• Nie Znaleziono Wyników

Bistability in voltage-biased normal-metal/insulator/superconductor/insulator/normal-metal structures

N/A
N/A
Protected

Academic year: 2021

Share "Bistability in voltage-biased normal-metal/insulator/superconductor/insulator/normal-metal structures"

Copied!
13
0
0

Pełen tekst

(1)

Bistability in voltage-biased normal-metal/insulator/superconductor/insulator/normal-metal

structures

I. Snyman1,2 and Yu. V. Nazarov3

1National Institute for Theoretical Physics, Private Bag X1, 7602 Matieland, South Africa 2Instituut-Lorentz, Universiteit Leiden, P.O. Box 9506, 2300 RA Leiden, The Netherlands 3Kavli Institute of Nanoscience, Delft University of Technology, 2628 CJ Delft, The Netherlands 共Received 27 August 2008; revised manuscript received 29 October 2008; published 15 January 2009兲

As a generic example of a voltage-driven superconducting structure, we study a short superconductor connected to normal leads by means of low transparency tunnel junctions with a voltage bias V between the leads. The superconducting order parameter⌬ is to be determined self-consistently. We study the stationary states as well as the dynamics after a perturbation. The system is an example of a dissipative driven nonlinear system. Such systems generically have stationary solutions that are multivalued functions of the system pa-rameters. It was discovered several decades ago that superconductors outside equilibrium conform to this general rule in that the order parameter as a function of driving may be multivalued. The main difference between these previous studies and the present work is the different relaxation mechanisms involved. This does not change the fact that there can be several stationary states at a given voltage. It can however affect their stability as well as the dynamics after a perturbation. We find a region in parameter space where there are two stable stationary states at a given voltage. These bistable states are distinguished by distinct values of the superconducting order parameter and of the current between the leads. We have evaluated共1兲 the multivalued superconducting order parameter⌬ at given V, 共2兲 the current between the leads at a given V, and 共3兲 the critical voltage at which superconductivity in the island ceases. With regards to dynamics, we find numerical evidence that only the stationary states are stable and that no complicated nonstationary regime can be induced by changing the voltage. This result is somewhat unexpected and by no means trivial, given the fact that the system is driven out of equilibrium. The response to a change in the voltage is always gradual even in the regime where changing the interaction strength induces rapid anharmonic oscillations of the order parameter. DOI:10.1103/PhysRevB.79.014510 PACS number共s兲: 74.40.⫹k, 74.78.Fk, 74.25.Fy, 74.78.Na

I. INTRODUCTION

Electron transport devices combining superconducting 共S兲, insulating 共I兲, and normal-metal 共N兲 elements are known as superconducting heterostructures. Often such heterostruc-tures are more than the sum of their parts.1,2Phenomena that

are not present in bulk S, I, or N systems appear when a device contains junction between these components. The fol-lowing examples are well known: 共1兲 the conductance of a high transparency NS junction does not equal the conduc-tance of the normal metal on its own, as one might naively expect. If the normal metal is free of impurities, the conduc-tance is higher than that of the normal metal.3This surprising

effect is due to a process known as Andreev reflection.4

Dur-ing Andreev reflection at an NS interface, an electron im-pinging on the interface from the N side is reflected back as a hole while a Cooper pair propagates away from the inter-face on the S side.共2兲 In Josephson junctions, the simplest of which is perhaps the SIS heterostructure,5 a dc current can

flow at zero-bias voltage. This happens when the supercon-ducting phase difference across the junction is nonzero.6

The above examples can be understood in terms of equi-librium properties of the heterostructure. When a supercon-ducting device is perturbed outside equilibrium, yet more interesting effects can occur,7for instance, oscillations under

stationary nonequilibrium conditions. An elementary ex-ample: if a Josephson junction is biased with a dc共i.e., fixed兲 voltage, an ac 共i.e., oscillating兲 current flows through the junction.6Another example of the kind has been investigated

in the context of cold Fermi gases in optical traps. In these systems, the interaction between atoms can be tuned and changed by means of a so-called Feshbach resonance. If the interaction is attractive, the gas forms a BCS condensate. Recent studies8,9 have considered what happens if the value

of the attractive pairing interaction is changed abruptly. It was discovered that, depending on the ratio between the ini-tial and final values of the interaction strength, the conden-sate order parameter can perform anharmonic oscillations that do not decay in time.

The initial motivation for the research presented in this paper came from the study of Keizer et al.,10where the

au-thors investigated the suppression of the superconducting or-der parameter by a voltage applied to a superconducting wire. It was assumed that ⌬ remains stationary. However, this assumption does not seem well justified: the stationary voltage could induce periodic oscillations of 兩⌬兩 or even richer chaotic dynamics. Thus prompted, we wanted to ad-dress the validity of this assumption for a decidedly simpler NISIN structure, namely, a short superconductor connected to normal leads by means of tunnel junctions. The structure is biased with a voltage V.

We require that 共1兲 the dominant energy relaxation mechanism in the superconductor is the tunneling of elec-trons to the leads, and共2兲 spatial variations in the supercon-ducting order parameter inside the superconductor are negli-gible. To meet the first requirement, the superconductor must have dimensions smaller than the inelastic-scattering length of quasiparticles. This is not an unrealistic requirement given

(2)

current experimental techniques. To meet the second require-ment, the superconductor should contain impurities or have an irregular shape so that the electron wave functions of the isolated island are isotropic on the scale of the superconduct-ing coherence length.11 Furthermore, the tunnel junctions

connecting it to the leads should have a bigger normal-state resistance than that of the superconductor proper. In this case, opening up the system by connecting leads does not reintroduce spatial anisotropy of wave functions inside the island.

The study of NISIN structures has a long history.12,13Our

study complements several previous studies.10,14,15 These dealt with quasi-one-dimensional superconducting wires be-tween normal leads. Setups where either the superconductor was impurity free or the transparency of the NS interfaces was high were considered. For these setups, spatial varia-tions in the order parameter, specifically the spatial gradient of the superconducting phase, can be large. Including these spatial variations in the description of the superconductor significantly complicates matters. Hence these studies fo-cused on numerical calculations and assumed that the super-conducting order parameter and all other quantities of inter-est were stationary. It should also be mentioned that asymmetric couplings, where the superconductor is coupled more strongly to one lead than the other, did not receive detailed analysis. The only asymmetric setup considered con-sisted of one interface with tunable transparency and the other perfectly transparent.15One of the main conclusions of

these studies is that, if the bias voltage is large enough, the system switches to the normal state. Some evidence for a bistable region, where, depending on the history of the sys-tem, either the superconducting or the normal state can occur at a given voltage, was reported.10

The absence of spatial variations in the system we study allows us to perform analytical calculations, provided we assume stationarity. Results are obtained for an arbitrary ra-tio of the coupling strengths to the leads. We derive transcen-dental equations relating the superconducting order param-eter to the bias voltage and derive an explicit formula for the current between the leads. As mentioned, the assumption of stationarity is however not a priori justified. As was seen in the examples mentioned at the beginning of this introduction, nonequilibrium conditions in superconductors often go hand in hand with nonstationary behavior of observable quantities. Indeed, the NISIN junction that we study is a nonlinear sys-tem subjected to a driving force 共and to damping兲. Nonlin-earity here means that the dynamical equations for one-particle Green’s functions are not linear in the Green’s functions. This is due to the existence of a nonzero super-conducting order parameter. The driving force is provided by the voltage共and the damping by tunneling of electrons from the island into the leads兲. Nonlinear driven systems 共think of the nonlinear pendulum兲 often have chaotic dynamics. The assumption of stationarity would miss this. We therefore supplement our analytical calculation with numerical calcu-lations that study the dynamics in real time.

Our main results are the following: the stationary states that we found analytically are stable. Furthermore, there is a parameter region where two different stationary states are stable at the same voltage.共This multivaluedness of the order

parameter is by no means a new phenomenon. It is a com-mon feature of superconductors outside equilibrium.16,17兲 For

a symmetric coupling to the source and drain leads, one of the two states we find is superconducting共characterized by a nonzero order parameter兲 and the other is normal. Since we are in the regime of high tunnel barriers, at a given voltage, the superconducting island allows less current to flow be-tween the leads than the island in the normal state.3 This

current is a directly measurable quantity and allows one to distinguish between superconducting and normal states. For some asymmetric couplings however, both the stable states are superconducting. We have calculated the current that flows between the leads at a given voltage and at arbitrary asymmetry of the coupling to the two leads. We find that the value of the current also allows one to distinguish between different stable superconducting states at a given voltage.

The time-dependent calculations revealed that, once the bias voltage becomes constant in time, the system always relaxes into one of the stationary states. Nonstationary be-havior of physical quantities always decays in time, unlike in the case of a dc-biased Josephson junction.共Despite it being a nonlinear system, a superconductor driven by a voltage is therefore fundamentally different from a nonlinear pendulum driven by an external force.兲 If the bias voltage is changed slowly, an initial stationary state evolves adiabatically. By changing the voltage slowly we have observed the expected hysteresis associated with the existence of two stable states at some voltages.

The study presented here is complementary to earlier studies of bulk nonequilibrium superconductors. There, the superconducting order parameter is very sensitive to the de-tails of the quasiparticle distribution function.18 Owing to

long inelastic relaxation times in the bulk superconductors, driving by either microwave radiation or tunnel quasiparticle injection may result in sufficient modification of the distri-bution function. General reasoning predicts multiple super-conducting states in this situation, and indeed they have been found in Refs. 16 and 17. The stability of these states strongly depend on the nature of the quasiparticle distribu-tion funcdistribu-tion and the relaxadistribu-tion mechanism.16,17,19Proper ac-count of superconducting fluctuations may be required to understand the transitions between the stable states and even-tually to recover adiabaticity of the superconductor dynamics in the low-frequency limit.20 In distinction from previous

work, we assume that the relaxation is provided by tunneling to/from the leads rather than by inelastic processes in the bulk. The relaxation time is therefore⯝ប/EThand may be-come comparable with inverse of the energy scales ⌬ and eV. The latter forbids the use of the Boltzmann equation for the distribution function implemented in earlier studies.16,17

Therefore our work is based on a Green’s function technique. The rest of the paper is structured as follows. In Sec.IIwe specify the model to be studied, and present the equations that determine its state. In Sec.IIIwe solve these equations analytically, assuming that the system is in a stationary state. We analyze the stationary states we find and calculate the I-V characteristic of the system. In Sec.IVwe establish that the stationary states are the only stable states of the dc-biased system. We do so by studying the dynamics of the system after a perturbation. In Sec. Vwe summarize our main re-sults.

(3)

II. MODEL

As stated in Sec.I, we consider a superconducting island connected to two normal leads by means of low transparency tunnel barriers. The superconducting order parameter is taken to be spatially isotropic inside the island. The physical requirements for this condition to hold have already been discussed in Sec. I. We assume that the dominant energy relaxation mechanism for the superconductor is tunneling of electrons to the leads. For given barrier transparencies, this restricts the size of the superconductor to less than the inelastic-scattering length of quasiparticles inside the super-conductor.

Our analysis of the system is based on the Keldysh Green’s function technique.21–23 We start our discussion of

the equations governing the system by defining the necessary Green’s functions.

A. Definition of Green’s functions

The Green’s functions are expectation values of products of the Heisenberg operators am共t兲 and am共t兲 that create

and annihilate electrons in levels of the isolated island. Here m labels single-particle levels. The ⫾ index accounts for Kramer’s degeneracy. As we are dealing with a problem in-volving superconductivity, all Green’s functions are 2⫻2 matrices in Nambu space. It is useful to define Nambu space matrices ␩j, j = 0 , . . . , 3 such that ␩0 is the identity matrix, and␩1,␩2, and␩3are the standard Pauli matrices. We also define matrices␩=共␩1⫾i␩2兲/2.

The retarded共R兲, Keldysh 共K兲, and advanced 共A兲 Green’s functions of each level are defined as24

Rm共t,t⬘兲 = − i␩3

兵am+共t兲,am+共t⬘兲其 兵am+共t兲,am−共t

兲其

兵am−共t兲,a m+共t

兲其 兵a m−共t兲,a m−共t

兲其

⫻␪共t − t

兲, 共2.1a兲 Km共t,t⬘兲 = − i␩3

关am+共t兲,am+共t⬘兲兴 关a m+共t兲,am−共t⬘兲兴

关am−共t兲,am+共t

兲兴 关am−共t兲,am−共t

兲兴

, 共2.1b兲 Am共t,t⬘兲 =␩3Rm共t⬘,t兲†␩3. 共2.1c兲 The Green’s functions are grouped into a matrix

Gm共t,t

兲 =

Rm共t,t⬘兲 Km共t,t⬘兲

0 Am共t,t

. 共2.2兲

This further 2⫻2 matrix structure is referred to as Keldysh space. As with Nambu space, it is useful to define matrices ␶j, j = 0 , . . . , 3. The matrixjis the same as the matrix␩jbut

now operating in Keldysh space. We also carry over the defi-nition of␶from Nambu space. A basis for the 4⫻4 matri-ces that results from combining Keldysh and Nambu indimatri-ces is constructed by means of a tensor product␶j丢␩k, with the ␶’s always acting in Keldysh space and the ␩’s in Nambu space.

The quantities that we calculate, namely, the order param-eter⌬共t兲 and the current I共t兲, are collective in the sense that

they result from the sum of the contributions of all the indi-vidual levels. Accordingly a formalism exists that does not require knowledge of the Green’s functions of individual lev-els but only the sums22,25–28

G共t,t

⬘兲 =

is

m

Gm共t,t⬘兲, G = G,R,K,A, 共2.3兲

which are known as quasiclassical Green’s functions. Here␦s

is the mean level spacing of the island.

We will work with the quasiclassical Green’s functions throughout the present section. The advantage of doing so is that the theory can be formulated with the least amount of clutter. When doing time-dependent numerics in Sec. IV

however, we find it more convenient to work with the Green’s functions of the individual levels.

B. Equations of motion

The equations that determine the Green’s functions can be derived from the circuit theory of nonequilibrium superconductivity.26–28Viewed as a matrix in time, Nambu,

and Keldysh indices, the Green’s function G satisfies the commutation relation29

关H − ⌺,G兴 = 0. 共2.4兲

Here H describes the dynamics of the isolated supercon-ductor: H共t,t⬘兲 =␶0丢␩3␦共t − t⬘兲关it− h共t兲兴, 共2.5a兲 h共t兲 =

−␮s共t兲 ⌬共t兲 ⌬共t兲 s共t兲

. 共2.5b兲

The matrix h共t兲 is a remnant of the Bogoliubov–de Gennes Hamiltonian.11Bearing in mind that we consider a

nonequi-librium setup, we must allow the order parameter ⌬共t兲 and the chemical potential␮s共t兲 of the superconductor to be time

dependent. Their values at each instant in time are deter-mined by imposing self-consistency.

The time derivative standing to the right of G in the term GH of Eq. 共2.4兲 can be shifted to act on the second time

argument of G at the cost of a minus sign, i.e.,

dt˜G共t,t˜兲共t˜ − t⬘兲 = −

dt˜t˜G共t,t˜兲共t˜ − t⬘兲 = −tG共t,t⬘兲.

共2.6兲 The self-energy contains a term corresponding to each lead, i.e.,

⌺ = ⌺共l兲+共r兲, 共2.7兲 l and r referring to the left and right leads, respectively. The leads act as reservoirs, broadening the island levels to a finite lifetime and determining their filling. The self-energy of lead j is共j兲= −ijG共j兲, where Green’s function G共j兲 of lead j is

defined similarly to the Green’s function of the supercon-ductor 关Eq. 共2.3兲兴 with the sum now running over states in

the lead. Here ⌫jis the tunneling rate from any island level

(4)

different levels to be the same.兲 The leads are large compared to the superconductor, and therefore Gj does not depend on

the state of the superconductor. Furthermore, since the leads are normal, the off-diagonal Nambu space matrix elements of the lead Green’s functions are zero. Explicitly then, the Green’s function for lead j = l , r has the form

G共j兲共t,t⬘兲 =

R 共j兲共t,t⬘兲 K共j兲共t,t⬘兲 0 A共j兲共t,t⬘兲

, 共2.8兲 with R共j兲共t,t⬘兲 =共t − t⬘兲3= − A共j兲共t,t⬘兲, 共2.9a兲 K共j兲共t,t

兲 = 2

j共t,t

兲 0 0 ␴j共t,t⬘兲

. 共2.9b兲

The function␴jdescribes the distribution of particles in lead

j. In general it is given byj共t,t

兲 =

dE 2␲e −iE共t−t⬘兲关1 − 2f j共E兲兴e−i关␾j共t兲−␾j共t⬘兲兴, 共2.10兲 where fj共E兲 is the filling factor of states at energy E in lead

j. The phasej sets the time-dependent chemical potential ␮j共t兲=tj共t兲 in lead j. The time-dependent bias voltage

be-tween the leads is

V共t兲 = 关l共t兲 −r共t兲兴/e, 共2.11兲

where e is the electron charge. It is convenient to define the total inverse lifetime or Thouless energy ETh=⌫l+⌫r and a

dimensionless symmetry parameter ␥=共⌫l−⌫r兲/ETh. For a perfectly symmetric coupling to the leads, ␥= 0 while ␥ =⫾1 corresponds to the island being coupled to only one of the two leads.

The commutator Eq. 共2.4兲 on its own is not enough to

specify G uniquely. Indeed what Eq.共2.4兲 says is that G has

the same eigenstates as H −⌺ but it does not say anything about the eigenvalues of G. Additional to Eq. 共2.4兲 there is

also a relation between the eigenvalues of G and those of H −⌺.30Let兩␭典 be a simultaneous eigenstate of H−⌺ and G,

such that its eigenvalue with respect to H −⌺ is ␭. Then its eigenvalue with respect to G is sgn关Im共␭兲兴. 共One can show that the eigenvalues of H −⌺ come in complex-conjugate pairs and that there are no purely real eigenvalues.兲 Hence G squares to unity, i.e.,

G2= I. 共2.12兲

C. Gauge invariance

At this point we have defined three different Fermi ener-gies, namely, that of the superconductor ␮s共t兲 and those of

the leads ␮j共t兲 with j=l,r. Since the reference point from

which energy is measured is arbitrary, there is some redun-dancy. This redundancy is encoded in a symmetry of the equations for the Green’s function and boils down to gauge invariance. Consider a transformation on the Green’s func-tion

G→ G˜ = UGU†, 共2.13a兲 U共t,t

兲 =␦共t − t

兲␶0丢exp关i␩3⌳共t兲兴. 共2.13b兲 As is easily verified, G˜ obeys equations of the same form as G, with chemical potentials and the order parameter trans-formed according to

j共t兲 →˜j共t兲 =j共t兲 +t⌳共t兲, j = s,l,r, 共2.14a兲

⌬共t兲 → ⌬˜共t兲 = ⌬共t兲exp关2i⌳共t兲兴. 共2.14b兲 When considering stationary solutions we will fix the gauge by demanding that⌬ is time independent. When considering nonstationary solutions we will fix the gauge such that the reference point from which chemical potentials are measured is halfway between the chemical potentials of the reservoirs, i.e.,␮r共l兲共t兲= +共−兲eV共t兲/2.

D. Self-consistency of

The value of the order parameter is set by the self-consistency condition ⌬共t兲 = gs

m 具am−共t兲am+共t兲典 = −g 2 Tr关␩−K共t,t兲兴, 共2.15兲 where g⬎0 is the dimensionless pairing interaction strength. This self-consistency equation suffers from the usual loga-rithmic divergence which requires regularization by intro-ducing a large energy cutoff Eco. We define ⌬0 as the order parameter of an isolated superconductor at zero temperature for given g and Eco,

⌬0= Eco sinh1 g ⇒1 g=

0 Eco dE

冑E

2+ 0 2. 共2.16兲 This definition then allows us to express ⌬ in Eq. 共2.15兲 in

terms of⌬0 rather than in terms of Ecoand g. E. Current and chemical potential

The current from the superconductor into reservoir j is31

Ij共t兲 =2eGj

dt

Tr兵␶−丢␩3关G共t,t⬘兲G 共j兲共t⬘,t兲 − G共j兲共t,t⬘兲G共t⬘,t兲兴其, Gl共r兲=关1 + 共− 兲␥兴 ETh ␦s e2 关ប兴. 共2.17兲

Here Gj is the tunneling conductance of the tunnel barrier

between lead j and the superconductor, and we have indi-cated in square brackets a factor ofប which equals unity in the units we use throughout the paper. The total rate of change in the charge in the superconductor is equal to the negative of the sum of the currents to the leads, i.e.,

(5)

d

dtQ共t兲 = Il共t兲 + Ir共t兲. 共2.18兲 The charge in the superconductor is related to the chemical potential␮sby means of the capacitance C of the

supercon-ductor so that␮s has to obey

1 e d dts共t兲 = C d dtQ共t兲. 共2.19兲

When the system is not stationary, this equation sets the value of ␮s共t兲 at each instant in time since dQ共t兲/dt can be

calculated directly from G共t,t

兲. F. Summary

In summary then, our task is to find the Green’s function G as defined in Eq. 共2.3兲 of the superconductor. In general,

the procedure for doing this is as follows: we make an ansatz for the order parameter⌬共t兲 and the chemical potentials共t兲.

We then diagonalize the operator H −⌺ 共that depends on ⌬ and␮兲. The Green’s function G is constructed in the eigen-basis of H −⌺, according to the prescription of Sec. II B. Subsequently we judge the correctness of the ansatz for⌬共t兲 and ␮共t兲 by inquiring whether Eqs. 共2.15兲 and 共2.19兲 are

satisfied.

III. STATIONARY SOLUTIONS

We consider a time-independent bias voltage between the left and right reservoirs. In this case the chemical potentials ␮land␮rof the reservoirs are time independent. We make

the ansatz that the chemical potential ␮s and the order

pa-rameter⌬ of the superconductor are also time independent. The Green’s function G共t,t

兲 only depends on the time dif-ference t − t

. It is convenient to work with the Fourier-transformed Green’s function G共E兲 which is related to G共t,t

兲 by

G共t,t

兲 =

dE 2␲e

−iE共t−t⬘兲G共E兲. 共3.1兲 It is also convenient to construct a traceless operator M = H −⌺−␶0丢␩0␮s with Keldysh structure

M =

MR MK 0 MA

. 共3.2兲

In the energy representation, the components of M have the explicit form MR共E兲 =

E + iETh −⌬ ⌬ⴱ − E − iE Th

, 共3.3a兲 MA共E兲 =

E − iETh −⌬ ⌬ⴱ − E + iE Th

, 共3.3b兲 MK共E兲 = 2iETh

共E兲 0 0 ␴共− E兲

, 共3.3c兲 ␴共E兲 =1 −␥ 2 ␴l共E兲 + 1 +␥ 2 ␴r共E兲. 共3.3d兲 We take the left and right leads to be in local zero-temperature equilibrium at Fermi energies ␮l=␮+ eV/2 and ␮r=␮− eV/2 so that the filling factors in both reservoirs is a

step function fj共E兲=共−E兲 and from Eq. 共2.10兲 follows ␴l共E兲 = sgn共E −− eV/2兲, 共3.4a兲 ␴r共E兲 = sgn共E −+ eV/2兲, 共3.4b兲

where␮is the average chemical potential 共␮r+␮l兲/2 in the

leads, in the gauge where the phase of the order parameter is time independent. The value of␮will later be determined by requiring self-consistency of the order parameter ⌬. The Green’s function G共E兲 obeys 关M共E兲,G共E兲兴=0. The retarded, advanced, and Keldysh components of this equation are

关MR共E兲,R共E兲兴 = 关MA共E兲,A共E兲兴 = 0, 共3.5a兲

MR共E兲K共E兲 + MK共E兲A共E兲 − R共E兲MK共E兲 − K共E兲MA共E兲 = 0.

共3.5b兲 With the aide of the prescription below Eq. 共2.11兲 for

choosing the eigenvalues of G, one then readily finds for the retarded and advanced Green’s functions

R共E兲 = 1 c共E兲

E + iETh −⌬ ⌬ⴱ − E − iE Th

, 共3.6a兲 A共E兲 = 1 c共− E兲

E − iETh −⌬ ⌬ⴱ − E + iE Th

, 共3.6b兲 c共E兲 =

共E + iETh兲2−兩⌬兩2. 共3.6c兲 The function c共E兲, which we will frequently encounter, is defined with branch cuts along the lines E=⫾兩⌬兩⫾x − iETh where x is real and positive. The branch with limE苸R→⫾⬁c共E兲/E=1 is taken. Considered as a function of

real E, the real part of c共E兲 is odd, and the imaginary part is even and positive so that

c共E兲= − c共− E兲. 共3.7兲

The real and imaginary parts of c共E兲 are plotted for real E in Fig.1.

Note that R共E兲2= A共E兲2=

0 as required by Eq.共2.12兲. In general the superconducting density of states is 共E兲 = Tr␩3关R共E兲−A共E兲兴/2s so that we find from the solutions

for R and A关Eq. 共3.6兲兴

共E兲 = 2 ␦s

Re

E + iETh

c共E兲

. 共3.8兲

The density of states for an isolated superconductor has sin-gularities at energies E =⫾兩⌬兩 of the form 1/

E2−兩⌬兩2. The coupling to the leads regularizes the singularities at an en-ergy scale of ETh. Furthermore, whereas the density of states of the isolated superconductor vanishes for energies 兩E兩 ⬍兩⌬兩, the coupling to the leads softens the gap so that there are some states for energies兩E兩⬍兩⌬兩 as shown in Fig. 2.

(6)

The next step is to solve Eq.共3.5b兲 for K共E兲. Here we use

the fact that R共E兲K共E兲A共E兲=−K共E兲, which follows from the requirement that G2= I 关Eq. 共2.12兲兴. Note also that R共E兲

= MR共E兲/c共E兲 and A共E兲=MA共E兲/c共−E兲. Using these

identi-ties and multiplying Eq.共3.5b兲 from the left by R共E兲, we find

K共E兲 = 1

c共E兲 + c共− E兲关MK共E兲 − R共E兲MK共E兲A共E兲兴. 共3.9兲 After some algebra we obtain

K共E兲 =

K

共1兲共E兲 K共2兲共E兲

− K共2兲共E兲K共1兲共− E兲

, 共3.10兲 where

K共1兲共E兲 =s共E兲共E兲 −

兩⌬兩2 E Re

1

c共E兲

关␴共E兲 +共− E兲兴, 共3.11a兲 K共2兲共E兲 = − ⌬ Re

1

c共E兲

关␴共E兲 −共− E兲兴iETh

E 关␴共E兲 +共− E兲兴

. 共3.11b兲 Having obtained K共E兲 we can find 兩⌬兩 and␮from the self-consistency condition 关Eq. 共2.15兲兴. Below we write the real

and imaginary parts of the self-consistency equation sepa-rately. The real part reads

0 =

dE

Re

1 c共E兲

共E兲 −共− E兲 2 − 1

冑E

2+ 0 2

, 共3.12兲 while the imaginary part reads

0 =

dE1 ERe

1

c共E兲

关␴共E兲 +共− E兲兴. 共3.13兲 These integrals can be done explicitly. We use the identities

0

E

dE

Re 1

c共E

= FR共E兲 − FR共0兲, 共3.14a兲

0 E dE

1 E

Re 1 c共E

兲= 1

冑E

Th2 +兩⌬兩2 FI共E兲, 共3.14b兲 where FR共E兲 = ln

E + iETh+ c共E兲 ⌬0

, 共3.15a兲 FI共E兲 = arctan

Re关c共E兲兴

冑E

Th2 +兩⌬兩2

. 共3.15b兲

Here the branch for which −␲/2⬍arctan共x兲⬍␲/2 is im-plied. Thus we obtain the transcendental equations

0 =共1 −␥兲FR

eV 2 +␮

+共1 +␥兲FR

␮− eV 2

, 共3.16a兲 0 =共1 −␥兲FI

eV 2 +␮

+共1 +␥兲FI

␮− eV 2

, 共3.16b兲 that determine兩⌬兩 and␮for given V, ETh, and␥. Below we solve these equations analytically in certain limiting cases and numerically for more general cases. Only the amplitude of⌬ is fixed by these equations. By choosing the appropriate gauge 关cf. Eq. 共2.14b兲兴, we can set the phase of ⌬ to any

value. In the rest of this section we therefore drop the abso-lute value notation, and take⌬ as real and positive.

Before explicitly finding ⌬ and ␮, we calculate the cur-rent from Eq.共2.17兲 and the solution for K共E兲. Assuming that

the self-consistency equation关Eq. 共3.13兲兴 is fulfilled, we find

that the current Irfrom the superconductor to the right lead is

equal to the negative of the current Il from the

supercon-ductor to the left lead, as it should. For the current I = Ir

= −Il from the left lead to the right lead, we find

3 2 1 0 1 2 3 0 1 2 3 4 5 E/|∆| δs 

FIG. 2. The density of states of the superconducting island关Eq. 共3.8兲兴 for finite Thouless energy 共solid line兲. The dashed line shows

the density of states of the isolated superconducting island with the same兩⌬兩 while the horizontal dot-dashed line shows the density of states of the normal island. A value of ETh= 0.1兩⌬兩 was used.

2 1 0 1 2 1 0 1 E/|∆| c( E )/ |∆ |

FIG. 1. The function c共E兲 as defined in Eq. 共3.6c兲 frequently

appears in expressions associated with stationary solutions. The solid line represents the real part and the dashed line the imaginary part. The Thouless energy was taken as ETh= 0.1兩⌬兩.

(7)

I =共1 −␥ 2兲eE Th 2

␮−eV/2 ␮+eV/2 dE共E兲 =GN e Re

c

␮+ eV 2

− c

␮− eV 2

. 共3.17兲 Here GNis the series conductance of the tunneling barriers to

the leads

GN=关Gl

−1 + Gr

−1−1, 共3.18兲

and Gl and Gr are the junction conductances given in Eq.

共2.17兲.

Now we investigate the transcendental equations 关Eqs. 共3.16兲兴 for␮and⌬. There are three parameters, namely, ETh, ␥, and V, which determine the solution. Two of these, ETh and␥, are fixed for a given device while the voltage V can be varied for a given device.共Recall that EThmeasures the over-all coupling to the leads while ␥ measures the degree of asymmetry between the two lead couplings.兲 Hence it is natural to specify values for EThand␥, and then consider⌬, ␮, and I as functions of V. In Fig.3 we show four curves of ⌬ versus V, each corresponding to a different choice of pa-rameters ETh and ␥. In Fig. 4 we show the corresponding curves of␮versus V.

Let us first note the general trend that increasing EThleads to a smaller order parameter. The reason for this is that ETh−1is the typical time an electron remains in the superconductor.

The shorter this time 共the larger ETh兲 the harder it is for electrons to form Cooper pairs, and superconductivity is in-hibited. Second, note that at large enough EThthe order pa-rameter is a decreasing function of V. We can therefore ob-tain the critical Thouless energy ETh共c兲 beyond which superconductivity vanishes by setting V to zero and asking how large can we make EThbefore ⌬ becomes zero.

In the case of V = 0, the self-consistency equations are solved by␮= 0 and ⌬ =

⌬0

1 − 2ETh ⌬0 ETh⬍ ⌬0/2 0 ETh⬎ ⌬0/2

. 共3.19兲

From this we conclude that the critical Thouless energy is ETh共c兲=⌬0/2.

Having established the range of EThin which supercon-ductivity persists, we now take a closer look at⌬ as a func-tion of V. We have chosen the parameters of the four solu-tions in Fig.3 to show all the different possible shapes that curve of⌬ versus V can take. We see that at a given voltage V there can be either zero, one, two, or three nonzero solu-tions⌬. We note that this behavior is qualitatively similar 共as is to be expected兲 to that of a driven bulk superconductor in which the electron-phonon interaction is responsible for relaxation.16,17 As explained in Sec. I, quantitative

differ-ences are accounted for by the different relaxation mecha-nism共tunneling兲 that operates in the present system.

To characterize the different types of curve, we consider V as a function of⌬ on the interval ⌬苸关0,⌬0

1 − 2ETh/⌬0兴. In curves of the type A in Fig.3, V is a monotonically decreas-ing function of ⌬. In contrast, curves of type B, C, and D have local extrema. A curve of type B has a local minimum at the left boundary ⌬=0 of the ⌬ interval on which the function V共⌬兲 is defined. Then the curve reaches a maximum at some intermediate value⌬1before dropping to zero at the right boundary⌬=⌬0

1 − 2ETh/⌬0. Curves C and D are dis-tinguished from curve B by the fact that V reaches a local maximum instead of a minimum at the left boundary⌬=0 of the ⌬ interval. There is another local maximum at interme-diate ⌬1 before V drops to zero at ⌬=⌬0

1 − 2ETh/⌬0. In curves of type C, the absolute maximum of V as a function of⌬ is at the intermediate value ⌬1while for curves of type D the absolute maximum of V is at⌬=0.

Next we ask how the ETh-␥ parameter space is divided into regions A, B, C, and D corresponding to the respective types of solution of the self-consistency equations. Specifi-cally, which regions share a mutual border? Assuming that the function V共⌬兲 changes smoothly as EThand␥are varied, the transitions A↔B, B↔C, C↔D, and D↔A are possible. The transition A↔C is not possible. Whenever one tries to smoothly deform a curve of type A in Fig. 3 to a curve of type C, one invariably reaches a curve of type B or D during an intermediate stage of the deformation. Similarly the tran-sition B↔D is impossible. A smooth deformation of a curve of type B into one of type D passes through an intermediate stage where the curve is of types A or C. To illustrate these ideas we consider a polynomial equation of the form

0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 ∆ / ∆0 eV /∆0 A A BB CC D D

FIG. 3. The order parameter⌬ versus voltage V, for given ETh and ␥. Curves A, B, C, and D, respectively, correspond to ETh = 0.35⌬0 and ␥=0.2, ETh= 0.2⌬0and ␥=0.075, ETh= 0.1⌬0 and ␥ = 0.1, and ETh= 0.01⌬0and␥=0.3. 0.0 0.5 1.0 1.5 2.0 0.0 0.2 0.4 0.6 µ/ ∆0 eV /∆0 A A BB C C D D

FIG. 4. The chemical potential ␮ versus voltage V, for given

EThand␥. Curves A, B, C, and D correspond to the same parameter values as in Fig.3.

(8)

V ˜ ⌬0 = V0 ⌬0 −1 6

⌬ ⌬0

6 −a 4

⌬ ⌬0

4 −b 2

⌬ ⌬0

2 . 共3.20兲

We ask: what are the respective regions of the a-b plane in which V˜ 共⌬兲 is a curve of type A, B, C, and D? Region A, where V˜ 共⌬兲 is of type A, is given by a⬎0, b⬎0 or a ⬍0, b⬎a2/4. Region B where V˜共⌬兲 is of type B consists of all points 共a,b兲 such that b⬍0. Region C consists of all points 共a,b兲 such that a⬍0 and 0⬍b⬍3a2/16. Region D consists of all points 共a,b兲 such that a⬍0 and 3a2/16⬍b ⬍a2/4. The regions and their borders are shown in the inset of Fig.5. The most pertinent feature of the figure is that the four distinct regions meet in the single point a = b = 0.

Based on a combination of numerical and analytical re-sults, we have concluded that the ETh-␥parameter space has a very similar topology to this polynomial example.共In prin-ciple it could have differed from the polynomial example by having disconnected regions of the same type, for instance two islands of region D with one embedded in a sea of re-gion A and the other in a sea of rere-gion C.兲 Figure 5 is a schematic diagram of how the ETh-␥parameter space is par-titioned into regions A, B, C, and D. The following features of the diagram are conjectures based on numerical evidence: 共1兲 the regions of types A, B, C, and D are simply connected. 共2兲 The border between regions A and D starts at the corner= 1, ETh= 0. Other features are deduced from analytical re-sults: 共1兲 the line ␥= 0, ETh⬍⌬0/2

2 belongs to region B. 共2兲 The line␥= 0,⌬0/2

2⬍ETh⬍⌬0/2 belongs to region A. 共3兲 For ETh⬎⌬0/2 the system is in the normal state while it is superconducting for ETh⬍⌬0/2. 共4兲 The border of regions D and C meets the border of region B and C at ETh= 0, ␥ = 0.

In region D, superconductivity can persist up to voltages that are large compared to⌬0. For given EThand␥there is however always a critical voltage Vc beyond which

super-conductivity ceases. 共This is a second-order phase transi-tion.兲 For the voltage Vc we have obtained the following

analytical result from Eq.共3.16兲. At finite␥and for EThthat is sufficiently small, Vcobeys the power law

V共c兲= ⌬0 2e

2ETh ⌬0 sec␲␭ 2

−1/␭ , ␭ =1 −兩␥兩 1 +兩␥兩. 共3.21兲 This power law is valid as long as V共c兲Ⰷ⌬0/e. It is from this result that we are able to conclude that the region of finite␥ and infinitesimal EThbelongs to region D.

Another analytical result can be obtained for the case of perfectly symmetric coupling to the leads, i.e.,␥= 0. In this case, Eq.共3.16b兲 is solved by␮= 0 and the relation between ⌬ and V can be stated as

eV =⌬0

1 + ⌬2 ⌬02

1 − 4ETh 2 /⌬ 0 2

1 −⌬ 2 ⌬02

2. 共3.22兲 This result is plotted for several values of EThin Fig.6. It is from this result that we are able to conclude that the line segment ␥= 0, 0⬍ETh⬍⌬0/2

2 belongs to region B while the line segment ␥= 0,⌬0/2

2⬍ETh⬍⌬0/2 belongs to re-gion A. The ETh→0 limit of Eq. 共3.22兲 can be obtained by

considering a bulk superconductor and assuming a quasipar-ticle distribution function n共E兲=关共−eV/2−E兲+共eV/2 − E兲兴/2. It is also worth noting that the same result is ob-tained for a T junction where the stem of the T is a super-conductor and the bar a voltage-biased dirty normal-metal wire.10

Finally, we consider the I-V curves associated with the solutions⌬ and␮of Figs.3and4. The results are shown in Fig.7. From these curves we can infer the results that will be obtained in an experiment in which the voltage V is swept adiabatically from zero to several⌬0/e and back to zero. In region A of parameter space, there is a single current

associ-    0.0 0.5 0.0 0.5 1.0 0 0 γ ETh/∆0 A A A A B B B B C C C C D D D D N N 1/2√2 1/2√2 Q Q a b

FIG. 5. Schematic diagram of the partitioning of the ETh-␥ pa-rameter space into regions where the curve of⌬ versus V is of the types A, B, C, and D共Fig.3兲. The regions A, B, C, and D meet at

point Q. The line ETh=⌬0/2 separates the normal and supercon-ducting regions of parameter space. The dots in the figure indicate the parameter values that correspond to the curves in Fig.3. The inset shows the regions A, B, C, and D in the parameter space a-b of the polynomial V˜ 共⌬兲 of Eq. 共3.20兲. The topology of the ETh-␥ parameter space of the superconductor in the region of point Q can be understood by considering the topology of the parameter space of the polynomial. 0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 ∆ / ∆0 eV /∆0

FIG. 6. The order parameter⌬ versus voltage V for symmetric coupling to the leads, i.e.,␥=0, according to Eq. 共3.22兲. Different

curves correspond to different ETh. From the top to bottom curve we took ETh/⌬0= .01, 0.14, 0.26, 1/2

2共⯝0.35兲, 0.42, and 0.47. The curve corresponding to ETh= 1/2

2⌬0is plotted thicker than the others. For smaller ETh, curves are of type B with two nonzero values for⌬ at some voltages. For larger ETh, curves are of type A, with at most one nonzero⌬ at every voltage.

(9)

ated with each voltage. At some voltage V+ of order ⌬0/e, the system makes a phase transition to the normal state but this does not lead to a discontinuity in the current versus voltage curve. In contrast, in regions B, C, and D, the current will make discontinuous jumps as the voltage is swept. Hys-teresis will also be observed. Suppose the device is in region B of parameter space, as V is swept from zero upward, a voltage V+is crossed where the current makes a finite jump. After the jump, the system is in the normal state and the current is GNV. 关GN is the normal-state conductance of the

setup, cf. Eq.共3.18兲.兴 On the backward sweep from V⬎V+to zero, the system remains normal when V+ is reached. At some voltage V⬍V+the current jumps from its value GNV− in the normal state to a smaller value, signaling the onset of superconductivity. The behavior of the system in region C of parameter space is similar. The upward sweep of the voltage produces a jump in the current at a voltage V+. After the jump the system is normal and the current is given by I = GNV. The difference from region B appears when the

volt-age is swept back from V+to zero. At some voltage smaller than V+ the current starts deviating from its value in the normal state but there is no discontinuous jump yet. Even so, the system has turned superconducting. When the jump in current now occurs at V⬍V+, the system switches between two different superconducting states. Finally, for parameters in region D, the voltage sweep produces results similar to that in region C. The difference between regions C and D is that in D the system also jumps between two superconduct-ing states at V+during the forward sweep.

IV. DYNAMICS

We concluded the previous section with a discussion of hysteresis in the current-voltage characteristic of the super-conducting island. The conclusions we drew rely on the as-sumption that after the system is perturbed by a change in the bias voltage, it relaxes into a stationary state. The validity of this assumption is by no means obvious since the system is driven共by the bias voltage兲 and the stationary state is not an equilibrium state. Frankly, our own initial expectation was that the presence of a bias voltage would cause the dynamics

of 兩⌬共t兲兩 to be quasiperiodic or chaotic. We therefore did numerical simulations in order to investigate the dynamics of 兩⌬共t兲兩 in the presence of a bias voltage. Our main result is this: suppose the bias voltage assumes the constant value Vf

for times t⬎tf, then共contrary to our original expectations兲 at

tⰇtf the superconductor will always be found in one of the

stationary states associated with Vf. This is true regardless of

the history of the system prior to t⬍tf. In particular, the time

dependence of the bias voltage prior to tf does not matter.

Nor does the state of the superconductor prior to tf matter.

Only when there is more than one nonzero stationary solu-tion associated with Vf does the history of the system have

any bearing on its final state. In this case, the history of the system determines which of the possible stationary states eventually becomes the final state of the superconductor. For slowly varying voltages, the predictions of Sec.IIIregarding hysteresis are confirmed. In this section we discuss the nu-merics that yielded the above results.

For the purpose of numerics we find it advantageous not to take the sum over levels of the Green’s function as we did in the previous sections. Instead we work with the Green’s functions of each individual level. The advantage of this scheme is that it allows us to work with ordinary differential equations. From these differential equations it is straightfor-ward to construct a time series in which the next element can be calculated if the present elements are known. As far as we can see, no such “local in time” update equations exist for the Green’s functions summed over levels. Naturally there are disadvantages to working with the individual level Green’s functions as well; the number of equations to be solved numerically is increased significantly. As a result the calculation is computationally expensive and therefore time-consuming.

The Green’s functions of the individual levels obey the equations

共Hm⌺兲Gm= Gm共Hm⌺兲 = I. 共4.1兲

Here Hm differs from the operator H that appeared in Eq.

共2.5兲 in that it contains the energy ␧mof level m. It is

explic-itly given by Hm共t,t⬘兲 =␶0丢␩3␦共t − t⬘兲关it− hm共t兲兴, 共4.2a兲 hm共t兲 =

m−␮s共t兲 ⌬共t兲 ⌬共t兲 s共t兲 − ␧m

. 共4.2b兲

The operator hm共t兲 is the time-dependent Bogoliubov–de

Gennes Hamiltonian.11The self-energy ⌺ is the same as in

Sec. II B.

We measure energies from a point halfway between the chemical potentials of the leads. As a result the phases␾j共t兲

that appear in the reservoir self-energies are ␾r共l兲共t兲

= +共−兲␾共t兲/2 whereis related to the voltage V by V共t兲 =⳵t共t兲/e.

We parametrize the Green’s functions in terms of a set of auxiliary functions. This eliminates some redundancies that are present due to symmetries of the equations of motion. We start by noting that, since the retarded and advanced Green’s functions are related by Eq. 共2.1c兲, we do not need to

con-0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0 eI / GN ∆0 eV /∆0 A A B B C C D D

FIG. 7. The current I through the superconductor versus the voltage V across it. Curves A, B, C, and D correspond to the re-spective parameter values quoted in Figs.3and4. The dashed line shows the current through the system in the absence of superconductivity.

(10)

sider both. We work with the retarded Green’s function. We define a matrix rm共t,兲 that is related to Rm共t,t−␶兲 by the

equation

Rm共t,t −兲 = i␩3rm共t,␶兲, 共4.3a兲

rm共t,兲 = rm共0兲共t,␶兲␩0+ irm共t,␶兲 ·␩. 共4.3b兲

Here the component rm共0兲共t,兲 of rm共t,␶兲 is a scalar function

whereas the other three components are grouped into a vec-tor rm共t,␶兲 such that

rm共t,兲 = 关rm共1兲共t,兲,rm共2兲共t,兲,rm共3兲共t,␶兲兴. 共4.4兲

The vector ␩=共␩1,␩2,␩3兲 contains the Pauli matrices in Nambu space. Before the voltage bias between the leads is established 共i.e., for tⱕ0 and all兲, the functions rm共0兲共t,␶兲

and rm共t,␶兲 are real. When the equations of motion 关Eq.

共4.1兲兴 for the retarded Green’s function are rewritten in terms

r共0兲and r, we find that their reality is preserved at all times. Next we consider the Keldysh Green’s function. In order to calculate the time evolution of the order parameter, we only need to know the Keldysh Green’s function at coincid-ing times. Here the parametrization

Km共t,t兲 = i␩3km共t兲 ·␩, 共4.5兲

in terms of a real vector,

km共t兲 = 关km共1兲共t兲,km共2兲共t兲,km共3兲共t兲兴, 共4.6兲

is respected by the initial condition and preserved by the equations of motion.

From the equations of motion 关Eq. 共4.1兲兴, we derive

dif-ferential equations d dtrm共t,兲 = bm共t兲rm共t,兲 − rm共t,兲bm共t −␶兲, 共4.7兲 d dtkm共t兲 + 2bm共t兲 ⫻ km共t兲 + 2EThkm共t兲 = 4EThfm共t兲, 共4.8兲 for the matrix rm共t,兲 and the vector km共t兲. Equation 共4.8兲

with ETh= 0 was studied in Refs.8and9. In these references the dynamics of the order parameter of an isolated supercon-ductor was calculated. We see that coupling the system to leads introduces two terms proportional to ETh. One关on the left-hand side of Eq. 共4.8兲兴 can be considered a damping

term and is proportional to km共t兲. The other 关on the

right-hand side of Eq.共4.8兲兴 can be considered a driving or source

term.

In Eq.共4.8兲, bm共t兲 is a matrix and bm共t兲 a vector such that

bm共t兲 = ibm共t兲 ·␩, 共4.9a兲

bm共t兲 = 关Re ⌬共t兲,− Im ⌬共t兲,s共t兲 − ␧m兴. 共4.9b兲

The equation for km共t兲 contains a source term 4EThfm共t兲. The

vector fm共t兲 is given by fm共t兲 =

0 ⬁ d关rm共0兲共t,兲s共t,兲 − rm共t,兲s共0兲共t,兲 − rm共t,␶兲 ⫻ s共t,␶兲兴. 共4.10兲

In this equation the scalar function s共0兲共t,␶兲 and the vector

s共t,␶兲 parametrize the Keldysh component of the self-energy

as follows:

K共t,t⬘兲 = 2ETh␩3s共t,␶兲, 共4.11a兲 s共t,兲 = s共0兲共t,␶兲␩0+ is共t,␶兲 ·␩. 共4.11b兲 Referring back to Sec.II, where⌺Kis expressed in terms of

the Fourier transform of the reservoir filling factors, we find explicitly s共0兲共t,␶兲 = 1 ␲P

1 ␶

cos ␾共t兲 −共t −␶兲 2 , 共4.12a兲 s共t,␶兲 = − ␥ ␲␶

0,0,sin ␾共t兲 −共t −␶兲 2

. 共4.12b兲

By imposing self-consistency, the order parameter⌬共t兲 is expressed in terms of the components of km共t兲 as

⌬共t兲 =gs

2 m=−

km共1兲共t兲 − ikm共2兲共t兲, 共4.13兲

where the number of levels on the island is 2⍀+1. This makes the differential equations nonlinear since they contain terms in which ⌬共t兲 multiplies km and r. We eliminate the

dimensionless pairing strength g and the mean level spacingsin favor of⌬eq, the order parameter of the island in equi-librium, by means of the equilibrium self-consistency rela-tion 2 gs =

m=−⍀ ⍀ 1 ␰m 2 ␲arctan ␰m ETh , 共4.14兲 where ␰m=

m2 +⌬eq2. 共4.15兲 As with ⌬共t兲, the chemical potentials共t兲 is determined

by a self-consistency equation. The chemical potential takes into account the work that must be performed against the electric field of the excess charge on the superconductor in order to add more charge. Thus␮s共t兲 is related to the charge

of the island by␮s共t兲=e关Q共t兲−Q0兴/C, where C is the capaci-tance of the island. In this equation Q0 represents the fixed positive background charge and Q共t兲 is the combined charge of all the electrons on the island. Since the differential Eqs. 共4.7兲 and 共4.8兲 only depend on the difference ␮s共t兲−s共t

−␶兲, the positive background charge need not be specified. The charge Q共t兲 is related to the Keldysh Green’s function. Indeed, the average number nm共t兲 of electrons 共with spin

degeneracy included兲 in level m at time t is given by nm共t兲

=兵1−i Tr关Km共t,t兲兴其/2. Hences共t兲 is related to km by the

(11)

s共t兲 −s共t −␶兲 =

e2 2Cm=−

km共3兲共t兲 − km共3兲共t −␶兲. 共4.16兲

Finally, we have to specify the initial conditions for rm共t,兲 and km共t兲. We will assume for our simulations that

the voltage between the reservoirs is zero and the system is in zero-temperature equilibrium for times t⬍0. The corre-sponding initial condition at t = 0 is

rm共0兲共0,␶兲 = −␪共␶兲e−ETh␶cos共␰m␶兲, 共4.17a兲

rm共0,␶兲 =␪共␶兲e−ETh␶ sin共␰m␶兲 ␰m 共− ⌬eq,0,␧m兲, 共4.17b兲 km共0兲 = 1 ␰m 2 ␲arctan ␰m ETh共⌬eq,0,−␧m兲. 共4.17c兲 We are now ready to study the time evolution of ⌬共t兲 when a nonzero bias voltage V共t兲 between the leads is present for times t⬎0. In the calculations we report on here, we worked with ETh= 0.069⌬0and three different␥, namely, ␥= 0.05, ␥= 0.1, and ␥= 0.2. These all correspond to points from regions C and D in the ETh-␥parameter space of Fig.5. Hence, for each of the parameter choices, there is a bias voltage interval关V, V+兴 where there are more than one non-zero stationary solutions for兩⌬兩. The three curves of station-ary 兩⌬兩 versus V, corresponding to the different parameter choices, are plotted in the top panel of Fig.8.

For given ETh and ␥ we did two numerical runs with different time-dependent voltages V共t兲. The two voltages are plotted as functions of time in the bottom panel of Fig.8. In the first run we start by rapidly establishing a bias voltage V1⬍V. Rapid here means dV/dtⰇ⌬0ETh/e. In this case V changes by an amount of order⌬0/e—the scale at which the stationary solution for 兩⌬兩 depends on V—in a time that is

short compared to the relaxation time ETh−1.共Slow refers to the opposite limit.兲 We then keep the voltage constant at V1for a length of time of several ETh. This time interval is long enough for any transient behavior induced by the rapid change in V共t兲 to disappear. We then slowly increase the bias voltage until we reach a bias voltage Vf苸关V, V+兴 for which more than one nonzero stationary solutions exist. In the sec-ond run we start by rapidly establishing a bias voltage V2 ⬎V+. We keep the voltage fixed at V2 for a time of several ETh−1. We then slowly decrease the voltage to Vf. The values of

V1, V2, and Vf were chosen V1= 0.83⌬0, V2= 1.76⌬0, and Vf= 1.34⌬0. The calculations were performed with 501 equally spaced levels with level spacing␦s= 0.018⌬0and the capacitance was chosen C = 0.1e2/⌬

0.

The resulting兩⌬兩 are plotted as functions of time in Fig.9. They first show that after the initial rapid change in the bias voltage the system always relaxes into a stationary state con-sistent with the new voltage. The relaxation takes a time of the order ETh−1. Second, if the system is in a stationary state and the bias voltage is changed slowly, then兩⌬共t兲兩 adiabati-cally tracks the stationary solution corresponding to the in-stantaneous value of the voltage. This is seen most clearly in Fig.10where we plot兩⌬共t兲兩 as a function of V共t兲 and com-pare this to the stationary兩⌬兩 vs constant V curves. Our pre-diction about hysteresis is confirmed. Systems with different histories end up in different stationary states at the same voltage bias. If the voltage is slowly swept from a small initial voltage to Vf苸关V, V+兴, a stationary state with a large value for兩⌬兩 is reached. If the voltage is swept from a large initial voltage to Vf苸关V, V+兴, a stationary state is reached that corresponds to a small value of 兩⌬兩. We must mention here that we observe some slow drift共too slow to be visible in Fig.9兲 in 兩⌬兩 after the voltage has reached Vf. The value of

兩⌬兩 seems to increase linearly at a rate d兩⌬兩/dt⬃10−4 0 2 . Within the numerical accuracy of the calculation, this is

neg-FIG. 8. Top panel: The stationary solutions for ⌬ vs the bias voltage V, corresponding to the parameters used in generating Fig.

9. The outer curve 共dashed兲 corresponds to ETh= 0.069⌬0 and ␥ = 0.2. The middle curve 共solid兲 corresponds to ETh= 0.069⌬0 and ␥=0.1. The inner curve 共dot dashed兲 corresponds to ETh= 0.069⌬0 and␥=0.05. The vertical lines indicate Vfand V2.共V1is beyond the left edge of the figure.兲 Bottom panel: the time dependence of the voltage. The upper共solid兲 curve corresponds to the solid curves of ⌬ vs t in Fig.9. The lower共dashed兲 curve corresponds to the dashed curves of⌬ vs t in Fig.9.

FIG. 9. The amplitude of the order parameter as a function of time. All curves are for ETh= 0.069⌬0. The top, middle, and bottom panels correspond to␥=0.05, ␥=0.1, and ␥=0.2, respectively. The dashed curves correspond to a voltage that is increased from V1 = 0.83⌬0to Vf= 1.34⌬0. The solid curves correspond to the voltage being decreased from V2= 1.76⌬0to Vf= 1.34⌬0. The vertical lines

indicate the time interval in which the voltage changes from either

V1or V2to Vf. The thin horizontal lines correspond to the stationary

(12)

ligible and we believe the drift is simply an artifact of the numerics.

In our data there is one exception to the rule of adiabatic evolution. In the middle panel of Fig. 9, 兩⌬共t兲兩 takes much longer that ETh−1to respond when the voltage is changed from V2to Vf. Hence,兩⌬共t兲兩 as a function of V共t兲 does not track the

stationary solution in this instance. The reason is the follow-ing: for a voltage V = V2, the only stationary solution has ⌬ = 0. When the voltage is decreased to Vf, a nonzero

station-ary solution for⌬ exists. However ⌬=0 is still a valid state, albeit unstable. The time it takes the system to diverge from the unstable state is not determined by ETh but rather by small numerical errors that perturb the unstable state.

One possible explanation for the observed stability of the stationary states is overdamping. According to this hypoth-esis, if we decrease the Thouless energy further, thereby de-creasing the damping, the stationary solutions will become unstable. Some evidence for the hypothesis might be visible in Fig. 9. After the voltage is changed rapidly, we might expect 兩⌬共t兲兩 to perform damped oscillations while relaxing to the new stationary state. However in Fig.9no such oscil-lations are visible, apparently implying that the relaxation rate is larger than the oscillation frequency. There is however another possible explanation for the lack of oscillatory be-havior after an abrupt change in V. The argument is that an abrupt change in V cannot be communicated to the system abruptly but only at a rate comparable to the damping rate ETh. This is because the superconductor learns of the change in voltage by the same mechanism as by which damping occurs, that is, by tunneling of particles between the leads and the island. Hence the response of the order parameter is always gradual.

How do we test whether overdamping hypothesis is true or false? Ideally we would have liked to repeat the above numerical calculation with a smaller value of EThand see if the stationary states are still stable. However, the value ETh = 0.069⌬0 that we used above is close to the smallest value for which we can do reliable numerics in reasonable time. Since we cannot make EThsmaller, we resolve the issue of overdamping as follows. We compare the dynamics of ⌬ after an abrupt change in the pairing interaction strength g at ETh= 0.069⌬0to the dynamics after a change in g at ETh= 0.32 We know that in the isolated system, 共ETh= 0兲 兩⌬兩 will per-form persistent oscillations.8,9The period of oscillation gives

a typical time scale for the internal dynamics of⌬. If, in the open system共i.e., ETh⫽0兲, we observe a few damped oscil-lations共the more the better兲 in 兩⌬兩 before the system relaxes to equilibrium, it means that damping occurs at a timescale larger than that of the internal dynamics of the supercon-ductor. In this case the hypothesis of overdamping is discred-ited.

In our numerical implementation of the above, we work with the following parameters: the initial pairing interaction is such that for t⬍0, ⌬=⌬i. The increased pairing interaction

strength corresponds to an equilibrium value of the order parameter ⌬f= 20⌬i. The persistent oscillations of 兩⌬共t兲兩 in

the isolated system are shown in the blue curve in Fig. 11. We repeat the calculation, which is now for a superconductor connected to leads. We use a Thouless energy ETh= 0.075⌬f.

The result for兩⌬共t兲兩 in the presence of leads is the solid curve in Fig.11. We see that兩⌬共t兲兩 eventually decays to a constant, as expected. The extent of the damping is such that several oscillations are completed within the decay time. Hence we conclude that the numerical results that we obtained previ-ously are outside the regime of overdamping. It follows that the lack of oscillatory behavior in Fig. 9 is due to the fact that the superconductor only gradually becomes aware of a change in the voltage.

V. CONCLUSION

We have studied a voltage-biased NISIN junction, i.e., a superconducting island connected to normal leads by means

FIG. 10. The amplitude of the order parameter兩⌬共t兲兩 as a func-tion of voltage V共t兲. The parameter values of the three panels are the same as those in Fig.9, i.e., all curves are for ETh= 0.069⌬0. The top, middle, and bottom panels correspond to␥=0.05, ␥=0.1, and ␥=0.2, respectively. The thick dashed curves correspond to a volt-age that is increased from V1= 0.83⌬0 to Vf= 1.34⌬0. The thick solid curves correspond to the voltage being decreased from V2 = 1.76⌬0to Vf= 1.34⌬0. The thin dashed lines represents the station-ary value of兩⌬兩 vs V, as calculated in Sec.IIIand plotted in Fig.8.

FIG. 11. The order parameter versus time after the pairing strength was increased from⌬i= 0.05⌬fto⌬fabruptly at t = 0. The

dashed curve is for an isolated superconductor while the solid curve is for a superconductor connected to leads. For this case a Thouless energy ETh= 0.075⌬f was used. The data was obtained using 501

equally spaced levels with level spacing ␦s= 0.02⌬f. The capaci-tance was chosen C = 0.1e2/⌬

Cytaty

Powiązane dokumenty

Zarówno SLD (Wniosek o stwierdzenie niezgodnoœci ustawy z Konstytucj¹ RP z dnia 2 lutego 2006 r.), jak i PO (Wniosek o stwierdzenie niezgodnoœci ustaw z Konstytucj¹

Onder 'rollend materieel op civiele vliegvelden' wordt verstaan alle voertuigen, aangedreven of niet aangedreven, die op het platform van een vliegveld voor de burgerluchtvaart

Verder is gekeken naar de door Figee toegepaste evenaar, die de hangkabels aan de spreader bevestigt, en de

Dokonując przeglądu dyskusji toczących się wokół pojęcia seksualizacji, rozwaŜanej na poziomie jednostkowym (w odniesieniu do funkcjonowania osoby lub grupy osób),

gov/data/realtime/hmi_igr/1024/latest.html, przeprowad ź tygodniow ą ob- serwacj ę dowolnie wybranej plamy słonecznej w celu wyznaczenia mo- mentu, w którym jej odległo ść od

W części analizowanych modlitwach po komunii mszału Pawła VI odwołujących się do proble­ mu zadatku zasadnicza myśl tej grupy modlitw nawiązuje do stwier­ dzenia,

W cyklu o Anne Shirley Lucy Maud Montgomery także częściowo powiela model obowiązujący w literaturze dla dziewcząt na przełomie XIX i XX wieku; wszakże bohaterka

Prawdziwość historii badacza może być zaś weryfiko- wana przez zdecydowanie większe grono osób, bowiem w odróżnieniu od narracji osób badanych jego narracja nie jest