1: %\documentstyle[aps,prbbib,preprint,epsfig]{revtex}
2: \documentclass{jltp}
3:
4: \usepackage{graphicx}
5:
6: %%%%%%%%%%%%%%%%%% START OF PAPER %%%%%%%%%%%%%%%%%%%%%%
7:
8:
9: % \draft command makes pacs numbers print
10: %\draft
11:
12: \title{Analytical solution for nonlinear Schr$\ddot{\hbox{o}}$dinger
13: vortex reconnection}
14: \author{Sergey Nazarenko and Robert West}
15: \address{Mathematics Institute, University of Warwick, Coventry, CV4 7AL, UK}
16: %\date{\today}
17:
18: \runninghead{S.V. Nazarenko and R.J. West}{Analytical solution for
19: nonlinear Schr$\ddot{\hbox{o}}$dinger vortex reconnection}
20:
21: \begin{document}
22:
23: \maketitle
24:
25: %%%%%%%%%%%%%%%%%%%%% ABSTRACT %%%%%%%%%%%%%%%%%%%%%%%%%
26:
27: \begin{abstract}
28: Analysis of the nonlinear Schr$\ddot{\hbox{o}}$dinger vortex reconnection
29: is given in terms of coordinate-time power series. The lowest order terms
30: in these series correspond to a solution of the linear
31: Schr$\ddot{\hbox{o}}$dinger equation and provide several interesting properties
32: of the reconnection process, in particular the non-singular character of
33: reconnections, the anti-parallel configuration of vortex filaments and a
34: square-root law of approach just before/after reconnections. The complete
35: infinite power series represents a fully nonlinear analytic solution in a finite
36: volume which includes the reconnection point, and is valid for finite time provided
37: the initial condition is an analytic function. These series solutions are
38: free from the periodicity artifacts and discretization error of the direct
39: computational approaches and they are easy to analyze using a computer algebra
40: program.
41:
42: PACS number: 67.40.Vs
43: \end{abstract}
44:
45: % insert suggested PACS numbers in braces on next line
46: %\pacs{67.40.Vs}
47:
48:
49: %%%%%%%%%%%%%%%%%% MAIN DOCUMENT %%%%%%%%%%%%%%%%%%%%%
50: \section{INTRODUCTION}
51:
52: Vortex solutions of the nonlinear Schr$\ddot {\hbox{o}}$dinger
53: (NLS) equation are of interest in nonlinear optics
54: \cite{{Akhmediev},{Snyder},{Swartzlander}}
55: and in the theory of Bose-Einstein condensates \cite{{Berloff}} (BEC).
56: The NLS equation is also often used to describe turbulence in superfluid
57: helium. \cite{gp} NLS is a nice model in this case
58: because the vortex quantization appears naturally in this model
59: and because
60: its large-scale limit is the compressible Euler equation
61: describing classical inviscid fluids. \cite{{Schwarz},{Ercolani}}
62: At short scales, the NLS equation
63: allows for a ``quantum uncertainty principle'' which allows
64: vortex reconnections without the need for a finite
65: viscosity or other dissipation. Numerically, NLS vortex
66: reconnection was studied by Koplik and Levine\cite{Koplik} and, more recently,
67: by Leadbeater {\em et al.}\cite{Leadbeater} and, for a non-local version of NLS
68: equation, by Berloff {\em et al.}\cite{Berloff} In applications to superfluid
69: turbulence, the NLS equation was directly computed by Nore {\em et al.}\cite{Nore}
70: Such cryogenic turbulence consists of repeatedly reconnecting vortex
71: tangles, with each reconnection event resulting in the generation of
72: Kelvin waves on the vortex cores \cite{Svistunov} and a
73: sound emission. \cite{Leadbeater}
74: These two small-scale processes are very hard to correctly compute
75: in direct simulations of 3D NLS turbulence due to
76: numerical resolution problems. A popular way to avoid this
77: problem is to compute vortex tangles by a Biot-Savart method
78: (derived from the Euler equation) and use a simple rule to reconnect
79: vortex filaments that are closer than some critical (``quantum'') distance.
80: This approach was pioneered by Schwarz \cite{Schwarz} and it has been further developed by
81: Samuels {\em et al.}\cite{Barenghi} In this case, it is important
82: to prescribe realistic vortex reconnection rules.
83: Therefore, elementary vortex reconnection events have
84: to be carefully studied and parameterized. Numerically, such a
85: study was performed by Leadbeater {\em et al.}, \cite{Leadbeater} the present paper
86: is devoted to the analytical study of these NLS vortex reconnection events.
87:
88: The analytical approach of this paper is based on expanding
89: a solution in powers of small distance from the
90: reconnection point, and small time measured from the
91: reconnection moment.
92: %
93: The idea is to exploit the fact that when vortex filaments are near reconnection,
94: the nonlinearity in the NLS equation is small. This smallness of the nonlinearity
95: just stems from the definition of vortices in NLS (curves where $\Psi=0$) and the
96: continuity of $\Psi$. Their core size is of the order of the distance over
97: which $\Psi\rightarrow 1$ (where $\Psi=1$ represents the background condensate).
98: Therefore, for vortices near reconnection, separated by a distance much smaller than
99: their core size, $\Psi$ is small provided it is continuous.
100: %
101: %The idea is to exploit the fact that
102: %when vortex filaments are near reconnection, that is within
103: %a distance of each other that is less than their
104: %core size, the nonlinearity in the NLS equation is small.
105: %This smallness of the nonlinearity just stems from the definition of vortices in NLS
106: %(curves where $\Psi=0$) and the continuity of $\Psi$. Their core size is
107: %of the order of the distance over which $\Psi\rightarrow 1$ (where $\Psi=1$ represents
108: %the background condensate).
109: %Therefore, at distances much smaller than the core size,
110: %close to the reconnection point of the two vortices, $\Psi$ is small provided it is continuous.
111: %
112: Thus, to the first approximation the
113: solution near the reconnection point can be described
114: by a linear solution which, already at this level, contains
115: some very important information about the reconnection process:
116: (1) that the reconnection proceeds smoothly without any singularity
117: formation, (2) that in the immediate vicinity of the
118: reconnection the vortices are strictly anti-parallel and
119: (3) just before the reconnection event the distance between the vortices
120: decreases as $|t|^{1/2}$, where $t$ is the time measured from the
121: reconnection moment. Note that result (1) could surprise those who draw their
122: intuition from vortex collapsing events in the Euler equation (which are believed
123: to be singular). On the other hand, results (2) and (3) are remarkably
124: similar to the numerical and theoretical results found for the Euler equation. \cite{{Pumir1},{Pumir2},{Pumir3}}
125:
126: In section II of this paper we examine the local analysis of the
127: reconnection process by deriving a linear solution and in section III
128: consider its properties. The linear solution describes many, but not all the
129: important properties of vortex reconnection. In particular, it cannot describe
130: solutions outside the vortex cores and, therefore, it cannot
131: describe the far-field sound radiation produced by the reconnection. On the other hand,
132: one can substitute the linear solution back into the NLS equation and
133: find the first nonlinear correction to this solution. Recursively
134: repeating this procedure, one can recover the fully nonlinear
135: solution in terms of infinite coordinate and time series. This derivation is
136: discussed in detail in section IV. The series produced are a general solution to a
137: Cauchy initial value problem.
138: Thus, by Cauchy-Kowalevski theorem, \cite{Cauchy} these series define an
139: analytic function (with a finite convergence radius) provided
140: the initial conditions are analytic. The generation of such a suitable
141: initial condition is addressed in section V. Our series representation
142: of the solution to the NLS equation is exact, and therefore will
143: include such properties as sound emission. However,
144: due to the finite radius of convergence of the analytic solution,
145: one is unable to observe a far-field sound emission directly.
146: In this paper, we use {\em Mathematica} to compute some examples of the fully nonlinear
147: solutions for the vortex reconnection. The results of which are presented in
148: section VI.
149:
150: Let us summarise the advantages and disadvantages that our
151: analytical solution has with respect to those being
152: computed via direct numerical simulations (DNS).
153: Firstly, our analytical solutions are obtained as a
154: general formula, simultaneously applicable for a broad
155: class of initial vortex positions and orientations.
156: Secondly, our analytical solutions are not affected by
157: any periodicity artifacts (which are typical in DNS using
158: spectral methods) or by discretization errors.
159: On the other hand, our analytical solutions are only
160: available for a finite distance from the vortex lines
161: (of the order of the vortex core size) because their
162: defining power series have a finite radius of convergence.
163:
164:
165: \section{LOCAL ANALYSIS OF THE RECONNECTION}
166:
167: Let us start with the defocusing NLS equation written in the
168: non-dimensional form,
169: \begin{equation}\label{NLS}
170: i\Psi_t+\Delta \Psi +(1-|\Psi|^2) \Psi =0.
171: \end{equation}
172: Suppose that in vicinity of the point ${\bf r}=(x,y,z)=(0,0,0)$
173: at $t=t_0$ we have $\Psi = \Psi_0$ such that
174: %
175: $Re \Psi_0 = z$, and $Im \Psi_0 = az+bx^2-cy^2$,
176: %
177: %\begin{eqnarray}
178: %Re \Psi_0 &=& z,\\
179: %Im \Psi_0 &=& az+bx^2-cy^2,
180: %\end{eqnarray}
181: where $a,b$ and $c$ are some positive constants. For such initial conditions
182: the geometrical location of the vortex filaments, $\Psi=0$,
183: is given by two intersecting straight lines,
184: %
185: $z=0$ and $y=\pm\sqrt{b/c}\, x$.
186: %
187: %\begin{equation}
188: %z=0, \;\;\;\; y= \pm \sqrt{b/c} \, x.
189: %\end{equation}
190:
191: In the small vicinity of the point ${\bf r} =0$, deep inside
192: the vortex core (where $\Psi_0 \approx 0$), we can ignore
193: the nonlinear term found in equation (\ref{NLS}). Further, by a simple
194: transformation $\Psi = \Phi e^{it}$ we can eliminate the
195: third term $\Psi$ and obtain
196: %
197: $i\Phi_t+\Delta \Phi = 0$.
198: %
199: %\begin{equation}
200: %i\Phi_t+\Delta \Phi = 0.
201: %\end{equation}
202: %
203: (This just corresponds to multiplying our solution by a phase, it does not alter
204: its properties, but does make the following analysis simpler).
205: It is easy to see that the initial condition has not changed under this transformation,
206: $\Psi_{0} = \Phi_{0}$. Advancing our system a small distance in time $t-t_0$, we find
207: %
208: $Re\,\Phi = Re\,\Psi_0 - (t-t_0) \,\Delta Im\,\Psi_0$ and
209: $Im\,\Phi = Im\,\Psi_0 + (t-t_0) \,\Delta Re\,\Psi_0$,
210: %
211: %\begin{eqnarray}
212: %Re \Phi &=& Re \Psi_0-(t-t_0) \,\Delta Im \Psi_0, \nonumber \\
213: %Im \Phi &=&Im \Psi_0 +(t-t_0) \,\Delta Re \Psi_0, \nonumber
214: %\end{eqnarray}
215: %
216: or
217: %
218: \begin{eqnarray}\label{linear}
219: Re\,\Phi &=& z-2(b-c)\, (t-t_0),\\
220: Im\,\Phi &=& az+bx^2-cy^2. \nonumber
221: \end{eqnarray}
222: %
223: For both $t-t_0<0$ and $t-t_0>0$ the set of vortex lines, $\Phi=0$,
224: is given by two hyperbolas. A bifurcation happens at
225: $t=t_0$ where these hyperbolas degenerate into the two intersecting
226: lines (see Fig. \ref{contour}). This bifurcation corresponds to
227: the reconnection of the vortex filaments. Thus, we have constructed
228: a local (in space and time) NLS solution corresponding to
229: vortex reconnection. Obviously, this solution corresponds to a
230: smooth function $\Phi$ at the reconnection point. It should be stressed that
231: this is not an assumption, but just the way in which we have chosen to
232: construct our solution. However, we do believe that this observed
233: smoothness is a common feature of NLS vortex
234: reconnection events. If this is true then all such reconnecting
235: vortices could locally be described by the presented solution as
236: the intersection of a hyperbola with a moving plane provides a generic
237: local bifurcation describing a reconnection in the case of smooth fields.
238: %
239: % Figure 1.
240: %
241: \begin{figure}[!Ht]
242: \centerline{\includegraphics[width=4.5in]{fig1c.eps}}
243: \caption{Linear solution Eq. (\ref{linear}) of the nonlinear Schr$\ddot{\hbox{o}}$dinger
244: equation for $a=1$, $b=3$, $c=2$ and $t_0=0$. Sub-figures (a), (b) and (c) show
245: the intersection of the real (plane) and imaginary (hyperbolic paraboloid) parts
246: of Eq. (\ref{linear}) at successive times \mbox{$t=-0.1$}, \mbox{$t=0.0$} and
247: \mbox{$t=0.1$} respectively. Sub-figures (d), (e) and (f) show the corresponding
248: lines of intersection where \mbox{$\Psi=0$}; reconnection occurs at
249: \mbox{$t=t_0=0$}.}
250: \label{contour}
251: \end{figure}
252:
253: \section{PROPERTIES OF THE VORTEX RECONNECTION}
254:
255: The local linear solution we have constructed (\ref{linear}) reveals
256: several important properties of the reconnection of NLS vortices.
257: \newline
258: 1. Whatever the initial orientation of the vortex filaments,
259: the reconnecting parts of these filaments align so that they approach
260: each other in an anti-parallel configuration. Indeed, according to
261: (\ref{linear}), the fluid velocity field
262: %
263: $
264: \vec{v} = \nabla \arctan ( [Im\,\Phi]/[Re\,\Phi] )
265: = \nabla \arctan ( [az + bx^2 -cy^2]/[z-2(b-c)(t-t_0)] )
266: $.
267: %
268: At the mid-point between the two vortices one finds
269: a velocity field consistent with an anti-parallel pair,
270: %
271: $
272: \vec{v} = 1/[2a^2(c-b)(t-t_0)] \vec{e}_z \neq 0.
273: $
274: %
275: (For a parallel configuration one would find $\vec{v} = 0$).
276: %Taking derivatives, one obtains a velocity field consistent with that
277: %of an anti-parallel vortex pair.
278: Amazingly similar anti-parallel configurations
279: have been observed in the numerical Biot-Savart simulations of thin
280: vortex filaments
281: in inviscid incompressible fluids.\cite{{Pumir1},{Pumir3}}
282: %Later, we
283: %will return to
284: %this point to discuss the possible origins of this similarity.
285: \newline
286: 2.
287: %It follows from Eq. (\ref{linear}) that t
288: The reconnecting parts of the
289: vortex filaments approach each other as $\sqrt{t-t_0}$.
290: %This can clearly
291: %be seen by considering for example a configuration similar to that found in
292: %Fig. \ref{contour}$d$. We know the vortex lines correspond to the real and
293: %imaginary parts of $\Phi$ being identically zero.
294: Indeed, setting $Re\,\Phi = Im\,\Phi = 0$ and $y=0$ in
295: (\ref{linear}) one
296: %can determine $z$ from the real part and substitute into
297: %the imaginary part to
298: obtains
299: %
300: $
301: x = \pm \sqrt{2a[(c/b)-1](t-t_0)}.
302: $
303: %
304: Exactly the same scaling behaviour, for approaching thin filaments in
305: incompressible fluids, has been given by the theory of Siggia and
306: Pumir \cite{{Pumir1},{Pumir2},{Pumir3}} and has been observed numerically
307: in Biot-Savart computations. \cite{{Pumir1},{Pumir3}}
308: \newline
309: 3. The nonlinearity plays a minor role in the late
310: stages of vortex reconnection in NLS.
311: This is a simple manifestation of the fact that in the
312: close spatio-temporal
313: vicinity of the reconnection point $\Psi \approx 0$,
314: so that the dynamics are almost linear.
315: \newline
316: This last property can be also reformulated as follows. No singularity is observed in the
317: process of reconnection according to the solution (\ref{linear}):
318: both the real and imaginary parts of $\Psi$ behave continuously
319: in space and time. This property is in drastic contrast to the
320: singularity formation found in vortex collapsing events described by
321: the Euler equation. Indeed, distinct from incompressible fluids,
322: no viscous dissipation is needed for the NLS vortices to reconnect.
323: Here, dispersion does the same job of breaking the topological constraints (related to
324: Kelvin's circulation theorem) as viscosity does in a normal fluid.
325: %\newline
326:
327:
328: \section{NONLINEAR SOLUTION}
329: We will now move on to consider the full NLS equation. We will use a recursion
330: relationship to compute the solution assuming that \mbox{$x,y,z\sim \epsilon$}
331: and \mbox{$t\sim \epsilon^{3}$} (for simplicity, we take $t_0=0$).
332: The solution we obtain will therefore be of the form
333: \mbox{$\Psi=\Psi^{(0)}+\Psi^{(1)}+\Psi^{(2)}+\cdots$}, where
334: \mbox{$\Psi^{(n)}\sim\epsilon^{n}$}.
335: The above $\epsilon$ scaling of $x$, $y$, $z$ and $t$ has been chosen to generate
336: a recursion relationship when substituted in the NLS equation (\ref{NLS}). Of course
337: we could have chosen a different $\epsilon$ dependence, however, as the final series representation
338: of our solution contains an infinite number of terms, this would just correspond to
339: the same solution but with a suitable re-ordering.
340:
341: Consider the NLS equation (\ref{NLS}). Firstly, we note that
342: \mbox{$\partial_{t}\sim \epsilon^{-3}$} and
343: \mbox{$\triangle\sim\epsilon^{-2}$} and therefore
344: %
345: $i\Psi_{t}^{(m)} \sim\epsilon^{m-3}$,
346: $\triangle\Psi^{(n)} \sim\epsilon^{n-2}$,
347: $\left[ |\Psi|^{2}\Psi \right]^{(p)} \sim\epsilon^{p}$ and
348: $\Psi^{(q)} \sim\epsilon^{q}$,
349: %
350: %\begin{eqnarray*}
351: % i\Psi_{t}^{(m)} &&\sim\epsilon^{m-3},\\
352: % \triangle\Psi^{(n)} &&\sim\epsilon^{n-2},\\
353: % \left[ |\Psi|^{2}\Psi \right]^{(p)} &&\sim\epsilon^{p},\\
354: % \Psi^{(q)} &&\sim\epsilon^{q},
355: %\end{eqnarray*}
356: %
357: where
358: %
359: %\begin{equation}
360: $
361: [|\Psi|^{2}\Psi]^{(p)}
362: = \sum_{i,j=1}^{p}\Psi^{*(i)} \Psi^{(j)} \Psi^{(p-i-j)}
363: $.
364: %\nonumber.
365: %\end{equation}
366: %
367: Matching the terms, by setting $m=n+1$ and $p=q=n-2$, and integrating we find
368: %
369: \begin{equation}\label{t-rec}
370: \Psi^{(n+1)}= \Psi_{0}^{(n+1)}
371: + i\int_{0}^{t} \left[\triangle\Psi^{(n)}
372: +\Psi^{(n-2)}-[|\Psi|^{2}\Psi]^{(n-2)}\right] dt,
373: \end{equation}
374: %
375: where \mbox{$\Psi_{0}^{(n)}$} are arbitrary $n^{th}$ order functions of
376: coordinate which appear as constants of integration with respect to time.
377: The full nonlinear solution of the Cauchy initial value problem can now
378: be obtained by matching $\Psi_{0}^{(n)}$ to the $n^{th}$ order components
379: of the initial condition at $t=0$ obtained via a
380: Taylor expansion in coordinate. Let us assume that the initial condition
381: is an analytic function so that it can be represented by power series
382: in coordinates with a non-zero volume of convergence. Then, by the
383: Cauchy-Kowalevski theorem, the function $\Psi$ will remain analytic for
384: non-zero time. In other words, the solution can also be represented
385: as a power series with a non-zero domain of convergence in space and time.
386: Remarkably, the recursion relation Eq. (\ref{t-rec}) is precisely the
387: means by which one can write down the terms of the power-series
388: representation of the fully nonlinear solution to the NLS equation, with
389: an arbitrary analytical initial condition $\Psi_{0}$.
390:
391:
392: \section{INITIAL CONDITION}
393: Our next step is to construct a suitable initial condition for our study of
394: reconnecting vortices. This initial condition will have to be formulated in
395: terms
396: of a power series. We start by formulating the famous line vortex
397: solution to the steady state NLS equation \cite{gp} in terms of a power
398: series. Substituting \mbox{$\Psi=Ae^{i\theta}$} into Eq. (\ref{NLS}), we
399: find
400: %
401: $ \triangle A - A|\nabla\theta|^{2} + A - A^{2}=0\nonumber$,
402: %
403: %\begin{equation}
404: % \triangle A - A|\nabla\theta|^{2} + A - A^{2}=0\nonumber,
405: %\end{equation}
406: %
407: where we have used the fact that \mbox{$\nabla\theta\cdot\nabla A = 0$} and
408: \mbox{$\triangle \theta=0$}. We can simplify this equation, since $A=A(r)$
409: and therefore, \mbox{$\triangle A = \frac{1}{r}\partial_{r}(r\partial_r A)$}.
410: However, we also note that \mbox{$|\nabla^{2}\theta |=\frac{1}{r^{2}}$} since
411: \mbox{$\partial_x \theta =-y/r^{2}$} and \mbox{$\partial_y \theta =-x/r^{2}$}.
412: Therefore, we have
413: %
414: $\frac{1}{r}\partial_{r}(r\partial_r A)-\frac{A}{r^{2}}-A^{3}+A=0$.
415: %
416: %\begin{equation}
417: % \frac{1}{r}\partial_{r}(r\partial_r A)-\frac{A}{r^{2}}-A^{3}+A=0\nonumber.
418: %\end{equation}
419: %
420: We will solve this equation using another recursive
421: method. We would like to get a solution of the form
422: $A= a_{0} + a_{1}r + a_{2}r^{2} + a_{3}r^{3} + \cdots = \sum_{n} A^{(n)}$.
423: (However, we can set $a_{0}$ to zero on physical grounds, since we require
424: $\Psi=0$ at $r=0$). As before
425: %
426: $\frac{1}{r}\partial_{r}(rA_{r}^{(m)}) = m^{2}a_{m}r^{m-2}
427: \sim \epsilon^{m-2}$,
428: $\frac{A^{(n)}}{r^{2}} = a_{n}r^{n-2}\sim \epsilon^{n-2}$,
429: $[A^{3}]^{(p)} \sim \epsilon^{p}$ and
430: $A^{(q)} = a_{q}r^{q}\sim \epsilon^{q}$,
431: %
432: %\begin{eqnarray*}
433: % &&\frac{1}{r}\partial_{r}(rA_{r}^{(m)}) = m^{2}a_{m}r^{m-2}
434: % \sim \epsilon^{m-2},\\
435: % &&\frac{A^{(n)}}{r^{2}} = a_{n}r^{n-2}\sim \epsilon^{n-2},\\
436: % &&[A^{3}]^{(p)} \sim \epsilon^{p},\\
437: % &&A^{(q)} = a_{q}r^{q}\sim \epsilon^{q},
438: %\end{eqnarray*}
439: %
440: where
441: %
442: %\begin{equation}\label{hmm1}
443: $
444: [A^{3}]^{(p)}=\sum_{i,j=1}^{p} A^{(i)}A^{(j)}A^{(p-i-j)}.
445: $
446: %\nonumber
447: %\end{equation}
448: %
449: Again, by matching powers of $r$ we can derive a recursion
450: relationship for $a_{n}$. Setting $m=n$ and $p=q=n-2$ we obtain
451: %
452: \begin{equation}\label{hmm2}
453: a_{n} = (f_{n-2}-a_{n-2})/(n^{2}-1)
454: \nonumber,
455: \end{equation}
456: %
457: where \mbox{$f_{p}=[A^{3}]^{(p)}/r^{p}$}.
458:
459: We should note that $a_{2n}=0$ for all $n$. Therefore, taking a power of r
460: out of our expansion for $A(r)$ we find,
461: %
462: \begin{equation}\label{squidge}
463: \Psi = A(r)r^{i\theta} = rg(r)e^{i\theta}\nonumber,
464: \end{equation}
465: %
466: where \mbox{$g(r)=g(r^{2})= a_{1} + a_{3}r^{2}+ a_{5}r^{4}+ \cdots =
467: \sum_{n=1}^{\infty}a_{2n-1}r^{2n-2} $}. Further, \mbox{$re^{i\theta}$} is
468: complex so we can write \mbox{$re^{i\theta}=x+iy$} and hence our prototype
469: solution, for a vortex pointing along the $z$-axis, is
470: %
471: $\Psi = (x+iy)g(x^{2}+y^{2})$.
472: %
473: %\begin{equation}
474: % \Psi = (x+iy)g(x^{2}+y^{2}).
475: %\end{equation}
476: %
477:
478: We can manipulate this prototype solution to get an initial condition for
479: our vortex reconnection problem. Our initial condition $\Psi_0$ will be made
480: up of two vortices, $\Psi_1$ and $\Psi_2$, a distance $2d$ and angle
481: $2\alpha$ apart. Following the example of others, [Koplik {\em et al.}, Ref. \onlinecite{Koplik}] and
482: [Leadbeater {\em et al.}, Ref. \onlinecite{Leadbeater}],
483: we take the initial condition to be the product of $\Psi_1$ and $\Psi_2$,
484: that is
485: %
486: $\Psi_0=\Psi_1 \Psi_2\nonumber$.
487: %
488: %\begin{equation}
489: % \Psi_0=\Psi_1 \Psi_2\nonumber.
490: %\end{equation}
491: %
492: One could argue that such an initial condition is rather special, as
493: two vortices found in close proximity would typically have already
494: distorted one another
495: in their initial approach. Nevertheless, such a configuration provides us
496: with a valuable insight into the dynamics of NLS vortex reconnections.
497:
498: Firstly, we would like the vortices in the $(x,y)$ plane. We can do this by
499: transforming our coordinates $x\rightarrow y$, $y\rightarrow z$
500: and $z\rightarrow x$. This will give us a vortex pointing along the $x$-axis
501: %
502: $\Psi = (y+iz)g(y^{2}+z^{2})\nonumber$.
503: %
504: %\begin{equation}
505: % \Psi = (y+iz)g(y^{2}+z^{2})\nonumber.
506: %\end{equation}
507: %
508: The vortex can now be rotated by angle $\alpha$ to the $x$-axis in the
509: $(x,y)$ plane via
510: %
511: $x \rightarrow x\cos\alpha -y\sin\alpha$ and
512: $y \rightarrow y\cos\alpha +x\sin\alpha$.
513: %
514: %\begin{eqnarray*}
515: % x \rightarrow& x\cos\alpha -y\sin\alpha,\\
516: % y \rightarrow& y\cos\alpha +x\sin\alpha.
517: %\end{eqnarray*}
518: %
519: Finally, we shift the whole vortex in the $z$ direction by a distance $d$
520: using \mbox{$z\rightarrow z-d$} we finally obtain
521: %
522: $\Psi_{1} = [y\cos\alpha+x\sin\alpha+i(z-d)]
523: g((y\cos\alpha+x\sin\alpha)^{2}+(z-d)^{2})$.
524: %
525: %\begin{equation}
526: % \Psi_{1} = [y\cos\alpha+x\sin\alpha+i(z-d)]\;
527: % g((y\cos\alpha+x\sin\alpha)^{2}+(z-d)^{2})\nonumber.
528: %\end{equation}
529: %
530: In a similar manner, $\Psi_2$ is a vortex at angle $-\alpha$ and shifted
531: by $-d$ in the $z$ direction,
532: %
533: $\Psi_{2} = [y\cos\alpha-x\sin\alpha+i(z+d)]
534: g((y\cos\alpha-x\sin\alpha)^{2}+(z+d)^{2})$.
535: %
536: %\begin{equation}
537: % \Psi_{2} = [y\cos\alpha-x\sin\alpha+i(z+d)]\;
538: % g((y\cos\alpha-x\sin\alpha)^{2}+(z+d)^{2})\nonumber.
539: %\end{equation}
540: %
541:
542:
543:
544: \section{RESULTS}
545: % Figure 2.
546: %
547: \begin{figure}[!Ht]
548: \centerline{\includegraphics[height=2in]{fig2c.eps}}
549: \caption{The initial condition for the prototype vortex solution is constructed
550: via an appropriate expansion for \mbox{$A=A(r)$}, Eq. (\ref{hmm2}).
551: Here we can see the expansion for \mbox{$A=A(r)$} truncated at three
552: different orders of \mbox{$n$}; \mbox{$n=5$} (dash-dot line), \mbox{$n=15$} (dashed line) and
553: \mbox{$n=21$} (solid line) with \mbox{$a_1=0.6$}. At higher orders one would see the
554: existence of a finite radius of convergence at \mbox{$r\approx 2.5$}.}
555: \label{Ar}
556: \end{figure}
557: %
558: % Figure 3.
559: %
560: \begin{figure}[!Ht]
561: \centerline{\includegraphics[width=4.5in]{fig3c.eps}}
562: \caption{Sub-figures (a) to (f) show the evolution of two initially separated vortices
563: in time \mbox{$t$}. This realization is for \mbox{$|\Psi|=0.1$}, \mbox{$a_1=0.6$},
564: \mbox{$d=0.6$} and \mbox{$\alpha=\pi/4$}. The reconnection and separation events
565: are clearly evident.}
566: \label{pics}
567: \end{figure}
568: %
569: It would time consuming to expand the analytical solution, derived in the previous section,
570: by hand. Thankfully, we can use a computer to perform the necessary algebra, and to derive
571: the hugh number of terms the recursive formulae generate.
572: What follows is an example solution of the reconnection of two initially separated vortices.
573:
574: Firstly, we need to consider the validity and accuracy of
575: our initial condition. Fig. \ref{Ar}, shows the prototype solution
576: $A=A(r)$ for a single vortex, at various different orders.
577: Increasing the order will obviously improve accuracy. However, one should
578: note that at higher order there is evidence of a finite radius of convergence
579: $r_c$. This will restrict the spatial region of validity for our full
580: t-dependent solution. Our prototype solution also has a dependence
581: on $a_1$.
582: In the following simulation we have chosen $a_1=0.6$ numerically
583: so that the properties of $A(r)$ match that of a NLS vortex.
584: It is evident that we cannot satisfy these properties completely (namely
585: $\Psi\rightarrow 1$ as $r \rightarrow \infty$) as our power series
586: diverges near $r_c$. Nevertheless, this does not present us with
587: a problem if we restrict ourselves to considering the evolution of
588: contours of $\Psi$, such as $|\Psi|< 1$, where $A(r)$ is realistically represented.
589: Further, it should be noted that sound radiation could in principle be visualized in
590: our solution by drawing contours of $|\Psi|$ close to unity. However, to have
591: an accurate representation, we would need to take a very large number of terms
592: in the series expansion, therefore the study of sound in our model is somewhat
593: harder than the analysis of the vortices themselves.
594: Of course the validity of the full $t$-dependent solution will be restricted,
595: in the spatial sense, by the initial condition's region of convergence.
596: The region of convergence will evolve, remaining finite during a
597: finite interval of time (by the Cauchy-Kowalevski theorem), but then may shrink to
598: zero afterwards.
599:
600: We will now discuss an example solution.
601: %One should note, however,
602: %that even using a computer it can be time consuming constructing
603: %a general high order solution due to the huge number
604: %of terms the recursive formulae generate.
605: %By specifying some
606: %of the initial parameters before hand, e.g. the separation
607: %distance $d$, the solution can be generated much more quickly.
608: As we only wish to demonstrate this method, we will not consider
609: a high order solution in this paper. In our example
610: simulation below, we used Mathematica to perform the
611: necessary algebra in generating a nonlinear solution up to
612: \mbox{$O(\epsilon^{6})$}. (One should note that although the
613: prototype solution (\ref{squidge}) for a single vortex has $a_{2n}=0$ for all $n$,
614: our initial condition is made up of two vortices, i.e. two series multiplied together.
615: Therefore, there will be cross terms of order $O(\epsilon^{6})$ in our initial condition).
616:
617: %At this low order it is important that we choose a suitable
618: %contour of \mbox{$|\Psi|$}, so that our initial condition is
619: %reasonably accurate. If we take
620: %$|\Psi|=0.1$ we will have a good initial condition, for
621: %\mbox{$|r|\le 1$} (see Fig. \ref{Ar}).
622:
623: Our choice of parameters will be $d=0.6$ and $\alpha=\pi/4$.
624: This corresponds to two vortices, initially separated by a distance 1.2,
625: at right angles to each other. Fig. \ref{pics} shows the evolution of the iso-surface
626: $|\Psi|=0.1$ in time, demonstrating reconnection and then separation. Examining this
627: solution in detail we can clearly see evidence of some of the properties
628: mentioned earlier - that of a smooth reconnection (the absence of singularity)
629: and the anti-parallel alignment of vortices prior to reconnection.
630: %Firstly, we see there is a smooth reconnection of the
631: %two vortices without any singularity formation. Secondly, in the vicinity
632: %of the reconnection the vortices are indeed anti-parallel.
633:
634:
635: \section{CONCLUSION}
636:
637: In this paper we presented a local analysis of the NLS reconnection processes.
638: We showed that many interesting properties of the reconnection can
639: already be seen at the linear level of the solution. We derived a recursion
640: formula Eq. (\ref{t-rec}) that gives the fully nonlinear solution of the
641: initial value problem in a finite volume around the reconnection point for
642: a finite period of time. In fact, formula (\ref{t-rec}) can describe a
643: much wider class of problems. Of interest, for example,
644: are solutions describing the creation or annihilation of
645: NLS vortex rings. This process is easily described by considering vortex
646: rings, at there creation/annihilation moment, as the
647: the intersection of a plane with the minimum of a paraboloid.
648: Further, this method of expansion around a reconnection point can be used
649: for other evolution equations, e.g. the Ginzburg-Landau equation.
650: These applications will be considered in future.
651: We wish to thank Robert Indik, Nicholas Ercolani and Yuri Lvov for
652: their many fruitful discussions.
653:
654: %%%%%%%%%%%%%%%% ACKNOWLEDGEMENTS %%%%%%%%%%%%%%%%%%%%
655:
656: %\acknowledgments
657:
658: %\section*{ACKNOWLEDGMENTS}
659:
660:
661: %%%%%%%%%%%%%%%%%%% REFERENCES %%%%%%%%%%%%%%%%%%%%%%%
662:
663: %\begin{references}
664: \begin{thebibliography}{9}
665:
666: \bibitem{Akhmediev} N.N. Akhmediev,
667: Opt. Quan. Elec. \textbf{30}, 535 (1998).
668:
669: \bibitem{Snyder} A.W. Snyder, L. Poladian and D.J. Mitchell,
670: Opt. Lett. \textbf{17} (11) 789 (1992).
671:
672: \bibitem{Swartzlander} G.A. Swartzlander and C.T. Law,
673: Phys. Rev. Lett. \textbf{69} (17), 2503 (1992).
674:
675: \bibitem{Berloff} N.G. Berloff and P.H. Roberts,
676: J. Phys. A \textbf{32} (30), 5611 (1999).
677:
678: \bibitem{gp} V.L. Ginzburg and L.P. Pitaevskii,
679: Sov. Phys. JETP \textbf{7}, 858 (1958).
680:
681: \bibitem{Schwarz} K.W. Schwarz,
682: Phys. Rev. B. \textbf{38} (4), 2398 (1988).
683:
684: \bibitem{Ercolani} N. Ercolani and R. Montgomery,
685: Phys. Lett. A. \textbf{180}, 402 (1993).
686:
687: \bibitem{Koplik} J. Koplik and H. Levine,
688: Phys. Rev. Lett. \textbf{71} (9), 1375 (1993).
689:
690: \bibitem{Leadbeater} M. Leadbeater, T. Winiecki, D.C. Samuels,
691: C.F. Barenghi and C.S. Adams, Phys. Rev. Lett. \textbf{86} (8), 1410
692: (2001).
693:
694: \bibitem{Nore} C. Nore, M. Abid and M.E. Brachet
695: Phys. Rev. Lett \textbf{78} (20), 3896 (1997).
696:
697: \bibitem{Svistunov} B.V. Svistunov,
698: Phys. Rev. B \textbf{52} (5), 3647 (1995).
699:
700: \bibitem{Barenghi} C.F. Barenghi, D.C. Samuels, G.H. Bauer and R.J.
701: Donnelly, Phys. Fluids \textbf{9} (9), 2631 (1997).
702:
703: \bibitem{Pumir1} A. Pumir and E. Siggia, Phys. Fluids A \textbf{2} (2),
704: 220 (1990).
705:
706: \bibitem{Pumir2} A. Pumir and E. Siggia,
707: Physica D \textbf{37}, 539 (1989).
708:
709: \bibitem{Pumir3} A. Pumir and E. Siggia,
710: Phys. Rev. Lett. \textbf{55} (17), 1749 (1985).
711:
712: \bibitem{Cauchy} R. Courant and D. Hilbert,
713: \emph{Methods of mathematical physics: partial differential equations 2},
714: (Interscience, London, 1965).
715:
716: \end{thebibliography}
717: %\end{references}
718:
719:
720:
721:
722: \end{document}
723:
724: