hep-th0205111/8.tex
1: %version modified by FP, 6 May
2: \documentstyle[eqsecnum,epsfig,aps,floats,preprint,delarray]{revtex}
3:  
4: \def\build#1_#2^#3{\mathrel{
5: \mathop{\kern 0pt#1}\limits_{#2}^{#3}}}
6:        
7: \def\baselinestretch{1.2}
8: \setlength{\oddsidemargin}{0.0cm}
9: \setlength{\textwidth}{16.5cm}
10: \setlength{\topmargin}{-.9cm}
11: \setlength{\textheight}{22.5cm}%
12: 
13: \newcommand{\be}{\begin{equation}}
14: \newcommand{\ee}{\end{equation}}
15: \newcommand{\bea}{\begin{eqnarray}}
16: \newcommand{\eea}{\end{eqnarray}}
17: 
18: %minore o circa uguale
19: \def\laq{~\raise 0.4ex\hbox{$<$}\kern -0.8em\lower 0.62
20: 
21: ex\hbox{$\sim$}~}
22: %maggiore o circa uguale
23: 
24: \def\gaq{~\raise 0.4ex\hbox{$>$}\kern -0.7em\lower 0.62
25: ex\hbox{$\sim$}~}
26: 
27: 
28: \def \pa {\partial}
29: \def \ra {\rightarrow}
30: 
31: \def \la {\lambda}
32: \def \La {\Lambda}
33: \def \Da {\Delta}
34: \def \b {\beta}
35: \def \a {\alpha}
36: \def \ap {\alpha^{\prime}}
37: \def \Ga {\Gamma}
38: \def \ga {\gamma}
39: \def \sg {\sigma}
40: \def \da {\delta}
41: \def \ep {\epsilon}
42: \def \r {\rho}
43: \def \om {\omega}
44: \def \Om {\Omega}
45: \def \noi {\noindent}
46: \def \ti {\tilde}
47: 
48: \def \wt {\widetilde}
49: \def \ep {\epsilon}
50: \def \fb {\overline \phi}
51: \def \rb {\overline \rho}
52: \def \sgb {\overline \sg}
53: \def \pb {\overline p}
54: \def \fbp {\dot{\fb}}
55: \def \va {\varphi}
56: 
57: 
58: \newcommand{\epsi}{e^\psi}
59: \newcommand{\rhm}{\rho_m}
60: \newcommand{\wtm}{\widetilde m_P}
61: \newcommand{\whh}{\widehat H}
62: 
63: \parskip 0.2cm
64: 
65: \begin{document}
66: \par
67: \begingroup
68: 
69: \begin{flushright}
70: IHES/P/02/09 \\ 
71: Bicocca-FT-02-03 \\
72: CERN-TH/2002-093
73: \end{flushright}
74: 
75: 
76: {\large\bf\centering\ignorespaces
77: Violations of the equivalence principle in a dilaton-runaway scenario
78: \vskip2.5pt}
79: {\dimen0=-\prevdepth \advance\dimen0 by23pt
80: \nointerlineskip \rm\centering
81: \vrule height\dimen0 width0pt\relax\ignorespaces
82: T. Damour$^{1}$, F. Piazza$^{2}$ and G. Veneziano$^{3,4}$
83: \par}
84: %\bigskip
85: {\small\it\centering\ignorespaces
86: ${}^{1}$
87: Institut des Hautes Etudes Scientifiques, 91440 
88: Bures-sur-Yvette, France \\
89: %\bigskip
90: ${}^{2}$
91: Dipartimento di Fisica, Universit\`a di Milano Bicocca \\
92: Piazza delle Scienze 3, I-20126 Milan, Italy \\
93: %\bigskip
94: ${}^{3}$
95: Theory Division, CERN, CH-1211 Geneva 23, Switzerland\\
96: %\bigskip
97: ${}^{4}$
98: Laboratoire de Physique Th\'eorique, Universit\'e Paris
99: Sud, 91405 Orsay, France\\
100: \par}
101: 
102: \par
103: \bgroup
104: \leftskip=0.10753\textwidth \rightskip\leftskip
105: \dimen0=-\prevdepth \advance\dimen0 by17.5pt \nointerlineskip
106: \small\vrule width 0pt height\dimen0 \relax
107: 
108: \begin{abstract}
109: We explore a version of the cosmological dilaton-fixing and decoupling mechanism  
110: in which  the dilaton-dependence of the low-energy effective action is extremized
111: for infinitely large values of the bare string coupling 
112: $g_s^2 = e^{\phi}$. We study the efficiency with which the dilaton $\phi$ runs away 
113: towards its ``fixed point'' at infinity during a primordial inflationary stage, and 
114: thereby approximately decouples from matter. The residual dilaton couplings are found 
115: to be related to the amplitude of the density fluctuations generated during inflation. 
116: For the simplest inflationary potential, $V (\chi) = \frac{1}{2} \, m_{\chi}^2 (\phi) 
117: \, \chi^2$, the residual dilaton couplings are shown to predict violations of the 
118: universality of gravitational acceleration near the $\Delta a / a \sim 10^{-12}$ level. 
119: This suggests that a modest improvement in the precision of equivalence principle tests 
120: might be able to detect the effect of such a runaway dilaton. Under some assumptions
121: about the coupling of the dilaton to dark matter and/or dark energy, the expected
122: time-variation of natural ``constants'' (in particular of the fine-structure
123: constant) might also be large enough to be within reach of improved experimental 
124: or observational data.
125: \end{abstract}
126: 
127:  \par\egroup
128: %\vskip2pc]
129: \thispagestyle{plain}
130: \endgroup
131: 
132: \pacs{PACS numbers: 11.25.-w,  04.80.Cc,  98.80.Cq }
133: 
134: \section{Introduction}\label{sec1}
135: 
136: All string theory models predict the existence of a scalar partner of the spin 2 
137: graviton: the dilaton $\phi$, whose vacuum expectation value (VEV) determines the 
138: string coupling constant $g_s = e^{\phi / 2}$ \cite{Witten}. At tree level,
139:  the dilaton is massless 
140: and has gravitational-strength couplings to matter which violate the equivalence 
141: principle \cite{TV88}. This is in violent conflict with present experimental tests of 
142: general relativity. It is generally assumed that this conflict is avoided because, 
143: after supersymmetry breaking, the dilaton might acquire a (large enough) mass (say 
144: $m_{\phi} \gtrsim 10^{-3} \, {\rm eV}$ so that observable deviations from Einstein's 
145: gravity are quenched on distances larger than a fraction of a millimeter). However, 
146: Ref.~\cite{DP94} (see also \cite{DN93}) has proposed a mechanism which can naturally 
147: reconcile a {\it massless} dilaton with existing experimental data. The basic idea of 
148: Ref.~\cite{DP94} was to exploit the string-loop modifications of the (four dimensional) 
149: effective low-energy action (we use the signature $-+++$)
150: \be
151: \label{eq1.1}
152: S = \int d^4 x \sqrt{\widetilde{g}} \left( \frac{B_g (\phi)}{\alpha'} \, \widetilde R + 
153: \frac{B_{\phi} (\phi)}{\alpha'} \, \lbrack 2 \widetilde{\build\Box_{}^{}} \phi -
154: (\widetilde{\nabla} \phi)^2 \rbrack - \frac{1}{4} \, B_F 
155: (\phi) \, \widetilde{F}^2 - V + \cdots \right) \, ,
156: \ee
157: i.e. the $\phi$-dependence of the various coefficients $B_i (\phi)$, $i = g , \phi , F, 
158: \ldots$ , given in the weak-coupling region ($e^{\phi} \to 0$) by series of the form
159: \be
160: \label{eq1.2}
161: B_i (\phi) = e^{-\phi} + c_0^{(i)} + c_1^{(i)} \, e^{\phi} + c_2^{(i)} \, e^{2\phi} + 
162: \cdots \, ,
163: \ee
164: coming from genus expansion of string theory: $B_i = \Sigma_n \, g_s^{2(n-1)} c_n^{(i)}$, 
165: with $n = 0,1,2,\ldots$. It was shown in \cite{DP94} that, if there exists a special 
166: value $\phi_m$ of $\phi$ which extremizes all the (relevant) coupling functions 
167: $B_i^{-1} (\phi)$, the cosmological evolution of the graviton-dilaton-matter system 
168: naturally drives $\phi$ towards $\phi_m$. This provides a mechanism for fixing a 
169: massless dilaton at a value where it decouples from matter (``Least Coupling 
170: Principle''). A simple situation where the existence of a universally extremizing 
171: dilaton value $\phi_m$ is guaranteed is that of $S$ duality, i.e. a symmetry $g_s 
172: \leftrightarrow 1/g_s$, or $\phi \rightarrow -\phi$ (so that $\phi_m = 0$).
173: 
174: It has been recently suggested \cite{V01} that the infinite-bare-coupling limit $g_s 
175: \rightarrow \infty$ ($\phi \rightarrow +\infty$) might yield smooth {\it finite} limits 
176: for all the coupling functions, namely
177: \be
178: \label{eq1.3}
179: B_i (\phi) = C_i + {\cal O} (e^{-\phi}) \, .
180: \ee
181: Under this assumption, the coupling functions are all extremized at infinity, i.e. 
182: $\phi_m = + \infty$. The late-time cosmology of models satisfying (\ref{eq1.3}) has 
183: been recently explored \cite{GPV}. In the ``large N''-type toy model of \cite{V01}
184: it would be natural to expect that the ${\cal O} (e^{-\phi})$ term in Eq.~(\ref{eq1.3})
185: be {\it positive}, so that $B_i (\phi)$ be {\it minimized} at infinity. This would
186: correspond to couplings  $\lambda_i (\phi) \sim  B_i^{-1} (\phi) 
187: = C_i^{-1} - {\cal O} (e^{-\phi})$ which are {\it maximized} at infinity. Note, however,
188:  that the most relevant cosmological
189: coupling for this work, the coupling  to the inflaton, $\lambda(\phi)$,
190:  contained in $V$ (see
191:  Eq. (\ref{eq2.10}) below) is closer to a $B_i$ than to its inverse. Thus $\lambda(\phi)$
192: is naturally {\it minimized} at infinity (see further discussion of this point below), a
193:  crucial property for the attractor mechanism of \cite{DP94,DN93}.
194: 
195: 
196: In this paper\footnote{The main results of this work have been recently
197: summarized in a short note \cite{prl}.} we shall consider in detail the early-time cosmology of models satisfying 
198: (\ref{eq1.3}). More precisely, our main aims will be: (i) to study the efficiency with 
199: which a primordial inflationary stage drives $\phi$ towards the ``fixed point'' at 
200: infinity $\phi_m = + \infty$ (thereby generalizing the work \cite{DV96} which 
201: considered the inflationary attraction towards a local extremum $\phi_m$), and (ii) to 
202: give quantitative estimates of the present violations of the equivalence principle (non 
203: universality of free fall, and variation of ``constants''). Our most important 
204: conclusion is that the runaway of the dilaton towards strong-coupling (under the 
205: assumption (\ref{eq1.3})) naturally leads to equivalence-principle violations which are 
206: rather large, in the sense of not being much smaller than the presently tested 
207: level $\sim 10^{-12}$. This gives additional motivation for the currently planned 
208: improved tests of the universality of free fall. Within our scenario, most of the other 
209: deviations from general relativity (``post-Einsteinian'' effects in gravitationally 
210: interacting systems: solar system, binary pulsars, ...) are too small to be of 
211: phenomenological interest. However, under some assumptions about the coupling of $\phi$
212: to dark matter and/or dark energy, the time variation of the natural ``constants'' 
213: (notably the fine-structure constant) predicted by our scenario might be large enough
214: to be within reach of improved experimental and/or observational data. The 
215: phenomenologically interesting conclusion that equivalence-principle violations are 
216: generically predicted to be rather large after inflation (in sharp contrast with the 
217: results of \cite{DV96}) is due to the fact that the attraction towards an extremum at 
218: infinity is much less effective than the attraction towards a (finite) local extremum 
219: as originally contemplated in \cite{DP94}. This reduced effectiveness was already 
220: pointed out in Ref.~\cite{DN93} within the context of equivalence-principle-respecting 
221: tensor-scalar theories (\`a la Jordan-Fierz-Brans-Dicke).
222: 
223: \section{Dilaton runaway}
224: 
225: In this section we study the dilaton's runaway 
226: during the various stages of 
227: cosmological evolution. We first show (subsection
228: \ref{sec2}) 
229: that, like in the case of a local extremum \cite{DV96}, inflation is 
230: particularly efficient in pushing $\phi$ towards the fixed point.
231: We will then argue (subsection
232: \ref{sec2b}) that the order of magnitude of the bare string coupling
233:  $e^{\phi} \simeq e^{c\varphi}$ does not suffer further appreciable changes 
234:  during all the subsequent evolution. 
235: 
236: 
237: \subsection{The inflationary period}\label{sec2}
238: 
239: Assuming some primordial inflationary stage driven by the potential energy of an
240: inflaton 
241: field $\widetilde{\chi}$, and taking into account generic couplings to the dilaton 
242: $\phi$, we consider an effective action of the form
243: \be 
244: \label{eq2.1}
245: S = \int d^4 x \sqrt{\widetilde g} \left( \frac{B_g (\phi)}{\alpha'} \, \widetilde{R} + 
246: \frac{B_{\phi} (\phi)}{\alpha'} \, \lbrack 2 \widetilde{\build\Box_{}^{}} \phi -
247: (\widetilde{\nabla} \phi)^2 \rbrack
248: - \frac{1}{2} \, 
249: B_{\chi} (\phi) (\widetilde{\nabla} \widetilde{\chi})^2 - \widetilde V 
250: (\widetilde{\chi} , \phi) \right) \, .
251: \ee
252: In this string-frame action, the dilaton dependence of all the functions $B_i (\phi)$, 
253: $\widetilde V (\widetilde{\chi} , \phi)$ is assumed to be of the form (\ref{eq1.2}). It 
254: is convenient to replace the ($\sigma$-model) string metric $\widetilde{g}_{\mu \nu}$ 
255: by the conformally related Einstein metric $g_{\mu \nu} = C \, B_g (\phi) \, 
256: \widetilde{g}_{\mu \nu}$, and the dilaton field by the variable
257: \be
258: \label{eq2.2}
259: \varphi = \int d \phi \left[ \frac{3}{4} \left( \frac{B'_g}{B_g} \right)^2 + 
260: \frac{B'_{\phi}}{B_g} + \frac{1}{2} \, \frac{B_{\phi}}{B_g} \right]^{\frac{1}{2}} \, , \;
261: B' \equiv \partial B/\partial \phi \; .
262: \ee
263: The normalization constant $C$ is chosen so that the string units coincide with the 
264: Einstein units when $\phi \rightarrow + \infty$: $C \, B_g (+\infty) = 1$. [Note that 
265: $C = 1/C_g$ in terms of the general notation of Eq.~(\ref{eq1.3}).] Introducing the 
266: (modified) Planck mass
267: \be
268: \label{eq2.3}
269: \widetilde{m}_P^2 = \frac{1}{4\pi G} = \frac{4}{C \alpha'}\, ,
270: \ee
271: and replacing also the inflaton by the dimensionless variable $\chi = C^{-1/2} \, 
272: \widetilde{m}_P^{-1} \, \widetilde{\chi}$, we end up with an action of the form
273: \be
274: \label{eq2.4}
275: S = \int d^4 x \sqrt{g} \left[ \frac{\widetilde{m}_P^2}{4} \, R - 
276: \frac{\widetilde{m}_P^2}{2} \, (\nabla \varphi)^2 - \frac{\widetilde{m}_P^2}{2} \, F 
277: (\varphi) (\nabla \chi)^2 - \widetilde{m}_P^4 \, V (\chi , \varphi) \right] \, ,
278: \ee
279: where
280: \be
281: \label{eq2.5}
282: F(\varphi) = B_{\chi} (\phi) / B_g (\phi) \, , \ V(\chi , \varphi) = C^{-2} \, 
283: \widetilde{m}_P^{-4} \, B_g^{-2} (\phi) \, \widetilde V (\widetilde{\chi} , \phi) \, .
284: \ee
285: In view of our basic assumption (\ref{eq1.3}), note that, in the strong-coupling limit 
286: $\phi \rightarrow +\infty$, $d\varphi / d \phi$ tends, according to Eq.~(\ref{eq2.2}), 
287: to the constant $(C_{\phi} / 2 C_g)^{1/2}$, while the dilaton-dependent factor 
288: $F(\varphi)$ in front of the inflaton kinetic term tends to the constant $C_{\chi} / 
289: C_g$. The toy model of Ref.~\cite{V01} suggests that the various (positive) constants $C_i$ in 
290: Eq.~(\ref{eq1.3}) are all largish and comparable to each other. We shall therefore 
291: assume that the various ratios $C_i / C_j$ are of order unity. The most important such 
292: ratio for the following is $c \equiv (2 C_g / C_{\phi})^{1/2}$ which gives the 
293: asymptotic behaviour of the bare string coupling as
294: \be
295: \label{eq2.6}
296: g_s^2 = e^{\phi} \simeq e^{c\varphi} \, .
297: \ee
298: In view of the fact that, in the strong-coupling limit we are interested in, the factor 
299: $F(\varphi)$ in Eq.~(\ref{eq2.4}) quickly tends to a constant, we can simplify our 
300: analysis (without modifying the essential physics) by replacing it by a constant (which 
301: can then be absorbed in a redefinition of $\chi$). Henceforth, we shall simply take 
302: $F(\varphi) = 1$. [See, however, the comments below concerning the self-regenerating
303: inflationary regime.]
304: 
305: Following \cite{DN93,D95} it is then useful to combine the Friedmann equations for the 
306: scale factor $a(t)$ during inflation ($ds^2 = -dt^2 + a^2 (t) \, \delta_{ij} \, dx^i \, 
307: dx^j$) with the equations of motion of the two scalar fields $\chi (t)$, $\varphi (t)$, 
308: to write an autonomous equation describing the evolution of the two scalars in terms of 
309: the parameter
310: \be
311: \label{eq2.7}
312: p = \int H dt = \int \frac{\dot a}{a} \, dt = \ln \, a + {\rm const}
313: \ee
314: measuring the number of $e$-folds of the expansion. For any multiplet of scalar fields, 
315: $\mbox{\boldmath$\varphi$} = (\varphi^a)$, this yields the simple equation 
316: \cite{DN93,D95}
317: \be
318: \label{eq2.8}
319: \frac{2}{3 - \mbox{\boldmath$\varphi$}'^2} \, \mbox{\boldmath$\varphi$}'' + 2 
320: \, \mbox{\boldmath$\varphi$}' = - \nabla_{\mbox{\boldmath$\varphi$}} \ln \vert V 
321: (\mbox{\boldmath$\varphi$}) \vert \, ,
322: \ee
323: where $\mbox{\boldmath$\varphi$}' \equiv d \mbox{\boldmath$\varphi$} / dp$, and where 
324: all operations on $\mbox{\boldmath$\varphi$}$ are covariantly defined in terms of the 
325: $\sigma$-model metric defining the scalar kinetic terms ($d\sigma^2 = \gamma_{ab} 
326: (\varphi) \, d\varphi^a \, d\varphi^b$). In our simple model (with $F(\varphi) = 1$), 
327: we have a flat metric $d\sigma^2 = d \varphi^2 + d\chi^2$. [Note that, when 
328: $\gamma_{ab} (\varphi)$ is curved the acceleration term $\mbox{\boldmath$\varphi$}''$ 
329: involves a covariant derivative.]
330: 
331: The generic solution of Eq.~(\ref{eq2.8}) is easily grasped if one interprets it as a 
332: mechanical model: a particle with position $\mbox{\boldmath$\varphi$}$, and 
333: velocity-dependent mass $m(\mbox{\boldmath$\varphi$}') = 
334: 2/(3-\mbox{\boldmath$\varphi$}'^2)$, moves, in the ``time'' $p = \ln a + {\rm cst}$, in 
335: the manifold $d \sigma^2$ under the influence of an external potential $\ln \vert V 
336: (\mbox{\boldmath$\varphi$})\vert$ and a constant friction force $-2 \, 
337: \mbox{\boldmath$\varphi$}'$. If the curvature of the effective potential $\ln \vert V 
338: (\mbox{\boldmath$\varphi$})\vert$ is sufficiently small the motion of 
339: $\mbox{\boldmath$\varphi$}$ rapidly becomes slow and friction-dominated:
340: \be
341: \label{eq2.9}
342: 2 \, \frac{d\mbox{\boldmath$\varphi$}}{dp} \simeq - \nabla_{\mbox{\boldmath$\varphi$}} 
343: \ln V (\mbox{\boldmath$\varphi$}) \, .
344: \ee
345: Eq.~(\ref{eq2.9}) is equivalent to the usual ``slow roll'' approximation.
346: 
347:  Consistently with our general assumption (\ref{eq1.3}), we consider potentials allowing
348: a strong-coupling expansion of the form:
349: \be
350: \label{genpot}
351: V (\chi , \varphi) = V_0(\chi) + V_1(\chi) e^{-c\varphi} + {\cal O}(e^{-2c\varphi}) \, ,
352: \ee
353: where $V_0(\chi)$ is a typical chaotic-inflation potential with 
354: $V_0(0)=0$, while $V_1(0) = v_1 \ge 0$ can possibly provide (if $v_1 >0$)
355:  the effective cosmological constant
356: driving   today's acceleration in the scenario of \cite{GPV}.
357: For the sake of simplicity we shall discuss mainly the ``factorized'' 
358: power-law case $V_0(\chi) \sim V_1(\chi) \sim \chi^n$ for which we can
359: conveniently write $V$ in the form:
360: \be
361: \label{eq2.10}
362: V (\chi , \varphi) = \lambda (\varphi) \, \frac{\chi^n}{n} \, ,
363: \ee
364:  with a dilaton-dependent coupling constant $\lambda (\varphi)$  of the form
365: \be
366: \label{eq2.11}
367: \lambda (\varphi) = \lambda_{\infty} (1 + b_{\lambda} \, e^{-c\varphi}) \, .
368: \ee
369: This example belongs to the class of the
370:  two-field inflationary potentials discussed in \cite{L90}. We have checked that
371: our results remain qualitatively the same for the more general potential (\ref{genpot})
372: provided that  $V_0(\chi)$ and $V_1(\chi)$ are not extremely diffent and given
373: the fact that $v_1$ is phenomenologically constrained to be very small.
374: [Note that, within the simplified model (\ref{eq2.11}), the ratio  $V_1(\chi)/V_0(\chi)$
375: is equal to the constant coefficient  $b_{\lambda}$.]
376: 
377: The universal (positive) constant $c$ appearing in the exponential $e^{-c\varphi}$
378:   is the same as in 
379: Eq.~(\ref{eq2.6}) [i.e. $c \equiv (2 C_g / C_{\phi})^{1/2}$, which is expected to be of 
380: order unity]. The coefficient $b_{\lambda}$ in (\ref{eq2.11}) is such that $b_{\lambda} 
381: \, e^{-c \varphi} \simeq b_{\lambda} \, e^{-\phi}$ roughly corresponds to a combination 
382: of terms $\sim \pm \, C_i^{-1} \, {\cal O} (e^{-\phi})$ coming from the strong-coupling 
383: asymptotics of several $B_i (\phi)$, Eq.~(\ref{eq1.3}) (see Eq.~(\ref{eq2.5})). In the 
384: toy model of \cite{V01} one would therefore expect $b_{\lambda}$ to be smallish. 
385: Anyway, we shall see that in final results only the ratios of such $b_i$ coefficients 
386: enter. More important than the magnitude of $b_{\lambda}$ is its sign. It is crucial 
387: for the present strong-coupling attractor scenario to assume that $b_{\lambda} > 0$, 
388: i.e that $\lambda (\varphi)$ reaches a {\it minimum} at strong-coupling, $\varphi 
389: \rightarrow + \infty$. Note again that this behaviour is consistent with the
390: simple ``large $N$''-type idea of \cite{V01} if we assimilate $\lambda (\varphi)$ to one
391: of the inverse couplings $B_i$ appearing in (\ref{eq1.1})
392:  (for instance $B_F \sim g_F^{-2}$, where $g_F$ is a 
393: gauge coupling), rather than to
394:  the coupling itself.
395: If the latter were the case, $\lambda (\varphi)$ would reach 
396: a {\it maximum} as $\phi \rightarrow + \infty$, and the attractor mechanism of 
397: \cite{DP94} would drive $\phi$ towards weak coupling ($\phi \rightarrow - \infty$).
398: However, the Einstein-frame $\phi$-dependence of $V(\chi)$ gets contributions from 
399: several $B_i^{\pm n}(\phi)$, Eq.~(\ref{eq2.5}), which might conspire to minimize it
400: at strong coupling. This feature is also probably necessary in order to solve
401: the cosmological-constant problem through some argument by which the vacuum at infinity
402: has vanishing energy density.
403: 
404: Substituting the potential (\ref{genpot}) into the slow roll equation (\ref{eq2.9})
405: and assuming (for simplicity) that $ V_1(\chi) e^{-c\varphi}$ is significantly smaller 
406: than $V_0(\chi)$ 
407: leads to a decoupled set of evolution equations for $\chi$ and $\varphi$ (where 
408: $V' \equiv \partial V / \partial \chi$):
409: \be
410: \label{eq2.12}
411: \frac{d \chi}{dp} = - \frac{1}{2} \, \frac{V'_0}{V_0} \, ,
412: \ee
413: \be
414: \label{eq2.13}
415: \frac{d \varphi}{dp} =  \frac{1}{2}  \, c \, e^{-c \varphi} \frac{V_1}{V_0} \, .
416: \ee
417: Given some ``initial'' conditions $\chi_{\rm in}$, $\varphi_{\rm in}$ (discussed below)
418: at some starting 
419: point, say $p=0$, the solution of Eqs.~(\ref{eq2.12}), (\ref{eq2.13}) is simply
420: 
421: \be
422: \label{eq2.14}
423: p =  \, 2 \int_{\chi}^{\chi_{\rm in}} d\bar\chi \,
424: \bar\chi \left( \frac{V_0(\bar\chi)}{\bar\chi \, V'_0(\bar\chi)}\right) \, ,
425: \ee
426: \be
427: \label{eq2.15}
428: e^{c \varphi} = e^{c \varphi_{\rm in}} +   \, \frac{c^2}{2} \, 
429: \int dp \, \frac{V_1(\chi (p))}{V_0(\chi (p))} \, ,
430: \ee
431: which simply become:
432: \be
433: \label{eq2.16}
434: p =  \frac{1}{n} \left( \chi_{\rm in}^2 - \chi^2 \right)  \, , \quad 
435: e^{c \varphi} + \frac{b_{\lambda} \, c^2}{2n} \, \chi^2 = {\rm const}. = e^{c 
436: \varphi_{\rm in}} + \frac{b_{\lambda} \, c^2}{2n} \, \chi_{\rm in}^2 \, ,
437: \ee
438: in the simplified case of eqs. (\ref{eq2.10}), (\ref{eq2.11}).
439: 
440: Equations (\ref{eq2.16}) show that, in order for the string coupling 
441:  $g_s^2 \simeq e^{c \varphi}$ to have reached large values at the end of inflation,
442: a large total number of e-folds must have occurred while the
443: (dimensionless) inflaton field $\chi$ decreases from a large initial value, to a
444: value of order unity (in Planck units). To get a quantitative estimate of the string coupling
445: at the end of inflation we need to choose the initial conditions 
446: $\chi_{\rm in}$, $\varphi_{\rm in}$. A physically reasonable way (which is further
447: discussed below) of choosing $\chi_{\rm in}$
448: is to start the classical evolution (\ref{eq2.12})-(\ref{eq2.16}) at the exit of the
449: era of self-regenerating inflation (see \cite{L90} and references therein).
450: We will now show how to relate the exit from self-regenerating inflation 
451:  to the size of density fluctuations generated by inflation.
452: 
453: 
454:  Let us recall (see \cite{L90} and references 
455: therein) that the density fluctuation $\delta \equiv \delta \rho / \rho$ on large scales 
456: (estimated in the one-field approximation where the inflaton $\chi$ is the main 
457: contributor) is obtained by evaluating the expression
458: \be
459: \label{eq2.19}
460: \delta (\chi) \simeq \frac{4}{3} \, \frac{1}{\pi} \left( \frac{2}{3} 
461: \right)^{\frac{1}{2}} \, \frac{V^{3/2}}{\partial_\chi V}
462: \ee
463: at the value $\chi = \chi_{\times}$, at which the physical scale
464: we are considering crossed the horizon outwards during inflation. For the scale
465: corresponding to  our present
466: horizon this usually corresponds to  a value  $\chi_{\times}(H_0)$  ($\chi_H$ for short)  
467: reached some $ 60$ $e$-folds before 
468: the end of slow-roll. From Ref.~\cite{L90}, $\chi_H \simeq 5 \sqrt n$ for the model 
469: (\ref{eq2.10}) (and with our modified definition of $\chi$). 
470: The numerical value of $\delta_H \equiv \delta (\chi_H)$ which 
471: is compatible with cosmological data (structure 
472: formation and cosmic microwave background) is $\delta_H \simeq 5 \times 10^{-5}$.
473:  In the model 
474: (\ref{eq2.10}) the function $\delta (\chi)$ defined by (\ref{eq2.19}) scales with 
475: $\chi$ as $\chi^{\frac{n+2}{2}}$. Putting together this information we obtain a
476: relation between  $\chi_{\rm in}$ and $\delta(\chi_{\rm in})$, which involves
477: the value of the observable horizon-size fluctuations 
478: $\delta_H \equiv \delta (\chi_H)$:
479: \be
480: \label{eq2.21}
481: \frac{\delta (\chi_{\rm in})}{\delta (\chi_H)} = \left( \frac{\chi_{\rm in}}{\chi_H} 
482: \right)^{\frac{n+2}{2}} ,
483: \ee
484: i.e.
485: \be
486: \label{eq2.22}
487: \chi_{\rm in} \simeq \chi_H \, \left(\frac{\delta_{\rm in}}{\delta_H} \right)^{ \frac{2}{n+2}} 
488: \simeq \, 5 \sqrt n \, 
489: \left(\frac{\delta_{\rm in}}{\delta_H} \right)^{ \frac{2}{n+2}} \, ,
490: \ee
491: where we introduced the short-hand notation 
492: $\delta_{\rm in} \equiv \delta (\chi_{\rm in})$.
493: 
494: 
495: Inserting Eq.~(\ref{eq2.22}) into Eq.~(\ref{eq2.15}) we then obtain the following 
496: estimate of the string coupling constant after inflation as a function of 
497: $\varphi_{\rm in}$ and $\delta(\chi_{\rm in})$
498: \be
499: \label{eq2.23}
500: e^{c \varphi_{\rm end}} -  e^{c \varphi_{\rm in}} \, \simeq \, \frac{c^2}{2}
501: \langle V_1/V_0 \rangle  p \, \sim \, \frac{c^2}{2n} \langle V_1/V_0 \rangle  
502: \chi_{\rm in}^2 \, \sim \, 
503: \frac{25 c^2}{2} \, \langle V_1/V_0 \rangle  \, 
504: \left(\frac{\delta_{\rm in}}{\delta_H} \right)^{ \frac{4}{n+2}} \, ,
505: \ee
506: where $\langle V_1/V_0 \rangle $ denotes the average value of $V_1/V_0$: $\langle V_1/V_0 \rangle  \equiv
507:  \int dp (V_1/V_0)/\int dp $ [note that this average ratio is equal to $b_{\lambda}$
508:  in the simplified model (\ref{eq2.11})].
509:  
510: To get a quantitative estimate of $e^{c \varphi_{\rm end}}$ we still need to 
511: estimate the value of $\delta (\chi_{\rm in})$ corresponding to the chosen 
512: ``initial'' value of the inflaton.  As we will now check,
513: taking for $\chi_{\rm in}$ the value corresponding to the exit from self-regenerating 
514: inflation corresponds simply to taking $\delta (\chi_{\rm in}) \sim 1$. Indeed, let us
515: first recall that, during inflation, each (canonically normalized) scalar field
516: (of mass smaller than the expansion rate $H$) undergoes typical quantum fluctuations
517: of order $H/(2 \pi)$, per Hubble time (see, e.g., \cite{L90}).
518:  This implies (for our dimensionless fields) that the value of $\chi$ at the exit
519:  from self-regeneration, say $\chi_{\rm ex}$, is characterized by
520:  ${\widehat H}_{\rm ex}/(2 \pi) \approx \lbrack \partial_\chi V/(2 V) \rbrack_{\rm ex}$,
521:  where ${\widehat H} \equiv H / \widetilde{m}_P $ is the dimensionless Hubble expansion 
522: rate and where the right-hand side (RHS) is the classical change of $\chi$ per Hubble time
523:  (corresponding to the RHS of Eq.(\ref{eq2.12})). Using Friedmann's equation 
524:  (in the slow-roll approximation) ${\widehat H}^2_{\rm ex} \approx
525:  (2/3) V(\chi_{\rm ex})$, it is easily seen that that the exit from
526:  self-regeneration corresponds to $\delta (\chi_{\rm ex}) \approx 4/3 \sim 1$.
527:  It is, a posteriori, physically quite reasonable to start using the classical evolution
528: system only when the (formal extrapolation) of the density fluctuation $\delta (\chi)$ 
529: becomes smaller than one.
530: 
531: Within some approximation (see \cite{L90}), one can implement the effect of the 
532: combined quantum fluctuations of $(\varphi, \chi)$ 
533:  by adding  random terms with r.m.s. values 
534: $\widehat H/2\pi$  on the right hand side of equations (\ref{eq2.12}) and (\ref{eq2.13}),
535: $d \chi/dp$ and $d \varphi/dp$ being precisely the shifts of the fields in a Hubble time.
536: The system of equations becomes thus of the Langevin-type 
537: \begin{equation}
538: \label{eq4.1}
539: \frac{d \chi}{d p} \, = 
540: \, - \frac{1}{2} \, \frac{V'_0}{V_0}\,  +\, \frac{\whh}{2\pi}\, \xi_1 \, ,
541: \end{equation}
542: \begin{equation}
543: \label{eq4.2}
544: \frac{d \varphi}{dp} \, = \, \frac{1}{2} \, c \, e^{-c \varphi} \frac{V_1}{V_0}\,  +\, \frac{\whh}{2\pi} \,
545: \xi_2 \, ,
546: \end{equation} 
547: where ${\widehat H} \approx [(2/3) V(\chi,\phi)]^{1/2}$ (in the slow-roll approximation) is the
548: dimensionless expansion rate, and where $\xi_1$ and $\xi_2$ are
549: (independent) normalized random white noises:
550: \be
551: \langle\xi_i(p_1)\, \xi_j(p_2) \rangle \, \, =\, \delta_{ij}\delta(p_1-p_2), \qquad i,j = 1,2 \, .
552: \ee
553: 
554: When the random force terms dominate the evolution
555: in either equation (\ref{eq4.1}) or equation (\ref{eq4.2}) the quasi-classical description 
556: (\ref{eq2.12}) , (\ref{eq2.13}) breaks down. 
557: The phase space of the sytem can thus be roughly divided into four regions 
558: according to whether the evolution of none, one or both of the two fields
559:  is dominated by quantum fluctuations. 
560: This is depicted in Figure 1 where such regions are delimited by dashed, 
561: thick curves in the case of a power-law potential (\ref{eq2.10}). 
562: 
563: \begin{figure}[t] \label{figure1}
564: \begin{center}
565: \includegraphics[height=9cm]{fig2.eps}
566: \vskip 3mm
567: \caption{\sl The phase space of the system is represented in the case of  
568: a power--law potential (\ref{eq2.10}) with $n=2$, $b_\lambda = 0.1$ and 
569: $\lambda_\infty = 10^{-10}$. The 
570: thick--dashed (red) curves delimitate the quantum behaviour of the two fields, 
571: the horizontal curve $\chi = \lambda_\infty^{-1/(n+2)} $  and the 
572: hyperbola-like curve $\chi = b_\lambda^{2/n} \lambda_\infty^{-1/n} e^{-2 c\varphi/n}$
573: being the limit of the quantum behaviour for $\chi$ and $\varphi$ respectively. 
574: In the white region both fields have a classical behaviour. The last 
575: ``fully classical'' trajectory has been represented by a thick (blue) curve.
576: The bright--gray regions are those where either the $\varphi$ or the $\chi$ evolution
577: are dominated by quantum fluctuations. The fully--quantum region 
578: is the dark--gray region on the top right.}
579: \end{center}
580: \end{figure}
581: 
582: Apart from factors of order one, the evolution of the inflaton $\chi$ is quasi-classical
583: in the region under the line $\chi = \lambda_\infty^{-1/(n+2)}$. 
584: In the chaotic inflationary models \cite{L90} 
585: such an inflaton's value corresponds to
586: the exit from the self-regenerating regime and to the beginning of the quasi-classical
587: slow-roll inflation. As mentioned above it 
588: also corresponds to a perturbation $\delta(\chi) \sim 1$.
589: Somewhat surprisingly, in the model at hand, however, the quasi-classical region for the 
590: inflaton evolution 
591: $\chi \lesssim \lambda_\infty^{-1/(n+2)}$ is  affected
592: by quantum fluctuations that still dominate the  evolution of $\varphi$
593: in  the region above the hyperbola-like 
594: curve $\chi = b_\lambda^{2/n} \lambda_\infty^{-1/n} e^{-2 c\varphi/n}$.
595: We must therefore study, in some detail, the evolution of the system 
596: in the presence of the noise term for $\varphi$ as in eq. (\ref{eq4.2}), assuming 
597: a quasi classical evolution for $\chi$. This is done in the appendix. 
598: The final result is that, if we start at $\chi \lesssim \lambda_\infty^{-1/(n+2)}$, 
599: (classical evolution for $\chi$)  
600: the average value of $e^{c\varphi}$ 
601: is {\it multiplicatively renormalized}, by a factor of order unity, with respect to 
602:  the classical trajectory $e^{c\varphi_{\rm cl}}$, given by solving equations
603:  (\ref{eq2.12}) - (\ref{eq2.13}), i.e.
604: $\langle e^{c\varphi} \rangle = {\cal O}(1) e^{c\varphi_{\rm cl}}$. One also finds
605: that the dispersion of $e^{c\varphi}$ around its average value is comparable to its
606: average value.
607: 
608: We shall not try to discuss here what happens in the self-regenerating region
609: $\chi \agt \chi_{\rm in} \sim \lambda_\infty^{-1/(n+2)}$. Let us recall that the simple decoupled
610: system (\ref{eq2.12}) , (\ref{eq2.13}) was obtained by neglecting the kinetic coupling term
611: $F(\varphi)$ in Eq.~(\ref{eq2.4}). If we were to consider a more general model, we would have
612: more coupling between $\chi$ and $\varphi$ and we would expect that (contrary to Fig. 1 which
613: exhibits a ``classical $\varphi$ region'' above the ``quantum $\chi$ line'') the evolution
614: in the self-regenerating region involves a strongly coupled system of Langevin equations.
615: Then, as discussed in \cite{L90}, solving such a system necessitates to give boundary conditions
616: on all the boundaries of the problem: notably for $\chi \to \infty$, but also for 
617: $\varphi \to +\infty$ and $\varphi \to -\infty$. We leave to future work such an
618: investigation (and a discussion of what are reasonable boundary conditions). In this work
619: we shall content ourselves with ``starting'' the evolution on the quantum $\chi$ boundary
620: line $\chi_{\rm in}$ with some value $\varphi = \varphi_{\rm in}$, assuming that
621: $ e^{c\varphi_{\rm in}}$ is smaller than the driving effect due to inflation, i.e. than
622: the RHS of Eq.~(\ref{eq2.23}). [This assumption is most natural in a work aimed at 
623: studying the ``attracting'' effect due to primordial inflation.]
624: 
625: 
626: Going back to our result (\ref{eq2.23}), we can now insert, 
627: according to the preceding discussion, the values $\delta_{\rm in}=1$ and 
628: $e^{c \varphi_{\rm in}} \ll e^{c \varphi_{\rm end}}$.
629: Finally, in the simplified model (\ref{eq2.10}), (\ref{eq2.11}), we get the estimate:
630:  \be
631: \label{eq2.23'}
632: e^{c \varphi_{\rm end}}  = {\cal O}(1) \cdot e^{c\varphi_{\rm cl,\, end}} \sim
633: {\cal O}(1) \cdot \frac{25 c^2}{2} \, b_{\lambda} \, 
634: \left(\delta_H\right)^{ -\frac{4}{n+2}} \, .
635: \ee
636: 
637: A more general analysis, based on the potential (\ref{genpot})
638:  leads to the same final result but with $n$ replaced
639:  by some average value of $\frac{\chi V_{0,\chi}}{V_0}$, and with $b_{\lambda}$ replaced by
640:  some average of the ratio $V_1/V_0$.
641:  Note that 
642: smaller values of the exponent $n$ lead to larger values of $e^{c \varphi_{\rm end}}$, i.e. 
643: to a more effective attraction towards the ``fixed point at infinity''. The same is true
644: if we take different exponents $n_0$ and $n_1$ (for $V_0$ and $V_1$ respectively) 
645: and assume $V_1(\chi_{\rm in})
646:  \gg V_0 (\chi_{\rm in})$ to hold as a result of $\chi_{\rm in} \gg 1$ and 
647: $n_1 > n_0$. Also note that, 
648: numerically, if we consider $n=2$, i.e. the simplest chaotic-inflation potential $V = 
649: \frac{1}{2} \, m_{\chi}^2 (\varphi) \, \chi^2$, Eq.~(\ref{eq2.23'}) involves the large 
650: number $12.5 \times \delta_H^{-1} \sim 2.5 \times 10^5$. In the case where $n=4$, i.e. $V = 
651: \frac{1}{4} \, \lambda (\varphi) \, \chi^4$, we have instead the number $12.5 \times 
652: \delta_H^{-2/3} \sim 0.92 \times 10^4$. To understand the phenomenological meaning of these 
653: numbers we need to relate $e^{c \varphi_{\rm end}}$ to the present, observable deviations 
654: from general relativity. This issue is addressed in Section \ref{sec3} after having argued
655: that the post-inflationary evolution of $\varphi$ is sub-dominant.
656: 
657:  
658: 
659: \subsection{Attraction of $\varphi$ by the subsequent cosmological evolution}\label{sec2b}
660: 
661: We have discussed above the efficiency with which inflation drives the dilaton towards a 
662: fixed point at infinity. We need to complete this discussion by estimating the effect 
663: of the many $e$-folds of expansion that took place between the end of inflation and the 
664: present time. To address this question, we need to study in more detail the 
665: coupling of a runaway dilaton to various types of matter, say a multi-component 
666: distribution of (relativistic or non-relativistic) particles. 
667: We work in the Einstein frame, with an action of the type
668: 
669: \begin{eqnarray}
670: S = \int d^4x \sqrt{g}\left[\frac{\widetilde{m}_P^2}{4}R-\frac{\widetilde{m}_P^2}{2}
671: (\nabla \varphi)^2  - \frac{1}{4} B_F(\varphi) F^2 + \dots \right]  
672: \qquad \qquad\\[2mm]
673:  \qquad   \qquad  \qquad  -  \sum_A \int m_A[\varphi(x_A)] \sqrt{-g_{\mu \nu}(x_A) dx_A^\mu dx_A^\nu}\, . 
674: \nonumber
675: \end{eqnarray}
676: Following \cite{TV88,DP94}, one 
677: introduces the crucial dimensionless quantity
678: \be
679: \label{eq3.1}
680: \alpha_A (\varphi) \equiv \frac{\partial \ln m_A (\varphi)}{\partial \, \varphi} \, ,
681: \ee
682: measuring the coupling of $\varphi$ to a particle of type $A$. [For consistency with previous 
683: work, we keep the notation $\alpha_A$ but warn the reader that this should not be confused 
684: with the various gauge coupling constants, often denoted $\alpha_i = g_i^2 / 4\pi$.] The 
685: quantity $\alpha_A$ determines the effect of cosmological matter on the evolution of $\varphi$ 
686: through the general equation \cite{DN93,DP94}
687: 
688: \be
689: \label{eq3.2}
690: \frac{2}{3 - \varphi'^2} \, \varphi'' + \left( 1 - \frac{P}{\rho} \right) \varphi' = - 
691: \sum_A \alpha_A (\varphi) \, \frac{\rho_A - 3 P_A}{\rho} \, ,
692: \ee
693: where the primes denote derivatives with respect to $ p = \ln a + {\rm const} $ and where
694: $\rho = \Sigma_A \, \rho_A$ and $P = \Sigma_A \, P_A$ 
695: are the total ``material''
696: energy density and pressure respectively, both obtained
697: as sums over the various components  
698: filling the universe \emph{at the exception of} the kinetic energy density and pressure 
699: of $\varphi$, $\rho_k = (\widetilde{m}_P^2 /2) (d\varphi/dt)^2 = 
700: (\widetilde{m}_P^2 /2) H^2 {\varphi'}^2 $ and 
701: % $
702: $P_k = \rho_k$. Accordingly, the Friedmann equation reads 
703: \be \label{friedman}
704: 3 H^2\, =\, \frac{2}{\widetilde{m}_P^2} \, \rho_{\rm tot} \, = \, 
705:  \frac{2 \, \rho}{ \widetilde{m}_P^2} + 
706: H^2 {\varphi'}^2.
707: \ee
708: 
709: Note that $\rho$ and $P$ may also account for the potential energy density and 
710: pressure of the scalar field,  $\rho_V = V(\varphi)$, $ P_V = - \rho_V $ and that one
711: can formally extend Eq. (\ref{eq3.2}) to the ``vacuum
712: energy'' component $V(\varphi)$  by associating to the potential 
713: $V(\varphi)$ the mass scale $m_V (\varphi) \equiv V(\varphi)^{\frac{1}{4}}$ which gives
714: $\alpha_V = \partial \ln m_V (\varphi) / \partial \varphi = \frac{1}{4} \, \partial 
715: \ln V(\varphi) / \partial \, \varphi$. Equation (\ref{eq2.8}) is then recovered in the limit
716: where the scalar field is the dominant component.
717: 
718: In the simple cases (which are quite frequent; at least as approximate cases) 
719: where one ``matter'' component, with known ``equation of state'' 
720: $P_A / \rho_A = w_A = {\rm const.} $,
721: dominates the cosmological density and pressure, Eq. (\ref{eq3.2}) yields an autonomous
722: equation for the evolution (with redshift) of $\varphi$. 
723: Using (\ref{friedman}) one  finds  
724: that the ``equation of 
725: state parameter'' $w_{\rm tot} \equiv P_{\rm tot}/ \rho_{\rm tot}$  corresponding to
726: the {\it total} energy and pressure (including now the kinetic contributions of $\varphi$;
727: i.e. $\rho_{\rm tot} = \rho + (\widetilde{m}_P^2/2) (d\varphi/dt)^2 \, , \, 
728: P_{\rm tot} = P + (\widetilde{m}_P^2/2) \, (d\varphi/dt)^2$) is given 
729: in terms of the ``matter'' equation-of-state parameter $w \equiv P/\rho$ by
730: \be
731: \label{eq3.2bis}
732: w_{\rm tot} = w  + \frac{1 - w}{3} (\varphi')^2 \, . 
733: \ee
734: The knowledge of $w_{\rm tot}$ then allows one to write explicitly the energy-balance equation
735: $ d \rho_{\rm tot} + 3 ( \rho_{\rm tot} + P_{\rm tot}) d \ln a = 0$, which is easily
736: solved in the simple cases where $w_{\rm tot}$ is (approximately) constant.
737: 
738: 
739: We see from Eq.~(\ref{eq3.2}) that, during the radiation era (starting, say, immediately 
740: after the end of inflation), i.e. when the universe is dominated by an ultra-relativistic gas 
741: ($\rho_A - 3 P_A = 0$), the ``driving force'' on the right-hand side of (\ref{eq3.2}) 
742: vanishes, so that $\varphi$ is not driven further away towards infinity. Actually, one should 
743: take into account both the ``inertial'' effect of the ``velocity'' $\varphi'$ acquired during 
744: the preceding inflationary driving of $\varphi$, and the integrated effect of the many ``mass 
745: thresholds'', $T_A \sim m_A$, when some component becomes non-relativistic (so that $\rho_A - 
746: 3 P_A \ne 0$). Using the results of \cite{DN93,DP94} one sees that, in our case, both these 
747: effects have only a small impact on the value of $\varphi$. Therefore, to a good approximation 
748: $\varphi \simeq \varphi_{\rm end}$ until the end of radiation era.
749: 
750: On the other hand, when the universe gets dominated by non-relativistic  matter, one 
751: gets a non-zero driving force in Eq.~(\ref{eq3.2}). In the slow roll approximation,
752: as the transient behaviour has died out, since $w=P/\rho$ gets negligible, we have simply
753: \be
754: \label{eq3.3'}
755: \varphi'_m = - \alpha_m (\varphi) \, ,
756: \ee
757: where $\varphi'_m$ stands the $\varphi$-velocity during matter domination.
758: and $\alpha_m 
759:  (\varphi)$ denotes the coupling (\ref{eq3.1}) to dark matter.
760: 
761: The coupling to dark matter,  $\alpha_m (\varphi)$, depends on the assumption one makes
762: about the asymptotic behaviour, at strong bare string coupling,  of the mass of the
763: WIMPs constituting the dark matter.
764: One natural looking, minimal assumption is that dark matter, like all visible types of 
765: matter, is coupled in 
766: a way which levels off at strong-bare-coupling, as in Eq.~(\ref{eq1.3}). In other words, one 
767: generally expects that $m_m (\varphi) \simeq m_m (+\infty) (1+b_m \, e^{-c \varphi})$ so that 
768: $\alpha_m (\varphi) \simeq -b_m \, c \, e^{-c \varphi}$. It is then easy to solve 
769: Eq.~(\ref{eq3.3'}), with initial conditions $\varphi_0 = \varphi_{\rm end}$, $\varphi'_0 = 0$ 
770: (inherited from radiation era) at the beginning of the matter era. But we shall not bother to 
771: write the explicit solution because it is easily seen that the smallness of $e^{-c 
772: \varphi_{\rm end}}$ guarantees that the ``driving force'' $\propto \alpha_m (\varphi)$ remains 
773: always so small that the ${\cal O} (10)$ $e$-folds of matter era until vacuum-energy 
774: domination (or until the present) have only a fractionally negligible effect on $\varphi$.
775: 
776: A more significant evolution of $\varphi$ during the matter era is provided
777: if, as first proposed in \cite{DGG} and taken up in \cite{SBM,OP}, 
778: dark matter couples much more strongly to 
779: $\varphi$ than ``ordinary'' matter. 
780: Such a stronger coupling to dark matter, which is not constrained by usual equivalence 
781: principle experiments, follows assuming more general quantum corrections in the 
782: dark matter sector of the theory, i.e. corrections such that the dark matter mass 
783: $m_m (\varphi)$, instead of
784: levelling off, either vanishes or keeps increasing at strong bare coupling:
785: $m_m(\varphi) \propto e^{c_m \varphi}$, so that $\alpha_m = c_m$ is a (negative or 
786: positive) constant.
787: In \cite{GPV} (but see also \cite{AT}) it has been shown that under the latter
788: assumption (i.e. with a positive coupling parameter $\alpha_m >0 $) 
789: the dilaton can play the role of quintessence, leading to a late-time cosmology
790: of accelerated expansion. By Eq. (\ref{eq3.3'}) we have
791: $\varphi = \varphi_{\rm end} - \alpha_m p $, where $p$ is now counted from the end of 
792: the radiation era. 
793: Given that about nine e-folds separate us from the end of the radiation era,
794: we see that such an evolution might (if  $|\alpha_m|$ is really of order unity)
795:  have a significant effect on the present value
796: of $\varphi$ (when compared with the value at the end of inflation, i.e.
797: $c \varphi_{\rm end} \sim \ln (1/\delta_H) \sim 10$). However, the running of 
798: $ \varphi$ during the matter era changes the standard recent cosmological picture
799: and is therefore constrained by observations.
800: In fact, by equation (\ref{eq3.2bis}), the total matter-era equation of state parameter $w_{\rm tot}$ 
801: in the presence of the dilaton reads $ w_{\rm tot} = (\varphi'_m)^2/3$. Accordingly, 
802: the matter density varies as $ \rho \propto a^{- 3 (1 + w_{\rm tot})} = 
803: a^{-( 3 +(\varphi'_m)^2 )}$, possibly affecting the
804: standard scenario of structure formation as well as the  global temporal
805: picture between now and the epoch of matter-radiation equality. 
806: The compatibility with phenomenology therefore puts constraints on the magnitude of
807: ${\varphi'_m}^2 = \alpha_m (\varphi)^2$.
808:  In  \cite{GPV}   $w_{\rm tot} < 0.1$, i.e.
809: $v \equiv {\varphi'_m}^2/ 0.3 < 1$,  was suggested to be  the maximal deviation one can 
810: roughly tolerate during the matter era, the establishment of a more precise bound being presently
811: under study \cite{AGUT}.
812:  We shall
813: therefore assume either that we are in the ``normal'' case  where the dilaton does not
814: couple more strongly to dark matter than to ordinary matter (so that 
815: $\alpha_m (\varphi) \simeq -b_m \, c \, e^{-c \varphi} \ll 1$), or that, if
816: it does, $ \alpha_m ^2 < 0.3$.
817: This leads to a displacement of $ \varphi$ during matter era 
818: smaller than the dispersion  $\varphi_{\rm end} - \varphi_{\rm cl, end} \sim 
819: \varphi_{\rm cl, end}$ produced by quantum fluctuations during inflation, Eq.~(\ref{eq2.23'}).
820: 
821: 
822: In this context one should also
823: consider the attraction effect of a negative pressure component, either in the form of
824: a $\varphi$-dependent vacuum energy (dilatonic quintessence) 
825: or in the form of any other, $\varphi$-independent  component (such as a ``genuine'' 
826: cosmological constant). Of course, the present recent 
827: ($z \lesssim 1$) accelerated expansion phase is very short (in ``$p$-time'') 
828: and sensible changes of the dilaton 
829: value since the end of matter domination are not expected. Still, it is crucial to  
830: estimate the present dilaton velocity $\varphi'_0$ 
831: since it is related to the cosmological variations of the coupling 
832: constants (see next section). 
833: In the general case where both non-relativistic matter and (possibly 
834: $\varphi$-dependent) vacuum energy density $V(\varphi)$ are present, the value
835: of $\varphi_0'$ predicted by our model is obtained by applying Eq. (\ref{eq3.2})
836: (in the slow-roll approximation):
837: \be \label{eq2.32bis}
838: (\Omega_m + \Omega_V)(1-w_0) \varphi'_0 = (\Omega_m + 2 \Omega_V ) \varphi_0' 
839: = - \Omega_m \alpha_m - 4 \Omega_V \alpha_V.
840: \ee 
841: 
842: In the above expression $\Omega_m $ and $\Omega_V$ are, respectively, 
843: the non-relativistic (dark) matter- and the 
844: vacuum-fraction of critical energy density ($\rho_c \equiv (3/2) {\widetilde{m}_P^2} H^2$),
845:  and  the already mentioned prescriptions   
846: $\alpha_V =  \frac{1}{4} \, \partial 
847: \ln V(\varphi) / \partial \, \varphi$, $P_V = -\rho_V = -V(\varphi)$ have been used.
848: 
849: The value of ${\varphi'_0}$ is therefore some combination of the values of $\alpha_m$ and
850: $\alpha_V$. We can have two classes of contrasting situations: In the first class,
851:  the dilaton couples
852: ``normally'' (i.e. weakly) both to dark matter and to dark energy, i.e. both 
853:  $\alpha_m \simeq -b_m \, c \, e^{-c \varphi} \ll 1$ and $\alpha_V \ll 1$ and 
854:  Eq.~(\ref{eq2.32bis}) implies $\varphi'_0 \ll 1$. In the second class, the dilaton
855:  couples more strongly to some type of dark matter or energy, i.e. either (or both) 
856:  $\alpha_m$  or/and   $\alpha_V$ is of order unity so that ${\varphi'_0} = {\cal O} (1)$.
857: The second case is realized in the scenario of \cite{GPV}. In the context of this 
858: scenario we have an exponential dependence of the potential on $\varphi$,
859: $ V(\varphi) \simeq V_1 e^{-c\varphi}$ so that $\alpha_V \simeq - (c/4)$ and
860: \be \label{eq2.34}
861: \varphi'_0 \, = \, \frac{c\, \Omega_V - \alpha_m \Omega_m}{2\Omega_V + \Omega_m} \, \lesssim  \,
862: \frac{c}{3}.
863: \ee 
864: The last inequality follows from the bound
865: $\alpha_m >c/2$ (which is a necessary condition to have positive
866: acceleration in the model \cite{GPV}) and the reasonable bound $\Omega_m > 0.25$. 
867:  
868:   In the present work, we
869:  wish, however, to be as independent as possible from specific assumptions (as the ones
870:  used in \cite{GPV}). Therefore, rather than insisting on specific (model-dependent) predictions
871:  for the present value of ${\varphi'_0}$ we wish to find the (model-independent) upper
872: bounds on the possible values of ${\varphi'_0}$ set by current observational data. 
873: There are several ways of getting such phenomenological bounds, because the existence
874: of a kinetic energy (and pressure) associated to $d \varphi / d t = H \varphi'$ has
875: several observable consequences. A rather secure bound can be obtained by relating
876: the value of $\varphi'$ to the  deceleration parameter 
877: $q \equiv - \ddot{a}a/\dot{a}^2$. In the general class of models that we consider,
878:  the cosmological
879: energy density and pressure have (currently) three significant contributions: dark matter
880: ( $\Omega_m = \rho_m/ \rho_c$), dark energy ($\Omega_V$) and the kinetic effect of a
881: scalar field ($\Omega_k = \rho_k/\rho_c$ with $\rho_k = (\widetilde{m}_P^2 /2) (d\varphi/dt)^2 = 
882: (\widetilde{m}_P^2 /2) H^2 {\varphi'}^2 $ so that $\Omega_k = \varphi'^2/3$).
883: We assume (consistently with recent cosmic background data) that the space curvature is
884: zero. Therefore we have the first relation 
885: \be \label{om1}
886:  \Omega_m + \Omega_V + \Omega_k = 1 = \Omega_m + \Omega_V + \varphi'^2/3 \, .
887: \ee 
888: The deceleration parameter is given by the general expression
889: $ 2 q = \Sigma_A \Omega_A ( 1 + 3 w_A)$. Using $ w_m = 0$, $ w_V = -1$ and $w_k = + 1$,
890: we get 
891: \be \label{q1}
892: 2 q  = \Omega_m - 2 \Omega_V + \frac{4}{3} \varphi'^2.
893: \ee
894: Using the relation (\ref{om1})  above to eliminate $\Omega_V$ we get the following expression for 
895: $\varphi'^2$ in terms of the observable quantities $q$ and $\Omega_m$
896: \be \label{q2}
897: \varphi'^2 = 1 + q - \frac{3}{2} \Omega_m.
898: \ee
899: 
900: The supernovae Ia data \cite{SNI} give a strict upper bound on the
901: present value $q_0$: $q_0<0$. A generous lower bound on the present value of
902: $\Omega_m$ is $\Omega_{m 0} > 0.2$ \cite{omegam}. 
903: Inserting these two constraints in Eq.(\ref{q2}) finally yields
904: the safe upper bound
905: \be \label{q3}
906:  {\varphi'_0}^2 < 0.7 \, , \; {\rm i.e.} \; \vert \varphi'_0 \vert < 0.84 \; .
907: \ee
908:  
909:  To summarize, 
910: quite different rates of evolution for the dilaton are possible. A very slow variation
911: is expected whenever dilaton couplings to both dark energy and dark matter follow
912: the ``normal'' behaviour (\ref{eq1.3}). Otherwise, dilaton variations on the Hubble 
913: scale are expected. However, cosmological observations set the strict
914: upper bound (\ref{q3}) on the present time variation of $\varphi$.
915: For the purpose of the present section (evaluating the current location of the dilaton)
916: these two alternatives do not make much difference because the vacuum-dominance
917: era has started less than about 0.7 e-folds away ($\ln (1+z_*)$ with $z_* < 1$).
918: Therefore, $\varphi$ did not have enough ``$p$-time'' , during vacuum dominance, to
919: move much, even if it is coupled to vacuum energy with $\alpha_V \simeq - (c/4) \sim 1$.
920: 
921:  Finally, we conclude from this analysis that, to a 
922: good approximation (and using the fact that the phenomenology of
923: the matter-era constrains the dark-matter couplings of the dilaton to be rather small),
924:  the value of $\varphi$ now is essentially given by the value $\varphi_{\rm 
925: end}$ at the end of inflation, i.e. by Eq.~(\ref{eq2.23}).
926: 
927: \section{Deviations from general relativity induced by a runaway dilaton}\label{sec3}
928: 
929: \subsection{Composition-independent deviations from general relativity}
930:  
931: 
932: The previous section has reached the conclusion that present deviations 
933: from general relativity are given, to a 
934: good approximation, by the values of the matter-coupling coefficients 
935: $\alpha_{A} (\varphi)$ 
936: given by Eq.~(\ref{eq3.1}) calculated at $\varphi \simeq \varphi_{\rm end}$ as given by 
937: Eq. (\ref{eq2.23'}). Let us now see the 
938: meaning of this result in terms of observable quantities.
939: 
940: Let us first consider the (approximately) composition-independent deviations from general 
941: relativity, i.e. those that do not essentially depend on violations of the equivalence 
942: principle. Most composition-independent gravitational experiments (in the solar system or in 
943: binary pulsars) consider the long-range interaction between objects whose masses are 
944: essentially baryonic (the Sun, planets, neutron stars). As argued in \cite{TV88,DP94} the 
945: relevant coupling coefficient $\alpha_A$ is then approximately universal and given by the 
946: logarithmic derivative of the QCD confinement scale $\Lambda_{\rm QCD} (\varphi)$, because the 
947: mass of hadrons is essentially given by a pure number times $\Lambda_{\rm QCD} (\varphi)$. [We 
948: shall consider below the small, non-universal, corrections to $m_A (\varphi)$ and 
949: $\alpha_A (\varphi)$ 
950: linked to QED effects and quark masses.] Remembering from Eq.~(\ref{eq1.1}) the 
951: fact that, in the string frame (where there is a fixed cut-off linked to the string mass 
952: $\widetilde{M}_s \sim (\alpha')^{-1/2}$) the gauge coupling is dilaton-dependent ($g_F^{-2} = 
953: B_F (\varphi)$), we see that (after conformal transformation) the Einstein-frame confinement 
954: scale has a dilaton-dependence of the form
955: \be
956: \label{eq3.5}
957: \Lambda_{\rm QCD} (\varphi) \sim C^{-1/2} \, B_g^{-1/2} (\varphi) \exp [- 8 \pi^2 \, b_3^{-1} 
958: \, B_F (\varphi)] \, \widetilde{M}_s \, ,
959: \ee
960: where $b_3$ denotes the one-loop (rational) coefficient entering the Renormalization Group 
961: running of $g_F$. Here $B_F (\varphi)$ denotes the coupling to the ${\rm SU} (3)$ gauge 
962: fields. For simplicity, we shall assume that (modulo rational coefficients) all gauge fields 
963: couple (near the string cut off) to the same $B_F (\varphi)$. This yields the following 
964: approximately universal dilaton coupling to hadronic matter
965: \be
966: \label{eq3.6}
967: \alpha_{\rm had} (\varphi) \simeq \left( \ln \left( \frac{\widetilde{M}_s}{\Lambda_{\rm QCD}} 
968: \right) + \frac{1}{2} \right) \frac{\partial \ln B_F^{-1} (\varphi)}{\partial \, \varphi} \, .
969: \ee
970: We recall that the quantity $\alpha_{\rm had}(\varphi)$, 
971: which measures the coupling of the
972: dilaton to hadronic matter, should not be confused with any "strong" gauge
973: coupling, $\alpha_s = g_s^2/4 \pi$.
974: Numerically, the coefficient in front of the R.H.S. of (\ref{eq3.6}) is of order 40. 
975: Consistently with our basic assumption (\ref{eq1.3}), we parametrize 
976: the $\varphi$ dependence of the gauge coupling $g_F^2 = B_F^{-1}$ as
977: \be
978: \label{eq3.7}
979: B_F^{-1} (\varphi) = B_F^{-1} (+ \infty) \, [1 - b_F \, e^{-c\varphi}] \, .
980: \ee
981: Note that, like $b_\lambda$ (see section \ref{sec2}), 
982: also $b_F$ is expected to be smallish [$\sim B_F^{-1} (+ \infty)$ or, equivalently,
983: $\sim C_F^{-1}$ in the notations of (\ref{eq1.3})] and typically the ratio 
984: $b_F/b_\lambda$ is of order unity.
985: We finally obtain 
986: \be
987: \label{eq3.8}
988: \alpha_{\rm had} (\varphi) \simeq 40 \, b_F \, c \, e^{-c\varphi} \, .
989: \ee
990: We can now insert the estimate (\ref{eq2.23'}) of the value of $\varphi$ reached because of the 
991: cosmological evolution. Neglecting the ${\cal O} (1)$ renormalization factor due to quantum noise,
992: we get the estimate
993: \be
994: \label{eq3.9}
995: \alpha_{\rm had} (\varphi_{\rm end}) \simeq 3.2 \, \frac{b_F}{b_{\lambda} \, c} \, 
996: \delta_H^{\frac{4}{n+2}} \, ,
997: \ee
998: \be
999: \label{eq3.10}
1000: \alpha_{\rm had}^2 (\varphi_{\rm end}) \simeq 10 \left(\frac{b_F}{b_{\lambda} \, c} \right)^2 
1001: \delta_H^{\frac{8}{n+2}} \, .
1002: \ee
1003: As said above, it is plausible to expect that the quantity $c$ (which is a ratio) and the 
1004: ratio $b_F / b_{\lambda}$ are both of order unity. This then leads to the numerical estimate 
1005: $\alpha_{\rm had}^2 \sim 10 \, \delta_H^{\frac{8}{n+2}}$, with $\delta_H \simeq 5 \times 
1006: 10^{-5}$. An interesting aspect of this result is that the expected present value of 
1007: $\alpha_{\rm had}^2$ depends rather strongly on the value of the exponent $n$ (which entered 
1008: the inflaton potential $V(\chi) \propto \chi^n$). In the case $n=2$ (i.e. $V(\chi) = 
1009: \frac{1}{2} \, m_{\chi}^2 \, \chi^2$) we have $\alpha_{\rm had}^2 \sim 2.5 \times 10^{-8}$, 
1010: while if $n=4$ ($V (\chi) = \frac{1}{4} \, \lambda \, \chi^4$) we have $\alpha_{\rm had}^2 \sim 
1011: 1.8 \times 10^{-5}$.
1012: 
1013:  How do these numbers compare to present (composition-independent) 
1014: experimental limits on deviations from Einstein's theory \cite{exp}? This question has been 
1015: addressed in the literature. Concerning solar-system (post-Newtonian) tests it was shown (see, 
1016: e.g., \cite{DEF92}) that the two main ``Eddington'' parameters $\gamma - 1$ and $\beta - 1$ 
1017: measuring post-Newtonian deviations from general relativity are linked as follows to the 
1018: dilaton coupling $\alpha_{\rm had} (\varphi)$:
1019: \be
1020: \label{eq3.11}
1021: \gamma - 1 = - 2 \, \frac{\alpha_{\rm had}^2}{1 + \alpha_{\rm had}^2} \simeq - 2 \, \alpha_{\rm 
1022: had}^2 \, ,
1023: \ee
1024: \be
1025: \label{eq3.12}
1026: \beta - 1 = \frac{1}{2} \, \frac{\alpha'_{\rm had} \, \alpha_{\rm had}^2}{(1 + \alpha_{\rm 
1027: had}^2)^2} \simeq \frac{1}{2} \, \alpha'_{\rm had} \, \alpha_{\rm had}^2 \, ,
1028: \ee
1029: where $\alpha'_{\rm had} \equiv \partial  \alpha_{\rm had} (\varphi) / \partial \varphi$.
1030: 
1031: {}From Eq.~(\ref{eq3.8}) we see that $\alpha'_{\rm had} \simeq -c \, \alpha_{\rm had}$, so that 
1032: the deviation $\beta - 1$ is ${\cal O} (\alpha_{\rm had}^3)$ and thereby predicted to be too 
1033: small to be phenomenologically interesting. This leaves $\gamma - 1 \simeq -2 \, \alpha_{\rm 
1034: had}^2$ as the leading observable deviation. The best current solar-system limit on $\gamma - 
1035: 1$ comes from Very Long Baseline Interferometry measurements of the deflection of radio waves 
1036: by the Sun and is (approximately) $\vert \gamma - 1 \vert \lesssim 2 \times 10^{-4}$, 
1037: corresponding to $\alpha_{\rm had}^2 \lesssim 10^{-4}$ (see \cite{exp} for reviews and 
1038: references). In addition to solar-system tests, we should also consider binary-pulsar tests 
1039: which provide another high-precision window on possible deviations from general relativity. 
1040: They have been analyzed in terms of the two quantities $\alpha_{\rm had}$ (denoted $\alpha$) 
1041: and $\alpha'_{\rm had}$ (denoted $\beta$) in \cite{DEF}. The final conclusion is that the 
1042: binary-pulsar limit on $\alpha_{\rm had}$ is of order $\alpha_{\rm had}^2 \lesssim 10^{-3}$.
1043: 
1044: At this stage it seems that the runaway scenario explored here is 
1045: leading to deviations from 
1046: general relativity which are much smaller than present experimental limits. However, we must 
1047: turn our attention to {\it composition-dependent} effects which turn out to be much more 
1048: sensitive tests.
1049: 
1050: \subsection{Composition-dependent deviations from general relativity}
1051: 
1052: Let us then consider situations where the non-universal couplings of the dilaton induce 
1053: (apparent) violations of the equivalence principle. Let us start by considering the 
1054: composition-dependence of the dilaton coupling $\alpha_A$, Eq.~(\ref{eq3.1}), i.e. the 
1055: dependence of $\alpha_A$ on the type of matter we consider.
1056: The definition of $\alpha_A$ is such that, at the Newtonian approximation, the interaction 
1057: potential between particle $A$ and particle $B$ is $-G_{AB} \, m_A \, m_B / r_{AB}$ where 
1058: \cite{DP94}
1059: \be
1060: \label{eq3.13}
1061: G_{AB} = G (1 + \alpha_A \, \alpha_B) \, .
1062: \ee
1063: Here, $G$ is the bare gravitational coupling constant entering the Einstein-frame action 
1064: (\ref{eq2.4}), and $\alpha_A = \alpha_A (\varphi)$ is the strength of the dilaton coupling to 
1065: $A$-particles, taken at the present (cosmologically determined) VEV of $\varphi$. The term 
1066: $\alpha_A \, \alpha_B$ comes from the additional attractive effect of dilaton exchange. Two 
1067: test masses, made respectively of $A$- and $B$-type particles will then fall in the 
1068: gravitational field generated by an external mass $m_E$ with accelerations differing by
1069: \be
1070: \label{eq3.14}
1071: \left( \frac{\Delta a}{a} \right)_{AB} \equiv 2 \, \frac{a_A - a_B}{a_A + a_B} \simeq 
1072: (\alpha_A - \alpha_B) \, \alpha_E \, .
1073: \ee
1074: We have seen above that in lowest approximation $\alpha_A \simeq \alpha_{\rm had}$ does not 
1075: depend on the composition of $A$. We need, however, now to retain the small 
1076: composition-dependent effects to $\alpha_A$ linked to the $\varphi$-dependence of QED and 
1077: quark contributions to $m_A$. This has been investigated in \cite{DP94} with the result
1078: \be
1079: \label{eq3.15}
1080: \left( \frac{\Delta a}{a} \right)_{AB} = \left( \frac{\alpha_{\rm had}}{40} \right)^2 \left[ 
1081: C_B \, \Delta \left( \frac{B}{M} \right) + C_D \, \Delta \left( \frac{D}{M} \right) + C_E \, 
1082: \Delta \left( \frac{E}{M} \right) \right]_{AB} \,  ,
1083: \ee
1084: where $(\Delta X)_{AB} \equiv X_A - X_B$, where $B \equiv N + Z$ is the baryon number, $D 
1085: \equiv N-Z$ the neutron excess, $E \equiv Z (Z-1) / (N+Z)^{1/3}$ a quantity linked to nuclear 
1086: Coulomb effects, and where $M \equiv m/u$ denotes the mass in atomic mass unit, $u = 
1087: 931.49432$ MeV. It is difficult (and model-dependent) to try to estimate the coefficients 
1088: $C_B$ and $C_D$. It was argued in \cite{DP94} that their contributions to (\ref{eq3.15}) is 
1089: generically expected to be sub-dominant with respect to the last contribution, $\propto C_E$, 
1090: which can be better estimated because it is linked to the $\varphi$-dependence of the 
1091: fine-structure constant $e^2 \propto B_F^{-1} (\varphi)$. This then leads to the numerical 
1092: estimate $C_E \simeq 3.14 \times 10^{-2}$ and a violation of the universality of free fall 
1093: approximately given by
1094: \be
1095: \label{eq3.16}
1096: \left( \frac{\Delta a}{a} \right)_{AB} \simeq 2 \times 10^{-5} \, \alpha_{\rm had}^2 \left[ 
1097: \left( \frac{E}{M} \right)_A - \left( \frac{E}{M} \right)_B \right] \, .
1098: \ee
1099: 
1100: The values of $B/M$, $D/M$ and $E/M$ have been computed in \cite{D96}. For mass-pairs that 
1101: have been actually used in recent experiments (such as Beryllium and Copper), as well as for 
1102: mass-pairs that are planned to be used in forthcoming experiments (such as Platinum and 
1103: Titanium) one finds: $(E/M)_{\rm Cu} - (E/M)_{\rm Be} = 2.56$, $(E/M)_{\rm Pt} - (E/M)_{\rm 
1104: Ti} = 2.65$. Using the average estimate $\Delta (E/M) \simeq 2.6$, we get from (\ref{eq3.16}) 
1105: and (\ref{eq3.10}) the estimate
1106: \be
1107: \label{eq3.17}
1108: \left( \frac{\Delta a}{a} \right) \simeq 5.2 \times 10^{-5} \, \alpha_{\rm had}^2 \simeq 5.2 
1109: \times 10^{-4} \left( \frac{b_F}{b_{\lambda} \, c} \right)^2 \, \delta_H^{\frac{8}{n+2}} 
1110: \, .
1111: \ee
1112: Note also (from (\ref{eq3.11})) the link between composition-dependent effects and 
1113: post-Newtonian ones
1114: \be
1115: \label{eq3.18}
1116: \left( \frac{\Delta a}{a} \right) \simeq - 2.6 \times 10^{-5} (\gamma - 1) \, .
1117: \ee
1118: As current tests of the universality of free fall (UFF) have put limits in the $10^{-12}$ 
1119: range (e.g. $(\Delta a / a)_{\rm Be \, Cu} = (-1.9 \pm 2.5) \times 10^{-12}$ from 
1120: \cite{Su94}), we see from Eq.~(\ref{eq3.18}) that this corresponds to limits on $\gamma - 1$ 
1121: or $\alpha_{\rm had}^2$ in the $10^{-7}$ range. Therefore tests of the UFF put much more 
1122: stringent limits on dilaton models than solar-system or binary-pulsar tests.
1123: 
1124: 
1125: If we insert the estimate $\delta_H \sim 5 \times 10^{-5}$ in (\ref{eq3.17}) we obtain a level 
1126: of violation of UFF due to a runaway dilaton which is
1127: \be
1128: \label{eq3.19}
1129: \frac{\Delta a}{a} \simeq 1.3 \left( \frac{b_F}{b_{\lambda} \, c} \right)^2 \times 
1130: 10^{-12} \quad \hbox{for} \ n=2 \, ,
1131: \ee
1132: \be
1133: \label{eq3.20}
1134: \frac{\Delta a}{a} \simeq 0.98 \left( \frac{b_F}{b_{\lambda} \, c} \right)^2 \times 
1135: 10^{-9} \quad \hbox{for} \ n=4 \, .
1136: \ee
1137: At face value, one is tempted to conclude that a scenario with $n=4$ (i.e. $V(\chi) \propto 
1138: \chi^4$) tends to be too weak an attractor towards $\varphi = + \infty$ to be naturally 
1139: compatible with equivalence-principle tests. [See, however, the discussion below.] On the 
1140: other hand, the simple scenario $n=2$ ($V(\chi) = \frac{1}{2} \, m_{\chi}^2 \, \chi^2$) is 
1141: quite appealing in that it naturally provides enough attraction towards $\varphi = + \infty$ 
1142: to be compatible with all existing experimental tests. At the same time it suggests 
1143: that a modest improvement in the precision of UFF experiments might discover a violation 
1144: caused by a runaway dilaton.
1145: 
1146: \subsection{Cosmological variation of ``constants''}
1147: 
1148: Let us now consider another possible deviation
1149: from General Relativity and the standard model: a possible variation 
1150: of the coupling constants, most notably of the fine structure 
1151: constant $e^2/\hbar c$ on which the strongest limits are available.
1152: We will discuss first the effects due to the cosmological time-variation of the 
1153: homogeneous component of $\varphi$ 
1154: and, in the next subsection, the possible spatial (and time) variations 
1155: due to  quantum fluctuations of $\varphi$
1156:  as they got amplified during inflation.
1157: 
1158:  Consistently with our previous assumptions we expect 
1159: $e^2 \propto B_F^{-1} (\varphi)$ so that, from (\ref{eq3.7}),
1160: \be
1161: \label{eq3.21}
1162: e^2 (\varphi) = e^2 (+ \infty) \, [1 - b_F \, e^{- c \varphi} ] \, .
1163: \ee
1164: The present logarithmic variation of $e^2$ (using again $dp = H \, dt ;  
1165: \varphi' = d \varphi / d p$) is thus
1166: given by
1167: \be
1168: \label{eq3.21bis}
1169: \frac{d \ln e^2}{H \, dt} = \frac{d \ln e^2}{dp} \simeq b_F \, c \, e^{-c \varphi} \, 
1170: \varphi'_0 \, ,
1171: \ee
1172: where the current value of $\varphi'$,  $\varphi'_0$, is given in general by Eq.~(\ref{eq2.32bis}).
1173: Using  Eq.~(\ref{eq3.8}), we can rewrite the result (\ref{eq3.21bis}) in terms of the 
1174: hadronic coupling:
1175: \be
1176: \label{eq3.21ter}
1177: \frac{d \ln e^2}{H \, dt} \simeq \frac{1}{40} \alpha_{\rm had} \varphi'_0 \, .
1178: \ee
1179: 
1180: As said in section \ref{sec2b}, we have basically two alternatives concerning the 
1181: current coupling of the dilaton
1182: to the dominant energy sources in the universe. 
1183: These two alternatives lead to drastically different predictions for the 
1184: current value of the rate of variation of the fine-structure constant. 
1185: We shall consider
1186: these two alternatives in turn.
1187: 
1188: In the conservative case where 
1189: the dilaton does not play any special role in the present accelerated phase of the 
1190: universe ($\alpha_V \simeq 0$) nor does it have any stronger coupling to dark matter 
1191: than to visible matter ($\alpha_m \simeq -b_m \, c \, e^{-c\varphi}$) the dilaton 
1192: ``velocity'' $\varphi'$ is exponentially suppressed ( so that, from (\ref{om1}),
1193: $ \Omega_V \simeq  1 - \Omega_m$) and 
1194: by Eq. (\ref{eq2.32bis}) one obtains
1195: \be \label{eq3.22'}
1196: \frac{d \ln e^2}{H \, dt} \, \simeq \,  - \frac{\Omega_m}{\Omega_m + 2 \Omega_V} \, b_F \, c\,  
1197: e^{-c\varphi}\alpha_m(\varphi) \, \simeq \, \frac{\Omega_m}{2-\Omega_m} \, b_F \, b_m \, c^2 \, 
1198: e^{-2 c\varphi}.
1199: \ee
1200: An indicative value for the ratio $\Omega_m /(\Omega_m + 2 \Omega_V) \simeq 
1201: \Omega_m /(2-\Omega_m)$, by taking for instance
1202: $\Omega_m = 0.3$, is $0.18$. 
1203: As above, it is useful to relate (\ref{eq3.22'}) to the estimate (\ref{eq3.8}) for $\alpha_{\rm had}$. 
1204: This yields
1205: \be
1206: \label{eq3.23}
1207: \frac{d \ln e^2}{H \, dt} \, \simeq \, \frac{1}{(40)^2} \, \frac{\Omega_m}{2-\Omega_m} \, \frac{b_m}{b_F} \, 
1208: \alpha_{\rm had}^2 \, .
1209: \ee
1210: In terms of the UFF level $\Delta a / a$ predicted by our model in (\ref{eq3.17}) we see also 
1211: that
1212: \be
1213: \label{eq3.24}
1214: \frac{d \ln e^2}{H \, dt} \, \simeq \, 
1215: 12 \, \frac{\Omega_m}{2-\Omega_m} \, \frac{b_m}{b_F} \, \frac{\Delta a}{a} \, .
1216: \ee
1217: Even  if the universe were completely dominated by dark matter 
1218: ($\Omega_m=1$) we see, assuming that $b_m/b_F$ is of order unity,
1219:  that current experimental limits on UFF 
1220: ($\Delta a / a \lesssim 
1221: 10^{-12}$) imply (within dilaton models) that $\vert d \ln e^2 / dt \vert \lesssim 10^{-11} \, 
1222: H \sim 10^{-21} \, {\rm yr}^{-1}$ (the sign of $d \ln e^2 / dt$ being given by the
1223: sign of $b_m/b_F$). This level of variation is much smaller than the current best 
1224: limit on the time variation of $e^2$, namely $\vert d \ln e^2 / dt \vert \lesssim 5 \times 
1225: 10^{-17} \, {\rm yr}^{-1} \sim 5 \times 10^{-7} \, H$, as obtained from an analysis of Oklo data 
1226: \cite{Oklo}. (Note that the assumption-dependent analysis of Ref. \cite{OPQCCV}
1227: gives a limit on the variation of $e^2$ which is strengthened by 
1228: about two orders of magnitude.)
1229: 
1230: 
1231: The situation, however, is drastically different if we consider the alternative case
1232: where the dilaton coupling to the current dominant energy sources does not tend to 
1233: triviality, as in the case of a $\varphi$-dependent vacuum energy
1234: $V(\varphi) = V_0 + V_1 e^{-c\varphi}$ when the first 
1235: term is zero or negligible.
1236: In such a case the dilaton shares a relevant part of the 
1237: total energy density and more significant (though still quite
1238: constrained by UFF data) variations of 
1239: the coupling constants are generally expected.
1240: A general expression for the dilaton ``velocity'' is given in eq.
1241: (\ref{q2}) in terms of observable quantities.  
1242: Using Eqs.~(\ref{q2}) and (\ref{eq3.21ter}) one can relate the expected variation 
1243: of the electromagnetic coupling constant to the hadronic coupling:
1244: \be \label{eq3.25'}
1245: \frac{d \ln e^2}{H \, dt} \simeq \pm \, \frac{\alpha_{\rm had}}{40}  
1246: \, \sqrt{  1 + q_0- 3\Omega_m/2  } 
1247:  \, .
1248: \ee 
1249: 
1250: We can also use the estimate (\ref{eq3.9}) relating
1251: ${\alpha}_{\rm had}$ to the density fluctuations generated during inflation.
1252: We obtain
1253: \be
1254: \label{eq3.26}
1255: \frac{d \ln e^2}{H \, dt} \simeq  
1256: \pm \, 8 \times 10^{-2}  \, \sqrt{  1 + q_0- 3\Omega_m/2  }
1257: \, \frac{b_F}{b_{\lambda} c} \,  \delta_H^{\frac{4}{n+2}}  \, .
1258: \ee 
1259: 
1260: However, in view of the theoretical uncertainties attached to the initial 
1261: conditions
1262: $\chi_{\rm in}$ and $\varphi_{\rm in}$ used in the estimate (\ref{eq3.9}), 
1263: as well as  the ones associated to the order unity ratio 
1264: $b_F / (b_{\lambda} c)$,
1265: it is more interesting
1266: to rewrite our prediction in terms of {\it observable} quantities. Using again
1267: the link Eq.~(\ref{eq3.17}) between ${\alpha}_{\rm had}$ and the observable
1268: violation of the universality of free fall (UFF) the 
1269: above result can be written in the form
1270: \be
1271: \label{eq3.26''}
1272: \frac{d \ln e^2}{H \, dt} \simeq \pm \, 3.5 \times 10^{-6} \, 
1273: \sqrt{  1 + q_0- 3\Omega_m/2  }
1274: \, \sqrt{ 10^{12} \frac{\Delta a}{a}} \, .
1275: \ee
1276: 
1277: 
1278: Note that the sign of the variation of $e^2$ is in general model-dependent (as it
1279: depends both on the sign of $b_F$ and the sign of $\varphi'_0$). Specific classes
1280: of models might, however, favour particular signs of $d e^2/d t$. For instance,  
1281:  from the point of view of \cite{V01} one would expect the ${\cal O} (e^{-\phi})$ terms
1282:  in Eq.~(\ref{eq1.3}) to be positive, which would then imply that $b_F$ is positive.
1283:  If we combine this information with the prediction Eq.~(\ref{eq2.34}) of the
1284: model \cite{GPV} implying that $\varphi'$ is also positive, we would reach the conclusion that 
1285: $e^2$ must be currently {\it increasing}. 
1286: 
1287: Independently of this question of the sign, we see that
1288: Eq.~(\ref{eq3.26''}) predicts an interesting link between the observational violation 
1289: of the UFF
1290: (constrained to $\Delta a/a \lesssim 10^{-12}$),
1291: and the current time-variation of the fine-structure constant. Contrary to the relation 
1292: (\ref{eq3.24}), obtained above under the alternative assumption about the dilaton dependence
1293: of the dominant cosmological energy, which predicted a  relation linear in $\Delta a/a$,
1294:  we have here a relation involving the square root of the UFF violation (such
1295: a relation is similar to the result of \cite{DGG} which concerned the time-variation of
1296: the Newton constant). 
1297: 
1298: The phenomenologically interesting consequence of Eq.~(\ref{eq3.26''})
1299: is to predict a time-variation of constants which may be large enough to be detected
1300: by high-precision laboratory experiments. Indeed, using $H_0 \simeq 66$ km/s/Mpc, and the
1301: plausible estimates
1302: $\Omega_m = 0.3$, $q_0 = -0.4$,
1303: Eq.~(\ref{eq3.26''}) yields the numerical estimate 
1304: $d \ln e^2 / dt \sim \pm 0.9 \times 10^{-16}
1305: \,   \sqrt{ 10^{12} \Delta a/a} \, {\rm yr}^{-1}$.
1306:  Therefore, the current bound on UFF violations
1307: ($\Delta a/a \sim 10^{-12}$) corresponds to the level $10^{-16}{\rm yr}^{-1}$,
1308: which is comparable to the planned sensitivity of currently developed cold-atom clocks
1309: \cite{salomon}. [Present laboratory bounds are at the  $10^{-14}{\rm yr}^{-1}$
1310: level \cite{prestage,salomon}.] Note that if we insert in Eq.~(\ref{eq3.26''}) 
1311: the secure bounds $\Omega_m > 0.2$ and $q_0 < 0$ (leading to the limit Eq.~(\ref{q3})), 
1312: we get as maximal estimate of the time
1313: variation of the fine-structure constant $d \ln e^2 / dt \sim \pm 2.0 \times 10^{-16}
1314: \,   \sqrt{ 10^{12} \Delta a/a} \, {\rm yr}^{-1}$.
1315: We note also that the upper limit on the variation of $e^2$ 
1316: given by the Oklo data, i.e. 
1317: $\vert d \ln e^2 / dt \vert \lesssim 5 \times 10^{-17} \, {\rm yr}^{-1}$ \cite{Oklo},
1318: ``corresponds''  to a violation of the UFF at the
1319: level  $ \sim  10^{-13}$. 
1320: 
1321: In this respect, it is interesting to consider not only the {\it present} variation of
1322: $e^2$ (the only one relevant for laboratory experiments), but also its variation
1323: over several billions of years. (We recall that the  Oklo phenomenon took place 
1324: about two billion years ago, and that astronomical observations constrain the variation
1325: of $e^2$ over the last ten billion years or so). In particular, an interesting question
1326: is to see whether our model could reconcile the Oklo limit (which corresponds to a 
1327: redshift $z \simeq 0.14$) with the recent claim \cite{Webb}
1328: of a variation $ \Delta e^2/ e^2 = ( -0.72 \pm 0.18 ) \times 10^{-5}$ around
1329: redshifts $ z \approx 0.5-3.5$ as proposed in \cite{SBM,OP}. 
1330: The only hope of reconciling the two results would be to
1331: allow for a faster variation of $e^2$ for redshifts $ z > 0.5$. Such recent redshifts have
1332: (apparently) been connected to a transition from matter dominance to vacuum dominance. Let us
1333: see whether taking into account this transition might allow for a large enough change of
1334: $e^2$ around redshifts $ z \approx 0.5-3.5$. We must clearly assume the ``strong coupling'' scenario
1335: $ \alpha_m = {\cal O}(1)$. In this scenario, the variation of $\varphi$
1336: during the matter era is given by Eq.(\ref{eq3.3'}). Neglecting, for simplicity, the transient 
1337: evolution effects localized around the matter-vacuum transition (and treating  both
1338: $\varphi'_m = - \alpha_m$ and $\varphi'_V = \varphi'_0$ as constants), the solution giving the 
1339: recent cosmological evolution of $ \varphi $ reads
1340: $ \varphi - \varphi_0 = - \varphi'_0 \ln (1+z)$ during the vacuum era, and
1341: $ \varphi - \varphi_0 = - \varphi'_0 \ln (1+z_*) - \varphi'_m \ln [ (1+z)/(1+z_*)]$ during
1342: the matter era (the index $0$ refers to the present epoch, i.e. $z=0$;  $z_*$ denotes 
1343: the transition redshift).
1344:  Inserting this change in Eq.(\ref{eq3.21}) leads to the following expression 
1345: for the cosmological change of the fine-structure constant:
1346: \be
1347: \label{3.26''}
1348: \frac{e^2 - e^2_0}{e^2_0} = - {\rm sign} (b_F) \, 3.5 \times 10^{-6} \,
1349: \lbrack \varphi'_0 \ln (1+z_*) + \varphi'_m \ln \frac{ 1+z}{1+z_*} \rbrack \,  
1350: \sqrt{ 10^{12} \frac{\Delta a}{a}}\, .
1351: \ee
1352: Here, we have written the result for the matter era. During the vacuum era the bracket is
1353: simply $[ \varphi'_0 \ln (1+z) ]$. Remembering that the absolute value of $\varphi'_m$ is 
1354: (like that of $\varphi'_V$)
1355: observationally constrained to be smaller than $\sqrt{0.3} \simeq 0.55$ (and that $\varphi'_0$
1356: is also constrained by $ \vert \varphi'_V \vert < 0.84$), we see that there is
1357: no way, within our model, to explain a variation of $e^2$ as large as 
1358: $ \Delta e^2/ e^2 = ( -0.72 \pm 0.18 ) \times 10^{-5}$ around redshifts $ z \approx 0.5-3.5$ 
1359: \cite{Webb}. In our model, even under the assumption that UFF is violated just below
1360: the currently tested level, such a change would have to correspond to a 
1361: value  $ \vert \varphi'_m \vert > 2$, entailing observationally unacceptable modifications
1362: of standard cosmology. [For instance, in the model \cite{GPV} a value as large as
1363:  $\alpha_m >1$ already leads to a pathological behaviour (``total dragging'') 
1364: where all the components scale like radiation.] This difficulty of reconciling 
1365: the Oklo limit with the claim of 
1366: \cite{Webb} was addressed in \cite{OP,SBM} within a different class of models, namely
1367: with a field $\phi$ 
1368: which {\it does not} couple universally to all gauge fields $F_{\mu \nu}$, as the
1369: dilaton $\varphi$ is expected to do. The fact that the field $\phi$ in \cite{OP} 
1370: (or $\psi$ in \cite{SBM}) is assumed to couple only to the electromagnetic 
1371: gauge field
1372: drastically changes our Eq. (\ref{eq3.17}) and allows one to satisfy the UFF limit
1373: $\Delta a/a \lesssim 10^{-12}$, for a stronger coupling of
1374: $\phi$ to electromagnetism than in our class of models, i.e. (in our notation)
1375: for a larger  $d \ln B_F(\varphi)/ d \varphi$. This explains why Ref. \cite{OP} 
1376: could construct some explicit (but fine-tuned) models in which all
1377: observational
1378: limits (UFF, Oklo,...) could be met and still allow for a variation of
1379: $e^2$ as strong as the claim[23].
1380: The maximal variation predicted by Eq.(\ref{3.26''})
1381: for redshifts  corresponding to the matter era
1382: (obtained when $\Delta a/a = 10^{-12}$ and  $\varphi'_m = \pm \sqrt{0.3}$; and assuming a smaller
1383: value of $\varphi'_0$ to be compatible with the Oklo constraint), is of order
1384: $ \Delta e^2/ e^2 = \pm 1.9 \times 10^{-6}$. This is only a factor $\sim 4$ below the claim \cite{Webb}
1385: and is at the level of their one sigma error bar. Therefore a modest improvement in the
1386: observational precision (accompanied by an improved control of systematics) will start to
1387: probe a domain of variation of constants which, according to our scenario, corresponds
1388: to an UFF violation smaller than the $10^{-12}$ level.
1389: 
1390: \subsection{Spatio-temporal fluctuations of the ``constants''}
1391: 
1392: We now turn to the second possible source of spatial-temporal variations 
1393: for $e^2$ in our model,
1394: the quantum fluctuations of the dilaton generated during inflation.  
1395: Within linear perturbation theory, the relevant calculation may
1396: be summarized  as follows.
1397: 
1398:  Consider a flat FRW universe 
1399: $ds^2 = -dt^2 + a(t)^2 \, \Sigma_i dx_i^2 $ . 
1400: The dilaton fluctuations 
1401: can be expanded in Fourier components $\delta \varphi_{\bf k}$ of given comoving
1402: momentum ${\bf k}$ as follows:
1403: \be
1404: \delta\varphi({\bf x},t) = \frac{1}{(2\pi)^{3/2}}\int d^3k \,
1405: \delta \varphi_{\bf k}(t) e^{i {\bf k x}},
1406: \ee
1407: where $t$ is the cosmological time.  
1408: Each Fourier mode $\delta \varphi_{\bf k}$ 
1409: ``leaves'' the horizon during inflation with an amplitude 
1410: $\sim \whh_{\rm ex}(k)/\sqrt{2 k^3}\, $ 
1411: \cite{lyth} where, by definition, $\whh_{\rm ex}(k)$ is the value of the dimensionless
1412: Hubble expansion rate as $k a^{-1}$ equals $H$ during inflation (note that we denote here
1413: $\whh_{\rm ex}$ what was denoted $\whh_{\times}$ above).
1414: Well after the exit ($k \ll a H$) the amplitude of each mode ``freezes out'',
1415: i.e. remains roughly constant, until it reenters the horizon during the 
1416: post-inflationary epoch 
1417: ($k a_{re}^{-1} \simeq H_{re}$). After re-entry  the amplitude starts to 
1418: damp out as $a^{-1}$. 
1419: For a given Fourier mode $\delta\varphi_{\bf k}(t)$, 
1420: the latter damping effect is described by the piecewise function 
1421: \begin{equation}
1422: f_z(k) \equiv \left\{ \begin{array}{llc}
1423: 1 & {\rm if} & \quad a_0^{-1} H_0^{-1} \, k \, < \, (z+1)^{1/2} \\[2mm]
1424: a_0^2 H_0^2 \, (z+1) k^{-2}& {\rm if} & \qquad (z+1)^{1/2} \, < \, 
1425: a_0^{-1} H_0^{-1}\, k \, < \, 10^2  \\[2mm]
1426: 10^{-2} a_0 H_0 \, (z+1) k^{-1}\qquad \qquad & {\rm if} &  \quad a_0^{-1} H_0^{-1}\, k \, 
1427: > \, 10^2 
1428: \end{array} \right.
1429: \ee
1430: Here the cosmological redshift $z=a_0/a(t)-1$ has been introduced in replacement of the 
1431: cosmological time
1432: $t$. The first case refers to  Fourier modes that have
1433:  not reentered yet at redshift $z$ 
1434: and whose amplitudes are still frozen. The second and third cases refer to  modes 
1435: that reenter during matter and radiation domination respectively.
1436: Putting all together, and assuming a gaussian probability distribution for the 
1437: perturbations, we have:
1438: \be \label{eq3.28f}
1439: \langle\delta\varphi_{\bf k} (t)^* \, \delta\varphi_{\bf k'} (t') \rangle  \, \, =\, 
1440: \frac{\whh_{\rm ex}^2(k)}{2k^3} \, f_z(k)\, f_{z'}(k)
1441: \, \delta^3({\bf k} - {\bf k'}) \, .
1442: \ee
1443: Possible spatial/temporal variations of $e^2$ induced by the 
1444: fluctuations of the dilaton will be given by
1445: \be \label{eq3.27}
1446: \left. \frac{\Delta^{\rm fluc} e^2}{e^2}\right|_{({\bf x}, t;\, {\bf x'}, t')} \, = \, 
1447: \frac{d \ln e^2}{d\varphi}\, \Delta^{\rm fluc} \varphi |_{({\bf x}, t;\, {\bf x'}, t')}
1448: \, ,
1449: \ee
1450: where the r.m.s $\Delta^{\rm fluc}\varphi$ between two events (${\bf x}, t$) and (${\bf x'}, t'$) 
1451: is defined as follows
1452: \begin{eqnarray} \nonumber
1453: \Delta^{\rm fluc} \varphi |_{({\bf x}, t;\, {\bf x'}, t')} ^2 \, 
1454: & \equiv & \, \langle\, [\delta\varphi({\bf x}, t) - \delta\varphi({\bf x'}, t')]^2 \rangle  \\[2mm]
1455: & = & \, \frac{\whh_{\rm ex}^2}{(2\pi)^3} \int \frac{d^3 k}{2\, k^3}\, 
1456: \left[f_z(k)^2 + f_{z'}(k)^2
1457: - 2 f_z(k) f_{z'}(k) \, e^{i {\bf k}({\bf x} - {\bf x'})} \right]\\[2mm]
1458: & = & \, \frac{\whh_{\rm ex}^2}{(2\pi)^2} \int_0^\infty \frac{d k}{k}\, 
1459: \left\{\left[f_z(k) - f_{z'}(k)\right]^2 + 2
1460: f_z(k) f_{z'}(k)\left[1- \, \frac{\sin k x} 
1461: {k x}\right]\right\}\, . \label{eq3.30}
1462: \end{eqnarray}
1463: Here, $x \equiv |{\bf x} - {\bf x'}|$ is the coordinate distance between the two events
1464: and, consistently with the slow-roll approximation, the Hubble expansion rate at
1465: exit has been assumed to be scale-invariant:
1466:  $\whh_{\rm ex}(k) \simeq \whh_{\rm ex}  \simeq 3 \times 10^{-5}$ . 
1467: 
1468: If one considers spatial fluctuations over terrestrial or 
1469: solar system proper length scales $l=a_0k^{-1} \ll H_0^{-1}$ at the present time 
1470: $t=t'= t_0$, the first square brackets in (\ref{eq3.30}) vanishes and one can expand the 
1471: sine function at small $kx$  
1472: obtaining 
1473: \be \label{eq3.28}
1474: \Delta^{\rm fluc} \varphi |_{l;\, z=0} \, \simeq  \frac{\whh_{\rm ex}}{2\pi} \, 
1475: \, \frac{H_0\, l}{\sqrt{3}} \qquad ; \, 
1476: \left. \frac{\Delta^{\rm fluc} e^2}{e^2}\right|_{l;\, z=0} \, \simeq \, 10^{-2} \, 
1477: \alpha_{\rm had}\, \whh_{\rm ex}\, H_0 \, l \, .
1478: \ee
1479: As expected, these variations are
1480: extremely small, ${\Delta^{\rm fluc} e^2}/{e^2}|_{l;\, z=0}\simeq 10^{-33} \, 
1481: l / {\rm km}$. 
1482: It is also interesting to compare dilaton fluctuations at different redshifts 
1483: along a comoving observer wordline. By putting $x\simeq 0$ in (\ref{eq3.30}) 
1484: the second term in the square brackets vanishes and one has:
1485: \be \label{eq3.34}
1486: \Delta^{\rm fluc} \varphi |_{z;\, x=0} \, \simeq \, \frac{\whh_{\rm ex}}{2\pi}
1487: \frac{1}{\sqrt{2}}
1488: \left[\log(1+z) - \frac{z}{1+z} + \frac{10^{-8}}{2} z^2\right]^{1/2} \simeq 
1489: \frac{\whh_{\rm ex}}{2\pi} \left[\frac{z}{2} - \frac{z^2}{3} + \dots\right] \, .
1490: \ee
1491: 
1492: It is slightly  more complicated  to compare dilaton fluctuations
1493:  between ``now'' and events 
1494: at redshift $z$ along a null ray. Expanding in powers of $z$ around $z=0$  
1495: one gets from (\ref{eq3.30}), after a straightforward calculation:
1496: \be \label{eq3.29} 
1497: \Delta^{\rm fluc} \varphi |_z  \,  \simeq \, \frac{\whh_{\rm ex}}{2\pi}\, 
1498: \left[\frac{1}{2}\sqrt{\frac{7}{3}}\, z - \frac{2}{\sqrt{21}}\, z^2 + \dots \, \right] .
1499: \ee
1500: Numerically, at redshift $z \sim 1$, the effects of dilatonic fluctuations 
1501: give  $\Delta^{\rm fluc}  \varphi |_{z=1} \sim \whh_{\rm ex}/(2 \pi) \sim 5 \times 10^{-6}$.
1502:  This is to be
1503: contrasted with the effects of the cosmic, 
1504: homogeneous evolution which yields $\Delta  \varphi |_{z=1} \simeq \alpha_m$.
1505: In the ``normal'' case where $\alpha_m \sim e^{-c \varphi} \sim \delta_H^{4/(n+2)}$,
1506: the two effects, though a priori unrelated, are related  in our  scenario , when $n=2$.
1507: Indeed, if $n=2$,  $\delta_H^{4/(n+2)} = \delta_H \sim 5 \times 10^{-5}$ is linked to
1508: $\whh_{\rm ex}/(2 \pi)$ via $ \delta(\chi_{\rm ex}) = A \whh_{\rm ex}/(2 \pi)$
1509: with $A = (8/3) V/ \partial_{\chi} V = (8/3) (\chi/n) \simeq 40/( 3 \sqrt{n}) \sim 10$.
1510: On the other hand, in the case where $\varphi$ is strongly coupled to dark matter, the
1511: homogeneous evolution $\Delta  \varphi |_{z=1} \simeq \alpha_m \sim 1$ is parametrically
1512: larger than the fluctuations $\Delta^{\rm fluc}  \varphi |_{z=1} \sim \whh_{\rm ex}/(2 \pi)$.
1513:  
1514: To conclude on this subsection, we see that the inhomogeneous space-time fluctuations
1515: of the fine-structure constant are typically too small to be observable (if the limits
1516: from UFF are already satisfied), being suppressed, relative to their natural values 
1517: $H_0 l\,, \, H_0 t$, by the  small factor $\alpha_{\rm had} \whh_{\rm ex} $. 
1518: 
1519: 
1520: \section{Summary and conclusion}
1521: We have studied the dilaton-fixing mechanism of \cite{DP94} within the context where the 
1522: dilaton-dependent low-energy couplings are extremized at $\varphi = + \infty$, i.e. for 
1523: infinitely large values of the bare string coupling $g_s^2 = e^{\phi} \simeq e^{c \varphi}$. 
1524: [The crucial coupling to the inflaton, say $\lambda(\varphi)$ in Eq.(\ref{eq2.10}), must be
1525: {\it minimized} at $ \varphi \to + \infty$; the other couplings can be either minimized
1526: or maximized there.]
1527: This possibility of a fixed point at infinity (in bare string coupling space) has been 
1528: recently suggested \cite{V01}, and its late-cosmological consequences have been explored in 
1529: \cite{GPV}. We found that a primordial inflationary stage, with inflaton potential $V(\chi) = 
1530: \lambda (\varphi) \, \chi^n / n$, was much less efficient in decoupling a dilaton with least 
1531: couplings at infinity than in the case where the least couplings are reached at a finite value 
1532: of $\varphi$ (as in \cite{DP94,DV96}). This reduced efficiency has interesting 
1533: phenomenological consequences. Indeed, it predicts much larger observable deviations from 
1534: general relativity. In the case of the simplest chaotic potential \cite{L90} $V(\chi) = 
1535: \frac{1}{2} \, m_{\chi}^2 (\varphi) \, \chi^2$, we find that, under the simplest assumptions 
1536: about the pre-inflationary state, this scenario predicts violations of the universality of 
1537: free fall (UFF) of order $\Delta a / a \sim 5 \times 10^{-4} \, \delta_H^2$ where $\delta_H$ 
1538: is the density fluctuation generated by inflation on horizon scales. The observed level of 
1539: large-scale density (and cosmic microwave background temperature) fluctuations fixes 
1540: $\delta_H$ to be around $5 \times 10^{-5}$ which finally leads to a prediction for a violation 
1541: of the UFF near the $\Delta a / a \sim 10^{-12}$ level. This is naturally compatible with 
1542: present experimental tests of the equivalence principle, and suggests that a modest 
1543: improvement in the precision of UFF tests might be able to detect a deviation linked to 
1544: dilaton exchange with a coupling reduced by the attraction towards the fixed point at 
1545: infinity. Because of the presence of unknown dimensionless
1546:  ratios $(c, b_F / b_\lambda)$ in 
1547: our estimates, and of quantum noise in the evolution of the dilaton,
1548:  we cannot give sharp quantitative estimates of $\Delta a / a$. However, we note 
1549: that dilaton-induced violations of the UFF have a rather precise signature with a 
1550: composition-dependence of the form (\ref{eq3.15}), with probable domination by the last 
1551: (Coulomb energy) term \cite{DP94}. As explored in \cite{D96} this signature is quite distinct 
1552: from UFF violations induced by other fields, such as a vector field. We note that the approved 
1553: Centre National d'Etudes Spatiales (CNES) mission MICROSCOPE \cite{Touboul} (to fly in 
1554: 2004) will explore the level $\Delta a / a \sim 10^{-15}$, while the planned National 
1555: Aeronautics and Space Agency (NASA) and European Space Agency (ESA) mission STEP (Satellite 
1556: Test of the Equivalence Principle)\cite{worden}
1557:  could explore the $\Delta a / a \sim 10^{-18}$ level. 
1558: Our scenario gives additional motivation for such experiments and suggests that they might find 
1559: a rather strong violation signal, whose composition-dependence might then be studied in detail 
1560: to compare it with Eq.~(\ref{eq3.15}).
1561: 
1562: In the case of inflationary potentials $V(\chi) \propto \chi^n$ with $n > 2$ our simplest 
1563: estimates predict a violation of the UFF of order $\Delta a / a \sim 5 \times 10^{-4} \, 
1564: \delta_H^{\frac{8}{n+2}}$ which is larger than $10^{-12}$. At face value this suggests that 
1565: existing UFF experimental data can be interpreted as favouring $n \leq 2$ over $n > 2$. 
1566: However, we must remember that our estimates have made several simplifying assumptions. It is 
1567: possible that the large quantum fluctuations of the inflaton in the self-regenerating regime 
1568: $\chi > \chi_{\rm in}$, with $\chi_{\rm in}$ defined by Eq.~(\ref{eq2.19}), can give more time 
1569: for $\varphi$ to run away towards large values, so that the effective value of $e^{c 
1570: \varphi_{\rm in}}$ to be used in Eq.~(\ref{eq2.23}) turn out to dominate the first term in the 
1571: R.H.S. that we have used for our estimates. We leave to future work a study of the system of
1572: Langevin equations describing  the coupled fluctuations of $\phi$ and $\chi$ during the
1573: self-regenerating regime.
1574: 
1575: 
1576: Finally let us note some other conclusions of our work.
1577: 
1578:   We recover the conclusion of previous 
1579: works on dilaton models that the most interesting experimental probes of a massless weakly 
1580: coupled dilaton are tests of the UFF. The composition-independent gravitational tests (solar-system, 
1581: binary-pulsar) tend to be much 
1582: less sensitive probes (as highlighted by the relations (\ref{eq3.18}), (\ref{eq3.23}) and 
1583: (\ref{eq3.24})).
1584: 
1585:  However, a possible exception concerns the time-variation of the
1586: coupling constants. Here the conclusion depends crucially on the assumptions made about the
1587: couplings of the dilaton to the cosmologically dominant forms of energy (dark matter and/or
1588: dark energy). If these couplings are of order unity (and as large as is phenomenologically 
1589: acceptable, i.e. so that $(\varphi'_0)^2 = 0.7$),
1590: the present time variation of the fine-structure constant is linked to the violation
1591: of the UFF by the relation $d \ln e^2 / dt \sim \pm 2.0 \times 10^{-16}
1592: \,   \sqrt{ 10^{12} \Delta a/a} {\rm yr}^{-1}$. [The most natural sign here being $+$, 
1593: i.e. $b_F > 0$, which corresponds to {\it smaller} $e^2$ in the past, 
1594: just as suggested by the claim \cite{Webb}.]
1595: Such a time variation might be observable
1596: ( if $\Delta a/a$ is not very much below its present upper bound $\sim 10^{-12}$) through
1597: the comparison of high-accuracy cold-atom clocks and/or via improved measurements of
1598: astronomical spectra.
1599: 
1600:  More theoretical work is needed to justify the basic assumption 
1601: (\ref{eq1.3}) of our scenario. In particular, it is crucial to investigate whether it is 
1602: natural to expect that the sign of the crucial coefficient $b_{\lambda}$ in Eq.~(\ref{eq2.11}) 
1603: be indeed {\it positive}. [Recall that the general mechanism of \cite{DP94} is an attraction 
1604: towards ``Least Couplings'' while Eq.~(\ref{eq1.3}) with ${\cal O}(e^{- \phi}) > 0$ leads
1605: to largest couplings at infinity.] Note in this respect that the sign of the other $b_i$'s is not 
1606: important as, once inflation has pushed $e^{c \varphi}$ to very large values $e^{c 
1607: \varphi_{\rm end}}$, the subsequent cosmological evolutions tend to be ineffective in further 
1608: displacing $\varphi$. 
1609: \acknowledgments
1610: It is a pleasure to thank Ian Kogan for suggesting a method to solve the (decoupled) Langevin
1611: equation for $\varphi$.
1612: GV   wishes to
1613: acknowledge the support of a ``Chaire Internationale Blaise
1614: Pascal", administered by the ``Fondation de L'Ecole Normale
1615: Sup\'erieure''. The work of F.P. was supported in part
1616: by INFN and MURST under contract 2001-025492, and by the European
1617: Commission TMR program HPRN-CT-2000-00131, in association to the
1618: University of Padova."
1619: 
1620: 
1621: 
1622: 
1623: 
1624: 
1625: 
1626: 
1627: 
1628: 
1629: \appendix 
1630: 
1631: \section{The stochastic evolution of the dilaton}
1632: 
1633: In this appendix we study the stochastic evolution of the dilaton $\varphi$ during 
1634: inflation as described by the Langevin-type equation (\ref{eq4.2}). 
1635: We restrict our attention to the region of 
1636: phase space where the evolution of the inflaton
1637: $\chi$ is classical, and to a power-law potential of the form (\ref{eq2.10}). 
1638: It follows that the inflaton evolves according to the classical slow-roll equation 
1639: (\ref{eq2.13}) whose solution reads
1640: \be \label{a0} 
1641: \chi^2 = \chi_{\rm in}^2 - n p,
1642: \ee 
1643: where $p$, the 
1644: parameter defined in (\ref{eq2.7}), is shifted in such a way that $p_{\rm in} \equiv 0$. 
1645: Equation  (\ref{eq4.2}) takes the form 
1646: \begin{equation} \label{a1}
1647: \frac{d \varphi}{dp} = \frac{1}{2} \, b_{\lambda} \, c \, e^{-c\varphi} +  \,
1648: \xi (p) \, ,
1649: \end{equation} 
1650: where $\xi(p)$ is a gaussian stochastic variable (GSV), with a ``time-dependent'' r.m.s. 
1651: amplitude $\whh (p) /2\pi$:
1652: \be \label{a3}
1653: \langle\xi(p_1)\, \xi(p_2) \rangle \, \, =\, \frac{\whh^2}{(2\pi)^2}\, \delta(p_1-p_2), 
1654: \ee
1655: [the relation with the normalized random white noise term of (\ref{eq4.2}) is
1656: $\xi(p) = \xi_2(p) \whh / 2\pi$].
1657: For any given source term $\xi(p)$, the formal solution of (\ref{a1}) reads
1658: \be \label{a4}
1659: e^{c\, \varphi(p)} \, =\, e^{c\, \varphi_{\rm in}} \, e^{c\, \eta(p)} + 
1660: \frac{b_\lambda c^2}{2}
1661: \int_0^p dp' e^{c\, [\eta(p) - \eta(p')]}\, ,
1662: \qquad
1663: \eta(p) \equiv \int_0^p dp'\, \xi(p')\, . 
1664: \ee
1665: Note that the classical solution in (\ref{eq2.16}), 
1666: $e^{c\, \varphi_{\rm cl}(p)} = e^{c\, \varphi_{\rm in}} + (b_\lambda c^2 /2)\,  p$ ,
1667: can be easily recovered in the 
1668: small noise limit $\xi(p) \rightarrow 0$, $\eta(p) \rightarrow 0$. 
1669: 
1670: It proves convenient to compare the true solution
1671:  to the classical one by studying the statistical 
1672: behaviour of the ratio $A(p) \equiv e^{c\, \varphi(p)}/ e^{c\, \varphi_{\rm cl}(p)}$.
1673: As we will show below, $\langle e^{c\, \eta(p)} \rangle  = {\cal O}(1)$. Moreover, we 
1674: are also assuming $e^{c\, \varphi_{\rm in}} = {\cal O}(1)$ or, at least,
1675: $e^{c\, \varphi_{\rm in}} \ll (b_\lambda c^2/2) p$ (see Section II for details)
1676: so that the leading contribution to the first equation in (\ref{a4}) is given by the integral, 
1677: and we have
1678: \be\label{a4'}
1679: A(p) \, \equiv \, e^{c\, \varphi(p)}/ e^{c\, \varphi_{\rm cl}(p)} \, \simeq \, 
1680: \frac{1}{p} 
1681: \int_0^p dp' e^{c\, [\eta(p) - \eta(p')]}\, .
1682: \ee 
1683: Since $\xi(p)$ is a GSV, also its integral $\eta(p)$ is a (centered) GSV. Moreover,
1684: if $x$ is a GSV with $\sigma_x^2 \, \equiv \, \langle x^2 \rangle  - \langle x \rangle ^2 \, = \, \langle x^2 \rangle $, 
1685: by Bloch's theorem $y= e^x$ is a new stochastic variable with 
1686: $\langle y \rangle  \, = \, \langle e^x \rangle  \, = \, e^{\langle x^2 \rangle /2}$ and 
1687: $\sigma_y^2 \, = \, e^{2 \langle x^2 \rangle } - e^{\langle x^2 \rangle }$.
1688: The average value of $A(p)$ thus reads 
1689: \be 
1690: \label{a5}
1691: \langle A(p) \rangle   \, \, \simeq \,
1692: \frac{1}{p} \int_0^p dp' \, e^{(c^2/2)\, \langle[\eta(p) - \eta(p')]^2 \rangle }\,. 
1693: \ee
1694: The exponent on the right hand side of the above equation can be estimated by 
1695: using (\ref{a3}) and the slow-roll approximation  
1696: $\whh^2 \simeq 2 V(\chi,\varphi)/3 = 2 \lambda(\varphi) \chi^n /3n \simeq 2 
1697: \lambda_\infty \chi^n/3n $. One gets
1698: \be \label{a6}
1699: \langle[\eta(p)-\eta(p')]^2 \rangle \, \, =\, 
1700: \frac{1}{(2\pi)^2} \int_{p'}^{p} \whh^2 dp''\, \simeq \, \frac{n}{2(n+2)}
1701: \left[\left(\frac{\chi(p')}{\chi_{\rm in}}\right)^{n+2}\!\!\! -\left( 
1702: \frac{\chi(p)}{\chi_{\rm in}}\right)^{n+2} \right]\, ,
1703: \ee
1704: where $\chi_{\rm in}$ is the  value at exit from self-regenerating inflation: 
1705: $\whh(\chi_{\rm in})/2\pi = n /(2\chi_{\rm in})$ (see Section II for more details).
1706: Since we are interested in evaluating (\ref{a5}) at the end of inflation, 
1707: $p = p_{\rm end} \simeq \chi_{\rm in}^2/n$, we can thus write
1708: \be \label{a8}
1709: \langle[\eta(p)-\eta(p')]^2 \rangle \, \, \simeq\, 
1710: \frac{n}{2(n+2)} \left[1-\frac{p'}{p_{\rm end}}\, \right]^{(n+2)/2}.
1711: \ee
1712: When evaluated at $p'=0$, the above formula gives  
1713: $\langle\eta(p)^2 \rangle  = n/(2(n+2))$. Thus the normalization factor to the initial condition in 
1714: (\ref{a4}) is of order one, as anticipated: $\langle e^{c\, \eta(p)} \rangle  =  e^{ \frac{1}{2} c \langle 
1715: \eta(p)^2 \rangle} = {\cal O}(1)$.
1716: {}From (\ref{a5}) and (\ref{a8}) we have:
1717: \begin{eqnarray} \nonumber
1718: \langle A(p_{\rm end})  \rangle  \, 
1719: &\simeq &\, \,  \frac{1}{p_{\rm end}} 
1720: \int_0^{p_{\rm end}} dp' \exp\left[\frac{c^2 n}{4(n+2)}\left(1-\frac{p'}{p_{\rm end}}
1721: \right)^{(n+2)/2}\right]   \\
1722: & = & \, 
1723: \int_0^1 \, \exp\left[\frac{c^2 n}{4(n+2)} 
1724: x^{(n+2)/2}\right] dx  \\ \nonumber
1725: & = & \,  \exp\left(\frac{c^2 n \theta}{4(n+2)}\right) = {\cal O}(1) ,
1726: \end{eqnarray}
1727: with $0 < \theta < 1$.
1728: 
1729:  We can estimate the dispersion of the same quantity by  expanding
1730: the exponential inside the integral (\ref{a5}) 
1731: in powers of $\xi(p)$:
1732: \be
1733: A(p) \, \simeq \, 
1734: 1 + \frac{c}{p} \int_0^p dp' \int_{p'}^p dp'' \xi(p'') + \frac{1}{2} \frac{c^2}{p} 
1735: \int_0^p dp'\left(\int_{p'}^p dp''\xi(p'')\right)^2 + \dots .
1736: \ee
1737: At lowest order in $\xi(p)$ the variance of the above quantity calculated at 
1738: $p = p_{\rm end}$ reads
1739: \begin{eqnarray} \nonumber
1740: \sigma^2_{A(p_{\rm end})}\,  & = &
1741: \left\langle \left(\frac{c}{p} \int_0^p dp' \int_{p'}^p dp'' \xi(p'')\right)^2\right \rangle \\
1742: &=& \frac{c^2}{p^2}  \int_0^p dp'  \int_0^p dp''\int_{{\rm max}(p',p'')}^p dp''' 
1743: \frac{\whh^2(p''')}{(2\pi)^2} \\ \nonumber
1744: &=& \frac{2 c^2}{p^2} \int_0^p dp' p' \int_{p'}^p dp'' \frac{\whh^2(p'')}{(2\pi)^2} .
1745: \end{eqnarray}
1746: As in equation (\ref{a6}) we can use the slow-roll approximation and obtain 
1747: \be
1748: \sigma^2_{A(p_{\rm end})}\  =  \
1749: \frac{c^2\, n}{n+2}\int_0^1 x\, (1-x)^{(n+2)/2} dx\ 
1750: = \ c^2 {\cal O}(1) \, . \nonumber
1751: \ee
1752: 
1753: 
1754: 
1755: 
1756: \begin{references}
1757: \bibitem{Witten}  E. Witten, {\it Phys. Lett. B} {\bf 149}, 351 (1984).
1758: \bibitem{TV88} T.R. Taylor and G. Veneziano, {\it Phys. Lett. B} {\bf 213}, 450 (1988).
1759: \bibitem{DP94} T. Damour and A.M. Polyakov, {\it Nucl. Phys. B} {\bf 423}, 532 (1994); {\it 
1760: Gen. Rel. Grav.} {\bf 26}, 1171 (1994).
1761: \bibitem{DN93} T. Damour and K. Nordtvedt, {\it Phys. Rev. Lett.} {\bf 70}, 2217 (1993); {\it 
1762: Phys. Rev. D} {\bf 48}, 3436 (1993).
1763: \bibitem{V01} G. Veneziano, {\it J. High Energy Phys.} {\bf 06}, 051 (2002).  
1764: 
1765: \bibitem{GPV} M. Gasperini, F. Piazza and G. Veneziano, {\it Phys. Rev. D} {\bf 65}, 023508
1766:  (2002).
1767: \bibitem{prl}  T. Damour, F. Piazza and G. Veneziano, {\it Phys. Rev. Lett.} {\bf 89}, 081601 (2002).
1768: \bibitem{DV96} T. Damour and A. Vilenkin, {\it Phys. Rev. D} {\bf 53}, 2981 (1996).
1769: \bibitem{D95} T. Damour, in {\it Susy 95}, Proceedings of the 1995 International Workshop on 
1770: Supersymmetry and Unification of Fundamental Interactions, edited by I.~Antoniadis and 
1771: H.~Videau (Editions Fronti\`eres, Gif-sur-Yvette, 1996), pp.~577-584.
1772: \bibitem{L90} A. Linde, {\it Particle Physics and Inflationary Cosmology}, (Harwood, Chur, 
1773: 1990).
1774: \bibitem{DGG} T. Damour, G.W. Gibbons and C. Gundlach, {\it Phys. Rev. Lett.} {\bf 64}, 123 
1775: (1990).
1776: \bibitem{AT} L. Amendola and D. Tocchini-Valentini, {\it Phys. Rev. D} {\bf 64}, 043509 (2001)
1777: \bibitem{AGUT} L. Amendola, M. Gasperini, C. Ungarelli and D. Tocchini-Valentini, 
1778: in preparation.
1779: \bibitem{SNI} S. Perlmutter et al., {\it Astrophys. J.} {\bf 517}, 565 (1999);
1780: A. Riess et al. {\it Astronom. J.} {\bf 116}, 1009 (1998).
1781: \bibitem{omegam} See, e.g., the review of global cosmological parameters (chap. 17) in
1782:  D.E. Groom et al., {\it European Physical Journal} {\bf C15}, 1 (2000); updated version
1783:  available on http://pdg.lbl.gov/.
1784: \bibitem{exp} D.E. Groom et al., {\it European Physical Journal} {\bf C15}, 1 (2000),
1785:  see Chapter~14 ``Experimental tests of gravitational theory'',
1786:  available on http://pdg.lbl.gov/; 
1787: and C.M.~Will, {\it Living Rev. Rel.} {\bf 4}, 4 (2001).
1788: \bibitem{DEF92} T. Damour and G. Esposito-Far\`ese, {\it Class. Quant. Grav.} {\bf 9}, 
1789: 2093 (1992).
1790: \bibitem{DEF} T. Damour and G. Esposito-Far\`ese, {\it Phys. Rev. D} {\bf 54}, 1474 (1996); 
1791: {\it Phys. Rev. D} {\bf 58}, 042001 (1998).
1792: \bibitem{D96} T. Damour, {\it Class. Quant. Grav.} {\bf 13}, A33 (1996).
1793: \bibitem{Su94} Y. Su et al., {\it Phys. Rev. D} {\bf 50}, 3614 (1994).
1794: \bibitem{Oklo} A.I. Shlyakhter, {\it Nature} {\bf 264}, 340 (1976); T. Damour and F. Dyson, 
1795: {\it Nucl. Phys. B} {\bf 480}, 37 (1996).
1796: \bibitem{OPQCCV}  K. A. Olive, M. Pospelov, Y. Z. Qian, A. Coc, M. Casse and 
1797: E. Vangioni-Flam, \emph{Constraints on the Variations of the Fundamental Couplings}, 
1798: hep-ph/0205269
1799: \bibitem{Gaspe} M. Gasperini, {\it Phys. Rev. D} {\bf 64} 043510 (2001). 
1800: \bibitem{Webb} J.K. Webb et al., {\it Phys. Rev. Lett.} {\bf 87}, 091301 (2001).
1801: \bibitem{SBM} H. B. Sandvik, J. D. Barrow and J. Magueijo, {\it Phys. Rev. Lett.} {\bf 88} 031302 (2002).
1802: \bibitem{OP} K. A. Olive and M. Pospelov,  {\it Phys. Rev. D} {\bf 65}, 85044 (2002).
1803: \bibitem{lyth} A. R. Liddle and D. H. Lyth, {\it Cosmological inflation and large scale structure}, 
1804: (Cambridge University Press, 2000).
1805: \bibitem{salomon}C. Salomon et al., in {\it Cold atom clocks on Earth and in space,  
1806: Proceedings of the 17$^{th}$ Int. Conf. on Atomic Physics}, E. Arimondo, M.
1807: Inguscio, ed., (World Scientific, Singapore, 2001), p 23;
1808: Y. Sortais et al., {\it Physica Scripta}, {\bf 95}, 50 (2001). 
1809: \bibitem{prestage} J. D. Prestage, R. L. Tjoelker, and L. Maleki, {\it
1810: Phys. Rev. Lett.}
1811: {\bf 74} 3511 (1995).
1812: \bibitem{Touboul} P. Touboul et al., {\it C.R. Acad. Sci. Paris} {\bf 2} (s\'erie IV) 1271 
1813: (2001).
1814: \bibitem{worden} P. W. Worden, in {\it Proceedings of the Seventh Marcel Grossmann Meeting
1815: on General Relativity}, edited by R. T. Jantzen and G. Mac Keiser (World Scientific, Singapore,
1816: 1996), pp 1569-1573.
1817: \end{references}
1818: 
1819: \end{document}
1820: 
1821: 
1822: 
1823: 
1824: 
1825: 
1826: 
1827: