1: \documentstyle[11pt,amsfonts]{article}
2: \renewcommand{\theequation}{\arabic{equation}}
3: \newcommand{\be}{\begin{equation}}
4: \newcommand{\ee}{\end{equation}}
5: \newcommand{\bea}{\begin{array}}
6: \newcommand{\ea}{\end{array}}
7: \newcommand{\beqa}{\begin{eqnarray}}
8: \newcommand{\eeqa}{\end{eqnarray}}
9: \newcommand{\bean}{\begin{eqnarray*}}
10: \newcommand{\eean}{\end{eqnarray*}}
11: \newcommand{\eqn}[1]{(\ref{#1})}
12: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
13: \newcommand{\del}{\partial}
14: \newcommand{\ao}{\mbox{\bf a}}
15:
16:
17: \newcommand{\nn}{\nonumber}
18:
19:
20: % A macro to raise things. Used in math and journal macros.
21: \def\up#1{\leavevmode \raise.16ex\hbox{#1}}
22:
23: \font\eightrm=cmr10
24:
25: \newcommand{\journal}[4]{{\sl #1 }{\bf #2} \up(19#3\up) #4}
26:
27: %%%%%%%% my style
28: \setlength{\textheight}{9.0in}
29: \setlength{\textwidth}{6.2in}
30: \setlength{\topmargin}{-0.375in}
31: \hoffset=-.5in
32: \renewcommand{\baselinestretch}{1.17}
33: \setlength{\parskip}{6pt plus 2pt}
34:
35: \newcommand{\gapproxeq}{\lower
36: .7ex\hbox{$\;\stackrel{\textstyle >}{\sim}\;$}}
37: \newcommand{\lapproxeq}{\lower .7ex\hbox{$\;\stackrel
38: {\textstyle <}{\sim}\;$}}
39:
40: % the following commands make the equations be numbered by section
41: %they must not be used with the chapter choice
42: \renewcommand{\theequation}{\thesection.\arabic{equation}}
43:
44: %those command define the appendix with correct numbering in report style
45: \newcounter{appendice}
46: \newcommand{\appendice}
47: {
48: \setcounter{equation}{0}
49: \renewcommand{\theequation}{\Alph{appendice}.\arabic{equation}}
50: \addtocounter{appendice}{1}
51: {\bf Appendix \Alph{appendice}}
52: }
53:
54: \def\thebibliography#1{{\bf REFERENCES\markboth
55: {REFERENCES}{REFERENCES}}\list
56: {[\arabic{enumi}]}{\settowidth\labelwidth{[#1]}\leftmargin\labelwidth
57: \advance\leftmargin\labelsep
58: \usecounter{enumi}}
59: \def\newblock{\hskip .11em plus .33em minus -.07em}
60: \sloppy
61: \sfcode`\.=1000\relax}
62: \let\endthebibliography=\endlist
63:
64: \def\BI{{\rm 1\!l}}
65:
66: \begin{document}
67:
68: %\vspace*{5mm}
69:
70:
71:
72: \centerline{ \LARGE Can classical wormholes stabilize the
73: brane-anti-brane system? }
74:
75: \vskip 2cm
76:
77: \centerline{ {\sc A. Pinzul$^{a}$ and A. Stern$^{b}$ } }
78:
79: \vskip 1cm
80: \begin{center}
81: {\it a) Department of Physics, Syracuse University,\\ Syracuse,
82: New York 13244-1130, USA\\}
83: {\it b) Department of Physics, University of Alabama,\\
84: Tuscaloosa, Alabama 35487, USA}
85:
86: \end{center}
87:
88:
89:
90:
91:
92:
93:
94:
95:
96: \vskip 2cm
97:
98:
99:
100:
101: \vspace*{5mm}
102:
103: \normalsize
104: \centerline{\bf ABSTRACT}
105: We investigate the static
106: solutions of Callan and Maldecena and Gibbons
107: to lowest order Dirac-Born-Infeld theory. Among them are charged
108: wormhole solutions
109: connecting branes to anti-branes. It is seen that
110: there are no such solutions when the separation
111: between the brane and anti-brane is smaller than some
112: minimum value. The minimum
113: distance coincides with the energy minimum, and depends monotonically on the charge. Making the charge
114: sufficiently large, such that the minimum separation is much bigger than $
115: \sqrt{\alpha '}$, may suppress known quantum processes leading to decay of the
116: brane-anti-brane system. For this to be possible the zeroth order wormhole
117: solutions should be reasonable approximations of solutions in the
118: full $D-$brane theory. With this in mind we address the question of whether
119: the zeroth order solutions are stable under inclusion of higher order
120: corrections to the Dirac-Born-Infeld action.
121:
122: \vspace*{5mm}
123:
124: \newpage
125: \scrollmode
126:
127: \section{Introduction}
128:
129: The Born-Infeld nonlinear description of electrodynamics\cite{bi}
130: and its subsequent
131: generalization to membranes \'a la Dirac\cite{Dirac} is of current
132: interest due to its role as an effective theory for
133: D$p-$branes. The associated Dirac-Born-Infeld (DBI) action appears at lowest order in the derivative expansion
134: for the effective D$p$-brane action.\cite{Fradkin:1985},\cite{Johnson}
135: The original Born-Infeld theory has a charged static solution, which
136: was generalized by Callan and Maldecena\cite{Callan:1997kz}
137: and Gibbons\cite{Gib} to families of static solutions
138: on the brane. The families are associated with orbits of the
139: $SO(1,1)$ group. The Lagrangian is invariant under $SO(1,1)$ and
140: can be used to label the orbits. One such orbit contains the
141: solution of Born
142: and Infeld. Another is a BPS solution representing a fundamental string attached to the
143: brane. Finally, there is a third family of solutions
144: corresponding to wormholes which connect the brane to an anti-brane a
145: finite distance away. Here we show
146: that there are no charged wormhole solutions having a separation
147: between the $p-$brane and anti$-p-$brane smaller than some
148: minimum value. The minimum separation
149: distance goes like $|Q|^{\frac 1{p-1}}$, $Q$ being the $U(1)$
150: charge. At minimum separation, the self-energy
151: is also a minimum, where there appears a cusp singularity in the
152: plot of the
153: energy versus separation distance.
154: In the quantum analysis of the brane-anti-brane system an instability is known to occur at distance scales of order $ \sqrt{\alpha '}$ due to
155: excitation of tachyonic modes.\cite{Banks:1995ch}
156: Then for sufficiently large charge, i.e.
157: \be |Q|^{\frac 1{p-1}}>> \sqrt{\alpha '}\;,
158: \ee such quantum processes may be suppressed, and it is
159: possible that
160: charged wormholes can stabilize the brane-anti-brane
161: system.
162:
163: For the above scenario to be correct, however, classical stability of
164: the wormhole solutions should be checked. This means $a)$ enlarging to
165: time dependent solutions, and investigating whether solutions are
166: stable with respect to perturbations about the
167: static solution. It also means $b)$ checking whether the solutions to the zeroth order
168: effective theory are a reasonable approximation of solutions to the
169: full effective D$p-$brane action. Here we shall only consider $b)$.
170: One signal that solutions may be unstable in the sence $b)$ is the presence of singularities in the
171: field strength, where the derivative expansion cannot be trusted. Such a singularity is present for the original
172: Born-Infeld solution, in that case at a single point, and for the
173: entire orbit of solutions connected to the Born-Infeld solution. Despite the
174: singularity, these
175: solutions are associated with a
176: finite self-energy. For the BPS case the
177: singularity occurs an infinite distance away from the brane, and
178: appears harmless. The wormhole-type solutions were originally
179: constructed by joining together two local solutions, obtained in the static gauge,
180: at the minimum circumference of the wormhole.\cite{Callan:1997kz},\cite{Gib}
181: A singularity in the field strength occurs along the throat - precisely at the minimum
182: circumference. The singularity in the field strength is a
183: coordinate singularity, which can be removed by going to another gauge.
184: Nevertheless, it is a signal that higher order corrections may not be negligible.
185:
186:
187: To check stability in the sence $b)$, we will rely on recent
188: computations of the derivative corrections to Born-Infeld theory.
189: The first order corrections to the action were obtained separately by Wyllard \cite{Wyl} and
190: Das, Mukhi and Suryanarayana\cite{das}. [Higher order corrections
191: seem currently out of reach.] Using their results we
192: carried out a stability check previously for the case of the original Born-Infeld
193: solution.\cite{Karatheodoris:2002bb} There we argued that the original
194: Born-Infeld solution is unstable under inclusion of
195: these first order corrections. More specifically, we numerically obtained
196: corrections to the zeroth order Born-Infeld solution, but found that
197: they give an infinitely large correction to the Lagrangian.
198: We give a simpler proof of the result here. Because the Lagrangian is $SO(1,1)$
199: invariant the result applies to the entire orbit of solutions
200: connected to the Born-Infeld solution. Concerning the BPS solution,
201: it is known that the such a solution is stable {\bf to all
202: orders} in the derivative expansion.\cite{Thorlacius:1997zd} We
203: verify that this is consistent with the first order results of \cite{Wyl} and
204: \cite{das}. We find that the stability analysis for the wormhole solutions
205: leads to the same results obtained for the Born-Infeld solution.
206: Namely, corrections to the zeroth order solution lead to an
207: infinitely large correction of the Lagrangian. In this case, we need
208: to rely on numerical computation for the result.
209:
210:
211: In section 2 we
212: give analytic expressions for the three families of solutions along with their
213: self-energies. Here we show that the charged wormhole solutions
214: have a minimum length.
215: The question of stability of the zeroth order solutions is addressed in section 3. In appendix A we write down the wormhole solution in another gauge, where
216: the field strength is singularity-free. In fact, it is a constant in
217: that gauge. In appendix B we use the results of \cite{Wyl} and
218: \cite{das} to obtain the first order corrections to
219: the field equations for the BPS case, and show that the answer agrees with \cite{Thorlacius:1997zd}.
220:
221: \section{Zeroth Order Solutions}
222: \setcounter{equation}{0}
223: \subsection{Dirac-Born-Infeld theory}
224:
225: We consider the $p$-dimensional brane embedded in a ten dimensional
226: space-time with flat metric $[\eta_{AB}]=$diag$(-1,1,...,1)$, $A,B,..=0,1,...,9$. We denote the
227: brane coordinates by $X^A$. They are functions of $p+1$ parameters
228: $\xi^\mu$, $\mu,\nu,...=0,1,...,p$. Additional degrees
229: of freedom on the brane are $U(1)$ potentials ${\cal A}_\mu(\xi)$.
230: The DBI action is written in terms of the $(p+1)\times(p+1)$ matrix
231: \be h_{\mu\nu} =\eta_{AB}\;\partial_\mu
232: X^A(\xi)\;\partial_\nu X^B(\xi)+2\pi\alpha'\; F_{\mu\nu}(\xi) \;, \label{giffh}\ee
233: where $\partial_\mu =\frac{\partial}{\partial \xi^\mu}$. The first
234: term is the induced metric on the brane, while $
235: F_{\mu\nu}=\partial_\mu{\cal A}_\nu-\partial_\nu{\cal A}_\mu$ is the $U(1)$
236: field strength. We assume the two-form contribution is absent
237: $B_{AB}=0$. The DBI action is\cite{bi},\cite{Dirac}
238: \be {\cal S}^{(0)}_{DBI} = \frac{T_p}{ g_s} \int d^{p+1}\xi\; {\cal
239: L}^{(0)}_{DBI} \;,\qquad {\cal L}^{(0)}_{DBI} = 1-\sqrt{-\det [
240: h_{\mu\nu}]}\;, \label{Dpact} \ee where
241: $T_p$ is the tension, which is expressable in terms of $\alpha'$
242: according to \be T_p ={(4\pi^2 \alpha ')^{-(p+1)/2} }\;,\label{Tp}\ee and
243: $g_s$ is the string coupling.
244: $ {\cal S}^{(0)}_{DBI}$ is invariant under diffeomorphisms on the brane $\xi^\mu \rightarrow
245: \xi'^\mu(\xi)\;,$ $U(1)$ gauge transformations ${\cal A}_\mu(\xi) \rightarrow
246: {\cal A}_\mu(\xi) +\partial_\mu\Lambda(\xi) $, as well as
247: ten-dimensional Poincar\'e transformations. Variations in $X^A(\xi)$
248: and ${\cal A}_\mu(\xi)$ lead to the equations of motion
249: \beqa \partial_\mu\biggl\{ \sqrt{-\det
250: h}\;(h^{\mu\nu}+ h^{\nu\mu}) \eta_{AB} \partial_\nu X^B\biggr\}&=& 0 \cr & &\cr
251: \partial_\mu\biggl\{ \sqrt{-\det
252: h}\;( h^{\mu\nu}- h^{\nu\mu})\biggr\}&=& 0 \;,
253: \label{xaeom}\eeqa respectively,
254: where $h^{\mu\nu}h_{\nu\rho} = \delta ^\mu_\rho$.
255:
256:
257:
258: The known families of spherically
259: symmetric static solutions\cite{Callan:1997kz},\cite{Gib} can be
260: classified in terms of $SO(1,1)$ orbits (we do this below), and they
261: describe different topologies embedded in the flat
262: ten-dimensional background. For one family of solutions, containing the original
263: Born-Infeld solution,
264: a time slice is ${\mathbb{R}}^p$ minus a point. We
265: could therefore describe it with the introduction of a delta function source to the right-hand-side
266: of (\ref{xaeom}). Another family
267: corresponds to a
268: brane and anti-brane connected by a wormhole. In that case one can patch together local solutions to
269: (\ref{xaeom}). The families of solutions are written in terms of two
270: integration constants $Q$ and $C$, $Q$ being the electric charge.
271: Locally, all solutions can be expressed in the static gauge, where one identifies $\xi^\mu$ with the first
272: $p+1$ brane coordinates $X^\mu$, $\mu=0,1,...,p$. The remaining $X^\alpha$, $
273: \alpha= p+1,p+2,...,9$, then denote normal coordinates, and
274: (\ref{giffh}) becomes
275: \be h_{\mu\nu} =\eta_{\mu\nu} +
276: \;\partial_\mu X^\alpha\partial_\nu X^\alpha +2\pi\alpha'\; F_{\mu\nu}\;, \label{hisg}\ee
277: The static solutions of \cite{Callan:1997kz},\cite{Gib} are for a radial electric field
278: with a single transverse mode excited. Choose the nonvanishing degrees of freedom to be ${\cal
279: A}_0(r)$ and $X_{p+1}(r)$, where $r$ is the radial coordinate on the
280: brane. Since the metric is diagonal, the resulting matrix $h$ is diagonal except for the $2\times 2$
281: submatrix with corresponding indices $\mu$ and $\nu$ equal to $0$ and $r$. That $2\times 2$
282: submatrix and its inverse are given by
283: \be \pmatrix{-1 & -f(r) \cr f(r) &1+g(r)^2 \cr} \qquad{\rm and}\qquad
284: \frac1{1+g(r)^2 -f(r)^2} \pmatrix{-1-g(r)^2 & -f(r) \cr f(r) &1
285: \cr}\;, \ee respectively, where $f(r)=2\pi\alpha'\partial_r{\cal
286: A}_0(r)$ and $g(r) =\partial_rX_{p+1}(r)$. Substituting into
287: the equations of motion (\ref{xaeom}) gives
288: \be \partial_r\biggl\{\frac {r^{p-1}f(r)}{\sqrt{1+g(r)^2 -f(r)^2}}\biggr\} =
289: \partial_r\biggl\{\frac {r^{p-1}g(r)}{\sqrt{1+g(r)^2 -f(r)^2}}\biggr\} =0 \;\ee
290: The solutions for $f(r)$ and $g(r)$ are
291: \be \frac{f(r)}Q = \frac{g(r)}C = \frac{1}{\sqrt{Q^2 -C^2
292: +r^{2p-2}}} \label{sln0} \ee
293: The integration constants $Q$ and $C$ have units of [length$]^{p-1}$.
294:
295: For the configurations (\ref{sln0}) the Lagrangian and equations of motion are invariant under the
296: $SO(1,1)$ transformation
297: \be \pmatrix{ f(r) \cr g(r)}\rightarrow \pmatrix{\cosh\nu &\sinh \nu\cr
298: \sinh\nu &\cosh \nu\cr}\pmatrix{ f(r) \cr g(r)}\label{so11}\ee
299: The integration constants transform in the same
300: way \be \pmatrix{ Q \cr C}\rightarrow \pmatrix{\cosh\nu &\sinh \nu\cr
301: \sinh\nu &\cosh \nu\cr}\pmatrix{ Q \cr C}\label{so11qc}\ee
302: There are then three kinds of orbits: $i)\;|Q|>|C|$, $ii)\;|Q|=|C|$ and $iii)\;|Q|<|C|$. $i)$ is connected to the original
303: Born-Infeld solution, $ii)$ is the BPS solution and $iii)$ is associated
304: with wormhole solutions. The orbits can be classified by their corresponding value for the spatial
305: integral of the Lagrangian density $ {\cal
306: L}^{(0)}_{DBI}$
307: \be L^{(0)}_{DBI}(r)=- \Omega_{p-1}\biggl\{\int^\infty_r dr
308: \frac{ r^{2p-2} } {\sqrt{ r^{2p-2} +Q^2-C^2}}-\int^\infty_0 drr^{p-1} \biggr\}
309: \;, \ee where $ \Omega_{p-1}$ is the volume of a unit
310: $p-1$ sphere. We removed a hole around the origin of
311: radius $r$ in the integration for the first integral, in order to
312: accommodate the different cases, as
313: their domains differ. The second integral is the vacuum subtraction. The first integral
314: can be expressed in terms of the one for $X_{p+1}(r)$, yielding
315: \be L^{(0)}_{DBI}(r)=\frac { \Omega_{p-1}}p
316: \biggl\{ r\sqrt{ r^{2p-2} +Q^2-C^2}\; +\; (C^2-Q^2)\frac{X_{p+1}(r)}C
317: \biggr\}
318: \label{lfbih} \ee For cases $i),\; ii)$ and $iii)$ we find that the
319: result (after setting $r$ to its appropriate value) is positive, zero and negative, respectively.
320:
321: \subsection{Analytic solutions}
322:
323: The right hand side of (\ref{sln0}) can be integrated to obtain
324: analytic expressions for the potential $ {\cal A}_{0}(r)$ and the
325: transverse displacement $
326: X_{p+1}(r)$. For this expand in powers of $ \frac
327: {r^{2p-2}}{Q^2-C^2}$ and integrate term by term. The result for the
328: indefinite integral
329: can be expressed in terms of a hypergeometric
330: function:\be
331: \frac {r}{\sqrt{Q^2-C^2}}\;
332: F\biggl(\frac 12,\frac 1 {2p-2};\frac {2p-1} {2p-2};- \frac
333: {r^{2p-2}}{Q^2-C^2} \biggr)\;,\label{indef}\ee for $ |\frac
334: {r^{2p-2}}{Q^2-C^2}|<1$.
335: Now set the limits of integration to be $r$ and $\infty$, with the
336: assumption that the potentials vanish at the latter. To evaluate
337: (\ref{indef}) for these limits we analytic
338: continue $F(a,b;c;z)$ using\cite{Bateman}
339: \beqa F(a,b;c;z)&=&\frac {\Gamma(c)\Gamma(b-a)}
340: {\Gamma(b)\Gamma(c-a)}(-z)^{-a}\;F(a,1-c+a;1-b+a;\frac1z)\cr
341: &+&\frac {\Gamma(c)\Gamma(a-b)}
342: {\Gamma(a)\Gamma(c-b)}(-z)^{-b}\;F(b,1-c+b;1-a+b;\frac1z)\label{idfhyg}\eeqa
343: where $|{\rm arg}(-z)|<\pi$. So (\ref{indef}) can be rewritten as
344: $$ \frac1{(2-p)r^{p-2}}\;F\biggl(\frac12,
345: \frac{p-2}{2p-2};\frac{3p-4}{2p-2};-\frac{Q^2-C^2}
346: {r^{2p-2}} \biggr)\;+\;
347: \frac{(Q^2-C^2)^{\frac{2-p}{2p-2}}}{\sqrt\pi}\;\Gamma\biggl(\frac{2p-1}{2p-2}\biggr)
348: \Gamma\biggl(\frac{p-2}{2p-2}\biggr)\;,
349: $$ since $F(a,0;c;z)=1$. For $p>2$, only the last term survives when evaluating
350: at $\infty$, which then get subtracted out after evaluating
351: between $r$ and $\infty$. For the transverse displacement
352: $X_{p+1}$ one gets
353: \be
354: X_{p+1}(r) = \frac{- C} { (p-2)r^{p-2}}\;F\biggl(\frac12,
355: \frac{p-2}{2p-2};\frac{3p-4}{2p-2};-\frac{Q^2-C^2}
356: {r^{2p-2}} \biggr), \qquad{\rm for}\;p>2\label{slfx}
357: \ee
358:
359: Next we examine the result for the three different orbits:
360:
361: Case $i)\;|Q|>|C|$.
362: The limit $C\rightarrow 0$ where the transverse mode is not excited gives the
363: original Born-Infeld solution\cite{bi}, while $|Q|> |C|>0$ yields a deformation
364: of the Born-Infeld solution where a spike protrudes from the brane. We
365: plot below the function $X_{p+1}(r) $ for $p=3$ on a two dimensional
366: spatial slice:
367:
368: \bigskip
369: \input epsf
370: \def\epsfsize#1#2{0.8#1}
371: \centerline{\hss{\epsffile{spike.eps}}}
372:
373: \medskip
374: \centerline{\qquad\qquad\qquad
375: ${\tt fig. 1}\qquad p=3\qquad |Q|=1\quad |C|=.8$}
376: \noindent
377: >From (\ref{slfx}) [and (\ref{idfhyg})] the maximum size of the spike is the
378: absolute value of
379: \be X_{p+1}(0) =\frac{C \; \Gamma(\frac{3p-4}{2p-2})
380: \Gamma(\frac{1}{2p-2}) } { (2-p)\sqrt{\pi} }\;(Q^2 -C^2)^{\frac{2-p}{2p-2}}\;\;,
381: \qquad{\rm for}\;p>2 \label{xo}\ee For the integral of the
382: Lagrangian density [C.f. (\ref{lfbih})]
383: one gets \be
384: L^{(0)}_{DBI}(0)=\frac{\Omega_{p-1} \;
385: \Gamma(\frac{3p-4}{2p-2} )\Gamma(\frac 1 {2p-2}) }{p(p-2)\sqrt{\pi}
386: }\;(Q^2 -C^2)^{ \frac p {2p-2} }\; >\;0\;,
387: \qquad{\rm for}\;p>2
388: \ee
389:
390: Case $ii) \;|Q|=|C|$. One
391: arrives at the BPS solution for this case, where (\ref{sln0}) reduce to the Coulumb solutions $f(r) = g(r) =
392: \frac{Q}{r^{p-1}}$. Since $F(a,b;c;0)=1$, (\ref{slfx}) reduces to \be
393: X_{p+1}(r) = \frac{- Q}{ (p-2)\; r^{p-2}}\;, \qquad {\rm
394: for}\;p> 2\ee and
395: the spike
396: becomes infinitely long,
397:
398: \bigskip
399: \input epsf
400: \def\epsfsize#1#2{0.8#1}
401: \centerline{\hss{\epsffile{bps.eps}}}
402:
403: \medskip
404: \centerline{\qquad\qquad \qquad
405: ${\tt fig. 2}\qquad p=3\qquad |Q|= |C|=1$}
406: \noindent
407: representing a fundamental string attached to the
408: brane.\cite{Callan:1997kz} In
409: this case the integral of the
410: Lagrangian density
411: $ L^{(0)}_{DBI}(0)$ goes to zero.
412: These solutions are BPS because they
413: preserve
414: half of the supersymmetries of the ground state solution.
415: Supersymmetries are present when the matrix \be\partial_\mu
416: X_{p+1}(\xi)\;[\Gamma^\mu,\Gamma^{p+1}]\;+2\pi\alpha'\;F_{\mu\nu}(\xi)\;[\Gamma^\mu,\Gamma^\nu]\label{susycd}\ee
417: is degenerate. $\Gamma^A$ are $\Gamma$ matrices for the
418: ten-dimensional background space, $[\Gamma^A,\Gamma^B] = 2\eta^{AB}$.
419: To see that this holds when $|Q|=|C|$, one observes that
420: (\ref{susycd}) is proportional to $(Q\Gamma^0 +
421: C\Gamma^{p+1})\Gamma^r$, whose square is
422: $(C^2-Q^2)\BI$. (\ref{susycd}) is then nilpotent when $|Q|=|C|$.
423:
424: Case $iii)\;|Q|<|C|$. Here one gets a finite diameter tube with a minimum radius $r_0 =
425: (C^2-Q^2)^{\frac1{2p-2}}$:
426:
427: \bigskip
428: \input epsf
429: \def\epsfsize#1#2{0.8#1}
430: \centerline{\hss{\epsffile{tube.eps}}}
431:
432: \centerline{\qquad\qquad\qquad
433: ${\tt fig. 3}\qquad p=3\qquad |Q|=1\quad |C|=1.2$}
434: \noindent
435: Both
436: $ g$ and $f$, and consequently also the electric field, are singular at
437: $r=r_0$. Nevertheless, ${\cal A}_0$ and $ X_{p+1}$ are not. From
438: the latter the
439: tube has a finite length. After
440: expressing $C$ in terms of $r_0$ and $Q$, it is
441: \be X_{p+1}(r_0) = \frac
442: {\sqrt{\pi(r_0^{2p-2}+Q^2)}}{ (2-p)\;r_0^{p-2}}\;\frac{
443: \Gamma(\frac{2p-1}{2p-2} )}{
444: \Gamma(\frac{p}{2p-2})}\;, \qquad{\rm for}\;p>2 \;,\label{xro}\ee
445: The domain of integration for the integral of the
446: Lagrangian density [C.f. (\ref{lfbih})] now goes from $r_0$ to
447: $\infty$. One gets \be
448: L^{(0)}_{DBI}(r_0)=-\frac{\Omega_{p-1} \;\sqrt{\pi}}{p(p-2)}
449: \;\frac{
450: \Gamma(\frac{2p-1}{2p-2} )}{
451: \Gamma(\frac{p}{2p-2})}\;r_0^p \;< \;0\;,
452: \qquad{\rm for}\;p>2
453: \ee
454:
455:
456: The static gauge breaks down at $r=r_0$, and so the above
457: solution is only local. A global solution was proposed by gluing
458: this one to the analogous solution on an anti-brane.\cite{Callan:1997kz},\cite{Gib} The
459: global solution
460: then represents a wormhole connecting the brane with an anti-brane a distance of
461: $2|X_{p+1}(r_0)|$ away, with a
462: throat of minimum radius $r_0$ .
463: The gluing of the two local solutions to form a wormhole occurs at
464: $r=r_0$, precisely where there
465: is a singularity
466: in the electric field, which might be a matter of concern.
467: On the other hand, the electric field singularity is a coordinate singularity, which is easily seen by
468: transforming to another gauge. Take for example the gauge where the
469: $r-$coordinate of the static gauge is replaced by $z=X_{p+1}$. In
470: the new gauge the
471: solution
472: is described by the inverse, call it $R(z)$, of the function
473: $X_{p+1}(r)$. Now the electrostatic field is in the $z-$direction,
474: and there are
475: coordinate singularities at the location of the brane (and
476: anti-brane).
477: We denote the electric field ($\times\; 2\pi\alpha'$) in the new gauge
478: by $E(z)$. It can be computed locally by
479: performing a coordinate transformation from the static gauge,
480: \be E(z) = \frac{\partial R}{\partial z}\; f(R(z)) = \frac 1{g(R(z))}\; f(R(z)) \ee
481: Substituting in the solution for $f$ and $g$ given in
482: (\ref{sln0}) gives a constant electric field
483: \be E(z) = Q/C\; \label{eqoc}\ee
484: So in this
485: coordinate frame there are no
486: singularities in the electric field (for $C\ne 0$).
487: In appendix A we write down the
488: action in this gauge and show that (\ref{eqoc}) solves the
489: corresponding equations of motion.
490:
491: >From (\ref{xro}) it follows that there is a minimum separation distance
492: between the brane and anti-brane for a fixed $Q$ (and $p>2$). It is equal to \be \min \; 2| X_{p+1}(r_0)| = \frac
493: {2 \sqrt{\pi(p-1)}} { (p-2)^{\frac{3p-4}{2p-2}}}\;\; \frac{
494: \Gamma(\frac{2p-1}{2p-2} )}{
495: \Gamma(\frac{p}{2p-2})}\; |Q|^{\frac 1{p-1}}
496: \;, \label{mindis}\ee and occurs when $C$ and $Q$ are
497: constrained by
498: \be C^2 = (p-1) Q^2\ee
499: Below we plot the separation distance versus
500: throat size for a fixed $Q$ when $p=3$:
501:
502: \bigskip
503: \input epsf
504: \def\epsfsize#1#2{0.8#1}
505: \centerline{\hss{\epsffile{sepvsro.eps}}}
506:
507: \medskip
508: \centerline{\qquad\qquad\qquad
509: ${\tt fig. 4}\qquad p=3\qquad Q=1\qquad 2|X_{p+1}(r_0)|\quad{\rm
510: vs.}\quad r_0 $}
511: \noindent
512:
513:
514: \subsection{Self-Energy}
515:
516: Concerning the energy, one can apply the canonical formalism starting
517: from the Lagrangian in (\ref{Dpact}). For this again assume the
518: static gauge and hence (\ref{hisg}). The Hamiltonian density is
519: \be {\cal H}^{(0)}_{DBI} =P^\alpha \dot X_{\alpha} + \Pi^\mu \dot {\cal A}_\mu - {\cal
520: L}^{(0)}_{DBI}\;,\ee where the dot denotes a
521: time derivative and
522: \beqa P^\alpha &=& -\frac12 \sqrt{-\det h}\; (h^{0\mu}+h^{\mu 0}) \partial_\mu X ^\alpha
523: \cr & &\cr
524: \Pi^\mu &=& -\pi\alpha' \sqrt{-\det h}\; (h^{\mu 0}-h^{0\mu})
525: \;,\label{mca}\eeqa are the momenta conjugate to $ X_\alpha$ and ${\cal
526: A}_\mu$, respectively. As usual, the momentum
527: conjugate to ${\cal
528: A}_0$ is constrained to be zero. After integrating by parts
529: \be {\cal H}^{(0)}_{DBI} =
530: P^\alpha \dot X_{\alpha} + \Pi^i F_{0i} +\sqrt{-\det h}-1 -\partial_i
531: \Pi^i {\cal A}_0 \;, \label{ham}\ee where $i=1,2,...,p$. The
532: coefficient of $ {\cal A}_0$ gives the Gauss law constraint. The
533: remaining terms are equal to
534: \be - \sqrt{-\det h}\; h^{00} \; -\;1\;\label{slfngydn} \ee
535: Although the Lagrangian density is invariant under the $SO(1,1)$
536: symmetry (\ref{so11}), the Hamiltonian density is not. Then unlike the
537: integral of the Lagrangian density, the integral of the energy density will not be
538: constant along the orbits of $SO(1,1)$.
539:
540: The
541: integral of (\ref{slfngydn}) gives the self-energy of the
542: DBI solutions (\ref{sln0}).
543: After removing a hole around the origin of
544: radius $r$ in the integration domain one gets
545: \be {\cal U}(r) = \frac
546: {T_p \Omega_{p-1}} {g_s}\; {\cal E}(r)\;,\qquad {\cal E}(r) =\int^\infty_r dr
547: \frac{( r^{2p-2} +{Q^2})} {\sqrt{ r^{2p-2} +Q^2-C^2}}-\int^\infty_0 drr^{p-1}
548: \;,\label{efbih} \ee where the factor $T_p/g_s$ comes from (\ref{Dpact}).
549: In the second integral we subtract off the total vacuum energy of
550: the brane. Note that we must restrict the lower limit in the
551: first integral to be greater than or equal to $ r_0$ for the case
552: $|C|>|Q|$. The result can be expressed in terms of $X_{p+1}(r)$:
553: \be {\cal E}(r)\; = \;-[(p-1)Q^2 +C^2]\; \frac {X_{p+1}(r)}{ p\;C} -
554: \frac rp
555: \sqrt{r^{2p-2} +Q^2-C^2} \label{Eitor}
556: \ee This gives a positive answer for the self-energy since ${X_{p+1}(r)}/{C}$ is negative for the solutions, while the second term vanishes
557: after evaluating at the minimum value of $r$ ($0$ for $|Q|\ge |C|$
558: and $r_0$ for $|C|>|Q|$).
559:
560:
561: For case $i)\;|Q|> |C|$, one gets the total self-energy of the
562: solution by setting $r$ in (\ref{Eitor}) to zero, which yields
563: \be {\cal U}(0)\; =\; \frac
564: {T_p \Omega_{p-1}} {g_s}\; \frac{ \Gamma(\frac{3p-4}{2p-2}
565: )\Gamma(\frac 1{2p-2}) } {p (p-2)\sqrt{\pi}} \;
566: \frac{ (p-1)Q^2 +C^2} { (Q^2 -C^2)^{\frac{p-2}{2p-2} } }\;, \qquad{\rm for}\;p>2
567: \label{spen}\ee For a fixed $Q$ it goes monotonically from the
568: Born-Infeld value \be {\cal U}_{\rm Born-Infeld}\; = \; \frac
569: {T_p \Omega_{p-1}} {g_s}\; \frac{p-1 } {p (p-2)\sqrt{\pi}} \; \Gamma\biggl(\frac{3p-4}{2p-2}
570: \biggr)\Gamma\biggl(\frac 1{2p-2}\biggr) \; { |Q|^{\frac{p}{p-1} } }\;, \qquad{\rm for}\;p>2
571: \label{biengy}\ee corresponding to $C=0$, to
572: infinity in the BPS limit, corresponding to $|C|\rightarrow |Q|$.
573: We plot below ${\cal E}$ for $Q=1$ and $p=3$:
574:
575: \bigskip
576: \input epsf
577: \def\epsfsize#1#2{0.8#1}
578: \centerline{\hss{\epsffile{energy1.eps}}}
579:
580: \medskip
581: \centerline{\qquad\qquad\qquad
582: ${\tt fig. 5}\qquad p=3\qquad Q=1\qquad {\cal E}\quad{\rm
583: vs.}\quad C\quad(C<Q) $}
584: \noindent
585:
586: For case $ii)\;|Q|=|C|$ the total self-energy is infinite. At large distances $|X_{p+1}|$,
587: the energy per unit length of the infinite string solution is
588: constant. From (\ref{Eitor}) it follows that \be \frac{d {\cal
589: U}}{d|X_{p+1}|}\rightarrow \frac
590: {T_p \Omega_{p-1}} {g_s}\; |Q|\;,\qquad{\rm as} \quad|X_{p+1}| \rightarrow \infty\ee
591:
592:
593: For case $iii)\;|Q|< |C|$, one gets the total self-energy by setting $r$ in (\ref{Eitor}) equal to $r_0$:
594: \be {\cal U}(r_0)\; = \frac
595: {T_p \Omega_{p-1}} {g_s}\;\frac
596: {\sqrt{\pi}} {p-2}\; \frac{
597: \Gamma(\frac{2p-1}{2p-2} )}{
598: \Gamma(\frac{p}{2p-2})} \; \biggl(\frac{Q^2}{r_0^{p-2}} + \frac {r_0^p}p \biggr)\;, \qquad{\rm for}\;p>2 \;\label{enxro}\ee
599: For a fixed $Q$, the minimum energy configuration occurs for $r_0^{2p-2} = (p-2)
600: Q^2$, corresponding to the minimum separation distance (\ref{mindis})
601: between the brane and anti-brane.
602: The minimum value for $ {\cal U}(r_0)$ is \be {\cal U}_{\rm min}\; = \frac
603: {T_p \Omega_{p-1}} {g_s}\;\; \frac
604: {2\sqrt{\pi}(p-1)} {p(p-2)^{\frac{3p-4}{2p-2}}}\; \frac{
605: \Gamma(\frac{2p-1}{2p-2} )}{
606: \Gamma(\frac{p}{2p-2})} \; |Q|^{\frac{p}{p-1}}\;, \qquad{\rm
607: for}\;p>2\label{wrhlen}\ee If $Q=0$ the minimum energy configuration
608: occurs when the brane and anti-brane coincide. For $Q\ne 0$ and a
609: separation distance greater than the minimum value
610: (\ref{mindis}), there are two possible solutions with different
611: throat sizes. The one
612: with a smaller throat is energetically favored. Call $ {\cal U}_0(X_{p+1})$ and $ {\cal U}_1(X_{p+1})$ the energy of the
613: thin and fat wormholes, respectively. For a large separation distance,
614: \beqa {\cal U }_0(X_{p+1}) &\rightarrow & \frac
615: {T_p \Omega_{p-1}} {g_s}\;\;|QX_{p+1}| \cr & &\cr
616: {\cal U}_1(X_{p+1})&\rightarrow & \frac
617: {T_p \Omega_{p-1}} {pg_s}\;\;\biggr\{\frac
618: {\sqrt{\pi}} {p-2}\; \frac{
619: \Gamma(\frac{2p-1}{2p-2} )}{
620: \Gamma(\frac{p}{2p-2})}\biggl\}^{1-p} \;|X_{p+1}|^p \;
621: \eeqa Upon plotting the
622: energy versus the separation distance one gets a double-valued
623: function, with a cusp at the minimum
624: separation, as is illustrated below for $Q=1$ and $p=3$:
625:
626: \bigskip
627: \input epsf
628: \def\epsfsize#1#2{0.8#1}
629: \centerline{\hss{\epsffile{energy2.eps}}}
630:
631: \medskip
632: \centerline{\qquad\qquad\qquad
633: ${\tt fig. 6}\qquad p=3\qquad Q=1\qquad {\cal E}\quad{\rm
634: vs.}\quad 2 |X_{p+1}(r_0)| $}
635:
636:
637:
638: The minimum energy solution for fixed $Q$
639: in case $i) \;|Q|> |C|$ was the original Born-Infeld solution, while
640: in case $iii)\;|Q|< |C|$ it corresponded to (\ref{wrhlen}). In both cases
641: the energy goes like $ |Q|^{\frac{p}{p-1}}$.
642: Assuming charge
643: conservation, such solutions are energetically unstable under fission into far separated solutions with total
644: charge equal to $Q$. It was however pointed out in \cite{Gib} that
645: fission may not be realized at the classical level for singular field configurations,
646: and the above solutions are of this type. Assuming fission does
647: occur, either classically or quantum mechanically the minimum energy configuration should
648: be an ensemble of far separated wormholes in case $iii)$ or
649: Born-Infeld solutions in case $i)$ with the fundamental
650: charge. In comparing $i)$
651: with $iii)$, the ratio $ {\cal U}_{\rm min}/ {\cal U}_{\rm Born-Infeld}$ is
652: less than one for $p\ge 4$, while it is greater than one for $p=3$.
653: Thus for $p\ge 4$, it is energetically more favorable for wormholes
654: to develop between a charged brane and equally charged
655: anti-brane than for Born-Infeld configurations to develop on the brane
656: and anti-brane. The opposite is true for $p=3$.
657:
658: \subsection{Thermodynamic considerations}
659:
660: Here we make a side remark concerning the thermodynamics of wormholes.
661: Once again, for case $iii)$ when the energy is greater
662: than the minimum, two types of wormholes with different thickness may be present. Say they
663: are in a heat bath with temperature $T$ and call $ \rho_0$ and
664: $ \rho_1$ the density of the
665: thin and fat wormholes, respectively. If one assumes they are
666: in dissipative and thermal equilibrium, then the ratio
667: of their densities at a temperature $T$ is given by \be \frac{\rho_1}{\rho_0} = \exp
668: \frac{{\cal U}_0(X_{p+1}) -{\cal U}_1(X_{p+1})}{k_B T}\ee
669:
670:
671:
672: \section{Inclusion of First Order Corrections}
673: \setcounter{equation}{0}
674:
675: Here we examine what happens to the zeroth order classical solutions upon
676: including the first order derivative corrections in the action. We
677: already checked in \cite{Karatheodoris:2002bb} that
678: the zeroth order Born-Infeld solution ($C=0$) does not survive upon
679: the inclusion of such corrections. More specifically, we numerically
680: found
681: a classical solution to the corrected field equations, but
682: it was associated with an infinite value
683: for the Lagrangian. Because as with zeroth order, the Lagrangian is $SO(1,1)$
684: invariant, the result that the of an infinite value for the Lagrangian follows for the entire orbit of solutions
685: connected to the $C=0$ solution; i.e. case $i)$. On the other hand,
686: the case $ii)$ BPS solution ($|Q|=|C|$) is stable upon inclusion of
687: the first order corrections, and just like at zeroth order,
688: the Lagrangian vanishes. In fact the BPS solution is known to survive to all
689: orders in the derivative expansion.\cite{Thorlacius:1997zd} We shall
690: verify that this result is consistent with the explicit expression for
691: the first order terms obtained in \cite{Wyl},\cite{das}. The
692: stability analysis for the wormhole solutions case $iii)$
693: leads to the same results we obtain for case $i)$.
694: Namely, corrections to the zeroth order solution lead to an
695: infinitely large correction of the Lagrangian.
696:
697: The first
698: order corrections were initially computed in \cite{Wyl},\cite{das} for the space-filling
699: D$9-$brane. A dimensional reduction could then
700: be performed to get the corrections to the DBI action (\ref{Dpact})
701: for an arbitrary D$p-$brane. We first briefly recall the results of
702: the dimensional reduction procedure at
703: zeroth order.\cite{Johnson} One starts with the Born-Infeld (BI) action ${\cal S}^{(0)}_{BI}$ for the space-filling
704: D$9-$brane. It is written in terms of a $10\times 10$ matrix $\tilde h$ with elements
705: \be \tilde h_{AB} =\eta_{AB} +2\pi\alpha'\; F_{AB}\;,\qquad A,B,...=0,1,...,9\label{hAB}\ee where $
706: F_{AB}=\partial_A{\cal A}_B-\partial_B{\cal A}_A$ is the ten
707: dimensional field
708: strength and we again assume a flat background metric
709: $\eta_{AB}$. $\tilde h$ in
710: (\ref{hAB}) can be obtained from $h$ in (\ref{giffh}) by assuming the
711: static gauge, which here means
712: $X^A=\xi^A$, for all $A$. ${\cal S}^{(0)}_{BI}$
713: is given by\be {\cal S}^{(0)}_{BI} = \frac{T_9}{ g_s} \int d^{10}\xi\; {\cal L}^{(0)}_{BI} \;,\qquad {\cal L}^{(0)}_{BI} = 1-
714: \sqrt{-\det [\tilde h_{AB}]} \label{biac}\;, \ee
715: and from (\ref{Tp}), $T_9 =1/{(4\pi^2 \alpha ')^5 }$.
716: In dimensional reduction
717: to the D$p-$brane, $9-p$ of the nine spatial directions are
718: `T-dualized'. Choose the T-dual directions to be $A=
719: \alpha= p+1,p+2,...,9$. One of the consequence of this procedure, is that the
720: gauge potentials ${\cal A}_\alpha $ in the T-dual directions get
721: replaced with the transverse modes $X_\alpha$ of the D$p-$brane
722: according to \be 2\pi\alpha'{\cal A}_\alpha \rightarrow X_\alpha\;.\ee
723: The fundamental degrees of
724: freedom are then
725: $X_\alpha$ and the remaining $p+1$ gauge potentials ${\cal A}_\mu$,
726: $\mu=0,1,...,p$, which are functions of the $p+1$
727: coordinates of the brane $\xi^\mu$. Then the nonvanishing matrix elements
728: of $\tilde h$ are $ \tilde h_{\mu\nu}$ and they are identical to $
729: h_{\mu\nu}$ of the DBI action written in the static gauge and
730: given in (\ref{hisg}).
731: Finally after performing integrations in the T-dual directions (\ref{biac}) gets replaced by \be {\cal S}^{(0)}_{BI} = \frac{T_p}{ g_s} \int d^{p+1}\xi\;\biggl( 1-\sqrt{-\det [\tilde h_{\mu\nu}]}\biggr) \ee
732: where the D$p-$brane
733: tension is again given by (\ref{Tp}), and one recovers (\ref{Dpact}) in the
734: static gauge.
735: So instead of working with $ h_{\mu\nu}$ as we did in the previous
736: section we could have started with the $10\times 10$ matrix $ \tilde h$.
737: Then for the static spherically symmetric solutions of the previous
738: section where just the $p+1$
739: transverse mode is excited, $$ 2\pi \alpha'F_{0i}=f \hat r_i\qquad 2\pi \alpha'F_{p+1\;i}=g \hat r_i \;,\qquad
740: i=1,2,...p\;,$$ where $\hat r$ is the unit vector in the radial direction and
741: spherical symmetry means $ f $ and $g $ are only functions of the
742: radial variable $r$. The $10\times 10$ matrix
743: $ \tilde h$ takes the form
744: \be \tilde h= \pmatrix{-1 & - f \hat r & & \cr
745: f \hat r &\BI_{p\times p} & g \hat r & \cr
746: &- g \hat r & 1 & \cr & & & &\BI_{(8-p)\times(8-
747: p)}\cr}\;\label{hinv1}\ee
748:
749:
750:
751:
752: The first
753: order corrections ${\cal S}^{(1)}_{BI}$ to the action $ {\cal
754: S}^{(0)}_{BI}$ of the space-filling D$9-$brane obtained in \cite{Wyl},\cite{das}
755: involve first
756: and second derivatives of the field strength
757: $F_{AB}$. They are contained in the rank-$4$ tensor
758: \be S_{ABCD} =2\pi \alpha'\partial_A \partial_B
759: F_{CD}
760: +(2\pi \alpha')^2 \tilde h^{EG}( \partial_A F_{CE} \partial_B
761: F_{DG} - \partial_A F_{D E} \partial_B
762: F_{CG} ) \;, \ee which is antisymmetric in the last two indices.
763: Here $\tilde h^{AB}\tilde h_{BC}= \delta^A_C$.
764: The total action is
765: $${\cal S}^{(0)}_{BI} + {\cal S}^{(1)}_{BI} = \frac{T_9}{ g_s} \int
766: d^{10}x\;\biggl\{ 1-
767: \sqrt{-\det [\tilde h_{AB}]} \;\biggl(1 +\frac \kappa 4 \Delta\biggl)\biggr\} \;, $$
768: \be \Delta =\; \tilde h^{AB} \tilde h^{CD}\tilde h^{EG} \tilde h^{IJ}(
769: S_{B CEG} S_{D AIJ} - 2 S_{GI BC}
770: S_{J E D A} )\;,\label{Dlta} \ee where $\kappa ={{
771: (2\pi\alpha')^2}\over {48}} $.
772: We again specialize to the case
773: where a single transverse mode [the $(p+1)^{\rm th}$ mode] is excited on
774: a $p\le 8$ brane, and consider static spherically symmetric fields. So the ansatz for $\tilde h$ is again (\ref{hinv1}).
775: Its determinant and inverse are given
776: by
777: \be \det\tilde h= -1 + f^2 - g^2 \label{deth}\ee and
778: \be \tilde h^{-1}=\frac1{\det\tilde h} \pmatrix{1+ g^2 & \hat r f &
779: - f g &\cr -\hat r f & -\BI + (f^2-g^2) P & \hat r g & \cr
780: - f g
781: &- \hat r g & f^2-1 & \cr & & &\det\tilde h \;\BI_{(8-p)\times(8- p)} \cr
782: }\;,\ee respectively. $P$ is the projection matrix $P_{ij} = \delta_{ij}
783: -\hat r_i \hat r_j$, satisfying $P_{ij} \hat r_j=0$ and $P_{ij}P_{jk}
784: =P_{ik}$. Some work shows that the nonvanishing components of $ S_{ABCD}$ are
785: \beqa
786: S_{ijk0}=- S_{ij0k}&=& \det \tilde h\; {H_f}'\; \hat r_i\hat r_j \hat
787: r_k -\frac1r \biggl(H_f+\frac fr\biggr)(P_{ik}\hat r_j
788: +P_{jk}\hat r_i) + \biggl(\frac fr \biggr)' P_{ij}\hat r_k \cr & &\cr
789: S_{ijk\;p+1}=- S_{ij\;p+1\;k}&=&\det \tilde h\;{ H_g}'\; \hat r_i\hat r_j \hat
790: r_k -\frac1r \biggl(H_g +\frac gr \biggr)(P_{ik}\hat r_j
791: +P_{jk}\hat r_i) + \biggl(\frac gr \biggr)' P_{ij}\hat r_k \cr & &\cr
792: S_{ijk\ell}&=& \frac{(\ln \det \tilde h)'}{2r}(P_{ik} \hat r_j \hat r_\ell
793: +
794: P_{j\ell} \hat r_i \hat r_k -P_{i\ell} \hat r_j \hat r_k-P_{jk} \hat r_i \hat r_\ell)\cr & &\cr & &
795: +\;\frac{1 +(\det\tilde h)^{-1}}{r^2} (P_{ik} P_{j\ell} -
796: P_{i\ell}P_{jk})
797: \;,
798: \eeqa
799: where \be H_f = \frac{f'}{\det\tilde h}\;,\qquad H_g =
800: \frac{g'}{\det\tilde h}
801: \ee
802: the prime here denoting derivatives in $r$. In addition we define
803: $$ H_k= f H_g - g H_f $$
804: Substituting into the formula in (\ref{Dlta}) for $\Delta$ gives
805: \beqa \Xi = - \frac14 (\det \tilde h)^2 \Delta &=& {H_f}'^2 - {H_g}'^2 +
806: {H_k}'^2 \cr& &\cr
807: & + &\frac{p-1}{r^2}\biggl\{\biggl(2+(\det\tilde h)^2 \biggr)\;({H_f}^2 - {H_g}^2 + {H_k}^2) \cr & & \cr
808: & +&\frac1{2} (\ln\det\tilde h)'^2 +\frac2{r} (\ln\det\tilde h)' -\frac1{r}
809: (\det\tilde h)' \cr& &\cr &
810: +&
811: \frac1{r^2}(1+\det\tilde h)\biggl( p+1+(p-2)\det\tilde h\biggr)\biggr\}\;\eeqa
812: So for the above ansatz the correction to the zeroth order
813: Lagrangian density ${\cal L}^{(0)}_{DBI}$ in (\ref{Dpact}) is
814: \be {\cal L}^{(1)}_{DBI} = \frac{ \kappa\;\Xi}{(-\det\tilde h)^{3/2}} \;
815: \label{radL}\ee
816:
817:
818: In obtaining the equations of motion one
819: must again write $f(r)$ and $g(r)$ in terms of potentials and extremize with
820: respect to the latter. As the general system is quite involved, below we shall restrict to functions
821: $f(r)$ and $g(r)$ which are related by a constant factor
822: \be \frac{f(r)}Q = \frac{g(r)}C\;, \ee as what occurred for the
823: zeroth order solutions (\ref{sln0}). This
824: set of configurations
825: respects the $SO(1,1)$
826: symmetry (\ref{so11}) and (\ref{so11qc}). Using
827: \be \frac{H_f(r)}Q =\frac{H_g(r)}C\;,\qquad H_k(r) = 0 \ee the
828: Lagrangian density simplifies, and it is $SO(1,1)$ invariant.
829: Once again there are three distinct orbits: $i)\;|Q|>|C|$, $ii)\;|Q|=|C|$ and $iii)\;|Q|<|C|$, and one expects that these orbits are classified by the corresponding value for the spatial
830: integral of the Lagrangian density, which is now $ {\cal
831: L}^{(0)}_{DBI} + {\cal
832: L}^{(1)}_{DBI}$. To compute the latter we only have to examine one
833: point on each of the orbits, which we do below.
834:
835: $i)\;|Q|>|C|$. A convenient point on this orbit is the purely
836: electrostatic case, where the transverse mode is suppressed:
837: $f(r)=2\pi\alpha' {\cal A}_0'(r)$ and $g(r)=0$. Substituting this into $ {\cal
838: L}^{(0)}_{DBI} + {\cal
839: L}^{(1)}_{DBI}$,
840: and varying with respect to ${\cal A}_0(r)$ gives the corrected Born-Infeld equation
841: $$ \frac{ \sqrt{-\det \tilde
842: h}} \kappa \biggl( r^{p-1}f-Q\sqrt{-\det \tilde
843: h} \biggr)$$
844: \be = \biggl[ \frac{2r^{p-1} H_f' } {(-\det\tilde
845: h)^{3/2}}\biggr]'' - \frac{ 3r^{p-1} f H_f'^2 } {(-\det\tilde h)^{3/2}}
846: -\frac{2(p-1)(3+f^4)(r^{p-3} H_f)'}{(-\det\tilde h)^{3/2}}\label{feco}\ee
847: $$+
848: \frac{(p-1)r^{p-5}f} {(-\det\tilde h)^{3/2}}\biggl\{ f^4(p-2 -r^2 H_f^2)+
849: f^2(-2p-3+4r^2 H_f^2)+ 3(-2p +6 +3r^2 H_f^2) \biggr\}$$
850: To get back the zeroth order equations set the left hand side equal
851: to zero. So the right hand side represents the
852: derivative corrections. The zeroth order Born-Infeld solution
853: satisfies (\ref{feco}) as $r\rightarrow\infty$, so the corrections are
854: negligible in this region. In \cite{Karatheodoris:2002bb},
855: starting with the zeroth order Born-Infeld
856: solution at $r\rightarrow\infty$, we used
857: (\ref{feco}) to numerically integrate to finite $r$. We found the
858: resulting corrections to $f(r)$ at finite $r$ to be small, and just
859: like at zeroth order, $f$ tends to $1$ as $r\rightarrow 0$. Call
860: $f_0(r)$ the zeroth solution, and $f_1(r)$ the correction it receives at first order. Below we
861: plot $f_0(r)$ and $f_0(r)+f_1(r)$ when
862: $Q=1,\;\; C=0$ and $p=3$:
863:
864: \bigskip
865: \input epsf
866: \def\epsfsize#1#2{0.8#1}
867: \centerline{\hss{\epsffile{frstordr.eps}}}
868:
869:
870: \medskip
871: \centerline{\qquad\qquad\qquad
872: ${\tt fig. 7}\qquad f_0\;\; {\tt and }\;\; f_0+f_1\;\;{\tt
873: vs}\;\;r$}
874: \centerline{\qquad\qquad\qquad
875: $\qquad\qquad{\tt for}\;\; p=3,\;\; Q=1,\;\;C=0 $}
876:
877: \noindent
878: In the above we set $\kappa=1$ which is equivalent to choosing the
879: scale for $r$. If the zeroth order Born-Infeld solution gives a reasonable
880: approximation to a classical solution in the full effective theory,
881: and one can apply the derivative expansion to get the latter, then higher order corrections should be small. In particular, we expect only a
882: small change in the value of the Lagrangian at the next order. If
883: one assumes
884: this to be the case
885: a Taylor expansion about the zeroth order solution gives
886: $$ \int d^{p}\xi\;[ {\cal
887: L}^{(0)}_{DBI} + {\cal
888: L}^{(1)}_{DBI}](f_0+f_1) $$ \be \approx \int d^{p}\xi\;[ {\cal
889: L}^{(0)}_{DBI}](f_0 ) + \int d^{p}\xi\;[{\cal
890: L}^{(1)}_{DBI}](f_0 ) +\int d^{p}\xi\;\frac{\delta {\cal
891: L}^{(0)}_{DBI}}{\delta f}\bigg|_{f=f_0 }\; f_1 \ee
892: The last term vanishes by the field equations, and so the first order
893: correction to the Lagrangian is $ \int d^{p}\xi\;[{\cal
894: L}^{(1)}_{DBI}](f_0 ) $. However, it is easy to check that $[{\cal
895: L}^{(1)}_{DBI}](f_0 )$ diverges near the origin as $1/r^{3p+1}$. So
896: the first order correction is not small;
897: Rather, $ \int d^{p}\xi\;[{\cal
898: L}^{(1)}_{DBI}](f_0 ) $ is infinite! This agrees with the
899: result found in \cite{Karatheodoris:2002bb}, and indicates that the Born-Infeld solution, and
900: indeed all case $i)$ solutions, are unstable under inclusion of first
901: order derivative corrections.
902:
903:
904:
905: $ii)\;|Q|=|C|$. This is the BPS case $g(r)=f(r)$. Here $ {\cal L}^{(1)}_{DBI}=0$, and so just like at zeroth
906: order the Lagrangian
907: vanishes. To find equations of motion we must first vary $f$ and $g$
908: (or more precisely $X_{p+1}$
909: and ${\cal A}_0$) separately and then impose the
910: BPS condition. We do this in Appendix B. (Actually, there we don't
911: impose the restriction of spherical symmetry.) The result is simply
912: \be \nabla^2\; \biggl\{ 1 + 2\kappa \;(\nabla^2)^2 \biggr\}\;{\cal
913: A}_0 =0 \;,\label{eqfbps}\ee with the same equation for $X_{p+1}$.
914: For the case of spherically symmetric solutions we can use
915: $\nabla^2= \frac1{r^{p-1}} \partial_r r^{p-1}\partial_r$.
916: Eq. (\ref{eqfbps}) says that the zeroth order solution is also valid
917: at first order. This agrees with \cite{Thorlacius:1997zd}, where it was shown that the
918: BPS solution is valid to all orders. The result (\ref{eqfbps}) thus
919: provides a check of the computations in \cite{Wyl},\cite{das}.
920:
921:
922:
923: $iii)\;|Q|<|C|$. A convenient point is the purely transverse case. Here the electric field vanishes:
924: $f(r)=0$ and $g(r)=X_{p+1}'(r)$. Substituting this in (\ref{radL})
925: and varying $X_{p+1}(r)$ gives
926: $$ \frac{ \sqrt{-\det\tilde
927: h}} \kappa \biggl( r^{p-1}g-C\sqrt{-\det\tilde
928: h} \biggr)$$
929: \be = \biggl[ \frac{2r^{p-1} H_g' } {(-\det\tilde
930: h)^{3/2}}\biggr]'' + \frac{ 3r^{p-1} g H_g'^2 } {(-\det\tilde h)^{3/2}}
931: -\frac{2(p-1)(3+g^4)(r^{p-3} H_g)'}{(-\det\tilde h )^{3/2}}\label{fectw}\ee
932: $$+
933: \frac{(p-1)r^{p-5}g} {(-\det\tilde h)^{3/2}}\biggl\{ g^4(p-2 +r^2 H_g^2)+
934: g^2(2p+3+4r^2 H_g^2)+ 3(-2p +6 -3r^2 H_g^2) \biggr\}$$
935: Again to get back the zeroth order equations set the left hand side equal
936: to zero, and so the right hand side represents the
937: derivative corrections. (\ref{fectw})
938: is also obtained by making the transformation
939: $f(r) \rightarrow ig(r)$ and $Q \rightarrow iC$ in (\ref{feco}), so
940: solving for real $g(r)$ in (\ref{fectw}) is equivalent to solving
941: for imaginary $f(r)$ in (\ref{feco}). As with case $i)$, starting with the zeroth order
942: solution at $r\rightarrow\infty$, we can use
943: (\ref{fectw}) to numerically integrate to finite $r$. We find that just
944: like at zeroth order, $g$ becomes singular at some finite $r$, which
945: appears to be slightly greater than $r_0$. We can then conclude that
946: the corrections cause the wormhole to become wider. Call
947: $g_0(r)$ the zeroth order solution, and $g_1(r)$ the correction it
948: receives at first order. Below we
949: plot $g_0(r)$ and $g_0(r)+ g_1(r)$ when
950: $C=1,\; Q=0$ and $p=3$:
951:
952: \bigskip
953: \input epsf
954: \def\epsfsize#1#2{0.8#1}
955: \centerline{\hss{\epsffile{wrmhl2.eps}}}
956:
957: \centerline{\qquad\qquad\qquad
958: ${\tt fig. 8} \qquad g_0\;\; {\tt and }\;\;g_0+ g_1\;\;{\tt
959: vs}\;\;r $}
960: \centerline{\qquad\qquad\qquad
961: $\qquad\qquad{\tt for}\;\; p=3,\;\; C=1,\;\;Q=0 $}
962:
963: \noindent
964: Again we set $\kappa=1$. Figure 8 shows that the correction $g_1$ to the
965: solution is small away
966: from the wormhole throat. On the other hand, the corresponding
967: correction to the Lagrangian density appears not to be small, as is
968: indicated below where we numerically compare $[{\cal
969: L}^{(0)}_{DBI}+ {\cal
970: L}^{(1)}_{DBI}](g_0+g_1) $ and $[ {\cal
971: L}^{(0)}_{DBI}](g_0) $:
972:
973: \bigskip
974: \input epsf
975: \def\epsfsize#1#2{0.8#1}
976: \centerline{\hss{\epsffile{lagrangdens.eps}}}
977:
978:
979: \medskip
980: \centerline{\qquad\qquad\qquad
981: ${\tt fig. 9} \qquad [ {\cal
982: L}^{(0)}_{DBI}](g_0) \; {\tt and }\; [ {\cal
983: L}^{(0)}_{DBI}+ {\cal
984: L}^{(1)}_{DBI}](g_0+g_1) \;\;{\tt vs}\;\;r $}
985: \centerline{\qquad\qquad\qquad
986: $\qquad\qquad{\tt for}\;\;
987: p=3,\;\; C=1,\;\;Q=0 $}
988:
989: \noindent
990: The lower curve in figure 9 is $ [ {\cal
991: L}^{(0)}_{DBI}+ {\cal
992: L}^{(1)}_{DBI}](g_0+g_1) $. We then expect large corrections to its
993: integral. In fact, we find that the numerical
994: integration of $ [ {\cal
995: L}^{(0)}_{DBI}+ {\cal
996: L}^{(1)}_{DBI}](g_0+g_1) $ fails to give a convergent result. So just
997: as in case $i)$, the integral of the Lagrangian density appears to
998: be ill-defined, indicating that the case $iii)$ solutions
999: are also unstable under inclusion of first
1000: order derivative corrections.
1001:
1002: \section{Conclusion}
1003:
1004: The preliminary indications here are that the classical wormhole
1005: solution may not be a reasonable approximation to a solution in the
1006: full $D-$brane theory. If so it therefore cannot prevent the
1007: decay of the brane-anti-brane system. More generally, it appears
1008: that the only
1009: solution that survives higher order derivative corrections may be the
1010: BPS solution. On the other hand, a more extensive analysis may be possible. For example, it would be interesting to drop the
1011: restriction to static solutions. Perhaps time dependent
1012: configurations can survive first order derivative
1013: corrections, or perhaps the zeroth order static solutions evolve to time
1014: dependent ones after including the higher order. To check this would require
1015: combining two separate stability analyses, which we referred to as $a)$ and $b)$
1016: in the introduction. A final but
1017: unpleasant (from a computational point of view) possibility is that
1018: solutions are recovered only after going beyond the first order.
1019: Moreover, all orders may be required, meaning the solutions may be non perturbative.
1020:
1021:
1022:
1023:
1024: \bigskip
1025:
1026: {\parindent 0cm{\bf Acknowledgement}}
1027:
1028: This work was supported in part by the joint NSF-CONACyT grant
1029: E120.0462/2000.
1030:
1031: \bigskip
1032: \noindent
1033: \appendice $\qquad$ {\bf Alternative Gauge}
1034:
1035: \setcounter{equation}{0}
1036:
1037:
1038:
1039: Here we reconstruct the zeroth order wormhole solutions starting with the
1040: action written in an alternative gauge. This gauge is obtained by replacing the
1041: $r-$coordinate of the static gauge by $z=X_{p+1}$ gauge. It has the
1042: advantage that it removes the coordinate singularity appearing at the midway
1043: point of the wormhole, and shows that there is
1044: a smooth solution connecting the brane to anti-brane. Coordinate
1045: singularities re-appear, however, at the location of the brane and anti-brane.
1046: We show that the electric field is well behaved in this gauge, and in
1047: fact is a constant.
1048:
1049: Consider the
1050: domain $S^{p-1}\times{\mathbb{R}}^2$, with local coordinate patches
1051: used to parametrize the
1052: D$p-$brane. Denote by $z$ one of the coordinates of ${\mathbb{R}}^2$, while the other
1053: corresponds to time $t$. We look for spherically symmetric static solutions with
1054: \be X_0 =t\;,\qquad X_1^2 + X_2^2 +\cdot\cdot\cdot+X_p^2 =
1055: R(z)^2\;,\qquad X_{p+1} =z \ee
1056: So for example for the
1057: D$3-$brane we can write \beqa X_1&=& R(z) \sin\theta \cos\phi \cr
1058: X_2&=& R(z) \sin\theta \sin\phi \cr
1059: X_3&=& R(z) \cos\theta \;, \eeqa
1060: using standard spherical coordinates $\theta$ and $\phi$,
1061: \be h_{tt} =-1\;,\qquad h_{\theta\theta}= R(z)^2\;,\qquad
1062: h_{\phi\phi}= R(z)^2\sin^2\theta \;,\qquad h_{zz}= R'(z)^2+1\;,\ee
1063: where here the prime denotes a derivative in $z$.
1064: Now introduce a $z-$dependent electrostatic field in the $z-$direction,
1065: leading to the off-diagonal components
1066: \be h_{tz}=-h_{zt} =- E(z) \ee
1067: In terms of the electrostatic potential ${\cal
1068: A}_0$ which we now write as a function of $z$, $E(z)=2\pi\alpha'\partial_z{\cal
1069: A}_0(z)$.
1070: After performing
1071: the angular integrations, the DBI Lagrangian $L_{DBI}^{(0)}$ (ignoring the vacuum term) will be proportional to $R(z)^2 \;\sqrt{R'(z)^2-E(z)^2+1}$.
1072: Generalizing to arbitrary $p$,
1073: \be L_{DBI}^{(0)}\; \propto \; R(z)^{p-1} \;\sqrt{R'(z)^2-E(z)^2+1}\ee
1074: Variations in the electrostatic potential and $R(z)$ give
1075: \be \biggl(\frac{ R(z)^{p-1} \;E(z)}{\sqrt{R'(z)^2-E(z)^2+1}}
1076: \biggr)^{'} =0 \ee \be R''(z) R(z) = (p-1)\; (R'(z)^2-E(z)^2 +1)\;, \ee respectively. From the first equation
1077: \be\bigg| \frac {E(z)}Q\bigg| = \sqrt{\frac{R'(z)^2 +1}{R(z)^{2p-2} +Q^2}}\;,\label{ezdir}\ee
1078: where $Q$ is an integration constant, and substituting into the second equation
1079: \be R''(z) = (p-1)\;R(z)^{2p-3}\; \frac{R'(z)^2 +1}{R(z)^{2p-2} +Q^2} \ee
1080: After integrating once
1081: \be R'(z) =\frac 1C \sqrt{ R(z)^{2p-2} - C^2+Q^2}\;,\label{rprm}\ee where
1082: $C$ is an integration constant. Since $R'(z)$ corresponds to
1083: $1/g(r)$, the result agrees with
1084: (\ref{sln0}). For the wormhole solution, $R(z)$ is nonsingular everywhere
1085: except at the location of the brane and anti-brane. At the
1086: midway-point on the wormhole, $R$ is a well-defined function of $z$,
1087: and is a minimum since
1088: \be R''(z_{mid}) = \frac{p-1}{r_0}\;\biggl(1-\frac {Q^2}{C^2}\biggr)\;,
1089: \qquad R(z_{mid})=r_0=(C^2-Q^2)^{\frac 1{2p-2}}\;, \ee
1090: and $|Q|<|C|$ for wormhole solutions.
1091: So now
1092: coordinate singularities appear at the brane and anti-brane, rather
1093: than at the midway-point on the wormhole. By substituting
1094: (\ref{rprm}) into (\ref{ezdir}) one gets that the electric field
1095: $E(z)$ in the $z-$direction is a constant
1096: \be | E(z)| = \bigg| \frac QC\bigg|\;, \ee which agrees with
1097: (\ref{eqoc}). It goes to one in the
1098: BPS limit, and $1/\sqrt{p-1}$ for the minimum energy wormhole. We
1099: conclude that in this
1100: coordinate frame there are no
1101: singularities in the electric field (for $C\ne 0$).
1102:
1103: \bigskip
1104: \noindent
1105: \appendice $\qquad$ {\bf First order BPS equations}
1106:
1107: \setcounter{equation}{0}
1108:
1109: Here we derive the first order BPS equation (\ref{eqfbps}). Unlike in
1110: section 3, we make
1111: no restriction to spherical symmetry. Our starting point is then not
1112: (\ref{hinv1}), but
1113: \be \tilde h= \pmatrix{-1 & - \vec f & & \cr
1114: \vec f &\BI_{p\times p} & \vec g & \cr
1115: &- \vec g & 1 & \cr & & & &\BI_{(8-p)\times(8-
1116: p)}\cr}\;\label{ghinv1}\;,\ee
1117: where $\vec f=2\pi \alpha'\vec \nabla {\cal A}_0$ and $\vec g=\vec \nabla X_{p+1}$ are vector fields on the D$p-$brane.
1118: The general BPS condition is $\vec f=\vec g$. Upon imposing this
1119: condition on $\tilde h^{-1}$ one gets
1120: \be \tilde h^{-1}|_{BPS}= \pmatrix{-1-\vec f^2 &-\vec f &
1121: \vec f^2 & &\cr \vec f & \BI_{p\times p} & -\vec f & \cr
1122: \vec f^2 & \vec f & -\vec f^2+1 & \cr& & & &\BI_{(8-p)\times(8-
1123: p)}\cr
1124: }\;,\ee while the only nonvanishing components of $ S_{ABCD}$ are
1125: \be S_{ijk0}|_{BPS}= S_{ijk\;p+1}|_{BPS}=- S_{ij0k}|_{BPS}=- S_{ij\;p+1\;k}|_{BPS}=
1126: \partial_i\partial_j f_k \ee Since the BPS action vanishes, we must impose
1127: $\vec f =\vec g$ {\it after} performing the variations in the action to find the BPS field equations, i.e. we
1128: must be allowed to perform separate variations of ${\cal A}_0$ and $ X_{p+1}$.
1129: By varying ${\cal A}_0$ and then setting $\vec f=\vec g$ , \beqa
1130: \delta
1131: \tilde h^{00}|_{BPS}&=& -2\;(1+\vec f^2) \; \vec f\cdot\delta\vec f
1132: \cr \delta\tilde h^{p+1\;p+1}|_{BPS}&= &-2\;\vec f^2\; \vec f\cdot\delta\vec f
1133: \cr \delta
1134: \tilde h^{0\;p+1}|_{BPS}&=& (1+2\vec f^2) \; \vec f\cdot\delta\vec f
1135: \; =\;\delta \tilde h^{p+1\;0}|_{BPS} \cr
1136: \delta \tilde h^{i0}|_{BPS}&=& (1+\vec f^2)\;\delta f_i+ \vec f
1137: \cdot\delta\vec f\;f_i\;=\;-\delta\tilde h^{0i}|_{BPS}\cr
1138: \delta\tilde
1139: h^{i\;p+1}|_{BPS}&= & -\vec f^2\;\delta f_i- \vec f
1140: \cdot\delta\vec f\;f_i\; =\;-\delta
1141: \tilde h^{p+1\;i}|_{BPS} \cr \delta
1142: \tilde h^{ij}|_{BPS}&=& - f_i\delta
1143: f_j + f_j\delta f_i \;, \eeqa while the nonvanishing components of $
1144: \delta S_{ABCD}$ are \beqa \delta S_{ijk0}|_{BPS}&=&-\delta S_{ij0k}|_{BPS}=
1145: \partial_i\partial_j\delta f_k
1146: + \delta S_{ijk\;p+1}|_{BPS} \cr & &\cr
1147: \delta S_{ijk\;p+1}|_{BPS}&=&-\delta S_{ij\;p+1\;k}|_{BPS}=
1148: \partial_i f_\ell \;\partial_j f_k\;\delta f_\ell
1149: - \tilde h^{0\ell} \partial_i \delta f_k\; \partial_j f_\ell
1150: +(i\rightleftharpoons j)\cr & &\cr \delta S_{ijk\ell}|_{BPS}&=&\partial_j f_\ell\;
1151: \partial_i \delta f_k+
1152: \partial_i f_k\;\partial_j \delta f_\ell -(i\rightleftharpoons j)
1153: \eeqa
1154: In evaluating $\delta \Delta|_{BPS}$ we can use $(\tilde h^{AB} S_{ijAB})|_{BPS}=0$.
1155: Then
1156: \beqa \delta \Delta|_{BPS}&=& -4\tilde h^{CD} S_{jiDA}\; \{ \delta
1157: \tilde h^{AB} S_{ijBC}
1158: + \tilde h^{AB} \delta S_{ijBC} +\tilde h^{AB} \delta \tilde h^{i\ell}S_{\ell jBC}
1159: \}\;|_{BPS}\cr & &\cr
1160: &=& -8\;
1161: \partial_i\partial_j f_k\;\partial_i\partial_j\delta f_k \eeqa
1162: Now substitute $\vec f=2\pi \alpha'\vec \nabla {\cal A}_0$ to obtain
1163: (\ref{eqfbps}) from the variation of ${\cal A}_0$ in ${\cal L}^{(0)}_{DBI} + {\cal
1164: L}^{(1)}_{DBI}$. One gets the same results from variations of $X_{p+1}$.
1165:
1166: \bigskip
1167:
1168: \bigskip
1169:
1170: \begin{thebibliography}{99}
1171:
1172:
1173: \bibitem{bi} M. Born and L. Infeld, Proc. Roy. Soc. London {\bf A144} 425 (1934).
1174:
1175: \bibitem{Dirac} P.A.M. Dirac, Proc. Roy. Soc. London {\bf A268} 57 (1962).
1176:
1177:
1178: \bibitem{Fradkin:1985}
1179: E.~S. Fradkin and A.~A. Tseytlin, Phys. Lett. {\bf B163} (1985) 123;
1180: %%CITATION = PHLTA,B163,123;%%.
1181: A.~Abouelsaood, C.~G. Callan, C.~R. Nappi, and S.~A. Yost, Nucl. Phys. {\bf B280} (1987) 599;
1182: %%CITATION = NUPHA,B280,599;%%.
1183: E.~Bergshoeff, E.~Sezgin, C.~N. Pope, and P.~K. Townsend, Phys. Lett. {\bf
1184: 188B} (1987) 70;
1185: %%CITATION = PHLTA,188B,70;%%.
1186: R.~G. Leigh, Mod.
1187: Phys. Lett. {\bf A4} (1989) 2767.
1188: %%CITATION = MPLAE,A4,2767;%%
1189: \bibitem{Johnson} For reviews, see
1190: C.~P.~Bachas,
1191: %``Lectures on D-branes,''
1192: arXiv:hep-th/9806199; A.~A.~Tseytlin,
1193: %``Born-Infeld action, supersymmetry and string theory,''
1194: arXiv:hep-th/9908105;
1195: C.~V.~Johnson,
1196: %``D-brane primer,''
1197: arXiv:hep-th/0007170; I.~V.~Vancea,
1198: %``Introductory lectures on D-branes,''
1199: arXiv:hep-th/0109029; R.~J.~Szabo,
1200: %``BUSSTEPP lectures on string theory: An introduction to string theory and D-brane dynamics,''
1201: arXiv:hep-th/0207142.
1202:
1203:
1204:
1205: \bibitem{Callan:1997kz}
1206: C.~G.~Callan and J.~M.~Maldacena,
1207: %``Brane dynamics from the Born-Infeld action,''
1208: Nucl.\ Phys.\ B {\bf 513}, 198 (1998).
1209:
1210:
1211:
1212: \bibitem{Gib}
1213: G.~W.~Gibbons,
1214: %``Born-Infeld particles and Dirichlet p-branes,''
1215: Nucl.\ Phys.\ B {\bf 514}, 603 (1998).
1216:
1217:
1218: \bibitem{Banks:1995ch}
1219: T.~Banks and L.~Susskind,
1220: %``Brane - Antibrane Forces,''
1221: arXiv:hep-th/9511194.
1222:
1223: \bibitem{Wyl}
1224: N.~Wyllard,
1225: %``Derivative corrections to D-brane actions with constant background fields,''
1226: Nucl.\ Phys.\ B {\bf 598}, 247 (2001); JHEP {\bf 0108}, 027 (2001).
1227:
1228: \bibitem{das} S.~R.~Das, S.~Mukhi and N.~V.~Suryanarayana,
1229: %``Derivative corrections from noncommutativity,''
1230: JHEP {\bf 0108}, 039 (2001).
1231: \bibitem{Karatheodoris:2002bb}
1232: G.~Karatheodoris, A.~Pinzul and A.~Stern,
1233: %``Fate of the Born-Infeld solution in string theory,''
1234: Mod. Phys. Lett. A {\bf 18}, 1681 (2003);
1235: arXiv:hep-th/0211033.
1236:
1237:
1238:
1239: \bibitem{Thorlacius:1997zd}
1240: L.~Thorlacius,
1241: %``Born-Infeld string as a boundary conformal field theory,''
1242: Phys.\ Rev.\ Lett.\ {\bf 80}, 1588 (1998).
1243:
1244: \bibitem{Bateman}
1245: H. Bateman, Higher transcendental functions, Bateman
1246: Manuscript Project,
1247: New York, McGraw-Hill, 1953-55.
1248:
1249:
1250: \end{thebibliography}
1251: \end{document}
1252:
1253: