Identification of odd-frequency superconducting pairing in Josephson junctions
Abstract
Optimal choice of spin polarization enables electron injection into the helical edge state at a precise position, despite the uncertainty principle, permitting access to specific non-local Green’s functions. We show, within 1D effective description, that this fact facilitates a direct identification of odd-frequency pairing through parity measurement (under frequency reversal) of the non-local differential conductance in a setup comprising Josephson junction on the helical edge state of a 2D topological insulator with two spin-polarized probes tunnel-coupled to the junction region. A 2D numerical simulation has also been conducted to confirm theoretical predictions as well as to demonstrate the experimental feasibility of the proposal.
pacs:
74.50.+r,74.78.Na,85.25.-j,85.25.Cp,85.65.+h,75.50.Xx,85.80.FiIntroduction: In a conventional superconductor, electrons form Cooper pairs due to a weak attraction mediated via lattice vibrations[1]. The pairing function between the two electrons within a Cooper pair can be classified into different symmetry classes by means of four permutation operators with respect to spin (), relative coordinate (), orbital index (), and time coordinate (), all of which can only have eigenvalues . Generally, BCS theory of single orbital superconductor () is instantaneous in time (). This leaves two choices for the combinations of eigenvalues of the operators and to satisfy the Berezinskii condition[2] for the pairing function, i.e., . These correspond to spin-singlet even-parity (e.g. -wave) and spin-triplet odd-parity[3] (e.g. -wave) symmetries, respectively. If one removes the constraint of instantaneous pairing, the possibility of temporal (non-local in time) Cooper pairing becomes feasible, and both can be possible, depending on whether the pairing function is symmetric or anti-symmetric under the permutation of relative time coordinates of the electrons forming the Cooper pair. Temporally symmetric pairing () does not give rise to any new symmetry classes other than those already described by BCS theory and is generally known as even-frequency (even-) pairing. However, temporally anti-symmetric pairing (), which is known as odd-frequency (odd-) pairing[4, 5, 6], opens up possibilities of two new symmetry classes (assuming ) that cannot be described by BCS theory, namely, spin-singlet odd-parity () and spin-triplet even-parity () pairings, consistent with Berezinskii condition. It can also be shown that this odd- (even-) pairing is also odd (even) under the sign change of energy[4], which gives the advantage of observing the behavior of this pairing in the energy domain rather than in the temporal domain.
On a historical note, the concept of odd- spin-triplet pairing traces back to Berezinskii’s 1974 proposal in 3He[2] and predictions in disordered systems[7, 8]. Balatsky and Abrahams later indicated the presence of odd- spin-singlet pairing in time-reversal and parity symmetry broken superconductors[9]. Subsequently, the odd- pairing was explored within the framework of a two-channel Kondo system[10], the 1D model[11], the 2D Hubbard model[12] and heavy fermion compounds[13]. The bulk odd- pairing has been indirectly hinted theoretically through the Majorana scanning tunneling microscope[14] and also experimentally using the Kerr effect[15, 16], and the paramagnetic Meissner effect[17, 18, 19, 20, 21]. Further, the evidence of bulk odd- pairing due to magnetic impurities has emerged in -wave superconductors[22, 23]. In Ref. [24], a measurement tool was introduced for directly detecting odd- pairing in bulk systems using time- and angle-resolved photoelectron fluctuation spectroscopy. However, this method requires advanced technology beyond the current capabilities of facilities. More recently, in Ref. [25], authors proposed a detection scheme to directly identify odd- pairing in bulk systems using the quasiparticle interference method in the presence of an external magnetic field. Initially, odd- pairing was regarded as an inherent bulk phenomenon[2, 9, 26] but was subsequently acknowledged to manifest in heterojunctions[27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55] and in the systems under the influence of time-dependent fields[56, 57]. Certain theoretical studies have indirectly identified odd- pairing in heterostructures. For instance, this was achieved through phase-tunable electron transport in topological Josephson junctions (JJ)[58], examining Josephson current characteristics on the surface of Weyl nodal loop semimetals[59], and analyzing current noise in JJ[60]. Experimental evidence of odd- pairing in heterostructures emerged through measurements of long-range supercurrents in magnetic JJ[61, 62]. These two experiments utilize a ferromagnetic material to create an odd- pairing effect, which is subsequently identified by measuring long-range superconducting pairing. In contrast, this letter investigates a configuration that inherently supports odd- pairing without the need for any ferromagnet/superconductor junction. Topological JJ at the edge of a two-dimensional (2D) quantum-spin-Hall insulator (QSHI) serves this purpose. It is important to mention here that induced superconductivity in the QSHI edge has already been experimentally demonstrated[63].
We demonstrate that it is possible to read off all the independent, spatially non-local Green’s functions by measuring differential conductance between two spin-polarized probes placed at the junction. The underlying interplay between spin-momentum locking of helical edge states (HES) and spin-polarization of the probes allows for this advantage. Experimentally, the injection of spin-polarized current into the HES of QSHI has already been achieved[64]. By using spin-polarized probes, one can inject (detect) electrons into (from) the helical edge with well-defined position and momentum simultaneously, which, in general, is prohibited due to uncertainty principle[65].To elaborate further, a spin- (spin-) electron injected at position in the helical edge can only tunnel into the right (left) moving edge mode owing to spin-momentum locking. Keeping the spin-polarization axes of the tunneling probes aligned parallel or anti-parallel to the spin-quantization axis of HES, we study the difference between non-local differential conductance from the left probe to right probe and from right probe to left probe i.e.
(1) |
where with , and denotes the applied voltage at the left (right) tunneling probe (). It can be shown that this quantity can be written as a product of odd- and even- pairing amplitudes and hence should be odd with respect to incident energy (and hence with respect to the applied voltage ). We show that this difference stems from the same origin as that of the odd- part of the non-local differential conductance, which can be accessed via bias reversal i.e. . This fact leads to
(2) |
From an experimental perspective, measuring the anti-symmetric behavior of each of these independent quantities and their equality confirms the existence of odd- pairing. In what follows, we will first present our results using an effective 1D analytic model of HES. Following that, we’ll provide a 2D simulation of a realistic setup, in line with HgTe/CdTe quantum well parameters, confirming our predictions.
Model and its motivation: The system, we are interested in, is a JJ of length (), realized in a helical edge state (HES) of a 2D QSHI[66, 67, 68, 69, 70] (1D Dirac fermions) which is proximitized to a conventional -wave superconductor. Schematic of the setup is shown in FIG. 1 (a), while FIG. 1 (b) shows the corresponding simplified cartoon representation of the same setup. The Bogoliubov-de Gennes (BdG) Hamiltonian characterizing such junctions can be written as[71, 72] where is the Nambu basis and
(3) |
Here, and are the Pauli matrices acting respectively on the spin basis and particle-hole basis, is the chemical potential of the HES, represents Fermi velocity of the HES, and, is the momentum operator. The spin-quantization axis of the HES is considered to be along -direction[71, 72]. The superconducting pairing potential is given by . Superconducting leads () are identified as with the corresponding superconducting phases such that the superconducting phase difference , in general.

Input | Polarization | Polarization | Output | Contributing | Input | Polarization | Polarization | Output | Contributing |
at | of () | of () | at | Green’s function | at | of () | of () | at | Green’s function |
e | e | e | e | ||||||
e | h | e | h | ||||||
e | h | e | h | ||||||
e | e | e | e |
Two spin-polarized tunneling probes[73] ( being closer to ) with spin-polarization axes oriented respectively along and , at angles and with respect to the spin-quantization axis of the HES (say, in the plane)111The azimuthal angle is of no consequence, e.g., it can also be assumed in the plane, are introduced at positions and respectively within the junction region (). For most of the paper, we will be only discussing the case for , unless otherwise stated. To keep the algebra simple while retaining the essential physical inputs, we model the probes as a spin-polarized 1D mode. The second quantized Hamiltonians of these probes can be expressed as:
(4) |
The tunneling Hamiltonian between the HES and the probes can be written as:
(5) |
with ( and ). is the tunneling strength between and at and is the overlap of the spinor part of the first quantized wave function of electrons in the probes and the HES. From this point onward, we shall consider . Note that this form of tunneling respects symmetry and hence cannot induce spin-flip scattering when is parallel or anti-parallel to the spin-quantization axis of the HES. Using the equation of motion approach for the Hamiltonian (where ), the scattering matrices can be expressed as,
(6) |
where are the corresponding incoming (outgoing) plane wave amplitudes on the HES and the probe. The scattering matrix for holes () can be determined by exploiting the particle-hole symmetry (for details, see Supplemental Material (SM) Section A) of the system. We are interested in studying charge transport between the two probes via the junction when a finite voltage bias is applied such that .
Note that, the spatial non-locality of the probes, along with the helical nature of the edge state, opens up the possibility of detecting only a hole at when an electron is injected into HES through provided the polarization of the probes is tuned such that and . This leads to the fact that the differential conductance between the probes is directly proportional to the square of the modulus of anomalous Green’s functions. This can be understood in terms of a simple process, as described below. If the spin polarization of is set to spin- (i.e., ), it can only inject a right-moving electron to the HES due to the spin-momentum locking of the edge states. In the absence of spin-flip scattering within the HES, this electron will either remain as a spin- electron after an even number of Andreev reflections (including the possibility of no Andreev reflection) or will convert to a spin- hole after an odd number of Andreev reflections when it reaches . Thus, will detect only a hole for . Along the same line of argument, it is straightforward to see that for , will detect only an electron. By definition, the probability amplitude of these non-local transmissions will be proportional to the corresponding anomalous Green’s functions, namely and , where and are the propagation amplitudes for spin up electron and spin down hole respectively from the left probe to the right probe and, represents the retarded (advanced) normal Green’s functions while () denotes the retarded (advanced) anomalous Green’s functions calculated at . Here retarded and advanced Green’s functions are defined in the appropriate frequency domain[38]. The above discussion is summarized in the form of TABLE. 1.
Using Landauer’s formula[75], we can calculate the differential conductance from to , given by and from to , given by , where (, , ). The difference between these two non-local differential conductances at energy (frequency) is given by:
(7) |
where we have used the property ( and ). Here we have used only retarded Green’s functions to express . However, a proof involving advanced Green’s functions follows in a similar way (See SM Section B). In Eq. (7), is anti-symmetric with respect to , and if measured, it can be treated as a direct signature of odd- pairing subjected to the condition that even- pairing within the junction is non-zero. In Eq. (7), and denote the odd-in- (anti-symmetric with respect to ) and even-in- (symmetric with respect to ) parts of the corresponding Green’s functions which are in turn directly related to odd- and even- pairing amplitudes respectively. Note that, the quantity in Eq. (7) can also be expressed as . The voltage configuration necessary for measuring and is discussed above Eq. (2). It is worth mentioning that the Green’s functions discussed above are not, in general, the free Green’s function of the JJ, i.e. the Green’s functions in the absence of tunneling probes. However, in the weak tunneling limit (), different inter-lead transmission amplitudes can be written in terms of free Green’s functions (for details see SM Section C). Next, we will proceed to put the above qualitative discussion on firm grounds by calculating the quantity from numerical analysis using a 2D model, and comparing the theoretical results discussed above (and also results presented in the SM).

Numerical simulation using 2D model: In this part, we conduct a numerical simulation of a 2D lattice model depicted in Fig. 1 using the KWANT package [76] to further analyse the effective 1D model discussed above. The system is mainly a QSH insulator which is described by a real space configuration of the Bernevig-Hughes-Zhang (BHZ) model[77] mapped on a square lattice of dimension with additional terms incorporated as needed. Here ) is the lattice spacing. Two superconducting leads are attached to the bottom side () and to the right side () of the system, which are defined by the -wave superconductivity-proximitized QSH insulators[78] having a finite phase difference () between them and thus, defining a 1D JJ extended over the left and top edges of the QSH region (see Fig. 1(a)). On the other hand, two spin-polarized normal leads, which can be described by two quantum anomalous Hall insulators carrying chiral edge states of opposite chiralities, are tunnel-coupled via tiny insulating sections to the left side (: having the -spin channel only) and to the top side (: having only the -spin channel) of the QSH region. The QAH state of a spin-polarized lead is obtained by applying an exchange field to the QSH system, which oppositely affects the effective bulk gaps for the two spin states, eventually, at sufficient strength, destroying the topological state for one of them [79]. A strong onsite ferromagnetic impurity term is added over a small section of dimension at the extreme bottom-right corner (denoted by ) of the QSH region to ensure that the current is allowed to propagate only through the left and top edges of the QSH region in-between and . A detailed insight into the 2D simulation is closely comparable to the same discussed in Ref. [80] along with a comprehensive understanding of how particle-hole symmetry can be incorporated into the lattice model [81]. For the numerical calculation, we consider the value of chemical potential () to be zero and the -wave superconducting pairing potential () as 2.0 meV. The values of the remaining parameters related to the BHZ model closely resemble those describing HgTe/CdTe quantum wells[77, 80, 81]. Finally, by conducting a numerical simulation with this setup, we compute the quantity for as shown in FIG. 2 (solid cream line). The insulating necks of and are considered to be identical.
To compare the numerical result with the theoretical result, one needs to estimate the values of , , and . To estimate the value of , we analyzed the positions of the peaks and dips of . By mapping the corresponding energy values to the ABS energies, we estimate where is the superconducting coherence length. Plugging this value of into theoretical calculations we numerically calculated the following quantity:
(8) |
for different values of and . Here is the number of numerical grids within the energy window . Keeping the tunneling strengths same, it turns out for (dashed brown line in FIG. 2). Although, this plot is a good fit to the numerical results, if we relax the condition of . then the value of can be even minimized giving rise to a better fit to the numerical results. It turns out, for and (black dot-dashed line in FIG. 2). Dependence of on the system parameters , and is shown in SM Section D.
Discussion: The above results rely on the fact that the spin-polarization axes of and can be adjusted exactly parallel or anti-parallel to the spin-polarization axis of HES. For any other values of and , will have contributions from interference effects between the spin channels and the electron-hole channels due to the presence of more than one Green’s function in the expressions of . These interference contributions will lead to the deviation from perfectly anti-symmetric features. Also, the detection of odd-frequency pairing through does not require the probes and to be placed at different positions. This is due to the fact that the information of position enters into the calculation as pure phase contributions to (, ).
In our proposed set-up, the size of the sample (i.e. the length of the junction ) should be large enough to accommodate two spin-polarized probes. In Ref. [64], the width of the spin-polarized probe tip is , which sets the minimum dimension of the quantum spin-Hall sample (length of the edge) to be . However, in the same experiment, the size of the sample (length of the edge) is . Thus, it should be possible to experimentally fabricate a similar sample to realize our proposal. The values of other model parameters, e.g. Fermi velocity , the chemical potential , and pairing term can be estimated from the experiment[63].
Lastly, note that is related to the ABS corresponding to the shuttling of a Cooper pair from left to right, while is related to the ABS corresponding to the shuttling of Cooper pair in the opposite direction. Thus, in the weak tunneling limit, at or , (when these two ABS become degenerate), the differential conductances and becomes equal, leading to the vanishing of .
Acknowledgements: S.P. and A.M. acknowledge the Ministry of Education, India, and IISER Kolkata for funding. V.A. would like to acknowledge Korea NRF (SRC Center for Quantum Coherence in Condensed Matter, Grant No. RS-2023-00207732) for funding.
Author contribution: S.P. and A.M. contributed equally to this work.
References
- Bardeen et al. [1957] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Theory of superconductivity, Phys. Rev. 108, 1175 (1957).
- Berezinskii [1974] V. Berezinskii, New model of the anisotropic phase of superfluid he3, JETP Lett 20 (1974).
- Sigrist and Ueda [1991] M. Sigrist and K. Ueda, Phenomenological theory of unconventional superconductivity, Rev. Mod. Phys. 63, 239 (1991).
- Linder and Balatsky [2019] J. Linder and A. V. Balatsky, Odd-frequency superconductivity, Rev. Mod. Phys. 91, 045005 (2019).
- Tanaka et al. [2012] Y. Tanaka, M. Sato, and N. Nagaosa, Symmetry and topology in superconductors–odd-frequency pairing and edge states–, Journal of the Physical Society of Japan 81, 011013 (2012).
- Cayao et al. [2020] J. Cayao, C. Triola, and A. M. Black-Schaffer, Odd-frequency superconducting pairing in one-dimensional systems, The European Physical Journal Special Topics 229, 545 (2020).
- Kirkpatrick and Belitz [1991] T. R. Kirkpatrick and D. Belitz, Disorder-induced triplet superconductivity, Phys. Rev. Lett. 66, 1533 (1991).
- Belitz and Kirkpatrick [1992] D. Belitz and T. R. Kirkpatrick, Even-parity spin-triplet superconductivity in disordered electronic systems, Phys. Rev. B 46, 8393 (1992).
- Balatsky and Abrahams [1992] A. Balatsky and E. Abrahams, New class of singlet superconductors which break the time reversal and parity, Phys. Rev. B 45, 13125 (1992).
- Emery and Kivelson [1992] V. J. Emery and S. Kivelson, Mapping of the two-channel kondo problem to a resonant-level model, Phys. Rev. B 46, 10812 (1992).
- Balatsky and Boncˇa [1993] A. V. Balatsky and J. Boncˇa, Even- and odd-frequency pairing correlations in the one-dimensional t-j-h model: A comparative study, Phys. Rev. B 48, 7445 (1993).
- Bulut et al. [1993] N. Bulut, D. J. Scalapino, and S. R. White, Effective particle-particle interaction in the two-dimensional hubbard model, Phys. Rev. B 47, 6157 (1993).
- Coleman et al. [1993] P. Coleman, E. Miranda, and A. Tsvelik, Possible realization of odd-frequency pairing in heavy fermion compounds, Phys. Rev. Lett. 70, 2960 (1993).
- Kashuba et al. [2017] O. Kashuba, B. Sothmann, P. Burset, and B. Trauzettel, Majorana stm as a perfect detector of odd-frequency superconductivity, Phys. Rev. B 95, 174516 (2017).
- Schemm et al. [2014] E. Schemm, W. Gannon, C. Wishne, W. Halperin, and A. Kapitulnik, Observation of broken time-reversal symmetry in the heavy-fermion superconductor upt3, Science 345, 190 (2014).
- Komendová and Black-Schaffer [2017] L. Komendová and A. M. Black-Schaffer, Odd-frequency superconductivity in measured by kerr rotation, Phys. Rev. Lett. 119, 087001 (2017).
- Di Bernardo et al. [2015a] A. Di Bernardo, Z. Salman, X. L. Wang, M. Amado, M. Egilmez, M. G. Flokstra, A. Suter, S. L. Lee, J. H. Zhao, T. Prokscha, E. Morenzoni, M. G. Blamire, J. Linder, and J. W. A. Robinson, Intrinsic paramagnetic meissner effect due to -wave odd-frequency superconductivity, Phys. Rev. X 5, 041021 (2015a).
- Bergeret et al. [2001a] F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Josephson current in superconductor-ferromagnet structures with a nonhomogeneous magnetization, Phys. Rev. B 64, 134506 (2001a).
- Alidoust et al. [2014] M. Alidoust, K. Halterman, and J. Linder, Meissner effect probing of odd-frequency triplet pairing in superconducting spin valves, Phys. Rev. B 89, 054508 (2014).
- Krieger et al. [2020] J. A. Krieger, A. Pertsova, S. R. Giblin, M. Döbeli, T. Prokscha, C. W. Schneider, A. Suter, T. Hesjedal, A. V. Balatsky, and Z. Salman, Proximity-induced odd-frequency superconductivity in a topological insulator, Phys. Rev. Lett. 125, 026802 (2020).
- Yokoyama et al. [2011] T. Yokoyama, Y. Tanaka, and N. Nagaosa, Anomalous meissner effect in a normal-metal–superconductor junction with a spin-active interface, Phys. Rev. Lett. 106, 246601 (2011).
- Perrin et al. [2020] V. Perrin, F. L. N. Santos, G. C. Ménard, C. Brun, T. Cren, M. Civelli, and P. Simon, Unveiling odd-frequency pairing around a magnetic impurity in a superconductor, Phys. Rev. Lett. 125, 117003 (2020).
- Kuzmanovski et al. [2020] D. Kuzmanovski, R. S. Souto, and A. V. Balatsky, Odd-frequency superconductivity near a magnetic impurity in a conventional superconductor, Phys. Rev. B 101, 094505 (2020).
- Kornich et al. [2021] V. Kornich, F. Schlawin, M. A. Sentef, and B. Trauzettel, Direct detection of odd-frequency superconductivity via time- and angle-resolved photoelectron fluctuation spectroscopy, Phys. Rev. Res. 3, L042034 (2021).
- Chakraborty and Black-Schaffer [2022] D. Chakraborty and A. M. Black-Schaffer, Quasiparticle interference as a direct experimental probe of bulk odd-frequency superconducting pairing, Phys. Rev. Lett. 129, 247001 (2022).
- Abrahams et al. [1995] E. Abrahams, A. Balatsky, D. J. Scalapino, and J. R. Schrieffer, Properties of odd-gap superconductors, Phys. Rev. B 52, 1271 (1995).
- Bergeret et al. [2001b] F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Long-range proximity effects in superconductor-ferromagnet structures, Phys. Rev. Lett. 86, 4096 (2001b).
- Bergeret et al. [2003] F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Manifestation of triplet superconductivity in superconductor-ferromagnet structures, Phys. Rev. B 68, 064513 (2003).
- 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).
- Eschrig et al. [2003] M. Eschrig, J. Kopu, J. C. Cuevas, and G. Schön, Theory of half-metal/superconductor heterostructures, Phys. Rev. Lett. 90, 137003 (2003).
- Volkov et al. [2006] A. F. Volkov, A. Anishchanka, and K. B. Efetov, Odd triplet superconductivity in a superconductor/ferromagnet system with a spiral magnetic structure, Phys. Rev. B 73, 104412 (2006).
- Yokoyama et al. [2007] T. Yokoyama, Y. Tanaka, and A. A. Golubov, Manifestation of the odd-frequency spin-triplet pairing state in diffusive ferromagnet/superconductor junctions, Phys. Rev. B 75, 134510 (2007).
- Tanaka et al. [2007a] Y. Tanaka, Y. Tanuma, and A. A. Golubov, Odd-frequency pairing in normal-metal/superconductor junctions, Phys. Rev. B 76, 054522 (2007a).
- Black-Schaffer and Balatsky [2012] A. M. Black-Schaffer and A. V. Balatsky, Odd-frequency superconducting pairing in topological insulators, Phys. Rev. B 86, 144506 (2012).
- Black-Schaffer and Balatsky [2013] A. M. Black-Schaffer and A. V. Balatsky, Proximity-induced unconventional superconductivity in topological insulators, Phys. Rev. B 87, 220506 (2013).
- Crépin et al. [2015] F. m. c. Crépin, P. Burset, and B. Trauzettel, Odd-frequency triplet superconductivity at the helical edge of a topological insulator, Phys. Rev. B 92, 100507 (2015).
- Burset et al. [2015] P. Burset, B. Lu, G. Tkachov, Y. Tanaka, E. M. Hankiewicz, and B. Trauzettel, Superconducting proximity effect in three-dimensional topological insulators in the presence of a magnetic field, Phys. Rev. B 92, 205424 (2015).
- Cayao and Black-Schaffer [2017] J. Cayao and A. M. Black-Schaffer, Odd-frequency superconducting pairing and subgap density of states at the edge of a two-dimensional topological insulator without magnetism, Phys. Rev. B 96, 155426 (2017).
- Hwang et al. [2018] S.-Y. Hwang, P. Burset, and B. Sothmann, Odd-frequency superconductivity revealed by thermopower, Phys. Rev. B 98, 161408 (2018).
- Cayao and Black-Schaffer [2018] J. Cayao and A. M. Black-Schaffer, Odd-frequency superconducting pairing in junctions with rashba spin-orbit coupling, Phys. Rev. B 98, 075425 (2018).
- Linder et al. [2010] J. Linder, A. Sudbø, T. Yokoyama, R. Grein, and M. Eschrig, Signature of odd-frequency pairing correlations induced by a magnetic interface, Phys. Rev. B 81, 214504 (2010).
- Linder et al. [2009] J. Linder, T. Yokoyama, A. Sudbø, and M. Eschrig, Pairing symmetry conversion by spin-active interfaces in magnetic normal-metal–superconductor junctions, Phys. Rev. Lett. 102, 107008 (2009).
- Pal and Benjamin [2021] S. Pal and C. Benjamin, Exciting odd-frequency equal-spin triplet correlations at metal-superconductor interfaces, Phys. Rev. B 104, 054519 (2021).
- Tamura et al. [2023] S. Tamura, Y. Tanaka, and T. Yokoyama, Generation of polarized spin-triplet cooper pairings by magnetic barriers in superconducting junctions, Phys. Rev. B 107, 054501 (2023).
- Tanaka and Golubov [2007] Y. Tanaka and A. A. Golubov, Theory of the proximity effect in junctions with unconventional superconductors, Phys. Rev. Lett. 98, 037003 (2007).
- Eschrig et al. [2007] M. Eschrig, T. Löfwander, T. Champel, J. Cuevas, J. Kopu, and G. Schön, Symmetries of pairing correlations in superconductor–ferromagnet nanostructures, Journal of Low Temperature Physics 147, 457 (2007).
- Volkov et al. [2003] A. F. Volkov, F. S. Bergeret, and K. B. Efetov, Odd triplet superconductivity in superconductor-ferromagnet multilayered structures, Phys. Rev. Lett. 90, 117006 (2003).
- Fominov et al. [2007] Y. V. Fominov, A. F. Volkov, and K. B. Efetov, Josephson effect due to the long-range odd-frequency triplet superconductivity in junctions with néel domain walls, Phys. Rev. B 75, 104509 (2007).
- Buzdin [2005] A. I. Buzdin, Proximity effects in superconductor-ferromagnet heterostructures, Rev. Mod. Phys. 77, 935 (2005).
- Tsintzis et al. [2019] A. Tsintzis, A. M. Black-Schaffer, and J. Cayao, Odd-frequency superconducting pairing in kitaev-based junctions, Phys. Rev. B 100, 115433 (2019).
- Asano and Tanaka [2013] Y. Asano and Y. Tanaka, Majorana fermions and odd-frequency cooper pairs in a normal-metal nanowire proximity-coupled to a topological superconductor, Phys. Rev. B 87, 104513 (2013).
- Kuzmanovski and Black-Schaffer [2017] D. Kuzmanovski and A. M. Black-Schaffer, Multiple odd-frequency superconducting states in buckled quantum spin hall insulators with time-reversal symmetry, Phys. Rev. B 96, 174509 (2017).
- Fleckenstein et al. [2018] C. Fleckenstein, N. T. Ziani, and B. Trauzettel, Conductance signatures of odd-frequency superconductivity in quantum spin hall systems using a quantum point contact, Phys. Rev. B 97, 134523 (2018).
- Tanaka and Tamura [2021] Y. Tanaka and S. Tamura, Theory of surface andreev bound states and odd-frequency pairing in superconductor junctions, Journal of Superconductivity and Novel Magnetism 34, 1677 (2021).
- Tanaka et al. [2007b] Y. Tanaka, A. A. Golubov, S. Kashiwaya, and M. Ueda, Anomalous josephson effect between even- and odd-frequency superconductors, Phys. Rev. Lett. 99, 037005 (2007b).
- Triola and Balatsky [2016] C. Triola and A. V. Balatsky, Odd-frequency superconductivity in driven systems, Phys. Rev. B 94, 094518 (2016).
- Triola and Balatsky [2017] C. Triola and A. V. Balatsky, Pair symmetry conversion in driven multiband superconductors, Phys. Rev. B 95, 224518 (2017).
- Cayao et al. [2022] J. Cayao, P. Dutta, P. Burset, and A. M. Black-Schaffer, Phase-tunable electron transport assisted by odd-frequency cooper pairs in topological josephson junctions, Phys. Rev. B 106, L100502 (2022).
- Dutta and Black-Schaffer [2019] P. Dutta and A. M. Black-Schaffer, Signature of odd-frequency equal-spin triplet pairing in the josephson current on the surface of weyl nodal loop semimetals, Phys. Rev. B 100, 104511 (2019).
- Seoane Souto et al. [2020] R. Seoane Souto, D. Kuzmanovski, and A. V. Balatsky, Signatures of odd-frequency pairing in the josephson junction current noise, Phys. Rev. Res. 2, 043193 (2020).
- Khaire et al. [2010] T. S. Khaire, M. A. Khasawneh, W. P. Pratt, and N. O. Birge, Observation of spin-triplet superconductivity in co-based josephson junctions, Phys. Rev. Lett. 104, 137002 (2010).
- Di Bernardo et al. [2015b] A. Di Bernardo, S. Diesch, Y. Gu, J. Linder, G. Divitini, C. Ducati, E. Scheer, M. G. Blamire, and J. W. Robinson, Signature of magnetic-dependent gapless odd frequency states at superconductor/ferromagnet interfaces, Nature communications 6, 8053 (2015b).
- Hart et al. [2014] S. Hart, H. Ren, T. Wagner, P. Leubner, M. Mühlbauer, C. Brüne, H. Buhmann, L. W. Molenkamp, and A. Yacoby, Induced superconductivity in the quantum spin hall edge, Nature Physics 10, 638 (2014).
- Brüne et al. [2012] C. Brüne, A. Roth, H. Buhmann, E. M. Hankiewicz, L. W. Molenkamp, J. Maciejko, X.-L. Qi, and S.-C. Zhang, Spin polarization of the quantum spin hall edge states, Nature Physics 8, 485 (2012).
- Das and Rao [2011] S. Das and S. Rao, Spin-polarized scanning-tunneling probe for helical luttinger liquids, Phys. Rev. Lett. 106, 236403 (2011).
- Fu and Kane [2009] L. Fu and C. L. Kane, Josephson current and noise at a superconductor/quantum-spin-hall-insulator/superconductor junction, Phys. Rev. B 79, 161408 (2009).
- 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).
- Kane and Mele [2005] C. L. Kane and E. J. Mele, Quantum spin hall effect in graphene, Phys. Rev. Lett. 95, 226801 (2005).
- Calzona and Trauzettel [2019] A. Calzona and B. Trauzettel, Moving majorana bound states between distinct helical edges across a quantum point contact, Phys. Rev. Res. 1, 033212 (2019).
- Keidel et al. [2020] F. Keidel, S.-Y. Hwang, B. Trauzettel, B. Sothmann, and P. Burset, On-demand thermoelectric generation of equal-spin cooper pairs, Phys. Rev. Res. 2, 022019 (2020).
- Mukhopadhyay and Das [2021] A. Mukhopadhyay and S. Das, Thermal signature of the majorana fermion in a josephson junction, Phys. Rev. B 103, 144502 (2021).
- Mukhopadhyay and Das [2022] A. Mukhopadhyay and S. Das, Thermal bias induced charge current in a josephson junction: From ballistic to disordered, Phys. Rev. B 106, 075421 (2022).
- Wadhawan et al. [2018] D. Wadhawan, K. Roychowdhury, P. Mehta, and S. Das, Multielectron geometric phase in intensity interferometry, Phys. Rev. B 98, 155113 (2018).
- Note [1] The azimuthal angle is of no consequence, e.g., it can also be assumed in the plane.
- Pershoguba et al. [2019] S. S. Pershoguba, T. Veness, and L. I. Glazman, Landauer formula for a superconducting quantum point contact, Phys. Rev. Lett. 123, 067001 (2019).
- Groth et al. [2014] C. W. Groth, M. Wimmer, A. R. Akhmerov, and X. Waintal, Kwant: a software package for quantum transport, New Journal of Physics 16, 063065 (2014).
- Bernevig et al. [2006] B. A. Bernevig, T. L. Hughes, and S.-C. Zhang, Quantum spin hall effect and topological phase transition in hgte quantum wells, Science 314, 1757 (2006), https://www.science.org/doi/pdf/10.1126/science.1133734 .
- Alicea [2012] J. Alicea, New directions in the pursuit of majorana fermions in solid state systems, Reports on progress in physics 75, 076501 (2012).
- Li et al. [2013] H. Li, L. Sheng, R. Shen, L. B. Shao, B. Wang, D. N. Sheng, and D. Y. Xing, Stabilization of the quantum spin hall effect by designed removal of time-reversal symmetry of edge states, Phys. Rev. Lett. 110, 266802 (2013).
- Adak et al. [2022a] V. Adak, K. Roychowdhury, and S. Das, Spin-polarized voltage probes for helical edge state: A model study, Physica E: Low-dimensional Systems and Nanostructures 139, 115125 (2022a).
- Adak et al. [2022b] V. Adak, A. Mukhopadhyay, S. J. De, U. Khanna, S. Rao, and S. Das, Chiral detection of majorana bound states at the edge of a quantum spin hall insulator, Phys. Rev. B 106, 045422 (2022b).