0709.1128/ms.tex
1: \documentclass[aps,prd,twocolumn,amssymb,eqsecnum,showpacs,nofootinbib]{revtex4}
2: %\documentclass[prd,nofootinbib,preprint]{revtex4}
3: \usepackage{graphicx}
4: \usepackage{amsmath}
5: \usepackage{amssymb}
6: \usepackage{bm}
7:  \usepackage{color}
8: 
9: \def\nn{\nonumber}
10: \def\be{\begin{equation}}
11: \def\ee{\end{equation}}
12: \def\ba{\begin{eqnarray}}
13: \def\ea{\end{eqnarray}}
14: \def\dper{d_\perp}
15: \def\Dp{{\rm D}p}
16: \def\rp{r_\perp}
17: \def\rpl{r_\parallel}
18: \def\Vp{V_\perp}
19: \def\Vpl{V_\parallel}
20: \def\lp{\ell_\perp}
21: \def\lpl{\ell_\parallel}
22: \def\lsim{\roughly<}
23: \def\gsim{\roughly>}
24: \def\vpl{v_\parallel}
25: \def\a{{\alpha^\prime}}
26: \def\dprime{\prime \prime}
27: \def\linf{L_{\rm inf}}
28: \def\Deltar{\Delta^2_{\cal R}(k)}
29: \def\Deltarz{\Delta^2_{\cal R}(k_0)}
30: %\def\mpl{{\rm p}}
31: \def\mpl{m_{\rm p}}
32:   \def\be{\begin{equation}}
33: \def\ee{\end{equation}}
34:  \def\bi{\begin{itemize}}
35:  \def\ei{\end{itemize}}
36:   \def\ben{\begin{enumerate}}
37: \def\een{\end{enumerate}}
38:   \def\bt{\begin{tabular}}
39: \def\et{\end{tabular}}
40: \def\bc{\begin{center}}
41: \def\ec{\end{center}}
42: \def\la{\label}
43: \def\kap{\kappa}
44: \def\gam{\gamma}
45: \def\bea{\begin{eqnarray}}
46: \def\eea{\end{eqnarray}}
47: \def\l{\left}
48: \def\r{\right}
49: \def\f{\frac}
50: \def\hub{{\cal H}}
51: \def\d{\partial}
52: \def\ddaoa{\frac{\ddot{a}}{a}}  %\Double-Dot A Over A
53: \def\dRoY{\left( \frac{\delta R}{Y} \right)}   % Delta R / Y
54: \newcommand{\koa}[1]{\frac{k^{#1}}{a^{#1}}}                     % (k/a)
55: \newcommand{\aok}[1]{\frac{a^{#1}}{k^{#1}}}                     % (a/k)
56: \newcommand{\pderv}[2]{\frac{\partial #1}{\partial #2}}
57: %\def\le{\left}
58: \def\ri{\right}
59: \def\fr{\frac}
60: \def\la{\label}
61: \def\kap{\kappa}
62: \def\gam{\gamma}
63: \def\omp{e^{\kappa\beta\phi}}
64: \def\omn{e^{-\kappa\beta\phi}}
65: \def\omph{e^{\frac{\kappa\beta\phi}{2}}}
66: \def\omnh{e^{-\frac{\kappa\beta\phi}{2}}}
67: \def\ompd{e^{2\kappa\beta\phi}}
68: \def\omnd{e^{-2\kappa\beta\phi}}
69: \def\om{\kappa\beta\phi}
70: \def\dom{\kappa\beta\varphi}
71: \def\fru{(1+f_{R})}
72: \def\frd{\frac{1}{(1+f_{R})}}
73: \def\frdh{\frac{1}{(1+f_{R})^{\frac{1}{2}}}}
74: \def\frru{\frac{f_{RR}}{(1+f_{R})}}
75: \def\frrd{\frac{(1+f_{R})}{f_{RR}}}
76: \def\tcr{\textcolor{red}}
77: \def\tcb{\textcolor{blue}}
78: \newcommand{\bes}{\begin{subequations}}
79: \newcommand{\ees}{\end{subequations}}
80: 
81: 
82: \begin{document}
83: 
84: \title{Adiabatic instability in coupled dark energy-dark matter models}
85: \author{Rachel Bean$^{1}$}
86: \author{\'Eanna \'E. Flanagan$^{1,2}$}
87: \author{Mark Trodden$^{3}$}
88: 
89: \affiliation{$^{1}$ Department of Astronomy, Cornell University, Ithaca, NY 14853, USA}
90: \affiliation{$^{2}$Laboratory for Elementary Particle Physics, Cornell University, Ithaca, NY 14853, USA.}
91: \affiliation{$^{3}$Department of Physics, Syracuse University, Syracuse, NY 13244, USA}
92: 
93: % ----------------------------------------------------------------------
94: %
95: % TIME OF DAY
96: %
97: \newcount\hh
98: \newcount\mm
99: \mm=\time
100: \hh=\time
101: \divide\hh by 60
102: \divide\mm by 60
103: \multiply\mm by 60
104: \mm=-\mm
105: \advance\mm by \time
106: \def\hhmm{\number\hh:\ifnum\mm<10{}0\fi\number\mm}
107: 
108: 
109: % ----------------------------------------------------------------------
110: 
111: 
112: \date{\today}
113: 
114: \begin{abstract}
115: 
116: We consider theories in which there exists a nontrivial coupling between the
117: dark matter sector and the sector responsible for the acceleration of the universe.
118: Such theories can possess an adiabatic regime in which the
119: quintessence field always sits at the minimum of its effective
120: potential, which is set by the local dark matter density.
121: We show that if the coupling strength is much larger than
122: gravitational, then the adiabatic regime is always
123: subject to an instability.  The instability, which can also be thought of as a type
124: of Jeans instability, is characterized by a negative sound speed squared of
125: an effective coupled dark matter/dark energy fluid, and results in the
126: exponential growth of small scale modes.
127: We discuss the role of the instability in specific coupled CDM and Mass Varying Neutrino (MaVaN) models
128: of dark energy, and clarify for these theories the regimes in which the instability
129: can be evaded due to non-adiabaticity or weak coupling.
130: 
131: \end{abstract}
132: \maketitle
133: 
134: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
135: \section{Introduction}
136: \label{intro}
137: 
138: In order for our cosmological models to provide an accurate fit to all
139: current observational data, it is necessary to postulate two dramatic
140: augmentations beyond the minimalist assumption of baryonic matter
141: interacting gravitationally through Einstein's equations. The first
142: assumption is that there must exist either new gravitational dynamics
143: or a new component of the cosmic energy budget -- {\it dark matter} --
144: that allows structure to form and accounts for weak lensing and
145: galactic rotation curves. The second assumption is that a further
146: dynamical modification or energy component -- {\it dark energy} --
147: exists, driving late-time cosmic acceleration.
148: 
149: In the first case, recent results~\cite{Clowe:2006eq} have added
150: significant support to an explanation in terms of particulate
151: cold dark matter (CDM), rather than a modification of gravity. In the case of
152: cosmic acceleration, however, the data remains consistent with a
153: simple cosmological constant, with modifications to gravity, or with
154: dark energy as the correct explanation.
155: 
156: A logical possibility is that the two dark sectors -- dark matter and
157: dark energy -- interact with each other or with the visible sector of
158: the theory
159: \cite{Damour:1990,Carroll:1998zi,Amendola:1999er,Bean:2000zm,Bean:2001ys,
160:  Chiba:2003ir,Majerotto:2004ji,Das:2005yj,Lee:2006za,Kesden:2006zb}.
161: In fact, a number of models have been proposed that exploit this idea
162: to address, among other things, the coincidence problem
163: \cite{Bean:2000zm,Bean:2001ys}. Further,
164: there exist classes of modified gravity models which may either be
165: mapped to interacting dark energy models, or closely approximated by
166: them over a broad range of dynamical interest \cite{Chiba:2003ir,Tsujikawa:2007gd}.
167: 
168: 
169: There are concerns, however, about coupling these two rather
170: differently behaving sectors. One concern is the presence of dynamical
171: attractors that produce a cosmic expansion history significantly
172: different from $\Lambda CDM$; see, for example, Refs.\
173: \cite{Amendola:1999er,Amendola:2006mr,Agarwal:2007}.
174: Another specific example is
175: the possibility of
176: instabilities that are not present for the uncoupled system
177: \cite{Afshordi:2005ym,Kaplinghat:2006jk,Bjaelde:2007ki}.
178: 
179: In this paper we perform a careful analysis of the viability of
180: coupled dark energy-dark matter models.
181: We consider implications of
182: such coupled theories on cosmological scales, both in terms of
183: homogeneous background expansion and the growth of linear
184: perturbations. We focus, in particular, on a
185: class of models in which there exists an {\it adiabatic regime} in
186: which the dark energy field instantaneously tracks the minimum of its
187: effective potential, as explored by Das, Corasaniti and
188: Khoury\cite{Das:2005yj}.
189: We show that if the coupling strength
190: is much larger than gravitational, then the adiabatic regime is always
191: subject to an instability.  The instability is characterized by a negative sound speed squared of
192: an effective coupled dark matter/dark energy fluid, and results in the
193: exponential growth of small scale modes.
194: We analyze a number of different models, and show that for these
195: models the instability
196: strongly constrains the region in parameter space that is compatible with observations.
197: A short version of our results was given in the recent paper \cite{letter}.
198: 
199: We improve on previous investigations of this instability
200: \cite{Afshordi:2005ym,Kaplinghat:2006jk,Bjaelde:2007ki} in a number of
201: ways.  First, we show that the instability occurs only for coupling strengths that are strong
202: compared to gravitational coupling, and can be evaded at weaker
203: couplings; this point was missed all in previous work.
204: Second, we give a simple intuitive explanation of the instability
205: as a type of Jeans instability.  Normally, for cosmological perturbations, Hubble damping
206: converts the exponential growth of gravitationally unstable modes into power law growth.
207: However, here the Hubble damping is ineffective, due to the fact that
208: the effective Newton's constant for the interaction of dark matter
209: with itself is much larger than the Newton's constant governing the
210: background cosmology.  The result is the exponential growth of perturbations.
211: Finally, we generalize previous treatments to allow an arbitrary coupling between
212: the dark energy sector and the visible sector, in addition to the
213: coupling between dark energy and dark matter.
214: 
215: 
216: 
217: In more detail, we consider models which are characterized by a
218: function $\alpha_c(\phi)$ governing the interaction between CDM and a
219: quintessence type scalar field $\phi$, and a function $\alpha_b(\phi)$
220: governing the interaction between visible matter (baryons) and $\phi$.
221: The effective Newton's constants
222: for the interaction of CDM with itself (cc), the interaction of CDM
223: with baryons (cb), and the interaction of baryons with themselves (bb)
224: are (see Appendix \ref{sec:Geff})
225: \bes
226: \label{Newtons}
227: \bea
228: G_{cc} &=& G \left [1 + 2 \mpl^2 \alpha_c'(\phi)^2 \right], \\
229: G_{cb} &=& G \left [1 + 2 \mpl^2 \alpha_c'(\phi) \alpha_b'(\phi)\right], \\
230: G_{bb} &=& G \left [1 + 2 \mpl^2 \alpha_b'(\phi)^2 \right].
231: \eea
232: \ees
233: These Newton constants are those of the Einstein frame in the short
234: wavelength limit, and include the effect of the scalar interaction
235: mediated by $\phi$.  Here $\mpl$ is the Planck mass.
236: 
237: The adiabatic instability occurs when $G_{cc} \gg G$, or equivalently
238: $\mpl \alpha_c' \gg 1$, when the scalar coupling between dark matter
239: particles is strong compared to the tensor coupling\footnote{This
240: requirement is admittedly a fine tuning and unnatural from the
241: viewpoint of effective field theory.}.
242: In the present day Universe, this regime is excluded by observations
243: if we assume that perturbations to the cosmological background value of
244: $\phi$ are in the linear regime, so that the parameters
245: (\ref{Newtons}) are constants (see below).
246: First, tests of general relativity in the Solar System constrain $\mpl
247: \alpha_b'$ to be small compared to unity.  Second, observations of
248: tidal disruption of satellite galaxies of the Milky Way provide the
249: constraint\footnote{Only the combination (\ref{Mark}) of the Newton's constants
250:   can be constrained by observations of the gravitational interactions
251:   of baryons and dark matter, since once cannot separately measure the
252:   densities and the Newton's constants.} \cite{Kesden:2006zb}
253: \be
254: \left| \frac{G_{bc}}{\sqrt{G_{cc} G_{bb}}} -1 \right| \alt 0.02.
255: \label{Mark}
256: \ee
257: Combining these constraints excludes the regime $G_{cc} \gg G$.
258: 
259: 
260: Nevertheless, it is still of interest to explore the adiabatic
261: instability, for a number of reasons.
262: First, the argument given above assumes a single type of CDM particle.
263: In more complicated models with two different CDM species, one coupled
264: and one uncoupled, the constraint can be evaded (see Secs.\
265: \ref{sec:twocomponent} and \ref{sec:MaVaN} below).  Second, the Newton's
266: constants can evolve as a function of
267: redshift, via the dependence of $\phi$ on redshift,
268: so present day observations $(z \approx 0)$ do not exclude the
269: occurrence of the instability at high redshifts.  Third, it is useful
270: to have multiple independent constraints on coupled models.
271: Finally, it is not necessarily true that perturbations to the
272: cosmological background value of $\phi$ are in the linear regime.  In particular, it
273: has been claimed that observational constraints on $\mpl
274: \alpha_b'$ can be evaded in chameleon models\footnote{In particular,
275: chameleon models in the regime $\mpl \alpha_b' \gg 1$
276: have been extensively explored \cite{Mota:2006ed}.}, due the
277: dependence of the local effective Newton's constants (\ref{Newtons})
278: on $\phi$ and hence on the local matter density \cite{Khoury:2003aq,Khoury:2003rn}.
279: 
280: 
281: 
282: 
283: The structure of the paper is as follows. In the next section, we
284: introduce the general class of models and the framework in which we
285: work.
286: In section~\ref{adiabatic} we discuss the
287: adiabatic regime, giving both local and nonlocal conditions for its
288: applicability.  We also derive the effective equation of state for
289: these models in the adiabatic regime; as previously noted in Ref.\
290: \cite{Das:2005yj}, superacceleration is a generic feature of these models.
291: Section \ref{instability} discusses the adiabatic instability that can
292: arise in the adiabatic regime.  We discuss two complementary ways of
293: understanding the instability, a hydrodynamic viewpoint and a Jeans
294: instability viewpoint.  We also derive the range of lengthscales over
295: which the instability operates, which was incomplete in earlier
296: analyses \cite{Afshordi:2005ym,Kaplinghat:2006jk}.
297: In
298: section~\ref{examples} we apply our general analysis to some
299: well-known coupled models. We perform analytic and numerical analyses of the evolution of
300: the background cosmology and of perturbations in both coupled CDM,
301: chameleon \cite{Khoury:2003aq,Khoury:2003rn}
302: and mass varying neutrino (MaVaN) \cite{Fardon:2003eh,Kaplan:2004dq,Fardon:2005wc} models, and identify
303: the regimes in which they are subject to the instability. We conclude in~\ref{conclusions} by summarizing
304: our findings and discussing their implications.
305: 
306: In appendix~\ref{visible} we generalize the derivation of the
307: instability to include baryonic matter, and in appendix~\ref{kinetic}
308: we generalize the derivation of the instability to regimes where
309: CDM cannot be treated as a pressureless fluid but instead must be
310: treated in terms of kinetic theory.
311: Finally, in appendix~\ref{sec:Geff} we derive and describe the Jeans
312: instability viewpoint on the instability.
313: 
314: A note on conventions: throughout this paper we use a metric signature
315: (-,+,+,+), and we define the reduced Planck mass by $\mpl \equiv
316: (8\pi G)^{-1/2} $.
317: 
318: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
319: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
320: \section{Models with a coupling between dark energy and dark matter}
321: \label{sec:web}
322: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
323: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
324: 
325: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
326: \subsection{General class of models}
327: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
328: 
329: We begin from the following action
330: \bea
331: S &=& S[g_{ab},\phi,\Psi_{\rm j}] \nonumber \\
332: &=& \int d^4x\sqrt{-g}
333: \left[ \frac{1}{2} \mpl^2 R
334: -\frac{1}{2} (\nabla \phi)^2
335:  - V(\phi)
336: \right] \nonumber \\
337: &&+ \sum_{\rm j} S_{\rm j}[e^{2 \alpha_{\rm j}(\phi)} g_{ab}, \Psi_{\rm j}],
338: \label{action0}
339: \eea
340: where $g_{ab}$ is the Einstein frame metric, $\phi$ is a scalar field
341: which acts as dark energy, and $\Psi_{\rm j}$ are the matter fields.
342: The functions $\alpha_{\rm j}(\phi)$ are coupling functions that determine the strength
343: of the coupling of the jth matter sector to the scalar field.
344: The matter sectors include cold dark matter with coupling function
345: $\alpha_c(\phi)$, and baryons with coupling function $\alpha_b(\phi)$.
346: 
347: This general action encapsulates many models studied in the
348: literature.
349: The case
350: of equal coupling to dark and visible matter,
351: $\alpha_{\rm j}(\phi) = \alpha(\phi)$ for all j\footnote{A more
352: precise characterization of equal coupling is $\alpha_{\rm j}'(\phi)
353: = \alpha'(\phi)$ for all j, since any constant term in $\alpha_{\rm
354:   j}(\phi)$ can be absorbed into a rescaling of all dimensionful
355: parameters in the action $S_{\rm j}$.},
356: corresponds to
357: scalar-tensor theories of gravity which have been extensively studied,
358: recently under the name of ``coupled quintessence''
359: \cite{Amendola:1999er}.
360: This class of theories includes the $f(R)$ modified gravity theories
361: \cite{Carroll:2003wy}.
362: Several authors have considered the case $\alpha_b(\phi)=0$, in which
363: the dark energy couples only to dark matter.
364: We note that this choice yields a
365: microphysical model
366: for Modified-Source Gravity~\cite{Carroll:2006jn}
367: in the adiabatic regime discussed in Sec.\ \ref{adiabatic} below
368: (in the approximation where one considers only a dark matter source
369: and neglects baryons).
370: 
371: 
372: There is good theoretical motivation for considering nontrivial and different
373: coupling functions $\alpha_{\rm j}(\phi)$, since
374: this is a generic prediction
375: of string theory and of higher dimensional models.
376: In fact, typically the moduli and dilation fields of string theory must be
377: massive today, because for massless fields it is difficult to satisfy the observational
378: constraints [Solar System tests and fifth force experiments that require
379:  $d\alpha_b/d\phi \alt 10^{-2} \mpl^{-1}$, equivalence principle
380: tests that require matter coupling functions for
381: different matter sectors aside from dark matter to be
382: very nearly the same] in a natural way because
383: of loop corrections.
384: However dynamical dark energy models require massless or nearly
385: massless fields so one is forced to confront the naturalness issue.
386: One possible solution, suggested by Damour and Polyakov
387: \cite{Damour:1994ya}, is that there is an attractor mechanism under
388: cosmological evolution that drives the theory to be very close to
389: general relativity, $\alpha_{\rm j}^\prime \to 0$ for all $j$.
390: It is possible that this mechanism does not work
391: perfectly, and that there are residual deviations in the form of
392: nontrivial matter couplings.
393: 
394: 
395: In addition, while equivalence principle tests strongly constrain
396: differences between the coupling functions for different types
397: of visible matter, the corresponding constraints on dark matter are
398: much weaker, as first pointed out by Damour, Gibbons and Gundlach \cite{Damour:1990}.
399: So there is good motivation for exploring models with various couplings
400: to dark matter.
401: 
402: 
403: If the jth sector consists of a fermion of mass $m$, the
404: corresponding action in Eq.\ (\ref{action0}) can be written as
405: \begin{eqnarray}
406: && S_{\rm j} =
407: \int d^4 x
408: \sqrt{- g} \bigg[ e^{p \alpha(\phi)} {\bar \Psi}_{\rm j} i \gamma^\mu \nabla_\mu
409:  \Psi_{\rm j} - e^{q \alpha(\phi) } m {\bar \Psi}_{\rm j} \Psi_{\rm
410:    j} \bigg], \nonumber\\
411: \label{eq:action10}
412: \end{eqnarray}
413: with $p=2$ and $q=3$.  Other possibilities for
414: $p$ and $q$ have been explored in the
415: literature.  For example, Farrar and Peebles \cite{Farrar:2003uw}
416: consider the case $p=0$, and Bean \cite{Bean:2001ys} considers the
417: case $p=q$.  All of these choices are equivalent\footnote{Up to
418:   redefining $\alpha$ by multiplying by a constant.} in the
419: nonrelativistic limit, when the fermion's rest masses dominate their
420: gravitational interactions.
421: The choice we make here is motivated by the fact that it satisfies the
422: equivalence principle within the jth sector, and also the fact that it
423: arises naturally from higher dimensional models.
424: 
425: 
426: 
427: The field equations resulting from the action (\ref{action0}) are
428: \bea
429: \mpl^2 G_{ab} &=& \nabla_a \phi \nabla_b \phi - \frac{1}{2} g_{ab}
430: (\nabla \phi)^2 - V(\phi) g_{ab} \nonumber \\
431: &&
432: + \sum_{\rm j}e^{4 \alpha_{\rm j}(\phi)} \left[ ({\bar \rho}_{\rm j} + {\bar
433:     p}_{\rm j}) u_{{\rm j}\,a} u_{{\rm j}\,b} + {\bar
434:     p}_{\rm j} g_{ab} \right],\ \ \
435: \label{ee0}
436: \eea
437: and
438: \be
439: \nabla_a \nabla^a \phi - V'(\phi) = \sum_{\rm j} \alpha_{\rm j}'(\phi)
440: e^{4 \alpha_{\rm j}(\phi)}
441: ({\bar \rho}_{\rm j} - 3 {\bar p}_{\rm j} ).
442: \label{eq:scalar10}
443: \ee
444: where the prime represents a derivative with respect to $\phi$. Here we treat the matter $\Psi_{\rm j}$ as a fluid with Jordan-frame density
445: ${\bar \rho}_{\rm j}$ and pressure ${\bar p}_{\rm j}$ and with a
446: 4-velocity $u_{{\rm j}\,a}$ normalized according to $g^{ab} u_{{\rm
447:     j}\,a} u_{{\rm j}\,b}=-1$ [see Appendix \ref{visible}].
448: 
449: For much of this paper we shall work in the approximation where we consider only the
450: gravitational effects of the dark matter and neglect the gravitational
451: effects of the baryons.  In this approximation, the sums over j in
452: Eqs.\ (\ref{ee0}) and (\ref{eq:scalar10}) reduce to a single term involving the dark
453: matter.  Dropping the index j for simplicity, and denoting the
454: CDM coupling function $\alpha_c(\phi)$ simply
455: by $\alpha(\phi)$, the field equations become
456: \bea
457: \mpl^2 G_{ab} &=& \nabla_a \phi \nabla_b \phi - \frac{1}{2} g_{ab}
458: (\nabla \phi)^2 - V(\phi) g_{ab} \nonumber \\
459: && + e^{4 \alpha(\phi)}
460: \left[ ({\bar \rho} + {\bar p}) u_a u_b + {\bar p} g_{ab} \right],
461: \label{ee0d}
462: \eea
463: and
464: \be
465: \nabla_a \nabla^a \phi - V'(\phi) = \alpha'(\phi) e^{4 \alpha(\phi)}
466: ({\bar \rho} - 3 {\bar p} ),
467: \label{eq:scalar10a}
468: \ee
469: These are are the standard equations for a
470: scalar-tensor cosmology.
471: 
472: For dark matter we have ${\bar
473:   p}=0$, and we define a rescaled density variable
474: \be
475: \rho \equiv e^{3 \alpha(\phi)} {\bar \rho}.
476: \label{newdensity}
477: \ee
478: With this new variable, we obtain
479: \be
480: \mpl^2 G_{ab} = \nabla_a \phi \nabla_b \phi - \frac{1}{2} g_{ab}
481: (\nabla \phi)^2 - V(\phi) g_{ab} + e^{\alpha(\phi)}
482: \rho u_a u_b,
483: \label{ee}
484: \ee
485: and
486: \be
487: \nabla_a \nabla^a \phi - V'(\phi) = \alpha'(\phi) e^{\alpha(\phi)}
488: \rho \ .
489: \label{eq:scalar1}
490: \ee
491: The scalar field equation (\ref{eq:scalar1}) can also be written as
492: \be
493: \nabla_a \nabla^a \phi - V_{\rm eff}'(\phi) = 0,
494: \label{eq:scalar2}
495: \ee
496: where the effective potential that includes the matter coupling is
497: \be
498: V_{\rm eff}(\phi) = V(\phi) + e^{\alpha(\phi)}
499: \rho.
500: \label{phieqn}
501: \ee
502: and the prime means derivative with respect to $\phi$ at fixed $\rho$.
503: The equations of motion for the fluid are [see Appendix \ref{visible}]
504: \be
505: \nabla_a ( \rho u^a) =0,
506: \label{eq:fluid1}
507: \ee
508: and
509: \be
510: u^b \nabla_b u^a  = - (g^{ab} + u^a u^b) \nabla_b \alpha \ .
511: \label{eq:fluid2}
512: \ee
513: 
514: 
515: Note that the equations of motion (\ref{ee}) -- (\ref{eq:fluid2}) of the theory
516: do not depend on the coupling function $\alpha_b(\phi)$ of the
517: scalar field to visible matter, in the approximation where we neglect
518: the baryons.
519: Our discussion of the instability in the following sections
520: will be valid for arbitrary $\alpha_b(\phi)$.
521: However we note that $\alpha_b(\phi)$ will
522: enter when we want to compare with observations, since the metric that
523: is measured is $e^{2 \alpha_b} g_{ab}$.  When discussing observations below we will
524: focus on the case $\alpha_b=0$, where the dark energy is coupled only to dark matter.
525: 
526: 
527: 
528: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
529: 
530: \section{The Adiabatic Regime}
531: \label{adiabatic}
532: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
533: 
534: The effective potential (\ref{phieqn}) governing the evolution of the
535: scalar field
536: is the sum of two terms, one arising from the original
537: potential $V(\phi)$, and the other arising from the coupling to the
538: energy density of the dark matter fluid. It is possible, for
539: appropriate choices of the potential and the coupling function, that
540: competition between these terms leads to a minimum of the effective
541: potential. Further, in some regimes, it may be possible for the
542: solution of the equation of motion for $\phi$ to
543: adiabatically\footnote{Note that throughout this paper, we use
544: ``adiabatic'' in the sense of ``gradual change'', not in the thermodynamic
545: sense of ``at constant entropy''.} in the  track
546: the position of this minimum.
547: That is to say, the timescale or lengthscale for $\phi$ to adjust
548: itself to the changing position of the minimum of the effective potential may be short
549: compared to the timescale or lengthscale over which the background
550: density is changing.
551: 
552: 
553: This {\it adiabatic regime} has been previously discussed for spatial
554: variations of $\phi$ in the interior of massive bodies in the so-called Chameleon field models
555: \cite{Khoury:2003aq,Khoury:2003rn}, and for the time variation of
556: $\phi$ in a cosmological context by Refs.\ \cite{Kaplinghat:2006jk,Das:2005yj}.
557: It also has been studied for the specific case of $f(R)$ modified gravity models
558: \cite{Navarro:2006mw,Faulkner:2006ub,Hu:2007nk}, for which the action
559: in the Einstein frame is of the general form (\ref{action0}).
560: In this section we review the adiabatic regime in a general context and give a careful
561: derivation of its domain of validity.
562: 
563: 
564: 
565: 
566: The adiabatic approximation consists of (i) omitting the d'Alembertian
567: term in Eq.\ (\ref{eq:scalar2}), which gives an algebraic equation for $\phi$
568: that one can solve to obtain $\phi$ as a function of the density $\rho$;
569: (ii) omitting the terms involving the gradient of $\phi$ from the field
570: equation (\ref{ee}).  The resulting equations are the same as those of
571: Modified-Source gravity~ \cite{Carroll:2006jn}, in the approximation
572: where one considers only a dark matter source and neglects baryons.
573: The equations can be written as general relativity
574: coupled to a fluid with an effective energy density $\rho_{\rm eff}$
575: and effective pressure $p_{\rm eff}$:
576: \be
577: \mpl^2 G_{ab} =
578: \left[ (\rho_{\rm eff} + p_{\rm eff}) u_a u_b + p_{\rm eff} g_{ab} \right] \ ,
579: \label{ee1}
580: \ee
581: where
582: \be
583: \rho_{\rm eff}(\rho) = e^{\alpha[\phi_{\rm m}(\rho)]} \rho
584: + V[\phi_{\rm m}(\rho)] \ ,
585: \label{rhoeff}
586: \ee
587: \be
588: p_{\rm eff}(\rho) = - V[\phi_{\rm m}(\rho)] \ .
589: \label{peff}
590: \ee
591: Here $\phi_{\rm m}(\rho)$ is the solution of the algebraic
592: equation
593: \be
594: V_{\rm eff}'(\phi)= V'(\phi) + \alpha'(\phi) e^{\alpha(\phi)} \rho =0
595: \label{eq:alg}
596: \ee
597: for $\phi$.
598: Eliminating $\rho$ between Eqs.\ (\ref{rhoeff}) and (\ref{peff}) gives
599: the equation of state $p_{\rm eff} = p_{\rm eff}(\rho_{\rm eff})$.
600: 
601: 
602: In the adiabatic regime, the matter and scalar field are tightly
603: coupled together and evolve as one effective fluid.  By taking the
604: divergence of the field equation (\ref{ee1}), we see that this fluid
605: obeys the usual fluid equations of motion with the given effective
606: equation of state.
607: In a cosmological context, the effective fluid description (\ref{ee1}) is valid
608: for the background cosmology and for linear (and, indeed, nonlinear) perturbations. Therefore,
609: the equation of state of perturbations is the same as that of the
610: background cosmology, or the so-called entropy perturbation vanishes.
611: 
612: 
613: 
614: 
615: 
616: \subsection{Condition for global validity of adiabatic approximation}
617: \label{domain}
618: Of course, the adiabatic approximation will not be a good one for all
619: choices of $V(\phi)$ and $\alpha(\phi)$, or in all physical
620: situations. Our goal in this subsection is to establish criteria under
621: which we can trust the adiabatic approximation.
622: Roughly speaking, the mass of the scalar field
623: associated with the minimum of the effective potential must be
624: sufficiently large.
625: More precisely, we define the effective mass as a function of density
626: $\rho$ by
627: \be
628: m_{\rm eff}^2(\rho) = \left. \frac{\partial^2 V_{\rm eff}}{\partial \phi^2}
629:   (\phi,\rho) \right|_{\phi = \phi_{\rm m}(\rho)},
630: \label{meffdef}
631: \ee
632: where the derivatives are taken at constant $\rho$, and $\phi_{\rm
633:   m}(\rho)$ is the value of the scalar field which minimizes the
634: effective potential at a given density $\rho$.
635: We assume that $m_{\rm eff}^2$ is positive, otherwise there is no
636: local minimum of the potential and an adiabatic regime does not arise.
637: For a perturbation with lengthscale or timescale ${\cal L}$ and density
638: $\rho$ to be in the adiabatic regime, it is necessary that
639: \be
640: {\cal L} \gg m_{\rm eff}^{-1}(\rho).
641: \label{adiabaticok}
642: \ee
643: We call this condition the {\it local adiabatic condition}.\footnote{It has also
644: been called the Compton condition \cite{Hu:2007nk}, since the RHS of
645: Eq.\ (\ref{adiabaticok}) is the effective Compton wavelength of the field.}
646: If the condition is satisfied everywhere in spacetime,
647: then the adiabatic approximation will be good everywhere.  However if
648: the condition is satisfied in a local region, it does not necessarily
649: follow that the adiabatic approximation is valid in that region, due
650: to non-local effects \cite{Khoury:2003aq,Khoury:2003rn,Hu:2007nk}.  This is
651: discussed further in Sec.\ \ref{sec:nonlocal} below.
652: 
653: 
654: We now derive a slightly more precise version of the local adiabatic criterion
655: (\ref{adiabaticok}).  For a given density $\rho_0$ we define
656: $\phi_0 = \phi_{\rm m}(\rho_0)$ and
657: $\delta \phi = \phi - \phi_0$.
658: We expand the potential as
659: \be
660: V(\phi) = V_0 + V_1 \delta \phi + \frac{1}{2} V_2 \delta \phi^2 +
661: O(\delta \phi^3) \ ,
662: \ee
663: and defining $W(\phi) = e^{\alpha(\phi)}$ we similarly expand
664: \be
665: W(\phi) = W_0 + W_1 \delta \phi + \frac{1}{2} W_2 \delta \phi^2 +
666: O(\delta \phi^3) \ .
667: \ee
668: The effective mass (\ref{meffdef}) is then given by
669: \be
670: m_{\rm eff}^2 \equiv \frac{\partial^2 V_{\rm eff}}{\partial \phi^2}(\phi_0,\rho_0) = V_2 + \rho_0 W_2 \ ,
671: \ee
672: and
673: the condition (\ref{eq:alg}) that the effective potential be minimized yields
674: \be
675: V_1 = -\rho_0 W_1.
676: \label{ident}
677: \ee
678: Also a short computation using the definition (\ref{eq:alg}) of the function $\phi_{\rm m}(\rho)$ gives
679: \be
680: \frac{d \phi_{\rm m}}{d\rho} = - \frac{W_1}{V_2 + \rho_0
681:   W_2} = - \frac{ \alpha_0^\prime e^{\alpha_0}}{m_{\rm eff}(\rho_0)^2} \ ,
682: \label{derivformula}
683: \ee
684: where $\alpha_0 = \alpha(\phi_0)$ and $\alpha_0^\prime = \alpha'(\phi_0)$.
685: We also define ${\cal L}$ to be the smallest lengthscale or timescale over which the density $\rho$
686: changes, so that $\nabla_a \nabla^a \rho \sim  \rho/{\cal L}^2$ and
687: $(\nabla \rho)^2 \sim \rho^2 / {\cal L}^2$.
688: 
689: 
690: 
691: 
692: The field equations are the Einstein equation
693: (\ref{ee}), the scalar field equation (\ref{eq:scalar1}), and the
694: fluid equations (\ref{eq:fluid1}) and (\ref{eq:fluid2}).
695: These equations are not all independent, since the Einstein equation
696: enforces conservation of the total stress energy tensor.  We will take
697: as the independent equations just the Einstein equation (\ref{ee}) and
698: the fluid equations (\ref{eq:fluid1}) and (\ref{eq:fluid2}), since the
699: scalar field equation can be derived from these.\footnote{Similarly,
700: the approximate form (\ref{eq:alg}) of the scalar field equation can be
701: obtained from the approximate form (\ref{ee1}) of the Einstein equation
702: together with the fluid equations.}
703: 
704: 
705: Therefore, to justify the
706: adiabatic approximation, it is sufficient to justify dropping the scalar
707: field derivative terms in the Einstein equation (\ref{ee}).
708: The
709: ratio of the scalar field gradient terms to the potential term
710: evaluated at the adiabatic solution is of order
711: \bea
712: \frac{ (\nabla \phi)^2}{V(\phi)} \sim
713: \frac{1}{V_0} \left( \frac{d \phi_{\rm
714:         m}}{d \rho} \right)^2 ( \nabla \rho )^2
715: \sim
716: \frac{1}{V_0 {\cal L}^2} \left( \frac{d \phi_{\rm
717:         m}}{d \ln \rho} \right)^2.\ \ \
718: \label{einsteincheck}
719: \eea
720: By combining Eqs.\ (\ref{ident}) and (\ref{derivformula}) we obtain
721: \be
722: \frac{d \phi_{\rm m}}{d \ln \rho} = \frac{V_1}{m_{\rm eff}^2},
723: \ee
724: and using this to eliminate one of the factors of $d \phi_{\rm m} / d
725: \ln \rho$ from Eq.\ (\ref{einsteincheck}) gives
726: \bea
727: \frac{ (\nabla \phi)^2}{V(\phi)} \sim
728: \frac{V_1}{V_0} \frac{d \phi_{\rm
729:         m}}{d \ln \rho} \left( \frac{1}{m_{\rm eff}^2 {\cal L}^2}
730:     \right)
731: \sim \frac{ d \ln V }{d \ln \rho}
732: \left( \frac{1} {m_{\rm eff}^2 {\cal L}^2 } \right).\ \ \ \
733: \label{einsteincheck1}
734: \eea
735: Thus, the adiabatic approximation will be valid whenever
736: \be
737: \frac{ d \ln V }{d \ln \rho} \left(
738: \frac{1} {m_{\rm eff}^2 {\cal L}^2 }\right)  \ll 1.
739: \label{einsteincheck1a}
740: \ee
741: Now, since the first factor on the left hand side
742: involves derivatives of logarithmic
743: factors, we expect this prefactor to generically be of order unity.
744: When this is true the condition (\ref{einsteincheck1a}) reduces to the
745: local adiabatic condition (\ref{adiabaticok}).
746: 
747: 
748: 
749: 
750: 
751: 
752: 
753: \subsection{Nonlocal condition for breakdown of adiabatic approximation}
754: \label{sec:nonlocal}
755: 
756: As mentioned above, the condition ${\cal L} \gg m_{\rm eff}^{-1}$ in a
757: local region is a necessary but not a sufficient condition for the validity of
758: the adiabatic approximation in that region.  This is because
759: the corrections to the adiabatic approximation are determined by a
760: wave equation obtained by perturbing Eq.\ (\ref{eq:scalar1}) whose solutions depend in a non-local
761: way on its sources.  The corrections in a given region can become
762: large due to a breakdown of the local adiabatic condition (\ref{adiabaticok})
763: that occurs elsewhere \cite{Hu:2007nk}.
764: 
765: For the special case of static, spherically symmetric
766: systems, and for a constant density object in a constant density background,
767: the chameleon field papers \cite{Khoury:2003aq,Khoury:2003rn}
768: derived a precise condition for the validity of the adiabatic approximation
769: inside the system, the so-called ``thin shell'' condition, which depends on the asymptotic value
770: of the potential.  The thin-shell condition is both a necessary and
771: sufficient condition, but it restricted to static situations.  Below
772: we will show that for several specific examples
773: the thin-shell condition and the adiabatic condition
774: (\ref{adiabaticok}) give the same predictions in order
775: of magnitude for the boundary of the adiabatic regime.
776: 
777: 
778: We now discuss an order-of-magnitude non-local criterion for static,
779: spherically symmetric situations that predicts when
780: the adiabatic approximation breaks down even when the local adiabatic condition
781: (\ref{adiabaticok}) is satisfied.
782: The condition is a generalization of a condition derived
783: by Hu \cite{Hu:2007nk} in the context of $f(R)$ modified gravity models, and is
784: also a generalization of the thin-shell condition of Refs.\ \cite{Khoury:2003aq,Khoury:2003rn}.
785: 
786: 
787: Consider a spherically symmetric density profile $\rho(r)$, which we
788: will assume for simplicity is monotonically decreasing as $r$
789: increases.  From this density profile we can compute the corresponding
790: adiabatic scalar field profile
791: \be
792: \phi_{\rm ad}(r) \equiv \phi_{\rm m}[\rho(r)].
793: \label{adprofile}
794: \ee
795: If the local adiabatic condition (\ref{adiabaticok}) is
796: satisfied for all $r$, then this adiabatic field is a good
797: approximation to the actual solution $\phi(r)$.  Suppose therefore
798: that the local adiabatic condition is violated\footnote{If the density
799:   goes to a constant at large $r$, then the local adiabatic condition
800:   is always satisfied as $r \to \infty$.} for some set of values
801: ${\bar r}_1 < r < {\bar r}_2$.  Then typically what occurs is
802: that there is a larger interval $r_1 < r < r_2$ with $r_1 < {\bar
803:   r}_1$ and $r_2 > {\bar r}_2$ on which the solution $\phi(r)$ differs
804: significantly from the adiabatic field profile (\ref{adprofile}).
805: For a given interval $(r_1,r_2)$,
806: significant deviations from the adiabatic approximation should occur in the vicinity of $r=r_1$ if
807: \be
808: \phi_2 - \phi_1 \agt \frac{ \alpha_2^\prime e^{\alpha_2}}{r_1} \int_{r_1}^{r_2} r^2 \left[ \rho(r) -
809: \rho_2\right],
810: \label{nonlocal}
811: \ee
812: where $\rho_i = \rho(r_i)$, $\phi_i = \phi_{\rm ad}(r_i)$, $\alpha_i =
813: \alpha(\phi_i)$ and $\alpha_i^\prime = \alpha^\prime(\phi_i)$, $i = 1,2$.
814: The criterion assumes that $r_2 \gg r_1$ and $\rho_2 \ll \rho_1$.
815: 
816: 
817: The criterion (\ref{nonlocal}) is a generalization of other criteria that have
818: appeared in the literature. First, if one assumes that the density
819: varies on a lengthscale of order $\sim r$, then using this to
820: approximate the integral in (\ref{nonlocal}) gives
821: \be
822: \phi_2 - \phi_1 \agt \alpha_2^\prime e^{\alpha_2} r_1^2 (\rho_1 -
823: \rho_2).
824: \label{nonlocal1}
825: \ee
826: Equation (\ref{nonlocal1}) is the condition derived in Ref.\ \cite{Hu:2007nk} in the context of
827: $f(R)$ models, although there the factor of $\alpha_2^\prime
828: e^{\alpha_2}$ was neglected.  The condition (\ref{nonlocal1}) was
829: found to reliably predict the onset of deviations from the adiabatic
830: profile in numerical solutions for the Solar System and the Galaxy\cite{Hu:2007nk}.
831: 
832: 
833: Second, one can derive from Eq.\ (\ref{nonlocal}) the
834: thin-shell condition of Refs.\ \cite{Khoury:2003aq,Khoury:2003rn}, up
835: to a factor of order unity.
836: Suppose one has a uniform density sphere of density $\rho_c$ and
837: radius $R_c$, embedded in a uniform density, infinite medium of
838: density $\rho_\infty$.  For this case the local adiabatic criterion is
839: satisfied everywhere except at the point of discontinuity of the
840: density at $r = R_c$, so we expect a breakdown of the adiabatic
841: approximation near this point.  Following Ref.\ \cite{Khoury:2003aq}
842: we assume that $\alpha^\prime = \beta/\mpl$ is a constant, and that we
843: are in the regime where $\alpha(\phi) \ll 1$, so $e^\alpha \approx
844: 1$.  If we apply the criterion (\ref{nonlocal}) to an interval
845: $(r_1,r_2)$ with $r_1 < R_c < r_2$, we obtain that
846: the adiabatic approximation should fail near $r = r_1$
847: if
848: \be
849: \frac{ \Delta \phi} {\beta \mpl \Phi_c} \agt \frac{R_c^3 - r_1^3}{r_1
850:   R_c^2},
851: \label{thinshell1}
852: \ee
853: where $\Phi_c \sim \rho_c R_c^2/\mpl^2$ is the Newtonian potential at
854: the center of the sphere.  If the left hand side is large compared to
855: unity, then Eq.\ (\ref{thinshell1}) will be satisfied for all values
856: of $r_1$ except very close to $r_1=0$, and it follows that the
857: adiabatic approximation will not apply in most of the interior of the
858: sphere.  On the other hand, if the left hand side of Eq.\
859: (\ref{thinshell1}) is small compared to unity, then it follows that
860: one would expect the adiabatic approximation to be valid throughout the interior of the
861: sphere except very close to the surface, in a thin shell of thickness
862: $\sim R_c \Delta \phi / (\beta \mpl \Phi_c)$.
863: Both of these conclusions agree in order of magnitude with those of
864: Refs.\ \cite{Khoury:2003aq,Khoury:2003rn} \footnote{This example shows
865: that the condition (\ref{nonlocal1}) derived in Ref.\ \cite{Hu:2007nk} is less
866: general than the condition (\ref{nonlocal}) derived here, since that
867: condition cannot be used to derive the thickness of the thin shell.}.
868: 
869: 
870: We now turn to the derivation of the non-local condition
871: (\ref{nonlocal}).
872: We start by noting that static
873: solutions of the equation of motion (\ref{eq:scalar2}) can be obtained
874: by extremizing the energy functional
875: \be
876: E = \int dr \, r^2 \, \left[ \frac{1}{2} \left( \frac{d \phi}{d r}
877:   \right)^2 + V(\phi) + e^{\alpha(\phi)} \rho(r) \right].
878: \ee
879: The basic idea \cite{Hu:2007nk} is that the adiabatic field profile (\ref{adprofile})
880: minimizes just the potential energy, and there some kinetic energy
881: cost for following this adiabatic profile.  When this kinetic energy cost becomes
882: sufficiently large, it becomes energetically favorable for the field
883: to switch to a different, non-adiabatic profile with a smaller kinetic
884: energy and with a larger potential energy, for a net gain in energy.
885: 
886: 
887: We now compute the total energies $E_{\rm ad}$ for the adiabatic
888: profile and $E_{\rm trial}$ for an alternative trial profile which is
889: qualitatively similar to numerical solutions \cite{Hu:2007nk}.  When
890: \be
891: E_{\rm trial} - E_{\rm ad} < 0
892: \label{condt0}
893: \ee
894: there is a net gain in energy for switching to the trial profile, and
895: so the adiabatic profile is no longer a good approximation.
896: The trial profile is simply $\phi_{\rm trial}(r) = \phi_2 = $ constant for
897: $r_1 + \Delta r \le r \le r_2$, and $\phi_{\rm trial}(r) = \alpha - \beta/r$ for
898: $r_1 \le r \le r + \Delta r$,
899: with $\alpha$ and $\beta$ chosen to satisfy
900: continuity at $r_1$ and $r_1 + \Delta r$, $\phi_{\rm trial}(r_1) = \phi_1$ and
901: $\phi_{\rm trial}(r_1 + \Delta r) = \phi_2$.  This trial profile depends on one
902: parameter, namely the width $\Delta r$ of the region in which the
903: field transitions from $\phi_1$ to $\phi_2$.
904: We obtain
905: \bea
906: \label{long}
907: E_{\rm trial} - E_{\rm ad} &=&
908: \int_{r_1}^{r_2} dr r^2 \left\{  - \frac{1}{2}  \phi_{\rm ad}'(r)^2   \right. \nn \\
909: &&   + \alpha_2^\prime e^{\alpha_2} \left[ \phi_{\rm trial}(r) - \phi_{\rm
910:       ad}(r) \right] \left[ \rho(r) -\rho_2 \right]  \bigg\} \nn \\
911: && + \frac{r_1 (r_1 + \Delta r) (\phi_2 - \phi_1)^2 }{2 \Delta r}.
912: \eea
913: In deriving this formula we have used an approximate Taylor expansion
914: of the effective potential about the point $\phi_2$ together with Eq.\ (\ref{eq:alg}):
915: \bea
916: V_{\rm eff}(\phi_2,\rho) - V_{\rm eff}(\phi,\rho)
917: &\approx& ( \phi_2 - \phi) \frac{\partial
918:   V_{\rm eff}}{\partial \phi}(\phi_2,\rho) \nn \\
919: &=&  \alpha_2^\prime e^{\alpha_2} ( \phi_2 - \phi) (\rho - \rho_2).\ \ \ \ \ \
920: \eea
921: 
922: Consider first the kinetic energy contributions, the first and third
923: lines of Eq.\ (\ref{long}).  The minimum kinetic energy for
924: any field configuration which satisfies $\phi(r_1) = \phi_1$ and $\phi(r_2) = \phi_2$ is
925: the energy of the Laplace equation solution with these boundary
926: conditions\footnote{This can be shown using the Schwarz inequality
927:   $(\int f g dr)^2 \le \int f^2 dr \, \int g^2 dr$, with $f = r
928:   \phi'(r)$ and $g = 1/r$.}, namely
929: $E_{\rm K,min} = (\Delta \phi)^2 r_1 r_2 / (2 (r_2 - r_1))$, where
930: $\Delta \phi = \phi_2 - \phi_1$.  Therefore we can write the kinetic energy of the
931: adiabatic field profile [the negative of the first line of Eq.\ (\ref{long})] as
932: $E_{\rm K,ad} = E_{\rm K,min} (1 + \varepsilon)$, where $\varepsilon$
933: is dimensionless and nonnegative.
934: We will assume that $\varepsilon \agt 1$, since $\varepsilon \ll 1$ would require
935: the adiabatic profile $\phi_{\rm ad}(r)$ to be very close to the $1/r$
936: profile, which would be a fine tuning.  For generic density profiles
937: we expect $\varepsilon \agt 1$.
938: If we now choose $\Delta r = 2 r_1 /
939: \varepsilon$, then it follows that the net gain in kinetic energy is
940: \be
941: \varepsilon \frac{E_{\rm K,min}}{2} \agt E_{\rm K,{\rm min}} \sim r_1 \Delta
942: \phi^2,
943: \label{KEgain}
944: \ee
945: using $r_2 \gg r_1$.
946: 
947: Turn now to the potential energy contributions, the second line of
948: Eq.\ (\ref{long}).  If we assume that the potential $V(\phi)$ is
949: monotonically decreasing, as in the Chameleon field models, then it
950: follows that $\phi_{\rm ad}(r)$ is an increasing function of $r$, so
951: that
952: \be
953: \phi_{\rm trial}(r) - \phi_{\rm ad}(r) \le \phi_2 - \phi_1.
954: \ee
955: This implies that the potential energy term in Eq.\ (\ref{long}) is
956: bounded above by
957: $$
958:  \alpha_2^\prime e^{\alpha_2} \Delta \phi \int_{r_1}^{r_2} r^2 \left[ \rho(r) -
959: \rho_2\right].
960: $$
961: Inserting this together with the estimate (\ref{KEgain}) of the kinetic
962: energy gain into Eq.\ (\ref{condt0}) yields the criterion (\ref{nonlocal}).
963: 
964: 
965: 
966: \subsection{Observed equation of state parameter}
967: 
968: 
969: In this section we derive the observed equation of state parameter
970: $w_{\rm obs}$ for these models, in the adiabatic regime, for the case $\alpha_b =0$.
971: Generically we have $w_{\rm
972:   obs} < -1$ corresponding to superacceleration, as previously noted
973: in Ref.\ \cite{Das:2005yj}.
974: 
975: The background cosmological evolution is given in the Einstein frame
976: by the equation
977: \be
978: 3 \mpl^2 H^2 = V + e^{\alpha} \frac{\rho_0}{a^3},
979: \label{eee1}
980: \ee
981: where $\rho_0$ is a constant, together with the evolution equation for
982: the scalar field.  Observations of the acceleration of the Universe
983: are fit to the model
984: \be
985: 3 \mpl^2 H^2 = \frac{\rho_1}{a^3} + \rho_{\rm DE}(a)
986: \label{eee2}
987: \ee
988: where $\rho_1 = e^{\alpha_0} \rho_0$ is the observed matter density
989: today\footnote{Note that in this model some fraction of the mass
990:   density in the $V$ term in Eq.\ (\ref{eee1}) will cluster, so
991:   measurements of mass density using clusters will not correspond
992:   exactly to measurements of the second term in Eq.\ (\ref{eee1}).  We
993:   will not address this issue here.}, $\alpha_0$ is the value of $\alpha$ today,
994: and $\rho_{\rm DE}$
995: is the inferred ``dark energy density''.
996: The equation of state parameter is then given by
997: \be
998: w_{\rm obs}(a) = -1 - \frac{1}{3} \frac{d \ln \rho_{\rm DE}}{d \ln a}.
999: \label{eee3}
1000: \ee
1001: Combining Eqs.\ (\ref{eee1}) -- (\ref{eee3}) and using Eq.\ (\ref{eq:alg})
1002: to eliminate the term proportional to $d\phi/da$ gives
1003: \be
1004: w_{\rm obs} = \frac{- V}{ V + \frac{e^{\alpha_0} \rho_0}{a^3}
1005:   (e^{\alpha - \alpha_0} -1)}.
1006: \ee
1007: Next we use Eq.\ (\ref{eq:alg}) again to obtain $\rho_0/a^3 = -
1008: e^{-\alpha} V'(\phi) / \alpha'(\phi)$, which gives
1009: \be
1010: w_{\rm obs} = \frac{-1}{1 - \frac{d \ln V}{d \alpha} ( 1 - e^{\alpha_0
1011:     - \alpha})}.
1012: \ee
1013: 
1014: 
1015: This formula has the property that $w_{\rm obs}=-1$ today for all
1016: models. Also expanding to first order about $a=1$ we obtain $1/w_{\rm
1017:   obs} = -1 + \ln(V/V_0)$, where $V_0$ is the value of $V$ today, so
1018: $w_{\rm obs} < -1$ in the past since $V = V[\phi_{\rm m}(\rho)]> V_0$ in the past.
1019: 
1020: 
1021: 
1022: 
1023: 
1024: \section{Adiabatic instability}
1025: \label{instability}
1026: 
1027: 
1028: In the adiabatic regime, the models (\ref{action0}) discussed here can
1029: exhibit instabilities on
1030: small scales characterized by a negative sound speed squared of the
1031: effective coupled fluid.
1032: This instability extends down to the
1033: smallest scales for which the adiabatic approximation is valid.
1034: Starting with a uniform fluid, the instability will give rise to
1035: exponential growth of small perturbations.  The final state of the
1036: coupled fluid is beyond the scope of this paper.
1037: Theories that exhibit this instability are typically ruled out as
1038: models of dark energy.
1039: 
1040: \subsection{Hydrodynamic viewpoint}
1041: \label{sec:hydro}
1042: 
1043: This adiabatic instability was first discovered by Afshordi,
1044: Zaldarriaga and Kohri \cite{Afshordi:2005ym} in a context slightly
1045: different to that considered here.  That context was the mass varying neutrino model of
1046: dark energy, where a dynamical dark energy model is obtained by
1047: coupling a light scalar field to neutrinos but not to dark matter.
1048: The instability was previously discussed in the context of the models
1049: considered here by Kaplinghat and Rajaraman \cite{Kaplinghat:2006jk}.
1050: In this paper we will generalize
1051: the treatment of the instability given in Ref.\ \cite{Kaplinghat:2006jk}.
1052: 
1053: A fairly simple form of the stability criterion can be obtained by
1054: writing the potential $V(\phi)$ as a function $V(\alpha)$ of the
1055: coupling function $\alpha(\phi)$ by eliminating $\phi$.
1056: This gives
1057: \be
1058: \rho_{\rm eff} = V + e^{ \alpha} \rho = V - \frac{
1059:   dV/d\phi}{d\alpha/d\phi} = V - \frac{dV}{d\alpha} \ ,
1060: \ee
1061: where we have used Eqs.\ ~(\ref{rhoeff}) and (\ref{eq:alg}).  The
1062: square $c_a^2$ of the adiabatic sound speed is then given by
1063: \be
1064: \frac{1}{c_a^2} = \frac{d \rho_{\rm eff}}{dp_{\rm eff}} = \frac{
1065:   d\rho_{\rm eff}/d\alpha} {dp_{\rm eff}/d\alpha} = \frac
1066: {\frac{d}{d\alpha} \left[ V - \frac{dV}{d\alpha} \right] }
1067: {\frac{d}{d\alpha} \left[ -V \right] } = -1 + \frac{ \frac{d^2 V}{d\alpha^2} }{
1068:   \frac{dV}{d\alpha}} \ .
1069: \label{soundspeed}
1070: \ee
1071: The system will be unstable if
1072: \be
1073: c_a^2 <0.
1074: \label{unstablec}
1075: \ee
1076: Equation (\ref{soundspeed}) gives a simple prescription for computing when a
1077: given theory will be unstable, assuming it is in the adiabatic
1078: regime.  This equation furnishes $c_a^2$ as a function of $\alpha$,
1079: which can be re-expressed as a function of $\phi$ using $\alpha =
1080: \alpha(\phi)$, and then as a function of density $\rho$ using $\phi =
1081: \phi_{\rm m}(\rho)$ from Eq.\ (\ref{eq:alg}).
1082: 
1083: More generally, outside of the adiabatic regime, the
1084: effective sound speed is the ratio of the local pressure and density
1085: perturbations
1086: $$
1087: c_s^2 = \frac{\delta p}{\delta \rho}
1088: $$
1089: and can differ from $c_a^2$.  In a cosmological context we have
1090: $c_a^2\equiv \dot{P}/\dot{\rho}$ and $c_s^2(k,a)\equiv \delta
1091: P(k,a)/\delta\rho(k,a)$, where $k$ is spatial wavenumber and $a$ is
1092: scale factor.  For most of this paper we will consider only the
1093: adiabatic regime in which $c_s^2 \to c_a^2$, although the more general
1094: regime will be probed in our numerical integrations in Sec.\ \ref{examples}.
1095: 
1096: 
1097: We now argue that the adiabatic sound speed (\ref{soundspeed}) is {\it always}
1098: negative in the adiabatic regime.
1099: In the definition (\ref{meffdef}) of the effective mass, we rewrite
1100: the $\phi$ derivative in terms of $\alpha$ derivatives using
1101: $\alpha = \alpha(\phi)$.
1102: Using the fact that the first derivative of $V_{\rm eff}$ with respect to $\phi$ vanishes at
1103: $\phi = \phi_{\rm m}(\rho)$, we obtain
1104: \be
1105: m_{\rm eff}^2 = \left( \frac{ d\alpha }{d \phi} \right)^2
1106: \frac{\partial ^2 V_{\rm eff}}{\partial \alpha^2},
1107: \ee
1108: where the derivatives are taken at constant $\rho$.
1109: Using Eq.\ (\ref{phieqn}) and then eliminating $\rho$ using Eq.\
1110: (\ref{eq:alg})
1111: gives
1112: \be
1113: m_{\rm eff}^2 = \left( \frac{ d\alpha }{d \phi} \right)^2
1114: \left[ \frac{d^2 V}{d\alpha^2} - \frac{d V}{d\alpha} \right].
1115: \label{meff2}
1116: \ee
1117: Simplifying using the expression (\ref{soundspeed}) for the sound speed squared
1118: gives
1119: \be
1120: m_{\rm eff}^2 = \left( \frac{ d\alpha }{d \phi} \right)^2
1121:  \frac{d V}{d\alpha} \frac{1}{c_s^2}.
1122: \label{meff3}
1123: \ee
1124: The adiabatic instability arises in the
1125: case $c_s^2<0$ and $m_{\rm eff}^2>0$. [When $m_{\rm eff}^2<0$ there is
1126: no adiabatic regime, in the sense that adiabatic solutions are subject
1127: to a tachyonic instability.  This instability
1128: has been discussed in the context of $f(R)$ gravity models in Refs.\
1129: \cite{Dolgov:2003px,Seifert:2007fr,Sawicki:2007tf}.]  It follows from Eqs.\
1130: (\ref{soundspeed}) and (\ref{meff3}) that the adiabatic instability
1131: occurs when
1132: \be
1133: \frac{dV}{d\alpha} < \frac{d^2 V}{d\alpha^2} < 0,
1134: \label{cc1}
1135: \ee
1136: or when
1137: \be
1138: \frac{dV}{d\alpha} < 0 < \frac{d^2 V}{d\alpha^2}.
1139: \label{cc2}
1140: \ee
1141: 
1142: However, the region in the $(V_{,\alpha},V_{,\alpha\alpha})$ parameter
1143: space defined by the union of the regions (\ref{cc1}) and (\ref{cc2}) is
1144: just the adiabatic regime; the adiabatic approximation requires
1145: both that $V_{,\alpha}<0$ so that the effective potential has a local
1146: extremum, and also that $m_{\rm eff}^2>0$ so that the local extremum is a
1147: local minimum.  It follows that the adiabatic instability condition
1148: (\ref{unstablec}) is satisfied throughout the adiabatic regime.
1149: This genericity of the instability has also been deduced by Kaplinghat
1150: and Rajaraman \cite{Kaplinghat:2006jk} using a different method.
1151: However, as we discuss in the following subsection, the instability
1152: actually occurs only if the coupling $\alpha'(\phi)$ is sufficiently
1153: large, a point missed in Refs.\ \cite{Afshordi:2005ym,Kaplinghat:2006jk}.
1154: 
1155: 
1156: \subsection{Scales over which the instability operates}
1157: \label{sec:scales}
1158: 
1159: For the instability to be relevant on some spatial scale ${\cal L}$, then we must have
1160: \be
1161: {\cal L} \gg m_{\rm eff}^{-1}
1162: \ee
1163: in order that the adiabatic approximation be valid, as discussed
1164: above.  There is also an upper bound on the range of spatial scales
1165: which comes about as follows.  When the instability is present,
1166: spatial Fourier modes with wavelength ${\cal L}$ grow exponentially on
1167: a timescale
1168: \be
1169: \tau \sim \frac{{\cal L}}{\sqrt{|c_s^2|}} \ ,
1170: \ee
1171: where $c_s^2$ is given by Eq.~(\ref{soundspeed}).  If this timescale
1172: is longer than the Hubble time $H^{-1}$, then the mode does
1173: not have time to grow and the instability is not relevant.
1174: Therefore the range of scales over which the instability operates is
1175: \be
1176: m_{\rm eff}^{-1} \ll {\cal L} \ll \frac{\sqrt{|c_s^2|}}{H} \ .
1177: \ee
1178: More generally,
1179: for a fluid of density $\rho$, if the instability is to be unmodified
1180: by the gravitational dynamics of the fluid, then the instability
1181: timescale must be shorter than the gravitational dynamical time, which
1182: from Eq.\ (\ref{ee1}) is $\sim \mpl/\sqrt{\rho_{\rm eff}(\rho)}$.
1183: Here $\rho_{\rm eff}(\rho)$ is the total mass density (\ref{rhoeff}) of the
1184: coupled dark matter-dark energy fluid.
1185: This gives the criterion
1186: \be
1187: m_{\rm eff}(\rho)^{-1} \ll {\cal L} \ll \frac{\mpl
1188:   \sqrt{|c_s^2(\rho)|}}{\sqrt{\rho_{\rm eff}(\rho)}} \ ,
1189: \label{range}
1190: \ee
1191: which determines the values of density $\rho$ and lengthscale ${\cal
1192:   L}$ for which the instability operates.
1193: The upper lengthscale can be rewritten using Eqs.\ (\ref{meff3}),
1194: (\ref{rhoeff}) and (\ref{eq:alg}) to give
1195: \be
1196: \frac{1}{m_{\rm eff}(\rho)} \ll  {\cal L} \ll \frac{\mpl
1197: |\alpha^\prime[\phi_{\rm m}(\rho)]|}{m_{\rm eff}(\rho)}
1198: \sqrt{\frac{1}{1 - \frac{1}{\frac{d\ln V}{d \alpha}}}}.
1199: \label{range1}
1200: \ee
1201: Here the quantity $d \ln V / d\alpha(\alpha)$
1202: on the right hand side is
1203: expressed as a function of $\phi$ using $\alpha = \alpha(\phi)$, and
1204: then as a function of the density using $\phi = \phi_{\rm m}(\rho)$.
1205: 
1206: At longer lengthscales, it is possible that the negative sound speed squared still
1207: engenders an instability, but determining this requires a stability
1208: analysis of the system in question including the effects of self
1209: gravity. The results might vary from one system
1210: to another.
1211: In this paper we will restrict attention to the regime (\ref{range}) where
1212: the presence of the instability can be easily diagnosed.
1213: 
1214: The factor in square brackets in Eq.\ (\ref{range1}) is always smaller
1215: than unity, since $V$ is assumed to be a decreasing function of $\phi$
1216: and hence also of $\alpha$.  It follows that the ratio of the maximum
1217: lengthscale ${\cal L}_{\rm max}$ to the minimum lengthscale ${\cal
1218:   L}_{\rm min}$ satisfies
1219: \be
1220: \frac{{\cal L}_{\rm max}}{{\cal L}_{\rm min}} \le \mpl
1221: |\alpha'[\phi_{\rm m}(\rho)]|.
1222: \ee
1223: Hence in order for there to be a nonempty regime in which the
1224: instability operates, the strength of the coupling of the field to the
1225: dark matter must be much stronger than gravitational strength,
1226: \be
1227: \mpl |\alpha' |\gg 1,
1228: \label{strongcoupling}
1229: \ee
1230: as discussed in the introduction.  In Sec.\ \ref{jeans} below we give a simple
1231: explanation for this requirement.
1232: In Sec.\ \ref{sec:MaVaN} below we give an example of a model where we confirm
1233: numerically instability is present at strong coupling but not when the
1234: coupling is weak.
1235: 
1236: 
1237: 
1238: \subsection{Jeans instability viewpoint}
1239: \label{jeans}
1240: 
1241: There are two different ways of describing and understanding the instability, depending
1242: on whether one thinks of the
1243: scalar-field mediated forces as being ``gravitational'' forces or ``pressure'' forces.
1244: From one point of view, that of the Einstein frame description, the
1245: instability is independent of gravity.  This can be seen from the
1246: equation of motion (\ref{ee1}); the instability is present even when
1247: the (Einstein-frame) metric perturbation due to the fluid can be
1248: neglected.  In the adiabatic regime the acceleration due the scalar
1249: field is a gradient of a local function of the density [cf.\ Eq.\
1250: (\ref{euler1}) below], which can be thought of as a pressure.  The net
1251: effect of the scalar interaction is to give a contribution to the
1252: specific enthalpy $h(\rho) = \int dp/\rho$ of any fluid which is independent
1253: of the composition of the fluid.
1254: If net sound speed squared of the
1255: fluid is negative, then there exists an instability in accord with our
1256: usual hydrodynamic intuition.
1257: 
1258: 
1259: From another point of view, however, that of the Jordan
1260: frame description, the instability involves gravity.
1261: The gravitational force in this frame is mediated partly by a tensor
1262: interaction and partly by a scalar interaction.
1263: The effective Newton's constant describing the interaction of dark matter with itself is
1264: \be
1265: G_{cc} = G \left[ 1
1266:   + \frac{2 \mpl^2 \alpha^\prime(\phi)^2 }{1 +
1267: \frac{m_{\rm eff}^2}{ {\bf k}^2 }
1268: } \right],
1269: \label{Gformula0}
1270: \ee
1271: where ${\bf k}$ is a spatial wavevector.
1272: This is Eq.\ (\ref{Gformula}) of Appendix \ref{sec:Geff} specialized to
1273: ${\rm i} = {\rm j} = c$, with $\alpha_c$ written just as $\alpha$, and
1274: specialized to $\alpha_b=0$ (since
1275: experiments tell us that $|\alpha_b'| \alt 10^{-2} \mpl^{-1}$ today).
1276: Here the 1 in the square brackets describes the tensor interaction and
1277: the second term the scalar interaction.
1278: At long lengthscales, $k \ll m_{\rm eff}/(\mpl |\alpha'|)$, the scalar interaction is
1279: suppressed and we have $G_{cc} \approx G$.  At short lengthscales,
1280: $k \gg m_{\rm eff}$, the scalar field is effectively massless and
1281: $G_{cc}$ asymptotes to a constant, $G_{cc} \approx G [1 + 2 \mpl^2 (\alpha^\prime)^2 ]$.
1282: However, when $\mpl |\alpha^\prime| \gg 1$ there is an intermediate range of
1283: lengthscales,
1284: \be
1285: \frac{m_{\rm eff}} {\mpl |\alpha^\prime|} \ll k \ll m_{\rm eff}
1286: \label{range3}
1287: \ee
1288: in which the effective Newton's constant increases linearly with
1289: $k^2$,
1290: \be
1291: G_{cc} \approx G \frac{2 \mpl^2 (\alpha^\prime)^2}{m_{\rm eff}^2}
1292: {\bf k}^2.
1293: \label{Gcc1}
1294: \ee
1295: A gravitational interaction with $G_{cc} \propto {\bf k}^2$ behaves just
1296: like a (negative) pressure in the hydrodynamic equations.  This
1297: explains why the the effect of the scalar interaction can be thought
1298: of as either pressure or gravity in the range of scales (\ref{range3}).
1299: Note that the range of scales (\ref{range3}) coincides with with the
1300: range (\ref{range1}) derived above, up to a logarithmic correction factor.
1301: 
1302: From this second, Jordan-frame point of view, the instability is
1303: simply a Jeans instability.  In a cosmological background with
1304: Einstein-frame metric $ds^2 = -dt^2 + a(t)^2 d{\bf x}^2$, the evolution equation for
1305: the CDM fractional density perturbation $\delta$ with comoving
1306: wavenumber $k_c$ on subhorizon scales in the adiabatic limit is \cite{Brax:2005ew,paperIII}
1307: \be
1308: {\ddot \delta} + 2 H {\dot \delta} - 4 \pi G_{\rm cc} e^\alpha \rho
1309: \delta = 0,
1310: \label{perts}
1311: \ee
1312: where $H = {\dot a}/a$.  Here $G_{cc}$ is given by the expression
1313: (\ref{Gformula0}) evaluated at the physical wavenumber $k= k_c/a$,
1314: and we have neglected photons and baryons.
1315: 
1316: 
1317: Now in the absence of the Hubble damping term in Eq.\ (\ref{perts}),
1318: the gravitational
1319: interaction described by the last term would cause an exponential
1320: growth of the mode, the usual Jeans instability of uniform fluid.
1321: Normally in a cosmological context, the Hubble damping term
1322: is present and the timescale $\sim 1/H$ associated with this term is
1323: of the same order as the timescale $1/ \sqrt{G \rho}$ associated with
1324: the gravitational interaction in the last term.  Because of this
1325: equality of timescales, the exponential growth is converted to power
1326: law growth by the Hubble damping.  In the present context, however,
1327: things work differently.  The gravitational constant governing the
1328: gravitational self-interaction of the mode is $G_{cc}(k)$ instead of
1329: $G$, and consequently the timescale associated with the last term in
1330: Eq. (\ref{perts}) is shorter than the Hubble damping time by
1331: a factor of
1332: \be
1333: \sim \sqrt{\frac{G_{cc} }{ G }} \sim \frac{ k \mpl |\alpha^{\prime}|  }{ m_{\rm
1334:   eff}} \gg 1.
1335: \ee
1336: Therefore the Hubble damping is ineffective and the Jeans instability
1337: causes approximate exponential growth rather than power law growth.
1338: 
1339: 
1340: The above discussion can also be cast in terms a scale-dependent sound speed instead of a scale
1341: dependent Newton's constant; this clarifies the relation to our previous
1342: discussion of Secs.\ \ref{sec:hydro} and \ref{sec:scales}.
1343: As a slight generalization, consider
1344: a fluid with an intrinsic sound speed $c_{\rm s,in}$.  Then
1345: the evolution equation (\ref{perts}) generalizes to (see Appendix \ref{visible})
1346: \be
1347: {\ddot \delta} + 2 H {\dot \delta} + \frac{ c_{\rm tot}(k)^2 k^2 }{a^2} \delta= 0,
1348: \label{pertsa}
1349: \ee
1350: where the effective total sound speed squared is
1351: \be
1352: c_{\rm tot}(k)^2 = c_{\rm s,in}^2 - \frac{4 \pi}{k^2} G e^\alpha \rho
1353: \left[ 1
1354:   + \frac{2 \mpl^2 \alpha^\prime(\phi)^2 }{1 +
1355: \frac{m_{\rm eff}^2}{ {\bf k}^2 }
1356: } \right].
1357: \label{ctot}
1358: \ee
1359: In the range of lengthscales (\ref{range3}) this squared sound speed is a
1360: constant, independent of $k$, as for a normal, hydrodynamic sound
1361: speed.  Outside of this range of scales, we have $c_{\rm tot}^2 \propto
1362: 1/k$ at both large and small $k$ (if the intrinsic sound speed can be neglected), describing a conventional
1363: gravitational interaction.
1364: 
1365: 
1366: 
1367: We reiterate that the existence of the range of scales (\ref{range3})
1368: in which Newton's constant scales linearly with $k^2$ depends on the
1369: assumption of strong coupling, $|\alpha'| \mpl \ll 1$.  If, instead,
1370: $|\alpha'| \mpl \alt 1$, the dependence of $G$ on $k$ is very close to
1371: that of standard gravity, and the instability reduces to
1372: the normal Jeans instability of a fluid, characterized in a
1373: cosmological context by power law growth.
1374: 
1375: 
1376: %\subsection{Kinetic theory description}
1377: \subsection{Domain of validity of fluid description of dark matter}
1378: 
1379: 
1380: Up till now we have described cold dark matter as a pressureless
1381: fluid.  However, at a more fundamental level, one should use a kinetic
1382: theory description based on the collisionless Boltzmann equation.  In
1383: the conventional $\Lambda$CDM framework, the fluid approximation
1384: breaks down at small scales, below the free-streaming lengthscale,
1385: and also in the nonlinear regime after violent relaxation has taken
1386: place in CDM halos \cite{Gunn}.  We now discuss how, in the models discussed here,
1387: the conventional picture for the fluid domain of validity is slightly modified.
1388: 
1389: 
1390: Let us denote by $\sigma$ the rms velocity of the dark matter particles.
1391: Consider a perturbation characterized by a wavelength $\lambda$ and
1392: wavenumber $k = 2 \pi / \lambda$.
1393: The characteristic growth or oscillation time associated with this perturbation is
1394: $
1395: \tau(k) \sim \lambda / \sqrt{ | c_{\rm tot}(k) |^2},
1396: $
1397: where the total effective sound speed $c_{\rm tot}$ is given by Eq.\ (\ref{ctot}).
1398: The distance traveled by a dark matter particle in this time is $d(k)
1399: \sim \sigma \tau(k)$, and the ratio of this distance to the size of
1400: the perturbation is
1401: \be
1402: \frac{d(k) }{\lambda} \sim \frac{\sigma \tau(k)}{\lambda} \sim
1403: \frac{\sigma}{\sqrt{|c_{\rm tot}(k)^2|}}.
1404: \label{ratio0}
1405: \ee
1406: When this dimensionless ratio is of order unity or larger,
1407: perturbations do not have time to grow before they are washed out by
1408: free streaming of the particles, and the fluid approximation breaks
1409: down.  Using the formula (\ref{ctot}) with $c_{\rm s,in}$ set to
1410: zero\footnote{Since we expect $c_{\rm s,in} \sim \sigma$, setting
1411: $c_{\rm s,in}$ to zero is only consistent in the regime $\sigma^2
1412: \ll | c_s^2|$.  However, for $\sigma^2 \agt |c_s^2|$, retaining the intrinsic sound speed in Eq.\
1413: (\protect{\ref{ratio0}}) does not change the final result
1414: (\protect{\ref{lambdaFS}}) for the free streaming scale in order of
1415: magnitude.},
1416: , we obtain
1417: \be
1418: \frac{d(k)}{\lambda} \sim \frac{ \sigma k}{ \sqrt{4 \pi G e^\alpha \rho}}
1419: \left[ 1
1420:   + \frac{2 \mpl^2 \alpha^\prime(\phi)^2 }{1 +
1421: \frac{m_{\rm eff}^2}{ {\bf k}^2 }
1422: } \right]^{-1/2}.
1423: \label{ratio1}
1424: \ee
1425: 
1426: 
1427: Now in the conventional CDM framework, the factor in the square
1428: brackets is unity, so the ratio (\ref{ratio1}) is proportional
1429: to $k$ and becomes large as $k\to\infty$.  Hence the fluid approximation breaks down on small scales,
1430: below the critical free-streaming lengthscale $\lambda_{\rm FS} \sim
1431: \sigma / \sqrt{G \rho}$.
1432: 
1433: In the present context things work a little differently due to the
1434: scale dependence of Newton's constant.
1435: We can rewrite Eq.\ (\ref{ratio1}) in the approximate form
1436: \begin{equation}
1437: \label{ratio2}
1438: \frac{d(k)}{\lambda} \sim \left\{ \begin{array}{ll}
1439:     \frac{\sigma}{\sqrt{|c_s^2|}} \frac{\sqrt{2} k \mpl | \alpha'|}{m_{\rm eff}} & \mbox{ $k \ll
1440:         \frac{m_{\rm eff}}{\sqrt{2} \mpl |\alpha'|}$,}\\
1441:        \frac{\sigma}{\sqrt{|c_s^2|}} & \mbox{
1442:         $\frac{m_{\rm eff}}{\sqrt{2} \mpl |\alpha'|} \ll k \ll m_{\rm
1443:           eff},$}\\
1444:          \frac{\sigma}{\sqrt{|c_s^2|}} \frac{k}{m_{\rm eff}} &
1445:         \mbox{ $ m_{\rm eff} \ll k, $}\\
1446:         \end{array} \right.
1447: \end{equation}
1448: where $c_s^2 = - 8 \pi G e^\alpha \rho \mpl^2 \alpha^{\prime\,2} / m_{\rm eff}^2$
1449: is the constant value of the second term in the
1450: expression (\ref{ctot}) for $c_{\rm tot}^2$ in the range of scales
1451: (\ref{range3}), or equivalently the sound speed discussed in Sec.\ \ref{sec:hydro}.
1452: We see that the ratio $d(k)/\lambda$ is proportional to $k$ at large
1453: scales and at small scales, but that in the intermediate range of
1454: scales it is a constant, so that the effect of free streaming is
1455: equally important for all the modes in this range.
1456: If we define the free streaming lengthscale $\lambda_{\rm FS}(\sigma)$
1457: to be the smallest lengthscale for which free streaming is unimportant, $d(k)/\lambda \alt 1$, then we
1458: obtain
1459: \begin{equation}
1460: \label{lambdaFS}
1461: \lambda_{\rm FS}(\sigma) \sim \left\{ \begin{array}{ll} \frac{\sigma}{\sqrt{|c_s^2|}}
1462:     \frac{ \mpl | \alpha'|}{m_{\rm eff}} & \mbox{ $\sigma^2 \agt
1463:         |c_s^2|$,}\\
1464:        \frac{\sigma}{\sqrt{|c_s^2|}} \frac{1}{m_{\rm eff}}& \mbox{
1465:         $\sigma^2 \alt |c_s^2|.$}\\
1466:         \end{array} \right.
1467: \end{equation}
1468: This lengthscale jumps discontinuously at $\sigma^2 \sim |c_s^2|$.
1469: 
1470: 
1471: 
1472: There are thus two different regimes that occur:
1473: \begin{itemize}
1474: 
1475: \item When $\sigma^2
1476: \ll |c_s^2|$, free streaming is important only at
1477: scales small compared to $1/m_{\rm eff}$ (for which the adiabatic
1478: approximation is invalid anyway).
1479: The fluid approximation is valid throughout the range of
1480: lengthscales (\ref{range3}), and so the adiabatic instability is present.
1481: This conclusion is confirmed by a kinetic theory analysis (see
1482: Appendix \ref{kinetic}), which shows
1483: that linearized perturbations of any
1484: homogeneous, isotropic initial particle distribution function are always unstable on scales
1485: that are in the regime (\ref{range3}), as long as $\sigma^2 \ll |c_s^2|$.
1486: 
1487: 
1488: 
1489: \item When $\sigma^2 \agt |c_s^2|$, free streaming becomes important
1490: and the fluid approximation breaks down throughout the range of scales
1491: (\ref{range3}).
1492: %\footnote{Normally for a fluid one
1493: %expects the sound speed to be of order the velocity
1494: %dispersion $\sigma$ of the particles.  This constraint does not apply
1495: %here since $c_s^2$ is a parameter that essentially arises from the
1496: %(scalar-mediated) gravitational interactions, and not from the
1497: %microphysics of the fluid.  Thus, it is possible to have $\sigma^2 \ll
1498: %|c_s^2|$ or $\sigma^2 \gg |c_s^2|$.}
1499: One expects the free streaming
1500: (also called Landau damping) to kill the instability.
1501: This is confirmed by our kinetic theory analysis of Appendix \ref{kinetic}:
1502: we show that for a Maxwellian distribution, the finite velocity
1503: dispersion stabilizes the coupled fluid whenever
1504: $\sigma^2 \ge | c_s^2|$, in agreement with the analysis of Ref.\ \cite{Afshordi:2005ym}.
1505: 
1506: \end{itemize}
1507: 
1508: 
1509: 
1510: Consider now the evolution of cosmological perturbation modes.  When
1511: will our analysis of instability apply?  First,
1512: the condition $\sigma^2 \le |c_s^2|$ is not
1513: very restrictive, since the CDM cools rapidly with the Universe's expansion, with
1514: temperature scaling as $(1+z)^2$ once the
1515: particles become non-relativistic.  For example, for CDM particles of
1516: mass $\sim $ GeV with weak scale cross sections, the CDM temperature
1517: is $\sim 10^{-4}$ K at decoupling \cite{Gunn}.
1518: However, after perturbations go nonlinear and
1519: violent relaxation takes place in CDM halos, the effective
1520: coarse-grained velocity dispersion becomes much larger.
1521: Hence, our analysis does not apply to modes that are in the nonlinear
1522: regime.  Our analysis will apply in the early Universe, before any
1523: modes have gone nonlinear.  It will also apply to large scale modes, even
1524: after smaller scale modes have gone nonlinear, since such large scale
1525: modes should still be well described by linear theory (see for example
1526: the qualitative arguments in chapter 28 of Peebles \cite{Peebles}).
1527: Our investigation of specific models later in the paper will focus on
1528: these large scale, linear regime modes.
1529: 
1530: We note that earlier investigations of the instability focused instead
1531: on small scale modes, below the free streaming scale $\lambda_{\rm
1532:   FS}$ \cite{Afshordi:2005ym,Kaplinghat:2006jk}.
1533: From the formula (\ref{lambdaFS}) we see that, for the models
1534: discussed here, either the adiabatic approximation is not valid on these
1535: small scales since $\lambda_{\rm FS} \alt 1/m_{\rm eff}$, or
1536: $\sigma^2 \agt |c_s^2|$ and the instability is killed by free streaming.
1537: 
1538: 
1539: 
1540: 
1541: 
1542: 
1543: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1544: \section{Examples of theories with adiabatic instability}
1545: \label{examples}
1546: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1547: In this section we discuss some specific classes of theories.
1548: 
1549: 
1550: \subsection{Exponential potential and constant coupling}
1551: \label{sec:exponential}
1552: 
1553: 
1554: We first consider theories with exponential potentials of the form
1555: \be
1556: V = V_0 e^{-\lambda \phi/\mpl} \ ,
1557: \label{expV}
1558: \ee
1559: with $\lambda>0$ and with linear coupling functions
1560: \be
1561: \alpha(\phi) = - \beta C \frac{\phi}{\mpl} \ ,
1562: \label{constantcoupling}
1563: \ee
1564: where $\beta = \sqrt{2/3}$ and $C$ is a constant.\footnote{The
1565:   notation in Eq.\ (\ref{constantcoupling}) is chosen such that $f(R)$
1566:   gravity theories correspond to $C=1/2$.}
1567: These theories have been previously studied in Ref.\ \cite{Amendola:1999er}.
1568: The effective potential is, from Eq.\ (\ref{phieqn}),
1569: \be
1570: V_{\rm eff}(\phi,\rho) = V_0 e^{-\lambda \phi/\mpl} + e^{-\beta C \phi/\mpl} \rho \ ,
1571: \ee
1572: and solving for the local minimum of this potential yields
1573: the relation between $\phi$ and $\rho$ in the adiabatic regime:
1574: \be
1575: e^{(\lambda -\beta C) \phi_{\rm m}(\rho)/\mpl} = \frac{ \lambda V_0 }{-\beta C
1576:   \rho} \ .
1577: \label{phimans}
1578: \ee
1579: Note that $C$ must be negative in order for
1580: the effective potential to have a local minimum and for an adiabatic
1581: regime to exist.  We will restrict attention to this case, and we
1582: define the dimensionless
1583: positive parameter $\gamma = - \lambda / \beta C$.
1584: The corresponding effective mass parameter is
1585: \be
1586: m_{\rm eff}^2 = \lambda^2 \mpl^{-2} V_0 \frac{1+\gamma}{\gamma} \left( \frac{
1587:     \rho}{\gamma V_0} \right)^{\frac{\gamma}{\gamma+1}} \ .
1588: \label{mass1}
1589: \ee
1590: 
1591: 
1592: Next, we compute the sound speed squared.  Using Eq.\
1593: (\ref{soundspeed}) we obtain
1594: \be
1595: c_s^2  =  - \frac{1}{1+\gamma} \ ,
1596: \label{cs21}
1597: \ee
1598: so this
1599: model is always unstable in the adiabatic regime.
1600: From Eqs.\ (\ref{expV}) and (\ref{phimans}) we also obtain
1601: \be
1602: \frac{\partial \ln V}{\partial \ln \rho} = \frac{\gamma}{1 + \gamma} \ .
1603: \label{logd}
1604: \ee
1605: We now insert the effective mass (\ref{mass1}), the sound speed
1606: squared (\ref{cs21}) and the logarithmic derivative
1607: (\ref{logd})
1608: into Eq.\ (\ref{einsteincheck1a})
1609: and into the second half of
1610: Eq.\ (\ref{range}).  This yields the range of spatial scales ${\cal
1611:   L}$ over which
1612: the instability operates for a given density $\rho$ to be
1613: ${\cal L}_{\rm min}(\rho) \ll {\cal L} \ll {\cal L}_{\rm max}(\rho)$,
1614: where
1615: \be
1616: {\cal L}_{\rm min}(\rho)^2 = \frac{\gamma^2 }{\lambda^2 (1 + \gamma)^2 }
1617: \frac{\mpl^2}{V_0}
1618: \left( \frac{ \gamma
1619:     V_0}{ \rho} \right)^{\frac{\gamma}{\gamma+1}}
1620: \label{Lmin2}
1621: \ee
1622: and ${\cal L}_{\rm max}(\rho)^2$ is a constant times this:
1623: \be
1624: {\cal L}_{\rm max}(\rho)^2 = \beta^2 C^2 {\cal L}_{\rm min}(\rho)^2 \ .
1625: \label{Lmax2}
1626: \ee
1627: Thus, there is a nonempty unstable regime only when $\beta |C| \gg 1$,
1628: ie with the scalar coupling is strong compared to the gravitational
1629: coupling, in agreement with the discussion in Sec.\ \ref{sec:scales}.
1630: 
1631: To see the effect of the instability more explicitly, we consider cosmological perturbations.
1632: The Einstein-frame FRW equation in the adiabatic limit is
1633: \be
1634: 3 \mpl^2 H^2 = V + e^\alpha \rho \ ,
1635: \label{cosmo1}
1636: \ee
1637: where $\rho \propto 1/a^3$.  This yields $a(t) \propto t^{2/(3 + 3 w_{\rm eff})}$, where
1638: the effective equation of state parameter is
1639: \be
1640: w_{\rm eff} = - \frac{1}{1 + \gamma} \ .
1641: \label{weff1}
1642: \ee
1643: In the strong coupling limit $|C| \to \infty$ that we specialize to here, $w_{\rm eff} \to -1$.
1644: Thus the adiabatic regime of this model with large $|C|$ is incompatible with
1645: observations in the matter dominated era, where we know $w_{\rm eff} \approx
1646: 0$ except for at small redshifts.
1647: Nevertheless, the model is still
1648: useful as an illustration of the instability.
1649: 
1650: 
1651: We find from Eqs.\ (\ref{Lmin2}), (\ref{Lmax2}) and (\ref{cosmo1})
1652: that the range of unstable scales is given by
1653: \be
1654: \frac{1}{\beta^2 C^2} \ll \frac{H^2 a^2}{k^2}  \ll \frac{1}{3(1+\gamma)} \ ,
1655: \ee
1656: where $k$ is comoving wavenumber.  This range of scales always lies
1657: just inside the horizon.  A given mode $k$ will evolve through this
1658: unstable region before it exits the horizon.
1659: 
1660: 
1661: Next we use the approximate form (\ref{Gcc1}) of Newton's constant in the
1662: perturbation evolution equation (\ref{perts}), and transform from $t$
1663: derivatives to $a$ derivatives.
1664: This gives
1665: \be
1666: \frac{d^2 \delta}{d a^2} + \frac{3}{a} \left( 1 - \frac{1}{2} \frac{d
1667:     \ln \rho_{\rm eff}}{d \ln \rho} \right) \frac{d \delta}{da} -
1668: \left( \frac{ (\alpha^\prime)^2 k^2 e^\alpha \rho}{m_{\rm eff}^2 H^2
1669:     a^4} \right) \delta =0 \ .
1670: \ee
1671: Specializing this equation to the exponential model using Eqs.\
1672: (\ref{constantcoupling}), (\ref{phimans}), (\ref{mass1}) and
1673: (\ref{cosmo1}) and taking the strong coupling limit
1674: $|C| \to \infty$ gives
1675: \be
1676: \frac{d^2 \delta}{d a^2} + \frac{3}{a} \frac{d \delta}{da} -
1677: \frac{k^2}{ H^2 a^4} \delta =0 \ .
1678: \ee
1679: In the strong coupling limit $H$ is approximately a constant, $H
1680: \approx H_0$, and the growing mode solution is
1681: \be
1682: \delta(a) \propto \frac{1}{a} K_1\left(\frac{k}{H_0 a}\right) \approx
1683: \sqrt{ \frac{\pi H_0}{2 k a}} \exp \left(-\frac{k}{H_0 a} \right) \ ,
1684: \ee
1685: where $K_1$ is the modified Bessel function.  The mode grows by a
1686: factor $\sim e$ when the scale factor changes from $a$ to $a + \Delta
1687: a$, where $\Delta a/a \sim a H_0 / k \ll 1$ for subhorizon
1688: modes.
1689: 
1690: A more detailed analysis of the cosmology of this model is given in
1691: Ref.\ \cite{paperIII}, but in the non-adiabatic regime $|C| \sim 1$ rather than the
1692: strong coupling regime $|C| \gg 1$ considered here.
1693: 
1694: 
1695: \subsection{Two component dark matter models}
1696: \label{sec:twocomponent}
1697: 
1698: We next consider models in which there are two dark matter sectors, a
1699: density $\rho_c$ which is not coupled to the scalar field, and a
1700: density $\rho_{co}$ which is coupled with coupling function (\ref{constantcoupling})
1701: and exponential potential (\ref{expV}).
1702: Both of these components are treated as pressureless fluids.
1703: The FRW equation for this model in the adiabatic limit is
1704: [cf.\ Eq.\ (\ref{cosmo1}) above]
1705: \be
1706: 3 \mpl^2 H^2 = V + e^\alpha \rho_{co} + \rho_c \ .
1707: \label{cosmo2}
1708: \ee
1709: Similar two component models have been considered by Farrar and Peebles
1710: \cite{Farrar:2003uw}.  This model is also similar to the mass varying
1711: neutrino model model \cite{Afshordi:2005ym,Bjaelde:2007ki}
1712: where the neutrinos play the role of the coupled component; see Sec.\
1713: \ref{sec:MaVaN} below.
1714: The first two terms on the right hand
1715: side of Eq.\ (\ref{cosmo2}) act like a fluid with equation of state
1716: parameter given by (\ref{weff1}), and in the strong coupling limit
1717: $|C| \gg 1$ this
1718: fluid acts like a cosmological constant.  Thus, the background
1719: cosmology can be made close to $\Lambda$CDM by taking $|C|$ to be
1720: large.
1721: 
1722: The fraction of dark matter which is coupled must be small in the
1723: limit of large coupling, $|C| \gg 1$.
1724: Denoting $\Omega_V = V / (3 \mpl^2 H^2)$,
1725: $\Omega_{co} = e^\alpha \rho_{co} / (3 \mpl^2 H^2)$ and
1726: $\Omega_c = \rho_c / (3 \mpl^2 H^2)$, we have $1 = \Omega_V +
1727: \Omega_{co} + \Omega_c$.  Also from Eq.\ (\ref{phimans}) it follows
1728: that, if the asymptotic adiabatic regime  has been reached, $\Omega_{co} = \gamma \Omega_V$, and we we obtain
1729: \be
1730: \Omega_{co} = \frac{\gamma}{1 + \gamma} (1 - \Omega_c) \ .
1731: \label{Omegaco}
1732: \ee
1733: Since $\Omega_c \sim 0.3$ today, and $\gamma \ll 1$ in the strong
1734: coupling limit we are considering, we must have $\Omega_{co} \ll 1$
1735: today.
1736: 
1737: 
1738: 
1739: The maximum and minimum lengthscales for the instability are still
1740: given by Eqs.\ (\ref{Lmin2}) and (\ref{Lmax2}), but with $\rho$ replaced by
1741: $\rho_{co}$.  Since $\rho_{\rm co}$ is approximately a constant in the
1742: strong coupling limit, these lengthscales are also constants.
1743: If the parameters of the model are chosen so that $\Omega_c \sim 1$
1744: today, then
1745: \be
1746: {\cal L}_{\rm max} \sim H_0^{-1}, \ \ \ \ {\cal L}_{\rm min} \sim
1747: \frac{H_0^{-1}}{\beta |C|} \ .
1748: \label{scales0}
1749: \ee
1750: The evolution equations for the fractional density perturbations
1751: $\delta_{\rm j} = \delta \rho_{\rm j} / \rho_{\rm j}$
1752: in the adiabatic limit on subhorizon scales
1753: are given by
1754: \be
1755: {\ddot \delta}_{\rm j} + 2 H {\dot \delta}_{\rm j} - 4\pi \sum_k
1756: G_{{\rm j}{\rm k}}
1757: \rho_{\rm k} e^{\alpha_{\rm k}} \delta_{\rm k}=0 \ ,
1758: \ee
1759: where the effective Newton's constants are given by Eq.\
1760: (\ref{Gformula}) with $\alpha_b$ set to zero.  Writing this out
1761: explicitly we obtain
1762: \bea
1763: {\ddot \delta}_c + 2 H {\dot \delta}_c &=& \frac{1}{2 \mpl^2} \rho_c \delta_c +
1764: \frac{1}{2 \mpl^2} e^\alpha \rho_{co} \delta_{co}, \\
1765: {\ddot \delta}_{co} + 2 H {\dot \delta}_{co} &=& \frac{1}{2 \mpl^2}
1766: \rho_c \delta_c \nonumber \\
1767: &&+
1768: \frac{1}{2 \mpl^2} \left[ 1 + \frac{2 \beta^2 C^2}{1 + \frac{m_{\rm
1769:         eff}^2 a^2}{k^2}} \right] e^\alpha \rho_{co}
1770: \delta_{co} \ . \nonumber \\
1771: \label{uns}
1772: \eea
1773: The condition for the instability to operate is that
1774: the timescale associated with the second
1775: term on the right hand side of Eq.\ (\ref{uns}) be short compared with
1776: $H^{-1}$, or
1777: \be
1778: \frac{\beta^2 C^2 k^2}{m_{\rm eff}^2 a^2} \rho_{co} e^\alpha \gg H^2
1779: \mpl^2 \ .
1780: \label{condt4}
1781: \ee
1782: Now the effective mass for this model is given by
1783: $m_{\rm eff}^2 = \beta^2 C^2 \mpl^{-2} (V + e^\alpha \rho_{co}) = 3
1784: \beta^2 C^2 \mpl^{-2} H^2 (\Omega_V + \Omega_{co})$.  Substituting this
1785: into Eq.\ (\ref{condt4}) and using Eq.\ (\ref{Omegaco}) gives the criterion
1786: $k/(a H) \gg 1$.  Therefore
1787: the instability should operate whenever modes are inside the horizon
1788: and in the range of scales (\ref{scales0}).
1789: 
1790: \begin{figure}[t]
1791: \begin{center}
1792: \includegraphics[width=3.5in]{fig1.ps}
1793: \caption{[Bottom] The two component coupled dark energy (CDE)
1794:   model, with exponential potential and coupling, with $\lambda =2$ and coupling $C=-20$ with $H_0=70 \, {\rm km} \, {\rm s}^{-1}\, {\rm Mpc}^{-1}$, $\Omega_{b}=0.05$,
1795:   $\Omega_{c}=0.2$, $\Omega_{co}=0.05$, and $\Omega_{V}=0.70$. At late times the scalar field finds the adiabatic minimum with asymptotic equation of state, and sound speed $= -1/(1+\gamma) = -0.89$, able to reproduce a viable background evolution consistent   with supernovae, CMB angular diameter distance and BBN expansion
1796:   history constraints. The figure shows the evolution of the effective equation of state, $w_{eff} = P_{tot}/\rho_{tot}=(2/3) (d\ln t / d\ln a) -1,$ (black full line), the adiabatic speed of sound, $c_{a}^{2}=\dot{P}/\dot{\rho}$ for all components (blue long dashed line) and for the coupled components only (green dot long dashed line), and effective speed of sound for $c_s^2=\delta P/\delta\rho$ at $k=0.01/Mpc$ for all components (red dot-dashed line)  and for the coupled components alone (magenta dotted line). The effective equation of state for a comparable $\Lambda$CDM model with $\Omega_{c}=0.25$, $\Omega_b=0.05$ and $\Omega_{\Lambda}=0.7$ is also shown (black dashed line). [Top] The growth of the fractional over-density $\delta=\delta\rho/\rho$ for $k=0.01/Mpc$ for the coupled CDM component, $\delta_{co}$, (red long dashed line) and uncoupled component, $\delta_{c}$, (black full line) in comparison to the growth for the $\Lambda$CDM model (black dashed line). At late times the adiabatic behavior triggers a dramatic increase in the rate of growth of both uncoupled and coupled components, leading to structure predictions inconsistent with observations.\label{fig1}}
1797: \end{center}
1798: \end{figure}
1799: 
1800: These expectations are confirmed by numerical integrations.
1801: In figure \ref{fig1}  we present a numerical analysis
1802: of such a two component model. We consider an exponential potential with $\lambda =2$ and strong
1803: coupling with $C=-20$, and typical cosmological parameters are assumed,
1804: $H_0=70 \, {\rm km} \, {\rm s}^{-1}\,{\rm Mpc}^{-1}$,
1805: $\Omega_{b}=0.05$, $\Omega_{c}=0.2$, $\Omega_{co}=0.05$, and
1806: $\Omega_{V}=0.7$. We fix initial conditions of $\phi/\mpl=10^{-3}$ and
1807: $\dot\phi=0$ at $a=10^{-10}$, although the dynamical attractor renders
1808: the final evolution largely insensitive to these choices, we have checked that $\phi/\mpl(a=10^{-10})=10^{-30}-1$ give the same evolution. We neglect the effect of the coupled CDM component peculiar velocity in the initial conditions (it is many orders of magnitude smaller than  the density perturbation) and assume
1809: that the coupled  and uncoupled CDM components have the same initial
1810: fractional density perturbations $\delta_c=\delta_{co}$, fixed by the
1811: usual adiabatic initial conditions. As shown in figure \ref{fig1}, the
1812: background evolution is entirely consistent with a $\Lambda$CDM like
1813: scenario, with $w_{\rm eff}=-0.69$ today
1814: and an asymptotic equation of state at
1815: late times, given by (\ref{weff1}), $w_{\rm eff} = -1/(1+\gamma)=-0.89$. The large
1816: coupling drives the evolution to an adiabatic regime at late times,
1817: with an adiabatic sound speed $c_{a}^{2}\rightarrow -1/(1+\gamma)$ as
1818: in (\ref{cs21}). This drives a rapid growth in over-densities once in the adiabatic regime so that although consistent with structure observations at early times, they are inconsistent once the accelerative regime has begun.
1819: 
1820: 
1821: 
1822: 
1823: 
1824: 
1825: In summary, these models provide a class of theories for which the
1826: background cosmology is compatible with observations, but which are
1827: ruled out by the adiabatic instability of the perturbations.
1828: 
1829: 
1830: \subsection{Chameleon models}
1831: 
1832: Next we study the so-called chameleon models \cite{Khoury:2003aq,Khoury:2003rn}
1833: defined by the potential
1834: \be
1835: V(\phi) = \lambda M^4 \left( \frac{M }{\phi} \right)^n \ ,
1836: \ee
1837: where $M$ is a mass scale and $n>0$ and $\lambda$ are dimensionless constants,
1838: together with the coupling function~(\ref{constantcoupling}).
1839: In these models it has been previously shown that the adiabatic regime
1840: is achieved in static solutions describing macroscopic bodies like
1841: the Earth, and that cosmological solutions in the adiabatic regime
1842: provide good models of dark energy \cite{Brax:2004qh,Brax:2004px,Brax:2005ew}.
1843: We now study under what conditions these models are
1844: unstable.
1845: 
1846: 
1847: The effective potential is, from Eq.\ (\ref{phieqn}),
1848: \be
1849: V_{\rm eff}(\phi,\rho) = \lambda M^4 \left( \frac{M }{\phi} \right)^n
1850:  + e^{-\beta C \phi/\mpl} \rho \ ,
1851: \ee
1852: and solving for the local minimum of this potential yields
1853: the relation between $\phi$ and $\rho$ in the adiabatic regime:
1854: \be
1855: x^{n+1} e^x = \frac{\rho_{\rm crit}}{\rho} \ ,
1856: \label{xeqn}
1857: \ee
1858: Here $x = -\beta C \phi_c(\rho)/\mpl$ is dimensionless and the critical density is
1859: \be
1860: \rho_{\rm crit} = n \lambda M^4 \left(\frac{-\beta C M}{\mpl}\right)^n \ .
1861: \ee
1862: As before the existence of a local minimum in the effective potential requires
1863: $C$ to be negative.
1864: We shall restrict attention to the regime
1865: \be
1866: \rho \gg \rho_{\rm crit}
1867: \label{highdensity}
1868: \ee
1869: since for models of dark energy $\rho_{\rm crit}$ will be of order the
1870: present day cosmological density.  In this
1871: regime the solution to Eq.\ (\ref{xeqn}) is approximately
1872: \be
1873: x \approx \left( \frac{\rho_{\rm crit} }{\rho} \right)^{\frac{1}{n+1}}.
1874: \ee
1875: The corresponding effective mass parameter is
1876: \be
1877: m_{\rm eff}^2 = (x + n + 1) n \lambda M^2 \left( \frac{-\beta C M}{x \mpl} \right)^{n+2},
1878: \label{mass2}
1879: \ee
1880: which in the regime (\ref{highdensity}) simplifies to
1881: \be
1882: m_{\rm eff}^2 = (n+1) (-\beta C)^2 \mpl^{-2} \rho_{\rm crit} \left( \frac{\rho}{\rho_{\rm crit}}
1883: \right)^{\frac{n+2}{n+1}}.
1884: \ee
1885: 
1886: Next, we compute the sound speed squared.  Using Eq.\
1887: (\ref{soundspeed}) we obtain
1888: \be
1889: \frac{1}{c_s^2}  = -1 -  \frac{n+1}{(-\beta C)} \frac{\mpl}{\phi},
1890: \label{cs213}
1891: \ee
1892: and since $C$ is negative, we see that this
1893: model is always unstable in the adiabatic regime.
1894: Inserting the effective mass (\ref{mass2}) and the sound speed squared (\ref{cs213})
1895: into Eq.\ (\ref{range}) we obtain the range of spatial scales ${\cal
1896:   L}$ over which
1897: the instability operates for a given density $\rho$:
1898: \be
1899: {\cal L}_{\rm min}(\rho)^2 \ll {\cal L}^2 \ll
1900: (\beta C)^2  {\cal L}_{\rm min}(\rho)^2,
1901: \label{range2}
1902: \ee
1903: where
1904: \be
1905: {\cal L}_{\rm min}^2(\rho) = \frac{\mpl^2}{ (n+1) (\beta C)^2 \rho_{\rm crit}} \left(
1906:   \frac{ \rho_{\rm crit} }{\rho} \right)^{\frac{n+2}{n+1}}.
1907: \ee
1908: We see that the range of unstable lengthscales is non-empty only if
1909: \be
1910: \beta |C| \gg 1,
1911: \label{ad34}
1912: \ee
1913: which as before is equivalent to the strong coupling condition (\ref{strongcoupling}).
1914: Note that the first of the two inequalities in
1915: Eq.\ (\ref{range2}) is equivalent to the ``thin-shell condition'' of Ref.\
1916: \cite{Khoury:2003rn}
1917: when the background value of the scalar field can be neglected.
1918: 
1919: 
1920: As for the exponential models of Sec.\ \ref{sec:exponential}, the
1921: effective equation of state $w_{\rm eff}$ is close to $-1$
1922: in the adiabatic regime for large coupling, assuming $\rho \gg \rho_{\rm crit}$.
1923: Therefore these models do not give an acceptable background cosmology for the matter dominated era
1924: in their adiabatic regime.  However, one can construct two component
1925: models analogous to those in Sec.\ \ref{sec:twocomponent} using the
1926: chameleon potential.  Those models give an acceptable background
1927: cosmology, but are then ruled out by the adiabatic instability.
1928: 
1929: \subsection{Mass Varying Neutrino (``MaVaN'') models }
1930: \label{sec:MaVaN}
1931: 
1932: The impact of adiabatic instabilities has been discussed extensively in the context of MaVaN models, in which the light mass of the neutrino and the recent accelerative era are twinned together through a scalar field coupling \cite{Fardon:2003eh,Kaplan:2004dq,Fardon:2005wc}.  The adiabatic instability was shown to be a concern in these models with the implication of forming compact localized regions of neutrinos after undergoing dramatic adiabatic collapse \cite{Afshordi:2005ym}.
1933: 
1934: 
1935: The action for the MaVaN models is of the form (\ref{eq:action10}) with $p=0$.
1936: Therefore the analysis in the earlier sections of this paper, which
1937: assumed $p=2$, does not directly apply to these models.  However it is
1938: straightforward to generalize our analysis to cover this case.
1939: Consistent with the general discussion in sections
1940: \ref{adiabatic}-\ref{instability}, the instability is present in these
1941: theories if the coupling is strong. Here we discuss examples of
1942: models which evade or are subject to the instability on cosmological
1943: scales.
1944: 
1945: Recently,  Ref.\ \cite{Bjaelde:2007ki} discussed a MaVaN scenario with a logarithmic potential and scalar dependent mass,
1946: \bea
1947: V(\phi) &= &V_0\log(1+\xi\phi) \
1948: \\
1949: \label{mavan1}
1950: m_\nu(\phi) &=& m_{\nu 0}\left(\frac{\phi_*}{\phi}\right) \ ,
1951: \eea
1952: where $\xi$ is a constant and $m_{\nu 0}$ and $\phi_*$ are the current values of the neutrino mass and the scalar field respectively.
1953: It was found in \cite{Bjaelde:2007ki} that this model can exhibit an instability in growth for an otherwise cosmologically viable background solution. We find, however, that this model can also allow stable solutions for identical fractional densities today as those studied in \cite{Bjaelde:2007ki}, for a wide range of parameter values. If the coupling, $m_\nu'/m_\nu$, is not large compared to $\mpl^{-1}$ at late times, the evolution never enters the adiabatic regime on cosmic scales. This translates in this model to $\phi_*$ not being significantly less than $\mpl$.
1954: 
1955: 
1956: 
1957: The MaVaN model evolves according to a coupled Klein Gordon equation,
1958: \bea
1959: \ddot{\phi} + 2H \dot\phi +a^2V'(\phi) &=& -a^2 (\rho'_\nu(\phi)-3P'_\nu(\phi)) \ .\ \ \ \ \ \ \label{KGnu}
1960: \eea
1961: Assuming no chemical potential, and $g$ spin states per neutrino species with momentum $p$ and mass $m$ ,the neutrino density and pressure are given by
1962: \bea
1963: a^4 \rho_\nu& = & {g (k_{B}T_{\nu}^{0})^{4} \over 2\pi^{2}} \int_{0}^{\infty} dq q^{2} (q^{2}+a^{2}\bar{m}_\nu^{2})^{1\over 2} f(q) \\
1964: a^4 P_\nu & = & {g (k_{B}T_{\nu}^{0})^{4} \over 2\pi^{2}} \int_{0}^{\infty} dq q^{2} {q^{2}\over 3(q^{2}+a^{2}\bar{m}_\nu^{2})^{1\over 2}} f(q) \ \  \ \ \
1965: \\
1966: f(q)&\approx& \left[\exp\left({q}\right)+1\right]^{-1} \ ,
1967: \eea
1968:  with $q \equiv ap/ k_{B}T_{\nu}^{0}$ and $\bar{m}_\nu\equiv m_\nu c^{2}/k_{B}T_{\nu}^{0}$.
1969: 
1970: In the relativistic regime, with $\bar{m}\ll 1$, the potential is negligible and the driving term, on the right hand side of  (\ref{KGnu}), can be calculated by doing a Taylor expansion to first order in $\bar{m}_\nu$,
1971: \bea
1972: a^4 (\rho_\nu-3P_\nu)
1973: & \approx &{g (k_{B}T_{\nu}^{0})^{4} \over 2\pi^{2}} \int_{0}^{\infty} dq q f(q) a^{2}\bar{m}_\nu^{2} \ \
1974: \\
1975: a^{2}(\rho_\nu-3P_\nu)
1976: &\approx&  \frac{10}{7\pi^2}\bar{m}_\nu(\phi)^{2}\rho_0 \ ,
1977: \eea
1978: so that
1979: \bea
1980: a^{2}(\rho'_\nu-3P'_\nu) &\approx& -\frac{2}{\phi}(\rho_\nu-3P_\nu) a^{2} \ ,
1981: \label{rhopres}
1982: \eea
1983: where $\rho_0\equiv 7\pi^2 g(k_BT_\nu^0)^4 /240$ would be the relativistic neutrino energy density per neutrino species today with temperature $T_{\nu}^{0}$.
1984: 
1985: 
1986: Putting (\ref{rhopres}) into (\ref{KGnu}) and neglecting the potential
1987: we find a power law attractor $\phi\propto \tau^{x}$ with $x=0.5$.
1988: The normalization of $\phi$ is wholly specified in the attractor by (\ref{KGnu}). Writing $\phi = \phi_i(\tau/\tau_i)^{0.5}$, and $a\propto\tau^{p}$, with $p=2/(1+3w_{\rm eff})$, we find
1989: \bea
1990: \phi &=& \left[\frac{80\rho_0m_{\nu 0}^{2}\phi_*^{2}}{7\pi^2(4p-1)}\right]^{0.25}\tau^{0.5} \ .
1991: \label{phinorm}
1992: \eea
1993: 
1994: When the neutrino is non-relativistic, if we again neglect the potential,
1995: \bea
1996: (\rho_\nu-3P_\nu) &=&  \frac{3H_{0}^{2}\mpl^2 \Omega_\nu}{a^3} \left(\frac{\phi_*}{\phi}\right)
1997: \\
1998: a^2(\rho'_\nu-3P'_\nu) &=&-\frac{3H_{0}^{2}\mpl^2 \Omega_\nu\phi_*}{a} \left(\frac{1}{\phi^{2}}\right) \ .
1999: \eea
2000: The Klein-Gordon equation has a solution $\phi \propto \tau^x$ with $x=(2-p)/3$, tending towards a cessation of growth in $\phi$ in the matter dominated era.
2001: \begin{figure}[t]
2002: \begin{center}
2003: \includegraphics[width=3.6in]{fig2.ps}
2004: \caption{[Top panels] Evolution of $m_\nu(\phi)$ and neutrino temperature (left), and associated equation of state and adiabatic sound speed (right) for MaVaN model described in the text with $m_{\nu 0}=0.312eV$ and $\phi_*=1.8\mpl$. [Bottom left panel] Scalar field evolution in the MaVaN scenario, for 3 values of $\phi_*\sim 10^{-3}\mpl$, $0.3\mpl$ and $1.8\mpl$ (full lines) showing the $\tau^{0.5}$ attractor while the neutrino is relativistic allowing late time evolution to be independent of initial conditions. The vacuum expectation value of the scalar field, if the field becomes adiabatic, (dashed lines) is also shown. For $\phi_* \gtrsim 10^{-2}\mpl$ the scalar field does not enter an adiabatic era on cosmological scales before now, and growth of perturbations remains well-behaved. For smaller $\phi_*$, for example $\phi_*\sim 10^{-3}\mpl$ shown, the evolution is adiabatic at late times, similar to that discussed in \cite{Bjaelde:2007ki}. [Bottom right panel] The resulting matter power spectrum from the coupled dark energy (CDE) model with $\phi_* = 1.8\mpl$ and $\Omega_\nu=0.02$ (full line) is very similar to that for $\Lambda$CDM (dashed line) with the same baryon fraction and $H_0$, $\Omega_m=0.3$, $\Omega_\nu=0.02$ when normalized at large scales.}\label{fig2}
2005: \end{center}
2006: 
2007: \end{figure}
2008: 
2009: 
2010: 
2011: If the neutrinos are non-relativistic (but, of course, assuming that
2012: they decoupled when they were relativistic)
2013: \bea
2014: (\rho_\nu-3P_\nu) &\approx& N_\nu \frac{n_0}{a^3} m_{\nu 0} \left(\frac{\phi_*}{\phi}\right)
2015: \\
2016: & \approx& \frac{180\zeta(3)}{7\pi^4}\frac{\rho_0}{a^{3}} N_\nu \bar{m}_{\nu0}\left(\frac{\phi_*}{\phi}\right) \ .
2017: \eea
2018: At late times, the complete effective potential is relevant, for which there exists a minimum at positive $\phi$ given by
2019: \bea
2020: \frac{\phi_{VEV}}{\phi_*}  = \frac{1}{a^{3}}\frac{\Omega_{\nu}}{\Omega_{pot}}\log(1+\xi\phi_*) \ .
2021: \eea
2022: 
2023: 
2024: We modified CAMB \cite{Lewis:2002ah} to investigate the evolution numerically.
2025: % corrected a misspelling in the following sentence, "tos a fiducial"
2026: % -> "to a fiducial"
2027: In figure \ref{fig2} we show the scalar field and neutrino mass evolution for $\xi = 10^{20}\mpl^{-1}$, $\phi_*=1.8\mpl$, and $m_{\nu 0}=0.312eV$ (giving $\Omega_\nu=0.02$), $\Omega_b=0.05$, $\Omega_c=0.23$, $H_0=70 $kms$^{-1}$Mpc$^{-1}$.  In the bottom left hand figure, the numerical evolution of scalar field is shown for $\phi_*=1.8\mpl$ along with two smaller values $\phi_*\sim 10^{-3}\mpl$ and $0.3\mpl$ for which the minimum of the effective potential is steeper. The background evolution obeys the $\phi\sim\tau^{0.5}$ attractor and normalization in (\ref{phinorm}), rendering it largely independent of the initial conditions. When the neutrino becomes non-relativistic the evolution slows and for small values of $\phi_*$ starts to track the VEV, as discussed in \cite{Bjaelde:2007ki}. However for larger values of $\phi_*$, the VEV is not reached until later times, and in addition, the effective potential minimum is shallow enough that the scalar field does not get fixed at the VEV automatically, enabling well-behaved growth. The bottom right hand figure shows the resultant matter power spectrum, which is very similar to a fiducial $\Lambda$CDM model with the same Hubble factor, baryon and neutrino density today when normalized to the same amplitude at large scales.
2028: 
2029: 
2030: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2031: \section{Conclusions}
2032: \label{conclusions}
2033: 
2034: If dark energy and dark matter are to fit into a coherent fundamental
2035: physics framework then there are likely couplings between them.
2036: Such couplings may have far reaching macroscopic implications on scales
2037: ranging from the solar system up to cosmological horizon. As we have
2038: discussed, these effects may lead to observationally distinctive
2039: characteristics, which may allow us to tease out the nature of the
2040: dark sector. However they may also give rise to catastrophic
2041: instabilities with which we may constrain the class of physically
2042: viable dark energy models.
2043: 
2044: In this paper we have  considered such theories in which there exists a nontrivial coupling between the
2045: dark matter sector and the sector responsible for the acceleration of the universe.
2046: 
2047: We have comprehensively analyzed an instability -- characterized by a
2048: negative sound speed squared of
2049: an effective coupled dark matter/dark energy fluid -- that exists
2050: whenever such theories enter an adiabatic regime in which the scalar
2051: field faithfully tracks the minimum of the effective potential, and
2052: the coupling strength is strong compared to gravitational strength.
2053: The adiabatic regime occurs when the relaxation time scale associated with
2054: the scalar field is much shorter than the Hubble time. We have
2055: demonstrated how this instability can be viewed from the
2056: alternative perspectives of the kinetic theory of dark matter and as a
2057: Jeans instability associated with modified Newton's constants.
2058: 
2059: We have established the conditions under which the adiabatic instability occurs, finding  a condition on the coupling, $|\alpha'|\mpl\gg1$, and have identified the time and length scales over which the instability is active in a given setting governed by the matter density.  These length scales can differ greatly dependent on whether one is considering galactic or cosmic densities.
2060: 
2061: Our work builds on previous analyses  of MaVaN
2062: \cite{Fardon:2003eh,Kaplan:2004dq,Fardon:2005wc,Afshordi:2005ym,Bjaelde:2007ki},
2063: chameleon \cite{Khoury:2003aq,Khoury:2003rn}, and general coupled
2064: scenarios \cite{Kaplinghat:2006jk}. Our numerical analyses of coupled
2065: CDM and MaVaN models reinforce our analytic findings. We show that the
2066: regime of adiabatic behavior agrees with the predictions based on coupling strength and length scale
2067: conditions mentioned above.  In the appropriate limits, our results
2068: reduce to the previous findings in a number of cases, while in several
2069: other cases we have provided corrected results. In particular,  we
2070: have shown that stable MaVaN models exist that evade the instability,
2071: if $m_\nu'/m_\nu$ is not large compared to $\mpl$.
2072: 
2073: 
2074: 
2075: 
2076: \acknowledgments
2077: 
2078: We would like to thank Ole Bjaelde,  Anthony Brookfield, Steen
2079: Hannestad, Carsten Van der Bruck, Ira Wasserman and Christoph
2080: Wetterich for useful discussions in the course of this work.
2081: We thank an anonymous referee for some helpful comments.
2082: RB's work is supported by National Science Foundation (NSF) grants AST-0607018 and PHY-0555216, EF's by
2083: NSF grants PHY-0457200 and PHY-0555216, and MT's by NSF grant
2084: PHY-0354990 and by the Research Corporation.
2085: 
2086: \appendix
2087: 
2088: 
2089: \section{Effect of normal matter on instability}
2090: \label{visible}
2091: 
2092: In the body of this paper
2093: we have neglected the effect of baryons and the other matter
2094: species, assuming that their
2095: density is much smaller than that of the dark matter.
2096: A natural
2097: question is does the instability persist in regimes where the density
2098: of visible matter
2099: density is comparable to or larger than that of the dark matter -- one
2100: might expect the instability to be killed by
2101: the pressure of the visible matter.  In general the answer
2102: depends on the coupling function $\alpha_b$ in the action
2103: (\ref{action0}), which governs the coupling of the scalar field $\phi$
2104: to normal matter.  The instability persists in two specific cases, (i)
2105: $\alpha_b =0$ where normal matter is uncoupled from $\phi$, and
2106: (ii) $\alpha_b = \alpha_c$, the case usually considered in which
2107: dark matter and normal matter couple the same way to the scalar field,
2108: respecting the equivalence principle.
2109: 
2110: To derive this result, we generalize the derivation given in the body of the paper
2111: from a single, pressureless fluid to a set of $N$ non-interacting
2112: fluids, each of which can have a pressure.
2113: We start from the action (\ref{action0}),
2114: and we define the Jordan frame metric that the ${\rm
2115:   j}$th sector couples minimally to by
2116: \be
2117: {\bar g}_{{\rm j}\,ab} = e^{2 \alpha_{\rm j}(\phi)} g_{ab} \ .
2118: \label{Jordan1}
2119: \ee
2120: The stress energy tensor ${\bar T}_{{\rm j}}^{\,ab}$ of the ${\rm
2121: j}$th sector is defined by
2122: \be
2123: S_{\rm j}[ {\bar g}_{{\rm j}\,ab} + \delta {\bar g}_{{\rm j}\,ab}, \Psi_{\rm
2124:     j}] = S_{\rm j}[ {\bar g}_{{\rm j}\,ab}, \Psi_{\rm j}] + \frac{1}{2}
2125: \int d^4 x \sqrt{- {\bar g}_{\rm j}} {\bar T}_{{\rm j}}^{\,ab}
2126: \delta {\bar g}_{{\rm j}\,ab} \ ,
2127: \ee
2128: and we assume it has the form of a perfect fluid
2129: \be
2130: {\bar T}_{{\rm j}}^{\,ab} = ( {\bar \rho}_{\rm j} + {\bar p}_{\rm
2131:   j}) {\bar u}_{\rm j}^{\,a} {\bar u}_{\rm j}^{\,b} + {\bar
2132:   p}_{\rm j} {\bar g}_{\rm j}^{\,ab} \ .
2133: \ee
2134: Here ${\bar u}_{\rm j}^{\,a}$ is the 4-velocity which is normalized
2135: according to ${\bar g}_{{\rm j}\,ab} {\bar u}_{\rm j}^{\,a}
2136: {\bar u}_{\rm j}^{\,b} = -1$.
2137: This perfect fluid assumption requires that the matter in the jth
2138: sector be barotropic, ie that its pressure be determined uniquely by its density,
2139: which in many regimes is a good approximation.
2140: 
2141: The equation of motion of the ${\rm j}$th fluid is
2142: \be
2143: {\bar \nabla}_{{\rm j}\,a} {\bar T}_{\rm j}^{\,ab} = 0 \ ,
2144: \ee
2145: where ${\bar \nabla}_{{\rm j}\,a}$ is the derivative operator
2146: determined by the metric ${\bar g}_{{\rm j}\,ab}$.
2147: The components of this equation perpendicular to and parallel to the
2148: 4-velocity are
2149: \be
2150: {\bar a}_{\rm j}^a = - \frac{1}{{\bar \rho}_{\rm j} + {\bar p}_{\rm
2151:     j}} ({\bar g}_{\rm j}^{ab} + {\bar u}_{\rm j}^a {\bar u}_{\rm
2152:   j}^b)  {\bar \nabla}_{{\rm j}\,b} {\bar p}_{\rm j} \ ,
2153: \ee
2154: where ${\bar a}_{\rm j}^a = {\bar u}_{\rm j}^{\,b} {\bar \nabla}_{{\rm
2155:     j}\,b} {\bar u}_{\rm j}^{\,a}$ is the Jordan frame 4-acceleration,
2156: and
2157: \be
2158: {\bar \nabla}_{{\rm j}\,a} \left[ ({\bar \rho}_{\rm j} + {\bar p}_{\rm
2159:     j}) {\bar u}_{\rm j}^{\,a} \right] = {\bar u}_{\rm j}^{\,a} {\bar
2160:   \nabla}_{{\rm j}\,a} {\bar p}_{\rm j} \ .
2161: \ee
2162: Next, we conformally transform these fluid
2163: equations
2164: using Eq.\ (\ref{Jordan1}) to
2165: write them in terms
2166: of the Einstein frame metric $g_{ab}$ and the Einstein-frame
2167: normalized 4-velocities $u_{\rm j}^{\,a}=e^{\alpha_{\rm j}}{\bar u}_{\rm j}^{\,a}$ which satisfy $g_{ab} u_{\rm
2168:   j}^{\,a} u_{\rm j}^{\,b} =-1$.  This gives
2169: \be
2170: a_{\rm j}^{\,a} = -
2171:  (g^{ab} + u_{\rm j}^a u_{\rm j}^b)  \left[ \frac{ \nabla_b {\bar p}_{\rm j}}
2172:   {{\bar \rho}_{\rm j} + {\bar p}_{\rm j}} + \nabla_b \alpha_{\rm j}\right] \ ,
2173: \label{fluid1}
2174: \ee
2175: where $a_{\rm j}^{\,a} = u_{\rm j}^{\,b} \nabla_b u_{\rm j}^{\,a}$ and
2176: \be
2177: e^{-3 \alpha_{\rm j}} \nabla_a \left[ e^{3 \alpha_{\rm j}} ({\bar
2178:       \rho}_{\rm j} + {\bar p}_{\rm j} ) u_{\rm j}^{\,a} \right] =
2179:   u_{\rm j}^a \nabla_a {\bar p}_{\rm j} \ .
2180: \label{fluid2}
2181: \ee
2182: 
2183: 
2184: 
2185: Next, we consider the non-relativistic limit of the fluid equations
2186: (\ref{fluid1}) and (\ref{fluid2})
2187: together with the adiabatic limit of
2188: the scalar field equation (\ref{eq:scalar10}).  We also
2189: neglect self-gravity in the Einstein frame, taking $g_{ab} \approx \eta_{ab}$.  This
2190: approximation should be valid on sufficiently small spatial scales.
2191: We get
2192: \be
2193: \frac{\partial}{\partial t} \left( e^{3 \alpha_{\rm j}} {\bar
2194:   \rho}_{\rm j} \right) + {\bf \nabla} \cdot \left( e^{3 \alpha_{\rm
2195:     j}} {\bar \rho}_{\rm j} {\bf v}_{\rm j} \right) =0 \ ,
2196: \ee
2197: \be
2198: \frac{ \partial {\bf v}_{\rm j}} {\partial t} + ( {\bf v}_{\rm j}
2199: \cdot {\bf \nabla} ) {\bf v}_{\rm j} = - \frac{ {\bf \nabla} {\bar
2200:     p}_{\rm j}}{{\bar \rho}_{\rm j}} - {\bf \nabla} \alpha_{\rm j} \ ,
2201: \label{euler1}
2202: \ee
2203: where $\alpha_{\rm j} = \alpha_{\rm j}(\phi)$ and $\phi = \phi({\bar
2204:   \rho}_{\rm k})$ is given by the equation
2205: \be
2206: V'(\phi) = - \sum_{\rm j} \alpha'_{\rm j}(\phi) e^{4 \alpha_{\rm j}(\phi)}
2207: {\bar \rho}_{\rm j} \ .
2208: \label{scalar12}
2209: \ee
2210: 
2211: 
2212: Next, we switch to using the rescaled density variables
2213: \be
2214: \rho_{\rm j} = e^{3 \alpha_{\rm j}} {\bar \rho}_{\rm j} \ .
2215: \ee
2216: We linearize the resulting equations about the background solution
2217: $\rho_{\rm j} = \rho_{{\rm j}0} = $ constant, ${\bf
2218:   v}_{\rm j} = 0$, $\phi = \phi_0$.  In other words we assume that the
2219: dark matter and
2220: visible matter densities are constant and that the fluids are not
2221: in relative motion.  We define $\alpha_{{\rm j}0} = \alpha_{\rm
2222:   j}(\phi_0)$ and $\alpha'_{{\rm j}0} = \alpha'_{\rm j}(\phi_0)$,
2223: and look for a solution
2224: \begin{eqnarray}
2225: \rho_{\rm j} &=& \rho_{{\rm j}0} + \delta \rho_{\rm
2226:   j},\\
2227: \phi &=& \phi_0 + \delta \phi \ .
2228: \end{eqnarray}
2229: From Eq.\ (\ref{scalar12}) we obtain that $\delta \phi = \sum_j \chi_{\rm j}
2230: \delta \rho_{\rm j}$, where
2231: \be
2232: \chi_{\rm j} = - \frac{ \alpha'_{{\rm j}0} e^{\alpha_{{\rm j}0}} }
2233: {V''(\phi_0) + \sum_{\rm k} \left[ \alpha''_{{\rm k}0} +
2234:     (\alpha'_{{\rm k}0})^2\right] e^{ \alpha_{{\rm k}0}}
2235:     \rho_{{\rm k}0}} \ .
2236: \label{chidef}
2237: \ee
2238: Next we assume that all the variables are proportional to $\exp
2239: \left[i {\bf k} \cdot {\bf x} - i \omega t \right]$, and that the
2240: velocities are proportional to ${\bf k}$.  This assumption will
2241: yield one particular set
2242: of modes of oscillation of the coupled fluids, but these modes
2243: will contain the instability if it is present.  The resulting
2244: eigenvalue equation for $\omega^2$ is
2245: \be
2246: \omega^2  \delta \rho_{\rm j} = {\bf k}^2 \Gamma_{{\rm j}{\rm k}} \delta
2247:         \rho_{\rm k} \ ,
2248: \label{evaleqn}
2249: \ee
2250: where the matrix $\Gamma_{{\rm j}{\rm k}}$, which plays the role of
2251: the effective squared sound speed, is
2252: \be
2253: \Gamma_{{\rm j}{\rm k}} = {\bar c}_{\rm j}^2 \delta_{{\rm j}{\rm k}} +
2254: (1 - 3 {\bar c}_{\rm j}^2) \rho_{{\rm j}0} \alpha'_{{\rm j}0}
2255: \chi_{\rm k} \ .
2256: \label{Gamma}
2257: \ee
2258: Here
2259: \be
2260: {\bar c}_{\rm j}^2 = \frac{d {\bar p}_{\rm j}}{d {\bar \rho}_{\rm j}}
2261: \ee
2262: is the physical (Jordan frame) squared sound speed of the ${\rm j}$th fluid.
2263: 
2264: 
2265: From the eigenvalue equation (\ref{evaleqn}), it follows that
2266: the system will be stable if and only if
2267: all of the eigenvalues of
2268: the matrix $\Gamma_{{\rm j}{\rm k}}$ are real and positive.
2269: For the case of a single fluid with ${\bar c}_{\rm j}^2=0$, this criterion
2270: reduces to the criterion (\ref{unstablec}) derived in the body of the paper.
2271: In general, there is a competition between the positive squared
2272: sound speeds of the fluids in the first term in Eq.\ (\ref{Gamma}), and the
2273: (possibly) negative squared sound speeds coming from the interaction
2274: with the scalar field given in the second term.
2275: 
2276: 
2277: We now specialize to two fluids, a dark matter fluid with zero sound
2278: speed, and a fluid describing the visible matter.  Taking ${\rm j}
2279: = 1 = c$ for the dark matter (CDM) and ${\rm j} = 2 = b$ for
2280: the visible matter (baryons), and defining
2281: \be
2282: \nu_{\rm j}  = (1 - 3 {\bar c}_{\rm j}^2) \rho_{{\rm j}0} \alpha'_{{\rm j}0}
2283: \label{nudef}
2284: \ee
2285: gives for the matrix
2286: \begin{equation}
2287: \Gamma_{{\rm j}{\rm k}} = \left( \begin{array}{cc}
2288: \nu_{c} \chi_c &
2289: \nu_{c} \chi_b   \\
2290: \nu_{b} \chi_c   &
2291: {\bar c}_b^2 + \nu_b \chi_b
2292: \end{array} \right) \ .
2293: \label{matrix1}
2294: \end{equation}
2295: Now if the system is stable, then both of the eigenvalues $\lambda_1$
2296: and $\lambda_2$ of the
2297: matrix must be real and nonnegative, and hence both the determinant
2298: $\lambda_1 \lambda_2$ and
2299: the trace $\lambda_1 + \lambda_2$ of the matrix must be nonnegative.
2300: Conversely, if either the determinant or the trace of the matrix is
2301: negative, then the system is unstable.
2302: The determinant is
2303: \begin{eqnarray}
2304: {\rm det} \, {\bf \Gamma} &=& {\bar c}_b^2 \nu_c
2305:  \chi_c
2306: = - \frac{ ( \alpha'_c)^2 e^{\alpha_c} {\bar c}_b^2 \rho_c}
2307: {\Upsilon
2308: } \ , \ \ \ \
2309: \label{detGamma}
2310: \end{eqnarray}
2311: where
2312: \bea
2313: \Upsilon &=& V'' + \left[ \alpha''_c + (\alpha'_c)^2 \right]
2314: e^{\alpha_{c}} \rho_c
2315: \nonumber \\ &&
2316: + \left[ \alpha''_b + (\alpha'_b)^2 \right] e^{\alpha_b} \rho_b
2317: \label{Upsilondef}
2318: \eea
2319: and where we have used Eqs.\ (\ref{chidef}) and (\ref{nudef}).
2320: In Eqs.\ (\ref{detGamma}) and (\ref{Upsilondef}), $\phi$ can be taken
2321: to be the function of $\rho_{c}$ and $\rho_b$ given by [cf.\ Eq.\
2322: (\ref{fluid2})]
2323: \be
2324: V'(\phi) = - \alpha'_c(\phi) e^{\alpha_c(\phi)}
2325: \rho_c - \alpha'_b(\phi) e^{\alpha_b(\phi)}
2326: \rho_b \ .
2327: \label{Phieqn4}
2328: \ee
2329: Equations (\ref{detGamma}) -- (\ref{Phieqn4}) allow us to determine
2330: which values of $\rho_c$ and $\rho_b$ satisfy the sufficient condition
2331: ${\rm det} \, {\bf \Gamma} < 0$ for instability,
2332: given the functions $V(\phi)$, $\alpha_c(\phi)$ and
2333: $\alpha_b(\phi)$.  Note that $\alpha_c(\phi)$ was denoted by
2334: $\alpha(\phi)$ in the body of the paper.
2335: 
2336: 
2337: 
2338: We now consider two special cases.  If $\alpha'_b=0$, so that
2339: the visible matter is not coupled to the scalar field, then
2340: eigenvalues of the matrix ${\bf \Gamma}$ are just ${\bar c}_{b}^2$ and
2341: $\nu_c \chi_c$.  The second of these
2342: eigenvalues coincides with the effective squared sound speed computed above in
2343: Eq.\ (\ref{soundspeed}).  Thus we recover the results of the body of
2344: the paper.
2345: 
2346: 
2347: The second special case is when $\alpha_b = \alpha_c$.
2348: In this case the instability criterion ${\rm det} \, {\bf \Gamma}<0$ again reduces to the criterion
2349: (\ref{soundspeed}) computed earlier for a single fluid, with the
2350: modification that the effective sound speed squared is now a function
2351: of the total density $\rho = \rho_c + \rho_b$
2352: rather than just of $\rho_c$.  If this
2353: total density is in the unstable regime, then the instability will
2354: persist despite the pressure of the visible matter.
2355: 
2356: Finally, one can also consider the effect of the scalar field on just
2357: the baryons, neglecting the dark matter.  The sound speed of the
2358: baryons gets an additional term, but it is a small correction unless
2359: the coupling is large, $\alpha_b^\prime \mpl \gg 1$, and observational
2360: tests of general relativity in the Solar System
2361: require $\alpha_b^\prime \mpl \ll 1$.
2362: 
2363: \section{Kinetic theory treatment of instability}
2364: \label{kinetic}
2365: 
2366: 
2367: In this appendix we describe dark matter using the collisionless
2368: Boltzmann equation, and specialize to the non-relativistic regime and
2369: to lengthscales small enough that self-gravity can be neglected.
2370: We show, first, that the instability is generic, occurring for any
2371: velocity distribution function with sufficiently small velocity
2372: dispersion, and, second, that for Maxwellian distributions
2373: the fluid is stabilized by free streaming once the velocity dispersion
2374: becomes larger than a critical value.
2375: 
2376: Starting from the general
2377: action (\ref{action0}), we specialize to the non-relativistic limit
2378: neglecting self-gravity ($g_{ab} \approx \eta_{ab}$), which will
2379: be valid on sufficiently small spatial scales.  The Jordan-frame
2380: metric is then
2381: $$
2382: ds^2 = e^{2 \alpha} (-dt^2 + d{\bf x}^2) \ .
2383: $$
2384: We assume that dark matter is composed of non-interacting,
2385: non-relativistic particles of mass $\mu$.
2386: We denote by $f(t,{\bf x},{\bf v})$ the one-particle distribution function,
2387: normalized so that the number of particles in the volume element
2388: $d^3x$ and in the velocity region $d^3v$ is
2389: $$
2390: f(t,{\bf x},{\bf v})  d^3x d^3v \ .
2391: $$
2392: Now the physical (Jordan-frame)
2393: volume element is $e^{3 \alpha} d^3 x$,
2394: so the Jordan frame mass density ${\bar \rho}$ [cf.\
2395: Eq.\ (\ref{ee0}) above] is given by
2396: $
2397: {\bar \rho} = \mu e^{-3 \alpha} \int d^3v f.
2398: $
2399: From Eq.\ (\ref{newdensity}) the rescaled density variable $\rho$ is then given
2400: by
2401: \be
2402: \rho = \mu \int d^3v f \ .
2403: \label{kineticrho}
2404: \ee
2405: With this notation the collisionless Boltzmann equation is [cf.\ Eq.\
2406: (\ref{euler1}) above with ${\bar p}_{\rm j}=0$]
2407: \be
2408: \frac{\partial f}{\partial t} + v^i \frac{\partial f}{\partial x^i} =
2409: \frac{\partial \alpha}{\partial x^i} \frac{\partial f}{\partial v^i} \ .
2410: \ee
2411: In the adiabatic limit, the variable $\alpha$ in this equation is
2412: given by $\alpha = \alpha(\phi)$, where $\phi$ is given in terms of
2413: $\rho$ by Eq.\ (\ref{eq:alg}), and $\rho$ is given in turn in terms
2414: of $f$ by Eq.\ (\ref{kineticrho}).
2415: 
2416: 
2417: We now linearize the Boltzmann equation about a homogeneous background
2418: solution,
2419: taking
2420: $
2421: f = f_0({\bf v}) + \delta f(t,{\bf x},{\bf v}).
2422: $
2423: and taking $\delta f \propto \exp[- i \omega t + i {\bf k} \cdot {\bf
2424:   x}]$.  Using the identity $d\alpha/d\rho = c_s^2/\rho$ from Eq.\
2425: (\ref{soundspeed}), and integrating over ${\bf v}$, gives the dispersion
2426: relation
2427: \be
2428: \int d^3 v \frac{ {\bf k} \cdot \frac{\partial f_0}{\partial {\bf
2429:       v}}}{\omega - {\bf k} \cdot {\bf v}} = - \frac{\rho_0}{\mu c_s^2} \ .
2430: \ee
2431: Here $\rho_0$ is the background density.
2432: Without loss of generality we take ${\bf k}$ to be in the
2433: $z$-direction, assuming $f_0$ is isotropic, and we define
2434: \be
2435: F(v) = \frac{\mu}{\rho_0} \int_{-\infty}^\infty dv_x
2436: \int_{-\infty}^\infty dv_y \, f_0(v_x,v_y,v) \ .
2437: \ee
2438: This distribution function is normalized so that $\int dv F =1$, and
2439: the dispersion relation can now be written as
2440: \be
2441: 1 + c_s^2 \int dv \, \frac{F'(v)}{\omega/k - v} =0 \ .
2442: \label{dispersion}
2443: \ee
2444: Given the distribution function $F(v)$ we can solve this equation to obtain
2445: $\omega = \omega(k)$, which is complex in general.
2446: 
2447: The integrand in Eq.\ (\ref{dispersion}) contains a singularity at $v =
2448: \omega/k$.  The derivation given here is incomplete since it does not
2449: provide a specification for how to deal with the singularity.
2450: However, just as in plasma physics \cite{Landau}, it is possible to give an
2451: alternative derivation based on Laplace transforms.  That alternative derivation yields the
2452: following specification: the integral over $v$ must be taken over the
2453: Landau contour, which runs along the real axis if ${\rm Im}(\omega) > 0$, but
2454: dips below the real axis to encircle the pole at $v = \omega/k$ if
2455: ${\rm Im}(\omega) \le 0$.
2456: 
2457: 
2458: We now specialize to the adiabatic regime where $c_s^2 < 0$, and we
2459: write $c_s^2 = - \beta^2$.  We write the complex frequency in terms of
2460: its real and imaginary parts, $\omega = \omega_r +  i \omega_i$, and
2461: we look for a solution $\omega = \omega(k)$ of the dispersion relation
2462: (\ref{dispersion}) with $\omega_i > 0$, corresponding to an unstable
2463: mode.  For such a solution the Landau contour is along the real
2464: axis which simplifies the analysis.
2465: 
2466: If the velocity dispersion of the distribution $F(v)$ is small, we can
2467: expand the denominator of the integrand in Eq.\ (\ref{dispersion}) as a power series in
2468: $v k / \omega$.  Integrating by parts and solving the resulting
2469: equation for $\omega^2$ gives
2470: \be
2471: \omega^2 = k^2 \left[ - \beta^2 + 3 \sigma^2 +
2472:   O\left(\frac{\sigma^4}{\beta^2}\right) \right] \ ,
2473: \label{omegaans}
2474: \ee
2475: where $\sigma^2 = \int dv v^2 F(v)$.
2476: This equation has a solution with positive imaginary part,
2477: $\omega_i>0$, consistent
2478: with the assumption used in its derivation.  Therefore
2479: the instability is generic, present for
2480: any velocity distribution, as long as $\sigma \ll \beta$.
2481: Equation (\ref{omegaans}) also suggests that the instability will be
2482: removed when $\sigma$ gets large, as argued in Ref.\ \cite{Afshordi:2005ym},
2483: since then the positive second term
2484: in the square brackets will overcome the negative first term.
2485: However, Eq.\ (\ref{omegaans}) is only valid in the regime $\sigma \ll
2486: \beta$, and so to investigate stabilization we must use an alternative
2487: method of computation.
2488: 
2489: Returning to the general dispersion relation (\ref{dispersion}), we obtain from
2490: its real and imaginary parts the equations
2491: \be
2492: 1 - k \beta^2 \int_{-\infty}^\infty dv \, \frac{ (\omega_r - k v )
2493:   F'(v)}{(\omega_r - k v)^2 + \omega_i^2} =0 \ ,
2494: \label{realpart}
2495: \ee
2496: and
2497: \be
2498: \int_{-\infty}^\infty dv \, \frac{
2499:   F'(v)}{(\omega_r - k v)^2 + \omega_i^2} =0 \ .
2500: \label{imagpart}
2501: \ee
2502: We now specialize to the Maxwellian velocity distribution
2503: \be
2504: F(v) = \frac{1}{\sqrt{2 \pi} \sigma} e^{ - \frac{ v^2}{2 \sigma^2}} \ .
2505: \ee
2506: For this case Eq.\ (\ref{imagpart}) has a unique solution for
2507: $\omega_r$, namely $\omega_r =0$.  This can be seen by splitting the
2508: integral into $v>0$ and $v<0$ contributions and substituting $v \to
2509: -v$ in the $v<0$ term.  This yields
2510: \be
2511: 0 =
2512: \omega_r \int_0^\infty dv
2513: \frac{ v^2 e^{- \frac{v^2}{2 \sigma^2}}}
2514: { \left[ (\omega_r - k v)^2 +
2515:       \omega_i^2 \right] \left[ (\omega_r + k v)^2 +
2516:       \omega_i^2 \right]} \ ,
2517: \ee
2518: and the result follows since the integrand is everywhere positive.
2519: Equation (\ref{realpart}) now simplifies to
2520: \be
2521: \frac{\sigma^2}{\beta^2} = \int_{-\infty}^\infty d{\bar v}
2522: \left(
2523: \frac{
2524: {\bar v}^2
2525: }
2526: {
2527: {\bar v}^2 + \kappa^2
2528: }
2529: \right)
2530: \, \frac{1}{\sqrt{2\pi}} e^{- \frac{{\bar v}^2}{2}} \ ,
2531: \label{finale}
2532: \ee
2533: where we have defined ${\bar v} = v / \sigma$ and $\kappa = \omega_i /
2534: (k \sigma)$.
2535: 
2536: Equation (\ref{finale})
2537: determines $\kappa$ and thence $\omega_i$ as a function of $\sigma/\beta$.  When
2538: $\kappa$ is large, expanding the factor in round brackets in the
2539: integrand as a power series in ${\bar v}/\kappa$
2540: gives the result (\ref{omegaans}) above.  As $\kappa$ decreases,
2541: $\sigma/\beta$ increases, until as $\kappa \to 0$, $\sigma/\beta$
2542: approaches the limiting value $\sigma/\beta =1$.  It can be seen that
2543: there are no solutions to Eq.\ (\ref{finale}) with $\sigma/\beta>1$,
2544: since the right hand side is a monotonic function of $\kappa^2$.
2545: Therefore, whenever\footnote{Note that the critical value of velocity
2546:   dispersion $\sigma=\beta$ is a factor of $\sqrt{3}$ larger than indicated by the
2547:   approximate formula (\ref{omegaans}) which was used in Ref.\ \cite{Afshordi:2005ym}.}
2548: \be
2549: \sigma > \beta \ ,
2550: \ee
2551: there are no unstable modes with $\omega_i > 0$.  In other words, the
2552: instability has been removed by the damping process associated with
2553: the finite dispersion (Landau damping, also called free streaming).
2554: The perturbation to the distribution function
2555: is proportional to the integrand in Eq.\ (\ref{finale}).  Therefore
2556: the nature of the damping is that interaction with the growing
2557: mode moves some particles from velocities $v \sim \beta$ to larger
2558: velocities, removing energy from the mode.
2559: 
2560: 
2561: 
2562: \section{Effective Newton's constant}
2563: \label{sec:Geff}
2564: 
2565: In this appendix we derive the formula (\ref{Gformula}) for the effective Newton's
2566: constant for the theory (\ref{action0}).
2567: 
2568: Since Newton's constant $G$ is dimensionful, we need to define the
2569: system of units used in measurements of $G$, as our final result will
2570: depend on the system of units chosen.  Here we are not concerned with
2571: changes of units in the usual sense where the ratios between the old
2572: and new standards of mass, length and time are constants, independent
2573: of space and time.  Rather, we are concerned with changes in the
2574: operational procedures for how units are defined, for which the ratios
2575: between the old and new standards can vary with space and time, as
2576: discussed by Dicke \cite{Dicke}.  Changes of units of this type
2577: encompass conformal transformations of the metric, but can be more
2578: general.  The necessity of specifying a system
2579: of units before discussing the space-time variation of dimensionful
2580: constants of nature is discussed in detail in Duff \cite{Duff}.
2581: 
2582: We choose to use systems of units which
2583: are determined completely by physics in the
2584: visible sector (which we label by ${\rm j}= b$ for baryonic), and are not
2585: determined by
2586: gravitational physics or by physics of the dark matter sector.
2587: For example, this is true of SI units.  Alternatively one could take
2588: one's standards of mass, length and time to be the mass, size, and
2589: light travel time across a Hydrogen atom.  Any system of units of this
2590: type will give rise to the formula (\ref{Gformula}) below for $G$.
2591: 
2592: We note that alternative choices are in principle possible.  For
2593: example, one could have a small fiducial black hole, and one could transport
2594: it adiabatically around the Universe to act as a standard.  Then one
2595: could define the standard of mass to be the mass of the black hole,
2596: while defining the standards of length and time in terms of Hydrogen
2597: atoms as above.  In this system of units $G$ would vary in space and
2598: time, but in a manner different to that described by the formula
2599: (\ref{Gformula}) below, and Planck's constant $\hbar$ would also vary in
2600: space and time.  Another possibility would be to define the units of
2601: mass in terms of the mass of the dark matter particle; this would also
2602: yield a different formula for $G$ whenever the dark matter coupling
2603: function $\alpha_c(\phi)$ differs from that of the baryons
2604: $\alpha_b(\phi)$.
2605: 
2606: We next discuss the quantities on which the formula for $G$ depends.  First,
2607: for the theory (\ref{action0}) with several different matter sectors,
2608: Newton's constant becomes a matrix whose i,j element governs the
2609: strength of the gravitational interaction between sector ${\rm i}$ and
2610: sector j.  Second, $G$ will depend on the scalar field
2611: $\phi$, and through this dependence become a function of space and
2612: time.  Here $\phi$ should be thought of as a background value of the
2613: scalar field; gravitational interactions that act as perturbations
2614: to this background, for which the perturbation to the
2615: scalar field can be treated linearly, have a strength described by the
2616: formula (\ref{Gformula}).  Finally, since the scalar contribution to
2617: the gravitational force will in general have a Yukawa profile, $G$
2618: will also depend on a spatial wave vector ${\bf k}$, which is measured
2619: in the units discussed above.
2620: 
2621: 
2622: 
2623: 
2624: Our formula for Newton's constant is
2625: \be
2626: G_{{\rm i}{\rm j}}(\phi,{\bf k}) = G e^{2 \alpha_b(\phi)} \left[ 1
2627:   + \frac{2 \mpl^2 \alpha_{\rm i}^\prime(\phi) \alpha_{\rm j}^\prime(\phi)}{1 +
2628: \frac{m_{\rm eff}^2(\phi,\rho)}{e^{2 \alpha_b(\phi)} {\bf k}^2 }
2629: } \right] \ .
2630: \label{Gformula}
2631: \ee
2632: Here on the right hand side, $G$ is a constant, related to $\mpl^2$ by
2633: $G = 1 / (8 \pi \mpl^2)$, and the effective mass $m_{\rm eff}(\phi,\rho)$
2634: is defined by
2635: \be
2636: m_{\rm eff}^2(\phi,\rho) = \frac{ \partial^2 V_{\rm
2637:     eff}(\phi,\rho)}{\partial^2 \phi} \ ,
2638: \label{meffgen}
2639: \ee
2640: where the effective potential is given by Eq.\ (\ref{phieqn}).  In the
2641: adiabatic regime this effective mass reduces to the mass (\ref{meffdef}).
2642: Several different aspects of this formula have appeared before in the
2643: literature.  The overall prefactor of $e^{2\alpha_b}$ is well known, and
2644: the formula for the case of one matter sector has been previously derived in
2645: the context of cosmological perturbation theory \cite{Amendola:2004a}.
2646: The $1$ in the square brackets describes the exchange of a graviton,
2647: and the second term in the square brackets describes the exchange of a
2648: scalar quantum, which couples to particles in the sector i with an
2649: amplitude proportional to $\alpha_{\rm i}'(\phi)$.
2650: The scalar coupling term
2651: vanishes in the long wavelength limit $k \to 0$ if the scalar
2652: field has a finite effective mass $m_{\rm eff}$.
2653: 
2654: 
2655: 
2656: 
2657: We now turn to the derivation of the formula (\ref{Gformula}).
2658: For the derivation it will be convenient to use a different system of
2659: units, which are defined as follows.
2660: The standards of mass, length and time are defined to be the mass,
2661: size and light travel time across a small fiducial black hole, which
2662: is transported adiabatically around the Universe to act as a
2663: reference.  Equivalent units\footnote{By equivalent units we mean
2664:   units that differ only by multiplication by constants.} can be
2665: defined by demanding that
2666: the speed of light $c$ and Planck's constant $\hbar$ be unity, and
2667: that lengths and times are measured using the Einstein frame metric
2668: $g_{ab}$. At the end of the derivation we will transform back to the
2669: visible-sector-based units.
2670: 
2671: Starting with the action (\ref{action0}), we perform the following
2672: steps.  We specialize the matter action in the jth sector to be that
2673: of a collection of point particles with masses $m^{({\rm
2674:     j})}_{\alpha_{\rm j}}$
2675: and worldlines $x^{({\rm j})a}_{\alpha_{\rm j}}(\lambda)$:
2676: \be
2677: S_{\rm j}[e^{2 \alpha_{\rm j}(\phi)} g_{ab}, \Psi_{\rm j}] =
2678: \sum_{\alpha_{\rm j}}
2679: m^{({\rm j})}_{\alpha_{\rm j}} \int d\lambda \sqrt{ - e^{2 \alpha_{\rm
2680:       j}} g_{ab}
2681:   \frac{d x^{({\rm j})a}_{\alpha_{\rm j}}}{d\lambda}   \frac{d
2682:     x^{({\rm j})b}_{\alpha_{\rm j}}}{d\lambda}} \ .
2683: \ee
2684: We specialize to the Newtonian limit, so that the Einstein-frame
2685: metric $g_{ab}$ is of the form $- (1 + 2 \Phi) dt^2 + (1 - 2 \Phi)
2686: d{\bf x}^2$.  We write the scalar field as $\phi + \delta \phi$, and
2687: expand to quadratic order in $\delta \phi$ and $\Phi$.  We assume that there is a
2688: background mass density $\rho$, so that the effective mass of the
2689: scalar field is given by Eq.\ (\ref{meffgen}).
2690: Finally, we compute from
2691: the action the energy $E$ for a static configuration, up to an
2692: additive constant.  This yields
2693: \bea
2694: E &=& \int d^3 x \left\{ \mpl^2 ({\bf \nabla} \Phi)^2 + \frac{1}{2} (
2695:   {\bf \nabla} \delta \phi)^2 + \frac{1}{2} m_{\rm eff}^2 \delta
2696: \phi^2 \right. \nn \\
2697: && + \sum_{\rm j} e^{\alpha_{\rm j}(\phi)} \left[ 1 + \Phi +
2698:       \alpha_{\rm j}^\prime(\phi) \delta \phi \right] \rho_{\rm j}  \bigg\} \ ,
2699: \eea
2700: where
2701: \be
2702: \rho_{\rm j} = \sum_{\alpha_{\rm j}} m^{({\rm j})}_{\alpha_{\rm j}} \delta^3[ {\bf x} - {\bf
2703:   x}^{({\rm j})}_{\alpha_{\rm j}} ] \ .
2704: \ee
2705: Now integrating out $\Phi$ and $\delta \phi$ yields
2706: \bea
2707: E &=& - \frac{1}{4 \mpl^2} \sum_{{\rm i},{\rm j}} \sum_{\alpha_1, \alpha_2, \ldots}
2708: \sum_{\beta_1, \beta_2, \ldots} e^{\alpha_{\rm i}(\phi) + \alpha_{\rm j}(\phi)}
2709: \frac{ m^{({\rm i})}_{\alpha_{\rm i}} m^{({\rm j})}_{\beta_{\rm
2710:       j}}}{r^{{\rm i}{\rm j}}_{\alpha\beta}}
2711: \nn \\
2712: && \times \left[ 1 + 2 \mpl^2
2713:       \alpha_{\rm i}'(\phi) \alpha_{\rm j}'(\phi) e^{-
2714:       m_{\rm eff} r^{{\rm i}{\rm j}}_{\alpha\beta} }\right] \ ,
2715: \label{energy2}
2716: \eea
2717: where $r^{{\rm i}{\rm j}}_{\alpha\beta} = | {\bf x}^{({\rm
2718:     i})}_{\alpha_{\rm i}} - {\bf x}^{({\rm j})}_{\beta_{\rm j}}|$.
2719: 
2720: Now we transform back to the visible sector or Jordan frame units.
2721: The energy ${\tilde E}$, distances ${\tilde r}^{{\rm i}{\rm
2722:     j}}_{\alpha\beta}$ and masses ${\tilde m}^{({\rm i})}_{\alpha_{\rm
2723:     i}}$ in those units are given by
2724: \bes
2725: \bea
2726: {\tilde E} &=& e^{-\alpha_b(\phi)} E,  \\
2727: {\tilde r}^{{\rm i}{\rm j}}_{\alpha\beta} &=& e^{\alpha_b(\phi)}
2728: r^{{\rm i}{\rm j}}_{\alpha\beta}, \\
2729: {\tilde m}^{({\rm i})}_{\alpha_{\rm i}} &=& e^{-\alpha_b(\phi)} \left[
2730:   e^{\alpha_{\rm i}(\phi)} m^{({\rm i})}_{\alpha_{\rm i}} \right] \ ,
2731: \label{jmass}
2732: \eea
2733: \ees
2734: where the factor in square brackets on the right hand side of Eq.\
2735: (\ref{jmass}) is the mass as measured in the Einstein-frame units.
2736: Substituting these unit transformations into the energy expression
2737: (\ref{energy2}) and transforming from position space to momentum space
2738: yields an expression for the energy from which we can
2739: finally read off the effective Newton's constant (\ref{Gformula}).
2740: 
2741: 
2742: 
2743: 
2744: \begin{thebibliography}{99}
2745: 
2746: \bibitem{Clowe:2006eq}
2747:  D.~Clowe et. al.,
2748:   arXiv:astro-ph/0608407.
2749: 
2750: \bibitem{Damour:1990}
2751:  T.~Damour, G.~W.~Gibbons and C.~Gundlach,
2752:  % "DARK MATTER, TIME VARYING G, AND A DILATON FIELD"
2753:  Phys. Rev. Lett. {\bf 64}, 123 (1990).
2754: 
2755: \bibitem{Carroll:1998zi}
2756:  S.~Carroll,
2757:  % "Quintessence and the rest of the world"
2758: Phys. Rev. Lett. {\bf 81}, 3067 (1998).
2759: 
2760: %\cite{Amendola:1999er}
2761: \bibitem{Amendola:1999er}
2762:   L.~Amendola,
2763:   %``Coupled quintessence,''
2764:   Phys.\ Rev.\ D {\bf 62}, 043511 (2000)
2765:   [arXiv:astro-ph/9908023].
2766:   %%CITATION = ASTRO-PH 9908023;%%
2767: 
2768: %\cite{Bean:2000zm}
2769: \bibitem{Bean:2000zm}
2770:   R.~Bean and J.~Magueijo,
2771:   % ``Dilaton-derived quintessence scenario leading naturally to the  late-time
2772:   %acceleration of the universe,''
2773:   Phys.\ Lett.\ B {\bf 517}, 177 (2001)
2774:   [arXiv:astro-ph/0007199].
2775:   %%CITATION = ASTRO-PH 0007199;%%
2776: 
2777:   %\cite{Bean:2001ys}
2778: \bibitem{Bean:2001ys}
2779:   R.~Bean,
2780:   %``Perturbation evolution with a non-minimally coupled scalar field,''
2781:   Phys.\ Rev.\ D {\bf 64}, 123516 (2001)
2782:   [arXiv:astro-ph/0104464].
2783:   %%CITATION = ASTRO-PH 0104464;%%
2784: 
2785: \bibitem{Majerotto:2004ji}
2786:   E.~Majerotto, D.~Sapone and L.~Amendola,
2787:   % "Supernovae type Ia data favor negatively coupled phantom energy"
2788:  arXiv:astro-ph/0410543
2789: 
2790: \bibitem{Das:2005yj}
2791:   S.~Das, P.~S.~Corasaniti and J.~Khoury,
2792: %  "Super-acceleration as signature of dark sector interaction"
2793:   Phys. Rev. D {\bf 73}, 083509 (2006).
2794: 
2795: \bibitem{Lee:2006za}
2796:   S.~Lee, G-C.~Liu and K-W.~Ng,
2797:   % "Constraints on the coupled quintessence from cosmic
2798:   %  microwave  background anisotropy and matter power spectrum"
2799:  Phys. Rev. D {\bf 73}, 083516 (2006).
2800: 
2801: %\cite{Kesden:2006zb}
2802: \bibitem{Kesden:2006zb}
2803:   M.~Kesden and M.~Kamionkowski,
2804:   %``Galilean Equivalence for Galactic Dark Matter,''
2805:   Phys.\ Rev.\ Lett.\  {\bf 97}, 131303 (2006)
2806:   [arXiv:astro-ph/0606566].
2807:   %%CITATION = PRLTA,97,131303;%%
2808: 
2809: 
2810: 
2811: %\cite{Chiba:2003ir}
2812: \bibitem{Chiba:2003ir}
2813:   T.~Chiba,
2814:   %``1/R gravity and scalar-tensor gravity,''
2815:   Phys.\ Lett.\ B {\bf 575}, 1 (2003)
2816:   [arXiv:astro-ph/0307338].
2817:   %%CITATION = ASTRO-PH 0307338;%%
2818: 
2819: %\cite{Tsujikawa:2007gd}
2820: \bibitem{Tsujikawa:2007gd}
2821:   S.~Tsujikawa,
2822:   %``Matter density perturbations and effective gravitational constant in
2823:   %modified gravity models of dark energy,''
2824:   Phys.\ Rev.\  D {\bf 76}, 023514 (2007)
2825:   [arXiv:0705.1032 [astro-ph]].
2826:   %%CITATION = PHRVA,D76,023514;%%
2827: 
2828: 
2829: \bibitem{Amendola:2006mr}
2830:   L.~Amendola, D.~Polarski, and S.~Tsujikawa,
2831:   %``Are f(R) dark energy models cosmologically viable?,''
2832:   [arXiv:astro-ph/0603703].
2833:   %%CITATION = ASTRO-PH 0603703;%%
2834: 
2835: 
2836: \bibitem{Agarwal:2007}
2837: N. ~Agarwal and R. ~Bean,
2838: %``The dynamical viability of scalar-tensor theories with a general coupling"
2839: [arXiv:0708.3967].
2840: 
2841: \bibitem{Afshordi:2005ym}
2842:  N.~Afshordi, M.~Zaldarriaga and K.~Kohri,
2843:  % "On the stability of dark energy with mass-varying neutrinos"
2844:  Phys. Rev. D {\bf 72}, 065024 (2005).
2845: 
2846: 
2847: \bibitem{Kaplinghat:2006jk}
2848:   M.~Kaplinghat and A.~Rajaraman,
2849:   %``Stable models of super-acceleration,''
2850:   arXiv:astro-ph/0601517.
2851:   %%CITATION = ASTRO-PH 0601517;%%
2852: 
2853: 
2854: 
2855: 
2856: 
2857: %\cite{Bjaelde:2007ki}
2858: \bibitem{Bjaelde:2007ki}
2859:   O.~E.~Bjaelde, A.~W.~Brookfield, C.~van de Bruck, S.~Hannestad, D.~F.~Mota, L.~Schrempp and D.~Tocchini-Valentini,
2860:   %``Neutrino Dark Energy -- Revisiting the Stability Issue,''
2861:   arXiv:0705.2018 [astro-ph].
2862:   %%CITATION = ARXIV:0705.2018;%%
2863: 
2864: \bibitem{letter}
2865: R.\ Bean, \'E.\ \'E.\ Flanagan, M. Trodden,
2866: %{\it The Adiabatic Instability on Cosmology's Dark Side},
2867: arXiv:0709.1124 [astro-ph].
2868: 
2869: \bibitem{Khoury:2003aq}
2870:  J.~Khoury, A.~Weltman,
2871:  % "Chameleon fields: Awaiting surprises for tests of gravity in space"
2872:  Phys. Rev. Lett. {\bf 93}, 171190 (2004).
2873: 
2874: \bibitem{Khoury:2003rn}
2875: J.~Khoury and A.~Weltman,
2876:  %"Chameleon cosmology"
2877: Phys. Rev. D {\bf 69}, 044026 (2004).
2878: 
2879: %\cite{Mota:2006ed}
2880: \bibitem{Mota:2006ed}
2881:   D.~F.~Mota and D.~J.~Shaw,
2882:   %``Strongly coupled chameleon fields: New horizons in scalar field theory,''
2883:   Phys.\ Rev.\ Lett.\  {\bf 97}, 151102 (2006)
2884:   [arXiv:hep-ph/0606204].
2885:   %%CITATION = PRLTA,97,151102;%%
2886: 
2887: %\cite{Fardon:2003eh}
2888: \bibitem{Fardon:2003eh}
2889:   R.~Fardon, A.~E.~Nelson and N.~Weiner,
2890:  %``Dark energy from mass varying neutrinos,''
2891:  JCAP {\bf 0410}, 005 (2004)
2892:   [arXiv:astro-ph/0309800].
2893:   %%CITATION = JCAPA,0410,005;%%
2894: 
2895: %\cite{Fardon:2005wc}
2896: \bibitem{Fardon:2005wc}
2897:   R.~Fardon, A.~E.~Nelson and N.~Weiner,
2898:   %``Supersymmetric theories of neutrino dark energy,''
2899:  JHEP {\bf 0603}, 042 (2006)
2900:   [arXiv:hep-ph/0507235].
2901:   %%CITATION = JHEPA,0603,042;%%
2902: 
2903: %\cite{Kaplan:2004dq}
2904: \bibitem{Kaplan:2004dq}
2905:   D.~B.~Kaplan, A.~E.~Nelson and N.~Weiner,
2906:   %``Neutrino oscillations as a probe of dark energy,''
2907:  Phys.\ Rev.\ Lett.\  {\bf 93}, 091801 (2004)
2908:   [arXiv:hep-ph/0401099].
2909:   %%CITATION = PRLTA,93,091801;%%
2910: 
2911: 
2912: 
2913: 
2914:   %\cite{Carroll:2003wy}
2915: \bibitem{Carroll:2003wy}
2916:   S.~M.~Carroll, V.~Duvvuri, M.~Trodden and M.~S.~Turner,
2917:   %``Is cosmic speed-up due to new gravitational physics?,''
2918:   Phys.\ Rev.\ D {\bf 70}, 043528 (2004)
2919:   [arXiv:astro-ph/0306438].
2920:   %%CITATION = ASTRO-PH 0306438;%%
2921: 
2922: 
2923: %\cite{Carroll:2006jn}
2924: \bibitem{Carroll:2006jn}
2925:   S.~M.~Carroll, I.~Sawicki, A.~Silvestri and M.~Trodden,
2926:   %``Modified-Source Gravity and Cosmological Structure Formation,''
2927:   New J.\ Phys.\  {\bf 8}, 323 (2006)
2928:   [arXiv:astro-ph/0607458].
2929:   %%CITATION = ASTRO-PH 0607458;%%
2930: 
2931: 
2932: %\cite{Damour:1994ya}
2933: \bibitem{Damour:1994ya}
2934:   T.~Damour and A.~M.~Polyakov,
2935:   %``String theory and gravity,''
2936:   Gen.\ Rel.\ Grav.\  {\bf 26}, 1171 (1994)
2937:   [arXiv:gr-qc/9411069].
2938:   %%CITATION = GRGVA,26,1171;%%
2939: 
2940: 
2941: %\cite{Farrar:2003uw}
2942: \bibitem{Farrar:2003uw}
2943:   G.~R.~Farrar and P.~J.~E.~Peebles,
2944:   %``Interacting dark matter and dark energy,''
2945:   Astrophys.\ J.\  {\bf 604}, 1 (2004)
2946:   [arXiv:astro-ph/0307316].
2947:   %%CITATION = ASJOA,604,1;%%
2948: 
2949: 
2950: \bibitem{Faulkner:2006ub}
2951:  T.~Faulkner, M.~Tegmark, E.~F.~Bunn and Y.~Mao,
2952:  % "Constraining f(R) gravity as a scalar tensor theory"
2953:  arXiv: astro-ph/0612569
2954: 
2955: \bibitem{Navarro:2006mw}
2956:  I.~Navarro and K.~Van Acoleyen,
2957:  % f(R) actions, cosmic acceleration and local tests of gravity"
2958:   arXiv:gr-qc/0611127
2959: 
2960: %\cite{Hu:2007nk}
2961: \bibitem{Hu:2007nk}
2962:   W.~Hu and I.~Sawicki,
2963:   %``Models of f(R) Cosmic Acceleration that Evade Solar-System Tests,''
2964:   arXiv:0705.1158 [astro-ph].
2965:   %%CITATION = ARXIV:0705.1158;%%
2966: 
2967: 
2968:   %\cite{Dolgov:2003px}
2969: \bibitem{Dolgov:2003px}
2970:   A.~D.~Dolgov and M.~Kawasaki,
2971:   % "Can modified gravity explain accelerated cosmic expansion?",
2972:   Phys. Lett. B {\bf 573}, 1 (2003)
2973:   [arxiv:astro-ph/0307285].
2974: 
2975:   %\cite{Seifert:2007fr}
2976: \bibitem{Seifert:2007fr}
2977:   M.~D.~Seifert,
2978:   % "Stability of spherically symmetric solutions in modified theories of gravity"
2979:   arXiv:gr-qc/0703060.
2980: 
2981:   %\cite{Sawicki:2007tf}
2982: \bibitem{Sawicki:2007tf}
2983:   I.~Sawicki and W.~Hu,
2984:   %``Stability of cosmological solution in f(R) models of gravity,''
2985:   arXiv:astro-ph/0702278.
2986:   %%CITATION = ASTRO-PH/0702278;%%
2987: 
2988: \bibitem{Gunn} Gunn, J.~E., Lee, B.~W.,
2989: Lerche, I., Schramm, D.~N., \& Steigman, G.\ 1978, Ap.\ J.\ {\bf 223},
2990: 1015 (1978).
2991: 
2992: \bibitem{Peebles} P.J.E. Peebles, {\it The Large Scale Structure of
2993: the Universe}, Princeton University Press, Princeton, New Jersey,
2994: 1980.
2995: 
2996: %\cite{Brax:2005ew}
2997: \bibitem{Brax:2005ew}
2998:   P.~Brax, C.~van de Bruck, A.~C.~Davis and A.~M.~Green,
2999:   %``Small Scale Structure Formation in Chameleon Cosmology,''
3000:   Phys.\ Lett.\  B {\bf 633}, 441 (2006)
3001:   [arXiv:astro-ph/0509878].
3002:   %%CITATION = PHLTA,B633,441;%%
3003: 
3004: \bibitem{paperIII} R.\ Bean, \'E.\ \'E.\ Flanagan, M. Trodden, {\it
3005:     Constraints on coupled dark matter-dark energy models}, in
3006:   preparation.
3007: 
3008: %\cite{Lewis:2002ah}
3009: \bibitem{Lewis:2002ah}
3010:   A.~Lewis and S.~Bridle,
3011:   %``Cosmological parameters from CMB and other data: a Monte-Carlo approach,''
3012:   Phys.\ Rev.\  D {\bf 66}, 103511 (2002)
3013:   [arXiv:astro-ph/0205436].
3014:   %%CITATION = PHRVA,D66,103511;%%
3015: 
3016: \bibitem{Brax:2004qh}
3017: P.\ Brax, C.\ van de Bruck, A.-C.\ Davis, J. Khoury and A. Weltman,
3018: astro-ph/0408415.
3019: 
3020: \bibitem{Brax:2004px}
3021: P.\ Brax, C.\ van de Bruck, A.-C.\ Davis, J. Khoury and A. Weltman,
3022: astro-ph/0410103.
3023: 
3024: \bibitem{Landau} L.~D.~Landau, J. Phys. USSR {\bf 10}, 25 (1946).
3025: 
3026: \bibitem{Dicke}
3027: R.\ H.\ Dicke,
3028: %{\it Mach's principle and invariance under transformation of units},
3029: Phys. Rev. {\bf 126}, 2163 (1961).
3030: 
3031: 
3032: \bibitem{Duff}
3033: M.\ J.\ Duff, {\it Comment on time variation of fundamental
3034: constants}, hep-th/0208093.
3035: 
3036: \bibitem{Amendola:2004a}
3037: L.\ Amendola, Phys. Rev. D {\bf 69}, 103524 (2004).
3038: 
3039: 
3040: 
3041: 
3042: 
3043: \end{thebibliography}
3044: 
3045: 
3046: \end{document}
3047: