nucl-th0601018/farhan2
1: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
2: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3: 
4: % Some other (several out of many) possibilities
5: %\documentclass[preprint,aps]{revtex4}
6: %\documentclass[preprint,aps,draft]{revtex4}
7: %\documentclass[prc]{revtex4}% Physical Review C
8: 
9: \usepackage{graphicx}% Include figure files
10: \usepackage{dcolumn}% Align table columns on decimal point
11: \usepackage{bm}% bold math
12: 
13: %\nofiles
14: 
15: \begin{document}
16: 
17: %\preprint{APS/123-QED}
18: 
19: \title{The strength of nuclear shell effects at $N=126$ in the r-process 
20: region}
21: 
22: \author{A.R. Farhan and M.M. Sharma}
23:  \affiliation{Physics Department, Kuwait University, Kuwait 13060}
24: 
25: 
26: \date{\today}% It is always \today, today,
27:              %  but any date may be explicitly specified
28: 
29: \begin{abstract}
30: We have investigated nuclear shell effects across the magic number
31: $N=126$ in the region of the r-process path. Microscopic calculations 
32: have been performed using the relativistic Hartree-Bogoliubov (RHB) 
33: approach within the framework of the Relativistic Mean-Field (RMF) 
34: theory for isotopic chains of rare-earth nuclei in the r-process 
35: region. The Lagrangian model NL-SV1 with the inclusion of the vector 
36: self-coupling of $\omega$ meson has been employed. The RMF results show 
37: that the shell effects at $N=126$ remain strong and exhibit only a slight 
38: reduction in the strength in going from the r-process path to the neutron 
39: drip line. This is in striking contrast to a systematic weakening 
40: of the shell effects at $N=82$ in the r-process region predicted 
41: earlier in the similar approach. In comparison the shell effects with 
42: microscopic-macroscopic mass formulae show a near constancy of 
43: shell gaps leading to strong shell effects in the region of 
44: r-process path to the drip line. A recent analysis of solar-system 
45: r-process abundances in a prompt supernova explosion model 
46: using various mass formulae including the recently introduced mass 
47: tables based upon Hartree-Fock-Bogoliubov method shows that 
48: whilst mass formulae with weak shell effects at $N=126$ give rise 
49: to a spread and an overproduction of nuclides near the third
50: abundance peak at $A \sim 190$, mass tables with droplet models showing 
51: stronger shell effects are able to reproduce the abundance features 
52: near the third peak appropriately. In comparison, several analyses of 
53: the second r-process peak at $A \sim 130$ have required weakened 
54: (quenched) shell effects at $N=82$. Our predictions in the RMF theory 
55: with NL-SV1, which exhibit weaker shell effects at $N=82$ and correspondingly
56: stronger shell effects at $N=126$ in the r-process region, support the 
57: conjecture that a different nature of the shell effects at the magic 
58: numbers may be at play in r-process nucleosynthesis of heavy nuclei.
59: 
60: 
61: \end{abstract}
62: 
63: \pacs{~21.10.Dr, 21.30.-x, 21.60.-n, 25.30.+k, 26.50.+x}% PACS, the Physics and Astronomy
64:                              % Classification Scheme.
65: %\keywords{Suggested keywords}%Use  showkeys class option if keyword
66:                               %display desired
67: \maketitle
68: 
69: \section{Introduction}
70: \label{intro}
71: About half of nuclei heavier than Fe are synthesized in the process
72: of rapid neutron capture (r-process) 
73: \cite{BBFH.57,Hille.78,Cowan.91,Kratz.93,Kratz.00}. 
74: In environments of high neutron densities and high temperatures, 
75: extremely neutron-rich nuclei with at least
76: 10-30 neutrons away from the stability line are produced.
77: These nuclei are highly unstable and experimentally 
78: inaccessible, especially those in the heavy mass region. 
79: The ensuing nuclei undergo a sequence of neutron capture 
80: accompanied by a spate of $\beta^-$ decays thus
81: leading to formation of heavy elements in nature. 
82: The r-process path passes through the magic numbers $N=$~50, 82 and 
83: 126 at different mass values. The synthesis of nuclei around these 
84: magic numbers is reflected vividly in the known nuclear abundance
85: peaks around  $A \sim$ 80, 130 and 190, respectively. 
86: 
87: The shell effects at the magic numbers play a significant role in
88: determining the r-process nuclear abundances \cite{Kratz.93}.
89: The question whether the shell effects near the r-process 
90: path are strong or do quench has become crucial to understanding 
91: the nucleosynthesis of heavy nuclei \cite{Pfeiffer.01}.
92: This question remains open to-date and the existing data do not suffice
93: to answer this question. This is due to the reason that r-process 
94: nuclei are extremely neutron rich and are not accessible experimentally. 
95: Moreover, on the basis of a few nuclei that are known in the 
96: extreme regions, it is not easy to make reliable predictions  in
97: the farther regions of the period table. Consequently, it is proving 
98: to be difficult to ascertain the nature of the shell effects 
99: in the vicinity of the r-process path. A knowledge about 
100: properties of these nuclei is, therefore, obtained from theoretical 
101: models. At the same time, new data in unknown regions are being
102: obtained experimentally. Such data can be of enormous value in
103: defining the nature of properties of nuclei in the extreme regions.
104: 
105: In principle, microscopic calculations within a reliable model 
106: would be attractive for the purpose. The primary condition on utility
107: of a microscopic framework should be its ability to reproduce features
108: and properties of nuclei in the known domains with better accuracy.
109: Calculations within a microscopic model for a considerably large 
110: number of nuclei can be cumbersome. However, with the progress in 
111: computing speeds, such a task is no more beyond one's reach. 
112: 
113: Heretofore, macroscopic-microscopic approaches have largely been used to 
114: calculate and extrapolate properties of nuclei in the inaccessible 
115: regions. Most prominent amongst these is the approach of the 
116: Finite-Range Droplet Model (FRDM) \cite{Moeller.95}. The mass formula 
117: FRDM has been obtained on the basis of extensive fits of more than 
118: a thousand known nuclei across the period table including those 
119: discovered at the periphery of the periodic table in the last decade. 
120: R-process calculations have been performed using the
121: binding energies (masses) and neutron separation energies from 
122: the FRDM in conjunction with $\beta$-decay properties obtained in the 
123: Quasi-Particle Random Phase Approximation (QRPA) \cite{Moeller.90}. 
124: Another mass formula that has also been employed extensively is based 
125: upon the Extended Thomas-Fermi model with Strutinsky Integral (ETF-SI) 
126: \cite{Abou.95}. Herein, the liquid drop (or the smooth) part is provided 
127: by the Extended Thomas-Fermi model and the shell corrections are 
128: superimposed thereupon using the method of the Strutinsky shell 
129: correction \cite{Stru.68}.
130: 
131: Employing the results of these two mass formulae extensive r-process
132: network chain calculations have been undertaken
133: \cite{Kratz.93,Kratz.00,Pfeiffer.01}. It was concluded that
134: due to strong shell effects (gaps) at $N=82$ and at $N=126$ in 
135: the region of the r-process path inherent in the mass models 
136: FRDM and ETF-SI, strong deficiencies (troughs) are obtained in 
137: reproducing solar system r-process abundances below the peaks 
138: $A \sim 130$ and $A\sim 190$ \cite{Kratz.93,Pfeiffer.01}.
139: A remedial measure was suggested by the Hartree-Fock-Bogoliubov 
140: (HFB) calculations \cite{Doba.96} using the Skyrme force SkP. 
141: This force is shown to quench (weaken) the strength of the
142: shell effects in the r-process region significantly, especially at $N=82$. 
143: This feature of HFB+SkP has proved to be useful in filling up the troughs
144: below $A \sim 130$ and $A\sim 190$ in the r-process abundance curve.
145: In this approach, the quenching effect is attributed mainly to a large
146: effective mass $m*=1$ of the force SkP. Inspired by the usefulness of
147: the quenching along the r-process path, contrary to the original feature
148: of both the FRDM and the ETF-SI, quenching has been introduced in the 
149: new variant of the mass formula, viz., ETF-SI~(Q) \cite{Pear.96}. 
150: The mass table ETF-SI~(Q) has been shown to be successful in removing 
151: the main deficiencies in the abundance curve \cite{Pfeiffer.97}.
152: These experiments with the mass formulae have indicated the need of
153: weaker shell effects along the r-process path.  A microscopic basis 
154: of such a requirement at the r-process path, however, needs to be
155: examined.
156: 
157: In a recent investigation of the r-process nucleosynthesis of heavy nuclei
158: using mass formulae based upon Hartree-Fock-Bogoliubov approach, it has been
159: shown \cite{Wanajo.04} that weak shell effects in microscopic mass formulae
160: result in a spread of abundance distribution near the $A \sim 130$ 
161: and $A \sim 190$ peaks. This has the consequence that large deviations are
162: observed compared to the solar-system abundances especially for the peak
163: about $A \sim 190$. In an earlier analysis \cite{Thielemann.94}, it
164: was also shown that in a realistic astrophysical scenario a mass model
165: without quenching at $N=126$ can fill up deficiencies (troughs) 
166: near $A \sim 175$ due to freeze-out effects. However, this seems to apply
167: to only the third peak in the abundance curve. Thus, the role of the shell
168: effects at $N=126$ in the r-process region is not yet clear. 
169:  
170: In our earlier work \cite{Sharma.02,Sharma.01,Farhan.03}, we investigated 
171: the behaviour of the shell effects at the magic number $N=82$ within
172: the Relativistic Hartree-Bogoliubov approach. Using the Lagrangian model 
173: of the nonlinear quartic coupling of $\omega$ meson in the RMF theory, 
174: it was shown that microscopic RMF calculations show a weakening 
175: of the shell effects at $N=82$ in going from the stability line to the 
176: r-process path. It was also shown that in going from the r-process 
177: path to the neutron drip line, the the shell gap diminishes 
178: to a vanishingly low value at a given isospin, resulting in a complete 
179: washout of the shell effects \cite{Sharma.02}. In the present work, 
180: we have investigated the behaviour and evolution of the shell effects 
181: at $N=126$ in the region of the r-process path within the framework of 
182: the RHB approach. The salient features of the formalism are discussed 
183: in Section~\ref{rmf} with emphasis on the shell properties of nuclei 
184: as inherent to the subject. Results of microscopic RMF calculations 
185: using two different Lagrangian models are presented in 
186: Section~\ref{results}. A comparison of the results is made with the 
187: predictions of various macroscopic-microscopic mass formulae and
188: influence of shell effects on the r-process nucleosynthesis is 
189: discussed. A discussion is presented of the possible consequences of
190: the shell effects on the r-process nucleosynthesis. A summary of the 
191: results is presented in the last section.
192: 
193: \section{The shell effects in nuclei}
194: \label{shell}
195: 
196: The shell effects constitute an important feature of nuclei and 
197: are known to manifest strongly in terms of the magic numbers. 
198: This is signified by a conspicuous presence of prominent kinks 
199: about the major magic numbers in two-neutron separation 
200: energies ($S_{2n}$) all along the stability line \cite{Bor.93}. 
201: This is a manifestation of the existence of large shell gaps at 
202: magic numbers in nuclei. The spin-orbit interaction plays a 
203: pivotal role in the creation of the shell gaps and consequently
204: of the magic numbers. Of late, there are indications that with extreme
205: isospin in light nuclei, re-adjustment of single-particle levels might 
206: lead to an emergence of new magic numbers \cite{Fridmann.05} other than those 
207: which are are hitherto established.  
208: 
209: The spin-orbit interaction and consequently how the shell effects 
210: behave in the extreme regions would play a significant role in carving 
211: out shell gaps in nuclei near the r-process path. 
212: In microscopic approaches such as nonrelativistic 
213: density-dependent Skyrme theory and the relativistic mean-field 
214: theory, the spin-orbit interaction is determined 
215: by data on a few nuclei. Whilst in the former, the spin-orbit 
216: interaction is added in the theory on an {\it ad hoc} basis and its 
217: strength is adjusted to known spin-orbit splittings in a few nuclei, 
218: it arises naturally in the RMF theory as a consequence of the 
219: Dirac-Lorentz structure of nucleons. This has shown much usefulness 
220: in explaining properties that involve shell effects, such as anomalous 
221: isotope shifts in stable nuclei, especially those associated to
222: the Pb chain \cite{SLR.93}. This feature of the shell effects has not 
223: been possible to attain in the Skyrme theory without undertaking 
224: significant alterations in the isospin dependence of the spin-orbit 
225: interaction \cite{RF.95}. On the other hand, the intrinsic form of the 
226: spin-orbit interaction in the RMF theory has been found to be 
227: advantageous over that in the nonrelativistic approach. The appropriate
228: isospin dependence of the spin-orbit interaction \cite{SLK.94}
229: has been found to be successful in reproducing the anomalous 
230: isotope shifts in Pb nuclei \cite{SLR.93} as well as in Sr and Kr 
231: isotopes \cite{LS.95}. Consequently, it is expected to have 
232: implications in predicting the shell strength in the extreme
233: regions of the r-process path.
234: 
235: The RMF theory \cite{SW.86,Rein.89,Ser.92,Ring.96} has shown an immense 
236: potential in being able to describe properties of nuclei along the 
237: stability line \cite{GRT.90,SNR.93} and for a large number 
238: of nuclei beyond the stability line. Most of the 
239: Lagrangian parameter sets are based upon reproduction of binding 
240: energies, charge radii and in some cases surface thickness of a 
241: few key nuclei \cite{Rein.89}. Various forces are obtained 
242: in such a way that spin-orbit splitting in some key nuclei such 
243: as $^{16}$O is reproduced reasonably well. It should, however, be 
244: pointed out that it does not necessarily ensure that shell gaps or 
245: shell effects at the major magic numbers are reproduced correctly. 
246: 
247: The shell effects were not considered explicitly in the initial 
248: developments in the RMF theory of finite nuclei. This problem was 
249: addressed in Ref. \cite{Sharma.00}, where shell effects were investigated in 
250: nuclei along the stability line with a view to see their influence on 
251: nuclei near r-process path or on some known ``waiting-point'' 
252: nuclei \cite{Sharma.02}. It was shown \cite{Sharma.00} that the otherwise 
253: successful RMF forces based upon non-linear self-coupling of $\sigma$-meson 
254: such as NL-SH \cite{SNR.93} overestimate the experimental shell gaps in 
255: nuclei along the stability line. In order to solve this problem, 
256: the Lagrangian model with the nonlinear scalar coupling of $\sigma$ meson
257: was extended with the inclusion of the nonlinear quartic coupling of 
258: the $\omega$ meson \cite{Sharma.00}. Consequently, shell effects in 
259: Ni and Sn isotopes at the stability line were reproduced well.
260: 
261: The shell effects along the stability line (may) have repercussions 
262: (as it seems to be the case for the RMF theory, but not 
263: unequivocally for the mass formulae) on the shell effects far 
264: away from it. The character of the shell effects, be it
265: strong or weak vis-a-vis experimental data along the line of stability 
266: is likely to extrapolate alike (unless major re-adjustments in 
267: single-particle scheme take place giving rise to unexpected pattern of
268: behaviour) in the unknown regions of the periodic 
269: table. A test case for this hypothesis was provided by the waiting-point 
270: nucleus $^{80}$Zn ($N=50$) which lies close to the r-process path at $N=50$.
271: It was shown \cite{Sharma.02} that forces such as NL-SH that 
272: overestimate the shell effects at the stability line overestimate the 
273: shell effects for the waiting-point nucleus $^{80}$Zn and also in r-process
274: nuclei at $N=82$. On the other hand, the force NL-SV1 based upon the 
275: vector self-coupling of $\omega$-meson, which reproduces the shell 
276: effects in nuclei at the stability line, is able to reproduce the available 
277: data on the waiting-point nucleus $^{80}$Zn \cite{Sharma.02}. A firmer
278: verification of predictability of various theories would be provided
279: by future experimental data in the extreme regions. In the present work,
280: we explore how this feature translates for the shell 
281: effects at $N=126$ in the r-process region.
282: 
283: \section{The Relativistic Mean-Field theory}
284: \label{rmf}
285: 
286: The RMF approach \cite{SW.86} is based upon the Lagrangian density 
287: which consists of fields due to various mesons interacting with 
288: nucleons. The mesons include the isoscalar scalar $\sigma$-meson, 
289: the isoscalar vector $\omega$-meson and the isovector vector $\rho$-meson.
290: The details of the formalism can be found in 
291: Refs. \cite{Sharma.02,Rein.89,Ring.96,GRT.90}.
292: 
293: The RMF Lagrangian that describes the nucleons as Dirac spinors 
294: moving in the meson fields is given by \cite{SW.86}
295: 
296: 
297: \begin{eqnarray}
298: {\cal L}&=& \bar\psi \left( \rlap{/}p - g_\omega\rlap{/}\omega -
299: g_\rho\rlap{/}\vec\rho\vec\tau - \frac{1}{2}e(1 - \tau_3)\rlap{\,/}A -
300: g_\sigma\sigma - M_N\right)\psi\nonumber\\
301: &&+\frac{1}{2}\partial_\mu\sigma\partial^\mu\sigma-U(\sigma)
302: -\frac{1}{4}\Omega_{\mu\nu}\Omega^{\mu\nu}+ \frac{1}{2}
303: m^2_\omega\omega_\mu\omega^\mu\\ &&+\frac{1}{4}g_4(\omega_\mu\omega^\mu)^2
304: -\frac{1}{4}\vec R_{\mu\nu}\vec R^{\mu\nu}+
305: \frac{1}{2} m^2_\rho\vec\rho_\mu\vec\rho^\mu -\frac{1}{4}F_{\mu\nu}F^{\mu\nu}
306: \nonumber
307: \end{eqnarray}
308: where $M_N$ is the bare nucleon mass and $\psi$ is its Dirac spinor. 
309: Nucleons interact with $\sigma$, $\omega$, and $\vec\rho$) mesons, 
310: with the masses being $m_\sigma$, $m_\omega$ and $m_\rho$ and the 
311: coupling constants being $g_\sigma$, $g_\omega$, $g_\rho$, respectively. 
312: The electromagetic interaction is
313: represented by the electromagnetic vector field $A^\mu$. 
314: 
315: The field tensors for the vector mesons are given as 
316: $\Omega_{\mu\nu}=\partial_\mu\omega_\nu-\partial_\nu\omega_\mu$ 
317: and by similar expressions for the $\rho$-meson and 
318: the photon. For a realistic description of nuclear properties a nonlinear
319: self-coupling 
320: \begin{eqnarray}
321: U(\sigma) = \frac{1}{2} m^2_\sigma \sigma^2_{} + \frac{1}{3}g_2\sigma^3_{} 
322: + \frac{1}{4}g_3\sigma^4 
323: \end{eqnarray}
324: for $\sigma$-mesons has been widely used. The non-linear vector self-coupling 
325: of $\omega$-meson \cite{Bod.91} as added earlier \cite{Sharma.00} in 
326: the Lagrangian with the non-linear scalar field is represented by the 
327: coupling constant g$_4$. 
328: 
329: In the lowest order of the quantum field theory, i.e., in the  mean-field 
330: approximation, the nucleons are assumed to move independently in the 
331: meson fields. The meson fields are replaced by their classical expectation 
332: values. The variational principle leads to the Dirac equation:
333: \begin{eqnarray}
334: \{ -i{\bf {\alpha}} \nabla + V({\bf r}) + \beta {m*} \}
335: ~\psi_{i} = ~\epsilon_{i} \psi_{i}
336: \end{eqnarray}
337: where $V({\bf r})$ represents the $vector$ potential:
338: \begin{eqnarray}
339: V({\bf r}) = g_{\omega} \omega_{0}({\bf r}) + g_{\rho}\tau_{3} {\bf {\rho}}
340: _{0}({\bf r}) + \frac{e(1-\tau_{3})}{2} {A}_{0}({\bf r})
341: \end{eqnarray}
342: and $S({\bf r})$ is the $scalar$ potential
343: \begin{eqnarray}
344: S({\bf r}) = g_{\sigma} \sigma({\bf r})
345: \end{eqnarray}
346: which defines the effective mass as:
347: \begin{eqnarray}
348: m^{\ast}({\bf r}) = m + S({\bf r})
349: \end{eqnarray}
350: The Klein-Gordon equations for the meson fields are time-independent
351: inhomogeneous equations with the nucleon densities as sources.
352: \begin{eqnarray}
353: \{ -\Delta + m_{\sigma}^{2} \}\sigma({\bf r})
354:  &=&-g_{\sigma}\rho_{s}({\bf r})
355: -g_{2}\sigma^{2}({\bf r})-g_{3}\sigma^{3}({\bf r})
356: \nonumber \\ 
357: \  \{ -\Delta + m_{\omega}^{2} \} \omega({\bf r})
358: &=& g_{\omega}\rho_{v}({\bf r}) + g_4 \omega^3({\bf r}) 
359: \nonumber \\ 
360: \  \{ -\Delta + m_{\rho}^{2} \}\rho({\bf r})
361: &=& g_{\rho} \rho_{3}({\bf r})
362: \nonumber \\ 
363: \  -\Delta A({\bf r}) = e\rho_{c}({\bf r})
364: \end{eqnarray}
365: 
366: The stationary state solutions $\psi_i$ are obtained from the coupled 
367: system of Dirac and Klein-Gordon equations. The ground-state of the 
368: nucleus is described by a Slater determinant $\vert\Phi >$ of single-particle
369: spinors $\psi_i$ (i = 1,2,....A). Solution of the Dirac equation 
370: is achieved by using the method of oscillator expansion \cite{GRT.90}. 
371: In the RMF approach, the pairing is included within the BCS scheme.
372: However, for the case of nuclei in the extreme regions of the r-process
373: path and drip lines, the Fermi energy is very close to the continuum and
374: many single-particle states couple to the continuum. Thus, the BCS method
375: of pairing provides a crude approximation of such cases. The Relativistic 
376: Hartree-Bogoliubov (RHB) approach based upon quasi-particle scheme provides 
377: an appropriate framework to deal with nuclei of such a nature.
378: 
379: \subsection{The Relativistic Hartree-Bogoliubov approach}
380: 
381: Nuclei that are known to show strong pairing correlations
382: are treated appropriately within the framework of the
383: RHB approach. Pairing correlations in the neighbourhood of 
384: the Fermi energy in case of nuclei near r-process path and drip 
385: line become even more important. Herein, the RHB approach provides
386: a suitable framework to deal with nuclei in the extreme regions.
387: 
388: It has been shown \cite{KR.91} that using Green's function 
389: techniques \cite{Go.58} a relativistic Hartree-Bogoliubov 
390: approach can be implemented using the Lagrangian as given above. 
391: Neglecting retardation effects and the Fock term, one obtains  
392: relativistic Dirac-Hartree-Bogoliubov (RHB) equations
393: \begin{equation}
394: \left(\begin{array}{cc} h & \Delta \\ -\Delta^* & -h^* \end{array}\right)
395: \left(\begin{array}{r} U \\ V\end{array}\right)_k~=~
396: E_k\,\left(\begin{array}{r} U \\ V\end{array}\right)_k,
397: \label{RHB} 
398: \end{equation}
399: where $E_k$ are quasiparticle energies and the coefficients $U_k$ and 
400: $V_k$ are four-dimensional Dirac spinors normalized as
401: \begin{equation}
402: \int ( U^+_k U^{}_{k'}~+~V^+_kV^{}_{k'}\, ) d^3r~=~\delta_{kk'}.
403: \end{equation} 
404: The average field
405: \begin{equation}
406: h~=~\mbox{\boldmath $\alpha.p$}~+~g_\omega\omega~+~ \beta(M+g_\sigma
407: \sigma)~-~\lambda
408: \label{h-field}
409: \end{equation}
410: contains the chemical potential $\lambda$. 
411: The meson fields $\sigma$ and $\omega$ are determined self-consistently 
412: from the Klein Gordon equations as done in the case of the 
413: RMF equations discussed above with the scalar density 
414: $\rho_s=\sum_k \bar V^{}_kV^{}_k$ and the baryon density 
415: $\rho_v=\sum_k V^+_kV^{}_k$.  The sum on $k$ is taken only over 
416: the particle states in the no-sea approximation. 
417: The pairing potential 
418: $\Delta$ in Eq. (\ref{RHB}) is given by
419: \begin{equation}
420: \Delta_{ab}~=~\frac{1}{2}\sum_{cd} V^{pp}_{abcd} \kappa_{cd}
421: \label{pair}
422: \end{equation}
423: The RHB equations (\ref{RHB}) are a set of four coupled 
424: integro-differential equations for the Dirac spinors $U(r)$ and 
425: $V(r)$ that are obtained self-consistently. The RHB calculations 
426: are performed by expanding fermionic and bosonic wavefunctions 
427: in 20 oscillator shells. For the pairing channel, we use the 
428: finite-range Gogny force D1S \cite{Berger.84}. The Gogny force is 
429: a sum of two Gaussians with finite range. It has been shown 
430: \cite{Berger.84} that the Gogny force is able to describe 
431: pairing properties of a large number of finite nuclei in the
432: medium and heavy mass regions. Details of the RHB theory can be 
433: found in Ref. \cite{KR.91}
434: 
435: 
436: \subsection{Lagrangian Models}
437: 
438: The Lagrangian model with the nonlinear scalar coupling of $\sigma$ meson
439: has been the widely used one for finite nuclei within the RMF theory. 
440: It has been successful in reproducing ground-state properties of 
441: nuclei at the stability line as well as of those far away from it. 
442: Here, we will consider the successful forces NL-SH \cite{SNR.93} 
443: and NL3 \cite{Lala.97} within this Lagrangian model. 
444: We will also employ the forces NL-SV1 and NL-SV2 \cite{Sharma.05} 
445: (see ref. \cite{Sharma.00} for the parameter sets) with the 
446: nonlinear vector self-coupling of $\omega$ meson. 
447: As mentioned above, forces NL-SV1 and NL-SV2 were constructed with the 
448: inclusion of the quartic vector coupling of $\omega$ meson, in order 
449: to solve the problem of strong shell effects with Lagrangian model with 
450: nonlinear scalar coupling of $\sigma$ meson \cite{Sharma.00}. 
451: The introduction of the non-linear coupling of $\omega$-meson also
452: softens the equation of state (EOS) of the nuclear matter significantly. 
453: This has the consequence that the maximum neutron star mass with such 
454: an EOS would show a better agreement with empirically observed 
455: values. A detailed discussion of the properties associated with the 
456: introduction of the nonlinear vector self-coupling of $\omega$ meson 
457: in the RMF theory will be presented elsewhere \cite{Sharma.05}. It will 
458: be shown \cite{Sharma.05} that the Lagrangian parameter set NL-SV1 is 
459: also able to improve upon the ground-state properties such as binding 
460: energies, charge radii and isotopes shifts of nuclei along the stability
461: line and far away from it as compared to those with NL-SH and NL3.
462: 
463: 
464: \section{Details of the calculations}
465: 
466: In RHB calculations, wavefunctions are expanded into an
467: harmonic-oscillator basis to solve the Dirac and the Klein-Gordan 
468: equations \cite{GRT.90}. For both the fermionic and bosonic fields 
469: a basis of 20 oscillator shells has been used. The pairing has 
470: been taken in the Bogoliubov approach. For, nuclei a few neutrons 
471: below and a few neutrons above a magic number are usually 
472: spherical, RHB calculations have been performed for a 
473: spherical configuration. For a comparative study of the shell 
474: effects, we have performed RHB calculations with the two Lagrangian 
475: models as discussed above. However, our focus is on investigation of 
476: the potential and predictive power of the new Lagrangian model with the 
477: vector self-coupling of $\omega$ meson vis-a-vis the scalar 
478: self-coupling of $\sigma$ meson.
479: 
480: With a view to investigate as to how the shell effects evolve in going
481: from the region of the stability line towards the r-process path and 
482: ultimately to the neutron drip line, we have selected even-even nuclei 
483: from the isotopic chains of Hf ($Z=72$) down to Xe ($Z=54$) across 
484: the neutron magic number $N=126$.  For our focus is on the behaviour of
485: the shell gap, nuclei relevant to the discussion are those with $N=124$, 
486: $N=126$ and $N=128$, so as to be able to calculate $S_{2n}$ values 
487: across the magic number $N=126$.
488: 
489: \section{Results and Discussion}
490: \label{results}
491: \subsection{Shell effects along the stability line}
492: First, we examine the known shell effects at $N=126$ at the stability line.
493: The experimental shell gap at $N=126$ is known only in a very few 
494: isotopic chains in this region. Here, the doubly closed nucleus 
495: $^{208}$Pb in the Pb chain provides a cardinal point in the study 
496: of the shell effects at $N=126$. 
497: Interestingly, though the nuclear landscape of the periodic table 
498: has been extended significantly due to sustained experimental 
499: efforts in the last decade towards synthesizing nuclei far beyond the 
500: stability line in the laboratory, the heaviest known Pb nucleus 
501: has reached only to $^{214}$Pb. 
502: 
503: We begin with the premise that the two-neutron separation energy 
504: at the magic number provides a reasonably good indicator of the shell gap.
505: Therefore, we calculate the shell gap at the magic number as
506: defined conveniently by
507: \begin{equation}
508: \Delta_S = S_{2n}(Z, N_0) - S_{2n}(Z, N_0+2),
509: \label{gap}
510: \end{equation}
511: where $S_{2n}(Z, N_0)$ denotes the 2-neutron separation energy of the
512: nucleus $(Z, N_0)$ with a magic neutron number $N_0$.
513: 
514: The shell gap in Pb nuclei obtained from the RMF approach using various
515: Lagrangian sets is shown in Fig.~1(a) and (b). The corresponding gap 
516: from various macroscopic-microscopic mass formulae is shown in Fig.~1(c).
517: The results are compared with the experimentally known data. 
518: The $S_{2n}$ values calculated in the RMF theory with the forces 
519: NL-SH and NL3 within Lagrangian model of the nonlinear scalar 
520: coupling of $\sigma$ meson are shown in Fig.~1(a). 
521: The difference between the data points at $N=126$ and $N=128$, 
522: as indicated by Eq.~\ref{gap}, manifests the shell gap at the
523: magic number $N=126$. A comparison with the experimental data points shows
524: that the shell gap from the force NL-SH underestimates the experimental
525: gap by $\sim 0.5$ MeV. On the other hand, the recent force NL3 shows a good 
526: agreement with the data. Here, we do not show the binding energy of 
527: $^{208}$Pb itself that is overestimated both by NL-SH and NL3 by 
528: about 2-3 MeV. However, the difference in the $S_{2n}$ values of the
529: neighboring nuclei turns out to be satisfactory in both the cases.
530: It is interesting to note that although NL-SH underestimates 
531: the shell gap at $N=126$ slightly, it is known to exhibit 
532: generally stronger shell effects for nuclei at and beyond the
533: stability line than those from NL3. This implies that a focus on a 
534: single or a few data points may not be deterministic as far as 
535: extrapolations (predictions) for nuclei beyond the stability line
536: are concerned. Consequently, the shell gap at $N=126$ in the Pb 
537: nucleus may not provide a successful conjecture as to whether the 
538: shell effects with either of the forces shall remain strong or 
539: weak in the domain that is far beyond the stability line.
540: 
541: 
542: % For one-column wide figures use
543: \begin{figure}
544: % Use the relevant command for your figure-insertion program
545: % to insert the figure file.
546: % For example, with the option graphics use
547: %\centering
548: \hspace{-0.5cm}
549: \resizebox{0.45\textwidth}{!}{%
550: %  \includegraphics{fig1.eps}
551:   \rotatebox{270}{\includegraphics{fig1.eps}}
552: }
553: % If not, use
554: \vspace{0cm}       % Give the correct figure height in cm
555: \caption{The two-neutron separation energies $S_{2n}$ for Pb isotopes
556: as obtained from (a) the forces with the Lagrangian model of the nonlinear
557: scalar self-coupling of $\sigma$ meson, (b) from the forces which
558: include the nonlinear quartic coupling of $\omega$ meson, and (c)
559: from the mass formulae FRDM \cite{Moeller.95}, ETF-SI \cite{Abou.95} 
560: and the recently obtained results from mass tables HFB-2 \cite{Goriely.02}.
561: The experimental data is shown by the solid circles.}
562: \label{fig:1}       % Give a unique label
563: \end{figure}
564: 
565: 
566: % For one-column wide figures use
567: \begin{figure*}
568: % Use the relevant command for your figure-insertion program
569: % to insert the figure file.
570: % For example, with the option graphics use
571: \centering
572: \resizebox{0.65\textwidth}{!}{%
573: %  \includegraphics{fig2.eps}
574:   \rotatebox{270}{\includegraphics{fig2.eps}}
575: }
576: % If not, use
577: \vspace{1cm}       % Give the correct figure height in cm
578: \caption{The two-neutron separation energies $S_{2n}$ for the isotopic
579: chains from Hf ($Z=72$) to Xe ($Z=54$) in going towards the r-process nuclei
580: and in approaching the neutron drip line, obtained from the forces with the 
581: Lagrangian model of the nonlinear scalar self-coupling of 
582: $\sigma$ meson (a) NL-SH \cite{SNR.93} and (b) NL3 \cite{Lala.97}. 
583: The $S_{2n}$ values from the forces which include the nonlinear 
584: quartic coupling of $\omega$ meson \cite{Sharma.00,Sharma.05}
585: are also shown in (c) with NL-SV2  and in (d) with NL-SV1.} 
586: \label{fig:2}       % Give a unique label
587: \end{figure*}
588: 
589: 
590: 
591: The $S_{2n}$ values obtained from the forces NL-SV1 and NL-SV2 with the 
592: Lagrangian model with the nonlinear quartic coupling of $\omega$ meson 
593: are shown in Fig.~1(b). A comparison with the experimental data shows
594: that NL-SV1 reproduces the shell gap at $N=126$ well, whereas it is
595: underestimated by $\sim 0.5$ MeV by the force NL-SV2. Again, a
596: seemingly paradoxical situation arises here. As in the case of NL-SH,
597: exhibiting generally stronger shell effects and yet underpredicting 
598: the shell gap at $N=126$ as shown in Fig.~1(a), the force NL-SV2 has 
599: also been shown to exhibit shell effects slightly stronger than those 
600: with NL-SV1 in the r-process region at $N=82$ \cite{Sharma.02}. 
601: This shows that reliance on a few data points may not be useful in 
602: predicting behaviour in the extreme regions. This will be shown in 
603: the latter parts of this paper, where we will discuss the shell 
604: effects in the region of the r-process path and the neutron drip line.
605: 
606: 
607: 
608: It is equally interesting to see as to how various mass formulae predict 
609: the shell gap at $N=126$ in Pb nuclei at the stability line.
610: In Fig.~1(c), the data points from the mass formula FRDM and ETF-SI
611: are shown. It may be remarked that various mass formulae including 
612: FRDM and ETF-SI have been obtained with a view to reproducing 
613: experimental data on more than a thousand nuclei. Generally, these 
614: mass tables have achieved a great success in reproducing a large set of 
615: experimental database through exhaustive fits over the periodic table 
616: than possibly a microscopic theory could ever do. However, as pointed 
617: out in the literature \cite{Moeller.95,Goriely.02}, 
618: discrepancies at the magic numbers do remain a significant drawback. 
619: 
620: The FRDM indicates a shell gap that is $\sim 1.5$ MeV 
621: smaller than the experimental one. On the other hand, ETF-SI 
622: underestimates the shell gap only by about 0.5 MeV. The undervaluation
623: of the shell gap with FRDM and ETF-SI at $N=126$ along the 
624: stability line is to be contrasted to the stronger shell effects 
625: due to these mass formulae when extrapolated in the extreme regions 
626: of the r-process path both at $N=82$ and $N=126$, than are suggested
627: for a successful reproduction of r-process abundances. Notwithstanding 
628: the need  of weaker shell effects along the r-process path, a new mass table 
629: based upon the Skyrme Hartree-Fock-Bogoliubov (HFB) approach has 
630: recently been produced \cite{Goriely.02}. The data points from the mass
631: table HFB-2 as shown in Fig.~1(c) underestimate the shell gap 
632: at $N=126$ significantly. It is worth pointing out that the 
633: mass formula HFB-2 \cite{Goriely.02} seems to have 
634: achieved a similar quality of fit across much of the periodic table. 
635: It is comparable to ETF-SI and FRDM, albeit with shell gaps that are 
636: predicted to be small in the r-process region, as we will see 
637: in Section~\ref{mass}.
638: 
639: 
640: \begin{figure*}
641: % Use the relevant command for your figure-insertion program
642: % to insert the figure file.
643: % For example, with the option graphics use
644: %\rotatebox{90}
645: \centering
646: %\vspace{0.5cm}
647: \resizebox{0.65\textwidth}{!}{%
648: %  \includegraphics{fig3.eps}
649:   \rotatebox{270}{\includegraphics{fig3.eps}}
650: }
651: % If not, use
652: \vspace{1cm}       % Give the correct figure height in cm
653: \caption{The two-neutron separation energies $S_{2n}$ for the isotopic
654: chains from Hf ($Z=72$) to Xe ($Z=54$) in going towards the r-process nuclei
655: and in approaching the drip line, as predicted by the mass models (a) FRDM 
656: \cite{Moeller.95} and (b) ETF-SI \cite{Abou.95}. The $S_{2n}$ values 
657: from the mass formula ETF-SI~(Q) \cite{Pear.96} that includes a quenching 
658: of the shell effects superimposed on ETF-SI are shown in (c). The results 
659: from the mass formula HFB-2 \cite{Goriely.02} are shown in (d).}
660: \label{fig:3}       % Give a unique label
661: \end{figure*}
662: 
663: 
664: \subsection{Shell effects near the r-process path - RMF theory}
665: 
666: The shell effects at $N=126$ for r-process nuclei play a crucial role in
667: determining r-process abundances around the peak at $A \sim 190$ 
668: \cite{Kratz.93}. For practical purposes r-process nuclei are 
669: defined to be those with $S_n \sim 2-4$ MeV. Thus, the r-process path 
670: is not strictly well defined and it does vary from model to model. 
671: However, it is generally accepted that nuclei with 
672: $Z \sim 64-69$ near $N=126$ fall along the r-process path.
673: 
674: 
675: 
676: 
677: The results of RHB calculations for the forces NL-SH and NL3 with the
678: Lagrangian model of nonlinear $\sigma$ coupling are shown in the upper 
679: panels (a) and (b) of Fig.~2, respectively. It is seen clearly that the
680: shell gap that is represented by the difference between the $S_{2n}$
681: values at $N=126$ and at $N=128$ shows a gradual decrease 
682: with an increase in isospin i.e., in going from the element 
683: Hf ($Z=72$) that is slightly above the r-process path to Xe ($Z=54$) 
684: that is near the drip line. This behaviour is similar for both
685: NL-SH and NL3. In absolute terms the shell gaps for NL-SH are 
686: slightly larger than those of NL3. However, in both the cases, 
687: the shell gap does not show a significant decrease in approaching 
688: the drip line, as is probably expected from a comparison with the 
689: corresponding behaviour of the shell gap at $N=82$ at the drip line 
690: \cite{Sharma.02}.
691: 
692: 
693: 
694: The $S_{2n}$ values obtained from the forces NL-SV1 and NL-SV2 with
695: the inclusion of quartic coupling of $\omega$ meson are shown in
696: the lower panels (c) and (d) of Fig.~2, respectively. For both the forces
697: NL-SV1 and NL-SV2, the shell gaps show a similar gradual decrease in going
698: towards the r-process path and the drip line. Qualitatively, the 
699: behaviour of both the Lagrangian models as portrayed in the upper
700: and the lower panels, respectively, is very similar. However, 
701: quantitatively, the shell gaps with NL-SV1 are smaller than both with
702: NL-SH and NL3. The force NL-SV2 provides slightly larger shell gaps than 
703: those  NL-SV1. 
704: 
705: 
706: All the parameters sets of both the Lagrangian models exhibit a slight
707: reduction in the shell strength in the region of the r-process path.
708: The difference lies only in the degree by which the shell effects are
709: reduced in going from the r-process path to the drip line. We will discuss
710: the comparative behaviour of the shell effects with various RMF models
711: in Section~\ref{compar}. 
712: 
713: 
714: 
715: \subsection{Shell effects near the r-process path - mass models}
716: \label{mass}
717: 
718: In the absence of and in essence rather infeasibility at present
719: of constructing a mass table based purely on microscopic calculations, 
720: masses from various mass tables based upon macroscopic-microscopic approach 
721: are used in r-process calculations. Here, we present the results of the two 
722: most elaborate mass formulae, the FRDM and the ETF-SI. 
723: 
724: We show the $S_{2n}$ values obtained from the FRDM and ETF-SI in the 
725: upper panels (a) and (b) of Fig.~3, respectively. In contrast to the 
726: microscopic calculations of Fig.~2, the shell gaps with FRDM and 
727: ETF-SI show nearly constant values in going from $Z=72$ to $Z=54$. 
728: On the other hand, a slight increase in the shell gap is visible 
729: with FRDM in going towards Xe. However, as the $S_{2n}$ values 
730: from FRDM for the nuclides with $N=128$ become negative below 
731: Nd $(Z=60)$, the apparent increase in the value of the shell gap for 
732: nuclei below Nd can be therefore be discounted. 
733: 
734: A constant shell gap is also displayed by ETF-SI as shown in Fig.~3(b).
735: The magnitude of the shell gap with ETF-SI, is, however, larger than 
736: that with FRDM by $\sim$ 0.8 MeV. On the lower end, i.e, for 
737: nuclides with $N=128$, the ETF-SI also shows negative $S_{2n}$ 
738: values below Nd $(Z=60)$. Thus, the drip line is reached
739: at $N=128$ for the elements below Nd, the behaviour very similar to 
740: that of FRDM. Moreover, with ETF-SI the $S_{2n}$ values for 
741: nuclides with the magic number $(N=126)$ are $\sim 1$ MeV 
742: larger than those of FRDM. The similar behaviour of the
743: shell effects with FRDM and ETF-SI and arrival of the drip line at a
744: similar location is not surprising, for the shell corrections 
745: superimposed on the smooth part in the two models are based upon 
746: the same prescription of the Strutinsky shell correction \cite{Stru.68}
747: 
748: \begin{figure*}
749: \centering
750: %\hspace{4cm}
751: \vspace{0.5cm}
752: \resizebox{0.60\textwidth}{!}{%
753: %  \includegraphics{fig4.eps}
754:   \rotatebox{270}{\includegraphics{fig4.eps}}
755: }
756: % If not, use
757: \vspace{0cm}       % Give the correct figure height in cm
758: \caption{The $S_{2n}$ values from NL-SV1 are compared to those from
759: (a) NL-SH and (b) NL3. The shell gap region is bounded by the two vertical
760: lines at $N=126$ and $N=128$.}
761: \label{fig:4}       % Give a unique label
762: \end{figure*}
763: 
764: 
765: Predictions from the mass formulae FRDM and ETF-SI in the extreme
766: regions of the r-process path have been used extensively for network
767: chain calculations of r-process nuclear abundances. Results of
768: calculations have shown that due to strong shell effects that are
769: prevalent  with FRDM and ETF-SI along the r-process path, there is
770: a significant deficiency (troughs) in the r-process abundances below
771: the $A \sim 130$ and $A \sim 190$ peaks \cite{Kratz.93}. Inspired by the 
772: Hartree-Fock-Bogoliubov (HFB) calculations with the Skyrme force
773: SkP \cite{Doba.96}, quenching was introduced in the ETF-SI mass formula, thus 
774: creating a new mass table ETF-SI~(Q) \cite{Pear.96}. The resulting 
775: r-process calculations \cite{Pfeiffer.97} with ETF-SI~(Q) have been 
776: able to fill up the deficiencies below the two peaks and results 
777: seem to be promising. This has put up a requirement (indirect)
778: of a weakening of the shell structure for nuclei near the r-process 
779: path \cite{Pfeiffer.97,Chen.95}. This feature that is clearly absent 
780: in the original mass formulae FRDM and ETF-SI has been introduced 
781: in the ETF-SI (Q) rather artificially. The ETF-SI~(Q) results in 
782: Fig.~3(c) show that the shell gaps remain the same as that with 
783: ETF-SI in going from Hf $(Z=72)$ to Er $(Z=68)$. However, the shell gap 
784: starts decreasing rapidly below Dy $(Z=66)$ in going towards Xe due 
785: to the quenching introduced therein. A microscopic basis of the 
786: aforesaid quenching introduced in the mass formula ETF-SI~(Q) 
787: is yet to be established.
788: 
789: Motivated by the quenching present in the HFB calculations with SkP, attempts
790: have been made to introduce the HFB approach in some mass tables. We show
791: in Fig.~3 (d) the results taken from the recently developed mass table 
792: HFB-2 \cite{Goriely.02} within the Skyrme Ansatz. 
793: The shell gap with HFB-2 is reduced for all the isotopic 
794: chains as compared to FRDM and ETF-SI.  A reduction in 
795: the shell gap with HFB-2 was also seen for Pb isotopes in Fig.~1(c). 
796: Though the shell gap is reduced vis-a-vis other mass formulae, 
797: it, however, remains constant in going from Hf $(Z=72)$ to 
798: Xe $(Z=54)$. There are no indications of an additional reduction 
799: in the shell strength near the drip line as compared to the region 
800: of the r-process path. Thus, the constancy of the shell gap
801: in going from the region of the r-process path to the drip line seems
802: to be the salient feature of the macroscopic-microscopic mass tables 
803: presented in Fig.~3. Here, the only exception is ETF-SI (Q), wherein 
804: the quenching was introduced by force. In comparison, the RMF results show
805: a slight reduction in the shell strength in going from the r-process path
806: to the drip line.
807: 
808: An earlier version of the HFB mass table, i.e., HFB-1 was 
809: constructed by replacing the BCS pairing by the Bogoliubov 
810: pairing scheme in Ref. \cite{Samyn.02}. In this work, it was shown 
811: that the behaviour of shell gaps far away from the stability line
812: does not depend much upon whether the BCS pairing or the Bogoliubov 
813: pairing is used. Accordingly, the shell gaps at $N=126$ 
814: with HFB-1 were found to be comparable to those with FRDM and 
815: that an introduction of Bogoliubov pairing did not result 
816: in a quenching of the shell effects \cite{Samyn.02}.
817: 
818: 
819: 
820: \subsection{A comparative analysis of the shell effects}
821: \label{compar}
822: In order to understand the comparative behaviour of the shell 
823: effects first in the RMF theory, we show in Fig.~4 the $S_{2n}$ 
824: values across the magic number $N=126$ from the force NL-SV1 and 
825: compare these with the other forces such as NL-SH and NL3. 
826: It may be reminded that although the force NL-SH has been 
827: found to be successful in reproducing binding energies, 
828: charge radii and deformation properties of a large number of
829: nuclei far away from the stability line, including the anomalous
830: isotope shifts in Pb nuclei, the shell effects with NL-SH were found
831: to be stronger as compared to the experimental data \cite{Sharma.00}.
832: In comparison, the force NL-SV1, having been able to describe the shell 
833: effects in Ni and Sn isotopes at the stability line, was also shown 
834: to be successful in reproducing the available data on the shell 
835: effects at the waiting-point nucleus $^{80}$Zn \cite{Sharma.02}. 
836: In view of this, we treat the force NL-SV1 as our benchmark. 
837: 
838: \begin{figure*}
839: \centering
840: %\hspace{4cm}
841: \vspace{0.5cm}
842: \resizebox{0.60\textwidth}{!}{%
843: %  \includegraphics{fig5.eps}
844:   \rotatebox{270}{\includegraphics{fig5.eps}}
845: }
846: % If not, use
847: \vspace{0cm}       % Give the correct figure height in cm
848: \caption{The $S_{2n}$ values from NL-SV1 are compared to those from the
849: mass formula (a) ETF-SI and with the quenched mass formula (b) ETF-SI~(Q).
850: The shell gap region is bounded by the two vertical lines at $N=126$ and 
851: $N=128$.}
852: \label{fig:5}       % Give a unique label
853: \end{figure*}
854: 
855: 
856: 
857: In Fig.~4(a), we compare the results of NL-SV1 to those from NL-SH. 
858: The variation in the shell gap is illustrated by the changing slopes
859: of the curves between the vertical bars at $N=126$ and $N=128$. 
860: A look at the difference in the $S_{2n}$ values and at the 
861: corresponding slope of the curve for Hf $(Z=72)$ shows that the 
862: shell gap with NL-SH is bigger than with NL-SV1.
863: As one progresses towards the r-process nuclei such as Er, Dy, Gd and
864: Sm, this difference in the shell gap between NL-SH and NL-SV1
865: increases with an increase in the isospin. One sees that the shell 
866: gaps with NL-SH remain stronger even in nuclei near the drip 
867: line such as Xe. In comparison, NL-SV1 shows a faster decrease in 
868: the shell gap in going from r-process nuclei to the drip line, a 
869: feature that has been called for for reproduction of r-process 
870: nuclear abundances.
871: 
872: 
873: The results from NL-SV1 are compared with those from the force
874: NL3 in Fig.~4(b). A comparison between the shell gaps 
875: from the two forces shows that beginning with Hf, 
876: the shell gap with NL3 is slightly larger than that with NL-SV1. 
877: This difference, however, increases slowly when one moves
878: from Hf to Xe. Thus, the shell effects with NL3 are slightly stronger
879: than those with NL-SV1. Additionally, the $S_{2n}$ values for nuclides with
880: $N=124$ are $\sim 0.5$ MeV higher with NL3 than NL-SV1, especially in the
881: region of r-process nuclei. This feature is similar to that of NL-SH 
882: vis-a-vis NL-SV1 as shown in Fig.~4(a). 
883: We compare in Fig.~5 the shell gaps from NL-SV1 to those from the
884: mass models (a) ETF-SI and from its variant (b) ETF-SI (Q) that embeds 
885: a quenching near the r-process region. For Hf ($Z=72$), the shell gap
886: with NL-SV1 is $\sim 1$ MeV larger than that of ETF-SI. However, as
887: one proceeds to the r-process nuclei and towards the drip line, 
888: this difference between NL-SV1 and ETF-SI decreases 
889: and then it reverses. For drip line nuclei near Xe ($Z=54$), 
890: the NL-SV1 shell gap is then smaller than that with ETF-SI. 
891: Whereas the shell gap with ETF-SI hardly shows any change in 
892: going from Hf to Xe, the shell gap with NL-SV1 does show a consistent 
893: decrease in going from the r-process to the drip line. 
894: 
895: 
896: We also compare the NL-SV1 predictions to those of ETF-SI~(Q) in
897: Fig.~5(b). From the nuclei of Hf $(Z=72)$ to about Sm $(Z=62)$, there 
898: is not much difference between the results of ETF-SI and ETF-SI~(Q).
899: Therefore, for these nuclei, a comparison of NL-SV1 shell gaps with 
900: ETF-SI~(Q) ones is similar to that with ETF-SI as shown in Fig.~5(b).
901: In the region of r-process nuclei $Z=64-68$, the shell gaps between
902: NL-SV1 and ETF-SI~(Q) are similar. However, due to an extra quenching 
903: added in ETF-SI~(Q), differences in the shell gaps of the two 
904: approaches begin appearing below Z=62. The shell gap with ETF-SI~(Q)
905: becomes especially small for nuclei near the drip line. The impact of the 
906: variations in the shell effects in nuclei can not be visualized 
907: without comprehensive r-process calculations. In the case of 
908: ETF-SI~(Q), such calculations do exist and have shown promising 
909: results. On the other hand, r-process calculations using the 
910: results from NL-SV1 are being planned currently.
911: 
912: \subsection{Models with strong shell effects}
913: 
914: The strength of the shell effects at the major magic numbers has been
915: a point of numerous discussions in respect of nuclear abundances 
916: \cite{Kratz.93,Kratz.00,Pfeiffer.01}. The focus has mostly been 
917: on the possibilities and potential capabilities 
918: of various mass formulae of macroscopic-microscopic origin. This
919: is evidently due to the fact that mass formulae have been able to
920: \begin{figure*}
921: \centering
922: %\hspace{4cm}
923: \vspace{0.5cm}
924: \resizebox{0.60\textwidth}{!}{%
925: %  \includegraphics{fig6.eps}
926:   \rotatebox{270}{\includegraphics{fig6.eps}}
927: }
928: % If not, use
929: \vspace{0cm}       % Give the correct figure height in cm
930: \caption{The $S_{2n}$ values from the FRDM are compared to those (a) from 
931: the force with strong shell effects NL-SH and (b) the mass formula with
932: strong shell effects ETF-SI.}
933: \label{fig:6}       % Give a unique label
934: \end{figure*}
935: produce large-scale nuclear binding energies and other data relevant 
936: for use in network chain calculations. The microscopic theories have 
937: not yet enjoyed this privilege with the exception of isolated case(s)
938: where nuclear masses for the purpose have been calculated meaningfully.
939: Here, we wish to take stock of the situation on the shell effects of
940: nuclei within various frameworks.
941: 
942: 
943: 
944: We compare results of the mass formulae FRDM and ETF-SI, both of which 
945: show strong shell effects along with those from the RMF force NL-SH 
946: also exhibits strong shell effects. The $S_{2n}$ values from FRDM and 
947: NL-SH are shown across the shell closure $N=126$ in Fig.~6(a). 
948: The parallel lines for $S_{2n}$ with FRDM from Hf to Xe in the
949: the region of the shell gap bounded by the vertical lines in Fig.~6(a)
950: indicate the constancy of the shell gap with FRDM as discussed above.
951: The shell gap as indicated by the $S_{2n}$ values from NL-SH and the ensuing
952: large slope of the curve between $N=126$ and $N=128$ demonstrates 
953: overly strong shell effects with the force NL-SH. In comparison, the shell
954: gap with FRDM is smaller than that with NL-SH by about 2 MeV for Hf nucleus.
955: This difference between NL-SH and FRDM persists even for r-process nuclei 
956: in the region of $Z=62-66$. Thus, the shell effects at $N=126$ with NL-SH 
957: are significantly stronger than those with FRDM. However, due to natural 
958: decrease in the shell gap with an increase in isospin in the RMF theory, 
959: the shell gap with NL-SH does become comparable to FRDM for nuclei near 
960: the drip line. 
961: 
962: The indication that the shell strength with NL-SH is strong 
963: appeared in Ref. \cite{SLHR.94}, where the shell effects with NL-SH 
964: were studied across $N=82$. In that work, it was shown that the shell 
965: effects at $N=82$ with NL-SH were as strong as those with FRDM, but 
966: not stronger than FRDM. It was also shown that there was a remarkable 
967: agreement between the ground-state properties of the isotopic chains 
968: of Zr from $^{112}$Zr through to $^{130}$Zr in the two approaches.
969: Nuclei in this region exhibited not only similar values of quadrupole 
970: deformation, but also similar shape transitions across the region 
971: with FRDM and NL-SH \cite{SLHR.94}. This led to a surmise that the 
972: shell effects and shell structure with NL-SH are very similar to 
973: those of FRDM. However, this does not seem to be the case in the 
974: region of $N=126$. 
975: 
976: Taking into consideration the results of network chain calculations 
977: performed thus far, we believe that r-process network chain 
978: calculations with the force NL-SH or any similar microscopic force 
979: exhibiting strong shell effects are not expected to be successful,
980: at least for the third peak at $A \sim 190$.
981: 
982: 
983: We compare the shell effects with FRDM  to those with ETF-SI in 
984: Fig.~6(b). The shell gaps with ETF-SI for nuclei in the vicinity 
985: of Hf are more than 1 MeV larger than those with FRDM. The strong
986: shell gaps with ETF-SI are maintained across the r-process region 
987: and a near constancy of the shell gap at $N=126$ both with FRDM and
988: ETF-SI is seen clearly. The difference in the shell gap between 
989: ETF-SI and FRDM, however, decreases in going from the r-process path
990: nuclei towards the drip line nuclei. The fact remains that with 
991: ETF-SI the shell effects are stronger than those with FRDM. 
992: 
993: Using macroscopic-microscopic mass formulae FRDM and ETF-SI, 
994: it was shown \cite{Kratz.93,Thielemann.94,Kratz.95} that due to the 
995: stronger shell gaps in ETF-SI, r-process network chain calculations 
996: lead to a greater deficiency (trough) in the r-process nuclear 
997: abundances  below $A \sim 190$ peak than with FRDM. Thus, the earlier
998: conclusion of much of these analyses has been that  stronger shell 
999: effects at $N=126$ do not seem to be conducive to reproducing nuclear
1000: abundances. However, it was shown in a later analysis 
1001: \cite{ Thielemann.94} that in a ``realistic'' astrophysical scenario,
1002: there is no stringent need for a quenching of the $N=126$ shell effects.
1003: The trough that appeared near $A \sim 175$ in earlier analyses could be 
1004: filled due to freeze-out effects even by using a mass model without quenching. 
1005: This raises the possibility that under appropriate conditions, mass formulae
1006: without a quenching of the shell strength at $N=126$ can be used
1007: successfully.
1008: 
1009: \begin{figure}
1010: \hspace{0cm}
1011: \resizebox{0.45\textwidth}{!}{%
1012: %  \includegraphics{fig7.eps}
1013:   \rotatebox{270}{\includegraphics{fig7.eps}}
1014: }
1015: % If not, use
1016: \vspace{0cm}       % Give the correct figure height in cm
1017: \caption{The shell gaps $\Delta_S$ (Eq.~\ref{gap}) at $N=126$ 
1018: as obtained from the RMF forces (a) NL-SH, NL-SV1 and NL-SV2 and 
1019: compared with those from ETF-SI~(Q). (b) The shell gaps from mass 
1020: formulae ETF-SI, FRDM and HFB-2 are shown. A comparison is also made 
1021: with the ETF-SI~(Q).}
1022: \label{fig:7}       % Give a unique label
1023: \end{figure}
1024: 
1025: \subsection{The $N=126$ shell gap in nuclei: RMF versus mass formulae}
1026: 
1027: The status of $N=126$ shell gaps $\Delta_S$ as defined in Eq.~\ref{gap} is 
1028: summarized in Fig.~7 for the RMF theory and for various mass formulae. 
1029: The shell gap at $N=126$ as obtained from the RMF calculations with NL-SH,
1030: NL-SV1 and NL-SV2 are shown in Fig.~7(a). All the microscopic
1031: calculations show a decreasing shell gap in going through
1032: the r-process region. This reduction is, however, a little stronger 
1033: as one moves towards the drip line.
1034: 
1035: A comparison shows that NL-SH exhibits shell gaps which are 
1036: consistently larger than those with NL-SV1 and NL-SV2. This characterizes
1037: the stronger shell effects of NL-SH vis-a-vis other forces. The shell gaps
1038: with NL-SV2 are slightly larger than those with NL-SV1. This was also
1039: shown to be the case for shell gaps at $N=82$ with NL-SV2 \cite{Sharma.02}. 
1040: 
1041: We compare the shell gaps with the RMF forces to those from the quenched
1042: mass formula ETF-SI~(Q) in Fig.~7(a). Evidently, the shell gaps with 
1043: ETF-SI~(Q) are much smaller than those with NL-SH. However, these are also
1044: smaller than those of the NL-SV1 and NL-SV2. The effect of the added 
1045: quenching in ETF-SI~(Q) is apparent for nuclei below $Z=64$, 
1046: where the shell gap is reduced significantly as compared to the nearly
1047: constant values maintained in ETF-SI (see Fig.~7(b)). For nuclei 
1048: below Nd $(Z=60)$, however, the quenching in the ETF-SI~(Q) is much 
1049: stronger than the weakening of the shell effects predicted by NL-SV1. 
1050: 
1051: The shell gaps from various mass formulae are compared in Fig.~7(b).
1052: Both ETF-SI and FRDM exhibit shell gaps which remain nearly constant
1053: in going from the region of the r-process path towards the drip line.
1054: Comparatively, ETF-SI shows shell effects that are stronger than
1055: those with FRDM in much of the region shown in the figure. As shown in 
1056: earlier calculations \cite{Pfeiffer.97}, the strength of the shell effects
1057: with ETF-SI was found to produce much larger deficiency below 
1058: $A \sim 190$ peak in the r-process abundances. 
1059: 
1060: The shell gaps with the new mass formula HFB-2 are compared with those
1061: from ETF-SI and FRDM in Fig.~7(b). The shell gaps with HFB-2 are nearly 
1062: constant as shown also by the other mass formulae. These are, however,
1063: systematically smaller than those of ETF-SI and FRDM. There is a tendency 
1064: of only a slight decrease in the shell gap in going towards the drip 
1065: line nuclei. It is interesting to note that the addition of a 
1066: Bogoliubov based pairing in the mass formula has not been found to 
1067: be sufficient to suppress further the shell gaps below the r-process 
1068: region. A comparison of the shell gaps from 
1069: ETF-SI, FRDM and HFB-2 with those from ETF-SI~(Q) shows that 
1070: the shell gaps from ETF-SI~(Q) are in striking contrast to all 
1071: the other mass tables. However, as mentioned earlier, this striking 
1072: difference has been brought about by the introduction of a quenching 
1073: based upon results of HFB+SkP calculations. This extra weakening
1074: of the shell gaps in the drip line region is not shown by any of 
1075: the widely used mass formulae. It has been reported in some calculations
1076: \cite{Pear.96, Kratz.95} that a weakening of the shell effects 
1077: in the r-process region to the drip line is required 
1078: for reproducing the r-process abundances around the peaks at 
1079: $A \sim 130 $ and $A \sim 190$. 
1080: 
1081: \subsection{The shell $N=126$ versus $N=82$}
1082: 
1083: We discussed the evolution of the shell gap at $N=82$ near the
1084: r-process path in the RMF theory in Ref. \cite{Sharma.02} in detail. 
1085: The shell gaps at $N=82$ were found to weaken in the region of 
1086: the r-process path. In going to the drip line, the shell effects
1087: at $N=82$ showed a substantial reduction in the shell strength, eventually
1088: leading to a complete disappearance of the shell gap for the drip line
1089: nuclei.
1090: 
1091: \begin{figure}
1092: \hspace{0cm}
1093: \resizebox{0.45\textwidth}{!}{%
1094: %  \includegraphics{fig8.eps}
1095:   \rotatebox{270}{\includegraphics{fig8.eps}}
1096: }
1097: % If not, use
1098: \vspace{0cm}       % Give the correct figure height in cm
1099: \caption{The $S_{2n}$ values from NL-SV1 are shown for nuclei 
1100: across the shell closure (a) $N=82$ and (b) $N=126$ in the 
1101: region of the r-process path. The evolution of the shell 
1102: gaps at $N=126$ is compared to that at $N=82$.}
1103: \label{fig:8}       % Give a unique label
1104: \end{figure}
1105: 
1106: 
1107: We show in Fig.~8 a comparison of the shell effects at $N=126$ 
1108: as obtained in the present work with those at $N=82$ \cite{Sharma.02}
1109: in the region of the r-process path to the drip line. In order to 
1110: visualize the shell gap at $N=82$, we show the $S_{2n}$ values 
1111: obtained with NL-SV1 across the magic number for nuclides of
1112: Sn $(Z=50)$ down to the drip line nuclides of Kr $(Z=36)$. 
1113: The shell gap at $N=82$ shows a strong decrease in going from 
1114: Cd $(Z=48)$ to Zr $(Z=40)$ in the region of the r-process path
1115: as mentioned above. This reduction in the shell gap at $N=82$ is 
1116: significantly faster than the corresponding shell gap at $N=126$ 
1117: (Fig.~8(b)). In the latter case, the shell gap shows a much slower 
1118: decrease in the region of the r-process nuclei from 
1119: Er $(Z=68)$ to Nd $(Z=60)$. The $N=126$ shell strength exhibits 
1120: a resilience to the change in isospin.
1121: 
1122: \subsection{The shell effects and r-process nucleosynthesis}
1123: 
1124: In the broader context of the shell effects and their implications on
1125: the r-process nucleosynthesis, it is pertinent to discuss recent
1126: r-process calculations which employ newly developed mass formulae
1127: HFB-2 and HFB-7 \cite{Wanajo.04}. In this work, effect of various 
1128: mass formulae on r-process nucleosynthesis has been studied using 
1129: astrophysical model of prompt supernova explosion from a collapsing 
1130: O-Ne-Mg core \cite{Wanajo.03}. In the mass formulae HFB-2 and HFB-7 
1131: which have been used, pairing has been taken into account by the 
1132: Bogoliubov method in a Skyrme density-functional approach. R-process
1133: calculations in this study have also been carried out using the nuclear
1134: masses from FRDM \cite{Moeller.95} and from the old droplet mass (DM)
1135: model \cite{Hilf.76}.
1136: 
1137: It has been shown \cite{Wanajo.04} that due to the shell effects which 
1138: are significantly reduced (weak) with HFB-2 and HFB-7, the abundance 
1139: curve in the third peak at $A \sim 190$ is spread leading to a shift 
1140: of the peak and consequently the valley at $A \sim 183$ is shifted 
1141: significantly to lower masses. The results show that by the use of 
1142: HFB-2 and HFB-7 masses, there are large deviations in the third peak as 
1143: compared to the solar-system r-process abundances. Thus, due to the
1144: weakness of the shell effects at $N=126$ near the r-process in the mass
1145: models HFB-2 and HFB-7, there is a significant overproduction of nuclei
1146: to the left of the third peak in the solar pattern.
1147: It is interesting to note that on using masses due to FRDM and DM,
1148: which are characterized by stronger shell effects at $N=126$, it has
1149: been shown \cite{Wanajo.04} that the abundance curve gives rise 
1150: to a sharp r-process peak at $A\sim 190$, in better agreement 
1151: with the solar pattern in the region. These results suggest that
1152: in the prompt supernova model considered in Ref. \cite{Wanajo.04}
1153: and conditions applicable therein, mass models/microscopic theories
1154: with stronger shell effects at $N=126$ would reproduce 
1155: the main features of the solar 
1156: r-process abundance curve around the third peak reasonably well. 
1157: It may be recalled that rather similar conclusion was reached in Ref.
1158: \cite{Thielemann.94}, wherein it was shown that a mass model
1159: without quenching at $N=126$ can fill up the trough at $A \sim 175$
1160: and reproduce the abundance curve near the third peak due to
1161: freeze-out effects.
1162: 
1163: In view of the different behaviour of the shell effects at $N=82$
1164: and $N=126$ as predicted by NL-SV1 in the RMF theory as shown in 
1165: Fig.~8, it is pertinent to discuss the results of the r-process 
1166: calculations of Ref. \cite{Wanajo.04} around the second peak at 
1167: $A \sim 130$. The results have shown that droplet mass models FDRM
1168: and DM with strong shell effects also at $N=82$ produce troughs 
1169: (underproduction) at $A \sim 115$ and $A \sim 140$ in the abundance
1170: curve. This shows that in contrast with $N=126$, stronger shell effects 
1171: at $N=82$ are not desirable for reproducing the solar abundance pattern.
1172: On the other hand, with HFB-2 and HFB-7 deficiencies in the
1173: r-process production below and above the second peak at $A \sim 130$
1174: are significantly remedied especially with HFB-2. However, the weak
1175: shell effects inherent in HFB-2 and HFB-7 have the consequence that 
1176: abundances around $A \sim 130$ are also spread out as opposed to the
1177: solar system r-process abundances. 
1178: 
1179: The results of Ref. \cite{Wanajo.04} indicate that within the model 
1180: employed in this work, the two peaks at $A \sim 130$ and 
1181: $A \sim 190$ require a different nature of the shell effects. For a
1182: successful reproduction of abundances near $A \sim 130$, shell effects
1183: at $N=82$ in the r-process region should not be strong, whereas
1184: the third peak seems to require moderate to strong shell effects
1185: at $N=126$ in contrast to $N=82$. Thus, in the model considered in
1186: Ref. \cite{Wanajo.04}, the analysis of the second and 
1187: the third peaks in the solar-system r-process abundance 
1188: curve suggests a different nature of the shell effects at $N=82$ 
1189: and at $N=126$.
1190: 
1191: Microscopic RHB calculations with the Lagrangian model NL-SV1 show
1192: two very different features for the shell effects at $N=82$ and 
1193: $N=126$, as depicted in Fig.~8. These features seem to be consistent
1194: with the picture that emerges from the results of Ref. \cite{Wanajo.04}.  
1195: It remains to be seen in future network chain calculations with 
1196: masses obtained from microscopic calculations with NL-SV1 whether
1197: the two different features of the shell effects exhibited 
1198: by NL-SV1 would suffice towards reproducing the r-process 
1199: abundances near the second and the third peak.
1200: 
1201: 
1202: \section{Conclusions}
1203: 
1204: We have investigated the shell effects in nuclei at the magic 
1205: neutron number $N=126$ in the region of the r-process 
1206: path using the relativistic Hartree-Bogoliubov approach within
1207: the relativistic mean-field theory. Two Lagrangian models, one with 
1208: the nonlinear scalar coupling of $\sigma$ meson and another one 
1209: that includes a nonlinear vector coupling of $\omega$ meson have 
1210: been considered. A comparison of the shell effects in the RMF theory
1211: has been made with the predictions of various mass formulae. It is shown 
1212: that the predictions of RMF calculations exhibit a slight reduction 
1213: of the shell gap in going from the r-process path to the neutron
1214: drip line irrespective of the shell strength exhibited by a Lagrangian
1215: parameter set along the stability line. This is slightly different
1216: from a near constancy of the shell gaps demonstrated by major mass formulae 
1217: in the region of the r-process path and the drip line. Consequences of 
1218: the shell strength on the r-process nucleosynthesis have been discussed.
1219: 
1220: It is shown that the Lagrangian force NL-SV1 with the vector 
1221: self-coupling of $\omega$ meson, which reproduces the shell gaps 
1222: along the stability line, shows that the shell effects at $N=126$ 
1223: exhibit only a marginal reduction in the shell strength in going from 
1224: the r-process path to the drip line. Consequenty, the shell 
1225: effects retain a strong character along the r-process path 
1226: at $N=126$. This is in striking contrast to the earlier
1227: results \cite{Sharma.02} with NL-SV1 that in the RMF theory 
1228: shell effects at $N=82$ exhibit a significant weakening of the strength
1229: in going from the r-process path to the neutron drip line. This shows 
1230: that different magic numbers may exhibit a different nature of the 
1231: shell effects in the extreme regions of the periodic table.
1232:  
1233: Analysis of the results of a recent r-process calculations \cite{Wanajo.04} 
1234: has suggested that stronger shell effects at $N=126$ and comparatively 
1235: weaker shell effects at $N=82$ are conducive to reproducing 
1236: the r-process abundances in the second and the third peak, respectively,
1237: in the solar-system r-process abundance curve. Our results exhibit
1238: features which are consistent with this analysis and support the
1239: conjecture that a different nature of the shell effects may be at 
1240: play in r-process nucleosynthesis of heavy nuclei.
1241: 
1242: A scrutiny of the shell effects with various RMF forces and with
1243: various mass formulae has shown that a given nature of the shell 
1244: effects in a known region may not be sufficient to calibrate the 
1245: shell structure in unknown regions. Accordingly, an extrapolation in 
1246: the unknown regions of the r-process path is frought with an 
1247: uncertainty.
1248: 
1249: 
1250: 
1251: \begin{acknowledgments}
1252: This work is supported by the Research project No. SP01/02 of 
1253: the Research Administration, Kuwait University.
1254: \end{acknowledgments}
1255: 
1256: \begin{thebibliography}{999}
1257: %
1258: % and use \bibitem to create references.
1259: %
1260: \bibitem{BBFH.57} E.M. Burbidge, G.R. Burbidge, A.A. Fowler and F. Hoyle,
1261: Rev. Mod. Phys. \textbf{29}, 547 (1957).
1262: 
1263: \bibitem{Hille.78} W. Hillebrandt, Space Sci. Rev., \textbf{21}, 639 (1978).
1264: 
1265: \bibitem{Cowan.91} J.J. Cowan, F.-K. Thielemann, and J.W. Truran, Phys.
1266: Rep. \textbf{208}, 267 (1991).
1267: 
1268: \bibitem{Kratz.93} K.-L. Kratz, J.P. Bitouzet,F.-K. Thielemann, P. M\"oller,
1269: and B. Pfeiffer, Astrophys. J. \textbf{403}, 216 (1993).
1270:  
1271: \bibitem{Kratz.00} K.-L. Kratz, B. Pfeiffer, F.-K. Thielemann, and 
1272: W.B. Walters, Hyp. Int., \textbf{129}, 185 (2000).
1273: 
1274: \bibitem{Pfeiffer.01} B. Pfeiffer, K.-L. Kratz, F.-K. Thielemann, and 
1275: W.B. Walters, Nucl. Phys. \textbf{A693}, 282 (2001).
1276: 
1277: \bibitem{Moeller.95} P. M\"oller, J. Nix, W. Swiatecki, At. Data Nucl. Data
1278: Tables \textbf{59}, 185 (1994).
1279: 
1280: \bibitem{Moeller.90} P. M\"oller and J. Randrup, Nucl. Phys. 
1281: \textbf{A514}, 1 (1990).
1282:  
1283: \bibitem{Abou.95} Y. Aboussir, J.M. Pearson, A.K. Dutta and F. Tondeur,
1284: At. Data Nucl. Data Tables \textbf{61}, 127 (1995).
1285: 
1286: \bibitem{Stru.68} V.M. Strutinsky, Nucl. Phys. \textbf{A122}, 1 (1968).
1287: 
1288: \bibitem{Doba.96} J. Dobaczewksi, W. Nazarewicz, T.R. Werner, J.F. Berger,
1289: C.R. Chinn, and J. Decharge, Phys. Rev. C \textbf{61}, 2809 (1996).
1290: 
1291: \bibitem{Pear.96} J.M. Pearson, R.C. Nayak, and S. Goriely, 
1292: Phys. Lett. B \textbf{387}, 455 (1996).
1293: 
1294: \bibitem{Pfeiffer.97} B. Pfeiffer, K.-L. Kratz, and F.-K. Thielemann,
1295: Z. Phys.  \textbf{A357}, 235 (1997).
1296: 
1297: \bibitem{Wanajo.04} S. Wanajo, S. Gorielly, M. Samyn and N. Itoh, 
1298: Astrophys. J. \textbf{606}, 1057 (2004).
1299: 
1300: \bibitem{Thielemann.94} F.K. Thielemann, K.L. Kratz, B. Pfeiffer, 
1301: T. Rauscher, L. van Wormer abd M.C. Wiescher, Nucl. Phys. \textbf{A570}, 
1302: 329c (1994).
1303: 
1304: \bibitem{Sharma.02} M.M. Sharma and A.R. Farhan, Phys. Rev. C \textbf{65}, 
1305: 044301 (2002).
1306: 
1307: \bibitem{Sharma.01} M.M. Sharma and A.R. Farhan, Nucl. Phys. \textbf{A688}, 
1308: 353c (2001).
1309: 
1310: \bibitem{Farhan.03} A.R. Farhan and M.M. Sharma, Nucl. Phys. \textbf{A719}, 
1311: 221c (2003).
1312: 
1313: \bibitem{Bor.93} C. Borcea, G. Audi, A.\,H. Wapstra, and P. Favaron, 
1314:                  Nucl. Phys. A \textbf{565}, 158 (1993).
1315: 
1316: \bibitem{Fridmann.05} J. Fridmann et al, Nature \textbf{435}, 922 (2005).
1317: 
1318: \bibitem{SLR.93} M.M. Sharma, G.A. Lalazissis, and P. Ring,
1319:     Phys. Lett. B \textbf{317}, 9 (1993.
1320:  
1321: \bibitem{RF.95} P.-G. Reinhard and H. Flocard, Nucl. Phys. 
1322: \textbf{A584}, 467 (1995).
1323: 
1324: \bibitem{SLK.94}M.M. Sharma, G.A. Lalazissis, J. K\"onig,
1325:     and P. Ring, Phys. Rev. Lett. \textbf{74}, 3744 (1994).
1326: 
1327: \bibitem{LS.95} G.A. Lalazissis and M.M. Sharma, 
1328:   Nucl. Phys. A \textbf{586}, 201 (1995).
1329: 
1330: \bibitem{SW.86} B.D. Serot and J.D. Walecka,
1331:     Adv. Nucl. Phys. \textbf{16}, 1 (1986).
1332: 
1333: \bibitem{Rein.89} P.G. Reinhard, Rep. Prog. Phys. \textbf{52}, 439 (1989).
1334: 
1335: \bibitem{Ser.92} B.D. Serot, Rep. Prog. Phys. \textbf{55}, 1855 (1992).
1336: 
1337: \bibitem{Ring.96} P. Ring, Prog. Part. Nucl. Phys. \textbf{37}, 193 (1996).
1338: 
1339: \bibitem{GRT.90}Y.K. Gambhir, P. Ring, and A. Thimet, Ann. Phys. (N.Y.) 
1340:                 \textbf{198}, 132 (1990).
1341: 
1342: \bibitem{SNR.93} M.M. Sharma, M.A. Nagarajan and P. Ring, Phys. Lett.
1343: B \textbf{312}, 377 (1993).
1344: 
1345: \bibitem{Sharma.00} M.M. Sharma, A.R. Farhan and S. Mythili, 
1346: Phys. Rev. C \textbf{61}, 054306 (2000).
1347: 
1348: \bibitem{Bod.91} A.R. Bodmer, Nucl. Phys. \textbf{A526}, 703 (1991).
1349: 
1350: \bibitem{Sharma.05} M.M. Sharma, in preparation (2005).
1351: 
1352: \bibitem{KR.91} H. Kucharek and P. Ring, Z. Phys. {\bf A339}, 23 (1991).
1353: 
1354: \bibitem{Go.58} L.P. Gorkov, Sov. Phys. JETP \textbf{7}, 505 (1958).
1355: 
1356: \bibitem{Berger.84} J.F. Berger, M. Girod and D. Gogny, 
1357: Nucl. Phys. {\bf A428}, 32 (1984).
1358: 
1359: \bibitem{Lala.97} G.A. Lalazissis, J. K\"onig, and P. Ring, Phys. Rev.
1360: C \textbf{55}, 540 (1997).
1361: 
1362: \bibitem{MN.92} P. M\"oller, J.R. Nix, Nucl. Phys. \textbf{A536}, 221 (1992).
1363: 
1364: \bibitem{Goriely.02} S. Goriely, M. Samyn, P.-H. Heenen, J.M. Pearson and
1365: F. Tondeur, Phys. Rev. C \textbf{66}, 024326 (2002).
1366: 
1367: \bibitem{Chen.95} B. Chen, J. Dobaczewski, K.-L. Kratz, K. Langanke, 
1368: B. Pffeifer, F.-K. Thielemann, and P. Vogel, 
1369: Phys. Lett. B \textbf{355}, 37 (1995).
1370: 
1371: \bibitem{Samyn.02} M. Samyn, S. Goriely, P.-H. Heenen, J.M. Pearson and
1372: F. Tondeur, Nucl. Phys. \textbf{A700}, 142 (2002).
1373: 
1374: \bibitem{SLHR.94} M.M. Sharma, G.A. Lalazissis, W. Hillebrandt, and 
1375: P. Ring, Phys. Rev. Lett. \textbf{72}, 1431 (1994). 
1376: 
1377: \bibitem{Kratz.95} K.-L. Kratz, in Nuclei in Cosmos III, AIP Conf. Proc.
1378: \textbf{327}, 113 (1995).
1379: 
1380: \bibitem{Wanajo.03} S. Wanajo, M. Tamamura, N. Itoh, K. Nomoto, 
1381: Y. Ishimaru, T.C. Beers, and S. Nozawa, Astrophys. J. \textbf{593}, 968 
1382: (2004).
1383: 
1384: \bibitem{Hilf.76} E.R. Hilf, H. von Groote, and K. Takahashi, in Proc. Third
1385: International Conference on Nuclei Far from Stability, (CERN), Geneva, p. 142.
1386: 
1387: \end{thebibliography}
1388: 
1389: \end{document}
1390: 
1391: 
1392: 
1393: 
1394: 
1395: 
1396: 
1397: