1: \documentclass[12pt]{article}
2: \usepackage{a4}\usepackage{epsf}
3: \newcommand{\NO}{Nielsen-Olesen }
4: \newcommand{\uv}{ultraviolet }
5: \newcommand{\be}{\begin{equation}}
6: \newcommand{\ee}{\end{equation}}
7: \newcommand{\bea}{\begin{eqnarray}}
8: \newcommand{\eea}{\end{eqnarray}}
9: \newcommand{\la}{\lambda}
10: \newcommand{\nn}{\nonumber}
11: \newcommand{\pa}{\partial}
12: \newcommand{\Tr}{{\rm Tr~}}
13: \newcommand{\Ref}[1]{(\ref{#1})}
14: \newcommand{\E}{{\cal E}}
15: \renewcommand{\P}{{\cal P}}
16:
17: \begin{document}
18: \title{Fermionic Vacuum Energy from a Nielsen-Olesen Vortex}
19: \author{
20: {\sc M. Bordag}\thanks{e-mail: Michael.Bordag@itp.uni-leipzig.de} \\
21: {\small and}\\
22: I. Drozdov\thanks{e-mail: Igor.Drosdow@itp.uni-leipzig.de}\\
23: \small University of Leipzig, Institute for Theoretical Physics\\
24: \small Augustusplatz 10/11, 04109 Leipzig, Germany}
25: \maketitle
26: \begin{abstract}
27: We calculate the vacuum energy of a spinor field in the background of
28: a Nielsen-Olesen vortex. We use the method of representing the
29: vacuum energy in terms of the Jost function on the imaginary
30: momentum axis. Renormalization is carried out using the heat kernel
31: expansion and zeta functional regularization. With this method well
32: convergent sums and integrals emerge which allow for an efficient
33: numerical calculation of the vacuum energy in the given case where
34: the background is not known analytically but only numerically. The
35: vacuum energy is calculated for several choices of the parameters
36: and it turns out to give small corrections to the classical
37: energy.
38: \end{abstract}
39: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
40: \section{Introduction}\label{Sec1}
41: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
42: Quantum corrections to classical background configurations are a topic
43: of continuing interest. At present it is stimulated by the observation
44: made in lattice calculations that the field configurations responsible
45: for confinement are dominated by monopole or string like
46: configurations. Another motivation comes from the stability analysis
47: of Z-strings with respect to fermionic fluctuations.
48:
49: During the last years quantum corrections to string like
50: configurations have been investigated quite actively. In
51: \cite{Bordag:1998tg} for the background of a finite radius magnetic
52: flux tube in QED the vacuum energy of a spinor was calculated. There
53: the method of representing the vacuum energy in terms of the Jost
54: function of the related scattering problem taken on the imaginary
55: momentum axis was applied. This method has been developed earlier in
56: \cite{Bordag:1996fv} for spherically symmetric scalar background
57: fields. The specific example considered in \cite{Bordag:1998tg} was a
58: homogenous magnetic field inside the flux tube. This investigation was
59: extended to more complicated profiles of the magnetic field inside the
60: flux tube in \cite{Scandurra:2000wr,Drozdov:2002um}. In
61: \cite{Dunne:1998kw} a magnetic background was considered which depends
62: only on one spatial coordinate. In addition, this dependence is of a
63: form that the corresponding wave equation has an explicit solution so
64: that the vacuum energy could be calculated quite easily. A more
65: general approach to fermionic vacuum energy was taken in
66: \cite{Fry:2001jw} where general bounds on the fermionic determinant
67: were obtained. Another approach was taken in
68: \cite{Pasipoularides:2000gg} where several profiles of the magnetic
69: background were considered and compared with the derivative
70: expansion. In \cite{Pasipoularides:2003jt} the limit of a strong
71: magnetic field was investigated in more detail.
72:
73: There is still an interest in backgrounds of infinitely thin strings
74: which constitute a singular background. Typical for these
75: configurations is the need to apply the method of self adjoint
76: extensions. The object to consider in such examples is the vacuum
77: energy density per unit volume rather than the 'global' vacuum energy
78: which is for a string the density per unit length. For recent
79: investigations see, for example, \cite{Langfeld:2002vy} and
80: \cite{Diakonov:1999gg}. The problem with infinitely thin strings is
81: that their classical energy is infinite. Also, there are additional
82: counter terms. In the language of heat kernel expansion there are
83: additional contributions to the heat kernel coefficients, for instance
84: coefficients with half integer number, which reside on the surface
85: where the singularity is located. Another related example was
86: considered in \cite{Scandurra:2000wr} where the background is a finite
87: radius flux tube with the magnetic field concentrated on the surface
88: of the string. The advantage of singular (and non smooth like in
89: \cite{Bordag:1998tg}) backgrounds can be seen in the usually quite
90: explicite formulas for the quantum fluctuations. So in
91: \cite{Langfeld:2002vy,Diakonov:1999gg,Scandurra:2000wr} only Bessel
92: functions appear and in \cite{Bordag:1998tg,Drozdov:2002um} hyper
93: geometric functions in addition.
94:
95: In general, physical background configurations should have a finite
96: energy, hence strings should have a non zero radius. A typical example
97: is the \NO string \cite{Nielsen:1973cs}. But here, not only the vacuum
98: fluctuations have to be calculated numerically, but even the
99: background itself. The problem appears to have a calculational scheme
100: which does not need explicite formulas and which allows for efficient
101: numerical evaluation. There the main problem comes from the \uv
102: divergencies. In analytical terms it is well known how to handle
103: them. First one has to introduce some intermediate regularization.
104: After that one has to subtract the counter terms and finally to remove
105: the regularization resulting now in a finite result. However,
106: consider the last step in zeta functional regularization. Here one has to
107: perform an analytical continuation. Or let us consider some cut-off
108: regularization, where one has to remove the contributions proportional to
109: non negative powers of the cut-off parameter. In principle such a
110: procedure can be done numerically (there are some examples) but this
111: is quite complicated and ineffective. It is better to transform the
112: expression for the vacuum energy in a way that the final removal of
113: the regularization can be performed analytically so that only well
114: convergent sums and integrals remain. Such a method had been developed
115: in \cite{Bordag:1995jz,Bordag:1996fv} and in
116: \cite{Bordag:1998tg,Bordag:2002sa} applied to strings of finite
117: radius. The method is based on a representation of the regularized
118: vacuum energy in terms of the Jost function of the related scattering
119: problem taken on the imaginary momentum axis. Another method, using
120: phase shifts and momenta on the real axis was used in
121: \cite{Graham:2002fi} (see also \cite{Graham:2002xq} and references
122: therein), mainly for spherically symmetric backgrounds. Also, using
123: phase shifts, the case of a color magnetic vortex was considered in
124: \cite{Diakonov:2002bx}. In a similar way in \cite{Groves:1999ks} the
125: vacuum energy for an electroweak string had been considered, where,
126: however, a step profile was taken in the final stage.
127:
128: A completely different method is worth to be mentioned. In
129: \cite{Gies:2001tj,Gies:2001zp} world line methods were applied to the
130: calculation of vacuum energy which have the advantage not to rely on
131: separation of variables and therefore to be applicable for much more
132: general background configurations.
133:
134: In the present paper we calculate the vacuum energy of a fermion in
135: the background of a \NO string. We use the method of representing the
136: regularized vacuum energy in terms of the Jost function on the
137: imaginary momentum axis which was developed in
138: \cite{Bordag:1996fv}. For the \NO vortex the background potential is
139: given only as a numerical solution of the corresponding equations of
140: motion. We use zeta functional regularization and determine the
141: counter terms from standard heat kernel expansion. The renormalized
142: vacuum energy is divided into two parts, the 'finite' and the
143: asymptotic ones, by subtraction and addition of some first terms of
144: the uniform asymptotic expansion of the Jost function which is
145: obtained using the Lipmann-Schwinger equation. In the asymptotic part
146: the analytic continuation in the regularization parameter is performed
147: analytically and well convergent double integrals remain. The 'finite'
148: part is represented as a well convergent sum and integral involving
149: the Jost function which is calculated from the numerical solutions of
150: the corresponding wave equation. Using these tools, the dependence on
151: the parameters of the considered model is investigated numerically.
152:
153: The paper is organized as follows. In the next section the basic
154: notations for the considered model are introduced, the renormalization
155: is discussed and the basic formulas for the representation of the
156: vacuum energy are given. In the third section the asymptotic part of
157: the Jost function is derived and the asymptotic part of the vacuum
158: energy as well. In the fourth section the 'finite' part of the vacuum
159: energy is derived and the convergence properties are
160: discussed. Sect. 5 contains the numerical part of the work and Sect. 6 the
161: conclusions.
162:
163: \noindent
164: We use units where $\hbar=c=1$.
165:
166:
167: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
168: \section{Basic notations}\label{Sec2}
169: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
170: The Abelian Higgs model contains a $U(1)$ gauge field, $A_\mu(x)$, and
171: a complex scalar field, $\Phi(x)$. The action is
172: %
173: \be\label{AH}S=\int d^4x \ \left(
174: -\frac14 F_{\mu\nu}^2
175: + \mid D_\mu \Phi |^2-\la\left(|\Phi|^2-\frac{\eta^2}{2} \right)^2\right),
176: \ee
177: %
178: where $D_\mu=\pa_\mu-i q A_\mu$ is the covariant derivative and
179: $F_{\mu\nu}=\pa_\mu A_\nu-\pa_\nu A_\mu$ is the field strength, for
180: more details see \cite{Achucarro:1999it}. The vacuum configuration is
181: given by $A_\mu=0$, $\Phi=\eta e^{ic}/\sqrt{2}$ where $c$ is some
182: constant and $\eta$ is the Higgs condensate. This configuration has
183: zero energy. A configuration with finite non zero energy must be at
184: spatial infinity in the vacuum manifold. Hence asymptotically the
185: gauge potential must be a pure gauge and the scalar field must tend to
186: a constant times an angular dependent phase. A configuration of such
187: type is the Nielsen-Olesen string \cite{Nielsen:1973cs}. In
188: cylindrical coordinates $(x,y,z)\to(r,\varphi,z)$ one makes the ansatz
189: %
190: \be\label{ansatz}\Phi=\frac{\eta}{\sqrt{2}} f(r) e^{i n \varphi}, \ \
191: \ q A_\varphi = n v(r),
192: \ee
193: %
194: where $A_\varphi$ is the angular component of the vector
195: potential. The profile functions satisfy the boundary conditions
196: %
197: \be\label{bc} f(0)=v(0)=0, \ \ \ \ \ \ f(r)\to 1, v(r)\to 1 \ \ {\rm
198: as} \ \ r\to\infty.
199: \ee
200: %
201: Here $n$ is the winding number and gives at once the magnetic flux in
202: the string. In the following we consider $n=1$ only. The equations of
203: motion imply
204: %
205: \bea\label{eom}f''(r)+\frac1r
206: f'(r)-\frac{1}{r^2}\left(1-v(r)\right)^2f(r)+
207: \la\eta^2\left(1-f(r)^2\right)f(r)&=& 0 ,
208: \nn \\
209: v''(r)-\frac1r v'(r)+q^2 \eta^2 f(r)^2\left(1-v(r)\right)&=& 0 .
210: \eea
211: %
212: After a rescaling, $r=\rho/(q\eta)$, these equations depend in fact
213: only on the combination $\beta=2\la/q^2$, which is at once the squared
214: ratio of the Higgs and vector masses. For $\beta=1$ the system
215: exhibits an additional symmetry and can be reduced to two first order
216: equations (Bogomolny equations). For $\beta<1$ the system is stable
217: for all values of the flux $n=1,2,\dots$, for $\beta>1$ only $n=1$ is
218: stable (a vortex with $n>1$ decays into $n$ vortexes with $n=1$).
219:
220: The system of equations \Ref{eom} together with the boundary
221: conditions \Ref{bc} does not have an analytical solution and one is
222: left with numerical methods. The simplest way is to set the
223: derivatives in zero, i.e. the constants $a$ and $b$ the expansion for
224: small $r$, $f(r)=a r +\dots$ and $v(r)=b r^2 +\dots$, and to numerically
225: integrate the equations from $r=0$ to larger values of $r$. One has to
226: adjust these constants in a way that for large $r$ the asymptotic
227: values $f=1$ and $v=1$ are approached. Examples are shown in Fig. 1
228: for two values of $\beta$ with $q\eta=1$. The corresponding values
229: of the derivatives are $(a,b)=(0.26817,0.17481)$ for $\beta=0.3$ and
230: $(a,b)=(0.79958,0.42848)$ for $\beta=6$. As the asymptotic values for
231: $r\to\infty$ are reached with exponential speed it is in fact
232: sufficient to consider $r$ up to $\approx 15$.
233: %
234: \begin{figure}[t] \label{fig1}
235: \epsfxsize=14cm
236: \epsffile{fig1.eps}
237: % \end{picture}
238: \caption{The profile functions of the Nielsen-Olesen vortex as
239: functions of the radius $r$ for $q\eta=1$.}
240: \end{figure}
241: %
242: sufficient to consider $r$ up to $\approx 15$.
243:
244: The classical energy (more exactly, the energy density per unit
245: length) of these configurations is given by
246: %
247: \bea\label{eclass}E_{\rm class}&=&\pi\int_0^\infty dr \ r \
248: \left( \frac{1}{q^2 r^2}v'(\rho)^2
249: +\eta^2 \left(
250: f'(\rho)^2
251: +\frac{1}{r^2}(1-v(\rho))^2 f(\rho)^2\right)
252: \right. \nn \\ && \left. \qquad \qquad \qquad
253: +\frac{\la}{2} \eta^4 (1-f(\rho)^2)^2\right).
254: \eea
255: %
256: We calculate the background for $q\eta=1$ and restore later the
257: general setting by inserting $\rho=q\eta r$ in \Ref{eclass}.
258:
259: The spinor is taken as a four component Dirac spinor with a
260: coupling to the background given by the Lagrange density
261: %
262: \be\label{Lspinor}L=-i \overline{\Psi} \left( i\gamma^\mu D_\mu-f_e\mid
263: \Phi \mid \right) \Psi .
264: \ee
265: %
266: This model is chosen for the reason of simplicity. It provides a
267: coupling of the spinor not only to the vector but also to the scalar
268: background and it is motivated by the Yukawa coupling in the fermionic
269: sector of the standard model. We note that the interaction in
270: \Ref{Lspinor} is gauge invariant and that the coupling constant $f_e$
271: is dimensionless. However, this model has a drawback. Since there is
272: only one spinor we are forced to take the absolute value,
273: $\mid\Phi\mid=\sqrt{(\Re \Phi)^2 +(\Im \Phi)^2}$, of the complex
274: scalar field (in the standard model there we don't need to do so). As a
275: consequence, the interaction is non polynomial. This is a
276: complication, for example if considering the vacuum energy together
277: with the dynamics of the background. Also, there is a need for an
278: additional counter term (see below). However, because we are not
279: going to consider the dynamics of the background, this complication
280: does not affect the calculation of the vacuum energy if we let enter
281: the additional counter terms into the classical energy with a coefficient
282: which is put equal to zero after the renormalization is carried out.
283: In addition, if we consider the same problem of calculating the vacuum
284: energy in the standard model then there is no such complication due to
285: its renormalizability.
286:
287: The background is static and due to the translational invariance in
288: direction along the z-axis and the rotational invariance around the z-axis
289: ($m=0,\pm1,\dots$ is the orbital angular momentum), the corresponding
290: momenta can be separated, $\Psi\to e^{-ip_0x^0+ip_3x^3}\left(\Psi_1
291: e^{-(m+1)\varphi}\atop \Psi_2 \ \ e^{-m\varphi}\right)$. After that
292: the Dirac equation decouples and one of the resulting pair of two
293: component equations may be represented as
294: %
295: \be\label{spinoreq}\left(\begin{array}{ccc}
296: p_0-\mu(r) & , & \frac{\pa}{\pa r}-\frac{m-v(r)}{r} \\
297: -\frac{\pa}{\pa r}-\frac{m+1-v(r)}{r}& , &p_0+\mu(r)\end{array}\right)
298: \left(\begin{array}{c}\Psi_1\\ \Psi_2 \end{array}\right)=0.
299: \ee
300: %
301: The other pair is obtained by changing the sign of $\mu$. Here
302: the notation $\mu(r)=f_e\frac{\eta}{\sqrt{2}} f(r)$ has been
303: introduced. With respect to the spinor this is a radius dependent mass
304: density. From its value at infinity we define the spinor mass,
305: $m_e=f_e\frac{\eta}{\sqrt{2}}$. For a constant mass density the
306: problem is the same as for a pure magnetic flux tube which appears as
307: a special case in this way.
308:
309: The vacuum energy of the spinor is given by the general formula
310: %
311: \be\label{Equant}\E_{\rm quant}=-\frac12 \sum_{(n)} e_{(n)}^{1-2s}.
312: \ee
313: %
314: Here $s$ is the zeta functional regularization parameter and $e_{(n)}$
315: are the one particle energies. For simplicity we dropped the
316: arbitrary constant $\mu$ which is usually introduced to adjust the
317: dimension in zeta functional regularization. The quantum numbers $(n)$
318: include the sign of the one particle energies (all enter with positive
319: sign), the spin $s_z$ (two projections which we can account for by
320: summing over the sign of $\mu(r)$), the orbital angular quantum number
321: $m$ and the radial quantum number $n_r$ (assuming for a moment the
322: system being inserted into a large cylinder). In addition there is the
323: momentum $p_3$ which can be integrated according to
324: %
325: \[\label{p3}\int_{-\infty}^\infty \frac{dp_3}{2\pi} \
326: \left(p_3^2+x^2\right)^{\frac12-s}
327: =
328: x^{2(1-s)}\frac{\Gamma(s-1)}{2\sqrt{\pi}\Gamma(s-\frac12)}
329: =
330: x^{2(1-s)}\frac{C_s}{4\pi s}
331: \]
332: %
333: with $C_s=1+s(2\ln 2-1)+\dots$ and we arrive at
334: %
335: \be\label{Equant1}\E_{\rm quant}=-\frac{C_s}{4\pi s}
336: \sum_{s_z, m, n_r} \left(e_{s_z, m, n_r}\right)^{2(1-s)}.
337: \ee
338: %
339: In order to get rid of the large cylinder we proceed as in
340: \cite{Bordag:1998tg}. We consider the cylindrical scattering problem
341: associated with Eq. \Ref{spinoreq} and define the Jost functions,
342: $f_m(k)$, with $p_0=\sqrt{m_e^2+k^2}$. We rewrite the sum over $n_r$
343: by an integral, tend the large cylinder to infinity and after dropping
344: the Minkowski space contribution we end up with
345: %
346: \be\label{Equant2}\E_{\rm quant}=\frac{C_s}{4\pi} \sum_{sgn
347: \mu}\sum_{m=-\infty}^\infty
348: \int_{m_e}^\infty dk \ \left(k^2-m_e^2\right)^{1-s} \ \frac{\pa}{\pa
349: k} \ln f_m(i k)
350: \ee
351: %
352: which is our expression for the regularized ground state energy. Here
353: the integration is turned to the imaginary axis, more specifically, it
354: goes along the cut resulting from $\left(k^2-m_e^2\right)^{1-s}$. In
355: order to fully exploit the symmetry we renumber the orbital momenta by
356: $\nu$ according to
357: %
358: \be\label{nu}m=\left\{\begin{array}{rr} \nu-\frac12 & (m\ge 0) \\[3pt]
359: -\nu-\frac12 & (m<0) \end{array}\right.
360: \ee
361: %
362: with $\nu=\frac12,\frac32,\dots$ and two signs of the orbital momentum
363: are to be taken into account. After rotation $k\to ik$ we define
364: $p=\sqrt{k^2-m_e^2}$ and for positive orbital momentum
365: Eq. \Ref{spinoreq} can be rewritten in the form
366: %
367: \be\label{speq}\left(\begin{array}{ccc}
368: ip-\mu(r) & , & \frac{\pa}{\pa r}-\frac{\nu-1/2-v(r)}{r} \\
369: -\frac{\pa}{\pa r}-\frac{\nu+1/2-v(r)}{r}& , &ip+\mu(r)\end{array}\right)
370: \left(\begin{array}{c}\Psi_1\\ \Psi_2 \end{array}\right)=0.
371: \ee
372: %
373: It can be seen (see below, Eqs. \Ref{eqg+} and \Ref{eqg-}) that for
374: the Jost function a change in the sign of $\mu$ corresponds to complex
375: conjugation and a change in the sign of the magnetic background,
376: $v(r)\to-v(r)$, corresponds to an exchange of positive and negative
377: orbital momenta. As a consequence, in Eq. \Ref{Equant2} we have a
378: factor of 4 and we take the real part of the half sum of positive and
379: negative orbital momenta.
380:
381: The next step is the renormalization. We use the standard heat kernel
382: expansion according to which the divergent part of the ground state
383: energy is given by (see, e.g., \cite{Bordag:1998tg} and we drop $a_0$)
384: %
385: \be\label{Ediv}E^{\rm div}=
386: \frac{m_e^2}{32\pi^2}\left(\frac1s+\ln\frac{4}{m_e^2}-1\right) a_1
387: -\frac{1}{32\pi^2}\left(\frac1s+\ln\frac{4}{m_e^2}-2\right)a_2 ,
388: \ee
389: %
390: where $a_i$ are the standard heat kernel coefficients. We define the
391: renormalized ground state energy by $s\to0$
392: %
393: \be\label{Eren} E_{\rm ren}=\E_{\rm quant}- E^{\rm div}
394: \ee
395: %
396: in the limit $s\to0$. The definition \Ref{Eren} of the
397: renormalized ground state energy is chosen in a way that $E_{\rm ren}$
398: vanishes if the spinor mass $m_e$ taken as a parameter independent
399: from the background becomes large. This is the so called 'large mass'
400: normalization condition. It has been discussed in detail in
401: \cite{Bordag:2000f,Bordag:1999vs} and in \cite{BGNV} it was shown to
402: be equivalent to the known 'no tadpole' condition.
403:
404: The heat kernel
405: coefficients can be calculated using known methods from the squared
406: Dirac operator,
407: %
408: \bea\label{Dsquared}
409: \left(iD\hspace{-8pt}/+\mu(r)\right)\left(iD\hspace{-8pt}/-\mu(r)\right)&=&
410: -D^2+\frac{q}{2}F_{\mu\nu}\sigma^{\mu\nu}
411: -if_e \gamma^\mu \frac{\pa}{\pa x^\mu}\mid\Phi\mid
412: -f_e^2\mid\Phi\mid^2 \nn \\
413: &\equiv&-D^2+V
414: \eea
415: %
416: with $\sigma^{\mu\nu}=\frac{i}{2}[\gamma^\mu,\gamma^\nu]$.
417: The general expressions for the relevant heat kernel coefficients are
418: %
419: \bea\label{allghkks}a_1&=&\Tr \int d^2x \ (-V), \nn \\
420: a_2&=&\Tr \int d^2x \left(
421: -\frac{1}{12}F_{\mu\nu}^2+\frac12 V^2-\frac16 \Delta V \right),
422: \eea
423: %
424: where the trace is over the gamma matrices. Inserting for $V$ the
425: result is \bea\label{hkks}a_1 &=& -8\pi\int\limits_0^\infty dr \ r \
426: \left(\mu(r)^2-m_e^2\right), \nn \\ a_2 &=& 8\pi\int\limits_0^\infty
427: dr \ r \ \left(\frac13 \frac{v'(r)^2}{r^2} + \frac12 \left(
428: \mu'(r)^2+(\mu(r)^2-m_e^2)^2\right) \right) \eea
429: %
430: (note that we consider densities per unit length of the string).
431: From Eq. \Ref{Ediv} we obtain the divergent part of the vacuum energy
432: in the form
433: %
434: \bea\label{Ediv2}\E^{\rm div}&=&-\frac{1-2s(\ln(m_e/2)+1)}{4\pi s}
435: \int\limits_0^\infty dr \ r \
436: \left(\frac13
437: \frac{v'(r)^2}{r^2}+\frac{f_e^2\eta^2}{4}f'(r)^2 \right. \nn \\
438: && \left. +\frac{f_e^4\eta^4}{8}\left(f(r)^4-1\right)\right)
439: -\frac{1}{4\pi}\int\limits_0^\infty dr \ r \ \frac{f_e^4\eta^4}{4}
440: \left(f(r)^2-1\right) +O(s).
441: \eea
442: %
443: The interpretation is as follows. From the $v'(r)^2$-term we have the
444: standard renormalization of the electric charge $q$ in the classical
445: action \Ref{eclass}. A renormalization of the scalar coupling
446: $\lambda$ absorbs the divergence proportional to $f(r)^4$. After that
447: the remaining freedom is in a change of the condensate $\eta$.
448: However, this is obviously insufficient to absorb the remaining parts
449: of $\E^{\rm div}$. In this way, in the classical energy an additional
450: structure must be present. As suggested from Eq. \Ref{Dsquared}, it
451: is necessary to introduce at last a term proportional to
452: $\left(\frac{\pa}{\pa x_\mu}\mid\Phi \mid\right)^2$ into the action
453: \Ref{AH} and into the classical energy \Ref{eclass} as well. It
454: should be noted that such a term is gauge invariant and that it has
455: the correct dimension. However, it represents another non polynomial
456: interaction. This is not surprising as the model given by
457: Eq. \Ref{Lspinor} itself contains a non polynomial interaction.
458: Accepting this we can finish the renormalization if we put the
459: coefficient in front of this term equal to zero after performing the
460: renormalization.
461:
462: In order to perform the limit $s\to0$ in Eq. \Ref{Eren} we need to
463: rewrite the regularized ground state energy. For this we define the
464: asymptotic Jost function as the part of its uniform asymptotic
465: expansion for $\nu\to\infty$, $z\equiv\frac{k}{\nu}$ fixed, which
466: includes all powers up to $\nu^{-3}$,
467: %
468: \be\label{lnfas} \ln f_\nu(ik)= \ln f^{\rm as}_\nu
469: +O\left(\frac{1}{\nu^4}\right).
470: \ee
471: %
472: Using $\ln f^{\rm as}_\nu$ we divide the energy into
473: %
474: \be\label{Eren2}E_{\rm ren}=\E_{\rm f}+ E^{\rm as}
475: \ee
476: %
477: with the 'finite' part,
478: %
479: \be\label{Ef}\E^{\rm f}= \frac{1}{\pi}\sum_{\nu=\frac12,\frac32,\dots}
480: \int_{m_e}^\infty dk \ \left(k^2-m_e^2\right) \ \frac{\pa}{\pa
481: k} \left( \ln f_\nu(i k)-\ln f^{\rm as}_\nu(i k) \right)
482: \ee
483: %
484: and the 'asymptotic' part,
485: %
486: \be\label{Eas}E^{\rm as}=\frac{1}{\pi}\sum_{\nu=\frac12,\frac32,\dots}
487: \int_{m_e}^\infty dk \ \left(k^2-m_e^2\right)^{2(1-s)} \ \frac{\pa}{\pa
488: k} \ln f^{\rm as}_\nu(i k) \ - \ E^{\rm div}.
489: \ee
490: %
491: The sum and the integral in $\E^{\rm f}$, Eq. \Ref{Ef}, are finite by
492: construction of the asymptotic Jost function. Therefore we could put
493: $s=0$ therein. In the asymptotic part, $E^{\rm as}$, Eq. \Ref{Eas}, the
494: analytic continuation in $s$ has to be still performed which will be
495: done in the next section.
496:
497: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
498: \section{The asymptotic part of the vacuum energy}\label{Sec3}
499: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
500: The asymptotic part of the Jost function can be derived from the
501: Lipmann-Schwinger equation just generalizing the procedure developed
502: in \cite{Bordag:1998tg}. The operator $\Delta\P$ there in Eq. (28)
503: reads now
504: %
505: \be\label{dP}\Delta\P=\left(
506: \begin{array}{ccc}\mu(r)&,&-\frac{v(r)}{r}\\-\frac{v(r)}{r}&,&-\mu(r)
507: \end{array}
508: \right).
509: \ee
510: %
511: Again, we have to perform iterations up to the fourth order in the
512: operator $\Delta\P$. Using the formulas given in
513: \cite{Bordag:1998tg} one arrives at the representation
514: %
515: \be\label{lnfas1}\ln f_{\nu}^{\rm as}(ik)=\int\limits_0^\infty
516: \frac{dr}{r}\sum_{n=1}^3\sum_{j=n}^{3n}X_{nj}\frac{t^j}{\nu^n}
517: \ee
518: %
519: with $t=\frac{1}{\sqrt{1+(\nu z)^2}}$. The coefficents are
520: %
521: \bea\label{Xnj}X_{11}&=&\frac12 v(r)^2+\frac12 r^2(\mu(r)^2-m_e^2),\nn \\
522: X_{13}&=&-\frac12 v(r)^2,\nn \\
523: X_{33}&=&\frac14 v(r)^2-\frac18 r^2 v'(r)^2-\frac18 v(r)^4-\frac14
524: r^2 v(r)^2(\mu^2-m_e^2)\nn \\
525: &&-\frac18 r^4 (\mu'(r)^2+(\mu^2-m_e^2)^2),\nn \\
526: X_{35}&=&-\frac{39}{16} v(r)^2+\frac18 r^2 v'(r)^2+\frac{3}{4} v(r)^4
527: +\frac34 r^2 v(r)^2(\mu^2-m_e^2)-\frac{3}{16}r^2 (\mu^2-m_e^2) ,\nn \\
528: X_{37}&=&\frac{35}{8}v(r)^2-\frac{5}{8} v(r)^4+\frac{5}{16}r^2
529: (\mu^2-m_e^2),\nn \\
530: X_{39}&=&-\frac{35}{16}v(r)^2.
531: \eea
532: %
533: The dependence on $v(r)$ is the same as in \cite{Bordag:1998tg} (where
534: it was denoted by $a(r)$), the dependence on $\mu(r)$ is new. In
535: Eq. \Ref{lnfas1} it had been integrated by parts in order to get the
536: shortest representation for the coefficients $X_{nj}$.
537:
538: In $\E^{\rm as}$, Eq. \Ref{Eas}, the analytic continuation in $s$ can
539: be performed by rewriting the sum over the orbital momenta $\nu$ by
540: integrals using the Abel-Plana formula in the form as given in the
541: Appendix, Eq. \Ref{APla}. From the first part, i.e., from the direct
542: integral over $\nu$, we get a contribution which just cancels $E^{\rm
543: div}$ in Eq. \Ref{Eas}. So we are left with the contribution from the
544: second part. Here we integrate over $k$ using the simple formula
545: \Ref{C3} and obtain
546: %
547: \be\label{Eas2}\E^{\rm as}=-\frac{C_s}{\pi} m_e^{2(1-s)}
548: \int_0^\infty\frac{dr}{r}\sum_{n,j} X_{nj}
549: \frac{\Gamma(2-s)\Gamma(s+\frac{j}{2}-1)}{\Gamma(j/2)} \ \Sigma_{nj}(rm_e)
550: \ee
551: %
552: with
553: %
554: \be\label{Signj}\Sigma_{nj}(x)=
555: \frac{1}{x^j}\int\limits_0^\infty\frac{d\nu}{1+e^{2\pi\nu}}
556: \frac1i \left(
557: \frac{(i\nu)^{j-n}}{\left(1+\left(i\nu\over x\right)^2\right)^{s+j/2-1}}-
558: \frac{(-i\nu)^{j-n}}{\left(1+\left(-i\nu\over x\right)^2\right)^{s+j/2-1}}
559: \right).
560: \ee
561: %
562: The difference in the right hand side results from the deformation of
563: the integration contour in the Abel-Plana formula which gets tight
564: around the cut. Therefore the integration starts effectively from
565: $\nu=x$.
566: This formula provides the best representation to perform the analytic
567: continuation in $s$. For $(n,j)=(1,1)$ we simply note
568: %
569: \be\label{Sig11}\Sigma_{11}(x)=\frac{2}{x^2}\int_x^\infty d\nu \
570: \frac{1}{1+e^{2\pi \nu}} \sqrt{\nu^2-x^2}.
571: \ee
572: %
573: Also for $(n,j)=(1,3)$ we can put $s=0$ immediately. However, we
574: prefer to integrate by parts in order to get a representation which is
575: in line with Eq. \Ref{Sig11}. In a similar way, integrating by parts,
576: we can proceed in all other contributions. So the final form is
577: %
578: \be\label{Sigall}\Sigma_{nj}(x)=\frac{2}{x^2}\int_x^\infty d\nu \ f_{nj}(\nu)\
579: \sqrt{\nu^2-x^2}
580: \ee
581: %
582: with
583: %
584: \[\label{fnj}
585: \begin{array}{rclrcl}
586: f_{11}(\nu)&=&\frac{1}{1+e^{2\pi \nu}} , &
587: f_{13}(\nu)&=&-\left(\frac{\nu}{1+e^{2\pi \nu}}\right)' , \\[4pt]
588: f_{33}(\nu)&=&\left(\frac{1}{\nu\left(1+e^{2\pi \nu}\right)}\right)',
589: &
590: f_{35}(\nu)&=&
591: \left(\frac{1}{\nu}\left(\frac{\nu}{1+e^{2\pi \nu}}\right)'\right)' , \\[4pt]
592: f_{37}(\nu)&=&\frac13
593: \left(\frac{1}{\nu}\left(\frac{1}{\nu}\left(\frac{\nu^3}{1+e^{2\pi
594: \nu}}\right)'\right)'\right)' , &
595: f_{39}(\nu)&=&\frac1{15}
596: \left(\frac{1}{\nu}\left(\frac{1}{\nu}\left(\frac{1}{\nu}\left(\frac{\nu^5}{1+e^{2\pi
597: \nu}}\right)'\right)'\right)'\right)'.
598: \end{array}
599: \]
600: %
601: Using these formulas in Eq. \Ref{Eas2} we can put $s=0$ there and
602: insert for the coefficients $X_{nj}$ using Eqs.\Ref{Xnj}. After
603: rearranging contributions we arrive at
604: %
605: \bea\label{Eas3}\E^{\rm as}&=&\frac{-2}{\pi}\int_0^\infty\frac{dr}{r^3}
606: \left(v(r)^2 h_1(r m_e)+r^2v'(r)^2 h_2(r m_e)+v(r)^4 h_3(r
607: m_e)\right. \nn \\ && \left. \qquad +
608: r^2 v(r)^2 (\mu(r)^2-m_e^2) h_4(r m_e)+r^2 (\mu(r)^2-m_e^2) h_5(r
609: m_e)\right. \nn \\ && \left. \qquad+
610: r^4 (\mu'(r)^2 +(\mu(r)^2-m_e^2)^2) h_6(r m_e) \right)
611: \eea
612: %
613: with
614: %
615: \be\label{hj}h_j(x)=\int_x^\infty d\nu \ f_j(\nu) \ \sqrt{\nu^2-x^2}
616: \ee
617: %
618: and
619: %
620: \bea\label{fj}f_1(\nu)&=&-f_{11}(\nu)-f_{13}(\nu)+\frac12 f_{33}(\nu)
621: -\frac{13}{8}f_{35}(\nu)+\frac{7}{4}f_{37}(\nu)
622: -\frac{5}{8}f_{39}(\nu) , \nn \\
623: f_2(\nu)&=&-\frac14 f_{33}(\nu)+\frac1{12} f_{35}(\nu), \nn \\
624: f_3(\nu)&=&-\frac14 f_{33}(\nu)+\frac12 f_{35}(\nu)-\frac14
625: f_{37}(\nu), \nn \\
626: f_4(\nu)&=&-\frac12 f_{33}(\nu)+\frac12 f_{35}(\nu), \nn \\
627: f_5(\nu)&=&- f_{11}(\nu)-\frac18 f_{35}(\nu)+\frac18 f_{37}(\nu), \nn \\
628: f_6(\nu)&=&-\frac14 f_{33}(\nu).
629: \eea
630: %
631: According to these functions we divide the asymptotic part of the
632: ground state energy,
633: %
634: \be\label{Eas4}\E^{\rm as}\equiv \sum_{j=1}^6 \E^{\rm as}_j ,
635: \ee
636: %
637: into parts which have the meaning, for instance, of the asymptotic
638: part of the vacuum energy resulting from the magnetic background,
639: $\E^{\rm as}_1$, $\E^{\rm as}_2$ and $\E^{\rm as}_3$, or from the
640: mixed contribution, $\E^{\rm as}_4$, etc. We finish this section with
641: the remark, that the integrals in Eq. \Ref{Eas3} are all finite for
642: the considered background of a \NO vortex due to corresponding
643: properties of the functions $h_i(x)$, their behavior at $x\to0$ for
644: instance.
645:
646: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
647: \section{The 'finite' part of the vacuum energy and numerical results}
648: \label{Sec4}
649: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
650: In order to calculate the finite part of the vacum energy defined by
651: Eq. \Ref{Ef} we have, first of all, to set up a numerical scheme for
652: the calculation of the Jost function. The Jost function is defined as
653: the coefficients in the asymptotics for large radius in the so called
654: regular solution of the wave equation \Ref{spinoreq},
655: %
656: \be\label{seq1}\Psi^{(\pm)}(r)\sim \frac12 \left\{
657: f_\nu^{(\pm)}(k) \Psi^{(\pm)}_{H^{(2)}}(kr)+
658: \overline{f}_\nu^{(\pm)}(k) \Psi^{(\pm)}_{H^{(1)}}(kr) \right\},
659: \ee
660: %
661: where
662: %
663: \be\label{seq1a}\Psi^{(\pm)}_{H^{(1,2)}}(kr)=\left(\begin{array}{c}
664: \sqrt{p_0+m_e} \ H^{(1,2)}_{\nu\pm(\frac12-\delta)}(kr) \\[6pt]
665: \pm\sqrt{p_0-m_e} \ H^{(1,2)}_{\nu\mp(\frac12+\delta)}(kr)
666: \end{array}\right)
667: \ee
668: %
669: ($p_0=\sqrt{m_e^2+k^2}$) are the two linear independent solutions at
670: $r\to\infty$. The signs $(\pm)$ correspond to the sign of the orbital
671: momentum according to Eq. \Ref{nu} and $H^{(1,2)}_\mu(z)$ are Hankel
672: functions. In general, $\delta$ is the dimensionless value of the
673: magnetic flux, i.e. it is $v(r\to\infty)$, which is equal to one in
674: our case.
675:
676: From Eq. \Ref{seq1}, the Jost function can be expressed in terms of the
677: solutions,
678: %
679: \be\label{jost1}f_\nu^{(\pm)}(k)=\frac{\pi r}{2i}\left(
680: \mp\sqrt{p_0-m_e} \psi_1(r) H^{(1)}_{\nu\mp(\frac12+\delta)}(kr)
681: +\sqrt{p_0+m_e} \psi_2(r) H^{(1)}_{\nu\pm(\frac12-\delta)}(kr) \right),
682: \ee
683: %
684: where $\psi_{1,2}(kr)$ are the upper and lower components of
685: $\Psi(r)$. Strictly speaking, as given by Eq. \Ref{jost1}, the
686: functions $f_\nu^{(\pm)}(k)$ depend on $r$ and only for $r\to\infty$
687: they tend to the Jost functions. However, for sufficiently large $r$
688: (larger than the scale of the background, say $r>15$ in the examples
689: shown in Fig. 1), Eq. \Ref{jost1} provides a good approximation. In
690: the considered case the background approaches its asymptotic values
691: with exponential speed. In the same way the difference between the
692: Jost function and the approximation \Ref{jost1} is small.
693:
694: The regular solutions $\Psi(r)$ of Eq. \Ref{spinoreq} are defined as
695: becoming proportional to the free solutions for $r\to0$ and can be
696: expressed in terms of Bessel function,
697: %
698: \be\label{regsol}\Psi(r)\sim
699: \left(\frac{k}{q}\right)^{\nu+\frac12}
700: \left(\begin{array}{c}
701: \sqrt{p_0+m_0} \ J_{\nu\pm\frac12}(qr) \\[6pt]
702: \pm \sqrt{p_0-m_0} \ J_{\nu\mp\frac12}(qr)
703: \end{array}\right)
704: \ee
705: %
706: with $q=\sqrt{k^2-m_e^2+m_0^2}$. We introduced for a moment $m_0=\mu(0)$
707: which corresponds to a more general case. In all examples considered
708: below we will have $m_0=0$. Note also the factor in front which
709: provides the correct normalization.
710:
711: In order to actually carry out a numerical integration of
712: Eq. \Ref{spinoreq} it is useful to change to functions with regular
713: initial values at $r=0$. At once, we must change to imaginary
714: momenta. So we substitute $k\to ik$ and we make the ansatz
715: %
716: \bea\label{ans1}\Psi^{(+)}=i^\nu \left(\begin{array}{c}
717: i\sqrt{p-im_0} \ \phi_1^{(+)} \\[6pt]
718: \sqrt{p+im_0} \ \phi_2^{(+)}
719: \end{array}\right) &,&
720: \Psi^{(-)}=i^\nu \left(\begin{array}{c}
721: \sqrt{p-im_0} \ \phi_1^{(+)} \\[6pt]
722: -i\sqrt{p+im_0} \ \phi_2^{(+)}
723: \end{array}\right).
724: \eea
725: %
726: The equations for $\phi^{(\pm)}$ are
727: %
728: \be\label{eq+}
729: \left(\begin{array}{ccc}
730: -(p+i\mu(r)) \ e^{i\theta} &,& \frac{\pa}{\pa r}-\frac{\nu-1/2-v(r)}{r}
731: \\
732: \frac{\pa}{\pa r}+\frac{\nu+1/2-v(r)}{r}&,&-(p-i\mu(r)) \ e^{-i\theta}
733: \end{array}\right)\phi^{(+)} =0
734: \ee
735: %
736: and
737: %
738: \be\label{eq-}
739: \left(\begin{array}{ccc}
740: -(p+i\mu(r)) \ e^{i\theta} &,& \frac{\pa}{\pa r}+\frac{\nu+1/2+v(r)}{r}
741: \\
742: \frac{\pa}{\pa r}-\frac{\nu-1/2+v(r)}{r}&,&-(p-i\mu(r)) \ e^{-i\theta}
743: \end{array}\right)\phi^{(-)} =0
744: \ee
745: %
746: with $e^{i\theta}=\sqrt{(p+im_e)/(p-im_0)}$.
747: Further we substitute
748: %
749: \be\label{ans+}\phi^{(+)}=\left(\frac{ q
750: r}{2}\right)^{\nu-\frac12}\frac{1}{\Gamma(\nu+\frac12)}
751: \left(\begin{array}{r} \frac{qr}{2\nu+1} \ g_1^{(+)} \\ g_2^{(+)}
752: \end{array}\right)
753: \ee
754: %
755: and
756: %
757: \be\label{ans-}\phi^{(-)}=\left(\frac{ q
758: r}{2}\right)^{\nu-\frac12}\frac{1}{\Gamma(\nu+\frac12)}
759: \left(\begin{array}{r} g_1^{(-)} \\ \frac{qr}{2\nu+1} \ g_2^{(-)}
760: \end{array}\right).
761: \ee
762: %
763: The boundary conditions at $r=0$ for these functions are
764: %
765: \be\label{bcg}g^{(\pm)}_{(1,2)}(r)_{\mid_{r=0}}=1
766: \ee
767: %
768: and they obey the equations
769: %
770: \be\label{eqg+}\left(\begin{array}{ccc}
771: -(p+i\mu(r)) \frac{qr}{2\nu+1} \ e^{i\theta} &,&
772: \frac{\pa}{\pa r}+\frac{v(r)}{r}
773: \\
774: \frac{\pa}{\pa r}+\frac{2\nu+1-v(r)}{r}&,&
775: -(p-i\mu(r)) \frac{2\nu+1}{qr} \ e^{-i\theta}
776: \end{array}\right)\left(\begin{array}{c}g_1^{(+)} \\
777: g_2^{(+)} \end{array}\right)=0,
778: \ee
779: %
780: \be\label{eqg-}\left(\begin{array}{ccc}
781: -(p+i\mu(r)) \frac{2\nu+1}{qr} \ e^{i\theta}&,&
782: \frac{\pa}{\pa r}+\frac{2\nu+1+v(r)}{r}
783: \\
784: \frac{\pa}{\pa r}-\frac{v(r)}{r}&,&
785: -(p-i\mu(r)) \frac{qr}{2\nu+1} \ e^{-i\theta}
786: \end{array}\right)\left(\begin{array}{c}g_1^{(-)} \\
787: g_2^{(-)} \end{array}\right)=0.
788: \ee
789: %
790: The above mentioned symmetries can be seen in this representation
791: explicitely. The Jost functions read now
792: %
793: \be\label{fg+}
794: f_\nu^{(+)}(ik)=\frac{r(q r/2)^{\nu-\frac12}}{\Gamma(\nu+\frac12)}
795: \left(\frac{qr}{2\nu+1}wg_1^{(+)}(r) K_{\nu-\frac12-\delta}(kr)
796: +w^{*} g_2^{(+)}(r) K_{\nu+\frac12-\delta}(kr)\right)
797: \ee
798: %
799: and
800: %
801: \be\label{fg-}f_\nu^{(-)}(ik)=\frac{r(q r/2)^{\nu-\frac12}}{\Gamma(\nu+\frac12)}
802: \left( wg_1^{(-)}(r) K_{\nu+\frac12+\delta}(kr)
803: + \frac{qr}{2\nu+1} w^{*} g_2^{(-)}(r) K_{\nu-\frac12+\delta}(kr)\right)
804: \ee
805: %
806: with $w=\sqrt{(p+im_e)(p-im_0)}$ and modified Bessel
807: functions. Eqs. \Ref{eqg+} and \Ref{eqg-} together with the boundary
808: conditions \Ref{bcg} can be solved easily numerically. We used the
809: package {\verb NDSolve } in \verb Mathematica .
810:
811: Next we need the asymptotic part of the Jost function. We use
812: Eq. \Ref{lnfas1} with the coefficients $X_{nj}$, Eq. \Ref{Xnj}. The
813: integrals over $r$ are convergent and can be calculated numerically
814: without any problems. Taking into account the mentioned symmetries we
815: drop all contributions which are odd under $v\to -v$ and we define the
816: subtracted logarithm of the Jost function as
817: %
818: \be\label{lnfsub}\ln f^{\rm sub}=
819: \frac12 \Re(\ln f^{(+)}_\nu(ik)+\ln f^{(-)}_\nu(ik))-\ln f_\nu^{\rm as}.
820: \ee
821: %
822: This expression we insert into $\E^{\rm f}$, Eq. \Ref{Ef}. After
823: integration by parts we arrive at
824: %
825: \be\label{Ef2}\E^{\rm f}=\sum_{\nu=\frac12,\frac32,\dots} \E^{\rm f}_\nu
826: \ee
827: %
828: with the contributions from the individual orbital momenta,
829: %
830: \be\label{Efnu}\E^{\rm f}_\nu= \int_{m_e}^\infty
831: dk \ \ \E^{\rm f}_\nu (k),
832: \ee
833: %
834: and the integrands
835: %
836: \be\label{Efnuk}\E^{\rm f}_\nu (k)= -\frac{2}{\pi}
837: k \ \ln f^{\rm sub}.
838: \ee
839: %
840: As illustration of the convergence we show $\E^{\rm f}_\nu (k)$
841: enhanced by a factor of $k^3 \nu^3$ in Fig. 2 for several first $\nu$.
842: It is seen that all curves approach the same limiting value which is
843: essentially given by the next term in the uniform asymptotic expansion
844: of the logarithm of the Jost function after that included into $\ln
845: f^{\rm as}$, Eq. \Ref{lnfas} taking into account that the order
846: $\frac{1}{\nu^4}$ is absent for symmetry reasons and the order
847: $\frac{1}{\nu^5}$ is the next one. In this way the integrals over $k$
848: are fast convergent.
849:
850: In Fig. 3 we show the contributions from the individual orbital
851: momenta. The corresponding sum in Eq. \Ref{Ef2} is also fast
852: convergent and it is sufficient to take a quite small number of the
853: first orbital momenta.
854: %
855: \begin{figure}[t] \label{fig2}
856: \epsfxsize=12cm
857: \epsffile{fig2.eps}
858: \caption{The integrand $\E^{\rm f}_\nu (k)$ in Eq. \Ref{Efnu}
859: multiplied by $k^3 \nu^3 $ as a function of $k$ for several first
860: values of $\nu$. In this figure and in the next two the parameters
861: are $\beta=0.3$, $q=0.5$, $f_e=1$, $\eta=1$.}
862: \end{figure}
863: %
864: %
865: \begin{figure}[t] \label{fig3}
866: \epsfxsize=12cm
867: \epsffile{fig3.eps}
868: \caption{The contributions from the individual orbital momenta
869: $\E^{\rm f}_\nu$ to $\E^{\rm f}$ in Eq. \Ref{Ef2}.}
870: \end{figure}
871:
872: We would like to add a note on bound states. They appear for imaginary
873: momenta, $k=i\kappa$, in Eq. \Ref{spinoreq}, in the region where
874: $p_0=\sqrt{m_e^2-\kappa^2}$ is real, $0<p_0<m_e$. For such momenta the
875: representation \Ref{jost1} for the Jost function can be rewritten
876: in the form
877: %
878: \bea\label{jostbs}f^{(\pm)}_\nu(i\kappa)&=&-\lim_{r\to\infty}r \
879: i^{-\nu\mp(\frac12-\delta)}
880: \left\{
881: \sqrt{-p_0+m_e} \Psi_1(r) K_{\nu\mp(\frac12+\delta)}(\kappa r)\nn
882: \right.\\ &&\left.
883: ~~~~~~~~~~~~~~~
884: +\sqrt{ p_0+m_e} \Psi_2(r) K_{\nu\pm(\frac12+\delta)}(\kappa r) \right\}.
885: \eea
886: %
887: The functions $\Psi_{1,2}$ are solutions of the Eqs. \Ref{spinoreq}
888: with the initial conditions \Ref{regsol}. As both, equation and
889: initial conditions are real for the considered momenta the solutions
890: are also real. In this way the expression in the figure brackets in
891: Eq. \Ref{jostbs} are real. Their zeros just determine the location of
892: the bound states. As an example we plot in Fig. 4 these functions for
893: the two lowest orbital momenta. For orbital momentum $m=0$ we see one
894: bound state, for negative, $m=-1$, none. The appearance of the bound
895: states can be explained easily. Without scalar potential, i.e., for a
896: constant $\mu(r)=m_e$ the spinor moves in a pure magnetic field and in
897: the state where the magnetic moment is antiparallel to the magnetic
898: field its coupling just compensates the lowest Landau level resulting
899: in a zero mode known from \cite{AharonovCasherj}. Switching on the
900: scalar potential the spinor feels an additional attraction and becomes
901: a bound state. As concerns the vacuum energy these bound states are
902: outside of the integration region in Eq. \Ref{Equant2} and therefore
903: they do not complicate the calculations. As shown in
904: \cite{Bordag:1996fv} they are accounted for automatically in the
905: representation of the vacuum energy in terms of the Jost function. The
906: only we have to take care of is a possible appearance of bound states
907: on the imaginary $k$-axis in Eq. \Ref{Equant2} above $m_e$. This would
908: require a zero of the Jost functions in representation \Ref{fg+} or
909: \Ref{fg-} for real $k$ there. But these are genuine complex
910: expressions and it can be shown that they do not have such zeros.
911:
912: \begin{figure}[t] \label{figbs}
913: \epsfxsize=12cm
914: \epsffile{figbs.eps}
915: \caption{The logarithm of the expressions in the figure brackets in
916: Eq. \Ref{jostbs}} as function of $\kappa$ for lowest orbital
917: momenta. The peak indicates a bound state. Its location is
918: $k=0.039232$.
919: \end{figure}
920:
921: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
922: \section{Numerical results}\label{Sec5}
923: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
924: We investigated numerically the vacuum energy for values of $\beta$
925: ranging from $\beta=0.3$ to $\beta=6$ for some choices of the remaining
926: parameters. The convergence properties of the sums and integrals
927: involved have been discussed in the previous section. Here, let us
928: first discuss the weight of the individual contributions.
929:
930: We start with the classical energy which is given by
931: Eq. \Ref{eclass}. It can be divided into three parts. The first one,
932: $E_{\rm mag}$, is the energy of the magnetic field. The second one,
933: $E_{\rm cov}$, results from the covariant derivative in Eq. \Ref{AH}
934: and is given by the second and the third contributions in
935: \Ref{eclass}. The third one, $E_{\rm scal}$, is the self energy of the
936: Higgs and is given by the fourth contributions in \Ref{eclass}. We
937: show in Table 1 two examples, one for smallest coupling of the scalar,
938: $\lambda=\beta q^2/2$ and the other for the largest value
939: considered. In both cases as well as in all intermediate ones, the
940: scalar self energy contribution is larger than the magnetic energy and
941: than $E_{\rm cov}$.
942:
943: \begin{table}
944: \begin{tabular}{|lllll|}\hline
945: & $E_{\rm mag}$ & $E_{\rm cov}$ & $E_{\rm scal}$ &$E_{\rm class}$\\
946: $\beta=0.3$ & 0.47913 & 1.50653 & 1.91653 & 3.90220\\
947: $\beta=6$ & 0.97224 & 2.64606 & 3.88896 & 7.50726\\\hline
948: \end{tabular}
949: \caption{Two examples for the classical energy and its constituent parts for
950: $q=0.5$, $\eta=1$.}
951: \end{table}
952:
953: Next we consider for the parameters $q=0.5$, $f_e=1$, $\eta=1$ the
954: constituent parts of the asymptotic part of the vacuum energy, i.e.,
955: the $\E^{\rm as}_j$ as defined in Eq. \Ref{Eas4}. These quantities are
956: shown in Table 2. As it can be seen the contributions $\E^{\rm as}_5$
957: and $\E^{\rm as}_6$ are dominating. They result from the scalar
958: background, see Eq. \Ref{Eas3}. But generally, all have at last a
959: numerical smallness of two orders of magnitude.
960:
961: \begin{table}
962: \begin{tabular}{|lllll|}\hline
963: & $\E^{\rm as}_1 $ & $\E^{\rm as}_2 $ & $\E^{\rm as}_3 $ & $\E^{\rm as}_4$ \\
964: $\beta=0.3$ & 0.00009139& -0.00006497& -5.5701 $10^{-8}$ &
965: 0.00001052\\
966: $\beta=6$& 0.0004916& -0.0003590& -1.1546 $10^{-6}$ & 0.00004194
967: \\\hline
968: & $\E^{\rm as}_5 $ & $\E^{\rm as}_6 $ & $\E^{\rm as}=\sum_{j=1}^6 \E^{\rm as}_j $ &\\
969: $\beta=0.3$ &
970: -0.004379& -0.0017591 &-0.0061022 &\\
971: $\beta=6$&-0.003898& -0.002483& -0.006208 &\\ \hline
972: \end{tabular}
973: \caption{ the constituent parts of the asymptotic part of the vacuum
974: energy for $q=0.5$, $f_e=1$, $\eta=1$.}
975: \end{table}
976:
977: Now, in Table 3, we represent the result of the calculations of the
978: complete vacuum energy for all considered examples. For two cases we
979: also represent the parts of the vacuum energy as a function of
980: $\beta$, see Figs. 5 and 6. From these results it can be seen that
981: there seems to be no general rule which part of the vacuum energy is
982: dominating and which not. Moreover, in dependence on the parameters
983: one part may be larger than other and smaller as well. It is even
984: impossible to say something definite about the sign of the vacuum
985: energy, it may change although in most cases it is negative.
986:
987: \begin{table}
988: \begin{tabular}{|llllllll|}\hline
989: $q$ &$\beta$& $f_e$&$\eta$& $E_{\rm class}$& $\E^{\rm as}_1 $ &
990: $\E^{\rm f} $ & $\E_{\rm ren}=\E^{\rm as} +\E^{\rm f} $ \\\hline
991: 0.5 & 0.3 & 1& 1 & 3.902 & -0.006102 & 0.02410 & 0.01800\\
992: 0.5 & 0.6 & 1& 1 & 4.549 & -0.006102 & 0.01590 & 0.009795\\
993: 0.5 & 1. & 1& 1 & 5.099 & -0.006104 & 0.01153 & 0.005421\\
994: 0.5 & 3.5 & 1& 1 & 6.710 & -0.006146 & 0.004224 & -0.001922\\
995: 0.5 & 6. & 1& 1 & 7.507 & -0.006208 & 0.001884 & -0.004325\\ \hline
996: 1. & 0.3 & 1& 1 & 2.465 & -0.005766 & 0.008686 & 0.002920\\
997: 1. & 0.6 & 1& 1 & 2.829 & -0.005665 & 0.005322 & -0.0003436\\
998: 1. & 1. & 1& 1 & 3.142 & -0.005594 & 0.003030 & -0.002565\\
999: 1. & 3.5 & 1& 1 & 4.093 & -0.005593 & -0.002588 & -0.008181\\
1000: 1. & 6. & 1& 1 & 4.591 & -0.005792 & -0.005077 & -0.01087\\ \hline
1001: 2. & 0.3 & 1& 1 & 2.105 & -0.003170 & -0.001883 & -0.005053\\
1002: 2. & 0.6 & 1& 1 & 2.399 & -0.002891 & -0.007285 & -0.01018\\
1003: 2. & 1. & 1& 1 & 2.652 & -0.002924 & -0.01182 & -0.01475\\
1004: 2. & 3.5 & 1& 1 & 2.652 & -0.002924 & -0.02451 & -0.02743\\
1005: 2. & 6. & 1& 1 & 3.861 & -0.006539 & -0.03022 & -0.03676\\ \hline
1006: 0.5 & 0.3 & 0.1& 0.1 & 0.03902 & -0.00006102 &-9.589 $10^{-6}$ &-0.00007061\\
1007: 0.5 & 0.6 & 0.1& 0.1 & 0.04549 & -0.00006102 & -0.00001214 & -0.00007316\\
1008: 0.5 & 1. & 0.1& 0.1 & 0.05099 & -0.00006104 & -0.00001421 & -0.00007525\\
1009: 0.5 & 3.5 & 0.1& 0.1 & 0.06710 & -0.00006146 & -0.00001975 & -0.00008121\\
1010: 0.5 & 6. & 0.1& 0.1 & 0.07507 & -0.00006208 & -0.00002216 & -0.00008425\\ \hline
1011: 0.5 & 0.3 & 0.1& 1 & 3.902 & -0.0003003 & -0.0009589 & -0.001259\\
1012: 0.5 & 0.6 & 0.1& 1 & 4.549 & -0.0005062 & -0.001214 & -0.001720\\
1013: 0.5 & 1. & 0.1& 1 & 5.099 & -0.0007110 & -0.001421 & -0.002132\\
1014: 0.5 & 3.5 & 0.1& 1 & 6.710 & -0.001423 & -0.001975 & -0.003398\\
1015: 0.5 & 6. & 0.1& 1 & 7.507 & -0.001814 & -0.002216 & -0.004030\\ \hline
1016: \end{tabular}
1017: \caption{The constituent parts of of the vacuum
1018: energy for all considered examples.}
1019: \end{table}
1020:
1021: \begin{figure}[t] \label{fig5}
1022: \epsfxsize=12cm
1023: \epsffile{fig5.eps}
1024: \caption{The vacuum energy as a function of $\beta$ for the $q=0.5$,
1025: $f_e=1$, $\eta=1$. In order to represent all quantities within one
1026: plot the classical energy is divided by 200.}
1027: \end{figure}
1028: \begin{figure}[t] \label{fig6}
1029: \epsfxsize=12cm
1030: \epsffile{fig6.eps}
1031: \caption{The same as in figure 5 but for $q=2$.}
1032: \end{figure}
1033:
1034:
1035: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1036: \section{Conclusions}\label{Sec6}
1037: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1038: In the present paper we calculated numerically the vacuum energy of a
1039: fermion in the background of a \NO string. As for renormalization we
1040: used standard zeta functional regularization and determined the
1041: counter terms from the first heat kernel coefficients (up to $a_2$).
1042: It turned out that from the renormalization a counter term appears
1043: which is not present in the initial action. It is gauge invariant and
1044: it has the correct dimension but it represents a non polynomial
1045: interaction. This has to be considered together with the non polynomial
1046: interaction present in the model itself as discussed in Sect. 2.
1047:
1048: The numerical investigations have been performed using methods
1049: developed in the papers \cite{Bordag:1996fv,Bordag:1998tg}. In the
1050: present paper the background is given purely numerically in difference
1051: to the previous papers where it had been given analytically, for
1052: example as a step function or some Gaussian profile. It has been
1053: demonstrated that the computational scheme used here is well suited to
1054: work with such backgrounds. This is a step forward to physically
1055: really interesting problems. In the considered model the stability of
1056: the background is given by topological arguments and for a realistic
1057: choice of the parameters the quantum contribution is small. This
1058: smallness has two sources. The one is the smallness of the coupling
1059: constants which appears in front of the quantum contribution relative
1060: to the classical one. The second one is a purely numerical smallness
1061: of about two orders of magnitude as discussed in Sect. 5. It is
1062: present even if the parameters and couplings are all of order one. It
1063: is obviously connected with the dimension and the \uv renormalization.
1064: So in \cite{Bordag:1995jz} it was found for a one dimensional
1065: background that in (1+1) dimension the vacuum energy is by one order
1066: of magnitude larger than for the same problem in (3+1)
1067: dimensions. Comparing formulas (29) and (32) in \cite{Bordag:1995jz}
1068: the interpretation is suggested that this difference is due to the one
1069: additional \uv subtraction in the (3+1) dimensional case.
1070:
1071: As seen in Sect. 5, for the dependence of the vacuum energy on the
1072: various parameters no general rule can be seen so far. Even the sign
1073: of the vacuum energy changes. The same applies to the relative weight
1074: of the individual contributions. So sometimes the asymptotic part is
1075: dominating, in other cases, however, the 'finite' part is larger. In
1076: general, they are of the same order. From this one can conclude only
1077: that in the given background there is no small parameter which could
1078: allow for some approximative scheme. So for instance, if the
1079: asymptotic part is dominating one could hope to get a good
1080: approximation by including higher orders of the uniform asymptotic
1081: expansion of the Jost function into the asymptotic part of the vacuum
1082: energy and neglect the 'finite' part of it which is the numerically
1083: much harder part of the problem. The examples in \cite{Bordag:2002sa}
1084: have been of a kind suggesting this way in contrast to the example in the
1085: present paper.
1086:
1087: For the considered model of a spinor in the background of the \NO
1088: vortex, due to its smallness, the vacuum energy has only a very small
1089: influence on the dynamics of the background. Hence in the considered
1090: case the vacuum energy is of limited physical importance. However, its
1091: calculation gave new insights into the structure of such calculations
1092: and demonstrated the power of the methods used. A next step could be
1093: to apply them to the $Z$ and electroweak strings where the stability
1094: is not guaranteed by topological arguments and where the stability
1095: issue is not finally settled with respect to the fermion contributions
1096: \cite{Achucarro:1999it,Groves:1999ks}.
1097: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1098: \section*{Acknowledgments}
1099: The authors are deeply indebted to V. Skalozub for valuable
1100: discussions especially in an early stage of the work. Further they
1101: thank V. Nikolaev for discussions on the numerical calculation of the
1102: background and D. Vassilevich for helpful discussions and
1103: suggestions. \\ One of the authors (I.D.) thanks the Graduiertenkolleg
1104: {\it Quantenfeldtheorie} at the University of Leipzig for support.
1105: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1106: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1107: \section*{Appendix}\label{App}
1108: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1109: The Abel-Plana formula used in Sec.2 reads
1110: %
1111: \be\label{APla}
1112: \sum\limits_{\nu=\frac12,\frac32,\dots}f(\nu)=
1113: \int\limits_{0}^{\infty}d \nu~f(\nu)
1114: +\int\limits_{0}^{\infty}{d \nu\over 1+e^{2\pi\nu}}{f(i\nu)-f(-i\nu)\over i}.
1115: \ee
1116: %
1117: The following formulas are used in the text,
1118: %
1119: \be\label{C3}
1120: \int\limits_m^\infty dk(k^2-m^2)^{1-s}{\partial\over\partial k}t^{j}
1121: =
1122: - m^{2-2s}
1123: {\Gamma(2-s)\Gamma(s+\frac{j}{2}-1)\over \Gamma(\frac{j}{2})}
1124: { \left({\nu\over mr }\right)^{j-n} \over
1125: \left(1+\left({\nu\over mr}\right)^{2}\right)^{s+\frac{j}{2}-1}}
1126: \ee
1127: %
1128: with $t=1/\sqrt{1+(kr/\nu)^2}$. They can be easily derived, see also
1129: (C3) and (C2) in \cite{Bordag:1998tg}.
1130:
1131: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1132:
1133: \begin{thebibliography}{10}
1134:
1135: \bibitem{Bordag:1998tg}
1136: M.~Bordag and K.~Kirsten.
1137: \newblock {\em Phys. Rev.}, 60:105019, 1999.
1138:
1139: \bibitem{Bordag:1996fv}
1140: M.~Bordag and K.~Kirsten.
1141: \newblock {\em Phys. Rev.}, D53:5753--5760, 1996.
1142:
1143: \bibitem{Scandurra:2000wr}
1144: Marco Scandurra.
1145: \newblock {\em Phys. Rev.}, D62:085024, 2000.
1146:
1147: \bibitem{Drozdov:2002um}
1148: I.~Drozdov.
1149: \newblock Vacuum polarization by a magnetic flux of special rectangular form.
1150: \newblock 2002.
1151: \newblock hep-th/0210282, Int.~J.~Mod.~Phys., to appear.
1152:
1153: \bibitem{Dunne:1998kw}
1154: Gerald~V. Dunne and Theodore~M. Hall.
1155: \newblock {\em Phys. Lett.}, B419:322--325, 1998.
1156:
1157: \bibitem{Fry:2001jw}
1158: M.~P. Fry.
1159: \newblock {\em Int. J. Mod. Phys.}, A17:936--945, 2002.
1160:
1161: \bibitem{Pasipoularides:2000gg}
1162: Pavlos Pasipoularides.
1163: \newblock {\em Phys. Rev.}, D64:105011, 2001.
1164:
1165: \bibitem{Pasipoularides:2003jt}
1166: Pavlos Pasipoularides.
1167: \newblock The strong magnetic field asymptotic behaviour for the
1168: fermion-induced effective energy in the presence of a magnetic flux tube.
1169: \newblock 2003.
1170: \newblock hep-th/0301192.
1171:
1172: \bibitem{Langfeld:2002vy}
1173: Kurt Langfeld, Laurent Moyaerts, and Holger Gies.
1174: \newblock {\em Nucl. Phys.}, B646:158--180, 2002.
1175:
1176: \bibitem{Diakonov:1999gg}
1177: Dmitri Diakonov.
1178: \newblock {\em Mod. Phys. Lett.}, A14:1725--1732, 1999.
1179:
1180: \bibitem{Nielsen:1973cs}
1181: Holger~Bech Nielsen and P.~Olesen.
1182: \newblock {\em Nucl. Phys.}, B61:45--61, 1973.
1183:
1184: \bibitem{Bordag:1995jz}
1185: M.~Bordag.
1186: \newblock {\em J. Phys.}, A28:755--766, 1995.
1187:
1188: \bibitem{Bordag:2002sa}
1189: M.~Bordag.
1190: \newblock {\em Phys. Rev.}, D67:065001, 2003.
1191:
1192: \bibitem{Graham:2002fi}
1193: Noah Graham, Robert~L. Jaffe, and Herbert Weigel.
1194: \newblock {\em Int. J. Mod. Phys.}, A17:846--869, 2002.
1195:
1196: \bibitem{Graham:2002xq}
1197: N.~Graham et~al.
1198: \newblock {\em Nucl. Phys.}, B645:49--84, 2002.
1199:
1200: \bibitem{Diakonov:2002bx}
1201: Dmitri Diakonov and Martin Maul.
1202: \newblock {\em Phys. Rev.}, D66:096004, 2002.
1203:
1204: \bibitem{Groves:1999ks}
1205: Martin Groves and Warren~B. Perkins.
1206: \newblock {\em Nucl. Phys.}, B573:449--500, 2000.
1207:
1208: \bibitem{Gies:2001tj}
1209: Holger Gies and Kurt Langfeld.
1210: \newblock {\em Int. J. Mod. Phys.}, A17:966--978, 2002.
1211:
1212: \bibitem{Gies:2001zp}
1213: Holger Gies and Kurt Langfeld.
1214: \newblock {\em Nucl. Phys.}, B613:353--365, 2001.
1215:
1216: \bibitem{Achucarro:1999it}
1217: Ana Achucarro and Tanmay Vachaspati.
1218: \newblock {\em Phys. Rept.}, 327:347--426, 2000.
1219:
1220: \bibitem{AharonovCasherj}
1221: Y.~Aharonov and A.~Casher.
1222: \newblock {\em Phys. Rev.}, A19:2461--2462, 1979.
1223:
1224: \bibitem{Bordag:2000f}
1225: M.~Bordag.
1226: \newblock {\em Comments on Atomic and Nuclear Physics, Comments on Modern
1227: Physics}, 1, part D:347--361, 2000.
1228:
1229: \bibitem{Bordag:1999vs}
1230: M.~Bordag, K.~Kirsten, and D.~Vassilevich.
1231: \newblock {\em Phys. Rev.}, D59:085011, 1999.
1232:
1233: \bibitem{BGNV}
1234: Michael Bordag, Alfred~Scharff Goldhaber, Peter van Nieuwenhuizen, and Dmitri
1235: Vassilevich.
1236: \newblock {\em Phys. Rev.}, D66:125014, 2002.
1237:
1238: \end{thebibliography}
1239: \end{document}
1240:
1241:
1242: