We extend the scattering theory of the Josephson effect to include a coupling of the Josephson junction to a gapless electron reservoir in the normal state. By opening up the system with a quasiparticle escape rate 1/τ, the supercurrent carried at zero temperature by an Andreev level at energy εA is reduced by a factor (2/π)arctan(2εAτ/). We make contact with recent work on “non-Hermitian Josephson junctions,” by comparing this result to different proposed generalizations of the Josephson effect to non-Hermitian Hamiltonians.

The thermodynamic properties of a quantum dot are affected by the coupling to a superconductor. The Josephson effect is a striking example, and a current will flow through the quantum dot in equilibrium if the quantum dot forms a weak link between two superconductors.1 This supercurrent I(ϕ) depends periodically on the phase difference ϕ of the pair potential in the superconductors. A dissipative electromagnetic environment may degrade the supercurrent via phase fluctuations.2–4 Here, we investigate an altogether different decay mechanism, the coupling of the quantum dot to a gapless electron reservoir (see Fig. 1).

FIG. 1.

Josephson junction, formed by a quantum dot that is tunnel coupled with rates Γ1 and Γ2 to two superconductors (gap Δ0 and phase difference ϕ). The quantum dot has an additional weak coupling (rate γ) to a gapless electron reservoir in the normal state. We seek the γ-dependence of the current–phase relationship I(ϕ).

FIG. 1.

Josephson junction, formed by a quantum dot that is tunnel coupled with rates Γ1 and Γ2 to two superconductors (gap Δ0 and phase difference ϕ). The quantum dot has an additional weak coupling (rate γ) to a gapless electron reservoir in the normal state. We seek the γ-dependence of the current–phase relationship I(ϕ).

Close modal

The reservoir is a source of dephasing because quasiparticles that enter it from the quantum dot return without any phase coherence. Such a dephasing mechanism for persistent currents was introduced by Büttiker5 and applied to superconducting circuits by several authors.6–13 Our main advance is that we obtain closed-form expressions for the supercurrent–phase relationship, for the case of a spatially uniform coupling rate to the electron reservoir.

Because the effective Hamiltonian of the quantum dot becomes non-Hermitian when we open it up to the reservoir, such a system is a simple example of a “non-Hermitian Josephson junction,” a topic of current interest.14–17 Our explicit expressions support Ref. 17 in the debate over the proper generalization of the current–phase relationship IdE/dϕ to complex eigenvalues E of the effective Hamiltonian.

In a scattering formulation,18 the dependence of the density of states ρ of the Josephson junction on the superconducting phase difference ϕ is given by the determinantal expression,
(1)
with ρ0 the ϕ-independent density of states of the junction when it is decoupled from the superconductors. The determinant contains the product of the Andreev reflection matrix RA from the superconductors and the scattering matrix SN of the junction in the normal state. The energy ε>0 is the excitation energy of Bogoliubov quasiparticles (electron–hole superpositions). The pair potential in the two superconductors, to the left and to the right of the junction, has amplitude Δ0 and phase ±ϕ/2.
The electron and hole degree of freedom introduces a block structure in the scattering matrices. The matrix SN is block diagonal,
(2)
the electron and hole blocks are uncoupled and related by particle-hole symmetry.
The Andreev reflection matrix has the block structure,
(3)
The Pauli matrix σy acts on the spin degree of freedom. The blocks rA and rA describe Andreev reflection from, respectively, electron to hole and hole to electron, in opposite spin bands. Andreev reflection happens with unit probability for energies ε<Δ0; any normal reflection at the normal–superconductor (NS) interface is incorporated into SN.
Substitution of Eqs. (2) and (3) in Eq. (1) gives the determinant,18 
(4)
(5)
The density of states determines the free energy F of the Josephson junction at temperature T,19 
(6)
The supercurrent through the Josephson junction then follows from the following relation:
(7)
It is convenient to extend the integration range in Eq. (6) to negative ε (so we no longer need to extract the imaginary part) and then to perform a partial integration,
(8)
with F0 the ϕ-independent contribution to the free energy.

We open up the Josephson junction by weakly coupling it to a gapless electron reservoir in the normal state. Quasiparticles enter the reservoir at a rate 1/τ2γ/, assumed to be spatially uniform in the junction region. This can be modeled by coupling to the reservoir via a spatially extended tunnel barrier.20,21 The reservoir is closed, and it does not drain any current, but particles that enter it return to the junction without any phase coherence, so they no longer contribute to the supercurrent.

The full scattering matrix of the normal region, including the scattering channels to the electron reservoir, is unitary, with SN(ε+iγ) a sub-unitary submatrix for the scattering channels to the superconductors. Similarly, the Andreev reflection matrix RA(ε+iγ) is now sub-unitary also for ε<Δ0.

Instead of Eq. (4), we have for the density of states the expression
(9)
The term ρ0 now also contains the ϕ-independent contributions from the electron reservoir. The free energy then follows from
(10)
We discuss three applications of this general formula.

As a first application, we consider the case that the two superconductors are coupled via a quantum dot containing a single resonant level, at energy εR relative to the Fermi level. We ignore Coulomb blockade effects, which will be less significant in the open system.

The electronic scattering matrix of the quantum dot has the following form:
(11)
with Γ1,Γ2 the tunnel rates into the left and right superconductor, and Γ=Γ1+Γ2. The normal-state transmission probability through the quantum dot at the Fermi level (ε=0) is given by the Breit–Wigner formula,22 
(12)
In the weak-coupling, near-resonant regime, when εR,ΓΔ0, the energy dependence of RA can be neglected and we may substitute α(ε)α(0)=i. The determinant (4) evaluates to
(13)
plus ϕ-independent terms. The prefactor g accounts for spin and possibly other degeneracies. The energy,
(14)
is the energy of the Andreev level in the closed system.23,24 There is no dependence on Δ0 in the weak-coupling regime.
The free energy (10) becomes
(15)
The resulting supercurrent is
(16)
with ψ(x) the digamma function and
(17)
the zero-temperature supercurrent of the closed system. The T0 limit of Eq. (16) (plotted in Fig. 2) is
(18)
so opening up the system to an electron reservoir at T=0 reduces the supercurrent by a factor (2/π)arctan(2εAτ/).
FIG. 2.

Zero-temperature current–phase relation of the quantum dot Josephson junction of Fig. 1, with and without the coupling to the electron reservoir. The plot is computed from Eq. (18) for the case TBW=1 of unit transmission probability at the Fermi level through the quantum dot.

FIG. 2.

Zero-temperature current–phase relation of the quantum dot Josephson junction of Fig. 1, with and without the coupling to the electron reservoir. The plot is computed from Eq. (18) for the case TBW=1 of unit transmission probability at the Fermi level through the quantum dot.

Close modal

The second application is a point contact junction of length L short compared to the superconducting coherence length ξ0=vF/Δ0. In this short-junction regime, the energy dependence of SN can be neglected relative to the energy dependence of RA, and we may evaluate SN at the Fermi level. We assume that time reversal symmetry is preserved; hence, σySNσy=SN.

The Andreev levels in the closed system then depend on the phase difference according to Ref. 18,
(19)
in terms of the mode-dependent transmission probabilities Tn[0,1] in the normal state (eigenvalues of the transmission matrix product tt). The number N counts the number of propagating electron modes through the point contact. Equation (19) is analogous to the result (14) for a quantum dot, with Δeff replaced by Δ0 and TBW replaced by Tn.
Instead of Eq. (13), we now have
(20)
and following the same steps as in the previous application, we find the supercurrent in the open system,
(21)
(22)
with In=(ge/)dεn/dϕ the zero-temperature supercurrent in the n-th mode of the closed system.

For the third and final application, we consider a superconductor–normal-metal–superconductor (SNS) junction of length Lξ0. Both states above and below Δ0 then contribute to the supercurrent.

It is convenient to transform the slowly converging integration over energies of Eq. (10) into a more rapidly converging sum over Matsubara frequencies,25 
(23)
We will restrict ourselves for this application to zero temperature, when the sum can be replaced by an integral,
(24)
We consider a ballistic single-mode junction. The electronic scattering matrix is
(25)
We linearize the momentum near the Fermi energy, k(ε)=kF+ε/vF.
The determinant (5) evaluates to
(26)
The sum over Matsubara frequencies decays on the scale of vF/L, which is much smaller than Δ0 in the long-junction regime. We may thus approximate α(iω)α(0)=i, when
(27)
resulting in the zero-temperature free energy,
(28)
The function Li2(x) is the dilogarithm.
The corresponding supercurrent is
(29)
see Fig. 3. As a check, in the limit γ0, we recover the known sawtooth ϕ-dependence,26–28,
(30)
van Wees et al.6 and Chang and Bagwell7 have qualitatively similar plots to Fig. 3, for different models of coupling to electron reservoirs that do not allow for a closed-form solution.
FIG. 3.

The same as Fig. 2, but now for a single-mode ballistic SNS junction in the long junction limit (length L large compared to the superconducting coherence length ξ0=vF/Δ0). The plot is computed from Eq. (29).

FIG. 3.

The same as Fig. 2, but now for a single-mode ballistic SNS junction in the long junction limit (length L large compared to the superconducting coherence length ξ0=vF/Δ0). The plot is computed from Eq. (29).

Close modal

The coupling of the Josephson junction to the electron reservoir pushes the poles of its scattering matrix into the lower half of the complex plane, down to E=±εAiγ. These poles can be considered as the complex eigenvalues of an effective non-Hermitian Hamiltonian H. The open Josephson junction thus provides a physical realization of the non-Hermitian Josephson effect studied in Refs. 14–17. Different generalizations have been proposed for the relation IdE/dϕ when E is complex. Let us compare these with our findings.

To be specific, we consider the weakly coupled quantum dot Josephson junction, with effective Hamiltonian,29–31 
(31)
One checks that the eigenvalues are ±εAiγ, with εA given by Eq. (14).
Li et al.15 and Cayao and Sato16 argue that only the real part of E contributes to the physical supercurrent, which in this case would imply no effect from the coupling to the reservoir. Shen et al.,17 in a remarkable recent paper, give instead the zero-temperature relation,
(32)
which reduces precisely to our result (17) (with g=2).32 

In summary, we have calculated how the coupling to a gapless electron reservoir in the normal state reduces the supercurrent through a Josephson junction. A simple answer is obtained for the model where the escape rate 1/τ2γ/ of quasiparticles into the reservoir is spatially uniform. At zero temperature, the reduction factor for a given Andreev level equals (2/π)arctan(2εAτ/). This applies to a weakly coupled quantum dot Josephson junction, or to a short point contact (length L smaller than the coherence length ξ0). A more complicated τ-dependence is obtained in a long junction (Lξ0), when also states above the gap contribute to the supercurrent.

I have benefited from discussions with A. R. Akhmerov, Yu. V. Nazarov, P.-X. Shen, and T. Vakhtel. This project has received funding from the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation programme (Advanced Grant No. 832256).

The author has no conflicts to disclose.

C. W. J. Beenakker: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
A. A.
Golubov
,
M. Yu.
Kupriyanov
, and
E.
Ilichev
, “
The current-phase relation in Josephson junctions
,”
Rev. Mod. Phys.
76
,
411
(
2004
).
2.
D. V.
Averin
, “
Adiabatic quantum computation with Cooper pairs
,”
Solid State Commun.
105
,
659
(
1998
).
3.
J. P.
Pekola
and
J. J.
Toppari
, “
Decoherence in circuits of small Josephson junctions
,”
Phys. Rev. B
64
,
172509
(
2001
).
4.
Y.
Makhlin
,
G.
Schön
, and
A.
Shnirman
, “
Dissipative effects in Josephson qubits
,”
Chem. Phys.
296
,
315
(
2004
).
5.
M.
Büttiker
, “
Small normal-metal loop coupled to an electron reservoir
,”
Phys. Rev. B
32
,
1846(R)
(
1985
).
6.
B. J.
van Wees
,
K.-M. H.
Lenssen
, and
C. J. P. M.
Harmans
, “
Transmission formalism for supercurrent flow in multiprobe superconductor-semiconductor-superconductor devices
,”
Phys. Rev. B
44
,
470(R)
(
1991
).
7.
L.-F.
Chang
and
P. F.
Bagwell
, “
Control of Andreev-level occupation in a Josephson junction by a normal-metal probe
,”
Phys. Rev. B
55
,
12678
(
1997
).
8.
N. A.
Mortensen
,
A.-P.
Jauho
, and
K.
Flensberg
, “
Dephasing in semiconductor-superconductor structures by coupling to a voltage probe
,”
Superlattice Microstruct.
28
,
67
(
2000
).
9.
T.
Gramespacher
and
M.
Büttiker
, “
Distribution functions and current-current correlations in normal-metal–superconductor heterostructures
,”
Phys. Rev. B
61
,
8125
(
2000
).
10.
M.
Belogolovskii
,
A.
Golubov
,
M.
Grajcar
,
M. Yu.
Kupriyanov
, and
P.
Seidel
, “
Charge transport across a mesoscopic superconductor–normal metal junction: Coherence and decoherence effects
,”
Physica C
357–360
,
1592
(
2001
).
11.
M.
Belogolovskii
, “
Phase-breaking effects in superconducting heterostructures
,”
Phys. Rev. B
67
,
100503(R)
(
2003
).
12.
B.
Béri
, “
Dephasing-enabled triplet Andreev conductance
,”
Phys. Rev. B
79
,
245315
(
2009
).
13.
T.
Domański
,
M.
Žonda
,
V.
Pokorný
,
G.
Górski
,
V.
Janiš
, and
T.
Novotný
, “
Josephson-phase-controlled interplay between correlation effects and electron pairing in a three-terminal nanostructure
,”
Phys. Rev. B
95
,
045104
(
2017
).
14.
V.
Kornich
, “
Current-voltage characteristics of the normal metal-insulator-PT-symmetric non-Hermitian superconductor junction as a probe of non-Hermitian formalisms
,”
Phys. Rev. Lett.
131
,
116001
(
2023
).
15.
C.-A.
Li
,
H.-P.
Sun
, and
B.
Trauzettel
, “
Anomalous Andreev spectrum and transport in non-Hermitian Josephson junctions
,”
Phys. Rev. B
109
,
214514
(
2024
).
16.
J.
Cayao
and
M.
Sato
, “
Non-Hermitian phase-biased Josephson junctions
,” arXiv:2307.15472 (
2023
).
17.
P.-X.
Shen
,
Z.
Lu
,
J. L.
Lado
, and
M.
Trif
, “
Non-Hermitian Fermi-Dirac distribution in persistent current transport
,”
Phys. Rev. Lett.
133
,
086301
(
2024
).
18.
C. W. J.
Beenakker
, “
Universal limit of critical-current fluctuations in mesoscopic Josephson junctions
,”
Phys. Rev. Lett.
67
,
3836
(
1991
).
19.
J.
Bardeen
,
R.
Kümmel
,
A. E.
Jacobs
, and
L.
Tewordt
, “
Structure of vortex lines in pure superconductors
,”
Phys. Rev.
187
,
556
(
1969
).
20.
M. R.
Zirnbauer
, “
Conductance fluctuations of disordered mesoscopic devices with many weakly coupled probes
,”
Nucl. Phys. A
560
,
95
(
1993
).
21.
P. W.
Brouwer
and
C. W. J.
Beenakker
, “
Voltage-probe and imaginary potential models for dephasing in a chaotic quantum dot
,”
Phys. Rev. B
55
,
4695
(
1997
).
22.
G.
Breit
and
E. P.
Wigner
, “
Capture of slow neutrons
,”
Phys. Rev.
49
,
519
(
1936
).
23.
C. W. J.
Beenakker
and
H.
van Houten
, “
Resonant Josephson current through a quantum dot
,” in
Single-Electron Tunneling and Mesoscopic Devices
, edited by
H.
Koch
and
H.
Lubbig
(
Springer
,
Berlin
,
1992
). Online at arXiv:cond-mat/0111505.
24.
I. A.
Devyatov
and
M. Yu.
Kupriyanov
, “
Resonant Josephson tunneling through S-I-S junctions of arbitrary size
,”
J. Exp. Theor. Phys.
85
,
189
(
1997
).
25.
P. W.
Brouwer
and
C. W. J.
Beenakker
, “
Anomalous temperature dependence of the supercurrent through a chaotic Josephson junction
,”
Chaos, Solitons Fractals
8
,
1249
(
1997
). Several misprints are corrected in the online version at arXiv:cond-mat/9611162.
26.
C.
Ishii
, “
Josephson currents through junctions with normal metal barriers
,”
Prog. Theor. Phys.
44
,
1525
(
1970
).
27.
J.
Bardeen
and
J. L.
Johnson
, “
Josephson current flow in pure superconducting-normal-superconducting junctions
,”
Phys. Rev. B
5
,
72
(
1972
).
28.
A. V.
Svidzinsky
,
T. N.
Antsygina
, and
E. N.
Bratus
, “
Concerning the theory of the Josephson effect in pure SNS junctions
,”
J. Low Temp. Phys.
10
,
131
(
1973
).
29.
T.
Meng
,
S.
Florens
, and
P.
Simon
, “
Self-consistent description of Andreev bound states in Josephson quantum dot devices
,”
Phys. Rev. B
79
,
224521
(
2009
).
30.
P.
Recher
,
Yu. V.
Nazarov
, and
L. P.
Kouwenhoven
, “
Josephson light-emitting diode
,”
Phys. Rev. Lett.
104
,
156802
(
2010
).
31.
D. O.
Oriekhov
,
Y.
Cheipesh
, and
C. W. J.
Beenakker
, “
Voltage staircase in a current-biased quantum-dot Josephson junction
,”
Phys. Rev. B
103
,
094518
(
2021
).
32.
To evaluate Eq. (32), one can work in a basis where the Hamiltonian (31) is diagonal, with eigenvalues ±εAiγ, and note that (d/dϕ)TrH=0, while dH/dϕ is real, so that (d/dϕ)ImTr(HlnH)=Tr(dH/dϕ)ImlnH=(dεA/dϕ)ImlnεAiγεAiγ=2(dεA/dϕ)arctan(εA/γ). This is at zero temperature. We have checked that the finite-T expression (16) also agrees with Ref. 17.
Published open access through an agreement with Leiden University