cond-mat0203208/odv.tex
1: \documentclass[aps,prb,twocolumn,showpacs,showkeys,groupedaddress,floatfix,amsmath,amssymb]{revtex4}
2: %\documentclass[aps,prb,preprint,showpacs,showkeys,groupedaddress,floatfix,amsmath,amssymb]{revtex4}
3: 
4: \usepackage{graphicx}
5: \usepackage{dcolumn}
6: \usepackage{bm}
7: 
8: 
9: \begin{document}
10: 
11: 
12: \title{Stationary phase slip state in quasi-one-dimensional rings.}
13: 
14: \author{D. Y. Vodolazov}
15: \author{B. J. Baelus}
16: \author{F. M. Peeters}
17: \email{peeters@uia.ua.ac.be} \affiliation{Departement Natuurkunde,
18: Universiteit Antwerpen (UIA), Universiteitsplein 1,  B-2610
19: Antwerpen, Belgium}
20: 
21: 
22: %\date{\today}
23: 
24: 
25: \begin{abstract}
26: The nonuniform superconducting state in a ring in which the order
27: parameter vanishing at one point is studied. This state is
28: characterized by a jump of the phase by $\pi$ at the point where
29: the order parameter becomes zero. In uniform rings such a state is
30: a saddle-point state and consequently unstable. However, for
31: non-uniform rings with e.g. variations of geometrical or physical
32: parameters or with attached wires this state can be stabilized and
33: may be realized experimentally.
34: \end{abstract}
35: 
36: \pacs{74.60Ec, 74.20.De, 73.23.-b} \keywords{phase slip,
37: mesoscopic superconductivity}
38: 
39: \maketitle
40: 
41: In recent years \cite{Berger,Horane,Berger1} the existence of a
42: single-connected state in a ring was proposed theoretically. Such
43: a state implies that the relation between the phase $\phi$ of the
44: order parameter $\psi=fe^{{\rm i}\phi}$, the current density $j$
45: and the magnetic flux $\Phi$ through the ring (which follows from
46: the single valuedness of $\psi$ - see for example Ref.
47: \cite{Tinkham})
48: \begin{equation}
49: \oint \frac{j}{f^2}ds=2\pi n-\Phi,
50: \end{equation}
51: is no longer valid (the flux is expressed in $\Phi_0/2\pi$, the
52: current density in $j_0=c\Phi_0/4\pi^2\lambda^2\xi$, $\lambda$ is
53: the London penetration length, $\xi$ the coherence length and
54: $\Phi_0$ is the quantum of magnetic flux). The reason is that the
55: order parameter in such a single-connected state is zero at one
56: point. Moreover, it was claimed that under certain conditions,
57: i.e. radius of the ring less than $\xi$ and the flux through the
58: ring is about $(n+1/2)\Phi_0$ this state may become metastable in
59: some range of magnetic fields \cite{Horane}.
60: 
61: In this paper we revisited this problem and we will show that a
62: state where the order parameter vanishes in one point is still
63: double-connected and that Eq. (1) is valid for such a state. We
64: propose to call such a state a one-dimensional vortex state (ODV
65: state), because like for an ordinary two-dimensional Abrikosov
66: vortex, there is a jump in the phase of the order parameter of
67: $\pi$ at the point where $\psi=0$. In contrast to a
68: two-dimensional Abrikosov vortex the 1D vortex is an unstable
69: structure in an uniform ring. However, if some inhomogeneities are
70: present in the ring (defects, nonuniform thickness or nonuniform
71: width of the ring, attached superconducting wires) the ODV
72: structure can be stabilized and may consequently be realized
73: experimentally.
74: 
75: Consider a uniform ring with thickness $d \lesssim \lambda$ and
76: width $w \lesssim \xi$. In addition, let the radius of the ring
77: $R$ be much larger than $w$. Under these conditions we can neglect
78: the screening effects and the problem is reduced to a
79: one-dimensional one. The distribution of the current density $j$
80: and the order parameter $\psi$ of the system at a temperature not
81: far from the critical temperature $T_c$ is described by the 1D
82: Ginzburg-Landau equation (plus the condition ${\rm div} {\bf
83: j}=0$)
84: \begin{subequations}
85: \begin{equation}
86:  \frac{d^2f}{ds^2}+f(1-f^2-p^2)=0, %2a
87: \end{equation}
88: \begin{equation}
89:  \frac{dj}{ds}=\frac{d}{ds}f^2p=0, %2b
90: \end{equation}
91: \end{subequations}
92: where the gauge-invariant momentum $p=\nabla \phi -A$ is scaled in
93: units of $\Phi_0/(2\pi\xi)$, the length of the ring is $L=2\pi R$
94: and the circular coordinate $s$ is in units of the coherence
95: length $\xi$. In these units, the magnetic field is scaled in
96: units of the second critical field, $H_{c2}$, and the magnetic
97: flux in $\Phi_0/2\pi$.
98: 
99: The coupled Eqs. (2a) and (2b) have to be solved with the boundary
100: condition $\psi(-L/2)=\psi(L/2)$. We  use the method proposed in
101: \cite{Langer} (see also \cite{Zhang}). These equations have the
102: first integral
103: \begin{equation}
104: \frac{1}{2}\left(\frac{df}{ds}\right)^2+\frac{f^2}{2}-\frac{f^4}{4} %3
105: +\frac{j^2}{2f^2}=E.
106: \end{equation}
107: From a formal point of view, Eq. (3) is nothing else then the
108: condition of "energy", $(E)$, conservation for some "particle"
109: with "coordinate" $f$ and "momentum" $j$. The role of "time" is
110: played \cite{Langer} by the circular coordinate $s$ . In Fig. 1 we
111: show the dependence of the "potential energy" of this system
112: \begin{equation}
113: U(f)=\frac{f^2}{2}-\frac{f^4}{4}+\frac{j^2}{2f^2}, %4
114: \end{equation}
115: for different values of $j$. Possible solutions of Eqs. (2a,b) are
116: in the region where confined "trajectories" of our virtual
117: particle exist. This is possible for currents $0<j\leq j_c$
118: ($j_c=\sqrt{4/27}j_0$ is the depairing current density ).
119: \begin{figure}[ht]
120: \includegraphics[width=0.48\textwidth]{fig1.eps}
121: \caption{Dependence of the "potential energy" $U(f)$ for different
122: values of the current $j$: 1- $j$=$0.01$, 2- $j$=$0.05$, 3-
123: $j$=$0.15$, 4- $j$=$j_c$=$\sqrt{4/27}$.}
124: \end{figure}
125: Using Eq. (3) we immediately can write the solution of Eqs. (2a,b)
126: for the ring
127: \begin{eqnarray}
128: \sqrt{2}s=\int_{t_0}^{t}\frac{dt}{\sqrt{(t-t_0)(t_1-t_0)(t_2-t_0)}}= \nonumber\\
129: \frac{2}{\sqrt{t_2-t_0}}F\left({\rm sin^{-1}}\sqrt{\frac{t-t_0} %5
130: {t_1-t_0}},\sqrt{\frac{t_1-t_0}{t_2-t_0}}\right),
131: \end{eqnarray}
132: where $t(s)=f^2(s)$, $F(\theta,m)$ is the elliptic integral of the
133: first kind and $t_0\leq t_1\leq t_2$ are the solutions of the
134: cubic equation
135: \begin{equation}
136: t^3-2t^2+4Et-2j^2=0. %6
137: \end{equation}
138: Using the boundary condition for $f$ we may conclude, that for a
139: given current there exist a maximum of three solutions. There are
140: two uniform solutions (in the points of the minimum $E=E_{min}$
141: and the maximum $E=E_{max}$ of the "potential energy"(4)) and one
142: nonuniform solution with energy $E_{min}<E<E_{max}$ which is
143: defined by the equation
144: \begin{equation}
145: L=\sqrt{\frac{8}{t_2-t_0}}K\left(\sqrt{\frac{t_1-t_0}{t_2-t_0}}\right), %7
146: \end{equation}
147: where $K(m)$ is the complete elliptic integral of the first kind.
148: Numerical analysis of Eq. (7) for $0<j<j_c$ shows that there is a
149: minimal value for the ring length $L_{min}$ for which there exists
150: a solution for this equation. When $j\to j_c$, $L_{min} \to
151: \infty$ and in the opposite limit $j \to 0$ one can show that
152: $L_{min} \to \pi$. The latter corresponds to a ring radius $R=1/2$
153: (or $\xi/2$ in dimensional units). For a radius $R>1/2$ metastable
154: states may exist and superconductivity is present for any value of
155: the magnetic flux (at least in our one-dimensional model)
156: \cite{Horane,Buzdin}.
157: 
158: In principle, Eqs. (5-7) define the nonuniform solution of Eqs.
159: (2a,b) for a ring. Unfortunately, even using the explicit Kordano
160: expressions for the roots of Eq. (6) the results are rather
161: complicated for arbitrary values of the current $j$. However a
162: tractable analytical solution is possible if we consider the case
163: of low currents $j \ll 1$ for which the roots of Eq. (6) are
164: simplified to
165: \begin{subequations}
166: \begin{equation}
167:  t_0 \simeq j^2/2E ,  %8a
168: \end{equation}
169: \begin{equation}
170:  t_1\simeq 1-\sqrt{1-4E} , %8b
171: \end{equation}
172: \begin{equation}
173: t_2\simeq 1+\sqrt{1-4E} . %8c
174: \end{equation}
175: \end{subequations}
176: After inserting these results into Eq. (7) we obtain the "energy".
177: For $L \gg \pi$ the "energy" $E\simeq 1/4$ is practically
178: independent of $L$ (e.g. for $L=8$ the difference $1/4-E(L=8)$ is
179: less than 0.002). Also in the limit $L-\pi \ll 1$ we found that
180: the "energy" $E$ is independent of $j$
181: \begin{equation}
182: E(L) \simeq \frac{1}{2}\left(\frac{L-\pi}{\pi/8}\right)^2. %9
183: \end{equation}
184: In the limit $j \ll 1$ we can also find \cite{comm1} the
185: dependence of $t(s)$ near the minimum point of $t(s)$
186: \begin{subequations}
187: \begin{eqnarray}
188: t(s)=2j^2+s^2/2, & L\gg \pi, s\ll 1 \label{a}%10a
189: \\
190: t(s)=j^2/2E+2Es^2, &  L-\pi \ll 1,  s\ll 1 \label{b} %10b
191: \end{eqnarray}
192: \end{subequations}
193: Using Eqs. (10a,b) it is easy to show, that in the limit $j \to 0$
194: the gauge-invariant momentum is given by
195: \begin{equation}
196: p(s)=j/t(s)=\frac{j}{|j|}(\pi+2\pi m)\delta(s), %11
197: \end{equation}
198: where $\delta(s)$ is the Dirac function and $m$ is a integer. As a
199: result the integral
200: \begin{eqnarray}
201: \lim\limits_{j\to 0}\oint \frac{j}{f^2}ds=\lim\limits_{j \to
202: 0}\oint p(s)ds= \nonumber
203: \\
204: =\phi(+\epsilon)-\phi(-\epsilon)=\Delta \phi=\pm %12
205: (\pi+2\pi m),
206: \end{eqnarray}
207: for arbitrary ring size \cite{comm2}. Combining Eq. (12) with Eq.
208: (1) it is easy to show that if $\Phi=(k+1/2)\Phi_0$ ($k$ is an
209: integer) a solution of Eq. (2a) can be found which vanishes at one
210: point.
211: 
212: In previous \cite{Berger,Horane,Berger1} studies of this state Eq.
213: (11) was not taken into account (only the absolute value of the
214: order parameter was found). Furthermore, in Ref. \cite{Horane} it
215: was claimed that for small rings there is a magnetic field region
216: in which such a state exist and is metastable. But we find that in
217: this region this state even does not exist (except in one point)
218: because Eq. (1) is not fulfilled!
219: 
220: The distribution of the order parameter as obtained in the present
221: paper may also be applied to a ring in which some part of the ring
222: consists of a normal metal. Then at the boundary between the
223: normal metal and the superconductor the condition $\psi \simeq 0$
224: is fulfilled (if the normal part is longer than $\xi$) and the
225: distribution of the density $|\psi(s)|^2$ coincides with the one
226: obtained in Ref. \cite{Horane} (but the phase $\phi(s)$ will be
227: different for our previous ring geometry). Besides we cannot call
228: such a state single-connected as in Ref.
229: \cite{Berger,Horane,Berger1} because the relation (1) is valid
230: even for this case. Therefore, it is better to call this state a
231: one-dimensional vortex state because like the two-dimensional
232: Abrikosov vortex there is a point where $|\psi|=0$ and the phase
233: of the order parameter exhibits a jump equal to $k\pi$ (as the
234: Abrikosov vortex with orbital momentum $2\pi k$). A numerical
235: calculation shows that, as in the case for an ordinary Abrikosov
236: vortex in superconductors where such a state is stable, the phase
237: jump is always equal to $\pi$, i.e. $k=1$.
238: \begin{figure}[h]
239: \includegraphics[width=0.5\textwidth]{fig2.eps}
240: \caption{Distribution of the absolute value (a) and phase (b) of
241: the order parameter at different values of the magnetic flux: 1-
242: $\Phi$=$0.45\Phi_0$, 2- $\Phi$=$0.49\Phi_0$, 3-
243: $\Phi$=$0.498\Phi_0$, 4- $\Phi$=$(0.5-0)\Phi_0$, 5-
244: $\Phi$=$(0.5+0)\Phi_0$, 6- $\Phi$=$0.502\Phi_0$, 7-
245: $\Phi$=$0.51\Phi_0$, 8- $\Phi$=$0.55\Phi_0$. Note that the order
246: parameter for $\Phi=\Phi_0/2+\alpha$ is the same as for
247: $\Phi=\Phi_0/2-\alpha$ when $\alpha<\Phi_0/2$. In the insets (a)
248: $f_{min}$ and (b) the phase difference $\Delta \phi$ near the
249: point $s/L=3/4$ are shown as function of $\Phi$.}
250: \end{figure}
251: 
252: Our numerical analysis of the time-dependent Ginzburg-Landau
253: equations showed that this state is completely unstable for a
254: uniform ring. However for rings with nonuniform width(thickness)
255: or attached wires a nonuniform distribution of the order parameter
256: becomes possible \cite{Berger,Berger1,Fink1,Fink2} and for small
257: rings with $L \sim \pi$ this state is realized in practice
258: (besides for the case of a ring, a ODV state can also exist in the
259: Wheatstone bridge - see Refs. \cite{Fink3,Ammann}). In Fig. 2 the
260: distribution of the absolute value and the phase of the order
261: parameter is shown for different values of the magnetic field for
262: a nonuniform ring \cite{comm3}. We used the following model where
263: the width of the ring was varied as
264: \begin{equation}
265: w(s)=1+w_0{\rm sin}\left(\frac{2\pi s}{L}\right), %13
266: \end{equation}
267: The parameters used are: $w_0=0.1$ and $L=3.25$. It is seen that
268: with increasing magnetic flux the magnitude of the order parameter
269: $f$ decreases in the thinner part of the ring (at $s/L=3/4$) and
270: becomes zero when $\Phi/\Phi_0=1/2$. In Ref. \cite{Berger1} it was
271: found that the ODV state is only possible at $\Phi=(n+1/2)\Phi_0$.
272: The reason is now clear - only at this values of the magnetic flux
273: the phase difference near the point where the order parameter is
274: zero will be compensated by the term $2\pi n-\Phi$ and the current
275: density will be equal to zero. It is interesting to note that the
276: current density is also zero for $\Phi=n\Phi_0$ even in the case
277: of a nonuniform ring, but for values of the flux where the ODV
278: state does not exist. At these values of the magnetic flux the
279: order parameter is uniform along the ring and the term $2\pi n$ is
280: completely compensated by the term $\Phi$ in Eq. (1).
281: 
282: \begin{figure}[hbtp]
283: \includegraphics[width=0.48\textwidth]{fig3.eps}
284: \caption{The current in the ring (with attached wire of length
285: $5\xi$) as function of the flux. The different curves are for
286: different circumference of the ring: 1- $L$=1, 2- $L$=2, 3- $L$=3,
287: 4- $L$=3.5, 5- $L$=4. For $L$=4 the ODV state does not exist in
288: the ring and hysteresis appears. In the inset the dependence of
289: $dj/d\Phi$ on the flux $\Phi$ is shown for rings with lengths
290: $L$=1, 2, 3, 3.5 (curves 1,2,3,4 respectively). Note that there
291: are two maxima in the range $(0,1)$. With increasing length the
292: two maxima merge into one and when one approaches the critical
293: length $dj/d\Phi$ diverges at $\Phi=\Phi_0/2$.}
294: \end{figure}
295: 
296: Above we showed that the ODV state can be realized by varying the
297: geometrical parameters of the ring. But there are two alternative
298: approaches to realize the ODV state in the ring experimentally.
299: First, it is possible to include an other phase in the ring. This
300: leads to the appearance of a weak link in the sample and if the
301: radius of the ring is less than some critical value (about $\xi$)
302: it also leads to the existence of the ODV state in the ring at
303: $\Phi=(n+1/2)\Phi_0$. We modelled this situation by introducing an
304: additional term $\rho(s)f$ in Eq. (2a), and, as an example, we
305: took $\rho(s)=-\alpha \delta(s)$ and $\alpha=1$ (it should be
306: noted that qualitatively the results do not depend on the specific
307: value of $\alpha$). We will not present the numerical results of
308: the modified Ginzburg-Landau equations but we found that they are
309: qualitatively similar to the behavior of $f$ and $\phi$ shown in
310: Fig. 2. When approaching the flux $\Phi=(n+1/2)\Phi_0$ the order
311: parameter reaches zero in the defect point and a jump in the phase
312: equal to $\pi$ occurs. Secondly, such a ODV state should also
313: appear in small rings with attached wire(s). As was shown in Refs.
314: \cite{Fink1,Fink2} an attached wire leads to a nonuniform
315: distribution of the order parameter in the ring. In Fig. 3 the
316: dependence of the current $j$ in such a ring on the flux through
317: the ring is shown (see also Ref. \cite{Fink2}). As for nonuniform
318: rings, where the parameters are chosen such that the ODV state may
319: exist, there is no hysteresis in such a system \cite{Berger1}.
320: \begin{figure}[t]
321: \includegraphics[width=0.48\textwidth]{fig4.eps}
322: \caption{The Gibbs free energy and the magnetization (right lower
323: inset) of the asymmetric ring (upper inset) as function of the
324: applied magnetic field.}
325: \end{figure}
326: Small inhomogeneities in rings gives us the unique possibility to
327: study states close to the ODV state because there is no sharp
328: transition from the state with finite order parameter to a state
329: with vanishing order parameter at one point (see Fig. 2). The
330: latter is, in some sense, a 'frozen' slip phase state. A small
331: variation of the flux through the ring near $\Phi=\Phi_0/2$ leads
332: to a change of $\Delta \phi$ by $2\pi$ (from $-\pi$ to $+\pi$ with
333: a concomitant change of the current from $-0$ to $+0$ - see Fig.
334: 2(b)). As a result an additional phase circulation $\oint \nabla
335: \phi ds$=$2\pi$ appears in the system, not because the magnetic
336: flux $\Phi$ is changed, but because the term $\oint j/f^2 ds$
337: changes from $-\pi$ to $+\pi$ (see Eq. (1)). When such a jump
338: occurs $\nabla \phi$ does not change in the ring, except near the
339: point where $f=0$, and as a result the current density in the
340: system changes continuously. In the usual case transitions of the
341: vorticity from a state with phase circulation $2\pi n$ to a state
342: with phase circulation $2\pi (n+1)$ leads to a jump in the current
343: {\it everywhere} in the ring because $\oint \nabla \phi ds$
344: changes by $2\pi$ and the order parameter and the current density
345: are finite {\it everywhere} in the system.
346: 
347: To supplement the above study we made also a numerical study of a
348: nonuniform ring of finite width and thickness by implementing our
349: previous finite difference solution of the coupled nonlinear
350: Ginzburg-Landau equations \cite{Baelus1}. As an example we took
351: the following parameters: outer radius of the ring $R_o=2\xi$,
352: radius of the hole $R_i=1.5\xi$, displacement of the hole from the
353: center $a=0.4\xi$, ring thickness $d=0.001\xi$ and Ginzburg-Landau
354: parameter $\kappa=2$.
355: 
356: In Fig. 4 the dependence of the Gibbs free energy $F$ and the
357: magnetization $M=-\partial F/\partial H$ of this system on the
358: magnetic field are shown. As in the case of our one-dimensional
359: ring these dependencies are reversible and there are magnetic
360: fields at which the magnetization is equal to zero (in the points
361: of local maximum and minimum of the free energy). We found that
362: the distribution of the order parameter and the phase in the ring
363: (see inset of Fig. 5) is similar to analogical distributions for
364: the above one-dimensional ring (Fig. 2) at low magnetic fields.
365: But different with the one-dimensional system the phase
366: circulation increase of $2\pi$ (between the points (2) and (3) in
367: Fig. 5) does not occur at the magnetic field value where the free
368: energy has a local maximum (and zero magnetization). With
369: decreasing width of the ring this point shifts towards the local
370: maximum in the free energy.
371: \begin{figure}[t]
372: \includegraphics[width=0.35\textwidth,angle=-90]{fig5.eps}
373: \caption{The Gibbs free energy $F(H)$ near the first maximum. In
374: the inset the phase and the order parameter distribution at
375: different values of the applied magnetic field (indicated by the
376: squares on the $F(H)$ curve) are shown. The phase was calculated
377: along the outer perimeter of the ring. The open circle on $F(H)$
378: indicates the position at which the vorticity increases from 0 to
379: 1.}
380: \end{figure}
381: It is interesting to note that for the system corresponding to
382: Fig. 4 the vortex enters through the thinnest part of the ring in
383: the case of the first three maxima in the free energy while for
384: the highest magnetic field maxima (i.e. $H/H_{c2}\simeq 4.6$) the
385: vortex enters through the thickest part of the ring (see Fig. 6).
386: Because the width of thicker part of the ring ($=0.9\xi$) is
387: considerable larger than the thinner part ($=0.1\xi$) and is of
388: order $\xi$ the ODV for $H/H_{c2}\simeq 4.6$ becomes the usual
389: two-dimensional vortex. As a result the order parameter is equal
390: to zero only in one point along the radial coordinate and the
391: circulation of the phase of the order parameter now is also a
392: function of this coordinate (see Fig. 6). When the vortex
393: enters/exits the ring at low magnetic fields there is also a slow
394: dependence of the order parameter on the radial coordinate along
395: the thinnest part of the ring. Therefore, we can conclude that in
396: a ring with finite width the one-dimensional vortex is transformed
397: into the usual two-dimensional vortex.
398: 
399: \begin{figure}[h]
400: \includegraphics[width=0.36\textwidth,angle=-90]{fig6.eps}
401: \caption{The same as Fig. 5 but now for the last maximum in Fig.
402: 4. In the inset a contourplot of the order parameter is shown for
403: different values of the magnetic field which are given by the
404: squares in the main figure. $L_{outer}$ is the vorticity as
405: calculated along the outer perimeter of the ring and $L_{inner}$
406: is the vorticity calculated along the perimeter of the hole.}
407: \end{figure}
408: 
409: The process of the appearance of a stable vortex in a ring with
410: finite width is similar to the creation of a phase slip in wires
411: of finite width in the presence of a transport current. In the
412: latter case the distribution of the gauge-invariant momentum is
413: not exactly uniform over the wire width. The maximum value of $p$
414: is obtained at the edges and as a result the order parameter
415: vanishes first in these points. When a phase slip is created the
416: distribution of the order parameter is not uniform over the width.
417: With decreasing wire width this nonuniformity decreases but it
418: will be uniform, strictly speaking, only in the limit $w \to 0$.
419: With increasing wire width the phase slip transforms to the
420: ordinary process of vortex/antivortex pair nucleating at the
421: edges, penetrating deep into the superconductor and annihilating.
422: In nonuniform rings the distribution of $p$ over the width is
423: nonuniform and in contrast to wires with transport current it is
424: nonsymmetric with respect to the ring width. Suppression of the
425: order parameter first occurs only on the external side (when
426: increasing the magnetic field) or on the internal side (when
427: decreasing the magnetic field). As in the case of an
428: one-dimensional ring we may call this state a stable ("frozen")
429: phase slip state if the width of the ring, where the vortex
430: penetrates, is less than $\xi$.
431: 
432: The ODV state of a nonuniform ring is very similar to the
433: saddle-point state as found in Ref. \cite{Baelus1} in case of an
434: uniform ring of finite width and in Ref. \cite{Schweigert} for the
435: case of a disk. Note that for an uniform ring with zero width the
436: nonuniform solution Eq. (6) corresponds to a saddle-point point of
437: $F(\psi)$. In Fig. 7 the gradual decrease of the hysteresis with
438: increasing displacement $a$ of the hole from the center of the
439: ring is shown. It is seen that with increasing $a$ the region
440: where metastable states exist decreases and ultimately vanishes
441: for some critical displacement $a_c$. The saddle point states
442: (only shown for $a=0$, by the thin full curve) approaches the
443: stable phase slip state when $a=a_c$.
444: \begin{figure}[t]
445: \includegraphics[width=0.36\textwidth,angle=-90]{fig7.eps}
446: \caption{The Gibbs free energy $F(H)$ at low magnetic fields for
447: different displacements $a$ of the hole from the center of the
448: ring. The thin solid curve (for $a=0$) corresponds to the saddle
449: point state.}
450: \end{figure}
451: In Ref. \cite{Baelus2} also nonuniform superconducting rings of
452: finite width were investigated (but with larger ring radius). It
453: was found that for low magnetic fields $F(H)$ is irreversible but
454: for high magnetic fields the dependence $F(H)$ was reversible. In
455: high fields the transition from a state with vorticity one to
456: another state occured through the same scenario as discussed
457: above. A similar state was also considered in Ref. \cite{Berger3}
458: in the framework of the linearized Ginzburg-Landau equations for a
459: nonuniform ring of finite width. They found that in such a system
460: the vortex may be stable in some magnetic field region. In low
461: magnetic field this vortex enter through the thinner part of the
462: ring and at high magnetic fields through the thicker part. Our
463: calculations generalizes this result to lower temperatures (i.e.
464: where the nonlinear Ginzburg-Landau equations have to be used) and
465: with inclusion of the non-zero demagnetization factor of the ring.
466: We found the dependence of the Gibbs free energy (and
467: magnetization) of this sample on the applied magnetic field. Our
468: results also allowed to compare the results of the one-dimensional
469: model with the full two-dimensional one.
470: 
471: The ODV (or stable phase slip) state may be observed by magnetic
472: experiments. Magnetic susceptibility is proportional \cite{Zhang}
473: to $dj/d\Phi$ and the ODV state exhibits some characteristic
474: peculiarities as was shown in the inset of Fig. 3. Magnetization
475: $M$ is proportional to the current $j$ and hence $M(H)$ is
476: reversible, for samples where the ODV state exists (see Fig. 3 and
477: inset of Fig. 5). Furthermore $M=0$ at $\Phi\simeq (n+1/2)\Phi_0$.
478: Alternatively, because at $\Phi=(n+1/2)\Phi_0$ there is a point in
479: the ring where the order parameter is equal to zero, this state
480: may be found by transport measurements. For example in Ref.
481: \cite{Liu} the dependence of the resistance of a hollow cylinder
482: with radius of order $\xi$ was studied at temperature $T<T_c$ far
483: from $T_c$. At $\Phi=\Phi_0/2$ the resistance $(R)$ exhibited a
484: maximum but the value of $R$ was a factor three less than the
485: normal state resistance $R_n$. If the cylinder has a nonuniform
486: thickness the ODV state can be realized as a stable state when
487: $\Phi=\Phi_0/2$ and it will lead to a resistive (but
488: superconducting) state even for very small currents and naturally
489: the resistance of such a state will be less than $R_n$ as was
490: found by Liu et al \cite{Liu}.
491: 
492: In conclusion, we studied the nonuniform state in a
493: superconducting ring in which the order parameter vanishes at one
494: point. It was shown that this state is characterized by a jump in
495: the phase of the order parameter by $\pi$ near the point $\psi=0$.
496: It allows, in correspondence with the ordinary two-dimensional
497: Abrikosov vortex, to call such a state a one-dimensional vortex.
498: This state is unstable in an uniform ring. In case of a nonuniform
499: ring (with variations of the geometrical ($d$, $w$) or physical
500: parameters ($\xi$, $\lambda$) along the ring) or for a ring with
501: an attached wire, this state may be realized in practise. For
502: rings with non-zero width the ODV state transforms into the usual
503: two-dimensional vortex. We also showed that this state is the
504: remnant of the saddle point state connecting two superconducting
505: states with different vorticity as found in a uniform ring. The
506: latter state becomes stable in a nonuniform ring.
507: 
508: The work was supported by the Flemish Science Foundation (FWO-Vl),
509: the "Onderzoeksraad van de Universiteit Antwerpen," the
510: "Interuniversity Poles of Attraction Program - Belgian State,
511: Prime Minister's Office - Federal Office for Scientific, Technical
512: and Cultural Affairs," and the European ESF-Vortex Matter. One of
513: us (D.Y.V.) is supported by a postdoctoral fellowship of FWO-Vl
514: and partially by a RFBR grant N01-02-16593. Discussions with Prof.
515: A. Geim and comments from Prof. J. Berger are gratefully
516: acknowledged.
517: 
518: \begin{references}
519: 
520: \bibitem{Berger} J. Berger and J. Rubinstein, Phys. Rev. Lett. {\bf
521: 75}, 320 (1995).
522: 
523: \bibitem{Horane} E. M. Horane, J. I. Castro, G. C. Buscaglia, and A.
524: Lopez,  Phys. Rev. B {\bf 53}, 9296 (1996).
525: 
526: \bibitem{Berger1} J. Berger and J. Rubinstein, Phys. Rev. B {\bf 56},
527: 5124 (1997).
528: 
529: \bibitem{Tinkham} M. Tinkham, Phys. Rev. {\bf 129}, 2413 (1963).
530: 
531: \bibitem{Langer} J. S. Langer and V. Ambegaokar, Phys. Rev. {\bf 164},
532:  498 (1967).
533: 
534: \bibitem{Zhang} X. Zhang and J. Price, Phys. Rev. B {\bf 55}, 3128
535: (1997).
536: 
537: \bibitem{Buzdin} A. Bezryadin, A. Buzdin, and B. Pannetier, Phys. Rev.
538: B {\bf 51}, 3718 (1995).
539: 
540: \bibitem{comm1}Numerical analysis shows that for large rings $L\gg \pi$ the solution
541: $t(s)=t_1-(t_1 -t_0)/{\rm cosh}^2(s\sqrt{(t_1-t_0)/2})$ found in
542: \cite{Langer} works very well for arbitrary currents $j<j_c$. The
543: reason is that for rings with $R \gg 1$ the "energy" $E$ is close
544: to $E_{max}$ and the roots $t_1$ and $t_2$ practically coincide.
545: As a result the situation is analogous to the one considered in
546: Ref. \cite{Langer}.
547: 
548: \bibitem{comm2} This result is a generalization of a result obtained
549: in Ref. \cite{Langer} for the saddle-point of Eq. (2) for the case
550: of an infinite long wire with vanishing transport current.
551: 
552: \bibitem{Fink1} H. J. Fink and V. Grunfeld, Phys. Rev. B {\bf 33},
553: 6088 (1986).
554: 
555: \bibitem{Fink2} H. J. Fink and V. Grunfeld, Phys. Rev. B {\bf 31},
556: 600 (1985).
557: 
558: \bibitem{Fink3} H. J. Fink, Phys. Rev. B {\bf 45}, 4799 (1992); {\it ibid.}
559: {\bf 48}, 3579 (1993).
560: 
561: \bibitem{Ammann} C. Ammann, P. Erd\"os, and S. B. Haley,
562: Phys. Rev. B {\bf 51}, 11739 (1995).
563: 
564: \bibitem{comm3} It is interesting to note that the results shown in
565: Fig. 2 correspond qualitatively to the time evolution of the order
566: parameter for an uniform ring in the case the vorticity increases
567: with one unit. The only difference is that for an uniform ring the
568: current density, generally speaking, is not equal to zero at the
569: point where the order parameter vanishes. The reason is that for
570: time-dependent processes the full current density is equal to the
571: sum of the superconducting $j_s$ and the normal $j_n$ parts. In
572: the point $f=0$, $j_s=0$ but $j_n\neq 0$.
573: 
574: \bibitem{Baelus1} B. J. Baelus, F. M. Peeters, and V. A.
575: Schweigert, Phys. Rev. B {\bf 63}, 144517 (2001).
576: 
577: \bibitem{Schweigert} V. A. Schweigert and F. M. Peeters, Phys.
578: Rev. Lett. {\bf 83}, 2409 (1999).
579: 
580: \bibitem{Baelus2} B. J. Baelus, F. M. Peeters, and V. A.
581: Schweigert, Phys. Rev. B {\bf 61}, 9734 (2000).
582: 
583: \bibitem{Berger3} J. Berger and J. Rubinstein, Phys. Rev. B {\bf 59},
584: 8896 (1999).
585: 
586: \bibitem{Liu} Y. Liu, Yu. Zadorozhny, M. M. Rosario, B. Y. Rock,
587: P. T. Carrigan, and H. Wang, Science {\bf 294}, 2332 (2001).
588: 
589: \end{references}
590: 
591: \end{document}
592: