1: \documentclass[aps,amsmath,prb,a4paper,showpacs]{revtex4}
2: \usepackage{graphicx}
3: %\usepackage{amsmath}
4: \newcommand{\bfm}[1]{\mbox{\boldmath${#1}$}}
5: %\usepackage[english]{babel}
6: \newcommand{\eqn}[1]{(\ref{#1})}
7: \newcommand{\UU}{\breve{{\cal{U}}}}
8: \newcommand{\GG}{\breve{G}}
9: \newcommand{\ovr}{\overline{R}}
10: \newcommand{\beq}{\begin{equation}}
11: \newcommand{\eneq}{\end{equation}}
12: \newcommand{\bea}{\begin{eqnarray}}
13: \newcommand{\enea}{\end{eqnarray}}
14: \newcommand{\bean}{\begin{eqnarray*}}
15: \newcommand{\eean}{\end{eqnarray*}}
16: \newcommand{\nn}{\nonumber}
17: \newcommand{\met}{\frac{1}{2}}
18: \newcommand{\drond}[1]{\frac{\partial}{\partial #1} }
19: \newcommand{\EE}{\mathbf{E}}
20: \newcommand{\pp}{\mathbf{p}}
21: \newcommand{\sig}{\mathbf{\sigma}}
22: \newcommand{\rr}{\mathbf{r}}
23: \newcommand{\CRE}[1]{C^{\dagger}_{#1}}
24: \newcommand{\DES}[1]{C_{#1}}
25: \newcommand{\LAG}[3]{L_{#1}^{#2}\left(#3 \right)}
26: \newcommand{\COE}{\mathcal{C}}
27: \newcommand{\h}{\hbar}
28: \newcommand{\bd}{\begin{displaymath}}
29: \newcommand{\ed}{\end{displaymath}}
30: \newcommand{\de}{\delta}
31: \newcommand{\f}{\varphi}
32: \newcommand{\ww}{\omega}
33: \newcommand{\DD}{\Delta}
34: \newcommand{\ug}{\underline{\hat{g}}}
35: \begin{document}
36: % The correct dates will be entered by Springer
37: %
38: \title{Coherent response of a low $T_c$ Josephson junction to an
39: ultrafast laser pulse}
40: \author { P.Lucignano $^{1,2}$}
41: \author{ G. Rotoli $^{1,3}$}
42: \author{ E.Santamato $^{1,2}$}
43: \author{A. Tagliacozzo $^{1,2}$}
44: \affiliation{$^1${\sl Coherentia} - Istituto Nazionale di Fisica della Materia
45: (INFM), Unit\'a di Napoli}
46: \affiliation{ $^2$ Dipartimento di Scienze Fisiche
47: Universit\`a di Napoli "Federico II ", Monte S.Angelo - via
48: Cintia, 80126 Napoli, Italy}
49: \affiliation{$^3$ Dipartimento di Energetica,
50: Universit\`a di L'Aquila, Localit\`a Monteluco, 67040 L'Aquila,
51: Italy}
52: %
53: \date{\today}
54: % The correct dates will be entered by Springer
55: %
56: %\mail{procolo@na.infn.it}
57: \begin{abstract}
58: By irradiating with a single ultrafast laser pulse a superconducting
59: electrode of a Josephson junction it is possible to drive
60: the quasiparticles (qp's) distribution strongly out of equilibrium. The
61: behavior of the Josephson device can, thus, be modified on a fast time
62: scale, shorter than the qp's relaxation time. This could be very
63: useful, in that it allows fast control of Josephson charge qubits and,
64: in general, of all Josephson devices.
65: If the energy released to the top layer contact $S1$ of the junction
66: is of the order of $\sim
67: \mu J$, the coherence is not degradated, because the perturbation is very fast.
68: Within the framework of the quasiclassical Keldysh Green's function
69: theory, we find that the order parameter of $S1$ decreases.
70: We study the perturbed dynamics of
71: the junction, when the current bias is close to the critical current,
72: by integrating numerically its classical equation of
73: motion. The optical
74: ultrafast pulse can produce switchings of the junction from
75: the Josephson state to the voltage state.
76: The switches can be controlled by tuning the laser
77: light intensity and the pulse duration of the Josephson junction.
78: \end{abstract}
79: %
80: \pacs{{74.50.+r}{}, {74.40.+k}{},{74.25.Gz}{}, {74.25.Fy}{}}
81: % end of PACS codes
82:
83: %end of abstract
84: %\titlerunning{Coherent response of a low $T_c $ Josphson junction ...}
85: %\authorrunning{P.Lucignano , {\sl et al. }}
86: %
87: \maketitle
88: %
89: %\onecolumn
90: \section{Introduction}
91: The characteristic frequency in the dynamics of a Josephson junction
92: (JJ) is the so-called Josephson plasma frequency $\omega _{pJ}$
93: (e.g.$10 \div 100 GHz$). Coupling of a JJ to a microwave field
94: leads to the well known lock-in conditions, which show up as Shapiro
95: steps in the I/V characteristic. On the other hand, photo-response to
96: radiation in a superconductor induces heat relaxation (bolometric
97: effect \cite{barone}) and non equilibrium generation of quasiparticles
98: (qp) \cite{testardi,parker}. Both phenomena are extensively studied
99: since they are relevant for the fabrication of fast and sensitive
100: detectors. The models used are phenomenological
101: \cite{owen,ivlev,ser,lindg,semenov}, mainly involving different
102: temperatures associated to separate distributions of electrons and
103: phonons out of equilibrium.\\
104: \begin{figure}[h]
105: % \hspace{-0.5cm}
106: \includegraphics[width=\columnwidth ]{gj.eps} \caption{Sketch
107: of the Josephson junction exposed to a laser radiation pulse.}
108: \label{fig0}
109: \end{figure}
110: Recently, laser light with pulses of femtosecond duration $ \tau _c
111: \in (10^{-14}s,10^{-13}s )$ has become available, as a source to test
112: the photo-response of a JJ \cite{lindg}. Ultrafast pulses can be
113: extremely useful, in that they allow studying an unexplored regime in
114: non equilibrium superconductivity. Indeed, photon absorption, by
115: creating electron-hole (e-h) pairs at very high energies, drives the
116: quasi-particle (qp) energy distribution out of equilibrium during the
117: time $\tau_c$. The qp non-equilibrium distribution depends on the
118: energy relaxation time parameter $\tau _E $, defined as the time by
119: which a `hot' electron is thermalized by repeated scatterings with
120: other electrons or phonons. The process involves generation of many
121: qp's during energy degradation, until the system relaxes back to the
122: equilibrium distribution function $ n_o (\omega )$. This time scale is
123: determined by the electron-electron interaction time $\tau_{e-e}$ and
124: the electron-phonon interaction time $\tau_{e-ph}$, which are strongly
125: material dependent \cite{kaplan}, ranging from $4\cdot10^{-7} s$ for
126: $Al$ to $1.5 10^{-10}s$ for $Nb$. In this work we analyze the
127: possibility that, keeping temperature quite low, ultrashort laser
128: radiation induces direct switches out of the Josephson conduction
129: state at zero voltage, due to coherent reduction of the critical
130: current $J_c$.
131:
132: There are many reasons for the switching from the zero to the
133: resistive state in a Josephson junction. Among these, thermal escape
134: \cite{thermal}, quantum tunneling \cite{quantum}, latching logic
135: circuits \cite{latching} and pulsed assisted escape\cite{pulse}. A
136: clear cut discrimination between different mechanisms can be difficult
137: to achieve. In our case quantum escape is ruled out because the
138: temperature is not expected to be low enough. Also, we assume that there
139: is no external circuit to induce switching and re-set of the zero voltage
140: state as in latching logic elements.
141:
142: Pulsed assisted escape is a generic term for a large class of
143: phenomena including in principle bolometric heating of the junction
144: which is re-set in relatively slow times \cite{testardi}. Production
145: of quasiparticles generated by $X$-ray radiation has been studied up
146: to recently\cite{barone2,ovchinnikov}. A cascade follows, which
147: increases the number of excitations and lowers their energy down to
148: the typical phonon energy $\omega _D$ in a duration time, which is of
149: the order of the nanoseconds. Subsequently, qp's decay by heating the
150: sample. However, the power of the laser can be reduced enough and
151: both the substrate and the geometry can be chosen such that the energy
152: released by the radiation on the junction can be small. On the other
153: hand, appropriate experimental conditions can make the time interval
154: between two pulses long enough, so that the bolometric response is
155: negligible.
156:
157: Generally speaking, junctions are more sensitive to pulses especially
158: when their harmonic content is close to $\omega_{pJ}$, but this is not
159: our case. In fact the laser carrier frequency ($\Omega \sim 100 THz$)
160: is quite high compared to $\omega _{pJ}$ and we consider the case
161: $\Omega >> \tau _c^{-1} > \omega_{pJ} > \tau _E^{-1}$, what implies that little
162: relaxation takes place during the duration of the pulse. $\tau _c $
163: should also be shorter than the pair-breaking time $\h/\Delta_o \sim 1
164: \div 5 ps$. Here $\Delta_o $ is the unperturbed gap parameter.
165: Our approach assumes that, on a time scale intermediate between the
166: pulse duration and the relaxation time $\sim \tau _{E}$, the order
167: parameter of the irradiated superconductor is sensitive to the
168: non-equilibrium qp distribution, which modulates it coherently till it
169: switches out of the zero voltage state.
170:
171: To analyze the dynamics of the order parameter and the way how the
172: latter affects the Josephson current, we adopt a non-equilibrium
173: formalism based on quasiclassical Green's functions \cite{rammer,belzig}.
174: The quasiclassical approach has been mostly used in the past in
175: connection with the proximity effect \cite{belzig}, as well as with
176: non-equilibrium due to other space inhomogeneity
177: conditions\cite{amin}. As far as we know, this is the first time that
178: its extension to non-equilibrium in time is applied to a coherent
179: response after an ultrafast laser irradiation.
180:
181: The quasiclassical approximation to the Gorkov equations, is obtained
182: by averaging over the period of the optical frequency $\Omega $, which
183: is a fast time scale\cite{noi}. Our equations include the physics of the
184: cascade process, which occurs when one focuses on the kinetics of the
185: qp diffusion. A kinetic equation approach to the steady state
186: non equilibrium qp distribution, including phonon scattering has
187: been developed in ref.\cite{scalapino}, for light irradiation,
188: mostly in the microwave range. The cascade regime is extensively
189: discussed in Ref.\cite{ovchinnikov}, however it will not be specifically
190: addressed here.
191:
192: Instead, if the switching of the irradiated superconductor due to the
193: ultrafast pulse takes place prior to the occurrence of qp relaxation,
194: an approximate solution of the dynamical equations can be derived,
195: which describes an istantaneous response of the order parameter.
196:
197: We take a low $T_c$ JJ with an $s$-wave order parameter as the reference
198: case (e.g. a high quality $Nb$ or $Al$ junction) and $T<<
199: T_c$. The optical penetration depth of the laser light $
200: \lambda_\delta$ in the topmost superconductor exposed to radiation
201: $S1$ is assumed to be shorter than its thickness, so that any
202: modification induced by the radiation field only involves $S1$ itself
203: \cite{testardi} (see. Fig.(\ref{fig0})). In a small size JJ the
204: spatial variation of the order parameter along the lateral dimension
205: of $S1$ is not taken into account, except when the qp diffusion
206: process cannot be ignored.
207:
208: We consider just one pulse of given duration $\tau _c$ which releases
209: the energy ${\cal{E}}$ per pulse, by exciting $e-h$ pairs and by
210: creating a non equilibrium distribution of qp's. A related
211: dimensionless quantity $q$, as defined in Eq. (\ref{qq}), parametrizes
212: the strength of the perturbation due to the radiation. The
213: perturbation is assumed to be small so that only the lowest order in
214: the expansion in $q$ is retained. This allows us to derive a
215: temporary reduction of the order parameter $\Delta $ induced by the
216: pulse, as shown in Fig.(\ref{fig1}). We do not give a detailed
217: description for the relaxation of the non-thermal qp distribution in
218: the irradiated superconductor. The self-energy terms corresponding to
219: this process require further analysis. According to the Eliashberg
220: formulation \cite{kaplan} these terms affect the quasiparticle
221: amplitude $Z(\omega)$ introducing changes in the phonon distribution
222: and retardation in the response. Nevertheless, we expect that these
223: self-energy terms become effective only on a longer time scale after
224: the laser pulse. Our equations pinpoint a non retarded evolution of the
225: order parameter prior to relaxation, which implies a reduction of the
226: critical current. This shows that the coherent modulation of the gap
227: parameter can produce switching of the junction out of the Josephson
228: state.
229:
230: The switches are studied numerically by solving the classical equation
231: fo motion of a current biased JJ with current $J$ close to the
232: critical curent $J_c$, during the excitation process. After the
233: switching the dissipation in not treated selfconsistently: a standard
234: dissipation, typical of thermal equilibrium, is assumed in the JJ
235: dynamics, by adding a conductance term in the numerical simulation.
236: We stress that the assumed model for dissipation determines the actual
237: qp branch of the I-V characteristics, but does not affect
238: substantially the switching probability. The switching from the
239: Josephson state to the voltage state in the parameter space
240: $(q,\tau_c,J/J_c)$ is reported in Fig.\ref{fig5}. Interestingly, we
241: find that for fixed value of $q$ and $J/J_c$ there is an optimum pulse
242: duration to achieve the strongest sensitivity of the junction to the
243: switching process. We show that this is due to the way how the non
244: equilibrium qp's distribution affects the pairing in $S1$. The paper
245: is organized as follows.
246:
247: In Section 2 we calculate the non equilibrium qp distributions
248: immediately after the pulse, prior to relaxation.
249:
250: In Section 3 we introduce the time dependent quasiclassical Keldysh
251: Green functions formalism extended to the time domain, using the
252: general frame given in appendix A. We calculate the correction to the
253: single particle
254: propagators to first order in $q$. A correction to the gap
255: and to the Josephson critical current follows due to the laser pulse.
256:
257: In Section 4 the dynamics of the JJ after the pulse is simulated
258: numerically. Finally, a summary of our results is given in Section 5.
259:
260: Appendix A collects the formulae of the quasiclassical approximation
261: in the non equilibrium Keldysh theory, which are used in the core of
262: the paper. In Appendix B we derive the kinetic equation for the non
263: equilibrium distribution function which drives the relaxation of the
264: system. The equations of motion for the quasiclassical retarded Green
265: function is reported in appendix C, while the equation of motion for the
266: advanced and Keldysh Green function can be derived in the same way .
267:
268:
269: \section{The non-equilibrium qp distribution}
270:
271: \subsection{Non-equilibrium electron-hole pair excitations induced by
272: optical irradiation}
273:
274: The optical frequencies ($\Omega \sim 100 THz$) building the
275: wavepacket of the laser pulse excite $e-h$ pairs at high energies. As
276: explained in the introduction the non-equilibrium arising from the
277: alteration of the qp distribution has a relaxation time $\tau _E$
278: which is long compared to the optical period: $\Omega \tau _E >>1 $.
279: In addition to this, the duration of the pulse $\tau_c\sim \omega _c
280: ^{-1} $ is even shorter than the pair breaking time, so that we expect
281: that, in our case, dissipative phenomena do not affect the coherence of the
282: superconductor on the time scale $\tau_c$.
283:
284: Qp's are generated as if the metal were normal, because
285: superconductivity doesn't play any role in their excitation at large
286: energies. They propagate according to something very much like the
287: free particle time-ordered Green's function $g_o $ (from now on we put
288: $\hbar =1 $):
289: \beq
290: g_o^T(k ;t-t' ) \equiv
291: -i \langle T[c_k(t) c_k^\dagger (t') ] \rangle = -i e^{-i\xi (t-t') }
292: \left [ ( 1- n_k ) \theta (t- t' ) - n_k \theta (- t+ t' ) \right ]\:.
293: \label{green}
294: \eneq
295: Here $\xi $ is the qp energy with momentum $k$ and is measured from
296: the chemical potential $\mu$. $n_k$ is the qp distribution function.
297: We assume that $e-h$ symmetry is conserved in the excitation process, so
298: that $\mu$ is not altered with respect to its equilibrium value.
299:
300: The equation of motion for the Green's function $\tilde g$ in the
301: presence of the radiation field is:
302: \beq
303: \left (i \frac{\partial}{\partial t} - \frac{1}{2m}[\vec \nabla _r -
304: \frac{e}{c}
305: \vec A(\vec r,\vec R , t)]^2 + \mu \right )\:
306: \tilde g (\vec r, \vec R ,t,t')=
307: \delta(r) \delta(t-t')
308: \label{bg1}\; ,
309: \eneq
310: $\vec r$ is the relative space coordinate, while $\vec R$ is the center of
311: mass coordinate.
312: The vector potential is a wave-packet centered at frequency $\Omega$
313: according to:
314: \beq
315: \vec{A}(\vec r, \vec{R},t)=\sum_{\pm}\sum _{\vec p} \vec{a}_\pm
316: (\vec p, \vec{R},t)e^{\mp i ( \vec {p}\cdot \vec{r} - \Omega t)} \; .
317: \label{potvet}
318: \eneq
319: Here $\vec{a}_{\pm}$ are slowly varying 'envelope' functions of $\vec R$
320: on the
321: size of the irradiated spot and on the time scale $\Omega^{-1}$.
322: We look for solutions of eq.(\ref{bg1}) in the form:
323: \beq
324: \tilde g (\vec{r}, \vec{R};t,t') = g(\vec{r}, \vec{R};t,t')
325: + \sum_{\pm}\sum _{\vec{p}}g ^{\pm}(\vec{p}, \vec{R},t,t')
326: e^{\mp i ( \vec {p}\cdot \vec{r} - \Omega t)}\;\; ,
327: \label{ciccia}
328: \eneq
329: where $g $ and $g^{\pm}$ are slowly varying functions of $
330: \vec{R}$ and $t$ on the same scales. A similar expansion can be
331: done w.r.to the variable $t'$. Following Eq. (\ref{ciccia}), a
332: decomposition of eq.(\ref{bg1}) into harmonics arises\cite{noi}. We define the
333: zero order harmonic equation as the one that does not contain
334: exponentials $e^{\pm i \Omega t}$. By averaging over a period $\Omega
335: ^{-1} $ we neglect harmonics of order two, or higher. This amounts to
336: include one photon excitation processes only, with released energy
337: ${\cal{E}}$. Some extra details can be found in Appendix A:
338: \beq
339: \left (i \frac{\partial}{\partial t} + \frac{1}{2m} \nabla ^2_r +
340: \frac{e^2}{mc^2} \sum _{\vec {p}',\vec{p}''} \vec{a}_+
341: (\vec {p}', \vec{R},t) \vec{a}_-
342: (\vec {p}'', \vec{R},t)e^{ i ( \vec {p}'-\vec{p}'') \cdot \vec{r} }\:
343: +\mu \right )
344: g (\vec r, \vec R ,t,t')=
345: \delta(r) \delta(t-t')
346: \nonumber\; .
347: \eneq
348:
349: Fourier transforming w.r.to $\vec r$ ( $\vec r \to \vec p \to \xi $)
350: we have:
351: \beq
352: \left (i \frac{\partial}{\partial t} - \left (\frac{p^2}{2m}
353: -\mu \right ) \right ) g (\vec {p}, \vec R ,t,t') +
354: \frac{e^2}{mc^2} \sum _{\vec {p}',\vec{p}''} \vec{a}_+
355: (\vec {p}', \vec{R},t) \vec{a}_-
356: (\vec {p}'', \vec{R},t)\:
357: g (\vec {p}'-\vec {p}''-\vec {p}, \vec R ,t,t')= \delta(t-t')
358: \label{bg21}\;.
359: \eneq
360: The radiation field generates and annihilates high energy
361: $e - h$ pairs. Hence we assume that the forcing term conserves the
362: total impulse, $ \vec {p}'+\vec {p}'' \sim 0 $, but
363: $\vec {p}'-\vec {p}''$ transfers an energy $2\xi $ to the electrons.
364: Therefore we take the coupling term in the Hamiltonian as:
365: \beq
366: \frac{e^2}{m c^2 } \vec{a}_+
367: (\vec {p}', \vec{R},t) \vec{a}_-
368: (\vec {p}'', \vec{R},t ) \to
369: q \:
370: \frac{\ww _c }{\sqrt{\pi}} e^{-\met \omega _c^2 t^2} \: \delta
371: (2 \xi + \xi ' ) \:\delta ( \vec {p}'+\vec {p}'' ) \: .
372: \label{defq}
373: \eneq
374: A Gaussian shaped time dependence has been chosen for the
375: pulse with half-width $\omega _c^{-1}$, while the space dependence
376: has been neglected for simplicity. In Eq. (\ref{defq}) the dimensionless
377: quantity appears:
378: \beq
379: q\sim\frac{e^2}{c }\: \frac
380: {\omega_D }{\Omega } \: \; \frac {{\cal{E}}} {m \Omega^2 R^2_o }\:,
381: \label{qq}
382: \eneq
383: Where $R_o$ is the laser spot (see Eq.\ref{brown}).
384: Here the number of excited $e-h$ pairs is
385: $\sim\Omega/\omega_D$, with $\ww_D$ the Debye energy.
386: Experiments \cite{testardi,lindg} show that the energy released by the
387: pulse can be very low, so that we will always expand in $q$. In fact,
388: while in the case of an $ rf $ radiation $q \sim 1 $, in the case of a
389: femtosecond laser pulse $q\sim 0.01\div 0.1 $, corresponding to
390: a fraction of $\mu J $ released per pulse on the superconducting
391: surface of $ \sim 100 \mu m ^2$.
392:
393: The zero
394: order harmonic equation, Eq. (\ref{bg21}), becomes :
395: \beq
396: \left (i \partial _t -\xi \right ) g (\xi; t,t' ) +
397: q \frac{\omega _c}{\sqrt{\pi}}
398: e^{-\met \ww _c^2 t^2 } g (- \xi; t,t' ) = \delta (t-t')
399: \label{bg22})\;.
400: \eneq
401:
402: To derive the non-equilibrium correction to the qp distribution
403: function, the kinetic equation \ref{ovA2} should be solved.
404: In place of this we proceed in this work in an heuristic way.
405: Our approach lacks mathematical rigour, but singles
406: out directly the role of the laser induced $e-h$ excitations
407: at frequencies $( \Omega -\omega _c , \Omega + \omega _c )$.
408: Our result is valid
409: in the limit of large $\xi $'s and zero temperature, before
410: relaxation takes place.
411:
412: We solve Eq(\ref{bg22}) for the retarded Green's function for $t>0$ and
413: $-t' \sim 0^+$ by truncating the Dyson equation to lowest order in $q$:
414: \bea
415: g^R (\xi; t,t') = g_o^R (\xi; t-t') - q \frac{\omega _c}{\sqrt{\pi}}
416: \int dt'' g_o^R(\xi; t,t'') e^{-\met \ww _c^2 {t''}^2 } g_o^R (- \xi;
417: t''-t' ) \nonumber\\ = g_o^R (\xi; t-t') +i q \frac{\omega
418: _c}{\sqrt{\pi}}
419: \int _{-\infty}^{+\infty} \frac{d\omega}{2\pi}
420: \frac{e^{-i \omega t}}{\omega- \xi +i0^+} \int _{0^+}^{+\infty}
421: dt \: e^{i (\ww - \xi ) t }
422: e^{-\met \ww _c^2 t^2 }\label{sopra}\;,
423: \enea
424: where $g_o^R (\ww ) = \{ \ww -\xi + i 0^+ \}^{-1} $ is the Fourier
425: transform of the retarded Green's function. The time integral can be
426: expressed in terms of the function $ w[z] \equiv e^{-z^2} \: erfc\:
427: (-iz) $. If we now approximate Eq. (\ref{sopra}) by evaluating $w[z]$
428: only at the pole and use the the integral representation of the step
429: function:
430: \beq
431: \theta (t)= \frac{e^{i z t}}{2 \pi i} \int _{-\infty}^{+\infty} d\ww
432: \frac{e^{- i \omega t}}{\omega-z} \;\;\; for \;\; \Im m z < 0\:\:\: ,
433: \eneq
434: the correction $ \delta g^R ( \xi , t)$ to $g^R$ for $\xi >> 0 $,
435: which includes the non equilibrium qp 's distribution,
436: is: \beq \delta g^R ( \xi , t, 0^-)= - \frac{q}{\sqrt {2}} e^{-i (\xi
437: -i0^+) t}\: \: w\left [- 2\xi /(\sqrt {2}\ww _c ) \right ] \:\: \theta
438: (t ) ,
439: \label{corrg}
440: \eneq
441: From eq.(\ref{corrg}) we obtain the time
442: ordered Green's function for $t>0 ,\: t' = 0^- $ and $\xi >0$:
443: \beq
444: g ( \xi , t>0 , 0^- )= (-i ) e^{- i (\xi -i0^+) t}\: \theta (t )\:
445: \left ( 1 - \frac{q}{\sqrt{2}} \:
446: \: \rho\left [ 2\xi /(\sqrt {2}\ww _c ) \right ] \right ) ,
447: \label{gcorr}
448: \eneq
449: with
450: \beq
451: \rho [ x ] \equiv e^{- x^2 } \frac{2 }{\sqrt \pi } \int _0^ { x } ds
452: \: e^{ s^2 } \:\: .
453: \eneq
454: $ \rho [ x ] $ increases linearly with $x$ and it decreases slowly, as
455: $1/x$, at large arguments. In Eq. (\ref{gcorr}) we have neglected $\Re
456: e [w] $ because $|\xi | >> 0 $.
457:
458: Eq. (\ref{gcorr}) is to be compared with the free propagating time
459: ordered Green's function of eq.(\ref{green}) for the same time
460: arguments. Comparison yields the amount by which the distribution
461: function is driven out of the equilibrium:
462: \beq
463: \delta n (\xi ) \approx \frac{q}{ \sqrt {2}}
464: \: \rho \left [ 2\xi /(\sqrt {2}\ww _c ) \right ] \;\;\; for \;\; |
465: \xi | >> \DD _o \:\:\: .
466: \label{dnx}
467: \eneq
468: Note that the expression of Eq. (\ref{dnx}) changes sign according to $
469: sign\; \{\xi \}$. This stems from the assumed $e-h$ symmetry. In turn
470: this implies that no charge imbalance occurs.
471:
472: Eq. (\ref{dnx}) can be considered as the non equilibrium distribution for
473: qp 's starting at the time of the pulse $t\sim 0^+ $.
474:
475: In the absence of relaxation, a change in the available qp density of
476: states follows. Because $(g^R(\xi ,\ww ) )^* = g^A(\xi ,\ww ) $ if
477: $\ww$ is real, the correction to the density of state $\delta \nu
478: (\xi)$ is:
479: \beq
480: \delta \nu (\xi) = -\frac{1}{\pi} \left (
481: \delta g^R - \delta g^A \right ) ( \xi , 0^+) \approx
482: \frac{q}{\pi \sqrt {2}}
483: \: \rho\left [ 2\xi /(\sqrt {2}\ww _c ) \right ]
484: \;\;\; for \;\; \xi > 0\:\:\: .
485: \label{dens}
486: \eneq
487: The first stages of the relaxation process involve the inelastic
488: diffusion of qp's in the medium which is qualitatively discussed in
489: the next section.
490:
491: \subsection{Inelastic diffusion of the qp's at initial times }
492: Let us discuss shortly what was neglected in the derivation of the
493: change in the equilibrium qp distribution $ \delta n(\xi ) $ given by
494: Eq.(\ref{dnx}).
495:
496: The single particle Green's function $g (p,R,t,t' ) $ is assumed to be
497: a slowly varying function of $ (t+t' )/2 = \overline t $ and a fast
498: varying function of $t -t'$. Fourier transforming w.r.to the latter
499: variable ( see Appendix A ) there is an $\omega $ dependence even in
500: the stationary case ( i.e. with no $\overline t$ dependence). This
501: $\omega $ dependence is determined by the frequency dependent
502: Eliashberg $e-ph $ coupling $\alpha ^2 F(\omega ) $ and is contained
503: in the $e-ph $ self-energy $M_{e-ph} (\omega ) $ \cite{grimvall}.
504: Accordingly, the complex qp renormalization parameter $ Z_n (\omega )
505: $ is defined by $[1- Z_n (\omega ) ] \omega = M_{e-ph} (\omega ) $.
506: In our derivation we have not included the self-energy, so that we are
507: implicitly taking $ Z_n (\omega ) \to 1 $, what applies for large
508: $\omega ( \sim \Omega ) $, prior to relaxation.
509:
510: Moreover, because the system is in the superconducting state, we
511: should have dealt with the corresponding superconducting parameter $
512: Z_s (\omega )$. The latter is derived
513: together with the complex gap parameter $\Delta ( \omega ,
514: \overline t ) $ with $\Delta ( \Delta_o , \overline t=0^- )=\Delta_o $
515: from the coupled Eliashberg equations ( we drop the overline on $t$ in
516: the following ).
517:
518: The procedure of averaging over the fast time scale $\Omega ^{-1}$
519: singles out two frequency components of $ \Delta ( \omega , t ) $ and
520: $ Z_s (
521: \omega , t ) $: $\omega = \Delta _o$ and $\omega=\Omega $ as a consequence retardation arises from frequencies up to $ 10\omega_D$ is neglected.
522: $\Omega$ is so large that $ Z_s $ and $ Z_n $ do not differ sizeably.
523: In fact, their real parts differ by a quantity of the order of $
524: (\omega _D \Delta _o /\Omega ^2 )^2 \ln (\omega _D / \Delta _o ) $. $
525: \Delta ( \Omega , t ) $ itself is expected to be so small that it can
526: be neglected alltogether. Indeed, in connection with eq. \ref{kcomp}
527: of Appendix A we do not discuss the self-energy terms. Of course this
528: approximation breaks down on the time scale of $e-ph $ relaxation.
529:
530: Let us now discuss the $t$ dependence. The
531: equation of motion for the qp distribution function $n(\vec{R},t)$ is
532: derived in Appendix B, where we take $n_T =0$, because we neglect
533: charge imbalance corrections.
534:
535: In averaging over a few optical periods the kinetic equation for
536: $\delta n _L $, the electric field $\vec E$ averages to zero. The qp
537: relaxation is governed by the collision integral $I[n(\vec{R},t);t]$
538: which describes the inelastic processes. In Ref.\cite{ovchinnikov} the
539: cascade of the $e-h$ excitations due to inelastic scattering is studied
540: in detail. Two stages occur. In the first stage e-e interactions
541: multiply the number of excited qp's in the energy range from $\Omega$
542: to $\omega _D $ which is taken as the cutoff energy of the pairing
543: interaction. This happens in a time interval short w.r.to the pulse
544: duration ($\sim 10 ^{-14} s $). In the second stage a much slower
545: relaxation process takes place, by which the energy of the qp's
546: reaches $\DD $. This process involves electron-phonon scattering on a
547: time scale $ \hbar \omega_D^2/\DD ^3
548: \sim 10^2 ns$ which is much larger than any time scale in our problem.
549: Here we will leave this stage aside. In the time interval we are
550: concerned with, we have little relaxation and the energies involved
551: are $\omega >> \DD $.
552:
553: According to Eq. (\ref{funh}) the distribution function prior to
554: relaxation deviates from the equilibrium value by the quantity $\delta
555: n_L = -2 \delta n (\xi ) $ given by Eq. (\ref{dnx}). There is no
556: explicit dependence on $\omega$ in our correction, becauise
557: retardation is nelgected. Still qp's diffuse in space inside the
558: junction over a characteristic distance $R_o \sim \sqrt{ D
559: \tau _{e-e} } $, where $D$ is the diffusion coefficient. Hence
560: \beq
561: \delta n_{L}(\xi, \vec{R},t)=
562: - 2 \frac{R_o^2}{R_o^2+ D t} \:
563: \delta n (\xi )\:
564: e^{\left \{-\frac{R^2}{R_o^2+\frac{D} t }\right\}} \:\:\: .
565: \label{brown}
566: \eneq
567: For relatively large times $ \tau_{e-ph}>>t >> R_o^2/D $ we will
568: ignore the spatial dependence by putting $R = 0$. This is the first
569: step of a perturbative analysis of the non equilibrium distribution
570: functions.
571:
572:
573: \section{Changes of the superconductive
574: properties on the time scale $\ww _c ^{-1} $}
575:
576: \subsection{The correction to the gap parameter}
577:
578: In this Subsection we derive the Keldysh Green's function in the
579: presence both of a time dependent gap $\DD (t)$ and of a
580: non-equilibrium qp distribution as given by Eq. (\ref{brown}). We
581: assume weak coupling superconductivity and we neglect here the
582: frequency dependence of the $e-ph$ coupling parameter $\lambda $. This
583: follows from the neglecting the retardation effects mentioned in
584: Sec. IIb on time scales much faster than the $e-ph $ relaxation
585: time. From the Keldysh Green's function $\hat g ^K$ (where the hat
586: denotes matrix representation in the Nambu space, see appendix A) we
587: recalculate the gap self-consistently, according to the formula:
588: \beq
589: \DD(t)=- \frac{\nu(0) \lambda}{4} \int_{-\infty}^{\infty}
590: <f^K({\vec p/|\vec p|},\omega,t)>_{\vec v_F} \;
591: d \omega \;.
592: \label{gap}
593: \eneq
594: The average over the direction of the momenta on the Fermi surface is
595: indicated. The Keldysh Green's function in thermal equilibrium is:
596: \beq
597: \hat g_o^K =\tanh \frac{\beta \ww }{2} \left(\hat g^R - \hat g^A\right)\:.
598: \label{gok}
599: \eneq
600: Out of equilibrium we use the definition:
601: \beq
602: \hat g^K = \hat g^R \hat h - \hat h \hat g^A \:\: .
603: \label{ggk}
604: \eneq
605: However, $ \hat h $ defined in appendix B is here diagonal, because we
606: assume that no charge imbalance arises. Hence, up to first order in $q$,
607: \beq
608: \delta \hat g ^K \approx n^0_L(\hat g^R_{ad}-\hat g^A_{ad})
609: +\delta n_L (\hat g_o^R- \hat g_o^A)\:.
610: \label{lingk2}
611: \eneq
612: Here $ n_L^o = \tanh (\beta \ww
613: /2) $ is the equilibrium distribution and $\hat g^R_{ad}-\hat
614: g^A_{ad}$ is introduced in Appendix C (see Eq.\ref{fg}) and is
615: discussed in the following.
616:
617: We now first derive the contribution coming from the second term of
618: Eq. (\ref{lingk2}). We start from the outset using Eq. (\ref{brown})
619: and performing the quasiclassical approximation. The latter involves
620: an energy integration:
621: \beq
622: \delta n_L \cdot \left ( \hat {g}_o^R - \hat {g}_o^A \right )
623: \equiv
624: \frac{i}{\pi}\int _{-\infty}^{+\infty}
625: d\;\xi \: \delta n_L (\xi ,t) \cdot \left ( \hat {g}_o^R (\xi, \omega
626: ,t) - \hat {g}_o^A (\xi, \omega ,t)\right )\:.
627: \label{qcl}
628: \eneq
629:
630: Using the equilibrium BCS functional forms, the Green's functions
631: appearing on the diagonal of $\hat{g}^{A/R}$ are \cite{gork}:
632: \bea
633: g^R(\xi , \omega)=\frac{u^2_{\xi}}{\omega- E_{\xi} +i 0^+} +
634: \frac{v^2_{\xi}}{\omega+ E_{\xi} +i 0^+}\:,\\
635: g^A(\xi , \omega)=\frac{u^2_{\xi}}{\omega- E_{\xi} -i 0^+} +
636: \frac{v^2_{\xi}}{\omega+ E_{\xi} -i 0^+}\:.
637: \enea
638: The equilibrium values for $u_{\xi}$ and $v_{\xi}$ are:
639: \beq u_{\xi}^0
640: = \left (\met( 1 + \frac{\xi }{E}) \right )^\met \: ,
641: \hspace*{1cm} v_{\xi}^0 = \left (\met( 1 - \frac{\xi }{E}) \right
642: )^\met \: ,
643: \eneq
644: with $E = \sqrt{{\xi}^2 + |\Delta_o|^2 }$. From now on we will drop
645: the subscript in the equilibrium gap parameter
646: (i.e. $\Delta\equiv\Delta_o$ if no time dependence is indicated
647: explicitely). Because the factor $\delta n_L(\xi,t) $ appearing in
648: eq.(\ref{qcl}), as given by Eq. (\ref{brown}) is odd w.r. to $\xi$
649: only the second term in $u^2$ and $v^2$ survives, when the integral in
650: Eq. (\ref{qcl}) is performed. Let us consider the case $\ww > \DD$
651: only and specialize Eq. (\ref{qcl}) to its diagonal part. According
652: to Eq. (\ref{brown}) we have:
653: \bea
654: \Re e\left \{\delta n_L \cdot g^R_o \right \} =
655: \Re e \left \{ - \frac{\sqrt {2} i}{\pi}\:
656: q(t) \: \int_{-\infty} ^{+\infty }d\:\xi \frac{ \xi }{2 E}
657: \: \rho\left [ 2\xi /(\sqrt {2}\ww _c ) \right ]
658: \left (\frac{1}{\omega- E+i 0^+} -
659: \frac{1}{\omega+ E+i 0^+}\right ) \right \}=
660: \nonumber \\
661: = - q(t) \:
662: 2\sqrt {2} \:
663: \rho\left [ 2(\ww ^2 -|\DD |^2 )^\met /(\sqrt {2}\ww _c )
664: \right ]\:.
665: \label{qqqq}
666: \enea
667: Here we have defined the function $q(t)$:
668: \beq
669: q(t) = q\: \:
670: \frac{\pi R_o^2}{R_o^2+ Dt} \:.
671: \eneq
672: Doing similarly for $g^A$ and subtracting, the imaginary part cancels:
673: \beq
674: \delta n_L \cdot \left ( {g}_o^R - {g}_o^A \right ) =
675: - q(t) 4 \sqrt {2} \: \rho\left [ 2(\ww ^2 -|\DD |^2 )^\met /(\sqrt
676: {2}\ww _c ) \right ] \; ,\;\; for \;\; \ww > \DD \:\:\: .
677: \eneq
678: Here the largest contribution of the non-equilibrium
679: excitations arises from $ \omega \sim \ww _c$. On the other hand
680: $ \ww _c$ can be larger or smaller than $\omega _D $.
681:
682: Now we evaluate the correction due to $\hat g^R_{ad}-\hat g^A_{ad}$.
683: In appendix C we show that an adiabatic solution of the motion
684: equation of $g^{R,A}$ is possible, in the sense that the functional
685: dependence on $\ww$ is the same as the equilibrium one, but the gap
686: parameter changes slowly with time (see Eq. (\ref{fg})):
687: \beq
688: \hat g_{ad}^{R(A)}=+(-)\frac{ \hat
689: M}{\sqrt{(\omega \pm i 0^+)^2-|\Delta (t) |^2}} \:
690: \label{fg1}
691: \eneq
692: with
693: \beq
694: \hat M=\left( \begin{array}{cc}
695: \omega & \Delta(t)\\
696: -\Delta(t) ^* &-\omega
697: \end{array}\right).
698: \eneq
699: This functional form for the $R/A$ functions is obtained if the $e-h$
700: symmetry is maintained and if one neglects the diffusion in space-time
701: which will be mainly important at intermediate times
702: \cite{ovchinnikov}.\\ This adiabatic approximation in the advanced and
703: retarded Green's functions allows us to write the Keldysh propagator in
704: the form:
705: \beq
706: g^K= g_{ad}^{K}- q(t) 4\sqrt {2}
707: \: \rho\left [ 2(\ww ^2 -|\DD |^2 )^\met /(\sqrt {2}\ww _c )
708: \right ] \:.\label{gggg}
709: \eneq
710: To calculate $f^K $ we resort to the analogous of eq.(\ref{ff}) which
711: is valid for $\hat g^K $: $ f^K = g^K \DD /\ww $. Hence we have:
712: \beq
713: f ^K = f_{ad}^K - q(t) 4 \sqrt {2} \: \frac{\DD}{\ww } \: \rho\left [
714: 2(\ww ^2 -|\DD |^2 )^\met /(\sqrt {2}\ww _c ) \right ] \;\;\; for \;\;
715: \ww >> \DD \:\:\: . \label{anomalo}
716: \eneq
717: Now we insert Eq. (\ref{anomalo}) into Eq. (\ref{gap}) and consider the
718: linear correction to the gap of the irradiated contact according to :
719: \[\Delta(t)= \Delta+\delta\Delta(t)\:.\]
720: Here $\delta\Delta(t)$ is the correction to the
721: unperturbed gap parameter $\DD$ due to
722: the radiation. In the limit of zero temperature, up to
723: first order in $q (t)$ and $\Delta/\omega_D$, $\delta \Delta(t)$ is given by:
724: \beq
725: \frac{\delta \DD (t)}{\Delta} =
726: -\frac{q(t)
727: 4 \sqrt {2}}{ln(2\omega_D/\Delta)-2+{\cal{O}}(\Delta^2/\omega_D^2)} \int_{\Delta }^{\omega_D} \; \frac{ d
728: \omega }{\ww} \: \rho\left [ 2(\ww ^2 -|\DD |^2 )^\met /(\sqrt {2}\ww
729: _c ) \right ] \: .
730: \label{delgap}
731: \eneq
732: \begin{figure}[ht]
733: % \hspace{-0.5cm}
734: \includegraphics[width=\columnwidth]{gap2.eps}
735: \caption{Variation of the gap immediately after the pulse vs
736: the inverse of duration of the pulse $\tau_c^{-1}=\omega_c$ in
737: units of $\Delta$. Our approximations break down for very low
738: ratios $\omega_c/\DD$ (long pulses). On the other hand, large
739: $\omega_c/\DD$ represent very short pulses: this situation is
740: unrealistic with the available optical devices.} \label{fig1}
741: \end{figure}
742: It is interesting to
743: note that the correction arising from the adiabatic dynamics has the
744: role of renormalizing the coupling $q(t)$ via the prefactor
745: $-[ln(2\omega_D/\Delta)-2]^{-1}$. This prefactor is negative because
746: $ln(2\omega_D/\Delta)>2$. Therefore Eq. (\ref{delgap}) shows that the
747: gap of the irradiated superconductor is decreased due to the non
748: equilibrium distribution of the qp excitations.
749:
750: Eq. (\ref{delgap}) has the same structure as Eq. (14) of ref.\cite{scalapino}.
751: There is a striking difference however.
752: The inverse square root singularity at the gap threshold,
753: which shows up in Eq.(14) of ref.\cite{scalapino}
754: does not appear in Eq. (\ref{delgap}).
755:
756: The inverse square root
757: singularity originates from the unperturbed density of
758: states at the excitation threshold and is usually present
759: in non-stationary superconductivity \cite{eliashberg}. It is responsible
760: for retardation and oscillating tails.
761: In our case, the gap threshold
762: plays a little role, because we do not
763: have extensive pair-breaking and q.p. generation
764: at energies $\sim \Delta $.
765: Hence, just a tail $ \sim 1/\omega $
766: survives in the integrand.
767:
768: In Fig.(\ref{fig1}) the variation of the gap immediately after the
769: pulse ($t\sim 0$), is plotted vs the inverse of the pulse duration
770: $\omega_c$ in units of $\Delta$, for different values of $\omega_D$.
771: Our approximations are not valid when the
772: pulse becomes too long (very low values of $\omega_c/\DD$).
773: For longer pulses the integrand in Eq. (\ref{delgap}) has a narrow peak
774: lined up at the gap threshold.
775: In this case the inverse square root singularity in the density of
776: states at the gap threshold is important and the adiabatic
777: approximation Eq.(\ref{gggg}) breaks down.
778:
779: For shorter pulses the peak becomes
780: broader and is centered at larger $\omega$'s. If the integration range is
781: small, the result is quite sensitive to the location of the peak
782: (see full line in Fig. (\ref{fig1})): most remarkably,
783: a minimum appears in the curve
784: when the pulse is rather long ($\omega_c/\omega_D<5$).
785: By contrast, the gap correction is rather
786: flat w.r.to changes of $\omega_c$ when $\omega_D / \Delta$ is larger
787: (broken and dotted line in Fig.\ref{fig1}).
788:
789:
790: \subsection{Correction to the Josephson current}
791: In this
792: subsection we derive the correction to the Josephson current arising
793: from the two terms of the anomalous propagator $f(R,t,t')$ given by
794: Eq. (\ref{anomalo}).
795: Eq.s(\ref{gggg},\ref{anomalo}) show that the non equilibrium Keldysh Green
796: functions of the irradiated superconductor can be separated into two
797: terms. The first one is what we call the 'adiabatic' contribution,
798: while the second one is strongly dependent on the non equilibrium qp's
799: distribution function and is first order in $q(t)$.
800: \\Within the linear response theory in the tunneling matrix element
801: $|T_o|^2$, the pair current at zero voltage is:
802: \bea
803: J (\vec R =0,t)=2e |T_o|^2 \int_{-\infty}^{\infty}\;dt'
804: e^{2ieV(t-t')}
805: \nonumber\\
806: \left ({f^>}_1^\dagger (0,t, t')\: f^{A}_2(0,t',t) +
807: {f^R}^\dagger_2 (0,t,t') \: f^<_1 (0 ,t',t)
808: \right )\;,
809: \label{jcur}
810: \enea
811: where $|T_o|^2$ is assumed to be independent of energy, for
812: simplicity. The current of Eq. (\ref{jcur})is evaluated at the junction
813: site, defined by $ \vec R = 0 $ and the irradiated
814: superconducting layer $S1$ is labeled by 1 here, while the
815: superconductor unexposed to radiation $S2$ is labeled by 2. The
816: perturbed Josephson current has an adiabatic term $J^{ad}(t)$
817: obtained by
818: inserting the first term of Eq. (\ref{anomalo}) into Eq. (\ref{jcur}),
819: plus a correction $\delta J(t)$
820: arising from the second term of Eq. (\ref{anomalo}). Using the
821: definitions ${f}^K = {f}^> + {f}^< $ and $ {f}^R - {f}^A = {f}^> -
822: {f}^< $ and expanding to lowest order in $q(t)$ the adiabatic critical
823: current is:
824: \bea
825: J^{ad}=J^{ad}_c(t) \sin(\varphi),\;\;\;\;\;\;
826: J^{ad}_c(t)=
827: \frac{\pi \h\Delta }{2 e R_N}\left(1 + \frac{\delta \DD(t)}{2\DD}
828: \right)\:,
829: \label{jos}
830: \enea
831: where $R_N$ is the normal resistence,
832: $\Delta$ is the unperturbed gap parameter of both contacts
833: (assuming $\Delta_1=\Delta e^{i \varphi}$ and $\Delta_2=\Delta$ in the
834: absence of laser perturbation), and $\delta\Delta(t)/\Delta$ is given by
835: Eq. (\ref{delgap}). Denoting by $\delta f^K$ the second term of
836: Eq. (\ref{anomalo}) the correction $ \delta J(t)$ is:
837: \beq
838: \delta J(t) =
839: e |T_o|^2 \int_{-\infty}^{\infty}\;d\ww \:
840: \left ( (\delta {f}^K(\omega+i e V) )^\dagger _1
841: {f}^A_2(\omega- i e V) + ({ {f}}^R(\omega+ i e V))^\dagger _2
842: ( \delta {f}^K (\omega- i e V))_1 \right )\:.\label{newjos}
843: \eneq
844: At zero temperature and $V = 0 $, this gives:
845: \bea
846: \delta J(t)=\frac{\pi \h }{e R_N} \int_{-\infty}^{\infty}\;d\ww \:
847: \frac{|\Delta|^2 q(t)}{\ww} \rho\left [ 2(\ww ^2 -|\DD |^2 )^\met /(\sqrt {2}\ww
848: _c ) \right ]\nonumber
849: \\ \left(
850: \frac{e^{-i\varphi}}{
851: \sqrt{(\omega+i0^+)^2-|\Delta |^2}}+\frac{e^{i \varphi}}{\sqrt{
852: (\omega-i0^+)^2-|\Delta |^2}}\right) \label{cosphi} \: .
853: \enea
854: which is zero for parity. This conclusion holds because we assume that
855: no charge imbalance occurs. If $V \neq 0 $ the contributions to the
856: integral evaluated in the complex plane are finite. \\$\delta J(t)$ is
857: a $cos \varphi$-like correction. In the unperturbed Josephson effect a
858: '$cos \varphi $' term only arises when $V>2 \DD$ \cite{barone}.
859: By contrast, our
860: calculation shows that a $cos \varphi$ term can arise in the Josephson
861: current with a small nonzero voltage in the presence of an ultrafast
862: laser pulse.
863:
864: \section{Classical dynamics of the irradiated junction}
865: In this section we integrate the classical equation of motion of the
866: irradiated junction numerically. Here we discuss the possibility that
867: the laser pulse induces switches of the junction from the zero voltage
868: state, to the resistive state. The characteristics of the Josephson
869: junction for a finite voltage, is obtained within the RCSJ
870: (resistively and capacitively shunted junction) model \cite{barone}.
871: The phase of the superconductor S2 is taken as the reference phase.
872:
873: In the absence of the pulse, the junction is biased by a current
874: constant in time $J_{b}$. As discussed in the previous Section, the
875: pulse activates the superconductor S1 by varying its gap dynamically
876: in time.
877:
878:
879: Consequently, a voltage $V$
880: arises at the junction, related to the dynamics of the phase
881: difference $\varphi (t) $. The latter solves the differential equation:
882: \bea
883: \ddot \varphi + Q_0^{-1} \dot \varphi + \frac { J_c^{ad}(t)}{J_c^o}
884: \sin \varphi(t)= \gamma
885: \label{mmm}\; ,
886: \enea
887: where $\gamma={J_{b}/J_c^o}$ and $J_c^o={(\pi\hbar \Delta)/(2eR_N)}$
888: is Josephson critical current of the unperturbed junction. The
889: time-dependent driving term is deduced from Eq. (\ref{jos}). We
890: assume $\omega_{pJ0} > \tau_E^{-1}$ ($\tau_E^{-1}(Nb)\sim 7 GHz$), where
891: $\omega_{pJ0}$ is the plasma frequency at zero bias. This condition is
892: satisfied for high quality $Nb$ junctions, where $\omega_{pJ0}$ is in
893: the range $40 GHz \div 120 GHz$ \cite{wallraff}, but it holds also if
894: we take into account the dependence of the plasma frequency on $\gamma
895: $: $\omega_{pJ}=\omega_{pJ0} (1 -
896: \gamma^2)^{(1/4)}$. At $\gamma = 0.98$ the term $(1
897: - \gamma^2)^{(1/4)} = 0.44$: this still gives a large plasma frequency
898: for the given range. In any case the plasma frequency changes
899: marginally when the energy is degradated into heat if $q$ is
900: small. Under this conditions the relaxation process occurs long after
901: the switch to resistive state.
902:
903: In Eq. (\ref{mmm}), $Q_0=\omega_{pJ0} R(\varphi) C$ is the quality
904: factor, where $R(\varphi)$ is the junction intrinsic resistance, which
905: is in general a non-linear function of the phase.
906: The dissipative $Q^{-1}_0\dot
907: \varphi$ term includes thermal incoherent pair breaking effects at
908: equilibrium. In the simulation we use both a constant junction
909: resistance $R$ and a patchwork model given by
910: \cite{Likharev}:
911: \begin{equation}
912: Q^{-1}(\dot\varphi)=Q_0^{-1}\frac{\omega_{pJ}}{\Delta}\frac{(\frac{\dot
913: \varphi \omega_{pJ}}{\Delta})^{N}}{1+(\frac{\dot \varphi \omega_{pJ}
914: }{\Delta})^{N}}
915: \label{NLD}
916: \end{equation}
917: with $N=16$ and $Q_0^{-1}=0.636$, which corresponds to a normal
918: resistance above the gap $R_N=(\omega_{pJ} \varphi_0)/(c J^0_c)$. In
919: general we ignore the direct dependence of $R(\varphi)$ on the
920: phase. By the way, $Q$ should also depend on the energy which is
921: released by each single laser pulse due to the incoherent pair
922: breaking process. However, under the hypothesis that this energy is
923: very low we assume that the quality factor, due to the optically
924: induced normal resistance of the sample, is constant within the
925: considered energy range.\\ Actually, in the presence of the pulse,
926: also the current contribution of Eq. (\ref{newjos}) should be plugged
927: into the l.h.s. of Eq. (\ref{mmm}). This current term depends on the
928: voltage $V=\dot \varphi$. However, in view of the fact that in this
929: work we are only concerned with the switching of the junction out of
930: the zero voltage state, we do not derive the full dynamics of the
931: phase self-consistently.
932:
933: \begin{figure}[ht]
934: \vspace{0.5cm} \hspace{-0.5cm} \includegraphics[width=\columnwidth
935: ]{Vg08.eps} \caption{Voltage behavior in time for
936: different energy released on the sample and the quality factor $Q =
937: 100$. The voltage is normalized to $V_0=\omega_{pJ} \varphi_0/c$.}
938: \label{voltaggio}
939: \end{figure}
940: \begin{figure}[ht]
941: \vspace{0.5cm} \hspace{-0.5cm} \includegraphics[width=\columnwidth
942: ]{carat.eps} \caption{(Color online) Voltage behavior in time with different
943: quality factors $Q$ in the linear conductance model (A,B)
944: and the non-linear conductance model as given by
945: Eq. (\ref{NLD}) (C). The voltage is normalized to $V_0=\omega_{pJ}
946: \varphi_0/c$.}
947: \label{voltaggio1}
948: \end{figure}
949: \begin{figure}[ht]
950: \vspace{0.5cm} \hspace{-0.5cm} \includegraphics[width=\columnwidth
951: ]{conds2.eps} \caption{Switching front in the parameter space
952: a)$(\omega_c/\Delta,q)$ at fixed current bias $\gamma = J_b/J^o_c $ and b)
953: $(\gamma,q)$ at fixed $\omega_c/\Delta$, for $Q = 100$ and
954: $\omega_D/\Delta=10$. The full curves have been added as a guide
955: for the eye: they mark the border between the zero voltage
956: (Josephson) state and the voltage state. $q$ is the coupling
957: strength due to the laser pulse.}
958: \label{fig5}
959: \end{figure}
960:
961: In Fig. (\ref{voltaggio}) we show the voltage just after the pulse for
962: different values of the released energy. The time evolution of the
963: voltage is sketched for some successfully induced switches. The
964: junction starts in the zero voltage state. At $\omega_{pJ}t=0^+$ it is
965: irradiated by the laser pulse. There are few oscillations at
966: frequency $\omega_{pJ}$ before the switching occurs, followed by an
967: overall increase of the oscillating voltage. The larger is $q$ the
968: faster is the switch. If no switch is induced the junction remains in
969: the zero voltage state: the phase and the voltage are weakly perturbed
970: by the radiation and show decaying plasma oscillations around their
971: equilibrium values.
972:
973: In Fig.(\ref{voltaggio1}) we show the approach to the gap voltage for
974: different $Q$ values and two different conductance models. Except
975: for the asymptotic trend, the curves for $Q=10$ (B) and $Q=100$(A) show a
976: similar behavior. The non-linear conductance gives rise to a more
977: pronounced shoulder in the curve after the first increase of the
978: voltage. The first phase oscillation at frequency $\omega_{pJ}$ are
979: largely independent of the dissipation model used.
980:
981: The switching of the junction out of the zero voltage state depends on
982: the bias current $J_{b}$, on the released energy per pulse $q$, and on
983: the pulse duration. In Fig.(\ref{fig5}) we sketch the switching front
984: in the parameter space a)$(\omega_c/\Delta,q)$ at fixed $\gamma$ and
985: b) $(\gamma,q)$ at fixed $\omega_c/\Delta$, for $Q = 100$ and
986: $\omega_D /\Delta=10$. For each point $(q,\omega_c/\Delta ,\gamma)$
987: we calculate $J_c^{ad}(t)$ from Eq.s(\ref{jos}, \ref{delgap}). Next we
988: plug the result into the equation of motion Eq. (\ref{mmm}). Numerical
989: simulation of the dynamics shows whether the junction is stable in the
990: zero voltage state, or it switches to a running state. The points of
991: the curve mark the frontier between the two behaviors. The full curve
992: is just a guide for the eye. The non-monotonicity of
993: $\delta\Delta(t)/\Delta$ with the pulse duration, appearing in
994: Fig.(\ref{fig1}) forces a similar behavior in
995: Fig.({\ref{fig5}}$a$). This means that the pulse duration can be
996: appropriately chosen, in order to optimize the junction switching with
997: the laser field. Indeed, if $\omega_c/\Delta$ is of the order of
998: $5\div 15$ a very small released energy is required for the switching
999: of the junction, because the order parameter is much depressed by the
1000: laser pulse in that range. Out of this range the shorter is the pulse
1001: the larger is the energy required. By contrast a slightly larger
1002: released energy is also required for longer pulses
1003: ($\omega_c/\Delta<5$). This is because longer pulses imply a more
1004: extended change in the qp distribution up to higher energies. As a
1005: consequence the maximum of the function $\rho$ of Eq.
1006: \ref{delgap} contributes less to the correction of the gap parameter.
1007: Nevertheless caution should be used in considering our results for
1008: longer pulses because of the neglected relaxation effects.
1009:
1010: \section{Conclusion}
1011: The effect of an ultrafast laser pulse on the superconducting
1012: coherence at a Josephson Junction allows studying an unexplored regime
1013: in non equilibrium superconductivity.
1014:
1015: Non-equilibrium in superconductivity is usually addressed in the
1016: context of one of the possible applications of Josephson junctions,
1017: that is radiation$/$particle detectors. Highly energetic radiation
1018: produces pair breaking and quasiparticles which, in turn, excite a
1019: large number of them, in a cascade process. Usually the setup is
1020: optimized such that the qp's can be collected and contribute to the
1021: current across the junction with a sharp signal. Losses are due to
1022: degradation of the released energy into heat during the relaxation
1023: process. To achieve optimum performance, the Josephson current is
1024: usually suppressed by applying a magnetic field. This picture has been
1025: discussed quantitatively by examining the quasiclassical kinetic
1026: equation for the non equilibrium qp distribution function
1027: \cite{ovchinnikov}.
1028:
1029: In this work we have concentrated on a quite different time scale: the
1030: one fixed by the duration of an ultrafast laser pulse. While the
1031: relaxation process mentioned above takes place on a time scale of $0.1
1032: \div 100 \; ns$, we have considered a laser perturbation lasting at
1033: most hundreds of femtoseconds. This type of tool can be quite valuable
1034: for future applications, because fast pulses in flux and gate voltages
1035: are extremely important when processing information in superconducting
1036: quantum computing devices (qubits) \cite{makhlin}. Indeed, finite rise
1037: and fall times of pulses may result in a significant error in
1038: dynamical computation schemes \cite{choi}. The carrier frequency of
1039: the laser is $\sim 100 THz$ and the optical radiation is expected to
1040: produce many $e-h$ pair excitations as it would occur in a normal
1041: metal. In our case, qp's do not have enough time to relax down to the
1042: typical phonon frequencies ($\sim \omega_D$) and to heat the sample
1043: before the stimulated switching occurs. We do not wish to collect qp's either,
1044: what requires a suitable geometry of the junction.
1045:
1046: Instead, we have addressed the question how the critical current for
1047: Josephson conduction $J_c$ can be coherently affected by a laser
1048: induced small perturbation with an unrelaxed non-equilibrium
1049: distribution of qp's, that is before the dissipative response sets
1050: in.
1051:
1052: Using the quasiclassical approach to non-equilibrium Keldysh Green's
1053: functions, we have shown that, if the temperature is very low, the
1054: order parameter of the irradiated superconductor can respond
1055: adiabatically to a weak perturbing signal. A non equilibrium
1056: distribution of qp's is generated and consequently $J_c$ is
1057: temporarily reduced (see Eq. (\ref{jos}) and Fig(\ref{fig1})). This
1058: reduction can drive switches of the junction out of the zero voltage
1059: state. In our approach the retardation effects which arise from the
1060: frequency dependence of the $e-ph$ coupling $\alpha^2 F(\omega)$ and
1061: from the actual features of $e-ph$ relaxation processes have been
1062: neglected. They came into play on a time scale longer then the
1063: duration of the laser pulse, $\omega _c ^{-1} \sim \tau _c $. Indeed
1064: in the equation of motion for the Keldysh Green's function reported in
1065: appendix B and C the role of the frequency dependent self-energy terms
1066: has not been discussed.
1067:
1068: The parameter which is
1069: related to the energy released by the radiation and describes the
1070: strength of the perturbation is $q$. In our case, the switches can be
1071: induced by pulses with $q \sim 0.05$, with relatively low values of
1072: $\omega_c/\Delta\sim 10$, by polarizing the junction very close to the
1073: critical current.
1074:
1075: In experiments on laser induced non-equilibrium effects in
1076: superconductors \cite{testardi} or Josephson junctions \cite{parker},
1077: the released energy is of few $\mu J$, which corresponds to values of
1078: $q$ between $0.25$ and $0.67$ for the given laser spot dimension.
1079: In our case for $q=0.05$ the switching occurs at 98\% of the critical
1080: current. Therefore, a coherent effect of the laser on the
1081: superconducting condensate is sufficiently large to be observed in sensitive
1082: experiments monitoring the escape rate \cite{thermal,quantum}.
1083: These experiments can appreciate very small variations of the critical
1084: current, if temperature is kept low and the released energy is
1085: sufficiently small, so that sizeable heating effects do not occur.
1086:
1087: We have also found a `$\cos \varphi $' contribution to the Josephson
1088: conduction due to the presence of the excited qp's (see
1089: Eq. (\ref{cosphi})). This term, which will be examined in detail
1090: elsewhere, vanishes at zero $V,T$, provided excitations do not
1091: generate charge imbalance. A similar term, proportional to the voltage
1092: $V$, can be derived also in the BCS theory of Josephson conduction
1093: \cite{barone} but it is identically zero as long as $V \le 2 \Delta /e
1094: $, because of the absence of qp's at zero temperature. This is not the
1095: case here, due to the presence of a non-equilibrium qp distribution.
1096:
1097: Our derivation of Eq.s(\ref{delgap},\ref{jos},\ref{cosphi}) assumes
1098: that no charge imbalance is created by the perturbation. This is
1099: because the radiation excites $e-h$ pairs and the pulse duration is
1100: short enough, so that pair breaking is very limited. The absence of
1101: charge imbalance is a crucial approximation in our solution
1102: scheme. This assumption allows us to keep the unperturbed functional
1103: form of the quasiclassical Green's functions and to insert a time
1104: dependent gap parameter $\Delta (t)$ in their expressions. $\Delta
1105: (t)$ follows the perturbation adiabatically and is determined by the
1106: non equilibrium $e-h$ pair distribution produced by the pulse. Charge
1107: imbalance corrections should be reconsidered, but they are usually
1108: expected to have a minor effect on the dynamics of the junction.
1109:
1110: To complete the picture, we have simulated the classical dynamics of
1111: the junction switching to the resistive state. This picture is only
1112: valid over few periods $2 \pi /\omega_{pJ}$ on time duration less than
1113: the electron-phonon relaxation time.
1114:
1115: Precursor oscillations can be seen in Fig. (\ref{voltaggio})
1116: at the Josephson plasma frequency $\omega _{pJ}$.
1117: Most remarkably the duration of the pulse can be optimized
1118: in order to induce controlled coherent switching
1119: at the minimum possible released energy ${\cal{E}}$.
1120:
1121: \begin{acknowledgments}
1122: We are indebted to B.Altshuler, A.Barone, D.Bercioux, L.N.Bulaevskii,
1123: F.Hekking, Yu.N.Ovchinnikov, G.Pepe and G.Schoen for useful discussions
1124: at various
1125: stages in the preparation of this paper.
1126: \end{acknowledgments}
1127: \appendix
1128: \section{Quasiclassical time dependent Green's function approach}
1129: The quasiclassical Green's function solve the Eilenberger form of the
1130: Gorkov equations in commutator form \cite{belzig}:
1131: \beq
1132: \left [ \left
1133: (\breve{G}_0^{-1} -\breve \Sigma -\breve \Delta \right ),
1134: \GG \right ]_ \otimes =0 \;.
1135: \label{gorkov}
1136: \eneq
1137: The matrix Green's function $\GG$ in the Keldysh space is:
1138: \beq
1139: \GG= \left( \begin{array}{cc}
1140: \hat G^R & \hat G^K\\
1141: \hat 0 &\hat G^A
1142: \end{array}\right) \:\: ,
1143: \eneq
1144: where, in turn, $\hat G^R,\hat G^A,\hat G^K$ are the retarded,
1145: advanced and Keldysh Green functions in the Nambu space:
1146: \beq
1147: \hat G^{(A,R,K)} =
1148: \left( \begin{array}{cc}
1149: g^{(A,R,K)} & f^{(A,R,K)}\\
1150: -f^{(A,R,K)\dagger} & - g^{(A,R,K)\dagger}
1151: \end{array}\right)\:\;.
1152: \eneq
1153: Here $f$ is the anomalous propagator and its Keldysh component
1154: defines the gap:
1155: \beq
1156: \Delta= \frac{\nu (0)\lambda}{4} \int_{-\infty}^{+\infty} d \omega\; <f^K>_{\vec v_F}
1157: \:. \label{eqgap}
1158: \eneq
1159: The average $<>_{\vec v _ F}$ denotes angular averaging over the Fermi
1160: surface.
1161: The gap matrix $\breve \DD$ is defined as:
1162: \beq
1163: \breve \Delta \equiv \left(
1164: \begin{array}{cc}
1165: \hat \Delta & 0\\ 0 &\hat \Delta
1166: \end{array}\right) \: ,
1167: \:\:\:\:
1168: \hat \Delta= \left(
1169: \begin{array}{cc}
1170: 0& -\Delta \\ \Delta^* &0
1171: \end{array}\right) \: . \label{eqgap1}
1172: \eneq
1173: The self-energy $\breve \Sigma $ includes elastic and inelastic
1174: scattering with impurities and gives rise to relaxation processes.
1175: The commutator is evaluated w.r. to the $\otimes$ operation, which
1176: implies integration over the intermediate variables according to:
1177: \beq
1178: \GG _0^{-1}\otimes \GG _0\equiv \int d2\: \GG _0^{-1}(1,2)
1179: \GG (2,1') \: ,
1180: \eneq
1181: where $1 \equiv (\vec r_1 ,t_1)$.
1182: The differential operator $\GG _0^{-1}(1,2)$ is
1183: \beq
1184: \GG _0^{-1}(1,2)=\left [ i\breve \sigma _3
1185: \partial _{t_1} -
1186: \frac{1}{2m} \left (\vec \nabla_{\vec{r}_1} -i \frac{e}{c} \breve {\sigma }_3
1187: \vec A(1)
1188: \right ) ^2 +(e\phi(1) -\mu) \breve I \right ]\delta(1-2)\;.
1189: \eneq
1190: Here
1191: \beq
1192: \breve \sigma_i \equiv \left(
1193: \begin{array}{cc}
1194: \hat \sigma_i & 0\\ 0 &\hat \sigma_i
1195: \end{array}\right) ,\;\;\;\;\breve I \equiv \left(
1196: \begin{array}{cc}
1197: \hat I & 0\\ 0 &\hat I
1198: \end{array}\right) \;\;\;\;,
1199: \eneq
1200: where $\hat{\sigma}_i (i=1,2,3)$ are the usual Pauli matrices and $\hat I$ the
1201: $2\times2$ unit matrix .\\
1202: The vector potential $ \vec A(1) $
1203: describes the laser radiation field of frequency
1204: $\Omega$:
1205: \beq
1206: \vec{ A}(1)=\sum_{\pm} \sum _{\vec{p}}\vec{a}_\pm
1207: (p,t)e^{\mp i (\vec{p}\cdot \vec{r}_1 - \Omega t)} \; ,
1208: \eneq
1209: where $\vec{a}_{\pm}$ can be slowly varying 'envelope' functions of
1210: space, on the
1211: light spot size $ R_o $ and on time, on the scale $\Omega^{-1}$. These
1212: are the reference space and time scales in the following.
1213: In the frame of the quasiclassical approximation, the original Green's
1214: functions $ G(1,2) \equiv G(\vec{r},t, \vec{r}',t') $ are
1215: assumed to be slowly varying function of the coordinate $\vec R
1216: =(\vec{r}+\vec{r}')/2 $ while they oscillate fast as functions of
1217: $\vec{r}-\vec{r}'$ on the scale of the Fermi wavelength $\lambda_F$.
1218: It is customary to rewrite also the time dependence in terms of the
1219: new variables $ \overline t = ( t +t')/2 $ and $ t-t' $ and to Fourier
1220: transform $w.r.to$ $\vec{r}-\vec{r}'$ and $ t-t' $, thus obtaining $
1221: G( \vec{p},\omega ,\vec{R},\overline t ) $.\\
1222: The motion equation for the Keldysh component of
1223: eq.(\ref{gorkov}) reads:
1224: \beq
1225: \left[ \hat G_0^{-1} -\Re e (\hat \Sigma)
1226: -\hat \Delta,\hat G^K \right ]_{\otimes}
1227: = \left[ \hat \Sigma ^K,\Re e ( \hat G)
1228: \right]_{\otimes} +\frac{i}{2}\left
1229: \{\hat \Sigma^K ,\hat A \right\}_{\otimes}
1230: -\frac{i}{2}\left\{\hat \Gamma , \hat G^K\right\}_{\otimes} \: ,
1231: \label{kcomp}
1232: \eneq
1233: where we have defined a quantity proportional to the density of states
1234: $\hat A= i(\hat G^R-\hat G^A)$ (not to be confused with the vector
1235: potential), and written down the imaginary and the real part of the
1236: self-energy, $\hat\Gamma=i(\hat\Sigma^R-\hat\Sigma^A)$ and $\Re e
1237: \hat\Sigma=\met (\hat \Sigma^R+\hat \Sigma^A)$, respectively, as well
1238: as the real part of the retarded/advanced Green's function $\Re e
1239: \hat G =\met (\hat G^R+\hat G^A)$ ($\{ ,\} $ denotes the anticommutator ).
1240: The next step is the gradient expansion of the $\otimes$ product (we
1241: drop the overline on $t$ in the following):
1242: \beq
1243: \hat C \otimes \hat B = \exp {1/2 (\partial^C_t \partial^B_\omega -
1244: \partial^C_\omega \partial^B_t) }
1245: \exp {1/2 (\partial_p^C \partial_R^B-\partial^C_R \partial_p^B )}
1246: \hat C \hat B \:,
1247: \eneq
1248: (here $\partial_p^C$ stands for the gradient w.r.to the impulse
1249: operating on $\hat C$ ), followed by the averaging of the result for
1250: $| \vec{p}|$ close to $p_F$, that is over the energies $p^2/2m-\mu
1251: \equiv\xi $ while the direction of $\vec{p}$, $\hat p $, is untouched:
1252: \beq
1253: \hat{g}(\hat{p},\omega ,\vec{R} , t)=\frac{i}{\pi}\int
1254: d\;\xi \:\: \hat{G}(\vec{p}, \omega ,\vec R , t) \:\: .
1255: \eneq
1256: This is justified, because $\lambda_F$ is much shorter both of the
1257: superconducting correlation length and of the spatial range of the
1258: laser spot ($e-h$ symmetry is assumed). Eventually $\hat g$ depends on
1259: $\hat p,\omega,R, t$. The average of eq.(\ref{kcomp}) over all
1260: directions in the Fermi surface, can be done if no external bias is
1261: applied and anisotropies of the diffusion and relaxation process are
1262: not expected. In the presence of radiation with optical frequency it
1263: is customary to average out the fast oscillating components with
1264: frequency $\Omega $
1265: \cite{landau}. Following eq.(\ref{potvet}), we expand the Green's
1266: functions similarly:
1267: \beq
1268: \hat{\underline g}(\omega,R,t) = {\hat g} (\omega,R,t) + \sum_{\pm}
1269: \hat{\chi} ^{\pm}
1270: (\omega,R,t) e^{\pm i \Omega t}\; .
1271: \eneq
1272: Here $\underline {\hat g }(\omega,R,t)$ is assumed to be the slowly
1273: varying part on the scale of the pulse duration $\tau _c$, while $\hat
1274: {\chi}^\pm $ are fast varying ones. All these functions are slowly
1275: varying functions of space as well, on the light spot size scale. A
1276: decomposition of Eq. (\ref{kcomp}) into harmonics arises. We are
1277: interested in the zero order harmonic equation, which shows a slow
1278: dynamics that can be followed coherently by the irradiated
1279: superconductor. Leaving the self-energy terms in Eq. (\ref{kcomp}) for
1280: the moment aside and dropping the superscript $^{(K)}$, we obtain
1281: \beq
1282: [\tau_3 a_+,\hat {\chi}^-]+[\tau_3 a_-,\hat {\chi}^+]
1283: +[(\omega \hat \tau_3 - \hat \Delta), \ug]
1284: -\met\{\hat\tau_3,\partial_t \ug\} +\met \{\partial_t
1285: \hat \Delta,\partial_\omega \ug\}-\frac{e^2}{2mc^2}
1286: \partial_t A^2\partial_\omega \ug = ... \: ,
1287: \label{gorfin}
1288: \eneq
1289: where the ellipsis refers to the missing self energy terms. The first
1290: order harmonic equations are:
1291: \beq
1292: \pm i \; \{\tau_3, \hat{\chi}^{\pm}\} +\frac{ e }{2mc \Omega}[a_\pm,\ug]=0\;,
1293: \eneq
1294: they show that the first two terms in Eq. (\ref {gorfin}) are ${\cal{O}} (
1295: \Omega ^{-1}) $ smaller than the others and can be neglected to lowest
1296: order. Hence the effective equations for the Green functions are
1297: \beq
1298: [(\omega \hat \tau_3 - \hat \Delta), \ug]
1299: -\met\{\hat\tau_3,\partial_t \ug\} +\met \{\partial_t
1300: \hat \Delta,\partial_\omega \ug\}-\frac{e^2}{2mc^2}
1301: \partial_t A^2\partial_\omega \ug = ... \: ,
1302: \label{gorfin2}
1303: \eneq
1304: The last three terms in Eq. (\ref {gorfin2}) include the time
1305: dependent non equilibrium dynamics that is absent in the case of a time
1306: independent approach.
1307: Retarded, advanced and Keldysh Green's function, they all
1308: satisfy analogous equations.
1309: \section{Kinetic equation for $n(\omega,r,t)$}
1310: One can linearize Eq. (\ref{gorfin2}) for Keldysh Green's function by
1311: posing $\hat g^K=\hat g^R \hat h - \hat h \hat g^A$. This yields the
1312: kinetic equation for the distribution function $n(\omega, R,t)$
1313: \cite{scalapino}. We
1314: neglect any variation in space and concentrate on the $t$ dependence
1315: here. The distribution matrix $\hat h$ is defined starting from the
1316: $n_L$ and $n_T$ functions according to:
1317: \beq
1318: \hat h= n_L \hat 1 + n_T \hat \sigma_3\:,
1319: \eneq
1320: or
1321: \beq \hat h =
1322: \left( \begin{array}{cc}
1323: n_L (E)+n_T(E)& 0\\
1324: 0&n_L (E) - n_T(E) \end{array}\right) \equiv \left( \begin{array}{cc}
1325: 1-2n(E)& 0\\
1326: 0&2n(-E)- 1\end{array}\right)\: .
1327: \label{funh}
1328: \eneq
1329: The second equality defines the relation with the qp distribution
1330: function $n (E)$. We always assume $e-h$ symmetry, so that
1331: $n(E)+n (-E ) =1 $ and $n_T = 0 $.
1332: In the equilibrium case one has $ n_o(E)=\frac{1}{e^{E/T}+1} $, so that:
1333: \beq
1334: n_{L(T)}^o= \met\left[tanh(\frac{E}{2T})+(-)tanh(\frac{E}{2T})\right]\:.
1335: \eneq
1336: Substituting Eq. (\ref{ggk})
1337: in Eq. (\ref{gorfin2}), we get the kinetic equation for the distribution
1338: matrix $\hat h $.In particular, in case there is no charge imbalance,
1339: the equation for the longitudinal component $n_L$ is
1340: \cite{ovchinnikov}:
1341: \beq
1342: \partial_ t n_L Tr((\hat g ^R \hat \tau_3- \hat \tau_3 \hat g^A))+
1343: \partial _\ww n_L Tr(\partial _t \hat \DD (\hat g ^R - \hat g^A))
1344: -\frac{e^2}{2 m c^2}\partial_t A^2 Tr
1345: \left(\partial_\omega n_L\left(\hat g^R-\hat
1346: g^A\right)\right)
1347: =
1348: -4 I[n_L(\ww)] \:, \label{ovA2}
1349: \eneq
1350: where $I[n_L(\ww)]$ is a collision integral which regulates the
1351: relaxation of the qp distribution toward equilibrium. Eq. (\ref{ovA2})
1352: is
1353: fully discussed in ref.\cite{ovchinnikov}.
1354:
1355: \section{Equation of motion of $ G^R $}
1356: We now write down the equations Eq. (\ref{gorfin2}) explicitly for the
1357: retarded Green's functions. We label the matrix components by $(i,j)\:
1358: (i,j=1,2 )$ and drop the superscript $^R $ everywhere. The matrix
1359: elements of Eq. (\ref{gorfin})in the Nambu space become :
1360: \bea
1361: &{\bf (1,1)\to }& \partial_t g - \DD f^\dagger + \DD^* f
1362: + \partial_t \Delta \partial_\omega f^\dagger+\partial_t \DD^*
1363: \partial _\omega f
1364: + \frac{e^2}{2 m c^2} \partial_t A^2 \partial_ \omega g= ...\nonumber\\
1365: &{\bf (2,2) \to }& \partial_t g^\dagger
1366: + \DD f^\dagger- \DD^* f
1367: + \partial_t \Delta \partial_\omega f^\dagger+ \partial_t \DD^*
1368: \partial _\omega f
1369: -\frac{e^2}{2 m c^2} \partial_t A^2 \partial_ \omega g^\dagger =
1370: ...\nonumber\\
1371: &{\bf (1,2) \to }& 2 \omega f - \DD g^\dagger- g \DD
1372: + \partial_t \DD \partial_\omega g^\dagger - \partial_\omega
1373: g
1374: \partial_t \DD + \frac{e^2}{2 m
1375: c^2}\partial _t A^2
1376: \partial _\omega f= ...\nonumber\\
1377: &{\bf (2,1) \to }& +2 \omega f^\dagger - \DD^* g - g^\dagger \DD^*
1378: + \partial_t \DD^* \partial_\omega g - \partial_\omega
1379: g^\dagger
1380: \partial_t \DD^*
1381: - \frac{e^2}{2 m c^2} \partial _t A^2 \partial _\omega f^\dagger = ...
1382: \nonumber\:\:\:\:\:\:.
1383: \enea
1384: The equilibrium result suggests that
1385: \beq
1386: f=\frac{\Delta(t)}{\omega} g \label{ff}
1387: \eneq
1388: solves Eq.$(1,2)$ except for terms $\propto\partial_t A^2$ which
1389: describe the relaxation at later times. Let us assume that this
1390: relation holds also in the non-equilibrium case. Then the formal
1391: solution, follows adiabatically the $t-$dependence of the gap parameter $\Delta$ by keeping an equilibrium-like shape:
1392: \beq
1393: g_{ad}=\frac{\omega}{\sqrt{\omega^2-|\Delta|^2(t)}},\;\;\;\;\;\;
1394: f_{ad}=\frac{\Delta(t)}{\sqrt{\omega^2-|\Delta|^2(t)}}
1395: \label{fg}
1396: \eneq
1397: This approximate solution is quite appealing, because it satisfies
1398: the equilibrium
1399: condition for $\Delta$ at $t\to \infty $. However it neglects
1400: diffusion in space-time which will be mainly important at later times
1401: w.r. to the pulse duration.
1402:
1403:
1404:
1405: %
1406: % and use \bibitem to create references.
1407: %
1408:
1409: \begin{thebibliography}{40}
1410: \bibitem{barone} A. Barone and G. Patern\'o,
1411: {\it Physics and applications of Josephson effect}, Wiley New York,
1412: 1982.
1413: \bibitem{testardi} L. R. Testardi, Phys. Rev. B {\bf 4}, 2189, 1971.
1414: \bibitem{parker} W. H. Parker and W. D. Williams, Phys. Rev. Lett. {\bf 29}, 924, 1972.
1415: \bibitem{owen} C. S. Owen and D. J. Scalapino, Phys. Rev. Lett. {\bf 28}, 1559, 1972.
1416: \bibitem{ivlev} R. A. Vardanyan and B. I. Ivlev, Sov. Phys. JEPT {\bf 38},
1417: 1156, 1974.
1418: \bibitem{ser}A.V. Sergeev, M.Yu. Reizer,
1419: J. Mod. Phys. B \textbf{6} 635 (1996), M.Yu. Reizer, Phys. Rev.
1420: B \textbf{39} 1602 (1989), A.V. Sergeev, M.Yu. Reizer, {Zh. Eks.
1421: Teor. Fiz.} \textbf{90} 1096 (1986) [{Sov. Phys. JETP} \textbf{63}
1422: 616 (1986)]
1423: \bibitem{lindg} M. Lindgren, M. Currie, C. Williams, T.Y. Hsiang,
1424: P.M. Fauchet, R. Sobolewski, S.H. Moffat, R.A. Hughes, J.S.
1425: Preston, and F.A. Hegmann {Appl. Phys. Lett.} \textbf{74} 853
1426: (1999)
1427: \bibitem{semenov}A.D. Semenov, R.S. Nebosis, Yu. Gousev, M.A.
1428: Heusinger, K.F. Renk, {Phys. Rev. B} \textbf{52} 581 (1995).
1429: \bibitem{kaplan}S.B.Kaplan, C.C.Chi, D.N.Langenberg, J.J.Chuang, S.Jafarey,
1430: D.J. Scalapino, Phys.Reb.B {\bf 14}, 4854, 1976.
1431: \bibitem{thermal} M. H. Devoret, J. M. Martinis, D. Esteve and J. Clarke,
1432: Phys. Rev. Lett. {\bf 53}, 1260, 1984.
1433: \bibitem{quantum} M. H. Devoret, J. M. Martinis and J. Clarke, Phys. Rev. Lett. {\bf 55}, 1908, 1985.
1434: \bibitem{latching} A.Moopenn, E.R. Arambula, M.J. Lewis, H.W. Chan,
1435: IEEE Trans. Appl. Supercond. 3, 2698, 1993,
1436: \bibitem{pulse} G. P. Pepe, G. Peluso, M.Valentino,
1437: A. Barone, L. Parlato, E. Esposito, C. Granata, M. Russo,
1438: C. De Leo and G. Rotoli Appl. Phys. Lett. {\bf 79}, 2770, 2001.
1439: \bibitem{barone2}Superconductive Particle Detectors,
1440: edited by A.Barone (World, Singapore, 1988); A. Barone, Nucl.Phys. B {\bf 44} 645 (1995).
1441: \bibitem{ovchinnikov}Yu.N.Ovchinnikov and V.Z.Kresin
1442: Eur.Phys.J.B {\bf 32}, 297, (2003), Yu.N.Ovchinnikov and V.Z.Kresin, Phys.Rev. B. {\bf 58},12416,(1998).
1443: \bibitem{rammer} A.I.Larkin and Yu.N.Ovchinnikov Zh.Eksp.Teor.Fiz {\bf 55},2262 (1968) [Sov.Phys. JETP {\bf 28}, 1200, 1969];
1444: J.Rammer, H.Smith, {Rev. Mod. Phys.},
1445: \textbf{58}, 323,(1986).
1446: \bibitem{belzig}W.Belzig, F.K.Wilhelm, C.Bruder, G.Schoen, E.D.Zaikin
1447: Superlattices and Microstructures {\bf 25}, 1251 (1999).
1448: \bibitem{amin}M.H.S.Amin, Phys.Rev.B {\bf 68 }, 054505, (2003).
1449: \bibitem{noi}preliminary report on some of this material has appeared on
1450: P.Lucignano, F.J.W.Hekking and A.Tagliacozzo, in ``Quantum computing and Quantum bits in mesoscopic
1451: systems'', A.J.Leggett, B.Ruggero, P.Sivestrini
1452: ed.s, Kluwer Academic, Plenum Publishers N.Y. (2004).
1453: \bibitem{grimvall} G.Grimvall 'The electron phonon interaction in metals', North Holland Publ. Co. Amsterdam 1981
1454: \bibitem{scalapino}J.J.Chang and D.J.Scalapino, Phys.Rev.B {\bf 15 },
1455: 2651, (1977).
1456: \bibitem{eliashberg} L.P.Gorkov and G.M.Eliashberg,Soviet Physics,JEPT{27},328
1457: (1968)
1458: \bibitem{gork} A.A.Abrikosov, L.P. Gor'kov, I.E.Dzyaloshinski, {\it Methods of quantum field theory in statistical physics};
1459: \bibitem{wallraff}A. Wallraff, A. Franz, A. V. Ustinov, V. V. Kurin,
1460: I. A. Shereshevsky and N. K. Vdovicheva, PhysicaB {\bf 284-288}, 575, 2000.
1461: \bibitem{Likharev} K. K. Likharev, {\it Dynamics of Josephson junctions and circuits}, Gordon and Breach Science Publishers, 1986.
1462: \bibitem{makhlin} Yu. Makhlin, G. Sch{\"o}n, and A. Shnirman,
1463: {Rev. Mod. Phys.} \textbf{73}, 357 (2001)
1464: \bibitem{choi} M.S. Choi, R. Fazio, J. Siewert, and C. Bruder,
1465: {Europhys. Lett.} \textbf{53}, 251 (2001)
1466: \bibitem{landau} L.D.Landau, E.M.Lifsits, Mechanics, Pergamon
1467: Press (1960).
1468: \end{thebibliography}
1469: \end{document}
1470:
1471:
1472:
1473:
1474:
1475: