cond-mat0112242/pp.tex
1: %\documentstyle[prb,aps,epsf]{revtex}
2: \documentstyle[multicol,prl,aps,epsf]{revtex}
3: 
4: %\def\top#1{\vskip #1\begin{picture}(290,80)(80,500)\thinlines \put(
5: %65,500){\line( 1, 0){255}}\put(320,500){\line( 0, 1){ 5}}\end{picture}}
6: %\def\bottom#1{\vskip #1\begin{picture}(290,80)(80,500)\thinlines \put(
7: %330,500){\line( 1, 0){255}}\put(330,500){\line( 0, -1){ 5}}\end{picture}}
8: 
9: \begin{document}
10: \title{Quantum interference and supercurrent
11: in multiple-barrier proximity structures}
12: \author{ Artem V. Galaktionov and Andrei D. Zaikin}
13: \address{Forshchungszentrum Karlsruhe, Institut f\"ur Nanotechnologie,
14: 76021, Karlsruhe, Germany\\
15: I.E. Tamm Department of Theoretical Physics, P.N. Lebedev Physics
16: Institute, Leninskii pr. 53, 119991 Moscow, Russia}
17: %\date{\today}
18: \maketitle
19: 
20: \begin{abstract}
21: We analyze an interplay between the proximity effect and quantum interference
22: of electrons in hybrid structures superconductor-normal metal-superconductor
23: which contain several insulating barriers. We demonstrate that the dc
24: Josephson current in these structures may change qualitatively due to quantum
25: interference of electrons scattered at different interfaces. In junctions with
26: few conducting channels mesoscopic fluctuations of the supercurrent are
27: significant and its amplitude can be strongly enhanced due to resonant
28: effects. In the many channel limit averaging over the scattering phase
29: effectively suppresses interference effects for systems with {\it two}
30: insulating barriers. In that case a standard quasiclassical approach
31: describing scattering at interfaces by means of Zaitsev boundary conditions
32: allows to reproduce the correct results. However, in systems with {\it three}
33: or more barriers the latter approach fails even in the many channel limit. In
34: such systems interference effects remain important in this limit as well. For
35: short junctions these effects result in additional suppression of the
36: Josephson critical current indicating the tendency of the system towards
37: localization. For relatively long junctions interference effects may -- on the
38: contrary -- enhance the supercurrent with respect to the case of independent
39: barriers.
40: 
41: \end{abstract}
42: 
43: 
44: \begin{multicols}{2}
45: 
46: 
47: \section{Introduction}
48: In recent years there has been a great deal of activity devoted to
49: both experimental and theoretical studies of mesoscopic superconducting-normal
50: ($SN$) hybrid structures\cite{lam,bel}). New
51: important phenomena such as anomalous Meissner screening,
52: re-entrant behavior of the
53: conductance, nonequilibrium-driven $\pi$-junction state and many others have
54: been discovered and thoroughly investigated. In some cases it was
55: found that an interplay between the proximity effect and
56: quantum interference of electrons in a normal metal play a significant
57: role at sufficiently low temperatures. For instance, interference of
58: electrons scattered at impurities in the normal layer may
59: strongly enhance the Andreev conductance $G_{A}$ of $SN$
60: systems leading to the so-called zero-bias anomaly\cite{VZK,HN,B}:
61: At low voltages
62: and temperatures $G_A$ turns out to depend linearly on the
63: $SN$ interface transmission $D$ in contrast to the standard result
64: $G_A\propto D^2$ obtained in the absence of interference effects
65: in the $N$-layer.
66: 
67: In this paper we will address a different -- although somewhat related --
68: problem. We will analyze the dc Josephson effect in $SNS$ systems which
69: contain several insulating barriers. In this case electrons scattered at
70: different barriers can interfere inside the junction. We will demonstrate that
71: this effect may lead to qualitative modifications of the supercurrent across
72: the junction. The most pronounced effect of quantum interference is expected
73: in $SNS$ systems with few conducting channels. This situation can be realized,
74: for instance, if two superconductors are connected via a carbon nanotube
75: \cite{Kas,Christ}. More conventional $SNS$ structures with many
76: conducting channels and several insulating barriers are also of
77: considerable interest, for instance in relation to possible
78: applications, see e.g. Ref. \onlinecite{Misha} and further references
79: therein. We will demonstrate that for such systems
80: quantum interference effects are also important
81: provided there exist more than two scatterers inside the junction.
82: 
83: A powerful tool for theoretical studies of mesoscopic
84: superconductivity is provided by the quasiclassical formalism of
85: the energy-integrated Eilenberger Green functions \cite{Eil}
86: (see also Refs. \onlinecite{lam,bel,lar,Albert} for a review).
87: The Eilenberger equations are, on one hand, much simpler than
88: the fully microscopic Gorkov equations and,
89: on the other hand, allow to correctly describe the system behavior at
90: distances much longer as compared to the Fermi wavelength $1/k_F$.
91: Since typical length scales in superconductors
92: (e.g. the coherence length $\xi_0$ or the London penetration
93: depth) are all several orders of magnitude greater than $1/k_F$, the
94: quasiclassical approach is usually an excellent approximation.
95: 
96: The quasiclassical equations cannot be applied only in the vicinity
97: of inter-metallic interfaces and barriers where rapid changes
98: of the system properties (at scales comparable to $1/k_F$) occur.
99: Fortunately, in many cases this problem can be circumvented by
100: matching the Eilenberger Green functions on both sides of the interface
101: with the aid of the proper boundary conditions. In order to derive
102: such conditions it is necessary to go beyond quasiclassics, however
103: under certain assumptions the final
104: result can be formulated
105: only in terms of the quasiclassical Eilenberger propagators. The derivation of
106: these boundary conditions was performed by Zaitsev\cite{zait}. Supplemented
107: by these boundary conditions, the Eilenberger quasiclassical formalism
108: was proven to be an extremely efficient tool for a quantitative description of
109: numerous inhomogeneous and hybrid superconducting structures.
110: 
111: 
112: 
113: An important ingredient of the derivation\cite{zait} is the
114: assumption that the
115: boundaries are situated sufficiently far from each other,
116: so that {\it interference effects} emerging from scattering at different
117: interfaces can be totally neglected. Under this assumption one arrives
118: at the nonlinear matching conditions involving the
119: third power of the quasiclassical propagators. These matching conditions
120: are expressed in terms of the interface transparency coefficients
121: for the electrons with different directions of the Fermi
122: momenta. It is also essential that Zaitsev boundary conditions do
123: not depend on scattering phases at the interface potentials.
124: 
125: Although for metallic structures containing one
126: interface one can indeed disregard interference effects,
127: in systems with several boundaries this is in general not
128: anymore possible. Hence, the applicability of
129: the nonlinear matching conditions\cite{zait} to multiple-barrier
130: systems requires additional analysis.
131: Some authors\cite{j1} argued that the standard quasiclassical approach
132: can break down in a multiple-interface geometry due to the problems with the
133: normalization of the Eilenberger functions.
134: 
135: In principle the above problem with boundary conditions can be avoided
136: within the approach based on the Bogolyubov-de Gennes
137: equations\cite{dG}. However, this approach, though frequently successful, may
138: also be technically inconvenient in complicated situations, for
139: instance because of a necessity to evaluate the energy eigenvalues
140: of the system and to perform summation over the energy spectrum
141: in the final results.
142: 
143: 
144: It is also possible to formulate an alternative quasiclassical
145: approach\cite{PZ} which allows to avoid the
146: abovementioned problems. Without going into details here let us just
147: mention that within the technique\cite{PZ} one deals with the quasiclassical
148: spinor amplitude $u,v$-functions which depend on
149: one coordinate and one time only and obey linear first order equations.
150: The Eilenberger Green functions are expressed via two
151: linearly-independent solutions of these equations
152: in a way that both the Eilenberger equations and
153: the normalization conditions are automatically satisfied. The
154: formalism\cite{PZ} -- just as the Eilenberger one -- can be formulated
155: both within the Matsubara and the Keldysh techniques and thus is suitable
156: both in equilibrium and non-equilibrium situations (see,
157: e.g. Ref. \onlinecite{GZ}) in various superconducting structures.
158: It is also important that very simple {\it linear} boundary conditions
159: for the quasiclassical amplitudes can be formulated at each of
160: the interfaces where electron scattering takes place. The number of
161: interfaces in the system is not restricted and the interference effects are
162: properly taken care of. Thus it is possible to take
163: advantage of the quasiclassical approximation and at the same time
164: to formulate general and simple boundary conditions without making
165: additional assumptions employed in Ref. \onlinecite{zait}.
166: 
167: Similar ideas have recently been put forward by Shelankov and
168: Ozana\cite{shel}. These authors also used linear matching conditions (obtained
169: by means of the scattering matrix approach) for the ``wave functions'' which
170: factorize the two-point Green functions. The next step\cite{shel} was to
171: construct quasiclassical one-point Green functions and formulate {\it
172: nonlinear} boundary conditions for such functions which would now adequately
173: include information about scattering on arbitrary number of ``knots''. Linear
174: boundary conditions were also used by Brinkman and Golubov\cite{brink} in a
175: calculation related to ours (see below).
176: 
177: 
178: In this paper, following Refs. \onlinecite{PZ,shel,brink}, we will
179: use simple linear boundary conditions in order to match the
180: quasiclassical amplitude functions at interfaces. However, unlike in
181: Ref. \onlinecite{shel}, we will avoid reformulating these boundary
182: conditions as nonlinear
183: ones for the Eilenberger Green functions. Rather we will directly
184: express the two-point Green functions and the expectation value of
185: the current operator in terms of the quasiclassical amplitudes.
186: We will then apply our method to the calculation of dc
187: Josephson currents in hybrid $SINI'S$ and $SINI'NI''S$ structures in
188: the clean limit and for arbitrary interface transmission coefficients
189: 
190: 
191: The interference of the scattering events at different interfaces manifests
192: itself in the expressions containing scattering phases $\phi$ at the interface
193: potentials. For the systems with two barriers (in our case
194: $SINI'S$-systems) with {\it many transmission channels} the
195: summation over their contributions is equivalent to effective
196: averaging over $\phi$. In this limit one can demonstrate that {\it after}
197: such averaging our result is equivalent one obtained from the
198: Eilenberger equations supplemented by the Zaitsev boundary
199: conditions. However, in the case of more than two barriers (i.e. for
200: $SINI'NI''S$ junctions) the dependence on
201: the scattering phases turns out to be much more essential. In this
202: situation the approach employing Zaitsev boundary conditions
203: turns out to fail also in the many channel limit where quantum interference
204: effects survive even after averaging over the scattering phases.
205: 
206: The paper is organized as follows. Our quasiclassical approach is
207: outlined in Sec. II. In Sec. III we apply this approach for the
208: analysis of the dc Josephson effect in $SINI'S$ structures with
209: arbitrary interface transmissions. The Josephson
210: current across $SINI'NI''S$ structures is evaluated in Sec. IV.
211: In Sec. V we present a brief discussion and summary of our results.
212: Some technical details of our calculation are relegated to Appendices.
213: 
214: 
215: \section{General method}
216: \subsection{Quasiclassical approximation}
217: The starting point of our analysis are the microscopic Gor'kov
218: equations\cite{AGD}. In what follows we will
219: assume that our system is uniform along the directions parallel to
220: the interfaces (coordinates $y$ and $z$).
221: Performing the Fourier transformation  of the normal
222: $G$ and anomalous $F^+$ Green function with respect to these coordinates
223: $$ G_{\omega_n}(\bbox{r}, \bbox{r'})=
224: \int\frac{d^2 \bbox{k_\parallel}}{(2\pi)^2}
225: G_{\omega_n} (x,x',\bbox{k_\parallel})e^{i
226: \bbox{k_\parallel}(\bbox{r_\parallel}- \bbox{r'_\parallel})}
227: $$
228: we express the Gor'kov equations in the following standard form
229: \begin{equation}
230: {\small \left( \begin{array}{cc} i\omega_n -\hat H & \Delta(x)\\ \Delta^*(x)&
231: i\omega_n +\hat H_c\end{array}\right)\left( \begin{array}{c} G_{\omega_n}
232: (x,x',\bbox{k_\parallel})\\F^+_{\omega_n} (x,x',\bbox{k_\parallel})
233: \end{array}\right)= \left(\begin{array}{c} \delta(x-x')\\ 0\end{array}
234: \right).} \label{start}
235: \end{equation}
236: Here $\omega_n=(2n+1)\pi T$ is the Matsubara frequency, and $\Delta(x)$ is the
237: superconducting order parameter. The Hamiltonian $\hat H$ in Eq.(\ref{start})
238: reads
239: \begin{equation}
240: \hat H=-\frac{1}{2m}\frac{\partial^2}{\partial x^2}+
241: \frac{\bbox{\tilde k^2_\parallel}}{2m}-\epsilon_F +V(x).
242: \label{H}
243: \end{equation}
244: Here $\bbox{\tilde k_\parallel}= \bbox{k_\parallel}-
245: \frac{e}{c}\bbox{A_\parallel}(x)$, $\epsilon_F$ is Fermi energy,
246: the term $V(x)$ accounts for the external
247: potentials (including the boundary potential). The Hamiltonian $\hat
248: H_c$ is obtained from $\hat H$ (\ref{H}) by inverting the sign of the electron
249: charge $e$. The above Hamiltonians can also include the self-energy
250: terms which, however, will not be considered below.
251: 
252: The quasiclassical approximation makes it possible to conveniently
253: separate fast oscillations of the Green functions due to the factor
254: $\exp(\pm ik_x x)$ from the
255: envelope of these functions changing at much longer scales as compared
256: to the atomic ones. Making use of this approximation for
257: two-component vector $\overline{\varphi}_\pm(x)\exp(\pm ik_x x)$ we
258: obtain
259: $$
260: \left( \begin{array}{cc} i\omega_n -\hat H & \Delta(x)\\ \Delta^*(x)&
261: i\omega_n +\hat H_c\end{array}\right)\overline{\varphi}_\pm(x)e^{\pm
262: ik_x x}
263: $$
264: \begin{equation}\simeq
265: e^{\pm ik_x x}\left( \begin{array}{cc} i\omega_n -\hat H^{a}_\pm &
266: \Delta(x)\\ \Delta^*(x)& i\omega_n +\hat H^{a}_{\pm c}\end{array}\right)
267: \overline{\varphi}_\pm(x),
268: \end{equation}
269: where we defined $k_x=\sqrt{k_F^2-k_\parallel^2}$ and
270: \begin{equation}
271: \hat H^{a}_\pm=\mp
272: iv_x\partial_x-\frac{e}{c}\bbox{A_\parallel}(x)\bbox{v_\parallel} +
273: \frac{e^2}{2m c^2}\bbox{A_\parallel}^2(x) +\tilde V(x).
274: \end{equation}
275: Here  $v_x=k_x/m$, $\tilde V(x)$ represents a slowly varying part of the
276: potential which {\it does not} include fast variations which may occur at
277: metallic interfaces. The latter will be accounted for by the boundary
278: conditions to be formulated below. But first let us briefly describe
279: the general structure of the Green functions obeying eq. (\ref{start}).
280: 
281: \subsection{Construction of the Green functions}
282: Consider the equation
283: \begin{equation}\left(
284: \begin{array}{cc} i\omega_n -\hat H^{a}_{\pm} & \Delta(x)\\ \Delta^*(x)&
285: i\omega_n +\hat H^{a}_{\pm c} \end{array}\right) \overline{\varphi}_\pm=0.
286: \label{appr}
287: \end{equation}
288: There exist two linearly
289: independent solutions $\overline{\varphi}_+$ of eq. (\ref{appr}). One such
290: solution (denoted below by $\overline{\varphi}_{+1}$) does not diverge
291: at $x\rightarrow +\infty$, the other solution $\overline{\varphi}_{+2}$
292: is well-behaved at $x\rightarrow -\infty$.
293: Similarly, two linearly independent solutions $\overline{\varphi}_{-1,2}$
294: do not diverge respectively at $x\rightarrow -\infty$ and
295: $x\rightarrow +\infty$.
296: 
297: A particular solution of the
298: Gor'kov equations (\ref{start}) can now be sought in the following form
299: 
300: \begin{eqnarray}
301: \left( \begin{array}{c} G_{\omega_n} (x,x',\bbox{k_\parallel})\\F^+_{\omega_n}
302: (x,x',\bbox{k_\parallel})\end{array}\right)=\overline{\varphi}_{+1}(x) g_1(x')
303: e^{ik_x(x-x')}+ \nonumber\\\overline{\varphi}_{-2}(x) g_2(x')
304: e^{-ik_x(x-x')}\quad \mbox{if}\: x>x' \label{xbp}
305: \end{eqnarray}
306: and
307: \begin{eqnarray}
308: \left( \begin{array}{c} G_{\omega_n} (x,x',\bbox{k_\parallel})\\F^+_{\omega_n}
309: (x,x',\bbox{k_\parallel})\end{array}\right)=\overline{\varphi}_{-1}(x) f_1(x')
310: e^{-ik_x(x-x')}+ \nonumber\\\overline{\varphi}_{+2}(x) f_2(x')
311: e^{ik_x(x-x')}\quad \mbox{if}\: x<x'\label{xmp}.
312: \end{eqnarray}
313: These functions satisfy Gor'kov equations at $x\neq x'$. The functions
314: $f_{1,2}(x)$ and $g_{1,2}(x)$ are determined with the aid of the
315: continuity condition for the Green functions at $x=x'$ and the condition
316: resulting from the integration of $\delta(x-x')$ in eq.(\ref{start}).
317: As a result we arrive at the linear equations
318: \begin{eqnarray}
319: && \overline{\varphi}_{+1}(x) g_1(x)+\overline{\varphi}_{-2}(x) g_2(x)=
320: \nonumber\\  &&\overline{\varphi}_{-1}(x) f_1(x)+\overline{\varphi}_{+2}(x)
321: f_2(x), \label{trr}\\ &&\frac{iv_x}{2}\Big[\overline{\varphi}_{+1}(x) g_1(x)-
322: \overline{\varphi}_{-2}(x) g_2(x)+ \nonumber\\&& \overline{\varphi}_{-1}(x)
323: f_1(x)-\overline{\varphi}_{+2}(x) f_2(x)\Big]=\left( \begin{array}{c}
324: 1\\0\end{array}\right),\nonumber
325: \end{eqnarray}
326: which can be trivially resolved.
327: 
328: 
329: For a homogeneous superconductor in the absence of the
330: magnetic field this procedure allows to immediately recover
331: the well known result
332: \begin{eqnarray}
333: & G_{\omega_n}(x,x')=-\frac{i}{v_x(1+\gamma^2)}\left( e^{ik_S|x-x'|}-
334: \gamma^2e^{-ik_S^*|x-x'|}\right),&\nonumber\\
335: &F^+_{\omega_n}(x,x')=\frac{\gamma e^{-i\chi}}{ v_x(1+\gamma^2)}\left(
336: e^{ik_S|x-x'|}+e^{-ik_S^*|x-x'|}\right),&\nonumber
337: \end{eqnarray}
338: where $\chi$ is the phase of the pairing potential,
339: $k_S=k_x+i\Omega_n/v_x$,
340: $\Omega_n=\sqrt{|\Delta|^2+\omega_n^2}$ and $
341: \gamma=\frac{|\Delta|}{\omega_n+\Omega_n}$.
342: Here for convenience we set  $\omega_n>0$.
343: 
344: In a non-homogeneous situation a general solution of the
345: Gor'kov equations takes the form
346: \begin{eqnarray}
347: &\left( \begin{array}{c} G_{\omega_n} (x,x')\\F^+_{\omega_n} (x,x')
348: \end{array}\right)= \left( \begin{array}{c}
349: G_{\omega_n} (x,x')\\F^+_{\omega_n} (x,x')\end{array}\right)_{part}+&\nonumber
350: \\ &[ l_1(x')\overline{\varphi}_{+1}(x) +l_2(x')\overline{\varphi}_{+2}(x)]
351: e^{ik_x x}+& \label{general} \\ &[l_3(x')\overline{\varphi}_{-1}(x)+
352: l_4(x')\overline{\varphi}_{-2}(x)]e^{-ik_x x}.&\nonumber
353: \end{eqnarray}
354: For systems which consist of several metallic layers the
355: particular solution is obtained with the aid of the procedure outlined above
356: provided both coordinates $x$ and $x'$ belong to the same layer. Should
357: $x$ and $x'$ belong to different layers, the particular solution is zero
358: because in that case the $\delta$-function in eq. (\ref{start}) fails.
359: The functions $l_{1,2,3,4}(x')$ in each layer should be derived from the proper
360: boundary conditions which we will now specify.
361: 
362: 
363: 
364: \subsection{Boundary conditions}
365: 
366: In what follows we shall assume interfaces to be non-magnetic. In this case
367: matching of the wave functions on the left and on the right side of a
368: potential barrier, respectively $A_1 \exp(ik_{1x}x)+ B_1 \exp(-ik_{1x}x)$ and
369: $A_2 \exp(ik_{2x}x)+ B_2 \exp(-ik_{2x}x)$, is performed in a standard way (see
370: e.g \onlinecite{LL}):
371: \begin{eqnarray}
372: && A_2=\alpha A_1+\beta B_1,\: B_2=\beta^*A_1 +\alpha^* B_1,\nonumber
373: \\
374: && |\alpha|^2-|\beta|^2=\frac{k_{1x}}{k_{2x}}. \label{scatt}
375: \end{eqnarray}
376: The reflection and transmission coefficients are given by
377: \begin{equation}
378: R=\left|\frac{\beta}{\alpha}\right|^2,\quad
379: D=1-R=\frac{k_{1x}}{k_{2x}|\alpha|^2}. \label{scatt2}
380: \end{equation}
381: For the sake of simplicity below we shall set $k_{1x}=k_{2x}$.
382: Since typical energies of interest, such as $\Delta$ and typical Matsubara
383: frequencies, are all much smaller than the magnitude of the interface
384: potentials, the relationships (\ref{scatt}) can be directly applied to
385: the two-element columns in eq. (\ref{general}). In this way we
386: uniquely determine the Green functions of our problem.
387: 
388: For an illustration let us consider a metallic layer with the left and right
389: boundaries located respectively at $x=d_1$ and $x=d_2$. We will also choose
390: the argument $x'$ inside this layer. As it was already explained, the
391: particular solution of the Gor'kov equations for $x<d_1$ or $x>d_2$ is
392: $G(x,x')=F^+(x,x')=0$, while it has the form (\ref{xbp}), (\ref{xmp}) if the
393: coordinate $x$ belongs to this layer. Thus at the left boundary we get
394: \begin{eqnarray}
395: &&\overline{\varphi}_{+2}(d_1) f_2(x') e^{-ik_x x'}+
396: l_1(x')\overline{\varphi}_{+1}(d_1)
397: +l_2(x')\overline{\varphi}_{+2}(d_1)=\nonumber\\
398: &&\alpha_1\left[l_1^L(x')\overline{\varphi}_{+1}^L(d_1)
399: +l_2^L(x')\overline{\varphi}_{+2}^L(d_1)\right]+\nonumber\\ &&\beta_1\left[
400: l_3^L (x')\overline{\varphi}_{-1}^L(d_1)+ l_4^L (x')\overline{\varphi}_{-2}^L
401: (d_1)\right]\label{leftb1}
402: \end{eqnarray}
403: and
404: \begin{eqnarray}
405: &&\overline{\varphi}_{-1}(d_1) f_1(x') e^{ik_x x'}+
406: l_3(x')\overline{\varphi}_{-1}(d_1)+
407: l_4(x')\overline{\varphi}_{-2}(d_1)=\nonumber\\ &&\beta_1^*
408: \left[l_1^L(x')\overline{\varphi}_{+1}^L(d_1)
409: +l_2^L(x')\overline{\varphi}_{+2}^L(d_1)\right] +\nonumber\\ &&\alpha_1^*
410: \left[ l_3^L (x')\overline{\varphi}_{-1}^L(d_1)+ l_4^L
411: (x')\overline{\varphi}_{-2}^L (d_1)\right].\label{leftb2}
412: \end{eqnarray}
413: The superscript $L$ labels the solutions in the layer
414: located at $x<d_1$. The above boundary conditions provide
415: four linear equations for the functions $l(x')$ with the source
416: term $f_{1,2}(x')\exp(\pm ik_x x')$. Similarly, with the aid of
417: eq. (\ref{xbp}) four boundary conditions at the right boundary $x=d_2$
418: can be established. Analogous procedure should be applied to other
419: interfaces.
420: 
421: 
422: 
423: 
424: \section{Josephson current in $\bbox{SINI'S}$ junctions}
425: 
426: We shall consider $SNS$ junctions composed of clean superconducting
427: ($S$) and normal ($N$) metals. We will assume that a thin insulating
428: layer ($I$) can be present at both $SN$ interfaces which, therefore, will
429: be characterized by arbitrary transparencies ranging from zero to one.
430: Specular reflection at both interfaces will be assumed. We also assume
431: that between interfaces electrons propagate ballistically and no
432: electron-electron or electron-phonon interactions are present in
433: the normal metal. For simplicity we will restrict our attention to the
434: case of identical superconducting electrodes with singlet
435: isotropic pairing. Furthermore, we shall neglect possible suppression
436: of the superconducting order parameter $\Delta$ in the electrodes
437: close to the $SN$ interface. This is a standard approximation which
438: is well justified in a large number of cases. The phase of the order
439: parameter is set to be $-\chi /2$ in the left electrode
440: and $\chi /2$ in the right one. The thickness of the normal layer
441: is denoted by $d$.
442: 
443: 
444: In order to evaluate the dc Josephson current across this structure
445: we shall follow the quasiclassical approach described
446: in the previous section. Technical details of our calculation are
447: presented in Appendix A. As a result we arrive at the expression
448: for the two point Green function in the normal layer. After that the
449: current density can be calculated from the standard formula
450: \begin{equation}
451:  J=\frac{ie}{m} T\sum_{\omega_n}\int
452: \frac{d^2k_\parallel}{(2\pi)^2}\left(\nabla_{x'}- \nabla_x\right)_{x'\to
453: x}G_{\omega_n} (x,x',\bbox{k_\parallel}). \label{ccuu}
454: \end{equation}
455: On can also rewrite the current in the form
456:  $J=J_+(\chi)-J_+^*(-\chi)$
457: where $J_+$ is defined by eq. (\ref{ccuu}) with positive
458: Matsubara frequencies $\omega_n >0$. Using (\ref{str}) and
459: omitting terms oscillating at atomic distances we obtain
460: \begin{equation}
461: J_+=2ie T\sum_{\omega_n>0} \int_{|k_\parallel|<k_F}
462: \frac{d^2k_\parallel}{(2\pi)^2}(V_1-U_2),
463: \end{equation}
464: or explicitly
465: \begin{equation}
466: J=4e T \sin\chi\sum_{\omega_n>0}\int_0^{k_F}
467: \frac{k_xdk_x}{2\pi} \frac{\sin\chi}{\cos\chi +W},
468: \label{J}
469: \end{equation}
470: where we defined
471: \begin{eqnarray}
472: W=\frac{4\sqrt{R_1R_2}}{D_1D_2}\frac{\Omega_n^2}{\Delta^2}\cos(2k_x d+\phi)
473: \nonumber
474: \\+\frac{\Omega_n^2
475: (1+R_1)(1+R_2)+ \omega_n^2D_1D_2}{D_1D_2\Delta^2}
476: \cosh\frac{2\omega_nd}{v_x}\label{W}\\
477: +\frac{2(1-R_1R_2)}{D_1D_2}
478: \frac{\Omega_n\omega_n}{\Delta^2}\sinh\frac{2\omega_nd}{v_x}.
479: \nonumber
480: \end{eqnarray}
481: Here $2k_xd+\phi$ is the phase of the product
482: $\alpha_2^*\beta_2\alpha_1^*\beta_1^*$. Eqs. (\ref{J}), (\ref{W})
483: provide a general expression for the dc Josephson current in $SINI'S$
484: structures valid for arbitrary transmissions $D_1$ and $D_2$ ranging
485: from zero to one. This expression is the central result of this
486: section.
487: 
488: We also note that the integral over $k_x$ in eq. (\ref{J}) can be
489: rewritten as a sum over independent conducting channels
490: \begin{equation}
491: \frac{{\cal A}}{2\pi}\int_0^{k_F}k_x dk_x(...) \to\sum_m^N (...),
492: \label{Dm}
493: \end{equation}
494: where ${\cal A}$ is the junction cross section. In this case
495: $D_{1,2}$ and $R_{1,2}$ may also depend on $m$. This would
496: correspond to different transmissions for different conducting channels.
497: 
498: Finally let us point out that in the limit of symmetrical
499: low transparent barriers $D_{1}=D_{2}\ll 1$
500: the problem was recently studied by Brinkman and Golubov\cite{brink}.
501: In the corresponding limit their result (eq. (8) of Ref.
502: \onlinecite{brink}) is similar -- although not fully equivalent --
503: to our eqs. (\ref{J}), (\ref{W}).
504: 
505: \subsection{Junctions with few conducting channels}
506: 
507: Let us first  analyze the above result for the case of one
508: conducting channel $N=1$. We observe that the first term in
509: eq. (\ref{W}) contains $\cos(2k_x d+\phi)$ which oscillates
510: at distances of the order of the Fermi wavelength. Provided at least
511: one of the barriers is highly transparent and/or
512: (for sufficiently long junctions $d \gtrsim \xi_0$) the temperature is high
513: $T \gg v_F/d$ this oscillating term is unimportant and can be
514: neglected. However, at lower transmissions of both barriers
515: and for relatively short
516: junctions $d \lesssim v_F/T$ this term turns out to be of the same order
517: as the other contributions to $W$ (\ref{W}). In this case
518: the supercurrent is sensitive to the exact positions of the discrete energy
519: levels inside the junction which can in turn vary considerably
520: if $d$ changes at the atomic scales $\sim 1/k_F$. Hence, one can
521: expect sufficiently strong sample-to-sample fluctuations of the
522: Josephson current even for junctions with nearly identical
523: parameters.
524: 
525: Let us first consider the limit of relatively short $SINI'S$
526: junctions in which case we obtain
527: \begin{equation}
528: I=\frac{e\Delta}{2}\frac{{\cal T}\sin \chi}{{\cal D}}
529: \tanh \left[\frac{{\cal D}\Delta }{2T}\right],
530: \label{I}
531: \end{equation}
532: where we defined
533: \begin{equation}
534: {\cal D}=\sqrt{1-{\cal T}\sin^2(\chi /2)}
535: \label{D1}
536: \end{equation}
537: and an effective normal transmission of the junction
538: \begin{equation}
539: {\cal T}=\frac{D_1D_2}{1+R_1R_2+2\sqrt{R_1R_2}\cos(2k_x d+\phi )}.
540: \label{T}
541: \end{equation}
542: Eq. (\ref{I}) has exactly the same functional form as the
543: result derived by Haberkorn {\it et al.} \cite{Hab}
544: for $SIS$ junctions with an arbitrary transmission
545: of the insulating barrier. This result is recovered from our eqs.
546: (\ref{I}), (\ref{T}) if we assume
547: e.g. $D_1 \ll D_2$ in which case the total transmission (\ref{T})
548: reduces to ${\cal T} \simeq D_1$.
549: 
550: As we have already discussed the total transmission ${\cal T}$ and,
551: hence, the Josephson current fluctuate depending on the exact position
552: of the bound states inside the junction. The resonant transmission is
553: achieved for $2k_xd +\phi=\pm\pi$, in which case we get
554: \begin{equation}
555: {\cal T}_{\rm res}=\frac{D_1D_2}{(1-\sqrt{R_1R_2})^2}.
556: \label{Tres}
557: \end{equation}
558: This equation demonstrates that for symmetric junctions $D_1=D_2$ at
559: resonance the Josephson current does not depend on the barrier
560: transmission at all. In this case ${\cal T}_{\rm res}=1$ and our
561: result (\ref{I}) coincides with the formula derived by Kulik and
562: Omel'yanchuk\cite{KO} for ballistic constrictions. In the limit
563: of low transmissions $D_{1,2} \ll 1$ we recover the standard
564: Breit-Wigner formula ${\cal T}_{\rm res}=4D_1D_2/(D_1+D_2)^2$
565: and reproduce the result obtained by Glazman and Matveev\cite{glazman}
566: for the problem of resonant tunneling through a single Anderson
567: impurity between two superconductors.
568: 
569: Note that our results (\ref{I}-\ref{T}) also support the conclusion
570: reached by Beenakker\cite{Ben} that the Josephson current across
571: sufficiently short junctions has a universal form and depends only on the
572: total scattering matrix of the weak link which can be evaluated in the
573: normal state. Although this conclusion is certainly correct in the
574: limit $d \to 0$, its applicability range depends significantly on the
575: physical nature of the scattering region. From eqs. (\ref{J}),
576: (\ref{W}) we observe that the result (\ref{I}), (\ref{D1}) applies
577: at $d \ll \xi_0$ not very close to the resonance. On the other hand,
578: at resonance the above result is
579: valid only under a more stringent condition
580: $d \ll \xi_0D_{\rm max}$, where we define $D_{\rm max}=$max$(D_1,D_2)$.
581: 
582: Now let us briefly analyze the opposite limit of sufficiently long junctions
583: $d\gg \xi_0$. Here we will restrict ourselves to the most interesting case
584: $T=0$. From eqs. (\ref{J}), (\ref{W}) we obtain
585: \begin{eqnarray}
586: && I=\frac{ev_x\sin\chi}{\pi d
587: z_1}\left[\frac{\arctan\sqrt{z_2/z_1}}{\sqrt{z_2/z_1}}\right],
588: \\&& z_1=\cos^2(\chi/2)+\frac{1}{D_1 D_2}\left( R_++2\sqrt{R_1 R_2}
589: \cos(2k_x d+\phi) \right), \nonumber
590: \\ && z_2=\sin^2(\chi/2)+\frac{1}{D_1 D_2}\left(
591: R_+-2\sqrt{R_1 R_2} \cos(2k_x d+\phi)\right),\nonumber
592: \end{eqnarray}
593: where $R_+=R_1+R_2$. For a fully transparent channel
594: $D_1=D_2=1$ the above expression
595: reduces to the well known Ishii-Kulik result\cite{Ishii,Kulik}
596: \begin{equation}
597: I=\frac{ev_x\chi}{\pi d},\quad -\pi<\chi<\pi,
598: \end{equation}
599: whereas if one transmission is small $D_1 \ll 1$ and $D_2 \approx 1$
600: we reproduce the result\cite{ZZh}
601: \begin{equation}
602: I=\frac{ev_xD_1\sin\chi}{2d}.
603: \label{Z}
604: \end{equation}
605: Provided the transmissions of both $NS$-interfaces are low $D_{1,2}\ll
606: 1$ we obtain in the off-resonant region
607: \begin{equation}
608: I=\frac{ev_x}{4\pi d}D_1D_2\sin\chi\Upsilon[2k_xd+\phi],
609: \end{equation}
610: where $\Upsilon[x]$ is a $2\pi$-periodic function defined as
611: \begin{equation}
612: \Upsilon[x]=\frac{x}{\sin{x}},\quad  -\pi<x<\pi.
613: \end{equation}
614: In the vicinity of the resonance $||2k_xd +\phi|-\pi
615:  |\lesssim D_{\rm max}$ the above result does not hold
616: anymore. Exactly at resonance $2k_xd +\phi=\pm \pi$ we get
617: \begin{equation}
618: I=\frac{ev_x\sqrt{D_1 D_2} \sin\chi}{4d
619: \left\{\cos^2\frac{\chi}{2}+\frac{1}{4}\left(
620: \sqrt{\frac{D_1}{D_2}}-\sqrt{\frac{D_2}{D_1}}\right)^2\right\}^{1/2}}.
621: \end{equation}
622: For a symmetric junction $D_{1,2}=D$ this formula yields
623: \begin{equation}
624: I=\frac{ev_xD \sin(\chi/2)}{2d},\quad  -\pi<\chi<\pi,
625: \end{equation}
626: while in a strongly asymmetric case $D_1 \ll D_2$ we again arrive
627: at the expression (\ref{Z}). This implies that at resonance the
628: barrier with higher transmission $D_2$ becomes effectively transparent
629: even if $D_2 \ll 1$. We conclude that for $D_{1,2} \ll 1$ the maximum
630: Josephson current is proportional to the product of transmissions
631: $D_1D_2$ off resonance, whereas exactly at resonance it is
632: proportional to the lowest of two transmissions $D_1$ or $D_2$.
633: 
634: We observe that both for short and long $SINI'S$ junctions interference
635: effects may enhance the Josephson effect or partially suppress it
636: depending on the exact positions of the bound states inside the junction. We
637: also note that in order to evaluate the supercurrent across $SINI'S$ junctions
638: it is in general {\it not} sufficient to derive the transmission probability
639: for the corresponding $NINI'N$ structure. Although the normal
640: transmission of the above structure is given by eq. (\ref{T}) for {\it all}
641: values of $d$, the correct expression for the Josephson current can be
642: recovered by combining eq. (\ref{T}) with the results\cite{Hab,Ben}
643: in the limit of short junctions $d \ll D\xi_0$ only. In this case one can
644: neglect suppression of the anomalous Green functions inside the normal layer
645: and, hence, the information about the normal transmission turns out to be
646: sufficient. On the contrary, for longer junctions the decay of Cooper pair
647: amplitudes inside the $N$-layer cannot be anymore disregarded. In this case
648: the supercurrent will deviate from the form (\ref{I}) even though the normal
649: transmission of the junction (\ref{T}) will remain unchanged. This deviation
650: becomes particularly pronounced for long junctions, i.e. for $d \gg \xi_0$ out
651: of resonance and for $d \gg D\xi_0$ at resonance.
652: 
653: The above analysis can trivially be generalized to the case of an
654: arbitrary number of independent conducting channels inside the
655: junction $N >1$. In that case the supercurrent is simply given by
656: the sum of the contributions from all the channels. Although
657: all these contributions have the same form, they are in general
658: not equal because the phase factors $2k_xd+\phi$ change randomly
659: for different channels. Accordingly, mesoscopic fluctuations
660: of the supercurrent should become smaller with increasing number
661: of channels and eventually disappear in the limit of large $N$.
662: In the latter limit the Josephson current is obtained by
663: averaging over all values of the phase $2k_xd+\phi$. The corresponding
664: results are presented below.
665: 
666: 
667: 
668: 
669: \subsection{Many channel limit}
670: 
671: Averaging over the phase factors $2k_xd+\phi$ is effectively performed by
672: integrating over directions of the electron momentum in eq. (\ref{J}). Since
673: the term in the expression for $W$ (\ref{W}) which contains $\cos(2k_x
674: d+\phi)$ oscillates very rapidly with changing $k_x$, averaging can be
675: performed by first integrating the current (\ref{J}) over the phase
676: $2k_xd+\phi$ and then integrating the result over $k_x$. We obtain
677: \begin{equation}
678: J=\frac{2}{\pi}ek_F^2 T \sin\chi\sum_{\omega_n>0}\int_0^1\mu
679: d\mu \frac{t_1(\mu )t_2(\mu )}{{\cal Q}^{1/2}(\chi , \mu)}.
680: \label{JN}
681: \end{equation}
682: Here and below we define $\mu =k_x/k_F$,
683: $
684: t_{1,2}(\mu)=D_{1,2}(\mu)/(R_{1,2}(\mu)+1)
685: $,
686: $t_\pm =t_1\pm t_2$ and
687: \begin{eqnarray}
688: \label{Q}
689: {\cal Q}=\left[t_1t_2\cos\chi+
690: \big(1+(t_1t_2+1)\frac{\omega_n^2}{\Delta^2} \big) \cosh\frac{2\omega_nd}{\mu
691: v_F}\right.\\
692: \left.+t_+\frac{\omega_n\Omega_n}{\Delta^2}
693: \sinh\frac{2\omega_nd}{\mu v_F}\right]^2-
694: (1-t_1^2)(1-t_2^2)\frac{\Omega_n^4}{\Delta^4}.
695: &\nonumber
696: \end{eqnarray}
697: The above equations fully determine the Josephson current in $SINI'S$
698: junctions in the many channel limit and at arbitrary transmissions of
699: specularly reflecting $SN$-interfaces.
700: 
701: Let us make use of this result in order to perform a direct comparison between
702: our analysis and the approach based on the Eilenberger equations supplemented
703: by Zaitsev boundary conditions. The corresponding calculation within the
704: latter approach is performed in Appendix B. It is interesting to observe that
705: for $SINI'S$ junctions this calculation yields exactly the same result
706: (\ref{JN}), (\ref{Q}) as obtained within our calculation after averaging over
707: the scattering phase $2k_xd +\phi$.
708: 
709: 
710: \begin{figure}
711: \centerline{\epsfclipon\epsfxsize=1.0\hsize\epsffile{intcurr.eps}} {\small
712: Fig. 1. The Josephson current density (\ref{Jour}) normalized by
713: $ek_F^2v_F/6\pi^2 d$ is plotted as a function of the phase difference $\chi$.
714: Here we assumed that the boundary is described by an effective potential
715: $U_0\delta(x\pm d/2)$ in which case one has $t=v_x^2/(2U_0^2+v_x^2)$. The
716: dependence $J(\chi )$ was evaluated for $2U_0^2/v_F^2=10^{-4},
717: 0.03,0.1,0.3,0.6$ (from top to bottom).}
718: \end{figure}
719: 
720: 
721: 
722: This observation allows to make an important conclusion concerning the
723: applicability of the quasiclassical analysis employing Zaitsev boundary
724: conditions for the Eilenberger propagators. The exact result for the Josephson
725: current in $SINI'S$ systems, eqs. (\ref{J}), (\ref{W}), cannot be recovered
726: within the latter approach because it essentially ignores interference effects
727: arising from electron scattering on two insulating barriers. At the same time,
728: in the limit of many conducting channels the scattering phase is effectively
729: averaged out. In this limit Zaitsev boundary conditions turn out to correctly
730: describe the supercurrent. It is also important to emphasize that the latter
731: conclusion applies for the systems with not more than two barriers. Below we
732: will analyze the supercurrent in $SNS$ structures with three insulating
733: barriers and will show that the approach based on Zaitsev boundary conditions
734: fails to provide correct results even in the limit of many conducting
735: channels.
736: 
737: 
738: 
739: 
740: But first let us present several limiting expressions for the sake of
741: completeness. We start from the limit of a sufficiently thick junction $d\gg
742: \xi_0$ and consider $T=0$. In this case we find
743: \begin{equation}
744: J=\frac{ek_F^2 v_F\sin\chi}{2 d\pi^2}\int_0^1  d\mu \mu^2
745: t_1t_2\sqrt{f_1f_2} F(\varphi, h),
746: \label{ell}
747: \end{equation}
748: where $F(\varphi,h)=\int_0^\varphi(1-h\sin^2\theta)^{-1/2}d\theta$ is the
749: incomplete elliptic integral, $\varphi =\arcsin (1/\sqrt{f_1})$,
750: $h=f_1+f_2-f_1f_2$ and
751: $$ f_1=\frac{1}{1-t_1t_2\cos^2\frac{\chi}{2}-
752: \frac{1}{2}\big[1-t_1t_2-\sqrt{(1-t_1^2)(1-t_2^2)}\big]},
753: $$
754: $$
755: f_2=\frac{1}{1-t_1t_2\sin^2\frac{\chi}{2}-
756: \frac{1}{2}\big[1-t_1t_2-\sqrt{(1-t_1^2)(1-t_2^2)}\big]}.
757: $$
758: For an $SINS$ junction ($t_2=1$) the above result yields (cf. Ref.
759: \onlinecite{svi})
760: $$
761: J=\frac{ek_F^2 v_F\sin\chi}{d\pi^2}\int_0^1 \frac{d\mu
762: \mu^2t_1}{\sqrt{1-t_1^2\cos^2\chi}}\arctan
763: \sqrt{\frac{1-t_1\cos\chi}{1+t_1\cos\chi}}.$$ The expression (\ref{ell}) also
764: simplifies in the case of a symmetric junction $t_1=t_2$
765: \begin{equation}
766: J=\frac{ek_F^2 v_F}{\pi^2
767: d}\int_0^1 \frac{\rho\mu^2 d\mu}{\sqrt{1+\rho^2}}
768: F\left(y,\frac{1}{1+\rho^2}\right),
769: \label{Jour}
770: \end{equation}
771: where
772: $$
773: y=\arccos\left[t\cos(\chi/2)\right],\;\;\;\;
774: \rho(\mu,\chi)=\frac{t^2\sin\chi}{2\sqrt{1-t^2}}.
775: $$
776: The current density $J$ (\ref{Jour}) is plotted in
777: Fig. 1 as a function of the Josephson phase $\chi$
778: for several values of the barrier transmission.
779: Note that in the case of small interface transparencies
780: the limit $T \to 0$ is effectively achieved at temperatures
781: much lower than $tv_F/d$.
782: 
783: 
784: 
785: 
786: Let us now proceed to the case of small transparencies of both interfaces
787: $t_{1,2}\ll 1$. In this limit the expression (\ref{Q}) takes the form
788: \begin{eqnarray}
789: {\cal Q}=\frac{\Omega_n^4}{\Delta^4} \left[ \sinh\frac{2\omega_n d}{\mu
790: v_F}+t_+\frac{\omega_n}{\Omega_n}\right]^2+\frac{\Omega_n^2}{\Delta^2}{\cal
791: P}(\mu, \chi), \label{qsm}
792: \end{eqnarray}
793: where
794: \begin{equation}
795: {\cal P}(\mu, \chi)= t_+^2(\mu)\cos^2(\chi/2)+
796: t_-^2(\mu)\sin^2(\chi/2).
797: \end{equation}
798: As we have already pointed out, the above result is not identical
799: to one presented in eq. (13) of Ref. \onlinecite{brink}
800: (see also \cite{Misha}). However, it
801: is easy to see that this difference does not affect the final expression
802: for the current in two important limits of short ($d\ll t\xi_0$)
803: and long ($d\gg t\xi_0$) junctions. Only in the intermediate case $d\sim
804: t\xi_0$ some deviations between our results and those of Ref.
805: \onlinecite{brink} are observed. This is demonstrated in Fig.2.
806: 
807: 
808: 
809: The case of short junctions $d\ll t\xi_0$ was already
810: studied in Ref. \onlinecite{brink}.
811: Therefore here we only present the asymptotic expression for the current at
812: $d\gg t\xi_0$
813: \begin{eqnarray}
814: J=\frac{ek_F^2v_F\sin\chi}{2\pi^2
815: d}\int_0^1 d\mu\mu^2 t_1t_2\ln(\epsilon_1/\epsilon_2),
816: \label{dras}
817: \end{eqnarray}
818: where $\epsilon_1=\mbox{min}\{\mu v_F/d,\Delta\}$, $\epsilon_2=\mu
819: v_F/(4d\sqrt{{\cal P}})$ for $T\ll tv_F/d$ and $\epsilon_2\simeq T$ for
820: $tv_F/d \ll T \ll \epsilon_1$.
821: The accuracy of the above formula is in general logarithmic, and it
822: becomes next to logarithmic in the limits $d\ll \xi_0$ or
823: $d\gg \xi_0$.
824: 
825: \begin{figure}
826: \centerline{\epsfclipon\epsfxsize100mm\epsffile{critic.eps}} {\small Fig. 2.
827: The maximum Josephson current $J_c$ of a symmetric $SINIS$-junction with low
828: transparency. The angular dependence of transparency is taken to be
829: $t(\mu)=t_0\mu^2$. The normal-state resistance of the structure is denoted by
830: $R_N$. The solid line represents eqs. (\ref{JN}), (\ref{qsm}) of our paper and
831: the dashed line represents eq. (13) of Ref. [18]. }
832: \end{figure}
833: 
834: 
835: 
836: \section{Josephson current in $\bbox{SINI'NI''S}$ junctions}
837: 
838: Let us now consider $SNS$ structure with three insulating barriers. As
839: before, two of them are located at $SN$ interfaces, and the third
840: barrier is inside the $N$-layer at a distance $d_1$ and $d_2$
841: respectively from the left and right $SN$ interfaces.
842: The transmission and reflection coefficients of this intermediate
843: barrier are denoted as $D_0$ and $R_0=1-D_0$, whereas the left and
844: the right barriers are characterized respectively by $D_1=1-R_1$ and
845: $D_2=1-R_2$.
846: 
847: The supercurrent is evaluated along the same lines as it was done in
848: Section 2 for the case of two barriers. A straightforward (although
849: somewhat lengthy and cumbersome) procedure yields the final result for the
850: Josephson current which is again expressed by eq. (\ref{J}),
851: where the function $W$ now takes the form
852: \begin{equation}
853: W=W_++W_-+W_{12} \label{wsum}
854: \end{equation}
855: These three contributions to the $W$-function depend respectively
856: on the sum of thicknesses $d_1$ and $d_2$, their difference and on these
857: two values separately. We find
858: \end{multicols}
859: %\top{-2.8cm}
860: \begin{eqnarray}
861: & W_{+}=\frac{4\sqrt{R_1R_2}}{D_0D_1D_2}\frac{\Omega_n^2}{\Delta^2}
862: \cos[2k_x (d_1+d_2)+\phi_1+\phi_2]+
863: \frac{2(1-R_1R_2)}{D_0D_1D_2}\frac{\Omega_n\omega_n}{\Delta^2}
864: \sinh\frac{2\omega_n(d_1+d_2)}{v_x} &\nonumber
865: \\ & +\frac{\Omega_n^2
866: (1+R_1)(1+R_2)+ \omega_n^2D_1D_2}{D_0D_1D_2\Delta^2}
867: \cosh \frac{2\omega_n(d_1+d_2)}{v_x};&
868: \\ & W_{-}=\frac{4R_0\sqrt{R_1R_2}}{D_0D_1D_2}\frac{\Omega_n^2}{\Delta^2}
869: \cos[2k_x (d_1-d_2)+\phi_1-\phi_2]+
870: \frac{2R_0(D_1-D_2)}{D_0D_1D_2}\frac{\Omega_n\omega_n}{\Delta^2}
871: \sinh\frac{2\omega_n(d_1-d_2)}{v_x}
872: &\nonumber
873: \\ & +\frac{R_0}{D_0}\left[1+\frac{2\Omega_n^2}{\Delta^2}
874: \frac{R_1+R_2}{D_1D_2} \right]\cosh \frac{2\omega_n(d_1-d_2)}{v_x};
875: \\ & W_{12}=\frac{4\sqrt{R_0R_1}}{D_0D_1}\cos[2k_x d_1+\phi_1]\left[
876: \frac{\Omega_n\omega_n}{\Delta^2}\sinh\frac{2\omega_n
877: d_2}{v_x}+\frac{1+R_2}{D_2}\frac{\Omega_n^2}{\Delta^2}
878: \cosh\frac{2\omega_n d_2}{v_x} \right]&\nonumber
879: \\  & +\frac{4\sqrt{R_0R_2}}{D_0D_2}\cos[2k_x d_2+\phi_2]\left[
880: \frac{\Omega_n\omega_n}{\Delta^2}\sinh\frac{2\omega_n
881: d_1}{v_x}+\frac{1+R_1}{D_1}\frac{\Omega_n^2}{\Delta^2}
882: \cosh\frac{2\omega_n d_1}{v_x} \right].&
883: \end{eqnarray}
884: %\bottom{-2.7cm}
885: \begin{multicols}{2}
886: Here we introduced two phases
887: \begin{eqnarray}
888: 2k_xd_{1}+\phi_{1}=\mbox{arg}\,\alpha_0^*\beta_0\alpha_{1}^*\beta_{1}^*,\\
889: 2k_xd_{2}+\phi_{2}=\mbox{arg}\,\alpha_0^*\beta_0^*\alpha_{2}^*\beta_{2},
890: \nonumber
891: \end{eqnarray}
892: related to the corresponding elements of the scattering matrices for all three
893: barriers. In contrast to the case of two barriers these phases cannot be
894: simultaneously removed by shifting $k_x$. The expression for the Josephson
895: current in $SINI'S$ junctions derived in the previous section can easily be
896: recovered if we set $D_0=1-R_0=1$. By setting $D_{1,2}=1-R_{1,2}=0$ in the
897: above equations we arrive at the result for the supercurrent in $SNINS$
898: systems derived in Ref. \onlinecite{ZZh}.
899: 
900: \subsection{One channel limit}
901: Let us first analyze the above general result in the limit of one conducting
902: channel. In the limit of short junctions $d\ll \xi_0D_{\rm max}$ we again
903: reproduce the result (\ref{I}) where the total effective transmission of the
904: normal structure with three barriers takes the form
905: \begin{equation}
906: {\cal T}=\frac{2t_1t_0t_2}{1+t_1t_0t_2+{\cal
907: C}(\varphi_{1,2},t_{0,1,2})},
908: \end{equation}
909: where
910: $$
911: {\cal C}=\cos\varphi_1\sqrt{(1-t_0^2)(1-t_1^2)}
912: +\cos\varphi_2\sqrt{(1-t_0^2)(1-t_2^2)}
913: $$
914: \begin{equation}
915: +(\cos\varphi_1\cos\varphi_2-t_0\sin\varphi_1\sin\varphi_2)
916: \sqrt{(1-t_1^2)(1-t_2^2)}.
917: \end{equation}
918: Here we define $t_{0,1,2}=D_{0,1,2}/(1+R_{0,1,2})$ and
919: $\varphi_{1,2}=2k_xd_{1,2}+\phi_{1,2}$. For later purposes let us also perform
920: averaging of this transmission over the phases $\varphi_1$ and $\varphi_2$. We
921: obtain
922: \begin{equation}
923: \langle {\cal T}\rangle=\frac{2 t_1t_0 t_2}{\sqrt{2  t_1t_0 t_2
924: +t_1^2t_0^2+t_1^2t_2^2+t_0^2t_2^2-t_1^2t_0^2t_2^2}}.
925: \label{3t}
926: \end{equation}
927: In particular, in the case of similar barriers with small
928: transparencies $D_{0,1,2}\approx D \ll 1$ the average normal
929: transmission of our structure is  $\langle {\cal T}\rangle \sim
930: D^{3/2}$. Suppression of the average transmission below the value
931: $\sim D$ is a result of destructive interference and indicates
932: the tendency of the system towards localization. Eq. (\ref{3t})
933: follows from an explicit integration, but it can also be
934: understood in simple terms.  Consider the square $0<\varphi_1<2\pi,
935: 0<\varphi_2<2\pi$. The main contribution to the average transmission
936: comes from the resonant region ${\cal T} \sim 1$. In the symmetric
937: case $t_{0,1,2}=t\ll 1$ this resonance occurs approximately along the
938: lines $\left[
939: \sqrt{(1+\cos\varphi_1)(1+\cos\varphi_2)}-t\right]^2\sim t^3$ in two quadrants
940: $\varphi_1,\varphi_2<\pi$ and $\varphi_1,\varphi_2>\pi$. In other
941: words, the resonant region is represented by two
942: hyperbola-like curves with characteristic widths $\sim
943: D^{3/2}$. This dependence of the average transmission is recovered from
944: the exact result (\ref{3t}).
945: 
946: Let us now proceed to the limit of a long junction $d_{1,2}\gg \xi_0$
947: and $T=0$. In the off-resonant region we find
948: \begin{equation}
949: I=\frac{ev_x D_1D_0D_2 \sin\chi}{8\pi d_1} {\cal B}(\varphi_{1,2}, d_2/d_1),
950: \label{I55}
951: \end{equation}
952: where
953: \begin{equation}
954: {\cal B}=\int_0^\infty \frac{dx}{\left[\cosh
955: x+\cos\varphi_1\right]\left[ \cosh (d_2 x/d_1)+\cos\varphi_2\right]}.
956: \label{B55}
957: \end{equation}
958: Evaluating this integral for $d_1=d_2$ we get
959: \begin{equation}
960: J=\frac{ev_x D_1D_0D_2 \sin\chi}{8\pi d_1}
961: \frac{\Upsilon[\varphi_1]-\Upsilon[\varphi_2]}{\cos\varphi_2-\cos\varphi_1}.
962: \end{equation}
963: This expression diverges at resonance (i.e. at $\varphi_1\simeq \pi$ or
964: $\varphi_2 \simeq \pi$) where it becomes inapplicable. In the resonant region
965: $\varphi_2 \simeq \pi$ we obtain
966: \begin{equation}
967: I=\frac{ev_x \sqrt{D_1D_0 D_2}\sin\chi}{4d
968: \sqrt{2(1+\cos\varphi_1)({\cal T}^{-1}-\sin^2(\chi/2))}}.
969: \end{equation}
970: 
971: 
972: \subsection{Many channel junctions}
973: As it was already discussed, in the many channel limit it is
974: appropriate to average the current over the scattering phases.
975: If the widths $d_1$ and $d_2$ fluctuate independently
976: on the atomic scale, averaging
977: over $\varphi_1$ and $\varphi_2$ can also be performed independently.
978: If $d_1$ and $d_2$ do not change on the atomic scale but
979: are incommensurate, independent averaging over the two phases can be performed
980: as well. The situation is different only for strictly
981: commensurate $d_1$ and $d_2$ in which case no independent averaging
982: can be fulfilled.
983: 
984: Let us first briefly discuss the latter situation of commensurate
985: $N$-layers. For simplicity we assume $d_1=d_2$, consider a symmetric situation
986: $D_1=D_2=D \ll 1$ and set the transparency of the intermediate interface to be
987: $D_0\gg D^2$. We will only present the result for the case of short junctions
988: $d\ll \xi_0D_{\rm max}$. We observe that the denominator in eq.
989: (\ref{J}), (\ref{wsum}) has a resonant structure as a function of
990: $\varphi_1+\varphi_2$. Integrating near the resonances we obtain
991: \begin{equation}
992: J= \frac{1}{2\pi}ek_F^2 T \sin\chi \int_0^1 d\mu \mu D(\mu)
993: \sum_{\omega_n>0}\frac{\Delta^2}{\Omega_n^2 \sqrt{{\cal R}(\chi,
994: \phi)}},
995: \end{equation}
996: where
997: \begin{equation}
998: {\cal R}=
999: (1+|\beta_0|^2\sin^2\phi)^2\left(1-\frac{\Delta^2\sin^2(\chi/2)}{
1000: \Omega_n^2(1+|\beta_0|^2\sin^2\phi )}\right) \label{newsh}
1001: \end{equation}
1002: and $|\beta_0|^2(\mu)=[1-D_0(\mu)]/D_0(\mu)$. An
1003: interesting feature of the expression (\ref{newsh}) is a dependence
1004: of the critical Josephson current on the scattering phase
1005: $\phi=(\phi_1-\phi_2)/2$. For instance, provided
1006: $|\beta_0|$ is large (the transparency of the intermediate layer is small),
1007: the critical current can vary from $\sim e k_F^2 |\Delta| D$ to $\sim e k_F^2
1008: |\Delta| D D_0$ with $\phi$ varying from 0 to $\pi/2$. One should also
1009: bear in mind that $\phi$ may depend on $\mu$. However, since
1010: the main contribution to the supercurrent comes predominantly
1011: from the electrons with the momenta perpendicular to the
1012: interfaces, we can estimate the current with $\phi$ corresponding to the
1013: forward direction. If $D_0\ll D^2$ and  $\sin^2\phi\gg D_0$ the current is
1014: given by eq. (\ref{newsh}). For $\phi=0$ (or
1015: $\phi=\pi$) a different expression follows
1016: \begin{equation}
1017: J=\frac{1}{2\pi} e k_F^2\sin\chi \Delta\tanh\frac{\Delta}{2T} \int_0^1 d\mu
1018: \mu \frac{D_0(\mu)}{D(\mu)}.
1019: \end{equation}
1020: 
1021: Now let us consider a more realistic
1022: situation of incommensurate $d_1$ and $d_2$ which allows for
1023: independent
1024: averaging over the scattering phases $\varphi_1$ and $\varphi_2$.
1025: Technically this procedure amounts to evaluating the integral
1026: of the expression $1/[t+\cos x\cos(\lambda x)]$ from $x=0$ to some
1027: large value $x=L$. At $\lambda=1$ the result of this
1028: integration is $L/\sqrt{t(1+t)}$.
1029: However, if $\lambda$ is irrational, the integral approaches the value
1030: $2LK(1/t^2)/\pi t$, where $K(h)=F(\pi/2,h)$ is the complete elliptic
1031: integral. This simple example illustrates our averaging procedure
1032: over two independent phases $x$ and $\lambda x$.
1033: 
1034: Let us assume that the transparencies of all three interfaces are
1035: small as compared to one. After
1036: averaging over $\varphi_1$ one arrives the expression which has a resonant
1037: dependence on $\varphi_2$ near $\varphi_2=\pi$. Expanding in powers
1038: of near this resonance
1039: with $\delta\varphi_2=\varphi_2-\pi$ and keeping the terms
1040: proportional to $\delta\varphi_2^2$ and $\delta\varphi_2^4$ we find
1041: \begin{eqnarray}&
1042: \left\langle\frac{1}{\cos\chi+W_++W_-+W_{12}}\right\rangle_{\varphi_1}=&
1043: \nonumber\\ & \frac{\Delta^2}{2\Omega_n^2} \left\{\frac{1}{D_1^{2}}+
1044: \frac{2\delta \varphi_2^2}{D_0D_1D_2}
1045: \left[1-\frac{2\Delta^2\sin^2(\chi/2)}{\Omega_n^2}\right]
1046: +\frac{\delta\varphi_2^4}{D_0^2D_2^2} \right\}^{-1/2}. &\nonumber
1047: \end{eqnarray}
1048: Then evaluating the integral over $\delta\varphi_2$
1049: we derive the final expression for the current
1050: \begin{equation}
1051: J=\frac{ek_F^2}{\pi^2}D_{\rm eff} \sin\chi T \sum_{\omega_n>0}\frac{\Delta
1052: ^2}{\Omega_n^2}K\left[ \frac{\Delta^2\sin^2(\chi/2)}{
1053: \Omega_n^2}\right],
1054: \label{kll}
1055: \end{equation}
1056: where we define the effective transmission
1057: \begin{equation}
1058: D_{\rm eff}=\int_0^1 \mu d\mu \sqrt{D_0D_1D_2}.
1059: \label{Deff}
1060: \end{equation}
1061: Hence, for similar barriers we obtain the dependence $J \propto
1062: D^{3/2}$ rather than $J \propto D$ (as it would be the case for independent
1063: barriers). The latter dependence would follow from the calculation
1064: based on Zaitsev boundary conditions for the Eilenberger propagators.
1065: We observe, therefore, that quantum interference effects {\it
1066: decrease} the Josephson current in systems with three insulating
1067: barriers. This is essentially quantum effect which cannot be recovered
1068: from Zaitsev boundary conditions even in the multichannel limit.
1069: This effect has exactly the same origin as a quantum suppression of
1070: the average normal transmission $\langle {\cal T} \rangle$ due to
1071: localization effects. Further limiting expressions for short junctions
1072: can be directly recovered from eq. (\ref{3t}).
1073: 
1074: We also note that the current-phase relation (\ref{kll}) deviates from
1075: a pure sinusoidal dependence even though all three transmissions are small
1076: $D_{0,1,2} \ll 1$.  At $T=0$ the critical
1077: Josephson current is reached at $\chi \simeq 1.7$ which is slightly higher
1078: than $\pi/2$. Although this deviation is quantitatively not very
1079: significant, it is nevertheless important as yet one more indication
1080: of quantum interference of electrons inside the junction.
1081: 
1082: Finally, let us turn to the limit of long junctions $d_{1,2}\gg\xi_0$. We
1083: again restrict ourselves to the case of low transparent interfaces. At high
1084: temperatures $T\gg v_F/2\pi d_{1,2}$  from eq. (\ref{J}),(\ref{wsum}) we get
1085: \begin{equation}
1086: J=\frac{eT  k_F^2}{\pi}\frac{\Delta^2\sin\chi}{\Delta^2+\pi^2 T^2}
1087: \int_0^1 d\mu \mu D_0D_1D_2e^{-\frac{d}{\xi(T)\mu}},
1088: \label{lds}
1089: \end{equation}
1090: where $d=d_1+d_2$ and $\xi(T)=v_F/(2\pi T)$. In this case the anomalous
1091:  Green function strongly decays deep in the normal layer. Hence,
1092: interference effects are not important and the interfaces can be
1093: considered as independent from
1094: each other. In the opposite limit $T\to 0$ (more precisely $T\ll Dv_F/d$),
1095: however, interference effects become important, and the current becomes
1096: proportional to $D^{5/2}$ rather than to $D^3$. Explicitly, at $T \to 0$ we
1097: get
1098: \begin{equation}
1099: J=\frac{e k_F^2 v_F\sin\chi}{16\pi^2\sqrt{d_1d_2}}\int_0^1d\mu \mu^2
1100: D_1D_2\sqrt{D_0}\ln D_0^{-1}.
1101: \label{inc}
1102: \end{equation}
1103: This expression is valid with the logarithmic accuracy and
1104: no distinction between $\ln D_0$, $\ln D_1$ or
1105: $\ln D_2$ should be made. We see that, in contrast to short junctions,
1106: in the limit of thick
1107: normal layers interference effects {\it increase} the Josephson
1108: current as compared to the case of independent barriers. The
1109: result (\ref{inc}), as well as one of eqs. (\ref{kll}) (\ref{Deff}) cannot be
1110: obtained from the Eilenberger approach
1111: supplemented by Zaitsev boundary conditions.
1112: 
1113: \section{Discussion and conclusions}
1114: Let us summarize our key results and observations.
1115: 
1116: In the present work we considered an interplay between the proximity
1117: effect and quantum interference of electrons in hybrid structures
1118: composed of normal metallic layers and superconductors. Quantum
1119: interference effects occur between electrons scattered at different
1120: metallic interfaces or other potential barriers and can strongly
1121: influence the supercurrent across the system.
1122: 
1123: The standard quasiclassical approach which describes scattering at interfaces
1124: by means of the nonlinear boundary conditions \cite{zait} for
1125: energy-integrated Eilenberger propagators -- while very efficient in numerous
1126: other situations -- is in general not suitable for the problem in question.
1127: Because of this reason we made use of an alternative quasiclassical approach
1128: which allows to investigate superconducting systems with more than one
1129: potential barrier and fully accounts for the interference effects. Within this
1130: approach scattering at boundaries is described with the aid of linear boundary
1131: conditions for quasiclassical amplitude functions. Electron propagation
1132: between boundaries is described by linear quasiclassical equations. Our
1133: approach is technically not equivalent to one based on the Bogolyubov-de
1134: Gennes equations. In particular, our method allows to explicitly construct
1135: two-point Green functions of the system and bypass such intermediate steps as
1136: finding an exact energy spectrum of the system with subsequent summation over
1137: the energy eigenvalues inevitable within the Bogolyubov-De Gennes approach. On
1138: the other hand, if needed, the full information about the energy bound states
1139: can easily be recovered within our technique by finding the poles of the Green
1140: functions in the Matsubara frequency plane.
1141: 
1142: Within our method we evaluated the dc Josephson current in $SNS$
1143: junctions containing two and three insulating barriers with
1144: arbitrary transmissions, respectively $SINI'S$ and
1145: $SINI'NI''S$ junctions. For the system with two barriers and few
1146: conducting channels we found
1147: strong fluctuations of the Josephson critical current
1148: depending on the exact position of the resonant level inside
1149: the junction. For short junctions $d \ll \xi_0D$ at
1150: resonance the Josephson current does not depend on the barrier
1151: transmission $D$ and is given by the standard Kulik-Omel'yanchuk
1152: formula \cite{KO} derived for ballistic weak links. In the limit
1153: of long $SNS$ junctions $d \gg \xi_0$ resonant effects may also
1154: lead to strong enhancement of the supercurrent, in this
1155: case at $T \to 0$ and at resonance the Josephson current is proportional to $D$
1156: and not to $D^2$ as it would be in the absence of interference
1157: effects.
1158: 
1159: It is also interesting to observe that, while the above
1160: results for few conducting channels cannot be obtained by means
1161: of the approach employing Zaitsev boundary conditions, in the many
1162: channel limit and for junctions with two barriers the latter
1163: approach {\it does} allow to recover correct results. This is
1164: because the contributions sensitive to the scattering phase
1165: are effectively averaged out during summation over conducting
1166: channels or, which is the same, during averaging of the current
1167: over the directions of the Fermi velocity.
1168: 
1169: Quantum interference effects turn out to be even more
1170: important in the proximity systems which contain three insulating
1171: barriers. In this case the quasiclassical approach based on Zaitsev
1172: boundary conditions fails even in the limit of many conducting
1173: channels. In that limit the Josephson current is {\it decreased}
1174: for short junctions ($J \propto D^{3/2}$) as compared to the
1175: case of independent barriers ($J \propto D$). This effect is caused by
1176: destructive interference of electrons reflected from different
1177: barriers and indicates the tendency of the system towards
1178: localization. In contrast, for long $SNS$ junctions with three
1179: barriers an interplay between quantum interference and proximity
1180: effect leads to enhancement of the Josephson current at $T \to 0$:
1181: We obtained the dependence $J \propto D^{5/2}$ instead of $J \propto
1182: D^3$ for independent barriers. We also discuss some further concrete
1183: results which turn out to be quite sensitive to the details of the model.
1184: 
1185: Finally, we note that in a very recent publication\cite{shel2} Ozana and
1186: Shelankov analyzed the applicability of the quasiclassical technique for the
1187: case of superconducting sandwiches with several insulating barriers. For such
1188: systems they also arrived at the conclusion that in the many channel limit the
1189: standard quasiclassical scheme based on the Eilenberger equations {\it and}
1190: Zaitsev boundary conditions effectively breaks down in the presence of more
1191: than two reflecting interfaces. These authors argued that in such cases this
1192: scheme disregards certain classes of interfering quasiclassical paths. This
1193: conclusion\cite{shel2} is similar to one reached in the present paper for
1194: $SNS$ structures. We would like to point out, however, that from our point of
1195: view the failure of the above scheme is not so much due to the
1196: quasiclassical approximation
1197: and/or normalization conditions employed within the Eilenberger formalism.
1198: The problem is rather in the boundary conditions\cite{zait} which
1199: disregard interference effects which occur in the structures with
1200: several interfaces/barriers with transmissions smaller than one.
1201: 
1202: 
1203: 
1204: 
1205: We would like to thank J.C. Cuevas, D.S. Golubev, A.A. Golubov, M. Eschrig,
1206: A. Shelankov, G. Sch\"on and U. Z\"ulicke for discussions and useful remarks.
1207: The work is part of the {\bf CFN} (Center for
1208: Functional Nanostructures) which is supported by the DFG (German Science
1209: Foundation). We also acknowledge partial support of RFBR under Grant No.
1210: 00-02-16202.
1211: 
1212: \end{multicols}
1213: 
1214: 
1215: \appendix
1216: 
1217: \section{}
1218: Let us consider an $SINI'S$ system and assume that the normal metal layer is
1219: located at $-d/2<x<d/2$. It is convenient to choose the coordinate $x'$ within
1220: the normal layer, $-d/2<x'<d/2$. Then a general solution of eq. (\ref{start})
1221: (decaying at $x \to -\infty$) in the left superconductor reads
1222: \begin{equation}
1223: \left( \begin{array}{c} G_{\omega_n} (x,x')\\F^+_{\omega_n}
1224: (x,x')\end{array}\right)=  \left( \begin{array}{c} 1\\-i
1225: e^{i\chi/2}\gamma^{-1}
1226: \end{array}\right)e^{\kappa x/v_x}e^{ik_x x} f(x')+
1227: \left( \begin{array}{c} 1\\i e^{i\chi/2}\gamma
1228: \end{array}\right)e^{\kappa x}e^{-ik_x x} g(x'). \label{ls}
1229: \end{equation}
1230: Here $\kappa=\Omega_n/v_x$. The solution in the right superconductor can found
1231: analogously. We get
1232: \begin{equation}
1233: \left( \begin{array}{c} G_{\omega_n} (x,x')\\F^+_{\omega_n}
1234: (x,x')\end{array}\right)=  \left( \begin{array}{c} 1\\i e^{-i\chi/2}\gamma
1235: \end{array}\right)e^{-\kappa x}e^{ik_x x} n(x')+
1236: \left( \begin{array}{c} 1\\-i e^{-i\chi/2}\gamma^{-1}
1237: \end{array}\right)e^{-\kappa x}e^{-ik_x x} r(x').
1238: \label{rs}
1239: \end{equation}
1240: The above solutions contain four unknown functions $f(x')$, $g(x')$, $n(x')$
1241: and $r(x')$. These functions should be found by matching (\ref{ls}),
1242: (\ref{rs}) with the solution of eq. (\ref{start}) in the normal layer. The
1243: latter has the form
1244: \begin{eqnarray}
1245: \left( \begin{array}{c} G_{\omega_n} (x,x')\\F^+_{\omega_n}
1246: (x,x')\end{array}\right)=\left(
1247: \begin{array}{c}-\frac{i}{v_x} e^{[ik_x-(\omega_n/v_x)]|x-x'|}\\0\end{array}
1248: \right)+e^{-\omega_nx/v_x}e^{ik_x x} h(x')\left(
1249: \begin{array}{c} 1\\0\end{array}\right)+ e^{\omega_nx/v_x}e^{-ik_x x} j(x')
1250: \left(
1251: \begin{array}{c} 1\\0\end{array}\right) \label{norm} \\
1252: +e^{\omega_nx/v_x}e^{ik_x x} k(x')\left(
1253: \begin{array}{c} 0\\1\end{array}\right)+e^{-\omega_nx/v_x}e^{-ik_x x} l(x')
1254: \left(
1255: \begin{array}{c} 0\\1\end{array}\right)\nonumber
1256: \end{eqnarray}
1257: and contain four additional unknown functions $h(x')$, $j(x')$,
1258: $k(x')$ and $l(x')$. The boundary conditions at two $NS$ interfaces
1259: provide eight equations which allow to uniquely determine all the
1260: above functions and, hence, the supercurrent in $SINI'S$ junctions.
1261: These equations are specified below.
1262: 
1263: Consider the left boundary. Making use of eq. (\ref{leftb1}) we find
1264: \begin{equation}
1265: \left( \begin{array}{c} h(x')e^{\frac{\omega_n d}{2v_x}}\\
1266: k(x')e^{-\frac{\omega_n d}{2v_x}} \end{array}\right) =\alpha_1
1267: \left(\begin{array}{c} 1\\-i e^{i\chi/2}\gamma^{-1}
1268: \end{array}\right)e^{-\kappa d/2}f(x')+\beta_1
1269: \left( \begin{array}{c} 1\\i e^{i\chi/2}\gamma
1270: \end{array}\right)e^{-\kappa d/2}g(x'), \label{f}
1271: \end{equation}
1272: while eq. (\ref{leftb2}) yields
1273: \begin{equation}
1274: \left( \begin{array}{c} -\frac{i}{v_x}e^{ik_x x'} e^{-\frac{\omega_n}{v_x}x'}
1275: e^{\frac{-\omega_n d}{2v_x}}\\ 0
1276: \end{array}\right) +\left(
1277: \begin{array}{c} j(x')e^{\frac{-\omega_n d}{2v_x}}\\ l(x')e^{\frac{\omega_n
1278: d}{2v_x}}
1279: \end{array}\right) =\alpha_1^* \left(\begin{array}{c} 1\\i e^{i\chi/2}\gamma
1280: \end{array}\right)e^{-\kappa d/2}g(x')+\beta_1^*
1281: \left( \begin{array}{c} 1\\-i e^{i\chi/2}\gamma^{-1}
1282: \end{array}\right)e^{-\kappa d/2}f(x'). \label{s}
1283: \end{equation}
1284: Similarly, applying the boundary conditions at the right interface one gets
1285: \begin{equation}
1286: \left( \begin{array}{c} 1\\i e^{-i\chi/2}\gamma
1287: \end{array}\right)e^{-\kappa d/2} n(x')=\alpha_2
1288: \left( \begin{array}{c} -\frac{i}{v_x}e^{-ik_x x'} e^{\frac{\omega_n}{v_x}x'}
1289: e^{\frac{-\omega_n d}{2v_x}}+e^{\frac{-\omega_n d}{2v_x}}h(x')\\
1290: e^{\frac{\omega_n d}{2v_x}}k(x') \end{array}\right)+\beta_2 \left(
1291: \begin{array}{c} e^{\frac{\omega_n d}{2v_x}}j(x')\\
1292: e^{-\frac{\omega_n d}{2v_x}}l(x') \end{array}\right), \label{t}
1293: \end{equation}
1294: and
1295: \begin{equation}
1296: \left( \begin{array}{c} 1\\-i e^{-i\chi/2}\gamma^{-1}
1297: \end{array}\right)e^{-\kappa d/2} r(x')=
1298: \alpha_2^* \left(
1299: \begin{array}{c} e^{\frac{\omega_n d}{2v_x}}j(x')\\
1300: e^{-\frac{\omega_n d}{2v_x}}l(x') \end{array}\right)+ \beta_2^* \left(
1301: \begin{array}{c} -\frac{i}{v_x}e^{-ik_x x'} e^{\frac{\omega_n}{v_x}x'}
1302: e^{\frac{-\omega_n d}{2v_x}}+e^{\frac{-\omega_n d}{2v_x}}h(x')\\
1303: e^{\frac{\omega_n d}{2v_x}}k(x') \end{array}\right). \label{fo}
1304: \end{equation}
1305: It is easy to see that the free terms in eqs. (\ref{f})-(\ref{fo})
1306: are
1307: \begin{equation}
1308: z_1(x')=-\frac{i}{v_x}e^{ik_x x'-(\omega_n x'/v_x)},\:
1309: z_2(x')=-\frac{i}{v_x}e^{-ik_x x'+(\omega_n x'/v_x)}.
1310: \end{equation}
1311: Eqs. (\ref{f})-(\ref{fo}) can be trivially resolved and we arrive at the
1312: solutions for the functions $h(x')$ and $j(x')$ (which are only
1313: needed in order to evaluate the current) with the structure
1314: \begin{equation}
1315: h(x')=U_1z_1(x')+U_2z_2(x'),\quad j(x')=V_1z_1(x')+V_2z_2(x'),\label{str}
1316: \end{equation}
1317: where $U_{1,2}$ and $V_{1,2}$ do not depend on $x'$.
1318: 
1319: \begin{multicols}{2}
1320: 
1321: 
1322: \section{}
1323: Consider the Eilenberger quasiclassical propagator which has a $2\times
1324: 2$ matrix structure in the Nambu space
1325: \begin{equation} \hat g(\bbox{p},\bbox{R},\omega_n)=\left(
1326: \begin{array}{cc} g(\bbox{p},\bbox{R},\omega_n) &
1327: if(\bbox{p},\bbox{R},\omega_n)\\ -if^+(\bbox{p},\bbox{R},\omega_n) &
1328: -g(\bbox{p},\bbox{R},\omega_n) \end{array} \right). \end{equation}
1329: Here $\bbox{p}$ is the electron momentum on the Fermi surface
1330: and $\bbox{R}$ is its coordinate.
1331: This quasiclassical propagator obeys the normalization condition
1332: $\hat g^2=1$ or, equivalently, $g^2+ff^+=1$. In addition, anomalous ($f,
1333: f^+$) and normal ($g$) Green functions obey important symmetry relations
1334: \begin{eqnarray}
1335: &f^{+*}(\bbox{p},\bbox{R},\omega_n)=f(-\bbox{p},\bbox{R},\omega_n)=
1336: f(\bbox{p},\bbox{R},-\omega_n), & \label{sym} \\ &
1337: g^*(\bbox{p},\bbox{R},\omega_n)=g(-\bbox{p},\bbox{R},\omega_n)=
1338: -g(\bbox{p},\bbox{R},-\omega_n).& \nonumber
1339: \end{eqnarray}
1340: The Eilenberger equations \cite{Eil} can be written in a concise matrix form as
1341: \begin{eqnarray}
1342: &i\bbox{v}_F\nabla \hat g+\hat\omega \hat g -\hat g\hat\omega=0,& \label{E}\\
1343: &\displaystyle\hat \omega
1344: =\bigl(i\omega_n+\frac{e}{c}\bbox{v}_F\bbox{A}\bigr)\hat\sigma_z
1345: -\hat\Delta+\frac{i}{2\tau}\langle \hat g\rangle + \frac{i}{2\tau_s}\langle
1346: \hat \sigma_z\hat g \hat\sigma_z\rangle,& \nonumber
1347: \end{eqnarray}
1348: where $\bbox{A}$ stands for the vector-potential; $\tau$ and $\tau_s$ are the
1349: elastic scattering time on nonmagnetic and paramagnetic impurities,
1350: respectively. The angular brackets denote averaging over the Fermi surface.
1351: The matrix $\hat\Delta (\bbox{R})$ incorporates the superconducting order
1352: parameter $\Delta (\bbox{R})$, $\hat\sigma_z$ is the Pauli matrix
1353: \begin{equation} \hat\Delta=\left( \begin{array}{cc}0 & \Delta\\-\Delta^* & 0
1354: \end{array}\right), \quad \hat\sigma_z=\left(
1355: \begin{array}{cc} 1 & 0\\ 0 & -1 \end{array}\right).  \end{equation}
1356: The current density $\bbox{j}(\bbox{R})$ is defined as follows
1357: \begin{equation}
1358: \bbox{j}(\bbox{R})= -2\pi i e TN(0)\sum_m\langle
1359: \bbox{v}_F(\bbox{p})g(\bbox{p},\bbox{R},\omega_n)\rangle
1360: \label{Jeil}
1361: \end{equation}
1362: $N(0)$ is the density of states at the Fermi energy per one spin direction.
1363: 
1364: The Eilenberger equations (\ref{E}) should be supplemented by Zaitsev boundary
1365: conditions at metallic interfaces. These conditions have the form\cite{zait}
1366: \begin{equation}
1367: \hat g_{a+}=\hat g_{a-}=\hat g_a,
1368: \label{bound1}
1369: \end{equation}
1370: $$\hat
1371: g_a\bigl[(1-D(\bbox{p}))(\hat g_{s+}+\hat g_{s-})^2+ (\hat g_{s+}-\hat
1372: g_{s-})^2\bigr]
1373: $$
1374: \begin{equation}
1375: =D(\bbox{p})\bigl[ \hat g_{s+}\hat g_{s-}-\hat g_{s-}\hat
1376: g_{s+}\bigr].
1377: \label{bound2}
1378: \end{equation}
1379: Here $\hat g_{s,a}(\bbox{p},\bbox{R},\omega_n)= [\hat
1380: g(\bbox{p},\bbox{R},\omega_n)\pm \hat g(\bbox{p}_r,\bbox{R},\omega_n)] /2$, by
1381: $\bbox{p}_r$ we denote the reflected momentum. The subscripts $\pm$ in
1382: eqs. (\ref{bound1}), (\ref{bound2}) stand for the expressions on the right (left)
1383: side of the interface, respectively. Finally, $D(\bbox{p})$ is the
1384: transparency coefficient of the boundary for the electron at the Fermi surface
1385: with the given direction of momentum.  Eq. (\ref{bound1})
1386: results in the current conservation at the boundary.
1387: 
1388: 
1389: Consider the Josephson current in a clean $SINI'S$ structure in
1390: the absence of the magnetic field. We assume that both $NS$ interfaces
1391: are specularly reflecting and are perpendicular to the
1392: $x$-axis. In this case the quasiclassical propagator depends on $p_x, x$ and
1393: $\omega_n$. Making use of the symmetry relations (\ref{sym}) the functions
1394: $g_{s,a}(p_x,x,\omega_n)=[g(p_x,x,\omega_n)\pm g(-p_x,x,\omega_n)]/2$ and
1395: similarly defined functions $f_{s,a},f^+_{s,a}$ can be parametrized as follows
1396: \begin{eqnarray}
1397: &g_s=b_{s1},\: f_s=b_{s2}-ib_{s3},\: f^+_{s}=b_{s2}+ib_{s3},& \label{b}\\
1398: &g_a=ib_{a1},\: f_a=b_{a3}+ib_{a2},\: f^+_{a}=-b_{a3}+ib_{a2},& \nonumber
1399: \end{eqnarray}
1400: The parameters $b_s, b_a$ are real and constitute two three-dimensional
1401: vectors
1402: $\bbox{b}_{s(a)}=(b_{s1(a1)},b_{s2(a2)},b_{s3(a3)})$. Combining
1403: the Eilenberger equations for $\hat g(\pm p_x,
1404: x,\omega_n)$, one easily finds
1405: \begin{equation} \frac{d \bbox{b}_s}{d x}=\bbox{M}{\bf
1406: \times}\bbox{b_a}, \quad \frac{d \bbox{b}_a}{d x}=-\bbox{M}{\bf
1407: \times}\bbox{b_s}. \label{main} \end{equation}
1408: Eqs. (\ref{main}) should be considered only for positive $p_x>0$ and $\omega_n>0$.
1409: The three-dimensional vector $\bbox{M}$ in (\ref{main}) is real and
1410: has the following components $M_1=2\omega/v_x$,
1411: $M_2=(\Delta+\Delta^*)/v_x$ and
1412: $M_3=i(\Delta-\Delta^*)/v_x$. Note that by introducing a complex
1413: vector $\bbox{z}=\bbox{b}_s+i\bbox{b}_a$ one can rewrite
1414: eqs. (\ref{main}) as $d
1415: \bbox{z}/dx=-i\bbox{M}\times \bbox{z}$. From the latter
1416: equation we conclude that $ \bbox{z}^2$ should be equal to a
1417: constant. From the normalization condition one finds $
1418: \bbox{z}^2=1$, or
1419: \begin{equation}
1420: \bbox{b}_s^2=1+\bbox{b}_a^2,\quad \bbox{b_s}\bbox{b_a}=0. \label{norm2}
1421: \end{equation}
1422: The boundary conditions (\ref{bound1}), (\ref{bound2}) take the form
1423: \begin{eqnarray}
1424: &&\bbox{b}_{a-}= \bbox{b}_{a+}=\bbox{b}_a, \label{zai2}\\
1425: &&\bigl[(\bbox{b}_{s+}-\bbox{b}_{s-})^2+
1426: (1-D)(\bbox{b}_{s+}+\bbox{b}_{s-})^2\bigr]\bbox{b}_a=2D \bbox{b}_{s+} \times
1427: \bbox{b}_{s-} \nonumber
1428: \end{eqnarray}
1429: 
1430: Assuming the pairing potential to be constant in the superconductors one
1431: easily recover the solution of the Eilenberger equations. For the left
1432: superconductor $x<-d/2$  we obtain
1433: \begin{equation} \bbox{b}_s=\bbox{e}_{M-}+\bbox{C_-}\exp(|\bbox M|x),\quad
1434: \bbox{b}_a = \bbox{b}_s \times \bbox{e}_{M-}. \label{ed}
1435: \end{equation}
1436: Here $\bbox{C_-}$ is an arbitrary vector perpendicular to $\bbox{M}$ and
1437: $\bbox{e}_{M-}$ is the unit vector in the direction of $\bbox{M}$
1438: \begin{equation}
1439: \bbox{e}_{M-}=\frac{1}{\sqrt{|\Delta|^2+\omega_n^2}}\left( \begin{array}{c}
1440: \omega_n\\ \Delta\cos(\chi/2)\\ \Delta\sin(\chi/2)
1441: \end{array}\right). \label{e}
1442: \end{equation}
1443: Analogously, for $x>d/2$ we have
1444: \begin{equation} \bbox{b}_s=\bbox{e}_{M+}+\bbox{C_+}\exp(-|\bbox M|x),\quad
1445: \bbox{b}_a = -\bbox{b}_s \times \bbox{e}_{M+},
1446: \end{equation}
1447: where vector $\bbox{e}_{M+}$ is given by eq.(\ref{e}) with the changed sign of
1448: $\chi$. Using these equations, from eq. (\ref{zai2}) one can establish the
1449: relation between $\bbox{b}_a$ and $\bbox{b}_s$ at the normal side of the
1450: interface near the left superconductor
1451: \begin{equation}
1452: \bbox{b}_a =t_1\bbox{b}_s\times\bbox{e}_{M-}.\label{bbb}
1453: \end{equation}
1454: Similarly, for the right boundary we get
1455: \begin{equation}
1456: \bbox{b}_a =t_2 \bbox{e}_{M+}\times \bbox{b}_s.\label{bbd}
1457: \end{equation}
1458: With the aid of these conditions one can easily find the Josephson current in
1459: $SINI'S$ junctions. What remains is to solve the Eilenberger equations
1460: (\ref{main}) in the normal metal. In the absence of external fields and
1461: impurities we have
1462: \begin{equation}
1463: \bbox{b}_s=\left( \begin{array}{c} C\\ L_+\cosh\tilde x+L_-\sinh\tilde x\\
1464: M_+\cosh\tilde x+M_-\sinh\tilde x \end{array}\right),
1465: \end{equation}
1466: \begin{equation}
1467: \bbox{b}_a=\left(
1468: \begin{array}{c} D\\ M_+\sinh\tilde x+M_-\cosh\tilde x\\ -L_+\sinh\tilde
1469: x-L_-\cosh\tilde x \end{array}\right),
1470: \end{equation}
1471: where $\tilde x=2\omega_n x/v_x$ and $C,D, L_\pm, M_\pm$ are constants
1472: determined from the normalization and boundary conditions. Finally, making use
1473: of eq. (\ref{Jeil}) we arrive at the result (\ref{JN}),(\ref{Q}).
1474: 
1475: 
1476: 
1477: 
1478: \begin{references}
1479: \bibitem{lam} C.J. Lambert and R. Raimondi, J. Phys. Cond. Mat. {\bf 10},
1480: 901 (1998).
1481: \bibitem{bel} W. Belzig, F.K. Wilhelm, C. Bruder, G. Sch\"on, and A.D. Zaikin,
1482:  Superlattices and Microstructures {\bf 25}, 1251 (1999).
1483: \bibitem{VZK} A.F. Volkov, A.V. Zaitsev, and T. Klapwijk, Physica C
1484: {\bf 210}, 21 (1993).
1485: \bibitem{HN} F.W.J. Hekking and Yu.V. Nazarov, Phys. Rev. Lett. {\bf
1486: 71}, 1625 (1993); Phys. Rev. B {\bf 49}, 6847 (1994).
1487: \bibitem{B} C.W.J. Beenakker, B. Rejaei, and J.A. Melsen, Phys. Rev. Lett. {\bf
1488: 72}, 2470 (1994); J.A. Melsen and C.W.J. Beenakker, Physica B {\bf
1489: 203}, 219 (1994).
1490: \bibitem{Kas} A. Yu. Kasumov {\it et al.}, Science {\bf 281}, 540 (1998).
1491: \bibitem{Christ} C. Schoenenberger {\it et al.}, unpublished.
1492: \bibitem{Misha} M. Yu. Kupriyanov, A. Brinkman, A.A. Golubov,
1493: M. Siegel, and H. Rogalla, Physica C {\bf 326-327}, 16 (1999).
1494: \bibitem{Eil} G. Eilenberger, Z. Phys. {\bf 214}, 195 (1968).
1495: \bibitem{lar} A.I. Larkin and Yu.N. Ovchinnikov, in
1496: {\it Nonequilibrium Superconductivity}, edited by D.N. Lan\-genberg and
1497: A.I. Lar\-kin (North-Holland, Amsterdam, 1986).
1498: \bibitem{Albert} A. Schmid, in
1499: {\it Nonequilibrium Superconductivity}, edited by K.E. Gray
1500: (Plenum, New York, 1981).
1501: \bibitem{zait} A.V. Zaitsev, Sov. Phys. JETP {\bf 59}, 1015 (1984) [Zh.
1502: Eksp. Teor. Fiz. {\bf 86}, 1742 (1984)].
1503: \bibitem{j1} M. Ashida, S. Aoyama, J. Hara, and K. Nagai, Phys. Rev. B {\bf
1504: 40}, 8673 (1989); Y. Nagato, S. Higashitani, K. Yamada, and K. Nagai,
1505: J. Low Temp. Phys. {\bf 103}, 1 (1996).
1506: \bibitem{dG} P.G. de Gennes, {\it Superconductivity of metals and
1507: alloys} (Benjamin, 1966).
1508: \bibitem{PZ} A.D. Zaikin and S.V. Panyukov, in {\it Nonequilibrium
1509: Superconductivity}, edited by V.L. Ginzburg (Nova Science Publ., New York, 1988),
1510: p. 137.
1511: \bibitem{GZ} U. Gunsenheimer and A.D. Zaikin, Phys. Rev. B
1512: {\bf 50}, 6317 (1994).
1513: \bibitem{shel} A. Shelankov and M. Ozana, Phys. Rev. B {\bf 61}, 7077
1514: (2000).
1515: \bibitem{brink} A. Brinkman and A.A. Golubov,  Phys. Rev. B {\bf 61}, 11297
1516: (2000).
1517: \bibitem{AGD} A.A. Abrikosov, L.P. Gor'kov, and I.Ye. Dzyaloshinkski.
1518: {\it Quantum Field Theoretical Methods in Statistical Physics}
1519: (Second Edition, Pergamon, Oxford, 1965).
1520: \bibitem{LL} L.D. Landau and E.M. Lifshtitz, {\it Quantum Mechanics}
1521: (Pergamon, Oxford, 1962).
1522: \bibitem{Hab} W. Haberkorn, H. Knauer, and J. Richter,
1523: Phys. Stat. Solidi (A) {\bf 47}, K161 (1978).
1524: \bibitem{KO} I.O. Kulik and A.N. Omel'yanchuk, Fiz. Nizk. Temp. {\bf
1525: 4}, 296 (1978) [Sov. J. Low Temp. Phys. {\bf 4}, 142 (1978)].
1526: \bibitem{glazman} L.I. Glazman and K.A. Matveev, Zh. Eskp.
1527: Teor. Fiz. Pis'ma Red. {\bf 49}, 570 (1989) [JETP Lett. {\bf 49}, 659
1528: (1989)].
1529: \bibitem{Ben} C.W.J. Beenakker, Phys. Rev. Lett. {\bf 67}, 3836 (1991).
1530: \bibitem{Ishii} C. Ishii, Progr. Theor. Phys. {\bf 44}, 1525 (1970).
1531: \bibitem{Kulik} I.O. Kulik, Zh. Eksp. Teor. Fiz. {\bf 57}, 1745 (1969)
1532: [Sov. Phys. JETP {\bf 30}, 944 (1970)].
1533: \bibitem{ZZh} A.D. Zaikin and G. F. Zharkov, Zh. Eksp. Teor. Fiz. {\bf
1534: 78}, 721 (1980) [Sov. Phys. JETP {\bf 51}, 364 (1980)].
1535: \bibitem{svi} A.A. Svidzinskii, {\it Spatially nonhomogeneous problems in the
1536: theory of superconductivity} (Nauka, Moscow, 1982).
1537: \bibitem{shel2} M. Ozana and A. Shelankov, Phys. Rev. B {\bf 65}, 014510
1538: (2002).
1539: 
1540: 
1541: 
1542: \end{references}
1543: \end{multicols}
1544: \end{document}
1545: