1: \documentclass[aps,showpacs,floatfix]{revtex4}
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: \usepackage{amssymb}
4: \usepackage{amsmath}
5: \usepackage{epsfig}
6: \usepackage{subfigure}
7:
8: \setcounter{MaxMatrixCols}{10}
9: %TCIDATA{OutputFilter=Latex.dll}
10: %TCIDATA{Version=4.00.0.2312}
11: %TCIDATA{LastRevised=Tuesday, September 13, 2005 15:10:50}
12: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
13: %TCIDATA{Language=American English}
14:
15: \begin{document}
16:
17: \title{The Power of General Relativity }
18: \author{Timothy Clifton}
19: \email{T.Clifton@damtp.cam.ac.uk}
20: \affiliation{DAMTP, Centre for Mathematical Sciences, University of Cambridge,
21: Wilberforce Road, Cambridge, CB3 0WA, UK}
22: \author{John D. Barrow}
23: \email{J.D.Barrow@damtp.cam.ac.uk}
24: \affiliation{DAMTP, Centre for Mathematical Sciences, University of Cambridge,
25: Wilberforce Road, Cambridge, CB3 0WA, UK}
26: \date{\today }
27: \pacs{95.30.Sf, 98.80.Jk, 04.80.Cc, 98.80.Bp, 98.80.Ft, 95.10.Eg}
28:
29: \begin{abstract}
30: We study the cosmological and weak-field properties of theories of gravity
31: derived by extending general relativity by means of a Lagrangian
32: proportional to $R^{1+\delta }$. This scale-free extension reduces to
33: general relativity when $\delta \rightarrow 0$. In order to constrain
34: generalisations of general relativity of this power class we analyse the
35: behaviour of the perfect-fluid Friedmann universes and isolate the
36: physically relevant models of zero curvature. A stable matter-dominated
37: period of evolution requires $\delta >0$ or $\delta <-1/4$. The stable
38: attractors of the evolution are found. By considering the synthesis of light
39: elements (helium-4, deuterium and lithium-7) we obtain the bound $%
40: -0.017<\delta <0.0012.$ We evaluate the effect on the power spectrum of
41: clustering via the shift in the epoch of matter-radiation equality. The
42: horizon size at matter--radiation equality will be shifted by $\sim 1\%$ for
43: a value of $\delta \sim 0.0005.$ We study the stable extensions of the
44: Schwarzschild solution in these theories and calculate the timelike and null
45: geodesics. No significant bounds arise from null geodesic effects but the
46: perihelion precession observations lead to the strong bound $\delta =2.7\pm
47: 4.5\times 10^{-19}$ assuming that Mercury follows a timelike geodesic. The
48: combination of these observational constraints leads to the overall bound $%
49: 0\leq \delta <7.2\times 10^{-19}$ on theories of this type.
50: \end{abstract}
51:
52: \maketitle
53:
54: \section{Introduction}
55:
56: There is a long history of considering generalisations of Einstein's theory
57: of general relativity which reduce to general relativity in the weak gravity
58: limit when the spacetime curvature, $R$, becomes small. Typically, these
59: studies consider a gravitational Lagrangian which augments the linear
60: Einstein-Hilbert Lagrangian by the addition of terms of quadratic or higher
61: order in $R,$ first considered by Eddington \cite{edd}; these additions may
62: also include terms in $\ln R$, \cite{gur}. More general extensions of
63: general relativity in this spirit have considered the structure of theories
64: derived from gravitational Lagrangians that are general analytic functions
65: of $R$, \cite{buch, kerner, BO, magg}. These choices produce theories which
66: can look like general relativity plus small polynomial corrections in the
67: appropriate limiting situations as $R$ becomes small. There has also been
68: interest in theories with corrections to general relativity that are $%
69: O(R^{-1})$ because of their scope to introduce cosmological deviations from
70: general relativity at late times which might mimic the effects of dark
71: energy on the Hubble flow \cite{tur, new}. We also know that theories derived
72: from a Lagrangian that is an analytic function of $R$ have an important
73: conformal relationship to general relativity with scalar field sources so
74: long as the trace of the energy-momentum tensor vanishes in the higher-order
75: gravity theory \cite{CB, maeda}. All these theories introduce corrections to
76: general relativity which come with a characteristic length scale that is
77: determined by the new coupling constant that couples the higher-order terms
78: to the Einstein-Hilbert part of the Lagrangian. In general, these theories
79: are mathematically complicated with 4$^{th}$-order field equations that can
80: exhibit singular perturbation behaviour unless care is taken to ensure that
81: the stationary action does not become maximal rather than minimal \cite{ruz,
82: BO}. and there are few interesting exact solutions other than those of
83: general relativity, which are particular solutions in vacuum and for
84: trace-free energy momentum tensor so long as the cosmological constant
85: vanishes \cite{BO}.
86:
87: In this paper we are going to consider a different type of generalisation of
88: Einstein's general relativity, in which no new scale is introduced. The
89: Lagrangian is proportional to $R^{n}$, and so general relativity is
90: recovered in the $n\rightarrow 1$ limit, from above or below. Particular
91: cases have been studied by Buchdahl \cite{Buc70} and Roxburgh \cite{rox}.
92: This gravitation theory has many appealing properties and, unlike
93: other higher-order gravity theories, admits simple
94: exact solutions for Friedmann cosmological models and exact static
95: spherically symmetric solutions which generalise the Schwarzschild
96: metric. As well as allowing comparison with observation these
97: solutions also provide an interesting testing
98: ground for new developments in gravitation theory such as particle production,
99: holography and gravitational thermodynamics. Furthermore,
100: this theory is of additional interest because it permits a very general
101: investigation of the nature of its behaviour in the vicinity of a
102: cosmological singularity which brings the behaviour of general
103: relativity into sharper focus. In another paper \cite{Bar05}, we show
104: that the counterpart of the Kasner anisotropic vacuum cosmology can be
105: found exactly and strong conclusions drawn about the presence or absence
106: of the chaotic behaviour found in the Mixmaster universe.
107:
108: The structure of this paper is as follows; in the next section we present
109: the gravitational action and field equations for the theory of gravity we
110: will be considering. A conformal relationship with general relativity
111: containing a scalar field in a Liouville (exponential) potential is then
112: outlined and the Newtonian limit of the field equations is investigated. The
113: rest of the paper is then split into two sections; the first investigates
114: the cosmology of the theory and the second investigates the static and
115: spherically symmetric weak field - in both cases our goal is to calculate
116: predictions for physical processes, the results of which can be compared
117: with observation. We use observational data from cosmology and the standard
118: solar system tests of general relativity to bound the allowed values of $n$,
119: the single defining parameter of the theory.
120:
121: In the cosmology section we consider Friedmann--Robertson--Walker universes.
122: We present the equivalent of the Friedmann equations, in this theory, and
123: find some power--law exact solutions. A dynamical systems approach is then
124: used to show the extent to which these solutions can be considered as
125: attractors of spatially flat universes at late times. After showing the
126: attractive properties of these solutions (with certain exceptions) we
127: proceed to predict the results of primordial nucleosynthesis and the form of
128: the power spectrum of perturbations in this theory. These predictions are
129: then compared to observation and used to constrain deviations from general
130: relativity.
131:
132: The static and spherically symmetric weak-field analysis follows. We present
133: the field equations and find the physically relevant exact solution to them.
134: A dynamical systems approach is then used to find the asymptotic attractor
135: of the general solution at large distances. This asymptotic form is then
136: perturbed and the linearised field equations are found and solved. The exact
137: solution is shown to be the relevant solution in this limit, when
138: oscillatory modes in the perturbed metric functions are set to zero. We find
139: the null and time-like geodesics for this spacetime to Newtonian and
140: post-Newtonian order. Predictions are then made for the outcomes of the
141: classical tests of general relativity in this theory; namely the bending of
142: light, the time-delay of radio signals and the perihelion precession of
143: Mercury. These predictions are then compared to observation and again used
144: to constrain deviations from general relativity.
145:
146: \section{Field equations}
147:
148: We consider here a gravitational theory derived from the Lagrangian density
149: \begin{equation}
150: \mathcal{L}_{G}=\frac{1}{\chi }\sqrt{-g}R^{1+\delta }, \label{action}
151: \end{equation}%
152: where $\delta $ is a real number and $\chi $ is a constant. The limit $%
153: \delta \rightarrow 0$ gives us the familiar Einstein--Hilbert Lagrangian of
154: general relativity and we are interested in the observational consequences
155: of $\left\vert \delta \right\vert $ $>0$.
156:
157: We denote the matter action as $S_{m}$ and ignore the boundary term.
158: Extremizing
159: \begin{equation*}
160: S=\int \mathcal{L}_{G}d^{4}x+S_{m},
161: \end{equation*}%
162: with respect to the metric $g_{ab}$ then gives \cite{Buc70}
163: \begin{multline}
164: \delta (1-\delta ^{2})R^{\delta }\frac{R_{,a}R_{,b}}{R^{2}}-\delta (1+\delta
165: )R^{\delta }\frac{R_{;ab}}{R}+(1+\delta )R^{\delta }R_{ab}-\frac{1}{2}%
166: g_{ab}RR^{\delta } \label{field} \\
167: -g_{ab}\delta (1-\delta ^{2})R^{\delta }\frac{R_{,c}R_{,}^{\ c}}{R^{2}}%
168: +\delta (1+\delta )g_{ab}R^{\delta }\frac{\Box R}{R}=\frac{\chi }{2}T_{ab},
169: \end{multline}%
170: where $T_{ab}$ is the energy--momentum tensor of the matter, and is defined
171: in terms of $S_{m}$ and $g_{ab}$ in the usual way. We take the quantity $%
172: R^{\delta }$ to be the positive real root of $R$ throughout this paper.
173:
174: \subsection{Conformal equivalence to general relativity}
175:
176: Rescaling the metric by the conformal factor $\Omega (r)=\Omega
177: _{0}R^{\delta }$ the vacuum field equations (\ref{field}) become
178: \begin{equation*}
179: \bar{G}_{ab}=\frac{3\delta ^{2}}{2}\frac{R_{,a}R_{,b}}{R^{2}}-\frac{3\delta
180: ^{2}}{4}\bar{g}_{ab}\bar{g}^{cd}\frac{R_{,c}R_{,d}}{R^{2}}-\frac{\delta }{%
181: 2(1+\delta )}\frac{\bar{g}_{ab}}{\Omega _{0}}\frac{R}{R^{\delta }},
182: \end{equation*}%
183: where $\bar{g}_{ab}=\Omega g_{ab}$ and other quantities with overbars are
184: constructed from the rescaled metric $\bar{g}_{ab}$.
185:
186: Making the definition of a scalar field
187:
188: \begin{equation*}
189: \phi \equiv \sqrt{\frac{3}{16\pi G}}\ln R^{\delta },
190: \end{equation*}
191:
192: these equations can be rewritten as
193: \begin{equation}
194: \bar{G}_{ab}=8\pi G\left( \phi _{,a}\phi _{,b}-\frac{1}{2}\bar{g}_{ab}(\bar{g%
195: }^{cd}\phi _{,c}\phi _{,d}+2 V(\phi ))\right) \label{conformalfield}
196: \end{equation}%
197: and
198: \begin{equation*}
199: \square \phi =\frac{dV}{d\phi },
200: \end{equation*}%
201: where $V(\phi )$ is given by
202: \begin{equation}
203: V(\phi )=\frac{\delta \ \text{sign}(R)}{16\pi G(1+\delta )\Omega _{0}}\exp
204: \left\{ {\sqrt{\frac{16\pi G}{3}}\frac{(1-\delta )}{\delta }\phi }\right\} .
205: \label{pot1}
206: \end{equation}%
207: The magnitude of the quantity $\Omega _{0}$ is not physically important and
208: simply corresponds to the rescaling of the metric by a constant quantity,
209: which can be absorbed by an appropriate rescaling of units. It is, however,
210: important to ensure that $\Omega _{0}>0$ in order to maintain the +2
211: signature of the metric. This result is a particular example of the general
212: conformal equivalence to general relativity plus a scalar field for
213: Lagrangians of the form $f(R)$, where $f$ is an analytic function found in
214: refs. \cite{CB, maeda}.
215:
216: \subsection{The Newtonian Limit}
217:
218: By comparing the geodesic equation to Newton's gravitational force law it
219: can be seen that, as usual,
220: \begin{equation}
221: \Gamma _{00}^{\mu }=\Phi _{,\mu } \label{ChrN}
222: \end{equation}%
223: where $\Phi $ is the Newtonian gravitational potential. All the other
224: Christoffel symbols have $\Gamma _{\;bc}^{a}=0$, to the required order of
225: accuracy.
226:
227: We now seek an approximation to the field equations (\ref{field}) that is of
228: the form of Poisson's equation; this will allow us to fix the constant $\chi
229: $. Constructing the components of the Riemann tensor from (\ref{ChrN}) we
230: obtain the standard results
231: \begin{equation}
232: R_{\;0\nu 0}^{\mu }=\frac{\partial ^{2}\Phi }{\partial x^{\mu }\partial
233: x^{\nu }}\qquad \text{and}\qquad R_{00}=\nabla ^{2}\Phi . \label{R00}
234: \end{equation}%
235: The $00$ component of the field equations (\ref{field}) can now be written
236: \begin{equation}
237: (1+\delta )R_{00}-\frac{1}{2}g_{00}R=\frac{\chi }{2}\frac{T_{00}}{R^{\delta }%
238: } \label{approx}
239: \end{equation}%
240: where terms containing derivatives of $R$ have been discarded as they will
241: contain third and fourth derivatives of $\Phi $, which will have no
242: counterparts in Poisson's equation. Subtracting the trace of equation (\ref%
243: {approx}) gives
244: \begin{equation}
245: (1+\delta )R_{00}=\frac{\chi }{2R^{\delta }}\left( T_{00}-\frac{1}{%
246: 2(1-\delta )}g_{00}T\right) \label{approx2}
247: \end{equation}%
248: where $T$ is the trace of the stress--energy tensor. Assuming a
249: perfect--fluid form for $T$ we should have, to first--order,
250: \begin{equation}
251: T_{00}\simeq \rho \qquad \text{and}\qquad T\simeq 3p-\rho \simeq -\rho .
252: \label{T00}
253: \end{equation}%
254: Substituting (\ref{T00}) and (\ref{R00}) into (\ref{approx2}) gives
255: \begin{equation*}
256: \nabla ^{2}\Phi \simeq \frac{\chi (1-2\delta )}{4(1-\delta ^{2})}\frac{\rho
257: }{R^{\delta }}.
258: \end{equation*}%
259: Comparison of this expression with Poisson's equation allows one to read off
260: \begin{equation}
261: \chi =16\pi G\frac{(1-\delta ^{2})}{(1-2\delta )}R_{0}^{\delta } \label{chi}
262: \end{equation}%
263: where $R_{0}$ is the value of the Ricci tensor at the time $G$ is measured.
264: It can be seen that the Newtonian limit of the field equations (\ref{field})
265: does not reduce to the usual relation $\nabla ^{2}\Phi \propto \rho $, but
266: instead contains an extra factor of $R^{\delta }$. This can be interpreted
267: as being the space--time dependence of Newton's constant, in this theory.
268: Such a dependence should be expected as the Lagrangian (\ref{action}) can be
269: shown to be equivalent to a scalar--tensor theory, after an appropriate
270: Legendre transformation \footnote{This equivalence to
271: scalar-tensor theories should not be taken to imply that bounds on the
272: Brans-Dicke parameter $\omega$ are immediately applicable to this
273: theory. It can be shown that a potential for the scalar-field can
274: have a non-trivial effect on the resulting phenomenology of the theory
275: \cite{Olmo}. Furthermore, the form of the perturbation to general
276: relativity that we are considering does not allow an expansion of the
277: corresponding scalar field of the form $\phi_0+\phi_1$ where $\phi_0$ is constant
278: and $\vert \phi_1 \vert << \vert \phi_0 \vert$, so that any
279: constraints obtained in a weak-field expansion of this sort cannot be
280: applied to this situation.} (see e.g, \cite{Mag94}). This type of Newtonian
281: gravity theory admits a range of simple exact solutions in the case where
282: the effective value of $G$ is a power-law in time even though the theory
283: is non-conservative and there is no longer an energy integral
284: \cite{newt}.
285:
286: \section{Cosmology}
287:
288: In this paper we will be concerned with the idealised homogeneous and
289: isotropic space--times described by the Friedmann--Robertson--Walker metric
290: with curvature parameter $\kappa $:
291: \begin{equation}
292: ds^{2}=-dt^{2}+a^{2}(t)\left( \frac{dr^{2}}{(1-\kappa r^{2})}+r^{2}d\theta
293: ^{2}+r^{2}\sin ^{2}\theta d\phi ^{2}\right) . \label{FRW}
294: \end{equation}%
295: Substituting this metric ansatz into the field equations (\ref{field}), and
296: assuming the universe to be filled with a perfect fluid of pressure $p$ and
297: density $\rho $, gives the generalised version of the Friedmann equations
298: \begin{align}
299: (1-\delta )R^{1+\delta }+3\delta (1+\delta )R^{\delta }\left( \frac{\ddot{R}%
300: }{R}+3\frac{\dot{a}}{a}\frac{\dot{R}}{R}\right) -3\delta (1-\delta
301: ^{2})R^{\delta }\frac{\dot{R}^{2}}{R^{2}}& =\frac{\chi }{2}(\rho -3p)
302: \label{Friedman1} \\
303: -3\frac{\ddot{a}}{a}(1+\delta )R^{\delta }+\frac{R^{1+\delta }}{2}+3\delta
304: (1+\delta )\frac{\dot{a}}{a}\frac{\dot{R}}{R}R^{\delta }& =\frac{\chi }{2}%
305: \rho
306: \label{Friedman2}
307: \end{align}%
308: where, as usual,
309: \begin{equation}
310: R=6\frac{\ddot{a}}{a}+6\frac{\dot{a}^{2}}{a^{2}}+6\frac{\kappa }{a^{2}}.
311: \label{R2}
312: \end{equation}%
313: It can be seen that in the limit $\delta \rightarrow 0$ these equations
314: reduce to the standard Friedmann equations of general relativity. A study of
315: the vacuum solutions to these equations for all $\kappa $ has been made by
316: Schmidt, see the review \cite{schmidt} and a qualitative study of the
317: perfect-fluid evolution for all $\kappa $ has been made by Carloni et al
318: \cite{Car04}. Various conclusions are also immediate from the general
319: analysis of $f(R)$ Lagrangians made in ref \cite{BO} by specialising them to
320: the case $f=R^{1+\delta }$. In what follows we shall be interested in
321: extracting the physically relevant aspects of the general evolution so that
322: observational bounds can be placed on the allowed values of $\delta $.
323:
324: Assuming a perfect-fluid equation of state of the form $p=\omega \rho $
325: gives the usual conservation equation $\rho \propto a^{-3(\omega +1)}$.
326: Substituting this into equations (\ref{Friedman1}) and (\ref{Friedman2}),
327: with $\kappa =0$, gives the power--law exact Friedmann solution for $\omega
328: \neq -1$
329: \begin{equation}
330: a(t)=t^{\frac{2(1+\delta )}{3(1+\omega )}} \label{power}
331: \end{equation}%
332: where
333: \begin{equation}
334: (1-2\delta )(2-3\delta (1+\omega )-2\delta ^{2}(4+3\omega ))=12\pi
335: G(1-\delta )(1+\omega )^{2}\rho _{c} \label{rhoc}
336: \end{equation}%
337: and $\rho _{c}$ is the critical density of the universe.
338:
339: Alternatively, if $\omega =-1$, there exists the de Sitter solution
340: \begin{equation*}
341: a(t)=e^{nt}
342: \end{equation*}%
343: where
344:
345: \begin{equation*}
346: 3(1-2\delta )n^{2}=8\pi G(1-\delta )\rho _{c}.
347: \end{equation*}
348:
349: The critical density (\ref{rhoc}) is shown graphically, in figure \ref%
350: {density}, in terms of the density parameter $\Omega _{0}=\frac{8\pi G\rho
351: _{c}}{3H_{0}^{2}}$ as a function of $\delta $ for pressureless dust ($\omega
352: =0$) and black-body radiation ($\omega =1/3$). It can be seen from the graph
353: that the density of matter required for a flat universe is dramatically
354: reduced for positive $\delta $, or large negative $\delta $. In order for
355: the critical density to correspond to a positive matter density we require $%
356: \delta $ to lie in the range
357: \begin{equation}
358: -\frac{\sqrt{73+66\omega +9\omega ^{2}}+3(1+\omega )}{4(4+3\omega )}<\delta <%
359: \frac{\sqrt{73+66\omega +9\omega ^{2}}-3(1+\omega )}{4(4+3\omega )}.
360: \label{range}
361: \end{equation}
362:
363: \begin{figure}[tbp]
364: \epsfig{file=density.eps,height=4cm}
365: \caption{\textit{Critical density, $\Omega_0$, as a function of $\protect%
366: \delta$. Solid line corresponds to pressure-less dust and dashed line to
367: black--body radiation.}}
368: \label{density}
369: \end{figure}
370:
371: \subsection{The dynamical systems approach}
372:
373: The system of equations (\ref{Friedman1}) and (\ref{Friedman2}) have been
374: studied previously using a dynamical systems approach by Carloni, Dunsby,
375: Capozziello and Troisi for general $\kappa $ \cite{Car04}. We elaborate on
376: their work by studying in detail the spatially flat, $\kappa =0$, subspace
377: of solutions. This allows us to draw conclusions about the asymptotic
378: solutions of (\ref{Friedman1}) and (\ref{Friedman2}) when $\kappa =0$ and so
379: investigate the stability of the power--law exact solution (\ref{power}) and
380: the extent to which it can be considered an attractor solution. By
381: restricting to $\kappa =0$ we avoid `instabilities' associated with the
382: curvature which are already present in general relativistic cosmologies.
383:
384: In performing this analysis we choose to work in the conformal time
385: coordinate
386: \begin{equation} \label{tau}
387: d \tau \equiv \sqrt{\frac{8 \pi \rho}{3 R^{\delta}}} dt.
388: \end{equation}
389:
390: Making the definitions
391: \begin{equation*}
392: x \equiv \frac{R^{\prime}}{R} \qquad \text{and} \qquad y \equiv \frac{%
393: a^{\prime}}{a},
394: \end{equation*}
395: where a prime indicates differentiation with respect to $\tau$, the field
396: equations (\ref{Friedman1}) and (\ref{Friedman2}) can be written as the
397: autonomous set of first order equations
398: \begin{align} \label{phase1}
399: x^{\prime}&= \frac{2- \delta (1+ 3 \omega)}{\delta^2 (1+\delta)}- \frac{%
400: \delta x^2}{2}-\frac{(4-\delta (1+3 \omega)) x y}{2 \delta}- \frac{2
401: (1-\delta) y^2}{\delta^2} \\
402: y^{\prime}&= -\frac{1}{\delta}+\frac{1}{2} (2+3 \delta) x y+\frac{(2+\delta
403: (1+3 \omega)) y^2}{2 \delta}.
404: \label{phase2}
405: \end{align}
406:
407: These coordinate definitions are closely related to those chosen by Holden
408: and Wands \cite{Hol98} for their phase-plane analysis of Brans-Dicke
409: cosmologies and allow us to proceed in a similar fashion.
410:
411: \subsubsection{Locating the critical points}
412:
413: The critical points at finite distances in the system of equations (\ref%
414: {phase1}) and (\ref{phase2}) are located at
415: \begin{equation} \label{critical1}
416: x_{1,2} = \pm \frac{1-3 \omega}{\delta \sqrt{(1+ \delta) (2-3 \omega)}}
417: \qquad \text{and} \qquad y_{1,2} = \pm \frac{1}{\sqrt{(1+\delta) (2-3 \omega)%
418: }}
419: \end{equation}
420: and at
421: \begin{equation} \label{critical2}
422: x_{3,4} = \mp \frac{3 \sqrt{2 (1+\omega)}}{\sqrt{(1+\delta) (2-3 \delta
423: (1+\omega)-\delta^2 (8+6 \omega))}} \qquad \text{and} \qquad y_{3,4} = \pm
424: \frac{\sqrt{2 (1+\delta)}}{\sqrt{2-3 \delta (1+\omega)-\delta^2 (8+6 \omega)}%
425: }.
426: \end{equation}
427:
428: The exact form of $a(t)$ at these critical points, and the stability of
429: these solutions, can be easily deduced. At the critical point $(x_{i},y_{i})$
430: the forms of $a(\tau )$ and $R(\tau )$ are given by
431: \begin{equation}
432: a(\tau )=a_{0}e^{y_{i}\tau }\qquad \text{and}\qquad R(\tau
433: )=R_{0}e^{x_{i}\tau }, \label{atau}
434: \end{equation}%
435: where $a_{0}$ and $R_{0}$ are constants of integration. In terms of $\tau $
436: the perfect-fluid conservation equation can be integrated to give
437: \begin{equation*}
438: \rho =\rho _{0}e^{-3(1+\omega )y_{i}\tau },
439: \end{equation*}%
440: where $\rho _{0}$ is another positive constant. Substituting into the
441: definition of $\tau $ now gives
442: \begin{equation*}
443: d\tau \propto e^{-\frac{3}{2}(1+\omega )y_{i}\tau -\frac{\delta }{2}%
444: x_{i}\tau }dt
445: \end{equation*}%
446: or, integrating,
447: \begin{equation}
448: t-t_{0}\propto \frac{1}{\frac{3}{2}(1+\omega )y_{i}+\frac{\delta }{2}x_{i}}%
449: e^{\frac{3}{2}(1+\omega )y_{i}\tau +\frac{\delta }{2}x_{i}\tau }. \label{t}
450: \end{equation}
451:
452: It can now be seen that if $3 (1+\omega) y_i+\delta x_i >0$ then $t
453: \rightarrow \infty$ as $\tau \rightarrow \infty$ and $t \rightarrow t_0$ as $%
454: \tau \rightarrow -\infty$. Conversely, if $3 (1+\omega) y_i+\delta x_i <0$
455: then $t \rightarrow t_0$ as $\tau \rightarrow \infty$ and $t \rightarrow
456: -\infty$ as $\tau \rightarrow -\infty$.
457:
458: In terms of $t$ time the solutions corresponding to the critical points at
459: finite distances can now be written as
460: \begin{equation*}
461: a(t)\propto (t-t_{0})^{\frac{2y_{i}}{3(1+\omega )y_{i}+\delta x_{i}}}\qquad
462: \text{and}\qquad R(t)\propto (t-t_{0})^{\frac{2x_{i}}{3(1+\omega
463: )y_{i}+\delta x_{i}}}.
464: \end{equation*}%
465: The critical points 1 and 2 can now been seen to correspond to $a\propto t^{%
466: \frac{1}{2}}$ and the points 3 and 4 correspond to (\ref{power}).
467:
468: In order to analyse the behaviour of the solutions as they approach infinity
469: it is convenient to transform to the polar coordinates
470: \begin{align*}
471: x &= \bar{r} \cos \phi \\
472: y &= \bar{r} \sin \phi.
473: \end{align*}
474: The infinite phase plane can then be compacted into a finite size by
475: introducing the coordinate
476: \begin{equation*}
477: r = \frac{\bar{r}}{1+\bar{r}}.
478: \end{equation*}
479: The equations (\ref{phase1}) and (\ref{phase2}) then become
480: \begin{multline} \label{phase3}
481: r^{\prime}=\frac{-1}{4 \delta^2 (1+\delta)} \Biggl( 4 (1-2 r) (\delta
482: (1+\delta) \sin \phi-(2-\delta (1+3 \omega)) \cos \phi) \\
483: - r^2 ((6-4 \delta+3 \delta^2+\delta^3-12 \delta \omega) \cos \phi
484: +(1+\delta) (2-2 \delta-\delta^2-2 \delta^3) \cos 3 \phi \\
485: - 2 \delta (3-\delta (1+3 \omega)+3 \cos 2 \phi) \sin \phi) \Biggr)
486: \end{multline}
487: and
488: \begin{multline} \label{phase4}
489: \phi^{\prime}=\frac{-1}{2 \delta^2 (1+\delta) (1-r) r} \Biggl( (2 \delta
490: (1+\delta) \cos \phi+2(2-\delta (1+3 \omega)) \sin \phi) (1-2 r) \\
491: -\Bigl( \delta (1+\delta) \cos \phi (1-3 \cos 2 \phi)-4 \sin \phi+4
492: (1-\delta)^2 \sin^3 \phi \\
493: +2 \delta (1+3 \omega+\delta (1+\delta) (1+2 \delta) \cos^2 \phi) \sin \phi
494: \Bigr) r^2 \Biggr).
495: \end{multline}
496:
497: In the limit $r\rightarrow 1$ ($\bar{r} \rightarrow \infty$) it can be seen
498: that critical points at infinity satisfy
499: \begin{equation*}
500: \sin \phi_i (\delta \cos \phi_i+\sin \phi_i) (\delta (1+2 \delta) \cos
501: \phi_i+2 (1-\delta) \sin \phi_i)=0
502: \end{equation*}
503: and so are located at
504: \begin{align} \label{5,6}
505: \phi_{5,(6)} &=0\qquad(+\pi) \\
506: \label{7,8}
507: \phi_{7,(8)} &=\tan^{-1} (-\delta)\qquad (+\pi) \\
508: \phi_{9,(10)} &=\tan^{-1} \left( -\frac{\delta (1+2 \delta)}{2 (1-\delta)}%
509: \right) \qquad(+\pi).
510: \label{9,10}
511: \end{align}
512: The form of $a(t)$ can now be calculated for each of these critical points
513: by proceeding as Holden and Wands \cite{Hol98}. Firstly, as $r \rightarrow 1$
514: equation (\ref{phase3}) approaches
515: \begin{align*}
516: r^{\prime}&\rightarrow \frac{1}{4 \delta^2} \Biggl( \delta (1+2 \delta (1+3
517: \omega)) \sin \phi_i-3 \delta \sin 3 \phi_i \\
518: &\qquad -(2-\delta (2+\delta))\cos \phi_i+(2-\delta (2+\delta+2 \delta^2))
519: \cos 3 \phi_i\Biggr) \\
520: &\equiv f(\phi_i)
521: \end{align*}
522: which allows the integral
523: \begin{equation*}
524: r-1= f(\phi_i) (\tau-\tau_0)
525: \end{equation*}
526: where the constant of integration, $\tau_0$, has been set so that $r
527: \rightarrow 1$ as $\tau \rightarrow \tau_0$. Now the definition of $x$
528: allows us to write
529: \begin{equation*}
530: \frac{R^{\prime}}{R}=\frac{r}{(1-r)} \cos \phi_i=-\frac{f(\phi_i)
531: (\tau-\tau_0)+1}{f(\phi_i) (\tau-\tau_0)} \cos \phi_i \rightarrow -\frac{%
532: \cos \phi_i}{f(\phi_i) (\tau-\tau_0)}
533: \end{equation*}
534: as $\tau \rightarrow \tau_0$. Integrating this it can be seen that
535: \begin{equation*}
536: R \propto \vert \tau-\tau_0 \vert^{-\frac{\cos \phi_i}{f(\phi_i)}} \qquad
537: \text{as} \qquad r \rightarrow 1.
538: \end{equation*}
539: Similarly,
540: \begin{equation*}
541: a \propto \vert \tau-\tau_0 \vert^{-\frac{\sin \phi_i}{f(\phi_i)}} \qquad
542: \text{as} \qquad r \rightarrow 1.
543: \end{equation*}
544:
545: The definition of $\tau$ (\ref{tau}) now gives
546: \begin{equation*}
547: d \tau \propto \vert \tau-\tau_0 \vert^{\frac{3}{2} (1+\omega) \frac{%
548: \sin\phi_i}{f(\phi_i)}+\frac{\delta}{2} \frac{\cos\phi_i}{f(\phi_i)}} dt
549: \end{equation*}
550: which integrates to
551: \begin{equation} \label{tinfty}
552: t-t_0 \propto - \frac{f(\phi_1)}{F(\phi_i)} \vert \tau-\tau_0 \vert^{-\frac{%
553: F(\phi_i)}{f(\phi_i)}}
554: \end{equation}
555: where
556: \begin{equation*}
557: F(\phi_i)=\frac{3 (1+\omega) \sin \phi_i +\delta \cos \phi_i-2 f(\phi_i)}{2}.
558: \end{equation*}
559: The location of critical points at infinity can now be written in terms of $%
560: t $ as the power--law solutions
561: \begin{equation} \label{inftysol}
562: R(t) \propto (t-t_0)^{\frac{\cos\phi_i}{F(\phi_i)}} \qquad \text{and} \qquad
563: a(t) \propto (t-t_0)^{\frac{\sin\phi_i}{F(\phi_i)}}.
564: \end{equation}
565:
566: Direct substitution of the critical points (\ref{5,6}), (\ref{7,8}) and (\ref%
567: {9,10}) into (\ref{inftysol}) gives
568: \begin{align*}
569: a_{5,6}(t)& \rightarrow \ constant \\
570: a_{7,8}(t)& \rightarrow \sqrt{t-t_{0}} \\
571: a_{9,10}(t)& \rightarrow (t-t_{0})^{\frac{\delta (1+2\delta )}{(1-\delta )}}
572: \end{align*}%
573: as $r\rightarrow 1$. Moreover, it can be seen from (\ref{tinfty}) that as $%
574: r\rightarrow 1$ and $\tau \rightarrow \tau _{0}$ so $t\rightarrow t_{0}$ as
575: long as $F(\phi _{i})/f(\phi _{i})<0$, as is the case for the stationary
576: points considered here (as long as the value of $\delta $ lies within the
577: range given by (\ref{range})).
578:
579: The exact forms of $a(t)$ at all the critical points are summarised in the
580: table below.
581:
582: \begin{center}
583: \begin{tabular}{c|c}
584: \textbf{Critical point} & \textbf{a(t)} \\ \hline
585: 1, 2, 7 and 8 & $t^{\frac{1}{2}}$ \\
586: 3 and 4 & $t^{\frac{2 (1+\delta)}{3 (1+\omega)}}$ \\
587: 5 and 6 & constant \\
588: 9 and 10 & $t^{\frac{\delta (1+2 \delta)}{(1-\delta)}}$%
589: \end{tabular}
590: \end{center}
591:
592: \subsubsection{Stability of the critical points}
593:
594: The stability of the critical points at finite distances can be established
595: by perturbing $x$ and $y$ as
596: \begin{equation}
597: x(r)=x_{i}+u(r)\qquad \text{and}\qquad y(r)=y_{i}+v(r) \label{lin}
598: \end{equation}%
599: and checking the sign of the eigenvalues, $\lambda _{i}$, of the linearised
600: equations
601: \begin{equation*}
602: u^{\prime }=\lambda _{i}u\qquad \text{and}\qquad v^{\prime }=\lambda _{i}v.
603: \end{equation*}
604:
605: Substituting (\ref{lin}) into equations (\ref{phase1}) and (\ref{phase2})
606: and linearising in $u$ and $v$ gives
607: \begin{align*}
608: u^{\prime }& =-\left( \delta x_{i}+\frac{(4-\delta (1+3\omega ))}{2\delta }%
609: y_{i}\right) u-\left( \frac{(4-\delta (1+3\omega ))}{2\delta }x_{i}+4\frac{%
610: (1+\delta )}{\delta ^{2}}y_{i}\right) v \\
611: v^{\prime }& =\frac{(2+3\delta )}{2}y_{0}u+\left( \frac{(2+3\delta )}{2}%
612: x_{i}+\frac{(2+\delta (1+3\omega ))}{\delta }y_{0}\right) v.
613: \end{align*}%
614: The eigenvalues $\lambda _{i}$ are therefore the roots of the quadratic
615: equation
616: \begin{equation*}
617: \lambda _{i}^{2}+B\lambda _{i}+C=0
618: \end{equation*}%
619: where
620: \begin{align*}
621: B& =-\frac{1}{2}(2+\delta )x_{i}-\frac{3}{2}(1+3\omega )y_{i} \\
622: C& =-\frac{\delta }{2}(2+3\delta )x_{i}^{2}-(2+\delta (1+3\delta
623: ))x_{i}y_{i}+\frac{1}{2\delta }(2-11\delta -6\omega (1-\delta )+9\delta
624: \omega ^{2})y_{i}^{2}.
625: \end{align*}%
626: If $B>0$ and $C>0$ then both values of $\lambda _{i}$ are negative, and we
627: have a stable critical point. If $B<0$ and $C>0$ both values of $\lambda
628: _{i} $ are positive, and the critical point is unstable to perturbations. $%
629: C>0$ gives a saddle-point.
630:
631: For points $1$ (upper branch) and $2$ (lower branch) this gives
632: \begin{equation*}
633: B= \mp \frac{(1+\delta (2+3 \omega)-3 \omega)}{\delta \sqrt{(1+\delta) (2-3
634: \omega)}} \qquad \text{and} \qquad C= - \frac{(1+4 \delta-3 \omega)}{\delta
635: (1+\delta)}
636: \end{equation*}
637: and for points $3$ (upper branch) and $4$ (lower branch)
638: \begin{equation*}
639: B = \pm \frac{3 (1-\omega (1+2 \delta))}{\sqrt{2 (1+\delta) (2-3 \delta
640: (1+\omega)-2 \delta^2 (4+3 \omega))}} \qquad \text{and} \qquad C= \frac{(1+4
641: \delta -3 \omega)}{\delta (1+\delta)}.
642: \end{equation*}
643:
644: The stability of the critical points at finite distances for a universe
645: filled with pressureless dust are therefore, for various different values of
646: $\delta $, given by
647:
648: \begin{center}
649: \begin{tabular}{c|cc|ccc}
650: \label{w=0} \textbf{Critical point} & \textbf{B} & \textbf{C} & $-\frac{%
651: \sqrt{73}+3}{16} <\delta <-\frac{1}{4}$ & $-\frac{1}{4} <\delta<0$ & $%
652: 0<\delta <\frac{\sqrt{73}-3}{16} $ \\ \hline
653: 1 & $-\frac{(1+2 \delta)}{\delta \sqrt{2 (1+\delta)}}$ & $-\frac{(1+4 \delta)%
654: }{\delta (1+\delta)}$ & Saddle & Stable & Saddle \\
655: 2 & $\frac{(1+2 \delta)}{\delta \sqrt{2 (1+\delta)}}$ & $-\frac{(1+4 \delta)%
656: }{\delta (1+\delta)}$ & Saddle & Unstable & Saddle \\
657: 3 & $\frac{3}{\sqrt{2 (1+\delta) (2-3 \delta-8 \delta^2)}}$ & $\frac{(1+4
658: \delta)}{\delta (1+\delta)}$ & Stable & Saddle & Stable \\
659: 4 & $-\frac{3}{\sqrt{2 (1+\delta) (2-3 \delta-8 \delta^2)}}$ & $\frac{(1+4
660: \delta)}{\delta (1+\delta)}$ & Unstable & Saddle & Unstable%
661: \end{tabular}
662: \end{center}
663:
664: and for a universe filled with black-body radiation are given by
665:
666: \begin{center}
667: \begin{tabular}{c|cc|c}
668: \label{w=1/3} \textbf{Critical point} & \textbf{B} & \textbf{C} & $-\frac{%
669: \sqrt{6}+1}{5} <\delta < \frac{\sqrt{6}-1}{5}$ \\ \hline
670: 1 & $-\frac{3}{\sqrt{1+\delta}}$ & $-\frac{4}{(1+\delta)}$ & Saddle \\
671: 2 & $\frac{3}{\sqrt{1+\delta}}$ & $-\frac{4}{(1+\delta)}$ & Saddle \\
672: 3 & $\frac{(1-\delta)}{\sqrt{(1+\delta) (1-2\delta-5 \delta^2)}}$ & $\frac{4%
673: }{(1+\delta)}$ & Stable \\
674: 4 & $-\frac{(1-\delta)}{\sqrt{(1+\delta) (1-2\delta-5 \delta^2)}}$ & $\frac{4%
675: }{(1+\delta)}$ & Unstable%
676: \end{tabular}
677: \end{center}
678:
679: Values of $\delta<-\frac{\sqrt{73+66\omega+9\omega^2}+3(1+\omega)}{4 (4+3
680: \omega)}$ and $\delta>\frac{\sqrt{73+66\omega+9\omega^2}-3(1+\omega)}{4 (4+3
681: \omega)}$ have not been considered here as they lead to negative values of $%
682: \rho_0$ for the solution (\ref{power}). (The reader may note the
683: difference here between the range of $\delta$ for which point 3 is a
684: stable attractor compared with the analysis of Carloni et. al.).
685:
686: Point 3 lies in the $y>0$ region and so corresponds to the expanding
687: power--law solution (\ref{power}). It can be seen from the table above that
688: this solution is stable for certain ranges of $\delta $ and a saddle-point
689: for others. In contrast, point 4, the contracting power--law solution (\ref%
690: {power}), is unstable or a saddle-point. The nature of the stability of
691: these points and the trajectories which are attracted towards them will be
692: explained further in the next section.
693:
694: A similar analysis can be performed for the critical points at infinity.
695: This time only the variable $\phi $ will be perturbed as
696: \begin{equation}
697: \phi (t)=\phi _{i}+q(t). \label{phipert}
698: \end{equation}%
699: The conditions for stability of the critical points are now that $r^{\prime
700: }>0$ and the eigenvalue $\mu $ of the linearised equation $q^{\prime }=\mu q$
701: satisfies $\mu <0$, in the limit $r\rightarrow 1$. If both of these
702: conditions are satisfied then the point is a stable attractor, if only one
703: is satisfied the point is a saddle-point and if neither are satisfied then
704: the point is repulsive.
705:
706: Substituting (\ref{phipert}) into (\ref{phase4}) and linearising in $q(t)$
707: gives, in the limit $r\rightarrow 1$,
708: \begin{align*}
709: q^{\prime }& =\frac{1}{4\delta ^{2}(1-r)}\left( (6(1-\delta )+\delta
710: ^{2}(1+2\delta ))\cos \phi _{i}-3(2(1-\delta )-\delta ^{2}(1+2\delta ))\cos
711: 3\phi _{i}-3\delta \sin \phi _{i}+9\delta \sin 3\phi \right) q \\
712: & \equiv \mu q.
713: \end{align*}%
714: The sign of $r^{\prime }$ in the limit $r\rightarrow 1$ can be read off from
715: (\ref{phase3}). The stability properties of each of the stationary points at
716: infinity can now be summarised in the table below
717:
718: \begin{center}
719: \begin{tabular}{c|cccc}
720: \label{w=0infty} \textbf{Critical point} & $-N_1 <\delta <-\frac{1}{2}$ & $-%
721: \frac{1}{2}<\delta<0$ & $0<\delta<\frac{1}{4}$ & $\frac{1}{4}<\delta <N_2 $
722: \\ \hline
723: 5 & Stable & Saddle & Unstable & Unstable \\
724: 6 & Unstable & Saddle & Stable & Stable \\
725: 7 & Unstable & Unstable & Saddle & Stable \\
726: 8 & Stable & Stable & Saddle & Unstable \\
727: 9 & Saddle & Stable & Stable & Saddle \\
728: 10 & Saddle & Unstable & Unstable & Saddle%
729: \end{tabular}
730: \end{center}
731:
732: where $N_1=\frac{\sqrt{73+66\omega+9\omega^2}+3(1+\omega)}{4 (4+3 \omega)}$
733: and $N_2=\frac{\sqrt{73+66\omega+9\omega^2}-3(1+\omega)}{4 (4+3 \omega)}$.
734:
735: \subsubsection{Illustration of the phase plane}
736:
737: Some representative illustrations of the phase plane are now presented.
738: Firstly, the compactified phase plane for a universe filled with
739: pressureless dust, $\omega =0$, and a value of $\delta =0.1$ is shown in
740: figure \ref{d0.1}.
741: \begin{figure}[tbp]
742: \epsfig{file=d01.eps,height=10cm}
743: \caption{\textit{Phase plane of cosmological solutions for $\protect\omega %
744: =0 $ and $\protect\delta =0.1$.}}
745: \label{d0.1}
746: \end{figure}
747: Figure \ref{d0.1} is seen to be split into three separate regions labelled
748: I, II and III. The boundaries between these regions are the sub-manifolds $%
749: R=0$. As pointed out by Carloni et. al. the plane $R=0$ is an invariant
750: sub-manifold of the phase space through which trajectories cannot pass.
751:
752: The equation for $R$ in a FRW universe, (\ref{R2}), can now be rewritten as
753: \begin{equation}
754: R=\frac{16\pi \rho }{\delta R^{\delta }}((1+\delta )y^{2}+\delta (1+\delta
755: )xy-1). \label{xy}
756: \end{equation}%
757: This shows that the boundary $R=0$ is given in terms of $x$ and $y$ by $%
758: (1+\delta )y^{2}+\delta (1+\delta )xy-1=0$ and that in region I the sign of $%
759: R$ must be opposite to the sign of $\delta $ in order to have a positive $%
760: \rho $. Similarly, in regions II and III, $R$ must have the same sign as $%
761: \delta $ in order to ensure a positive $\rho $.
762:
763: It can be seen that regions II and III are symmetric under a rotation of $%
764: \pi $ and a reversal of the direction of the trajectories. As region II is
765: exclusively in the semi--circle $y\geqslant 0$ all trajectories confined to
766: this region correspond to eternally expanding (or expanding and
767: asymptotically static) universes. Similarly, region III is confined to the
768: semi--circle $y \leqslant 0$ and so all trajectories confined to this region
769: correspond to eternally contracting (or contracting and asymptotically
770: static) universes. Region I, however, spans the $y=0$ plane and so can have
771: trajectories which correspond to universes with both expanding and
772: contracting phases. In fact, it can be seen from figure \ref{d0.1} that, for
773: $\delta=0.1$ all trajectories in region I are initially expanding and
774: eventually contracting.
775:
776: It can be seen from figure \ref{d0.1} that in region I the only stable
777: attractors are, at early times, the expanding point 10 and at late times the
778: contracting point 9. (By `attractors at early times' we mean the critical
779: points which are approached if the trajectories are followed backwards in
780: time). Both of these points correspond to the solution
781:
782: \begin{equation*}
783: a\propto t^{\delta \frac{(1+2\delta )}{(1-\delta )}}
784: \end{equation*}
785: which describes a slow evolution independent of the matter content of the
786: universe. Notably, region I only has stable attractor points, at both early
787: and late times, which have been shown to correspond to $t\rightarrow $%
788: constant; region I therefore does not have an asymptotic attractor when $%
789: t\rightarrow \infty $, for the range of $\delta $ being considered. In
790: region II the only stable attractors can be seen to be the static point 5 at
791: some early finite time, $t_{0}$, and the expanding matter-driven expansion
792: described by point 3 as $t\rightarrow \infty $. Conversely, in region III
793: the only stable attractors are the contracting point 4 for $t\rightarrow
794: -\infty $ and the static point 6 for $t\rightarrow t_{0}$.
795:
796: Figure \ref{d-0.1} shows the compactified phase plane for a universe
797: containing pressureless dust and having $\delta =-0.1$.
798: \begin{figure}[tbp]
799: \epsfig{file=dm012.eps,height=10cm}
800: \caption{\textit{Phase plane of cosmological solutions for $\protect\omega %
801: =0 $ and $\protect\delta =-0.1$.}}
802: \label{d-0.1}
803: \end{figure}
804: Figure \ref{d-0.1} is split into three separate regions in a similar way to
805: figure \ref{d0.1}, with the boundary between the regions again corresponding
806: to $R=0$ and is given in terms of $x$ and $y$ by (\ref{xy}). Regions II and
807: III again correspond to expanding and contacting solutions, respectively.
808: Region I, still has point 10 as the early-time attractor and point 9 as the
809: late-time attractor, but now has all trajectories initially contracting and
810: eventually expanding. There are still no stable attractors in Region I which
811: correspond to regions where $t\rightarrow \infty $. Region II now has point
812: 7 as an early-time stable attractor solution and point 1 as a late-time
813: stable attractor solution, corresponding to $a\rightarrow t^{\frac{1}{2}}$
814: as $t\rightarrow \infty $. Point 3, which was the stable attractor at late
815: times when $\delta =0.1$ is now no longer located in Region II and can
816: instead be located in region I where it is now a saddle-point in the phase
817: plane. Interestingly, the value of $\delta $ for which point 3 ceases to
818: behave as a stable attractor ($\delta =0$) is exactly the same value of $%
819: \delta $ at which the point moves from region II into region I; so as long
820: as point 3 can be located in region I, it is the late-time stable attractor
821: solution and as soon as it moves into region I it becomes a saddle-point. At
822: this same value of $\delta ,$ point 1 ceases to be a saddle-point and
823: becomes the late-time stable attractor for region II, so that region II
824: always has a stable late-time attractor where $t\rightarrow \infty $. Region
825: III behaves in a similar way to the description given for region II above,
826: under a rotation of $\pi $ and with the directions of the trajectories
827: reversed.
828:
829: Phase planes diagrams for $\omega =0$ with values of $\delta $ other than $%
830: 0.1$ and $-0.1$, but still within the range
831:
832: \begin{equation*}
833: -\frac{\sqrt{73+66\omega +9\omega ^{2}}+3(1+\omega )}{4(4+3\omega )}<\delta <%
834: \frac{\sqrt{73+66\omega +9\omega ^{2}}-3(1+\omega )}{4(4+3\omega )},
835: \end{equation*}%
836: look qualitatively similar to those above with some of the attractor
837: properties of the critical points being exchanged as they pass each
838: other. Notably, for $\delta <-\frac{1}{4}$ point 3 returns to region II and once
839: again becomes the stable late-time attractor for trajectories in that
840: region. The points that are the stable attractors for any particular value
841: of $\delta $ can be read off from the tables in the previous section.
842:
843: Universes filled with perfect fluid black-body radiation also retain
844: qualitatively similar phase-plane diagrams to the ones above; with the
845: notable difference that the point 3 is always located in region II and is
846: always the late-time stable attractor of that region. This can be seen
847: directly from the Ricci scalar for the solution (\ref{power}) which is given
848: by
849: \begin{equation*}
850: R=\frac{3\delta (1+\delta )}{t^{2}}
851: \end{equation*}%
852: and can be seen to have the same sign as $\delta $, for $\delta >-1$, and so
853: is always found in region II.
854:
855: For a spatially-flat, expanding FRW universe containing black-body radiation
856: we therefore have that (\ref{power}) is the generic attractor as $%
857: t\rightarrow \infty $. Similarly, for a spatially-flat, expanding,
858: matter-dominated FRW universe (\ref{power}) is the attractor solution as $%
859: t\rightarrow \infty $; except when $-\frac{1}{4}<\delta <0$, in which case
860: it is point 1 ($a\propto t^{\frac{1}{2}}$).
861:
862: If we require a stable period of matter domination, during which $a(t)\sim
863: t^{\frac{2}{3}}$, we therefore have the theoretical constraint $\delta >0$
864: (or $\delta <-\frac{1}{4}$). Such a period is necessary for structure to
865: form through gravitational collapse in the post-recombination era of the
866: universe's expansion.
867:
868: The effect of a non--zero curvature, $\kappa \neq 0$, on the cosmological
869: dynamics is similar to the general relativistic case. The role of negative
870: curvature ($\kappa =-1$) can be deduced by noting that its effect is similar
871: to that of a fluid with $\omega =-1/3$. The solution (\ref{power}) is
872: unstable to any perturbation away from flatness and will diverge away from $%
873: \kappa =0$ as $t\rightarrow \infty $. This is usually referred to as the
874: `flatness problem' and can be seen to exist in this theory from the analysis
875: of Carloni et. al. \cite{Car04}.
876:
877: \subsection{Physical consequences}
878:
879: The modified cosmological dynamics discussed in the last section lead to
880: different predictions for the outcomes of physical processes, such as
881: primordial nucleosynthesis and CMB formation, compared to the standard
882: general-relativistic model. The relevant modifications to these physical
883: processes, and the bounds that they can impose upon the theory, will be
884: discussed in this section. We will use the solutions (\ref{power}) as they
885: have been shown to be the generic attractors as $t\rightarrow \infty $
886: (except for the case $-\frac{1}{4}<\delta <0$ when $\omega =0$, which has
887: been excluded as physically unrealistic on the grounds of structure
888: formation).
889:
890: \subsubsection{Primordial nucleosynthesis}
891:
892: We find that the temperature-time adiabat during radiation domination for
893: the solution (\ref{power}) is given by the exact relation
894: \begin{equation}
895: t^{2(1+d)}=\frac{A}{T^{4}} \label{T(t)}
896: \end{equation}%
897: where, as usual (with units $\hbar =c=1=k_{B}$),
898: \begin{equation*}
899: \rho =\frac{g\pi ^{2}}{15}T^{4}
900: \end{equation*}%
901: where $g$ is the total number of relativistic spin states at temperature $T$%
902: . The constant $A$ can be determined from the generalised Friedmann equation
903: (\ref{Friedman2}) and is dependent on the present day value of the Ricci
904: scalar, through equation (\ref{chi}). (This dependence is analogous to the
905: dependence of scalar--tensor theories on the evolution of the non--minimally
906: coupled scalar, as may be expected from the relationship between these
907: theories \cite{Mag94}). As a first approximation we assume the universe to
908: have been matter dominated throughout its later history; this allows us to
909: write
910: \begin{equation}
911: A=\left( \frac{45(1-2\delta )(1-2\delta -5\delta ^{2})}{32(1-\delta )g\pi
912: ^{3}G}\right) \left( \frac{2(1+\delta )}{3H_{0}}\right) ^{2\delta }
913: \label{A}
914: \end{equation}%
915: where $H_{0}$ is the value of Hubble's constant today and we have used the
916: solution (\ref{power}) to model the evolution of $a(t)$. Adding a recent
917: period of accelerated expansion will refine the constant $A$, but in the
918: interests of brevity we exclude this from the current analysis.
919:
920: As usual, the weak-interaction time is given by
921: \begin{equation*}
922: t_{wk}\propto \frac{1}{T^{5}}.
923: \end{equation*}%
924: The freeze--out temperature, $T_{f}$, for neutron--proton kinetic
925: equilibrium is then defined by
926: \begin{equation*}
927: t(T_{f})=t_{wk}(T_{f}),
928: \end{equation*}%
929: hence the freeze--out temperature in this theory, with $\delta \neq 0$, is
930: related to that in the general-relativistic case with $\delta =0$, $%
931: T_{f}^{GR}$, by
932: \begin{equation}
933: T_{f}=C(T_{f}^{GR})^{\frac{3(1+\delta )}{(3+5\delta )}} \label{freeze}
934: \end{equation}%
935: where
936: \begin{equation}
937: C=\left( \frac{(1-\delta )}{(1-2\delta )(1-2\delta -5\delta ^{2})}\right) ^{%
938: \frac{1}{2(2+5\delta )}}\left( \frac{45}{32g\pi ^{3}G}\right) ^{\frac{\delta
939: (1+\delta )}{2(3+5\delta )}}\left( \frac{3H_{0}}{2(1+\delta )}\right) ^{%
940: \frac{\delta }{(3+5\delta }}. \label{freeze2}
941: \end{equation}%
942: The neutron--proton ratio, $n/p$, is now determined at temperature $T$ when
943: the equilibrium holds by
944: \begin{equation*}
945: \frac{n}{p}=\exp \left( -\frac{\Delta m}{T}\right) .
946: \end{equation*}%
947: where $\Delta m$ is the neutron-proton mass difference. Hence the
948: neutron--proton ratio at freeze-out in the $R^{1+\delta }$ early universe is
949: given by
950: \begin{equation*}
951: \frac{n}{p}=\exp \left( -\frac{\Delta m}{C(T_{f}^{GR})^{1-\varepsilon }}%
952: \right) ,
953: \end{equation*}%
954: where
955: \begin{equation*}
956: \varepsilon \equiv \frac{2\delta }{3+5\delta }.
957: \end{equation*}
958:
959: The frozen--out $n/p$ ratio in the $R^{1+\delta }$ theory is given by a
960: power of its value in the general relativistic case, $\left( \frac{n}{p}%
961: \right) _{GR}\approx 1/7$, by
962: \begin{equation*}
963: \frac{n}{p}=\left( \frac{n}{p}\right) _{GR}^{C(T_{f}^{GR})^{\varepsilon }}.
964: \end{equation*}%
965: It is seen that when $C(T_{f}^{GR})^{\varepsilon }>1$ ($\delta <0$) there is
966: a smaller frozen-out neutron-proton ratio that in the general-relativistic
967: case and consequently a lower final helium-4 abundance than in the standard
968: general-relativistic early universe containing the same number of
969: relativistic spin states. This happens because the freeze-out temperature is
970: lower than in general relativity. The neutrons remain in equilibrium to a
971: lower temperature and their slightly higher mass shifts the number balance
972: more towards the protons the longer they are in equilibrium. Note that a
973: reduction in the helium-4 abundance compared to the standard model of
974: general relativity is both astrophysically interesting and difficult to
975: achieve (all other variants like extra particle species \cite{shv, nus},
976: anisotropies \cite{ht, jb1, jb2, jb3}, magnetic fields \cite{mag, YM},
977: gravitational waves \cite{jb1, jb2, skew}, or varying $G$ \cite{bmaeda,
978: newt, clift}, lead to an \textit{increase} in the expansion rate and in the
979: final helium-4 abundance). Conversely, when $C(T_{f}^{GR})^{\varepsilon }<1$
980: ($\delta >0$) freeze-out occurs at a higher temperature than in general
981: relativity and a higher final helium-4 abundance fraction results. The final
982: helium-4 mass fraction $Y$ is well approximated by
983: \begin{equation}
984: Y=\frac{2n/p}{(1+n/p)}.
985: \end{equation}
986:
987: It is now possible to constrain the value of $\delta $ using observational
988: abundances of the light elements. In doing this we will use the results of
989: Carroll and Kaplinghat \cite{Car02} who consider nucleosynthesis with a
990: Hubble constant parametrised by
991: \begin{equation*}
992: H(T)=\left( \frac{T}{1MeV}\right) ^{\alpha }H_{1}.
993: \end{equation*}%
994: Our theory can be cast into this form by substituting
995: \begin{equation*}
996: \alpha =\frac{2}{(1+\delta )}
997: \end{equation*}%
998: and
999: \begin{equation*}
1000: H_{1}=\frac{(1+\delta )}{2}A^{-\frac{1}{2(1+\delta )}}(1MeV)^{\frac{2}{%
1001: (1+\delta )}},
1002: \end{equation*}%
1003: so, taking $g=43/8$, $G=6.72\times 10^{-45}MeV^{-2}$ and $H_{0}=1.51\times
1004: 10^{-39}MeV$ \cite{Ben03}, this can be rewritten as
1005: \begin{equation*}
1006: H_{1}=\frac{(1+\delta )}{2}\left( \frac{7.96\times 10^{-43}(1-\delta )}{%
1007: (1-2\delta )(1-2\delta -5\delta ^{2})}\right) ^{\frac{1}{2(1+\delta )}%
1008: }\left( \frac{2.23\times 10^{-39}}{(1+\delta )}\right) ^{\frac{\delta }{%
1009: (1+\delta )}}MeV.
1010: \end{equation*}%
1011: Carroll and Kaplinghat use the observational abundances inferred by Olive
1012: et. al. \cite{Oli00}
1013: \begin{align*}
1014: 0.228\leqslant Y_{P}& \leqslant 0.248 \\
1015: 2\leqslant 10^{5}\times & \frac{D}{H}\leqslant 5 \\
1016: 1\leqslant 10^{10}\times & \frac{^{7}Li}{H}\leqslant 3
1017: \end{align*}%
1018: to impose the constraint
1019: \begin{equation*}
1020: H_{1}=H_{c}\left( \frac{T_{c}}{MeV}\right) ^{-\alpha }
1021: \end{equation*}%
1022: where $H_{c}=2.6\pm 0.9\times 10^{-23}MeV$ at $T_{c}=0.2MeV$ for $%
1023: 0.5\leqslant \eta _{10}\leqslant 50$, or $H_{c}=2.0\pm 0.3\times 10^{-23}$
1024: for $1\leqslant \eta _{10}\leqslant 10$ and $\eta _{10}$ is $10^{10}$ times
1025: the baryon to photon ratio.
1026:
1027: These results can now be used to impose upon $\delta$ the constraints
1028: \begin{equation*}
1029: -0.017 \leqslant \delta \leqslant 0.0012,
1030: \end{equation*}
1031: for $0.5 \leqslant \eta_{10} \leqslant 50$, or
1032: \begin{equation}
1033: -0.0064 \leqslant \delta \leqslant 0.0012,
1034: \end{equation}
1035: for $1 \leqslant \eta_{10} \leqslant 10$.
1036:
1037: \subsubsection{Horizon size at matter--radiation equality}
1038:
1039: The horizon size at the epoch of matter--radiation equality is of great
1040: observational significance. During radiation domination cosmological
1041: perturbations on sub--horizon scales are effectively frozen. Once matter
1042: domination commences, however, perturbations on all scales are allowed to
1043: grow and structure formation begins. The horizon size at matter--radiation
1044: equality is therefore frozen into the power spectrum of perturbations and is
1045: observable. Calculation of the horizon sizes in this theory proceeds in a
1046: similar way to that in Brans--Dicke theory \cite{Lid98}.
1047:
1048: In making an estimate of the horizon sizes in $R^{1+\delta }$ theory we will
1049: use the generalised Friedmann equation, (\ref{Friedman2}), in the form
1050: \begin{equation}
1051: H^{2}+\delta H\frac{\dot{R}}{R}-\frac{\delta R}{6(1+\delta )}=\frac{8\pi
1052: G(1-\delta )}{3(1-2\delta )}\frac{R_{0}^{\delta }}{R^{\delta }}\rho .
1053: \label{Fred}
1054: \end{equation}
1055: Again, we assume the form (\ref{power}) to model the evolution of the scale
1056: factor during the epoch of matter domination. This gives the results
1057: \begin{align*}
1058: a(t)& =a_{0}\left( \frac{t}{t_{0}}\right) ^{\frac{2(1+\delta )}{3}} \\
1059: H_{0}& =\frac{2(1+\delta )}{3t_{0}} \\
1060: \rho _{m}& =\frac{3H_{0}^{2}}{16\pi G}\frac{(1-2\delta )(2-3\delta -8\delta
1061: ^{2})}{(1-\delta )(1+\delta )^{2}}\frac{a_{0}^{3}}{a^{3}} \\
1062: R(t)& =\frac{4(1+5\delta +4\delta ^{2})}{3t^{2}}
1063: \end{align*}%
1064: during the matter-dominated era. In order to simplify matters we assume the
1065: above solutions to hold exactly from the time of matter--radiation equality
1066: up until the present (neglecting the small residual radiation effects and
1067: the very late time acceleration). Substituting them into (\ref{Fred}) along
1068: with $\rho _{eq}=2\rho _{meq}$ at equality we can then solve for $H_{eq}$ to
1069: first order in $\delta $ to find
1070: \begin{equation}
1071: \frac{a_{eq}H_{eq}}{a_{0}H_{0}}\simeq \sqrt{2}\sqrt{1+z_{eq}}^{\frac{%
1072: 1-2\delta }{1+\delta }}(1-2.686\delta )+O(\delta ^{2}) \label{Heq}
1073: \end{equation}%
1074: where $z_{eq}$ is the redshift at matter radiation equality and $H$ has been
1075: treated as an independent parameter. The value of $1+z_{eq}$ can now be
1076: calculated in this theory as
1077: \begin{equation}
1078: 1+z_{eq}=\frac{\rho _{r0}}{\rho _{m0}}. \label{zeq}
1079: \end{equation}%
1080: Taking the present day temperature of the of the microwave background as $%
1081: T=2.728\pm 0.004K$ \cite{Fix96} gives
1082: \begin{equation}
1083: \rho _{r0}=3.37\times 10^{-39}MeV^{4} \label{rhor0}
1084: \end{equation}%
1085: where three families of light neutrinos have been assumed at a temperature
1086: lower than that of the microwave background by a factor $(4/11)^{\frac{1}{3}%
1087: } $. Using the same values for $G$ and $H_{0}$ as above we than find from
1088: the above expression for $\rho _{m}$ that
1089: \begin{equation}
1090: \rho _{m0}=2.03\times 10^{-35}\frac{(1-2\delta )(2-3\delta -8\delta ^{2})}{%
1091: (1-\delta )(1+\delta )^{2}}MeV^{4}. \label{rhom0}
1092: \end{equation}%
1093: Substituting (\ref{rhor0}), (\ref{rhom0}) and (\ref{zeq}) into (\ref{Heq})
1094: then gives the expression for the horizon size at equality, to first order
1095: in $\delta $, as
1096: \begin{equation}
1097: \frac{a_{eq}H_{eq}}{a_{0}H_{0}}\simeq 155(1-19\delta )+O(\delta ^{2}).
1098: \end{equation}%
1099: This expression shows that the horizon size at matter--radiation equality
1100: will be shifted by $\sim 1\%$ for a value of $\delta \sim 0.0005$. This
1101: shift in horizon size should be observable in a shift of the peak of the
1102: power spectrum of perturbations, compared to its position in the standard
1103: general relativistic cosmology. Microwave background observations,
1104: therefore, allow a potentially tight bound to be derived on the value of $%
1105: \delta $. This effect is analogous to the shift of power--spectrum peaks in
1106: Brans--Dicke theory (see e.g. \cite{Lid98}, \cite{Che99}).
1107:
1108: A full analysis of the spectrum of perturbations in this theory requires a
1109: knowledge of the evolution of linearised perturbations as well as a
1110: marginalization over other parameters which can mimic this effect (e.g.
1111: baryon density). Such a study is beyond the scope of the present work.
1112:
1113: \section{Static and spherically--symmetric solutions}
1114:
1115: In order to test the $R^{n}$ gravity theory in the weak-field limit by means
1116: of the standard solar-system tests of general relativity we need to find the
1117: analogue of the Schwarzschild metric in this generalised theory and use it
1118: to describe the gravitational field of the Sun. In the absence of any matter
1119: the field equations (\ref{field}) can be written as
1120: \begin{equation}
1121: R_{ab}=\delta \left( \frac{R_{;}^{\ cd}}{R}-(1-\delta )\frac{R_{,}^{\
1122: c}R_{,}^{\ d}}{R^{2}}\right) \left( g_{ac}g_{bd}+\frac{1}{2}\frac{(1+2\delta
1123: )}{(1-\delta )}g_{ab}g_{cd}\right) . \label{staticfield}
1124: \end{equation}%
1125: We find that an exact static spherically symmetric solution of these field
1126: equations is given in Schwarzschild coordinates by the line--element
1127: \begin{equation}
1128: ds^{2}=-A(r)dt^{2}+\frac{dr^{2}}{B(r)}+r^{2}(d\theta ^{2}+\sin ^{2}\theta
1129: d\phi ^{2}) \label{Chan}
1130: \end{equation}%
1131: where
1132: \begin{align*}
1133: A(r)& =r^{2\delta \frac{(1+2\delta )}{(1-\delta )}}+\frac{C}{r^{\frac{%
1134: (1-4\delta )}{(1-\delta )}}} \\
1135: B(r)& =\frac{(1-\delta )^{2}}{(1-2\delta +4\delta ^{2})(1-2\delta (1+\delta
1136: ))}\left( 1+\frac{C}{r^{\frac{(1-2\delta +4\delta ^{2})}{(1-\delta )}}}%
1137: \right)
1138: \end{align*}%
1139: and $C=$ constant. This solution is conformally related to the $Q=0$ limit
1140: of the one found by Chan, Horne and Mann for a static spherically--symmetric
1141: space--time containing a scalar--field in a Liouville potential \cite{Cha95}%
1142: . It reduces to Schwarzschild in the limit of general relativity: $\delta
1143: \rightarrow 0$.
1144:
1145: In order to evaluate whether or not this solution is physically relevant we
1146: will proceed as follows. A dynamical systems approach will be used to
1147: establish the asymptotic attractor solutions of the field equations (\ref%
1148: {staticfield}). The field equations will then be perturbed around these
1149: asymptotic attractor solutions and solved to first order in the perturbed
1150: quantities. This linearised solution will then be treated as the physically
1151: relevant static and spherically--symmetric weak--field limit of the field
1152: equations (\ref{staticfield}) and compared with the exact solution (\ref%
1153: {Chan}).
1154:
1155: \subsection{Dynamical system}
1156:
1157: The dynamical systems approach has already been applied to a situation of
1158: this kind by Mignemi and Wiltshire \cite{Mig89}. We present a brief summary
1159: of the relevant points of their work in the above notation; for a
1160: comprehensive analysis the reader is referred to their paper.
1161:
1162: Taking the value of sign$(R)$ from (\ref{Chan}) as sign$(-\delta (1+\delta
1163: )/(1-2\delta (1+\delta )))$ and making a suitable choice of $\Omega _{0}$
1164: allows the scalar-field potential (\ref{pot1}) to be written as
1165: \begin{equation}
1166: V(\phi )=-\frac{3\delta ^{2}}{8\pi G(1-2\delta (1+\delta ))}\exp \left( {%
1167: \sqrt{\frac{16\pi G}{3}}\frac{(1-\delta )}{\delta }\phi }\right) .
1168: \label{pot}
1169: \end{equation}
1170:
1171: In four dimensions Mignemi and Wiltshire's choice of line--element
1172: corresponds to
1173: \begin{equation}
1174: d\bar{s}^{2}=e^{2U(\xi )}\left( -dt^{2}+\bar{r}^{4}(\xi )d\xi ^{2}\right) +%
1175: \bar{r}^{2}(\xi )(d\theta ^{2}+\sin ^{2}\theta d\phi ^{2}) \label{metric}
1176: \end{equation}%
1177: which, after some manipulation, gives the field equations (\ref%
1178: {conformalfield}) as
1179: \begin{align}
1180: \zeta ^{\prime \prime }& =-\frac{2c_{1}^{2}(1-\delta )^{2}+6\delta ^{2}\eta
1181: ^{\prime }{}^{2}-24\delta ^{2}\eta ^{\prime }\zeta ^{\prime }-2(1-2\delta
1182: -8\delta ^{2})\zeta ^{\prime }{}^{2}}{1-2\delta +4\delta ^{2}}-e^{2\zeta }
1183: \label{wilt1} \\
1184: \eta ^{\prime \prime }& =\frac{(1-2\delta -8\delta ^{2})(c_{1}^{2}(1-\delta
1185: )^{2}+3\delta ^{2}\eta ^{\prime }{}^{2}-12\delta ^{2}\eta ^{\prime }\zeta
1186: ^{\prime }-(1-2\delta -8\delta ^{2})\zeta ^{\prime }{}^{2})}{3\delta
1187: ^{2}(1-2\delta +4\delta ^{2})}+\frac{(1-2\delta (1+\delta ))}{3\delta ^{2}}%
1188: e^{2\zeta } \label{wilt2}
1189: \end{align}%
1190: and
1191: \begin{equation}
1192: e^{2\eta }=-\frac{(1-2\delta (1+\delta ))}{3\delta ^{2}(1-2\delta +4\delta
1193: ^{2})}\left( c_{1}^{2}(1-\delta )^{2}+3\delta ^{2}\eta ^{\prime
1194: }{}^{2}-12\delta ^{2}\eta ^{\prime }\zeta ^{\prime }-(1-2\delta -8\delta
1195: ^{2})\zeta ^{\prime }{}^{2}+(1-2\delta +4\delta ^{2})e^{2\zeta }\right)
1196: \label{wilt3}
1197: \end{equation}%
1198: where
1199: \begin{align*}
1200: \zeta (\xi )& =U(\xi )+\log \bar{r}(\xi ) \\
1201: \eta (\xi )& =-\frac{(1-2\delta (1+\delta ))}{3\delta ^{2}}U(\xi )+2\log
1202: \bar{r}(\xi )-\frac{(1-\delta )^{2}}{3\delta ^{2}}c_{1}\xi +\text{constant.}
1203: \end{align*}%
1204: Primes denote differentiation with respect to $\xi $ and $c_{1}$ is a
1205: constant of integration.
1206:
1207: Defining the variables $X$, $Y$ and $Z$ by
1208: \begin{equation*}
1209: X=\zeta ^{\prime }\qquad Y=\eta ^{\prime }\qquad Z=e^{\zeta }
1210: \end{equation*}%
1211: the field equations (\ref{wilt1}) and (\ref{wilt2}) can then be written as
1212: the following set of first-order autonomous differential equations
1213: \begin{align}
1214: X^{\prime }& =-\frac{2c_{1}^{2}(1-\delta )^{2}+6\delta ^{2}Y^{2}-24\delta
1215: ^{2}XY-2(1-2\delta -8\delta ^{2})X^{2}}{1-2\delta +4\delta ^{2}}-Z^{2}
1216: \label{X'}
1217: \\
1218: \label{Y'}
1219: Y^{\prime }& =\frac{(1-2\delta -8\delta ^{2})(c_{1}^{2}(1-\delta
1220: )^{2}+3\delta ^{2}Y^{2}-12\delta ^{2}XY-(1-2\delta -8\delta ^{2})X^{2})}{%
1221: 3\delta ^{2}(1-2\delta +4\delta ^{2})}+\frac{(1-2\delta (1+\delta ))}{%
1222: 3\delta ^{2}}Z^{2} \\
1223: Z^{\prime }& =XZ. \label{Z'}
1224: \end{align}%
1225: (The reader should note the different definition of $Z$ here to that of
1226: Mignemi and Wiltshire). As identified by Mignemi and Wiltshire, the only
1227: critical points at finite values of $X$, $Y$ and $Z$ are in the plane $Z=0$
1228: along the curve defined by
1229: \begin{equation*}
1230: c_{1}^{2}(1-\delta )^{2}+3\delta ^{2}Y^{2}-12\delta ^{2}XY-(1-2\delta
1231: -8\delta ^{2})X^{2}=0.
1232: \end{equation*}%
1233: These curves are shown as bold lines in figure \ref{finitephase}, together
1234: with some sample trajectories from equations (\ref{X'})and (\ref{Y'}). From
1235: the definition above we see that the condition $Z=0$ is equivalent to $\bar{r%
1236: }e^{U}=0$. Whilst we do not consider trajectories confined to this plane to
1237: be physically relevant we do consider the plot to be instructive as it gives
1238: a picture of the behaviour of trajectories close to this surface and
1239: displays the attractive or repulsive behaviour of the critical points, which
1240: can be the end points for trajectories which could be considered as
1241: physically meaningful.
1242: \begin{figure}[tbp]
1243: \epsfig{file=finitephase2.eps,height=8cm}
1244: \caption{\textit{The $Z=0$ plane of the phase space defined by $X$,$Y$ and $%
1245: Z $ for $\protect\delta =0.1$ and $c_{1}=0.5$. The bold lines show the
1246: critical points in this plane and the diagonal lines show the unphysical
1247: trajectories confined to this plane. The dotted line is $Y=2X$ and separates
1248: the critical points where $\protect\xi \rightarrow \infty $ from the points
1249: where $\protect\xi \rightarrow -\infty $ }}
1250: \label{finitephase}
1251: \end{figure}
1252: The dotted line in figure \ref{finitephase} corresponds to the line $Y=2X$
1253: and separates two different types of critical points. The critical points
1254: with $Y>2X$ can be seen to be repulsive to the trajectories in the $Z=0$ plane
1255: and correspond to the limit $\xi \rightarrow -\infty $. Conversely, the
1256: points with $Y<2X$ are attractive and correspond to the limit $\xi
1257: \rightarrow \infty $. As $Z=\bar{r}e^{U},$ all critical points of this type
1258: in the $Z=0$ plane correspond either to naked singularities, $\bar{r}%
1259: \rightarrow 0$, or regular horizons, $\bar{r}\rightarrow $constant.
1260:
1261: The two bold lines in figure \ref{finitephase} are the points at which the
1262: surface defined by
1263: \begin{equation*}
1264: c_1^2 (1-\delta)^2+3 \delta^2 Y^2 -12 \delta^2 X Y -(1-2 \delta-8\delta^2)
1265: X^2 +(1-2 \delta+4\delta^2) Z^2 =0
1266: \end{equation*}
1267: crosses the $Z=0$ plane. This surface splits the phase space into three
1268: separate regions between which trajectories cannot move. These regions are
1269: labelled $I$, $II$ and $III$ in figure \ref{finitephase}. It can be seen
1270: from (\ref{wilt3}) that trajectories are confined to either regions $I$ or $%
1271: II$, for the potential defined by (\ref{pot}). If we had chosen the opposite
1272: value of sign$(R)$ in (\ref{pot1}) then trajectories would be confined to
1273: region $III$. We will show, however, that region $III$ does not contain
1274: solutions with asymptotic regions in which $\bar{r} \rightarrow \infty$ and
1275: so is of limited interest for our purposes.
1276:
1277: In order to find the remaining critical points it is necessary to analyse
1278: the sphere at infinity. This can be done by making the transformation
1279: \begin{equation*}
1280: X=\rho \sin \theta \cos \phi \qquad Y=\rho \sin \theta \sin \phi \qquad
1281: Z=\rho \cos \theta
1282: \end{equation*}%
1283: and taking the limit $\rho \rightarrow \infty $. The set of equations (\ref%
1284: {X'}), (\ref{Y'}) and (\ref{Z'}) then give
1285: \begin{multline*}
1286: \frac{d\theta }{d\tau }\rightarrow -\frac{\cos \theta }{24\delta
1287: ^{2}(1-2\delta +4\delta )}(6\delta ^{2}\cos \phi (3-3\delta (2-9\delta
1288: )+(1-\delta (2+11\delta ))\cos 2\theta ) \\
1289: -(3-3\delta (4-\delta (15-22\delta -32\delta ^{2}))+(5-\delta (20-\delta
1290: (3+34\delta +32\delta ^{2})))\cos 2\theta )\sin \phi \\
1291: -2(18\delta ^{2}(1-\delta (2+7\delta ))\cos 3\phi -(1-\delta (4+\delta
1292: (9-26\delta +32\delta ^{2})))\sin 3\phi )\sin ^{2}\theta )
1293: \end{multline*}%
1294: and
1295: \begin{multline*}
1296: \frac{d\phi }{d\tau }\rightarrow -\frac{1}{24\delta ^{2}(1-2\delta +4\delta
1297: ^{2})}(6\delta ^{2}(1-\delta (2+41\delta )-5(1-\delta (2+5\delta ))\cos
1298: 2\theta )\text{cosec}\theta \sin \phi \\
1299: +2((1-\delta (4+\delta (9-26\delta +32\delta ^{2})))\cos 3\phi +18\delta
1300: ^{2}(1-\delta (2+7\delta ))\sin 3\phi )\sin \theta \\
1301: -2\cos \phi (4(1-2\delta (2-\delta (3-2\delta -4\delta ^{2})))\text{cosec}%
1302: \theta -(7-\delta (28+\delta (15-2\delta (43+128\delta ))))\sin \theta ))
1303: \end{multline*}%
1304: where $d\tau =\rho d\xi $. These equations can be used to plot the positions
1305: of critical points and trajectories on the sphere at infinity. The result of
1306: this is shown in figure \ref{infinityphase}. Once again, these trajectories
1307: do not correspond to physical solutions in the phase space but are
1308: illustrative of trajectories at large distances and help to show the
1309: attractive or repulsive nature of the critical points.
1310: \begin{figure}[tbp]
1311: \epsfig{file=infinityphase2.eps,height=8cm}
1312: \caption{\textit{The surface at infinity of the phase space defined by $X$,$%
1313: Y $ and $Z$ for $\protect\delta =0.1$. The shaded areas show where regions $I
1314: $ and $II$. Region $III$ is unshaded.}}
1315: \label{infinityphase}
1316: \end{figure}
1317: The surface at infinity has eight critical points, labelled $A$-$H$ in
1318: figure \ref{infinityphase}. Points $A$ and $B$ are the end-points of the
1319: trajectory that goes through the origin in figure \ref{finitephase} and are
1320: located at
1321: \begin{equation*}
1322: \theta =\frac{\pi }{2},\qquad \text{and}\qquad \phi _{1,(2)}=\cos
1323: ^{-1}\left( \frac{-6\delta ^{2}}{\sqrt{1-4\delta -12\delta ^{2}+32\delta
1324: ^{3}+100\delta ^{4}}}\right) (+\pi )
1325: \end{equation*}%
1326: or, in terms of the original functions in the metric (\ref{metric}),
1327: \begin{equation*}
1328: \bar{r}\rightarrow (\xi -\xi _{1})^{\frac{3\delta ^{2}}{(1-2\delta -8\delta
1329: ^{2})}}\qquad \text{and}\qquad e^{U}\rightarrow (\xi -\xi _{1})^{\frac{%
1330: 3\delta ^{2}}{(1-2\delta -8\delta ^{2})}},
1331: \end{equation*}%
1332: where $\xi _{1}$ is a constant of integration. The points $A$ and $B$
1333: therefore both correspond to $\xi \rightarrow \xi _{1}$ and hence to $\bar{r}%
1334: \rightarrow 0$.
1335:
1336: Points $C$, $D$, $E$ and $F$ are the four end points of the two curves in
1337: figure \ref{finitephase} and therefore correspond to $\xi \rightarrow \infty$
1338: or $-\infty$ and $\bar{r} \rightarrow 0$ or constant.
1339:
1340: The remaining points, $G$ and $H$, are located at
1341: \begin{equation*}
1342: \phi _{1,(2)}=\frac{\pi }{4}(+\pi )\qquad \text{and}\qquad \theta =\frac{1}{2%
1343: }\cos ^{-1}\left( -\frac{1-2\delta +10\delta ^{2}}{3-6\delta +6\delta ^{2}}%
1344: \right)
1345: \end{equation*}%
1346: or
1347: \begin{equation}
1348: \bar{r}^{\frac{(1-2\delta +4\delta ^{2})}{(1-\delta )^{2}}}\rightarrow \pm
1349: \sqrt{\frac{(1-2\delta -2\delta ^{2})}{(1-2\delta +4\delta ^{2})}}\frac{1}{%
1350: (\xi -\xi _{2})}\qquad \text{and}\qquad e^{\frac{(1-2\delta +4\delta ^{2})U}{%
1351: 3\delta ^{2}}}\rightarrow \pm \sqrt{\frac{(1-2\delta -2\delta ^{2})}{%
1352: (1-2\delta +4\delta ^{2})}}\frac{1}{(\xi -\xi _{2})}, \label{GH}
1353: \end{equation}%
1354: where $\xi _{2}$ is an integration constant, the positive branch corresponds
1355: to point $H$ and the negative branch to point $G$. These points are,
1356: therefore, the asymptotic limit of the exact solution (\ref{Chan}) and
1357: correspond to $\xi \rightarrow \xi _{2}$ and hence $\bar{r}\rightarrow
1358: \infty $.
1359:
1360: Whilst it may initially appear that trajectories are repelled from the point
1361: $H$, this is only the case in terms of the coordinate $\xi $. In terms of
1362: the more physically relevant quantity $\bar{r},$ the point $H$ is an
1363: attractor. This can be seen from the first equation in (\ref{GH}). Taking
1364: the positive branch here it can be seen that $\bar{r}$ increases as $\xi $
1365: decreases. So, in terms of $\bar{r}$ the points $G$ and $H$ are both
1366: attractors, as $\bar{r}\rightarrow \infty $.
1367:
1368: We can now see that in region $I$ all trajectories appear to start at
1369: critical points corresponding to either $\bar{r} \rightarrow 0$ or constant
1370: and end at point $G$ where $\bar{r} \rightarrow \infty$. Region $II$ appears
1371: to share the same features as region $I$ with all trajectories starting
1372: at either $\bar{r} \rightarrow 0$ or constant and ending at $H$ where $\bar{r%
1373: } \rightarrow \infty$. Region $III$ has no critical points corresponding to $%
1374: \bar{r} \rightarrow \infty$ and so all trajectories both begin and end on
1375: points corresponding to $\bar{r} \rightarrow 0$ or constant.
1376:
1377: Therefore the only solutions with an asymptotic region in which $\bar{r}%
1378: \rightarrow \infty $ exist in regions $I$ and $II$ where the potential can
1379: be described by equation (\ref{pot}). Furthermore, all trajectories in
1380: regions $I$ and $II$ appear to be attracted to the solution
1381: \begin{equation}
1382: ds^{2}=-\bar{r}^{\frac{6\delta ^{2}}{(1-\delta )^{2}}}dt^{2}+\frac{%
1383: (1-2\delta +4\delta ^{2})(1-2\delta -2\delta ^{2})}{(1-\delta )^{4}}d\bar{r}%
1384: ^{2}+\bar{r}^{2}(d\theta ^{2}+\sin ^{2}\theta d\phi ^{2}),
1385: \label{asymptotic}
1386: \end{equation}%
1387: which is the asymptotic behaviour of the solution found by Chan, Horne and
1388: Mann \cite{Cha95}. We therefore conclude that all solutions with an
1389: asymptotic region in which $\bar{r}\rightarrow \infty $ are attracted
1390: towards the solution (\ref{asymptotic}) as $\bar{r}\rightarrow \infty $.
1391:
1392: Rescaling the metric back to the original conformal frame we therefore
1393: conclude that the generic asymptotic attractor solution to the field
1394: equations, (\ref{staticfield}), is
1395: \begin{equation}
1396: ds^{2}=-r^{2\delta \frac{(1+2\delta )}{(1-\delta )}}dt^{2}+\frac{(1-2\delta
1397: +4\delta ^{2})(1-2\delta -2\delta ^{2})}{(1-\delta )^{2}}dr^{2}+r^{2}(d%
1398: \theta ^{2}+\sin ^{2}\theta d\phi ^{2}) \label{asymptotic2}
1399: \end{equation}%
1400: as $r\rightarrow \infty $. It reduces to Minkowski in the $\delta
1401: \rightarrow 0$ limit of general relativity.
1402:
1403: \subsection{Linearised solution}
1404:
1405: We now proceed to find the general solution, to first order in
1406: perturbations, around the background described by (\ref{asymptotic2}).
1407: Writing the perturbed line-element as
1408: \begin{equation}
1409: ds^{2}=-r^{2\delta \frac{(1+2\delta )}{(1-\delta )}}(1+V(r))dt^{2}+\frac{%
1410: (1-2\delta +4\delta ^{2})(1-2\delta -2\delta ^{2})}{(1-\delta )^{2}}%
1411: (1+W(r))dr^{2}+r^{2}(d\theta ^{2}+\sin ^{2}\theta d\phi ^{2})
1412: \label{asymptotic3}
1413: \end{equation}%
1414: and making no assumptions about the order of $R$ the field equations (\ref%
1415: {staticfield}) become, up to first order in $V$ and $W$,
1416: \begin{multline}
1417: \frac{\delta (1+2\delta )(1+2\delta ^{2})}{(1-\delta )^{2}r^{2}}+\frac{%
1418: (1+2\delta ^{2})}{(1-\delta )}\frac{V^{\prime }}{r}-\frac{\delta (1-2\delta )%
1419: }{2(1-\delta )}\frac{W^{\prime }}{r}+\frac{V^{\prime \prime }}{2} \label{rr}
1420: \\
1421: =\frac{\delta (1+2\delta )}{2}\frac{R^{\prime }{}^{2}}{R^{2}}-\frac{\delta
1422: (1+2\delta )}{2(1-\delta )}\frac{R^{\prime \prime }}{R}-\frac{3\delta }{%
1423: 4(1-\delta )}\frac{R^{\prime }}{R}V^{\prime }-\frac{\delta (1+2\delta
1424: )(2+\delta )}{2(1-\delta )^{2}r}\frac{R^{\prime }}{R}+\frac{\delta
1425: (1+2\delta )}{4(1-\delta )}\frac{R^{\prime }}{R}W^{\prime },
1426: \end{multline}%
1427: \begin{multline}
1428: \frac{\delta (1+2\delta )(1-2\delta -2\delta ^{2})}{(1-\delta )^{2}r^{2}}-%
1429: \frac{\delta (1+2\delta )}{(1-\delta )}\frac{V^{\prime }}{r}+\frac{(2-\delta
1430: +2\delta ^{2})}{2(1-\delta )}\frac{W^{\prime }}{r}-\frac{V^{\prime \prime }}{%
1431: 2} \label{tt} \\
1432: =-\frac{3\delta }{2}\frac{R^{\prime }{}^{2}}{R^{2}}+\frac{3\delta }{%
1433: 2(1-\delta )}\frac{R^{\prime \prime }}{R}+\frac{\delta (1+2\delta )(2-\delta
1434: +2\delta ^{2})}{2(1-\delta )^{2}r}\frac{R^{\prime }}{R}+\frac{\delta
1435: (1+2\delta )}{4(1-\delta )}\frac{R^{\prime }}{R}V^{\prime }-\frac{3\delta }{%
1436: 4(1-\delta )}\frac{R^{\prime }}{R}W^{\prime }
1437: \end{multline}%
1438: and
1439: \begin{multline}
1440: -\frac{2\delta (3-4\delta +2\delta ^{2}+8\delta ^{3})}{(1-\delta )^{2}r^{2}}+%
1441: \frac{2(1-2\delta +4\delta ^{2})(1-2\delta -2\delta ^{2})}{(1-\delta
1442: )^{2}r^{2}}W-\frac{V^{\prime }}{r}+\frac{W^{\prime }}{r} \label{thth} \\
1443: =-\delta (1+2\delta )\frac{R^{\prime }{}^{2}}{R^{2}}+\frac{\delta (1+2\delta
1444: )}{(1-\delta )}\frac{R^{\prime \prime }}{R}+\frac{\delta (4-\delta +2\delta
1445: ^{2}+4\delta ^{3})}{(1-\delta )^{2}r}\frac{R^{\prime }}{R}+\frac{\delta
1446: (1+2\delta )}{2(1-\delta )}\frac{R^{\prime }}{R}V^{\prime }-\frac{\delta
1447: (1+2\delta )}{2(1-\delta )}\frac{R^{\prime }}{R}W^{\prime }.
1448: \end{multline}
1449:
1450: Expanding $R$ to first order in $V$ and $W$ gives
1451: \begin{equation} \label{R}
1452: R=-\frac{6 \delta (1+\delta)}{(1-2\delta-2 \delta^2)} \frac{1}{r^2}+R_1
1453: \end{equation}
1454: where
1455: \begin{multline} \label{R1}
1456: R_1= \frac{2 (1+\delta+\delta^2)}{(1-2\delta-2\delta^2)} \frac{W}{r^2} -%
1457: \frac{2 (1-\delta) (1+2 \delta^2)}{(1- 2\delta-2\delta^2)(1-2 \delta+4
1458: \delta^2)} \frac{V^{\prime}}{r} \\
1459: +\frac{(1-\delta)(2-\delta+2 \delta^2)}{(1- 2\delta-2\delta^2)(1-2 \delta+4
1460: \delta^2)} \frac{W^{\prime}}{r}- \frac{(1-\delta)^2}{(1-
1461: 2\delta-2\delta^2)(1-2 \delta+4 \delta^2)} V^{\prime\prime}.
1462: \end{multline}
1463:
1464: Substituting (\ref{R}) into the field equations (\ref{rr}), (\ref{tt}) and (%
1465: \ref{thth}) and eliminating $R_1$ using (\ref{R1}) leaves
1466: \begin{multline*}
1467: \frac{(1+\delta+\delta^2)(5-12\delta+12 \delta^2+4 \delta^3)}{3 (1-\delta)^2
1468: (1+\delta)} \frac{W}{r} +\frac{(16-47\delta+76 \delta^2-34 \delta^3-16
1469: \delta^4+32 \delta^5)}{6 (1-\delta^2) (1-2\delta+4 \delta^2)} W^{\prime} \\
1470: -\frac{(1+\delta +7 \delta^2-19 \delta^3+44 \delta^4+20 \delta^5)}{3
1471: (1-\delta^2) (1-2\delta+4 \delta^2)} \frac{Y}{r} -\frac{(8-15 \delta+18
1472: \delta^2 +16 \delta^3)}{6 (1+\delta) (1-2\delta+4\delta^2)} Y^{\prime} \\
1473: = -\frac{(1-2\delta-2\delta^2)(5-12\delta+12\delta^2+4\delta^3)}{12
1474: (1-\delta^2) (1+\delta)} \frac{\psi}{r} -\frac{(1-2 \delta-2\delta^2)}{4
1475: (1-\delta^2)} \psi^{\prime},
1476: \end{multline*}
1477: \begin{multline*}
1478: -\frac{(1+2 \delta) (1+\delta+\delta^2) (3-4 \delta+4 \delta^2)}{3
1479: (1-\delta)^2 (1+\delta)} \frac{W}{r} -\frac{(1+2\delta)(2-\delta+2
1480: \delta^2)(3 -4\delta+4\delta^2)}{6 (1-\delta^2) (1-2\delta+4\delta^2)}
1481: W^{\prime} \\
1482: +\frac{(3-2\delta +17\delta^2 -4 \delta^3+40 \delta^2)}{3 (1-\delta^2) (1-2
1483: \delta+4 \delta^2)} \frac{Y}{r}+ \frac{(6-\delta+2 \delta^2+20 \delta^3)}{ 6
1484: (1+\delta) (1-2\delta+4\delta^2)} Y^{\prime} \\
1485: = \frac{(1+2 \delta) (1-2\delta-2 \delta^2) (3-4\delta+4 \delta^2)}{12
1486: (1-\delta)^2 (1+\delta)} \frac{\psi}{r} + \frac{(1+2 \delta)(1-2
1487: \delta-2\delta^2)}{12 (1-\delta^2)} \psi^{\prime}.
1488: \end{multline*}
1489: and
1490: \begin{multline*}
1491: -\frac{2 (8-8 \delta+3 \delta^2+10 \delta^3-28 \delta^4-12 \delta^5)}{3
1492: (1-\delta^2) (1+\delta) (1-2\delta-2 \delta^2)} \frac{W}{r} - \frac{(13-22
1493: \delta+12 \delta^2+26 \delta^3-56 \delta^4)}{3 (1-\delta^2) (1-2 \delta-2
1494: \delta^2)(1-2 \delta +4 \delta^2)} W^{\prime} \\
1495: +\frac{2 (4+9 \delta^2+8 \delta^3-12\delta^4)}{3 (1+\delta) (1-2
1496: \delta-2\delta^2)(1-2\delta+4 \delta^2)} \frac{Y}{r} + \frac{(5 -4\delta-4
1497: \delta^2+12 \delta^3)}{3 (1+\delta) (1-2 \delta-2\delta^2)(1-2\delta+4
1498: \delta^2)} Y^{\prime} \\
1499: = \frac{(5-4 \delta-4 \delta^2+12 \delta^3)}{6 (1-\delta)^2 (1+\delta)}
1500: \frac{\psi}{r} +\frac{(1+2 \delta)}{6 (1-\delta^2)} \psi^{\prime}
1501: \end{multline*}
1502: where $Y= r V^{\prime}$ and $\psi=r^3 R_1^{\prime}$, subject to the
1503: constraint (\ref{R1}).
1504:
1505: For $-\frac{(7+3 \sqrt{21})}{20}<\delta<-\frac{(7-3\sqrt{21})}{20}$ the
1506: general solution to this first order set of coupled equations is given, in
1507: terms of $V$ and $W$, by
1508: \begin{align} \label{V}
1509: V(r) &= c_1 V_1(r)+c_2 V_2(r)+ c_3 V_3(r) +\text{constant} \\
1510: W(r) &= -c_1 V_1(r)+c_2 W_2(r)+c_3 W_3(r)
1511: \end{align}
1512: where
1513: \begin{align*}
1514: V_1 &=-r^{-\frac{(1-2 \delta+4 \delta^2)}{(1-\delta)}} \\
1515: V_2 &= \frac{ (1+2 \delta) r^{-\frac{(1-2 \delta+4 \delta^2)}{2 (1-\delta)}}%
1516: }{2 (2-3 \delta+12 \delta^2+16 \delta^3)} \left( (1+2 \delta)^2 \sin (A \log
1517: r) + 2 A (1-\delta) \cos(A \log r) \right) \\
1518: W_2 &= r^{-\frac{(1-2 \delta+4 \delta^2)}{2 (1-\delta)}} \sin (A \log r) \\
1519: V_3 &= \frac{ (1+2 \delta) r^{-\frac{(1-2 \delta+4 \delta^2)}{2 (1-\delta)}}%
1520: }{2 (2-3 \delta+12 \delta^2+16 \delta^3)} \left( (1+2 \delta)^2 \cos (A \log
1521: r) - 2 A (1-\delta) \sin(A \log r) \right) \\
1522: W_3 &= r^{-\frac{(1-2 \delta+4 \delta^2)}{2 (1-\delta)}} \cos (A \log r)
1523: \end{align*}
1524: and
1525: \begin{equation*}
1526: A= -\frac{\sqrt{7-28 \delta+36 \delta^2-16 \delta^3-80 \delta^4}}{2
1527: (1-\delta)}.
1528: \end{equation*}
1529: The extra constant in (\ref{V}) is from the integration of $Y$ and can be
1530: trivially absorbed into the definition of the time coordinate. The above
1531: solution satisfies the constraint (\ref{R1}) without imposing any conditions
1532: upon the arbitrary constants $c_1$, $c_2$ and $c_3$.
1533:
1534: It can be seen by direct comparison that the constant $c_1$ is linearly
1535: related to the constant $C$ in (\ref{Chan}) by a factor that is a function
1536: of $\delta$ only. The constants $c_2$ and $c_3$ correspond to two new
1537: oscillating modes.
1538:
1539: \subsection{Physical consequences}
1540:
1541: In order to calculate the classical tests of metric theories of gravity
1542: (i.e. bending and time-delay of light rays and the perihelion precession of
1543: Mercury) we require the static and spherically symmetric solution to the
1544: field equations (\ref{field}). Due to the complicated form of these
1545: equations we are unable to find the general solution; instead we propose to
1546: use the first--order solution around the generic attractor as $r\rightarrow
1547: \infty $. This method should be applicable to gravitational experiments
1548: performed in the solar system as the gravitational field in this region can
1549: be considered weak and we will be considering experiments performed at large
1550: $r$ (in terms of the Schwarzschild radius of the massive objects in the
1551: system). To this end we will use the solution found at the end of the
1552: previous subsection. We choose to arbitrarily set the constants $c_{2}$ and $%
1553: c_{3}$ to zero - this removes the oscillatory parts of the solution, and
1554: hence ensures that the gravitational force is always attractive. This
1555: considerable simplification of the solution also allows a straightforward
1556: calculation of both null and timelike geodesics which can be used to compute
1557: the outcomes of the classical tests in this theory.
1558:
1559: \subsubsection{Solution in isotropic coordinates}
1560:
1561: Having removed the oscillatory parts of the solution we are left with the
1562: part corresponding to the exact solution (\ref{Chan}). Making the coordinate
1563: transformation
1564: \begin{equation*}
1565: r^{(1-2 \delta+4 \delta^2)/(1-d)}= \left(1-\frac{C}{4 \hat{r}^{\sqrt{\frac{%
1566: (1-2 \delta+4 \delta^2)}{(1-2 \delta-2 \delta^2)}}}} \right)^2 \hat{r}^{%
1567: \sqrt{\frac{(1-2 \delta+4 \delta^2)}{(1-2 \delta-2 \delta^2)}}}
1568: \end{equation*}
1569: the solution (\ref{Chan}) can be transformed into the isotropic coordinate
1570: system
1571: \begin{equation} \label{iso}
1572: ds^2=-A(\hat{r}) d t^2+B(\hat{r}) ( d\hat{r}^2+ \hat{r}^2 (d
1573: \theta^2+\sin^2\theta d \phi^2))
1574: \end{equation}
1575: where
1576: \begin{equation*}
1577: A(\hat{r})= \hat{r}^{\frac{2 \delta (1+2 \delta)}{\sqrt{(1-2
1578: \delta-2\delta^2)(1-2\delta+4 \delta^2)}}} \left(1+\frac{C}{4 \hat{r}^{\sqrt{%
1579: \frac{(1-2 \delta+4 \delta^2)}{(1-2 \delta-2 \delta^2)}}}} \right)^2 \left(1-%
1580: \frac{C}{4 \hat{r}^{\sqrt{\frac{(1-2 \delta+4 \delta^2)}{(1-2 \delta-2
1581: \delta^2)}}}} \right)^{-\frac{2 (1+4 \delta)}{(1-2 \delta+4 \delta^2)}}
1582: \end{equation*}
1583: and
1584: \begin{equation*}
1585: B(\hat{r})= \hat{r}^{-2+2\frac{(1-\delta)}{\sqrt{(1-2
1586: \delta-2\delta^2)(1-2\delta+4 \delta^2)}}} \left(1-\frac{C}{4 \hat{r}^{\sqrt{%
1587: \frac{(1-2 \delta+4 \delta^2)}{(1-2 \delta-2 \delta^2)}}}} \right)^{\frac{4
1588: (1-\delta)}{(1-2 \delta+4 \delta^2)}},
1589: \end{equation*}
1590: which is, to linear order in $C$,
1591: \begin{equation*}
1592: A(\hat{r})= \hat{r}^{\frac{2 \delta (1+2 \delta)}{\sqrt{(1-2
1593: \delta-2\delta^2)(1-2\delta+4 \delta^2)}}} \left(1+ \frac{(1-\delta) (1-2
1594: \delta)}{(1-2 \delta+4 \delta^2)}\frac{C}{\hat{r}^{\sqrt{\frac{(1-2 \delta+4
1595: \delta^2)}{(1-2 \delta-2 \delta^2)}}}} \right)
1596: \end{equation*}
1597: and
1598: \begin{equation*}
1599: B(\hat{r})= \hat{r}^{-2+2\frac{(1-\delta)}{\sqrt{(1-2
1600: \delta-2\delta^2)(1-2\delta+4 \delta^2)}}} \left(1- \frac{(1-\delta)}{(1-2
1601: \delta+4 \delta^2)}\frac{C}{\hat{r}^{\sqrt{\frac{(1-2 \delta+4 \delta^2)}{%
1602: (1-2 \delta-2 \delta^2)}}}} \right).
1603: \end{equation*}
1604:
1605: \subsubsection{Newtonian limit}
1606:
1607: We first investigate the Newtonian limit of the geodesic equation in order
1608: to set the constant $C$ in the solution (\ref{iso}) above. As usual, we have
1609: \begin{equation*}
1610: \Phi _{,\mu }=\Gamma _{\;00}^{\mu }
1611: \end{equation*}%
1612: where $\Phi $ is the Newtonian gravitational potential. Substituting in the
1613: isotropic metric (\ref{iso}) this gives
1614: \begin{align}
1615: \nabla \Phi & =\frac{\nabla A(\hat{r})}{2B(\hat{r})} \label{motion} \\
1616: & =\frac{\delta (1+2\delta )\hat{r}^{1-2\sqrt{\frac{1-2\delta -2\delta ^{2}}{%
1617: 1-2\delta +4\delta ^{2}}}}}{\sqrt{(1-2\delta -2\delta ^{2})(1-2\delta
1618: +4\delta ^{2})}}-\frac{(1-\delta )(1-8\delta +4\delta ^{2})C\hat{r}^{1-\frac{%
1619: 3(1-2\delta )}{\sqrt{(1-2\delta -2\delta ^{2})(1-2\delta +4\delta ^{2})}}}}{2%
1620: \sqrt{(1-2\delta -2\delta ^{2})(1-2\delta +4\delta ^{2})^{3}}}+O(C^{2}).
1621: \end{align}
1622:
1623: The second term in the expression goes as $\sim \hat{r}^{-2+O(\delta ^{2})}$
1624: and so corresponds to the Newtonian part of the gravitational force. The
1625: first term, however, goes as $\sim \hat{r}^{-1+O(\delta ^{2})}$ and has no
1626: Newtonian counterpart. In order for the Newtonian part to dominate over the
1627: non-Newtonian part we must impose upon $\delta $ the requirement that it is
1628: at most
1629: \begin{equation*}
1630: \delta \sim O\left( \frac{C}{r}\right) .
1631: \end{equation*}%
1632: If $\delta $ were larger than this then the non-Newtonian part of the
1633: potential would dominate over the Newtonian part, which is clearly
1634: unacceptable at scales over which the Newtonian potential has been measured
1635: and shown to be accurate.
1636:
1637: This requirement upon the order of magnitude of $\delta$ allows (\ref{motion}%
1638: ) to be written
1639: \begin{equation} \label{force}
1640: \nabla \Phi = \frac{\delta}{\hat{r}^{1+O(C^2)}}-\frac{C}{2 \hat{r}^{2+O(C^2)}%
1641: }+O(C^2)
1642: \end{equation}
1643: where expansions in $C$ have been carried out separately in the coefficients
1644: and the powers of $\hat{r}$ of the two terms.
1645:
1646: Comparison of (\ref{force}) with the Newtonian force law
1647: \begin{equation*}
1648: \nabla \Phi_N=\frac{G m}{r^2}
1649: \end{equation*}
1650: allows the value of $C$ to be read off as
1651: \begin{equation*}
1652: C=-2 G m +O(\delta).
1653: \end{equation*}
1654:
1655: \subsubsection{Post--Newtonian limit}
1656:
1657: We now wish to calculate, to post--Newtonian order, the equations of motion
1658: for test particles in the metric (\ref{iso}). The geodesic equation can be
1659: written in its usual form
1660: \begin{equation*}
1661: \frac{d^{2}x^{\mu }}{d\lambda ^{2}}+\Gamma _{\;ij}^{\mu }\frac{dx^{i}}{%
1662: d\lambda }\frac{dx^{j}}{d\lambda }=0,
1663: \end{equation*}%
1664: where $\lambda $ can be taken as proper time for a timelike geodesic or as
1665: an affine parameter for a null geodesic. In terms of coordinate time this
1666: can be written
1667: \begin{equation}
1668: \frac{d^{2}x^{\mu }}{dt^{2}}+\left( \Gamma _{\;ij}^{\mu }-\Gamma _{\;ij}^{0}%
1669: \frac{dx^{\mu }}{dt}\right) \frac{dx^{i}}{dt}\frac{dx^{j}}{dt}=0.
1670: \label{motion2}
1671: \end{equation}%
1672: We also have the integral
1673: \begin{equation}
1674: g_{ij}\frac{dx^{i}}{dt}\frac{dx^{j}}{dt}=S \label{motion3}
1675: \end{equation}%
1676: where $S=-1$ for particles and $0$ for photons.
1677:
1678: Substituting (\ref{iso}) into (\ref{motion2}) and (\ref{motion3}) gives, to
1679: the relevant order, the equations of motion
1680: \begin{multline}
1681: \frac{d^{2}\mathbf{x}}{dt^{2}}=-\frac{Gm}{r^{2}}\left( 1+\left\vert \frac{d%
1682: \mathbf{x}}{dt}\right\vert ^{2}\right) \mathbf{e_{r}}+4\frac{G^{2}m^{2}}{%
1683: r^{3}}\mathbf{e_{r}}+4\frac{Gm}{r^{2}}\mathbf{e_{r}}\cdot \frac{d\mathbf{x}}{%
1684: dt}\frac{d\mathbf{x}}{dt} \label{geo1} \\
1685: -\frac{\delta }{r}\left( 1-\left\vert \frac{d\mathbf{x}}{dt}\right\vert
1686: ^{2}\right) \mathbf{e_{r}}-4\frac{\delta ^{2}}{r}\mathbf{e_{r}}+\delta \frac{%
1687: Gm}{r^{2}}\mathbf{e_{r}}+O(G^{3}m^{3})
1688: \end{multline}%
1689: and
1690: \begin{equation}
1691: \left\vert \frac{d\mathbf{x}}{dt}\right\vert ^{2}=1-4\frac{Gm}{r}+\frac{S}{%
1692: r^{2\delta }}-2S\frac{Gm}{r^{1+2\delta }}+O(G^{2}m^{2}). \label{geo2}
1693: \end{equation}%
1694: (In the interests of concision we have excluded the $O(\delta ^{2})$ terms
1695: from the powers of $r$, the reader should regard them as being there
1696: implicitly). The first three terms in equation (\ref{geo1}) are identical to
1697: their general-relativistic counterparts. The next two terms are completely
1698: new and have no counterparts in general relativity. The last term in
1699: equation (\ref{geo1}) can be removed by rescaling the mass term by $%
1700: m\rightarrow m(1+\delta )$; this has no effect on the Newtonian limit of the
1701: geodesic equation as any term $Gm\delta $ is of post--Newtonian order.
1702:
1703: \subsubsection{The bending of light and time delay of radio signals}
1704:
1705: From equation (\ref{geo2}) it can be seen that the solution for null
1706: geodesics, to zeroth order, is a straight line that can be parametrised by
1707: \begin{equation*}
1708: \mathbf{x}=\mathbf{n}(t-t_{0})
1709: \end{equation*}%
1710: where $\mathbf{n}\cdot \mathbf{n}=1$. Considering a small departure from the
1711: zeroth order solution we can write
1712: \begin{equation*}
1713: \mathbf{x}=\mathbf{n}(t-t_{0})+\mathbf{x}_{1}
1714: \end{equation*}%
1715: where $\mathbf{x}_{1}$ is small. To first order, the equations of motion (%
1716: \ref{geo1}) and (\ref{geo2}) then become
1717: \begin{equation}
1718: \frac{d^{2}\mathbf{x}}{dt^{2}}=-2\frac{Gm}{r^{2}}\mathbf{e_{r}}+4\frac{Gm}{%
1719: r^{2}}(\mathbf{n}\cdot \mathbf{e_{r}})\mathbf{n} \label{geo3}
1720: \end{equation}%
1721: and
1722: \begin{equation}
1723: \mathbf{n}\cdot \frac{d\mathbf{x}}{dt}=-2\frac{Gm}{r}. \label{geo4}
1724: \end{equation}
1725:
1726: Equations (\ref{geo3}) and (\ref{geo4}) can be seen to be identical to the
1727: first-order equations of motions for photons in general relativity. We
1728: therefore conclude that any observations involving the motion of photons in
1729: a stationary and spherically symmetric weak field situation cannot tell any
1730: difference between general relativity and this $R^{1+\delta }$ theory, to
1731: first post--Newtonian order. This includes the classical light bending and
1732: time delay tests which should measure the post-Newtonian parameter $\gamma $ to be one
1733: in this theory, as in general relativity.
1734:
1735: \subsubsection{Perihelion precession}
1736:
1737: In calculating the perihelion precession of a test particle in the geometry (%
1738: \ref{iso}) it is convenient to use the standard procedures for computing the
1739: perturbations of orbital elements (see \cite{Sma53} and \cite{Rob68}). In
1740: the notation of Robertson and Noonan \cite{Rob68} the measured rate of
1741: change of the perihelion in geocentric coordinates is given by
1742: \begin{equation}
1743: \frac{d\tilde{\omega}}{dt}=-\frac{p\mathcal{R}}{he}\cos \phi +\frac{\mathcal{%
1744: J}(p+r)}{he}\sin \phi \label{per}
1745: \end{equation}%
1746: where $p$ is the semi--latus rectum of the orbit, $h$ is the
1747: angular--momentum per unit mass, $e$ is the eccentricity and $\mathcal{R}$
1748: and $\mathcal{J}$ are the components of the acceleration in radial and
1749: normal to radial directions in the orbital plane, respectively. The radial
1750: coordinate, $r$, is defined by
1751: \begin{equation}
1752: r\equiv \frac{p}{(1+e\cos \phi )} \label{r}
1753: \end{equation}%
1754: and $\phi $ is the angle measured from the perihelion. We have, as usual,
1755: the additional relations
1756: \begin{equation*}
1757: p=a(1-e^{2})
1758: \end{equation*}%
1759: and
1760: \begin{equation}
1761: h\equiv \sqrt{Gmp}\equiv r^{2}\frac{d\phi }{dt}. \label{dr}
1762: \end{equation}
1763:
1764: From (\ref{geo1}), the components of the acceleration can be read off as
1765: \begin{equation}
1766: \mathcal{R}=-\frac{Gm}{r^{2}} -\frac{Gm}{r^2} v^2 +4\frac{Gm}{r^{2}}v_{\mathcal{R}}^{2}+4\frac{%
1767: G^{2}m^{2}}{r^{3}}-\frac{\delta }{r}+\frac{\delta }{r}v^{2}-4\frac{\delta
1768: ^{2}}{r} \label{Raccel}
1769: \end{equation}%
1770: and
1771: \begin{equation}
1772: \mathcal{J}=4\frac{Gm}{r^{2}}v_{\mathcal{R}}v_{\mathcal{J}} \label{Jaccel}
1773: \end{equation}%
1774: where we now have the radial and normal-to-radial components of the velocity
1775: as
1776: \begin{align*}
1777: v_{\mathcal{R}}& =\frac{eh}{p}\sin \phi \\
1778: v_{\mathcal{J}}& =\frac{h}{p}(1+e\cos \phi )
1779: \end{align*}%
1780: and $v^{2}=v_{\mathcal{R}}^{2}+v_{\mathcal{J}}^{2}$. In writing (\ref{Raccel}%
1781: ), the last term of (\ref{geo1}) has been absorbed by a rescaling of $m$, as
1782: mentioned above.
1783:
1784: The expressions (\ref{Raccel}) and (\ref{Jaccel}) can now be substituted
1785: into (\ref{per}) and integrated from $\phi=0$ to $2 \pi$, using (\ref{r})
1786: and (\ref{dr}) to write $r$ and $dr$ in terms of $\phi$ and $d\phi$. The
1787: perihelion precession per orbit is then given, to post--Newtonian accuracy,
1788: by the expression
1789: \begin{equation} \label{precession}
1790: \Delta \tilde{\omega} = \frac{6 \pi G m}{a(1-e^2)}-\frac{2\pi\delta}{e^2}
1791: \left( e^2-1-\frac{(1+4 \delta) a (1-e^2)}{Gm} \right).
1792: \end{equation}
1793: The first term in (\ref{precession}) is clearly the standard general
1794: relativistic expression. The second term is new and contributes to leading
1795: order the term
1796: \begin{equation*}
1797: \frac{2 \pi a}{Gm} \left( \frac{1-e^2}{e^2} \right) \delta.
1798: \end{equation*}
1799:
1800: Comparing the prediction (\ref{precession}) with observation is a
1801: non-trivial matter. The above prediction is the highly idealised precession
1802: expected for a timelike geodesic in the geometry described by (\ref{iso}).
1803: If we assume that the geometry (\ref{iso}) is a good approximation to the
1804: weak field for a static Schwarzschild--like mass then it is not trivial to
1805: assume that the timelike geodesics used to calculate the rate of perihelion
1806: precession (\ref{precession}) are the paths that material objects will
1807: follow. Whilst we are assured from the generalised Bianchi identities \cite%
1808: {Mag94} of the covariant conservation of energy--momentum, $%
1809: T_{\;\;;b}^{ab}=0 $, and hence the geodesic motion of an ideal fluid of
1810: pressureless dust, $u^{i}u_{\;;i}^{j}=0$, this does not ensure the geodesic
1811: motion of extended bodies. This deviation from geodesic motion is known as
1812: the Nordvedt effect \cite{Nor68} and, whilst being zero for general
1813: relativity, is generally non--zero for extended theories of gravity. From
1814: the analysis so far it is also not clear how orbiting matter and other
1815: nearby sources (other than the central mass) will contribute to the geometry
1816: (\ref{iso}).
1817:
1818: In order to make a prediction for a physical system such as the solar
1819: system, and in the interests of brevity, some assumptions must be made. It
1820: is firstly assumed that the geometry of space--time in the solar system can
1821: be considered, to first approximation, as static and spherically symmetric.
1822: It is then assumed that this geometry is determined by the Sun, which can be
1823: treated as a point-like Schwarzschild mass at the origin, and is isolated
1824: from the effects of matter outside the solar system and from the background
1825: cosmology. It is also assumed that the Nordvedt effect is negligible and
1826: that extended massive bodies, such as planets, follow the same timelike
1827: geodesics of the background geometry as neutral test particles.
1828:
1829: In comparing with observation it is useful to recast (\ref{precession}) in
1830: the form
1831: \begin{equation*}
1832: \Delta \tilde{\omega}=\frac{6\pi Gm}{a(1-e^{2})}\lambda
1833: \end{equation*}%
1834: where
1835: \begin{equation*}
1836: \lambda =1+\frac{a^{2}(1-e^{2})^{2}}{3G^{2}m^{2}e^{2}}\delta .
1837: \end{equation*}%
1838: This allows for easy comparison with results which have been used to
1839: constrain the standard post-Newtonian parameters, for which
1840: \begin{equation*}
1841: \lambda =\frac{1}{3}(2+2\gamma -\beta ).
1842: \end{equation*}%
1843: The observational determination of the perihelion precession of Mercury is
1844: not clear cut and is subject to a number of uncertainties; most notably the
1845: quadrupole moment of the Sun (see e.g. \cite{Pir03}). We choose to use the
1846: result of Shapiro et. al. \cite{Sha76}
1847: \begin{equation}
1848: \lambda =1.003\pm 0.005 \label{Shapiro}
1849: \end{equation}%
1850: which for standard values of $a$, $e$ and $m$ \cite{All63} gives us the
1851: constraint
1852: \begin{equation}
1853: \delta =2.7\pm 4.5\times 10^{-19}.
1854: \end{equation}%
1855: In deriving (\ref{Shapiro}) the quadrupole moment of the Sun was assumed to
1856: correspond to uniform rotation. For more modern estimates of the
1857: anomalous perihelion advance of Mercury see \cite{Pir03}.
1858:
1859: \section{Conclusions}
1860:
1861: We have considered here the modification to the gravitational Lagrangian $%
1862: R\rightarrow R^{1+\delta }$, where $\delta $ is a small rational number. By
1863: considering the idealised Friedmann--Robertson--Walker cosmology and the
1864: static and spherically symmetric weak field situations we have been able to
1865: determine suitable solutions to the field equations which we have used to
1866: make predictions of the consequences of this gravity theory for
1867: astrophysical processes. These predictions have been compared to
1868: observations to derive a number of bounds on the value of $\delta $.
1869:
1870: Firstly, we showed that for a spatially-flat, matter-dominated universe the
1871: attractor solution for the scale factor as $t\rightarrow \infty $ is of the
1872: form $a(t)\propto t^{\frac{1}{2}}$ if\ $-\frac{1}{4}<\delta <0$. This is
1873: unacceptable as sub--horizon scale density perturbations do not grow in a
1874: universe described by a scale factor of this form. We therefore have the
1875: constraints
1876: \begin{equation}
1877: \delta \geqslant 0\qquad (\text{or}\quad \delta <-1/4),
1878: \end{equation}%
1879: in which case the attractor solution for the scale factor as $t\rightarrow
1880: \infty $ changes to that of the exact solution $a(t)\propto t^{\frac{%
1881: 2(1+\delta )}{3}}$.
1882:
1883: Secondly, we showed that the modified expansion rate during primordial
1884: nucleosynthesis alters the predicted abundances of light elements in the
1885: universe. Using the inferred observational abundances of Olive et. al. \cite%
1886: {Oli00} we were able to impose upon $\delta $ the constraints
1887: \begin{equation}
1888: -0.017\leqslant \delta \leqslant 0.0012,
1889: \end{equation}%
1890: for $0.5\leqslant \eta _{10}\leqslant 50$, or
1891: \begin{equation}
1892: -0.0064\leqslant \delta \leqslant 0.0012,
1893: \end{equation}%
1894: for $1\leqslant \eta _{10}\leqslant 10$.
1895:
1896: Next, we considered the horizon size at the time of matter--radiation
1897: equality. After showing that the horizon size is different in this theory to
1898: its counterpart in general-relativistic cosmology we discussed the
1899: implications for microwave background observations. This argument runs in
1900: parallel to that of Liddle et. al. for the Brans--Dicke cosmology \cite%
1901: {Lid98}. The horizon size at matter--radiation equality will be shifted by $%
1902: \sim 1\%$ for a value of $\delta \sim 0.0005.$
1903:
1904: Finally, we investigated the static and spherically symmetric weak--field
1905: situation. We calculated the null and timelike geodesics of the space--time
1906: to post--Newtonian accuracy. We then showed that null geodesics are, to the
1907: required accuracy, identical in this theory to those in the Schwarzschild
1908: solution of general relativity. The light bending and radio time--delay
1909: tests should, therefore, yield the same results as in general relativity, to
1910: the required order.
1911:
1912: Our prediction for the perihelion precession of Mercury gave us our tightest
1913: bounds on $\delta $. Assuming that Mercury follows timelike geodesics of the
1914: space--time we used the results of Shapiro et. al. \cite{Sha76} to impose
1915: upon $\delta $ the constraint
1916: \begin{equation}
1917: \delta =2.7\pm 4.5\times 10^{-19}.
1918: \end{equation}%
1919: This constraint is due to the unusual feature of the static and
1920: spherically--symmetric space--time that as $r\rightarrow \infty $ it is
1921: asymptotically attracted to a form that is not Minkowski spacetime, but
1922: reduces to Minkowski spacetime as $\delta \rightarrow 0$.
1923:
1924: Combining the above results we therefore have that $\delta $ should be
1925: constrained to lie within the range
1926: \begin{equation}
1927: 0\leq \delta <7.2\times 10^{-19}.
1928: \end{equation}
1929: This is a remarkably strong observational constraint upon deviations of this kind from
1930: general relativity.
1931:
1932: \section{Acknowledgements}
1933:
1934: We would like to thank Robert Scherrer, David Wiltshire and Spiros
1935: Cotsakis for helpful comments and suggestions. TC is supported by the PPARC.
1936:
1937: \begin{thebibliography}{99}
1938: \bibitem{edd} A.S. Eddington, \textit{The Mathematical Theory of Relativity,
1939: 2nd. edn. }Cambridge UP, Cambridge, (1924).
1940:
1941: \bibitem{gur} V.T. Gurovich, Sov. Phys. JETP \textbf{46}, 193 (1977).
1942:
1943: \bibitem{buch} H. Buchdahl, J. Phys. A \textbf{12}, 1229 (1979).
1944:
1945: \bibitem{kerner} Kerner, R. Gen. Rel. Gravn. \textbf{14}, 453 (1982).
1946:
1947: \bibitem{BO} J.D. Barrow and A.C. Ottewill, J. Phys. A \textbf{16}, 2757
1948: (1983).
1949:
1950: \bibitem{magg} G. Magnano, M. Ferraris and M. Francaviglia, Gen. Rel. Gravn.
1951: \textbf{19}, 465 (1987).
1952:
1953: \bibitem{tur} S. M. Carroll, A. De Felice, V. Duvvuri, D. A. Easson, M.
1954: Trodden, and M. S. Turner, Phys. Rev. D \textbf{71}, 063513 (2005).
1955:
1956: \bibitem{new} S. Nojiri and S. D. Odintsov, Phys. Rev. D \textbf{68},
1957: 123512 (2003).
1958:
1959: \bibitem{CB} J.D. Barrow and S. Cotsakis, Phys. Lett. B \textbf{214} 515
1960: (1994).
1961:
1962: \bibitem{maeda} K-I. Maeda, Phys. Rev. D \textbf{39}, 3159 (1989)
1963:
1964: \bibitem{ruz} T.V. Ruzmaikina and A.A. Ruzmaikin, Sov. Phys. JETP \textbf{30
1965: }, 372 (1970).
1966:
1967: \bibitem{Buc70} H. A. Buchdahl, Mon. Not. R. Astron. Soc. \textbf{150}, 1
1968: (1970).
1969:
1970: \bibitem{rox} I. Roxburgh, Gen. Rel. Gravn. \textbf{8}, 219 (1977).
1971:
1972: \bibitem{Bar05} J. D. Barrow and T. Clifton, gr-qc/0509085.
1973:
1974: \bibitem{Olmo} G. J. Olmo, Phys. Rev. D \textbf{72}, 083505 (2005).
1975:
1976: \bibitem{Mag94} G. Magnano and M. Sokolowski, Phys. Rev. D \textbf{50}, 5039
1977: (1994).
1978:
1979: \bibitem{schmidt} H.J. Schmidt, gr-qc/0407095.
1980:
1981: \bibitem{newt} J.D. Barrow, Mon. Not. R. astr. Soc., \textbf{282}, 1397
1982: (1996).
1983:
1984: \bibitem{Car04} S. Carloni, P. K. S. Dunsby, S. Capoziello and A, Troisi,
1985: gr-qc/0410046.
1986:
1987: \bibitem{Hol98} D. J. Holden and D. Wands, Class. Quant. Grav. \textbf{15},
1988: 3271 (1998).
1989:
1990: \bibitem{shv} V.F. Shvartsman, Sov. Phys. JETP Lett. \textbf{9}, 184 (1969).
1991:
1992: \bibitem{nus} G. Steigman, hep-ph/0501100.
1993:
1994: \bibitem{ht} S.W. Hawking and R.J. Tayler, Nature \textbf{209}, 1278 (1966).
1995:
1996: \bibitem{jb1} J.D. Barrow, Mon. Not. Roy. astr. Soc., \textbf{175}, 359
1997: (1976).
1998:
1999: \bibitem{jb2} J.D. Barrow, Mon. Not. Roy. astr. Soc., \textbf{178}, 625
2000: (1977).
2001:
2002: \bibitem{jb3} J.D. Barrow, Mon. Not. Roy. astr. Soc., \textbf{211}, 221
2003: (1984).
2004:
2005: \bibitem{mag} J.D. Barrow, P.G. Ferreira, J. Silk, Phys. Rev. Lett. \textbf{%
2006: 78}, 3610 (1997).
2007:
2008: \bibitem{YM} J.D. Barrow, Y. Jin, and K-I. Maeda, Phys. Rev. D
2009: \textbf{72}, 103512 (2005).
2010:
2011: \bibitem{skew} J.D. Barrow, Phys. Rev. D \textbf{55}, 7451, (1997).
2012:
2013: \bibitem{bmaeda} J.D. Barrow and K-I. Maeda, Nucl. Phys. B \textbf{341}, 294
2014: (1990).
2015:
2016: \bibitem{clift} T. Clifton, J.D. Barrow and R. Scherrer, Phys. Rev. D
2017: \textbf{71}, 123526 (2005).
2018:
2019: \bibitem{Car02} S. M. Carroll and M. Kaplinghat, Phys. Rev. D \textbf{65},
2020: 063507 (2002).
2021:
2022: \bibitem{Ben03} C. L. Bennett et. al., Astrophys. J. Suppl. \textbf{148}, 1
2023: (2003).
2024:
2025: \bibitem{Oli00} K. A. Olive, G. Steigman and T. P. Walker, Physics Reports
2026: \textbf{333}, 389 (2000).
2027:
2028: \bibitem{Lid98} A. R. Liddle, A. Mazumdar and J. D. Barrow, Phys. Rev. D
2029: \textbf{58} 027302 (1998).
2030:
2031: \bibitem{Fix96} D. J. Fixsen, E. S. Cheng, J. M. Gales, J. C. Mather, R. A.
2032: Shafer and E. L. Wright, Astrophys. J. \textbf{473}, 576 (1996).
2033:
2034: \bibitem{Che99} X. Chen and M. Kamionkowski, Phys. Rev. D \textbf{60},
2035: 104036 (1999).
2036:
2037: \bibitem{Cha95} K. C. K. Chan, J. H. Horne and R. B. Mann, Nucl. Phys.
2038: \textbf{B447}, 441 (1995).
2039:
2040: \bibitem{Mig89} S. Mignemi and D. L. Wiltshire, Class. Quantum Grav. \textbf{%
2041: 6}, 987 (1989).
2042:
2043: \bibitem{Sma53} W. M. Smart, \textit{Celestial Mechanics}, Longman Green,
2044: London (1953).
2045:
2046: \bibitem{Rob68} H. P. Robertson and T. W. Noonan, \textit{Relativity and
2047: Cosmology}, Saunders, Philadelphia (1968).
2048:
2049: \bibitem{Nor68} K. Nordvedt, Phys. Rev. \textbf{169}, 1014 (1968); K.
2050: Nordvedt, Phys. Rev. \textbf{169}, 1017 (1968).
2051:
2052: \bibitem{Pir03} S. Pireaux, J. P. Rozelot and S. Godier, Astrophys. Space
2053: Sci. \textbf{284}, 1159 (2003).
2054:
2055: \bibitem{Sha76} I. I. Shapiro, C. C. Counselman III and R. W. King, Phys.
2056: Rev. Lett. \textbf{36}, 555 (1976).
2057:
2058: \bibitem{All63} C. W. Allen, \textit{Astrophysical quantities}, The Athlone
2059: Press, London (1965).
2060:
2061: \end{thebibliography}
2062:
2063: \end{document}
2064: