This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Dark Andreev States in Superconductors

Andrey Grankin and Victor Galitski Joint Quantum Institute, Department of Physics, University of Maryland, College Park, MD 20742, USA
Abstract

The conventional Bardeen-Cooper-Schrieffer (BCS) model of superconductivity assumes a frequency-independent order parameter, which allows a relatively simple description of the superconducting state. In particular, its excitation spectrum readily follows from the Bogoliubov-de-Gennes (BdG) equations. A more realistic description of a superconductor is the Migdal-Eliashberg theory, where the pairing interaction, the order parameter, and electronic self-energy are strongly frequency dependent. This work combines these ingredients of phonon-mediated superconductivity with the standard BdG approach. Surprisingly, we find qualitatively new features such as the emergence of a shadow superconducting gap in the quasiparticle spectrum at energies close to the Debye energy. We show how these features reveal themselves in standard tunneling experiments. Finally, we also predict the existence of additional high-energy bound states, which we dub “dark Andreev states.”

Bardeen-Cooper-Schrieffer theory of superconductivity (Bardeen et al., 1957) and Bogoliubov-de Gennes equations have proven to be relatively simple and reliable tools to describe a variety of conventional superconductors. A key simplifying assumption of this approach is that the superconducting order parameter is energy-independent (Altland and Simons, 2010). While it is clearly not the case in any real superconductor, many features such as the quasiparticle spectrum, thermodynamic and electromagnetic properties (Altland and Simons, 2010; Bardeen et al., 1957) appear insensitive to this approximation. It is reasonable, because the omitted energy dependence normally does not affect low-energy physics. In contrast, accurate determination of the transition temperature is sensitive to the details of the phonon dispersion, the dynamical screening of Coulomb interaction, and the structure of the superconducting gap. The latter can be obtained from the Migdal-Eliashberg (ME) equations (Eliashberg, 1960; Marsiglio, 2020) which are integral equations in both energy and momentum.

This Migdal-Eliashberg theory Marsiglio et al. (1988) predicts a non-trivial structure of the superconducting gap as a function of real frequency. In particular, it was shown that for superconductivity mediated by the Einstein phonon modes (Marsiglio, 2018; Mirabi et al., 2020), the gap function has sharp resonance-like features. In the weak coupling regime, these features can be well approximated by the Lorentz function centered at the Debye frequency. Another interesting real-frequency behavior of the gap function is discussed in (Christensen and Chubukov, 2021) where the authors predict the possibility of formation of frequency-domain vortices in the presence of phonon-induced attraction and Coulomb repulsion.

In this work, we study how a frequency-dependent order parameter affects the quasiparticle spectrum and tunneling properties of a superconductor. We assume that the frequency dependence has a Lorentz shape, which corresponds to optical-phonon-mediated pairing (Marsiglio, 2020). We solve the corresponding generalized BdG equations and show that an additional gap emerges in the qusiparticle spectrum at higher energies. In the case of an SNS junction, we also find that additional Andreev in-gap high-energy states can form. We use analogy with quantum optics to interpret these high-energy peaks in terms of “dark” resonance features (Fleischhauer and Lukin, 2000; Lukin et al., 1999; Fleischhauer and Lukin, 2002) of the BdG Hamiltonian with a frequency-dependent order parameter. This suggests a speculation that these finite-energy states could potentially be used to reliably store quantum information.

Refer to caption
Figure 1: a) Effective proximity system which induces a Lorentz-shape frequency dependence of the order parameter. Fictitious flat-band superconductor shown in blue and a normal-state metal is shown in orange. b) Quasiparticle spectrum of the BdG equation with the frequency-dependent order parameter. Additional band gaps are formed close to the characteristic frequency of the order parameter frequency dependence. Blue solid lines correspond to the eigenenergies of Eq. (3). Orange dashed stands for the conventional BCS gap (±ϕ0\pm\phi_{0}) and dashed green lines Ek=ω0E_{k}=\omega_{0} and Ekω0+ϕ0E_{k}\approx\omega_{0}+\phi_{0} correspond to the additional gap due to the frequency-dependence of the order parameter. In the simulation we assumed ϕ0=ω0/10\phi_{0}=\omega_{0}/10. c) Schematic representation of the four-state system equivalent to the on-shell quasi-electron branch of the BdG Hamiltonian Eq. (4) in the limit of weak coupling t0t\rightarrow 0.
Refer to caption
Figure 2: Andreev reflection coefficient as function of the incident electron energy. As result of the additional band gap the additional Andreev reflection peak is observed at energies close to ω0\omega_{0}. Blue solid line corresponds to the BdG equation with the frequency-dependent anomalous self-energy. Orange dashed stands for the conventional BdG assuming no frequency dependence of the order parameter. Inset shows the local density of states of the Green’s function Eq. (1). The gap is chosen such that ϕ0=ω0/10\phi_{0}=\omega_{0}/10.

In this work we consider the conventional s-wave superconductor but keeping a complete realistic frequency dependence of the order parameter. We restrict our discussion to the mean-field level and rely on the Bogoliubov-de Gennes approach, which requires a straightforward generalization. Consider an electron gas with the creation (annihilation) operators ψ𝐤,σ=,\psi^{\dagger}_{{\bf k,\sigma=\shortuparrow,\downarrow}} (ψ𝐤,σ)(\psi_{{\bf k},\sigma}), where 𝐤{\bf k} denotes the electron momentum. For the two-component spinor Ψ𝐤,n={ψ𝐤,(iεn),ψ𝐤,(iεn)}T\Psi_{{\bf k},n}=\{\psi_{{\bf k},\downarrow}\left(i\varepsilon_{n}\right),\psi_{-{\bf k},\uparrow}^{\dagger}\left(-i\varepsilon_{n}\right)\}^{T}, the Green function is defined as: 𝒢^𝐤(iεn)Ψ𝐤,nΨ𝐤,n\hat{{\cal G}}_{{\bf k}}(i\varepsilon_{n})\equiv-\langle\Psi_{{\bf k},n}\otimes\Psi_{{\bf k},n}^{\dagger}\rangle, where εn=(2n+1)π/β\varepsilon_{n}=\left(2n+1\right)\pi/\beta is the standard the fermionic Matsubara frequency, β\beta is the inverse temperature, and nn\in\mathbb{Z}. Neglecting the momentum dependence of the self-energy, the Green function of an interacting Fermi gas can be written as (Marsiglio, 2020; Schrieffer, 2018):

𝒢^𝐤1(iεn)=iεn𝒵nτ^0ϕnτ^1ξkτ^3,\hat{{\cal G}}_{{\bf k}}^{-1}\left(i\varepsilon_{n}\right)=i\varepsilon_{n}\mathcal{Z}_{n}\hat{\tau}_{0}-\phi_{n}\hat{\tau}_{1}-\xi_{k}\hat{\tau}_{3}, (1)

where ξk=k2/2mμ\xi_{k}=k^{2}/2m-\mu, mm is the electron mass, μ\mu is the chemical potential, τi\tau_{i} are Pauli matrices. 𝒵n\mathcal{Z}_{n} denotes the inverse quasiparticle residue obtained from the odd part of the normal-state self-energy. The off-diagonal matrix element ϕn\phi_{n} stands for the anomalous self energy. 𝒵n\mathcal{Z}_{n} and ϕn\phi_{n} naturally appear in both intrinsic and proximity-induced superconductors (Chubukov et al., 2020; Liu et al., 2019). The BdG equations correspond to 𝒢^𝐤1(iε)χ=0\hat{{\cal G}}_{{\bf k}}^{-1}\left(i\varepsilon\right)\chi=0, which parameterically determines the sought-after quasiparticle dispersion (here, χ\chi is a Nambu spinor).

To calculate the quasiparticle spectrum, we need an explicit form of the frequency dependence of the anomalous self-energy. In what follows, we consider the aforementioned Lorentz-shaped form (parameterized by its amplitude ϕ0\phi_{0} and the characteristic frequency ω0\omega_{0})

ϕn=ϕ0ω02εn2+ω02.\phi_{n}=\phi_{0}\,\frac{\omega_{0}^{2}}{\varepsilon_{n}^{2}+\omega_{0}^{2}}. (2)

As demonstrated in Refs. (Marsiglio, 2018; Mirabi et al., 2020) this solution naturally appears in superconductors, where pairing is induced by an optical phonon mode with the Einstein spectrum at weak coupling. We now consider the spectrum of quasiparticles in Eq. (1).

To develop intuition, it is instructive to consider an auxiliary setup, which involves a fictitious flat-band superconductor with a frequency-independent gap proximity coupled to a normal metal as shown in Fig. 1. As we show, a proper choice of parameters in this setup gives rise to a Green function, which replicates Eq. 1 with the order parameter (2). The advantage of this construction is that its BdG Hamiltonian below involves only standard, frequency-independent parameters:

𝒢^k1(iεn)\displaystyle\hat{{\cal G}}_{k}^{\prime-1}\left(i\varepsilon_{n}\right) =iεnτ^0σ^0H^BdG\displaystyle=i\varepsilon_{n}\hat{\tau}_{0}\hat{\sigma}_{0}-\hat{H}_{\text{BdG}} (3)
HBdG=\displaystyle H_{\text{BdG}}= ξkτ^3(σ^3+1^)2ω0τ^1(1^σ^3)2+tτ^3σ^1,\displaystyle\xi_{k}\hat{\tau}_{3}\frac{\left(\hat{\sigma}_{3}+\hat{1}\right)}{2}-\omega_{0}\hat{\tau}_{1}\frac{\left(\hat{1}-\hat{\sigma}_{3}\right)}{2}+t\hat{\tau}_{3}\hat{\sigma}_{1}, (4)

Here τ^i\hat{\tau}_{i} represents Pauli matrices in the Nambu space and σ^\hat{\sigma} parametrizes an effective two-band model: (σ^3+1^)/2\left(\hat{\sigma}_{3}+\hat{1}\right)/2 projects on the normal metal fermion modes and (1^σ^3)/2\left(\hat{1}-\hat{\sigma}_{3}\right)/2 projector on the fictitious flat-band superconductor. The order parameter in the latter is set the characteristic phonon frequency, ω0\omega_{0}. We note that we assumed no intrinsic order parameter for the original fermions. The coefficient tt denotes the tunneling amplitude between the superconductor and the metal. The auxiliary superconductor can be integrated-out generating both the anomalous and normal self energies. By choosing t=ϕ0ω0t=\sqrt{\phi_{0}\omega_{0}} we can exactly match frequency dependence of the order parameter to reproduce Eq. 2. The corresponding 𝒵\mathcal{Z}-factor in Eq. (1) is equal to 𝒵n=[1+ϕ0ω0/(ω02+εn2)]1\mathcal{Z}_{n}=\left[1+\phi_{0}\omega_{0}/(\omega_{0}^{2}+\varepsilon_{n}^{2})\right]\approx 1, for ϕ0ω0\phi_{0}\ll\omega_{0}. This procedure effectively replaces the integrating out the bosonic Einstein-phonon degree of freedom (which generates the frequency dependence of the gap in the physical setup) with the integrating out the degrees of freedom of the auxiliary flat-band superconductor. Note that this picture is introduced for the purposes of illustration only. All results can reproduced in the original model directly. As we discuss in the supplementary material, the inverse quasiparicle residue, 𝒵n\mathcal{Z}_{n} cannot be set identically to one because it would result in an unstable spectrum.

The quasiparticle spectrum can now be found by diagonalizing the static BdG Hamiltonian HBdGH_{\text{BdG}} given in Eq. (4). The result is shown on Fig.  1 (b) and it has two band gaps. First, we observe the conventional BCS-like band gap at low energies [ϕ0,ϕ0][-\phi_{0},\phi_{0}]. It can be obtained by e.g. neglecting the frequency-dependence of the self-energy in Eq. (1), which reduces it to the textbook case. The second band gap at high energies is a specific feature of the two-band system as defined in Eq. (3). As a result of the flatness of one of the dispersion relations, the avoided crossing forms a band gap close to the frequency, ω0\omega_{0}. The value of this second band gap can be readily obtained analytically from Eq. (3):

ϕ2=ω02(ω0(4ϕ0+ω0)+ϕ0+ω02)ω0ϕ0,\phi_{2}=\sqrt{\frac{\omega_{0}}{2}\left(\sqrt{\omega_{0}(4\phi_{0}+\omega_{0})}+\phi_{0}+\frac{\omega_{0}}{2}\right)}-\omega_{0}\approx\phi_{0}, (5)

where the approximate sign corresponds to the limit ϕ0ω0\phi_{0}\ll\omega_{0}, which must hold for weak coupling.

We now define the local density of states (LDOS) of the electron gas as 1π𝑑ξkIm(G𝐤(ω+i0+))1,1\frac{-1}{\pi}\int d\xi_{k}\text{Im}\left(G_{{\bf k}}\left(\omega+i0^{+}\right)\right)_{1,1}, where G𝐤G_{{\bf k}} is the Green function 3 analytically continued to real frequencies. As the direct consequence of the band gap the LDOS is strongly depleted at energies [ω0,ω0+ϕ0]\approx[\omega_{0},\omega_{0}+\phi_{0}] as shown on inset in Fig. 2. We note that the exactly zero density of states is a feature of the flat-band dispersion of the auxiliary superconductor/Einstein phonons. However, as we discuss in the SM, introduction of a finite curvature to the phonon dispersion would still lead to a significant depletion of the density of states. As we discuss below this secondary gap can host additional Andreev (Sauls, 2018) reflection peaks, observable in metal-superconductor heterostructures.

We now explore how the additional sharp Lorentz-like features of the gap function affect the superconducting proximity effect. In order to describe the transmission and reflection of quasiparticles, we employ the Blonder-Tinkham-Klapwijk (BTK) formalism (Blonder et al., 1982). We consider a heterostructure consisting of normal and superconducting metals (NS). Following (Blonder et al., 1982), we consider the scattering of an incident electron off of the barrier. The strength of the proximity effect can be characterized by the probability for an electron to scatter into a hole-type excitation.

Within the BTK theory the boundary condition is given by:

ψN(0)\displaystyle\vec{\psi}_{N}\left(0\right) =ψS(0),\displaystyle=\vec{\psi}_{S}\left(0\right), (6)
z2mNψN(0)\displaystyle\frac{\partial_{z}}{2m_{N}}\vec{\psi}_{N}\left(0\right) =z2mSψS(0)+HψS(0).\displaystyle=\frac{\partial_{z}}{2m_{S}}\vec{\psi}_{S}\left(0\right)+H\vec{\psi}_{S}(0). (7)

where mSm_{S} and mNm_{N} are the effective electron masses and HH is the δ\delta-barrier height. Performing the analytic continuation and replacing iϵnω+i0+i\epsilon_{n}\rightarrow\omega+i0^{+}, the wavefunctions on superconducting side satisfy the equation G^𝐤1(ω+i0+)ψS=0\hat{G}_{{\bf k}}^{-1}(\omega+i0^{+})\vec{\psi}_{S}=0. On the normal side the equation is the same with the substitution ϕ(ω+i0+)0\phi(\omega+i0^{+})\rightarrow 0 and 𝒵(ω+i0+)1\mathcal{Z}(\omega+i0^{+})\rightarrow 1. In the following we do not explicitly write 0+0^{+} for shortness. The normal-state solution representing an incident electron and reflected electron and hole components is:

ψN=[10]eikez+[rN0]eikez+[0rA]eikhz,\vec{\psi}_{N}=\begin{bmatrix}1\\ 0\end{bmatrix}e^{ik_{\text{e}}z}+\begin{bmatrix}r_{N}\\ 0\end{bmatrix}e^{-ik_{\text{e}}z}+\begin{bmatrix}0\\ r_{A}\end{bmatrix}e^{ik_{\text{h}}z}, (8)

where rNr_{N} and rAr_{A} denote the reflection amplitude in the electron and hole channels respectively and the electron/hole momenta are given by ke/h=2m(μ±ω)k_{\text{e}/\text{h}}=\sqrt{2m\left(\mu\pm\omega\right)}. Analogously we find the solution for the quasi-electrons and quasi-holes propagating in the superconductor:

ψS\displaystyle\vec{\psi}_{S} Cqe[1η+]eikqez+Cqh[1η]eikqhz,\displaystyle\approx C_{\text{qe}}\begin{bmatrix}1\\ \eta_{+}\end{bmatrix}e^{ik_{\text{qe}}z}+C_{\text{qh}}\begin{bmatrix}1\\ \eta_{-}\end{bmatrix}e^{-ik_{\text{qh}}z}, (9)

where CqeC_{\text{qe}} and CqhC_{\text{qh}} are the corresponding amplitudes of the quasi-electron and quasi-holes and we denoted the corresponding coherence factors as η±=ϕ(ω)/(𝒵(ω)ω±𝒵2(ω)ω2ϕ2(ω))\eta_{\pm}=\phi(\omega)/\left(\mathcal{Z}(\omega)\omega\pm\sqrt{\mathcal{Z}^{2}(\omega)\omega^{2}-\phi^{2}(\omega)}\right). In the quasiclassical limit the quasielectron and quasihole momenta kqe,kqhk_{\text{qe}},k_{\text{qh}} can be taken to be equal to the corresponding Fermi momenta.

Upon solving the set of equations Eqs. (6-9) in the quasi-classical Belzig et al. (1999) limit and assuming mS=mNm_{S}=m_{N} we find the Andreev reflection coefficient to be:

rA\displaystyle r_{\text{A}} =ηη+η+Z2(ηη+),\displaystyle=\frac{\eta_{-}\eta_{+}}{\eta_{-}+Z^{2}\left(\eta_{-}-\eta_{+}\right)}, (10)
rN\displaystyle r_{N} =Z(i+Z)η+Z2(ηη+)(ηη+)\displaystyle=\frac{-Z\left(i+Z\right)}{\eta_{-}+Z^{2}\left(\eta_{-}-\eta_{+}\right)}\left(\eta_{-}-\eta_{+}\right) (11)

where the normalized barrier height is defined as ZmH/2mμZ\equiv mH/\sqrt{2m\mu}. We note that the conventional Andreev reflection can be obtained from Eq. (10) by simply assuming frequency-independent 𝒵\mathcal{Z} and ϕ\phi. In the limit of 𝒵=0\mathcal{Z}=0 the Andreev reflection coefficient is given by rA=η+r_{A}=\eta_{+}. We therefore find that in order to have a strong Andreev reflection the condition ϕ𝒵ω\phi\gg\mathcal{Z}\omega should be satisfied. The latter condition is always satisfied at very low frequencies leading to the conventional Andreev reflection (Sauls, 2018) result |rA|1\left|r_{\text{A}}\right|\approx 1. However it can also be satisfied at large frequencies if the frequency-dependent order parameter Δ(ω)ϕ(ω)/𝒵(ω)\Delta\left(\omega\right)\equiv\phi\left(\omega\right)/\mathcal{Z}\left(\omega\right) is larger than the frequency Δ(ω)ω\Delta\left(\omega\right)\gg\omega. In this case the reflection is up to a possible phase factor identical to the low-energy case. As shown in Fig. 2, this scenario is realized for the Lorentz-like order parameter introduced above in the frequency range close to the resonance ωω0\omega\sim\omega_{0}.

Finite-energy bound states –

We now consider the possibility of having the finite-energy Andreev bound states (Sauls, 2018) in the junction consisting of two superconductors separated by a metallic region with the two boundaries located at L/2L/2 and L/2-L/2. We assume the phase difference between two superconductors to be γ\gamma. In order to find the bound states we now follow the same procedure as for outlined above for the Andreev reflection but matching solutions at the two boundaries simultaneously. The general solution can be obtained analytically but it is too cumbersome and we therefore consider some simple limiting case. In particular as we demonstrate in the supplementary material, in the limit ϕ0ω0\phi_{0}\ll\omega_{0} and L0L\rightarrow 0 there are two additional in-gap bound states ω[ω0,ω0+ϕ0]\omega\in[\omega_{0},\omega_{0}+\phi_{0}] with the energies ωα\omega_{\alpha} (both positive and negative):

ωα=ω0+12{1+αcosγ2},\omega_{\alpha}=\omega_{0}+\frac{1}{2}\left\{1+\alpha\cos\frac{\gamma}{2}\right\}, (12)

with α=±\alpha=\pm. We thus find the energy of these bound states is of the order of the characteristic frequency of the order parameter frequency dependence. For both intrinsic and proximity superconductors ω0\omega_{0} can be expected to be of the order of several THz. This implies the existence of such bound states can potentially be probed by means of laser excitation. Their response is equivalent to a two-level system thus making them a good candidate for realization of solid state qubits.

Interpretation as "dark" resonance –

We now discuss the interpretation of the additional Andreev reflection peak in terms of the so-called “dark” resonance. The concept of dark resonance is extensively studied within the field of quantum optics (Lukin et al., 1999). It is based on the existence of “slowly”-evolving superposition states in a quantum system which are decoupled from the “fast” e.g. environment modes. For example, such optical phenomena as the Electromagnetically Induced Transparency (EIT) (Fleischhauer and Lukin, 2000) are based on the existence of a dark resonance in a driven three-level system. The on-shell BdG Green’s function Eq. (3), can be expressed as follows:

G^𝐤𝐤(ω)(ω+i0+)=(ωξ±(ω)0t00ω+ξ±(ω)0tt0ωω00tω0ω)\hat{G}_{{\bf k}\rightarrow{\bf k}(\omega)}^{\prime}\left(\omega+i0^{+}\right)=\left(\begin{array}[]{cccc}\omega-\xi_{\pm}\left(\omega\right)&0&t&0\\ 0&\omega+\xi_{\pm}\left(\omega\right)&0&-t\\ t&0&\omega&\omega_{0}\\ 0&-t&\omega_{0}&\omega\end{array}\right)

where ξ±(ω)=±((t2ω2)2ω2ω02)/(ω2ω02)\xi_{\pm}\left(\omega\right)=\pm\sqrt{\left(\left(t^{2}-\omega^{2}\right)^{2}-\omega^{2}\omega_{0}^{2}\right)/(\omega^{2}-\omega_{0}^{2})} is the on-shell quasi-electron and quasi-hole dispersions. By construction the Green function has the flat-band superconducting degree of freedom, which can be considered “slow.” By “dark” we thus define states, which have projection onto the auxiliary degrees of freedom only. Let us now find the quasi-electron and quasi-hole coherence vectors corresponding to the on-shell Green’s function Eq. (3, 4). The latter is schematically shown in Fig. 1 (c) in the limit of t0t\rightarrow 0. At frequencies ωω0\omega\approx\omega_{0} one of the eigenstates of the auxiliary degrees of freedom crosses ω=0\omega=0 therefore being degenerate with the electron branch. The corresponding eigenvector is readily found to be given by [0,0,1,1]T/2\approx[0,0,1,1]^{T}/\sqrt{2}. Thus this vector only has "slow" non-propagating (flat-band) components and are decoupled from the other degrees of freedom.

Refer to caption
Figure 3: Differential tunneling conductance as function of bias voltage (expressed in frequency units) for different values of the barrier height: Z=0Z=0 solid blue, Z=1Z=1 dashed orange and Z=3Z=3 dot-dashed green. Inset shows the same at low frequencies. The assumed electron-phonon coupling is λ=0.3\lambda=0.3 (see supplementary material). The normal-state resistance is denoted as RN=(Z2+1)/(2ν0e2vF𝒜)R_{N}=(Z^{2}+1)/(2\nu_{0}e^{2}v_{F}\mathcal{A}), where 𝒜\mathcal{A} is the contact surface area.
Einstein phonon model –

Finally we demonstrate that key results and conclusions of the toy model, involving the auxiliary superconductor, hold in the physical setup of interest, where the pairing interaction is induced by the optical phonon mode. More specifically we now consider the gap function induced by the interaction with the optical phonon mode with the propagator 𝒟𝐪(iνm)=2ω0/(ω02+νm2),\mathcal{D_{{\bf{q}}}}(i\nu_{m})=-2\omega_{0}/(\omega_{0}^{2}+\nu_{m}^{2}), where νm2πm/β\nu_{m}\equiv 2\pi m/\beta, mm\in\mathbb{Z}. As we demonstrate in the Supplementary material, the matrix-valued self-energy Σ^(iϵn)\hat{\Sigma}(i\epsilon_{n}) is found by solving the Dyson’s equation. The latter reduces to two coupled equations for the inverse quasiparticle residue 𝒵n\mathcal{Z}_{n} and the gap function ϕn\phi_{n} in case when the phonon propagator is momentum-independent: Σ^(iϵn)=(1𝒵n)iϵnτ^0+ϕnτ^1\hat{\Sigma}(i\epsilon_{n})=(1-\mathcal{Z}_{n})i\epsilon_{n}\hat{\tau}_{0}+\phi_{n}\hat{\tau}_{1}. The approximate analytical form of the phonon-induced self-energy can be obtained in the limit of weak electron-phonon coupling. As was shown in (Marsiglio, 2018; Mirabi et al., 2020), the imaginary-axis dependence of the order parameter has the Lorentz form ϕnω02/(ω02+ϵn2)\phi_{n}\propto\omega_{0}^{2}/(\omega_{0}^{2}+\epsilon_{n}^{2}) and 𝒵n1{\cal Z}_{n}\approx 1. This is thus in agreement with the assumed gap-frequency behavior in Eq. (1). At finite electron-phonon coupling strength, the gap frequency dependence deviates from the Lorentz form and additional resonances at frequencies nω0n\omega_{0} with n=1,2,3n=1,2,3\ldots (Mirabi et al., 2020) emerge. However, the sharp features of the gap function, reminiscent to the pure Lorentz case, remain even at finite coupling strength Marsiglio (2020); Mirabi et al. (2020); Marsiglio (2018). We now study how these features affect the Andreev reflection from the NS boundary. More precisely, we consider the differential tunneling conductance which expresses through the reflection coefficients Eqs. (10, 11) as dI/dV1+|rA|2|rN|2dI/dV\propto 1+|r_{A}|^{2}-|r_{N}|^{2} Blonder et al. (1982). Both reflection coefficients are found numerically by solving the complete set of Migdal-Eliashberg equations. The result of numerical calculation is shown in Fig. 3 for different values of the barrier height ZZ. We find the prominent additional reflection peaks close to the Debye frequency, which correspond to the dark Andreev resonances. Note that the exact solution for the phonon case involves imaginary self-energy contributions, which give rise to a finite life-time of the superconducting quasiparticles (absent in the toy model). However, these complications do not appear to affect the qualitative picture and the signatures of the dark Andreev states are preserved.

Conclusions & outlook –

In this work, we studied the quasiparticle properties of a superconductor with a frequency-dependent order parameter. When the latter has resonant features, we find an additional depletion of the high-energy density of states. We provide a physically equivalent two-band picture with an additional superconducting band gap emerging at high energies. For the NS junction, the band gap leads to additional Andreev reflection peaks at high energies. In the case of an SNS junction, we find Andreev bound states within the high-energy band gap. We provide an interpretation of these phenomena in terms of the “dark” resonance of the BdG Hamiltonian. We expect that the predicted phenomena should be accessible in experiment in superconductors with optical-phonon-mediated pairing (for example in MgB2MgB_{2} Mazin et al. (2002), K3C60K_{3}C_{60} Zhang et al. (1991), etc) and also proximity systems involving flat-band materials Balents et al. (2020). This work also suggests a number of follow up ideas, at the intersection of superconductivity and quantum optics: e.g., the possibility of control of the dark Andreev states by external laser driving. Furthermore, it would be interesting to explore the role of frequency dependence of the order parameter on the quasiparticle spectrum in topological superconductors and whether dark Majorana-like states are possible.

Acknowledgements –

This work was supported by the National Science Foundation under Grant No. DMR-2037158, the U.S. Army Research Office under Contract No. W911NF1310172, and the Simons Foundation. The authors are grateful to Jay Sau for an illuminating discussion.

References

  • Bardeen et al. (1957) J. Bardeen, L. N. Cooper,  and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957).
  • Altland and Simons (2010) A. Altland and B. D. Simons, Condensed matter field theory (Cambridge university press, 2010).
  • Eliashberg (1960) G. Eliashberg, Sov. Phys. JETP 11, 696 (1960).
  • Marsiglio (2020) F. Marsiglio, Annals of Physics 417, 168102 (2020), eliashberg theory at 60: Strong-coupling superconductivity and beyond.
  • Marsiglio et al. (1988) F. Marsiglio, M. Schossmann,  and J. P. Carbotte, Phys. Rev. B 37, 4965 (1988).
  • Marsiglio (2018) F. Marsiglio, Physical Review B 98, 024523 (2018).
  • Mirabi et al. (2020) S. Mirabi, R. Boyack,  and F. Marsiglio, Physical Review B 101, 064506 (2020).
  • Christensen and Chubukov (2021) M. H. Christensen and A. V. Chubukov, Physical Review B 104, L140501 (2021).
  • Fleischhauer and Lukin (2000) M. Fleischhauer and M. D. Lukin, Phys. Rev. Lett. 84, 5094 (2000).
  • Lukin et al. (1999) M. D. Lukin, S. F. Yelin, M. Fleischhauer,  and M. O. Scully, Phys. Rev. A 60, 3225 (1999).
  • Fleischhauer and Lukin (2002) M. Fleischhauer and M. D. Lukin, Physical Review A 65, 022314 (2002).
  • Schrieffer (2018) J. R. Schrieffer, Theory of superconductivity (CRC press, 2018).
  • Chubukov et al. (2020) A. V. Chubukov, A. Abanov, I. Esterlis,  and S. A. Kivelson, Annals of Physics 417, 168190 (2020).
  • Liu et al. (2019) C.-X. Liu, J. D. Sau, T. D. Stanescu,  and S. Das Sarma, Phys. Rev. B 99, 024510 (2019).
  • Sauls (2018) J. Sauls, “Andreev bound states and their signatures,”  (2018).
  • Blonder et al. (1982) G. E. Blonder, M. Tinkham,  and T. M. Klapwijk, Phys. Rev. B 25, 4515 (1982).
  • Belzig et al. (1999) W. Belzig, F. K. Wilhelm, C. Bruder, G. Schön,  and A. D. Zaikin, Superlattices and microstructures 25, 1251 (1999).
  • Mazin et al. (2002) I. Mazin, O. Andersen, O. Jepsen, O. Dolgov, J. Kortus, A. A. Golubov, A. Kuz’menko,  and D. Van Der Marel, Physical review letters 89, 107002 (2002).
  • Zhang et al. (1991) F. C. Zhang, M. Ogata,  and T. M. Rice, Phys. Rev. Lett. 67, 3452 (1991).
  • Balents et al. (2020) L. Balents, C. R. Dean, D. K. Efetov,  and A. F. Young, Nature Physics 16, 725 (2020).

Supplemental Material for:
Dark Andreev States in Superconductors

Andrey Grankin, Victor Galitski

Joint Quantum Institute, Department of Physics, University of Maryland, College Park, MD 20742, USA

I Derivation of the Green’s function in the toy model

In this section we trace-over the auxiliary fermionic degrees of freedom and derive the Green’s function of the superconductor. We start with the BDG Hamiltonian:

HBdG=\displaystyle H_{\text{BdG}}= ξkτ^3(σ^3+1^)2ω0τ^1(1^σ^3)2+tτ^3σ^1,\displaystyle\xi_{k}\hat{\tau}_{3}\frac{\left(\hat{\sigma}_{3}+\hat{1}\right)}{2}-\omega_{0}\hat{\tau}_{1}\frac{\left(\hat{1}-\hat{\sigma}_{3}\right)}{2}+t\hat{\tau}_{3}\hat{\sigma}_{1}, (13)

The quantum-mechanical action corresponding to Eq. (13) reads:

S=nΨ~𝐤𝒢𝐤1(iϵn)Ψ~𝐤,S=-\sum_{n}\tilde{\Psi}_{{\bf k}}^{\dagger}{\cal G}_{{\bf k}}^{{}^{\prime}-1}(i\epsilon_{n})\tilde{\Psi}_{{\bf k}}, (14)

where Ψ~\tilde{\Psi} is the extended Bogolyubov spinor and 𝒢𝐤1(iϵn)=iϵnHBdG{\cal G}_{{\bf k}}^{\prime-1}(i\epsilon_{n})=i\epsilon_{n}-H_{\text{BdG}}. The partitioning function is related to Eq. (14) Z=𝑑eSZ=\int d\ldots e^{-S}. We now perform the Gaussian integral over the auxiliary degrees of freedom parametrized by the projector Q(1^σ^3)/2Q\equiv\left(\hat{1}-\hat{\sigma}_{3}\right)/2. The projector onto the remaining fermionic degrees of freedom reads 𝒫=𝒬{\cal P}={\cal I}-{\cal Q}. The Green’s function is straightforwardly found to be:

𝒢𝐤1(iϵn)\displaystyle{\cal G}_{{\bf k}}^{-1}\left(i\epsilon_{n}\right) =iϵnξkτ^3t2τ^31iϵn+ω0τ^1τ^3\displaystyle=i\epsilon_{n}-\xi_{k}\hat{\tau}_{3}-t^{2}\hat{\tau}_{3}\frac{1}{i\epsilon_{n}+\omega_{0}\hat{\tau}_{1}}\hat{\tau}_{3}
=iϵnξkτ^3+t2τ^3iϵnω0τ^1ϵn2+ω02τ^3\displaystyle=i\epsilon_{n}-\xi_{k}\hat{\tau}_{3}+t^{2}\hat{\tau}_{3}\frac{i\epsilon_{n}-\omega_{0}\hat{\tau}_{1}}{\epsilon_{n}^{2}+\omega_{0}^{2}}\hat{\tau}_{3}
=iϵnξkτ^3+t2iϵn+ω0τ1ϵn2+ω02\displaystyle=i\epsilon_{n}-\xi_{k}\hat{\tau}_{3}+t^{2}\frac{i\epsilon_{n}+\omega_{0}\tau_{1}}{\epsilon_{n}^{2}+\omega_{0}^{2}}

We thus find the additional normal- and anomalous self-energies t2iϵn/(ϵn2+ω02),-t^{2}i\epsilon_{n}/(\epsilon_{n}^{2}+\omega_{0}^{2}), t2ω0/(ϵn2+ω02)-t^{2}\omega_{0}/(\epsilon_{n}^{2}+\omega_{0}^{2}). Using tϕ0ω0t\equiv\sqrt{\phi_{0}\omega_{0}} as discussed in the main text we find

𝒵n=(1+ϕ0ω0ϵn2+ω02).{\cal Z}_{n}=\left(1+\frac{\phi_{0}\omega_{0}}{\epsilon_{n}^{2}+\omega_{0}^{2}}\right).

I.1 Instability in the absence of 𝒵n{\cal Z}_{n}

Let us now consider a superconductor with the frequency-dependent order parameter and no renormalization of the normal-state self energy:

𝒢^𝐤1(iεn)=iεnτ^0ϕ0ω02ϵn2+ω02τ^1ξkτ^3,\hat{{\cal G}}_{{\bf k}}^{-1}\left(i\varepsilon_{n}\right)=i\varepsilon_{n}\hat{\tau}_{0}-\phi_{0}\frac{\omega_{0}^{2}}{\epsilon_{n}^{2}+\omega_{0}^{2}}\hat{\tau}_{1}-\xi_{k}\hat{\tau}_{3}, (15)

We now find the spectrum of this Greens function and demonstrate that it is unstable.

G^𝐤1(ω+i0+)=(ω+i0+)τ^0ϕ0ω02ω02(ω+i0+)2τ^1ξkτ^3\hat{G}_{{\bf k}}^{-1}\left(\omega+i0^{+}\right)=\left(\omega+i0^{+}\right)\hat{\tau}_{0}-\phi_{0}\frac{\omega_{0}^{2}}{\omega_{0}^{2}-\left(\omega+i0^{+}\right)^{2}}\hat{\tau}_{1}-\xi_{k}\hat{\tau}_{3}
detG^𝐤1(ω+i0+)=0\det\hat{G}_{{\bf k}}^{-1}\left(\omega+i0^{+}\right)=0

The eigenvalues ωi\omega_{i} are shown in Fig. 4. We find that that the spectrum becomes complex with both positive and negative imaginary values. This corresponds to an unstable system. We thus conclude that the frequency-dependence of the gap implies a presence of a non-zero quasiparticle residue function 𝒵n\mathcal{Z}_{n}.

Refer to caption
Figure 4: Real (blue) and imaginary (orange) parts of the eigenvalues of the BdG Green’s function Eq. (I.1). The non-vanishing imaginary parts correspond to the eigenvalues merging close to the ω0\omega_{0} frequency.

II Derivation of the gap function at weak coupling

Dyson equation for the matrix-valued self energy reads:

Σ^𝐤(iεn)=g2Vβ𝐪,m𝒟𝐤𝐤(iεniϵn){τ^3G^𝐤(iϵn)τ^3}\hat{\Sigma}_{{\bf k}}(i\varepsilon_{n})=-\frac{g^{2}}{V\beta}\sum_{{\bf q},m}{\cal D}_{{\bf k-k^{\prime}}}(i\varepsilon_{n}-i\epsilon_{n^{\prime}})\left\{\hat{\tau}_{3}\hat{G}_{{\bf k^{\prime}}}(i\epsilon_{n^{\prime}})\hat{\tau}_{3}\right\} (16)

The propagator of the Einstein phonon mode reads

𝒟𝐤𝐤(iεniϵn)=2ω0ω02(iεniϵn)2,{\cal D}_{{\bf k-k^{\prime}}}(i\varepsilon_{n}-i\epsilon_{n^{\prime}})=\frac{-2\omega_{0}}{\omega_{0}^{2}-(i\varepsilon_{n}-i\epsilon_{n^{\prime}})^{2}},

where ω0\omega_{0} is the Debye frequency. Taking into account the fact that the phonon propagator does not depend on momentum we can simplify Eq. (16):

Σ^(iεn)=g2ν0βm𝒟(iεniϵn)𝑑ξk{τ^3G^𝐤(iϵn)τ^3}.\hat{\Sigma}(i\varepsilon_{n})=-\frac{g^{2}\nu_{0}}{\beta}\sum_{m}{\cal D}(i\varepsilon_{n}-i\epsilon_{n^{\prime}})\int_{-\infty}^{\infty}d\xi_{k^{\prime}}\left\{\hat{\tau}_{3}\hat{G}_{{\bf k^{\prime}}}(i\epsilon_{n^{\prime}})\hat{\tau}_{3}\right\}. (17)

The momentum integral 𝑑ξk\int d\xi_{k^{\prime}} can be taken analytically by employing the Green’s function ansatz Eq. (1).

II.0.1 Weak-coupling result

Let us now explicitly find the gap function. Let us first write the complete set of Eliashberg equations as obtained from Eq. (17):

Zn\displaystyle Z_{n} =1+λβmω02ω02+(εnϵn)2πϵnεnZmϵn2Zm2+ϕm2\displaystyle=1+\frac{\lambda}{\beta}\sum_{m}\frac{\omega_{0}^{2}}{\omega_{0}^{2}+\left(\varepsilon_{n}-\epsilon_{n^{\prime}}\right)^{2}}\frac{\pi\frac{\epsilon_{n^{\prime}}}{\varepsilon_{n}}Z_{m}}{\sqrt{\epsilon_{n^{\prime}}^{2}Z_{m}^{2}+\phi_{m}^{2}}} (18)
ΔnZn\displaystyle\Delta_{n}Z_{n} =λβmω02ω02+(εnϵn)2πΔmϵn2+Δm2.\displaystyle=\frac{\lambda}{\beta}\sum_{m}\frac{\omega_{0}^{2}}{\omega_{0}^{2}+\left(\varepsilon_{n}-\epsilon_{n^{\prime}}\right)^{2}}\frac{\pi\Delta_{m}}{\sqrt{\epsilon_{n^{\prime}}^{2}+\Delta_{m}^{2}}}. (19)

where we defined Δmϕm/Zm\Delta_{m}\equiv\phi_{m}/Z_{m}. At very weak coupling we can neglect the non-linearity of both equations or equivalently set ϵn2+Δm2|ϵn|\sqrt{\epsilon_{n^{\prime}}^{2}+\Delta_{m}^{2}}\approx\left|\epsilon_{n^{\prime}}\right|. In this case Eq. (19) has logarithmic singularity on the righthand side. Following (Marsiglio, 2018) we can separate singular and regular terms in a straightforward way:

ΔnZn\displaystyle\Delta_{n}Z_{n} =λβm(ω02ω02+(εnϵn)2ω02ω02+(εn)2)πΔm|ϵn|\displaystyle=\frac{\lambda}{\beta}\sum_{m}\left(\frac{\omega_{0}^{2}}{\omega_{0}^{2}+\left(\varepsilon_{n}-\epsilon_{n^{\prime}}\right)^{2}}-\frac{\omega_{0}^{2}}{\omega_{0}^{2}+\left(\varepsilon_{n}\right)^{2}}\right)\frac{\pi\Delta_{m}}{\left|\epsilon_{n^{\prime}}\right|}
+λβm(ω02ω02+(εn)2)πΔm|ϵn|\displaystyle+\frac{\lambda}{\beta}\sum_{m}\left(\frac{\omega_{0}^{2}}{\omega_{0}^{2}+\left(\varepsilon_{n}\right)^{2}}\right)\frac{\pi\Delta_{m}}{\left|\epsilon_{n^{\prime}}\right|}

The first term to the righthand side is now regular while the second one is singular. It is therefore leading at very weak coupling λ0\lambda\rightarrow 0. In this case we get:

Δnω02ω02+(εn)2\Delta_{n}\propto\frac{\omega_{0}^{2}}{\omega_{0}^{2}+\left(\varepsilon_{n}\right)^{2}}

We note that in the limit λ0\lambda\rightarrow 0 the anomalous self-energy has the same functional formMirabi et al. (2020); Marsiglio (2018). Analytical continuation requires some caution and the result will be shown in the section below.

II.1 Analytic continuation

Following the formulation derived in Marsiglio (2018); Mirabi et al. (2020); Marsiglio et al. (1988) we now perform the so-called iterative analytic continuation procedure of the self-energy Eq. (16). Let us now use the spectral decomposition of fermionic and bosonic

𝒟(iεniϵn)\displaystyle{\cal D}\left(i\varepsilon_{n}-i\epsilon_{n^{\prime}}\right) =1π𝑑xIm{DR(x)}x(iεniϵn),\displaystyle=\frac{1}{\pi}\int dx\frac{\text{Im}\left\{D^{R}\left(x\right)\right\}}{x-(i\varepsilon_{n}-i\epsilon_{n^{\prime}})}, (20)
𝒢^𝐤(iϵn)\displaystyle\hat{{\cal G}}_{{\bf k}}(i\epsilon_{n^{\prime}}) =1π𝑑yIm{G^𝐤R(y)}yiϵn.\displaystyle=\frac{1}{\pi}\int dy\frac{\text{Im}\left\{\hat{G}_{{\bf k}}^{R}\left(y\right)\right\}}{y-i\epsilon_{n^{\prime}}}. (21)

where DRD^{R} and G^R\hat{G}^{R} stand for the retarded Green’s functions. By substituting these expressions into Eq. (16) and taking explicitly the Matsubara-frequency sum over ϵn\epsilon_{n^{\prime}} we find:

Σ^(iεn)=\displaystyle\hat{\Sigma}(i\varepsilon_{n})=
g2ν0dξkπ2𝑑x𝑑yIm{DR(x)}Im{τ^3G^𝐤R(y)τ^3}yiεnx(nx+fy),\displaystyle-g^{2}\nu_{0}\int_{-\infty}^{\infty}\frac{d\xi_{k}}{\pi^{2}}\int dxdy\frac{\text{Im}\left\{D^{R}\left(x\right)\right\}\text{Im}\left\{\hat{\tau}_{3}\hat{G}_{{\bf k}}^{R}\left(y\right)\hat{\tau}_{3}\right\}}{y-i\varepsilon_{n}-x}\left(n_{x}+f_{y}\right),

where nxn_{x} and fyf_{y} are respectively bosonic and fermionic occupation numbers. Analytic continuation is readily performed by replacing iεnω+i0+i\varepsilon_{n}\rightarrow\omega+i0^{+}.

Σ^(ω+i0+)=\displaystyle\hat{\Sigma}\left(\omega+i0^{+}\right)=
g2ν0dξkπ2𝑑x𝑑yIm{DR(x)}Im{τ^3G^𝐤R(y)τ^3}yxωi0+(nx+fy)\displaystyle-g^{2}\nu_{0}\int_{-\infty}^{\infty}\frac{d\xi_{k}}{\pi^{2}}\int dxdy\frac{\text{Im}\left\{D^{R}\left(x\right)\right\}\text{Im}\left\{\hat{\tau}_{3}\hat{G}_{{\bf k}}^{R}\left(y\right)\hat{\tau}_{3}\right\}}{y-x-\omega-i0^{+}}\left(n_{x}+f_{y}\right)

The integral over yy can now be taken explicitly. For that we use the following facts: retarded Green’s function only has poles in the lower part of the complex plane.

Σ^(ω+i0+)=\displaystyle\hat{\Sigma}\left(\omega+i0^{+}\right)=
g2ν0dξkπ𝑑xIm{DR(x)}{τ^3G^𝐤R(x+ω)τ^3}(nx+fx+ω)\displaystyle-g^{2}\nu_{0}\int_{-\infty}^{\infty}\frac{d\xi_{k}}{\pi}\int dx\text{Im}\left\{D^{R}\left(x\right)\right\}\left\{\hat{\tau}_{3}\hat{G}_{{\bf k}}^{R}\left(x+\omega\right)\hat{\tau}_{3}\right\}\left(n_{x}+f_{x+\omega}\right)
g2ν0βn𝑑ξk𝒟(ωiεn)τ^3𝒢^𝐤(iεn)τ^3\displaystyle-\frac{g^{2}\nu_{0}}{\beta}\sum_{n}\int_{-\infty}^{\infty}d\xi_{k}{\cal D}\left(\omega-i\varepsilon_{n}\right)\hat{\tau}_{3}\hat{{\cal G}}_{{\bf k}}\left(i\varepsilon_{n}\right)\hat{\tau}_{3}

The right-hand side of this expression contains both the Matsubara and the real-frequency Green’s function. We can now estimate the gap-frequency function in the limit of very weak coupling. We will be interested in the following quantity (see main text) Δ(ω+i0+)ϕ(ω+i0+)/Z(ω+i0+)\Delta\left(\omega+i0^{+}\right)\equiv\phi\left(\omega+i0^{+}\right)/Z\left(\omega+i0^{+}\right).

Let us now derive the real-axis λ0\lambda\rightarrow 0 expression for the anomalous self-energy. By taking into account Im{DR(x)}=π(δ(ω+ω0)δ(ωω0))\text{Im}\left\{D^{R}\left(x\right)\right\}=\pi\left(\delta\left(\omega+\omega_{0}\right)-\delta\left(\omega-\omega_{0}\right)\right) at low we find:

Σ^(ω+i0+)=\displaystyle\hat{\Sigma}\left(\omega+i0^{+}\right)=
+g2ν0𝑑ξk{τ^3G^𝐤R(ω0+ω)τ^3}(nω0+fω0+ω)\displaystyle+g^{2}\nu_{0}\int_{-\infty}^{\infty}d\xi_{k}\left\{\hat{\tau}_{3}\hat{G}_{{\bf k}}^{R}\left(\omega_{0}+\omega\right)\hat{\tau}_{3}\right\}\left(n_{\omega_{0}}+f_{\omega_{0}+\omega}\right)
g2ν0𝑑ξk{τ^3G^𝐤R(ωω0)τ^3}(nω0+fω0+ω)\displaystyle-g^{2}\nu_{0}\int_{-\infty}^{\infty}d\xi_{k}\left\{\hat{\tau}_{3}\hat{G}_{{\bf k}}^{R}\left(\omega-\omega_{0}\right)\hat{\tau}_{3}\right\}\left(n_{-\omega_{0}}+f_{-\omega_{0}+\omega}\right)
g2ν0βn𝑑ξ𝒟(ωiεn)τ^3𝒢^𝐤(iεn)τ^3\displaystyle-\frac{g^{2}\nu_{0}}{\beta}\sum_{n}\int_{-\infty}^{\infty}d\xi{\cal D}\left(\omega-i\varepsilon_{n}\right)\hat{\tau}_{3}\hat{{\cal G}}_{{\bf k}}\left(i\varepsilon_{n}\right)\hat{\tau}_{3}

The first term is exponentially suppressed at Tω0T\ll\omega_{0}. The second term is non-singular and moreover in the T0T\rightarrow 0 limit it only has contributes at ω>ω0\omega>\omega_{0}. The logarithmically-divergent contribution in the λ0\lambda\rightarrow 0 is given by the second term. In this limit the anomalous self-energy obeys:

ϕ(ω+i0+)\displaystyle\phi(\omega+i0^{+}) g2ν0βn𝑑ξk𝒟(ωiεn)ϕnεn2𝒵n2+ϕn2+ξk2\displaystyle\approx-\frac{g^{2}\nu_{0}}{\beta}\sum_{n}\int_{-\infty}^{\infty}d\xi_{k}{\cal D}\left(\omega-i\varepsilon_{n}\right)\frac{\phi_{n}}{\varepsilon_{n}^{2}\mathcal{Z}_{n}^{2}+\phi_{n}^{2}+\xi_{k}^{2}}
=λπdωn2πω02ω02(ωiϵn)2ϕnεn2𝒵n2+ϕn2\displaystyle=\lambda\pi\int\frac{d\omega_{n}}{2\pi}\frac{\omega_{0}^{2}}{\omega_{0}^{2}-(\omega-i\epsilon_{n})^{2}}\frac{\phi_{n}}{\sqrt{\varepsilon_{n}^{2}\mathcal{Z}_{n}^{2}+\phi_{n}^{2}}}

Under the same assumptions as in Sec. (II.0.1) we find:

ϕ(ω+i0+)ω02ω02(ω+i0+)2.\phi(\omega+i0^{+})\propto\frac{\omega_{0}^{2}}{\omega_{0}^{2}-(\omega+i0^{+})^{2}}.

II.2 Normal state self-energy in the weak-coupling limit

We now consider the diagonal terms of the normal-state self-energy in the limit of weak coupling.

II.2.1 Real-frequency

Σ(ω+i0+)=\displaystyle\Sigma\left(\omega+i0^{+}\right)=
2πig2ν0θ(ωω0)ig2ν0𝑑εn2ω0sign(ϵn)ω02(ωiεn)2\displaystyle-2\pi ig^{2}\nu_{0}\theta\left(\omega-\omega_{0}\right)-ig^{2}\nu_{0}\int d\varepsilon_{n}\frac{2\omega_{0}\text{sign}(\epsilon_{n})}{\omega_{0}^{2}-\left(\omega-i\varepsilon_{n}\right)^{2}}
=2πig2ν0θ(ωω0)ig2ν0(4iarctanh(ωω0)+2πθ(ωω0))\displaystyle=-2\pi ig^{2}\nu_{0}\theta\left(\omega-\omega_{0}\right)-ig^{2}\nu_{0}\left(-4i\text{arctanh}\left(\frac{\omega}{\omega_{0}}\right)+2\pi\theta\left(\omega-\omega_{0}\right)\right)
=πiλω0θ(ωω0)λω0arctanh(ωω0)\displaystyle=-\pi i\lambda\omega_{0}\theta\left(\omega-\omega_{0}\right)-\lambda\omega_{0}\text{arctanh}\left(\frac{\omega}{\omega_{0}}\right)

III High-energy Andreev Bound states

We now derive the energy of Andreev bound states appearing at high energy due to the additional band gap. Our consideration is essentially very similar to the derivation of the Andreev reflection coefficient in the main text. In particular we now respectively denote the wave functions of the right and left superconductors as ψS,R\vec{\psi}_{S,R} and ψS,L\vec{\psi}_{S,L}. Solving the equation of motion in all three media we find :

ψS,R\displaystyle\vec{\psi}_{\text{S,R}} =[1eiγϕ(ω)𝒵(ω)ω+Z2(ω)ω2ϕ2(ω)]CqeReikqez+[1eiγϕ(ω)𝒵(ω)ωZ2(ω)ω2ϕ2(ω)]CqhReikqhz,\displaystyle=\begin{bmatrix}1\\ \frac{e^{i\gamma}\phi(\omega)}{\mathcal{Z}(\omega)\omega+\sqrt{Z^{2}(\omega)\omega^{2}-\phi^{2}(\omega)}}\end{bmatrix}C_{\text{qe}}^{R}e^{ik_{\text{qe}}z}+\begin{bmatrix}1\\ \frac{e^{i\gamma}\phi(\omega)}{\mathcal{Z}(\omega)\omega-\sqrt{Z^{2}(\omega)\omega^{2}-\phi^{2}(\omega)}}\end{bmatrix}C_{\text{qh}}^{R}e^{-ik_{\text{qh}}z}, (22)
ψS,L\displaystyle\vec{\psi}_{\text{S,L}} =[1ϕ(ω)𝒵(ω)ω+Z2(ω)ω2ϕ2(ω)]CqeLeikqez+[1ϕ(ω)𝒵(ω)ωZ2(ω)ω2ϕ2(ω)]CqhLeikqhz,\displaystyle=\begin{bmatrix}1\\ \frac{\phi(\omega)}{\mathcal{Z}(\omega)\omega+\sqrt{Z^{2}(\omega)\omega^{2}-\phi^{2}(\omega)}}\end{bmatrix}C_{\text{qe}}^{L}e^{-ik_{\text{qe}}z}+\begin{bmatrix}1\\ \frac{\phi(\omega)}{\mathcal{Z}(\omega)\omega-\sqrt{Z^{2}(\omega)\omega^{2}-\phi^{2}(\omega)}}\end{bmatrix}C_{\text{qh}}^{L}e^{ik_{\text{qh}}z}, (23)
ψN\displaystyle\vec{\psi}_{\text{N}} =[αN+0]eikez+[αN0]eikez+[0β+]eikhz+[0β]eikhz.\displaystyle=\begin{bmatrix}\alpha_{N}^{+}\\ 0\end{bmatrix}e^{ik_{\text{e}}z}+\begin{bmatrix}\alpha_{N}^{-}\\ 0\end{bmatrix}e^{-ik_{\text{e}}z}+\begin{bmatrix}0\\ \beta^{+}\end{bmatrix}e^{-ik_{\text{h}}z}+\begin{bmatrix}0\\ \beta^{-}\end{bmatrix}e^{ik_{\text{h}}z}. (24)

In Eq. (22) we introduced an additional global phase factor eiγe^{i\gamma} due to the possible global phase difference between the two superconductors. The problem can be solved numerically or in the limit when the thickness of the normal metal is very thin L0L\rightarrow 0. In this case we just match boundaries between two superconductors:

ψS,L(0)\displaystyle\vec{\psi}_{\text{S,L}}\left(0\right) ψS,R(0)\displaystyle\approx\vec{\psi}_{\text{S,R}}\left(0\right) (25)
zψS,L(0)\displaystyle\partial_{z}\vec{\psi}_{\text{S,L}}\left(0\right) zψS,R(0)\displaystyle\approx\partial_{z}\psi_{\text{S,R}}\left(0\right) (26)

These two equations define the spectrum of the bound states. The bound state energy ωα\omega_{\alpha} is readily obtained from the equation:

ϕ2(ωα)Z(ωα)2ωα2+ϕ2(ωα)cosγ=0\phi^{2}\left(\omega_{\alpha}\right)-Z\left(\omega_{\alpha}\right)^{2}\omega_{\alpha}^{2}+\phi^{2}(\omega_{\alpha})\cos\gamma=0

By using the Lorentz-form expressions of ϕ\phi and 𝒵\mathcal{Z} in the limit ϕ0ω0\phi_{0}\ll\omega_{0} we find the bound state energies Eq. (9) of the main text.

IV Spectral function

We now study the spectral function of the Green’s function. It is defined as:

V Density of states calculation

In this section we consider the density of states of the superconductor with the BdG Green’s function

G^k1(ω+i0+)=(ω+i0+)4{ξk0t00ξk0tt0ξk/αω00tω0ξk/α},\hat{G}_{k}^{-1}\left(\omega+i0^{+}\right)=(\omega+i0^{+}){\cal I}_{4}-\begin{Bmatrix}\xi_{k}&0&t&0\\ 0&-\xi_{k}&0&-t\\ t&0&\xi_{k}/\alpha&\omega_{0}\\ 0&-t&\omega_{0}&-\xi_{k}/\alpha\end{Bmatrix}, (27)

This Green’s function is almost the same as Eq. 3 in the main text except for instead of the flat-band superconductor we take one with the dispersion ξk/α\xi_{k}/\alpha where α\alpha denotes the mass ratio parameter. The evolution of density of states as function of α\alpha is shown in Fig. 5. We therefore observe a significant density of states depletion for α3\alpha\gg 3.

Refer to caption
Figure 5: Local density of states in Eq. 27 as function of the mass ratio parameter α\alpha.