1: % 29.02.2000; final
2: \documentstyle[12pt,fleqn,rotate,epsfig]{article}
3: \setlength{\oddsidemargin}{+.2cm}
4: \setlength{\evensidemargin}{-.2cm} \setlength{\topmargin}{-.3cm}
5: \setlength{\textwidth}{16.cm} \setlength{\textheight}{21cm}
6: \setlength{\mathindent}{0cm}
7: \newcommand{\be}{\begin{equation}}
8: \newcommand{\ee}{\end{equation}}
9: \newcommand{\bea}{\begin{eqnarray}}
10: \newcommand{\eea}{\end{eqnarray}}
11: \newcommand{\nn}{\nonumber\\}
12: \newcommand{\php}{\phantom{'}\!\!}
13:
14:
15: \renewcommand{\appendix}{
16: % \renewcommand{\section}{
17: % \newpage\thispagestyle{plain}
18: % \secdef\Appendix\sAppendix}
19: \setcounter{section}{0}
20: \renewcommand{\thesection}{\Alph{section}}
21: }
22:
23: %
24: %
25: %+++++++++++++++++++++++++++++++++++++++++++++++
26: %
27: %
28: \begin{document}
29: %
30: \title{Equilibration within a semiclassical off-shell transport
31: approach\footnote{supported by GSI Darmstadt}}
32: %
33: \author{W. Cassing and S. Juchem\\
34: Institut f\"ur Theoretische Physik, Universit\"at Giessen\\
35: 35392 Giessen, Germany}
36: %
37: %
38: \maketitle
39: %
40: %
41: \begin{abstract}
42: Equilibration times for nuclear matter configurations -- modelling
43: intermediate and high energy nucleus-nucleus collisions -- are
44: evaluated within the semiclassical off-shell transport approach
45: developed recently. The transport equations are solved for a
46: finite box in coordinate space employing periodic boundary
47: conditions. The off-shell transport model is shown to give proper
48: off-shell equilibrium distributions in the limit $t \rightarrow
49: \infty$ for the nucleon and $\Delta$-resonance spectral functions.
50: We find that equilibration times within the off-shell approach are
51: only slightly enhanced as compared to the on-shell limit for the
52: momentum configurations considered.
53: \end{abstract}
54:
55: \vspace{2cm}
56: %
57: %
58: PACS: 24.10.Cn; 24.10.-i; 25.70.-z
59: \\ Keywords: Many-body theory;
60: Nuclear-reaction models and methods; Low and intermediate energy
61: heavy-ion reactions
62:
63: \newpage
64: \section{Introduction}
65: The dynamical description of strongly interacting systems out of
66: equilibrium nowadays is dominantly based on transport theories and
67: efficient numerical recipies have been set up for the solution of
68: the coupled channel transport equations \cite{URQMD,CB99} (and
69: Refs. therein). These transport approaches have been derived
70: either from the Kadanoff-Baym equations \cite{kb62} in Refs.
71: \cite{pd841,Bot,Mal,ph95,gl98} or from the hierarchy of connected
72: equal-time Green functions \cite{Zuo,Wang1} in Refs.
73: \cite{CaWa,CNW} by applying a Wigner transformation and
74: restricting to first order in the derivatives of the phase-space
75: variables ($X, P$). Whereas theoretical formulations of off-shell
76: quantum transport have been limited to the formal level for a
77: couple of years \cite{Bot,ph95,Rudy1,Rudy2} only recently a
78: tractable semiclassic form has been derived for testparticles in
79: the eight dimensional phase space of a particle
80: \cite{ca1,ca2,Leupold}.
81:
82: Whereas in Refs. \cite{ca1,ca2} we have
83: investigated the off-shell transport approach with respect to
84: nucleus-nucleus collisions at GANIL, SIS and AGS energies, we here
85: concentrate on equilibration phenomena relative to the on-shell
86: dynamics by imposing periodic boundary conditions for the system
87: confined to a box of size $V= L^3$, where $L$ denotes the length
88: of the cubic box. We, furthermore, compare the equilibrium nucleon
89: and $\Delta$ distribution functions ($t \rightarrow \infty$)
90: to the statistical model (SM)
91: employing the same spectral functions as in the transport
92: approach. For related studies at higher bombarding energies or
93: energy densities within on-shell transport approaches we refer the
94: reader to Refs. \cite{Brat2000,Brav1,Brav2,Brav3,Solfr99}.
95:
96: \section{Extended semiclassical transport equations}
97: We briefly recall the basic equations for Green functions and
98: particle self energies that are exploited in the derivation of
99: off-shell transport equations in the semiclassical limit.
100:
101: %
102: The general starting point for the derivation of a transport
103: equation for particles with a finite and dynamical width are the
104: Dyson-Schwinger equations for the retarded and advanced Green
105: functions $S^{ret}$, $S^{adv}$ and for the non-ordered Green
106: functions $S^{<}$ and $S^{>}$ \cite{kb62}. In the case of scalar
107: bosons -- which is considered in the following for simplicity --
108: these Green functions are defined by
109: %
110: \bea \hspace{2.0cm} i \, S^{<}_{xy} & := & \; < \,
111: \Phi^{\dagger}(y) \; \Phi(x) \, > \, , \; \; i \, S^{>}_{xy} := \;
112: <\, \Phi(x) \; \Phi^{\dagger}(y) \, > \, , \nonumber\\ i \,
113: S^{ret}_{xy} & := & \phantom{-} \: \Theta (x_0 - y_0) \, < \, [ \,
114: \Phi(x) \, , \, \Phi^{\dagger}(y) \, ] \, > \, , \nonumber\\ i \,
115: S^{adv}_{xy} & := & - \: \Theta (y_0 - x_0) \, < \, [ \, \Phi(x)
116: \, , \, \Phi^{\dagger}(y) \, ] \, > \, . \eea
117: %
118: They depend on the space-time coordinates $x,y$ as indicated by
119: the indices $\cdot_{xy}$. The Green functions are determined via
120: Dyson-Schwinger equations by the retarded and advanced self
121: energies $\Sigma^{ret},\Sigma^{adv}$ and the collisional self
122: energy $\Sigma^{<}$:
123: %
124: %
125: %
126: \bea \hat{S}_{0x}^{-1} \; S_{xy}^{ret} \; \: = \; \: \delta_{xy}
127: \; \: + \; \: \Sigma_{xz}^{ret} \: \odot \: S_{zy}^{ret} \; ,
128: \label{dsret_spatial} \eea
129: %
130: \bea \hat{S}_{0x}^{-1} \; S_{xy}^{adv} \; \: = \; \: \delta_{xy}
131: \; \: + \; \: \Sigma_{xz}^{adv} \: \odot \: S_{zy}^{adv} \; ,
132: \label{dsadv_spatial} \eea
133: %
134: \bea \hat{S}_{0x}^{-1} \; S_{xy}^{<} \; \: = \; \:
135: \Sigma_{xz}^{ret} \: \odot \: S_{zy}^{<} \; \: + \; \:
136: \Sigma_{xz}^{<} \: \odot \: S_{zy}^{adv} \: , \label{kb_spatial}
137: \eea \\
138: %
139: where Eq. (\ref{kb_spatial}) is the well-known Kadanoff-Baym
140: equation. Here $\hat{S}^{-1}_{0x}$ denotes the (negative)
141: Klein-Gordon differential operator which is given for bosonic
142: field quanta of (bare) mass $M_0$ by $\hat{S}^{-1}_{0x} = -
143: (\partial^{\mu}_{x} \partial^{x}_{\mu} + M^2_0 )$; $\delta_{xy}$
144: represents the four-dimensional $\delta$-distribution $\delta_{xy}
145: \equiv \delta^{(4)}(x-y)$ and the symbol $\odot$ indicates an
146: integration (from $-\infty$ to $\infty$) as well as a summation
147: over all discrete intermediate variables (cf. \cite{ph95,ca1}).
148:
149: \subsection{The semiclassical limit}
150: For the derivation of a semiclassical transport equation one now
151: changes from a pure space-time formulation into the
152: Wigner-representation. The theory is then formulated in terms of
153: the center-of-mass variable $X = (x+y)/2$ and the momentum $P$,
154: which is introduced by Fourier-transformation with respect to the
155: relative space-time coordinate $(x-y)$. In any semiclassical
156: transport theory one, furthermore, keeps only contributions up to
157: the first order in the space-time gradients. After carrying-out
158: these two steps the Dyson-Schwinger equations
159: (\ref{dsret_spatial})-(\ref{kb_spatial}) become
160: %
161: \bea \left[ \, P^2 \, - \, M^2_0 \, + \, i P^{\mu}
162: \partial^{X}_{\mu} \, \right] \: S^{ret}_{XP} \; = \; 1 \: + \: (
163: \, 1 \, - \, i \, \Diamond \, ) \, \{ \, \Sigma^{ret}_{XP} \, \}
164: \: \{ \, S^{ret}_{XP} \, \} \: , \label{dsret_wignerfo} \eea
165: %
166: \bea \left[ \, P^2 \, - \, M^2_0 \, + \, i P^{\mu}
167: \partial^{X}_{\mu} \, \right] \: S^{adv}_{XP} \; = \; 1 \: + \: (
168: \, 1 \, - \, i \, \Diamond \, ) \, \{ \, \Sigma^{adv}_{XP} \, \}
169: \: \{ \, S^{adv}_{XP} \, \} \: , \label{dsadv_wignerfo} \eea
170: %
171: \bea \left[ \, P^2 \, - \, M^2_0 \, + \, i P^{\mu}
172: \partial^{X}_{\mu} \, \right] \: S^{<}_{XP} \; = \; ( \, 1 \, - \,
173: i \, \Diamond \, ) \, \left[ \: \{ \, \Sigma^{ret}_{XP} \, \} \:
174: \{ \, S^{<}_{XP} \, \} \; + \;
175: % ( \, 1 \, - \, i \, \Diamond \, ) \,
176: \{ \, \Sigma^{<}_{XP} \, \} \: \{ \, S^{adv}_{XP} \, \} \: \right]
177: , \label{kb_wignerfo} \eea\\
178: %
179: %
180: %
181: %
182: %
183: where the operator $\Diamond$ is defined as \cite{ph95,ca1}
184: %
185: \bea
186: \Diamond \, \{ \, F_{1} \, \} \, \{ \, F_{2} \, \}
187: \; := \; \frac{1}{2} \left( \frac{\partial F_{1}}{\partial
188: X^{\mu}} \: \frac{\partial F_{2}}{\partial P_{\mu}} \; - \;
189: \frac{\partial F_{1}}{\partial P_{\mu}} \: \frac{\partial
190: F_{2}}{\partial X^{\mu}} \right) .\label{poissonoperator} \eea\\
191: %
192: It is a four-dimensional generalization of the well-known
193: Poisson-bracket. Starting from (\ref{dsret_wignerfo}) and
194: (\ref{dsadv_wignerfo}) one obtains algebraic relations between the
195: real and the imaginary part of the retarded Green functions. On
196: the other hand eq. (\ref{kb_wignerfo}) leads to a 'transport
197: equation' for the Green function $S^{<}$ \cite{ca1}.
198:
199: To this aim one separates all retarded and advanced quantities --
200: Green functions and self energies -- into real and imaginary
201: parts,
202: %
203: \bea S_{XP}^{ret,adv} \; = \;
204: %G_{XP}
205: Re S^{ret}_{XP} \; \mp \; \frac{i}{2} \, A_{XP} \; , \qquad
206: \Sigma_{XP}^{ret,adv} \; = \; Re \Sigma^{ret}_{XP} \; \mp \;
207: \frac{i}{2} \, \Gamma_{XP} \; . \label{ret_sep} \eea\\
208: %
209: The imaginary part of the retarded propagator is given (up to a
210: factor 2) by the normalized spectral function
211: %
212: \bea A_{XP} \: = \: i \left[ \, S_{XP}^{ret} \: - \: S_{XP}^{adv}
213: \, \right] \: = \: - 2 \, Im \, S^{ret}_{XP} \; , \qquad \qquad
214: \qquad \int \frac{d P_0^2}{4 \pi} \, A_{XP} \; = \; 1 \; ,
215: \label{spectralfunction} \eea\\
216: %
217: while the imaginary part of the self energy corresponds to half
218: the particle width $\Gamma_{XP}$. By separating the complex equations
219: (\ref{dsret_wignerfo}) and (\ref{dsadv_wignerfo}) into their real
220: and imaginary contributions we obtain an algebraic equation
221: between the real and the imaginary part of $S^{ret}$,
222: %
223: \bea Re S^{ret}_{XP} \; = \; \frac{P^{2} \: - \: M_{0}^{2} \: - \:
224: Re \Sigma^{ret}_{XP}}{\Gamma_{XP}} \; A_{XP} \, .
225: \label{dispersion} \eea\\
226: %
227: In addition we gain an algebraic solution for the spectral
228: function as
229: %
230: \bea A_{XP} \; = \; \frac{ \Gamma_{XP} } {( \, P^2 \, - \,
231: M_{0}^{2} \, - \, Re \Sigma^{ret}_{XP} )^{2} \: + \:
232: \Gamma_{XP}^{2}/4} \; , \label{alg_spectral} \eea\\
233: %
234: while the real part of the retarded propagator is given by
235: %
236: \bea Re S^{ret}_{XP} \; = \; \frac{P^{2} \: - \: M_{0}^{2} \: - \:
237: Re \Sigma^{ret}_{XP}} {( \, P^2 \, - \, M_{0}^{2} \, - \, Re
238: \Sigma^{ret}_{XP} )^{2} \: + \: \Gamma_{XP}^{2}/4} \, .
239: \label{alg_realpart} \eea \\
240: %
241: Furthermore, the (Wigner-transformed) Kadanoff-Baym equation
242: (\ref{kb_wignerfo}) allows for the construction of a transport
243: equation for the Green function $S^{<}$. When separating the real
244: and the imaginary contribution of this equation we find i) a
245: generalized transport equation,
246: %
247: \bea \Diamond \, \{ \, P^{2} &-& M_{0}^{2} \: - \: Re
248: \Sigma^{ret}_{XP} \, \} \; \{ \, S^{<}_{XP} \, \} \; - \; \Diamond
249: \, \{ \, \Sigma^{<}_{XP} \, \} \; \{ Re S^{ret}_{XP} \, \}
250: \nonumber\\[0.2cm] &=& \frac{i}{2} \: \left[ \: \Sigma^{>}_{XP} \:
251: S^{<}_{XP} \; - \;
252: \Sigma^{<}_{XP} \: S^{>}_{XP} \: \right],
253: \label{general_transport} \eea\\
254: %
255: and ii) a generalized mass-shell constraint
256: %
257: \bea [ \, P^{2} &-& M_{0}^{2} \: - \: Re \Sigma^{ret}_{XP} \, ] \;
258: S^{<}_{XP} \; - \; \Sigma^{<}_{XP} \; Re S^{ret}_{XP}
259: \nonumber\\[0.2cm] &=& \frac{1}{2} \, \Diamond \; \{ \,
260: \Sigma^{<}_{XP} \, \} \; \{ \, A_{XP} \, \} \; - \; \frac{1}{2} \,
261: \Diamond \; \{ \, \Gamma_{XP} \, \} \; \{ \, S^{<}_{XP} \, \} \; .
262: \label{general_massshell} \eea\\
263: %
264: Besides the drift term (i.e. $\Diamond \{P^2 - M^2_0\} \{ S^{<} \}
265: = - P^{\mu} \partial^{X}_{\mu} S^{<} )$ and the Vlasov term (i.e.
266: $-\Diamond \{ Re \Sigma^{ret} \} \{ S^{<} \} $) a third
267: contribution appears on the l.h.s. of (\ref{general_transport})
268: (i.e. $-\Diamond \{\Sigma^{<} \} \{Re S^{ret} \}$), which vanishes
269: in the quasiparticle limit and incorporates -- as shown in
270: \cite{ca1,ca2,Leupold} -- the off-shell behaviour in the particle
271: propagation which has been neglected so far in transport
272: studies\footnote{This also holds true for the recent numerical
273: off-shell simulations in Refs. \cite{Effe1,Effe}}. The r.h.s. of
274: (\ref{general_transport}) consists of a collision term with its
275: characteristic gain ($\sim \Sigma^{<} S^{>}$) and loss ($\sim
276: \Sigma^{>} S^{<}$) structure, where scattering and decay processes of
277: particles are described.
278:
279:
280: Within the specific term ($\Diamond \{\Sigma^{<} \} \{Re S^{ret}
281: \}$) a further modification is necessary. According to Botermans
282: and Malfliet \cite{Bot} the collisional self energy $\Sigma^{<}$
283: should be replaced by $S^{<} \cdot \Gamma / A$ to gain a
284: consistent first order gradient expansion scheme. The replacement
285: is allowed since the difference between these two expressions
286: $(\Sigma^< - S^< \cdot \Gamma/A)$ can be shown to be of first
287: order in the space-time gradients itself \cite{ca1}. Furthermore,
288: this substitution is required to get rid of the inequivalence
289: between the general transport equation (\ref{general_transport})
290: and the general mass shell constraint (\ref{general_massshell})
291: \cite{Knoll}.
292:
293: Finally, the general transport equation (in first order gradient
294: expansion) reads \cite{ca1,ca2,Leupold}
295: %
296: \setlength{\mathindent}{-0.5cm} \bea A_{XP} \, \Gamma_{XP} &&
297: \!\!\!\!\! \!\!\!\!\! \left[ \, \Diamond \; \{ \, P^2 - M_0^2 - Re
298: \Sigma^{ret}_{XP} \, \} \; \{ \, S^<_{XP} \, \} \: - \:
299: \frac{1}{\Gamma_{XP}} \; \Diamond \; \{ \, \Gamma_{XP} \, \} \; \{
300: \, ( \, P^2 - M_0^2 - Re \Sigma^{ret}_{XP} \, ) \, S^<_{XP} \, \}
301: \, \right] \nonumber\\[0.2cm] && \; = \; i \, \left[ \,
302: \Sigma^>_{XP} \: S^<_{XP} \: - \: \Sigma^<_{XP} \: S^>_{XP} \,
303: \right]. \label{trans_approx} \eea \\
304: \setlength{\mathindent}{0.5cm}
305: %
306: Its formal structure is fixed by the approximations applied,
307: however, its physical contents is fully determined by the
308: different self energies, i.e. $Re \Sigma_{XP}^{ret}, \Gamma_{XP},
309: \Sigma_{XP}^<$ that have to be specified in
310: addition.
311:
312:
313: In order to obtain an approximate solution to the transport
314: equation (\ref{trans_approx}) a testparticle ansatz is used for
315: the Green function $S^{<}$, more specifically for the real and
316: positive semidefinite quantity $F_{XP} \; = A_{XP} N_{XP} = \; i
317: \, S^{<}_{XP}$,
318: %
319: \bea F_{XP} \; \sim \; \sum_{i=1}^{N} \; \delta^{(3)} ({\vec{X}}
320: \, - \, {\vec{X}}_i (t)) \; \; \delta^{(3)} ({\vec{P}} \, - \,
321: {\vec{P}}_i (t)) \; \; \delta(P_0 - \, \epsilon_i(t)) \: .
322: \label{testparticle} \eea
323: %
324: In the most general case (where the self energies depend
325: on four-momentum $P$, time $t = X_0$ and the spatial coordinates
326: $\vec{X}$) the equations of motion for the testparticles read \cite{ca2}
327: %
328: %\setlength{\mathindent}{-0.2cm}
329: \bea \label{eomr} \frac{d {\vec X}_i}{dt} \! & = & \! \phantom{- }
330: \frac{1}{1 - C_{(i)}} \, \frac{1}{2 \epsilon_i} \: \left[ \, 2 \,
331: {\vec P}_i \, + \, {\vec \nabla}_{P_i} \, Re \Sigma^{ret}_{(i)} \,
332: + \, \frac{ \epsilon_i^2 - {\vec P}_i^2 - M_0^2 - Re
333: \Sigma^{ret}_{(i)}}{\Gamma_{(i)}} \: {\vec \nabla}_{P_i} \,
334: \Gamma_{(i)} \: \right],
335: \\[0.3cm]
336: %
337: %
338: \label{eomp} \frac{d {\vec P}_i}{d t} \! & = & \! -
339: \frac{1}{1-C_{(i)}} \, \frac{1}{2 \epsilon_{i}} \: \left[ {\vec
340: \nabla}_{X_i} \, Re \Sigma^{ret}_i \: + \: \frac{\epsilon_i^2 -
341: {\vec P}_i^2 - M_0^{2} - Re \Sigma^{ret}_{(i)}}{\Gamma_{(i)}} \:
342: {\vec \nabla}_{X_i} \, \Gamma_{(i)} \: \right],
343: \\[0.3cm]
344: %
345: %
346: \label{eome} \frac{d \epsilon_i}{d t} \!
347: \setlength{\mathindent}{-0.5cm} & = & \! \phantom{- } \frac{1}{1 -
348: C_{(i)}} \, \frac{1}{2 \epsilon_i} \: \left[ \frac{\partial Re
349: \Sigma^{ret}_{(i)}}{\partial t} \: + \: \frac{\epsilon_i^2 - {\vec
350: P}_i^2 - M_0^{2} - Re \Sigma^{ret}_{(i)}}{\Gamma_{(i)}} \:
351: \frac{\partial \Gamma_{(i)}}{\partial t} \right], \eea\\
352: %
353: where the notation $F_{(i)}$ implies that the function is taken at
354: the coordinates of the testparticle, i.e. $F_{(i)} \equiv
355: F(t,\vec{X}_{i}(t),\vec{P}_{i}(t),\epsilon_{i}(t))$.
356:
357: In (\ref{eomr}-\ref{eome}) a common multiplication factor
358: $(1-C_{(i)})^{-1}$ appears, which contains the energy derivatives
359: of the retarded self energy
360: %
361: \bea \label{correc} C_{(i)} \: = \: \frac{1}{2 \epsilon_i} \left[
362: \frac{\partial}{\partial \epsilon_i} \, Re \Sigma^{ret}_{(i)} \: +
363: \: \frac{\epsilon_i^2 - {\vec P}_i^2 - M_0^2 - Re
364: \Sigma^{ret}_{(i)}}{\Gamma_{(i)}} \: \frac{\partial }{\partial
365: \epsilon_i} \, \Gamma_{(i)} \right] \: , \eea
366: %
367: which yields a shift of the system time $t$ to the 'eigentime' of
368: particle $i$ defined by $\tilde{t}_{i} = t /(1-C_{(i)})$. The
369: derivatives with respect to the 'eigentime', i.e. $d \vec{X}_i / d
370: \tilde{t}_i$, $d \vec{P}_i / d \tilde{t}_i$ and $d \epsilon_i / d
371: \tilde{t}_i$ then emerge without this renormalization factor for
372: each testparticle $i$ when neglecting the explicit time dependence
373: of $C_{(i)}$ in line with the semiclassical approximation scheme.
374: The role and the importance of this correction factors have been
375: studied in Ref. \cite{ca2} for a four-momentum-dependent 'trial'
376: potential and we refer the reader to the latter analysis for more
377: details.
378:
379: Following Ref. \cite{ca1} we take $M^{2} = P^2 - Re \Sigma^{ret}$
380: as an independent variable instead of the energy $P_0$.
381: Eq. (\ref{eome}) then
382: turns to
383: %
384: \bea \label{eomm} \frac{dM_i^2}{dt} \; = \; \frac{M_i^2 -
385: M_0^2}{\Gamma_{(i)}} \; \frac{d \Gamma_{(i)}}{dt} \eea
386: %
387: for the time evolution of the test-particle $i$ in the invariant
388: mass squared as derived in Refs. \cite{ca1,ca2}. We mention that
389: corresponding equations of motion have been derived by Leupold in
390: the nonrelativistic limit \cite{Leupold}.
391:
392: \subsection{Generalized collision terms for bosons and fermions}
393: %
394: The collision term of the Kadanoff-Baym equation can only be
395: worked out in more detail by giving explicit approximations for
396: $\Sigma^{<}$ and $\Sigma^{>}$. We recall the formulation and
397: result from Ref. \cite{ca2} that is based on Dirac-Brueckner
398: theory, i.e. $$ I_{coll}(X,\vec{P},M^2) = Tr_2 Tr_3 Tr_4 A(X,{\vec
399: P},M^2) A(X,{\vec P}_2, M_2 ^2) A(X,{\vec P}_3, M_3 ^2) A(X,{\vec
400: P}_4, M_4 ^2) $$ $$
401: %\{
402: |T(({\vec P},M^2) + ({\vec P}_2,M_2^2) \rightarrow ({\vec
403: P}_3,M_3^2) + ({\vec P}_4,M_4^2))|_{{\cal A,S}}^2 \; \;
404: \delta^{(4)}({P} + {P}_2 - {P}_3 - {P}_4) $$
405: \be
406: \label{Icoll} \hspace{1cm} [ \, N_{X{\vec P}_3 M_3^2} \, N_{X
407: {\vec P}_4 M_4^2} \, {\bar f}_{X {\vec P} M^2} \, {\bar f}_{X
408: {\vec P}_2 M_2^2} \: - \: N_{X{\vec P} M^2} \, N_{X {\vec P}_2
409: M_2^2} \, {\bar f}_{X {\vec P}_3 M_3^2} \, {\bar f}_{X {\vec P}_4
410: M_4^2} \, ]
411: %\}
412: \ee with
413: \be
414: \label{pauli} {\bar f}_{X {\vec P} M^2} = 1 + \eta \, N_{X {\vec
415: P} M^2} \ee and $\eta = \pm 1$ for bosons/fermions, respectively.
416: The indices ${\cal A,S}$ stand for the antisymmetric/symmetric
417: matrix element of the in-medium off-shell scattering amplitude $T$
418: in case of fermions/bosons. In Eq. (\ref{Icoll}) the trace over
419: particles 2,3,4 reads explicitly for fermions
420: \be
421: \label{trace} Tr_2 = \sum_{\sigma_2, \tau_2} \frac{1}{(2 \pi)^4}
422: \int d^3 P_2 \frac{d M^2_2}{2 \sqrt{\vec{P}^2_2+M^2_2}}, \ee where
423: $\sigma_2, \tau_2$ denote the spin and isospin of particle 2. In
424: case of bosons we have instead
425: \be
426: \label{trace2} Tr_2 = \sum_{\sigma_2, \tau_2} \frac{1}{(2 \pi)^4}
427: \int d^3 P_2 \frac{d P_{0,2}^2}{2}, \ee since here the spectral
428: function $A_B$ is normalized as
429: \be
430: \label{sb} \int \frac{d P_0^2}{4 \pi} A_B(X,P) = 1 \ee whereas for
431: fermions we have
432: \be
433: \label{sb1} \int \frac{d P_0}{2 \pi} A_F(X,P) = 1. \ee We mention
434: that the spectral function $A_F$ in case of fermions in
435: (\ref{Icoll}) is obtained by considering only particles of
436: positive energy and assuming the spectral function to be identical
437: for spin 'up' and 'down' states (cf. Ref. \cite{ca2}).
438:
439: Neglecting the 'gain-term' in Eq. (\ref{Icoll}) one recognizes
440: that the collisional width $\Gamma_{coll}$ of the particle in the
441: rest frame is given by
442: \be
443: \label{gcoll} \Gamma_{coll}(X,\vec{P},M^2) = Tr_2 Tr_3 Tr_4 \;
444: %\{
445: |T(({\vec P},M^2) + ({\vec P}_2,M_2^2) \rightarrow ({\vec
446: P}_3,M_3^2) + ({\vec P}_4,M_4^2))|_{{\cal A,S}}^2 \ee $$ A(X,{\vec
447: P}_2,M_2^2) A(X,{\vec P}_3,M_3^2) A(X,{\vec P}_4, M_4^2) \; \;
448: \delta^4(P + P_2 - P_3-P_4) \ N_{X {\vec P}_2 M_2^2} \, {\bar
449: f}_{X {\vec P}_3 M^2_3} \, {\bar f}_{X {\vec P}_4 M^2_4} \,
450: %\}
451: , $$ where as in Eq. (\ref{Icoll}) local off-shell transition
452: amplitudes enter for the transitions $P + P_2 \rightarrow P_3 +
453: P_4$. We note that the extension of Eq. (\ref{Icoll}) to inelastic
454: scattering processes (e.g. $NN \rightarrow N\Delta$) or ($\pi N
455: \rightarrow \Delta$ etc.) is straightforward when exchanging the
456: elastic transition amplitude $T$ by the corresponding inelastic
457: one and taking care of Pauli-blocking or Bose-enhancement for the
458: particles in the final state. The relation of the quantity
459: $\Gamma_{XP}$ to the collisional width $\Gamma_{coll}$ is given by
460: $\Gamma_{XP} = 2 P_0 (\Gamma_{decay}(XP) + \Gamma_{coll}(XP))$,
461: where the particle decay width $\Gamma_{decay}$ in the medium
462: might also be different compared to the vacuum.
463:
464: Thus the transport approach determines the particle spectral
465: function dynamically via (\ref{gcoll}) -- with respect to the collisional
466: width $\Gamma_{coll}$ -- for all hadrons if the
467: in-medium transition amplitudes $T$ are known in their full
468: off-shell dependence. Since this information is not available for
469: configurations of hot and dense matter, a couple of assumptions
470: and numerical approximation schemes have to be invoked in actual
471: applications so far.
472:
473: As in Refs. \cite{ca1,ca2} the following dynamical calculations are
474: based on the conventional HSD transport approach
475: \cite{CB99,Ehehalt} -- in which $Re \Sigma^{ret}_{XP}$ is
476: specified for the hadrons -- however, the equations of motion for
477: the testparticles are extended to (\ref{eomr},\ref{eomp},\ref{eomm}).
478: For further details on the elastic and inelastic transition rates we
479: refer the reader to Ref. \cite{ca2}.
480:
481: \subsection{Comment on particle number conservation}
482: In previous derivations of the off-shell transport equations
483: one has started from a formulation of the non-equilibrium theory
484: in space-time representation $(x,x')$ and then changed into a
485: phase-space representation via Fourier transformation with respect
486: to $(x-x')$ \cite{Bot,ca1,ca2,Leupold}.
487: The semiclassical limit then has been introduced by assuming gradients
488: in $X=(x+x')/2$ to be small \cite{Leupold} for $Re \Sigma^{ret}_{XP}$ and $S^<_{XP}$.
489: Here we argue that for reasons of symmetry in phase-space also the
490: four-momentum derivatives in $P$ have to be small to achieve a
491: proper semiclassical limit as inherent in the formulation of the
492: transport equation (\ref{trans_approx}) in
493: terms of the generalized Poisson-bracket (\ref{poissonoperator}).
494:
495: We briefly demonstrate in the following lines that instead of a coordinate-space
496: representation one may formulate the theory equivalently in
497: momentum-space and then transform to phase space:
498: The momentum-dependent (two-point) functions are given as
499: %
500: %
501: \bea F_{p_{1\php} \, p_{1'}} \: = \:
502: \int \! d^{4}\!x_1 \! \int d^{4}\!x_{1'} \; \;
503: e^{i ( p_{1\php}x_{1\php} - \, p_{1'} x_{1'})}
504: \; \; F_{x_{1\php} \, x_{1'}}.
505: \label{fourier_transformation}
506: \eea\\
507: %
508: The evolution of the retarded and advanced Green functions $S^{ret},
509: S^{adv}$ and the Green function $S^<$ turns to
510: %
511: \bea
512: \hat{S}_{0 \, p_{1\php}}^{-1} \; S_{p_{1\php} \, p_{1'}}^{ret,adv}
513: \; & = & \:
514: (2 \pi)^4 \; \delta^{(4)}(p_{1\php} - p_{1'})
515: \; \: +\: \:
516: \int \frac{d^{4}p_2 }{(2 \pi)^{4}} \; \;
517: \Sigma_{p_{1\php} \, p_{2\php}}^{ret,adv} \; \;
518: S_{p_{2\php} \, p_{1'}}^{ret,adv} \; , \\
519: %
520: \hat{S}_{0 \, p_{1\php}}^{-1} \; S_{p_{1\php} \, p_{1'}}^{<}
521: \; & = & \:
522: \int \frac{d^{4}p_2 }{(2 \pi)^{4}} \; \; \left[ \;
523: \Sigma_{p_{1\php} \, p_{2\php}}^{ret} \; S_{p_{2\php} \, p_{1'}}^{<}
524: \; \: + \; \:
525: \Sigma_{p_{1\php} \, p_{2\php}}^{<} \; \ S_{p_{2\php} \, p_{1'}}^{adv}
526: \; \right] \: ,
527: \label{kb_momentum}
528: \eea \\
529: %
530: with the 'kinetic' operator in momentum-space
531: $\hat{S}_{0 \, p_1}^{-1} = (p_1^2-M^2_0)$ in the case of
532: relativistic scalar bosons with (bare) mass $M_0$.
533: Obviously, the equations in momentum-space are formally equivalent
534: to the equations in coordinate-space, i.e. the convolution integrals
535: in the coordinate $x_2$ are replaced by convolution integrals in
536: the momentum $p_2$.
537: When transforming from momentum- to phase-space via
538: a Fourier-transformation with respect to the four-momentum
539: coordinate $(p_1-p_{1'})$
540: %
541: \bea
542: F_{XP} \: = \: \int \frac{d^{4}(p_1\!-\!p_{1'})}{(2 \pi)^4} \; \;
543: e^{- i X (p_{1\php} - p_{1'})} \; \; F_{p_{1\php} \, p_{1'}}
544: \label{wigner_transformation_2}
545: \eea\\
546: %
547: (with $P=(p_1+p_{1'})/2$) we gain again the familiar equations in
548: Wigner-representation:
549: %
550: \bea
551: \hat{S}_{0 \, XP}^{-1} \; S_{XP}^{ret,adv}
552: \; & = & \; 1
553: \; \: + \: \:
554: e^{- i \Diamond} \; \; \Sigma_{XP}^{ret,adv} \; S_{XP}^{ret,adv} \\[0.5cm]
555: %
556: \hat{S}_{0 \, XP}^{-1} \; S_{XP}^{<}
557: \; & = & \:
558: e^{- i \Diamond} \;
559: \left[ \; \Sigma_{XP}^{ret} \; \;
560: S_{XP}^{<}
561: \; \: + \; \:
562: \Sigma_{XP}^{<} \; \;
563: S_{XP}^{adv} \; \right] \: ,
564: \label{kb_wigner}
565: \eea \\
566: %
567: with the 'kinetic' operator in phase-space
568: $\hat{S}_{0 \, XP}^{-1} = (P^2 + i P^{\mu} \partial^{X}_{\mu}
569: - \frac{1}{4} \partial^{X}_{\mu} \partial_{X}^{\mu} - M^2_0)$.
570: Here we have used the following relation for the convolution
571: integrals in momentum-space
572: %
573: \bea
574: \int \frac{d^{4}(p_1\!-\!p_{1'})}{(2 \pi)^4} \; \;
575: e^{-i X (p_{1} - p_{1'})} \;
576: \int \frac{d^{4}p_2}{(2 \pi)^4} \; \;
577: \; F_{1, \, p_{1\php} \, p_{2\php}} \; F_{2, \, p_{2\php} \, p_{1'}}
578: \; \; = \; \; e^{-i \Diamond} \; \; F_{1, \, PX} \; \;
579: F_{2, \, PX},
580: \label{wigner_convolution}
581: \eea\\
582: %
583: with the Poisson-operator $\Diamond$ defined in (\ref{poissonoperator}).
584:
585: The semiclassical limit -- along the conventional line of arguments --
586: now is achieved by assuming gradients of the self energies $\Sigma^{ret,adv}$ in
587: $P$ to be small. This assumption is especially well taken for systems
588: with dominantly short range interactions since the different self energies
589: $\Sigma^{ret,adv}(p,p')$ become smooth functions in momentum-space
590: such that a restriction to first order gradients in the momentum
591: $P=(p+p')/2$ can be more easily justified. Note again that
592: the four-dimensional Poisson-bracket (\ref{poissonoperator}) is
593: symmetric in the space-time and the four-momentum derivatives.
594: Thus a phase-space gradient expansion requires all gradients
595: to be small contrary to the assumptions made in Ref. \cite{Leupold}.
596:
597: As a consequence mixed gradients of second order as
598: $\partial^2/(\partial t \partial p_0)$
599: have to be neglected in a consistent truncation scheme of first
600: order. As shown by Botermans and Malfliet in the review \cite{Bot} the particle
601: number conservation then holds strictly.
602: Only when keeping special terms of second order
603: $(\sim \partial^2/(\partial t \partial p_0)$
604: unphysical pecularities may appear, that lead to a violation of particle
605: number conservation in contradiction to the full theory (2)-(4).
606:
607:
608:
609: \section{Infinite nuclear matter problems}
610: In case of infinite nuclear matter problems the solution of the
611: transport equations simplifies since all spatial gradients
612: (with respect to $\vec{X}$) vanish. This implies that the momentum
613: coordinates of the testparticles $\vec{P}_i$ are constant
614: according to (19) in between collisions.
615:
616: The initial conditions of the problem then are fully specified by an initial
617: bombarding energy per nucleon, that defines the relative shift in
618: momentum of two Fermi spheres with a Fermi momentum $P_F$ = 260 MeV/c,
619: the number of nucleons $N$ and the total volume of the box
620: $V=L^3$. Alternatively, one can also characterize the system by an
621: average density $\rho =N/V$ and energy density $\epsilon$ (cf.
622: Ref. \cite{Brat2000}). We use $L = 10$ fm and assume an equal
623: number of neutrons and protons.
624:
625: In general (for $t \rightarrow \infty$) the stationary solution of
626: Eq. (\ref{Icoll}) for a fermion $h$ is given by
627: %
628: \begin{equation}
629: \label{equilibrium}
630: F_h(\vec{X}, \vec{P},M^2)
631: \; = \; \frac{A_h({\vec{X}},{\vec{P}},M^2)}{\exp((E-\mu_h)/T)-\eta}
632: \end{equation}
633: with $E=\sqrt{{\vec{P}}^2 + M^2}$ and $\eta = \pm 1$ for bosons/fermions,
634: respectively,
635: while $A_h$ denotes the spectral function (\ref{alg_spectral}),
636: $T$ the temperature of the system and $\mu_h$ the chemical
637: potential for the hadron. This situation corresponds to the grand canonical
638: ensemble of quantum statistical mechanics. In case of infinite nuclear matter
639: problems the dependence on $\vec{X}$ vanishes additionally.
640:
641: \subsection{Numerical results}
642: As an example for equilibration phenomena we show in Fig.
643: \ref{bild1} the time evolution of the quadrupole moment (involving
644: all baryons $B$)
645: \be
646: \label{quad} Q_2(t) = \sum_B \frac{g_B}{(2 \pi)^4} \int d^3 X \int d^3 P
647: \, \frac{d M^2}{2 \sqrt{{\vec P}^2 + M^2}} \ (2P_z^2-P_x^2-P_y^2)\
648: F_B(\vec{X},t,\vec{P},M^2), \ee for $\rho=\rho_0$ and initial
649: bombarding energies of 0.1 A GeV and 1 A GeV, respectively.
650: In (\ref{quad}) the degeneracy factor is $g_B=4$ for nucleons and $g_B$= 16
651: for $\Delta$-resonances. The
652: solid lines result from the off-shell calculation while the
653: dashed lines result from the on-shell limit. As can be seen from
654: Fig. \ref{bild1} the off-shell results are practically identical
655: to the on-shell limits except for the very long time behaviour, where the
656: off-shell limit needs some more time to achieve equilibrium.
657: However, when defining an equilibration time $\tau$
658: by a drop of the observable by the factor $e^{-1}$ we find no
659: sizeable effect from the off-shell propagation on $\tau$.
660:
661: The equilibrium distributions in the nucleon transverse mass $M_t
662: = \sqrt{p_t^2 + M^2}$ are shown in Fig. \ref{bild2} for initial
663: bombarding energies of 0.1 A GeV (upper part) and 1 A GeV (lower
664: part). Again the solid lines denote the off-shell results from the
665: transport calculations while the on-shell spectra are displayed in
666: terms of dashed lines. Both limits (within the statistics)
667: give the same temperature $T \approx$ 97 MeV for 1.0 A GeV and $T
668: \approx$ 26 MeV for 0.1 A GeV as can be extracted from the high
669: energy tail of the transverse mass spectra. Differences can only
670: be found for $m_T \leq M_0$ since the finite width in the nucleon
671: spectral function only can show up in the off-shell case. This
672: component is quite small at 0.1 A GeV since the collisional width
673: of nucleons at density $\rho_0$ and temperature $T \approx$ 26 MeV
674: is about 8 MeV.
675:
676: Without explicit display we mention that the time evolution of the
677: nucleon spectral function in the invariant mass $M$ becomes broad in
678: the initial nonequilibrium phase of the reaction and approaches the
679: equilibrium distribution (\ref{equilibrium}) roughly within the
680: equilibration times from Fig. \ref{bild1}. Since the width of the
681: nucleon spectral function in equilibrium at 0.1 A GeV (GANIL
682: energy) is rather small ($\Gamma \approx $ 8 MeV) we concentrate on
683: the energy of 1 A GeV (SIS energy) in the following, where the
684: nucleon collisional width $\Gamma_{coll}$ at a temperature of 97
685: MeV is about 40 MeV and roughly 20\% of the baryons are excited
686: $\Delta$ resonances\footnote{Note, that in collisions of finite systems
687: at this bombarding energy the $\Delta$-abundancy is slightly lower due
688: to a rapid expansion of the system and the additional compressional energy
689: stored in the system in the high density phase.}.
690:
691: As demonstrated above, for the equilibrated system we can extract
692: a temperature $T$ by
693: fitting the particle spectra with the Bolzmann distribution
694: \begin{eqnarray}
695: {d^3N_i\over dp^3} \sim \exp(-E_i/T), \label{Boltz}\end{eqnarray}
696: where $E_i=\sqrt{p_i^2+m_i^2}$ is the energy of particle $i$. We
697: note that at the temperatures of interest here, the Bose and Fermi
698: distributions are practically identical to a Boltzmann
699: distribution. We find that in equilibrium the spectra of all
700: particles (nucleons, $\Delta$'s and pions) can be characterized by
701: a single temperature $T$ (cf. e.g. Ref. \cite{Brat2000}).
702:
703: \subsection{Comparison to the statistical model}
704: In order to investigate the equilibrium behaviour of hadron matter
705: we compare our transport (box) calculations with a simple
706: Statistical Model (SM) for an Ideal Hadron Gas where the
707: system is described by a grand canonical ensemble of
708: non-interacting fermions and bosons in equilibrium at temperature
709: $T$. All baryon and meson species considered in the transport
710: model ($N, \Delta, \pi$) also are included in the
711: statistical model.
712:
713: We recall that in the SM particle multiplicities
714: $n_i$ and energy densities $\varepsilon_i$ for particles with
715: spectral functions $A_i$ are given by
716: \begin{eqnarray}
717: && n_i ={g_i \over (2\pi \hbar)^3} \int \frac{dM}{2 \pi}
718: \int\limits_0^\infty {A_i(M)/ 4\pi p^2 dp \over \exp\left[(E_i -
719: B_i\mu_B - S_i \mu_S)/T\right] - \eta}, \label{Nth} \\ &&
720: \varepsilon_i ={g_i \over (2\pi \hbar)^3} \int \frac{dM}{2 \pi}
721: \int\limits_0^\infty {A_i(M)\ 4\pi E_i p^2 dp \over \exp\left[(E_i
722: - B_i\mu_B - S_i\mu_S)/T\right]- \eta}, \label{Enth}
723: \end{eqnarray}
724: where $E_i = \sqrt{p^2+M^2}$ is the energy of particle $i$, $B_i$
725: is the baryon charge, $S_i$ is the strangeness, and $g_i$ is the
726: spin-isospin degeneracy factor, while $\eta=\pm 1$ for
727: bosons/fermions, respectively. In Eqs. (\ref{Nth}),(\ref{Enth})
728: $\mu_B$ and $\mu_S$ are the baryon and strangeness chemical
729: potentials. The energy density $\varepsilon$, baryon density
730: $\rho$ and strange particle density of the whole system in
731: equilibrium then is given as
732: \begin{eqnarray}
733: && \varepsilon = \sum\limits_i \varepsilon_i (T,\mu_B,\mu_S)
734: \label{3eq_en} \\ && \rho = \sum\limits_i B_i \ n_i
735: (T,\mu_B,\mu_S) \label{3eq_rhoB} \\ && \rho_S = \sum\limits_i S_i
736: \ n_i (T,\mu_B,\mu_S) \, \equiv \, 0 . \label{3eq_rhoS}
737: \end{eqnarray}
738: As 'input' for the SM we use the same $\varepsilon, \rho$
739: and $\rho_S$ as in the box calculations and we obtain the
740: thermodynamical parameters -- $T, \mu_B, \mu_S$ -- by solving the
741: system of nonlinear equations (\ref{3eq_en}),(\ref{3eq_rhoB}) and
742: (\ref{3eq_rhoS}).
743:
744: We now turn to a comparison of the equilibrium distributions in
745: mass for nucleons and $\Delta$'s, i.e. $dN_N/dM$ and
746: $dN_\Delta/dM$, from the box calculations with those from the SM,
747: which are obtained by integration (summation) over momentum. We
748: first discuss the on-shell limit where the nucleon spectral
749: function is represented by a $\delta$-function around the bare
750: mass while the $\Delta$ spectral function -- as implemented in the
751: transport approach -- is given by (in the $\Delta$ rest frame)
752: \begin{equation}
753: \label{delta} A_\Delta(M) \; = \; \frac{ 2 M^2 \Gamma_{\pi
754: N}(M)}{(M^2-M_{\Delta0}^2)^2 + M^2 \Gamma^2_{\pi N}(M)}
755: \end{equation}
756: with
757: \begin{equation}
758: \label{delta2} \Gamma_{\pi N}(M) = \Gamma_0 (\frac{q}{q_R})^3 \
759: (\frac{q_R^2 + \delta^2} {q^2 + \delta^2})^3 ,
760: \end{equation}
761: where $q_R$ denotes the pion momentum in the $\Delta$ rest frame
762: of the resonance and $q$ is the corresponding pion three-momentum
763: at invariant mass $M$, i.e.
764: \begin{equation}
765: q^2 = \frac{(M^2 - (M_N+M_\pi)^2) (M^2 - (M_N-M_\pi)^2)}{4 M^2}.
766: \end{equation}
767: The quantity $\delta$ in (\ref{delta2}) is fixed by \cite{Teis}
768: \begin{equation}
769: \delta^2 = (M - M_N - M_\pi)^2 + \frac{\Gamma_0^2}{4}
770: \end{equation}
771: with $\Gamma_0 = 120$ MeV to achieve a good
772: description for the $\pi N \rightarrow \Delta$
773: reaction in vacuum. Recall that in this case we have
774: $\Gamma_{XP}/(2 M) = \Gamma_{decay} = \Gamma_{\pi N}$ in the
775: $\Delta$ rest frame, i.e. $\Gamma_{coll}$ = 0.
776:
777: In the upper part of Fig. \ref{bild3} we compare the asymptotic ($t
778: \rightarrow \infty$) distributions for nucleons ($N$) and
779: $\Delta$'s in the on-shell limit from the transport (box)
780: calculation (solid histograms) with the result from the SM at a
781: temperature $T$ = 97 MeV employing the $\Delta$ spectral function
782: (\ref{delta}). Since the differences are hardly visible, we
783: conclude that the transport calculation reproduces the result from
784: the SM, where the thermodynamical parameters are determined by
785: energy and baryon number conservation. We mention that we have
786: discarded strangeness in this comparison, since kaons and hyperons
787: are very scarce at this energy and have been switched off in both
788: models. Since the $\Delta$ width (\ref{delta2}) is zero below the
789: $\pi N$ threshold, the $\Delta$ mass distribution only can extend
790: above $M_{\pi} + M_N$.
791:
792: In case of the off-shell calculation we obtain somewhat different
793: mass distributions for nucleons and $\Delta$'s from the box
794: calculation, which are given in terms of the solid histograms in
795: the lower part of Fig. \ref{bild3}. Here the nucleon spectral
796: function becomes very broad and also the $\Delta$ distribution
797: extends below the $\pi N$ threshold in vacuum. The nucleon
798: spectral function is broadened due to the elastic ($NN \rightarrow
799: NN$) and inelastic ($NN \rightarrow N\Delta$) scattering
800: processes, which can roughly be described by a collisional width
801: $\Gamma_{coll}$ of 40 MeV in the nucleon spectral function
802: \begin{equation}
803: \label{nucleon} A_N(M) \; = \; \frac{2 M^2 \Gamma_{coll}}
804: {(M^2-M_{N0}^2)^2 + M^2 \Gamma_{coll}^2},
805: \end{equation}
806: as shown by the left dashed line in the lower part of Fig.
807: \ref{bild3}. The collisional width in case of nucleons is related to $\Gamma_{XP}$
808: as $2P_0 \Gamma_{coll}=\Gamma_{XP}$ which reduces in the rest
809: system of the particle to $2M \Gamma_{coll}=\Gamma_{XP}$.
810:
811: The $\Delta$ mass spectrum in this case is more difficult to
812: interprete. This is due to the fact that in the decay $\Delta
813: \rightarrow \pi N^*$ the $\Delta$ may decay to a pion and an
814: off-shell nucleon $N^*$ according to the nucleon equilibrium
815: spectral function $A_N(M)$, which essentially shifts the $\pi N^*$
816: threshold accordingly. However, taking into account this change in
817: width due to the in-medium $\pi N^*$ decay, only, the low mass
818: part of the $\Delta$ distribution is underestimated considerably.
819: Here the 'collisional' channels $\Delta N \rightarrow \Delta N$
820: and $\Delta N \rightarrow NN$ are much more important. We mention
821: that the latter reaction is described by the extended detailed
822: balance relation of Ref. \cite{Wolf} while the elastic
823: differential cross section is taken the same as for $NN$
824: scattering (as a function of the momentum transfer). Since
825: especially the $\Delta N$ absorption reaction depends on mass $M$
826: and momentum $p$ of the baryons explicitly, it is not
827: straightforward to present analytical formulas since final state
828: Pauli blocking for the nucleons leads to a highly nonlinear
829: problem. We thus have extracted the $\Delta$ collisional width
830: $\Gamma^\Delta_{coll}(p,M)$ from the transport calculation
831: explicitly by calculating the number of $\Delta N$ unblocked
832: collisions per unit time as a function of the $\Delta$ momentum
833: $p$ and mass $M$. Using ($p = |\vec{P}|$)
834: \begin{equation}
835: \label{delta3} \frac{\Gamma_{XP}}{2 P_0} = \Gamma_{tot}(p,M) = \Gamma_{\pi N^*}(M) +
836: \Gamma^\Delta_{coll}(p,M) ,
837: \end{equation}
838: where $\Gamma_{\pi N^*}$ denotes the $\Delta$ in-medium $\pi N^*$
839: width averaged over the nucleon equilibrium distribution function
840: (cf. lower part of Fig. 3) and integrating over momentum with the
841: appropriate Fermi function (for $T$ = 97 MeV) we obtain the right
842: dashed line in the lower part of Fig. \ref{bild3} that describes
843: the box $\Delta$- distribution rather well. Thus the transport
844: calculation at finite density $\rho_0$ and temperature $T$ is
845: consistent with the SM when employing the same physical baryon
846: spectral functions. One might have expected this equivalence in
847: equilibrium due to energy and baryon number conservation, however,
848: the actual numerical result then may be regarded as a consistency
849: test of the numerical implementation schemes for the various
850: elastic and inelastic reaction channel in the medium for off-shell
851: particle propagation.
852:
853: The mass integrated number of $\Delta$'s in the off-shell case is
854: larger by a few \% as compared to the on-shell case in equilibrium
855: due to the low mass tail of the $\Delta$ distribution, however,
856: the high mass tails of the $\Delta$'s are identical within the
857: statistics achieved as demonstrated in Fig. \ref{bild4}. On the
858: other hand, the number of pions is practically identical in both
859: cases since the low mass tail of the $\Delta$'s is dominantly due
860: to $\Delta N$ reactions and only to a lower extent to the $\pi
861: N^*$ decay as discussed above.
862:
863:
864:
865: \section{Summary}
866: In this work we have employed the semiclassical off-shell approach
867: from Refs. \cite{ca1,ca2} to study equilibration phenomena in
868: intermediate and high energy nucleus-nucleus collisions in a
869: finite box with periodic boundary conditions. The semiclassical
870: off-shell transport approach describes the virtual propagation of
871: particles in the invariant mass squared $M^2$ besides the
872: conventional propagation in the mean-field potential given by the
873: real part of the self energy. The imaginary part of the retarded
874: self energy $Im \Sigma^{ret}_{XP} = -1/2 \Gamma_{XP} = - P_0
875: (\Gamma_{coll}(XP)
876: + \Gamma_{decay}(XP))$ -- apart from decay contributions to the
877: width -- is determined by the collision integrals
878: themselves and 'evaluated' within the transport approach dynamically.
879:
880: Our explicit calculations demonstrate that the off-shell
881: propagation of nucleons has practically no sizeable effect on
882: equilibration times especially at lower bombarding energies; more
883: importantly, the off-shell dynamics even lead to a slight increase
884: of the equilibration time for kinetic equilibrium as noticed early by
885: Danielewicz \cite{pd841} (cf. also Ref. \cite{CGreiner}).
886: Furthermore, we have demonstrated that the off-shell HSD approach
887: reproduces the 'proper' spectral functions with respect to the
888: statistical model (in case of a grand canonical ensemble) for nucleons and
889: $\Delta$'s that in equilibrium are given analytically once the
890: collisional width $\Gamma_{coll}(\vec{P}, M)$ is known as a
891: function of the 3-momentum $\vec{P}$ and invariant mass $M$.
892:
893: %
894: %
895: \vspace{1cm} \noindent The authors like to acknowledge stimulating
896: discussions with E.L. Bratkovskaya, C. Greiner and S. Leupold
897: throughout this study.
898:
899:
900:
901: %----------------------------------------------------------------------
902: \begin{thebibliography}{999}
903: \bibitem{URQMD}
904: S. Bass, M. Belkacem, M. Bleicher et al.,
905: Prog. Part. Nucl. Phys. 41 (1998) 255.
906: \bibitem{CB99}
907: W. Cassing and E. L. Bratkovskaya, Phys. Rep. 308 (1999) 65.
908: \bibitem{kb62} L. P. Kadanoff and G. Baym,
909: {\it Quantum statistical mechanics}, Benjamin, New York, 1962.
910: \bibitem{pd841}
911: P. Danielewicz, Ann. Phys. (N.Y.) 152 (1984) 239; {\it ibid.} 305.
912: \bibitem{Bot}
913: W. Botermans and R. Malfliet, Phys. Rep. 198 (1990) 115.
914: \bibitem{Mal}
915: R. Malfliet, Prog. Part. Nucl. Phys. 21 (1988) 207.
916: \bibitem{ph95}
917: P. A. Henning, Nucl. Phys. A 582 (1995) 633; Phys. Rep. 253 (1995) 235.
918: \bibitem{gl98}
919: C. Greiner and S. Leupold, Ann. Phys. (N.Y.) 270 (1998) 328.
920: \bibitem{Zuo}
921: S. J. Wang, W. Zuo and W. Cassing, Nucl. Phys. A 573 (1994) 245.
922: \bibitem{Wang1}
923: S. J. Wang and W. Cassing, Ann. Phys. (N.Y.) 159 (1985) 328.
924: \bibitem{CaWa}
925: W. Cassing and S. J. Wang, Z. Phys. A 337 (1990) 1.
926: \bibitem{CNW}
927: W. Cassing, K. Niita and S. J. Wang, Z. Phys. A 331 (1988) 439.
928: \bibitem{Rudy1}
929: R. Malfliet, Nucl. Phys. A 545 (1992) 3.
930: \bibitem{Rudy2}
931: R. Malfliet, Phys. Rev. B 57 (1998) R11027.
932: \bibitem{ca1} W. Cassing and S. Juchem, Nucl.
933: Phys. A 665 (2000) 385.
934: \bibitem{ca2} W. Cassing and S. Juchem, nucl-th/9910052, Nucl.
935: Phys. A, in print.
936: \bibitem{Leupold} S. Leupold, nucl-th/9909080 Nucl. Phys. A, in print.
937: \bibitem{Brat2000}
938: E. L. Bratkovskaya, W. Cassing, C. Greiner et al., nucl-th/0001008.
939: \bibitem{Brav1}
940: M. Belkacem, M. Brandstetter, S.A. Bass et al.,
941: Phys. Rev. C 58 (1998) 1727.
942: \bibitem{Brav2}
943: L.V. Bravina, M.I. Gorenstein, M. Belkacem et al.,
944: Phys. Lett. B 434 (1998) 379;
945: L.V. Bravina, M. Brandstetter, M.I. Gorenstein et al.,
946: J. Phys. G 25 (1999) 351.
947: \bibitem{Brav3}
948: L.V. Bravina, E.E. Zabrodin, M.I. Gorenstein et al.,
949: Phys. Rev. C 60 (1999) 024904.
950: \bibitem{Solfr99}
951: J. Sollfrank, U. Heinz, H. Sorge, N. Xu, Phys. Rev. C 59 (1999) 1637.
952: \bibitem{Effe1} M. Effenberger, E. L. Bratkovskaya and U. Mosel,
953: Phys. Rev. C 60 (1999) 44614.
954: \bibitem{Effe} M. Effenberger and U. Mosel, Phys. Rev. C 60 (1999) 51901.
955: \bibitem{Knoll} Yu. B. Ivanov, J. Knoll, and D. N. Voskresensky, nucl-th/9905028.
956: \bibitem{Ehehalt}
957: W. Ehehalt and W. Cassing, Nucl. Phys. A 602 (1996) 449.
958: \bibitem{Teis} S. Teis, W. Cassing, M. Effenberger et al., Z. Phys. A356 (1997) 421.
959: \bibitem{Wolf} Gy. Wolf, W. Cassing and U. Mosel, Nucl. Phys. A 552 (1993) 549.
960: \bibitem{CGreiner} C. Greiner, K. Wagner and P.G. Reinhard,
961: Phys. Rev. C 49 (1994) 1693.
962: \end{thebibliography}
963: %
964: %
965: %----------------------------------------------------------------------
966: \begin{figure}[h]
967: {\vspace*{-1.0 cm} \hspace*{0.0 cm}
968: {\epsfig{file=q2_t.eps,height=20 cm,width=15 cm}}} \caption{The
969: quadrupole moment in momentum space (\protect\ref{quad}) for an
970: infinite nuclear matter problem in a finite box with periodic
971: boundary conditions characterized by bombarding energies of 0.1
972: A GeV and 1 A GeV at density $\rho = \rho_0$. The solid lines
973: present the results for the off-shell calculations while the
974: dashed lines are obtained in the on-shell limit.
975: \label{bild1}}
976: \end{figure}
977:
978: \begin{figure}[h]
979: {\vspace*{-1.0 cm} \hspace*{0.0 cm} {\epsfig{file=mt.eps,height=20
980: cm,width=15 cm}}} \caption{The transverse mass distribution for an
981: infinite nuclear matter problem at bombarding energies of 0.1 A
982: GeV (upper part) and 1 A GeV (lower part) at density $\rho =
983: \rho_0$. The solid lines present the results for the off-shell
984: calculations while the dashed lines are obtained in the on-shell
985: limit.
986: \label{bild2}}
987: \end{figure}
988:
989: \begin{figure}[h]
990: {\vspace*{-3.0 cm} \hspace*{0.0 cm}
991: {\epsfig{file=dndm.eps,height=20 cm,width=15 cm}}} \caption{The
992: differential distribution in mass for nucleons and $\Delta$
993: resonances at equilibrium for an infinite nuclear matter problem
994: at an initial bombarding energy of 1 A GeV at density $\rho =
995: \rho_0$. The solid histograms present the results from the
996: transport (box) calculations for the on-shell limit (upper part)
997: and off-shell limit (lower part), respectively. The dashed lines
998: ('thermo') are obtained from the statistical model at temperature $T= $ 97
999: MeV employing the spectral functions from the transport approach
1000: (see text).
1001: \label{bild3}}
1002: \end{figure}
1003: %
1004: \begin{figure}[h]
1005: {\vspace*{-1.0 cm} \hspace*{1.5 cm}
1006: {\epsfig{file=dndmdel.eps,height=15 cm,width=12 cm}}}
1007: \caption{Comparison of the differential distribution in mass for
1008: $\Delta$ resonances at equilibrium for the same infinite
1009: nuclear matter problem as in Fig. 3. The solid histogram presents
1010: the results from the transport calculations for the off-shell
1011: limit while the dashed histogram stems from the on-shell
1012: calculation.
1013: \label{bild4}}
1014: \end{figure}
1015: %
1016: \end{document}
1017: #!/bin/csh -f
1018: # Uuencoded gz-compressed .tar file created by csh script uufiles
1019: # For more info (11/95), see e.g. http://xxx.lanl.gov/faq/uufaq.html
1020: # If you are on a unix machine this file will unpack itself: strip
1021: # any mail header and call resulting file, e.g., figures.uu
1022: # (uudecode ignores these header lines and starts at begin line below)
1023: # Then say csh figures.uu
1024: # or explicitly execute the commands (generally more secure):
1025: # uudecode figures.uu ; gunzip figures.tar.gz ;
1026: # tar -xvf figures.tar
1027: # On some non-unix (e.g. VAX/VMS), first use editor to change filename
1028: # in "begin" line below to figures.tar-gz , then execute
1029: # uudecode figures.uu
1030: # gzip -d figures.tar-gz
1031: # tar -xvf figures.tar
1032: #
1033: uudecode $0
1034: chmod 644 figures.tar.gz
1035: gunzip -c figures.tar.gz | tar -xvf -
1036: rm $0 figures.tar.gz
1037: exit
1038:
1039: