0808.2208/dgp.tex
1: \documentclass[twocolumn,prd,amsmath,superscriptaddress,nofootinbib,floatfix]{revtex4}
2: 
3: \usepackage{amsmath}
4: %\usepackage{bm}
5: \usepackage{graphics}
6: \usepackage{color}
7: %\newcommand{\wh}[1]{\textcolor{blue}{[{\bf WH}: #1]}}
8: %\newcommand{\wf}[1]{\textcolor{red}{[{\bf WF}: #1]}}
9: 
10: \topmargin -0.5in
11: 
12: \def\gsim{\;\rlap{\lower 2.5pt
13:  \hbox{$\sim$}}\raise 1.5pt\hbox{$>$}\;}
14: \def\lsim{\;\rlap{\lower 2.5pt
15:  \hbox{$\sim$}}\raise 1.5pt\hbox{$<$}\;}
16: 
17: \def\be{\begin{equation}}
18: \def\ee{\end{equation}}
19: \def\bea{\begin{eqnarray}}
20: \def\eea{\end{eqnarray}}
21: \def\nn{\nonumber}
22: \def\bt{\bm{\theta}}
23: \def\cf{\mathcal{F}}
24: \def\ch{\mathcal{H}}
25: 
26: %
27: \begin{document}
28: 
29: \title{Challenges to the DGP Model from Horizon-Scale Growth
30: and Geometry}
31: 
32: %
33: \author{Wenjuan Fang}
34: \affiliation{Department of Physics, Columbia University, New York,
35: NY 10027}
36: \author{Sheng Wang}
37: \affiliation{Kavli Institute for Cosmological Physics, Enrico Fermi
38: Institute, University of Chicago, Chicago, IL 60637}
39: \author{Wayne Hu}
40: \affiliation{Kavli Institute for Cosmological Physics, Enrico Fermi
41: Institute, University of Chicago, Chicago, IL 60637}
42: \affiliation{Department of Astronomy and Astrophysics, University of
43: Chicago, Chicago, IL 60637}
44: \author{Zolt\'{a}n Haiman}
45: \affiliation{Department of Astronomy, Columbia University, New York,
46: NY 10027}
47: \author{Lam Hui}
48: \affiliation{Department of Physics, Columbia University, New York,
49: NY 10027}
50: \author{Morgan May}
51: \affiliation{Brookhaven National Laboratory, Upton, NY 11973}
52: 
53: %
54: \date{\today}
55: 
56: %
57: \begin{abstract}
58: We conduct a Markov Chain Monte Carlo study of the
59: Dvali-Gabadadze-Porrati (DGP) self-accelerating braneworld scenario
60: given the cosmic microwave background (CMB) anisotropy, supernovae and
61: Hubble constant data by implementing an effective dark energy
62: prescription for modified gravity into a standard Einstein-Boltzmann
63: code.  We find no way to alleviate the tension between distance measures
64: and horizon scale growth in this model.  Growth alterations due to
65: perturbations propagating into the bulk appear as excess CMB anisotropy
66: at the lowest multipoles.  In a flat cosmology, the maximum likelihood
67: DGP model is nominally a $5.3\sigma$ poorer fit than $\Lambda$CDM.
68: Curvature can reduce the tension between distance measures but only at
69: the expense of exacerbating the problem with growth leading to a
70: $4.8\sigma$ result that is dominated by the low multipole CMB
71: temperature spectrum.  While changing the initial conditions to reduce
72: large scale power can flatten the temperature spectrum, this also
73: suppresses the large angle polarization spectrum in violation of recent
74: results from WMAP5.  The failure of this model highlights the power of
75: combining growth and distance measures in cosmology as a test of gravity
76: on the largest scales.
77: \end{abstract}
78: 
79: %
80: \maketitle
81: 
82: %
83: \section{Introduction}
84: 
85: On the self-accelerating branch of the Dvali-Gabadadze-Porrati (DGP)
86: braneworld model \cite{DGP}, cosmic acceleration arises from a
87: modification to gravity at large scales rather than by introducing a
88: form of mysterious dark energy with negative pressure \cite{De01}.  In
89: this model our universe is a ($3+1$)-dimensional brane embedded in an
90: infinite Minkowski bulk with differing effective strengths of gravity.
91: The relative strengths define a crossover scale on the brane beyond
92: which ($4+1$)-dimensional gravity and bulk phenomena become important.
93: 
94: A number of theoretical and observational problems of the
95: self-accelerating branch of DGP have been recently uncovered. This
96: branch suffers from pathologies related to the appearance of ghost
97: degrees of freedom
98: \cite{LuPR03,NicR04,Ko04,GoKS06,ChGNP06,DefGI06,Dv06,KoyS07}. Ghosts
99: can lead to runaway excitations when coupled to normal modes and
100: their existence can invalidate the self-accelerating background
101: solution as well as linear perturbations around it.
102: 
103: 
104: Observational problems also arise if one posits that ghosts and strong
105: coupling do not invalidate the idea of self-acceleration itself, {\it
106: i.e.}~that our Hubble volume behaves in a manner that is perturbatively
107: close on scales near the horizon to the modified Friedmann equation
108: specified on this branch.  These problems fall into two classes, those
109: due to the background expansion history and those due to the growth of
110: structure during the acceleration epoch.
111: 
112: In a flat spatial geometry the DGP model adds only one degree of freedom
113: to fit acceleration, the crossover scale.  Like $\Lambda$CDM, the extra
114: degree of freedom can be phrased as the matter density relative to the
115: critical density.  Remarkably, with this one parameter, the $\Lambda$CDM
116: model can fit three disparate sets of distance measures: the local
117: Hubble constant and baryon acoustic oscillations, the relative distances
118: to high-redshift supernovae (SNe), and the acoustic peaks in the cosmic
119: microwave background (CMB).  The flat DGP model cannot fit these
120: observations simultaneously and the significance of the discrepancy
121: continues to grow.  The addition of spatial curvature can help alleviate
122: some, but not all, of this tension \cite{FaG06,MaM06,SoSH07}.
123: 
124: On intermediate scales that encompass measurements of the large scale
125: structure of the Universe, the early onset of modifications to the
126: expansion also imply a substantial reduction in the linear growth of
127: structure which is itself slightly reduced by the modification to
128: gravity \cite{LuSS04,KoM06}.  If extended to the mildly nonlinear regime
129: where strong coupling effects may alter growth, this reduction is in
130: substantial conflict with weak lensing data \cite{WaHMH07}.
131: 
132: On scales approaching the crossover scale, which are probed by the
133: CMB, the unique modification to gravity in the DGP model itself
134: strongly alters the growth rate.  Here the propagation of
135: perturbations into the bulk requires a ($4+1$)-dimensional
136: perturbation framework \cite{De02}.  The approximate iterative
137: solutions introduced in Rfn.~\cite{SaSH07} have been recently
138: verified to be sufficiently accurate by a more direct calculation
139: \cite{CaKSS08} but still are computationally too expensive for Monte
140: Carlo explorations of the DGP parameter space.  More recently, a
141: ($3+1$)-dimensional effective approach dubbed the parameterized
142: post-Friedmann (PPF) framework \cite{Hu08,HuS07} has been developed
143: that accurately encapsulates modified gravity effects with a closed
144: effective dark energy system.  The PPF approach enables standard
145: cosmological tools such as an Einstein-Boltzmann linear theory
146: solver to be applied to the DGP model.
147: 
148: In this {\it Paper}, we implement the PPF approach to DGP and
149: conduct a thorough study of the tension between CMB distance, energy
150: density, and growth measures, along with the SNe and local distance
151: measures, across an extended DGP parameter space.  We find that even
152: adding epicycles to the DGP model does not significantly improve the
153: agreement with the data. Curvature, while able to alleviate the
154: problem with distances, exacerbates the problem with growth.
155: Changing the initial power spectrum to remove excess power in the
156: temperature spectrum destroys the agreement with recent polarization
157: measurements from the five-year Wilkinson Microwave Anisotropy Probe
158: (WMAP) \cite{Noetal08}.
159: 
160: The outline of the paper is as follows.  In \S \ref{sec:theory} we
161: review the impact of the DGP modifications on distance measures and
162: the growth of structure emphasizing an effective dark energy PPF
163: approach that is detailed in Appendix A and compared with a
164: growth-geometry splitting approach in Appendix B.  We present the
165: results of the likelihood analysis in \S \ref{sec:results} and
166: discuss these results in \S \ref{sec:discuss}.
167: 
168: %
169: \section{DGP Distance and Growth Phenomenology}
170: \label{sec:theory}
171: 
172: In the DGP model, gravity remains a metric theory on the brane in a
173: background that is statistically homogeneous and isotropic.  Its
174: modifications therefore are confined to the field equations that relate
175: the metric to the matter.  Once the metric is obtained, all the usual
176: implications for the propagation of light from distant sources and the
177: motion of matter remain unchanged.  In this section, we review the DGP
178: modifications to the background metric, or expansion history, and the
179: gravitational potentials, or linear metric perturbations.  We cast these
180: modifications in the language of an effective dark energy contribution
181: under ordinary gravity \cite{Hu08,HuS07,Ba07,KuS07,CaCM07,JaZ07,BeZ08}.
182: 
183: %
184: \subsection{Background Evolution}
185: \label{subsec:background}
186: 
187: That the DGP model is a metric theory in a statistically homogeneous
188: and isotropic universe imposes a background
189: Friedmann-Robertson-Walker (FRW) metric on the brane.  The FRW
190: metric is specified by two quantities, the evolution of the scale
191: factor $a$ and the spatial curvature $K$.  The DGP model modifies
192: the field equation, {\it i.e.}~the Friedmann equation, relating the
193: evolution of the scale factor $H=a^{-1} da/dt$ to the matter-energy
194: content.  On the self-accelerating branch it becomes (see, {\it
195: e.g.}, Rfns.~\cite{De01,DeLRZA02})
196: %
197: \begin{equation}
198: H^2  = \left(\sqrt{\frac{8 \pi G}{3}\sum_i
199: \rho_i+\frac{1}{4r_c^2}}+\frac{1}{2r_c}\right)^2-\frac{K}{a^2},
200: \label{eqn:modfriedmann}
201: \end{equation}
202: %
203: where $r_c$ is the crossover scale and subscript $i$ labels the true
204: matter-energy components of the universe to be distinguished below from
205: the effective dark energy contribution.
206: 
207: On the other hand, given the metric the matter evolves in the same way
208: as in ordinary gravity
209: %
210: \begin{equation}
211: \dot{\rho}_i=-3aH(\rho_i+P_i).
212: \label{eqn:densityevolution}
213: \end{equation}
214: %
215: Overdots represent derivatives with respect to conformal time $\eta
216: = \int dt/a$.  From these relations, one sees that as $a \rightarrow
217: \infty$, $H \rightarrow r_c^{-1}$ and the Universe enters a de Sitter
218: phase of accelerated expansion.
219: 
220: In the limit that $r_c \rightarrow \infty$ the ordinary Friedmann
221: equation is recovered.  The effect of a finite $r_c$ compared with the
222: Hubble scale can be encapsulated in a dimensionless parameter
223: $\Omega_{r_c}$ much like the usual contributions of the density
224: $\Omega_i = 8\pi G\rho_i/3H_0^2$ and curvature $\Omega_K = -K/H_0^2$ to
225: the expansion rate
226: %
227: \begin{equation}
228: \Omega_{r_c} \equiv \frac{1}{4r_c^2H_0^2} .
229: \end{equation}
230: %
231: The modified Friedmann equation (\ref{eqn:modfriedmann}) today then
232: becomes the constraint equation
233: %
234: \begin{equation}
235: 1=\left(\sqrt{\Omega_{r_c}} +\sqrt{\Omega_{r_c}+\sum_i
236: \Omega_i}~\right)^2+\Omega_K.
237: \label{eqn:normDGP}
238: \end{equation}
239: %
240: 
241: It is convenient and instructive to recast the impact of $r_c$ as an
242: effective dark energy component.  With the same background evolution,
243: the effective dark energy will have an energy density of
244: %
245: \begin{equation}
246: \rho_e \equiv \frac{3}{8 \pi G} \left(H^2 +\frac{K}{a^2}\right)-
247: \sum_i \rho_i.
248: \end{equation}
249: %
250: Conservation of its energy-momentum tensor is guaranteed by the
251: Bianchi identities and requires its ``equation of state'' $w_e
252: \equiv P_e/\rho_e$ to be given by Eq.~(\ref{eqn:densityevolution})
253: %
254: \begin{equation}
255: w_e=\frac{\sum_i(\rho_i+P_i)}{3(H^2 + a^{-2}K)/8\pi G + \sum_i
256: \rho_i}-1.
257: \label{eqn:we}
258: \end{equation}
259: %
260: If we define $\Omega_e$ in the same way as $\Omega_i$, the usual
261: constraint condition applies $\sum_i\Omega_i +\Omega_e + \Omega_K=1$.
262: Comparing it to Eq.~(\ref{eqn:normDGP}), we obtain the following
263: relationship between $\Omega_e$ and $\Omega_{r_c}$
264: %
265: \begin{equation}
266: \Omega_e=2\sqrt{\Omega_{r_c}(1-\Omega_K)}.
267: \label{eqn:omegae}
268: \end{equation}
269: %
270: 
271: Given $w_e$ and $\Omega_e$, we can now describe the background evolution
272: of the DGP cosmology by using the ordinary Friedmann equation for $H$.
273: Likewise the Hubble parameter specifies the comoving radial distance
274: %
275: \begin{equation}
276: D(z)=\int_0^z\frac{dz'}{H(z')}\,,
277: \end{equation}
278: %
279: and the luminosity distance
280: %
281: \begin{equation}
282: d_L(z)=\frac{(1+z)}{\sqrt{-\Omega_K}H_0}\sin{\left(\sqrt{-\Omega_K}H_0\,
283: D(z)\right)}\,,
284: \label{eqn:luminositydistance}
285: \end{equation}
286: %
287: as usual.
288: 
289: The effective dark energy for DGP has quite a different equation of
290: state from that of a cosmological constant. For example, when there
291: is no curvature, $w_e$ starts at ${-1/(1+\Omega_m)}$ at the present,
292: approaches $-1/2$ in the matter dominated regime and $-1/3$ in the
293: radiation dominated regime (see Fig.~\ref{fig:we}).
294: 
295: With the same values of parameters, DGP will have a larger amount of
296: effective dark energy at a given redshift as $w_e$ is always larger than
297: $-1$, and the universe will expand at a larger rate than in
298: $\Lambda$CDM.  This reduces the absolute comoving radial distance to a
299: distant source.  For relative distance measures like the SNe, the
300: reduction in $H_0 d_L(z)$ can be compensated by lowering the {\it
301: fractional} contribution of matter through $\Omega_m$.  The same is not
302: true for absolute distances, such as those measured by the CMB, baryon
303: oscillations, and Hubble constant, if the overall matter contribution to
304: the expansion rate $\Omega_m H_0^2 a^{-3}$ remains fixed.
305: 
306: %
307: \begin{figure}[t!]
308: \resizebox{90mm}{!}{\includegraphics{weff.eps}}
309: \caption{Equation of state of the effective dark energy $w_e$ for the
310: self-accelerating DGP model with $\Omega_m=0.26$ and $\Omega_K=0$.}
311: \label{fig:we}
312: \end{figure}
313: %
314: 
315: 
316: %
317: \subsection{Structure Formation in the Linear Regime}
318: 
319: The same methodology of introducing an effective dark energy component
320: for the background expansion history applies as well to the linear
321: metric perturbations that govern the evolution of large scale structure.
322: 
323: To define the effective dark energy, one must first parameterize the
324: solutions to the ($4+1$)-dimensional equations involving metric
325: perturbations in the brane as well as the bulk \cite{De02}.  Three
326: simplifications aid in this parameterization \cite{HuS07}.  The first is
327: that at high redshift the effect of $r_c$ and the extra dimension goes
328: away rapidly [see Eq.~(\ref{eqn:modfriedmann})].  The parameterization
329: needs only to be accurate between the matter dominated regime and the
330: present.
331: 
332: Likewise, the predictions of $\Lambda$CDM for the relationships
333: between the matter and baryon density at recombination and
334: morphology of the CMB acoustic peaks remain unchanged.  As a
335: consequence, the shape of the CMB acoustic peaks still constrains
336: the physical cold dark matter and baryon density $\Omega_c h^2$ and
337: $\Omega_b h^2$ as usual.  With these as fundamental high redshift
338: parameters, in a flat spatial geometry, the DGP degree of freedom
339: $r_c$ is then specified by either
340: $\Omega_m=1-\Omega_e=\Omega_c+\Omega_b$ or $H_0$.  We shall see
341: below that in a flat universe the competing requirements of CMB and
342: SNe distance measures on $r_c$ will slightly shift the values of
343: $\Omega_c h^2$ and $\Omega_b h^2$ to reach a compromise at the
344: expense of the goodness of fit.
345: 
346: The second simplification is that well below the horizon, perturbations
347: on the brane are in the quasi-static regime where time derivatives can
348: be neglected in comparison with spatial gradients and propagation
349: effects into the bulk are negligible.  This allows the equations to be
350: simply closed on the brane by a modified Poisson equation (\ref{eqn:qs})
351: that can be recast as arising from the anisotropic stress of the
352: effective dark energy \cite{LuSS04,KoM06}.
353: 
354: The final simplification is that on scales above the horizon, the impact
355: of the bulk perturbations on the brane becomes scale free and depends
356: only on time \cite{SaSH07} through dimensionless combinations of $H$,
357: $r_c$ and $K$.  Furthermore on these scales, generic modifications to
358: gravity are fully defined by the Friedmann equation and the anisotropic
359: stress of the effective dark energy \cite{HuE98,Be06}.
360: 
361: As shown in Rfn.~\cite{HuS07} and detailed in Appendix A,
362: interpolation between these two limits leads to a simple PPF
363: parameterization of DGP on all linear scales.  This parameterization
364: has been verified to be accurate at a level substantially better
365: than required by cosmic variance by a direct computation of bulk
366: perturbations \cite{CaKSS08}.  With such a parameterization,
367: efficient Einstein-Boltzmann codes such as CAMB \cite{camb} can be
368: modified to calculate the full range of CMB anisotropy, as is done
369: in this paper.
370: 
371: The result of this calculation is that compared with $\Lambda$CDM,
372: the growth of structure in the DGP model during the acceleration
373: epoch is suppressed.  In the quasi-static regime, this suppression
374: is mostly due to the higher redshift extent of the acceleration
375: epoch discussed in the previous section (see also Fig.~\ref{fig:we})
376: with a small component from the effective anisotropic stress or
377: modification to the Poisson equation \cite{LuSS04,KoM06}.  On scales
378: approaching $r_c$, the effect of the anisotropic stress becomes much
379: more substantial due to the perturbations propagating into the bulk
380: \cite{SaSH07} resulting in an even stronger integrated Sachs-Wolfe
381: (ISW) effect in the CMB anisotropy power at the lowest multipoles
382: \cite{SoSH07}.
383: 
384: %
385: \section{Constraints from Current Observations}
386: \label{sec:results}
387: 
388: In this section, we employ Markov Chain Monte Carlo (MCMC) techniques to
389: explore constraints on the DGP parameter space from current
390: observations, and compare them with the successful $\Lambda$CDM model.
391: The data we use are: the Supernovae Legacy Survey (SNLS) \cite{SNLS},
392: the CMB anisotropy data from the five-year WMAP \cite{WMAP5} for both
393: temperature and polarization (TT + EE + TE), and the Hubble constant
394: measurement from the Hubble Space Telescope (HST) Key Project \cite{H0}.
395: 
396: We use the public MCMC package \textit{CosmoMC} \cite{cosmomc}, with a
397: modified version of CAMB for DGP described in Appendix A, to sample
398: the posterior probability distributions of model parameters.  The MCMC
399: technique employs the Metropolis-Hastings algorithm \cite{Metropolis53,
400: Hastings70} for the sampling, and the Gelman and Rubin $R$ statistic
401: \cite{GeR92} for the convergence test. We conservatively require
402: $R-1<0.01$ for the eight chains we run for each model, and this
403: generally gave us $\sim 5000$ independent samples.
404: 
405: The SNe magnitude-redshift relation probes the relative luminosity
406: distance between the low and high redshift sample.  The luminosity
407: distance itself is completely determined by the background expansion
408: history through Eq.~(\ref{eqn:luminositydistance}).  On the other
409: hand the power spectra of the CMB anisotropy probes not only the
410: background expansion, but also the growth of structure.  To better
411: separate the two types of information, we also consider a canonical
412: scalar field (``quintessence'' or QCDM) model with the same
413: expansion history as DGP. This model is defined by the equation of
414: state parameter $w_e$ and density parameter $\Omega_e$ in
415: Eqs.~(\ref{eqn:we}) and (\ref{eqn:omegae}).
416: 
417: With only scalar perturbations in consideration, our basic parameter
418: set is chosen to be $\{\Omega_b h^2, \Omega_c h^2, \theta_s, \tau,
419: n_s, A_s\}$, which in turn stand for the density parameters of
420: baryons and cold dark matter, angular size of the sound horizon at
421: recombination, optical depth from reionization
422: (assumed to be instantaneous), spectra index of the primordial
423: curvature fluctuation and its amplitude at $k_*=0.002$ Mpc$^{-1}$,
424: {\it i.e.}, $\Delta_{\zeta}^2=A_s(k/k_*)^{n_s-1}$.  Also, we follow
425: Rfn.~\cite{WMAP5}, and include $A_{SZ}$, with flat prior of
426: $0<A_{SZ}<2$, to account for the contributions to the CMB power
427: spectra from Sunyaev-Zeldovich fluctuations. The lensing effect on
428: the CMB is neglected. Note that in all the three models we
429: considered, {\it i.e.}~self-accelerating DGP, $\Lambda$CDM and QCDM,
430: we have the same parameter sets and priors except that we restrict
431: the DGP parameter space to $H_0 r_c>1.08$ so that metric
432: fluctuations remain well behaved [see Eq.~(\ref{eqn:gsh})]. We
433: apply this prior to the QCDM model as well for a fair comparison. In
434: practice, the excluded models are strongly disfavored by the data
435: and the prior is only necessary for numerical reasons.
436: %\wf{ For reference, here are all the priors: $0\leq A_{SZ}
437: %\leq 2, 0.1\leq \ln{(10^{10} A_s)}\leq 4, 0.1\leq n_s \leq 2, -0.3\leq
438: %\Omega_K \leq 0.3, 0.01\leq \tau \leq 0.8, 0.5 \leq \theta_s \leq 10,
439: %0.01 \leq \Omega_c h^2\leq 0.99, 0.01 \leq \Omega_b h^2 \leq0.2$ (base
440: %params); $40\leq H_0 \leq 100, 10\leq Age \leq 20, \Omega_e \geq 0,
441: %z_{ri}\geq 0$(derived params), additionally $H_0r_c>1.08$ for DGP and QCDM.}
442: 
443: %
444: \subsection{Flat Models}
445: 
446: We start with the minimal parameterization of a flat universe with
447: scale free initial conditions.  In this case the three model classes
448: $\Lambda$CDM, DGP and QCDM all have only one parameter that
449: describes acceleration.  In the chain parameters this is $\theta_s$
450: but can be equivalently defined as the derived parameters $H_0$ or
451: $\Omega_m$. The constraints on the three model classes are given in
452: Table \ref{tab:fmean} for the means and marginalized errors on
453: various parameters. Table \ref{tab:fbf} shows the best-fit values of
454: the parameters and the corresponding likelihoods, which serve as a
455: ``goodness of fit'' criterion.
456: 
457: %
458: \begin{table}
459: \caption{\label{tab:fmean}Mean and marginalized errors for various
460: parameters of the self-accelerating DGP, QCDM with the same expansion
461: history as the DGP and $\Lambda$CDM models from SNLS + WMAP5 + HST, {\it
462: assuming a flat universe}. The first 6 parameters are directly varied
463: when running the Markov Chains, while the others are derived parameters,
464: as are in the following tables.}
465: \begin{tabular}{|c |c |c |c |}
466: \hline
467: parameters &  DGP  &  QCDM &  $\Lambda$CDM \\
468: \hline
469: $100\Omega_b h^2$    &2.36$\pm$0.07    &2.32$\pm$0.07    &2.25$\pm$0.06   \\
470: $\Omega_c h^2$       &0.090$\pm$0.005  &0.090$\pm$0.005  &0.109$\pm$0.005 \\
471: $100\theta_s$        &1.042$\pm$0.003  &1.041$\pm$0.003  &1.040$\pm$0.003 \\
472: $\tau$               &0.10$\pm$0.02    &0.10$\pm$0.02    &0.09$\pm$0.02   \\
473: $n_s$                &1.00$\pm$0.02    &0.99$\pm$0.02    &0.96$\pm$0.01   \\
474: $\ln{[10^{10}A_s]}$  &3.02$\pm$0.05    &3.05$\pm$0.04    &3.18$\pm$0.04   \\
475: \hline
476: $H_0$                &$66\pm2$         &65$\pm$2         &72$\pm$2        \\
477: $\Omega_m$           &$0.26\pm0.02$    &0.26$\pm$0.02    &0.26$\pm$0.02   \\
478: $\Omega_{r_c}$       &$0.136\pm0.009$  &..               &..              \\
479: \hline
480: \end{tabular}
481: \end{table}
482: %
483: 
484: %
485: \begin{table}
486: \caption{\label{tab:fbf}Parameters and the likelihood values at the
487: best-fit point of the self-accelerating DGP, QCDM with the same
488: expansion history as the DGP and $\Lambda$CDM models fitting to SNLS +
489: WMAP5 + HST, {\it assuming a flat universe}.}
490: \begin{tabular}{|c |c |c |c |}
491: \hline
492: parameters &  DGP  &  QCDM &  $\Lambda$CDM \\
493: \hline
494: $100\Omega_b h^2$     &2.37    &2.32    &2.26   \\
495: $\Omega_c h^2$        &0.0888  &0.0907  &0.110  \\
496: $100\theta_s$         &1.04    &1.04    &1.04   \\
497: $\tau$                &0.0954  &0.0998  &0.0825 \\
498: $n_s$                 &0.998   &0.983   &0.959  \\
499: $\ln {[10^{10}A_s]}$  &3.01    &3.06    &3.18   \\
500: \hline
501: $H_0$                 &66.0    &65.1    &71.6   \\
502: $\Omega_m$            &0.258   &0.269   &0.258  \\
503: $\Omega_{r_c}$        &0.138   &..      &..     \\
504: \hline
505: $-2\ln L$             &2805.8  &2797.6  &2777.8 \\
506: \hline
507: \end{tabular}
508: \end{table}
509: %
510: 
511: First, we compare the constraints on the QCDM model with those on
512: $\Lambda$CDM. The differences are expected to reflect those between
513: the background expansion histories of the DGP and $\Lambda$CDM
514: models. In spite of the clustering effects on the largest scales of
515: a quintessence dark energy, the difference between QCDM and
516: $\Lambda$CDM is completely encoded in the equation of state of their
517: dark energy components given the fixed sound speed of quintessence.
518: 
519: Constraints from the SNe magnitudes come from the dimensionless
520: luminosity distance $H_0 d_L(z)$ [see
521: Eq.~(\ref{eqn:luminositydistance})], once the unknown absolute
522: magnitude is marginalized.  In order to match the predictions for
523: $H_0 d_L(z)$ of a flat $\Lambda$CDM model, we would expect that the
524: QCDM model has a smaller $\Omega_m$ to compensate the larger $w_e$
525: its dark energy has.  However, lowering $\Omega_m$ also shortens
526: more of the distance to the last scattering surface and hence
527: increases the angular size of the sound horizon (see \S
528: \ref{subsec:background}).
529: 
530: The physical scale of the sound horizon can partially compensate and
531: is controlled by $\Omega_c h^2$ and $\Omega_b h^2$ but these
532: parameters are also well measured by the shape of the peaks and can
533: only be slightly adjusted at a cost to the goodness of fit.  Thus
534: the parameter ranges in Table \ref{tab:fmean} for $\Omega_c h^2$
535: decrease and $\Omega_b h^2$ increase slightly which both have the
536: effect of decreasing the angular size of the horizon while
537: $\Omega_m$ remains nearly unchanged.  This compromise between the
538: energy density and distance constraints results in tension between
539: the CMB and SNe data.  This tension shows up as a difference between
540: the $-2\ln L$ values of these two models: $2\ln L(\Lambda{\rm
541: CDM})-2\ln L({\rm QCDM})\simeq 20$.
542: 
543: The QCDM model also favors a larger $n_s$ and a smaller $A_s$.  This
544: is a consequence of the larger ISW effect in the QCDM model due to a
545: slower growth rate.  Tilting the spectrum can compensate for the
546: excess power in the low-$\ell$ modes (as shown by the
547: near-coincidence of the short-dashed and dashed curves in
548: Fig.~\ref{fig:fcmpTT}).
549: 
550: %
551: \begin{figure}[t!]
552: \resizebox{90mm}{!}{\includegraphics{f_cmpTT.eps}}
553: \caption{\label{fig:fcmpTT} Predictions for the power spectra of the
554: CMB temperature anisotropies $C_{\ell}^{\rm TT}$ of the best-fit DGP
555: (solid), QCDM with the same expansion history as DGP (short-dashed),
556: and $\Lambda$CDM (dashed, coincident with QCDM at low $\ell$) models
557: obtained by fitting to SNLS + WMAP5 (both temperature and
558: polarization) + HST, {\it assuming a flat universe}. Bands represent
559: the 68\% and 95\% cosmic variance regions for the DGP model.  Points
560: represent WMAP5 measurements; note that noise dominates over cosmic
561: variance for $\ell \gtrsim 500$.}
562: \end{figure}
563: %
564: 
565: Next we compare the constraints on the DGP model with those on QCDM.
566: Since the two models have the same expansion history, the differences
567: are entirely caused by the differing growth rates.  Due to the
568: propagation of perturbations into the bulk for scales near $r_c$ and the
569: opposite effect of dark energy clustering in QCDM, there is a
570: substantially stronger ISW effect in the first few multipoles of the CMB
571: anisotropy power \cite{SoSH07}.  Since this effect is only important on
572: the largest angular modes in the CMB TT power spectra which is further
573: limited by the large cosmic variance, the parameter ranges for these two
574: models do not differ significantly.  Nonetheless from
575: Fig.~\ref{fig:fcmpTT}, it is clear that the best-fit DGP model over
576: predicts the low-$\ell$ modes anisotropy, though as before, $n_s$ and
577: $A_s$ adjustments try to reduce the primordial perturbations on large
578: scales.  This leads to DGP being an even worse fit than QCDM with
579: $2\ln L({\rm QCDM})-2\ln L({\rm DGP})\simeq 8$.
580: 
581: When DGP is compared to $\Lambda$CDM, this brings the change in $-2\ln
582: L$ for the maximum likelihood parameters to $\simeq 28$, where $\sim
583: 70\%$ is driven by the background expansion, while $\sim 30\%$ by the
584: dynamical effects on structure growth.
585: 
586: %
587: \subsection{Adding in Curvature}
588: \label{subsec:omk}
589: 
590: From our analysis in the above section, the flat DGP model is a poor
591: fit to the current observations mostly because it cannot
592: simultaneously satisfy the geometrical requirements of the relative
593: luminosity distances of the SNe and the angular size of the sound
594: horizon at recombination with a single parameter.  Since curvature
595: has more of an effect on high redshift distance measures, the
596: tension in the distance measures can be alleviated by including
597: $\Omega_K$ in the parameter space \cite{MaM06,SoSH07}.  Our results
598: are given in Table \ref{tab:nfmean} and Table \ref{tab:nfbf}.
599: 
600: %
601: \begin{table}
602: \caption{\label{tab:nfmean} Mean and marginalized errors for various
603: parameters of the self-accelerating DGP, QCDM with the same expansion
604: history as the DGP and $\Lambda$CDM models from SNLS + WMAP5 + HST, {\it
605: allowing curvature}.}
606: \begin{tabular}{|c |c |c |c |}
607: \hline
608: parameters &  DGP  &  QCDM &  $\Lambda$CDM \\
609: \hline
610: $100\Omega_b h^2$     &2.37$\pm$0.07    &2.34$\pm$0.07    &2.25$\pm$0.06    \\
611: $\Omega_c h^2$        &0.096$\pm$0.006  &0.098$\pm$0.006  &0.108$\pm$0.006  \\
612: $100\theta_s$         &1.043$\pm$0.003  &1.042$\pm$0.003  &1.040$\pm$0.003  \\
613: $\tau$                &0.09$\pm$0.02    &0.09$\pm$0.02    &0.09$\pm$0.02    \\
614: $\Omega_K$            &0.019$\pm$0.008  &0.027$\pm$0.008  &-0.004$\pm$0.009 \\
615: $n_s$                 &1.00$\pm$0.02    &0.99$\pm$0.02    &0.96$\pm$0.01    \\
616: $\ln {[10^{10}A_s]}$  &3.02$\pm$0.05    &3.06$\pm$0.04    &3.18$\pm$0.04    \\
617: \hline
618: $H_0$                 &74$\pm$4         &77$\pm$5         &70$\pm$4         \\
619: $\Omega_m$            &0.22$\pm$0.03    &0.20$\pm$0.03    &0.27$\pm$0.03    \\
620: $\Omega_{r_c}$        &0.15$\pm$0.01    &..               &..               \\
621: \hline
622: \end{tabular}
623: \end{table}
624: %
625: 
626: %
627: \begin{table}
628: \caption{\label{tab:nfbf} Parameters and likelihood values at the
629: best-fit point of the self-accelerating DGP, QCDM with the same
630: expansion history as the DGP and $\Lambda$CDM models fitting to SNLS +
631: WMAP5 + HST, {\it allowing curvature}.}
632: \begin{tabular}{|c |c |c |c |}
633: \hline
634: parameters &  DGP  &  QCDM &  $\Lambda$CDM \\
635: \hline
636: $100\Omega_b h^2$     &2.38    &2.36    &2.27     \\
637: $\Omega_c h^2$        &0.0937  &0.0960  &0.107    \\
638: $100\theta_s$         &1.04    &1.04    &1.04     \\
639: $\tau$                &0.0887  &0.0914  &0.0884   \\
640: $\Omega_K$            &0.0189  &0.0268  &-0.00553 \\
641: $n_s$                 &0.996   &0.992   &0.959    \\
642: $\ln {[10^{10}A_s]}$  &3.02    &3.05    &3.18     \\
643: \hline
644: $H_0$                 &73.8    &78.3    &69.8     \\
645: $\Omega_m$            &0.216   &0.195   &0.266    \\
646: $\Omega_{r_c}$        &0.149   &..      &..       \\
647: \hline
648: $-2\ln L$             &2800.8  &2787.2  &2777.5   \\
649: \hline
650: \end{tabular}
651: \end{table}
652: %
653: 
654: With curvature, the $-2\ln L$ of the maximum likelihood $\Lambda$CDM
655: model almost has no improvement. This is consistent with the results
656: of Rfn.~\cite{Ko08}, who found strong limits on curvature in
657: $\Lambda$CDM, by fitting WMAP5 data combined with SNe or HST.  As
658: expected the maximum likelihood model in the QCDM space improves in
659: $-2\ln L$ by $\sim10$. The QCDM model needs an open universe to
660: increase the distance to last scattering to compensate the smaller
661: $\Omega_m$, consistent with the findings by \cite{MaM06}.  Even with
662: this additional freedom, distance measures remain in slight tension
663: due to the Hubble constant since lowering $\Omega_m$ with $\Omega_m
664: h^2$ well determined by the CMB implies a higher Hubble constant
665: \cite{SoSH07}.  The allowed amount of this shift is limited by the
666: HST Key Project constraint of $H_0=72 \pm 8$ km s$^{-1}$ Mpc$^{-1}$.
667: For QCDM the baryon acoustic oscillation constraint would already
668: disfavor such a shift \cite{detection,hutsib,percival,gch08} but its
669: application to DGP requires cosmological simulations of the strong
670: coupling regime. Note that with the distance tension partially
671: removed, the shifts in $\Omega_c h^2$ and $\Omega_b h^2$ are
672: reduced.
673: 
674: The lowering of $\Omega_m$, in addition to the large equation of state
675: parameter of the quintessence, also causes matter domination to
676: terminate at an earlier redshift, and leads to a stronger ISW effect in
677: the QCDM model, as can be seen in Fig.~\ref{fig:nfcmpTT}, again with a
678: partial compensation from $n_s$ and $A_s$.  The net difference of $-2\ln
679: L$ values of the QCDM model compared to $\Lambda$CDM is $\simeq 10$,
680: 50\% smaller than before but still a significantly poorer fit.
681: 
682: %
683: \begin{figure}[t!]
684: \resizebox{90mm}{!}{\includegraphics{nf_cmpTT.eps}}
685: \caption{\label{fig:nfcmpTT} Predictions for the power spectra of the
686: CMB temperature anisotropies $C_{\ell}^{\rm TT}$ of the best-fit DGP
687: (solid), QCDM with the same expansion history as DGP (short-dashed), and
688: $\Lambda$CDM (dashed) models obtained by fitting to SNLS + WMAP5 (both
689: temperature and polarization) + HST, {\it allowing curvature}.}
690: \end{figure}
691: %
692: 
693: The situation is even worse for DGP.  Here the enhancement of the ISW
694: effect at low $\Omega_m$ is even more substantial.  Thus the mean value
695: of $\Omega_K$ is smaller than the optimal one for the distance
696: constraints in QCDM.  Even adjusting the other parameters to give the
697: maximum likelihood model shown in Fig.~\ref{fig:nfcmpTT}, the poor fit
698: is noticeable at the low multipoles.  For example the probability of
699: obtaining a quadrupole as extreme as the observations from the DGP
700: maximum likelihood model is $\sim 1\%$ compared with $\sim 6\%$ for the
701: $\Lambda$CDM maximum likelihood model.  The net difference in $-2\ln L$
702: by including curvature as a parameter is only $\sim 5$ in DGP showing
703: only marginal evidence for curvature in the model at best.
704: 
705: Moreover, the difference from $\Lambda$CDM remains substantial with
706: $-2\Delta\ln L \simeq 23$.  Since the difference of $-2\Delta\ln L$ from
707: QCDM can be attributed to the ISW effect, $\sim 40\%$ of this difference
708: is driven by the background expansion, and $\sim 60\%$ by the
709: dynamical effects on structure growth.
710: 
711: %
712: \subsection{Changing the Initial Power}
713: 
714: The adjustments of $A_s$ and $n_s$ in the above examples suggest that
715: perhaps a more radical change in the initial power spectrum can bring
716: DGP back in agreement with the data.  For example, one can sharply
717: reduce large scale power in the temperature spectrum by cutting off the
718: initial power spectrum on large scales, below a wavenumber $k_{\rm
719: min}$.  While this is a radical modification and has no particular
720: physical motivation, it is useful to check whether such a loss of power
721: could satisfy the joint temperature and polarization constraints.
722: 
723: CMB polarization at these scales arises from reionization which
724: occurs at a substantially higher redshift than DGP modifications
725: affect.  A finite polarization requires not only ionization but also
726: large scale anisotropy at this epoch.  Eliminating initial power on
727: these scales eliminates the polarization as well.  The EE power in
728: the low multipoles has now been measured at 4-$5\sigma$ level
729: \cite{Noetal08} leading to a significant discrepancy if the large
730: scale power is removed in the model.
731: 
732: Given that including this parameter does not improve the fit, instead of
733: adding it to the MCMC parameter space, we illustrate its effects on the
734: maximum likelihood DGP model in \S \ref{subsec:omk}.
735: 
736: By maximizing the likelihood for this model to the TT power alone,
737: we find the best agreement is obtained at $k_{\rm min}=8\times
738: 10^{-4}$ Mpc$^{-1}$, with $-2\Delta \ln L^{\rm TT} \simeq -12$
739: compared to $k_{\rm min}=0$. Model predictions with this $k_{\rm
740: min}$ are plotted in Fig.~\ref{fig:TTtrunc}, together with the WMAP
741: 5 year data.
742: 
743: With this power truncation the over prediction problem for the
744: low-$\ell$ TT power is alleviated, but the EE polarization power on
745: large angular scales is significantly reduced (see
746: Fig.~\ref{fig:EEtrunc}).  When the polarization data of TE + EE are
747: included, we find $-2\Delta \ln L^{\rm all} \simeq 6$ at $k_{\rm
748: min}=8\times 10^{-4}$ Mpc$^{-1}$, and when $k_{\rm min}$ is varied,
749: the combined data does not favor any positive values of $k_{\rm
750: min}$ as shown in Fig.~~\ref{fig:Ltrunc}.  We conclude that though
751: the DGP model can be made a better fit to the TT power spectrum with
752: a large scale cut-off, polarization measurements are now
753: sufficiently strong to rule out this possibility.  While we have
754: only included instantaneous reionization models, changing the
755: ionization history to have an extended high redshift tail can only
756: exacerbate this problem by increasing EE power in the $\ell \sim
757: 10-30$ regime \cite{MorH08}.
758: 
759: %
760: \begin{figure}[t!]
761: \resizebox{90mm}{!}{\includegraphics{TTtrunc.eps}}
762: \caption{\label{fig:TTtrunc} Predictions for the power spectra of the
763: CMB temperature anisotropies $C_{\ell}^{\rm TT}$ of the best-fit DGP
764: model as found in \S \ref{subsec:omk} without cutting off any large
765: scale primordial perturbations (solid) and with a cut-off scale of
766: $k_{\rm min}=8\times 10^{-4}$ Mpc$^{-1}$ (dotted) -- the best-fit scale
767: obtained when fitting to the WMAP 5 year TT data alone, while all other
768: parameters are fixed at their best-fit values with $k_{\rm min}=0$.}
769: \end{figure}
770: %
771: 
772: %
773: \begin{figure}[t!]
774: \resizebox{90mm}{!}{\includegraphics{EEtrunc.eps}}
775: \caption{\label{fig:EEtrunc} Predictions for the power spectra of the
776: CMB E-mode polarization $C_{\ell}^{\rm EE}$ of the best-fit DGP model as
777: found in \S \ref{subsec:omk} without cutting off any large scale
778: primordial perturbations (solid) and with a cut-off scale of $k_{\rm
779: min}=8\times 10^{-4}$ Mpc$^{-1}$ (dotted) -- the best-fit scale obtained
780: when fitting to the WMAP 5 year TT data alone, while all other
781: parameters are fixed their best-fit values.  Note here, according to
782: Rfn.~\cite{Noetal08}, the reionization feature at the lowest-$\ell$
783: modes is preferred by the data through $\Delta \chi^2=19.6$.}
784: \end{figure}
785: %
786: 
787: %
788: \begin{figure}[t!]
789: \resizebox{90mm}{!}{\includegraphics{Ltrunc.eps}}
790: \caption{\label{fig:Ltrunc} The total log-likelihood of TT + TE + EE
791: as a function of the cut-off scale $k_{\rm min}$ of the primordial
792: perturbations, shown as the difference from its value at $k_{\rm
793: min}=0$.  We find that when polarization is included, the combined
794: data does not favor any positive values of $k_{\rm min}$. The local
795: minimum shows up at the scale of $k_{\rm min}=8\times 10^{-4}$
796: Mpc$^{-1}$, which is the one favored by the TT data alone. Note all
797: other parameters are fixed at their best-fit values of the DGP model
798: as found in \S \ref{subsec:omk}.}
799: \end{figure}
800: %
801: 
802: %
803: \section{Discussion}
804: \label{sec:discuss}
805: 
806: We have conducted a thorough Markov Chain Monte Carlo likelihood study
807: of the parameter space available to the DGP self-accelerating braneworld
808: scenario given CMB, SNe and Hubble constant data.  To carry out
809: this study, we have introduced techniques for characterizing modified
810: gravity and non-canonical dark energy candidates with the public
811: Einstein-Boltzmann code CAMB that are of interest beyond the DGP
812: calculations themselves.
813: 
814: We find no way to alleviate substantially the tension between
815: distance measures and the growth of horizon scale fluctuations that
816: impact the low multipole CMB temperature and its relationship to the
817: polarization.  In particular we show that the maximum likelihood
818: flat DGP model is a poorer fit than $\Lambda$CDM by $2\Delta \ln L =
819: 28$, nominally a $(2\Delta\ln L)^{1/2}=5.3\sigma$ result.
820: Interestingly, a substantial ($\sim 30\%$) contribution comes from
821: the change in the growth near the crossover scale where
822: perturbations leak into the bulk.
823: 
824: Adding in spatial curvature to the model can bring the distance measures
825: back in agreement with the data but only at the cost of exacerbating
826: the problem with growth.  The net result is that the maximum likelihood
827: only improves by $2\Delta \ln L = 5$ with one extra parameter and the
828: difference from $\Lambda$CDM is $2\Delta \ln L=23$, which is still
829: $4.8\sigma$ discrepant with most of the difference arising from the
830: changes to growth.
831: 
832: Furthermore, while the excess power at large angles can be reduced by
833: changing the initial power spectrum to eliminate large scale power,
834: existing WMAP5 CMB polarization measurements already forbid this
835: possibility.  Specifically, by introducing any finite cut off to the
836: initial power spectrum to flatten the temperature power spectrum, the
837: global likelihood decreases.
838: 
839: While it is still possible that the resolution of the ghost and strong
840: coupling issues of the theory can alter these consequences, it is
841: difficult to see how they can do so without altering the very mechanism
842: that makes it a candidate for acceleration without dark energy.  The
843: failure of this model highlights the power of combining growth and
844: distance measures in cosmology as a test of gravity on the largest
845: scales.
846: 
847: %
848: \begin{acknowledgments}
849: 
850: We thank K. Koyama and Y.S. Song for useful conversations. This work
851: was supported in part by the NSF grant AST-05-07161, by the
852: Initiatives in Science and Engineering (ISE) program at Columbia
853: University, and by the Pol\'anyi Program of the Hungarian National
854: Office for Research and Technology (NKTH).  SW and WH were supported
855: by the KICP under NSF PHY-0114422 WH was further supported by
856: U.S.~Dept. of Energy contract DE-FG02-90ER-40560 and the David and
857: Lucile Packard Foundation. LH was supported by the DOE grant
858: DE-FG02-92-ER40699, and thanks Alberto Nicolis for useful
859: discussions and Tai Kai Ng at the Hong Kong University of Science
860: and Technology for hospitality. MM was supported by the DOE grant
861: DE-AC02-98CH10886. Computational resources were provided by the
862: KICP-Fermilab computer cluster.
863: 
864: \end{acknowledgments}
865: 
866: %
867: \vfill
868: \bibliographystyle{physrev}
869: \bibliography{my}
870: 
871: %
872: \onecolumngrid
873: \appendix
874: \begin{center}
875:   {\bf APPENDIX A}
876: \end{center}
877: 
878: In Appendix A, we give details of the modification to CAMB employed
879: in the main text to calculate the modified growth of perturbations
880: in the DGP model under the PPF prescription.  For perturbations of
881: both the metric and matter in the various gauges, we use the same
882: notations as in Rfn.~\cite{Hu08}.
883: 
884: %
885: \subsection{PPF Description of Modified Gravity in the Linear Regime}
886: 
887: Structure formation in the linear regime for a certain class of modified
888: gravity theories can be equivalently described as an effective dark
889: energy component under ordinary gravity.  The requirements of this class
890: are that members remain metric theories in a statistically homogeneous
891: and isotropic universe where energy-momentum is conserved.  This class
892: includes the self-accelerating branch of the DGP model.
893: 
894: Following Rfns.~\cite{HuS07, Hu08}, scalar perturbations for the DGP
895: self-accelerating scenario can be parameterized by three free
896: functions, $g(\eta,k)$, $f_{\zeta}(\eta)$, and $f_G(\eta)$, and one
897: free parameter $c_{\Gamma}$.  We shall see in \S \ref{subsec:ppfDE}
898: that this is equivalent to specifying relationships between the
899: density perturbation, velocity and anisotropic stress of the
900: effective dark energy that close the conservation laws of the
901: effective dark energy.
902: 
903: The metric ratio or anisotropic stress parameter $g(\eta,k)$ is
904: defined as
905: %
906: \begin{equation}
907: \Phi_+ \equiv g(\eta,k) \Phi_- - {4\pi G \over H^2 k_H^2}
908: P_T\Pi_{T}\,,
909: \end{equation}
910: %
911: where $\Phi_+ \equiv (\Phi+\Psi)/2$, $\Phi_- \equiv (\Phi - \Psi)/2$,
912: with $\Phi=\delta g_{ij}/2g_{ij}$, $\Psi=\delta g_{00}/2g_{00}$ as the space-space and time-time metric
913: perturbations in the Newtonian gauge,
914:  $k_H=(k/aH)$, $\Pi$ stands for the
915: anisotropic stress, and the subscript ``$T$'' denotes sum over all the
916: true components. When anisotropic stresses of the true components are
917: negligible, $g$ just parameterizes deviations from GR in the metric
918: ratio of $\Phi_+$ to $\Phi_-$.
919: 
920: There are two key features of the PPF parameterization and these
921: define the two additional functions $f_{\zeta}(\eta)$ and
922: $f_G(\eta)$. The first is that the curvature perturbation in the
923: total matter comoving gauge $\zeta$ is conserved up to order $k_H^2$
924: in the super-horizon (SH) regime in the absence of non-adiabatic
925: fluctuations and background curvature
926: %
927: \begin{equation}
928: \lim_{k_H \ll 1} {\dot{\zeta} \over aH} = -{\Delta P_T - {2\over 3}
929: c_K P_T \Pi_T \over \rho_T + P_T} - {K \over k^2} k_H V_T + {1 \over
930: 3} c_K f_\zeta(\eta) k_H V_T \,, \label{eqn:sh}
931: \end{equation}
932: %
933: where $f_{\zeta}(\eta)$ parameterizes the relationship between the
934: metric and the matter.  The second is that in the quasi-static (QS)
935: regime, one recovers a modified Poisson equation with a potentially
936: time dependent effective Newton constant
937: \begin{equation}
938: \lim_{k_{H}\gg 1} \Phi_- = {4\pi G \over c_K k_H^2 H^2}{\Delta_T
939: \rho_T + c_K P_T \Pi_T \over  1+f_G(\eta)} \,. \label{eqn:qs}
940: \end{equation}
941: %
942: Note that even if $f_G=0$, the Poisson equation for $\Psi$ may also
943: be modified by a non-zero $g(\eta,k)$. In the above two equations,
944: $\Delta$, $V$ and $\Delta P\,(\neq P\Delta)$ are density, velocity,
945: pressure perturbations in the total matter comoving gauge, and
946: $c_K=1-3K/k^2$. To bridge the two regimes, an intermediate quantity
947: $\Gamma$ is introduced,
948: %
949: \begin{equation}
950: \Phi_-+\Gamma = {4\pi G \over c_K k_H^2 H^2} (\Delta_T \rho_T + c_K
951: P_T \Pi_T)\,,
952: \label{eqn:phim}
953: \end{equation}
954: %
955: and an interpolating equation is adopted to make sure it dynamically
956: recovers the behavior specified by Eqs.~(\ref{eqn:sh}) and
957: (\ref{eqn:qs})
958: %
959: \begin{equation}
960: (1 + c_\Gamma^2 k_H^2) \left[{\dot{\Gamma} \over aH} + \Gamma +
961: c_\Gamma^2 k_H^2 (\Gamma - f_G \Phi_-)\right] =
962: S\,.
963: \label{eqn:dgamma}
964: \end{equation}
965: %
966: Here the free parameter $c_{\Gamma}$ gives the transition scale between
967: the two regimes in terms of the Hubble scale, and the source term $S$ is
968: given by,
969: %
970: \begin{equation}
971: S={\dot{g}/(aH) -2 g \over g+1} \Phi_- +  {4\pi G \over (g+1)k_H^2
972: H^2} \left\{ g \left[{\dot{(P_T\Pi_T)}\over aH} + P_T\Pi_T\right] -
973: \left[ (g + f_\zeta + g f_\zeta)(\rho_T+P_T) - (\rho_e+P_e) \right]
974: k_H V_T \right\}\,.
975: \label{eqn:source}
976: \end{equation}
977: By interpolating between two exact behaviors specified by functions of
978: time alone $f_\zeta(\eta)$ and $f_G(\eta)$, the PPF parameterization 
979: is  both simple and general.  Moreover, the one function of time and scale
980: $g(\eta,k)$ also interpolates between two well defined functions of time alone
981: for many models, including DGP.  The same is not true of parameterizations
982: that involve the effective anisotropic stress alone or the metric functions directly.
983: 
984: %
985: 
986: %
987: \subsection{PPF Parameterization for DGP}
988: 
989: The PPF parameterization for the self-accelerating DGP is given in
990: Rfn.~\cite{HuS07}, which we summarize as the follows.  On super-horizon
991: scales, the iterative scaling solution developed in Rfn.~\cite{SaSH07}
992: is well described by
993: %
994: \begin{equation}
995: g_{\rm SH}(\eta)={ 9 \over 8 H r_c -1} \left( 1 + {0.51 \over H r_c
996: - 1.08} \right) \,, \label{eqn:gsh}
997: \end{equation}
998: %
999: and $f_\zeta(\eta)=0.4 g_{\rm SH}(\eta)$, while in the QS regime,
1000: the solution is parameterized by \cite{KoM06}
1001: \begin{equation}
1002: g_{\rm QS}(\eta) = -{1\over 3}\left[ 1-2 H r_c\left( 1 + {\dot{H}
1003: \over 3aH^2}\right) \right]^{-1}\,,
1004: \end{equation}
1005: %
1006: and $f_G(\eta)= 0$.  On an arbitrary scale in the linear regime, $g$
1007: is then interpolated by
1008: %
1009: \begin{equation}
1010: g(\eta,k) = { g_{\rm SH} +  g_{\rm QS}(c_{g}k_H)^{n_{g}} \over 1+
1011: (c_{g}k_H)^{n_{g}}}\,,
1012: \end{equation}
1013: %
1014: with $c_g=0.14$ and $n_g=3$. This fitting formula has been shown to
1015: give an accurate prediction for the evolution of $\Phi_-$ according
1016: to the dynamical scaling solution \cite{HuS07}. The transition scale
1017: for $\Gamma$ is set to be $c_\Gamma=1$, {\it i.e.}~at the horizon
1018: scale.  We note here that the solutions developed in
1019: Rfns.~\cite{KoM06,SaSH07} are for flat universes, so strictly
1020: speaking, the above PPF parameterization only works for DGP with
1021: $\Omega_K \rightarrow 0$.  However for the small curvatures that are
1022: allowed by the data, we would expect its effect on structure
1023: formation to be small and arise from terms such as $H^2 \rightarrow
1024: H^2+K/a^2$ (see, {\it e.g.}, Rfn.~\cite{GiSK08}).  Given the cosmic
1025: variance of the low-$\ell$ multipoles, these corrections should have
1026: negligible impact on the results.
1027: 
1028: %
1029: \subsection{``Dark Energy'' Representation of PPF}
1030: \label{subsec:ppfDE}
1031: 
1032: By comparing the equations that the PPF quantities satisfy with their
1033: counterparts in a dark energy system under general relativity, we obtain
1034: the following relations for the perturbations of PPF's corresponding
1035: effective dark energy.  These relations act as the closure conditions
1036: for the stress energy conservation equations of the effective dark
1037: energy.  The first closure condition is a relationship between the PPF
1038: $\Gamma$ variable and the components of the stress energy tensor of the
1039: effective dark energy
1040: %
1041: \begin{equation}
1042: \rho_{e}\Delta_{e} + 3(\rho_{e}+P_e) {V_{e}-V_{T}\over k_{H} } +
1043: c_K P_e\Pi_{e} = -{k^2 c_K \over 4\pi G a^2} \Gamma \,.
1044: \end{equation}
1045: The second closure condition is a relationship for the anisotropic
1046: stress
1047: %
1048: \begin{equation}
1049: P_e\Pi_e=-\frac{k_H^2H^2}{4 \pi G}g\Phi_-\,.
1050: \label{eqn:pie}
1051: \end{equation}
1052: %
1053: Stress energy conservation then defines the velocity perturbation
1054: %
1055: \begin{equation}
1056: { V_{e}-V_{T}\over k_{H}} =-{H^2 \over 4\pi G (\rho_{e} + P_e) }
1057: {g+1 \over F} \left[ S - \Gamma - {\dot{\Gamma} \over aH} +
1058: f_{\zeta}{4\pi G (\rho_{T}+P_T) \over H^2}{V_{T}\over k_{H}}
1059: \right] \,,
1060: \label{eqn:Ve}
1061: \end{equation}
1062: %
1063: with
1064: %
1065: \begin{equation}
1066: F = 1 + {12 \pi G a^2 \over k^2 c_K} (g+1)(\rho_T + P_T)\,.
1067: \end{equation}
1068: %
1069: For details of all the derivations, see Rfn.~\cite{Hu08}.  The PPF
1070: equation for $\Gamma$ then replaces the continuity and Navier-Stokes
1071: equations for the dark energy.  Note that the effective dark energy
1072: pressure perturbation, which obeys complicated dynamics to
1073: enforce the large and small scale behavior, is not used.
1074: 
1075: %
1076: \subsection{Modifying CAMB to Include PPF}
1077: 
1078: Given the dark energy representation of PPF, we only need to modify
1079: the parts in CAMB where dark energy perturbations appear explicitly
1080: or are needed, in addition to its equation of state which will be
1081: specified by the desired background expansion. These include the
1082: Einstein equations and the source term for the CMB temperature
1083: anisotropy. Since CAMB adopts the synchronous gauge, in this
1084: section, we will express everything in the same gauge.
1085: 
1086: The two Einstein equations used in CAMB are,
1087: %
1088: \begin{eqnarray}
1089: {\dot{h_L}\over 2k} - c_K k_H \eta_T &=& {4\pi G \over k_H H^2}
1090: (\rho_{T}\delta_{T}^{s} + \rho_{e} \delta_{e}^{s} )\,,\\
1091: k\dot{\eta_{T}}- K {( \dot{h_L} + 6\dot{\eta_{T}}) \over 2k} &=&
1092: {4\pi G a^2} \left[ (\rho_{T}+P_T)v_{T}^s +
1093: (\rho_{e}+P_e)v_{e}^s \right] \,,
1094: \end{eqnarray}
1095: %
1096: where $h_L$, $\eta_T$ are metric perturbations, $\delta$, $v$ are
1097: density and velocity perturbations, and superscript ``s'' labels the
1098: synchronous gauge. Here, we only need to provide $\delta_e^s$ and
1099: $v_e^s$.  Given the gauge transformation relation for velocity
1100: %
1101: \begin{equation}
1102: V=v^s+k\alpha\,,
1103: \end{equation}
1104: %
1105: with $\alpha \equiv (\dot{h_L} + 6\dot{\eta_{T}}) / 2k^2$, the
1106: following expression for $v_e^s$ is easily obtained from
1107: Eq.~(\ref{eqn:Ve})
1108: %
1109: \begin{equation}
1110: (\rho_e+P_e)v_e^s = (\rho_e+P_e)v_T^s - \frac{k_H H^2}{4 \pi G}
1111: \frac{(g+1)}{F}  \left[ S - \Gamma - \frac{\dot{\Gamma}}{aH} +
1112: f_{\zeta}\frac{4\pi G(\rho_T+P_T)(v_T^s + k \alpha )}{k_H H^2}
1113: \right]\,.
1114: \label{eqn:ve}
1115: \end{equation}
1116: %
1117: In order to calculate $v_e^s$, we need to evaluate $\alpha$, which we
1118: find, with the help of the first closure condition and the two Einstein
1119: equations, to be given by
1120: %
1121: \begin{equation}
1122: k \alpha = k_H \eta_T + \frac{4 \pi G}{c_K k_H H^2} \left[\rho_T
1123: \delta_T^s + \frac{3(\rho_T+P_T)v_T^s}{k_H}\right] - \frac{4 \pi G
1124: }{k_H H^2}P_e \Pi_e - k_H \Gamma \,.
1125: \end{equation}
1126: %
1127: Here, the anisotropic stress is gauge-independent, and is given by
1128: Eq.~(\ref{eqn:pie}), where $\Phi_-$ is given by gauge-transforming
1129: the density perturbation in Eq.~(\ref{eqn:phim}) according to
1130: %
1131: \begin{equation}
1132: \rho \Delta = \rho \delta^s - \dot{\rho} {v_T^s \over k}\,.
1133: \end{equation}
1134: %
1135: In addition to $\alpha$, we need also to specify $S$ and $\dot{\Gamma}$
1136: in order to get $v_e^s$. Given $\Phi_-$, $S$ can be calculated by
1137: gauge-transforming $V_T$ in Eq.~(\ref{eqn:source}), and $\dot{\Gamma}$
1138: then follows from Eq.~(\ref{eqn:dgamma}).  Provided $v_e^s$ and
1139: $P_e\Pi_e$, $\delta_e^s$ can be obtained by gauge-transforming the first
1140: closure condition
1141: %
1142: \begin{equation}
1143: \rho_e\delta_e^s=- c_K P_e \Pi_e - 3(\rho_e + P_e) {v_e^s \over k_H}
1144: -\frac{c_K k_H^2 H^2}{4 \pi G} \Gamma\,.
1145: \end{equation}
1146: %
1147: 
1148: The source term for the CMB temperature anisotropy is given by
1149: \cite{ZaSB97}
1150: %
1151: \begin{equation}
1152: S_T(\eta, k)={\cal G}\left(\Delta_{T0}+2 \dot{\alpha} +{\dot{v_b^s}
1153: \over k}+{\Sigma \over 4 {\bar b}} +{3\ddot{\Sigma}\over 4k^2 {\bar
1154: b}}\right) + e^{-\kappa}(\dot{\eta_T}+\ddot{\alpha}) +\dot{\cal
1155: G}\left({v_b^s \over k}+\alpha +{3\dot{\Sigma}\over 2k^2 {\bar
1156: b}}\right) +{3 \ddot{\cal G}\Sigma \over 4k^2 {\bar b}}\,,
1157: \end{equation}
1158: %
1159: where the visibility function ${\cal G} \equiv -\dot{\kappa}
1160: \exp(-\kappa)$, with $\kappa(\eta)$ the optical depth from today to
1161: $\eta$ for Thomson scattering, $\Delta_{T \ell}$ is the multipole of the
1162: temperature anisotropy, $v_b^s$ is the baryon velocity, $\Sigma \equiv
1163: \Delta_{T2} -12 _2\Delta_{P2}$, with $_2\Delta_{P2}$ the quadrupole of
1164: the polarization anisotropy (for more explicit definitions for
1165: $\Delta_{T \ell}$ and $_2\Delta_{P \ell}$, see Rfn.~\cite{ZaSB97}), and
1166: $\bar{b}^2=c_K$.  Here we need to specify $\dot{\alpha}$ and
1167: $\ddot{\alpha}$. $\dot{\alpha}$ is given by the Einstein equation
1168: %
1169: \begin{equation}
1170: \dot{\alpha}+2{\dot{a} \over a} \alpha - \eta_T = -{8 \pi G \over
1171: k_H^2 H^2} (P_T \Pi_T + P_e \Pi_e) \,,
1172: \end{equation}
1173: %
1174: Differentiating this equation with respect to $\eta$ also gives us
1175: $\ddot{\alpha}$, which will need the derivative of the effective dark
1176: energy's anisotropic stress.  From energy-momentum conservation and
1177: the first closure condition, we obtain
1178: %
1179: \begin{equation}
1180: \dot{(P_e \Pi_e)}={\dot{h_L} \over 2} {{(\rho_e + P_e)} \over c_K} -
1181: \frac{k_H^2 H^2}{4 \pi G} \dot{\Gamma} + {aH \over c_K}  \left\{
1182: \delta \rho_e^s + (\rho_e+P_e) v_e^s \left[ {6 \over k_H} + k_H - {3
1183: \over k}\left({\dot{a} \over a }+{\dot{H} \over H }\right)\right]
1184: \right\}\,,
1185: \end{equation}
1186: %
1187: which completes our modification of CAMB.
1188: 
1189: \onecolumngrid
1190: \appendix
1191: \begin{center}
1192:   {\bf APPENDIX B}
1193: \end{center}
1194: 
1195: In Appendix B, we briefly contrast the PPF prediction for the CMB
1196: temperature power spectrum with an attempt to approximate the DGP
1197: prediction via the parameter-splitting technique
1198: \cite{ZHS05,WaHMH07}. The technique works by splitting the dark
1199: energy equation of state $w$ into two separate parameters, with one,
1200: $w_{\rm geometry}$, determining geometric distances and the other,
1201: $w_{\rm growth}$, determining the growth of structure. We choose
1202: $w_{\rm geometry} = -0.7$ and $w_{\rm growth} = -0.57$, which are
1203: obtained respectively by fitting to $H(z)$ (for $z = 0$ to $10$) and
1204: fitting to the quasi-static growth factor (for $z = 0$ to $2$)
1205: according to the best-fit flat DGP model (Table \ref{tab:fbf}). The
1206: result is shown in Fig. \ref{fig:paramsplit}. One can see that
1207: parameter-splitting falls short of the PPF prediction on large
1208: angular scales. This illustrates the importance of correctly
1209: modeling the perturbation growth on horizon scales. The
1210: parameter-split of $w$ does not contain enough freedom to describe
1211: perturbation growth in both the sub-horizon (quasi-static) and
1212: super-horizon regimes.
1213: 
1214: %
1215: \begin{figure}[t!]
1216: \resizebox{90mm}{!}{\includegraphics{paramsplit.eps}}
1217: \caption{\label{fig:paramsplit} A comparison of the PPF prediction
1218: (upper solid line) with the prediction by parameter-splitting (lower
1219: dashed line) for the DGP model. }
1220: \end{figure}
1221: %
1222: 
1223: \end{document}
1224: