1: \documentclass[aps,preprint,nofootinbib,10ppt]{revtex4}
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: \usepackage{amsfonts}
4: \usepackage{amsmath}
5: \usepackage{amssymb}
6: \usepackage{graphicx,epsfig}
7: \usepackage{color}
8: \usepackage[colorlinks=false, pdfstartview=FitV, linkcolor=blue, citecolor=blue, urlcolor=blue]{hyperref}
9:
10:
11: \begin{document}
12:
13: \title{Realistic Equations of State for the Primeval Universe}
14: \author{R. Aldrovandi}
15: \email{ra@ift.unesp.br}
16: \affiliation{Instituto de F\'{\i}sica Te\'{o}rica, Universidade Estadual Paulista. Rua
17: Pamplona 145, CEP 01405-900, S\~{a}o Paulo, SP, Brazil}
18: \author{R. R. Cuzinatto}
19: \email{rcuzin@phys.ualberta.ca}
20: \affiliation{Theoretical Physics Institute, University of Alberta, Edmonton, Alberta,
21: Canada, T6G 2J1.}
22: \author{L. G. Medeiros}
23: \email{leogmedeiros@gmail.com}
24: \affiliation{Centro Brasileiro de Pesquisas F\'{\i}sicas. Rua Xavier Sigaud 150, CEP
25: 22290-180, Rio de Janeiro, RJ, Brazil}
26: \pacs{PACS number}
27:
28: \begin{abstract}
29: Early universe equations of state including realistic interactions between
30: constituents are built up. Under certain hypothesis, these equations are
31: able to generate an inflationary regime prior to the nucleosynthesis period.
32: The resulting accelerated expansion is intense enough to solve the flatness
33: and horizon problems. In the cases of curvature parameter $\kappa $ equal to
34: $0$ or $+1$, the model is able to avoid the initial singularity and offers a
35: natural explanation for why the universe is in expansion. All the results
36: are valid only for a matter-antimatter symmetric universe.
37: \end{abstract}
38:
39: \volumeyear{2008}
40: \volumenumber{number}
41: \issuenumber{number}
42: \eid{identifier}
43: \date[Date text]{date}
44: \received[Received text]{date}
45: \revised[Revised text]{date}
46: \accepted[Accepted text]{date}
47: \published[Published text]{date}
48: \startpage{1}
49: \endpage{}
50: \maketitle
51:
52: %\preprint{IFT-P/?}
53:
54: %\affiliation{Instituto de F\'{\i}sica
55: %Te\'{o}rica, Universidade Estadual Paulista. Rua Pamplona 145, CEP
56: %01405-900, S\~{a}o Paulo, SP, Brazil}
57:
58: \section{Introduction\label{sec-Intro}}
59:
60: The gravitational field, as described by General Relativity, couples to all
61: types of energy: rest masses, kinetic terms and interaction terms.
62: Relativistic cosmology is, in consequence, deeply concerned with such
63: sources. The kinetic terms and the rest-masses are currently taken into
64: account by assuming ideal cosmic fluids constituted by ultrarelativistic
65: and/or non-relativistic matter. Solutions of this type are found in standard
66: texts \cite{nar,wein} and -- when multi-component fluids are considered --
67: in several papers, e.g., \cite{Solutions}. Interaction terms are commonly
68: used only in perturbative models. In fact, using the Boltzmann equation in a
69: Friedmann-Robertson-Walker (FRW) background, the inhomogeneities both in the
70: cosmic microwave background (CMB) and in the matter content \cite{Dod,kolb}
71: are studied, for comparison with the observational data \cite%
72: {Wmap,LargeScale}. Notwithstanding, interactions are not considered as
73: direct sources of gravitation in this line of research \cite{artigo 1}.
74:
75: Of course, some well-known proposals do consider interaction processes as
76: direct sources of gravitation. They are usually related to the accelerated
77: expansion regimes: present-day dynamics \cite{SuNo} or inflation \cite%
78: {Liddle}. Nevertheless, these theories do not actually consider the
79: fundamental interactions (electromagnetic, weak and strong) between
80: particles in the source constituents. Indeed, in the standard inflationary
81: approaches \cite{Guth,AlbreSten,Linde} an accelerated expansion is obtained
82: through self-interaction processes of scalar \textit{inflaton} fields $\phi
83: _{i}\left( x\right) $. This self-interaction is chosen so as to produce just
84: those features which are necessary to describe the early accelerated regime
85: \cite{Liddle}. The phenomenological explanations for the present-day
86: acceleration include, among others, \emph{(i)} the quintessence models \cite%
87: {CalDaveStein,Stein,Zlatev,FraRos,PeRa}, which roughly follow the same lines
88: of the inflationary theory; \emph{(ii)} models of matter and dark-energy
89: unification via Chapligyn-like equations of state (EOS) \cite{Orfeu} or
90: through equations of the Van der Walls type \cite{Capozzielo,Kremer}; \emph{%
91: (iii)} models of mass-varying-neutrino type, which couple neutrinos to a
92: quintessence scalar field \cite{Motta}; \emph{(iv)} models introducing
93: interactions in the energy conservation equation \cite{Gabi,NelsonPinto}.
94:
95: A formal procedure has been recently proposed for including the
96: (fundamental) interactions as direct sources of gravitation in the
97: cosmological context \cite{artigo 1}. Imported from equilibrium statistical
98: mechanics, this formalism allows the construction of realistic equations of
99: state for both relativistic \cite{Dashen,Reichl} and non-relativistic
100: systems \cite{Pat,Beth}. Our objective here is to find primeval cosmic fluid
101: EOS taking into account \textit{physically realistic} interaction processes
102: between the constituent particles, in addition to their kinetic and
103: rest-mass terms. In particular, we shall examine their effect on the scale
104: factor evolution. It will be shown that under certain hypothesis and
105: approximations an early accelerated regime can be obtained as a consequence
106: of these interacting processes.
107:
108: The idea of building realistic equations of state considering
109: interaction between elementary particles in the primeval universe is
110: not new. Actually, during the late 1970's and the beginning os the
111: 80's a series of papers by Bugrii, Trushevsky and Beletsky
112: \cite{ucraone,ucratwo,ucrathree,ucrafour} discussed the construction
113: of the high-energy EOS and their application to
114: the pre-nucleosynthesis universe.\footnote{%
115: The authors are thankful to an unknown referee for calling their attention
116: to these works.} Nevertheless, their treatment is different from the one
117: developed here as we will make clear some pages ahead.
118:
119: The paper is organized as follows. Section \ref{sec-Pns} reviews the general
120: features of the early universe in its standard presentation \cite%
121: {nar,wein,Dod,kolb,Liddle}. Section \ref{sec-SisInt} presents some results
122: from equilibrium statistical theory of interacting systems. Specifically,
123: the coefficients appearing in the perturbative fugacity expansions are
124: expressed in terms of the scattering matrix operator $\hat{S}$. In addition,
125: the matrix $S_{2}$ describing the two-particle scattering is associated to
126: the experimentally observable phase-shifts. The goal of Section \ref%
127: {sec-EoS_Pns} is to construct realistic EOS\ for the pre-nucleosynthesis
128: universe and discuss the hypothesis and approximations undertaken. In
129: Section \ref{sec-Conseq},\ the more direct cosmological consequences coming
130: from these equations are examined, including the effect of driving an
131: accelerated expansion (inflationary era). Section \ref{sec-Final} contains
132: some final comments. Details of a too technical nature, as well as some
133: data, have been relegated to appendices.
134:
135: %%%%%%%%%%%%%%%%%%%%%%%%
136:
137: \section{General features of the pre-nucleosynthesis universe \label{sec-Pns}%
138: }
139:
140: %%%%%%%%%%%%%%%%%%%%%%%%
141:
142: The period of the Universe evolution going under the name ``Early Universe''
143: covers many different and physically significant events. Indeed, it is usual
144: to consider both the matter-radiation decoupling ($kT_{\gamma }\sim 0.1 $ $%
145: eV $) and the electroweak transition ($kT_{\gamma }\sim 100$ $GeV$)\ as
146: belonging to that period. This work is concerned with a specific interval
147: within it, namely the pre-nucleosynthesis period (PNS). We shall define PNS
148: as the period immediately before the nucleosynthesis of the light elements,
149: when the nucleons (protons and neutrons) are in thermodynamical equilibrium
150: with the rest of the cosmic fluid, i.e., $kT_{\gamma }\gtrsim 20$ $MeV$. So,
151: in principle, any energy value larger than $20$ $MeV$ belongs to PNS;
152: nevertheless, it will be enough to restrict our working frame to energies of
153: a few hundreds of $MeV$. In consequence, PNS will in what follows actually
154: mean the interval $20$ $MeV\leq kT\lesssim 300$ $MeV$.
155:
156: Present-day universe is dominated by a dark energy component $\Lambda $\
157: plus a non-relativistic contribution formed by ordinary baryons $b$\ and
158: dark-matter $dm$; the ultrarelativistic components -- photons $\gamma $\ and
159: neutrinos $\nu $\ -- are quite negligible. As one goes back in time,
160: however, this situation changes drastically. Indeed, the evolution equations
161: for ultrarelativistic and the nonrelativistic matter \cite{nar,wein} in
162: terms of the expansion parameter,
163: \begin{equation}
164: \rho _{\gamma ,\nu }=\rho _{\gamma 0,\nu 0}\left( \frac{a_{0}}{a}\right) ^{4}%
165: \text{ \ \ \ \ and \ \ \ }\rho _{b,DM}=\rho _{b0,DM0}\left( \frac{a_{0}}{a}%
166: \right) ^{3}, \label{C1_10}
167: \end{equation}%
168: together with the Friedmann equations,%
169: \begin{equation}
170: \left( \frac{\dot{a}}{a}\right) ^{2}=\frac{8\pi G}{3}\rho -\frac{\kappa }{%
171: a^{2}}+\frac{\Lambda }{3}\text{ \ \ \ and \ \ \ }\left( \frac{\ddot{a}}{a}%
172: \right) =-\frac{4\pi G}{3}\left( \rho +3p\right) +\frac{\Lambda }{3},
173: \label{C1_11}
174: \end{equation}%
175: show that the ultrarelativistic contribution overcomes that of the other
176: components at energies corresponding to $kT_{\gamma }\sim 1$ $eV$.\footnote{%
177: We will see in Section \ref{sec-EoS_Pns}\ that this statement must be
178: qualified when the interaction processes are taken into account.} The
179: cosmological period dominated by ideal ultrarelativistic particles is
180: usually called the \textit{radiation era}.
181:
182: To find out the energetically relevant constituents during the PNS period we
183: must notice that, as we turn to the past and the energy $kT$\ increases,
184: pair-production processes become more and more frequent, generating a large
185: variety of particles. The main scenario is that of ultrarelativistic
186: particles generating non-relativistic ones. If we restrict ourselves to a
187: few hundreds of $MeV$, the relevant particles are:
188:
189: \begin{itemize}
190: \item \textit{fundamental bosons:} photons $\left( \gamma \right) $;
191:
192: \item \textit{leptons:\ }electrons $\left( e^{-}\right) $, positrons $\left(
193: e^{+}\right) $, muons $\left( \mu ^{-}\right) $, antimuons $\left( \mu
194: ^{+}\right) $, electronic and muonic neutrinos $\left( \nu _{e},\nu _{\mu
195: }\right) $ and electronic and muonic antineutrinos $\left( \bar{\nu}_{e},
196: \bar{\nu}_{\mu }\right) $;
197:
198: \item \textit{hadrons:\ }pions $\left( \pi ^{+},\pi ^{-},\pi ^{0}\right) $,
199: kaons $(K^{+},K^{-},K^{0},\bar{K}^{0})$, nucleons $\left( p,n\right) $ and
200: antinucleons $\left( \bar{p},\bar{n}\right) $.
201: \end{itemize}
202:
203: Concerning this list, it is important to emphasize that: \textit{(i)} the
204: particles taken into account are those with rest-mass bellow $1$ $GeV$; and,
205: \textit{(ii)} the hadrons considered are only those which are stable by the
206: strong interaction. Furthermore, we assume matter-anti-matter symmetry.
207:
208: Among the particles cited above, those that are ultrarelativistic at $20$ $%
209: MeV$\ are $\gamma $, $e^{\pm }$, $\nu _{e}$, $\nu _{\mu }$, $\bar{\nu}_{e}$
210: and $\bar{\nu}_{\mu }$; the remaining ones are non-relativistic. By
211: arguments given in classical texts \cite{nar,wein,kolb}, the chemical
212: potentials involved in the PNS reactions are zero, i.e., $\mu _{i}\simeq 0$
213: for every species ($i=\gamma $, $e^{\pm }$, $\mu ^{\pm }$, $\pi ^{\pm }$, $%
214: K^{\pm }$, ...).
215:
216: Thermal equilibrium is warranted as long as every reaction rate $\Gamma
217: _{i}(kT)$ of each given component $i$ with all the other particles is much
218: larger than the universe expansion rate. This last, in turn, is measured by
219: the Hubble function $H(kT)$. Therefore, when the condition%
220: \begin{equation}
221: \Gamma (T)\gg H(T) \label{C1_22}
222: \end{equation}%
223: is satisfied, thermodynamical \textit{quasi-static expansion} holds. In
224: fact, this may be verified for each variety by remembering that $\Gamma
225: =n\sigma \left\vert \vec{v}\right\vert $, where $n$ is the target-particle
226: density and $\sigma \left\vert \vec{v}\right\vert $\ is the interaction
227: cross-section times the relative velocity of the particles. The presence of
228: thermal and chemical equilibria in PNS is extremely convenient, as it
229: enables us to adopt, in the comoving reference frame, the usual statistical
230: mechanics on the $\mathbf{E}^{3}$ manifold (the phase space measure is given
231: simply by $\frac{1}{\left( 2\pi \right) ^{3}}\iint d^{3}xd^{3}p$).
232:
233: The energy density turning up in (\ref{C1_11}) is \cite{nar,kolb}
234: \begin{equation}
235: \rho _{UR}=\rho _{\gamma }+\rho _{e^{-}}+\rho _{e^{+}}+\rho _{\nu _{e}}+\rho
236: _{\nu _{\mu }}+\rho _{\bar{\nu}_{e}}+\rho _{\bar{\nu}_{\mu }}=\frac{9}{2}\,
237: \rho _{\gamma }=\frac{9\pi ^{2}}{30}(kT)^{4} \label{C1_17}
238: \end{equation}%
239: and can be used to obtain the Hubble function as a function of $kT$:%
240: \begin{equation}
241: H(kT)=\left( \frac{\dot{a}}{a}\right) =\sqrt{\frac{4\pi ^{3}}{5}} \, \frac{%
242: (kT)^{2}}{m_{Pl}}, \label{C1_18}
243: \end{equation}%
244: where $m_{Pl}\equiv G^{-1/2}=1.221\times 10^{22}$ $MeV$. The relation
245: between the scale factor and the energy $kT$ is determined by comparing (\ref%
246: {C1_17})\ to (\ref{C1_10}):%
247: \begin{equation}
248: kT=\frac{A}{a} \,\, . \label{C1_19}
249: \end{equation}%
250: $A$ is a constant which depends on the present--day values $T_{0}$ and $%
251: a_{0} $. Substituting this last equation into (\ref{C1_18}) and solving the
252: resulting differential equation -- imposing the initial condition $a(t=0)=0$%
253: \ -- leads to%
254: \begin{equation}
255: a(t)=A\left( \frac{16\pi ^{3}}{5}\right) ^{1/4}\left( \frac{t}{m_{Pl}}%
256: \right) ^{1/2}. \label{C1_20}
257: \end{equation}%
258: From (\ref{C1_19}) follows then the relation between the energy $kT$ and the
259: cosmological time $t$:
260: \begin{equation}
261: kT=\left( \frac{16\pi ^{3}}{5}\right) ^{-1/4}\left( \frac{t}{m_{Pl}}\right)
262: ^{-1/2} \label{C1_21}
263: \end{equation}%
264: The last four equations determine the evolution of the primeval universe in
265: a radiation--dominated era.
266:
267: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
268:
269: \section{Statistics of interacting systems\label{sec-SisInt}}
270:
271: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
272:
273: It was discussed elsewhere \cite{artigo 1} how to construct equations of
274: state for an interacting system of particles within the standard ensemble
275: formalism of statistical mechanics, and how this could be applied to a
276: simple model relevant to cosmology. In what follows, we will present a brief
277: review of the information needed here.\ We also indicate Refs. \cite{Dashen}%
278: , \cite{Beth}, \cite{mayer} and \cite{LeeYang} for further information.
279:
280: The grand canonical partition function $\Theta $\ is expressed as
281: \begin{equation}
282: \Theta (z,V,T)=\sum\limits_{N=0}^{\infty }Q_{N}(V,T)z^{N}, \label{s1}
283: \end{equation}%
284: where $V$ is the volume, $T$ is the temperature, $z=e^{\mu /kT}$ is the
285: fugacity and $Q_{N}(V,T)$ is the N-particle canonical partition function.
286: The chemical potential $\mu =\mu ^{NR}+m$ is composed by the rest mass $m$
287: plus the non-relativistic chemical potential $\mu ^{NR}$. This last is that
288: usually found in standard statistical mechanics texts, e.g. \cite{Pat}.
289: Here, we are concerned only with one-component systems. For the cases of
290: more constituents, see Appendix \ref{ap-Multicomp}.
291:
292: The thermodynamical quantities -- pressure $p$, energy density $\rho $,
293: numerical density $n$, etc. -- are related to the grand canonical potential
294: \begin{equation}
295: \Omega (z,T)\equiv \frac{1}{V}\ln \Theta (z,V,T)=\sum\limits_{N=1}^{\infty
296: }b_{N}z^{N}, \label{s2}
297: \end{equation}%
298: written in terms of the \textit{cluster integrals} $b_{N}$. The
299: thermodynamical limit has already been taken in Eq. (\ref{s2}), so that the $%
300: b_{N}$ are functions of the temperature solely (the $V$ dependence
301: disappears). According to Dashen, Ma and Bernstein \cite{Dashen}, the
302: cluster integrals are calculated from the $S$-matrix as follows:
303: \begin{equation}
304: b_{N}-b_{N}^{(0)}=\frac{g_{N}}{V}\int \frac{e^{-\beta E}}{4\pi i}Tr\left(
305: \hat{A}\hat{S}^{-1}\frac{\overleftrightarrow{\partial }}{\partial E}\hat{S}%
306: \right) _{\tilde{c}_{N}}dE, \label{s3}
307: \end{equation}%
308: where $\beta =1/kT$, $g_{N}$ counts the degeneracy coming from internal
309: degrees of freedom, $\hat{A}$ symmetrizes (antisymmetrizes) the bosonic
310: (fermionic) states, $\hat{S}$ is the scattering matrix operator and $%
311: b_{N}^{(0)}$ is the cluster integral of the non-interacting quantum system.
312: The subscript $\tilde{c}_{N}$ represents all the $N$-particle connected
313: diagrams\footnote{%
314: An $N$-particle connected diagram is a graphic representation of $N$ balls
315: linked directly or indirectly by lines which represent the correlations
316: coming from interactions or statistical effects.} in which the interaction
317: occurs at least once. In (\ref{s3}), it was used the short-cut
318: \begin{equation}
319: \hat{S}^{-1}\frac{\overleftrightarrow{\partial }}{\partial E}\hat{S}\equiv
320: \hat{S}^{-1}\frac{\partial \hat{S}}{\partial E}-\frac{\partial \hat{S}^{-1}}{%
321: \partial E}\hat{S}. \label{s4}
322: \end{equation}%
323: By using (\ref{s2}) $p$, $n$ and $\rho $ are determined as series in the
324: fugacity:
325: \begin{equation}
326: \frac{p(z,T)}{kT}=\sum\limits_{N=1}^{\infty }b_{N}z^{N}~;\text{ \ }%
327: n(z,T)=\sum\limits_{N=1}^{\infty }Nb_{N}z^{N};\text{ \ }\rho
328: (z,T)=-\sum\limits_{N=1}^{\infty }\frac{\partial b_{N}}{\partial \beta }%
329: z^{N}. \label{D32A}
330: \end{equation}%
331: This is the parametric form of the equations of state.
332:
333: An alternative description is given by the pressure $p(n,T)$ and the energy
334: density $\rho (n,T)$ written in terms of $n\left( z,T\right) $. They are
335: obtained by inversion of series $n\left( z,T\right) $ and substitution into $%
336: p(z,T)$ and $\rho (z,T)$. This results in the \textit{virial expansion}:
337: \begin{equation}
338: \frac{p(n,T)}{kT}=\sum_{l=1}^{\infty }a_{l}(T)n^{l};\ \ \ \frac{\rho (n,T)}{%
339: \left( kT\right) ^{2}}=\sum_{l=1}^{\infty }c_{l}(T)n^{l}, \label{D33}
340: \end{equation}%
341: with $a_{l}$ and $c_{l}$ representing the virial coefficients for the
342: pressure and the energy density. These coefficients are completely
343: determined by the $b_{N}$. Appendix \ref{ap-Multicomp} presents the explicit
344: forms of $a_{l}$ and $c_{l}$ for a two-component system.
345:
346: In practice, it is not possible to really sum any of the series (\ref{D32A}-%
347: \ref{D33}). Therefore, the choice of using $p$ and $\rho $ in terms of $z$
348: or $n$ depends on the perturbative characteristics of each series and on the
349: system under study. We will return to this question in Section \ref%
350: {sec-EoS_Pns}.
351:
352: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
353:
354: \subsection{The second coefficient of the fugacity expansion}
355:
356: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
357:
358: The elastic interaction between two particles is decomposed into
359: rotation--invariant sectors, so that each part depends only on their
360: relative distance $r$ (central interaction):
361: \begin{equation}
362: \hat{V}^{a}(r)\equiv \hat{H}^{a}-\hat{H}^{(0)a}, \label{Rcinquenta
363: oito}
364: \end{equation}
365: index $a$ standing for the other types of invariance: spin, isospin, charge
366: conjugation, etc.\footnote{%
367: \, For instance, proton-neutron interaction ($pn$) related to total spin $S$%
368: , total angular momentum $J$ and orbital angular momentum $L$ can be
369: represented by a central interaction operator such as $\hat{V}%
370: _{^{2S+1}L_{J}}^{pn}(r)$.} The scattering matrix $S_{2}^{a}$\ depends then
371: only on the energy and the angle between the initial and final momenta. In
372: this angular momentum representation, the $\hat{S}_{2}^{a}$\ operator can be
373: written solely in terms of the phase shifts $\delta _{l}^{a}(k)$:
374: \begin{equation}
375: \left\langle k^{\prime }l^{\prime }m^{\prime }\right\vert \hat{S}%
376: _{2}^{a}\left\vert klm\right\rangle =e^{2i\delta _{l}^{a}(k)}\delta
377: _{k^{\prime },k}\delta _{l^{\prime },l}\delta _{m^{\prime },m}.
378: \label{Rsetenta quatro}
379: \end{equation}
380:
381: The symmetrization performed by operator $\hat{A}$\ in (\ref{s3}) must
382: account for the complete state associated with $a$. It is possible to
383: transfer this symmetrization instruction to the index $a$. Once this is
384: done, the cluster integral of the two-particle system is written as%
385: \begin{equation}
386: b_{2}-b_{2}^{(0)}=\sum\limits_{a}\frac{g_{2}^{a}}{V}\int \frac{dE}{4\pi i}%
387: e^{-\beta E}Tr\left[ \left( \hat{S}_{2}^{a}\right) ^{-1}\frac{%
388: \overleftrightarrow{\partial }}{\partial E}\hat{S}_{2}^{a}\right] ,
389: \label{Rsetenta oito}
390: \end{equation}%
391: where $g_{2}^{a}$ counts the degeneracy degree of the duly symmetrized
392: states.
393:
394: If we carry out a coordinate transformation to the center of mass, the state
395: will behave just like a free state (plane wave), that is, it will obey the
396: free Hamiltonian. In order to explore this fact, we remember some properties
397: of the relativistic invariant $s$:%
398: \begin{align}
399: \omega ^{2}& \equiv s=\left( p_{1\mu }+p_{2\mu }\right)
400: ^{2}=m_{1}^{2}+m_{2}^{2}+2p_{1\mu }p_{2}^{\mu } \notag \\
401: & =E_{1}^{2}+E_{2}^{2}+2E_{1}E_{2}-\left( \vec{p}_{1}+\vec{p}_{2}\right)
402: ^{2}=E^{2}-\vec{P}^{2}, \label{Rsetenta nove}
403: \end{align}%
404: $m_{1}$ and $m_{2}$ are the masses of the two particles,\ $E=E_{1}+E_{2}$ is
405: the sum of their energy, and the momentum of the center of mass is $\vec{P}%
406: \equiv \vec{p}_{1}+\vec{p}_{2}$. Since $E$ and $\vec{P}$ are the energy and
407: the momentum of the two-particle cluster, $\omega $ may be understood as the
408: \textit{mass} of this cluster and encapsulates all the information about the
409: interaction. As the interaction calculated by (\ref{Rsetenta oito}) does not
410: depend on the coordinates of the center of mass, the integration and the
411: differential operator with respect to $E$ are reexpressed as%
412: \begin{equation}
413: b_{2}-b_{2}^{(0)}=\sum\limits_{a}\frac{g_{2}^{a}}{V}\int d^{3}R\int \frac{%
414: d^{3}P}{\left( 2\pi \right) ^{3}}\int \frac{d\omega }{4\pi i}e^{-\beta \sqrt{%
415: \vec{P}^{2}+\omega ^{2}}}Tr\left[ \left( \hat{S}_{2}^{a}\right) ^{-1}\frac{%
416: \overleftrightarrow{\partial }}{\partial \omega }\hat{S}_{2}^{a}\right] .
417: \label{Roitenta quatro}
418: \end{equation}%
419: The next step is to integrate on $d^{3}P$:
420: \begin{equation}
421: b_{2}-b_{2}^{(0)}=\sum\limits_{a}\frac{g_{2}^{a}}{2\beta \pi ^{2}}%
422: \int\limits_{M}^{\infty }\omega ^{2}K_{2}(\beta \omega )\frac{1}{4\pi i}Tr%
423: \left[ \left( \hat{S}_{2}^{a}\right) ^{-1}\frac{\overleftrightarrow{\partial
424: }}{\partial \omega }\hat{S}_{2}^{a}\right] d\omega , \label{s5}
425: \end{equation}%
426: where $M=m_{1}+m_{2}$ is the minimum value assumed by $\omega $; $%
427: K_{2}(\beta \omega )$ is the modified Bessel function. Using (\ref{Rsetenta
428: quatro}), the trace in the angular momentum representation is
429:
430: \begin{equation}
431: Tr\left[ \left( \hat{S}_{2}^{a}\right) ^{-1}\frac{\overleftrightarrow{%
432: \partial }}{\partial \omega }\hat{S}_{2}^{a}\right] =4i\sum\limits_{l=0}^{%
433: \infty }\left( 2l+1\right) \frac{\partial \delta _{l}^{a}(\omega )}{\partial
434: \omega } \,\, . \label{Rnoventa um}
435: \end{equation}%
436: Equation (\ref{s5}) becomes:%
437: \begin{equation}
438: b_{2}-b_{2}^{(0)}=\frac{1}{2\beta \pi ^{3}}\sum\limits_{a}\sum%
439: \limits_{l=0}^{\infty }g_{2}^{a}\left( 2l+1\right) \int\limits_{M}^{\infty
440: }\omega ^{2}K_{2}(\beta \omega )\frac{\partial \delta _{l}^{a}(\omega )}{%
441: \partial \omega }d\omega . \label{Rnoventa dois A}
442: \end{equation}%
443: This equation, which reduces to the Beth~--~Uhlenbeck \cite{Beth,Pat} case
444: in the non-relativistic limit, is the fundamental tool to obtain the EOS for
445: the pre-nucleosynthesis universe.
446:
447: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
448:
449: \section{Equations of state for the pre-nucleosynthesis universe \label%
450: {sec-EoS_Pns}}
451:
452: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
453:
454: The EOS to be built here are supposedly realistic because they account for
455: the interactions. Nevertheless, we will not actually consider all the four
456: fundamental interactions.
457:
458: The gravitational interaction is accounted for only through Einstein (more
459: specifically, Friedmann) equations. We will neglect any possible change in
460: the description of statistical mechanics due to the curved background of the
461: cosmic manifold.\footnote{%
462: \ Ref. \cite{AlBelPe}\ discusses this subject for the de Sitter solution.}
463:
464: The electromagnetic interaction, though responsible for the thermalization
465: of the charged particles, will not be relevant to the primeval EOS, since:
466: (i) the shielding effect due to the existence of opposite charges makes it
467: possible to consider the effective interaction as of short range, and the
468: fluid as neutral; (ii) the coupling constant of the electromagnetic
469: interaction is relatively small ($e^{2}\sim 1/137$). The mean interaction
470: energy between charged particles is nearly two orders of magnitude lesser
471: than their mean kinetic energies \cite{nar}. The weak interaction, though
472: responsible for the thermalization of the neutral particles, is of
473: short-range, and its intensity is much smaller \cite{Grif} than the strong
474: interaction. It will be neglected. Due to its high coupling constant, and
475: despite its short-range, the strong (``hadronic'') interaction is dominant
476: in the PNS period. As a matter of fact, it is the only one that will be
477: taken into account (see Appendix \ref{ap-Espalha}).
478:
479: Justified by these considerations, we divide the relevant particles for the
480: PNS in three categories:
481:
482: \begin{enumerate}
483: \item Ideal ultrarelativistic particles: $\gamma ,e^{-},e^{+},\nu _{e},\nu
484: _{\mu },\bar{\nu}_{e},\bar{\nu}_{\mu }$;
485:
486: \item Ideal relativistic particles: $\mu ^{-},\mu ^{+}$;
487:
488: \item Interacting relativistic particles: $\pi ^{+},\pi ^{-},\pi
489: ^{0},K^{+},K^{-},K^{0},\bar{K}^{0},p,\bar{p},n,\bar{n}$.
490: \end{enumerate}
491:
492: The equations of state for such a system of particles is the sum of the
493: contributions of each species to quantities $p$ and $\rho $, keeping in mind
494: what has been said in Section \ref{sec-SisInt} whenever interactions are
495: important. Both $p$ and $\rho $\ can be written as functions of $(z,T)$ or
496: in terms of $(n,T)$ -- Eqs. (\ref{D32A}-\ref{D33}). Thus, there are four
497: pairs $\left\{ p,\rho \right\} $\ given by the combination of $p(z,T)$, $%
498: p(n,T)$, $\rho (z,T)$\ and $\rho (n,T)$. We have to decide which combination
499: is the most suitable.
500:
501: There are strong theoretical arguments (based on the series convergences)
502: and outstanding experimental indications \cite{artigo 1,Hirshefeld} that the
503: virial expansion $p(n,T)$\ is the most convenient form for the pressure. The
504: choice between $\rho (z,T)$ and $\rho (n,T)$ is more subtle. We shall keep
505: the form $\rho (z,T)$ since it maintains the probabilistic notion inherited
506: by statistical mechanics. In fact, the energy density is a mean weighted by
507: the Boltzmann factor, and in the case of the grand canonical ensemble it
508: reads%
509: \begin{equation}
510: \rho =\frac{E}{V}=\frac{1}{\Theta (V,T,z)}\sum\limits_{a,r}\frac{E_{a}}{V}%
511: e^{-\beta E_{a}}z^{N_{r}}. \label{s7}
512: \end{equation}%
513: Comparing this equation to (\ref{s2}) and (\ref{D32A}) one sees that only $%
514: \rho (z,T)$\ preserves the probabilistic character order by order.
515:
516: Hence, the EOS for the relativistic particles (interacting or not) will be
517: formed by the virial series for the pressure and the fugacity series for the
518: energy density: $\left\{ p(n,T),\rho (z,T)\right\} $. Only the first terms
519: of the expansions will be considered, as only two-by-two interactions can be
520: calculated or experimentally measured. Following the classification above,
521: the pressure and energy density will be
522: \begin{eqnarray}
523: p &=&p_{Had}+p_{R}+p_{UR}=p_{Had}+p_{R}+\frac{\rho _{UR}}{3}, \label{C5_3}
524: \\
525: \rho &=&\rho _{Had}+\rho _{R}+\rho _{UR}, \label{C5_4}
526: \end{eqnarray}%
527: where the labels $Had$, $R$ and $UR$ represents the hadrons, the ideal
528: relativistic particles and the ideal ultrarelativistic particles.
529:
530: %%%%%%%%%%%%%%%%%
531:
532: \subsection{EOS for the hadrons \label{sec-EoS_Had}}
533:
534: %%%%%%%%%%%%%%%%%
535: Pions $\pi ^{+},\pi ^{-},\pi ^{0}$, kaons $K^{+},K^{-},K^{0},\bar{K}^{0}$\
536: and nucleons $N = p,\bar{p},n,\bar{n}$\ exhibit spin, isospin and charge
537: conjugation symmetries under strong interactions (Appendix \ref{ap-Espalha}%
538: ). For this reason, the set of hadrons can be treated as an interacting
539: system of three components -- $\pi $, $K$\ and $N$\ -- whose EOS are derived
540: from (\ref{D41c}) and (\ref{D42a}). The energy density will be
541: \begin{eqnarray}
542: \rho _{Had}(z_{\pi },z_{K},z_{N},kT) &=&\left( kT\right) ^{2}\left[ \dot{b}%
543: _{1\pi }z_{\pi }+\dot{b}_{1K}z_{K}+\dot{b}_{1N}z_{N}+\right. \notag \\
544: &&\left. +\dot{b}_{2\pi \pi }z_{\pi }^{2}+\dot{b}_{2KK}z_{K}^{2}+\dot{b}%
545: _{2NN}z_{N}^{2}+\right. \notag \\
546: &&\left. +\dot{b}_{2\pi K}z_{\pi }z_{K}+\dot{b}_{2\pi N}z_{\pi }z_{N}+\dot{b}%
547: _{2KN}z_{K}z_{N}\right] , \label{C5_6}
548: \end{eqnarray}%
549: where the dot indicates differentiation with respect to $kT$. To this point
550: we have not used the fact (already discussed) that $\mu _{\pi }\simeq \mu
551: _{K}\simeq \mu _{N}\simeq 0$ during the pre-nucleosynthesis period, and
552: consequently $z_{\pi }\simeq z_{K}\simeq z_{N}\simeq 1$. The pressure is
553: \begin{eqnarray}
554: p_{Had}(n_{\pi },n_{K},n_{N},kT) &=&kT\left( a_{1\pi }n_{\pi
555: }+a_{1K}n_{K}+a_{1N}n_{N}+\right. \notag \\
556: &&\left. +a_{2\pi \pi }n_{\pi }^{2}+a_{2KK}n_{K}^{2}+a_{2NN}n_{N}^{2}+\right.
557: \notag \\
558: &&\left. +a_{2\pi K}n_{\pi }n_{K}+a_{2\pi N}n_{\pi
559: }n_{N}+a_{2KN}n_{K}n_{N}\right) , \label{C5_8}
560: \end{eqnarray}%
561: with the numerical densities given by
562: \begin{eqnarray}
563: n_{\pi }(z_{\pi },z_{K},z_{N},kT) &=&b_{1\pi }z_{\pi }+2b_{2\pi \pi }z_{\pi
564: }^{2}+b_{2\pi K}z_{\pi }z_{K}+b_{2\pi N}z_{\pi }z_{N}, \label{C5_9a} \\
565: n_{K}(z_{\pi },z_{K},z_{N},kT) &=&b_{1K}z_{K}+2b_{2KK}z_{K}^{2}+b_{2\pi
566: K}z_{\pi }z_{K}+b_{2KN}z_{K}z_{N}, \label{C5_9b} \\
567: n_{N}(z_{\pi },z_{K},z_{N},kT) &=&b_{1N}z_{N}+2b_{2NN}z_{N}^{2}+b_{2\pi
568: N}z_{\pi }z_{N}+b_{2KN}z_{K}z_{N}. \label{C5_9c}
569: \end{eqnarray}%
570: The explicit form of the ideal and interaction-related terms are given in
571: Appendix \ref{ap-Multicomp}. The rest mass adopted are: $m_{\pi }=0.1396$ $%
572: GeV$, $m_{K}=0.4957$ $GeV$ and $m_{N}=0.93826$ $GeV$.
573:
574: In principle, the term $b_{2}$\ should contain all the interaction processes
575: involving the two particles to which it refers. That is, it should describe
576: the elastic and inelastic scatterings as well as the bound states. Actually,
577: it is possible to argue that, in the construction of the EOS for the
578: pre-nucleosynthesis period, bound states and inelastic processes are
579: negligible when compared to the elastic channels.
580:
581: Taking into account just the elastic processes, the six interaction terms ($%
582: \pi \pi $, $\pi K$, $\pi N$, $KK$, $KN$ and $NN$) are given by equations of
583: type (\ref{Rnoventa dois A}). The explicit expression of the cluster
584: integrals are obtained with the help of all the phase-shift data sets -- see
585: Appendix \ref{ap-Espalha} for an example. It is necessary to perform an
586: integration involving the modified Bessel function $K_{2}$ and derivatives
587: of the phase-shifts. The six integrations have been done numerically with
588: the software \emph{Mathematica 5.0}. Before that, we carefully put back the
589: constants $c$ and $\hbar $ in such a way that the length and energy units
590: were $fm$ and $GeV$, respectively.
591:
592: The higher limits of the integrals are different for each $b_{2}$, since the
593: phase-shift data sets are obtained in different energy intervals. Still,
594: most of them lie in the interval from $1$ $GeV$ to $2.5$ $GeV$. As the
595: relevant temperature values for our studies are in the range $20$ $MeV\leq
596: kT\lesssim 300$ $MeV$, the chosen values for the\ higher limits are good
597: enough to include all the main contribution from the integral kernels.
598: Besides, the Bessel function $K_{2}$ assures the fast decrease of the kernel
599: values. We consider only angular momenta $l=0,1$ and $2$, i.e., the $S$ , $P$
600: and $D$ contributions (Appendix \ref{ap-Espalha}). Once the coefficients $%
601: b_{1}$ and $b_{2}$ have been calculated, it is straightforward to obtain the
602: numerical densities (\ref{C5_9a}), (\ref{C5_9b}) and (\ref{C5_9c}), and the
603: equations of state for the hadrons, Eqs. (\ref{C5_6}) and (\ref{C5_8}).
604:
605: The approach described above is not the only one for building hadronic EOS
606: at few hundreds of $MeV$. There is also the treatment called Hadron
607: Resonance Gas (HRG) \cite{Munzinger,Andro,tawfik2,Cheng} describing the
608: Fireballs created in the Heavy Ion Colliders (SPS -- Super Proton Synchroton
609: -- and RHIC -- Relativistic Heavy Ion Collider). This technique agrees
610: qualitatively with the Lattice QCD calculations \cite{Cheng} and it
611: reproduces the abundance of the observed particles relatively well \cite%
612: {Andro}. The HRG modeling is equivalent to ours as long as the ideal terms
613: are concerned, but the procedure of including the interaction is rather
614: different. The Hadron Resonance Gas accounts for interaction by introducing
615: a excluded volume term \emph{a la} Van der Walls \cite{Yen} and by
616: considering the resonance contributions \cite{Munzinger,Andro}.\footnote{%
617: The HRG model is similar to the Hagedorn's statistical bootstrap model \cite%
618: {Hag} in the aspect that both cases include the hadronic interactions
619: primary through the resonances.} On the other hand, our work accounts for
620: interactions through the phase shifts. Because of this, other effects --
621: besides the ones related to resonances and hard cores (excluded volumes) --
622: are considered. Reference \cite{Delta} analyzes these features in detail for
623: the case of $\Delta $ resonance in the $\pi N$ interaction. It is worth to
624: emphasize that in our approach the resonances are only characteristics of
625: the elastic scattering between stable hadrons; the resonances are not taken
626: as particles.
627:
628: In connection with what we said in the Introduction: the construction of
629: hadronic EOS through the S-Matrix was already implemented in Refs. \cite%
630: {ucraone,ucratwo}. However, the only processes considered there (e.g. \cite%
631: {ucraone}) are those including the nucleons (scattering $NN$). The present
632: work is more constructive and complete in this sense: all the important
633: statistical quantities -- pressure, energy density and numerical density --
634: are obtained from the partition function which accounts for all the relevant
635: processes of interaction two by two (scattering $\pi \pi $, $\pi K$, $\pi N$%
636: , $KK$, $KN$ and $NN$).
637:
638: %%%%%%%%%%%%%%%%%%%%%%
639:
640: \subsection{EOS for the ideal particles}
641:
642: %%%%%%%%%%%%%%%%%%%%%%
643: The ideal (non-interacting) particles are divided into two categories:
644: ultrarelativistic ($\gamma , e^{-},\allowbreak e^{+}, \nu _{e}, \nu _{\mu },
645: \bar{\nu}_{e}, \bar{\nu}_{\mu }$) and relativistic ($\mu ^{-},\mu ^{+}$).
646: The ultrarelativistic sector is relatively simple, and has been discussed in
647: Section \ref{sec-Pns}. The calculation for the relativistic sector is
648: analogous to what we have done for the hadrons. The energy density, for
649: instance, is
650: \begin{equation}
651: \rho _{R}(z_{\mu },kT)=\left( kT\right) ^{2}\left[ \dot{b}_{1\mu
652: }^{(0)}z_{\mu }+\dot{b}_{2\mu }^{(0)}z_{\mu }^{2}\right] . \label{C5_28}
653: \end{equation}%
654: The $b_{1\mu }^{(0)}$ and $b_{2\mu }^{(0)}$ are the ideal terms determined
655: by (\ref{Aj11}) and (\ref{Aj12}), with $m_{\mu }=0.10566$ $GeV$. As in the
656: hadronic case, we set $z_{\mu }\simeq 1$ during the PNS. The pressure is%
657: \begin{equation}
658: p_{R}(n_{\mu },kT)=kT\left[ a_{1\mu }n_{\mu }+a_{2\mu }n_{\mu }^{2}\right]
659: \text{ \ \ \ \ with \ \ \ } n_{\mu }(z_{\mu },kT)=b_{1\mu }^{(0)}z_{\mu
660: }+2b_{2\mu }^{(0)}z_{\mu }^{2} \, . \label{C5_30}
661: \end{equation}
662:
663: %%%%%%%%%%%%%%%%
664:
665: \subsection{The complete EOS}
666:
667: %%%%%%%%%%%%%%%%
668:
669: The complete EOS for the energy density $\rho $ is obtained by substituting
670: Eqs.(\ref{C5_6}), (\ref{C1_17}) and (\ref{C5_28}) into (\ref{C5_4}). The
671: plot of $\rho(kT)$ is shown in Figure \ref{fig1}. The three curves exhibit
672: similar features: all are positive and increase monotonically. The
673: ultrarelativistic components dominate up to $\approx 0.12$ $GeV$. From $0.12$
674: $GeV$ to $0.16$ $GeV$, the other particles (mainly $\pi $ and $\mu $) become
675: important. At $0.275$ $GeV$ the ultrarelativistic, ideal-relativistic, and
676: interacting sectors correspond respectively to $25\%$, $33\%$ and $42\%$ of
677: the total energy density. %%%%%%%%%%%%%%%%%%%%%%%%%%%%
678: %%%%%%%%%% FIGURA %%%%%%%%%%
679: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
680:
681: \begin{figure}[ht]
682: \begin{center}
683: \includegraphics[height=6.4cm, width=8.5cm]{densidade_de_energia.eps}
684: \end{center}
685: \caption{Plot of the energy density $\protect\rho $ in $GeV/fm^{3}$ as a
686: function of the thermal energy $kT$ given in $GeV$. Three curves are
687: presented: the complete energy density which includes the effects of all the
688: relevant particles and their interactions (full line); energy density of all
689: particles (ultrarelativistic, relativistic and hadronic) without considering
690: interactions, i.e., all the particles are taken as ideal ones (dashed line);
691: energy density for the ultrarelativistic particles (dotted line).}
692: \label{fig1}
693: \end{figure}
694: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
695:
696: \bigskip
697:
698: The complete EOS for the pressure $p$ comes from the substitution of Eqs.( %
699: \ref{C5_8}), (\ref{C1_17}) and (\ref{C5_30}) into (\ref{C5_3}). Figure \ref%
700: {fig2} shows the plot of function $p=p(kT)$.%
701: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
702: %%%%%%%%%% FIGURA %%%%%%%%%%
703: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
704:
705: \begin{figure}[ht]
706: \begin{center}
707: \includegraphics[height=6.4cm, width=8.5cm]{pressao.eps}
708: \end{center}
709: \caption{Plot of the pressure $p$ in $GeV/fm^{3}$ as a function of the
710: energy $kT$ given in $GeV$. Interactions are accounted for only in the
711: complete case (full line). The conclusion is: interaction drags the pressure
712: to negative values, pointing to cosmic acceleration.}
713: \label{fig2}
714: \end{figure}
715: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
716:
717: \bigskip
718:
719: It can be seen in Figure \ref{fig2} that, up to $0.14$ $GeV$, the three
720: curves for the pressure present similar behavior: as in the energy density
721: case, all are positive--valued and increase monotonically. From this point
722: on the interaction contributions grow, and there is a sudden change in the
723: slope for the complete system. The curve mitigates its increase rate,
724: reaches a maximum at $0.175$ $GeV$, and then starts a steep fall. It becomes
725: negative at around $0.2$ $GeV$ and continues to decrease at a high rate. A
726: comparative graphic analysis of the effects on $\rho $ and $p$ coming from
727: the different kinds of interactions and scatterings reveals that the
728: relevant processes for the determination of $b_{2}$ in scales of $kT\lesssim
729: 300$ $MeV$ are the scatterings $\pi \pi $, $\pi K$ and $\pi N$ related to
730: the $S$ and $P$ partial waves. Therefore, it is fair enough to take into
731: account only $l=0,1$ and $2$ when calculating $b_{2}$, just as we did.
732:
733: A last point we would like to discuss is the validity range of the
734: expansions $\rho (kT)$ and $p(kT)$. This validity is threatened by three
735: factors: \textbf{(I)} an eventual phase transition due to instabilities in
736: the function $p(kT)$; \textbf{(II)} the non-convergence of the pressure and
737: energy density series, which would not allow the perturbative approach
738: adopted here; \textbf{(III)} the occurrence of the deconfinement of hadrons
739: in quarks and gluons, which would change the nature of the particles. Let us
740: comment on these issues:
741:
742: \textbf{(I)} There are two necessary and sufficient conditions that assure
743: the stability of the usual thermodynamical systems \cite{Callen}, namely:
744: \begin{equation}
745: \frac{1}{T}\left( \frac{\partial \rho }{\partial T}\right) _{n_{i}}\geq 0%
746: \text{ \ \ \ and \ \ \ }n_{i}\left( \frac{\partial p}{\partial n_{i}}\right)
747: _{T}\geq 0 \, , \label{s8}
748: \end{equation}%
749: with $i=\pi $, $K$, $N$ or $\mu $.
750:
751: The function $\rho $\ depends only on $kT$\ ($z_{i}=1$ for every species $i$%
752: ), while $p$\ depends on $kT$\ and also on the densities $n_{i}$\ of the
753: four types of particles. Thus, the second condition in (\ref{s8})\ splits
754: into four others, all required to guarantee the system stability. From
755: Figure \ref{fig1}, it is easily seen that the curve $\rho (kT,\{z_{i}\})$
756: for the complete set always satisfies (\ref{s8}).\footnote{$\left\{
757: z_{i}\right\} $ represents the set of the four fugacities, all of which
758: equals one.}
759:
760: Notwithstanding, the same does not occur with $p\left( kT,\{n_{i}\}\right) $%
761: . Indeed, as the numerical densities are positive and increasing functions
762: of $kT$, Figure \ref{fig2} makes clear that the pressure does not obey\
763: condition (\ref{s8}); i.e., for some energy value higher than $0.14$ $GeV$
764: the EOS for the pressure becomes unstable. In principle, this instability is
765: a strong indication of a phase transition and points to a breakdown of the
766: EOS validity. However, this argument is based on the usual situation of a
767: thermodynamical system: a gas confined in a recipient of finite volume. By
768: its very nature, the universe cannot be assumed to have the same
769: characteristics of such a simple and controlled environment. First: the
770: universe has no frontiers. In this case, the surface pressure (which is
771: different from the internal pressure) does not exists. And in many cases are
772: precisely the surface effects which generate the phase transitions in an
773: ordinary thermodynamical system. Second: in the context of cosmology,
774: \textit{the (internal) pressure is a direct source of the gravitational field%
775: } \cite{artigo 1}. This particular feature produces rather different effects
776: from those engendered by the pressure in usual thermal systems, for which
777: one would expect the reduction in the volume as the internal pressure
778: increases. This notion, spelled by the second relation in (\ref{s8}), is of
779: great importance, since it prevents the ordinary system from disappearing:
780: if $\partial p/\partial V$\ were positive, nothing would avoid the collapse
781: of the gas. On the other hand, in the context of cosmology, an increase of
782: pressure increases the gravitational attractive effect, and in this sense, a
783: cosmological system with $\partial p/\partial V<0$ is unstable, i.e., if the
784: universe is not expanding it tends to collapse. We want to emphasize the
785: distinction between the effects of the internal pressure in the two types of
786: systems: in ordinary equilibrium thermodynamics, an increase of $p$ leads to
787: an increase of the volume $V$ of the recipient containing the gas because
788: the shocks of the particles against the walls are more and more frequent;
789: but in cosmology the increase of $p$ strengthens the gravitational field and
790: produces a net tendency to the reduction of the volume. Hence, it is
791: reasonable to say that the stability criterion (\ref{s8}) cannot be directly
792: applied in cosmology, and cannot be used to rule out the above EOS. From the
793: microscopic point of view, the energy density fluctuations $\delta \rho$
794: grow fast in regions where $\partial p/\partial \rho \equiv c_{s}^2<0$ ($%
795: c_{s}$ is named the sound velocity in the media) and the system becomes
796: unstable. For the usual thermodynamical systems this strongly suggests phase
797: transition. Our system, the standard cosmological model, is not of this
798: type, though. The universe expands in a rate determined by the relation
799: between $p$ and $\rho$. This complicates the analyzes and we can not affirm
800: that $c_{s}^2<0$ necessarily leads to a phase transition of the system. The
801: correct treatment of the evolution of $\delta \rho$ should be done along
802: with the perturbations in the space-time geometry -- FRW metrics. We leave
803: these investigations for the future.
804:
805: \textbf{(II)} The convergence of the expansions for $p$ and $\rho $ is
806: related with the formal aspect of validity of these series and the
807: correction of the perturbative treatment. This issue is far from trivial,
808: since is not possible to sum all the terms for any realistic system in a
809: pure thermodynamical context, let alone in the cosmological framework. The
810: virial coefficients $a$, or the cluster integrals $b$, depend on the sum of
811: the connected diagrams representing mutual interactions. And it is extremely
812: difficult to foresee the behavior of the result when $N$ particles interact.
813: So, the answer to the question concerning the convergence of $p$ and $\rho $
814: is inaccessible. Nevertheless, the authors proposed elsewhere \cite{artigo 1}
815: a toy-model for the interactions in the early universe, computed the
816: perturbed EOS (until third order) and showed that there is a good indication
817: of convergence for the pressure series. So, we will assume in this work the
818: validity of truncating the equations for $p$, $\rho$ and $n$ in the second
819: order terms.
820:
821: \textbf{(III)} The QCD coupling constant diminishes as the energy $kT$\
822: increases. So, one could expect that at a certain critical temperature $%
823: kT_{c}$, its value becomes sufficiently low to allow for the deconfinement
824: of the hadronic matter, giving rise to a system composed by quarks and
825: gluons. If such a transition takes place during the PNS, our model ceases to
826: be valid beyond $kT_{c}$: the EOS have been calculated assuming that the
827: fundamental particles are hadrons, not quarks and gluons. According to the
828: lattice QCD calculations \cite{Lattice}, the critical temperature is
829: situated between $150$ and $180$ $MeV$. From an experimental perspective,
830: recent results from the RHIC \cite{RHIC} indicate that a sort of transition
831: occurs at these temperature, but it is not the deconfinement in a
832: quark-gluon plasma \cite{QGP}. The experiments run in RHIC found what seems
833: to be a new state for the nuclear matter. This state would correspond to a
834: perfect fluid (without viscosity) identified as a CGC (\textit{Color Glass
835: Condensate}) \cite{CGC}. Unlike the quark-gluon plasma, the CGC has non
836: negligible correlations, i.e., the nuclear interaction processes are
837: relevant.
838:
839: There is no doubt left by the experiments that the hadronic matter in the
840: heavy ion colliders undergoes a phase transition in energies around $170$ $%
841: MeV$. But it is possible that this phenomenon does not occur in the
842: primordial universe (or that it occurs in different energy scales), due to
843: the various differences between the laboratory and the early universe.
844: Specifically, the temporal scale of the events in the possible cosmological
845: QCD transition ($10^{-5}s$) is quite different from the scale of the
846: transition in the accelerators ($10^{-23}s$) \cite{Vega}. Other differences
847: -- previously cited -- are the non-existence of border in the cosmological
848: system and the role of pressure as direct source of gravitation. It may
849: happen that the new state of nuclear matter (CGC) discovered in RHIC is
850: affected by the absence of borders or by the expansion of the universe. Even
851: though we do not have strong arguments against the cosmological QCD
852: transition, we will assume the hypothesis that it does not occur in the PNS
853: period.
854:
855: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
856:
857: \section{Cosmological consequences \label{sec-Conseq}}
858:
859: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
860:
861: Let us admit that the proposed EOS, Eqs.~(\ref{C5_3}) and (\ref{C5_4}), are
862: valid during all the PNS period. A first application to cosmology can done
863: through the usual parametrization
864: \begin{equation}
865: p(kT)=\omega (kT)\rho (kT). \label{C5_32}
866: \end{equation}%
867: Substituting (\ref{C5_32}) into the second Friedmann equation (\ref{C1_11}),
868: and neglecting $\Lambda $, we find
869: \begin{equation}
870: \left( \frac{\ddot{a}}{a}\right) = - \, \frac{4\pi G}{3}\left[ 1+3\omega (kT)%
871: \right] \rho (kT). \label{C5_33}
872: \end{equation}%
873: As the energy density is a positive and increasing function of $kT$, the
874: information about the acceleration rate depends on the parametric function $%
875: \omega (kT)$\ solely. Isolating this quantity in (\ref{C5_32})\ and using
876: the explicit forms of $\rho (kT)$ and $p(kT)$, the plot of Figure \ref{fig3}
877: obtains. %%%%%%%%%%%%%%%%%%%%%%%%%%%%
878: %%%%%%%%%% FIGURA %%%%%%%%%%
879: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
880: %\begin{tabular}{l}
881: \begin{figure}[ht]
882: \begin{center}
883: \includegraphics[height=6.4cm, width=8.5cm]{omega.eps}
884: \end{center}
885: \caption{Plot of $\protect\omega $ as a function of $kT$. Two curves are
886: shown: the complete $\protect\omega $ function, including all the particles
887: and relevant interactions (full line); ideal $\protect\omega $ function,
888: which computes all the particles but not the mutual interactions (dashed
889: line).}
890: \label{fig3}
891: \end{figure}
892: %\end{tabular}
893: %%%%%%%%%%%%%%%%%%%%%%%%%%%
894: It shows that the complete and the ideal curves for $\omega $ are
895: practically superimposed until $0.11$ $GeV$. From this value on, the
896: complete $\omega $ curve begins to decrease, becoming negative at about $0.2$
897: $GeV$. On the other hand, the ideal $\omega $ curve remains positive and
898: nearly constant up to $0.3$ $GeV$. The effect of the behavior of the
899: complete $\omega $\ function on (\ref{C5_33}) is to change the universe's
900: expansion rate between $0.1$ $GeV<kT<0.3$ $GeV$. As we go back in time, the
901: universe passes continuously from a decelerated dynamics (pretty close to a
902: radiation dominated era) to an accelerated stage. Comparison of the two
903: lines in Figure \ref{fig3} shows that this acceleration is due to the
904: hadronic interaction processes.
905:
906: The next step is to determine the scale factor $a(kT)$ and the function $%
907: t(kT)$\ associating cosmic time and temperature (or thermal energy). We
908: shall use the conservation equation,%
909: \begin{equation}
910: \frac{d\rho }{da}=-\frac{3}{a}(p+\rho ), \label{Rtwenty three}
911: \end{equation}%
912: and the first Friedmann equation (\ref{C1_11}). We obtain $a(kT)$ by
913: integrating (\ref{Rtwenty three}):
914: \begin{equation}
915: a(kT)=a_{Pns0}\exp \left[ -\int\limits_{kT_{Pns0}}^{kT}\frac{\frac{d\rho }{%
916: d\left( kT^{\prime }\right) }}{3(p+\rho )}d\left( kT^{\prime }\right) \right]
917: , \label{Rtwenty fiveA}
918: \end{equation}%
919: where the label $Pns0$ indicates that the quantity is calculated at the end
920: of the pre-nucleosynthesis period which, in energy, is $kT_{Pns0}=20MeV$.
921: The time function $t(kT)$\ is obtained by the combination of (\ref{Rtwenty
922: three}) and (\ref{C1_11}) ($c$ is re-inserted):
923: \begin{equation}
924: t_{Pns0}-t=\int\limits_{kT_{Pns0}}^{kT}\frac{\frac{d\rho }{d\left(
925: kT^{\prime }\right) }}{3(p+\rho )\sqrt{\frac{8\pi G}{3c^{2}}\rho -\frac{%
926: \kappa c^{2}}{a^{2}}}}d\left( kT^{\prime }\right) \text{,\ \ \ \ \ }\frac{%
927: 8\pi G}{3c^{2}}=\frac{996\text{ }fm^{3}}{(ms)^{2}GeV}\,\,\,.
928: \label{Rtwenty six}
929: \end{equation}
930:
931: Equations (\ref{Rtwenty fiveA}) and (\ref{Rtwenty six}) describe the
932: evolution of the pre-nucleosynthesis universe. A careful analysis of these
933: equations reveals that both diverge when $p= -\, \rho $\ and, according to
934: the complete EOS (with interactions included), this happens at $kT_{c}\simeq
935: 0.2767$ $GeV$ (critical temperature).\footnote{%
936: Not to be confounded with the deconfinement temperature, also usually called
937: ``critical temperature''.} The scale factor always decreases as $kT$\
938: increases. The decreasing rate $\frac{da}{d\left( kT\right) }$ is not so
939: well-behaved: it varies a lot and, as $kT$ approaches $kT_{c}$, $\frac{da}{%
940: d\left( kT\right) }\rightarrow -\infty $. This means that the \textit{scale
941: factor tends to zero when} $kT\rightarrow kT_{c}$. In order to study the
942: time function around $kT_{c}$ we have to particularize for the three values
943: of $\kappa $:\
944:
945: \begin{enumerate}
946: \item For $\kappa =0$, the time interval $\left\vert
947: t_{Pns0}-t_{c}\right\vert \rightarrow \infty $ when $kT\rightarrow kT_{c}$,
948: because $\rho (kT) $ and $\frac{d\rho }{d\left( kT\right) }$ are always
949: positive. In a plane space-section universe, the time interval that would
950: elapse from the initial singularity till $t_{Pns0}$\ is infinity, and that
951: singularity is never attained.
952:
953: \item For $\kappa =1$, we see from (\ref{C1_11})\ that it exists a value of $%
954: kT$\ close to $kT_{c}$\ where $\dot{a}=0$. This value, indicated by $kT_{B}$%
955: , is close to $kT_{c}$\ to $39$ decimal digits. Defining $a_{0}$\ as the
956: present--day value of the scale factor, and $a_{B}=a(kT_{B})$\ as the
957: minimum value of the scale factor, then it is found that $a_{B}\simeq
958: 1.7\times 10^{-24}a_{0}$.\footnote{%
959: This calculation considers matter dominance in the interval $kT_{\gamma
960: 0}\simeq 2\times 10^{-4}eV<kT_{\gamma }<$ $1$ $eV$ and radiation dominance
961: in $1$ $eV<kT_{\gamma }<$ $20$ $MeV$, with the initial condition $%
962: a(kT_{\gamma 0})=a_{0}$.} Thus, in an universe of spherical space-section,
963: the initial singularity does not exist; the universe reaches a minimum size
964: measured by $a_{B}$. Recall that recent cosmological data \cite{PDG}
965: slightly favor $\kappa =1$ (as $\Omega _{T0}\simeq 1.003$). Combining this
966: type of solution that presents a minimum size to the requirement that the
967: closed model presents also a maximum size, we obtain the so called \textit{%
968: eternal universe}, or \textit{bouncing universe \cite{Boucing}}. The time
969: interval from the minimum radius till the end of the PNS period is
970: determined by using (\ref{Rtwenty six}) and the values of $H_{0}$ and $%
971: \Omega _{T0}$; it is $\left\vert t_{Pns0}-t_{B}\right\vert \simeq 2.28$ $ms$.
972:
973: \item For $\kappa =-1$, the time interval $\left\vert
974: t_{Pns0}-t_{c}\right\vert \rightarrow 2.28$ $ms$ when $kT\rightarrow kT_{c}$%
975: . Therefore, for a hyperbolic space-section, the initial singularity is
976: reached in a finite time.
977: \end{enumerate}
978:
979: Notice that despite the different possible evolutions, the universe always
980: presents a \textit{maximum temperature} $kT_{c}$.
981:
982: The statistical bootstrap model by Hagedorn also predicts a maximum
983: temperature $kT_{H}$ for the universe; but the mechanisms that lead to it
984: are distinct from the causes for our critical temperature $kT_{c}$. In the
985: context of Hagedorn's model, the increase in the total energy $E$\ is
986: responsible for the rise in the kinetic energy and also for the increase in
987: the number of kinds of particles in the system. As $E$\ rises it is more
988: advantageous to produce new particles than to increase the temperature of
989: the system. This ultimately results in an infinite limit for the energy ($%
990: E\rightarrow \infty $)\ and the pressure ($p\rightarrow \infty $) at a
991: finite value of the temperature ($kT_{H}<\infty $). On the other hand, $%
992: kT_{c}$\ is the temperature associated to the equality $p=-\,\rho $ in our
993: cosmological model with interacting particles -- see comments below Eq. (\ref%
994: {Rtwenty six}); $kT_{c}$\ appears due to the structure of the equations (\ref%
995: {C1_11}) and (\ref{Rtwenty three}).\
996:
997: %%%%%%%%%%%%%%%%%%%
998:
999: \subsection{The inflationary regime}
1000:
1001: %%%%%%%%%%%%%%%%%%%
1002:
1003: Function $\omega (kT)$ evolves continuously from $1/3$ (radiation era; $%
1004: kT\leq 0.02$ $GeV$) and tends to $-1$ (as $kT\rightarrow kT_{c}$). The
1005: inflationary regime (primeval accelerated expansion period) occurs when $%
1006: \omega (kT)<-1/3$. On the other hand, our model requires that $\omega
1007: (kT)>-1 $. We match these requirements by defining the inflationary period
1008: as that for which $-1<\omega (kT)<-\frac{1}{3}$. According to Figure \ref%
1009: {fig3}, this corresponds to the energy interval
1010: \begin{equation}
1011: 0.2356\text{ }GeV\simeq kT_{accel}<kT<kT_{c}\simeq 0.2767GeV. \label{C5_41}
1012: \end{equation}%
1013: An early acceleration is required to solve some cosmological problems --
1014: horizon, flatness, origin of the inhomogeneities, etc. -- without imposing
1015: specific initial conditions \cite{Guth,kolb,Liddle}. An accelerated regime,
1016: however, must exhibit some special features to actually rule out these
1017: problems. We now show that our model indeed eliminates the horizon and
1018: flatness problems.
1019:
1020: The \textit{horizon problem} is the lack of causal connection between
1021: regions far apart which nevertheless exhibit similar physical
1022: characteristics. The region of causal connection is quantified by the Hubble
1023: radius $R\equiv (aH)^{-1}$, the maximal distance particles can travel during
1024: an $e$-fold increase of the scale factor, an increase by the factor $e\simeq
1025: 2.78$ \cite{Dod}. The standard Bib Bang model tells us that, from the time $%
1026: t_{Pns0}$\ until the present day, the maximum commoving distance between two
1027: causal connected regions grew $9$ orders of magnitude:%
1028: \begin{equation}
1029: \frac{R_{0}}{R_{Pns0}}=\frac{a_{Pns0}H_{Pns0}}{a_{0}H_{0}}\sim 1.4\times
1030: 10^{9}. \label{C5}
1031: \end{equation}%
1032: Let $kT_{hor}$\ be the energy value above which all the universe's content
1033: is in causal contact. In order to solve the horizon problem in the context
1034: of our model, we need to show that there is a $kT_{hor}<kT_{c}$ such that%
1035: \begin{equation}
1036: \frac{R_{Pns0}}{R_{hor}}=\frac{a\left( kT_{hor}\right) H(kT_{hor})}{%
1037: a_{Pns0}H_{Pns0}}\sim 7.2\times 10^{-10}. \label{C5_44}
1038: \end{equation}%
1039: This is done by using Eqs.(\ref{C1_11}) and (\ref{Rtwenty fiveA}). Indeed,
1040: for all the values of $\kappa $, the $kT_{hor}$\ satisfying (\ref{C5_44})
1041: respects $\left( kT_{c}-kT_{hor}\right) \simeq 10^{-35}$.
1042:
1043: Analysis of the Hubble radius shows that, from the beginning of the
1044: deceleration ($kT_{accel}\simeq 0.2356$ $GeV$) till nowadays, $R$ has
1045: increased $1.75\times 10^{10}$. But from $kT_{hor}$ to $kT_{accel}$ the
1046: Hubble radius diminished $1.8\times 10^{10}$. And for values of time smaller
1047: than that corresponding to $kT_{hor}$\ all the universe was in causal
1048: contact, allowing the thermalization of the cosmic fluid.
1049:
1050: The \textit{flatness problem} may be treated quantitatively through the
1051: first Friedmann equation%
1052: \begin{equation}
1053: 1-\Omega _{T}=-\frac{\kappa c^{2}}{a^{2}H^{2}}=\Omega _{\kappa },\text{\ \ \
1054: \ \ }\Omega _{T}\equiv \frac{\rho }{\rho _{c}}\equiv \frac{8\pi G\rho }{%
1055: 3c^{2}H^{2}}. \label{C1_47}
1056: \end{equation}%
1057: According to recent data $\Omega _{T0}\simeq 1.003$. Now, $aH$ decreases
1058: with time if the universe is of the radiation-dominated or matter-dominated
1059: types (the usual cases). This means that $\left\vert \Omega _{\kappa
1060: }\right\vert $\ increases with time in radiation and matter-dominated
1061: models. If this is so, in the early universe $\left\vert \Omega _{\kappa
1062: }\right\vert \ll $ $10^{-3}$ and we are led to the question: Why was the
1063: total density value $\rho $ so close to the\ critical density $\rho _{c}$ in
1064: the initial instants of the universe when $\kappa =\pm 1$? This necessary
1065: fine-tuning in the initial condition for $\rho $ is the flatness problem.
1066:
1067: The standard model establishes that the energy density at the end of the
1068: pre-nucleosynthesis period $\rho _{Pns0}$ is close to $\rho _{c}$\ within a
1069: precision of $21$ decimal digits. Hence, to solve the flatness problem one
1070: needs to show that there is a value $kT_{flat}<kT_{c}$ such that%
1071: \begin{equation}
1072: \left\vert \Omega _{\kappa }\right\vert _{flat}=\frac{1.5\times
1073: 10^{-21}a_{Pns0}^{2}H_{Pns0}^{2}}{a^{2}(kT_{flat})H^{2}(kT_{flat})}\sim
1074: 10^{-3}. \label{C5_42}
1075: \end{equation}%
1076: The value $kT_{flat}$\ verifying (\ref{C5_42}) is found to be $\left(
1077: kT_{c}-kT_{flat}\right) \simeq 10^{-35}$\ (for $\kappa =\pm 1$). From $%
1078: kT_{flat}$ to the end of the PNS, the scale factor increases $12$\ orders of
1079: magnitude. And this large variation happens in a time interval of $2.25$ $ms$%
1080: , by Eq. (\ref{Rtwenty six}). The duration of the accelerated expansion is
1081: much smaller: $0.26$ $ms$. This makes clear that the high inflationary rate
1082: -- $a(kT)$ increase of $10$ orders of magnitude in $0.26$ $ms$\ -- rapidly
1083: flattens the universe.
1084:
1085: The values (of the thermal energy, of the scale factor) characterizing the
1086: PNS period in what concerns the inflationary issues are collected in Table 1.
1087:
1088: \begin{equation*}
1089: \begin{tabular}{|c|c|c|c|c|c|}
1090: \hline
1091: & $\Delta t(ms)$ & $\frac{a}{a_{Pns0}}$ & $\frac{R}{R_{Pns0}}$ & $\left\vert
1092: \Omega _{\kappa }\right\vert $ & $k\left\vert T-T_{c}\right\vert (GeV)$ \\
1093: \hline
1094: $kT_{Pns0}$ & $0$ & $1$ & $1$ & $1.5\times 10^{-21}$ & $\simeq 0.2567$ \\
1095: \hline
1096: $kT_{accel}$ & $1.99$ & $5.0\times 10^{-2}$ & $8.0\times 10^{-2}$ & $%
1097: 9.55\times 10^{-24}$ & $\simeq 0.0411$ \\ \hline
1098: $kT_{hor}$ & $2.25$ & $1.9\times 10^{-12}$ & $1.44\times 10^{9}$ & $10^{-3}$
1099: & $10^{-35}$ \\ \hline
1100: $kT_{flat}$ & $2.25$ & $1.9\times 10^{-12}$ & $1.44\times 10^{9}$ & $10^{-3}$
1101: & $10^{-35}$ \\ \hline
1102: \end{tabular}%
1103: \end{equation*}
1104:
1105: Table 1: Resum\'{e} of the main results concerning the behavior of our model
1106: in the PNS period and the accelerated regime. $kT_{Pns0}=0.020$ $GeV$ is the
1107: temperature at the end of the PNS; $kT_{acel}$ is the temperature below
1108: which there is no acceleration; $kT_{hor}$ is the temperature at which the
1109: causality problem is solved; and $kT_{flat}$ is the temperature that rules
1110: out the fine-tuning in $\rho $. Definitions: $\Delta t=t_{Pns0}-t$ and $%
1111: a_{Pns0}\equiv 1$.
1112:
1113: \bigskip
1114:
1115: %%%%%%%%%%%%%%%%%%%%
1116:
1117: \section{Final remarks \label{sec-Final}}
1118:
1119: %%%%%%%%%%%%%%%%%%%%
1120:
1121: This work presents equations of state (EOS) for the early universe which
1122: include the interactions among the constituent particles. We argue that the
1123: dominant processes affecting the pre-nucleosyntesis period (PNS) are those
1124: involving the strong interaction, the only which has been considered. Total
1125: EOS accounting for photons, leptons and hadrons have been built, using a
1126: phenomenological description of the hadronic interaction and assuming
1127: thermodynamical equilibrium. Assuming the hypothesis of no deconfinement,
1128: some interesting points have turned up.
1129:
1130: First, the interactions naturally drove the pressure to negative values. The
1131: de Sitter model has taught us long ago that an exotic equation of state, $%
1132: p=-\rho $, in which the pressure is negative, would give rise to an
1133: accelerated expanding universe. Therefore, our results could explain
1134: the primeval acceleration regime, as an alternative scenario to
1135: (scalar field) inflation. Indeed, our model is able to connect this
1136: initial accelerated stage with a decelerated expansion of the type
1137: expected for a radiation-dominated universe. In addition, the
1138: acceleration generated in our interacting cosmic fluid has been
1139: shown to be intense enough to solve the horizon and flatness
1140: problems.
1141:
1142: Another noticeable feature has been obtained: for the models with $\kappa =0$
1143: or $\kappa =1$ the initial singularity is avoided, and a natural explanation
1144: for the expansion of the universe emerges. In fact, either in an eternal
1145: universe (compatible with $\kappa =1$) or in a universe with a beginning
1146: (corresponding to $\kappa =0$), the primeval acceleration produced by the
1147: strong interaction is capable of engendering an expansion which evolves to a
1148: decelerated type and continue its dynamics. Nevertheless, let us emphasize
1149: that these results have been derived for a symmetric universe: the quantity
1150: of matter is assumed to equal that of antimatter.
1151:
1152: Such effects are actually more general than the results of the specific
1153: model proposed. Indeed, in order for then to be valid it is enough to
1154: introduce in the FRW cosmology a continuous parameter $\lambda $\ (related
1155: or not to interactions) responsible for the passage from $p(\lambda )=\frac{%
1156: \rho (\lambda )}{3}$ to the exotic EOS\,\,\, $p(\lambda )=-\rho (\lambda )$
1157: in such a way that $\frac{d\rho }{d\lambda }>0$ at all times. This type of
1158: fitting links smoothly the accelerated and decelerated expansion periods via
1159: the parametric equation $p=\omega \rho $ in which $\omega $ varies between $%
1160: 1/3$ and $-1$. This interval automatically eliminates the ghost models($%
1161: \omega \leq -1$) \cite{ghost}.
1162:
1163: Besides the mechanisms engendering primeval acceleration, there are
1164: other important subjects related to the pre-nucleosynthesis period
1165: to be discussed. Amongst them, we highlight two: the generation of
1166: the initial perturbations of the energy density content and the
1167: matter-antimatter asymmetry. The initial perturbations might be
1168: generated\ in the context of our model through the introduction of
1169: statistical fluctuation processes \cite{mag}; this possibility shall
1170: be investigated in the future. The problem of the matter-antimatter
1171: asymmetry is more involved: as it is well known, the inflation wipes
1172: out all the traces of an eventual initial asymmetry and this demands
1173: the existence of a mechanism driving a post-inflationary
1174: matter-antimatter asymmetry. Moreover, a necessary condition to
1175: produce an excess of particles over anti-particles is the system to
1176: be out of the thermodynamical equilibrium \cite{Sak}. This exigence
1177: prevents our model to describe this asymmetry once all the numerical
1178: densities are obtained from the statistical mechanics in thermal
1179: equilibrium. This limitation could be overcome if we consider
1180: out-of-equilibrium phase-transition in a post-inflationary stage. A
1181: mechanism of this kind was proposed in Ref. \cite{ucrafour} and will
1182: the investigated in another opportunity.
1183:
1184: From a wider perspective, this works calls attention to the fact that
1185: particle interactions are a direct source of gravitation \cite{artigo 1}.
1186: When one ignores this truth, simplifications result in the cosmological
1187: models; but these are perhaps too large to enable a proper description of
1188: the real universe. The role of the fundamental interactions in cosmology
1189: opens new paths to the study of the universe evolution, and even if our
1190: complete EOS are not valid during all the PNS period, the central idea may
1191: be applied to model other eras. Up to this point, we would not dare to
1192: affirm that the inclusion of hadronic interactions is more appropriated than
1193: the scalar fields to generate the inflation. Many issues solved by the
1194: inflationary theory have been left untouched here. But the interaction
1195: scheme is certainly more suitable from the theoretical point of view: they
1196: have a clear physical interpretation.
1197:
1198: %========================================%
1199: %<<<<<<<<<<< ACKNOWLEDGMENTS >>>>>>>>>>>%
1200: %========================================%
1201:
1202: \section*{\protect\small Acknowledgements}
1203:
1204: L. G. M. is grateful to Funda\c{c}\~{a}o de Amparo \`{a} Pesquisa do
1205: Estado de S\~ao Paulo (FAPESP) and Funda\c{c}\~{a}o de Amparo \`{a}
1206: Pesquisa do Estado do Rio de Janeiro (FAPERJ), Brazil; R. A. and R.
1207: R. C. (grant 201375/2007-9) are thank to Conselho Nacional de
1208: Pesquisas
1209: (CNPq), Brazil. R. R. C. and L. G. M. also thank Instituto de F\'{\i}sica Te%
1210: \'{o}rica, Universidade Estadual Paulista, Brazil, where this work was
1211: initiated. Finally, R. R. C. would like to thank Prof. V. P. Frolov and Dr.
1212: A. Zelnikov for the kind hospitality extended him at University of Alberta.
1213:
1214: \appendix
1215: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
1216:
1217: \section{Multi-component systems \label{ap-Multicomp}}
1218:
1219: A multi-component system is a set with more than one type of particle or
1220: conserved quantum number. The grand canonical partition function $\Theta $
1221: for a $B$-components system is, in analogy to (\ref{s1}),%
1222: \begin{equation}
1223: \Theta (z_{1},..,z_{B},V,T)=\sum\limits_{N_{1}=0}^{\infty
1224: }...\sum\limits_{N_{B}=0}^{\infty
1225: }Q_{N_{1},...,N_{B}}(V,T)z_{1}^{N_{1}}...z_{B}^{N_{B}}. \label{D37}
1226: \end{equation}%
1227: (See Refs. \cite{Smith Lain 81,Osborn 77}.) The grand canonical potential $%
1228: \Omega $\ as a function of the fugacities (and the temperature) is, then,%
1229: \begin{equation}
1230: \Omega (z_{1},...,z_{B},T)=\frac{1}{V}\ln \Theta
1231: (z_{1},...,z_{B},V,T)=\sum\limits_{N_{1}=0}^{\infty
1232: }...\sum\limits_{N_{B}=0}^{\infty
1233: }b_{N_{1},...,N_{B}}z_{1}^{N_{1}}..z_{B}^{N_{B}}, \label{D39}
1234: \end{equation}%
1235: where $b_{N_{1},...,N_{B}}$ are the $B$-component cluster integrals, with $%
1236: b_{0,...,0}\equiv 0$ ($N_{1}=N_{2}=...=N_{B}=0$).
1237:
1238: The construction of the equations of state -- pressure $%
1239: p=p(z_{1},..,z_{B},T) $, energy density $\rho =\rho (z_{1},..,z_{B},T)$ and
1240: numerical densities $n_{1}(z_{1},..,z_{B},T)$, $n_{2}(z_{1},..,z_{B},T)$%
1241: ,..., $n_{B}(z_{1},..,z_{B},T)$ -- is done through the usual mapping from
1242: statistical mechanics to thermodynamics, namely
1243: \begin{subequations}
1244: \begin{align}
1245: \frac{p(z_{1},...,z_{B},T)}{kT}& \equiv \Omega
1246: =\sum\limits_{N_{1}=0}^{\infty }...\sum\limits_{N_{B}=0}^{\infty
1247: }b_{N_{1},...,N_{B}}z_{1}^{N_{1}}...z_{B}^{N_{B}}, \label{D41a} \\
1248: n_{i}(z_{1},...,z_{B},T)& \equiv z_{i}\left( \frac{\partial \Omega }{%
1249: \partial z_{i}}\right) _{V,T,\left\{ z_{B}\right\} \neq
1250: z_{i}}=\sum\limits_{N_{1}=0}^{\infty }...\sum\limits_{N_{B}=0}^{\infty
1251: }N_{i}b_{N_{1},...,N_{B}}z_{1}^{N_{1}}...z_{B}^{N_{B}}, \label{D41b} \\
1252: \rho (z_{1},...,z_{B},T)& \equiv kT^{2}\left( \frac{\partial \Omega }{%
1253: \partial T}\right) _{V,\left\{ z_{B}\right\} }=\left( kT\right)
1254: ^{2}\sum\limits_{N_{1}=0}^{\infty }...\sum\limits_{N_{B}=0}^{\infty }\frac{%
1255: \partial b_{N_{1},...,N_{B}}}{\partial \left( kT\right) }%
1256: z_{1}^{N_{1}}...z_{B}^{N_{B}}. \label{D41c}
1257: \end{align}%
1258: Symbol $\left\{ z_{B}\right\} $ is the set of fugacities, from $z_{1}$ to $%
1259: z_{B}$. As in the one-component case, these is another form for the EOS:
1260: that corresponding to the virial expansion, with the pressure and the energy
1261: density in terms of the numerical densities. It is:
1262: \end{subequations}
1263: \begin{subequations}
1264: \begin{gather}
1265: \frac{p(n_{1},...,n_{B},T)}{kT}=\sum\limits_{l_{1}=0}^{\infty
1266: }...\sum\limits_{l_{B}=0}^{\infty
1267: }a_{l_{1},...,l_{B}}(T)n_{1}^{l_{1}}...n_{B}^{l_{B}}, \label{D42a} \\
1268: \frac{\rho (n_{1},...,n_{B},T)}{\left( kT\right) ^{2}}=\sum%
1269: \limits_{l_{1}=0}^{\infty }...\sum\limits_{l_{B}=0}^{\infty
1270: }c_{l_{1},...,l_{B}}(T)n_{1}^{l_{1}}...n_{B}^{l_{B}}. \label{D42b}
1271: \end{gather}%
1272: The virial coefficients $a_{l_{1},...,l_{B}}$ and $c_{l_{1},...,l_{B}}$ are
1273: determined by the clusters integrals $b_{N_{1},...,N_{B}}$. For a
1274: 2-component case, the first virial coefficients are
1275: \end{subequations}
1276: \begin{subequations}
1277: \begin{align}
1278: & \left. a_{0,0}=0,\text{\ \ \ \ \ \ \ }a_{0,1}=a_{1,0}=1,\right.
1279: \label{ApenA1} \\
1280: & \left. a_{2,0}=\frac{-b_{2,0}}{b_{1,0}^{2}},\text{\ \ }a_{0,2}=\frac{%
1281: -b_{0,2}}{b_{0,1}^{2}},\text{\ \ }a_{1,1}=\frac{-b_{1,1}}{b_{1,0}b_{0,1}}%
1282: ,\right. \text{\ } \label{ApenA2}
1283: \end{align}%
1284: \end{subequations}
1285: \begin{subequations}
1286: \begin{align}
1287: & \left. c_{0,0}=0,\text{\ \ \ \ \ }c_{1,0}=\frac{\dot{b}_{1,0}}{b_{1,0}},%
1288: \text{\ \ }c_{0,1}=\frac{\dot{b}_{0,1}}{b_{0,1}},\right. \label{ApenA3} \\
1289: & \left. c_{2,0}=-\dot{a}_{2,0},\text{\ \ }c_{0,2}=-\dot{a}_{0,2},\text{\ \ }%
1290: c_{1,1}=-\dot{a}_{1,1}.\right. \label{ApenA4}
1291: \end{align}%
1292: The dot $^{\cdot }$\ indicates differentiation with respect to $kT$.
1293: Equation (\ref{Rnoventa dois A}) for a 2-component system reads:
1294: \end{subequations}
1295: \begin{subequations}
1296: \begin{eqnarray}
1297: b_{2,0}-b_{2,0}^{(0)} &=&\frac{1}{2\beta \pi ^{3}}\sum\limits_{a}\sum%
1298: \limits_{l=0}^{\infty }g_{2,0}^{a}\left( 2l+1\right)
1299: \int\limits_{2m_{X}}^{\infty }\omega ^{2}K_{2}(\beta \omega )\left[ \frac{%
1300: \partial \delta _{l}^{a}(\omega )}{\partial \omega }\right] _{2,0}d\omega ,
1301: \label{Aj8a} \\
1302: b_{0,2}-b_{0,2}^{(0)} &=&\frac{1}{2\beta \pi ^{3}}\sum\limits_{a}\sum%
1303: \limits_{l=0}^{\infty }g_{0,2}^{a}\left( 2l+1\right)
1304: \int\limits_{2m_{Y}}^{\infty }\omega ^{2}K_{2}(\beta \omega )\left[ \frac{%
1305: \partial \delta _{l}^{a}(\omega )}{\partial \omega }\right] _{0,2}d\omega ,
1306: \label{Aj8b} \\
1307: b_{1,1}-b_{1,1}^{(0)} &=&\frac{1}{2\beta \pi ^{3}}\sum\limits_{a}\sum%
1308: \limits_{l=0}^{\infty }g_{1,1}^{a}\left( 2l+1\right)
1309: \int\limits_{m_{X}+m_{Y}}^{\infty }\omega ^{2}K_{2}(\beta \omega )\left[
1310: \frac{\partial \delta _{l}^{a}(\omega )}{\partial \omega }\right]
1311: _{1,1}d\omega , \label{Aj8c}
1312: \end{eqnarray}%
1313: where $m_{X}$ and $m_{Y}$ are the masses of the first and second components.
1314: The ideal terms are determined from the free-system \cite{Ruben1} as:
1315: \end{subequations}
1316: \begin{subequations}
1317: \begin{align}
1318: b_{1,0}& =\frac{g_{X}}{\Lambda _{X}^{3}(\beta )},\text{\ \ \ \ }b_{0,1}=%
1319: \frac{g_{Y}}{\Lambda _{Y}^{3}(\beta )},\text{\ \ \ \ }b_{1,1}^{(0)}=0~;
1320: \label{Aj11} \\
1321: b_{2,0}^{(0)}& =\frac{(\pm 1)g_{X}}{2\Lambda _{X}^{3}(2\beta )},\text{\ \ }%
1322: b_{0,2}^{(0)}=\frac{(\pm 1)g_{Y}}{2\Lambda _{Y}^{3}(2\beta )}, \label{Aj12}
1323: \end{align}%
1324: where $g_{X}$ and $g_{Y}$ are the number of internal degrees of freedom of
1325: the free-systems; the upper (lower) signs refer to the Bose-Einstein
1326: (Fermi-Dirac) statistics; and, $\Lambda (j\beta )$ is the relativistic
1327: thermal wave-length\
1328: \end{subequations}
1329: \begin{equation}
1330: \Lambda _{X,Y}^{3}(j\beta )=\frac{\left( 2\pi \right) ^{3}j\beta }{4\pi
1331: m_{X,Y}^{2}K_{2}(j\beta m_{X,Y})}. \label{Aj13}
1332: \end{equation}%
1333: The generalization for more than two components is straightforward.
1334:
1335: %%%%%%%%%%%%%%%%%%%%%%%
1336:
1337: \section{Hadronic scattering\label{ap-Espalha}}
1338:
1339: %%%%%%%%%%%%%%%%%%%%%%%
1340:
1341: Elastic hadronic scattering processes can be described by the phase shifts $%
1342: \delta $ in the context of the partial--wave formalism. In this approach,
1343: one uses the spectroscopic classification and separates the
1344: \textquotedblleft elastic\textquotedblright\ process involving interactions
1345: among pions, kaons and nucleons in six types.\footnote{%
1346: The quotation marks were included to indicate that, besides the truly
1347: elastic processes $\left( \pi ^{-}+p\rightarrow \pi ^{-}+p\right) $, we are
1348: considering those with charge conjugation $\left( \pi ^{-}+p\rightarrow \pi
1349: ^{0}+n\right) $. The isospin symmetries $I$ are also taken into account.}
1350: They are: $\pi \pi $, $\pi K$, $\pi N$, $KK$, $KN$ and $NN$. There is a set
1351: of phase shifts $\delta $ for each of such processes. The $\delta $'s are
1352: directly or indirectly determined from the experimental data. We will
1353: discuss in detail how this is done in the case of the pion-pion scattering.
1354:
1355: The \textquotedblleft elastic\textquotedblright\ pion-pion scattering
1356: depends on the energy, the total orbital angular momentum $L$ and the total
1357: isospin $I$.\ It is then convenient to introduce the notation $\delta _{L,I}$%
1358: \ for each phase shift. As a scattering of identical particles with integral
1359: spin, the associated two-pions total state (orbital angular momentum state
1360: plus isospin state) must be symmetric. We will take $I=0$, $1$, $2$\ and
1361: restrict the analyses to $L=0$, $1$, $2$ (partial waves $S$, $P$ and $D$).
1362: This restriction is not arbitrary: it is imposed by the experimental data
1363: that are available. Hence, the relevant phase shifts are those in Table B.1.
1364:
1365: \begin{equation*}
1366: \begin{tabular}{|c|c|c|}
1367: \hline
1368: $S\text{-Wave}$ & $P\text{-wave}$ & $D\text{-wave}$ \\ \hline
1369: \begin{tabular}{c}
1370: $\delta _{0,0}\text{\ \ with\ \ }g=1$ \\
1371: $\delta _{0,2}\text{ \ with\ \ }g=5$%
1372: \end{tabular}
1373: & $\delta _{1,1}\text{ \ with\ \ }g=9$ &
1374: \begin{tabular}{c}
1375: $\delta _{2,0}\text{ \ with\ \ }g=5\text{ }$ \\
1376: $\delta _{2,2}\text{ \ with\ \ }g=25$%
1377: \end{tabular}
1378: \\ \hline
1379: \end{tabular}%
1380: \end{equation*}
1381: Table B.1: Phase shifts $\delta _{L,I}$ for $\pi \pi $-scattering. The
1382: degeneracy degree $g$ accounts for the bosonic symmetry and the projections
1383: of orbital angular momentum $\left( 2L+1\right) $ and of isospin $\left(
1384: 2I+1\right) $.
1385:
1386: \bigskip
1387:
1388: The data for the phase shifts $\delta _{0,0}$ and $\delta _{1,1}$ are found
1389: in Ref. \cite{EstaMartin,Rosselet}, which give the results for the
1390: extensively repeated and measured scattering $\pi ^{-}p\rightarrow \pi
1391: ^{-}\pi ^{+}n$. The data for $\delta _{0,2}$, $\delta _{2,0}$ and $\delta
1392: _{2,2}$ \cite{EstaMartin,FrogPeter} are determined through modeling based on
1393: the Roy equation \cite{Roy,Basdevant}. The $S$-wave data are shown in Figure %
1394: \ref{fig4}.
1395:
1396: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
1397: %%%%%%%%%% FIGURA %%%%%%%%%%
1398: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
1399: %\begin{tabular}{l}
1400: \begin{figure}[ht]
1401: \begin{center}
1402: \includegraphics[height=6cm, width=8cm]{Delta_00.eps} %
1403: \includegraphics[height=6cm, width=8cm]{Delta_02.eps}
1404: \end{center}
1405: \caption{Experimental data for the $S$-wave $\protect\pi \protect\pi $\
1406: scattering and the corresponding fitted curves for the functions $\protect%
1407: \delta _{L,I}\left( \protect\omega \right) $.}
1408: \label{fig4}
1409: \end{figure}
1410: %\end{tabular}
1411: %%%%%%%%%%%%%%%%%%%%%%%%%%%
1412:
1413: The fits for each one of the phase shifts\ were done using the software
1414: \textit{Origin }$6.1$ via the polynomial regression method or the non-linear
1415: least squares fitting. The experimental uncertainties were considered. The
1416: best-fit phase shifts are given by the expressions
1417: \begin{equation}
1418: \delta _{0,0}\left( \omega \right) =-1.13+4.36\omega -0.11\omega
1419: ^{2}-4.52\omega ^{3}+3.76\omega ^{4}, \label{ApF1}
1420: \end{equation}%
1421: and%
1422: \begin{equation}
1423: \delta _{0,2}\left( \omega \right) =-0.03+1.23\omega -6.67\omega
1424: ^{2}+12.62\omega ^{3}-13.39\omega ^{4}+8.57\omega ^{5}-3.11\omega
1425: ^{6}+0.48\omega ^{7}. \label{ApF2}
1426: \end{equation}%
1427: Analogously, we can obtain the fitted curves $\delta _{L,I}=\delta
1428: _{L,I}\left( \omega \right) $\ for partial waves $P$ and $D$. The functions $%
1429: \delta _{L,I}\left( \omega \right) $\ -- such as (\ref{ApF1}) and (\ref{ApF2}%
1430: ) -- are differentiated with respect to the energy and substituted in
1431: equations (\ref{Aj8a}-\ref{Aj8c}) for the cluster integrals $b$ which, in
1432: turn, are used to obtain the $p$ and $\rho $ equations of state for the PNS
1433: universe.
1434:
1435: The other scatterings ($\pi K$, $\pi N$, $KK$, $KN$ and $NN$)\ are treated
1436: in the same way. The experimental data used in the description of the
1437: pion-kaon scattering are found in Ref. \cite{EstaCarnegie} and they are good
1438: enough only to analyze the values $L=0$, $1$ -- $S$-waves and $P$-waves. In
1439: the case of the pion-nucleon scattering, the relevant reference is \cite{CNS}
1440: and we study $L=0$, $1$, $2$ ($S$, $P$ and $D$-waves). The kaon-kaon
1441: scattering data are (indirectly) obtained from \cite{FrogPeter,KamiLesnLois}
1442: and \cite{FurnLesn}\ using the separable potential formalism \cite%
1443: {ActaPolonia,KamiLesn}; these data refer solely to the $S$-wave $K\bar{K}$
1444: scattering (we suppose that the processes $K\bar{K}$ and $KK$\ are
1445: identical, which means that the processes are independent of charge
1446: conjugation $C$. The kaon-nucleon phase shifts (for $S$, $P$ and $D$-waves)
1447: are in Ref. \cite{CNS} -- once again we admit independence under $C$. The
1448: same Ref. \cite{CNS} presents the nucleon-nucleon data for $S$, $P$ and $D$%
1449: -waves and, in these cases, in addition to the $C$-independence, it is
1450: assumed independence on the isospin $I$: the proton-proton phase shifts are
1451: very similar to the neutron-neutron ones.
1452:
1453: \bigskip
1454:
1455: \begin{thebibliography}{99}
1456: \bibitem{nar} V. Narlikar, \textit{Introduction to Cosmology,} 2nd ed.,
1457: Cambridge University Press, 1993.
1458:
1459: \bibitem{wein} S. Weinberg, \textit{Gravitation and Cosmology. Principles
1460: and Aplications of the General Theory of Relativity}, John Wiley and Sons,
1461: New York, 1972.
1462:
1463: \bibitem{Solutions} R. Aldrovandi, R. R. Cuzinatto and L. G. Medeiros,
1464: Foundations of Physics \textbf{36}, 1736 (2006); [gr-qc/0508073].
1465:
1466: \bibitem{Dod} S. Dodelson, \textit{Modern Cosmology: Anisotropies and
1467: Inhomogeneities in the Universe}, Academic Press, 2003.
1468:
1469: \bibitem{kolb} E. W. Kolb and M. S. Turner, \textit{The Early Universe,}
1470: Perseus Books, 1994.\textit{\ }
1471:
1472: \bibitem{Wmap} D. N. Spergel et. al., Astrophys. J. Suppl. \textbf{170} 377
1473: (2007); [astro-ph/0603449].
1474:
1475: \bibitem{LargeScale} M. Colless et al., MNRAS, \textbf{328}, 1039 (2001).
1476:
1477: \bibitem{artigo 1} R. Aldrovandi, R. R. Cuzinatto, L. G. Medeiros, \textit{%
1478: Interacting Constituents in Cosmology}, to appear in \textit{Int. J. Mod.
1479: Phys. }\textbf{D}; [gr-qc/0705.1369].
1480:
1481: \bibitem{SuNo} D. Rapetti, S. W. Allen, M. A. Amin, R. D. Blandford, Mon.
1482: No. Ro. Ast. Soc. \textbf{375,} 1510 (2007); [astro-ph/0605683] -- S.
1483: Permutter et al, Nature \textbf{391}, 51 (1998); [astro-ph/ 9712212] -- A.
1484: G. Riess et al., The Astronomical Journal \textbf{116}, 1009 (1998);
1485: [astro-ph/9805201].
1486:
1487: \bibitem{Liddle} A. R. Liddle e D. H. Lyth, \textit{Cosmological Inflation
1488: and Large-Scale Structure, } Cambridge University Press, 2000.
1489:
1490: \bibitem{Guth} A. H. Guth, Phys. Rev. \textbf{D23}, 357 (1981).
1491:
1492: \bibitem{AlbreSten} A. Albrecht \& P. J. Steinhardt,\textit{\ } Phys. Rev.
1493: Lett. \textbf{48}, 1220 (1982).
1494:
1495: \bibitem{Linde} A. D. Linde, Phys. Lett. \textbf{B108}, 389 (1982).
1496:
1497: \bibitem{CalDaveStein} R. R. Caldwell, R. Dave \& P. J. Steinhardt, Phys.
1498: Rev. Lett. \textbf{80}, 1582 (1998); [astro-ph/9708069].
1499:
1500: \bibitem{Stein} P. J. Steinhardt, L. Wang \& I. Zlatev,\textit{\ }Phys. Rev.
1501: \textbf{D59}, 123504 (1999); [astro-ph/9812313].
1502:
1503: \bibitem{Zlatev} P. J. Steinhardt, L. Wang \& I. Zlatev,\textit{\ }Phys.
1504: Rev. Lett. \textbf{82}, 896 (1999); [astro-ph/9807002].
1505:
1506: \bibitem{FraRos} U. Fran\c{c}a \& R. Rosenfeld, JHEP \textbf{210}, 015
1507: (2002); [astro-ph/0206194].\textit{\ }
1508:
1509: \bibitem{PeRa} B. Ratra \& P. J. E. Peebles, Phys. Rev. \textbf{D37}, 3406
1510: (1988).
1511:
1512: \bibitem{Orfeu} M.C. Bento, O. Bertolami, A.A. Sen, Phys. Rev. \textbf{D66}
1513: 043507 (2002); [gr-qc/0202064] -- M.C. Bento, O. Bertolami, Anjan Ananda
1514: Sen; Phys. Rev. \textbf{D70} 083519 (2004); [astro-ph/0407239].
1515:
1516: \bibitem{Capozzielo} S. Capozziello, S. Carloni, A. Troisi, \textit{Recent
1517: Research Developments in Astronomy \& Astrophysics} - \textbf{RSP/AA/21}
1518: (2003); [astro-ph/0303041] -- S. Capozzielo et. al., Phys. Lett. \textbf{A299%
1519: }, 494 (2002).
1520:
1521: \bibitem{Kremer} G. M. Kremer, Gen. Rel. Grav. \textbf{36}, 1423 (2004);
1522: [gr-qc/0401060].
1523:
1524: \bibitem{Motta} A. Brookfield, C. van de Bruck, D.F. Mota, D.
1525: Tocchini-Valentini, Phys. Rev. Lett. \textbf{96}, 061301 (2006) and Phys.
1526: Rev. \textbf{D73}, 083515 (2006).
1527:
1528: \bibitem{Gabi} L. Amendola, G. C. Campos, R. Rosenfeld, Phys. Rev. \textbf{%
1529: D75}, 083506 (2007); [ astro-ph/0610806].
1530:
1531: \bibitem{NelsonPinto} N. Pinto-Neto, B. M. O. Fraga, to appear in General
1532: Relativity and Gravitation (2007); [gr-qc/0711.3602].
1533:
1534: \bibitem{Dashen} R. Dashen, S-K. Ma \& J. Bernstein, Phys. Rev. \textbf{187}%
1535: , 345 (1969). Errata -- Phys. Rev. \textbf{A6}, 851 (1972).
1536:
1537: \bibitem{Reichl} L. E. Reichl, \textit{A Modern Course in Statistical Physics%
1538: }, University of Texas Press, Austin, 1980.
1539:
1540: \bibitem{Pat} R.K. Pathria, \textit{Statistical Mechanics,} 2nd. ed.,
1541: Butterworth Heinemann, Oxford, 1996.
1542:
1543: \bibitem{Beth} E. Beth \& G. E. Uhlenbeck, Physica \textbf{3}, 729 (1936) e
1544: Physica \textbf{4}, 915 (1937).
1545:
1546: \bibitem{ucraone} A. I Bugrii \& A. A. Trushevskii, Astrophysics \textbf{13}%
1547: , 195 (1978).
1548:
1549: \bibitem{ucratwo} A. I Bugrii \& A. A. Trushevskii, Zh. Eksp. Teor. Fiz.,
1550: \textbf{73}, 3 (1977) -- In Russian.
1551:
1552: \bibitem{ucrathree} Y. U. Beletsky, A. I Bugrii \& A. A. Trushevskii, Z.
1553: Phys. \textbf{C10}, 317 (1981).
1554:
1555: \bibitem{ucrafour} Y. U. Beletsky, A. I Bugrii \& A. A. Trushevskii,
1556: Astrophysics \textbf{24}, 107 (1986).
1557:
1558: \bibitem{AlBelPe} R. Aldrovandi, J. P. Beltr\'{a}n Almeida and J. G.
1559: Pereira, Int. J. Mod. Phys. \textbf{D13} 2241 (2004); [gr-qc/0405104].
1560:
1561: \bibitem{mayer} J. E. Mayer, J. Chem. Phys. \textbf{5}, 67; J. E. Mayer \&
1562: P. G. Ackermann, J. Chem. Phys. \textbf{5}, 74 (1937).
1563:
1564: \bibitem{LeeYang} T. D. Lee \& C. N. Yang, Phys. Rev. \textbf{113} 1165;
1565: \textbf{116}, 25 (1959).
1566:
1567: %\bibitem{Mezbacher} E. Merzbacher, \textit{Quantum Mechanics}, 3rd. ed.,
1568: %John Wiley and Sons, New York, 1998.
1569:
1570: \bibitem{Grif} D. Griffiths, \textit{Introduction to Elementary Particles},
1571: John Wiley \& Sons, New York, 1987.
1572:
1573: \bibitem{Hirshefeld} J. O. Hirschfelder, C. F. Curtiss \& R. B. Bird, \textit{%
1574: Molecular Theory of Gases and Liquids}, John Wiley and Sons, New
1575: York, 1954.
1576:
1577: \bibitem{Munzinger} P. Braun-Munzinger et al, \textit{Review for QGP3};
1578: [nucl-th/0304013].
1579:
1580: \bibitem{Andro} A. Andronic et al, Nucl. Phys. \textbf{A772}, 167 (2006);
1581: [nucl-th/0511071]
1582:
1583: \bibitem{tawfik2} A. Tawfik, Phys. Rev. \textbf{D71}, 54502 (2005);
1584: [hep-ph/0412336].
1585:
1586: \bibitem{Cheng} M. Cheng et al; [hep-lat/0710.0354].
1587:
1588: \bibitem{Yen} G. D. Yen et al, Phys. Rev. \textbf{C56}, 2210 (1997);
1589: [nucl-th/9711062].
1590:
1591: \bibitem{Hag} R. Hagedorn, Supp. Nuovo Cimento \textbf{3}, 147 (1965).
1592:
1593: \bibitem{Delta} W. Weinhold et al, Phys. Lett. \textbf{B433}, 236 (1998);
1594: [nucl-th/9710014].
1595:
1596: \bibitem{Callen} H.B. Callen, \textit{Thermodynamics and an Introduction to
1597: Thermostatistics}, 2d ed., John Wiley and Sons, New York, 1985.
1598:
1599: \bibitem{Lattice} F. Karsch, Nucl. Phys. \textbf{A698}, 199 (2002).
1600:
1601: \bibitem{RHIC} Relativistic Heavy Ion Collider, site -
1602: http://www.bnl.gov/rhic/.
1603:
1604: \bibitem{QGP} D. H. Rischke, Progress in Particle and Nucl. Phys. \textbf{52}%
1605: , 197 (2004); [nucl-th/0305030].
1606:
1607: \bibitem{CGC} E. Iancu \& R. Venugopalan,\textit{\ Review for QGP3};
1608: [hep-ph/0303204].
1609:
1610: %\bibitem{Tawfik} A. Tawfik, \textit{Black Box QGP}, (2006); [hep-ph/0604037].
1611:
1612: \bibitem{Vega} D. Boyanovsky et al, Ann. Rev. Nucl. Part. Sci. \textit{56},
1613: 441, (2006); [hep-ph/0602002].
1614:
1615: \bibitem{PDG} Particle Data Group, site - http://pdg.lbl.gov/.
1616:
1617: \bibitem{Boucing} M. Novello, \textit{V Brazilian School of Cosmology and
1618: Gravitation}, World Scientific, Cingapura (1987) -- M. Novello, \textit{VII
1619: Brazilian School of Cosmology and Gravitation}, Editions Fronti\`{e}res,
1620: Cingapura (1993) -- M. Novello, J. M. Salim, \textit{The stability of a
1621: bouncing universe}, (2003); [hep-th/0305254].
1622:
1623: \bibitem{ghost} Sean M. Carroll, Mark Hoffman \& Mark Trodden, Phys. Rev.
1624: \textbf{D68}, 023509 (2003); [astro-ph/0301273] -- Robert R.
1625: Caldwell, Marc Kamionkowski \& Nevin N. Weinberg, Phys. Rev. Lett.
1626: \textbf{91}, 071301(2003); [astro-ph/0302506] -- V. K. Onemli, R. P.
1627: Woodard, Phys. Rev. \textbf{D70}, 107301 (2004); [gr-qc/0406098].
1628:
1629: \bibitem{mag} J. Magueijo \& L. Pogosian, Phys. Rev. \textbf{D67},
1630: 043518 (2003); [astro-ph/0211337].
1631:
1632: \bibitem{Sak} A. D. Sakharov, Soviet Physics Journal of Experimental
1633: and Theoretical Physics (JETP) \textbf{5} 24 (1967).
1634:
1635: \bibitem{Smith Lain 81} C. R. Smith, R. Inguva \& K. D. Lain,\textit{\ }%
1636: Phys. Rev. \textbf{A23}, 3285 (1981).
1637:
1638: \bibitem{Osborn 77} T. A. Osborn, Phys. Rev. \textbf{A16}, 334 (1977).
1639:
1640: \bibitem{Ruben1} R. Aldrovandi, J. Gariel \& G. Marcilhacy, \textit{On the
1641: pre-nucleosynthesis cosmological period}, Rev. Ci\"{o}enc. Ex. Nat., \textbf{%
1642: 5} 133 (2003); [gr-qc/0203079].
1643:
1644: \bibitem{EstaMartin} P. Estabrooks \& A. D. \ Martin, Nucl. Phys. \textbf{B79%
1645: }, 301 (1974).
1646:
1647: \bibitem{Rosselet} L. Rosselet et al, Phys. Rev. \textbf{D15}, 574 (1977).
1648:
1649: \bibitem{FrogPeter} C. D. Froggatt \& J. L. Petersen, Nucl. Phys. \textbf{%
1650: B129}, 89 (1977).
1651:
1652: \bibitem{EstaCarnegie} P. Estabrooks et al, Nucl. Phys. \textbf{B133}, 490
1653: (1978).
1654:
1655: \bibitem{CNS} Center for Nuclear Studies (Said Program), site:
1656: http://gwdac.phys.gwu.edu/: online data.
1657:
1658: \bibitem{ActaPolonia} L. Le\'{s}niak, Acta Physica Polonica \textbf{B8},
1659: 1835 (1996).
1660:
1661: \bibitem{KamiLesn} R. Kami\'{n}ski \& L. Le\'{s}niak, Phys. Rev. \textbf{D50}%
1662: , 3145 (1994).
1663:
1664: \bibitem{KamiLesnLois} R. Kami\'{n}ski, L. Le\'{s}niak \& B. Loiseau, Phys.
1665: Lett. \textbf{B413}, 130 (1997).
1666:
1667: \bibitem{FurnLesn} A. Furman \& L. Le\'{s}niak, Phys. Lett. \textbf{B538},
1668: 266 (2002).\textbf{\ }
1669:
1670: \bibitem{Roy} S. M. Roy, Phys. Lett. \textbf{B36}, 353 (1971).
1671:
1672: \bibitem{Basdevant} J. L. Basdevant, C. D. Froggatt \& J. L. Petersen, Nucl.
1673: Phys. \textbf{B72}, 413 (1974).
1674: \end{thebibliography}
1675:
1676: \end{document}
1677: