1: \documentclass[12pt]{iopart}
2: \input{amssym}
3: \usepackage{epsfig}
4: \global\arraycolsep=2pt
5:
6: \newcommand{\be}{\begin{equation}}
7: \newcommand{\bnabla}{\mbox{\boldmath{$\nabla$}}}
8: \newcommand{\ee}{\end{equation}}
9: \newcommand{\bea}{\begin{eqnarray}}
10: \newcommand{\eea}{\end{eqnarray}}
11: \newcommand{\ben}{\begin{equation*}}
12: \newcommand{\een}{\end{equation*}}
13: \newcommand{\bean}{\begin{eqnarray*}}
14: \newcommand{\eean}{\end{eqnarray*}}
15:
16:
17: \begin{document}
18:
19: \title{Exact zero-point interaction energy between cylinders}
20:
21:
22: \author{F.D. Mazzitelli$^{(a)}$, D.A.R. Dalvit$^{(b)}$, and F.C. Lombardo$^{(a)}$ }
23:
24:
25: \address{$^{(a)}$Departamento de F\'{\i}sica J.J. Giambiagi, Facultad de Ciencias Exactas y
26: Naturales, Universidad de Buenos Aires - Ciudad Universitaria, Pabell\'on I, 1428 Buenos Aires, Argentina}
27:
28: \address{$^{(b)}$Theoretical Division, MS B213, Los Alamos National Laboratory, Los Alamos, NM 87545, USA}
29:
30: \date{\today}
31:
32: \begin{abstract}
33: We calculate the exact Casimir interaction energy between two
34: perfectly conducting, very long, eccentric cylindrical shells
35: using a mode summation technique. Several limiting cases of the
36: exact formula for the Casimir energy corresponding to this
37: configuration are studied both analytically and numerically. These
38: include concentric cylinders, cylinder-plane, and eccentric
39: cylinders, for small and large separations between the surfaces.
40: For small separations we recover the proximity approximation,
41: while for large separations we find a weak logarithmic decay of the Casimir
42: interaction energy, typical of cylindrical geometries.
43: \end{abstract}
44:
45: \pacs{03.70+k, 12.20.-m, 04.80.Cc}
46:
47: \maketitle
48:
49: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
50: \section{Introduction}
51:
52: Almost 60 years ago \cite{Casimir1948}, Casimir discovered an
53: interesting macroscopic consequence of the zero point fluctuations
54: of the electromagnetic field: an attractive force between
55: uncharged parallel conducting plates. Since then, the dependence
56: of the Casimir force with the geometry of the conducting surfaces
57: has been the subject of several works \cite{reviews}. For many
58: years, the only practical way to compute the Casimir energy for
59: non planar configurations was the so called proximity force
60: approximation (PFA) \cite{Derjaguin1957}. This approximation is
61: valid for surfaces whose separation is much smaller than typical
62: local curvatures. Due to the high precision experiments performed
63: since 1997 \cite{exp}, there has been a renewed interest in the
64: geometry dependence of the Casimir force, and in particular in the
65: calculations of the corrections to the proximity approximation.
66:
67: In the last years there have been a number of attempts to compute
68: the Casimir forces beyond the PFA, using for example semiclassical
69: \cite{sem, Mazzitelli2003} and optical \cite{opt} approximations, and
70: numerical path-integral methods \cite{num}.
71: Large deviations from PFA for corrugated plates have been reported \cite{Genet2003},
72: and in recent months the Casimir
73: energy has been computed exactly for several configurations of
74: experimental interest, as the case of a sphere in front of a
75: plane, and a cylinder in front of a plane
76: \cite{Bulgac2006,Emig2006,Bordag2006,Gies2006}. As first suggested
77: in \cite{Dalvit2004}, the latter configuration is intermediate
78: between the sphere-plane and the plane-plane geometries, and may
79: shed light on the longstanding controversy about thermal
80: corrections to the Casimir force. There is an ongoing experiment
81: to measure precisely the Casimir force for this geometry
82: \cite{Brown-Hayes2005}.
83:
84: The configuration of two eccentric cylinders is of experimental relevance too
85: \cite{Dalvit2004,Mazzitelliqftext,Dalvit2006}. Although parallelism
86: is as difficult as for the plane-plane configuration, the fact
87: that the concentric configuration is an unstable equilibrium
88: position opens the possibility of measuring the derivative of the
89: force using null experiments. Up to now, the Casimir interaction
90: energy between two cylindrical shells has been computed
91: semiclassically and exactly in the concentric case
92: \cite{Mazzitelli2003,Saharian2006}, and using the proximity
93: approximation in the eccentric situation
94: \cite{Dalvit2004,Mazzitelliqftext}. In principle, one could
95: consider experimental configurations in which a very thin metallic
96: wire is placed inside a larger hollow cylinder. In this case, a
97: more accurate determination of the Casimir force is needed. The
98: aim of this paper is to describe in detail the derivation of the exact Casimir
99: interaction energy for eccentric cylinders, initially reported by us in \cite{Dalvit2006},
100: and to compute analytically different limiting cases of relevance for Casimir force
101: measurements in this configuration. To this end we will
102: use the mode summation technique combined with the argument
103: theorem in order to write the Casimir energy as a contour integral
104: in the complex plane \cite{Nesterenko1998}.
105:
106: The paper is organized as follows. In Section 2 we derive an
107: expression for the Casimir interaction energy for any
108: configuration invariant under translations in one of the spatial
109: dimensions. When properly subtracted, this expression reduces to an
110: integral over the imaginary axis, and is similar to expressions
111: for the Casimir energy derived using path integrals or scattering
112: methods. In Section 3 we derive the exact formula for the
113: interaction energy between eccentric cylinders and we
114: analyze some particular cases of the exact formula. We first show
115: the known results for the concentric case obtained from
116: the exact formulation, and that it is possible to derive the interaction energy for the
117: cylinder-plane configuration in the appropriate limit. In Section 4 we
118: consider the exact formula in the limit of quasi concentric cylinders
119: of arbitrary radii. We discuss two opposite limits of
120: this exact formula: large and small separations between the
121: metallic surfaces. In the first limit, we find that the Casimir
122: energy between a thin wire contained in a hollow cylinder has a
123: weak logarithmic decay as the ratio between the outer and inner radii becomes
124: very large. In the second limit, we recover previous results
125: obtained using PFA for quasi concentric cylinders.
126: Finally, Section 5 contains the conclusions of our work.
127:
128: %%%%%%%%%%%%%
129:
130: \section{Casimir energy as a contour integral}
131:
132: The Casimir energy for a system of conducting shells can be
133: written as
134: \begin{equation}
135: E_c= \frac{1}{2} \sum_p(w_p-\tilde w_p) ,
136: \label{ecasmodes}
137: \end{equation}
138: where $w_p$ are the eigenfrequencies of the electromagnetic field
139: satisfying perfect conductor boundary conditions on the surfaces
140: of the conductors, and $\tilde w_p$ are those corresponding to the
141: reference vacuum (conductors at infinite separation). Throughout this
142: paper we use units $\hbar=c=1$. The subindex
143: $p$ denotes the set of quantum numbers associated to each
144: eigenfrequency. Introducing a cutoff for high frequency modes
145: $E_{c}(\sigma)={1\over 2}\sum_p(e^{-\sigma w_p} w_p-e^{-\sigma
146: \tilde w_p} \tilde w_p)$,
147: the Casimir energy $E_{c}$ is the limit of $E_{c}(\sigma)$ as
148: $\sigma\rightarrow 0$. For simplicity we choose an exponential
149: cutoff, although the explicit form is not relevant.
150:
151: Let us consider a general geometry with translational invariance
152: along the $z-$axis (as for example very long and parallel waveguides
153: of arbitrary sections). The transverse electric (TE) and
154: transverse magnetic (TM) modes can be described in terms of two
155: scalar fields with adequate boundary conditions. In cylindrical
156: coordinates, the modes of each scalar field will be of the form
157: $h_{n, k_z}(t,r,\theta,z)=e^{(-iw_{n, k_z}t+ik_z
158: z)}R_n(r,\theta)$, where the eigenfrequencies are $w_{n,
159: k_z}=\sqrt{k_z^2+\lambda^2_n}$, and $\lambda_n$ are the
160: eigenvalues of the two dimensional Laplacian
161: \begin{equation}
162: \left(\frac{\partial^2}{\partial
163: r^2}+\frac{1}{r}\frac{\partial}{\partial r}+
164: \frac{1}{r^2}\frac{\partial^2}{\partial\theta^2}+\lambda_n^2\right
165: ) R_n(r,\theta)=0.
166: \end{equation}
167: The set of quantum numbers $p$ is given by
168: $(n, k_z)$. For very long cylinders of length $L$ we can replace
169: the sum over $k_z$ by an integral. The result is
170: \begin{equation}
171: E_{c} (\sigma) = {L \over 2}\int_{-\infty}^{\infty}{dk_z\over
172: 2\pi} \sum_{n}\left (\sqrt{k_z^2+\lambda_{n}^2}e^{-\sigma
173: \sqrt{k_z^2+\lambda_{n}^2}} - \sqrt{k_z^2+\tilde \lambda_{n}^2}
174: e^{-\sigma \sqrt{k_z^2+\tilde\lambda_{n}^2}} \right) \; .
175: \label{exs}
176: \end{equation}
177: From the argument theorem it follows that
178: \begin{equation}
179: {1\over 2\pi i} \int_{C} \,d\lambda \; \lambda \; e^{-\sigma \lambda} {d\over d\lambda}
180: \ln f(\lambda)=\sum_i \lambda_i \; e^{-\sigma \lambda_i} \; ,
181: \end{equation}
182: where $f(\lambda)$ is an analytic function in the complex $\lambda$ plane within the closed contour
183: ${C}$, with simple zeros at $\lambda_1, \lambda_2, \dots$ within ${C}$.
184: We use this result to replace the sum over $n$ in Eq.(\ref{exs}) by a contour integral
185: \begin{equation}
186: E_c (\sigma)={L\over 4\pi i}\int_{-\infty}^{\infty} {dk_z\over 2\pi}
187: \int_{C} d\lambda \sqrt{k_z^2+\lambda^2}
188: e^{-\sigma \sqrt{k_z^2+ \lambda^2}}{d\over d\lambda}
189: \ln \left( \frac{F}{F_{\infty}} \right) ,
190: \end{equation}
191: where $F$ is a function that vanishes at $\lambda_n$ for all $n$
192: (and $F_{\infty}$ vanishes at $\tilde\lambda_n$).
193:
194: %%%%%%%%%%%%
195: %Figure 1
196: %%%%%%%%%%%%
197: \begin{figure}
198: \centerline{\psfig{figure=njp.fig1.eps,height=5cm,width=9cm,angle=0}}
199: \caption{Geometrical configuration studied in this paper. Two perfectly conducting eccentric
200: cylinders of radii $a<b$, length $L$, and eccentricity $\epsilon$ interact via the Casimir force.
201: The figure on the right shows the polar coordinates $(r,\theta)$ and $(\rho,\varphi)$ of any point $P$
202: between the eccentric cylinders used for the determination of the classical eigenvalues for this
203: configuration.}
204: \label{fig1}
205: \end{figure}
206:
207:
208: In the rest of this section we will consider the particular
209: configuration of two eccentric cylinders with circular sections of
210: radii $a$ and $b$, respectively. We will denote the eccentricity
211: of the configuration by $\epsilon$ (see Fig. 1).
212: The geometrical dimensionless parameters $\alpha \equiv b/a$
213: and $\delta \equiv \epsilon/a$ fully characterize the eccentric cylinder configuration.
214: It is worth emphasizing that the results of this section can be
215: trivially extended to more general configurations, as long as they
216: are translationally invariant along one spatial dimension. It is
217: convenient to compute the difference between the energy of the
218: system of two eccentric cylinders and the energy of two isolated
219: cylindrical shells of radii $a$ and $b$,
220: \begin{equation}
221: E_{12}(\sigma)=E_c(\sigma) - E_1(\sigma,a)-E_1( \sigma,b) \; ,
222: \end{equation}
223: where
224: \begin{equation}
225: E_1(\sigma,a) ={L\over 4\pi i} \int_{-\infty}^{\infty} {dk_z\over 2\pi}
226: \int_{C} d\lambda \sqrt{k_z^2+ \lambda^2}
227: e^{-\sigma \sqrt{k_z^2+\lambda^2}}{d\over d\lambda}
228: \ln \left( \frac {F_{1{\rm cyl}}(a)}{F_{1{\rm cyl}}(\infty)} \right) \; .
229: \end{equation}
230: Here $F_{1{\rm cyl}}(a)$ is a function that vanishes at the
231: eigenfrequencies of an isolated cylindrical shell of radius $a$.
232: Therefore
233: \begin{equation}
234: E_{12}(\sigma) = {L\over 4\pi i}\int_{-\infty}^{\infty} {dk_z\over 2\pi}
235: \int_{C} d\lambda \sqrt{k_z^2+\lambda^2}
236: e^{-\sigma \sqrt{k_z^2+\lambda^2}}{d\over d\lambda} \ln M(\lambda) \; ,
237: \end{equation}
238: where
239: \begin{equation}
240: M= \frac{F}{F_{\infty}} \frac{F_{1{\rm cyl}}(\infty)^2} {F_{1{\rm cyl}}(a)F_{1{\rm cyl}}(b)} \; .
241: \label{Mformal}
242: \end{equation}
243:
244: %%%%%%%%%%%%
245: %Figure 2
246: %%%%%%%%%%%%
247: \begin{figure}
248: \centerline{\psfig{figure=njp.fig2.eps,height=6cm,width=5cm,angle=0}}
249: \caption{Contour for the integration in the complex plane.}
250: \label{fig2}
251: \end{figure}
252:
253:
254:
255: To proceed we must choose a contour for the integration in the
256: complex plane. In order to compute $E_{c} (\sigma),
257: E_1(\sigma,a)$, and $E_1(\sigma,b)$ separately, an adequate
258: contour is a circular segment ${C}_{\Gamma}$ and two straight line
259: segments forming an angle $\phi$ and $\pi -\phi$ with respect to
260: the imaginary axis (see Fig. 2). The nonzero angle $\phi$ is
261: needed to show that the contribution of ${C}_{\Gamma}$ vanishes in
262: the limit $\Gamma\rightarrow \infty$ when $\sigma > 0$. For the
263: rest of the contour, the divergences in $E_{c}(\sigma)$ are
264: cancelled out by those of $E_1(\sigma,a)$ and $E_1(\sigma,b)$, as
265: in the case of concentric cylinders \cite{Mazzitelli2003}.
266: Therefore, in order to compute $E_{12}(\sigma)$ we can set
267: $\phi=0$ and $\sigma =0$, and the contour integral reduces to an
268: integral on the imaginary axis. We find
269: \begin{equation}
270: E_{12}= -{L \over 2\pi}\int_{-\infty}^{\infty} {dk_z\over 2\pi} ~{\rm Im}
271: \left\{\int_0^{\infty} dy \sqrt{k_z^2-y^2} {d\over dy}\ln M(iy)
272: \right\} \; . \label{xx}
273: \end{equation}
274: As we will see, $M(iy)$ is a real function - hence, the integral
275: over $y$ in Eq. (\ref{xx}) is restricted to $y > k_z$. After some straightforward
276: steps one can re-write this equation as
277: \begin{equation}
278: E_{12}={L\over 4\pi} \int_{0}^{\infty} dy \ y\ln M(iy) \; .
279: \label{xxx}
280: \end{equation}
281: As we have already mentioned, a similar expression can be derived
282: for conductors of arbitrary shape, as long as there is
283: translational invariance along the $z$-axis. It is worth noting
284: that the structure of this expression is similar to the ones
285: derived recently for the cylinder-plane geometry using path
286: integrals \cite{Emig2006,Bordag2006}, and for the sphere-plane
287: geometry using the Krein formula \cite{Bulgac2006}.
288:
289: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
290:
291: \section{The exact formula}
292:
293: In this section we derive the exact formula for the Casimir energy
294: between eccentric cylinders. We proceed in two steps: we first
295: find the function $F$ with zeroes at the eigenfrequencies for the
296: geometric configuration. Then we obtain an explicit expression for
297: the function $M$, which involves a definition of the Casimir energy as a difference
298: between the energy of the actual configuration and a configuration
299: with very large and separated conductors.
300:
301: %%%%%
302:
303: \subsection{The classical eigenvalues}
304:
305: The solution of the Helmholtz equation in the annular region
306: between eccentric cylinders has been considered in the framework
307: of classical electrodynamics, fluid dynamics, and reactor physics,
308: among others \cite{Singh1984,Balseiro1950}. The eigenvalues have
309: been computed using different approaches, as for instance
310: conformal transformations that map the eccentric annulus onto a
311: concentric one. As the two dimensional Helmholtz equation is not
312: conformally invariant, the transformed equation has coordinate
313: dependent coefficients and has to be solved numerically
314: \cite{Hine1971}. As is well known, it is very difficult to compute
315: the Casimir energy from the numerical eigenfrequencies. It is more
316: efficient to use the procedure outlined in Section 2, that only
317: needs a function $F$ with zeroes at the eigenvalues. Although for
318: the eccentric annulus this function has been previously found in
319: the literature \cite{Singh1984,Balseiro1950}, for the benefit of
320: the reader we include here a derivation of this result.
321:
322: The electromagnetic field inside an eccentric waveguide can be
323: described in terms of TM and TE modes. The TM modes are
324: characterized by a vanishing $z$ component of the magnetic field,
325: $B_z=0$. The other components of the electromagnetic field can be
326: derived from the $z$ component of the electric field,
327: $E_z(r,\theta,z,t)=E(r,\theta) e^{-i \omega t + i z k_z}$, with
328: \begin{equation}
329: E(r,\theta)=\sum_m\left[A_m J_m(\lambda r)+B_m N_m(\lambda
330: r)\right]e^{im\theta} ,
331: \label{E}
332: \end{equation}
333: where $w^2=k_z^2+\lambda^2$, and $(r,\theta)$ are polar
334: coordinates with origin at the center of the outer cylinder (see Fig. 1).
335: One can also describe the $z$ component of the electric field
336: using polar coordinates $(\rho , \varphi) $
337: with origin at the center of the inner cylinder (see Fig. 1),
338: \begin{equation}
339: \bar E(\rho,\varphi)=\sum_n\left[\bar A_n J_n(\lambda \rho)+\bar
340: B_n N_n(\lambda \rho)\right]e^{in\varphi}\, . \label{barE}
341: \end{equation}
342: The perfect conductor boundary conditions imply that the
343: $z$ component of the electric field must vanish on the cylindrical
344: shells
345: \begin{eqnarray}
346: A_m J_m(\lambda b)+B_m N_m(\lambda b)&=& 0 ,\nonumber\\
347: \bar A_n J_n(\lambda a)+\bar B_n N_n(\lambda a) &=& 0\,\, ,
348: \label{ABdir}
349: \end{eqnarray}
350: i.e., the functions $E$ and $\bar E$ satisfy Dirichlet boundary
351: conditions on the surfaces. The coefficients of the series in Eqs.(\ref{E}) and (\ref{barE})
352: can be related to one another using the addition theorem for Bessel functions
353: \begin{equation}
354: e^{im\varphi}{\cal C}_m(\lambda\rho)=\sum_p e^{ip\theta}{\cal
355: C}_p(\lambda r)J_{p-m}(\lambda\epsilon)\,\, , \label{addition1}
356: \end{equation}
357: where ${\cal C}_m$ denotes either $J_m$ or $N_m$. Indeed, as at any
358: point $P$ in the annulus region one must have $E(P)=\bar E(P)$, it
359: is possible to show that
360: \begin{eqnarray}
361: \bar A_n &=& \sum_m A_m J_{n-m}(\lambda\epsilon) , \nonumber\\
362: \bar B_n &=& \sum_m B_m J_{n-m}(\lambda\epsilon) .
363: \label{relAB}
364: \end{eqnarray}
365: Combining Eqs. (\ref{ABdir}) and (\ref{relAB}) one obtains the linear, homogeneous
366: system of equations
367: \begin{equation}
368: \sum_m A_m \left[
369: \frac{J_n(\lambda a)}{N_n(\lambda a)} - \frac{J_m(\lambda b)}{N_m(\lambda b)}
370: \right] J_{n-m}(\lambda \epsilon) = 0 .
371: \end{equation}
372: The solution of this linear system of equations is non trivial only if
373: ${\rm det} [ Q^{\rm TM}_{mn} ] =0$, where
374: \begin{equation}
375: Q^{\rm TM}_{mn}(a,b,\epsilon)= \left[J_n(\lambda a) N_m(\lambda b)
376: -J_m(\lambda b) N_n(\lambda a)\right] J_{n-m}(\lambda
377: \epsilon)\,\, .
378: \end{equation}
379: This equation defines the allowed values for $\lambda$, and
380: therefore defines the eigenfrequencies of the TM modes.
381:
382: The TE modes can be treated in the same fashion. For these modes
383: the $z$ component of the electric field vanishes in the annulus
384: region, $E_z=0$. The perfect conductor boundary conditions imply
385: that the normal component of the magnetic field should vanish on
386: the conducting shells, so now we must impose Neumann boundary
387: conditions on the surfaces. The eigenvalues $\lambda$ for the TE
388: modes are the solutions of ${\rm det} [Q^{\rm TE}_{mn}]=0$, where
389: \begin{equation}
390: Q^{\rm TE}_{mn}(a,b,\epsilon)= \left[J'_n(\lambda a) N'_m(\lambda
391: b) -J'_m(\lambda b) N'_n(\lambda a)\right] J_{n-m}(\lambda
392: \epsilon).
393: \end{equation}
394:
395: In the concentric limit $\epsilon =0$ we have
396: $J_{n-m}(0)=\delta_{nm}$, the two matrices $Q^{\rm TE}_{mn}$ and
397: $Q^{\rm TM}_{mn}$ become diagonal, and the equations for the
398: eigenvalues are those of the concentric case
399: \cite{Mazzitelli2003}. In what follows we will use these matrices
400: to define the function $M$ that enters in Eq.(\ref{xxx}).
401:
402: %%%%%%%%%%
403:
404: \subsection{The function $M$}
405:
406: Roughly speaking, the function $M$ that determines the Casimir
407: energy through Eq.(\ref{xxx}) is the ratio of the function associated to
408: the actual geometric configuration and the one associated to a
409: configuration in which the conducting surfaces are very far away from each other.
410: As the last configuration is not univocally defined, we will use
411: this freedom to choose a particular one that simplifies the
412: calculation. It turns to be convenient to subtract a configuration
413: of two cylinders with very large (and very different) radii, but
414: with the same eccentricity as that of the configuration of interest.
415:
416: We start considering the Dirichlet modes.
417: To compute $F_{1{\rm cyl}}(a)$ in Eq.(\ref{Mformal}) we note that the
418: eigenfrequencies $\lambda$ for the geometry of a single cylinder of radius $a$
419: surrounded by a larger one of radius $R$ are defined by the equations
420: \begin{eqnarray}
421: && J_n(\lambda a)=0\, ,\nonumber\\
422: && J_n(\lambda a) N_n(\lambda R) - J_n(\lambda R) N_n(\lambda
423: a)=0\; . \label{condtm}
424: \end{eqnarray}
425: The first equation defines the eigenfrequencies in the region
426: $r<a$ and the second one gives the eigenfrequencies of the modes
427: in the region $a<r<R$. $F_{1{\rm cyl}}(a)$ is the product of these
428: two relations for all values of $n$, evaluated on the imaginary
429: axis ($\lambda=i y\equiv i\beta/a$). Namely,
430: \begin{eqnarray}
431: F_{1{\rm cyl}}(a)&=&\prod_n J_n(\lambda a)[J_n(\lambda a)N_n(\lambda R)-J_n(\lambda R)
432: N_n(\lambda a)]\nonumber \\
433: & \equiv & J(a) {\rm det}[ Q^{\rm TM}(a,R,0)] \; ,
434: \end{eqnarray}
435: where we have introduced the notation $J(a) \equiv \prod_n
436: J_n(\lambda a)$ to simplify the formulas below. The function
437: $F_{1{\rm cyl}}(\infty)$ has the same expression, but replacing
438: $a$ by $R_1$, with $R_1$ very large but smaller than $R$. Using
439: the asymptotic expansion of the modified Bessel functions it is
440: easy to prove that $F_{1{\rm cyl}}(a) / F_{1{\rm cyl}}(\infty)
441: \simeq 2 \beta I_n(\beta) K_n(\beta) R_1/a$. The functions $F$ and
442: $F_{\infty}$ in Eq.(\ref{Mformal}) are given by
443: \begin{eqnarray}
444: F &=& J(a) {\rm det}[Q^{\rm TM}(a,b,\epsilon)] {\rm det}[ Q^{\rm TM}(b,R,0)]\nonumber \\
445: &=& \frac {J(a)}{J(b)}{\rm det}[Q^{\rm TM}(a,b,\epsilon)] F_{1{\rm cyl}}(b),
446: \label{efe} \\
447: F_{\infty} &=& \frac {J(R_1)}{J(R_2)} {\rm det}[Q^{\rm
448: TM}(R_1,R_2,\epsilon)] F_{1{\rm cyl}}(\infty),
449: \label{efeinfinito}
450: \end{eqnarray}
451: where $R_1<R_2<R$. As we already stressed, in order to define $F_{\infty}$
452: we consider a configuration of two eccentric cylinders of large radii $R_1<R_2$ and with the same
453: eccentricity $\epsilon$ of the original configuration.
454: Evaluating the determinant in Eq.(\ref{efe}) on the imaginary axis one obtains
455: \begin{eqnarray}
456: {\rm det}[Q^{\rm TM}(a,b,\epsilon)] &=& {\rm det} \left[
457: \frac{2}{\pi} I_{n-m} \left( \beta\frac{\epsilon}{a} \right)
458: [K_n(\beta) I_m(\alpha\beta)\right. \nonumber \\
459: &-&\left. (-1)^{m+n}I_n(\beta)
460: K_m(\alpha\beta)] \right]. \label{detcomp}
461: \end{eqnarray}
462: Using again the asymptotic expansions of the Bessel functions one
463: gets $ {\rm det}[Q^{\rm TM}(R_1,R_2,\epsilon)] \propto a I_{n-m}
464: (\beta \epsilon / a) \frac{e^{\beta(R_2-R_1)/a}}{2\sqrt{R_1R_2}\beta}$. The
465: equations above can be combined to obtain
466: \begin{eqnarray}
467: M^{\rm TM}(\beta) &=&\, {\rm det}\, \left[I_{n-m}(\beta \epsilon/a)
468: \frac{I_m(\alpha\beta)}{I_n(\alpha\beta)} \left[1 -
469: (-1)^{m+n}\frac{ I_n(\beta)K_m(\alpha\beta)}{K_n(\beta)
470: I_m(\alpha\beta)}\right]\right ]\nonumber \\
471: &\times & {\rm det}\,
472: I_{nm}^{-1} \left( \beta \frac{\epsilon}{a} \right),
473: \label{Mdfin}
474: \end{eqnarray}
475: where $ I_{nm}^{-1}(\beta \epsilon / a)$ denotes the inverse
476: matrix of $ I_{n-m}(\beta \epsilon /a)$ and $\alpha \equiv b/a$.
477: Computing explicitly the determinant one can show that the factor
478: $I_m(\alpha\beta) / I_n(\alpha\beta)$ cancels out. Moreover,
479: writing $M^{\rm TM}$ as a single determinant we get
480: \begin{equation}
481: M^{\rm TM}(\beta)={\rm det} [\delta_{np}-A_{np}^{\rm TM}],
482: \end{equation}
483: with
484: \begin{equation}
485: A_{np}^{\rm TM}=(-1)^{n}\frac{I_n(\beta)}{K_n(\beta)} \sum_m
486: (-1)^m \frac{K_m(\alpha\beta)}{I_m(\alpha\beta)} I_{n-m} \left(
487: \beta\frac{\epsilon}{a} \right) I_{mp}^{-1} \left(
488: \beta\frac{\epsilon}{a} \right). \label{atmfin}
489: \end{equation}
490: The addition theorem for the modified Bessel functions, ${\cal
491: C}_m(u\pm v)=\sum_p {\cal C}_{m\mp p}(u)J_{p}(v)$ \cite{foot},
492: implies that $I_{mp}^{-1}(\beta \epsilon/a)=(-1)^{m-p}
493: I_{m-p}(\beta \epsilon / a)$. Finally, the elements of the matrix
494: $A^{\rm TM}$ read
495: \begin{equation}
496: A_{np}^{\rm TM}= \frac{I_n(\beta)}{K_n(\beta)}
497: \sum_m \frac{K_m(\alpha\beta)}{I_m(\alpha\beta)}
498: I_{n-m} \left( \beta\frac{\epsilon}{a} \right) I_{p-m}\left( \beta\frac{\epsilon}{a} \right),
499: \label{Atm}
500: \end{equation}
501: where we omitted a global factor $(-1)^{n+p}$ because it does not contribute to the determinant.
502:
503: The analysis for the TE modes is straightforward, the main
504: difference being that $Q^{\rm TE}(a,b,\epsilon)$ contains
505: derivatives of those Bessel functions that do not depend on the
506: eccentricity. Therefore, following similar steps it is possible to
507: show that
508: \begin{equation}
509: M^{\rm TE}(\beta)={\rm det}[\delta_{np}-A_{np}^{\rm TE}],
510: \end{equation}
511: where
512: \begin{equation}
513: A_{np}^{\rm TE}= \frac{I'_n(\beta)}{K'_n(\beta)}
514: \sum_m \frac{K'_m(\alpha\beta)}{I'_m(\alpha\beta)}
515: I_{n-m}\left( \beta\frac{\epsilon}{a} \right) I_{p-m}\left( \beta\frac{\epsilon}{a} \right).
516: \label{Ate}
517: \end{equation}
518: The function $M$ for the electromagnetic field is the product $M=M^{\rm TE}M^{\rm TM}$,
519: and therefore the interaction energy is the sum of the TE and TM contributions
520: \begin{eqnarray}
521: E_{12}&=&\frac{L}{4\pi a^2}\int_0^{\infty}d\beta\, \beta\, \ln
522: M(\beta)=\frac{L}{4\pi a^2}\int_0^{\infty}d\beta\, \beta\, \ln
523: M^{\rm TE}(\beta) \nonumber \\
524: &+&\frac{L}{4\pi a^2}\int_0^{\infty}d\beta\, \beta\, \ln M^{\rm
525: TM}(\beta)=E^{{\rm TE}} + E^{{\rm TM}} . \label{e12tmte}
526: \end{eqnarray}
527:
528: In order to calculate the exact Casimir interaction energy one
529: needs to perform a numerical evaluation of the determinants. We
530: find that as $\alpha$ approaches smaller values, larger matrices
531: are needed for ensuring convergence. Moreover, for increasing
532: values of the eccentricity $\epsilon$ it is necessary to include
533: more terms in the series defining the coefficients $A_{np}^{\rm
534: TE, TM}$. In Fig. 3 we plot the interaction energy difference
535: $\Delta E= E_{12} -E_{12}^{\rm cc}$ between the eccentric
536: ($E_{12}$) and the concentric ($E_{12}^{\rm cc}$) configurations
537: as a function of $\alpha$ for different values of
538: $\delta$. As we will show below, these numerical results interpolate between the
539: PFA and the asymptotic behavior for large $\alpha$. Fig. 4 shows
540: the complementary information, with the Casimir energy as a
541: function of $\delta$ for various values of $\alpha$, showing
542: explicitly that the concentric equilibrium position is unstable.
543:
544: %%%%%%%%%%%%
545: %Figure 3
546: %%%%%%%%%%%%
547:
548: \begin{figure}
549: \centerline{\psfig{figure=njp.fig3.eps,height=6cm,width=9cm,angle=0}}
550: \caption{Exact Casimir interaction energy difference $|\Delta E|$ between the eccentric and concentric configurations as a function of $\alpha=b/a$ for different values of $\delta=\epsilon/a$.
551: Here $\Delta E = E_{12} - E_{12}^{\rm cc}$.
552: Energies are measured in units of $L/4 \pi a^2$. These results interpolate between the $(\alpha-1)^{-5}$ behavior
553: for $\alpha \rightarrow 1$, and the $(\alpha^4 \log \alpha)^{-1}$ behavior for $\alpha \gg 1$.}
554: \label{fig3}
555: \end{figure}
556:
557:
558:
559:
560: %%%%%%%%%%%%
561: %Figure 4
562: %%%%%%%%%%%%
563:
564: \begin{figure}
565: \centerline{\psfig{figure=njp.fig4.eps,height=6cm,width=9cm,angle=0}}
566: \caption{Exact Casimir interaction energy difference $\Delta E$ between the eccentric and concentric configurations as a function of $\delta=\epsilon/a$ for different values of $\alpha=b/a$. Energies are measured
567: in units of $L/4 \pi a^2$. The maximum at $\delta=0$ shows the instability of the concentric equilibrium position.}
568: \label{fig4}
569: \end{figure}
570:
571:
572:
573:
574: %%%%%%%%%%%%%%%%%
575:
576: \subsection{Concentric Cylinders}
577:
578: The exact Casimir interaction between concentric cylinders \cite{Mazzitelli2003} can be
579: obtained as a particular case of the exact formulas (\ref{Atm}), (\ref{Ate}).
580: In the concentric limit $\epsilon =0$, the matrices that appear in
581: the definition of $M^{\rm TE}$ and $M^{\rm TM}$ become diagonal and the Casimir energy reads
582: \cite{Mazzitelli2003}
583: \begin{equation}
584: E_{12}^{\rm cc} = {L \over 4\pi a^2} \int_{0}^{\infty} d\beta \
585: \beta\ln M^{\rm cc}(\beta), \label{cc}
586: \end{equation}
587: where
588: \begin{equation}
589: M^{\rm cc}(\beta)=\prod_n \left[1-{I_n(\beta)K_n(\alpha
590: \beta)\over I_n(\alpha \beta)K_n(\beta)}\right]
591: \left[1-{I'_n(\beta)K'_n(\alpha \beta) \over I'_n(\alpha
592: \beta)K'_n(\beta)}\right] . \label{Mcc}
593: \end{equation}
594: The first factor corresponds to Dirichlet (TM) modes and the
595: second one to Neumann (TE) modes.
596:
597: The proximity limit $\alpha - 1\ll 1$ has already been analyzed
598: for the concentric case \cite{Mazzitelli2003}. In order to
599: compute the concentric Casimir interaction energy in this limit
600: it was necessary to perform the summation over all values of $n$.
601: As expected, the resulting value is equal to the one obtained via
602: the proximity approximation, namely
603: \begin{equation}
604: E_{12, {\rm PFA}}^{\rm TE, cc} = E_{12, {\rm PFA}}^{\rm TM, cc} = \frac{1}{2} E_{12, {\rm PFA}}^{\rm EM, cc} =
605: - \frac{\pi^3 L}{720 a^2} \; \frac{1}{(\alpha-1)^3}.
606: \end{equation}
607: Here $E^{\rm EM, cc}$ denotes the full electromagnetic Casimir interaction energy in the concentric configuration.
608:
609: In the large $\alpha$ limit one can show that only the $n=0$ term contributes to the
610: interaction energy
611: \begin{equation}
612: E_{12}^{\rm cc} \approx {L \over 4\pi b^2} \int_{0}^{\infty} dx \ x \left[\ln
613: \left(1 - \frac{I_0(\frac{x}{\alpha}) K_0(x)}{I_0(x) K_0(\frac{x}{\alpha})}\right) + \ln
614: \left(1 - \frac{I'_0(\frac{x}{\alpha}) K'_0(x)}{I'_0(x) K'_0(\frac{x}{\alpha})}\right)\right]. \end{equation}
615: Using the small argument behavior of the Bessel functions it is easy to prove
616: that, in the limit $\alpha \gg 1$, the TM mode contribution dominates, giving
617: \begin{equation} E_{12}^{\rm cc}
618: \approx - {L \over 4\pi b^2\ln\alpha} \int_{0}^{\infty} dx \
619: x \frac{K_0(x)}{I_0(x)} \approx - {1.26 L \over 8\pi
620: b^2\ln\alpha}. \label{cclargea}
621: \end{equation}
622: Note that the modulus of the energy decreases logarithmically with
623: $\alpha$. Fig. 5 depicts the exact Casimir interaction energy
624: between concentric cylinders as a function of $\alpha$, for values
625: that interpolate between the above mentioned limiting cases.
626:
627: It is worth noticing that, while for small values of $\alpha$ both
628: TM and TE modes contribute with the same weight to the interaction
629: energy, the TM modes dominate in the large $\alpha$ limit.
630:
631: %%%%%%%%%%%%
632: %Figure 5
633: %%%%%%%%%%%%
634:
635: \begin{figure}
636: \centerline{\psfig{figure=njp.fig5.eps,height=6cm,width=9cm,angle=0}}
637: \caption{Modulus of the Casimir interaction energy in the
638: concentric case as a function of $\alpha=b/a$. Energies are
639: measured in units of $L/4 \pi a^2$. These results interpolate
640: between the $(\alpha-1)^{-3}$ behavior for $\alpha \rightarrow 1$,
641: and the $(\alpha^2 \log \alpha)^{-1}$ behavior for $\alpha \gg
642: 1$.} \label{fig5}
643: \end{figure}
644:
645:
646: \subsection{A cylinder in front of a plane}
647:
648: It is interesting to see how the Casimir energy for the
649: cylinder-plane configuration is contained as a particular case of
650: the exact formula derived in Section {\it 3.2}. To do that let us
651: consider a cylinder of radius $a$ in front of an infinite plane,
652: and let us denote by $H$ the distance between the center of the
653: cylinder and the plane. The eccentric cylinders formula should
654: reproduce the cylinder-plane Casimir energy in the limit
655: $b,\epsilon \rightarrow\infty$ keeping $H=b-\epsilon$ fixed (see
656: Fig. 1).
657:
658: We note that for $x\gg h >1$ the ratio of Bessel functions appearing in Eq.(\ref{Atm}) can be approximated by
659: \begin{equation}
660: \frac{I_{m-n}(x)I_{m-p}(x)}{I_m(x+h)}\simeq I_{m-n-p}(x-h) .
661: \label{prodI}
662: \end{equation}
663: This is trivially true for fixed $m$ and large values of $x$, as can be seen from the
664: large argument expansion of the Bessel functions. Moreover, using the uniform expansion of
665: the Bessel functions, it can be shown that Eq.(\ref{prodI}) is also valid in the large $m$ limit,
666: for all values of $x$. Therefore we approximate
667: \begin{eqnarray}
668: \sum_m\frac{K_m(x+h)}{I_m(x+h)}I_{m-n}(x)I_{m-p}(x)& \simeq & \sum_m
669: K_m(x+h)I_{m-p-n}(x-h) \nonumber \\ &=&K_{n+p}(2h),
670: \label{addition}
671: \end{eqnarray}
672: where in the last equality we used the addition theorem of Bessel
673: functions. Inserting this result (with $x\equiv \beta\epsilon/a$
674: and $h\equiv \beta H/a$) in Eq.(\ref{Atm}) we get
675: \begin{equation}
676: A_{np}^{\rm TM}\simeq \frac{I_n(\beta)}{K_n(\beta)}K_{n+p}(2\beta H/a)
677: \equiv A_{np}^{\rm TM, c-p},
678: \end{equation}
679: which coincides with the known result for the Dirichlet matrix elements for the cylinder-plane geometry
680: \cite{Emig2006,Bordag2006}. The TE modes can be analyzed in the same fashion. Using that for large $x$
681: \begin{equation}
682: \frac{K'_m(x)I_m(x)}{K_m(x)I'_m(x)}\simeq -1\,\, ,
683: \end{equation}
684: one can prove that
685: \begin{eqnarray}
686: \sum_m\frac{K'_m(x+h)}{I'_m(x+h)}I_{m-n}(x)I_{m-p}(x)& \simeq &
687: -\sum_m K_m(x+h)I_{m-p-n}(x-h)\nonumber \\
688: &=& -K_{n+p}(2h).
689: \label{addition2}
690: \end{eqnarray}
691: Therefore, from Eq.(\ref{Ate}) we obtain
692: \begin{equation}
693: A_{np}^{\rm TE}\simeq -\frac{I'_n(\beta)}{K'_n(\beta)}K_{n+p}(2\beta
694: H/a) \equiv A_{np}^{\rm TE, c-p},
695: \end{equation}
696: which is the result for the TE modes in the cylinder-plane geometry \cite{Emig2006,Bordag2006}.
697:
698:
699: \section{Quasi-concentric cylinders}
700:
701: We will now consider a situation in which the eccentricity of the
702: configuration is much smaller than the radius of the inner
703: cylinder, i.e., $\delta = \epsilon /a \ll 1$. As discussed in
704: \cite{Dalvit2004} this configuration may be relevant for
705: performing null experiments to look for extra gravitational
706: forces. Note that we do not assume that the radius of the inner
707: and outer cylinder are similar, so the proximity approximation is
708: in general not valid for this configuration.
709:
710: As described in the previous section, when $\epsilon =0$ the matrix
711: defining the eigenfrequencies is diagonal. When considering a
712: small non-vanishing eccentricity, the behavior of the Bessel
713: functions for small arguments $I_{m-n}(\beta \delta)\sim (\beta
714: \delta)^{n-m}$ suggests that one only needs to use matrix elements
715: near the diagonal. Using this idea, we will approximate the
716: Casimir interaction energy by keeping only terms proportional to $I_0,
717: I_1$ and $I_1^2$. In this approximation, the matrices
718: $\delta_{np}-A_{np}^{\rm TM}$ and $\delta_{np}-A_{np}^{\rm TE}$ become
719: tridiagonal matrices, and the $\epsilon-$dependent part of the Casimir energy
720: will be quadratic in the eccentricity.
721:
722: We will describe in detail the case of the Dirichlet modes; the treatment of Neumann modes is
723: similar. To order $O(\delta^2)$ the only non-vanishing elements of the matrix $A_{np}^{\rm TM}$ are
724: \begin{eqnarray}
725: A_{n,n}^{\rm TM}&\simeq &\frac{I_n(\beta)}{K_n(\beta)}\left[
726: \frac{K_n(\alpha\beta)}{I_n(\alpha\beta)}I_0^2(\delta\beta) +
727: \frac{K_{n-1}(\alpha\beta)}{I_{n-1}(\alpha\beta)}I_1^2(\delta\beta)
728: +\frac{K_{n+1}(\alpha\beta)}{I_{n+1}(\alpha\beta)}I_1^2(\delta\beta)\right],\nonumber\\
729: A_{n,n+1}^{\rm TM}&\simeq &\frac{I_n(\beta)}{K_n(\beta)}\left[
730: \frac{K_n(\alpha\beta)}{I_n(\alpha\beta)}
731: +\frac{K_{n+1}(\alpha\beta)}{I_{n+1}(\alpha\beta)}\right]I_0(\delta\beta)I_1(\delta\beta),\nonumber\\
732: A_{n+1,n}^{\rm TM}&\simeq
733: &\frac{I_{n+1}(\beta)}{K_{n+1}(\beta)}\left[
734: \frac{K_n(\alpha\beta)}{I_n(\alpha\beta)}
735: +\frac{K_{n+1}(\alpha\beta)}{I_{n+1}(\alpha\beta)}\right]I_0(\delta\beta)I_1(\delta\beta).
736: \label{tri}
737: \end{eqnarray}
738: We split the matrix $A^{\rm TM}$ into three terms, $A^{\rm
739: TM}={\cal D}^{\rm TM, cc}+{\cal D}^{\rm TM} + {\cal N}^{\rm TM}$,
740: where ${\cal D}^{\rm TM, cc}$ is the diagonal matrix corresponding
741: to the concentric case, ${\cal D}^{\rm TM}$ the diagonal part of
742: the matrix that depends on $\delta$, and ${\cal N}^{\rm TM}$ is
743: the non-diagonal part of the matrix. The non-vanishing matrix
744: elements are
745: \begin{eqnarray}
746: {\cal D}_{n,n}^{\rm TM, cc} &=& \frac{I_n(\beta)}{K_n(\beta)} \frac{K_n(\alpha\beta)}{I_n(\alpha\beta)},\nonumber\\
747: {\cal D}_{n,n}^{\rm TM} &=&A_{n,n}^{\rm TM}-{\cal D}_{n,n}^{\rm TM, cc},\nonumber\\
748: {\cal N}_{n,n+1}^{\rm TM} &=&A_{n,n+1}^{\rm TM} , \quad {\cal N}_{n+1,n}^{\rm TM} =
749: A_{n+1,n}^{\rm TM} .
750: \label{tri2}
751: \end{eqnarray}
752: Note that although ${\cal D}^{\rm TM}=O(\delta^2)$ and ${\cal N}^{\rm TM}=O(\delta)$, both give quadratic
753: contributions
754: to the determinant. Up to this order we have
755: \begin{eqnarray}
756: \ln {\rm det}\,[1-A^{\rm TM}]&\simeq& \ln {\rm det}\,[1-{\cal D}^{\rm TM, cc}] +
757: \ln {\rm det} \left[ 1-\frac{{\cal D}^{\rm TM}}{1-{\cal D}^{\rm TM, cc}} \right]\nonumber \\ &+&
758: \ln {\rm det} \left[ 1-\frac{{\cal N}^{\rm TM}}{1-{\cal D}^{\rm TM, cc}} \right].
759: \label{dettri}
760: \end{eqnarray}
761: The first term is associated to the interaction energy between
762: concentric cylinders $E_{12}^{\rm cc}$, studied in Section {\it 3.3}, and
763: being $\delta$-independent does not
764: contribute to the force between eccentric cylinders.
765: The second term can be easily evaluated
766: \begin{equation}
767: \ln {\rm det} \left[ 1-\frac{{\cal D}^{\rm TM}}{1-{\cal D}^{\rm
768: TM, cc}} \right] \simeq\ln\left( 1-{\rm tr}\, \frac{{\cal D}^{\rm
769: TM}}{1-{\cal D}^{\rm TM, cc}}\right)\simeq -\sum_n\frac{{\cal
770: D}^{\rm TM}_{n,n}}{1-{\cal D}_{n,n}^{\rm TM, cc}} . \label{diagpart}
771: \end{equation}
772: To compute the last term in Eq.(\ref{dettri}) we use that the
773: determinant of an arbitrary tridiagonal matrix $T$ of dimension
774: $p$ can be calculated using the recursive relation
775: ${\rm det} [ T_{ \{ p\} } ] = T_{p,p} {\rm det} [T_{ \{ p-1\} } ] -
776: T_{p,p-1} T_{p-1, p} \, {\rm det}[ T_{ \{ p-2\} } ]$,
777: where $T_{\{ k\} }$ denotes the submatrix formed by the first $k$
778: rows and columns of $T$. Up to quadratic order in $\delta$ we
779: obtain
780: \begin{eqnarray}
781: \ln {\rm det} \left[ 1-\frac{{\cal N}^{\rm TM}}{1-{\cal D}^{\rm
782: TM, cc}} \right] &\simeq &\ln\left( 1- \sum_n \frac{A_{n, n+1}^{\rm
783: TM} \; A_{n+1, n}^{\rm TM}}{(1-{\cal D}_{n,n}^{\rm TM, cc})(1-{\cal
784: D}_{n+1,n+1}^{\rm TM, cc})} \right) \nonumber \\ &\simeq & -\sum_n \frac{A_{n,
785: n+1}^{\rm TM} \; A_{n+1, n}^{\rm TM}}{(1-{\cal D}_{n,n}^{\rm TM,
786: cc})(1-{\cal D}_{n+1,n+1}^{\rm TM, cc})}. \label{nondiagpart}
787: \end{eqnarray}
788: Putting all together, the TM part of the Casimir interaction energy between quasi-concentric
789: cylinders can be written
790: as
791: \begin{equation}
792: E_{12}^{\rm TM} = E_{12}^{{\rm TM, cc}} - \frac{L \epsilon^2}{4
793: \pi a^4} \sum_n \int_0^{\infty} d\beta \; \beta^3
794: \frac{1}{1-{\cal D}^{\rm TM, cc}_{n,n}} \left[ {\cal D}^{\rm TM}_n +
795: \frac{{\cal N}^{\rm TM}_n}{1-{\cal D}^{\rm TM, cc}_{n+1,n+1}} \right].
796: \label{finalqcc}
797: \end{equation}
798: Here
799: \begin{eqnarray}
800: {\cal D}^{\rm TM}_n & \equiv & \frac{ {\cal D}^{\rm TM, cc}_{n,n}}{2} + \frac{I_n(\beta)}{4 K_n(\beta)} \left[
801: \frac{K_{n-1}(\alpha\beta)}{I_{n-1}(\alpha \beta)} +
802: \frac{K_{n+1}(\alpha \beta)}{I_{n+1}(\alpha \beta)} \right] , \nonumber \\
803: {\cal N}^{\rm TM}_n & \equiv & \frac{I_n(\beta) I_{n+1}(\beta)}{4
804: K_n(\beta) K_{n+1}(\beta)} \left[ \frac{K_{n}(\alpha
805: \beta)}{I_{n}(\alpha \beta)} + \frac{K_{n+1}(\alpha
806: \beta)}{I_{n+1}(\alpha \beta)} \right]^2 . \label{newd}
807: \end{eqnarray}
808: The corresponding formulas for the TE modes can be obtained from
809: these ones by replacing the Bessel functions by their derivatives
810: with respect to the argument.
811:
812: The expression for the Casimir energy for quasi-concentric
813: cylinders derived in this section is far simpler than the exact
814: formulas Eqs.(\ref{Atm}), (\ref{Ate}). It is very useful for the
815: analytical and numerical evaluation of the Casimir energy in the
816: different limiting cases we will study below: the large distance
817: limit ($a \ll b$), for which one obtains a logarithmic decay of
818: the energy, and the small distance limit ($a \simeq b$), where the
819: proximity approximation holds. The first case is very simple to
820: handle because the energy is dominated by the lowest modes, while
821: the second case is much more involved.
822:
823: \subsection{Large distances: logarithmic decay}
824:
825: When the ratio of the outer and the inner radii $\alpha = b/a$ is
826: much larger than one, the exact Casimir energy is dominated by the
827: lowest term $n=0$ in the summation. Moreover, it can be shown that
828: the contribution of the Dirichlet modes is much larger than the
829: contribution of the Neumann modes. Therefore, from
830: Eq.(\ref{e12tmte}) we get, in the limit $\alpha \rightarrow
831: \infty$,
832: \begin{equation}
833: E_{12}^{\infty} \simeq {L \over 4\pi a^2} \int_{0}^{\infty} d\beta
834: \; \beta \ln(1-A_{00}^{\rm TM}(\beta)) \simeq -{L \over 4\pi
835: a^2\alpha^2}\int_{0}^{\infty} dx \; x \; A_{00}^{\rm TM} \left(
836: \frac{x}{\alpha} \right) , \label{large}
837: \end{equation}
838: where
839: \begin{equation} A_{00}^{\rm TM}\left( \frac{x}{\alpha}\right) \simeq
840: \frac{I_0(\frac{x}{\alpha})}{K_0(\frac{x}{\alpha})}\left[
841: \frac{K_0(x)}{I_0(x)}I_0^2\left(\frac{\delta x}{\alpha}\right) +
842: 2\frac{K_1(x)}{I_1(x)}I_1^2\left(\frac{\delta
843: x}{\alpha}\right)\right]\, .
844: \end{equation}
845: Using the small argument expansion of the Bessel functions it is
846: easy to see that
847: \begin{equation}
848: A_{00}^{\rm TM}\left(\frac{x}{\alpha} \right)\simeq
849: \frac{1}{\ln\alpha}\left[\frac{K_0(x)}{I_0(x)}+
850: \frac{\delta^2x^2}{2\alpha^2}\left(\frac{K_0(x)}{I_0(x)}+\frac{K_1(x)}{I_1(x)}\right)\right].
851: \label{a00}
852: \end{equation}
853: In this expression, valid when $a,\epsilon \ll b$, we
854: kept the leading terms proportional to $(\alpha^2\ln\alpha )^{-1}$
855: and only the subleading terms that depend on the eccentricity.
856: Inserting Eq.(\ref{a00}) into Eq.(\ref{large}) and computing
857: numerically the integrals we find
858: \begin{equation}
859: E_{12}^{\infty} \simeq -{L \over 8\pi b^2\ln\alpha}
860: \left(1.26+3.33 \frac{\epsilon^2}{b^2} \right),
861: \label{resultlarge}
862: \end{equation}
863: where the first term is the concentric contribution
864: $E_{12}^{\infty, {\rm cc}}$ derived before (see
865: Eq.(\ref{cclargea})). It is worth to note that Eqs. (\ref{a00})
866: and (\ref{resultlarge}) have been derived under the assumption
867: $\ln\alpha\gg 1$, and therefore are valid for extremely large
868: values of $\alpha$. For intermediate values $\alpha\gg 1,
869: \ln\alpha =O(1)$, the interaction energy is also dominated by the
870: Dirichlet $n=0$ term. The final result is still of the form given
871: in Eq.(\ref{resultlarge}), with numerical coefficients that depend
872: logarithmically on $\alpha$. In Fig. 6 we plot the ratio between
873: the exact Casimir interaction energy difference $\Delta E \equiv
874: E_{12} - E_{12}^{\rm cc}$ and its asymptotic expression $\Delta
875: E_{\infty} \equiv E_{12}^{\infty} - E_{12}^{\infty,{\rm cc}}$ as
876: a function of $\alpha$. As mentioned, extremely large values of
877: $\alpha$ are needed in order for the ratio of energies to
878: asymptotically approach 1. From Eq.(\ref{resultlarge}) we see that
879: the force between cylinders in the limit $a,\epsilon\ll b$ is
880: proportional to $L \epsilon /b^4\ln(b/a)$. The weak logarithmic
881: dependence on the ratio $b/a$ is characteristic of the cylindrical
882: geometry (see also \cite{Emig2006,Bordag2006}), and it is also
883: found in the electrostatic counterpart of the Casimir energy, that
884: we briefly analyze next.
885:
886: %%%%%%%%%%%%
887: %Figure 6
888: %%%%%%%%%%%%
889:
890: \begin{figure}
891: \centerline{\psfig{figure=njp.fig6.eps,height=6cm,width=9cm,angle=0}}
892: \caption{Ratio of the exact $\Delta E$ and asymptotic $\Delta E_{\infty}$ Casimir energy differences in the limit of small eccentricity $\epsilon \ll a$.
893: In the $\alpha \rightarrow \infty$ limit, the Casimir energy difference between eccentric
894: and concentric configurations decays logarithmically as $(\alpha^4 \log \alpha)^{-1}$.}
895: \label{fig6}
896: \end{figure}
897:
898: The electrostatic capacity for the system of two eccentric cylinders is given by
899: \begin{equation}
900: C=\frac{2\pi\epsilon_0L}{\ln [Y+\sqrt{Y^2-1}]}, \label{capacity}
901: \end{equation}
902: where $Y=(a^2+b^2-\epsilon^2)/2ab$, and $\epsilon_0$ is the permittivity of vacuum.
903: Therefore, the electrostatic force between cylinders kept at a fixed potential difference $V$ is
904: \begin{equation}
905: F_{\rm elec}=\frac{\epsilon}{ab}\frac{\pi\epsilon_0 V^2L}{\sqrt{Y^2-1}\ln^2[Y+\sqrt{Y^2-1}]} .
906: \label{electroforce}
907: \end{equation}
908: In the quasi-concentric case we can set $\epsilon =0$ in the
909: definition of $Y$. In the large $\alpha$ limit we get
910: \begin{equation}
911: F_{\rm elec} \simeq \frac{L \epsilon}{b^2\log ^2 \left( \frac{b}{a} \right) } .
912: \end{equation}
913: Just as in the Casimir case, in the limit $a\ll b$ the Coulomb force decays logarithmically with
914: the ratio $a/b$.
915:
916: \subsection{Small distances: the proximity approximation}
917:
918: The proximity limit for concentric cylinders has been reviewed in
919: Sec. {\it 3.3}; the case of a cylinder in front of a plane has
920: been considered in detail in \cite{Bordag2006}. In this section we
921: extend these results to the case of quasi-concentric cylinders.
922:
923: We will concentrate on calculating the Casimir interaction energy
924: difference $\Delta E \equiv E_{12} - E_{12}^{\rm cc}$ between the
925: eccentric ($E_{12}$) and the concentric ($E_{12}^{\rm cc}$)
926: configurations. As the small distance limit is dominated by the
927: large-$n$ modes, the key point in the derivation of the PFA from
928: the exact expression of the Casimir energy is the use of the
929: uniform approximation for the Bessel functions. In the large $n$
930: limit, and to leading order in $\alpha -1$ one has
931: \begin{eqnarray}
932: \frac{I_n(\beta)}{K_n(\beta)}\frac{K_n(\alpha\beta)}{I_n(\alpha\beta)}&
933: \simeq & e^{-2n(\alpha-1)h(x)},\nonumber \\
934: \frac{I_n(\beta)}{K_{n}(\beta)}\frac{K_{n\pm
935: 1}(\alpha\beta)}{I_{n\pm 1}(\alpha\beta)}& \simeq &
936: e^{-2n(\alpha-1)h(x)}\left[\frac{1+h(x)}{x}\right]^{\pm 2},
937: \label{unifapprox}
938: \end{eqnarray}
939: where $\beta = n x$ and $h(x)=\sqrt{1+x^2}$. Inserting these
940: approximations in Eq.(\ref{newd}) we get
941: \begin{equation}
942: {\cal D}^{\rm TM}_n(n x)=\frac {e^{-2 n (\alpha-1)h(x)}}{2} \left[
943: 1+ \frac{1}{2} \left( \frac{1+h(x)}{x} \right)^{2} +
944: \frac{1}{2}\left( \frac{1+h(x)}{x} \right)^{-2} \right].
945: \end{equation}
946:
947: %%%%%%%%%%%%
948: %Figure 7
949: %%%%%%%%%%%%
950: \begin{figure}
951: \centerline{\psfig{figure=njp.fig7.eps,height=6cm,width=9cm,angle=0}}
952: \caption{Ratio of the exact and PFA Casimir interaction energy differences $\Delta E = E_{12} -E_{12}^{\rm cc}$
953: between eccentric ($E_{12}$) and concentric ($E_{12}^{\rm cc}$) cylinders in the limit of small eccentricity
954: $\epsilon \ll a$. The curve EM denotes the full electromagnetic Casimir energy.}
955: \label{fig7}
956: \end{figure}
957:
958: The contribution to the interaction energy coming from the
959: diagonal part of the matrix can be written as (see
960: Eq.(\ref{finalqcc}))
961: \begin{equation}
962: \Delta E_{\cal D}^{\rm TM}=-\frac{L\delta^2}{2\pi a^2}
963: \sum_{n,k\geq 1} \int_0^\infty d\beta \; \beta^3 \; {\cal D}^{\rm
964: TM}_n ({\cal D}_{n,n}^{\rm TM, cc})^{k-1} , \label{ED}
965: \end{equation}
966: where we replaced the sum over all integers $n$ by twice the sum
967: over the positive integers (the term $n=0$ gives a subleading
968: contribution for small $\alpha -1$). Inserting the uniform expansions
969: into Eq.(\ref{ED}) and changing variables in the integral we obtain
970: \begin{eqnarray}
971: \Delta E^{\rm TM}_{\cal D}&=& -\frac{L\delta^2}{4\pi a^2}\sum_{n,k\geq 1}n^4\int_0^\infty
972: dx \; x^3 e^{-2n(\alpha-1)h(x)k} \nonumber \\
973: && \times \left[ 1+\frac{1}{2}\left( \frac{1+h(x)}{x} \right)^{2}+
974: \frac{1}{2}\left( \frac{1+h(x)}{x} \right)^{-2} \right] .
975: \label{ED2}
976: \end{eqnarray}
977: To leading order in $\alpha -1$ the sum over $n$ gives
978: $\sum_{n} n^4 e^{-2n(\alpha-1)h(x)}\simeq 24/ [2 h(x)(\alpha-1)k]^5$.
979: Next we perform first the sum over $k$ and then the integral over
980: $x$. We get
981: \begin{equation}
982: \Delta E^{\rm TM}_{\cal D} = -\frac{3}{8}\frac{L\delta^2\zeta(5)}{\pi a^2 (\alpha - 1)^5} ,
983: \label{ED3}
984: \end{equation}
985: where $\zeta(x)$ is the Riemann zeta function.
986:
987: The evaluation of the non diagonal contribution to the Casimir
988: energy can be done using similar steps, starting from
989: Eq.(\ref{nondiagpart}). In the proximity limit, we can approximate
990: ${\cal D}_{n+1,n+1}^{\rm TM, cc}$ by ${\cal D}_{n,n}^{\rm TM, cc}$ in
991: the denominator of that equation. Therefore, using
992: Eq.(\ref{finalqcc})we write the non diagonal contribution to the
993: energy as
994: \begin{equation}
995: \Delta E^{\rm TM}_{\cal ND} \simeq -\frac{L\delta^2}{2\pi
996: a^2}\int_0^\infty d\beta \; \beta^3 \sum_{n,k\geq 1} {\cal N}^{\rm TM}_n
997: \;k \; ( {\cal D}_{n,n}^{\rm
998: TM, cc})^{k-1}. \label{END2}
999: \end{equation}
1000: Now we use the uniform expansion for the Bessel functions in
1001: Eq.(\ref{newd}) to obtain
1002: \begin{eqnarray}
1003: {\cal N}^{\rm TM}_n &\simeq & \frac{e^{-4n(\alpha -1)h(x)}}{2} \nonumber \\
1004: && \times \left[ 1+\frac{1}{2}\left( \frac{1+h(x)}{x} \right)^{2}
1005: + \frac{1}{2}\left( \frac{1+h(x)}{x} \right)^{-2} \right]
1006: \label{an+1} .
1007: \end{eqnarray}
1008: Replacing Eq.(\ref{an+1}) into Eq.(\ref{END2}) we get
1009: \begin{eqnarray}
1010: \Delta E^{\rm TM}_{\cal ND} &=& -\frac{L\delta^2}{8\pi a^2}\sum_{n,k\geq1} n^4 \int_0^{\infty} dx \; x^3 \;
1011: e^{-2n(k+1)(\alpha-1)h(x)}\; k \nonumber \\
1012: &&
1013: \times \left[
1014: 2 + \left( \frac{1+h(x)}{x} \right)^{2} + \left( \frac{1+h(x)}{x} \right)^{-2}
1015: \right].
1016: \end{eqnarray}
1017: As before, we first compute the sum over $n$, and expand the
1018: result to leading order in $\alpha-1$. The sum over $k$ can be
1019: calculated using that $\sum_{k\geq 1}\frac{k}{(k+1)^5}=\zeta (4)-\zeta(5)$.
1020: Finally, we compute analytically the remaining integrals to get
1021: \begin{equation}
1022: \Delta E^{\rm TM}_{\cal ND}=-\frac{3}{8}\frac{L\delta^2}{\pi a^2(\alpha-1)^5}(\zeta(4)-\zeta(5)).
1023: \end{equation}
1024: The contribution of the Dirichlet modes to the Casimir interaction
1025: energy in the limit $\alpha \rightarrow 1$ is therefore
1026: \begin{equation}
1027: \Delta E^{\rm TM}=\Delta E^{\rm TM}_{\cal D} + \Delta E^{\rm TM}_{\cal ND}=
1028: -\frac{L\delta^2}{a^2(\alpha-1)^5}\frac{\pi^3}{240}.
1029: \end{equation}
1030: It can be shown that the contribution of the TE modes to the
1031: interaction energy in the short distance limit is equal to that of
1032: the TM modes, as expected from the parallel plate configuration.
1033: Indeed, the uniform expansion for the ratio of Bessel functions is
1034: equal to the expansion for the derivatives, i.e.,
1035: \begin{eqnarray}
1036: \frac{I_n'(\beta)}{K_n'(\beta)}\frac{K_n'(\alpha\beta)}{I_n'(\alpha\beta)}&
1037: \simeq & e^{-2n(\alpha-1)h(x)},\nonumber\\
1038: \frac{I_n'(\beta)}{K_{n}'(\beta)}\frac{K_{n\pm
1039: 1}'(\alpha\beta)}{I_{n\pm 1}'(\alpha\beta)}& \simeq &
1040: e^{-2n(\alpha-1)h(x)}\left[\frac{1+h(x)}{x}\right]^{\pm 2},
1041: \label{unifapprox2}
1042: \end{eqnarray}
1043: and therefore all calculations can be repeated without changes.
1044:
1045: The final result for the Casimir interaction energy difference in the small distance approximation is
1046: \begin{equation}
1047: \Delta E_{\rm PFA}^{\rm TE} = \Delta E_{\rm PFA}^{\rm TM} = \frac{1}{2} \Delta E_{\rm PFA}^{\rm EM} =
1048: - \frac{\pi^3 L \epsilon^2}{240 a^4 (\alpha-1)^5} ,
1049: \end{equation}
1050: where $\Delta E^{\rm EM}$ denotes the full electromagnetic Casimir
1051: energy difference between eccentric and concentric configurations.
1052: Fig. 7 depicts the ratio of the exact Casimir energy difference
1053: $\Delta E$ and the PFA limit for the almost concentric cylinders
1054: configuration. As evident from the figure, PFA agrees with the
1055: exact result at a few percent level only for $\alpha$ close to
1056: unity, and then it noticeably departs from the PFA prediction. The
1057: resulting PFA expression for the attractive Casimir force between
1058: quasi-concentric cylinders reads
1059: \begin{equation}
1060: F^{\rm PFA}= \frac{\pi^3}{60}\frac{\epsilon L}{a^4(\alpha -1)^5},
1061: \end{equation}
1062: that reproduces the result previously obtained in \cite{Dalvit2004}.
1063:
1064:
1065:
1066: \section{Conclusions}
1067:
1068: We have derived an exact formula for the Casimir interaction energy
1069: between eccentric cylinders using a mode summation technique. This formula
1070: is written as an integral of the determinant of an infinite dimensional matrix, and it
1071: reproduces as a particular case the interaction energy between concentric
1072: cylinders, and as a limiting case the energy in the cylinder-plane
1073: geometry. In the quasi-concentric case, the infinite dimensional
1074: matrix becomes tridiagonal, and hence much easier to deal with
1075: than the exact formula when performing analytic and numerical calculations.
1076: We have carried out the numerical evaluation of the Casimir interaction
1077: energy using both the exact and tridiagonal formulas, and studied different
1078: limiting cases of relevance for Casimir force measurements.
1079:
1080: The large and small distance limits were analyzed. In the former case,
1081: the Casimir energy is dominated by the lowest modes, and shows a weak
1082: logarithmic decay, typical of cylindrical geometries. In the latter case,
1083: the Casimir energy is dominated by the highest modes, and the exact formula
1084: reproduces the proximity approximation. We found that the first order correction
1085: ($\alpha -1 \ll 1$) to PFA for the quasi-concentric cylinders has the form
1086: $\Delta E / \Delta E_{\rm PFA} = 1 + s (\alpha-1) + O((\alpha-1)^2)$, where the coefficient
1087: of the linear curvature correction is positive, $s>0$, both for TE and TM modes.
1088: This contrasts with the first order corrections to PFA in the cylinder-plane configuration,
1089: where the linear curvature correction to TM modes is positive, while the one for TE
1090: modes is negative \cite{Bordag2006}.
1091:
1092: The exact Casimir force computed in this paper, in particular for the
1093: quasi-concentric configuration, offers a qualitatively different
1094: approach for implementing new experiments to measure the Casimir force and to
1095: search for extra-gravitational forces in the micrometer and nanometer scales,
1096: since it opens the possibility of measuring the derivative of the force using (Cavendish-like) null experiments.
1097:
1098: \ack
1099: We are grateful to R. Onofrio and J. Von Stecher for fruitful
1100: discussions. We thank A. L\'opez D\'avalos for pointing
1101: Ref.\cite{Balseiro1950} to us. The work of F.C.L. and F.D.M. is
1102: supported by UBA, Conicet and ANPCyT (Argentina).
1103:
1104: \section*{References}
1105: \begin{thebibliography}{[99]}
1106:
1107: \bibitem{Casimir1948}
1108: H.B.G. Casimir, Proc. K. Ned. Akad. Wet. B {\bf 51}, 793 (1948).
1109:
1110: \bibitem{reviews}
1111: G. Plunien, B. M\"uller, and W. Greiner, Phys. Rep. \textbf{134},
1112: 87 (1986); P. Milonni, {\it The Quantum Vacuum} (Academic Press,
1113: San Diego, 1994); V. M. Mostepanenko and N. N. Trunov, {\it The
1114: Casimir Effect and its Applications} (Clarendon, London, 1997); M.
1115: Bordag, {\it The Casimir Effect 50 Years Later} (World Scientific, Singapore, 1999);
1116: M. Bordag, U. Mohideen, and V. M. Mostepanenko, Phys. Rep.
1117: \textbf{353}, 1 (2001); K. A. Milton, {\it The Casimir Effect:
1118: Physical Manifestations of the Zero-Point Energy} (World
1119: Scientific, Singapore, 2001); S. Reynaud {\it et al.}, C. R. Acad.
1120: Sci. Paris \textbf{IV-2}, 1287 (2001); K. A. Milton, J. Phys. A:
1121: Math. Gen. \textbf{37}, R209 (2004); S.K. Lamoreaux, Rep. Prog.
1122: Phys. \textbf{68}, 201 (2005).
1123:
1124: \bibitem{Derjaguin1957}
1125: B.V. Derjaguin and I.I. Ibrikosova, Sov. Phys. JETTP {\bf 3}, 819 (1957);
1126: B.V. Derjaguin, Sci. Am. {\bf 203}, 47 (1960);
1127: J. Blocki, J. Randrup, W.J. Swiatecki, and C.F. Tsang, Ann. Phys. {\bf 105}, 427 (1977).
1128:
1129: \bibitem{exp}
1130: S.K. Lamoreaux, Phys. Rev. Lett. \textbf{78}, 5 (1997); U.
1131: Mohideen and A. Roy, Phys. Rev. Lett. \textbf{81}, 4549 (1998);
1132: B.W. Harris, F. Chen, and U. Mohideen, Phys. Rev. A \textbf{62},
1133: 052109 (2000); T. Ederth, Phys. Rev. A {\bf 62}, 062104 (2000);
1134: H.B. Chan, V.A. Aksyuk, R.N. Kleiman, D.J. Bishop,
1135: and F. Capasso, Science \textbf{291}, 1941 (2001); H. B. Chan,
1136: V.A. Aksyuk, R.N. Kleiman, D.J. Bishop, and F. Capasso, Phys. Rev.
1137: Lett. \textbf{87}, 211801 (2001); G. Bressi, G. Carugno, R.
1138: Onofrio, and G. Ruoso, Phys. Rev. Lett. \textbf{88}, 041804
1139: (2002); D. Iannuzzi, I. Gelfand, M. Lisanti, and F. Capasso, Proc.
1140: Nat. Ac. Sci. USA \textbf{101}, 4019 (2004); R.S. Decca, D. L\'opez,
1141: E. Fischbach, and D.E. Krause, Phys. Rev. Lett. \textbf{91},
1142: 050402 (2003); R.S. Decca {\it et al.}, Phys. Rev. Lett.
1143: \textbf{94}, 240401 (2005); R.S. Decca {\it et al.}, Annals of
1144: Physics \textbf{318}, 37 (2005).
1145:
1146: \bibitem{sem}M. Schaden and L. Spruch, Phys. Rev. A {\bf 58}, 935 (1998).
1147:
1148: \bibitem{Mazzitelli2003}
1149: F.D. Mazzitelli, M.J. S\'anchez, N.N. Scoccola, and J. von
1150: Stecher, Phys. Rev. A {\bf 67}, 013807 (2003).
1151:
1152: \bibitem{opt}
1153: R. L. Jaffe and A. Scardicchio, Phys. Rev. Lett. {\bf 92}, 070402 (2004)
1154:
1155: \bibitem{num}
1156: H. Gies, K. Langfeld, and L. Moyaerts, J. High Energy Phys. \textbf{6}, 18 (2003).
1157:
1158: \bibitem{Genet2003} C. Genet, A. Lambrecht, Paulo A. Maia Neto, and S. Reynaud,
1159: Europhys. Lett. {\bf 62}, 484 (2003); R.B. Rodrigues, Paulo A. Maia Neto, A. Lambrecht,
1160: and S. Reynaud, Phys. Rev. Lett. {\bf 96}, 100402 (2006).
1161:
1162: \bibitem{Bulgac2006}
1163: A. Bulgac, P. Magierski, and A. Wirzba, Phys. Rev. D {\bf 73}, 025007 (2006).
1164:
1165: \bibitem{Emig2006}
1166: T. Emig, R.J. Jaffe, M. Kardar, and A. Scardicchio, Phys. Rev. Lett. {\bf 96}, 080403 (2006).
1167:
1168: \bibitem{Bordag2006}
1169: M. Bordag, Phys. Rev. D {\bf 73}, 125018 (2006).
1170:
1171: \bibitem{Gies2006}
1172: H. Gies and K. Klingmuller, Phys. Rev. Lett. {\bf 96}, 220401 (2006).
1173:
1174: \bibitem{Dalvit2004}
1175: D.A.R. Dalvit, F.C. Lombardo, F.D. Mazzitelli, and R. Onofrio, Europhys. Lett. {\bf 68}, 517 (2004).
1176:
1177: \bibitem{Brown-Hayes2005}
1178: M. Brown-Hayes, D.A.R. Dalvit, F.D. Mazzitelli, W.J. Kim, and R. Onofrio,
1179: Phys. Rev. A {\bf 72}, 052102 (2005).
1180:
1181: \bibitem{Mazzitelliqftext}F.D. Mazzitelli, in {\it Quantum Field
1182: Theory Under the Influence of External Conditions}, K.A. Milton
1183: (editor), Rinton Press, Princeton (2004).
1184:
1185: \bibitem{Dalvit2006} D.A.R. Dalvit, F.C. Lombardo, F.D. Mazzitelli, and R. Onofrio,
1186: Phys. Rev. A {\bf 74}, 020101(R) (2006).
1187:
1188: \bibitem{Saharian2006} A.A. Saharian, ICTP Report IC/2000/14, arXiv:hep-th/0002239;
1189: A.A. Saharian and A.S. Tarloyan, arXiv:hep-th/0603144.
1190:
1191: \bibitem{Nesterenko1998} See, for example, V.V. Nesterenko and I.G. Pirozhenko,
1192: Phys. Rev D {\bf 57}, 1284 (1998).
1193:
1194: \bibitem{Singh1984}
1195: G.S. Singh and L.S. Kothari, J. Math. Phys. {\bf 25}, 810 (1984).
1196:
1197: \bibitem{Balseiro1950}
1198: J.A. Balseiro, Revista de la Union Matem\'atica Argentina, v. XIV, 118 (1950).
1199:
1200: \bibitem{Hine1971}
1201: M.J. Hine, J. Sound Vib. {\bf 15}, 295 (1971).
1202:
1203: \bibitem{foot} This is a particular case of Eq.(\ref{addition1}), valid for $\vert v\vert < \vert u\vert$.
1204:
1205: \end{thebibliography}
1206:
1207: \end{document}
1208: