hep-th0102002/br.tex
1: 
2: \documentstyle[12pt,a4]{article}
3: \renewcommand{\baselinestretch}{1.5}
4: \newcommand{\be}{\begin{equation}}
5: \newcommand{\ee}{\end{equation}}
6: \newcommand{\unop}{1\!{\rm I}}
7: 
8: \title{Classical solutions in the Einstein-Born-Infeld-Abelian-Higgs model}
9: \author{Y. Brihaye, B. Mercier\\
10: Facult\'e des Sciences, Universit\'e de Mons-Hainaut,\\
11: B-7000 MONS, Belgium.}
12: \date{\ }
13: \begin{document}
14: \begin{titlepage}
15: \maketitle
16: \thispagestyle{empty}
17: \begin{abstract}
18: We consider the classical equations of the
19: Born-Infeld-Abelian-Higgs model (with and without coupling to gravity)
20: in an axially symmetric ansatz. 
21: A numerical analysis of the equations reveals that the (gravitating)
22: Nielsen-Olesen vortices are smoothly deformed by the Born-Infeld
23: interaction, characterized by a coupling constant $\beta^2$,
24: and that these solutions cease to exist at a critical value of $\beta^2$.
25: When the critical value is approached, the length of the magnetic field
26: on the symmetry axis becomes infinite.
27: \end{abstract}
28: \vfill
29: \end{titlepage}
30: \newpage
31: \section{Introduction}
32: Recently, various gauge field theories were studied in which the 
33: dynamics of the gauge fields is governed by a Born-Infeld (BI) lagrangian
34: \cite{bi}. The relationship between string theory and BI lagrangian
35: obtained in \cite{tsey1} started an important activity in the topic
36: which is reviewed e.g. in \cite{pol, tsey2}.
37: 
38: Among the various aspects of BI field theory, the construction 
39: of classical solutions occupies a large place. Many investigations
40: were carried out in Abelian \cite{mns} as well as in non-Abelian
41: \cite{kg,gpss} models. Often, the pattern of classical solutions
42: is enriched by the occurence of the BI-term. The case of the SU(2)
43: gauge theory is typical of this type: no static finite
44: energy solutions exist with the minimal Yang-Mills action
45: but solutions exist when the BI term is included \cite{kg}.
46: Lately, various Born-Infeld solitons have been considered
47: in the context of gravity (see. e.g. \cite{tri,wsk}).
48: 
49: 
50: Perhaps the simplest model containing solitons for which the
51: effects of the BI term may be studied is the Abelian-Higgs model
52: \cite{no} in which vortex-solutions are available.
53: Several aspects of BI-vortex solutions have been  considered so far, 
54: namely the construction of self dual equations from the BI
55: lagrangian \cite{shir} and the construction of electrically
56: charged vortex solutions \cite{mns}.
57: Very interesting applications of the vortex solutions are cosmic
58: strings \cite{vil} which are believed to 
59: have played a role in the early phase of the Universe; it is therefore natural to also consider the 
60: effect of gravitation on the BI-vortex solutions.
61: 
62: Here we consider the Abelian-Higgs-Born-Infeld model coupled to gravity
63: and we preform a detailed numerical analysis of the equations in the
64: axially symmetric ansatz.
65: We find that, when the various coupling constants 
66: of the Einstein-Abelian-Higgs equations
67: are such that a magnetically charged
68: soliton exists (i.e. for $\beta^2 = \infty$ where $\beta^2$ denotes the
69: BI coupling constant), then the solution is smoothly deformed 
70: by the BI interaction. When the parameter  $\beta^2$ approaches
71: a critical value, the length of the
72: magnetic field of the solution on its
73: symmetry axis becomes infinite and the solution disappears.
74: 
75: The paper is organized as follows~: in Sect. 2 we describe the
76: Lagrangian, the ansatz and the classical equations. Several known
77: solutions and some of their features are summarized in Sect. 3.
78: Sect. 4 contains our results on the evolution of these 
79: solutions in presence of the Born-Infeld interaction.
80: %%%%%%%%%%%%%%%%%%%%%%%
81: %%%%%%%%%%%%%%%%%%%%%%%%
82: 
83: \section{The equations.}
84: We consider the gravitating 
85: Born-Infeld-Abelian Higgs model in four dimensions. It is
86: described by the action \cite{bi,no}
87: \be
88: S = \int d^4x \sqrt{-g}{\cal L}
89: \ee
90: \be
91: \label{lagran}
92: {\cal L} = 
93:    {1\over 2} D_{\mu} \phi D^{\mu} \phi^{\ast} 
94: +  \hat \beta^2 (1 - {\cal R}) 
95: - {\lambda \over 4} (\phi^{\ast} \phi-v^2)^2 
96: + {1\over{16 \pi G}} R
97: \ee
98: with 
99: \be
100:   {\cal R} = \sqrt{1 + \frac{F_{\mu\nu}F^{\mu\nu}}{2 \hat \beta^2}
101:          -\frac{(F_{\mu\nu}\tilde F^{\mu\nu})^2}{16 \hat \beta^4}
102:                    }
103: \ee
104: where $D_{\mu} = \nabla_{\mu} - i e A_{\mu}$ is the gauge
105: covariant derivative, $A_{\mu}$ is the gauge potential
106: of the U(1) gauge symmetry,
107: $F_{\mu\nu}$ is the
108: corresponding field strength and $\phi$ is a complex scalar field with vacuum
109: expectation value $v$. 
110: The Born-Infeld coupling constant
111:  is noted $\hat \beta^2$. The geometry is introduced as usual
112: by means of the Ricci scalar $R$. We use the same
113: notations as in \cite{verbin}.
114: 
115: \par In the following we study the classical equations
116: of the above field theory in the static case and within the
117:  cylindrically symmetric ansatz. The
118: metric and matter fields are parametrized in terms of four functions of the cylindrical radial variable  $r$  :
119: \be
120: ds^2 = N^2(r) dt^2 - dr^2 -  L^2(r) d\varphi^2-N^2(r) dz^2
121: \ee
122: \be
123: \phi = vf(r) e^{i n\varphi}
124: \ee
125: \be
126: A_{\mu} dx^{\mu} = {1\over e} (n-P(r)) d\varphi
127: \ee
128: Here $N, L,P,f$ are the  radial functions of $r$, $n$ is an integer
129: indexing the vorticity of the Higgs field around the $z$-axis. 
130: 
131: 
132: 
133: In the process of establishing the classical equations
134: in terms of the radial functions  it appears convenient
135: to define  dimensionless coupling constants ~:
136: \be
137: \alpha = \frac{e^2}{\lambda} \ \ , \ \  \gamma = 8\pi G v^2 \ \ , 
138: \ \ \beta^2 = \frac{\hat \beta^2}{\lambda v^4} \ \ ,
139: \ee
140: and to rescale the radial variable $r$ and the function $L(r)$ according to
141: \be
142: x = \sqrt{\lambda v^2}r \ \ , \ \ L(x) = \sqrt{\lambda v^2}  L(r)
143: \ee
144: 
145: 
146: \par 
147: With these redefinitions (and in the units $\hbar = c = 1$)
148: the diagonal components of the energy-momentum tensor read
149: \begin{eqnarray}
150: {\cal T}^0_0 &=& \lambda v^4(\epsilon_s+\epsilon_v+\epsilon_w+u)\\
151: {\cal T}^r_r &=& \lambda v^4(-\epsilon_s+\epsilon_v-\tilde\epsilon_v+\epsilon_w+u)\\
152: {\cal T}^{\varphi}_{\varphi} &=& \lambda v^4(\epsilon_s+\epsilon_v-\tilde \epsilon_v-\epsilon_w
153: +u)\\
154: {\cal T}^z_z &=& {\cal T}^0_0
155: \end{eqnarray}
156: where the different densities are
157: expressed in the dimensionless variable and
158: functions~:
159: \be
160: \epsilon_s = {1\over 2}(f')^2\quad , \quad \epsilon_w = {P^2f^2\over {2L^2}}\quad , \quad u = {1\over 4} (1-f^2)^2
161: \ee
162: \be
163: \epsilon_v = \beta^2({\cal R}-1)\quad , \quad \tilde \epsilon_v = {(P')^2\over
164: {\alpha L^2 {\cal R}}}\quad , \quad {\cal R} \equiv \sqrt{1+{(P')^2\over{\alpha\beta^2L^2}}}
165: \ee
166: the prime indicates the derivative with respect to $x$.
167: The corresponding components of the Ricci tensor are given explicitely
168: e.g. in \cite{verbin}.
169: 
170: Then, after an algebra, the classical equations 
171: corresponding to the above lagrangian
172: and ansatz can be shown to be consistent and lead
173:  to the following
174: system of four equations for the functions $N,L,P,f$~:
175: \be
176: \label{eq1}
177: {(LNN')'\over{N^2L}} = \gamma \left(
178: \beta^2 (1-{\cal R}) 
179: + \frac{P'^2}{\alpha L^2} \frac{1}{{\cal R}}
180: -{1\over 4}(1-f^2)^2
181: \right)
182: \ee
183: \be
184: \label{eq2}
185: {(N^2L')'\over{N^2L}} = \gamma \left( 
186: \beta^2 (1-{\cal R}) 
187: - {P^2f^2 \over{L^2}} 
188: - {1\over 4} (1-f^2)^2\right)
189: \ee
190: \be
191: \label{eq3}
192: {L\over{N^2}} (\frac{N^2 P'}{L{\cal R}})' = \alpha f^2 P
193: \ee
194: \be
195: \label{eq4}
196: {(N^2Lf')'\over{N^2L}} = f(f^2-1) + f {P^2\over {L^2}}
197: \ee
198: %%with
199: %%\be
200: %%          {\cal R} = \sqrt{1 + \frac{(P')^2}{\alpha \beta^2 L^2}}
201: %%\ee
202: 
203: The problem posed by these differential equations is 
204: further specified by the following set of boundary conditions
205: \begin{eqnarray}
206: \label{cn} N(0) &=& 1\qquad , \qquad N'(0) = 0\\
207: \label{cl} L(0) &=& 0 \qquad , \qquad L'(0) = 1\\
208: \label{cp} P(0) &=& n \qquad , \qquad \lim_{x\rightarrow \infty} P(x) = 0 \label{condp} \\
209: \label{cf} f(0) &=& 0 \qquad , \qquad \lim_{x\rightarrow \infty}f(x) = 1
210: \end{eqnarray}
211: which are necessary to guarantee the regularity of the configuration 
212: on the symmetry axis
213: and the finiteness of the inertial mass  defined next.
214: 
215: 
216: 
217: The classical solutions can be characterized, namely,
218:  by their magnetic field
219: \be
220: \label{magnetic}
221: B(r) = -{1\over{eL(r)}} {dP\over{dr}}(r) = -{\gamma\over{\alpha}}
222: ({e\over{8\pi G}}) {P'(x)\over{L(x)}},
223: \ee
224: and by their inertial mass and Tolman mass by unit length \cite{verbin}
225: \begin{eqnarray}
226: {\cal M}_{in} &=& \int \sqrt{-g_3} \ {\cal T}^0_0 \ dx_1dx_2\\
227: {\cal M}_{t_0} &=& \int \sqrt{-g_4} \ 
228: ({\cal T}^0_0 - {\cal T}^r_r- {\cal T}^{\varphi}_{\varphi}- {\cal T}^z_z) \ dx_1dx_2
229: \end{eqnarray}
230: which in turn can be reduced to one dimensional integrals~:
231: \begin{eqnarray}
232: G{\cal M}_{in} &=& {\gamma\over 8} M_{in}\\
233: &=& {\gamma\over 8} \int^{\infty}_0 dx NL(\epsilon_s + \epsilon_w + \epsilon_v + u).\\
234: G{\cal M}_{t_0}&=& {\gamma\over 8} M_{t_0}\\
235: &=& {\gamma\over 8} \int^{\infty}_0 dx N^2L (\tilde \epsilon_v-\epsilon_v-u).
236: \end{eqnarray}
237: In the flat case ($\gamma = 0$) the classical energy $E$
238: of the solutions
239: is obtained directly in terms of the inertial mass~: 
240: $E= (2 \pi v^2) M_{in}$. 
241: 
242: 
243: \vspace{0.5cm}
244: 
245: \section{Some special cases.}
246: In this section we briefly 
247: review a few particular solutions of the system above
248: and comment on their main properties.
249: 
250: \subsection{Nielsen-Olesen vortices.}
251: 
252: We first consider the equations 
253: for the case $\gamma=0, \beta=\infty$.
254: This case corresponds to the flat Maxwell-Higgs system; Eqs. 
255: (\ref{eq1}), (\ref{eq2}) are 
256: solved by $N(x) = 1, L(x) = x$. The solutions 
257: of the remaining equations are embeddings  of the celebrated
258: two-dimensional Nielsen-Olesen vortices \cite{no} into 
259: the three-dimensional space. 
260: They exist for all values of $\alpha$ and $n$
261:  and possess the following behaviour around the origin
262: \be
263: \label{noor}
264: P(x) = n - \frac{a_0}{2} x^2+O(x^{2+2n})
265:  \ \ \ , \ \ \ 
266: f(x) = b_0 x^n + O(x^{n+2})
267: \ee
268: and asymptotically
269: \be
270: \label{nobord}
271: P(x) = a_{\infty} e^{-x \sqrt{\alpha}}
272: \ \ \ , \ \ \ 
273: f(x) = 1-f_{\infty} e^{-x \sqrt{2}}  \ \ .
274: \ee
275: The parameters $a_0, b_0, a_{\infty}, f_{\infty}$ are determined numerically. For fixed $n$,
276: the classical energy, say $E(\alpha,n)$, of the solution
277: decreases monotonically when $\alpha$ increases. The
278: value $\alpha=2$ is very particular because first order equations exist
279: \cite{taubes} which imply (\ref{eq3}), (\ref{eq4})
280: (they are usually called self dual equations).
281: \par In relation to the self dual case, $\alpha=2$, the Nielsen-Olesen
282: solutions enjoy the following remarkable property :
283: \begin{eqnarray}
284: \label{sd}
285: {1\over n} E(\alpha,n)&>& E(\alpha,1)\qquad {\rm for} \qquad \alpha< 2\nonumber\\
286: {1\over n} E(\alpha,n) &=& E(\alpha,1)\qquad {\rm for} \qquad \alpha=2\nonumber\\
287: {1\over n} E(\alpha,n) &<& E(\alpha,1)\qquad {\rm for} \qquad \alpha > 2
288: \end{eqnarray}
289: as illustrated by Fig. 1 for $n=1,2$. This supports the idea 
290: \cite{rebbi} that, in the domain $\alpha > 2$, the n-vortice solutions
291: (with $n>2$) are stable against decay into $n$
292: separated solutions with unit vorticity (remembering the exponential decay
293: (\ref{nobord}) of the solutions under consideration).
294: 
295: \vspace{0.5cm}
296: 
297: \subsection{ The Melvin solution.}
298: \par Cancelling the Higgs field and the Higgs potential in the initial
299: lagrangian, we are left with the Einstein-Maxwell lagrangian. The
300: corresponding equations can be obtained from (\ref{eq1})-(\ref{eq4}) after
301: an appropriate rescaling of $x$ and $L$~:
302: \be
303: x\rightarrow {x\over{\sqrt{\gamma}}}\quad , \quad L \rightarrow
304: \sqrt{\gamma\over{\alpha}} L
305: \ee
306: and setting $\gamma=0$ afterwards
307: (then Eq.(\ref{eq4}) decouples). 
308: A solution of these equation was discovered by
309: Melvin \cite{melvin}. It is namely characterized by 
310: the following (asymptotic) power law behaviours of the various fields   
311: \be
312: N(x\rightarrow \infty) = Ax^{2\over 3}\quad , \quad L(x\rightarrow \infty)
313: = B x^{-{1\over 3}}\quad , \quad P(x\rightarrow \infty) = Cx^{-{2\over 3}}
314: \ee
315: where $A,B,C$ are constants (the numerical evaluation gives $A\approx
316: 2.25, B\approx 0.44)$. The corresponding masses are $M_{in} = M_{t_0}
317: = 2$.
318: 
319: \vspace{0.5cm}
320: 
321: \subsection{Gravitating vortices} 
322: \par The solutions corresponding to the Maxwell case
323: (i.e. $\beta^2 = \infty$)  were studied at length in \cite{verbin,bl}. We just mention
324: that global solutions (i.e. defined for $x\in [0,\infty])$ exist only on a 
325: finite domain in the (quarter of) plane $\alpha,\gamma$.
326: More precisely they exist for
327: \be
328: \label{domain}
329: 0 \leq \gamma \leq \gamma_{c}(\alpha,n)
330: \ee
331: %%%On Fig. 2, we present a plot of the critical values %%$\gamma_{cr}(\alpha,n)$ for $n=1$ and $n=2$.
332: The function $\gamma_c(\alpha,n)$ increases monotonically with $\alpha$
333: and, around the self-dual value $\alpha= 2$ it roughly behaves accordingly to
334: $\gamma_c(\alpha,n) = 2/n + \kappa(\alpha-2) \ , \ \kappa \approx 0.15$.  
335: In fact, for fixed $\alpha$, there exist two branches of solutions
336: for the set of $\gamma$ in the domain (\ref{domain}).
337: Following \cite{verbin} we refer  to the two branches as to  
338: the string and the Melvin branches. 
339: 
340: For $\gamma > \gamma_{c}(\alpha,n)$, solutions  still do
341: exist but they are closed 
342: \cite{verbin}; i.e. they are only defined for $x\in [0,x_{max}]$.
343:  We will not
344: be concerned with these type of solutions in the present paper.
345: 
346: \section{Numerical results.}
347: We solved numerically the system (\ref{eq1})-(\ref{eq4}) for numerous values
348: of the coupling constants $\alpha,\beta,\gamma$, studying in particular the
349: response of the three types of solutions mentioned above to
350: the Born-Infeld interaction. We now discuss then results in the same
351: order as before.
352: 
353: \subsection{The Nielsen-Olesen-Born-Infeld vortices}
354: We start by considering $\gamma = 0$. The formula  
355: (\ref{nobord}) still holds for $\beta < \infty$ and,
356: as expected, the Nielsen-Olesen solutions are smoothly deformed by the 
357: Born-Infeld term.  Fig. 2
358: shows the classical energy 
359: by means of $M_{in}(\alpha,\beta,n)/n$ as a function of $\beta^2$
360: for several values of $\alpha$ and $n$. It clearly indicates that the inertial
361: mass depends only little on $\beta$.
362: The corresponding values $E_m$ in the Maxwell limit ($\beta^2 = \infty$)
363: are indicated nearby. The main feature demonstrated by this figure
364: is that the solutions exist only for a finite range of the parameter $\beta$,
365: i.e. for  
366: \be
367: \beta_{c}(\alpha,n)< \beta < \infty
368: \ee
369: confirming the result of \cite{mns}. 
370: An in-depth analysis about the reasons
371: for this critical phenomenon is provided by Fig. 3 where
372: the  numerical evaluation of the parameters $a_0, b_0$ for the solutions are
373: plotted as functions of $\beta^2$. Our numerical analysis strongly indicates
374: that 
375: \be
376: \label{lim}
377: \lim_{\beta\rightarrow \beta_{c}} a_0 = \infty
378: \ee
379: so that, very likely, no solutions exist for $\beta < \beta_{c}$.
380:  This critical behaviour is, however, not reflected by the value of the
381: classical energy and by the parameter $b_0$ which remain finite for 
382: $\beta\rightarrow \beta_{c}$.
383: \par Physically, this phenomenon is 
384: reflected by the magnetic field of the solution. 
385: Indeed, from the definition (\ref{magnetic}),
386:  it is immediate to see that $B(0)$
387: (i.e. the value of the magnetic field on the symmetry axis of the solution) is directly 
388: related to $a_0$; Eq. (\ref{lim}) can therefore be rephased in saying that the magnetic
389: field on the z-axis becomes infinite when the critical value of $\beta$ is approached.
390:  This phenomenon is further illustrated  by Fig. 4 where the
391:  magnetic function $B(x)$ is compared for several values of $\beta^2$
392:  in the case $\alpha=1, n=1$. 
393: 
394: Refering to the relations (\ref{sd}), a natural question is to 
395: analyse how the domain of stability of the
396: $n>2$-vortices (i.e. $2<\alpha<\infty$) develops with the Born-Infeld interaction. 
397: An important element of  answer is given by Fig. 1. 
398: Clearly the effect of the Born-Infeld term is that the 
399: stable phase appears for lower values of $\alpha$. 
400: In terms of the Higgs particle mass $M_H = \lambda v$
401: (remember $\alpha = {e^2\over{\lambda}}$)
402: this means that the stable phase persists
403: up to a  higher value of the Higgs particle mass.
404: 
405: 
406: 
407: 
408: \par This result motivates an  analysis
409: of the effect of the Born-Infeld term in the
410: Georgi-Glashow model for multi-monopole solutions. 
411: It is known  that in this model, self dual equations exist
412: for $\lambda=0$ (i.e. for a massless Higgs particle) and that
413: multimonopoles solutions are unstable for $\lambda > 0$ \cite{manton}.
414: 
415: 
416: It is likely that the Born-Infeld
417: interaction generates a small domain of the Higgs mass
418: parameter where the mass of the two-monopole
419: is lower than twice the mass of the single monopole, thus leading to a stable phase.
420: Such a phenomenon was observed recently in the gravitating Georgi-Glashow model \cite{hkk}.
421: Due to gravity, a region of the parameter space exists where the multimonopoles seem to be stable against decay into monopoles.
422: 
423: 
424: 
425: \subsection{The Melvin-Born-Infeld Solution.}
426: \par The Melvin solution gets smoothly deformed by the Born-Infeld term. Decreasing
427: $\beta^2$ we observed that the solution disappears for $\beta^2_{c}\approx 6.4$; again
428: the phenomenon is related to the fact that the value $B(0)$ tends to infinity for
429: $\beta^2\rightarrow \beta^2_{c}$. The profiles of $N(x), L(x)$ and of the magnetic
430: field for the Melvin solution and its deformation for $\beta^2 = 7$ are compared on Fig. 5 (the dotted line for $B$ has $B(0)\approx 40$).
431: 
432: 
433: \subsection{The Born-Infeld gravitating strings.}
434: The study of the equations (\ref{eq1}-\ref{eq4}) for generic values
435: of $\alpha, \beta, \gamma$ is a vast task. Here, we have attempted to
436: confirm the features which occured for the flat case. 
437: We limit ourselves to the case $\alpha = 2$ but we expect that the
438: results   are qualitatively
439: the same for the other values of this coupling constant. 
440: Our numerical analysis of the equations
441:  confirms that, for $\gamma$ fixed, the solutions
442: on the string branch as well as the ones on the Melvin branch
443: terminate at finite values of $\beta^2$.
444: Fig. 6 shows the domains of existence of the two types of solutions
445: in the $\beta^2, \gamma$ plane.
446: The figure clearly reveals that the value $\beta_c^2$, where the 
447: string branch terminates (the lower curve) varies only little with
448: $\gamma$ and that is is smaller than its counterpart,
449: say $\tilde \beta^2_c$, for the Melvin branch
450: (upper curve).
451: 
452: Taking e.g.  $\alpha=2, \gamma=1$ we find
453: \be
454:      \beta^2_{c} \approx 0.4   \ \ \ , \ \ \ 
455:       \tilde \beta^2_{c} \approx 5.1
456: \ee
457: respectively for the solution of the string and of the Melvin branches.
458: The inertial mass deviates only slightly from the  exact value
459: $M_{in}=2$ in the Maxwell limit, the Tolman mass decreases slowly
460: from $M_{to}\approx 0.62$ in the Maxwell limit to $M_{to}\approx 0.4$ 
461: for $\beta^2 = \tilde \beta^2_{c}$. 
462: 
463: For both branches, the value $B(0)$ of the magnetic field
464: on the z-axis becomes very large (probably infinite
465: but our numerical analysis
466: does not allow us to be absolutely sure) 
467: when the critical value of $\beta^2$ is approached.
468: Other quantities, e.g. the inertial and Tolman masses remain finite.
469: 
470: We finally mention that the string branch 
471: (which exists for $0\leq \gamma \leq
472: 2$ in the case $\alpha = 2$) extends slightly into the region $\gamma > 2$
473: when the Born-Infeld term is present.
474: 
475: 
476: \section{Final remarks}
477: To finish, we would like to present a generalization
478: of the result in \cite{shir} to the gravitating case.
479: The self-dual potential and the corresponding equations
480: (i.e.  setting $\lambda=e^2/2$ in Eq.(\ref{lagran}) and $\alpha = 2$ in 
481: Eqs.(\ref{eq1})-(\ref{eq2})) can be replaced by a  $\beta^2$-depending one
482: for which first order equations exist which imply the classical (second order) 
483: axially-symmetric equations. 
484: In  the reduced  variables, this can be achieved by the following substitution
485: \be
486: \label{potsd}
487: %%  V(\phi) = \hat \beta^2 \Biggl(   1- 
488: %%%\sqrt{ 1 - {e^2 \over 2 \hat \beta^2} (\phi^{\ast} \phi-v^2)^2 }
489: %%%                         \Biggl)
490: V(f) = \frac{1}{4}(f^2-1)^2 \longrightarrow 
491:   V_{sd}(f) =  \beta^2 \Biggl(   1- 
492: \sqrt{ 1 - {1 \over 2  \beta^2} (f^2-1)^2 }
493:                          \Biggl)
494: \ee
495: With this potential the self dual equations read
496: \be
497: N(x)=1 \ \ , \ \ L'= 1 + \frac{\gamma}{2}( P(1-f^2) + \theta) \ \  , \ \  
498: \frac{P'}{L} = \frac{1}{f} \frac{\partial V_{sd}}{\partial f} \ \ , \ \ 
499: f' = \frac{P \ f}{L}
500: \ee
501: and imply the full axially symmetric equations of the theory.
502: Further exact solutions and algebraic properties 
503:  can be obtained with this modified theory.
504: 
505: 
506: %%%%%%%%****************************
507: %%%%%%%% J'AI COMMENCE CI-DESSUS
508: %%%%%%%%**************************
509: %%%%%%%%*************************
510: \vskip 2 cm 
511: {\bf Acknowledgements}\hfill\break
512:  One of us (Y.~B.) gratefully acknowledges conversations 	
513:  with B.~Hartmann and with F.A.~Schaposnik. 
514:                                                        
515: 
516: \newpage
517: \begin{thebibliography}{7}
518: \bibitem{bi}M. Born and M. Infeld, Proc. R. Soc. London A 143 (1934) 410.
519: \bibitem{tsey1} E.S. Fradkin and A.A. Tseytlin, Phys. Lett. 163b (1985) 123.
520: \bibitem{pol} J. Polchinski, {\it Tasi Lectures on D-Branes}, hep-th/9611050.
521: \bibitem{tsey2} A.A. Tseytlin, {Born-Infeld action, supersymmetry and string theory},              hep-th/9908105.  
522: \bibitem{mns} E. Moreno, C. Nunez, F.A. Schaposnik, Phys. Rev. D58 (1998) 369.
523: \bibitem{kg} D. Gal'tsov and R. Kerner, Phys. Rev. Lett. 84 (2000) 5955. 
524: \bibitem{gpss} N. Grandi, R.L. Pakman, F.A. Schaposnik and G. Silva, Phys. Rev. D472 (2000) 125002.
525: \bibitem{tri} P.K. Tripathy, Phys. Lett. B458 (1999) 252, Phys. Lett. B463
526: (1999) 1.
527: \bibitem{wsk} M. Wirschins, A. Sood and J. Kunz, hep-th/0004130, Phys. Rev.
528: D (to appear)
529: \bibitem{no} H.B. Nielsen and P. Olesen, Nucl. Phys. B61 (1973) 45.
530: \bibitem{shir}K. Shiraishi and S. Hirenzaki, Int. Journ. Mod. Phys. A6 (1991) 2635.
531: \bibitem{vil} A. Vilenkin, Phys. Rep. 121 (1985) 263.
532: \bibitem{verbin} M. Christensen, A.L. Larsen and Y. Verbin, Phys. Rev.
533: D60 (1999) 12501-2.
534: \bibitem{taubes} C. H. Taubes, Commun. Math. Phys. 72 (1980) 277.
535: \bibitem{rebbi} L. Jacobs and C. Rebbi, Phys. Rev. B 19 (1979) 4486.
536: \bibitem{melvin} M. A. Melvin, Phys. Lett. 8 (1964) 65.
537: 
538: \bibitem{bl} Y. Brihaye and M. Lubo, Phys. Rev. D62  (2000) 085004. 
539: \bibitem{manton} N. S. Manton, Nucl. Phys. B 126 (1977) 525. 
540: 
541: \bibitem{hkk} B. Hartmann, J. Kunz and B. Kleihaus, hep-th:/0009195,
542:  Phys. Rev. Lett. (to appear).
543: 
544: \end{thebibliography}
545: 
546: \newpage
547: \centerline{Figures Captions}
548: \begin{description}
549: \item [\ ] {\bf{Figure 1}}\\
550: The ratio  $E/n$ is plotted as a function of $\alpha$
551: for the Maxwell and Born-Infeld vortices and for $n=1,2$. 
552: \item [\ ] {\bf{Figure 2}}\\
553: The ratio $E/n$ for the Born-Infeld vortices is plotted
554: as a function of $\beta^2$ for different values of $\alpha$ and $n$.
555: On each branch, $E_m$ gives the energy of the Maxwell solution.
556: \item [\ ] {\bf{Figure 3}}\\
557: The parameters $a_0,b_0$ entering in the solutions via Eq. (\ref{noor})
558: are reported as functions of $\beta^2$ for different values of
559: $\alpha$ and of $n$.
560: \item [\ ] {\bf{Figure 4}}\\
561: The profiles of the magnetic field $B(x)$ is presented for
562: several values of $\beta^2$ approaching the critical value 
563: $\beta_c^2\approx 0.318$.
564: \item [\ ] {\bf{Figure 5}} \\
565: The profiles of the functions $L,N,B$ for Melvin solution
566: (solid lines) and Melvin-Born-Infeld solution (dotted lines).
567: On the dotted line $B(0) \approx 42$.
568: \item [\ ] {\bf{Figure 6}} \\
569: The domain of the gravitating string and Melvin branches of solutions
570: for $\alpha = 2$.
571: 
572: \end{description}
573: 
574: 
575: 
576: 
577: \end{document}
578: