nucl-th0001008/part1
1: %-------------------------------------------------------------------
2: % 05.01.00 , revised version from 09.02.00; accepted 14.03.00
3: \documentstyle[12pt,epsfig]{article}
4: \topmargin -48pt
5: \def\baselinestretch{1.2}
6: \textwidth  160mm
7: \textheight 235mm
8: \oddsidemargin  2mm
9: \evensidemargin 2mm
10: \begin{document}
11: %----------------------------------------------------------------------
12: \title{Aspects of thermal and chemical equilibration
13: of hadronic matter
14: \thanks{Work supported by BMBF, GSI Darmstadt and DFG.}}
15: \author{E. L. Bratkovskaya, W. Cassing, C. Greiner,\\
16: M. Effenberger, U. Mosel and A. Sibirtsev \\
17: {\normalsize Institut f\"{u}r Theoretische Physik, Universit\"{a}t Giessen}\\
18: {\normalsize 35392 Giessen, Germany}}
19: \date{}
20: \maketitle
21: \begin{abstract}
22: We study thermal and chemical equilibration in 'infinite' hadron matter
23: as well as in finite size relativistic nucleus-nucleus collisions using
24: a BUU cascade transport model that contains resonance and string
25: degrees-of-freedom.  The 'infinite' hadron matter is simulated within a
26: cubic box with periodic boundary conditions.  The various equilibration
27: times depend on baryon density and energy density and are much shorter
28: for particles consisting of light quarks then for particles including
29: strangeness.  For kaons and antikaons the chemical equilibration time
30: is found to be larger than $\simeq$ 40 fm/c for all baryon and energy
31: densities considered.  The inclusion of continuum excitations, i.e.
32: hadron 'strings', leads to a limiting temperature of $T_s\simeq$ 150
33: MeV. We, furthermore, study the expansion of a hadronic fireball after
34: equilibration. The slope parameters of the particles after expansion
35: increase with their mass; the pions leave the fireball much faster then
36: nucleons and accelerate subsequently heavier hadrons by rescattering
37: ('pion wind').  If the system before expansion is close to the limiting
38: temperature $T_s$, the slope parameters for all particles after
39: expansion practically do not depend on (initial) energy and baryon
40: density.  Finally, the equilibration in relativistic nucleus-nucleus
41: collision is considered.  Since the reaction time here is much shorter
42: than the equilibration time for strangeness, a chemical equilibrium of
43: strange particles in heavy-ion collisions is not supported by our
44: transport calculations.  However, the various particle spectra can
45: approximately be described within the blast model.
46: \end{abstract}
47: \vspace{2mm}
48: 
49: \noindent
50: PACS: 25.75.-q, 13.60.Le, 13.60.Rj, 21.65.+f, 24.10.Pa
51: 
52: \noindent
53: Keywords: Relativistic heavy-ion collisions, Meson production, Baryon
54: production, Nuclear Matter, Thermal and statistical models,
55: Equilibration
56: 
57: %---------------------------------------------------------------------
58: \newpage
59: \section{Introduction}
60: 
61: Nucleus-nucleus collisions at relativistic and ultrarelativistic
62: energies are studied experimentally and theoretically to obtain
63: information about the properties of hadrons at high density and/or
64: temperature as well as  about the phase transition to a new state of
65: matter, the quark-gluon plasma (QGP).  In the latter deconfined partons
66: are the essential degrees of freedom that resolve the underlying
67: structure of hadrons \cite{QM}.  Whereas the early 'big-bang' of the
68: universe most likely evolved through steps of kinetic and chemical
69: equilibrium, the laboratory 'tiny bangs' proceed through phase-space
70: configurations that initially are far from an equilibrium phase and
71: then evolve by fast expansion. These 'specific initial conditions' --
72: on the theoretical side -- have lead to a rapid development of
73: nonequilibrium quantum field theory and nonequilibribrium kinetic
74: theory \cite{BotMal90,Henning}.  Presently, semiclassical transport
75: models are widely used as approximate solutions to these theories and
76: practically are an essential ingredient in the experimental data
77: analysis. For recent reviews we refer the reader to Refs.
78: \cite{Ko,Bass,Cass99}.
79: 
80: On the other hand, many observables from strongly interacting systems
81: are dominated by many-body phase space such that spectra and abundances
82: look 'thermal'.  It is thus tempting to characterize the experimental
83: observables by global thermodynamical quantities like 'temperature',
84: chemical potentials or entropy \cite{BM,Satz,Sollfrank,Spieles,Cleymans}.
85: We note, that even the use of macroscopic models like hydrodynamics
86: \cite{Hydro,Rischke} employs as basic assumption the concept of local
87: thermal and chemical equilibrium. The crucial question, however, how
88: and on what timescales a global thermodynamic equilibrium can be
89: achieved, is presently a matter of debate. Thus nonequilibrium
90: approaches have been used in the past to address the problem of
91: timescales associated to global or local equilibration
92: \cite{Rafelski,Cass90,Lang91,Bl93,Brav1,Brav2,Brav3,Solfr99}.  In view
93: of the increasing 'popularity' of thermodynamic analyses a thorough
94: microscopic reanalysis of this questions appears necessary especially
95: for nucleus-nucleus collisions at ultrarelativistic energies that aim
96: at 'detecting' a phase transition to the QGP.
97: 
98: In this paper we study equilibration phenomena in 'infinite' hadronic
99: matter using a microscopic transport model that contains both hadron
100: resonance and string degrees-of-freedom.  With this investigation we
101: want to provide insight into the reaction dynamics by the use of
102: cascade-like models and also point out some of their limitations.  The
103: 'infinite' hadronic matter is modelled by initializing the system
104: solely by nucleonic degrees of freedom through a fixed baryon density
105: and energy density, while confining it to a cubic box and imposing
106: periodic boundary conditions during the propagation in time.  We,
107: furthermore, then study the expansion of the hadronic fireball after
108: equilibration to investigate the changes in hadron spectra during the
109: rapid explosion as well as related equilibration phenomena in realistic
110: nucleus-nucleus collisions for light and heavy systems.
111: 
112: Our paper is organized as follows: In Section 2 we briefly describe the
113: approach employed in our investigations, specify the initial conditions
114: for a finite box with periodic boundary conditions, present our
115: numerical results and extract various (hadronic) equilibration times as
116: well as thermodynamical properties for different initial conditions.
117: Section 3 is devoted to the expansion dynamics of the equlibrated
118: fireball and a discussion of the related physical phenomena. In Section
119: 4 we analyse reactions of colliding finite light and heavy systems and
120: compare our result to a blast model. Section 5 concludes our study
121: with a summary.
122: 
123: %---------------------------------------------------------------------
124: \section{Equilibration and limiting temperature}
125: 
126: To investigate the equilibration phenomena addressed above we perform
127: microscopic calculations using the Boltzmann-Uehling-Uhlenbeck (BUU)
128: model of Refs. \cite{Effe99gam,EffePhD}. This model is based on the
129: resonance concept of nucleon-nucleon and meson-nucleon interactions at
130: low invariant energy $\sqrt{s} \ $ \cite{TeisZP97}, adopting all
131: resonance parameters from the Manley analysis \cite{Manley}.  The high
132: energy collisions -- above $\sqrt{s}$ = 2.6~GeV for baryon-baryon
133: collisions and $\sqrt{s}$ = 2.2~GeV for meson-baryon collisions -- are
134: described by  the LUND string fragmentation model FRITIOF
135: \cite{FRITIOF}. This aspect is similar to that used in the HSD approach
136: \cite{Cass99,Ehehalt,Brat98,Geiss} and the UrQMD code \cite{Bass}. For
137: a detailed description of the underlying model at low energy we refer
138: the reader to Ref.~\cite{EffePhD}.
139: 
140: For later discussions it is essential to realize that the code respects
141: detailed balance only for reactions of the type $1 \leftrightarrow 2 +
142: 3$ and approximately for $1 + 2 \leftrightarrow 3 + 4$
143: \footnote{In the latter case small violations of detailed balance are
144: due to the treatment of $t$-channel and background contributions.}
145: where the numbers $1, \ldots, 4$ are any reaction partners. This
146: implies that in particular at high energies, where the string degrees
147: of freedom with their decay to many ($> 2$) final particles becomes
148: important, detailed balance is violated. We will discuss the
149: consequences of this violation, which is inherent in all such transport
150: codes, at the appropriate points in the following sections.
151: 
152: \subsection{A box with periodic boundary conditions}
153: 
154: In order to study 'infinite' hadronic matter problems we confine the
155: particles in a cubic box with periodic boundary conditions for their
156: propagation similar to a recent box calculation within the UrQMD model
157: \cite{Brav1}. We specify the initial conditions, i.e. baryon density
158: $\rho$, strange particle density $\rho_S$ and energy density
159: $\varepsilon$ as follows: first the initial system is fixed to $N_p=80$
160: protons and $N_n=80$ neutrons, which are randomly distributed in a
161: cubic box of volume $V$. The 3-momenta $\vec p_i$ of the nucleons in a
162: first step are randomly distributed inside a Fermi-sphere of radius
163: $p_F$ = 0.26~GeV/c (at $\rho_0$) and in a second step  boosted by $\pm
164: \beta_{cm}$ by a proper Lorentz transformation. Thus the initial baryon
165: density $\rho$ is fixed as $\rho=A/V$, $A=N_p+N_n$. The strange
166: particle density is set to zero as in related heavy-ion experiments
167: while the energy density is defined as $\varepsilon = E/V$, where $E$
168: is the total energy of all nucleons
169: \begin{eqnarray}
170: E=\sum\limits_i^A\sqrt{p_i^2+m_N^2}.
171: \label{energy_i}
172: \end{eqnarray}
173: The  boost velocity $\beta_{cm}$ is related to the initial energy density
174: $\varepsilon$ (excluding Fermi motion) as
175: \begin{eqnarray}
176: \beta_{cm}=\sqrt{1-{\rho^2 m_N^2\over \varepsilon^2} }
177: \label{betta}\end{eqnarray}
178: using
179: \begin{eqnarray}
180: \varepsilon = \gamma_{cm} \rho m_N \label{vargam}
181: \end{eqnarray}
182: with $\gamma_{cm}= 1/\sqrt{1-\beta_{cm}^2}$. Recall that $\rho_0 m_N \simeq
183: 0.15$~GeV/fm$^3$ so that an energy density $\varepsilon \simeq
184: 1.5$~GeV/fm$^3$ at density $\rho_0$ corresponds to $\gamma_{cm}\simeq
185: 10$, i.e. the SPS energy $T_{lab}\simeq 185$~A$\cdot$GeV. We thus start
186: with a 'true' nonequilibrium situation in order to mimique the  initial
187: stage in a relativistic heavy-ion collision. The initial phase
188: represents two interpenetrating, (ideally) infinitely extended fluids
189: of cold nuclear matter.
190: 
191: We now propagate all particles in the box in the cascade mode (without
192: mean-field potentials) using periodic boundary conditions, i.e.
193: particles moving out of the box are reinserted at the opposite side
194: with the same momentum. The phase-space distribution of particles then
195: can change due to elastic collisions,  resonance and string production
196: and their decays to mesons and baryons again. We recall that we include
197: all baryon resonances up to an invariant mass of 2 GeV and meson
198: resonances up to the $\phi$-meson. According to the initial conditions
199: for $\varepsilon$ and $\rho$ the factor $\gamma_{cm}$ in (\ref{vargam})
200: determines if strings are excited in the very first collisions. This is
201: the case for $\gamma_{cm} > 1.4$ where the early equlibration stages
202: are dominated by string formation and decay.
203: 
204: \subsection{Chemical equilibration}
205: 
206: Figure \ref{Fig1} shows the time evolution of the various particle
207: abundances (nucleons $N$,  $\Delta$,  $\Lambda$, $\pi$, $\eta$,  $K^+$
208: and $K^-$ mesons) for density $\rho=\rho_0$ (left panel) at different
209: energy densities $\varepsilon =1.1, 0.52$ and 0.22 GeV/fm$^3$ and for
210: $\rho=3\rho_0$ (right panel) at $\varepsilon =3.4, 1.57$ and
211: 0.66~GeV/fm$^3$.  These initial conditions correspond to bombarding
212: energies $T_{lab}$ per nucleon of roughly 100, 20 and 2 A$\cdot$GeV,
213: respectively. In Fig.~\ref{Fig1} (as well as in
214: Figs.~\ref{Fig2},\ref{Fig3}) we count all particles which are
215: 'hadronized', i.e.\ produced by string decay after a  formation time of
216: $\tau_F=0.8$~fm/c in their rest frame.
217: 
218: After several fm/c the number of nucleons decreases due to inelastic
219: collisions that produce either baryon resonances or additional mesons.
220: The number of $\Delta$-resonances grows up to a maximum in a few fm/c,
221: since a lot of $\Delta$'s are produced in the first $NN$ collisions;
222: their number subsequently decreases with time due to their decay and
223: excitation of further resonances or due to reabsorption.  The numbers
224: of $\pi$'s  and $\eta$'s increase very fast and reach the equilibrium
225: value within a few fm/c whereas the strange particles ($K^+,
226: K^-,\Lambda$) require a much longer time for equilibration.
227: 
228: In Fig. \ref{Fig2} we present the time evolution of the particle ratios
229: $\pi/N$, $\Delta/N$, $\Lambda/N$, $K^+/\pi^+$, $K^-/\pi^-$, $\eta/\pi$
230: for density $\rho=\rho_0$ at energy densities $\varepsilon = 1.1, 0.52$
231: and 0.2 GeV/fm$^3$, while Fig.~\ref{Fig3} shows the same particle ratios
232: for density $\rho=3\rho_0$ at energy densities $\varepsilon = 3.4,
233: 1.57$ and 0.66 GeV/fm$^3$, respectively. The left panels in both plots
234: correspond to the full time scale as in Fig.~\ref{Fig1} (up to 1000 fm/c),
235: whereas the right panels present in more detail the initial phase (up
236: to 30 fm/c).  We use the same scale for the $y$-axis on the right and
237: left panels, so one can easily see that the $\pi/N$, $\Delta/N$ ratios
238: reach the equilibrium values very fast especially at low energy
239: density since the string degrees of freedom here play a minor role and
240: pion production basically emerges through $\Delta$ resonance decay.
241: 
242: The meson-pion ratios ($K^+/\pi^+$, $K^-/\pi^-$, $\eta/\pi$) at high
243: energies show a decrease in the first few fm/c and then an increase
244: again up to the equilibrium values. This is due to the fact that the
245: bulk of the strange mesons is produced very early (at high energy
246: density) through string formation and decay whereas most of the pions
247: appear later, with a delay of several fm/c, as a result of the decay of
248: heavy vector mesons (e.g. $\rho$ and $\omega$). From the right panels
249: of Figs.~\ref{Fig2} and \ref{Fig3} one can see that in the initial
250: stage the particle ratios containing strange to nonstrange particles --
251: $K^+/\pi^+$, $K^-/\pi^-$, $\Lambda/N$ -- are still far off chemical
252: equilibrium for all energies and densities and the equilibration takes
253: up to a few hundred fm/c depending on the energy and baryon density.
254: 
255: For the higher energies the initial particle production proceeds via
256: the formation and decay of string excitations. This leads in particular
257: to a very early onset of strange particles (mainly kaons and hyperons)
258: within the first fm/c either due to the initial strings or due to
259: secondary or ternary baryon-baryon, meson-baryon and meson-meson
260: induced string-like interactions (see the right panels of Figs.
261: \ref{Fig2} and \ref{Fig3}). In Ref. \cite{Geiss} it was shown that
262: these early secondary and ternary reactions can contribute up to about
263: 50 $\%$ of the total strange particles obtained in a Pb~+~Pb reaction at
264: CERN SPS energies and thus explain the factor of 2 in the observed
265: relative strangeness enhancement compared to p+p reactions. This,
266: however, does not imply that chemical equilibrium for the dominant
267: strange particles has been achieved in this reaction, as our analysis
268: clearly shows. In the later stages, when the system has become, more or
269: less, isotropic in momentum space, strange particles can only be
270: further produced by low energy hadronic reactions, which, however, have
271: a considerable threshold and are thus strongly suppressed. This
272: explains the long chemical equilibration times for the strange
273: particles first demonstrated by Koch, M\"uller and Rafelski
274: \cite{Rafelski}.
275: 
276: In order to define an overall chemical equilibration time
277: we perform a fit to
278: the particle abundances $N(t)$ for pions and kaons as
279: \begin{eqnarray}
280: N(t) = N_{eq} \left(1 - \exp(-t/\tau_{eq})\right),
281: \label{taueq}\end{eqnarray}
282: where $N_{eq}$ is the equilibrium limit.  The equilibration time $\tau_{eq}$
283: thus corresponds to the time $t$ when $\simeq 63$\% of $N_{eq}$ is achieved.
284: 
285: Figure \ref{Fig4} shows the equilibration time $\tau_{eq}$ versus
286: energy density for $\pi$ and $K^+$ mesons at different baryon densities
287: of $1/3\rho_0, \rho_0, 3\rho_0$ and $6\rho_0$. We find that the
288: equilibration time for pions scales as $\tau_{eq}^\pi \sim 1/\rho$ or
289: $\Gamma_\pi \sim \rho$, thus we present the curve only for baryon
290: density $\rho_0$.  Whereas $\tau_{eq}^\pi$ slowly grows with
291: energy-density, $\tau_{eq}^K$ falls steeply with $\varepsilon$. This
292: marked difference is due to the fact that, on one hand, the kaon
293: production rate increases dramatically with $\sqrt{s}$ whereas that of
294: the pions, on the other hand, is more flat. With increasing energy thus
295: more strange particles are produced through strings especially from the
296: primary collisions with high $\sqrt{s}$ and the chemical equilibration
297: is achieved faster.
298: 
299: In Fig.~\ref{Fig4} we have considered an 'ideal' situation, i.e. hadron
300: matter at fixed energy and baryon density.  In realistic heavy-ion
301: collisions the system goes through the different stages due to
302: interactions and expansion. However, as follows from Fig. \ref{Fig4},
303: the equilibration time for strangeness is larger than 40~fm/c for all
304: energy and baryon densities. Thus in realistic nucleus-nucleus
305: collisions the chemical equilibration of strange particles requires
306: also a time above~40 fm/c which is considerably larger than the actual
307: reaction time of a few 10 fm/c or less (cf. Section 4).
308: 
309: The particle abundances used to extract $\tau_{eq}$ in Fig. \ref{Fig4}
310: have been calculated without any in-medium potentials.
311: In fact, the introduction of attractive potentials (especially for
312: $K^-$) will lower the hadronic thresholds and thus increase the
313: scattering rate between strange and nonstrange hadrons, whereas the
314: $K^+$ feels some repulsive potential and the trend goes in the opposite
315: way.  According to our calculations such in-medium modifications (in
316: line with Ref. \cite{Cass99}) give a correction to the $K^+$
317: equilibration times by atmost 10 \% and shortens the $K^-$
318: equilibration times up to 20 \% at density $\rho_0$.
319: 
320: \subsection{Thermal equilibration and limiting temperature}
321: 
322: In this subsection we investigate the approach to thermal equilibration.
323: This is initially driven by the very early string phase on the momentum
324: equilibration of the hadronic degrees of freedom, when the system is
325: still very far from equilibrium and the energy density is sufficiently
326: high. This one can see by looking at the quadrupole moment
327: ${<}Q_2{>}=<2p_z^2-p_x^2-p_y^2>$ of the momentum distribution of all hadrons
328: involved. In the left panel of Fig.~\ref{Fig5} we present the time
329: evolution of the quadrupole moment $<Q_2>$ for density $\rho=\rho_0$ at
330: energy densities $\varepsilon=0.22$, 0.3, 0.52, 0.8, 1.1 and 1.6
331: GeV/fm$^3$. In order to take into account the string contributions we
332: have counted here all particles even within the formation time.  The
333: thin solid lines indicate exponential fits of the form
334: \begin{eqnarray}
335: <Q_2>(t) \simeq A_1\exp(-t/\tau_{short})+ A_2\exp(-t/\tau_{long})
336: \label{Q2}\end{eqnarray}
337: with two equilibration times $\tau_{short}$ and $\tau_{long}$.
338: 
339: The right panel of Fig.~\ref{Fig5} shows $\tau_{short}$ and
340: $\tau_{long}$ versus energy density $\varepsilon$. Whereas
341: $\tau_{short}\simeq 5$~fm/c is roughly independent on $\varepsilon$ the
342: 'hadronic' equilibration time $\tau_{long}$ increases with energy
343: density. These results have to be interpreted as follows: in the
344: initial nonequilibrium phase the string degrees of freedom are excited
345: and decay according to many-body phase on a short time scale
346: $\tau_{short}$. The string decays reduce the initial quadrupole moment
347: (at high energy density) in time by a significant factor of about
348: $3-4$. One can understand the result obtained for $\tau_{short}$ in a
349: rather simple way.  Due to our prescription of the initialization of
350: the system the first strings on average are formed  after the time
351: $\tau_{coll} \approx 1/((\rho/2) \sigma_{NN } \langle v_{NN} \rangle)
352: \approx 3-4$~fm/c for $\rho = \rho_0$.  The strings then decay within
353: their formation time $\tau_F \approx 0.8$~fm/c  giving rise to a
354: significant production of transversal momentum.  One should point out,
355: that according to these arguments $\tau_{short} $ approximately scales
356: like $1/\rho $.  Due to Lorentz contraction $\tau_{short}$ is thus
357: considerably smaller in a real heavy-ion collision. Hence, string
358: decays provide a very efficient source for a strong decrease in
359: longitudinal momentum and production of transverse momentum in the very
360: early stage of an ultrarelativistic heavy-ion collision. A decrease
361: (increase) of the formation time $\tau_F$ to 0.5 fm/c (1.5 fm/c)
362: changes $\tau_{short}$ on the scale of 20\%, only.
363: 
364: After string decay, however, the emerging hadronic system still has
365: significantly larger longitudinal than transverse momenta -- the ratio
366: increases with energy density $\varepsilon$ -- and low energy hadronic
367: reactions are less effective in transfering longitudinal to transverse
368: momentum or simply in production of mesons.  This explains the increase
369: of $\tau_{long}$ with $\varepsilon$ in simple terms.
370: 
371:  From the above analysis it follows that after typical relaxation times
372: of $\tau_{short} \approx$ 5~fm/c the momentum unisotropy
373: of hot and dense matter has dropped to ${\rm e}^{-1}$ such that one might
374: describe the system by simple global thermodynamical variables like
375: temperature etc. This thermal equilibrium has to be contrasted with the
376: chemical equilibrium which -- as we have shown in the preceding
377: subsection -- is reached only after much longer times ($\geq 40$ fm/c
378: for strange particles, for example).
379: 
380: For the equilibrated system we can extract a temperature $T$ by fitting the
381: particle spectra with the Bolzmann distribution
382: \begin{eqnarray}
383: {d^3N_i\over dp^3} \sim \exp(-E_i/T),
384: \label{Boltz}\end{eqnarray}
385: where $E_i=\sqrt{p_i^2+m_i^2}$ is the energy of particle $i$.  We note
386: that at the temperatures of interest here, the Bose and Fermi
387: distributions are practically identical to a Boltzmann distribution. We
388: find that in equilibrium  the spectra of all particles can be
389: characterized by one single temperature $T$. This is demonstrated in
390: Fig.~\ref{Fig6} where we show the spectra of nucleons ($N$), pions ($\pi$)
391: and kaons ($K^+$) as a function of the kinetic energy $E-m$ for
392: $\rho=\rho_0$ at energy densities $\varepsilon=0.52, 0.8$ and 1.6
393: GeV/fm$^3$ (left panel) and for $\rho=3\rho_0$ at energy densities
394: $\varepsilon=0.66, 1.57$ and 2.85 GeV/fm$^3$ (right panel). Here we have
395: averaged the spectra from 950 fm/c to 1000 fm/c in order to decrease the
396: numerical fluctuations. The spectra of $N, \pi, K^+$ here can be
397: fitted with a single temperature $T$ which increases with the energy
398: density $\varepsilon$ for both baryon densities $\rho_0$ and $3\rho_0$.
399: We note explicitly that the slope of the equilibrium particle spectra
400: does not depend on the formation time $\tau_F$.
401: 
402: 
403: In Fig.~\ref{Fig7} we display the energy density $\varepsilon$ versus
404: temperature $T$ for different baryon densities $\rho$:  $1/3\rho_0$
405: (open down triangles), $\rho_0$ (full squares), $3\rho_0$ (full dots),
406: $6\rho_0$ (full up triangles). In order to compare calculations for
407: different baryon densities we have subtracted the baryon energy density
408: at rest, i.e. $\simeq m_N\rho$ (except for Fermi motion). As seen from
409: Fig.~\ref{Fig7} the temperature grows with energy density up to a
410: limiting value reminiscent of a 'Hagedorn' temperature \cite{Hagedorn}.
411:  From our detailed investigations we obtain for the limiting temperature
412: $T_s \simeq 150\pm 5$~MeV which practically does not depend on baryon
413: density.  Such a singular behavior of $\varepsilon(T)$ for $T\simeq
414: T_s$ has also been found in the box calculations in Ref.~\cite{Brav1}
415: for $\rho=\rho_0$.  Our limiting temperature is slightly higher than
416: that in Ref.~\cite{Brav1} ($T_s = 130 \pm 10$~MeV)  due to the different
417: number of degrees of freedom; the model \cite{Brav1} contains more
418: resonances and uses a different threshold for string excitations.
419: Thus, there is some phenomenological sensitivity to the hadronic zoo of
420: particles and string thresholds employed in the model.
421: 
422: In Fig. \ref{Fig8} we show the excitation function for the ratio of
423: string energy density to the energy density of the whole system
424: $\varepsilon_{string}/\varepsilon$ at $\rho=\rho_0$ when the system has
425: equilibrated for long times.  If the equilibrated system is very dense,
426: lower energy strings are still continuously being excited and thus --
427: because of their subsequent decay -- the strings constitute a
428: stationary portion of the total energy of the system.  The relative
429: ratio in the energy density increases with $\varepsilon$ up to a
430: saturation value of $\simeq 16$\% and then stays essentially constant. This
431: reflects that the system reaches a limiting temperature, since the
432: relative amount of string excitations compared to resonance excitations
433: does not change any more, whereas the number of strings as well as the
434: number of hadrons produced increases with $\varepsilon$. This fact one
435: might have guessed since the string production rate in equilibrium
436: depends only on the temperature $T$ characterizing the Bose/Fermi
437: distributions in the collision terms. In addition, this constant
438: fraction, of course, also intrinsically depends on the excitation
439: threshold and on the chosen decay (or formation) time $\tau_F$ of the
440: strings.
441: 
442: As pointed out above, the string degrees of freedom play an essential
443: role for particle production at high bombarding energies since they
444: describe the continuum excitations of the system.  The number of
445: strings created is especially high at the first stages of the
446: collision, when the energy of baryon-baryon interactions is close to
447: the initial energy $\sqrt{s}$.  It decreases with time to some constant
448: value which corresponds to the equilibrium state. Because of this
449: string-dominance one now has to worry about possible consequences of a
450: violation of detailed balance for these degrees of freedom. As already
451: pointed out earlier, all hadronic cascade-type approaches use the
452: phenomenological string picture in order to describe quantitatively
453: energetic (soft or semi-hard) inelastic reactions above some specified
454: $\sqrt{s}$-threshold. In such binary hadronic reactions typically many
455: hadronic particles and resonances are produced, the number depending on
456: the incident energy $\sqrt{s}$. The `back reaction' of these particles
457: produced from decay of an excited string (or two strings in the LUND
458: model) leading to the formation of only two energetic hadrons again is
459: not considered as it is statistically suppressed  and difficult to
460: describe. On the other hand, in an `infinite' matter calculation these
461: back reactions have to be taken into account in order to
462: allow for the principle of detailed balance. This is not done here as
463: it is technically difficult to handle; it thus represents a potential
464: `Achilles heal' in a thermodynamic analysis. However, for simulating a
465: heavy-ion collision this deficiency is not of any major importance
466: since the excitation of strings happens in the first moment of the
467: reaction when the phase space is still widely open and no back reaction
468: can occur.
469: 
470: \subsection{Comparison to the statistical model}
471: 
472: In order to investigate the equilibrium behavior of hadron matter we
473: also compare our transport (box) calculations with a simple Statistical
474: Model (SM) for an Ideal Hadron Gas (IHG) where the system is described
475: by a grand canonical ensemble of non-interacting fermions and bosons in
476: equilibrium at temperature $T$.  All baryon and meson species
477: considered in the transport model \cite{Effe99gam} also have been
478: included in the statistical model. Our main objective here is to
479: compare our results with the Hagedorn bootstrap picture of hadronic
480: matter \cite{Hagedorn}.
481: 
482: We recall that in  the SM particle multiplicities $n_i$ and energy
483: densities $\varepsilon_i$ are given by
484: \begin{eqnarray}
485: && n_i ={g_i \over (2\pi \hbar)^3} \int\limits_0^\infty
486: {4\pi p^2 dp \over \exp\left[(E_i - B_i\mu_B - S_i \mu_S)/T\right]\pm 1},
487:  \label{Nth} \\
488: && \varepsilon_i ={g_i \over (2\pi \hbar)^3}
489: \int\limits_0^\infty {4\pi E_i p^2 dp \over
490: \exp\left[(E_i - B_i\mu_B - S_i \mu_S)/T\right]\pm 1},
491: \label{Enth}
492: \end{eqnarray}
493: where $E_i = \sqrt{p^2+m_i^2}$ is the energy of particle $i$, $B_i$ is
494: the baryon charge, $S_i$ is the strangeness, and $g_i$ is the
495: spin-isospin degeneracy factor.  In Eqs. (\ref{Nth}),(\ref{Enth})
496: $\mu_B$ and $\mu_S$ are the baryon and strangeness chemical potentials.
497: Here we neglect the electric chemical potential ($\mu_n=\mu_p=\mu_B$)
498: since we consider an isospin symmetric system.  Note, however, that in
499: realistic collisions of heavy-ions (like Au~+~Au) this reduction is no
500: longer fully appropriate.
501: For particles with finite spectral width we include in Eqs.
502: (\ref{Nth}),(\ref{Enth}) the spectral functions $\rho_i(m) $ with the
503: same parametrization for the width as in the transport model,
504: \begin{eqnarray}
505: && n_i ={g_i \over (2\pi \hbar)^3} \int \rho_i(m) dm
506: \int\limits_0^\infty
507: {4\pi p^2 dp \over \exp\left[(E_i - B_i\mu_B - S_i \mu_S)/T\right]\pm 1},
508: \label{Nthsf} \\
509: && \varepsilon_i ={g_i \over (2\pi \hbar)^3} \int \rho_i(m) dm
510: \int\limits_0^\infty {4\pi E_i p^2 dp \over
511: \exp\left[(E_i - B_i\mu_B - S_i \mu_S)/T\right]\pm 1}.
512: \label{Enthsf}\end{eqnarray}
513: 
514: 
515: The energy density $\varepsilon$, baryon density $\rho$ and strange
516: density of the hole system in equilibrium then given as
517: \begin{eqnarray}
518: && \varepsilon =  \sum\limits_i \varepsilon_i (T,\mu_B,\mu_S)  \label{3eq_en} \\
519: && \rho = \sum\limits_i B_i \ n_i (T,\mu_B,\mu_S) \label{3eq_rhoB} \\
520: && \rho_S = \sum\limits_i S_i \ n_i (T,\mu_B,\mu_S)
521: \, \equiv \, 0 . \label{3eq_rhoS}
522: \end{eqnarray}
523: 
524: As 'input' for the SM we use the same $\varepsilon, \rho$ and $\rho_S$
525: as in the box calculations and  we obtain the thermodynamical
526: parameters -- $T, \mu_B, \mu_S$ -- by solving the system of nonlinear
527: equations (\ref{3eq_en}),(\ref{3eq_rhoB}) and (\ref{3eq_rhoS}).
528: 
529: Within the SM we find that the temperature increases continuously
530: with energy density since the continuum excitations, i.e. the string degrees
531: of freedom, are not included (full dots  in Fig.~\ref{Fig9}),
532: whereas the box calculation with strings gives the limiting temperature
533: (full squares in Fig.~\ref{Fig9}).  Both curves in Fig.~\ref{Fig9}
534: have been calculated for density $\rho_0$.
535: 
536: To reproduce qualitatively our box result within the SM we have to include
537: continuum excitations in the statistical model, i.e.
538: a Hagedorn mass spectrum for strings as defined by \cite{Hagedorn}
539: \begin{eqnarray}
540: \rho^{str}(m) = {\rho_0^{str} \over m^3} \exp(m/T_H),
541: \label{Hagden}\end{eqnarray}
542: where $T_H$ denotes the 'Hagedorn' temperature. For $T_H$ we use the
543: temperature $T_s$ as obtained from the box calculations, i.e.
544: $T_H=T_s\simeq 150$~MeV. In (\ref{Hagden}) $\rho_0^{str}$ is a fit
545: parameter additionally to  $T, \mu_B$ and $\mu_S$ to reproduce
546: $\varepsilon(T)$ from the box calculations. The string multiplicities
547: $n_i^{str}$  are given by
548: \begin{eqnarray}
549: && n_i^{str} ={1 \over (2\pi \hbar)^3} \int\limits_{m_{min}}^\infty
550: \rho_i^{str}(m) dm \int\limits_0^\infty
551: {4\pi p^2 dp \over \exp\left[(E_i - B_i\mu_B - S_i \mu_S)/T\right]\pm 1},
552: \label{Nstr} \end{eqnarray}
553: where the lowest mass in the string excitation ($m_{min}$) is defined
554: by the string threshold in the transport model:  $m_{min}=2.6-m_N$~GeV
555: for baryon strings and $m_{min} =2.2$~GeV for meson strings.  In our
556: transport model we include the following strings $i$:  baryon strings
557: $B=1,S=0,-1,-2,-3$, anti-baryon strings $B=-1, S=0,1,2,3$, meson
558: strings $B=0, S=0,1$ and anti-meson strings $B=0, S=-1$.
559: 
560: Before going over to the actual analysis we point out that the limiting
561: temperature $T_s$ from our string model involves somewhat different
562: physics assumptions than the Hagedorn model at temperature $T_H$. $T_s$
563: should not really be identified with the 'Hagedorn' temperature $T_H$,
564: though close similarities exist. In the Hagedorn picture and for
565: temperatures close to $T_H$ the abundance of `normal' hadrons or known
566: resonances stays constant with increasing energy density whereas the
567: number and energy density of the (hypothetical) bootstrap excitations
568: diverges for $T\rightarrow T_H$. The Hagedorn model thus assumes
569: `particles' of mass $m\to\infty$ to be populated for $T\to T_H$, that
570: dynamically can be formed in collisions of high mass hadrons for
571: $t\to\infty$.  In contrast, our string model does not include energetic
572: string-string interactions that might produce more massive strings.
573: (There exist some phenomenological recipes how to incorporate such
574: interactions \cite{Sailer}.) The 'high mass' strings decay to hadrons
575: and, because of the detailed balance problem discussed in the last
576: subsection, are only repopulated by binary hadron-hadron or
577: hadron-string interactions, so that their internal energy is limited
578: and the low-energy hadronic degrees of freedom are overpopulated. This
579: leads to the saturation of string-energy to total energy (observed in
580: Fig.~\ref{Fig8}) to a value of $\simeq 0.16$ in contrast to the value
581: of 1 in the Hagedorn model.
582: 
583: This, however, does not imply a fundamental inconsistency for the
584: overall properties of the system. In perfect chemical equilibrium, like
585: in the Hagedorn model, more strings (or hypothetical resonances) would
586: be excited which, for lower temperatures (e.g.\ in a nearly isentropic
587: expansion of the system like in heavy-ion collisions), would
588: immediately decay into a large number of hadronic particles. The
589: violation of detailed balance in our case thus physically describes an
590: overpopulation of hadronic particles only in stationary equilibrium.
591: The important point, however, is the observation that in either
592: description the system at equilibrium can not exceed the critical
593: temperature $T_s$.
594: 
595: As seen in Fig.~\ref{Fig9} we achieve agreement of the extended SM and
596: our box calculations from Fig.~\ref{Fig8} by choosing $T_H \approx T_s$
597: in Eq.\ (\ref{Hagden}). In addition, from the extended SM we can also
598: define thermodynamical parameters such as the baryon chemical potential
599: $\mu_B$. In Fig.~\ref{Fig10} we present the resulting $T-\mu_B$
600: correlation, i.e. temperature $T$ versus baryon chemical potential
601: $\mu_B$, at fixed baryon densities (in the box calculations) of $\rho =
602: 1/3\rho_0, \rho_0$ and various energy densities. The open triangles and
603: squares (connected by the dashed lines) show the result of the SM
604: without strings at densities $1/3\rho_0$ and $\rho_0$, respectively,
605: whereas the full triangles and squares (connected by the solid lines)
606: correspond to the thermodynamical fit of the box calculations (at
607: $1/3\rho_0$ and $\rho_0$) including string excitations.
608: The errorbars indicate the uncertainty in the extraction of $\mu_B$ in
609: the SM; they become larger when the system is closer to $T_H$ due to
610: the divergence in the energy density integral (\ref{3eq_en}). The
611: arrow at $\mu_B=0$  indicates the temperature $T_s=150$~MeV from
612: our box calculations. The full dots in Fig.~\ref{Fig10} correspond to
613: chemical freeze-out points extracted in a thermodynamical model from
614: hadron abundances \cite{BM}; the open dots are the thermal
615: freeze-out points from the momentum spectra of hadrons and two-particle
616: correlations as taken from Ref.~\cite{HeinzQM97}.
617: 
618: Our calculations here are for nuclear matter densities $1/3\rho_0$ and
619: $\rho_0$ whereas the freeze-out points have been extracted from heavy-ion
620: data; the comparison thus can be only qualitative.  However, one can see
621: the general tendency: if the continuum excitations (strings) are not
622: included in the thermodynamical analysis, one can 'extract' much larger
623: temperatures at high energy density simply due to the limited number of
624: degrees of freedom involved in the model analysis. In this respect our
625: box result  is more in line with the thermal ('kinetic') freeze-out
626: analysis from Ref.~\cite{HeinzQM97} than with the thermodynamical
627: analysis from Ref.~\cite{BM} that is based on particle ratios and thus
628: on chemical freeze-out. The point to make is that
629: at higher temperatures, like e.g. the ones obtained
630: for a `chemical' freeze-out in Ref.~\cite{BM},
631: the consideration of continuum excitations does make a thermodynamical
632: analysis much less certain than at lower temperatures, like e.g.
633: at `thermal' (or kinetic) freeze-out as in Ref. ~\cite{HeinzQM97},
634: where the continuum excitations do not play any significant role.
635: 
636: In this context we have to mention, furthermore, that a combined
637: experimental analysis of particle spectra and HBT radii favors
638: even lower freeze-out temperatures (below 100 MeV \cite{Nix,Heinz99}).
639: For these freeze-out conditions the pion density (for fixed charge)
640: drops below $\sim 10^{-2}$~fm$^3$, i.e. the average distance between
641: two pions (of different charge) becomes large than $\sim 4.6$~fm,
642: which in turn is large compared to their classical interaction radius
643: $r_I=\sqrt{\sigma_{\pi\pi}/\pi}$ at all relative momenta between the
644: two pions.  Since thermal freeze-out temperatures of 90-100 MeV at SPS
645: energies can be considered as a lower bound, the 'experimental' points
646: in Fig.~\ref{Fig10} have to be taken with care.
647: 
648: 
649: %---------------------------------------------------------------------
650: \section{Expanding hadronic fireballs}
651: 
652: In realistic nucleus-nucleus collisions the system rapidly expands
653: after the possible formation of a hot hadronic fireball.  The final
654: hadronic spectra can be changed substantially during this expansion
655: phase, i.e.  the temperature extracted from the experimentally observed
656: slopes of the spectra also contains information about the nuclear
657: expansion dynamics.
658: 
659: To investigate the expansion of the hadronic fireball we initialize the
660: system in a box with periodic boundary conditions -- as described above
661: -- and propagate the system up to 500 fm/c, when equilibrium is
662: reached. Afterwards we let the system expand without boundary
663: conditions. Even though this is an idealized description of the
664: expansion phase during a heavy-ion collision we hope to learn from this
665: scenario how the expansion stage changes the picture of perfect thermal
666: equilibrium (for an analysis of an actual collision see the discussion
667: in the following section).
668: 
669: In Fig.~\ref{Fig11} we present the time evolution of the various
670: particle abundances (nucleons $N$,  $\Delta$,  $\Lambda$, $\pi$, $K^+$
671: and $K^-$ mesons) during the expansion for density $\rho=\rho_0$ (left
672: panel) at different energy densities $\varepsilon =0.22, 0.3$ and 1.1
673: GeV/fm$^3$ and for density $\rho=1/3\rho_0$ at $\varepsilon =0.84$
674: GeV/fm$^3$ (upper part in the right panel), for $\rho=\rho_0$ at
675: $\varepsilon =1.6$ GeV/fm$^3$ (middle part in the right panel) and for
676: $\rho=3 \rho_0$ at $\varepsilon =3.4$ GeV/fm$^3$ (lower part in the
677: right panel).
678: The number of stable particles ($N, \Lambda, \pi, K^+, K^-$) increases
679: during the expansion up to some asymptotic value due to string and
680: heavy resonance decay as well as inelastic interactions.
681: One can see that the asymptotic values are reached after a few 10 fm/c from
682: the beginning of the expansion (depending on the initial
683: energies and baryon densities) which is comparable to the actual
684: reaction time in heavy-ion collisions (cf. Section 4).
685: 
686: In Fig.~\ref{Fig12}  we show the spectra of nucleons ($N$), pions
687: ($\pi$) and kaons ($K^+$) versus the kinetic energy $E-m$ for
688: $\rho=\rho_0$ at energy densities $\varepsilon=0.22, 0.3$ and
689: 1.1~GeV/fm$^3$  before the expansion -- averaged over time from 450
690: fm/c to 500~fm/c -- (left panel) and after the expansion -- averaged from
691: 580 fm/c to 600~fm/c -- (right panel). For completeness in
692: Fig.~\ref{Fig13} we present the result for $\rho=1/3\rho_0$ at
693: $\varepsilon=0.84$ GeV/fm$^3$ (upper part), for $\rho=\rho_0$ at
694: $\varepsilon=1.6$ GeV/fm$^3$ (middle part) and for $\rho=3\rho_0$ at
695: $\varepsilon=3.4$ GeV/fm$^3$ (lower part). In the left panels the
696: systems are in equilibrium; the $N, \pi, K^+$ spectra show a common
697: temperature $T$ whereas after the expansion the slopes of the particle
698: spectra are different; the nucleon  spectra are much harder than the
699: pion spectra, i.e. the apparent temperature of particles (after the
700: expansion) increases with the mass $m$.  This effect is illustrated in
701: Fig.~\ref{Fig14}, where we show the apparent slope $T$ versus $m$ for
702: $\pi, K^+, N$ for $\rho=\rho_0$ at different energy densities:
703: $\varepsilon = 0.2$~GeV/fm$^3$ (full up triangles), $\varepsilon =
704: 0.3$~GeV/fm$^3$ (full squares), $\varepsilon = 1.1$~GeV/fm$^3$ (full
705: dots), $\varepsilon = 1.6$~GeV/fm$^3$ (full diamonds); and for
706: $\rho=3\rho_0$ at $\varepsilon=3.4$~GeV/fm$^3$ (open down triangles).
707: The arrow indicates the limiting temperature $T_s=150$~MeV before the
708: expansion.  One can see from Figs.~\ref{Fig12}--\ref{Fig14} that the
709: 'expansion' temperature of particles increases also with the energy
710: density.  However, if during the equilibration phase the system reaches
711: $T_s$, the 'expansion' temperatures for different particles show a
712: universal behaviour, i.e. practically do not depend on the energy
713: $\varepsilon$ as well as on the baryon density $\rho$. This phenomenon
714: is due to the fact that close to $T_s$ the initial hadron velocity
715: distributions, reflected in the particle momentum profile, become
716: similar for all $\varepsilon$ and $\rho$ in equilibrium.
717: 
718: In order to investigate the origin of the enhancement in the particle slope
719: during the expansion we have performed several illustrative calculations:
720: at 500 fm/c -- after the system has achieved equilibrium -- we
721: i) let all resonances and strings decay; in this case we find that the
722: slopes do not change as compared to the equilibrium phase,
723: ii) we let the system expand without interactions (allowing only decays)
724: and find that the slopes slightly decrease in comparison to the
725: equilibrium phase.
726: Both examples indicate that the slope enhancement stems basically from
727: multiple interactions of the particles in the initial stages of
728: the expansion phase.
729: 
730: For analyzing the expansion flow phenomena we have performed a fit of
731: the particle spectra (after expansion) using the blast model of Siemens
732: and Rasmussen \cite{Siemens}. In this model all particle spectra are
733: described by a universal formula with common thermal freeze-out
734: parameters, i.e. a temperature $T$ of the fireball and a radial-flow
735: velocity $\beta$:
736: \begin{eqnarray}
737: {d^3N_i\over dp^3} = A_i \exp\left(-{\gamma E_i\over T}\right)
738: \left[{\sinh\alpha_i\over \alpha_i} \left(\gamma+{T\over E_i}\right) -
739: {T\over E_i}\cosh\alpha_i\right], \label{flow}
740: \end{eqnarray}
741: where $\gamma=(1-\beta^2)^{-1/2}, \ \alpha=\gamma \beta p_i/T$. Here
742: $E_i, p_i$ are the total energy and momentum of the considered particle
743: $i$ while $A_i$ are normalization factors.
744: 
745: We now try to describe the final particle spectra after the expansion
746: by Eq.~(\ref{flow}) with common freeze-out parameters $T$ and $\beta$.
747: In Fig.~\ref{Fig15} we show the result of our least-squares fit, using
748: the MINUIT method \cite{MINUIT}, for the energy densities $\varepsilon
749: =0.3$~GeV/fm$^3$ (upper part) and 1.1~GeV/fm$^3$ (lower part) at
750: $\rho=\rho_0$.  The left panel shows the contour plots for the
751: parameter errors in the $T-\beta$ plane (for the $\chi_{optimal}^2+1$
752: level); the dot-dashed lines stand for
753: nucleons ($N$), the solid lines for pions ($\pi$) and the dashed lines
754: for kaons ($K^+$). The full symbols indicate the 'best' values for $T$
755: and $\beta$ according to the $\chi^2$ criteria (squares for $N$, dots
756: for $\pi$ and triangles for $K^+$). The thin solid lines in the right
757: panel demonstrate the fit of the particle spectra within the optimal
758: parameters from MINUIT.
759: 
760: Since the particle spectra cover  several orders of magnitude and the
761: low energy points contribute to $\chi^2$ with a larger weight than
762: those at high energy, we use the logarithmic $\chi^2$ method to give a
763: higher weight to the tail of the spectra in the fitting procedure,
764: i.e.\  we minimize $\chi_{\ln}^2=\sum\limits_i (\ln f(x_i) - \ln
765: f_0(x_i))^2$, where $f_0$ represent the 'experimental' data (i.e. the
766: results of our box calculations), $f$ is the value of the fit
767: (\ref{flow}) at point $x_i$.
768: 
769: One can see from Fig.~\ref{Fig15} that the 'best' parameters $T$ and
770: $\beta$ (as well as the contours for the parameter errors) are
771: quite different for $N$, $\pi$ and $K^+$ especially for
772: $\varepsilon=0.3$~GeV/fm$^3$.  So we do not find (within the 'optimal'
773: $\chi^2$) common freeze-out parameters for all spectra simultaneously.
774: This is similar to an analysis of experimental spectra by Peitzmann et
775: al. \cite{Peitzmann}.  For all particles we obtain different
776: values for  $\beta$ and much lower temperatures $T$ than that of the
777: initial fireball:  $T_{in}=107$~MeV for $\varepsilon=0.3$~GeV/fm$^3$ and
778: $T_{in}=145$~MeV for $\varepsilon=1.1$~GeV/fm$^3$.
779: 
780: Since especially the pion spectra contain large contributions from
781: resonance decays at low $E-m$, we have also performed fits excluding
782: the pion spectra for $E-m \le 0.4$ GeV. This procedure essentially
783: gives lower $\beta$ parameters and higher values for $T$ (open circle
784: in the upper panel). However, the low energy cut-off is an additional
785: parameter that allows to 'extend' the ($\beta, T)$ values to a wider
786: range.
787: 
788: Thus our analysis indicates that the final particle spectra do not
789: allow a reliable reconstruction of freeze-out parameters within the
790: collective flow model (\ref{flow}). The parameters $T$ and $\beta$
791: obtained from the fit are very sensitive to the low energy shape of the
792: hadron spectra or low energy cut-off applied since this region
793: contributes with the largest weight to the $\chi^2$ minimization. On
794: the other hand, a global 'eye' fit with the parameters given by the
795: 'star' for all hadrons considered gives a quite reasonable overall
796: description of the spectra (dashed lines, r.h.s of Fig.~\ref{Fig15}),
797: too. A very accurate deduction of one single overall fit for all
798: hadrons by a common temperature ($T$) and flow velocity ($\beta $)
799: parameter (see, e.g., \cite{HeinzQM97} and references therein) seems to
800: us thus rather ambiguous. In particular, such an analysis may indicate
801: that thermal equilibrium has been reached to a much larger extent than
802: is actually true.
803: 
804: The particle flow effect due to the expansion is demonstrated in
805: Fig.~\ref{Fig16} where we show the velocity distributions $dN/d\beta$
806: for nucleons ($N$), pions ($\pi$) and kaons ($K^+$) for $\rho=1/3 \rho_0$
807: at  $\varepsilon=0.84$~GeV/fm$^3$ (upper part), for $\rho=\rho_0$ at
808: $\varepsilon=1.1$~GeV/fm$^3$ (middle part) and for $\rho=3\rho_0$ at
809: $\varepsilon=3.4$~GeV/fm$^3$ (lower part).  The left panel shows the
810: $dN/d\beta$ distribution at equilibrium whereas the right panel
811: corresponds to $dN/d\beta$ after the expansion phase.  On can see that
812: the average velocity of the particles decreases with the mass; the
813: pions are much faster than the nucleons. They thus leave the reaction
814: zone  at the initial stage of the ongoing and rapidly evolving
815: expansion with a higher velocity and accelerate the slower hadrons that
816: 'feel' the 'pion wind' by the multiple interactions \cite{Pang}.
817: We recall that the pion density is very high especially at high
818: $\varepsilon$ such that practically all other hadrons are shifted in
819: direction of large $\beta$. The same effect is shown in
820: Figs.~\ref{Fig12},\ref{Fig13} by the enhancement of the slope of the
821: nucleon spectra due to the expansion.
822: 
823: 
824: %---------------------------------------------------------------------
825: \section{Reactions of finite systems}
826: 
827: In this Section we turn to realistic nucleus-nucleus collisions with
828: the BUU transport model. We have learned from our analysis in the
829: previous Section that even by starting from an idealized scenario of
830: perfect thermal equilibrium, a rapid expansion stage makes the
831: extraction of one global temperature $T$ and one global expansion
832: parameter $\beta$ quite ambiguous. We thus expect this to become even
833: worse for the true situation of a relativistic heavy-ion collision,
834: where a perfect equilibrium state at some intermediate stage cannot
835: really be assumed. Also, we note that even for a very heavy system like
836: Pb+Pb the fraction of effective surface layer ($\sim 4\pi R_{eff}^2
837: \lambda$) to total volume is still quite sizeable, resulting in a
838: continuous emission or evaporation of particles from the outer layers
839: before a global freeze-out of bulk particles occurs. (For Pb+Pb
840: reactions at SPS energies -- combining a hydrodynamical evolution with
841: a nonequilibrium picture of surface emission -- it has indeed been
842: shown that at least 25$\%$ of all particles are continuously evaporated
843: before a global freeze-out has occurred \cite{Dumitru}.)
844: 
845: In Fig.~\ref{Fig17} we show the time evolution of the particle
846: abundances (nucleons $N$,  $\Delta$, $\Lambda$, $\pi$, $\eta$,  $K^+$)
847: for central collision of the light system $^{12}$C~+~$^{12}$C (upper
848: part) and heavy system $^{197}$Au~+~$^{197}$Au and
849: $^{208}$Pb~+~$^{208}$Pb (lower part) at the low energy of 1 A$\cdot$GeV
850: (left panel) and the high energy of 100 and 160 A$\cdot$GeV (right
851: panel).  The number of nucleons decreases in a few fm/c due to the
852: inelastic collisions, whereas the number of $\Delta$-resonances
853: increases accordingly. At low energy the $\eta$-mesons and strange
854: particles (we disregard strange particles for C~+~C at 1 A$\cdot$GeV
855: due to the low statistics) appear with a delay of a few fm/c due to the
856: fact that they are basically produced from resonance decays (the same
857: as pions) or from secondary pion-baryon collisions, whereas at high
858: energy they appear earlier due to the primary production mechanism
859: through the string formation and decay.
860: 
861: As seen from Fig~\ref{Fig17} the reaction time $\tau_{reac}$ for
862: 1~A$\cdot$GeV is $\sim 20-30$~fm/c, whereas for high energies
863: $\tau_{reac}$ is shorter due to a faster expansion -- $\tau_{reac}
864: \simeq 10-20$~fm/c.  It has been shown in Section 2.2 that the chemical
865: equilibration of hadronic matter under 'ideal' conditions (box without
866: expansion) requires a quite long time, e.g. the equilibration time
867: $\tau_{eq}$ for strange particles has been found to be larger than 40
868: fm/c for all energies and densities (cf.  Fig.~\ref{Fig4}).  In
869: realistic central nucleus-nucleus collisions, such as Au~+~Au, the
870: system expands rapidly (depending on the energy) after the compression
871: and formation of the hadronic fireball. The number of interactions,
872: which is the dynamical origin for equilibration, decreases
873: correspondingly very fast with time; after a few 10 fm/c the
874: particles are moving practically freely. Thus, the reaction time even
875: for central Au~+~Au collisions is much shorter than the time required
876: for strangeness equilibration: $\tau_{reac} \ll \tau_{eq}$.
877: 
878: The thermal equilibration time in this energy range around 0.25
879: GeV/fm$^3$, as obtained from the box calculation, is about 5--7 fm/c
880: (see Fig. \ref{Fig5}). Notice, however, that this calculation --
881: because of its periodic boundary conditions -- probably underestimates
882: the equilibration time. Indeed, studies of the longitudinal and
883: transverse temperatures ($T_L$ and $T_T$, resp.)
884: \cite{Lang91,EffePhD} have shown that full thermal equilibrium is
885: reached only in the very late expansion phase, when the density $\rho$
886: has dropped already below its saturation value. After a period of 10
887: fm/c (after first contact) one still finds $T_L \approx 1.5\, T_T$, i.e.
888: an anisotropy of about 40 \%, considerably more than indicated in
889: Fig.~\ref{Fig5}. At the bombarding energy of 160 A$\cdot$GeV we find
890: a rapid decrease of the quadrupole moment in the momentum space of all
891: hadrons by about a factor of 3 at the scale of 5 fm/c leading to
892: longitudinally expanding matter. In view of Fig.~\ref{Fig5} this
893: stretched ellipsoid in momentum space becomes isotropic only on the
894: scale  of 10--20 fm/c since low energy hadronic reactions are less
895: effective for equilibration. Since this 'hadronic' equilibration time
896: is larger than the reaction time for Pb~+~Pb at 160 A$\cdot$GeV a
897: substantial anisotropy remains in the hadron momentum distributions
898: after the collision.
899: 
900: 
901: Contrary to the box case the $d^3N/dp^3$ spectra for realistic
902: nucleus-nucleus collisions do not follow the simple exponential
903: behaviour (\ref{Boltz}) due to the strong longitudinal expansion;
904: especially at high bombarding energy the particle spectra show the
905: specific 'banana' shape (reflecting the $pp$ spectra at high energies).
906: In order to exclude this simple dynamical effect related to the
907: longitudinal expansion, we present in Fig.~\ref{Fig18} (right panel)
908: the transverse mass spectra $1/m_T^2 dN/dm_T$ at mid-rapidity ($-0.5\le
909: y_{cm} \le 0.5$) versus $m_T-m$ for central Au~+~Au collisions at 1
910: A$\cdot$GeV (upper part) and for central Pb~+~Pb at 160 A$\cdot$GeV
911: (lower part) calculated at the end of the reaction. The $m_T$-spectra
912: show an exponential behaviour \cite{Brat_mt} (excluding small $m_T$),
913: however, with different slopes which can not be associated directly
914: with a temperature of a hot fireball formed at the intermediate stages
915: of the reaction.
916: 
917: In line with Section 3 we have performed a fit within the blast model
918: (\ref{flow}) within an interval of unit rapidity around midrapidity
919: using MINUIT.  The results of the fit are displayed in
920: Fig.~\ref{Fig18}. The full symbols (squares for $N$, dots for $\pi$ and
921: triangles for $K^+$) correspond to the 'best' values for $T$ and
922: $\beta$ according to the $\chi^2$ criteria. The thin solid lines in the
923: right panel demonstrate the fit of the $m_T$ spectra within the optimal
924: fit parameters (we obtain a smaller $\chi^2$ within the linear $\chi^2$
925: method ($\chi^2=\sum\limits_i ( f(x_i) - f_0(x_i))^2$), which provides
926: a better description of the low $m_T$ spectra).
927: 
928: Similar to Section 3 (cf. Fig.~\ref{Fig15}) we obtain (within the
929: 'optimal' $\chi^2$ criterium) quite different freeze-out parameters for
930: $N, \pi$ and $K^+$ spectra. In order to exclude the influence of
931: $\Delta$- (and other resonance) decays on the pion spectra and to
932: investigate the sensitivity of the freeze-out parameters to the low
933: energy cuts applied, we performed a fit of the particle spectra using
934: the following cut-offs: $m_T-m > 0.2$~GeV (open symbols) and $m_T-m >
935: 0.4$~GeV (open symbols with crosses inside) for Au~+~Au at 1
936: A$\cdot$GeV;  $m_T-m > 0.4$~GeV (open symbols) and $m_T-m
937: > 0.5$~GeV (open symbols with crosses inside) for Pb~+~Pb at 160
938: A$\cdot$GeV.  As seen from the left panel of Fig.~\ref{Fig18} the
939: implementation of the low $m_T$ cut-off leads to a substantial shift of
940: the 'optimal' MINUIT parameters $\beta$ and $T$ especially for pions.
941: The $\beta, T$ values for the different spectra move towards to each
942: other when discarding the low $m_T$ points. For Au~+~Au at 1
943: A$\cdot$GeV our $\beta$ and $T$ parameters agree with those extracted
944: by the TAPS collaboration \cite{Averbeck} using the blast model (star
945: in the upper left plot). Here we have to mention that the cut-off
946: $m_T-m > 0.4$~GeV has been applied in the experimental analysis, too
947: \cite{Averbeck}. For the Pb~+~Pb spectra at 160 A$\cdot$GeV our
948: freeze-out parameters are similar to those from K\"ampfer
949: \cite{Kaempfer} ($T=120$~MeV, $\beta=0.43$; star in the lower left plot).
950: The dashed lines in the right panel of Fig.~\ref{Fig18} show
951: the fit to the particle $m_T$ spectra for the $\beta,T$ values
952: corresponding to the 'stars' from the left panel. Again this 'eye' fit
953: gives a reasonable description of the spectra (except of the very low
954: $m_T$ part).
955: 
956: Here we have to mention again that the extraction of freeze-out
957: parameters from the experimental data is very sensitive to the
958: details of the thermodynamical model applied as well as to the
959: observables considered. For example, the analysis of SIS data at
960: 1.0 GeV from Ref. \cite{Cleymans} gives thermal freeze-out parameters --
961: $T\simeq 52$ MeV and $\beta \simeq 0.4$.
962: At SPS energies the chemical freeze-out temperature extracted in Ref.
963: \cite{BM99} from the thermal-analysis of particle ratios is $T\simeq
964: 168$ MeV, whereas the analysis of particle spectra and two-particle
965: correlations (HBT data) \cite{Nix,Heinz99} provides a much lower
966: thermal freeze-out temperature $T\simeq 90-95$~MeV. For a survey
967: different freeze-out parameters the reader is referred to
968: Fig. 4 of Ref.~\cite{Redlich}).
969: 
970: In view of the various uncertainties inherent in the extraction of the
971: thermal freeze-out parameters we conclude that a full, i.e.\ thermal
972: and chemical, thermodynamical equilibrium at freeze-out cannot be
973: deduced from such an analysis.
974: 
975: 
976: %---------------------------------------------------------------------
977: \section{Summary}
978: 
979: In this paper we have performed a systematic study of equilibration
980: phenomena and equilibrium properties of 'infinite' hadronic matter as
981: well as of relativistic nucleus-nucleus collisions using a BUU
982: transport model that contains resonance and string degrees-of-freedom.
983: The 'infinite' hadron matter is modelled  by initializing the system at
984: fixed baryon density, strange density and energy density by confining
985: it in a cubic box with periodic boundary conditions.
986: 
987: We have shown that the equilibration times $\tau_{eq}$ for different
988: particles depend on baryon density and energy density. The time
989: $\tau_{eq}$ for non-strange particles is much shorter than for
990: particles including strangeness; for kaons and antikaons the
991: equilibration time is found to be larger than $\simeq$ 40 fm/c for all
992: baryon and energy densities considered. The overall abundance of the
993: dominant strange particles (kaons and $\Lambda$'s) being produced and
994: obtained within the BUU cascade model for heavy-ion collisions can
995: therefore not be described by assuming a perfect chemical equilibrium
996: as strangeness is typically still undersaturated to a quite large
997: extent. We mention taht transport model calculations like ours can
998: describe the yield and spectra of the produced nonstrange hadrons as
999: well as $K^+, K^-, \Lambda$ yields quite well at SPS energies
1000: \cite{Cass99,Geiss}.  On the other hand, at AGS energies the measured
1001: $K^+/\pi^+$ ratio in central Au~+~Au collisions is underestimated by
1002: about 30\% \cite{CassQM}.  However, we have to point out that the more
1003: exotic strange particles (like the measured antihyperon yields of
1004: Ref.~\cite{WA97}) can by far not be explained within such standard
1005: hadronic multiple channel reactions.  These hadronic data possibly
1006: point towards new physics.
1007: 
1008: We have, furthermore, shown that thermal equilibrium is established
1009: quickly, within about 5 fm/c at SIS energies and samewhat larger times
1010: at high energies.  The inclusion of continuum excitations, i.e. hadron
1011: 'strings', leads to a limiting temperature of $T_s \simeq 150$~MeV in
1012: our transport approach which practically does not depend on the baryon
1013: density and energy.  We have compared our results with the
1014: statistical model (SM), which contains the same degrees of freedom and
1015: the same spectral functions of particles as our transport model. We
1016: found that the limiting temperature behaviour can be reproduced in the
1017: statistical model only after including continuum excitations of the
1018: Hagedorn type, otherwise the fireball temperature extracted from the
1019: particle abundances and spectra is overestimated substantially.
1020: 
1021: Close to the critical temperature $T_s$, the hadronic energy densities
1022: can increase to a couple of GeV/fm$^3$. From lattice QCD calculations
1023: one expects that a phase transition to a potentially deconfined QGP
1024: state should occur. Referring to the limiting temperature $T_s\approx
1025: 150 $ MeV obtained, a QGP should be revealed and clearly distinguished from
1026: a hadronic state of matter if one can unambiguously prove the
1027: existence of an equilibrated and thermal phase of strongly interacting
1028: matter with temperatures exceeding, e.g., 200 MeV.  The best candidates
1029: are electromagnetic probes, either direct photons or dileptons. On the
1030: other hand these are also `contaminated' by hadronic background and/or
1031: preequilibrium physics. So far no thermal electromagnetic source with
1032: temperatures larger or equal than 200 MeV has been clearly identified.
1033: 
1034: We have also studied the expansion of the equilibrated hadronic
1035: fireball and found that the slope parameters of the particles after
1036: expansion increase with their mass; the pions leave the fireball much
1037: faster than nucleons and accelerate heavier hadrons by rescattering
1038: ('pion wind'). If the system before expansion is close to the limiting
1039: temperature $T_s$, the slope parameters for all particles after
1040: expansion practically do not depend on energy and baryon density. This
1041: is due to the fact that the particle velocity distributions in
1042: equilibrium do not change any more for $T \approx T_s$. We have fitted
1043: the resulting spectra within the blast model of Siemens and Rasmussen.
1044: Our analysis shows a strong sensitivity of the $(\beta, T)$ parameters
1045: on the spectral shape at low energy (or a low energy cut-off) so that
1046: no reliable parameter determination can be reported. However, a global
1047: 'eye' fit with 'average' $(\beta, T)$ parameters describes the data
1048: reasonably well.
1049: 
1050: Additionally, we have considered the  equilibration in realistic
1051: nucleus-nucleus collisions of light (C~+~C) and heavy (Au~+~Au and
1052: Pb~+~Pb) systems. The $(\beta, T)$ parameters extracted from our
1053: calculations for Au~+~Au at 1 A$\cdot$GeV agree with those extracted from
1054: the TAPS collaboration \cite{Averbeck} and for Pb~+~Pb at 160 A$\cdot$GeV
1055: with the parameters from Ref.~\cite{Kaempfer}. Here the reaction time
1056: is a few 10~fm/c and decreases with the initial energy due to the fast
1057: expansion.  Since the reaction time is much shorter than the
1058: equilibration time for strangeness, a chemical equilibrium of strange
1059: particles in heavy-ion collisions is not supported by our transport
1060: calculations. Although again simple fits within the blast model provide
1061: a decent parametrization of our transport results for the differential
1062: particle spectra (Fig.~\ref{Fig18}) a deduction of global parameters
1063: for thermal freeze-out is again found to be rather ambiguous,
1064: especially when considering also the lower momentum contributions of
1065: the various particle spectra.
1066: 
1067: \section*{Acknowledgements}
1068: The authors are grateful for valuable discussions with V.~Metag,
1069: H.~Oeschler and H.~St\"ocker.
1070: 
1071: %---------------------------------------------------------------------
1072: %\newpage
1073: \begin{thebibliography}{999}
1074: \bibitem{QM}
1075:     QUARK MATTER '96, Nucl. Phys. A 610 (1997) 1;
1076:     QUARK MATTER '97, Nucl. Phys. A 638 (1998) 1;
1077:     QUARK MATTER '99, Nucl. Phys. A (1999), in print.
1078: \bibitem{BotMal90}
1079:     W. Botermans and R. Malfliet, Phys. Rep. 198 (1990) 115.
1080: \bibitem{Henning}
1081:     P.A. Henning, Phys. Rep. 253 (1995) 235.
1082: \bibitem{Ko}
1083:     G. Q. Li and C. M. Ko, J. Phys. G 22 (1997) 1673.
1084: \bibitem{Bass}
1085:     S. A. Bass et al., Prog. Part. Nucl. Phys. 42 (1998) 279;
1086:     J. Phys. G 25 (1999) 1859.
1087: \bibitem{Cass99}
1088:     W. Cassing and E. L. Bratkovskaya, Phys. Rep. 308 (1999) 65.
1089: \bibitem{BM}
1090:        P. Braun-Munzinger, J. Stachel, J.P. Wessels, and N. Xu,
1091:         Phys. Lett. B 344 (1995)~43;
1092:        P. Braun-Munzinger, J. Stachel, J.P. Wessels, and N. Xu,
1093:        Phys. Lett. B 365 (1996)~1;
1094:     J. Stachel, Nucl. Phys. A 654 (1999) 119.
1095: \bibitem{Satz}
1096:     J. Cleymans and H. Satz, Z. Phys. C 57 (1993) 135.
1097: \bibitem{Sollfrank}
1098:      J. Sollfrank, M. Gazdzicki, U. Heinz and J. Rafelski,
1099:      Z. Phys. C  61 (1994) 659;
1100:      F. Becattini, M. Gazdzicki and J. Sollfrank,
1101:      Eur. Phys. J. C 5 (1998) 143.
1102: \bibitem{Spieles}
1103:      C. Spieles, H. St\"ocker and C. Greiner,
1104:      Eur. Phys. J. C 2 (1998) 351.
1105: \bibitem{Cleymans}
1106:     J. Cleymans, H. Oeschler, and K. Redlich,
1107:     nucl-th/9809027; J. Phys. G 25 (1999) 281.
1108: \bibitem{Hydro}
1109:     H. St\"ocker and W. Greiner, Phys. Rep. 137 (1986) 277.
1110: \bibitem{Rischke}
1111:     U. Ornik et al., Phys. Rev. C 54 (1996) 1381;
1112:     S. Bernard et al., Nucl. Phys. A 605 (1996) 566;
1113:     J. Sollfrank et al., Phys. Rev. C 55 (1997) 392.
1114: \bibitem{Rafelski}
1115:     P. Koch, B. M\"uller, and J. Rafelski, Phys. Rep. 142 (1986) 167.
1116: \bibitem{Cass90}
1117:     W. Cassing, V. Metag, U. Mosel, and K. Niita,
1118:     Phys. Rep. 188 (1990) 363.
1119: \bibitem{Lang91}
1120:     A. Lang, B. Bl\"attel, W. Cassing, V. Koch, U. Mosel, and K. Weber,
1121:     Z. Phys. A 340 (1991) 287.
1122: \bibitem{Bl93}
1123:     B. Bl\"attel, V. Koch, and U. Mosel,
1124:     Rep. Progr. Phys. 56 (1993) 1.
1125: \bibitem{Brav1}
1126:     M. Belkacem, M. Brandstetter, S.A. Bass et al.,
1127:     Phys. Rev. C 58 (1998) 1727.
1128: \bibitem{Brav2}
1129:     L.V. Bravina, M.I. Gorenstein, M. Belkacem et al.,
1130:     Phys. Lett. B 434 (1998) 379;
1131:     L.V. Bravina, M. Brandstetter, M.I. Gorenstein et al.,
1132:     J. Phys. G 25 (1999) 351.
1133: \bibitem{Brav3}
1134:     L.V. Bravina, E.E. Zabrodin, M.I. Gorenstein et al.,
1135:     Phys. Rev. C 60 (1999) 024904.
1136: \bibitem{Solfr99}
1137:     J. Sollfrank, U. Heinz, H. Sorge, N. Xu, Phys. Rev. C 59 (1999) 1637.
1138: \bibitem{TeisZP97}
1139:     S. Teis, W. Cassing, M. Effenberger, A. Hombach, U. Mosel,
1140:     and Gy. Wolf, Z. Phys. A 356 (1997) 421.
1141: \bibitem{Effe99gam}
1142:     M. Effenberger, E.L. Bratkovskaya, and U. Mosel,
1143:     Phys. Rev. C 60 (1999) 044614.
1144: \bibitem{EffePhD}
1145:     M. Effenberger, Ph.D. Thesis, Univ. of Giessen, 1999;
1146:     http://theorie.physik.uni-giessen.de/ftp.html.
1147: \bibitem{Manley}
1148:        D. M. Manley and E. M. Saleski, Phys. Rev. D 45 (1992) 4002.
1149: \bibitem{FRITIOF}
1150:     B. Anderson, G. Gustafson and Hong Pi, Z. Phys. C 57 (1993) 485.
1151: \bibitem{Ehehalt}
1152:     W. Ehehalt and W. Cassing, Nucl. Phys. A 602 (1996) 449.
1153: \bibitem{Brat98}
1154:     E. L. Bratkovskaya and W. Cassing, Nucl. Phys. A 619 (1997) 413.
1155: \bibitem{Geiss}
1156:     J. Geiss, W. Cassing, and C. Greiner, Nucl. Phys. A 644 (1998) 107.
1157: \bibitem{Hagedorn}
1158:     R.Hagedorn, Suppl. Novo Cimento 3 (1965) 147;
1159:     Suppl. Novo Cimento 6 (1965) 311;
1160:     R.Hagedorn and J. Ranft, Suppl. Novo Cimento 6 (1968) 169.
1161: \bibitem{Sailer}
1162:     K. Sailer et al., J. Phys. G 17 (1991) 1005.
1163: \bibitem{HeinzQM97}
1164:     U. Heinz, Nucl. Phys. A 638 (1998) 357c.
1165: \bibitem{Nix}
1166: 	J.R. Nix, Phys. Rev. C 58 (1998) 2303;
1167: 	J.R. Nix et al., nucl-th/9801045.
1168: \bibitem{Heinz99}
1169:      B. Tomasik, U.A. Wiedemann and U. Heinz, nucl-th/9907096.
1170: \bibitem{Siemens}
1171:     P.J. Siemens and J.O. Rasmussen, Phys. Lett. 42 (1979) 880.
1172: \bibitem{MINUIT}
1173:     F. James and M. Roos, Com. Phys. Commun. 10 (1975) 343.
1174: \bibitem{Peitzmann}
1175:     T. Peitzmann et al., Phys. Rev. Lett. 83 (1999) 926.
1176: \bibitem{Pang}
1177:     M. Gyulassy, private communication.
1178: \bibitem{Brat_mt}
1179:     E.L. Bratkovskaya, W. Cassing and U. Mosel,
1180:     Phys. Lett.  B 424 (1998) 244.
1181: \bibitem{Dumitru}
1182:     A. Dumitru, C. Spieles, H. St\"ocker and C. Greiner,
1183:     Phys. Rev. C56 (1997) 2202.
1184: \bibitem{Averbeck}
1185:     R. Averbeck, nucl-ex/9803001.
1186: \bibitem{Kaempfer}
1187:      B. K\"ampfer, J. Phys. G 23 (1997) 2001.
1188: \bibitem{BM99}
1189:      P. Braun-Munzinger, I. Heppe and J. Stachel,
1190:      Phys. Lett. B 465 (1999) 15.
1191: \bibitem{Redlich}
1192: 	J. Cleymans and K. Redlich, Phys. Rev. C 60 (1999) 054908.
1193: \bibitem{CassQM}
1194: 	W. Cassing, Nucl. Phys. A 661 (1999) 468c.
1195: \bibitem{WA97}
1196:     E. Andersen et al. (WA97 Collaboration), Phys. Lett. B433 (1998) 209;
1197:     J. Phys. G 25 (1999) 171; J. Phys. G 25 (1999) 181.
1198: \end{thebibliography}
1199: 
1200: %\end{document}
1201: %---------------------------------------------------------------------
1202: \newpage
1203: 
1204: \begin{figure}[t]
1205: \phantom{a}\vspace*{-2cm}
1206: \centerline{\psfig{figure=fig1box.eps,width=16cm}}
1207: \vspace*{-2.5cm}
1208: \caption{Time evolution of the various
1209: particle abundances (nucleons $N$,  $\Delta$,  $\Lambda$, $\pi$,
1210: $\eta$,  $K^+$ and $K^-$ mesons) for density $\rho=\rho_0$ (left
1211: panel) at different energy densities $\varepsilon =1.1, 0.52$ and
1212: 0.22 GeV/fm$^3$ and for $\rho=3\rho_0$ (right panel) at
1213: $\varepsilon =3.4, 1.57$ and 0.66 GeV/fm$^3$.}
1214: \label{Fig1}
1215: \end{figure}
1216: 
1217: 
1218: \begin{figure}[t]
1219: \phantom{a}\vspace*{-2cm}
1220: \centerline{\psfig{figure=fig2box.eps,width=16cm}}
1221: \vspace*{-2.5cm}
1222: \caption{Time evolution of particle ratios $\pi/N$, $\Delta/N$,
1223: $\Lambda/N$, $K^+/\pi^+$, $K^-/\pi^-$, $\eta/\pi$ for density
1224: $\rho=\rho_0$ at energy densities $\varepsilon = 1.1, 0.52$ and 0.2
1225: GeV/fm$^3$. The left panel shows the time scale up to 1000 fm/c,
1226: whereas the right panel demonstrates the initial stage up to 30 fm/c.}
1227: \label{Fig2}
1228: \end{figure}
1229: 
1230: \begin{figure}[t]
1231: \phantom{a}\vspace*{-2cm}
1232: \centerline{\psfig{figure=fig3box.eps,width=16cm}}
1233: \vspace*{-2.5cm}
1234: \caption{Time evolution of particle ratios $\pi/N$, $\Delta/N$,
1235: $\Lambda/N$, $K^+/\pi^+$, $K^-/\pi^-$, $\eta/\pi$ for density $\rho=3\rho_0$
1236: at energy densities $\varepsilon = 3.4, 1.57$ and 0.66 GeV/fm$^3$.
1237: The left panel shows the time scale up to 1000 fm/c, whereas the right
1238: panel demonstrates the initial stage up to 30 fm/c.}
1239: \label{Fig3}
1240: \end{figure}
1241: 
1242: 
1243: \begin{figure}[t]
1244: \centerline{\psfig{figure=fig4box.eps,width=16cm}}
1245: \vspace*{-2.5cm}
1246: \caption{Equilibration time $\tau_{eq}$ versus energy density $\varepsilon$
1247: for $\pi$ and $K^+$ mesons at different baryon densities $1/3\rho_0,
1248: \rho_0, 3\rho_0$ and $6\rho_0$.}
1249: \label{Fig4}
1250: \end{figure}
1251: 
1252: \begin{figure}[t]
1253: \centerline{\psfig{figure=fig5box.eps,width=16cm}} \vspace*{-2.5cm}
1254: \caption{Left panel: time evolution of the quadrupole moment $<Q_2>$
1255: for density $\rho=\rho_0$ at energy densities $\varepsilon=0.22, 0.3,
1256: 0.52, 0.8, 1.1$ and 1.6 GeV/fm$^3$. The thin solid lines indicate the
1257: fit of $<Q_2>$ by two exponentials (\ref{Q2}). The parameters
1258: $\tau_{short}$ and $\tau_{long}$ are shown in the right panel as a
1259: function of the energy density $\varepsilon$.}
1260: \label{Fig5}
1261: \end{figure}
1262: 
1263: \begin{figure}[t]
1264: \centerline{\psfig{figure=fig6box.eps,width=16cm}}
1265: \vspace*{-2.5cm}
1266: \caption{The spectra of nucleons ($N$), pions ($\pi$)
1267: and kaons ($K^+$) as a function of the kinetic energy $E-m$ for
1268: $\rho=\rho_0$ at energy densities $\varepsilon=0.52, 0.8$ and 1.6
1269: GeV/fm$^3$ (left panel) and for $\rho=3\rho_0$ at energy densities
1270: $\varepsilon=0.66, 1.57$ and 2.85 GeV/fm$^3$ (right panel).}
1271: \label{Fig6}
1272: \end{figure}
1273: 
1274: \begin{figure}[t]
1275: \phantom{a}\vspace*{-2cm}
1276: \centerline{\psfig{figure=fig7box.eps,width=16cm}}
1277: \vspace*{-1.5cm}
1278: \caption{The energy density  $\varepsilon-m_N\rho$ versus equilibrium
1279: temperature $T$ for different baryon densities $\rho$:  $1/3\rho_0$
1280: (open down triangles), $\rho_0$ (full squares), $3\rho_0$ (full dots),
1281: $6\rho_0$ (full up triangles). }
1282: \label{Fig7}
1283: \end{figure}
1284: 
1285: 
1286: \begin{figure}[t]
1287: \centerline{\psfig{figure=fig8box.eps,width=16cm}} \vspace*{-2.5cm}
1288: \caption{The excitation function for the ratio of string energy density
1289: to the energy density of the hole system
1290: $\varepsilon_{string}/\varepsilon$ at $\rho=\rho_0$.}
1291: \label{Fig8}
1292: \end{figure}
1293: 
1294: \begin{figure}[t]
1295: \phantom{a}\vspace*{-2cm}
1296: \centerline{\psfig{figure=fig9box.eps,width=16cm}}
1297: \vspace*{-1.5cm}
1298: \caption{The energy density versus equilibrium temperature $T$ for
1299: baryon density $\rho=\rho_0$. The full dots  correspond to the
1300: statistical model (SM) without strings, the full squares show our
1301: box calculations including string degrees of freedom, while the
1302: solid line shows the result from the extended SM including a Hagedorn
1303: mass spectrum for strings.}
1304: \label{Fig9}
1305: \end{figure}
1306: 
1307: \begin{figure}[t]
1308: \phantom{a}\vspace*{-1.5cm}
1309: \centerline{\psfig{figure=fig10box.eps,width=16cm}} \vspace*{-3.5cm}
1310: \caption{The $T-\mu_B$ phase correlation, i.e. temperature $T$ versus
1311: baryon chemical potential $\mu_B$. The open triangles and squares
1312: (connected by the dashed lines) show the result of the statistical
1313: model without strings (standard SM) fitted to out box calculations at
1314: densities $1/3\rho_0$ and $\rho_0$, respectively, whereas the full
1315: triangles and squares (connected by the solid lines) correspond to the
1316: thermodynamical fit of the box calculations (at $1/3\rho_0$ and
1317: $\rho_0$) including string excitations (extended SM). The arrow at
1318: $\mu_B=0$ indicates the limiting temperature $T_s=150$~MeV from our box
1319: calculations.  The full dots correspond to the chemical freeze-out
1320: points from Ref.~\protect\cite{BM} while the open dots are the thermal
1321: freeze-out points from Ref.~\protect\cite{HeinzQM97}. }
1322: \label{Fig10}
1323: \end{figure}
1324: 
1325: \begin{figure}[t]
1326: \phantom{a}\vspace*{-0.5cm}
1327: \centerline{\psfig{figure=fig11box.eps,width=16cm}}
1328: \vspace*{-2.5cm}
1329: \caption{Time evolution of the various particle abundances (nucleons
1330: $N$,  $\Delta$,  $\Lambda$, $\pi$, $K^+$ and $K^-$ mesons) during the
1331: expansion (starting at $t=500$~fm/c) for density $\rho=\rho_0$
1332: (left panel) at different energy densities $\varepsilon =0.22, 0.3$ and
1333: 1.1 GeV/fm$^3$ and for density $\rho=1/3\rho_0$ at $\varepsilon =0.84$
1334: GeV/fm$^3$ (upper part in the right panel), for $\rho=\rho_0$ at
1335: $\varepsilon =1.6$ GeV/fm$^3$ (middle part in the right panel) and for
1336: $\rho=3 \rho_0$ at $\varepsilon =3.4$ GeV/fm$^3$ (lower part in the
1337: right panel).}
1338: \label{Fig11}
1339: \end{figure}
1340: 
1341: \begin{figure}[t]
1342: \centerline{\psfig{figure=fig12box.eps,width=16cm}}
1343: \vspace*{-2.5cm}
1344: \caption{The spectra of nucleons ($N$), pions ($\pi$) and kaons ($K^+$)
1345: versus the kinetic energy $E-m$ for $\rho=\rho_0$ at
1346: $\varepsilon=0.22, 0.3$ and 1.1~GeV/fm$^3$  before the expansion
1347: (left panel) and after the expansion (right panel).}
1348: \label{Fig12}
1349: \end{figure}
1350: 
1351: \begin{figure}[t]
1352: \centerline{\psfig{figure=fig13box.eps,width=16cm}}
1353: \vspace*{-2.5cm}
1354: \caption{The spectra of nucleons ($N$), pions ($\pi$) and kaons ($K^+$)
1355: versus the kinetic energy $E-m$ before (left panel) and after (right
1356: panel) expansion for $\rho=1/3\rho_0$ at $\varepsilon=0.84$ GeV/fm$^3$
1357: (upper part), for $\rho=\rho_0$ at $\varepsilon=1.6$ GeV/fm$^3$ (middel
1358: part) and for $\rho=3\rho_0$ at $\varepsilon=3.4$ GeV/fm$^3$ (lower part). }
1359: \label{Fig13}
1360: \end{figure}
1361: 
1362: \begin{figure}[t]
1363: \centerline{\psfig{figure=fig14box.eps,width=16cm}}
1364: \vspace*{-2.5cm}
1365: \caption{The spectral slope $T$ after expansion versus the hadron mass $m$
1366: for $\pi, K^+, N$ at $\rho=\rho_0$ and different
1367: energy densities: $\varepsilon = 0.2$~GeV/fm$^3$ (up full triangles),
1368: $\varepsilon = 0.3$~GeV/fm$^3$ (full squares),
1369: $\varepsilon = 1.1$~GeV/fm$^3$ (full dots),
1370: $\varepsilon = 1.6$~GeV/fm$^3$ (full diamonds);
1371: for $\rho=3\rho_0$ at $\varepsilon=3.4$~GeV/fm$^3$ (down open triangles).
1372: The arrow indicates the limiting temperature $T_s\simeq 150$~MeV
1373: before the expansion.}
1374: \label{Fig14}
1375: \end{figure}
1376: 
1377: \begin{figure}[t]
1378: \phantom{a}\vspace*{-3cm}
1379: \centerline{\psfig{figure=fig15box.eps,width=16cm}}
1380: \vspace*{-4cm}
1381: \caption{Left panel: the contour plots for the parameter errors in the
1382: $T-\beta$ plane; dot-dashed lines: nucleons ($N$), solid lines:
1383: pions ($\pi$) and dashed lines: kaons ($K^+$) for the energy
1384: densities $\varepsilon =0.3$ (upper part) and 1.1~GeV/fm$^3$ (lower
1385: part) at $\rho=\rho_0$. The full symbols indicate the 'optimal'
1386: parameters $T$ and $\beta$ (squares for $N$, dots for $\pi$ and
1387: triangles for $K^+$).
1388: The open dot (upper left plot) reflects $\beta,T$ for the pion spectra
1389: including the cut-off $E-m >0.4$ GeV.
1390: Right panel: the full symbols (squares
1391: for $N$, dots for $\pi$ and triangles for $K^+$) are the box
1392: calculations (for the same $\rho$ and $\varepsilon$ as in the left
1393: panel).  The thin solid lines show the fit of the particle spectra with the
1394: 'optimal' $T$ and $\beta$ parameters from the left panel; the dashed lines
1395: correspond to the 'eye' fit with the average $\beta$ and $T$ parameters
1396: given by the stars from the left panel.}
1397: \label{Fig15}
1398: \end{figure}
1399: 
1400: \begin{figure}[t]
1401: \centerline{\psfig{figure=fig16box.eps,width=16cm}}
1402: \vspace*{-2.5cm}
1403: \caption{The velocity distributions $dN/d\beta$ for nucleons ($N$),
1404: pions ($\pi$) and kaons ($K^+$) for $\rho=1/3 \rho_0$ at
1405: $\varepsilon=0.84$~GeV/fm$^3$ (upper part), for $\rho=\rho_0$ at
1406: $\varepsilon=1.1$~GeV/fm$^3$ (middle part) and for $\rho=3\rho_0$
1407: at  $\varepsilon=3.4$~GeV/fm$^3$ (lower part). The left panel shows
1408: $dN/d\beta$ at equilibrium whereas the right panel corresponds to
1409: $dN/d\beta$ after the expansion phase.}
1410: \label{Fig16}
1411: \end{figure}
1412: 
1413: \begin{figure}[t]
1414: \centerline{\psfig{figure=fig17box.eps,width=16cm}}
1415: \vspace*{-2.5cm}
1416: \caption{Time evolution of the particle abundances (nucleons $N$,
1417: $\Delta$,  $\Lambda$, $\pi$, $\eta$,  $K^+$) for central C~+~C
1418: collisions (upper part) and Au~+~Au and Pb~+~Pb collisions (lower part)
1419: at 1 A$\cdot$GeV (left panel) and 160 A$\cdot$GeV (right panel).}
1420: \label{Fig17}
1421: \end{figure}
1422: 
1423: \begin{figure}[t]
1424: \phantom{a}\vspace*{-2.5cm}
1425: \centerline{\psfig{figure=fig18box.eps,width=16cm}}
1426: \vspace*{-6cm}
1427: \caption{Left panel: the full symbols indicate the 'optimal' parameters
1428: $T$ and $\beta$ (squares for $N$, dots for $\pi$ and triangles for
1429: $K^+$) obtained by exploring the full $m_T$ spectra in the fitting
1430: procedure. The open symbols correspond to $\beta,T$ for the cut-off $m_T-m
1431: >0.2$~GeV (upper part) and 0.4 GeV (lower part); the open symbols with
1432: crosses inside indicate the $\beta,T$ parameters
1433: for the cut-off $m_T-m >0.4$~GeV
1434: (upper part) and 0.5 GeV (lower part).  The stars corresponds to the
1435: $\beta,T$ parameters from Ref.  \protect\cite{Averbeck} for Au~+~Au at
1436: 1 A$\cdot$GeV and from Ref. \protect\cite{Kaempfer} for Pb~+~Pb at 160
1437: A$\cdot$GeV.  Right panel: Transverse mass ($m_T$) spectra of nucleons
1438: ($N$), pions ($\pi$) and kaons ($K^+$) for central Au~+~Au collisions
1439: at 1 A$\cdot$GeV (upper part) and  central Pb~+~Pb collisions at
1440: 160~A$\cdot$GeV (lower part) for $-0.5\le y_{cm} \le 0.5$:  the full
1441: symbols are the transport calculation; the thin solid lines show the
1442: $m_T$ spectra for the 'optimal' $T$ and  $\beta$ parameters from the
1443: left panel; the dashed lines show the fit with the $\beta,T$ values
1444: corresponding to the 'stars' from the left panel. }
1445: \label{Fig18}
1446: \end{figure}
1447: 
1448: \end{document}
1449: 
1450: #!/bin/csh -f
1451: # this uuencoded Z-compressed .tar file created by csh script  uufiles
1452: # for more information, see e.g. http://xxx.lanl.gov/faq/uufaq.html
1453: # if you are on a unix machine this file will unpack itself:
1454: # strip off any mail header and call resulting file, e.g., figures.uu
1455: # (uudecode ignores these header lines and starts at begin line below)
1456: # then say        csh figures.uu
1457: # or explicitly execute the commands (generally more secure):
1458: #    uudecode figures.uu ;   uncompress figures.tar.Z ;
1459: #    tar -xvf figures.tar
1460: # on some non-unix (e.g. VAX/VMS), first use an editor to change the
1461: # filename in "begin" line below to figures.tar_Z , then execute
1462: #    uudecode figures.uu
1463: #    compress -d figures.tar_Z
1464: #    tar -xvf figures.tar
1465: #
1466: uudecode $0
1467: chmod 644 figures.tar.Z
1468: zcat figures.tar.Z | tar -xvf -
1469: rm $0 figures.tar.Z
1470: exit
1471: 
1472: