1: \documentstyle[aps]{revtex}
2: \tolerance=1000
3: \draft
4: \setlength{\oddsidemargin}{0.1 cm}
5: \setlength{\topmargin}{-2 cm}
6: \setlength{\textheight}{24.5 cm}
7: \setlength{\textwidth}{16 cm}
8: \def\baselinestretch{2.0}
9: \begin{document}
10: \title{Generator coordinate method calculations of one-nucleon removal
11: reactions on $^{40}$Ca}
12: \author{M.V. Ivanov, M.K. Gaidarov, and A.N. Antonov}
13: \address{Institute of Nuclear Research and Nuclear Energy,
14: Bulgarian Academy of Sciences, Sofia 1784, Bulgaria}
15: \author{C. Giusti}
16: \address{Dipartimento di Fisica Nucleare e Teorica,
17: Universit\`a di Pavia,\\
18: Istituto Nazionale di Fisica Nucleare, Sezione di Pavia,
19: Pavia, Italy}
20: \maketitle
21:
22: \begin{abstract}
23: An approach to the Generator Coordinate Method (GCM) using
24: Skyrme-type effective forces and Woods-Saxon construction
25: potential is applied to calculate the single-particle proton and
26: neutron overlap functions in $^{40}$Ca. The
27: relationship between the bound-state overlap functions and the
28: one-body density matrix has been used. These overlap functions
29: are applied to calculate the cross sections of one-nucleon
30: removal reactions such as ($e,e'p$), ($\gamma,p$) and ($p,d$) on
31: $^{40}$Ca on the same theoretical footing. A consistent description
32: of data for the different reactions is achieved.
33: The shapes of the experimental cross sections for transitions to the
34: $3/2^{+}$ ground state and the first $1/2^{+}$ excited state of the
35: residual nuclei are well reproduced by the overlap functions obtained
36: within the GCM. An additional spectroscopic factor accounting for
37: correlations not included in the overlap function must be applied to
38: the calculated results to reproduce the size of the experimental
39: cross sections.
40:
41: \end {abstract}
42: \vspace{1cm}
43:
44: \section{Introduction}
45: Experiments on nuclear reactions accompanied by one-nucleon
46: removal from $^{40}$Ca (e.g. \cite{Roos75,Kra90,Lap93,Abe92}) have
47: accumulated much spectroscopic information on its nucleon-hole
48: spectral density function and, generally, on the single-particle
49: aspects of nuclear structure. From the theoretical point of view
50: two topics in the analyses of these processes are of significant
51: interest and have been mainly studied: the reaction mechanism and
52: the ground-state correlation effects. The latter can be
53: successfully considered by using the unique relationship between
54: the overlap functions (OF) related to bound states of the
55: $(A-1)$-particle system and the one-body density matrix (ODM) of
56: the $A$-particle system in its ground state \cite{Vn93}. This
57: makes it possible to investigate the effects of the various types
58: of nucleon-nucleon correlations included in the ODM on the
59: bound-state proton and neutron overlap functions.
60:
61: In our recent works \cite{Gai99,Gai2000} a consistent study of
62: overlap functions and one-nucleon removal reactions on $^{16}$O
63: using different correlation methods has been carried
64: out and the comparison with the experimental data has been
65: performed. In our previous calculations we used methods which
66: account mainly for short-range and tensor nucleon-nucleon
67: correlations. It is desirable, however, to take into
68: consideration also correlations originating from the collective
69: motion of the nucleons. This was partially done for the $^{16}$O
70: nucleus in \cite{Gai2000}. In this respect, the various
71: applications of the Generator Coordinate Method to nuclear
72: problems \cite{Chr86,Ant,Iva2000} have shown its efficiency as a
73: potential source of information on nucleon-nucleon correlations in
74: nuclei. The results on the one- and two-body density and momentum
75: distributions, occupation probabilities and natural orbitals
76: obtained within the GCM using various construction potentials
77: \cite{Iva2000} have shown that the nucleon-nucleon correlations
78: accounted for in this method are different from the short-range
79: ones and are rather related to the collective motion of the
80: nucleons. It was pointed out that these correlations are also
81: important in calculations of other single-particle properties,
82: such as one-body overlap functions, which are necessary in the
83: calculations of cross-sections of one-nucleon removal reactions.
84:
85: It is known that, in general, when going from light to heavier
86: systems the collectivity becomes stronger. Because of its
87: undeformed closed-shell structure, the medium-heavy $^{40}$Ca
88: nucleus is one of the few nuclei for which microscopic
89: calculations can be performed. For instance, only recently the
90: Fermi hypernetted chain theory has been extended to describe the
91: ground-state properties of $^{40}$Ca \cite{Sa96}. Therefore, it
92: is very attractive to probe its "doubly-magic" character also in
93: calculations of different types of one-nucleon removal reactions.
94: As a first step differential cross sections for the
95: $^{40}$Ca($p,d$) pick-up reaction have been calculated in
96: \cite{Di97} with overlap functions obtained from the ODM in the
97: Jastrow correlation method (JCM) \cite{Sto96}. Although only
98: short-range correlations have been included in the OF, it was
99: shown that the angular distributions obtained are in a qualitative
100: agreement with the empirical ($p,d$) data for the transition to
101: the ground state of the residual nucleus. These results have
102: clearly pointed out the necessity of inclusion of another kind of
103: NN correlations, namely the long-range correlations,
104: corresponding to collective degrees of freedom. Better agreement
105: with the experimental $^{16}$O($e,e'p$) and $^{16}$O($\gamma,p$)
106: data was achieved in \cite{Gai2000}, where it was concluded that
107: the long-range correlations affect the spectroscopic factors
108: causing an additional depletion of the quasihole states.
109:
110: The main aim of the present work is to study the effects of the
111: nucleon-nucleon correlations included in the Generator Coordinate
112: Method on the behaviour of the bound-state proton and neutron
113: overlap functions in $^{40}$Ca and of the related one-nucleon
114: removal reaction cross sections. We first calculate both proton
115: and neutron single-particle overlap functions and spectroscopic
116: factors on the basis of the corresponding proton (with included
117: Coulomb interaction) and neutron one-body density matrices of the
118: $^{40}$Ca nucleus obtained with GCM using the relationship
119: between the ODM's and the overlap functions. Second, these
120: bound-state OF are used to calculate the cross sections of the
121: $^{40}$Ca($e,e'p$), $^{40}$Ca($\gamma,p$) and $^{40}$Ca($p,d$)
122: reactions. Thus it becomes possible to investigate the role of
123: the correlations related to collective nucleon motions and
124: accounted for in the GCM on the overlap functions and one-nucleon
125: removal cross sections in $^{40}$Ca also in comparison with data.
126:
127: In Section II we give the basic theoretical relationships
128: necessary to obtain the one-body density matrix in GCM and the
129: bound-state overlap functions by means of the asymptotic
130: behaviour of ODM. The results of the calculations of overlap
131: functions and cross sections of ($e,e'p$), ($\gamma,p$) and
132: ($p,d$) reactions on $^{40}$Ca are presented and discussed in
133: Section III. Section IV contains the concluding remarks.
134:
135: \section{The theoretical scheme}
136: \subsection{The GCM ground-state one-body density matrix}
137: We start from a standard GCM-type $A$-particle wave function
138: $\Psi$ when one real generator coordinate $x$ is considered
139: \cite{Gri57}, i.e.
140: %
141: \begin{equation}
142: {\Psi}(\{{\bf{r}}_i\})=
143: \int_{0}^{\infty}{f(x){\Phi}(\{{\bf{r}}_i\},x)dx}\,,\;\;i=1,...,A.
144: \label{eq:psi}
145: \end{equation}
146: %
147: In Eq. (\ref{eq:psi}) ${\Phi}(\{{\bf{r}}_i\},x)$ is the generating
148: function, $f(x)$ is the generator or weight function and $A$ is
149: the mass number of the nucleus.
150:
151: The application of the Ritz variational principle ${\delta}E=0$
152: leads to the Hill-Wheeler integral equation \cite{Gri57,Hil53} for
153: the weight function and the energy of the system:
154: %
155: \begin{equation}
156: \int_{0}^{\infty}{[{\cal H}(x,x')-EI(x,x')]f(x')dx'}=0\,,
157: \label{eq:HW}
158: \end{equation}
159: %
160: where
161: %
162: \begin{equation}
163: {\cal H} (x,x')=\langle {\Phi}(\{{\bf{r}}_i\},x)|\hat{H}|{\Phi}
164: (\{{\bf{r}}_i\},x')\rangle\,
165: \label{eq:ek}
166: \end{equation}
167: %
168: and
169: %
170: \begin{equation}
171: I(x,x')=\langle
172: {\Phi}(\{{\bf{r}}_i\},x)|{\Phi}(\{{\bf{r}}_i\},x')\rangle \,
173: \label{eq:ok}
174: \end{equation}
175: %
176: are the energy and overlap kernels, respectively, and $\hat{H}$
177: is the Hamiltonian of the system. Solving the Hill-Wheeler
178: equation (\ref{eq:HW}) one can obtain the solutions
179: ${f}_0,{f}_1,...$ for the weight functions which correspond to
180: the eigenvalues of the energy ${E}_0,{E}_1...$.
181:
182: The corresponding ground-state one-body density matrix is given
183: by (see, e.g. \cite{Ant}):
184: %
185: \begin{equation}
186: \rho({\bf{r}},{{\bf{r'}}})=\int\!\!\int
187: {f}_{0}(x){f}_{0}(x')I(x,x'){\rho}(x,x',{\bf{r}},{\bf{r'}})dxdx'
188: \label{eq:obdm}
189: \end{equation}
190: %
191: with
192: %
193: \begin{equation}
194: {\rho}(x,x',{\bf{r}},{{\bf{r'}}})=4\sum_{{\lambda},{\mu}=1}^{A/4}
195: {{{(N^{-1}(x,x'))}_{{\mu}{\lambda}}}{{\varphi}_{\lambda}^{*}}({\bf{r}},x)
196: {{\varphi}_{\mu}}({{\bf{r'}}},x')}\,,
197: \label{eq:nmatr}
198: \end{equation}
199: %
200: where $\varphi_{\lambda}({\bf r},x)$ are the single-particle wave
201: functions corresponding to a given construction potential by means
202: of which the generating single Slater determinant wave function
203: ${\Phi}(\{{\bf{r}}_i\},x)$ is built. The matrix
204: $N_{\lambda\mu}^{-1}(x,x^{\prime })$ in Eq. (\ref{eq:nmatr}) is
205: the inverse matrix of
206: %
207: \begin{equation}
208: {N}_{{\lambda}{\mu}}(x,x')=
209: \int_{}^{}{{\varphi}_{\lambda}^{*}({\bf{r}},x){\varphi}_{\mu}({\bf{r}},x')}
210: d{\bf{r}}\,.
211: \label{eq:lamu}
212: \end{equation}
213: %
214:
215: As known, the results of the GCM calculations depend on the type
216: of the construction potential used to define the single-particle
217: wave functions $\varphi_{\lambda}({\bf r},x)$. In the present
218: work we choose the Woods-Saxon (WS) potential as a construction
219: potential with the diffuseness parameter as a generator
220: coordinate. The GCM scheme, which has been already adopted in
221: \cite{Gai2000}, gives a very good agreement with the data for the
222: $^{16}$O($e,e'p$) and $^{16}$O($\gamma,p$) reaction cross
223: sections. In our present calculations the Skyrme-type effective force
224: $SkM^{*}$ \cite{Bar82} is used, with parameters which give
225: realistic binding energy of $^{40}$Ca obtained from the
226: Hill-Wheeler equation. The agreement of the calculated
227: basic nuclear characteristics with their empirical values
228: obtained in \cite{Iva2000}, proved once more the reliability of
229: these effective forces, which are used also in the present study.
230:
231: \subsection{The overlap functions for the proton and neutron bound states}
232: For a correct calculation of the cross section of nuclear
233: reactions with one-neutron or one-proton removal from the target
234: nucleus, the corresponding overlap functions for the neutron and
235: proton bound states must be used in the reaction amplitudes. The
236: construction of the ground-state ODM of $^{40}$Ca from the
237: Generator Coordinate Method calculations makes it possible to
238: apply the procedure for extracting single-particle overlap
239: functions \cite{Vn93}. Here we would like to remind that the
240: single-particle overlap functions are defined by the overlap
241: integrals between eigenstates of the $A$-particle and the
242: $(A-1)$-particle systems:
243: %
244: \begin{equation}
245: \phi _{\alpha }({\bf r})=\langle \Psi _{\alpha }^{(A-1)}|a({\bf
246: r})|\Psi ^{(A)}\rangle ,
247: \label{eq:OF}
248: \end{equation}
249: %
250: where $a({\bf r})$ is the annihilation operator for a nucleon
251: with spatial coordinate ${\bf r}$ (spin and isospin operators are
252: implied). In the mean-field approximation $\Psi ^{(A)}$ and $\Psi
253: _{\alpha }^{(A-1)}$ are single Slater determinants, and the
254: overlap functions are identical with the mean-field
255: single-particle wave functions, while in the presence of correlations both
256: $\Psi ^{(A)}$ and $\Psi _{\alpha }^{(A-1)}$ are complicated superpositions of
257: Slater determinants. In general, the overlap functions
258: $(\ref{eq:OF})$ are not orthogonal. Their norm defines the
259: spectroscopic factor
260: %
261: \begin{equation}
262: S_{\alpha }=\langle \phi _{\alpha }|\phi _{\alpha }\rangle .
263: \label{eq:spf}
264: \end{equation}
265: %
266: The normalized overlap function associated with the state $\alpha
267: $ then reads
268: %
269: \begin{equation}
270: \tilde{\phi}_{\alpha }({\bf r})=S_{\alpha }^{-1/2}\phi _{\alpha
271: }({\bf r}). \label{eq:NOF}
272: \end{equation}
273: %
274: The one-body density matrix (e.g. Eq. $(\ref{eq:obdm})$) can be
275: expressed in terms of the overlap functions in the form:
276: %
277: \begin{equation}
278: \rho ({\bf r},{\bf r^{\prime }})=\sum_{\alpha }\phi _{\alpha }^{*}({\bf r}%
279: )\phi _{\alpha }({\bf r^{\prime }})=\sum_{\alpha }S_{\alpha }\tilde{\phi}%
280: _{\alpha }^{*}({\bf r})\tilde{\phi}_{\alpha }({\bf r^{\prime }}).
281: \label{eq:rho}
282: \end{equation}
283: %
284:
285: It is known (e.g. \cite{Mah91}) that the overlap function
286: associated with the bound state ($\alpha \equiv nlj$) of the
287: $(A-1)$- or $(A+1)$-nucleon system is an eigenstate of a
288: single-particle Schr\"{o}dinger equation in which the mass
289: operator plays the role of a potential. Due to its finite range,
290: the asymptotic behaviour of the radial part of the neutron
291: overlap functions for the bound states of the $(A-1)$-system is
292: given by \cite{Vn93,Bang85}:
293: %
294: \begin{equation}
295: \phi_{nlj}(r)\rightarrow C_{nlj}\exp(-k_{nlj}r)/r,
296: \label{eq:nasym}
297: \end{equation}
298: %
299: where $k_{nlj}$ is related to the neutron separation energy
300: %
301: \begin{equation}
302: k_{nlj}=\frac {\sqrt{2m\epsilon_{nlj}}}{\hbar}, \;\;\;
303: \epsilon_{nlj}=E_{nlj}^{(A-1)}-E_{0}^{A}.
304: \label{eq:decay}
305: \end{equation}
306: %
307: For proton bound states, due to an additional long-range part
308: originating from the Coulomb interaction, the asymptotic
309: behaviour of the radial part of the corresponding proton overlap
310: functions reads
311: %
312: \begin{equation}
313: \phi_{nlj}(r)\rightarrow C_{nlj}\exp[-k_{nlj} r-\eta \ln
314: (2k_{nlj}r)]/r,
315: \label{eq:pasym}
316: \end{equation}
317: %
318: where $\eta $ is the Coulomb (or Sommerfeld) parameter and $k_{nlj}$ in
319: (\ref{eq:decay}) contains in this case the mass of the proton and the proton
320: separation energy.
321:
322: Taking into account Eqs. (\ref{eq:rho}) and (\ref{eq:nasym}), the
323: lowest ($n=n_{0}$) neutron bound-state $lj$-overlap function is
324: determined by the asymptotic behaviour of the associated partial
325: radial contribution of the one-body density matrix
326: $\rho_{lj}(r,r^{\prime})$ ($r^{\prime}=a\rightarrow \infty $) as
327: %
328: \begin{equation}
329: \phi _{n_{0}lj}(r)={\frac{{\rho _{lj}(r,a)}}{{C_{n_{0}lj}~\exp
330: (-k_{n_{0}lj}\,a})/a}}~,
331: \label{eq:nof}
332: \end{equation}
333: %
334: where the constants ${C_{n_{0}lj}}$ and ${k_{n_{0}lj}}$ are
335: completely determined by $\rho_{lj}(a,a)$. In this way the
336: separation energy $\epsilon_{n_{0}lj}$ and the spectroscopic
337: factor $S_{n_{0}lj}$ can be determined as well. Similar
338: expression for the lowest proton bound-state overlap function can
339: be obtained having in mind its proper asymptotic behaviour
340: (\ref{eq:pasym}).
341:
342: The applicability of this theoretical procedure has been
343: demonstrated in Refs. \cite{Gai99,Gai2000,Sto96}. In particular,
344: it has been shown \cite{Gai2000} that the substantial realistic
345: inclusion of short-range as well as tensor correlations in the
346: ODM of $^{16}$O constructed within the Green Function Method
347: \cite{Po96} and, as a consequence, in the extracted overlap
348: functions, leads to a fair and consistent description of the
349: experimental cross sections of the $^{16}$O$(e,e^{\prime}p)$ and
350: $^{16}$O$(\gamma,p)$ reactions. The importance of collective long-range
351: correlations on the additional spectroscopic factors extracted in
352: comparison with data has been also pointed out in \cite{Gai2000}.
353:
354: In this work we use as a basis the GCM results for the correlated
355: states of $^{40}$Ca obtained in \cite{Iva2000}, where
356: correlations due to collective nucleon motion are properly taken
357: into account. The procedure to calculate the overlap functions
358: from the one-body density matrix outlined above is performed
359: numerically for $^{40}$Ca within the GCM. In this respect we
360: would like to note that an important condition of this procedure
361: is the exponential asymptotics of the overlap functions at
362: $r^{\prime}\rightarrow \infty$ (see Eqs. (\ref{eq:nasym}) and
363: (\ref{eq:pasym})), which is related to the correct asymptotics of
364: $\rho_{lj}(r,r^{\prime})$ at $r^{\prime}\rightarrow \infty$. This
365: condition is fulfilled in our GCM approach by using realistic
366: single-particle Woods-Saxon wave functions in the generating
367: single Slater determinant wave function
368: ${\Phi}(\{{\bf{r}}_i\},x)$ in Eq. (\ref{eq:psi}). The exponential
369: decay of the partial radial ODM is with a decay constant
370: $k_{nlj}$ related to the separation energy (Eq.
371: (\ref{eq:decay})). The question about the correct asymptotic
372: behaviour of the extracted overlap functions will be discussed
373: also in Section III.
374:
375: \section{Results and discussion}
376: In the present paper the GCM numerical calculations have been
377: performed using Woods-Saxon construction potential with
378: diffuseness parameter as a generator coordinate. The values of
379: the radius and the depth of the construction potential have been
380: fixed. Following \cite{BH91}, the value of the depth used for
381: both neutron and proton cases has been taken to be 50 MeV, while
382: the corresponding values of the potential radius are 1.30 fm for
383: the neutron and 1.24 fm for the proton case, respectively. The
384: effective $SkM^{*}$ interaction \cite{Bar82} is used, with
385: parameter set values ($t_{0}=-2645, t_{1}=410, t_{2}=-135, t_{3}=
386: 15595, {\sigma}=1/6$), which give the realistic binding energy of
387: $^{40}$Ca obtained from the Hill-Wheeler equation (\ref{eq:HW}).
388: A discretization procedure is performed in order to solve this
389: integral equation using a set of regular mesh points with steps
390: as well as ranges of values of the generator coordinate which
391: lead to the convergence of the results, i.e. the latter do not
392: change after decreasing the step size or increasing the range of
393: the generator coordinate value. The range of variation of the
394: diffuseness parameter turns out to be within the interval
395: 0.49$\div $0.9072 fm. The ground-state energy and the excitation
396: energy of the first monopole $0^{+}$ state of $^{40}$Ca derived
397: from the procedure (340.07 MeV and 20.4 MeV) are close to the
398: corresponding experimental values (342.06 MeV and 20.0 MeV
399: \cite{Voi85}, respectively). Likewise, in \cite{Iva2000} the same
400: GCM scheme gives a value for the excitation energy of the $0^{+}$
401: breathing state in $^{16}$O which is in a good agreement with the
402: result obtained in \cite{Bri63}.
403:
404: The one-body density matrix (\ref{eq:obdm}) constructed within the
405: GCM approach has been applied to calculate overlap functions
406: related to the quasihole states in the $^{40}$Ca nucleus. In our
407: method the extracted overlap functions for a given orbital
408: momentum $l$ are the same for the different $j=l\pm 1/2$.
409: Hereafter the overlap functions for the neutron and proton $1d$
410: and $2s$ states will be of particular interest: transitions to
411: the $3/2^{+}$ ground state and the first $1/2^{+}$ excited state
412: of the corresponding residual nuclei observed in
413: $^{40}$Ca($e,e'p$), $^{40}$Ca($\gamma,p$) and $^{40}$Ca($p,d$)
414: reactions will be considered. The squared overlap functions
415: $|r\phi(r)|^2$ for the neutron and proton $1d$ and $2s$ quasihole
416: states are illustrated in Fig. \ref{fig:ovp}. They have the
417: correct asymptotic behavior related to the exponential decrease
418: of the ODM in its asymptotic region, being different for protons
419: and neutrons. The latter is in accordance with the experimental
420: proton and neutron separation energies. As known, the one-neutron
421: separation energy for $^{40}$Ca is almost twice larger than the
422: one-proton one and, therefore, the asymptotic tail of the GCM
423: density matrix is dominated by the contribution from the proton
424: overlap between the $^{40}$Ca and $^{39}$K ground-state wave
425: functions, where the role of the Coulomb interaction is important.
426:
427: The comparison of $|r\phi(r)|^2$ calculated with the GCM overlap
428: functions and with the shell-model Woods-Saxon wave functions for
429: the neutron bound states is given in Fig. \ref{fig:comp}. Small
430: differences between the two functions can be seen for both $1d$
431: and $2s$ states. The same is valid also for the $1s$ and $1p$
432: quasihole states, not presented in the figure. This result
433: justifies the approximation of using shell-model orbitals instead
434: of overlap functions in calculating the nucleon knockout cross
435: section for these nuclear states. Nevertheless, as it has been
436: pointed out in previous works \cite{Gai99,Gai2000}, the
437: calculated cross sections of one-nucleon removal reactions are
438: very sensitite to the differences in the corresponding overlap
439: functions.
440:
441: The values of the neutron and proton separation energies derived
442: from the procedure mentioned above are listed in Table 1. It can
443: be seen that the separation energies obtained within the GCM are
444: very close to the experimental values for the $1d$ state of
445: $^{40}$Ca and are anyhow better than the JCM result for
446: $\epsilon_{n}$ in the $2s$ state. In addition, we present the
447: separation energy values obtained within the mean-field
448: approximation (the Hartree-Fock method \cite{MS91}). We should
449: note that the differences between the latter and those from both
450: correlation methods are dependent on whether the nucleons are in
451: a valence orbital or not. Here we would like to note also that the
452: use of Woods-Saxon wave functions (instead of e.g.
453: harmonic-oscillator wave functions as in the JCM calculation),
454: which have the proper exponential-type decay, leads to a correct
455: asymptotic behaviour of the extracted overlap functions with
456: satisfactory separation energies of the quasihole states.
457:
458: The values of the spectroscopic factors deduced from the
459: numerical procedure turn out to be very close to unity. This is
460: not surprising, because tensor and short-range nucleon-nucleon
461: correlations are not included in the GCM. As known, these correlations
462: are responsible for the bulk part of the depletion of the occupied
463: states \cite{Gai99}. The values of the spectroscopic
464: factors derived from the Jastrow correlation method \cite{Sto96}, where
465: short-range correlations are included, are 0.892 for the $1d$ state and
466: 0.956 for the $2s$ state, respectively.
467:
468: The reduced cross sections for the $^{40}$Ca($e,e'p$) reaction as
469: a function of the missing momentum $p_{m}$, i.e. the magnitude of
470: the recoil momentum of the residual nucleus, for the transitions
471: to the $3/2^{+}$ ground state and the first $1/2^{+}$ excited
472: state [at excitation energy $E_{x}$=2.522 MeV] of $^{39}$K are
473: displayed in Figs. \ref{fig:eepgr} and \ref{fig:eepex},
474: respectively. Calculations have been done with the code DWEEPY
475: \cite{DWEEPY}, which is based on a nonrelativistic distorted wave
476: impulse approximation (DWIA) description of the nucleon knockout
477: process and includes final-state interactions and Coulomb
478: distortion of the electron waves \cite{Oxford}. In the figures
479: the results obtained with the overlap functions generated by the
480: GCM are compared with the data from \cite{Kra90} in parallel and
481: perpendicular kinematics. In parallel kinematics the momentum of
482: the outgoing proton is fixed and is taken parallel or
483: antiparallel to the momentum transfer. Different values of the
484: missing momentum are then obtained by varying the electron
485: scattering angle. In perpendicular kinematics the outgoing proton
486: energy and the momentum transfer are kept constant and different
487: values of the missing momentum are obtained by varying the angle
488: of the outgoing proton.
489:
490: A fair agreement with the shape of the experimental cross section
491: is achieved for both transitions and kinematics with the overlap
492: functions emerging from the ODM calculated within the GCM, which
493: includes realistic correlations corresponding to single-particle
494: and collective motion modes. The OF's already include the
495: spectroscopic factors. In order to reproduce the size of the
496: experimental cross section, an additional reduction factor has
497: been applied to the theoretical results. As in our previous
498: analysis of the $^{16}$O($e,e'p$) reaction \cite{Gai2000}, this
499: factor can be considered as a further spectroscopic factor,
500: reflecting the correlations not included in the OF which
501: correspondingly cause depletion of the quasihole states. The same
502: reduction factor has been applied to the cross sections
503: calculated with the OF for both parallel and perpendicular
504: kinematics, namely 0.55 for the $3/2^{+}$ ground state in Fig.
505: \ref{fig:eepgr} and 0.50 for the $1/2^{+}$ excited state in Fig.
506: \ref{fig:eepex}. Small variations around these values do not
507: change significantly the comparison with the data. However, for
508: the ground-state transition in perpendicular kinematics a
509: reduction factor about 20\% lower would give a better agreement
510: with data. A different spectroscopic factor for this transition
511: in the two kinematics is found also in the data analysis of
512: \cite{Kra90}, where in perpendicular kinematics the spectroscopic
513: factor is 25\% lower than the one determined in parallel
514: kinematics. This discrepancy is related in \cite{Kra90} to the
515: basic ingredients of the theory: the optical potential, the
516: bound-state wave function, the electron distortion and the
517: influence of meson-exchange currents (MEC). Our total
518: spectroscopic factors (the norm of OF's multiplied by the
519: additional reduction factor mentioned above) with the GCM, 0.54
520: for $3/2^{+}$ and 0.49 for $1/2^{+}$, are comparable with the
521: "experimental" ones, 0.65 and 0.52, determined under parallel
522: kinematical conditions in the analysis of \cite{Kra90}, where the
523: calculations are performed within the same DWIA framework and
524: with the same optical potential, but where phenomenological
525: single-particle wave functions are used with some parameters
526: adjusted to the data. We emphasize that in the present analysis
527: the overlap functions theoretically calculated on the basis of the
528: ODM of $^{40}$Ca within the GCM do not contain free parameters.
529: Our spectroscopic factors are also in a reasonable agreement with
530: those extracted from the fully relativistic theoretical analysis
531: of the $^{40}$Ca($e,e'p$) reaction in \cite{Udi93}, where,
532: however, only a part of the ($e,e'p$) data from \cite{Kra90} is
533: considered, while somewhat larger spectroscopic factors (around
534: 0.7--0.8) are obtained in the relativistic analyses of
535: \cite{Jin92,Jo96}.
536:
537: In Fig. \ref{fig:eepgr} for the
538: $^{40}$Ca($e,e'p$)$^{39}$K$_{g.s.}$ reaction we present also the
539: results obtained with the phenomenological Woods-Saxon wave
540: function. The WS wave function is also able to give a good
541: description of the data with a reduction factor of 0.6625,
542: somewhat larger than the one applied to the reduced cross section
543: computed with GCM overlap function. Also with the WS wave
544: function a reduction factor lower by about 20\% would give a
545: better description of the data in perpendicular kinematics.
546:
547: Fig. \ref{fig:gp60} shows the angular distribution of the
548: $^{40}$Ca($\gamma,p$)$^{39}$K$_{g.s.}$ reaction at
549: $E_{\gamma}$=60 MeV. In the figure the results given by the sum
550: of the one-body and of the two-body seagull currents are compared
551: with the contribution given by the one-body current, which roughly
552: corresponds to the DWIA treatment based on the direct knockout
553: (DKO) mechanism. Both contributions of DKO and MEC are
554: consistently evaluated in the theoretical framework of
555: ref.\cite{Benenti}. The results obtained with the OF from GCM for
556: the ground state transition and with the phenomenological WS wave
557: function are compared in the figure. In order to check the
558: consistency in the description of different one-proton removal
559: reactions, the calculated cross sections have been multiplied by
560: the same reduction factors obtained from the analysis of
561: ($e,e'p$) data, i.e. 0.55 with GCM and 0.6625 with WS. The
562: differences between the two curves are considerable and larger
563: than in the ($e,e'p$) reaction. This result might have been
564: expected, since it is well known that ($\gamma,p$) results are
565: extremely sensitive to the theoretical ingredients adopted in the
566: calculation, in particular for the bound states.
567:
568: The GCM calculation lies well below the data in DWIA, but a
569: reasonable agreement with the size and the shape of the
570: experimental cross section is obtained when MEC are added. Thus,
571: the important role played by MEC \cite{Gai2000,Benenti} is
572: confirmed here also for the $^{40}$Ca($\gamma,p$) reaction. The
573: WS calculation are larger that the ones with the GCM. This result
574: can be understood also from Fig. \ref{fig:eepgr}, where at the
575: high values of the missing momentum, that are explored in the
576: ($\gamma,p$) reaction, the ($e,e'p$) reduced cross sections
577: calculated with WS wave function are larger than the ones
578: obtained with the OF from the GCM. Thus, in DWIA the WS result is
579: closer to the data, in particular for the lowest angles, but when
580: MEC are added it overshoots the experimental cross section.
581:
582: Therefore, although both calculations with the GCM and WS wave
583: functions are able to give a good description of the
584: $^{40}$Ca($e,e'p$) data for the transition to the $3/2^{+}$
585: ground state of $^{39}$K, the ($\gamma,p$) results presented in
586: Fig. \ref{fig:gp60} for the same transition show that the GCM
587: overlap function leads to a better and more consistent description
588: of data for the $(e,e'p)$ and $(\gamma,p)$ reactions. This result
589: suggests proper accounting for the nucleon correlation effects in
590: the framework of the GCM.
591:
592: In Fig. \ref{fig:pd} the differential cross section of the
593: $^{40}$Ca($p,d$) reaction for the transition to the $3/2^{+}$
594: ground state in $^{39}$Ca, calculated at incident proton energy
595: $E_{p}$=65 MeV with the neutron GCM overlap function, is given and
596: compared with the experimental data \cite{Roos75}. It has been
597: obtained using the Distorted Wave Born Approximation (DWBA) with
598: zero-range approximation for the $p$-$n$ interaction inside the
599: deuteron. It is well known that the cross section of the ($p,d$)
600: reaction is more sensitive to the reaction mechanism adopted than
601: to the choice of the bound-state wave function. In this respect
602: quasifree nucleon knockout reactions are more suitable for our
603: investigation. Nevertheless, within the DWBA for such a pick-up
604: process as the ($p,d$) reaction, it can be seen that the shape of
605: the angular distribution is well reproduced by the GCM. Applying
606: the total spectroscopic factor of 0.54 for $3/2^{+}$ state to the
607: calculated cross section, a reasonable agreement with the size of
608: the experimental cross section is achieved. The role of the
609: additional depletion or reduction of the spectroscopic factors
610: due to correlations not included in the OF has been already seen
611: on the examples of the previous reactions considered. Although
612: some differences between the calculated and experimental angular
613: distributions exist, we would like to emphasize that, in
614: principle, overlap functions extracted from realistic one-body
615: density matrices have to be used as a tool for an accurate
616: description of one-nucleon removal reactions.
617:
618: \section{Conclusions}
619: In summary, we have calculated single-particle overlap functions
620: and the spectroscopic factors corresponding to low-lying
621: (quasihole) $(A-1)$-nucleon states on the base of the one-body
622: density matrix of $^{40}$Ca by considering its asymptotic (large
623: $r$) region and using the Generator Coordinate Method. These
624: calculations have been performed within an approach in which a
625: Woods-Saxon construction potential and a Skyrme-type effective
626: force are used. A consistent analysis of the cross sections of
627: one-nucleon removal reactions on $^{40}$Ca by means of the same
628: overlap functions is given. In contrast to standard DWIA
629: analyses, where phenomenological single-particle wave functions
630: were used with some parameters fitted to the data, in this paper
631: the results have been obtained with theoretically calculated
632: overlap functions which do not include free parameters. The
633: differences between the GCM overlap functions and the shell model
634: WS wave functions for both hole- and particle- states are analyzed
635: as a test for the role of correlations inherent to our approach.
636:
637: We have found that the calculated cross sections are generally in
638: good agreement with the experimental data. The quality of the
639: agreement, however, is sensitive to details of the overlap
640: functions. The important role of the additional reduction
641: spectroscopic factor applied to the present calculations is
642: pointed out. The fact that the reduction factor is consistent in
643: different nucleon removal reactions and also that this result is
644: obtained both for $^{40}$Ca in the present work and for $^{16}$O
645: in \cite{Gai2000} gives a more profound theoretical meaning to
646: this parameter. It can be interpreted as the spectroscopic factor
647: accounting for correlations not included in the Generator
648: Coordinate Method. The values of the total spectroscopic factors
649: obtained in the present analysis are in reasonable agreement with
650: those given by previous theoretical investigations.
651:
652: Apart from the short-range and tensor correlations studied in
653: previous papers \cite{Gai99,Gai2000}, in this work we looked into
654: the role of correlations caused by the collective nucleon motion.
655: Exploring the GCM, a consistent picture for all three $(e,e'p)$,
656: $(\gamma,p)$ and $(p,d)$ reactions on $^{40}$Ca was achieved on
657: the same theoretical ground. The results indicate that the
658: effects of NN correlations taken into account within our approach
659: and which are of long-range type are of significant importance
660: for the correct analysis of the processes considered.
661:
662: Finally, we would like to emphasize that the theoretical method to
663: calculate the cross sections of one-nucleon removal reactions by
664: means of single-particle overlap functions for the bound states
665: presented in this paper is, in principle, valid for all kind of
666: nuclei. The particular GCM scheme employed in our work is
667: applicable, however, only to nuclei with equal numbers of protons
668: and neutrons but to both closed and non-closed shell nuclei.
669: Calculations of similar reactions on open $s$-$d$ shell nuclei
670: are in progress.
671:
672: \acknowledgments We acknowledge the financial support given by
673: the Bulgarian National Science Foundation under Contracts
674: $\Phi$--809 and $\Phi$--905.
675:
676: \begin{references}
677:
678: \bibitem{Roos75} P.G. Roos, S.M. Smith, V.K.C. Cheng, G. Tibell, A.A.
679: Cowley, and R.AA. Riddle, Nucl. Phys. {\bf A255}, 187 (1975).
680:
681: \bibitem{Kra90}
682: J. Kramer, Ph.D. thesis, Universiteit van Amsterdam, 1990.
683:
684: \bibitem{Lap93} L. Lapik\'{a}s, Nucl. Phys. {\bf A553}, 297c
685: (1993).
686:
687: \bibitem{Abe92} C. Van den Abeele, D. Ryckbosch, J. Ryckbosch, J. Dias, L.
688: Van Hoorebeke, R. Van de Vyver, J.-O. Adler, B.-E. Andersson, L. Isaksson,
689: H. Ruijter, and B. Schr\"{o}der, Phys. Lett. B {\bf 296}, 302 (1992).
690:
691: \bibitem{Vn93} D. Van Neck, M. Waroquier, and K. Heyde, Phys. Lett. B {\bf
692: 314}, 255 (1993).
693:
694: \bibitem{Gai99} M.K. Gaidarov, K.A. Pavlova, S.S. Dimitrova, M.V. Stoitsov,
695: A.N. Antonov, D. Van Neck, and H. M\"{u}ther, Phys. Rev. C. {\bf 60},
696: 024312 (1999).
697:
698: \bibitem{Gai2000} M.K. Gaidarov, K.A. Pavlova, A.N. Antonov, M.V. Stoitsov,
699: S.S. Dimitrova, M.V. Ivanov, and C. Giusti, Phys. Rev. C {\bf 61}, 014306
700: (2000).
701:
702: \bibitem{Chr86} A.N. Antonov, Chr.V. Christov, and I.Zh. Petkov,
703: Nuovo Cim. {\bf 91A}, 119 (1986); A.N. Antonov, I.S. Bonev, Chr.V.
704: Christov, and I.Zh. Petkov, Nuovo Cim. {\bf 100A}, 779 (1988); A.N.
705: Antonov, I.S. Bonev, and I.Zh. Petkov, Bulg. J. Phys. {\bf 18}, 169 (1991);
706: A.N. Antonov, I.S. Bonev, Chr.V. Christov, and I.Zh. Petkov, Nuovo Cim.
707: {\bf 103A}, 1287 (1990).
708:
709: \bibitem{Ant}
710: A.N. Antonov, P.E. Hodgson, and I.Zh. Petkov, {\it Nucleon Correlations in
711: Nuclei} (Springer--Verlag, Berlin, 1993); {\it Nucleon Momentum and Density
712: Distributions in Nuclei} (Clarendon Press, Oxford, 1988).
713:
714: \bibitem{Iva2000} M.V. Ivanov, A.N. Antonov, and M.K. Gaidarov,
715: Int. J. Mod. Phys. E {\bf 9}, No.4, 339 (2000).
716:
717: \bibitem{Sa96} F. Arias de Saavedra, G. Co', A. Fabrocini, and S. Fantoni,
718: Nucl. Phys. {\bf A605}, 359 (1996); F. Arias de Saavedra, G. Co', and M.M.
719: Renis, Phys. Rev. C {\bf 55}, 673 (1997); A. Fabrocini, F. Arias de
720: Saavedra, and G. Co', Phys. Rev. C {\bf 61}, 044302 (2000).
721:
722: \bibitem{Di97} S.S. Dimitrova, M.K. Gaidarov, A.N. Antonov, M.V. Stoitsov,
723: P.E. Hodgson, V.K. Lukyanov, E.V. Zemlyanaya, and G.Z. Krumova, J. Phys. G {\bf
724: 23}, 1685 (1997).
725:
726: \bibitem{Sto96} M.V. Stoitsov, S.S. Dimitrova, and A.N. Antonov, Phys. Rev. C
727: {\bf 53}, 1254 (1996).
728:
729: \bibitem{Gri57} J.J. Griffin and J.A. Wheeler, Phys. Rev. {\bf 108}, 311
730: (1957).
731:
732: \bibitem{Hil53} D.L. Hill, J.A. Wheeler, Phys. Rev. {\bf 89}, 1102
733: (1953).
734:
735: \bibitem{Bar82} J. Bartel, P. Quentin, M. Brack, C. Guet, and
736: H.B. H{\aa}kansson, Nucl. Phys. {\bf A386}, 79 (1982).
737:
738: \bibitem{Mah91} C. Mahaux and R. Sartor, Adv. Nucl. Phys. {\bf 20}, 1 (1991).
739:
740: \bibitem{Bang85} J.M. Bang, F.A. Gareev, W.T. Pinkston, and J.S. Vaagen, Phys.
741: Rep. {\bf 125}, 253 (1985).
742:
743: \bibitem{Po96} A. Polls, H. M\"{u}ther, and W.H. Dickhoff, {\it Proceedings
744: of Conference on Perspectives in Nuclear Physics at Intermediate
745: Energies}, Trieste, 1995, edited by S. Boffi, C. Ciofi degli
746: Atti, and M.M. Giannini, (World Scientific, Singapore, 1996),
747: p.308.
748:
749: \bibitem{BH91} H.P. Blok and J.H. Heisenberg, in {\it Computational Nuclear
750: Physics 1, Nuclear Structure}, Edited by K. Langanke, J.A.
751: Maruhn, S.E. Koonin (Springer-Verlag, Berlin, 1991), p. 190.
752:
753: \bibitem{Voi85} N.A. Voinova-Eliseeva, I.A. Mitropolsky, LINP 1104, 1985
754: (Leningrad).
755:
756: \bibitem{Bri63} D.M. Brink and G.F. Nash, {\it Nucl. Phys.} {\bf 40}, 608
757: (1963).
758:
759: \bibitem{MS91} C. Mahaux and R. Sartor, Nucl. Phys. {\bf A528},
760: 253 (1991).
761:
762: \bibitem{DWEEPY} C. Giusti and F.D. Pacati, Nucl. Phys. {\bf A473}, 717 (1987);
763: Nucl. Phys. {\bf A485}, 461 (1988).
764:
765: \bibitem{Oxford} S. Boffi, C. Giusti, F.D. Pacati, and M. Radici,
766: {\it Electromagnetic Response of Atomic Nuclei, Oxford Studies in Nuclear
767: Physics} (Clarendon Press, Oxford, 1996).
768:
769: \bibitem{Udi93} J.M. Udias, P. Sarriguren, E. Moya de Guerrra, E.
770: Garrido, and J.A. Caballero, Phys. Rev. C {\bf 48}, 2731 (1993).
771:
772: \bibitem{Jin92} Y. Jin, D.S. Onley, and L.E. Wright, Phys. Rev. C {\bf
773: 45}, 1311 (1992).
774:
775: \bibitem{Jo96} J.L. Johansson, H.S. Sherif, and G.M. Lotz, Nucl. Phys.
776: {\bf A605}, 517 (1996).
777:
778: \bibitem{Benenti} G. Benenti, C. Giusti, and F.D. Pacati, Nucl. Phys.
779: {\bf A574}, 716 (1994).
780:
781: \bibitem{Schwandt}
782: P. Schwandt, H.O. Meyer, W.W. Jacobs, A.D. Bacher, S.E. Vigdor, M.D.
783: Kaitchuck, and T.R. Donoghue, Phys. Rev. C {\bf 26}, 55 (1982).
784:
785: \bibitem{Ful69} C.B. Fulmer et al., Phys. Rev. {\bf 181}, 1565
786: (1969).
787:
788: \bibitem{Becc69} F.D. Becchetti and G.W. Greenlees, Phys. Rev.
789: {\bf 182}, 1190 (1969).
790: \end{references}
791: \newpage
792: %%%%%%%%%%%%%%%%%%%%%%%%%%%%% table %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
793:
794: \noindent {\bf Table 1:} Neutron ($\epsilon_{n}$) and proton
795: ($\epsilon_{p}$) separation energies (in MeV) calculated on the
796: basis of the one-body density matrix for $^{40}$Ca within the
797: GCM. Comparison is made with the Hartree-Fock (HF) \cite{MS91}
798: single-particle energies, JCM results \cite{Sto96} and with the
799: experimental data (EXP) \cite{MS91}. The HF and EXP values are
800: given for the $d_{3/2}$ state.
801: \vspace{1cm}
802:
803: \begin{center}
804: \begin{tabular}{cccccccccc}
805: \hline\hline
806: & & \multicolumn{4}{c}{$\epsilon_{n}$} & & \multicolumn{3}{c}{$\epsilon_{p}$} \\
807: \cline{3-6} \cline{8-10}
808: nl & & GCM & HF & JCM & EXP & & GCM & HF & EXP \\
809: \hline
810: 1d & & 15.84 & 17.52 & 24.75 & 15.64 & & 8.69 & 10.33 & 8.33 \\
811: 2s & & 14.15 & 19.51 & 13.07 & 18.19 & & 7.39 & 12.42 & 10.94\\
812: \hline\hline
813: \end{tabular}
814: \end{center}
815:
816: \newpage
817: %%%%%%%%%%%%%%%%%%%%%%%%%%%%% figures %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
818: \begin{figure}
819: \caption[]{$|r\phi(r)|^2$ for the neutron and proton $1d$ (a) and
820: $2s$ (b) quasihole states in $^{40}$Ca obtained with GCM. The
821: normalization is: $\int \phi^{2}(r)r^{2}dr=S$. \label{fig:ovp} }
822: \end{figure}
823:
824: \begin{figure}
825: \caption[]{$|r\phi(r)|^2$ for the neutron $1d$ and $2s$ quasihole
826: states in $^{40}$Ca obtained with GCM overlap function (solid
827: line) and Woods-Saxon single-particle wave function
828: $|r\varphi(r)|^2$ (dashed line). All curves are normalized to
829: unity. \label{fig:comp} }
830: \end{figure}
831:
832: \begin{figure}
833: \caption[]{Reduced cross section of the $^{40}$Ca($e,e'p$)
834: reaction as a function of the missing momentum $p_{\mathrm m}$
835: for the transition to the $3/2^{+}$ ground state of $^{39}$K in
836: parallel (a) and perpendicular (b) kinematics. The incident
837: electron energy is 460 MeV in case (a), 483.2 MeV in case (b) and
838: the outgoing proton energy is 100 MeV. The optical potential is
839: taken from ref. \cite{Schwandt}. The overlap function is derived
840: from the ODM of GCM (solid line). The dashed line is calculated
841: with the WS wave function. The experimental data (full circles)
842: are taken from ref.\cite{Kra90}. The theoretical results have been
843: multiplied by the reduction factors given in the text.
844: \label{fig:eepgr} }
845: \end{figure}
846:
847: \begin{figure}
848: \caption[]{The same as in Fig. 3 for the transition to the first
849: $1/2^{+}$ excited state at 2.522 MeV of $^{39}$K. The theoretical
850: results have been multiplied by the reduction factor 0.50.
851: \label{fig:eepex} }
852: \end{figure}
853:
854: \begin{figure}
855: \caption[]{Angular distribution of the $^{40}$Ca($\gamma,p$)
856: reaction for the transition to the $3/2^{+}$ ground state of
857: $^{39}$K at $E_\gamma = 60$ MeV. The optical potential is taken
858: from ref. \cite{Schwandt}. The separate contributions given by
859: the one-body current (DWIA) and the final result given by the sum
860: of the one-body and the two-body seagull current (DWIA+MEC) are
861: shown. Line convention is as in Fig. \ref{fig:eepgr}. The
862: experimental data (triangles) are taken from ref.\cite{Abe92}.
863: The theoretical results have been multiplied by the reduction
864: factors given in the text, consistent with the analysis of
865: ($e,e'p$) data. \label{fig:gp60} }
866: \end{figure}
867:
868: \begin{figure}
869: \caption[]{Differential cross section for the $^{40}$Ca($p,d$)
870: reaction at incident proton energy $E_{p}=65$ MeV for the
871: transition to the $3/2^{+}$ ground state in $^{39}$Ca. The neutron
872: overlap function is derived from the ODM of GCM (solid line). The
873: proton and deuteron optical potentials are taken from refs.
874: \cite{Ful69} and \cite{Becc69}, respectively. The experimental
875: data \cite{Roos75} are given by the full circles. The theoretical
876: results have been multiplied by the reduction factor 0.55.
877: \label{fig:pd} }
878: \end{figure}
879: \end{document}
880: