nucl-th0410103/ee.tex
1: \documentclass{elsart}
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: %\usepackage{epsfig}
4: %\usepackage{psfig}
5: \usepackage{graphicx}
6: %\usepackage{graphics}
7: \usepackage{xspace}
8: \usepackage{amssymb}
9: \usepackage{amsmath}
10: \usepackage{latexsym}
11: %\usepackage{natbib}
12: \usepackage{mathrsfs}
13: %
14: \newcommand{\ove}{\overline}
15: \newcommand{\sla}{\mskip 1.mu /\mskip-9mu}
16: \newcommand{\diff}{{\rm\,d}}
17: \newcommand{\bint}{\mskip .5mu \int \mskip-18mu}
18: \newcommand{\ds}{\displaystyle}
19: %
20: \def\ll{\mbox{\boldmath $\lambda$}}
21: \def\l{\mbox{\boldmath $l$}}
22: \def\x{\mbox{\boldmath $x$}}
23: \def\z{\mbox{\boldmath $z$}}
24: \def\s{\mbox{\boldmath $s$}}
25: \def\h{\mbox{\boldmath $h$}}
26: \def\S{\mbox{\boldmath $S$}}
27: \def\N{\mbox{\boldmath $N$}}
28: \def\L{\mbox{\boldmath $L$}}
29: %\def\r{\mbox{{\bf r}}}
30: \def\R{\mbox{\boldmath $R$}}
31: \def\p{\mbox{\boldmath $p$}}
32: \def\P{\mbox{\boldmath $P$}}
33: \def\q{\mbox{\boldmath $q$}}
34: \def\v{\mbox{\boldmath $v$}}
35: \def\k{\mbox{\boldmath $k$}}
36: \def\T{\mbox{\boldmath $T$}}
37: \def\e{\mbox{\boldmath $e$}}
38: \def\N{\mbox{\boldmath $N$}}
39: \def\J{\mbox{\boldmath $J$}}
40: \def\A{\mbox{\boldmath $A$}}
41: \def\B{\mbox{\boldmath $B$}}
42: \def\ss{\mbox{\boldmath $\sigma$}}
43: \def\ta{\mbox{\boldmath $\tau$}}
44: \def\dd{\mbox{\boldmath $\nabla$}}
45: \def\ga{\mbox{\boldmath $\gamma$}}
46: \def\Ps{\mbox{\boldmath $\Psi$}}
47: \def\Fi{\mbox{\boldmath $\Phi$}}
48: \def\al{\mbox{\boldmath $\alpha$}}
49: \def\r{\mbox{\boldmath $r$}}
50: \def\rf{\mbox{\boldmath $r^{ \prime}$}}
51: \def\pf{\mbox{\boldmath $p^{ \prime}$}}
52: \def\sf{\mbox{\boldmath $s^{ \prime}$}}
53: %\def\pm{\mbox{\boldmath $p_{\mathrm m}$}}
54: \def\j{\mbox{$\jmath$}}
55: \def\ep{\mbox{$e^{\prime}$}}
56: \def\mcg{\mbox{$\mathcal{G}$}}
57: \def\mcv{\mbox{$\mathcal{V}$}}
58: \def\mcs{\mbox{$\mathcal{S}$}}
59: \def\mct{\mbox{$\mathcal{T}$}}
60: \def\mcm{\mbox{$\mathcal{M}$}}
61: \def\mch{\mbox{$\mathcal{H}$}}
62: \def\mcu{\mbox{$\mathcal{U}$}}
63: \def\mcd{\mbox{$\mathcal{D}$}}
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
66: \begin{document}
67: \begin{frontmatter}
68: \title{Antisymmetrized Green's function approach to $(e,e')$ reactions
69: with a realistic nuclear density}
70: 
71: \author[pv]{F.~Capuzzi},
72: \author[pv]{C.~Giusti},
73: \author[pv]{F.~D.~Pacati}
74: \address[pv]{Dipartimento di Fisica Nucleare e Teorica,
75: Universit\`{a} di Pavia and \\
76: Istituto Nazionale di Fisica Nucleare, Sezione di Pavia, I-27100
77: Pavia, Italy}
78: \author[so]{D.~N.~Kadrev}
79: \address[so]{Institute for Nuclear Research and Nuclear Energy,
80: Sofia 1784, Bulgaria}
81: 
82: \date{\today}
83: 
84: \begin{abstract}
85: A completely antisymmetrized Green's function approach to the
86: inclusive quasielastic $(e,e')$ scattering, including a realistic
87: one-body density, is presented. The single particle Green's function
88: is expanded in terms of the eigenfunctions of the nonhermitian
89: optical potential. This allows one to treat final state interactions
90: consistently in the inclusive and in the exclusive reactions.
91: Nuclear correlations are included in the one-body density. Numerical
92: results for the response functions of $^{16}$O and
93: $^{40}$Ca are presented and discussed.
94: \end{abstract}
95: 
96: \begin{keyword}
97: Electron scattering \sep Many-body theory
98: 
99: \PACS 25.30.Fj \sep 24.10.Cn
100: \end{keyword}
101: 
102: \end{frontmatter}
103: 
104: \section{Introduction \label{sec.intro}}
105: 
106: The one-body mechanism gives a natural interpretation of the
107: inclusive electron scattering in the quasielastic region. However,
108: in order to explain the experimental data of the separated
109: longitudinal and transverse responses more complicated mechanisms
110: are needed. A review of the experimental data and their possible
111: explanations can be found in Ref. \cite{book}. Thereafter, only a few
112: experimental papers were published \cite{batesca,csfrascati}, while
113: new experiments with high resolution are planned at JLab
114: \cite{jlabpro}.
115: 
116: Many papers were published in order to explain the problems raised by
117: the separation, i.e., the apparent lack of strength in the
118: longitudinal response and the apparent excess of strength in the
119: transverse one. Among them, the more recent ones are concerned with
120: the contribution to the inclusive cross section of meson exchange
121: currents and isobar excitations \cite{Sluys,Cenni,Amaro,Amaro2}, with
122: the effect of correlations \cite{Fabrocini,Co}, and the use of a
123: relativistic framework in the calculations \cite{Amaro,meucci}.
124: 
125: At present, however, a consistent and simultaneous description of the
126: longitudinal and transverse response functions is not available. A
127: possible solution could be the combined effect of two-body currents
128: and tensor correlations \cite{Leidemann,Fabrocini,Sick}.
129: 
130: A peculiar problem of inclusive electron scattering is the treatment
131: of final state interactions (FSI), since they are essential to
132: explain the exclusive reaction with one-nucleon emission, which is
133: the dominant process contributing to the inclusive reaction in the
134: quasielastic region. The large absorption due, e.g., to the
135: imaginary part of the optical potential, produces a loss of flux
136: that is appropriate for the exclusive process, but inconsistent for
137: the inclusive one, where the total flux must be conserved. This
138: conservation is preserved in the Green's function approach considered
139: here. This result was originally derived by arguments based on the
140: multiple scattering theory \cite{hori} and then by means of the
141: Feshbach projection operator formalism
142: \cite{chinn,bouch,capuzzi,capma,meucci}.
143: 
144: The spectral representation of the single-particle (s.p.) Green's
145: function, based on a biorthogonal expansion in terms of the
146: eigenfunctions of the nonhermitian optical potential, allows one to
147: perform explicit calculations and to treat FSI consistently in the
148: inclusive and in the exclusive reactions. In previous papers
149: \cite{capuzzi,meucci} the Green's function approach was used both in
150: a nonrelativistic and in a relativistic framework to perform explicit
151: calculations of the inclusive response functions.
152: 
153: Two issues are the main goal of this paper: a completely
154: antisymmetrized presentation of the Green's function approach, and
155: the effect of nuclear correlations, which are included in the model
156: through the use of a realistic one-body density matrix (ODM). In
157: the previous application of the method \cite{capuzzi,meucci}
158: correlations were neglected. The response functions were given by a
159: sum over the residual nucleus states restricted to be s.p. one-hole
160: states in the target. A pure shell model was assumed for the nuclear
161: structure and therefore the overlap functions between the target and
162: the residual nucleus were given by phenomenological s.p. wave
163: functions with a shell-model spectroscopic factor. In the present
164: work we are able to include partial occupation numbers, and
165: therefore correlations, through the natural expansion of the ODM.
166: 
167: The definitions and the main properties of the quantities involved in
168: the model are given in Section~\ref{sec.def}. In Sec.~\ref{sec.green}
169: the antisymmetrized Green's function approach is developed and the
170: inclusive cross section is expressed in terms of the ODM. In
171: Sec.~\ref{sec.spectral} the Green's function is calculated in terms of
172: the spectral representation related to the optical potential. 
173: In Sec.~\ref{sec.density} some models describing the ODM including 
174: correlations are briefly reviewed.
175: In Sec.~\ref{sec.results} the results of calculations for $^{16}$O and
176: $^{40}$Ca are reported and compared with some experimental data.
177: Summary and conclusions are drawn in Sec.~\ref{conc}.
178: 
179: 
180: \section{Definitions and main properties \label{sec.def}}
181: 
182: In this Section we collect most of the definitions which will be
183: useful in the treatment of the inclusive $(e,e')$ reaction, using the
184: Green's functions approach along the lines of
185: Refs.~\cite{capuzzi,capma,meucci}.
186: 
187: 
188: \subsection{One-body density, Green's functions and related
189: quantities \label{ssec:2.1}}
190: 
191: We deal in this paper with three different systems: the $Z$--proton
192: residual nucleus, the ($Z+1$)--proton target nucleus, and the
193: ($Z-1$)--proton system obtained removing a proton from the residual
194: nucleus. 
195: The contribution of the neutrons, here disregarded in order to simplify 
196: the formalism, can be introduced in an obvious way.
197: Let $T$ and $H$ be the kinetic energy and the Hamiltonian
198: operators acting in the Hilbert spaces
199: $\mathcal{H}^{Z},\mathcal{H}^{Z+1}$, and $\mathcal{H}^{Z-1}$,
200: corresponding to the three given systems and written in the Fock's
201: formalism. Here we are interested only in the following eigenvectors
202: of $H$: i) the eigenvectors $|n\rangle$, representing the bound
203: states of the residual nucleus with energy $\epsilon_n$, ii) the
204: eigenvector $|\psi_0\rangle$, representing the initial state, which
205: is the ground state of the target nucleus with energy $E_0$, and
206: iii) the eigenvectors $|\psi_f\rangle$, representing all the final
207: (bound or unbound) states of the target nucleus, including
208: $|\psi_0\rangle$. These eigenvectors are properly orthonormalized.
209: Relatedly, we define the hole overlaps, referred to $|\psi_0\rangle$,
210: \begin{equation} \label{eq:2.1}
211: \langle\r |\phi_n \rangle \equiv \langle n |a_{\mathbf {r}} |\psi_0
212: \rangle,
213: \end{equation}
214: and the particle overlaps, referred to $|n\rangle$,
215: \begin{equation}
216: \langle\r |\chi_{f,n} \rangle \equiv \langle n |a_{\mathbf {r}}
217: |\psi_f \rangle, \label{eq:2.2}
218: \end{equation}
219: where $a_{\mathbf {r}}$ annihilates a proton at the point $\r$. The
220: corresponding annihilation operator of a proton of momentum $\p$ is
221: \begin{equation}
222: a_{\mathbf{p}} \equiv (2\pi)^{-3/2} \int\diff\r \exp(-i\p \cdot \r) \
223: a_{\mathbf{r}}. \label{eq:2.3}
224: \end{equation}
225: For sake of simplicity, we consider spinless protons, we assume that
226: $|\psi_0 \rangle$ and $|n\rangle$ are not degenerate, and we omit any
227: degeneracy index in the unbound states $|\psi_f \rangle$. When
228: referring to quantities related to $n = 0$, the index 0 will be
229: usually suppressed in the following Sections.
230: 
231: \subsubsection{The one-body density matrix}
232: 
233: The one-body density matrix related to $|\psi_0 \rangle$ is defined
234: as
235: \begin{equation}
236: \langle\r|K|\rf\rangle = \langle\psi_0| a_{\mathbf {r'}}^{\dagger}
237: a_{\mathbf {r}}|\psi_0 \rangle . \label{eq:2.4}
238: \end{equation}
239: It is real and symmetric by the exchange $\r \longleftrightarrow
240: \rf$, since $|\psi_0 \rangle$ is not degenerate. The corresponding
241: operator $K$ in $L^2({\bf R}^3)$ is nonnegative, self-adjoint, and
242: has a finite trace equal to $Z$+1. The eigenvalue equation
243: \begin{equation}
244: K|u_\nu\rangle = n_\nu|u_\nu\rangle \label{eq:2.5}
245: \end{equation}
246: defines the natural orbitals $u_\nu(\r) = \langle\r|u_\nu\rangle$ and
247: the related occupation numbers $n_\nu$, which satisfy the relations
248: \begin{equation}
249: 0 \leq n_\nu \leq 1 \, ,\, \, \sum_\nu n_\nu = Z + 1. \label{eq:2.6}
250: \end{equation}
251: The operator $K$ is invertible in the subspace of $L^2({\bf R}^3)$
252: spanned by the orbitals $u_\nu$, with $n_\nu \neq 0$. If $n_\nu = 0$
253: is not an eigenvalue of $K$, it is necessarily an accumulation point
254: of eigenvalues. Thus $K$ is fully invertible, but $K^{-1}$ is a
255: rather complicate operator.
256: 
257: 
258: \subsubsection{The one-body Green's function}
259: 
260: At complex energies $z$, the one-body Green's function related to
261: $|n\rangle$ is the sum of the particle and hole Green's functions:
262: \begin{equation}
263: \langle \r |\mcg_n(z)|\rf\rangle = \langle \r
264: |\mcg_n^{(p)}(z)|\rf\rangle + \langle \r |\mcg_n^{(h)}(z)|\rf\rangle
265: \label{eq:2.7}
266: \end{equation}
267: with
268: \begin{equation}
269: \langle \r |\mcg_n^{(p)}(z)|\rf\rangle = \langle n |a_{\r} (z - H +
270: \epsilon_n)^{-1} a_{\rf}^{\dagger} |n\rangle , \label{eq:2.8}
271: \end{equation}
272: \begin{equation}
273: \langle \r |\mcg_n^{(h)}(z)|\rf\rangle = \langle n |a_{\rf}^{\dagger}
274: (z + H - \epsilon_n)^{-1} a_{\r} |n\rangle . \label{eq:2.9}
275: \end{equation}
276: These Green's functions are symmetric by the exchange $\r
277: \longleftrightarrow \rf$, since $|n\rangle$ is not degenerate.
278: 
279: At real energies $E$, we consider the retarded Green's function
280: instead of the time-ordered one, since it is more convenient in the
281: formalism developed below. It is defined as
282: \begin{equation}
283: \mcg_n(E) = \lim_{\eta \rightarrow +0} \mcg_n(E+i\eta).
284: \label{eq:2.10}
285: \end{equation}
286: In this equation, as well as in all the next equations involving
287: $\eta$, the limit is understood in the weak sense, i.e., it must be
288: performed after inclusion into a scalar product between normalizable
289: vectors. Henceforth, the symbol of limit will be usually understood.
290: 
291: 
292: \subsubsection{The spectral function}
293: 
294: The Green's function $\mcg_n(z)$ is analytic in the complex plane
295: except for cuts and poles on the real axis. Hence, it can be written
296: as
297: \begin{equation}
298: \mcg_n(z) = \int_{-\infty}^{+\infty} \diff E' \, \frac {\mcs_n(E')}
299: {z-E'}, \label{eq:2.11}
300: \end{equation}
301: where
302: \begin{equation}
303: \mcs_n(E) = \frac {1} {2\pi i} [\mcg_n(E-i\eta) - \mcg_n(E+i\eta)] =
304: \frac {1} {2\pi i} [\mcg_n^{\dagger}(E) - \mcg_n(E)] \label{eq:2.12}
305: \end{equation}
306: is the spectral function, which satisfies the sum rule
307: \begin{equation}
308: \int_{-\infty}^{+\infty} \diff E \, \mcs_n(E) = 1. \label{eq:2.13}
309: \end{equation}
310: Since $\mcs_n(E)$ is nonnegative, Eq.~(\ref{eq:2.11}) and
311: (\ref{eq:2.13}) show that
312: \begin{equation}
313: \mcg_n(z)|\phi \rangle = 0 \, \, \Longrightarrow \, |\phi\rangle = 0
314: \end{equation}
315: for $z$ complex. Hence, $\mcg_n(z)$ is fully invertible.
316: 
317: \subsubsection{The self-energy}
318: 
319: At complex energies $z$, the Green's function can be written as
320: \begin{equation}
321: \mcg_n(z) = \frac {1} {z - \mct -\mcm_n(z)} , \label{eq:2.15}
322: \end{equation}
323: where $\mct$ is the one-body kinetic energy and $\mcm_n$ the one-body
324: self-energy. The related Hamiltonian $h_n(z)$ is defined as
325: \begin{equation}
326: h_n(z) \equiv \mct + \mcm_n(z) = z - [\mcg_n(z)]^{-1}.
327: \label{eq:2.14}
328: \end{equation}
329: Note that $\mcg_n(z)$ is fully invertible, so that $\mcm_n(z)$ has no
330: mathematical drawbacks related to undue restrictions of its domain.
331: In contrast, $\mcg_n^{(p)}(z)$ is only partially invertible if 1 is
332: an eigenvalue of the density matrix associated with $|n\rangle$. This
333: produces a restriction of the domain of the related self-energy
334: $\mcm_n^{(p)}(z)$ which may lead, e.g., to an incorrect Dyson
335: equation. The same drawback affects $\mcm_n^{(h)}(z)$ if 0 is an
336: eigenvalue of the density matrix.
337: 
338: The analyticity of $\mcg_n(z)$ induces similar analyticity properties
339: into $h_n(z)$ and $\mcm_n(z)$. They are analytic in the complex plane
340: except for cuts and poles on the real axis, which are different from
341: those of $\mcg_n(z)$.
342: 
343: The one-body self-energy at real energies $\mcm_n(E)$ is defined as
344: \begin{equation}
345: \mcm_n(E) = \lim_{\eta \rightarrow +0} \mcm_n(E+i\eta)
346: \label{eq:2.16}
347: \end{equation}
348: In the Appendix B of \cite{capma3} it is proved that $\mcg_n(E)$ is
349: fully invertible, except for the values of $E$ corresponding to the
350: poles of $\mcm_n(E)$, and that holds the relation
351: \begin{equation}
352: \mcg_n(E) = \frac {1} {E - \mct - \mcm_n(E) + i \eta} .
353: \label{eq:2.17}
354: \end{equation}
355: The related Hamiltonian $h_n(E)$ is defined as
356: \begin{equation}
357: h_n(E) \equiv \mct + \mcm_n(E) = E - [\mcg_n(E)]^{-1} .
358: \label{eq:2.17a}
359: \end{equation}
360: 
361: \subsection{Self-energy and extended projection operators
362: \label{ssec:2.2}}
363: 
364: In his pioneering papers~\cite{fesh,fesh2} Feshbach introduced a
365: projection operator formalism to obtain a closed expression for the
366: theoretical optical-model potential. Here, we use an analogous
367: procedure for the self-energy, based on different projection
368: operators. More details can be found in Ref.~\cite{capma3}.
369: 
370: Let us introduce the vectors $\alpha_{\mathbf {r}}|n \rangle$, where
371: \begin{equation}
372: \alpha_{\mathbf {r}} \equiv a_{\mathbf {r}} + a_{\mathbf
373: {r}}^{\dagger} \, . \label{eq:2.18}
374: \end{equation}
375: They belong to the Hilbert space $\mch^{(Z+1)} \oplus \mch^{(Z-1)}$
376: and form an orthonormal set. In fact one has
377: \begin{equation}
378: \langle n|\alpha_{\mathbf {r}} \alpha_{\mathbf {r'}}|n \rangle =
379: \langle n|a_{\mathbf {r}} a_{\mathbf {r'}}^{\dagger} + a_{\mathbf
380: {r'}}^{\dagger} a_{\mathbf {r}} |n \rangle = \delta (\r - \rf),
381: \label{eq:2.19}
382: \end{equation}
383: where we have used the property that $\langle n| a_{\mathbf{r}
384: }^{\dagger} a_{\mathbf {r'}}|n \rangle$ is symmetric since $|n
385: \rangle$ is not degenerate. Therefore the operators
386: \begin{equation}
387: P_n = \int \diff \r \, \alpha_{\mathbf {r}}|n \rangle \langle n|
388: \alpha_{\mathbf {r}} = \int \diff \p \, \alpha_{\mathbf {p}} |n
389: \rangle \langle n| \alpha_{\mathbf {p}} \ , \, \, Q_n = 1 - P_n
390: \label{eq:2.20}
391: \end{equation}
392: are projection operators in $\mch^{(Z+1)} \oplus \mch^{(Z-1)}$. Such
393: operators, which will have an important role in Sec.~\ref{sec.green},
394: are called {\lq\lq extended projection operators\rq\rq} in order to
395: distinguish them from the Feshbach's ones. For sake of simplicity, we
396: define here $P_n$ in the simpler case of spinless particles. The
397: necessary changes to include spin and isospin variables can be found
398: in the Appendix A of Ref.~\cite{capma3}.
399: 
400: One can easily check that the correspondence
401: \begin{equation}
402: P_n |\phi \rangle \longleftrightarrow \langle n| \alpha_{\mathbf {r}}
403: |\phi \rangle , \label{eq:2.22}
404: \end{equation}
405: \begin{equation}
406: P_n O P_n \longleftrightarrow \langle n| \alpha_{\mathbf {r}} O
407: \alpha_{\mathbf {r'}}|n \rangle , \label{eq:2.23}
408: \end{equation}
409: where $|\phi \rangle$ and $O$ are vectors and operators in
410: $\mch^{(Z+1)} \oplus \mch^{(Z-1)}$, defines an isomorphism between
411: the Hilbert space of the vectors $P_n |\phi \rangle$
412: and $L^2({\bf R}^3)$. This means that every property
413: involving vectors and operators of one space implies that the same
414: property holds in the other space.
415: 
416: The above defined isomorphism is useful to develop a more synthetic
417: formalism for the quantities defined in Sec.~\ref{ssec:2.1}. To this
418: extent, we introduce the Hamiltonian-type operator $\hat H_n$, defined
419: by the relations
420: \begin{equation}
421: \hat H_n = H - \epsilon_n \quad \mathrm{in} \quad \mch^{(Z+1)} \quad
422: , \quad \hat H_n = \epsilon_n - H \quad \mathrm{in} \quad
423: \mch^{(Z-1)} , \label{eq:2.24}
424: \end{equation}
425: and extended to the whole space $\mch^{(Z+1)} \oplus \mch^{(Z-1)}$ by
426: linearity. Moreover, we introduce the related quantities
427: \begin{eqnarray}
428: \hat G_n(z) & = & \frac {1} {z - \hat H_n} \quad , \quad \hat G_n(E)
429: =
430: \hat G_n(E+i\eta) , \nonumber \\
431: \hat S_n(E) & = & \delta(E - \hat H_n) = \frac {1} {2\pi i} (\hat
432: G_n^{\dagger}(E) - \hat G_n(E)). \label{eq:2.25}
433: \end{eqnarray}
434: By the definition of $\delta(E - \hat H_n)$ one has
435: \begin{equation}
436: \int _{-\infty}^{+\infty} \diff E \, \hat S_n(E) = 1. \label{eq:2.26}
437: \end{equation}
438: 
439: Due to the orthogonality between vectors related to different numbers
440: of protons and due to the symmetry properties of the particle and
441: hole components of $\mcg_n(z)$, Eq.~(\ref{eq:2.7}) is reduced to the
442: simple expression
443: \begin{equation}
444: \langle \r |\mcg_n(z)|\rf\rangle = \langle n |\alpha_{\mathbf {r}}
445: \hat G_n(z) \alpha_{\mathbf {r'}}|n\rangle . \label{eq:2.27}
446: \end{equation}
447: Thus, $\mcg_n(z)$ is the one-body operator isomorphically
448: corresponding to the many-body operator $P_n\hat G_n(z)P_n$. 
449: Due to Eqs.~(\ref{eq:2.12}) and (\ref{eq:2.25}) it follows that $\mcs_n(E)$
450: corresponds to $P_n\delta(E-\hat H_n)P_n$. Therefore, one has
451: \begin{equation}
452: \langle \r |\mcs_n(E)|\rf\rangle = \langle n |\alpha_{\mathbf {r}}
453: \delta (E - \hat H_n) \alpha_{\mathbf {r'}}|n\rangle . \label{eq:2.28a}
454: \end{equation}
455: As the
456: handling of equations is often simpler when one deals with many-body
457: operators of this type, we introduce, in keeping with the definitions
458: of Sec.~\ref{ssec:2.1}, the many-body definitions of the Green's
459: function, the spectral function and the self-energy:
460: \begin{eqnarray}
461: G_n(z) & =& P_n\hat G_n(z)P_n \quad , \quad G_n(E) = P_n\hat
462: G_n(E)P_n \, , \\
463: S_n(E) & =& P_n\hat S_n(E)P_n = P_n \, \delta(E - \hat H_n)P_n =
464: \frac{1}{2\pi i} [G_n^{\dagger}(E) - G_n(E)], \\
465: h_n(z) & = & P_n\left(z - [\hat G_n(z)]^{-1}\right) P_n \, , \,
466: h_n(E) = P_n\left(E - [\hat G_n(E)]^{-1}\right) P_n \, , \\
467: M_n(z) & = & h_n(z) - P_n \hat T_n P_n \quad , \quad M_n(E) =
468: h_n(E) - P_n \hat T_n P_n \, , \label{eq:2.31}
469: \end{eqnarray}
470: where
471: \begin{equation}
472: \hat T_n \equiv \int \diff \k \, \alpha_{\mathbf {k}} |n \rangle
473: \frac {\k^2} {2m} \langle n |\alpha_{\mathbf {k}} , \label{eq:2.32}
474: \end{equation}
475: is the many-body operator which corresponds isomorphically to the
476: one-body kinetic energy $\mct$.
477: 
478: The closed expression of the many-body self-energy is deduced in
479: Secs.~3.1--3.3 of Ref.~\cite{capma3} and is:
480: \begin{equation}
481: M_n(z) = P_n[\hat H_n - \hat T_n + \hat H_n Q_n (z - Q_n \hat H_n
482: Q_n)^{-1} Q_n \hat H_n]P_n \, . \label{eq:2.33}
483: \end{equation}
484: For a two-body potential
485: \begin{equation}
486: V = \frac {1} {4} \int \diff\r \diff\s \diff\rf \diff\sf \,
487: a_{\mathbf {r}}^{\dagger} a_{\mathbf {s}}^{\dagger} {\mathcal V}
488: (\r,\s,\rf,\sf) a_{\mathbf {s'}} a_{\mathbf {r'}} , \label{eq:2.34}
489: \end{equation}
490: the corresponding one-body self-energy $\mcm_n(z)$ is the sum of the
491: static and dynamical parts
492: \begin{equation}
493: \langle \r| \mcm_n^{(S)}|\rf \rangle =\langle n|J_{\mathbf {r}}
494: a_{\mathbf {r'}}^{\dagger} + a_{\mathbf {r'}}^{\dagger} J_{\mathbf
495: {r}}|n\rangle , \label{eq:2.35}
496: \end{equation}
497: \begin{equation}
498: \langle \r| \mcm_n^{(D)}|\rf \rangle =\langle n|(J_{\mathbf {r}} +
499: J_{\mathbf {r}}^{\dagger})Q_n (z - Q_n \hat H_n Q_n)^{-1}
500: Q_n(J_{\mathbf {r'}} + J_{\mathbf {r'}}^{\dagger})|n\rangle ,
501: \label{eq:2.36}
502: \end{equation}
503: with
504: \begin{equation}
505: J_{\mathbf {r}} \equiv [a_{\mathbf {r}} , V] . \label{eq:2.37}
506: \end{equation}
507: 
508: The structure of $\mcm_n(z)$ is identical to that of the Feshbach's
509: potential, but its one-body expression is different. For instance,
510: the Feshbach's potential is not symmetric for the exchange $\r
511: \longleftrightarrow \rf$. In contrast, $\mcm_n(z)$ is symmetric, as
512: one can check working out the right-hand side of Eqs.~(\ref{eq:2.35})
513: and (\ref{eq:2.36}).
514: 
515: 
516: \section{Green's function approach to inclusive $(e,e')$ reactions
517: \label{sec.green}}
518: 
519: \subsection{Hadronic tensor \label{ssec:3.1} }
520: 
521: In order to avoid complications of minor interest in the present
522: context, we omit recoil corrections, use the one-photon exchange
523: approximation with one-body currents, and consider only spinless
524: point-like protons.
525: 
526: The inclusive cross section for the quasielastic $(e,e')$ scattering
527: on a nucleus of $Z+1$ protons is given by
528: \begin{equation}
529: \sigma = K (2\epsilon_{\mathrm L} R_{\mathrm L} + R_{\mathrm T}) ,
530: \label{eq:3.1}
531: \end{equation}
532: where $K$ is a kinematical factor and
533: \begin{equation}
534: \epsilon_{\mathrm L} = -\frac {q^2} {|\q|^2} \left( 1 - 2 \frac
535: {|\q|^2} {q^2} \tan^2 \frac {\theta} {2} \right)^{-1} \label{eq:3.2}
536: \end{equation}
537: measures the polarization of the virtual photon. In
538: Eq.~(\ref{eq:3.2}), $\theta$ is the scattering angle of the electron
539: and $q^2 = \omega^2 - |\q|^2$, where ($\omega,\q$) is the
540: four-momentum transfer. All nuclear structure information is contained
541: in the longitudinal and transverse response functions, defined by
542: \begin{eqnarray}
543:  R_{\mathrm L}(\omega,\q) & = & W^{00}(\omega,\q), \nonumber \\
544:  R_{\mathrm T}(\omega,\q) & = & W^{11}(\omega,\q) + W^{22}(\omega,\q)
545: \label{eq:3.3}
546: \end{eqnarray}
547: in terms of the diagonal components of the hadronic tensor
548: \begin{eqnarray}
549: W^{\mu\mu}(\omega,\q) & = & \langle \psi_0| J^{\mu\dagger}(\q)
550: \delta(\omega + E_0 - H) J^{\mu}(\q)|\psi_0\rangle \nonumber \\
551:  & = & \bint\sum_{\textrm {f}}\langle \psi_0| J^{\mu\dagger}(\q)
552:  |\psi_f\rangle\langle \psi_f|J^{\mu}(\q)|\psi_0\rangle
553:  \delta(\omega + E_0 - E_f).
554: \label{eq:3.4}
555: \end{eqnarray}
556: Here $J^{\mu}(\q)$ is the one-body nuclear charge-current operator,
557: $|\psi_0\rangle$ is the initial state of the ($Z+1$)-protons target
558: nucleus of energy $E_0$ and $|\psi_f\rangle$ is the corresponding
559: final state of energy $E_f=E_0+\omega$. The target ground state
560: $|\psi_0\rangle$ is assumed to be nondegenerate. The sum runs over
561: the target bound states and the scattering states corresponding to a
562: proton scattered from the residual nucleus in a bound state
563: $|n\rangle$ or in an unbound state $|\epsilon \rangle$. For sake of
564: simplicity, the degeneracy indices are suppressed. All the states are
565: properly antisymmetrized. Accordingly, the nuclear Hamiltonian $H$
566: and the current components $J^{\mu}(\q)$ are understood as second
567: quantization operators.
568: 
569: 
570: \subsection{Energy sum rules for the incoherent and the coherent
571: contributions \label{ssec:3.2}}
572: 
573: Acting on $|\psi_0\rangle$ the scalar component
574: \begin{equation}
575: J^0(\q) = \int \diff \r \, \exp (i\q\cdot\r) \, a_{\mathbf
576: {r}}^{\dagger} a_{\mathbf {r}} \label{eq:3.5}
577: \end{equation}
578: yields the wave function
579: \begin{equation}
580: \sum_{i=1}^{Z+1} J_i^0(\q,\r_i) \psi_0 (\r_1,...,\r_{Z+1}) \, , \,
581: J_i^0(\q,\r_i) = \exp (i\q\cdot\r_i). \label{eq:3.6}
582: \end{equation}
583: Similar expressions are obtained for the other components $J^\mu$
584: of the current operator. The hadronic tensor of Eq.~(\ref{eq:3.4}) is
585: the sum of the incoherent contribution
586: \begin{equation}
587: W_{\mathrm{inc}}^{\mu\mu}(\omega,\q) = (Z+1) \langle \psi_0|
588: J_1^{\mu\dagger}(\q)\delta(\omega + E_0 - H) J_1^{\mu}(\q)|\psi_0
589: \rangle \label{eq:3.7}
590: \end{equation}
591: and the coherent one
592: \begin{equation}
593: W_{\mathrm{coh}}^{\mu\mu}(\omega,\q) = (Z+1) \sum_{i=2}^{Z+1}\langle
594: \psi_0| J_1^{\mu\dagger}(\q)\delta(\omega + E_0 - H)
595: J_i^{\mu}(\q)|\psi_0 \rangle . \label{eq:3.8}
596: \end{equation}
597: 
598: The non-energy-weighted electromagnetic sum rule deals with the
599: integrated strength
600: \begin{equation}
601: \Sigma^{\mu\mu}(\q) = \int_{-0}^{\infty}\diff \omega \,
602: W^{\mu\mu}(\omega,\q) = \langle \psi_0
603: |J^{\mu\dagger}(\q)J^{\mu}(\q)|\psi_0 \rangle , \label{eq:3.9}
604: \end{equation}
605: where the lower integration limit means that the integral includes
606: the $\delta$--singularity at $\omega$=0 (full sum rule inclusive of
607: the elastic contribution). In the scalar case one has
608: \begin{equation}
609: \Sigma_{\mathrm{inc}}^{00}(\q) = Z + 1 , \label{eq:3.10}
610: \end{equation}
611: \begin{equation}
612: \Sigma_{\mathrm{coh}}^{00}(\q) = (Z + 1) \sum_{i=2}^{Z+1} \langle
613: \psi_0 |J_1^{0\dagger}(\q)J_i^{0}(\q)|\psi_0 \rangle .
614: \label{eq:3.11}
615: \end{equation}
616: 
617: In the Fermi gas model $\Sigma_{\mathrm{coh}}^{00}(\q)$ is negligible
618: for $|\q| > 2k_{\mathrm F}$, where $k_{\mathrm F}$ is the Fermi
619: momentum. This is still approximately true in presence of
620: correlations [Refs.~\cite{fabro,jourd}]. Therefore one has
621: \begin{equation}
622: \Sigma^{00}(\q) \simeq (Z + 1) \quad \mathrm{for} \quad |\q| > 2
623: k_{\mathrm F} . \label{eq:3.12}
624: \end{equation}
625: 
626: 
627: \subsection{Projection operator method \label{ssec:3.3}}
628: 
629: The approach developed in this Section differs form previous
630: treatments \cite{chinn,capuzzi,capma} for three reasons.
631: 
632: i) We want to include the effects of the final state interaction in
633: terms of the self-energy, rather than in terms of the Feshbach
634: optical potential or of the particle self-energy. In fact, the full
635: self-energy is a more fundamental quantity, has no mathematical
636: drawbacks, and is more closely related to the empirical optical-model
637: potentials. Therefore, we shall not use the Feshbach's projection
638: operators as in Ref.~\cite{capuzzi}, but the extended ones of
639: Eq.~(\ref{eq:2.20}), already used in Ref.~\cite{capma}.
640: 
641: ii) We want to include the interference between different channels
642: $|n\rangle$, neglected in Ref.~\cite{capma}, which is useful to
643: express the hadronic tensor in terms of the local potential
644: equivalent to the self-energy. This requires some changes in the
645: approach of Ref.~\cite{capma}, which give in practice the same result
646: when the interference effects are disregarded.
647: 
648: iii) We want to introduce different approximations for the elastic
649: and inelastic contributions to the hadronic tensor, when dealing with
650: the dependence on the state of the residual nucleus.
651: 
652: In order to keep the treatment as simple as possible, we shall only
653: refer to the scalar component $W^{00}(\omega,\q)$, with $J^0(\q)$
654: written in the momentum representation, i.e.
655: \begin{equation}
656: J^{0}(\q) = \int \diff\p \, a_{\mathbf {p}}^{\dagger} a_{\mathbf
657: {p-q}} . \label{eq:3.13}
658: \end{equation}
659: Disregarding the upper indices of the hadronic tensor and inserting
660: the completeness relation of the residual-nucleus states, Eq.~(\ref{eq:3.4})
661: yields
662: \begin{equation}
663: W(\omega,\q) = \bint\sum_n \int\diff \p \, \langle \psi_0| a_{\mathbf
664: {p-q}}^{\dagger} |n\rangle\langle n| a_{\mathbf {p}} \ \delta(\omega
665: + E_0 - H) J^{0}(\q)|\psi_0\rangle , \label{eq:3.14}
666: \end{equation}
667: where the sum is understood over $|n\rangle$ and $|\epsilon\rangle$.
668: Equation (\ref{eq:3.14}) was obtained in Ref.~\cite{capma} in a more
669: complicated way. Here, the same result is recovered with the
670: insertion of a completeness relation into the second quantization
671: expression of $J^{0}(\q)$ and has the same generality, i.e. can be
672: applied to every one-body charge-current operator.
673: 
674: As was done in previous papers, we shall now disregard the
675: contribution of the continuum states of the residual nucleus. Although
676: this approximation is correct at the energy and momentum transfers
677: considered here, we shall recover the continuum contribution in
678: Sec.~\ref{sssec:3.4.3}. Equation (\ref{eq:3.14}) is approximated with
679: \begin{equation}
680: W(\omega,\q) = \sum_n \mathrm{Re} \int\diff \p \, \langle \psi_0|
681: a_{\mathbf {p-q}}^{\dagger} |n\rangle\langle n| a_{\mathbf {p}} \
682: \delta(\omega + E_0 - H) J^{0}(\q)|\psi_0\rangle , \label{eq:3.15}
683: \end{equation}
684: where the real part has been extracted in order to restore the real
685: nature of $W(\omega,\q)$, which can be lost after the approximation.
686: This is equivalent to introduce the truncated completeness
687: symmetrically in $J^{0\dagger}(\q)$ and in $J^{0}(\q)$.
688: 
689: Now, the treatment of Ref.~\cite{capma} is slightly modified in order
690: to obtain expressions directly related to the self-energy, according
691: to the above items i) and ii). This requires the introduction of the
692: vectors $(a_{\mathbf {p}}^{\dagger} + a_{\mathbf {p}})|n\rangle$,
693: instead of $a_{\mathbf {p}}^{\dagger}|n\rangle$, and the use of
694: Eq.~(\ref{eq:2.24}) in order to feature $\hat H_n$. To this extent, we
695: add to the second factor of Eq.~(\ref{eq:3.15}), written as
696: \begin{equation}
697: \langle n| a_{\mathbf {p}} \ \delta(\omega + E_0 - \epsilon_n - \hat
698: H_n) J^{0}(\q)|\psi_0\rangle , \label{eq:3.15ab}
699: \end{equation}
700: the term
701: \begin{equation}
702: \langle n|a_{\mathbf {p}}^{\dagger}\, \delta(\omega + E_0 -\epsilon_n
703: - \hat H_n)J^{0}(\q)|\psi_0\rangle = \langle n|a_{\mathbf
704: {p}}^{\dagger}\, \delta(\omega + E_0 - 2\epsilon_n +
705: H)J^{0}(\q)|\psi_0\rangle, \label{eq:3.19c}
706: \end{equation}
707: which is null, since it is a scalar product between states with a
708: different number of protons. Moreover, we add to the terms
709: (\ref{eq:3.15ab}) and (\ref{eq:3.19c}) the corresponding ones with
710: $\omega$ replaced by $-\omega$, i.e.
711: \begin{equation}
712: \langle n|a_{\mathbf {p}}\, \delta(-\omega + E_0 - \epsilon_n - \hat
713: H_n) J^{0}(\q)|\psi_0\rangle = \langle n|a_{\mathbf {p}}\,
714: \delta(-\omega + E_0 - H) J^{0}(\q)|\psi_0\rangle, \label{eq:3.19b}
715: \end{equation}
716: which is null for $\omega > 0$ because the $\delta$-function is not
717: fed with the eigenfunctions of $H$, and
718: \begin{eqnarray}
719: \langle n|a_{\mathbf {p}}^{\dagger}\, \delta(-\omega + E_0 -
720: \epsilon_n &-& \hat H_n)J^{0}(\q)  |\psi_0\rangle = \nonumber\\
721: \langle n| a_{\mathbf{p}}^{\dagger}\, \delta(&-&\omega + E_0 -
722: 2\epsilon_n + H)J^{0}(\q)|\psi_0\rangle, \label{eq:3.19a}
723: \end{eqnarray}
724: which is null for the same reason as the term (\ref{eq:3.19c}). The
725: terms (\ref{eq:3.19b}) and (\ref{eq:3.19a}) are
726: necessary to obtain a null sum rule for the interference
727: contribution, as will be apparent below. We emphasize that the term 
728: (\ref{eq:3.19b}) can be added only for $\omega \neq 0$.
729: 
730: After addition of the null terms (\ref{eq:3.19c})--(\ref{eq:3.19a}),
731: Eq.~(\ref{eq:3.15}) reads
732: \begin{eqnarray}
733: W(\omega,\q) = A(\omega,\q) + A(-\omega,\q) \label{eq:3.16}
734: \end{eqnarray}
735: with
736: \begin{eqnarray}
737:  A(\omega,\q) &=& \sum_n \mathrm{Re} \int\diff \p \, \langle \psi_0|
738: a_{\mathbf {p-q}}^{\dagger} |n\rangle \nonumber \\
739: & \times & \langle n| (a_{\mathbf {p}} +
740: a_{\mathbf {p}}^{\dagger})\ \delta(\omega + E_0 - \epsilon_n - \hat
741: H_n) J^{0}(\q)|\psi_0\rangle . \label{eq:3.17}
742: \end{eqnarray}
743: In Eq.~(\ref{eq:3.16}), as well as in the following analogous
744: expressions, it is understood that the $\delta$--singularity at
745: $\omega= 0$ must be retained only in $A(\omega,\q)$ to avoid a double
746: counting. Therefore, one has
747: \begin{equation}
748: \int_{-0}^{\infty} \diff\omega \, W(\omega,\q) =
749: \int_{-\infty}^{\infty} \diff\omega \, A(\omega,\q). \label{eq:3.18}
750: \end{equation}
751: 
752: Now we use the extended projection operators, defined in
753: Eq.~(\ref{eq:2.20}),
754: \begin{equation}
755: P_n =\int \diff\p \, \alpha_{\mathbf {p}}|n\rangle\langle
756: n|\alpha_{\mathbf {p}} \quad \mathrm{and}\quad Q_n = 1 - P_n
757: \label{eq:3.20}
758: \end{equation}
759: to separate in Eq.~(\ref{eq:3.17}) a direct ($A^{\mathrm D}$) and an
760: interference ($A^{\mathrm I}$) term, such that
761: \begin{equation}
762: A(\omega,\q) = A^{\mathrm D}(\omega,\q) + A^{\mathrm I}(\omega,\q)
763: \label{eq:3.21}
764: \end{equation}
765: with
766: \begin{eqnarray}
767: A^{\mathrm D}(\omega,\q) &=& \sum_n \mathrm{Re} \int\diff \p \,
768: \langle \psi_0| a_{\mathbf {p-q}}^{\dagger} |n\rangle  \nonumber \\
769: & \times & \langle n|(\alpha_{\mathbf {p}} P_n \, \delta(\omega + E_0
770: -\epsilon_n -\hat H_n) P_n J^{0}(\q)|\psi_0\rangle , \label{eq:3.22a}
771: \end{eqnarray}
772: \begin{eqnarray}
773: A^{\mathrm I}(\omega,\q) &=& \sum_n \mathrm{Re} \int\diff \p \,
774: \langle \psi_0| a_{\mathbf{p-q}}^{\dagger} |n\rangle  \nonumber \\
775: & \times & \langle n|(\alpha_{\mathbf {p}} P_n \, \delta(\omega + E_0
776: -\epsilon_n -\hat H_n) Q_n J^{0}(\q)|\psi_0\rangle . \label{eq:3.22b}
777: \end{eqnarray}
778: Accordingly, the hadronic tensor $W(\omega,\q)$ is decomposed into a
779: direct and an interference contribution:
780: \begin{equation}
781: W^{\mathrm D}(\omega,\q) = A^{\mathrm D}(\omega,\q) + A^{\mathrm
782: D}(-\omega,\q) \label{eq:3.23a}
783: \end{equation}
784: \begin{equation}
785: W^{\mathrm I}(\omega,\q) = A^{\mathrm I}(\omega,\q) + A^{\mathrm
786: I}(-\omega,\q) \label{eq:3.23b}
787: \end{equation}
788: We remark that $W^{\mathrm I}$ does not contribute to the sum rule,
789: due to the relation $P_nQ_n=0$ and to the relation
790: \begin{equation}
791: \int_{-\infty}^{\infty} \diff\omega \, \delta(\omega + E_0 -
792: \epsilon_n - \hat H_n) = \int_{-\infty}^{\infty} \diff E \, \delta(E
793: - \hat H_n) = 1 , \label{eq:3.24}
794: \end{equation}
795: which yields
796: \begin{equation}
797: \int_{-0}^{\infty} \diff\omega \, W^{\mathrm I}(\omega,\q) =
798: \int_{-\infty}^{\infty} \diff \omega \, A ^{\mathrm I}(\omega,\q) = 0
799: . \label{eq:3.25}
800: \end{equation}
801: 
802: Note that the terms (\ref{eq:3.19b}) and (\ref{eq:3.19a}), which yield
803: the contribution of the negative energies to the second integral, are
804: essential to fulfill Eq.~(\ref{eq:3.25}).
805: 
806: We emphasize that the insertion of the terms
807: (\ref{eq:3.19c})--(\ref{eq:3.19a}) does not change the expression of
808: $W(\omega,\q)$, but produces an effect on its decomposition into
809: $W^{\mathrm D}(\omega,\q)$ and $W^{\mathrm I}(\omega,\q)$. As a
810: consequence, the present decomposition is slightly different form
811: that of Ref.~\cite{capma}. The term~(\ref{eq:3.19c}), absent in
812: Ref.~\cite{capma}, modifies the decomposition at very low energies,
813: i.e. at $\omega \leq 2\epsilon_n - E_0 - \epsilon_0^{A-1}$, where
814: $\epsilon_0^{A-1}$ is the ground state energy of the ($Z-1$)--body
815: Hamiltonian. The term~(\ref{eq:3.19b}) produces no effects. The term
816: (\ref{eq:3.19a}), which is also present in Ref.~\cite{capma}, differs
817: here by a shift of $2(E_o - \epsilon_n)$ in the value of the
818: argument. This shift is the only difference from the treatment of
819: Ref.~\cite{capma} at the energies $\omega$ considered in this paper.
820: The changes introduced here influence only the separation between the
821: direct and the interference contributions, allowing a more
822: appropriate treatment of the latter, which in Ref.~\cite{capma} is
823: totally neglected. The effects of these changes on $W^{\mathrm
824: D}(\omega,\q)$ are in practice negligible at high momentum transfers, 
825: as it will be seen in Sec.~\ref{sssec:3.4.5}.
826: 
827: 
828: \subsection{Direct contribution to the hadronic tensor
829: \label{ssec:3.4}}
830: 
831: \subsubsection{One-body expression of the incoherent and coherent
832: contributions \label{sssec:3.4.1}}
833: 
834: Using (\ref{eq:3.20}), Eq.~(\ref{eq:3.22a}) is expressed in terms of
835: vectors and operators of a s.p. Hilbert space as
836: \begin{equation}
837: A^{\mathrm D}(\omega,\q) = \sum_n \mathrm{Re} \langle \phi_n|
838: j^{0\dagger}(\q) \mcs_n(\omega_n)|\Phi_n(\q)\rangle \, , \, \omega_n
839: = \omega + E_0 - \epsilon_n, \label{eq:3.26}
840: \end{equation}
841: where $|\phi_n\rangle$ and $\mcs_n$ represent the hole overlaps and
842: the one-body spectral functions, defined by Eqs.~(\ref{eq:2.1}) and
843: (\ref{eq:2.28a}), respectively, and
844: \begin{equation}
845: \langle \p| j^{0}(\q)|\phi_n\rangle = \langle \p - \q|\phi_n\rangle =
846: \langle n|a_ {\mathbf {p-q}}|\psi_0\rangle , \label{eq:3.27}
847: \end{equation}
848: \begin{equation}
849:  \langle \p| \Phi_n(\q)\rangle = \langle n|a_ {\mathbf {p}}
850:  J^{0}(\q)|\psi_0\rangle .
851: \label{eq:3.28}
852: \end{equation}
853: 
854: Using the relation
855: \begin{equation}
856: [a_{\mathbf {p}},J^{0}(\q)] = a_ {\mathbf {p-q}} \label{eq:3.29}
857: \end{equation}
858: in Eq.~(\ref{eq:3.26}), the contribution to $|\Phi_n(\q)\rangle$ due
859: to the many-body current $J^{0}(\q)$ is split into the contribution of
860: a single proton and that of all the other residual $Z$ protons, as
861: \begin{equation}
862: \langle \p| \Phi_n(\q)\rangle = \langle \p | j^{0}(\q)|\phi_n\rangle
863: + \langle \p |\delta \Phi_n(\q)\rangle , \label{eq:3.30}
864: \end{equation}
865: where
866: \begin{eqnarray}
867: & & \langle \p |\delta \Phi_n(\q)\rangle \equiv \langle n
868: |J^{0}(\q)
869: a_ {\mathbf {p}}|\psi_0\rangle \nonumber \\
870:  = \sqrt {Z+1} & \int & \diff \p_2 ... \diff \p_{Z+1}
871: \langle n|\p_2 ... \p_{Z+1}\rangle\langle\p,\p_2 ... \p_{Z+1}|
872: \sum_{i=2}^{Z+1} J_i^{0}(\q)|\psi_0 \rangle . \label{eq:3.31}
873: \end{eqnarray}
874: From Eq.~(\ref{eq:3.30}), we express $W^{\mathrm D}(\omega,\q)$ as the
875: sum of the terms
876: \begin{eqnarray}
877: W_{\mathrm{inc}}^{\mathrm D}(\omega,\q) & = &
878: A_{\mathrm{inc}}^{\mathrm D}(\omega,\q) + A_{\mathrm{inc}}^{\mathrm
879: D} (-\omega,\q), \nonumber \\
880: A_{\mathrm{inc}}^{\mathrm D}(\omega,\q) & = & \sum_n \mathrm{Re}
881: \langle \phi_n |j^{0\dagger}(\q) \mcs_n (\omega_n) j^{0}(\q)|
882: \phi_n\rangle \label{eq:3.32}
883: \end{eqnarray}
884: and
885: \begin{eqnarray}
886: W_{\mathrm{coh}}^{\mathrm D}(\omega,\q) & = &
887: A_{\mathrm{coh}}^{\mathrm D}(\omega,\q) + A_{\mathrm{coh}}^{\mathrm
888: D} , (-\omega,\q), \nonumber \\
889: A_{\mathrm{coh}}^{\mathrm D}(\omega,\q) & = & \sum_n \mathrm{Re}
890: \langle \phi_n |j^{0\dagger}(\q)
891: \mcs_n(\omega_n)|\delta\Phi_n(\q)\rangle . \label{eq:3.33}
892: \end{eqnarray}
893: The decomposition (\ref{eq:3.30}) is exactly equivalent to that of
894: Eqs.~(\ref{eq:3.7}) and (\ref{eq:3.8}). In keeping with this
895: correspondence, we call $W_{\mathrm{inc}}^{\mathrm D}$ and
896: $W_{\mathrm{coh}}^{\mathrm D}$ the {\lq\lq incoherent\rq\rq} and
897: {\lq\lq coherent\rq\rq} contributions to $W^{\mathrm D}$.
898: 
899: 
900: \subsubsection{Dependence on the state of the residual nucleus}
901: 
902: No information is available concerning the spectral functions
903: $\mcs_n(E)$ related to the excited states of the residual nucleus.
904: Therefore, we are obliged to express $\mcs_n(E)$ in terms of
905: $\mcs_0(E)$. According to previous papers
906: \cite{fesh3,moniz,moniz2}, we assume: i) $\mcs_n(E)$ only
907: differs from $\mcs_0(E)$ by an energy shift $\delta\epsilon_n$, ii)
908: the contributions of the various shifts can be taken into account by
909: means of a single average shift $\tilde \epsilon$. Thus, in
910: Eqs.~(\ref{eq:3.26}), (\ref{eq:3.32}), and (\ref{eq:3.33}) we set
911: \begin{equation}
912: \mcs_n(\omega + E_0 - \epsilon_n) \simeq \mcs_0(\tilde\omega) \, ,
913: \, \tilde\omega = \omega + E_0 - \epsilon_0 - \tilde\epsilon .
914: \label{eq:3.34}
915: \end{equation}
916: The choice of $\tilde\epsilon$ will be discussed later.
917: 
918: 
919: \subsubsection{Inclusion of the unbound states of the residual
920: nucleus \label{sssec:3.4.3} }
921: 
922: For a single state $|\epsilon\rangle$ of the continuous spectrum, one
923: can define neither an extended projection operator nor a Feshbach's
924: one. This is the reason why the states $|\epsilon\rangle$ are usually
925: neglected in the treatments based on the Green's function approach.
926: Even if the continuum contribution is negligible at the energy and
927: momentum transfers considered here, the lack of completeness
928: resulting if one considers only the bound states $|n\rangle$ appears
929: unsatisfactory from a conceptual point of view, mainly because the
930: sum rule is not exactly fulfilled. This drawback can be easily
931: eliminated if Eq.~(\ref{eq:3.34}) is used. In fact, from
932: Eq.~(\ref{eq:3.26}), by expliciting the scalar products and using
933: Eqs.~(\ref{eq:3.27}) and (\ref{eq:3.28}), we have
934: \begin{eqnarray}
935: A^{\mathrm D}(\omega,\q) & = & \sum_n \mathrm{Re} \int \diff\p
936: \diff\p' \langle \psi_0 |a_{\mathbf {p-q}}^{\dagger}|n\rangle \langle
937: n|a_{\mathbf {p'}} J^{0}(\q) |\psi_0\rangle \langle \p|
938: \mcs_0(\tilde\omega)|\p'\rangle , \nonumber \\ \tilde\omega & = &
939: \omega + E_0 -\epsilon_0 - \tilde\epsilon . \label{eq:3.35}
940: \end{eqnarray}
941: 
942: Here, it is quite natural to recover the contribution of the
943: continuous spectrum after making the substitution
944: \begin{equation}
945: \sum_n |n \rangle \langle n| \longrightarrow \sum_n |n \rangle
946: \langle n| + \int \diff\epsilon \, |\epsilon\rangle \langle\epsilon|
947: = 1 \label{eq:3.36}
948: \end{equation}
949: which yields
950: \begin{eqnarray}
951: A^{\mathrm D}(\omega,\q) & = & \mathrm{Re} \int \diff\p \diff\p'
952: \langle \psi_0 |a_{\mathbf {p-q}}^{\dagger} a_{\mathbf {p'}}
953: J^{0}(\q) |\psi_0\rangle \langle \p| \mcs_0(\tilde\omega)|\p'\rangle
954: \nonumber \\ & = & \mathrm{Re}\,\mathrm{Tr} [\, {\overline
955: K}_{\mathbf {q}} \mcs_0(\tilde\omega)], \label{eq:3.37}
956: \end{eqnarray}
957: where
958: \begin{eqnarray}
959: \langle \p'|\,{\overline K}_{\mathbf {q}} |\p\rangle \equiv \langle
960: \psi_0 |a_{\mathbf {p-q}}^{\dagger} a_{\mathbf {p'}} J^{0}(\q)
961: |\psi_0\rangle
962: %\nonumber \\
963:  = \int \diff \k \,\langle \psi_0 |a_{\mathbf {p-q}}^{\dagger}
964: a_{\mathbf {p'}}a_{\mathbf {k}}^{\dagger} a_{\mathbf
965: {k-q}}|\psi_0\rangle . \label{eq:3.38}
966: \end{eqnarray}
967: Analogously, using Eq.~(\ref{eq:3.29}), one has
968: \begin{equation}
969: A_{\mathrm {inc}}^{\mathrm D}(\omega,\q) = \mathrm{Re}\,\mathrm{Tr}
970: [K_{\mathbf {q}} \mcs_0(\tilde\omega)] \label{eq:3.39}
971: \end{equation}
972: with
973: \begin{equation}
974: \langle \p'| K_{\mathbf {q}} |\p\rangle = \langle \psi_0 | a_{\mathbf
975: {p-q}}^{\dagger} a_{\mathbf {p'-q}}|\psi_0\rangle \label{eq:3.40}
976: \end{equation}
977: and
978: \begin{equation}
979: A_{\mathrm {coh}}^{\mathrm D}(\omega,\q) = \mathrm{Re}\,\mathrm{Tr}
980: [\Delta K_{\mathbf {q}} \mcs_0(\tilde\omega)] \label{eq:3.41}
981: \end{equation}
982: with
983: \begin{equation}
984: \langle \p'| \Delta K_{\mathbf {q}} |\p\rangle = \int \diff \k \,
985: \langle \psi_0 | a_{\mathbf {p-q}}^{\dagger} a_{\mathbf
986: {k}}^{\dagger}a_{\mathbf {k-q}} a_{\mathbf {p'}}|\psi_0\rangle .
987: \label{eq:3.42}
988: \end{equation}
989: 
990: Since $K_{\mathbf {q}}$ and $\mcs_0$ are nonnegative (hermitian)
991: operators and $K_{\mathbf {q}}$ is self-adjoint, $\mathrm{Tr}
992: [K_{\mathbf {q}} \mcs_0(\tilde\omega)]$ is nonnegative, as it follows
993: easily by expressing the trace on the basis of the eigenvectors of
994: $K_{\mathbf {q}}$. Thus one has
995: \begin{equation}
996: A_{\mathrm {inc}}^{\mathrm D}(\omega,\q) = {\mathrm {Tr}} 
997: \, [K_{\mathbf {q}} \mcs_0( \tilde \omega )] \ge 0 \, . \label{eq:3.45}
998: \end{equation}
999: 
1000: The expression of
1001: \begin{equation}
1002: W^{\mathrm D}(\omega,\q) = A^{\mathrm D}(\omega,\q) + A^{\mathrm
1003: D}(-\omega,\q) ,
1004: \end{equation}
1005: obtained from Eq.~(\ref{eq:3.37}), and the corresponding expressions
1006: of $W_{\mathrm{inc}}^{\mathrm D}(\omega,\q)$ and
1007: $W_{\mathrm{coh}}^{\mathrm D}(\omega,\q)$, are essentially the same
1008: as those obtained in Secs.~9.1.1 and 9.1.3 of Ref.~\cite{capma}. This
1009: can be easily checked noting that the spectral function $\mcs_0(E)$ is
1010: equal to the particle (hole) spectral function at positive (negative)
1011: energies. The only differences from the results of Ref.~\cite{capma}
1012: are the insertion of the average energy $\tilde \epsilon$ and an
1013: energy shift in $A_{\mathrm {inc}}^{\mathrm D}(-\omega,\q)$ and in
1014: $A_{\mathrm {coh}}^{\mathrm D}(-\omega,\q)$.
1015: 
1016: The problem of defining a projection operator related to a single
1017: continuous eigenvalue $\epsilon$ of the residual nucleus is of
1018: strictly mathematical nature. It is solved in the Appendix
1019: associating with $\epsilon$ a set of approximate eigenvectors
1020: depending on an index $\eta$ which describes the accuracy of the
1021: approximation (increasing for $\eta \rightarrow +0$). They can
1022: replace the exact eigenvectors $|\epsilon \rangle$ in the whole
1023: formalism for two reasons. i) In the limit for $\eta \rightarrow +0$
1024: they satisfy a completeness relation analogous to Eq.~(\ref{eq:3.36}).
1025: ii) One can associate with $\epsilon$ a projection operator since the
1026: approximate eigenvectors are normalizable.
1027: 
1028: 
1029: 
1030: \subsubsection{Non-energy-weighted sum rules}
1031: 
1032: We shall prove here that the approximation of Eq.~(\ref{eq:3.34}),
1033: relating the spectral functions $\mcs_n(E)$ to $\mcs_0(E)$, has no
1034: effect on the sum rules, both for the incoherent and the coherent
1035: contributions to the hadronic tensor. Since in Eq.~(\ref{eq:3.25}) we
1036: proved that the interference term $W^{\mathrm I}(\omega,\q)$ does not
1037: contribute to the sum rule, Eqs.~(\ref{eq:3.37}), (\ref{eq:3.39}), and
1038: (\ref{eq:3.41}) should satisfy the same sum rules as the exact
1039: hadronic tensor.
1040: 
1041: Remembering Eq.~(\ref{eq:3.10}) and using Eq.~(\ref{eq:2.13}), i.e.
1042: \begin{equation}
1043: \int_{-\infty}^{+\infty} \diff E \, \mcs_0(E) = 1 \, ,
1044: \label{eq:3.46}
1045: \end{equation}
1046: Eq.~(\ref{eq:3.37}) yields
1047: \begin{eqnarray}
1048: \Sigma^{\mathrm D}(\q) & \equiv & \int_{-0}^{\infty} \diff \omega \,
1049: W^{\mathrm D}(\omega,\q) = \int_{-\infty}^{\infty} \diff \omega \,
1050: A^{\mathrm D}(\omega,\q) \nonumber \\
1051: & = & {\mathrm {Re}}\, {\mathrm {Tr}} \, \left[\, {\overline
1052: K}_{\mathbf {q}} \int_{-\infty}^{\infty} \diff \tilde \omega \,
1053: \mcs_0(\tilde \omega) \right] = {\mathrm {Re}}\,{\mathrm {Tr}} \,
1054: [\,{\overline K}_{\mathbf {q}}]. \label{eq:3.47}
1055: \end{eqnarray}
1056: By Eqs.~(\ref{eq:3.38}) and (\ref{eq:3.13}), one has
1057: \begin{equation}
1058: {\mathrm {Tr}} \, [\,{\overline K}_{\mathbf {q}}] = \int\diff \p
1059: \diff \k \, \langle \psi_0| a_{\mathbf {p-q}}^{\dagger}a_{\mathbf
1060: {p}}a_{\mathbf {k}}^{\dagger} a_{\mathbf {k-q}}|\psi_0\rangle =
1061: \langle \psi_0|J^{0\dagger}(\q) J^{0}(\q)|\psi_0\rangle
1062: \label{eq:3.48}
1063: \end{equation}
1064: and so
1065: \begin{equation}
1066: \Sigma^{\mathrm D}(\q) =\langle \psi_0|J^{0\dagger}(\q)
1067: J^{0}(\q)|\psi_0\rangle , \label{eq:3.49}
1068: \end{equation}
1069: according to the exact sum rule of Eq.~(\ref{eq:3.9}). Likewise, using
1070: Eqs.~(\ref{eq:3.39}) and (\ref{eq:3.40}), one has
1071: \begin{equation}
1072: \Sigma_{\mathrm {inc}}^{\mathrm D}(\q) \equiv \int_{-0}^{\infty}
1073: \diff \omega \, W_{\mathrm {inc}}^{\mathrm D}(\omega,\q) = {\mathrm
1074: {Tr}} \, [K_{\mathbf {q}}] = {\mathrm {Tr}} \, [K] = Z + 1,
1075: \label{eq:3.50}
1076: \end{equation}
1077: according to the exact sum rule of Eq.~(\ref{eq:3.10}). It follows
1078: that also the sum rule for the coherent part is in accordance with
1079: Eq.~(\ref{eq:3.11}).
1080: 
1081: 
1082: \subsubsection{Practical approximations for the direct contribution to 
1083: the hadronic tensor \label{sssec:3.4.5} }
1084: 
1085: We have previously expressed the direct contribution $W^{\mathrm
1086: D}(\q)$ to the hadronic tensor as the sum of the incoherent and
1087: coherent parts
1088: \begin{equation}
1089: W_{\mathrm {inc}}^{\mathrm D}(\omega,\q) = A_{\mathrm {inc}}^{\mathrm
1090: D}(\omega,\q) + A_{\mathrm {inc}}^{\mathrm D}(-\omega,\q) ,
1091: \label{eq:3.51}
1092: \end{equation}
1093: \begin{equation}
1094: W_{\mathrm {coh}}^{\mathrm D}(\omega,\q) = A_{\mathrm {coh}}^{\mathrm
1095: D}(\omega,\q) + A_{\mathrm {coh}}^{\mathrm D}(-\omega,\q) .
1096: \label{eq:3.52}
1097: \end{equation}
1098: In this Subsection, we profit from some results of Ref.~\cite{capma}
1099: to simplify $W^{\mathrm D}(\omega,\q)$ in view of practical
1100: calculations. We notice that the terms of Eqs.~(\ref{eq:3.51}) and
1101: (\ref{eq:3.52}) at positive (negative) arguments correspond to the
1102: particle (hole) contributions of the quoted reference.
1103: 
1104: In the calculations based on the Green's function approach, the
1105: coherent part of the hadronic tensor is usually disregarded since it
1106: is too difficult to evaluate. Consequently, only the sum rule
1107: $\Sigma_{\mathrm {inc}}^{\mathrm D}(\q) = Z + 1$ of (\ref{eq:3.50}) is
1108: relevant. This is considered correct at high momentum transfers,
1109: schematically for $q > k_{\mathrm F}$. In Sec.~10 of
1110: Ref.~\cite{capma}, it was shown that this approximation is excessive
1111: even for uncorrelated systems, since only $A_{\mathrm {coh}}^{\mathrm
1112: D}(\omega,\q)$ is negligible at $q > k_{\mathrm F}$, whereas
1113: $A_{\mathrm {coh}}^{\mathrm D}(-\omega,\q)$ is still sizable for
1114: $k_{\mathrm F} < q < 2k_{\mathrm F}$. Fortunately, the latter term is
1115: largely cancelled by $A_{\mathrm {inc}}^{\mathrm D}(-\omega,\q)$ and
1116: this produces the simplification
1117: \begin{equation}
1118: W^{\mathrm D}(\omega,\q) = A_{\mathrm {inc}}^{\mathrm D}(\omega,\q)
1119: \quad \mathrm{at} \quad q > k_{\mathrm F}. \label{eq:3.53}
1120: \end{equation}
1121: 
1122: The energy shifts that are introduced here in the arguments of $A_{\mathrm
1123: {inc}}^{\mathrm D}(-\omega,\q)$ and $A_{\mathrm {coh}}^{\mathrm
1124: D}(-\omega,\q)$ do not influence the cancellation found in
1125: Ref.~\cite{capma}.
1126: 
1127: Due to this cancellation, the sum rule for $W^{\mathrm D}(\omega,\q)$
1128: is deprived of the contribution of $A_{\mathrm {inc}}^{\mathrm
1129: D}(-\omega,\q)$, which is positive as shown in Eq.~(\ref{eq:3.45}).
1130: Thus, the sum rule for the total hadronic tensor, which coincides
1131: with that of $W^{\mathrm D}(\omega,\q)$ since the interference term
1132: does not contribute, reduces to
1133: \begin{equation}
1134: \Sigma (\q) < Z+1 \quad \mathrm{for} \quad k_{\mathrm F} < q <
1135: 2k_{\mathrm F}. \label{eq:3.54}
1136: \end{equation}
1137: 
1138: 
1139: \subsection{Inclusion of the interference term \label{ssec:3.5}}
1140: 
1141: The interference term $W^{\mathrm I}(\omega,\q)$, defined in
1142: Eqs.~(\ref{eq:3.22b}) and (\ref{eq:3.23b}), gives no contribution in
1143: absence of final state correlations, since in this case $\delta(E -
1144: \hat H_n)$ does not connect the projection operators $P_n$ and $Q_n$.
1145: In the Green's function approach $W^{\mathrm I}(\omega,\q)$ is
1146: usually disregarded also in presence of final-state correlations,
1147: since it does not seem reducible to a one-body expression. As already
1148: noticed, this approximation does not affect the sum rule. We give
1149: here an approximated expression of $W^{\mathrm I}(\omega,\q)$, which
1150: is reducible to a one-body expression and does not modify the sum
1151: rule. This result has
1152: a conceptual relevance and will be useful to introduce the empirical
1153: optical-model potential.
1154: 
1155: Here, we are only interested in the interference term $A^{\mathrm
1156: I}(\omega,\q)$ to be added to Eq.~(\ref{eq:3.53}). Therefore, we start
1157: from the expression (\ref{eq:3.23b}) of $W^{\mathrm I}(\omega,\q)$ and
1158: disregard the contribution $A^{\mathrm I}(-\omega,\q)$. Then, we
1159: introduce an approximation which allows the reduction of $A^{\mathrm
1160: I}(\omega,\q)$ to a one-body expression and, finally, we retain only
1161: its incoherent part.
1162: 
1163: We insert into the expression (\ref{eq:3.22b}) of $A^{\mathrm
1164: I}(\omega,\q)$ the Eq.~(\ref{eq:2.25}), i.e.,
1165: \begin{equation}
1166: \delta(E - \hat H_n) = \frac {1} {2 \pi i} [\hat G_n^{\dagger}(E) -
1167: \hat G_n(E)] , \label{eq:3.55}
1168: \end{equation}
1169: with
1170: \begin{equation}
1171:  \hat G_n(E) \equiv \frac {1} {E - \hat H_n + i \eta} .
1172: \label{eq:3.56}
1173: \end{equation}
1174: Then, we use the approximated relations
1175: \begin{equation}
1176:  P_n \hat G_n(E) Q_n J^0(\q)|\psi_0\rangle = - P_n \hat G_n(E)P_n
1177: M'_n(E) J^0(\q)|\psi_0\rangle,
1178: \label{eq:3.57}
1179: \end{equation}
1180: \begin{equation}
1181:  P_n \hat G_n^{\dagger}(E) Q_n J^0(\q)|\psi_0\rangle = - P_n \hat
1182:  G_n^{\dagger}(E)P_n (M'_n)^{\dagger}(E) J^0(\q)|\psi_0\rangle,
1183: \label{eq:3.58}
1184: \end{equation}
1185: where $M'_n(E)$ is the energy derivative of the many-body self-energy
1186: of Eq.~(\ref{eq:2.31}), obtaining
1187: \begin{eqnarray}
1188: A^{\mathrm I}(\omega,\q) & =& - \frac {1} {\pi} \sum_n \mathrm{Re}
1189: \int \diff \p \, \langle \psi_0 |a_{\mathbf {p-q}}^{\dagger}|n
1190: \rangle \nonumber \\
1191: & \times & \langle n| \alpha_{\mathbf {p}}\frac {[\hat G_n^{\dagger}
1192: (M'_n)^{\dagger}] (\omega_n) - (\hat G_n M'_n)(\omega_n)} {2i} J^0
1193: (\q)| \psi_0\rangle \label{eq:3.59}
1194: \end{eqnarray}
1195: with $\omega_n = \omega + E_0 - \epsilon_n$.
1196: 
1197: Equations (\ref{eq:3.57}) and (\ref{eq:3.58}) have the same structure
1198: as Eq.~(59) of Ref.~\cite{Cap} and can be deduced with the same method
1199: replacing the Hamiltonian $H$ by $\hat H_n$ and the Feshbach's
1200: projection operators by the extended ones. As remarked in the quoted
1201: reference, the Eqs.~(\ref{eq:3.57}) and (\ref{eq:3.58}) must be used
1202: inside the matrix elements of Eq.~(\ref{eq:3.59}) and hold in the
1203: region of the quasielastic peak at intermediate and high energies.
1204: 
1205: Operating as in Sec.~\ref{sssec:3.4.1} and retaining only the
1206: incoherent part of Eq.~(\ref{eq:3.59}), we obtain 
1207: \begin{eqnarray}
1208:  A_{\mathrm{inc}}^{\mathrm{I}}(\omega,\q) = - \frac {1} {\pi}
1209:  \sum_n & \mathrm{Re} &
1210: \langle \phi_n |j^{0\dagger}(\q) \frac {[\mcg_n^{\dagger}
1211: (\mcm'_n)^{\dagger}]
1212: (\omega_n) - (\mcg_n \mcm'_n)(\omega_n)} {2i} \nonumber \\
1213: & \times & j^0 (\q)|\phi_n\rangle, \label{eq:3.60}
1214: \end{eqnarray}
1215: where $\mcm_n(E)$ is the one-body expression of the self-energy.
1216: Using the relation
1217: \begin{equation}
1218:  \mcg_n(E)= \int_{-\infty}^{+\infty} \diff E' \, \frac {\mcs_n(E')}
1219:  {E- E' + i\eta} ,
1220: \label{eq:3.61}
1221: \end{equation}
1222: deduced from Eq.~(\ref{eq:2.11}), and the relation [see Eq.~(\ref{eq:2.17a})]
1223: \begin{equation}
1224:  \mcm_n(E) = E- [\mcg_n(E)]^{-1} - \mct ,
1225: \label{eq:3.62}
1226: \end{equation}
1227: Eq.~(\ref{eq:3.34}) is extended to the relations
1228: \begin{eqnarray}
1229: &\mcg_n(\omega+E_0-\epsilon_n) \simeq \mcg_0(\tilde \omega) ,
1230: \,\,\,\,\,\, \mcm_n(\omega+E_0-\epsilon_n) \simeq \mcm_0(\tilde
1231: \omega)\, , \nonumber \\
1232: & \tilde \omega = \omega + E_0 - \epsilon_0 - \tilde \epsilon .
1233: \label{eq:3.63}
1234: \end{eqnarray}
1235: Therefore, the contribution of the continuous spectrum can be included
1236: as in Sec.~\ref{sssec:3.4.3} to yield
1237: \begin{eqnarray}
1238:  A_{\mathrm {inc}}^{\mathrm I}(\omega,\q) = - \frac {1}{\pi}
1239: \mathrm{Re} \, \mathrm{Tr} \,\left[ K_{\mathbf {q}} \left( \frac
1240: {[\mcg_0^{\dagger} (\mcm'_0)^{\dagger}] (\tilde \omega) - (\mcg_0
1241: \mcm'_0)(\tilde \omega)} {2i} \right) \right] . \label{eq:3.64}
1242: \end{eqnarray}
1243: Making explicit the real part of Eq.~(\ref{eq:3.64}), we have
1244: \begin{eqnarray}
1245: A_\mathrm{inc}^\mathrm{I}(\omega,\q) &=& - \frac{1}{2\pi}
1246: \mathrm{Tr} \, \left[  K_{\mathbf{q}}  \left(  \frac
1247: {[\mcg_0^{\dagger} (\mcm'_0)^{\dagger} +
1248: (\mcm'_0)^{\dagger}\mcg_0^{\dagger}] (\tilde \omega)}
1249: {2i} \right. \right. \nonumber \\
1250: & - & \left. \left. \frac {(\mcg_0 \mcm'_0 + \mcm'_0 \mcg_0)(\tilde
1251: \omega)} {2i} \right) \right] . \label{eq:3.65}
1252: \end{eqnarray}
1253: where $\mcm'_0(E)$ and $\mcg_0(E)$ are symmetrically arranged.
1254: 
1255: Equation (\ref{eq:3.64}) implies
1256: \begin{equation}
1257: \int_{-\infty}^{+\infty} \diff \omega \, A_{\mathrm {inc}}^{\mathrm
1258: I} (\omega,\q) = 0 , \label{eq:3.66}
1259: \end{equation}
1260: in keeping with Eq.~(\ref{eq:3.25}). In fact, denoting by $C_{\infty}$
1261: the large circle with center in the origin, one has
1262: \begin{equation}
1263: \int_{-\infty}^{+\infty} \diff E\, [\mcg_0^{\dagger} (E)
1264: (\mcm'_0)^{\dagger}(E) - \mcg_0(E) \mcm'_0(E)] = \int_{C_{\infty}}
1265:  \diff z \, \mcg_0(z)\mcm'_0(z) = 0 ,
1266: \label{eq:3.67}
1267: \end{equation}
1268: where the first equality is due to the analyticity of
1269: $\mcg_0(z)\mcm'_0(z)$ in the complex plane, except for cuts and poles
1270: on the real axis, and the latter one follows from the fast decrease
1271: of $\mcg_0(z)\mcm'_0(z)$ at infinity (more than $1/|z|$).
1272: 
1273: 
1274: 
1275: \subsection{Practical approximation for the total hadronic tensor
1276: \label{ssec:3.6}}
1277: 
1278: In this Subsection we consider the total hadronic tensor as obtained
1279: by the sum of the direct contribution of Eq.~(\ref{eq:3.53}) with the
1280: interference term of Eq.~(\ref{eq:3.65}):
1281: \begin{equation}
1282: W(\omega,\q) = A_{\mathrm {inc}}^{\mathrm D}(\omega,\q) + A_{\mathrm
1283: {inc}}^{\mathrm I}(\omega,\q) . \label{eq:3.68}
1284: \end{equation}
1285: Using in Eq.~(\ref{eq:3.45}), the relation (see Eq.~(\ref{eq:2.12}))
1286: \begin{equation}
1287: \mcs_0(E) = \frac {1} {2\pi i} [\mcg_0^{\dagger}(E) - \mcg_0(E)] ,
1288: \label{eq:3.69}
1289: \end{equation}
1290: one has
1291: \begin{equation}
1292: W(\omega,\q) = \frac {1} {2\pi i} \mathrm{Tr} \, \left[ K_{\mathbf q}
1293: \left( \mcg_{\mathrm {eff}}^{\dagger}(\tilde \omega) - \mcg_{\mathrm
1294: {eff}}(\tilde \omega)\right) \right] \, , \, \tilde \omega = \omega +
1295: E_0 - \epsilon_0 - \tilde \epsilon , \label{eq:3.70}
1296: \end{equation}
1297: where
1298: \begin{equation}
1299: \mcg_{\mathrm {eff}}(E) = [1 - \mcm'_0(E)]^{1/2} \, \mcg_0(E) [1 -
1300: \mcm'_0(E)]^{1/2} . \label{eq:3.71}
1301: \end{equation}
1302: We call $\mcg_{\mathrm {eff}}(E)$ \lq\lq effective Greeen's
1303: function\rq\rq.
1304: 
1305: Strictly, Eq.~(\ref{eq:3.71}) should involve in a symmetrical way the
1306: factor $1 - \mcm'_0(E)/2$, instead of the square root operator. The
1307: replacement by the latter has been done to feature the more
1308: fundamental quantity $1 - \mcm'_0(E)$, related to the effective
1309: $\omega$--mass (see Ref.~\cite{mabor}), and is justified by the fact
1310: that $\mcm'_0(E)$ is small in the energy region of interest. For
1311: instance, in the scattering $p-^{40}$Ca the approximation is very
1312: good at proton energies beyond 50 MeV (see Fig. 2 of
1313: Ref.~\cite{tornow}, where the result has been obtained using the
1314: dispersion relation specific of the self-energy). Concerning the
1315: precise meaning of the square root operator of Eq.~(\ref{eq:3.71}), we
1316: remark that it is trivial if $\mcm'_0(E)$ is local (as supposed in
1317: many models) and that in any case the square root can be defined
1318: without problems as a power series for $\|\mcm'_0(E)\| \leq 1$.
1319: 
1320: The effective Green's function shows an interesting property. While
1321: the energy derivative of $\mcg_0(E)$ satisfies the relation
1322: \begin{equation}
1323: \mcg'_0(E) = - \mcg_0(E) [1 - \mcm'_0(E)] \mcg_0(E), \label{eq:3.72}
1324: \end{equation}
1325: the derivative of $\mcg_{\mathrm {eff}}(E)$, performed with the help
1326: of the previous equation and disregarding the contribution of
1327: $\mcm''_0(E)$, as allowed by the quasilinear behavior of $\mcm_0(E)$
1328: (see again Fig. 2 of Ref.~\cite{tornow}), yields
1329: \begin{equation}
1330: \mcg'_{\mathrm {eff}}(E) \simeq [1 - \mcm'_0(E)]^{1/2}\, \mcg'_0(E)
1331: [1 - \mcm'_0(E)]^{1/2}= - \mcg_{\mathrm {eff}}^2(E) .
1332: \label{eq:3.73}
1333: \end{equation}
1334: This is the typical relation satisfied by the Green's function of an
1335: energy independent Hamiltonian. This property really holds. In fact
1336: $\mcg_{\mathrm {eff}}(E)$ is invertible, since $[1 -
1337: \mcm'_0(E)]^{-1/2}$ can be defined as a power series and, therefore,
1338: one can define the related self-energy $\mcm_{\mathrm {eff}}(E)$ and
1339: the Hamiltonian $h_{\mathrm {eff}}(E)$ as in Eq.~(\ref{eq:2.17a}):
1340: \begin{equation}
1341: \mcm_{\mathrm {eff}}(E) = h_{\mathrm {eff}}(E) - \mct \, , \,
1342:  h_{\mathrm {eff}}(E) =E - \mcg_{\mathrm {eff}}^{-1}(E) .
1343: \label{eq:3.74}
1344: \end{equation}
1345: Note that the energy derivative
1346: \begin{equation}
1347: \mcm'_{\mathrm {eff}}(E) = 1 + \mcg_{\mathrm {eff}}^{-1}(E)
1348: \mcg'_{\mathrm {eff}}(E) \mcg_{\mathrm {eff}}^{-1}(E) \label{eq:3.75}
1349: \end{equation}
1350: is approximately equal to zero due to Eq.~(\ref{eq:3.73}). This means
1351: that the hadronic tensor can be reduced to a one-body expression by
1352: means of the Green's function associated to an effective self-energy,
1353: nearly energy independent.
1354: 
1355: Equation (\ref{eq:3.70}) holds under the same conditions as
1356: Eqs.~(\ref{eq:3.53}) and (\ref{eq:3.65}), i.e. at intermediate and
1357: high energies, near the quasielastic peak and for $q > k_{\mathrm
1358: F}$. Now, we arrange Eq.~(\ref{eq:3.70}) in order to feature the role
1359: of the scalar operator $j^0(\q;\r) = \exp(i \q \cdot \r)$. This is
1360: included in $K_{\mathbf q}$ as
1361: \begin{equation}
1362: \langle \r|K_{\mathbf q} |\rf\rangle = \exp(-i \q \cdot \r) K(\r,\rf)
1363: \exp(i \q \cdot \rf) = \langle \r|j^{0\dagger}(\q)Kj^0(\q)|\rf\rangle
1364: . \label{eq:3.76}
1365: \end{equation}
1366: Permuting the factors under the trace symbol and using the property
1367: \begin{equation}
1368: \mathrm{Tr}\,[O^{\dagger}] = {\overline{\mathrm{Tr}\,[O]}} ,
1369: \end{equation}
1370: Eq.~(\ref{eq:3.70}) can equivalently be written as
1371: \begin{eqnarray}
1372: &W(\omega,\q) = \mathrm{Tr} \, [Kj^{0\dagger}(\q)\mcs_{\mathrm
1373: {eff}} (\tilde\omega) j^0(\q)]
1374:  = -\frac {1} {\pi} \mathrm{Im}\, \mathrm{Tr}\, [Kj^{0\dagger}(\q)
1375: \mcg_{\mathrm {eff}}(\tilde\omega)j^0(\q)] \, ,  \nonumber \\
1376: & \tilde\omega = \omega + E_0 -\epsilon_0 - \tilde\epsilon,
1377: \label{eq:3.77}
1378: \end{eqnarray}
1379: where
1380: \begin{equation}
1381: \mcs_{\mathrm {eff}}(\tilde\omega) = \frac {1} {2\pi i}
1382: [\mcg_{\mathrm {eff}}^{\dagger}(\tilde\omega) - \mcg_{\mathrm
1383: {eff}}(\tilde\omega)] . \label{eq:3.78}
1384: \end{equation}
1385: 
1386: So far, we have considered only the scalar component $J^0(\q)$ of the
1387: current in the momentum representation. The extension to an arbitrary
1388: one-body operator
1389: \begin{equation}
1390: O(\q) = \int\diff\r\diff\rf a_{\mathbf r}^{\dagger} \, O(\q;\r,\rf)
1391: a_{\mathbf r'} \label{eq:3.79}
1392: \end{equation}
1393: is performed on sight. Thus, Eq.~(\ref{eq:3.77}) is generalized to all
1394: the diagonal components of the hadronic tensor, provided that only
1395: one-body currents are considered:
1396: \begin{eqnarray}
1397: & W^{\mu\mu}(\omega,\q)  =  \mathrm{Tr} \, [Kj^{\mu\dagger}(\q)
1398: \mcs_{\mathrm {eff}}(\tilde\omega) j^{\mu}(\q)]
1399:  = -\frac {1} {\pi} \mathrm{Im}\, \mathrm{Tr}\, [Kj^{\mu\dagger}(\q)
1400: \mcg_{\mathrm {eff}}(\tilde\omega)j^{\mu}(\q)] ,  \nonumber \\
1401: & \tilde\omega = \omega + E_0 -\epsilon_0 - \tilde\epsilon,
1402: \label{eq:3.80}
1403: \end{eqnarray}
1404: 
1405: 
1406: \subsection{Energy shift prescriptions \label{ssec:3.7}}
1407: 
1408: In the literature two types of energy shifts have been proposed to
1409: connect the spectral functions $\mcs_n(E)$.
1410: 
1411: a) Kinetic energy prescription~\cite{hori,chinn,capuzzi,capma}:
1412: \begin{equation}
1413: \mcs_n(\omega + E_0 -\epsilon_n) = \mcs_0(\omega + E_0 -\epsilon_0 -
1414: \delta\epsilon_n) \, , \, \delta\epsilon_n = \epsilon_n - \epsilon_0,
1415: \label{eq:3.81}
1416: \end{equation}
1417: which produces an energy shift equal to $\epsilon_n - \epsilon_0$ in
1418: the expression of the hadronic tensor. This approximation keeps the
1419: value of the argument of the spectral function corresponding to the
1420: kinetic energy of the emitted proton. This explains the name.
1421: 
1422: b) Total energy prescription~\cite{capma}:
1423: \begin{equation}
1424: \mcs_n(\omega + E_0 -\epsilon_n) = \mcs_0(\omega + E_0 -\epsilon_0 -
1425: \delta\epsilon_n) \, , \, \delta\epsilon_n = 0, \label{eq:3.82}
1426: \end{equation}
1427: which yields no energy shift in the expression of the hadronic
1428: tensor. This approximation keeps the value of the total energy of the
1429: system given by the proton and the residual nucleus in the state
1430: $|n\rangle$.
1431: 
1432: The same prescriptions are applied to $\mcg_n(E)$ and $\mcm_n(E)$
1433: according to Eqs.~(\ref{eq:3.61}), and (\ref{eq:3.62}).
1434: 
1435: The total energy prescription has the merit of being simple, since it
1436: requires no shift. In contrast, the kinetic energy prescription leads
1437: to an expression of the hadronic tensor which involves a convolution
1438: of $\mcs_0(E)$ with the hole spectral function related to
1439: $|\psi_0\rangle$~\cite{capma}.
1440: 
1441: The kinetic energy prescription is the one that has been adopted by
1442: most authors [Refs.~\cite{hori,chinn,capuzzi}]. Moreover, it is
1443: closely related to an expression proposed for high energy transfers
1444: [Refs.~\cite{benhar,ciofi,benhar2,sick,benhar3,benhar4}]. However,
1445: we decided for an intermediate choice, due to the reasons explained
1446: below.
1447: 
1448: Using the relations (see Eqs.~(\ref{eq:2.12}) and (\ref{eq:2.17}))
1449: \begin{equation}
1450: \mcs_n(E) = \frac {1} {2\pi i} [ \mcg_n^{\dagger}(E) - \mcg_n(E)]\, ,
1451: \, \mcg_n(E) = \frac {1} {E - \mct - \mcm_n(E) + i \eta} ,
1452: \label{eq:3.83}
1453: \end{equation}
1454: the spectral function $\mcs_n(E)$ is decomposed as
1455: \begin{equation}
1456: \mcs_n(E) = \mcs_n^{\mathrm {el}}(E) + \mcs_n^{\mathrm {in}}(E)
1457: \label{eq:3.84}
1458: \end{equation}
1459: with
1460: \begin{equation}
1461: \mcs_n^{\mathrm {el}}(E) = \eta \, \mcg_n^{\dagger}
1462: (E+i\eta)\mcg_n(E+i\eta) , \label{eq:3.85}
1463: \end{equation}
1464: \begin{equation}
1465: \mcs_n^{\mathrm {in}}(E) = - \mcg_n^{\dagger} (E) \mcm_n^{\mathrm
1466: I}(E)\mcg_n(E) \, , \, \mcm_n^{\mathrm I} = \frac {1} {2 i}
1467: [\mcm_n(E) - \mcm_n^{\dagger}(E)] . \label{eq:3.86}
1468: \end{equation}
1469: The limit for $\eta \rightarrow 0$ is understood, as usual, in
1470: Eq.~(\ref{eq:3.85}).
1471: 
1472: The pioneering paper of Ref.~\cite{hori} has shown that for $E > 0$
1473: $\mcs_n^{\mathrm {el}}(E)$ yields the elastic contribution, due to
1474: the overlaps $\langle n|a_{\mathbf r}|\psi_{\mathbf{k},n}\rangle$,
1475: where $|\psi_{\mathbf{k},n}\rangle$ describes a proton of kinetic
1476: energy $k^2/2m = E$ scattered by the residual nucleus in the state
1477: $|n\rangle$. The asymptotic condition of incident plane wave imposes
1478: that $\mcs_n^{\mathrm {el}}(E)$ must be related to $\mcs_0^{\mathrm
1479: {el}}(E)$ so as to preserve $\k$ and hence the argument $E$. This is
1480: in favour of Eq.~(\ref{eq:3.81}) rather than Eq.~(\ref{eq:3.82}). In
1481: contrast, $\mcs_n^{\mathrm {in}}(E)$ gives the inelastic
1482: contribution, due to the overlaps $\langle n|a_{\mathbf
1483: r}|\psi_{\mathbf{k},m}\rangle$, with $m \neq n$, which do not contain
1484: the plane wave. Therefore, the requirement of preserving $\k$ is not
1485: so stringent in this case, and prevails the exigency of conserving in
1486: Eq.~(\ref{eq:3.86}) the size of $\mcm_n^{\mathrm I}(E)$, which is
1487: mainly determined by the number of the inelastic channels which are
1488: open when a proton of kinetic energy $k^2/2m$ is scattered by
1489: $|n\rangle$. This number depends on the total energy $k^2/2m +
1490: \epsilon _n$. Hence, Eq.~(\ref{eq:3.81}) does not correctly yield the
1491: thresholds of the inelastic processes, and Eq.~(\ref{eq:3.82}) is
1492: favoured to relate $\mcm_n^{\mathrm I}(E)$ to $\mcm_0^{\mathrm
1493: I}(E)$. Since the real and the imaginary parts of the self-energy
1494: fulfill a dispersion relation (see Sec.~5.4 of Ref.~\cite{capma2}),
1495: the total energy prescription must be applied to $\mcm_n(E)$ and hence
1496: to $\mcg_n(E)$ and $\mcs_n^{\mathrm {in}}(E)$.
1497: 
1498: Due to the above reasons, we adopt the kinetic energy prescription
1499: for $\mcs_n^{\mathrm {el}}(E)$ and the total energy one for
1500: $\mcs_n^{\mathrm {in}}(E)$, i.e.,
1501: \begin{equation}
1502: \mcs_n(\omega + E_0 - \epsilon_n) = \mcs_0^{\mathrm {el}}(\omega +
1503: E_0 - \epsilon_0 - \bar \epsilon) + \mcs_0^{\mathrm {in}}(\omega +
1504: E_0 - \epsilon_0) . \label{eq:3.87}
1505: \end{equation}
1506: Here, $\bar \epsilon$ is taken equal to the average excitation energy
1507: of the residual nucleus given in terms of the energies $\epsilon_n$
1508: and the spectroscopic factors $\lambda_n$, as
1509: \begin{equation}
1510: \bar \epsilon = \frac {\sum_n \lambda_n (\epsilon_n - \epsilon_0)}
1511: {\sum_n \lambda_n} . \label{eq:3.88}
1512: \end{equation}
1513: According to Eq.~(\ref{eq:3.87}), the effective spectral function
1514: $\mcs_{n,\mathrm {eff}}(E)$, defined as in Eq.~(\ref{eq:3.78}), and
1515: which differs from $\mcs_n(E)$ by corrections involving only
1516: $\mcm'_n(E)$, is related to the ground state effective spectral
1517: function $\mcs_{\mathrm {eff}}(E)$ of Eq.~(\ref{eq:3.78}) as
1518: \begin{equation}
1519: \mcs_{n,\mathrm {eff}}(\omega + E_0 - \epsilon_n) = \mcs_{\mathrm
1520: {eff}}^{\mathrm {el}}(\omega + E_0 - \epsilon_0 - \bar \epsilon) +
1521: \mcs_{\mathrm {eff}}^{\mathrm {in}}(\omega + E_0 - \epsilon_0),
1522: \label{eq:3.89}
1523: \end{equation}
1524: where
1525: \begin{equation}
1526: \mcs_{\mathrm {eff}}^{\mathrm {el}}(E) = \eta \, \mcg_{\mathrm
1527: {eff}}^{\dagger} (E+i\eta) \mcg_{\mathrm {eff}}(E+i\eta),
1528: \label{eq:3.90}
1529: \end{equation}
1530: \begin{equation}
1531: \mcs_{\mathrm {eff}}^{\mathrm {in}}(E) = - \mcg_{\mathrm
1532: {eff}}^{\dagger} (E) \mcm_{\mathrm {eff}}^{\mathrm I}(E)
1533: \mcg_{\mathrm {eff}}(E) \, , \, \mcm_{\mathrm {eff}}^{\mathrm I}(E)
1534: \equiv \frac {1} {2i} (\mcm_{\mathrm {eff}}
1535:  - \mcm_{\mathrm {eff}}^{\dagger}) ,
1536: \label{eq:3.91}
1537: \end{equation}
1538: and $\mcm_{\mathrm {eff}}(E)$ is the effective mass operator defined
1539: in Eq.~(\ref{eq:3.74}). Therefore, Eq.~(\ref{eq:3.80}) becomes
1540: \begin{equation}
1541: W^{\mu\mu} ( \omega , \q) = {\mathrm {Tr}} \, \big\{ Kj^{\mu \dagger}
1542: (\q) \big[\mcs_{\mathrm {eff}}^{\mathrm {el}} (\omega + E_0
1543: -\epsilon_0 - \bar\epsilon) + \mcs_{\mathrm {eff}}^{\mathrm {in}}
1544: (\omega + E_0 -\epsilon_0)\big] j^\mu (\q) \big\}. \label{eq:3.92}
1545: \end{equation}
1546: 
1547: 
1548: \subsection{Hadronic tensor in terms of the local equivalent
1549: potential \label{ssec:3.8}}
1550: 
1551: We have expressed the hadronic tensor in terms of the self-energy,
1552: rather than in terms of the Feshbach's potential as done in
1553: Ref.~\cite{capuzzi}, since the latter has a quite complicated
1554: spatial nonlocality~\cite{capma2} and its kernel is not symmetric in
1555: the coordinate representation. These drawbacks make problematic to
1556: find a local equivalent potential to be compared with the empirical
1557: optical-model potential. In contrast, the self-energy is symmetric,
1558: has a simpler nonlocal structure, and is considered the theoretical
1559: mean field most closely related to the empirical
1560: potential~\cite{capma2}.
1561: 
1562: Here, we want to show that the hadronic tensor can be expressed in
1563: terms of the Perey-Buck local potential, phase equivalent to the
1564: self-energy. To this extent, we investigate the connection between
1565: the Green's function $\mcg_0(E)$ of the self-energy $\mcm_0(E)$ and 
1566: that related to
1567: its local equivalent potential. We start with a simple model of
1568: $\mcm_0(E)$, that is widely used to calculate the self-energy by means
1569: of dispersion relations (see Sec.~4.4 of Ref.~\cite{mabor}). It
1570: takes into account the fact that the nonlocality in the self-energy
1571: is gathered mainly in its statical part, which is hermitian, whereas
1572: the non-hermitian dynamical part has a weaker nonlocal structure.
1573: Thus, one sets
1574: \begin{equation}
1575: \mcm_0(E) = \mcu + \mcd(E), \label{eq:3.93}
1576: \end{equation}
1577: where $\mcd(E)$ is non-hermitian and local, whereas the energy
1578: independent term $\mcu$ is hermitian and has the Frahn-Lemmer
1579: nonlocal structure
1580: \begin{equation}
1581: \mcu (\r , \rf) = H(|\r - \rf|) \, U\left(\frac {1} {2}|\r + \rf|
1582: \right). \label{eq:3.94}
1583: \end{equation}
1584: The precise shape of the nonlocality form factor $H$ is not relevant
1585: for the next equations. Setting
1586: \begin{equation}
1587: F(\q^2) \equiv \int \diff \s \, \exp (-i\q\cdot\s) H(s) ,
1588: \label{eq:3.95}
1589: \end{equation}
1590: the Perey-Buck potential~\cite{PB62,cap-lectnot} is defined by the implicit equation
1591: \begin{equation}
1592: \mcv_{\mathrm L}(E) = U\, F\big( E-\mcv_{\mathrm L} (E)\big) +
1593: \mcd(E) . \label{eq:3.96}
1594: \end{equation}
1595: 
1596: Every eigenvector $|\chi_{\mathrm L}(E)\rangle$ of $\mcv_{\mathrm L}
1597: (E)$ is related to the corresponding eigenvector $|\chi(E)\rangle$ of
1598: $\mcm_0(E)$ by the phase conserving relation
1599: \begin{equation}
1600: |\chi_{\mathrm L}(E)\rangle = f(E)|\chi(E)\rangle , \label{eq:3.97}
1601: \end{equation}
1602: where the inverse Perey factor $f(E)$~\cite{fied,cap-lectnot} multiplies by 
1603: the function
1604: \begin{equation}
1605: f(E;r) = \big[ 1 + U(r)F'\big( E-\mcv_{\mathrm L} (E;r)\big)
1606: \big]^{1/2} . \label{eq:3.98}
1607: \end{equation}
1608: Higher-order corrections~\cite{cap-lectnot}, involving surface terms,
1609: have been disregarded in Eqs.~(\ref{eq:3.96}) and (\ref{eq:3.98}).
1610: Analogously, the Green's functions related to $\mcv_{\mathrm L}$ and
1611: $\mcm_0$
1612: \begin{equation}
1613: \mcg_{\mathrm L} (E) \equiv [E - \mct - \mcv_{\mathrm L} (E) +
1614: i\eta]^{-1} , \label{eq:3.99}
1615: \end{equation}
1616: \begin{equation}
1617: \mcg_0 (E) \equiv [E - \mct - \mcm_0 (E) + i\eta]^{-1}, \label{eq:3.100}
1618: \end{equation}
1619: are connected by
1620: \begin{equation}
1621: \mcg_{\mathrm L} (E) = f(E) \mcg_0(E) f(E) . \label{eq:3.101}
1622: \end{equation}
1623: The energy derivatives of the two sides of Eq.~(\ref{eq:3.96}) yield
1624: \begin{equation}
1625: 1 - \mcm'_0 (E) = f(E) [1 - \mcv'_{\mathrm L} (E)] f(E) .
1626: \label{eq:3.102}
1627: \end{equation}
1628: This relation is analogous to the relation among the effective masses
1629: $m^*$, $m_k$, and $m_\omega$ in the infinite nuclear matter (Sec.~3.6
1630: of \cite{mabor}).
1631: 
1632: In Eqs.~(\ref{eq:3.71}), (\ref{eq:3.74}), and (\ref{eq:3.78}), we have
1633: defined the effective Green's function of $\mcg_0$ and the linked
1634: Hamiltonian, self-energy, and spectral function as
1635: \begin{eqnarray}
1636: \mcg_{\mathrm {eff}}(E) & = & \, [1 - \mcm'_0(E)]^{1/2} \, \mcg_0 (E) [1
1637: - \mcm'_0(E)]^{1/2} \, , \nonumber \\
1638:  h_{\mathrm {eff}}(E) & = & \, E - \mcg_{\mathrm {eff}}^{-1}(E)\, ,
1639: \nonumber \\
1640: \mcm_{\mathrm {eff}}(E) & = & \, h_{\mathrm {eff}}(E) - \mct\, ,
1641: \nonumber \\
1642: \mcs_{\mathrm {eff}}(E) & = & \, \frac {1} {2\pi i} [\mcg_{\mathrm
1643: {eff}}^{\dagger}(E) - \mcg_{\mathrm {eff}}(E)] . \label{eq:3.103}
1644: \end{eqnarray}
1645: Likewise, the corresponding quantities related to $\mcg_{\mathrm L}
1646: (E)$ and $\mcv_{\mathrm L} (E)$ are defined as
1647: \begin{eqnarray}
1648: \mcg_{\mathrm {L,eff}}(E) & = & \, [1 - \mcv'_{\mathrm L}(E)]^{1/2}
1649: \, \mcg_{\mathrm L} (E) [1 - \mcv'_{\mathrm L}(E)]^{1/2} \, ,
1650: \nonumber
1651: \\
1652:  h_{\mathrm {L,eff}}(E) & = & \, E - \mcg_{\mathrm {L,eff}}^{-1}(E)\,
1653: ,
1654:  \nonumber \\
1655: \mcv_{\mathrm {L,eff}}(E) & = & \, h_{\mathrm {L,eff}}(E) - \mct\, ,
1656: \nonumber \\
1657: \mcs_{\mathrm {L,eff}}(E) & = & \, \frac {1} {2\pi i} [\mcg_{\mathrm
1658: {L,eff}}^{\dagger}(E) - \mcg_{\mathrm {L,eff}}(E)] . \label{eq:3.104}
1659: \end{eqnarray}
1660: Inserting Eq.~(\ref{eq:3.102}) in the first Eq.~(\ref{eq:3.103}) and
1661: then using Eq.~(\ref{eq:3.101}), one has
1662: \begin{equation}
1663: \mcg_{\mathrm {eff}} (E) = \mcg_{\mathrm {L,eff}} (E),
1664: \label{eq:3.105}
1665: \end{equation}
1666: and then
1667: \begin{equation}
1668: h_{\mathrm {eff}} (E) = h_{\mathrm {L,eff}} (E) \, , \, \mcm_{\mathrm
1669: {eff}} (E) = \mcv_{\mathrm {L,eff}} (E)\, , \, \mcs_{\mathrm {eff}}
1670: (E) = \mcs_{\mathrm {L,eff}} (E) . \label{eq:3.106}
1671: \end{equation}
1672: We emphasize that this invariance property, which will be very useful
1673: below, is the mere consequence of having introduced into the hadronic
1674: tensor the contribution of the interference between different
1675: channels.
1676: 
1677: As a consequence of Eqs.~(\ref{eq:3.105}) and (\ref{eq:3.106}), the
1678: expression (\ref{eq:3.92}) of the hadronic tensor reads
1679: \begin{eqnarray}
1680: W^{\mu\mu} ( \omega , \q) = & {\mathrm{Tr}} \, \left\lbrace
1681: Kj^{\mu \dagger} (\q) \left[  \mcs_{\mathrm {L,eff}}^{\mathrm{el}}
1682: (\omega + E_0 -\epsilon_0 - \bar\epsilon) \right. \right. \nonumber\\
1683: & + \left. \left. \mcs_{\mathrm{L,eff}}^{\mathrm{in}} (\omega + E_0
1684: -\epsilon_0 ) \right]  j^\mu (\q) \right\rbrace \label{eq:3.107}
1685: \end{eqnarray}
1686: with
1687: \begin{equation}
1688: \mcs_{\mathrm {L,eff}}^{\mathrm {el}}(E) \equiv \eta \, \mcg_{\mathrm
1689: {L,eff}}^{\dagger} (E+i\eta) \mcg_{\mathrm {L,eff}}(E+i\eta) ,
1690: \label{eq:3.108}
1691: \end{equation}
1692: \begin{equation}
1693: \mcs_{\mathrm {L,eff}}^{\mathrm {in}}(E) = - \mcg_{\mathrm
1694: {L,eff}}^{\dagger} (E) \mcv_{\mathrm {L,eff}}^{\mathrm I}(E)
1695: \mcg_{\mathrm {L,eff}}(E) \, , \label{eq:3.109}
1696: \end{equation}
1697: where $\mcv_{\mathrm {L,eff}}^{\mathrm I}(E)$ is the antihermitian
1698: part of $\mcv_{\mathrm {L,eff}}(E)$. Using the first two
1699: Eqs.~(\ref{eq:3.104}) and Eq.~(\ref{eq:3.99}), one readily obtains
1700: $h_{\mathrm {L,eff}}(E)$ in terms of $h_{\mathrm {L}}(E)$:
1701: \begin{equation}
1702: h_{\mathrm {L,eff}}(E) = [1 - \mcv'_{\mathrm {L}}(E)]^{-1/2}
1703: [h_{\mathrm {L}}(E) - \mcv'_{\mathrm {L}}(E) E] [1 - \mcv'_{\mathrm
1704: {L}}(E)]^{-1/2} . \label{eq:3.111}
1705: \end{equation}
1706: Therefore, in principle, the hadronic tensor, as well as its elastic
1707: and inelastic parts, can be calculated in terms of quantities related
1708: to $\mcv_{\mathrm {L}}(E)$.
1709: 
1710: 
1711: 
1712: 
1713: \subsection{Practical calculations \label{ssec:3.9}}
1714: 
1715: The local potential $\mcv_{\mathrm {L}}(E)$, phase equivalent to the
1716: self-energy, can be identified with the empirical optical-model
1717: potential. Therefore, the symbol $\mcv_{\mathrm {L}}(E)$ will denote
1718: in the following the empirical potential.
1719: 
1720: The use of two different energy shifts to treat the elastic and
1721: inelastic parts of the hadronic tensor represents a complication. In
1722: order to overcome this difficulty, it is sufficient to find a simple
1723: way of calculating the elastic part.
1724: 
1725: We only deal with positive values of $E$, since we consider only
1726: scattering states. Let $|\chi_{\mathrm {L,eff}}^{(-)}(E)\rangle$
1727: denote the eigenvector of $h_{\mathrm {L,eff}}^{\dagger}(E)$, related
1728: to the eigenvalue $E$, and satisfying the condition of incoming
1729: scattering wave. The right hand side of Eq.~(\ref{eq:3.108}) can be
1730: handled to obtain
1731: \begin{equation}
1732: \mcs_{\mathrm {L,eff}}^{\mathrm{el}}(E) = \sum |\chi_{\mathrm
1733: {L,eff}}^{(-)}(E)\rangle \langle \chi_{\mathrm{L,eff}}^{(-)}(E)| ,
1734: \label{eq:3.112}
1735: \end{equation}
1736: where the sum refers to the understood degeneracy indices. The
1737: nontrivial proof of Eq.~(\ref{eq:3.112}) can be found in Sec.~4.12
1738: of Ref.~\cite{capma2}. We remark that this proof extends, without any
1739: modification, to a general Green's function and its self-energy.
1740: 
1741: Equation (\ref{eq:3.112}) involves the eigenvectors $|\chi_{\mathrm
1742: {L,eff}}^{(-)}(E)\rangle$ of $h_{\mathrm {L,eff}}^{\dagger}(E)$,
1743: which is a complicated Hamiltonian, but can be expressed in terms of
1744: the corresponding eigenvectors $|\chi_{\mathrm {L}}^{(-)}(E)\rangle$
1745: of $h_{\mathrm {L}}^{\dagger}(E)$. In fact, by a simple substitution
1746: in the eigenvalue equation for $h_{\mathrm {L,eff}}^{\dagger}(E)$ and
1747: using Eq.~(\ref{eq:3.111}), one checks that these eigenvectors are
1748: related by the phase-conserving relation:
1749: \begin{equation}
1750: |\chi_{\mathrm {L,eff}}^{(-)}(E)\rangle = [ 1 - (\mcv'_{\mathrm
1751: {L}})^{\dagger}(E)]^{1/2} |\chi_{\mathrm{L}}^{(-)}(E) \rangle.
1752: \label{eq:3.113}
1753: \end{equation}
1754: Therefore, Eq.~(\ref{eq:3.112}) reads
1755: \begin{equation}
1756: \mcs_{\mathrm {L,eff}}^{\mathrm{el}}(E) = \sum [ 1 - (\mcv'_{\mathrm
1757: {L}})^{\dagger}(E)]^{1/2} |\chi_{\mathrm {L}}^{(-)}(E)\rangle \langle
1758: \chi_{\mathrm{L}}^{(-)}(E)| [ 1 - (\mcv'_{\mathrm {L}})(E)]^{1/2} .
1759: \label{eq:3.114}
1760: \end{equation}
1761: 
1762: From this equation, the elastic contribution to the hadronic tensor
1763: can be calculated with the same difficulty as for the calculation of
1764: an integrated single-proton knockout. In contrast, no similar way for
1765: directly calculating the inelastic contribution is available. In
1766: order to overcome this difficulty, the hadronic tensor can be
1767: written as
1768: \begin{equation}
1769: W^{\mu\mu} (\omega, \q) = W_1^{\mu\mu} (\omega, \q) + W_2^{\mu\mu}
1770: (\omega, \q) \label{eq:3.115}
1771: \end{equation}
1772: with
1773: \begin{eqnarray}
1774: & & W_1^{\mu\mu} (\omega , \q) = {\mathrm {Tr}} \, \big [ Kj^{\mu
1775: \dagger} (\q) \mcs_{\mathrm {L,eff}} (E) j^\mu (\q) \big]
1776:  \nonumber \\ & = & - \frac {1} {\pi} {\mathrm {Im}}
1777: {\mathrm {Tr}} \big\{ Kj^{\mu \dagger} (\q)[ 1 - (\mcv'_{\mathrm
1778: {L}})(E)]^{1/2} \mcg_{\mathrm {L}} (E)[ 1 - (\mcv'_{\mathrm
1779: {L}})(E)]^{1/2}j^\mu (\q)\big\} , \label{eq:3.116}
1780: \end{eqnarray}
1781: and
1782: \begin{equation}
1783: W_2^{\mu\mu} (\omega , \q) = {\mathrm {Tr}} \, \big\{ Kj^{\mu
1784: \dagger} (\q) [\mcs_{\mathrm {L,eff}}^{\mathrm {el}} (E -
1785: \bar\epsilon) - \mcs_{\mathrm {L,eff}}^{\mathrm {el}} (E)] j^\mu (\q)
1786: \big\} , \label{eq:3.117}
1787: \end{equation}
1788: where $E = \omega + E_0 -\epsilon_0$.
1789: 
1790: In the following, $W_2^{\mu\mu} (\omega , \q)$ will be calculated
1791: using Eq.~(\ref{eq:3.114}), and $W_1^{\mu\mu} (\omega , \q)$ will be
1792: obtained from the method of Ref.~\cite{capuzzi}, based on the
1793: spectral representation of the s.p. Green's function.
1794: 
1795: 
1796: \section{Spectral representation of the hadronic tensor}
1797: \label{sec.spectral}
1798: 
1799: 
1800: In this Section we consider the spectral representation of the s.p.
1801: Green's function which allows practical calculations of the hadron
1802: tensor of Eq.~(\ref{eq:3.116}). Due to the complex nature of
1803: $\mcv_{\mathrm L}(E)$, the spectral representation of $\mcg_{\mathrm
1804: L}(E)$ involves a biorthogonal expansion in terms of the
1805: eigenfunctions of $h_{\mathrm L}(E)$ and $h_{\mathrm
1806: L}^{\dagger}(E)$. We consider the incoming wave scattering solutions
1807: of the eigenvalue equations, i.e.,
1808: \begin{eqnarray}
1809: & & \left(\mathcal{E} - h_{\mathrm L}^{\dagger}(E)\right) \mid
1810: \chi_{\mathrm L,\mathcal{E}}^{(-)}(E)\rangle = 0 \ , \label{eq.inco1}
1811: \\
1812: & & \left(\mathcal{E} - h_{\mathrm L}(E)\right) \mid \tilde
1813: {\chi}_{\mathrm L,\mathcal{E}}^{(-)}(E)\rangle = 0 \ .
1814: \label{eq.inco2}
1815: \end{eqnarray}
1816: The choice of incoming wave solutions is not strictly necessary, but
1817: it is convenient in order to have a closer comparison with the
1818: treatment of the exclusive reactions, where the final states fulfill
1819: this asymptotic condition and are the eigenfunctions
1820: $\mid\chi_{\mathrm L,E}^{(-)}(E)\rangle$ of $h_{\mathrm
1821: L}^{\dagger}(E)$.
1822: 
1823: The eigenfunctions of Eqs.~(\ref{eq.inco1}) and (\ref{eq.inco2})
1824: satisfy the biorthogonality condition
1825: \begin{equation}
1826: \langle\chi_{\mathrm L,\mathcal{E}}^{(-)}(E)\mid \tilde
1827: {\chi}_{\mathrm L,\mathcal{E}'}^{(-)}(E)\rangle = \delta
1828: \left(\mathcal{E} - \mathcal{E}' \right) \ , \label{bicon}
1829: \end{equation}
1830: and, in absence of bound eigenstates, the completeness relation
1831: \begin{equation}
1832: \int_0^{\infty} \diff \mathcal{E}\mid\tilde {\chi}_{\mathrm
1833: L,\mathcal{E}}^{(-)}(E)\rangle\langle \chi_{\mathrm
1834: L,\mathcal{E}}^{(-)}(E)\mid =1 . \label{eq.comple}
1835: \end{equation}
1836: 
1837: Equations (\ref{bicon}) and (\ref{eq.comple}) are the mathematical
1838: basis for the biorthogonal expansions. The contribution of possible
1839: bound state solutions of Eqs.~(\ref{eq.inco1}) and (\ref{eq.inco2})
1840: can be disregarded in Eq.~(\ref{eq.comple}) since their effect on the
1841: hadron tensor is negligible at the energy and momentum transfers
1842: considered in this paper.
1843: 
1844: Using Eqs.~(\ref{eq.comple}) and (\ref{eq.inco2}), one obtains the
1845: spectral representation
1846: \begin{equation}
1847: \mcg_{\mathrm L}(E) = \int_0^{\infty} \diff \mathcal{E}\mid\tilde
1848: {\chi}_{\mathrm L,\mathcal{E}}^{(-)}(E)\rangle
1849: \frac{1}{E-\mathcal{E}+i\eta} \langle\chi_{\mathrm
1850: L,\mathcal{E}}^{(-)}(E)\mid \ . \label{eq.sperep}
1851: \end{equation}
1852: Therefore, $W_1^{\mu\mu} (\omega , \q)$ (Eq.~(\ref{eq:3.116})) can be
1853: written, after expanding the ODM $K$ in terms of the natural
1854: orbitals, as
1855: \begin{eqnarray}
1856: W_1^{\mu\mu}(\omega , \q) = -\frac{1}{\pi} \sum_{\nu} \mathrm{Im}
1857: \bigg[  \int_0^{\infty}  & \diff \mathcal{E} & \frac{1}{\omega + E_0 -
1858: \epsilon_0-\mathcal{E}+i\eta} \nonumber\\
1859: & \times &  T_{\nu}^{\mu\mu} (\mathcal{E} ,\omega + E_0 -
1860: \epsilon_0) \bigg]   \
1861: , \label{eq.pracw}
1862: \end{eqnarray}
1863: where
1864: \begin{eqnarray}
1865: T_{\nu}^{\mu\mu}(\mathcal{E} ,E) & = & n_{\nu}\langle u_{\nu} \mid
1866: j^{\mu\dagger}(\q) \sqrt{1-\mcv'_{\mathrm L}(E)}
1867: \mid\tilde{\chi}_{\mathrm L,\mathcal{E}}^{(-)}(E)\rangle \nonumber \\
1868: & \times &
1869:  \langle\chi_{\mathrm L,\mathcal{E}}^{(-)}(E)\mid
1870:  \sqrt{1-\mcv'_{\mathrm L}(E)} j^{\mu}(\q)\mid u_{\nu} \rangle \ .
1871: \label{eq.tprac}
1872: \end{eqnarray}
1873: The quantities $|u_{\nu}\rangle$ and $n_\nu$ are the natural orbitals
1874: and the occupation numbers, defined in Eq.~(\ref{eq:2.5}), and come
1875: from the natural expansion of the ODM $K$ of Eq.~(\ref{eq:3.116}). The
1876: limit for $\eta \rightarrow +0$, understood before the integral of
1877: Eq.~(\ref{eq.pracw}), can be calculated exploiting the standard
1878: symbolic relation
1879: \begin{equation}
1880: \lim_{\eta \rightarrow +0} \frac{1}{E-\mathcal{E}+i\eta} =
1881: \mathcal{P} \left(\frac{1}{E-\mathcal{E}}\right) - i \pi \delta
1882: \left(E-\mathcal{E}\right) \ , \label{eq.princ}
1883: \end{equation}
1884: where $\mathcal{P}$ denotes the principal value of the integral.
1885: Therefore, Eq.~(\ref{eq.pracw}) reads
1886: \begin{eqnarray}
1887: W_1^{\mu\mu} (\omega , \q) & = & \sum_{\nu} \Bigg[ \mathrm{Re}
1888: T_{\nu}^{\mu\mu} (\omega + E_0 - \epsilon_0, \omega + E_0 -
1889: \epsilon_0)
1890: \nonumber \\
1891: & - & \frac{1}{\pi} \mathcal{P} \int_0^{\infty} \diff \mathcal{E}
1892: \frac{1}{\omega + E_0 -\epsilon_0-\mathcal{E}} \mathrm{Im}
1893: T_{\nu}^{\mu\mu} (\mathcal{E},\omega + E_0 - \epsilon_0) \Bigg] \ ,
1894: \label{eq.finale}
1895: \end{eqnarray}
1896: which separately involves the real and imaginary parts of
1897: $T_{\nu}^{\mu\mu}$.
1898: 
1899: The contribution coming from Eq.~(\ref{eq:3.117}) can be calculated
1900: using Eq.~(\ref{eq:3.114}), i.e. as
1901: \begin{eqnarray}
1902: W_{2}^{\mu\mu}(\omega, \q) & = & \sum _{\nu}n_{\nu}\big [\langle
1903: u_{\nu} | j^{\mu\dagger}(\q) \sqrt{1-\mcv'_{\mathrm L}(\bar E)}
1904: |{\chi}_{\mathrm L,\bar E}^{(-)} (\bar E)\rangle \nonumber \\
1905: &\times& \langle\chi_{\mathrm L,\bar E}^{(-)}(\bar E)|
1906: \sqrt{1-\mcv'_{\mathrm L}(\bar E)} j^{\mu}(\q)| u_{\nu} \rangle
1907: \nonumber \\
1908: & - & \langle u_{\nu} | j^{\mu\dagger}(\q) \sqrt{1-\mcv'_{\mathrm
1909: L}(E)} |{\chi}_{\mathrm L, E}^{(-)}(E)\rangle \nonumber \\
1910: &\times& \langle\chi_{\mathrm L,E}^{(-)}(E)|
1911: \sqrt{1-\mcv'_{\mathrm L}(E)} j^{\mu}(\q)| u_{\nu} \rangle \big ] \
1912: , \label{eq.w2}
1913: \end{eqnarray}
1914: with $E = \omega + E_0 - \epsilon_0$ and $\bar E = E -\bar \epsilon$.
1915: In this equation, only the optical potential wave functions of
1916: Eq.~(\ref{eq.inco1}) appear everywhere, and the calculation is similar
1917: to that of the integral of the exclusive cross section, but for the
1918: different energy values which are involved and the presence of
1919: occupation numbers and natural orbitals.
1920: 
1921: Some remarks on Eq.~(\ref{eq.finale}) are in order. Let us disregard
1922: the square root correction, due to interference effects, and the
1923: minor contribution of the integral over the energy. One has
1924: \begin{equation}
1925: W_1^{\mu\mu} (\omega , \q) \simeq \sum_{\nu} n_{\nu} \mathrm{Re}
1926: \big [ \langle \chi_{\mathrm L,E}^{(-)}(E) | j^{\mu}(\q) |
1927: u_{\nu}\rangle \langle u_{\nu}| j^{\mu\dagger}(\q) |{\tilde
1928: \chi}_{\mathrm L,E}^{(-)}(E)\rangle \big ] . \label{eq:165}
1929: \end{equation}
1930: Let us compare Eq.~(\ref{eq:165}) with the corresponding elastic
1931: contribution due to $\mcs_{\mathrm {L}}^{\mathrm{el}}(E)$, given by
1932: Eq.~(\ref{eq:3.114}) without the square root corrections, i.e.
1933: \begin{eqnarray}
1934: W_1^{{\mathrm{el}} \, \mu\mu} (\omega , \q) = \sum_{\nu} n_{\nu}
1935: \langle \chi_{\mathrm L,E}^{(-)}(E) | j^{\mu}(\q) | u_{\nu}\rangle
1936: \langle u_{\nu}| j^{\mu\dagger}(\q) |{\chi}_{\mathrm
1937: L,E}^{(-)}(E)\rangle . \label{eq:166}
1938: \end{eqnarray}
1939: In Eq.~(\ref{eq:166}) the attenuation of the strength, mathematically
1940: due to the imaginary part of $\mcv_{\mathrm L}^{\dagger}(E)$, is
1941: related to the flux lost toward the inelastic channels. In the
1942: inclusive response this attenuation must be compensated by a
1943: corresponding gain due to the inelastic contribution to $W_1^{\mu\mu}
1944: (\omega , \q)$. In the description provided by Eq.~(\ref{eq:165}),
1945: including both elastic and inelastic contributions, the attenuation
1946: of strength of the first factor in the square bracket is compensated
1947: by the second factor, where the imaginary part of $\mcv_{\mathrm
1948: L}(E)$ has the effect of increasing the strength. We want to stress
1949: here that in the Green's function approach it is just the imaginary
1950: part of the optical potential which accounts for the redistribution
1951: of the strength among different channels.
1952: 
1953: The main difference of the present approach with the one of
1954: Ref.~\cite{capuzzi} is that now we are able to include in the initial
1955: states the effect of correlations. In Ref.~\cite{capuzzi}, indeed,
1956: the initial states were calculated as the overlaps between the target
1957: nucleus and the residual nucleus described as a hole state,
1958: corresponding to the separation energy taken from phenomenology, or
1959: computed through an independent-particle model. Here, the initial
1960: states are described through a realistic ODM, including correlations.
1961: However, this goal is obtained using a constant separation energy,
1962: i.e. the one corresponding to the ground state of the residual
1963: nucleus, and this produces an undue shift of the cross section. The
1964: contribution of Eq.~(\ref{eq:3.117}) is added in order to compensate
1965: the shift.
1966: 
1967: 
1968: \section{Realistic one-body density and correlations
1969: \label{sec.density}}
1970: 
1971: The main issue of this paper is, together with a new and completely
1972: antisymmetrized presentation of the Green's function approach, to
1973: investigate the effect of nuclear correlations in the inclusive
1974: quasielastic electron scattering. Correlations are included in the
1975: ODM, which is expressed within the natural orbital representation
1976: \cite{Low55} in the form
1977: \begin{equation}
1978: K(\mathbf{r},\mathbf{r}') \equiv \langle\r|K|\rf\rangle = 
1979:  \sum_\nu n_\nu u^*_\nu(\r') u_\nu(\r).
1980: \label{eq:density}
1981: \end{equation}
1982: 
1983: The calculations presented in the next Section are done and compared 
1984: for different realistic density
1985: matrices which include the contribution of short-range and tensor
1986: correlations and are obtained within different approaches. In
1987: particular, we consider the Jastrow correlation method (JCM)
1988: \cite{SAD93}, the correlated basis function (CBF) method
1989: \cite{NDD+97}, and the Green's function method (GFM) \cite{PMD96}.
1990: 
1991: %\subsection{The Jastrow correlation method}
1992: 
1993: In Ref.~\cite{SAD93} the ODM is obtained within the JCM in its low-order
1994: approximation (LOA).
1995: The JCM incorporates the nucleon-nucleon (NN)
1996: short-range correlations (SRC) starting from the ansatz for 
1997: the wave-function of $A$ fermions \cite{Jas55}:
1998: \begin{equation}
1999: \Psi ^A(\mathbf{r}_1,\mathbf{r}_2,\ldots ,\mathbf{
2000: r}_{A})=(C_{A})^{-1/2} \prod_{1\leq i<j\leq A} F(\mid \mathbf{
2001: r}_{i}-\mathbf{r}_{j}\mid ) \Phi^{A}( \mathbf{r}_1,\mathbf{
2002: r}_2,\ldots ,\mathbf{r}_{A}), \label{eq:jas1}
2003: \end{equation}
2004: where $C_A$ is a normalization constant and $\Phi^A$ is a single
2005: Slater determinant wave function built from harmonic-oscillator (HO) s.p.
2006: wave functions which depend on the oscillator parameter
2007: $\alpha_{osc.}$, having the same value for both protons and neutrons.
2008: Only central correlations are included in the correlation factor
2009: $F(r)$, which is state-independent and has a simple Gaussian form
2010: \begin{equation} \label{eq:jas2}
2011: F(r) = 1- \exp (-\beta ^{2}r^{2}),
2012: \end{equation}
2013: where the correlation parameter $\beta $ determines the healing
2014: distance. The LOA \cite{GGR71} keeps all terms up to the second order
2015: in $(F-1)$ and the first order in $(F^2-1)$ in such a way that the
2016: normalization of the density matrices is ensured order by order.
2017: Under the above assumptions analytical expressions for the ODM and
2018: corresponding natural orbitals have been obtained in \cite{SAD93}.
2019: The values of the parameters $\alpha_{osc.}$ and $\beta $ have been
2020: obtained \cite{SAD93} phenomenologically by fitting the experimental
2021: elastic form factor data. Thus, in the present calculations the
2022: following values of the parameters are used: $\alpha_{osc.}=0.59$
2023: fm$^{-1}$, $\beta =1.43$ fm$^{-1}$ for $^{16}$O and
2024: $\alpha_{osc.}=0.52$ fm$^{-1}$, $\beta =1.21$ fm$^{-1}$ for
2025: $^{40}$Ca, respectively.
2026: 
2027: %\subsection{The CBF theory}
2028: 
2029: In the CBF method the trial $A$-particle wave function has the form
2030: \begin{equation} \label{eq:vn1}
2031: \Psi^A (x_{1},...,x_{A})=\mathcal{S}\left[ \prod_{i<j=1}^{A} F
2032: (x_{i},x_{j})\right] \Phi^A (x_{1},...,x_{A}),
2033: \end{equation}
2034: where $x_i$ are particle coordinates which contain spatial, spin, and
2035: isospin variables, $\mathcal{S}$ is a symmetrization operator, and
2036: $\Phi^A$ is an uncorrelated (Slater determinant) wave function
2037: normalized to unity and describing a closed-shell spherical system.
2038: The correlation factor $F$ is generally written as
2039: \begin{equation} \label{eq:vn2}
2040: F(x_{i},x_{j})=\sum_{n}h_{n}(|\mathbf{r}_{i}-{\bf
2041: r}_{j}|)\hat{\mathcal{O}}_n
2042: \end{equation}
2043: with basic two-nucleon operators $\hat{\mathcal{O}}_n$ inducing
2044: central, spin-spin, tensor and spin-orbit correlations, either with
2045: or without isospin exchange.
2046: The ODM generated by a CBF-type correlated wave function for $^{16}$O
2047: has been constructed in \cite{NDD+97}. The LOA,
2048: which keeps terms up to the first order of the function
2049: $h(x_{i},x_{j};x_{i}',x_{j}')= F(x_{i}',x_{j}') F(x_{i},x_{j})-1$, has
2050: been used. The s.p. orbits entering the Slater determinant
2051: $\Phi^A$ are taken from a Hartree-Fock calculation with the
2052: Skyrme-III effective force. The correlation factor $F(x_{1},x_{2})$
2053: obtained in \cite{PWP92} by variational Monte Carlo calculations with
2054: Argonne $v_{14}$ NN forces has been used. The two-nucleon correlation
2055: factors were restricted to the central, spin-isospin and
2056: tensor-isospin operators. 
2057: 
2058: 
2059: %\subsection{The Green function approach}
2060: 
2061: For a nucleus like $^{16}$O with $J=0$ ground state angular momentum
2062: the ODM can easily be separated into submatrices
2063: of a given orbital angular momentum $l$ and total angular momentum
2064: $j$. Within the GFM \cite{DM92,MPD95} the ODM in
2065: momentum representation can be evaluated from the imaginary part of
2066: the s.p. Green's function by integrating
2067: \begin{equation} \label{eq:m1}
2068: K_{lj}(k_1,k_2)=\int_{-\infty }^{\varepsilon_{F}}dE\frac{1}{\pi}
2069: \text{Im} (\mathcal{G}_{lj}(k_1,k_2;E)),
2070: \end{equation}
2071: 
2072: where the energy variable $E$ corresponds to the energy difference
2073: between the ground state of the $A$ particle system and the energies
2074: of the states in the $(A-1)$-particle system (negative $E$ with large
2075: absolute value correspond to high excitation energies of the residual
2076: system) and $\varepsilon _{F}$ is the Fermi energy. The
2077: s.p. Green's function $\mathcal{G}_{lj}$ is obtained from
2078: the solution of the Dyson equation
2079: \begin{eqnarray}\label{eq:m2}
2080: \mathcal{G}_{lj}(k_1,k_2;E)&= & \mathcal{G}_{lj}^{(0)}(k_1,k_2;E)+
2081: \int dk_3 \int dk_4 \mathcal{G}_{lj}^{(0)} (k_1,k_3;E) \nonumber \\
2082: & \times & \Delta\Sigma_{lj}(k_3,k_4;E) \mathcal{G}_{lj}(k_4,k_2;E),
2083: \end{eqnarray}
2084: where $\mathcal{G}^{(0)}$ refers to the Hartree-Fock propagator and
2085: $\Delta\Sigma _{lj}$ represents contributions to the real and
2086: imaginary part of the irreducible self-energy, which go beyond the
2087: Hartree-Fock approximation of the nucleon self-energy used to derive
2088: $\mathcal{G}^{(0)}$. The results for the ODM have been analyzed in
2089: \cite{PMD96} in terms of the natural orbitals and the occupation
2090: numbers in $^{16}$O. Within the natural orbital
2091: representation they can be determined by diagonalizing the ODM of 
2092: the correlated system.
2093: The numerical results from \cite{PMD96} show that the ODM can be
2094: described quite accurately in terms of four natural orbitals for each
2095: partial wave $lj$. 
2096: 
2097: 
2098: \section{Results and discussion \label{sec.results}}
2099: 
2100: The model presented in this paper has been applied to evaluate the response
2101: functions of the inclusive quasielastic electron scattering for the
2102: nuclei $^{16}$O and $^{40}$Ca. For the nucleus $^{16}$O, no data are
2103: available for the separated longitudinal and transverse response
2104: functions. However, we can compare the results produced by several
2105: realistic ODM, obtained within different correlation methods. For the
2106: nucleus $^{40}$Ca, experimental data are available and we can be
2107: compare our results with them.
2108: 
2109: Calculations are performed in a nonrelativistic approach including
2110: the contributions of both $W_1$ (Eq.~(\ref{eq.finale})) and $W_2$
2111: (Eq.~(\ref{eq.w2})), but disregarding the second term in $W_1$. The
2112: evaluation of this term, which contains the principal value, requires
2113: the integration over all the eigenfunctions of the continuum spectrum
2114: of the optical potential and represents a quite complicate task.
2115: Moreover, it would be useless in a nonrelativistic approach where the
2116: contribution of this term is very small~\cite{capuzzi}. In $W_1$ and
2117: $W_2$ the sum over all the natural orbitals $u_\nu$ is involved and
2118: partial occupation numbers $n_\nu$ are included in the model. All
2119: the results presented in the following are normalized to the number
2120: of nucleons in the nucleus.
2121: The s.p. energies $\epsilon_n$ and the spectroscopic factors
2122: $\lambda_n$ we have used in the calculation of the 
2123: $W_2$ contribution are those corresponding to the overlap states obtained 
2124: in Ref.~\cite{SAD96} for JCM, in \cite{NDD+97} for CBF, and in \cite{GPD+97}
2125: for GFM, following the method of Ref.~\cite{NWH93}.
2126: 
2127: The other ingredients in the matrix elements of Eqs.~(\ref{eq.tprac})
2128: and (\ref{eq.w2}) are the same as those taken in our previous
2129: application of the Green's function approach to the inclusive
2130: electron scattering (Ref.~\cite{capuzzi}) and in our treatment of the
2131: exclusive one-nucleon knockout, which is based on the distorted wave
2132: impulse approximation and was widely and successfully applied to the
2133: analysis of $(e,e'p)$ data~\cite{book,DWEEPY}. The one-body nuclear
2134: current operator is given by the nonrelativistic approximation of
2135: Ref.~\cite{McVoy}, $\tilde \chi$ and ${\chi}$ are eigenfunctions of a
2136: phenomenological spin-dependent optical potential and of its
2137: hermitian conjugate, determined through a fit to elastic
2138: nucleon-nucleus scattering data including cross sections and
2139: polarizations~\cite{Schwandt}. This allows a consistent treatment of
2140: FSI in the inclusive and in the exclusive electron scattering. The
2141: role of FSI in the Green's function approach was already investigated
2142: in Ref.~\cite{capuzzi} in a nonrelativistic framework and, more
2143: recently, in Ref.~\cite{meucci} in a relativistic framework and will
2144: not be discussed later on in this paper.
2145: 
2146: %%%%%%%%%%%%%%% Fig. 1%%%%%%%%%%%%%%%%%%%
2147: \begin{figure}[ht]
2148: \begin{center}
2149: \includegraphics[height=15cm,width=12cm]{Fig1.eps}
2150: \vskip -0.1cm
2151: \caption {Longitudinal (upper panel) and transverse
2152: (lower panel) response functions for the $^{16}$O$(e,e')$ reaction at
2153: $q = 400$ MeV$/c$. The curves indicated by JCM $W_1$ and JCM
2154: $W_1+W_2$ give the contributions of $W_1$ and of the sum $W_1+ W_2$
2155: (see Eqs.~(\ref{eq.finale}) and (\ref{eq.w2})) with the ODM of the
2156: JCM~\cite{SAD93}. SM is obtained using the SM prescription of
2157: Ref.~\cite{capuzzi} with the phenomenological s.p. wave functions of
2158: Ref.~\cite{ES}. SM $W_1+W_2$ is the sum of the two terms $W_1$ and
2159: $W_2$ where the ODM is replaced by the SM prescription
2160: (Eq.~(\ref{eq.SM2})) with the phenomenological s.p. wave functions. SM
2161: HO $W_1$ gives the contribution of $W_1$ in the SM with harmonic
2162: oscillator wave functions.} \label{fig1}
2163: \end{center}
2164: \end{figure}
2165: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2166: A numerical example for the longitudinal and transverse response
2167: functions of the $^{16}$O$(e,e')$ reaction at $q = 400$ MeV$/c$ is
2168: given in Fig.~\ref{fig1}. The curves called JCM $W_1$ and JCM
2169: $W_1+W_2$ give the contribution of $W_1$ and of the sum $W_1+W_2$ and
2170: are calculated with the ODM within the JCM in its LOA, 
2171: where only short-range correlations are included. The
2172: SM curve corresponds to the prescription adopted in our previous
2173: applications of the Green's function approach~\cite{capuzzi}, where
2174: correlations are neglected and the components of the hadronic tensor
2175: are given by
2176: \begin{equation}
2177: W^{\mu\mu} (\omega , \q) = \sum_{n} \mathrm{Re} T_{n}^{\mu\mu}
2178: (\omega + E_0 - \epsilon_n, \omega + E_0 - \epsilon_n), \label{eq.SM}
2179: \end{equation}
2180: with
2181: \begin{eqnarray}
2182: T_{n}^{\mu\mu}(E,E) & = & \lambda_{n}\langle \varphi_{n} \mid
2183: j^{\mu\dagger}(\q) \sqrt{1-\mcv'_{\mathrm L}(E)}
2184: \mid\tilde{\chi}_{\mathrm L,E}^{(-)}(E)\rangle \nonumber \\
2185: & \times &
2186:  \langle\chi_{\mathrm L,E}^{(-)}(E)\mid
2187:  \sqrt{1-\mcv'_{\mathrm L}(E)} j^{\mu}(\q)\mid \varphi_{n} \rangle \
2188: . \label{eq.SM1}
2189: \end{eqnarray}
2190: A pure shell model (SM) is assumed for the nuclear structure: the sum
2191: is over all the occupied states in the SM and a unitary spectral
2192: strength ($\lambda_n=1$) is taken for each s.p. state $\varphi_n$. In
2193: the calculations for the SM curve the wave functions $\varphi_n$ are
2194: taken from a phenomenological Woods-Saxon potential ~\cite{ES} and
2195: $\epsilon_n$ are the experimental excitation energies of the states
2196: $n$.
2197: 
2198: The difference between the correlated JCM $W_1+W_2$ and the
2199: uncorrelated SM results indicates that correlations give a
2200: redistribution of the strength (the total strength is conserved in
2201: all the calculations presented in this paper) and a reduction of the
2202: response functions by $\sim 10\%$ in the peak region. Part of the
2203: difference is however due to the different methods. If the sum
2204: $W_1+W_2$ is calculated in a SM approach, where the ODM is replaced
2205: by
2206: \begin{equation}
2207: \langle\r|K|\rf\rangle = \sum_{n} \lambda_n \varphi^*_n(\r')
2208: \varphi_n(\r), \label{eq.SM2}
2209: \end{equation}
2210: the corresponding result SM $W_1+W_2$, which is also displayed in
2211: the figure, lies between the SM and JCM $W_1+W_2$ curves, and the
2212: reduction produced by SRC in the peak region is
2213: no more than $\sim 5\%$. 
2214: The different position of the peaks of the curves JCM $W_1+W_2$ 
2215: and SM is mainly due to the following reasons. i) In the JCM $W_1+W_2$ 
2216: curve the inelastic part of the spectral function $\mcs_n$ is calculated using 
2217: the total energy prescription of Eq.~(\ref{eq:3.82}) which does not produce a
2218: shift, differently from the kinetic energy prescription (\ref{eq:3.81}) 
2219: adopted for the SM curve. The effect of these different prescriptions is
2220: indicated by the distance between the peaks of SM $W_1+W_2$ and SM. 
2221: ii) In the cases JCM $W_1+W_2$ and SM the calculations
2222: use different values of $\epsilon_n$. The resulting effect is indicated 
2223: by the distance between the peaks of JCM $W_1+W_2$ and SM $W_1+W_2$.
2224: 
2225: The term $W_2$ gives a shift of the calculated response functions
2226: that is important to determine the position of the peak and to bring
2227: it close to the SM result. In the previous application of
2228: Ref.~\cite{capuzzi} the SM result was able to give a fair description
2229: of the size and shape of the experimental longitudinal response of
2230: the $^{12}$C$(e,e')$ reaction, in a range of momentum transfer
2231: between 400 and 550 MeV$/c$. In contrast, the experimental transverse
2232: response was generally underestimated. For the transverse response,
2233: however, an important contribution might be given by two-body
2234: meson-exchange currents, which are not included in the present model
2235: based on the s.p. Green's function.
2236: 
2237: The last curve drawn in Fig.~\ref{fig1} gives the contribution of
2238: $W_1$ calculated in the SM approach (Eq.~(\ref{eq.SM2})) and with the
2239: HO single-particle wave functions used in the
2240: calculation of the ODM with the JCM (SM HO $W_1$). The comparison
2241: with the JCM $W_1$ result indicates that correlation effects are
2242: still within $\sim 5\%$, but in this case they enhance the response
2243: functions in the peak region. Phenomenological Woods-Saxon wave
2244: functions certainly represent a more reliable basis for a SM
2245: calculation. A calculation with HO wave functions, however, may allow
2246: a more consistent comparison with the correlated ODM of the JCM and
2247: may give an idea of the uncertainty produced by different s.p. wave
2248: functions.
2249: 
2250: The comparison between the uncorrelated SM and the correlated JCM
2251: results shows that the effects of correlations depend on the
2252: uncorrelated result that is considered for the comparison. The
2253: effects of SRC are however small and within the
2254: range of uncertainty produced by the choice of the theoretical
2255: ingredients. From this point of view our results are in substantial
2256: agreement with those of Ref.~\cite{Co}, where the effects of
2257: SRC on the response functions of the
2258: quasielastic electron scattering were investigated in a different
2259: model. In Fig.~\ref{fig1} correlation effects have a similar behavior
2260: on the longitudinal and transverse responses, while in Ref.~\cite{Co}
2261: correlations act differently on the two responses. This apparent
2262: discrepancy might be explained by the different models used here and
2263: in Ref.~\cite{Co}, by the small effects produced by correlations, and
2264: by the uncertainties related to the other ingredients of the
2265: calculations.
2266: 
2267: 
2268: %%%%%%%%%%%%%%% Fig. 2%%%%%%%%%%%%%%%%%%%
2269: \begin{figure}[ht]
2270: \begin{center}
2271: \includegraphics[height=15cm,width=12cm]{Fig2.eps}
2272: \vskip -0.1cm
2273: \caption {Longitudinal response functions for the
2274: $^{16}$O$(e,e')$ reaction at $q = 400$ MeV$/c$. The contributions of
2275: $W_1$ and of the sum $W_1+ W_2$ with the ODM of the
2276: CBF~\cite{NDD+97} and of the GFM~\cite{PMD96} are displayed in the
2277: upper and lower panels, respectively. The SM result is the same as in
2278: Fig.~\ref{fig1}.} \label{fig2}
2279: \end{center}
2280: \end{figure}
2281: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2282: The longitudinal response functions of the $^{16}$O$(e,e')$ reaction
2283: at $q = 400$ MeV$/c$ obtained with the ODM within the
2284: CBF~\cite{NDD+97} and the GFM~\cite{PMD96} are shown in Fig.
2285: \ref{fig2}. In both calculations the ODM includes both short-range
2286: and tensor correlations. The two density matrices, which are obtained
2287: within different correlation methods, give similar results. In
2288: comparison with the SM curve, correlations produce a redistribution
2289: of the strength and a reduction of the response in the peak region by
2290: $\sim 15\%$. In comparison with the SM $W_1+ W_2$ result (that is
2291: not shown Fig.~\ref{fig2}, but is given in Fig.~\ref{fig1}) the
2292: reduction is within $10\%$, but anyhow larger than for the JCM
2293: density matrix, where only SRC are included. The
2294: term $W_2$ gives a shift of the response function that brings the
2295: position of the maximum toward the SM result.
2296: 
2297: %%%%%%%%%%%%%%% Fig. 3%%%%%%%%%%%%%%%%%%%
2298: \begin{figure}[ht]
2299: \begin{center}
2300: \includegraphics[height=8cm,width=12cm]{Fig3.eps}
2301: \vskip -0.1cm
2302: \caption {Longitudinal response functions for the
2303: $^{40}$Ca$(e,e')$ reaction at $q = 450$ MeV$/c$. The contributions of
2304: $W_1$ and of the sum $W_1+W_2$ with the ODM of the
2305: JCM~\cite{SAD93} are displayed together with the SM and SM $W_1+
2306: W_2$ results produced by the phenomenological s.p. wave functions of
2307: Ref.~\cite{ES}. The Saclay data (open squares) are from
2308: Ref.~\cite{meziani}, the MIT-Bates (black circles) are from
2309: Ref.~\cite{batesca}. } \label{fig3}
2310: \end{center}
2311: \end{figure}
2312: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2313: The longitudinal response functions of the $^{40}$Ca$(e,e')$ reaction
2314: at $q = 450$ MeV$/c$ is shown in Fig.~\ref{fig3}. The JCM $W_1$ and
2315: JCM $W_1+W_2$ results are compared with the SM and SM $W_1+W_2$ ones
2316: and with the available data~\cite{meziani,batesca}. The main
2317: difference between the different results is in the position of the
2318: maximum. The contribution of $W_2$ shifts the response by $\sim$ 15
2319: MeV at larger values of the energy transfer. The difference between 
2320: the SM $W_1+W_2$ and SM results and between the JCM $W_1+W_2$
2321: and SM $W_1+W_2$ results indicates the sensitivity of the position of the
2322: peak to the different prescriptions used for the energy shifts and to
2323: the values of $\epsilon_n$ used in the calculations. These effects
2324: are more important in a heavier nucleus like $^{40}$Ca than in
2325: $^{16}$O. In comparison with data, our results are closer to the
2326: Saclay data.
2327: 
2328: 
2329: 
2330: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2331: \section{Summary and conclusions \label{conc}}
2332: 
2333: In this paper we have presented a completely antisymmetrized Green's
2334: function approach to inclusive quasielastic electron scattering.
2335: The main goals are the following.
2336: 
2337: i) To express the final state interaction in terms of the
2338: self-energy, rather than in terms of the Feshbach optical potential
2339: as was done in previous papers, since the mass operator is more
2340: closely related to the empirical optical-model potentials.
2341: Therefore, we have developed a theoretical approach based on extended 
2342: projection operators, as in Ref.~\cite{capma}.
2343: 
2344: ii) To include an approximated treatment of the interference between
2345: different reaction channels, which is usually disregarded. 
2346: This has required some modifications of the treatment of Ref.~\cite{capma},
2347: resulting in a different separation of the hadronic tensor into a 
2348: direct and an interference part. The latter one is essential to explain the
2349: replacement of the self-energy with the empirical potential.
2350: 
2351: iii) To separate the elastic and inelastic contributions to the
2352: hadronic tensor, in order to introduce different approximations
2353: concerning the dependence on the state of the residual nucleus.
2354: 
2355: iv) To include in the calculation the effects of correlations in the
2356: target nucleus through the use of a realistic one-body density
2357: matrix.
2358: 
2359: The method has been applied to the reactions $^{16}$O($e,e'$) and
2360: $^{40}$Ca($e,e'$). Results produced by density matrices including
2361: short-range as well as tensor correlations have been compared. The
2362: effects of short range correlations are small (within $\sim 5\%$) and
2363: within the range of uncertainty related to the choice of other
2364: ingredients of the calculation (e.g. the s.p. wave functions). This
2365: result is in substantial agreement with previous
2366: investigations~\cite{Co}, but a similar effect is found in the
2367: present work on the longitudinal and the transverse response
2368: functions. Stronger effects are obtained when also tensor
2369: correlations are included. For the $^{16}$O($e,e'$) reaction at $q =
2370: 450$ MeV$/c$ the height of the peak is reduced by $\sim 10\%$ when
2371: both types of correlations are considered. A correction dependent on
2372: the s.p. energies is necessary [the term $W_2$ in Eq.~(\ref{eq.w2})]
2373: in order to determine the position of the peak. 
2374: The comparison of the present results
2375: with the available experimental data for the nucleus $^{40}$Ca gives
2376: a better agreement, in the longitudinal response, with the Saclay
2377: data than with the MIT-Bates ones.
2378: 
2379: In order to get deeper insight into the dependence of the inclusive
2380: electron scattering on correlations, one should also compute the
2381: contribution of two-body currents and their interference with tensor
2382: correlations, that seems to give a larger
2383: effect\cite{Leidemann,Fabrocini,Sick}. In order to accomplish this
2384: task, however, a new approach, based on the two-particle Green's
2385: function and including the two-body density matrix, should be
2386: developed.
2387: 
2388: 
2389: \appendix
2390: \section{APPENDIX}
2391: 
2392: A real number $a$ belongs to the continuous spectrum of a
2393: self-adjoint operator $A$ if and only if:
2394: 
2395: i) it does not belong to the discrete spectrum,
2396: 
2397: ii) for every $\eta>0$, one can find a normalizable vector
2398: $|\psi_{a,\eta}\rangle$ and a constant $C_a$ such that
2399: \begin{equation}
2400: \| (a - A)|\psi_{a,\eta}\rangle \|^2 \leq C_a \||\psi_{a,\eta}\rangle
2401: \|^2 \eta . \label{eq:a.1}
2402: \end{equation}
2403: 
2404: The vectors $|\psi_{a,\eta}\rangle$ are called {\lq\lq approximate
2405: eigenvectors of $A$ related to the eigenvalue $a$\rq\rq} [see
2406: Sec.~8.1 of Ref~\cite{richt}]. Now, we construct a set of
2407: approximate eigenvectors of the Hamiltonian $H$ of the residual
2408: nucleus, which, added to the exact bound eigenvectors $|n\rangle$,
2409: satisfy a completeness relation in the limit $\eta \rightarrow +0$.
2410: We remark that this property is not an automatic consequence of
2411: Eq.~(\ref{eq:a.1}).
2412: 
2413: Let $\mch_c$ be the Hilbert subspace spanned by the non-normalizable
2414: eigenvectors $|\epsilon\rangle$ of the continuous spectrum of $H$
2415: which, for simplicity, is assumed to coincide with $[0,+\infty)$. Let
2416: $\{|u_k\rangle\}$ be a basis in $\mch_c$ consisting of normalizable
2417: vectors such that $\langle\epsilon|u_k\rangle$ are functions of
2418: $\epsilon$ bounded, continuous, and everywhere different from zero.
2419: This can be realized weighing the eigenvectors $|\epsilon\rangle$ by
2420: means of a complete orthonormal set of functions having the same
2421: properties. We set
2422: \begin{equation}
2423: |\epsilon,k;\eta\rangle \equiv \sqrt{\frac {\eta} {\pi}} (\epsilon -
2424: H + i\eta)^{-1} |u_k\rangle . \label{eq:a.2}
2425: \end{equation}
2426: Note that the vectors $|\epsilon,k;\eta\rangle$ and $|n\rangle$ are
2427: orthogonal.
2428: 
2429: \underline{Theorem 1}. The vectors $|\epsilon,k;\eta\rangle$ are
2430: approximate eigenvectors of $H$.
2431: 
2432: \underline{Proof}. One has
2433: \begin{equation}
2434: \lim_{\eta \rightarrow +0} \| |\epsilon,k;\eta\rangle \|^2 =
2435: |\langle\epsilon|u_k\rangle |^2 \neq 0 \label{eq:a.3}
2436: \end{equation}
2437: since
2438: \begin{equation}
2439:  \| |\epsilon,k;\eta\rangle \|^2 = \int_0^{\infty} \diff \, \epsilon'
2440:  L(\epsilon - \epsilon'; \eta) |\langle\epsilon'|u_k\rangle |^2 \, ,
2441: \,
2442:  L(x; \eta) \equiv \frac {\eta} {\pi} \frac {1} {x^2 + \eta^2} ,
2443: \label{eq:a.4}
2444: \end{equation}
2445: and, in the distributional sense of the limit,
2446: \begin{equation}
2447: \lim_{\eta \rightarrow +0} L(x; \eta) = \delta(x) . \label{eq:a.5}
2448: \end{equation}
2449: Moreover, the following inequality holds:
2450: \begin{equation}
2451:  \|(\epsilon - H) |\epsilon,k;\eta\rangle \|^2 =
2452:  \frac {\eta} {\pi}\langle u_k| \frac {(\epsilon - H)^2} {(\epsilon -
2453: H)^2 +
2454:  \eta^2} |u_k \rangle \leq \frac {\eta} {\pi}\langle u_k|u_k \rangle
2455: =
2456:  \frac {\eta} {\pi} .
2457: \label{eq:a.6}
2458: \end{equation}
2459: Therefore, one has
2460: \begin{equation}
2461: \lim_{\eta \rightarrow +0} \frac {\|(\epsilon - H)
2462: |\epsilon,k;\eta\rangle \|^2} {\||\epsilon,k;\eta\rangle \|^2} = 0
2463: \label{eq:a.7}
2464: \end{equation}
2465: and Eq.~(\ref{eq:a.1}) is satisfied. $\square$
2466: 
2467: \vskip 1cm
2468: 
2469: \underline{Theorem 2}. In the limit for $\eta \rightarrow +0$, one
2470: has the completeness relation
2471: \begin{equation}
2472: \sum_n |n\rangle \langle n| + \lim_{\eta \rightarrow +0} \sum_k
2473: \int_0^{\infty} \diff \epsilon \, |\epsilon,k;\eta\rangle \langle
2474: \epsilon,k;\eta| = 1 , \label{eq:a.8}
2475: \end{equation}
2476: where the convergence is understood in the weak sense, i.e., inside a
2477: scalar product between normalizable states.
2478: 
2479: \underline{Proof}. Due to the orthogonality between the vectors
2480: $|\epsilon,k;\eta\rangle$ and $|n\rangle$, it is sufficient to prove
2481: Eq.~(\ref{eq:a.8}) inside the scalar product between vectors
2482: $|\phi\rangle$ and $|\chi\rangle$ belonging to $\mch_c$, where both
2483: states $\{|\epsilon\rangle\}$ and $\{|u_k\rangle\}$ are complete.
2484: Thus the sum over $n$ does not contribute, and one considers only
2485: \begin{eqnarray}
2486: & & \sum_k \int_0^{\infty}\diff \epsilon \, \langle
2487: \phi|\epsilon,k;\eta\rangle \langle \epsilon,k;\eta|\chi\rangle =
2488: \frac {\eta} {\pi} \int_0^{\infty} \diff \epsilon \, \langle \phi|
2489: \frac {1} {(\epsilon - H)^2 +\eta^2}
2490: |\chi \rangle \nonumber \\
2491: & & = \int_0^{\infty}\diff \epsilon \, \int_0^{\infty}\diff
2492: \epsilon' \, L(\epsilon - \epsilon';\eta) \langle
2493: \phi|\epsilon'\rangle
2494: \langle\epsilon'|\chi\rangle \nonumber \\
2495: & & =\int_0^{\infty}\diff \epsilon' \, \langle \phi|\epsilon'\rangle
2496: \langle\epsilon'|\chi\rangle \int_{-\epsilon'}^{\infty} \diff x \,
2497: L(x;\eta) , \label{eq:a.9}
2498: \end{eqnarray}
2499: where in the last step we have exchanged the integrals, according to
2500: the Fubini theorem, since $|L(\epsilon - \epsilon';\eta) \langle
2501: \phi|\epsilon'\rangle \langle\epsilon'| \chi\rangle|$ is Lebesgue
2502: summable in $\epsilon$ and $\epsilon'$. Observing that
2503: \begin{eqnarray}
2504: |\langle \phi|\epsilon'\rangle \langle\epsilon'|\chi\rangle
2505: \int_{-\epsilon'}^{\infty} \diff x \, L(x;\eta) | & < & \, |\langle
2506: \phi|\epsilon'\rangle \langle\epsilon'|\chi\rangle
2507: \int_{-\infty}^{\infty} \diff x \, L(x;\eta) | \nonumber \\ & = &
2508: \langle \phi|\epsilon'\rangle \langle\epsilon'|\chi\rangle ,
2509: \label{eq:a.10}
2510: \end{eqnarray}
2511: where the last term is Lebesgue summable and does not depend on
2512: $\eta$, we can use the dominated convergence theorem to take the
2513: limit for $\eta\rightarrow +0$ within the first integral of the last
2514: term in Eq.~(\ref{eq:a.9}). Thus, using Eq.~(\ref{eq:a.5}), one
2515: obtains
2516: \begin{eqnarray}
2517: & & \lim_{\eta\rightarrow +0}\sum_k \int_0^{\infty}\diff \epsilon \,
2518: \langle \phi|\epsilon,k;\eta\rangle\langle
2519: \epsilon,k;\eta|\chi\rangle = \int_0^{\infty}\diff \epsilon'
2520: \,\langle\phi|\epsilon'\rangle \langle\epsilon'|\chi\rangle
2521: \int_{-\epsilon'}^{\infty} \diff x \,
2522: \delta(x) \nonumber \\
2523: & & =\int_0^{\infty}\diff \epsilon' \, \langle\phi|\epsilon'\rangle
2524: \langle\epsilon'|\chi\rangle = \langle\phi|\chi\rangle \, , \,
2525: \forall \, |\phi\rangle, |\chi\rangle \in \mch_c .\quad \square
2526: \label{eq:a.11}
2527: \end{eqnarray}
2528: 
2529: \vskip 1cm
2530: 
2531: Substituting the completeness relation of Eq.~(\ref{eq:3.14}) by
2532: (\ref{eq:a.8}) and introducing the projection operators
2533: \begin{equation}
2534: P(\epsilon,k;\eta) \equiv \int \diff \p \, \alpha_{\mathbf p}
2535: |\epsilon,k;\eta\rangle^{\mathrm N} \, ^{\mathrm N}\langle
2536: \epsilon,k;\eta|\alpha_{\mathbf p} \, , \, |\epsilon,k;\eta
2537: \rangle^{\mathrm N} \equiv \frac {|\epsilon,k;\eta\rangle}
2538: {\||\epsilon,k;\eta\rangle \|}, \label{eq:a.12}
2539: \end{equation}
2540: the contribution of the continuous spectrum is recovered adding to
2541: Eq.~(\ref{eq:3.22a}) the term
2542: \begin{eqnarray}
2543: & & \lim_{\eta\rightarrow +0}\sum_k \int\diff \epsilon \, \mathrm
2544: {Re} \int \diff \p \diff \pf \, \langle \psi_0|a_{\mathbf
2545: {p-q}}^{\dagger}|
2546: \epsilon,k;\eta\rangle \nonumber \\
2547: & \times & ^{\mathrm N}\langle\epsilon,k;\eta|\alpha_{\mathbf p}
2548: \delta (\omega + E_0 -\epsilon - \hat H_{\epsilon})\alpha_{\mathbf
2549: p'} | \epsilon,k;\eta\rangle^{\mathrm N} \, \,
2550: \langle\epsilon,k;\eta|a_{\mathbf p'} J^0(\q)|\psi_0\rangle ,
2551: \label{eq:a.13}
2552: \end{eqnarray}
2553: where $\hat H_{\epsilon}$ is defined analogously to $\hat H_n$ (see
2554: Eq.~(\ref{eq:2.24})), i.e. as
2555: \begin{equation}
2556: \hat H_{\epsilon} = H - \epsilon \quad \mathrm{in} \quad \mch^{(Z+1)}
2557: \quad , \quad \hat H_{\epsilon} = \epsilon - H \quad \mathrm{in}
2558: \quad \mch^{(Z-1)} . \label{eq:a.14}
2559: \end{equation}
2560: Thus, using the approximation
2561: \begin{equation}
2562: ^{\mathrm N}\langle\epsilon,k;\eta|\alpha_{\mathbf p} \delta (\omega
2563: + E_0 -\epsilon - \hat H_{\epsilon})\alpha_{\mathbf p'}
2564: |\epsilon,k;\eta\rangle^{\mathrm N} \simeq \langle \p| \mcs_0 (\omega
2565: + E_0 -\epsilon_0 - \tilde\epsilon) |\pf\rangle, \label{eq:a.15}
2566: \end{equation}
2567: analogous to Eq.~(\ref{eq:3.34}), and using again Eq.~(\ref{eq:a.8})
2568: one recovers Eqs.~(\ref{eq:3.37})--(\ref{eq:3.42}).
2569: 
2570: 
2571: \begin{ack}
2572: 
2573: We would like to thank A. N. Antonov for useful
2574: discussions and continuous encouragement. We are grateful to 
2575: S. S. Dimitrova, H. M\"uther and D. Van Neck for the results 
2576: of the ODM's we have used.
2577: 
2578: \end{ack}
2579: 
2580: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2581: \begin{thebibliography}{}
2582: \bibitem{book}
2583: S.~Boffi, C.~Giusti, F.~D.~Pacati, and M.~Radici,
2584: \textit{Electromagnetic Response of Atomic Nuclei}, Oxford Studies in
2585: Nuclear Physics, Vol. 20, Clarendon Press, Oxford, 1996; S.~Boffi,
2586: C.~Giusti, and F.~D.~Pacati, Phys. Rep. \textbf{226} (1993) 1.
2587: 
2588: \bibitem{batesca}
2589: C.~F.~Williamson, \textit{et al.}, Phys. Rev. C \textbf{56} 
2590: (1997) 3152.
2591: 
2592: \bibitem{csfrascati}
2593: M.~Anghinolfi, \textit{et al.}, Nucl. Phys. \textbf{A602}, (1996) 405.
2594: 
2595: \bibitem{jlabpro}
2596: J.~P.~Chen, S.~Choi, and Z.~E.~Meziani, spokespersons, JLab experiment
2597: E-01-016.
2598: 
2599: \bibitem{Sluys}
2600: V.~Van der Sluys, J.~Ryckebusch, and M.~Waroquier, Phys. Rev. C
2601: \textbf{51}, (1995) 2664.
2602: 
2603: \bibitem{Cenni}
2604: R.~Cenni, F.~Conte, and P.~Saracco, Nucl. Phys. \textbf{A623}, 
2605: (1997) 391.
2606: 
2607: \bibitem{Amaro}
2608: J.~E.~Amaro, M.~B.~Barbaro, J.~A.~Caballero, T.~W.~Donnelly, and
2609: A.~Molinari, Phys. Rep. \textbf{368}, (2002) 317; Nucl. Phys.
2610: \textbf{A723} (2003) 181.
2611: 
2612: \bibitem{Amaro2}
2613: J.~E.~Amaro, M.~B.~Barbaro, J.~A.~Caballero, and F.~Kazemi Tabatabaei,
2614: Phys. Rev. C \textbf{68} (2003) 014604.
2615: 
2616: \bibitem{Fabrocini}
2617: A.~Fabrocini, Phys. Rev. C \textbf{55} (1997) 338.
2618: 
2619: \bibitem{Co}
2620: G.~Co' and A.~M.~Lallena, Ann. Phys. \textbf{287} (2001) 101.
2621: 
2622: \bibitem{meucci}
2623: A.~Meucci, F.~Capuzzi, C.~Giusti, and F.~D.~Pacati, Phys. Rev. C
2624: \textbf{67} (2003) 054601.
2625: 
2626: \bibitem{Leidemann}
2627: W.~Leidemann and G.~Orlandini, Nucl. Phys. \textbf{A506} (1990) 447.
2628: 
2629: \bibitem{Sick}
2630: I.~Sick, in \textit{Nuclear Theory}, Heron Press Science Series,
2631: Sofia, 2002, p. 16.
2632: 
2633: \bibitem{hori}
2634: Y.~Horikawa, F.~Lenz, and N.~C.~Mukhopadhyay, Phys. Rev. C
2635: \textbf{22} (1980) 1680.
2636: 
2637: \bibitem{chinn}
2638: C.~R.~Chinn, A.~Picklesimer, and J.~W.~Van Orden, Phys. Rev. C
2639: \textbf{40} (1989) 790; Phys. Rev. C \textbf{40} (1989) 1159.
2640: 
2641: \bibitem{bouch}
2642: P.~M.~Boucher and J.~W.~Van Orden, Phys. Rev. C \textbf{43}
2643: (1991) 582.
2644: 
2645: \bibitem{capuzzi}
2646: F.~Capuzzi, C.~Giusti, and F.~D.~Pacati, Nucl. Phys. \textbf{A524}
2647:  (1991) 681.
2648: 
2649: \bibitem{capma}
2650: F.~Capuzzi and C.~Mahaux, Ann. Phys. (N.Y.) \textbf{254} (1997) 130.
2651: 
2652: \bibitem{fesh}
2653: H.~Feshbach, Ann. Phys. (N.Y.) \textbf{5} (1958) 357.
2654: 
2655: \bibitem{fesh2}
2656: H.~Feshbach, Ann. Phys. (N.Y.) \textbf{19} (1962) 287.
2657: 
2658: \bibitem{capma3}
2659: F.~Capuzzi and C.~Mahaux, Ann. Phys. (N.Y.) \textbf{245} (1996) 147.
2660: 
2661: \bibitem{fabro}
2662: A.~Fabrocini and S.~Fantoni, Nucl. Phys. \textbf{A503} (1989) 375.
2663: 
2664: \bibitem{jourd}
2665: J.~Jourdan, Nucl. Phys. \textbf{A603} (1996) 117.
2666: 
2667: \bibitem{fesh3}
2668: H.~Feshbach, \textit{Theoretical Nuclear Physics: Nuclear Reactions},
2669: Wiley, New York, 1992, and references contained therein.
2670: 
2671: \bibitem{moniz}
2672: E.~J.~Moniz, Phys. Rev. \textbf{184} (1969) 1154.
2673: 
2674: \bibitem{moniz2}
2675: E.~J.~Moniz, I.~Sick, R.~R.~Whitney, J.~R.~Ficenec, R.~D.~Kephart,
2676: and W.~P.~Trower, Phys. Rev. Lett. \textbf{26} (1971) 445.
2677: 
2678: \bibitem{tornow}
2679: W.~Tornow, Z.~P.~Chen, and J.~P.~Delaroche, Phys. Rev. C \textbf{42}
2680: (1990) 693 .
2681: 
2682: \bibitem{Cap}
2683: F.~Capuzzi, Nucl. Phys. \textbf{A554} (1993) 362.
2684: 
2685: \bibitem{benhar}
2686: O.~Benhar, A.~Fabrocini, S.~Fantoni, G.~A.~Miller, V.~R.~
2687: Pandharipande, and I.~Sick, Phys. Rev. C \textbf{44} (1991) 2328.
2688: 
2689: \bibitem{ciofi}
2690: C.~Ciofi degli Atti and S.~Simula, Phys. Lett. B \textbf{325}
2691: (1994) 276.
2692: 
2693: \bibitem{benhar2}
2694: O.~Benhar, A.~Fabrocini, S.~Fantoni, and I.~Sick, Nucl. Phys. A
2695: \textbf{579} (1994) 493.
2696: 
2697: \bibitem{sick}
2698: I.~Sick, S.~Fantoni, A.~Fabrocini, and O.~Benhar, Phys. Lett. B
2699: \textbf{323} (1994) 267.
2700: 
2701: \bibitem{benhar3}
2702: O.~Benhar, A.~Fabrocini, S.~Fantoni, and I.~Sick, Phys. Lett. B
2703: \textbf{343} (1995) 47.
2704: 
2705: \bibitem{benhar4}
2706: O.~Benhar, A.~Fabrocini, S.~Fantoni, V.~R.~Pandharipande, S.~C.~
2707: Pieper, and I.~Sick, Phys. Lett. B \textbf{359} (1995) 8.
2708: 
2709: \bibitem{mabor}
2710: C.~Mahaux, P.~F.~Bortignon, R.~A.~Broglia, and C.~H.~Dasso,
2711: Phys. Rep. \textbf{120} (1985) 1.
2712: 
2713: \bibitem{capma2}
2714: F.~Capuzzi and C.~Mahaux, Ann. Phys. (N.Y.) \textbf{239} (1995) 57.
2715: 
2716: \bibitem{PB62} F.~G.~Perey and B.~Buck, Nucl. Phys. \textbf{32} (1962) 353.
2717: 
2718: \bibitem{cap-lectnot}
2719: F.~Capuzzi, in \textit{Nuclear Optical Model Potential}, Lecture Notes
2720: in Physics, Vol. \textbf{55}, eds. J. Ehlers, \textit{et al.},
2721: Springer, Berlin, 1976, p. 20.
2722: 
2723: \bibitem{fied} H.~Fiedeldey, Nucl. Phys. \textbf{77} (1966) 149.
2724: 
2725: \bibitem{Low55} P.~-~O.~L\"owdin, Phys. Rev. \textbf{97} (1955) 1474.
2726: 
2727: \bibitem{SAD93} M.~V.~Stoitsov, A.~N.~Antonov, and S.~S.~Dimitrova,
2728: Phys. Rev. C \textbf{47} (1993) R455; Phys. Rev. C \textbf{48}
2729: (1993) 74; Z. Phys. A \textbf{345} (1993) 359.
2730: 
2731: \bibitem{NDD+97} D.~Van Neck, L.~Van Daele, Y.~Dewulf, and
2732: M.~Waroquier, Phys. Rev. C \textbf{56} (1997) 1398.
2733: 
2734: \bibitem{PMD96} A.~Polls, H.~M\"uther and W.~H.~Dickhoff,
2735: \textit{Proceedings of Conference on Perspectives in Nuclear Physics
2736: at Intermediate Energies}, Trieste, 1995, edited by S. Boffi, C.
2737: Ciofi degli Atti, and M.M. Giannini, World Scientific, Singapore,
2738: 1996, p. 308.
2739: 
2740: \bibitem{Jas55} R.~Jastrow, Phys. Rev. \textbf{98} (1955) 1479.
2741: 
2742: \bibitem{GGR71} M.~Gaudin, J.~Gillespie, and G.~Ripka, Nucl. Phys.
2743: \textbf{A176} (1971) 237; M.~Dal R\'{i}, S.~Stringari, and
2744: O.~Bohigas, Nucl. Phys. \textbf{A376} (1982) 81.
2745: 
2746: \bibitem{PWP92} S.~C.~Pieper, R.~B.~Wiringa, and V.~R.~Pandharipande,
2747: Phys. Rev. C \textbf{46} (1992) 1741.
2748: 
2749: \bibitem{DM92} W.~H.~Dickhoff and H.~M\"uther, Rep. Prog. Phys.
2750: \textbf{55} (1992) 1947.
2751: 
2752: \bibitem{MPD95} H.~M\"uther, A.~Polls, and W.~M.~Dickhoff, Phys. Rev.
2753: C \textbf{51} (1995) 3040; H.~M\"uther, G.~Knehr, and A.~Polls,
2754: Phys. Rev. C \textbf{52} (1995) 2955.
2755: 
2756: \bibitem{SAD96} M.~V.~Stoitsov, A.~N.~Antonov, and S.~S.~Dimitrova,
2757: Phys. Rev. C \textbf{53} (1996) 1254.
2758: 
2759: \bibitem{GPD+97} M.~K.~Gaidarov, K.~A.~Pavlova, S.~S.~Dimitrova,
2760: M.~V.~Stoitsov, A.~N.~Antonov, D.~Van Neck, and H.~Muether, Phys.
2761: Rev. C \textbf{60} (1999) 024312.
2762: 
2763: \bibitem{NWH93} D.~Van Neck, M.~Waroquier, and K.~Heyde, Phys. Lett.
2764: B \textbf{314} (1993) 255.
2765: 
2766: \bibitem{DWEEPY}
2767: C.~Giusti and F.~D.~Pacati, Nucl. Phys. \textbf{A473} (1987) 717;
2768: Nucl. Phys. \textbf{A485} (1988) 461.
2769: 
2770: \bibitem{McVoy}
2771: K.~W.~McVoy and L.~Van Hove, Phys. Rev. \textbf{125} (1962) 1034.
2772: 
2773: \bibitem{Schwandt}
2774: P.~Schwandt, H.~O.~Meyer, W.~W.~Jacobs, A.~D.~Bacher, S.~E.~Vigdor,
2775: M.~D.~Kaitchuck, and T.~R.~Donoghue, Phys. Rev. C \textbf{26}
2776: (1982) 55.
2777: 
2778: \bibitem{ES}
2779: L.~R.~B.~Elton and A.~Swift, Nucl. Phys. \textbf{A94} (1967) 52.
2780: 
2781: \bibitem{meziani}
2782: Z.~E.~Meziani, \textit{et al.}, Phys. Rev. Lett. \textbf{52}
2783: (1984) 2130; \textbf{54} (1985) 1233.
2784: 
2785: \bibitem{richt}
2786: R.~D.~Richtmyer, \textit{Principles of Advanced Mathematical Physics},
2787: Vol. \textbf{1}, Springer-Verlag, New York, 1978.
2788: 
2789: \end{thebibliography}
2790: 
2791: \end{document}
2792: