hep-ph0404172/2d3.tex
1: \documentclass[prd,showkeys,showpacs]{revtex4}
2: %\documentclass[12pt,graphicx,floats]{article}
3: \usepackage{graphicx}
4: \renewcommand{\textfraction}{0}
5: \def\be{\begin{equation}}
6: \def\ee{\end{equation}}
7: \def\bea{\begin{eqnarray}}
8: \def\eea{\end{eqnarray}}
9: \def\Bphi{\mbox{\boldmath $\Phi$}}
10: \def\Bpi{\mbox{\boldmath $\pi$}}
11: \def\hphi{\mbox{\boldmath $\hat\Phi$}}
12: \def\Bx{\mbox{\boldmath $x$}}
13: \def\Bxi{\mbox{\boldmath $\xi$}}
14: \def\Bk{\mbox{\boldmath $k$}}
15: \def\Bp{\mbox{\boldmath $p$}}
16: \def\Bal{\mbox{\boldmath $\alpha$}}
17: \def\Bet{\mbox{\boldmath $\beta$}}
18: \def\Te{{\cal T}}
19: \def\t{\tau}
20: \def\z{\eta}
21: \def\dt{\partial_\tau}
22: \def\dz{\partial_\eta}
23: \def\dtt{\partial_{\tau\tau}}
24: \def\dzz{\partial_{\eta\eta}}
25: \begin{document}
26: %
27: \title{The topological approach to baryon-antibaryon and meson production in rapidly expanding Bjorken rods.}
28: %
29: \author{G. Holzwarth\footnote{e-mail: holzwarth@physik.uni-siegen.de}}
30: %
31: \address{Siegen University, 57068 Siegen, Germany} 
32: %
33: \vspace{3cm}
34: %
35: \begin{abstract}\noindent
36: 
37: The topological approach to baryon-antibaryon production in the chiral phase transition
38: is numerically simulated for rapidly expanding hadronic systems. For that purpose the
39: dynamics of the effective chiral field is implemented on a space - rapidity 
40: lattice. The essential features of evolutions from initial 'hot' configurations into 
41: final ensembles of (anti-)baryons embedded in the chiral condensate are studied in proper
42: time of comoving frames. Typical times for onset and completion of the roll-down and  
43: exponents for the growth of correlations are extracted. Meson and baryon-antibaryon yields 
44: are estimated. For standard assumptions about initial coherence lengths they are compatible 
45: with experimental results.  
46: \end{abstract} 
47: 
48: \pacs{11.10.Lm,11.27.+d,25.75.-q,64.60.Cn,75.40.Mg}
49:   
50: \keywords{ Chiral phase transition, topological textures, heavy-ion collisions, antibaryon production}
51: \maketitle
52: 
53: %\newpage
54: \section{Introduction}
55: 
56: The topological approach to baryon structure and dynamics in the framework of an effective 
57: action for mesonic chiral fields has achieved a number of
58: remarkable successes.  The soliton concept~\cite{Skyrme} for baryons provides an impressive
59: account of spectrum and properties of baryon resonances (essentially without numerous 
60: 'missing resonances')~\cite{Schwesinger}, with predictive power that recently has even led to
61: the first indications for pentaquarks~\cite{Nakano}. Model-independent
62: relations between T-matrix elements for meson-baryon scattering~\cite{Hayashi} and explicit results for specific 
63: channels are well supported by experimental data~\cite{Mattis}. 
64: The matrix element of the axial singlet current related to the spin content of the proton is naturally 
65: of the observed order of magnitude~\cite{Brodsky}.
66: The 'unexpected' behaviour recently found~\cite{Jones} in the ratio of electric and magnetic
67: proton form factors has been predicted in this approach long ago~\cite{Ho96}. 
68: The underlying chiral effective action is profoundly based on the
69: $1/N_c$-expansion of QCD~\cite{Witten}, preserving all  relevant symmetries. Efforts to
70: include next-to-leading order quantum corrections have brought substantial improvement 
71: as expected~\cite{Mouss}.  
72: 
73: The manifestations of a chiral phase transition pose another natural challenge for an effectice
74: theory with a ground state that is characterized by spontaneously broken symmetry.  
75: The possible formation of disoriented domains~\cite{Bjorken} during the growth of the chiral condensate
76: has been in the focus of interest for some time. But signatures in terms of anomalous multiplicity ratios 
77: for differently charged pions have not been observed~\cite{Bearden}, in accordance with theoretical 
78: conclusions~\cite{Gavin,HoKl02}. Anomalies in anti-baryon production were very early recognized
79: as possible signals for interesting dynamics~\cite{DeGrand} in that phase transition, and the concept 
80: to consider baryons as topological solitons in a chiral condensate should lead to quite definite 
81: expectations for this process. 
82: 
83: Meanwhile, in relativistic heavy-ion collisions at RHIC, very high energy densities are being 
84: produced in extended spatial regions which are essentially baryon free and well separated in rapidity from the 
85: nuclear slabs receding from the collision volume. 
86: The experimental values found in the central rapidity region for the ratio of the integrated
87: ${\bar p}$ to $\pi^-$ yields 
88: lies between 0.065 and 0.075 \cite{phenix}. This is still too close
89: to the thermal equilibrium ${\bar p}/\pi^-$ ratio (for a typical plasma temperature of $\Te \sim 200$ MeV),
90: \be
91: {\bar p}/\pi^- \sim 2 \exp((m_\pi-m_p)/\Te)=0.035
92: \ee
93:  to constitute a clear indication 
94: for interesting underlying physics. Still, although the experimental result does not look very 
95: exciting, it still poses a constraint for the possible validity of the soliton concept,
96: because any conceivable dynamical production process must be able to produce a comparable number.
97: 
98: In the topological approach the number of baryon-antibaryon pairs produced during the 
99: chiral phase transition depends on two factors: the first is the modulus $|\rho|$ of the average winding density
100: present in the initial 'hot' field configuration. In analogy to applications in cosmology~\cite{Kibble} and
101: condensed matter systems~\cite{Zurek} this quantity is closely related to the coherence length for the local
102: orientations of the chiral field $\Bphi$. Without detailed knowledge about the initial field configurations
103: this coherence length enters as a parameter and takes away stringent predictive power from the approach.
104: However, different conjectures about the nature of the initial field ensemble suggest typical ranges for the
105: coherence lengths which then may be discriminated by the experimentally observed abundancies.    
106: 
107: The second factor is the reduction of the initially present total $n_i=\int |\rho| dV$ through the dynamical 
108: ordering process, which finally leads to the formation of stable soliton structures embedded in the topologically
109: trivial ordered chiral condensate of the 'cold' system. The solitons or antisolitons evolve from
110: topological obstacles which are met by the aligning field orientations, and  develop into their
111: stable 'cold' form during the course of the evolution. At the end, the same integral $n_f=\int |\rho| dV$
112: counts the number of finally surviving nontrivial separate structures, so it is identified with the number of
113: baryons and antibaryons created in the process. The decrease of $n$ during the roll-down 
114: is reasonably well represented by
115: a power law $(\tau/\tau_0)^{-\gamma}$ and the exponent $\gamma$ can be measured in numerical simulations.
116: Evidently, the initial time $\tau_0$ which marks the onset of the evolution, enters here as a second
117: parameter which further reduces the predictive power of the approach. Fortunately, it turns out that $\gamma$
118: is rather small, so the dependence on $\tau_0$ is only weak. 
119: 
120: Measuring $\gamma$ and the time $\tau_f$ when the roll-down is completed, presents a typical task for
121: numerical simulations once the equation of motion (EOM) which governs the field evolutions is implemented
122: on a lattice. The underlying effectice chiral action is known from other applications, so no additional
123: parameters enter at this point. In condensed matter applications, a phase transition is generally driven
124: by an externally imposed quench, or by a dissipative term included in the EOM. In cosmology or in our
125: present heavy-ion application it is the rapid expansion of the hot volume which drives the cooling process. 
126: This expansion is efficiently implemented~\cite{Huang} by transforming to rapidity -
127: proper-time coordinates, i.e. by boosting to the local comoving frame. This is especially convenient if we consider
128: a system that expands only in one (longitudinal) direction with its transverse scales unchanged, (the Bjorken rod).
129: The resulting dilution of the longitudinal gradients drives the system towards its global minimum. However, as
130: there is no genuine dissipation in the system, the total energy approaches a constant which resides in the 
131: chiral fluctuations around the global minimum. Thus,  the simulations also allow to estimate pion- or sigma-
132: meson abundancies.
133: 
134: Naturally, before the field configurations can roll down towards the global minimum, the potential 
135: $V(\Phi^2,\Te)$ which underlies
136: the EOM must have changed from the 'hot' chirally symmetric form to its 'cold' symmetry-violating form.
137: But, during the early stages, the evolutions are dominated by local
138: aligning of the field orientation $\hphi$. During this phase the form of the potential is 
139: not important. So its time dependence can be replaced by a sudden quench where the 'hot' field 
140: configuration is exposed to the 'cold' potential $V(\Phi^2,\Te=0)$, from the outset at initial time $\tau_0$ . 
141: In the following, for definiteness we make use of this sudden quench approximation, (although the simulations,
142: of course, allow to study other cases as well).   
143: 
144: For the sake of simplicity we first discuss all relevant features for the case of the 2-dimensional 
145: $O(3)$-model, with only one spatial dimension transverse to the longitudinal rapidity coordinate.   
146: Except for computational complexity the extension to the 3-dimensional $O(4)$-field presents no essential
147: new features. The effective action, its transformation to the Bjorken frame, and the resulting EOM
148: are presented in section II. It is important for the choice of the initial ensemble of field configurations
149: that it allows in a convenient way to monitor the initial coherence lengths because they are the crucial
150: parameters for the final baryon-antibaryon multiplicities. We choose an isotropic Gaussian random ensemble of 
151: field fluctuations in momentum space which is characterized by a temperature-like parameter to be able to 
152: compare with other approaches. Of course, this is not necessary. In fact, even at initial time $\tau_0$
153: the longitudinally expanding
154: Bjorken rod need not be an isotropic system, and it may be physically justified to distinguish already in the
155: initial ensemble two different, longitudinal and transverse correlation lengths. This is easy to incorporate, 
156: but in section III we present initial conditions which are locally isotropic.
157: 
158: As discussed elsewhere~\cite{Ho03} stable solitons shrink in a spatially expanding frame. Therefore, lattice
159: implementations of their dynamics will necessarily involve lattice artifacts after some time. These are 
160: discussed in section IV. They can be isolated and subtracted from the physically interesting quantities.
161: 
162: In section V the essential features of typical evolutions are discussed. Estimates for the times of onset
163: and completion of the roll-down are obtained, and the dynamical exponents for the growth of correlation lengths and
164: decrease of defect number are established and compared. The spectrum of the fluctuations remaining after the
165: roll-down is considered and finally the mesonic and baryonic multiplicities are obtained. 
166: 
167: 
168: The extension to the physically interesting 3+1-dimensional $O(4)$-field is discussed in section VI. The 
169: topological generalization is well known, the additional transverse dimension is of little influence for the 
170: growth exponents. However, the coupling constants in the effective action here are related to physical
171: quantities, so they are known (except for some uncertainty concerning the $\sigma$-mass), and the results can
172: be compared with experimentally determined abundance ratios.
173: 
174: Of course, it would be desirable to obtain a very definite answer whether the topological approach to
175: antibaryon production in a chiral phase transition is validated or ruled out by the data. However, with our poor
176: knowledge about the initial conditions in the hot plasma after a heavy-ion collision, we cannot expect 
177: much more than allowed ranges for the relevant parameters, which hopefully overlap with standard ideas
178: about coherence lengths and formation times. 
179: 
180: 
181: 
182: \section{The effective action in the Bjorken frame.}
183: 
184: For simplicity we first discuss the 2+1 dimensional $O(3)$ model. It is defined in terms of the dimensionless 
185: $3$-component field $\Bphi = \Phi \hphi$ with unit-vector field $\hphi$, ($\hphi
186: \cdot \hphi =1$), and modulus ('bag'-)field $\Phi$,
187: with the following lagrangian density in $2+1$ dimensions $(x,z,t)$
188: 
189: \be\label{lag}
190: {\cal{L}} = f_\pi^2\left({\cal{L}}^{(2)}+{\cal{L}}^{(4)}+{\cal{L}}^{(0)}\right)
191: \ee
192: ($f_\pi^2$ is an overall constant of dimension [mass$^1$], so the physical fields $f_\pi\Bphi$ are of 
193: mass-dimension [mass$^{1/2}$]).
194: The second-order part ${\cal{L}}^{(2)}$ comprises the kinetic terms of the linear $\sigma$ model 
195: 
196: \be \label{L2}
197: {\cal{L}}^{(2)}=\frac{1}{2} \partial_\mu \Bphi \partial^\mu \Bphi,
198: \ee
199: ${\cal{L}}^{(4)}$ is the four-derivative 'Skyrme'-term (which involves only the unit-vector field $\hphi$)
200: defined in terms of the topological current $\rho_\mu$ 
201: \be\label{top}
202: \rho^\mu = \frac{1}{8 \pi} \epsilon^{\mu \nu \rho} \hphi \cdot
203: ( \partial_\nu \hphi \times \partial_\rho \hphi ) ,
204: \ee
205: (which satisfies $\partial_\mu \rho^\mu = 0$)
206: \be \label{L4}
207: {\cal{L}}^{(4)}=-\lambda\ell^2  \:\varrho_\mu\varrho^\mu=
208: -\frac{2\lambda\ell^2}{(8\pi)^2}\left[ (\partial_\mu \hphi \partial^\mu
209: \hphi)^2 -(\partial_\mu \hphi \partial_\nu \hphi)(\partial^\mu \hphi\partial^\nu \hphi) \right], 
210: \ee
211: and ${\cal{L}}^{(0)}$ contains the $\Phi^4$ potential and an explicit symmetry-breaker in 3-direction
212: \be
213: \label{pot}
214: {\cal{L}}^{(0)}=-V(\Bphi,\Te)=-\frac{1}{\ell^2}\left( \frac{\lambda}{4}\left( \Bphi^2-f(\Te)^2 \right)^2
215: - H \Phi_3 \right)  -\mbox{const.}
216: \ee
217: with dimensionless coupling constants $\lambda$ and $H$, and
218: \be\label{f2}
219: f^2(\Te)= f_0^2(\Te)-\frac{H}{\lambda f_0(\Te)}.
220: \ee
221:  This choice ensures that the global 
222: minimum of the potential $V(\Bphi,\Te)$ is always located at $\Bphi_0=(0,0,f_0(\Te))$. Generically, the function 
223: $f_0^2(\Te)$ decreases from $f_0^2=1$ at $\Te=0$ towards zero for large $\Te$.  
224: The constant in the potential (\ref{pot})
225: is chosen such that the value of the potential $V$ at $\Bphi=0$ is independent of $\Te$, (given by the constant
226: $V(0,\Te)=(\lambda +2H)/(4\ell^2)$), 
227: and at the $(\Te=0)$-minimum $\Bphi=\Bphi_0=(0,0,1)$ we have $V(\Bphi_0,\Te=0)=0$. 
228: 
229: The masses of the $\pi$- and $\sigma$-fluctuations $(\pi_1,\pi_2, f_0+\sigma)$ around this minimum  are
230: \be\label{masses}
231: m_\pi^2=\frac{H}{\ell^2 f_0},~~~~~~~~~~~~m_\sigma^2=\frac{2\lambda f_0^2}{\ell^2} + m_\pi^2.
232: \ee 
233: Without explicit symmetry breaking, $H=0$, we assume that $f^2(\Te)$ changes sign at $\Te=\Te_c$, such that
234: $\Bphi_0=(0,0,0)$  and $m_\sigma^2=m_\pi^2=m^2=\lambda |f^2| / \ell^2$ for $\Te>\Te_c$. 
235: 
236: The parameter $\ell$ (with dimension of a length) which we have separated out from the coupling constants of
237: potential and Skyrme term
238: can be absorbed into the spatial coordinates $\Bx$. So it characterizes the spatial radius 
239: of stable extended solutions (which scales like $1/\sqrt{f^2}$). As $\ell$ simply sets the spatial scale,
240: it could be put equal to one, as long as no other (physical or artificial) length  scales are relevant. For
241: lattice implementations, however, the lattice constant $a$ and the size of the lattice $(Na)$
242: set (usually unphysical) scales. 
243: To avoid artificial scaling violations we have to ensure that the size of physical structures (like solitons) 
244: is large as compared to the lattice constant $a$ and small as compared to the lattice size $Na$. So,
245: for numerical simulations we have to choose $1\ll \ell/a \ll N$. 
246: It has been shown in ref.\cite{HoKl01} that for solitons which extend over more than at least 4-5 lattice units the 
247: energy $E_B$ is independent of $\ell/a$.
248: So, in the following we will adopt $\ell/a\sim 5$ as sufficiently large. This appears also as 
249: physically reasonable, if we consider
250: typical lattice constants of 0.2 fm and baryon radii of about 1 fm. On the other hand this will require lattice
251: sizes of at least $N \sim 50$ to avoid boundary effects for the structure of individual solitons.
252: Unfortunately, in the Bjorken frame which we shall use in the following, the longitudinal extension of
253: stable solitons shrinks like $\tau^{-1}$ as function of proper time $\tau$. 
254: This means that after times of order $\ell$ the simulations will be
255: influenced by lattice artifacts, which may even dominate for large times. 
256: 
257: 
258: For rapid expansion in (longitudinal) $z$-direction we perform the transformation from $(z,t)$ to
259: locally comoving frames ($\eta$,$\tau$) with proper time $\tau$ and rapidity $\eta$, defined through
260: \begin{eqnarray}\label{trafo}
261: t&=&\tau \cosh\eta,~~~~~~~~~~\tau~=~\sqrt{t^2-z^2},~~~~~~~~~~~~~
262: \partial_t~=~\cosh\eta\; \dt -\frac{\sinh\eta}{\tau}\dz \nonumber\\
263: z&=&\tau \sinh\eta,~~~~~~~~~~\eta~=~\mbox{atanh}\left(\frac{z}{t}\right),~~~~~~~~~~~~
264: \partial_z~=~-\sinh\eta\; \dt +\frac{\cosh\eta}{\tau}\dz.
265: \end{eqnarray}
266: Inserting (\ref{trafo}) into (\ref{L2}) and (\ref{L4}) 
267: leaves the form of ${\cal{L}}^{(2)}$ and ${\cal{L}}^{(4)}$ invariant, with $\partial_t$ replaced by
268: $\dt$, and $\partial_z$ replaced by $\frac{1}{\tau}\dz$. 
269: The specific structure of the Skyrme term again eliminates all terms with four $\tau$- or $\eta$-derivatives. 
270: For the effective action we take the integration boundaries from $-\infty$ to $+\infty$ 
271: for rapidity $\eta$ and for the transverse coordinate $x$.
272: The 3-dimensional space-time volume element $dx\:dz\:dt$ is replaced by $\tau \:dx \:d\eta \:d\tau$. 
273: Therefore, in a separation
274: of the action ${\cal S}$ in kinetic terms $T$, gradient terms $L$, and the potential $U$, 
275: \be \label{S}
276: {\cal S}=\int d\tau \int_{-\infty}^{+\infty}\!{\cal L} \;d\eta \:dx  
277: =\int(T_{\bot}+T_{\|}-L_{\bot}-L_{\|}-U) d\tau
278: \ee
279: the longitudinal $\|$-terms involving rapidity gradients carry a factor $1/\tau$, while all other 
280: terms carry a factor $\tau$. 
281: So we have
282: \bea
283: T_{\bot}&=&\tau \int \left\{\frac{1}{2} (\dt\Bphi \dt\Bphi)
284: +\frac{\lambda\ell^2}{(4\pi)^2}\left[ \frac{\Bphi}{\Phi^3}\cdot(\dt \Bphi\times
285: \partial_x \Bphi) \right]^2\right\}\;d\eta \:dx,\label{int1} \\
286: T_{\|}&=&\frac{1}{\tau} \int \left\{
287: \frac{\lambda\ell^2}{(4\pi)^2}\left[ \frac{\Bphi}{\Phi^3}\cdot(\dt \Bphi\times
288: \dz \Bphi)\right]^2\right\}\;d\eta \:dx,\label{int2}\\
289: L_{\bot}&=&\tau \int \left\{\frac{1}{2} (\partial_x\Bphi \partial_x\Bphi)
290: \right\}\;d\eta \:dx,\label{int3}\\
291: L_{\|}&=&\frac{1}{\tau} \int \left\{\frac{1}{2} (\dz\Bphi \dz\Bphi)
292: +\frac{\lambda\ell^2}{(4\pi)^2}\left[ \frac{\Bphi}{\Phi^3}\cdot(\dz \Bphi\times
293: \partial_x \Bphi) \right]^2\right\}\;d\eta \:dx, \label{int4}\\
294: U&=&\tau \int \left\{\frac{\lambda}{4\ell^2}\left( \Bphi^2-f^2 \right)^2
295: - \frac{H}{\ell^2} \Phi_3 +\mbox{const.}\right\}\;d\eta \:dx.  \label{int5}
296: \eea
297: Variation of ${\cal S}$ with respect to $\Bphi$ leads to the equation of motion (EOM).
298: 
299: The contributions of ${\cal{L}}^{(4)}$ to the longitudinal and transverse parts $T_{\|}$ and $T_{\bot}$
300: of the kinetic energy cause certain numerical difficulties for the implementation of the EOM on a lattice.
301: They require at every timestep the inversion of matrices which depend on gradients of the unit-vectors $\hphi$,
302: which multiply first and second time-derivatives of the chiral field. This can be troublesome in areas
303: where the unit-vectors are aligned, and can be poorly defined in regions where the unit-vectors vary
304: almost randomly for next-neighbour lattice points (i.e. for initially random configurations, or near the center
305: of defects). In any case, stabilizing the evolutions requires extremely small timesteps and leads to
306: very time-consuming procedures. Although these problems can be handled, we have compared the results
307: with evolutions where the kinetic energy is taken from ${\cal{L}}^{(2)}$ alone.
308: For coupling strenghths $\lambda\ell^2$ within reasonable limits,
309: we find that the resulting differences do not justify the large additional expense caused by the fourth-order
310: kinetic contributions. Evidently, the reason is, that the EOM determines the field-velocities (depending on
311: the functional form of the kinetic energy) in such a way that the numerical value of the total kinetic energy
312: is not very sensitive to its functional form. We therefore use in the following an effective action where
313: the kinetic terms (\ref{int1}) and (\ref{int2}) are replaced by
314: \be\label{Tkin}
315: T_{\bot}=\frac{\tau}{2}\int(\dt\Bphi \dt\Bphi)\,d\eta \,dx,~~~~~~~~~~~~~~~T_{\|}=0.
316: \ee 
317: With this simplification the EOM is 
318: \be \label{EOM}
319: \frac{1}{\tau}\partial_\tau\Bphi+\partial_{\tau\tau}\Bphi-\partial_{xx}\Bphi-
320: \frac{1}{\tau^2}\partial_{\eta\eta}\Bphi+\frac{\lambda}{\ell^2}(\Phi^2-f^2)\Bphi
321: -\frac{H}{\ell^2} \hat{\mbox{\boldmath $e$}}_3 +\frac{\lambda\ell^2}{\tau^2}\frac{\delta \rho_0^2}{\delta\Bphi}=0.
322: \ee
323: This form has the big advantage that we can make use of the geometrical meaning of the winding density
324: $\rho_0$ as the area of a spherical triangle, bounded by three geodesics on a 2-dimensional spherical surface.
325: In closed form it is expressed through the unit-vectors pointing to its corners, and does not involve gradients.
326: So this allows for a very accurate and fast lattice implementation of the last term in the EOM.
327: 
328: \section{Initial configurations} 
329: 
330: We assume that at an initial proper time $\tau_0$ the system consists of a hadronic fireball 
331: with energy density $\varepsilon_0$ stored in a random ensemble
332: of hadronic field fluctuations. Subsequently, for $\tau > \tau_0$, it is subject to EOM (\ref{EOM}). 
333: The initial condition and the symmetry of the action imply
334: boost invariance, i.e. the system looks the same in all locally comoving frames, so it is 
335: sufficient to consider its dynamics in a rapidity slice of size $\Delta\eta$ near midrapidity $\eta=0$,
336: which constitutes a section of the initially created Bjorken rod with transverse extension 
337: ${\cal A}$. The energy $E=T+L+U$ in this slice then is given by an  $\eta$-integral 
338: which extends over the finite rapidity interval $\Delta\eta$ and represents the energy 
339: contained in a comoving volume ${\cal V}= \tau \Delta\eta {\cal A}$.
340: Due to the symmetry of the initial condition this comoving volume grows with increasing proper time $\tau$ into
341: spatial regions with high energy density, therefore $E$ contains contributions which increase with $\tau$. 
342: The average energy density $\varepsilon=E/{\cal V}$ satisfies $d\varepsilon/d\tau \leq 0$.
343: 
344: 
345: 
346: For numerical simulations we implement the configurations $\Bphi(x,\eta,\tau)$ on a rectangular 
347: lattice $(x,\eta) = (ia,jb)$ ($i,j=1...N$) with lattice constants $a$ for the transverse coordinate and $b$ for
348: the rapidity lattice.
349: We define the initial configurations $\Bphi_{ij}$  at the lattice sites ($i,j$) as Fourier transforms
350: of configurations $\tilde {\Bphi}_{kl}$ on a momentum lattice 
351: \be  \label{Fourier}
352: \Bphi_{ij}=\frac{1}{N}\sum_{k,l=-N/2+1}^{N/2} 
353: \frac{1}{2}\left( e\:^{\mbox{\small i}\frac{2\pi}{N}(i\cdot k +j\cdot l)} 
354: \tilde{\Bphi}_{kl}+ c.c.\right),
355: \ee
356: with $\tilde{\Bphi}^*_{kl}=\tilde{\Bphi}_{-k-l}$. Inversely, the real parts $\Bal_{kl}$ and the imaginary parts 
357: $\Bet_{kl}$ of $\tilde {\Bphi}_{kl}$ are obtained from the real configuration
358: $\Bphi_{ij}$ through
359: \bea \label{albet}
360: \Bal_{kl}& =&\frac{1}{N}\sum_{i,j=1}^{N}\cos\frac{2\pi}{N}(i k +j l)\;\Bphi_{ij}=\Bal_{-k-l}\\
361: \Bet_{kl} &=&-\frac{1}{N}\sum_{i,j=1}^{N}\sin\frac{2\pi}{N}(i k +j l)\;\Bphi_{ij}=-\Bet_{-k-l},
362: \eea
363: so we obtain the spectral power $P_{pq}$ of the configurations (or a specific component of it) 
364: at any time $\tau$ from 
365: \be
366: P_{pq}=\tilde {\Bphi}_{kl}\cdot \tilde {\Bphi}_{kl}^*=\Bal_{kl}\cdot\Bal_{kl}+\Bet_{kl}\cdot\Bet_{kl}
367: \ee
368: for any transverse or longitudinal momentum $(p,q)= \frac{2\pi}{aN}(k,l)$, 
369: for $(k,l = -N/2+1,..., N/2)$. 
370: 
371: For the initial configurations at $\tau=\tau_0$ the real and imaginary parts of each of the three 
372: components of $\tilde {\Bphi}_{kl}$  
373: at each momentum-lattice point ($p,q$) are chosen 
374: randomly from a Gaussian deviate $G_{kl}(\tilde{\Phi})$ with $kl$-dependent width $\sigma_{kl}$,
375: \be \label{Gauss}
376: G_{kl}(\tilde{\Phi})=\frac{1}{\sqrt{2\pi \sigma_{kl}^2}} \exp\left(-\frac{\tilde\Phi^2}{2\sigma_{kl}^2}\right),
377: ~~~~\mbox{with}~~~~~
378: %\sigma_{kl}^2=\frac{2\sigma_0^2}{\pi m \sqrt{\Te_\|\Te_\bot}}
379: %\left(\frac{2\pi}{a}\right)\left(\frac{2\pi}{b\tau_0}\right)
380: %\exp\left(-\frac{p^2}{2 m \Te_\bot}-\frac{q^2}{2 m \Te_\|}\right),
381: %\ee
382: %(or, relativistically,
383: %\be
384: \sigma_{kl}^2=\frac{\sigma_0^2}{Z}\exp\left(-\frac{\sqrt {p^2+q^2+m^2}}\Te\right),
385: \ee
386: with normalization $Z$ chosen in such a way that
387: \be\label{modsum}
388: \sum_{k,l=-N/2+1}^{N/2}\sigma_{kl}^2 = N^2\sigma_0^2.
389: \ee
390: (In the continuum limit $(a\rightarrow 0$, $N\rightarrow\infty)$ we have 
391: $Z=\frac{\Te^2}{2\pi}\left(1+\frac{m}{\Te}\right)e^{-m/\Te}.)$
392: 
393: In other words, we choose a Boltzmann distribution for the average occupation numbers 
394: $n_{kl} =\langle\langle \tilde\Phi_{kl}\tilde\Phi_{kl}^*\rangle\rangle
395:  = \sigma_{kl}^2$ for each field component, as for relativistic (non-interacting)
396: particles with mass $m$. Here the mass $m^2$ is defined by the absolute value 
397: \be\label{mass}
398: m^2(\Te)=\frac{\lambda}{\ell^2} |f^2(\Te)|
399: \ee 
400: for the fluctuations around $\Bphi=0$ in the symmetric potential (\ref{pot}) at the initially high
401: temperature $\Te=\Te_0$, where $f^2(\Te)$ is negative. 
402: The amplitude $\sigma_0^2$ plays the role of a fugacity
403: \be
404: \sigma_0^2=\exp(-\mu/\Te)
405: \ee
406: for negative chemical potential $\mu$. In the temperature range which we consider $(0.05<a\Te<0.8)$
407: (cf. fig.(\ref{fig3})) a suitable value for $\mu$ is $a\mu\sim -0.6$. (With this choice the 
408: average amplitude of the chiral field is not subject to abrupt deviations from its initial value 
409: immediately after the onset of the dynamical evolution).  
410: 
411: We assume isotropy of the initial ensemble with respect to rotations in $O(3)$-space such that
412: the three components of the field fluctuations $\tilde\Phi^\alpha_{kl}$ ($\alpha=1,2,3$)
413: have the same average square amplitude $\sigma_{kl}^2$. By picking each component independently
414: at each point $(k,l)$ from the Gaussian ensemble, different components are uncorrelated and equal components 
415: at different points (on the momentum lattice) are also uncorrelated,   
416: \bea
417: \langle\langle \tilde\Phi^\alpha_{kl} \tilde\Phi^{\beta *}_{k'l'} \rangle\rangle &=&
418: \langle\langle \alpha^\alpha_{kl} \alpha^\beta_{k'l'} \rangle\rangle +
419: \langle\langle \beta^\alpha_{kl} \beta^\beta_{k'l'} \rangle\rangle \nonumber \\
420: &=&\sigma_{kl}^2 \delta_{\alpha\beta}
421: \left( \frac{1}{2}(\delta_{k k'}\delta_{l l'} +\delta_{-k k'}\delta_{-l l'})
422: +\frac{1}{2}(\delta_{k k'}\delta_{l l'} -\delta_{-k k'}\delta_{-l l'})\right) 
423: =\sigma_{kl}^2 \delta_{\alpha\beta}\delta_{k k'}\delta_{l l'}.
424: \eea
425: Together with (\ref{Fourier}) this leads to the fluctuation in the real field configurations
426: \be \label{fluc}
427: \langle\langle \Phi^\alpha_{ij}\Phi^{\beta }_{ij} \rangle\rangle 
428: =\delta_{\alpha\beta}\frac{1}{N^2}\sum_{k,l=-N/2+1}^{N/2} \!\!\! \sigma_{kl}^2
429: = \delta_{\alpha\beta} \; \sigma_0^2
430: \ee
431: which is, of course, independent of the lattice point $(i,j)$. Its magnitude is controlled by the constant
432: $\sigma_0^2$ in (\ref{Gauss}). It should be sufficiently small to keep the amplitudes of the average
433: initial fluctuations small. 
434: On the lattice the upper limit for the momenta $p$,$q$ is $\frac{\pi}{a}$, (i.e. $k,l=N/2$). So, as long as 
435: \be \label{LamT}
436: \Te \ll \frac{\pi}{a},
437: \ee
438: the lattice cut-off (upper limit momentum) imposed by the finite lattice constant 
439: is unimportant because the corresponding states are almost unoccupied. 
440: Note that periodicity and antisymmetry of the imaginary parts  in (\ref{albet}) requires
441: that  $\Bet_{kl}$ vanishes if both $k$ and $l$ are multiples of $N/2$. With the condition
442: (\ref{LamT}) satisfied, this holds with good accuracy also for the initial configuration picked 
443: randomly from the ensemble (\ref{Gauss}). 
444:  
445:  The average number of topological defects in a random ensemble of vector configurations
446:  is closely related to the characteristic angular coherence  length  in that ensemble.
447:  Therefore, it will be necessary to measure the (equal-time) correlation functions
448:  for the unit-vector fields $\hphi$ for the evolving ensembles.
449:  In order to have an analytical result at least for the initial configurations
450:  (where length and orientation of the 3-vectors are uncorrelated), it is easier to consider
451:  the correlations among the full vectors $\Bphi$.
452: Therefore, we define normalized transverse and longitudinal correlation functions
453:  \bea \label{corrfct}
454:  C_\bot(i)&=&\frac{1}{3\sigma_0^2N^2}
455:  \left[\langle\langle\sum_{m,n=1}^{N}\Bphi_{mn}\cdot\Bphi_{m+i,n}\;\rangle\rangle
456:  -\frac{1}{N^2}\langle\langle\sum_{m,n=1}^{N}\Bphi_{mn}\;\rangle\rangle \cdot 
457:  \langle\langle\sum_{k,l=1}^{N}\Bphi_{kl}\;\rangle\rangle\right],\nonumber \\
458:  C_\|(i)&=&\frac{1}{3\sigma_0^2N^2}
459:  \left[\langle\langle\sum_{m,n=1}^{N}\Bphi_{mn}\cdot\Bphi_{m,n+i}\;\rangle\rangle
460:  -\frac{1}{N^2}\langle\langle\sum_{m,n=1}^{N}\Bphi_{mn}\;\rangle\rangle \cdot 
461:  \langle\langle\sum_{k,l=1}^{N}\Bphi_{kl}\;\rangle\rangle\right],
462:  \eea
463:  with transverse coherence  lengths $R_\bot$ and longitudinal (dimensionless) coherence  rapidity $R_\|$
464:  defined through
465:  \be \label{cut}
466:  C(i)< \frac{1}{e}~~~~~~~~~~\mbox{for}~~~~ i > \frac{R_\bot}{a},~~~~\mbox { or}~~~ i > \frac{R_\|}{b}
467:  \ee
468:  respectively.
469:  For the initial ensemble (\ref{Gauss}) the correlations
470:  are, of course, isotropic on the lattice, i.e. 
471:  \be \label{initcorr}
472:  \frac{R_\bot}{a}=\frac{R_\|}{b}=\frac{R_0}{a}
473:  \ee
474:  with initial spatial coherence  length $R_0$.
475:  In the continuum limit $(a\rightarrow 0$, $N\rightarrow\infty)$, we obtain
476:  $C(r)$ as function of the spatial distance $r$ (or rapidity $\eta=r (b/a)$)
477:  \be  \label{corrinit}
478:  C(r)=
479:  \frac{e^{-\frac{m}{\Te}\left(\sqrt{1+r^2\Te^2}-1\right)}}{\left(1+r^2\Te^2\right)^{3/2}}
480:  \left(\frac{1+\frac{m}{\Te}\sqrt{1+r^2\Te^2}}{1+\frac{m}{\Te}}\right).
481:  \ee
482:  Specifically, putting $m=0$, the coherence  length $R$ as defined in (\ref{cut}) is  
483:  \be \label{Rm0}
484:  R=\frac{\sqrt{e^{2/3}-1}}{\Te} \approx \frac{0.97}{\Te}. 
485:  \ee
486:  This allows to put limits on the range of temperatures which can be reasonably represented
487:  on the lattice. Typically, for lattice size of $N\sim 100$, $\Te$ should lie within the range  
488:  from about  0.02 to about 0.8 inverse lattice units. For smaller values the inital coherence  
489:  length already covers more than half of the lattice so almost no defects will fit on the lattice,
490:  for larger values the correlation lengths approach the lattice constant.
491:   It may be noted that with (\ref{mass}), for $(\ell/a)\sim 5$ and $(a\Te)\sim 0.1$, 
492:   the ratio $m/\Te$ is not very small, so generally we expect appreciable deviations from the 
493:  $\Te^{-1}$ scaling in (\ref{Rm0}), (e.g. for $(\ell/a)=4$ we find $(R/a)\sim (a\Te)^{-0.8}$, 
494:  cf. fig.\ref{fig3}).
495:  
496:  During the evolution in the Bjorken frame the correlations rapidly become anisotropic. 
497:  We then conveniently define an average coherence  
498:  length $\bar R$ through
499:  \be \label{avcorr}
500:  \frac{a^2}{{\bar R}^2}=\frac{1}{2}\left( \frac{a^2}{R_\bot^2}+\frac{b^2}{R_\|^2}\right).
501:  \ee 
502: This may be compared to the coherence  radius obtained from the angular-averaged correlation function 
503: \be \label{avcorrfct}
504: \bar C(r)
505: =\frac{1}{3\sigma_0^2N^2}
506: \left[\langle\langle\sum_{m,n=1}^{N}\sum_{i,j}\Bphi_{mn}\cdot\Bphi_{m+i,n+j}\;\rangle\rangle
507: -\frac{1}{N^2}\langle\langle\sum_{m,n=1}^{N}\Bphi_{mn}\;\rangle\rangle \cdot 
508:  \langle\langle\sum_{k,l=1}^{N}\Bphi_{kl}\;\rangle\rangle\right]
509: \ee 
510: where the $i,j$-sum indicates an average over all lattice points in a narrow circular ring with radius $r$ 
511: around the lattice point $mn$.
512: 
513: The essential characteristics of the evolutions are not very sensitive to the choice of the 
514: initial time derivatives. (They can as well be put to zero.) The equations of motion very quickly establish
515: appropriate velocities. Of course, the absolute value of the total energy depends on that choice.
516: For the simulations presented in the following we construct in analogy to the initial 
517: configurations (\ref{Fourier}) an initial ensemble of time derivatives 
518: through
519: \be
520: (\partial_\tau\Bphi)_{ij}=\frac{\mbox{i}}{N}\sum_{k,l=-N/2+1}^{N/2} \frac{\omega_{kl}}{2} 
521: \left( e\:^{\mbox{\small i}\frac{2\pi}{N}(i\cdot k +j\cdot l)} \tilde{\Bphi}_{kl}- c.c.\right),
522: ~~~~~~\mbox{with}~~~~~
523: \omega_{kl}=\sqrt{\left(\frac{2\pi}{aN}k\right)^2+\left(\frac{2\pi}{a N}l\right)^2 + m^2}.
524: \ee 
525: The Fourier coefficients $\tilde{\Bphi}_{kl} $ again are picked randomly from 
526: the same Gaussian deviate (\ref{Gauss}).
527: 
528: 
529: \section {Shrinking solitons in comoving frames}
530: Let $\Bphi^{(s)}(x,z)$ be a static soliton solution of the model (\ref{lag}) in its $(x,z)$ rest 
531: frame, which minimizes the static energy $E=L+U$ with a finite value for the 
532: soliton energy $E=E_0$. 
533: After the transformation to the Bjorken frame, the configuration 
534: $\Bphi_\tau^{(s)}(x,\eta)=\Bphi^{(s)}(x,\tau\eta)$ then describes a static solution of the action in the
535: comoving $(x,\eta)$ frame at proper time $\tau$ (where $\partial_z$ is replaced by $(1/\tau) \partial_\eta$),
536: for the same value of $E_0$. It represents a soliton with the same finite radius 
537: in transverse $x$-direction as before,
538: but with its radius in longitudinal $\eta$-direction shrinking like $1/\tau$ with increasing proper time $\tau$.
539: The total energy $E_0$ of this shrinking soliton is, of course, independent of $\tau$.
540: (Naturally, this consideration strictly applies 
541: only to the adiabatic case, where $\tau$ is considered as a parameter. In the dynamical ordering process 
542: the evolution of the solitons towards their static form may appreciably lag behind the actual progress 
543: of proper time.)
544: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
545: \begin{figure}[h]
546: \centering
547: \includegraphics[width=9.5cm,height=13cm,angle=-90]{fig1a.ps}
548: \includegraphics[width=9.5cm,height=13cm,angle=-90]{fig1b.ps}
549: \caption{Soliton configuration after a typical evolution on a 50$\times$50 lattice (for $\lambda=1, 
550: \ell/a=4, \sigma_0=0.2, H=0.2, a\Te=0.2$) at time $\tau/\tau_0=1000$, i.e. long after completion
551: of the roll-down. The bag-field $|\Bphi|$ of the solitons (upper part) is sqeezed longitudinally to 
552: lattice-unit size; the positive or negative winding densities (lower part) are located at the 
553: center of the bags.} 
554: \label{fig1}
555: \end{figure}
556: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
557: For lattice implementations,
558: with the typical spatial radius of the stable solitons given by $\ell$,  
559: the longitudinal extension of the solitons for times $\tau \gg \ell $ has shrunk
560: down to (dimensionless rapidity-)lattice-unit size and longitudinally adjacent solitons no longer interact. 
561: In transverse direction, however, the solitons develop their stable size of $\ell/a$ lattice units,
562: they keep interacting, attracting close neighbours or annihilating with overlapping antisolitons
563: (cf. fig.\ref{fig1}).
564: 
565: 
566: For solitons shrinking longitudinally down to lattice-unit size the energy will begin to deviate 
567: from the value $E_0$ as soon as the longitudinal extent covers merely a few lattice units.
568: To get an approximate idea for the energy limit 
569: let us assume that a single separate soliton finally degenerates into a transverse string of $2\ell+1$
570: lattice points, on which $|\Bphi|$ varies from nearly zero (in its center) to the surrounding vacuum 
571: configuration $\Bphi_0=(0,0,1)$, i.e.   $\Bphi=(0,0, |i|/\ell)$ for $-\ell\leq i\leq\ell$, on that string
572: of lattice points. Then we find for the contributions of a single soliton
573: \be\label{shrilat}
574: L^{(2)}_\bot\sim\frac{\tau}{\ell},~~~~~~~~~~ L^{(2)}_\|\sim\frac{\ell}{\tau},
575:  ~~~~~~~~~~U\sim\lambda\frac{\tau}{\ell}.
576: \ee
577: So, apparently, solitons shrinking on a lattice contribute to the energy terms which rise linearly with
578: proper time $\tau$ which (as lattice artifact) will dominate the total energy for large $\tau$.
579: 
580: We expect the winding density of the squeezed defect to be located on $\nu$ lattice squares 
581: near its center. This implies for the fourth-order term
582: \be\label{e4lim}
583: L^{(4)}_\|=\frac{\lambda\ell^2/\nu}{\tau}.
584: \ee
585: The winding density is determined by the orientation of the field unit-vectors alone, so it sufficient
586: to consider the unit-vectors $\hphi$. We expect
587: the squeezed defect to consist of just one unit-vector $\hphi=(0,0,-1)$ at the soliton center  
588: looking into the direction opposite to all surrounding unit-vectors $(0,0,1)$. That lattice point is the
589: top of four adjacent rectangular triangles (with the diagonals connecting the four nearest neighbour points as 
590: bases) which together cover an area of two lattice squares. So we expect
591: a winding density $\rho=1/\nu$ with $\nu\sim 2$. 
592:  
593: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
594: \begin{figure}[h]
595: \centering
596: \includegraphics[width=11cm,height=17cm,angle=-90]{fig2.ps}
597: \caption{\label{fig2} Potential energy $U$, kinetic energy $T^{(2)}$, transverse and longitudinal (second-order)
598: gradient terms $L_\bot^{(2)}$ and $L_{||}^{(2)}$, the number of defects $n$, and the average length of the 
599: chiral field $|\Bphi|$, for a typical evolution after a sudden quench (for $\lambda=1, 
600: \ell/a=4, \sigma_0=0.2, H=0.2, a\Te=0.2, N=50$). For comparison, the straight lines given by eqs.(\ref{linear}) 
601: with  (\ref{C2}) are included.} 
602: \end{figure}
603: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
604: 
605: This dominance of lattice artifacts for $\tau \gg \ell $ is illustrated in fig.(\ref{fig2})
606: which shows a typical evolution on a $50\times 50$ lattice for $\ell=4$.
607: The total winding number is $B=-1$. After the roll-down the number of solitons stabilizes at $n=9$
608: (cf. fig.\ref{fig1}).
609: Apparently, both $U$ and $L^{(2)}_\bot$,  approach a linearly rising limit for $\tau \gg 10^2$,
610: approximately like $\sim \frac{1}{2}n (\tau/\ell)$ which dominates the total energy, but does not
611: affect the (essentially constant) kinetic energy. Longitudinal contributions drop off like $\tau^{-1}$,
612: so they are irrelevant. 
613: 
614: It appears from fig.(\ref{fig2}) that for this evolution the roll-down (where the average of $\Phi$
615: aproaches the vacuum value $\Phi=1$) takes place during the time interval $2\ell <\tau < 4 \ell$,
616: i.e. long before artificial lattice effects dominate the energy. It is also by the end of the roll-down
617: that the number of created defects stabilizes. So we would conclude that results
618: obtained from lattice simulations for baryon-antibaryon production during the chiral phase transition 
619: in a rapidly expanding chiral gas are not severely affected by lattice artifacts.
620: On the other hand, to follow the evolutions beyond the end of the roll-down, which comprise
621: small '$\sigma$' and '$\pi$'-oscillations of $\Bphi$ around the true vacuum, interfering with small oscillations
622: of the bag profiles (resonances), will require to subtract the lattice artifacts.    
623: 
624: \section{Evolution until freeze-out}
625: 
626: In this chapter we will follow typical evolutions of the chiral field after a sudden quench 
627: in more detail and try to analyze their characteristic features up to the end of the roll-down.
628: 
629: Immediately before the sudden quench at $\tau=\tau_0$ the initial ensemble is prepared as described
630: in section II. The average length of one component of the chiral field is given by $\sigma_0$ (cf. eq. (\ref{fluc})),
631: the potential in (\ref{pot}) is characterized by a negative value of $f^2$. 
632: So, for sufficiently small $\sigma_0^2$ we have at $\tau=\tau_0$
633: \be \label{Uth0}
634: U_0= \tau_0\frac{\lambda}{4\ell^2} \int\left(f^4+2 |f^2|\langle\langle\Bphi^2\rangle\rangle\right) 
635: dx\:d\eta =(C_0+C_2){\cal V}_0,
636: \ee
637: where ${\cal V}_0=\tau_0\Delta\eta {\cal A}$ is the inital volume of the Bjorken slice, and the constants are
638: \be\label{const1}
639: C_0=\frac{\lambda}{4\ell^2}f^4, ~~~~~~~~~~~~C_2=\frac{3\lambda}{2\ell^2}|f^2| \sigma_0^2.
640: \ee
641: For the derivatives at lattice points $(i,j)$ Eq.(\ref{Fourier}) implies
642: \be  \label{deriv}
643: (\partial_x\Bphi)_{ij}=\frac{\mbox{i}}{N}\sum_{k,l=-N/2+1}^{N/2} \left(\frac{2\pi k}{aN}\right)
644: e\:^{\mbox{\small i}\frac{2\pi}{N}(i\cdot k +j\cdot l)} \tilde{\Bphi}_{kl},
645: ~~~~~~(\partial_\eta\Bphi)_{ij}=\frac{\mbox{i}}{N}\sum_{k,l=-N/2+1}^{N/2} \left(\frac{2\pi l}{bN}
646: \right) e\:^{\mbox{\small i}\frac{2\pi}{N}(i\cdot k +j\cdot l)} \tilde{\Bphi}_{kl}.
647: \ee
648: Again replacing the integrands in (\ref{int3}),(\ref{int4}) by ensemble averages leads to
649: the second-order gradients contribution at $\tau=\tau_0$
650: \be \label{L20}
651: L_0^{(2)}=C^{(2)}\;{\cal V}_0.
652: \ee
653: For the constant $C^{(2)}$ we have in the continuum limit
654: \be \label{C2}
655: C^{(2)}=\frac{9}{2}\sigma_0^2 \Te^2 \left(1+\frac{m^2}{3\Te^2(1+\frac{m}{\Te})}\right).
656: \ee
657: Similarly, one may obtain a rough estimate for $L^{(4)}$ averaged over the initial ensemble
658: by replacing in (\ref{L4}) the unitvectors $\hphi$ by $\Bphi / \sigma_0$.  
659: 
660: 
661: During the very early phase of an evolution in proper time 
662: the initially random ensemble of fluctuations will essentially stay random.
663: This means that the integrals in (\ref{int3})-(\ref{int5}) 
664: will remain constant, given by their initial values. 
665: Therefore, the time dependence of the different 
666: contributions (\ref{int3})-(\ref{int5}) to the total energy is given by the kinematical factors
667: ($\tau/\tau_0$) or ($\tau_0/\tau$) alone, with the integrals approximated
668: by replacing the integrands through their averages in the initial ensemble.
669:      
670: After the quench, $f^2$ is positive, so for sufficiently small $\sigma_0^2$ we have
671: \be \label{Uth}
672: U= \tau\frac{\lambda f^2}{4\ell^2} \int\left(f^2-2\langle\langle\Bphi^2\rangle\rangle\right) 
673: dx\:d\eta =\frac{\tau}{\tau_0}(C_0-C_2){\cal V}_0,
674: \ee
675: and
676: \be \label{linear}
677: L_{\bot}^{(2)}=\frac{\tau}{\tau_0} \;C^{(2)}\;{\cal V}_0,~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
678: L_{\|}^{(2)}=\frac{\tau_0}{\tau}\;C^{(2)}\;{\cal V}_0.
679: \ee
680: 
681: 
682: In fig.(\ref{fig2}) both straight lines, (\ref{linear}) with (\ref{C2}), are included 
683: for comparison. It may be observed that the integral $L_{\|}^{(2)}$ involving the
684: longitudinal gradients follows the straight line decrease almost until the onset of the roll-down.
685: This means that the rapidity gradients basically stay random. On the other hand, the integral
686: $L_{\bot}^{(2)}$  follows the linear rise only for about one unit of proper time after the onset
687: of the evolution. Already near $\tau/\tau_0 \sim 2$, the transverse gradients 
688: are strongly affected by the dynamics  and interfere with the kinetic energy. 
689: Due to the relative factor of $1/\tau^2$ of $L_{\|}$ as compared to $L_{\bot}$ the dynamics
690: quickly gets dominated by the transverse gradients alone, such that
691: the average kinetic energy follows the average transverse-gradient energy $L^{(2)}_\bot$,
692: while the rapidity gradients (in $L_{\|}^{(2)}$ and $L_{\|}^{(4)}$) which decrease 
693: like $1/\tau$ are no longer relevant for the overall dynamical evolution.
694:  
695: Disregarding rapidity gradients altogether, the EOM (\ref{EOM}) reduces to
696: \be
697: \frac{1}{\tau}\partial_\tau\Bphi+\partial_{\tau\tau}\Bphi-\partial_{xx}\Bphi
698:  -m^2\Bphi=0,
699: \ee
700: which describes wave propagation in transverse direction, $A(\tau) \exp(ipx)$. 
701: Here the mass $m^2$ again characterizes the fluctuations around $\Bphi=0$, 
702: \be
703: m^2=\lambda f^2/\ell^2
704: \ee
705: so $m^2$ is negative for negative $f^2$ (where $\Bphi=0$ is the stable minimum),
706: and it is positive  for potentials which actually do have a 
707: lower symmetry-breaking minimum. The amplitudes $A(\tau)$
708: generically are Bessel functions,
709: \bea \label{waves}
710: A(\tau) &\sim& J_0(\tau\sqrt{p^2-m^2})~~~~~~~~~~~\mbox{for}~~~~ p^2-m^2 >0, 
711: \nonumber \\
712: A(\tau) &\sim& I_0(\tau\sqrt{m^2-p^2})~~~~~~~~~~~\mbox{for}~~~~ p^2-m^2 <0 .         
713: \eea 
714: For large values of their arguments the amplitudes of $J_0$ decrease like $1/\sqrt{\tau}$, 
715: while $I_0$ contains exponentially rising parts.
716: Modes with large transverse wave numbers contribute most to $L^{(2)}_\bot$.
717: Therefore, with their amplitudes decreasing like $1/\sqrt{\tau}$, the kinematical
718: factor $\tau$ in $L^{(2)}_\bot$ is compensated. So we expect that the linear rise of $L^{(2)}_\bot$ ends 
719: as soon as the dynamics is dominated by the transverse gradients and is followed by a phase where 
720: \be
721: \langle\langle \,T \,\rangle\rangle\sim\langle\langle L^{(2)}_\bot\rangle\rangle\sim \mbox{const.}|_\tau.
722: \ee
723: For negative $m^2$ no amplification occurs. After the quench, however, when $f^2$ has become positive,  
724: a few modes with small transverse wave numbers will start to get amplified.
725: Typically, for wave numbers $p=2\pi k/N$, with $k$ integer ($0\leq k\leq N/2$), 
726: waves with $k/N < \sqrt{\lambda}f/(2\pi\ell)$ get amplified, e.g. the lowest three or four out of $N=100$ for
727: $\ell\sim 5$ (for $\lambda=1$ and $f^2=1$). At first, the rate of amplification is slow because the exponential
728: rise is compensated by a decreasing function for small arguments in $I_0(x)$. These low-$k$ modes do not
729: contribute much to $L^{(2)}_\bot$. In fact, the $k=0$ mode, which experiences the largest rate of
730: amplification, does not contribute at all. 
731: 
732: While the amplification effect is not very pronounced for $L^{(2)}_\bot$, the few slowly exponentially 
733: rising contributions from the lowest-momentum transverse waves cause a noticeable rise of the
734: condensate $\langle\langle \Phi^2 \rangle\rangle$ after some time. This enters into the fluctuating part $C_2$ 
735: of the potential $U$ and drives it away from its linear rise given by (\ref{Uth}).
736: Then also the fourth-order terms in the potential become important and the dynamical evolution 
737: subsequently is dominated by the local potential. 
738: This initiates the roll-down of the field configuration at the majority of the lattice points into the
739: true vacuum $\Bphi_0=(0,0,1)$. The transition into the symmetry-violating configuration takes place,
740: with formation of bags and solitons in those regions where the winding density happens to be high.
741: 
742: To estimate the time $\tau_1$ for the onset of the roll-down we consider the $k=0$ mode with
743: amplitude $I_0(\tau m)$. Amplification of this amplitude by a factor $e$ in the time interval from
744: $\tau_0$ to $\tau_1$ requires
745: \be \label{tau1}
746: \ln I_0(\tau_1 m)= 1+ \ln I_0(\tau_0 m).
747: \ee
748: The r.h.s. depends only very weakly on $\tau_0$, as long as $(\tau_0 m)\leq 1$. In fact, $(\tau_1 m)$ varies
749: only from 2.26 to 2.55 for  $0\leq(\tau_0 m)\leq 1$. So, for convenience we simply take $(\tau_1 m)\approx 5/2$ if 
750: $(\tau_0 m)$ is of the order of 1 or less. Otherwise, for larger values of $(\tau_0 m)$, $\tau_1$ has to be obtained more
751: accurately from (\ref{tau1}). Typically, therefore, the transition from the gradient-dominated 
752: to the potential-dominated phase, happens near
753: \be\label{onset}
754: \tau_1 \sim \frac{5\,\ell}{2\sqrt{\lambda f^2}}.
755: \ee
756: However, up to this time $\tau_1$ of the onset of the roll-down, i.e. throughout the whole gradient-dominated phase 
757: the potential plays no significant role.
758: The overall evolution proceeds practically independently
759: from the (positive or negative) value of $f^2$ in the $\Phi^4$-potential (\ref{pot}). 
760: This also implies that the quenchtime
761: (the timescale for changes in $f^2$) is irrelevant as long as it is smaller than the time
762: during which the gradient terms dominate the evolution and it justifies the use of the sudden quench
763: approximation where we impose the 'cold' ($\Te=0$) potential from the outset at $\tau>\tau_0$.
764:  
765: With $\ell/\tau_0 > 1$, the ratio $(\tau_1/\tau_0)^2$ is sufficiently large to render all longitudinal
766: (rapidity) gradients unimportant as compared to the potential. 
767: This means that during the subsequent roll-down different rapidity slices become effectively decoupled,
768: and begin to evolve independently from each other, 
769: while in longitudinal directions the solitons contract to lattice unit size.
770: Within these rapidity slices, $\pi$- and $\sigma$-modes propagate transversely, 
771: and eventual further annihilations of soliton-antisoliton pairs take place
772: while the transverse shapes of the sqeezed bags are established.
773: 
774: By the end of the roll-down the remaining nontrivial and sufficiently separate structures
775: have essentially reached their stable form. Apart from small fluctuations, the integral 
776: (or sum) over the absolute values of the winding density $n=\int \!|\rho| \;dx dz$ then
777: stabilizes and counts the (integer) number of these defects. Therefore we identify
778: the end-of-the-roll-down time with the (chemical) freeze-out time $\tau_f$
779: when the numbers of baryons and antibaryons created are fixed. A rough estimate for $\tau_f$
780: may be obtained if we follow the further amplification of the amplitudes $I_0(\tau m)$ of the 
781: $k=0$ modes beyond $\tau_1$. For large arguments the increase in $I_0(\tau m)$ is mainly due to the
782: exponential $\exp(\tau m)$, so we obtain
783: \be\label{freeze}
784: \tau_f \sim \tau_1 + \frac{\ell}{\sqrt{\lambda f^2}}\ln\left(\frac{\Phi(\tau_f)}{\Phi(\tau_1)}\right)
785: \ee 
786: For a typical amplification ratio of 5 to 10 during roll-down we then find an approximate freeze-out time
787: of
788: \be\label{tauf}
789: \tau_f \sim \frac{4\ell}{\sqrt{\lambda f^2}}.
790: \ee
791: This certainly represents a lower limit for the duration of the
792: roll-down, because the increasing $\Phi^4$ contributions to the potential will slow down the
793: symmetry-breaking motion. The numerical simulations confirm this simple argument and indicate
794: that $(m \tau_f) \sim 4-5$ provides a reasonably accurate estimate for the freeze-out time
795: (as long as $(\tau_0 m) \leq 1$). 
796: After the quench, when $f_0^2$ has assumed its $(\Te=0)$-value $f_0^2=1$, we may neglect the 
797: small contributions of the explicit symmetry-breaking $H$ to $f^2$ and to the $\sigma$-mass
798: $m_\sigma^2$ in eqs.(\ref{f2}) and (\ref{masses}) and rewrite (\ref{tauf}) in the form
799: \be\label{taufreeze}
800: \frac{\tau_f}{\tau_0}\approx \frac{4\sqrt{2}}{\tau_0 m_\sigma}.
801: \ee
802: 
803: The typical example for an evolution given in fig.\ref{fig2} shows how during the roll-down
804: the configurations pick up an appreciable amount of kinetic energy  
805: until the potential starts to deviate from its linear rise and interferes 
806: with $\langle\langle \,T\rangle\rangle$. Subsequently, $\langle\langle U\rangle\rangle$ starts to pick up
807: the unphysical linearly rising lattice contributions (\ref{shrilat}) of the shrinking solitons, 
808: while the time-averaged $\langle\langle \,T\rangle\rangle$ remains basically constant. 
809: As the heavy solitons carry no kinetic energy, $\langle\langle \,T\rangle\rangle$ then resides
810: in small transversely propagating fluctuations which eventually are emitted as $\sigma$ and $\pi$ mesons.
811: 
812: 
813: \subsection{Correlation lengths and defect numbers}
814: In contrast to the integer net-baryon number $B=\int \!\rho \;d^2x$, the integral (or lattice sum) over the 
815: absolute values of the local winding density $|\rho|$ 
816: \be \label{defnum}
817: n=\int |\rho| \; dxdz
818: \ee
819: generally is not integer. 
820: The ensemble average of $n$ is closely related to the coherence  length $R$ for the 
821: field unit-vectors in the statistical ensemble of $O(3)$-field configurations. If an $O(n)$-field
822: is implemented on a $d$-dimensional cubic lattice with lattice constant $a$, then
823: the field orientations on the vertices of a sublattice with lattice unit $R/a$ can
824: be considered as statistically independent. Then the average 
825: $\langle\langle \; n \; \rangle\rangle$ expected on an $N^d$ lattice is 
826: \be\label{Kibble}
827: \langle\langle \; n \; \rangle\rangle=\nu_d \; (aN/R)^d
828: \ee
829: where $\nu_d$ is the average fraction of the surface of the sphere ${\bf S}^d$ covered by the image of the 
830: sublattice unit (this is the very definition of a winding density). 
831: The number $\nu_d$ can be estimated for different manifolds~\cite{Kibble}. 
832: For the map (compactified-)${\bf R}^2\rightarrow {\bf S}^2$
833: defined by the unit-vectors $\hphi(x,z)$ of the $O(3)$-field in $d=2$ dimensions it is $\nu_2=1/4$,
834: (i.e. $1/2^{d+1}$ for each of two triangles which make up each square sublattice unit cell). 
835: 
836: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
837: \begin{figure}[h]
838: \centering
839: \includegraphics[width=7cm,height=8.5cm,angle=-90]{fig3a.ps}
840:  \includegraphics[width=7cm,height=8.5cm,angle=-90]{fig3b.ps}
841: \caption{\label{fig3}Initial (measured) coherence  lengths $R$ and number of defects $n$ 
842: as functions of initial temperature $\Te$, measured for five random initial
843: configurations for each temperature, on an $N=100$ lattice, with $m$ put to zero (left), 
844: and with $m=1/\ell$ for $\ell=4$ (right). (All quantities in lattice units $a$).}
845: 
846: \includegraphics[width=10cm,height=12cm,angle=-90]{fig5.ps}
847: \caption{\label{fig5} Crosses show the measured coherence  lengths $R_\bot, R_\|$ and the average ${\bar R}$ as obtained from
848: the definition (\ref{avcorr}), as functions of proper time $\tau$. For $\tau>2$ they are parametrized by
849: power laws with exponents 0.15, 0.75, and 0.25, respectively. The measured defect number $n$ (full line) 
850: is compared to the statistical result (\ref{Kibble}) with $\nu=1/6$ (crosses connected by lines).
851: ($N=100$,$\Te=0.2$,$\sigma_0=0.2$,$\ell=4$,$H=0.1$,$\lambda=1$).}
852: \end{figure}
853: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% 
854: 
855: 
856: Inserting our result (\ref{Rm0}) (obtained for $m=0$) into this estimate for $d=2$ dimensions, leads to 
857:  $ \langle\langle \; n \; \rangle\rangle _{\tau=\tau_0} \propto \Te^2$.
858:  In fig.\ref{fig3}a this is compared with numbers $n$ and coherence  lengths $R$ measured 
859:  for several initial configurations on an $N\times N$
860:  lattice for different temperatures (for $N=100$ and mass $m=0$). Evidently, the finiteness of the lattice 
861:  causes a small systematic deviation from this $\Te^2$-dependence, especially for large values of 
862:  $\langle\langle \, n \, \rangle\rangle$, as the coherence  length approaches the lattice constant. 
863:  The measured numbers $\langle\langle \; n \; \rangle\rangle$ 
864:  follow (\ref{Kibble}) with satisfactory accuracy for $\nu\approx 1/5$. 
865:  In fig.\ref{fig3}b the same comparison is shown for non-vanishing mass $m=1/\ell$, for $\ell/a=4$.
866:  For $m\ne 0$ the coherence  length $R$ can be obtained from (\ref{corrinit}) and compared to the 
867:  measured values. Fig.\ref{fig3}b shows that they are reasonably well described
868:  by $R\propto\Te^{-0.8}$. The corresponding measured numbers $n$ follow (\ref{Kibble}) with good
869:  accuracy for $\nu\approx 1/6$. Of course, small changes of $\nu$ could be absorbed into a slightly
870:  redefined coherence  length (note that the correlation function (\ref{corrinit}) for $m=0$ does not decrease 
871:  exponentially). We shall, however, keep the definition (\ref{cut}).
872: 
873:     
874: 
875: The above considerations apply to random configurations which need not contain any
876: fully developed solitons but may consist of only small fluctuating local winding densities
877: which cover small fractions of the image sphere.   
878: However, if the configurations 
879: finally have evolved into an ensemble of well-separated solitons or antisolitons embedded in a 
880: topologically trivial vacuum with only small fluctuations in the local winding density, then the 
881: integral (\ref{defnum}) counts the number of these embedded baryons-plus-antibaryons. We therefore adopt the
882: notion 'number of defects' for $\langle\langle \; n \; \rangle\rangle$, irrespective whether
883: configurations comprise only small local winding densities, or partial or complete solitons.   
884: 
885: 
886: 
887: For a typical evolution (see e.g. fig.\ref{fig2}) the number of defects
888: measured as function of proper time shows a slow decrease which follows approximately
889: a power law 
890: \be\label{powlaw}
891: n\sim n_{(\tau=\tau_0)}\left(\frac{\tau}{\tau_0}\right)^{-\gamma} .
892: \ee
893: By the end of the
894: roll-down at freeze-out time $\tau_f$ this decrease levels off and $n$ settles near the constant 
895: which counts the number of the finally surviving fully developed solitons-plus-antisolitons 
896: (cf. figs.\ref{fig1}). The decrease in $n$ reflects the slow increase
897: in the average coherence  length ${\bar R}$ up till the end of the roll-down. 
898: The longitudinal coherence  length $R_\|$ 
899: grows very slowly because rapidity gradients are suppressed with $1/\tau$ in the Bjorken frame. 
900: This leads to an effective decoupling of field-vectors in longitudinal direction
901: and subdues the drive for aligning field orientations in adjacent rapidity bins. On the other hand,
902: the transverse coherence  length $R_\bot$ grows rapidly. 
903: For $R_\bot\gg R_\|$, the average radius ${\bar R}$ obtained from (\ref{avcorr}) is dominated by $R_\|$.
904: A typical example is shown in fig.\ref{fig5} for an evolution which starts at $\tau_0/a=1$. 
905: The average ${\bar R}$ grows with an exponent of $\alpha \approx 0.25$. 
906: The statistical argument in (\ref{Kibble}) then leads to $n\sim \tau^{-2\alpha}$, with $2\alpha=\gamma\approx 0.5$. 
907: This is slightly steeper
908: than the measured decrease in $n$. But as the growth in the coherence  radii sets in only after one
909: or two units of $\tau$ after the onset of the evolution, the final number of surviving defects is
910: reasonably well reproduced by the statistical expression (\ref{Kibble}) 
911: (we adopt $\nu=1/6$ from fig.\ref{fig3}b). 
912: Altogether, we typically find exponents $\gamma\sim 0.4\pm 0.05$
913: for the decrease (\ref{powlaw}) of the number of defects. Then, with
914:  (\ref{taufreeze}) for the typical freeze-out time, we have
915: \be \label{decrease}
916: \langle\langle \; n \; \rangle\rangle_{|\tau=\tau_f}=\langle\langle \; n \; \rangle\rangle_{|\tau=\tau_0}
917: \left(\frac{\tau_0\;m_\sigma}{4\sqrt{2}}\right)^{0.4}
918: \ee
919: for the reduction of  the number of defects  from its initial value at the onset of the evolution until
920: the end of the roll-down. With $(\tau_0\,m_\sigma)$ of the order of 0.5 to 1, 
921: we find reduction factors
922: of 1/3 to 1/2, which is not even one order of magnitude. So, this is not a dramatic result. The reason
923: is, evidently, that in the expanding Bjorken frame 
924: the gradient coupling in rapidity direction quickly gets suppressed.
925: 
926: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
927: \begin{figure}[h]
928: \centering
929: \includegraphics[width=9cm,height=14cm,angle=-90]{fig4.ps}
930: \caption{\label{fig4}The numbers of defects $n$ for two different values $H=0$ and $H=1.0$ for the strength of the 
931: explicit symmetry breaking, each for five evolutions on a $100\times 100$ lattice ($\Te=0.3,\lambda=1,\ell=5,
932: \sigma_0=0.2$). The arrows point to the freeze-out times $\tau_f\approx 22$, and $\tau_f\approx 11$, respectively,
933: where the average lengths $|\Bphi|$ of the chiral field vectors have reached  $|\Bphi|=1$. 
934: The initial values of $n$ lie within a band from 145 to 185, 
935: they all end (for both values of $H$) in a band from 32 to 52, which corresponds to reduction
936: factors of about 1/4.  }
937: \end{figure}
938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
939: 
940: 
941: It should be noted that all numerically measured exponents are independent of the
942: choice of the lattice constants, because scaling $x\rightarrow ax$, (i.e.$\ell\rightarrow a\ell$), 
943: $\eta\rightarrow b\eta$, $\tau\rightarrow (a/b)\tau$ leaves the EOM (\ref{EOM}) invariant.
944: The length unit $a$ only
945: serves to define the resolution with which the spatial structure of the field configurations is analyzed
946: and all physical results should be independent of this scale. On the other hand, the initial time $\tau_0$
947: denotes the physical point in time when the system begins its evolution in terms of hadronic degrees 
948: of freedom with a sudden or rapid quench in the relevant potential. So, physical  results generally will
949: depend on $\tau_0$, as is evident from the reduction factor obtained in (\ref{decrease}). 
950: 
951: 
952: Small explicit symmetry breaking ($H\ne 0$) accelerates the decrease of $n$
953: during roll-down, but at the same time it reduces the freeze-out time, such that the final number of $n$
954: remains essentially unaffected by small non-zero values of $H$.
955: Fig.\ref{fig4} shows a number of evolutions for two different
956: strengths $H$ of explicit symmetry breaking. 
957: 
958: The same is true if additional damping is introduced into the 
959: EOM (\ref{EOM}) by adding a term $\kappa\partial_\tau\Bphi$ with damping constant $\kappa$ to account for the
960: fact that the field fluctuations are actually emitted from the expanding Bjorken rod, carrying away energy.
961: Through this dissipative dynamics the evolutions are slowed down, the roll-down times may be retarded by 
962: an order of magnitude, but the overall reduction factor in the number of surviving defects remains 
963: unaffected. Of course, all fluctuations then are damped away during the course of the evolution, and the integrals
964: (\ref{int3}) to (\ref{int5}) finally are determined by the remaining ensemble of squeezed solitons alone, 
965: while the kinetic energy goes to zero.
966: 
967: \subsection{Meson spectrum}
968: 
969: For times long after the roll-down the average kinetic energy  $\langle\langle \; T \; \rangle\rangle$,
970: and the potential energy parts $\langle\langle \; L_\bot+U \; \rangle\rangle$ (after subtraction of the
971: linearly rising (lattice) contributions from the squeezed solitons (\ref{shrilat})) converge towards the same constant
972: $E_f/2$. Their sum $E_f$ represents the average total energy stored in the mesonic field fluctuations 
973: after the roll-down. Both averages show residual fluctuations around their smooth background 
974: with opposite phases, such that their sum $E_f$ is smooth. Analysing the spectral density of  
975: either $\langle\langle \; T \; \rangle\rangle$, or $\langle\langle \; L_\bot+U \; \rangle\rangle$,
976: (after subtracting the background), tells us about the spectral distribution of pions and 
977: $\sigma$-mesons that will eventually be emitted from the expanding Bjorken rod. 
978: 
979: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
980: \begin{figure}[h]
981: \centering
982: \includegraphics[width=8cm,height=14cm,angle=-90]{fig6.ps}
983: \caption{\label{fig6} Spectral density $n(\omega)$ of the residual fluctuations in the 
984: average kinetic energy $\langle\langle \; T \; \rangle\rangle$ for times long after the roll-down,
985: ($N=80$, $\ell/a=4$, $H=0.1$). The arrows point to the lowest (double-)frequencies (\ref{omega}) for $\pi$- and
986: $\sigma$-mesons (see text). The green line is the exponential $\exp(-12a\omega)$.}
987: \end{figure}
988: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
989: 
990: We consider the Fourier-transforms 
991: \be
992: c(\omega)+i s(\omega)=\int_{\tau_a}^{\tau_b} \langle\langle \; T(\tau) -{\bar T}(\tau)\; \rangle\rangle
993:  e^{i \omega \tau} d\tau,
994: \ee
995: where the integral covers times long after the roll-down, e.g. $\tau_a/\tau_0\sim 100 $, $\tau_b/\tau_0\sim 1000$,
996: and ${\bar T}(\tau)$ subtracts the smooth background. The absolute value, $\epsilon(\omega)= \sqrt{c^2+s^2}$,
997: represents a spectral energy density, from which we may extract the spectral particle number density 
998: \be
999: n(\omega)=\frac{\epsilon(\omega)}{\omega} = \sum_{ij}n_{ij}^{(\pi)}\;\delta(\omega-2\omega_{ij}^{(\pi)})+
1000: \sum_{ij}n_{ij}^{(\sigma)}\;\delta(\omega-2\omega_{ij}^{(\sigma)}) + \cdot\cdot\cdot  .
1001: \ee
1002: The $ij$-sum with $i,j=0,1,2,...N/2$ covers all frequencies on the lattice for pions and 
1003: $\sigma$-mesons with masses $m_\pi$ and $m_\sigma$ given in (\ref{masses})
1004: \be \label{omega}
1005: \omega_{ij}^{(\pi/\sigma)}=\sqrt{\left(\frac{2\pi}{aN}i\right)^2+
1006: \left(\frac{2\pi}{a N}j\right)^2 + m_{\pi/\sigma}^2} ~~.
1007: \ee  
1008: Generically, $T(\tau)$ contains contributions $\sim (\cos(\omega_{ij}^{(\pi)}\tau))^2$ from the pionic
1009: fluctuations, and $\sim (\cos(\omega_{ij}^{(\sigma)}\tau)+ c)^2$ from the $\sigma$-fluctuations around some
1010: nonvanishing average $c$. Therefore, the spectral functions $\epsilon(\omega)$ and $n(\omega)$ will, in addition
1011: to the double frequencies $2\omega_{ij}^{(\sigma)}$, also contain contributions for the $\sigma$-mesons 
1012: at the single frequencies $\omega_{ij}^{(\sigma)}$. 
1013: 
1014: 
1015: Figure \ref{fig6} shows the spectral density $n(\omega)$
1016: as obtained from the residual fluctuations in the average kinetic energy. 
1017: The long vertical arrows point to the first four $2\omega_{ij}^{(\pi)}$
1018: pionic frequencies (\ref{omega}) for $ij=00,10,20,30$, with $m_\pi^2=H/\ell^2$, ($H=0.1,\ell/a=4,f_0=1$).
1019: It may be seen that the overwhelming part of the strength resides in the lowest and first excited pionic
1020: modes. The strength decreases rapidly with excitation energy, approximately like $\exp(-12a\omega)$. 
1021: The same is true for the strength of the $\sigma$-modes. (The short arrows in fig.\ref{fig6}
1022: point to the first three modes
1023: with $ij=00,10,20$, with single frequencies $\omega_{ij}^{(\sigma)}$ and double frequencies 
1024: $2\omega_{ij}^{(\sigma)}$). However, the number density $\sum n_{ij}^{(\sigma)}$ for the sigmas, 
1025: which we may extract from the strength located at the double frequencies
1026: $2\omega_{ij}^{(\sigma)}$, is only about $5\%$ of the pionic strength residing in the first three pionic modes.
1027: For an order-of-magnitude estimate of the pionic multiplicities we therefore ignore the $\sigma$-contributions.
1028: 
1029: 
1030: \subsection{Meson and baryon multiplicities}
1031: 
1032: To obtain a simple estimate for the energy $E_f$ finally available for meson production
1033: we consider the time of the onset of the roll-down $\tau_1$ in (\ref{onset}) which marks the transition from 
1034: the gradient-dominated to the potential-dominated phase. At this time, for sufficiently small $\sigma_0^2$, 
1035: the total energy is dominated by the linearly rising term $(\tau/\tau_0)C_0 {\cal V}_0$ in the potential
1036: (\ref{Uth}). With the onset of the roll-down the average potential 
1037: $\langle\langle \; U \; \rangle\rangle$ starts to deviate from this linear rise and bends
1038: down to interfere with $\langle\langle \; T \; \rangle\rangle$ and $\langle\langle \; L \; \rangle\rangle$,
1039: (cf. fig. \ref{fig2}). In the numerical simulations the large-time limit of 
1040: $\langle\langle \; U \; \rangle\rangle$ and $\langle\langle \; L \; \rangle\rangle$ is masked 
1041: by the (lattice-artificial) rise of the soliton contributions. But the asymptotic 
1042: $\langle\langle \; T \; \rangle\rangle$ is free of these artifacts and (apart from residual fluctuations)
1043: approaches a constant value, which  
1044: is well represented by the linearly rising $\frac{1}{2}(\tau/\tau_0)C_0 {\cal V}_0$ taken at $\tau=\tau_1$. 
1045: Approximating $\tau_1$ by $5/(\sqrt{2}m_\sigma)$ as given in (\ref{onset}), we then have (with $f^2=1$)
1046: \be \label{Ef}
1047: E_f = f_\pi^2\frac{\tau_1}{\tau_0}C_0{\cal V}_0\approx f_\pi^2\frac{5 m_\sigma}{8\sqrt{2}}(ab N^2),
1048: \ee
1049: where we have again neglected the small contribution of the explicit symmetry-breaking $H$ to the
1050: $\sigma$-mass.
1051: Within this level of accuracy we can also ignore that about $30\%$ of the pions 
1052: carry the energy $\omega_{10}^{(\pi)}$ (instead of $m_\pi$), 
1053: and obtain the pion multiplicity $n_\pi$ from dividing (\ref{Ef}) by $m_\pi$,
1054: \be
1055: n_\pi=\frac{5}{8\sqrt{2}}\frac{m_\sigma}{m_\pi} f_\pi^2 (ab N^2).
1056: \ee
1057: This number may be compared with the baryon-plus-antibaryon multiplicity given in (\ref{decrease}). 
1058: We use for $\langle\langle \; n \; \rangle\rangle_{|\tau=\tau_0}$ the statistical result (\ref{Kibble}),
1059: with initial spatial coherence  length $R_0$. Then we have
1060: \be
1061: \langle\langle \; n \; \rangle\rangle = \nu\left(\frac{aN}{R_0}\right)^2
1062: \left(\frac{\tau_0\;m_\sigma}{4\sqrt{2}}\right)^\gamma .
1063: \ee 
1064: The last factor relies on the estimates (\ref{onset}) and (\ref{tauf}) for the times $\tau_1$ and $\tau_f$,
1065: which are valid as long as $(\tau_0 m)\leq 1$, (otherwise they have to be obtained more accurately from
1066: (\ref{tau1}) and (\ref{freeze})).
1067: The evolutions described above have been performed for initial configurations selected with netbaryon
1068: number $B=0$. So, the average number $n_{\bar p}$ of antibaryons created during the phase transition 
1069: is $\langle\langle \; n \; \rangle\rangle/2$. With typical values $\nu\sim 1/4$, $\gamma\sim 0.4$
1070: we find for the multiplicity ratio of antibaryons to pions
1071: \be
1072: n_{\bar p}/n_\pi \approx 0.14 \frac{a}{b}\frac{m_\pi }{m_\sigma f_\pi^2}\frac{(\tau_0 m_\sigma)^\gamma}{R_0^2}.
1073: \ee
1074: With an overall energy scale $f_\pi^2$ of the order of the pion mass $m_\pi$, and $R_0$ of the
1075: order of $m^{-1}=\sqrt{2} m_\sigma^{-1}$, this ratio is 
1076: \be
1077: n_{\bar p}/n_\pi \approx 0.07 (\frac{a}{b}m_\sigma) (\tau_0 m_\sigma)^\gamma .
1078: \ee
1079: The ratio $a/b$ of the spatial and rapidity lattice constants which appears in this result has a 
1080: physical meaning: according to (\ref{initcorr}) it is equal to the ratio of the (transverse) spatial 
1081: coherence  length $R_0$ and the (longitudinal) rapidity coherence  distance $R_{\| 0}$
1082: in the initial configuration. Naturally, this ratio is of the order of $\tau_0$.
1083:  So, for initial times $\tau_0$ typically of the order of the inverse
1084: $\sigma$-mass we find antibaryon-to-pion multiplicity ratios of the order of $0.05$ to $0.1$.
1085: 
1086: 
1087: \section{Generalization to 3-d O(4)}
1088: 
1089: For the generalization to the 3+1 dimensional O(4)-model we keep the parametrization as given in eqs.(\ref{lag}),
1090: (\ref{L2}), and (\ref{pot}). In this case $f_\pi^2$ is an overall constant of dimension [mass$^2$], 
1091: so the physical fields $f_\pi\Bphi$ are of mass-dimension one.
1092: The winding density is no longer given by (\ref{top}),
1093: but we keep ${\cal{L}}^{(4)}$ as defined by the second equality in eq.(\ref{L4}). Conventionally, the strength of
1094: the  ${\cal{L}}^{(4)}$-term in (\ref{L4}) is given in terms of the Skyrme parameter $e$ as
1095: \be\label{Skyrme}
1096: \frac{2\lambda\ell^2}{(8\pi)^2} \Rightarrow \frac{1}{4e^2f_\pi^2}.
1097: \ee
1098: In this case the typical spatial radius of a stable skyrmion in its rest frame is mainly determined
1099: by the balance between ${\cal{L}}^{(2)}$ and ${\cal{L}}^{(4)}$, so it is of the order of $(ef_\pi)^{-1}$.
1100: 
1101: For the map (compactified-)${\bf R}^3\rightarrow {\bf S}^3$ defined by the unit vectors of 
1102: the $O(4)$-field in 3 spatial dimensions 
1103: the statistical result (\ref{Kibble}) for the average number of defects found on a $(aN)^3$ lattice 
1104: for initial configurations with coherence  length $R_0$ generalizes as
1105: \be\label{Kibble3}
1106: \langle\langle \; n \; \rangle\rangle|_{\tau=\tau_0}= \frac{5}{2^{3+1}}\; (aN/R_0)^3.
1107: \ee
1108: (We again use $a=b\tau_0$ for the lattice constants).
1109: The factor 5 counts the number of 3-simplices (tetraeders) which make up a cubic sub-lattice cell of size $R_0^3$,
1110: the factor $(1/2^{d+1})$ with $d=3$ is the (absolute value of the) average surface area  covered by the image
1111: of one 3-simplex on the image sphere ${\bf S}^3$.  So the factor $5/16$ counts the average 'number of defects'
1112: associated with a cubic lattice cell with lattice constant given by the initial coherence  length $R_0$.  
1113: A certain arbitrariness in the
1114: definition of the coherence  length may translate into modifications of this factor $5/16$;
1115: (e.g. for a random lattice of 3-simplices the factor 5 is replaced by ($24\pi^2/35 \sim 6.8$) \cite{Kibble}).
1116: In any case we do not expect order-of-magnitude changes in this factor as compared to the $d=2$ case, where we
1117: had $2/2^{2+1}$.
1118: 
1119: However, through the cubic power the result is now very sensitive to the actual value of $R_0$ in the
1120: initial ensemble. Different concepts about the physical nature of the initial configurations will imply 
1121: quite different ways to arrive at the appropriate initial coherence  lengths $R_0$. 
1122: For an initial ensemble which is characterized by a temperature $\Te$ we could proceed
1123: as in (\ref{corrinit}) and relate $R_0$ to the temperature, or to the mass 
1124: $m^2(\Te)=\lambda |f^2(\Te)|/\ell^2$ of the field fluctuations;
1125: but it has also been suggested \cite{Ellis}
1126: to tie $R_0$ to the parton density which makes it independent of the temperature concept.
1127: So, for the moment it seems appropriate to keep the initial coherence  length $R_0$ as a parameter.   
1128: 
1129: Adding a second transverse dimension does not change the result (\ref{avcorr}) for the average of the
1130: transverse and longitudinal coherence  lengths. The growth in the resulting ${\bar R}$ again is dominated 
1131: by the slow increase of the longitudial coherence  length $R_\|$ which is unaffected by additional
1132: transverse dimensions. The estimates (\ref{onset}) and (\ref{tauf}) for the times $\tau_1$ 
1133: and $\tau_f$ of the onset and end of the roll-down also remain unaffected, as they only rely on 
1134: the amplitudes $A(\tau)$ of the transverse waves (\ref{waves}), 
1135: irrespective of the number of spatial dimensions. (In this case, ${\cal{L}}^{(4)}$ now
1136: contributes to $L_\bot$ with a term containing four transverse gradients, acting on the direction of the
1137: O(4) field. The roll-down, however, takes place in areas which are topologically trivial, i.e. with small
1138: angular gradients, so we do not expect a strong effect on the roll-down times. 
1139: Within the approximations which led to (\ref{decrease}), we then find
1140: for the average number of baryons+antibaryons present after the roll-down
1141: \be\label{defects}
1142: \langle\langle \; n \; \rangle\rangle_{|\tau=\tau_f}=\frac{5}{16}\; \left(\frac{a N}{R_0}\right)^3
1143: \left(\frac{\tau_0\;m_\sigma}{4\sqrt{2}}\right)^{3\alpha}
1144: \ee
1145: with $\alpha\sim 0.2$ to 0.25. 
1146: We denote the transverse area $(aN)^2$ of the Bjorken rod by ${\cal A}$, and replace the ratio $(a/b)$ 
1147: of the lattice constants again by the initial time $\tau_0$. Then we obtain for the 
1148: rapidity density of antiprotons ($n_{\bar p}=\frac{1}{4}\langle\langle \; n \; \rangle\rangle$)
1149: \be\label{pbar}
1150: \frac{d n_{\bar p}}{d \eta}=\frac{5}{64}\; \frac{\tau_0{\cal A}}{R_0^3}
1151: \left(\frac{\tau_0 m_\sigma}{4\sqrt{2}}\right)^{3\alpha}.
1152: \ee
1153: (At this point we count all baryons as nucleons, assuming that excitational fluctuations and rotations
1154: contribute to the surrounding pionic fluctuations).
1155: Although the strength of the Skyrme term does not appear explicitly in (\ref{pbar}), the presence of the
1156: ${\cal{L}}^{(4)}$-term is essential for the formation of the solitons, because during the evolution 
1157: it transforms the
1158: sum over the absolute values of average winding densities into the average number of fully developed
1159: solitons and antisolitons. So it is hidden in the growth law characterized by $\alpha$. 
1160: 
1161: For meson production we adopt the considerations which led to the estimate (\ref{Ef}). Counting again all
1162: mesons as zero momentum pions we have 
1163: \be \label{npi}
1164: n_\pi=\frac{E_f}{m_\pi} = \frac{f_\pi^2}{m_\pi}\frac{\tau_1}{\tau_0}C_0{\cal V}_0
1165: =f_\pi^2\frac{5}{8\sqrt{2}}\frac{m_\sigma}{m_\pi}({\cal A}\,b N).
1166: \ee
1167: The rapidity density of negatively charged pions $n_{\pi^-}=\frac{1}{3}n_\pi$ then is
1168: \be \label{piminus}
1169: \frac{d n_{\pi^-}}{d \eta}= f_\pi^2\frac{5}{24\sqrt{2}}\frac{m_\sigma}{m_\pi}{\cal A}.
1170: \ee
1171: In a heavy-ion collision the transverse area ${\cal A}$ of the Bjorken rod will correspond to the 
1172: spatial overlap of the colliding relativistic nuclear slabs. As we have assumed spatially homogeneous 
1173: initial conditions we have to consider slabs with constant nucleon (area-)density. In order to account for the 
1174: number A of nucleons contained in one slab, its radius must be taken as $r_0$A$^{1/2}$,
1175: with $r_0\approx 1.2$ fm. Then, as function of centrality, $(d n_{\pi^-}/d \eta)$ is directly 
1176: proportional to the number of participants $N_p$, which is one of the basic experimental results in
1177: relativistic heavy-ion collisions. For central collisions of A-nucleon slabs we have $N_p = 2$A, so we find
1178: for the $\pi^-$-rapidity density per $N_p/2$ participants 
1179: \be \label{ppp}
1180: \frac{1}{N_p/2}\frac{d n_{\pi^-}}{d \eta}= \frac{5 \pi}{24\sqrt{2}}\frac{m_\sigma}{m_\pi}(r_0 f_\pi)^2.
1181: \ee
1182: This is an interesting result because all parameters have been absorbed into physical quantities. 
1183: There are, however, several caveats: We have used for this result the form of the potential (\ref{pot}) {\it after} 
1184: the quench (only this enters into the calculations). This means that differences between the average potential energy 
1185: {\it before} the quench (\ref{Uth0}) and immediately {\it after} the quench (\ref{Uth}) are left out. 
1186: However, this difference is of the order $\sigma_0^2$, 
1187: which has been neglected in (\ref{ppp}) anyway. But generally, the result (\ref{ppp}) should be 
1188: considered as a lower limit. It should further be noted
1189: that the result (\ref{ppp}) depends linearly on the time $\tau_1$ for the onset of the roll-down.
1190: The definition of $\tau_1$ in (\ref{onset}) is not very stringent and may be subject to changes by
1191: $\pm 20\%$. The estimate (cf. eq.(\ref{onset})) we used for the time $\tau_1$ 
1192: required $(\tau_0 m_\sigma)\leq\sqrt{2}$ , 
1193: i.e. with $m_\sigma \sim 3-5$ fm$^{-1}$, the initial time $\tau_0$ should not exceed $0.3-0.5$ fm.
1194: Another unsatisfactory feature of the homogeneous Bjorken rod is that the inhomogeneity
1195: in the nucleon (area-)density of real relativistic nuclear slabs with transverse radii $r_0$A$^{1/3}$ 
1196: has to be represented through radii $r_0$A$^{1/2}$ for homogeneous slabs.
1197: 
1198: The experimental value \cite{phenix} for the $\pi^-$-rapidity density per $N_p/2$
1199: lies between 1.2 and 1.5 for $N_p$ increasing up to 350.
1200: With $r_0 f_\pi=0.57$, $m_\sigma/m_\pi\approx 5-8$, our result (\ref{ppp}) leads to
1201: \be \label{pipp}
1202: \frac{1}{N_p/2}\frac{d n_{\pi^-}}{d \eta}\approx 0.75-1.2 ~.
1203: \ee
1204:  In the light of the reservations discussed above, this is quite satisfactory.
1205:  
1206:  In contrast to the parameter-free pion multiplicities, the result for baryon-antibaryon creation
1207: depends on two parameters: the time $\tau_0$ when the initial hadronic field ensemble is established 
1208: and begins its expansion, and the initial coherence  length $R_0$ within that ensemble.
1209:  From  (\ref{pbar}) and  (\ref{piminus}) we have (with $\alpha\sim 0.2$ to 0.25)
1210: \be\label{ratio}
1211: \frac{n_{\bar p}}{n_{\pi^-}}\approx 
1212: 0.15 \left(\frac{m_\pi m_\sigma}{f_\pi^2}\right)\frac{(\tau_0 m_\sigma)^{3\alpha+1}}{(R_0 m_\sigma)^3}.
1213: \ee
1214: The experimental value for the ratio of integrated ${\bar p}$ to $\pi^-$ multiplicities 
1215: lies between 0.065 and 0.085 \cite{phenix} for varying numbers of participants.
1216: With $m_\pi/f_\pi^2=3.0$ fm, and a typical $\sigma$-mass of $m_\sigma \approx$ 3 fm$^{-1}$,
1217: the experimentally observed multiplicity ratios are reproduced if $R_0$ and $\tau_0$ (both in [fm]) satisfy
1218: \be \label{rule}
1219: R_0 \approx (3 \tau_0)^{\alpha+1/3}.
1220: \ee
1221: For initial times in the range  0.2 $\leq \tau_0\leq$ 0.5 the dependence on $\alpha$  
1222: is very weak and the coherence  length varies in the range 0.7  $\leq R_0 \leq$ 1.2 (all in [fm]).
1223: These values are certainly within the limits of conventional assumptions. Interpreted in terms of
1224: a thermodynamic equilibrium ensemble, $R_0\sim 1\mbox{fm}\sim \Te^{-1}$ implies the standard
1225: estimate $\Te\sim 200$ MeV for the chiral phase transition. With our choice $a/b=\tau_0=R_0/R_{\| 0}$ for the ratio
1226: of the spatial and rapidity lattice constants, the initial time $\tau_0 \approx 1/3$ fm resulting from (\ref{rule})
1227: for $R_0=1$ fm then means that in the initial ensemble the initial rapidity coherence  distance $R_{\| 0}$ 
1228: extends over three units of rapidity.
1229:       
1230: \section{Conclusion}
1231: 
1232: We have presented numerical simulations of the dynamical evolution which chiral field configurations
1233: undergo in a rapidly expanding spatial volume. Starting at an initial time $\tau_0$ from a random hadronic
1234: field ensemble with restored chiral symmetry, we follow its ordering process and roll-down into the
1235: global potential minimum with spontaneously broken chiral symmetry. In accordance with standard concepts of
1236: heavy-ion physics we have considered one-dimensional longitudinal expansion of an essentially baryon-free
1237: region of high energy density, as it may be realized in the aftermath of an ultra-relativistic collision 
1238: of heavy ions for central rapidities. 
1239: 
1240: Performed on a space-rapidity lattice in proper time of comoving frames, 
1241: such simulations are very powerful instruments which
1242: allow to investigate a multitude of interesting features related to the chiral phase transition.    
1243: We have concentrated here on the topological aspects which are directly related to baryon-antibaryon
1244: multiplicity as a sensitive signal for the phase transition. Mesonic abundancies could be analysed as well,
1245: both for $\pi$ and $\sigma$ mesons (or any other elementary fluctuations included in the chiral field).
1246: Not only their spectra can be obtained, but from the instantaneous configurations the spectral power of their
1247: momentum distribution could be extracted at every point in time.  
1248: 
1249: 
1250: The method is not restricted to thermally equilibrated initial ensembles  with global or local temperature;
1251: inhomogenities and anisotropy in the correlation lengths could be implemented naturally. Surface effects could be 
1252: investigated by suitable boundary conditions. This may be interesting with respect to the $A$-dependence of
1253: spectra and multiplicities. Here we have applied only standard periodic conditions. 
1254: The one-dimensional expansion could be replaced by anisotropic or spherically symmetric expansion, which may be
1255: of specific interest in cosmological applications.  
1256: We have used the sudden quench approximation, which could be replaced by any desired time-dependence of
1257: the chiral potential with arbitrary quench times. 
1258: We have selected ensembles with conserved net-baryon number $B=0$ or very small $B$. 
1259: Any other choice would be possible, and it appears as a peculiarly attractive feature to
1260: study evolutions in ensembles with high net-baryon density, either fixed or in the form of grandcanonical
1261: ensembles. The method is well suited to analyse distribution, growth and realigning of domains with 
1262: disoriented chiral condensate as has been shown previously in purely dissipative dynamics~\cite{HoKl02}.
1263: The generalization to SU(3)-fields appears most interesting, to learn about strangeness production
1264: in terms of baryonic and kaonic abundance ratios.  
1265:  
1266: Evidently, the method opens up a wide field of applications. Unfortunately, however, we know very little
1267: about the nature and characteristics of the initial ensemble which enters crucially into all physical results.
1268: So, in our present analysis of antibaryon and pion multiplicities, the experimental data do not
1269: allow to draw definite conclusions about the validity of the topological approach, because the results depend 
1270: on two initial coherence lengths, the spatial $R_0$ and rapidity $R_{\| 0}$, 
1271: (which for an isotropic initial ensemble are related by $R_0=R_{\| 0}\tau_0$). We can only conclude that
1272: conventional assumptions about these quantities lead to results which are compatible with experimentally 
1273: detected multiplicities. So, luckily, the mechanism is not ruled out. On the other hand,
1274: an assumption like $\tau_0=R_0$ ( which would imply that the correlations have
1275: grown with the speed of light from a pointlike origin ) is ruled out: it would overestimate the abundance
1276: ratio in (\ref{ratio}) by a factor of 5.  
1277:  
1278:  
1279: \section{Acknowledgement}
1280: The author appreciates helpful discussions with J. Klomfass, H. Walliser, and H. Weigel. 
1281: 
1282: \begin{thebibliography}{99}
1283: \itemsep=-0.1cm
1284: 
1285: \bibitem{Skyrme} T.H.R. Skyrme, {\it{Proc.R.Soc.}} {\bf A260} (1961) 127;\\
1286:  
1287:  \bibitem{Schwesinger} B. Schwesinger, {\it{Nucl.Phys.}} {\bf A537} (1992) 253;\\
1288:  H. Weigel, {\it{Int.J.Mod.Phys.}} {\bf A11} (1996) 2419;\\
1289:  Y.S. Oh, D.P. Min, M. Rho, and N.N. Scoccola, {\it{Nucl.Phys.}} {\bf A534} (1991) 493;\\
1290: 
1291: \bibitem{Nakano} M.Chemtob, {\it{Nucl.Phys.}} {\bf B256} (1985) 600;\\
1292: H. Walliser, {\it{Nucl.Phys.}} {\bf A548} (1992) 649;\\
1293: D. Diakonov, V. Petrov, and M.V. Polyakov, {\it{Z.Phys.}} {\bf A359} (1997) 305;\\
1294: H. Weigel, {\it{Eur.Phys.J.}} {\bf A2} (1998) 391;\\
1295: T. Nakano et al., [LEPS Collaboration], {\it{Phys.Rev.Lett.}} {\bf 91} (2003) 012002;\\
1296: 
1297: 
1298: \bibitem{Hayashi} A. Hayashi, G. Eckart, G. Holzwarth, and H. Walliser, {\it{Phys.Lett.}} {\bf 147B} 
1299:  (1984) 5;\\
1300:  M.P. Mattis, and M.E. Peskin, {\it{Phys.Rev.}} {\bf D32} (1985) 58;\\ 
1301:  
1302:  \bibitem{Mattis} M.P. Mattis and M. Karliner, {\it{Phys.Rev.}} {\bf D31} (1985) 2833;\\
1303:  G. Eckart, A. Hayashi, and G. Holzwarth, {\it{Nucl.Phys.}} {\bf A448} (1986) 732;\\
1304:  
1305:  \bibitem{Brodsky}S. Brodsky, J. Ellis, and M. Karliner, {\it{Phys.Lett.}} {\bf B206} (1988) 309;\\
1306:  R. Johnson, N.W. Park, J.Schechter, V. Soni, and H. Weigel {\it{Phys.Rev.}} {\bf D42} (1990) 2998;\\
1307:  
1308: \bibitem{Jones} M.K. Jones et al., {\it{Phys.Rev.Lett.}} {\bf 84} (2000) 1398;\\ 
1309:  
1310: \bibitem{Ho96} G. Holzwarth, {\it{Z.Physik}} {\bf A356} (1996) 339;\\ 
1311:  
1312: 
1313: \bibitem{Witten}E. Witten, {\em Nucl. Phys.} {\bf B223} (1983) 422,433;\\
1314:  
1315: \bibitem{Mouss}B. Moussallam, {\it Ann. Phys.(NY)} {\bf 225} (1993) 264;\\
1316: F. Meier and H. Walliser, {\it Phys.Reports} {\bf 289} (1997) 383;\\
1317: 
1318: \bibitem{Bjorken} A.A. Anselm, {\em Phys. Lett.} {\bf B217} (1988) 169; \\
1319: A.A. Anselm and M.G. Ryskin, {\em Phys. Lett.} {\bf B266} (1991) 482;\\
1320: J.P. Blaizot and A. Krzywicki, {\em Phys. Rev.} {\bf D46} (1992) 246;
1321:  {\em Phys. Rev.} {\bf D50} (1994) 442;\\
1322: J.D. Bjorken, {\em Int. J. Mod. Phys.} {\bf  A7} (1992) 4189;\\
1323: S.Gavin, {\em Nucl. Phys.} {\bf A590} (1995) 163c;\\
1324: 
1325: 
1326: \bibitem{Bearden} I.G. Bearden et al.,  {\it Phys.Rev.} {\bf C65} (2002) 044903;\\
1327: M.M. Aggarwal et al., {\it Phys.Rev.} {\bf C65} (2002) 054912;\\
1328: T.K. Nayak et al., {\it Pramana} {\bf 57} (2001) 285-300;\\
1329: T.C. Brooks et al.,  {\it Phys.Rev.} {\bf D61} (2000) 032003;\\
1330: 
1331: 
1332: \bibitem{Gavin}
1333: S. Gavin, A. Gocksch, and R.D. Pisarski, {\it Phys.Rev.Lett.} {\bf 72} (1994) 2143;\\
1334: 
1335: \bibitem{HoKl02}G. Holzwarth and J. Klomfass, {\it Phys.Rev.} {\bf D66} (2002) 045032;\\
1336: 
1337: 
1338: \bibitem{DeGrand} T.A. DeGrand, {\it Phys. Rev.} {\bf D30} (1984) 2001;\\ 
1339: J. Ellis, U. Heinz, and H. Kowalski, {\it Phys. Lett.} {\bf B233} (1989) 223;\\
1340: J.I. Kapusta and A.M. Srivastava, {\it Phys.Rev.} {\bf D52} (1995) 2977;\\
1341: J. Dziarmaga and M. Sadzikowski,{\it Phys.Rev.Lett.} {\bf 82} (1999) 4192;\\
1342: 
1343: 
1344: \bibitem{phenix}K. Adcox et al., [PHENIX Collaboration], {\it Phys.Rev.Lett.} {\bf 88} (2002) 242301;\\ 
1345: 
1346: \bibitem{Kibble} T.W.B. Kibble, {\it J. Phys.} {\bf A9} (1976) 1387;\\
1347:  N.H. Christ, R. Friedberg and T.D. Lee, {\it Nucl.Phys.} {\bf B202} (1982) 89;\\
1348: 
1349: \bibitem{Zurek} W.H. Zurek, {\it Nature} {\bf 317} (1985) 505;\\
1350: G.E. Volovik, {\it Exotic Properties of Superfluid
1351: He$^3$} (World Scientific, Singapore 1992);\\ 
1352: M. Hindmarsh and T.W.B. Kibble, {\it Rep. Prog. Phys.} {\bf 58} (1995) 477;\\
1353: W.H. Zurek, {\it Phys. Rept.} {\bf 276} (1996) 177;\\
1354: P. Laguna and W.H. Zurek, {\it Phys. Rev. Lett.} {\bf 78} (1997) 2519;\\
1355: R.H. Brandenberger, {\it Pramana} {\bf 51} (1998) 191;\\
1356: G.J. Stephens, {\it{Phys.Rev.}} {\bf D61} (2000) 085002;\\
1357: 
1358: \bibitem{Huang} Z. Huang and X.-N. Wang, {\it Phys.Rev.} {\bf D49} (1994) 4335; \\
1359: F. Cooper, Y. Kluger, E.Mottola, and J.P. Paz, {\it Phys.Rev.} {\bf D51} (1995) 2377; \\
1360: M.A. Lampert, J.F. Dawson, and F. Cooper, {\it Phys.Rev.} {\bf D54} (1996) 2213; \\
1361: J. Randrup, {\it Phys. Rev.Lett.} {\bf 77}(1996) 1226; {\it Nucl.Phys.} {\bf A616} (1997) 531; \\
1362: T.C. Petersen and J. Randrup, {\it Phys.Rev.} {\bf C61} (2000) 024906; \\
1363: O. Scavenius, A. Dumitru, and A.D. Jackson, {\it Phys.Rev.Lett.} {\bf 87} (2001) 182302;\\
1364: 
1365: 
1366: \bibitem{Ho03} G. Holzwarth,{\it Phys.Rev.} {\bf D68} (2003) 016008; \\
1367: 
1368: \bibitem{HoKl01} G. Holzwarth and J. Klomfass,{\it Phys.Rev.} {\bf D63} (2001) 025021; \\
1369: 
1370: \bibitem{Ellis}J. Ellis and H. Kowalski, {\it Phys. Lett.} {\bf B214} (1988) 161.
1371: \end{thebibliography}
1372: \end{document}
1373: 
1374: 
1375: 
1376: 
1377: 
1378: