0807.3577/ba2.tex
1: \documentclass[12pt]{iopart}
2: \usepackage{iopams}
3: \usepackage{epsfig}   
4: \usepackage{graphics}
5: \usepackage[T1]{fontenc}
6: 
7:  \newcommand{\W}{{\sf \bfseries {\Large [}Was:~{\Large [}}}
8:  \newcommand{\N}{{\sf \bfseries {\Large ]}Now:~{\Large [}}}
9:  \newcommand{\I}{{\sf \bfseries {\Large [}Insert:~{\Large [}}}
10:  \newcommand{\R}{{\sf \bfseries {\Large [}Remove:~{\Large [}}}
11:  \newcommand{\C}{{\sf \bfseries {\Large [}Comment:~{\Large [}}\sf}
12:  \newcommand{\F}{\normalfont {\sf \bfseries {\Large ]]} }}
13:   \newcommand{\RW}{{\sf \bfseries {\Large [}Re-written:~{\Large [}}}
14: 
15: 
16: \newcommand{\av}[1]{\langle{#1}\rangle}
17: 
18: 
19: \begin{document} 
20: 
21: \title{Apparent and average acceleration of the Universe}
22: 
23: \author{Krzysztof Bolejko$^{1,2}$ and Lars Andersson$^{3,4}$}
24: 
25: \address{$^1$School of Physics, University of Melbourne, VIC 3010, Australia}
26: \address{$^2$Nicolaus Copernicus Astronomical Center, Bartycka 18, 00-716 Warsaw, Poland}
27: \address{$^3$Department of Mathematics, University of Miami, Coral Gables, FL 33124, USA}
28: \address{$^4$Albert Einstein Institute, Am M\"uhlenberg 1, D-14467 Golm, Germany}
29: 
30: \ead{\mailto{bolejko@camk.edu.pl}, \mailto{larsa@math.miami.edu}}
31: 
32: \begin{abstract}
33: In this paper we consider the relation between the volume deceleration parameter
34: obtained within the Buchert averaging scheme and the deceleration parameter
35: derived from the supernova observation. This work was motivated by recent
36: findings that showed that there are models which despite $\Lambda=0$ have 
37: volume deceleration parameter $q^{vol} < 0$. This opens the
38: possibility that backreaction and averaging effects may be used 
39: as an interesting alternative explanation to the dark
40: energy phenomenon.  
41: 
42: We have calculated $q^{vol}$ in some Lema\^itre--Tolman models. For those models
43: which are chosen to be realistic and which fit the supernova data, we find
44: that $q^{vol} > 0$, while those models which we have been able to find which
45: exhibit $q^{vol} < 0$ turn out to be unrealistic. This indicates that 
46: care must be exercised in relating the deceleration parameter 
47: to observations. 
48: \end{abstract}
49: 
50: \noindent{\it Keywords}: dark energy theory, supernova type Ia, superclusters and voids
51: 
52: \pacs{98.80-k, 95.36.+x, 98.65.Dx}
53: 
54: \section{Introduction}
55: Accelerated expansion, modeled by a positive cosmological constant, 
56: is an essential element of the current standard cosmological model 
57: of the Universe. The accelerated expansion was originally motivated by 
58: supernova observations \cite{sn} and is supported by many other
59: types of cosmological observations. Observational data is, in modern cosmology, 
60: analyzed almost exclusively within the 
61: framework of homogeneous and isotropic Friedmann models \cite{grs}. This analysis
62: leads to the Concordance model, which 
63: provides a remarkably precise fit to cosmological observations.
64: In this situation, if the 
65: Ehlers-Geren-Sachs theorem \cite{EGS} and `almost EGS
66: theorem' \cite{SME} are invoked\footnote{These theorems imply that if anisotropies in the cosmic microwave background radiation are small for all fundamental observers then the Universe is locally almost spatially homogeneous and isotropic. However, as shown in \cite{NUWL99} the almost Robertson--Walker geometry also requires the smallness of the Weyl curvature.}, 
67: then it seems that an assumption of large scale homogeneity of
68: the Universe can be justified. This on the other hand implies that the Universe must be filled
69: with dark energy which currently drives the acceleration of the Universe.
70: 
71: However the Concordance model is not the only one which can fit cosmological 
72: observations. Anti-Copernican inhomogeneous models which assume
73: the existence of a local Gpc scale void also fit cosmological observations
74: \cite{ide} (see \cite{C07} for a review).
75: Moreover, on small and medium scales our Universe is not
76: homogeneous. Therefore, one may ask 
77: whether Friedmann models can describe our
78: Universe correctly. In particular, it is important to ask 
79: what is the best way to fit a homogeneous model to a realistic
80: and inhomogeneous Universe. This problem, known as the fitting problem,
81: was considered by Ellis and Stoeger \cite{ES87}.
82: In considering the fitting problem, it becomes apparent that 
83: a homogeneous model fitted to inhomogeneous data can evolve quite differently 
84: from  the real Universe.
85: The difference between evolution of homogeneous
86: models and an inhomogeneous Universe is caused by backreaction effects, due
87: to the nonlinearity of the Einstein equation. 
88: Unfortunately, in the standard approach, the backreaction is rarely
89: taken into account -- in most cases when
90: modelling our Universe on a local scale Newtonian 
91: mechanics is employed and on large scales the Friedmann
92: equations (or linear perturbations of Friedmann background) are used \cite{P93}. 
93: Such an approach to cosmology is often encouraged by the ``no--go'' theorem
94: which states that the Universe can be very accurately described
95: by the conformal Newtonian metric perturbed about a spatially flat background,
96: even if $\delta \rho / \rho \gg 0$. In such a case the backreaction 
97: is negligible \cite{IW06,KAF06}.
98: However, the results obtained by van Elst and Ellis \cite{EE98} and
99: recently by Kolb, Marra and Matarrese \cite{KMM08}
100: show that the application of ``no-go'' theorem is limited.
101: Therefore, one should be aware that in the absence of an analysis of the
102: backreaction and other effects caused by inhomogeneities in the universe, 
103: there remains the possibility that the observed accelerated expansion of the Universe is
104: only apparent \cite{El08}. 
105: The direct study of the dynamical effects of inhomogeneities is difficult.
106: Due to the nonlinearity of the Einstein equations, the solution of the Einstein equations for the homogeneous matter distribution leads in principle to a different description  
107: of the Universe than an average of a inhomogeneous solution to the exact
108: Einstein equations (even though inhomogeneities when averaged  
109: over a sufficiently large scale might tend to be zero).
110: 
111: Neither the analysis of the evolution of a
112: general matter distribution nor the numerical evolution of cosmological models
113: employing the full Einstein equations are available at the level of
114: detail which would make them useful for this problem. 
115: There are currently several different approaches which attempt to take 
116: backreaction effects into account. 
117: One approach is based on exact solutions -- see for example
118: \cite{ha}. 
119: Another, and more popular approach is based on averaging.
120: 
121: In the averaging approach to backreaction, one considers a solution to 
122: the Einstein equations for a general matter distribution and 
123: then an average of various observable quantities is taken. 
124: If a simple volume average is considered then such an attempt leads
125: to the Buchert equations \cite{B00}. The Buchert equations are very
126: similar to the Friedmann equations except for the backreaction term which is
127: in general nonvanishing, if inhomogeneities 
128: are present.
129: For a review on backreaction and the
130: Buchert averaging scheme the reader is referred to
131: \cite{SR2006,B08}.  
132: Within this framework and using spherically symmetric
133: inhomogeneous models  
134: Nambu and Tanimoto \cite{NT}, Paranjape and Singh \cite{PS}, Kai, Kozaki, Nakao, Nambu, and Yoo
135: \cite{KKNNY}, Chuang, Gu, and Hwang \cite{CGH},
136: provided explicit examples that one can obtain negative values of the volume deceleration parameter
137: even if $\Lambda = 0$.
138: Another interesting example was presented by R\"as\"anen
139: \cite{SR2006,R06b} where 
140: it was shown that the total volume deceleration parameter 
141: of two isolated and locally decelerating regions
142: can also be negative.
143: 
144: 
145: There are however important ambiguities in the application of an averaging
146: procedure. The average itself not
147: only depends on a choice of volume but also on a choice of time slicing. 
148: This is very crucial in
149: cosmology. Once inhomogeneities are present the age of the Universe is not
150: everywhere the same. Namely, the big bang in inhomogeneous models is not a
151: single event, so the average taken over a hypersurface of constant cosmic
152: time $t$ is different from the average taken over a hypersurface of constant
153: age of the Universe $t - t_B$ \cite{SR2004b}.
154: %
155: Moreover, the results of the averaging procedure vary 
156: if the discrepancy between the average
157: cosmic time and the local time is introduced (the local time is the time
158: which is measured by local clocks; the cosmic time is the time which appears
159: in the averaged homogeneous model). This phenomenon was studied by Wiltshire
160: \cite{Wa}, and has been used in an ambitious alternative concordance
161: model. The model proposed by Wiltshire introduces some additional assumptions which allow 
162: to some extent a comparison of averaged quantities with observations.
163: Such a comparison shows quite good agreement with observations, \cite{LNW08}. Thus, while serious
164: fundamental questions remain concerning Wiltshire's approach, it is another
165: example of an approach where one does not need  dark energy
166: to fit cosmological observations.
167: 
168: The averaging procedure is also gauge-dependent.
169: For example using different gauge one can obtain that the backreaction mimics
170: not dark energy but dark matter \cite{KKM}.
171: The averaging schemes, therefore, in the literature have been criticized, and their
172: inherent ambiguities (and in some cases obscurity) have been discussed,
173: cf. e.g. \cite{IW06}.  A key point is that it is far from obvious if the
174: average quantities, such as the acceleration of the averaged universe are 
175: really the quantities which are measured in astronomical observations. 
176: In particular, an operational analysis is to a large extent lacking in the
177: discussions of averaging. Thus, it is important to test the
178: averaging procedures with the exact and inhomogeneous solutions of the
179: Einstein equations. Within exact models each quantity can easily be
180: calculated and then compared with its averaged counterpart.  This paper aims
181: to perform such an analysis within the Lema\^itre--Tolman model.
182: 
183: 
184: The structure of this paper is as follows. Buchert's averaging procedure
185: is presented in section \ref{BSsec}, and some background on the  
186: Lema\^itre--Tolman model is given in section \ref{LTsec}. The
187: volume and distance deceleration parameters are introduced
188: in section \ref{DPsec}. 
189: Finally, in section  \ref{QAsec}, we discuss the relation between the 
190: deceleration parameters, supernova observations and models of cosmic
191: structures.
192: 
193: 
194: \section{The Buchert scheme}\label{BSsec}
195: 
196: 
197: If the averaging procedure is applied to the Einstein equations, then for
198: irrotational and pressureless matter the following equations are obtained
199: \cite{B00}
200: 
201: \begin{eqnarray}
202: && 3 \frac{\ddot{a}}{a} = - 4 \pi G \av{\rho} + \mathcal{Q},
203: \label{bucherteq1} \\ &&  3 \frac{\dot{a}^2}{a^2} = 8 \pi G \av{ \rho}	-
204: \frac{1}{2} \av{ \mathcal{R} } - \frac{1}{2} \mathcal{Q}, \label{bucherteq2} \\
205: &&  \mathcal{Q} \equiv \frac{2}{3}\left( \av{{\Theta^2}} - \av{ \Theta }^2 \right)
206: - 2 \av{ \sigma^2},
207: \label{qdef}
208: \end{eqnarray}
209: where $\av{ \mathcal{R} }$ is an average of the spacial Ricci scalar $^{(3)}
210: \mathcal{R}$, $\Theta$ is the scalar of expansion, $\sigma$ is the shear
211: scalar, and $\av{\ }$ is the volume average over the hypersurface of constant time:
212: $\av{A} = ( \int d^3x \sqrt{-h} )^{-1} \int d^3x \sqrt{-h} A $. The scale factor $a$ is
213: defined as follows:
214: 
215: \begin{equation}
216: a = (V/V_0)^{1/3},
217: \label{aave}
218: \end{equation}
219: where V$_0$ is an initial volume.
220: 
221: 
222: Equations (\ref{bucherteq1}) and (\ref{bucherteq2}) are very similar to the
223: Friedmann equations, where Q=0, and $\rho$ and $\mathcal{R}$ depend on time
224: only.  In fact, they are 
225: kinematically
226: equivalent with a Friedmann model that has an additional scalar field    
227: source \cite{BLA06}. 
228: However the Buchert equation do not form a closed system.
229: To close these equation one has to introduce some further assumptions \cite{B00}.
230: As can be seen from  (\ref{qdef}) if the dispersion of expansion is
231: large, Q can be large as well and one can get acceleration ($\ddot{a}>0$)
232: without employing the cosmological constant.
233: 
234: 
235: \section{The Lema\^itre--Tolman model}\label{LTsec}
236: 
237: 
238: The Lema\^itre--Tolman model\footnote{The pressure free and irrotational solution 
239: of the Einstein equations for spherically symmetric space-time is often
240: called the Tolman, Tolman--Bondi, or Lema\^itre--Tolman--Bondi model.
241: However, it is more justified to refer to this solution
242: as to the Lema\^itre--Tolman model (cf. \cite{K97}).}
243:  \cite{LT} is a spherical symmetric,
244: pressure free and irrotational solution of the Einstein equations.  Its
245: metric is of the following form
246: 
247: \begin{equation}
248: {\rm d}s^2 =  c^2{\rm d}t^2 - \frac{R'^2(r,t)}{1 + 2 E(r)}\ {\rm d}r^2 -
249: R^2(t,r) {\rm d} \Omega^2, \label{ds2}
250: \end{equation}
251: where $ {\rm d} \Omega^2 = {\rm d}\theta^2 + \sin^2 \theta {\rm
252: d}\phi^2$. Because of the signature $(+, -, -, -)$, the $E(r)$ function must
253: obey $E(r) \ge - 1/2.$ Prime $'$ denotes $\partial_r$.
254: 
255: The Einstein equations reduce, in $\Lambda=0$ case, to the following two
256: \begin{equation}\label{den}
257: \kappa \rho(r,t) c^2 = \frac{2M'(r)}{R^2(r,t) R'(r,t)},
258: \end{equation}
259: \begin{equation}\label{vel}
260: \frac{1}{c^2}\dot{R}^2(r,t) = 2E(r) + \frac{2M(r)}{R(r,t)},
261: \end{equation}
262: \noindent where $M(r)$ is another arbitrary function and $\kappa = 8 \pi
263: G/c^4$. Dot $\dot{}$ denotes $\partial_t$.
264: 
265: When $R' = 0$ and $M' \ne 0$, the density becomes infinite. This happens at
266: shell crossings. This is an additional singularity to the Big Bang that
267: occurs at $R = 0, M' \neq 0$.  By setting the initial conditions
268: appropriately the shell crossing singularity can be avoided (see
269: \cite{HL1985} for detail discussion).
270: 
271: Equation (\ref{vel}) can be solved by simple integration:
272: 
273: \begin{equation}\label{evo}
274: \int\limits_0^R\frac{d\tilde{R}}{\sqrt{2E + \frac{2M}{\tilde{R}}}} = c \left[t- t_B(r)\right],
275: \end{equation}
276: where $t_B$ appears as an integration constant and is an arbitrary function
277: of $r$. This means that the big bang is not a single event as in the
278: Friedmann models, but occurs at different times at different distances from
279: the origin.
280: 
281: The scalar of the expansion is equal to
282: 
283: \begin{equation}
284: \Theta =  \frac{\dot{R}'}{R'} + 2 \frac{\dot{R}}{R}.
285: \label{theta}
286: \end{equation}
287: 
288: The shear tensor is of the following form:
289: 
290: \begin{equation}
291:  \sigma^\alpha{}_\beta = \frac{1}{3} \left( \frac{\dot{R}'}{R'} -
292:   \frac{\dot{R}}{R} \right) {\rm diag} (0, 2, -1, -1),
293: \end{equation}
294: thus $\sigma^2 \equiv (1/2) \sigma_{\alpha \beta} \sigma^{\alpha \beta}
295: =(1/3) (\dot{R}'/R' - \dot{R}/R)^2$.
296: 
297: The spacial Ricci scalar in the Lema\^itre--Tolman is equal to
298: 
299: \begin{equation}
300: ^{(3)} \mathcal{R} = - \frac{4}{R^2} \left( E + \frac{E' R}{R'} \right)
301: \end{equation}
302: 
303: 
304: \section{The apparent and average acceleration}\label{DPsec}
305: 
306: The deceleration parameter within the Friedmann models is defined as
307: 
308: \begin{equation}
309: q =  - \frac{\ddot{a}a}{\dot{a}^2},
310: \label{decparflrw}
311: \end{equation}
312: where $a$ is the scale factor.
313: By analogy we can define the deceleration parameter which is based on the
314: averaging scheme. Substituting (\ref{aave}) into (\ref{decparflrw}) and using
315: (\ref{bucherteq1}) and (\ref{bucherteq2}) we get
316: 
317: \begin{equation}
318: q^{vol} = - \frac{- 4 \pi G \av{ \rho } + \mathcal{Q}}{8 \pi G \av{ \rho }  -
319: \frac{1}{2} \av{ \mathcal{R} } - \frac{1}{2} \mathcal{Q}}.
320: \label{decparave}
321: \end{equation}
322: We refer to this deceleration parameter as the {\it volume deceleration parameter}, $q^{vol}$
323: since it is positive when the second derivative of volume is negative
324: and negative when the second derivative of volume is positive (and of sufficiently large value).
325: 
326: On the other hand one can introduce a deceleration parameter
327: defined relative to the distance. Within homogeneous models the
328: distance  to a given redshift is larger for accelerating models than for
329: decelerating ones.
330: Taylor
331: expanding the luminosity distance in the Friedmann model we obtain 
332: \begin{eqnarray}
333: && D_L =	\left. \frac{ d D_L}{d z} \right|_{z=0} z + \frac{1}{2}
334: \left. \frac{ d^2 D_L}{d z^2} \right|_{z=0} z^2 + \mathcal{O}(z^3) \nonumber
335: \\ && = \frac{c}{H_0} z + \frac{c}{2 H_0} (1 - q) z^2 + \mathcal{O}(z^3).
336: \label{dlfl}
337: \end{eqnarray}
338: Employing a similar procedure in the case of 
339: the Lema\^itre--Tolman model we get 
340: \begin{equation}
341: D_L =  \frac{cR'}{\dot{R}'} z  + \frac{c}{2} \frac{R'}{\dot{R}'} \left( 1 +
342: \frac{R' \ddot{R}'}{\dot{R}'^2} + \frac{cR''}{R' \dot{R}'} -
343: \frac{c \dot{R}''}{\dot{R}'^2} \right) z^2 + \mathcal{O}(z^3),
344: \label{dllt}
345: \end{equation}
346: Thus by comparing (\ref{dllt}) with  (\ref{dlfl}), the Hubble and the
347: deceleration parameter in the Lema\^itre--Tolman model can be defined as
348: \begin{equation}
349: H^{dis}_0 = \frac{\dot{R}'}{R'}, \quad q^{dis}_0 =  - \frac{R' \ddot{R}'}{\dot{R}'^2} - \frac{c R''}{R' \dot{R}'} +
350: \frac{c \dot{R}''}{\dot{R}'^2}.
351: \label{qddf}
352: \end{equation}
353: The above quantities are defined at the origin ($r=0$). However, following the Partovi and Mashhoon \cite{PM84} we can extend the above quantities to any $r$. Then, the coefficients of Taylor expansion are
354: 
355: 
356: \begin{eqnarray}
357:  \frac{ d D_L}{d z}  = && 2 R + \dot{R} \frac{dt}{dz} + R' \frac{dr}{dz} \nonumber \\
358: \frac{ d^2 D_L}{d z^2} = && 2 R + 4 \dot{R} \frac{dt}{dz} + 4 R' \frac{dr}{dz} 
359: + \ddot{R} \left(\frac{dt}{dz}\right)^2 + 2 \dot{R}' \frac{dt}{dz} \frac{dr}{dz} 
360: + R'' \left(\frac{dr}{dz}\right)^2 \nonumber \\
361: && + \dot{R} \frac{d^2t}{dz^2} + R' \frac{d^2r}{dz^2},
362: \end{eqnarray}
363: and we obtain
364: 
365: \begin{equation}
366: H^{dis} =  \left( \frac{ d D_L}{d z} \right)^{-1}, \quad q^{dis} =  1 -H^{dis} \frac{ d^2 D_L}{d z^2}.
367: \label{qdis}
368: \end{equation}
369: We refer to this deceleration parameter as the {\it distance deceleration parameter}.
370: Although of physical importance is the luminosity distance
371: and its ability of fitting the supernova data, the $q^{dis}$ is
372: of great usefulness. It allows us, 
373: without solving the geodesic equations, to easily check whether  
374: a considered model can be use to fit supernova data.
375: As we will see in the next section, models which fit supernova data have 
376: at least in some regions $q^{dis}<0$.
377: 
378: 
379: \section{Connection between deceleration parameter and observations}\label{QAsec}
380: 
381: Let us first focus on supernova observations. There is already a considerable
382: literature on 
383: inhomogeneous models which are able to fit the supernova observations without the
384: cosmological constant
385: \cite{ide}.
386: We shall examine four such models in this section. 
387: For each of these models we shall calculate 
388: the volume and distance deceleration parameters
389: and compare with each other.
390: The four models to be considered present a very good fit to
391: supernova data. The supernova data consists of 182 supernovae from 
392: the Riess gold sample \cite{R07}. The $\chi^2$ test for models 1-4 is
393: respectively 183.6, 184.3, 164.7, and 178.5 (for comparison
394: the $\chi^2$ of fitting the $\Lambda$CDM model is 165.3).
395: The residual Hubble diagram for
396: these models is presented in figure \ref{f1}.
397: The deceleration parameters for models 1-4 are presented in figure
398: \ref{f2}. Left panel presents the distance deceleration parameter [as defined
399: by (\ref{qdis}) - where $dt/dz$ and $dr/dz$ were calculated for the radial geodesic]. 
400: The distance deceleration parameter is positive at the origin, but soon becomes negative.
401: Moreover, a very similar shape is obtained if instead 
402: $q^{dis}$ [as defined by (\ref{qdis})] $q^{dis}_0$ [as defined by (\ref{qddf})] is used.
403: Thus, $q^{dis}$ (or even $q^{dis}_0$, if treated as a function of $r$) can be regarded as a useful test to check if
404: a given model is able to fit supernova data.
405: However, the most significant is that the volume deceleration parameter
406: which is presented in the right panel of figure \ref{f2} is strictly positive.
407: Thus, the ability of reproducing the supernova data does not
408: require that the volume deceleration parameter is negative. This raises the question whether the average acceleration has 
409: any relation with the observed acceleration of the Universe; and if yes, are models with
410: average acceleration also able to fit supernova data?
411: 
412: 
413: \begin{figure}
414: \begin{center}
415: \includegraphics[scale=0.7]{fg1.eps}
416: \caption{The Residual Hubble diagram for models 1-4.  The black dashed line 
417: presents $\Delta {\rm m}$ for the $\Lambda$CDM model.}
418: \label{f1}
419: \end{center}
420: \end{figure}
421: 
422: \begin{figure}
423: \includegraphics[scale=0.65]{fg2a.eps} 
424: \includegraphics[scale=0.65]{fg2b.eps}
425: \caption{The distance deceleration parameter (left panel) and the volume
426: deceleration parameter (right panel) for models 1-4.}
427: \label{f2}
428: \end{figure}
429: 
430: 
431: Let us now focus on models of cosmic structures.  It was recently shown that
432: using a perturbative approach, backreaction cannot explain the apparent
433: acceleration \cite{pab}. 
434: However, because of large density
435: fluctuations within cosmic structures, results obtained in terms of the
436: perturbation framework might be questionable.  Moreover, 
437: in view of the fact 
438: that there are known examples of exact inhomogeneous models with negative
439: volume deceleration parameter and $\Lambda = 0$, it is worthwhile to check
440: if realistically evolving models of cosmic structures can have negative
441: values of deceleration parameter.
442: First, let us consider a model of galaxy clusters with the Navarro-Frenk-White
443: density distribution \cite{NFW} (left panel of figure \ref{f3}). 
444: Although, the NFW profile describes virialized systems\footnote{The Lema\^itre--Tolman model which evolve from smooth density profile at last scattering
445: to a high value profile like the NFW profile is always characterized 
446: by a collapse -- central region within this models are at the current instant collapsing. Thus such systems cannot be considered as virialized systems.}
447: the use of this profile will prove to be very instructive.
448: The average deceleration parameter $q^{vol}$ for model 5 is presented in the right panel of
449: figure \ref{f3}.  As can be seen in this case the deceleration parameter is
450: positive (curve 5a). However, it is possible to modify this model so 
451: that the 
452: $q^{vol}$ becomes negative -- curve 5b in the right panel of figure \ref{f3}.  
453: This was obtained by choosing the $E$ function which is of large positive value
454: (for details see Appendix). However,  after such a 
455: modification this model becomes unrealistic.  Specifically, 
456: the age of the Universe in this model becomes unrealistically small. The bang time
457: function $t_B$ in this model is of large amplitude, around $11.44 \times
458: 10^9$ y. This means that the actual age of the Universe in this model is
459: approximately a few hundreds of thousand years.
460: 
461: 
462: \begin{figure}
463: \includegraphics[scale=0.65]{fg3a.eps} 
464: \includegraphics[scale=0.65]{fg3b.eps}
465: \caption{The current density distribution (left panel) and deceleration
466: parameter (right panel) for model 5.}
467: \label{f3}
468: \end{figure}
469: 
470: 
471: 
472: Now let us examine the volume deceleration parameter within models of cosmic
473: voids and superclusters.  Figure \ref{f4} presents density distribution of
474: realistically evolving cosmic structures (void -- curve 6, supercluster -- curve 7). 
475: It can be seen from the right panel of 
476: figure \ref{f4} that the 
477: volume deceleration parameter within these models is positive.
478: As above, we can modify our models in such a way that the volume
479: deceleration parameter is negative, but again this leads to a very large amplitude of
480: $t_B$. For example, in model 8 whose density and the volume deceleration parameter are presented in figure \ref{f5}\footnote{
481: Employing a model of qualitatively similar features as model 8, Hossain \cite{H07}
482: showed that the observer situated at the origin 
483: in order to successfully employ the Friedmann model has to assume the existence
484: of dark energy.} the volume deceleration parameter is negative. 
485: However, the bang time function in model 8 is of amplitude $\approx 11 \times 10^9$ y,
486: which leads to unrealistically small age of the Universe.
487: 
488: 
489: \begin{figure}
490: \includegraphics[scale=0.65]{fg4a.eps}
491:  \includegraphics[scale=0.65]{fg4b.eps}
492: \caption{The current density distribution (left panel) and volume deceleration
493: parameter (right panel) for models of cosmic structures (models 6, 7).}
494: \label{f4}
495: \end{figure}
496: 
497: \begin{figure}
498: \includegraphics[scale=0.65]{fg5a.eps}  
499: \includegraphics[scale=0.65]{fg5b.eps}  
500: \caption{The current density distribution (left panel) and
501: deceleration parameter (right panel) for model 8.}
502: \label{f5}
503: \end{figure}  
504: 
505: 
506: 
507: \section{Conclusions}\label{concl}
508: In this paper we have studied 
509: the relation between the volume deceleration parameter
510: obtained within the Buchert averaging scheme and the deceleration parameter
511: derived from the observations of supernovae. 
512: This work was motivated by recent
513: results showing there there are models which despite $\Lambda=0$
514: and average expansion rate is accelerating, i.e.
515: $\ddot{a} > 0$ [where $a$ is defined by relation (\ref{aave})].
516: This opens the
517: possibility that backreaction and averaging effects may be used 
518: as an interesting alternative explanation to the dark
519: energy phenomenon.  
520: 
521: We have compared the quantities obtained within the exact and
522: inhomogeneous models with their average counterparts. 
523: We focused on the
524: supernova observations and models of cosmic structures. For this purpose the
525: Lema\^itre--Tolman model was employed. It was showed numerically that the averaging of
526: models which fit the supernova observations does not lead to volume
527: acceleration ($\ddot{a} < 0$ for these 
528: averaged models and hence $q^{vol} >0$).
529: It was also shown that
530: realistically evolving models of cosmic structures have also $q^{vol} >0$.
531: It was possible to modify these model in such a way that after the
532: averaging $q^{vol} <0$. This was obtained by choosing $E$ function of positive amplitude
533: - as was recently proved by Sussman \cite{S08} this is a necessary condition to obtain 
534: $q^{vol} <0$. However, in models with realistic density distribution,
535: in such a cases, $E \gg 1 \gg M/R \approx 10^{-7} - 10^{-6}$, hence as seen from (\ref{evo}) $t_B \approx t$
536: (to remind $c \times 10^{10}$y $\approx 3$ Gpc). Thus, within such models
537: the age of the Universe is unrealistically small.
538: 
539: Our analysis has been performed in the limited class of Lema\^itre--Tolman models, which due to their spherical symmetry are arguably too
540: simple to give a full understanding of averaging and backreaction problems. However, within this class, we conclude that the volume  deceleration parameter $q^{vol}$ 
541: is not a quantity which can be directly related to observations.
542: 
543: It is possible that the volume deceleration parameter $q^{vol}$ becomes
544: negative only after averaging over the scales which are larger than 100 Mpc. On such
545: large scales the structure of the Universe becomes too complicated to be
546: fully described by spherically symmetric models. However,
547: it is intriguing that models which fit the supernova observations
548:  and for which the distance deceleration parameter, $q^{dis}$,
549: is negative have still $q^{vol}>0$. This suggest
550: that the volume  deceleration $q^{vol}$ 
551: does not have a clear interpretation in terms of observable quantities. 
552: It does not, of course, mean that averaging and backreaction effects 
553: cannot potentially be employed to explain the phenomenon of dark energy. 
554: However, our work here indicates that such a 
555: potential solution of the dark energy problem should be based upon different
556: methods than those related to volume deceleration parameter.
557: Rather than showing that $q^{vol}<0$ the averaging approach should explain
558: observations -- reproduce correct values of distance
559: to supernovae, correct shape of the CMB power spectrum, etc.
560: An interesting,  quasi-Friedmannian approach, was recently suggested in \cite{LABKC08}.
561: In this approach backreaction is modeled in terms of 
562: the morphon field \cite{BLA06}.
563: In such a case a Universe is describe by a homogeneous model with the spatial
564: curvature being just a function of time. As shown in \cite{LABKC08}
565: such approach lead to an agreement with supernova and CMB data
566: without the need for dark energy, but requires $q^{vol}<0$.
567: 
568: 
569: \ack
570: 
571: We would like to thank Henk van Elst and Thomas Buchert for useful discussions and comments.
572: KB would like to thank Peter and Patricia Gruber 
573: and the International Astronomical Union for the PPGF Fellowship,
574: the support of the Polish Astroparticle Network (621/E-78/SN-0068/2007)
575: is also acknowledged.
576: KB is also grateful to LA and the Albert Einstein Institute, where part of this research was carried out, for their hospitality. 
577: LA acknowledges the support of the NSF with grants 
578: DMS-0407732 and DMS-0707306 to the University of Miami.
579: 
580: \appendix
581: 
582: 
583: \section{Model specification}
584: 
585: 
586: \setcounter{section}{1}
587: 
588: 
589: There are three arbitrary functions of the radial coordinate  
590: in the Lema\^itre--Tolman. However only two functions are 
591: independent and the third one is specified by the choice of the
592: radial coordinate. Models considered in this paper are defined as follows:
593: 
594: \begin{enumerate}
595: 
596: \item
597: Model 1 and 2
598: 
599: The radial coordinate is chosen as the present day value of the
600: areal distance $r:=R_0$. Models 1 and 2 are specified
601: by the present day density distribution and the time bang function.
602: The density distribution is parametrized by
603: 
604: \begin{equation}
605: \rho(t_0,r) = \rho_b \left[ 1 + \delta_{\rho} - \delta_{\rho} \exp \left( - \frac{r^2}{\sigma^2} \right) \right],
606: \label{rhofl}
607: \end{equation}
608: where $\rho_b = \Omega_m \times  (3H_0^2)/(8\pi G)$, $\Omega_m = 0.27$, $H_0 = 70$ km s$^{-1}$ Mpc$^{-1}$.
609: In model 1 $\rho_{\delta} = 1.9$, $\sigma = 0.9$ Mpc, and 
610: in model 2 $\rho_{\delta} = 1.5$, $\sigma = 0.5$ Mpc.
611: In these models the big bang is assumed to occur simultaneously at every point, i.e. $t_B = 0$.
612: The functions $M$ and $E$ are then calculated and using 
613: eqs. (\ref{den}) and (\ref{evo}) respectively.
614: 
615: The time instants as well as background density $\rho_b$ 
616: in all models (1--8) is chosen as density of a Friedmann model ($\Omega_m = 0.27$, $H_0 = 70$ km s$^{-1}$)
617: and time instants are calculated using the following formula \cite{P93}:
618: 
619: \begin{equation}
620: t(z) =  \frac{1}{H_0} \int\limits_{z}^{\infty} \frac{d \tilde{z}}{(1+\tilde{z}) \sqrt{ \Omega_{mat} (1+\tilde{z})^3 + \Omega_K (1+\tilde{z})^2} },
621: \label{tz}
622: \end{equation}
623: where $\Omega_K = 1 - \Omega_{m}$.
624: The last scattering instant ($t_{LS}$) is set to take place when $z = 1089$ and the current instant ($t_{0}$) when $z=0$ --- $t_{LS} = 4.98 \times 10^5$ y, and $t_0 = 11.4421 \times 10^9$ y.
625: 
626: 
627: \item
628: Model 3 and 4
629: 
630: As above the radial coordinate is chosen  as the present day value of the areal distance $r:=R_0$. These two models are defined by the current expansion rate, 
631: and assumption that $\rho(t_0,r) = \rho_b$. The expansion rate is parametrized using
632: 
633: \begin{equation}
634: H_T(t_0,r) = \frac{\dot{R}}{R} = H_0 \left[ 1 - \delta_{H} + \delta_{H} \exp \left( - \frac{r^2}{\sigma^2} \right) \right],
635: \label{expfl}
636: \end{equation}
637: where $H_0 \delta_{H} = 9.6$ km s$^{-1}$ Mpc$^{-1}$, $\sigma =0.6$ Mpc,
638: and $H_0 \delta_{H} = 12$ km s$^{-1}$ Mpc$^{-1}$, $\sigma =1.2$ Mpc
639: for model 3 and 4 respectively.
640: In these models density is assumed to be homogeneous at the current epoch.
641: The function $M$ is then calculated using the above relation and eq. (\ref{vel}).
642: It should be noted that the $H_T$ is one of several generalization 
643: of the Hubble constant, which in the Friedmann model
644: is $H_0 = \dot{a}/a$. Apart from the transverse Hubble parameter, $H_T$, one can also define the radial  Hubble parameter, $H_R$, [see eq. (\ref{qddf})],
645: and the volume Hubble parameter defined as $H_V = (1/3) \Theta = H_R + 2 H_T$.
646: 
647: 
648: \item
649: Model 5a
650: 
651: The radial coordinate is chosen as the present day value of the areal distance, i.e. $r:=R_0$.
652: The model is defined by density distributions given at the present instant
653: and at last scattering.
654: The density distribution at the current instant is parametrized by
655: 
656: \begin{equation}
657:  \rho(t_0,r) = \rho_{b} \frac{\delta}{(r/r_s)(1+r/r_s)^2} ,
658: \label{nfwp}
659: \end{equation}
660: where $\delta = 28 170$ and $r_s = 191 kpc$. This is a 
661: Navarro, Frenk, and White galaxy cluster profile \cite{NFW}.
662: As can be seen this profile is singular at the origin but
663: this problem can be overcome by matching the NFW profile
664: with a singular--free profile as $f(r) = - a r^2 + b$.
665: 
666: The density profile at last scattering is assumed to be homogeneous, thus the areal distance at last scattering is:
667: 
668: \begin{equation}
669: R (t_{LS},r) = \left( \frac{M}{\kappa \rho_{LS} c^2} \right)^{1/3}.
670: \end{equation}
671: The function $M(r)$ is then calculated from eq. (\ref{den}).
672: Function $E$ can be calculated by subtracting solutions
673: of (\ref{evo}) for $t_{LS}$ and $t_0$ (for details see \cite{KH02}).
674: The function $E$ is presented in the left panel of figure \ref{figA1}.
675: 
676: 
677: 
678: \item 
679: Model 5b 
680: 
681: The radial coordinate is chosen as the present day value of the
682: areal distance, $r:=R_0$.
683: The model is defined by density distribution given by (\ref{nfwp})
684: and $E$ of the following form
685: 
686: \begin{equation}
687: E (r) = 10^3 \sin \left(10^{-3} r {\rm Mpc}^{-1} \right).
688: \end{equation}
689: This profile is presented in the left panel of figure \ref{figA1}
690: and the bang time function $t_B$ in the right panel.
691: 
692: \item
693: Models 6 and 7
694: 
695: The radial coordinate is chosen as the value of the
696: areal distance at last scattering instant, $r:=R_{LS}$.
697: Model 6 and 7 are defined by the assumption that $t_B=0$
698: and the density distribution,
699: which at last scattering is of the following form
700: 
701: \begin{equation}
702: \rho(t_{LS},r) = \rho_b \left( 1 - \delta \exp{(-a \ell^2 r^2)}
703: + \gamma  \exp{\left[-\left(\frac{\ell r - c}{d}\right)^2 \right]} \right), 
704: \end{equation}
705: where  $\ell = 1/kpc$;
706: $\delta = 1.2 \times 10^{-3}$ and $2 \times 10^{-3}$
707: for model 6 and 7 respectively;
708: $\gamma = 14.62 \times 10^{-4}$ and $8.03 \times 10^{-4}$
709: for model 6 and 7 respectively;
710: $a = 0.01$ and $0.04$ for model 6 and 7 respectively;
711: $c = 18$ and $12$ for model 6 and 7 respectively;
712: and $d=6$ and $5$ for model 6 and 7 respectively.
713: The bang time function for both these models is $t_B = 0$.
714: The mass function, $M(r)$ is calculated
715: from eq. (\ref{den}), and the function $E(r)$ is calculated from eq. (\ref{evo}).
716: 
717: 
718: \item
719: Model 8
720: 
721: The radial coordinate is chosen  as a present day value of the
722: areal distance: $r:=R_0$. Density distribution is of the following
723: form
724: 
725: \begin{equation}
726: \rho(t_{0},r) = 6.2 \rho_b \exp{(-4 \times 10^{-8} (\ell r)^2)}, 
727: \end{equation}
728: and the function $E$ is
729: 
730: \begin{equation}
731: E(r) =  \left(\frac{H_0}{c} r \right)^2 \exp{(10^{-3} \ell r)}, 
732: \end{equation}
733: which except for $[\exp{(10^{-3} \ell r)}]$ is the same 
734: as $E(r)$ profile in the empty Universe.
735: This profile is presented in the left panel of figure \ref{figA2}
736: and the bang time function $t_B$ in the right panel.
737: 
738: 
739: 
740: 
741: \end{enumerate}
742: 
743: \begin{figure}
744: \includegraphics[scale=0.65]{fgA1a.eps}  
745: \includegraphics[scale=0.65]{fgA1b.eps}  
746: \caption{Left panel presents the function $E(r)$  for models 5a and 5b.
747: Please note that the y-scale in the upper part of the left panel is
748: different than in the lower part. Right panel bang time function 
749: for model 5b.}
750: \label{figA1}
751: \end{figure}  
752: 
753: 
754: 
755: 
756: 
757: 
758: \begin{figure}
759: \includegraphics[scale=0.65]{fgA2a.eps}  
760: \includegraphics[scale=0.65]{fgA2b.eps}  
761: \caption{The function $E(r)$ (left panel) and $t_B(r)$ (right panel) for model 8.}
762: \label{figA2}
763: \end{figure}  
764: 
765: 
766: \section*{References}
767: 
768: \begin{thebibliography}{99}
769: 
770: \bibitem{sn} 
771: Perlmutter S \etal 1997 {\it Astrophys. J}. {\bf 483} 565;
772: Riess A G \etal 1998 {\it Astrophys. J}. {\bf 116} 1009
773: Perlmutter S \etal 1999 {\it Astrophys. J}. {\bf 517} 565
774: 
775: \bibitem{grs} 
776: Hinshaw G \etal 2008 submitted to {\it Astrophys. J. Suppl. Ser.}
777: ({\it Preprint} arXiv:0803.0732);
778: Tegmark M \etal 2006 {\it Phys. Rev.} D {\bf 74} 123507;
779: Percival W J  \etal 2002 {\it Mon. Not. Roy. Astr. Soc.} {\bf 337} 1068 
780: 
781: \bibitem{EGS} 
782: Ehlers J, Geren P and Sachs R K 1968 {\it J. Math. Phys}.  {\bf 9} 1344
783: 
784: \bibitem{SME}
785: Stoeger W R, Maarteens R and Ellis G F R 1995 {\it Astrophys. J}. {\bf 443} 1
786: 
787: \bibitem{NUWL99}
788: Nilsson U S, Uggla C, Wainwright J and Lim W C 1999 {\it Astrophys. J}. {\bf 521} L1
789: 
790: 
791: \bibitem{ide} 
792: D\k{a}browski M P and Hendry M A 1998
793: {\it Astrophys. J}. {\bf 498}, 67;
794: C\'el\'erier M N 2000 {\it Astron. Astrophys.} {\bf 353} 63;
795: Iguchi H, Nakamura T and Nakao K 2002 {\it Prog. Theor. Phys}. {\bf 108} 809;
796: God\l{}owski W, Stelmach J and Szyd\l{}owski M 2004 {\it Class. Q. Grav}. {\bf 21} 3953;
797: Alnes H and Amarzguioui M 2006 {\it Phys. Rev}. D {\bf 74} 103520;
798: Alnes H. and Amarzguioui M and Gr\o n \O~ 2006 {\it Phys.Rev.} D {\bf 73} 083519;
799: Chung D J H and Romano A E 2006 {\it Phys Rev}. D {\bf 74} 103507;
800: Alnes H and Amarzguioui M 2007 {\it Phys. Rev}. D {\bf 75} 023506;
801: Enqvist K and Mattsson T 2007 {\it J. Cosmol. Astropart. Phys}. JCAP07(2007)02019;
802: Alexander S, Biswas T, Notari A and Vaid D 2007 {\it Preprint} arXiv:0712.0370;
803: Biswas T, Mansouri R and Notari A 2007 {\it J. Cosmol. Astropart. Phys}. JCAP12(2007)017;
804: Brouzakis N, Tetradis N and Tzavara E 2007 {\it J. Cosmol. Astropart. Phys}. JCAP0702(2007)013;
805: Brouzakis N, Tetradis N and Tzavara E 2008 {\it J. Cosmol. Astropart.
806: Phys}. JCAP0704(2008)008;
807: Marra V, Kolb E W, Matarrese S and Riotto A 2007 {\it Phys. Rev}. D {\bf 76} 123004;
808: Biswas T  and Notari A  2008 {\it J. Cosmol. Astropart. Phys}. JCAP06(2008)021;
809: Bolejko K 2008 {\it PMC Physics\/} PMCA02(2008)01 ({\it Preprint} astro-ph/0512103);
810: Garc\'ia-Bellido J and Haugb\o elle T 2008 {\it J. Cosmol. Astropart. Phys}. JCAP04(2008)03;
811: Enqvist K 2008 {\it Gen. Rel. Grav}. {\bf 40} 451;
812: Khosravi Sh, Kourkchi E, Mansouri R and Akrami Y 2008 {\it Gen. Rel. Grav.} {\bf 40} 1047;
813: Bolejko K and Wyithe J S B 2008  {\it Preprint} arXiv:0807.2891;
814: Yoo C-M, Kai T and Nakao K-i 2008 {\it Preprint} arXiv:0807.0932
815: 
816: \bibitem{C07}
817: C\'el\'erier M N 2007 {\it New Adv. Phys.} {\bf 1} 29
818: 
819: \bibitem{ES87} 
820: Ellis G F R and Stoeger W 1987 {\it Class. Quant. Grav}. {\bf 4} 1697
821: 
822: \bibitem{P93}
823: Peebles P J E 1993 {\it Principles of physical cosmology}
824: (Princeton: Princeton University Press)
825: 
826: \bibitem{IW06}
827: Ishibashi A and Wald R M 2006 {\it Class. Q. Grav.} {\bf 23} 235
828: 
829: \bibitem{KAF06}
830: Kasai M, Asada H and Futamase T 2006 {\it Prog. Theor. Phys}. {\bf 115} 827
831: 
832: \bibitem{EE98}
833: van Elst H and Ellis G F R 1998 {\it Class. Quant. Grav}. {\bf 15} 3545
834: 
835: \bibitem{KMM08} 
836: Kolb E W, Marra V and Matarrese S 2008 {\it Preprint} arXiv:0807.0401 
837: 
838: \bibitem{El08}
839: Ellis G F R 2008 {\it Nature} {\bf 452} 158
840: 
841: \bibitem{ha}
842: Hellaby C 1988 {\it Gen. Rel. Grav}. {\bf 20} 1203;
843: Lu T H C and Hellaby C 2007 {\it Class. Quant. Grav.} {\bf 24} 4107;
844: McClure M and Hellaby C 2007 {\it Preprint} arXiv:0709.0875
845: 
846: \bibitem{B00} 
847: Buchert T 2000 {\it Gen. Rel. Grav.} {\bf 32} 105
848: 
849: \bibitem{SR2006} 
850: R\"as\"anen S 2006 {\it J. Cosmol. Astropart. Phys.} JCAP11(2006)003
851: 
852: \bibitem{B08} 
853: Buchert T 2008 {\it Gen. Rel. Grav.} {\bf 40} 467
854: 
855: \bibitem{NT} 
856: Nambu Y and Tanimoto M 2005 {\it Preprint} gr-qc/0507057 
857: 
858: 
859: \bibitem{PS} 
860: Paranjape A and Singh T P 2006 {\it Class. Quant. Grav}. {\bf 23} 6955
861: 
862: \bibitem{KKNNY} 
863: Kai T, Kozaki H, Nakao K, Nambu Y and Yoo C M 2007 
864: {\it Prog. Theor. Phys}. {\bf 117} 229
865: 
866: \bibitem{CGH} 
867: Chuang C H, Gu J A and Hwang W Y P 2008 {\it Class. Quant. Grav}. {\bf 25} 175001
868: 
869: 
870: \bibitem{R06b} 
871: R\"as\"anen S 2006 {\it Int. J. Mod. Phys.} D {\bf 15} 2141
872: 
873: 
874: \bibitem{SR2004b}
875: R\"as\"anen S 2004 {\it J. Cosmol. Astropart. Phys.} JCAP11(2004)010
876: 
877: \bibitem{Wa}
878: Wiltshire  D L 2007 {\it New J. Phys}. {\bf 9} 377;
879: Wiltshire D L 2007 {\it Phys. Rev. Lett}. {\bf 99}  251101
880: 
881: \bibitem{LNW08} 
882: Leith B M, Ng S C C and Wiltshire D L 2008 {\it Astrophys. J}. {\bf 672} L91 
883: 
884: \bibitem{KKM}
885: Khosravi S, Kourkchi E and Mansouri R 2007 {\it Preprint} arXiv:0709.2558
886: 
887: \bibitem{BLA06}
888: Buchert T, Larena J and  Alimi J-M 2006 {\it Class. Q. Grav.} {\bf 23} 6379
889: 
890: \bibitem{K97} 
891: Krasinski A 1997 {\it Gen. Rel. Grav}. \textbf{29} 637;
892: Krasinski A 1997 {\it Gen. Rel. Grav}. \textbf{29} 931;
893: Pleba\'nski J and Krasi\'nski A 2006 {\it  Introduction to general relativity and cosmology} (Cambridge: Cambridge University Press)
894: 
895: \bibitem{LT} 
896: Lema\^{\i}tre G 1933 {\it Ann. Soc. Sci. Bruxelles\/} A \textbf{53} 51
897: (1933) (reprinted in 1997 {\it Gen. Rel. Grav}. \textbf{29} 641);
898: Tolman R C 1934 {\it Proc. Nat. Acad. Sci. USA}	 \textbf{20} 169
899: (reprinted in 1997 {\it Gen. Rel. Grav}. \textbf{29} 935)
900: 
901: \bibitem{HL1985} 
902: Hellaby C and Lake K 1985 {\it Astrophys. J}. {\bf 290} 381
903: (plus errata in 1986 {\it Astrophys. J}. {\bf 300} 461)
904: 
905: \bibitem{PM84} 
906: Partovi M H and Mashhoon B 1984 {\it Astrophys. J}. {\bf 276} 4;
907: 
908: \bibitem{R07}
909: Riess A G et al. 2007 {\it Astrophys. J}. \textbf{659} 98
910: 
911: \bibitem{pab}
912: Li N, Schwarz D J 2007 {\it Preprint} arXiv:0710.5073;
913: Paranjape A and Singh T P 2008 {\it Preprint} arXiv:0806.3497 
914: 
915: \bibitem{NFW} 
916: Navarro J F, Frenk C S and White S D M 1996 {\it Astrophys. J}. {\bf 462} 563 
917: 
918: \bibitem{H07}
919: Hossain G M 2007 {\it Preprint} arXiv:0709.3490 
920: 
921: \bibitem{S08}
922: Sussman R A 2008 {\it Preprint} arXiv:0807.1145
923: 
924: \bibitem{LABKC08} 
925: Larena J, Alimi J-M, Buchert T, Kunz M and Corasaniti P-S 2008 {\it Preprint} arXiv:0808.1161
926: 
927: 
928: \bibitem{KH02}
929: Krasi\'nski A and Hellaby C 2002 {\it  Phys. Rev}. D {\bf 65} 023501 
930: 
931: 
932:  \end{thebibliography}
933: 
934: \end{document}
935: 
936: 
937: