nucl-th0405027/nz.tex
1: \documentclass[12pt]{elsart}
2: \usepackage{amssymb}
3: \usepackage{graphicx}
4: \parindent 0 cm
5: \begin{document}
6: \begin{frontmatter}
7: \title{
8: Projected multicluster model with Jastrow and linear state dependent 
9: correlations for $12 \leq A \leq 16$ nuclei.
10: }
11: \author[address1]{E. Buend\'{\i}a}
12: \ead{buendia@ugr.es},
13: \author[address1]{F. J. G\'alvez}
14: \ead{galvez@ugr.es},
15: \author[address2]{A. Sarsa\corauthref{cor1}}
16: \ead{fa1sarua@uco.es} 
17: \corauth[cor1]{Corresponding author}
18: \address[address1]{
19: Departamento de F\'{\i}sica Moderna, Facultad de Ciencias\\
20: Universidad de Granada, E-18071 Granada, Spain}
21: \address[address2]{
22: Departamento de F\'{\i}sica, Campus de Rabanales\\
23: Universidad de C\'ordoba, E-14071 C\'ordoba, Spain}
24: 
25: \begin{abstract}
26: 
27: Variational wave functions based on a Margenau-Brink cluster model
28: with short range and state dependent correlations, and angular
29: momentum projection are obtained for some nuclei with $12 \leq A \leq
30: 16$.  The calculations have been carried out starting from the
31: nucleon-nucleon interaction by using the Variational Monte Carlo
32: method.  The configuration used consists of three alpha clusters
33: located at the apexes of an equilateral triangle, and an additional
34: cluster, not necessarily of alpha type, forming a tetrahedron.
35: This cluster is located at the top of its height.
36: Short-range and state dependent correlations are included by means of
37: a central Jastrow factor and a linear operatorial correlation factor
38: respectively.  Angular momentum projection is performed by using the
39: Peierls-Yoccoz operators.  Optimal structures are obtained for all the
40: nuclei studied.  Some aspects of our methodology have been tested by
41: comparing with previous calculations carried out without short range
42: correlations.  The binding energy, the root mean square radius, and
43: the one- and two-body densities are reported.  The effects of
44: correlations on both the energy and the nucleon distribution are
45: analyzed systematically.
46: 
47: \end{abstract}
48: 
49: \begin{keyword}
50: Nuclear structure; Cluster models; Variational Monte Carlo; $N\neq Z$;
51: $v_4$ forces
52: \PACS 21.60.-n; 21.60.Gx; 21.60.Ka; 27.20.+n
53: %21.60.-n Nuclear structure models and methods
54: %21.60.Gx Cluster models
55: %21.60.Ka Monte Carlo models
56: %27.20.+n 6(less-than-or-equal-to)A(less-than-or-equal-to)19
57: \end{keyword}
58: 
59: \end{frontmatter}
60: 
61: \newpage
62: 
63: \section{Introduction}
64: \label{introduction}
65: 
66: The joint use of short-range dynamic correlations with model wave
67: functions including relevant aspects of the nuclear structure
68: constitutes the most commonly used scheme to describe nuclear bound
69: states with realistic or semi realistic interactions.  Short range
70: correlations are essential elements in the wave function because, as
71: it is well known, any of the so-called realistic or semi-realistic
72: parameterizations of the nuclear potential presents a strong
73: short-range repulsive core.  On the other hand, the formation of
74: different kind of clusters in the nuclei can be understood as a
75: collective movement of the nucleons governed by the medium and long
76: range part of the nuclear potential.  Therefore, for an accurate
77: description of the nuclear states, it is convenient to consider both
78: aspects in any variational approach to the nuclear bound states using
79: this type of interactions.  In principle, short range correlations are
80: mainly governed by the nucleon-nucleon interaction while medium and
81: long range effects depend on the particular nuclear state.  However,
82: and in a more careful approach, the final form of the short range
83: correlations will depend on the model wave function giving rise to a
84: non negligible dependence of the correlations on the nucleus.  
85: 
86: A direct way to include both short range and medium and long range
87: correlations is by using Jastrow type correlation factors, but
88: the calculation of the expectation values becomes very cumbersome,
89: especially when state dependent correlations are included. There exist
90: several methods to evaluate these expectation values as those based on
91: cluster expansions \cite{bishop78,bosca88}, the
92: Fermi-HiperNetted-Chain method \cite{fabro98,fabro00} or statistical
93: methods such as the Variational Monte Carlo
94: \cite{carlson88,pieper90}. The Coupled Cluster method allows to
95: incorporate both type of correlations
96: \cite{zabo81,feldmeier98,heisen99}. In this way it is possible to
97: understand how the different correlation mechanisms are incorporated
98: \cite{bishop90,guar96}
99: 
100: Alpha cluster models, or cluster models in general, have been widely
101: applied in microscopic descriptions of bound and scattering states of
102: nuclear systems \cite{dedu97,dude97}.  Variational wave functions
103: built within this framework constitute an appropriate scheme for
104: nuclei such as $^8$Be and $^{12}$C, that present a clear cluster
105: structure.  The use of wave functions including the possibility of
106: formation of alpha cluster structures or any other kind of grouping of
107: nucleons improves the description of these nuclei and their neighbours
108: with respect to simple mean field approximations
109: \cite{arima72,michel98}.
110: 
111: Multi cluster models have been used in microscopic calculations,
112: i.e. without effective cluster-cluster interactions, based on the
113: Generator Coordinate Method for some nuclei between A=12 and A=16
114: \cite{dude96,desc02}.  In these works a Volkov nucleon-nucleon
115: potential was used \cite{volk65}.  Other results of microscopic
116: multicluster calculations based on the stochastic variational method
117: have been reported \cite{vst95,oasv00} for some nuclei using the
118: Minnesota potential.  Neither of these potentials presents a strongly
119: repulsive short range part and, therefore, short range correlations do
120: not play a significant role.  On the other hand, previous studies of
121: alpha clustering based on nuclear potentials with a strongly repulsive
122: core have been mainly restricted to spin-isospin saturated nuclei
123: \cite{gmn01,bgps02}.
124: 
125: The aim of this work is to study the ground state of some $p$-shell
126: nuclei, $A\neq 4n$, including clustering effects and short range and
127: state dependent correlations, starting from $v_4$--type
128: nucleon-nucleon interactions.  The nucleon clustering is described in
129: terms of model wave functions based on a generalized Margenau-Brink
130: model as in \cite{dude96}.  Short range correlations are included by
131: means of a Jastrow factor and the dependence on the spin and isospin
132: exchange channels is included by using a linear state dependent
133: correlation factor.  Angular momentum projection is carried out in
134: order to obtain variational wave functions that are eigenfunctions of
135: the total angular momentum operator.  The calculations are performed
136: by means of the Variational Monte Carlo method.
137: 
138: Here we extend a previous work \cite{bgps02} to the $A\neq 4n$ case.
139: This generalization is not straightforward because the angular
140: momentum projection involves a spin mixing not present in spin and
141: isospin saturated nuclei.  In this paper we present an analytical
142: reduction of the different expectation values for these nuclei,
143: obtaining expressions suitable for the Variational Monte Carlo method.
144: By using this scheme, the computing time is hardly increased with
145: respect to the spin-isospin saturated case.  We apply the method to
146: the ground state of $^{13}$C, $^{14}$C, $^{14}$N, and $^{15}$N. 
147: The results obtained are also valid for the mirror nuclei
148: $^{13}$N, $^{14}$O and $^{15}$O because the electrostatic energy has
149: been not considered in the minimization process.  A systematic
150: analysis of the effects of the different correlations mechanisms
151: included in the wave functions on the total energy and on the
152: contribution of the different channels is carried out.  One and two
153: body densities are reported and the effect of the correlations are
154: discussed.
155: 
156: The scheme of this work is as follows. In Section 2 the variational wave
157: function and the analytical reduction of the expectation values leading to 
158: a form appropriate for the Variational Monte Carlo method are detailed.
159: In Section 3 we report and discuss the main results here obtained. 
160: The conclusions of the present work can be found in Section 4.
161: 
162: \section{Wave function}
163: \label{wavefunction}
164: 
165: The variational trial wave function used in this work is 
166: \begin{equation}
167: \Psi^{\pm}_{JKM}(1,2,\ldots, A)=
168: F_{\mathcal J}(1,\ldots,A)
169: F_{\mathcal L}(1,\ldots,A) 
170: \Phi^{\pm}_{JKM}(1,\ldots,A)
171: \end{equation}
172: 
173: This structure has been used in previous studies of spin and isospin
174: saturated nuclei \cite{bgps02,bgps01}.  It consists of a central
175: Jastrow correlation factor $F_{\mathcal J}$, a linear correlation
176: factor $F_{\mathcal L}$ that can include state dependent correlations,
177: and a model wave function $\Phi^{\pm}_{JKM}$ that is antisymmetric and has
178: the proper values of the total angular momentum and parity. 
179: 
180: The Jastrow factor depends only on the distance between pair of
181: nucleons
182: \begin{equation}
183: F_{\mathcal J}(1,\ldots,A)=\prod_{i<j}^A f(r_{ij}). 
184: \end{equation}
185: The linear factor is defined as
186: \begin{equation}
187: F_{\mathcal L}(1,\ldots,A)=\sum_{i<j}^A g(i,j)  
188: \end{equation}
189: where the function $g(i,j)$ depends on the radial and intrinsic 
190: degrees of freedom of particles, $i,j$. This is the only part of the
191: trial wave function where state dependent correlations are present
192: explicitly.  Here we employ the same parameterization for the
193: correlation functions $g(i,j)$ and $f(r)$, used in previous works
194: \cite{gmn01,bgps02,bgps01} that has shown to provide good
195: results
196: \begin{equation}
197: g(i,j)=\sum_{k=1}^{4}g^{(k)}(r_{ij}){\bf P}^{(k)}(i,j),
198: \end{equation}
199: where
200: \[
201: {\bf P}^{(1)}(i,j)=1,~~~
202: {\bf P}^{(2)}(i,j)=\frac{1}{2}\left(1+
203: \vec{\sigma}_i\cdot\vec{\sigma}_j\right)
204: \]
205: \begin{equation}
206: {\bf P}^{(3)}(i,j)=\frac{1}{2}\left(1+
207:   \vec{\tau}_i\cdot\vec{\tau}_j\right),~~~ {\bf P}^{(4)}(i,j)={\bf
208:   P}^{(2)}(i,j){\bf P}^{(3)}(i,j). 
209: \end{equation}
210: 
211: This operatorial dependence of the correlation factor is the same
212: as that of the nucleon-nucleon interactions considered in this work.
213: The functions $g^{(k)}(r)$, $k=1,..,4$, and $f(r)$ are parameterized
214: as a linear combination of Gaussians
215: \begin{equation}
216: g^{(k)}(r)=\sum_{m=1}^{M} a_m^{(k)}~e^{-b_m r^2},~~~
217: f(r)=1+\sum_{n=1}^{N} c_n~e^{-d_n r^2}.  
218: \end{equation}
219: 
220: The new aspects of treating $A\neq 4n$ nuclei with respect to spin and
221: isospin saturated ones are originated in the angular momentum
222: projection.  Therefore we shall focus here on the model part of the
223: wave function and on the angular momentum projection.  The correlation
224: factors are treated as in the spin and isospin saturated case.
225: 
226: The model wave function used here is based on a generalization
227: of the Margenau-Brink model.  Instead of using only alpha-particle
228: like nucleon clusters, more general groupings are allowed giving rise
229: to a  multicluster description \cite{dude96,vst95}. 
230: Within the molecular viewpoint of the Margenau-Brink scheme,
231: the model wave function is obtained starting from the following functions
232: \begin{equation}
233: \Phi_{\vec{{\bf C}}}(1,2,\ldots, A)=
234: {\mathcal  A} \left\{ 
235: \Phi_1(x_1,\ldots,x_{k_1})
236: \ldots \Phi_n(x_{k_{n-1}-1},\ldots,x_A) 
237: \right\}
238: \label{clusgen}
239: \end{equation}
240: where $\vec{\bf C}\equiv\left\{\vec{c}_k\right\}_{k=1}^n$ is a set of
241: parameters that represent the centers of the clusters, and ${\mathcal
242: A}$ is the corresponding antisymmetrizer.  In this work the
243: arrangement of the nucleons, shown in Fig. \ref{fig1}, consists of three
244: $\alpha$ clusters and a fourth incomplete cluster that can be made of
245: one, two or three nucleons depending on the nucleus under study.  
246: 
247: \begin{figure}[ht]
248: \begin{center}
249: \includegraphics[scale=0.90]{fig01.ps}
250: \caption{\label{fig1}
251: Cluster description of the nuclei in terms of three alpha particles and
252: a general {\em s} incomplete cluster with 1,2 or 3 nucleons}
253: \end{center}
254: \end{figure}
255: 
256: For this configuration, the general form of the function 
257: given in Eq. (\ref{clusgen}) reduces to 
258: \begin{equation}
259: \Phi_{\vec{{\bf C}}}(1,2,\ldots, A)=
260: {\mathcal  A} \left\{ \left[\prod_{m=1}^3
261: \Phi_{\alpha_m}(x_{4m-3},\ldots,x_{4m}) \right]
262: \Phi_s(x_{13},\ldots,x_A)
263: \right\}
264: \label{cluspart}
265: \end{equation}
266: where  $\Phi_{\alpha_m}$ stands for the wave function of an
267: alpha particle centered at $\vec{c}_m$, and $\Phi_s$ represents the
268: incomplete cluster wave function centered at $\vec{c}_s$.  
269: 
270: In this work the  $\Phi_{\alpha_m}$ functions are
271: taken to be Slater determinants built from harmonic oscillator single
272: particle orbitals centered at $\vec{c}_m$
273: \begin{equation}
274: \phi_{\beta, \, \vec{c}}(\vec{r})=
275: \left(\frac{\beta^2}{\pi}\right)^{3/4} \, 
276: e^{-\frac{1}{2} \beta^2 (\vec{r}-\vec{c})^2}
277: \end{equation}
278: 
279: The oscillator parameter, $\beta$, is the same for all of
280: the alpha clusters.  For the incomplete cluster wave function another
281: Slater determinant centered at $\vec{c}_s$ is employed also built from
282: $s$-wave harmonic oscillator single particle orbitals.  The oscillator
283: parameter in this case is, in general, different to that for the
284: $\alpha$ cluster wave function.  The importance of using a different
285: harmonic oscillator parameter will be discussed.  With these choices
286: for the cluster wave functions, the model wave function of the $A$
287: nucleons is a Slater determinant.  In general this function is not
288: eigenfunction of parity or total angular momentum operators.
289: 
290: The following linear combinations 
291: \begin{equation}
292: \label{paridad}
293: \Phi_{\vec{{\bf C}}}^{\pm}(1,2,\ldots, A)= 
294: \Phi_{\vec{{\bf C}}}(1,2,\ldots, A) \pm
295: \Phi_{-\vec{{\bf C}}}(1,2,\ldots, A) 
296: \end{equation}
297: have definite parity.  Model wave functions with the total angular
298: momentum of the state under study can be obtained from
299: Eq. (\ref{paridad}) by using the Peierls-Yoccoz projection operators
300: \cite{peyo57}
301: \begin {eqnarray}
302: \Phi^{\pm}_{JKM}(1,\ldots, A)
303: & = & \frac{2J+1}{8\pi^2}
304: \int d\Theta {\mathcal D}_{MK}^{J*}(\Theta){\bf R}(\Theta) 
305: \Phi^{\pm}_{\vec{\bf C}}(1,\ldots, A)
306: \end{eqnarray}
307: where ${\bf R}(\Theta)$ is the rotation operator, ${\mathcal
308: D}_{MK}^{J*}(\Theta)$ is the rotation matrix and $\Theta$ represents the
309: Euler angles. The quantum number $J$ gives the total angular momentum,
310: $K $ is its projection along the nuclear $z$ axis and $M $ the
311: projection along the $Z$ axis of the laboratory fixed frame. The
312: projection within this scheme is carried out by rotating the intrinsic
313: state and integrating over all angles weighted by the rotation
314: matrix. 
315: 
316: The function $\Phi_{\vec{\bf C}}(1,\ldots, A)$ in Eq. (\ref{cluspart})
317: is the generator function of the model wave functions.  Note that we
318: have removed the parametric dependence of the model wave function on
319: the position of the centers, $\vec{\bf C}$, in order to simplify the
320: notation. The distances between the clusters, $R_c$ and $R_d$, are
321: determined variationally.
322: 
323: The action of the rotation operator on the generator function is now
324: described in detail.  As we have mentioned before, this is the source
325: of the new methodological aspects originated by the fact that the
326: nuclear states are not spin and isospin saturated.  We do not need to
327: consider here the correlation factors because they are rotationally
328: invariant.  The generator function is a  Slater determinant.  The
329: action of the rotation operator on it leads to a
330: linear combination of Slater determinants.  If the Slater determinant
331: is spin and isospin saturated this linear combination contains only
332: one Slater determinant that also is spin and isospin saturated,
333: containing the same single-particle orbitals.  The only difference is
334: that, after rotation, these orbitals depend on the rotated
335: coordinates.  This was exploited previously to study $A = 4n$ nuclei
336: \cite{bgps01}.  When the nuclei are not spin or isospin saturated the
337: rotation gives rise to a mixing of spin states.
338: 
339: When the incomplete shell consists of one nucleon,
340: as for example in the ground state of $^{13}$C, 
341: the action of the rotation operator can be written as follows
342: \begin {eqnarray}
343: {\bf R}(\Theta)\Phi_{\vec{\bf C},\beta s_{\beta}t_{\beta}} & = & 
344: \sum_{s_i=\pm\frac{1}{2}}{\mathcal D}_{s_{\beta},s_i}^{\frac{1}{2}}(\Theta)
345: \overline{\Phi}_{\vec{\bf C},\beta s_i t_{\beta}}
346: \end{eqnarray}
347: where $\beta$ stands for the spatial quantum numbers of the orbital of the
348: incomplete cluster, and $s_\beta$ and $t_\beta$ are the third component 
349: of  spin and isospin, respectively.
350: The over line indicates that the Slater determinant must be evaluated on
351: the rotated coordinates.
352: Therefore, and concerning to the spin dependence of the state, the effect
353: of the rotation is to mix the two possible spin projections of the orbital
354: in the incomplete cluster. The weight of each component is given by the
355: matrix element of the rotation matrix.
356: 
357: When there are two extra nucleons the result of the rotation can be written
358: as follows
359: \begin {eqnarray}
360: {\bf R}(\Theta)
361: \Phi_{\vec{\bf C},\beta s_{\beta}t_{\beta},\beta s_{\gamma}t_{\gamma}} 
362: & = & 
363: \sum_{s_i,s_j=\pm\frac{1}{2}}
364: {\mathcal D}_{s_{\beta},s_i}^{\frac{1}{2}}(\Theta)
365: {\mathcal D}_{s_{\gamma},s_j}^{\frac{1}{2}}(\Theta)
366: \overline{\Phi}_{\vec{\bf C},\beta s_i t_{\beta},\beta s_j t_{\gamma}}
367: \nonumber\\
368:  & = & \sum_{s_i,s_j=\pm\frac{1}{2}} \sum_{S=0,1} 
369: \langle\frac{1}{2}\frac{1}{2}s_{\beta}s_{\gamma}
370: | S,s_{\beta}+s_{\gamma}\rangle\\
371: & &
372: \langle\frac{1}{2}\frac{1}{2}s_i s_j| S,s_i+s_j\rangle
373: {\mathcal D}^{S}_{s_{\beta}+s_{\gamma},s_i+s_j}(\Theta)
374: \overline{\Phi}_{\vec{\bf C},\beta s_i t_{\beta},\beta s_j t_{\gamma}}
375: \nonumber
376: \end{eqnarray}
377: where ($\beta s_{\beta}t_{\beta}$) and ($\beta s_{\gamma}t_{\gamma}$) 
378: stand for the quantum numbers of the orbitals of the incomplete shell.
379: Note that we have considered the same spatial dependence for both 
380: single particle orbitals. 
381: Therefore, if one is dealing with two extra protons ($^{14}$O) or two
382: extra neutrons ($^{14}$C) with the two possible spin orientations the
383: term  $S=1$ vanishes.
384: Only in the case of one proton and one neutron outside closed shell 
385: ($^{14}$N) both total spin components will contribute.
386: 
387: Finally, the case of three nucleons outside closed shell ($^{15}$N and
388: $^{15}$O ) is a conjugate configuration to that of one nucleon outside
389: closed shell and it is handled in the same way.
390: 
391: The values allowed for $J$ and $K$ are governed by the symmetry group
392: of the system. i.e. by the spatial positions of the centers of the
393: clusters.  For the nuclei here considered the group is C$_{3v}$.  The
394: spin of the extra cluster must be also considered in determining the
395: possible values of $K$. If $M_S$ is the total spin third component the
396: allowed $K$ values are given by the selection rule $|K-M_S|=3n$, with
397: $n$ a positive integer \cite{dude96}, and, for any $K$, $ J \geq K$
398: and the parity is $\pi = (-1)^{J \pm S}$.  The energy grows with $K$,
399: providing different rotational bands.  In this work we are concerned
400: only with the ground state, therefore we shall restrict ourselves to
401: $K=1$ for $^{14}$N and $K=0$ for all the rest.  For one and three
402: extra nucleons $M_S=1/2$ and the ground state is $(1/2)^+$, and for
403: two extra nucleons there are two possibilities; i) both nucleons are
404: protons or neutrons $M_S=0$ and the state is $0^+$, ii) one nucleon is
405: a proton and the other a neutron $M_S=0,1$, and the $1^+$ ground state
406: must be constructed with $M_S = 1$ and $K=1$.
407: 
408: In order to compute the expectation value of the Hamiltonian in the
409: projected wave function it is convenient to use the following expression
410: \cite{bgps01,brin66}
411: \begin{equation}
412: \langle \Psi^{\pm}_{JKM}|{\bf H}|\Psi^{\pm}_{JKM}\rangle
413: =\frac{2J+1}{8\pi^2}
414: \int {\mathcal D}^{J*}_{MK}(\Theta)
415: \langle \Phi^{\pm}_{\vec{\bf C}}|
416: F_{\mathcal L} F_{\mathcal J}\, {\bf H} 
417: F_{\mathcal J} F_{\mathcal L}\,{\bf R}(\Theta)
418: |\Phi^{\pm}_{\vec{\bf C}}\rangle
419: \end{equation}
420: 
421: Let us focus on the spin-isospin configuration of the nuclear state.
422: Note that, because of the rotational invariance property of the
423: Hamiltonian, only the ket is rotated remaining the bra on its original
424: configuration. This is important because it determines the
425: configurations that gives non zero contribution to the integral when
426: projected onto the bra.  The action of the rotation operator is to
427: produce a linear combination of configurations containing the original
428: one.  One needs to analyze all of them to determine if, after the
429: action of the spin--isospin operators of $F_{\mathcal L}$ and the
430: Hamiltonian, the original configuration is obtained.  As a result,
431: only the original configuration appearing after rotation contributes
432: with both central and state dependent correlation factors, except
433: except for incomplete clusters made of one proton and one neutron with
434: $S=0$, that we have not studied here, for which two of the
435: configuration appearing after rotation give non zero contribution.
436: Note that the weight factor must be included when doing the integral
437: in all of the cases.  The treatment of state dependent correlations in
438: terms of the intermediate states is not modified with respect to the
439: case of spin and isospin saturated nuclei \cite{bgps00,bgps03}.
440: 
441: \section{Results}
442: \label{results}
443: 
444: First we will test the new methodological aspects implemented in this work
445: by comparing with the results of Dufour and Descouvemont \cite{dude96}
446: obtained by using a different computational scheme.  
447: We will employ for the test both the same nucleon-nucleon interaction 
448: (the Volkov V7 potential), and the same wave function as in \cite{dude96}.
449: It is worth to point out that the correlation factor is not needed
450: because the interaction does not present a strongly repulsive core.
451: 
452: In Table \ref{table1} we show for the ground state and some excited
453: states of the nuclei studied in this work the binding energy and the
454: root mean square radius, $\langle r^2 \rangle^{1/2}$.  As can be seen
455: from the table, both set of results are in a very good agreement.  The
456: spin-orbit interaction is not included in our work and therefore one
457: can not compare directly the results for nuclei with an odd number of
458: nucleons.  For these nuclei we have compared with the average value of
459: the states $1/2^-$ and $3/2^-$ of \cite{dude96}.  This average gives a
460: value that it is very close to the Monte Carlo result of this work,
461: specially for $^{13}$C where the spin-orbit splitting is smaller than
462: in $^{15}$N.  From this test it can be concluded that, for $A\neq 4n$,
463: the angular momentum projection scheme of this work 
464: provide reliable results. 
465: 
466: \begin{table}
467: \caption{\label{table1} Binding energy and root mean square radius,
468: $\langle r^2\rangle^{1/2}$, for different nuclear states calculated in
469: this work (mc) as compared with the results of Dufour and Descouvemont
470: (dd) \cite{dude96}.  Both calculations have been performed by using
471: the Volkov V7 interaction \cite{volk65} and the same variational wave
472: function without correlations.  The inverse of the oscillator
473: parameter, $\beta^{-1}$, and the distances between the clusters, $R_c$ and 
474: $R_d$, are also included.  The energies are in MeV, and $\langle
475: r^2\rangle^{1/2}$, $\beta^{-1}$, and $R_c$ and $R_d$, in fm.  The
476: statistical error in the Monte Carlo calculation is indicated in
477: parentheses.  The Coulomb energy has been included in the total energy.
478: }
479: \begin{tabular}{lllllll}
480: \hline
481: $^{A}$X$(K,J^{\pi})$                  & $\beta^{-1}$ & $R_c, R_d$
482: & $E_{mc}$  & $E_{dd}$ & $\langle r^2\rangle^{1/2}_{mc}$ 
483: & $\langle r^2\rangle^{1/2}_{dd}$ \\
484: \hline
485: $^{12}$C$(0,0^+)$                     & 1.38         & 2.65 
486: &  86.49(4)  & 86.7    & 2.31(7)      & 2.31 \\
487: $^{12}$C$(3,3^-)$                     & 1.38         & 3.14 
488: &  76.41(4)  & 76.5    & 2.49(9)      & 2.49 \\
489: $^{13}$C$(\frac{1}{2},\frac{1}{2}^-)$ & 1.39         & 2.29,2.114 
490: &  88.99(7)  &  89.6   & 2.25(9)      & 2.25 \\
491: $^{14}$C$(0,0^+)$                     & 1.39         & 2.26,2.057 
492: &  102.26(6) &  102.5  & 2.26(7)      & 2.26 \\
493: $^{15}$N$(\frac{1}{2},\frac{1}{2}^-)$ & 1.35         & 1.84,1.887 
494: &  119.37(7) &  121.9  & 2.15(11)     & 2.15 \\
495: $^{16}$O$(0,0^+)$                     & 1.34         & 1.49,2.409 
496: &  147.83(5) &  148.0  & 2.18(3)      & 2.18 \\
497: $^{16}$O$(3,3^-)$                     & 1.37         & 2.24,1.958 
498: &  129.46(10)&  129.8  & 2.27(10)     & 2.26 \\
499: \hline
500: \end{tabular}
501: \end{table}
502: 
503: The ground state of these nuclei has been studied in this work by
504: using a semi realistic potential.
505: We have used the modified Afnan-Tang nuclear potential
506: MS3 \cite{afta68,guar81}.  This is a $v_4$ type interaction with a
507: strongly repulsive core.  It gives meaningless results when used with
508: non correlated trial wave functions.  Thus, in order to analyze the effects
509: of nuclear correlations with respect to the non-correlated case, it is
510: more convenient to use an interaction with a less repulsive short
511: range part as the Brink-Boeker BB1 force \cite{brbo67}.
512: 
513: The ground state energy and the root mean square radius $\langle r^2
514: \rangle^{1/2}$ for different nuclei calculated from a number of trial
515: wave functions by using the BB1 and the MS3 interactions are reported
516: in Tables \ref{table2} and \ref{table2b}, respectively.  The optimal
517: parameters of the trial wave functions are also shown.  The notation
518: is as follows: MB stands for a non correlated trial wave function, JL
519: includes central Jastrow and linear state independent correlations and JLO is a
520: wave function with central Jastrow and state dependent linear
521: correlations.  In the JLO approach we have used the same non-linear
522: parameters as in JL, i.e. the variational freedom is restricted only
523: to the linear parameters of the different operatorial channels.  This
524: scheme has shown to work properly for spin and isospin saturated
525: nuclei \cite{bgps02,bgps01} in such a way that the loss of energy due
526: to this partial optimization was very small.  This is convenient
527: because when state dependent correlations are included, two things
528: happens; first the calculation becomes slower, and second, the
529: statistical error increases.  Therefore it
530: is very convenient, from a computational point of view, that the
531: non-linear parameters can be well determined by means of a state
532: independent optimization. Note that the linear parameters are computed  by
533: solving a generalized eigenvalue problem and then only a long run is
534: required to fix them.  The expectation value of the Coulomb energy
535: $E_c$, not included in the total binding energy, is reported separately.
536:  For the results shown in this work we
537: have used $2^8 \times 10^5$ ($2^6 \times 10^5$) moves per-nucleon with
538: state independent (state dependent) correlated wave functions.
539: 
540: \begin{table}
541: \caption{\label{table2} Ground state energies calculated by using
542: different trial wave functions without correlations (MB), with state
543: independent correlations (JL) and with linear state dependent
544: correlations (JLO) for the BB1 Brink-Boeker potential.  Energies are
545: in MeV, $\langle r^2\rangle^{1/2}$ in fm, $\beta_1, \beta_2$ in
546: fm$^{-1}$ and $R_c, R_d$, in fm.  The statistical error is shown in
547: parentheses.  The Coulomb energy is not included in the total energy.
548: }
549: \begin{tabular}{lllllll}
550: \hline
551: $^{A}$X$(K,J^{\pi})$ 
552: & WF            & $\beta_1, \beta_2$ & $R_c, R_d$ 
553: & $E$           & $E_{c}$              & $\langle r^2\rangle^{1/2} $ \\
554: \hline
555: & MB            & 0.70                 & 3.4 
556: & $-$80.01(4)   & 7.197(1)             & 2.63(4) \\
557: $^{12}$C$(0,0^+)$
558: & JL            & 0.72                 & 3.5
559: & $-$112.36(4)  & 7.417(1)             & 2.53(7)\\
560: & JLO           & 0.72                 & 3.5       
561: & $-$117.68(11) & 7.397(1)             & 2.53(7)\\
562: \hline
563: & MB            & 0.68, 0.59           & 3.5, 3.0
564: & $-$78.29(6)       & 7.057(1)             & 2.71(9) \\
565: $^{13}$C$(\frac{1}{2},\frac{1}{2}^-)$ 
566: & JL            & 0.72, 0.54           & 3.4, 3.0 
567: & $-$112.65(7)  & 7.558(1)             & 2.53(8)\\
568: & JLO           & 0.72, 0.54           & 3.4, 3.0 
569: & $-$119.8(2) & 7.613(2)             & 2.52(15)\\
570: \hline
571: & MB            & 0.69, 0.56           & 3.2, 2.5 
572: & $-$86.36(5)   & 7.363(1)             & 2.64(6) \\
573: $^{14}$C$(0,0^+)$
574: & JL            & 0.74, 0.58           & 3.1, 2.8 
575: & $-$122.93(8) & 7.836(1)             & 2.47(5)\\
576: & JLO           & 0.74, 0.58           & 3.1, 2.8 
577: & $-$131.75(13) & 7.854(1)             & 2.46(8)\\
578: \hline
579: & MB            & 0.68, 0.57           & 3.2, 2.8 
580: & $-$85.09(6)   & 9.849(1)             & 2.65(8) \\
581: $^{14}$N$(1,1^+)$
582: & JL            & 0.71, 0.57           & 3.0, 2.5 
583: & $-$121.68(7) & 10.438(1)             & 2.47(7)\\
584: & JLO           & 0.71, 0.57           & 3.0, 2.5 
585: & $-$131.8(2) & 10.381(2)             & 2.48(10)\\
586: \hline
587: & MB            & 0.66, 0.56           & 3.0, 2.5 
588: & $-$97.69(10)    & 9.948(1)             & 2.65(9) \\
589: $^{15}$N$(\frac{1}{2},\frac{1}{2}^-)$ 
590: & JL            & 0.74, 0.63           & 2.7, 2.4
591: & $-$139.55(10) & 10.821(1)            & 2.39(9)\\
592: & JLO           & 0.74, 0.63           & 2.7, 2.4 
593: & $-$152.0(4) & 10.837(5)            & 2.38(18)\\
594: \hline
595: & MB            & 0.66                 & 2.9, 2.4 
596: & $-$118.70(5)  & 13.470(1)            & 2.60(3) \\
597: $^{16}$O$(0,0^+)_{C_{3v}}$
598: & JL            & 0.76                 & 2.8, 2.4 
599: & $-$166.92(6)  & 14.516(1)            & 2.36(3)\\
600: & JLO           & 0.76                 & 2.8, 2.4 
601: & $-$179.46(10) & 14.515(2)            & 2.35(5)\\
602: \hline
603: & MB            & 0.67                 & 2.8 
604: & $-$118.52(5)  & 13.456(1)            & 2.60(3) \\
605: $^{16}$O$(0,0^+)_t$
606: & JL            & 0.74                 & 2.6 
607: & $-$166.66(6)  & 14.446(2)            & 2.37(4)\\
608: & JLO           & 0.74                 & 2.6
609: & $-$180.61(8)  & 14.552(2) & 2.35(5)\\
610: \hline
611: \end{tabular}
612: \end{table}
613: 
614: \begin{table}
615: \caption{\label{table2b} Ground state energies calculated by using
616: different trial wave functions without correlations (MB), with state
617: independent correlations (JL) and with linear state dependent
618: correlations (JLO) for the modified Afnan-Tang MS3 potential.
619: Energies are in MeV, $\langle r^2\rangle^{1/2}$ in fm $\beta_1,
620: \beta_2$ in fm$^{-1}$ and $R_c, R_d$, in fm.  The statistical error is
621: shown in parentheses.  The Coulomb energy is not included in the total
622: energy.
623: }
624: \begin{tabular}{lllllll}
625: \hline
626: $^{A}$X$(K,J^{\pi})$ 
627: & WF            & $\beta_1, \beta_2$ & $R_c, R_d$ 
628: & $E$           & $E_{c}$              & $\langle r^2\rangle^{1/2} $ \\
629: \hline
630: $^{12}$C$(0,0^+)$
631: & JL            & 0.70                 & 3.5 
632: & $-$74.54(5)  & 7.571(1)             & 2.48(4)\\
633: & JLO           & 0.70                 & 3.5 
634: & $-$87.2(4)    & 7.440(2)             & 2.49(15)\\ 
635: \hline
636: $^{13}$C$(\frac{1}{2},\frac{1}{2}^-)$ 
637: & JL            & 0.70, 0.46           & 3.3, 3.1 
638: & $-$73.37(10)  & 7.833(1)             & 2.47(8)\\
639: & JLO           & 0.70, 0.46           & 3.3, 3.1 
640: & $-$88.6(6)    & 7.864(1)             & 2.44(13)\\
641: \hline
642: $^{14}$C$(0,0^+)$
643: & JL            & 0.69, 0.48           & 3.4, 3.0 
644: & $-$77.52(7)   & 7.840(1)             & 2.50(5)\\
645: & JLO           & 0.69, 0.48           & 3.4, 3.0 
646: & $-$94.6(3)    & 7.840(1)             & 2.44(10)\\
647: \hline
648: $^{14}$N$(1,1^+)$
649: & JL            & 0.69, 0.54           & 3.2, 2.8 
650: & $-$81.95(9)   & 10.699(1)             & 2.42(7)\\
651: & JLO           & 0.69, 0.54           & 3.3, 3.8 
652: & $-$99.3(4)    & 10.865(3)             & 2.37(10)\\
653: \hline
654: $^{15}$N$(\frac{1}{2},\frac{1}{2}^-)$ 
655: & JL            & 0.67, 0.54           & 3.2,2.8 
656: & $-$91.77(12)    & 10.701(1)            & 2.45(9) \\
657: & JLO           & 0.67, 0.54           & 3.2, 2.8 
658: & $-$112.6(6)     & 10.878(3)          & 2.39(15)\\
659: \hline
660: $^{16}$O$(0,0^+)$
661: & JL            & 0.71                 & 2.7 
662: & $-$114.46(7) & 14.827(1)            & 2.32(3)\\
663: & JLO           & 0.71                 & 2.7 
664: & $-$135.6(3) & 15.036(2)            & 2.27(7)\\
665: \hline
666: \end{tabular}
667: \end{table}
668: 
669: The wave functions used in this work includes two different
670: oscillator parameters, one for the complete clusters and another
671: for the incomplete one. 
672: This gives rise to an improvement in the energy of about 3 or 4 MeV
673: when the incomplete cluster is made of one or two nucleons.
674: The improvement is noticeably reduced if the incomplete cluster
675: contains three nucleons.
676: The smaller value for the oscillator parameter of the incomplete cluster is 
677: due to the fact that the nucleons are more localized in the alpha particle
678: cluster than in the incomplete cluster.
679: In general we have obtained oscillator parameters that vary between those 
680: of $^{12}$C and $^{16}$O.
681: 
682: With respect to the optimum parameters of the inter-cluster distances,
683: we have obtained that the distance between the centers of the complete
684: clusters is bigger than the distance between the incomplete cluster
685: and an alpha-particle cluster.  The total energy is not very sensitive
686: to variations of the inter-cluster distances in the neighbourhood of
687: the equilibrium values.  We have indicated such situation by giving
688: these distances with only one decimal digit.  Finally and, as it could
689: be expected, when moving from $A=12$ to $A=15$ the optimal values of
690: the variational parameters tend to those of $^{16}$O.  This is the
691: case for all of the interactions and wave functions analyzed in this
692: work. It is remarkable that the ground state energy of $^{16}$O
693: obtained with the $C_{3v}$ symmetry is practically the same as the
694: one obtained with a tetrahedral symmetry.
695: 
696: In general, the effect of the correlations is to reduce the average
697: size of the nucleus.  Therefore, the optimum values in the model wave
698: function will depend on the presence, or not, of the correlation factor.
699: The modification with respect to the non-correlated
700: wave function is roughly proportional in all of the parameters in such
701: a way that nucleon correlations give rise to an isotropic contraction
702: of the nucleus.
703: 
704: It is interesting to point out the importance of correlations in the
705: binding energy of $^{12}$C and $^{14}$C as compared with $^{13}$C and
706: $^{14}$N, respectively. With both interactions, $^{12}$C is more
707: bounded than $^{13}$C with central correlations, but state dependent
708: correlations reverse this situation, obtaining a difference of 1 and 2
709: MeV with the MS3 and BB1 interaction, respectively. The behaviour of
710: the nuclear binding energy of $^{14}$C and $^{14}$N is different with
711: both potentials. With the BB1 interaction, and without correlations,
712: $^{14}$C is slightly more bounded than $^{14}$N. The difference in
713: their binding energy decreases with the use of central correlations
714: and is zero with state dependent correlations. However, with the MS3
715: potential, $^{14}$N is 4.5 MeV more bounded than $^{14}$C with central
716: and state dependent correlations. The reason of this different
717: behaviour lies in the contribution of the Bartlett and Heisenberg
718: channels of the MS3 interaction, that are null in the BB1
719: potential. Finally it is also worth mentioning here that  we have
720: obtained a negligible effect of the state dependent correlations on
721: the Coulomb energy,  which depends basically on the
722: parameters of the model wave function.
723: 
724:  
725: The correlations increase the binding energy by a quantity which grows
726: with the number of nucleons $A$.  In order to get a deeper insight into
727: the coupling between correlations and the particular nucleus we report
728: in Table \ref{table3} the increment in energy per number of pairs of
729: nucleons.  For example the increase in the binding energy per nucleon
730: pair when state independent correlations are included with respect to
731: the uncorrelated model is given by 
732: \[
733: \Delta_{\rm JL-MB }=
734:  \frac{2}{A(A-1)} \left( E_{\rm JL} -E_{\rm MB} 
735: \right)
736: \]
737: where $E_{\rm JL}$ ($E_{\rm MB}$) is the energy in the JL (MB) model.
738: The quantity $\Delta_{\rm JLO-JL }$ is defined in a similar way.
739: As it can be seen, the increment per number of pairs is roughly constant
740: for all of the nuclei considered, specially $\Delta_{\rm{JLO-JL}}$,
741: that accounts for the effect of state dependent correlations.
742: The increment due to state dependent correlations in the MS3
743: potential is practically twice the increment in the BB1 case.
744: 
745: \begin{table}
746: \caption{\label{table3}
747: Increase in the binding energy per number of nucleon pairs due to the
748: inclusion of different correlation factors for the nuclei studied
749: in this work.
750: In parentheses is indicated the nuclear interaction.
751: The increment is in MeV per number of nucleon pairs.
752: The error is in the last figure.
753: }
754: \begin{tabular}{llll}
755: \hline
756: $^{A}$X$(K,J^{\pi})$ 
757: & $\Delta_{\rm{JL-MB}}$(BB1) &  $\Delta_{\rm{JLO-JL}}$(BB1) 
758: & $\Delta_{\rm{JLO-JL}}$(MS3) \\
759: \hline
760: $^{12}$C$(0,0^+)$                     & -0.49 & -0.08 & -0.19 \\
761: $^{13}$C$(\frac{1}{2},\frac{1}{2}^-)$ & -0.44 & -0.09 & -0.19 \\
762: $^{14}$C$(0,0^+)$                     & -0.40 & -0.10 & -0.19 \\
763: $^{14}$N$(1,1^+)$                     & -0.40 & -0.11 & -0.19 \\
764: $^{15}$N$(\frac{1}{2},\frac{1}{2}^-)$ & -0.40 & -0.12 & -0.19 \\
765: $^{16}$O$(0,0^+)(C_{3v})$             & -0.40 & -0.12 & -0.18 \\
766: \hline
767: \end{tabular}
768: \end{table}
769: 
770: A more detailed analysis of the effect of the state dependent
771: correlations on the energy can be done by looking at the contribution
772: of the kinetic energy and of the different channels of the potential
773: energy. In Fig. 2 we plot the differences between these quantities
774: calculated with the JL and JLO wave functions for both the BB1 and the
775: MS3 interactions. Both the kinetic energy and the energy of the Wigner
776: channel rise with state dependent correlations for both
777: potentials. This increase is more important for the kinetic energy
778: with the MS3 potential than with the BB1 one, whereas the opposite
779: holds for the energy of the Wigner energy. For the BB1 potential, the
780: Majorana channel is the responsible for the decrease in the
781: ground state energy when state dependent correlations are considered. For
782: the MS3 interaction, the effect on the Majorana channel is practically
783: canceled with that on the kinetic and Wigner energies, and the
784: Bartlett and Heisenberg channels make the nuclei more bounded.  The
785: contribution of these two channels is very close and is nearly
786: independent of the nucleus considered.
787: 
788: \begin{figure}[ht]
789: \begin{center}
790: \includegraphics[scale=1.25]{fig02.ps}\hspace{0.1cm}
791: \caption{\label{fig2} Increase in the total energy, the expectation
792: values of the kinetic energy and the different channels of the
793: interacting potential when state dependent are included with respect
794: to the JL approximation.  In the left hand panel we plot the results
795: for the BB1 potential and in the right hand one for the MS3
796: potential. The lines are for guiding the eyes
797: }
798: \end{center}
799: \end{figure}
800: 
801: The one-- and two-- body densities give the spatial distribution of
802: the nucleons in the nuclei.
803: Here we have calculated these densities to analyze the effect of the
804: different correlation mechanisms introduced in the variational
805: wave functions.
806: In Fig.  \ref{fig3} we show the one body nuclear density 
807: calculated with the JL wave function for all of the
808: nuclei here studied and the two interactions considered.
809: As it can be seen, the qualitative behaviour is similar for both
810: potentials, with a higher value of the maximum as the number of
811: nucleons increases.
812: It is also worth pointing out that as $A$ increases, the density tends
813: to that of $^{16}$O.
814: It is for this nuclei and the MS3 interaction where this density is 
815: more separated from the others.
816: 
817: \begin{figure}[ht]
818: \begin{center}
819: \includegraphics[scale=1.25]{fig03.ps}\hspace{0.1cm}
820: \caption{\label{fig3}
821: One body density for all the the nuclei studied in this work
822: calculated with the JL wave function.
823: In the left hand panel we plot the results for the BB1 potential and
824: in the right hand one for the MS3 potential.
825: }
826: \end{center}
827: \end{figure}
828: 
829: The effect of the state dependent correlations on the one--body density
830: for these nuclei is studied in Fig. \ref{fig4}, where we plot the
831: difference between the single particle density obtained with the JL
832: and the JLO wave functions.  The first noticeable fact is that the
833: general behaviour is different for the two interactions used here.
834: Thus at short distances state dependent correlations tend to increase
835: the density with the BB1 interaction and the opposite happens with the MS3
836: potential, except for $^{12}$C for which a negative region at short
837: distances appears.  In addition, for the BB1 potential, the effect of
838: the operatorial correlations is roughly independent of the nucleus
839: while for the MS3 potential effects of the operatorial correlations
840: show a more accused dependence on the nucleus.
841: 
842: \begin{figure}[ht]
843: \begin{center}
844: \includegraphics[scale=1.25]{fig04.ps}\hspace{0.1cm}
845: \caption{\label{fig4}
846: Effect of the state dependent correlations on the one body density
847: for the different nuclei considered in this work.
848: In the left hand panel we plot the results for the BB1 potential and
849: in the right hand one for the MS3 potential.
850: }
851: \end{center}
852: \end{figure}
853: 
854: The effects of correlations are more important on the two body density
855: than in the one body density.  In Fig. \ref{fig5} we plot the two body
856: density obtained from the state independent correlated wave function
857: JL for all of the nuclei studied and the two interactions considered
858: in this work.  The behaviour of this density is very similar for both
859: potentials, although the effect of the nuclear core is much more
860: important in the MS3 potential.  The main difference is that with the
861: MS3 interaction shorter distances are favoured with respect to the BB1
862: potential.  At distances between 2 and 3.5 fm the differences among
863: the nuclei considered are more important, with bigger values as the
864: number of nucleons increases from $^{12}$C to $^{16}$O.  This can be
865: understood as a progressive filling of the incomplete cluster that
866: gives rise to a larger number of particles at these intermediate
867: distances.
868: 
869: \begin{figure}[ht]
870: \begin{center}
871: \includegraphics[scale=1.25]{fig05.ps}\hspace{0.1cm}
872: \caption{\label{fig5}
873: Two body density for all the the nuclei studied in this work
874: calculated with the JL wave function.
875: In the left hand panel we plot the results for the BB1 potential and
876: in the right hand one for the MS3 potential.
877: }
878: \end{center}
879: \end{figure}
880: 
881: Finally, the effect of including state dependent correlations on this
882: density is studied in Fig. \ref{fig6} where we plot the difference
883: between the two--body density calculated from the JL and JLO wave
884: functions.  As it was the case for the one body density, the effect of
885: state dependent correlations is roughly independent of the nucleus
886: when the BB1 potential is used and a more accused dependence is
887: observed for the MS3 interaction. For both potentials, state dependent
888: correlations bring together nucleons with respect to the JL case.
889: 
890: \begin{figure}[ht]
891: \begin{center}
892: \includegraphics[scale=1.25]{fig06.ps}\hspace{0.1cm}
893: \caption{\label{fig6}
894: Effect of the state dependent correlations on the two body density
895: for the different nuclei considered in this work.
896: In the left hand panel we plot the results for the BB1 potential and
897: in the right hand one for the MS3 potential.
898: }
899: \end{center}
900: \end{figure}
901: 
902: \section{Conclusions}
903: \label{conclusions}
904: 
905: Variational Monte Carlo calculations for $p$-shell, $A\neq 4n$, nuclei
906: starting from the nucleon-nucleon interaction have been presented.
907: The ground state energy and the one and two body densities have been
908: calculated.  The variational wave function consists of three factors:
909: a central Jastrow term, a spin-isospin dependent linear term and a model wave
910: function.  The model wave function is based on a cluster model
911: allowing for the formation of different kind of nucleon clusters with
912: centers at fixed positions.  The Peierls-Yoccoz projection operators
913: have been used in order to obtain trial wave functions with the proper
914: values of the angular momentum.  This work extend previous ones
915: carried out for spin and isospin saturated nuclei.
916: 
917: The present scheme has shown to be appropriate for describing two
918: important and complementary aspects of the nuclear dynamics as the
919: short range correlations and the formation of nucleon clusters.  The
920: former is induced by the short range repulsive part of the nuclear
921: potential while the later is a collective effect due to the medium and
922: long range part of the interaction.
923: 
924: In this work, an analytical reduction of the expectation values for
925: $A\neq 4n$ nuclei is presented. 
926: The use of the Peierls-Yoccoz projection operators introduces new 
927: features when the nuclei are not spin and isospin saturated.
928: Here we obtain a final form of the expectation values
929: which is specially suited for the Variational Monte Carlo calculation.
930: This is done for both state independent and state dependent correlation
931: factors.
932: As a result the different expectation values can be computed with
933: no significant extra computational cost with respect to the case of spin
934: and isospin saturated nuclei.
935: 
936: The scheme is applied to several nuclei with $12 \leq A \leq 16$.
937: The methodology has been first tested against previous works using
938: a completely different scheme of calculation.
939: Then results obtained by using two different nucleon-nucleon potentials 
940: including a repulsive core at short distances and state-dependent 
941: interaction channels have been reported.
942: The binding energies and the root mean square radius along with the
943: optimal parameters of the wave functions 
944: are shown for the different nuclei and states considered here. 
945: The effect of the different correlation mechanisms included in the trial
946: wave function on the energy and on the equilibrium geometries is
947: discussed.
948: The importance of using different oscillator parameters for the different
949: kind of nucleon clusters is shown.
950: The effect of the correlations on the different interaction channels
951: is analyzed in terms of the number of nucleons.
952: Finally one and two body densities obtained for the nuclei here studied
953: with several approximations of the wave functions are reported and
954: discussed.
955: 
956: \section*{Acknowledgements}
957: 
958: This work has been partially supported by the Ministerio de Ciencia y
959: Tecnolog\'{\i}a and FEDER under contract BFM2002-00200, and by the
960: Junta de Andaluc\'{\i}a.
961: 
962: \begin{thebibliography}{35}
963: \expandafter\ifx\csname furl\endcsname\relax
964: \def\url#1{\texttt{#1}}\fi
965: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
966: 
967: \bibitem{bishop78}
968: R. F. Bishop, C. Howes, J. M. Irvine and M. Modarres, 
969: J Phys G 4 (1978) 1709, L89 and L123.  
970: 
971: \bibitem{bosca88} 
972: M. C. Bosc\'a and R. Guardiola, 
973: Nucl. Phys. A 476 (1988) 471 and A 489 (1988) 45 
974: 
975: \bibitem{fabro98}
976: A.  Fabrocini, F. Arias de Saavedra, G. Co' and P. Folgarait, 
977: Phys. Rev. C 57 (1998) 1668 
978: 
979: \bibitem{fabro00}
980: A. Fabrocini, F. Arias de Saavedra and G. Co', 
981: Phys. Rev C 61  (2000) 044302 
982: 
983: \bibitem{carlson88}
984: J. Carlson, 
985: Phys Rev C36, 2026 (1987) and C38 (1988) 1879 
986: 
987: \bibitem{pieper90}
988: S. C. Pieper, R. B. Wiringa and V. R. Pandharipande, 
989: Phys. Rev. Lett. 64 (1990) 364 
990: 
991: \bibitem{zabo81}
992: J. G. Zabolitzky, Phys. Lett. B 100  (1981) 5.  
993: 
994: \bibitem{feldmeier98}
995: H. Feldmeier, T. Neff, R. Roth and J. Schnack, 
996: Nucl. Phys. A 632 (1998) 61
997: 
998: \bibitem{heisen99}
999: J. H. Heisenberg and B. Mihaila, 
1000: Phys. Rev. C 59 (1999) 1440.  
1001: 
1002: \bibitem{bishop90}
1003: R. F. Bishop, M. F. Flynn, M. C. Bosc\'a, E. Buend\'ia and R. Guardiola,
1004: Phys. Rev. C 42 (1990)  1341.  
1005: 
1006: \bibitem{guar96}
1007: R. Guardiola, P. I. Moliner, J. Navarro,
1008: R. F. Bishop, A. Puente and N. R. Walet, 
1009: Nucl Phys A 609 (1996)  218.
1010: 
1011: 
1012: 
1013: \bibitem{dedu97}
1014: P. Descouvemont, M. Dufour,
1015: Nucl. Phys. A 621 (1997) 311c
1016: 
1017: \bibitem{dude97}
1018: M. Dufour, P. Descouvemont,
1019: Phys. Rev. C 56 (1997) 1831
1020: 
1021: \bibitem{arima72}
1022: A. Arima, H. Horiuchi, K. Kubodera and N. Takigawa, 
1023: Adv. Nucl. Phys. 5 (1972) 345.
1024: 
1025: \bibitem{michel98}
1026: F. Michel, S. Okkubo and G. Reidemeister, 
1027: Prog. Theor. Phys. Suppl. 132 (1998) 7.
1028: 
1029: \bibitem{dude96}
1030: M. Dufour, P. Descouvemont,
1031: Nucl. Phys. A 605 (1996) 160
1032: 
1033: \bibitem{desc02}
1034: P. Descouvemont,
1035: Nucl. Phys. A 709 (2002) 275
1036: 
1037: \bibitem{volk65} 
1038: A.B. Volkov,
1039: Nucl. Phys. 74 (1965) 33
1040: 
1041: \bibitem{vst95}
1042: K. Varga, Y. Suzuki, I. Tanihata,
1043: Phys. Rev. C 52 (1995) 3013
1044: 
1045: \bibitem{oasv00}
1046: Y. Ogawa, K. Arai, Y. Suzuki, K. Varga,
1047: Nucl. Phys. A 673 (2000) 122
1048: 
1049: \bibitem{gmn01}
1050: R. Guardiola, I. Moliner, M.A. Nagarajan, 
1051: Nucl. Phys. A 670 (2001) 393
1052: 
1053: \bibitem{bgps02}
1054: E. Buend\'{\i}a, F.J. G\'alvez, J. Praena, A. Sarsa,
1055: Nucl. Phys. A 710 (2002) 29
1056: 
1057: \bibitem{bgps01}
1058: E. Buend\'{\i}a, F.J. G\'alvez, J. Praena, A. Sarsa,
1059: J. Phys. G: Nucl. Part. Phys. 27 (2001) 2211
1060: 
1061: \bibitem{peyo57} 
1062: R.E. Peierls, J. Yoccoz, 
1063: Proc. Roy. Soc. (London) A 70 (1957) 381.
1064: 
1065: \bibitem{brin66}
1066: D.M. Brink, in: Proc. Int. School of Physics, Enrico Fermi 36, Academic
1067: Press, New York, 1966, p. 247.
1068: 
1069: \bibitem{bgps00}
1070: E. Buend\'{\i}a, F.J. G\'alvez, J. Praena, A. Sarsa,
1071: J. Phys. G 26 (2000) 1795
1072: 
1073: \bibitem{bgps03} 
1074: E. Buend\'{\i}a, F.J. G\'alvez, J. Praena, A. Sarsa, in: Horizon
1075: in World Physics, Nova Science Publisher Inc., New York, 2003, p. 15.
1076: 
1077: \bibitem{afta68} 
1078: I.R. Afnan, Y.C. Tang,
1079: Phys. Rev. 175 (1968) 1337.
1080: 
1081: \bibitem{guar81} 
1082: R. Guardiola, in: Recent Progress in Many-Body Theories, in Lectures
1083: Notes in Physics Vol 142, 1981, p. 398
1084: 
1085: \bibitem{brbo67} 
1086: D.M. Brink, B. Boeker,
1087: Nucl. Phys. A 91 (1967) 1.
1088: 
1089: 
1090: \end{thebibliography}
1091: 
1092: \end{document}
1093: