1: \documentclass[pre,twocolumn,amsmath,amssymb]{revtex4}
2: \usepackage[T1]{fontenc}
3: \usepackage[latin1]{inputenc}
4: \usepackage{bm}
5: \usepackage[dvips]{graphicx}
6: \usepackage{mathrsfs}
7: \usepackage{epsfig}
8: \usepackage{psfrag}
9: %
10: \def\vec{\bm}
11: \let\SAVEDnabla=\nabla
12: \def\nabla{\bm \SAVEDnabla}
13: \def\i{i}
14: \def\e{e}
15: \def\d{\mathrm{d}}
16: \def\M{\mathscr M}
17: \def\U{\mathscr U}
18: \def\L{\mathscr L}
19: \newcommand{\ord}{{\cal O}}
20: \newcommand{\bea}{\begin{eqnarray}}
21: \newcommand{\ea}{\end{eqnarray}}
22: %
23: \begin{document}
24: %
25: %\title{Scattering of sound waves by a vortex and the Iordanskii force}
26: \title{Global singularity-free solution of the Iordanskii force problem}
27: \author{Markus Flaig and Uwe R. Fischer}
28: \affiliation{Eberhard-Karls-Universit\"at T\"ubingen, Institut f\"ur
29: Theoretische Physik\\
30: Auf der
31: Morgenstelle 14, D-72076 T\"ubingen, Germany}
32: \date{\today}
33: %
34: \begin{abstract}
35: We present a derivation of the transverse
36: force acting on a hydrodynamic vortex in
37: the presence of an incoming sound wave
38: from a global solution of the scattering problem,
39: using the method of matched asymptotic expansions.
40: %The transverse part of this
41: %force is intimately linked to the long disputed Iordanskii discussed
42: %in the context of superfluidity.
43: The solution presented includes a detailed treatment of the interaction
44: of the incident wave with the vortex core, and is free from
45: the singularities in the momentum exchange between vortex and sound wave
46: %scattering cross section
47: %for the partial wave shifts
48: %in the forward scattering direction,
49: %peculiarities
50: which have led to contradictory results for the value of
51: the transverse force in the literature.
52: \end{abstract}
53: \maketitle
54: %
55: \section{\label{S:intro}Introduction}
56: %
57: The transverse force on a quantized vortex
58: %in a scalar superfluid like HeII
59: due to phonon scattering, first found
60: %by Lifshitz and Pitaevskii~\cite{lif58} and
61: by Iordanskii~\cite{ior66}, has been controversially discussed in
62: particular since the work of Thouless, Ao and Niu, which predicted
63: that it in fact would vanish \cite{transverse,grisha3,son97, wex,
64: she98,grishaiord,sto00,son01,tho01,zhang}.
65: Based on general properties of
66: {\em superfluid} order and the Berry phase associated with the vortex motion,
67: they came to the conclusion that
68: the total transverse force
69: is proportional to the superfluid density alone
70: and given solely by the superfluid Magnus force, and not,
71: as the existence of the Iordanskii force would imply,
72: proportional to the {\em total density}, normal plus superfluid.
73: Recent calculations based on the vector potential
74: due to the Berry phase of quasiparticles lead to the same conclusion of
75: a vanishing Iordanskii force \cite{zhang}.
76:
77: It therefore seems in order to reinvestigate the assumptions
78: underlying calculations of the
79: Iordanskii force based on classical (but two-fluid) hydrodynamics
80: (see, e.g., \cite{son97,sto00}), where
81: %for low temperatures the computation of
82: the transverse force
83: %is reduced to the calculation of
84: stems from the momentum transfer between
85: asymmetrically scattered sound waves and the vortex
86: \cite{pit59,fet64}. %,cle68}.
87: The existing treatments %conventionally
88: use various simplifications
89: for a treatment of this problem, which constitute sources for
90: possible ambiguities and problems in the final result obtained
91: for the value of the transverse force. First of all, the authors
92: usually consider point vortices. However,
93: for small distances to the vortex axis, the corresponding
94: velocity field cannot be taken literally since then the
95: fluid velocity would diverge and the Mach number of
96: the problem would become infinite. It is not \textit{a priori}
97: clear that replacing the point vortex by a realistic vortex core would
98: not affect the oscillatory motion of the vortex core
99: due to the incident wave, and thus the momentum
100: transfer between vortex and sound wave.
101: Second, %as is done by many authors \cite{sto00},
102: using simplified scalar wave equations instead of the full
103: linearized hydrodynamic equations,
104: one does not take into account the effects due the interaction
105: of the incident wave with the vortex core leading to its
106: oscillatory motion accurately \cite{son97}.
107: Finally, working, as is conventional,
108: with an (unphysical) infinitely extended plane wave results in mathematical
109: difficulties (due to the non-convergence of sums over partial wave
110: shifts in the scattering cross section \cite{wex}).
111: We will show that no such singularities
112: appear in the momentum transfer between sound wave and vortex, provided
113: one is using a realistically shaped incident wave of a general form.
114: % which does not need to be that of a plane wave.
115:
116: In the following, we present a study of the scattering problem which avoids
117: using the simplifications discussed in the above.
118: In particular, we derive a {global} solution
119: for the coupled hydrodynamic equations
120: using the method of matched asymptotic
121: expansions, matching the solution for the vortical region, with a realistic
122: core of finite extension, smoothly to the far-field wave region
123: solution, assuming a general incoming wave instead of a plane wave.
124: As a result, we obtain a regular expression
125: for the transverse momentum exchange between sound field and vortex.
126: %, containing partial wave amplitudes of the incident wave up to second order.
127: %and a momentum exchange between vortex and sound field
128: When interpreted in terms of two-fluid hydrodynamics, this momentum exchange
129: exactly corresponds to the Iordanskii force.
130:
131: %
132: \section{\label{S:basic}Basic hydrodynamic equations}
133: %
134: Our starting point are the Euler and continuity
135: equations for an ideal fluid described in terms of density $\rho$,
136: velocity $\vec v$ and barotropic pressure functional $P = P (\rho)$:
137: %
138: \begin{subequations} \label{E:hydro}
139: \begin{align}
140: \label{E:continuity}
141: \frac{\partial \rho}{\partial t} +
142: \nabla \cdot ( \rho \vec v ) &=
143: 0, \\
144: \label{E:euler}
145: \frac{\partial \vec v}{\partial t} +
146: ( \vec v \cdot \nabla ) \vec v &=
147: -\frac{1}{\rho} \nabla P.
148: \end{align}
149: \end{subequations}
150: %
151: In order to consider small perturbations $(\rho_1, \vec v_1)$ on a given
152: background solution $(\rho_0, \vec v_0)$, which we require to satisfy
153: the hydrodynamic equations, we introduce the following
154: %ans\"atze for the
155: expansions of velocity and density:
156: %
157: \begin{equation} \label{E:expansion}
158: %\begin{split}
159: \vec v =
160: \vec v_0 +
161: \vec v_1 , \qquad
162: \rho =
163: \rho_0 +
164: \rho_1
165: %+\ord (\rho_2)
166: .
167: %\end{split}
168: \end{equation}
169: %where %the subscripts denote the order of the perturbation, i.e.,
170: %``0'' stands for the background solution, ``1'' for the %first order in the
171: %perturbation. % and so forth \cite{note}.
172: %
173: Inserting these ans\"atze into~\eqref{E:hydro},
174: and equating terms of first order
175: on both sides results in the following equations:
176: %
177: \begin{subequations} \label{E:lin-hydro}
178: \begin{gather}
179: \label{E:lin-continuity}
180: \frac{\d \rho_1}{\d t} + \nabla \cdot ( \rho_0 \vec v_1) +
181: \rho_1 \nabla \cdot \vec v_0 = 0, \\
182: \label{E:lin-euler}
183: \frac{\d \vec v_1}{\d t} +
184: (\vec v_1 \cdot \nabla) \vec v_0 =
185: - \nabla\left(\frac{c^2}{\rho_0} \rho_1\right);
186: \end{gather}
187: \end{subequations}
188: %
189: where the sound speed $c$ is as usual defined by
190: $c^2 = \partial P/\partial
191: \rho$ and
192: $\d / \d t = (\partial/\partial t + \vec v_0 \cdot
193: \nabla)$ denotes the convective derivative.
194: We now switch from $(\rho_1, \vec v_1)$ to another set of
195: variables $(\psi, \vec \xi)$ via the following
196: decomposition of the perturbations \cite{ber04}
197: %
198: \begin{equation} \label{E:switch}
199: \begin{split}
200: \rho_1 &=
201: -\frac{\rho_0}{c^2}
202: \frac{\d \psi}{\d t},\\
203: \vec v_1 &=
204: \nabla \psi + \vec \xi.
205: \end{split}
206: \end{equation}
207: %
208: The velocity perturbation is decomposed into a
209: part with zero vorticity, $\nabla\psi$, where the {\em pressure potential}
210: $\psi$ is regular, and a part comprising the vorticity in the sound
211: wave, $\vec\xi$.
212:
213: Our linearized hydrodynamic equations~\eqref{E:lin-hydro} may
214: %, after some algebra,
215: then be rewritten as
216: %
217: \begin{subequations} \label{E:berg}
218: \begin{gather}
219: \frac{\d}{\d t} \left(
220: \frac{1}{c^2}
221: \frac{\d \psi}{\d t} \right) =
222: \frac{1}{\rho_0}
223: \nabla \cdot [ \rho_0 ( \nabla \psi + \vec \xi )],
224: \label{E:berg-1} \\
225: \frac{\d \vec \xi}{\d t} +
226: ( \vec \xi \cdot \nabla ) \vec v_0 -
227: \nabla \psi \times \vec \omega_0 =
228: 0. \label{E:berg-2}
229: %\frac{c^2}{\rho_0} \nabla \left( \frac{\rho_0}{c^2} \right)
230: %\frac{\d \psi}{\d t}.
231: \end{gather}
232: \end{subequations}
233: %
234: The above coupled pair of equations has first been derived by Bergliaffa
235: et.\,al.\,\,using Clebsch potentials~\cite{ber04}. Eq.~\eqref{E:berg-1}
236: with the vortical part of the fluid velocity perturbations,
237: $\vec \xi$, set to zero, was derived in \cite{unr81,pie90}.
238: %and also App.~\ref{A:geometry} of thepresent paper).
239: %From Eq.~\eqref{E:berg-2} we may infer
240: It can be shown ~\cite{ber04}, that in
241: the case where the frequency $\varOmega$ of the sound wave is much
242: greater than the background vorticity $\vec \omega_0 = \nabla \times
243: \vec v_0$, there exist solutions where $\vec \xi$ is smaller by a
244: factor $\vec \omega_0/\varOmega$ than $\psi$.
245: %;
246: %$\vec \xi \approx -\nabla \psi \times \vec \omega_0/\varOmega$;
247: %which means that $\vec \xi$ is small compared to $\nabla \psi$
248: %and can consequently be neglected in~\eqref{E:berg-1}.
249: This leaves us with
250: only one equation for the determination of the scalar quantity $\psi$
251: %in the region
252: when the {\em quasiclassical} condition $\omega_0/\varOmega\ll 1$
253: is fulfilled; see the analysis of the {\em wave region}
254: in the section to follow.
255: %Although the simplifying {\em quasiclassical} condition of large
256: %frequency/momentum of the sound wave will not be assumed
257: %in the present paper, we will find it nevertheless advantageous
258: %to work with the equations~\eqref{E:berg} instead of~\eqref{E:lin-hydro}.
259:
260: We study an axisymmetric vortex, which may be
261: described by the following ansatz for the background solution,
262: employing a cylindrical system of coordinates $(r, \phi,
263: z)$ in what follows:
264: %
265: \begin{equation} \label{E:background}
266: \begin{split}
267: \rho_0 =
268: \rho_0(r), \qquad %\\
269: \vec v_0 = v_0(r) \, \vec e_\phi.
270: \end{split}
271: \end{equation}
272: %
273: The radial velocity $v_0$ is linked to the vorticity $\vec \omega_0 =
274: \omega_0 \, \vec e_z$ (which we consider to be a fixed quantity)
275: by the relation
276: %
277: \begin{equation} \label{E1:vorticity}
278: v_0(r) =
279: \frac 1r
280: \int_0^r
281: r' \omega_0(r') \, \d r'.
282: \end{equation}
283: %
284: We require the vorticity for large $r$ to decay faster
285: than any power of $r$. This will later on allow us to use an approach
286: based on the method of matched asymptotic expansions \cite{fed,fls99}.
287: For the velocity this assumption implies that, for $r \rightarrow
288: \infty$, it is given by $v_0 = \gamma/r$ (here and below we ignore
289: exponentially small terms), with
290: $\gamma = \int_0^\infty
291: r' \omega_0(r') \, \d r'$, which is the circulation divided by $2\pi$.
292: While the continuity
293: equation~\eqref{E:continuity} is automatically satisfied by the
294: ansatz~\eqref{E:background}, the density $\rho_0$ must be determined
295: such that the Euler equation~\eqref{E:euler} be fullfilled, which then
296: leads to the following equation:
297: %
298: \begin{equation} \label{density}
299: %\frac{d\rho_0}{dr} =
300: \rho_0'=
301: \frac{\rho_0}{c^2}
302: \frac{v_0^2}{r},
303: \end{equation}
304: %
305: where the prime denotes derivation with respect to $r$.
306: We immediately see that density variations are of second
307: order with respect to the Mach number $\M$ which is defined as $\M
308: = \U/c_\infty$, with $\U$ being some typical velocity in the vortex
309: %(for example, we may set $\U = \mathrm{max} \, |v_0|$)
310: and $c_\infty = c(r \rightarrow \infty)$ the sound speed for the fluid at
311: rest at infinity. We require the Mach number to be a small quantity ($\M \ll
312: 1$), which allows for an expansion of the solution
313: of the hydrodynamic equations in $\M$ in the section to follow.
314: %
315: \section{Matched asymptotic expansions}
316: %
317: To carry out the matching procedure, we
318: divide the plane into two regions: An inner,
319: \textit{vortical} region characterized by $r \simeq \L$, in which the
320: vorticity is concentrated, and an outer, \textit{wave} region,
321: where $r$ is of the order of the typical wavelength
322: $\lambda=2\pi \varOmega/c_\infty$ of the incoming wave,
323: which we take within the matched asymptotic expansion procedure
324: to be $\lambda \simeq \M^{-1} \L$. After having developed suitably
325: scaled non-dimensional equations for each region, the solutions in
326: both region are expanded in the Mach number and matched where the
327: regions overlap.
328:
329: In the vortical region we rescale the spatial variable $\vec r$ according
330: to $\vec r \rightarrow \L \vec r$ and the velocity $\vec v_0$ according
331: to $\vec v_0 \rightarrow \U \vec v_0$. The typical time scale of the
332: problem is given by the inverse of the frequency $\varOmega^{-1}
333: \simeq \L/\U$.
334: Consequently, the appropriate scaling for the time
335: variable $t$ is $t \rightarrow (\L/\U) t$, and the
336: frequency $\varOmega$ takes the scaled form $(\U/\L) \varOmega$. Thus we
337: have $|\vec \omega_0| / \varOmega = \ord(1)$, which means, in the light
338: of the remarks in the previous section, that we should replace
339: $\vec \xi$ by $\vec \xi/\L$. With these scalings, and using
340: Eq.~\eqref{density}, our Eq.~\eqref{E:berg-1} becomes:
341: %
342: \begin{equation} \label{non}
343: \nabla \cdot (\nabla \psi + \vec \xi ) =
344: \M^2 \left[ \frac{\d^2 \psi}{\d t^2} -
345: \frac{v_0^2}{r} (\partial_r \psi + \xi_r ) \right] +
346: \ord(\M^4),
347: \end{equation}
348: %
349: while Eq.~\eqref{E:berg-2} stays formally the same.
350:
351: In the wave region, the appropriately scaled spatial variable is given by $\vec
352: R = \M \vec r$. Furthermore, we define the velocity potential $\varPsi$
353: in the wave region as $\varPsi (\vec R) = \psi(\vec r)$. Since in the
354: wave region the velocity is one order in the Mach number smaller than
355: in the vortical region, we introduce the wave-region velocity $\vec V_0
356: = \M^{-1} \vec v_0 = \frac{\gamma}{R} \, \vec e_\phi$. Now, in the
357: wave region, Eq.~\eqref{non} takes on the following form:
358: %
359: \begin{equation} \label{NON}
360: (\nabla^2 - \partial_t^2) \varPsi =
361: 2 \M^2 ( \vec V_0 \cdot \nabla ) \partial_t \varPsi +
362: \ord(\M^4).
363: \end{equation}
364: %
365: Here and in what follows,
366: we use the convention that the nabla operator acting on a
367: wave-region quantity corresponds to a derivative with respect to
368: %the scaled position variable
369: $\vec R$.
370: The equation for $\vec \xi$ is not needed because we
371: restrict ourselves to solutions where in the wave region
372: the vortical part of the velocity perturbation, $\vec
373: \xi$, is exponentially small.
374:
375: In order to decouple the equations in the vortical region, we split
376: $\vec \xi$ into an irrotational and a divergence free part according
377: to \bea
378: \vec \xi = \nabla \eta + \vec e_z \times \nabla \zeta,
379: \ea
380: where $\eta$ and $\zeta$ are uniquely specified by demanding that $\eta,
381: \zeta \rightarrow 0$ as $r \rightarrow \infty$. Our Eq.~\eqref{non}
382: then becomes
383: %
384: \begin{multline} \label{Non}
385: \nabla^2
386: ( \psi + \eta ) =
387: \M^2
388: \left( - i \varOmega + \vec v_0 \cdot \nabla \right)^2
389: \psi \\
390: - \M^2 \,
391: \frac{v_0^2}{r}
392: [ \partial_r (\psi + \eta) - r^{-1} \partial_\phi \zeta ] +
393: \ord(\M^4).
394: \end{multline}
395: %
396: This Poisson equation may be solved with standard
397: methods. Using a relation following from
398: Eq.~\eqref{E:berg-2}
399: \bea
400: \frac{\partial \vec \xi}{\partial t}
401: -\vec v_0 \times (\nabla\times \vec \xi)
402: +\nabla (\vec v_0 \cdot \vec \xi)
403: -(\nabla \psi +\vec \xi)\times \vec \omega_0
404: = 0 ,\nonumber\\ \label{vecidentity}
405: \ea
406: and operating with the curl,
407: we arrive %, after quite some manipulation,
408: at the following equation:
409: %
410: \begin{multline} \label{vort}
411: \left(- i \varOmega +
412: \frac{v_0}{r}
413: \frac{\partial}{\partial \phi} \right)
414: \nabla^2 \zeta
415: -\frac{\omega_0'}{r}\frac{\partial \zeta}{\partial \phi}
416: \\
417: = -\omega_0' \frac{\partial}{\partial r} (\psi + \eta) -
418: \omega_0 \nabla^2 (\psi + \eta) .
419: \end{multline}
420: %
421: Note that here the fact that
422: $\nabla \cdot \vec v_0 = 0$ has been used. The above equation
423: allows us to calculate $\zeta$
424: after $\psi + \eta$ has been calculated using~\eqref{Non}.
425: %
426: %\section{\label{S:calc}Calculation of the solution}
427: %
428:
429: We now formally expand the general solution in the Mach number $\M$
430: and then solve the resulting equations to the relevant order in the
431: Mach number. In the vortical region the expansion takes on the following form
432: %
433: \begin{equation} \label{expansion}
434: \begin{split}
435: \psi &= \psi^{(0)} + \M \psi^{(1)} + \M^2 \psi^{(2)} + \ldots, \\
436: \eta &= \eta^{(0)} + \M \eta^{(1)} + \M^2 \eta^{(2)} + \ldots, \\
437: \zeta &= \zeta^{(0)} + \M \zeta^{(1)} + \M^2 \zeta^{(2)} + \ldots;
438: \end{split}
439: \end{equation}
440: %
441: while in the wave region we have
442: %
443: \begin{equation} \label{Expansion}
444: \varPsi = \varPsi^{(0)} + \M \varPsi^{(1)} + \M^2 \varPsi^{(2)} + \ldots.
445: \end{equation}
446: %
447: We restrict our analysis to order $\ord(\M^2)$, because at higher
448: orders the calculations become very involved.
449: %
450: \subsection{The solution to $\ord(\M^0)$}
451: %
452: At this order, the solution in the wave region is determined by the
453: ordinary wave equation $(\nabla^2 + k^2) \varPsi^{(0)} = 0$,
454: where the nondimensional wave number $k$ is defined as $k =
455: \varOmega$. We expand the general solution to zeroth order, corresponding
456: to our incoming wave, in the following way into partial waves of order
457: $\ell$:
458: %
459: \begin{equation}
460: \varPsi^{(0)} = \sum_{\ell = -\infty}^\infty A_\ell \,
461: J_{|\ell|}(kR) \, e^{i (\ell \phi - \varOmega t)}, \label{Psi(0)}
462: \end{equation}
463: %
464: where the $J_{|\ell|}$ are Bessel functions.
465: %The existence of terms
466: %irregular at the origin (e.g. Neumann functions) % $N_n$)
467: %is ruled out by the fact that to this order there are no terms
468: %in the vortical region for them to match onto, as we will show below.
469: %As
470: %already mentioned in the introduction, we will, however, consider
471: %the general case of an arbitrary wave as the zeroth-order solution in
472: %the wave region.
473: The form of $\varPsi^{(0)}$ in the vortical region may
474: be determined by expanding the Bessel functions for small $R$, yielding
475: %
476: \bea
477: \label{asym}
478: \varPsi^{(0)} & = &
479: A_0 e^{-i\varOmega t} + \M kr \sum_{\ell = \pm 1} \frac{A_\ell}{2}
480: e^{i (\ell \phi - \varOmega t)} \nonumber\\
481: & & +\M^2 \frac{k^2 r^2}{4} \left( \sum_{\ell = \pm 2}
482: \frac{A_\ell}{2} e^{i (\ell\phi - \varOmega t)} -
483: A_0 e^{-i\varOmega t}
484: \right) \nonumber\\
485: & & + \ord(\M^3).
486: \ea
487: %
488: %This asymptotic form can be used to carry out the matching with the
489: %vortical region.
490: This will be needed later to carry out the matching.
491:
492: In the vortical region, Eq.~\eqref{Non} to this order gives the
493: ordinary Poisson equation $\nabla^2 (\psi^{(0)} + \eta^{(0)}) = 0$,
494: with the following general solution
495: %
496: \begin{equation}
497: \psi^{(0)} + \eta^{(0)} =
498: \sum_{\ell = -\infty}^\infty
499: a_\ell \,
500: (kr)^{|\ell|} \,
501: e^{i (\ell \phi - \varOmega t)}.
502: \end{equation}
503: %
504: In the wave region, $r^{|\ell|} = \ord(\M^{-|\ell|})$, ruling out the
505: existence of terms with $|\ell| > 0$, since these would
506: be of \textit{negative} order in the Mach number. Thus, $\psi
507: ^{(0)} + \eta^{(0)} = a_0$.
508:
509: A decomposition of $\zeta^{(0)}$ into azimuthal modes according to
510: $\zeta^{(0)}=\sum_{\ell=-\infty}^\infty\zeta_\ell^{(0)}
511: (r)\,e^{i(\ell\phi-\varOmega t)}$ reduces
512: Eq.~\eqref{vort} to
513: %
514: \begin{equation} \label{rayleigh}
515: \left[
516: \frac{\d^2}{\d r^2} +
517: \frac 1r \frac{\d}{\d r} -
518: \frac{\ell^2}{r^2} +
519: \frac{\ell \omega_0'}{r \varOmega - \ell v_0}
520: \right] \zeta_\ell^{(0)} (r) =
521: 0.
522: \end{equation}
523: %
524: We restrict ourselves to the %non-resonant
525: case where there are no critical layers present,
526: $\varOmega \neq \ell v_0(r)/r$ for all $\ell, r$~\cite{fls99}.
527: %the frequency of the sound wave $\varOmega$ does not correspond to
528: %any of the eigenmodes of the vortex, that is
529: In this case, the solutions of this radial Rayleigh equation are
530: known to behave for large $r$ as $\zeta_\ell^{(0)} (r) \rightarrow \alpha_\ell\left(
531: r^{|\ell|} + \beta_\ell r^{-|\ell|} \right)$, immediately giving $\alpha_\ell =
532: 0$ for all $\ell$.
533: Furthermore, using identity (\ref{vecidentity}) and
534: $v_1^{(0)}=\nabla\psi^{(0)}+ \vec \xi^{(0)}=0$ as well as
535: $\nabla\psi^{(0)} =-\nabla \eta^{(0)}$,
536: %and $\vec \xi^{(0)} = \nabla \eta^{(0)}$,
537: we can derive the equation
538: $-i\Omega \eta^{(0)} +\frac{v_0}r \frac{\partial\eta^{(0)}}{\partial\phi }
539: =$ const., from which it follows
540: that $\eta^{(0)}=C e^{i(\Omega r/v_0) \phi}+$ const. Therefore,
541: as $\eta^{(0)}$ should be uniquely defined at each point in space,
542: $C$=0 in the presently considered %non-resonant
543: case; %of $\varOmega \neq \ell v_0(r)/r$;
544: $\eta^{(0)}$ thus equals a constant, which is zero because
545: it must vanish at large $r$.
546:
547: %There are thus no terms present in the vortical region
548: %which could match to Neumann functions in the wave region, verifying
549: %the statement made above after Eq. (\ref{Psi(0)}).
550: Matching the solution in the vortical region
551: with the wave region gives $a_0 = A_0$, thus
552: %completely specifying the solution as
553: %
554: \begin{equation} \label{E:zeroth}
555: \psi^{(0)} = A_0 \, e^{-i \varOmega t}, \quad
556: \eta^{(0)} = \zeta^{(0)} = 0.
557: \end{equation}
558: %
559: To this lowest (zeroth) order in the Mach number, there is thus no
560: perturbative flow in the vortical region,
561: %$\vec v_1^{(0)} = \nabla \psi^{(0)} + \vec \xi^{(0)} =0$;
562: but since $\psi^{(0)} \neq 0$, there are %pressure and
563: density perturbations present.
564: %
565: \subsection{The solution to $\ord(\M)$}
566: %
567: In the wave region, we have once more the ordinary wave equation
568: $(\nabla^2 + k^2) \varPsi^{(1)} = 0$. We require the solution
569: to this equation to be causal; therefore we write
570: %
571: \begin{equation} \label{hankel}
572: \varPsi^{(1)} =
573: \sum_{\ell = -\infty}^\infty %\left[
574: B_\ell H_\ell^{(1)} (kR)\,
575: %C_n H_n^{(2)} (kR) \right]
576: e^{i (\ell \phi - \varOmega t)},
577: \end{equation}
578: where the Hankel functions of the first kind $H_\ell^{(1)}$
579: correspond to outgoing cylindrical waves.
580: %(we omitted containing Hankel
581: %functions of the second kind $H_n^{(2)}$ correspond to ingoing ones.
582: %In order to obtain a causal solution (an outgoing wave),
583: %the terms containing Hankel functions of the second kind therefore have
584: %to be discarded: $C_n = 0$ for all $n$. Actually, $B_n = 0$ for all
585: %$n$ also, since d
586: Due to the irregular behavior of the Hankel functions at the
587: origin, none of the terms in the above sum
588: matches to the solution in the vortical region,
589: as becomes clear from Eq. (\ref{Asym}) below; therefore, $B_\ell = 0$.
590: Thus, $\varPsi^{(1)} = 0$ in the wave region,
591: reflecting the simple fact that there
592: is no sound radiation to linear order in
593: the Mach number into the wave region.
594:
595: In the vortical region, $\psi^{(1)} + \eta^{(1)}$ is once more
596: determined by the ordinary Poisson equation, giving $\psi^{(1)} +
597: \eta^{(1)} = \sum_{\ell = -1}^1 b_\ell \, (kr)^{|\ell|} \, e^{i (\ell \phi -
598: \varOmega t)}$, where the sum is limited by the fact that higher-order
599: terms would be of negative order in the Mach number.
600: The quantity $\zeta^{(1)}$ ist determined by the following equation:
601: %
602: \begin{multline} \label{E:guess}
603: \left[ \left(-i\Omega + \frac{v_0}{r} \frac{\partial}{\partial \phi}
604: \right)\nabla^2 -\frac{\omega_0'}r
605: \frac{\partial}{\partial \phi} \right] \zeta^{(1)}
606: \\
607: = -k\omega_0' \sum_{\ell = \pm 1} b_\ell e^{i (\ell \phi - \varOmega t)}.
608: \end{multline}
609: %
610: Using the ansatz $\zeta_p^{(1)} = f(r) \sum_{\ell = \pm 1} b_\ell
611: e^{i (\ell \phi - \varOmega t)}$ and $\nabla^2 v_0
612: = \partial_r (r^{-1} \partial_r (r v_0))= \omega_0'$,
613: we immediately see that the equation above is solved once one chooses
614: $f= -i v_0$. The asymptotics of the solution thus obtained
615: for $r \rightarrow \infty$ is given by
616: %
617: \begin{equation}
618: \zeta_\mathrm{p}^{(1)} \rightarrow
619: -i \frac{\gamma}{r}
620: \sum_{\ell = \pm 1} b_\ell e^{i (\ell \phi - \varOmega t)}.
621: \end{equation}
622: %
623: One gets the full solution to $\zeta^{(1)}$ by adding the general
624: solution $\zeta_\mathrm{h}^{(1)}$ of the homogeneous Rayleigh equation to
625: $\zeta_\mathrm{p}^{(1)}$. Since $\zeta_\mathrm{p}^{(1)} \rightarrow 0$ for $r
626: \rightarrow \infty$, $\zeta_\mathrm{h}^{(1)}\rightarrow 0$
627: for $r\rightarrow\infty$ as well.
628: %yielding $\zeta^{(1)} = \zeta_\mathrm{p}^{(1)}$. This can be shown by u
629: Using the homogeneous version of Eq. (\ref{E:guess}), which can be
630: transformed to a Rayleigh equation like Eq. (\ref{rayleigh});
631: the same argument as given after Eq. (\ref{rayleigh}) then yields
632: $\zeta_\mathrm{h}^{(1)}=0$. Since we have
633: %
634: \begin{multline}
635: \vec e_z \times \nabla \zeta^{(1)} \rightarrow
636: i \frac{\gamma}{r^2}
637: \sum_{\ell = \pm 1} b_\ell (\vec e_\phi + i \ell \vec e_r )
638: e^{i (\ell \phi - \varOmega t)} \\
639: = \nabla \left[\frac{\gamma}{r}
640: \sum_{\ell = \pm 1} \ell\, b_\ell\, e^{i (\ell \phi - \varOmega t)}
641: \right]
642: \end{multline}
643: %
644: for $r \rightarrow \infty$, the asymptotics of $\eta^{(1)}$ for $r
645: \rightarrow \infty$ can be determined to be $\eta^{(1)} \rightarrow
646: -\frac{\gamma}{r} \sum_{\ell = \pm 1} \ell\, b_\ell\,
647: e^{i (\ell \phi - \varOmega t)}$, thus giving
648: %
649: \begin{equation} \label{Asym}
650: \psi^{(1)} \rightarrow
651: b_0 +
652: \sum_{\ell = \pm 1} b_\ell \left[ kr + %(k r)^{|\ell|} +
653: \ell \frac{\gamma}{r} \right] e^{i (\ell \phi - \varOmega t)}
654: \end{equation}
655: %
656: for $r \rightarrow \infty$. By comparison with the solution in the
657: wave region, we obtain $b_0 = 0$ and $b_{\pm 1} = A_{\pm 1}/2$.
658: %Now the velocity perturbation $\vec v_1^{(1)}$ is already fully
659: %determined, and it is thus not necessary to solve for
660: %$\eta^{(1)}$.
661:
662: The rotational part of $\vec v_1^{(1)}$,
663: $\vec e_z\times \nabla \zeta^{(1)}$, describes the
664: oscillatory motion of the vortex core induced by the sound wave
665: \cite{leb00}, which may be seen as follows.
666: Near the origin, we may expand the incoming wave
667: into a plane-wave form
668: %, with propagation direction $\vec e_k$ ,
669: $\varPsi^{(0)} \approx \psi_0 e^{i(\vec k\cdot \vec R -\Omega t)}$,
670: for which the expansion coefficients are given by
671: %$A_\ell = \psi_0 e^{- i \ell \phi_k} i^{|\ell|}$.
672: $A_\ell = \psi_0 e^{- i \ell \phi_k} \,i^{|\ell|}$, with
673: %$\psi_0 =\psi_\mathrm{inc}(r = 0)$, and
674: $\phi_k$ the angle which the propagation
675: vector $\vec k = k ( \cos \phi_k,
676: \sin \phi_k, 0 )$ makes with the $x$ axis. We therefore have
677: $\vec e_z \times \nabla \zeta^{(1)}\approx \vec e_z \times
678: (\frac12 \psi_0 \omega_0 \vec e_k) e^{-i\Omega t}$. Then, using
679: $\vec v_0 \approx \frac12 (\vec \omega_0 \times \vec r) $ near the origin,
680: $\vec e_z \times (\frac12 \psi_0 \omega_0 \vec e_k) e^{-i\Omega t}
681: \approx - \psi_0 (e^{-i\Omega t} \vec e_k\cdot \nabla) \vec v_0 $.
682: We see that the term $\vec e_z \times \nabla \zeta^{(1)}$
683: causes a displacement $\vec r_0(t) \equiv
684: \psi_0 e^{-i\Omega t} \vec e_k$ of the vortex line.
685: %, since $\vec v_0 (\vec r) -(\vec r_0 \cdot \nabla) \vec v_0
686: %\simeq \vec v_0(\vec r - \vec r_0)$.
687: It follows for the resulting displacement
688: velocity $\frac\partial{\partial t} {\vec r}_0(t)
689: = -i \Omega \psi_0 e^{-i\Omega t}\vec e_k
690: = \nabla \varPsi^{(0)}|_{\vec r =0}$, so
691: that the vortex indeed moves
692: with the velocity in the incident sound wave to this order
693: in the Mach number.
694: % for a further discussion.)
695: %
696: \subsection{The solution to $\ord(\M^2)$}
697: %
698: In the vortical region,
699: $\nabla^2 \left( \psi^{(2)} + \eta^{(2)} \right) =
700: -k^2 A_0 e^{-i\Omega t}$,
701: %
702: with the following solution:
703: %
704: \begin{equation}
705: \psi^{(2)} + \eta^{(2)} =
706: -\frac{A_0 e^{-i\Omega t}}{4} k^2 r^2 +
707: \sum_{\ell = -2}^2
708: c_\ell \, (k r)^{|\ell|} \, e^{i (\ell \phi - \varOmega t)}.
709: \end{equation}
710: %
711: %Since terms that are in the wave region of negative order in the Mach
712: %number have to be exluded, $c_\ell = 0$ for $n > 2$.
713: Comparison
714: with~\eqref{asym} leads to $c_0 = c_{\pm 1} = 0$ and
715: $c_{\pm 2} = A_{\pm 2}/8$.
716: %
717: %This suffices to determine the solution in the wave region to second
718: %order; we need not calculate $\eta^{(2)}$ and $\zeta^{(2)}$.
719: In the wave region, $\varPsi^{(2)}$ obeys a forced wave equation:
720: %
721: \begin{equation} \label{E:forced}
722: ( \nabla^2 + k^2 ) \varPsi^{(2)} =
723: -2 i k \gamma R^{-2} \partial_\phi \varPsi^{(0)}.
724: \end{equation}
725: %
726: In the appendix, it is shown that the solution to this equation
727: may be written as a contour integral.
728: %, using the Schl\"afli
729: %representation for the Bessel functions.
730: The result is (with $\cal C$ denoting the %\textit{Schl\"afli contour}
731: integration contour shown in the appendix):
732: %
733: \begin{equation}
734: \begin{split} \label{solution}
735: \varPsi_\mathrm{p}^{(2)}
736: & =\mathrm{sgn}(\ell)\, i \frac{k \gamma}{2 \pi} A_\ell
737: \sum_{\ell = -\infty}^\infty
738: \varPhi_\ell(R) e^{i ( \ell \phi - \varOmega t )}
739: \end{split}
740: \end{equation}
741: where the expansion coefficients are given by
742: \begin{equation}
743: \begin{split}
744: \varPhi_\ell(R) =
745: \int_{\cal C}
746: \left( \ln t + i \frac{\pi}{2} \right)
747: \exp \left[i \frac{kR}{2}\left(t - \frac1t \right) \right]
748: \frac{\d t}{t^{|\ell|+1}}.
749: \end{split}
750: \end{equation}
751: %
752: We can write the general solution to
753: Eq.~(\ref{E:forced}) as $\varPsi^{(2)} = \varPsi_\mathrm{p}^{(2)}
754: + \varPsi_\mathrm{h}^{(2)}$, with $\varPsi_\mathrm{h}^{(2)}$
755: being the general solution of the homogeneous equation.
756: For large $R$, $\varPsi_\mathrm{p}^{(2)}$ corresponds to an outgoing wave
757: (cf. the appendix). Therefore, $\varPsi_\mathrm{h}^{(2)}$ should also
758: contain only the outgoing wave:
759: %
760: \begin{equation}
761: \varPsi_\mathrm{h}^{(2)} =
762: \sum_{\ell = -\infty}^\infty %\left[
763: C_\ell \, H_\ell^{(1)} (kR)
764: %+ E_n \, H_1^{(2)} (kR) \right]
765: e^{i (\ell \phi - \varOmega t)}.
766: \end{equation}
767: %
768: % This means that $E_\ell = 0$ for a causal solution.
769: %The remaining terms are (for $n \neq 0$) of the order
770: %$\ord(\M^{-n})$ in the vortical region; thus $D_n = 0$ for $n > 2$.
771: The remaining coefficients are determined by matching to the solution in
772: the vortical region, giving $C_\ell= 0$ for $\ell\neq \pm 1$ and
773: %
774: \begin{equation}
775: C_{\pm 1} =
776: \pm i \frac{\pi}{4}
777: k \gamma A_{\pm 1}.
778: \end{equation}
779: %
780: %This completes the determination of the solution to second order.
781: The solution to quadratic order in the Mach number,
782: $\varPsi^{(2)}$, is thus made up of two parts. The first
783: part, $\varPsi_\mathrm{p}^{(2)}$, may be interpreted to represent
784: the interaction
785: of the incident wave with the long-range velocity field of the vortex,
786: while the second, $\varPsi_\mathrm{h}^{(2)}$ stands for the interaction of the
787: wave with the vortex core \cite{fls99}.
788: %
789: \section{Force on the vortex from phonon scattering}
790: %
791: In this section we will return to the original (dimensional)
792: variables of section~\ref{S:basic}. Far from the vortex, our solution
793: %which we did derive in the last section,
794: takes on the following asymptotical form:
795: $\psi = \sum_\ell \psi_\ell (r) \, e^{i (\ell \phi - \varOmega t)}$ where,
796: using the asymptotic form of the
797: expansion coefficients $\varPhi_\ell(R)$ in
798: the far-field solution \eqref{solution},
799: %in Eq. (\ref{solution})
800: derived in the appendix [Eq. (\ref{E:asymptotics})],
801: we have
802: %
803: \begin{equation} \label{form}
804: \psi_\ell =
805: \frac{e^{-i \frac{\pi}{2} |\ell| -i \frac{\pi}{4}}}{\sqrt{2\pi kr}}
806: A_\ell \left\{
807: (1 + 2i\delta_\ell ) e^{i kr} +
808: e^{-i kr + i \pi |\ell| + i \frac{\pi}{2}}
809: \right\}.
810: \end{equation}
811: %
812: Here, the phase shifts $\delta_\ell$ are given by
813: %
814: \begin{equation} \label{shifts}
815: \delta_\ell =
816: \begin{cases}
817: \mp \frac{\pi}{4} \frac{k \gamma}{c_\infty} & ; \ell = \pm 1 , \\
818: -\mathrm{sgn}(\ell)
819: \frac{\pi}{2} \frac{k \gamma}{c_\infty} & ; \ell \neq \pm 1,
820: \end{cases}
821: \end{equation}
822: which are idential to those obtained by Sonin \cite{son97}.
823: %
824: %Far from the vortex
825: To evaluate the momentum transfer between sound wave and vortex,
826: we consider the time-averaged momentum-flux tensor, which reads
827: \cite{son97,sto00,stonePRE}:
828: %
829: %\begin{equation}
830: %\langle \varPi_{ij} \rangle =
831: %\left \langle
832: %\left( \frac{\partial P}{\partial \rho} \right)_{0}
833: %\rho_2 +
834: %\left( \frac{\partial^2 P}{\partial \rho^2} \right)_{0}
835: %\frac{\rho_1^2}{2}
836: %\right \rangle
837: %\delta_{ij}
838: %+ \rho_0
839: %\left \langle v_{1i} v_{1j} \right \rangle.
840: %\end{equation}
841: %
842: %The Gibbs-Duhem-Relation $\d P = \rho \, \d \mu$ tells us that
843: %$\frac{\partial P}{\partial \rho} = \rho \frac{\partial \mu}{\partial
844: % \rho} = c^2$ and $\frac{\partial^2 P}{\partial \rho^2} =
845: %\frac{\partial \mu}{\partial \rho} + \rho \frac{\partial^2
846: % \mu}{\partial \rho^2}$. Inserting the expansion~\eqref{E:expansion}
847: %into~\eqref{E:lin-euler} and equating second-order terms yields:
848: %
849: %\begin{equation}
850: %\frac{v_1^2}{2} =
851: %-\left[
852: %\left( \frac{\partial \mu}{\partial \rho} \right)_{0}
853: %\rho_2 +
854: %\left( \frac{\partial^2 \mu}{\partial \rho^2} \right)_{0}
855: %\frac{\rho_1^2}{2}
856: %\right].
857: %\end{equation}
858: %
859: %Now we are able to express the momentum-flux tensor in terms of known
860: %quantities:
861: %
862: \begin{equation} \label{E:flux}
863: \langle \varPi_{ij} \rangle =
864: \left( \frac{c_\infty^2}{\rho_\infty}
865: \frac{\langle \rho_1^2 \rangle}{2} -
866: \rho_\infty \frac{\langle v_1^2 \rangle}{2}
867: \right)
868: \delta_{ij} +
869: \rho_\infty \langle v_{1i} v_{1j} \rangle,
870: \end{equation}
871: where $\rho_\infty$ is the constant density far from the vortex.
872: %
873: We have to insert into this expression
874: the time-averaged expressions for the density
875: and the velocity expressed in terms of $\psi$ and $\vec \xi$ according
876: to~\eqref{E:switch}. Since we are looking at the region far from
877: the vortex, we may use the simplified expressions:
878: %
879: \begin{align}
880: \rho_1&=
881: -\frac{\rho_\infty}{c_\infty^2}
882: {\rm Re} \, [ \partial_t \psi ] =
883: \frac{i\rho_\infty k}{2 c_\infty} (\psi^* -\psi ), \\
884: v_{1i} &=
885: {\rm Re} \, [ \partial_i \psi ] =
886: \frac 12 ( \partial_i \psi^* + \partial_i \psi ) ,
887: \end{align}
888: %
889: which, when inserted into the time-averaged momentum-flux
890: tensor~\eqref{E:flux}, yield
891: %
892: \begin{align} \label{tensor}
893: \langle \varPi_{ij} \rangle &=
894: \frac{\rho_\infty}{4}
895: %\left(
896: [\partial_i \psi^* \partial_j \psi +
897: \partial_i \psi \partial_j \psi^*
898: %+2{\mathrm{Re}}[ \partial_i \psi^* \partial_j \psi]
899: +( k^2 |\psi|^2 -
900: |\nabla \psi|^2) \delta_{ij}
901: %\right)
902: ].
903: \end{align}
904: %
905: This comprises a general expression for the time-averaged
906: momentum-flux tensor in the far-region where $\vec v_0 \approx 0$.
907:
908: The projection $ F_{\phi_0}$ of the force on the vortex per
909: unit length $\vec{ F}$ in the direction of the unit vector
910: $\vec n_0 = (\cos \phi_0, \sin \phi_0, 0)$ is given by
911: %
912: \begin{equation}
913: F_{\phi_0} =
914: \langle \vec{ F} \cdot \vec n_0 \rangle =
915: -\oint \langle \varPi_{ij} \rangle n_i \, \d a_j.
916: \end{equation}
917: %
918: %We choose a cylindrical surface as our contour of integration, which
919: %means that $\vec a = \vec e_r$.
920: Inserting~\eqref{tensor} into the above expression,
921: and choosing a cylindrical surface around the vortex as
922: our contour of integration,
923: we get%
924: \begin{equation}
925: F_{\phi_0} =
926: -\frac{\rho_\infty}{4}
927: \int_{-\pi}^\pi \d \phi \, r
928: \cos ( \phi - \phi_0 ) \,
929: ( \partial_r \psi^\ast \partial_r \psi +
930: k^2 \psi^* \psi ).
931: \end{equation}
932: %
933: Making use of (\ref{form}), and performing the angular integration, we find
934: %
935: %
936: \begin{equation} \label{force}
937: F_{\phi_0} =
938: \rho_\infty k
939: %\sum_{n = -\infty}^\infty
940: \!\sum_{\ell=-\infty}^\infty \!
941: {\rm Re} \left[
942: i^{|\ell| + 1 - |\ell + 1|} e^{i \phi_0} A_\ell^* A_{\ell + 1}
943: ( \delta_\ell - \delta_{\ell + 1}) \right].
944: \end{equation}
945: %
946: Employing the obtained phase shifts for sound-vortex scattering,
947: Eq.~\eqref{shifts}, %in this expression,
948: we arrive at the following final expression for the force:
949: %
950: \begin{equation} \label{E:result}
951: F_{\phi_0} =
952: \frac{\rho_\infty k^2}{c_\infty} \frac \pi 4 \gamma
953: \, {\rm Re} \left[
954: e^{i \phi_0}
955: \sum_{\ell = 0}^1
956: (A_\ell^* A_{\ell+1} - A_{-\ell-1}^* A_{-\ell})
957: \right].
958: \end{equation}
959: The above equation represents the general result for the force on the vortex
960: from phonon scattering, containing partial waves up to order $|\ell|=2$,
961: valid for an arbitrary incident wave form.
962:
963: Note that there are no singularities in \eqref{E:result} present
964: due to partial wave summation, because we have
965: assumed an arbitrary incoming wave form right from the start.
966: Only now, {\em after}
967: obtaining the general result for the force in terms of the
968: expansion coefficients $A_\ell$, we use a plane wave
969: approximation for the incoming wave near the origin.
970: %, in the case that the incident wave %$\psi_\mathrm{inc}$
971: %varies only slowly over the scale of a few wavelengths near the origin.
972: %
973: Then, with $A_\ell = \psi_0 e^{- i \ell \phi_k} \,i^{|\ell|}$,
974: %
975: \begin{equation} \label{magnus}
976: F_{\phi_0} =
977: -\frac{\rho_\infty k^2}{c_\infty} \pi \gamma |\psi_0|^2
978: \sin (\phi_0 - \phi_k).
979: \end{equation}
980: %
981: %For an incident wave that does not vary smoothly near the origin (for
982: %example, a very narrow beam), the expression for the force will
983: %usually be different and then has to be calculated by the general
984: %expression~\eqref{E:result}.
985: Using the following expression for the momentum density
986: of a plane sound wave,
987: $\vec j = |\psi_0|^2 \frac{\rho_\infty k}{2 c_\infty} \vec k$
988: ~\cite{sto00}, Eq. \eqref{magnus} may easily be generalized
989: to yield the parallel and transverse components of the force,
990: %
991: \begin{equation} \label{E:transverse}
992: \vec{ F}_{||} = 0, \quad
993: \vec{ F}_\perp =
994: -\vec \varGamma \times \vec j,
995: \end{equation}
996: %
997: where $\vec \varGamma = 2\pi {\gamma}\vec e_z$ and $\vec j$
998: is the momentum density of the wave at the
999: location of the vortex. The coordinate invariant form of~\eqref{E:transverse}
1000: is obviously not restricted to the case where the vortex axis points into the
1001: $z$-direction. For a vortex in a superfluid we have to
1002: replace $\vec j$ by the thermally averaged momentum
1003: density of the phonons $\vec j_\mathrm{ph}$, which in terms of the
1004: quantities of two-fluid hydrodynamics reads $\vec
1005: j_\mathrm{ph} = \rho_n ( \vec v_n - \vec v_s )$, where
1006: $\vec v_n$ and $\vec v_s$ are normal and superflow
1007: velocities, respectively. For a vortex
1008: which is at rest with respect to the superfluid condensate ($\vec v_L =
1009: \vec v_s$) this gives
1010: %
1011: \begin{equation}
1012: \vec{F}_\mathrm{Iordanskii} =
1013: \rho_n [ \vec \varGamma \times ( \vec v_L -\vec v_n)],
1014: \end{equation}
1015: %
1016: which is the Iordanskii force.
1017: We remark that
1018: the arguments presented in \cite{sto00} suggest that
1019: the above form of the force remains invariant
1020: (to the considered order in the Mach number), also
1021: if the vortex moves with respect to the superflow.
1022: %
1023: \section{Conclusion}
1024: %
1025: We have solved the problem of scattering of sound waves by a vortex
1026: using an expansion into partial waves and the method of matched
1027: asymptotic expansions, up to the relevant quadratic order in the Mach number.
1028: %The leading order solution for each partial wave was expressed
1029: %analytically, from which in turn
1030: %the asymptotic behaviour at large distances from the vortex
1031: %could be calculated.
1032: To the best of our knowledge, the
1033: analytical form of the solution for the scattered wave
1034: in the form of a Schl\"afli contour integral, Eq. (\ref{solution}),
1035: has not been given yet in the literature.
1036: %Far from the vortex, the solution does only
1037: %depend on the total circulation and not on details of the
1038: %vortex core structure. %which is what one would expect.
1039: By evaluating the momentum-flux tensor far from the vortex, we derived
1040: a general expression for the force exerted by the sound wave on the vortex.
1041: The force depends only on the lowest partial waves ($|\ell| \le 2$)
1042: and is thus completely determined by the shape of the incident wave
1043: in a small region around the vortex axis with a diameter of the order of a few wavelengths
1044: (an analogous result was obtained by Shelankov for %the special case of
1045: an incident wave of beam-shaped form~\cite{she98}).
1046: We demonstrated that the usage of a
1047: ``physical'' incident wave of arbitrary shape (and finite extent)
1048: removes the mathematical difficulties which plagued previous work
1049: employing the partial wave method, where
1050: the total transverse cross section, obtained after performing a summation over
1051: all partial waves, contained singularities (was not convergent)
1052: \cite{son97,transverse,wex}.
1053: The resulting global, singularity-free solution of the scattering problem
1054: we presented
1055: then leads, in conjunction with conventional two-fluid hydrodynamics,
1056: to the Iordanskii force on the vortex, %which
1057: confirming its existence as a result of the asymmetrical scattering
1058: of sound waves at a vortex.
1059:
1060: Direct experimental confirmation of the Iordanskii
1061: force has so far been elusive.
1062: A conceivable experimental procedure to verify
1063: the existence of the Iordanskii force in a superfluid
1064: is to study the influence of sound pulses
1065: in Bose-Einstein condensates, created by Bragg scattering
1066: \cite{BECPhonons}, on the vortex.
1067: Using Bose-Einstein condensates has, inter alia,
1068: the advantage of being
1069: able to create sound wave pulses in an accurately controlled manner, with
1070: variable large momenta of the order of
1071: $c/\gamma$. We point out that our approach predicts the transverse force
1072: correctly also if the incoming wave is not a plane wave,
1073: which will in general be the case in the inhomogeneously
1074: trapped condensates. If the
1075: vortex is pinned by an optically created external potential to fix it in
1076: one position, and, shortly after the sound wave has
1077: been created, the pinning potential is turned off, % \cite{BECPinning},
1078: one should in principle
1079: be able to observe the transverse displacement of
1080: the vortex due to the sound wave scattered by it.
1081:
1082: \bigskip
1083:
1084: %
1085: \appendix
1086:
1087: \section{\label{A:forced}Forced wave equation}
1088:
1089: In this appendix, we derive the solution~\eqref{solution} to the
1090: forced wave equation~\eqref{E:forced}. Employing the ansatz \eqref{solution},
1091: %
1092: %\begin{equation}
1093: %\varPsi_\mathrm{p}^{(2)} =
1094: %\sum_{\ell = -\infty}^\infty
1095: %\varPhi_\ell (R) \,
1096: %e^{i (\ell \phi - \varOmega t)},
1097: %\end{equation}
1098: %
1099: Eq.~\eqref{E:forced} reduces to
1100: %
1101: \begin{equation} \label{E:Forced}
1102: \left[
1103: R^2 \frac{\d^2}{\d R^2} +
1104: R \frac{\d}{\d R} + k^2 R^2 - \ell^2
1105: \right]
1106: \varPhi_\ell =
1107: %2 \ell A_\ell k \gamma
1108: -4|\ell| \pi i\,
1109: J_{|\ell|} (kR).\vspace*{-2.5em}
1110: \end{equation}
1111: %
1112: \begin{center}
1113: \begin{figure}[b]
1114: \centerline{\epsfig{file=contour.eps,width=0.385\textwidth}}
1115: %\includegraphic\end{center}
1116: \caption{\label{F:contour}Schl\"afli contour of integration}
1117: \end{figure}
1118: \end{center}
1119: We write the solution to the above inhomogeneous Bessel
1120: equation %(\ref{E:Forced})
1121: as a contour integral.
1122: On the right-hand side, using the Schl\"afli integral representation for
1123: the Bessel functions \cite{Schlaefli}, with the contour $\cal C$ shown in
1124: Fig.~\ref{F:contour}, we express $J_{|\ell|}(kR)$ as
1125: %
1126: \begin{equation} \label{E:schlaefli}
1127: J_{|\ell|} (kR) =
1128: \frac{1}{2 \pi i}
1129: \int_{\cal C}
1130: \exp \left[\frac{kR}{2}\left(t - \frac1t \right) \right]
1131: \frac{\d t}{t^{|\ell| + 1}}.
1132: \end{equation}
1133: %
1134: On the left-hand side, inspired by the above Schl\"afli
1135: integral representation, we try the following ansatz:
1136: %
1137: \begin{equation} \label{E:ansatz}
1138: \varPhi_\ell =
1139: %\frac{1}{2 \pi i}
1140: \int_{\cal C}
1141: g_\ell(t) \,
1142: \exp \left[\frac{kR}{2}\left(t - \frac1t \right) \right]
1143: \frac{\d t}{t^{|\ell|+1}}.
1144: \end{equation}
1145: %
1146: %
1147: %
1148: After partially integrating the left-hand side and using the well known
1149: recursion relation $J_{\ell-1}(x) + J_{\ell+1}(x) = \frac{2\ell}{x}
1150: J_\ell(x)$ for
1151: the Bessel functions on the right-hand side, we find that the function
1152: $g_\ell(t)$ is determined by $\frac{dg_\ell}{dt}
1153: = %-\mathrm{sgn} (\ell) \, A_\ell k \gamma \,
1154: t^{-1}$. If we choose the integration constant such that
1155: %
1156: \begin{equation}
1157: g_\ell(t) =
1158: %-\mathrm{sgn} (\ell) \, A_\ell k \gamma \,
1159: \ln t + i \frac{\pi}{2} ,
1160: \end{equation}
1161: %
1162: then the $\varPhi_\ell$ correspond to causal solutions
1163: of~\eqref{E:Forced}. This may easily be verified by applying the
1164: method of steepest descent, yielding the following asymptotics:
1165: %
1166: \begin{equation} \label{E:asymptotics}
1167: \varPhi_\ell\rightarrow
1168: %-\mathrm{sgn}(\ell) \, i k \gamma A_\ell
1169: -2\pi \sqrt{\frac{\pi}{2k R}}
1170: \exp \left[ i \left(kR - |\ell| \frac \pi 2 - \frac \pi 4
1171: \right) \right]
1172: \end{equation}
1173: %
1174: for $R \rightarrow \infty$.
1175: Notice that $\varPsi_\mathrm{p}^{(2)}$ is
1176: regular at the origin, $\varPsi_\mathrm{p}^{(2)} \rightarrow 0$ as $R
1177: \rightarrow 0$, as can easily be confirmed by expanding the integrand for
1178: small $R$ and integrating.
1179: %Moreover, we can easily convince
1180: %ourselves that the $\varPhi_n$ are regular at the origin.
1181: %
1182:
1183:
1184: \begin{thebibliography}{999}
1185: %\bibitem{lif58}E.\,M. Lifshitz and L.\,P. Pitaevskii, Sov. Phys. JETP
1186: % \textbf 6, 418 (1958).
1187: \bibitem{ior66}S.\,V. Iordanskii, Sov. Phys. JETP \textbf{22}, 160 (1966).
1188: \bibitem{transverse} D.\,J. Thouless, P. Ao, and Q. Niu,
1189: %: {\em Transverse Force on a Quantized Vortex in a Superfluid},
1190: Phys. Rev. Lett. {\bf 76}, 3758 (1996).
1191: \bibitem{grisha3} G.\,E. Volovik,
1192: %: {\em Three nondissipative forces on
1193: %a moving vortex line in superfluids and superconductors},
1194: JETP Lett. {\bf 62}, 65 (1995);
1195: %\bibitem{grishacomment} G. E. Volovik: {\em Comment on
1196: %``Transverse Force on a Quantized Vortex in a Superfluid''},
1197: Phys. Rev. Lett. {\bf 77}, 4687 (1996).
1198: \bibitem{son97}E.\,B. Sonin, Phys. Rev. B \textbf{55}, 485 (1997);
1199: Phys. Rev. Lett. {\bf 81}, 4276 (1998).
1200: \bibitem{wex}C. Wexler, Phys. Rev. Lett. \textbf{79}, 1321 (1997);
1201: %\bibitem{wexlerthou}
1202: C. Wexler and D. J. Thouless,
1203: %: {\em Scattering of Phonons by a Vortex in a Superfluid},
1204: %cond-mat/9804118, submitted to
1205: Phys. Rev. B {\bf 58}, R8897 (1998).
1206: \bibitem{she98}A.\,L. Shelankov, Europhys. Lett. \textbf{43}, 623 (1998).
1207: \bibitem{grishaiord} G.\,E. Volovik,
1208: %: {\em Vortex versus spinning
1209: %string: Iordanski\v{\i} force and gravitational Aharonov-Bohm effect},
1210: %cond-mat/9804308,
1211: JETP Lett. {\bf 67}, 881 (1998).
1212: \bibitem{sto00}M. Stone, Phys. Rev. B \textbf{61}, 11780 (2000).
1213: \bibitem{tho01} D.\,J. Thouless, M.\,R. Geller, W.\,F. Vinen, J.-Y. Fortin,
1214: and S.\,W. Rhee, Phys. Rev. B \textbf{63}, 224504 (2001).
1215: \bibitem{son01}E.\,B. Sonin, {\sf preprint} cond-mat/0104221,
1216: published in
1217: {\em Vortices in Unconventional Superconductors and Superfluids},
1218: R.\,P. H\"ubener, N. Schopohl, and G.\,E. Volovik (Eds.), Springer,
1219: Berlin 2002, pp. 119.
1220: %\bibitem{Kopnin} For a general review of the vortex force problem in
1221: %superfluids and superconductors, see N.\,B. Kopnin, Rep. Prog. Phys.
1222: %{\bf 65}, 1633 (2002).
1223: \bibitem{zhang} C. Zhang, A.\,M. Dudarev, and Q. Niu,
1224: %{\sf preprint} cond-mat/0507125.
1225: Phys. Rev. Lett. {\bf 97}, 040401 (2006).
1226: \bibitem{pit59}L.\,P. Pitaevskii, Sov. Phys. JETP \textbf 8, 888 (1959).
1227: \bibitem{fet64}A.\,L. Fetter, Phys. Rev. \textbf{136}, A1488 (1964).
1228: %\bibitem{cle68}R.\,M. Cleary, Phys. Rev. \textbf{175}, 587 (1968).
1229: \bibitem{stonePRE} M. Stone, Phys. Rev. E {\bf 62}, 1341 (2000).
1230: \bibitem{ber04} S.\,E.\,P. Bergliaffa, K. Hibberd,
1231: M. Stone, and M. Visser, Physica D \textbf{191}, 121 (2004).
1232: \bibitem{unr81}W.\,G. Unruh, Phys. Rev. Lett. \textbf{46}, 1351 (1981).
1233: \bibitem{pie90}A.\,D. Pierce, J. Acoust. Soc. Am. \textbf{87}, 2292
1234: (1990).
1235: \bibitem{fed}M.\,V. Fedoryuk (Ed.), \textit{Partial Differential
1236: Equations V}, Springer, Berlin 1998.
1237: %\bibitem{arw95}G.\,B. Arfken und H.\,J. Weber, \textit{Mathematical
1238: % Methods for Physicists}. Academic Press, New York 1995.
1239: %\bibitem{aha59}Y. Aharonov und D. Bohm, Phys. Rev. \textbf{115}, 485
1240: % (1959).
1241: \bibitem{fls99}R. Ford und S.\,G. Llewellyn Smith,
1242: J. Fluid. Mech. \textbf{386}, 305 (1999).
1243: \bibitem{leb00}S. Leblanc, J. Fluid. Mech. \textbf{414}, 315 (2000).
1244: %\bibitem{fis02}U.R. Fischer und M. Visser, Phys. Rev. Lett
1245: % \textbf{88}, 110201 (2002).
1246: \bibitem{BECPhonons} D. M. Stamper-Kurn, A.\,P. Chikkatur, A. G\"orlitz,
1247: S. Inouye, S. Gupta, D.\,E. Pritchard, and W. Ketterle,
1248: %D. M. Stamper-Kurn {\it et al.},
1249: %: {\em Excitation of Phonons
1250: %in a Bose-Einstein Condensate by Light Scattering},
1251: Phys. Rev. Lett. {\bf 83}, 2876 (1999).
1252: %\bibitem{BECPinning} The pinning of a whole vortex lattice by an
1253: %optical lattice potential superimposed onto a rotating
1254: %Bose-Einstein condensate has been recently demonstrated by
1255: %S. Tung, V. Schweikhard, and E.\,A. Cornell, {\sf preprint}
1256: %cond-mat/0607697.
1257: \bibitem{Schlaefli} L. Schl\"afli, Math. Ann. {\bf 3}, 134 (1871).
1258: \end{thebibliography}
1259: \end{document}
1260:
1261:
1262: \section{\label{A:geometry}Some vortex geometry}
1263: %
1264: Using the continuity equation~\eqref{E:continuity}, Eq.~\eqref{E:berg-1}
1265: may be transformed to give Unruh equation~\cite{unr81}
1266: %
1267: \begin{equation} \label{E:unruh}
1268: \left( \frac{\partial}{\partial t} + \nabla \cdot \vec v_0 \right)
1269: \frac{\rho_0}{c^2}
1270: \left( \frac{\partial}{\partial t} + \vec v_0 \cdot \nabla \right)
1271: \psi =
1272: \nabla \cdot [\rho_0 (\nabla \psi + \vec \xi)].
1273: \end{equation}
1274: %
1275: %We now consider the limes of $\varOmega \rightarrow \infty$, because
1276: %only the sound waves with high frequency are capable of ``seeing'' the
1277: %details of the velocity variations in the vortex.
1278: We now consider sound waves with large frequency $\varOmega \gg |\vec
1279: v_0|$, for which we may set $\vec \xi = 0$. By introducing an
1280: effective metric via the definition
1281: %
1282: \begin{equation} \label{E:effective}
1283: \sqrt{-g} g^{\mu \nu} \equiv
1284: \frac{\rho_0}{c^2}
1285: \begin{pmatrix}
1286: 1 & \vec v_0^T \\
1287: \vec v_0 & \vec v_0 \vec v_0^T - c^2 \vec 1_3
1288: \end{pmatrix}.
1289: \end{equation}
1290: %
1291: our equation~\eqref{E:unruh} may be written in a covariant manner as
1292: $\partial_\mu ( \sqrt{-g} g^{\mu \nu} \partial_\nu \psi ) = 0$. Thus
1293: sound waves with large frequency appear to be moving in an effective
1294: (and generally curved) space time described by the
1295: metric~\eqref{E:effective}. This observation allows us to take over
1296: many useful concepts known in the realm of general relativity. For
1297: example the velocity variations in the vortex may be measured using
1298: the curvature scalar $R$ which quantifies the strength of
1299: curvature. Neglecting density variations, this quantity turns out
1300: to be given by
1301: %
1302: \begin{equation} \label{E:curvature}
1303: R =
1304: \frac{[v_0(r) - r v_0'(r)]^2}{2 \, \mbox{$c_\infty$}^2 r^2}.
1305: \end{equation}
1306: %
1307: If in this formula we set $r \simeq \L$ with $\L$ the vortex core
1308: size, then $R \simeq \M^2/\L^2$. Due to the smallness of the Mach
1309: number, the curvature of the effective space time is also small. In
1310: the eikonal approximation, it can be shown that the acoustic geometry
1311: of the long-range velocity field makes the vortex act as a weak
1312: gravitational lens~\cite{fis02}.
1313: %
1314:
1315: In the appendix it is shown that the velocity field of the vortex
1316: %probed by sound waves with large frequenzy $\varOmega \gg |\vec
1317: %\omega_0|$
1318: might be described in terms of an effective curved space time. The
1319: corresponding Riemann scalar $R$, which may be used to quantify the
1320: strength of local curvature, turns out to be of the order $R \simeq
1321: \M^2/\L^2$, with $\L$ the size of the vortex core (e.g. the region
1322: around the vortex axis where the vorticity is appreciably different
1323: from zero). In this paper we
1324: will take the wavelength $\lambda = 2 \pi \varOmega / c_\infty$ of the
1325: incident wave to be of the order of $\lambda \simeq 1/\sqrt R \simeq
1326: \M^{-1} \L$. Such sound waves will be insensitive to the details of
1327: the vortex core structure, but their wavelength is still small enough
1328: for them to effectively ``see'' the acoustic geometry generated by the
1329: long-range velocity field of the vortex. (It should
1330: be noted, however, that the small effects due to the finite curvature
1331: of the acoustic space time are unimportant for the determination
1332: of the transverse force).
1333:
1334: Technically, this appears two be an application of
1335: the method of matched asymptotic expansions mentioned above; but since
1336: there are no complications present in our problem, the matching can be
1337: carried out naively, without making recurse to the more sophisticated
1338: tools of the MAE method. (For more information, see~\cite{fed}.)
1339:
1340: \begin{equation}
1341: %\begin{split}
1342: \label{solution}
1343: \varPsi_\mathrm{p}^{(2)}
1344: = \sum_{n = -\infty}^\infty \Phi_n(R) \exp [\i ( n \phi - \varOmega t ) ]
1345: \end{equation}
1346: %& =
1347: %\exp [\i ( n \phi - \varOmega t ) ]\\ & \quad \times
1348: where the expansion coefficients are given by
1349: \begin{equation}
1350: \Phi_n(R) =
1351: \i \frac{k \gamma}{2 \pi} %\sum_{n = -\infty}^\infty
1352: \mathrm{sgn}(n) A_n \int_C
1353: \left( \ln t + \i \frac{\pi}{2} \right)
1354: \exp \left[\i \frac{kR}{2}\left(t - \frac1t \right) \right]
1355: \frac{\d t}{t^{|n|+1}}.
1356: \end{equation}
1357:
1358: \bibitem{note} We note that the velocity needs to be
1359: expanded up to first order in the fluctuations only,
1360: while the density perturbation should generally be expanded
1361: to second order; which is needed for a correct derivation of the
1362: energy-momentum tensor of the excitations %, Eq. (\ref{E:flux}) below
1363: \cite{son97,sto00,stonePRE}.