1: \documentclass[pra]{revtex4}
2: \usepackage{graphicx}
3:
4: \begin{document}
5:
6:
7: \title{Running-phase state in a Josephson washboard potential}
8:
9: \author{G.~S.~Paraoanu}
10: \affiliation{NanoScience Center and Department of Physics, University of Jyv\"askyl\"a,
11: P.O.~Box 35 (YFL), FIN-40014 University of Jyv\"askyl\"a, FINLAND}
12:
13:
14: \begin{abstract}
15:
16:
17: We investigate the dynamics of the phase variable of an ideal underdamped Josephson junction in switching
18: current
19: experiments. These experiments have provided the first evidence for macroscopic quantum tunneling in large
20: Josephson junctions and are currently used for state read-out of superconducting qubits.
21: We calculate the shape of the resulting macroscopic wavepacket and find that
22: the propagation of the wavepacket long enough after a switching event leads to an average
23: voltage increasing linearly with time.
24:
25:
26:
27: \end{abstract}
28:
29: \maketitle
30:
31: %\section{Introduction}
32:
33:
34:
35: The dynamics of a large underdamped Josephson junction characterized by a
36: capacitance $C$ and Josephson energy $E_{J}$ can be described by the motion of a particle in a
37: washboard potential $U(\gamma )= E_{J}(1 - \cos\gamma ) + I\bar\Phi_0 \gamma $.
38: The particle has $C$ as the mass, the flux $\bar{\Phi}_{0}\gamma$ as the coordinate
39: and the charge on the capacitor $Q$ as the canonically conjugate momentum. Here $\gamma$ is the
40: phase difference of the superconducting order parameter across the junction and $\bar \Phi_{0} =
41: \Phi_{0} /2\pi = \hbar/2e$ is the flux quanta divided by $2\pi$. Much attention has been given since the
42: discovery of the Josephson effect to the switching dynamics of the junction in the thermal activation
43: regime and in the macroscopic tunneling (MQT) regime. Surprisingly, while the description of the state
44: of the junction before a switching event and calculations of the corresponding probability has been a topical issue
45: for many decades, what happens with the quantum state of the junction after tunneling did not receive that
46: much attention. It is argued \cite{tinkham,switch} that the junction ends up in a running-phase state,
47: with the voltage $Q/C$ increasing until it becomes sufficiently large so that the transport could be
48: done through quasiparticle excitations.
49: However, a quantum mechanical description of this state is missing. Much of what we understand about
50: the running-wave state, for instance
51: the physics of the retrapping current, comes from assuming a quasi-classical dynamics.
52:
53: In this paper we give an explicit formula for the macroscopic wavefunction of an ideally underdamped
54: junction after a MQT switching event.
55: If the switching probability is exponential,
56: which is the case for all the theoretical models and also confirmed experimentally,
57: one expects \cite{suppl} the following expression for the dynamics of the wavefunction
58: \begin{equation}
59: |\Psi (t)\rangle = e^{-\Gamma t /2}e^{-i\omega_0 t}|\Psi_{0}\rangle + |\Psi_{out}(t)\rangle.
60: \label{psii}
61: \end{equation}
62: In this equation, $\Psi_{0}$ is the initial state, coresponding to a bound state inside one of the
63: metastable wells, while $\Psi_{out}(t)$ is the wavefunction of the particle corresponding to states in the
64: continuum, outside the well (Fig. \ref{tfig}). This expression gives indeed an exponentially decreasing
65: probability for
66: the particle to be inside the well, with lifetime $\Gamma ^{-1}$.
67: In the following, we are interested in the structure of $\Psi_{out}(t)$.
68:
69: To solve this problem, the standard approach is to start with a wavepacket localized initially in one of the
70: metastable wells, and then expand it and evolve it in the eigenfunctions of the full Hamiltonian.
71: This procedure works for simple potentials \cite{deltapotential}, but even in these cases the solutions are
72: complicated. Fortunately, unlike problems in scattering theory, in condensed
73: matter the frequent situation is that we do not need an exact solution of the Scr\"odinger problem for tunneling,
74: but rather we are interested in the most generic features of it. In most cases in solid state physics,
75: tunneling is simply
76: treated as a process that annihilates a particle on some mode of a solid and creates one on another mode.
77: We will approach our problem in the same spirit \cite{cond}.
78: A good approximation in MQT is that no other state within the well is involved
79: with the exception of the state with energy $\omega_0$ in which the system is prepared, $|\Psi_0 \rangle$;
80: therefore one can write a reduced Hamiltonian of the form
81: \begin{equation}
82: H = \hbar\omega_0 |\Psi_0 \rangle\langle \Psi_0 | + \int \hbar\epsilon |\psi_\epsilon \rangle\langle
83: \psi_\epsilon | + \int d\epsilon \left[ k(\omega_0, \epsilon ) |\Psi_0 \rangle\langle
84: \psi_\epsilon |
85: + k(\epsilon , \omega_0 ) |\psi_\epsilon \rangle\langle \Psi_0 |\right], \label{hammi}
86: \end{equation}
87: where by $\{\psi_\epsilon \}$ we denote the continuum of eigenvectors outside the barrier.
88: We then write the wavefunction in the form
89: \begin{equation}
90: |\Psi (t) \rangle = a (t) e^{-i\omega_0 t} |\Psi_0\rangle + \int d\epsilon b (\epsilon, t) e^{-i\epsilon t}
91: |\psi_\epsilon \rangle ,
92: \label{lapl}
93: \end{equation}
94: with $a(0) = 1$, $b (0) = 0$. Inserting this expression in the Schr\"odinger equation we get
95: an integro-differential equation for $a (t)$. The Laplace transform of this equation reads
96: \begin{equation}
97: {\cal L} [a] (s) = \frac{1}{s + {\cal L}[\chi ](s)},\label{laplace}
98: \end{equation}
99: where
100: \begin{equation}
101: \chi (t) = \frac{1}{\hbar ^2}\int d\epsilon |k (\epsilon ,\omega_0 )|^2e^{i (\omega_0 - \epsilon )t}.
102: \end{equation}
103: In general, the tunneling matrix element $k(\omega_0, \epsilon)$
104: depend on the energies $\epsilon$ and they are determined by
105: the overlap of the left and right wavefunctions under the barrier
106: \cite{cond}.
107: We notice that since typically the lifetime of the metastable states is much larger than the
108: oscillation period in the well (in other words the last term in the Hamiltonian is a perturbation),
109: the states $\{|\epsilon \rangle \}$ which contribute effectively to tunneling
110: are located in a relatively small energy interval compared to the plasma oscillation frequency,
111: therefore the shape of these states under the barrier
112: is approximately identical. We can then take
113: the tunneling matrix element as being a complex constant; but since we
114: will be interested exclusively in the outgoing component, the
115: relative phase between the wavefunction inside the well and that outside
116: will not play any role. We have confirmed this assumption also by
117: expanding the initial wavefunction in terms of the WKB solution
118: of the washboard potential calculated in \cite{josephson}.
119: Therefore
120: we take $k = \hbar\sqrt{\Gamma /2\pi}$ real;
121: we obtain ${\cal L}[\chi] (s) = \Gamma /2$= constant, which turns out to be
122: the decay probability of the system. Indeed, the inverse Laplace transform of Eq. (\ref{laplace}) gives
123: precisely the classical exponential decay law
124: \begin{equation}
125: a (t) = e^{-\Gamma t/2}.
126: \end{equation}
127: The outgoing wavepacket becomes
128: \begin{equation}
129: |\Psi_{out} (t) \rangle = -i\sqrt{\frac{\Gamma }{2\pi}}\int d\epsilon \frac{e^{(-\Gamma /2 + i \omega_0 )t}- e^{-i\epsilon t}}{\Gamma /2 + i (\omega_{0} -\epsilon )}
130: |\psi _\epsilon^{\pm} \rangle .
131: \label{q}
132: \end{equation}
133: To conclude this derivation, we find that with the identification $k = \hbar\sqrt{\Gamma /2\pi}$ the Hamiltonian
134: Eq. (\ref{hammi}) becomes a model Hamiltonian for decay in the continuum which can be solved exactly, with
135: solution given by Eqs. (\ref{psii}) and (\ref{q}). A similar type of model Hamiltonian has been obtained in
136: \cite{twopotential,kurizkikofman}.
137: One can show, using the properties of the
138: Lorentz distribution, that
139: these wavefunctions are correctly normalized, as explained above.
140:
141: Let us now single out one component of the wave $\Psi_{out} (t)$, namely
142: \begin{equation}
143: \Psi^{\rightarrow}_{out} = i\sqrt{\frac{\Gamma }{2\pi}}\int d\epsilon \frac{e^{-i\epsilon t}}
144: {\Gamma /2 + i (\omega_{0} -\epsilon )}
145: |\psi _\epsilon^{+} \rangle .
146: \end{equation}
147: We first notice that the normalization of the total function $|\Psi_{out}\rangle$ is such that
148: $\langle \Psi_{out}|\Psi_{out}\rangle = 1 - \exp (-\Gamma t)$, which reflects correctly the fact that
149: the probability of finding the particle outside comes from an exponential decay law,
150: while that of $|\Psi^{\rightarrow}_{out}\rangle$
151: is such that $\langle \Psi^{\rightarrow}_{out}|\Psi^{\rightarrow}_{out}\rangle = 1$. In the following
152: we will see that $|\Psi^{\rightarrow}_{out} \rangle$ plays indeed a special role.
153: \begin{figure}[htb]
154: \includegraphics[width=84truemm]{tfigure.eps}
155: \caption{Tunneling out of one of the metastable wells of the washboard potential. }
156: \label{tfig}
157: \end{figure}
158: To move on, we notice that part of the expression for the outgoing phase contains a term
159: which decays exponentially on a time scale $\Gamma^{-1}$. These terms are associated with the
160: fast components of the localized wavefunction which would escape first.
161: Although a calculation that includes these terms is no doubt interesting, especially for the problem of
162: non-exponential decay rates \cite{deltapotential},
163: in what follows
164: we will regard them as transient oscillatory effects whose presence will be difficult to assess experimentally
165: anyway, and we will neglect their contribution.
166: In the WKB approximation, far enough from the classical turning point, the eigenvalues
167: $\{ |\psi _\epsilon \rangle \}$ have the form (up to a normalization factor and constant phase factors
168: due to matching to the region left of the classical turning point) of incoming and outgoing scattering states
169: \begin{equation}
170: \psi^{\pm}_\epsilon (\gamma ) \approx \sqrt{\frac{e}{C\bar\Phi _0 V_{\epsilon}(\gamma )}}\exp\left[\pm i\frac{C}{e}
171: \int_{\gamma_0}^{\gamma}V_{\epsilon }(\varphi)d\varphi \right] ,\label{wkb}
172: \end{equation}
173: where
174: \begin{equation}
175: V_{\epsilon}(\gamma ) = \frac{\hbar\omega_p}{\sqrt{2}e}\sqrt{\frac{2e\epsilon + I \gamma }
176: {I_{c0}} + 1 - \cos\gamma }.
177: \end{equation}
178: The physical meaning of this voltage is that it corresponds to the (classical) energy accumulated
179: on the capacitor when the phase difference across the junction is $\gamma$ and the initial energy
180: of the system is $\epsilon$; indeed, $CV^{2}_{\epsilon}(\gamma )/2 = \hbar\epsilon - U(\gamma )$.
181: We now use the fact that for values of $\gamma$ outside the well and far enough from the classical turning
182: point the inequality $|\hbar\epsilon - \hbar\omega_{0}|\ll \hbar\omega_0 - U(\gamma )$ holds.
183: We can therefore take $V_\epsilon (\gamma ) = V_{\omega_0}(\gamma )\stackrel{\rm not}{=}
184: V_{0}(\gamma )$ in the denominator of Eq. (\ref{wkb})
185: and approximate the exponent as
186: \begin{equation}
187: V_{\epsilon} (\gamma ) \approx V_{0}(\gamma )\left[1 + \frac{\epsilon - \omega_{0}}{2(
188: \omega_{0}- \hbar^{-1}U( \gamma ))}\right].
189: \end{equation}
190: With these approximations, using Eqs. (\ref{q}) and (\ref{wkb}) we can performe the integral over
191: the angular frequencies $\epsilon$; as a result, the contribution of the in-going scattering states in zero,
192: while the out-going scattering states build up a wavepacket of the form
193: \begin{equation}
194: \Psi^{\rightarrow}_{out}(\gamma ,t) = \frac{{\cal N}}{\sqrt{V_{0}(\gamma )}}
195: \exp\left[\frac{i}{\hbar}\int_{\gamma_{0}}^{\gamma }
196: C \tilde{V}(\varphi )\bar\Phi_0 d\varphi -
197: \left(i\omega_{0} +
198: \frac{\Gamma}{2}\right)\left(t-\int_{\gamma_0}
199: ^{\gamma}
200: \frac{\bar \Phi_{0}d\varphi }{V_{0}(\varphi)}\right)\right]
201: \Theta\left[t - \int_{\gamma_0}^{\gamma}
202: \frac{\bar \Phi_{0}d\varphi }{V_{0}(\varphi)}\right].
203: \label{out}
204: \end{equation}
205: Here ${\cal N}$ is a normalization factor which can be obtained through $\int_{-\infty}^{\infty}
206: d(\bar\Phi_{0}\gamma )|\Psi^{\rightarrow}_{out}(\gamma ,t)|^2 = 1$
207: with the mention that we make a negligible error by extending the integral to $-\infty$,
208: {\it i.e.} before the well region (where the actual values are exponentially small). The voltage
209: $\tilde{V}(\gamma )$ is defined as
210: $\tilde{V}(\gamma )= V_{0}(\gamma ) - \hbar\omega_0 /C V_{0}(\gamma )\approx V_{0}(\gamma )$.
211: It is interesting to see also what happens with the rest of the components of $|\Psi_{out}\rangle$.
212: Although they do contribute to the normalization as discussed before, they are decaying both in time and
213: away from
214: the barrier as $\exp [-\frac{\Gamma}{2}\int_{\gamma_0}^{\gamma}V_{0}^{-1}(\varphi)
215: \bar \Phi_{0}d\varphi ]$
216: which, as we will see below, would give far from the barier a factor of $\exp [-
217: \Gamma /2(t + \sqrt{2\gamma}/\omega_{p})]$. It is clear that these terms can be neglected starting
218: roughly from a time $\Gamma^{-1}$.
219: The wavefunction Eq. (\ref{out}) contains all the information about the dynamical evolution of the state of the
220: circuit containing the Josephson junction, and it is the main result of this paper.
221:
222: In a typical experiment,
223: the voltage across the junction is
224: monitored by a voltmeter at room temperature. A fundamental issue is to find a microscopic mechanism for the junction-voltmeter interaction and
225: a suitable theory of quantum measurement that would
226: model the collapse of the wavefunction; this is however beyond the scope of this paper.
227: Still, Eqs. (\ref{out}) and (\ref{prob}) give a quite clear qualitative
228: picture of what happens: the
229: particle rolls down the washboard potential with a quasi-classical speed given by energy conservation
230: $CV^{2}_{0}(\gamma )/2 = \hbar\omega_0 - U(\gamma )$. Quantum mechanics enters in the picture through the tunneling rate; we expect the results
231: of the measurements
232: to have a spread dermined by $\Gamma$.
233: One can assume that the measurement projects the outgoing state onto eigenvalues of the voltage
234: operator; therefore the
235: probability of recording the value $V$ at the moment $t$ will be given by the standard
236: quantum mechanics recipe
237: \begin{equation}
238: P(V, t) =
239: \frac{1}{2\pi\hbar}\vert \int_{-\infty}^{\infty} d(\bar\Phi_0\gamma )\Psi_{out}^{*}(\gamma ,t)\exp{(iVC\gamma /2e)} \vert ^2.
240: \label{prob}
241: \end{equation}
242: As an example, had the outside the well potential $U$ been
243: zero, we would have gotten for the charge $CV$, by performing the integration in Eq. (\ref{prob}), a
244: standard Cauchy-Breit-Wigner distribution centered around $CV_0$ and full width at half maximum
245: $\Gamma\hbar /V_0$
246: \begin{equation}
247: P(CV)= \frac{1}{\pi}\frac{\Gamma\hbar}{2 V_0}\left[ (CV-CV_0)^2 +
248: \left( \frac{\Gamma \hbar}{2V_0}\right)^2 \right]^{-1}.
249: \end{equation}
250:
251: %A better way to measure the voltage would be however through a less invasive method, such as
252: %coupling the junction capacitively to an RF-SET read-out.
253:
254: %Therefore the same phenomena happens in Eq. [\ref{out}]: there will be a spreading of the values
255: %measured by the voltmeter around the instantaneous classical values $V_{0}(\gamma (t) )$, where
256: %$\gamma (t)$ is given by solving the equation $\int_{\gamma_0}^{\gamma (t)}
257: %V^{-1}_{0}(\varphi ) \bar\Phi_0 d\varphi = t$.
258:
259: Considering again the case of a junction with a washboard potential $U(\gamma )$, we notice that
260: a good approximation is $U (\gamma ) \approx E_{J}\gamma $. This comes from the fact that switching is
261: typically observed at values of the bias current close to the critical current of the junction, as well
262: as from the observation that for times larger than $\Gamma ^{-1}$ the wavepacket is concentrated at
263: large values of $\gamma \gg 1$, in which case the $\cos\gamma$ term in the potential is negligible.
264: In other words, the particle gets soon so fast that the "speed bumps" created by the Josephson
265: effect are not slowing it down significantly. This can be checked {\it a posteriori}. A first observation
266: is that the relevant quantity for the dynamics of the center of the wavepacket is the argument of the
267: $\Theta $ function; the condition that this argument vanishes sets the maximum value
268: of $|\Psi_{out}|^2$ and gives a phase $\omega_{p}t/2\gg 1$ for $t$ larger than $\Gamma^{-1}$.
269: A legitimate concern is whether the wavefunction does not spread faster than it moves downwards. This
270: is not the
271: case, as we will see below: the spread of the wavefunction increases linearly with time, while the
272: average coordinate (phase) is advancing as $t^2$. With these observations, the normalization constant
273: can be calculated, and the outgoing wavepacket becomes
274: \begin{equation}
275: \psi^{\rightarrow}_{out} (\gamma ,t) = \sqrt{\frac{\Gamma}{\sqrt{2\gamma}\bar\Phi_{0}\omega_{p}}}\exp\left[i\frac{\hbar\omega_{p}
276: \gamma^{3/2}}{6\sqrt{2}E_{c}} - \left(i\frac{\omega_0}{\omega_p} + \frac{\Gamma}{2\omega_p}\right)\left(
277: t\omega_{p}-\sqrt{2\gamma}\right)\right]\Theta\left(t\omega_p - \sqrt{2\gamma}\right).\label{super}
278: \end{equation}
279: A plot of the wavefunction is given in Fig. \ref{wf}.
280:
281: \begin{figure}[htb]
282: \includegraphics[width=84truemm]{wavefunction.eps}
283: \caption{Contourplot $(\gamma, t)$ of the modulus of the running-phase wavefunction Eq. (\ref{super}) for $\omega_p = 60 \Gamma = 30$ GHz.}
284: \label{wf}
285: \end{figure}
286:
287: The average phase (flux) corresponding to this wavepacket can be obtained
288: \begin{equation}
289: \langle \gamma \rangle (t)= \omega_{p}^2(t^{2}/2 - \Gamma ^{-1}t + \Gamma^{-2}),\label{av}
290: \end{equation}
291: and we notice that the dominant term is quadratic in t.
292: The spread of the flux variable is given by (we keep only the dominant term here)
293: \begin{equation}
294: \sqrt{\langle \gamma^2\rangle (t) - \langle \gamma \rangle^2(t)} = \omega_{p}^{2}\Gamma^{-1}t.
295: \end{equation}
296: To get the average voltage we can use Ehrenfest theorem; we obtain
297: \begin{equation}
298: \langle V(t) \rangle = \bar\Phi_{0}\omega_{p}^{2}(t -\Gamma^{-1}). \label{avv}
299: \end{equation}
300: The dominant term for the voltage is linear in time and satisfies the classical energy
301: conservation $C\langle V\rangle ^2(t) /2 = \hbar \omega_{0} - U(\langle \gamma \rangle (t))\approx I_{0}\langle
302: \gamma \rangle (t)$.
303:
304: Let us now analyze what happens in typical switching current experiments, as they are done now in the
305: context of superconducting qubits \cite{qubit}: the bias current of the junction is increased
306: fast to a value that allows tunneling, it is kept there for a time $0 < \tau < \Gamma^{-1}$, then it is lowered
307: to a value that suppresses tunneling. This value has to be large enough so that the experimentalist can get a
308: reliable reading of voltage on the quasiparticle branch if the junction has switched; in practice, it can still satisfy
309: $I\approx I_0$. Although the change
310: of the bias current has a major effect with respect to tunneling through the barrier, where the
311: the tunneling rate decreases exponentially with the height of the barrier, from the point of view of the
312: structure of the running-phase state it amounts only to a modification of the parameter $\Gamma$.
313: Finally,
314: the current is put to zero and, after waiting long enough for retrapping to occur, the whole cycle can be
315: repeated. In our model, the essential physics is that after the time $\tau$, the tunneling matrix element $t$ is zero,
316: therefore the system evolves only under the action of $H_0$. The wavefunction is "cut" into two separate pieces,
317: one which is (almost) the bound state $\Psi_0$ inside the well, the other being the wavepacket in the continuum
318: which evolves as
319: \begin{equation}
320: |\Psi_{out} (t) \rangle = -i\sqrt{\frac{\Gamma }{2\pi}}\int d\epsilon \frac{e^{(-\Gamma /2 + i \omega_0 )\tau }
321: e^{-i\omega_{j} (t-\tau )}
322: - e^{-i\epsilon t}}{\Gamma /2 + i (\omega_{0} -\epsilon )}
323: |\psi _\epsilon^{\pm}\rangle ,
324: \end{equation}
325: with normalization $\langle \Psi_{out} (t)|\Psi_{out} (t) \rangle = 1- e^{-\Gamma \tau}$. Now, for $t-\tau >
326: \Gamma^{-1}$ we can see
327: that the outgoing function consists of two consecutive (separated by the time $\tau$) and dephased (with $\omega_0 \tau $)
328: outgoing wavepackets with the structure of $|\Psi^{\rightarrow}_{out} (t)\rangle$
329: which propagate at the same speed across the phase coordinate $\gamma$. The second wavepacket, which has a probability amplitude
330: smaller by a factor of $e^{-\Gamma /2\tau}$, results from the
331: waves localized near the barrier during the time $\tau$ when tunneling was in progress.
332: After integration over energy, we get
333: \begin{equation}
334: |\Psi_{out} (t) \rangle = e^{-(\Gamma /2 + i\omega_0)\tau }|\Psi^{\rightarrow}_{out} (t -\tau )\rangle
335: - |\Psi^{\rightarrow}_{out} (t)\rangle ,\label{xx}
336: \end{equation}
337: where $|\Psi^{\rightarrow}_{out} (t)\rangle$ is given by Eqs. (\ref{out}) and (\ref{super}). To check that
338: the normalization
339: $\langle \Psi_{out} (t)|\Psi_{out} (t) \rangle = 1- e^{-\Gamma \tau}$ remains valid, we notice that
340: in the region of overlap of the two wavepackets, which coincides with the domain where
341: $|\Psi^{\rightarrow}_{out}(t-\tau )\rangle$
342: is finite $t-\tau > \int_{\gamma_0}^{\gamma}\bar \Phi_{0} V_{0}^{-1}(\varphi)d\varphi $, there exists a very
343: simple relation between them:
344: $|\Psi^{\rightarrow}_{out} (t) \rangle = \exp[-i\omega_0\tau -\Gamma \tau /2]|\Psi^{\rightarrow}_{out} (t-\tau )\rangle$.
345: Using this property and the previous expressions Eqs. (\ref{av}) and (\ref{avv}), we can calculate the average
346: phase and voltage on the state Eq. (\ref{xx})
347: \begin{equation}
348: \langle \gamma \rangle (t) = \frac{1}{2}\omega_{p}^{2}[1- e^{-\Gamma \tau}]t^{2} -
349: e^{-\Gamma \tau}\omega_{p}^{2}\tau t - \frac{1}{2}e^{-\Gamma \tau}\omega_{p}^{2},
350: \end{equation}
351: and
352: \begin{equation}
353: \langle V \rangle (t) = \bar\Phi_{0}\omega_{p}^{2}t\left(1- e^{-\Gamma \tau }\right)
354: -\bar\Phi_{0}\omega_{p}^{2}e^{-\Gamma \tau}\tau .
355: \end{equation}
356: In Fig. \ref{prediction} we present a plot of the average voltage as a function of the time $\tau$.
357: We see that for values of $\tau$ of the same order or larger than the lifetime $\Gamma^{-1}$
358: the average voltage at t flattens, reflecting the fact that the junction has switched, as in the case
359: of Eq. (\ref{avv}).
360: \begin{figure}[htb]
361: \includegraphics[width=84truemm]{prediction.eps}
362: \caption{The average voltage as a function of the time $\tau$ for $\omega_p = 30$ GHz, $\Gamma = 1$ ns
363: and $t = 10$ ns.}
364: \label{prediction}
365: \end{figure}
366:
367: For designing an experiment to test these predictions, several remarks should be made.
368: In the case of real junctions, the Josephson energy and the plasma frequency can be reduced
369: by using a SQUID configuration and by adding capacitors in parallel with the junctions.
370: This makes the time evolution of the switching state slower and therefore easier to detect.
371: %Finally, we notice as before that we do not need to prepare the system at t=0 in the state corresponding
372: %to $I=0$; the tunneling can be suppresses to a good approximation on the time scale of the experiment
373: %(at most 40 ns as we will see below) also with a suitably large value of the critical current.
374: An important limitation on time comes from the fact that as soon as the voltage reaches the quasiparticle
375: branch (at twice the value of the gap) our analysis is not valid. The other limitation is technological:
376: even with a good dilution refrigerator,
377: thermalizing the junction is very difficult at low temperatures. With a good high-power refrigerator
378: with base temperature of about 5 mK, we assume an optimistic value of 10 mK for the effective temperature
379: of the electrons. This temperature corresponds to a crossover angular frequency of 8.66 GHz between the MQT
380: and the thermal
381: activation transition.
382: A plasma frequency of $\omega_p = 30$ GHz (zero bias current) will thus keep us safely in the MQT regime when
383: the current is raised up to about half a percent close to the critical current, according to the formula that
384: gives the plasma oscillation frequency at a finite bias current \cite{tinkham,switch}. For Nb, with gap of 1.4 meV, this
385: corresponds to a time of approximately 10 ns, as given by Eq. (\ref{avv}). A voltage increase on this timescale
386: can be detected with standard experimental techniques
387: Suppose now that we choose to work at currents about 5\% less than the critical current.
388: We still have to satisfy the condition $t>\Gamma^{-1}$; an inspection of the formula that gives
389: the tunneling rate for underdamped junctions (see {\it e.g.} \cite{switch}) shows that switching rates of about 500 MHz and more
390: (with the restriction $\Gamma \ll \omega_p$) can be achieved for $E_J /\omega_p$ of the order of 30, values which can be obtained easily with
391: large junctions.
392:
393:
394:
395: %In our calculation we have worked with an idealized model of underdamped junction, therefore
396: %neglecting completely the effect of dissipation. Although
397: %overdamped real junctions can be fabricated, the quality factor is typically limited to values of
398: %the order of 100. In the standard
399: %RCSJ model
400: %\begin{equation}
401: %C\frac{d^2}{dt^2}\left({\bar \Phi_0} \phi \right) = I - I_{0}\sin( \phi ) -\frac{1}{R}\frac{d}{dt}
402: %\left({\bar \Phi_0} \phi\right) + I_n (t) \label{rcsj}
403: %\end{equation}
404: %where we have expressed everything in term of the "flux" coordinate $ \bar \Phi_0\phi $.
405: %The value of the resistance $R$ can be determined experimentally; typical values are in the range of 100-200$\Omega$.
406: %Classically, we notice that dissipation becomes important when the friction force becomes of the order of magnitude
407: %of the force applied; the limitation is that the maximum current through the resistor when the junction reaches the
408: %quasiparticle threshold should be less than the critical current of the junction. For Nb, this gives a minimum
409: %critical current of the order of 10-20$\mu$A, which can be achieved in large junctions.
410: %A more delicate discussion refers to the role of finite temperature.
411: %The temperature manifests itself
412: %through fluctuations in the biasing current with white-noise correlations
413: %$\langle I_{n}(t)I_{n}(0)) \rangle = 2K_{B}TR^{-1}\delta (t)$;
414: %the average of this noise is zero, $\langle I_{n}\rangle = 0 $.
415: %We notice that although the condition for MQT is fulfilled, this does not guarantee that the evolution will
416: %preserve the purity of the state. Consider the quantized version of Eq. [\ref{rcsj}], the so-called quantum Lagrange equation. We introduce
417: %the analog of de Broglie thermal wavelength, which in our case has dimensions of flux:
418: %\begin{equation}
419: %\Phi_{T} = \frac{h}{\sqrt{2\pi CK_{B}T}} = \frac{2}{\sqrt{\pi}}\Phi_{0}\sqrt{\frac{E_{c}}{K_{B}T}}
420: %\end{equation}
421: %For an initial superposition of two states, the decoherence time is typically set by \cite{zurek}
422: %\begin{equation}
423: %\tau _{D} = RC \left( \frac{2\pi \Phi_{T}}{\Phi_{0}\delta \phi}\right)^2 = \frac{8\pi e^2 R}{K_{B}T(\delta\phi)^2}
424: %\end{equation}
425: %Assuming that this result sets correctly the orders of magnitude also for our problem,
426: %we see that the out-going state and the state inside the barrier will decohere fast.
427:
428: %To estimate the quantity $\delta\phi$ we consider a junction biased so that the MQT is experimentally observable
429: %(less than 1\% of the critical current). Then $\delta\phi$ is the phaseshift of an eigenfunction when the energy
430: %changes with $K_B T$. This gives $8\pi /(\delta \phi)^2 \approx 10^4$, resulting in a characteristic
431: %decoherence time $\tau_D \approx 10ns$. This is large enough for the pure-state quantum behavior of the particle to be
432: %observed. It is also consistent with the values obtained from Ramsey interference experiments with superconducting
433: %quantum bits.
434:
435: %For most of the existing experiments with switching junctions, $\Gamma ^-1$ is in fact larger than the time it takes
436: %for the switching voltage to reach the quasiparticle branch. Even if our calculations are not valid in this limit
437: %because of transient effects we see that there will be anyway a growing outgoing component
438: %of the wavepacket moving towards the quasiparticle branch.
439: %We see that in the case of real-life experiment the environment decides whether the junction has switched or not
440: %faster than the time scale $\Gamma ^-1$ at which switching itself takes place. In classical terms, this
441: %sounds almost paradoxical: the information that the junction has switched or not becomes available faster than
442: %the switching event itself. It is as if the apparatus provides the result of the measurement of a process
443: %before the process happens.
444:
445: %The time that the particle spends in the running-phase state, before reaching the
446: %quasiparticle gap, is quite short. This is due to the fact
447: %that for underdamped junctions the speed of the particle down the washboard potential increases fast. This makes
448: %ny measurement of the runing-phase state a challenging experimental problem but even so, oscilloscopes
449: %and amplifiers with bandwidths of tens of GHz exist and could be used for the observation of the pulses.
450: %However, a much better experimental strategy is to reduce the plasma frequency by using instead of
451: %a single junction a dc-SQUID system in a magnetic field. This will suppress $\omega_p$ to almost
452: %any suitable value (in practice one-two orders of magnitude). Finally, to keep the quality factor above 1,
453: %a good solution is to add a capacitor in parallel with the SQUID. With this arrangement, the exponential
454: %running-phase time can be set in the range of tens of nanoseconds, thus enabling the experimental
455: %observation.
456:
457:
458:
459:
460: %\section{Conclusion}
461:
462:
463:
464: %\section{Acknowledgements}
465:
466: G.~S.~P. was supported by an EU Marie Curie Fellowship (HPMF-CT-2002-01893); this work is also
467: part of the SQUBIT-2 project (IST-1999-10673), the
468: Academy of Finland TULE No.7205476, and the
469: Center of Excellence in Condensed Matter and Nuclear Physics at
470: the University of Jyv\"askyl\"a.
471:
472:
473:
474: \begin{thebibliography}{99}
475:
476: \bibitem{tinkham} M.~Tinkham, {\it Introduction to Superconductivity}, 2nd ed. (McGraw-Hill
477: Inc., New York, 1996).
478:
479: \bibitem{switch}J.~Clarke, A.~N.~Cleland, M.~H.~Devoret, D.~Esteve, and J.~M.~Martinis,
480: Science {\bf 239}, 992 (1988).
481:
482: \bibitem{suppl} A.~J.~Leggett, Suppl. Progr. Theor. Phys. {\bf 69}, 80 (1980).
483:
484: \bibitem{deltapotential} R.~G.~Winter, Phys. Rev. {\bf 123}, 1503 (1961);
485: C.~B.~Chiu, E.~C.~G.~Sudarshan, and
486: B.~Misra, Phys. Rev. D {\bf 16}, 520 (1977); S.~De~Leo and P.~P.~Rotelli, quant-ph/0401145.
487:
488: \bibitem{cond} J.~Bardeen, Phys.~Rev.~Lett. {\bf 6}, 67 (1961);
489: M.~Galperin, D.~Segal, and A.~Nitzan, J.~Chem.~Phys. {\bf 111}, 1569 (1999).
490:
491: \bibitem{josephson}K.~S.~Chow, D.~A.~Browne, and V.~Ambegaokar, Phys. Rev. B {\bf 37}, 1624 (1988);
492: S.~Takagi, {\it Macroscopic Quantum Tunneling} (Cambridge University Press, Cambridge, 2002).
493:
494: \bibitem{twopotential} S.~A.~Gurvitz and G.~Kalbermann, Phys.~Rev.~Lett. {\bf 59}, 262 (1987);
495: S.~A.~Gurwitz, Phys.~Rev.~A {\bf 38}, 1747 (1988).
496:
497: \bibitem{kurizkikofman} A.~Barone, G.~Kurizki, and A.~G.~Kofman, Phys.~Rev.~Lett. {\bf 92}, 200403
498: (2004).
499:
500: \bibitem{qubit}
501: %Y.~Nakamura, Y.~A.~Pashkin, and J.~S.~Tsai,
502: %Nature, {\bf 398}, 786 (1999);
503: %J.~E.~Mooij, T.~P.~Orlando, L.~Levitov, L.~Tian,
504: %C.~H.~van~der~Val, and S.~Lloyd, Science {\bf 285}, 1036 (1999);
505: %J.~R.~Friedman, V.~Patel, W.~Chen, S.~K.~Tolpygo, and J.~E.~Lukens, Nature {\bf 406},
506: %43 (2000);
507: D.~Vion, A.~Aassime, A.~Cottet, P.~Joyez, H.~Pothier, C.~Urbina,
508: D.~Esteve, and M.~H.~Devoret, Science {\bf 296}, 886 (2002); I.~Chiorescu, Y.~Nakamura,
509: C.~J.~P.~M.~Harmans, and J.~E.~Mooij, Science {\bf 299}, 1869 (2003); J.~Claudon, F.~Balestro, F.~W.~J.~Hekking, and
510: O.~Buisson, Phys. Rev. Lett. {\bf 93}, 187003 (2004); J.~Sjostrand, J.~Walter,
511: D.~Haviland, H.~Hansson, A.~Karlhede, cond-mat/04066510.
512: %; Yu.~A.~Paskin, T.~Yamamoto,
513: %O.~Astafiev, Y.~Nakamura, D.~V.~Averin, and J.~S.~Tsai, Nature {\bf 421}, 823 (2003).
514:
515:
516:
517: %\bibitem{zurek} W.~H.~Zurek, in {\it Frontiers of Nonequilibrium Quantum Statistical Physics},
518: %NATO ASI vol. B 135, pp. 145-149, G.~T.~Moore and M.~O.~Scully, eds., (Plenum 1986); quant-ph/0302044.
519:
520:
521: \end{thebibliography}
522:
523: \end{document}
524: