nucl-th0109075/npa.tex
1: \documentstyle[preprint,aps,epsfig]{revtex}
2: \begin{document} 
3: \preprint{Nuclear Physics A (2001) in press.}
4: \title{Probing Mechanical and Chemical Instabilities in Neutron-Rich Matter} 
5: \bigskip 
6: \author{\bf Bao-An Li\footnote{email: Bali@astate.edu}, Andrew T. Sustich, Matt Tilley and Bin Zhang} 
7: \address{Department of Chemistry and Physics\\
8: P.O. Box 419, Arkansas State University\\
9: State University, Arkansas 72467-0419, USA}
10: \maketitle
11: 
12: \maketitle 
13: \begin{abstract}
14: The isospin-dependence of mechanical and chemical instabilities is  
15: investigated within a thermal and nuclear transport model using 
16: a Skyrme-type phenomenological equation of state for neutron-rich matter. 
17: Respective roles of the nuclear mean field and the 2-body stochastic 
18: scattering on the evolution of density and isospin fluctuations in either 
19: mechanically or chemically unstable regions of neutron-rich matter are investigated. 
20: It is found that the mean field dominates overwhelmingly the fast growth of both 
21: fluctuations, while the 2-body scattering influences significantly the later growth of the 
22: isospin fluctuation only. The magnitude of both fluctuations decreases with the increasing 
23: isospin asymmetry because of the larger reduction of the attractive isoscalar mean 
24: field by the stronger repuslive neutron symmetry potential in the more neutron-rich matter.
25: Moreover, it is shown that the isospin fractionation happens later, but grows faster 
26: in the more neutron-rich matter. Implications of these results to current 
27: experiments exploring properties of neutron-rich matter are discussed.\\
28: {\bf PACS} numbers: 25.70.-z, 25.75.Ld., 24.10.Lx\\
29: {\bf Key Words:} {\it Isospin physics, equation of state, neutron-rich matter, 
30: fluctuations and instabilities}
31: \end{abstract}  
32: \newpage 
33: \section{Introduction}
34: Explosion mechanisms of supernova and related properties of neutron stars are 
35: among the most interesting topics of modern nuclear astrophysics.
36: Investigations into these questions rely critically on the knowledge about the 
37: equation of state of isospin asymmetric nuclear matter. 
38: Thus, one of the most important subjects in nuclear physics is the 
39: study of the isospin-dependence of the nuclear equation of state 
40: using nuclear reactions induced by neutron-rich nuclei or radioactive 
41: beams\cite{rib,ria,li01}. In these reactions transient states of nuclear 
42: matter with sufficiently high isospin asymmetries as well as large thermal and 
43: compressional excitations can be created. The rapid progress in recent 
44: experiments with rare isotopes has therefore made the study of novel properties of 
45: extremely isospin-asymmetric nuclear matter possible. The planned Rare Isotope 
46: Accelerator ({\rm RIA}) will further enhance the exploration of this new frontier 
47: dramatically\cite{ria}. 
48: 
49: Prospects for discovering new physics in neutron-rich matter have generated 
50: much interest in the nuclear science 
51: community\cite{li01,li97,li98,dasgupta,tsang01,gary,sherry}. In particular, 
52: the isospin-dependence of the nuclear equation of state ({\rm EOS}), 
53: reflected mainly in the density dependence of the symmetry energy,
54: has recently received much attention. This is because the isospin-dependence 
55: of the nuclear {\rm EOS} is among the most 
56: important but very poorly known properties of neutron-rich 
57: matter\cite{wir88,brown00,hor00,dito01,lom01,bom01}. 
58: It is very important to the mechanisms of Type II supernova explosions and 
59: neutron-star mergers. It also determines the proton fraction 
60: and electron chemical potential in neutron stars at $\beta$ equilibrium. These 
61: quantities consequently influence the cooling rate of protoneutron stars and 
62: the possibility of kaon condensation in dense stellar 
63: matter\cite{lat91,bom94,sum94,lee96,pra97}. Moreover, the isospin-dependence of 
64: the nuclear {\rm EOS} also determines the stability boundaries of asymmetric
65: nuclear matter. These boundaries and their isospin-dependence subsequently
66: affects the multifragmentation mechanism in nuclear reactions at intermediate 
67: energies. Multifragmentation is thought to occur in symmetric nuclear matter 
68: due to the growth of instabilities triggered by density fluctuations, such as 
69: the Coulomb, surface and volumetric instabilities\cite{moretto}. While in 
70: isospin-asymmetric nuclear matter, it is interesting to study not only the
71: isospin-dependence of the above instabilities, but also new mechanisms for nuclear
72: multifragmentation, such as the chemical instability.
73: 
74: In this work we shall explore the isospin-dependence of both mechanical and chemical 
75: instabilities using a Skyrme-type phenomenological {\rm EOS} 
76: for neutron-rich matter within both a thermal model and a nuclear transport model.
77: The paper is organized as follows. In Section 2, we shall first outline the 
78: isospin-dependent nuclear {\rm EOS} that we use in this work, and then investigate 
79: relevant thermodynamical properties of neutron-rich matter using a thermal model. 
80: In Section 3, we shall establish boundaries of the mechanical and chemical 
81: instabilities and their isospin-dependence. Then, in Section 4 we study the 
82: dynamical growth of mechanical and chemical instabilities and associated phenomena within a
83: nuclear transport model. Finally, a summary and outlook are given in Section 5.
84: 
85: \section{Equation of State and Thermodynamical Properties of Neutron-Rich Matter}
86: To understand the thermodynamic properties of isospin-asymmetric nuclear matter, 
87: we must have a good knowledge of its EOS. The isospin-dependent part of the 
88: nuclear {\rm EOS} is critical for many unique properties of asymmetric matter. 
89: At present, nuclear many-body theories predict vastly different 
90: isospin-dependent nuclear {\rm EOS} depending on both the calculation 
91: techniques and the bare two-body and/or three-body interactions employed, 
92: see e.g., \cite{wir88,brown00,hor00}. 
93: Therefore, we use here a Skyrme-type phenomenological {\rm EOS} for asymmetric
94: nuclear matter. With proper parameterizations, it allows us to sample 
95: predictions by different many-body theories and to calculate
96: many thermodynamical properties analytically. 
97: 
98: Various studies (e.g., \cite{bom91,hub93}) have shown that the 
99: energy per nucleon $e(\rho,\delta)$ in asymmetric nuclear matter 
100: of density $\rho$ and isospin asymmetry parameter
101: \begin{equation} 
102: \delta\equiv (\rho_n-\rho_p)/(\rho_n+\rho_p) 
103: \end{equation}
104: can be approximated very well by a parabolic function in $\delta$.
105: At zero temperature, the $e(\rho,\delta)$ can be parameterized as 
106: \begin{equation}\label{ieos}
107: e(\rho,\delta)= \frac{a}{2}+\frac{b}{1+\sigma}u^{\sigma}+\frac{3}{5}e_F^0u^{2/3}
108: +S(\rho)\cdot\delta^2.
109: \end{equation}
110: In the above $u\equiv \rho/\rho_0$ is the reduced density, $e_F^0$ is the Fermi 
111: energy and $a=-123.6$ MeV, $b=70.4$ MeV and $\sigma=2$ corresponding to a stiff 
112: nuclear {\rm EOS} of isospin-symmetric nuclear matter. The last term is the 
113: symmetry energy whose density dependence is currently rather 
114: uncertain\cite{wir88,brown00,hor00,dito01,lom01,bom01}. We adopt
115: here a form of
116: \begin{equation}
117: S(\rho)=S_0(\rho_0)\cdot u^{\gamma},
118: \end{equation}
119: which was used by Heiselberg and Hjorth-Jensen in their recent 
120: studies of neutron stars\cite{hei00} with $S_0(\rho_0)=30$ MeV and $\gamma$ 
121: as a free parameter. We note here that a value of about $\gamma=0.6$ was obtained
122: by fitting to the result of variational many-body calculations\cite{hei00,akm97}. 
123: The $\gamma$ parameter is directly related to the $K_{sym}$ parameter of 
124: asymmetric nuclear matter via 
125: \begin{equation}
126: K_{\rm sym}\equiv 9\rho_0^2\frac{\partial^2 S(\rho)}{\partial \rho^2}|_{\rho
127: =\rho_0}=9S_0(\rho_0)\cdot\gamma(\gamma-1).
128: \end{equation}
129: The values of $K_{sym}$ predicted by many-body theories scatter
130: from about $-400$ {\rm MeV} to $+466$ {\rm MeV}(e.g. \cite{bom91}). 
131: Experimental values extracted from studying giant monopole 
132: resonances of asymmetric nuclei do not constrain the $K_{\rm sym}$ 
133: parameter either. The reported experimental values of $K_{sym}$ 
134: are between $-566\pm 1350$ MeV and $34\pm 159$ MeV\cite{shl93}.
135: 
136: Shown in Fig. 1 are the equations of state with $\gamma=$ 0.5 and 1
137: corresponding to $K_{sym}=$ -68 MeV and 0, respectively. The saturation points
138: of asymmetric nuclear matter with different $\delta$ are linked with the dashed lines 
139: to guide the eye. These two $\gamma$ parameters lead to rather different
140: saturation points especially for neutron-rich matter. The main features of the
141: equation of state obtained with $\gamma=0.5$ and $\gamma=1$ are surprisingly
142: similarly to those based on the Skyrme Hartree-Fock (SHF) and the Relativistic 
143: Mean Field (RMF) models\cite{oya98}, respectively. 
144: 
145: It is well known that the symmetry energy has a kinetic and a potential 
146: contribution
147: \begin{equation} S(\rho)=\frac{3}{5}e_{\rm F}^0u^{\frac{2}{3}}(2^{\frac{2}{3}}-1)
148: +V_2 
149: \end{equation} 
150: where $V_2$ is the potential contribution. The corresponding symmetry 
151: potential energy density is  
152: \begin{equation} 
153: W_{asy}=V_2\rho\delta^{2}, 
154: \end{equation} 
155: and the single-particle potential  
156: $v_{\rm asy}^{q}$ can be obtained from 
157: \begin{equation}\label{vasy} 
158: v_{\rm asy}^{q}=\frac{\partial W_{asy}}{\partial \rho_{q}}
159: =\pm(S_0u^{\gamma}-12.7u^{2/3})\delta +
160: (S_0(\gamma-1)u^{\gamma}+4.2u^{\frac{2}{3}})\delta^{2}
161: \end{equation} 
162: where ``+" and ``-" are for $q=neutron$ and $q=proton$, respectively. 
163: Shown in Fig.\ 2 are the symmetry energy and the corresponding symmetry potential
164: with three $\gamma$ parameters. The magnitude of repulsive (attactive) symmetry 
165: potentials for neutrons (protons) increases with both the density $\rho$ and 
166: isospin asymmetry $\delta$. The most important effect of $v_{asy}^q$ is to 
167: cause the migration of neutrons (protons) from relatively high (low) to 
168: low (high) density regions leading to the isospin fractionation (distillation)
169: effect, and this effect grows with increasing $\delta$. 
170: 
171: For asymmetric nuclear matter at a finite temperature $T$, 
172: the nucleon chemical potential $\mu_q$ corresponding to the {\rm EOS} 
173: of Eq.\ \ref{ieos} is given by\cite{jaqaman1}
174: \begin{equation}\label{muq} 
175: \mu_q=au+bu^{\sigma}+v_{\rm asy}^{q}+T\left[{\rm ln}(\frac{\lambda_T^3\rho_q}{2}) 
176: +\sum_{n=1}^{\infty}\frac{n+1}{n}b_n(\frac{\lambda_T^{3}\rho_q}{2})^n\right], 
177: \end{equation} 
178: where $\lambda_T=\left[2\pi\hbar^{2}/(m_q T)\right]^{1/2}$ is the thermal 
179: wavelength of a nucleon. The coefficients $b_n$ are obtained from mathematical 
180: inversion of the Fermi distribution function\cite{jaqaman1}. 
181: From the above chemical potentials for neutrons and protons, the global pressure 
182: for asymmetric nuclear matter can be obtained from the 
183: Gibbs-Duhem relation 
184: \begin{equation}\label{gibbs} 
185: \frac{\partial P}{\partial \rho}=\frac{\rho}{2}\left[(1+\delta) 
186: \frac{\partial \mu_n}{\partial \rho}+(1-\delta)
187: \frac{\partial \mu_p}{\partial \rho}\right]. 
188: \end{equation} 
189: We separate the result according to 
190: \begin{equation} 
191: P=P_0+P_{asy}+P_{kin}, 
192: \end{equation} 
193: where $P_0$ is the isospin-independent nuclear interaction contribution 
194: \begin{equation} 
195: P_0=\frac{1}{2}a\rho_0u^{2}+b\frac{\sigma\rho_0}{\sigma+1}u^{\sigma+1}, 
196: \end{equation} 
197: $P_{kin}$ is the kinetic contribution 
198: \begin{equation} 
199: P_{{\rm kin}}=T\rho\{1+\frac{1}{2}
200: \sum_{n=1}^{\infty} b_n(\frac{\lambda_T^{3}\rho}{4})^n\left[(1+\delta)^{n+1} 
201: +(1-\delta)^{n+1}\right]\}, 
202: \end{equation}  
203: and $P_{asy}$ is the contribution from the isospin-dependent 
204: nuclear interaction $V_2$ 
205: \begin{equation} 
206: P_{asy}=(S_0\gamma\rho_0u^{\gamma+1}-8.5\rho_0u^{\frac{5}{3}})\delta^{2}. 
207: \end{equation} 
208: The total pressure $P$ depends on the isospin asymmetry $\delta$ through
209: both the kinetic pressure $P_{kin}$ and the asymmetry pressure $P_{asy}$.
210: We compare the relative values of the three partial pressures in Fig.\ 3 for a
211: typical temperature of 5 MeV with $\gamma=0.5$ and $\delta=$0.2, 0.4 and 0.6, 
212: respectively. It is seen that the pressure increases with the increasing isospin 
213: asymmetry and $P_{asy}$ has an appreciable contribution to the total pressure.
214: The isothermal spinodal (ITS) line moves toward lower densities as the isospin 
215: asymmetry increases. Fig.\ 4 shows in more detail the interplay of temperature T,
216: isospin asymmetry $\delta$ and the $\gamma$ parameter. In each window, 
217: the isospin asymmetry $\delta$ goes from 0 to 1 (bottom to top) with an increment of 0.2. 
218: By increasing the isospin asymmetry the pressure increases as if the temperature
219: is increasing. It is also seen that the $\gamma$ parameter has a small
220: effect on the total pressure especially at high temperature because of the 
221: dominating role of the kinetic pressure. 
222:  
223: \section{Isospin-Dependence of Mechanical and Chemical Instabilities}
224: Based on the {\rm EOS} obtained from various microscopic theories 
225: and phenomenological models\cite{lat78,bar80,muller,liko97,baran98,baran00,cat01}, 
226: it has long been predicted that isospin-asymmetric nuclear matter under certain 
227: conditions can be mechanically or chemically unstable, i.e.,  
228: \begin{equation}\label{mech}  
229: \left(\frac{\partial P}{\partial \rho}\right)_{T,\delta}\leq 0~~~({\rm mechanical}),
230: \end{equation}
231: or
232: \begin{equation}
233: \left(\frac{\partial \mu_n}{\partial \delta}\right)_{P,T}\leq 0~~~({\rm chemical}). 
234: \end{equation} 
235: In these unstable regions, small fluctuations in $\rho$ or $\delta$ respectively 
236: are expected to grow. 
237: 
238: Mechanical instabilities have been well studied in the past for
239: isospin-symmetric systems, here we concentrate on probing its 
240: isospin-dependence. Results presented in Fig. 3 and 4 allow us to explore 
241: effects of the temperature, density and $\gamma$ parameter on not only 
242: the pressure itself, but also how these factors influence the mechanical 
243: instability. The latter happens in the region where 
244: the slope of the pressure with respect to density is negative, 
245: as defined in Eq. \ref{mech}. From results shown in Fig.\ 3 and 4, it is 
246: easy to relate the increase in temperature with a decreasing 
247: probability to observe a mechanical instability. It is seen that the 
248: mechanical instability region shrinks when either the 
249: temperature or the isospin asymmetry increases. Above a critical 
250: temperature $T_c$ determined by the condition
251: \begin{equation} 
252: \left(\frac{\partial P}{\partial\rho}\right)_{T,\delta}= 
253: \left(\frac{\partial^{2}P}{\partial\rho^{2}}\right)_{T,\delta}=0,
254: \end{equation} 
255: the pressure increases monotonically with density, and the mechanical 
256: instability disappears. To be more quantitatively, we show in Fig. 5, 
257: the critical temperature as a function of the isospin asymmetry 
258: $\delta$ with a $\gamma$ parameter of 0.5. It is interesting to 
259: note that the critical temperature decreases with increasing
260: isospin asymmetry. Thus one expects to see a shrinking mechanical instability 
261: region for increasingly more neutron-rich systems. It is important to stress here
262: that the critical temperature $T_c$ discussed above is only associated with the
263: disappearance of mechanical instabilities. For isospin-asymmetric nuclear matter, 
264: as we shall discuss in detail in the following, the system can still be chemically
265: unstable against density fluctuations and undergo fragmentation above 
266: the critical temperature $T_c$. 
267: 
268: To study the chemical instability and its isospin dependence we show in 
269: Fig. 6 the chemical potential isobars $\mu(T,P,\delta)$ as a function of 
270: $\delta$ for neutrons and protons with a  parameter $\gamma=0.5$ at a typical
271: temperature of 5 MeV and pressure of 0.005, 0.02, 0.06, 0.12, 0.2 and 0.8 
272: MeV/fm$^3$, respectively. We note that for corresponding pressures, 
273: the isobars for neutrons and protons begin from the same chemical potential 
274: at $\delta=0$ as one expects. The most interesting feature in this plot is 
275: that there exists an envelope of pressures ($0.005$ MeV/fm$^3$ $<p<$ 0.8 
276: MeV/fm$^3$) inside of which chemical instability occurs over a range of 
277: $\delta$ where  
278: \begin{equation} 
279: \left(\frac{\partial \mu_n}{\partial \delta}\right)\leq 0
280: \end{equation}
281: or
282: \begin{equation}
283: \left(\frac{\partial \mu_p}{\partial \delta}\right)\geq 0.
284: \end{equation} 
285: Inside this region, the system is unstable against isospin fluctuations. 
286: For instance, if several neutrons have migrated into a region of chemical 
287: instability due to some statistical or dynamical fluctuations in a reaction 
288: process, the isospin asymmetry parameter $\delta$ in the region will 
289: increase and the energy of that region will decrease because of its lower neutron 
290: chemical potential. To minimize the total energy of the system, 
291: it is then favorable for the system to have even more neutrons move to the 
292: chemically unstable region, thus leading to the further growth of the isospin 
293: fluctuation. Results at other temperatures show similar features. 
294: However, the boundary of the instability region shrinks as the temperatures T 
295: increase, i.e., for increasing T, chemical instability occurs for much 
296: less isospin asymmetric matter. The temperature dependence and its origin 
297: will be more clearly illustrated in the following. 
298:  
299: Having discussed separately the mechanical and chemical instabilities, we now 
300: compare their relative boundaries in the configure space of 
301: $(T,\rho,\delta)$. The Gibbs-Duhem relation of Eq.\ \ref{gibbs} can be 
302: used to find boundaries of mechanically unstable regions 
303: (ITS: isothermal spinodal) in the
304: $\rho-\delta$ plane for given temperatures. 
305: Since the chemical instability condition has to be evaluated at constant 
306: pressures, the following Maxwellian relation 
307: \begin{equation}\label{trans} 
308: \left(\frac{\partial \mu_n}{\partial \delta}\right)_{T,P}
309: =\left(\frac{\partial \mu_n}{\partial \delta}\right)_{T,\rho}
310: -\left(\frac{\partial \mu_n}{\partial \rho}\right)_{T,\delta} 
311: \cdot\left(\frac{\partial P}{\partial \rho}\right)^{-1}_{T,\delta} 
312: \cdot\left(\frac{\partial P}{\partial \delta}\right)_{T,\rho} 
313: \end{equation} 
314: is used to find boundaries of the chemically unstable 
315: regions (DS: diffusive spinodal).
316: Shown in Fig.\ 7 are the boundaries of the mechanical (thick lines) and 
317: chemical (think lines) instabilities in the $\rho-\delta$ plane with $\gamma=0.5$ 
318: at $T=5$ (lower window), 10 (middle window) and 15 MeV (upper window), 
319: respectively. It is seen that the diffusive spinodal enveloping the region of 
320: mechanical instability extends further out into 
321: the plane; the two regions of mechanical and chemical 
322: instabilities do no overlap. As the temperature increases, both instabilities
323: become less prominent over a more narrow range of densities at smaller isospin 
324: asymmetries. It is thus possible to observe phenomena 
325: due to the chemical instability even in reactions using stable nuclei at 
326: intermediate energies, which can attain local temperatures up to 20 MeV and 
327: isospin asymmetries up to about 0.4\cite{li01}.
328: The above features are independent of the parameters of the {\rm EOS} 
329: and are in good agreement with those based on more microscopic many-body 
330: theories\cite{lat78,bar80,muller}. In particular, we found that the variation of 
331: $\gamma$ parameter has very little effect on the instability boundaries. 
332: Therefore, in the following a constant of $\gamma=0.5$ is used. The results 
333: shown in Fig.\ 7 not only provide the motivations but also serve as a guidance
334: for our following transport model simulations.
335: 
336: \section{Evolution of Density and Isospin Fluctuations in Mechanically or Chemically
337: Unstable Neutron-Rich Matter}
338: Density and isospin fluctuations are expected to grow in the mechanically and/or
339: chemically unstable regions of asymmetric nuclear matter. Information about 
340: the respective growth rates of these fluctuations and how they depend on the 
341: isospin-asymmtry of nuclear matter is currently still rather rare. This 
342: information is vitally important for the structure and stability of both neutron 
343: stars and radioactive nuclei as well as mechanisms of nuclear multifragmentation 
344: in reactions induced by neutron-rich nuclei. In the following, we study the 
345: evolution of fluctuations in isospin-asymmetric nuclear matter initialized in 
346: the mechanically or chemically unstable regions within an isospin-dependent transport 
347: model IBUU\cite{li97,li98}. We note that Baran et al have used a similar approach 
348: \cite{baran98} in studying thermodynamical instabilities which can be identified
349: either as a mechanical or chemical instability\cite{baran00}.
350: 
351: The IBUU model uses consistently the isospin-dependent 
352: {\rm EOS} and the corresponding potentials as in the thermal model outlined 
353: in Section 2. Moreover, scatterings among neutrons and protons are fully 
354: isospin-dependent in terms of their total and differential cross sections 
355: as well as the Pauli blockings. Nucleons are initialized using a Boltzmann 
356: distribution function in a cubic box of length $L_{box}$ with periodic 
357: boundary conditions. The box is further divided into cells of 1 fm$^3$ volume
358: in which the average density $\rho_{cell}$ and isospin asymmetry $\delta_{cell}$
359: are evaluated. We used $10^{4}$ test particles per nucleon in evaluating the
360: $\rho_{cell}$ and $\delta_{cell}$. 
361: 
362: Shown in Fig. 8 is an illustration of the evolution of a system initialized 
363: in the mechanically unstable region with 
364: $T_i=5$ MeV, $\delta_i=0.6$ and $\rho_i=0.05$ fm$^{-3}$. For the purpose of 
365: this illustration a small box of $L_{box}=$10 fm is used.  
366: The most interesting feature shown here is the gradually increasing isospin 
367: fractionation. This is indicated by the spreading of the initial system into 
368: regions with $\rho\leq \rho_i$ and $\delta\geq\delta_i$ and where 
369: $\rho\geq \rho_i$ but $\delta\leq\delta_i$. The degree of isospin fractionation 
370: can be quantified by using the ratio $(N/Z)_{gas}/(N/Z)_{liquid}$, 
371: where $(N/Z)_{gas}$ and $(N/Z)_{liquid}$ is the isospin asymmetry of the low 
372: $(\rho/\rho_0\leq 1/8)$ and high $(\rho/\rho_0> 1/8)$ density regions, 
373: respectively. Shown in Fig.\ 9 are the degree of isospin fractionation as a 
374: function of time for a system initialized with $T_i=$5 MeV at a density of 
375: $\rho_i=0.05$ $fm^{-3}$. As seen from Fig.\ 7, the system is mechanically 
376: unstable with an isospin asymmetry $\delta$ of 0.2 and 0.6, while it is 
377: chemically unstable with $\delta=0.9$. As one expects from the $\delta$ 
378: dependence of the symmetry potential $v_{asy}$, the isospin fractionation 
379: grows faster with the increasing $\delta$.
380: It is also interesting to note that the isospin fractionation happens 
381: later as the $\delta_i$ increases.
382: 
383: The variations of the density $\rho$ and isospin asymmetry $\delta$ with 
384: respect to their initial values can be quantified by using, respectively, 
385: \begin{equation}
386: \sigma_d(t)\equiv(\bar{\rho^2}-\rho_i^2)^{1/2}
387: \end{equation}
388: and 
389: \begin{equation}
390: \sigma_{\delta}(t)\equiv(\bar{\delta^2}-\delta_i^2)^{1/2},
391: \end{equation}
392: where the average is over all cells. Shown in Fig.\ 10 and Fig. 11 
393: are the reduced variation in 
394: isospin asymmetry $\sigma_{\delta}(t)-\sigma_{\delta}(0)$ and 
395: density $\sigma_d(t)/\rho_i$ as a function of time for a system initialized 
396: at $T_i=5$ MeV, $\rho_i=$0.05 fm$^{-3}$ and $\delta_i=$0.2, 0.6 and 0.9,
397: respectively. As a reference, results are also shown for a 
398: system initialized with $T_i=15$ MeV and $\delta_i=0.9$ with the dash-dot 
399: lines (with this initial conditions the system is both mechanically and 
400: chemically stable as shown in Fig.\ 7). As one expects, both the isospin 
401: and density fluctuations stay almost constant and there is no isospin 
402: fractionation at all for this system. While for a system initialized in the 
403: mechanically (chemically) unstable region with $\delta_i=$0.2 and 0.6 
404: ($\delta_i=$0.9), it is interesting to see that both the isospin and 
405: density fluctuations grow {\it faster} with the {\it decreasing} 
406: isospin asymmetry $\delta_i$. This finding is consistent with that found by
407: Baran et al in ref. \cite{baran98}. Thus, the more 
408: neutron-rich system in either the mechanically or chemically unstable regions
409: is more stable against the growth of both density and isospin fluctuations. 
410: This is also responsible for the later start of the isospin fractionation in 
411: the more neutron-rich systems as shown in Fig.\ 9. 
412: 
413: Why is the more neutron-rich matter more stable against both the density and 
414: isospin fluctuations? In the following we try to answer not only this question, but 
415: also explore seeds of fluctuations and their isospin dependence, 
416: and investigate which aspects of nuclear dynamics are important in governing the 
417: growth of fluctuations. The evolution dynamics of asymmetric nuclear matter is 
418: governed by the isospin-dependent nuclear mean filed and stochastic nuclear scatterings. 
419: It is necessary to study their respective roles in generating the fluctuations. 
420: We start with investigating the role of nucleon-nucleon scatterings.
421: In our approach, first-order effects of 2-body stochastic scatterings are included 
422: through the collision integral of the BUU equation. Second-order effects from the 
423: explicitly stochastic correction to the collision integral as in the Boltzmann-Langevin 
424: model are negelected. Thus, t here are two main seeds of fluctuations in our approach, i.e., 
425: the initial numerical fluctuations from the random sampling of the initial state and the later 
426: 2-body stochastic nuclear scatterings. 
427: Both of them may lead to the growth of fluctuations by propagating  
428: through the nuclear mean field in the early and later stages of the evolution, 
429: respectively. The magnitude of the initial numerical fluctuations in both density 
430: and isospin asymmetrty is clearly shown in the upper window of Fig. 8. The seeds of 
431: fluctuations due to 2-body stochastic nuclear scatterings are determined by the 
432: isospin-dependent collision dynamics in asymmetric nuclear matter. 
433: Most importantly, the experimental neutron-neutron cross section, 
434: as shown in Fig.\ 12, is only about $1/3$ of that for neutron-proton 
435: collisions\cite{nndata}, and also the final state for neutron-neutron 
436: scatterings is more strongly Pauli blocked in the more neutron-rich matter. 
437: To be more quantitative, we show in Fig.\ 13 the Pauli blocking rate 
438: as a function of time. As the $\delta_i$ increases from 0.2 to 0.9 the 
439: Pauli blocking rate increases from about 35\% to 50\%. Therefore, there are
440: significantly less nucleon-nucleon collisions in more neutron-rich nuclear
441: matter. Shown in Fig.\ 14 are the average number of 
442: possible (before using the Pauli blocking) nucleon-nucleon collisions 
443: per nucleon $N_{try}/A$ and the successful ones ($N_{coll}/A$) as a function 
444: of time. The value of $N_{try}/A$ reflects directly the isospin-dependent 
445: elementary nucleon-nucleon cross sections shown in Fig.\ 12. 
446: Before about 40 fm/c, there is essentially no collision 
447: because of the strong Pauli blocking when the phase space nonuniformity due to the 
448: initial fluctuation is still small. Later, nuclear scatterings become important, moreover, 
449: they are more frequent with the decreasing isospin asymmetry. With $\delta_i=0.2$ each nucleon 
450: can make about 4 scatterings up to the time of 100 fm/c, while it can only make about 1 
451: collision with $\delta_i=0.9$. Because of the combined effects of both the Pauli blocking and 
452: the isospin-dependent elementary cross sections, the number of nuclear 
453: scatterings in nuclear matter with $\delta_i=0.9$ is consequently only 
454: about 1/6 of that with $\delta_i=0.2$, contributing to the smaller isospin 
455: and density fluctuations in the more neutron-rich matter. 
456: To separate the respective roles of the nuclear mean field and the 2-body scatterings, 
457: we have performed studies by turnning off the nucleon-nucleon collisions in the model. 
458: As shown in the upper window of Fig.\ 15,  the collisional seeds of fluctuations  
459: lead to the significant growth of the isospin fluctuation in the later stage of the evolution.
460: However, they have very little effect on the growth of density fluctuations as shown in the lower window. 
461: A comparison of the results obtained with and without the collision integral 
462: indicates that the growth of fluctuations are overwhelmingly dominated by the nuclear mean field.
463: This is in agreement with the analysis based on the linear response theory by Baran et al\cite{baran98}. 
464: The observed isospin-dependence of the fluctuations can be understood from the interplay 
465: between the {\it attractive isoscalar} mean field and the {\it repulsive symmetry potential for neutrons}. 
466: The symmetry potential, as shown in Eq.\ \ref{vasy}, is repulsive for neutrons and attractive for protons 
467: and their magnitudes increase with the increasing isospin asymmetry. Thus, the resultant attractive mean 
468: field is weaker in the more neutron-rich matter. Because of the small number of scatterings, particularly 
469: in the early stage of the evolution, the growth of fluctuations is thus mainly determined by the strength and 
470: sign of the resultant nuclear mean field according to the linear response theory. Based on the latter, 
471: the magnitude of both fluctuations can grow larger with the increasing strength of the attractive resultant 
472: mean field, and thus also with the decreasing isospin asymmetry $\delta$. 
473: 
474: \section{Summary and Outlook}
475: In summary, utilizing a phenomenological nuclear equation of state within 
476: a thermal model, first we explored the isospin dependence of the 
477: chemical and mechanical instabilities in asymmetric nuclear matter. The diffusive 
478: spinodal is found to extend further out into the $(\delta-\rho)$ plane 
479: than the isothermal spinodal. It is shown that the isospin dependence 
480: of the nuclear equation of state plays a key role in determining the 
481: thermodynamical properties of asymmetric nuclear matter. We also investigated the 
482: evolution of density and isospin fluctuations in mechanically and 
483: chemically unstable asymmetric nuclear matter within a nuclear 
484: transport model. Both the isospin and density fluctuations are found to 
485: decrease with the increasing isospin asymmetry of the system. 
486: Respective roles of the nuclear mean field and the 2-body stochastic 
487: scattering on the evolution of density and isospin fluctuations in either 
488: mechanically or chemically unstable regions of neutron-rich matter are investigated. 
489: It is found that the mean field dominates overwhelmingly the fast growth of both 
490: fluctuations, while the 2-body scattering influences significantly the later growth of the 
491: isospin fluctuation only. We have also shown that the isospin fractionation happens later, 
492: but grows much faster in the more neutron-rich matter. 
493: 
494: It is well known that more neutron-rich systems are less bound and have smaller
495: saturation densities as illustrated by the saturation lines in Fig. 1. However, 
496: fluctuations and their growth are also important 
497: for determining the final state of a neutron-rich system, such as in the 
498: projectile fragmentation in producing exotic beams\cite{fri00}. Our results above 
499: indicate that fluctuations actually have a compensating role to the lower 
500: binding energy in stabilizing neutron-rich system where fluctuations  
501: are smaller and do not grow as fast as in symmetric ones. 
502: Moreover, our results on the isospin fractionation accompanying the evolution 
503: of fluctuations indicate that the configuration of a more 
504: dense, isospin-symmetric region surrounded by a more isospin-asymmetric gas 
505: as in halo nuclei is a natural result of the isospin-dependent nuclear dynamics. 
506: Furthermore, we expect that the multifragmentation and isospin fractionation 
507: in nuclear reactions induced by neutron-rich nuclei to happen on longer time 
508: scales compared to symmetric reactions of the same masses. These expectations 
509: can be tested by measuring products of multifragmentation. In 
510: particular, the measurement of neutron-neutron, proton-proton as well as 
511: fragment-fragment correlation functions, in comparative studies of isospin 
512: symmetric and asymmetric nuclear reactions will be very 
513: useful\cite{robert,bauer}. 
514: 
515: \section{Acknowledgement}
516: We would like to thank V. Baran, M. Colonna, M. Di Toro, A. Evans, J.B. Natowitz, M.B. Tsang 
517: and S.J. Yennello for useful discussions. 
518: This work was supported in part by the National Science Foundation Grant 
519: No. 0088934 and Arkansas Science and Technology Authority Grant No. 00-B-14.
520:                                                        
521: 
522: \newpage \begin{thebibliography}{99} 
523: 
524: \bibitem{rib}A. Mueller and B. Sherril, Annu. Rev. Nucl. Part. Sci.
525: {\bf 43}, 529 (1993); P.G. Hansen, A.S. Jensen and B. Jonson, Ann. Rev. Nucl.
526: Part. Sci. {\bf 45}, 591 (1995); I. Tanihata, Prog. of Part. and Nucl. Phys., 
527: {\bf 35} (1995) 505; W. Nazarewicz, B. Sherril, I. Tanihata and 
528: P. Van Duppen, Nucl. Phys. News {\bf 6}, 17 (1996).
529: 
530: \bibitem{ria}RIA Physics White Paper, 2000, Eds. R. Casten, W. Nazarewicz et al..
531: 
532: \bibitem{li01}Isospin Physics in Heavy-Ion Collisions at Intermediate Energies,
533: Eds. Bao-An Li and W. Udo Schr\"oder, ISBN 1-56072-888-4, 
534: Nova Science Publishers, Inc (2001, New York).
535: 
536: \bibitem{li97}B.A. Li et al., Phys. Rev. Lett. {\bf 76}, 4492 (1996); 
537: {\it ibid}, {\bf 78}, 1644 (1997); B.A. Li, Phys. Rev. Lett. {\bf 85}, 4221 (2000); B.A. Li, 
538: Nucl. Phys. {\bf A681}, 434c (2001).
539: 
540: \bibitem{li98}B.A. Li, C.M. Ko and W. Bauer, 
541: topical review, Int. Jou. Mod. Phys. E{\bf 7}, 147 (1998).
542: 
543: \bibitem{dasgupta}J. Pan and S. Das Gupta, Phys. Rev. C{\bf 57}, 1839 (1998)
544: 
545: \bibitem{tsang01}H.S. Xu et al., Phys. Rev. Lett. {\bf 85}, 716 (2000); 
546: M.B. Tsang et al, Phys. Rev. Lett. {\bf 86}, 5023 (2001). 
547: 
548: \bibitem{gary}G.D. Westfall, Nucl., Phys. {\bf A681}, 343c (2001).
549: 
550: \bibitem{sherry}S.J. Yennello, Nucl., Phys. {\bf A681}, 317c (2001).
551: 
552: \bibitem{wir88}R.B. Wiringa, V. Fiks and A. Fabrocini, 
553: Phys. Rev. C{\bf 38}, 1010 (1988).
554: 
555: \bibitem{brown00}B.A. Brown, Phys. Rev. Lett. {\bf 85}, 5296 (2000).
556: 
557: \bibitem{hor00}C.J. Horowitz et al., Phys.Rev. C{\bf 63}, 025501 (2001).
558: 
559: \bibitem{dito01}M. Colonna et al., Phys. Lett. {\bf B428}, 1 (1998); 
560: M. Di Toro et al., Nucl. Phys. {\bf A681}, 426c (2001).
561: 
562: \bibitem{lom01}U. Lombardo and W. Zuo in ref. \cite{li01}.
563: 
564: \bibitem{bom01}I. Bombaci in ref. \cite{li01}.
565: 
566: \bibitem{lat91}J.M. Lattimer, C.J. Pethick, M. Prakash and P. Haensel, 
567: Phys. Rev. Lett. {\bf 66}, 2701 (1991).
568: 
569: \bibitem{bom94}I. Bombaci, T.T.S. Kuo and U. Lombardo, 
570: Phys. Rep. {\bf 242}, 165 (1994).
571: 
572: \bibitem{sum94}K. Sumiyoshi and H. Toki, Astro. Phys. Journal, {\bf 422}, 
573: 700 (1994); K. Sumiyoshi, H. Suzuki and H. Toki,
574: Astronomy and Astrophysics, {\bf 303}, 475 (1995).
575: 
576: \bibitem{lee96}C.-H. Lee, Phys. Rep. {\bf 275}, 255 (1996).
577: 
578: \bibitem{pra97}M. Prakash et al., Phys. Rep. {\bf 280}, 1 (1997); 
579: J. Lattimer and M. Prakash, Phys. Rep. {\bf 333}, 121 (2000).
580: 
581: \bibitem{moretto}L.G. Moretto and G.J. Wozniak, 
582: Ann. Rev. Nucl. Part. Sci. {\bf 43} (1993) 123.  
583: 
584: \bibitem{bom91}I. Bombaci and U. Lombardo, 
585: Phys. Rev. C{\bf 44}, 1892 (1991).
586: 
587: \bibitem{hub93}H. Huber, F. Weber and M..K. Weigel, 
588: Phys. Lett. {\bf B317}, 485 (1993); Phys. Rev. C{\bf 50}, R1287 (1994).
589: 
590: \bibitem{hei00}H. Heiselberg and M. Hjorth-Jensen, 
591: Phys. Rep. {\bf 328}, 237 (2000).
592: 
593: \bibitem{akm97}A. Akmal and V.R. Pandharipande, Phys. Rev. C{\bf 56}, 2261 (1997);
594: A. Akmal, V.R. Pandharipande and D.C. Ravenhall, Phys. Rev. C{\bf 58}, 1804 (1988).
595: 
596: \bibitem{shl93}S. Shlomo and D.H. Youngblood, Phys. Rev. C{\bf 47}, 529 (1993).
597: 
598: \bibitem{oya98}K. Oyamatsu, I. Tanihata, Y. Sugahara, K. Sumiyoshi and H. Toki,
599: Nucl. Phys. {\bf A634}, 3 (1998).
600: 
601: \bibitem{jaqaman1}H.R. Jaqaman, Phys. Rev. C{\bf 39}, 169 (1988); 
602: {\it ibid} C{\bf 40}, 1677 (1989).  
603: 
604: \bibitem{lat78}J.M. Lattimer and D.G. Ravenhall, 
605: Astr. Jour., {\bf 223}, 314 (1978). 
606: 
607: \bibitem{bar80}M. Barranco and J. R. Buchler, Phys. Rev. C{\bf 22}, 1729 (1980).
608: 
609: \bibitem{muller}H. M\"uller and B.D. Serot, Phys. Rev. C {\bf 52}, 2072 (1995).
610: 
611: \bibitem{liko97}B.A. Li and C.M. Ko, Nuc. Phys. A {\bf 618}, (1997) 498.   
612: 
613: \bibitem{baran98}V. Baran, M. Colonna, M. Di Toro and A.B. Larionov, Nucl. Phys. {\bf A632}, 287 (1998).
614: 
615: \bibitem{baran00} V. Baran, M. Colonna, M. Di Toro and V. Greco, Phys. Rev. Lett. {\bf 86}, 4492 (2001).
616: 
617: \bibitem{cat01} D. Catalano, G. Giansiracusa and U. Lombardo, 
618: Nucl. Phys. {\bf A681}, 390c (2001).
619: 
620: \bibitem{nndata}G. Alkahzov et al., Nucl. Phys. {\bf A280}, 365 (1977).
621: 
622: \bibitem{fri00}W.A. Friedman, M.B. Tsang, D. Bazin and W.G. Lynch,
623: preprint MSUCL-1167, July 2000.
624: 
625: \bibitem{robert}R. Ghetti et al., Nucl. Phys. {\bf A674}, 277 (2000);
626: Phys. Rev. C{\bf 62}, 037603 (2000).
627: 
628: \bibitem{bauer}W. Bauer, C.K. Gelbke and S. Pratt, 
629: Ann. Rev. Nucl. Part. Sci. {\bf 42}, 77 (1992).
630: 
631: 
632: \end{thebibliography}  
633: 
634: \newpage 
635: 
636: \begin{figure}[htp] 
637: \centering \epsfig{file=balifig1.eps,width=12cm,height=12cm,angle=-90} 
638: \caption{The equation of state of isospin-asymmetric nuclear matter with the
639: isospin asymmetry $\delta$ spans between 0 and 1 with a $\gamma$ parameter
640: of 0.5 (lower window) and 1 (upper window), respectively.} 
641: \label{fig1} 
642: \end{figure} 
643: 
644: \begin{figure}[htp] 
645: \centering \epsfig{file=balifig2.eps,width=12cm,height=12cm,angle=-90} 
646: \caption{Symmetry energy (upper window) and potential (lower window) as a 
647: function of reduced density with a $\gamma$ parameter of 0.5, 1.0 and 1.5, 
648: respectively. The symmetry potential is plotted for an isospin asymmetry 
649: $\delta=0.2$.} 
650: \label{fig2} 
651: \end{figure} 
652: 
653: \begin{figure}[htp] 
654: \centering \epsfig{file=balifig3.eps,width=12cm,height=12cm,angle=-90} 
655: \caption{Total and partial pressures as a function of density, $\rho$, for 
656: isospin asymmetry $\delta=0.2,0.4$ and $0.6$ with a temperature $T=5$ MeV 
657: and $\gamma=0.5$. The isothermal spinodal line (ITS) is indicated with arrows.} 
658: \label{fig3} 
659: \end{figure} 
660: 
661: \begin{figure}[htp] 
662: \centering \epsfig{file=balifig4.eps,width=12cm,height=12cm,angle=-90} 
663: \caption{Total pressure P as a function of density $\rho$ for 
664: isospin asymmetry $\delta=0.0,0.2,0.4,0.6,0.8$ and $1.0$ (from bottom to top)
665: with a temperature $T=5, 10$ and $15$ MeV and $\gamma=0.5, 1.0$ and 1.5, 
666: respectively.} 
667: \label{fig4} 
668: \end{figure} 
669: 
670: \begin{figure}[htp] 
671: \centering \epsfig{file=balifig5.eps,width=12cm,height=12cm,angle=-90} 
672: \caption{Critical temperature as a function of isospin asymmetry 
673: $\delta$.} 
674: \label{fig5} 
675: \end{figure}  
676: 
677: \begin{figure}[htp] 
678: \centering \epsfig{file=balifig6.eps,width=12cm,height=12cm,angle=-90} 
679: \caption{Chemical potential isobars, $\mu$ for neutrons and protons as a function 
680: of isospin asymmetry, $\delta$ at temperature, T=5 MeV and density parameter, 
681: $\gamma=0.5$. The isobars are at pressures of $P=0.005, 0.02, 0.06, 0.12, 0.20$ and $0.80$ 
682: MeV/fm$^3$, respectively.} 
683: \label{fig6}  
684: \end{figure} 
685: 
686: \begin{figure}[htp] 
687: \centering \epsfig{file=balifig7.eps,width=15cm,height=13cm,angle=-90} 
688: \caption{Boundaries of the mechanical (thick lines) and chemical (thin lines)
689: in the density-isospin asymmetry plane at a temperature of 5, 10 and 15 MeV, 
690: respectively. }
691: \label{fig7} 
692: \end{figure} 
693: 
694: \begin{figure}[htp] 
695: \centering \epsfig{file=balifig8.eps,width=15cm,height=13cm,angle=-90} 
696: \caption{An illustration of the evolution of asymmetric nuclear matter
697: in the density-isospin asymmetry plane. Each point in the scatter plots 
698: represent one cell of 1 fm$^3$ volume in a cubic box of side length 
699: $L_{box}=10$ fm.}
700: \label{fig8}
701: \end{figure}   
702: 
703: \begin{figure}[htp] 
704: \centering \epsfig{file=balifig9.eps,width=12cm,height=12cm,angle=-90} 
705: \caption{Evolution of the degree of isospin fractionation in a cubic box of side 
706: length $L_{box}=30$ fm and density 0.05 fm$^{-3}$, temperature $T_i=5$ MeV 
707: and $\delta_i=$0.2, 0.6 and 0.9, respectively.}   
708: \label{fig9}
709: \end{figure} 
710: 
711: \begin{figure}[htp] 
712: \centering \epsfig{file=balifig10.eps,width=12cm,height=12cm,angle=-90} 
713: \caption{Evolution of the reduced density fluctuation in a cubic box of side 
714: length $L_{box}=30$ fm and density 0.05 fm$^{-3}$. The dash-dot lines 
715: are calculated with $T_i=15$ MeV and $\delta_i=0.9$; while the solid 
716: lines are calculated with $T_i=5$ MeV and $\delta_i=$0.2, 0.6 and 0.9, 
717: respectively.}
718: \label{fig10}   
719: \end{figure}  
720:  
721: \begin{figure}[htp] 
722: \centering \epsfig{file=balifig11.eps,width=12cm,height=12cm,angle=-90} 
723: \caption{Evolution of the reduced isospin fluctuation for the same systems
724: as in Fig.\ 10.}
725: \label{fig11}   
726: \end{figure}  
727: 
728: \begin{figure}[htp] 
729: \centering \epsfig{file=balifig12.eps,width=12cm,height=12cm,angle=-90} 
730: \caption{Experimental neutron-neutron and neutron-proton cross sections 
731: as a faction of beam energy.}
732: \label{fig12}
733: \end{figure}  
734: 
735: \begin{figure}[htp] 
736: \centering \epsfig{file=balifig13.eps,width=12cm,height=12cm,angle=-90} 
737: \caption{The Pauli blocking rate as a function of time in a cubic box of 
738: $L_{box}=30$ fm with density 0.05 fm$^{-3}$, temperature $T_i=5$ MeV and 
739: $\delta_i=$0.2, 0.6 and 0.9, respectively.}
740: \label{fig13}
741: \end{figure} 
742: 
743: \begin{figure}[htp] 
744: \centering \epsfig{file=balifig14.eps,width=12cm,height=12cm,angle=-90} 
745: \caption{The average number of possible nucleon-nucleon collisions 
746: per nucleon (a) and the average number of successful nucleon-nucleon 
747: collisions per nucleon (b) as a function of time for the same systems as 
748: in Fig. 13.}
749: \label{fig14}
750: \end{figure} 
751: 
752: \begin{figure}[htp] 
753: \centering \epsfig{file=balifig15.eps,width=12cm,height=12cm,angle=-90} 
754: \caption{Evolution of isospin (upper window) and density (lower window)
755: fluctuations as a function of time. The solid and long dashed lines are
756: results of full calculations, while the dotted and dash-dot lines are 
757: results obtained by turnning off the 2-body scatterings.}
758: \label{fig15}
759: \end{figure} 
760: 
761: \end{document} 
762: 
763: 
764: 
765: 
766: 
767: 
768: