Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2021 Jan 11;118(3):e2019063118. doi: 10.1073/pnas.2019063118

Topological metals and finite-momentum superconductors

Noah F Q Yuan a,b,1,2, Liang Fu b,1,2
PMCID: PMC7826329  PMID: 33431686

Significance

Our work uncovers an interesting connection between topological band theory and superconductivity. While BCS superconductivity at zero magnetic field is largely insensitive to electron wavefunction, in this work we reveal a rich phase diagram of finite-momentum superconductivity in the presence of spin–orbit splitting and Zeeman fields, where electron pairing depends closely on the topology of spin texture on the Fermi surface.

Keywords: topology, spin–orbit coupling, superconductivity

Abstract

We show that the Zeeman field can induce a topological transition in two-dimensional spin–orbit-coupled metals and, concomitantly, a first-order phase transition in the superconducting state involving a discontinuous change of Cooper pair momentum. Depending on the spin–orbit coupling strength, we find different phase diagrams of two-dimensional (2D) superconductors under in-plane magnetic field.


A fundamental concept in the theory of metals is Fermi surface, the locus of gapless electronic Bloch states in momentum space. While Bloch states in conventional metals are spin degenerate, the degeneracy is lifted by a magnetic field via a Zeeman effect and in noncentrosymmetric crystals by spin–orbit coupling (SOC). In both scenarios, the Fermi surface becomes spin split. When attractive interaction is present at low energy, pairing instability of the Fermi surface turns a metal into a superconductor. The interplay between spin–orbit and Zeeman splitting has interesting consequences for superconductivity, as shown in many previous works (118). Recent discovery of superconductivity in a variety of two-dimensional spin–orbit-coupled materials, including transition metal dichalcogenides (19) and strontium titanate films (20), adds additional venues for further investigation of this important problem.

In this paper, we take a fresh look at spin–orbit-coupled metals and superconductors through the lens of wavefunction topology. We characterize the topology of electron wavefunctions on spin-split Fermi surfaces and establish a correspondence between topological metals in two dimensions and topological crystalline insulators in three dimensions. Applying an in-plane magnetic field to spin–orbit-coupled two-dimensional (2D) metals can induce a topological phase transition, characterized by a change of spin texture on the Fermi surface and π phase shift in quantum oscillation. When the metal becomes superconducting at low temperature, the field-induced topological transition of Fermi surface is found to strongly impact electron pairing. We present phase diagrams of 2D superconductors under in-plane magnetic fields for various SOC strengths.

This work is organized as follows. We start with a case study of 2D Rashba systems under an in-plane magnetic field, which induces a topology change of spin texture on the Fermi surface. We then define a general set of topological invariants for 2D metals having any space–time parity symmetry in terms of quantized π Berry phase on spin-nondegenerate Fermi surface. Finally, we examine the impact of Fermi surface spin splitting on superconductivity and show that the field-induced topological transition of the Fermi surface can cause a change in pairing from intrapocket to interpocket, leading to a first-order phase transition in the finite-momentum superconducting state at finite field. Importantly, the intrapocket pairing state evolves smoothly from the zero-momentum Bardeen–Cooper–Schrieffer (BCS) state at zero field, while the interpocket pairing state evolves smoothly from the finite-momentum Fulde–Ferrell–Larkin–Ovchinnikov (FFLO) state at field beyond the Pauli limit. By combining microscopic calculation, symmetry analysis, Ginzburg–Landau theory, and physical argument, we obtain a global phase diagram of spin–orbit-coupled superconductors under Zeeman field.

Rashba Systems

We consider a single-component 2D electron gas with SOC under magnetic field

H(k)=k22mμ+g(k)σ+σB, [1]

where k=(kx,ky) is the 2D momentum, m is effective mass, μ is chemical potential, g(k) is the SOC vector, B is the Zeeman energy due to the in-plane magnetic field, and Pauli matrices σ=(σx,σy,σz) denote spin.

As a concrete example, we consider Rashba SOC g=αRk×z^ with Rashba coefficient αR. At B=0, in the energy eigenstate the electron’s spin is tied with its momentum due to Rashba SOC. As a result, two concentric Fermi circles are present, with helical spin textures of the same chirality. The shape and spin configuration of both Fermi surfaces evolve with the in-plane magnetic field. As B increases, the two pockets approach each other and deform into ovals known as Cartesian ovals (Fig. 1A). The electron’s spin lies within the xy plane and winds by 2π around both inner and outer Fermi surfaces. At a critical field B=αRkFΔso with kF=2mμ, the two ovals touch each other at a point kPkFz^×B^, where a twofold spin degeneracy arises. The resulting Fermi surface is a single self-intersecting curve known as a limaçon of Pascal (Fig. 1B). As the field increases further, these two ovals disconnect again and move away from each other (Fig. 1C). Now, the spin winding number on each Fermi surface is 0, and the spin configuration resembles more the Zeeman-dominated case. The inner Fermi surface shrinks further and eventually disappears at sufficiently high fields.

Fig. 1.

Fig. 1.

(A–C) Fermi surfaces in normal phase. Yellow and blue colors denote inner and outer Fermi surfaces, respectively, and arrows denote electron spins. Shaded small disks in A and C denote paired electrons in corresponding superconducting phases. (D–F) Bogouliubov Fermi segments in corresponding superconducting phase, represented by maxima of zero-energy electronic density of states (DOS) ρ=1πIm[tr(Gτe)], where G is Gor’kov Green’s function, and τe=diag(1,0) acts in the particle-hole space. We choose parameters m=1,μ=10,αR=1,Δ=1, broadening η=0.01, and magnetic field (A and D) B=3, (B and E) B=Δso=20, and (C and F) B=7.

The merging of two pockets and the change of spin winding number at B=Δso mark another type of Fermi surface topological transition. It is fundamentally different from Lifshitz transitions that reconnect Fermi contours through a saddle point in the energy dispersion. Here instead, the original Dirac point at k=0 moves to kP at the Fermi level and results in Fermi surface touching at B=Δso. The dispersion around kP takes the form of an overtitled and anisotropic Dirac cone,

H(kP+p)=vFpx+αR(pxσypyσx), [2]

where vF=2μ/m is the Fermi velocity. Moreover, the topology of Fermi surface in momentum space is unchanged before and after the transition, and the DOS remains finite throughout, unlike the van Hove singularity resulting from saddle points.

Topology of Metals

Unlike Lifshitz transition associated with Fermi surface geometry, what we uncovered in Rashba systems involves the topology of quantum wavefunction on a spin nondegenerate Fermi surface. To characterize the wavefunction topology, we introduce a general topological invariant for 2D systems having any parity symmetry, i.e., invariant under any transformation that reverses the orientation of the space–time manifold, including time-reversal T, reflection M (xx), and the combined operation of twofold rotation (xx,yy) and time-reversal C2T. We define the Berry phase φ=Ek=μAkdk as the topological invariant of a Fermi surface, where Ak=iψkkψk is the Berry connection and ψk is the wavefunction with energy Ek. One can transform the Berry phase φ into an integral of Berry curvature Ωk=(k×Ak)z on the k-space section enclosed by the Fermi contour; namely, φ=Ek<μΩkd2k. Under the parity symmetry, we find energy dispersion is even Ek=EPk while Berry curvature is odd Ωk=ΩPk, where P can be T,M or C2T. As a result, φ=φ (mod 2π), and the Berry phase would be either zero or quantized to π, if inside the enclosed section there are even or odd numbers of Dirac points, respectively. At a Dirac point, two bands cross and the Berry curvature becomes singular. Due to the parity symmetry, Dirac points are restricted to be at time-reversal-invariant points (P=T) or on the reflection-invariant lines (P=M) or anywhere in the momentum space (P=C2T).

This quantized Berry phase therefore serves as a Z2 topological invariant. Generally speaking, Zeeman splitting of Fermi surface results in 0 Berry phase, while spin–orbit splitting results in π Berry phase on each Fermi surface enclosing a time-reversal-invariant momentum. In the 2D Rashba model, the Z2 topological distinction continues to hold in the presence of a Zeeman field, which preserves reflection and C2T. We can also include higher-order SOC such as hexagonal warping (21) to break C2T symmetry while preserving reflection M(xx), so that the SOC vector reads g=αRk×z^+w(k+3+k3)z^, where k±=kx±iky. In this case, the Dirac point is on the kx=0 line as restricted by M, and only magnetic field Bx can drive a topological transition of Fermi surfaces.

We refer to Fermi surfaces having quantized Berry phase π and 0 as topological and trivial, respectively. It is interesting to note that Fermi surfaces with quantized π Berry phase are the hallmark of surface states in 3D topological (crystalline) insulators protected by parity symmetry—T, M, or C2T (2226). While these topological surface states are known as “half” of 2D metals, we now turn this viewpoint the other way. The 2D topological metals, defined as having spin-nondegenerate Fermi surfaces with π Berry phase, can be viewed as a “sum” of topological surfaces. Thus follows a correspondence between 2D topological metals and 3D dimensional topological (crystalline) insulators.

Since the sum of Berry phases over all Fermi surfaces must be zero in any 2D metal with parity symmetry, topological metals have an even number of spin-nondegenerate Fermi surfaces with π Berry phase. A transition from a topological metal to a trivial one generally involves the touching of two Fermi surfaces at a band degeneracy point to enable the change of Berry phase from π to 0 on each Fermi surface. We thus conclude that an in-plane magnetic field generally induces a topological phase transition in spin–orbit-coupled 2D metals, provided that parity symmetry (e.g., C2T or M) is present.

The presence of a spin–orbit-split Fermi surface leads to beatings in quantum oscillation phenomena, as observed in semiconductor heterostructures and noncentrosymmetric metals. The Berry phase change from π to 0 across the topological transition can be further detected by analyzing the phase shift of quantum oscillation as a function of in-plane magnetic field (27, 28). We also note recent works on topological characterization of 3D metals under Zeeman fields using Chern numbers on Fermi surfaces (27, 29).

When attractive interaction is present, metals with spin-split Fermi surfaces may become unstable to pairing at low temperature. The competition between spin–orbit and Zeeman splitting, which drives the topological transition in the normal state, also significantly impacts the superconducting state, which we now turn to.

Finite-Momentum Superconductivity

We consider clean superconductors with a local attraction and an energy gap that is small compared to Fermi energy. In this case, superconductivity at zero field is the conventional BCS type. Increasing temperature T and/or in-plane magnetic field B drives superconducting to normal transition. The superconducting order parameter near the transition is determined by the pair susceptibility (9, 10, 17)

χ(q,B,T)=N0logωDTs=±δχs(q,B,T), [3]

where q is Cooper pair momentum. The first term is the q=0 pair susceptibility in BCS theory, where N0 is the total DOS of Fermi surfaces and ωD is the Debye frequency. The second term is the correction due to Fermi surface spin splitting and finite Cooper pair momentum:

δχs=FSsdk|v|ϕQ+sE+2πTcos2θ2+ϕQ+sE2πTsin2θ2 [4]

is contribution from inner (s=+) or outer (s=) Fermi surface, ϕ(x)=Reψ1+ix2ψ12, ψ is the digamma function, v is electron velocity, Q=vq is the depairing energy of finite-momentum pairing, E±=|h+|±|h| is the Zeeman depairing energy of inter- (+) or intrapocket () Cooper pairs, θ=h+,h is the angle between h±=B+g(12q±k), and g(k) is the SOC vector.

In the presence of Rashba SOC, the inner (s=+) and outer (s=) Fermi surfaces at B=0 have different DOS given by Ns=12N0(112sλ) with λ=Δso/μ, respectively. As shown in refs. 1 and 2, when ΔsoΔ0, this DOS asymmetry (often neglected) is important in determining Cooper pair momentum and critical fields at low temperatures. In this work, we always take into account DOS asymmetry N+N when SOC is present.

From pair susceptibility, the in-plane critical field Bc(T) is given by vχmax=1, where v is the attractive interaction strength in s-wave channel and χmax is the maximum of χ(q) among all q. In this way we also determine the Cooper pair momentum q near Bc.

Depending on the magnetic field B and Rashba coupling energy Δso in comparison to the BCS gap Δ0=2ωDe1/N0v, we find three phases at T=0 near the upper critical field: the normal phase (N) and finite-momentum superconducting phases whose Cooper pairs are dominantly interpocket (I) and intrapocket (II), respectively. In the limit of vanishing SOC Δso=0, phase I reduces to the well-known FFLO state, which occurs at magnetic field above the Pauli limit BPΔ0/2 (30, 31). Here the Zeeman splitting of the Fermi surface favors interpocket pairing between majority and minority spin states. In the opposite limit of large SOC ΔsoΔ0, the helical spin texture of the spin–orbit-split Fermi surface favors intrapocket pairing (II). This phase can survive magnetic fields much larger than the Pauli limit and is destroyed only when Zeeman energy becomes comparable to SOC and distorts the Fermi surface significantly (1, 2).

By calculating χ(q,B,T=0), we obtain the phase diagram shown in Fig. 2A. Phases I and II are both finite-momentum superconductors indistinguishable by symmetry. They are separated by a first-order quantum phase transition, where the Cooper pair momentum changes abruptly mainly due to the change of pairing from interpocket to intrapocket. As shown in Fig. 1A, the topological metal (B<Δso) has helical spin configuration of Fermi surfaces and hence electrons from the same Fermi surface can have opposite spins, which results in the intrapocket pairing (II) of Fig. 1D. On the contrary, the trivial metal (B>Δso) as shown in Fig. 1C has Zeeman spin configuration where two Fermi surfaces are spin polarized along opposite directions, and hence electrons from different Fermi surfaces can have opposite spins, leading to interpocket pairing (I) in Fig. 1F. Here our argument is based on Fermi surface spin configuration controlled by magnetic field. In the superconducting phase, the Cooper pair momentum can also alter Fermi surface spin configuration. Quantitative calculations taking into account both effects show that the exact phase boundary between I and II at T=0 is at Δso/B0.7. The first-order phase boundary between I and II ends at a tricritical point (Δso*,B*) with Δso*0.7Δ0,B*Δ0, where three phases (I, II, N) meet. Remarkably, near (Δso*,B*) the phase boundary between superconducting phases I and II is a straight line Δso/B0.7 and closely follows the one between topological and trivial metal in the absence of superconductivity. This finding demonstrates the direct impact of topology on finite-momentum pairing in spin–orbit-coupled metals.

Fig. 2.

Fig. 2.

(A) Phase diagram in the Δso-B plane at zero temperature, where N is the normal phase and I and II denote superconducting phases whose Cooper pairs are mainly inter- and intrapocket, respectively. Black (red) lines denote second- (first-)order phase transitions. (B) Pair susceptibility at the first-order (red) and second-order (black) phase transitions when Δso=0.67Δ0, where q0=Δ0/vF. Dots are from numerical calculations and lines are from interpolation of dots. We set μ=50Tc to include DOS asymmetry.

We further consider low-energy Bogouliubov quasiparticles of the superconducting phase under a magnetic field. As the field strength increases, the energy gap closes and zero-energy quasiparticles form a Fermi surface that is marked different from the normal-state Fermi surface. At B<Δso (Fig. 1D), two Bogouliubov–Fermi segments are formed by the inner pocket, while the outer pocket is fully gapped. At B>Δso (Fig. 1F), four Bogouliubov–Fermi segments are formed by both inner and outer Fermi surfaces. Such Bogouliubov–Fermi segments as shown in Fig. 1 DF can be measured experimentally via scanning tunneling microscope spectroscopy (32) or quasiparticle interference. In terms of topology, we find Berry phases along zero-energy contours of Bogouliubov quasiparticles are zero, and Bogouliubov–Fermi surfaces are trivial in our model. As a result, within the superconducting phase, topological transitions cannot be realized, and there are either first-order phase transitions (Δso<Δso*) or no phase transitions (Δso>Δso*), as shown in the phase diagram Fig. 2A.

Our results on finite-momentum superconductivity at T=0 have important implications for the global phase diagram as a function of temperature and magnetic field, which is plotted in Fig. 3 for different SOC strengths. Without SOC, phases I (FFLO) and II (BCS) are separated by a first-order phase transition line in the T-B plane, which starts at (T=0,B=BP) and ends at a finite-temperature tricritical point (T*=0.56Tc,B*=0.6Δ0) where phases I, II, and N meet (33, 34) (Fig. 3A).

Fig. 3.

Fig. 3.

Phase diagram in the T-B plane with different SOC strengths. Black (red) lines denote second- (first-)order phase transitions. We set (B) μ=10Tc and (C and D) μ=50Tc to include DOS asymmetry, which drives the tricritical point (T*,B*) in A to a critical point (T*,B*) in B.

In the presence of SOC, a small field BΔso displaces the centers of inner and outer Fermi pockets to opposite momenta ±k0±B. Then, pairing within the inner (outer) Fermi pocket would lead to Cooper pair momentum ±q±2k0, respectively. Importantly, due to the difference in DOS on the two pockets, the pairing susceptibilities χ(q) and χ(q) are generally unequal. The larger of the two sets the Cooper pair momentum near the superconducting transition temperature. This argument shows that the Cooper pair momentum q is linearly proportional to B in the weak-field regime. The resulting finite-momentum superconductor is characterized by intrapocket pairing and evolves smoothly out of the BCS state at B=0 and hence is different from the conventional FFLO phase at B>BP.

The linear coupling between Cooper pair momentum and in-plane magnetic field can be also deduced at a formal level from the Ginzburg–Landau free energy in terms of the real-space order parameter ψ(r), F=drf with

f=ψ*αψ+β|ψ|4. [5]

The coefficient of the quadratic term α can be expanded in powers of the wavevector q. Up to fourth order, it takes the following form dictated by symmetry,

α=a0+a1q2+a2q4(b0+b1q2)qΛB, [6]

where ΛB is an odd-in-B vector. Here, in addition to even-order terms a0,a1,a2, odd terms b0,b1 may be allowed in spin–orbit-coupled systems, which are invariant under joint rotation of Cooper pair momentum and the Zeeman field. At weak field, ΛBB. By minimizing α, we find the induced Cooper pair momentum q near normal-superconducting transition is proportional to B:

q=b02a1ΛBB. [7]

Since the new terms are odd under inversion I:qq,BB or reflection Mz:qq,BB, they exist in systems with broken I and Mz or equivalently in the following 15 point groups: Dn,Cnv,Cn,D2d,S4,C1(n=2,3,4,6). The direction of ΛB and hence the induced Cooper pair momentum depend on the point group symmetry. Fig. 4 shows the direction of q for point groups Dn,Cnv, and D2d, respectively. For 2D Rashba systems, the induced Cooper pair momentum at weak field can be obtained by calculating a1 and b0 using BCS theory: q=2αRB×z^/vF2 (4). Note that the DOS asymmetry must be included to obtain a nonzero q.

Fig. 4.

Fig. 4.

Cooper pair momentum q is determined by magnetic field B under different point groups. In D2d, directions of q,B form a mirror pair with respect to the mirror plane denoted by a dashed line.

By numerically calculating the susceptibility χ(q) as a function of B, we locate the normal-superconducting phase boundary, i.e., the upper critical field curve Bc(T). Without SOC (b0=b1=0), the Bc(T) curve is divided into two parts by a tricritical point (T*,B*), which corresponds to a0=a1=0 and a2>0. The Cooper pair momentum at the onset of superconductivity changes from q=0 at B<B* to q=|a1|/2a2 at B>B*.

When SOC is present, due to b0,b10 the Cooper pair momentum is already nonzero at weak field. Our microscopic calculation shows that for both small and large Rashba couplings, the Cooper pair momentum at Tc changes smoothly with the field. In other words, there is no tricritical point on the Bc(T) curve. On the other hand, a perturbatively small SOC strength cannot eliminate the strong first-order transition between BCS and FFLO states at low temperature. Therefore, for small SOC strength, we expect the phase diagram shown in Fig. 3B. Since at B0 the superconducting phases I and II both have finite-momentum Cooper pairs and share the same symmetry, the first-order transition between them starts at (T=0,BBP) (Fig. 2A) and ends at a critical point (T*,B*), which is located inside the superconducting phase and away from the Bc(T) curve (Fig. 3B).

As SOC strength increases, the critical point (T*,B*) moves to a higher field and lower temperature. At certain SOC strength Δso=Δso*, a zero-temperature tricritical point (0,B*) arises (Fig. 3C), a direct result of the topological transition of the normal-state Fermi surface. Near the zero-temperature tricritical point (0,B*), Eq. 6 also applies. In a small range of SOC strength Δso*ΔsoΔso**1.1Δso*, only phase II exists at zero temperature, while a short first-order line between phases I and II remains at finite temperature. Finally, for Δso>Δso**, the entire superconducting region is phase II with interpocket pairing (19) as shown in Fig. 3D. Putting all these results together, we arrive at a global phase diagram of 2D superconductors under an in-plane magnetic field, for different SOC strengths.

Conclusion

To conclude, we show that the interplay between SOC and Zeeman effect leads to additional normal and superconducting phases in metals. The Fermi surface transition in normal phase is of topological origin and drives a first-order phase transition within the finite-momentum superconducting state. Depending on the SOC strength, we find different phase diagrams of 2D superconductors under an in-plane magnetic field.

Acknowledgments

We thank Joe Checkelsky, Aravind Devarakonda, and Susan Stemmer for stimulating discussions. This work is supported by the Department of Energy Office of Basic Energy Sciences, Division of Materials Sciences and Engineering under Award DE-SC0010526. L.F. is partly supported by a Simons Investigator award from the Simons Foundation.

Footnotes

The authors declare no competing interest.

This article is a PNAS Direct Submission.

Data Availability.

All study data are included in this article.

References

  • 1.Barzykin V., Gor’kov L. P., Inhomogeneous stripe phase revisited for surface superconductivity. Phys. Rev. Lett. 89, 227002 (2002). [DOI] [PubMed] [Google Scholar]
  • 2.Michaeli K., Potter A. C., Lee P. A., Superconducting and ferromagnetic phases in SrTiO3/LaAlO3 oxide interface structures: Possibility of finite momentum pairing. Phys. Rev. Lett. 108, 117003 (2012). [DOI] [PubMed] [Google Scholar]
  • 3.Samokhin K. V., Paramagnetic properties of noncentrosymmetric superconductors: Application to CePt3Si. Phys. Rev. Lett. 94, 027004 (2005). [DOI] [PubMed] [Google Scholar]
  • 4.Dimitrova O., Feigel’man M. V., Theory of a two-dimensional superconductor with broken inversion symmetry. Phys. Rev. B 76, 014522 (2007). [Google Scholar]
  • 5.Samokhin K. V., Upper critical field in noncentrosymmetric superconductors. Phys. Rev. B 78, 224520 (2008). [Google Scholar]
  • 6.Aoyama K., Sigrist M., Model for magnetic flux patterns induced by the influence of in-plane magnetic fields on spatially inhomogeneous superconducting interfaces of LaAlO3-SrTiO3 bilayers. Phys. Rev. Lett. 109, 237007 (2012). [DOI] [PubMed] [Google Scholar]
  • 7.Agterberg D. F., Novel magnetic field effects in unconventional superconductors. Physica C 387, 13–16 (2003). [Google Scholar]
  • 8.Kaur R. P., Agterberg D. F., Sigrist M., Helical vortex phase in the noncentrosymmetric CePt 3 Si. Phys. Rev. Lett. 94, 137002 (2005). [DOI] [PubMed] [Google Scholar]
  • 9.Zwicknagl G., Jahns S., Fulde P., Critical magnetic field of ultra-thin superconducting films and interfaces. J. Phys. Soc. Jpn. 86, 083701 (2017). [Google Scholar]
  • 10.Lu M., et al. , Evidence for two-dimensional Ising superconductivity in gated MoS2. Science 350, 1353–1357 (2015). [DOI] [PubMed] [Google Scholar]
  • 11.Saito Y., et al. , Superconductivity protected by spin-valley locking in ion-gated MoS2. Nat. Phys. 12, 144–149 (2016). [Google Scholar]
  • 12.Xi X., et al. , Evidence of Ising pairing in superconducting NbSe2 atomic layers. Nat. Phys. 12, 139 (2016). [Google Scholar]
  • 13.Zhou B. T., Noah F., Yuan Q., Jiang H.-L., Law K. T., Ising superconductivity and Majorana fermions in transition-metal dichalcogenides. Phys. Rev. B 93, 180501 (2016). [Google Scholar]
  • 14.Zhou B. T., et al. , Magnetic field driven nodal topological superconductivity in monolayer transition metal dichalcogenides. Commun. Phys. 1, 40 (2018). [Google Scholar]
  • 15.Bulaevskii L. N., Guseinov A. A., Rusinov A. I., Superconductivity in crystals without symmetry centers. Sov. Phys. JETP 44, 1243 (1976). [Google Scholar]
  • 16.Gor’kov L. P., Rashba E. I., Superconducting 2D system with lifted spin degeneracy: Mixed singlet-triplet state. Phys. Rev. Lett. 87, 037004 (2001). [DOI] [PubMed] [Google Scholar]
  • 17.Frigeri P. A., Agterberg D. F., Koga A., Sigrist M., Superconductivity without inversion symmetry: MnSi versus CePt3Si. Phys. Rev. Lett. 92, 097001 (2004). [DOI] [PubMed] [Google Scholar]
  • 18.Frigeri P. A., Agterberg D. F., Sigrist M., Spin susceptibility in superconductors without inversion symmetry. New J. Phys. 6, 115 (2004). [Google Scholar]
  • 19.Devarakonda A., et al. , Clean 2D superconductivity in a bulk van der Waals superlattice, Science 370, 231–236 (2020). [DOI] [PubMed] [Google Scholar]
  • 20.Schumann T., et al. , Possible signatures of mixed-parity superconductivity in doped polar SrTi O 3 films. Phys. Rev. B 101, 100503 (2020). [Google Scholar]
  • 21.Fu L., Hexagonal warping effects in the surface states of the topological insulator Bi2Te3. Phys. Rev. Lett. 103, 266801 (2009). [DOI] [PubMed] [Google Scholar]
  • 22.Fu L., Kane C. L., Mele E. J., Topological insulators in three dimensions. Phys. Rev. Lett. 98, 106803 (2007). [DOI] [PubMed] [Google Scholar]
  • 23.Hsieh T. H., et al. , Topological crystalline insulators in the SnTe material class. Nat. Commun. 3, 982 (2012). [DOI] [PubMed] [Google Scholar]
  • 24.Fang C., Fu L., New classes of three-dimensional topological crystalline insulators: Nonsymmorphic and magnetic. Phys. Rev. B 91, 161105 (2015). [Google Scholar]
  • 25.Shiozaki K., Sato M., Topology of crystalline insulators and superconductors. Phys. Rev. B 90, 165114 (2014). [Google Scholar]
  • 26.Fang C., Fu L., New classes of topological crystalline insulators having surface rotation anomaly. Sci. Adv. 20, eaat2374 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Alexandradinata A., Wang C., Duan W., Glazman L., Revealing the topology of Fermi-surface wave functions from magnetic quantum oscillations. Phys. Rev. X 8, 011027 (2018). [Google Scholar]
  • 28.Wang C., Duan W., Glazman L., Alexandradinata A., Landau quantization of nearly degenerate bands and full symmetry classification of Landau level crossings. Phys. Rev. B 100, 014442 (2019). [Google Scholar]
  • 29.Sun S., Song Z., Weng H., Dai X., Topological metals induced by the Zeeman effect. Phys. Rev. B 101, 125118 (2020). [Google Scholar]
  • 30.Fulde P., Ferrell R. A., Superconductivity in a strong spin-exchange field. Phys. Rev. 135, A550–A563 (1964). [Google Scholar]
  • 31.Larkin A. I., Ovchinnikov Y. N., Nonuniform state of superconductors. Sov. Phys. JETP 20, 762 (1965). [Google Scholar]
  • 32.Yuan N. F. Q., Fu L., Zeeman-induced gapless superconductivity with a partial Fermi surface. Phys. Rev. B 97, 115139 (2018). [Google Scholar]
  • 33.Maki K., Tsuneto T., Pauli paramagnetism and superconducting state. Progr. Theor. Phys. 31, 945–956 (1964). [Google Scholar]
  • 34.Maki K., Effect of Pauli paramagnetism on magnetic properties of high-field superconductors. Phys. Rev. 148, 362 (1966). [Google Scholar]

Associated Data

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

Data Availability Statement

All study data are included in this article.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES