1: \documentclass[twocolumn,showpacs,prb,preprintnumbers,amsmath,amssymb,floatfix]{revtex4}
2:
3: \usepackage{epsfig,psfrag}
4: \usepackage{dcolumn}
5: \usepackage{bm}
6: \usepackage{graphicx}
7: \usepackage{color}
8:
9: \begin{document}
10:
11: \title{Superconducting nonequilibrium transport through a weakly interacting
12: quantum dot}
13: \author{L.~Dell'Anna, A.~Zazunov, and R.~Egger}
14: \affiliation{Institut f\"ur Theoretische Physik, Heinrich-Heine-Universit\"at,
15: D-40225 D\"usseldorf, Germany}
16: \date{\today}
17:
18: \begin{abstract}
19: We study the out-of-equilibrium current through an interacting
20: quantum dot modelled as an Anderson impurity contacted by
21: two BCS superconductors held at fixed voltage bias.
22: In order to account for multiple Andreev reflections, we develop a
23: Keldysh Green's function scheme perturbative
24: in the dot's interaction strength. We find an unexpected enhancement of the
25: current due to repulsive interactions for small to intermediate lead-to-dot couplings.
26: \end{abstract}
27: \pacs{73.63.-b, 74.45.+c, 74.50.+r}
28:
29: \maketitle
30:
31: \section{Introduction}
32:
33: Superconducting transport through low-dimensional nanoscale structures
34: is currently attracting considerable interest. Gate-tunable Josephson
35: currents through nanowire-based quantum dots have been reported,\cite{vandam}
36: and similar setups have been realized using (short)
37: carbon nanotubes\cite{tube,buitelaar} and
38: metallofullerene molecules.\cite{fulleren}
39: Nonequilibrium transport in such systems contacted
40: by superconducting electrodes has been a particular focus of recent
41: experimental effort,\cite{buitelaar,fulleren,jorgensen,lieber,jarillo,christian,lindelof}
42: mainly caused by an interesting interplay between interaction
43: effects (on the quantum dot) and superconducting correlations
44: (due to the electrodes).
45: One remarkable consequence is the observation of an
46: `even-odd' effect (as a function of the dot's occupation number)
47: in the conductance.\cite{christian,lindelof}
48: This effect is presumably caused by the absence or presence of Kondo correlations.
49: In this paper, we analyze superconducting transport through an interacting
50: quantum dot for the simplest case of a
51: single spin-degenerate level with repulsive on-site interaction energy $U>0$
52: (Anderson impurity), contacted by two wide $s$-wave BCS superconducting electrodes
53: with identical gap $\Delta$.
54: For simplicity, we assume that both lead-to-dot couplings (hybridizations) are equal,
55: $\Gamma_L=\Gamma_R=\Gamma$, and consider
56: the two electrodes held at potential difference (voltage bias) $V$.
57: Under a systematic perturbative expansion in the interaction strength $U$,
58: we compute the $I-V$ characteristics, in particular for the interesting
59: subgap regime $eV<2\Delta$, where multiple Andreev reflection (MAR)
60: processes provide the dominant transport mechanism.
61: A theory of coherent MAR has been originally developed for
62: superconducting point contacts,\cite{mar}
63: with the essential assumption that charging interaction effects
64: inside the contact can be neglected.
65: The problem of resonant MAR through a noninteracting quantum level has been treated
66: in Refs.~\onlinecite{ingerman,cuevas,wendin}.
67:
68: While the interacting problem in equilibrium has been
69: theoretically studied by many authors,\cite{siano}
70: the corresponding nonequilibrium problem is more difficult and
71: far less understood. Previous approaches can be broadly grouped in three classes.
72: (i) By ignoring MAR processes in the Coulomb blockade regime,
73: additional side-peaks in the differential conductance
74: at $eV=2 \Delta$ and $2(\Delta + U)$ were predicted,\cite{kang}
75: reflecting the singularity of the BCS spectral density of the leads.
76: (ii) Different mean-field schemes have been proposed,\cite{cuevas,avishai,slave}
77: based on slave-boson or Hubbard-Stratonovich-path-integral approaches.
78: These calculations predict an overall suppression of the current by the
79: interactions. This suppression is obtained only
80: for sufficiently repulsive interactions, while there
81: is no interaction effect for weak interaction.\cite{avishai}
82: (iii) A Fermi liquid approach
83: valid in the deep Kondo limit has been proposed.\cite{alfredo}
84: Here we do not discuss the Kondo regime, but instead focus on the
85: limit of weak interactions, $U/ \Gamma<1$, where a
86: controlled perturbative expansion in the small parameter $U/\Gamma$
87: is possible. Note that this approach still allows for arbitrary
88: ratio $\Gamma/\Delta$. Such calculations have been carried out
89: for normal-conducting ($\Delta=0$) electrodes recently,\cite{normal,hamasaki,eg}
90: and we here generalize them to superconducting electrodes.
91: The case $U< \Gamma$ is of experimental relevance for
92: the understanding of superconducting transport through quantum dots
93: or molecules with good lead-to-dot couplings.
94:
95: The structure of the remainder of this article is as follows.
96: In Sec.~\ref{sec2}, we discuss our perturbation theory approach
97: to superconducting transport through an Anderson dot, and its
98: numerical implementation. Results for the current-voltage
99: characteristics are shown and discussed in Sec.~\ref{sec3}.
100: The appendix contains qualitative arguments for the
101: current enhancement found at $\Gamma<\Delta$, based on an
102: evaluation of the Josephson current. We often set $\hbar=e=1$.
103:
104: \section{Perturbative approach to superconducting transport}
105: \label{sec2}
106:
107: We consider the canonical Anderson impurity model, $H=H_D+H_T
108: +H_L+H_R$, where a single-level dot with spinful fermion $d_\sigma$ ($H_D$) is
109: tunnel-coupled ($H_T$) to left/right superconducting reservoirs $H_{L/R}$
110: held at chemical potential difference $eV$. The isolated dot corresponds to
111: ($n_\sigma=d^\dagger_\sigma d^{}_\sigma=0,1$)
112: \begin{equation} \label{model}
113: H_D= E_0(n_\uparrow +n_\downarrow) + U n_\uparrow n_\downarrow
114: = \epsilon_0 (n_\uparrow+n_\downarrow)
115: -\frac{U}{2}(n_\uparrow - n_\downarrow)^2 .
116: \end{equation}
117: The `noninteracting' model below is taken to contain the interaction level
118: shift $\epsilon_0=E_0+U/2$ of the bare level $E_0$.
119: The leads are described by a pair of $s$-wave BCS Hamiltonians
120: in the standard wide-band limit.
121: We are interested in the $V\ne 0$ case,
122: and take the same real-valued gap parameter $\Delta>0$ for both electrodes.
123: Using the Nambu vector $\Psi^T_{j,k}=(\psi_{j,k,\uparrow},
124: \psi^\dagger_{j,-k,\downarrow})$ for electrons in lead $j=L/R$, we thus have
125: (we put $\hbar=e=1$ in intermediate steps)
126: \begin{equation} \label{h2}
127: H_{j} = \sum_k \Psi_{jk}^\dagger \left (
128: (k^2/2m -\epsilon_F)\sigma_z + \Delta \sigma_x \right) \Psi_{jk},
129: \end{equation}
130: with Pauli matrices $\sigma_{i}$ ($\tau_{i}$)
131: in Nambu (Keldysh) space.
132: Using the Nambu vector $d=(d^{}_\uparrow, d^\dagger_\downarrow)^T$
133: and $\Gamma=\pi\nu_0 |t_0|^2$ for
134: (normal) lead density of states $\nu_0$, the lead-dot coupling is
135: \begin{equation}\label{h3}
136: H_T = t_0 \sum_{k,j=L/R=\pm} \Psi^\dagger_{jk}
137: \sigma_z e^{\pm i\sigma_z V t/2} d + {\rm h.c.},
138: \end{equation}
139: where the voltage $V$ enters via the time-dependent phase.
140: We now define the Keldysh-Nambu Green's function for the dot fermion,
141: \begin{equation}\label{green}
142: G^{ss'}_{\alpha\alpha'}(t,t') = -i\langle {\hat T}_C [d_\alpha (t_s)
143: d^\dagger_{\alpha'}(t_{s'}) ]\rangle,
144: \end{equation}
145: where $\alpha,\alpha'=1,2$ ($s,s'=1,2$) are Nambu (Keldysh) indices,
146: and $\hat T_C$ is the time-ordering operator along the Keldysh
147: contour. Accordingly, $t_s$ denotes a time
148: taken on branch $s$ of the Keldysh contour.
149: It is convenient to use the Fourier decomposition \cite{zazu}
150: \begin{equation}\label{greennew}
151: G^{ab}(t,t') = \sum_{n, m = -\infty}^{\infty}
152: \int_F \frac{d \omega}{2 \pi} \,
153: e^{-i \omega_n t + i \omega_m t'} \, G^{ab}_{nm}(\omega) ,
154: \end{equation}
155: where $a,b=1,\ldots,4$ denotes Nambu-Keldysh indices defined
156: by $a=\alpha+2(s-1)$, and $\omega_n = \omega + n V$
157: for $\omega$ within the 'fundamental' domain
158: $F\equiv [-V/2,V/2]$.
159: For fixed $\omega\in F$, the Dyson equation
160: for the full Green's function $\check{G}$
161: (the check notation refers to the Keldysh-Nambu structure),
162: \begin{equation}\label{g1}
163: \check{G}^{-1} = \check{G}_{0}^{-1} - \check{\Sigma},
164: \end{equation}
165: then becomes a matrix equation suitable for numerical inversion. Here,
166: interaction effects are encoded in the self-energy $\check{\Sigma}$.
167: After integrating out the lead fermion degrees of freedom,
168: the noninteracting Green's function $\check{G}_0$ is
169: \begin{equation} \label{g0}
170: \check{G}_{0,nm}^{-1}(\omega) =
171: (\omega_n-\epsilon_0 \sigma_z) \tau_z \delta_{nm}
172: - \Gamma \sum_{j=L/R} {\check \gamma}_{j,nm}(\omega).
173: \end{equation}
174: The self-energy due to tracing out the respective lead
175: is given by the Nambu matrix
176: \begin{eqnarray} \label{selfen}
177: && \check{\gamma}_{j=L/R=\pm,nm}(\omega) = \\ \nonumber
178: && \left( \begin{array}{cc}
179: \delta_{nm} \, \check{X}(\omega_n \mp V/2) &
180: \delta_{m, n \mp 1} \, \check{Y}(\omega_n \mp V/2) \\
181: \delta_{m, n \pm 1} \, \check{Y}(\omega_n \pm V/2) &
182: \delta_{nm} \, \check{X}(\omega_n \pm V/2)
183: \end{array} \right)
184: \end{eqnarray}
185: with Keldysh matrices $\check{Y}(\omega) = - \Delta\check{X}(\omega)/\omega$
186: and
187: \[
188: \check{X}(\omega) = \left\{ \begin{array}{ll}
189: - \frac{\omega}{\sqrt{\Delta^2 - \omega^2}} \tau_z ,
190: & |\omega|<\Delta\\
191: \frac{i|\omega|}{\sqrt{\omega^2-\Delta^2}} \left( \begin{array}{cc}
192: 2 f_\omega - 1 & - 2 f_\omega \\ 2 f_{-\omega} & 2 f_\omega - 1
193: \end{array} \right) , & |\omega|>\Delta \end{array}\right.
194: \]
195: where $f_\omega=1/(1+e^{\omega/k_B T})$ is the Fermi function.
196: The steady-state dc current through the left/right junction
197: then follows as
198: \begin{equation} \label{dc_current}
199: I_{L/R} = \mp 2\Gamma\ {\rm Re} \sum_{nm} \int_F \frac{d \omega}{2 \pi}
200: {\rm tr}\left[ \sigma_z\check{\gamma}_{L/R,nm}(\omega)
201: \check{G}_{mn}(\omega) \right]^{+-},
202: \end{equation}
203: where the trace is only over Nambu space, and
204: $(+-)$ refers to the $(12)$ Keldysh component.
205: Current conservation, $I_L=I_R=I$, is fulfilled for all results below.
206:
207: \subsection{First order}
208:
209:
210: \begin{figure}[h!]
211: \scalebox{0.25}{\includegraphics{fig1a}}
212: \hspace{1cm}
213: \scalebox{0.25}{\includegraphics{fig1b}}
214: \caption{ \label{f1} Electron-electron interaction
215: self-energy diagrams taken into account in this paper:
216: (a) First order (left). (b) Second order (right). }
217: \end{figure}
218:
219:
220: Since the exact self-energy $\check{\Sigma}$ is not known, we proceed in a
221: perturbative fashion, starting with the first-order self-energy
222: in Fig.~\ref{f1}(a), made self-consistent by using the
223: full $\check{G}$ in the diagram.
224: It is convenient to introduce the four-point vertex, cp.~also
225: Ref.~\onlinecite{normal},
226: \begin{equation}\label{lambda}
227: \Lambda_{abcd} = \frac{1}{2}\Big(\delta_{ab}\tau_{cd}+\delta_{cd}\tau_{ab}-\delta_{ad}\tau_{cb}
228: -\delta_{cb}\tau_{ad}\Big),
229: \end{equation}
230: where $\tau=\sigma_0 \tau_z = {\rm diag}(1,1,-1,-1)$. Only 8 out
231: of the possible 256 entries of the tensor $\Lambda$ are nonzero,
232: with values $\pm 1$.
233: With this convention, the complete first-order self-energy is given by
234: \begin{equation}\label{self1}
235: \Sigma^{(1) ab}_{n,n+m} =
236: i\, {U} \Lambda_{abcd} \sum_{n'} \int_F\frac{d\omega'}{2\pi}
237: \, G^{dc}_{n',n'+m}(\omega'),
238: \end{equation}
239: where we use the sum convention for $c,d$.
240: Note that this self-energy is independent of $(n,\omega)$, but still
241: depends on the `off-diagonal' frequency index $m$.
242: In time representation, $m\ne 0$ contributions come with phase factors
243: $e^{\pm 2im V t}$ and thus correspond to anomalous (pairing) correlations.
244: The presence of the off-diagonal harmonics in the self-energy is
245: a consequence of the coherent MAR transport regime considered here.
246: The role of coherence is particularly important
247: for the interplay between MAR processes and charging effects in a quantum dot
248: with a relatively strong coupling to the leads.
249:
250: Let us at that stage briefly compare the first-order
251: self-consistent approach based on Eqs.~(\ref{g1}) and (\ref{self1})
252: to the mean-field approximation of Ref.~\onlinecite{avishai}.
253: The latter effectively considers only the time averaged components
254: of the self-energy (\ref{self1}) corresponding to the two first terms of Eq.~(\ref{lambda}),
255: thereby discarding all harmonics with $m\ne 0$ and exchange terms in Eq.~(\ref{self1}).
256: The resulting self-energy contributions taken into account in
257: Ref.~\onlinecite{avishai} correspond to $\gamma_+\tau_0$ and
258: $\gamma_-\tau_z$. The scalar constants $\gamma_\pm$
259: can be written as
260: \begin{eqnarray}\nonumber
261: \gamma_+ &=& i\frac{U}{2}\sum_{n'}
262: \int_F\frac{d\omega'}{2\pi}\,{\rm Tr} \left(
263: \tau_z \check{G}_{n' n'}(\omega') \right),
264: \\ \label{zaikin1}
265: \gamma_-&=& i\frac{U}{2}\sum_{n'} \int_F\frac{d\omega'}{2\pi}
266: \,{\rm Tr}\left( \check{G}_{n'n'}(\omega')\right),
267: \end{eqnarray}
268: where the trace is over both Nambu and Keldysh spaces.
269: Under this approximation, there is no interaction effect on the
270: current $I(V)$ below some critical value $U_c$. In fact,
271: nontrivial stable solutions $\gamma_\pm\ne 0$
272: for the self-consistency equation
273: (\ref{zaikin1}) exist only for $U>U_c$,\cite{avishai} where $U_c$
274: depends on $V,\Gamma$ and $\Delta$.
275: The symmetry-broken phase with $\gamma_-\ne 0$ corresponds to
276: a spin-polarized dot, and one then finds a Coulomb blockade
277: suppression of the current.\cite{avishai}
278: However, for normal leads, serious problems with spin-polarized
279: out-of-equilibrium mean-field solutions
280: for the Anderson dot have been identified recently,\cite{zarand}
281: and those arguments also apply to the superconducting case.
282: A typical value is $U_c\approx\Gamma$ for $V\approx \Gamma \approx \Delta/2$,
283: and we shall always limit ourselves to $U<U_c$
284: where no such problems arise.
285: In our calculations, the actual value of $U_c$ follows from the numerical
286: solution of the self-consistency problem, and we can thereby ensure that
287: no spin-polarized solutions are present. In contrast to the mean-field
288: scheme of Ref.~\onlinecite{avishai}, however, the full
289: first-order self-consistent
290: approach generates sizeable interaction corrections even for
291: small $U$, see Sec.~\ref{sec3}. These corrections are not
292: just a matter of numerical accuracy but reflect the importance
293: of on-dot pairing terms.
294:
295: \subsection{Second order}
296:
297: To go beyond the self-consistent first-order approximation
298: given by Eq.~(\ref{self1}), we have to evaluate the
299: second-order diagram shown in Fig.~\ref{f1}(b).
300: Due to the large numerical effort in evaluating this diagram, we restrict
301: ourselves to a non-selfconsistent scheme at this point, i.e., we use the
302: solution $\check{G}$ of the first-order problem to evaluate
303: $\check{\Sigma}^{(2)}$.
304: As is well known,\cite{baym,hershfield} under such a scheme
305: current conservation is only ensured
306: for the particle-hole symmetric case, $\epsilon_0=0$, see also
307: Ref.~\onlinecite{zazu}. We therefore show second-order
308: results only for $\epsilon_0=0$.
309: The second-order Nambu-Keldysh self-energy reads in time representation
310: (sum convention)
311: \begin{eqnarray} \label{self2}
312: \Sigma^{(2) ab}(t,t') &=& \frac{U^2}{2}
313: \Lambda_{afgh} \Lambda_{ebcd} \\ \nonumber &\times&
314: G^{fe}(t,t') G^{dg}(t',t) G^{hc}(t,t').
315: \end{eqnarray}
316: To avoid numerically expensive frequency convolutions, it is convenient to
317: first compute the Green's function in time representation
318: according to Eq.~(\ref{greennew}),
319: then evaluate the self-energy in Eq.~(\ref{self2}), and
320: finally transform this result back to frequency
321: space to use it in the Dyson equation (\ref{g1}).
322: Notice that Eq.~(\ref{self2}) corresponding to the skeleton
323: diagram in Fig.~\ref{f1}(b) represents only a part
324: of all possible second-order contributions.
325: The rest is given by the time-local piece
326: \[
327: -U^2\Lambda_{abcd}
328: \Lambda_{efgh}\int dt^{\prime\prime}\,{G}^{de}(t,t^{\prime\prime})\,
329: {G}^{hg}(t^{\prime\prime},t^{\prime\prime})\,
330: {G}^{fc}(t^{\prime\prime},t) ,
331: \]
332: which has already been taken into account by our self-consistent
333: first-order solution.
334:
335: \subsection{Calculation of current}
336:
337: The numerical implementation of the above perturbative approach is straightforward.
338: To evaluate the current $I(V)$ from Eq.~(\ref{dc_current}),
339: we partition the frequency summations into windows of width $V$, and impose
340: a bandwidth cutoff $\omega_c$, such
341: that $|\omega_n| < \omega_c$. In our calculations, we use $\omega_c=10\Delta$,
342: but the precise choice is not critical.
343: We then discretize the fundamental frequency domain $F$ with a step-size $\delta\omega$,
344: and use a
345: fast Fourier transform routine to switch between time and frequency
346: representations. (The efficient evaluation of the second-order self-energy
347: requires to employ the time representation, while the Dyson equation needs
348: the frequency representation.)
349: Typically, we found $\delta \omega = 0.005 \Delta$
350: sufficient for convergence. The matrix inversion in Eq.~(\ref{g1}) is then performed
351: for each $\omega \in F$ separately,
352: involving matrix dimensions of the order of $\omega_c/|V|$.
353: We refer to Ref.~\onlinecite{zazu} for further details of the numerical
354: implementation in the related case of a phonon-mediated interaction.
355:
356: In a first step, we solve the first-order self-consistent
357: problem posed by Eqs.~(\ref{g1}), (\ref{g0}) and (\ref{self1}).
358: This solution proceeds iteratively, where the stability or instability
359: of the solution for $\check{G}$ is checked carefully by probing small deviations
360: around it. The iterative solution can in fact
361: be carried out with very modest computational effort,
362: and quickly converges to a unique solution (as long as $U<U_c$).
363: In a second step, we then use this converged first-order
364: Green's function to evaluate
365: $\check{\Sigma}^{(2)}$ according to Eq.~(\ref{self2}),
366: and to finally compute the current from Eq.~(\ref{dc_current}).
367: The second-order calculation is
368: quite time-consuming for low bias voltage,
369: where many MAR orders need to be taken into account, and we
370: therefore restricted our
371: calculations to $eV/\Delta\geq 0.2$. As consistency
372: check for our numerical code, we have reproduced known results
373: for $U=0$, see Refs.~\onlinecite{ingerman,cuevas,wendin},
374: and the corresponding perturbative-in-$U$ results for normal-conducting
375: leads ($\Delta=0$),
376: see Refs.~\onlinecite{normal,hamasaki}.
377: We have also reproduced the respective
378: results of Ref.~\onlinecite{avishai} when
379: implementing their approximations.
380: As additional check, Green's function sum rules,
381: such as ${\rm tr}
382: \left[ \tau_z \sigma_z \check{G}(t,t) \right] = 0$ at coinciding times,
383: have been verified.
384:
385: \section{Results and discussion}\label{sec3}
386:
387: Next we discuss numerical results obtained
388: under the perturbative approach described in Sec.~\ref{sec2}.
389: All results are for $T=0$, and unless noted otherwise, we put $U/\Gamma=0.5$
390: which is sufficiently small to ensure $U<U_c$ for all investigated $V/\Delta$
391: and $\Gamma/\Delta$ but large enough to produce significant interaction
392: corrections to the $I-V$ characteristics.
393: We will focus on the most interesting subgap regime, $eV<2\Delta$.
394: The excess current $I_{exc}=\lim_{V\to\infty} [I(V,\Delta)-I(V,\Delta=0)]$
395: has also been computed. The interaction contribution $\delta I_{exc}$
396: to this quantity turns out to be generally small, similar to what
397: is found for the case of phonon-mediated interactions.\cite{zazu}
398: Remarkably, this interaction correction is positive for $\Gamma\alt
399: \Delta$, pointing towards a current enhancement. This trend is
400: quite generic and discussed next.
401:
402: \begin{figure}[ht!]
403: \scalebox{0.66}{\includegraphics{fig2.eps}}
404: \caption{\label{f2}
405: Interaction correction to the current (currents are always plotted
406: in units of $e\Delta/h$) from the self-consistent first-order
407: approach, for $U/\Gamma=\Gamma/\Delta=0.5$ and various $\epsilon_0/\Delta$.
408: The inverse voltage scale is taken to compare with standard MAR features.
409: Inset: Full $I-V$ curves for same parameters.}
410: \end{figure}
411:
412:
413: \begin{figure}[ht!]
414: \scalebox{0.65}{\includegraphics{fig3.eps}}
415: \caption{\label{f3}
416: Same as Fig.~\ref{f2} but for $\epsilon_0=0$. The dashed
417: curve gives the first-order self-consistent result, while
418: the solid curve includes also the second-order contribution.}
419: \end{figure}
420:
421: \begin{figure}[ht!]
422: \scalebox{0.65}{\includegraphics{fig4.eps}}
423: \caption{\label{f4}
424: Same as Fig.~\ref{f3} but for $\Gamma=2\Delta$.
425: Inset: $I-V$ curve from second-order perturbation theory and for $U=0$.}
426: \end{figure}
427:
428: Let us start by showing results obtained from the first-order self-consistent
429: scheme (i.e., without the second-order self-energy). In that case, by
430: virtue of self-consistency, we have
431: the freedom to vary $\epsilon_0$ without spoiling current
432: conservation.\cite{baym,hershfield}
433: Representative numerical results
434: for the voltage-dependent interaction correction to the current, $I(U)-I(U=0)$,
435: are shown for $\Gamma/\Delta=0.5$ in Fig.~\ref{f2}.
436: For all results shown here, we have $U<U_c$, and
437: the approach of Ref.~\onlinecite{avishai}
438: would not yield any interaction correction.
439: However, we find significant interaction effects for $U<U_c$
440: within the full first-order self-consistent approach.
441: These effects are due to the time-dependent ($m\ne 0$)
442: parts of the self-energy (\ref{self1}), which contain
443: pairing order parameters on the dot.
444: For instance, at the symmetric point $\epsilon_0=0$, we
445: find by perturbation theory in $U$ that $\gamma_\pm=0$, but the
446: complex-valued $m=1$ pairing term
447: \begin{equation}\label{delta}
448: \delta \equiv \Sigma_{n,n+1}^{(1),12}
449: \end{equation}
450: stays finite. This $\omega$-independent off-diagonal Nambu component of
451: the self-energy, absent in the normal ($\Delta=0$)
452: case, describes the effect of
453: interactions on the proximity-induced pairing
454: correlation on the dot. At the mean-field level, $\gamma_-$
455: is dominant for large $U$, while terms like $\delta$ dominate for small $U$.
456: Similar contributions with $|m|>1$ exist and are kept
457: in our self-consistent first-order calculations, but they turn out to be
458: significantly smaller.
459:
460: Quite remarkably, we find $I(U)>I(U=0)$
461: for most voltages and/or dot level energies $\epsilon_0$, pointing
462: to an {\sl enhancement of the MAR-mediated current by repulsive interactions}.
463: We have persistently found this
464: unexpected feature throughout the parameter regime
465: $\Gamma<\Delta$, also when including the second-order self-energy,
466: see below. A similar (but weaker) enhancement can be
467: found analytically for the critical Josephson current
468: of this system in equilibrium, see Appendix.
469: The current enhancement is reminiscent yet different from
470: the `antiblockade' behavior due to dynamical Coulomb blockade
471: effects on MAR transport discussed in Ref.~\onlinecite{alf3}.
472: It is also consistent with the crossover
473: from current enhancement to decrease with growing $\Gamma$
474: for phonon-mediated interactions and normal-conducting leads.\cite{eg}
475:
476: To illustrate the role of the second-order contribution for $U/\Gamma=0.5$,
477: we now focus on the symmetric case $\epsilon_0=0$, first taking again
478: $\Gamma/\Delta=0.5$. The results of the first- and second-order calculations
479: are compared in Fig.~\ref{f3}.
480: Notice that the second-order correction, which is the leading
481: time-nonlocal term in the perturbative expansion, becomes more
482: and more important when lowering the voltage.
483: In agreement with the conclusion drawn from the
484: first-order self-consistent calculation shown in Fig.~\ref{f2},
485: a clear enhancement
486: of the current by interactions can be observed
487: for a broad range of voltages. This enhancement is especially pronounced
488: for voltages slightly below the odd MAR
489: peaks located at $eV =2\Delta/(2n+1)$.
490: As indicated by our results for larger $\Gamma$, see Fig.~\ref{f4} for
491: the case $\Gamma/\Delta=2$, the interaction-induced enhancement of the
492: current is restricted to small $\Gamma/\Delta$.
493: For larger $\Gamma/\Delta$, the
494: current instead is weakly suppressed by interactions.
495: The same pronounced
496: MAR peak structure as in Fig.~\ref{f3} can be observed
497: in the pair order parameter
498: $\delta$ defined in Eq.~(\ref{delta}), whose absolute value is shown in Fig.~\ref{f5}.
499: The fact that the characteristic MAR features still appear at $eV=2\Delta/n$ for
500: the interacting case indicates that, at least for small $U$,
501: the number of Andreev reflections is not affected by the Coulomb interaction.
502: This is
503: in contrast to the inelastic MAR picture for phonon-mediated interactions,
504: as discussed in Ref.~\onlinecite{zazu}.
505: Moreover, our results indicate that
506: the Andreev quasiresonances\cite{ingerman,wendin} are not
507: shifted away from the gap subharmonics.
508:
509: \begin{figure}[ht!]
510: \scalebox{0.65}{\includegraphics{fig5.eps}}
511: \caption{\label{f5} Absolute value of $\delta$, see Eq.~(\ref{delta}),
512: in units of $\Delta$, for the parameters in Fig.~\ref{f3}.}
513: \end{figure}
514:
515: As follows from our numerical analysis,
516: the magnitude of the difference current is mainly determined by
517: the first harmonics of interaction-mediated pairing, Eq. (\ref{delta}),
518: and can be roughly approximated as
519: \begin{equation}
520: I(U)-I(0) \sim (e / \hbar) \, | \delta| \Delta / \Gamma ~.
521: \label{estimation}
522: \end{equation}
523: This can be seen, for instance, from comparison of the curves in
524: Figs.~\ref{f5} and \ref{f2} (for $\epsilon_0 = 0$).
525: A similar expression for the difference current, Eq. (\ref{estimation}),
526: is also obtained from a simple Fermi golden rule
527: calculation, by analyzing the dynamics of
528: Andreev (subgap) states at very low voltages, $e V \ll \Delta$.
529: The corresponding correction to the transition rate from Andreev
530: states into the continuum is determined by the imaginary part of the
531: Andreev state self-energy $\Sigma_{A}(\omega) = \delta
532: \sqrt{\Delta^2 - \omega^2}/\Gamma$,
533: while the total escape probability leading to the difference current
534: is given by an expression similar to Eq. (3) in Ref.~\onlinecite{alfredo}.
535:
536: The above results also suggest that as a function of the ratio $\Gamma/\Delta$,
537: there should be a crossover from enhancement to suppression of
538: the current around $\Gamma/\Delta\approx 1$.
539: This is what we get when fixing the voltage and changing $\Gamma/\Delta$.
540: In Fig.~\ref{f6}, we have chosen $V=0.6\Delta$, where the self-consistent
541: first-order approximation gives the main contribution, and plotted
542: $I(U)-I(0)$ either for fixed $U/\Gamma=0.5$ (solid line), or for
543: fixed $U/\Delta=0.25$ (dashed line).
544: The two curves both cross zero approximately at the same value, $\Gamma
545: \simeq \Delta$.
546:
547: \begin{figure}[ht!]
548: \scalebox{0.65}{\includegraphics{fig6.eps}}
549: \caption{\label{f6}
550: $I(U)-I(0)$ versus $\Gamma/\Delta$ at $V=0.6\Delta$,
551: $\epsilon_0=0$,
552: fixing $U/\Gamma=0.5$ (solid line) or $U/\Delta=0.25$ (dashed line),
553: from self-consistent first-order perturbation theory.}
554: \end{figure}
555:
556: In conclusion, we have presented a theory exploring
557: the effect of weak interactions
558: on superconducting transport through a quantum dot.
559: Employing second-order perturbation theory, valid for $U<\Gamma$,
560: we find an unexpected enhancement of the subgap current against
561: its noninteracting value when the hybridization $\Gamma$
562: is smaller than the BCS gap parameter $\Delta$.
563: The perturbation theory scheme pursued in this paper offers
564: controlled results in one corner of the parameter regime,
565: and in contrast to previous mean-field theories, we predict
566: significant interaction corrections even for weak interactions.
567:
568: \acknowledgments
569: We thank A. Levy Yeyati, T. Martin, and V. Shumeiko for discussions.
570: This work was supported by the EU HYSWITCH and INSTANS networks.
571:
572: \begin{appendix}
573: \section{Josephson current}
574:
575: In this appendix, we briefly show that the interaction-induced
576: current enhancement found in the $I-V$ curves for $\Gamma<\Delta$
577: discussed in Sec.~\ref{sec3} also appears in the equilibrium
578: Josephson current-phase relation (where $\phi$ is the phase difference
579: across the dot) for the same model.
580: We consider the corresponding first-order self-consistent theory
581: in equilibrium, for simplicity at $\epsilon_0=0$ only.
582: For small $U$, no polarization is present, $\gamma=0$, and
583: only the proximity-induced mean-field parameter $\delta=U\langle
584: d^\dagger_\uparrow d^\dagger_\downarrow\rangle$ gives an
585: effect. Assuming real-valued $\delta$,
586: the $T=0$ self-consistency equation reads
587: \begin{equation}\label{selfa}
588: \delta=U\int_{-\infty}^{\infty} \frac{d\omega}{2\pi}
589: \frac{\beta_\omega-\delta}{\alpha_\omega^2+
590: \left(\beta_\omega-\delta\right)^2}.
591: \end{equation}
592: where
593: \[
594: \alpha_\omega=\omega\left(1+\frac{\Gamma}{\sqrt{\omega^2+\Delta^2}}\right),
595: \quad \beta_\omega=\frac{\Gamma\Delta\cos(\phi/2)}{\sqrt{\omega^2+\Delta^2}}.
596: \]
597: The Josephson current is then given as
598: \begin{equation}\label{jos}
599: I =\frac{e\Gamma\Delta}{\pi\hbar}\sin(\phi/2)
600: \int_{-\infty}^{\infty} \frac{d\omega}{\sqrt{\omega^2+\Delta^2}}
601: \frac{\beta_\omega-\delta}{\alpha_\omega^2+\left(\beta_\omega-\delta\right)^2}
602: \end{equation}
603: The presence of $\delta$ in Eq.~(\ref{jos})
604: generally causes two counteracting effects: there is a
605: decrease of $I$ due to the numerator, but an increase due to the
606: appearance of $\delta$ in the denominator.
607: Which of these is more important can only be clarified by detailed
608: calculation.
609: We present analytical evaluations valid for
610: $|\delta|\ll \Gamma|\cos(\phi/2)|$, separately for the regimes $\Gamma /\Delta
611: \ll 1$ and $\Gamma/\Delta\gg 1$.
612:
613: Let us first discuss $\Gamma/\Delta\ll 1$, where
614: Eq.~(\ref{selfa}) yields
615: \begin{eqnarray}\label{azx}
616: \frac{\delta}{U} &\simeq& \frac{\Delta}{2|\Gamma\cos(\phi/2)-\delta|} \\
617: &\times& \nonumber
618: \left( \frac{2\Gamma\cos(\phi/2) \cos^{-1}\big|\frac{
619: \Gamma\cos(\phi/2)-\delta}{\Delta}\big|}{\pi \sqrt{\Delta^2-\left[
620: \Gamma\cos(\phi/2)-\delta\right]^2}} -\frac{ \delta}{\Delta}\right).
621: \end{eqnarray}
622: The interaction correction
623: to the Josephson current now follows from Eq.~(\ref{jos}),
624: \[
625: \delta I\simeq \frac{2e\delta}{\hbar}\tan(\phi/2)\left(
626: \frac{1}{2} f_1(\phi)\,{\rm sgn}\cos{(\phi/2)} -\frac{\delta}{U}\right).
627: \]
628: with
629: \[
630: f_1(\phi)=\frac{1+2 (\Gamma/\Delta) |\cos(\phi/2)|}
631: {[1+(\Gamma/\Delta)|\cos(\phi/2)|]^2}=1+O(\Gamma^2/\Delta^2).
632: \]
633: In the extreme limit $\Gamma/\Delta\to 0$,
634: Eq.~(\ref{azx}) yields $\delta=
635: (U/2){\rm sgn}\cos(\phi/2)$, and then $\delta I=0$.
636: Inspection of Eq.~(\ref{azx}) for finite $\Gamma/\Delta\ll 1$
637: shows however that $|\delta|<U/2$ for $\phi \ne \pi$.
638: As a result, the interaction current to the Josephson current
639: for $\Gamma\ll\Delta$ turns out to be positive, albeit
640: numerically small. Using Eq.~(\ref{azx}) we find
641: \[
642: \delta I\simeq \frac{eU}{\hbar \pi}\frac{\Gamma}
643: {\Delta}\sin(\phi/2)\,{\rm sgn}\cos{(\phi/2)}.
644: \]
645: We believe that this effect is related to
646: the enhancement of the current at finite bias $V$ in the regime
647: $\Gamma\ll \Delta$ discussed in Sec.~\ref{sec3}.
648:
649: On the other hand, for $\Gamma\gg \Delta$, Eq.~(\ref{selfa}) is solved by
650: $\delta\simeq U\Delta \ln(\Gamma/\Delta)\cos(\phi/2)/(\pi\Gamma)$,
651: and the lowest-order interaction correction to
652: the Josephson current is
653: \begin{eqnarray*}
654: \delta I &\simeq& \frac{2e\delta}{\hbar}\tan(\phi/2)
655: \left( \frac{\Delta}{2\Gamma} f_2(\phi)\,
656: {\rm sgn}\cos(\phi/2)
657: -\frac{\delta}{U}\right)
658: \end{eqnarray*}
659: with $f_2(\phi)=1+\cos^2(\phi/2) /4 $.
660: Due to the large $\ln(\Gamma/\Delta)$ factor appearing now in $\delta$,
661: the Josephson current will in general be decreased by interactions for
662: $\Gamma\gg \Delta$.
663:
664: We therefore find the same qualitative picture as for the nonequilibrium
665: current in Sec.~\ref{sec3}: The equilibrium Josephson current can also
666: be slightly increased by weak repulsive interactions for weak hybridization,
667: $\Gamma/\Delta\ll 1$, but is decreased in the opposite limit.
668: However, the increase of the Josephson current
669: for weak hybridization turns out to be much smaller than
670: for the corresponding nonequilibrium current.
671:
672: \end{appendix}
673:
674:
675: \begin{thebibliography}{99}
676:
677: \bibitem{vandam} J. van Dam, Yu.V. Nazarov, E.P.A.M. Bakkers, S. De Franceschi,
678: and L.P. Kouwenhoven, Nature {\bf 442}, 667 (2006).
679:
680: \bibitem{tube} A.Yu. Kasumov {\sl et al.}, Science {\bf 284}, 1508 (1999);
681: A. Morpurgo, J. Kong, C.M. Marcus, and H. Dai, Science {\bf 286}, 263 (1999);
682: M.R. Buitelaar, T. Nussbaumer, and C. Sch\"onenberger, Phys. Rev. Lett.
683: {\bf 89}, 256801 (2002);
684: J.-P. Cleuziou, W. Wernsdorfer, V. Bouchiat, T. Ondarcuhu, and M. Monthioux,
685: Nature Nanotechnology {\bf 1}, 53 (2006).
686:
687: \bibitem{buitelaar} M.R. Buitelaar, W. Belzig, T. Nussbaumer,
688: B. Babic, C. Bruder, and C. Sch\"onenberger, Phys. Rev. Lett. {\bf 91}, 057005 (2003).
689:
690: \bibitem{fulleren} A.Yu. Kasumov, K. Tsukagoshi, M. Kawamura, T. Kobayashi,
691: Y. Aoyagi, K. Senba, T. Kodama, H. Nishikawa, L. Ikemoto,
692: K. Kikuchi, V.T. Volkov,
693: Yu.A. Kasumov, R. Deblock, S. Gu{\'e}ron, and H. Bouchiat,
694: Phys. Rev. B {\bf 72}, 033414 (2005).
695:
696: \bibitem{jorgensen} H.I. Jorgensen, K. Grove-Rasmussen, T. Novotny,
697: K. Flensberg, and P.E. Lindelof, Phys. Rev. Lett. {\bf 96}, 207003 (2006).
698:
699: \bibitem{lieber} J. Xiang, A. Vidan, M. Tinkham,
700: R.M. Westervelt, and C.M. Lieber,
701: Nature Nanotechnology {\bf 1}, 208 (2006).
702:
703: \bibitem{jarillo} P. Jarillo-Herrero, J.A. van Dam, and L.P. Kouwenhoven,
704: Nature {\bf 439}, 953 (2006).
705:
706: \bibitem{christian} A. Eichler, M. Weiss, S. Oberholzer,
707: C. Sch\"onenberger, A. Levy Yeyati, J.C. Cuevas, and A. Martin-Rodero,
708: Phys. Rev. Lett. {\bf 99}, 126602 (2007).
709:
710: \bibitem{lindelof} T. Sand-Jespersen, J. Paaske, B.M. Andersen,
711: K. Grove-Rasmussen, H.I. Jorgensen, M. Aagesen, C.B. Sorensen,
712: P.E. Lindelof, K. Flensberg, and J. Nygard,
713: Phys. Rev. Lett. {\bf 99}, 126603 (2007).
714:
715: \bibitem{mar} E.N. Bratus, V.S. Shumeiko, and G. Wendin, Phys. Rev. Lett. {\bf 74}, 2110 (1995);
716: D.V. Averin and D. Bardas, {\sl ibid.} {\bf 75}, 1831 (1995);
717: J.C. Cuevas, A. Martin-Rodero, and A. Levy Yeyati,
718: Phys. Rev. B {\bf 54}, 7366 (1996).
719:
720:
721: \bibitem{ingerman}
722: \AA. Ingerman, G. Johansson,
723: V.S. Shumeiko, and G. Wendin, Phys. Rev. B {\bf 64}, 144504 (2001);
724: J. Lantz, V.S. Shumeiko, E.N. Bratus, and G. Wendin,
725: {\sl ibid.} {\bf 65}, 134523 (2002).
726:
727: \bibitem{cuevas} A. Levy Yeyati, J.C. Cuevas, A. Lopez-Davalos,
728: and A. Martin-Rodero, Phys. Rev. B {\bf 55}, R6137 (1997).
729:
730: \bibitem{wendin} G. Johansson, E.N. Bratus,
731: V.S. Shumeiko, and G. Wendin, Phys. Rev. B {\bf 60}, 1382 (1999).
732:
733: \bibitem{siano}
734: L. Glazman and K.A. Matveev,
735: JETP Lett. {\bf 49}, 659 (1989); A.V. Rozhkov and D.P. Arovas,
736: Phys. Rev. Lett. {\bf 82}, 2788 (1999); E. Vecino, A. Martin-Rodero, and
737: A. Levy Yeyati, Phys. Rev. B {\bf 68}, 035105 (2003);
738: F. Siano and R. Egger, Phys. Rev. Lett. {\bf 93}, 047002 (2004);
739: M.S. Choi, M. Lee, K. Kang, and
740: W. Belzig, Phys. Rev. B {\bf 70}, 020502(R) (2004);
741: C. Karrasch, A. Oguri, and V. Meden, {\sl ibid.} {\bf 77}, 024517 (2008).
742:
743: \bibitem{kang} K. Kang, Phys. Rev. B {\bf 57}, 11891 (1998);
744: Physica E {\bf 5}, 36 (1999); S.Y. Liu and X.L. Lei, Phys. Rev. B {\bf 70},
745: 205339 (2004).
746:
747: \bibitem{avishai} Y. Avishai, A. Golub, and A.D. Zaikin,
748: Phys. Rev. B {\bf 63}, 134515 (2001).
749:
750: \bibitem{slave} F.S. Bergeret, A. Levy Yeyati, and A. Martin-Rodero,
751: Phys. Rev. B {\bf 74}, 132505 (2006); Y. Avishai, A. Golub, and
752: A.D. Zaikin, {\sl ibid.} {\bf 67}, 041301(R) (2003).
753:
754: \bibitem{alfredo} A. Levy Yeyati, A. Martin-Rodero, and E. Vecino,
755: Phys. Rev. Lett. {\bf 91}, 266802 (2003).
756:
757: \bibitem{normal} T. Fujii and K. Ueda, J. Phys. Soc. Jpn.
758: {\bf 74}, 127 (2005).
759:
760: \bibitem{hamasaki} M. Hamasaki, Condensed Matter Physics {\bf 10}, 235 (2007).
761:
762: \bibitem{eg}
763: R. Egger and A.O. Gogolin, cond-mat/0712.0750 (to appear in Phys. Rev. B).
764:
765: \bibitem{zazu} A. Zazunov, R. Egger, C. Mora, and T. Martin,
766: Phys. Rev. B {\bf 73}, 214501 (2006).
767:
768: \bibitem{zarand}
769: B. Horv{\'a}th, B. Lazarovits, O. Sauret, and G. Zar{\'a}nd,
770: cond-mat/0712.0296.
771:
772: \bibitem{baym} G. Baym and L.P. Kadanoff, Phys. Rev. {\bf 124}, 287 (1961);
773: {\sl ibid.} {\bf 127}, 1391 (1962).
774:
775: \bibitem{hershfield} S. Hershfield, J.H. Davies, J. W. Wilkins, Phys. Rev. B
776: {\bf 46}, 7046 (1992).
777:
778: \bibitem{alf3}
779: A. Levy Yeyati, J.C. Cuevas, and A. Martin-Rodero, Phys. Rev. Lett. {\bf 95}, 056804 (2005).
780:
781:
782: \end{thebibliography}
783: \end{document}
784: