1: \documentclass [12pt] {book}
2: %\topmargin 0in \textheight 9in \oddsidemargin 0in
3:
4:
5: \usepackage{graphicx}
6: \usepackage{amsmath}
7: \usepackage{amsfonts}
8: \usepackage{amssymb}
9:
10:
11: \def\f{{\rm p}}
12: \def\vd{{(5)}}
13: \def \vu {{(5)}\!}
14: \def \D {\mbox{D}}
15: \def \curl {\mbox{curl}\,}
16: \def \ep {\varepsilon}
17: \def \cu{{\cal U}}
18: \def \cq{ Q}
19: \def \cp{ \Pi}
20: \def \bcq { \bar{Q}}
21: \def \bcp { \bar{\Pi}}
22: \def\bea{\begin{eqnarray}}
23: \def\eea{\end{eqnarray}}
24: \newcommand{\be}{\begin{equation}}
25: \newcommand{\ee}{\end{equation}}
26: \newcommand{\ba}{\begin{eqnarray}}
27: \newcommand{\ea}{\end{eqnarray}}
28: \newcommand{\half}{\frac{1}{2}}
29: \newcommand{\pre}{\frac{1}{4\pi\alpha'}}
30: \newcommand{\four}{^{(4)}}
31: \newcommand{\five}{\tilde}
32: \newcommand{\mb}[1]{\mbox{\boldmath $#1$}}
33: \evensidemargin 0in
34: %\renewcommand{\baselinestretch}{1.5}
35: \textwidth 6.0in
36:
37: \begin{document}
38: \pagestyle{empty}
39: \input{intestazione.tex}
40: \pagestyle{headings}
41: \pagenumbering{roman}
42: \tableofcontents
43: \chapter{Foreword}
44: \footnote{This is a rough and informal view of quantum gravity,
45: using \cite{R1}.}Just one year after the discovery of general
46: relativity, Einstein pointed out that quantum effects must lead to
47: modifications in the theory of general relativity \cite{E1}. The
48: problems of compatibility between gravity and quantum mechanics were further
49: considered by Heisenberg \cite{H1}. He realized that since the
50: gravitational coupling constant is dimensional, a theory which
51: quantizes gravity will have serious problems. Indeed in the
52: seventies, t'Hooft and Veltman as well as Deser and Van
53: Nieuwenhuizen \cite{tH1}, confirmed that the quantum theory of
54: gravity coupled with matter has non-renormalizable divergences.
55: This was disappointing for a quantum theory of gravity. Indeed it
56: showed that the old techniques of quantization could not work for
57: general relativity. Soon after, Hawking \cite{H2} discovered that
58: a black hole is not cold, but thanks to quantum effects emits
59: radiation with temperature
60: \begin{equation}
61: T=\frac{\hbar c^3}{8\pi kGM}.
62: \end{equation}
63: Then Unruh \cite{U1} proposed a relation between accelerated
64: observers, quantum theory, gravity and thermodynamics. This
65: suggested new profound relations between quantum and classical
66: worlds. The idea was that a full quantum gravity description must
67: reproduce these relations.
68:
69: A connection between classical gravity and quantum field theory
70: was first seen in the seventies.
71: In the late sixties indeed, Veneziano \cite{V1}, trying to understand why the theoretically
72: expected amplitude divergences did not appear when high
73: energy elastic scattering of baryons produced
74: particles with higher spins, introduced a duality between the Mandelstam
75: variables $s$ and $t$ of the process \cite{IZ}. In this
76: way the divergences in one channel for higher spins, were
77: exactly cancelled out by the other channel.
78: The Veneziano theory had some peculiarities
79: such as the existence of
80: a massless spin-two particle, which did not
81: correspond to any renormalizable quantum theory. This particle was
82: then recognized as the graviton.
83: Indeed several years later, people realized that the phenomenological
84: model of Veneziano could be derived from a more fundamental
85: theory. The idea was to quantize one dimensional objects (strings)
86: instead of point-like objects (particles). Interpreting the massless spin-two
87: particle as the graviton, it was
88: conjectured that string theory leads to a non-divergent theory of
89: quantum gravity.
90: As well as producing exciting insights,
91: the introduction of strings implied also the existence of extra spatial dimensions, which was
92: at that time considered to be a problem.
93:
94: In the eighties Polyakov \cite{P1} showed that the extra
95: dimensions were linked to the necessity of keeping the classical
96: symmetries for this action at the quantum level.
97:
98: Later works showed also that string theory provides a consistent
99: theory at perturbed level. The aim of this theory was to unify all
100: the forces without introducing phenomenological coupling
101: constants.
102:
103: Another important concept was developed by t'Hooft \cite{tH2} and
104: promoted by Susskind \cite{S1}. They proposed that the information
105: of a physical state in the interior of a region can be represented
106: on the region's boundary and is limited by the area of this
107: boundary. This was motivated to explain why the entropy of a black
108: hole (which contains all the information of
109: this object) depends only on the area of its horizon and not on
110: the volume inside the horizon. This principle was called
111: the ``Holographic principle''. The holography was a tentative way to connect
112: classical physics with quantum physics in a purely geometrical
113: way. In 1998 Maldacena \cite{M1} applied this concept in string
114: theory. Soon after, Witten \cite {W2} clarified the holographic
115: features of anti-deSitter spacetimes which was the basis for the
116: Maldacena work. The concept was that a conformal quantum field
117: theory (CFT) on a boundary of a spacetime with anti-deSitter
118: background can be described as a classical gravitational theory.
119: This was later called the AdS/CFT correspondence.
120:
121: A sociological event was also born with string theory. A dialogue
122: between General Relativity
123: and Quantum Physics communities reemerged.
124:
125: But, as Rovelli explains \cite{R1}, there are still profound
126: divergences. From the point of view of the General Relativity
127: community, quantum field theory is problematic. Up to now indeed,
128: quantum physics with gravity is consistent only for fixed
129: spacetime backgrounds. This implies that it is inadequate for a
130: full understanding of a dynamical theory of spacetime. On the
131: other side, for the Quantum Mechanics community, General
132: Relativity is only a low energy limit of a much more complex
133: theory, and thus cannot be taken too seriously as an indication of
134: the deep structure of Nature.
135:
136: A bridge between the two approaches must be found. One possible way is to study the
137: dynamics of gravitational theories inspired by quantum gravity.
138:
139: If one follows the direction of string theory and the
140: holographic principle, one must find a mechanism to reduce the dimensions of the
141: spacetime to four, at least at low energies.
142:
143: One possibility was introduced in 1999 by Randall and Sundrum
144: \cite{RS}, where the extra dimension is infinitely large and the
145: matter is trapped on a four-dimensional submanifold called the
146: ``brane''. The graviton is instead localized on the brane only at
147: low energies, reproducing the correct Newtonian limit. This is
148: possible thanks to a negative cosmological constant which away from
149: the brane induces an anti-deSitter spacetime.
150:
151: In this thesis I will discuss the astrophysical and cosmological implications
152: of this mechanism for simple phenomenological set ups.
153:
154: I use the signature $(-,+,+,+)$ and $(-,+,+,+,+)$ for the
155: four and five dimensional Lorentzian manifold, the natural
156: units $c=\hbar=1$ and the definition of the Ricci tensor
157: $R^\alpha{}_{\mu\alpha\nu}=R_{\mu\nu}$, where $R_{\alpha\mu\nu\beta}$ is the
158: Riemann tensor.
159:
160: \mainmatter
161: \newpage
162: \chapter{Extra dimensions: motivations}
163:
164: To give a flavour of how extra dimensions appear in modern
165: theoretical physics, I will describe simple examples in the
166: context of String Theory and Holography. Since this chapter must
167: be seen as only as a motivation for the study of extra
168: dimensions, I will not directly connect these theories with
169: the Randall-Sundrum mechanism. I will anyway discuss throughout
170: the thesis, the influences that these theories have in the
171: Randall-Sundrum-type models.
172:
173: \section{String Theory}
174:
175: Currently the most promising theory for the unification of all the
176: forces seems to be superstring theory. There are actually five
177: anomaly-free perturbative string theories which are: type I, type
178: IIA, type IIB, SO(32) and $E_8\times E_8$ heterotic theories
179: \cite{GSW,PO}. These are all supersymmetric and require in general
180: ten dimensions. Another interesting theory is supergravity. In
181: eleven dimensions it is unique and it has been proved that its
182: compactification in ten dimensions reproduces the low energy limit
183: of the type IIA superstring theory. This was the starting point to
184: conjecture that a more general theory, whose low energy limit is
185: the eleven-dimensional supergravity \cite{SU}, is actually the
186: ultimate theory. This concept was also reinforced thanks to the
187: evidence that the superstring theories are related to each other
188: by dualities. The ultimate theory was called M-Theory
189: \cite{MT1,MT2,MT3,MT4}.
190:
191: String theory contains an infinite tower of massive states,
192: corresponding to the oscillations of the string. The massless
193: states are separated from the massive ones by a gap of energy of
194: order $1/\sqrt{\alpha'}$. $T=1/{2\pi\alpha'}$ is called the string
195: tension, which is usually taken to be close to the Planck scale.
196: This is the only arbitrary parameter of the theory. In the limit
197: of an infinite tension, the massless excitations decouple from
198: the massive ones and the theory is described by the low energy
199: limit which in the effective action contains the usual
200: Einsteinian gravity. The low energy limit, more geometrically, is
201: dictated by the comparison between the curvature radius of the spacetime and the
202: string length. This means that the low energy limit is for
203: $R^{-1/2}\gg \sqrt{\alpha'}$ where $R$ is the Ricci scalar.
204:
205: \subsection{Classical string action and equation of motion}
206:
207: \subsubsection{String action in flat background}
208:
209: In this section I give a taste of how the extra dimensions arise in string theory.
210: A complete treatment of bosonic and
211: supersymmetric string theories can be found in \cite{GSW,PO}.
212:
213: The motion of a free falling relativistic particle is described by
214: the maximum spacetime length from an event $P_1$ to $P_2$
215: \cite{Wald}. The spacetime length is described by the integral \be
216: {\cal L}=-m\int^{P_2}_{P_1} ds, \ee which depends on the path,
217: where $m$ is the mass of the particle and \be
218: ds^2=g_{\alpha\beta}dx^\alpha dx^\beta, \ee where
219: $g_{\alpha\beta}$ is the spacetime metric.
220:
221: Supposing the particle is massive ($m>0$), we can define the
222: proper time $\tau$ such that \be \frac{ds}{d\tau}=-1. \ee With
223: this parameter we can rewrite the total length in terms of the
224: particle four velocity $u^\alpha=dx^\alpha/d\tau$, as \be {\cal
225: L}=-m\int^{\tau_2}_{\tau_1}\sqrt{-u^\alpha u_\alpha} d\tau. \ee
226:
227: Now we introduce a one-dimensional object instead of a point-like
228: one, i.e. a string. With the same logic we can say that the motion
229: of a string is described by the extremal area covered from a curve
230: $S_1$ to $S_2$. So the action for a string is \be {\cal
231: A}=-\frac{1}{2\pi\alpha'}\int^{S_2}_{S_1} dA, \ee where
232: $T=1/2\pi\alpha'$ is the tension of the string.
233:
234: Since the string is a two-dimensional spacetime object, we can
235: describe its motion as a two-dimensional sheet. This sheet can be
236: described by two internal coordinates. We use the proper time
237: $ \tau $ and the proper spatial length $ \sigma $. Of course the
238: metric which describes this sheet in spacetime is (with signature
239: $(-,+)$) \be h_{ab}{}^{\mu\nu}=\partial_a x^\mu \partial_b x^\nu,
240: \ee where the Latin letters $a,b,..$ are the internal coordinates
241: $(\sigma,\tau)$. This is just the generalization of the point-like
242: four velocity $u^\alpha=\partial_\tau x^\alpha$. Indeed we have
243: this simple scheme: \ba
244: &\mbox{particle}\rightarrow\mbox{string}\cr\nonumber\cr\nonumber
245: &\sqrt{-u^\alpha
246: u_\alpha}\rightarrow \sqrt{-\mbox{det}(h_{ab}{}^\alpha{}_\alpha)}
247: =\sqrt{-\mbox{det}(h_{ab})}.\nonumber\cr\nonumber \ea Then finally
248: the action for a string is \be\label{ng} {\cal
249: A}=-\frac{1}{2\pi\alpha'}\int^{\tau_2,\sigma_2}_{\tau_1,\sigma_1}
250: \sqrt{-\mbox{det}(h_{ab})}d\sigma d\tau. \ee This is called the
251: Nambu-Goto action. From now on $\mbox{det}(h_{ab})=h$ and the
252: surface with boundaries $(\sigma_1,\tau_1)$ and
253: $(\sigma_2,\tau_2)$ is denoted by $M$.
254:
255: Since we are interested in the quantum theory of strings, we need
256: to manipulate the action so that it appears in a more ``linear"
257: form. In this way we can use the standard quantization rules for
258: non-linear sigma models. To do that we introduce an auxiliary
259: metric $\gamma^{ab}(\sigma,\tau)$ and rewrite eq. (\ref{ng})
260: as \be\label{polact} {\cal A}=-\frac{1}{4\pi\alpha'}\int_M d\tau
261: d\sigma \sqrt{-\gamma}\gamma^{ab}h_{ab}. \ee Then we can see under
262: which conditions this action is equivalent to the Nambu-Goto one.
263: If $\gamma_{ab}$ is an auxiliary metric, then the variation of the
264: action with respect to it must vanish. Varying the action we get
265: the equation \be h_{ab}=\frac{1}{2}\gamma_{ab}\gamma^{cd}h_{cd},
266: \ee from which follows
267: \be
268: \sqrt{-h}=\sqrt{-\gamma}\gamma_{ab}h^{ab}. \ee
269:
270: The action (\ref{polact}) is called the Polyakov action. This
271: action is invariant under the worldsheet ($(\sigma,\tau)$-space)
272: diffeomorphism
273: \begin{eqnarray}
274: &\delta x^\mu=\pounds_{\xi}x^\mu=\xi^a\partial_a x^\mu\ ,\\
275: &\delta \gamma^{ab}=\pounds_{\xi}\gamma^{ab}=\xi^c\partial_c
276: \gamma^{ab}- 2\gamma^{c(a}\partial_c\xi^{b)},
277: \end{eqnarray}
278: and the two-dimensional Weyl invariance \be \delta
279: \gamma^{ab}=\omega \gamma^{ab}. \ee
280: Here $\xi^a$ and $\omega$
281: are small vector and scalar parameters.
282: Of course we have also the
283: spacetime Poincar\'e invariance
284: \be \delta
285: x^\mu=\omega^{\mu\nu}x_\nu+b^\mu\ , \ee
286: where $\omega^{\mu\nu}$ represent a rotation and $b^\mu$ a translation in spacetime.
287:
288: Now $\gamma^{ab}$ is actually the metric for the gravitational
289: field on the worldsheet, so we can use it to raise or lower the
290: worldsheet indices.
291:
292: In general in the action (\ref{polact}) we could add a term which
293: describes the field equation for it. In two dimensions (the
294: worldsheet) the only possibility is to add a cosmological
295: constant, since the Ricci scalar is a total derivative, and so
296: therefore does not affect the equations of motion. However this
297: breaks the Weyl invariance, which is a key point to quantize the
298: string, and so it is set to zero.
299:
300: Since the action is invariant under variations of $\gamma^{ab}$,
301: the energy-momentum tensor associated with it,
302: \be
303: T_{ab}=\partial_a x^\mu\partial_b x_\mu-\half
304: \gamma_{ab}\partial^cx^\mu\partial_cx_\mu\ , \ee
305: must vanish. In
306: particular its trace is zero. This is a consequence of the
307: conformal invariance (Weyl) of the action. We will see that, in
308: general, quantum mechanically the trace of this tensor does not
309: vanish. In order to keep this symmetry we will need in general
310: extra-dimensions.
311:
312: Thanks to the classical Weyl and Poincar\'e symmetries we can locally choose the
313: worldsheet metric. We use the Minkowski one \be
314: \gamma_{ab}=\eta_{ab}=\mbox{diag}(-1,1). \ee Since strings are
315: one-dimensional objects they can be open or closed. We will see
316: that in the first case we have some boundary conditions to impose.
317: The variation of the action with respect to the fields
318: $x^\mu(\tau,\sigma)$ is \be -4\pi\alpha'\delta{\cal A}=\int_M
319: d^2\sigma \partial_a x^\mu \partial^a \delta x^\mu, \ee where we
320: used $\delta\partial_a x^\mu\simeq\partial_a\delta x^\mu$ and
321: $d\sigma d\tau=d^2\sigma$.
322:
323: Integrating by parts we get
324: \be -4\pi\alpha'\delta{\cal A}=\int_M
325: d^2\sigma \partial _a (\eta^{ab}\partial _b x^\mu \delta
326: x_\mu)-\int_M d^2\sigma \delta x_\mu \Box x^\mu\ . \ee
327: If the string
328: is closed the first integral is zero. So the equations of motion
329: are \be \label{eqm} \Box
330: x^\mu=(\partial^2_\tau-\partial^2_\sigma)x^\mu(\tau,\sigma)=0. \ee
331: For an open string the first integral is \be \int_M d^2\sigma
332: \partial _a (\eta^{ab}\partial _b x^\mu \delta
333: x_\mu)=\oint_{\partial M}d\sigma_a\eta^{ab}\partial_b x^\mu\delta
334: x_\mu. \ee
335: We can choose $\partial M$ to be a space-like boundary,
336: so that $d\sigma_a=d\tau\delta^\sigma_a$. Normalizing the length
337: of the string such that $0\leq \sigma\leq\pi$, in order to have
338: zero variation we can consider a combination of the following
339: possibilities for the ends of the string:
340:
341: {\it Neumann boundary conditions} \ba
342: \partial_\sigma x^\mu\Big|_\pi=0,\ \ \partial_\sigma x^\mu\Big|_0=0.
343: \ea
344:
345: {\it Dirichlet boundary conditions} \ba
346: x^\mu\Big|_\pi=\mbox{const}\Rightarrow\delta x^\mu\Big|_\pi=0,\ \
347: x^\mu\Big|_0=\mbox{const}\Rightarrow\delta x^\mu\Big|_0=0. \ea
348:
349: \subsubsection{General string action}
350:
351: The string action can be generalized in the following way
352: \cite{GSW}:
353: \be\label{coupled} {\cal S}=-\pre \int d^2\sigma
354: \left[\sqrt{-\gamma}\gamma^{ab}g_{\mu\nu}(x^\alpha)\partial_a
355: x^\mu\partial_b x^\nu+\epsilon^{ab}B_{\mu\nu} (x^\alpha)\partial_a
356: x^\mu \partial_b x^\nu\right]. \ee
357: The coupling functions
358: $g_{\mu\nu}$ and $B_{\mu\nu}$ can be identified as the background
359: spacetime graviton and antisymmetric tensor fields in which the
360: string is propagating. $\epsilon^{ab}$ is the two-dimensional
361: Levi-Civita symbol. The coupling of these tensors with the string
362: fields is well justified also at the quantum level. At quantum
363: level we also get a massless scalar field in the spectrum. Such a
364: field in general breaks the Weyl invariance. Fradkin and Tseytlin
365: \cite{F1} have suggested that one should add to the string theory
366: action the renormalizable but not Weyl invariant term \be
367: S_{dil}=\frac{1}{4\pi}\int d^2\sigma
368: \sqrt{\gamma}R^{(2)}\Phi(x^\alpha), \ee where $R^{(2)}$ is the
369: two-dimensional Ricci scalar and the scalar field $\Phi(x^\alpha)$
370: is called the dilaton. Now it is true that at a classical level
371: this term breaks the conformal symmetry. Since string theory is a
372: quantum theory, we actually can require the weak condition that
373: the Weyl anomaly is cancelled at least at quantum level.
374:
375: From now we use the following definition of total string action
376: \be
377: I[x,\gamma]={\cal S}+S_{dil}\ .
378: \ee
379: In the next
380: section we show how the quantization of this action leads to the Weyl anomalies.
381:
382:
383: \subsection{Quantization of the string action}
384:
385: As an example of how extra-dimensions appear in string theory we first consider
386: the simplest background in which $\Phi=0$.
387:
388: A modern concept of quantization is to introduce the partition function via the path
389: integral
390: \be
391: Z=\int {\cal D}\gamma(\sigma){\cal D}x(\sigma)e^{-I[x,\gamma]}\ ,
392: \ee
393: where ${\cal D}f$ means the integration over all the possible functions $f$.
394:
395: Since Weyl symmetry of the classical action we can restrict ourselves in considering
396: \be
397: \gamma_{ab}=e^\phi\eta_{ab}\ .
398: \ee
399: In this way we can try to make explicit the measure ${\cal D}\gamma(\sigma)$ as the integration
400: of all the possible reparameterizations of the worldsheet metric. This is possible by introducing
401: auxiliary fields that basically fix the gauge for any choice of $\phi$ under integration.
402: These fields, called
403: Faddeev-Popov ghosts ($b,c$), are anticommuting \cite{GSW}.
404: Using the fact that the partition function can be rewritten as
405: \be
406: Z=\int {\cal D}\phi(\sigma)\int {\cal D}x(\sigma){\cal D}b(\sigma) {\cal D}c(\sigma)
407: e^{-I[x,b,c]}\ ,
408: \ee
409: where now the effective action becomes, in complex coordinates ($ds^2=dzd\overline z$)
410: \be
411: I[x,b,c]=I[x,\gamma]-\frac{1}{2\pi\alpha'}\int d^2\sigma \left[b_{zz}\nabla_{\overline z}
412: c^z+ \mbox{c.c.} \right]\ ,
413: \ee
414: the ghost $c^z$($c^{\overline z}$) is a holomorphic (antiholomorphic) vector and the
415: antighost $b_{zz}$($b_{\overline z\overline z}$) is an holomorphic (antiholomorphic) quadratic
416: differential.
417:
418: The effective action $I[x,b,c]$ is classically conformal invariant.
419: In order to keep this invariance at the quantum level, we have to show under which conditions
420: the product of the measures ${\cal D}x{\cal D}b {\cal D} c$ and $I[x,b,c]$
421: are quantum Weyl invariant. Considering the
422: quantum fluctuations of the string and ghost fields as gaussian, because of their
423: dependence on the
424: worldsheet metric, we obtain, under rescaling
425: $\eta_{ab}\rightarrow e^{\xi} \eta_{ab}$, \cite{Ginsparg2}
426: \be\label{poly}
427: {\cal D}_{e^{\xi}\eta}x {\cal D}_{e^\xi\eta}[\mbox{ghost}]=\exp\left(\frac{D-26}{48\pi}S_L(\xi)
428: \right){\cal D}_{\eta}x{\cal D}_\eta[\mbox{ghost}]\ ,
429: \ee
430: where $S_L(\xi)$ is known as the Liouville action
431: (see \cite{Ginsparg2} for the explicit form) and
432: $D$ is the spacetime dimension.
433: Then in order to keep Weyl invariance at the quantum level, the first requirement is that
434: $D=26$.
435:
436: Considering now a non vanishing $\Phi$, to see under what conditions $I[x,b,c]$ is quantum scale invariant,
437: one can calculate the quantum trace of the renormalized energy-momentum tensor of
438: the string, defined as
439: \be
440: \gamma^{ab}\frac{\delta \ln Z}{\delta \gamma_{ab}}\Big|_{\rm ren}=
441: i\gamma^{ab}\langle T_{ab}\rangle\ .
442: \ee
443: The renormalized result is \cite{C1}
444: \ba 2\pi
445: \langle T^a{}_a\rangle=\beta^\Phi\sqrt{-\gamma}R^{(2)}+\beta_{\mu\nu}^g\sqrt{\gamma}\gamma^{ab}
446: \partial_ax^\mu\partial_b
447: x^\nu+\beta_{\mu\nu}^B \epsilon^{ab}\partial_ax^\mu\partial_b
448: x^\nu, \ea
449: where $\beta^\Phi$, $\beta^g$ and $\beta^B$ are local
450: functionals of the coupling functions $\Phi$, $g_{\mu\nu}$ and
451: $B_{\mu\nu}$, and in the limit $R^{(2)-1}\sqrt{\alpha'}\ll
452: 1$,
453: \ba \beta^g_{\mu\nu}=
454: R_{\mu\nu}+2\nabla_\mu\nabla_\nu\Phi-\frac{1}{4}H_{\mu\rho\sigma}
455: H^{\rho\sigma}_\nu+O(\alpha')\label{noncri1}\\
456: \beta^B_{\mu\nu}=\nabla_\lambda H^\lambda_{\mu\nu}-2(\nabla_\lambda\Phi)H^\lambda_{\mu\nu}+
457: O(\alpha')\\
458: \beta^\Phi=\frac{D-D_c}{48\pi^2}+\frac{\alpha'}{16\pi^2}\left\{4(\nabla\Phi)^2-4\nabla^2\Phi-
459: R+\frac{1}{12}H^2\right\}
460: +O(\alpha'^2).\label{noncri2} \ea
461: Here
462: $H_{\mu\nu\lambda}=3\nabla_{[\mu}B_{\nu\lambda]}$, and $D_c=26$ is the number of
463: critical dimensions. In the beta function for the Weyl
464: anomaly $\beta^\Phi$, the leading term for bosonic strings was
465: discovered by Polyakov \cite{P1} in a similar way as we did in finding (\ref{poly}).
466: The graviton part was found by Friedan et
467: al. \cite{F2} and the $H$-fields by Witten \cite{W1} and
468: Curtright and Zachos \cite{C2}. In order to keep the Weyl symmetry at quantum level, we
469: have to make $\beta^\Phi$, $\beta^g$ and $\beta^B$ vanish.
470:
471: It is important to note that since the coefficient of
472: $R^{(2)}\Phi$ is smaller by a factor $\alpha'$ than the other
473: couplings, its {\it classical} contribution is of the same order
474: as the one-loop {\it quantum contribution} of the $g_{\mu\nu}$ and
475: $B_{\mu\nu}$ couplings. This is because $R^{(2)}\Phi$ is scale
476: non-invariant at the classical level, while the other couplings
477: only lose scale-invariance at the quantum level. It is very simple
478: to prove that $\beta^\Phi$ is actually a constant in space time.
479: Indeed applying the Bianchi identities to $\beta^g$ and $\beta^B$
480: we get \be \nabla_\mu \beta^\Phi=0. \ee Therefore once we have solved the
481: equations for $\beta^{G,B}$, $\beta^\Phi$ is determined up to a
482: constant.
483:
484: From eqs. (\ref{noncri1}-\ref{noncri2}) we can already argue that
485: string theory
486: does not have to live in critical
487: dimensions. This kind of String theory is called non-critical
488: string theory \cite{PO}. In the $\beta^\Phi$ function, even if
489: the correction to the critical theory with $D=D_c$ is at first
490: order in $\alpha'$, it is actually possible to solve the equation
491: consistently to the one-loop correction. Myers \cite{M2} indeed
492: found that at least for the bosonic string in flat spacetime, where
493: $D_c=26$, a consistent solution is possible. This is
494: \ba
495: B_{\mu\nu}=0\ ,\ \Phi(x^\mu)=V_\mu x^\mu, \ea
496: where
497: \ba V_\mu
498: V^\mu=\frac{26-D}{6\alpha'}. \ea
499: This solution is compatible with
500: any dimension of the spacetime. There are not any other
501: anomaly-free non-critical string theories known.
502:
503: Introducing supersymmetry in the string
504: action, which is a natural way to consider matter fields, one can prove
505: that $D_c=10$
506: instead of $D_c=26$ in eq. (\ref{noncri2}) \cite{C1}.
507: The bosonic sector of the supersymmetric string still satisfies
508: equations (\ref{noncri1}-\ref{noncri2}). These can be described by an
509: effective action \cite{M3}
510: \ba S_D=\frac{1}{2\kappa^2_D}\int d^D x
511: \sqrt{-g}e^{-\Phi}\left[ R+(\nabla\Phi)^2-\frac{1}{12}
512: H^2-\frac{2(D-10)}{3}\right].
513: \ea
514: Since the $\beta^\Phi$ function is at
515: order $\alpha'$, we can consistently consider the $\alpha'$
516: corrections to the $\beta^g_{\mu\nu}$ function as well. This
517: calculation can be done in heterotic string theory, which seems
518: to contain the gauge group of the Standard Model \cite{G1}. If we
519: choose a background with $B_{\mu\nu}=0$, the effective action then
520: becomes \cite{BoDe:85}
521: \be S_D=\frac{1}{2\kappa^2_D}\int d^Dx
522: \sqrt{-g}e^{-\Phi}\left[
523: R+(\nabla\Phi)^2-\frac{2(D-10)}{3}-\frac{\alpha'}{8}
524: \left\{L_{GB}-(\nabla\Phi)^4 \right\}\right], \ee
525: where
526: $L_{GB}=R^{\mu\nu\alpha\beta}R_{\mu\nu\alpha\beta}-4R^{\mu\nu}R_{\mu\nu}+R^2$
527: is the Gauss-Bonnet term.
528:
529: Applying a conformal transformation
530: $g_{\mu\nu}\rightarrow e^{\frac{4}{D-2}\Phi}g_{\mu\nu}$ we can rewrite the
531: action in a more familiar form \cite {GBE,M3}
532: \ba
533: S_D=\frac{1}{2k^2_D}\int d^Dx \sqrt{-g}[
534: R&-&\frac{4}{D-2}(\nabla\Phi)^2-\frac{2(D-10)}{3}
535: e^{\frac{4}{D-2}\Phi}+\cr\nonumber
536: &+&\frac{\alpha'}{8}e^{-\frac{4}{D-2}\Phi}(L_{GB}+4\frac{D-4}{(D-2)^3}(\nabla\Phi)^4)].
537: \ea
538: This frame is called the Einstein frame. At this level it is
539: explicit that string theory contains, as effective low energy
540: theory, the generalization of the Einstein theory in extra
541: dimensions, called the
542: Lovelock theory of gravity \cite{Lov}, when $\Phi$ is constant.
543:
544: \section{Holography}
545: The idea of Holography is to relate quantum physics on a spacetime boundary with
546: classical geometrical properties of the spacetime.
547:
548: Initially the aim of this theory was to understand the quantum physics of black holes.
549: Indeed when a black hole is formed,
550: classical and quantum physics encounter each other at its boundary, the horizon.
551: In particular all the quantum degrees of
552: freedom live holographically on the horizon. This arises from the study of the entropy
553: of a black hole. We know that the entropy is
554: the measure of the degrees of freedom of a physical state. Applying covariant quantization and
555: the analogies between thermodynamics and
556: black hole physics, Bekenstein and then Hawking \cite{holo1,holo2,holo3,H2} discovered
557: that the entropy of a black hole
558: is proportional to its area $A$:
559: \begin{equation}
560: S=\frac{A}{4G_N}.
561: \end{equation}
562: Quantum mechanically this was a breakthrough. Indeed quantum field theory (QFT)
563: gives an estimation of entropy
564: proportional to the volume and not to the area of a physical object.
565: This seems to say that QFT lives only on the horizon. Consider a spherical
566: region $\Gamma$ of volume $V$ in an asymptotically flat spacetime.
567: The boundary $\delta \Gamma$ has area $A$. The maximal entropy is
568: defined by
569: \begin{equation}
570: S_{max}=\ln N_{states}
571: \end{equation}
572: where $N_{states}$ is the total number of possible states of $\Gamma$.
573: Since we are considering gravitational objects, the number of states should
574: be linked with the degrees of freedom of the spacetime.
575: Suppose we consider that, at each fixed time, the space can be quantized. This means
576: that we divide it into elementary cells of width $\alpha$. Each of these cells store
577: the local information of a physical state. Moreover, assuming that for each cell
578: we have $m$ possible states, we get approximately
579: \begin{equation}
580: N_{states}=m^{V/\alpha^3}\ .
581: \end{equation}
582: This implies that the maximum entropy is
583: \begin{equation}
584: S_{max}\propto V.
585: \end{equation}
586: Now we consider a static star of energy $E$ and radius $R$. A static equilibrium implies that
587: the energy of the star $E$ must be bounded,
588: $E<M$. Where $M=R/2$ is the maximum mass one can fit in a static sphere or radius $R$.
589: The second law of thermodynamics tells us that an increase in energy implies an increment of
590: entropy. The maximum entropy will
591: then be reached when the black hole is formed. Indeed after the formation of a black hole,
592: all the degrees of freedom inside
593: the horizon are causally disconnected from the exterior, so that they
594: cannot be taken into account. This means that the maximum entropy is the
595: black hole entropy
596: \begin{equation}
597: S_{max}=\frac{A}{4G_N}.
598: \end{equation}
599: If we believe that we can calculate the entropy of a star by simple QFT calculations, we
600: implicitly break unitarity
601: of the quantum theory, losing predictability \cite{predict}. This is because there is a
602: transition of the number of states
603: from the collapsing star ($\ln N_{states}\propto V$) to the black hole ($\ln N_{states}
604: \propto A$). If instead
605: we accept that the maximum entropy of a
606: spatial region is proportional to the area of its boundary, rather then its volume, then we
607: can retain unitarity in the collapse process.
608: This is how the holographic principle was first formulated. In the extended version this
609: principle states that
610: for any Lorentzian
611: manifold it is possible to find a submanifold (screen) where all the quantum degrees of
612: freedom are present \cite{Bousso}.
613: The AdS/CFT correspondence \cite{M1} is a specific and explicit example. Here the screen
614: is identified as the AdS boundary.
615: In this holographic correspondence, a quantum conformal field theory on the screen,
616: can be
617: described as a boundary effect of a classical geometrical theory of spacetime.
618: If we believe seriously in this principle,
619: the evidence
620: that our world is described by quantum degrees of freedom, makes the study of gravity
621: in higher dimensions encouraging. In the following
622: I describe how the relationship between quantum physics and classical
623: gravity arises in an anti de Sitter manifold (AdS), which is
624: the geometry in which the Randall-Sundrum mechanism is constructed.
625:
626: \subsection{AdS/CFT correspondence}
627: This test of the holographic principle was developed in particular string theories.
628: Here it has been shown
629: that on $D$-dimensional AdS backgrounds there is a correspondence between a classical
630: perturbed $D$-dimensional super-gravity theory and a
631: $(D-1)$-dimensional super-conformal field theory.
632: This duality was then argued to be present also
633: in deformations of AdS, such as the generalization of the Randall-Sundrum model, where the
634: quantum theory of the boundary was not necessarily
635: a conformal field theory (for a very nice introduction to the topic see \cite{padilla}).
636: In order to have the flavour of this
637: duality, I give a simple
638: example of AdS/CFT correspondence. Here I review the result that the zero
639: point energy of a CFT theory is actually described
640: by the boundary energy of an anti de Sitter spacetime \cite{BK}.
641:
642: \subsubsection{The Brown-York tensor}
643: To define the energy of a manifold boundary we are going to use the definition of the
644: quasi-local energy given in \cite{BY}. The idea
645: is very simple and clear. We make an analogy with a classical non-relativistic system.
646: Suppose this system has an action $S$.
647: It satisfies the Hamilton-Jacobi equation $H=-\partial S/\partial t$,
648: where $H$ is the Hamiltonian of the system which describes the
649: energy and $t$ is the time. We would like now to generalize this concept for a gravitational
650: system. Take a $D$ dimensional manifold
651: $M$ which can be locally described by a product of a $D-1$ dimensional space $\Sigma$ and
652: a real line interval. The
653: boundary $\partial \Sigma$ need not be simply connected. The product
654: of $\partial \Sigma$ with the real line
655: orthogonal to
656: $\Sigma$ will be denoted by $B$. By analogy with classical mechanics, the quasi-local
657: energy associated with the spacelike
658: hypersurface $\Sigma$, is defined as minus the variation of the action with respect to a unit
659: increase in proper time separation between
660: $\partial \Sigma$ and its neighboring $D-2$ surface, as measured orthogonally to $\Sigma$
661: at $\partial \Sigma$. This basically measures
662: the variational rate of the action on $B$. In order to define this splitting of spacetime
663: we will use the ADM \cite{ADM}
664: decomposition which has a global splitting (at least in the absence of singularities or null
665: surfaces). Here we can naturally define an
666: Hamiltonian and therefore an energy.
667:
668: The spacetime metric is $g_{\mu\nu}$ and $n^\mu$ is the outward pointing spacelike unit
669: normal to the boundary $B$. The metric and
670: the extrinsic curvature of $B$ are denoted respectively by $\gamma_{\mu\nu}$ and
671: $\Theta_{\mu\nu}$.
672: Now denote by $u^\mu$ the future pointing timelike unit normal to a family of spacelike
673: hypersurfaces $\Sigma$ that
674: foliate the spacetime. The metric and the extrinsic curvature of $\Sigma$ are given by the
675: spacetime tensors $h_{\mu\nu}$ and
676: $K_{\mu\nu}$, respectively. $h^\mu_\nu$ is also the projection tensor on $\Sigma$. The
677: spatial coordinates $i,j,...=0,...,D-2$ are adapted
678: coordinates on $\Sigma$. We then define the momentum $P^{ij}$ conjugate to the spatial
679: metric $h_{ij}$. The ADM decomposition is
680: simply
681: \begin{equation}
682: ds^2=g_{\mu\nu}dx^\mu dx^\nu=-N^2dt^2+h_{ij}(dx^i+V^idt)(dx^j+V^jdt),
683: \end{equation}
684: where $N$ is the lapse function and $V^i$
685: the shift vector.
686:
687: For our purposes we use the fact that the foliation $\Sigma$ is orthogonal to $B$, which implies
688: $(u\cdot n)\Big|_B=0$. Because at this restriction,
689: the metric at $B$ can be decomposed as
690: \begin{equation}
691: \gamma_{\mu\nu}dx^\mu dx^\nu=-N^2dt^2+\sigma_{ab}(dx^a+V^adt)(dx^b+V^bdt),
692: \end{equation}
693: where $a,b,...=0,...,D-3$ are adapted coordinates on $\partial \Sigma$.
694:
695: For general relativity coupled to matter, consider first the action suitable for fixation
696: of the metric on the boundary \cite{York}
697: \begin{equation}
698: S=\frac{1}{2 \kappa}\int_M d^Dx \sqrt{-g} R+\frac{1}{\kappa}\int^{t_2}_{t_1}d^{D-1}x\sqrt{h}K-
699: \frac{1}{\kappa}\int_Bd^{D-1}x\sqrt{-\gamma}\Theta+S^m\ ,
700: \end{equation}
701: where $S^m$ is the matter action, including a possible cosmological term. Now the variation
702: of this action gives
703: \begin{eqnarray}
704: \delta S&=&(\mbox{terms giving the equations of motion})\cr
705: &+&(\mbox{boundary terms coming from the matter action})\cr &+&\int^{t_2}_{t_1}d^{D-1}
706: x P^{ij}\delta h_{ij}+\int_Bd^{D-1}x\pi^{ij}\delta\gamma_{ij}.
707: \end{eqnarray}
708: Here $\pi^{ij}$ is the conjugate momentum to $\gamma_{ij}$. We assume that the
709: matter action contains no derivatives of the metric. For the gravitational variables, the
710: boundary three-metric $\gamma_{ij}$ is
711: fixed on $B$, and the hypersurface metric $h_{ij}$ is fixed in $t_1$ and $t_2$. Since we are
712: interested in the variation of the action,
713: this will have an ambiguity in its definition. The ambiguity in $S$ is taken into account by
714: subtracting an arbitrary function of the
715: fixed boundary data. Thus we define the action
716: \begin{equation}
717: A=S-S^0,
718: \end{equation}
719: where $S^0$ is a functional of $\gamma_{ij}$. The variation in $A$ just differs from the
720: variation of $S$ by the term:
721: \begin{equation}\label{S0}
722: -\delta S^0=-\int_B d^{D-1}x\frac{\delta S^0}{\delta \gamma_{ij}}\delta\gamma_{ij}=-
723: \int_Bd^{D-1}x\pi_0^{ij}\delta \gamma_{ij},
724: \end{equation}
725: where it is clear that $\pi_0^{ij}$ is a function only of $\gamma_{ij}$. We call the
726: classical action $S_{cl}$ the action evaluated
727: on the classical solution of $A$. This will be of course a functional of the fixed boundary
728: data consisting of $\gamma_{ij}$,
729: $h_{ij}(t_1)$, $h_{ij}(t_2)$ and the matter fields. Then the variation of $S_{cl}$ among the
730: possible classical solutions give
731: \begin{eqnarray}
732: &\delta S_{cl}=(\mbox{terms involving variations in the matter fields})\cr \nonumber
733: &+\int^{t_2}_{t_1}d^{D-1}
734: x P^{ij}_{cl}\delta h_{ij}+\int_Bd^{D-1}x(\pi^{ij}_{cl}-\pi_{0}^{ij})\delta\gamma_{ij}.
735: \end{eqnarray}
736: The generalization of the Hamilton-Jacobi equation
737: for the momentum are the equations
738: \begin{equation}
739: P^{ij}_{cl}\Big|_{t_2}=\frac{\delta S_{cl}}{\delta h_{ij}(t_2)}.
740: \end{equation}
741: The generalization of the energy equation is
742: \begin{equation}\label{BY}
743: T^{ij}=\frac{2}{\sqrt{-\gamma}}\frac{\delta S_{cl}}{\delta\gamma_{ij}}\ ,
744: \end{equation}
745: which is the quasi-local stress-energy tensor for gravity.
746:
747: \subsubsection{Holographic zero point energy of AdS}
748:
749: In general, if the spacetime is not asymptotically flat the Brown-York tensor (\ref{BY})
750: diverges at infinity. For asymptotically AdS spacetime,
751: there is a resolution of this difficulty. If we believe there is an
752: holographic duality between AdS and CFT, the
753: fact that the Brown-York tensor diverges, it can be interpreted as the standard ultraviolet
754: divergences of
755: quantum field theory, and may be classically removed
756: by adding local counter-terms to the action. These subtractions depend only on the intrinsic
757: geometry of the boundary. Once
758: we
759: renormalize the energy in this way, we should be able via classical and quantum renormalization
760: schemes to obtain the same
761: zero point energy of the spacetime. The idea is to renormalize the stress-energy tensor by
762: adding a finite
763: series
764: of boundary invariants to the classical action.
765: The essential terms are fixed uniquely by requiring finiteness of the stress tensor. For
766: simplicity we will work it out in three dimensions,
767: but this is true in higher dimensions as well \cite{BK}.
768: In particular, global $\mbox{AdS}_5$, with an $S^3\times R$ boundary, has a positive
769: mass \cite{positive}.
770: This result is beautifully explained via the proposed
771: duality with a boundary CFT. The dual super Yang-Mills (SYM) theory on a sphere has a Casimir
772: energy that precisely matches this spacetime mass.
773:
774: The action we are considering now is the three-dimensional Einstein-Hilbert with negative
775: cosmological constant $\Lambda=-1/l^2$
776: and boundary terms
777: \begin{equation}
778: S=\frac{1}{2 \kappa}\int_M d^3x \sqrt{-g} \left(R-\frac{1}{l^2}\right)+
779: \frac{1}{\kappa}\int^{t_2}_{t_1}d^2x\sqrt{h}K-
780: \frac{1}{\kappa}\int_Bd^2x(\sqrt{-\gamma}\Theta)+S_{ct},
781: \end{equation}
782: where $S_{ct}$ is the counter-term action added in order to obtain a finite stress tensor.
783: In order
784: to preserve the equation
785: of motion, this extra term must be a boundary term. In particular it lives in $B$.
786: The Brown-York tensor is then
787: \begin{equation}
788: T^{ij}=\frac{1}{\kappa}\left[\Theta^{ij}-\Theta\gamma^{ij}+\frac{2}{\sqrt{-\gamma}}\frac{\delta
789: S_{ct}}{\delta\gamma_{ij}}\right],
790: \end{equation}
791: with everything calculated on the classical solution. Since $S_{ct}$ must only
792: depend on the boundary geometry, it describes a local
793: gravitational action in two dimensions. In particular, we wish to cancel out the divergences
794: coming from the negative energy from the
795: cosmological constant. Therefore the most general form is
796: \begin{equation}
797: S_{ct}=-\int_B\frac{1}{l}\sqrt{-\gamma}d^2 x\ \Rightarrow\ T^{ij}=\frac{1}{\kappa}\left[\Theta^{ij}-
798: \Theta\gamma^{ij}-\frac{1}{l}\gamma^{ij}\right].
799: \end{equation}
800: Now consider $\mbox{AdS}_3$ spacetime in light-cone coordinates
801: \begin{equation}\label{ads}
802: ds^2=\frac{l^2}{r^2}dr^2-r^2dx^+dx^-.
803: \end{equation}
804: In this case $T^{ij}=0$. What we would like to prove is that a dual CFT is living on the
805: surface $ds^2=-r^2dx^+dx^-$ with $r$
806: eventually taken to infinity (AdS boundary). If this is true, asymptotically conformal
807: perturbations of AdS
808: correspond to excitations of CFT fields on its boundary.
809: In particular, on the CFT side even if classically $T^i_i=0$, because quantum excitation
810: of the vacuum, $\gamma^{ij}<T_{ij}>\neq 0$. The scale of
811: this energy is linked to the renormalization scale. Connecting it with the
812: gravitational scale $l$ of AdS, we identify the Brown-York tensor on the boundary
813: of AdS as corresponding to the quantum CFT energy-momentum tensor on flat background.
814: In \cite{Brohe} Brown and Henneaux proved that a perturbation of the AdS metric with
815: asymptotic
816: AdS behaviour must have the following expansion
817: \begin{eqnarray}
818: \delta g_{+-}=O(1),\ \ \delta g_{++}=O(1),\ \ \delta g_{--}=O(1)\\
819: \delta g_{rr}=O\left(\frac{1}{r^4}\right),\ \ \delta g_{+r}=O\left(\frac{1}{r^3}\right),\ \
820: \delta g_{-r}=
821: O\left(\frac{1}{r^3}\right).
822: \end{eqnarray}
823: We rewrite this expansion using the following diffeomorphism
824: \begin{eqnarray}
825: x^+\rightarrow x^+-\xi^+-\frac{l^2}{2r^2}\partial^2_-\xi^-\ ,\label{diff1}\\
826: x^-\rightarrow x^--\xi^--\frac{l^2}{2r^2}\partial^2_+\xi^+\ ,\\\
827: r\rightarrow r+\frac{r}{2}(\partial_+\xi^++\partial_-\xi^-)\ ,\label{diff2}
828: \end{eqnarray}
829: where $\xi^\pm=\xi^\pm(x^\pm)$, which yields at
830: first order
831: \begin{equation}
832: ds^2=\frac{l^2}{r^2}dr^2-r^2dx^+dx^--\frac{l^2}{2}\partial^3_+\xi^+ (dx^+)^2-
833: \frac{l^2}{2}\partial^3_- \xi^-(dx^-)^2.
834: \end{equation}
835: Then we obtain the following non-zero components for the Brown-York tensor
836: \begin{equation}\label{grav}
837: \delta_{\xi^{\pm}}T_{\pm\pm}=-\frac{l}{2k}\partial_\pm^3\xi^\pm.
838: \end{equation}
839: This therefore represents the zero point energy of the boundary of AdS
840: due to perturbations of this spacetime that preserve the AdS symmetries on the boundary.
841: Now we study the
842: CFT side. The boundary of AdS is just a Minkowski spacetime with metric
843: \begin{equation}
844: ds^2=-r^2dx^+dx^-.
845: \end{equation}
846: A conformal theory is such that under conformal transformation of the type
847: $g_{\alpha\beta}\rightarrow \Omega \tilde g_{\alpha\beta}$, it is
848: classically invariant; then its energy momentum tensor satisfies $T^{\alpha\beta}=\tilde
849: T^{\alpha\beta}$. In particular this means that
850: classically $T^\alpha_\alpha=0$. If we now introduce a conformal transformation under
851: the diffeomeorphism (\ref{diff1}-\ref{diff2}), when
852: $r\rightarrow \infty$ we get quantum mechanically (starting again from a zero energy
853: momentum tensor)
854: \footnote{It
855: is beyond the scope of this thesis to prove this result; a very good introduction can be found
856: in \cite{ginsparg}.}
857: \begin{equation}\label{cft}
858: \delta_{\xi^\pm} T_{\pm\pm}=-\frac{c}{24\pi}\partial_\pm^3\xi^\pm.
859: \end{equation}
860: This term comes from the commutation rules.
861: In particular since this does not depend on the particular
862: energy-momentum tensor used, it is called the zero point energy.
863: The constant $c$ is called the
864: central charge and encode the renormalization scale of
865: the theory. Surprisingly eqs. (\ref{grav}) and (\ref{cft}) coincide if we measure
866: the scale of the renormalization with the AdS
867: scale, or $c=12\pi l/\kappa$. This means that the quantum vacuum energy, due
868: to the excitation of the fields in the CFT side, is nothing else than the residual
869: energy due to classical gravitational perturbations of AdS that preserve the AdS boundary symmetries.
870:
871: \newpage
872:
873: \chapter{A dynamical alternative to compactification}
874:
875: The possibility that spacetime has more than four dimensions is an
876: old idea. Originally the extra dimensional spaces were
877: thought to be compactified. This kind of model is usually
878: called Kaluza-Klein compactification (see for example \cite{KK}).
879: Here, to have a correct Newtonian and Standard Model limit, the
880: size of the extra dimensions must be less than the Electroweak
881: scale ($ \sim 1\ \mbox{TeV}^{-1}$). These theories were introduced to
882: incorporate geometrically scalar and vector fields in the
883: projected four-dimensional gravity, thus giving a geometrical interpretation of electromagnetism.
884:
885: Later, extra dimensions arose very naturally in string theory and associated phenomenology.
886:
887: In this thesis I will discuss only models with an infinitely large extra
888: dimension. In particular, I am interested in models where
889: gravity is dynamically localized at low energies \cite{RS}.
890: Another type of localization is also possible \cite{DGP,GRS}, in which
891: gravity has the correct Newtonian limit
892: at short distances. This arises from introducing an induced
893: gravity on the four-dimensional submanifold. In this way, at short
894: distances, the induced effects dominate, generating an effective
895: four-dimensional Einstein gravity consistent with Newtonian gravity
896: experiments. At large distances, by contrast, the
897: extra-dimensional gravity becomes more and more important. This
898: modifies, at large distances, the effective four-dimensional
899: gravity.
900:
901: The fact that particles can be trapped gravitationally in a submanifold,
902: was argued already by Visser \cite{Visser}. He realized that
903: a particle living in a five-dimensional spacetime with warped time component,
904: \begin{equation}\label{w1}
905: ds^2=-e^{-2\Phi(\xi)}dt^2+d{\vec x}\cdot d\vec x+d\xi^2,
906: \end{equation}
907: ($\xi$ is the extra-dimensional coordinate) is exponentially
908: trapped by the scalar field $\Phi(\xi)>0$. In 1999 Randall and
909: Sundrum \cite{RS} used a similar mechanism to localize gravity. In
910: order to do that, the whole four-dimensional metric must be warped.
911: In this way a curved background supports a ``bound state'' of the
912: higher-dimensional graviton with respect to the co-dimensions.
913: So although space is indeed infinite in extent, the
914: graviton is confined to a small region within this space.
915:
916: Suppose for simplicity we have only one extra dimension
917: \footnote{The mechanism is general even with more than one co-dimension.}.
918: Moreover suppose the five-dimensional gravity is governed by the Einstein-Hilbert action.
919: We then consider a domain wall with
920: vacuum energy $\lambda$ where, as a macroscopic approximation, all the matter fields live.
921: The action is then
922: \begin{equation}\label{act}
923: S=\frac{1}{\tilde\kappa^2}\int d^5x\sqrt{-\tilde
924: g}\{\tilde R-2\tilde\Lambda\}+\int d^4x\sqrt{-g}\{{\cal L}_m-2\lambda\},
925: \end{equation}
926: where $\tilde \kappa$ and $\tilde g_{AB}$ are respectively the five
927: dimensional Planck mass and metric, $\tilde\Lambda$ and $\tilde R$ are the five
928: dimensional negative cosmological constant and Ricci scalar, ${\cal L}_m$ and
929: $g_{\mu\nu}$ are the four-dimensional Lagrangian for matter and
930: the four-dimensional metric. The three-dimensional submanifold is
931: called the ``brane'' and the total spacetime will be
932: called the ``bulk''. We use the notation $A,B,...=0,...,4$ for
933: the bulk coordinates and $\mu,\nu,...=0,...,3$ for the brane
934: coordinates. The role of the five-dimensional cosmological
935: constant is very important. Physically, we can understand that
936: such a term will induce a positive ``pressure''. This localizes the
937: gravitons produced in the four-dimensional brane on the brane
938: itself. Indeed, as we shall see, the warp factor
939: is produced by the cosmological constant.
940:
941: In order to simplify the system we introduce
942: $Z_2$-symmetry in the co-dimension. This can be also motivated from the orbifold structure
943: appearing in M-theory to
944: incorporate heterotic string theory. In this way M-theory contains the
945: gauge groups of the standard model of particle physics \cite{HW}.
946:
947: We start with an empty brane, ${\cal L}_m=0$. We would like to find a warped
948: solution which has, as a slice, a Minkowski
949: spacetime. In the next section we will introduce the
950: general formalism in the presence of matter, then we will study small perturbations
951: around Minkowski, obtaining the correct Newtonian limit.
952: We start with the metric:
953: \begin{equation}\label{ansatz}
954: ds^2=e^{-2\sigma(y)}\gamma_{\mu\nu}(x^\alpha)dx^\mu dx^\nu+dy^2,
955: \end{equation}
956: where $y$ is the extra-dimensional coordinate and
957: $\gamma_{\mu\nu}(x^\alpha)$ is the four-dimensional metric. In order to
958: solve the variational problem $\delta S=0$, we have two ways. The first one is
959: to solve the Einstein equations together with junction
960: conditions between the brane and the bulk. Since the brane is a singular object, with this
961: approach we need to involve the theory of
962: distributions. We will discuss this technique in the next section and study it extensively
963: in the last chapter and Appendix \ref{Appjc}.
964: Here we follow another way. Instead of introducing a
965: thin shell (brane), we consider a spacetime with a boundary at
966: $y=0$. This means that we will allow the four-dimensional
967: metric to vary only on the boundary, whereas we will keep it fixed in the
968: bulk as Minkowski slices, in such a way that the global spacetime is AdS.
969: If we would like to
970: implement the $Z_2$ symmetry, we can just make a copy of the
971: same spacetime on the two sides of the boundary. In this way we
972: don't have to deal with distributions. Now if we substitute the
973: ansatz (\ref{ansatz}) into the action (\ref{act}), we are left
974: with a variational problem for the scalar field $\sigma(y)$ in the
975: bulk and a variational problem for the metric $\gamma_{\mu\nu}$ on
976: the brane. Then we use the constraint
977: $\gamma_{\mu\nu}=\eta_{\mu\nu}$. The
978: Lagrangian for gravity will be \cite{Wald}\footnote{I use the notation $\partial_y f= f'$.}
979: \begin{equation}
980: \sqrt{-\tilde g}\tilde R=12e^{-4\sigma}(\sigma')^2+2(K\sqrt{-g})'+e^{-2\sigma}{\cal R}\ ,
981: \end{equation}
982: where ${\cal R}$ is the Ricci scalar associated with $\gamma_{\mu\nu}$ and $K={1\over 2}
983: g^{\mu\nu}\pounds_n g_{\mu\nu}$,
984: is the trace of the extrinsic curvature orthogonal to the brane. Using this notation we get
985: \begin{equation}
986: \int d^5x\sqrt{-\tilde g}\tilde R=12\int d^5x \{e^{-4\sigma}(\sigma')^2+e^{-2\sigma}
987: {\cal R}\}+2\int d^4x K \sqrt{-g},
988: \end{equation}
989: where the boundary is at $y=0$. The last term is called the Gibbons-Hawking term \cite{GW}.
990:
991: Since the degrees of freedom of the four-dimensional Ricci scalar and the degrees of freedom
992: of the scalar field $\sigma$ are now
993: completely separate, we can set ${\cal R}=0$ before the variation.
994: Considering only the first integral, since $\delta \sigma=0$ on the boundary we have
995: \begin{equation}
996: 12\delta\int d^5x e^{-4\sigma}(\sigma')^2=-12\int d^5x \{2\sigma''-(4\sigma')^2\}e^{-4\sigma}
997: \delta\sigma.
998: \end{equation}
999: The variation of the five dimensional cosmological constant is
1000: \begin{equation}
1001: -2\delta\int d^5x\sqrt{-\tilde g}\tilde\Lambda=12\int d^5x \frac{2}{3}\tilde\Lambda
1002: e^{-4\sigma}\delta\sigma.
1003: \end{equation}
1004: Combining the two variations we get the equation
1005: \begin{equation}
1006: \sigma''-2(\sigma')^2=\frac{\tilde\Lambda}{3}.
1007: \end{equation}
1008: This equation has two separate branches, one is the trivial
1009: \begin{equation}
1010: \sigma'=\pm\sqrt{-\frac{\tilde\Lambda}{6}}\ ,
1011: \end{equation}
1012: and the other is
1013: \begin{equation}
1014: \sigma'=\pm\sqrt{1+\tanh \left(2\sqrt{-\frac{\tilde\Lambda}{6}}y+C\right)}.
1015: \end{equation}
1016: To implement the $Z_2$-symmetry we have to cut and paste the solution from one side to the
1017: other of the boundary. The second solution does not have parity. Moreover, the Einstein tensor
1018: describing this system contains another constraint equation
1019: hidden in the integration process we are using here.
1020: One can show that the second solution does not satisfy this additional constraint (see the $\tilde G_{yy}$ equation in \ref{second}
1021: with $e^{-\sigma(y)}=a(y)$, and considering the static case).
1022: This implies that we can only use the first solution
1023: \begin{equation}\label{jcRS}
1024: \sigma'=\sqrt{-\frac{\tilde\Lambda}{6}}\mbox{sgn}(y).
1025: \end{equation}
1026: Here we introduced the sign function $\mbox{sgn}(y)$, in order to describe the full spacetime.
1027: In particular we choose the positive
1028: branch of the square root. This is because we want a decreasing warp factor.
1029: Now we have to make the variations of the boundary action with
1030: respect to $g_{\mu\nu}$ vanish, keeping $\partial_y g_{\mu\nu}$ fixed. We have
1031: the combination
1032: \begin{equation}
1033: \frac{1}{\tilde\kappa^2}\int
1034: d^4x\sqrt{-g}\left\{ K_{\mu\nu}-g_{\mu\nu}K-\frac{\tilde\kappa^2\lambda}{2}
1035: g_{\mu\nu}\right\} \delta g_{\mu\nu},
1036: \end{equation}
1037: using the constraint $\gamma_{\mu\nu}=\eta_{\mu\nu}$, this boundary action vanishes for
1038: \begin{equation}\label{jcRS2}
1039: \sigma'(0)=\frac{\tilde\kappa^2\lambda}{6}.
1040: \end{equation}
1041: Combining (\ref{jcRS}) with (\ref{jcRS2}) we get the fine tuning
1042: \begin{equation}\label{min}
1043: \frac{\tilde \kappa^4\lambda^2}{6}+\tilde\Lambda=0.
1044: \end{equation}
1045: We can now write down the global geometry. This is an AdS spacetime in the bulk and
1046: a Minkowski
1047: spacetime on the brane:
1048: \begin{equation}\label{Minkmetric}
1049: ds^2=e^{-2\sqrt{-\tilde\Lambda/6}\mid y\mid}\eta_{\mu\nu}dx^\mu dx^\nu+dy^2.
1050: \end{equation}
1051: In the next paragraph we show how
1052: generalize this model in the presence of matter.
1053: Moreover we will see that if the fine tuning (\ref{min}) does not hold, an
1054: effective four-dimensional cosmological constant appears.
1055:
1056: \section{The Randall-Sundrum braneworld scenario}
1057: The mechanism provided by Randall-Sundrum can be generalized to include
1058: matter. We will follow
1059: \cite{SMS}. In this work a geometrical approach is used. We suppose that
1060: the bulk spacetime is governed by the Einstein
1061: equations and then we impose the junction conditions.
1062:
1063: The normal to the brane is $n^A$. The induced metric on
1064: the brane is $g_{AB}= \tilde g_{AB}-n_A n_B$. The Gauss equation
1065: which relates the $D-$dimensional Riemann tensor to the
1066: $(D-1)$-dimensional one is\footnote{I will use the notation with
1067: tilde for five-dimensional objects and without for
1068: four-dimensional ones.}
1069: \begin{equation}\label{gauss}
1070: R^A{}_{BCD}=\five R^E{}_{FGH}\ g^{A}_{~E}\ g^{F}_{~B}\ g^{G}_{~C}\ g^{H}_{~D}\ +K^A{}_C K_{BD}
1071: -K^A{}_D K_{BC}
1072: \end{equation}
1073: where
1074: \begin{equation}
1075: K_{AB}=\frac{1}{2}\pounds_n \tilde g_{AB}
1076: \end{equation}
1077: is the extrinsic curvature of the brane. Then we have the Codacci equations
1078: \begin{equation}\label{codacci}
1079: g^{B}_{~C}\nabla_B K^{C}{}_A- g^{B}_{~A}\nabla_B K=\five R_{CD}n^C g^{D}_{~A},
1080: \end{equation}
1081: where we use the notation $K^A{}_A=K$. From the Gauss equation we
1082: can build up the four-dimensional Einstein tensor
1083: \ba
1084: G_{AB}&=&\five G_{CD}\ g^C_{~A}\ g^D_{~B}\ +\five R_{CD}\ n^C\ n^D\ g_{AB}+KK_{AB}\cr
1085: &-&K^C{}_A\ K_{BC}-
1086: \frac{1}{2}g_{AB}(K^2-K^{AB}\ K_{AB})-\tilde
1087: E_{AB}\ ,
1088: \ea
1089: where we introduced
1090: \begin{equation}
1091: \tilde E_{AB}\equiv\five R^C{}_{DFG}\ n_C\ n^F\ g^{D}_{~A}\ g^{G}_{~B}.
1092: \end{equation}
1093: Now we require that the bulk spacetime is governed by the five-dimensional Einstein equations
1094: \begin{equation}\label{einstein5}
1095: \tilde G_{AB}=\tilde\kappa^2 \tilde T_{AB}\ ,
1096: \end{equation}
1097: where $\tilde T_{AB}$ is the five dimensional energy-momentum tensor. We
1098: then use the decomposition of the Riemann tensor
1099: into Weyl curvature, Ricci tensor and scalar curvature,
1100: \begin{equation}\label{decRie}
1101: \five R_{ABCD}=\frac{2}{3}\left(\tilde g_{A[C}\five R_{D]B}-\tilde g_{B[C}
1102: \five R_{D]A}\right)-\frac{1}{6}\tilde g_{A[C}\tilde g_{D]B}\five R+
1103: \five C_{ABCD}\ ,
1104: \end{equation}
1105: where $\tilde C_{ABCD}$ is the Weyl tensor or the traceless part
1106: of the Riemann tensor. Finally the four-dimensional Einstein
1107: tensor in adapted coordinates to the brane is
1108: \ba
1109: G_{\mu\nu}
1110: &=& {2 \tilde \kappa^2 \over 3}\left[T_{\mu\nu}
1111: +\left(T_{\rho\sigma}n^\rho n^\sigma-{1 \over 4}T^\rho_{~\rho}\right)
1112: g_{\mu\nu} \right]\cr
1113: &+& KK_{\mu\nu}
1114: -K^{~\sigma}_{\mu}K_{\nu\sigma} -{1 \over 2}g_{\mu\nu}
1115: \left(K^2-K^{\alpha\beta}K_{\alpha\beta}\right) - \tilde {\cal E}_{\mu\nu},
1116: \label{4dEinstein}
1117: \ea
1118: where
1119: \begin{equation}
1120: \tilde {\cal E}_{AB} \equiv \tilde C^C_{~DEF}n_C n^E\
1121: g_A^{~D}\ g_B^{~F},
1122: \end{equation}
1123: and it is traceless. From the Codacci equations (\ref{codacci}) and the five
1124: dimensional Einstein equations (\ref{einstein5}) we get
1125: \begin{equation}\label{quasicon}
1126: g^{~B}_C\nabla_B K^{~C}_A- g^{~B}_A\nabla_B K=\five T_{CD}n^C g^{~D}_A.
1127: \end{equation}
1128: Up to this point we have not introduced a braneworld. A braneworld is
1129: a four-dimensional hypersurface which can be described by the
1130: equation $y=0$ where $y$ is the extra-dimensional
1131: coordinate. Close to the brane we can use Gaussian normal
1132: coordinates such that the metric has the form
1133: \begin{equation}
1134: ds^2=g_{\mu\nu}dx^\mu dx^\nu+dy^2.
1135: \end{equation}
1136: The energy momentum tensor is
1137: \begin{equation}
1138: \tilde T_{AB}=-\frac{\tilde\Lambda}{\tilde \kappa^2} g_{AB}+\delta(y)\left(-\lambda g_{AB}+
1139: T_{AB}\right)\ ,
1140: \end{equation}
1141: where we assume that, at macroscopic level, the matter ($T_{AB}$), is defined only on the brane.
1142: This is
1143: mathematically achieved by introducing a
1144: Dirac distribution $\delta(y)$. $\lambda$ is the vacuum energy of the brane which
1145: coincides with its tension for $T_{AB}=0$\footnote{The tension of a hypersurface is defined as its
1146: extrinsic curvature.}. In particular
1147: we also suppose that there is not a energy-momentum exchange between bulk and brane, or
1148: \begin{equation}
1149: T_{AB}\ n^A=0.
1150: \end{equation}
1151: As I will show in Appendix \ref{Appjc}, reasonable junction conditions require that the singular
1152: behaviour of the metric is encoded in the singular behaviour of the Lie
1153: derivative of the extrinsic curvature along the normal. From that one can get
1154: Israel's junction conditions \cite{israel},
1155: \begin{eqnarray} \label{jcsms}
1156: \left[g_{\mu\nu}\right]
1157: &=&0\,,
1158: \nonumber\\
1159: \left[K_{\mu\nu}\right]&=& -\tilde\kappa^2
1160: \Bigl(T_{\mu\nu}-\frac{1}{3}g_{\mu\nu}(T-\lambda)\Bigr),
1161: \end{eqnarray}
1162: where $[X]:=\lim_{y \to +0}X - \lim_{y \to -0}X=X^+-X^-$.
1163: Substituting then the junction conditions (\ref{jcsms}) into (\ref{4dEinstein}), we get the
1164: effective four-dimensional gravitational
1165: equations
1166: \begin{eqnarray}
1167: G_{\mu\nu}=-\Lambda g_{\mu\nu}
1168: + 8 \pi G_N T_{\mu\nu}+\frac{48\pi G_N}{\lambda}\,S_{\mu\nu}
1169: -{\cal E}_{\mu\nu}\,, \label{eq:effective}
1170: \end{eqnarray}
1171: where
1172: \begin{eqnarray}
1173: {\cal E}_{\mu\nu}&=&{1\over 2}\left[\tilde {\cal E}_{\mu\nu}\right]\,,\\
1174: \Lambda &=&\frac{1}{2}\left(\tilde\Lambda
1175: +\frac{1}{6}\tilde\kappa^4\,\lambda^2\right)\,,
1176: \label{Lamda4}\\
1177: G_N&=&{\tilde \kappa^4\,\lambda\over48 \pi}\,,
1178: \label{GNdef}\\
1179: S_{\mu\nu}&=&
1180: -\frac{1}{4} T_{\mu\alpha}T_\nu^{~\alpha}
1181: +\frac{1}{12}T T_{\mu\nu}
1182: +\frac{1}{8}g_{\mu\nu}T_{\alpha\beta}T^{\alpha\beta}-\frac{1}{24}
1183: g_{\mu\nu}T^2\ .
1184: \label{pidef}
1185: \end{eqnarray}
1186: From the effective equations (\ref{eq:effective}), the localization of gravity is clearly
1187: obtainable in the limit
1188: $\lambda\rightarrow\infty$, $\tilde\kappa\rightarrow 0$, such that $G_N$ remains finite,
1189: and ${\cal E}_{\mu\nu}\rightarrow 0$. Indeed
1190: in this limit we recover Einsteinian gravity. The fact that $G_N>0$ constrains
1191: $\lambda$ to be positive \cite{CGKT}.
1192:
1193: We now study the matter conservation
1194: equations. Since $T_{AB}n^A=0$ from (\ref{quasicon}) and using the
1195: junction conditions
1196: \begin{equation}
1197: \nabla^\mu T_{\mu\nu}=0.
1198: \end{equation}
1199: This implies that matter follows the standard conservation equations of general relativity.
1200: Thanks to the
1201: four-dimensional Bianchi identities we get differential equations for the projection of the
1202: Weyl tensor, ${\cal E}_{\mu\nu}$. These effective gravitational
1203: equations do not in general close the system.
1204: Indeed the full determination of the projected Weyl tensor requires in general the study of the
1205: five-dimensional gravitational problem.
1206: As we will see, we can sometimes
1207: get around that by introducing particular symmetries for the metric. Then this formalism
1208: becomes a very
1209: powerful tool.
1210:
1211: \section{1+3 formalism in the braneworld}
1212: In the last section we commented that the four-dimensional effective equations are not in
1213: general closed. This is because the tensor ${\cal E}_
1214: {\mu\nu}$ is not in general completely constrained by the projected equations. However,
1215: often four-dimensional symmetries constrain sufficiently
1216: the five-dimensional geometry, and families of solutions can be found.
1217: To understand better this concept it is very useful to
1218: use the formalism developed by Maartens (see for example \cite{maacov}) which I explain in the
1219: following. In particular since we will be
1220: interested only in perfect fluids, we can drastically simplify all the equations, using the
1221: perfect fluid energy-momentum
1222: tensor
1223: \begin{equation}\label{decT}
1224: T_{\mu\nu}=\rho u_\mu u_\nu+ p h_{\mu\nu},
1225: \end{equation}
1226: where $u^\mu$ is the unit four-velocity of the matter ($u^\mu u_\mu=-1$) and $h_{\mu\nu}$
1227: is the space-like metric that projects
1228: orthogonal to $u^\mu$ ($h_{\mu\nu}=g_{\mu\nu}+u_\mu u_\nu$). $\rho$
1229: and $p$ are respectively the energy density
1230: and the pressure of the perfect fluid.
1231: In the same way as (\ref{decT}), we decompose the local correction $S_{\mu\nu}$ and the
1232: non-local one ${\cal E}_{\mu\nu}$. ${\cal E}_{\mu\nu}$
1233: gives a non-local contribution in the sense that it codifies all the bulk geometrical
1234: back-reaction on the brane.
1235: The decomposition of the matter-correction is
1236: \be
1237: S_{\mu\nu}={{1\over12}}\rho\left[\rho u_\mu u_\nu +\left(\rho+2
1238: p\right)h_{\mu\nu}\right]\,.
1239: \ee
1240: Using the fact that ${\cal E}_{\mu\nu}$ is traceless we can decompose it as
1241: \footnote{ $\kappa^2=8\pi G_N$ .}
1242: \be
1243: -{1\over \kappa^2} {\cal E}_{\mu\nu} = \cu\left(u_\mu u_\nu+{ {1\over3}}
1244: h_{\mu\nu}\right)+ {\cq_\mu} u_{\nu} + {\cq_\nu} u_{\mu}+\cp_{\mu\nu}\, ,
1245: \ee
1246: where we introduced an effective ``dark" radiative
1247: energy-momentum on the brane, with energy density $\cal U$,
1248: pressure ${\cal U}/3$, momentum density $Q_\mu$ and anisotropic
1249: stress ${\Pi}_{\mu\nu}$.
1250:
1251: The brane-world corrections can conveniently be consolidated into
1252: an effective total energy density, pressure, momentum density and
1253: anisotropic stress \cite{m}:
1254: \begin{eqnarray}
1255: \rho^{\rm eff} &=& \rho\left(1 +\frac{\rho}{2\lambda} +
1256: \frac{\cal U}{\rho} \right)\,,\label{reff} \\ \label{peff}
1257: p^{\rm eff } &=& p + \frac{\rho}{2\lambda}
1258: (2p+\rho)+\frac{\cal U}{3}\;, \\ q^{\rm eff }_\mu &=& Q_\mu\;, \\
1259: \label{e:pressure2} \pi^{\rm eff }_{\mu\nu} &=& \Pi_{\mu\nu}\;.
1260: \end{eqnarray}
1261: Note that nonlocal bulk effects can contribute to effective
1262: imperfect fluid terms even when the matter on the brane has
1263: perfect fluid form: there is in general an effective momentum
1264: density and anisotropic stress induced on the brane by
1265: the 5D graviton.
1266:
1267: The effective total equation of state and sound speed follow from
1268: eqs.~(\ref{reff}) and (\ref{peff}) as
1269: \bea
1270: w^{\rm eff} &\equiv & {p^{\rm eff}\over\rho^{\rm eff}} =
1271: {w+(1+2w)\rho/2\lambda+ \cu/3\rho \over 1+\rho/2\lambda +\cu/\rho}
1272: \,,\label{vh1}\\ c_{\rm eff}^2 &\equiv & {\dot{p}^{\rm
1273: eff}\over\dot{\rho}^{\rm eff}}= \left[c_{\rm s}^2+{\rho+p \over
1274: \rho+\lambda} +{4\cu\over 9(\rho+p)(1+\rho/\lambda)}\right]
1275: \left[1+ {4\cu\over 3(\rho+p)(1+\rho/\lambda)}\right]^{-1}
1276: \, \label{vh2}
1277: \eea
1278: where $w=p/\rho$ and $c_{\rm s}^2=\dot p/\dot\rho$. We also used the notation
1279: $\dot f=d f/d\tau$ where $\tau$ is the proper
1280: time of the perfect fluid.
1281:
1282: \subsection{Conservation equations}
1283: ${\cal E}_{\mu\nu}$, the projection of the bulk Weyl tensor on the
1284: brane, encodes corrections from the 5D graviton effects (often called Kaluza-Klein or KK
1285: modes). From
1286: the brane-observer viewpoint, the energy-momentum corrections in
1287: $S_{\mu\nu}$ are local, whereas the KK corrections in ${\cal
1288: E}_{\mu\nu}$ are nonlocal, since they incorporate 5D gravity wave
1289: modes. These nonlocal corrections cannot be determined purely from
1290: data on the brane. In the perturbative analysis
1291: which leads
1292: to the corrections in the gravitational potential,
1293: the KK modes that generate this correction are
1294: responsible for a nonzero ${\cal E}_{\mu\nu}$; this term is what
1295: carries the modification to the weak-field field equations. An equivalent picture is that
1296: these modes arise as a geometrical bulk
1297: back-reaction to the variations of
1298: the matter fields on the brane. We can see how the matter can source these modes as follows.
1299:
1300: The standard conservation equations
1301: \begin{equation}
1302: \nabla^\mu T_{\mu\nu}=0
1303: \end{equation}
1304: together with the four-dimensional Bianchi identities applied to the effective
1305: four-dimensional gravitational equations
1306: (\ref{eq:effective}) lead to the following non-local equations
1307: \begin{equation}\label{non-local}
1308: \nabla^\mu {\cal E}_{\mu\nu}=\frac{6 \kappa^2}{\lambda}\nabla^\mu S_{\mu\nu}.
1309: \end{equation}
1310: It is clear that the projection of the Weyl tensor, which encodes the curvature of
1311: the co-dimension, can be sourced (as
1312: a back-reaction) by the variation of the matter fields.
1313:
1314: It is useful to rewrite the local and non-local conservation equation in a $1+3$ formalism.
1315: Introducing the covariant projected Levi-Civita
1316: tensor $\ep_{abc}$, the spatial covariant
1317: derivative $D_a$ (where $D_a S^{b...}{}_{...c}=h^e{}_a h^b{}_f...h^g{}_c \nabla_e S^{f...}
1318: {}_{...g}$),
1319: the volume expansion $\Theta=\nabla^\alpha u_\alpha$, the
1320: proper time derivative $\dot S^{a...}{}_{...b}=u^\alpha
1321: \nabla_\alpha S^{a...}{}_{...b}$, the acceleration $A_a=\dot u_a$, the shear
1322: $\sigma_{ab}=D_{(a} u_{b)}-\Theta/3 h_{ab}$ and the
1323: vorticity $\omega_a=-{1 \over 2}\mbox{curl}\
1324: u_a$, we get the local conservation equations
1325: \begin{eqnarray}
1326: &&\dot{\rho}+\Theta(\rho+p)=0\,,\label{pc1}\\ && \D_ a
1327: p+(\rho+p)A_ a =0\,,\label{pc2}
1328: \end{eqnarray}
1329: and the nonlocal equations
1330: \begin{eqnarray}
1331: && \dot{\cu}+{{4\over3}}\Theta{\cu}+\D^ a\cq_ a+2A^ a\cq_
1332: a+\sigma^{ ab }\cp_{ ab }=0\,, \label{pc1'}\\&& \dot\cq_{
1333: a}+{{4\over3}}\Theta\cq_ a +{{1\over3}}\D_
1334: a{\cu}+{{4\over3}}{\cu}A_ a +\D^ b\cp_{ ab }+A^ b\cp_{ ab
1335: }+\sigma_{ a}{}^b\cq_b-\ep_{a}{}^{bc}\omega_b\cq_c
1336: \nonumber\\&&~~{} =-{(\rho+p)\over\lambda} \D_a\rho\,.\label{pc2'}
1337: \end{eqnarray}
1338: The non-local equations do not contain evolution equations for the anisotropic part of the
1339: projected Weyl tensor $\Pi_{ab}$.
1340: This makes the system of ``brane'' equations not closed. However if we
1341: ask for spatially homogeneous and isotropic solutions, the system becomes closed.
1342:
1343: \section{Newtonian limit: localization of gravity}\label{secNewton}
1344:
1345: Equations (\ref{eq:effective})
1346: show that in the limit $\lambda\rightarrow \infty$ and in the case of a
1347: bulk with ${\cal E}_{\mu\nu}=0$,
1348: we recover
1349: General Relativity. In this section we discuss the correction to the Newtonian
1350: potential due to a point-like particle. In order to do that we have to study the full five
1351: dimensional
1352: equations because such a perturbation is due to purely
1353: five dimensional effects. In doing that we follow \cite{DD}.
1354:
1355: We start by considering the unperturbed Randall-Sundrum braneworld.
1356: The solution for the metric in Gaussian normal
1357: coordinates is given by eq. (\ref{Minkmetric}), i.e.
1358: \be
1359: ds^2=dy^2+e^{-2\mid y\mid/l}\eta_{\mu\nu}dx^\mu dx^\nu\ ,
1360: \ee
1361: where ${l}$ is the curvature scale of the AdS spacetime, ${l}=\sqrt{-6/\tilde \Lambda}$.
1362: In order to solve the perturbations of this metric it is
1363: better to use a conformally Minkowskian coordinate system.
1364: Consider only one side of the spacetime, say $y\geq 0$. In this
1365: side, we can use the transformation of coordinates
1366: \be
1367: e^{y/l}dy=dz,
1368: \ee
1369: and then the metric transforms to
1370: \be
1371: ds^2=\left(\frac{l}{z}\right)^2\eta_{AB}dx^A dx^B.
1372: \ee
1373: In these coordinates the brane is at $z={l}$.
1374:
1375: We now perturb the Minkowskian metric. We also expect that,
1376: due to these perturbations, the position of the brane will change, generating a ``bending''.
1377: The metric is
1378: \be\label{metan}
1379: ds^2=\left(\frac{l}{z}\right)^2\left(\eta_{AB}+\gamma_{AB}\right)dx^A dx^B,
1380: \ee
1381: where $\gamma_{AB}$ is a perturbation and the position of the brane is at
1382: \be
1383: z={l}+\xi,
1384: \ee
1385: where $\xi\ll {l}$ is a function describing the bending. We have now the freedom
1386: to choose the transversal gauge
1387: $\gamma_{Az}=0$, and in addition we can use the traceless gauge
1388: $\gamma^\mu{}_\mu=\partial_\rho \gamma^\rho{}_\mu=0$.
1389: In particular we expect that the bending function will depend only on the four-dimensional
1390: brane coordinates, since the bending is due
1391: to a four-dimensional perturbation, therefore we require $\partial_z \xi=0$. Under these
1392: conditions, the unit normal to the brane
1393: is
1394: \be
1395: n^A=-\frac{z}{l}\left(\delta^A_z-\delta^A_\mu\partial^\mu \xi\right).
1396: \ee
1397: From this we can calculate the extrinsic curvature to first order in perturbations
1398: \footnote{$\partial_{\mu\nu}=\partial_\mu
1399: \partial_\nu$.}
1400: \be K^A{}_B=-\frac{1}{2} \tilde g ^{AC}\pounds_n g_{CB}\simeq \delta^\nu{}_B
1401: \delta^A{}_\mu\left(\frac{1}{l}\eta^\mu{}_\nu-\partial^\mu{}_\nu\xi
1402: -\frac{1}{2}\partial_z \gamma^\mu{}_\nu\right). \ee From the
1403: junction conditions (\ref{jcsms}) and using the fine tuning
1404: (\ref{min}) we get on the brane $\Sigma$
1405: \be\label{pluba}
1406: \frac{\tilde\kappa^2}{2}\left(T_{\mu\nu}-{1 \over 3}
1407: g_{\mu\nu}T\right)=\left(\partial_{\mu\nu} \xi-\frac{1}{2}\partial_z
1408: \gamma_{\mu\nu}\right)\Big|_\Sigma. \ee
1409: Taking now the trace and using the
1410: traceless condition for the perturbation we get
1411: \be
1412: -\frac{\tilde\kappa^2}{6} T=\Box_4 \xi\ . \ee
1413: Using a pointlike
1414: particle with mass $M$ \be T_{00}=M\delta(\vec r),\ \ \ \
1415: T_{0i}=T_{ij}=0, \ee we obtain \be\label{xi}
1416: \xi=-\frac{\tilde\kappa^2}{24\pi}\frac{M}{r}. \ee Applying this to
1417: (\ref{pluba}) we get the boundary conditions \be \label{boco}
1418: \partial_z\gamma_{00}\Big|_\Sigma=-{2\tilde\kappa^2 M\over 3}\delta (\vec r)\quad,\quad
1419: \partial_z\gamma_{0i}\Big|_\Sigma=0\quad,\quad
1420: \partial_z\gamma_{ij}\Big|_\Sigma=-{\tilde\kappa^2 M\over 3}\delta (\vec r)\delta_{ij}-
1421: {\tilde\kappa^2 M\over
1422: 12\pi}\partial_{ij}{1\over r}\,.
1423: \ee
1424: Now the procedure is to solve the bulk equations and then impose the boundary conditions
1425: (\ref{boco}).
1426: The bulk equations are nothing else than the linearization
1427: of the Einstein equations
1428: \be
1429: \tilde G_{AB}=-\tilde \Lambda \tilde g_{AB},
1430: \ee
1431: that, using the metric ansatz (\ref{metan}), give \cite{5}
1432: \be
1433: \Box_4 \gamma_{\mu\nu}+\partial_z^2\gamma_{\mu\nu}-\frac{3}{z}\partial_z\gamma_{\mu\nu}=0.
1434: \ee
1435: Finally the general solution of these equations in the static case is a
1436: superposition of Fourier modes~:
1437: \be
1438: \gamma_{\mu\nu}(x^\mu,z)=\int\! {d^3\vec
1439: k\over(2\pi)^{{3\over2}}} e^{i \vec k\,.\vec r}\hat\gamma_{\mu\nu}(\vec k, z)
1440: \ee
1441: with
1442: \be
1443: \hat\gamma_{\mu\nu}(\vec k, z)=z^2
1444: \left[e_{\mu\nu}^{(1)}(\vec k)H_2^{(1)}(ikz)+e_{\mu\nu}^{(2)}(\vec k)H_2^{(2)}(ikz)\right]
1445: \ee
1446: where (because of the traceless conditions) the polarization tensors $e_{\mu\nu}^{(1,2)}
1447: (\vec k)$ are transverse and traceless,
1448: and where
1449: $H_2^{(1,2)}(ikw)$ are the Hankel functions of first and second kind of order $2$.
1450: The junction conditions (\ref{boco}) therefore determine a combination of the polarization
1451: tensors such that
1452: \begin{eqnarray}
1453: e_{00}^{(1)}(\vec k)H_1^{(1)}(ik{l})+e_{00}^{(2)}(\vec
1454: k)H_1^{(2)}(ik{l})=-{2\tilde\kappa^2 M\over 3}
1455: {1\over(2\pi)^{3\over2}}{1\over ik{l}^2}\ ,\cr e_{0i}^{(1)}(\vec
1456: k)H_1^{(1)}(ik{l})+e_{0i}^{(2)}(\vec k)H_1^{(2)}(ik{l})=0\ ,\cr e_{ij}^{(1)}(\vec k)H_1^{(1)}
1457: (ik{l})+e_{ij}^{(2)}(\vec k)H_1^{(2)}(ik{l})=-{\tilde\kappa^2
1458: M\over 3} {1\over(2\pi)^{3\over2}}{1\over ik{l}^2}\left(\delta_{ij}-{k_ik_j\over4\pi k^2}
1459: \right)\,.
1460: \end{eqnarray}
1461: We are now interested in the solution which converges in the limit
1462: $z\rightarrow\infty$. This is possible by choosing
1463: $e_{\mu\nu}^{(2)}=0$. Then we have the following complete solution
1464: \be \gamma_{\mu\nu}(\vec
1465: r,z)=\int\!{d^3k\over(2\pi)^{3\over2}}\,e^{i\vec k\,.\vec
1466: r}\hat\gamma_{\mu\nu}(\vec k, z)\quad
1467: ,\quad\hat\gamma_{\mu\nu}(\vec k,z)= {\tilde\kappa^2 M\over3{l}(2\pi)^{3\over2}} \,z^2\,
1468: {K_2(kz)\over k{l}\,K_1(k{l})}\,c_{\mu\nu}\ , \ee with $c_{00}=2$, $c_{0i}=0$ and
1469: $c_{ij}=\delta_{ij}-k_ik_j/4\pi k^2$, and where $K_\nu(z)$ is the
1470: modified Bessel function defined as
1471: $K_\nu(z)=i{\pi\over2}e^{i\nu{\pi\over2}}H_\nu^{(1)}(iz)$. Near
1472: the brane this metric reduces to, setting $\epsilon={z\over{l}}-1$,
1473: \be \label{gamma}
1474: \hat\gamma_{\mu\nu}(\vec k,\epsilon)=
1475: {\tilde\kappa^2 M{l}\over3(2\pi)^{3\over2}}
1476: \left\{{K_2(k{l})\over k{l}\,K_1(k{l})}-\epsilon-{k{l}\over4}\epsilon^2\left[{K_2(k{l})
1477: \over K_1(k{l})}-3{K_0(k{l})\over K_1(k{l})}\right]+{\cal O}(\epsilon^3)\right\}c_{\mu\nu}
1478: . \ee
1479: The
1480: appearance of Dirac distributions in the expansion of
1481: $\gamma_{\mu\nu}(\vec r,z)$ does not however necessarily mean that
1482: $\gamma_{\mu\nu}(\vec r,z)$ is singular at $\vec r=0$ as the sum
1483: may be regular. We now see what the perturbed metric
1484: on the brane is. We know that the brane is at $z={l}+\xi$.
1485: Therefore if we want to expand the metric around $z={l}$ we
1486: have
1487: \be ds^2\Big|_{\Sigma}=\left[\frac{l}{{l}+\xi}\left(\eta_{AB}+\gamma_{AB}\right)dx^A
1488: dx^B\right]\Big|_{\Sigma}\simeq\left(\eta_{AB}+h_{AB}\right) dx^A dx^B, \ee where \be
1489: h_{AB}=\gamma_{AB}\Big|_{\Sigma}-2\frac{\xi}{l}\eta_{AB}. \ee
1490: Therefore
1491: the actual perturbed four-dimensional metric is \be \label{actual}
1492: g_{\mu\nu}=\eta_{\mu\nu}+h_{\mu\nu}. \ee The Newtonian potential
1493: is then $h_{00}$. If we now Fourier transform $h_{00}$ considering
1494: (\ref{xi}) and (\ref{gamma}), we get
1495: \be \hat h_{00}(\vec
1496: k)=\hat{\tilde h}_{00}(\vec
1497: k)=\hat\gamma_{00}\Big|_\Sigma+2{\hat\xi\over{l}}=
1498: {\tilde\kappa^2 M\over k^2{l}(2\pi)^{3\over2}}\left[1-{2k{l}\over3}{K_0(k{l})\over
1499: K_1(k{l})}\right]\,. \ee
1500: Taking the inverse-Fourier transform
1501: and integrating over angles, we obtain, setting $\alpha=r/{l}$,
1502: \be h_{00}(\vec r)={\tilde\kappa^2 M\over4\pi{l}}{1\over
1503: r}\left(1+{4\pi\over3}{\cal K}_\alpha\right)\quad\hbox{with}\quad
1504: {\cal K}_\alpha=\lim_{\epsilon\to0}\int_0^{+\infty}\!
1505: du\,\sin(u\alpha)\,{K_0(u)\over K_1(u)}e^{-\epsilon u}\,. \ee
1506: We have a short and long distance limit,
1507: $\lim_{\alpha\to0}{\cal K}_\alpha=\alpha^{-1}={l}/r$,
1508: $\lim_{\alpha\to\infty}{\cal K}_\alpha=\pi/2\alpha^2=\pi ({l}/r)^2/2$. We hence recover
1509: that at short distances the
1510: correction to Newton's law is ${l}/r$, whereas at
1511: distances large compared with the AdS curvature scale ${l}$,
1512: the correction is reduced by another
1513: ${l}/r$ factor, in agreement with \cite{RS,bunchNew}
1514: \be\label{weakexp} \lim_{r/{l}\to\infty}h_{00}(\vec
1515: r)={2G_NM\over r}\left[1+{2\over3}\left({{l}\over
1516: r}\right)^2\right]\,, \ee where, as before, we identify the
1517: Newtonian constant as $8\pi G_N=\tilde\kappa^2/l$. This
1518: expansion can be taken as a ``quality-test'' that solutions of the
1519: four-dimensional effective field equations have to pass to lead to
1520: a regular bulk. Indeed, as we will discuss in the next chapter,
1521: stellar solutions which do not have the weak regime expansion
1522: (\ref{weakexp}), probably lead to a non-regular Cauchy horizon for
1523: the bulk \cite{static}.
1524:
1525: \subsection{Quantum correction to the Newtonian potential and holographic interpretation}
1526:
1527: \footnote{To show the importance of studying braneworld scenarios from the holographic
1528: point of view, I will
1529: occasionally insert sections on the quantum side of the correspondence (e.g. AdS/CFT
1530: and its deformations). The main idea of this thesis
1531: is to study the classical astrophysics and cosmology of braneworlds. Therefore these
1532: sections aim to give the reader only a
1533: flavour of the
1534: quantum counterpart of the correspondence.} Following
1535: \cite{DL} we show that the four-dimensional quantum gravity correction to the
1536: Newtonian law
1537: corresponds to the classical correction (\ref{weakexp}) found in the context of the
1538: braneworld scenario. This is a test of the
1539: deformed AdS/CFT
1540: correspondence at perturbative level.
1541:
1542: We start with the linearized Einstein equations in four dimensions, using the metric
1543: (\ref{actual}),
1544: where $h_{\mu\nu}$ is a perturbation of the Minkowskian metric, the harmonic gauge
1545: $\partial_\mu g^{\mu\nu}=0$ implies
1546: \be
1547: \partial_\mu(h^{\mu\nu}-{1\over 2} \eta^{\mu\nu}h)=0.
1548: \ee
1549: Defining $\tilde h_{\mu\nu}=h^{\mu\nu}-{1\over 2} \eta^{\mu\nu}h$
1550: we get the
1551: linearized Einstein equations
1552: \be
1553: \Box \tilde h_{\mu\nu}=-16\pi G_N T_{\mu\nu},
1554: \ee
1555: so that in Fourier space
1556: \be
1557: \tilde h_{\mu\nu}(p)=16\pi G_N {1\over p^2}T_{\mu\nu}(p).
1558: \ee
1559: For quantum corrections, the total perturbation will be of form
1560: \begin{equation}
1561: \tilde h_{\mu\nu}=h^c_{\mu\nu}+h^q_{\mu\nu}\ ,
1562: \end{equation}
1563: where the superscript $c$ means ``classical'' and the superscript $q$ means ``quantum".
1564: The quantum spin-two tensor $h^q_{\mu\nu}$
1565: must vanish with the vanishing of the
1566: classical perturbations. We have
1567: then two other
1568: ingredients, the graviton propagator and the self-energy of the gravitons.
1569: These two objects must be represented by tensors of rank 4. This is because they have
1570: to be applied to the classical perturbation of
1571: the metric and they have to reproduce a rank 2 tensor (the quantum perturbation of the metric).
1572:
1573: We start with the propagator, calling it
1574: $D^{\alpha\beta\gamma\delta}$. Taking two points $x$ and
1575: $x'$ at the same proper time, the propagator describes the short-distance quantum effects
1576: of the products $h_{\alpha\beta}(x)h_{\gamma\delta}(x')$. This
1577: implies that it should be proportional to $\delta(x-x')$. Another
1578: ingredient is that it must depend on the unperturbed metric. In
1579: particular since it must be local, it cannot depend on the derivatives
1580: of the background metric. So in general the propagator will depend
1581: only on the products $\eta^{\alpha\beta}\eta^{\gamma\delta}$.
1582: Bearing in mind that $D_{\mu\nu\alpha\beta}\sim
1583: h_{\mu\nu}h_{\alpha\beta}$, by symmetry we can already guess the
1584: form of this propagator as
1585: \be D^{\mu\nu\alpha\beta}={1 \over
1586: 2}\delta(x-x')\left(\eta^{\mu\alpha}\eta^{\nu\beta}+\eta^{\mu\alpha}\eta^{\nu\alpha}+
1587: \lambda\eta^{\mu\nu}\eta^{\alpha\beta}\right),
1588: \ee
1589: where $\lambda$ is a constant that must be determined.
1590:
1591: For the Hamiltonian constraint to be non-singular, the propagator in Fourier
1592: space becomes \cite{BdW}
1593: \be D^{\mu\nu\alpha\beta}(p^2)=\frac{1}{2
1594: p^2}\left(\eta^{\mu\alpha}\eta^{\mu\beta}+\eta^{\mu\alpha}\eta^{\nu\alpha}-
1595: \eta^{\mu\nu}\eta^{\alpha\beta}\right).
1596: \ee
1597: Up to now we have
1598: \be
1599: h_q^{\mu\nu}=D^{\mu\nu\alpha\beta}\Pi_{\alpha\beta\gamma\delta}h_c^{\gamma\delta},
1600: \ee
1601: where $\Pi_{\alpha\beta\gamma\delta}$ is the graviton self
1602: energy. Now in Fourier space the self energy is a function of the
1603: momentum $p_\alpha$ of the graviton and of the unperturbed metric
1604: (we still are at the first order in quantum corrections). In
1605: particular since the tensor is a rank four tensor and
1606: the metric is dimensionless, the self energy must have all the
1607: possible combinations of $p_\alpha$ products of order four. This is
1608: because in general we will have at least a term $p_\alpha p_\beta
1609: p_\gamma p_\delta$. If we would like to preserve the classical
1610: isometries at quantum level, we can also use the Slanov-Ward
1611: gravitational identity \cite{SWi}
1612: \be p_\mu p_\nu
1613: D^{\mu\nu\alpha\beta}\Pi_{\alpha\beta\gamma\delta}D^{\gamma\delta\rho\sigma}=0.
1614: \ee
1615: This at first linearized level determines the self-energy up to
1616: two functions $\Pi_1(p)$ and $\Pi_2(p)$. Combining the classical
1617: and the one-loop quantum results at the linearized level, we get
1618: \ba
1619: h_{\mu\nu}&=&16 \pi \frac{G_N}{p^2}\left[T_{\mu\nu}-{1\over
1620: 2}\eta_{\mu\nu}T(p)\right]\cr\nonumber &-&16\pi
1621: G_N\left[2\Pi_2(p)T_{\mu\nu}(p)+\Pi_1(p)\eta_{\mu\nu}T(p)\right].
1622: \ea
1623: The actual form of the $\Pi_i$'s depend on the theory we
1624: consider. However for any massless theory in four dimensions,
1625: after cancelling the infinities with the appropriate counterterms,
1626: the finite remainder must have the form \cite{CLR} \be\label{self}
1627: \Pi_i(p)=32\pi G_N\left(a_i\ln \frac{p^2}{\mu^2}+b_i\right), \ee
1628: where $\mu$ is an arbitrary subtraction mass linked with the
1629: renormalization energy. Using now a point source
1630: $T_{00}=M\delta(\vec r)$, we obtain the perturbed time component
1631: of the metric
1632: \be g_{00}=-(1-\frac{2 G_N M}{r}-\frac{2\alpha G_N^2
1633: M}{r^3}), \ee
1634: where $\alpha=4\cdot 32\pi(a_1+a_2)$. Explicit
1635: calculations of the self-energy (\ref{self}) for different spins
1636: give \cite{DL,245Duff}
1637: \be a_i(s=1)=4a_i(s=1/2)=12a_i
1638: (s=0)=\frac{1}{120(4\pi)^2}(-2,3). \ee
1639: Considering the number of
1640: particle species of spin $s$ going around the loop $N_s$ and
1641: considering that for a single CFT these numbers can be rewritten
1642: in terms of the dimensionality of the gauge group of the CFT ($N$)
1643: \be (N_1,N_{1/2},N_0)=(N^2,4N^2,6N^2), \ee
1644: we have that the
1645: one-loop correction to the Newtonian potential is
1646: \be
1647: V(r)=\frac{G_N M}{r}\left(1+\frac{2N^2 G_N}{3\pi r^2}\right). \ee
1648: Using the AdS/CFT relation \cite{M1} $N^2=\pi{l}^3/2\tilde\kappa^2$ we get exactly
1649: the result (\ref{weakexp}),
1650: which was found considering a five dimensional classical
1651: braneworld scenario.
1652:
1653: \newpage
1654: \chapter{Stars in the braneworld}
1655:
1656: The first step in considering infinitely large extra dimensions is
1657: the discovery that gravity can be localized at low energies \cite{RS}.
1658: This basically gives the strongest constraint for such theories.
1659: The second step is therefore to see how deviations from general relativity may
1660: explain the nature of our Universe.
1661:
1662: The natural scenario,
1663: where deviations from general relativity occur, is cosmology.
1664: Indeed cosmological implications of these braneworld models have been
1665: extensively investigated (see e.g. the review~\cite{maacov} for
1666: further references). But this is not all. Significant deviations from Einstein's theory
1667: in fact occur also in astrophysics. Indeed
1668: very compact objects and gravitational collapse to black holes, can leave
1669: traces of the extra dimensions.
1670: For example, when an horizon
1671: forms, even if the high-energy
1672: effects eventually become disconnected
1673: from the outside region on the brane, they could leave a
1674: signature on the brane \cite{collapse}.
1675:
1676: In addition to local high-energy effects,
1677: there are also nonlocal corrections arising from the imprint on
1678: the brane of Weyl curvature in the bulk, i.e. from 5-dimensional
1679: graviton stresses. These nonlocal Weyl stresses arise on the brane
1680: whenever there is inhomogeneity in the density; the inhomogeneity
1681: on the brane generates Weyl curvature in the bulk which
1682: `backreacts' on the brane. Note that we can have these nonlocal Weyl
1683: stresses even if the density is homogeneous \cite{static}.
1684:
1685: The high-energy (local) and bulk graviton stress (nonlocal)
1686: effects combine to significantly alter the matching problem on the
1687: brane, compared with the general relativistic case. For spherical
1688: compact or collapsing objects (uncharged and non-radiating), matching in general
1689: relativity shows that the asymptotically flat exterior spacetime
1690: is Schwarzschild. High-energy corrections to the pressure,
1691: together with Weyl stresses from bulk gravitons, mean that on the
1692: brane, matching no longer leads to a Schwarzschild exterior in
1693: general \cite{static, collapse}. These stresses also mean that the matching conditions do
1694: not have unique solution on the brane \cite{static}; knowledge of the
1695: 5-dimensional Weyl tensor is needed as a minimum condition for
1696: uniqueness.
1697:
1698: \section{The static star case}
1699:
1700: \footnote{I will base this section essentially on \cite{static}.}In
1701: this section we consider the simplest case of a static spherical
1702: star with uniform density. We find an exact interior solution,
1703: thus generalizing the Schwarzschild interior solution of general
1704: relativity. We show that the general relativity compactness limit
1705: given by $GM/R<{4\over9}$ is reduced by high-energy 5-dimensional
1706: gravity effects. The existence of neutron stars allows us to put a
1707: lower bound on the brane tension, which is stronger than the bound
1708: from big bang nucleosynthesis, but weaker than the bound from
1709: experiments probing Newton's law on sub-millimetre scales. We also
1710: give two different exact exterior solutions, both of which satisfy
1711: the braneworld matching conditions and field equations and are
1712: asymptotically Schwarzschild, but neither of which is the
1713: Schwarzschild exterior. One of these solutions is the
1714: Reissner-N\"ordstrom-type solution found in~\cite{dmpr}, in which
1715: there is no electric charge, but instead a Weyl `charge' arising
1716: from bulk graviton tidal effects. The other is a new solution.
1717: Both of these exterior solutions carry the imprint of bulk
1718: graviton stresses, and each has an horizon on the brane which is
1719: larger than the Schwarzschild horizon.
1720:
1721: Both of our solutions (i.e. the full solution, interior plus exterior) are consistent
1722: braneworld solutions, but we do not know the bulk solutions of
1723: which they are boundaries. In fact, no exact 5-dimensional
1724: solution for astrophysical brane black holes is known, and the
1725: uniform star case is even more complicated.
1726:
1727: We have seen in the section (\ref{secNewton}) that the Newtonian potential on the brane is
1728: modified as follow
1729: \begin{equation}\label{pert}
1730: \Phi={GM\over r}\left(1+{2l^2\over3r^2}\right)\,,
1731: \end{equation}
1732: where $l$ is the curvature scale of AdS. This result assumes that the bulk
1733: perturbations are bounded in conformally Minkowski coordinates,
1734: and that the bulk is nearly AdS. It is not clear whether there
1735: is a covariant way of uniquely characterizing these perturbative
1736: results~\cite{DD}, and therefore it remains unclear what the
1737: implications of the perturbative results are for very dense stars
1738: on the brane. However, it seems reasonable to conjecture that the
1739: bulk should be asymptotically AdS, and that its Cauchy horizon
1740: should be regular. Then perturbative results suggest that on the
1741: brane, the weak-field potential should behave as in
1742: eq.~(\ref{pert}). In fact, perturbative analysis also constrains
1743: the weak-field behaviour of other metric components on the
1744: brane~\cite{bunchNew}, as well as the nonlocal stresses on the
1745: brane induced by the bulk Weyl tensor~\cite{ssm}.
1746: This is also supported at non-linear level if one assumes a
1747: bounded bulk in conformally Minkowskian coordinates.
1748: In this case indeed, it is possible to integrate numerically simple models of
1749: homogeneous and isotropic stars \cite{wiseman}.
1750:
1751: \subsection{Field equations and matching conditions}\label{sectionField}
1752:
1753: We have already discussed that the local and nonlocal extra-dimensional modifications to
1754: Einstein's equations on the brane may be consolidated into an
1755: effective total energy-momentum tensor:
1756: \begin{equation}
1757: G_{\mu\nu}=\kappa^2 T^{\rm eff}_{\mu\nu}\,, \label{6'}
1758: \end{equation}
1759: where $\kappa^2=8\pi G_N$ and the bulk cosmological constant is
1760: chosen so that the brane cosmological constant vanishes. The
1761: effective total energy density, pressure, anisotropic stress and
1762: energy flux for a fluid are given by eqs. (\ref{reff}-\ref{e:pressure2}).
1763:
1764: From big bang nucleosynthesis constraints, $\lambda \gtrsim
1765: 1$~MeV$^4$, but a much stronger bound arises from null results of
1766: sub-millimetre tests of Newton's law\footnote{From the definition (\ref{GNdef}) and the fine-tuning (\ref{min}) we get $\lambda=3/4\pi (l^2 G_N)^{-1}$.
1767: Table-top tests of Newton's laws currently find no deviation down to about $0.2$ mm. This implies from (\ref{pert}) that $l\lesssim 0.2$ mm, or
1768: $\lambda \gtrsim
1769: 10$~TeV$^4$.}: $\lambda \gtrsim
1770: 10$~TeV$^4$.
1771:
1772: The local effects of the bulk, arising from the brane extrinsic
1773: curvature, are encoded in the quadratic terms, $\sim
1774: (T_{\mu\nu})^2/\lambda$, which are significant at high energies,
1775: $\rho\gtrsim\lambda$. The nonlocal bulk effects, arising from the
1776: bulk Weyl tensor, are carried by nonlocal energy density ${\cal
1777: U}$, nonlocal energy flux ${Q}_\mu$ and nonlocal anisotropic
1778: stress ${\Pi}_{\mu\nu}$. Five-dimensional graviton stresses are
1779: imprinted on the brane via these nonlocal Weyl terms.
1780:
1781: Static spherical symmetry implies ${Q}_\mu=0$ and
1782: \begin{equation}\label{p}
1783: {\Pi}_{\mu\nu}={\Pi}(r_\mu r_\nu-{\textstyle{1\over3}}
1784: h_{\mu\nu})\,,
1785: \end{equation}
1786: where $r_\mu$ is a unit radial vector. For static spherical symmetry, the
1787: conservation equations (\ref{pc1}-\ref{pc2'}) reduce to
1788: \begin{eqnarray}
1789: &&\D_\mu p+(\rho+p)A_\mu=0\,,\label{c1}\\ && {\textstyle{1\over3}}
1790: \D_\mu{\cal U}+ {\textstyle{4\over3}}{\cal U}A_\mu+\D^\nu {\Pi}_{\mu\nu}=
1791: -\frac{(\rho+p)}{\lambda} \D_\mu\rho\, .
1792: \label{c2}
1793: \end{eqnarray}
1794: In static coordinates the metric is
1795: \begin{eqnarray}\label{m}
1796: ds^2=-A^2(r)dt^2+B^2(r)dr^2+r^2d\Omega^2\,,
1797: \end{eqnarray}
1798: and eqs.~(\ref{6'})--(\ref{m}) imply
1799: \begin{eqnarray}
1800: &&{1\over r^2}-{1\over B^2}\left({1\over r^2}-{2\over r}{B'\over
1801: B}\right) = 8\pi G_N\rho^{\rm eff}\,,\label{f1}\\ &&-{1\over
1802: r^2}+{1\over B^2}\left({1\over r^2}+{2\over r}{A'\over A}\right) =
1803: 8\pi G_N\left(p^{\rm eff}+ {2\over 3}{\Pi}
1804: \right)\,,\label{f2}\\&&p'+{A'\over A}(\rho+p)=0\,,\label{f3}\\
1805: &&{\cal U}'+4{A'\over A} {\cal U}+ 2{\Pi}'+2{A'\over A}{\Pi}+
1806: {6\over r}{\Pi}=-3\frac{(\rho+p)}{\lambda}\rho'\,. \label{f4}
1807: \end{eqnarray}
1808:
1809: The exterior is characterized by
1810: \begin{equation}
1811: \rho=0=p\,,~~{\cal U}={\cal U}^+\,,~{\Pi}={\Pi}^+\,,
1812: \end{equation}
1813: so that in general $\rho^{\rm eff}$ and $p^{\rm eff}$ are nonzero
1814: in the exterior: there are in general Weyl stresses in the
1815: exterior, induced by bulk graviton effects. These stresses are
1816: radiative, since their energy-momentum tensor is traceless. The system of equations
1817: for the exterior is not closed until a further condition is given
1818: on ${\cal U}^+$, ${\Pi}^+$ (e.g., we could impose ${\Pi}^+=0$ to close the system).
1819: In other words, from a brane
1820: observer's perspective, there are many possible static spherical
1821: exterior metrics, including the simplest case of Schwarzschild
1822: (${\cal U}^+=0={\Pi}^+$).
1823:
1824: The interior has nonzero $\rho$ and $p$; in general, ${\cal U}^-$
1825: and ${\Pi}^-$ are also nonzero, since by eq.~(\ref{f4}), {\em
1826: density gradients are a source for Weyl stresses in the interior}.
1827: For a uniform density star, we can have ${\cal U}^-=0={\Pi}^-$,
1828: but nonzero ${\cal U}^-$ and/ or ${\Pi}^-$ are possible,
1829: subject to eq.~(\ref{f4}) with zero right-hand side.
1830:
1831: From eq.~(\ref{f1}) we obtain
1832: \begin{eqnarray}
1833: B^2(r)=\left[1-\frac{2Gm(r)}{r}\right]^{-1}\,,
1834: \end{eqnarray}
1835: where the mass function is
1836: \begin{eqnarray}
1837: m(r)=4\pi\int^r_a\rho^{\rm eff}(r')r'^2dr'\,,
1838: \end{eqnarray}
1839: and $a=0$ for the interior solution, while $a=R$ for the exterior
1840: solution.
1841:
1842: The Israel matching conditions at the stellar surface
1843: $\Sigma$ give~\cite{israel}
1844: \begin{eqnarray}
1845: [G_{\mu\nu}r^\nu]_{\Sigma}=0\,,
1846: \end{eqnarray}
1847: where $[f]_\Sigma\equiv f(R^+)-f(R^-)$. By the brane field
1848: equation~(\ref{6'}), this implies $[T^{\rm
1849: eff}_{\mu\nu}r^\nu]_\Sigma=0$, which leads to
1850: \begin{eqnarray} \label{m2}
1851: \left[p^{\rm eff}+ {2\over 3}{\Pi}\right]_\Sigma=0\,.
1852: \end{eqnarray}
1853: Even if the physical pressure vanishes at the surface, the
1854: effective pressure is nonzero there, so that in general a radial
1855: stress is needed in the exterior to balance this effective
1856: pressure.
1857:
1858: The general relativity limit of eq.~(\ref{m2}) implies
1859: \begin{equation}\label{p0}
1860: p(R)=0\,.
1861: \end{equation}
1862: This can also be obtained from a slightly different
1863: point of view. Consider the conservation equation (\ref{f3})
1864: \be
1865: p'+(\rho+p)\frac{A'}{A}=0.
1866: \ee
1867: The junction conditions (see Appendix \ref{Appjc}) require that on the stellar
1868: surface the metric and
1869: its first derivative along the orthogonal direction to the surface must be continuous and
1870: non-singular.
1871: This implies that $A'/A$ is a continuous and non-singular function across the boundary of
1872: the star. The model we
1873: are going to consider is such that
1874: \be
1875: \rho=\tilde\rho\ \theta(r-R)\ ,\ p=\tilde p\ \theta(r-R)\ ,
1876: \ee
1877: where the functions with tilde are continuous and non-singular functions and
1878: \ba
1879: \theta(r-R)=\begin{cases} 1\ \mbox{for}\ r<R\cr 0\ \mbox{for}\ r>R\end{cases}\ ,
1880: \ea
1881: is the Heaviside function. Now we consider the first derivative of the pressure
1882: \be
1883: p'=\tilde p'\ \theta(r-R)+\tilde p\ \delta(r-R)\ ,
1884: \ee
1885: where $\delta(r-R)$ is the Dirac distribution. In order to balance this distribution
1886: in eq. (\ref{f3}), we have two possibilities. The first one is that $p(R)=0$ and the
1887: second is that the metric becomes singular\footnote{Even if we relax the standard junction
1888: conditions we are left with the non trivial problem of defining the product of distributions
1889: $\delta(r-R)\theta(r-R)$.}. If we consider the junction conditions, requiring $A'/A$
1890: to be continuous and non-singular, the second choice becomes unavailable. Then the only
1891: physically sensible model is with $p(R)=0$. We have only used the
1892: conservation equations (\ref{f3}) and the geometrical junction conditions, valid for
1893: both general relativity and braneworld effective gravity.
1894: This implies that unlike general relativity, where (\ref{p0}) and (\ref{m2})
1895: correspond to the same
1896: constraint, they become separate conditions in the braneworld.
1897:
1898: Using $p(R)=0$, the constraint (\ref{m2}) becomes
1899: \begin{equation}\label{m3}
1900: 3\frac{\rho^2(R)}{\lambda}+{\cal U}^-(R) +2 {\Pi}^-(R)= {\cal U}^+(R)
1901: +2 {\Pi}^+(R)\,.
1902: \end{equation}
1903: Equation~(\ref{m3}) gives the matching condition for any static
1904: spherical star with vanishing pressure at the surface. If there
1905: are no Weyl stresses in the interior, i.e. ${\cal U}^-=0= {\Pi}^-$,
1906: and if the energy density is non-vanishing at the surface,
1907: $\rho(R)\neq0$, then there must be Weyl stresses in the exterior,
1908: i.e. the exterior cannot be Schwarzschild. Equivalently, {\em if
1909: the exterior is Schwarzschild and the energy density is nonzero at
1910: the surface, then the interior must have nonlocal Weyl stresses}.
1911:
1912:
1913: \subsection {Braneworld generalization of exact uniform-density solution}
1914:
1915: Here we are interested in the most simple model of a compact star (such as
1916: for example a neutron star \cite{gravitation}). In this model the density is considered
1917: constant inside the star and zero outside and the geometry is isotropic.
1918: This of course implies that $\rho'=0$ everywhere. Moreover, consistently with the
1919: isotropy of the star, we set ${\Pi}^-=0$.
1920:
1921: Equation~(\ref{f3}) integrates in the interior for
1922: $\rho=$\,const to give
1923: \begin{equation}
1924: A^-(r)={\alpha\over \rho+p(r)}\,,
1925: \end{equation}
1926: where $\alpha$ is a constant.
1927: Eq. (\ref{f4}) reads
1928: \be
1929: {\cal U}'+4\frac{A'}{A}{\cal U}=0\ ,
1930: \ee
1931: with solution
1932: \begin{equation}\label{u}
1933: {\cal U}^-(r)={\beta\over \left[A^-(r)\right]^4}\,,
1934: \end{equation}
1935: where $\beta$ is a constant. The matching condition in
1936: eq.~(\ref{m3}) then reduces for a {\em uniform} star to
1937: \begin{equation}\label{m4}
1938: 3\frac{\rho^2}{\lambda}+{\beta\over\alpha^4}\rho^4= {\cal U}^+(R) +2
1939: {\Pi}^+(R)\,.
1940: \end{equation}
1941: It follows that in general the exterior of a uniform star cannot be
1942: Schwarzschild.
1943:
1944: Combining eqs. (\ref{f2},\ref{f3}) we obtain the generalization of the Tolman-Oppenheimer-
1945: Volkoff
1946: equation of hydrostatic equilibrium for the braneworld
1947: \be \label{hydro}
1948: \frac{dp}{dr}=-(p+\rho)\frac{G_Nm(r)+4\pi r^3 p^{\rm eff}}{r(r-2G_Nm(r))},
1949: \ee
1950: where the interior mass function is
1951: \begin{eqnarray}\label{mass}
1952: m^-(r)=M\left[1+\frac{3M}{8\pi\lambda R^3}\right]\left({r\over R}
1953: \right)^3+4\pi\frac{\beta}{\alpha^4}\int^r_0 r'{}^2{(p(r')+\rho)^4} dr'\,,
1954: \end{eqnarray}
1955: with $M=4\pi R^3\rho/3$, and the effective pressure is
1956: \be
1957: p^{\rm eff}=p+\frac{\rho}{2\lambda}(2p+\rho)+\frac{\beta}{3\alpha^4}(p+\rho)^4.
1958: \ee
1959: Eq. (\ref{hydro}) can be analytically solved if and only if
1960: $\beta=0$. Indeed in this case it is possible to separate the variables $p$ and $r$.
1961:
1962: Here we are interested in typical astrophysical stars. Their density
1963: $\rho\sim 10^{-3}{}\mbox{GeV}^4$ is much smaller then the tension of the brane
1964: $\lambda\sim 10\mbox{~TeV}^4$. Non-local corrections are produced
1965: by back-reaction of the five-dimensional gravitational field on the brane. Therefore we
1966: naively expect
1967: that their order of magnitude is much smaller than the local effects.
1968: This is also supported by numerical models \cite{wiseman}.
1969: Since we expect ${\cal U}^-\ll \rho^2/\lambda$, we assume in the following $\beta=0$
1970: and
1971: then we can analytically solve (\ref{hydro}).
1972:
1973: With uniform density and ${\cal U}^-=0= {\Pi}^-$, we have the
1974: case of purely local (high-energy) modifications to the general
1975: relativity uniform-density solution, i.e. to the Schwarzschild
1976: interior solution~\cite{exact}.
1977:
1978: We can now calculate the pressure. Considering that
1979: \begin{eqnarray}
1980: B^-(r)={1\over\Delta(r)}\,,
1981: \end{eqnarray}
1982: the pressure is given by
1983: \begin{eqnarray} \label{3}
1984: {p(r)\over\rho}=\frac{[{\Delta(r)}-{\Delta(R)}] (1+\rho/\lambda)}
1985: {[3{\Delta(R)}-{\Delta(r)}]+[3{\Delta(R)}-2{\Delta(r)}]\rho/
1986: \lambda}\,,
1987: \end{eqnarray}
1988: where
1989: \begin{eqnarray}
1990: \Delta(r)=\left[1-{2G_NM\over r}\left({r\over R}\right)^3\left\{ 1+
1991: {\rho\over 2\lambda }\right\}\right]^{1/2}\,.
1992: \end{eqnarray}
1993: In the general relativity limit, $\lambda^{-1}\to 0$, we regain
1994: the known exact solution~\cite{exact}. The high-energy corrections
1995: considerably complicate the exact solution.
1996:
1997: Since $\Delta(R)$ must be real, we find {\em an astrophysical
1998: lower limit on $\lambda$, independent of the Newton-law and
1999: cosmological limits:}
2000: \begin{equation}\label{al}
2001: {\lambda}\geq \left(\frac{G_NM}{R-2G_NM}\right)\rho~~\mbox{for all
2002: uniform stars}\,.
2003: \end{equation}
2004: In particular, since $\lambda,\rho>0$, $\Delta(R)^2>0$ implies $R>2GM$, so that the Schwarzschild
2005: radius is still a limiting radius, as in general relativity.
2006: Taking a typical neutron star (assuming uniform density) with
2007: $\rho\sim 10^{9}$~MeV$^4$ and $M\sim 4\times 10^{57}$~GeV, we find
2008: \begin{eqnarray}\label{al'}
2009: \lambda > 5\times 10^{8}~\mbox{MeV}^4\,.
2010: \end{eqnarray}
2011: This is the astrophysical limit, below which stable neutron stars
2012: could not exist on the brane. It is much stronger than the
2013: cosmological nucleosynthesis constraint, but much weaker than the
2014: Newton-law lower bound. Thus stable neutron stars are easily
2015: compatible with braneworld high-energy corrections, and the
2016: deviations from general relativity are very small. If we used the
2017: lower bound in eq.~(\ref{al'}) allowed by the stellar limit, then
2018: the corrections to general relativistic stellar models would be
2019: significant, as illustrated in fig. (\ref{g1}).
2020:
2021: \begin{figure}
2022: \begin{center}
2023: \includegraphics[width=6cm,height=7cm,angle=-90]{g1.ps}
2024: \end{center}
2025: \caption{ \label{g1} Qualitative comparison of the pressure $p(r)$,
2026: in general relativity (upper curve), and in a braneworld model
2027: with $\lambda= 5\times 10^{8}$~MeV$^4$ (lower curve).}
2028: \end{figure}
2029:
2030:
2031: We can also obtain an upper limit on compactness from the
2032: requirement that $p(r)$ must be finite. Since $p(r)$ is a
2033: decreasing function, this is equivalent to the condition that
2034: $p(0)$ is finite and positive, which gives the condition
2035: \begin{eqnarray}\label{comp}
2036: \frac{G_NM}{R} \leq {4\over 9}\left[{1+7\rho/4\lambda +
2037: 5\rho^2/8\lambda^2
2038: \over(1+\rho/\lambda)^2(1+\rho/2\lambda)}\right]\,.
2039: \end{eqnarray}
2040: It follows that high-energy braneworld corrections {\em reduce}
2041: the compactness limit of the star. For the stellar bound on
2042: $\lambda$ given by eq.~(\ref{al'}), the reduction would be
2043: significant, but for the Newton-law bound, the correction to the
2044: general relativity limit of $4\over9$ is very small. The lowest
2045: order correction is given by
2046: \begin{equation}\label{comp2}
2047: {G_NM\over R}\leq \frac{4}{9}\left[1-\frac{3\rho}{4\lambda}
2048: +O\left({\rho^2\over\lambda^2}\right)\right]\,.
2049: \end{equation}
2050: For $\lambda\sim10$~TeV$^4$, the minimum allowed by
2051: sub-millimetre experiments, and $\rho\sim 10^{9}$~MeV$^4$, the
2052: fractional correction is $\sim 10^{-16}$.
2053:
2054:
2055: \subsection{Two possible non-Schwarzschild exterior solutions}
2056:
2057:
2058: The system of equations satisfied by the exterior spacetime on the
2059: brane is not closed. Essentially, we have two independent unknowns
2060: ${\cal U}^+$ and ${\Pi}^+$ satisfying one equation, i.e.
2061: eq.~(\ref{f4}) with zero right-hand side. Even requiring that the
2062: exterior must be asymptotically Schwarzschild does not lead to a
2063: unique solution. Further investigation of the 5-dimensional
2064: solution is needed in order to determine what the further
2065: constraints are. We are able to find two exterior solutions for a
2066: uniform-density star (with
2067: ${\cal U}^-=0$) that are consistent with all equations
2068: and matching conditions on the brane, and that are asymptotically
2069: Schwarzschild.
2070:
2071: The first is the Reissner-N\"ordstrom-like solution given
2072: in~\cite{dmpr}, in which a tidal Weyl charge plays a role similar
2073: to that of electric charge in the general relativity
2074: Reissner-N\"ordstrom solution. We stress that there is {\em no}
2075: electric charge in this model: nonlocal Weyl effects from the 5th
2076: dimension lead to an energy-momentum tensor on the brane that has
2077: the same form as that for an electric field, but without any
2078: electric field being present. The formal similarity is not
2079: complete, since the tidal Weyl charge gives a {\em positive}
2080: contribution to the gravitational potential, unlike the negative
2081: contribution of an electric charge in the general relativistic
2082: Reissner-N\"ordstrom solution.
2083:
2084: The braneworld solution is~\cite{dmpr}
2085: \begin{eqnarray}
2086: &&\left(A^+\right)^2=\left(B^+\right)^{-2}=1-\frac{2G_N{\cal
2087: M}}{r}+\frac{q}{r^2}\,,\label{rn1} \\&& {\cal U}^+=-\frac{{\Pi}^+}{2} = \frac{4}{3}\pi G_N
2088: q\lambda \,{1\over r^4}\,,\label{rn2}
2089: \end{eqnarray}
2090: where the matching conditions imply
2091: \begin{eqnarray}
2092: q&=&-3G_NMR\,{\rho\over\lambda}\,,\\ {\cal M}&=& M \left(
2093: 1-\frac{\rho}{\lambda }\right)\,, \\ \alpha&=& \rho\Delta (R)\,.
2094: \end{eqnarray}
2095: Note that the Weyl energy density in the exterior is {\em
2096: negative}, so that 5-dimensional graviton effects lead to a
2097: strengthening of the gravitational potential (this is discussed
2098: further in~\cite{dmpr,ssm}). Since ${\cal M}>0$ is required for
2099: asymptotic Schwarzschild behaviour, we have a slightly stronger
2100: condition on the brane tension:
2101: \begin{equation}
2102: \lambda>\rho \,.
2103: \end{equation}
2104: However, this still gives a weak lower limit, $\lambda>
2105: 10^{9}~\mbox{MeV}^4$. In this solution the horizon is at
2106: \begin{equation}
2107: r_{\rm h}=G_N{\cal M}\left[1+\left\{1+\left({3R\over
2108: 2G_NM}-2\right){\rho\over \lambda} +{\rho^2\over
2109: \lambda^2}\right\}^{1/2}\right].
2110: \end{equation}
2111: Expanding this exact expression shows that the horizon is slightly
2112: beyond the general relativistic Schwarzschild horizon:
2113: \begin{equation}
2114: r_{\rm h}=2G_NM\left[1+{3(R-2G_NM)\over
2115: 4G_NM}\,{\rho\over\lambda}\right] +O\left({\rho^2\over
2116: \lambda^2}\right)
2117: >2GM\,.
2118: \end{equation}
2119: The exterior curvature invariant ${\cal R}^2=
2120: R_{\mu\nu}R^{\mu\nu}$ is given by
2121: \begin{eqnarray}\label{r2}
2122: {\cal R}= 8\pi G_N\left({\rho\over \lambda}\right)^2\left({R\over
2123: r}\right)^4\,.
2124: \end{eqnarray}
2125: Note that for the Schwarzschild exterior, ${\cal R}=0$.
2126:
2127: The second exterior is a new solution. Like the above solution, it
2128: satisfies the braneworld field equations in the exterior, and the
2129: matching conditions at the surface of the uniform-density star. It
2130: is given by
2131: \begin{eqnarray}
2132: \left(A^+\right)^2&=&1-\frac{2G_N{\cal N}}{r}\,, \label{n1} \\
2133: \left(B^+\right)^{-2}&=&\left(A^+\right)^2
2134: \left[1+\frac{C}{\lambda(r-{\textstyle{3\over2}} G_N{\cal
2135: N})}\right]\,,\label{n2} \\ {\cal U}^+&=& {2\pi G_N^2{\cal
2136: N}C\over(1-3G_N{\cal N}/2r)^2}\,{1\over r^4}\,,\label{n3}\\ {\Pi}^+&=& \left({2\over3}-
2137: {r \over G_N{\cal N}}\right){\cal U}^+\,.
2138: \label{n4}
2139: \end{eqnarray}
2140: From the matching conditions:
2141: \begin{eqnarray}
2142: {\cal N} &=& M\left[\frac{1+2\rho/\lambda}{1+3G_NM\rho/
2143: R\lambda)}\right]\,,\\ C &=& 3G_NM\rho\left[ \frac{1-3G_NM/2R}{ 1
2144: +3G_NM\rho/R\lambda}\right]\,,\\ \alpha &=&{\rho\Delta(R)\over (1+3
2145: G_NM\rho/R\lambda)^{1/2}}\,.
2146: \end{eqnarray}
2147: The horizon in this new solution is at
2148: \begin{equation}
2149: r_{\rm h}=2G_N{\cal N}\,,
2150: \end{equation}
2151: which leads to
2152: \begin{equation}
2153: r_{\rm h}=2G_NM\left[1+\left({2R-3G_NM\over 2R}\right){\rho\over
2154: \lambda}\right]+O\left({\rho^2\over\lambda^2}\right)>2G_NM\,.
2155: \end{equation}
2156: The curvature invariant is
2157: \begin{eqnarray}
2158: {\cal R}&=&\sqrt{{\textstyle{3\over2}}}RC\left({4\pi R\over
2159: 3M}\right)^2{(1-8G_N{\cal N}/3r+2G_N^2{\cal N}^2 /r^2)^{1/2}\over
2160: 1-3G_N{\cal N}/2r}\times\nonumber\\ &&~~{}\times \left({\rho\over
2161: \lambda}\right)^2\left({R\over r}\right)^3 \,.
2162: \end{eqnarray}
2163: Comparing with eq.~(\ref{r2}), it is clear that these two
2164: solutions are different. The difference in their curvature
2165: invariants is illustrated in fig. (\ref{g3}).
2166:
2167: \begin{figure}
2168: \begin{center}
2169: \includegraphics[width=6cm,height=7cm,angle=-90]{g3.ps}
2170: \end{center}
2171: \caption{ \label{g3} Qualitative behaviour of the curvature
2172: invariant ${\cal R}^2$: the upper curve is the
2173: Reissner-N\"ordstrom-like solution given by eqs.~(\ref{rn1}) and
2174: (\ref{rn2}); the lower curve is the new solution given by
2175: eqs.~(\ref{n1})--(\ref{n4}) ($\lambda= 5\times 10^{8}$~MeV$^4$).}
2176: \end{figure}
2177:
2178:
2179: \subsection{Interior solution with Weyl contribution}
2180:
2181: We now consider the case where the contribution of the projected Weyl tensor
2182: is important in the interior or when $\beta\neq 0$.
2183:
2184: Following \cite{m} we can calculate the tidal acceleration on the brane measured by
2185: a co-moving observer with four velocity $u^A$.
2186: Its modulus is
2187: \ba
2188: \tilde A=-\lim_{y\rightarrow 0}n_A\tilde R^A_{BCD}u^B n^C u^D\ .
2189: \ea
2190: Using the decomposition of the Riemann tensor (\ref{decRie}) and recalling that
2191: $T_{AB}n^A=0$, we have
2192: \ba
2193: \tilde A=\kappa^2{\cal U}+\frac{\tilde \Lambda}{6}.
2194: \ea
2195: Now if ${\cal U}<0$ the localization of the gravitational field near the brane is enhanced
2196: reinforcing the Newtonian potential.
2197: In this case therefore a negative Weyl energy contributes to binding the star to
2198: the brane. Since this effect is independent of the brane tension, it makes
2199: the star more stable than
2200: in the general relativistic case. We see it in a very special limiting case when
2201: the exterior is Schwarzschild. In this case
2202: (${\cal U}^+=0={\Pi}^+$), eqs.~(\ref{u}) and (\ref{m4})
2203: imply that the interior must have dark radiation
2204: density:
2205: \begin{eqnarray}
2206: {\cal U}^-(r)=-\left(\frac{4\pi G_N}{\rho}\right)^2[\rho+p(r)]^4\,.
2207: \end{eqnarray}
2208: It follows that the mass function in eq.~(\ref{mass}) becomes
2209: \begin{eqnarray}
2210: m^-(r)&=& M\left(1+\frac{\rho}{2\lambda
2211: }\right) \left({r\over R} \right)^3
2212: -\frac{6\pi}{\lambda\rho^2}\int^r_0 [\rho+p(r')]^4 r'^2
2213: dr'\,,
2214: \end{eqnarray}
2215: which is {\em reduced} by the negative Weyl energy density,
2216: relative to the solution in the previous section and to the
2217: general relativity solution. The effective pressure is given by
2218: \begin{equation}\label{pschw}
2219: p^{\rm eff}=p-{\rho\over2\lambda}(2+6w+4w^2+w^3)\,,
2220: \end{equation}
2221: where $w=p/\rho$. Thus $p^{\rm eff}<p$, so that 5-dimensional
2222: high-energy effects reduce the pressure in comparison with general
2223: relativity. This means, as we already anticipated, that this star is more
2224: stable than the general relativistic one.
2225:
2226:
2227: \subsection{Unique exterior solution: a conjecture}
2228:
2229: We found that the Schwarzschild solution is no longer the unique asymptotically
2230: flat vacuum exterior; in general, the exterior carries an imprint
2231: of nonlocal bulk graviton stresses. The exterior is not uniquely
2232: determined by matching conditions on the brane, since the
2233: 5-dimensional metric is involved via the nonlocal Weyl stresses.
2234: We demonstrated this explicitly by giving two exact exterior
2235: solutions, both asymptotically Schwarzschild. Each exterior which
2236: satisfies the matching conditions leads to different bulk metrics.
2237: Without any exact or
2238: approximate 5-dimensional solutions to guide us, we do not know
2239: how the properties of the bulk metric, and in particular its
2240: global properties, will influence the exterior solution on the
2241: brane.
2242:
2243: Guided by perturbative analysis of the static weak field
2244: limit~\cite{RS,bunchNew,ssm,DD}, we make the following conjecture:
2245: {\em if the bulk for a static stellar solution on the brane is
2246: asymptotically AdS and has regular Cauchy horizon, then the
2247: exterior vacuum which satisfies the matching conditions on the
2248: brane is uniquely determined, and agrees with the perturbative
2249: weak-field results at lowest order.} An immediate implication of
2250: this conjecture is that the exterior is not Schwarzschild, since
2251: perturbative analysis shows that there are nonzero Weyl stresses
2252: in the exterior~\cite{ssm} (these stresses are the manifestation
2253: on the brane of the massive Kaluza-Klein bulk graviton modes). In
2254: addition, the two exterior solutions that we present would be
2255: ruled out by the conjecture, since both of them violate the
2256: perturbative result for the weak-field potential,
2257: eq.~(\ref{pert}).
2258:
2259: The static problem is already complicated, so that analysis of
2260: dynamical collapse on the brane will be very difficult. However,
2261: the dynamical problem could give rise to more striking features.
2262: Energy densities well above the brane tension could be reached
2263: before horizon formation, so that high-energy corrections could be
2264: significant. We expect that these corrections, together with the
2265: nonlocal bulk graviton stress effects, will leave a non-static,
2266: but transient, signature in the exterior of collapsing matter.
2267: This what the next section considers.
2268:
2269: \section{Gravitational collapse}
2270:
2271: \footnote{This section is based on \cite{collapse}.}The study of gravitational collapse
2272: in general relativity is
2273: fundamental to understanding the behaviour of the theory at high
2274: energies. The Oppenheimer-Snyder model still provides a
2275: paradigmatic example that serves as a good qualitative guide to
2276: the general collapse problem in general relativity. It can be solved analytically,
2277: as it simply assumes a collapsing homogeneous dust cloud of finite
2278: mass and radius, described by a Robertson-Walker metric and
2279: surrounded by a vacuum exterior. In general relativity, this exterior is
2280: necessarily static and given by the Schwarzschild
2281: solution~\cite{s}. In other theories of gravity that differ from
2282: general relativity at high energies, it is natural to look for similar examples.
2283: In this section we analyze an Oppenheimer-Snyder-like collapse in the
2284: braneworld scenario, in order to shed light on some fundamental differences
2285: between collapse in general relativity and on the brane.
2286:
2287: Braneworld gravitational collapse is complicated by a number of
2288: factors. The confinement of matter to the brane, while the
2289: gravitational field can access the extra dimension, is at the root
2290: of the difficulties relative to Einstein's theory, and this is
2291: compounded by the gravitational interaction between the brane and
2292: the bulk. Matching conditions on the brane are more complicated to
2293: implement, and one also has to impose regularity and
2294: asymptotic conditions on the bulk, and it is not obvious what
2295: these should be.
2296:
2297: In general relativity, the Oppenheimer-Snyder model of collapsing dust has a Robertson-Walker
2298: interior matched to a Schwarzschild exterior. We show that even
2299: this simplest case is much more complicated on the brane. However,
2300: it does have a striking new property, which may be part of the
2301: generic collapse problem on the brane. The exterior is not
2302: Schwarzschild, and nor could we expect it to be, as discussed
2303: in the previous section, but the exterior is not even {\em static}, as shown by our
2304: no-go theorem. The reason for this lies in the nature of the
2305: braneworld modifications to general relativity.
2306:
2307: The dynamical equations for the projected Weyl tensor (\ref{non-local})
2308: are the complete set of
2309: equations on the brane. They are not closed, since ${\cal
2310: E}_{\mu\nu}$ contains 5D degrees of freedom that cannot be
2311: determined on the brane. However, using only the 4D
2312: projected equations, we prove a no-go theorem valid for the full
2313: 5D problem: {\em given the standard matching conditions on the
2314: brane, the exterior of a collapsing homogeneous dust cloud cannot be static}.
2315: We are not able to determine the non-static exterior metric, but
2316: we expect on general physical grounds that the non-static
2317: behaviour will be transient, so that the exterior tends to a
2318: static form.
2319:
2320: The collapsing region in general contains dust and also energy
2321: density on the brane from KK stresses in the bulk (``dark radiation"). We show that in
2322: the extreme case where there is no matter but only collapsing
2323: homogeneous KK energy density, there is a unique exterior which is
2324: static for physically reasonable values of the parameters. Since
2325: there is no matter on the brane to generate KK stresses, the KK
2326: energy density on the brane must arise from bulk Weyl curvature.
2327: In this case, the bulk could be pathological. The collapsing KK
2328: energy density can either bounce or form a black hole with a 5D
2329: gravitational potential, and the exterior is of the Weyl-charged
2330: de\,Sitter type \cite{dmpr}, but with
2331: no mass.
2332:
2333: \subsection{Gravitational collapse: a no-go theorem}\label{star}
2334:
2335: A spherically symmetric collapse region has a Robertson-Walker metric
2336: \begin{equation}\label{1}
2337: ds^2=-d\tau^2+a(\tau)^2(1+ {\textstyle {1\over4}}
2338: kr^2)^{-2}\left[dr^2+r^2d\Omega^2\right]\, ,
2339: \end{equation}
2340: where $\tau$ is the proper time of the perfect fluid. This implies that the
2341: four velocity is $u^\alpha=\delta^\alpha_\tau$, so that
2342: \be
2343: T^\alpha_\beta=\mbox{diag}\left[-\rho(\tau),p(\tau),p(\tau),p(\tau)\right]\ .
2344: \ee
2345: Moreover the symmetries of the geometry force the four-dimensional Einstein tensor
2346: to have the following properties
2347: \ba
2348: G^r_r=G^\theta_\theta=G^\phi_\phi\ ,
2349: \ea
2350: and to have the off-diagonal terms identically zero. This implies
2351: \ba
2352: Q_\mu=0=\Pi_{\mu\nu}\ .
2353: \ea
2354: Whether or not this model has a regular bulk solution is not possible
2355: to determine here. However if the metric (\ref{1}) applies throughout the brane (i.e. there is
2356: no collapse and no exterior region) we can fully integrate the five dimensional equations
2357: \cite{bdl} obtaining a regular bulk. We will see this in the last chapter dedicated to cosmology.
2358:
2359: In this case the only non-local non-zero equation left is
2360: \be
2361: \dot{\cal U}+\frac{4}{3}\Theta{\cal U}=0\ ,
2362: \ee
2363: where $\Theta=3\ \dot a/a$. Then
2364: \be
2365: {\cal U}=\frac{C}{\lambda a^4}\ ,
2366: \ee
2367: where $C$ is an integration constant.
2368:
2369: Then from the $G^0_0$ components of (\ref{eq:effective}) we get the modified Friedmann equation
2370: \begin{eqnarray} \label{evol}
2371: \frac{\dot{a}^2}{a^2}= {\textstyle{8\over3}}\pi G_N\rho\left(1 +
2372: \frac{\rho}{2\lambda}\right) + \frac{C}{\lambda a^4} -\frac{k}{a^2}+
2373: {\textstyle{1\over3}} {\Lambda}\, .
2374: \end{eqnarray}
2375: The $\rho^2$ term,
2376: which is significant for $\rho\gtrsim \lambda$, is the high-energy
2377: correction term, following from $S_{\mu\nu}$. Standard Friedmann
2378: evolution is regained in the limit $\lambda^{-1} \to 0$.
2379:
2380: At this point we would like to use this geometry as the interior of a collapsing region and try to
2381: match it with a static exterior. Since we are in comoving coordinates, the boundary of the
2382: collapsing region is described by the implicit function
2383: \be
2384: \Phi=r-r_0=0\ ,
2385: \ee
2386: so that its normal unit vector is $n_\alpha=a(1+kr^2/4)^{-1}\delta_\alpha^r$.
2387: From the junction conditions (\ref{Appjc2})
2388: \ba\label{jc}
2389: \left[G^r_r\right]=\left[p^{\rm eff}\right]=0\ ,
2390: \ea
2391: where again $\left[f\right]:=f(r_0^+)-f(r_0^-)$, and $r_0^\pm$ is respectively the limit
2392: from the exterior and the interior to the boundary. Using the same discussion of par.
2393: (\ref{sectionField}) we can argue also that
2394: \be
2395: \left[p\right]=0\ ,
2396: \ee
2397: so that for a vacuum exterior $p^\pm=0$. Therefore the collapsing region must be dust.
2398:
2399: In this case in the interior
2400: \be
2401: p^{\rm eff}=\frac{\rho^2}{2\lambda}+\frac{\cal U}{3}\ ,
2402: \ee
2403: implying that the exterior can not be static.
2404:
2405: In the following we find a ``measure" of the non-static behaviour of the exterior solution.
2406: This is encoded in a Ricci anomaly. As we will see in the final section this anomaly can be
2407: interpreted holographically as the Weyl anomaly due to the Hawking evaporation for a collapsing
2408: body.
2409:
2410: The conservation equations (\ref{pc1}) for a dust cloud gives
2411: \ba
2412: \rho=\rho_0\left(\frac{a_0}{a}\right)^3 ,
2413: \ea
2414: where $a_0$ is
2415: the epoch when the cloud started to collapse. The proper radius
2416: from the centre of the cloud is $R(\tau)=r a(\tau)/(1+ {\textstyle
2417: {1\over4}} kr^2)$. We call the collapsing boundary surface $\Sigma$,
2418: which has as a proper radius $R_\Sigma(\tau)= r_0 a(\tau)/(1+
2419: {\textstyle {1\over4}} kr_0^2)$.
2420:
2421: We can rewrite the modified Friedmann equation on the interior
2422: side of $\Sigma$ as
2423: \begin{equation}\label{geo1}
2424: \dot{R}^2= \frac{2G_NM}{R}+ 3\frac{G_NM^2}{4\pi\lambda R^4}+ \frac{Q}{\lambda
2425: R^2}+ E+ \frac{\Lambda}{3} R^2\,,
2426: \end{equation}
2427: where the ``physical mass" $M$ (total energy per proper star
2428: volume) and the total ``tidal charge" $Q$ are
2429: \begin{equation}\label{mq}
2430: M={\textstyle{4\over3}}\pi a_0^3 r_0^3\frac{\rho_0}{(1+ {\textstyle
2431: {1\over4}}kr_0^2)^3},\ \ Q=C\frac{r_0^4}{(1+ {\textstyle
2432: {1\over4}}kr_0^2)^4},
2433: \end{equation}
2434: and the ``energy" per unit mass is given by
2435: \begin{equation}\label{en}
2436: E=-\frac{kr_0^2}{(1+ {\textstyle {1\over4}}kr_0^2)^2}>-1 \,.
2437: \end{equation}
2438:
2439: Now we assume that the exterior is static, and satisfies the
2440: standard 4D junction conditions. Then we check whether this
2441: exterior is physical consistent by imposing the modified Einstein
2442: equations (\ref{eq:effective}) for vacuum, i.e.\ for $T_{\mu\nu} =0 =
2443: S_{\mu\nu}$. The standard 4D Israel matching conditions, which we
2444: assume hold on the brane, require that the metric and the
2445: extrinsic curvature of $\Sigma$ be continuous. The extrinsic
2446: curvature is continuous if the metric is continuous and if $\dot
2447: R$ is continuous~\cite{s}. We therefore need to match the metrics
2448: and $\dot R$ across $\Sigma$.
2449:
2450: The most general static spherical metric that could match the
2451: interior metric on $\Sigma$ is
2452: \begin{eqnarray}
2453: ds^2= -F(R)^2A(R)dt^2 +\frac{dR^2}{A(R)} +R^2d\Omega^2,\label{s1}
2454: \end{eqnarray}
2455: where
2456: \be \label{A}
2457: A(R) =1-2G_Nm(R)/R.
2458: \ee
2459:
2460: We need two conditions to determine the functions $F(R)$ and
2461: $m(R)$. Now $\Sigma$ is a freely falling surface in both metrics.
2462: Therefore the first condition comes from the comparison of the geodesic
2463: equations of the two metrics. The geodesic equation for a radial trajectory seen
2464: from the exterior is calculated as follows. For a time-like geodesic we have
2465: \be\label{geodesic}
2466: \frac{ds^2}{d\tau^2}=-1=-F(R)^2A(R)\dot t^2+\frac{\dot R^2}{A}\ .
2467: \ee
2468: Now since we are requiring that the exterior metric is static, it exist a
2469: time-like Killing vector $\xi^\alpha$ such that
2470: \ba
2471: \pounds_\xi g_{\alpha\beta}&=\xi_{\alpha}{}_;{}_\beta+\xi_{\beta}{}_;{}_\alpha=0\ ,\
2472: \xi^\alpha \xi_\alpha<0\nonumber
2473: \ea
2474: where $\pounds$ denotes the Lie derivative. If $u^\alpha$ is a geodesic vector,
2475: $u^\alpha\nabla_\alpha u_{\beta}=0$, we have that
2476: \be\label{tildeE}
2477: \tilde E^{1/2}=-\xi^\alpha u_{\alpha}=-F(R)^2A(R)\dot t\ ,
2478: \ee
2479: is a constant along the geodesic motion, or $u^\alpha\nabla_\alpha \tilde E=0$.
2480: Using (\ref{geodesic}) and (\ref{tildeE}), the radial geodesic
2481: equation for the exterior metric gives
2482: $\dot{R}^2=-A(R)+\tilde{E}/F(R)^2$.
2483: Comparing this with eq.~(\ref{geo1}) gives one condition. The
2484: second condition is easier to derive if we change to null
2485: coordinates\footnote{These coordinates, denoted by $(v,r,\theta,\phi)$ are
2486: such that the light-cones are described by the equation $v=\mbox{const}$.}. The
2487: exterior static metric, with
2488: $dv=dt+dR/[F(1-2Gm/R)]$, becomes
2489: \begin{eqnarray}
2490: ds^2&=& -F^2Adv^2 +2FdvdR+R^2d\Omega^2\,.\label{s1'}
2491: \end{eqnarray}
2492: The interior Robertson-Walker metric takes the form~\cite{tesi}
2493: \begin{eqnarray}
2494: ds^2&=& -\tau_{,v}^2\frac{1-(k+\dot{a}^2)R^2/a^2}{1-kR^2/a^2} dv^2
2495: +2 \tau_{,v} \frac{dvdR}{\sqrt{1-kR^2/a^2}}
2496: +R^2d\Omega^2\,,\label{s1''}
2497: \end{eqnarray}
2498: where
2499: \be
2500: d\tau=\tau_{,v}dv+\frac{1+{1\over4}kr^2}{r\dot
2501: a-1+{1\over4}kr^2} dR\ .
2502: \ee
2503: Comparing eqs.~(\ref{s1'}) and (\ref{s1''})
2504: on $\Sigma$ gives the second condition. The two conditions
2505: together imply that $F$ is a constant, which we can take as
2506: $F(R)=1$ without loss of generality (choosing $\tilde E=E+1$), and
2507: that
2508: \begin{equation}\label{s3}
2509: m(R)= M+\frac{3M^2}{8\pi\lambda R^3}+ \frac{Q}{2G_N\lambda R}+ \frac{\Lambda
2510: R^3}{6G_N}\,.
2511: \end{equation}
2512: In the limit $\lambda^{-1}\to 0$, we recover the 4D general relativity
2513: Schwarzschild-de\,Sitter solution. However, we note that the above
2514: form of $m(R)$ violates the weak-field perturbative limit in
2515: eq.~(\ref{pert}), and this is a symptom of the problem with a
2516: static exterior. Equations~(\ref{s1}) and (\ref{s3}) imply that
2517: the brane Ricci scalar is
2518: \begin{eqnarray}\label{m1}
2519: R^\mu{}_\mu=4\Lambda+\frac{9G_N M^2}{2\pi\lambda R^6}\,.
2520: \end{eqnarray}
2521: However, the modified Einstein equations (\ref{eq:effective}) for a vacuum exterior imply that
2522: \ba
2523: R_{\mu\nu}&=&\Lambda g_{\mu\nu}- {\cal E}_{\mu\nu}\\
2524: R^\mu{}_\mu &=& 4\Lambda\,. \label{vac}
2525: \ea
2526: Comparing with eq.~(\ref{m1}), we see that a static exterior is
2527: only possible if $M/\lambda=0\,$. We can therefore interpret $R^\mu{}_\mu$
2528: as a kind of ``potential energy" that must be released from the star during the collapse, due
2529: only to braneworld effects. We emphasize
2530: that this no-go result does not require any assumptions on the
2531: nature of the bulk spacetime.
2532:
2533: In summary, we have explored the consequences for
2534: gravitational collapse of braneworld gravity effects, using the
2535: simplest possible model, i.e.\ an Oppenheimer-Snyder-like collapse on a
2536: generalized Randall-Sundrum type brane. Even in this simplest
2537: case, extra-dimensional gravity introduces new features. Using
2538: only the projected 4D equations, we have shown, independent of the
2539: nature of the bulk, that the exterior vacuum on the brane is
2540: necessarily {\em non-static}. This contrasts strongly with general relativity,
2541: where the exterior is a static Schwarzschild spacetime. Although
2542: we have not found the exterior metric, we know that its non-static
2543: nature arises from (a)~5D bulk graviton stresses, which transmit
2544: effects nonlocally from the interior to the exterior, and (b)~the
2545: non-vanishing of the effective pressure at the boundary, which
2546: means that dynamical information on the interior side can be
2547: conveyed outside. Our results suggest that gravitational collapse
2548: on the brane may leave a signature in the exterior, dependent upon
2549: the dynamics of collapse, so that astrophysical black holes on the
2550: brane may in principle have KK hair.
2551:
2552: We expect that the non-static exterior will be transient and partially {\em
2553: non-radiative}, as follows from a perturbative study of non-static
2554: compact objects, showing that the Weyl term ${\cal E}_{\mu\nu}$ in
2555: the far-field region falls off much more rapidly than a radiative
2556: term~\cite{ssm}. It is reasonable to assume that the exterior
2557: metric will become static at late times and tend to Schwarzschild,
2558: at least at large distances.
2559:
2560:
2561: \subsection{Gravitational collapse of pure Weyl energy}
2562:
2563: The one case that escapes the no-go theorem is $M=0$. In general relativity,
2564: $M=0$ would lead to vacuum throughout the spacetime, but in the
2565: braneworld, there is the tidal KK stress on the brane, i.e.\ the
2566: $Q$-term in eq.~(\ref{geo1}). The possibility of black holes
2567: forming from KK energy density was suggested in~\cite{dmpr}. The
2568: dynamics of a Friedmann universe (i.e.\ without exterior),
2569: containing no matter but only KK energy density (``dark
2570: radiation") has been considered in~\cite{bcg}. In that case, there
2571: is a black hole in the Schwarzschild-AdS bulk, which sources
2572: the KK energy density. Growth in the KK energy density corresponds
2573: to the black hole and brane moving closer together; a singularity
2574: on the brane can arise if the black hole meets the brane. Here we
2575: investigate the collapse of a bound region of homogeneous KK
2576: energy density within an inhomogeneous exterior. It is not clear
2577: whether the bulk black hole model may be modified to describe this
2578: case, and we do not know what the bulk metric is. However, we know
2579: that there must be 5D Weyl curvature in the bulk, and that the
2580: bulk could be pathological, with a more severe singularity than
2581: Schwarzschild-AdS. Even though such a bulk would be unphysical
2582: (as in the case of the Schwarzschild black string \cite{blackstring}), it is
2583: interesting to explore the properties of a brane with collapsing
2584: KK energy density, since this idealized toy model may lead to
2585: important physical insights into more realistic collapse with
2586: matter and KK energy density.
2587:
2588: The exterior is static and unique, and given by the Weyl-charged
2589: de\,Sitter metric
2590: \begin{equation}\label{wd}
2591: ds^2=-Adt^2+\frac{dR^2}{A}+R^2d\Omega^2\,,~M=0\,,
2592: \end{equation}
2593: if $A>0$ ($A$ is given by (\ref{A}) together with the definition (\ref{s3})). For $Q=0$ it is de\,Sitter, with horizon
2594: $H^{-1}=\sqrt{3/\Lambda}$. For $\Lambda=0$ it is the special case
2595: $M=0$ of the solutions given in~\cite{dmpr}, and the length scale
2596: $H_Q^{-1}=\sqrt{|Q|/\lambda}$ is an horizon when $Q>0$; for $Q<0$,
2597: there is no horizon. As we show below, the interplay between these
2598: scales determines the characteristics of collapse.
2599:
2600: For $\Lambda=0$, the exterior gravitational potential is
2601: \begin{equation}
2602: \Phi=\frac{Q}{2\lambda R^2}\,,
2603: \end{equation}
2604: which has the form of a purely 5D potential when $Q>0$. When
2605: $Q<0$, the gravitational force is repulsive. We thus take $Q>0$ as
2606: the physically more interesting case, corresponding to {\em
2607: positive} KK energy density in the interior. However we note the
2608: remarkable feature that $Q>0$ also implies {\em negative} KK
2609: energy density in the exterior:
2610: \begin{equation}
2611: -{\cal E}_{\mu\nu}u^\mu u^\nu=\left\{ \begin{array}{ll}
2612: +3Q/(\lambda R_\Sigma^4)\,, & R<R_\Sigma\,, \\ %&\\
2613: -Q/(\lambda R^4)\,, & R> R_\Sigma\,. \end{array}\right.
2614: \end{equation}
2615: Negativity of the exterior KK energy density in the general case
2616: with matter has been previously noted~\cite{ssm,dmpr}.
2617:
2618: The boundary surface between the KK ``cloud" and the exterior has
2619: equation of motion $\dot{R}^2=E-V(R)$, where $V=A-1$.
2620:
2621: For $\Lambda=0$, the cases are:\\
2622: $\bf Q>0$:\, The cloud collapses
2623: for all $E$, with horizon at $R_{\rm
2624: h}=H_Q^{-1}=\sqrt{Q/\lambda}$. For $E<0$, given that $E>-1$, the
2625: collapse can at most start from rest at
2626: \be
2627: R_{\rm
2628: max}=\sqrt{\frac{Q}{\lambda|E|}}>R_{\rm h}\ .
2629: \ee
2630: {$\bf Q<0$}:\, It follows that if
2631: $E>0$, there is no horizon, and the cloud bounces at
2632: \be
2633: R_{\rm
2634: min}=\sqrt{|Q|/(\lambda E)}\ .
2635: \ee
2636:
2637: For $\Lambda> 0$, the potential is given by
2638: \be
2639: V/ V_{\rm c}=-\left(\frac{R}
2640: {R_{\rm c}}\right)^2\left[1+\epsilon(R_{\rm c}/ R)^4\right]\,,
2641: \ee
2642: where
2643: $V_{\rm c}={H/ H_Q}\,,$ $R_{\rm c}=1/\sqrt{HH_Q}\,,$ and
2644: $\epsilon={\rm sgn}\,Q$ (see fig. (\ref{fig3})). The horizons are given by
2645: \begin{equation}
2646: \left(R_{\rm h}^\pm\right)^2= \frac{R_{\rm c}^2}{2V_{\rm c}} \left[1\pm
2647: \sqrt{1-4\epsilon V_{\rm c}^2}\right]\,.\label{hor}
2648: \end{equation}
2649: If $\epsilon>0$ there may be two horizons; then $R_{\rm h}^-$ is
2650: the black hole horizon and $R_{\rm h}^+$ is a modified de\,Sitter
2651: horizon. When they coincide the exterior is no longer static, but
2652: there is a black hole horizon. If $\epsilon<0$ there is always one
2653: de\,Sitter-like horizon, $R_{\rm h}^+$.
2654: \begin{figure}[t]
2655: \begin{center}
2656: \includegraphics[width=8cm,height=4cm,angle=0]{L.eps}
2657: \end{center}
2658: \caption{ \label{fig3} The potential $V(R)$ for $\Lambda> 0$, with
2659: $R$ given in units of $R_{\rm c}$ and $V$ given in units of
2660: $V_{\rm c}$.}
2661: \end{figure}
2662: {$\bf Q>0$}:\, The potential has a
2663: maximum $-2V_{\rm c}$ at $R_{\rm c}$. If $E>-2V_{\rm c}$ the cloud
2664: collapses to a singularity. For $V_{\rm c}>{1\over2}$, i.e.\
2665: $Q>3\lambda/4\Lambda$, there is no horizon, and a naked
2666: singularity forms. For for $V_{\rm c}={1\over2}$ there is one
2667: black hole horizon $R_{h}^{-}=R_{h}^{+}=H^{-1}/\sqrt{2}$. If
2668: $E\leq -2V_{\rm c}$, then eq.~(\ref{en}) implies $V_{\rm
2669: c}<{1\over2}$, so there are always two horizons in this case.
2670: Either the cloud collapses from infinity down to $R_{\rm min}$ and
2671: bounces, with $R_{\rm min}<R_{\rm h}^+$ always, or it can at most
2672: start from rest at $R_{\rm max} (>R_{\rm h}^-)$, and collapse to
2673: a black hole, where ($\epsilon=1$)
2674: \begin{eqnarray}
2675: R^2_{\stackrel{\scriptstyle \rm min}{\rm max}} =\frac{R_{\rm c}^2}{2V_{\rm c}}
2676: \left[-E \pm \sqrt{E^2- 4\epsilon V_{\rm c}^2}\right]
2677: \,.\label{mini}
2678: \end{eqnarray}
2679: {$\bf Q<0$}:\, The potential is
2680: monotonically decreasing, and there is always an horizon, $R_{\rm
2681: h}^+$. For all $E$, the cloud collapses to $R_{\rm min} (<R_{\rm
2682: h}^+)$, and then bounces, where $R_{\rm min}$ is given by
2683: eq.~(\ref{mini}) with $\epsilon=-1$.
2684:
2685: In summary we have analyzed the idealized collapse of homogeneous KK energy
2686: density whose exterior is static and has purely 5D gravitational
2687: potential. The collapse can either come to a halt and bounce, or
2688: form a black hole or a naked singularity, depending on the
2689: parameter values. This may be seen as a limiting idealization of a
2690: more general spherically symmetric but inhomogeneous case. The
2691: case that includes matter may be relevant to the formation of
2692: primordial black holes in which nonlinear KK energy density could
2693: play an important role.
2694:
2695: \subsection{Holographic limit for $\lambda$ via Hawking process}
2696:
2697: Recently it has been suggested that Hawking evaporation of a four dimensional black hole
2698: may be described holographically by the five-dimensional classical black hole metric \cite{hawholo}.
2699: Here we see how this correspondence works in the case of an Oppenheimer-Snyder collapse.
2700: It seems that the Hawking effect does not have memory of
2701: the collapsing process \cite{bida}, and we expect that the conclusions
2702: we obtain are general.
2703:
2704: A Schwarzschild black hole produces quantum
2705: mechanically a trace anomaly for the vacuum energy-momentum tensor. This is due
2706: to the gravitational energy via curvature of spacetime which excites the vacuum.
2707: We discuss the analogy between the
2708: quantum stress-tensor anomaly and the Ricci anomaly found for a dust cloud
2709: gravitational collapse, (\ref{m1}).
2710:
2711: In order to simplify the problem we will calculate explicitly the conformal anomaly only
2712: for a massless scalar field in a curved background. We will then comment on the general
2713: vacuum excitations and give a measure for the tension of the brane using the correspondence.
2714:
2715: \subsubsection{Green function for a massless scalar field}
2716:
2717: Here we will basically follow \cite{bida}.
2718:
2719: A general local Lagrangian for a scalar field $\Phi$ coupled to gravity is
2720: \be\label{lagmassless}
2721: {\cal L}=\frac{1}{2}\sqrt{-g}\left[\partial_\mu \Phi\partial^\mu \Phi-(m^2-\xi R)\Phi^2\right]\ ,
2722: \ee
2723: where $\xi$ is an a dimensionless constant, $R$ is the Ricci scalar and $m$ is the mass
2724: of the field. The classical equations follow as
2725: \be\label{evolscalar}
2726: \left[\Box + m^2-\xi R\right]\Phi=0\ .
2727: \ee
2728: Since in the following we will consider vacuum excitations,
2729: we make the Lagrangian invariant under conformal transformation of the metric. The only non-trivial possibility is for
2730: $m=0$ and $\xi=1/6$ in four dimensions. Now as usual we define the momentum
2731: \be
2732: \pi=\frac{\partial {\cal L}}{\partial(\partial_t\Phi)}\ ,
2733: \ee
2734: and the quantum commutation rules are
2735: \ba
2736: \left[\Phi(t,x),\Phi(t,x')\right]&=&0\ ,\cr
2737: \left[\pi(t,x),\pi(t,x')\right]&=&0\ ,\cr
2738: \left[\Phi(t,x),\pi(t,x')\right]&=&i\delta^3(x-x')/\sqrt{-g}\ .\nonumber
2739: \ea
2740: If we define the Green function as the two-point correlation function
2741: \be
2742: iG_F(x-x')=\langle 0|T(\Phi(x)\Phi(x'))|0\rangle\ ,
2743: \ee
2744: where $T$ is the temporal order operator and $|0\rangle$ is the vacuum, we have
2745: \be\label{scalarGreen}
2746: \left[\Box -\frac{1}{6} R\right]G_F(x-x')=-\frac{\delta^4(x-x')}{\sqrt{-g}}\ .
2747: \ee
2748: In order to define a meaningful concept of a particle in curved space, we have to work only
2749: with local quantities. This means that we consider, as proper particles, only the high
2750: frequencies in the Fourier space of $\Phi$. Therefore we will be interested in having
2751: a solution for (\ref{scalarGreen}) only
2752: in the limit $x\rightarrow x'$.
2753:
2754: Introducing the Riemann coordinates $y^\mu=x^\mu-x'^\mu$, we have \cite{kre,pet}
2755: \ba
2756: g_{\mu\nu}&=\eta_{\mu\nu}+\frac{1}{3}R_{\mu\alpha\nu\beta}y^\alpha y^\beta-
2757: \frac{1}{6}R_{\mu\alpha\nu\beta}{}_{;\gamma}y^\alpha y^\beta y^\gamma\cr
2758: &+\left[\frac{1}{20}R_{\mu\alpha\nu\beta}{}_{;\gamma\delta}+\frac{2}{45}R_{\alpha\mu\beta\lambda}
2759: R^\lambda{}_{\gamma\nu\delta}\right]y^\alpha y^\beta y^\gamma y^\delta+...\ ,\nonumber
2760: \ea
2761: where all the coefficients are evaluated at $y^\alpha=0$.
2762:
2763: We can now solve (\ref{scalarGreen}) around $y^\alpha=0$
2764: in Fourier space and apply an anti-Fourier-transform, to obtain
2765: \be
2766: G_F(x)=\frac{-i}{(4\pi)^2\sqrt{-g}}\int^\infty_0i ds (is)^{-2}F(x,s)\ ,
2767: \ee
2768: where
2769: \be
2770: F(x,s)=1+a(x)(is)^2\ ,
2771: \ee
2772: and
2773: \be
2774: a(x)=\frac{1}{180}\left[R_{\alpha\beta\gamma\delta}R^{\alpha\beta\gamma\delta}-R^{\alpha\beta}
2775: R_{\alpha\beta}+\Box R\right]\ .
2776: \ee
2777:
2778: \subsubsection{Trace anomaly and the Hawking effect}\label{holocollapse}
2779:
2780: In semiclassical gravity the gravitational field is treated classically and the matter
2781: fields are
2782: treated quantum mechanically. In this case what actually sources the gravitational field is the
2783: expectation value of the energy-momentum tensor,
2784: \be\label{R}
2785: R_{\mu\nu}-{1\over 2}R g_{\mu\nu}+\Lambda_R g_{\mu\nu}=8\pi G_R \langle T_{\mu\nu}\rangle\ ,
2786: \ee
2787: where the subscript $R$ indicates that the cosmological and Newton constants must be
2788: renormalized. It is possible to prove that for a conformal theory we have
2789: $G_R=G_N$, whereas $\Lambda_R$ must be experimentally evaluated \cite{bida}
2790: \footnote{In general the left hand side can have additional higher curvature and derivative
2791: terms
2792: due to the renormalization process. It is possible to find exactly the expression of these
2793: corrections up to multiplicative constants that must be evaluated experimentally \cite{bida}.
2794: We do not really need in the following the equation (\ref{R}), but we wish to comment that
2795: such corrections are not divergent-free, and therefore allow the energy-momentum tensor to be
2796: non-conserved. Since we believe in the Einstein principle that the stress tensor is conserved
2797: by gravitational identities (e.g. Bianchi) and not by additional constraints introduced by hand,
2798: we set these constants zero.}. Now
2799: the classical energy-momentum tensor
2800: is connected to the variation of the action $S_m$ of matter,
2801: \be
2802: \frac{2}{\sqrt{-g}}\frac{\delta S_m}{\delta g^{\mu\nu}}=T_{\mu\nu}\ .
2803: \ee
2804: Given the introduction of quantum degrees of freedom, the semiclassical energy-momentum tensor
2805: differs from the classical one.
2806: This implies that we should find an effective action to encode the
2807: quantum degrees freedom as well. Calling it $W$ we can define
2808: \be
2809: \frac{2}{\sqrt{-g}}\frac{\delta W}{\delta g^{\mu\nu}}=\langle T_{\mu\nu}\rangle\ .
2810: \ee
2811: From path integral formalism, the $n$-point correlation function can be obtained by
2812: the generating functional
2813: \be
2814: Z[J]=\int{\cal D}[\Phi]\exp\left[iS_m[\Phi]+i\int J(x)\Phi(x)\sqrt{-g}d^4x\right]\ ,
2815: \ee
2816: where ${\cal D}[\Phi]$ indicates the integration over all the functions $\Phi$ and
2817: \be
2818: \left(\frac{\delta^n Z}{\delta J(x^1)...\delta J(x^n)}\right)\Big|_{J(x^i)=0}=\langle \mbox{out}
2819: ,0|T(\Phi(x^1)...\Phi(x^n))|0,\mbox{in}\rangle\ .
2820: \ee
2821: Now we have \cite{schwinger}
2822: \be\label{schwinger}
2823: \delta Z[0]=i\int {\cal D}[\Phi]\delta S_m e^{iS_m[\Phi]}=i\langle \mbox{out},0|
2824: \delta S_m|0,\mbox{in}\rangle\ ,
2825: \ee
2826: and therefore we obtain
2827: \be
2828: \frac{2}{\sqrt{-g}}\frac{\delta Z[0]}{\delta g^{\mu\nu}}=i\langle \mbox{out},0|
2829: T_{\mu\nu}|0,\mbox{in}\rangle\ .
2830: \ee
2831: Then
2832: \be
2833: \frac{2}{\sqrt{-g}}\frac{\delta\ln Z[0]}{\delta g^{\mu\nu}}=i\frac{\langle \mbox{out},0|
2834: T_{\mu\nu}|0,\mbox{in}\rangle}{\langle \mbox{out},0|0,\mbox{in}\rangle}=
2835: i\langle T_{\mu\nu}\rangle\ .
2836: \ee
2837: This means that we can identify $Z[0]=e^{iW}$.
2838:
2839: Integrating by parts the action for a massless scalar field with Lagrangian (\ref{lagmassless}),
2840: we get\footnote{The boundary term is taken to be zero.}
2841: \be
2842: S_m=-\frac{1}{2}\int d^4 x\sqrt{-g}\Phi \left(\Box -\frac{1}{6}R\right)\Phi=
2843: -\frac{1}{2}\int d^4 x d^4 y\sqrt{-g}\Phi(x) K_{xy}\Phi(y)\ ,
2844: \ee
2845: where
2846: \be
2847: K_{xy}=\left(\Box-\frac{1}{6}R\right)\delta^4(x-y)\sqrt{-g}\ ,
2848: \ee
2849: so that
2850: \be
2851: K_{xy}^{-1}=-G_F(x,y)\ .
2852: \ee
2853: If we change now the variable in (\ref{schwinger}), using
2854: \be
2855: \Phi'(x)=\int d^4y K_{xy}^{1/2}\Phi(y)\ ,
2856: \ee
2857: we obtain
2858: \be
2859: Z[0]\propto (\mbox{det} K_{xy}^{1/2})^{-1}=\sqrt{\mbox{det}(-G_F)}\ .
2860: \ee
2861: Then
2862: \be
2863: W=-i\ln Z[0]=-\frac{1}{2}i\mbox{tr}[\ln(-G_F)]\ .
2864: \ee
2865: We are here interested in the trace of the energy-momentum tensor, which is
2866: \be
2867: \frac{2}{\sqrt{-g}}g^{\mu\nu}\frac{\delta W}{\delta g^{\mu\nu}}=g^{\mu\nu}\langle T_{\mu\nu}\rangle
2868: =\langle T\rangle\ .
2869: \ee
2870: After laborious calculations \cite{duff} and using renormalization techniques,
2871: one obtains the following trace anomaly for the energy-momentum
2872: tensor
2873: \be\label{traceanomaly}
2874: \langle T\rangle=\frac{1}{2880\pi^2}\left[R_{\alpha\beta\gamma\delta}R^{\alpha\beta\gamma\delta}
2875: -R_{\alpha\beta}R^{\alpha\beta}+\Box R\right]\ .
2876: \ee
2877: We now specialize this anomaly to the Schwarzschild background \cite{chrifu}, obtaining
2878: \be
2879: \langle T\rangle =\frac{G_N^2}{60\pi^2}\frac{M^2}{R^6}\ .
2880: \ee
2881: These calculations are made having in mind that the vacuum is only described by a scalar field,
2882: but for more general fields one can find \cite{chridu}
2883: \be\label{anomalyTq}
2884: \langle T\rangle =-u\frac{G_N^2}{60\pi^2}\frac{M^2}{R^6}\ .
2885: \ee
2886: where
2887: \be
2888: u=12N_1-N_0-14N_{1/2}\ ,
2889: \ee
2890: and $N_{0,1/2,1}$ are respectively the number of species of spins $0,1/2,1$ of the theory considered.
2891:
2892: Now we go back to the gravitational collapse in the braneworld. We can interpret the anomaly
2893: (\ref{m1}) of the Ricci scalar as due to a non-zero energy momentum tensor for the exterior
2894: which is extracting energy (via evaporation) from the collapsing object\footnote{This
2895: in fact is what happens quantum mechanically \cite{bida}.}. Indeed
2896: if we are not too close to the body and we consider only short periods in the evaporation
2897: process, as in standard
2898: semiclassical calculations, we can neglect the
2899: back-reaction on the metric and consider this process as nearly time-independent \cite{DeWitt}.
2900:
2901: The braneworld calculation tells us that in order to release the potential energy
2902: of eq. (\ref{m1}), we must have an effective non-zero energy momentum tensor with trace
2903: \be
2904: T^{\rm eff}=-\frac{9}{16\pi^2\lambda}\frac{M^2}{R^6}\ .
2905: \ee
2906: But this is the same quantum anomaly as (\ref{anomalyTq})
2907: if we identify
2908: \be
2909: \lambda=\frac{135}{4uG_N^2}\ .
2910: \ee
2911: If we take this analogy seriously, then we have an indirect measure for the tension of the brane
2912: \be
2913: \lambda\sim 10 M_p^4 u^{-1}\ ,
2914: \ee
2915: where $M_p\sim 10^{19}\ \mbox{GeV}$ is the Planck mass. Moreover we obtain
2916: information on the theory which describes the Hawking process. Indeed since $\lambda$ is
2917: positive we should have
2918: \be\label{emp}
2919: N_1>\frac{N_0}{12}+\frac{7}{6}N_{1/2}\ .
2920: \ee
2921: As pointed out by \cite{Emparan}, this is incompatible with a SYM theory with ${\cal N}=4$ at
2922: large $N$, which is the quantum counterpart of the AdS/CFT correspondence \cite{M1}.
2923: This means that this process could be a new test of the holographic principle.
2924:
2925: \newpage
2926: \chapter{Cosmology in a generalized braneworld}
2927:
2928: \footnote{This chapter is based on \cite{barcelo, GS1, GS2, DG}.}In recent decades,
2929: developments in cosmology have been strongly
2930: influenced by high-energy physics. A remarkable example of this
2931: is the inflationary scenario and all its variants.
2932: This influence has been growing and becoming more and more important.
2933:
2934: It is a general belief that Einstein gravity is a
2935: low-energy limit of a quantum theory of gravity which is still unknown.
2936: Among promising candidates we have string theory, which suggests that
2937: in order to have a ghost-free action, quadratic curvature corrections to
2938: the Einstein-Hilbert action must be proportional to the Gauss-Bonnet
2939: term~\cite{BoDe:85}. An example has been given already in the first chapter.
2940: This term also plays a fundamental role in Chern-Simons
2941: gravitational theories~\cite{Cha:89}.
2942: However, although being a string-motivated scenario, the RS model and
2943: its generalizations~\cite{SMS} do not include these terms.
2944: From a geometric point of view, the combination of the
2945: Einstein-Hilbert and Gauss-Bonnet term constitutes, for 5D spacetimes,
2946: the most general Lagrangian producing second-order field
2947: equations~\cite{Lov} (see also~\cite{lovelock}).
2948:
2949: These facts provide a strong motivation for the study of braneworld
2950: theories including a Gauss-Bonnet term.
2951: Recent investigations of this issue have shown~\cite{MeOl:00}
2952: that the metric for a vacuum 3-brane (domain wall) is, up to a redefinition
2953: of constants, the warp-factor metric of the RS scenarios. This is because AdS is conformally
2954: Minkowskian and the Gauss-Bonnet term is topological on the boundary. The existence
2955: of a KK zero-mode localized on the
2956: 3-brane producing Newtonian gravity at low energies, has also been demonstrated
2957: ~\cite{local} (see
2958: also~\cite{Ne:01a,ChNeWe:01}). This can be simply proved by considering that higher
2959: curvature terms in the action can produce only
2960: higher order corrections to the Newtonian gravity.
2961: Properties of black hole solutions in
2962: AdS spacetimes have been studied in~\cite{bhads,Ca:01}. The cosmological
2963: consequences of these scenarios are less well understood. This issue has
2964: been studied in~\cite{AbMo:01}. However, only simple ans\"atze for
2965: the 5D metric (written in Gaussian coordinates as
2966: in~\cite{bdl}) were considered, e.g. the separability of the
2967: metric components in the time and extra-dimension coordinates.
2968: One can see that this assumption is too strong even in
2969: RS cosmological scenarios~\cite{bdl}, where it leads to a
2970: very restrictive class of cosmological models, not representative of
2971: the true dynamics.
2972: Other results with higher-curvature terms in braneworld scenarios
2973: are considered in~\cite{others}.
2974:
2975: In this chapter we study the equations governing the dynamics of
2976: FRW cosmological models in braneworld
2977: theories with a Gauss-Bonnet (GB) term (Lanczos gravity \cite{Lanczos}). In doing that
2978: we study the cosmological behaviour of shells
2979: (or branes) that are thin but still of a finite
2980: thickness $T$. In this way we want to shed some light on how
2981: the zero thickness is attained in the presence of GB interactions.
2982: This limit has been studied for Einstein gravity in \cite{Mounaix}.
2983: Thick shells in the context of GB interactions
2984: have been already studied in \cite{Corradini} and \cite{Giovannini}, but
2985: with a focus on different aspects than those here.
2986: The conclusion of our analysis here is twofold.
2987: On the one side, our results show that there is a generalized Friedmann equation
2988: \cite{Charmousis} that can be found by using
2989: a completely general procedure, in which the energy density of the brane
2990: in the thin-limit is related to the averaged density. However,
2991: considering specific geometric configurations, one can find another form for
2992: the Friedmann equation, such as in \cite{GS1},
2993: with a procedure in which
2994: the energy density of the brane in the thin-limit comes from
2995: the value of the boundary density in the thick-brane model.
2996: We also argue that the information lost when treating a
2997: real thin shell
2998: as infinitely thin, is in a sense larger in Lanczos
2999: gravity than in the analogous situation in standard General Relativity.
3000:
3001: Let us explain further this last point. From a physical point of view,
3002: in the process of passing from the notion of function to that of
3003: distribution, one loses information. Many different series of functions
3004: define the same limiting distribution. For example, the series
3005: %
3006: \begin{equation}
3007: f_T(y)=\left\{
3008: \begin{array}[c]{l}
3009: 0 ~~~{\rm for}~~~ |y| > T/2\,, \\
3010: \mbox{} \\
3011: {1 \over T} ~~~{\rm for}~~~ |y| < T/2\,;
3012: \end{array}\right.
3013: ~~~~~~~~
3014: g_T(y)=\left\{
3015: \begin{array}[c]{l}
3016: 0 ~~~{\rm for}~~~ |y| > T/2\,, \\
3017: \mbox{} \\
3018: {12 y^2 \over T^3} ~~~{\rm for}~~~ |y| < T/2\,,
3019: \end{array}\right.
3020: \end{equation}
3021: %
3022: define the same limiting Dirac delta distribution. The distribution
3023: only takes into account the total conserved area delimited by the
3024: series of functions. The gravitational field equations relate geometry
3025: with matter content. If we take the matter content to have some
3026: distributional character, the geometry will acquire also a
3027: distributional character. When analyzing the thin-limit of branes in
3028: Einstein gravity, by constructing series or families of solutions
3029: parametrized by their thickness, we observe that the divergent parts
3030: of the series of functions that describe the density-of-matter profile
3031: transfer directly to the same kind of divergent parts in the
3032: description of the associated geometry. Very simple density profiles
3033: [like the above function $f_T(y)$] are associated with very simple
3034: geometric profiles, and vice-versa. However, when considering
3035: Lanczos gravity this does not happen. The divergent parts
3036: of the series describing the matter density and the geometry are
3037: inequivalent. A simple density profile does not correspond to a very
3038: simple geometric profile and vice-versa; on the contrary, we observe
3039: that they have some sort of complementary behaviour. This result
3040: leads us to argue that the distributional description of the
3041: cosmological evolution of a brane in Lanczos gravity
3042: is hiding important aspects of the microphysics, not present when
3043: dealing with pure Einstein gravity. Also, we find that for simple models
3044: of the geometry, one can make compatible the two seemingly distinct
3045: generalized Friedmann equations found in the literature.
3046: We will then extend these concepts for a domain wall formed by a scalar field.
3047:
3048:
3049: %---------------------------------------------------------------------
3050: \section{Static thick shells in Einstein and Lanczos gravity}
3051: \label{S:static}
3052: %---------------------------------------------------------------------
3053:
3054: %---------------------------------------------------------------------
3055: \subsection{Einstein gravity}
3056: %---------------------------------------------------------------------
3057:
3058: To fix ideas and notation let us first describe the simple case
3059: in which we have a static thick brane in an AdS
3060: bulk. We take an ansatz for the metric of the form
3061: %
3062: \begin{eqnarray}
3063: ds^2= e^{-2A(y)} \eta_{\mu\nu}dx^\mu dx^\nu +dy^2\,,
3064: \label{brane}
3065: \end{eqnarray}
3066: %
3067: where $\eta_{\mu\nu}$ is the four-dimensional Minkowski metric.
3068: Comparing with the formulas given in Appendix~\ref{appa} this means
3069: taking $a(t,y)=n(t,y)=\exp(-2A(y))$, and $b(t,y)=1$.
3070: The energy-momentum tensor has the form
3071: %
3072: \begin{eqnarray}
3073: \tilde\kappa^2\, \tilde T_{AB} = \rho u_{A}u_{B} + p_Lh_{AB} + p_Tn_{A}n_{B}\,,
3074: \label{tAB}
3075: \end{eqnarray}
3076: %
3077: where $u_A=(-e^{2A},\mb{0},0)$ and $n_A=(0,\mb{0},1)$. Here, $\rho$, $p_L$
3078: and $p_T$ represent respectively the energy density, the longitudinal pressure
3079: and the transverse pressure, and are taken to depend only on $y$.
3080: The Einstein equations $\tilde G_{AB}=-\tilde\Lambda \tilde g_{AB}+\tilde\kappa^2 \tilde T_{AB}$
3081: with a negative cosmological constant,
3082: $\tilde\Lambda\equiv -6/l^2$,
3083: result in the following independent equations for the metric function
3084: $A(y)$:
3085: %
3086: \begin{eqnarray}
3087: &&3A''-6A'^2=\rho-{6 \over l^2}\,, \label{eins1} \\
3088: &&6A'^2=p_T+{6 \over l^2}\,, \label{eins2} \\
3089: &&p_L=-\rho\,.
3090: \label{E-equations}
3091: \end{eqnarray}
3092: %
3093: For convenience, we will hide the $\tilde\kappa^2$ dependence inside the matter magnitudes,
3094: $\rho=\tilde\kappa^2\rho_{\rm true}$, etc.
3095: We also consider a $Z_2$-symmetric geometry around $y=0$.
3096: The brane extends in thickness from $y=-T/2$ to $y=+T/2$. Outside
3097: this region $\rho=p_T=0$, so we have a purely AdS spacetime:
3098: $A(y)=-y /l+b$ for $y\in (-\infty,-T/2)$ and $A(y)=y /l+b$ for
3099: $y\in (T/2,+\infty)$. The junction conditions at $y=-T/2,+T/2$ [see Appendix~\ref{appa}]
3100: tell us that
3101: %
3102: \begin{eqnarray}
3103: && A(-T^-/2)=A(-T^+/2),~~~~ A(T^-/2)=A(T^+/2) \,, \\
3104: && A'(-T^-/2)=A'(-T^+/2),~~~~ A'(T^-/2)=A'(T^+/2)\,.
3105: \label{E-junctions}
3106: \end{eqnarray}
3107: %
3108: From this and using (\ref{eins2}), we deduce that the
3109: transversal pressure is zero at the brane boundaries
3110: $p_T(-T/2)=p_T(T/2)=0$. Since we are imposing
3111: $Z_2$-symmetry with $y=0$ as {\em fixed point}, hereafter we
3112: will only specify the value of the different functions in the interval
3113: $(-T/2,0)$.
3114:
3115: The function $A'$ is odd and therefore interpolates
3116: from $A'(-T/2)=-1/l$ to $A'(0)=0$. If in addition
3117: we impose that the null-energy condition $\rho+p_T=3A'' \geq 0$ be satisfied
3118: everywhere inside the brane, then $p_T$ has to be a negative and monotonically
3119: decreasing function from $p_T(-T/2)=0$ to $p_T(0)=-6/l^2$.
3120: This condition will turn out to be fundamental in defining
3121: a thin-shell limit.
3122:
3123: By isolating $A''$ from equations (\ref{eins1}) and (\ref{eins2})
3124: we can relate the
3125: total bending of the geometry on passing through the brane
3126: with its total $\rho+p_T$
3127: %
3128: \begin{eqnarray}
3129: {6 \over l} =3A'\Big|_{-T/2}^{T/2}=\int_{-T/2}^{T/2} (\rho+p_T)\; dy.
3130: \end{eqnarray}
3131: %
3132: At this stage of generality,
3133: one can create different one-parameter families of thick-brane versions
3134: of the Randall-Sundrum thin brane geometry, by parameterizing each
3135: member of a given family by its thickness $T$. The only requirement needed
3136: to do this is that the value of the previous integral must be kept fixed independently
3137: of the thickness of the particular thick-brane geometry. Thus, each particular
3138: family can be seen as a regularization of Dirac's delta distribution.
3139:
3140:
3141: We can realize that, provided the condition $\rho+p_T \geq 0$ is satisfied,
3142: there exists a constant $C$, independent of the thickness $T$, such that
3143: $p_T<C$, that is, the profile for $p_T$ is bounded and will not
3144: diverge in the thin-shell limit.
3145: Therefore, in the limit in which the thickness of the branes goes to
3146: zero, $T\rightarrow 0$, the integral of $p_T$ goes to zero with the
3147: thickness. (Strictly speaking, the thin-shell limit is reached when
3148: $T/l\rightarrow 0$ but throughout this paper we are going to
3149: maintain $l$ as a finite constant.) Instead, the profile of $\rho$ has to
3150: develop arbitrarily large values in order to fulfil
3151: %
3152: \begin{eqnarray}
3153: {6 \over l} =\lim_{T\rightarrow 0} \int_{-T/2}^{T/2} \rho\;dy.
3154: \label{average}
3155: \end{eqnarray}
3156: %
3157:
3158:
3159: In the thin-shell limit, we can think of the Einstein equations as
3160: providing a relation between the characteristics of the density
3161: profile and the shape of the internal geometry. A very complicated
3162: density profile will be associated with very complicated function
3163: $A(y)$. Physically we can argue that when a shell becomes very thin,
3164: one does not care about its internal structure and, therefore, one
3165: tries to describe it in the most simple terms. But what exactly is
3166: the meaning of simple? Here we will adopt two different definitions of
3167: simple: The first is to consider that the internal density is
3168: distributed homogeneously throughout the shell when the shell becomes
3169: very thin; the second is to consider that the profile for $A'$ is
3170: such that it interpolates from $A'(-T/2)=-1/l$ to $A'(0)=0$ through
3171: a straight line, or what is the same, that the internal profile of
3172: $A''$ is constant. Again, we require this for very thin shells. This
3173: geometric prescription is equivalent to asking for a constant internal
3174: scalar curvature, since $R=8A''-20A'^2$, and for every thin shell the term
3175: $A'^2$ is negligible relative to the constant $A''$
3176: term. Hereafter, we will use interchangeably the terms {\em straight
3177: interpolation} or {\em constant curvature} for these models. In
3178: building arbitrarily thin braneworld models, one needs the profiles
3179: for the internal density $\rho$ and the internal $A''$ to acquire
3180: arbitrarily high values (they will become distributions in the limit of
3181: strictly zero thickness). In the first of the two simple models
3182: described, the simplicity applies to the divergent parts of the
3183: matter content; in the second, the simplicity applies to the divergent parts of the geometry.
3184: From the physical point of view advocated in the introduction,
3185: simple profiles are those that do not involve losing information in the
3186: process of taking the limit of strictly zero thickness.
3187:
3188: We analyze each case independently.
3189:
3190: %---------------------------------------------------------------------
3191: \subsubsection{Constant density profile}
3192: %---------------------------------------------------------------------
3193:
3194: We first define for convenience $z\equiv y/T$ as a scale invariant
3195: coordinate inside the brane. Then,
3196: mathematically, the idea that the density profile, which we will assume
3197: to be analytic inside the brane for simplicity, becomes constant
3198: in the thin-shell limit, can be expressed as follows:
3199: %
3200: \begin{eqnarray}
3201: \rho(z)=\sum_n \beta_n(T) \; z^{2n},
3202: \label{constant-density-est}
3203: \end{eqnarray}
3204: %
3205: where
3206: %
3207: \begin{eqnarray}
3208: \lim_{T\rightarrow 0} T ~ \beta_n(T)\rightarrow 0,
3209: ~~~~\forall n\neq 0;~~~~
3210: \lim_{T\rightarrow 0} T ~ \beta_0(T)\rightarrow \rho_b = {\rm constant}\,.
3211: \label{condi}
3212: \end{eqnarray}
3213: %
3214: For these density profiles, the Einstein equations in the
3215: thin-shell limit tell us that
3216: %
3217: \begin{eqnarray}
3218: 3A''=\beta_0(T)-{6 \over l^2}+6A'^2\,.
3219: \end{eqnarray}
3220: %
3221: From here we get the profile for $A'$:
3222: %
3223: \begin{eqnarray}
3224: A'=\sqrt{{\beta_0(T) \over 6}-{1 \over l^2}}
3225: \tan \left(2\sqrt{{\beta_0(T) \over 6}-{1 \over l^2}}\; y \right)\, .
3226: \label{apri}
3227: \end{eqnarray}
3228: %
3229: Notice that this expression only makes sense for $\beta_0(T)>6/l^2$,
3230: but this is just the regime we are interested in. We have to impose now
3231: the boundary condition $A'(T/2)=1/l$ on the previous expression (\ref{apri}),
3232: %
3233: \begin{eqnarray}
3234: {1 \over l}=\sqrt{{\beta_0(T) \over 6}-{1 \over l^2}}
3235: \tan \left(\sqrt{{\beta_0(T) \over 6}-{1 \over l^2}}\; T \right).
3236: \end{eqnarray}
3237: %
3238: In this manner, we have implicitly determined the form of the function
3239: $\beta_0(T)$. In the limit
3240: in which $T\rightarrow 0$ with $T\beta_0(T)\rightarrow \rho_b$,
3241: we find the relation
3242: %
3243: \begin{eqnarray}
3244: {6 \over l}=\rho_b.
3245: \end{eqnarray}
3246: %
3247: This condition is just what we expected from the average
3248: condition (\ref{average}).
3249:
3250:
3251:
3252:
3253:
3254:
3255: %---------------------------------------------------------------------
3256: \subsubsection{Straight interpolation}
3257: %---------------------------------------------------------------------
3258:
3259:
3260: In this case, the mathematical idea that in thin-shell limit
3261: the profile for $A'$ corresponds to a straight interpolation,
3262: can be formulated as
3263: %
3264: \begin{eqnarray}
3265: A''(z)=\sum_n \gamma_n(T) \; z^{2n},
3266: \label{straight-interpolation}
3267: \end{eqnarray}
3268: %
3269: with
3270: %
3271: \begin{eqnarray}
3272: \lim_{T\rightarrow 0} T ~ \gamma_n(T)\rightarrow 0,
3273: ~~~~\forall n\neq 0;~~~~
3274: \lim_{T\rightarrow 0} T ~ \gamma_0(T)\rightarrow {2 \over l}\ .
3275: \label{straight-interpolation-cond}
3276: \end{eqnarray}
3277: %
3278: For these geometries, we find that the associated profiles for
3279: $p_T$ and $\rho$ in the thin-shell limit have the form,
3280: %
3281: \begin{eqnarray}
3282: &&p_T=-{6 \over l^2}\left(1 - 4 z^2 \right) + \omega_1(T,z) \,, \\
3283: &&\rho={6 \over l T}+{6 \over l^2}\left(1 - 4 z^2 \right) +
3284: \omega_2(T,z)\,,
3285: \end{eqnarray}
3286: %
3287: where here and throughout this chapter, $\omega_n(T,z)$ denotes functions
3288: that vanish in the limit $T\rightarrow 0$.
3289: Now, from this density profile we can see that
3290: %
3291: \begin{eqnarray}
3292: \lim_{T\rightarrow 0} \int_{-T/2}^{T/2} \rho\;dy = {6 \over l},
3293: \end{eqnarray}
3294: %
3295: as we expected. Moreover, we can see that the boundary value of the
3296: density satisfies $T\rho\Big|_{T/2} \rightarrow 6/l$ in the thin
3297: shell limit, which is the same condition satisfied by the averaged
3298: density, $T \langle \rho \rangle \rightarrow 6/l$.
3299:
3300: An additional interesting observation for what follows is the following.
3301: The set of profiles that yield constant density in the thin-shell limit
3302: (\ref{constant-density-est}) and straight interpolation for the
3303: geometric profile (\ref{straight-interpolation}) coincide. Therefore,
3304: in the thin-shell limit one can assume at the same time a constant
3305: internal structure for the density and a straight-interpolation for
3306: the geometry.
3307:
3308:
3309:
3310:
3311:
3312:
3313:
3314:
3315: %---------------------------------------------------------------------
3316: \subsection{Lanczos gravity}\label{EinsteinGB}
3317: %---------------------------------------------------------------------
3318:
3319: Let us move now to the analysis of the same ideas in the presence
3320: of the Gauss-Bonnet term. The field equations are now \cite{Lanczos}
3321: \begin{eqnarray}
3322: \tilde G_{AB}+\alpha \tilde H_{AB}=-\tilde \Lambda \tilde g_{AB}+\tilde \kappa^2 \tilde T_{AB}
3323: \,, \label{fieldeq}
3324: \end{eqnarray}
3325: where $\tilde H_{AB}$ is the Lanczos tensor \cite{Lanczos}:
3326: %
3327: \begin{eqnarray}
3328: \tilde H_{AB}=2\tilde R_{ACDE} \tilde R_{B}^{\; CDE}-4\tilde R_{ACBD}\tilde R^{CD}
3329: -4\tilde R_{AC} \tilde R_{B}^{\; C}+2\tilde R\tilde R_{AB}-\frac{1}{2}\tilde g_{AB}L_{GB}
3330: \,, \label{habterm}
3331: \end{eqnarray}
3332: %
3333: where $\sqrt{-\tilde g}L_{GB}$ is the Gauss-Bonnet Lagrangian density,
3334: \begin{eqnarray}
3335: L_{GB} = \tilde R^2 -4\tilde R^{AB}\tilde R_{AB}+\tilde R^{ABCD}\tilde R_{ABCD} \,.
3336: \end{eqnarray}
3337: For the ansatz (\ref{brane}) we obtain (see Appendix~\ref{appa})
3338: %
3339: \begin{eqnarray}
3340: &&3A''(1 - 4\alpha A'^2)-6A'^2(1 - 2\alpha A'^2)=\rho-{6 \over l^2},
3341: \label{EGB-equations-rho} \\
3342: &&6A'^2(1 - 2\alpha A'^2)=p_T+{6 \over l^2}, \label{EGB-equations-pt} \\
3343: &&p_L=-\rho.
3344: \label{EGB-equations-pl}
3345: \end{eqnarray}
3346: %
3347: The junction conditions for the geometry are the same as before, eq.
3348: (\ref{E-junctions}), implying again the vanishing of the transversal
3349: pressure at the boundaries, $p_T=0$.
3350:
3351: In the outside region the solution is a pure AdS spacetime
3352: but with a modified length scale
3353: %
3354: \begin{eqnarray}\label{modlength}
3355: {1 \over \tilde l} \equiv \sqrt{ {1 \over 4\alpha}
3356: \left(1-\sqrt{1- {8\alpha \over l^2}}\right) }.
3357: \end{eqnarray}
3358: %
3359: Now, isolating $A''$ from~(\ref{EGB-equations-rho}) and~(\ref{EGB-equations-pt}),
3360: we can relate the total bending of the geometry on passing through the brane
3361: with the integral of $\rho+p_T$ :
3362: %
3363: \begin{eqnarray}
3364: {6 \over \tilde l}\left(1- {4 \over 3}{\alpha \over {\tilde l}^2}\right)
3365: =(3A'-4 \alpha A'^3)\Big|_{-T/2}^{T/2}=\int_{-T/2}^{T/2} (\rho+p_T)\; dy.
3366: \end{eqnarray}
3367: %
3368: Again, if the condition $\rho+p_T \geq 0$ is fulfilled throughout the brane
3369: we have that in the thin shell limit,
3370: %
3371: \begin{eqnarray}
3372: {6 \over \tilde l}\left(1- {4 \over 3}{\alpha \over {\tilde l}^2}\right)
3373: =\lim_{T \rightarrow 0} \int_{-T/2}^{T/2} \rho\; dy.
3374: \label{gb-average-condition}
3375: \end{eqnarray}
3376: %
3377: At this point we can pursue this analysis in the two simple
3378: cases of constant density profile and straight interpolation.
3379: %
3380:
3381: %---------------------------------------------------------------------
3382: \subsubsection{Constant density profile}
3383: %---------------------------------------------------------------------
3384:
3385:
3386:
3387: Following the same steps as before for a constant density profile
3388: (\ref{constant-density-est})-(\ref{condi}), the equation that one
3389: has to solve in the thin-shell limit is
3390: %
3391: \begin{eqnarray}
3392: 3A''(1-4\alpha A'^2)=\beta_0(T)-{6 \over l^2}+6A'^2(1-2\alpha A'^2)\,.
3393: \end{eqnarray}
3394: %
3395: Introducing the notation $B\equiv A'$ we reduce this equation to the following
3396: integral
3397: %
3398: \begin{eqnarray}
3399: y={1 \over 4\alpha } \int_0^B {(4\alpha B^2-1) \; dB
3400: \over B^4 - {1 \over 2\alpha} B^2
3401: - {1 \over 2\alpha} \left({\beta_0(T) \over 6}-{1 \over l^2}\right)}\,.
3402: \end{eqnarray}
3403: %
3404: The result of performing the integration is
3405: %
3406: \begin{eqnarray}
3407: y={1 \over 2}
3408: \left[
3409: {1 \over \sqrt{-R_{-}}}\tan^{-1}\left({B \over \sqrt{-R_{-}}} \right)
3410: -{1 \over \sqrt{R_{+}}}\tanh^{-1}\left({B \over \sqrt{R_{+}}}\right)
3411: \right],
3412: \end{eqnarray}
3413: %
3414: where
3415: %
3416: \begin{eqnarray}
3417: R_{\pm}={1 \over 4\alpha }
3418: \left[ 1\pm \sqrt{1+ 8\alpha \left({\beta_0(T) \over 6}-{1 \over l^2}\right) }
3419: \right]\,.
3420: \end{eqnarray}
3421: %
3422: Again, by imposing the boundary condition
3423: %
3424: \begin{eqnarray}
3425: {T \over 2}={1 \over 2}
3426: \left[
3427: {1 \over \sqrt{-R_{-}}}\tan^{-1}\left({1 \over \tilde l \sqrt{-R_{-}}} \right)
3428: -
3429: {1 \over \sqrt{R_{+}}}\tanh^{-1}\left({1 \over \tilde l \sqrt{R_{+}}}\right)
3430: \right],
3431: \end{eqnarray}
3432: %
3433: we find the appropriate form for $\beta_0(T)$. With a lengthy but
3434: straightforward calculation, we can check that in the limit $T\rightarrow 0$,
3435: $\beta_0(T)\rightarrow \infty$, we have
3436: %
3437: \begin{eqnarray}
3438: T\beta_0(T)\rightarrow
3439: {6 \over \tilde l}\left(1- {4 \over 3}{\alpha \over {\tilde l}^2}\right),
3440: \end{eqnarray}
3441: %
3442: in agreement with condition (\ref{gb-average-condition}).
3443:
3444: Using this same asymptotic expansion, we can see that, in the
3445: thin-shell limit, the profile for $A'(y)$ satisfies
3446: %
3447: \begin{eqnarray}
3448: A'(y)-{4 \alpha\over 3}A'(y)^3={1 \over 3} \beta_0(T)\; y.
3449: \end{eqnarray}
3450: %
3451: Recursively, one can create a Taylor expansion for $A'(y)$.
3452: The first two terms are given by
3453: %
3454: \begin{eqnarray}
3455: A'(y)&=&{1 \over 3} \beta_0(T)\; y+
3456: {4 \alpha \over 81} \beta_0(T)^3\; y^3+ {\cal O}(y^5)\cr&=&
3457: {1 \over 3} T \;\beta_0(T)\; z+
3458: {4 \alpha \over 81}T^3 \; \beta_0(T)^3\; z^3 + {\cal O}(z^5).
3459: \end{eqnarray}
3460: %
3461: By differentiating this expression we find
3462: %
3463: \begin{eqnarray}
3464: A''(y)={1 \over 3} \beta_0(T)+{4 \alpha \over 27}T^2 \; \beta_0(T)^3\; z^2
3465: +{\cal O}(z^4).
3466: \end{eqnarray}
3467: %
3468: Now, contrary to what happens in Einstein theory, this profile
3469: does not correspond to the set considered in the straight interpolation before
3470: (see
3471: fig. (\ref{F:fig1})).
3472: By looking at (\ref{straight-interpolation}) we can identify
3473: %
3474: \begin{eqnarray}
3475: \gamma_0(T)\equiv{1 \over 3} \beta_0(T), ~~~~
3476: \gamma_1(T)\equiv{4 \alpha \over 27} T^2 \; \beta_0(T)^3.
3477: \end{eqnarray}
3478: %
3479: Then, we can see that
3480: %
3481: \begin{eqnarray}
3482: \lim_{T \rightarrow 0} T\gamma_0(T)=
3483: {2 \over \tilde l}\left(1- {4 \over 3}{\alpha \over {\tilde l}^2}\right)
3484: \neq {2 \over \tilde l}, ~~~~
3485: \lim_{T \rightarrow 0} T\gamma_1(T)=
3486: {32 \alpha \over \tilde l^3}
3487: \left(1- {4 \over 3}{\alpha \over {\tilde l}^2}\right)^3 \neq 0 .
3488: \end{eqnarray}
3489: %
3490: The coefficients $\gamma_n$ do not satisfy the conditions in
3491: (\ref{straight-interpolation-cond}). Therefore, the scalar
3492: curvature does not have a constant profile as does the energy density.
3493: %====================================================
3494: \begin{figure}[htb]
3495: \vbox{
3496: \hfil
3497: \scalebox{0.5}{\includegraphics{constant-density1.eps}}
3498: \hfil
3499: }
3500: \bigskip
3501: \caption{\label{F:fig1} Family of constant density profiles with decreasing
3502: thickness and associated geometric profile for $A'$ and $A''$.}
3503: \end{figure}
3504: %====================================================
3505:
3506:
3507:
3508:
3509: %---------------------------------------------------------------------
3510: \subsubsection{Straight interpolation}
3511: %---------------------------------------------------------------------
3512:
3513: As in the Einstein case, the straight interpolation profile for
3514: $A''$ corresponds to
3515: %
3516: \begin{eqnarray}
3517: A''(z)=\sum_n \gamma_n(T) \; z^{2n},
3518: \label{straight-interpolation2}
3519: \end{eqnarray}
3520: %
3521: with
3522: %
3523: \begin{eqnarray}
3524: \lim_{T\rightarrow 0} T ~ \gamma_n(T)\rightarrow 0,
3525: ~~~~\forall n\neq 0;~~~~
3526: \lim_{T\rightarrow 0} T ~ \gamma_0(T)\rightarrow {2 \over \tilde l}.
3527: \label{straight-interpolation2-cond}
3528: \end{eqnarray}
3529: %
3530: From here we can deduce the associated profiles for $p_T$ and $\rho$
3531: by substituting in (\ref{EGB-equations-rho}) and (\ref{EGB-equations-pt}).
3532:
3533: In the limit $T \rightarrow 0$, the dominant part in the density profile
3534: is
3535: %
3536: \begin{eqnarray}
3537: \rho=3\gamma_0(T)(1-4\alpha \gamma_0(T)^2 \; T^2 \; z^2).
3538: \end{eqnarray}
3539: %
3540: Identifying
3541: %
3542: \begin{eqnarray}
3543: \beta_0(T) \equiv 3\gamma_0(T)\,, ~~~~
3544: \beta_1(T) \equiv -12 \alpha T^2 \; \gamma_0(T)^3\,,
3545: \end{eqnarray}
3546: %
3547: we find that
3548: %
3549: \begin{eqnarray}
3550: \lim_{T \rightarrow 0} T \; \beta_0(T) = {6 \over \tilde l}, ~~~~
3551: \lim_{T \rightarrow 0} T \beta_1(T) \neq 0\,.
3552: \end{eqnarray}
3553: %
3554: Therefore, even in the thin-shell limit, a straight interpolation
3555: in the geometry does not correspond to a constant density profile (see fig.
3556: (\ref{F:fig2})).
3557: In the presence of a Gauss-Bonnet term it is not compatible to
3558: impose a simple description for the
3559: interior density profile and for the geometric warp factor
3560: at the same time.
3561: In the limit of strictly zero thickness (distributional limit), one will
3562: unavoidably lose some information on the combined matter-geometry system.
3563: %
3564: %====================================================
3565: \begin{figure}[htb]\label{straightfig}
3566: \vbox{
3567: \hfil
3568: \scalebox{0.5}{\includegraphics{straight-int1.eps}}
3569: \hfil
3570: }
3571: \bigskip
3572: \caption{\label{F:fig2} Plot of the straight interpolation profile
3573: for the geometric factor $A'$ and its associated density profile.}
3574: \end{figure}
3575: %====================================================
3576:
3577: To finish this section let us make an additional observation.
3578: From expressions (\ref{straight-interpolation2}) and
3579: (\ref{straight-interpolation2-cond}),
3580: we can see that
3581: %
3582: \begin{eqnarray}
3583: A''= \frac{2}{T} A'\Big|^{}_{T/2} + \nu(T,z) ~~~~\mbox{with}~~~~~
3584: \lim_{T\rightarrow 0} T\nu(T,z) = 0 \,. \label{propsi}
3585: \end{eqnarray}
3586: Using this property in~(\ref{EGB-equations-rho},\ref{EGB-equations-pt}),
3587: %
3588: \begin{eqnarray}
3589: \rho +p_T=3A''(1 - 4 \alpha A'^2)={6 \over T} A'(T/2)(1 - 4 \alpha A'^2)\ ,
3590: \end{eqnarray}
3591: %
3592: and evaluating at $y=T/2$, we find
3593: %
3594: \begin{eqnarray}
3595: T\rho\Big|_{T/2}=6A'(1 - 4 \alpha A'^2)\Big|_{T/2}.
3596: \label{boundary-value}
3597: \end{eqnarray}
3598: %
3599: We can see that contrary to what happens in the Einstein case, this condition
3600: is different from the averaged condition (\ref{gb-average-condition}),
3601: %
3602: \begin{eqnarray}
3603: T\langle \rho \rangle=\lim_{T \rightarrow 0} \int_{T/2}^{-T/2} \rho\; dy=
3604: (3A'-4 \alpha A'^3)\Big|_{-T/2}^{T/2}=
3605: 6A'(1-{4 \over 3} \alpha A'^2)\bigg|_{T/2}.
3606: \label{average3}
3607: \end{eqnarray}
3608: %
3609: Therefore, the averaged density and the boundary density are
3610: different, and this is independent of the brane thickness. For this simple
3611: model, in thin-shell limit one can define two different internal
3612: density parameters characterizing the thin brane.
3613: One represents
3614: the total averaged internal density, and can be defined as
3615: %
3616: \begin{eqnarray}
3617: \rho_{\rm av} \equiv \lim_{T \rightarrow 0} T \langle \rho \rangle.
3618: \end{eqnarray}
3619: %
3620: The other represents an internal density parameter calculated by
3621: extrapolating to the interior the value of the density on the
3622: boundary. This density can be defined as
3623: %
3624: \begin{eqnarray}
3625: \rho_{\rm bv} \equiv \lim_{T \rightarrow 0} T \rho\Big|_{T/2}.
3626: \end{eqnarray}
3627: %
3628: The junction conditions for a thin shell
3629: \be
3630: \rho=\frac{6}{\tilde l}\left(1-\frac{4}{3}\frac{\alpha}{\tilde l^2}\right)
3631: \ee
3632: given in \cite{Low:2000pq},
3633: corresponds to the averaged condition (\ref{gb-average-condition}), or
3634: (\ref{average3})
3635: and therefore relates the total bending of the geometry
3636: in passing through the brane to its total averaged density.
3637: Instead, the particular condition analyzed for the boundary value
3638: of $\rho\Big|_{T/2}$ in eq. (\ref{boundary-value}) yields in the thin-shell
3639: limit the junction condition \cite{Germani}
3640: \be
3641: \rho=\frac{6}{\tilde l}\left(1-4\frac{\alpha}{\tilde l^2}\right)\ .
3642: \ee
3643: This condition is only considering information about the boundary value
3644: of the density and not about its average.
3645:
3646: In summary, what this analysis suggests is that in the presence
3647: of the Gauss-Bonnet term, we can not ignore the interior
3648: structure of the brane, by modelling it by a simple model, even
3649: in the thin-shell limit. This point was first made in general terms by \cite{DeruDole}, and we have
3650: made it explicit. We will see again this feature
3651: in the next section on the cosmological dynamics of thick shells.
3652:
3653:
3654:
3655: %---------------------------------------------------------------------
3656: \section{Dynamical thick shells in Einstein and Lanczos gravity}
3657: \label{S:dynamic}
3658: %---------------------------------------------------------------------
3659:
3660: We use the class of spacetime metrics given in (\ref{themetric}),
3661: which contain a FRW metric in
3662: every hypersurface $\{y={\rm const}.\}$, with a matter content described by
3663: an energy-momentum tensor of the form (\ref{tAB}). We consider
3664: the additional assumption of a {\em static} fifth dimension: $\dot{b} = 0$.
3665: We can rescale the coordinate $y$ in such a way that $b=1$.
3666: Then the line element (\ref{themetric}) becomes
3667: %
3668: \begin{equation}
3669: ds^2=-n^2(t,y)dt^2+a^2(t,y)h_{ij}dx^idx^j+dy^2\, .
3670: \end{equation}
3671: %
3672: In Appendix~\ref{appa} we show that the $\{ty\}-$component of the
3673: Lanczos field equations, for the case with a well-defined
3674: limit in Einstein gravity, leads to the equation~(\ref{keyrel}).
3675: In our case it implies the following relation:
3676: %
3677: \begin{equation}
3678: n(t,y)=\xi(t)\dot a(t,y)\,. \label{0y}
3679: \end{equation}
3680: %
3681: The remaining field equations can be written in the form
3682: given in~(\ref{ttcom},\ref{ijcom},\ref{yycom}). In our case they
3683: become\footnote{The coupling constant $\alpha$ used here is one half the
3684: one used in \cite{GS1}.}
3685: %
3686: \begin{eqnarray}
3687: \left[a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\right]' =
3688: \frac{1}{6}(a^4)'\rho \,, \label{ttc}
3689: \end{eqnarray}
3690: \begin{eqnarray}
3691: \frac{n'}{n}\frac{\dot{a}}{a'}
3692: \left[a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\right]'
3693: - {\left[a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\right]^{\displaystyle{\cdot}}\,}' =
3694: 2\dot{a}a'a^2p_L\,, \label{ijc}
3695: \end{eqnarray}
3696: %
3697: %
3698: \begin{eqnarray}
3699: \left[a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\right]^\cdot =
3700: -\frac{1}{6}(a^4)^{\displaystyle{\cdot}} p_T \,, \label{yyc}
3701: \end{eqnarray}
3702: %
3703: where now $\Phi$ is given by
3704: %
3705: \begin{eqnarray}
3706: \Phi=\frac{\dot a^2}{n^2 a^2}+\frac{k}{a^2}-\frac{a'^2}{a^2}
3707: = H^2 +\frac{k}{a^2} -\frac{a'^2}{a^2} \,,
3708: \end{eqnarray}
3709: %
3710: and we define the Hubble function
3711: associated with each $y={\rm const}.$ slice as
3712: %
3713: \begin{eqnarray}
3714: H(t,y)\equiv \frac{\dot a}{n a}\,.
3715: \end{eqnarray}
3716: %
3717: With the assumption $\dot{b}=0$, the field equation (\ref{ijc}) leads to a
3718: conservation equation for matter of the standard form
3719: [see eq.(\ref{5ceq})]:
3720: \begin{eqnarray}
3721: \dot{\rho} = -3\frac{\dot{a}}{a}(\rho+p_L) \,. \label{mconeq}
3722: \end{eqnarray}
3723:
3724:
3725:
3726: Following the approach in the static scenario, we consider here the situation
3727: in which there is a $Z_2$ symmetry and a fixed proper thickness $T$
3728: for the brane. Then one has to solve separately the equations for
3729: the {\it bulk}
3730: ($|y|>T/2$) and the equations for the thick {\it brane} ($|y|<T/2$). The first
3731: step has already been done, and the result is~\cite{GS1}:
3732: \begin{eqnarray}
3733: \Phi+2\alpha\Phi^2+\frac{1}{l^2} = \frac{M}{a^4}\,,~~~~
3734: \mbox{for $|y|>\textstyle{T\over2}$}\,,
3735: \end{eqnarray}
3736: where, as we show in Appendix \ref{B2}, $M$ is a constant that can be identified
3737: with the mass of a black
3738: hole present in the bulk. Once the solution inside the thick brane has been
3739: found, one has to impose the junction conditions~(\ref{cindm},\ref{cincm}) at
3740: $y=\pm T/2$.
3741: %
3742:
3743: The first thing we can deduce from the junction conditions is that the
3744: quantity $\Phi$ is continuous across the two boundaries $y=\pm T/2\,.$
3745: But in general, its transversal derivative, $\Phi'$, will be discontinuous.
3746: Then, using equation (\ref{yyc}) it follows that the transversal pressure
3747: has to be zero on the boundary, $p_T(t,\pm T/2)=0$. At the same time, from
3748: (\ref{yyc}) we deduce that we must always have
3749: %
3750: \begin{eqnarray}
3751: a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\bigg|_{y=\pm T/2}=M\,.
3752: \label{intm}
3753: \end{eqnarray}
3754: %
3755: On the other hand, using again the relation~(\ref{keyrel}), we find that
3756: %
3757: \begin{eqnarray}
3758: H' = -\frac{a'}{a}H~~~~\Rightarrow~~~~\Phi' = -2\frac{a'}{a}
3759: \left(\Phi + \frac{a''}{a}\right)\,,
3760: \end{eqnarray}
3761: %
3762: and then, expanding (\ref{ttc}), we obtain
3763: %
3764: \begin{eqnarray}
3765: (1+4\alpha\Phi)\frac{a''}{a} = \Phi -\frac{2}{l^2} -\frac{1}{3}\rho
3766: \,.
3767: \label{mult}
3768: \end{eqnarray}
3769: %
3770:
3771:
3772:
3773:
3774:
3775:
3776: In the limit $T \rightarrow 0$, the profiles of the density $\rho$ and of
3777: $a''$ diverge, so that these dominant terms in expression~(\ref{mult})
3778: have to be equated. This results in
3779: %
3780: \begin{eqnarray}
3781: (1+4\alpha\Phi)\left({a' \over a}\right)'= -\frac{1}{3}\rho \,.
3782: \label{00-thin-limit}
3783: \end{eqnarray}
3784: %
3785: In what follows we consider the analysis of the Einstein and
3786: Lanczos theories separately.
3787:
3788:
3789: %---------------------------------------------------------------------
3790: \subsection{Einstein gravity}
3791: %---------------------------------------------------------------------
3792:
3793:
3794: In Einstein gravity it is not difficult to write down an equation describing
3795: the dynamics of every layer in the interior of a thick shell.
3796: To that end we take $\alpha=0$ in the equations above.
3797: By integrating (\ref{ttc}) over the interval $(-T/2,y_\ast)$,
3798: and using (\ref{intm},\ref{00-thin-limit}), we arrive at
3799: %
3800: \begin{eqnarray}
3801: \left(H^2 +{k \over a^2} + {1 \over l^2} \right)&=&
3802: \left({a' \over a}\right)^2+
3803: {M \over a^4}+{1 \over 6 a^4}\int_{-T/2}^{y_\ast} (a^4)'\rho\, dy\cr&=&
3804: {1 \over 36} \left(\int_{-y_\ast}^{y_\ast} \rho\, dy \right)^{2}+
3805: {M \over a^4}+{1 \over 6a^4}\int_{-T/2}^{y_\ast} (a^4)'\rho\, dy\,.
3806: \label{layer-Einstein}
3807: \end{eqnarray}
3808: %
3809: A particular layer of matter inside the shell, located at $y=y_\ast$,
3810: can be seen as separating an internal spacetime from a piece of
3811: external spacetime. From the previous equation, we can see that the
3812: cosmological evolution of each layer $y=y_\ast$ in the thick shell
3813: depends on the balance between the integrated density beyond the layer
3814: (external spacetime), and a weighted contribution of the integrated
3815: density in the internal spacetime. Therefore, the dynamics of each
3816: shell layer will be influenced by the particular characteristics of
3817: the internal density profile inside the shell. However, by looking at
3818: this same equation, we can see that the dynamics of the boundary layer
3819: $y_\ast=-T/2$ are only influenced by the total integrated density
3820: throughout the shell:
3821: %
3822: \begin{eqnarray}
3823: \left(H^2 +{k \over a^2} + {1 \over l^2}-
3824: {M \over a^4} \right)\bigg|_{-T/2}=
3825: {1 \over 36}\left(\int_{-T/2}^{T/2} \rho dy\right)^{2}=
3826: {1 \over 36}\left(T\langle \rho \rangle \right)^{2}={1 \over 36}\rho_{\rm av}^2 .
3827: \end{eqnarray}
3828: %
3829: This is the modified Friedmann equation for the cosmological evolution
3830: of the brane \cite{bdl}.
3831:
3832:
3833: %---------------------------------------------------------------------
3834: %\subsubsection{Constant density profile}
3835: %---------------------------------------------------------------------
3836:
3837: In the same manner as with static shells, we analyze
3838: the case in which the density profile tends to a time-dependent value
3839: in the thin-shell limit:
3840: %
3841: \begin{eqnarray}
3842: \rho =\beta_0(T,t)+ \omega_3(T,t,z), ~~~~
3843: \lim_{T \rightarrow 0} T \beta_0(T,t)=\rho_{\rm av}(t)\,, ~~~~
3844: \lim_{T \rightarrow 0} T\omega_3(T,t,z) =0\,.
3845: \label{constant-density}
3846: \end{eqnarray}
3847: %
3848: When $\alpha=0$, eq. (\ref{00-thin-limit}) tells us that if the
3849: density profile depends only on $t$ in the thin-shell limit, then, in
3850: this same limit, the divergent part of the geometry $(a'/a)'$ is also
3851: constant through the brane interior, describing what we called before a
3852: straight interpolation. A simple density profile amounts to a simple
3853: and equivalent geometric profile, and vice-versa. In this same case, but
3854: including the Gauss-Bonnet term, $\alpha \neq 0$, the geometrical factor $(a'/a)'$
3855: will exhibit a non-trivial profile in $y$, even in the thin-shell limit, as discussed
3856: in more detail in the next subsection.
3857:
3858:
3859: In the case in which $\rho$ depends only on time,
3860: eq. (\ref{layer-Einstein}) reads
3861: %
3862: \begin{eqnarray}
3863: \left.\left(H^2 +{k \over a^2} + {1 \over l^2} \right)\right|_{y_\ast}=
3864: \left.\left({a' \over a}\right)^2\right|_{y_\ast}+ {M \over a^4(y_\ast)}+
3865: \frac{1}{6}\rho(t)\left.\left(1+ {a^4_{T/2} \over a^4}\right)\right|_{y_\ast}\,.
3866: \end{eqnarray}
3867: %
3868: From Equation (\ref{00-thin-limit}) we deduce that
3869: %
3870: \begin{eqnarray}
3871: \left.{a' \over a}\right|_{y_\ast}={1 \over 6}\rho_{\rm av} -
3872: {1 \over 3}\int_{-T/2}^{y_\ast} \rho \, dy
3873: =- {1 \over 3}\rho_{\rm av} z_\ast\,,
3874: \end{eqnarray}
3875: %
3876: and integrating we obtain
3877: %
3878: \begin{eqnarray}
3879: a(t,y)=a_0(t)\exp\left(-{1 \over 6}\rho_{\rm av}(t)Tz^{2} \right)\ .
3880: \end{eqnarray}
3881: %
3882: (Remember that $z\equiv y/T$.)
3883: Therefore, at the lowest order in $T$ we have an equation for the
3884: internal geometry of the form
3885: %
3886: \begin{eqnarray}
3887: H_0^2+{k \over a_0^2}+{1 \over l^2}={1 \over 9}\rho_{\rm av}^2 z^2 +
3888: {1 \over 36}\rho_{\rm av}^2(1-4z^2)+\frac{M}{a_0^4}
3889: ={1 \over 36}\rho_{\rm av}^2+\frac{M}{a_0^4},
3890: \end{eqnarray}
3891: %
3892: which is exactly the standard braneworld generalized Friedmann equation~\cite{bdl}.
3893:
3894: %---------------------------------------------------------------------
3895: \subsection{Lanczos gravity}
3896: %---------------------------------------------------------------------
3897:
3898: In the general Lanczos case, eq. (\ref{00-thin-limit}) can be
3899: written as
3900: %
3901: \begin{eqnarray}
3902: \left[1+4\alpha\left(H^2+{k \over a^2}\right)\right]\left({a' \over a}\right)'
3903: -4\alpha \left({a' \over a}\right)^2 \left({a' \over a}\right)'
3904: =-{1 \over 3}\rho.
3905: \label{gb-expanded}
3906: \end{eqnarray}
3907: %
3908: Integrating between $-T/2$ and $T/2$ yields
3909: %
3910: \begin{eqnarray}
3911: 2\left[1+4\alpha\left(H^2+{k \over a^2}\right)\right]
3912: \left({a' \over a}\right)\bigg|_{T/2}
3913: -{8\alpha \over 3}\left({a' \over a}\right)^3 \bigg|_{T/2}
3914: =-{1 \over 3}\langle \rho \rangle T=-{1 \over 3}\rho_{\rm av}\,.
3915: \label{gb-blowup}
3916: \end{eqnarray}
3917: %
3918: The boundary equation (\ref{intm}) can be written as
3919: %
3920: \begin{eqnarray}
3921: \left.\left(H^2+{k \over a^2}\right)\right|_{T/2}-
3922: \left.\left({a' \over a}\right)^2\right|_{T/2}
3923: +2\alpha\left[\left(H^2+{k \over a^2}\right)-\left({a' \over a}\right)^2
3924: \right]^2_{T/2}-\cr-{M \over a^4(T/2)}+{1 \over l^2}=0\,. \label{bcegb}
3925: \end{eqnarray}
3926: %
3927: This is a quadratic equation for $(a'/a)^2\Big|_{T/2}$, with solutions
3928: %
3929: \begin{eqnarray}
3930: \left({a' \over a}\right)^2\bigg|_{T/2}={1 \over 4 \alpha}
3931: \left.\left[1+4\alpha\left(H^2+{k \over a^2}\right)\right|_{T/2}
3932: \pm
3933: \sqrt{1+{8\alpha \over l^2}-{8\alpha M \over a^4}} \right]\,.
3934: \end{eqnarray}
3935: %
3936: From these two roots we take only the minus sign, as it is the only
3937: one with a well-defined limit when $\alpha$ tends to zero.
3938: By squaring (\ref{gb-blowup}) and substituting the above solution
3939: we arrive at a cubic equation for $H^2+k/a^2$, first found
3940: in~\cite{Charmousis}. This cubic equation has a real root
3941: that can be expressed as~\cite{Gregory-Padilla}
3942: %
3943: \begin{eqnarray}
3944: H^2+{k \over a^2}={1 \over 8 \alpha}
3945: \left[
3946: (\sqrt{\lambda^2+\zeta^3}+\lambda)^{2/3}
3947: +
3948: (\sqrt{\lambda^2+\zeta^3}-\lambda)^{2/3}
3949: -2
3950: \right]\,, \label{feqchar}
3951: \end{eqnarray}
3952: %
3953: where
3954: %
3955: \begin{eqnarray}
3956: \lambda \equiv \sqrt{\alpha \over 2} \rho_{\rm av}\,,
3957: ~~~~
3958: \zeta \equiv \sqrt{1+8\alpha V(a) } \equiv
3959: \sqrt{1+{8\alpha \over l^2}-{8\alpha M \over a^4}}\,.
3960: \end{eqnarray}
3961: %
3962: In addition, we need the conservation equation
3963: %
3964: \begin{eqnarray}
3965: \dot \rho = 3 H (\rho+p_L)\,,
3966: \end{eqnarray}
3967: %
3968: which is valid for each section $y=y_*$, and in particular,
3969: for the boundary, $y=T/2$.
3970: This equation can be averaged to give
3971: %
3972: \begin{eqnarray}
3973: T\langle \dot \rho \rangle = 3 \langle H T (\rho+p_L) \rangle = 3 H\Big|_{T/2}
3974: (T \langle \rho \rangle+T \langle p_L \rangle)+{\cal O}(T),
3975: \end{eqnarray}
3976: %
3977: or written in another way,
3978: %
3979: \begin{eqnarray}
3980: \dot \rho_{\rm av}= 3 H\Big|_{T/2}(\rho_{\rm av}+p_{{\rm av}L})\ .
3981: \end{eqnarray}
3982: %
3983: This happens because
3984: %
3985: \begin{eqnarray}
3986: H(t,y)\rightarrow H_0(t,y_0)+{\cal O}(T)
3987: \end{eqnarray}
3988: %
3989: for any $y_0 \in[-T/2,T/2]$, which we have taken as $y_0=T/2$ for convenience.
3990:
3991: We analyze now the simple case of a constant density profile
3992: (\ref{constant-density}). For consistency with the $T \rightarrow 0$
3993: case, we know that
3994: %
3995: \begin{eqnarray}
3996: a(t,y)=a_0(t)[1+T{\tilde a}(t,z)]+ {\cal O}(T^2)\,,
3997: \end{eqnarray}
3998: and therefore, from~(\ref{0y}),
3999: \begin{eqnarray}
4000: n(t,y)= \xi(t)\left[a_0\left(1+T\tilde{a}(t,z)\right)\right]^{\displaystyle{\cdot}}\,.
4001: \end{eqnarray}
4002: %
4003: In the same limit, eq. (\ref{00-thin-limit}) becomes
4004: %
4005: \begin{eqnarray}
4006: {\tilde a}_{,zz}= -{1\over 3}
4007: {\beta_0(T,t)T \over \left[1+4\alpha
4008: \left(H_0^2+{k \over a_0^2}-{\tilde a}_{,z}^2 \right)\right]}.
4009: \end{eqnarray}
4010: %
4011: (Here the subscript $,z$ denotes differentiation with respect to $z$.)
4012: A necessary condition to have a straight interpolation for the geometry
4013: is that ${\tilde a}(t,z)=b(t)Z(z)$. To check whether or not a simple density
4014: profile corresponds to a simple geometrical profile, we can therefore try
4015: to solve this equation by separation of variables.
4016: It is not difficult to see that in order to find a solution
4017: with a well defined Einstein limit, we need
4018: %
4019: \begin{eqnarray}
4020: b(t) = \mu \,,
4021: ~~~~ \mu^{-1}\beta_0(T,t)T = \mu^{-1}\beta_0(T)T = \rho_{\rm av}= {\rm constant},
4022: ~~~~ H_0^2+{k \over a_0^2}=\Lambda_4\,,
4023: \end{eqnarray}
4024: %
4025: where $\mu$ is a constant that can be absorbed into the function $Z(z)$,
4026: and we can take it to be $\mu=1$.
4027: In this way we recover the AdS and dS solutions
4028: for the brane (depending on the sign of the effective four-dimensional
4029: cosmological constant). To find the specific $y$ profile, we have to solve
4030: %
4031: \begin{eqnarray}
4032: Z_{,zz}= -{1\over 3}
4033: {\rho_{\rm av} \over \left[1+4\alpha
4034: \left(\Lambda_4-Z_{,z}^2\right)\right]}.
4035: \end{eqnarray}
4036: %
4037: This equation can be integrated to give
4038: %
4039: \begin{eqnarray}
4040: (1+4\alpha\Lambda_4)Z_{,z}-{4\alpha \over 3}Z_{,z}^3=-{1\over 3}\rho_{\rm av}\, z\,.
4041: \end{eqnarray}
4042: %
4043: For our purposes the specific solution of this cubic equation is
4044: not important. What we want to point out, is that the solution does not
4045: correspond to a straight interpolation as happened in the Einstein case.
4046: So, in general, simple solutions for the matter profile lead to non-trivial
4047: profiles for the scalar curvature even in the thin-shell limit.
4048:
4049: Taking a simple model for the
4050: geometry, the straight interpolation model,
4051: %
4052: \begin{equation}
4053: a(t,z)=a_0(t)-\frac{1}{2}b(t)z^2T\ ,
4054: \label{req}
4055: \end{equation}
4056: %
4057: we deduce the density profile,
4058: using eq. (\ref{gb-expanded}),
4059: %
4060: \begin{eqnarray}
4061: \lim_{T \rightarrow 0}T \rho=
4062: 3\left[1+4\alpha\left(H_0^2+{k \over a_0^2}\right)\right]
4063: \left({b \over a_0}\right)-12\alpha\left({b \over a_0}\right)^3 z^2.
4064: \end{eqnarray}
4065: %
4066: As in the static case, even for very small thickness the density
4067: profile has now a non-trivial structure.
4068: We observe that
4069: %
4070: \begin{eqnarray}
4071: {a' \over a}\bigg|_{T/2}= -{1 \over 2}{b \over a_0} + {\cal O}(T) \,,
4072: ~~~~
4073: \left({a' \over a}\right)'= {2 \over T}\left[ {a' \over a}\bigg|_{T/2}
4074: +{\cal O}(T)\right]\,. \label{extraas}
4075: \end{eqnarray}
4076: %
4077: The second relation and eq.~(\ref{propsi}) coincide in the thin-shell
4078: limit. Therefore, evaluating~(\ref{gb-expanded}) on $y=T/2$, we obtain
4079: %
4080: \begin{eqnarray}
4081: \left[1+4\alpha\left(H^2+{k \over a^2}\right)\right]\left.\left({a' \over a}\right)
4082: \right|_{T/2}
4083: -4\alpha \left.\left({a' \over a}\right)^3\right|_{T/2}
4084: =-{1 \over 6}T \rho\Big|_{T/2}=-{1 \over 6}\rho_{\rm bv}\,. \label{cubic2}
4085: \end{eqnarray}
4086: %
4087: Following the same steps as before, but using this condition
4088: instead of eq. (\ref{gb-blowup}), we arrive at a cosmological generalized
4089: Friedmann equation \cite{GS1} different from that in~\cite{Charmousis} in its form
4090: and in the fact that it depends on the quantity associated with the boundary value of
4091: the energy density, $\rho_{\rm bv}$, instead of the value associated with the
4092: average of the energy density, $\rho_{\rm av}$. Remarkably, the cubic equation
4093: that results from combining the last equation~(\ref{cubic2}) with the boundary
4094: condition~(\ref{bcegb}), becomes in this case linear. That is, the coefficients of the
4095: terms quadratic and cubic in $H^2+k/a^2$ vanish~\cite{GS1}. The modified
4096: Friedmann equation found in this case is:
4097: %
4098: \begin{eqnarray}\label{nonconv}
4099: H^2+{k \over a^2}=
4100: {1 \over \left(1+{8\alpha \over l^2}-{8\alpha M \over a^4}\right)}
4101: {1 \over 36} \rho_{\rm bv}^2 +{1 \over 4 \alpha}\left(
4102: \sqrt{1+{8\alpha \over l^2}-{8\alpha M \over a^4}}-1 \right)\,.
4103: \end{eqnarray}
4104: %
4105: In contrast with the modified Friedmann equation~(\ref{feqchar}), which
4106: was obtained by using a completely general procedure, in order to obtain
4107: this equation we had to use a procedure which required an
4108: extra assumption, namely equation~(\ref{extraas}). Hence it will
4109: not work for profiles of the metric function $a(t,z)$ that do not
4110: satisfy these requirements or equivalent ones. On the other hand,
4111: by looking at the developments here presented, we can conclude that the
4112: different results found in the literature
4113: for the dynamics of a distributional shell have their origin in the
4114: additional internal richness introduced in the brane by the presence
4115: of the GB term.
4116:
4117: We have analyzed and compared how the thin-shell limit of static and
4118: cosmological braneworld models is attained in Einstein and
4119: Lanczos gravitational theories. We have seen that the
4120: generalized Friedmann equation proposed in~\cite{Charmousis} is always
4121: valid and relates the dynamical behaviour of the shell's boundary with
4122: its total internal density (obtained by integrating transversally the
4123: density profile). Instead, the generalized Friedmann equation
4124: proposed in~\cite{GS1} relates the dynamical behaviour of the
4125: shell's boundary with the boundary value of the density within the
4126: brane. This equation is not always valid, only for specific
4127: geometrical configurations.
4128:
4129: Einstein's equations in these models transfer the divergent
4130: contributions of the thin-shell internal density profile to the
4131: structure of the internal geometry in a faithful way. If we do not know
4132: the internal structure of the shell, we can always model it in simple
4133: terms by assuming an (almost) constant density profile and an (almost)
4134: constant internal curvature. However, the GB term makes it incompatible
4135: to have both magnitudes (almost) constant. If the density is (almost)
4136: constant, then the curvature is not, and vice-versa. Therefore
4137: the particular structure of the Lanczos theory
4138: introduces important microphysical features into the matter-geometry
4139: configurations, beyond those in Einstein gravity, that are hidden in the
4140: distributional limit. We see this more clearly in the next section.
4141:
4142: \section{Smooth flat brane model}
4143:
4144: In this section, using an adequate definition of the brane stress-energy tensor,
4145: we confirm the results obtained previously for a Gauss-Bonnet brane model, extending
4146: the straight interpolation case.
4147: To do so, we use an approach
4148: directly based on the field
4149: equations for a smooth flat brane.
4150:
4151: Suppose we can solve the problem of a five-dimensional scalar field with the metric
4152: \be
4153: ds^2=-e^{-2A_B(y)}d\eta^2+dy^2\ ,
4154: \ee
4155: where $d\eta^2$ is the four-dimensional Minkowski metric. Now the field equations read (see
4156: par. (\ref{EinsteinGB})):
4157: \ba
4158: 3A''_B\left(1-4\alpha A'^2_B\right)-p_B^t=\rho_B\ ,
4159: 6A'^2_B(1-2\alpha A'^2_B)-\frac{6}{l^2}=p^t_B\ ,\nonumber
4160: \ea
4161: where $A'_B(y)$ is a family of solutions parameterized by $B$, subject to the boundary
4162: condition
4163: \be
4164: \lim_{y\rightarrow\infty}A_B'(y)=\frac{1}{\tilde l}\ ,
4165: \ee
4166: ($\tilde l$ is defined in (\ref{modlength})), $\rho_B$ is the energy-density of the matter and
4167: $p_B^t$ is the transversal pressure in the $y$ direction. Moreover we have the ``total bending"
4168: condition
4169: \be
4170: \int^\infty_{-\infty}A''_B(y)dy=\frac{2}{\tilde l}\ .
4171: \ee
4172: Now the domain wall limit implies that
4173: \be
4174: \lim_{B\rightarrow\infty}A'_B(y)=\frac{1}{\tilde l}\ \mbox{for}\ \mid y\mid\geq T/2\ ,
4175: \ee
4176: where $T$ defines a ``proper thickness"\footnote{This is not uniquely defined, and can be
4177: for example interpreted as the proper variance of the distribution $A''_B(y)$,
4178: \be
4179: \sigma_B=\left[\int^\infty_{-\infty} A''_B(y)y^2dy/\int^\infty_{-\infty} A''_B(y)dy\right]^{1/2}
4180: \ .\nonumber
4181: \ee}
4182: of the smooth model and is eventually taken to zero.
4183: From it we have
4184: \be\label{totalbending}
4185: \lim_{B\rightarrow\infty}\int^{T/2}_{-T/2}A''_B(y)dy=\frac{2}{\tilde l}\ ,
4186: \ee
4187: and
4188: \be
4189: \lim_{B\rightarrow \infty}p_B^t=0\ \mbox{for}\ \mid y\mid\geq T/2\ .
4190: \ee
4191: The ``total bending" junction condition follows as
4192: \be
4193: \rho_{\rm av}=\lim_{T\rightarrow 0}T\langle \rho\rangle=\lim_{T\rightarrow 0,B\rightarrow
4194: \infty}\int^{T/2}_{-T/2}\rho_B(y)dy=\frac{6}{\tilde l}\left(1-\frac{4\alpha}{3\tilde l^2}
4195: \right)\ .
4196: \ee
4197: We now define another possibility for the junction conditions that we call ``holographic"
4198: junction
4199: conditions. From the integral (\ref{totalbending}) we have, using the average theorem for
4200: integrals,
4201: \be
4202: \lim_{B\rightarrow\infty}\int^{T/2}_{-T/2} A''_B(y)dy=\lim_{B\rightarrow\infty}A''_B(y_s)
4203: T=\frac{2}{\tilde l}\ ,
4204: \ee
4205: where $\mid y_s\mid\leq T/2$ and we call the hypersurface $y=y_s$ the ``screen".
4206: Therefore we have
4207: \be
4208: \lim_{B\rightarrow\infty}TA''_B(y_s)=\frac{2}{\tilde l}=2\lim_{B\rightarrow\infty}
4209: A'_B(T/2)\ .
4210: \ee
4211: Now considering the straight interpolation case we have that $y_s\sim T/2$ for $T\rightarrow 0$, then
4212: \be
4213: \rho_s=\lim_{T\rightarrow 0, B\rightarrow\infty} T\rho_B(y_s)=\frac{6}{\tilde l^2}\left(
4214: 1-\frac{4\alpha}{\tilde l^2}\right)\ .
4215: \ee
4216: This confirms that, even in the smooth model,
4217: \be
4218: \rho_{\rm av}\neq \rho_s\ .
4219: \ee
4220: \subsection{A simple explicit example}
4221:
4222: Consider the following family of solutions
4223: \be
4224: A'_B(y)=\frac{1}{\tilde l}\tanh(By)\ ,
4225: \ee
4226: for which
4227: \be
4228: \rho_{\rm av}=\frac{6}{\tilde l^2}\left(1-\frac{4\alpha}{3\tilde l^2}\right)\ .
4229: \ee
4230: To find the screen, we can try to solve the equation
4231: \be
4232: \frac{2}{\tilde l}=TA''_B(y_s)\ ,
4233: \ee
4234: for $y_s$. This has a solution
4235: \be
4236: y_s=\frac{1}{B}\tanh^{-1}\sqrt{1-\frac{2}{BT}}\ .
4237: \ee
4238: Then we have
4239: \be
4240: A'_B(y_s)=\frac{1}{\tilde l}\sqrt{1-\frac{2}{TB}}\ ,
4241: \ee
4242: so that
4243: \be
4244: \lim_{B\rightarrow\infty}A'_B(y_s)=\lim_{B\rightarrow\infty}A'_B(T/2)=\frac{1}{\tilde l}\ .
4245: \ee
4246: We can therefore conclude
4247: \be
4248: \rho_s=\frac{6}{\tilde l}\left(1-\frac{4\alpha}{\tilde l^2}\right)\ .
4249: \ee
4250:
4251: \section{Holographic description of the dark radiation term}
4252:
4253: The holographic principle has been applied extensively to cosmological cases in the original
4254: Randall-Sundrum model (see \cite{padilla} for a review). At the time of writing this
4255: thesis, this concept
4256: has also been applied to the Gauss-Bonnet case (see e.g. \cite{Padilla,Gregory-Padilla}).
4257: Since this field is
4258: relatively new, in this section we concentrate on the more simple Randall-Sundrum scenario.
4259:
4260: Here we show an example, without going too much into details, of how one can interpret,
4261: holographically,
4262: the Weyl contribution to the non-conventional Friedmann equation
4263: (\ref{nonconv}), in the case
4264: $\alpha=0$, when the Gauss-Bonnet term is switched off\footnote{See also \cite{SV} for a different perspective. Here
4265: the quantum mechanical properties of the 5D Schwarzschild-AdS black hole have been used.}. Splitting the energy density into the
4266: matter energy density ($\rho$)
4267: and brane tension ($\lambda$),
4268: $\rho_{\rm bv}=\rho+\lambda$, the modified Friedmann equation, replacing all the constants,
4269: reads (see also sec. \ref{star})
4270: \begin{equation}
4271: H^2=\frac{8\pi G_N}{3}\rho\left(1+\frac{\rho}{2\lambda}\right)+\frac{M}{a^4}-\frac{k}{a^2}
4272: +\frac{1}{3}\Lambda\ .
4273: \end{equation}
4274: In order to apply the AdS/CFT description, we will set $k=\Lambda=0$. We follow again
4275: \cite{bida}.
4276:
4277: We consider a four-dimensional cosmological model with a Friedmann geometry,
4278: \begin{equation}
4279: ds^2=a(\eta)^2\left[-d\eta^2+d\vec{x}\cdot d\vec{x}\right]\ .
4280: \end{equation}
4281: Since the Universe is accelerating (or decelerating) it produces particles by quantum
4282: vacuum fluctuations. These particles will
4283: back-react on the metric, producing a non-isotropic perturbation of the type
4284: \be
4285: ds^2=a(\eta)^2\left[-d\eta^2+\sum_{i=1}^3\left[1+h_i(\eta)\right] (dx^i)^2\right]\ ,
4286: \ee
4287: where $\mbox{max}\mid h_i(\eta)\mid\ll 1$. For simplicity we will use the constraint
4288: $\sum_{i=1}^3 h_i(\eta)=0$. Considering
4289: the vacuum as a massless scalar field $\phi(x)$, we can decompose it in Fourier space as
4290: \be
4291: \phi(x)=\int d^3k \left[a_k u_k(x)+a_k^\dag u^*_k(x)\right]\ ,
4292: \ee
4293: where $\dag$ means the Hermitian conjugate and $*$ the complex conjugate.
4294:
4295: Given the symmetries of the problem, we can use the following separation of variables,
4296: \be
4297: u_k=(2\pi)^{-3/2}e^{ik\cdot x}\frac{\chi_k(\eta)}{a(\eta)}\ .
4298: \ee
4299: Now the evolution equation (\ref{evolscalar}) in the conformal case and in conformal
4300: time reduces to
4301: \be
4302: \frac{d^2\chi_k}{d\eta^2}-\sum^3_{i=1}k_i^2\chi_k=0\ .
4303: \ee
4304: If we also impose the orthonormality of the $u_k$,$u^*_k$,
4305: we have the conditions
4306: \be
4307: \chi_k\partial_\eta\chi^*_k-\chi^*_k\partial_\eta\chi_k=i\ .
4308: \ee
4309: Since we would like an asymptotically flat spacetime, we also impose
4310: \be
4311: \lim_{\eta\rightarrow\infty}h_i(\eta)=0\ .
4312: \ee
4313: The normalized positive frequency solution for $\eta\rightarrow -\infty$ is
4314: \be
4315: \chi_k^{\rm in}(\eta)=(2k)^{-1/2}e^{-ik\eta}\ .
4316: \ee
4317: Then we can write the integral equation
4318: \be\label{chi}
4319: \chi_k(\eta)=\chi_k^{\rm in}(\eta)+k^{-1}\int_{-\infty}^\eta V_k (\eta')\sin\left[k(\eta-\eta')
4320: \right]
4321: \chi_k(\eta')d\eta'\ ,
4322: \ee
4323: with $V_k=\sum_{i=1}^3 k_i^2 h_i(\eta)$. The production of particles will be in the
4324: late region ($\eta\rightarrow\infty$),
4325: and
4326: \be
4327: \chi_k^{\rm out}(\eta)=\alpha_k\chi^{\rm in}_k(\eta)+\beta_k\chi^{\rm in\ *}_k(\eta)\ ,
4328: \ee
4329: where the Bogolubov coefficients to first order in $V_k$ are
4330: \ba
4331: \alpha_k &=& 1+i\int^\infty_{-\infty}\chi^{\rm in\ *}_k(\eta)V_k(\eta)\chi_k(\eta)d\eta\cr
4332: \beta_k &=& -i\int^\infty_{-\infty}\chi^{\rm in}_k(\eta)V_k(\eta)\chi_k(\eta)d\eta\
4333: .\nonumber
4334: \ea
4335: The number of particles created per proper volume ($n$) will be linked to the probability
4336: of non-zero values for $\chi_k$ in the
4337: output region. Then
4338: \be
4339: n=(2\pi a)^{-3}\int\mid\beta_k\mid^2d^3k\ ,
4340: \ee
4341: with an associated energy density
4342: \be
4343: \rho=(2\pi)^{-3}a^{-4}\int \mid \beta_k\mid^2 k d^3 k\ .
4344: \ee
4345: The second-order approximation in $V_k$ gives
4346: \be
4347: \rho=-(3840\pi^2 a^4)^{-1}\int^\infty_{-\infty}d\eta_1\int^{\infty}_{-\infty}
4348: d\eta_2\ln\left[2i(\eta_1-\eta_2)\right]
4349: \sum_i(\partial^3_\eta h_i(\eta_1))(\partial^3_\eta h_i(\eta_2))\ .
4350: \ee
4351: Physically the perturbation will be a composition of damped oscillating functions of the form
4352: \be
4353: h_i(\eta)=e^{-\alpha\eta^2}\cos(\beta\eta^2+\delta_i)\ ,
4354: \ee
4355: where $\delta_i-\delta_{i+1}=2/3\pi$. Then we finally have
4356: \be
4357: \rho=\frac{1}{2880\pi}\frac{(\alpha^2+\beta^2)^{1/2}}{\alpha^3 a^4}\ .
4358: \ee
4359: If we now associate
4360: \be
4361: \frac{1}{2880\pi}\frac{(\alpha^2+\beta^2)^{1/2}}{\alpha^3}=M\ ,
4362: \ee
4363: we have an holographic description of the projected Weyl tensor on the brane. Since this
4364: quantum
4365: effect is due only to the geometry, independently of the matter content, it is physically
4366: reasonable
4367: to connect it with the projected Weyl tensor, which is a purely geometrical quantity
4368: dependent only on the bulk
4369: curvature.
4370:
4371: \newpage
4372: \chapter{Conclusions}
4373:
4374: In this thesis I studied the Randall-Sundrum mechanism of localization of gravity in the presence
4375: of an infinitely large extra dimension, from the astrophysical and cosmological point of view.
4376: I focused on the exact basic models. As a
4377: motivation for studying these scenarios in light of the holographic principle,
4378: I showed their relations with known quantum
4379: solutions in four-dimensional gravity.
4380:
4381: \subsection*{Astrophysics}
4382: I investigated how 5-dimensional gravity can affect static
4383: stellar solutions on the brane. I found exact braneworld
4384: generalizations of the uniform-density stellar solution, and used
4385: this to estimate the local (high-energy) effects of bulk gravity.
4386: I derived astrophysical lower limits on the brane tension.
4387: I also found that the star is less
4388: compact than in general relativity. The smallness of high-energy corrections to
4389: stellar solutions flows from the fact that $\lambda$ is well above
4390: the energy density $\rho$ of stable stars. However nonlocal
4391: corrections from the bulk Weyl curvature (5-dimensional graviton
4392: effects) have qualitative implications that are very different
4393: from general relativity.
4394:
4395: The Schwarzschild solution is no longer the unique asymptotically
4396: flat vacuum exterior; in general, the exterior carries an imprint
4397: of nonlocal bulk graviton stresses. The exterior is not uniquely
4398: determined by matching conditions on the brane, since the
4399: 5-dimensional metric is involved via the nonlocal Weyl stresses.
4400: I demonstrated this explicitly by giving two exact exterior
4401: solutions, both asymptotically Schwarzschild. Each exterior which
4402: satisfies the matching conditions leads to a bulk metric, which
4403: could in principle be determined locally by numerical integration.
4404: Without any exact or
4405: approximate 5-dimensional solutions to guide us, we do not know
4406: how the properties of the bulk metric, and in particular its
4407: global properties, will influence the exterior solution on the
4408: brane.
4409:
4410: Guided by perturbative analysis of the static weak field
4411: limit, I made the following conjecture:
4412: {\em if the bulk for a static stellar solution on the brane is
4413: asymptotically AdS and has regular Cauchy horizon, then the
4414: exterior vacuum which satisfies the matching conditions on the
4415: brane is uniquely determined, and agrees with the perturbative
4416: weak-field results at lowest order.} An immediate implication of
4417: this conjecture is that the exterior is not Schwarzschild, since
4418: perturbative analysis shows that there are nonzero Weyl stresses
4419: in the exterior (these stresses are the manifestation
4420: on the brane of the massive Kaluza-Klein bulk graviton modes). In
4421: addition, the two exterior solutions that I present would be
4422: ruled out by the conjecture, since both of them violate the
4423: perturbative result for the weak-field potential.
4424:
4425: The static problem is already complicated, so that analysis of
4426: dynamical collapse on the brane can be very difficult. However,
4427: the dynamical problem gives rise to more striking features.
4428: Energy densities well above the brane tension could be reached
4429: before horizon formation, so that high-energy corrections could be
4430: significant.
4431: In this direction I explored the consequences for
4432: gravitational collapse of braneworld gravity effects, using the
4433: simplest possible model, i.e.\ an Oppenheimer-Snyder-like collapse on a
4434: generalized Randall-Sundrum type brane. Even in this simplest
4435: case, extra-dimensional gravity introduces new features. Indeed using
4436: only the projected 4D equations, I have shown, independent of the
4437: nature of the bulk, that the exterior vacuum on the brane is
4438: necessarily {\em non-static}. This contrasts strongly with GR,
4439: where the exterior is a static Schwarzschild spacetime. Although
4440: I have not found the exterior metric, I know that its non-static
4441: nature arises from (a)~5D bulk graviton stresses, which transmit
4442: effects nonlocally from the interior to the exterior, and (b)~the
4443: non-vanishing of the effective pressure at the boundary, which
4444: means that dynamical information on the interior side can be
4445: conveyed outside. My results suggest that gravitational collapse
4446: on the brane may leave a signature in the exterior, dependent upon
4447: the dynamics of collapse, so that astrophysical black holes on the
4448: brane may in principle have KK hair.
4449:
4450: I expect that the non-static property of the exterior will be transient and {\em
4451: non-radiative}, as follows from a perturbative study of non-static
4452: compact objects, showing that the Weyl term ${\cal E}_{\mu\nu}$ in
4453: the far-field region falls off much more rapidly than a radiative
4454: term. It is reasonable to assume that the exterior
4455: metric will tend to be static at late times and tend to Schwarzschild,
4456: at least at large distances.
4457:
4458: Moreover I showed that this non-static behaviour is due to an anomaly of the Ricci scalar.
4459: Indeed assuming a static exterior,
4460: I found that in the absence of the cosmological constant, it does not vanish as one must expect
4461: from
4462: a vacuum solution. This means that there is a ``potential energy" stored on the boundary of
4463: the star. This energy must
4464: be released in some way producing a time signature in the exterior. Since this anomaly resembles
4465: very much the Weyl anomaly due to the Hawking process, it is reasonable to conjecture
4466: that the time signal can be
4467: holographically reproduced by an ``evaporation process". Using this holographic point of view
4468: I derived an indirect measure for the brane vacuum energy, $\lambda\sim 10 M_p^4 u^{-1}$,
4469: where $u$ depends on the number of fields involved in the Hawking process.
4470:
4471: \subsection*{Cosmology}
4472:
4473: I analyzed and compared how the thin-shell limit of static and
4474: cosmological braneworld models is attained in Einstein and
4475: Lanczos gravitational theories. I showed that the
4476: generalized Friedmann equation proposed in~\cite{Charmousis} is always
4477: valid and relates the dynamical behaviour of the shell's boundary with
4478: its total internal density (obtained by integrating transversally the
4479: density profile). By contrast, the generalized Friedmann equation
4480: proposed in~\cite{Germani} relates the dynamical behaviour of the
4481: shell's boundary with the boundary value of the density within the
4482: brane. This equation is not always valid, but only for specific
4483: geometrical configurations.
4484:
4485: Einstein equations in these models transfer the divergent
4486: contributions of the thin-shell internal density profile to the
4487: structure of the internal geometry in a faithful way. If one does not know
4488: the internal structure of the shell, one can always model it in simple
4489: terms by assuming an (almost) constant density profile and an (almost)
4490: constant internal curvature. However, the Gauss-Bonnet term makes it incompatible
4491: to have both magnitudes (almost) constant. If the density is (almost)
4492: constant, then the curvature is not, and vice-versa. Therefore, one
4493: can say that the particular structure of the Lanczos theory
4494: introduces important microphysical features to the matter-geometry
4495: configurations beyond those in Einstein gravity, that are hidden in the
4496: distributional limit.
4497:
4498: Studying a smooth flat brane model I generalized these conclusions for a more physical model.
4499: In the particular example I gave, one can introduce two types of junction condition
4500: which relate the
4501: ``total bending" of the brane with the matter content. The one I called the
4502: ``total bending"
4503: junction condition that relates the bending to the total matter inside the brane. The second
4504: junction condition relates instead the matter content in a particular
4505: hypersurface called the ``screen",
4506: with the total bending. In this screen all the information of the
4507: total bending is stored. I call this a ``holographic" junction condition.
4508:
4509: Independently of the gravitational
4510: theory used, I showed how to derive the modified Friedmann equation and how it is related
4511: to the black hole solution of the theory.
4512:
4513: In particular for the simplest case of the
4514: Randall-Sundrum scenario I showed how to interpret holographically the black hole mass using quantum particle
4515: production with an FRW geometry.
4516:
4517: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4518:
4519:
4520: \appendix
4521: \newpage
4522: \chapter{Junction conditions}\label{Appjc}
4523:
4524: In this appendix we describe the junction conditions of a non-null hypersurface $\Sigma$,
4525: following
4526: \cite{Synge}.
4527: We use the principle of the continuity of geodesics across any
4528: hypersurface. Since the geodesic equation involves at most first-order derivatives of the metric,
4529: we require the continuity of the metric together with its first derivative.
4530:
4531: If $n$ is the unit normal
4532: vector of the hypersurface $\Sigma$, it is always possible, at least locally, to define Gaussian
4533: normal coordinates such that $x^n$ is the adapted coordinate to $n$ and
4534: $a,b,c...$ are the adapted coordinates to $\Sigma$. Then the following are continuous quantities:
4535: \be
4536: g_{\alpha\beta},\ g^{\alpha\beta}\ ,\ \partial_\gamma g_{\alpha\beta}\ ,\ \partial_{\gamma}
4537: g^{\alpha\beta}\ ,
4538: \ \partial_{an}g^{\alpha\beta}\ ,\ \partial_{ab}
4539: g_{\alpha\beta}\ ,
4540: \ \partial_{ab}g^{\alpha\beta}{}\ .
4541: \ee
4542: It follows that the $G^n_\alpha$ component must be continuous.
4543: In a covariant form,
4544: \be\label{Appjc2}
4545: [G^\alpha{}_\beta n^\beta]\Big|_{\Sigma}=0\ ,
4546: \ee
4547: where $[f]\Big|_\Sigma\equiv f\Big|_{\Sigma^+}-f\Big|_{\Sigma^-}$,
4548: and $\Sigma^\pm$ are respectively
4549: the outer and inner face of $\Sigma$. Equivalently
4550: \ba
4551: [g_{\alpha\beta}]\Big|_\Sigma = 0\ ,\
4552: [K_{\alpha\beta}]\Big|_\Sigma = 0\ ,\label{jcK}
4553: \ea
4554: where $K_{\alpha\beta}=\pounds_n g_{\alpha\beta}/2$ is the extrinsic curvature of $\Sigma$.
4555:
4556: We now work out the so-called Israel junction conditions for thin shells in this case.
4557: Since we are interested in $Z_2$-symmetric branes, this considerably simplifies the calculations.
4558:
4559: Suppose we have two parallel hypersurfaces $\Sigma_1$ and $\Sigma_2$. We again
4560: introduce Gaussian normal coordinates, where the coordinate $x^n$ defines the orthogonal
4561: direction of both $\Sigma_1$ and $\Sigma_2$ . Moreover we fix the point $x^n=0$ as the centre
4562: point, and we call $T$ the proper distance between the hypersurfaces.
4563: In these coordinates, the $Z_2$ symmetry means
4564: \ba
4565: g_{\alpha\beta}(x^n)&=&g_{\alpha\beta}(-x^n)\ ,\cr
4566: K_{\alpha\beta}(x^n)&=&-K_{\alpha\beta}(-x^n)\ . \nonumber
4567: \ea
4568: Therefore from (\ref{jcK}) we have
4569: \ba
4570: K_{\alpha\beta}\Big|_{\Sigma_1}=-K_{\alpha\beta}\Big|_{\Sigma_2}\ ,
4571: \ea
4572: or equivalently, making explicit only the $x^n$ dependence of $K_{\alpha\beta}$,
4573: \be
4574: K_{\alpha\beta}(T/2)=-K_{\alpha\beta}(-T/2)\ .
4575: \ee
4576: We consider the incremental ratio in the thin shell limit ($T\rightarrow 0$),
4577: \be
4578: \lim_{T\rightarrow 0}\frac{K_{\alpha\beta}(T/2)-K(-T/2)}{T}=\partial_{x^n}K_{\alpha\beta}(0)\ .
4579: \ee
4580: This behaves like a Dirac delta function. It is a straightforward exercise to show
4581: that if $\Phi(x^n)$ is a smooth test function, then
4582: \be
4583: \int^{T/2}_{-T/2}\partial_{x^n}K_{\alpha\beta} \Phi(x)dx=2K_{\alpha\beta}(0) \Phi(0)\ .
4584: \ee
4585: Therefore in the thin limit, $K_{\alpha\beta}$ ``jumps" or more rigorously\footnote{Since
4586: in this limit $\Sigma_1\rightarrow \Sigma_2$, $\Sigma$ denotes one of the two equivalent
4587: hypersurfaces.}
4588: \be
4589: [K_{\alpha\beta}]=2K_{\alpha\beta}\Big|_{\Sigma}\ ,
4590: \ee
4591: where here we define $[f]=f(0^+)-f(0^-)$.
4592:
4593: \newpage
4594: \chapter{5D geometry}
4595: \section{Lanczos gravity}\label{appa}
4596: In this appendix we present the main geometrical quantities and field
4597: equations associated with the 5D metric
4598: \begin{eqnarray}
4599: ds^2 = \tilde g_{AB}dx^Adx^B = -n^2(t,y)dt^2 + a^2(t,y)h_{ij}(x^k)dx^idx^j + b^2(t,y)dy^2\,,
4600: \label{themetric}
4601: \end{eqnarray}
4602: where $h_{ij}$ is the metric of the three-dimensional maximally symmetric
4603: surfaces $\{y=\mbox{const.\}}$, whose spatial curvature is parametrized
4604: by $k=-1,0,1$. A particular representation of $h_{ij}$ is
4605: \begin{eqnarray}
4606: h_{ij}dx^idx^j = \frac{1}{\left(1+\frac{k}{4}r^2\right)^2}\left(dr^2 +
4607: r^2d\Omega^2_2\right)\,,
4608: \end{eqnarray}
4609: where $d\Omega^2_2$ is the metric of the 2-sphere. The metric~(\ref{themetric})
4610: contains as particular cases the metrics used in this thesis.
4611:
4612: The non-zero components of the Einstein tensor $\tilde G_{AB}$ corresponding to this line
4613: element are given by ($\dot{Q}=\partial_tQ$, $Q'=\partial_yQ$):
4614: \begin{eqnarray}\label{second}
4615: \tilde G_{tt} & = & 3\left\{ n^2\Phi + \frac{\dot{a}}{a}\frac{\dot{b}}{b}-
4616: \frac{n^2}{b^2}\left[\frac{a''}{a}-\frac{a'}{a}\frac{b'}{b}\right]\right\}
4617: \,, \cr
4618: \tilde G_{ty} & = & 3\left( \frac{\dot{a}}{a}\frac{n'}{n}+\frac{a'}{a}\frac{\dot{b}}{b}
4619: -\frac{\dot{a}'}{a}\right) \,, \cr
4620: \tilde G_{ij} & = & \frac{a^2}{b^2}h_{ij}\left\{\frac{a'}{a}\left(\frac{a'}{a}
4621: +2\frac{n'}{n}\right)-\frac{b'}{b}\left(\frac{n'}{n}+2\frac{a'}{a}\right)
4622: +2\frac{a''}{a} + \frac{n''}{n}\right\}
4623: \cr
4624: & - & \frac{a^2}{n^2}h_{ij}\left\{\frac{\dot{a}}{a}\left(\frac{\dot{a}}{a}-
4625: 2\frac{\dot{n}}{n}\right) - \frac{\dot{b}}{b}\left(\frac{\dot{n}}{n}
4626: -2\frac{\dot{a}}{a}\right) + 2\frac{\ddot{a}}{a}+\frac{\ddot{b}}{b} \right\}
4627: - kh_{ij}\,, \cr
4628: \tilde G_{yy} & = & 3\left\{-b^2\Phi + \frac{a'}{a}\frac{n'}{n} - \frac{b^2}{n^2}
4629: \left[\frac{\ddot{a}}{a}-\frac{\dot{a}}{a}\frac{\dot{n}}{n}\right]\right\} \,,
4630: \end{eqnarray}
4631: where
4632: \begin{equation}
4633: \Phi(t,y) = \frac{1}{n^2}\frac{\dot{a}^2}{a^2}-\frac{1}{b^2}\frac{a'^2}{a^2}
4634: +\frac{k}{a^2}\,. \label{defphi}
4635: \end{equation}
4636: Apart from the metric and the Einstein tensor, the field equations in
4637: Lanczos gravity~(\ref{fieldeq}) contain a term quadratic in the
4638: curvature, namely $\tilde H_{AB}$ [see eq.~(\ref{habterm})].
4639: The non-zero components of this tensor can be written as
4640: \begin{eqnarray}
4641: \tilde H_{tt} & = & 6\Phi\left[
4642: \frac{\dot{a}}{a}\frac{\dot{b}}{b}+\frac{n^2}{b^2}\left(
4643: \frac{a'}{a}\frac{b'}{b}-\frac{a''}{a} \right) \right]\,, \nonumber \\
4644: \tilde H_{ty} & = &
4645: 6\Phi\left(\frac{\dot{a}}{a}\frac{n'}{n}+\frac{a'}{a}\frac{\dot{b}}{b}
4646: -\frac{\dot{a}'}{a}\right)\,, \nonumber \\
4647: \tilde H_{ij} & = & 2a^2h_{ij}\left\{ \Phi\left[\frac{1}{n^2}\left(\frac{\dot{n}}{n}
4648: \frac{\dot{b}}{b} - \frac{\ddot{b}}{b}\right) - \frac{1}{b^2}\left(
4649: \frac{n'}{n}\frac{b'}{b}-\frac{n''}{n}\right)\right] \nonumber \right. \\
4650: & + & \frac{2}{a^2bn}\left[ \frac{\dot{a}^2\dot{b}\dot{n}}{n^4}
4651: +\frac{a'^2b'n'}{b^4} + \frac{\dot{a}a'}{b^2n^2}\left(
4652: b'\dot{n}-\dot{b}n'\right) \right] \nonumber \\
4653: & - & 2\left[ \frac{1}{n^2}\frac{\ddot{a}}{a}\left(
4654: \frac{1}{n^2}\frac{\dot{a}}{a}
4655: \frac{\dot{b}}{b} + \frac{1}{b^2}\frac{a'}{a}\frac{b'}{b}\right)-\frac{1}{b^2}
4656: \frac{a''}{a}\left(\frac{1}{n^2}\frac{\dot{a}}{a}\frac{\dot{n}}{n}
4657: +\frac{1}{b^2}\frac{a'}{a}\frac{n'}{n}\right)\right] \nonumber \\
4658: & + & \left. \frac{2}{b^2n^2}\left[\frac{\ddot{a}}{a}\frac{a''}{a}-
4659: \frac{\dot{a}^2}{a^2}\frac{n'^2}{n^2}-\frac{a'^2}{a^2}\frac{\dot{b}^2}{b^2}
4660: -\frac{\dot{a}'}{a}\left(\frac{\dot{a}'}{a}-2\frac{\dot{a}}{a}\frac{n'}{n}
4661: -2\frac{a'}{a}\frac{\dot{b}}{b}\right) \right] \right\}\,, \nonumber \\
4662: \tilde H_{yy} & = & 6\Phi\left[\frac{a'}{a}\frac{n'}{n}+\frac{b^2}{n^2}\left(
4663: \frac{\dot{a}}{a}\frac{\dot{n}}{n}-\frac{\ddot{a}}{a}\right)\right]\,.
4664: \end{eqnarray}
4665:
4666: In this thesis we consider the situation in which a thick brane is embedded
4667: in the five-dimensional spacetime described by (\ref{themetric}), whose boundaries
4668: are located at $y=\mbox{const.}$ hypersurfaces. Consider the usual junction
4669: conditions at a hypersurface $\Sigma_{y_c}\equiv\{y=y_c\}$,
4670: that is, the continuity of the induced metric, $g_{AB}=\tilde g_{AB}-n_An_B$ and the
4671: extrinsic curvature, $K_{AB} = -g^{C}_{(A}g^D_{B)}\nabla^{}_C n^{}_D$,
4672: of $\Sigma_{y_c}$:
4673: \begin{eqnarray}
4674: n(t,y^+_c) = n(t,y^-_c) \,, ~~~~ a(t,y^+_c) = a(t,y^-_c) \,, \label{cindm}
4675: \end{eqnarray}
4676: \begin{eqnarray}
4677: \frac{n'(t,y^+_c)}{b(t,y^+_c)} = \frac{n'(t,y^-_c)}{b(t,y^-_c)} \,, ~~~~
4678: \frac{a'(t,y^+_c)}{b(t,y^+_c)} = \frac{a'(t,y^-_c)}{b(t,y^-_c)} \,. \label{cincm}
4679: \end{eqnarray}
4680:
4681:
4682: We assume a matter content described by an energy-momentum tensor
4683: of the form
4684: \begin{eqnarray}
4685: \tilde \kappa^2\, \tilde T_{AB} = \rho u_{A}u_{B} + p_Lh_{AB} + p_Tn_{A}n_{B}\,,
4686: \end{eqnarray}
4687: where
4688: \begin{eqnarray}
4689: u_A=(-n(t,y),\mb{0},0)\,,~~~~
4690: h_{AB} = \tilde g_{AB} + u_Au_B - n_An_B \,,~~~~
4691: n_A=(0,\mb{0},b(t,y))\,,
4692: \end{eqnarray}
4693: where $\rho\,,$ $p_L\,,$ and $p_T$ denote, respectively, the energy density and
4694: the longitudinal and transverse pressures with respect to the
4695: observers $u^A$. They are functions of $t$ and $y$. The energy-momentum
4696: conservation equations, $\nabla_A \tilde T^{AB}=0$, reduce to:
4697: \begin{eqnarray}
4698: \dot{\rho} = -\frac{\dot{b}}{b}(\rho+p_T) -3\frac{\dot{a}}{a}(\rho+p_L)\,,
4699: \label{5ceq}
4700: \end{eqnarray}
4701: \begin{eqnarray}
4702: p_T' = -3\frac{a'}{a}(p_T-p_L)-\frac{n'}{n}(\rho+p_T) \,.
4703: \end{eqnarray}
4704:
4705: The $\{ty\}$-component of the field equations for the metric
4706: (\ref{themetric}) in Lanczos gravity [eq.~(\ref{fieldeq})] has the
4707: form
4708: \begin{eqnarray}
4709: \left(1+4\alpha\Phi\right)\left(\frac{\dot{a}}{a}\frac{n'}{n}+
4710: \frac{a'}{a}\frac{\dot{b}}{b}-\frac{\dot{a}'}{a}\right) = 0\,. \label{tycom}
4711: \end{eqnarray}
4712: If we discard the possibility $1+4\alpha\Phi=0$ by restricting ourselves
4713: to models with a well-defined limit in Einstein gravity ($\alpha\rightarrow
4714: 0$), we see that the metric functions must satisfy
4715: \begin{eqnarray}
4716: \dot{a}' = \frac{n'}{n}\dot{a} + \frac{\dot{b}}{b}a' \,. \label{keyrel}
4717: \end{eqnarray}
4718: Using this consequence of the $\{ty\}$-component, we can rewrite the
4719: remains components of $\tilde G_{AB}$ and $\tilde H_{AB}$ as
4720: \begin{eqnarray}
4721: \tilde G_{tt} = \frac{3n^2}{2a^3a'}\left(a^4\Phi\right)'\,,~~~~
4722: \tilde G_{yy} = - \frac{3b^2}{2a^3\dot{a}}\left(a^4\Phi\right)^{\displaystyle{\cdot}} \,,
4723: \end{eqnarray}
4724: \begin{eqnarray}
4725: \tilde G_{ij} = \frac{1}{2\dot{a}a'}h_{ij}\left\{\frac{\dot{b}}{b}\frac{a'}{\dot{a}}
4726: \left(a^4\Phi\right)^{\displaystyle{\cdot}} + \frac{n'}{n}\frac{\dot{a}}{a'}
4727: \left(a^4\Phi\right)' - \left(a^4\Phi\right)^{\displaystyle{\cdot}}{}' \right\} \,,
4728: \end{eqnarray}
4729: \begin{eqnarray}
4730: \tilde H_{tt} = \frac{3n^2}{2a^3a'}\left(a^4\Phi^2\right)'\,,~~~~
4731: \tilde H_{yy} = - \frac{3b^2}{2a^3\dot{a}}\left(a^4\Phi^2\right)^{\displaystyle{\cdot}} \,,
4732: \end{eqnarray}
4733: \begin{eqnarray}
4734: \tilde H_{ij} = \frac{1}{2\dot{a}a'}h_{ij}\left\{\frac{\dot{b}}{b}\frac{a'}{\dot{a}}
4735: \left(a^4\Phi^2\right)^{\displaystyle{\cdot}} + \frac{n'}{n}\frac{\dot{a}}{a'}
4736: \left(a^4\Phi^2\right)' - \left(a^4\Phi^2\right)^{\displaystyle{\cdot}}{}' \right\} \,.
4737: \end{eqnarray}
4738:
4739: Then the field equations~(\ref{fieldeq}) for the metric~(\ref{themetric})
4740: are equivalent to eq. (\ref{keyrel}) and
4741: \begin{eqnarray}
4742: \left[a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\right]' =
4743: \frac{1}{6}(a^4)'\rho \,, \label{ttcom}
4744: \end{eqnarray}
4745: \begin{eqnarray}
4746: \frac{\dot{b}}{b}\frac{a'}{\dot{a}}
4747: \left[a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\right]^{\displaystyle{\cdot}}
4748: + \frac{n'}{n}\frac{\dot{a}}{a'}
4749: \left[a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\right]'-\cr
4750: - {\left[a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\right]^{\displaystyle{\cdot}}{}}' =
4751: 2\dot{a}a'a^2p_L\,, \label{ijcom}
4752: \end{eqnarray}
4753: \begin{eqnarray}
4754: \left[a^4\left(\Phi+2\alpha\Phi^2+\frac{1}{l^2}\right)\right]^{\displaystyle{\cdot}} =
4755: -\frac{1}{6}(a^4)^{\displaystyle{\cdot}} p_T \,. \label{yycom}
4756: \end{eqnarray}
4757: Introducing (\ref{ttcom}) and (\ref{yycom}) into (\ref{ijcom}), we obtain
4758: the conservation equation~(\ref{5ceq}).
4759:
4760: \section{Bulk geometry in static coordinates} \label{B2}
4761:
4762: In this section we show how to derive the modified Friedmann equation for any gravitational theory\footnote{See also \cite{Barcelo-edge} for
4763: the closed Friedmann brane case ($k=+1$).}.
4764: A Friedmann brane at $y=0$ in a bulk metric
4765: \be \label{matric1}
4766: ds^2=-n^2(\tau,y)d\tau^2+a^2(\tau,y)d\vec x\cdot d\vec x+ b(\tau,y) dy^2
4767: \ee
4768: is locally equivalent to a Friedmann brane moving geodesically in a black-hole-type metric
4769: \be\label{metricBH}
4770: ds^2=-f(R)dT^2+R^2d\vec x\cdot d\vec x + \frac{dR^2}{f(R)}\ .
4771: \ee
4772: The form of $f(R)$ is determined by solving the field equations of the theory.
4773:
4774: In this picture we imagine the brane as a hypersurface that is moving towards or away
4775: from a black-hole with an expansion factor $a(\tau,y(\tau))=R(T(\tau))$. At
4776: each fixed radial distance from the black-hole, $dR=0=da$, one has
4777: \be
4778: da=\dot ad\tau+a'dy=0\ \Rightarrow\ dy^2=\left(\frac{\dot a}{a'}\right)^2 d\tau^2\ .
4779: \ee
4780: Substituting into (\ref{matric1}), we obtain
4781: \be
4782: ds^2=-\left(n^2-\frac{\dot a^2}{a'^2}b^2\right)d\tau^2\ .
4783: \ee
4784: We suppose that the brane is a hypersurface in the black-hole background in geodesic
4785: motion. Therefore its four-velocity in static coordinates is
4786: \be
4787: u_Adx^A=-\sqrt{f(R(T))+\dot R(T)^2}dT+\frac{\dot R(T)}{f(R(T))}dR=-d\tau\ ,
4788: \ee
4789: where $\tau$ is the proper time on the brane. For $dR=0$ we have
4790: \be\label{tau}
4791: -\sqrt{f(R)+\dot R^2}dT=d\tau\ .
4792: \ee
4793: Substituting into (\ref{metricBH}), we have
4794: \be\label{BH}
4795: ds^2=-\frac{f(R(T))}{f(R(T))+\dot R(T)^2}d\tau^2\ .
4796: \ee
4797: Equating now eq. (\ref{tau}) and (\ref{BH}), using $R(T(\tau))=a(\tau,y(\tau))$, we obtain
4798: \be
4799: n^2-\left(\frac{\dot a b}{a'}\right)^2=\frac{f(a)}{f(a)+\dot a^2}\ .
4800: \ee
4801: Since $\tau$ is the proper time, $n(\tau,0)=1$, and
4802: \be
4803: f(a)=\left(\frac{a'}{b}\right)^2-\dot a^2\ .
4804: \ee
4805: Therefore defining the Hubble rate $H=\dot a/a$ and the function $H_y=a'/ba$, we obtain
4806: \be
4807: \frac{f(a)}{a^2}=H_y^2-H^2\ ,
4808: \ee
4809: where $H_y$ must be found by junction conditions that are dependent on the gravity theory
4810: chosen. In particular, for the Randall-Sundrum-type model, we have
4811: \be
4812: f(a)=k-\frac{M}{a^2}\ ,
4813: \ee
4814: and we find \cite{Kr:99,Id:00}
4815: \be
4816: H^2+\frac{k}{a^2}=H_y^2+\frac{M}{a^4}\ ,
4817: \ee
4818: where $M/a^4$ is the dark radiation term. Instead for Lanczos gravity, we have
4819: \be
4820: f(a) = k +
4821: \frac{a^2}{2\alpha}\left(1-\sqrt{1+\frac{4\alpha M}
4822: {3a^4}+\frac{2}{3}\alpha\Lambda}\right)\ ,
4823: \ee
4824: leading to \cite{GS1}
4825: \be
4826: H^2+\frac{k}{a^2}=H_y^2+\frac{1}{2\alpha}\left(1-\sqrt{1+\frac{4\alpha M}
4827: {3a^4}+\frac{2}{3}\alpha\Lambda}\right)\ .
4828: \ee
4829: Local equivalence of the metrics (\ref{matric1}) and (\ref{metricBH}) has been proved in the
4830: Randall-Sundrum model \cite{BH} and in the Lanczos model \cite{Charmousis}.
4831:
4832: \newpage
4833: \begin{thebibliography}{99}
4834:
4835:
4836:
4837: \bibitem{AbMo:01} B. Abdesselam and N. Mohammedi, Phys. Rev. {\bf D65}, 084018 (2002).
4838:
4839: \bibitem{bhads} S.~O. Alexeyev, Gravitation \& Cosmology {\bf 3}, 161 (1997);
4840: A. Chakrabarti and D.~H. Tchrakian, Phys. Rev. {\bf D65},
4841: 024029 (2002).
4842:
4843: \bibitem{KK}
4844: T. Appelquist, A. Chodos and P. G. O. Freund, {\it Modern Kaluza-Klein theories}
4845: (Addison-Wesley, Reading, USA, 1987).
4846:
4847:
4848: \bibitem{ADM}
4849: R. Arnowitt, S. Deser and C. W. Misner, in {\it Gravitation: An Introduction to Current
4850: Research}, edited by L. Witten (Wiley, New York,
4851: 1962).
4852:
4853: \bibitem{BK}
4854: V. Balasubramanian and P. Kraus, Commun. Math. Phys. {\bf 208}, 413 (1999).
4855:
4856: \bibitem{barcelo}
4857: C. Barcel\'o, C. Germani and C. F. Sopuerta, Phys. Rev. {\bf D68}, 104007 (2003).
4858:
4859: \bibitem{Barcelo-edge}
4860: C.~Barcel\'o and M.~Visser,
4861: %``Living on the edge: Cosmology on the boundary of anti-de Sitter space,''
4862: Phys.\ Lett.\ B {\bf 482}, 183 (2000).
4863:
4864: \bibitem{bdl}
4865: P. Binetruy, C. Deffayet, U. Ellwanger and D. Langlois, Phys. Lett. {\bf B477}, 285 (2000).
4866:
4867: \bibitem{bida}
4868: N. D. Birrell and P. C. W. Davies, {\it Quantum fields in curved
4869: space} (Cambridge University Press, Cambridge, 1994).
4870:
4871: \bibitem{BH}
4872: P. Bowcock, C. Charmousis and R. Gregory, Class. Quant. Grav. {\bf 17}, 4745 (2000);
4873: S. Mukohyama, T. Shiromizu and K. Maeda,
4874: Phys. Rev. {\bf D62}, 024028 (2000); Erratum: ibid., {\bf D63}, 029901 (2001).
4875:
4876: \bibitem{Bousso}
4877: R. Bousso, JHEP {\bf 06}, 028 (1999).
4878:
4879: \bibitem{bcg}
4880: P.\ Bowcock, C.\ Charmousis, and R.\ Gregory, Class.\ Quantum
4881: Grav.\ {\bf 17}, 4745 (2000).
4882:
4883: \bibitem{holo1}
4884: J. D. Bekenstein, Nuovo Cim. Lett. {\bf 4}, 737 (1972).
4885:
4886: \bibitem{holo2}
4887: J. D. Bekenstein, Phys. Rev. {\bf D7}, 2333 (1973).
4888:
4889: \bibitem{holo3}
4890: J. D. Bekenstein, Phys. Rev. {\bf D9}, 3292 (1974).
4891:
4892: \bibitem{Bruck}
4893: P.~Brax and C.~van de Bruck,
4894: %``Cosmology and brane worlds: A review,''
4895: Class.\ Quant.\ Grav.\ {\bf 20}, R201 (2003).
4896:
4897: \bibitem{Brohe}
4898: J.D. Brown and M. Henneaux, Commum. Math. Phys. {\bf 104}, 207 (1986).
4899:
4900: \bibitem{BY}
4901: J. D. Brown and J. W. York, Jr., Phys. Rev. {\bf D47}, 1407 (1993).
4902:
4903: \bibitem{collapse}
4904: M. Bruni, C. Germani and R. Maartens, Phys. Rev. Lett. {\bf 87}, 231302 (2001).
4905:
4906: \bibitem{C1}
4907: C. G. Callan, D. Friedan, E. J. Martinec and M. J. Perry, Nucl. Phys. {\bf B262}, 593 (1985).
4908:
4909: \bibitem{CamSop}
4910: A.~Campos and C.~F.~Sopuerta,
4911: %``Evolution of cosmological models in the brane-world scenario,''
4912: Phys.\ Rev.\ {\bf D63}, 104012 (2001).
4913:
4914: \bibitem{245Duff}
4915: D. M. Capper and M. J. Duff, Nucl. Phys. {\bf B82}, 147 (1974); D. M. Capper, Nuovo Cim.
4916: {\bf A25}, 29 (1975).
4917:
4918: \bibitem{CLR}
4919: D. M. Capper, G. Leibbrand and M. Ramon Medrano, Phys. Rev. {\bf D8}, 4320 (1973).
4920:
4921: \bibitem{SWi}
4922: D. M. Capper and M. Ramon Medrano, Phys. Rev. {\bf D9}, 1641 (1974).
4923:
4924: \bibitem{Cha:89} A.~H.~Chamseddine, Phys. Lett. {\bf B233}, 291 (1989);
4925: F.~M\"uller-Hoissen, Nucl. Phys. {\bf B346}, 235 (1990).
4926:
4927: \bibitem{blackstring}
4928: A. Chamblin, S. W. Hawking and H. S. Reall, Phys. Rev. {\bf D61}, 065007 (2000).
4929:
4930: \bibitem{Charmousis}
4931: C.~Charmousis and J.~F.~Dufaux,
4932: %``General Gauss-Bonnet brane cosmology,''
4933: Class.\ Quant.\ Grav.\ {\bf 19}, 4671 (2002).
4934:
4935: \bibitem{ChNeWe:01} Y.~M. Cho and I.~P. Neupane, arXiv: hep-th/0112227 (2001);
4936: Y.~M. Cho, I.~P. Neupane, and P.~S. Wesson, Nucl. Phys. {\bf B621}, 388 (2002).
4937:
4938: \bibitem{chrifu}
4939: S. M. Christensen and S. A. Fulling, Phys. Rev. {\bf D15}, 2088
4940: (1977).
4941:
4942: \bibitem{chridu}
4943: S. M. Christensen and M. J. Duff, Nucl. Phys. {\bf B154}, 301
4944: (1977); see also M. J. Duff, Class. Quant. Grav. {\bf 11}, 1387 (1994) and references therein.
4945:
4946: \bibitem{Corradini}
4947: O.~Corradini and Z.~Kakushadze,
4948: %``Localized gravity and higher curvature terms,''
4949: Phys.\ Lett.\ {\bf B494}, 302 (2000).
4950:
4951: %\bibitem{distribution}
4952: %T. M. Cover and J. A. Thomas.
4953: %{\it Elements of Information Theory},
4954: %(John Wiley and Sons, 1991).
4955:
4956: \bibitem{Friedmann} C.~Cs$\acute{\mbox{a}}$ki, M.~Graesser, C.~Kolda
4957: and J.~Terning,
4958: Phys.\ Lett.\ {\bf B462}, 34 (1999);
4959: J.~M.~Cline, C.~Grojean and G.~Servant,
4960: %``Cosmological expansion in the presence of an extra dimension,''
4961: Phys.\ Rev.\ Lett.\/ {\bf 83}, 4245 (1999).
4962: %[arXiv:hep-th/];
4963:
4964: \bibitem{CGKT}
4965: C. Csaki, M. Graesser, C. Kolda and J. Terning, Phys. Lett. {\bf B462}, 34 (1999).
4966:
4967: \bibitem{C2}
4968: T. Curtright and C. Zachos, Phys. Rev. Lett. {\bf 53}, 1799 (1984).
4969:
4970: \bibitem{dmpr}
4971: N. K. Dadhich, R. Maartens, P. Papadopoulos and V. Rezania,
4972: Phys. Lett. {\bf B487}, 1 (2000).
4973:
4974: \bibitem{Davis}
4975: S.~C.~Davis,
4976: %``Generalised Israel junction conditions for a Gauss-Bonnet brane world,''
4977: Phys.\ Rev.\ {\bf D67}, 024030 (2003).
4978:
4979: \bibitem{DD}
4980: N. Deruelle and T. Dolezel, Phys.\ Rev.\ {\bf D64}, 103506 (2001).
4981:
4982: \bibitem{DeruDole}
4983: N. Deruelle and T. Dolezel, Phys. Rev. {\bf D62}, 103502 (2000).
4984:
4985: \bibitem{5}
4986: N. Deruelle, T. Dolezel and J. Katz, Phys. Rev. {\bf D63}, 083513 (2001).
4987:
4988: \bibitem{lovelock} N. Deruelle and L. Farina-Busto, Phys. Rev. {\bf D41}, 3696 (1990);
4989: G.~A.~M. Marugan, Phys. Rev. {\bf D46}, 4320 (1992);
4990: 4340 (1992).
4991:
4992: \bibitem{DG}
4993: N. Deruelle and C. Germani, arXiv: gr-qc/0306116.
4994:
4995: \bibitem{Deruelle-Madore}
4996: N.~Deruelle and J.~Madore,
4997: %``On the quasi-linearity of the Einstein- Gauss-Bonnet gravity field
4998: %equations,''
4999: arXiv: gr-qc/0305004.
5000:
5001: \bibitem{BdW}
5002: B. DeWitt, Phys. Rev. {\bf 160}, 113 (1967).
5003:
5004: \bibitem{DeWitt}
5005: B. S. DeWitt, Phys. Rep. {\bf 19C}, 295 (1975).
5006:
5007: \bibitem{DGP}
5008: G. Dvali, G. Gabadadze and M. Porrati, Phys. Lett. {\bf B485}, 208 (2000).
5009:
5010: \bibitem{MT1}
5011: M. J. Duff, Int. J. Mod. Phys. {\bf A11}, 5623 (1996).
5012:
5013: \bibitem{DL}
5014: M. J. Duff and J. T. Liu, Phys. Rev. Lett. {\bf 85}, 2052 (2000); Class. Quant. Grav. {\bf 18},
5015: 3207 (2001); see also
5016: M. J. Duff, Phys. Rev. {\bf D9}, 1837 (1974).
5017:
5018: \bibitem{duff}
5019: M. J. Duff, Nucl. Phys. {\bf B125}, 334 (1977).
5020:
5021: \bibitem{E1}
5022: A. Einstein, Preussiche Akademie der Wissenschaften (Berlin)
5023: Sitzungsberichte, pg. 688 (1916).
5024:
5025: \bibitem{Emparan}
5026: R. Emparan, A. Fabbri and N. Kaloper, JHEP {\bf 0208}, 043 (2002).
5027:
5028: \bibitem{F1}
5029: E. S. Fradkin and A. A. Tseytlin,
5030: %``Effective Field Theory From Quantized Strings,''
5031: Phys.\ Lett.\ {\bf B158}, 316 (1985).
5032:
5033: \bibitem{F2}
5034: D. Friedan, Phys. Rev. Lett. {\bf 45}, 1057 (1980); L. Alvarez-Gaum\'e and D. Z. Freedman,
5035: Phys. Lett. {\bf 94B}, 171 (1980);
5036: L. Alvarez-Gaum\'e, D. Z. Freedman and S. Mukhi, Ann. Phys. {\bf 134}, 85 (1981).
5037:
5038: \bibitem{V2}
5039: S. Fubini and G. Veneziano, Ann. Phys. {\bf 63}, 12 (1971).
5040:
5041: \bibitem{bunchNew}
5042: J. Garriga and T. Tanaka, Phys. Rev. Lett. {\bf 84}, 2778 (2000);
5043: C. Csaki, J. Erlich, T. J. Hollowood and Y. Shirman, Nucl. Phys. {\bf B581}, 309 (2000);
5044: S. B. Giddings, E. Katz and L. Randall, JHEP
5045: {\bf 3}, 023 (2000); C. Csaki, J. Erlich and T. J. Hollowood, Phys. Rev. Lett. {\bf 84},
5046: 5932 (2000); C. Csaki, J. Erlich and
5047: T. J. Hollowood,
5048: Phys. Lett. {\bf B481}, 107 (2000); I. Ya. Aref'eva, M. G. Ivanov,
5049: W. Muck, K. S. Viswanathan and I. V. Volovich, Nucl. Phys. {\bf B590}
5050: 273 (2000); Z. Kakushadze, Phys. Lett. {\bf B497}, 125 (2000).
5051:
5052: \bibitem{static}
5053: C. Germani and R. Maartens, Phys. Rev. {\bf D64}, 124010 (2001).
5054:
5055: \bibitem{GS1}
5056: C. Germani and C. F. Sopuerta, Phys. Rev. Lett. {\bf 88}, 231101 (2002).
5057:
5058: \bibitem{GS2}
5059: C. Germani and C. F. Sopuerta, Astrophys. Space Sci. {\bf 283}, 487 (2003).
5060:
5061: \bibitem{Germani}
5062: C.~Germani and C.~F.~Sopuerta,
5063: %``String inspired braneworld cosmology,''
5064: Phys.\ Rev.\ Lett.\ {\bf 88}, 231101 (2002); ibid.,
5065: %[arXiv:hep-th/0202060];
5066: %%CITATION = HEP-TH 0202060;%%
5067: %----------------------------------------------------------------
5068: %\cite{Germani:2002vs}
5069: %\bibitem{Germani}
5070: %C.~Germani and C.~F.~Sopuerta,
5071: %``Varying fundamental constants from a string-inspired brane world model,''
5072: Astrophys.\ Space Sci.\ {\bf 283}, 487 (2003);
5073: %[arXiv:hep-th/0210086].
5074: %%CITATION = HEP-TH 0210086;%%
5075: %----------------------------------------------------------------
5076: %----------------------------------------------------------------
5077: %\cite{Kim:2000pz}
5078: %\bibitem{Kim:2000pz}
5079: J.~E.~Kim, B.~Kyae and H.~M.~Lee,
5080: %``Various modified solutions of the Randall-Sundrum model with the Gauss-Bonnet interaction,''
5081: Nucl.\ Phys.\ {\bf B582}, 296 (2000),
5082: Erratum: ibid.,\ {\bf B591}, 587 (2000);
5083: %[arXiv:hep-th/0004005].
5084: %%CITATION = HEP-TH 0004005;%%
5085: %----------------------------------------------------------------
5086: %\cite{Abdesselam:2001ff}
5087: %\bibitem{Abdesselam:2001ff}
5088: B.~Abdesselam and N.~Mohammedi,
5089: %``Brane world cosmology with Gauss-Bonnet interaction,''
5090: Phys.\ Rev.\ {\bf D65}, 084018 (2002).
5091: %[arXiv:hep-th/0110143].
5092: %%CITATION = HEP-TH 0110143;%%
5093:
5094: \bibitem{GW}
5095: G. W. Gibbons and S. W. Hawking, Phys. Rev. {\bf D15}, 2752 (1977).
5096:
5097: \bibitem{Giovannini}
5098: M.~Giovannini,
5099: %``Thick branes and Gauss-Bonnet self-interactions,''
5100: Phys.\ Rev.\ {\bf D64}, 124004 (2001).
5101:
5102: \bibitem{Ginsparg2}
5103: P. Ginsparg and G. Moore, arXiv: hep-th/9304011.
5104:
5105: \bibitem{ginsparg}
5106: P. Ginsparg, arXiv: hep-th/9108028.
5107:
5108: \bibitem{Gravanis}
5109: E.~Gravanis and S.~Willison,
5110: %``Israel conditions for the Gauss-Bonnet theory and the Friedmann equation on the brane
5111: %universe,''
5112: Phys.\ Lett.\ {\bf B562}, 118 (2003).
5113:
5114: \bibitem{GSW}
5115: M. B. Green, J. H. Schwartz and E. Witten, {\it Superstring theory}, volumes I \& II,
5116: (Cambridge University Press, Cambridge, 1987).
5117:
5118: \bibitem{Gregory-Padilla}
5119: J.~P.~Gregory and A.~Padilla,
5120: %``Braneworld holography in Gauss-Bonnet gravity,''
5121: arXiv: hep-th/0304250.
5122:
5123: \bibitem{GRS}
5124: R. Gregory, V. A. Rubakov and S. M. Sibiryakov, Phys. Rev. Lett. {\bf 84}, 5928 (2000).
5125:
5126: \bibitem{G1}
5127: D. Gross, J. Harvey, E. Martinec and R. Rohn, Phys. Rev. Lett. {\bf 54}, 502 (1985).
5128:
5129: \bibitem{GBE}
5130: D. J. Gross and J. H. Sloan, Nucl. Phys. {\bf B291}, 41 (1987).
5131:
5132: \bibitem{H2}
5133: S. W. Hawking, Nature {\bf 248}, 30 (1974); Comm. Math. Phys. {\bf 43}, 199 (1975).
5134:
5135: \bibitem{predict}
5136: S. W. Hawking, Phys. Rev. {\bf D14}, 2460 (1976).
5137:
5138: \bibitem {H1}
5139: W. Heisenberg, Z. Physics {\bf 110}, 251 (1938).
5140:
5141: \bibitem{tH1}
5142: G. t'Hooft, Nucl. Phys. {\bf B62}, 444 (1973);
5143: G. t'Hooft and M. Veltman, Ann. Inst. H. Poincar\'e, {\bf 20}, 69 (1974);
5144: S. Deser and P. V. Nieuwenhuizen, Phys. Rev. {\bf D10}, 401 (1974); ibid.,
5145: Phys. Rev. {\bf D10}, 411 (1974).
5146:
5147: \bibitem{tH2}
5148: G. t'Hooft, arXiv: gr-qc/9310026.
5149:
5150: \bibitem{HW}
5151: P. Horava and E. Witten, Nucl. Phys. {\bf B460}, 506 (1996).
5152:
5153: \bibitem{positive}
5154: G. T. Horowitz and R. C. Myers, Phys. Rev. {\bf D59}, 026005 (1999).
5155:
5156: \bibitem{Id:00} D. Ida, JHEP {\bf 09}, 14 (2000).
5157:
5158: \bibitem{israel}
5159: W. Israel, Nuovo Cim. {\bf 44B}, 1 (1966).
5160:
5161: \bibitem{IZ}
5162: C. Itzykson and J.-B. Zuber, {\it Quantum Field Theory} (McGraw-Hill, Singapore, 1980).
5163:
5164: \bibitem{others}
5165: J.~E.~Kim, B.~Kyae and H.~M.~Lee, Phys.\ Rev.\ {\bf D62}, 045013 (2000);
5166: N.~E. Mavromatos and J. Rizos, Phys. Rev. {\bf D62}, 124004 (2000);
5167: I. Low and A. Zee, Nucl. Phys. {\bf B585}, 395 (2000);
5168: S. Nojiri and S.~D. Odintsov, Phys. Lett. {\bf B494}, 135 (2000);
5169: S. Nojiri, S.~D. Odintsov, and S. Ogushi, Phys. Rev. {\bf D65}, 023521
5170: (2002);
5171: M. Giovannini, Phys. Rev. {\bf D64}, 124004 (2001);
5172: I.~P. Neupane, Class. Quant. Grav. {\bf 19}, 1167 (2002).
5173:
5174: \bibitem{exact}
5175: D. Kramer, H. Stephani, M. MacCallum and E. Herlt, {\em Exact
5176: Solutions of Einstein's Field Equations} (Cambridge University Press, Cambridge, 1980).
5177:
5178: \bibitem{Kr:99} P. Kraus, JHEP {\bf 12}, 011 (1999).
5179:
5180: \bibitem{kre}
5181: E. Kreysig, {\it Introduction to differential geometry and
5182: Riemanian geometry} (University of Toronto Press, Toronto, 1968).
5183:
5184: \bibitem{Lanczos}
5185: C. Lanczos,
5186: Z. Phys. {\bf 73} 147 (1932); ibid., Ann. Math. {\bf 39}, 842 (1938).
5187:
5188: \bibitem{Lidsey}
5189: J.~E.~Lidsey,
5190: %``Inflation and Braneworlds,''
5191: arXiv: astro-ph/0305528.
5192: %%CITATION = ASTRO-PH 0305528;%%
5193: %----------------------------------------------------------------
5194: %\cite{Lidsey:2003sj}
5195: \bibitem{Lidsey2}
5196: J.~E.~Lidsey and N.~J.~Nunes,
5197: %``Inflation in Gauss-Bonnet brane cosmology,''
5198: Phys.\ Rev.\ {\bf D67}, 103510 (2003).
5199: %[arXiv:astro-ph/0303168].
5200:
5201: \bibitem{Lov}
5202: D. Lovelock, J. Math. Phys. {\bf 12}, 498 (1971).
5203:
5204: \bibitem{Low:2000pq}
5205:
5206: I.~Low and A.~Zee,
5207: %``Naked singularity and Gauss-Bonnet term in brane world scenarios,''
5208: Nucl.\ Phys.\ {\bf B585}, 395 (2000);
5209: N.~E.~Mavromatos and J.~Rizos,
5210: %``String inspired higher-curvature terms and the Randall-Sundrum scenario,''
5211: Phys.\ Rev.\ {\bf D62}, 124004 (2000);
5212: N.~E.~Mavromatos and J.~Rizos,
5213: %``Exact solutions and the cosmological constant problem in dilatonic domain wall
5214: %higher-curvature string gravity,''
5215: Int.\ J.\ Mod.\ Phys.\ {\bf A18}, 57 (2003);
5216: I.~P.~Neupane,
5217: %``Consistency of higher derivative gravity in the brane background,''
5218: JHEP {\bf 0009}, 040 (2000);
5219: P.~Binetruy, C.~Charmousis, S.~C.~Davis and J.~F.~Dufaux,
5220: %``Avoidance of naked singularities in dilatonic brane world scenarios with a
5221: %Gauss-Bonnet term,''
5222: Phys.\ Lett.\ {\bf B544}, 183 (2002);
5223: C.~Charmousis and J.~F.~Dufaux,
5224: %``General Gauss-Bonnet brane cosmology,''
5225: Class.\ Quant.\ Grav.\ {\bf 19}, 4671 (2002);
5226: K.~A. Meissner and M. Olechowski, Phys. Rev. Lett. {\bf 86}, 3708 (2001).
5227:
5228: \bibitem{m}
5229: R. Maartens, Phys. Rev. {\bf D62}, 084023 (2000).
5230:
5231: \bibitem{mwbh}
5232: R. Maartens, D. Wands, B.A. Bassett and I. P. C. Heard, Phys. Rev.
5233: {\bf D62}, 041301 (2000).
5234:
5235: \bibitem{maacov}
5236: R. Maartens, Prog. Theor. Phys. Suppl. {\bf 148}, 213 (2003).
5237:
5238: \bibitem{M1}
5239: J. Maldacena, Adv. Theor. Math. Phys. {\bf 2}, 231 (1998); Int. J. Theor. Phys.
5240: {\bf 38}, 113 (1999).
5241:
5242: \bibitem{MeOl:00} K.~A. Meissner and M. Olechowski, Phys. Rev. Lett. {\bf 86}, 3708 (2001).
5243:
5244: \bibitem{local} K.~A. Meissner and M. Olechowski, Phys. Rev. {\bf D65}, 064017 (2002);
5245: J.~E.~Kim and H.~M.~Lee, Nucl. Phys. {\bf B602}, 346 (2001);
5246: ibid., {\bf B619} 763 (2001).
5247:
5248: \bibitem{M3}
5249: R. R. Metsaev and A. A. Tseytlin, Nucl. Phys. {\bf B293}, 385 (1987).
5250:
5251: \bibitem{gravitation}
5252: C. W. Misner, K. S. Thorne and J. A. Wheeler, {\it Gravitation} (W. H. Freeman,
5253: New
5254: York, 1998).
5255:
5256: \bibitem{Mounaix}
5257: P.~Mounaix and D.~Langlois,
5258: %``Cosmological equations for a thick brane,''
5259: Phys.\ Rev.\ {\bf D65}, 103523 (2002).
5260:
5261: \bibitem{msm}
5262: S.\ Mukohyama, T.\ Shiromizu and K.\ I.\ Maeda, Phys. Rev. {\bf
5263: D62}, 024028 (2000).
5264:
5265: \bibitem{M2}
5266: C. Myers, Phys. Lett. {\bf B199}, 371 (1987).
5267:
5268: \bibitem{Myers}
5269: R.~C.~Myers,
5270: %``Higher Derivative Gravity, Surface Terms And String Theory,''
5271: Phys.\ Rev.\ {\bf D36}, 392 (1987).
5272:
5273: \bibitem{Ne:01a} I.~P. Neupane, Phys. Lett. {\bf B512}, 137 (2001).
5274:
5275: \bibitem{Ca:01} S. Nojiri and S.~D. Odintsov, Phys. Lett. {\bf B521}, 87 (2001); R.-G.
5276: Cai, Phys. Rev. {\bf D65},
5277: 084014 (2002).
5278:
5279: \bibitem{pet}
5280: A. Z. Petrov, {\it Einstein spaces} (Pergamon, Oxford, 1969).
5281:
5282: \bibitem{padilla}
5283: A. Padilla, Ph.D. thesis (Durham University, 2002), arXiv: hep-th/0210217.
5284:
5285:
5286: \bibitem{Padilla}
5287: A.~Padilla,
5288: %``Surface terms and the Gauss-Bonnet Hamiltonian,''
5289: arXiv: gr-qc/0303082.
5290: %%CITATION = GR-QC 0303082;%%
5291:
5292: \bibitem{P1}
5293: A.M. Polyakov, Phys. Lett. {\bf 103B}, 207 (1981); ibid., Phys. Lett. {\bf 103B}, 211 (1981).
5294:
5295: \bibitem{PO}
5296: J. Polchinski, {\it String Theory} volumes I \& II, (Cambridge University Press,
5297: Cambridge, 1998).
5298:
5299: \bibitem{tesi}
5300: M.\ M.\ Prats, PhD thesis (University of Barcelona, 1995).
5301:
5302: \bibitem{RS}
5303: L. Randall and R. Sundrum, Phys. Rev. Lett. {\bf 83}, 3370 (1999).
5304:
5305: \bibitem{R1}
5306: C. Rovelli, arXiv: gr-qc/0006061.
5307:
5308: \bibitem{SV}
5309: I. Savonie and E. Verlinde, Phys. Lett. {\bf B507}, 305 (2001).
5310:
5311: \bibitem{ssm}
5312: M. Sasaki, T. Shiromizu and K. Maeda, Phys. Rev. {\bf D62},
5313: 024008 (2000).
5314:
5315: \bibitem{schwinger}
5316: J. Schwinger, Proc. Nat. Acad. Sci. (USA) {\bf 37}, 452 (1951).
5317:
5318: \bibitem{SMS}
5319: T. Shiromizu, K. Maeda and M. Sasaki, Phys. Rev. {\bf D62}, 024012 (2000).
5320:
5321: \bibitem{MT3}
5322: J.H. Schwarz, Nucl. Phys. Proc. Suppl. {\bf B55}, 1 (1997).
5323:
5324: \bibitem{MT4}
5325: A. Sen, arXiv: hep-th/9802051.
5326:
5327: \bibitem{Synge}
5328: J. L. Synge, {\it Relativity: the general theory} (North Holland, Amsterdam,
5329: 1971).
5330:
5331: \bibitem{s}
5332: H.\ Stephani, {\it General Relativity} (Cambridge University
5333: Press, Cambridge, 1990).
5334:
5335: \bibitem{S1}
5336: L. Susskind, J. Math. Phys. {\bf 36}, 6377 (1995).
5337:
5338: \bibitem{hawholo}
5339: T. Tanaka, Prog. Theor. Phys. Suppl. {\bf 148}, 307 (2003);
5340: R. Emparan, A. Fabbri and N. Kaloper, JHEP {\bf 0208}, 043 (2002);
5341: R. Casadio, arXiv: hep-th/0302171.
5342:
5343: \bibitem{MT2}
5344: P. K. Townsend, arXiv: hep-th/9612121.
5345:
5346: \bibitem{U1}
5347: W. G. Unruh, Phys. Rev. {\bf D14}, 870 (1976).
5348:
5349: \bibitem{V1}
5350: G. Veneziano, Nuovo Cimento {\bf 57A}, 190 (1968).
5351:
5352: \bibitem{Visser}
5353: M. Visser, Phys. Lett. {\bf B159}, 22 (1985).
5354:
5355: \bibitem{Wald}
5356: R. M. Wald, {\it General Relativity} (University of Chicago Press, Chicago, 1984).
5357:
5358: \bibitem{wiseman}
5359: T. Wiseman, Phys. Rev. {\bf D65}, 124007 (2002).
5360:
5361: \bibitem{W1}
5362: E. Witten, Comm. Math. Phys. {\bf 92}, 455 (1984).
5363:
5364: \bibitem{W2}
5365: E. Witten, Adv. Theor. Math. Phys. {\bf 2}, 253 (1998).
5366:
5367: \bibitem{SU}
5368: E. Witten, Nucl. Phys. {\bf B443}, 85 (1995).
5369:
5370: \bibitem{York}
5371: J. W. York, Found. Phys. {\bf 16}, 249 (1986).
5372:
5373: \bibitem{BoDe:85} B.~Zwiebach, Phys. Lett. {\bf B156}, 315 (1985);
5374: D.~G. Boulware and S. Deser, Phys. Rev. Lett. {\bf55}, 2656 (1985).
5375:
5376:
5377:
5378: %----------------------------------------------------------------
5379: \end{thebibliography}
5380: %----------------------------------------------------------------
5381:
5382:
5383:
5384:
5385: %----------------------------------------------------------------
5386: \end{document}
5387: %----------------------------------------------------------------
5388:
5389:
5390:
5391:
5392: \end{document}
5393: