SOC is naturally present due to electronic gates and the presence of a substrate and can be tuned experimentally. The presence of Rashba spin-orbit breaks the chiral symmetry that protects the nodal superconducting phase, and as a result, the nodal points are generally gapped. However, when the in-plane field is aligned along the direction, a lower crystalline symmetry protects the nodal phase [37]. We study the boundary states in the crystalline phase as well as the current-phase relation in a Josephson-junction geometry. Our results indicate that while the vertical mirror symmetry protects exponentially localized states at the boundary transformed by the symmetry, the current phase relation exhibits a trivial periodicity in the presence of Rashba spin-orbit.
The plan of the paper is as follows. We begin in Sec. II with an analysis of the low energy momentum-space Hamiltonian and its related symmetries. In Sec. III we introduce a toy model on a triangular lattice which reduces to the continuum Hamiltonian close to the point. We discuss the topological properties of this model with and without Rashba spin-orbit and study the stability of the boundary modes in a ribbon geometry. The physics of a Josephson junction fabricated out of such a material is discussed in Sec. IV.
II Continuum Model
An Ising superconductor such as monolayer NbSe2 subjected to an in-plane magnetic field of magnitude , with a superconducting pairing is governed by the following Bogoliubov-de-Gennes (BdG) Hamiltonian,
(1) | |||||
where is the kinetic energy term with being the chemical potential. The Ising SOC is unique to this class of materials, and pins the electron spins perpendicular to the plane. The form of is constrained by the crystalline symmetry point group which includes a mirror reflection plane (with normal along the direction), a three-fold rotational symmetry and a vertical mirror (with normal along the direction). The strong Ising SOC protects superconductivity in the presence of an in-plane magnetic field which can exceed the Pauli limit. The parameter denotes the angle the in-plane magnetic field makes with the axis. determines the strength of Rashba SOC, typically present in experimental setups, and can be tuned by gating or by appropriate choice of substrate.
In the absence of Rashba SOC i.e. when , the in-plane direction of is immaterial. When the BdG spectrum has twelve nodal points on the high symmetry lines along which the Ising SOC vanishes. This nodal superconducting phase is accompanied by the presence of Majorana flat bands [3, 7, 32, 33], as well as an energy phase relation that depends on the momentum transverse to the current direction, with a periodicity for the momenta lying between each pair of nodal points [36].
In this work we analyze the effect of Rashba SOC on the nodal superconducting phase, the fate of its boundary modes, and the Josephson current phase relation. To this end, we work in a parameter regime where and there are twelve nodal points in the absence of Rashba SOC.
II.1 Family of 1D Hamiltonians and symmetries
The origin of the nodal points can be understood by analyzing the family of Hamiltonian obtained by setting as a parameter, . In the absence of Rashba SOC, i.e. when , this model has a particle-hole symmetry given by
(2) |
with . While the magnetic field explicitly breaks time-reversal symmetry, the model has an emergent modified time-reversal (TR) symmetry,
(3) |
with which is a combination of time-reversal symmetry and basal plane mirror symmetry . The family of Hamiltonians therefore lies in class BDI of the Altland-Zirnbauer classification [38]. The presence of the nodal points can therefore be understood as a series of topological phase transitions tuned by the parameter as explained in Ref. 36. Next, we introduce a Rashba SOC as given in Eq. (1) which consists of two parts. The first term breaks the modified time-reversal symmetry while the second term breaks particle hole symmetry of the effective model, leaving in class A. However, when the field is oriented along the -axis, i.e. when , the system has a residual vertical mirror symmetry plane, defined by,
(4) |
with . We will show below that this 1D Hamiltonian realizes a crystalline insulating phase associated with gapless edge states which are localized along the direction and propagate along direction.
III Lattice Model
To gain further insight into the topological phase and the nature of its boundary modes, we study a tight-binding model presented in [39, 40] that captures the key features of the topological superconducting phase and exhibits the same low energy physics in the continuum limit .

The lattice model consists of a nearest neighbor hopping:
(5) |
where denotes the spin, spans all the nearest neighbors and is the on-site chemical potential. In the continuum limit, , this reduces to the kinetic energy term of Eq. (1) when we set and . Similarly, Ising SOC is modeled as a nearest-neighbor hopping with alternating signs (shown in Fig. 1), that reflect the symmetry. Note that the sign is opposite for the two spins.
(6) |
where for , respectively, and the lattice vectors are: and .
The in-plane magnetic field
(7) |
Lastly, the Rashba term can be written as follows
(8) |
In momentum space the lattice Hamiltonian, eq. (5)-(8) take the following form
where the kinetic energy term is
(10) |
Here and correspond to the Ising and Rasbha SOCs respectively and have the following form in the lattice model,
and
The strength of each term is suitably chosen to give the same low energy Hamiltonian as Eq. (1).
Figure 2 shows the dispersion of the two lower energy bands of the lattice model given by Eq. (III), in the vicinity of the point. The Rashba SOC breaks the chiral symmetry thus generically lifting the nodal points resulting in a fully gapped phase. However, when the magnetic field is aligned along the direction, the system has a residual vertical mirror symmetry , which protects the nodal points along the symmetry line. The system therefore realizes a nodal crystalline phase, as we show below. Due to the breaking of chiral symmetry, the nodes are shifted away from zero energy.


III.1 Symmetries and topological classification
To gain further insight into the topological aspects of the lattice model and the origin of the nodal points in the spectrum we consider the family of lattice Hamiltonians obtained by treating in eq. (III) as a parameter.
As discussed in Sec. II, Rashba SOC lifts the chiral symmetry thus leaving the family of lattice Hamiltonians in class A. However, when the magnetic field is aligned along the the resulting hamiltonian is symmetric under vertical mirror, , and is gapped except for 4 discrete values of . For values of between these nodal points the system realizes a one-dimensional topological crystalline phase [41, 42].
At , the Hamiltonian is mapped onto itself under reflection. In these reflection symmetric momenta, the energy levels have a well-defined reflection eigenvalue. The reflection eigenvalues of the two occupied levels, labeled as are and corresponding to and , respectively, where is the projector onto the two lowest energy levels at a given momentum . These are continuously connected to the negative energy states in the absence of Rashba SOC. The reflection eigenvalues define a index given by:
(13) |
Figure 3 shows the value of the reflection eigenvalues of the two occupied bands at the two reflection symmetric momenta, , as a function of parameter . Solid and dashed gray lines indicate the spectra along the and lines, respectively. The closing and reopening of the band gap is accompanied by a topological phase transition; i.e. a change in sign of (for the gap closing at ) and (for the gap closing at ).
III.2 Bulk boundary correspondence
To examine the bulk boundary correspondence for the nodal crystalline phase, we study the tight-binding lattice Hamiltonian on a ribbon-like geometry with open boundary conditions in the (non self reflecting) direction, and periodic boundary conditions in the direction. This makes a good quantum number and allows us to write the Hamiltonian in the ribbon geometry as an effective chain for a given :
(14) |
These terms correspond to kinetic energy, Ising SOC, Rashba SOC, in-plane field and superconductivity respectively, a detailed expression is given in Appendix A.
Diagonalizing the Hamiltonian in Eq. (14), we obtain the eigenvalues and the corresponding wavefunctions for the bulk and edge states. Figure 4 (a) shows the BdG spectrum in the Ribbon geometry for . Eigenvalues corresponding to states localized on the open direction boundary are shown in red. The two pairs of nodal points at and are accompanied by the appearance of zero energy states which are localized on the open direction boundary (marked in red).
Figure 4(b) shows the BdG spectra for . When the Rashba term is switched on, only one pair of nodal points survive, which are shifted away from zero energy. In Fig. 4(b) these are located at and . The two degenerate midgap states connecting this pair of nodes (shown in red) live on the (non-self-reflecting) x-boundary and have a non-zero dispersion as a function of (the conjugate momenta for the direction parallel to the boundary). The degeneracy between the boundary mode is protected by the mirror symmetry . In Appendix C we show that any perturbation breaking this reflection symmetry splits these edge states, see Fig. 3.
Figure 5 shows the spatial profile of the degenerate midgap states for a fixed between the pairs of nodes, for different values of Rashba spin-orbit (indicated by the markers). Here indicates the position along the chain. Solid lines indicate a best fit to an exponential law , showing that the states remain exponentially localized even for finite . Importantly, unlike the case, for finite the edge states do not satisfy the Majorana condition . This observation is also consistent with the analytic derivation of the boundary mode in the continuum limit, see Appendix B. The inset in Fig. 5 shows the best fit of the decay length with increasing Rashba SOC.


IV Josephson Junction
To study the Josephson energy-phase relation we close the finite ribbon into a torus-like geometry by adding a weak link between the first and last sites of the effective 1D chain, see Fig. 6. All hopping terms across the weak link acquire a phase and are also attenuated by a factor proportional to the strength of the insulating barrier. Varying the phase difference across the junction for a given allows to obtain the energy phase relation.

Fig. 7 shows for the mid-gap states at a fixed that lie between pair of nodal points. For this value of corresponds to a topological non-trivial phase of class BDI with winding . Conversely, with this value corresponds to a crystalline topological phase with , see Fig. 3. In the absence of Rashba SOC, shown in Fig. 7 (a), the Josephson energy exhibits a periodicity similar to the continuous model studied in Ref. 36, with the energy levels crossing zero at and . This indicates the presence of Majorana edge states localized in the vicinity of the weak link which decay exponentially into the bulk.
When the Rashba SOC is finite, shown in Fig. 7 (b), we find that is no longer symmetric about the line i.e. it is shifted away from zero. Moreover, an energy gap opens at in the presence of a Rashba SOC as is clear from the inset in Fig. 7 (b). Hence, in the presence of Rashba SOC the Josephson energy-phase relation has a periodicity for all values. This is consistent with the observation that the exponentially localized boundary states are not Majorana modes.

V Conclusion
We have studied the effect of Rashba spin-orbit coupling on the nodal superconducting phase of an Ising superconductor. This nodal phase was predicted in monolayer TMD’s such as NbSe2 in the presence of an in-plane field which exceeds the Pauli limit [3, 7]. The presence of Rashba SOC breaks the chiral symmetry and generally lifts the nodal points, resulting in a fully gapped state. However, when the magnetic field is aligned along the line the system has a residual mirror symmetry which protects the nodal points at . The system therefore realizes a nodal crystalline phase, characterized by states exponentially localized at the (non-self-reflecting) x-boundary, which disperse parallel to the boundary, provided that the x-boundary preserves the crystalline symmetry. However, we find that even in the presence of exponentially localized boundary states, the current phase relation in a Josephson junction becomes trivial and follows a periodicity.
We note that Rashba spin-orbit coupling is typically present in experimental setups and can be controlled using gates and by changing substrates. This gives an experimental knob to tune in and out of the topological phase, thus changing the periodic current phase relation to the trivial .
VI Acknowledgements
The authors would like to thank Hadar Steinberg and Ganapathy Murthy for fruitful discussions. D.M. acknowledges support from the Israel Science Foundation (ISF) (grant No. 1884/18). M.K. and D.M. acknowledge support from the ISF, (grant No. 1251/19).
Appendix A Effective Hamiltonian for a chain
We study the tight binding lattice hamiltonian in a ribbon-like geometry with open boundary conditions in the direction and periodic boundary conditions in the direction. Treating as a parameter, the resulting model describes a family of chains along the direction. Below we setting the lattice parameter . The resulting family of 1D chains is described by (14) with the terms corresponding to kinetic energy , Ising SOC and Rashba SOC are all dependent on , and involve terms which couple nearest as well as next nearest-neighbors:
(15) | |||||
(16) | |||||
(17) | |||||
The in-plane magnetic field arises from on-site terms,
(18) |
Note that in all the terms above, we have suppressed the index for the creation (annihilation) operators. However, since the superconducting term couples particle and hole components, it is written as
(19) |
Appendix B Derivation of boundary modes in the continuum limit of the effective 1D model
We consider the continuum limit of the effective 1D Hamiltonian obtained from (1) by treating as a parameter. Following a similar analysis as in Ref. 43, we focus on the regime of strong Ising spin-orbit coupling where the magnetic field, superconductivity, and Rashba SOC can be treated as weak perturbations. We, therefore, consider initially the following bare Hamiltonian:
(20) |
setting aside the gap-opening terms such as magnetic field, SC, and the transverse Rashba term. Here is the momentum of the 1D system and the Pauli matrices operate on the spin basis. For simplicity we consider the for which .
The eigenvalues of the bare 1D Hamiltonian (20) are given by:
(21) |
and the Fermi points that satisfy are located at and . Fig. 8 shows the spectrum of the bare Hamiltonian in the limit . In addition, we have 3 gap-opening perturbations:
(22) | |||||
(23) | |||||
(24) |

where we have introduced the spinor notation . Next, we linearize the spectrum close to the Fermi points, see Fig. 8. The fields then take the form,
(25) | |||
(26) |
where and are slowly varying, and the spin-orbit eigenvectors of the bare Hamiltonian (20) are given by , , with and . Note that for strong Ising SOC the spins are aligned along the direction .
Ignoring strongly oscillatory terms, the kinetic energy can be written as
(28) | |||||
and the gap-opening terms then become,
(29) | |||||
(30) | |||||
(31) |
The Hamiltonian separates into two decoupled subsystems which we label the “external” (e) and “internal” (i) branches and with the respective Hamiltonians:
(32) | |||||
(33) |
In what follows we will drop the subscript for brevity.
We solve for a semi-infinite wire with a boundary at . We make the following ansatz for the zero mode, with where with and two possible values for the inner branch and . Reincorporating the oscillatory phases and expressing the zero mode solutions in terms of the original basis we find,
(38) | |||
(43) |
with and
(48) | |||
(53) |
This allows us to construct a zero mode that satisfies the boundary conditions at namely which is:
(62) | |||||
(67) |
We note that in the absence of Rashba SOC, , and the boundary mode indeed satisfies the Majorana condition namely . However, at finite this condition is no longer met and the boundary state is no longer a Majorana mode.
Appendix C Breaking of reflection symmetry
In the family of 1D chains with open boundary conditions (14), the degeneracy of the mid-gap states is protected by Mirror symmetry , i.e. the edge states are reflected onto each other under this symmetry. Breaking the symmetry by adding a local potential on one of the two edges will lift the degeneracy. This situation is shown in Fig. 9.

Conversely, in the absence of Rashba SOC, i.e. when the degeneracy is protected by the Chiral symmetry of the 1D chain for fixed . Consequently, the degeneracy is not lifted even in the presence of a local chemical potential.
References
- Zhou et al. [2016] B. T. Zhou, N. F. Q. Yuan, H.-L. Jiang, and K. T. Law, Ising superconductivity and majorana fermions in transition-metal dichalcogenides, Physical Review B 93, 180501(R) (2016).
- Ilić et al. [2017] S. Ilić, J. S. Meyer, and M. Houzet, Enhancement of the Upper Critical Field in Disordered Transition Metal Dichalcogenide Monolayers, Physical Review Letters 119, 117001 (2017).
- He et al. [2018] W.-Y. He, B. T. Zhou, J. J. He, N. i. Q. Yuan, T. Zhang, and K. T. Law, Magnetic field driven nodal topological superconductivity in monolayer transition metal dichalcogenides, Communications Physics 1, 40 (2018).
- Sosenko et al. [2017] E. Sosenko, J. Zhang, and V. Aji, Unconventional superconductivity and anomalous response in hole-doped transition metal dichalcogenides, Phys. Rev. B 95, 144508 (2017).
- Nakamura and Yanase [2017] Y. Nakamura and Y. Yanase, Odd-parity superconductivity in bilayer transition metal dichalcogenides, Phys. Rev. B 96, 054501 (2017).
- Möckli and Khodas [2018] D. Möckli and M. Khodas, Robust parity-mixed superconductivity in disordered monolayer transition metal dichalcogenides, Physical Review B 98, 144518 (2018).
- Fischer et al. [2018] M. H. Fischer, M. Sigrist, and D. F. Agterberg, Superconductivity without inversion and time-reversal symmetries, Phys. Rev. Lett. 121, 157003 (2018).
- Smidman et al. [2017a] M. Smidman, M. B. Salamon, H. Q. Yuan, and D. F. Agterberg, Superconductivity and spin-orbit coupling in non-centrosymmetric materials: a review, Reports on Progress in Physics 80, 036501 (2017a).
- Yuan et al. [2014] N. F. Q. Yuan, K. F. Mak, and K. T. Law, Possible topological superconducting phases of , Phys. Rev. Lett. 113, 097001 (2014).
- Oiwa et al. [2018] R. Oiwa, Y. Yanagi, and H. Kusunose, Theory of superconductivity in hole-doped monolayer , Phys. Rev. B 98, 064509 (2018).
- Hsu et al. [2017] Y.-T. Hsu, A. Vaezi, M. H. Fischer, and E.-A. Kim, Topological superconductivity in monolayer transition metal dichalcogenides, Nature Communications 8, 14985 (2017).
- Wang et al. [2018] L. Wang, T. O. Rosdahl, and D. Sticlet, Platform for nodal topological superconductors in monolayer molybdenum dichalcogenides, Phys. Rev. B 98, 205411 (2018).
- Oiwa et al. [2019] R. Oiwa, Y. Yanagi, and H. Kusunose, Time-reversal symmetry breaking superconductivity in hole-doped monolayer mos2, Journal of the Physical Society of Japan 88, 063703 (2019), https://doi.org/10.7566/JPSJ.88.063703 .
- Sohn et al. [2018] E. Sohn, X. Xi, W.-Y. He, S. Jiang, Z. Wang, K. Kang, J.-H. Park, H. Berger, L. Forró, K. T. Law, J. Shan, and K. F. Mak, An unusual continuous paramagnetic-limited superconducting phase transition in 2D NbSe 2, Nature Materials 17, 504 (2018).
- Wang et al. [2012] Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, Electronics and optoelectronics of two-dimensional transition metal dichalcogenides, Nature Nanotechnology 7, 699 (2012).
- Geim and Grigorieva [2013] A. K. Geim and I. V. Grigorieva, Van der Waals heterostructures, Nature 499, 419 (2013).
- Lu et al. [2015] J. M. Lu, O. Zheliuk, I. Leermakers, N. F. Q. Yuan, U. Zeitler, K. T. Law, and J. T. Ye, Evidence for two-dimensional Ising superconductivity in gated MoS2, Science 350, 1353 (2015).
- Ugeda et al. [2016] M. M. Ugeda, A. J. Bradley, Y. Zhang, S. Onishi, Y. Chen, W. Ruan, C. Ojeda-Aristizabal, H. Ryu, M. T. Edmonds, H.-Z. Tsai, A. Riss, S.-K. Mo, D. Lee, A. Zettl, Z. Hussain, Z.-X. Shen, and M. F. Crommie, Characterization of collective ground states in single-layer NbSe2, Nature Physics 12, 92 (2016).
- Saito et al. [2016] Y. Saito, Y. Nakamura, M. S. Bahramy, Y. Kohama, J. Ye, Y. Kasahara, Y. Nakagawa, M. Onga, M. Tokunaga, T. Nojima, Y. Yanase, and Y. Iwasa, Superconductivity protected by spin-valley locking in ion-gated mos2, Nature Physics 12, 144 (2016).
- Xi et al. [2016] X. Xi, Z. Wang, W. Zhao, J.-H. Park, K. T. Law, H. Berger, L. Forró, J. Shan, and K. F. Mak, Ising pairing in superconducting nbse2 atomic layers, Nature Physics 12, 139 (2016).
- Costanzo et al. [2016] D. Costanzo, S. Jo, H. Berger, and A. F. Morpurgo, Gate-induced superconductivity in atomically thin mos2 crystals, Nature Nanotechnology 11, 339 (2016).
- Dvir et al. [2018] T. Dvir, F. Massee, L. Attias, M. Khodas, M. Aprili, C. H. L. Quay, and H. Steinberg, Spectroscopy of bulk and few-layer superconducting NbSe2 with van der Waals tunnel junctions, Nature Communications 9, 598 (2018).
- de la Barrera et al. [2018] S. C. de la Barrera, M. R. Sinko, D. P. Gopalan, N. Sivadas, K. L. Seyler, K. Watanabe, T. Taniguchi, A. W. Tsen, X. Xu, D. Xiao, and B. M. Hunt, Tuning ising superconductivity with layer and spin-orbit coupling in two-dimensional transition-metal dichalcogenides, Nature Communications 9, 1427 (2018).
- Xing et al. [2017] Y. Xing, K. Zhao, P. Shan, F. Zheng, Y. Zhang, H. Fu, Y. Liu, M. Tian, C. Xi, H. Liu, J. Feng, X. Lin, S. Ji, X. Chen, Q.-K. Xue, and J. Wang, Ising superconductivity and quantum phase transition in macro-size monolayer nbse2, Nano Letters 17, 6802 (2017).
- Hamill et al. [2021] A. Hamill, B. Heischmidt, E. Sohn, D. Shaffer, K.-T. Tsai, X. Zhang, X. Xi, A. Suslov, H. Berger, L. Forró, F. J. Burnell, J. Shan, K. F. Mak, R. M. Fernandes, K. Wang, and V. S. Pribiag, Two-fold symmetric superconductivity in few-layer nbse2, Nature Physics 10.1038/s41567-021-01219-x (2021).
- Yuan et al. [2016] N. F. Q. Yuan, B. T. Zhou, W.-Y. He, and K. T. Law, Ising superconductivity in transition metal dichalcogenides (2016), arXiv:1605.01847 [cond-mat.supr-con] .
- Liu et al. [2018] Y. Liu, Z. Wang, X. Zhang, C. Liu, Y. Liu, Z. Zhou, J. Wang, Q. Wang, Y. Liu, C. Xi, M. Tian, H. Liu, J. Feng, X. C. Xie, and J. Wang, Interface-Induced Zeeman-Protected Superconductivity in Ultrathin Crystalline Lead Films, Physical Review X 8, 021002 (2018).
- woo Cho et al. [2020] C. woo Cho, J. Lyu, T. Han, C. Y. Ng, Y. Gao, G. Li, M. Huang, N. Wang, J. Schmalian, and R. Lortz, Distinct nodal and nematic superconducting phases in the 2d ising superconductor nbse2 (2020), arXiv:2003.12467 [cond-mat.supr-con] .
- Cho et al. [2022] C.-w. Cho, J. Lyu, L. An, T. Han, K. T. Lo, C. Y. Ng, J. Hu, Y. Gao, G. Li, M. Huang, N. Wang, J. Schmalian, and R. Lortz, Nodal and nematic superconducting phases in monolayers from competing superconducting channels, Phys. Rev. Lett. 129, 087002 (2022).
- Frigeri et al. [2006] P. A. Frigeri, D. F. Agterberg, I. Milat, and M. Sigrist, Phenomenological theory of the s-wave state in superconductors without an inversion center, The European Physical Journal B 54, 435 (2006).
- Kuzmanović et al. [2022] M. Kuzmanović, T. Dvir, D. LeBoeuf, S. Ilić, M. Haim, D. Möckli, S. Kramer, M. Khodas, M. Houzet, J. S. Meyer, M. Aprili, H. Steinberg, and C. H. L. Quay, Tunneling spectroscopy of few-monolayer in high magnetic fields: Triplet superconductivity and ising protection, Phys. Rev. B 106, 184514 (2022).
- Zhao and Wang [2013] Y. X. Zhao and Z. D. Wang, Topological classification and stability of fermi surfaces, Phys. Rev. Lett. 110, 240404 (2013).
- Matsuura et al. [2013] S. Matsuura, P.-Y. Chang, A. P. Schnyder, and S. Ryu, Protected boundary states in gapless topological phases, New Journal of Physics 15, 065001 (2013).
- Galvis et al. [2014] J. A. Galvis, L. Chirolli, I. Guillamón, S. Vieira, E. Navarro-Moratalla, E. Coronado, H. Suderow, and F. Guinea, Zero-bias conductance peak in detached flakes of superconducting 2- probed by scanning tunneling spectroscopy, Phys. Rev. B 89, 224512 (2014).
- Nayak et al. [2021] A. K. Nayak, A. Steinbok, Y. Roet, J. Koo, G. Margalit, I. Feldman, A. Almoalem, A. Kanigel, G. A. Fiete, B. Yan, Y. Oreg, N. Avraham, and H. Beidenkopf, Evidence of topological boundary modes with topological nodal-point superconductivity, Nature Physics 17, 1413 (2021).
- Seshadri et al. [2022] R. Seshadri, M. Khodas, and D. Meidan, Josephson junctions of topological nodal superconductors, SciPost Phys. 12, 197 (2022).
- Shaffer et al. [2020] D. Shaffer, J. Kang, F. J. Burnell, and R. M. Fernandes, Crystalline nodal topological superconductivity and bogolyubov fermi surfaces in monolayer , Phys. Rev. B 101, 224503 (2020).
- Altland and Zirnbauer [1997] A. Altland and M. R. Zirnbauer, Nonstandard symmetry classes in mesoscopic normal-superconducting hybrid structures, Phys. Rev. B 55, 1142 (1997).
- Haim et al. [2022] M. Haim, A. Levchenko, and M. Khodas, Mechanisms of in-plane magnetic anisotropy in superconducting , Phys. Rev. B 105, 024515 (2022).
- Smidman et al. [2017b] M. Smidman, M. Salamon, H. Yuan, and D. Agterberg, Superconductivity and spin–orbit coupling in non-centrosymmetric materials: a review, Reports on Progress in Physics 80, 036501 (2017b).
- Hughes et al. [2011] T. L. Hughes, E. Prodan, and B. A. Bernevig, Inversion-symmetric topological insulators, Phys. Rev. B 83, 245132 (2011).
- Chiu et al. [2013] C.-K. Chiu, H. Yao, and S. Ryu, Classification of topological insulators and superconductors in the presence of reflection symmetry, Phys. Rev. B 88, 075142 (2013).
- Laubscher and Klinovaja [2021] K. Laubscher and J. Klinovaja, Majorana bound states in semiconducting nanostructures, Journal of Applied Physics 130, 081101 (2021), https://doi.org/10.1063/5.0055997 .