nucl-th0007019/text
1: %%%%%%%%%% espcrc1.tex %%%%%%%%%%
2: %
3: % $Id: espcrc1.tex 1.1 1999/07/26 10:27:59 Simon Exp spepping $
4: %
5: \documentclass[12pt,twoside]{article}
6: \usepackage{fleqn,espcrc1,epsfig}
7: 
8: % change this to the following line for use with LaTeX2.09
9: % \documentstyle[12pt,twoside,fleqn,espcrc1]{article}
10: 
11: % if you want to include PostScript figures
12: %\usepackage{graphicx}
13: % if you have landscape tables
14: %\usepackage[figuresright]{rotating}
15: 
16: % put your own definitions here:
17: %   \newcommand{\cZ}{\cal{Z}}
18: %   \newtheorem{def}{Definition}[section]
19: %   ...
20: %\newcommand{\ttbs}{\char'134}
21: %\newcommand{\AmS}{{\protect\the\textfont2
22: %  A\kern-.1667em\lower.5ex\hbox{M}\kern-.125emS}}
23: 
24: % add words to TeX's hyphenation exception list
25: %\hyphenation{author another created financial paper re-commend-ed Post-Script}
26: 
27: % declarations for front matter
28: \title{Thermal and chemical equilibration of hadronic matter
29: \thanks{Work supported by BMBF, GSI Darmstadt and DFG.}}
30: 
31: \author{E. L. Bratkovskaya, W. Cassing, C. Greiner,
32: 	 M. Effenberger, U. Mosel and A. Sibirtsev\\[3mm]
33: %\address{Institut f\"{u}r Theoretische Physik, Universit\"{a}t Giessen,
34: %	  35392 Giessen, Germany}}
35: Institut f\"{u}r Theoretische Physik, Universit\"{a}t Giessen,
36: 	  35392 Giessen, Germany}
37: \begin{document}
38: 
39: % typeset front matter
40: \maketitle
41: 
42: \begin{abstract}
43: We study thermal and chemical equilibration in 'infinite' hadron matter
44: as well as in finite size relativistic nucleus-nucleus collisions using
45: a BUU cascade transport model with resonance and string
46: degrees-of-freedom.  The 'infinite' hadron matter is simulated within a
47: cubic box employing periodic boundary conditions.  The various equilibration
48: times depend on baryon density and energy density and are much shorter
49: for particles consisting of light quarks then for particles including
50: strangeness.  For kaons and antikaons the chemical equilibration time
51: is found to be larger than $\simeq$ 40 fm/c for all baryon and energy
52: densities considered.  The inclusion of continuum excitations, i.e.
53: hadron 'strings', leads to a limiting temperature of $T_s\simeq$ 150 MeV.
54: \end{abstract}
55: 
56: \section{INTRODUCTION}
57: 
58: Nucleus-nucleus collisions at relativistic and ultrarelativistic
59: energies are studied experimentally and theoretically to obtain
60: information about the properties of hadrons at high density and/or
61: temperature as well as  about the phase transition to a new state of
62: matter, the quark-gluon plasma (QGP).  In the latter deconfined partons
63: are the essential degrees of freedom that resolve the underlying
64: structure of hadrons \cite{QM}.  Whereas the early 'big-bang' of the
65: universe most likely evolved through steps of kinetic and chemical
66: equilibrium, the laboratory 'tiny bangs' proceed through phase-space
67: configurations that initially are far from an equilibrium phase and
68: then evolve by fast expansion. These 'specific initial conditions' --
69: on the theoretical side -- have lead to a rapid development of
70: nonequilibrium quantum field theory and nonequilibribrium kinetic
71: theory \cite{BotMal90,Henning}.  Presently, semiclassical transport
72: models are widely used as approximate solutions to these theories and
73: practically are an essential ingredient in the experimental data
74: analysis. For recent reviews we refer the reader to Refs.
75: \cite{Ko,Bass,Cass99}.
76: 
77: On the other hand, many observables from strongly interacting systems
78: are dominated by many-body phase space such that spectra and abundances
79: look 'thermal'.  It is thus tempting to characterize the experimental
80: observables by global thermodynamical quantities like 'temperature',
81: chemical potentials or entropy \cite{BM,Satz,Sollfrank,Spieles,Cleymans}.
82: We note, that even the use of macroscopic models like hydrodynamics
83: \cite{Hydro,Rischke} employs as basic assumption the concept of local
84: thermal and chemical equilibrium. The crucial question, however, how
85: and on what timescales a global thermodynamic equilibrium can be
86: achieved, is presently a matter of debate. Thus nonequilibrium
87: approaches have been used in the past to address the problem of
88: timescales associated to global or local equilibration
89: \cite{Rafelski,Cass90,Lang91,Bl93,Brav1,Brav2,Brav3,Solfr99}.  In view
90: of the increasing 'popularity' of thermodynamic analyses a thorough
91: microscopic reanalysis of this questions appears necessary especially
92: for nucleus-nucleus collisions at ultrarelativistic energies that aim
93: at 'detecting' a phase transition to the QGP.
94: 
95: In this contribution we discuss equilibration phenomena in 'infinite'
96: hadronic matter using a microscopic transport model that contains both
97: hadron resonance and string degrees-of-freedom.  With this
98: investigation we want to provide insight into the reaction dynamics by
99: the use of cascade-like models and also point out some of their
100: limitations.  The 'infinite' hadronic matter is modelled by
101: initializing the system solely by nucleonic degrees of freedom through
102: a fixed baryon density and energy density, while confining it to a
103: cubic box and imposing periodic boundary conditions during the
104: propagation in time.
105: 
106: %---------------------------------------------------------------------
107: \section{EQUILIBRATION AND LIMITING TEMPERATURE}
108: 
109: To investigate the equilibration phenomena addressed above we perform
110: microscopic calculations using the Boltzmann-Uehling-Uhlenbeck (BUU)
111: model of Refs. \cite{Effe99gam,EffePhD}. This model is based on the
112: resonance concept of nucleon-nucleon and meson-nucleon interactions at
113: low invariant energy $\sqrt{s} \ $ \cite{TeisZP97}, adopting all
114: resonance parameters from the Manley analysis \cite{Manley}.  The high
115: energy collisions -- above $\sqrt{s}$ = 2.6~GeV for baryon-baryon
116: collisions and $\sqrt{s}$ = 2.2~GeV for meson-baryon collisions -- are
117: described by  the LUND string fragmentation model FRITIOF
118: \cite{FRITIOF}. This aspect is similar to that used in the HSD approach
119: \cite{Cass99,Ehehalt,Brat98,Geiss} and the UrQMD code \cite{Bass}. For
120: a detailed description of the underlying model at low energy we refer
121: the reader to Ref.~\cite{EffePhD}.
122: 
123: \subsection{A box with periodic boundary conditions}
124: 
125: In order to study 'infinite' hadronic matter problems we confine the
126: particles in a cubic box with periodic boundary conditions for their
127: propagation similar to a recent box calculation within the UrQMD model
128: \cite{Brav1}. We specify the initial conditions, i.e. baryon density
129: $\rho$, strange particle density $\rho_S$ and energy density
130: $\varepsilon$ as follows: first the initial system is fixed to $N_p=80$
131: protons and $N_n=80$ neutrons, which are randomly distributed in a
132: cubic box of volume $V$. The 3-momenta $\vec p_i$ of the nucleons in a
133: first step are randomly distributed inside a Fermi-sphere of radius
134: $p_F$ = 0.26~GeV/c (at $\rho_0$) and in a second step  boosted by $\pm
135: \beta_{cm}$ by a proper Lorentz transformation. Thus the initial baryon
136: density $\rho$ is fixed as $\rho=A/V$, $A=N_p+N_n$. The strange
137: particle density is set to zero as in related heavy-ion experiments
138: while the energy density is defined as $\varepsilon = E/V$, where $E$
139: is the total energy of all nucleons
140: \begin{eqnarray}
141: E=\sum\limits_i^A\sqrt{p_i^2+m_N^2}.
142: \label{energy_i}
143: \end{eqnarray}
144: The  boost velocity $\beta_{cm}$ is related to the initial energy density
145: $\varepsilon$ (excluding Fermi motion) as
146: \begin{eqnarray}
147: \beta_{cm}=\sqrt{1-{\rho^2 m_N^2\over \varepsilon^2} }
148: \label{betta}\end{eqnarray}
149: using
150: $\varepsilon = \gamma_{cm} \rho m_N$
151: with $\gamma_{cm}= 1/\sqrt{1-\beta_{cm}^2}$. Recall that $\rho_0 m_N \simeq
152: 0.15$~GeV/fm$^3$ so that an energy density $\varepsilon \simeq
153: 1.5$~GeV/fm$^3$ at density $\rho_0$ corresponds to $\gamma_{cm}\simeq
154: 10$, i.e. the SPS energy $T_{lab}\simeq 185$~A$\cdot$GeV. We thus start
155: with a 'true' nonequilibrium situation in order to mimique the  initial
156: stage in a relativistic heavy-ion collision. The initial phase
157: represents two interpenetrating, (ideally) infinitely extended fluids
158: of cold nuclear matter.
159: 
160: We now propagate all particles in the box in the cascade mode (without
161: mean-field potentials) using periodic boundary conditions, i.e.
162: particles moving out of the box are reinserted at the opposite side
163: with the same momentum. The phase-space distribution of particles then
164: can change due to elastic collisions,  resonance and string production
165: and their decays to mesons and baryons again. We recall that we include
166: all baryon resonances up to an invariant mass of 2 GeV and meson
167: resonances up to the $\phi$-meson. According to the initial conditions
168: for $\varepsilon$ and $\rho$ the factor $\gamma_{cm}$
169: determines if strings are excited in the very first collisions. This is
170: the case for $\gamma_{cm} > 1.4$ where the early equlibration stages
171: are dominated by string formation and decay.
172: 
173: \subsection{Chemical equilibration}
174: 
175: Figure \ref{Fig1} shows the time evolution of the various particle
176: abundances (nucleons $N$,  $\Delta$,  $\Lambda$, $\pi$ and  $K^+$
177: mesons) for density $\rho=\rho_0$ (left panel) at
178: energy density $\varepsilon =0.22$ GeV/fm$^3$ and for
179: $\rho=3\rho_0$ (right panel) at $\varepsilon =0.66$~GeV/fm$^3$.
180: These initial conditions correspond to bombarding
181: energy $T_{lab}$ per nucleon of roughly 2 A$\cdot$GeV.
182: In Fig.~\ref{Fig1} we count all particles which are
183: 'hadronized', i.e.\ produced by string decay after a  formation time of
184: $\tau_F=0.8$~fm/c in their rest frame.
185: \begin{figure}[h]
186: \centerline{\psfig{figure=cris1.eps,width=13cm}}
187: \vspace*{-5mm}
188: \caption{Time evolution of the various
189: particle abundances (nucleons $N$,  $\Delta$,  $\Lambda$, $\pi$
190: and  $K^+$ mesons) for density $\rho=\rho_0$ (left
191: panel) at energy density $\varepsilon =0.22$ GeV/fm$^3$ and for
192: $\rho=3\rho_0$ (right panel) at $\varepsilon =0.66$ GeV/fm$^3$.}
193: \label{Fig1}
194: \end{figure}
195: 
196: 
197: After several fm/c the number of nucleons decreases due to inelastic
198: collisions that produce either baryon resonances or additional mesons.
199: The number of $\Delta$-resonances grows up to a maximum in a few fm/c,
200: since a lot of $\Delta$'s are produced in the first $NN$ collisions;
201: their number subsequently decreases with time due to their decay and
202: excitation of further resonances or due to reabsorption.  The numbers
203: of $\pi$'s  and $\eta$'s increase very fast and reach the equilibrium
204: value within a few fm/c whereas the strange particles ($K^+,
205: K^-,\Lambda$) require a much longer time for equilibration.
206: 
207: For the higher energies the initial particle production proceeds via
208: the formation and decay of string excitations. This leads in particular
209: to a very early onset of strange particles (mainly kaons and hyperons)
210: within the first fm/c either due to the initial strings or due to
211: secondary or ternary baryon-baryon, meson-baryon and meson-meson
212: induced string-like interactions.
213: In Ref. \cite{Geiss} it was shown that
214: these early secondary and ternary reactions can contribute up to about
215: 50 $\%$ of the total strange particles obtained in a Pb~+~Pb reaction at
216: CERN SPS energies and thus explain the factor of 2 in the observed
217: relative strangeness enhancement compared to p+p reactions. This,
218: however, does not imply that chemical equilibrium for the dominant
219: strange particles has been achieved in this reaction, as our analysis
220: clearly shows. In the later stages, when the system has become, more or
221: less, isotropic in momentum space, strange particles can only be
222: further produced by low energy hadronic reactions, which, however, have
223: a considerable threshold and are thus strongly suppressed. This
224: explains the long chemical equilibration times for the strange
225: particles first demonstrated by Koch, M\"uller and Rafelski
226: \cite{Rafelski}.
227: 
228: In order to define an overall chemical equilibration time
229: we perform a fit to
230: the particle abundances $N(t)$ for pions and kaons as
231: \begin{eqnarray}
232: N(t) = N_{eq} \left(1 - \exp(-t/\tau_{eq})\right),
233: \label{taueq}\end{eqnarray}
234: where $N_{eq}$ is the equilibrium limit.  The equilibration time $\tau_{eq}$
235: thus corresponds to the time $t$ when $\simeq 63$\% of $N_{eq}$ is achieved.
236: 
237: Figure \ref{Fig4} shows the equilibration time $\tau_{eq}$ versus
238: energy density for $\pi$ and $K^+$ mesons at different baryon densities
239: of $1/3\rho_0, \rho_0, 3\rho_0$ and $6\rho_0$. We find that the
240: equilibration time for pions scales as $\tau_{eq}^\pi \sim 1/\rho$ or
241: $\Gamma_\pi \sim \rho$, thus we present the curve only for baryon
242: density $\rho_0$.  Whereas $\tau_{eq}^\pi$ slowly grows with
243: energy-density, $\tau_{eq}^K$ falls steeply with $\varepsilon$. This
244: marked difference is due to the fact that, on one hand, the kaon
245: production rate increases dramatically with $\sqrt{s}$ whereas that of
246: the pions, on the other hand, is more flat. With increasing energy thus
247: more strange particles are produced through strings especially from the
248: primary collisions with high $\sqrt{s}$ and the chemical equilibration
249: is achieved faster.
250: \begin{figure}[h]
251: \centerline{\psfig{figure=cris2.eps,width=10cm}}
252: \vspace*{-5mm}
253: \caption{Equilibration time $\tau_{eq}$ versus energy density $\varepsilon$
254: for $\pi$ and $K^+$ mesons at different baryon densities $1/3\rho_0,
255: \rho_0, 3\rho_0$ and $6\rho_0$.}
256: \label{Fig4}
257: \end{figure}
258: 
259: In Fig.~\ref{Fig4} we have considered an 'ideal' situation, i.e. hadron
260: matter at fixed energy and baryon density.  In realistic heavy-ion
261: collisions the system goes through the different stages due to
262: interactions and expansion. However, as follows from Fig. \ref{Fig4},
263: the equilibration time for strangeness is larger than 40~fm/c for all
264: energy and baryon densities. Thus in realistic nucleus-nucleus
265: collisions the chemical equilibration of strange particles requires
266: also a time above~40 fm/c which is considerably larger than the actual
267: reaction time of a few 10 fm/c or less \cite{ThermoNPA}.
268: 
269: The particle abundances used to extract $\tau_{eq}$ in Fig. \ref{Fig4}
270: have been calculated without any in-medium potentials.
271: In fact, the introduction of attractive potentials (especially for
272: $K^-$) will lower the hadronic thresholds and thus increase the
273: scattering rate between strange and nonstrange hadrons, whereas the
274: $K^+$ feels some repulsive potential and the trend goes in the opposite
275: way.  According to our calculations such in-medium modifications (in
276: line with Ref. \cite{Cass99}) give a correction to the $K^+$
277: equilibration times by atmost 10 \% and shortens the $K^-$
278: equilibration times up to 20 \% at density $\rho_0$.
279: 
280: \subsection{Thermal equilibration and limiting temperature}
281: 
282: In this subsection we investigate the approach to thermal equilibration.
283: For the equilibrated system we can extract a temperature $T$ by fitting the
284: particle spectra with the Boltzmann distribution
285: \begin{eqnarray}
286: {d^3N_i\over dp^3} \sim \exp(-E_i/T),
287: \label{Boltz}\end{eqnarray}
288: where $E_i=\sqrt{p_i^2+m_i^2}$ is the energy of particle $i$.  We note
289: that at the temperatures of interest here, the Bose and Fermi
290: distributions are practically identical to a Boltzmann distribution. We
291: find that in equilibrium  the spectra of all particles can be
292: characterized by one single temperature $T$ \cite{ThermoNPA}.
293: 
294: 
295: In the left panel of Fig.~\ref{Fig7} we display the
296: temperature $T$ versus energy density $\varepsilon$
297: for different baryon densities $\rho$:  $1/3\rho_0$
298: (open down triangles), $\rho_0$ (full squares), $3\rho_0$ (full dots),
299: $6\rho_0$ (full up triangles). In order to compare calculations for
300: different baryon densities we have subtracted the baryon energy density
301: at rest, i.e. $\simeq m_N\rho$ (except for Fermi motion). As seen from
302: Fig.~\ref{Fig7} the temperature grows with energy density up to a
303: limiting value reminiscent of a 'Hagedorn' temperature \cite{Hagedorn}.
304:  From our detailed investigations we obtain for the limiting temperature
305: $T_s \simeq 150\pm 5$~MeV which practically does not depend on baryon
306: density.  Such a singular behavior of $\varepsilon(T)$ for $T\simeq
307: T_s$ has also been found in the box calculations in Ref.~\cite{Brav1}
308: for $\rho=\rho_0$.  Our limiting temperature is slightly higher than
309: that in Ref.~\cite{Brav1} ($T_s = 130 \pm 10$~MeV)  due to the different
310: number of degrees of freedom; the model \cite{Brav1} contains more
311: resonances and uses a different threshold for string excitations.
312: Thus, there is some phenomenological sensitivity to the hadronic zoo of
313: particles and string thresholds employed in the model.
314: \begin{figure}[h]
315: \centerline{\psfig{figure=cris3.eps,width=13cm}}
316: \vspace*{-5mm}
317: \caption{Left panel: equilibrium temperature $T$ versus the energy
318: density  $\varepsilon-m_N\rho$ for different baryon densities $\rho$:
319: $1/3\rho_0$ (open down triangles), $\rho_0$ (full squares), $3\rho_0$
320: (full dots), $6\rho_0$ (full up triangles).
321: Right panel: equilibrium temperature $T$ versus the energy density for
322: baryon density $\rho=\rho_0$. The full dots  correspond to the
323: statistical model (SM) without strings, the full squares show our
324: box calculations including string degrees of freedom, while the
325: solid line shows the result from the extended SM including a Hagedorn
326: mass spectrum for strings.}
327: \label{Fig7}
328: \end{figure}
329: 
330: 
331: In order to investigate the equilibrium behavior of hadron matter we
332: also compare our transport (box) calculations with a simple Statistical
333: Model (SM) for an Ideal Hadron Gas where the system is described
334: by a grand canonical ensemble of non-interacting fermions and bosons in
335: equilibrium at temperature $T$.  All baryon and meson species
336: considered in the transport model \cite{Effe99gam} also have been
337: included in the statistical model \cite{ThermoNPA}.
338: 
339: 
340: Within the SM we find that the temperature increases continuously with
341: energy density since the continuum excitations, i.e. the string degrees
342: of freedom, are not included (full dots  in the right panel of
343: Fig.~\ref{Fig7}), whereas the box calculation with strings gives the
344: limiting temperature (full squares in Fig.~\ref{Fig7}).  Both curves in
345: Fig.~\ref{Fig7} have been calculated for density $\rho_0$.
346: 
347: To reproduce qualitatively our box result within the SM we have to include
348: continuum excitations in the statistical model, i.e.
349: a Hagedorn mass spectrum for strings \cite{Hagedorn} (for details
350: see \cite{ThermoNPA}).
351: For the 'Hagedorn' temperature $T_H$ we use the temperature $T_s$ as
352: obtained from the box calculations, i.e. $T_H=T_s\simeq 150$~MeV.
353: As seen in right panel of Fig.~\ref{Fig7} we achieve agreement of the
354: extended SM and our box calculations.
355: However, we point out that the limiting
356: temperature $T_s$ from our string model involves somewhat different
357: physics assumptions than the Hagedorn model at temperature $T_H$. $T_s$
358: should not really be identified with the 'Hagedorn' temperature $T_H$,
359: though close similarities exist. In the Hagedorn picture and for
360: temperatures close to $T_H$ the abundance of `normal' hadrons or known
361: resonances stays constant with increasing energy density whereas the
362: number and energy density of the (hypothetical) bootstrap excitations
363: diverges for $T\rightarrow T_H$. The Hagedorn model thus assumes
364: `particles' of mass $m\to\infty$ to be populated for $T\to T_H$, that
365: dynamically can be formed in collisions of high mass hadrons for
366: $t\to\infty$.  In contrast, our string model does not include energetic
367: string-string interactions that might produce more massive strings.
368: 
369: \section{SUMMARY}
370: 
371: In this contribution we have performed a systematic study of equilibration
372: phenomena and equilibrium properties of 'infinite' hadronic matter as
373: well as of relativistic nucleus-nucleus collisions using a BUU
374: transport model that contains resonance and string degrees-of-freedom.
375: The 'infinite' hadron matter is modelled  by initializing the system at
376: fixed baryon density, strange density and energy density by confining
377: it in a cubic box with periodic boundary conditions \cite{ThermoNPA}.
378: 
379: We have shown that the equilibration times $\tau_{eq}$ for different
380: particles depend on baryon density and energy density. The time
381: $\tau_{eq}$ for non-strange particles is much shorter than for
382: particles including strangeness; for kaons and antikaons the
383: equilibration time is found to be larger than $\simeq$ 40 fm/c for all
384: baryon and energy densities considered. The overall abundance of the
385: dominant strange particles (kaons and $\Lambda$'s) being produced and
386: obtained within the BUU cascade model for heavy-ion collisions can
387: therefore not be described by assuming a perfect chemical equilibrium
388: as strangeness is typically still undersaturated to a quite large
389: extent. We mention that transport model calculations like ours can
390: describe the yield and spectra of the produced nonstrange hadrons as
391: well as $K^+, K^-, \Lambda$ yields quite well at SPS energies
392: \cite{Cass99,Geiss}.  On the other hand, at AGS energies the measured
393: $K^+/\pi^+$ ratio in central Au~+~Au collisions is underestimated by
394: about 30\% \cite{CassQM}.  However, we have to point out that the more
395: exotic strange particles (like the measured antihyperon yields of
396: Ref.~\cite{WA97}) can by far not be explained within such standard
397: hadronic multiple channel reactions.  These hadronic data possibly
398: point towards new physics.
399: 
400: We have, furthermore, shown that thermal equilibrium is established
401: quickly, within about 5 fm/c at SIS energies and samewhat larger times
402: at high energies.  The inclusion of continuum excitations, i.e. hadron
403: 'strings', leads to a limiting temperature of $T_s \simeq 150$~MeV in
404: our transport approach which practically does not depend on the baryon
405: density and energy.  We have compared our results with the
406: statistical model (SM), which contains the same degrees of freedom and
407: the same spectral functions of particles as our transport model. We
408: found that the limiting temperature behaviour can be reproduced in the
409: statistical model only after including continuum excitations of the
410: Hagedorn type, otherwise the fireball temperature extracted from the
411: particle abundances and spectra is overestimated substantially.
412: 
413: Close to the critical temperature $T_s$, the hadronic energy densities
414: can increase to a couple of GeV/fm$^3$. From lattice QCD calculations
415: one expects that a phase transition to a potentially deconfined QGP
416: state should occur. Referring to the limiting temperature $T_s\approx
417: 150 $ MeV obtained, a QGP should be revealed and clearly distinguished from
418: a hadronic state of matter if one can unambiguously prove the
419: existence of an equilibrated and thermal phase of strongly interacting
420: matter with temperatures exceeding, e.g., 200 MeV.  The best candidates
421: are electromagnetic probes, either direct photons or dileptons. On the
422: other hand these are also `contaminated' by hadronic background and/or
423: preequilibrium physics. So far no thermal electromagnetic source with
424: temperatures larger or equal than 200 MeV has been clearly identified.
425: This might happen at RHIC energies in central Au~+~Au collisions which
426: are expected to be studied soon.
427: 
428: 
429: %---------------------------------------------------------------------
430: %\newpage
431: \begin{thebibliography}{999}
432: \bibitem{QM}
433:     QUARK MATTER '96, Nucl. Phys. A 610 (1997) 1;
434:     QUARK MATTER '97, Nucl. Phys. A 638 (1998) 1;
435:     QUARK MATTER '99, Nucl. Phys. A 669 (1999) 1.
436: \bibitem{BotMal90}
437:     W. Botermans and R. Malfliet, Phys. Rep. 198 (1990) 115.
438: \bibitem{Henning}
439:     P.A. Henning, Phys. Rep. 253 (1995) 235.
440: \bibitem{Ko}
441:     G. Q. Li and C. M. Ko, J. Phys. G 22 (1997) 1673.
442: \bibitem{Bass}
443:     S. A. Bass et al., Prog. Part. Nucl. Phys. 42 (1998) 279;
444:     J. Phys. G 25 (1999) 1859.
445: \bibitem{Cass99}
446:     W. Cassing and E. L. Bratkovskaya, Phys. Rep. 308 (1999) 65.
447: \bibitem{BM}
448:        P. Braun-Munzinger, J. Stachel, J.P. Wessels, and N. Xu,
449:         Phys. Lett. B 344 (1995)~43;
450:        Phys. Lett. B 365 (1996)~1;
451:     J. Stachel, Nucl. Phys. A 654 (1999) 119.
452: \bibitem{Satz}
453:     J. Cleymans and H. Satz, Z. Phys. C 57 (1993) 135.
454: \bibitem{Sollfrank}
455:      J. Sollfrank, M. Gazdzicki, U. Heinz and J. Rafelski,
456:      Z. Phys. C  61 (1994) 659;
457:      F. Becattini, M. Gazdzicki and J. Sollfrank,
458:      Eur. Phys. J. C 5 (1998) 143.
459: \bibitem{Spieles}
460:      C. Spieles, H. St\"ocker and C. Greiner,
461:      Eur. Phys. J. C 2 (1998) 351.
462: \bibitem{Cleymans}
463:     J. Cleymans, H. Oeschler, and K. Redlich,
464:     nucl-th/9809027; J. Phys. G 25 (1999) 281.
465: \bibitem{Hydro}
466:     H. St\"ocker and W. Greiner, Phys. Rep. 137 (1986) 277.
467: \bibitem{Rischke}
468:     U. Ornik et al., Phys. Rev. C 54 (1996) 1381;
469:     S. Bernard et al., Nucl. Phys. A 605 (1996) 566;
470:     J. Sollfrank et al., Phys. Rev. C 55 (1997) 392.
471: \bibitem{Rafelski}
472:     P. Koch, B. M\"uller, and J. Rafelski, Phys. Rep. 142 (1986) 167.
473: \bibitem{Cass90}
474:     W. Cassing, V. Metag, U. Mosel, and K. Niita,
475:     Phys. Rep. 188 (1990) 363.
476: \bibitem{Lang91}
477:     A. Lang, B. Bl\"attel, W. Cassing, V. Koch, U. Mosel, and K. Weber,
478:     Z. Phys. A 340 (1991) 287.
479: \bibitem{Bl93}
480:     B. Bl\"attel, V. Koch, and U. Mosel,
481:     Rep. Progr. Phys. 56 (1993) 1.
482: \bibitem{Brav1}
483:     M. Belkacem, M. Brandstetter, S.A. Bass et al.,
484:     Phys. Rev. C 58 (1998) 1727.
485: \bibitem{Brav2}
486:     L.V. Bravina, M.I. Gorenstein, M. Belkacem et al.,
487:     Phys. Lett. B 434 (1998) 379;
488:     L.V. Bravina, M. Brandstetter, M.I. Gorenstein et al.,
489:     J. Phys. G 25 (1999) 351.
490: \bibitem{Brav3}
491:     L.V. Bravina, E.E. Zabrodin, M.I. Gorenstein et al.,
492:     Phys. Rev. C 60 (1999) 024904.
493: \bibitem{Solfr99}
494:     J. Sollfrank, U. Heinz, H. Sorge, N. Xu, Phys. Rev. C 59 (1999) 1637.
495: \bibitem{TeisZP97}
496:     S. Teis, W. Cassing, M. Effenberger, A. Hombach, U. Mosel,
497:     and Gy. Wolf, Z. Phys. A 356 (1997) 421.
498: \bibitem{Effe99gam}
499:     M. Effenberger, E.L. Bratkovskaya, and U. Mosel,
500:     Phys. Rev. C 60 (1999) 044614.
501: \bibitem{EffePhD}
502:     M. Effenberger, Ph.D. Thesis, Univ. of Giessen, 1999;
503:     http://theorie.physik.uni-giessen.de/ftp.html.
504: \bibitem{Manley}
505:        D. M. Manley and E. M. Saleski, Phys. Rev. D 45 (1992) 4002.
506: \bibitem{FRITIOF}
507:     B. Anderson, G. Gustafson and Hong Pi, Z. Phys. C 57 (1993) 485.
508: \bibitem{Ehehalt}
509:     W. Ehehalt and W. Cassing, Nucl. Phys. A 602 (1996) 449.
510: \bibitem{Brat98}
511:     E. L. Bratkovskaya and W. Cassing, Nucl. Phys. A 619 (1997) 413.
512: \bibitem{Geiss}
513:     J. Geiss, W. Cassing, and C. Greiner, Nucl. Phys. A 644 (1998) 107.
514: \bibitem{ThermoNPA}
515: 	E. L. Bratkovskaya, W. Cassing, C. Greiner,
516: 	M. Effenberger, U. Mosel and A. Sibirtsev,
517: 	Nucl. Phys. A 675 (2000) 661.
518: \bibitem{Hagedorn}
519:     R.Hagedorn, Suppl. Novo Cimento 3 (1965) 147;
520:     Suppl. Novo Cimento 6 (1965) 311;
521:     R.Hagedorn and J. Ranft, Suppl. Novo Cimento 6 (1968) 169.
522: \bibitem{CassQM}
523: 	W. Cassing, Nucl. Phys. A 661 (1999) 468c.
524: \bibitem{WA97}
525:     E. Andersen et al.,
526:     J. Phys. G 25 (1999) 171; J. Phys. G 25 (1999) 181.
527: \end{thebibliography}
528: 
529: \end{document}
530: %---------------------------------------------------------------------
531: #!/bin/csh -f
532: # this uuencoded Z-compressed .tar file created by csh script  uufiles
533: # for more information, see e.g. http://xxx.lanl.gov/faq/uufaq.html
534: # if you are on a unix machine this file will unpack itself:
535: # strip off any mail header and call resulting file, e.g., figures.uu
536: # (uudecode ignores these header lines and starts at begin line below)
537: # then say        csh figures.uu
538: # or explicitly execute the commands (generally more secure):
539: #    uudecode figures.uu ;   uncompress figures.tar.Z ;
540: #    tar -xvf figures.tar
541: # on some non-unix (e.g. VAX/VMS), first use an editor to change the
542: # filename in "begin" line below to figures.tar_Z , then execute
543: #    uudecode figures.uu
544: #    compress -d figures.tar_Z
545: #    tar -xvf figures.tar
546: #
547: uudecode $0
548: chmod 644 figures.tar.Z
549: zcat figures.tar.Z | tar -xvf -
550: rm $0 figures.tar.Z
551: exit
552: 
553: