1: \documentclass[aps,prb,amsmath,twocolumn]{revtex4}
2: %\documentclass[aps,prb,twocolumn]{revtex4}
3: \usepackage{bm}
4: \usepackage{graphicx}
5: \begin{document}
6: \title{Coulomb Effects on Electron Transport
7: and Shot Noise in Hybrid Normal-Superconducting Metallic Structures}
8:
9: \author{Artem V. Galaktionov and Andrei D. Zaikin}
10: \affiliation{Forschungszentrum Karlsruhe, Institut f\"ur Nanotechnologie,
11: 76021, Karlsruhe, Germany\\
12: I.E. Tamm Department of Theoretical Physics, P.N.
13: Lebedev Physics Institute, 119991 Moscow, Russia}
14: %\date{}
15:
16: \begin{abstract}
17: We analyze the effect of electron-electron interactions on Andreev
18: current and shot noise in diffusive hybrid structures composed of
19: a normal metal attached to a superconductor via a weakly
20: transmitting interface. We demonstrate that at
21: voltages/temperatures below the Thouless energy of a normal metal
22: Coulomb interaction yields a reduction of both Andreev current and
23: its noise by a constant factor which essentially depends
24: on the system dimensionality. For quasi-1d structures this factor
25: is $\sim N_{\rm Ch}^{8/g}$, where $N_{\rm Ch}$ and $g$ are
26: respectively the number of conducting channels and dimensionless
27: conductance of a normal metal. At voltages above the Thouless
28: energy the interaction correction to Andreev current and shot
29: noise acquires an additional voltage dependence which turns out to be a
30: power-law in the quasi-1d limit.
31:
32: \end{abstract}
33: \maketitle
34:
35:
36: \section{Introduction}
37: It is well established that low temperature charge transport
38: through an interface between a normal metal and a superconductor
39: (NS) is dominated by Andreev reflection \cite{And}: An electron
40: with energy below the superconducting gap $\Delta$ enters the
41: superconductor from the normal metal, forms a Cooper pair together
42: with another electron, while a hole goes back into the normal metal.
43: In addition to Andreev
44: reflection electrons can be scattered at the interface potentials
45: and/or impurities. The combination of ``normal'' and Andreev
46: reflection mechanisms is essential for proper understanding of
47: transport phenomena in proximity structures composed of
48: superconducting and normal metals. This applies both to
49: equilibrium phenomena, such as dc Josephson effect in diffusive
50: SNS hybrids \cite{dG,ALO,ZZh,KL,Dubos} and Meissner effect in NS
51: systems \cite{Z,BBS}, and to non-equilibrium effects, such as
52: dissipative transport of subgap electrons across NS interfaces \cite{BTK}.
53:
54: For weakly transmitting NS interfaces the corresponding Andreev conductance
55: $G_A$ turns out to be rather small as a second order effect in the interface
56: transmission. At the same time interplay between disorder and interference
57: effects in the normal metal may strongly enhance Andreev conductance at
58: sufficiently low energies \cite{VZK,HN,Ben,Zai} leading to the
59: so-called zero-bias anomaly (ZBA) on the current-voltage characteristics.
60: In the case of diffusive metals Andreev conductance grows
61: as
62: \begin{equation}
63: G_A \propto 1/\sqrt{T},\;\;\;\; G_A \propto 1/\sqrt{V}
64: \label{ZBA1}
65: \end{equation}
66: with decreasing temperature and applied voltage $V$.
67:
68: What is the effect of Coulomb interaction on the Andreev conductance of
69: NS structures? In the limit of weakly transmitting NS interfaces
70: this question was addressed in Ref. \onlinecite{Zai} within a
71: simple capacitive model and the Coulomb blockade of Andreev
72: reflection was predicted at sufficiently low energies. Huck {\it
73: et al.} \cite{HHK} studied the problem by modeling the effect of
74: interactions by means of an effective electromagnetic environment
75: with an impedance $Z(\omega )$. Similarly to the the case of
76: normal tunnel junctions \cite{SZ,IN}, for the Ohmic environment
77: $Z(\omega )=R$ Coulomb interaction yields a power-law ZBA which,
78: being combined with Eq. (\ref{ZBA1}), yields \cite{HHK}
79: \begin{equation}
80: G_A \propto T^{8/g-1/2},\;\;\;\; G_A \propto V^{8/g-1/2}
81: \label{ZBA2}
82: \end{equation}
83: respectively for $eV <T$ and $eV>T$. Here and below we define the
84: dimensionless conductance as $g=R_q/R$, where $R_q=h/e^2$ is the quantum
85: resistance unit.
86:
87: While the approach \cite{HHK} takes into account Coulomb interaction
88: within a phenomenological scheme of the so-called $P(E)$-theory \cite{SZ,IN},
89: the question remains if and under which
90: conditions this scheme is sufficient to account for the effect of
91: electron-electron interactions in a disordered normal metal attached to a
92: superconductor. In this paper we develop a microscopic analysis of this
93: problem and evaluate the interaction correction to Andreev conductance
94: of diffusive NS structures. We demonstrate that both the form and
95: the magnitude of this interaction correction essentially depend on the
96: dimensionality of the sample as well as on the relation between temperature
97: and applied voltage to the relevant Thouless energy of the structure.
98: In particular, we predict that at sufficiently low energies the interaction
99: correction {\it saturates} to a constant value, thus
100: providing a temperature- and voltage-independent renormalization of
101: Andreev conductance. This saturation effect was recently
102: observed \cite{Takayanagi} in NS structures composed of a superconductor and
103: multi-walled carbon nanotubes (MWNT).
104:
105: The structure of the paper is as follows. In Sec. II we will
106: define the model for the NS system and outline the formalism of a
107: non-linear $\sigma$-model which will be used for our analysis.
108: Sec. III contains the derivation of the general expression for the
109: part of the action responsible for Andreev processes in our
110: system. With the aid of this expression in Sec. IV we will find
111: the Andreev current in the presence of electron-electron
112: interactions. Its analysis in a number of important limiting cases
113: is presented in Sec. V. In Sec. VI we will establish a general
114: relation between the current and shot noise in the presence of
115: electron-electron interactions. In Sec. VII we will briefly
116: discuss our key observations and compare our results with recent
117: experimental findings \cite{Takayanagi}.
118:
119:
120:
121: \section{Model and basic formalism}
122:
123: Throughout the paper we shall consider the hybrid structure depicted in
124: Fig. 1. A piece of a normal metal of a simple rectangular shape and of
125: length $L$ is attached to a bulk superconductor via an insulating
126: interface and to a big metallic reservoir N$'$ via a highly conducting
127: interface. The transversal dimensions of the normal metal are $L_y$ and
128: $L_z$, hence, the cross-section of the NS interface $\Gamma$ is just the
129: product $\Gamma =L_yL_z$. In what follows it will be convenient for us to
130: assume that the transmission of a tunnel barrier at this interface is
131: sufficiently low, so that its resistance $R_t$ exceeds that of a normal
132: metal $R\equiv L/\sigma \Gamma $,
133: \begin{equation}
134: R_t \gg R ,
135: \label{tN}
136: \end{equation}
137: where $\sigma =2e^2DN_0$ is the standard Drude conductivity with
138: $e$ being the electron charge, $D=v_F l/3$ standing for the diffusion
139: coefficient, and $N_0$ representing the density of states per spin
140: direction at the Fermi surface. In addition we
141: will assume that the normal metal is shorter than the dephasing
142: length $L_\varphi$,
143: \begin{equation}
144: L \ll L_\varphi .
145: \label{Lphi}
146: \end{equation}
147:
148:
149: \begin{figure}
150: \includegraphics[width=8.cm]{fig1.eps}
151: \caption{NS hybrid structure under consideration.}
152: \end{figure}
153:
154: Our theoretical analysis is based on the Keldysh representation
155: of a non-linear $\sigma$-model\cite{KA,FLS} which is formulated in
156: terms of the effective action
157: \begin{equation}
158: S=S_0+S_\Gamma ,
159: \end{equation}
160: where
161: \begin{eqnarray}
162: && S_0=\frac{i\pi N_0}{4}{\rm Tr}\,\left[ D\left( \bm
163: \partial \check g \right)^2 + 4i\left(i\check \tau_z
164: \partial_t+\check\Delta -e\check V\right)\check
165: g\right], \label{saq2}\\ &&
166: S_{\Gamma}=-\frac{i\pi}{4e^2R_t\Gamma}{\rm Tr}_\Gamma\,\check
167: g_-\check g_+. \label{saq3}
168: \end{eqnarray}
169: Here $\check g (\bm{R}, t_1,t_2)$ is the $4\times 4$ matrix that
170: depends on one coordinate and two time variables. It obeys
171: the normalization condition
172: \begin{equation}
173: \int dt'\check g(\bm{R}, t_1,t') \check g(\bm{R}, t',t_2)=\delta(t_1-t_2).
174: \label{norm}
175: \end{equation}
176: The product of matrices $\check g$ in Eq. (\ref{saq2}) should be
177: understood as a convolution, cf. Eq. (\ref{norm}). The trace in
178: Eq. (\ref{saq2}) is taken over the space and time variables and it
179: is accompanied by the summation over matrix indices.
180:
181: The term $S_{\Gamma}$ (\ref{saq3}) follows directly from the
182: Kupriyanov-Lukichev boundary conditions \cite{KL} which account for
183: tunneling of electrons between N- and S-metals. Spatial integration in
184: this term is restricted to the insulating interface, subscripts $\mp$
185: denote matrices taken at the left- and right-hand sides of the interface.
186: Other matrices in Eq. (\ref{saq2}) are defined as follows
187: \begin{eqnarray}
188: && \check\tau_z=\left( \begin{array}{cc}\hat\tau_z &0\\0& \hat\tau_z
189: \end{array}\right),\quad \hat\tau_z=\left( \begin{array}{cc}1
190: &0\\0& -1 \end{array}\right), \label{pmat}
191: \\ && \check\Delta(\bm{R},t)=\left( \begin{array}{cc}\hat\Delta &0\\0&
192: \hat\Delta\end{array}\right),\; \hat\Delta= \left( \begin{array}{cc}0
193: &\Delta(\bm{R},t)\\ -\Delta^*(\bm{R},t)&0
194: \end{array}\right).
195: \nonumber
196: \end{eqnarray}
197: The operator $\bm\partial $ acts as
198: \begin{eqnarray}
199: && \bm\partial \check g= \nabla\check g-i\frac{e}{c}
200: \left[\bm{A}\check\tau_z,\check g\right]\equiv \nonumber\\ &&
201: \nabla\check g(t,t') -i\frac{e}{c} \bm{A}(t)\check\tau_z \check
202: g(t,t')+\check g(t,t') i\frac{e}{c} \bm{A}(t')\check\tau_z.
203: \label{partial}
204: \end{eqnarray}
205: Here $\bm{A}(\bm{R},t)$ is the vector potential. In Eq.
206: (\ref{partial}) spatial arguments are suppressed for brevity.
207:
208: Finally, the $\check V$ term depends on the Hubbard-Stratonovich
209: fields $V_1$ and $V_2$ defined on the two branches of the Keldysh
210: contour. These fluctuating fields emerge after the standard
211: decoupling procedure in the term describing electron-electron
212: interactions. We define
213: \begin{equation}
214: \check V(\bm{R},t)=\left(\begin{array}{cc}V^+\hat 1& \frac{1}{2}
215: V^-\hat 1
216: \\ \frac{1}{2} V^-\hat 1 & V^+\hat 1
217: \end{array} \right),
218: \end{equation}
219: where $\hat 1$ is $2\times2$ unity matrix and $V^+=(V_1+V_2)/2,\,
220: V^-=V_1-V_2$ define respectively "classical" and "quantum"
221: fluctuating fields. If an external voltage $V$ is applied to the
222: system, it should simply be added to the classical component
223: $V^+$.
224:
225: It is necessary to supplement the action (\ref{saq2}),
226: (\ref{saq3}) describing fermionic degrees of freedom, by the
227: action for the bosonic fields $V^\pm$. The latter action
228: determines the correlators for these fields which will be
229: specified below.
230:
231: \section{Andreev action}
232:
233:
234: Let us first define the matrices $\check g_{\pm}$ in the absence of both
235: electron-electron interactions and tunneling between metals. For the
236: normal metal one has $\check g_+(\bm{R},t,t')=\check g_0(t,t')$, where
237: \begin{eqnarray} && \check g_0=\left(\begin{array}{cc} \hat g^R_0&
238: \hat g^K_0\\ 0& \hat g^A_0
239: \end{array}\right),\quad \hat
240: g^{R,A}_0(t,t')=\pm\hat\tau_z\delta(t-t'), \nonumber\\ && \hat
241: g^{K}_0(t,t')=-\frac{2i T}{\sinh[ \pi T(t-t')]}
242: \hat\tau_z.\label{equil}
243: \end{eqnarray}
244: The Fourier transform of the Keldysh matrix $\hat g^{K}_0$ reads
245: $\hat g^{K}_0(\epsilon)=2\tanh[\epsilon/2T] \hat\tau_z$. In what
246: follows we will consider the values of voltages and temperatures
247: much smaller than the superconducting gap $eV,T \ll \Delta$.
248: Hence, it will be sufficient for our purposes to specify the
249: matrix $\check g_-$ for a superconducting electrode only at
250: energies much smaller than $\Delta$. Under this condition we have
251: \begin{equation}
252: \check g_-(t,t')=\left(\begin{array}{cccc}0& -i&0 &0\\ i&0 &0 &0\\0&
253: 0&0&-i\\ 0& 0 &i &0
254: \end{array} \right)\delta(t-t').
255: \end{equation}
256:
257: Let us turn on electron-electron interactions. While in a
258: large superconducting electrode their influence on the matrix
259: $\check g_-$ can be safely neglected, the matrix $\check g_+$ in
260: the normal metal gets modified as \cite{SZ,KA}
261: \begin{equation}
262: \check g_+(\bm{R},t_1,t_2)=e^{-i\check K(\bm{R},t_1)} \check
263: g_0(t_1,t_2)e^{i\check K(\bm{R}, t_2)}. \label{gt}
264: \end{equation}
265: Here the matrix $\check K$ has the structure
266: \begin{equation}
267: \check K=\left(\begin{array}{cc} \varphi^+ \hat\tau_z&
268: \varphi^-\hat\tau_z/2\\ \varphi^-\hat\tau_z/2& \varphi^+ \hat\tau_z
269: \end{array}\right), \label{sf}
270: \end{equation}
271: and the phase fields $\varphi^{\pm}$ satisfy the equations
272: \begin{eqnarray}
273: && (Dq^2 + i\omega)\varphi^-(\bm{q},\omega)+eV^-(\bm{q}, \omega)=0,
274: \nonumber
275: \\&& (Dq^2 -i\omega) \varphi^+(\bm{q},\omega)-eV^+(\bm{q},
276: \omega)=\nonumber\\&& -Dq^2\varphi^-(\bm{q},\omega)\coth\frac{\omega}{2T}.
277: \label{phdef}
278: \end{eqnarray}
279: One can easily check that the part of the variation of $S_0$
280: (\ref{saq2}) linear both in $\check g$-variations and in the
281: fields $V^\pm$ vanishes under the transformation
282: (\ref{gt})-(\ref{phdef}). These equations are sufficient to
283: account for the effect of Coulomb interactions for the problem
284: in question.
285:
286: Now let us include tunneling between metals into consideration. In
287: order to eliminate the dependence of the matrix $\check g_+$ on
288: the fluctuating fields let us permute the factors $\exp(\pm
289: i\check K)$ in the term $S_\Gamma $ (\ref{saq3}) to act on the
290: matrix $\check g_-$. After this transformation we obtain
291: \begin{equation}
292: S_\Gamma=-\frac{i\pi}{4e^2R_t \Gamma} {\rm Tr}_\Gamma\,\check Q_-\check
293: g_+, \label{bt}
294: \end{equation}
295: where
296: \begin{eqnarray}
297: && \check Q_-(\bm{R},t, t')=\left(\begin{array}{cc} \hat A(\bm{R},t)& \hat
298: B(\bm{R},t) \\ \hat B(\bm{R},t)& \hat A(\bm{R},t)
299: \end{array}\right)\delta(t-t'),\nonumber\\
300: && \hat A=\left(\begin{array}{cc} 0& -ie^{2i\varphi^+}\cos \varphi^-
301: \\ie^{-2i\varphi^+}\cos\varphi^-& 0\end{array}\right),\label{mqs}\\
302: && \hat B=\left(\begin{array}{cc} 0& e^{2i\varphi^+}\sin \varphi^-
303: \\e^{-2i\varphi^+}\sin\varphi^-& 0\end{array}\right).\nonumber
304: \end{eqnarray}
305:
306: What remains is to evaluate the deviations of $\check g_+$ from
307: its normal state value (\ref{equil}) due to tunneling of Cooper
308: pairs from the superconducting electrode into the normal metal. It
309: is convenient to employ the following parameterization
310: \begin{equation} {\cal Q}=\check u\check g_+\check u, \quad
311: \check u=\check u^{-1}= \left(\begin{array}{cc} \delta(t-t')\hat 1&
312: -\frac{i T}{\sinh[ \pi T(t-t')]}\hat 1 \\ 0 & -\delta(t-t')\hat 1
313: \end{array}\right),\label{mdiag}
314: \end{equation}
315: which brings the matrix $\cal Q$ to the diagonal form in the absence of the
316: proximity effect. The latter effect can can then be interpreted in terms of
317: fluctuations of $\cal Q$ parameterized as
318: \begin{equation}
319: {\cal Q}=e^{-{\cal W}/2}\check\sigma_z\check\tau_z e^{{\cal
320: W}/2}=\check\sigma_z\check\tau_z \left(1+{\cal W}+{\cal
321: W}^2/2+\ldots\right)
322: \label{WWW}
323: \end{equation}
324: with
325: \begin{equation}
326: \check\sigma_z=\left( \begin{array}{cc}\hat 1 &0\\0&-\hat 1
327: \end{array}\right).
328: \end{equation}
329: The matrix ${\cal W}$ is in turn parameterized via the Pauli
330: matrices $\hat\tau_{x,y,z}$ as
331: \begin{equation}
332: {\cal W}=\left( \begin{array}{cc} w_1\hat\tau_x+w_2\hat\tau_y & w_0\hat 1+
333: w_3\hat\tau_z
334: \\ \overline w_0\hat 1+ \overline w_3\hat\tau_z &
335: \overline w_1\hat\tau_x+\overline w_2\hat\tau_y\end{array} \right).
336: \end{equation}
337: The functions $w_i$, and $\overline w_i$ can be expanded in
338: the eigenfunctions $\psi$ of the diffusion equation
339: \begin{equation}
340: w_i(x,y,z)=\sum_{\bm{q}} w_i(\bm{q})\psi_{\bm{q}}(x,y,z),
341: \end{equation}
342: obeying the boundary conditions $d\psi /dx|_{x=0}=0$, $\psi
343: |_{x=L}=0$ and impenetrable boundary conditions at $y= 0, L_y$ and
344: $z= 0, L_z$. We choose
345: \begin{eqnarray}
346: && \psi_{\bm{q}}(x,y,z)\propto \cos q_x x \cos q_y y \cos q_z z,
347: \nonumber\\ && q_x=\frac{\pi}{L}\left( l+\frac{1}{2}\right),\;
348: q_y=\frac{\pi}{L_y}m ,\; q_z=\frac{\pi}{L_z}n,\nonumber\\ && l,m,n=0,1,...
349: \label{dc2}
350: \end{eqnarray}
351: The normalization constant is determined by the condition $\int dx dy dz
352: \psi_{\bm{q}}^2(\bm{r})=1$.
353:
354: Using the effective action (\ref{saq2}), we can find all non-zero
355: averages for the products of the functions $w_i$, and $\overline
356: w_i$. Below we will only need the following averages \cite{FLS}
357: \begin{eqnarray}
358: && \left\langle w_i(\bm{q},\epsilon_1,\epsilon_2)w_i(-
359: \bm{q},\epsilon_3,\epsilon_4) \right\rangle=\nonumber\\&& -\frac{1}{\pi
360: N_0}\frac{(2\pi)^2\delta(\epsilon_1-\epsilon_4)\delta(\epsilon_2-
361: \epsilon_3)}{Dq^2-i(\epsilon_1+\epsilon_2)},\nonumber\\ && \left\langle
362: \overline w_i(\bm{q},\epsilon_1,\epsilon_2)\overline w_i(-
363: \bm{q},\epsilon_3,\epsilon_4) \right\rangle=\nonumber\\&& -\frac{1}{\pi
364: N_0}\frac{(2\pi)^2\delta(\epsilon_1-\epsilon_4)\delta(\epsilon_2-
365: \epsilon_3)}{Dq^2+i(\epsilon_1+\epsilon_2)},\label{avp}
366: \end{eqnarray}
367: which correspond to the Cooperons $(i=1,2)$.
368:
369: Now we are ready to evaluate the contribution of Andreev processes
370: to the effective action. In the weak tunneling limit it is
371: sufficient to expand $\exp (iS )$ up to the second order in
372: $S_\Gamma$ and then to average the resulting expression over the
373: fermionic degrees of freedom with subsequent re-exponentiation of
374: the result. In the limit $R_t\gg R$, the amplitude of the
375: anomalous Green function penetrating into the normal metal remains
376: much smaller than unity. Hence, in Eq. (\ref{WWW}) it is
377: sufficient to retain only the linear terms in ${\cal W}$. The
378: resulting Andreev contribution to the action then takes the form
379: \begin{eqnarray}
380: && \delta S_A=-\frac{i}{32}\left( \frac{\pi}{e^2 R_t
381: \Gamma}\right)^2\label{sot}\\ && \left\langle{\rm Tr}_{\Gamma_1}
382: {\cal Q}_- ( \Gamma_1)\check \sigma_z\check\tau_z{\cal
383: W}(\Gamma_1)\cdot {\rm Tr}_{\Gamma_2} {\cal Q}_- (
384: \Gamma_2)\check\sigma_z\check\tau_z{\cal
385: W}(\Gamma_2)\right\rangle.\nonumber
386: \end{eqnarray}
387: The space integration is performed here along the same interface
388: $\Gamma$, but independently for the terms containing $\Gamma_1$ and $\Gamma_2$.
389: The matrix ${\cal Q}_-$ is defined by the relation
390: ${\cal Q}_-=\check u \check Q_- \check u$. Employing Eqs. (\ref{mqs}),
391: (\ref{mdiag}) and performing the averaging with the aid of Eqs. (\ref{avp}),
392: we arrive at the final expression for $\delta S_A [\varphi^\pm ]$ which will
393: be extensively used in our subsequent analysis. We find
394:
395: \begin{widetext}
396: \begin{eqnarray}
397: && \delta S_A=\frac{i\pi R D}{2e^2R_t^2\Gamma L}\sum_{\bm{q}}\int\frac{d E
398: d\omega }{(2\pi)^2}\int_\Gamma d\bm{r}\int_\Gamma d\bm{r'}\int d t\int d
399: t'
400: e^{i\omega(t-t')}\psi_{\bm{q}}(\bm{r})\psi_{\bm{q}}(\bm{r'})\times\nonumber\\
401: && \bigg\{
402: \frac{Dq^2}{(Dq^2)^2+4E^2}\left[\tanh\frac{E+(\omega/2)}{2T}-
403: \tanh\frac{E-(\omega/2)}{2T}
404: \right]
405: \bigg[\coth\frac{\omega}{2T}e^{-2i(\varphi^+(\bm{r},t)-\varphi^+(\bm{r'},t'))}
406: \sin[\varphi^-(\bm{r},t)] \sin[\varphi^-(\bm{r'},t')] \nonumber\\ &&+
407: \sin\left[2\varphi^+(\bm{r},t)-2\varphi^+(\bm{r'},t')\right]
408: \cos[\varphi^-(\bm{r},t)] \sin[\varphi^-(\bm{r'},t')] \bigg]\nonumber
409: \\
410: && -\frac{2iE}{(Dq^2)^2+4E^2}\left[\tanh\frac{E+(\omega/2)}{2T}+
411: \tanh\frac{E-(\omega/2)}{2T} \right]
412: \sin\left[2\varphi^+(\bm{r},t)-2\varphi^+(\bm{r'},t')\right]
413: \cos[\varphi^-(\bm{r},t)]\sin[\varphi^-(\bm{r'},t')]\nonumber
414: \\&& -\frac{Dq^2}{(Dq^2)^2+4E^2}e^{2i(\varphi^+(\bm{r},t)-
415: \varphi^+(\bm{r'},t'))}
416: \cos[\varphi^-(\bm{r},t)-\varphi^-(\bm{r'},t')]\bigg\}. \label{finalaction}
417: \end{eqnarray}
418: \end{widetext}
419:
420: \section{Andreev current in the presence of interactions}
421:
422: Let is first evaluate Andreev current across the NS interface.
423: Substituting Eq. (\ref{finalaction}) into the formal expression for the current
424: \begin{equation}
425: I=e \left\langle \int_\Gamma d^2 r\frac{\delta}{\delta
426: \varphi^-(\bm{r})} \delta S_A \right\rangle
427: \label{curdef}
428: \end{equation}
429: and
430: performing averaging over the fluctuating fields $\varphi^\pm$ we
431: arrive at the general result for the Andreev current
432: \begin{widetext}
433: \begin{eqnarray}
434: && I=\frac{\pi R D}{2e R_t^2\Gamma L}\sum_{\bm{q}}\int\frac{dE
435: d\omega}{(2\pi)^2}\int_{\Gamma} d^2 r d^2 r'\psi_{\bm{q}}(\bm{r})
436: \psi_{\bm{q}}(\bm{r'}) \left[ \tanh\left(
437: \frac{E+eV-(\omega/2)}{2T}\right)- \tanh\left(
438: \frac{E-eV+(\omega/2)}{2T}\right)\right]\times \nonumber
439: \\&&
440: \frac{Dq^2}{\left(Dq^2\right)^2+4E^2} P(\omega,\bm{r},
441: \bm{r'})\frac{1-e^{-2eV/T}}{1-e^{(\omega-2eV)/T}}. \label{main}
442: \end{eqnarray}
443: \end{widetext}
444: Here the function
445: \begin{eqnarray}
446: && P(\omega,\bm{r},\bm{r'})=\int dt e^{J(t,\bm{r},\bm{r'}) +i\omega
447: t},\nonumber
448: \\ && e^{J(t,\bm{r},\bm{r'})}= \left\langle e^{2i\varphi_{2}(\bm{r},t)-
449: 2i\varphi_{1}(\bm{r'},0)} \right\rangle \label{J}
450: \end{eqnarray}
451: accounts for the effect of interactions described by the fluctuating phase
452: fields $\varphi_{1,2} =\varphi^+\pm (\varphi^-/2)$ defined on the two
453: branches of the Keldysh contour. Averaging in Eq. (\ref{J}) is
454: performed in the absence of the external voltage. Note that the form of
455: this function is reminiscent to that used in the standard $P(E)$-theory
456: \cite{SZ,IN} and the result for the Andreev current (\ref{main}) looks
457: somewhat similar to the expression derived within the framework of the
458: latter theory in Ref. \onlinecite{HHK}. In contrast to the latter work,
459: however, our approach and resulting Eqs. (\ref{main}), (\ref{J}) fully
460: account (i) for electron
461: coherence across the whole NS structure and (ii) for the spatial
462: dependence, introduced both by the functions $\psi_{\bm{q}}(\bm{r})$,
463: describing fermionic fluctuations, and by the function
464: $J(t,\bm{r},\bm{r'})$, describing bosonic (phase) fluctuations.
465:
466: These fluctuations depend on the dimensionality of the normal
467: sample. For our purposes it will be sufficient to employ the
468: standard random phase approximation (RPA)
469: and to express the correlators of the fluctuating
470: Hubbard-Stratonovich fields via an effective $d$-dimensional
471: dielectric function $\varepsilon(\omega,k)$ of our system. One
472: finds \cite{GZ}
473: \begin{eqnarray}
474: && \left\langle e V^+(t,\bm{r}) e
475: V^+(0,0)\right\rangle=\nonumber\\ && -\int \frac{d\omega d^d
476: k}{(2\pi)^{d+1}}{\rm Im}\left( V^*(k,\omega) \right)\coth
477: \left(\frac{\omega}{2T}\right) e^{-i\omega t+i\bm{k
478: r}},\nonumber\\ && \left\langle e V^+(t,\bm{r}) e
479: V^-(0,0)\right\rangle=\nonumber\\ && i\int \frac{d\omega d^d
480: k}{(2\pi)^{d+1}}V^*(k,\omega) e^{-i\omega t+i\bm{k r}}\nonumber,\\
481: && \left\langle e V^-(t,\bm{r}) e
482: V^-(0,0)\right\rangle=0.\label{aver}
483: \end{eqnarray}
484: Here we have defined
485: \begin{equation}
486: V^*(k,\omega)=\frac{1}{V_0^{-1}(k)+\frac{k^2}{4\pi
487: e^2}\left[\varepsilon(k,\omega)-1\right]},\label{vstar}
488: \end{equation}
489: where the term $V_0(k)$ stands for the unscreened Coulomb
490: interaction:
491: \begin{eqnarray}
492: V_0(k)=\frac{4\pi e^2}{k^2},\quad d=3,\nonumber
493: \\ V_0(k)=\frac{2\pi e^2}{k},\quad d=2,\nonumber
494: \\V_0(k)=e^2\ln\left(1+\frac{1}{\Gamma k^2}\right),\quad d=1.
495: \label{usc}
496: \end{eqnarray}
497: In the latter equation we have assumed $L_y\sim L_z \sim
498: \sqrt{\Gamma}$. In the three-dimensional case we obviously have
499: $V^*(k,\omega)=V_0(k)/\varepsilon(k,\omega)$.
500:
501: Note, that in the case of lower dimensions $d=1,2$ the term
502: $V_0^{-1}(k)$ also accounts for the electric field energy outside
503: the conductors. The corresponding interaction can be screened,
504: provided there are metallic gates near the film/wire. For
505: instance, in the case $d=2$ one has
506: \begin{equation}
507: V_0^{scr}(k)=\frac{2\pi e^2}{k}\left(1-e^{-bk}\right),
508: \end{equation}
509: where $b$ is the distance between the film and the gate electrode.
510:
511: In order to evaluate the average for the phase fields (\ref{J})
512: one can combine Eqs. (\ref{phdef}), (\ref{aver}) with the standard
513: Drude form of the dielectric function for the normal metal
514: \begin{equation}
515: \varepsilon(k,\omega)=1+\frac{4\pi \sigma}{-i\omega+Dk^2}. \label{rpa}
516: \end{equation}
517:
518: The resulting expressions for the function (\ref{J}) obtained in this way
519: remain applicable at frequencies above the corresponding Thouless energy
520: of the sample but should fail at lower frequencies, in which case
521: discrete wave vectors $\sim 1/L$ become significant. In order to
522: cure this complication it is
523: useful to depart from the representation of plane waves in Eq.
524: (\ref{aver}) and to employ a different set of the (volume-normalized) basis
525: functions $\Omega_{m}(\bm{r})$ obeying the equation
526: \begin{equation}
527: \nabla^2 \Omega_{m}(\bm{r})+ \lambda_m^2\Omega_{m}(\bm{r})=0
528: \end{equation}
529: with the corresponding eigenvalues $\lambda_m$. At the
530: impenetrable interfaces the functions $\Omega_{m}(\bm{r})$ satisfy
531: the boundary condition $\nabla_n \Omega_{m}(\bm{r})=0$. One should
532: also require the functions
533: $\Omega_{m}(\bm{r})$ to vanish at the interface with a large
534: metallic reservoir $x=L$.
535:
536: The corresponding generalization of the dielectric function
537: (\ref{rpa}) can be achieved by evaluating the response of a finite sample,
538: characterized by the current density $\bm{j}(\bm{r})=\sigma
539: \bm{E}(\bm{r})-D\nabla \rho(\bm{r})$ with $\bm{E}$ and $\rho$ standing
540: respectively for the
541: electric field and for the charge density. One gets \cite{SG}
542: \begin{eqnarray}
543: && \varepsilon_{\alpha
544: \beta}(\bm{r},\bm{r'},\omega)=\delta_{\alpha\beta}\delta(\bm{r}-\bm{r'})
545: +\\ && \frac{4\pi i}{\omega}\sigma_D
546: \left(\delta_{\alpha\beta}\delta(\bm{r}-\bm{r'})-\nabla_{\alpha}\nabla'_{\beta}
547: d(\bm{r},\bm{r'},\omega)\right), \nonumber
548: \end{eqnarray}
549: where
550: \begin{equation}
551: d(\bm{r},\bm{r'},\omega)=\sum_{m}\frac{\Omega_{m}(\bm{r})
552: \Omega_{m}(\bm{r'})}{\lambda_m^2-\frac{i\omega}{D}}.
553: \end{equation}
554:
555: Implementing these modifications also into Eqs. (\ref{aver}) and employing
556: Eqs. (\ref{phdef}) we arrive at the following general result
557: \begin{eqnarray}
558: && J(t,\bm{r},\bm{r'})=4\sum_{m}\int\frac{d\omega }{(2\pi)}\,{\rm
559: Im}\left[ \frac{V^*(\lambda_m,\omega)}{
560: (D\lambda_m^2-i\omega)^2}\right]\times\nonumber\\&&
561: \bigg[\left(\coth\frac{\omega}{2T} \cos \omega t -i \sin \omega
562: t\right)\Omega_m(\bm{r})\Omega_m(\bm{r'})- \label{jdf2}\\ &&
563: \frac{1}{2}\coth\frac{\omega}{2T} \left( \Omega_m(\bm{r})\Omega_m(\bm{r})+
564: \Omega_m(\bm{r'})\Omega_m(\bm{r'}) \right)\bigg]. \nonumber
565: \end{eqnarray}
566: The latter formula results in
567: $J(t-(i/T),\bm{r},\bm{r'})=J(-t,\bm{r},\bm{r'})$ leading to the relation
568: $P(-E,\bm{r},\bm{r'})=e^{-E/T}P(\omega,\bm{r},\bm{r'})$ as in the standard
569: $P(E)$ theory. Accordingly, at $T \to 0$ Eq. (\ref{main}) reduces to
570: \begin{eqnarray}
571: && \frac{dI}{dV}=\frac{D R}{2\pi R_t^2\Gamma L}
572: \sum_{\bm{q}}\int\limits_{0^-}^{2|eV|}dE \int_{\Gamma} d^2 r d^2
573: r'\psi_{\bm{q}}(\bm{r}) \psi_{\bm{q}}(\bm{r'})\nonumber \\ && P(E,\bm{r},
574: \bm{r'}) \frac{Dq^2}{\left(Dq^2\right)^2+(2|eV|-E)^2}. \label{ea}
575: \end{eqnarray}
576: At frequencies exceeding the sample Thouless energy it is possible to
577: perform averaging of Eq. (\ref{jdf2}) over fluctuations depending on
578: $\bm{r}+\bm{r'}$. Then one gets
579: \begin{eqnarray}
580: && J(t,|\bm{r}|)=4\int\frac{d\omega d^n k}{(2\pi)^{n+1}}\,{\rm Im}\left[
581: \frac{V^*(k,\omega))}{ (Dk^2-i\omega)^2}\right]\times\nonumber\\&&
582: \left[\coth\frac{\omega}{2T} (\cos \omega t\cos\bm{k}\bm{r}-1 )-i \sin
583: \omega t\cos\bm{k}\bm{r}\right]. \label{jdf}
584: \end{eqnarray}
585:
586:
587: \section{Results}
588:
589: \subsection{Non-interacting limit}
590:
591: Let us first reproduce the results for Andreev current in the
592: non-interacting limit. In this case we have $P(\omega,r)=2\pi
593: \delta(\omega)$ and the standard expressions for the $I-V$ curve
594: are readily recovered. In the limit $eV \gg T$ one finds
595: \begin{eqnarray}
596: && \frac{dI}{dV}=\frac{R}{R_t^2}, \quad {\rm if}\;
597: |eV|\ll \epsilon_{\rm Th}; \nonumber
598: \\ && \frac{dI}{dV}=\frac{R}{2R_t^2}\sqrt{\frac{\epsilon_{\rm Th}}{|eV|}}
599: , \quad {\rm if}\; |eV| \gg
600: \epsilon_{\rm Th}.
601: \label{ZBA}
602: \end{eqnarray}
603: Here and below $\epsilon_{\rm Th}=D/L^2$ is the Thouless energy.
604:
605: In the linear regime $eV \ll T$ for the Andreev conductance $G_A\equiv dI/dV$
606: we obtain
607: \begin{eqnarray}
608: && G_A^{(0)}=\frac{R}{R_t^2}, \quad {\rm if}\;
609: T\ll \epsilon_{\rm Th},
610: \label{ZBA0}
611: \\ && G_A^{(0)}=\frac{(2^{3/2}-1)\zeta(3/2)}{4\sqrt{\pi}}\frac{R}{R_t^2}
612: \sqrt{\frac{\epsilon_{\rm Th}}{T}}
613: , \quad {\rm if}\; T \gg \epsilon_{\rm Th},\nonumber
614: \end{eqnarray}
615: where $(2^{3/2}-1)\zeta(3/2)/4\sqrt{\pi}\approx 0.67$.
616:
617: Now let us turn to the effect of electron-electron interactions.
618: Both the value and the form of the interaction correction to the
619: Andreev conductance essentially depend on the effective
620: dimensionality of fluctuations of the phase fields. Depending on
621: the voltage and temperature the leading contribution to the
622: interaction correction is given either by the modes with $l, m,
623: n\gg 1$ (cf. Eq. (\ref{dc2})) or by the modes with $m$ (or $n$, or
624: both) being equal to zero. Correspondingly, below we will
625: distinguish between effectively 3d-, quasi-2d- and quasi-1d-type
626: of behavior of the interaction correction.
627:
628: \subsection{Effect of interactions: Bulk limit}
629:
630: We start from the case of 3d fluctuations. For convenience we shall assume
631: $L\gtrsim L_y \gtrsim L_z$ and consider the limit $eV \gg T$.
632: At sufficiently
633: high voltages $eV\gg D/L_z^2$ from Eq. (\ref{jdf}) (with $d=3$) we obtain
634: \begin{eqnarray}
635: && J(t,r)=\frac{e^2}{\pi^2 \sigma r}\int\limits_0^\infty \frac{\cos
636: (2Dt\omega/r^2)}{\omega}\left(
637: 1-e^{-\sqrt{\omega}}\cos\sqrt{\omega}\right) d\omega\nonumber\\ &&
638: -\frac{ie^2}{2\pi\sigma r} {\rm erf}\left(\frac{r}{2\sqrt{D|t|}}
639: \right){\rm sgn}\, t -\kappa ,
640: \end{eqnarray}
641: where $ {\rm erf}(x)=2\int_0^x \exp(-t^2) d t/\sqrt{\pi}$ and
642: \begin{equation}
643: \kappa=4\int\frac{d\omega d^3 k}{(2\pi)^{4}}\,{\rm Im}\left[
644: \frac{V^*(k,\omega)}{ (Dk^2-i\omega)^2}\right]\rm{sign}\,\omega.
645: \end{equation}
646: Making use of the condition $P(\omega, r)=0$ for $\omega <0$, we have for
647: $\omega\ge 0$
648: \begin{eqnarray}
649: && P(\omega, r)=e^{-\kappa} 2\pi\delta(\omega) +\nonumber\\ &&\frac{2e^2
650: e^{-\kappa} }{\pi\sigma r\omega}\left[ 1-e^{-r\sqrt{\omega/2D}}\cos \left(
651: r\sqrt{\omega/2D}\right)\right]. \label{per}
652: \end{eqnarray}
653: Thus we obtain
654: \begin{equation}
655: \frac{dI}{dV}=\frac{Re^{-\kappa }}{2 R_t^2}\left(\sqrt{\frac{\epsilon_{\rm
656: Th}}{|eV|}} +\frac{3.64 \Gamma}{g L^2} \right). \label{as3}
657: \end{equation}
658:
659: One observes that the relative ``weight'' of the second term in the
660: brackets of (\ref{as3}) grows with voltage as
661: $(1/g)\sqrt{eV/\epsilon_{\rm Th}}$. This form is consistent with
662: the standard square-root energy dependence of the interaction
663: correction to the density of states. In the second term in Eq.
664: (\ref{as3}) this dependence is exactly compensated by ZBA
665: (\ref{ZBA}) resulting in the voltage-independent contribution
666: $\sim R/gR_t^2$. This compensation will not occur in other limits
667: to be considered below.
668:
669: The value of $\kappa$ in the bulk limit can be estimated as
670: \begin{equation}
671: \kappa \equiv \kappa_3 \sim \frac{e^2}{\sigma
672: \sqrt{D}}\int\limits_0^{1/\tau}\frac{d\omega}{\sqrt{\omega}}\sim
673: \frac{1}{(p_Fl)^2}.\label{33}
674: \end{equation}
675: Here $\tau=l/v_F$ is the elastic scattering time.
676:
677: While the result (\ref{as3}) holds at high voltages, at lower
678: values $eV \lesssim \epsilon_{\rm Th}$ -- as it was already
679: discussed above -- the Andreev conductance saturates to a constant
680: value
681: \begin{equation}
682: \frac{dI}{dV}=\frac{Re^{-\kappa}}{R_t^2}\label{satt}
683: \end{equation}
684: with the voltage-dependent correction to this formula of order
685: $(eV)^2/g\epsilon_{\rm Th}^2$.
686:
687: Note that in the bulk limit the net effect of
688: electron-electron interactions is parametrically small and negative in the
689: admissible range of voltages $eV < 1/\tau$. Indeed, having in mind that
690: our calculation is performed in the limit $p_Fl \gg 1$ and making use of
691: the estimate (\ref{33}), in Eq. (\ref{as3}) one can expand $\exp (-\kappa
692: )$ to the first order in $\kappa$ and observe that the corresponding
693: negative correction exceeds the positive one by a factor $\sim
694: 1/\sqrt{|eV|\tau}$.
695:
696: Let us also point out that in order to obtain the Andreev
697: conductance in the voltage range $eV \ll T$ it is sufficient to
698: simply substitute $T$ instead of $eV$. As in the non-interacting
699: limit (cf. Eqs. (\ref{ZBA}) and (\ref{ZBA0})), this procedure will
700: yield the correct form of the interaction correction up to an
701: unimportant numerical prefactor of order one.
702:
703:
704:
705:
706:
707:
708:
709: \subsection{Effect of interactions: Quasi-2d structures}
710:
711:
712: Let us now turn to low-dimensional fluctuations. We first consider
713: systems which can be treated as quasi-2d ones, i.e. we assume that
714: both the length and the width of the normal metal strongly exceed
715: its thickness $L\sim L_y \gg L_z$. We shall address two different
716: physical situations corresponding to the absence and to the
717: presence of massive metallic gate electrodes in the vicinity of
718: the sample.
719:
720: In the absence of the gate electrodes we can use the 2d expression
721: for $V_0(k)$ (\ref{usc}), combined with Eq. (\ref{vstar}). We
722: first evaluate the constant $\kappa$ which now consists of two
723: contributions,
724: \begin{equation}
725: \kappa = \kappa_3 +\kappa_2,
726: \end{equation}
727: where the first one is again determined by high frequencies ($\omega
728: \gtrsim D/L_z^2$) and small scales (cf. Eq. (\ref{33})), while the second
729: one emerges from integrating over the parameter region $ 2\pi\sigma L_z
730: q\gg \omega$ and $Dq^2\ll \omega$ in Eq. (\ref{jdf2}) for smaller
731: frequencies $\omega \lesssim D/L_z^2$
732: \begin{equation}
733: \kappa_2 =\frac{8 L_y}{\pi g L}\ln\frac{L_y}{L_z}\ln\frac{\omega_0
734: \Gamma}{D}, \quad \omega_0=\frac{(2\pi \sigma L_z)^2}{D}.
735: \end{equation}
736: It is obvious from the above expressions that the contribution
737: $\kappa_3$ exceeds $\kappa_2$ for thicker films, i.e. as long as
738: $L_z$ remains bigger than the mean free path $l$. In the opposite
739: case, however, the contribution of 3d fluctuations $\kappa_3$ is
740: negligible, and in the expression for $\kappa_2$ one should
741: substitute $1/\tau$ instead of $D/L_z^2$ in the arguments of the
742: logarithms. Then one finds
743: \begin{equation}
744: \kappa \simeq \kappa_2 =\frac{2 L_y}{\pi g L}
745: \ln\frac{L_y^2}{D\tau}\ln\frac{\omega_0^2 \tau
746: L_y^2}{D}.
747: \label{22}
748: \end{equation}
749: We observe that the above expressions for the renormalization
750: constant $\kappa$ contain double logarithmic factors which are
751: rather typical for 2d structures. Similar double logarithmic
752: dependencies emerge in voltage- and temperature-dependent
753: contributions to be analyzed below.
754:
755: Let us again stick to the limit $eV \gg T$. At voltages
756: $D/L_y^2\ll eV\ll D/L_z^2$ we obtain
757: \begin{equation}
758: \frac{dI}{dV}=\frac{R}{2R_t^2}\sqrt{\frac{\epsilon_{\rm
759: Th}}{|eV|}}e^{-\kappa}
760: \left[1+ \frac{2 L_y}{\pi g L}\ln
761: \frac{|eV| L_y^2}{D} \ln\frac{\omega_0^2L_y^2}{|eV|D}\right].\label{2dng}
762: \end{equation}
763: Note that in contrast to the expression for $\kappa$ (\ref{22})
764: the voltage-dependent part of the interaction correction
765: (\ref{2dng}) does not involve the upper frequency cutoff $\sim
766: 1/\tau$. At lower voltages $eV\lesssim \epsilon_{\rm Th}$ the
767: differential conductance again saturates to the value (\ref{satt})
768: with $\kappa$ defined in Eq. (\ref{22}). The above results do not
769: coincide with ones obtained in Ref. \onlinecite{FLS} for the same
770: physical situation.
771:
772: It is worth pointing out that even though the combination $L_y/gL$ in the
773: above expressions always remains small, in some cases this smallness can be
774: compensated by large logarithms, and the formal expansion in $1/g$ can become
775: insufficient. In this situation the $I-V$ curve can be evaluated with the aid
776: of the following expression
777: \begin{equation}
778: P(E)=2\pi\delta(E)e^{-\Phi(0)} +2\pi \frac{d}{dE}e^{-\Phi(E)},
779: \end{equation}
780: where
781: \begin{equation}
782: \Phi(E)=\frac{e^2}{\pi^2\sigma L_z
783: }\ln\frac{1}{\sqrt{\epsilon_{\rm Th}^2+E^2}\tau}
784: \ln\frac{\omega_0^2 \tau}{\sqrt{\epsilon_{\rm Th}^2+E^2}}.
785: \end{equation}
786: This function also enters into the expression for the density of states
787: of normal 2d films \cite{KA}
788: \begin{equation}
789: N(\epsilon)=N_0 e^{-\Phi(E)/4}.
790: \end{equation}
791: We also note that in the case of thicker films $L_z \gg l$ our results
792: are consistent with the expression for the density of states \cite{Grab}.
793:
794: In the presence of massive gate electrodes close to our sample
795: the fluctuating electromagnetic fields do not extend outside its volume, and
796: the term $V^*$ in Eq. (\ref{jdf2}) gets modified. In this limit we find
797: \begin{equation}
798: \frac{dI}{dV}=\frac{R}{2R_t^2}\sqrt{\frac{\epsilon_{\rm
799: Th}}{|eV|}}e^{-\kappa}
800: \left[1+ \frac{ 2 L_y}{\pi g L}\ln^2
801: \frac{|eV| L_y^2}{D}\right], \label{f2}
802: \end{equation}
803: where
804: \begin{equation}
805: \kappa= \frac{ 2 L_y}{\pi g L}\ln^2 \frac{L_y^2}{\tau D}
806: \end{equation}
807:
808: If one considers the limit $T \gg eV$, it suffices to substitute $0.56 T$
809: instead of $eV$ in Eqs. (\ref{2dng}), (\ref{f2}) , and one will arrive at
810: the expression for the linear Andreev conductance $G_A(T)$ valid with the
811: logarithmic accuracy.
812:
813: \subsection{Effect of interactions: Quasi-1d structures}
814:
815:
816: Let us finally turn to quasi-1d structures $L \gg L_y\sim L_z$ and
817: examine the effect of electron-electron interactions on the
818: Andreev conductance in the absence of screening gates. Similarly
819: to the 2d case, there also exists the parameter region $Dq^2\ll
820: \omega$ and $\sigma \Gamma q^2\log(1/L_z^2q^2)\gg \omega$ , which
821: gives the leading contribution to $J(t)$. We obtain
822: \begin{eqnarray}
823: && J(t)=16\int\limits_{\epsilon_{\rm Th}}^\infty
824: \int\limits_{q_{\rm min}}^\infty \frac{dq}{2\pi}\frac{e^2}{\sigma
825: \Gamma \omega q^2}\times\nonumber\\ &&
826: \left\{\coth\frac{\omega}{2T}\left[ \cos(\omega t)-1 \right]
827: -i\sin(\omega t)\right\}.
828: \end{eqnarray}
829: Since $1/q$ cannot exceed $L$, one should choose
830: \begin{equation}
831: q_{\rm min}\sim L^{-1}+\sqrt{\frac{\omega}{\sigma \Gamma\ln[\sigma
832: \Gamma/\omega L_z^2]}}.
833: \end{equation}
834: Performing the $q$-integration one finds
835: \begin{eqnarray}
836: && J(t)=\frac{8}{g}\int\limits_{\epsilon_{\rm
837: Th}}^\infty\frac{d\omega}{\omega\left(1+ \sqrt{\omega
838: RC}\right)}\nonumber \\ && \left\{\coth\frac{\omega}{2T}\left[
839: \cos(\omega t)-1 \right] -i\sin(\omega t)\right\}, \label{mpe}
840: \end{eqnarray}
841: where
842: \begin{equation}
843: C=\frac{L}{\ln[L^2/\Gamma ]}
844: \label{Cap}
845: \end{equation}
846: is the capacitance of a long cylinder. The expression (\ref{mpe})
847: is rather similar to that obtained within the
848: $P(E)$-theory\cite{HHK}. There are at least two differences,
849: however. One of them is that here the combination $1+\sqrt{\omega
850: RC}$ enters in the denominator of Eq. (\ref{mpe}), while in the
851: model \cite{HHK} one has $1+\omega^2R^2C_J^2$, where $C_J$ is the
852: capacitance of a tunnel barrier at the NS interface. This
853: capacitance is neglected in our analysis and can be trivially
854: restored, if needed. Another -- more important -- difference is
855: that the frequency integral in Eq. (\ref{mpe}) cannot be extended
856: below the Thouless energy. This low frequency cut-off results in
857: the saturation of the interaction correction at small voltages and
858: temperatures. This effect is not contained in the standard
859: $P(E)$-theory which does not keep track of quantum coherence
860: of electrons entering an effective environment and which also does
861: not account for the sample geometry.
862:
863: \begin{figure}
864: \includegraphics{fig2.eps}
865: \caption{The zero temperature differential Andreev conductance
866: $G_A=dI/dV$ of quasi-1d diffusive NS systems. We have set
867: $RC\epsilon_{\rm Th}=0.01$. The dimensionless parameter $8/g$
868: equals to 0, 0.1, 0.5, 1.1 (from top to bottom). }
869: \end{figure}
870:
871: In the limit $T \to 0$ the asymptotic expression for the function $J(t)$ at
872: $t\gg RC$ reads
873: \begin{equation}
874: J(t)=\frac{8}{g}\left[\ln\left(\epsilon_{\rm Th} RC\right) -{\rm
875: ci}(|t|\epsilon_{\rm Th}) +i{\rm sign}\,t\,{\rm
876: si}(|t|\epsilon_{\rm Th})\right],
877: \end{equation}
878: where $ {\rm ci}\,t=-\int_t^\infty dz \cos z/z,\; {\rm
879: si}\,t=-\int_t^\infty dz \sin z/z$. In particular, for $t\ll
880: 1/\epsilon_{\rm Th}$ we get
881: \begin{equation}
882: J(t)=-\frac{8 }{g}\left[\ln\left(\frac{e^\gamma t}{RC}
883: \right)+i\frac{\pi}{2} {\rm sign}\,t \right],
884: \end{equation}
885: where $\gamma \simeq 0.577...$ is the Euler constant. At longer
886: times $t\gg 1/\epsilon_{\rm Th}$ the function $J(t)$ approaches
887: the constant, $J(t)=-\kappa$, where
888: \begin{equation}
889: \kappa \simeq \kappa_1 =\frac{8}{g}\ln\frac{1}{RC\epsilon_{\rm
890: Th}}= \frac{8}{g}\ln\frac{e^2p_F^2 \Gamma \ln (L^2/\Gamma)}{\pi^2
891: v_F} . \label{11}
892: \end{equation}
893:
894: The $I-V$ curve can now be easily evaluated. It is clear from Eq.
895: (\ref{mpe}) that in the limit of high voltages $eV \gg 1/RC$ the response
896: of an effective environment (normal metal) will be that of an
897: $RC$-transmission line. In this limit the interaction correction to the Andreev
898: conductance will behave analogously to that studied for single electron
899: tunneling within the $P(E)$-theory \cite{IN}. At lower voltages
900: $\epsilon_{\rm Th}\ll eV\ll 1/RC$ we find
901: \begin{eqnarray}
902: && \frac{dI}{dV}=\frac{R}{2R_t^2}\sqrt{\frac{\epsilon_{\rm Th}}{|eV|}}
903: \Big[ e^{-\kappa}+
904: \\ && \frac{\sqrt{\pi}e^{-8\gamma/g}}{\Gamma(\frac{8}{g}+\frac{1}{2})}\left(
905: (2|eV|RC)^{8/g} -(\epsilon_{\rm Th} RC)^{8/g}
906: \right)\Big].\nonumber
907: \end{eqnarray}
908: In the limit of large conductances $g \gg 1$ this expression
909: can also be rewritten in a simpler form
910: \begin{equation}
911: \frac{dI}{dV}=\frac{R}{2R_t^2}\sqrt{\frac{\epsilon_{\rm
912: Th}}{|eV|}}e^{-\kappa}\bigg[1 +\frac{8}{g} \ln
913: \frac{|eV|}{\epsilon_{\rm Th}}\bigg].\label{f1}
914: \end{equation}
915: At smaller voltages $eV < \epsilon_{\rm Th}$ we again observe the
916: saturation of the differential conductance to the value
917: (\ref{satt}) where $\kappa$ is now determined by Eq. (\ref{11}).
918:
919: The above results remain practically the same also in the presence
920: of screening electrodes in the vicinity of the sample. In that
921: case for $g \gg 1$ we again recover the logarithmic correction
922: (\ref{f1}) at large voltages while in the low voltage limit $eV\ll
923: \epsilon_{\rm Th}$ we find
924: \begin{equation}
925: \frac{dI}{dV}=\frac{Re^{-\kappa}}{R_t^2}\left[1+\frac{32}{15g}
926: \left(\frac{eV}{\epsilon_{\rm Th}}\right)^2\right].
927: \end{equation}
928: As before, in order to obtain the linear Andreev conductance $G_A(T)$ in
929: the regime $T \gg eV$ with the logarithmic accuracy, in the above
930: expressions it suffices to substitute $0.56 T$ instead of $eV$.
931:
932: \begin{figure}
933: \includegraphics[width=8cm]{fig3.eps}
934: \caption{The (normalized) value of the interaction correction to
935: the conductance $\delta G_A=G_A-G_A^{(0)}$ as a function of
936: voltage for $L_y=0.3L, L_z=0.1L$. One observes the saturation
937: regime at $eV \lesssim \epsilon_{\rm Th}$, the 1d logarithmic
938: regime (\ref{f1}) at $eV \gtrsim \epsilon_{\rm Th}$ which
939: eventually crosses over to the 3d regime (\ref{as3}) at $eV\sim
940: 10^5 \epsilon_{\rm Th}$. }
941: \end{figure}
942:
943: The differential conductance of quasi-1d diffusive NS structures
944: in the presence of electron-electron interactions is displayed in
945: Fig. 2 for several values of the parameter $8/g$. At larger
946: voltages one observes a competition between two types of ZBA --
947: conductance enhancement due to quantum interference of electrons
948: \cite{VZK,HN,Ben,Zai} and conductance suppression due to
949: electron-electron interactions. For small conductances $g < 16$
950: the second effect prevails and $dI/dV$ increases with voltage. At
951: large conductances $g > 16$ Coulomb effects become weaker and
952: $dI/dV$ shows the opposite trend. For smaller voltages $eV <
953: \epsilon_{\rm Th}$ the Andreev conductance saturates to the
954: voltage-independent value (\ref{satt}), (\ref{11}) for all values
955: of $g$. In this regime $dI/dV$ is always suppressed by
956: electron-electron interactions and this effect becomes stronger
957: with decreasing $g$.
958:
959: Note, that the quasi-1d logarithmic behavior (\ref{f1}) is rather
960: robust and can manifest itself for a wide range of voltages even
961: if the sample length $L$ is not really much bigger than its width
962: and thickness. An example of such a behavior is presented in Fig.
963: 3 for the sample with $L_y=0.3L, L_z=0.1L$. In this particular
964: case the logarithmic dependence of Eq. (\ref{f1}) remains
965: applicable within a very wide voltage range $\epsilon_{\rm
966: Th}\lesssim eV \lesssim 10^5\epsilon_{\rm Th}$, while the 3d
967: regime (\ref{as3}) may set in only at huge voltages $eV>10^5
968: \epsilon_{\rm Th}$.
969:
970:
971: \section{Current noise}
972:
973: Our effective action analysis also allows to study fluctuations of
974: Andreev current in diffusive NS structures in the presence of
975: electron-electron interactions. Here we restrict our attention to
976: the current noise. The noise spectrum is expressed in terms of the
977: current operators $\hat I$ in a standard way as
978: \begin{equation}
979: {\cal S}(t,t')=\frac{1}{2}\left\langle \hat I(t)\hat I(t')+\hat I(t')\hat
980: I(t)\right \rangle-\left\langle\hat I \right\rangle^2.
981: \end{equation}
982: This expression can easily be translated into
983: \begin{equation}
984: {\cal S}(t,t')=-ie^2\left\langle \int_\Gamma d^2 r \int_\Gamma d^2
985: r' \frac{\delta }{\delta \varphi^-(\bm{r},t)} \frac{\delta
986: }{\delta \varphi^-(\bm{r'},t')}\delta S_A \right\rangle ,
987: \label{defnoise}
988: \end{equation}
989: where $\delta S_A$ is defined in Eq. (\ref{finalaction}).
990: Employing this general expression for the Andreev action together
991: with the definitions of current and noise (\ref{curdef}),
992: (\ref{defnoise}) one can derive the following relationship
993: \begin{equation}
994: {\cal S}(\omega,V,T)=e\sum_{\pm}\coth\left( \frac{\omega\pm 2
995: eV}{2T}\right)I\left(\frac{\omega}{2e}\pm V,T\right),
996: \label{nc}
997: \end{equation}
998: which remains valid irrespective of the dimensionality at
999: frequencies and voltages smaller than both the inverse charge
1000: relaxation time $1/RC$ and the inverse elastic scattering time
1001: $1/\tau$.
1002:
1003: Note that the analogous expression between the current and the
1004: noise spectrum was derived in Ref. \onlinecite{LL} for normal
1005: tunnel barriers with interactions treated by means of the
1006: $P(E)$-theory and in Ref. \onlinecite{SBL} within a more general
1007: theoretical framework. Here Eq. (\ref{nc}) was derived for the
1008: Andreev current and the Andreev noise spectrum in diffusive NS
1009: structures in the presence of electron-electron interactions
1010: treated within a microscopic theory. Eq. (\ref{nc}) has exactly
1011: the same form as that \cite{LL,SBL}, one should only substitute
1012: $2e$ instead of $e$ for the charge of the charge carriers. In the
1013: zero frequency limit Eq. (\ref{nc}) reduces to the generalized
1014: Schottky relation
1015: \begin{equation}
1016: {\cal S}(\omega,V,T)=2e\coth (eV/T)I(V,T),
1017: \end{equation}
1018: previously obtained in Ref. \onlinecite{PBH} in the absence of
1019: interactions.
1020:
1021: Combining Eq. (\ref{nc}) with the results derived in the previous
1022: section one can fully describe the effect of Coulomb interaction
1023: on Andreev current noise in diffusive NS structures. Hence, there
1024: is no need to go into further details here. Let us only mention
1025: that -- similarly to the Andreev conductance -- at
1026: voltages/frequencies below $\epsilon_{\rm Th}$ we expect shot
1027: noise to be suppressed by the factor $\exp (-\kappa )$. At larger
1028: voltages $eV > \epsilon_{\rm Th}$ and in the low frequency limit
1029: the shot noise for quasi-1d NS structures should scale as
1030: \begin{equation}
1031: {\cal S} \propto |V|^{8/g-1/2}.
1032: \label{pl}
1033: \end{equation}
1034:
1035: \section{Discussion}
1036:
1037: The main observations of the present work can be summarized as
1038: follows. In hybrid NS structures electron-electron interactions
1039: generate two types of corrections to Andreev current. One of them
1040: is just a voltage-independent renormalization of the $I-V$ curve
1041: by the factor $\exp (-\kappa )$, while the other in general
1042: depends on voltage and/or temperature. The net effect of
1043: electron-electron interactions is always a reduction of the
1044: Andreev conductance. This reduction approaches its maximum value
1045: at temperatures and voltages below the Thouless energy of the
1046: sample. In this low energy regime the tunnel barrier and a normal
1047: metal act as a single coherent scatterer, whereas at energies above
1048: $\epsilon_{\rm Th}$ they can be treated as independent ones in the spirit
1049: of the $P(E)$-theory.
1050:
1051: The magnitude of Coulomb effects essentially depends on the system
1052: dimensionality and is characterized by the dimensionless parameter
1053: $\kappa$. In 3d systems we find $\kappa \sim 1/p_F^2l^2 \ll 1$,
1054: i.e. in this case Coulomb suppression of the Andreev conductance
1055: is weak except for very disordered samples. In the 2d case
1056: $\kappa$ is proportional to the parameter $1/p_F^2lL_z$ (which is
1057: again small) but is enhanced by the two logarithmic factors (cf.
1058: Eq. (\ref{22})), i.e. the net effect of electron-electron
1059: interactions is not necessarily weak in this case.
1060:
1061: This effect becomes even more pronounced in quasi-1d systems in
1062: which case the value $\kappa$ is defined by Eq. (\ref{11}).
1063: Obviously, the Coulomb suppression of the Andreev conductance can
1064: be very strong for such systems. Combining Eqs. (\ref{ZBA}),
1065: (\ref{ZBA0}) and (\ref{satt}) with Eq. (\ref{11}), and having in
1066: mind that for typical values of the Fermi velocity in metals one
1067: has $e^2/v_F \sim 1$, we arrive at a very simple estimate for the
1068: Andreev conductance in the presence of electron-electron
1069: interactions:
1070: \begin{equation}
1071: G_A/G_A^{(0)} \sim 1/N_{\rm Ch}^{8/g},
1072: \label{estim}
1073: \end{equation}
1074: where $N_{\rm Ch} \sim p_F^2\Gamma$ is the number of conducting
1075: channels in the normal wire. In Eq. (\ref{estim}) for simplicity
1076: we have disregarded an additional weak (logarithmic) dependence on
1077: $L$ and $\Gamma$, cf. Eq. (\ref{11}). Within this accuracy, Eq.
1078: (\ref{estim}) demonstrates that Coulomb suppression of the Andreev
1079: conductance in quasi-1d diffusive samples is determined by the
1080: number of conducting channels to the power $8/g$. We believe this
1081: prediction can be directly tested in future experiments.
1082:
1083: Recently the authors \cite{Takayanagi} have experimentally
1084: investigated the $I-V$ curves for hybrid systems composed of
1085: multi-walled carbon nanotubes (MWNT) attached to a superconducting
1086: electrode. The measurements \cite{Takayanagi} performed below the
1087: superconducting critical temperature $T_C$ have revealed
1088: power-law dependencies of the differential conductance on voltage
1089: and temperature $dI/dV \propto T^{\alpha_S}$ and $dI/dV \propto
1090: V^{\alpha_S}$. At smaller voltages $V < V_{\rm sat} \sim 1$ mV the
1091: conductance was found to saturate to a constant value. Also at
1092: $T>T_C$ the power-law dependencies have been
1093: observed $dI/dV \propto T^{\alpha_N}$ with the value $\alpha_N$
1094: somewhat smaller than $\alpha_S$.
1095:
1096: It is interesting to compare our theoretical predictions with these
1097: experimental results. The effect of conductance saturation
1098: observed in Ref. \onlinecite{Takayanagi} just matches with our
1099: prediction (\ref{estim}) and the value $eV_{\rm sat}$ appears to
1100: be in a good agreement with the Thouless energy $\epsilon_{\rm
1101: Th}$ estimated for MWNT \cite{Takayanagi}. Also the observed
1102: power-law dependencies are consistent with our theoretical
1103: picture. Identifying $\alpha_S=8/g-1/2$ and making use of the
1104: experimental value $\alpha_S \simeq 1.28$, we can estimate the
1105: effective dimensionless conductance for MWNT as $g \approx 4.5$.
1106: Above $T_C$ for the superconducting electrode one also expects the
1107: power-law dependence of the conductance on temperature
1108: \cite{SZ,IN}. Neglecting the impedance of the bulk electrode one
1109: would trivially get $\alpha_N = 2/g \approx 0.45$ which is
1110: slightly smaller than the measured values $\alpha_N \approx 0.77$
1111: close to $T_C$ and $\alpha_N \approx 0.55$ at higher temperatures.
1112: It is clear, however, that above $T_C$ the impedance of the
1113: superconducting electrode $R_S$ becomes non-zero and should also
1114: be taken into account. Accordingly, the value $\alpha_N$ should be
1115: modified as $\alpha_N = 2/g +2/g_S$, where $g_S=R_q/R_S$. This
1116: additional contribution to $\alpha_N$ could further improve the
1117: quantitative agreement between our theory and the experimental
1118: observations \cite{Takayanagi}.
1119:
1120: Finally, we would like to point out that it would also be
1121: interesting to experimentally study the effect of Coulomb
1122: interactions on the shot noise in quasi-1d hybrid NS systems,
1123: like, e.g., ones investigated in Ref. \onlinecite{Takayanagi}.
1124: According to the results obtained in Sec. VI, at low voltages one
1125: expects suppression of the shot noise by the factor $\sim
1126: N_{\rm Ch}^{8/g}$ while at higher voltages the power-law
1127: dependence (\ref{pl}) should be recovered. Simultaneous current
1128: and shot noise measurements could help to experimentally verify
1129: the general relation (\ref{nc}) between the noise spectrum and
1130: Andreev current in hybrid NS structures.
1131:
1132:
1133:
1134:
1135: \centerline{\bf Acknowledgments}
1136:
1137: We are grateful to D.S. Golubev for useful discussions. We also
1138: thank the authors of Ref. \onlinecite{Takayanagi} for
1139: communicating their results to us prior to publication.
1140:
1141: \begin{thebibliography}{}
1142:
1143: \bibitem{And} A.F. Andreev, Zh. Eksp. Teor. Fiz. {\bf 46}, 1823 (1964)
1144: [Sov. Phys. JETP {\bf 19}, 1228 (1964)].
1145: \bibitem{dG} P.G. de Gennes, {\it Superconductivity of Metals and Alloys}
1146: (Benjamin, New York, 1966).
1147: \bibitem{ALO} L.G. Aslamazov, A.I. Larkin, and Yu.N. Ovchinnikov,
1148: Zh. Eksp. Teor. Fiz. {\bf
1149: 55}, 323 (1968) [Sov. Phys. JETP {\bf 28}, 171 (1968)].
1150: \bibitem{ZZh} A.D. Zaikin and G.F. Zharkov, Fiz. Nizk. Temp. {\bf 7}, 375
1151: (1981) [Sov. J. Low Temp. Phys. {\bf 7}, 181 (1981)].
1152: \bibitem{KL} M.Yu. Kupriyanov and V.F. Lukichev, Zh. Eksp. Teor. Fiz. {\bf
1153: 94}, 139 (1988) [Sov. Phys. JETP {\bf 67}, 1163 (1988)].
1154: \bibitem{Dubos} P. Dubos, H. Courtois, B. Pannetier, F.K. Wilhelm,
1155: A.D. Zaikin, and G. Sch\"on, Phys. Rev. B {\bf 63}, 064502 (2001).
1156: \bibitem{Z} A.D. Zaikin. Solid State Commun. {\bf 41}, 533 (1982).
1157: \bibitem{BBS} W. Belzig, C. Bruder, and G. Sch\"on, Phys. Rev.
1158: B {\bf 53}, 5727 (1996).
1159: \bibitem{BTK} G.E. Blonder, M. Tinkham, and T.M. Klapwijk, Phys. Rev. B {\bf 25},
1160: 4515 (1982).
1161: \bibitem{VZK} A.F. Volkov, A.V. Zaitsev, and T.M. Klapwijk, Physica C {\bf 210}, 21
1162: (1993).
1163: \bibitem{HN} F.W.J. Hekking and Yu.V. Nazarov, Phys. Rev. Lett. {\bf 71}, 1625 (1993);
1164: Phys. Rev. B {\bf 49}, 6847 (1994).
1165: \bibitem{Ben} C.W.J. Beenakker, B. Rejaei, and J.A. Melsen, Phys. Rev. Lett.
1166: {\bf 72}, 2470 (1994).
1167: \bibitem{Zai} A.D. Zaikin, Physica B {\bf 203}, 255 (1994).
1168: \bibitem{HHK} A. Huck, F.W.J. Hekking, and B. Kramer, Europhys. Lett. {\bf
1169: 41}, 201 (1998).
1170: \bibitem{SZ} G. Sch\"on and A.D. Zaikin, Phys. Rep. {\bf 198}, 237 (1990).
1171: \bibitem{IN} G.-L. Ingold and Yu.V. Nazarov,
1172: in {\it Single Charge Tunneling}, ed. by H. Grabert and M. Devoret, NATO
1173: ASI Series B, vol. 294, pp. 21-107 (Plenum, 1992).
1174: \bibitem{Takayanagi} I. Takesue, T. Akazaki, S. Miyadai, N. Kobayashi,
1175: A. Tokita, M. Nomura, J. Haruyama, and H. Takayanagi, Physica E {\bf 24},
1176: 32 (2004).
1177: \bibitem{KA} A. Kamenev and A. Andreev, Phys. Rev.
1178: B {\bf 60}, 2218 (1999).
1179: \bibitem{FLS} M.V. Feigel'man, A.I. Larkin, and M.A. Skvortsov, Phys. Rev.
1180: B {\bf 61}, 12361 (2000).
1181: \bibitem{GZ} D.S. Golubev and A.D. Zaikin, Phys. Rev. B {\bf 59}, 9195
1182: (1999).
1183: \bibitem{SG} R.A. Serota and B. Goodman, Mod. Phys. Lett. B {\bf 13}, 649
1184: (1999).
1185: \bibitem{Grab} J. Rollb\"uhler and H. Grabert, Phys. Rev.
1186: Lett. {\bf 87}, 126804 (2001).
1187: \bibitem{LL} H. Lee and L.S. Levitov, Phys. Rev. B {\bf 53}, 7383 (1996).
1188: \bibitem{SBL} E.V. Sukhorukov, G. Burkard, and D. Loss, Phys. Rev. B
1189: {\bf 63}, 125315 (2001).
1190: \bibitem{PBH} F. Pistolesi, G. Bignon, and F.W.J. Hekking, Phys. Rev. B
1191: {\bf 69}, 214518 (2004).
1192: \end{thebibliography}
1193: \end{document}
1194: