1: \documentclass[prd,showpacs,preprint]{revtex4}
2:
3: \usepackage{amsmath}
4: \usepackage{amsfonts}
5: \usepackage{amssymb}
6: \usepackage{latexsym}
7: \usepackage[dvips]{graphicx}
8: \usepackage{graphics}
9: \usepackage{color}
10:
11:
12: \begin{document}
13:
14:
15: \title{Particle--Like Description in Quintessential Cosmology}
16: \author{Marek Szyd{\l}owski}
17: \email{uoszydlo@cyf-kr.edu.pl}
18: \affiliation{Astronomical Observatory, Jagiellonian University,\\ Orla 171, 30-244 Krak\'ow, Poland}
19: \author{Wojciech Czaja}
20: \email{czaja@oa.uj.edu.pl}
21: \affiliation{Astronomical Observatory, Jagiellonian University,\\ Orla 171, 30-244 Krak\'ow, Poland}
22: \date{\today}
23:
24: \begin{abstract}
25: Assuming equation of state for quintessential matter: $p = w(z)\rho$, we analyse dynamical behaviour
26: of the scale factor in FRW cosmologies. It is shown that its dynamics is formally equivalent to that of
27: a classical particle under the action of 1D potential $V(a)$. It is shown that Hamiltonian method can be
28: easily implemented to obtain a classification of all cosmological solutions in the phase space as well as
29: in the configurational space. Examples taken from modern cosmology \cite{kambenalc,corasaniti} illustrate
30: the effectiveness of the presented approach. Advantages of representing dynamics as a 1D Hamiltonian flow,
31: in the analysis of acceleration and horizon problems, are presented. The distant supernovae type Ia data
32: are used to reconstruct the expansion scenario. The inverse problem of reconstructing the Hamiltonian
33: dynamics (i.e. potential function) from the luminosity distance function $d_{L}(z)$ for supernovae is also
34: considered.
35: \end{abstract}
36:
37: \pacs{98.80.Bp, 98.80.Cq, 11.25.-w}
38:
39: \maketitle
40:
41: \section{Introduction}
42:
43: The Newtonian analogue to the Friedmann-Robertson-Walker models (hereafter FRW) was first
44: considered by Milne and McCrea who used classical mechanics concepts to describe the expansion of
45: the universe \cite{milne,mccrea}. This approach has some difficulties connected with description of
46: an infinite homogenous Newtonian cosmology \cite{harrison}. Pressure, in contrast to Newtonian theory,
47: has only relativistic character \cite{ellis}. In this way, a quintessential universe has
48: no Newtonian analogues. However, pressure seems to be play important role in the evolution of the real
49: universe. Recently found accelerated expansion of our universe is due to the presence of a certain vacuum
50: energy with the negative pressure equation of state \cite{perlmutter}-\cite{schmidt}. This type of energy is
51: called dark energy and equation of state, in the general quintessential form, is usually postulated in the
52: analysis of dynamics \cite{caldwell}.
53:
54: Lima et al. \cite{lima} considered the problem of reduction of FRW cosmologies with the equation of
55: state $p \propto \rho$ to the problem of classical particle under 1-dimensional homogenous potential.
56: We generalize this description for the FRW quintessential cosmologies.
57:
58: Recent observations of supernovae (SNIa) have revealed that this dark energy dominates the evolution
59: of the universe. The most natural candidate to represent dark energy seems to be the cosmological constant.
60: However, it is necessary to introduce a fine tunning of 120 orders of magnitude in order to obtain the
61: agreement with observations. Another popular idea is the so--called quintessence, a self-interacting scalar
62: field \cite{sahni}, but the quintessence program suffers from the fine tunning of microphysical parameters.
63: In this work, we discuss the possibility that the dark energy is characterized by the equation of state
64: \begin{equation}
65: p = w(a) \rho + 0 = w(a(z)) \rho + 0,
66: \label{eqstate}
67: \end{equation}
68: where $z + 1 = a^{-1}$ is the relation between redshift $z$ and the scale factor $a(t)$ in FRW models.
69:
70: The main aim of this paper is to reduce the FRW dynamics (with the equation of state in the form
71: (\ref{eqstate})) to the form of the Newtonian equation of motion for a unit mass particle in 1D potential.
72: Therefore, we extend classical results \cite{milne}-\cite{harrison} to the quintessential cosmologies.
73: On the other hand, the question concerning the reconstruction of the equation of motion (the same concerns
74: reconstruction form of equation of state \cite{sahni}) from the observational data becomes natural.
75: We consider problem of reconstructing the potential function for the particle--like description of dynamics.
76:
77: Organization of our paper is the following. In section 2, different methods of reducing the FRW
78: quintessential cosmology (Q.C.) to the 1D Hamiltonian flow is presented. Some applications of this formalism
79: are given in section 3. In section 2, we present simple analysis methods of classical cosmological problems
80: such as the horizon problem, acceleration or cosmological constant problems. Finally, in section 4 the
81: problem of reconstructing dynamics (potential function) from $d_{L}(z)$ in a model independent manner
82: (as far as possible) is investigated.
83:
84: \section{Hamiltonian dynamics of the Q.C. models}
85:
86: We consider dark energy description in terms of single $w(a(z))$ but it is always
87: possible to treat the equation of state of dark energy in terms of total pressure versus total density
88: ratio. Moreover, in any case our treatment can be generalized to the case of many-component of
89: non-interacting fluids.
90:
91: After assuming the quintessential fluid with the equation of state (\ref{eqstate}), where $p$ is the pressure
92: and $\rho$ the energy density, the dynamics of homogenous and isotropic models can be described by the
93: following set of equations
94: \begin{equation}
95: \frac{\ddot{a}}{a} = -\frac{1}{6}(1 + 3 w(a)) \rho + \frac{\Lambda}{3},
96: \label{dyn1}
97: \end{equation}
98: \begin{equation}
99: \frac{d \rho}{d t} = - 3 \frac{\dot{a}}{a}(1 + w(a)) \rho.
100: \label{dyn2}
101: \end{equation}
102: Equation (\ref{dyn1}) is the Raychaudhuri equation while eq. (\ref{dyn2}) is the continuity equation for
103: fluid with density $\rho$ and equation of state (\ref{eqstate}), $\dot{} \equiv \frac{d}{dt}$, $t$
104: -- cosmological time.
105:
106: The first integral of system (\ref{dyn1})-(\ref{dyn2}) is the Friedmann equation
107: \begin{equation}
108: \rho = 3 \frac{\dot{a}^{2}}{a^{2}} + 3 \frac{k}{a^{2}} - \Lambda,
109: \label{friedmann1}
110: \end{equation}
111: where $H = d(\ln{a})/dt$ is the Hubble function, $\Lambda$ is the cosmological constant and $k$ is the
112: curvature index.
113:
114: After substitution (\ref{friedmann1}) into (\ref{dyn1}) we obtain
115: \begin{equation}
116: \ddot{a} + \psi(a) \dot{a}^{2} + \kappa(a) = 0,
117: \label{dyn3}
118: \end{equation}
119: where
120: \begin{equation}
121: \psi(a) \equiv \frac{1 + 3 w(a)}{2a},
122: \label{psi}
123: \end{equation}
124: \begin{equation}
125: \kappa(a) \equiv \frac{1}{2}\Bigg[(1 + 3 w(a)) \frac{k}{a} - (1 + w(a)) \Lambda a\Bigg].
126: \label{kappa}
127: \end{equation}
128: There are two different methods of reducting (\ref{dyn3}) to the form of the Newtonian equation
129: of motion (i.e.\ of elimination of the $\psi(a)\dot{a}^{2}$ type term from (\ref{dyn3})).
130:
131: \noindent
132: {\bf 1.} The reparametrization of the time variable
133: \begin{equation}
134: t \rightarrow \tau:\hspace{10mm} dt = \varphi(a(\tau)) d\tau,
135: \label{reparam1}
136: \end{equation}
137: \begin{equation}
138: \varphi \equiv \exp \int_{}^{a}\psi(a)da
139: \label{reparam2}
140: \end{equation}
141: reduces basic equation (\ref{dyn3}) to
142: \begin{equation}
143: a'' \equiv \frac{d^{2}a}{d\tau^{2}} = -\frac{\partial V}{\partial a},
144: \label{reparam3}
145: \end{equation}
146: where
147: \begin{equation}
148: V(a) = \int_{}^{a}\kappa(a)\varphi^{2}(a)da
149: \label{reparam4}
150: \end{equation}
151: is the potential function for system (\ref{reparam3}). Finally, we obtain the dynamics reduced to the
152: Hamiltonian flow in the 1--dimensional configurational space.
153: \begin{equation}
154: \mathcal{H}(a,\dot{a}) = \frac{\dot{a}^{2}}{2} + V(a).
155: \label{reparam5}
156: \end{equation}
157: It can be checked that $\mathcal{H} = E = {\rm const}.$ is a constant of motion provided that energy density $\rho$
158: satisfies continuity condition (\ref{dyn2}). Therefore, trajectories of the system in the phase space
159: $(a,\dot{a})$ lie on the energy level $\mathcal{H} = E = {\rm const}.$ (in the case of the vacuum cosmology we have
160: $E = 0$). Of course, the Hamiltonian constraint should be consistent with the form of first integral
161: (\ref{friedmann1}). Hence, from (\ref{friedmann1}) we obtain
162: $$\frac{\dot{a}^{2}}{2} + V(a) = \frac{\dot{a}^{2}}{2} + \frac{k}{2} - \frac{\rho a^{2}}{6} - \frac{\Lambda a^{2}}{6},$$
163: i.e. the potential for the system (\ref{reparam5}) is
164: \begin{equation}
165: 2 V(a) = \varphi^{2}(a)\Bigg(k - \frac{\rho}{3}a^{2} - \frac{\Lambda a^{2}}{3}\Bigg) \equiv - \varphi^{2}(a) \frac{\rho_{{\rm eff}}}{3} a^{2}.
166: \label{reparam6}
167: \end{equation}
168: Now, the physical trajectories lie on the zero--energy level $\mathcal{H} = E = 0$ which coincides with the
169: form of the first integral, because now the curvature and $\Lambda$--terms are included into the effective
170: density $\rho_{{\rm eff}}$. Formally, the curvature term can be absorbed into the potential function by
171: postulating the curvature fluid for which $w_{k} = -1/3$, $\rho_{k} = -3k/a^{2}$ (as well as cosmological
172: constant term, where $p_{\Lambda} = - \rho_{\Lambda}$, $\rho_{\Lambda} = \Lambda$).
173:
174: From (\ref{reparam5}) and (\ref{reparam6}) we can see that both Hamiltonians
175: $$\mathcal{H} = \frac{1}{2}\Bigg(\frac{da}{dt}\Bigg)^{2} + V(q), \hspace{5mm} \bar{\mathcal{H}} = \frac{1}{2}\Bigg(\frac{da}{d \tau}\Bigg)^{2} + \varphi^{2}(a) V(q)$$
176: reproduce, in the consistent way the same Friedmann equations and they also reproduce equivalent equation of
177: motion, because the gauge freedom in choosing the lapse function (i.e. freedom in reparametrization
178: of time $t \rightarrow \tau$:~$dt = \varphi d\tau$: $\bar{\mathcal{H}} = \varphi^{2}(a) \mathcal{H}$).
179:
180: Let us consider special case of constant $w$. We have
181: $$\psi(a) = \frac{1+3w}{2}a^{-1},\hspace{10mm}\varphi = a^{\frac{1+3w}{2}},$$
182: $$V(a) = \Bigg(\frac{1}{2}k-\frac{\Lambda}{6}a^{2}+\frac{\rho_{0}}{a^{1+3w}}\Bigg)a^{1+3w},$$
183: $$\mathcal{H} \rightarrow \bar{\mathcal{H}} = a^{1+3w}\Bigg(\frac{1}{2}\dot{a}^{2}+V(a)\Bigg).$$
184: Classical equations of motion are
185: \begin{equation}
186: \dot{a} = \frac{\partial}{\partial \dot{a}}(N \mathcal{H}),\hspace{10mm} \ddot{a} = -\frac{\partial}{\partial a}(N \mathcal{H}),
187: \label{reparam7}
188: \end{equation}
189: where
190: $$N = a^{1+3w}.$$
191: The lapse function $N$ plays the role of a Lagrange multiplier and upon its variation we obtain Hamiltonian
192: constraint $\mathcal{H} = 0$.
193:
194: The cases of constant quintessential coefficient $w$ are important in applications. The forms of
195: potential function for different kinds of matter and flat model are presented in Table \ref{diffmatt}.
196: \begin{table}[!h]
197: \caption{The forms of potential function for different kinds of matter and flat model}
198: \begin{center}
199: \begin{ruledtabular}
200: \begin{tabular}{c|c|c|c|c|c|c|c}
201: &stiff mat.&rad.&dust&string&top. def.&$\Lambda$&phantom\\
202: \hline
203: $w$ &$1$&$1/3$&$0$&$-1/3$&$-2/3$&$-1$&$-4/3$\\
204: $V$&$\propto a^{-4}$&$\propto a^{-2}$&$\propto a^{-1}$&const.&$\propto a$&$\propto a^{2}$&$\propto a^{3}$\\
205: \end{tabular}
206: \end{ruledtabular}
207: \end{center}
208: \label{diffmatt}
209: \end{table}
210: Of course the presented formalism can be simply generalized to the case of any mixture of noninteracting
211: multifluids with the equation of state $p_{i} = w_{i}\rho_{i}$ \cite{szydlowski}. For our purpose it is
212: useful to put the dynamical equations into a new form by using dimensionless quantities
213: \begin{equation}
214: x \equiv \frac{a}{a_{0}},\hspace{5mm} T \equiv |H_{0}|t,\hspace{5mm}\Omega_{i,0} = \frac{\rho_{i,0}}{\rho_{{\rm cr},0}}
215: \label{reparam8}
216: \end{equation}
217: with
218: $$H = \frac{\dot{a}}{a},\hspace{5mm}\rho_{{\rm cr},0} = 3H_{0}^{2}.$$
219: The subscript $0$ means here that the quantity with this subscript is evaluated today (at time $t_{0}$).
220: Additionaly, we define $\Omega_{k} = -3k/6H_{0}^{2}$ and $\Omega_{\Lambda} = \Lambda/3H_{0}^{2}$.
221: Then the Hamiltonian is
222: \begin{equation}
223: \mathcal{H} = \frac{\dot{x}^{2}}{2} + V(x),
224: \label{reparam9}
225: \end{equation}
226: where
227: \begin{equation}
228: V(x) = - \frac{1}{2}\Omega_{k,0} - \frac{1}{2}\sum_{i}\Omega_{i,0}x^{1-3w_{i}}
229: \label{reparam10}
230: \end{equation}
231: which should be considered on the zero--energy level.
232:
233: The basic dynamical equations are then rewritten as
234: \begin{gather}
235: \frac{\dot{x}^{2}}{2} + V(x) = 0,\nonumber \\
236: \ddot{x} = \frac{1}{2}\sum_{i}\Omega_{i,0}(1-3w_{i})x^{-3w_{i}}.
237: \label{reparam11}
238: \end{gather}
239: In the general case, potential function can be obtained from $\rho_{{\rm eff}}$ i.e.
240: \begin{equation}
241: V(a) = - \frac{\rho_{{\rm eff}}(a)}{6}a^{2},
242: \label{reparam12}
243: \end{equation}
244: where for the quintessence matter we have
245: \begin{equation}
246: \begin{split}
247: \rho_{{\rm eff}} &= \rho_{\Lambda}+\rho_{k}+\rho_{0}a^{-3}\exp{\Bigg\{-3\int^{a}\frac{w(a)}{a}da\Bigg\}} \\
248: \rho_{{\rm eff}} &= \Lambda - 3k(1+z)^{2} \\
249: & \quad + \rho_{0}(1+z)^{3}\exp{\Bigg\{3\int^{a}\frac{w(a(z))}{1+z}dz\Bigg\}}.
250: \end{split}
251: \label{reparam13}
252: \end{equation}
253:
254: \noindent
255: {\bf 2.} In the second approach to the FRW dynamics representation by the Newtonian equation of motion, we
256: redefine the position variable $a$, namely we define X in the following way
257: \begin{equation}
258: a \rightarrow X \equiv a^{D(a)},
259: \label{redef1}
260: \end{equation}
261: where $D(a)$ is chosen in such a way that the term with $\dot{a}^{2}$ is absent in (\ref{dyn3}). Hence
262: we obtain new variable
263: \begin{equation}
264: X = \int^{a}\varphi(a)da,\hspace{5mm} D(a) = \log_{a}{\int^{a}\varphi(a)da}
265: \label{redef2}
266: \end{equation}
267: and the equation of motion takes the form
268: \begin{equation}
269: \ddot{X} = -\kappa(a(X))\varphi(a(X)) = -\frac{\partial V}{\partial X}.
270: \label{redef3}
271: \end{equation}
272: In the special case of constant $w$ we have
273: $$D(w) = \log_{a}{\int^{a}a^{\frac{1+3w}{2}}da} = \log_{a}{a^{\frac{3}{2}(1+w)}} = \frac{3}{2}(1+w),$$
274: and because of the obvious formula
275: \begin{equation}
276: V(a(X)) = \int^{a}\kappa(a)\varphi(a)dX,
277: \label{redef4}
278: \end{equation}
279: in the considered case, we obtain
280: \begin{equation}
281: \varphi = a^{\frac{1+3w}{2}},\hspace{10mm} X = a^{\frac{3}{2}(1+w)}.
282: \label{redef5}
283: \end{equation}
284: $$\mathcal{H}(X,\dot{X}) = \frac{\dot{X}^{2}}{2} + V(X) \equiv 0,$$
285: \begin{equation}
286: \begin{split}
287: V(X) & = \frac{3}{2}(1+w)\Bigg\{\frac{k}{2} - \frac{\Lambda}{6}X^{\frac{2}{D}} + \frac{\rho_{0}}{X^{\frac{1+3w}{D}}}\Bigg\}X^{\frac{1}{D}(1+3w)} \\
288: & = \frac{3}{2}(1+w) \Bigg\{\frac{k}{2}X^{\frac{2(D-1)}{D}} - \frac{\Lambda}{6}X^{2} + \rho_{0}\Bigg\}.
289: \end{split}
290: \label{redef6}
291: \end{equation}
292:
293: Let us observe that due to the new variable $X$, the considered system assumes the very simple form.
294: It can be now considered on the energy level $\mathcal{H} = E > 0$, $E = \frac{3}{2}(1+w)\rho_{0}$ and
295: $V(X) = \frac{k}{2}X^{2(1-\frac{1}{D})} - \frac{\Lambda}{6}X^{2}$. This system was analyzed on the phase
296: plane $(X,\dot{X})$ many years ago from the point of its structural stability. It is interesting that there
297: exists its natural generalization to the case of quintessential cosmology.
298:
299: Let us now emphasize some advantages of the considered approach. In general, the choice of
300: \begin{equation}
301: X \equiv \int^{a}\sqrt{a}\exp{\bigg(\frac{3}{2}\int^{a}\frac{w(a')}{a'}da'\bigg)}da
302: \label{redef7}
303: \end{equation}
304: makes it possible to reduce Q.C. to the nonlinear particle mechanics. There are many advantages of using
305: language of nonlinear mechanics. Let us consider some of them.
306:
307: \noindent
308: {\bf 1.}
309: Representation of dynamics as a one-dimensional Hamiltonian flow allows us to make the classification of
310: possible evolution paths in the phase space as well as in the configuration space. Then the discussion of the
311: existence and stability of critical points can be performed based on the geometry of the potential function.
312: It is so because the stability of critical points is determined by the Hessian
313: ($\partial^{2}\mathcal{H}/\partial x^{i}\partial y^{i}$). In our case the Hamiltonian function assumes the
314: very simple form characteristic for simple mechanical systems for which the Lagrangian function is natural,
315: i.e. quadratic in velocities $\mathcal{L} = \frac{1}{2}g_{\alpha\beta}\dot{q}^{\alpha}\dot{q}^{\beta}-V(q)$.
316: In our case, $g_{\alpha\beta} = {\rm const}.$ Then the characteristic equation for linearization of the
317: system is $\lambda^{2} + \rm{det}A = 0$, where $\lambda_{i}$ are eigenvalues of the linearization matrix $A$ of
318: the Hamiltonian system. Therefore, the only possible critical points in a finite domain of the phase space
319: are centres (${\rm det}A > 0$) or saddles (${\rm det}A < 0$). Of course, we can explore dynamics given by the
320: canonical equations. Denoting $x = a$, $y = \dot{a} = \dot{x}$ we obtain 2D dynamical system in the form
321: \begin{gather}
322: \dot{x} = \frac{\partial \mathcal{H}}{\partial y} = y, \nonumber \\
323: \dot{y} = - \frac{\partial \mathcal{H}}{\partial x} = - \frac{\partial V}{\partial x}.
324: \label{2ddyn}
325: \end{gather}
326:
327: We can observe that trajectories are integrable in quadratures. Namely, from the Hamiltonian constraint
328: $\mathcal{H} = E = 0$ we obtain the integral
329: \begin{equation}
330: t - t_{0} = \int_{a_{0}}^{a}\frac{da}{\sqrt{-2V(a)}},
331: \label{2dint}
332: \end{equation}
333: and for some specific forms of the potential function we can obtain the exact solutions.
334:
335: \noindent
336: {\bf 2.}
337: It is possible to make the classification of qualitative evolution paths by analyzing the characteristic
338: curve which represents the boundary equation in the configuration space. For this purpose we consider the
339: equation of the zero velocity, $\dot{a} = 0$, which represents the boundary of the domain admissible for
340: motion $\mathcal{D}_{E} = \{a\in \mathbb{R}_{+}: 2(E-V) \geqslant 0 \}$. By considering the boundary of
341: $\mathcal{D}_{E}$ given by the condition
342: $$\partial \mathcal{D}_{E} = \{a: V(a) = 0\},$$
343: full qualitative classification of evolutional paths can be performed.
344:
345: \noindent
346: {\bf 3.}
347: We can find the domains of cosmic acceleration as well as the domains for which the horizon problem is
348: solved. Because $\ddot{a} = -dV/da$, one can easily see that acceleration of the universe takes place if
349: $V(a)$ is a decreasing function of its argument. The condition $\partial V/\partial a|_{a=a_{0}}$ determines
350: the static critical point on the phase plane.
351:
352: Another interesting question concerns the horizon problem. It is easy to prove the following criterion
353: of avoiding this problem. The FRW cosmological model does not have an event horizon near the singularity
354: if $\dot{a}(t)c^{-1}$ tends to a constant while $a(t)$ tends to zero \cite{weinberg}.\\
355:
356: When all events whose coordinates in the past are located beyond some distance $d_{H}$ then can never
357: communicate with the observer at the coordinate $r = 0$ (in R-W metric). We can define the distance
358: $d_{H}$ as the past event horizon distance
359: $$d_{H}(t) = a(t) \int_{t_{0}}^{t}\frac{dt'}{a(t')}c = a(t)I.$$
360: Of course, whenever $I$ diverges as $t_{0} \rightarrow 0$, there is no past event horizon in the spacetime
361: geometry \cite{weinberg}. Then it is in principle possible to receive signals from sufficiently early
362: universe from any comoving particle like a typical galaxy. It the $t'$ integral converges for
363: $t_{0} \rightarrow 0$ then our communication with observer at $r=0$ is limited by what Rindler has called a
364: particle horizon.
365: The particle horizon will be present if energy density is growing faster then $a^{-2-\epsilon}$ as
366: $a \rightarrow 0$ ($\epsilon \geq 0$)\cite{weinberg}. Therefore if $\dot{a}>a^{-\epsilon/2}$ ($\dot{a}>A$)
367: then there is a particle horizon in the past.
368:
369: Let $c^{-1}\dot{a} < A$ ($c={\rm const}.$ is assumed but the corresponding theorem can be established for
370: the case of variable c(t) \cite{szydlowski}).
371: Then $I \geqslant \frac{1}{A}\int_{0}^{a_{0}}\frac{da}{a} = \frac{1}{A}(\ln{a_{0}}+\infty)$. Therefore,
372: $I$ diverges as $a \rightarrow 0$ and there exists no past horizon if the velocity of expansion factor
373: $\dot{a} \leqslant A$ is bounded. Our investigation of the particle horizon is independent
374: of any specific assumption about the behaviour of $a(t)$ near the singularity or of specific form of the
375: equation of state. If one assumes a linear equation of state $p = w \rho$ and $w = {\rm const}.$, then
376: Friedmann's equations imply the following behaviour for $a(t) \backsimeq (t)^{\frac{2}{3(\gamma+1)}}$ near
377: the singularity $t = 0$. The~integral $I$ would thus diverge only if $\gamma < -1/3$, i.e. only if the
378: pressure $p$ becomes negative. This is the condition for solving the horizon problem and it is identical to
379: that for the solution of the flatness problem ($8 \pi G \rho/3$ term will dominate the curvature term in a
380: long time evolution). We can also show here that the integral $\int \frac{dt}{a(t)}$ would diverge only if
381: the pressure of the cosmological fluid takes negative values in the general case of $w(a(t))$.
382: Conservation condition can be rewritten in the form
383: $$\frac{dp}{dt} = \frac{d}{dt}[\rho a^{3}(1+w(a))].$$
384: We can werify that boundedness of $\dot{a}(t)$ means also that $\rho a^{2}$ remains bounded near the
385: singularity (see FRW eq.). Therefore, $\rho a^{3} \rightarrow 0$ as $t \rightarrow 0$. By integrating both
386: sides of the equation written above from $0$ to $t$ we obtain
387: $$-3\int_{0}^{t}pa^{2}\dot{a}dt = a^{3}\rho(t) \geqslant 0.$$
388: Consequently, $w(a)$ must assume negative value without any specific assumption about the equation of
389: state.
390:
391: The above criterion can be now formulated in the language of the phase space variables.
392: If $V(a) \leqslant {\rm const}.$ (it does not diverge at singularity) as $a \rightarrow 0$ then there exists
393: no past horizon of the particle.
394:
395: \noindent
396: {\bf 4.}
397: Another natural generalization of the presented formalism consists in including anisotropy in simple Bianchi
398: models (Bianchi I or Bianchi V models) or VSL models with variable velocity of light of the
399: Albrecht--Maguejo--Barrow type.
400:
401: In the case of model BI ($k = 0$) or BV ($k = -1$), basic equation assumes the generalized form
402: \begin{equation}
403: \frac{\ddot{a}}{a} = -\frac{2}{3}\sigma^{2} - \frac{1}{6}(1 + 3w(a))\rho + \frac{\Lambda c^{2}}{3},
404: \label{anis1}
405: \end{equation}
406: where $c(a) = c_{0}a^{n}$ (in the AMB parametrization).
407:
408: Equation (\ref{anis1}) has generalized first integral
409: \begin{equation}
410: \Bigg(\frac{\dot{a}}{a}\Bigg)^{2} = \frac{\rho}{3} - \frac{kc^{2}}{a^{2}} + \frac{\sigma^{2}}{3} + \frac{\Lambda c^{2}}{3},
411: \label{anisint}
412: \end{equation}
413: where $\sigma^{2} = \frac{1}{2}\sigma_{ab}\sigma^{ab}$ is the shear scalar which can be used to reduce
414: dynamical equations to form (\ref{dyn3}), but now $\psi(a)$ preserves its form whereas
415: $\kappa(a)~\rightarrow~\bar{\kappa}(a)$ takes the new one which builds the new potential function.
416: \begin{equation}
417: \begin{split}
418: \bar{\kappa}(a) &= \frac{c^{2}(a)}{2}\Bigg[(1+3w(a))\frac{k}{a} - (1+w(a))\Lambda a \Bigg] \\
419: & \quad + \frac{\sigma^{2} a}{2}(1-w(a)), \\
420: V(a) & \rightarrow \bar{V}(a) = \int^{a}\bar{\kappa}(a)\varphi^{2}(a)da.
421: \end{split}
422: \label{anis2}
423: \end{equation}
424:
425: \noindent
426: {\bf 5.}
427: Let us now consider a class of FRW cosmologies with dissipation in the form of bulk viscosity. Owing to
428: the assumed spacetime FRW symmetries the shear viscosity vanishes, and we deal only with bulk viscosity
429: dissipation. It was given by Weinberg who classified the physical significance of bulk viscosity effects
430: within the framework of general relativity.
431:
432: The presence of bulk viscosity is equivalent to introducing effective pressure
433: \begin{equation}
434: p_{{\rm eff}} = w(a(z))\rho - 3 \xi H,
435: \label{viscos1}
436: \end{equation}
437: where $\xi$ is the bulk viscosity coefficient (constant for the sake of simplicity).
438:
439: If we introduce new variable $X$ following previously described scheme, we obtain the generalization of
440: (\ref{redef3}) in the form
441: \begin{equation}
442: \ddot{X} - \alpha \dot{X} + \frac{\partial V}{\partial X} = 0.
443: \label{viscos2}
444: \end{equation}
445: Equation (\ref{viscos2}) has the form of nonlinear oscillator equation with the damping force
446: $\propto \dot{X}$ and the exciting force $F(X) = -\frac{\partial V}{\partial X}$. In the following, this
447: system will be analyzed qualitatively by using methods of dynamical system. Therefore, on the basis of
448: the corresponding reduction, the particle--like description may be given if effects of bulk viscosity are
449: included.
450:
451: \noindent
452: {\bf 6.}
453: If we specialize the standard situation in cosmology with $\Lambda = 0$ and fluid which satisfy weak energy
454: condition $p + \rho \geqslant 0$ in which we set $p = w\rho$ and $w = {\rm const}.$ then the conservation
455: equation gives $\rho \propto a^{-3(1+w)}$ and the $\rho/3$ term will dominate the curvature therm $ka^{-2}$
456: at large $a$ so long as the matter stress violate strong energy condition $\rho + 3p < 0$ and obey
457: $p + \rho \geqslant 0$, that is if $-1 \leqslant w \leqslant -1/3$. This is what we shall mean by the
458: flatness problem. The scale factor then evolves as $a(t) \propto t^{\frac{2}{3(w+1)}}$ if $w > -1$ or
459: $\exp{(H_{0}t)}$, $H_{0}$ constant if $w = -1$.
460:
461: In the general case of variable $w(a(z))$ quintessential model will provide a solution of flatness problem
462: if zero curvature solution is representing late time attractor in the phase plane. This is possible in terms
463: of potential function if $\dot{a} > {\rm const}.$ as $a >> 1$ or
464: \begin{equation}
465: -C^{2}/2 < V < 0
466: \label{solvflat}
467: \end{equation}
468: This is the condition oposite to that for the solution of horizon problem $V < -C^{2}/2$ but it is formulate
469: rather for large $a$ than for small $a$.
470: Let us notice that if $\ddot{a} > 0$ as $a >> 1$ then $\dot{a} > Ct$ and period of accelerated expansion in
471: long term behaviour can solve the flatness problem.
472:
473: Let us consider the case of non-zero cosmological constant added to the potential function. Then in order
474: to explain why it does not totally dominate matter term $\rho/3$ at large $a$ we would need the presence of
475: fluid with subnegative pressure, i.e. with $w < -1$ such that a weak energy condition is violated
476: $p + \rho < 0$. This is what we shall mean by the cosmological constant problem. If we assume that $\rho$
477: is decreasing function of scale factor the solution of cosmological constant problem leads to contradiction
478: $\dot{a} < 0$.
479: In the general case we have following condition for solving $\Lambda$-problem
480: \begin{equation}
481: \rho_{{\rm eff}} > \Lambda \Leftrightarrow \frac{d}{da}\Big(\frac{V}{a^{2}}\Big) > 0 \Leftrightarrow
482: |V| > \Lambda a^{-2}
483: \label{solvlambda}
484: \end{equation}
485: i.e. modulus of potential function is growing faster than $a^{-2}$ at large $a$.
486:
487: Let us notice that for $w = {\rm const}.$ above condition leads to the identical condition to that to solve
488: the horizon problem $-1 \leqslant w \leqslant -1/3$. It is consequence of the fact that corresponding
489: conditions are valid at different domains $a \rightarrow \infty$ and $a \rightarrow 0$.
490:
491:
492:
493:
494: \section{Applications}
495:
496: We assume now that quintessential coefficient $w(a(z)) \equiv p/\rho$ is the analytical function
497: of~$z$
498: \begin{equation}
499: w(a(z)) = w(0) + \frac{dw}{dz}\Bigg|_{0}z + \frac{1}{2}\frac{d^{2}w}{dz^{2}}\Bigg|_{0}z^{2} + \dotso
500: \label{wrozw}
501: \end{equation}
502: Let $\gamma_{0} = w(0)$ and
503: $$\gamma_{1} = \frac{dw}{dz}\Bigg|_{0},\hspace{5mm} \gamma_{2} = \frac{1}{2}\frac{d^{2}w}{dz^{2}}\Bigg|_{0},
504: \hspace{2mm} \dotso, \hspace{2mm} \gamma_{i} = \frac{1}{i!}\frac{d^{i}w}{dz^{i}}\Bigg|_{0}.$$
505: We consider models with vanishing $\Lambda$--term because conception of quintessence is treated as an
506: alternative to the cosmological constant. The potential function of the dynamical system describing the FRW
507: model with equation of state (\ref{wrozw}) is given by
508: \begin{equation}
509: \begin{split}
510: V & = \int^{a}\kappa(a)\varphi^{2}(a)da = \frac{k}{2}a\exp{\Bigg\{3 \int^{a}\frac{w(a)}{a}da\Bigg\}} \\
511: & = \frac{k}{2}(1+z)^{-1}\exp{\Bigg\{(-3) \int^{z}\frac{w(a(z))}{(1+z)}dz\Bigg\}}.
512: \end{split}
513: \label{potqc}
514: \end{equation}
515: For example, in the case of mixture of noninteracting dust and fluid for which
516: $p_{x} = w_{x}\rho_{x}$, where $w_{x} = {\rm const}.$, we have
517: \begin{equation}
518: V(a) = \frac{k}{2}a\Bigg(\frac{a^{3w_{x}}}{\frac{\rho_{0m}}{\rho_{0x}}a^{3w_{x}} + 1}\Bigg),
519: \label{potqc1}
520: \end{equation}
521: here we consider quintessential matter for which
522: $$\rho_{x} = \rho_{0x} a^{-3(1+w_{x})}, \hspace{5mm} \rho_{0x} = {\rm const}.$$
523: and dust matter for which
524: $$\rho_{m} = \rho_{0m} a^{-3}, \hspace{5mm} \rho_{0m} = {\rm const}.$$
525: If for quintessential fluid $w_{x} = w_{x}(a)$, then the potential function takes the more general form
526: \begin{equation}
527: V(a) = \frac{k}{2}a\Bigg[\frac{\exp{\Big(3\int^{a}\frac{w_{x}(a)}{a}da}\Big)} {\frac{\rho_{0m}}{\rho_{0x}}
528: \exp{\Big(3\int^{a}\frac{w_{x}(a)}{a}da}\Big) + 1}\Bigg].
529: \label{potqc2}
530: \end{equation}
531:
532:
533: After substituting (\ref{wrozw}) into (\ref{potqc}) we obtain the general form of the potential function
534: which can be useful for a classification of cosmological models in the configurational space
535: \begin{equation}
536: \begin{split}
537: V(a(z)) &= \frac{k}{2}(1+z)^{-(1+3\gamma_{0})} \\
538: &\times \exp{\Bigg\{(-3)\int^{z}\frac{(\gamma_{1}z + \gamma_{2}z^{2} + \dotso)}{(1+z)}dz\Bigg\}}.
539: \end{split}
540: \label{potqc4}
541: \end{equation}
542: Let us consider a few special cases of (\ref{potqc4}).
543:
544: \noindent
545: {\bf A.}
546:
547: Let $\gamma_{0} = -\frac{1}{3}$ (like for a string) which determines the boundary the strong
548: energy condition violation. Additionaly, we consider expansion of $w(z)$ up to the second order term. Then
549: we obtain from the integral
550: \begin{equation}
551: \int \frac{\gamma_{1}z + \gamma_{2}z^{2}}{1+z}dz = (\gamma_{1}-\gamma_{2})z + \frac{\gamma_{2}}{2}z^{2}
552: - (\gamma_{1}-\gamma_{2})\ln{(1+z)}
553: \label{potqc5}
554: \end{equation}
555: the exact form of the potential function
556: \begin{equation}
557: V(a(z)) = \frac{k}{2} e^{-3(\gamma_{1}-\gamma_{2})z - \frac{3}{2}\gamma_{2}z^{2}}(1+z)^{3(\gamma_{1}-\gamma_{2})},
558: \label{potqc6}
559: \end{equation}
560: or for any $\gamma_{0}$
561: \begin{equation}
562: V(a(z))=\frac{k}{2}e^{-3(\gamma_{1}-\gamma_{2})z-\frac{3}{2}\gamma_{2}z^{2}}(1+z)^{-(1+3\gamma_{0})+3(\gamma_{1}-\gamma_{2})}.
563: \label{potqc7}
564: \end{equation}
565: \begin{figure}[!ht]
566: \begin{center}
567: \includegraphics[scale=0.6, angle=0]{1.eps}
568: \caption{Diagram of V(z) (formula (\ref{potqc7})) for the case of $k = \pm 1$,
569: $\gamma_{0} = -1/3$, $\gamma_{1} = -2/3$ and various $\gamma_{2}$
570: ($\gamma_{2}=-100,-10,-1,-0.1,0.1,1,10,100$)}
571: \label{wyk1}
572: \end{center}
573: \end{figure}
574:
575: Of course, the trajectories of the considered system lie in the domain admissible for motion defined as
576: $$\mathcal{D}_{E} = \{a \in \mathbb{R}_{+}: 2(E-V(a)) \geqslant 0, E = \frac{\rho_{0}}{6}\},$$
577: $$\rho = \rho_{0} a^{-3} \exp{\Bigg(3 \int \frac{w(a)}{a}da \Bigg).}$$
578: Therefore, equation of the boundary $\partial \mathcal{D}_{E}$ is
579: $$V=E \Leftrightarrow \ \frac{k}{2}e^{-3(\gamma_{1}-\gamma_{2})z-\frac{3}{2}\gamma_{2}z^{2}}
580: = E (1+z)^{(1+3\gamma_{0})-3(\gamma_{1}-\gamma_{2})}.$$
581: If we substitute $\gamma_{0} = -\frac{1}{3}$, then we obtain the potential function expressed in terms of
582: a single parameter $\gamma_{1}/\gamma_{2} \equiv x$, namely
583: $$\Bigg(\frac{1+z}{e^{z}}\Bigg)^{(x-1)} = \Bigg(\frac{2E}{k}\Bigg)^{\frac{1}{3\gamma_{2}}}e^{z^{2}/2}
584: = \bar{E}e^{z^{2}/2}.$$
585: Finally, we obtain
586: \begin{equation}
587: x(z) = 1 + \frac{\ln{\bar{E}} + z^{2}/2}{\ln{(1+z)}-z},
588: \label{potqc8}
589: \end{equation}
590: where $\bar{E} = (2E/k)^{1/3\gamma_{2}}$.\\
591: \begin{figure}[!ht]
592: \begin{center}
593: \includegraphics[scale=0.6, angle=0]{2.eps}
594: \caption{Diagram of dependence $x(z)$ for classification of models with potential (\ref{potqc7})
595: ($\bar{E}~=~0.1,0.6,1,2,10$)}
596: \label{wyk2}
597: \end{center}
598: \end{figure}
599: The plot of $x(z)$ for different $\bar{E}$ is shown in Fig. \ref{wyk2}.
600:
601: Let us note that if $\rho_{0} > 0$
602: ($E > 0$), then there is no boundary for $k = -1$, whereas if $\rho_{0} < 0$ ($E < 0$), there is
603: no boundary for $k = +1$ of the domain admissible for motion. Finally, we consider the evolution path
604: as a level of $x = {\rm const}.$ and then we classify all evolutions modulo the quantitative properties of
605: their dynamics \cite{robertson}.
606:
607: \noindent
608: {\bf B.}
609:
610: $\gamma_{2} = 0$, $\gamma_{0} \neq 0$, i.e. we consider quintessential model with pressure
611: $p~=~\gamma_{0}\rho + (\gamma_{1}z)\rho$. Then we obtain the potential in the form (Fig.\ref{wyk3})
612: \begin{equation}
613: V(a(z)) = \frac{k}{2}e^{-3\gamma_{1}z}(1+z)^{-(1+3\gamma_{0})+3\gamma_{1}}.
614: \label{potqc9}
615: \end{equation}
616: \begin{figure}[!ht]
617: \begin{center}
618: \includegraphics[scale=0.6, angle=0]{3.eps}
619: \caption{Diagram of the potential function $V(z)$ for the case {\bf B},
620: $p~=~(\gamma_{0}+\gamma_{1}z)\rho$, $\gamma_{0} = -1/3$, $\gamma_{1} = -2/3\beta$
621: ($\beta = -10,-1,0.1,0.4,1,10$), $k = \pm 1$}
622: \label{wyk3}
623: \end{center}
624: \end{figure}
625: Let us consider the boundary of the configuration space given by $V(z)$. Then $\gamma_{0}$ can be expressed
626: as a function of $z$ in the following way
627: \begin{equation}
628: \gamma_{0}(z) = (-\frac{1}{3}+\gamma_{1}) - \frac{\gamma_{1}z}{\ln{(1+z)}} + \frac{c/3}{\ln{(1+z)}},
629: \label{potqc10}
630: \end{equation}
631: where $z = a^{-1}-1$, $c = -\ln{|\frac{2E}{k}|}$.
632: The plot of $\gamma_{0}(z)$ for different $\gamma_{1}$ is shown in Fig. \ref{wyk4} for $\rho_{0} > 0$
633: and $k = +1$.
634: \begin{figure}[!ht]
635: \begin{center}
636: \includegraphics[scale=0.6, angle=0]{4.eps}
637: \caption{Diagram of boundary curve $\partial \mathcal{D}_{E}$ usefull in classification of the model
638: in the configurational space (formula (\ref{potqc10})). We put here $\gamma_{1} = -2/3\beta$,
639: ($\beta = -10,-5,-1,1,5,10$), $c = -1$}
640: \label{wyk4}
641: \end{center}
642: \end{figure}
643: If we put $\gamma_{0} = -1$ (cosmological constant) then we can obtain classification of cosmological models
644: with cosmological constant in terms of levels of $\gamma_{1} = {\rm const}.$, where
645: \begin{equation}
646: \gamma_{1}(z) = \frac{-\frac{2}{3}\ln{(1+z)}-c/3}{\ln{(1+z)}-z}.
647: \label{potqc11}
648: \end{equation}
649: The plot of $\gamma_{1}(z(a))$, for different $c$, is shown in Fig. \ref{wyk5}.
650: \begin{figure}[!ht]
651: \begin{center}
652: \includegraphics[scale=0.6, angle=0]{5.eps}
653: \caption{The plot of $\gamma_{1}(z(a))$ (formula (\ref{potqc11})) for different signs of c
654: ($c = -5,-1.5,0,0.15,2,5,10$)}
655: \label{wyk5}
656: \end{center}
657: \end{figure}
658:
659:
660: \noindent
661: {\bf C.}
662:
663: $p = w_{x}\rho_{x}$, $w_{x} = {\rm const}.$
664: Let us consider quintessential matter in the form of noninteracting mixture of dust and matter described
665: by the equation of state $p = w_{x}\rho_{x}$, where $w_{x} = {\rm const}.$ In this case, classification of
666: all evolutional paths can be given in terms of potential function (Fig. \ref{wyk6a})
667: \begin{equation}
668: V(a) = \frac{k}{2}a\Bigg(\frac{a^{3w_{x}}}{\frac{\rho_{0m}}{\rho_{0x}}a^{3w_{x}} + 1}\Bigg),
669: \label{potqc12}
670: \end{equation}
671: where
672: $$\rho_{x} = \rho_{0x} a^{-3(1+w_{x})}, \hspace{5mm} \rho_{m} = \rho_{0m} a^{-3}.$$
673: \begin{figure}[!ht]
674: \begin{center}
675: \includegraphics[scale=0.6, angle=0]{6a.eps}
676: \caption{Diagram of the potential function $V(a)$ for the case {\bf C}, $p~=~w_{x}\rho_{x}$, for different
677: ($w_{x}=-\frac{4}{3},-1,-\frac{2}{3},-\frac{1}{3},1,\frac{1}{3}$), $k = \pm 1$}
678: \label{wyk6a}
679: \end{center}
680: \end{figure}
681: The boundary curve $V(a) = E$ can be used to classify all evolutional paths in the configurational
682: space, namely
683: \begin{equation}
684: g(a) = 3w_{x}(a) = \frac{\ln{\bigg(\frac{A}{a - A \frac{\rho_{0m}}{\rho_{0x}}}\bigg)}}{\ln{a}},
685: \label{potqc13}
686: \end{equation}
687: where $A = 2E/k$.\\
688: The plot of $g(a)$, for different $A$, is shown in Fig. \ref{wyk6}. \\
689: \begin{figure}[!ht]
690: \begin{center}
691: \includegraphics[scale=0.6, angle=0]{6.eps}
692: \caption{Diagram of $g(a)$ for classification of evolution in the configurational space for the case
693: of constant $w_{x}$ ($A~=~0.45,0.55,1,2$), $\frac{\rho_{0m}}{\rho_{0x}}=1$}
694: \label{wyk6}
695: \end{center}
696: \end{figure}
697:
698: \noindent
699: {\bf D.}
700:
701: Since the observations of the supernovae of type Ia indicate that the universe must be today in an
702: accelerated expansion the nature of the fluid responsible for such a behaviour has been object of many
703: studies. While the most obvious candidate for such component is the vacuum energy the posibility that the
704: dark energy might be described by the Chaplygin gas is seriously suggested \cite{kambenalc}.
705:
706: The Chaplygin gas has an interesting motivation connected with the string theory. If we consider a d--brane
707: configuration in the d+2 Nambu-Goto action, the employment of the light-cone parametrization leads to the
708: action of a Newtonian fluid with the equation of state $p = -A/\rho$, whose symmetries are the same
709: as those of the Poincare' group. Hence, the relativistic character of the action is somehow hidden in the
710: equation of state (for review see R. Jackiv, A particle field theorist's lectures on supersymmetric,
711: non-abelian fluid mechanics and d-branes physics).
712:
713: Let us consider FRW model with fluid in the form of generalized Chaplygin gas for which equation of state
714: is given by
715: \begin{equation}
716: p = -\frac{A}{\rho^{\alpha}},\hspace{5mm} 0 \leqslant \alpha < 1.
717: \label{chapeqst}
718: \end{equation}
719: The energy-momentum conservation implies that Chaplygin gas density depends on the scale factor as
720: \begin{equation}
721: \rho = \bigg(A + \frac{B}{a^{3(1+\alpha)}}\bigg)^{\frac{1}{1+\alpha}},
722: \label{chapdens}
723: \end{equation}
724: where $A$ and $B$ are constants and the Chaplygin gas corresponds to the case $\alpha = 1$; $A$ is a positive
725: constant because sound velocity of Chaplygin gas is $v_{s}^{2}/c^{2} = A^{2}/\rho^{2}$.
726:
727: Equation (\ref{chapdens}) interpolates smoothly between a dust dominated phase, where
728: $\rho \propto a^{-3}$, and a De Sitter phase, where $p = -\rho$, through an intermediate regime described by
729: the Zeldovich stiff matter $p = \rho$.
730:
731: Following our previous consideration, the dynamics is given by the hamiltonian
732: $$\mathcal{H} = \frac{\dot{a}^{2}}{2} + V(a) \equiv 0,$$
733: where
734: \begin{equation}
735: V(a) = -\frac{1}{6}\rho_{{\rm eff}}a^{2} =
736: -\frac{1}{6}\bigg(A + \frac{B}{a^{3(1+\alpha)}}\bigg)^{\frac{1}{1+\alpha}}a^{2} + \frac{k}{2}.
737: \label{chapham}
738: \end{equation}
739: The dependence of $V(a)$ for $\alpha = 1$ is illustrated on the Fig. \ref{wyk7}
740: \begin{figure}[!ht]
741: \begin{center}
742: \includegraphics[scale=0.6, angle=0]{7.eps}
743: \caption{Diagram of the potential function $V(a)$ for the case {\bf D} (formula (\ref{chapham})),
744: $k = A = B = \alpha = 1$.}
745: \label{wyk7}
746: \end{center}
747: \end{figure}
748: The domain admissible for motion of the system with generalized Chaplygin gas is
749: $$\mathcal{D}_{0} = \{a: V(a) \leqslant 0\}.$$
750: The boundary curve $\partial \mathcal{D}_{0}$ can be used to classify possible evolution paths in
751: the configurational space in the following way. We consider constant levels of $A(a)$ relation given by
752: \begin{equation}
753: A(a) = \bigg(\frac{3k}{a^{2}}\bigg)^{1+\alpha} - \frac{B}{a^{3(1+\alpha)}}.
754: \label{chapa}
755: \end{equation}
756: The zero velocity curve for $k = +1$ are shown on Fig. \ref{wyk8} and Fig. \ref{wyk9}.
757: \begin{figure}[!ht]
758: \begin{center}
759: \includegraphics[scale=0.6, angle=0]{8.eps}
760: \caption{Diagram of $A(a)$ for classification of possible evolution paths for the case
761: of constant $k=+1$, $\alpha = 0.5$ and different $B$ ($B~=~0,0.6,1,2,5$)}
762: \label{wyk8}
763: \end{center}
764: \end{figure}
765: \begin{figure}[!ht]
766: \begin{center}
767: \includegraphics[scale=0.6, angle=0]{9.eps}
768: \caption{Diagram of $A(a)$ for classification of possible evolution paths for the case
769: of constant $k=+1$, $B = 2$ and different $\alpha$ ($\alpha~=~0,0.5,0.75,1$)}
770: \label{wyk9}
771: \end{center}
772: \end{figure}
773: \begin{figure}[!ht]
774: \begin{center}
775: \includegraphics[scale=0.35, angle=270]{10.eps}
776: \caption{The phase portrait $(a,x)$ for case {\bf D} ($k = A = B = \alpha = 1$). The shaded region is a
777: region of accelerated expansion of the universe.}
778: \label{wyk10}
779: \end{center}
780: \end{figure}
781:
782: It is interesting that the existence and character of the critical points of the considered system
783: $\dot{a} = x,\hspace{3mm} \dot{x} = -\frac{\partial V}{\partial a}$ depends on the geometry of the potential
784: function.
785:
786: It can be easily to shown that in our case on a finite region of the phase plane $(a, x)$ only saddle points
787: are admitted because $\partial^{2} V/\partial a^{2} < 0$ at the critical point $a = a_{0}$,
788: $\partial V/\partial a|_{a=a_{0}} = 0$. Therefore, all points are hiperbolic ($TrA = 0$) and the system is
789: structurally stable.
790:
791: In terms of $V(a)$ the domains of accelerating trajectories can be easily found, namelly if
792: $\partial V/\partial a < 0$ then the system starts to accelerate. Because the diagram of $V(a)$ is upper
793: convex, the static critical point will separate regions without acceleration from the domain in the phase
794: space where trajectories accelerate. One can shown that the critical value of $a$ is
795: $$a = a_{crit} = \Bigg(\frac{B}{2A}\Bigg)^{\frac{1}{3(1+\alpha)}}.$$
796: At this point the diagram $V(a)$ has the maximum.
797:
798: \noindent
799: {\bf E.}
800:
801: Interesting formulas for $w(a)$ were already proposed by Corasaniti and Copeland \cite{corasaniti}.
802: They considered a broad class of tracking potentials for scalar fields, namely $V(\phi) \propto
803: \phi^{-\alpha}$ -- inverse power low potential (INV) \cite{zlatev},
804: $V(\phi) \propto \phi^{-\alpha} \exp{(\phi^{2}/2)}$ -- supergravity potential (SUGRA) \cite{brax},
805: $V(\phi) \propto \exp{(-\alpha \phi)} + \exp{(\beta \phi)}$ (2EXP) \cite{barreiro}, The Skordis model
806: \cite{albrecht} and Copeland-Nunes-Rosati model (CNR) \cite{copeland}.
807:
808: \begin{figure}[!ht]
809: \begin{center}
810: \includegraphics[scale=0.4, angle=270]{ww.eps}
811: \caption{The dependence of the equation of state factor $w(a)$ versus the scale factor $a/a_{0}$ for
812: different models of potential of scalar field.}
813: \label{wfig}
814: \end{center}
815: \end{figure}
816:
817: \begin{figure}[!ht]
818: \begin{center}
819: \includegraphics[scale=0.4, angle=270]{vv.eps}
820: \caption{The dependence of the potential of the Hamiltonian system against the scale factor $a/a_{0}$ for
821: a different class of tracking potentials.}
822: \label{vfig}
823: \end{center}
824: \end{figure}
825:
826: First, we aply our method to construct a potential function of the corresponding Hamiltonian dynamical
827: system. It assumes the following form
828: \begin{align}
829: V(a)&=-\frac{\rho_{0}}{6}a^{-1}\exp{\Bigg(3\int_{a}^{1}\frac{w(a)}{a}da\Bigg)}\nonumber \\
830: &=-\frac{\rho_{0}}{6a}\exp{\Bigg\{3\Bigg(F_{1}\int_{a}^{1}\frac{f_{r}(a)}{a}da+F_{2}\int_{a}^{1}\frac{f_{m}(a)}{a}da+F_{3}(1-a)\Bigg)\Bigg\}},
831: \label{revpot}
832: \end{align}
833: where
834: \begin{equation}
835: \int_{a}^{1}\frac{f_{r,m}}{a}da = \sum_{n=0}^{\infty}[{\rm Ei}(-\beta n x)_{r,m}-{\rm Ei}(1)],
836: \label{eifunc}
837: \end{equation}
838: where $x=a-a_{c}$, $\beta=1/\Delta$, Ei -- exponent integral function
839: and coefficients $F_{1}$, $F_{2}$, $F_{3}$ are determined by the condition that $w(a)$ takes on the
840: respective values of $w^{r}$, $w^{m}$ and $w^{0}$, during radiation epoch ($a=a_{r}$), matter domination
841: ($a=a_{m}$) as well as today ($a=a_{0}=1$).
842:
843: The coresponding forms of $w(a)$ function for a different class of potentials are illustrated in
844: Fig.~\ref{wfig}.
845:
846: The function $f_{r,m}(a)$ has the following form \cite{corasaniti}
847: \begin{equation}
848: f_{r,m}(a)=\frac{1}{1+\exp{[-(a-a_{c}^{r,m})/\Delta_{r,m}]}},
849: \label{frmform}
850: \end{equation}
851: where the corresponding values of coefficients $a_{c}^{r,m}$ are taken from Table I in ref \cite{corasaniti}.
852:
853: After substitution formulas (\ref{frmform}) to (\ref{revpot}) we obtain different forms of
854: potentials for a different class of models (see Fig.~\ref{vfig}). From this figure we can observe that in
855: all cases a potential function is of the same type (upper convex), like for the Chaplygin gas. Therefore
856: the phase portraits determined from the potential functions of the systems are topologically equivalent.
857: From the physical point of view it means that considered models can be seriously treated as candidates for dark energy
858: description.
859:
860: Let us notice that the dynamical system
861: $$\dot{a} = x, \hspace{5mm} \dot{x} = -\frac{\partial V}{\partial a}$$
862: can be transformed to $(z,x)$ variables ($a = (1+z)^{-1}$) and then we obtain
863: $$\frac{dz}{d\tau} = -(1+z)^{2}x,\hspace{5mm} \frac{dx}{d\tau} = (1+z)^{2}\frac{\partial V}{\partial z}.$$
864: Therefore after the reparametrization of time along trajectories $\tau \rightarrow \eta: d\eta = -(1+z)^{2}d\tau$
865: (now $\eta$ will be a decreasing function of time variable $\tau$) we obtain the dynamical system
866: \begin{equation}
867: \frac{dz}{d\eta} = x, \hspace{5mm} \frac{dx}{d\eta} = -\frac{\partial V}{\partial z}.
868: \label{qcdynzxeta}
869: \end{equation}
870: Of course, system (\ref{qcdynzxeta}) can be analysed in terms of dynamical systems, i.e. in terms of the
871: method of qualitative analysis of differential equations on the phase plane $(z,x)$. System
872: (\ref{qcdynzxeta}) has the first integral in the form
873: \begin{equation}
874: \Bigg(\frac{dz}{d\eta}\Bigg)^{2} - 2(E - V(z)) = 0.
875: \label{qcdynzxetaint}
876: \end{equation}
877: For the considered system only two types of critical points can appear, namelly centres or saddle points.
878: If ${\rm det}A = \frac{\partial^{2} V}{\partial z^{2}}\big|_{z=z_{0}}$ is negative, the diagram of the
879: potential function $V(a(z))$ has maxima; they correspond to the saddle point. On the other hand, if the
880: diagram of the potential function $V(a(z))$ has minima, they correspond to centres. It is important that we
881: can discuss the stability of critical points based only on the geometry of the potential function. It is easy
882: to check that at the critical points of the system appearing at $a_{0}~=~{\rm const}.
883: (z_{0} = {\rm const}.)$, $\frac{\partial V}{\partial a}\big|_{a_{0}} = 0$
884: ($\frac{\partial V}{\partial z}\big|_{z_{0}} = 0$) we have $\kappa(a) = 0$, and then
885: \begin{equation}
886: \begin{split}
887: \frac{\partial^{2} V}{\partial a^{2}}\Bigg|_{a=a_{0}}
888: &= \varphi^{2}\frac{3}{2}\frac{k}{a}\frac{dw}{da}\Bigg|_{a=a_{0}}
889: = \frac{3}{2}\frac{\varphi^{2}}{a}k\frac{dw}{da}\Bigg|_{w(a)=-1/3} \\
890: &= -\frac{3}{2}\varphi^{2}k(1+z)^{3}\frac{dw}{dz}\Bigg|_{z=z_{0}}.
891: \end{split}
892: \label{qcdynzxeta1}
893: \end{equation}
894: Therefore, if $k \frac{dw}{dz}(z_{0})$ is positive, only saddles points can appear which guarantee the
895: structural stability of the system.
896:
897: For the case {\bf B} $\frac{dw}{dz} = \gamma_{1}$, the above condition means that $\gamma_{1}k > 0$.
898: The phase portraits for $\gamma_{0} = -1/3$, and various $\gamma_{1}$, are presented on Fig. \ref{B},
899: \ref{B1},\ref{B2} and \ref{B3}.
900:
901: Let us note that critical points are located at
902: \begin{equation}
903: z_{0} = -\Bigg(\frac{1}{3 \gamma_{1}} + \frac{\gamma_{0}}{\gamma_{1}}\Bigg).
904: \label{qccrpt}
905: \end{equation}
906: If we put, for example, $\gamma_{0} = -1$ (cosmological constant term) then $\gamma_{1} > -\frac{3}{2}$
907: critical points can only exist on a finite domain of the phase plane because $z > -1$.
908:
909: The idea of structural stability comes from Andronov and Pontriagin \cite{andronow}. A~dynamical
910: systems $S$ are said to be structurally stable if their dynamical behaviour remains qualitatively
911: (modulo homeomorphizm preserving orientation of trajectories) the same (equivalent) under small perturbations.
912: Structural stability is sometimes considered as a precondition for the ``real existence''.
913: Structurally stable dynamical systems on the 2D compact space (for example on the plane with circle at
914: infinity) form open and dense subsets in the space of all dynamical systems on the plane. Therefore, 2D
915: structurally unstable models seem to be nonadequate for describing real proceses because of measurement
916: errors.
917:
918: The main aim of the qualitative analysis of differential equations is not to find, and then to analyze,
919: individual solutions but rather to investigate space of all possible solutions for all
920: admissible initial conditions. A property is believed to be ``realistic'' if it can be attributed to large
921: (rather typical than exceptional) subsets of models within the space of all possible solutions, or if it
922: possesses a certain stability, i.e. if it is shared by a slighty perturbed model. There is a wide opinion
923: among specialists that realistic models should be structurally stable, or even stronger, that everything
924: that exists should possess a kind of structural stability.
925:
926: From the physical point of view it is interesting to answer the question: are the trajectories for which
927: acceleration of the universe takes place, distributed in a typical or exceptional way? How are trajectories
928: with interesting properties distributed in the phase plane? For example, the acceleration condition
929: $\ddot{a} = - \partial V/ \partial a > 0$ is satisfied if $V(a)$ is a decreasing function of $a$.
930: For us it is important that one should be easily able to deduce this from the geometry of the potential
931: function only.
932:
933: In the phase space, the area of acceleration is determined by the condition that
934: \begin{equation}
935: \frac{\ddot{a}}{a} = -\frac{1}{6}(1+3w)\rho > 0,\hspace{5mm}
936: \frac{\partial V}{\partial a} + \psi(a)(a')^{2} < 0.
937: \label{accond1}
938: \end{equation}
939: (\ref{accond1}) can be rewritten to the form which could be usefull in the analysis of the ``probability
940: of acceleration'' which can be defined as a measure of the space of those initial conditions that lead
941: to accelerating universes
942: \begin{align}
943: &-\frac{\partial}{\partial a} \ln{(E-V)}+\frac{1+3w(a)}{a} < 0, \nonumber \\
944: \intertext{or}
945: &(1+z)^{2}\frac{\partial}{\partial z}\ln{(E-V(z))}+(1+3w(a(z))(1+z) < 0.
946: \label{accond2}
947: \end{align}
948:
949: For example, domains of acceleration in the configurational space, for the case {\bf D}, as well as the
950: domain of $\ddot{a} > 0$ ($\frac{\partial V}{\partial a} < 0$) in the phase plane are presented on
951: Fig.~\ref{wyk10}.
952:
953:
954: Let us now consider the presence or absence of the particle horizon in the past. Good news from our earlier
955: discussion is that this property can be detected from the shape of the diagram of the potential function of
956: the system, namelly if $V(z)$ goes to a constant (zero is included) as $z \rightarrow \infty$ then we obtain
957: a model without the horizon.
958:
959: In the next section it will be demonstrated how we can answer the question about the horizon in the past on
960: the base of $V(z)$ taken from the observations.
961:
962:
963: \section{Inverse problem in Quintessential Cosmology}
964:
965: The presented formalism gives us a natural base to discuss the redshift magnitude relation $m(z)$ for SNIa
966: supernovae observational data. But on the other hand, because the Hubble function is related to the luminosity
967: distance, it is possible to determine both the quintessence parameter $w(z) = p/\rho$ and the potential
968: of the dynamical system $V(z)$. Therefore, the equation of state as well as the whole dynamics can be
969: reconstructed provided that the luminosity function $d_{L}(z)$ is known from observations. It is called
970: the inverse problem in dynamics of quintessential cosmology.
971:
972: As an example of constructing observables from the considered dynamics let us consider the
973: luminosity--distance relation $d_{L}(z)$ for quintessential models.
974:
975: If a light source of redshift $z$ is located at a radial coordinate $r_{1}$ (taken from R--W metric), its
976: luminosity distance $d_{L}$, its angular diameter distance $d_{A}$ and its proper motion distance are given by
977: \begin{equation}
978: d_{L}(z) = (1+z)a_{0}r_{1},\hspace{5mm} d_{A} = \frac{a_{0}r_{1}}{1+z},\hspace{5mm} d_{M} = a_{0}r_{1},
979: \label{invlumdist1}
980: \end{equation}
981: where $r_{1}$ calculated from metric gives
982: \begin{equation}
983: \begin{split}
984: \varphi(r_{1}) &= \int^{a_{0}}_{\frac{a_{0}}{1+z}}\frac{da}{a\dot{a}} =
985: \frac{1}{a_{0}}\int_{0}^{z}\frac{dz'}{H(z')} \\
986: &= \begin{cases} \sin^{-1}{r_{1}} & \textrm{when } k=+1 \\
987: r_{1} & \textrm{when } k=0 \\
988: \sinh{r_{1}} & \textrm{when } k=-1 \end{cases}
989: \end{split}
990: \label{invrad1}
991: \end{equation}
992: Here $a_{0}$ is the present value of the radius of the universe. The above equation can also be written in
993: the form of a single compact equation as
994: \begin{equation}
995: \frac{d_{L}(z)}{1+z} = \frac{1}{\sqrt{\kappa}}\zeta \Bigg(\sqrt{\kappa}\int^{z}_{0}\frac{dz'}{H(z')}\Bigg),
996: \label{invlumdist2}
997: \end{equation}
998: where
999: $$
1000: \begin{array}{lll}
1001: \zeta(q) = \sin{q} & \textrm{with } \kappa=\Omega_{k,0} & \textrm{when } \Omega_{k,0}>0, \\
1002: \zeta(q) = \sinh{q} & \textrm{with } \kappa=-\Omega_{k,0} &\textrm{when }\Omega_{k,0}<0, \\
1003: \zeta(q) = q & \textrm{with } \kappa=1 &\textrm{when }\Omega_{k,0}=0.
1004: \end{array}
1005: $$
1006: Thus for a given $\mathcal{M}$ (absolute magnitude) and $H(z)$ equation (\ref{invlumdist2}) gives the
1007: predicted value of $m(z)$ (observed magnitude) at a given $z$.
1008:
1009: By using the $\kappa$--corrected effective magnitudes $m_{i}$ which have also been corrected for the
1010: light curve width--luminosity relation and the galactic extinction, and by using the same standard errors
1011: $\sigma_{2,i}$ and $\sigma_{m_{i}}^{{\rm eff}}$ of the supernova with redshift $z_{i}$ as used by
1012: Perlmutter et al. we compute $\chi^{2}$ according to
1013: \begin{equation}
1014: \chi^{2} = \sum_{i}\frac{[m_{i}^{{\rm eff}}-m(z_{i})]^{2}}{\sigma_{z_{i}}^{2}+\sigma_{m_{i}\ {\rm eff}}^{2}}.
1015: \label{invhikw}
1016: \end{equation}
1017: The best fit parameters are obtained by minimizing equation (\ref{invhikw}).
1018:
1019: Luminosity distance and angular distance depend sensitively on the present densities of various energy
1020: components and their equations of state. For this reason, the magnitude--redshift relation for distant
1021: standard candles has been proposed as a potential test for cosmological models.
1022:
1023: In our formalism $H(z)$ can be immediately taken from the first integral of the dynamical equation and then
1024: we obtain
1025: \begin{equation}
1026: \frac{d_{L}(z)}{1+z} = \frac{1}{\sqrt{\kappa}}\zeta \Bigg(\sqrt{\kappa}\int^{z}_{0}
1027: \frac{e^{-3/2\int\frac{w(a(z'))}{1+z'}dz'}dz}{(1+z)^{3/2}\sqrt{2(E-V(z))}}\Bigg).
1028: \label{invlumdist3}
1029: \end{equation}
1030: Formula (\ref{invlumdist3}) limits the determination of the luminosity distance because it depends on
1031: quintessential parameter $w(z)$ through a multiple integral relation that smears out detailed information
1032: about $w(z)$. If $w(a(z))$ can be expanded as the Taylor series following (\ref{wrozw}) then we obtain the
1033: simplest formula without the double integration
1034: \begin{equation}
1035: \frac{d_{L}(z)}{1+z} = \frac{1}{\sqrt{\kappa}}\zeta \Bigg(\sqrt{\kappa}\int^{z}_{0}
1036: \frac{\exp{\{-\frac{3}{2}\sum \gamma_{i}z'^{i} dz'\}}}{\sqrt{2(1+z')^{3(1+\gamma_{0})}(E-V(z'))}}\Bigg).
1037: \label{invlumdist4}
1038: \end{equation}
1039: Many authors \cite{saini} assume that a quite accurate luminosity distance may be obtained and then examined
1040: to answer the question of whether the equation of state of the expanding universe can be determined uniquely.
1041:
1042: Our idea is more general. To determine the structure and evolution of an astrophysical system of the universe,
1043: the equation of state is usually necessary. By equation of state of the universe we mean the relation
1044: between the total energy density of cosmic matter and the total pressure. However, the equation of state
1045: relevant to the universe has not yet been established. Our idea is to reconstruct it from the form of the
1046: potential of the system $V(a(z))$.
1047:
1048: Let us consider, for simplicity, a flat model for which the Hubble parameter is related to luminosity
1049: distance by the relation
1050: \begin{equation}
1051: H(z) = \Bigg[\frac{d}{dz}\Bigg(\frac{d_{L}(z)}{1+z}\Bigg)\Bigg]^{-1}.
1052: \label{invhubbl}
1053: \end{equation}
1054: Then it is possible to determine the quintessence parameter
1055: \begin{equation}
1056: w(z) = -1 -\frac{2}{3}H(1+z)\frac{d^{2}}{dz^{2}}\Bigg(\frac{d_{L}(z)}{1+z}\Bigg).
1057: \label{invquintess1}
1058: \end{equation}
1059: Here the term $-1$ is established from the condition that for stationary solution $H(z) = {\rm const}.$
1060: $\frac{d^{2}}{dz^{2}}(\frac{d_{L}(z)}{1+z})$ will vanish but such a solution can appear on the phase plane
1061: $(H,\rho)$ only as an intersection trajectory of the flat model and the boundary of the weak energy
1062: condition $\rho+p=0$.
1063:
1064: Equation (\ref{invquintess1}) can be rewritten in the form
1065: \begin{align}
1066: p &= - \rho - 3\xi(z)H, \\
1067: \intertext{where}
1068: \xi(z) &= \frac{2}{3}H^{2}(z)(1+z)\frac{d^{2}}{dz^{2}}\Bigg(\frac{d_{L}(z)}{1+z}\Bigg). \nonumber
1069: \label{invquintess2}
1070: \end{align}
1071: The dependence $p(H)$ manifests the presence of bulk viscosity effects in the model
1072: $\xi(z) = -\frac{1}{3}\frac{\partial p}{\partial H}$. Therefore, from $d_{L}(z)$ we obtain the reconstruction
1073: of $w(z)$ as a mixture of the cosmological constant term and bulk viscosity. However, in determining $w(z)$
1074: there is an inherent limitation because the value $w(z)$ is poorly resolved and no usefull constraint can be
1075: obtained concerning its time variation. Of course, the value of $w(0)$ and $dw/dz$ with accuracy $0.1$
1076: and $0.15$ respectively will be obtained from SNAP3 but for the reconstruction of $w(z)$ higher derivatives
1077: can be useful. If we measure $\frac{dw^{i}}{dz}(0)$, $i \geqslant 2$, then in principle, the reconstruction
1078: equation of the equation of state is possible, because then
1079: $$w(z) = w(0) +\frac{dw}{dz}(0)z + \frac{1}{2}\frac{d^{2}w}{dz^{2}}(0)z^{2} + \ldots,$$
1080: where $\frac{d^{n}w}{dz^{n}}(0)$ is determined from the recurrence formula, namelly
1081: \begin{equation}
1082: \begin{split}
1083: \frac{d^{n}w}{dz^{n}}(0) & = \frac{2}{3}\frac{d^{n}}{dz^{n}}\Bigg|_{0}(\ln{H}) + \frac{2}{3}(1+z)\frac{d^{n+1}}{dz^{n+1}}\Bigg|_{0}(\ln{H}), \\
1084: w(0) & = - 1 + \frac{2}{3}\frac{d}{dz}\Bigg|_{0}(\ln{H}).
1085: \end{split}
1086: \label{invquintess3}
1087: \end{equation}
1088: Then $w(z)$ can be obtained from $H(z)$ in the following way
1089: \begin{equation}
1090: \begin{split}
1091: H(z) & \mapsto \ln{H(z)} \mapsto \forall n \frac{d^{n+1}}{dz^{n+1}}(\ln{H}) \\
1092: & \mapsto \frac{d^{n}w}{dz^{n}}(0) \mapsto w(z) = \sum_{i/1}^{\infty}\frac{1}{i!}\frac{d^{i}w}{dz^{i}}(0)z^{i} + w(0).
1093: \end{split}
1094: \label{invquintess4}
1095: \end{equation}
1096: At present this idea can not be realized but the idea of reconstructing dynamics requires the knowledge of
1097: $V(z)$ which can be calculated from the Hamiltonian constraint
1098: \begin{equation}
1099: \begin{split}
1100: V(a(z)) &\equiv - \frac{\rho_{{\rm eff}}a^{2}(z)}{6} = - \frac{1}{2} H^{2}(z)a^{2}(z) \\
1101: &= - \frac{1}{2} \Bigg[\frac{1}{(1+z)\frac{d}{dz}\big(\frac{d_{L}(z)}{1+z}\big)}\Bigg]^{2}.
1102: \end{split}
1103: \label{invpot}
1104: \end{equation}
1105: On the other hand, the knowledge of $V(z)$ gives us information how the horizon problem can be solved.
1106: Reconstruction of the potential function from SNIa data is presented on Fig.\ref{snpot}. We obtain the plot
1107: of the potential function from fitting
1108: \begin{equation}
1109: \begin{split}
1110: \frac{d_{L}(z)}{1+z} &= \int_{0}^{z}\Bigg[A_{0}+A_{1}(1+z')+A_{2}(1+z')^{2}+{} \\
1111: &+ A_{3}(1+z')^{3}+A_{4}(1+z')^{4}\Bigg]^{-\frac{1}{2}}dz'
1112: \end{split}
1113: \label{fitfun}
1114: \end{equation}
1115: function to the SNIa observational data.
1116:
1117: \begin{figure}[!ht]
1118: \begin{center}
1119: \includegraphics[scale=0.5]{snpot.ps}
1120: \caption{The plot of the potential function $V(z)$ has been obtained from equation (\ref{invpot}) and from
1121: fitting formula (\ref{fitfun}) to the SNIa observational data}
1122: \label{snpot}
1123: \end{center}
1124: \end{figure}
1125:
1126: We can observe that as $z \rightarrow \infty$, $V(z)$ as calculated from (\ref{invpot}) goes to $-\infty$.
1127: This can be treated as an empirical evidence of the presence horizon in the past. Let us note that other
1128: problems of the standard cosmology can be discussed analogously basing on information about $V(z)$ and its
1129: geometry.
1130:
1131: \section{Conclusions}
1132:
1133: In this work a class of FRW models with quintessence matter is examined in the context of the present
1134: acceleration of the universe. Our results are the following.
1135:
1136: \begin{enumerate}
1137: \item
1138: We have given a mathemathical background for discussing physical content of quintessential cosmology.
1139: Its dynamics is reduced to the dynamics of the unit mass point particle in 1D potential. Then different
1140: physical properties, like acceleration of the universe, existence of horizon, can be formulated only in terms
1141: of the potential function of the system. The proof of the corresponding condition is quite general. In
1142: particular, it is independent of any specific assumption about the behaviour of the scale factor near the
1143: singularity (such as the assumption of power low behaviour) or a specific form of equation of state.
1144: \item
1145: The dynamics is formulated in the hamiltonian formalism and the full classification of possible evolutions
1146: in the phase plane as well as in configurational space is given. In the near future, it will be possible to
1147: obtain, from SNAP3, the exact value of $\gamma_{0}$ and $\gamma_{1}$ appearing in the equation of state
1148: $p = (\gamma_{0} + \gamma_{1}z)\rho$, and then we could automatically answer the question about the
1149: horizon.
1150: \item
1151: The effectiveness of our treatment of dynamics of quintessential models in terms of single-particle mechanics
1152: is demonstrated for a broad class of tracking potentials. We obtain topological equivalence of the phase
1153: portraits (for this case) with the dynamical system obtained for the potential function reconstructed from
1154: SNIa data.
1155: \item
1156: The idea of reconstructing dynamics instead of quintessential coefficient is considered. It is what we called
1157: inverse problem in quintessential cosmology. We demonstrate that the reconstructed potential function of the
1158: system produces the particle horizon in the past and solves the flatness problem.
1159: \end{enumerate}
1160:
1161: \begin{acknowledgments}
1162: Authors are very greatefull to prof M. Heller and dr W. Godlowski for discussion and comments.
1163: \end{acknowledgments}
1164:
1165:
1166: \begin{thebibliography}{99}
1167: \bibitem{milne}
1168: E.A. Milne, Quart. J. Math. {\bf 5}, 64 (1934).
1169: \bibitem{mccrea}
1170: W.H. McCrea, E.A. Milne, Quart. J. Math. {\bf 5}, 73 (1934).
1171: \bibitem{mccrea1}
1172: W.H. McCrea, Proc. Roy. Soc. London A206, 562 (1951).
1173: \bibitem{harrison}
1174: E.R. Harrison, Ann. Phys. 35, 437 (1965).
1175: \bibitem{ellis}
1176: G.F.R. Ellis, (1973) In Cargese Lectures in Physics 6. E. Schatzman, ed. (Gordon and Breach, New York) p. 17
1177: \bibitem{perlmutter}
1178: S. Perlmutter et al., (The Supernova Cosmology Project) Nature {\bf 391}, 51 (1998);
1179: P.M. Garnavich et al. Astrophys. J. {\bf 509}, 94 (1998).
1180: \bibitem{riess}
1181: A.G. Riess et al., Astrophys. J. {\bf 116}, 1009 (1998).
1182: \bibitem{schmidt}
1183: B.P. Schmidt, Astrophys. J. {\bf 507}, 46 (1998).
1184: \bibitem{caldwell}
1185: R.R. Caldwell, R. Dave, P.J. Steinhardt, Phys. Rev. Lett. 80, 1582 (1998).
1186: \bibitem{lima}
1187: J.A.S. Lima, J.A.M. Moreira and J. Santos, Gen. Relat. Grav. {\bf 30}, 425 (1998).
1188: \bibitem{sahni}
1189: V. Sahni, The cosmological constant problem and quintessence, astro-ph/0202076;
1190: A. Starobinsky, JETP Lett. 68, 757 (1998).
1191: \bibitem{weinberg}
1192: S. Weinberg, ``Gravitation and Cosmology'', Wiley, New York, (1972).
1193: \bibitem{szydlowski}
1194: M. Szydlowski, M.P. Dabrowski, A. Krawiec, Phys. Rev. D {\bf 66} 064003 (2002).
1195: \bibitem{kambenalc}
1196: A. Kamenshchik, U. Moschella, V. Pasquier, Phys. Lett. B, 511, 256 (2001);
1197: M.C. Bento, O. Bertolami, A.A. Sen, gr-qc/0202064 (2002); Phys. Rev. D {\bf 67} 063003;
1198: J.S. Alcaniz, D. Jain, A. Dev, astro-ph/0210476 (2002);
1199: R. Bean, O. Dor, astro-ph/0301308.
1200: \bibitem{corasaniti}
1201: P.S. Corasaniti, E.J. Copeland, Phys. Rev. D {\bf 67} 063521 (2003).
1202: \bibitem{zlatev}
1203: I. Zlatev, L. Wang and P.J. Steinhardt, Phys. Rev. Lett. {\bf 82} 896 (1999).
1204: \bibitem{brax}
1205: P. Brax, J. Martin, Phys. Lett. B {\bf 468}, 40 (1999).
1206: \bibitem{barreiro}
1207: T. Barreiro, E.J. Copeland and N.J. Nunes, Phys. Rev. D {\bf 61} 127301 (2000).
1208: \bibitem{albrecht}
1209: A. Albrecht and C. Skordis, Phys. Rev. Lett. {\bf 84} 2076 (1999).
1210: \bibitem{copeland}
1211: E.J. Copeland, N.J. Nunes and F. Rosati, Phys. Rev. D {\bf 62} 123503 (2000).
1212: \bibitem{saini}
1213: T.D. Saini, S. Raychaudhury, V. Sahni and A.A. Starobinsky, Phys. Rev. Lett. {\bf 85}, 1162 (2000);
1214: D. Huterer, M.S. Turner, Phys. Rev. D {\bf 60}, 081301 (1999);
1215: T. Nakamara, T. Chiba, Mon. Nat. Astron. Soc. {\bf 306}, 696 (1999).
1216: \bibitem{robertson}
1217: H.P. Robertson, Rev. Mod. Phys. 5, 62 (1933).
1218: \bibitem{andronow}
1219: A.A. Andronov, L.S. Pontriagin, Systems grossiers. Dokl. Akad. Nauk SSSR {\bf 14} 247-250 (1937).
1220: \end{thebibliography}
1221:
1222: \onecolumngrid
1223:
1224: \newpage
1225:
1226: \begin{figure}[!hp]
1227: \begin{center}
1228: \includegraphics[scale=0.3, angle=270]{A.ode.ps}
1229: \caption{The phase portrait $(z,x)$ for case {\bf A} $p = (\gamma_{0} + \gamma_{1}z + \gamma_{2}z^{2})\rho$
1230: ($k = 1$, $\gamma_{2} = -1$)}
1231: \label{A}
1232: \end{center}
1233: \end{figure}
1234:
1235: \begin{figure}[!hp]
1236: \begin{center}
1237: \includegraphics[scale=0.3, angle=270]{A1.ode.ps}
1238: \caption{The phase portrait $(z,x)$ for case {\bf A} $p = (\gamma_{0} + \gamma_{1}z + \gamma_{2}z^{2})\rho$
1239: ($k = 1$, $\gamma_{2} = 1$)}
1240: \label{A1}
1241: \end{center}
1242: \end{figure}
1243:
1244: \begin{figure}[!hp]
1245: \begin{center}
1246: \includegraphics[scale=0.3, angle=270]{A2.ode.ps}
1247: \caption{The phase portrait $(z,x)$ for case {\bf A} $p = (\gamma_{0} + \gamma_{1}z + \gamma_{2}z^{2})\rho$
1248: ($k = -1$, $\gamma_{2} = -1$)}
1249: \label{A2}
1250: \end{center}
1251: \end{figure}
1252:
1253: \begin{figure}[!hp]
1254: \begin{center}
1255: \includegraphics[scale=0.3, angle=270]{A3.ode.ps}
1256: \caption{The phase portrait $(z,x)$ for case {\bf A} $p = (\gamma_{0} + \gamma_{1}z + \gamma_{2}z^{2})\rho$
1257: ($k = -1$, $\gamma_{2} = 1$)}
1258: \label{A3}
1259: \end{center}
1260: \end{figure}
1261:
1262: \begin{figure}[!hp]
1263: \begin{center}
1264: \includegraphics[scale=0.3, angle=270]{B.ode.ps}
1265: \caption{The phase portrait $(z,x)$ for case {\bf B} $p = (\gamma_{0} + \gamma_{1}z)\rho$
1266: ($k = 1$, $\gamma_{0}=-1/3$, $\gamma_{1}=-2/3$)}
1267: \label{B}
1268: \end{center}
1269: \end{figure}
1270:
1271: \begin{figure}[!hp]
1272: \begin{center}
1273: \includegraphics[scale=0.3, angle=270]{B1.ode.ps}
1274: \caption{The phase portrait $(z,x)$ for case {\bf B} $p = (\gamma_{0} + \gamma_{1}z)\rho$
1275: ($k = 1$, $\gamma_{0}=-1/3$, $\gamma_{1}=2/3$)}
1276: \label{B1}
1277: \end{center}
1278: \end{figure}
1279:
1280: \begin{figure}[!hp]
1281: \begin{center}
1282: \includegraphics[scale=0.3, angle=270]{B2.ode.ps}
1283: \caption{The phase portrait $(z,x)$ for case {\bf B} $p = (\gamma_{0} + \gamma_{1}z)\rho$
1284: ($k = -1$, $\gamma_{0}=-1/3$, $\gamma_{1}=-2/3$)}
1285: \label{B2}
1286: \end{center}
1287: \end{figure}
1288:
1289: \begin{figure}[!hp]
1290: \begin{center}
1291: \includegraphics[scale=0.3, angle=270]{B3.ode.ps}
1292: \caption{The phase portrait $(z,x)$ for case {\bf B} $p = (\gamma_{0} + \gamma_{1}z)\rho$
1293: ($k = -1$, $\gamma_{0}=-1/3$, $\gamma_{1}=2/3$)}
1294: \label{B3}
1295: \end{center}
1296: \end{figure}
1297:
1298: \begin{figure}[!hp]
1299: \begin{center}
1300: \includegraphics[scale=0.3, angle=270]{C.ode.ps}
1301: \caption{The phase portrait $(a,x)$ for case {\bf C} $p = w_{x}\rho_{x}$,
1302: $k = 1$, $\rho_{0m}/\rho_{0x} = 1$, $w_{x} = 1$ }
1303: \label{C}
1304: \end{center}
1305: \end{figure}
1306:
1307: \begin{figure}[!hp]
1308: \begin{center}
1309: \includegraphics[scale=0.3, angle=270]{C1.ode.ps}
1310: \caption{The phase portrait $(a,x)$ for case {\bf C} $p = w_{x}\rho_{x}$,
1311: $k = 1$, $\rho_{0m}/\rho_{0x} = 1$, $w_{x} = -1$}
1312: \label{C1}
1313: \end{center}
1314: \end{figure}
1315:
1316: \begin{figure}[!hp]
1317: \begin{center}
1318: \includegraphics[scale=0.3, angle=270]{C2.ode.ps}
1319: \caption{The phase portrait $(a,x)$ for case {\bf C} $p = w_{x}\rho_{x}$,
1320: $k = -1$, $\rho_{0m}/\rho_{0x} = 1$, $w_{x} = 1$ }
1321: \label{C2}
1322: \end{center}
1323: \end{figure}
1324:
1325: \begin{figure}[!hp]
1326: \begin{center}
1327: \includegraphics[scale=0.3, angle=270]{C3.ode.ps}
1328: \caption{The phase portrait $(a,x)$ for case {\bf C} $p = w_{x}\rho_{x}$,
1329: $k = -1$, $\rho_{0m}/\rho_{0x} = 1$, $w_{x} = -1$}
1330: \label{C3}
1331: \end{center}
1332: \end{figure}
1333:
1334: \end{document}
1335:
1336: