1: \documentstyle[aps,preprint]{revtex}
2: %\documentstyle[12pt]{article}
3: \renewcommand{\baselinestretch}{1.5}
4: \textheight=8.1 true in
5: \textwidth=5.9 true in
6: \begin{document}
7: \begin{center}
8: {\Large{\bf Liquid-gas phase transition and its order in finite nuclei}}\\
9:
10: \vskip 1.0cm
11: Tapas Sil$^1$, S. K. Samaddar$^1$, J. N. De$^2$ and S. Shlomo$^3$\\
12: $^1$Saha Institute of Nuclear Physics, 1/AF Bidhannagar, Kolkata 700064, India\\
13: $^2$Variable Energy Cyclotron Centre, 1/AF Bidhannagar, Kolkata 700064, India\\
14: $^3$The Cyclotron Institute, Texas A$\&$M University, College Station,\\
15: Texas 77845,USA
16: \end{center}
17:
18:
19:
20: \begin{abstract}
21: The liquid-gas phase transition in finite nuclei is studied in a heated
22: liquid-drop model where the drop is assumed to be in thermodynamic
23: equilibrium with the vapour emanated from it. Changing pressure along
24: the liquid-gas coexistence line of the systems, symmetric or asymmetric,
25: suggests that the phase transition is a continuous one. This is further
26: corroborated from the study of the thermal evolution of the entropy
27: at constant pressure.
28: \end{abstract}
29:
30: \vskip 1.0cm
31: PACS Number(s): 25.70.-z, 21.65.+f, 24.10.Pa
32: \newpage
33: %%%%%%%%%%%%%%%%%%%
34: The study of liquid-gas phase transition in finite nuclear systems is
35: of considerable contemporary interest
36: \cite{poc,hau,de1,nat1,lee,ell1,gul,das,paw}. Experimental analyses of
37: the accumulated data on multifragmentation and caloric curves show
38: compelling evidence of such a transition. Phase transitions are
39: normally signalled by peaks in the specific heat at constant volume
40: $C_V$ with rise in temperature. Theoretical models of different
41: genres, such as the microcanonical \cite{gro} or the canonical
42: \cite{bon} description of multifragmentation, the lattice-gas model
43: \cite{gul,das,sam} or even the microscopic treatment in a relativistic
44: \cite{sil} or a nonrelativistic \cite{de1} Thomas-Fermi framework
45: support such a structure in the heat capacity. A clear idea about
46: the subtle details of the liquid-gas phase transition in finite
47: nuclei, however, has not emerged yet. Confusion remains about whether
48: the system has evolved dynamically through the critical point
49: \cite{ell1,nat2,rad}; a coherent picture about the
50: order of the phase transition is also missing. Analyses
51: of the EOS group \cite{ell2} and the ISiS group \cite{kle} in the scaling
52: model give strong circumstantial evidence for a continuous
53: (second order) phase
54: transition. Calculations performed in a mean field model \cite{lee}
55: also lead to a similar conclusion.
56: Predictions from the lattice-gas models
57: \cite{gul,pan} are, however, compatible with a first order transition.
58:
59: Symmetric infinite nuclear matter is effectively a one-component
60: system; with heating, it undergoes a first order phase transition.
61: On the other hand, in case of two-component asymmetric nuclear matter,
62: as was shown in a comprehensive analysis by M$\ddot u$ller and
63: Serot \cite{mul} in a relativistic mean-field framework,
64: the separate conservation of the neutron and
65: proton number densities leads the system to a continuous phase transition
66: over a finite temperature interval.
67: This is also supported in the nonrelativistic calculations by Kolomietz
68: {\it et al} \cite{kol}.
69: Unlike symmetric nuclear matter, even
70: a finite symmetric nucleus ($N=Z$) behaves like a two-component system
71: as the Coulomb interaction lifts the isospin degeneracy. A finite
72: nucleus is then expected to undergo a continuous liquid-gas phase
73: transition, if at all. The conflicting predictions from the
74: previous model analyses do not have realistic inputs of nuclear
75: physics as relevant for a quantum system of interacting fermions
76: with a short-range nuclear force as also with the long-range Coulomb
77: interaction. In this communication, we focus on the nature of
78: the phase transition once specific features concerning an atomic
79: nucleus are properly taken into account.
80:
81: For our study, we choose a representative system, namely,
82: Rhenium with $A$= 186 and $Z$= 75. Such a nucleus is likely to be
83: formed from the reaction $^{124}Sn+^{124}Sn$ at energies
84: around 50 MeV per nucleon \cite{xu} after some nucleons have left
85: out from the reaction zone as preequilibrium particles. We also investigate
86: $^{150}Re$ to explore the isospin asymmetry effects in phase transition.
87: The nucleus is viewed as a spherical liquid drop with asymmetry
88: $X_0$ defined as $(N_0 - Z_0)/A_0$ where $A_0= N_0+Z_0$, is the mass
89: number of the total system. At a finite temperature, we assume the
90: depleted nucleus to be enveloped by its own vapour and the system
91: to be in complete thermodynamic equilibrium conserving the total
92: number of neutrons and protons. The nucleon distributions in the
93: liquid and gas are assumed to be uniform in each phase.
94: This definition allows
95: to explore the liquid-gas coexistence region for a finite system
96: in close analogy with bulk nuclear matter.
97:
98: In absence of a well-defined way to write the energy density functional of a
99: finite nucleus in terms of volume, surface, symmetry and Coulomb terms, we
100: write the free energy of the nuclear system at temperature $T$ in the single
101: phase as
102: \begin{equation}
103: F=A_0f_{nm}(\rho,X_0)+F_C+F_{surf}
104: \end{equation}
105: where $f_{nm}(\rho,X_0)$ is the free energy per particle of infinite
106: nuclear matter at density $\rho$ with asymmetry $X_0$ at the same
107: temperature $T$, $F_C$ the Coulomb free energy and $F_{surf}$ the
108: temperature and asymmetry dependent surface free
109: energy. In the liquid-gas coexistence region, the free energy is given by
110: \begin{equation}
111: F_{co}=F^l+F^g,
112: \end{equation}
113: where the liquid free energy is,
114: \begin{equation}
115: F^l= A^l f_{nm}(\rho^l,X^l)+F_C^l+ F_{surf}^l,
116: \end{equation}
117: and the free energy of the emanated gas is
118: \begin{equation}
119: F^g= A^g f_{nm}(\rho^g,X^g)+F_C^g.
120: \end{equation}
121: Here $A^l$ and $A^g$ are the number of nucleons in the liquid and the gas
122: phase, $\rho^l$, $\rho^g$ and $X^l$, $X^g$ are the corresponding density
123: and asymmetry. The free energy of infinite nuclear matter is evaluated in
124: the finite temperature Thomas-Fermi framework with a modified
125: Seyler-Blanchard interaction \cite{de2}. The Coulomb free energies for the
126: liquid and the gas, $F_C^l$ and $F_C^g$ are calculated corresponding to a
127: uniform charged sphere and a spherical-shell, respectively. For simplicity,
128: their mutual interaction is neglected. The surface
129: free energy of the liquid part $F_{surf}^l$ is taken as \cite{lev}
130: \begin{equation}
131: F_{surf}^l=\sigma(X^l,T)(A^l)^{2/3},
132: \end{equation}
133: where the surface energy coefficient is
134: \begin{equation}
135: \sigma(X,T)=[\sigma(X=0)-a_sX^2][1+1.5 T/T_c][1-T/T_c]^{3/2}
136: .
137: \end{equation}
138: Here, $\sigma(X=0)=18$ MeV, $a_s=28.66$ MeV and $T_c$, the critical
139: temperature of the symmetric nuclear matter is 15 MeV. The surface energy
140: coefficient decreases with density; we neglect its density dependence for
141: the liquid part. Since the gas density is very low, its surface energy is
142: neglected. The total surface energy of the liquid is
143: \begin{equation}
144: E_{surf}^l=F_{surf}^l+TS_{surf}^l
145: \end{equation}
146: where the total surface entropy is given by
147: \begin{equation}
148: S_{surf}^l=-\left(\frac{\partial F_{surf}^l}{\partial T}\right)_V
149: .
150: \end{equation}
151: The total entropy $S_0$ of the system is then calculated as
152: \begin{equation}
153: S_0=A^ls_{nm}^l(\rho^l,X^l)+A^gs_{nm}^g(\rho^g,X^g)+S_{surf}^l
154: .
155: \end{equation}
156: The per particle entropy $s_{nm}$ of homogeneous nuclear matter is obtained
157: from the standard mean-field prescription.
158:
159: The chemical potentials $\mu_n^l$ and $\mu_z^l$ for neutron and proton in
160: the liquid are given by
161: \begin{eqnarray}
162: \mu_n^l&=&\frac{\partial F^l}{\partial N^l},\nonumber\\
163: \mu_z^l&=&\frac{\partial F^l}{\partial Z^l}.
164: \end{eqnarray}
165: The liquid pressure is obtained from
166: \begin{equation}
167: P=\rho^2\frac{\partial (F/A)}{\partial \rho}.
168: \end{equation}
169: Similar equations follow for the gas phase.
170: For thermodynamic equlibrium between the liquid and the gas, the two
171: chemical potentials and the pressures in both phases must be the same,
172: i.e, $\mu_n^l=\mu_n^g,\;
173: \mu_z^l=\mu_z^g$ and $P^l=P^g$.
174: Along the coexistence line, the mass of the liquid drop changes. For a
175: chosen $A^l$, the quantities $\rho^l$, $\rho^g$, $X^l$, $X^g$ and $A^g$ are
176: determined by exploiting the three thermodynamic equlibrium conditions and
177: the constraints of baryon number and the total isospin conservation:
178: \begin{eqnarray}
179: \rho_{n,z}&=&\lambda\rho_{n,z}^l + (1-\lambda)\rho_{n,z}^g\nonumber,\\
180: \rho X_0&=&\lambda\rho^lX^l+(1-\lambda)\rho^gX^g.
181: \label{iso}
182: \end{eqnarray}
183: Here $\rho$ is the average nucleon density and $\lambda$ is the liquid
184: volume fraction.
185:
186:
187: The isotherms for the nucleus $^{186}Re$ at $T=7,8,9$ and 10 MeV are
188: displayed in Fig.1. For comparison, the isotherm for nuclear matter with
189: asymmetry same as that of $^{186}Re$ is also shown at $T=10$ MeV. The
190: difference between the isotherms for the infinite and the finite system is
191: not insignificant. Though the Coulomb and surface have opposing effects on
192: the pressure, the former wins over the latter at this temperature. The
193: liquid-gas coexistence lines for $^{186}Re$ for the four temperatures
194: mentioned are shown by the dotted lines. It is seen that the pressure
195: changes along the coexistence line, as seen earlier in the case of
196: asymmetric nuclear matter \cite{mul}; the slope of the coexistence lines
197: also increases with temperature. It is further noted that at a given
198: temperature and asymmetry, because of the Coulomb effect, the slope of the
199: coexistence line of a finite nucleus is more compared to that of asymmetric
200: nuclear matter. The variation of pressure along the coexistence line is a
201: pointer to a continuous phase transition. Unlike nuclear matter,
202: the coexistence lines do not extend from the pure gas phase to the pure
203: liquid phase. At relatively lower temperatures, it is found that
204: as the system expands, the size of the liquid drop depletes
205: and reaches a minimum
206: mass beyond which no thermodynamic equilibrium is possible;
207: at higher temperatures,
208: with compression the liquid drop attains a limiting mass which decreases with
209: increasing temperature as is evident from the figure. As an example, the
210: minimum liquid-drop mass at $T=7$ MeV is $A^l=24$ (marked as
211: A in the figure); at $T=10$ MeV,
212: the limiting liquid-drop mass is $A^l=130$ (marked as B). This implies that at
213: lower temperatures, a gas of a finite number of nucleons when compressed
214: start nucleating with a minimum mass for the seed in order to remain in
215: thermodynamic equilibrium. Similarly, at higher temperatures, for the
216: coexisting finite system, the evaporated gas should contain a minimum
217: number of nucleons.
218:
219: The isospin fractionation along the coexistence line for the nucleus
220: $^{186}Re$ at $T=8 $ MeV is shown in Fig.2. The system has proton fraction
221: $Y_0=0.403$ (defined as $Y=Z/A$).
222: As the system prepared in the gaseous phase is compressed, the
223: two-phase region is encountered at the point A with the emergence of a
224: minimum liquid mass ($A^l=20$) at the point B with
225: a proton fraction $Y_B$ larger than
226: $Y_0$. With further compression,
227: the gas phase depletes from
228: A to C while the liquid phase grows from B to D attaining the total
229: mass $A_0$ and proton fraction $Y_0$. During compression, the proton
230: fractions in both phases decrease, but the total proton fraction $Y_0$
231: remains fixed, as dictated by the conservation of the total isospin
232: given by Eq.(\ref{iso}).
233: It is evident from the figure that the gas phase is more neutron-rich
234: compared to the total system while the liquid phase is comparatively
235: neutron-deficient as also observed experimentally \cite{xu}.
236: This feature becomes more prominent with increasing
237: liquid mass. In order to explore the asymmetry effect on isospin
238: fractionation in phase transition, the calculated results for the symmetric
239: system $^{150}Re$ are also displayed in the figure. Contrary to the
240: relatively neutron-rich nucleus $^{186}Re$ ,
241: here it is found that the gas phase is
242: proton-rich. In symmetric nuclei, since $\mu_z$ is greater than $\mu_n$
243: because of Coulomb interaction, the separation of the gas phase behaves
244: more like that of proton-rich nuclear matter. The occurrence of liquid-gas
245: phase transition in symmetric medium-heavy nuclei should then lead to
246: preponderance of proton-rich isotopes in energetic heavy-ion collisions and
247: this can be tested in experiments.
248:
249: The heat capacity per particle $C_V$ for $^{186}Re$
250: at constant volume (defined as $\left(\frac{d(E^*/A_0)}{dT}\right)_V$,
251: where $E^*$ is the total excitation energy of the
252: system) is displayed in Fig.3 at a representative volume $V=10V_0$ which
253: can be interpreted as a {\it freeze-out} volume. Here $V_0$ is the normal
254: volume of the nucleus calculated with the radius parameter $r_0=1.16$ fm.
255: A very broad bump in $C_V$
256: with a maximum at $T\sim 10 $ MeV is seen. The system then corresponds
257: to a liquid part with $A^l$ around 80, the rest of the nucleons
258: being in the gas phase. This is
259: contrary to the results in the microscopic mean-field calculations obtained
260: earlier \cite{de1,sil} at around the same freeze-out volume, where a much
261: sharper peak was observed when the system
262: just vaporises completely. At very high temperature, the heat capacity
263: saturates at 1.5 corresponding to a pure classical gas.
264:
265: The thermal evolution of the entropy per particle for the nucleus
266: $^{186}Re$ at a constant pressure $P=0.06$ MeV fm$^{-3}$ is shown by the
267: solid line in figure 4. In contrast to a first order phase transition where
268: the entropy at constant pressure exhibits a discontinuity at a particular
269: temperature (the phase transition temperature), here the entropy change is
270: continuous. The noticeable rise in entropy in the temperature range 8.4 to
271: 10.3 MeV is the manifestation of a liquid-gas phase transition in this
272: temperature domain for the chosen pressure.
273: The dashed line in the figure corresponds to the
274: entropy evolution for the symmetric nucleus $^{150}Re$. This is quite
275: similar to that of the asymmetric nucleus considered, but markedly
276: different from that of symmetric nuclear matter \cite{mul}
277: which shows a sharp discontinuity at the transition temperature.
278: The striking similarity between the two results shows that the finite size
279: and Coulomb effects are the dominant factors in determining the nature of
280: phase transition in finite nuclei.
281: The continuous change of entropy at a constant pressure plead in favour of the
282: characterisation of the liquid-gas phase transition in atomic nuclei,
283: symmetric or asymmetric, as a continuous one.
284:
285: The liquid-gas phase transition in finite nuclei with explicit conservation
286: of the baryon number and the total isospin has been investigated in this
287: communication in a heated liquid drop model.
288: The peaked structure in the heat capacity,
289: though broad, signals the occurrence of a liquid-gas phase transition. From
290: the evolution of entropy at constant pressure, one sees that the transition
291: occurs over a range of temperatures; this strongly suggests that the
292: liquid-gas phase transition in a finite nuclear system is continuous.
293: The simplified assumptions in the model may affect the results somewhat
294: quantitatively, but the general qualitative features are expected
295: to remain unaltered. The thermodynamic concepts may not be very
296: meaningful when the number of particles in one of the phases is very
297: small, still this model serves as a window to understand the basic
298: features of liquid-gas phase transition in finite systems.
299:
300: \newpage
301: \begin{thebibliography}{99}
302: \bibitem{poc} J. Pochodzalla {\it et al},
303: Phys. Rev. Letts. {\bf 75}, 1040 (1995).
304: \bibitem{hau} J. A. Hauger {\it et al},
305: Phys. Rev. Lett. {\bf 77}, 235 (1996).
306: \bibitem{de1} J. N. De, S. Dasgupta, S. Shlomo, and S. K. Samaddar,
307: Phys. Rev. {\bf C55}, R1641 (1997).
308: \bibitem{nat1} J. B. Natowitz {\it et al},
309: Phys. Rev. {\bf C65}, 034618 (2002).
310: \bibitem{lee} S. J. Lee, and A. Z. Mekjian, Phys. Rev. {\bf C63}, 044605
311: (2001).
312: \bibitem{ell1} J. B. Elliot {\it et al} Phys. Rev. Letts. {\bf 88}, 042701
313: (2002).
314: \bibitem{gul} F. Gulminelli, and P. Chomaz, Phys. Rev. Letts. {\bf 82}, 1402
315: (1999).
316: \bibitem{das} S. Dasgupta, A. Z. Mekjian, and M. B. Tsang, (to be published
317: in Adv. in Nucl. Phys. eds, J. Negele, and E. Vogt).
318: \bibitem{paw} P. Pawlowski, Phys. Rev. {\bf C65}, 044615 (2002).
319: \bibitem{gro} D. H. E. Gross, Rep. Prog. Phys. {\bf 53}, 605 (1990).
320: \bibitem{bon} J. P. Bondorf, A. S. Botvina, A.S. Illjinov, I. N. Mishustin,
321: and K. Sneppen, Phys. Rep. {\bf 257}, 130 (1995).
322: \bibitem{sam} S. K. Samaddar, and S. Dasgupta, Phys. Rev.
323: {\bf C61}, 034610 (2000).
324: \bibitem{sil} Tapas Sil, B. K. Agrawal, J. N. De, and S. K. Samaddar,
325: Phys. Rev. {\bf C63}, 054604 (2001).
326: \bibitem{nat2} J. B. Natowitz, K. Hagel, Y. Ma, M. Murray, L. Qin, S.
327: Shlomo, R. Wada, and J.Wang, nucl-ex/0206010.
328: \bibitem{rad} Al. H. Raduta, Ad. R. Raduta, P. Chomaz, and F. Gulminelli,
329: Phys. Rev. {\bf C65}, 034606 (2002).
330: \bibitem{ell2} J. B. Elliot {\it et al} Phys. Letts. {\bf B381}, 35 (1996).
331: \bibitem{kle} M. Kleine Berkenbusch {\it et al}, Phys. Rev. Letts.{\bf 88},
332: 022701 (2002).
333: \bibitem{pan} J. Pan, S. Dasgupta, and M. Grant, Phys. Rev. Letts., {\bf
334: 80}, 1182 (1998)
335: \bibitem{mul} H. M\"uller, and B. D. Serot, Phys. Rev. {\bf C 52},
336: 2072 (1995).
337: \bibitem{kol} V. M. Kolomietz, A. I. Sanzhur, S. Shlomo, and S. A. Firin,
338: Phys. Rev. {\bf C 64}, 024315 (2001).
339: \bibitem{xu} H. S. Xu {\it et al}, Phys. Rev. Letts., {\bf 85}, 716 (2000).
340: \bibitem{de2} J. N. De, N. Rudra, Subrata Pal, and S. K. Samaddar,
341: Phys. Rev. {\bf C 63}, 780 (1996).
342: \bibitem{lev} S. Levit, and P. Bonche, Nucl. Phys. {\bf A437}, 426 (1985).
343: \end{thebibliography}
344: %%%
345: \newpage
346: \centerline{\bf Figure Captions}
347: \begin{itemize}
348:
349: \item[Fig.\ 1] The isotherms for the system $^{186}Re$ at different
350: temperatures as labelled on each curve are shown as full lines.
351: The liquid-gas coexistence lines are shown by the dotted lines.
352: The long-dashed line refers to the isotherm at $T=$ 10 MeV for
353: nuclear matter with asymmetry same as that of $^{186}Re$. The corresponding
354: coexistence line is shown by the filled circles.
355:
356: \item[Fig.\ 2] Evolution of proton fraction ($Y$) along the liquid-gas
357: coexistence line at $T=$ 8 MeV for the systems as shown.
358:
359:
360: \item[Fig.\ 3] The specific heat capacity $C_V$ as a function of
361: temperature for $^{186}Re$ in a freeze-out volume 10$V_0$.
362:
363:
364: \item[Fig.\ 4] Entropy per particle as a function of temperature at
365: a constant pressure for the systems $^{186}Re$ and $^{150}Re$.
366:
367:
368: \end{itemize}
369: \end{document}
370: