1: %final version; 16.12.99
2: %--------------------------------------------------------------------
3: \documentstyle[12pt,fleqn,rotate,epsfig]{article}
4:
5: \textwidth 16.5cm \textheight 23cm
6: \oddsidemargin 1mm
7: \evensidemargin 1mm
8: %\topmargin=1cm
9: \topmargin -40pt
10: %--------------------------------------------------------------------
11:
12: \begin{document}
13:
14: \title{Excitation functions of hadronic observables from SIS to RHIC
15: energies\thanks{supported by BMBF and GSI Darmstadt} }
16:
17: \author{W. Cassing, E. L. Bratkovskaya and S. Juchem \\[5mm]
18: {\normalsize Institut f\"{u}r Theoretische Physik,}\\
19: {\normalsize Universit\"{a}t Giessen,}
20: {\normalsize 35392 Giessen, Germany} }
21:
22: \date{ }
23: \maketitle
24:
25: \begin{abstract}
26: We calculate excitation functions for various dynamical quantities
27: as well as experimental observables from SIS to RHIC energies
28: within the HSD transport approach which is based on string, quark,
29: diquark ($q, \bar{q}, qq, \bar{q}\bar{q}$) and hadronic degrees of
30: freedom without including any explicit phase transition to a
31: quark-gluon plasma (QGP). It is argued that the failure of this
32: more 'conventional' approach in comparison to experimental data
33: should indicate the presence of a different phase which might be
34: either attributed to space-time regions of vanishing scalar quark
35: condensate ($<q\bar{q}>$ = 0) or to the presence of a QGP phase
36: with strongly interacting partons. We study the $K/\pi$ ratio, the
37: low mass dilepton enhancement in the invariant mass regime from
38: 0.2 -- 1.2 GeV as well as charmonium suppression for central
39: Au~+~Au collisions as a function of the bombarding energy and
40: present predictions for these observables as well as hadron
41: rapiditiy distributions at RHIC energies. Whereas all observables
42: studied within HSD show smooth increasing/decreasing excitation
43: functions, the experimental $K^+/\pi^+$ ratio indicates a maximum
44: at 11 A$\cdot$GeV (or above) which is interpreted as a signature
45: for a chirally restored phase in the course of the reaction.
46: \end{abstract}
47:
48: \vspace{1cm} \noindent PACS: 24.10.-i; 25.75.-q; 11.30.Rd;
49: 13.60.-r
50:
51: \noindent Keywords: Nuclear-reaction models and methods;
52: Relativistic heavy-ion collisions; Chiral symmetries; Meson
53: production
54:
55:
56: \newpage
57:
58:
59: %------------------------------------------------------------------------
60: \section{Introduction}
61:
62:
63: The ultimate goal of relativistic nucleus-nucleus collisions is to
64: reanalyze the early 'big-bang' under laboratory conditions and to
65: find the 'smoking gun' for a phase transition from the initial
66: quark-gluon plasma (QGP) to a phase characterized by an
67: interacting hadron gas \cite{Harris,Shuryak93,QM99}. Any
68: theoretical approach might describe such a phase transition
69: starting from the partonic side with strongly interacting quarks
70: and gluons \cite{Geiger3,Bernd} or from the hadronic side by
71: involving hadronic degrees of freedom \cite{Sorge99,Basrev}, i.e.
72: hadrons with proper self-energies or spectral functions at high
73: baryon density or temperature. It remains to be seen which
74: approach will prove to be more successful, economic and
75: transparent.
76:
77: Nucleus-nucleus collisions with initial energies per nucleon of
78: $\sqrt{s}$ = 200~GeV or $\approx$21.5~A$\cdot$TeV will be
79: available soon at the Relativistic-Heavy-Ion-Collider (RHIC) in
80: Brookhaven. In central collisions of Au~+~Au here energy densities
81: above 5 GeV/fm$^3$ are expected such that the critical energy
82: density for a QGP phase should be overcome in considerable
83: space-time volumes where the relevant degrees of freedom are
84: partons (quarks and gluons). Parton cascade calculations have been
85: used so far \cite{Geiger3,Geiger1,Wang} to estimate the energy
86: densities and particle production yields in violent reactions at
87: $\sqrt{s}$ = 200~GeV, an order of magnitude higher than at SPS
88: energies ($\sqrt{s} \approx$ 20 GeV). Intuitively one expects that
89: the initial nonequilibrium phase of a nucleus-nucleus collision at
90: RHIC energies should be described by parton degrees of freedom
91: whereas hadrons are only formed (by 'condensation') at a later
92: stage of the reaction which might be a couple of fm/c from the
93: initial contact of the heavy ions. Thus parton cascade
94: calculations -- including transitions rates from perturbative QCD
95: -- should be adequate for all initial reactions involving a large
96: 4-momentum transfer between the consituents since QCD is well
97: tested in its short distance properties. The question, however,
98: remains to which extent the parton calculations can be
99: extrapolated to low $Q^2$ where hadronic scales become important.
100: As a rough estimate one can employ here the average mass of vector
101: mesons, the nucleon and its first excited state, which gives
102: $Q^2_{crit} \approx$ 1 GeV$^2$. On the other hand, using the
103: uncertainty relation this implies time scales of $\Delta t<$ 0.2
104: fm/c = $Q_{crit}^{-1}$ or relative separations of partons $\Delta
105: r <$ 0.2 fm, which are small compared to the hadronic size or
106: average life time of the $\rho, \Delta, etc.$ in free space or the
107: formation time of hadrons $t_F \approx 0.7-0.8$ fm/c as used in
108: the HSD transport approach \cite{Cass99,Ehehalt,Jgeiss}.
109:
110: Turning the argument around, a nonequilibrium hadronic approach
111: involving a time scale of 0.7-0.8 fm/c cannot tell anything about
112: shorter times because the uncertainty relation does not allow to
113: distinguish states which are separated in mass by less than
114: $\approx$ 300 MeV = $t_F^{-1}$, which is the $N-\Delta$ mass
115: difference. Thus one faces the problem that neither the parton
116: description nor a nonequilibrium hadronic model should be valid
117: for times 0.2 fm/c $\leq t \leq$ 0.7-0.8 fm/c in individual
118: hadronic reactions, which corresponds to the nonperturbative
119: formation time of the hadronic wavefunction. This regime of the
120: 'soft' QCD physics is presently not well understood and
121: appropriate dynamical models are urgently needed. In the HSD
122: approach these intermediate times are described by color neutral
123: strings, where the leading quarks and 'diquarks' in a baryonic
124: string (or quarks and antiquarks in a mesonic string etc.) are
125: allowed to rescatter again with hadronic cross sections divided by
126: the number of constituent quarks and antiquarks in the hadrons,
127: respectively \cite{Jgeiss}.
128:
129: Furthermore, the question of chiral symmetry restoration at high
130: baryon density and/or high temperature is of fundamental interest,
131: too \cite{Harris,Shuryak93}. Whereas lattice QCD calculations at
132: zero baryon chemical potential indicate that a restoration of
133: chiral symmetry goes along with the deconfinement phase transition
134: at some critical temperature $T_c$, the situation is less clear
135: for finite baryon density where QCD sum rule studies indicate a
136: linear decrease of the scalar quark condensate $<\bar q q>$ --
137: which is nonvanishing in the vacuum due to a spontaneous breaking
138: of chiral symmetry -- with baryon density $\rho_B$ towards a
139: chiral symmetric phase characterized by $<\bar{q} q>$ = 0. This
140: decrease of the scalar condensate is expected to lead to a change
141: of the hadron properties with density and temperature, i.e. in a
142: chirally restored phase the hadrons might become approximately
143: massless \cite{Brown} or at least vector and axial vector currents
144: should become equal \cite{Kochr,Zahed}; the latter implies that
145: the $\rho$ and $a_1$ spectral functions should become identical.
146: Since the scalar condensate $<q\bar{q}>$ is not a direct
147: observable, its manifestations should be found in different
148: hadronic abundancies and spectra.
149:
150: Nowadays, our knowledge about the hadron properties at high
151: temperature or baryon density is based on heavy-ion experiments from
152: BEVALAC/SIS to SPS energies where hot and dense nuclear
153: systems are produced on a timescale of a few fm/c. As mentioned
154: above, the information from ultrarelativistic nucleus-nucleus
155: collisions at RHIC, i.e. initial $\sqrt{s} = $ 200 GeV per
156: nucleon, will be available soon \cite{QM99}. However, any
157: conclusions about the properties of hadrons in the nuclear
158: environment are based on the comparison of experimental data with
159: nonequilibrium kinetic transport theory
160: \cite{Cass99,Stoecker,Bertsch,Cass90,Koreview}. As a genuine
161: feature of transport theories there are two essential ingredients:
162: i.e. the {\it baryon (and meson) scalar } and {\it vector
163: self-energies} as well as {\it in-medium elastic} and {\it
164: inelastic cross sections} for all hadrons involved. Whereas in the
165: low-energy regime these 'transport coefficients' can be calculated
166: in the Dirac-Brueckner approach starting from the bare
167: nucleon-nucleon interaction \cite{Bot,Mal,T3} this is no longer
168: possible at high baryon density $(\rho_B \geq 2$-$3 \rho_0)$ and
169: high temperature, since the number of independent hadronic degrees
170: of freedom increases drastically and the interacting hadronic
171: system should enter a phase with $<q\bar{q}> \approx$ 0
172: \cite{Brown,Kochr,GEB,GEB1,Koch} as mentioned before. As a
173: consequence the hadron self-energies or spectral functions in the
174: nuclear medium will change substantially especially close to the
175: chiral phase transition and transport theoretical studies should
176: include the generic properties of QCD in line with nonperturbative
177: computations on the lattice \cite{Lat1,Lat2,Lat3,Lat4}.
178:
179: In this work we concentrate on excitation functions of hadronic
180: observables from SIS to RHIC energies with the aim to find out
181: optimal experimental conditions to search for 'traditional'
182: phenomena such as strangeness enhancement (Section 3), low mass
183: dilepton enhancement (Section 4) or charmonium suppression in
184: nucleus-nucleus collisions (Section 5). Our studies are performed
185: within the HSD transport approach that has been described in Refs.
186: \cite{Ehehalt,Jgeiss} and been tested for $p+p$, $p+A$ and $A+A$
187: collisions from the SIS to SPS energy regime \cite{Cass99}. Actual
188: predictions for hadron rapidity spectra and $J/\Psi$ suppression
189: will be presented in Section 6 while Section 7 concludes the study
190: with a summary.
191:
192: \section{Theoretical considerations}
193:
194: In this Section we briefly recall the ingredients of the covariant
195: transport theory, that has been denoted as {\bf H}adron-{\bf
196: S}tring-{\bf D}ynamics (HSD) \cite{Ehehalt}, which formally can be
197: written as a coupled set of transport equations for the
198: phase-space distributions $f_h (x,p)$ of hadron $h$
199: \cite{Ehehalt,Cass,Weber1}, i.e.
200: \begin{eqnarray}
201: && \left\{ \left( \Pi_\mu-\Pi_\nu\partial_\mu^p U_h^\nu
202: - M_h^*\partial^p_\mu U_h^S \right)\partial_x^\mu
203: + \left( \Pi_\nu \partial^x_\mu U^\nu_h+
204: M^*_h \partial^x_\mu U^S_h\right) \partial^\mu_p \right\}
205: f_h(x,p)\nonumber\\
206: && = \sum_{h_2 h_3 h_4\ldots} \int d2 d3 d4 \ldots
207: [G^{\dagger}G]_{12\rightarrow 34\ldots}
208: \delta^4_{\Gamma}(\Pi +\Pi_2-\Pi_3-\Pi_4 \ldots ) \nonumber\\
209: && \times \left\{ f_{h_3}(x,p_3)f_{h_4}(x,p_4)\bar{f}_h(x,p)
210: \bar{f}_{h_2}(x,p_2)\right. \nonumber\\
211: && -\left. f_h(x,p)f_{h_2}(x,p_2)\bar{f}_{h_3}(x,p_3)
212: \bar{f}_{h_4}(x,p_4) \right\} \ldots\ \ .
213: \label{Ehg24} \end{eqnarray} In Eq.~(\ref{Ehg24}) $U_h^S(x,p)$ and
214: $U_h^\mu(x,p)$ denote the real part of the scalar and vector
215: hadron self-energies, respectively, while $[G^+G]_{12\rightarrow
216: 34\ldots} \delta^4_{\Gamma} (\Pi +\Pi_2-\Pi_3-\Pi_4\ldots )$ is
217: the 'transition rate' for the process $1+2\rightarrow 3+4+\ldots$
218: which is taken to be on-shell in the semiclassical limit adopted.
219: The hadron quasi-particle properties in (\ref{Ehg24}) are defined
220: via the mass-shell constraint \cite{Weber1},
221: \begin{equation}
222: \delta (\Pi_\mu\Pi^\mu-M_h^{*2} ) \ \ ,
223: \label{Ehg25}\end{equation} with effective masses and momenta (for
224: a hadron of bare mass $M_h$ and momentum $p^\mu$) given by
225: \begin{eqnarray}
226: M_h^* (x,p)&=&M_h + U_h^{{S}}(x,p) \nonumber \\
227: \Pi^\mu (x,p)&=&p^\mu-U^\mu_h (x,p)\ \ ,
228: \label{Ehg26}\end{eqnarray}
229: while the phase-space factors
230: \begin{equation}
231: \bar{f}_h (x,p)=1 \pm f_h (x,p) \label{Ehg26a}\end{equation} are
232: responsible for fermion Pauli-blocking or Bose enhancement,
233: respectively, depending on the type of hadron in the final/initial
234: channel. The dots in Eq.~(\ref{Ehg24}) stand for further
235: contributions to the collision term with more than two hadrons in
236: the final/initial channels. The transport approach (\ref{Ehg24})
237: is fully specified by $U_h^S(x,p)$ and $U_h^\mu(x,p)$ $(\mu
238: =0,1,2,3)$, which determine the mean-field propagation of the
239: hadrons, and by the transition rates $G^\dagger G\ \delta^4
240: (\ldots )$ in the collision term, that describe the scattering and
241: hadron production/absorption rates.
242:
243: In Ref. \cite{Ehehalt} the scalar and vector mean-fields $U_h^S$
244: and $U^\mu_h$ have been determined in the mean-field limit from an
245: effective hadronic Lagrangian density ${\cal L}_H$ that has been
246: fitted to the equation of state of nucleonic matter as resulting from
247: the Nambu-Jona-Lasinio (NJL) model. Without going through the
248: arguments again we show in Fig. \ref{Fig1} the nucleon scalar
249: ($U_S$) and negative vector potential ($-U_0$) as a function of
250: the nuclear density $\rho$ and relative momentum ${\bf p}$ of the
251: nucleon with respect to the nuclear matter rest frame. Whereas the
252: vector potential increases practically linearly with density (at
253: low momentum ${\bf p}$) the scalar potential saturates with
254: density such that the nucleon effective mass $M^* = M_0+U_S$
255: almost drops to zero for $\rho \geq$ 0.6 fm$^{-3}$. Both
256: potentials decrease rather fast in magnitude with momentum ${\bf
257: p}$ and practically vanish above a few GeV/c.
258:
259: In Fig. \ref{Fig2} the real part of the potential
260: \begin{eqnarray}
261: U_{SEP} &=& U_0(\rho_0,{\bf p}) + \sqrt{{\bf p}^2 + (M_N +
262: U_S)^2} - \sqrt{{\bf p}^2 + M_N^2} \label{Ehg23}\end{eqnarray} is
263: shown again as a function of $\rho$ and ${\bf p}$. Whereas we see
264: a net attraction for momenta $|{\bf p}| \leq$ 0.5~GeV/c up to
265: densities of $\approx$ 0.3~fm$^{-3}$, the net potential becomes
266: repulsive for higher momenta, reaches a maximum repulsion at
267: $|{\bf p}| \approx$ 1 GeV/c and then drops again with $|{\bf p}|$.
268: We mention that at density $\rho_0$ the potential $U_{SEP}$
269: compares well with the potential from the data analysis of Hama et
270: al. \cite{Hama} as well as Dirac-Brueckner computations from
271: \cite{Mal} up to a kinetic energy $E_{kin}$ of 1 GeV
272: \cite{Ehehalt}. The formula (\ref{Ehg23}) reduces to the familiar
273: expression for the Schroedinger equivalent potential (Eq. (3.16)
274: of Ref. \cite{Cass99}) in the low density limit.
275:
276:
277: In our transport calculations we include nucleons, $\Delta$'s,
278: N$^*$(1440), N$^*$(1535), $\Lambda$, $\Sigma$ and $\Sigma^*$
279: hyperons, $\Xi$'s and $\Omega$'s as well as their antiparticles.
280: In a first approximation we assume here that all baryons (made out
281: of light $(u,d)$ quarks) have the same scalar and vector
282: self-energies as the nucleons while the hyperons pick up a factor
283: 2/3 according to the light quark content and $\Xi$'s a factor of
284: 1/3, respectively.
285:
286: In the HSD approach the high energy inelastic hadron-hadron
287: collisions are described by the FRITIOF model \cite{LUND}, where
288: two incoming hadrons emerge the reaction as two excited color
289: singlet states, i.e. 'strings'. According to the Lund model
290: \cite{LUND} a string is characterized by the leading constituent
291: quarks of the incoming hadron and a tube of color flux is supposed
292: to be formed connecting the rapidly receding string-ends. In the
293: HSD approach baryonic ($qq-q$) and mesonic ($q-\bar{q}$) strings
294: are considered with different flavors ($q = u,d,s$). In the
295: uniform color field of the strings virtual $q\bar{q}$ or
296: $qq\bar{q}\bar{q}$ pairs are produced causing the tube to fission
297: and thus to create mesons or baryon-antibaryon pairs. The
298: production probability $P$ of massive $s\bar{s}$ or
299: $qq\bar{q}\bar{q}$ pairs is suppressed in comparison to light
300: flavor production ($u\bar{u}$, $d\bar{d}$) according to a
301: Schwinger-like formula \cite{Schwinger}
302: \begin{eqnarray}
303: {P(s\bar{s}) \over P(u\bar{u})} = \gamma_s = \exp\left(-\pi
304: {m_s^2-m_q^2\over 2\kappa}, \right)
305: \label{schwinger}
306: \end{eqnarray}
307: with $\kappa\approx 1$~GeV/fm denoting the string tension. Thus in
308: the Lund string picture the production of strangeness and
309: baryon-antibaryon pairs is controlled by the constituent quark and
310: diquark masses. Inserting the constituent quark masses $m_u=0.3$~GeV
311: and $m_s=0.5$ GeV a value of $\gamma_s \approx 0.3$ is
312: obtained. While the strangeness production in proton-proton
313: collisions at SPS energies is reasonably well reproduced with this
314: value, the strangeness yield for p~+~Be collisions at AGS energies
315: is underestimated by roughly 30\% (cf. \cite{Jgeiss}). For that
316: reason the relative factors used in the HSD model are
317: \cite{Jgeiss}
318: \begin{eqnarray}
319: u:d:s:uu = \left\{
320: \begin{array}{ll}
321: 1:1:0.3:0.07 &, \mbox{at SPS to RHIC energies} \\ 1:1:0.4:0.07 &,
322: \mbox{at AGS energies} ,
323: \end{array}
324: \right. \label{HSD-supp}
325: \end{eqnarray}
326: with a linear transition of the strangeness suppression factor
327: $\gamma_s = s:u$ as
328: a function of $\sqrt{s}$ in between.
329:
330: Additionally a fragmentation function $f(x,m_t)$ has to be
331: specified, which is the probability distribution for hadrons with
332: transverse mass $m_t$ to acquire the energy-momentum fraction $x$ from
333: the fragmenting string,
334: \begin{eqnarray}
335: f(x,m_t)\approx {1 \over x} (1-x)^a \exp\left(-bm_t^2/x \right),
336: \end{eqnarray}
337: with $a=0.23$ and $b=0.34$~GeV$^{-2}$ \cite{Jgeiss}.
338:
339: We recall that the LUND model \cite{LUND} includes partonic
340: diffractive scattering and mini-jet production as well
341: \cite{PYTHIA}. The latter phenomena are not important at SPS
342: energies and below, however, become appreciable at RHIC energies.
343: In this respect the HSD approach dynamically also includes the
344: hard partonic processes as far as quarks and antiquarks are
345: involved. However, it does not employ hard gluon-gluon processes
346: beyond the level of 'string phenomenology'. This has to be kept in
347: mind with respect to the predictive power of the model at RHIC
348: energies and beyond.
349:
350: The medium modifications due to the hadron self-energies,
351: furthermore, require to introduce some conserving approximations
352: in the collision terms in line with the modified quasi-particle
353: properties. Since these in-medium modifications -- related to 'low
354: momentum physics' -- are not of primary interest in this study we
355: discard an explicit discussion here and refer the reader to Ref.
356: \cite{Ehehalt}.
357:
358:
359:
360: \subsection{The scalar condensate}
361:
362: The scalar quark condensate $<q\bar{q}>$ is viewed as an order
363: parameter for the restoration of chiral symmetry at high baryon
364: density and temperature. A model independent relation for the
365: scalar quark condensate at finite (but small) baryon density and
366: temperature has been given by Drukarev and Levin \cite{Drukarev},
367: i.e
368: \begin{equation}
369: \label{condens} \frac{<q\bar{q}>}{<q\bar{q}>_V} = 1 -
370: \frac{\rho_S}{f_\pi^2 m_\pi^2} \left[\Sigma_{\pi} + m \frac{d}{dm}
371: \left(\frac{E(\rho)}{A}\right) \right],
372: \end{equation}
373: where $<q\bar{q}>_V$ denotes the vacuum condensate, $\Sigma_\pi \approx$
374: 45 MeV is the pion-nucleon $\Sigma$-term, $f_\pi$ and $m_\pi$ the
375: pion decay constant and pion mass, respectively, while $E(\rho)/A$ is
376: the binding energy per nucleon. Since for low
377: densities the scalar density $\rho_S$ in (\ref{condens}) may be
378: replaced by the baryon density $\rho_B$, the change in the scalar
379: quark condensate starts linearly with $\rho_B$ and is reduced by a
380: factor 1/3 at saturation density $\rho_0$. A simple linear
381: extrapolation then would indicate that at $\rho_B \approx 3
382: \rho_0$ a restoration of chiral symmetry might be achieved in
383: heavy-ion collisions.
384:
385: A reasonable estimate for the scalar condensate in dynamical
386: calculations has been suggested by Friman et al. \cite{Toneev98},
387: \begin{equation}
388: \frac{<q\bar{q}>}{<q\bar{q}>_V} = 1 - \frac{\Sigma_\pi}{f_\pi^2
389: m_\pi^2}\rho_S - \sum\limits_h{\sigma_h \rho_S^h \over f_\pi^2
390: m_\pi^2}, \label{condens2} \end{equation}
391: where $\sigma_h$ denotes
392: the $\sigma$-commutator of the relevant mesons $h$. For pions and
393: mesons made out of light quarks and antiquarks we use $\sigma_h =
394: m_\pi/2$ whereas for mesons with a strange (antistrange) quark we
395: adopt $\sigma_h = m_\pi/4$ according to the light quark content.
396: Within the same spirit the $\Sigma$-term for hyperons is taken as
397: 2/3 $\Sigma_{\pi} \approx$ 30 MeV while for $\Xi$'s we use 1/3
398: $\Sigma_{\pi} \approx$ 15 MeV.
399:
400: The scalar density of mesons (of type $h$) is given by
401: \begin{equation}
402: \label{mesons} \rho_S^h = \frac{(2s+1) (2t+1)}{(2\pi)^3} \int d^3
403: {\bf p} \frac{m_h}{\sqrt{{\bf p}^2 + m_h^2}} f_h({\bf r},{\bf
404: p};t),
405: \end{equation}
406: with $f_h$ denoting the meson phase-space distribution
407: of species $h$. In (\ref{mesons}) $s,t$ denote the spin and
408: isospin quantum numbers, respectively. We note that the scalar
409: density of baryons is calculated in line with (\ref{mesons}) by
410: replacing the mass and momentum by the effective quantities in
411: (\ref{Ehg26}).
412:
413: The actual numerical result for the space-time dependence of the
414: scalar condensate (\ref{condens2}) is shown in Figs. \ref{Fig3}
415: and \ref{Fig4} for central Au~+~Au collisions at 6~A$\cdot$GeV and
416: 20~A$\cdot$GeV, respectively. Here the condensate is divided by
417: the vacuum condensate $<q\bar{q}>_V$ such that the nonperturbative
418: vacuum is characterized by a value of 1. Furthermore, the
419: $z$-direction has been stretched by the Lorentz-factors
420: $\gamma_{cm}$ to compensate for Lorentz contraction, while
421: negative numerical values for the condensate have been suppressed.
422: According to (\ref{condens}), (\ref{condens2}) the scalar
423: condensate is reduced inside the approaching nuclei by about 35\%;
424: this reduction becomes more pronounced when the nuclei achieve
425: full overlap. As seen from Fig. \ref{Fig3} already at 6
426: A$\cdot$GeV there is a substantial space-time region of vanishing
427: scalar condensate for 6~fm/c~$\leq t \leq 13$~fm/c, where the
428: conventional hadronic picture is not expected to hold anymore. The
429: space-time volume of vanishing quark condensate slightly increases
430: for 20 A$\cdot$GeV (Fig. \ref{Fig4}), however, the increase is
431: only moderate and resembles very much the situation for the
432: space-time integral of high density baryon matter (cf. Fig. 1 in
433: Ref. \cite{brat97}).
434:
435: We mention that at higher bombarding energies the 4-volume ($x =
436: (t,{\bf r})$)
437: \begin{eqnarray}
438: V(\alpha)= \int d^3{\bf r} dt \ \Theta(\alpha -
439: \frac{<q\bar{q}>_x}{<q\bar{q}>_V}) , \label{volume} \end{eqnarray}
440: that counts the fraction of the scalar condensate below the value
441: of $\alpha$ (0 $\leq \alpha \leq$ 0.3) is essentially determined
442: by the pion density whereas below about 20 A$\cdot$GeV the
443: dominant contribution stems from the baryons (cf. also Ref.
444: \cite{Toneev98}). Since the pion density can be considered as a
445: measure of the vacuum 'temperature', a chiral order transition
446: below about 20 A$\cdot$GeV is dominated by the baryon density
447: while especially at SPS or even RHIC energies a chiral order
448: transition is due to 'temperature'.
449:
450: The question now arises, if there are proper experimental
451: observables that allow to trace down such type of phase transition
452: (or cross over). When gating on central collisions of Au~+~Au (or
453: Pb~+~Pb) such phenomena should show up in their excitation
454: functions. We note that in a pure hadronic transport approach we
455: expect a smooth behaviour of practically all observables with
456: bombarding energy due to an increase of thermal excitation energy.
457: This is not so obvious for the HSD approach where a gradual
458: transition from hadronic excitations to strings and quark (or
459: diquark) degrees of freedom is involved. In fact, in Ref.
460: \cite{Sahu99} it has been claimed that the excitation function of
461: transverse and elliptic proton flow together with the transverse
462: $p_t$ spectra of protons suggest a transition from {\it hadron} to
463: {\it string} matter at about 5 A$\cdot$GeV. Here, however, we
464: concentrate on meson abundancies and spectra and refer the reader
465: to Ref. \cite{Sahu99} for the collective dynamical aspects and to
466: Ref. \cite{brat99c} for the thermal properties of the theory that
467: involves a limiting ('Hagedorn') temperature of $T_S \approx$ 150
468: MeV due to the continuum string excitations.
469:
470:
471: \section{Meson production}
472:
473: To present a general overview on the meson abundancies in
474: nucleus-nucleus collisions we show in Fig. \ref{Fig5} the meson
475: multiplicities for central collisions of Au~+~Au from SIS to RHIC
476: energies. All multiplicities for $\pi^+, \eta, K^+, K^-, \phi$ as
477: well as $J/\Psi$ mesons show a monotonic increase with bombarding
478: energy which is only very steep at 'subthreshold' energies, i.e.
479: at bombarding energies per nucleon below the threshold in free
480: space for $NN$ collisions. At higher bombarding energies the meson
481: abundancies group according to their quark content, i.e. the
482: multiplicities are reduced (relative to $\pi^+$) by about a factor
483: of 4--5 for a strange quark, a factor of $\approx$ 50 for
484: $s\bar{s} \equiv \phi$ and a factor of $\approx 2\cdot 10^{5}$ for
485: $c\bar{c} \equiv J/\Psi$. We mention that in these calculations
486: the meson rescattering and reabsorption processes have been taken
487: into account; this reduces the $J/\Psi$ cross section at RHIC
488: energies by about a factor of 10 (cf. Section 5). At
489: 'subthreshold' energies the $\phi$ multiplicity turns out to be
490: almost the same as the antikaon multiplicity, but then rises less
491: steeply with bombarding energy. Apart from the $\phi$ excitation
492: function -- that still has to be controlled experimentally -- we
493: thus find no pronounced change in the shape of the meson
494: abundancies up to RHIC energies where data are expected to come up
495: soon.
496:
497: We recall that our investigations on strangeness production up to
498: 2~A$\cdot$GeV \cite{Cass97a,Brat97b} have given some evidence for
499: attractive antikaon potentials in the medium while for kaons only
500: a very moderate repulsive potential was suggested \cite{Brat97b};
501: $\eta$ mesons apparently do not show sensible in-medium effects
502: according to the studies in Ref. \cite{Brat98a,Brat98b} in
503: comparison to the available experimental spectra. At AGS and SPS
504: energies, on the other hand, the potential effects on kaon and
505: antikaon abundancies have been found to be only very low
506: \cite{Cass99} such that meson potentials have been discarded in
507: Fig. \ref{Fig5} for the overview.
508:
509: Since the meson abundancies show no sudden change in the
510: excitation function, we turn to particle ratios. Here strangeness
511: enhancement has been suggested for more than 2 decades to possibly
512: qualify as a sensible probe for a QGP phase (cf. Ref. \cite{SQM98}
513: for a recent overview). Here we examine the $K^+/\pi^+$ ratio at
514: midrapidity in central Au~+~Au (or Pb~+~Pb) collisions where
515: experimental data are now available from SIS to SPS energies
516: \cite{SQM98,E866,Alard}. We recall that detailed comparisons of
517: pion and kaon rapidity distributions and transverse momentum
518: spectra to the available data have been presented in Refs.
519: \cite{Cass99,Jgeiss} such that we directly can continue with the
520: corresponding excitation function.
521:
522: In order to discuss the strangeness production over the complete
523: energy range from SIS to RHIC energies we show in Fig. \ref{Fig6}
524: the calculated $K^+/\pi^+$ ratio at midrapidity (-0.5 $\leq y_{cm}
525: \leq$ 0.5) for central Au~+~Au collisions in comparison to the
526: experimental data. This ratio experimentally is substantially
527: lower at SPS energies ($\approx 13.5\%$) compared to AGS energies
528: ($\approx 19\%$). At SPS energies this ratio is only enhanced by a
529: factor 1.75 for central Pb~+~Pb collisions relative to p~+~p
530: reactions and has to be compared to the factor $\approx$ 3 at AGS.
531: Such a decrease of the scaled kaon yield from AGS to SPS energies
532: is not described by the HSD transport model (without kaon
533: self-energies) which shows a monotonic increase with bombarding
534: energy similar to $pp$ reactions (open circles). Furthermore, the
535: higher temperatures and particle densities at SPS energies tend to
536: enhance the $K^+/\pi^+$ yield closer to its thermal equilibrium
537: value of $\approx 20-25\%$ \cite{BM} at chemical freezout and
538: temperatures of $T\approx 150$ MeV.
539:
540: Our findings have to be compared to results obtained by other
541: microscopic approaches. Here only the RQMD model very recently
542: \cite{Wang99} provides a study partly comparable to that of Ref.
543: \cite{Jgeiss}, however, excluding p~+~A reactions to control the
544: amount of $K^+$ production by rescattering. Whereas kaon
545: production in p~+~p reactions is comparable to our results in
546: \cite{Jgeiss} (input of the transport model), the RQMD model
547: yields a higher $K^+/\pi^+$ ratio in Au~+~Au, Pb~+~Pb collisions
548: at all energies due to a higher kaon yield from rescattering. The
549: latter can be attributed to 'high mass strange resonances' that
550: have been incorporated to describe the low energy kaon production
551: via resonance production and decay \cite{Sorge96} (s-channels).
552: Since these 'high mass resonances' can be repopulated in
553: resonance-resonance scattering, a rather high $\sqrt{s}$ is
554: concentrated in a single degree of freedom for a short time. Such
555: 'hot spots' then lead to a higher $K^+/\pi^+$ ratio in A~+~A
556: reactions especially at AGS energies. While the $K^+/\pi^+$ ratio
557: can be reasonably described at AGS energies in Au~+~Au reactions,
558: this ratio is overestimated significantly at SPS energies (cf.
559: Fig. 4 of Ref. \cite{Wang99}). Thus also the RQMD model does not
560: describe the relative decrease of the $K^+/\pi^+$ ratio from 11
561: A$\cdot$GeV to 160 A$\cdot$GeV. The results of the RQMD
562: calculations at SIS energies are not known to the authors.
563:
564:
565:
566: \section{Dilepton production}
567:
568: Electromagnetic decays to virtual photons ($e^+ e^-$ or $\mu^+ \mu^-$
569: pairs) have been suggested long ago to serve as a possible signature
570: for a phase transition to the QGP
571: \cite{Shuryak,Kaj,Ruuskanen,Cleym,Uheinz} or to be an ideal probe for
572: vector meson spectroscopy in the nuclear medium. As pointed out in
573: Refs. \cite{Zahed,Kling} the isovector current-current correlation
574: function is proportional to the imaginary part of the $\rho$-meson
575: propagator and also to the dilepton invariant mass spectra. Dileptons
576: are particularly well suited for an investigation of the violent phases
577: of a high-energy heavy-ion collision because they can leave the
578: reaction volume essentially undistorted by final-state interactions.
579: Indeed, the dilepton studies in heavy-ion collisions by the DLS
580: Collaboration at the BEVALAC \cite{ro88} and by the
581: CERES~\cite{CERES,Ullrich}, HELIOS~\cite{HELIOS,HELI2},
582: NA38~\cite{NA38} and NA50 Collaborations~\cite{NA50} at SPS energies
583: have found a vivid interest in the nuclear physics community.
584:
585:
586: We recall that the question of chiral symmetry restoration does
587: not necessarily imply that vector meson masses have to drop with
588: baryon density or temperature \cite{KoKoch}. Actually, chiral
589: symmetry restoration (ChSR) only demands that the isovector
590: current-current correlation function and the axial vector
591: current-current correlation function (dominated by the chiral
592: partner of the $\rho$, the $a_1$-meson) should become identical at
593: high $\rho_B$ or temperature $T$, respectively, because there
594: should be no more differences between left- and right-handed
595: particles or equivalently vector and axial vector currents
596: \cite{Zahed,KoKoch}. Thus also a strong broadening of the $\rho$-
597: as well as the $a_1$-spectral function and their mixing in the
598: medium \cite{Rapp,friman,RappNPA,Rapp99} can be considered as a
599: signature for ChSR which, however, is not easy to detect
600: experimentally.
601:
602: In Ref. \cite{Cass99} we have demonstrated that the present
603: experimental data on the low mass dilepton enhancement at SPS
604: energies can be described equally well within the 'dropping
605: $\rho$-mass' scenario as well as within the 'melting' $\rho$
606: picture, which implies a large spreading in mass of the $\rho$
607: spectral function due to its couplings to baryons and/or mesons.
608: The situation at SIS/BEVALAC energies is more 'puzzling' since
609: here the low mass dilepton enhancement is neither described in the
610: dynamical spectral function approach \cite{Brat98b} nor in the
611: 'dropping mass' scheme \cite{Brat98c}, though the $pp$ dilepton
612: data from the DLS collaboration \cite{DLSpp} are reproduced within
613: the known sources rather well from 1 -- 5 GeV bombarding energy
614: \cite{Brat99a,Ernst}. In short, the dynamical origin of the low
615: mass dilepton enhancement is not yet understood.
616:
617: Here we propose to investigate the excitation function of low mass
618: dilepton enhancement in central Au~+~Au collisions. As discussed in
619: Refs. \cite{Cass99,Rapp99} this excess of dileptons is most probably
620: due to the isovector ($\rho$) channel, i.e. the imaginary part of the
621: isovector current-current correlation function which, however, mixes
622: with the axial vector current-current correlation function at finite
623: temperature and baryon density \cite{Zahed,Rapp99}.
624:
625: Since the $\rho$-meson spectral function is of primary interest,
626: it is important to have some information about the actual baryon
627: densities that the $\rho$-meson experiences during its propagation
628: and decay in central Au~+~Au collisions. This information is
629: displayed in Fig. \ref{Fig7} -- as resulting from the HSD
630: transport calculation -- for central reactions at 2, 5, 10, 20, 40
631: and 160 A$\cdot$GeV. Here the meson-baryon production channels and
632: baryon-baryon production channels are summed up by the solid lines
633: and denoted as ($\pi B \rightarrow \rho, BB \rightarrow \rho$).
634: The meson-meson production channels such as $\pi \pi \rightarrow
635: \rho, a_1 \rightarrow \pi \rho \ $ etc. are summed up in the
636: dashed histograms indicated as $\pi \pi \rightarrow \rho$
637: according to the dominating channel. We find that especially the
638: initial $BB$ production channels produce $\rho$-mesons at very
639: high densities close to 2 $\rho_0 \gamma_{cm}$ where $\gamma_{cm}$
640: is the Lorentz factor in the cms. However, these production
641: channels are much less abundant than the meson-meson channels
642: which extend over a larger time span (for the heavy system
643: Au~+~Au) and essentially occur at much lower baryon density,
644: respectively. This effect becomes even more pronounced with
645: increasing bombarding energy, i.e. 40 -- 160 A$\cdot$GeV, where
646: most of the $\rho$-mesons are produced at baryon densities below 2
647: $\rho_0$. Since in Fig. \ref{Fig7} the relative abundancy of
648: $\rho$-mesons is displayed as a function of the baryon density at
649: the production (formation) point, the time averaged value of the
650: density is even lower due to a fast expansion of the hadronic
651: fireball. Thus to probe on average high baryon densities by
652: $\rho$-mesons one should step down in energy to 2 -- 5 A$\cdot$GeV
653: in order to optimize effects due to the coupling to baryons.
654:
655: A general overview of dilepton mass spectra in central ($b$=2 fm)
656: Au~+~Au collisions from 2 A$\cdot$GeV to 21.5 A$\cdot$TeV is
657: presented in Fig. \ref{Fig8}, where the upper part corresponds to
658: the case of vacuum spectral functions for all mesons (cf. Fig.
659: 8.20 of Ref. \cite{Cass99}), while the lower part is calculated by
660: employing the $\rho$ spectral function from Rapp et al.
661: \cite{RappNPA,CasRap}. Whereas for the free meson spectral
662: function one observes essentially an increase of the dilepton
663: production channels with bombarding energy without any substantial
664: change in the spectral shape (except for an increasing peak from
665: the $\phi$; cf. Fig. 5), the in-medium calculations yield almost
666: exponential mass spectra above $M \geq$ 0.4 GeV with small peaks from
667: vacuum $\omega$ and $\phi$ decays at $M \approx 0.78$ GeV and 1.02~GeV.
668:
669: We mention that for the vacuum spectral functions (upper part of
670: Fig. 8) the shape of the dilepton spectra (for $M_{e^+e^-} \geq$
671: 0.15 GeV) is due to the superposition of $\eta, \eta', \omega$ and
672: $a_1$ Dalitz decays and direct vector meson decays ($\rho, \omega,
673: \phi$), where all mesons may also be produced in meson-baryon and
674: meson-meson collisions. The increase in dilepton yield (for
675: $M_{e^+e^-} \geq $ 0.15 GeV) with bombarding energy thus is due to
676: an enhanced production of $\eta, \eta', \rho, \omega, \phi, a_1$
677: mesons. Their relative abundance from SIS to RHIC energies does
678: not scale directly with the charged particle multiplicity which is
679: dominated by protons, $\pi^{\pm}$ and $K^{\pm}$ (cf. Fig. 5).
680: However, above about 50 -- 100 A$\cdot$GeV the meson ratios do not
681: change very much (cf. Figs. 5,6) while the charged particle
682: multiplicity becomes dominated by $\pi^{\pm}$ and $K^{\pm}$. Thus from
683: SPS to RHIC energies the low mass dilepton yield should
684: approximately be proportional to the charged particle
685: multiplicity, too.
686:
687: The relative change in the dilepton spectra is quantitatively displayed
688: in Fig. \ref{Fig9} -- again for central ($b$= 2 fm) collisions of
689: Au~+~Au -- for the case of free meson spectral functions (solid lines) and
690: the spectral function from Rapp et al. \cite{RappNPA,CasRap} (dashed
691: lines). In line with the discussion above the most prominent spectral
692: changes are expected at rather low bombarding energies from
693: 2--10~A$\cdot$GeV, where the enhancement in the invariant mass range from
694: 0.3--0.6~GeV is about a factor of 4. Note that the $\omega$ and
695: $\phi$ vacuum decays show clear peaks on top of the 'exponential'
696: continuum, which will have to be identified experimentally. We mention
697: that in all our calculations we have incorporated an experimental
698: ('optimistic') mass resolution of $\Delta M$ = 10 MeV for the dilepton
699: invariant mass.
700:
701: \section{Charmonium production and suppression}
702:
703: Matsui and Satz \cite{matsui} have proposed that a suppression of the
704: $J/\Psi$ yield in ultra-relativistic heavy-ion collisions is a
705: plausible signature for the formation of the quark-gluon plasma because
706: the $J/\Psi$ should dissolve in the QGP due to color screening. This
707: suggestion has stimulated a number of heavy-ion experiments at CERN SPS
708: to measure the $J/\Psi$ via its dimuon decay. Indeed, these experiments
709: have shown a significant reduction of the $J/\Psi$ yield when going from
710: proton-nucleus to nucleus-nucleus collisions \cite{NA38}.
711: Especially for Pb~+~Pb at 160~A$\cdot$GeV an even more dramatic
712: reduction of $J/\Psi$ has been reported by the NA50 Collaboration
713: \cite{NA50,gonin,NA5099}.
714:
715: To interpret the experimental results, various models based on
716: $J/\Psi$ absorption by hadrons have been also proposed (cf. Refs.
717: \cite{Cass99,Vogt99,Seattle98} for recent reviews) that do not
718: involve the assumption of a QGP phase transition. The role of
719: comover dissociation is presently again heavily debated
720: \cite{Bernd,Seattle98} especially since theoretical calculations
721: for $J/\Psi$-meson dissociation cross sections differ by up to a
722: factor of 50 \cite{Bernd,Haglin,Bernd96,Blaschke,DimaSatz}. The
723: problem is even more complicated since the $J/\Psi$ meson is not
724: created instantly as a hadronic state and there is also a
725: substantial ($\approx$ 35 \%) feeding from the $\chi_c -\gamma$
726: decays. Moreover, it is expected that the $c\bar{c}$ pair is first
727: produced in a color-octet state together with a gluon
728: ('pre-resonance state') and that this more extended configuration
729: has a larger interaction cross section with baryons and mesons
730: before the $J/\Psi$ singlet state, $\Psi'$ or $\chi_c$ finally
731: emerges after some formation time $\tau_c$. Additionally, the
732: dissociation on mesons of formed $J/\Psi$'s will differ from
733: $\chi_c$ due to their different thresholds with respect to the
734: $D\bar{D}$ channel as well as for the $\Psi^{\prime}$. Since
735: experimental information on the various charmonium-meson cross
736: sections -- especially at low relative momenta -- will be hard to
737: obtain, the excitation function of charmonium suppression in
738: central nucleus-nucleus collisions might be exploited to obtain
739: additional information on the absorption scenarios. One expects
740: that quite below the bombarding energy necessary for the formation
741: of a QGP phase the charmonium absorption should be entirely due to
742: dissociation with hadrons; any additional suppression due to color
743: screening then will show up in a more rapid suppression with the
744: incident energy or with the energy density achieved
745: \cite{SatzQM99}.
746:
747: In this Section we calculate the excitation function for $J/\Psi$
748: suppression within two absorption scenarios, i.e. the 'early' and
749: 'late' comover dissociation models, which have been explored in detail
750: by our group before at SPS energies \cite{CaKo97JPsi,Cass97d,Geiss2}.
751: The 'early' comover absorption scenario is based on the idea, that a $c
752: \bar{c}$ pair is created in the initial 'hard' phase of the
753: nucleus-nucleus collision, where the string density is very high, and
754: the $c\bar{c}$ pair is dissolved in the color electric field of
755: neighboring strings. Since in the HSD approach the information on the
756: string density as well as the string space-time extension is available,
757: the absorption model has only a single parameter, i.e. the average
758: transverse dimension of an extending string. In Ref. \cite{Geiss2} a
759: string radius of 0.2 fm was found to describe simultaneously the data
760: for p~+~A and S~+~U at 200 A$\cdot$GeV from NA38 \cite{NA38} and for
761: Pb~+~Pb at 160 A$\cdot$GeV from NA50 \cite{NA50} when adopting the
762: conventional charmonium-nucleon dissociation cross section of 6 mb. It
763: has been also speculated that the overlap of strings due to percolation
764: might describe the phase transition to a QGP \cite{Satz99}.
765:
766: In the 'late' comover scenario the additional charmonium suppression is
767: due to charmo\-nium-meson scattering to $D\bar{D}$ with an average
768: charmonium-meson cross section of $\approx$ 3 mb \cite{Cass97d}. Cross
769: sections of this order have been calculated by Haglin in Ref.
770: \cite{Haglin} within a meson-exchange model and thus might appear not
771: unrealistic. In our present calculation we refer to the model II of
772: Ref. \cite{Cass97d} including a 'pre-resonance' charmonium life time of
773: 0.3--0.5~fm/c which is supported (within the errorbars) by a more
774: recent analysis of charmonium suppression as a function of the Feynman
775: variable $x_F$ in p~+~A reactions \cite{pAdata} from He et al.
776: \cite{Huefner} and Kharzeev et al. \cite{Dima99}. We recall that within
777: the model II of Ref. \cite{Cass97d} the $J/\Psi$ suppression data for
778: p~+~A and S~+~U at 200 A$\cdot$GeV from NA38 \cite{NA38} and for Pb~+~Pb
779: at 160 A$\cdot$GeV from NA50 \cite{NA50} have been described very well
780: when adopting a 'pre-resonance'-nucleon cross section of 6 mb and a
781: $J/\Psi$-nucleon cross section of 3--4~mb in line with the data on
782: $J/\Psi$ photoproduction \cite{Photo}. Independent dynamical studies on
783: charmonium suppression within the UrQMD model \cite{Gerland,Spieles}
784: later on have lead to very similar conclusions.
785:
786: Since the comparison of our calculations at SPS energies has been
787: performed to data taken in 1995 and before we show in Fig.
788: \ref{Fig10} a comparison of the 'early' comover model (dashed
789: line) \cite{Geiss2} and the 'late' comover model II \cite{Cass97d}
790: with the more recent data from NA50 \cite{NA5099,NA50QM} for
791: Pb~+~Pb at 160 A$\cdot$GeV using the explicit numbers from Refs.
792: \cite{Cass97d,Geiss2}\footnote{This comparison is necessary since
793: the 1995 data have been rescaled in \cite{NA50QM} and our
794: calculations are reported inconsistently in the more recent
795: presentations of this topic \cite{NA50QM,BM99}.}. The 'early'
796: comover absorption model here gives a little too low suppression
797: at high $E_T$ whereas the 'late' comover absorption model is still
798: in line with the more recent data from 1996 and 1998 with minimum
799: bias (open triangles and open circles). The data from 1996 (full
800: squares), that show a (much debated) two-step behaviour, do not
801: agree with the explicit $E_T$ dependence from our calculations;
802: however, the 1996 minimum bias data (open triangles) well match
803: for $E_T \geq$ 40 GeV whereas the $J/\Psi$ suppression is slightly
804: overestimated at lower $E_T$. We do not comment on the highest
805: $E_T$ data points from 1998 with minimum bias since our earlier
806: analysis did not extend to these specific events.
807:
808: The question now arises, if the excitation function for $J/\Psi$
809: suppression in central Au~+~Au collisions might show some unusual
810: behaviour within the two scenarios discussed above or how they
811: might be disentangled. In order to achieve the same suppression
812: factor in central collisions within the 'early' comover model we
813: have increased the string absorption radius from $r_s$ = 0.2 fm to
814: 0.22 fm to get the same value for the $J/\Psi$ survival factor
815: $S_{J/\Psi}$ at 160 A$\cdot$GeV. We note that the total $J/\Psi$
816: multiplicity shown in Fig. \ref{Fig5} has been calculated within
817: the 'late' comover model. It drops by almost 3 orders of magnitude
818: when decreasing the bombarding energy from 160 A$\cdot$GeV to 20
819: A$\cdot$GeV. At the lowest energy considered here the
820: experimental $J/\Psi$ signal will be very hard to measure; it is
821: hopeless within the present experimental setups. Nevertheless, it
822: is worth exploring theoretically if some unusual excitation
823: function might be found. Our results are displayed in Fig.
824: \ref{Fig11} (l.h.s.) for the 'early' absorption model (open
825: circles) and for the 'late' comover model (full circles); both
826: models practically do not differ in their excitation functions and
827: show a very smooth decrease of $S_{J/\Psi}$ from 0.4 to 0.3 with
828: increasing bombarding energy. The net absorption by baryons is
829: dominant in both scenarios, however, differs in magnitude due to
830: the simple fact that in the 'early' (string) absorption model
831: there is less suppression by baryons since the absorption by
832: strings competes at early times. In the 'late' comover model
833: there is more absorption by baryons because the mesons are formed
834: at later stages and not competitive in the early phase; their
835: relative contribution is lower as for strings accordingly.
836: However, these individual contributions cannot be distinguished
837: experimentally and thus are 'irrelevant'.
838:
839: The r.h.s. of Fig. 11 shows the survival probability $S_{J/\Psi}$
840: in the 'late' comover model for central Au + Au collisions from
841: 0.160 A$\cdot$TeV to 21.5 A$\cdot$TeV, respectively, where we have gated on
842: $J/\Psi$'s in the rapidity interval -1 $\leq y_{cm} \leq$ 1. The
843: solid line stands for the total $J/\Psi$ survival probability
844: while the dashed line displays the relative absorption on baryons
845: and the dotted line the relative dissociation on mesons. Whereas
846: the dissociation on baryons is practically constant with
847: bombarding energy, the absorption on mesons increases in line with
848: the higher meson densities achieved with increasing $E/A$. We note
849: that the 'early' comover model leads to a similar total absorption
850: within the numerical accuracy.
851:
852: We have to mention that neither the 'early' nor the 'late' comover
853: model might be realized in nature exclusively and both absorption
854: processes should occur within the same reaction with probably
855: different weights. Since we do not find a substantial difference
856: for both scenarios also a linear combination of both absorption
857: models, i.e. decreasing $r_s$ as well as the charmonium-meson
858: cross section accordingly, will lead to a similar excitation
859: function. This also holds for the relative suppression on the
860: transverse energy $E_T$ (cf. Fig. \ref{Fig10}).
861:
862: Inspite of the rather disappointing perspectives to disentangle
863: the 'late' and 'early' comover models experimentally at the full
864: range of SPS energies, the excitation function of $J/\Psi$ still
865: might show some discontinuity in $E/A$ experimentally, which could
866: rule out the two models studied here and indicate a transition to
867: a QGP phase. This subject is taken up in the next Section again
868: with respect to the dependence of $S_{J/\Psi}$ on the transverse
869: energy $E_T$ produced in Au + Au collisions at RHIC energies.
870:
871:
872:
873: \section{Predictions for RHIC energies}
874:
875: As noted in the introduction one expects that the initial
876: nonequilibrium phase of a nucleus-nucleus collision at RHIC
877: energies should be described by parton degrees of freedom, whereas
878: hadrons are only formed (by 'condensation') at a later stage of
879: the reaction and interact until freeze out. Thus parton cascade
880: calculations should be adequate for all initial reactions
881: involving a large 4-momentum transfer between the constituents,
882: while hadron cascades should be appropriate in the final hadronic
883: expansion phase. We suggest that the dynamics in between the
884: partonic and hadronic phase might be described by quarks
885: (diquarks) and strings as e.g. implemented in the HSD approach.
886:
887: The practical question is, however, if nonequilibrium partonic and
888: hadron/string models can be distinguished at all, i.e. do they
889: lead to different predictions for experimental observables? In
890: fact, first applications of the parton cascade model developed by
891: Geiger \cite{Geiger1,VNI} to nucleus-nucleus collisions at SPS
892: energies \cite{Geiger} have suggested that a reasonable
893: description of the meson and baryon rapidity distributions can
894: also be achieved on the basis of partonic degrees of freedom.
895: However, in the latter calculations the extrapolation of the
896: strong coupling constant to low $Q^2$ has been overestimated as
897: discovered recently by Bass and M\"uller \cite{Bass99c}. This
898: finding invalidates the detailed predictions and comparisons
899: within the hybrid models VNI+UrQMD or VNI+HSD presented in Ref.
900: \cite{QMRHIC} that have been tailored to describe the dynamics at
901: RHIC or even LHC energies.
902:
903: We start with $pp$ collisions at $\sqrt{s}$ = 200 GeV. The
904: calculated results for the proton, $\pi^+$ and $K^+$ rapidity
905: distributions in the cms are shown in Fig. \ref{Fig12} (upper
906: part) for both models, which are denoted individually by the
907: labels VNI and HSD in obvious notation. On the level of $pp$
908: collisions we find only minor differences between the two kinetic
909: models. The parton cascade shows a slightly higher amount of
910: proton stopping as the HSD model (l.h.s.) and as a consequence a
911: slightly higher production of $\pi^+$ and $K^+$ mesons (r.h.s.),
912: because the energy taken from the relative motion of the leading
913: baryons is converted to the production of mesons. It is presently
914: unclear which of the two approaches will be closer to experiment;
915: a proper description of $pp$ data will be a necessary step before
916: performing reliable extrapolations to nucleus-nucleus collisions.
917:
918: Inspite of this missing experimental information we directly step
919: towards central collisions ($b \leq$ 1.5 fm) for Au~+~Au at
920: $\sqrt{s}$ = 200~GeV. The calculated results for the net proton
921: (here $p-\bar{p}$), antiproton, $\pi^+$ and $K^+$ rapidity
922: distributions in the cms are shown in Fig. \ref{Fig12} (lower
923: part). In the HSD scenario essentially 'comover' scattering occurs
924: with a low change of the meson rapidity distribution. Thus the
925: meson rapidity distributions are roughly the same as for $pp$
926: collisions. Also note that at midrapidity the net baryon density
927: $\sim N_p - N_{\bar{p}}$ is practically zero, however, even at
928: midrapidity at lot of baryons appear that are produced together
929: with antibaryons. Thus also mesons (especially $c \bar{c}$ pairs)
930: encounter a lot of baryons and antibaryons on their way to the
931: continuum. Whereas the HSD approach predicts a vanishing net
932: baryon density at midrapidity, other recent models -- that combine
933: high and low energy transport concepts -- suggest a sizeable net
934: proton density for $y_{cm} \approx$ 0 \cite{Ko99}.
935:
936: The amount of higher order hadronic rescattering processes at RHIC
937: energies is depicted in Fig. \ref{Fig13} (lower right part) as
938: emerging from the HSD calculation, where the number of
939: baryon-baryon ($BB$) and meson-baryon collisions ($mB$) is shown
940: as a function of the invariant energy $\sqrt{s}$. We mention that
941: quark-baryon and diquark-baryon collisions are counted here as
942: $mB$ or $BB$ collisions, respectively. Apart from the initial
943: small peak at $\sqrt{s}$ = 200 GeV a substantial amount of
944: intermediate and low energy rescattering processes with maxima at
945: 2.5 GeV and 1.8 GeV are found, which essentially stand for flavor
946: exchange processes, multiple pion production in $mB$ and $BB$
947: collisions as well as secondary strangeness production channels.
948: For comparison the corresponding $\sqrt{s}$ distributions are also
949: displayed for bombarding energies of 2, 11 and 160 A$\cdot$GeV,
950: respectively, showing a dominance of low energy $BB$ and $mB$
951: collisions. The latter reactions occur at energy scales where
952: perturbative QCD is no longer applicable. This has to be kept in
953: mind additionally when comparing to $pp$ and $pA$ reactions at
954: $\sqrt{s}$ = 200 GeV.
955:
956: We return to the question of charmonium suppression at RHIC
957: energies since it is expected that one might probe increasing
958: energy densities also with increasing centrality of the collision,
959: where the latter can be correlated with the transverse energy
960: $E_T$ produced in a collision event. As argued e.g. by Satz
961: \cite{SatzQM99} the survival factor $S_{J/\Psi}$ then should show
962: steps as a function of $E_T$ due to the melting of first the
963: $\chi_c$ and then the $J/\Psi$ in a QGP phase. As seen from Fig.
964: 10 there are no pronounced steps in the $E_T$ dependence of
965: $J/\Psi$ suppression in the data for Pb + Pb at SPS energies
966: according to the authors point of view; this situation might
967: change at RHIC energies.
968:
969: Using the 'late' comover model described in Section 5 we have
970: calculated the $J/\Psi$ survival factor $S_{J/\Psi}$ as a function
971: of the transverse energy $E_T$ in the cms rapidity window [-1,1]
972: for Au + Au at $\sqrt{s}$ = 200 GeV on an event by event basis
973: covering all impact parameters from $b$ = 0 to 13 fm. The resulting
974: correlation of $S_{J/\Psi}$ with $E_T$ is shown in Fig. 14 and
975: indicates a smooth decrease with centrality (or increasing $E_T$)
976: reaching an average survival probability of $\approx$ 0.1 for the
977: most central events (cf. Fig. 11, r.h.s.). This result can be
978: understood as follows: According to our calculations the net
979: $J/\Psi$ dissociation by mesons in central Au + Au collisions at
980: the SPS is $\approx 16\%$ (cf. Fig. 11) while the rapidity
981: distribution of negatively charged particles ($h^-$) at
982: midrapidity here is about 180. At the RHIC energy we get a
983: corresponding $h^-$ rapidity density at midrapidity of $\approx$
984: 450 (cf. Fig. 12) which is higher by a factor of 2.5. Simply
985: multiplying the $J/\Psi$ meson absorption at the SPS of $16 \%$ by
986: the factor 2.5 we obtain about $40 \%$ for central collisions at
987: RHIC energies, which together with $\approx 52 \%$ of absorption
988: on baryons gives a survival probability of $8 \%$. The actual
989: numerical results in Fig. 14 indicate that this simple estimate
990: works quite well. On the other hand, if the $h^-$ rapidity density
991: is found to be lower (higher) experimentally, we expect
992: corresponding changes in the $J/\Psi$ suppression for central
993: events if the 'late' comover absorption model holds true. This
994: dependence might well be tested experimentally in the near future
995: to possibly falsify the comover dissociation model.
996:
997:
998: \section{Summary}
999:
1000: In this work we have performed a systematic analysis of hadron
1001: production in central Au~+~Au collisions from SIS to RHIC energies
1002: within the HSD transport approach. We have concentrated here on the
1003: 'classical' signatures, i.e. strangeness and low mass dilepton
1004: enhancement as well as charmonium suppression. For all observables our
1005: calculations give a monotonic increase (for the ratio $K^+/\pi^+$ and
1006: charmonium suppression) or decrease (for the low mass dilepton
1007: enhancement) with bombarding energy, respectively. So far, experimental
1008: data are available only in a limited range of bombarding energies or at
1009: a single energy, respectively. We have pointed out that the relative
1010: maximum indicated by the experimental data in the $K^+/\pi^+$ ratio at
1011: about 10 A$\cdot$GeV (or higher?) is not reproduced within the transport
1012: approach that is based on quark, diquark, string and hadronic degrees
1013: of freedom. We speculate that at AGS energies this failure might be
1014: attributed to a restoration of chiral symmetry in a sufficiently large
1015: space-time volume (cf. Figs. \ref{Fig3} and \ref{Fig4}).
1016:
1017: The enhancement of low mass dileptons in the range 0.3 GeV $\leq
1018: M_{e^+e^-} \leq$ 0.6 GeV is most pronounced at lower bombarding
1019: energies of 2--5 A$\cdot$GeV within our calculations since here the
1020: space-time volume for densities above 2 $\rho_0$ is very large such
1021: that a majority of $\rho$-mesons probes the high density phase of the
1022: reaction (cf. Fig. \ref{Fig5}). With increasing bombarding energy the
1023: average density -- which a $\rho$-meson experiences -- drops
1024: substantially such that high energy nucleus-nucleus collisions are not
1025: well suited for in-medium $\rho$ spectroscopy.
1026:
1027: The suppression of charmonium (here $J/\Psi$) increases smoothly
1028: with bombarding energy and centrality of the reaction within the
1029: 'early' and 'late' comover absorption scenarios. Unfortunately,
1030: both scenarios cannot be distinguished by means of the excitation
1031: function since they give approximately the same survival
1032: probability $S_{J/\Psi}$ with bombarding energy (cf. Fig. 11). In
1033: the transport approach the smooth increase of charmonium
1034: absorption with bombarding energy is easy to understand: a major
1035: fraction of $J/\Psi$'s is anyhow dissociated by baryons which
1036: basically are of the same number at all energies considered here;
1037: only the relative collisional energy changes. The additional
1038: absorption by 'early' strings or 'late' hadrons increases smoothly
1039: with bombarding energy since the string and hadron density
1040: increases accordingly. At RHIC energies this additional
1041: suppression mechanism leads to a $J/\Psi$ suppression of about
1042: 90\% in central Au~+~Au collisions even without employing an
1043: explicit formation of a QGP. We note, however, that the charmonium
1044: suppression shows a smooth dependence on the transverse energy
1045: $E_T$ (cf. Fig. 14); any gradual steps of $S_{J/\Psi}$ with $E_T$
1046: due to a melting of the $\chi_c$ or the $J/\Psi$ at higher energy
1047: density would indicate a new suppression mechanism which might be
1048: attributed to color screening in a QGP phase \cite{SatzQM99}.
1049:
1050:
1051: \vspace{1cm} The authors acknowledge inspiring discussions with J.
1052: Aichelin, S. A. Bass, G. E. Brown, C. Greiner, M. Gyulassy, C. M.
1053: Ko, U. Mosel, R. Rapp, H. Satz, H. Sorge, H. St\"ocker, J.
1054: Wambach, X.-N. Wang and K. Werner.
1055:
1056:
1057: \begin{thebibliography}{999}
1058: \bibitem{Harris}
1059: J. W. Harris and B. M\"uller, Annu. Rev. Nucl. Part. Sci. 46 (1996) 71.
1060: \bibitem{Shuryak93}
1061: E. V. Shuryak, Rev. Mod. Phys. 65 (1993) 1.
1062: \bibitem{QM99}
1063: Quark Matter'99, Nucl. Phys. A (1999) , in press.
1064: \bibitem{Geiger3}
1065: K. Geiger and B. M\"uller, Nucl. Phys. B 369 (1992) 600.
1066: \bibitem{Bernd}
1067: B. M\"uller, nucl-th/9906029, Nucl. Phys. A, in press.
1068: \bibitem{Sorge99}
1069: H. Sorge, nucl-th/9906051, Nucl. Phys. A, in press.
1070: \bibitem{Basrev}
1071: S. A. Bass et al., Prog. Part. Nucl. Phys. 41 (1998) 225;
1072: J. Phys. G 25 (1999) R1.
1073: \bibitem{Geiger1}
1074: K. Geiger, Phys. Rep. 258 (1995) 237.
1075: \bibitem{Wang}
1076: X.-N. Wang, Phys. Rep. 280 (1997) 287.
1077: \bibitem{Cass99}
1078: W. Cassing and E. L. Bratkovskaya, Phys. Rep. 308 (1999) 65.
1079: \bibitem{Ehehalt}
1080: W. Ehehalt and W. Cassing, Nucl. Phys. A 602 (1996) 449.
1081: \bibitem{Jgeiss}
1082: J. Geiss, W. Cassing, and C. Greiner, Nucl. Phys. A 644 (1998) 107.
1083: \bibitem{Brown}
1084: G. E. Brown and M. Rho, Phys. Rev. Lett. 66 (1991) 2720.
1085: \bibitem{Kochr}
1086: V. Koch, Int. J. Mod. Phys. E 6 (1997) 203.
1087: \bibitem{Zahed}
1088: J. V. Steele, H. Yamagishi, and I. Zahed,
1089: Phys. Lett. B 384 (1996) 255; Phys. Rev. D 56 (1997) 5605.
1090: \bibitem{Stoecker}
1091: H. St\"ocker and W. Greiner, Phys. Rep. 137 (1986) 277.
1092: \bibitem{Bertsch}
1093: G. F. Bertsch and S. Das Gupta, Phys. Rep. 160 (1988) 189.
1094: \bibitem{Cass90}
1095: W. Cassing, V. Metag, U. Mosel, and K. Niita,
1096: Phys. Rep. 188 (1990) 363.
1097: \bibitem{Koreview}
1098: C. M. Ko and G. Q. Li, J. Phys. G 22 (1996) 1673.
1099: \bibitem{Bot}
1100: W. Botermans and R. Malfliet, Phys. Rep. 198 (1990) 115.
1101: \bibitem{Mal}
1102: R. Malfliet, Prog. Part. Nucl. Phys. 21 (1988) 207.
1103: \bibitem{T3}
1104: A. Faessler, Prog. Part. Nucl. Phys. 30 (1993) 229.
1105: \bibitem{GEB}
1106: G. E. Brown, Prog. Theor. Phys. 91 (1987) 85.
1107: \bibitem{GEB1}
1108: G. E. Brown, C. M. Ko, Z. G. Wu, and L. H. Xia,
1109: Phys. Rev. C 43 (1991) 1881.
1110: \bibitem{Koch}
1111: V. Koch and G. E. Brown, Nucl. Phys. A 560 (1993) 345.
1112: \bibitem{Lat1}
1113: B. Peterson, Nucl. Phys. B 30 (1992) 66.
1114: \bibitem{Lat2}
1115: F. Karsch, Nucl. Phys. B 34 (1993) 63.
1116: \bibitem{Lat3}
1117: J. Engels, Phys. Lett. B 252 (1990) 625;
1118: J. Engels, F. Karsch and K. Redlich, Nucl. Phys. B 435 (1995) 295.
1119: \bibitem{Lat4}
1120: F. Karsch and E. Laermann, Phys. Rev. D 50 (1994) 6954.
1121: \bibitem{Cass}
1122: W. Cassing and U. Mosel, Prog. Part. Nucl. Phys. 25 (1990) 235.
1123: \bibitem{Weber1}
1124: K. Weber, B. Bl\"attel, W. Cassing, H.-C. D\"onges, V. Koch,
1125: A. Lang, and U. Mosel, Nucl. Phys. A 539 (1992) 713.
1126: \bibitem{Hama}
1127: S. Hama, B. C. Clark, E. D. Cooper, H. S. Sherif and R. L. Mercer,
1128: Phys. Rev. C 41 (1990) 2737.
1129: \bibitem{LUND}
1130: B. Nilsson-Almqvist and E. Stenlund, Comp. Phys. Comm. 43 (1987) 387;
1131: B. Anderson, G. Gustafson and Hong Pi, Z. Phys. C 57 (1993) 485.
1132: \bibitem{Schwinger}
1133: J. Schwinger, Phys. Rev. 83 (1951) 664
1134: \bibitem{PYTHIA}
1135: H.-U. Bengtsson and T. Sj\"ostrand,
1136: Computer Physics Commun. 46 (1987) 43.
1137: \bibitem{Drukarev}
1138: E. G. Drukarev and E. M. Levin, Nucl. Phys. A 511 (1990) 679.
1139: \bibitem{Toneev98}
1140: B. Friman, W. N\"orenberg and V. D. Toneev,
1141: Eur. Phys. J. A 3 (1998) 165.
1142: \bibitem{brat97}
1143: E. L. Bratkovskaya and W. Cassing, Nucl. Phys. A 619 (1997) 413.
1144: \bibitem{Sahu99}
1145: P. K. Sahu, W. Cassing, U. Mosel, and A. Ohnishi, nucl-th/9907002.
1146: \bibitem{brat99c}
1147: E. L. Bratkovskaya, W. Cassing, C. Greiner et al.,
1148: preprint UGI-99-32.
1149: \bibitem{Cass97a}
1150: W. Cassing, E. L. Bratkovskaya, U. Mosel, S. Teis, and A. Sibirtsev,
1151: Nucl. Phys. A 614 (1997) 415.
1152: \bibitem{Brat97b}
1153: E. L. Bratkovskaya, W. Cassing and U. Mosel, Nucl. Phys. A 622 (1997) 593.
1154: \bibitem{Brat98a}
1155: E. L. Bratkovskaya, W. Cassing and U. Mosel, Phys. Lett. B 424 (1998) 244.
1156: \bibitem{Brat98b}
1157: E. L. Bratkovskaya, W. Cassing, R. Rapp, and J.Wambach,
1158: Nucl. Phys. A 634 (1998) 168.
1159: \bibitem{SQM98}
1160: {\it Strangeness in Quark Matter 1998}, J. Phys. G 25 (1999) 143.
1161: \bibitem{E866}
1162: C. A. Ogilvie for the E866 and E819 Collaboration, Nucl. Phys. A 638
1163: (1998) 57c; J. Phys. G 25 (1999) 159.
1164: \bibitem{Alard}
1165: L. Ahle et al., Nucl. Phys. A 610 (1996) 139c;
1166: Phys. Rev. C 57 (1998) R466; Phys. Rev. C 60 (1999) 044904.
1167: \bibitem{BM}
1168: P. Braun-Munzinger, J. Stachel, J. P. Wessels, and N. Xu,
1169: Phys. Lett. B 344 (1995) 43;
1170: Phys. Lett. B 365 (1996) 1.
1171: \bibitem{Wang99}
1172: F. Wang, H. Liu, H. Sorge, N. Xu, and J. Yang, nucl-th/9909001.
1173: \bibitem{Sorge96}
1174: H. Sorge, Z. Phys. C 67 (1995) 479; Phys. Rev. C 52 (1995) 3291;
1175: Nucl. Phys. A 630 (1998) 522c.
1176: \bibitem{Shuryak}
1177: E. Shuryak, Phys. Lett. B 78 (1978) 150;
1178: Sov. J. Nucl. Phys. 28 (1978) 408.
1179: \bibitem{Kaj}
1180: K. Kajantie, J. Kapusta, L. McLerran, and A. Mekjian,
1181: Phys. Rev. D 34 (1986) 2746.
1182: \bibitem{Ruuskanen}
1183: P. V. Ruuskanen, Nucl. Phys. A 544 (1992) 169c.
1184: \bibitem{Cleym}
1185: J. Cleymans, K. Redlich and H. Satz, Z. Phys. C 52 (1991) 517.
1186: \bibitem{Uheinz}
1187: U. Heinz and K. S. Lee, Phys. Lett. B 259 (1991) 162.
1188: \bibitem{Kling}
1189: F. Klingl, N. Kaiser and W. Weise, Nucl. Phys. A 624 (1997) 527.
1190: \bibitem{ro88}
1191: G. Roche et al., Phys. Rev. Lett. 61 (1988) 1069;
1192: C. Naudet et al., Phys. Rev. Lett. 62 (1989) 2652;
1193: G. Roche et al., Phys. Lett. B 226 (1989) 228.
1194: \bibitem{CERES}
1195: G. Agakichiev et al., Phys. Rev. Lett. 75 (1995) 1272.
1196: \bibitem{Ullrich}
1197: G. Agakichiev et al., Phys. Lett. B 422 (1998) 405.
1198: \bibitem{HELIOS}
1199: M. A. Mazzoni, Nucl. Phys. A 566 (1994) 95c;
1200: M. Masera, Nucl. Phys. A 590 (1995) 93c.
1201: \bibitem{HELI2}
1202: T. {\AA}kesson et al., Z. Phys. C 68 (1995) 47.
1203: \bibitem{NA38}
1204: C. Baglin et al., Phys. Lett. B 220 (1989) 471; B 251 (1990) 465;
1205: B 270 (1991) 105; B 345 (1995) 617;
1206: S. Ramos, Nucl. Phys. A 590 (1995) 117c.
1207: \bibitem{NA50}
1208: M. Gonin et al., Nucl. Phys. A 610 (1996) 404c.
1209: \bibitem{KoKoch}
1210: C. M. Ko, V. Koch and G. Q. Li, Ann. Rev. Nucl. Part. Sci. 47 (1997) 505.
1211: \bibitem{Rapp}
1212: R. Rapp, G. Chanfray and J. Wambach, Phys. Rev. Lett. 76 (1996) 368.
1213: \bibitem{friman}
1214: B. Friman and H. J. Pirner, Nucl. Phys. A 617 (1997) 496.
1215: \bibitem{RappNPA}
1216: R. Rapp, G. Chanfray and J. Wambach, Nucl. Phys. A 617 (1997) 472.
1217: \bibitem{Rapp99}
1218: R. Rapp and J. Wambach, nucl-th/9909229, to appear in Adv. Nucl. Phys.
1219: \bibitem{Brat98c}
1220: E. L. Bratkovskaya and C. M. Ko, Phys. Lett. B 445 (1999) 265.
1221: \bibitem{DLSpp}
1222: W. K. Wilson et al., Phys. Rev. C 57 (1998) 1865.
1223: \bibitem{Brat99a}
1224: E. L. Bratkovskaya, W. Cassing, M. Effenberger, and U. Mosel,
1225: Nucl. Phys. A 653 (1999) 301.
1226: \bibitem{Ernst}
1227: C. Ernst, S. A. Bass, M. Belkacem et al., Phys. Rev. C 58 (1998) 447.
1228: \bibitem{CasRap}
1229: W. Cassing, E. L. Bratkovskaya, R. Rapp, and J. Wambach,
1230: Phys. Rev. C 57 (1998) 916.
1231: \bibitem{matsui}
1232: T. Matsui and H. Satz, Phys. Lett. B 178 (1986) 416.
1233: \bibitem{gonin}
1234: M. Gonin et al., Nucl. Phys. A 610 (1996) 404c.
1235: \bibitem{NA5099}
1236: M. C. Abreu et al., Phys. Lett. B 450 (1999) 456.
1237: \bibitem{Vogt99}
1238: R. Vogt, Phys. Rep. 310 (1999) 197.
1239: \bibitem{Seattle98}
1240: X.-N. Wang and B. Jacak, Eds.,{\it Quarkonium production
1241: in high-energy nuclear collisions}, World-Scientific, 1998.
1242: \bibitem{Haglin}
1243: K. Haglin, nucl-th/9907034.
1244: \bibitem{Bernd96}
1245: S. G. Matinyan and B. M\"uller, Phys. Rev. C 58 (1998) 2994.
1246: \bibitem{Blaschke}
1247: K. Martins, D. Blaschke and E. Quack, Phys. Rev. C 51 (1995) 2723.
1248: \bibitem{DimaSatz}
1249: D. Kharzeev and H. Satz, Phys. Lett. B 334 (1994) 155.
1250: \bibitem{SatzQM99} H. Satz, Quark Matter'99, hep-ph/9908339, Nucl.
1251: Phys. A, in press.
1252: \bibitem{CaKo97JPsi}
1253: W. Cassing and C. M. Ko, Phys. Lett. B 396 (1997) 39.
1254: \bibitem{Cass97d}
1255: W. Cassing and E. L. Bratkovskaya, Nucl. Phys. A 623 (1997) 570.
1256: \bibitem{Geiss2}
1257: J. Geiss, C. Greiner, E.L. Bratkovskaya, W. Cassing, and U. Mosel,
1258: Phys. Lett. B 447 (1999) 31.
1259: \bibitem{Satz99}
1260: H. Satz, Nucl. Phys. A 642 (1998) 130; M. Nardi and H. Satz, Phys. Lett. B 442
1261: (1998) 14.
1262: \bibitem{pAdata}
1263: M. J. Leitch, E866 Collaboration, Quark Matter'99, Nucl. Phys.
1264: A, in press.
1265: \bibitem{Huefner}
1266: Y. B. He, J. H\"ufner and B. Z. Kopeliovich, hep-ph/9908243.
1267: \bibitem{Dima99}
1268: D. Kharzeev and R. L. Thews, Phys. Rev. C 60 (1999) 041901.
1269: \bibitem{Photo}
1270: U. Camerini et al., Phys. Rev. Lett. 35 (1975) 483; B.
1271: Gittelman et al., Phys. Rev. Lett. 35 (1975) 1616;
1272: S. Aid et al., Nucl. Phys. B 472 (1996) 3.
1273: \bibitem{Gerland}
1274: L. Gerland et al., Phys. Rev. Lett. 81 (1998) 762.
1275: \bibitem{Spieles}
1276: C. Spieles et al., Eur. Phys. J. C 5 (1998) 349;
1277: Phys. Rev. C 60 (1999) 054901.
1278: \bibitem{NA50QM}
1279: C. Ciolo for the NA50 Collaboration, Quark Matter'99,
1280: Nucl. Phys. A, in press.
1281: \bibitem{BM99}
1282: P. Braun-Munzinger, nucl-ex/9909014, Nucl. Phys. A, in press.
1283: \bibitem{VNI}
1284: K. Geiger, Comp. Phys. Comm. 104 (1997) 70.
1285: \bibitem{Geiger}
1286: K. Geiger and D. K. Srivastava, Phys. Rev. C 56 (1997) 2718.
1287: \bibitem{Bass99c}
1288: S. A. Bass and B. M\"uller, Preprint DUKE-TH-99-192.
1289: \bibitem{QMRHIC}
1290: S. A. Bass, M. Bleicher, W. Cassing et al.,
1291: nucl-th/9907090, Nucl. Phys. A, in press.
1292: \bibitem{Ko99}
1293: B. Zhang, C. M. Ko, B. A. Li, and Z. Lin, nucl-th/9907017.
1294: \end{thebibliography}
1295:
1296: %----------------------------------------------------------
1297: \newpage
1298:
1299: \begin{figure}[t]
1300: \phantom{a}\vspace*{-2.5cm}
1301: \centerline{\psfig{figure=fig1exf.eps,width=16cm}}
1302: \vspace*{-5cm}
1303: \caption{The nucleon scalar ($U_S$) and negative vector potential ($-U_0$)
1304: as a function of the nuclear density $\rho$ and relative momentum
1305: $p$ of the nucleon with respect to the nuclear matter rest frame
1306: as implemented in the HSD transport approach
1307: \protect\cite{Ehehalt}.}
1308: \label{Fig1}
1309: \end{figure}
1310:
1311: \begin{figure}[t]
1312: \phantom{a}\vspace*{-3cm}
1313: \centerline{\psfig{figure=fig2exf.eps,width=16cm}} \vspace*{-5cm}
1314: \caption{The potential $U_{SEP}$ (5) -- as resulting from the
1315: nucleon scalar ($U_S$) and vector potential ($U_0$) in Fig. 1 --
1316: as a function of the nuclear density $\rho$ and relative momentum
1317: $p$ of the nucleon with respect to the nuclear matter rest frame.}
1318: \label{Fig2}
1319: \end{figure}
1320:
1321:
1322: \begin{figure}[t]
1323: \phantom{a}\vspace*{-3cm}
1324: \centerline{\psfig{figure=fig3exf.eps,width=16cm}} \vspace*{-2cm}
1325: \caption{The scalar quark condensate $<q\bar{q}(x,0,z;t)>$ for
1326: central Au~+~Au collisions at 6 A$\cdot$GeV divided by the vacuum
1327: condensate $<q\bar{q}>_V$ such that the nonperturbative vacuum is
1328: characterized by a value of 1. The $z$-direction has been
1329: stretched by the Lorentz-factor $\gamma_{cm}$ to compensate for
1330: Lorentz contraction, while negative numerical values for the
1331: condensate have been suppressed.} \label{Fig3}
1332: \end{figure}
1333:
1334:
1335: \begin{figure}[t]
1336: \phantom{a}\vspace*{-2.5cm}
1337: \centerline{\psfig{figure=fig4exf.eps,width=16cm}}
1338: \vspace*{-2cm}
1339: \caption{Same as Fig. 3 for central collisions of Au~+~Au at 20 A$\cdot$GeV.}
1340: \label{Fig4}
1341: \end{figure}
1342:
1343:
1344:
1345: \begin{figure}[t]
1346: \phantom{a}\vspace*{-2.5cm}
1347: \centerline{\psfig{figure=fig5exf.eps,width=16cm}}
1348: \vspace*{-5cm}
1349: \caption{The meson ($\pi^+, \eta, K^+, K^-, \phi$ and $J/\Psi$)
1350: multiplicities from the HSD approach for central collisions of $Au
1351: + Au$ from 200 A MeV to 21.5 A$\cdot$TeV.}
1352: \label{Fig5}
1353: \end{figure}
1354:
1355:
1356: \begin{figure}[t]
1357: \phantom{a}\vspace*{-2.5cm}
1358: \centerline{\psfig{figure=fig6exf.eps,width=16cm}}
1359: \vspace*{-5cm}
1360: \caption{The $K^+/\pi^+$ ratio at midrapidity from central Au~+~Au ($Pb
1361: + Pb$) collisions from 1 A$\cdot$GeV to 21.5 A$\cdot$TeV. The open circles
1362: show the results from HSD for $pp$ collisions while the open
1363: squares are obtained for central Au~+~Au reactions. The
1364: experimental data from Refs. \protect\cite{QM99,E866,Alard} are
1365: displayed in terms of the full circles.}
1366: \label{Fig6}
1367: \end{figure}
1368:
1369:
1370: \begin{figure}[t]
1371: \phantom{a}\vspace*{-2.5cm}
1372: \centerline{\psfig{figure=fig7exf.eps,width=16cm}}
1373: \vspace*{-2cm}
1374: \caption{The differential $\rho$-meson distribution versus baryon density
1375: $\rho/\rho_0$ (at the $\rho$ creation point) for central
1376: collisions of Au~+~Au at 2, 5, 10, 20, 40, and 160 A$\cdot$GeV from
1377: the HSD approach. The production channels involving two baryons or
1378: a meson and a baryon (denoted by $BB \to \rho, \pi B \to \rho$)
1379: are summed up in the solid histograms whereas the dashed
1380: histograms stand for the sum of the meson production channels
1381: (denoted by $\pi\pi \to \rho$) which are dominated by the pion
1382: annihilation channel.}
1383: \label{Fig7}
1384: \end{figure}
1385:
1386:
1387: \begin{figure}[t]
1388: \phantom{a}\vspace*{-2.5cm}
1389: \centerline{\psfig{figure=fig8exf.eps,width=16cm}} \vspace*{-2cm}
1390: \caption{The differential dilepton multiplicity $d n_{e^+e^-}/dM$
1391: for central collisions of Au~+~Au for 2, 3, 5, 10, 20, 50, 160,
1392: and 21500 A$\cdot$GeV. Upper part: HSD calculations involving the
1393: 'free' $\rho$-meson spectral function: lower part: HSD calculation
1394: involving the in-medium $\rho$ spectral function from Rapp et al.
1395: \protect\cite{RappNPA,CasRap}.} \label{Fig8}
1396: \end{figure}
1397:
1398:
1399: \begin{figure}[t]
1400: \phantom{a}\vspace*{-2.5cm}
1401: \centerline{\psfig{figure=fig9exf.eps,width=16cm}} \vspace*{-3cm}
1402: \caption{The differential dilepton multiplicity $d n_{e^+e^-}/dM$
1403: for central collisions of Au~+~Au at 2, 10, 50, and 160
1404: A$\cdot$GeV from the HSD calculations involving the 'free'
1405: $\rho$-meson spectral function (solid lines) and the in-medium
1406: $\rho$ spectral function from Rapp et al.
1407: \protect\cite{RappNPA,CasRap} (dashed lines) for comparison.}
1408: \label{Fig9}
1409: \end{figure}
1410:
1411:
1412:
1413: \begin{figure}[t]
1414: \phantom{a}\vspace*{-2.5cm}
1415: \centerline{\psfig{figure=fig10exf.eps,width=16cm}}
1416: \vspace*{-5cm}
1417: \caption{The $J/\Psi$ suppression (in terms of the $\mu^+ \mu^-$ decay
1418: branch relative to the Drell-Yan background from 2.9 -- 4.5 GeV
1419: invariant mass) as a function of the transverse energy $E_T$ in
1420: Pb~+~Pb collisions at 160 A$\cdot$GeV. The solid line stands for the
1421: HSD result within the 'late' comover absorption scenario from Ref.
1422: \protect\cite{Cass97d} while the dashed line results from the
1423: 'early' string absorption scenario from Ref. \protect\cite{Geiss2}
1424: involving a transverse string radius $r_s$ = 0.2 fm. The full dots
1425: stand for the NA50 data from 1995 \protect\cite{NA50}, the full
1426: squares for the 1996 data \protect\cite{NA5099}, the open
1427: triangles for the 1996 data with minimum bias
1428: \protect\cite{NA5099} while the open circles represent the 1998
1429: data \protect\cite{NA50QM}. Note that the 1995 data have been
1430: rescaled in $E_T$ as compared to Ref. \protect\cite{NA50}; the
1431: same rescaling has been adopted to the calculations from Refs.
1432: \protect\cite{Cass97d,Geiss2} which had been compared to the
1433: data from \protect\cite{NA50}.}
1434: \label{Fig10}
1435: \end{figure}
1436:
1437:
1438: \begin{figure}[t]
1439: \phantom{a}\vspace*{-2.5cm}
1440: \centerline{\psfig{figure=fig11exf.eps,width=16cm}} \vspace*{-5cm}
1441: \caption{The $J/\Psi$ survival factor $S^{J/\Psi}$ (in terms of
1442: the $\mu^+ \mu^-$ decay branch relative to the Drell-Yan
1443: background from 2.9 -- 4.5 GeV invariant mass normalized to the
1444: same quantity for $pd$ reactions) as a function of the bombarding
1445: energy in central Au~+~Au collisions from 20 to 160 A$\cdot$GeV
1446: (l.h.s.) and from 160 A$\cdot$GeV to 21.5 A$\cdot$TeV (r.h.s.).
1447: The solid line (full dots) stands for the HSD result within the
1448: 'late' comover absorption scenario from Ref.
1449: \protect\cite{Cass97d} while the solid line (open circles) results from
1450: the 'early' string absorption scenario from Ref.
1451: \protect\cite{Geiss2} involving a transverse string radius $r_s$ =
1452: 0.22 fm in order to match both absorption scenarios at 160
1453: A$\cdot$GeV. The dashed lines stand for the relative fraction of
1454: $J/\Psi$ dissociations with baryons while the dotted lines stand
1455: for the 'early' comover string dissociation (open circles) and
1456: 'late' comover meson dissociation (full dots), respectively. The
1457: total suppression factors (full lines) are practically identical
1458: for both scenarios from 20 to 160 A$\cdot$GeV (l.h.s.). This also
1459: holds for the energy range from 160 A$\cdot$GeV to 21.5
1460: A$\cdot$TeV within the numerical accuracy achieved. }
1461: \label{Fig11}
1462: \end{figure}
1463:
1464:
1465: \begin{figure}[t]
1466: \phantom{a}\vspace*{-2.5cm}
1467: \centerline{\psfig{figure=fig12exf.eps,width=16cm}} \vspace*{-5cm}
1468: \caption{The rapidity distribution of protons (upper left part) from
1469: $pp$ collisions at $\sqrt{s}$ = 200 GeV from the HSD approach
1470: (solid histogram) in comparison to the prediction from the parton
1471: cascade VNI \protect\cite{VNI} (dashed histogram); the upper right
1472: part shows the same comparison for the $\pi^+$ and $K^+$ rapidity
1473: distributions, respectively. The lower part of the figure displays
1474: the HSD predictions for central ($b \leq$ 1.5 fm) Au~+~Au at
1475: $\sqrt{s}$ = 200 GeV per nucleon; (l.h.s.) net proton
1476: ($p-\bar{p}$) and $\bar{p}$ rapidity distribution, (r.h.s.)
1477: $\pi^+$ and $K^+$ rapidity distributions.} \label{Fig12}
1478: \end{figure}
1479:
1480: \begin{figure}[t]
1481: \phantom{a}\vspace*{-2.5cm}
1482: \centerline{\psfig{figure=fig13exf.eps,width=16cm}} \vspace*{-5cm}
1483: \caption{The number of baryon-baryon (solid histograms) and
1484: meson-baryon collisions (dashed histograms) as a function of the
1485: invariant energy $\sqrt{s}$ for central ($b \leq$ 1.5 fm) Au~+~Au
1486: collisions at bombarding energies of 2, 11, 160 A$\cdot$GeV and 21
1487: A$\cdot$TeV, respectively, from the HSD model. The arrows indicate the
1488: string thresholds for $mB$ and $BB$ collisions of 2.3 GeV and 2.6
1489: GeV, respectively.} \label{Fig13}
1490: \end{figure}
1491:
1492: \begin{figure}[t]
1493: \phantom{a}\vspace*{-2.5cm}
1494: \centerline{\psfig{figure=fig14exf.eps,width=16cm}} \vspace*{-5cm}
1495: \caption{The $J/\Psi$ survival factor $S_{J/\Psi}$ as a function
1496: of the transverse energy $E_T$ in the rapidity interval (-1 $\leq
1497: y_{cm} \leq$ 1) in Au + Au collisions at $\sqrt{s}$ = 200 GeV in
1498: the 'late' comover model. The solid line represents the result for
1499: $J/\Psi$ dissociation on nucleons and mesons, whereas the dashed
1500: line and the dotted line correspond to the absorption on baryons
1501: and mesons, respectively. The error bars in the figure are due to
1502: statistics only. } \label{Fig14}
1503: \end{figure}
1504:
1505: \end{document}
1506: #!/bin/csh -f
1507: # Uuencoded gz-compressed .tar file created by csh script uufiles
1508: # For more info (11/95), see e.g. http://xxx.lanl.gov/faq/uufaq.html
1509: # If you are on a unix machine this file will unpack itself: strip
1510: # any mail header and call resulting file, e.g., figures.uu
1511: # (uudecode ignores these header lines and starts at begin line below)
1512: # Then say csh figures.uu
1513: # or explicitly execute the commands (generally more secure):
1514: # uudecode figures.uu ; gunzip figures.tar.gz ;
1515: # tar -xvf figures.tar
1516: # On some non-unix (e.g. VAX/VMS), first use editor to change filename
1517: # in "begin" line below to figures.tar-gz , then execute
1518: # uudecode figures.uu
1519: # gzip -d figures.tar-gz
1520: # tar -xvf figures.tar
1521: #
1522: uudecode $0
1523: chmod 644 figures.tar.gz
1524: gunzip -c figures.tar.gz | tar -xvf -
1525: rm $0 figures.tar.gz
1526: exit
1527:
1528: