1: \documentclass{article}
2: \usepackage{epsfig}
3:
4: \renewcommand{\baselinestretch}{1.33}
5: \newcommand{\be}{\begin{equation}}
6: \newcommand{\ee}{\end{equation}}
7:
8: \begin{document}
9: \title{Dynamics of point Josephson junctions in a microstrip line}
10: \author{ J.G. Caputo and L. Loukitch \\
11: {\normalsize \it Laboratoire de Math\'ematiques, INSA de Rouen} \\
12: {\normalsize \it B.P. 8, 76131 Mont-Saint-Aignan cedex, France }
13: \date{\today}}
14:
15: \maketitle
16: \begin{abstract}
17: We model the dynamics of point Josephson junctions in a 1D microstrip
18: line using a wave equation with delta distributed sine nonlinearities.
19: The model is suitable for both low T$_c$ and high T$_c$ systems (0 and
20: $\pi$ junctions).
21: For a single junction in the line, we found two limiting behaviors: the
22: ohmic mode where the junction acts as a pure resistor which stops
23: waves and separates the cavity and the junction mode where the wave
24: is homogeneous throughout the strip. This classification allows to bound the
25: IV curves of the system. Two junctions in a strip give generally ohmic
26: modes and combined junction/ohmic modes and yield information about the
27: behavior with an array with many junctions. Finally we use this analysis
28: to understand the many junction case for 0 and $\pi$ junctions and the
29: effect of an external magnetic field.
30: \end{abstract}
31:
32: \section{Introduction}
33:
34: Type I Superconductors are characterized by the phase of a complex
35: number, the macroscopic order parameter. Josephson \cite{josephson}
36: showed that when
37: two such superconductors are very close, electrons can tunnel from one
38: to the other according to the relations
39: \be\label{joseph}
40: V = {\Phi_0 \over 2 \pi}{d \phi \over dt}~~, ~~ I = s J_c
41: \sin(\phi)~,\ee
42: where $V$ and $I$ are respectively the voltage and current across
43: the barrier, $\phi$ is the difference between the phases in
44: the two superconductors, $s$ is the contact surface, $J_c$ is the
45: critical current density and
46: $\Phi_0$ is the flux quantum.
47: The two Josephson relations together with Maxwell's equations
48: imply the modulation of DC current by
49: an external magnetic field in the static regime (SQUIDs) and the
50: conversion of AC current into microwave radiation.
51:
52:
53: The ratio of the electromagnetic flux to
54: the flux quantum gives the Josephson length, the typical distance of
55: variation of the phase $\lambda_J = { \Phi_0 \over L J_c}$.
56: Depending whether their length $l$ is larger or not than $\lambda_J$,
57: Josephson junctions exhibit significantly different behaviors.
58: Small junctions for which $l < \lambda_J$ are used as SQUIDS to measure
59: magnetic fields in the static regime or as signal mixers where the
60: steep IV characteristic at the gap is used to downshift an
61: unknown high frequency signal by combining it with a known local
62: oscillator signal.
63: In this latter context Salez et al \cite{Salez} suggested to
64: use several unequally spaced junctions
65: , instead of just one, to get a better frequency response
66: over a large frequency band. Long Josephson junctions
67: for which $l >> \lambda_J$ are described by the 1D
68: sine-Gordon equation
69: \be\label{sg1d}
70: \phi_{tt} - \phi_{xx} + \sin (\phi)=0 ,\ee
71: which admit the kink solutions varying from 0 to $2 \pi$.
72: Long junctions have a characteristic dependence
73: of the maximum critical current in the presence of magnetic field
74: because of the screening of the applied field by the supercurrent.
75: In the dynamical regime long junctions are used as microwave
76: generators but provide small output power and poor impedance
77: matching to the transmission line. To improve on this, series arrays
78: of junctions have been used, with success since they have been
79: shown to synchronize
80: in the presence of external forcing \cite{bcsl99} \cite{vbsl02} giving
81: rise to an $N^2$ power output for $N$ junctions. Such behavior
82: has been understood both from the quantum point of view \cite{as01}
83: and the Resistive Shunted Junction (RSJ) equivalent circuit model
84: \cite{fp02}. The cavity surrounding the junctions
85: stores radiation which can increase the output power of
86: the device.
87:
88: The mathematical description of such a system is a wave equation for
89: the cavity coupled to the 2D version of the sine-Gordon equation
90: (\ref{sg1d}) via interface conditions.
91: In \cite{bcf02,bc02} one of the authors
92: studied this coupling for a long junction with a small surrounding passive
93: region. When the cavity is present
94: only at the junction tip and the miss-match at the interface is
95: not too large, the kink crosses it unaffected and adapts its
96: speed. When the passive region surrounds the junction, it sets the
97: kink speed. This ballistic kink motion will be absent for small
98: junctions so in the present work we consider the
99: coupling of a large cavity to one or a few small junctions. We want to
100: understand how junctions can talk to each other via the surrounding
101: passive region. Two approximations are assumed, we consider
102: the 1D limit where transverse modes of the microstrip do not play a role
103: and neglect the phase variation inside the junctions, in other words
104: we neglect the junction length. This is valid for an array of short
105: junctions whose length remains smaller than the wave length of the
106: radiation.
107:
108: This approach can be applied to high Tc superconductors where due to
109: a different symmetry of the wave function of the Cooper pairs one has
110: current phase relations that involve the higher harmonics \cite{lcti03}.
111: This different symmetry can also induce a negative coupling in (\ref{joseph})
112: leading to so-called $\pi$ junctions. Normal (0) and $\pi$ junctions can
113: be associated to form 0-$\pi$ junctions which have semifluxon solutions
114: that vary between 0 and $\pi$ and are attached to the discontinuity.
115: In the static regime the maximum DC current going through the system
116: is strongly modulated for a short junction like for a SQUID while it
117: is constant for small field when the junction is longer \cite{l04}. The
118: IV characteristics give resonances that are similar to zero field steps
119: \cite{l04,gsgk04}. Note however the difference between the ballistic motion
120: of a kink in a (0) long Josephson junction and the hopping of semi-fluxons
121: in a (0-$\pi$) junction. As for 0 junctions, (0-$\pi$) junctions can
122: be organized as arrays \cite{r03,sg04} and these have interesting paramagnetic
123: behavior.
124:
125: The model we introduce allows to describe accurately the coupling between
126: small junctions in such an array. Both 0 and $\pi$ junctions can be considered
127: and disorder can be added. For simplicity we have considered the large
128: damping limit where it is possible to develop the analysis. We studied in
129: detail the cases of one and two junctions that remain fairly simple
130: and used our conclusions for the case of many junctions. For a single
131: junction, we found two limiting behaviors: the ohmic mode where the junction
132: acts as a pure resistor which stops waves and separates the cavity and
133: the junction mode where the wave is homogeneous throughout the microstrip.
134: After introducing the model is section 2, we consider these limiting
135: behaviors in sections 3,4 and 5.
136: These allow to understand in detail the features of the IV characteristics
137: (section 6). The behavior of two junctions shown in section 7
138: follows generally ohmic modes
139: and combined junction/ohmic modes. Using this, in section 8 we
140: generalize the study to
141: many junctions embedded in a microstrip and understand the influence of
142: the sign of the current density (0 or $\pi$ junctions) and external
143: magnetic field. Conclusions are presented in section 9.
144:
145:
146: \section{The model}
147:
148: The device we model shown in the top panels of Fig. \ref{f1} (top view
149: on the left and section on the right) is a narrow microstrip in
150: which are embedded a number of small junctions in the window
151: design. We proceed from the description of the two superconductors
152: as two inductance arrays connected by capacitive elements (resp. and
153: a resistor and nonlinear Josephson element) in the passive (resp. junction)
154: region. The equivalent circuit is shown in the bottom panel of Fig.
155: \ref{f1}. In the following subsections we show that the continuum limit
156: yields an inhomogeneous 2D sine-Gordon equation which we reduce to 1D
157: when the microstrip is narrow. For small junctions we then justify our delta
158: function approach (\ref{sgd}).
159:
160:
161: \subsection{RSJ model for distributed Josephson junctions}
162:
163: A simple mathematical model of the window Josephson junction is to
164: describe
165: each superconductor by an array of inductances $L$. The coupling
166: elements
167: between two adjacent nodes in each array are, a capacitor $C$,
168: resistance $R$
169: and Josephson current $I_c$ \cite{Likharev}. The Kirchoff laws at
170: each couple of nodes $(i_b,i_t)$ in the bottom and top superconducting
171: layers can be
172: combined to give the relation expressing the conservation of currents
173: at node $i$ in the device
174: \begin{equation}
175: \label{discr}
176: C_i \ddot \Phi_i + \sum_j \frac {\Phi_i - \Phi_j}{L} +
177: I_i^c \sin \frac {\Phi_i}
178: {\Phi_0} + \frac {\dot \Phi_i}{R_i} = 0 \; ,
179: \end{equation}
180: where $\Phi = \Phi_t - \Phi_b$ (resp. $\Psi = \Psi_t - \Psi_b$),
181: is the phase difference between the two superconductors
182: in the junction (resp. passive) part, the summation $\sum_j$ is
183: applied to
184: the nearest neighbors, $I_i^c$ is the critical current which is non
185: zero
186: only in the junctions and $R_i$ is the resistance associated to the
187: current of quasi-particles again non zero
188: only in the junctions. In agreement with experiments, we have
189: assumed the same surface inductance for
190: the whole sample but have taken into account the variation of the
191: capacity
192: $C$ and resistance $R$ from the junctions to the rest of the cavity.
193: Note that equation (\ref{discr}) is a discretisation of Maxwell's
194: equations
195: (wave equation part) and Josephson constitutive equations (sine term),
196: assuming an electric field normal to the plates, a magnetic field in
197: the
198: junction plane and perfect symmetry
199: between the top and bottom superconducting layers. We can now obtain
200: the model
201: in the continuum limit, more suitable for analysis.
202:
203: \subsection{Continuum limit}
204:
205: The continuum version of the system (\ref{discr})
206: can be derived by introducing the following quantities
207: per unit area ($a^2$) of elementary cells
208: of length $a$.
209: \begin{equation}
210: \label{normcjr}
211: \overline {C_{J}}=\frac {C_{J}} {a^2},~~~j_i^c= \frac {I_i^c} {a^2},~~~
212: \overline{r_i} = R_i a^2.~~~
213: \end{equation}
214: We normalize the phases by the flux quantum $\Phi_0$,
215: \begin{equation}
216: \label{normphi}
217: \varphi_i=\frac {\Phi_i}{\Phi_0}
218: \end{equation}
219: and introduce the Josephson characteristic length
220: \be \label{lambdaj}
221: \lambda_J^2=\frac {\Phi_0}{j_c L_J} .
222: \ee
223: Notice that in this 2D problem the inductance associated to each cell
224: is equal to the branch inductance $L$.
225: To differentiate the junction regions from the passive regions we
226: introduce the indicator function $g_i$ such that $g_i=1$ in the
227: junctions and $0$ elsewhere. We can then write
228: \be\label{indic}
229: j_i^c = g_i j^c~~,~~
230: {1\over \overline{r_i}}=g_i{1\over \overline{r}}
231: ~~{\rm and} ~~
232: {\bar C_i}= g_i{\bar C_J} +(1-g_i){\bar C_I}
233: \ee
234: where $j^c$, $\overline{r}$, ${\bar C_J}$ and ${\bar C_I}$
235: are respectively the critical current densities, the quasi
236: particle resistance in the junction, the capacitance per unit area in
237: the junction and the capacitance per unit area in
238: the passive region.
239:
240: We substitute relations (\ref{normcjr}) - (\ref{indic}) in
241: (\ref{discr})
242: and obtain
243: \begin{equation} \label{discre2}
244: L {\overline C_I} \ddot \varphi_i
245: + \sum_j \frac {\varphi_i - \varphi_j}{a^2}
246: + g_i \left [ L({\overline C_J}-{\overline C_I}) \ddot \varphi_i
247: + {1 \over \lambda_J^2} \sin \varphi_i
248: +\frac {L}{\overline{r}} \dot \varphi_i \right ] = 0 \; ,
249: \end{equation}
250: which in the continuum limit yields
251: \begin{equation} \label{cont1}
252: L{\overline C_I} \varphi_{tt}
253: -\Delta \varphi
254: + {1 \over \lambda_J^2}g(x,y) \left [
255: L \lambda_J^2 ({\overline C_J}-{\overline C_I}) \varphi_{tt}
256: + \sin \varphi
257: + \frac {L \lambda_J^2}{\overline{r}} \varphi_t \right ]=0 .
258: \end{equation}
259: To obtain the final equation we rescale time by the inverse of the
260: plasma frequency $\omega_I^{-1}= \lambda_J \sqrt{ L {\overline C_I}} $
261: so that $\tilde t = \omega_I t$. As unit of space we use the
262: Josephson length,
263: $ \tilde x = x/\lambda_J,~\tilde y =y/\lambda_J $.
264: The final equation is then
265: \be
266: \label{insg}
267: \varphi_{tt} -\Delta \varphi
268: + g(x, y)( c \varphi_{tt} + \sin \varphi + \alpha \varphi_{t})=0 ~,
269: \ee
270: where the
271: tildes have been omitted for simplicity
272: and where the coefficients $c$ and $\alpha$ are
273: \be\label{calpha}
274: c = {{\overline C_J}\over{\overline C_I}} -1~~,~~
275: \alpha = {\lambda_J \over {\bar r} }\sqrt{L \over {\overline C_I}}\ee
276:
277: The boundary conditions represent the input of an external
278: current $I$ or magnetic field $b$ on the device. They are
279: \be\label{overlap}
280: \varphi_x|_{x=0}=b -(1-\nu){I\over 2 w}~~,~~
281: \varphi_x|_{x=l}=b +(1-\nu){I\over 2 w}~~,~~
282: \varphi_y|_{y=0}= -\nu {I\over 2l}~~,~~
283: \varphi_y|_{y=w}= \nu {I\over 2l},\ee
284: where $0<\nu<1$. The case $\nu=1$ corresponds to a pure
285: inline current feed and $\nu=0$ to a pure overlap current feed.
286: We will mostly consider the latter case in the discussion.
287:
288:
289: \subsection{Reduction to a 1D problem}
290:
291:
292: When the microstrip is narrow and the junctions occupy the
293: whole width we can assume that $g$ is independent of $y$.
294: We can then write
295: \be \varphi(x,y,t) = \nu I {y^2 \over 2s}
296: + \sum^\infty_{n=0} \phi_n(x,t) \cos \left({n \pi y \over w}\right),
297: \ee
298: where the first term takes care of the current feed and the second
299: satisfies homogeneous Neumann boundary conditions for $y=0,w$.
300:
301:
302: The calculations detailed in \cite{bcf02} show that for
303: a narrow width and small current only the first mode $\phi_0$
304: needs to be taken into account. We then obtain the following
305: equation for $\phi_0$ where the 0 has been dropped for
306: simplicity
307: \be\label{final1d}
308: {\phi}_{tt} -{\phi}_{xx}
309: + g(x)( c {\phi}_{tt} + \sin {\phi} + \alpha {\phi}_{t})=\nu{I \over s}
310: ~,
311: \ee
312: with the boundary conditions
313: ${{\phi}_x}_{x=0}= -(1-\nu){I l \over 2 s}$ ~~
314: ${{\phi}_x}_{x=l}= (1-\nu){I l \over 2 s}$.
315: In the following we will write
316: \be\label{curdens}
317: j = {I \over s},\ee
318: and assume an overlap current feed ($\nu=1$) except in section 7.
319:
320: \subsection{The delta function model}
321:
322: As the area of the junction is reduced the total super-current is
323: reduced and tends to zero. To describe situations where the area
324: is small so that the variations of the phase can be neglected in the
325: junction but the supercurrent is significant we introduce the
326: following delta function model.
327: Introduce the function $g(x) = d_a g_h(x)$\\
328: $g_h(x) = \left\{
329: \begin{array}{l r}
330: 1/(2h) & a-h<x<a+h \\
331: 0 & {\rm elsewhere} \\
332: \end{array} \right.$
333:
334: \begin{equation} \label{sggh}
335: \phi_{tt}- \phi_{xx} + d_a g_h(x) \left(\sin(\phi)
336: + \alpha\phi_t + c \phi_{tt} \right) =j.
337: \end{equation}
338: In the following we assume $\phi$ to be $C^2$ in $x$ and $t$ and
339: $\phi_{ttx}$ to be continuous.
340: We now show that equation (\ref{sggh}) converges when
341: $h\rightarrow 0$ to
342: \begin{equation} \label{sg2}
343: \phi_{tt}- \phi_{xx} + d_a \delta(x-a)\left(\sin(\phi)
344: + \alpha\phi_t + c \phi_{tt} \right) = j.
345: \end{equation}
346: For that introduce the Taylor expansion of $\phi$ near $x=a$
347: \begin{equation}\label{taylor}
348: \phi(x,t)=\phi(a,t) + (x-a)\phi_x(a,t) + O((x-a)^2) .
349: \end{equation}
350: In the following for simplicity of writing we will omit the
351: $t$ and $x$ dependencies. We integrate (\ref{sggh}) from $a-h$ to $a+h$
352: (see details in appendix: Continuum limit). The
353: first term
354: $$\int_{a-h}^{a+h} \phi_{tt} dx \rightarrow O ~~{\rm when} ~~h
355: \rightarrow 0, $$
356: the second one
357: $$\int_{a-h}^{a+h} \phi_{xx} dx=\phi_x(a+h)-\phi_x(a),$$
358: and the third one
359: $$\lim_{h\rightarrow 0}{d_a \over 2 h}\int_{a-h}^{a+h}
360: (\sin(\phi)+ \alpha\phi_t + c \phi_{tt})dx = d_a
361: \left( \sin(\phi(a))+ \alpha\phi_t(a) + c \phi_{tt} (a) \right),$$
362: when $h \rightarrow 0$ which is consistent with the model
363: \begin{equation} \label{sgcd}
364: \phi_{tt}- \phi_{xx} + d_a \delta(x-a) \left( \sin(\phi) + \alpha\phi_t
365: +
366: c \phi_{tt} \right)= j,
367: \end{equation}
368: together with the boundary conditions
369: $\phi_x\left (x=0\right)=0~,~~~
370: \phi_x\left (x=l,t\right)=0.$
371:
372: In the following we will consider the case $c=0$ for simplicity. We
373: assume the
374: same capacity in the junction and in the passive region
375: so that the final equation is
376: \be\label{sgd}
377: \phi_{tt}- \phi_{xx} + d_a \delta(x-a) \left( \sin(\phi) + \alpha\phi_t
378: \right)= j.
379: \ee
380: This equation can be considered as a 1D model of a point Josephson
381: junction embedded in a microstrip cavity.
382:
383: \section{The single junction case}
384:
385: The partial differential equation (\ref{sgd}) involves a distribution,
386: however
387: its solution is differentiable except at the junction points.
388: To solve it numerically we integrate the equation
389: over reference intervals using the finite volume approach.
390: The resulting system of ordinary differential equations is then solved
391: using the DOPRI5 4th and 5th order Runge-Kutta variable step
392: integrator developed by Hairer and Norsett \cite{hairer}.
393: Details of the numerical code are given in the second part of the appendix.
394:
395: We now consider the two limiting cases of a static solution and
396: high voltage for which simple expressions of
397: the solution can be obtained. In particular these have been used to
398: validate our numerical procedure.
399:
400: \subsection {Static behavior}
401:
402: We assume $\phi_t = \phi_{tt} \equiv 0$ in (\ref{sgd}) and get
403: \begin{eqnarray}
404: \label{statique}
405: -\phi_{xx} + \delta (x-a) d_a \sin (\phi) = j \\
406: \phi_x|_{x=0,l} =0
407: \end{eqnarray}
408: For $x\neq a$, $-\phi_{xx}=j $ so that for
409: $x<a$, $\phi= \phi_l \equiv -\frac{j}{2}x^2 + C_l x + D_l$ and for
410: $x>a$, $\phi= \phi_r \equiv -\frac{j}{2}x^2 + C_r x + D_r$.
411: Integrating (\ref{statique}) over the whole domain we get the
412: conservation of current
413: \be\label{conscur} d_a \sin(\phi(a))= j l,\ee
414: so that no solution exists for $|jl| > d_a$.
415:
416: The solution is obtained as usual by assuming continuity of the
417: phase and using the jump condition on the gradient obtained
418: by integrating on a small domain including
419: the junction
420: \be\label{jump}[\phi_x]_{a^-}^{a^+}=d_a \sin(\phi(a)).\ee
421: Using these two constraints we obtain the static solution
422: \begin{equation}
423: \phi_s(x)= \left\{\begin{array}{lr}
424: \frac{j}{2}(a - x)(a + x) + \arcsin \left ({j l \over d_a}\right) &
425: x<a\\
426: \frac{j}{2}((a - x)(x+a+2l) + \arcsin\left({j l \over d_a} \right) &
427: a<x
428: \end{array}\right .\end{equation}
429:
430: \subsection{High voltage behavior}
431:
432: Since the Josephson current is bounded, its influence in (\ref{sgd})
433: will decrease as the current $j $ is increased. In this limit
434: of high current the phase is growing very fast so that $\phi_t$
435: is close to the average voltage $V\equiv \left<\phi_t\right>_t$.
436: We can write approximately $\phi(x,t)= V t +\phi_v(x)$ so that
437: we can average (\ref{sgd}) assuming $\left<\phi_t(x,t)\right>_t = V$,
438: $\left<\sin(\phi(a))\right>_t=0$ and $\left<\phi_{tt}\right>_t=0$,
439: yielding
440: \be \label{hvolt}
441: -{\phi_v}{xx} + \delta\left(x-a\right)d_a \alpha V= j ,\ee
442: with the boundary conditions ${\phi_v}_x|_{x=0,l}=0$.
443: Integrating this equation over the whole domain
444: we get $V= {j l \over d_a \alpha}$.
445:
446: Using the same constraints as in the static case (continuity of the
447: phase and jump of its gradient at the junction) we get up to
448: a constant the high voltage solution $\phi_v$
449: \begin{equation}
450: \phi_v(x,t)= \left\{\begin{array}{lr}
451: {j\over 2 }(a-x)(a+x) + Vt + C & x<a \\
452: {j\over 2 }(a-x)(a+x-2l)+ Vt + C & a<x
453: \end{array}\right .\end{equation}
454: Notice that this and precedent solutions have the same spatial
455: behavior;
456: $\phi_s-\phi_v = Vt + C$. We will not find this result with more
457: junctions
458: (except in particulary cases).
459:
460: \subsection{Fourier representation }
461:
462: Because of the homogeneous Neumann boundary conditions one can write the
463: solution as a cosine Fourier series
464: \begin{equation}\label{four}
465: \phi(x,t) = \sum_{n=0}^{+\infty} A_n(t)\cos\left(\frac{n\pi x}{l}
466: \right).
467: \end{equation}
468: The solution of (\ref{sgd}) has a discontinuous first spatial
469: derivative to that the amplitude of the mode $A_n$ decreases as $1/n^2$
470: \cite{carslaw}. This means that in general many modes are needed
471: for the description of the solution. However since the system is
472: close to a linear one we will see that these modes provide insight into
473: the limiting behaviors.
474:
475:
476: Plugging (\ref{four}) into (\ref{sgd}) and projecting we get the
477: evolution of the modes,
478: \begin{equation}\label{a0}
479: l {A_0}_{tt} + d_a (\sin(\phi_a) + \alpha \phi_{at})
480: -j l =0 ,\end{equation}
481: \begin{equation}\label{an}
482: {A_n}_{tt} + \left({n \pi \over l}\right)^2 A_n
483: + {2 d_a \over l } c_n ( \sin(\phi_a) + \alpha \phi_{at})
484: =0 , \end{equation}
485: where we have introduced the coupling coefficients
486: associated to the junction at $x=a$,
487: $c_n=\cos\left(\frac{n\pi a}{l} \right)$.
488: Equations (\ref{a0}) and (\ref{an}) are coupled through the
489: term $\sin(\phi_a) + \alpha {\phi_a}_t$. Due to this particular
490: feature, two limiting behaviors can be expected.
491: \begin{itemize}
492: \item The ohmic mode for which
493: \be\label{omode}
494: d_a \alpha \phi_a|_t - j l=0 ~.\ee
495: Then the equations for the $A_n$ have a right hand side so that
496: the cavity modes are driven by the junction.
497:
498: \item The junction mode such that
499: \be \label{jmode}
500: d_a ( \sin(\phi_a) + \alpha {\phi_a}_t) -j l =0 ~.\ee
501: In this case the equations for the $A_n$ have a constant right hand
502: side
503: ($= j l$). We will show that in this case the cavity is synchronized
504: with the junction.
505: \end{itemize}
506:
507: The equations giving the evolution of $\phi_a$ in these two
508: limiting behaviors can be solved yielding for the ohmic mode
509: \be\label{som}
510: \phi_a = { j l \over d_a \alpha} t + C_1 ~, \ee
511: where $C_1$ is a constant and a voltage
512: \be\label{vohm}
513: V_o = { j l \over d_a \alpha}.\ee
514: For the junction mode the
515: equation can be solved using separation of
516: variables and we get
517: \be \label{sjm} \phi_a =2\arctan\left\{
518: {d_a \over j I } \left[ \alpha V_j \tan
519: \left( {V_j \over 2 } \left(t+C_2\right)
520: \right) +1 \right] \right\}~~,\ee
521: where the junction mode voltage $V_j$ is
522: \be\label{vjon}
523: V_j = {1 \over \alpha}\sqrt{\left( {j l \over d_a}\right)^2 -1 }\ee
524: where $C_2$ is a constant.
525:
526: Fig. \ref{f2} shows these two limiting behaviors for a junction
527: placed in $a=-1.57$ with $d_a=1$ $l=10$ and a current $j\approx
528: 0.1374$.
529: The numerical solution is presented with the crosses while
530: the analytical estimates (\ref{som}) and (\ref{sjm}) are shown
531: in continuous line.
532: The left panel corresponds to the ohmic mode and one can
533: see the excellent agreement of the numerical solution with the
534: line given by (\ref{som}). The right panel presents
535: the junction mode and the extension of the estimate (see below), again
536: the
537: agreement is very good.
538: Notice that for the same time interval the variation of $\phi_t(a,t)$
539: is different for the two panels of Fig. \ref{f2} so that we will get
540: a different average voltage for these two behaviors.
541:
542:
543: \section{The ohmic mode}
544:
545: In this case $\phi_t(a,t)=V_o={j l \over \alpha d_a}$. To get
546: the field we need to solve on each subdomain $[0,a[$ and $ ]a,l]$
547: the following problems
548: $$\begin{array}{cc}
549: \left\{\begin{array}{l}
550: \phi_{tt}-\phi_{xx}= j \\
551: \left .\phi_{x}\right|_{0}=0,~
552: \left .\phi_{t}\right|_{a}=\frac{j l }{\alpha d_a}
553: \end{array}\right.&
554: \left\{\begin{array}{l}
555: \label{onde1}
556: \phi_{tt}-\phi_{xx}= j \\
557: \left .\phi_{t}\right|_{a}=\frac{j l }{\alpha d_a},~
558: \left .\phi_{x}\right|_{l}=0 .
559: \end{array}\right.\\
560: \end{array}$$
561: Consider the left subdomain $[0,a[$, we introduce
562: $\phi\left(x,t\right)=v\left(x,t\right)+\frac{ j l t}{\alpha d_a}$
563: so that if $\phi$ solves the left problem then $v$ solves
564: \begin{eqnarray}
565: \label{onde2}
566: v_{tt}-v_{xx}= j \\
567: \left. v_x \right |_{0}=0~\left. v_t\right|_{a}=0
568: \end{eqnarray}
569: A particular solution of (\ref{onde2}) is $v_l=
570: -\frac{j}{2}x^2 + D_l$ where
571: $D_l$ is an integration constant.
572: Then $w= v -v_l$ solves the homogeneous wave equation
573: $$\partial_{tt} w-\partial_{xx} w=0 , $$
574: with $w_x|_0=0$ and $w_t|_a=0$.
575: We can assume without loss of generality that
576: $w\left(x,0\right)=0$.
577: The condition $w_x|_0 =0$ imposes a solution
578: of the type
579: \begin{equation}
580: w\left(x,t\right)=a_l\sin\left(V_l t\right)
581: \cos\left(V_l x\right) ,
582: \end{equation}
583: where $V_l$ is a constant. Also, $w_t|_{a}=0$ imposes a
584: condition on $V_l$ because
585: $$ w_t\left(x=a,t\right) =a_l \sin\left(V_l t\right) \cos(V_l a)=0$$
586: is possible for all $t$'s only if $\cos\left(V_l a\right)=0$. So
587: that we get $$V_l={ 2k_l+1 \over 2 }\frac{\pi}{a}.$$
588: The problem is identical for $x \in \left]a,l\right]$
589: and using similar arguments we
590: get $v_r=-\frac{j}{2}\left(x-l\right)^2 + D_r$ and
591: $$V_r={2k_r+1 \over 2} \frac{\pi}{a-l} .$$
592: We obtain a solution for the equation (\ref{sgd}).
593: The phase on the whole domain is then
594: \begin{equation}
595: \label{totalcos}
596: \phi\left(x,t\right) = \left\{ \begin{array}{l r}
597: \frac{j l t}{\alpha d_a} + a_l \sin\left(V_l t\right)\cos\left(V_l
598: x\right) + \frac{j}{2} (a-x)(a+x) & x<a \\
599: \frac{j l t}{\alpha d_a} + a_r \sin\left(V_r t\right)\cos\left(V_r
600: \left(x-l \right) \right) + \frac{j}{2} (a-x)(a+x-2l) & a<x
601: \end{array} \right. \end{equation}
602: From the problem (\ref{sgd}) integrating the equation on a small
603: interval
604: centered on $x=a$ and using (\ref{omode}) we get the jump condition
605: $$[\phi_x]_{a^-}^{a^+} -j l
606: = d_a\sin(\phi(a,t)) = d_a\sin (\frac{j l t}{\alpha d_a}).$$
607: On the other hand from (\ref{totalcos}) we get
608: $$\begin{array}{ccl}
609: \left[\phi_x\right]_{a^-}^{a^+} -j l &=& a_l
610: \sin\left(V_l t\right)
611: \sin \left( V_l a\right)-a_r V_r \sin\left(V_r t\right)
612: \sin ( V_r\left(a-l\right)) \\
613: & \equiv & C_1 \sin\left(V_l t\right)-C_2 \sin\left(V_r t\right)\\
614: \end{array}$$
615: Since $C_1$ and $C_2$ are independent of $t$, this implies that
616: $\sin(V_l t)$, $\sin(V_r t)$ and $\sin({j l t\over \alpha d_a})$ have
617: the
618: same period. We obtain,
619: \be \label{VleftVright}
620: { 2k_l+1 \over 2 }\frac{\pi}{a}= {j l \over \alpha d_a}=
621: {2k_r+1 \over 2} \frac{\pi}{a-l}=V,\ee
622: if $a_l$ and $a_r$ are not equal to zero. Then one can write the
623: solution as
624: \be\label{leftaright}
625: \phi(x,t) = \left\{ \begin{array}{l r}
626: \phi_v(x,t) + a_l \sin\left(V t\right)\cos\left(V x\right) & x<a \\
627: \phi_v(x,t) + a_r \sin\left(V t\right)\cos\left(V \left(x-l \right)
628: \right) & a<x.
629: \end{array}\right .\ee
630: where $ \phi_v$ is the high voltage solution. Because of this the
631: instantaneous
632: voltage $\phi_t$ has a discontinuous first derivative at $x=a$.
633:
634:
635: Assuming that $\alpha=1$ we see that the consistency equation
636: (\ref{VleftVright})
637: implies
638: $$(a-l) 2 k_l - 2a k_r = l.$$
639: There are very few solutions for $a$ rational et $l$ integer and for
640: most $a$ there are none. Let us look at other possibilities. Notice
641: first
642: that space homogeneous functions ($a_l=0$ or $a_r=0$) are solutions.
643: Having said this
644: it is easy to see that one can obtain left only or right only
645: oscillations. \\
646: left only: $\phi(x,t) = \left\{ \begin{array}{l r}
647: \phi_v(x,t) + a_l \sin(V t)\cos(V x) & x<a \\
648: \phi_v(x,t) & a<x
649: \end{array} \right .$\\
650: right only:
651: $\phi(x,t) = \left\{ \begin{array}{l r}
652: \phi_v(x,t) & x<a \\
653: \phi_v(x,t) + a_r \sin(V t)\cos\left(V\left(x-l \right)\right) & a<x
654: \end{array} \right .$\\
655: Again the instantaneous voltage at the junction has a discontinuous
656: first derivative.
657:
658: Fig. \ref{f3} shows the instantaneous voltage as a function of $x$ for
659: different times for two such ohmic modes, the left only oscillation
660: (left panel)
661: and right and left oscillations (right panel). The left panel shows
662: $\phi_t(x,t)$ for 60 values of time equidistributed between 0 and 10.
663: The parameters are the same as for Fig. \ref{f2}. In this case we have
664: $k_l=1$
665: so that about 1.5 periods exist to the left of $x=a$. Notice how the
666: amplitude
667: of $\phi_t$ is practically constant and equal to the average voltage
668: $V=j l =1.37$ on the right of $x=a$. The right panel corresponds to a
669: right and left oscillation for $a=2$ and we show three consecutives
670: times,
671: $t=0, 0.125$ and $0.250$. Here ${ 2 k_l+1 \over 2 k_r +1} = {a \over
672: l-a}=3/7$
673: so that we get about 1 period to the left of $x=a$ and two periods to
674: the right of $x=a$. We show for comparison
675: $V=\frac{\pi}{2}\left(=j l \right)$ the average voltage. Notice in both
676: cases
677: the discontinuity of $\phi_t$ at $x=a$.
678:
679:
680: Let us show that this is consistent with
681: the whole equation (\ref{sgd}), for the intensity $I=V/ \alpha d_a$.
682: Consider the left only solution
683: \begin{equation}
684: \label{gauchecos}
685: \phi\left(x,t\right)= \left \{ \begin{array}{l r}
686: Vt+ \frac{a_l}{V}\sin(V t)\cos(Vx)-\frac{j}{2}[x^2-a^2]& x < a \\
687: Vt-\frac{j}{2}\left[(x-l)^2+(a-l)^2\right]& x > a
688: \end{array} \right .
689: \end{equation}
690: with, $V=\frac{2k_l+1}{2}\frac{\pi}{a}$.
691: On the interval $[0,a[$ (respectively $]a,l]$) $\phi$ satisfies
692: the linear wave equation
693: $\phi_{tt}-\phi_{xx} = j$ with the boundary condition
694: $\phi_x|_{0}=0$ (respectively $\phi_x|_{l}=0$).
695: We now integrate (\ref{sgd}) on $[0,l]$ to obtain
696: \be
697: \int_{0}^a \phi_{tt}dx+\int_a^{l}\phi_{tt}dx +
698: \int_{0}^{l}\phi_{xx}dx+d_a\sin\left(Vt\right)+\alpha d_a V = jl, \ee
699: so that
700: $$-a_l \sin(V t)\left[\sin(Vx)\right]_0^a + 0 + 0 +
701: d_a\sin\left(V t\right) = 0.$$
702: The end result is
703: $$(-1)^{k_l+1} a_l \sin(V t)=d_a\sin(V t)$$
704: which implies $a_l=(-1)^{k_l+1} d_a$ and the equation is
705: balanced.
706:
707: \section{The junction mode}
708:
709: The expression (\ref{sjm}) a priori defined only for a given time
710: interval
711: can be $C^{\infty}$ extended for all times using the continuation
712: $$f \left(t+k\frac{2\alpha \pi}{\sqrt{(j/d_a)^2-1}} \right)=
713: f \left(t\right)+k 2\pi ~,$$
714: where $k$ is an integer.
715: The pseudo-period $T$ of $\phi_a$ is $T=\frac{2\alpha
716: \pi}{\sqrt{(j/d_a)^2-1}}$
717: so the voltage is $V=\frac{2\pi}{T}={1 \over
718: \alpha}\sqrt{(j/d_a)^2-1}$.
719: This can also be written
720: \be\label{ji}
721: j = {d_a \over l} \sqrt{\left(\alpha V\right)^2+1}\ee
722: So we deduce from $A^{''}_0(t)=0$ and the voltage expression
723: (\ref{vjon})
724: that $A_0(t)=Vt+C$.
725:
726: Let us now calculate the contribution of the
727: Fourier modes and
728: give a solution to the whole equation (\ref{an}). From (\ref{an})
729: and the previous equality, we obtain:
730: \be \label{a_pjunc}
731: A_p^{''}\left(t\right) + \left(\frac{p \pi}{l} \right)^2
732: A_p\left(t\right) = - 2 j c_p
733: \ee
734: The solution of this equation, $A_p$ is:
735: \be \label{solA_pjonc}
736: A_p \left(t\right)= \gamma_p \cos \left(\frac{p \pi}{l}t \right)
737: + \beta_p\sin \left(\frac{p \pi}{l}t \right)
738: -\frac{2 j l^2}{(p \pi)^2} c_p
739: \ee
740: Now we have the expression of all the Fourier modes.
741: We know that $\phi(a,t) = \sum_{n=0}^{+\infty}A_n(t) c_n^a$.
742: Substituting the Fourier modes (\ref{solA_pjonc}) into
743: this expression we obtain:
744: \be \label{pjmi}
745: \phi(a,t)=Vt+C+\sum_{p=1}^{+\infty}\left[\gamma_p \cos \left(\frac{p
746: \pi}{l}t \right)
747: + \beta_p \sin \left(\frac{p \pi}{l}t \right)-\frac{2 j l^2}
748: {(p \pi)^2} c_p\right]c_p ,
749: \ee
750: The series in the last term is uniformly convergent so that the term
751: can be
752: collected in the form of a constant yielding the final simpler
753: expression
754: \begin{equation}\label{pjm} \phi(a,t)=Vt+C_2+\sum_{p=1}^{+\infty}
755: \left[\gamma_p c_p\cos \left(\frac{p \pi}{l}t \right)
756: + \beta_p c_p \sin \left(\frac{p \pi}{l}t \right)\right]
757: \end{equation}
758: Introduce $\phi(a,t)-Vt+C_2 = u(t)$. We can choose $C_2$ as $\int_0^T
759: u(t) dt =0$.
760: Notice that $u$ is a $C^{\infty}$ and $T$-periodic function.
761: We can make Fourier projection on it.
762: $$u(t) = \sum_{p=1}^{+\infty}\left[\gamma_p c_p\cos \left(\frac{p
763: \pi}{l}t \right)
764: + \beta_p c_p \sin \left(\frac{p \pi}{l}t \right)\right] $$
765: Until $c_p \neq 0$, we have
766: \begin{eqnarray}
767: \gamma_p = \frac{2}{c_p T}\int_0^T u(t) \cos\left(\frac{p
768: \pi}{l}t\right) dt \\
769: \beta_p = \frac{2}{c_p T}\int_0^T u(t) \sin\left(\frac{p
770: \pi}{l}t\right) dt
771: \end{eqnarray}
772: We can make the following remarks
773: \begin{enumerate}
774: \item $V=\frac{2\pi}{T}=\frac{n\pi}{l}$ so $\frac{1}{T}=\frac{n}{2l}$
775: and $u$ is $\frac{2l}{n}$ periodic.
776: \item If $V=\frac{n\pi}{l}$ the terms of the Fourier series that aren't
777: zero are the multiples of $n$.
778: \item If $c_n=0$ then the junction modes for $V=\frac{n\pi}{l}$ can't
779: exist.
780: For example when the junction is in the middle of the circuit, then
781: $c_n^a=0$
782: for $n=2k+1$ and junction modes exist only for $V=\frac{2k\pi}{l}$
783: where $k$ is
784: an integer (see left panel of figure \ref{f5}).
785: \item When $|c_n|$ is very small, the circuit in the junction mode
786: stores
787: a lot of energy but is unstable. We will see this in the next section.
788: \end{enumerate}
789: Now remember (\ref{four}), we can give the explicit solution of the
790: phase
791: in the $n$ th junction mode, up to a constant
792: \begin{equation} \label{modjonc}
793: \phi(x,t) = {n \pi \over l} t + \sum_{p=1}^{+\infty}
794: \left [\gamma_p \cos \left(\frac{p n \pi}{l}t\right) +
795: \beta_p\sin \left(\frac{p n \pi}{l}t \right) \right ]
796: \cos\left(\frac{p n\pi}{l} x \right) ~.
797: \end{equation}
798: Notice that the instantaneous voltage $\phi_t$ has a continuous
799: derivative at the junction position $x=a$.
800: Fig. \ref{f4} shows the instantaneous voltage $\phi_t(x,t)$ as a
801: function of $x$ for 60 consecutive times for the same parameters as
802: in Fig. \ref{f2}. Notice how the solution oscillates over the whole
803: domain and that one cannot see the junction position. This type of
804: motion is similar to the transparency observed in the reflection
805: coefficient of a stripe \cite{lamb} when an integer number of
806: half-periods
807: of the wave "fits" in the stripe.
808:
809: \section{Single junction case: IV characteristics}
810:
811: For a large enough time the system finds a stationary state where the
812: energy provided by the input current is balanced by dissipation. This
813: state
814: is described by the IV characteristic curve. This is measured
815: on real devices and it is therefore very important to understand its
816: features.
817:
818: To compute it we fix the current and run the system for about
819: $t_1=3000$
820: time units so that
821: all transients have died off. Then we compute the average voltage
822: \be\label{avolt}
823: \left<\phi_t(a,t)\right>_t\equiv
824: {\phi(a,t_1+T)-\phi(a,t_1)\over T}\ee
825: at the junction. The value $T$ is about 2000. The space average of the
826: average voltage in time
827: $\left<\frac{1}{l}\int_{0}^{l} \phi_t (x,t) dx\right>_t$
828: is compared to this quantity to ensure that the
829: system is completely thermalized. Only then is the voltage recorded.
830:
831: The IV curves are computed for 99 steps of increasing current $j$ from
832: 0 to 4
833: starting with an initial condition $\phi(x,0) \equiv 0$ and
834: $\phi_t(x,0) \equiv 0$. We then decrease the current from 4 to 0.
835: All plots show these two curves for increasing and decreasing current.
836:
837: \subsection{Bounds for IV curves}
838:
839: We have isolated two characteristic behaviors of the circuit. These
840: allow
841: us to bound the IV curves for a single junction in a
842: microstrip. Fig. \ref{f5} shows two IV curves for
843: a centered junction $a=5$ (left panel) and an off-centered
844: junction $a=9$ in a microstrip of total length $l=10$.
845: For each case we plot the limiting behaviors given by the
846: ohmic mode and junction mode voltages (\ref{vohm}) and (\ref{vjon}).
847: $V, j$ are bounded by the relations
848: $$ d_a \alpha V \leq jl \leq d_a \sqrt{(\alpha V)^2 + 1}~,~{\rm or}~~
849: {1 \over \alpha}\sqrt{\left({jl \over d_a}\right)^2-1}\leq V \leq
850: {jl \over d_a \alpha}$$
851: The ohmic modes, (resp. the junction modes) only occur
852: for voltages such that $V=V_l=\frac{2 k_l +1}{2}\frac{\pi}{a}$ or
853: $V=V_r = \frac{2 k_r + 1}{2}\frac{\pi}{a-l}$ (resp. $V={k\pi \over l}$).
854: The voltage (horizontal) axis is labeled with the indices corresponding
855: in each case to the junction modes. The ohmic modes are marked with
856: ticks but not labelled.
857: We always go from an ohmic mode
858: to a junction mode and vice versa. As the current is increased the
859: junction mode ceases to exist and the
860: system then jumps to the closest ohmic mode. Notice
861: the hysteresis obtained between the curve for
862: increasing $j$ and the one for decreasing $j$.
863: Both the junction voltage and ohmic mode voltages are fixed by the length of
864: the system. On the other hand the current for which one
865: obtains these voltages depends on the damping $\alpha$
866: and strength of the junction $d_a$. For fixed $d_a$, an decrease of
867: $\alpha$ will tilt the IV curves towards the horizontal voltage axis and
868: this will separate the cavity modes and cause large voltage jumps from
869: a cavity mode to a far away ohmic mode and
870: hysteresis. Conversely an increase of $\alpha$ will make the
871: IV curves more vertical so that the voltage jumps and hysteresis
872: will be reduced.
873: Note also that as in the static case, a magnetic field has no
874: effect on the IV curves for a single junction.
875:
876: When the junction is centered as in the left
877: panel of Fig. \ref{f5} the ohmic modes (resp. junction modes)
878: correspond to the
879: odd $V=(2 k + 1) \pi/l$ (resp. even $V=2 k \pi/l$) cavity modes so the
880: voltage
881: interval between a junction mode and its corresponding ohmic mode
882: neighbors is constant. This does not happen when the junction is
883: off-centered as in the right panel of Fig. \ref{f5} so that one
884: can get very sharp resonances connecting an ohmic mode to its
885: junction mode neighbor like the ones for $n=2,3$ and 4. The resonance
886: for $n=3$ was obtained by decreasing current.
887: For larger values of $n$ the voltage separation becomes larger so that
888: the resonances are softer.
889:
890:
891: \subsection{Study of resonances}
892:
893: We have been able to bound the IV curves for our system. Now let us
894: discuss the fine structure of the resonances, why some are sharp
895: (see resonances 3,4 and 6 in Fig. \ref{f5}) while others are not.
896: To explain this one should consider the energy of the system and
897: the Fourier modes. Multiplying (\ref{sgd}) by $\phi_t$ and integrating
898: from $0$ to $l$ we obtain the work equation
899: \be \label{work}
900: \partial_t \left[{1 \over 2}\int_{0}^{l} \phi_{t}^2 + \phi_{x}^2 dx
901: + d_a \left(1-\cos\phi(a,t)\right) \right] =
902: j \int_{0}^{l} \phi_t dx -d_a \alpha\phi_t^2 (a,t).
903: \ee
904: We can write (\ref{work}) as
905: $\partial_t E_t\equiv \partial_t (E_p + E_j) = P_{dc} + P_q $ where
906: $E_p = {1 \over 2}\int_{0}^{l} ( \phi_{t}^2 + \phi_{x}^2) dx$
907: is the energy in the passive region, $E_j= d_a
908: \left(1-\cos\phi(a,t)\right) $ the Josephson energy,
909: $P_{dc} = j \int_{0}^{l} \phi_t dx$ is the power given by the direct
910: current
911: and $P_q= d_a \alpha\phi_t^2 (a,t)$ is power dissipated by the
912: quasiparticles.
913: In figure \ref{f6}, we plot from top to bottom $E_t,E_p,P_q$ and $E_j$
914: for
915: a centered junction $a=5$ (bottom panels) and an off-centered junction
916: $a=9$ (top panels). The left panels show the junction mode $n=4$
917: and the right panels the corresponding ohmic mode for the same current.
918: In the junction mode the quasi-particle term $P_q$ oscillates strongly
919: and the Josephson energy $E_j$ is very anharmonic. In the ohmic mode
920: $P_q$ is almost constant and $E_j$ is sinusoidal. In all cases the
921: energy in the cavity $E_p$ is much larger than the junction energy
922: $E_j$.
923: When the junction is off-centered (top left panel) the energy stored in
924: the cavity is much larger. This is due to the smallness of the coupling
925: coefficient $c_n$ as we now show.
926:
927:
928: For that consider the phase $\phi(a,t)$ at the junction as given by
929: (\ref{sjm}). It is independent of the junction position $a$.
930: Since we are in the junction mode $V=n\pi/l$ we write from
931: (\ref{modjonc})
932: $$\phi(a,t) = {n \pi \over l} t + \sum_{p=1}^{+\infty}
933: \left [\gamma_p \cos \left(\frac{p n \pi}{l}t\right) +
934: \beta_p\sin \left(\frac{p n \pi}{l}t \right) \right ]
935: c_{np}. $$
936: The first statement implies that the terms $\gamma_p c_{np}$
937: and $\beta_p c_{np}$ are
938: constants independent of the junction position $a$ so that
939: $\gamma_p$ and $\beta_p$ are proportional to $1/c_{np}$.
940: As a consequence if $c_{np}$ is small the junction mode
941: of index $n$ corresponds to a large amplitude in
942: the mode $A_{np}$. The system
943: accumulates this energy in the passive region and the resonance is very
944: sharp.
945:
946:
947: Fig. \ref{f6} shows a plot as a function of time, in each panel,
948: from top to bottom of the total energy $E_t$, the energy in the
949: passive region $E_p$, the power dissipated by the quasi-particles $P_q$
950: and
951: the Josephson energy $E_j$ for
952: a centered junction $a=5$ (bottom panels) and an off-centered junction
953: $a=9$ (top panels).
954: The left panels show the junction mode $n=4$
955: and the right panels the corresponding ohmic mode for the same current.
956: The coupling coefficient is small $c_4=0.3$ in the top panel
957: while it is maximum $c_4=1$ in the bottom panel where the junction
958: is centered. Therefore one expects a sharp resonance and a highly
959: excited passive cavity for $a=9$ as opposed to $a=5$.
960: This is indeed the case as one can see from the plot of $E_p$ in the
961: left bottom and left top panels. The Josephson energy $E_j$ and the
962: power dissipated by the junction are of the same order.
963: In the ohmic regime this accumulation of energy in the cavity is
964: absent. There is a small difference in $E_p$ between the two
965: configurations because they have different lengths of passive regions.
966: Note that the total energy is practically independent of time and so
967: is the dissipation. There is an exact balance between the Josephson
968: energy and the energy in the cavity $E_p$.
969:
970:
971:
972: \section{The many junction case}
973:
974: Assuming a general situation with magnetic field, a mixed current feed
975: and many junctions in the micro-strip we can generalize our model to
976: \be\label{manyj}
977: \phi_{tt}- \phi_{xx} + \sum_j d_j \delta(x-a_j) (\sin(\phi) +
978: \alpha\phi_t)=\nu j \ee
979: with the boundary conditions $\phi_x|_{x = 0}= b-{jl \over 2}(1-\nu)$
980: and $\phi_x|_{x = l}= b+{jl \over 2}(1-\nu)$.
981: If $\nu=1$,
982: In the model (\ref{manyj}) $\nu=1$ (resp. $\nu=0$) corresponds to
983: an overlap (resp. inline) current feed.
984:
985: \subsection{Static case}
986:
987: We have seen that for a single junction the maximum current $j l = d_a$
988: is reached for any value of the magnetic field. This is not true for
989: two junctions (a SQUID) where the maximum critical current if obtained
990: for
991: $b=0 ~~{\rm mod.} {2 \pi n \over a_2-a_1}~~$ where $n$ is an integer
992: \cite{cg04}. Static solutions for many junctions in a 1D microstrip
993: will be discussed in a forthcoming paper \cite{many}.
994:
995: \subsection{High voltage case}
996:
997: As in the case of a single junction, when the stationary
998: state is reached we can assume $\phi_{t}=V$ and average out the
999: $\phi_{tt}$ and $\sin(\phi)$ terms yielding the following
1000: boundary value problem for the time independent part $\phi_v$ of $\phi$
1001: \begin{eqnarray} \label{hvolov}
1002: - {\phi_v}_{xx} + \sum_{k=1}^n d_k \delta (x-{a_k})\alpha V = \nu j ,\\
1003: {\phi_v}_{x} |_{0,l}=b \mp {jl \over 2}(1-\nu),
1004: \end{eqnarray}
1005: On each interval separating the junctions ${\phi_v}_{xx} = - \nu j$ so
1006: that ${\phi_v}$ is a second degree polynomial
1007: ${\phi_v}(x) =-(\nu j/2)x^2 + b x + d$.
1008:
1009:
1010: Integrating (\ref{hvolov}) from $0$ to $l$ yields the voltage
1011: $$V = \frac{j l}{\alpha\sum_{k=1}^n d_k}.$$
1012: Integrating (\ref{hvolov}) over small domains containing the
1013: junctions we obtain as usual the jump conditions in the derivatives
1014: $$\left[ {\phi_v}_x \right]_{a_i^-}^{a_i^+} =
1015: \frac{d_i j l}{\sum_{k=1}^n d_k}.$$
1016:
1017: Using these remarks we can calculate $\phi_v$ for any number of
1018: junctions
1019: (for simplicity one can assume $\phi_v(a_1)=0$). Let's introduce
1020: $S(x)$, the
1021: spatial curvature of $\left<\phi\right>_t$. For two junctions we get
1022: \begin{equation}
1023: \phi_v(x) =
1024: \left\{\begin{array}{l r}
1025: \frac{-\nu j}{2}(x^2 -a_1^2) + \left(b-{jl \over 2}(1 -
1026: \nu)\right)(x-a_1)
1027: & x \leq a_1 \\
1028: \frac{-\nu j}{2}(x^2 -a_1^2) + \left({d_1 jl \over d_1+d_2} +
1029: b-{jl \over 2}(1 - \nu)\right)(x-a_1) & a_1<x<a_2 \\
1030: \frac{-\nu j}{2}(x^2 - a_2^2) + \left(b-{jl \over 2}(1+\nu)\right)
1031: (x-a_1) + S_2 & a_2 \leq x \\
1032: \end{array} \right .
1033: \end{equation}
1034: where $S_2 = \frac{-\nu j}{2}(a_2^2 -a_1^2) +
1035: \left({d_1 jl \over d_1+d_2} + b-{jl \over 2}(1 - \nu)\right)(a_2-a_1)$
1036: is the phase difference between the second and first junction
1037: ($S_2 \equiv \phi(a_2,t) -\phi(a_1,t)$). The phase is then given by
1038: \begin{equation}\label{CurveHV}
1039: \phi(x,t) = \phi_v(x) + Vt + C,
1040: \end{equation}{manyj}
1041: where $C$ is an integration constant. Notice that it is possible to
1042: find $\phi_v(x)$ for an arbitrary number of junctions and compute the
1043: phase difference $\phi_v(a_i)=S_i \equiv \phi(a_i,t)-\phi(a_{i-1},t)$.
1044:
1045: In Fig. \ref{f7} we show $\phi(x,t)$ for 6 consecutive values of
1046: time separated by $\delta t=0.1$ for a system of two junctions
1047: located respectively at $a_1=2$ and $a_2=6.2$ in a domain of
1048: length $l=10$ with current $j=1$ fed in overlap geometry ($\nu=1$).
1049: One can see that the behavior
1050: is close to what is predicted by the above expression.
1051:
1052: \subsection{Fourier representation}
1053:
1054: We can use the high voltage solution to simplify (\ref{manyj}).
1055: For that take $\psi = \phi + \phi_v$ so that (\ref{manyj}) becomes
1056: \begin{eqnarray}\label{simplj}
1057: \psi_{tt}-\psi_{xx}+\sum_{k=1}^n d_k \delta (x-{a_k})\left (
1058: \sin(\psi-S_k) + \alpha \psi_t-{jl \over \sum_{k=1}^n d_k} \right)=0,\\
1059: \left. \psi_x \right|_{x=0;l} = 0.
1060: \end{eqnarray}
1061: The transformation using the high voltage solution concentrates the
1062: current on the junctions. In this representation inline or overlap
1063: configurations differ by the terms $S_k$.
1064:
1065:
1066: Let us go back with Fourier series. Because of the homogeneous Neumann
1067: boundary conditions on (\ref{simplj}), we can decompose $\psi$ on
1068: a cosine Fourier series
1069: $$\psi(x,t) = \sum_{n=0}^{+\infty} A_n(t) \cos\left({n\pi x \over
1070: l}\right)$$
1071: and obtain the following equations for the modes $An$
1072: \begin{eqnarray}\label{manya0}
1073: l A_0^{''} + \sum_{j=1}^n d_j \left(\sin(\psi_j - S_j) + \alpha
1074: \psi_{jt}
1075: -{jl \over \sum_{k=1}^n d_k}\right)=0, \\
1076: A_n^{''} + \left({n \pi \over l}\right)^2 A_n + {2 \over l}
1077: \sum_{j=1}^n d_j c^j_n \left(\sin(\psi_j - S_j) + \alpha \psi_{jt}
1078: -{jl \over \sum_{k=1}^n d_k}\right)=0,
1079: \end{eqnarray}
1080: where $c^j_n = \cos\left({n\pi a_j\over l}\right),~~\psi_j \equiv
1081: \psi(a_j,t),~~
1082: \psi_{jt} \equiv \partial_t \psi(a_j,t)$.
1083:
1084: \section{IV curves for two or more junctions}
1085:
1086: \subsection{2 symmetrically placed junctions, limiting behaviors.}
1087:
1088: In a symmetric circuit, we have $a_2=l-a_1$ and $d_1=d_2$. For
1089: simplicity we assume no magnetic field so that
1090: $b=0$ too. With this symmetry, it is easy to show that
1091: $c_{2k}^2=c_{2k}^1$
1092: and $c_{2k+1}^2=-c_{2k+1}^1$.
1093: It is important to remark that when $b=0$, $S_2 = 0$. So there is no
1094: phase shift between the two junctions. Let us now go back
1095: to the Fourier modes, (\ref{manya0}) becomes:
1096: \begin{equation}\label{A_0_2jsym}
1097: d_1 \left( \sin(\psi_1) + \alpha \psi_{1t}+\sin(\psi_2)+ \alpha
1098: \psi_{2t}\right) -jl = - l A_0^{''}.
1099: \end{equation}
1100: For $p=2k$,
1101: \begin{eqnarray*}\label{A_2k_2j}
1102: A_{2k}^{''} + \left(\frac{2k \pi}{l}\right)^2A_{2k} &=&
1103: -\frac{2 c_{2k}^1}{l}[d_1 (\sin(\psi_1)+ \alpha \psi_{1t}+
1104: \sin(\psi_2)+\alpha \psi_{2t})-jl] \\
1105: &=& 2 c_{2k}^1 A_0^{''}
1106: \end{eqnarray*}
1107: We want to show that the following equation for $A_{2k+1}$ is decoupled
1108: from the system of Fourier modes
1109: \begin{equation}\label{A_2k+1_2j}
1110: A_{2k+1}^{''}+\left(\frac{(2k+1) \pi}{l}\right)^2A_{2k+1}=
1111: \frac{-2 c_{2k+1}^1}{l}d_1 (\sin(\psi_1) - \sin(\psi_2)
1112: + \alpha (\psi_{1t} - \psi_{2t})).\end{equation}
1113: For that we group together the even terms and odd terms of the
1114: Fourier series of $\psi_{1t}$ and $\psi_{2t}$ and define
1115: $s_o(t)\equiv\sum^{+\infty}_{k=0}
1116: c_{2k}^1 A_{2k}(t)$ and
1117: $s_e(t) \equiv \sum^{+\infty}_{k=0} c_{2k+1}^1 A_{2k+1}(t)$.
1118: Notice that:
1119: $$\psi(a_1,t)=s_o(t)+s_e(t)~~{\rm and }~~\psi(a_2,t)=s_o(t)-s_e(t)$$
1120: Replacing $\psi_1$ and $\psi_2$ in equation (\ref{A_2k+1_2j}),
1121: we obtain:
1122: \begin{equation}\label{decoupl_2j}
1123: A_{2k+1}^{''}+\left(\frac{(2k+1) \pi}{l}\right)^2A_{2k+1}=
1124: \frac{-4 c_{2k+1}^1}{l}d_1 \left(\cos(s_e)\sin(s_o) -
1125: \alpha s_e^{'}\right)
1126: \end{equation}
1127: This equations show that the system performs like a single centered
1128: junction with different coupling coefficients. So we can balance
1129: the Fourier modes in the junction mode: with initial conditions
1130: $\forall k$, $A_{2k+1}(0)=0$, $A^{'}_{2k+1}(0)=0$ we have
1131: $S_2(0)=0$ and as a consequence $A_{2k+1}(t) \equiv 0$,
1132: $\forall t \in \Re^+$. This shows that the equations for the
1133: odd terms are decoupled from the system.
1134:
1135: {\bf Junctions modes.}\\
1136: As we see, the system (\ref{manya0}) behaves like a single
1137: junction: the odd terms are equal to zero and the even terms
1138: give the junction modes
1139: (see (\ref{sjm})):
1140: $$\phi_1=\phi_2=2\arctan\left\{{d_1 \over jl } \left [\tan
1141: \left({1 \over 2 \alpha} V_j \left(t+C_2\right)\right) V_j + 1\right]
1142: \right\},$$
1143: with $V_j = \sqrt{\left({jl \over d_1}\right)^2-1} = \frac{n \pi}{l}$.
1144:
1145: Notice that the two junctions are perfectly synchronized. In a
1146: non-symmetric situation, junctions modes for both junctions
1147: are particular cases.
1148:
1149: {\bf Ohmic modes.}\\
1150: Choosing $S_2 = 0$ allow us to show again analytical solutions.
1151: Assuming $b=0$, there are three types of solution, given by different
1152: oscillations of the phase:
1153: \begin{enumerate}
1154: \item The phase oscillates between the junctions so that $V=
1155: \frac{(2k+1)\pi}{a_2-a_1}$. Due to the symmetry of the problem, we
1156: cannot find solutions with $V=\frac{2k\pi}{a_2-a_1}$.
1157: \item The phase oscillates outside the junctions:
1158: $V=\frac{(2 p + 1)\pi}{2 a_1}$.
1159: \item The phase oscillates in all the passive parts so that
1160: $V=\frac{(2k+1)\pi}{a_2-a_1}=\frac{(2p+1)\pi}{2 a_1}$. This is
1161: only possible if $a_2-a_1$ is a fraction of $l-a_2$.
1162: \end{enumerate}
1163: To conclude in the symmetric case, we know limit IV curves with exact
1164: solutions. See the left panel of Fig.\ref{f11}. In the general case, we cannot
1165: find exact solutions, we have near ohmic mode solutions and peaks at the
1166: expected resonant voltages $V=\frac{n \pi}{l}$.
1167:
1168: \subsection{The non-symmetric cases: IV curves }
1169:
1170: The IV curve obtained for a circuit of two junctions placed in
1171: a non symmetric fashion confirms the existence of ohmic modes
1172: at the expected voltages. It is still possible to characterize the
1173: IV curves using ohmic modes and junction modes even if there are
1174: only approximate. Without symmetry and when $S_2=0~[\pi]$ we can obtain
1175: exactly any one of ohmic modes discussed in previous section. In
1176: addition to the symmetric case there two new possibilities where $V$ can
1177: be computed:
1178: \begin{enumerate}
1179: \item The phase oscillates between the junctions and to the
1180: right of $a_1$. See the left panel of Fig.\ref{f8} as an example.
1181: \item The phase oscillates to the left of $a_2$ and between the
1182: junctions.
1183: \end{enumerate}
1184: In Fig.\ref{f8}, the middle panel shows the second point of the
1185: symmetric case and the right panel illustrates the case where
1186: the phase oscillates between the junctions for the non-symmetric case.
1187:
1188: Peaks appear at the junction modes $V=\frac{k\pi}{l}$, but they are
1189: in general smaller than in the symmetric cases.
1190: Fig.\ref{f9} shows the IV curves for two non symmetrically placed
1191: junctions and indeed one can see that the height of the resonances
1192: is lower than one for two decoupled junctions (the upper curve in
1193: the three panels). By varying the junction critical current
1194: (the coefficients $d_1$ and $d_2$ ) we show that the resonances
1195: are linked to one of the two junctions. To understand
1196: this phenomenon, go back to the coupling coefficients $c_n^j$.
1197: For the second resonance in Fig. \ref{f9}:
1198: $|c^1_2|=1$, $|c^2_2|=0.73$ so that the junctions are strongly coupled to
1199: the system. This is like in the symmetric case but it is an exception. The IV
1200: curve shows that they are together in junction mode (or very close).
1201: Consider now the third resonance. The coupling coefficients are $c^1_3
1202: = 0$ and $c^2_3 = 0.90$ so that the junction $a_1$ is decoupled from
1203: the system, we can assume ohmic behavior for it and $a_2$ is in
1204: junction mode. The fourth resonance corresponds to the reverse
1205: situation, $c^1_4=1$, $c^2_4=0.06$ where $a_2$ is decoupled.
1206:
1207: To confirm this, we plot in Fig. \ref{f10}, $\phi_1$ and $\phi_2$
1208: for the resonances $3$ and $4$ with $d_1=0.7$, $d_2=0.3$
1209: (left panel of Fig. \ref{f9}) and $d_1=d_2=0.5$ (middle panel
1210: of Fig. \ref{f9}).
1211: Together with the numerical results we plot the junction mode solution
1212: (\ref{sjm}) in dashed line. For the top panel (third resonance) one
1213: can see that the junction $a_2$ follows closely this behavior as opposed
1214: to $a_1$ which is in ohmic mode. The situation is reversed for the
1215: fourth resonance as shown in the bottom panels.
1216: For resonances 2, 3 and 4, the value of the critical current $d_i~,i=1,2$ does
1217: not modify appreciably the behavior of the phase at the junctions
1218: (remark that $S_2 \neq 0$ but stay small).
1219:
1220: For a general resonance the junctions will work in different modes,
1221: one will be closer to a junction mode while the other will be closer to
1222: an ohmic mode. We call this mode a resistor-junction mode.
1223:
1224: {\bf Resistor-junction modes.}\\
1225: Resistor-junction modes exist if the circuit has one junction heavily
1226: coupled while the other is decoupled. In this case, the system behaves almost
1227: like a single junction circuit.
1228: This is why these modes are associed to voltages
1229: $V = \frac{k\pi}{l}$.
1230:
1231: If the junction $a_i$ is in ohmic mode, then $[\phi_x]_{a_i^-}^{a_i^+}=
1232: d_k \sin(Vt+C_1) + C_k$. So that we can only have only one harmonic
1233: in the circuit. Whereas a junction
1234: mode exists only with an infinity of harmonics. So the system reaches a
1235: constructive equilibrium.
1236: For example in all the panels of Fig. \ref{f10}, we plot the exact junction
1237: mode to compare. The observation of the phases in the junction mode
1238: indicates that it lingers more around $(2k+1)\pi/2$ than the
1239: expression (\ref{sjm}). This gives for the IV curves
1240: of Fig. \ref{f9} resonances that are slightly higher than expected.
1241: This summarizes the influence of one junction on the other.
1242:
1243: \subsection{$\pi$-junctions and magnetic field effects}
1244:
1245: We now consider the situation where the critical current $d_i$ for
1246: some $i$ can be negative. Then the ith junction is a so-called $\pi$ junction
1247: for which the Josephson relation is given by $\sin(\phi +\pi)$
1248: instead of $\sin(\phi)$. In this case we will show that the dynamics
1249: is very different than the one where all the $d_i$ are positive.
1250: Therefore combining $\pi$ and standard junctions provides a way
1251: to change the resonant frequencies of the array.
1252: Another important parameter which we have not varied up to now is the
1253: magnetic field $b$. We will see that a change of $b$ changes the IV
1254: characteristic in a way that is similar to changing the critical
1255: current density of some of the junctions in the array. This is
1256: another way to tune the array to resonate on specific
1257: frequencies.
1258:
1259:
1260: We now illustrate these specific points on some examples.
1261: Consider an array of $N$ point junctions described by
1262: $$\phi_{tt}- \phi_{xx} + \sum_{j=1}^{N} d_j \delta(x-a_j) (\sin(\phi) +
1263: \alpha\phi_t + c \phi_{tt})=\nu j~,$$
1264: with the boundary conditions $\phi_x|_{x = 0}= b+{jl \over 2}(1-\nu)$
1265: and $\phi_x|_{x = l}= b-{jl \over 2}(1-\nu)$.
1266: First note that if $d_i<0$ for all i's then we can take
1267: $\psi= \phi+\pi$ to get us back to the standard case $d_i>0$. We therefore
1268: obtain the same IV curve as when all the $d_i>0$.
1269:
1270: An interesting case corresponds to $N=2$. Let us consider the effect
1271: of an additional magnetic field $b'$ on the IV curve. For that
1272: introduce $\psi(x,t)=\phi(x,t)+f(x)$ where $f(x)= b'(x-a_1)$.
1273: The equation verified by $\psi$ is
1274: $$\psi_{tt}- \psi_{xx}
1275: + d_1 \delta(x-a_1) (\sin(\psi + f_1) + \alpha\psi_t )
1276: + d_2 \delta(x-a_2) (\sin(\psi + f_2) + \alpha\psi_t )
1277: = \nu j, $$
1278: with the boundary conditions
1279: $\psi_x|_{x = 0}= b+ b'+{jl \over 2}(1-\nu)$
1280: and $\psi_x|_{x = l}= b+b'-{jl \over 2}(1-\nu)$.
1281: When $f_1=0$ and $f_2=2 \pi$ ie when $b'= 2 \pi / (a_2-a_1)$
1282: we obtain the same equation for
1283: $\phi$ and $\psi$ apart from the boundary conditions.
1284: The IV curves
1285: for the two magnetic fields $b=0$ and $b= 2 n \pi / (a_2-a_1)$
1286: where $n$ is an integer are then identical. Now if one chooses
1287: a magnetic field $b= \pi / (a_2-a_1)$ then the phases are shifted
1288: by $\pi$ one from the other and this is like changing the sign of
1289: one of the $d_i$.
1290: In practice
1291: these results only makes sense for magnetic fields for which the
1292: size of the junctions can be neglected. This is the limit of our
1293: approach.
1294:
1295:
1296: Fig. \ref{f11} shows two IV curves for two symmetrically placed junctions
1297: $a_1=2$ et $a_2=8$ such that
1298: $b=0$ $d_1=d_2=0.5$ corresponding to standard junctions
1299: (left panel) and $b=0$ $d_1=-d_2=0.5$ corresponding to a standard junction
1300: and a $\pi$ junction (right panel).
1301: We observe in the left panel resonances (junction modes) every
1302: $\frac{2k\pi}{l}$ exactly as for a centered junction. In the right
1303: panel, resonances (junction modes) are found every
1304: $\frac{(2k+1)\pi}{l}$. The $\pi$ phase shift between
1305: the two junctions implies that odd harmonics are highly coupled and
1306: even harmonics are decoupled from the system.
1307: We could have obtained the same
1308: change in the nature of the resonances by using the standard array (left panel)
1309: and changing the magnetic field to $b= \pi/(a_2-a_1)$. As
1310: the field is changed one then goes continuously from an IV curve
1311: with only even resonances to an IV curve with only odd resonances.
1312:
1313:
1314: Now let us consider how these results obtained for $N=2$ affect the
1315: behavior of an array of $N>3$ junctions. If there is a $d$ such that
1316: $a_{i+1}-a_i= n_i d$ where $n_i$ is an integer for all $i$ then
1317: the IV curves for $b=0$ will be the same as for $b= 2\pi/d$.
1318: If one chooses $b= \pi/d$ the junctions such that $f_i=\pi ~{\rm mod.}~ 2 \pi$
1319: will be reversed while the ones for which $f_i=2\pi ~{\rm mod.}~2 \pi$
1320: will not be affected. This is similar to a flute which is tuned
1321: by changing nodes.
1322:
1323:
1324: \section{Conclusion}
1325:
1326: We have introduced a simple and general model using a wave equation
1327: with delta distributed sine nonlinearities to describe the
1328: dynamics of point Josephson junctions in a 1D microstrip. This
1329: can be extended to $\pi$ junctions for which the current density can
1330: be either positive or negative in the domain. We can also apply it
1331: to situations where the current density in each junction is different.
1332:
1333:
1334: For a single junction we have shown two limiting behaviors, the ohmic
1335: mode where the junction behaves as a resistor driven by the cavity, separating
1336: waves and the junction mode where the Josephson element is driving the cavity.
1337: These limiting behaviors have allowed us to bound the IV curves and
1338: understand the observed resonances. When another junction is present
1339: in the cavity, this simple classification is changed in
1340: general and we observe ohmic modes and combined junction/ohmic modes. The
1341: existence or not of the nth resonance is connected to the value of
1342: the Fourier coefficient $c_n^a = \cos{n \pi a /l}$.
1343:
1344: These results carry over to the case of an array with many junctions where
1345: it is possible to choose a voltage where one of the junctions is inactive and
1346: another is in junction mode. We also use the analysis to understand the
1347: effect of an external magnetic field and the influence of having 0 or $\pi$
1348: junctions. We believe that such a device composed of normal and $\pi$
1349: junctions placed at specific locations in a microstrip
1350: can be used to great advantage for specific applications like
1351: resonators and frequency mixers.
1352:
1353:
1354: \begin{thebibliography}{99}
1355:
1356: \bibitem{josephson} B. D. Josephson, Phys. Lett. {\bf 1}, 251, (1962).
1357:
1358: \bibitem{Likharev} K. Likharev, {\em Dynamics of Josephson junctions
1359: and
1360: circuits}, Gordon and Breach, (1986).
1361:
1362: \bibitem{Barone} A. Barone and G. Paterno, {\em Physics and
1363: Applications
1364: of the Josephson effect}, J. Wiley, (1982).
1365:
1366: \bibitem{Salez} M. H. Chung and M. Salez, in Proc. 4th European
1367: Conference on
1368: applied superconductivity, EUCAS 99, 651, (1999).
1369:
1370: \bibitem{bcf02} A. Benabdallah, J. G. Caputo and N. Flytzanis, Physica
1371: D 161, 79-101, (2002)
1372:
1373: \bibitem{bc02} A. Benabdallah and J. G. Caputo, J. of Applied Physics,
1374: {\bf 92}, nb. 7, (2002).
1375:
1376: \bibitem{lcti03} T. Lindstrom et al, Phys. Rev. Lett. {\bf 90}, 117021, (2003).
1377:
1378: \bibitem{l04} N. Lazarides, Phys. Rev. B {\bf 69}, 212501, (2004).
1379:
1380: \bibitem{gsgk04} E. Goldobin et al, Phys. Rev. Lett. {\bf 92}, 570051, (2004).
1381:
1382: \bibitem{r03} G. Rotoli, Phys. Rev. B {\bf 68}, 052505, (2003).
1383:
1384: \bibitem{sg04} H. Susanto and S. A. van Gils, Phys. Rev. B {\bf 69}, 92507,
1385: (2004).
1386:
1387: \bibitem{as01} E. Almaas and D. Stroud, Phys. Rev. B, {\bf 63}, 144522,
1388: , (2001).
1389:
1390: \bibitem{fp02} G. Filatrella and N. F. Pedersen, Physica C, 372-376,
1391: 11,
1392: (2002).
1393:
1394: \bibitem{bb91} S. P. Benz and C. J. Burroughs, Appl. Phys. Lett. {\bf
1395: 58},
1396: (1991).
1397:
1398: \bibitem{bcsl99} P. Barbara, A. B. Cawthorne, S. V. Shitov and C. J.
1399: Lobb,
1400: Phys. Rev. Lett. {\bf 82}, 1963, (1999).
1401:
1402: \bibitem{vbsl02} S. V. Vasilic, P. Barbara, S. V. Shitov and C. J.
1403: Lobb,
1404: Phys. Rev. B {\bf 65}, 180503, (2002).
1405:
1406: \bibitem{hairer} E. Hairer, S. P. Norsett and G. Wanner. S{\it olving
1407: ordinary differential equations I} (Springer-Verlag, 1987).
1408:
1409: \bibitem{carslaw} H. S. Carslaw, {\it An introduction to the theory of
1410: Fourier series and integrals }, 3rd revised edition, Dover (1950).
1411:
1412: \bibitem{lamb} H. Lamb, {\it Elements of soliton theory}, Wiley,
1413: (1983).
1414:
1415: \bibitem{report} L. Maurice, {\it Statique de jonctions Josephson dans
1416: un
1417: microstrip}, Memoire de DEA, Insa de Rouen, (2003).
1418:
1419: \bibitem{Boussaha} M.F. Boussaha, {\it D\'eveloppement d'un
1420: d\'emonstrateur
1421: de r\'ecepteur h\'et\'erodyne submillim\'etrique ultra-large bande \`a
1422: base de jonctions Supraconducteur-Isolant-Supraconducteur (SIS),
1423: refroidi \`a 4,2K.}, PhD thesis, Universit\'e de
1424: Paris 6, (2003).
1425:
1426: \bibitem{cg04} J. G. Caputo and Y. Gaididei, Physica C, {\it Two point
1427: Josephson
1428: junctions in a superconducting stripline: static case.},
1429: Physica C {\bf 402}, 160-173, (2004).
1430:
1431: \bibitem{many} J. G. Caputo and L. Loukitch, in preparation.
1432:
1433: \end{thebibliography}
1434:
1435: \section{Appendix}
1436:
1437: \subsection{Continuum limit.}
1438:
1439: We integrate (\ref{sggh}) from $a-h$ to $a+h$
1440:
1441: The first term yields
1442: \begin{eqnarray*}
1443: \int_{a-h}^{a+h} \phi_{tt} dx & = &
1444: \int_{a-h}^{a+h} \phi_{tt}(a)+(x-a)\phi_{ttx}(a)+O(|x-a|^2) dx\\
1445: & = & 2h\phi_{tt}(a) + 4h^2\phi_{ttx}(a) + O(h^3)\\
1446: \end{eqnarray*}
1447: The second one
1448: $$\int_{a-h}^{a+h} \phi_{xx} dx=\phi_x(a+h)-\phi_x(a-h)$$
1449: and the third one
1450: \begin{eqnarray*}
1451: \int_{a-h}^{a+h}\left [\sin(\phi)+ \alpha\phi_t + c \phi_{tt} \right]
1452: dx & = & \int_{a-h}^{a+h}\sin\left(\phi(a)+(x-a)\phi_x(a)+
1453: O_1(|x-a|^2)\right)\\
1454: & & + \alpha \phi_t(a)+ \alpha (x-a)\phi_{xt}(a) + O_2(|x-a|^2)\\
1455: & & + c\phi_{tt}(a) + c(x-a)\phi_{xtt}(a) + O_3(|x-a|^2)dx\\
1456: & = & \int_{a-h}^{a+h}\sin\left(\phi(a)+(x-a)\phi_x(a)
1457: + O_1(|x-a|^2)\right)dx\\
1458: & & + \alpha \phi_t(a) + \alpha h \phi_{xt}(a)
1459: + c\phi_{tt}(a)+ ch \phi_{xtt}(a) + O_4(h^2)\\
1460: \end{eqnarray*}
1461: We now expand the sine term
1462: \begin{eqnarray*}
1463: \sin\left(\phi(a)+(x-a)\phi_x(a)\right)&=&
1464: \sin(\phi(a))\cos((x-a)\phi_x(a))\\
1465: &&+\sin((x-a)\phi_x(a))\cos(\phi(a))
1466: \end{eqnarray*}
1467: so that
1468: \begin{eqnarray*}
1469: &&\int_{a-h}^{a+h}\sin\left(\phi(a)+(x-a)\phi_x(a)\right)\\
1470: &=&\int_{a-h}^{a+h}\sin(\phi(a))\cos((x-a)\phi_x(a))+
1471: \sin((x-a)\phi_x(a))\cos(\phi(a))dx\\
1472: &=&\sin(\phi(a)){2 \over \phi_x(a)} \sin(h\phi_x(a))-0
1473: \end{eqnarray*}
1474: Consider the limit of this term when $h\rightarrow 0$.
1475: We have:
1476: $$ d_a {1 \over 2 h}{2 \over \phi_x(a)}\sin(h\phi_x(a)) \rightarrow d_a,$$
1477: so that the equation for the point junction is
1478: $$ \phi_x(a+h)-\phi_x(a-h) + d_a(\sin(\phi(a))+ \alpha\phi_t(a) +
1479: c \phi_{tt} (a))=j,$$
1480: which is consistent with the model (\ref{sgcd}).
1481: $$\phi_{tt}- \phi_{xx} + d_a\delta(x-a) \left( \sin(\phi) +
1482: \alpha\phi_t +
1483: c \phi_{tt} \right)=j $$
1484:
1485: \subsection{Method of lines}
1486:
1487: The basis of the method is to discretize the spatial part of the
1488: operator and keep the temporal part as such. We thereby transform
1489: the partial differential equation into a system of ordinary
1490: differential equations. This method allows to increase the
1491: precision of the approximation in time and space independently and
1492: easily. In our case the operator is a distribution so that the
1493: natural way to give it meaning is to integrate it over a volume.
1494: We therefore choose as space discretisation the finite volume
1495: approximation where the operator is integrated over reference volumes.
1496: The value of the function is assumed constant in each volume.
1497:
1498:
1499: As solver for the system of differential equations, we use
1500: the Runge-Kutta method of order 4-5 introduced by Dormand and Prince
1501: implemented as the Fortran code DOPRI5 by Hairer and Norsett
1502: \cite{hairer}
1503: which enables to control the local error by varying the time-step.
1504:
1505:
1506: We first transform (\ref{sgd}) into a system of first order partial
1507: differential equations
1508: We write $\psi(x,t)=\phi_t(x,t)$.
1509: \begin{equation} \label{sysdiscret}
1510: \left\{ \begin{array}{r c l}
1511: \psi(x,t) & = & \phi_t(x,t) \\
1512: \psi_t(x,t) & = & \phi_{xx}(x,t)-\delta(x-a) (d_a \sin(\phi(x,t)) +
1513: \alpha \psi(x,t))+j
1514: \end{array} \right.
1515: \end{equation}
1516: with the boundary conditions :
1517: $\phi_x|_{l \over 2}=\phi_x|_{-{l \over 2}}=0$.
1518:
1519: For simplicity we will describe the implementation of
1520: the finite volume discretisation in the case of a single junction.
1521: We introduce reference
1522: volumes $V_k$ whose centers we call $x_k$, $1\leq k \leq nn$.
1523: The discretisation points are placed such that
1524: the point $x_{ng+1}$ is at the junction, ($x_{ng+1}=a$).
1525: We thus define $x_k$ and $V_k$ using the following identities
1526: $$V_k=\left ]x_k-{h_g \over 2},x_k+{h_g \over 2}
1527: \right[~,~~~~0<k<ng+1$$
1528: with $(ng+1)h_g=a$
1529: $$V_k=\left ]x_k-{h_d \over 2},x_k+{h_d \over 2}
1530: \right[~,~~~~ng+1<k<nn+1$$
1531: with $(nn-ng)h_d=l-a$. Finally at the junction, $k=ng+1$
1532: $$V_{k_{ng+1}}=\left ]x_{ng+1}-{h_g \over 2},
1533: x_{ng+1}+{h_d \over 2} \right[.$$
1534: $nn$, $ng$ and $nd$ are respectively the total number of discretisation
1535: points, the number of points to the left of the junction and the number
1536: of points to the right.
1537:
1538: For a fixed t, we assume $\phi(x,t)$ to be constant on each volume $V_k$,
1539: so that
1540: $$\int_{x_k-{h \over 2}}^{x_k+{h \over 2}} \phi(x,t) dx =
1541: h \phi(x_k,t)~{\rm,~~with}~h=hg~{\rm or}~h=hd$$
1542: Integrating over $V_k$ yields:
1543: \begin{enumerate}
1544: \item In the linear part of the PDE : $0<k<nn+1$ and $k \neq ng+1$:
1545: \begin{equation} \label{spacediscr}
1546: \left \{ \begin{array}{r c l}
1547: \psi(x_k,t) & = & \phi_t(x_k,t) \\
1548: \psi_t(x_k,t) & = &
1549: \frac{\phi(x_{k+1},t)-2\phi(x_k,t)+\phi(x_{k-1},t)}{h^2} + j
1550: \end{array} \right.
1551: \end{equation}
1552: with $h=hg$ for $0<k<ng+1$ or $h=hd$ for $k>ng+1$. We recognize the usual
1553: discretisation of the second derivative.
1554:
1555: \item At the junction: $k=ng+1$, we obtain
1556: \begin{eqnarray*}\label{disconti}
1557: \int_{x_{ng+1}-{h_g \over 2}}^{x_{ng+1}+{h_d \over 2}}
1558: \delta(x-a)\left(d_a \sin(\phi(x,t)) + \alpha\phi_t(x,t)\right)=\\
1559: d_a \sin(\phi(x_{ng+1},t)) + \alpha\phi_t(x_{ng+1},t)
1560: \end{eqnarray*}
1561: So that the final system is:
1562: \begin{equation} \label{joncdiscr}
1563: \left \{ \begin{array}{r c l}
1564: \psi(x_{ng+1},t) & = & \phi_t(x_{ng+1},t) \nonumber\\
1565: \psi_t(x_{ng+1},t) & = & \frac{4}{hg+hd}\left (\frac{\phi(x_{ng+2},t)-
1566: \phi(x_{ng+1},t)}{hg/2} -\frac{\phi(x_{ng+1},t)-\phi(x_{ng},t)}
1567: {hd/2} \right)\\
1568: & & -\frac{2}{hg+hd}\left(d_a \sin(\phi(x_{ng+1},t))
1569: +\alpha\phi_t(x_{ng+1},t)\right) + j
1570: \end{array} \right.
1571: \end{equation}
1572: \end{enumerate}
1573: The ODE system (\ref{spacediscr}, \ref{joncdiscr}) is then integrated
1574: numerically using the DOPRI5 integrator.
1575:
1576: \begin{figure}
1577: %\centerline{ \epsfig{file=figs/microstrip.eps,height=7 cm,width=10 cm,angle=0}}
1578: \centerline{ \epsfig{file=f1.eps,height=7 cm,width=10 cm,angle=0}}
1579: \caption{Left top pannel, a schematic drawing of a narrow
1580: 2D microstrip line containing
1581: two small Josephson junctions, right top panel section at one of the junction.
1582: The bottom panel presents the equivalent 1D electric circuit.}
1583: \label{f1}
1584: \end{figure}
1585:
1586: \begin{figure}
1587: %\centerline{\epsfig{file=figs/phiajoncohm.ps,height=14 cm,width=5.5cm,angle=270}}
1588: \centerline{\epsfig{file=f2.ps,height=14 cm,width=5.5cm,angle=270}}
1589: \caption{plot of $\phi_a(t)$ for $j = 0.137414$ for a junction position
1590: $a=3.43$, $d_a=1$ and length $l=10$. The left panel shows
1591: the ohmic mode regime and the
1592: right panel the junction mode regime. The numerical solution of
1593: (\ref{sgd}) is shown with the crosses while the analytical estimates
1594: (\ref{som}) and (\ref{sjm}) are in dashed line. }
1595: \label{f2}
1596: \end{figure}
1597:
1598: \begin{figure}
1599: %\centerline{\epsfig{file=figs/ohm1jonc.ps,height=15 cm,width=5cm,angle=270}}
1600: \centerline{\epsfig{file=f3.ps,height=15 cm,width=5cm,angle=270}}
1601: \caption{Instantaneous voltage $\phi_t\left(x,t\right)$
1602: as a function of $x$ for different times for the ohmic mode: left
1603: only oscillation for the junction shown in Fig. \ref{f2} placed at
1604: $a=3.43$ (left panel) and left and right oscillation for a junction placed at
1605: $a=2$ (right panel). The average voltage is indicated in both figures, $V=j l
1606: =1.37$ in the left panel (same as in Fig. \ref{f2})
1607: and $V=\frac{\pi}{2}\left(=j l \right)$ on the right panel.
1608: The other parameters are $\alpha=1$, and $d_a=1$.}
1609: \label{f3}
1610: \end{figure}
1611:
1612: \begin{figure}
1613: %\centerline{\epsfig{file=figs/phitjonc.ps,height=8 cm,width=5cm,angle=270}}
1614: \centerline{\epsfig{file=f4.ps,height=8 cm,width=5cm,angle=270}}
1615: \caption{Instantaneous voltage $\phi_t(x,t)$ in the
1616: junction mode as
1617: a function of $x$ for 60 values of time equidistributed between 0 and
1618: 10.
1619: The parameters are the same as for Fig. 1. The average voltage $V$ is
1620: indicated on the vertical axis and the junction's position is indicated
1621: on horizontal axis.}
1622: \label{f4}
1623: \end{figure}
1624:
1625:
1626: \begin{figure}
1627: %\centerline{\epsfig{file=figs/a0_a4fig.ps,height=15. cm,width=6cm,angle=270}}
1628: \centerline{\epsfig{file=f5.ps,height=15. cm,width=6cm,angle=270}}
1629: \caption{Two IV characteristic for a single junction placed in the
1630: center of the microstrip at $a=5$
1631: (left panel) and near the right edge at $a=9$ (right panel).
1632: Only the positions of the junction modes are indicated in units of
1633: $\frac{\pi}{10}$. The other ticks correspond to the ohmic modes.
1634: For each panel two curves are presented, the top one is for increasing
1635: current
1636: while the bottom one is for a decreasing current showing a clear
1637: hysteresis.
1638: The minimum value of the current is 0, the maximum is 4 and
1639: the stepping is $\frac{4}{99}$. The other parameters are
1640: $d_a=\alpha=1$.
1641: }
1642: \label{f5}
1643: \end{figure}
1644:
1645: \begin{figure}
1646: %\centerline{\epsfig{file=figs/nrja0_a4.ps,height=16 cm,width=9 cm,angle=270}}
1647: \centerline{\epsfig{file=f6.ps,height=16 cm,width=9 cm,angle=270}}
1648: \caption{Plot as a function of time, in each panel, from top to bottom
1649: of the total energy $E_t$, the energy in the passive region $E_p$, the
1650: power dissipated by the quasi-particles $P_q$ and the Josephson energy
1651: $E_j$ for a centered junction $a=5$ (bottom panels) and an off-centered
1652: junction $a=9$ (top panels). The left panels show the junction mode $n=4$
1653: and the right panels the corresponding ohmic mode for the same current.
1654: The averages of $E_t$ and $E_j$ are shown on the $y$ axis.}
1655: \label{f6}
1656: \end{figure}
1657:
1658:
1659: \begin{figure}
1660: %\centerline{\epsfig{file=figs/hautvolt2j.ps,height=6.5 cm,width=4.5 cm,angle=270}}
1661: \centerline{\epsfig{file=f7.ps,height=6.5 cm,width=4.5 cm,angle=270}}
1662: \caption{High voltage case: plot of $\phi(x,t)$ as a function of
1663: $x$ pour 6 consecutive times separated by $\Delta t=0.1$. The
1664: two junctions are located at $a_1=2$ and $a_2=6.2$. The other
1665: parameters are $l=10, j=1,$ and $\alpha =0.5$.}
1666: \label{f7}
1667: \end{figure}
1668:
1669: \begin{figure}
1670: %\centerline{\epsfig{file=figs/modohmic.ps,height=16 cm,width=6. cm,angle=270}}
1671: \centerline{\epsfig{file=f8.ps,height=16 cm,width=6. cm,angle=270}}
1672: \caption{Three types of ohmic modes obtained for two junctions
1673: in a microstrip. The positions of the junctions are given
1674: in the pictures. We plot $\phi_t\left(x,t\right)$ $\forall x \in [0,l]$
1675: for six consecutive times. The parameters of equation
1676: (\ref{sgd}) are $\lambda=0.5$ and $\alpha=1$.
1677: All figures have been obtained with the same current $j l=2.6927937$.}
1678: \label{f8}
1679: \end{figure}
1680:
1681: \begin{figure}
1682: %\centerline{\epsfig{file=figs/a0b1.23fig.ps, height=16 cm, width=6 cm,angle=270}}
1683: \centerline{\epsfig{file=f9.ps, height=16 cm, width=6 cm,angle=270}}
1684: \caption{IV characteristic curves for two junctions placed in $a_1=5$
1685: and $a_2=6.2$. The critical current densities are $d_1=0.7,~d_2=0.3$ for
1686: the left panel, $0.5,~0.5$ for the middle panel and $0.3,~0.7$ for the
1687: right panel. We plot on all figures $j l=\sqrt{V^2+1}$, $j l =V$,
1688: $j l =d_1\sqrt{V^2+1}+d_2 V$ and $j l =d_2\sqrt{V^2+1}+d_1V$.}
1689: \label{f9}
1690: \end{figure}
1691:
1692: \begin{figure}
1693: %\centerline{\epsfig{file=figs/mix.ps,height=15 cm,width=10 cm,angle=270}}
1694: \centerline{\epsfig{file=f10.ps,height=15 cm,width=10 cm,angle=270}}
1695: \caption{Plots of the phases at the junctions $\phi(a_1,t)$ and
1696: $\phi(a_2,t)$ as a function of time for the IV curves in the
1697: two left panels of Fig. \ref{f9}. The left panels are
1698: for $d_1=0.7,d_2=0.3$ and the right panels for $d_1=d_2=0.5$.
1699: The top panels show the resonance of order 3 while the bottom
1700: panels show the resonance of order 4. Note how the junctions
1701: switch between ohmic mode and junction mode. The analytic
1702: expressions for the corresponding junction modes (\ref{sjm})
1703: are shown in dashed line.}
1704: \label{f10}
1705: \end{figure}
1706:
1707: \begin{figure}
1708: \centerline{
1709: %\epsfig{file=figs/a-3b3_0-0_0-pi.ps,height=16 cm,width= 6 cm,angle=270}}
1710: \epsfig{file=f11.ps,height=16 cm,width= 6 cm,angle=270}}
1711: \caption{IV curves for two symmetrically placed junctions
1712: $a_1=2$ et $a_2=8$.The scale on the $x$ axis is in units of
1713: $\frac{\pi}{l}$. In the left panel the junctions are identical while
1714: in the right panel the second one is a $\pi$-junction.
1715: Note the resonance on the even harmonics $\frac{2k\pi}{l}$ in the left
1716: panel and on the odd harmonics $\frac{(2k+1)\pi}{l}$ in the right panel.
1717: We could have obtained the same change of IV curves for the
1718: same array of standard junctions by an appropriate choice of the
1719: magnetic field (see text).}
1720: \label{f11}
1721: \end{figure}
1722:
1723: \end{document}
1724:
1725: