1: \documentclass[nofootinbib,floatfix,prd,preprintnumbers,superscriptaddress]{revtex4}
2: %\documentclass{iopart}
3: \usepackage{graphicx}
4: \usepackage{amsfonts}
5: \usepackage{amssymb}
6: \usepackage{amsbsy}
7: \usepackage{amsmath}
8: \usepackage{latexsym}
9: %\usepackage{subfloat}
10: \usepackage{bm}
11: %\usepackage{subfigure}
12: \usepackage{hyperref}
13: %\usepackage[margin=0.75in,a4paper]{geometry}
14:
15: \newcommand*{\K}{\mathcal{K}}
16: \setlength\arraycolsep{2pt}
17: \newcommand*{\di}{\partial}
18: \newcommand*{\rh}{r_\text{h}}
19: \newcommand*{\CHECK}{\,[\text{\textsc{check}}]}
20: \newcommand*{\rb}{r_\text{b}}
21: \newcommand*{\thatb}{\hat{t}_\text{b}}
22: \newcommand*{\rhatb}{\hat{r}_\text{b}}
23: \newcommand*{\zhatb}{\hat{z}_\text{b}}
24: \newcommand*{\rhohat}{\hat{\rho}}
25: \newcommand*{\fb}{f_\text{b}}
26: \newcommand*{\nuc}{\nu_\text{c}}
27: \newcommand*{\ac}{a_\text{c}}
28: \newcommand*{\nuh}{\nu_\text{h}}
29: \newcommand*{\zb}{z_\text{b}}
30: \newcommand*{\tb}{t_\text{b}}
31: \newcommand*{\vb}{v_\text{b}}
32: \newcommand*{\obs}{\text{obs}}
33: \newcommand*{\z}{\hat{z}}
34: \renewcommand*{\c}{\text{c}}
35: \newcommand*{\f}{\hat{f}}
36: \newcommand*{\GR}{\text{\tiny{GR}}}
37: \newcommand*{\KK}{\mathcal{E}}
38: \newcommand*{\Weyl}{\mathcal{E}}
39: \newcommand*{\KIS}{\text{\tiny{KIS}}}
40: \newcommand*{\p}{\hat{p}}
41: \newcommand*{\etahat}{\hat{\eta}}
42: \newcommand*{\chihat}{\hat{\chi}}
43: \renewcommand*{\t}{\hat{t}}
44: \renewcommand*{\a}{\hat{a}}
45: \renewcommand*{\b}{\text{b}}
46: \renewcommand*{\k}{\hat{k}}
47: \newcommand*{\n}{\hat{n}}
48: \renewcommand*{\r}{\hat{r}}
49: \newcommand*{\BBN}{\rho_\text{BBN}}
50: \renewcommand*{\H}{\hat{H}}
51: \newcommand*{\DOT}[2]{{\bm{{#1} \cdot {#2}}}}
52: \newcommand*{\approaches}[2]{\xrightarrow[#2]{\,\,\,{#1}\,\,\,}}
53: \newcommand*{\PS}{\text{\tiny{PS}}}
54: \newcommand*{\CI}{\text{\tiny{CI}}}
55: \newcommand*{\eff}{\text{\tiny{eff}}}
56: \newcommand*{\fiveD}{\text{\tiny{5D}}}
57: \newcommand*{\prim}{\text{\tiny{inf}}}
58:
59: \begin{document}
60:
61: \preprint{arXiv:0705.1685 [astro-ph]}
62:
63: \title{Scalar perturbations in braneworld cosmology}
64:
65: \author{Antonio Cardoso}%
66: \email{antonio.cardoso.AT.port.ac.uk}%
67: \affiliation{Institute of Cosmology \& Gravitation, University of
68: Portsmouth, Portsmouth~PO1~2EG, UK}
69:
70: \author{Takashi Hiramatsu}
71: \email{hiramatsu.AT.resceu.s.u-tokyo.ac.jp} \affiliation{Research
72: Center for the Early Universe (RESCEU), School of Science,
73: University of Tokyo, 7-3-1 Hongo, Bunkyo, Tokyo 113-0033, Japan}
74:
75: \author{Kazuya Koyama}
76: \email{kazuya.koyama.AT.port.ac.uk}%
77: \affiliation{Institute of Cosmology \& Gravitation, University of
78: Portsmouth, Portsmouth~PO1~2EG, UK}
79:
80: \author{Sanjeev S.~Seahra}
81: \email{sanjeev.seahra.AT.port.ac.uk} %
82: \affiliation{Institute of Cosmology \& Gravitation, University of
83: Portsmouth, Portsmouth~PO1~2EG, UK}
84:
85: \date{May 11, 2007}
86:
87: \begin{abstract}
88:
89: We study the behaviour of scalar perturbations in the
90: radiation-dominated era of Randall-Sundrum braneworld cosmology by
91: numerically solving the coupled bulk and brane master wave
92: equations. We find that density perturbations with wavelengths less
93: than a critical value (set by the bulk curvature length) are
94: amplified during horizon re-entry. This means that the radiation era
95: matter power spectrum will be at least an order of magnitude larger
96: than the predictions of general relativity (GR) on small scales.
97: Conversely, we explicitly confirm from simulations that the spectrum
98: is identical to GR on large scales.
99: %The comoving value of the critical scale is less than 10 AU for
100: % reasonable parameters,
101: Although this magnification is not relevant for the cosmic microwave
102: background or measurements of large scale structure, it
103: will have some bearing on the formation of primordial black holes
104: in Randall-Sundrum models.
105:
106: \end{abstract}
107:
108: \maketitle
109:
110: \section{Introduction}\label{sec:introduction}
111:
112: Our view of cosmology has been revolutionized by the notion that the
113: universe may be a lower-dimensional object embedded in some
114: higher-dimensional space. A particularly simple realization of this
115: idea is furnished by the Randall-Sundrum (RS) braneworld model
116: \cite{Randall:1999vf}, which postulates that our observable universe
117: is a thin 4-dimensional hypersurface residing in 5-dimensional
118: anti-de Sitter (AdS) space (see
119: Refs.~\cite{Langlois:2002bb,Maartens:2003tw,Brax:2004xh} for
120: reviews). Ordinary matter degrees of freedom are assumed to be
121: confined to the brane, while gravitational degrees of freedom are
122: allowed to propagate in the full 5-dimensional bulk. The warping of
123: AdS space allows us to recover ordinary general relativity (GR) at
124: distances greater than the curvature radius of the bulk $\ell$.
125: Current laboratory tests of Newton's law constrain $\ell$ to be less
126: than around $0.1\,\text{mm}$ \cite{Kapner:2006si}.
127:
128: The cosmological implications of the Randall-Sundrum scenario have
129: been extensively studied. It is well known that the Friedmann
130: equation governing the expansion of the brane universe differs from
131: general relativity by a correction of order $\rho/\sigma$, where
132: $\rho$ is the density of brane matter and $\sigma \gtrsim
133: (\text{TeV})^4$ is the brane tension. The magnitude of this
134: correction defines the ``high-energy'' regime of braneworld
135: cosmology as the era when $\rho \gtrsim \sigma$ or equivalently
136: $H\ell \gtrsim 1$, where $H$ is the Hubble parameter. At high
137: energies, the RS Friedmann equation implies $H \propto \rho$, which
138: results in dynamics significantly different from the standard $H^2
139: \propto \rho$ expansion law.
140:
141: The major outstanding issue in RS cosmology is the behaviour of
142: cosmological perturbations \cite{Mukohyama:2000ui,Kodama:2000fa,
143: vandeBruck:2000ju, Koyama:2000cc, Langlois:2000ph}. The equations of
144: motion governing fluctuations of the model are found to differ from
145: GR in two principal ways at early times: First, they acquire
146: $\mathcal{O}(\rho/\sigma)$ high-energy corrections similar to those
147: found in the Friedmann equation. By themselves, such corrections
148: are not difficult to deal with: they just modify the second-order
149: ordinary differential equations (ODEs) governing perturbations in
150: GR. But the second type of modification is more problematic:
151: Perturbations on the brane are also coupled to fluctuations of the
152: 5-dimensional bulk geometry, which are collectively known as the
153: ``Kaluza Klein'' (KK) degrees of freedom of the model. The KK modes
154: are governed by master partial differential equations (PDEs) defined
155: throughout the AdS bulk \cite{Mukohyama:2000ui,Kodama:2000fa}. In
156: the original work of Randall \& Sundrum \cite{Randall:1999vf}, the
157: brane had a simple Minkowski geometry and the KK mode master wave
158: equations were solvable via separation of variables. But the motion
159: of the brane in the cosmological case breaks the time-translation
160: symmetry of the bulk, which makes a simple separable solution
161: unattainable in most cases. The exception is the case of a brane
162: undergoing de Sitter inflation, where many analytic and
163: semi-analytic results are now available
164: \cite{Koyama:2004ap,Koyama:2005ek,Hiramatsu:2006cv,deRham:2007db,Koyama:2007as}.
165:
166: Many authors have considered various schemes to solve the
167: perturbation problem without dealing with the KK master PDEs
168: directly. The most straightforward approach is to simply set the
169: KK degrees of freedom to zero while retaining
170: $\mathcal{O}(\rho/\sigma)$ corrections, which we refer to as the
171: ``4-dimensional effective theory''. There are several
172: approximation methods that attempt to move beyond this simplest
173: effective theory in the case of tensor and scalar perturbations
174: \cite{Gordon:2000pt,Langlois:2000iu,Brax:2001qd,Bridgman:2001mc,
175: Koyama:2001ct,Leong:2001qm,%
176: Leong:2002hs,Easther:2003re,Battye:2003ks,Kobayashi:2004wy,Battye:2004qw,%
177: Koyama:2004cf}. These generally consider the behaviour of
178: fluctuations during certain limiting regimes (i.e, high-energy or
179: low-energy), but the only known way of tackling the problem on all
180: scales simultaneously is by direct numerical solution of the
181: equations of motion.
182:
183: The problem of tensor perturbations is somewhat simpler than the
184: scalar case. The reason is that in linear theory, tensor modes
185: are pure gravitational degrees of freedom that do not couple to
186: matter fluctuations (provided that the matter anisotopic stress is
187: neglected). In the braneworld context, this means that there are
188: no brane-confined tensor degrees of freedom. The perturbation
189: problem formally reduces to solving a wave equation in the bulk
190: with boundary conditions imposed on a moving brane. Numeric
191: solutions
192: have been obtained by several groups %
193: \cite{Hiramatsu:2003iz,Hiramatsu:2004aa,Hiramatsu:2006bd,Kobayashi:2005dd,Kobayashi:2005jx,%
194: Kobayashi:2006pe,Ichiki:2003hf,Seahra:2006tm}. Interestingly, the
195: effective theory predicts a blue spectrum (i.e.~excess power) in
196: the stochastic gravitational wave background at frequencies
197: corresponding to modes that enter the Hubble horizon in the
198: high-energy regime. However, once KK effects are incorporated via
199: numeric simulations, one finds that the true spectrum is flat and
200: virtually identical to GR. That is, the magnitudes of the
201: $\mathcal{O}(\rho/\sigma)$ and KK effects are nearly the same, but
202: they act in opposite sense and end up canceling each other out.
203: Physically, the reason that the KK modes cause a suppression of
204: the amplitude at high frequencies is that they efficiently radiate
205: gravitational wave energy from the brane into the bulk.
206:
207: The purpose of the paper is to numerically solve for the behaviour
208: of scalar perturbations in the radiation-dominated regime of
209: braneworld cosmology. (Numeric analysis of the scalar perturbations
210: during inflation has already been done in
211: Refs.~\cite{Hiramatsu:2006cv,Koyama:2007as}.) Unlike the tensor
212: case, there are several scalar degrees of freedom residing on the
213: brane, such as the density contrast $\delta\rho/\rho$. The problem
214: reduces to the solution of a bulk wave equation coupled to a master
215: boundary field satisfying its own second order ODE on the moving
216: brane. Recently, two of the numerical codes developed to deal with
217: tensor perturbations \cite{Hiramatsu:2003iz,Seahra:2006tm} have been
218: generalized to handle boundary degrees of freedom
219: \cite{Hiramatsu:2006cv,Cardoso:2006nh}. We use both codes in this
220: paper, which gives us the ability to confirm the consistency of our
221: numeric results via two independent algorithms. We are ultimately
222: interested in finding the matter transfer function in the radiation
223: era, and also determining the relative influence of KK and
224: high-energy effects on the density perturbations. Heuristically, we
225: may expect the KK modes to amplify high-energy/small-scale density
226: perturbations, which is the opposite effect from the tensor case.
227: The reason is that we know that the gravitational force of
228: attraction in the RS model is stronger than in GR on scales less
229: than $\ell$ \cite{Randall:1999vf,Garriga:1999yh}. This implies
230: that modes with a physical wavelength smaller than $\ell$ during
231: horizon crossing will be amplified due to the KK enhancement of the
232: gravitational force. However, this physical reasoning needs to be
233: tested with numeric simulations.
234:
235: The layout of the paper is as follows: The background geometry of
236: the RS cosmology we consider is described in \S\ref{sec:background}.
237: The formulae governing the gauge invariant scalar perturbations are
238: given in \S\ref{sec:scalar perturbations}. Also in that section, we
239: derive new analytic approximations for various perturbation
240: variables at very high energies. In \S\ref{sec:numeric}, we present
241: the numeric algorithms that we use to solve the coupled system of
242: bulk and brane wave equations. Our results for the matter transfer
243: function are also contained in that section. Finally, we summarize
244: and discuss the implications of our work in \S\ref{sec:discussion}
245:
246: \section{Background geometry and brane dynamics}\label{sec:background}
247:
248: As discussed in the Introduction, the Randall-Sundrum model we will
249: be considering consists of a thin brane, which is realized as
250: singular 4-surface, residing in 5-dimensional anti-de Sitter space.
251: In Poincar\'e coordinates, the bulk metric is given by the line
252: element
253: \begin{equation}\label{eq:5D metric 1}
254: ds_5^2 = \frac{\ell^2}{z^2} (-d\tau^2 + \delta_{ij} dx^i dx^j +dz^2).
255: \end{equation}
256: Here, $\ell$ is the curvature scale of the bulk. The boundary
257: of this spacetime given by a brane is described parametrically by
258: \begin{equation}
259: \tau = \tau_\b(\eta), \quad z = z_\b(\eta).
260: \end{equation}
261: The parameter $\eta$ is defined to be the conformal time along the
262: brane, which leads to the following induced line element
263: \begin{equation}
264: d\eta^2 = d\tau_\b^2 - dz_\b^2, \quad ds_\b^2 = a^2(-d\eta^2 +
265: d\mathbf{x}^2) = -dt^2 + a^2 d\mathbf{x}^2.
266: \end{equation}
267: Here, we have identified the scale factor as $a(\eta) =
268: \ell/z_\b(\eta)$ and $t$ as the cosmic time. Now, we define $u^a =
269: dx^a_\b/dt$ to be the 5-velocity field of comoving observers on the
270: brane, while $n^a$ is the brane normal. These vector fields can then
271: be used to construct directional derivatives parallel and orthogonal
272: to the brane:
273: \begin{subequations}
274: \begin{align}
275: \di_u & = u^a \di_a = \frac{1}{a} \left(\sqrt{1+H^2\ell^2}
276: \frac{\di}{\di \tau} -
277: H\ell \frac{\di}{\di z} \right)= \frac{\di}{\di t} = \frac{1}{a}
278: \frac{\di}{\di\eta}, \\ \di_n & = n^a
279: \di_a = \frac{1}{a} \left( -H\ell \frac{\di}{\di \tau} +
280: \sqrt{1+H^2\ell^2} \frac{\di}{\di z} \right),
281: \end{align}
282: \end{subequations}
283: respectively. Here, $H$ is the usual Hubble parameter
284: \begin{equation}
285: H = \left( \frac{\di_u r}{r} \right)_\b = \frac{1}{a} \frac{da}{dt},
286: \quad \frac{r}{\ell} = \frac{\ell}{z},
287: \end{equation}
288: where we have introduced the alternative bulk coordinate $r$. We
289: assume that the matter on the brane has a fluid-like stress-energy
290: tensor with density $\rho$ and pressure $p$:
291: \begin{equation}
292: T_{ab} = (\rho+p) u_a u_b + p h_{ab}, \quad g_{ab} = h_{ab} + n_a
293: n_b.
294: \end{equation}
295: The junction conditions are given in terms of the extrinsic
296: curvature $K_{ab}$ of the brane,
297: \begin{equation}\label{eq:extrinsic}
298: K_{ab} = h_a{}^c \nabla_c n_b, \qquad 2K_{ab} = -\kappa_5^2
299: [T_{ab} - \tfrac{1}{3} h_{ab} (T-\sigma) ],
300: \end{equation}
301: where $\kappa_5^2 = 1/M_5^3$ is the gravity-matter coupling in 5
302: dimensions and $\sigma$ is the brane tension. This yields
303: \begin{equation}
304: \left( \frac{\di_n r}{r} \right)_\b =
305: -\frac{\kappa_5^2(\rho+\sigma)}{6} =
306: -\frac{\sqrt{1+H^2\ell^2}}{\ell}.
307: \end{equation}
308: We assume the standard Randall-Sundrum fine-tuning
309: \begin{equation}
310: \kappa_5^2 = \kappa_4^2 \ell = 8\pi G \ell = \frac{6}{\sigma\ell},
311: \end{equation}
312: to obtain the Friedmann equation
313: \begin{equation}\label{eq:Friedmann}
314: H^2 = \frac{1}{\ell^2} \frac{\rho}{\sigma} \left(
315: 2 + \frac{\rho}{\sigma} \right) = \frac{8\pi G}{3} \rho \left(
316: 1 + \frac{\rho}{2\sigma} \right).
317: \end{equation}
318: Note the high-energy $\mathcal{O}(\rho/\sigma)$ correction to the
319: expansion rate. Finally, note that we have the usual conservation
320: of stress-energy on the brane; i.e.,
321: \begin{equation}\label{eq:conservation}
322: h^{ab} \nabla_a T_{bc} = 0, \qquad \frac{d\rho}{dt} = -3(1+w)\rho H, \qquad w =
323: \frac{p}{\rho}.
324: \end{equation}
325:
326: \section{Scalar perturbations}\label{sec:scalar perturbations}
327:
328: \subsection{Gauge invariant bulk perturbations}\label{sec:bulk
329: perturbations}
330:
331: We now turn our attention to the perturbation of the cosmology
332: introduced in \S\ref{sec:background}. It has been shown in
333: Refs.~\cite{Mukohyama:2000ui,Kodama:2000fa} that scalar-type
334: perturbations of the bulk geometry (\ref{eq:5D metric 1}) are
335: governed by a single gauge invariant master variable $\Omega$. We
336: summarize those results in this subsection.
337:
338: First, we introduce a harmonic basis with mode functions\footnote{To
339: describe the perturbations in the RS model we mostly follow the
340: notation and conventions of \citet{Kodama:2000fa}, which are based
341: on the 4-dimensional formalism of \citet{Kodama:1985bj}.}
342: \begin{equation}
343: Y = e^{i\mathbf{k} \cdot \mathbf{x}}, \qquad Y_i = -\frac{1}{k}
344: \di_i Y, \qquad Y_{ij} = \frac{1}{k^2} \di_i \di_j Y +
345: \frac{1}{3} \delta_{ij} Y.
346: \end{equation}
347: Using these mode functions, we write the perturbed bulk metric as
348: \begin{equation}\label{eq:pertubed bulk metric}
349: ds_5^2 = (q_{\alpha\beta} + f_{\alpha\beta})dx^\alpha dx^\beta +
350: \frac{2r}{\ell} f_\alpha Y_i
351: dx^\alpha dx^i + \frac{r^2}{\ell^2} \left[ (1+2H_L)\delta_{ij} +
352: 2H_T Y_{ij} \right] dx^i dx^j,
353: \end{equation}
354: where
355: \begin{equation}
356: q_{\alpha\beta} dx^\alpha dx^\beta = \frac{\ell^2}{z^2} (-d\tau^2 + dz^2).
357: \end{equation}
358: Implicit integration over the wavevector $\mathbf{k}$ is understood
359: in equations like (\ref{eq:pertubed bulk metric}). A priori, the
360: particular form of the perturbation variables $f_{\alpha\beta}$,
361: $f_\alpha$, $H_L$ and $H_T$ depend on the choice of gauge. However,
362: one can construct gauge invariant combinations as follows
363: \begin{equation}
364: F = H_L + \frac{1}{3} H_T + \frac{X^\alpha D_\alpha r}{r}, \qquad
365: F_{\alpha\beta} = f_{\alpha\beta} + D_\alpha X_\beta + D_\beta X_\alpha,
366: \end{equation}
367: where
368: \begin{equation}
369: X_\alpha = \frac{r}{k\ell} \left( f_\alpha + \frac{r}{k\ell} D_\alpha H_T
370: \right),
371: \end{equation}
372: and $D_\alpha$ is the covariant derivative associated with the
373: 2-metric $q_{\alpha\beta}$. The gauge invariant quantities are then
374: given in terms of the master variable $\Omega = \Omega(t,z)$
375: \begin{equation}
376: F = \frac{1}{6r} \left( D^\mu D_\mu - \frac{2}{\ell^2} \right)
377: \Omega, \quad F_{\alpha\beta} = \frac{1}{r} \left[ D_\alpha D_\beta -
378: \frac{1}{3} q_{\alpha\beta} \left(
379: 2 D^\mu D_\mu - \frac{1}{\ell^2} \right) \right]
380: \Omega.
381: \end{equation}
382: The bulk master variable satisfies the following wave equation
383: \begin{equation}\label{eq:bulk wave equation}
384: 0 = -\frac{\di^2\Omega}{\di \tau^2} + \frac{\di^2\Omega}{\di z^2}
385: + \frac{3}{z} \frac{\di\Omega}{\di z} + \left( \frac{1}{z^2} -
386: k^2 \right) \Omega.
387: \end{equation}
388:
389: \subsection{Gauge invariant brane perturbations}\label{sec:brane
390: perturbations}
391:
392: In addition to perturbations of the bulk metric, we must also
393: consider perturbations of the brane metric. These are
394: parameterized as
395: \begin{equation}
396: ds_\b^2 = (1-2\alpha Y)dt^2 - 2a\beta Y_i dt \, dx^i +
397: a^2 \left[ (1+2h_L)\delta_{ij} +
398: 2h_T Y_{ij} \right] dx^i dx^j.
399: \end{equation}
400: As in 4 dimensions, the gauge-dependant variables
401: $(\alpha,\beta,h_L,h_T)$ can be used to construct gauge-invariant
402: quantities
403: \begin{equation}
404: \Phi = h_L + \frac{1}{3} h_T - \frac{Ha}{k} \sigma_g, \qquad
405: \Psi = \alpha - \frac{1}{k} \frac{d}{dt} (a\sigma_g),
406: \end{equation}
407: where
408: \begin{equation}
409: \sigma_g = \frac{a}{k} \frac{dh_T}{dt} - \beta.
410: \end{equation}
411:
412: As demonstrated in Ref.~\cite{Shiromizu:1999wj}, the effective
413: Einstein equation on the brane is in general given by
414: \begin{equation}\label{eq:effective Einstein equation}
415: {}^{(4)}G_{ab} = \kappa_4^2 T_{ab} + \kappa_5^2 \Pi_{ab} -
416: \Weyl_{ab}, \qquad \Pi_{ab} = -\frac{1}{4} T_{ac}T^{c}{}_b +
417: \frac{1}{12} T T_{ab} +
418: \frac{1}{8} T_{cd} T^{cd} h_{ab} - \frac{1}{24} T^2
419: h_{ab}.
420: \end{equation}
421: Here, $\Weyl_{ab}$ represents the projection of the electric part
422: of the bulk Weyl tensor onto the brane. Notice that $\Weyl_{ab}$
423: satisfies
424: \begin{equation}\label{eq:Weyl conservation}
425: h^{ab} \Weyl_{ab} = 0, \qquad \nabla_a (\kappa_5^2 \Pi^{ab} - \Weyl^{ab} ) =
426: 0,
427: \end{equation}
428: and vanishes in the background geometry discussed in
429: \S\ref{sec:background}. When perturbing (\ref{eq:effective
430: Einstein equation}), one can treat $\Weyl_{ab}$ as an additional
431: fluid source with a radiation-like equation of state. Hence, we
432: parameterize the perturbations of $T_{ab}$ and $\Weyl_{ab}$ as
433: \begin{subequations}
434: \begin{align}
435: \delta T_0{}^0 & = -\delta\rho \, Y, & \delta
436: \Weyl_{0}{}^0 & = \kappa_4^2 \delta\rho_\Weyl Y, \\
437: \delta T_i {}^0 & = -k Y_i \delta q , & \delta
438: \Weyl_i{}^0 & = \kappa_4^2 k Y_i \delta q_\Weyl, \\
439: \delta T_i{}^j & = \delta p Y \delta_i{}^j +
440: k^2 \delta\pi Y_i{}^j, & \delta \Weyl_i{}^j & =
441: -\kappa_4^2 (\tfrac{1}{3} \delta\rho_\Weyl
442: Y \delta_i{}^j + k^2 \delta\pi_\Weyl
443: Y_i{}^j).
444: \end{align}
445: \end{subequations}
446: Henceforth, we will assume that the matter anisotropic stress
447: vanishes $\delta\pi = 0$. The Weyl fluid perturbations
448: $(\delta\rho_\Weyl,\delta q_\Weyl,\delta \pi_\Weyl)$ are the
449: ``Kaluza-Klein'' (KK) degrees of freedom of the model alluded
450: in the Introduction, since they represent the effects of bulk
451: geometry fluctuations on the brane. The matter perturbation
452: variables above are not gauge invariant, but the following
453: quantities are:
454: \begin{subequations}
455: \begin{align}
456: \rho \Delta & = \delta\rho - 3H\delta q, & a(\rho+p)V & =
457: -k \delta q - (\rho + p) a \sigma_g, & \Gamma & = \delta
458: p - c_s^2 \delta\rho, \\ \rho \Delta_\Weyl & = \delta\rho_\Weyl
459: - 3H\delta q_\Weyl, &
460: a(\rho+p)V_\Weyl & = -k \delta q_\Weyl - (\rho + p) a \sigma_g.
461: \end{align}
462: \end{subequations}
463: Here, we have defined the sound speed as $c_s^2 = \delta p /\delta
464: \rho$. Since the Weyl fluid has the equation of state $\delta
465: p_\Weyl = \tfrac{1}{3}\delta\rho_\Weyl$, there is no KK entropy
466: perturbation $\Gamma_\Weyl = 0$. Furthermore, the KK anisotropic
467: stress is automatically gauge invariant. Armed with these
468: definitions, it is possible to derive the following
469: gauge-invariant equations from the perturbation of
470: Eq.~(\ref{eq:effective Einstein equation}):
471: \begin{subequations}
472: \begin{align}
473: \Phi & = \frac{4 \pi G \rho a^2}{k^2} \left[ \left(1 + \frac{\rho}
474: {\sigma} \right) \Delta
475: + \Delta_\Weyl \right], \\
476: H\Psi - \frac{d\Phi}{dt} & = \frac{4\pi G (\rho+p) a}{k}
477: \left[ \left(1 + \frac{\rho}{\sigma} \right) V + V_\Weyl \right],
478: \\ \Phi + \Psi & = -8\pi G a^2 \delta\pi_\Weyl. \label{eq:KK anisotropic stress}
479: \end{align}
480: \end{subequations}
481: The first of these gives the high-energy $\mathcal{O}(\rho/\sigma)$
482: and KK corrections to the Poisson equation on the brane. The second
483: equation gives the modifications of the standard equation governing
484: the evolution of the peculiar velocity field. The third equation
485: demonstrates how the KK-modes can give rise to an anisotropic stress
486: on the brane.
487:
488: Additional equations can be obtained from perturbing the equation
489: representing the conservation of matter stress energy on the
490: brane: $\delta(h^{ab}\nabla_a T_{bc})=0$. These are
491: \begin{subequations}\label{eq:stress energy conservation
492: perturbed}
493: \begin{align}
494: \frac{1}{a} \frac{d}{dt} (aV) & = \frac{k}{a}\Psi + \frac{k}{a}
495: \frac{\Gamma+c_s^2\rho\Delta}{\rho(1+w)}, \\ \frac{1}{a^3}
496: \frac{d}{dt} (a^3\rho\Delta) & = -\frac{k}{a}
497: \rho(1+w)\left(1-\frac{3a^2}{k^2} \frac{dH}{dt} \right) V -
498: 3\rho(1+w)\left(\frac{d\Phi}{dt} - H\Psi\right).
499: \end{align}
500: \end{subequations}
501: Finally, we can also perturb the righthand expression in
502: Eq.~(\ref{eq:Weyl conservation}) to get two more equations, but
503: these are not sufficient to close the effective Einstein equations
504: (\ref{eq:effective Einstein equation}) on the brane. Hence we can
505: only go so far with this effective Weyl fluid description; a
506: complete treatment must describe how the brane degrees of freedom
507: presented here are coupled to the bulk degree of freedom $\Omega$.
508:
509: \subsection{Perturbation of the junction conditions}
510:
511: The means to connect the bulk perturbations of \S\ref{sec:bulk
512: perturbations} and the brane perturbations of \S\ref{sec:brane
513: perturbations} is the perturbation of the junction conditions
514: (\ref{eq:extrinsic}). This yields several results, the most
515: important of which is that $\Omega$ satisfies a boundary condition
516: on the brane
517: \begin{equation}\label{eq:boundary condition}
518: \left[ \di_n \Omega + \frac{1}{\ell} \left(1 +
519: \frac{\rho}{\sigma} \right) \Omega + \frac{6\rho a^3}{\sigma
520: k^2} \Delta \right]_\b = 0.
521: \end{equation}
522: The perturbation of (\ref{eq:extrinsic}) also gives the KK gauge
523: invariants in terms of $\Omega$:
524: \begin{subequations}
525: \begin{align}
526: \kappa_4^2 \rho \Delta_\Weyl & = \frac{k^2}{\ell a^3} \left[ \frac{k^2 + 3H^2 a^2}{3 a^2} \Omega
527: - H \di_u \Omega \right]_\b, \\
528: \kappa_4^2 a(\rho + p) V_\Weyl & = \frac{k^3}{3\ell a^2} \left[ H\Omega - \di_u \Omega
529: \right]_\b, \\ \kappa_4^2 \delta\pi_\Weyl & = \frac{1}{2\ell a^3}
530: \left[ \frac{1}{3} k^2 \Omega -H\di_u\Omega - \frac{3(\rho+p)}{\ell\sigma}
531: \di_n \Omega + \di_u^2 \Omega \right]_\b.
532: \end{align}
533: \end{subequations}
534: These reproduce the results first derived in
535: Ref.~\cite{Deffayet:2002fn}. Using these equations, we can find
536: explicit formulae for the gauge invariants defined above in terms of
537: $\Delta$ and $\Omega_\b(\eta) = \Omega(\tau_\b(\eta),z_\b(\eta))$:
538: \begin{subequations}\label{eq:explicit gauge invariants}
539: \begin{align}
540: \label{eq:explicit Phi} \Phi & = \frac{3a^2\rho(\rho+\sigma)}{k^2 \ell^2 \sigma^2}
541: \Delta + \left( \frac{3H^2 a^2+k^2}{6\ell a^3} \right) \Omega_\b
542: - \frac{H}{2\ell a^2} \frac{d\Omega_\b}{d\eta}, \\ \label{eq:explicit Psi} \Psi & =
543: -\frac{3\rho a^2 (3w\rho+4\rho+\sigma)}{k^2\ell^2\sigma^2} \Delta -
544: \left[ \frac{(3w+4)\rho^2}{2\ell^3 a\sigma^2} + \frac{(5+3w)\rho}{2\ell^3
545: a\sigma} +
546: \frac{k^2}{3\ell a^3} \right] \Omega_\b + \frac{3H}{2\ell a^2} \frac{d\Omega_\b}{d\eta}
547: - \frac{1}{2\ell a^3} \frac{d^2\Omega_\b}{d\eta^2}, \\ \label{eq:explicit V} V & =
548: \frac{3wHa}{k(1+w)} \Delta - \frac{1}{k(1+w)}
549: \frac{d\Delta}{d\eta} + \frac{kH}{2\ell a^2} \Omega_\b -
550: \frac{k}{2\ell a^3}
551: \frac{d\Omega_\b}{d\eta}.
552: \end{align}
553: \end{subequations}
554: Other quantities of interest are the curvature perturbation on
555: uniform density slices,
556: \begin{equation}\label{eq:explicit zeta}
557: \zeta = \Phi - \frac{Ha V}{k} + \frac{\Delta}{3(1+w)} = \left[ \frac{1}{3}
558: - \frac{3\rho a^2(w\sigma - \sigma -\rho)}{k^2\ell^2\sigma^2} \right]
559: \frac{\Delta}{1+w} + \frac{Ha}{k^2(1+w)} \frac{d\Delta}{d\eta}
560: + \frac{k^2}{6\ell a^3} \Omega_\b,
561: \end{equation}
562: and the curvature perturbation on comoving slices,
563: \begin{equation}
564: \mathcal{R}_c = \Phi - \frac{Ha V}{k} =
565: - \frac{3\rho a^2(w\sigma - \sigma -\rho)}{k^2\ell^2\sigma^2}
566: \frac{\Delta}{1+w} + \frac{Ha}{k^2(1+w)} \frac{d\Delta}{d\eta}
567: + \frac{k^2}{6\ell a^3} \Omega_\b,
568: \end{equation}
569: Hence, if we have knowledge of the density contrast $\Delta$ and the
570: value of $\Omega$ on the brane, we can obtain the behaviour of all
571: the other gauge invariants via algebra and differentiation.
572:
573: The above formulae can be easily manipulated to find a wave equation
574: for $\Delta$:
575: \begin{subequations}\label{eq:brane wave equation}
576: \begin{gather}
577: \frac{d^2 \Delta}{d\eta^2} + (1+3c_s^2-6w) Ha \frac{d\Delta}{d\eta} +
578: \left[c_s^2 k^2 + \frac{3\rho a^2}{\sigma\ell^2} A +
579: \frac{3\rho^2 a^2}{\sigma^2\ell^2} B \right] \Delta = -\frac{k^2\Gamma}{\rho} + \frac{k^4(1+w)\Omega_\b}{3\ell a^3}, \\
580: A = 6c_s^2 -1 -8w+3w^2, \qquad
581: B = 3c_s^2-9w-4.
582: \end{gather}
583: \end{subequations}
584: The above ODE, the bulk wave equation (\ref{eq:bulk wave equation})
585: and the boundary condition (\ref{eq:boundary condition}) comprise a
586: closed set of equations for $\Delta$ and $\Omega$.
587:
588: Before moving on, we have three remarks: First, note that in the
589: low energy universe, we can neglect $\mathcal{O}(\rho^2/\sigma^2)$
590: terms. Then, the $\Delta$ wave equation reduces to
591: \begin{equation}
592: \frac{d^2 \Delta}{d\eta^2} + (1+3c_s^2-6w) Ha \frac{d\Delta}{d\eta} +
593: \left[c_s^2 k^2 + 4\pi G\rho a^2 (6c_s^2 - 1 - 8w +3w^2)
594: \right] \Delta \approx -\frac{k^2\Gamma}{\rho} + \frac{k^4(1+w)\Omega_\b}{3\ell
595: a^3}.
596: \end{equation}
597: Setting $\Omega_\b = 0$ yields the standard 4-dimensional
598: dynamical equation for $\Delta$; hence, we recover GR at low
599: energies. Second, we note that from the conservation of
600: stress-energy on the brane (\ref{eq:stress energy conservation
601: perturbed}), we must have
602: \begin{equation}\label{eq:zeta conservation}
603: \frac{d\zeta}{d\eta} = -\frac{kV}{3} -
604: \frac{Ha\Gamma}{\rho(1+w)}.
605: \end{equation}
606: We can easily verify that Eqns.~(\ref{eq:explicit V}),
607: (\ref{eq:explicit zeta}) and (\ref{eq:brane wave equation}) imply
608: that (\ref{eq:zeta conservation}) is satisfied identically, which
609: is a good consistency check of our formulae. Finally, the entropy
610: perturbation can be simplified when the brane matter is either a
611: perfect fluid or a scalar field:
612: \begin{equation}
613: \Gamma = (\Upsilon - c_s^2) \rho \Delta, \quad \Upsilon = \begin{cases}
614: c_s^2, & \text{perfect fluid}, \\
615: 1, & \text{scalar field}. \end{cases}
616: \end{equation}
617: For the rest of this paper, we will be considering the perfect
618: fluid case.
619:
620: \subsection{Asymptotic behaviour in the high-energy radiation-dominated
621: regime}\label{sec:asymptotic}
622:
623: To model the perturbations in the very early universe, we assume
624: radiation domination and $\rho \gg \sigma$; or, equivalently, $H\ell
625: \gg 1$. Under these circumstances, the following approximations
626: hold:
627: \begin{equation}\label{eq:high-energy approximations}
628: w = 1/3, \quad a \approx a_0 (k\eta)^{1/3}, \quad \rho \approx \sigma H\ell,
629: \quad \di_n \approx -\di_u = -\di_t.
630: \end{equation}
631: The last relationship can be used to re-cast the boundary condition
632: on the bulk master variable $\Omega$ into a 1st order ordinary
633: differential equation for $\Omega_\b$. Hence, the equation of
634: motion for $\Delta$ and boundary condition become
635: \begin{equation}
636: 0 \approx \frac{d^2\Delta}{d\eta^2} + \left( \frac{k^2}{3} - \frac{4a_0 k^{1/3}}{3\ell \eta^{2/3}} -
637: \frac{2}{\eta^2} \right) \Delta - \frac{4k^3}{9\ell a_0^3\eta}
638: \Omega_\b, \quad \frac{d\Omega_\b}{d\eta} \approx \frac{1}{3\eta} \left( 1 + \frac{3a_0
639: k^{1/3} \eta^{4/3}}{\ell} \right) \Omega_\b + \frac{2\ell a_0^3}{k} \Delta.
640: \end{equation}
641: By differentiating the first equation we can derive a decoupled
642: third order equation for $\Delta$, which can be solved via power
643: series methods. The solutions for $\Delta$ and $\Omega$ are a
644: superposition of three linearly independent modes:
645: \begin{equation}
646: \Delta = \sum_{i=1}^3 \mathcal{A}_i \Delta^{(i)}, \quad
647: \Omega_\b = \sum_{i=1}^3 \mathcal{A}_i \Omega_\b^{(i)},
648: \end{equation}
649: where the $\mathcal{A}_i$ are constants. At leading order in
650: $(k\eta)$, these mode functions are given by
651: \begin{subequations}\label{eq:high-energy solutions}
652: \begin{align}
653: \text{dominant growing mode:} & & \Delta^{(1)} & \approx \tfrac{4}{3} (k\eta)^2
654: , &
655: \Omega_\b^{(1)} & \approx a_0^3 k^{-2} \ell (k\eta)^3
656: , \\
657: \text{subdominant growing mode:} & &
658: \Delta^{(2)} & \approx (k\eta)^{4/3},
659: & \Omega_\b^{(2)} & \approx -\tfrac{7}{2} a_0^3 k^{-2} \ell (k\eta)^{1/3}
660: ,\\
661: \text{decaying mode:} & & \Delta^{(3)} & \approx \tfrac{10}{3}
662: k\ell a_0^{-1} (k\eta)^{-1}, &
663: \Omega_\b^{(3)} & \approx -20 a_0^2 k^{-1} \ell^2.
664: \end{align}
665: \end{subequations}
666: As the labels on the left suggest, the growth of density
667: perturbations for the dominant growing mode is faster than for the
668: subdominant growing mode on superhorizon scales. The density
669: contrast of the third mode decays in the high-energy regime.
670:
671: We can calculate the behaviour of $\zeta$ and $\Psi$ for the
672: dominant growing mode by substituting the full series solutions
673: for $\Delta^{(1)}$ and $\Omega^{(1)}$ into (\ref{eq:explicit Psi})
674: and (\ref{eq:explicit zeta}), which yields:
675: \begin{equation}\label{eq:zeta Psi approx}
676: \zeta^{(1)} \approx 1, \qquad \Psi^{(1)} \approx -4 \zeta^{(1)}.
677: \end{equation}
678: We see that the dominant growing mode curvature perturbation and
679: Newtonian potential are conserved on superhorizon scales. Also
680: note that (\ref{eq:zeta Psi approx}) implies
681: \begin{equation}\label{eq:Delta approx}
682: \Delta^{(1)} = \frac{4}{27}\left( \frac{k}{Ha} \right)^2 \zeta^{(1)}.
683: \end{equation}
684: Eqs.~(\ref{eq:zeta Psi approx}) and (\ref{eq:Delta approx}) are
685: similar to the standard superhorizon growing mode results for a
686: radiation dominated universe in general relativity,
687: \begin{equation}\label{eq:GR relations}
688: \Psi^{(1)}_\GR \approx -\Phi^{(1)}_\GR \approx -\frac{3(1+w)}
689: {3w+5} \zeta^{(1)}_\GR = -\frac{2}{3} \zeta^{(1)}_\GR, \quad
690: \Delta_\GR^{(1)} \approx \frac{(3w+1)^2(1+w)}{2(3w+5)} \left(
691: \frac{k}{Ha} \right)^2 \zeta^{(1)}_\GR =
692: \frac{4}{9}\left( \frac{k}{Ha} \right)^2
693: \zeta^{(1)}_\GR,
694: \end{equation}
695: but the numerical factors are different.\footnote{It is
696: interesting to note that it is impossible to find an effective
697: equation of state parameter $w_\eff$ to make the general GR
698: formulae given in (\ref{eq:GR relations}) compatible with
699: (\ref{eq:zeta Psi approx}) and (\ref{eq:Delta approx}). Some
700: authors \cite{Sendouda:2006nu} have previously tried to describe
701: the high-energy radiation epoch of RS cosmology with an effective
702: fluid with $w_\eff = 5/3$, but we see that this would predict
703: $\Delta^{(1)} \approx 24/5 (k/Ha)^2 \zeta^{(1)}$ and $\Psi^{(1)}
704: \approx -4 \zeta^{(1)}/5$, which is in clear conflict with the
705: correct results (\ref{eq:zeta Psi approx}) and (\ref{eq:Delta
706: approx}).
707: %KK
708: This is due to a large Weyl anisotropic stress that modifies the GR
709: relationship between $\Psi$ and $\Phi$. }
710:
711: Finally, we note that the method described in the subsection fails
712: to yield an approximate solution for the other metric potential
713: $\Phi^{(1)}$. The reason is that when the full series expansions
714: for $\Delta^{(1)}$ and $\Omega_\b^{(1)}$ are substituted into
715: (\ref{eq:explicit Phi}), the leading order contribution vanishes,
716: leaving a result that is the same order as the error in the original
717: approximations (\ref{eq:high-energy approximations}). We must
718: therefore rely on numeric simulations to determine the asymptotic
719: behaviour of $\Phi$.
720:
721: \section{Numeric analysis}\label{sec:numeric}
722:
723: \subsection{Dimensionless parameters and integration algorithms}
724:
725: Our goal in the section is the solution of the system comprised of
726: the bulk wave equation (\ref{eq:bulk wave equation}) and the brane
727: wave equation (\ref{eq:brane wave equation}) subject to the boundary
728: condition (\ref{eq:boundary condition}). For the rest of the paper,
729: we will restrict ourselves to the case of a radiation-dominated
730: brane with $w = 1/3$. To perform the analysis, it is necessary to
731: make the various quantities in the equations dimensionless. To do
732: so, we define the ``*'' epoch as the moment in time when a mode with
733: wavenumber $k$ enters the Hubble horizon
734: \begin{equation}
735: k = H_* a_*, \qquad H_* = H(\eta_*), \qquad a_* = a(\eta_*),
736: \qquad z_* = z_\b(\eta_*).
737: \end{equation}
738: Then, we introduce dimensionless/normalized quantities as
739: \begin{equation}
740: \hat\Omega = \frac{\Omega}{a_*\ell^3}, \qquad \rhohat =
741: \frac{\rho}{\sigma}, \qquad \H = H\ell, \qquad \k = \H_* = \frac{k\ell}{a_*}
742: = kz_* =
743: \sqrt{\rhohat_*(\rhohat_* + 2)}, \qquad \a = \frac{a}{a_*}.
744: \end{equation}
745: Another important era is the critical epoch when $\H_c = H_c\ell =
746: 1$ and the radiation density has its critical value $\rhohat_c =
747: \sqrt{2}-1$. We define the critical epoch as the transition between
748: high and low energy regimes. The ratio of the wavenumber of any
749: given mode to the wavenumber $k_\c = H_\c a_\c$ of the critical mode
750: that enters the horizon at the critical epoch is
751: \begin{equation}
752: \frac{k}{k_\c} = \sqrt{\rhohat_*(\rhohat_*+2)} \left(
753: \frac{\sqrt{2}-1}{\rhohat_*} \right)^{1/4},
754: \end{equation}
755: where $a_\c$ is the value of the scale factor at the critical
756: epoch. Generally speaking, we call modes with $k > k_\c$
757: ``supercritical'' and modes with $k<k_\c$ ``subcritical''. The
758: scale defined by the critical mode in today's universe (with scale
759: factor $a_0$) is given by
760: \begin{equation}\label{eq:scale today}
761: \frac{a_0}{k_\c} = 1.4 \times
762: 10^{12} \left( \frac{\ell}{0.1\,\text{mm}} \right)^{1/2} \left( \frac{g_\c}{100}
763: \right)^{1/12} \text{m},
764: \end{equation}
765: where $g_\c$ is the number of relativistic degrees of freedom in the
766: matter sector at the critical epoch. For $\ell = 0.1\,\text{mm}$,
767: this corresponds to a scale of $\sim 10$ astronomical units (AU),
768: which is incredibly tiny by cosmological standards. Finally, if we
769: normalize all coordinates by $z_*$,
770: \begin{equation}
771: \hat\tau = \tau/z_*, \quad \hat z = z/z_*, \quad \hat\eta =
772: \eta/z_*,
773: \end{equation}
774: we find that the entire system of master equations is characterized
775: by the dimensionless matter density at horizon crossing $\rhohat_*$.
776:
777: Once the system has been reduced into the dimensionless form, we
778: have two independent numerical codes that can be used to solve for
779: $\Delta$ and $\hat\Omega_\b$. The first is the pseudo-spectral (PS)
780: method used in Ref.~\cite{Hiramatsu:2006cv, Hiramatsu:2006bd}
781: and the second is the characteristic integration (CI) algorithm developed in
782: Ref.~\cite{Cardoso:2006nh}\footnote{To apply the CI method as
783: described in \cite{Cardoso:2006nh}, the bulk wave equation needs to
784: be mapped into a canonical form via the change of variable
785: $\hat\Omega = (z_*/z)^{3/2}\phi$.}; detailed descriptions of each
786: method can be found in the respective papers. There is one technical
787: difference between the two algorithms that is worth highlighting
788: here: Namely, each code solves for $\hat\Omega$ in different
789: regions of the 5-dimensional spacetime, as shown in
790: Fig.~\ref{fig:compuational domains}. One implication of this is
791: that the initial data for each code needs to be specified at
792: different places. For the PS code, one needs to choose $\hat\Omega$
793: and $\di\hat\Omega/\di\hat\tau$ on an initial spacelike hypersurface
794: $\di\mathcal{M}^-_\PS$ of constant $\hat\tau$, while for the CI
795: algorithm one needs $\hat\Omega$ only on an initial null
796: hypersurface $\di\mathcal{M}^-_\CI$ intersecting the brane.
797: \begin{figure}
798: \begin{center}
799: \includegraphics{f1.eps} \caption{The
800: computational domains employed by the pseudo-spectral (PS) and
801: characteristic integration (CI) methods. The future and past
802: boundaries $\di\mathcal{M}_\PS^\pm$ of the PS domain $\mathcal{M}_\PS$
803: are lines of constant $\tau$, while for the CI method they are null
804: rays. In the PS method, one needs a constant $z$ regulator brane to
805: render the computational domain finite. In principle, the position of
806: the regulator is free, but it should be placed to the right of the
807: point $P$ in order to be outside the causal past of the physical brane.
808: Initial conditions for the two methods are placed on
809: $\di\mathcal{M}_\PS^-$ and $\di\mathcal{M}_\CI^-$,
810: respectively.\label{fig:compuational domains}}
811: \end{center}
812: \end{figure}
813:
814: What initial conditions should we actually use? The calculations of
815: \S\ref{sec:asymptotic} suggest that at sufficiently early times, the
816: dynamics of the system are well described by three distinct modes
817: (\ref{eq:high-energy solutions}). If our initial data surface is
818: positioned within this early era, it follows that generic choices of
819: initial conditions will excite some superposition of these three
820: modes. However, after a short period of time the dominant growing
821: mode will overtake the contributions from the subdominant growing
822: and decaying modes. Hence, we expect that the late time dynamics of
823: our model will be insensitive to the choice of initial conditions,
824: provided that the initial data hypersurface is oriented at an early
825: enough time. Since we do not expect the initial conditions to
826: matter very much, we make the simple choices
827: \begin{subequations}\label{eq:initial conditions}
828: \begin{align}
829: \text{PS initial conditions: } & & \Delta(O) & = \mathcal{N} \a_i^6,
830: & \hat\di_\eta \Delta(O) & = 6\mathcal{N}\H_i\a_i^7, & \hat\Omega(\di\mathcal{M}^-_\PS) & =
831: 0, & \hat\di_\tau \hat\Omega (\di\mathcal{M}^-_\PS) & = 6 \mathcal{N}\rhohat_* \a_i^6 /
832: \k^2 \H_i,
833: \\ \text{CI initial conditions: } & & \Delta(O) & = \mathcal{N}\a_i^6,
834: & \hat\di_\eta \Delta(O) & = 6\mathcal{N}\H_i\a_i^7, & \hat\Omega(\di\mathcal{M}^-_\CI) & =
835: 0, & &
836: \end{align}
837: \end{subequations}
838: where $\hat\di_\eta = \di/\di\etahat$, $\hat\di_\tau =
839: \di/\di\hat\tau$, and an ``$i$'' subscript denotes the initial
840: value of the scale factor and Hubble parameter. Here,
841: $\mathcal{N}$ is a normalization constant that we will often
842: select to make $\zeta = 1$ at the initial time. These initial
843: conditions are motivated by the fact that we expect $\Delta
844: \propto \a^6 \gg \hat\Omega_\b$ for the dominant growing mode at
845: very early times, which can certainly be satisfied by setting
846: $\hat\Omega = 0$ on the initial data surface. For the PS method,
847: the initial time derivative of $\hat\Omega$ is selected to satisfy
848: the boundary condition (\ref{eq:boundary condition}) at the
849: initial time. Note that the initial conditions for the two
850: methods are actually incompatible due to the different locations
851: of the initial surface. But as we have argued above and will see
852: below, this difference should make no difference to the final
853: results. (We will test this assumption in \S\ref{sec:initial
854: data} below.)
855:
856: \subsection{Typical waveforms}
857:
858: We now turn to the results of our simulations. In
859: Fig.~\ref{fig:compare} we show the output of the PS and CI codes
860: for a typical simulation of a supercritical mode with $\rhohat_* =
861: 50$. As expected we have excellent agreement between the two
862: codes, despite the fact that they use different initial conditions
863: (\ref{eq:initial conditions}). In fact, we have confirmed that the
864: codes give essentially identical results for a wide range of
865: parameters, which gives us confidence in our numerical methods.
866: It is also reassuring that the simulation results closely match
867: the analytic predictions of \S\ref{sec:asymptotic} on
868: high-energy/superhorizon scales. Note that for all simulations,
869: we recover that $\Delta$ and $\zeta$ are phase-locked plane waves,
870: \begin{equation}
871: \Delta(\eta) \propto \cos \frac{k\eta}{\sqrt{3}}, \quad \Delta(\eta)
872: \approx 4 \zeta(\eta),
873: \end{equation}
874: at sufficiently late times $k\eta \gg 1$, which is actually the same
875: behaviour as seen in GR, where we have the following exact
876: solutions for the growing mode density contrast and curvature
877: perturbation during radiation domination:
878: \begin{subequations}\label{eq:GR Delta zeta}
879: \begin{align}
880: \Delta^{(1)}_\GR & = \mathcal{A} \left( \frac{\sqrt{3}}{k\eta}\sin \frac{k\eta}{\sqrt{3}} -
881: 4 \cos \frac{k\eta}{\sqrt{3}} \right) \approx \mathcal{A} \begin{cases} \frac{4}{9}
882: k^2\eta^2, & k\eta \ll 1, \\ -4 \cos \frac{k\eta}{\sqrt{3}}, &
883: k\eta \gg 1, \end{cases} \\
884: \zeta^{(1)}_\GR & = \mathcal{A} \left( \frac{2\sqrt{3}}{k\eta}\sin \frac{k\eta}{\sqrt{3}} -
885: \cos \frac{k\eta}{\sqrt{3}} \right) \approx \mathcal{A}
886: \begin{cases} 1, & k\eta \ll 1, \\ -\cos \frac{k\eta}{\sqrt{3}}, &
887: k\eta \gg 1, \end{cases}
888: \end{align}
889: \end{subequations}
890: where $\mathcal{A}$ is a constant.
891: \begin{figure}
892: \begin{center}
893: \includegraphics{f2.eps}
894: \caption{Comparison between typical results of the PS and CI
895: codes for various brane quantities (\emph{left}); and the
896: typical behaviour of the bulk master variable (\emph{right})
897: as calculated by the CI method. Very good agreement between
898: the two different numerical schemes is seen in the left panel,
899: despite the fact that they use different
900: initial conditions. We also see excellent consistency between the
901: simulations and the approximations developed in
902: \S\ref{sec:asymptotic} for the behaviour of the system on
903: high-energy/superhorizon scales
904: [c.f.~Eqs.~(\ref{eq:high-energy solutions})--(\ref{eq:Delta approx})].
905: Also note that on subhorizon scales, $\Delta$ and $\zeta$
906: undergo simple harmonic oscillations, which is consistent with the behaviour in
907: GR (\ref{eq:GR Delta zeta}). The bulk profile demonstrates
908: our choice of initial conditions: We see that the bulk master variable $\hat\Omega$
909: is essentially zero during the early stages of the simulation,
910: and only becomes ``large'' when the mode crosses the horizon.
911: \label{fig:compare}}
912: \end{center}
913: \end{figure}
914: \begin{figure}
915: \begin{center}
916: \includegraphics{f3.eps} \caption{Typical behaviour of
917: various brane gauge invariants for modes entering the horizon
918: in subcritical (\emph{upper panels}) and supercritical
919: (\emph{lower panels}) epochs. For the plots on the left, the
920: grey straight lines give the analytic expectations
921: (\ref{eq:high-energy solutions}) for the behaviour of $\Delta$ and
922: $\hat\Omega_\b$ on large scales and high energies. The central plots show the
923: comoving and uniform density curvature perturbations,
924: $\mathcal{R}_c$ and $\zeta$ respectively. As in conventional
925: cosmology, these are conserved and equal on superhorizon
926: scales. Also note that the amplitude of $\zeta$ after horizon
927: crossing is larger for the mode with larger $k$.
928: The plots on the right show the two gauge invariant metric
929: perturbations on the brane. Before the critical epoch we have
930: that $|\Psi| \gg |\Phi|$, while afterwards we recover the GR
931: result $\Phi \approx -\Psi$. Finally, in the high-energy regime
932: we see $\Psi \approx -4\zeta$, in line with the approximation
933: (\ref{eq:zeta Psi approx}). \label{fig:typical}}
934: \end{center}
935: \end{figure}
936:
937: In Fig.~\ref{fig:typical}, we show the simulated behaviour of
938: several different gauge invariants on the brane for given
939: subcritical ($k = 0.25 k_\c$) and supercritical ($k = 10 k_\c$)
940: modes. These plots illustrate several points that are generic to
941: all of our simulations: The curvature perturbations
942: $\mathcal{R}_c$ and $\zeta$ are conserved on superhorizon scales
943: for all cases, just as in general relativity. This useful fact
944: allows us to define the primordial amplitude of a given
945: perturbation as the value of $\zeta$ or $\mathcal{R}_c$ before
946: horizon crossing. For the metric perturbations $\Phi$ and $\Psi$,
947: we find that for supercritical and superhorizon scales
948: \begin{equation}
949: |\Psi| \gg |\Phi|, \quad \Psi \approx -4\zeta = \text{constant},
950: \quad (\text{for $a \lesssim a_\c$ and $k \lesssim Ha$}).
951: \end{equation}
952: This implies that the Weyl anisotropic stress
953: $\kappa_4^2\delta\pi_\KK \approx -\Psi/a^2$ is relatively large in
954: the high-energy superhorizon regime.
955:
956: Finally, Fig.~\ref{fig:subcritical} illustrates how the ordinary
957: superhorizon behaviour of perturbations in GR is recovered for
958: modes entering the Hubble horizon in the low energy era. We see
959: how $\Delta$, $\Phi$ and $\Psi$ smoothly interpolate between the
960: non-standard high-energy behaviour to the usual expectations given
961: by (\ref{eq:GR relations}). Also shown in this plot is the
962: behaviour of the KK anisotropic stress, which steadily decays
963: throughout the simulation. This is characteristic of all the
964: cases we have studied.
965: \begin{figure}
966: \begin{center}
967: \includegraphics{f4.eps} \caption{The simulated behaviour of
968: an extremely subcritical $k \ll k_\c$ mode on superhorizon scales. On the left we show
969: how the $\Delta$ gauge invariant switches from the high-energy behaviour
970: predicted in Eq.~(\ref{eq:Delta approx}) to the familiar GR result
971: $(\ref{eq:GR relations})$ as the universe expands through the
972: critical epoch. We also show how the KK anisotropic stress $\kappa_4^2 \delta\pi_\KK$
973: steadily decays throughout the simulation, which is typical of
974: all the cases we have investigated. On the right, we show the
975: metric perturbations $\Phi$ and $\Psi$ as well as the curvature perturbation
976: $\zeta$. Again, note how the GR result $\Phi \approx -\Psi \approx -2\zeta/3$
977: is recovered at low energy. \label{fig:subcritical}}
978: \end{center}
979: \end{figure}
980:
981: \subsection{Enhancement factors and the transfer
982: function}\label{sec:transfer}
983:
984: If we examine the curvature perturbations plots in
985: Fig.~\ref{fig:typical} in detail, we see that the amplitude of
986: $\zeta$ increases during horizon crossing. Furthermore, the degree
987: of enhancement seems to increase with increasing $k$. This is quite
988: different from the behaviour of the perturbations of a radiation
989: dominated universe in ordinary GR (\ref{eq:GR Delta
990: zeta}), where the asymptotic amplitude of $\zeta$ is the same before and
991: after horizon crossing. Hence, in the braneworld case we see an
992: enhancement in the amplitude of perturbations that is not present in
993: conventional theory.
994:
995: What is responsible for this enhancement? As discussed in the
996: Introduction, there are actually two separate effects to consider:
997: First, there is the modification of the universe's expansion at high
998: energies and the $\mathcal{O}(\rho/\sigma)$ corrections to the
999: perturbative equations of motion (\ref{eq:explicit gauge
1000: invariants})--(\ref{eq:brane wave equation}). Second, there is the
1001: effect of the bulk degrees of freedom encapsulated by the bulk
1002: master variable $\Omega$ (or, equivalently, the KK fluid variables
1003: $\Delta_\KK$, $V_\KK$ and $\delta\pi_\KK$). To separate out the two
1004: effects, it is useful to introduce the 4-dimensional effective
1005: theory discussed above where all $\mathcal{O}(\rho/\sigma)$
1006: corrections to GR are retained, but the bulk effects are removed by
1007: artificially setting $\Omega = 0$. In the case of radiation
1008: domination, we obtain equations for the effective theory density
1009: contrast $\Delta_\eff$ and curvature perturbation $\zeta_\eff$ from
1010: (\ref{eq:explicit zeta}) and (\ref{eq:brane wave equation}) with
1011: $\Omega_\b = 0$:
1012: \begin{subequations}
1013: \begin{align}
1014: 0 & = \frac{d^2\Delta_\eff}{d\eta^2} + \left( \frac{k^2}{3} -
1015: \frac{4\rho a^2}{\sigma\ell^2} - \frac{18\rho^2
1016: a^2}{\sigma^2\ell^2}\right) \Delta_\eff, \\
1017: \zeta_\eff & = \left( \frac{1}{4}
1018: + \frac{3 \rho a^2}{2 \sigma k^2 \ell^2} + \frac{9\rho^2
1019: a^2}{4\sigma^2k^2\ell^2}\right)\Delta_\eff + \frac{3Ha}
1020: {4k}\frac{d\Delta_\eff}{d\eta}.
1021: \end{align}
1022: \end{subequations}
1023: These in conjunction with the Friedmann equation
1024: (\ref{eq:Friedmann}) and the conservation of stress energy
1025: (\ref{eq:conservation}) give a closed set of ODEs on the brane
1026: that describe all of the $\mathcal{O}(\rho/\sigma)$ corrections to
1027: GR.
1028:
1029: In Fig.~\ref{fig:effective}, we plot the predictions of GR, the
1030: effective theory, and the full 5-dimensional simulations for the
1031: behaviour of $\zeta$ and $\Delta$ for a supercritical mode. Since
1032: in any given model we expect the primordial value of the curvature
1033: perturbation to be fixed by inflation, it makes physical sense to
1034: normalize the waveforms from each theory such that
1035: \begin{equation}\label{eq:zeta normalization}
1036: \zeta_\fiveD \approx \zeta_\eff \approx \zeta_\GR \approx 1,
1037: \quad a \ll a_*.
1038: \end{equation}
1039: When this is enforced we see that the effective theory predicts a
1040: larger final amplitude for the density perturbation than GR.
1041: Furthermore, the final amplitude in the 5-dimensional simulation
1042: is larger than both of the other theories. From this we can infer
1043: that both $\mathcal{O}(\rho/\sigma)$ and KK effects induce
1044: enhancement in the amplitude of perturbations. This is in
1045: contrast to the situation for tensor perturbations in the
1046: Randall-Sundrum model, where modification of the expansion serves
1047: to increase the amplitude of gravitational waves while the bulk
1048: effects tend to decrease it
1049: \cite{Hiramatsu:2003iz,Hiramatsu:2004aa,Kobayashi:2005dd,Kobayashi:2005jx,%
1050: Kobayashi:2006pe,Ichiki:2003hf,Seahra:2006tm}.
1051: \begin{figure}
1052: \begin{center}
1053: \includegraphics{f5.eps} \caption{A comparison of the
1054: behaviour of the curvature perturbation $\zeta$ (\emph{left}) and the density
1055: perturbation $\Delta$ (\emph{right}) in the full 5-dimensional theory including
1056: KK contributions (5D), the effective 4-dimension theory including $\mathcal{O}(\rho/\sigma)$
1057: corrections (eff), and ordinary general relativity (GR). The waveforms for each theory are normalized
1058: such that $\zeta = 1$ on superhorizon scales. We can clearly see that given the
1059: same primordial values of the curvature perturbation, the final amplitude of the
1060: density perturbation in the 5D theory $\mathcal{C}_\fiveD$ is larger than in the
1061: effective theory $\mathcal{C}_\eff$, which in turn is larger than the GR
1062: value $\mathcal{C}_\GR = 4$.\label{fig:effective}}
1063: \end{center}
1064: \end{figure}
1065:
1066: The amount of enhancement should be a function of the frequency of
1067: the mode, since we expect extremely subcritical modes $k \ll k_\c$
1068: to behave as in GR. To test this, we can define a set of
1069: ``enhancement factors'', which are functions of $k$ that describe
1070: the relative amplitudes of $\Delta$ after horizon crossing in the
1071: various theories. As in Fig.~\ref{fig:effective}, let the final
1072: amplitudes of the density perturbation with wavenumber $k$ be
1073: $\mathcal{C}_\fiveD(k)$, $\mathcal{C}_\eff(k)$ and
1074: $\mathcal{C}_\GR(k)$ for the 5-dimensional, effective and GR
1075: theories, respectively, given that the normalization (\ref{eq:zeta
1076: normalization}) holds. Then, we define enhancement factors as
1077: \begin{equation}
1078: \mathcal{Q}_\eff(k) =
1079: \frac{\mathcal{C}_\eff(k)}{\mathcal{C}_\GR(k)}, \quad \mathcal{Q}_\KK(k) =
1080: \frac{\mathcal{C}_\fiveD(k)}{\mathcal{C}_\eff(k)}, \quad \mathcal{Q}_\fiveD(k) =
1081: \frac{\mathcal{C}_\fiveD(k)}{\mathcal{C}_\GR(k)}.
1082: \end{equation}
1083: It follows that $\mathcal{Q}_\eff(k)$ represents the
1084: $\mathcal{O}(\rho/\sigma)$ enhancement to the density perturbation,
1085: $\mathcal{Q}_\KK(k)$ gives the magnification due to KK modes, while
1086: $\mathcal{Q}_\fiveD(k)$ gives the total 5-dimensional amplification
1087: over the GR case. These enhancement factors are shown in the left
1088: panel of Fig.~\ref{fig:spectra}. We can see that they all increase
1089: as the scale is decreased, and that they all approach unity for $k
1090: \rightarrow 0$. Since $\mathcal{Q} = 1$ implies no enhancement of
1091: the density perturbations over the standard result, this means we
1092: recover general relativity on large scales. For all wavenumbers we
1093: see $\mathcal{Q}_\eff
1094: > \mathcal{Q}_\KK > 1$, which implies that the amplitude magnification due
1095: to the $\mathcal{O}(\rho/\sigma)$ corrections is always larger than
1096: that due to the KK modes. Interestingly, the $\mathcal{Q}$-factors
1097: appear to approach asymptotically constant values for large $k$:
1098: \begin{equation}
1099: \mathcal{Q}_\eff(k) \approx 3.0,
1100: \quad \mathcal{Q}_\KK(k) \approx
1101: 2.4,
1102: \quad \mathcal{Q}_\fiveD(k) \approx
1103: 7.1, \quad k \gg k_\c.
1104: \end{equation}
1105: It is difficult to know if these are the true asymptotic limits for
1106: $k \rightarrow \infty$ due to the limitations of computing power.
1107: \begin{figure}
1108: \begin{center}
1109: \includegraphics{f6.eps} \caption{Density perturbation
1110: enhancement factors (\emph{left}) and transfer functions
1111: (\emph{right}) from simulations, effective theory, and general
1112: relativity. All of the $\mathcal{Q}$ factors monotonically
1113: increase with $k/k_\c$, and we see that the $\Delta$ amplitude
1114: enhancement due to $\mathcal{O}(\rho/\sigma)$ effects $\mathcal{Q}_\eff$ is
1115: generally larger than the enhancement due to KK effects
1116: $\mathcal{Q}_\KK$. For asymptotically small scales $k \gg k_\c$,
1117: the enhancement seems to level off. The transfer functions in
1118: the right panel are evaluated at a given subcritical epoch in the
1119: radiation dominated era. The $T$ functions show how, for a fixed
1120: primordial spectrum of curvature perturbations
1121: $\mathcal{P}^\prim_\zeta$, the effective theory predicts excess
1122: power in the $\Delta$ spectrum $\mathcal{P}_\Delta \propto T^2
1123: \mathcal{P}^\prim_\zeta$ on supercritical/subhorizon scales
1124: compared to the GR result. The excess small-scale power is even
1125: greater when KK modes are taken into account, as shown by
1126: $T_\fiveD(k;\eta)$. \label{fig:spectra}}
1127: \end{center}
1128: \end{figure}
1129:
1130: In cosmological perturbation theory, transfer functions are very
1131: important quantities. They allow one to transform the primordial
1132: spectrum of some quantity set during inflation into the spectrum
1133: of another quantity at a later time. In this sense, they are
1134: essentially the Fourier transform of the retarded Green's function
1135: for cosmological perturbations. There are many different transfer
1136: functions one can define, but for our case it is useful to
1137: consider a function $T(k)$ that will tell us how the initial
1138: spectrum of curvature perturbations $\mathcal{P}_\zeta^\prim$ maps
1139: onto the spectrum of density perturbations $\mathcal{P}_\Delta$ at
1140: some low energy epoch within the radiation era characterized by
1141: the conformal time $\eta > \eta_\c$. It is customary to normalize
1142: transfer functions such that $T(k;\eta) \approaches{k}{0} 1$,
1143: which leads us to the following definition
1144: \begin{equation}
1145: T(k;\eta) = \frac{9}{4} \left[ \frac{k}{H(\eta)a(\eta)}
1146: \right]^{-2} \frac{\Delta_k(\eta)}{\zeta^\prim_k}.
1147: \end{equation}
1148: Here, $\zeta^\prim_k$ is the primordial value of the curvature
1149: perturbation and $\Delta_k(\eta)$ is the maximum amplitude of the
1150: density perturbation in the epoch of interest. As demonstrated in
1151: Fig.~\ref{fig:subcritical}, we know that we recover the GR result
1152: in the extreme small scale limit
1153: \begin{equation}
1154: \Delta_k(\eta) \approaches{k}{0} \frac{4}{9} \left[ \frac{k}{H(\eta)a(\eta)}
1155: \right]^{2} \zeta^\prim_k, \quad (a > a_c),
1156: \end{equation}
1157: which gives the transfer function the correct normalization. In
1158: GR, the transfer function is accurately given by
1159: \begin{equation}\label{eq:GR transfer function}
1160: T_\GR(k;\eta) \approx
1161: \begin{cases}
1162: 1, & k < 3Ha, \\ (3Ha/k)^2, & k > 3Ha.
1163: \end{cases}
1164: \end{equation}
1165: In the righthand panel of Fig.~\ref{fig:spectra}, we show the
1166: transfer functions derived from GR, the effective theory and the
1167: 5-dimensional simulations. As expected, the $T(k;\eta)$ for each
1168: formulation match one another on subcritical scales $k < k_\c$.
1169: However, on supercritical scales we have $T_\fiveD > T_\eff >
1170: T_\GR$.
1171:
1172: Note that if we are interested in the transfer function at some
1173: arbitrary epoch in the low-energy radiation regime $Ha \gg k_c$, it
1174: is approximately given in terms of the enhancement factor as
1175: follows:
1176: \begin{equation}\label{eq:5D transfer function}
1177: T_\fiveD(k;\eta) \approx
1178: \begin{cases}
1179: 1, & k < 3Ha, \\ (3Ha/k)^2 \mathcal{Q}_\fiveD(k), & k > 3Ha,
1180: \end{cases}
1181: \end{equation}
1182: Now, the spectrum of density fluctuations at any point in the
1183: radiation era is given by
1184: \begin{equation}
1185: \mathcal{P}_\Delta(k;\eta) = \frac{16}{81} T^2(k;\eta)
1186: \left( \frac{k}{Ha} \right)^4 \mathcal{P}^\prim_\zeta(k).
1187: \end{equation}
1188: Using the enhancement factor results in Fig.~\ref{fig:spectra} with
1189: Eqns.~(\ref{eq:GR transfer function}) and (\ref{eq:5D transfer
1190: function}), we see that the RS matter power spectrum (evaluated in
1191: the low-energy regime) is $\sim 50$ times bigger than the GR
1192: prediction on supercritical scales $k \sim 10^3 k_\c$.
1193:
1194: \subsection{Alternate initial data}\label{sec:initial data}
1195:
1196: Before concluding, we wish to briefly revisit the issue of initial
1197: conditions. We have argued above that the existence of a dominant
1198: growing mode at high-energy and on superhorizon scales means that
1199: the details of initial data for our simulations are unimportant. To
1200: some degree, the fact that the PS and CI algorithms produce
1201: essentially identical results is a good confirmation of this, since
1202: each method uses a different prescription for initial data.
1203:
1204: But what if we were to use different classes of initial data? For
1205: example, we could modify the CI initial data such that the bulk
1206: field has constant amplitude along the initial data surface:
1207: \begin{equation}
1208: \hat\Omega(\di\mathcal{M}_\CI^-) = \hat\Omega_i =
1209: \text{constant}.
1210: \end{equation}
1211: The waveforms generated by such a choice are shown in
1212: Fig.~\ref{fig:init data}. Though the overall amplitude and early
1213: time behaviour of the signal seems to be sensitive to the initial
1214: value of $\hat\Omega$, the ratio of the final amplitude of $\Delta$
1215: to the initial value of $\zeta$ is the same for each choice of
1216: initial data. We have confirmed that this is also true for other choices
1217: of $\hat\Omega(\di\mathcal{M}_\CI^-)$, including the case when the
1218: initial data is a simple sinusoid.\footnote{Sinusoidal initial data
1219: was used in Ref.~\cite{Seahra:2006tm} to test the insensitivity of
1220: the final spectrum of the stochastic background of gravitational
1221: waves to initial conditions in RS models.} Ultimately, it is the
1222: ratio $\Delta(k\eta \gg 1)/\zeta(k\eta \ll 1)$ that is relevant to
1223: the transfer function used to transcribe the predictions of
1224: inflation into observable quantities, hence we can be confident that
1225: our principal results hold for reasonable modifications of initial
1226: data.
1227: \begin{figure}
1228: \begin{center}
1229: \includegraphics{f7.eps} \caption{Simulated waveforms for
1230: $\hat\Omega =$ constant initial data in the CI method. Here,
1231: $(\hat\Omega/\Delta)_i$ is the ratio of the bulk field to the density
1232: perturbation at the initial time; this ratio is zero for the other simulations
1233: in this paper. The early time behaviour of $\Delta$ and $\hat\Omega_\b$ shows some
1234: sensitivity to $(\hat\Omega/\Delta)_i$, but otherwise the waveforms are
1235: identical to one another up to an overall amplitude scaling. Indeed, if we normalize
1236: the simulation results by the initial value of $\zeta$ we find that
1237: all of the $\Delta$ waveforms are coincident at late times, which means that
1238: the enhancement factor $\mathcal{Q}_\fiveD$ and transfer function
1239: $T_\fiveD$ are insensitive to $(\hat\Omega/\Delta)_i$. Note that
1240: at early times we have $\hat\Omega_\b \propto a$, which matches
1241: the behaviour of the subdominant growing mode from Eq.~(\ref{eq:high-energy solutions}).
1242: This confirms our expectations: varying the initial data causes the other subdominant modes
1243: to be excited to various degrees, but the dominant growing mode always ``wins''
1244: at late times.\label{fig:init data}}
1245: \end{center}
1246: \end{figure}
1247:
1248: \section{Discussion and implications}\label{sec:discussion}
1249:
1250: In this paper, we have written down the equations of motion for
1251: generic gauge-invariant scalar cosmological perturbations in the RS
1252: braneworld model in a form suitable for numeric analysis
1253: (\S\ref{sec:scalar perturbations}). We have developed new analytic
1254: approximations for the behaviour of fluctuations on high-energy
1255: superhorizon scales (\S\ref{sec:asymptotic}) for a radiation
1256: dominated brane. We have applied two different numerical algorithms
1257: to solve the equations in the early radiation-dominated universe
1258: (\S\ref{sec:numeric}). Our numerical results show that the amplitude
1259: of modes which enter the Hubble horizon during the high-energy
1260: regime gets enhanced over the the standard GR result. Conversely,
1261: modes which enter at low energy do not show any late-time deviations
1262: from standard theory, as seen in Fig.~\ref{fig:subcritical}. We
1263: only recover standard results for the metric perturbations, such as
1264: $\Psi \approx -\Phi$, at low energies.
1265:
1266: Our simulations confirm the common sense expectation that all
1267: tangible effects from the fifth dimension are on scales smaller than
1268: a critical value $k_c^{-1}$, whose physical size today is given by
1269: Eq.~(\ref{eq:scale today}). To parameterize the degree of
1270: amplification as a function of scale, we introduced so-called
1271: ``enhancement factors'' and transfer functions in
1272: \S\ref{sec:transfer}. These show that the degree of enhancement
1273: over the GR case seems to reach a constant value at high $k$; the
1274: amplification of the subhorizon matter power spectrum is $\sim 50$
1275: for $k \sim 10^3 k_c$. We presented analytic arguments and
1276: numerical evidence that our results are robust against modifications
1277: of the initial data for simulations (\S\ref{sec:initial data}).
1278:
1279: As discussed in \S\ref{sec:introduction} and \S\ref{sec:transfer},
1280: the enhancement of perturbations upon horizon crossing can be
1281: attributed to two effects: namely, the $\mathcal{O}(\rho/\sigma)$
1282: corrections to the background dynamics and the influence of the KK
1283: modes. Both of these give roughly equal contributions to the
1284: enhancement (Fig.~\ref{fig:spectra}). One could have anticipated
1285: the KK enhancement on simple physical grounds by arguing as follows:
1286: In Ref.~\cite{Garriga:1999yh} it was shown that the gravitational
1287: force of attraction between two bodies in the RS one brane model is
1288: \begin{equation}
1289: \text{gravitational force} \sim \begin{cases}
1290: {\kappa_4^2}/{r^2}, & r \gg \ell, \\
1291: {\kappa_5^2}/{r^3} = {\kappa_4^2 \ell}/{r^3}, &
1292: r \ll \ell.
1293: \end{cases}
1294: \end{equation}
1295: %The $1/r^4$ correction term is the direct effect of the KK degrees
1296: %of freedom on Newton's law.
1297: %KK
1298: That is, the Newtonian force becomes 5-dimensional on small scales.
1299: %KK
1300: This implies that the gravitational
1301: force is stronger than usual on scales $r \lesssim \ell$. It
1302: then follows that the self-gravity of perturbative modes that
1303: enter the horizon at high energies $H\ell \gtrsim 1$ will be
1304: greater than those which enter at low energies $H\ell \lesssim 1$.
1305: Therefore, we should expect that the amplitude of small-scale
1306: modes to be magnified over the amplitude of large-scale modes
1307: during horizon crossing in braneworld cosmology, which is exactly
1308: what we have seen in our simulations.\footnote{However, we should note
1309: that this simple argument neglects the influence of the KK
1310: anisotropic stress, which we know is non-negligible at high
1311: energies.}
1312:
1313: The amplitude enhancement of perturbations is important on comoving
1314: scales $\lesssim 10 \,\text{AU}$, which are far too small to be
1315: relevant to present-day/cosmic microwave background measurements of
1316: the matter power spectrum. However, it may have an important bearing
1317: on the formation of compact objects such as primordial black holes
1318: and boson stars at very high energies, i.e. the greater
1319: gravitational force of attraction in the early universe will create
1320: more of these objects than in GR (different aspects of primordial
1321: black holes in RS cosmology in the context of various effective
1322: theories have been considered in
1323: Refs.~\cite{Guedens:2002km,Guedens:2002sd,Majumdar:2002mr,Clancy:2003zd,%
1324: Sendouda:2003dc,Sendouda:2004hz,Sendouda:2006nu}). We know that the
1325: abundance of primordial black holes can be constrained by big bang
1326: nucleosythesis and observations of high-energy cosmic rays, so it
1327: would be interesting to see if the kind of enhancement of density
1328: perturbations predicted in this paper can be used to derive new
1329: limits on Randall-Sundrum cosmology. We leave this issue for future
1330: work.
1331:
1332: \begin{acknowledgments}
1333: AC is supported by FCT (Portugal) PhD fellowship SFRH/BD/19853/2004.
1334: KK and SSS are supported by PPARC (UK).
1335: \end{acknowledgments}
1336:
1337: \bibliography{ms}
1338:
1339: \end{document}
1340: