1:
2: %\documentclass[12pt]{article}
3: \documentclass[preprint,aps,draft,showpacs]{revtex4}
4:
5: \usepackage{epsfig}
6:
7: \begin{document}
8:
9: \title{The fully self-consistent charge-exchange quasiparticle random phase
10: approximation and its application to
11: the isobaric analog resonances}
12:
13: \author{S. Fracasso and G. Col\`o}
14:
15: \affiliation{Dipartimento di Fisica dell'Universit\`a degli Studi\\
16: and INFN, Sezione di Milano, via Celoria 16\\
17: 20133 Milano, Italy}
18:
19: \date{\today}
20:
21: \begin{abstract}
22: A microscopic model aimed at the description of charge-exchange nuclear excitations
23: along isotopic chains which include open-shell systems, is developed. It consists
24: of quasiparticle random phase approximation (QRPA) made on top of
25: Hartree-Fock-Bardeen-Cooper-Schrieffer (HF-BCS). The calculations are performed by using the Skyrme
26: interaction in the particle-hole channel and a zero-range, density-dependent
27: pairing force in the particle-particle channel. At variance with the (many) versions
28: of QRPA which are available in literature, in our work special emphasis is put
29: on the full self-consistency. Its importance, as well as the role played by the charge-breaking
30: terms of the nuclear Hamiltonian, like the Coulomb interaction, the
31: charge symmetry and charge independence breaking (CSB-CIB) forces
32: and the electromagnetic spin-orbit, are elucidated by means of numerical
33: calculations of the isobaric analog resonances (IAR). The theoretical energies of
34: these states along the chain of the Sn isotopes agree
35: well with the experimental data in the stable isotopes. Predictions for
36: unstable systems are presented.
37: \pacs{24.30.Cz, 21.60.Jz, 21.30.Fe, 25.40.Kv, 21.10.Sf, 27.60.+j}
38:
39: \end{abstract}
40:
41: \maketitle
42:
43: \section{INTRODUCTION}
44:
45: The significant lack of knowledge concerning many properties of the charge-exchange
46: nuclear excitations contrasts markedly with their importance for nuclear structure
47: and the impact which they have on many interesting physical phenomena.
48:
49: The charge-exchange transitions involve a change in $N$ and $Z$ of the nucleus,
50: keeping $A$ fixed. They can take place spontaneously, like in the well-known case of
51: $\beta$-decay, or be induced by external fields when in a nuclear reaction a given
52: amount of excitation energy $\Delta E$ and angular momentum $\Delta J$ is released to the
53: nucleus. The spectra of charge-exchange reactions, like ($p$,$n$) or ($^{3}$He,$t$), are
54: characterized by the emergence of collective isovector (i.e., $\Delta T$=1) giant
55: resonances (IVGRs) in analogy with the non charge-exchange case~\cite{Harakeh}.
56: However, a unifying picture of these $\Delta T_z$=$\pm$1 states is still, to a large
57: extent, missing. For instance, the $\Delta L$=0 charge-exchange isovector giant monopole
58: resonance (IVGMR) is one of the most elusive nuclear states, despite a long series
59: of experiments aimed at its identification~\cite{IVGMR}; at the same time, its
60: knowledge would be important for the determination of the ground state isospin
61: mixing. Also the higher multipoles, that is, the charge-exchange dipole, quadrupole
62: and octupole resonances, are basically unknown. This is mainly due to the lack of
63: really selective probes: in particular, the separation of the electric (i.e.,
64: $\Delta S$=0 or ``non spin-flip'') and magnetic (i.e., $\Delta S$=1 or ``spin-flip'') modes
65: is far from being trivial. On the other hand, a systematic pattern of the energy and
66: collectivity of these states would shed light on the strong uncertainties
67: concerning the isovector part of the nucleon-nucleon ({\itshape NN}) effective interaction and
68: the symmetry part of the nuclear equation of state.
69:
70: It has to be mentioned that knowing the properties of the nuclear charge-exchange
71: states allows also to attack other kinds of problems outside the realm of nuclear
72: structure. These states enter the description of the double $\beta$-decay, and
73: the need of a reliable theory of this process is a longstanding problem. More
74: generally, all the weak interaction processes within atomic nuclei involve
75: charge-exchange transitions as far as charged currents are involved. We have in
76: mind many processes which are of interest for neutrino physics, like the interaction
77: of these peculiar particles with nuclei, or for astrophysics, like the mechanisms
78: which are responsible for the evolution of neutron stars or the $\beta$-decay of isotopes which lie on
79: the $r$-process path of stellar nucleosynthesis.
80:
81: A significant exception to the unsatisfactory ignorance of the charge-exchange IVGRs
82: is provided by the availability of many experimental data on the isobaric analog
83: resonance (IAR) and the Gamow-Teller resonance (GTR). The IAR is the simplest
84: charge-exchange transition, in which a neutron is changed into a proton without
85: any other variation of the quantum numbers (that is, $\Delta J$=$\Delta L$=$\Delta S$=0).
86: The corresponding operator is
87: \begin{equation}
88: {\hat O}_{{\mbox{\upshape IAR}}} \equiv \sum_{i=1}^A t_-(i),
89: \label{op_iar}
90: \end{equation}
91: namely it is the usual Fermi, or isospin-lowering, operator. In the Gamow-Teller case,
92: the transition is accompanied by a spin-flip ($\Delta L$=0, $\Delta J$=$\Delta S$=1),
93: and the operator is
94: \begin{equation}
95: {\hat O}_{{\mbox {\upshape GTR}}} \equiv \sum_{i=1}^A \vec\sigma t_-(i).
96: \end{equation}
97: Many data coming at an early stage from ($p$,$n$), and later from other reactions, have
98: shown that these resonances can be systematically identified in the isotopes with neutron
99: excess (in which the corresponding $t_+$ transitions are Pauli-blocked). The IAR
100: consists of a single, very narrow peak, whereas the GTR manifests itself with a broad bump
101: and can also be fragmented in different peaks. Experimentally, when the incident
102: projectile energy is increased, the excitation of the GTR is favoured over the IAR; this
103: experimental fact has allowed to establish that the strengths of the spin-independent and
104: the spin-dependent
105: components of the effective {\itshape NN} interaction have different
106: behaviour as a function of
107: the energy.
108:
109: From this rather general introduction, the motivation for microscopic calculations
110: of the charge-exchange states in nuclei is already evident. We must
111: add that one of the main present interests in
112: nuclear physics is the understanding of the limits of nuclear stability, and of the
113: exotic, very neutron-rich (or proton-rich) nuclei, that is, of the systems with different values
114: of $N-Z$ than those which characterize the valley of stability.
115: The experimental evidences about the isospin properties
116: of exotic nuclei are still rather scarce. In order to make predictions in this delicate sector
117: accurate calculations are called for, which do not make approximations by neglecting
118: terms of the nuclear Hamiltonian in an uncontrolled way.
119:
120: For nuclei with mass up to $A\sim$50, the shell model (SM) calculations can be rather
121: successful and are indeed performed, also in the cases of interest for
122: applications. The agreement with the experimental findings (like the GT strength
123: and/or the $\beta$-decay half-life) can be quite good~\cite{sm}.
124: However, these calculations become too demanding, or impossible, for heavier nuclei.
125: Also, they have trouble if the space must be large enough so to account for high-energy
126: transitions; these transitions can be induced, for instance, by neutrinos which
127: follow a supernova explosion. In Ref.~\cite{volpe1} it has been shown that for energies
128: above 50 MeV the SM calculations may underestimate the strength of the charge-exchange
129: transitions.
130:
131: The alternative choice is a mean-field based calculation which employs an effective
132: {\itshape NN} interaction. In this case, the ground state of the parent ($N$,$Z$) system is obtained
133: by means of a Hartree-Fock (HF) calculation, extended to Hartree-Fock-Bardeen-Cooper-Schrieffer
134: (HF-BCS) or Hartree-Fock-Bogoliubov (HFB) in the case of open-shell nuclei where pairing
135: is relevant. In the two cases, respectively, the charge-exchange excited states
136: in the $(N\mp 1, Z\pm 1)$ isobars can be obtained within the framework of the linear
137: response theory, that is, by using the random phase approximation (RPA)
138: or its extension to the pairing case, namely the quasiparticle RPA (QRPA). These are
139: well-known theories, whose general features can be found in many textbooks. However,
140: there are only few examples, if any, of fully self-consistent QRPA calculations --- which
141: constitute the proper scheme for the analysis of long isotopic chains extending towards
142: the drip lines. In fact, self-consistency is a crucial issue if the calculations are
143: required to have predictive power far from the experimentally known regions of the mass
144: table. Moreover, as we discuss below, self-consistency plays a special role if the
145: isospin symmetry and its breaking enters the discussion. We repeat here that
146: self-consistency means that the residual particle-hole ({\itshape p-h}) and particle-particle ({\itshape p-p})
147: residual forces, which enter the QRPA equations (cf. Sec. II), are derived from the same
148: energy functional from which the HF-BCS of HFB equations describing the ground state
149: are obtained.
150:
151: The first attempt of self-consistent QRPA on top of HFB is found in Ref.~\cite{engel99}.
152: The Skyrme zero-range force and a zero-range pairing interaction are used, respectively,
153: in the mean-field and in the pairing channel to solve the HFB equations in coordinate
154: space (cf. also Ref.~\cite{hfb_doba}). The associated QRPA equations are solved in the canonical
155: basis. The method is applied to the calculation of Gamow-Teller $\beta$-decay half-lives.
156: These 1$^+$ states are known to be sensitive only to the $T$=0 component of the residual {\itshape p-p}
157: interaction, if pairing is described by means of a zero-range force. In
158: Ref.~\cite{engel99} it is assumed that, since this $T$=0 pairing does not manifest itself
159: in the HFB ground state of nuclei with $N$ different from $Z$ by a few units, one is
160: free to introduce it within QRPA in a completely different way than the $T$=1 pairing,
161: without any constraint related to self-consistency. The authors have employed a finite-range
162: interaction with free parameters: the overall strength is fitted to reproduce some selected
163: $\beta$-decay experimental findings. The same approach is used in Ref.~\cite{bender02}
164: to analyze the performance of existing Skyrme parametrizations in the case of the GT resonances,
165: and to correlate it with their ability to reproduce the values of
166: the empirical Landau parameters of infinite matter.
167:
168: In the present paper, we would like to discuss the implementation of a fully self-consistent
169: charge-exchange QRPA by putting emphasis on aspects which were not considered in
170: Refs.~\cite{engel99,bender02}. A first aspect is the issue of isospin invariance. We
171: show that the $T$=1 component of the residual {\itshape p-p} force can be fixed by exploiting
172: this invariance. Our hypothesis is supported by the absence of strong evidences coming from
173: literature which point to a clear
174: need to differentiate the strengths of the
175: three components of the $T$=1 pairing. Within this assumption, we show that we can
176: obtain results for the IAR which are quite satisfactory when compared with experiment.
177: The IAR is a serious benchmark for every theoretical model, because of its intimate
178: relationship with the isospin symmetry (cf., e.g., Ref.~\cite{auerbach_report}).
179: In fact, if the whole Hamiltonian $H$ commuted with
180: isospin, and if one were able to solve $H$ exactly, the resulting IAR would be degenerate
181: with the parent ground state. Many of the approximation schemes which are commonly used to
182: solve the nuclear many-body problem destroy this property of the Hamiltonian.
183: HF and HF-BCS belong to this category and introduce a spurious isospin breaking
184: (as soon as $N\neq Z$ in case of HF).
185: Instead, it has been demonstrated that self-consistent RPA and QRPA calculations
186: restore the isospin symmetry and eliminate any spuriousity~\cite{iso_restore}, being
187: in this sense ``good'' symmetry-preserving approximations. Consequently, only
188: within their framework it is possible to assess the relative importance of
189: the physical contributions which are responsible for an explicit isospin breaking: the Coulomb force,
190: the electromagnetic spin-orbit, and the other
191: charge-symmetry breaking (CSB) and charge-independence breaking (CIB) terms in the
192: nuclear Hamiltonian. The study of these issues in the case of the IAR for the
193: open-shell isotopes
194: is an original feature of the present work.
195: Since we do not go beyond QRPA, we cannot discuss
196: the (narrow) width of the IAR. The extensions of RPA and of the normal, non charge-exchange
197: QRPA, intended to take into account the coupling with more complex configuration and
198: therefore to describe the spreading width of the resonances, are described
199: elsewhere (see the references quoted in Sec. II).
200:
201: In our work, we employ zero-range forces. We are not aware of self-consistent calculations
202: of charge-exchange states done by using finite-range interactions like Gogny. On the other
203: hand, in recent papers the relativistic mean-field (RMF) effective Lagrangians, based on
204: the description of nucleons as Dirac particles which interact by means of the exchange
205: of effective mesons, have been used for the calculation of the IAR and the
206: GTR~\cite{paar0304}, as well as of $\beta$-decay rates~\cite{niksic05}. The RMF
207: description of the ground state and of the excited nuclear states emerges from
208: rather different ingredients than those which characterize the non-relativistic
209: mean-field. It is known that the isovector channel of the {\itshape NN} interaction, and
210: consequently the symmetry part of the energy functional, are quantitavely
211: not the same, generally speaking, in the two cases (the symmetry energy at saturation
212: and its derivatives are generally larger in the relativistic case). In the relativistic
213: calculations of the spin and isospin excitations the pion-exchange is very important; but
214: this degree of freedom is not present in the ground state description because of
215: parity conservation. On the other hand, in the case of RMF the spin-orbit is
216: automatically considered, at variance with the non-relativistic case. Finally, we
217: are not aware of attempts to include CSB and CIB forces in the RMF calculations.
218: All this should be kept in mind when comparing our results with those of~\cite{paar0304}.
219:
220: \section{THEORETICAL FRAMEWORK}
221:
222: As mentioned in the previous Section, charge-exchange RPA and QRPA are well-known and described
223: in textbooks. We try here to recall only the basic elements, or some details which are
224: useful for the following discussion.
225:
226: In the case of charge-exchange RPA, self-consistent calculations have been available
227: for many years. In particular, the first application to the case of the IAR can be found in
228: Ref.~\cite{auerbach_proc}. Extensive calculations of the response to different
229: multipole operators, made by using the coordinate
230: space formulation of RPA with proper treatment of the particle continuum, but dropping
231: for simplicity some terms of the residual interaction, are reported in~\cite{auerbach}.
232: As we have recalled in the Introduction,
233: it is well-known that mean-field calculations of this kind cannot reproduce the total width
234: of the resonances, but only the escape width if the continuum is correctly taken into
235: account. The spreading width, associated with the coupling of the simple {\itshape p-h} configurations
236: to the more complex states, of two particle-two hole (2$p$-2$h$) character, can be described
237: only by diagonalizing the effective Hamiltonian in a larger model space than the one of
238: RPA. A microscopic model suited for this purpose has been developed in~\cite{adachi,colo94}.
239: In~\cite{colo98} the importance of CSB and CIB forces for the IAR width has been studied.
240:
241: In the case of the QRPA, most of the charge-exchange calculations performed so far make use
242: of two separable {\itshape p-h} and {\itshape p-p} residual interactions (having, as a rule, the same functional
243: form and two different overall parameters $g_{ph}$ and $g_{pp}$),
244: as in the pioneering work by J.~A. Halbleib and
245: R.~A. Sorensen~\cite{halbleib}, where the formalism has been developed for the first time.
246:
247: We start by solving the HF-BCS equations in coordinate space by using a radial mesh
248: extending up to 20 fm (with a step of 0.1 fm). The
249: HF equations contain the Skyrme {\itshape NN} interaction and we have chosen in this work the
250: parametrization SLy4~\cite{chabanat}, which has been determined by trying to retain many of the
251: advantageous features of the previous versions of the Skyrme force, as well as by fitting
252: the equation of state of pure neutron matter obtained by means of realistic
253: forces. This latter characteristic should justify its use for systems outside
254: the valley of stability. The BCS equations are solved, as usual, in a limited space:
255: only the levels which correspond to the 82--126 neutron shell are included.
256: The pairing force that we have used is of the type
257: \begin{equation}
258: V=V_0 \left( 1- \left( {\varrho \left( {\vec r_1+\vec r_2\over 2} \right)
259: \over \varrho_c} \right)^\gamma \right)\cdot\delta(\vec r_1-\vec r_2).
260: \label{ppforce}
261: \end{equation}
262: The parameter $\gamma$ is fixed to one for the sake of simplicity. With the same spirit,
263: $\varrho_c$ is set at 0.16 fm$^{-3}$. The strength $V_0$ has been determined by requiring
264: a reasonably good agreement between the theoretical and empirical values of the pairing
265: gaps $\Delta$ along the whole series of isotopes under study. This agreement, when $V_0$ is
266: equal to our adopted value of 680 MeV~fm$^{3}$, is shown in Fig.\ \ref{average_delta}.
267: We notice in this context that a rather similar pairing force, having
268: $V_0$=625 MeV~fm$^{3}$, has been used independently
269: by other groups to carry out large-scale, systematic calculations of the pairing gaps
270: and of the rotational bands (see~\cite{duguet} and references therein).
271: It is known that the HFB treatment is more coherent than the HF-BCS one; however, qualitatively
272: important differences between the results of the two methods show up only in the case of
273: weakly bound nuclei, which will not be considered in the present study.
274:
275: When the ground state is obtained, together with the filled or partially occupied
276: states lying within the pairing window, a number of unoccupied states (which have occupation
277: factors $v^2$ strictly equal to zero) are calculated by using spherical box boundary conditions.
278: This means that our continuum is discretized. For
279: every value of $(l,j)$, we calculate unoccupied states with six increasing values of $n$.
280: The dimension of the space has been checked by looking at the results for the energy and
281: the strength of the IAR, which have been found to be stable when we enlarge the space,
282: by considering in some cases up to ten increasing values of $n$. We have checked that
283: also the $N-Z$ sum rule is accurately reproduced. In this configuration space, the QRPA matrix equation
284: written on the basis made up with the two quasiparticle states having good angular momentum
285: and parity $J^\pi$, reads
286: \begin{equation}
287: \left( \begin{array}{cc} A & B \\ -B & -A \end{array} \right)
288: \left( \begin{array}{c} X^{(n)} \\ Y^{(n)} \end{array} \right) = E_n
289: \left( \begin{array}{c} X^{(n)} \\ Y^{(n)} \end{array} \right).
290: \end{equation}
291: In this formula, $E_n$ is the energy of the $n$-th QRPA state in the parent nucleus,
292: while $X^{(n)}$, $Y^{(n)}$ are the corresponding forward- and backward-amplitudes.
293: The matrices $A$ and $B$, in the angular momentum coupled representation, have the explicit form
294: \begin{eqnarray}
295: A_{pn,p'n'} & = & (E_p+E_n)\delta_{pp'}\delta_{nn'} + \nonumber \\
296: & & + V^{(J)}_{pnp'n'} (u_{p}u_{n}u_{p'}u_{n'}+v_{p}v_{n}v_{p'}v_{n'}) + \nonumber \\
297: & & + W^{(J)}_{pnp'n'} (u_{p}v_{n}u_{p'}v_{n'}+v_{p}u_{n}v_{p'}u_{n'}), \nonumber \\
298: B_{pn,p'n'} & = & - V^{(J)}_{pnp'n'} (u_{p}u_{n}v_{p'}v_{n'}+v_{p}v_{n}u_{p'}u_{n'}) + \nonumber \\
299: & & + W^{(J)}_{pnp'n'} (u_{p}v_{n}u_{p'}v_{n'}+v_{p}u_{n}v_{p'}u_{n'}).
300: \end{eqnarray}
301: Here, the indices $p$ and $p'$ ($n$ and $n'$) refer to proton (neutron) quasiparticles.
302: $E$ is their energy, whereas $u$ and $v$ are the usual BCS occupation factors. $V^{(J)}$ and
303: $W^{(J)}$ indicate respectively the coupled
304: {\itshape p-p} and {\itshape p-h} matrix elements. The {\itshape p-h} matrix elements are derived from
305: the Skyrme part of the energy functional: all the terms are considered, including the two-body
306: spin-orbit.
307:
308: The {\itshape p-p} matrix elements, when consistently derived from the energy functional, are
309: those of the bare force (\ref{ppforce}): in fact, no rearrangement terms show up if we do
310: not impose any dependence on the anomalous density in the force itself. The isospin invariance
311: that we have assumed,
312: demands that the $T$=1 component of the pairing force is the same in the three channels:
313: neutron-neutron, proton-neutron and proton-proton. In the present case, since we
314: deal with the Sn isotopes which have magic proton number, there is no proton pairing in the
315: ground state. Also, we have neglected proton-neutron
316: pairing in the ground state: in fact, this may be important
317: only in nuclei having $N\sim Z$ and we
318: have considered Sn isotopes in the
319: range 104$\le A \le$132. However, the proton-neutron $T$=1 pairing force enters the
320: QRPA equations (in the $V$ matrix elements) and we can say that we have
321: preserved the self-consistency in the pairing channel, in the same way as in the {\itshape p-h}
322: one.
323:
324: The CSB and CIB forces are included in our HF-BCS iterative procedure. These forces are
325: parametrized as in Ref.~\cite{sagawa}, where they have been cast in a form similar to
326: that of the Skyrme interaction. They had been already employed, under the form of a Yukawa
327: function, in Ref.~\cite{suzuki} and they have been shown to reproduce well the correct
328: mass number dependence of the Coulomb displacement energies, as well as
329: a number of values of isospin mixing in the ground state. Finally, they turned out to be important
330: to account for the IAR width in $^{208}$Pb ~\cite{colo98}. For all these reasons, we use these
331: parametrizations in the present work. Because of their
332: operatorial form, they do not add any contribution to the {\itshape p-h} force in
333: RPA or QRPA. The electromagnetic spin-orbit is quite small: consequently,
334: the associated energy shift has been added
335: to the HF-BCS results using first-order perturbation theory.
336:
337: \section{RESULTS}
338:
339: The systematic trend of the IAR energies in the nuclei we have considered, $^{104\mbox{--}132}$Sn,
340: is plotted in Fig.\ \ref{syst_trend}. The energies are obtained within QRPA, by including all the terms
341: mentioned in the previous Section: only the proton-rich
342: $^{104,106}$Sn have been calculated using quasiparticle Tamm-Dancoff approximation (QTDA)
343: because of QRPA instabilities.
344: Our findings are compared with the experimental energies quoted in
345: Ref.~\cite{pham95}, where the results of the ($^3$He,$t$) reaction performed at
346: an incident beam energy of 200 MeV are reported. It can be immediately
347: realized that the agreement is fairly good. The difference between theory and experiment
348: is typically $\approx$ 200 keV in the series of isotopes which have been measured,
349: namely $^{112\mbox{--}124}$Sn (with the exception of the two extremes $^{112}$Sn and $^{124}$Sn
350: where this difference is larger).
351: It is remarkable that another microscopic, self-consistent model like RMF --- which starts
352: from a quite different description of the nuclear mean-field and its oscillations as
353: already stressed in the Introduction --- produces a similar numerical outcome~\cite{paar0304}.
354: The results for the IAR in the unstable nuclei do constitute a useful guideline
355: for possible future experiments.
356:
357: Concerning the results in the ($N+1$,$Z-1$) channel, unfortunately
358: few experimental measurements are available
359: for a comparision with our model. The only exception is the case of $^{120}$Sn.
360: In Fig.\ \ref{In_channel} we plot for this nucleus the response to the IVGMR operator,
361: \begin{equation}
362: {\hat O}_{{\mbox {\upshape IVGMR}}} \equiv \sum_{i=1}^A r^2_i t_+(i),
363: \label{op_ivgmr}
364: \end{equation}
365: as a function of the energy with respect to the ground state of $^{120}$In.
366: The continuous curve has been obtained by averaging the QRPA discrete strength
367: distribution with a 1 MeV width Lorentzian curve.
368: We can compare our results with three experiments carried out by means of
369: different nuclear reactions. By using ($\pi^-,\pi^0$) at 165
370: MeV~\cite{erell}, ($^{13}$C, $^{13}$N) at 50 MeV/A~\cite{berat} and
371: ($^{7}$Li, $^{7}$Be) at 350 MeV~\cite{annakkage} it has been shown, more or less
372: ambiguously, that a 0$^+$ state should lie, respectively, at 16.0$\pm$2.2 MeV, 14.7$\pm$1 MeV and
373: 17.0$\pm$1.6 MeV. In our calculation most of the strength is found indeed
374: in the energy region between 12 and 20 MeV. Our main peak seems
375: compatible with the ($^{7}$Li, $^{7}$Be) result.
376:
377: Coming back to the case of the IAR, we analyze in more detail our results
378: in order to clarify the most important features of our theoretical description.
379: Firstly, in analogy with the conclusion drawn in Ref.~\cite{paar0304}, we may show that
380: also in the present case the consistent treatment of pairing correlations is
381: very important. In Fig.\ \ref{pn} we display three different results obtained
382: for the IAR strength distribution in $^{114}$Sn. Not only the residual proton-neutron
383: pairing force plays a crucial role to concentrate the IAR in a single peak; it
384: also affects the IAR energy in an important way, that is, it induces a downward
385: shift of about 500 keV. In the whole isotopic
386: series we have studied, the peak associated with the IAR exhausts typically
387: a percentage between 95\% and 98\% of the $N-Z$ sum rule.
388: Only in the isotope $^{108}$Sn the IAR is found to be split in
389: two peaks.
390:
391: Having assessed the importance of the proton-neutron residual pairing, we have also
392: tried to understand the role played by various other correlations present in
393: our model. For this purpose, we display in Fig.\ \ref{contrib} results for the
394: IAR energy in $^{120}$Sn obtained by making different approximations.
395: The first number on the left side refers to a simple TDA calculation, without any
396: pairing, without the spin-orbit term in the residual {\itshape p-h} force, and without
397: the other terms which have been often neglected (electromagnetic
398: spin-orbit, CSB and CIB). This would be the simplest possible calculation,
399: analogous to that performed for many closed-shell nuclei in the previous literature.
400: The inclusion of RPA ground state correlations do not affect very much the
401: IAR, as it is expected for a nucleus which has neutron excess; the effect is
402: larger if we move towards the neutron-deficient isotopes. Pairing correlations
403: are more important. We have discussed above that they have to be included
404: consistently (we repeat that a calculation with pairing only in the ground state would lead
405: to a too high, and fragmented, IAR): moreover, their inclusion shifts the IAR
406: downwards by about 150 keV. At this stage, the QRPA result would differ from
407: the experimental finding by about 500 keV. This would be approximately true for all the
408: stable isotopes. The two-body spin-orbit have a
409: non negligible effect (about 100 keV) in pushing the IAR energy towards
410: the experimental value. Even more important, from a quantitative point of view, are in this case
411: the CSB and CIB forces which are inserted in the HF-BCS calculation (the
412: fact that they have opposite sign has been already remarked~\cite{auerbach_report}).
413: Finally, we have included for the sake of
414: completeness the one-body electromagnetic spin-orbit. This term has also been
415: calculated long time ago (see, e.g, p. 494 of Ref.~\cite{nolen}) and
416: it is known to have, as a rule, an effect of only few tenths of keV on the Coulomb
417: displacement energies. Because of its $j$-dependence, it may become
418: significant in the case of pure transitions associated with large angular
419: momentum, as it has been stressed in~\cite{ekman04}. We should add that
420: we have checked that the contributions stemming from the CSB, CIB and the
421: electromagnetic spin-orbit are almost constant over the isotopic chain. In this
422: sense, the numbers presented in Fig.~\ref{contrib} are considered as
423: typical. As far as the two-body spin-orbit is concerned, in the middle of
424: the chain the associated repulsive contribution is maximum; at the extremes of the
425: chain it becomes smaller or even attractive (for instance, in $^{132}$Sn we
426: find an attractive contribution associated with the diagonal $h_{11/2}$ matrix
427: element).
428:
429: Since many Skyrme parametrizations are available in the market, we would like
430: to mention that our results are not very sensitive to the choice of a specific
431: set. In fact, we have seen that the IAR energy of $^{120}$Sn varies by
432: less than 100 keV if we calculate it either using the force SLy4, or
433: SIII~\cite{beiner} or SGII~\cite{sgii}. We have also performed a
434: calculation using the recently introduced SkO' interaction~\cite{skop}, in view
435: of the possibility of testing it in the next future on the systematics of spin
436: states. In this case, the variation of the energy, with respect to the result
437: obtained by using SLy4, is somewhat larger~\cite{note_sc}. Also the
438: effect of varying the pairing strength $V_0$ has been
439: considered, and we refer to Fig.\ \ref{pairing_effect} for the results obtained
440: in the case of $^{116}$Sn (i.e., the isotope in the middle of the 50--82 neutron shell).
441: We can consider as satisfactory that variations of $V_0$ in the
442: range $\approx$ 650--710 MeV~fm$^{3}$, which
443: lead to sizeable ($\approx$20\%) variations of $\Delta$, do not seriously affect
444: the energy of the IAR. We can quite generally conclude that
445: the choice of parameters, both of our {\itshape p-h} and {\itshape p-p} forces,
446: do not seriously affect our main conclusions on IAR.
447:
448: \section{CONCLUSIONS}
449:
450: Very few examples of microscopic, fully self-consistent charge-exchange
451: QRPA calculations exist (in contrast with the non charge-exchange case).
452: This has motivated the present work, in which we have developed the method
453: and analyzed some specific issues: the relation between the isospin invariance
454: and the self-consistency in the pairing channel, and the role of the
455: usually neglected contributions in the mean-field. We have applied our scheme
456: for the calculation of the IAR along the chain of the Sn isotopes. Only
457: calculations based on RMF are available for this case. We find that our
458: non-relativistic model can account quite well for the experimental results.
459:
460: We plan to extend our calculations, and make further analysis of the charge-exchange
461: states. This will be done for different multipolarities, both in the non spin-flip and
462: spin-flip sectors. It is hoped that the comparison with experimental data, and
463: with the outcome of other microscopic models, can be instrumental to fix
464: rather general problems. In fact, as stressed in our Introduction,
465: many uncertainties plague the isovector channel of the effective {\itshape NN}
466: interaction, and consequently the symmetry part of the nuclear equation of
467: state.
468:
469: A possible improvement of our model consists in changing the description of
470: the nuclear ground state, which may be calculated within full HFB instead
471: of HF-BCS. This could allow a better description in the case, for instance,
472: of weakly bound systems. Another open problem is the consideration of the
473: role played by the proton-neutron pairing. Literature
474: reflects the existence of many different thoughts about this
475: interesting issue; a full microscopic QRPA
476: calculation in the case in which the particles do not have a definite
477: charge state may probably be at present too demanding. Finally,
478: we should mention that the extension beyond mean-field of theories
479: like ours remains to be done.
480:
481: \begin{thebibliography}{9}
482:
483: \bibitem{Harakeh} M.~N. Harakeh and A.~M. Van Der Woude, {\em Giant Resonances: Fundamental
484: High-Frequency Modes of Nuclear Excitations} (Oxford University Press, 2001);
485: P.~F. Bortignon, A. Bracco, R.~A. Broglia, {\em Giant Resonances. Nuclear Structure at
486: Finite Temperature} (Harwood Academic, New York 1998).
487:
488: \bibitem{IVGMR} T. Ichihara, M. Ishihara, H. Ohnuma, T. Niizeki, Y. Satou, H. Okamura,
489: S. Kubono, M.~H. Tanaka, and Y. Fuchi, Phys. Rev. Lett. {\bf 89} (2002) 142501 and
490: references therein; J. Guillot {\em et al.}, submitted to Phys. Rev. C.
491:
492: \bibitem{sm} P.~B. Radha, D.~J. Dean, S.~E. Koonin, K. Langanke, and P. Vogel,
493: Phys. Rev. {\bf C56} (1997) 3079; E. Caurier, K. Langanke, G. Martinez-Pinedo, and
494: F. Nowacki, Nucl. Phys. {\bf A653} (1999) 439.
495:
496: \bibitem{volpe1} C. Volpe, N. Auerbach, G. Col\`o, T. Suzuki, and N. Van Giai, Phys. Rev.
497: {\bf C62} (2000) 015501.
498:
499: \bibitem{engel99} J. Engel, M. Bender, J. Dobaczewski, W. Nazarewicz, and R. Surman,
500: Phys. Rev. {\bf C60} (1999) 014302.
501:
502: \bibitem{hfb_doba} J. Dobaczewski, H. Flocard, and J. Treiner, Nucl. Phys. {\bf A422}
503: (1984) 103.
504:
505: \bibitem{bender02} M. Bender, J. Dobaczewski, J. Engel, and W. Nazarewicz, Phys. Rev.
506: {\bf C65} (2002) 054322.
507:
508: \bibitem{auerbach_report} N. Auerbach, Phys. Rep. {\bf 98} (1983) 273.
509:
510: \bibitem{iso_restore} C.~A. Engelbrecht and R.~H. Lemmer, Phys. Rev. Lett.
511: {\bf 24} (1970) 607; A.~M. Lane and J. Martorell, Ann. Phys. (N.Y.) {\bf 129}
512: (1980) 273.
513:
514: \bibitem{paar0304} D. Vretenar, N. Paar, T. Nik\v{s}i\`{c}, and P. Ring, Phys. Rev.
515: Lett. {\bf 91} (2003) 262502; N. Paar, T. Nik\v{s}i\`{c}, D. Vretenar, and P. Ring, Phys. Rev.
516: {\bf C69} (2004) 054303.
517:
518: \bibitem{niksic05} T. Nik\v{s}i\`{c}, T. Marketin, D. Vretenar, N. Paar, and P. Ring, Phys. Rev.
519: {\bf C71} (2005) 014308.
520:
521: \bibitem{auerbach_proc} N. Auerbach, N. Van Giai, and A. Yeverechyahu, in {\em Highly
522: Excited States in Nuclear Reactions}, edited by H. Ikegami and M. Muraoka (RCNP,
523: Osaka, 1980), p. 623.
524:
525: \bibitem{auerbach} N. Auerbach and A. Klein, Nucl. Phys. {\bf A395} (1983) 77.
526:
527: \bibitem{adachi} S. Adachi and S. Yoshida, Nucl. Phys. {\bf A462} (1987) 61.
528:
529: \bibitem{colo94} G. Col\`o, N. Van Giai, P.F. Bortignon, and R.A. Broglia,
530: Phys. Rev. {\bf C50} (1994) 1496.
531:
532: \bibitem{colo98} G. Col\`o, H. Sagawa, N. Van Giai, P.~F. Bortignon, and T. Suzuki,
533: Phys. Rev. {\bf C57} (1998) 3049.
534:
535: \bibitem{halbleib} J.~A. Halbleib and R.~A. Sorensen, Nucl. Phys. {\bf A98} (1967) 542.
536:
537: \bibitem{chabanat} E. Chabanat, P. Bonche, P. Haensel, J. Meyer, and R. Schaeffer,
538: Nucl. Phys. {\bf A627} (1997) 710.
539:
540: \bibitem{duguet} T. Duguet, P. Bonche, P.-H. Heenen, and J. Meyer,
541: Phys. Rev. {\bf C65} (2002) 014311.
542:
543: \bibitem{sagawa} H. Sagawa, N. Van Giai, and T. Suzuki, Phys. Lett.
544: {\bf B353} (1995) 7
545:
546: \bibitem{suzuki} T. Suzuki, H. Sagawa, and N. Van Giai, Phys. Rev. {\bf C47} (1993) R1360
547:
548: \bibitem{pham95} K. Pham, J. J\"{a}necke, D.~A. Roberts, M.~N. Harakeh,
549: G.~P.~A. Berg, S. Chang, J. Liu, J. Stephenson, B.~F. Davis,
550: H. Akimune, and M. Fujiwara, Phys. Rev. {\bf C51} (1995) 526.
551:
552: \bibitem{erell} A. Erell, J. Alster, J. Lichtenstadt, M.~A. Moinester,
553: J.~D. Bowman, M.~D. Cooper, F. Irom, H.~S. Matis, E. Piasetzky, and
554: U. Sennhauser, Phys. Rev. {\bf C34} (1986) 1822.
555:
556: \bibitem{berat} C. B\'erat, M. Buenerd, J.~Y. Hostachy, P. Martin, J. Barrette,
557: B. Berthier, B. Fernandez, A. Miczaika, A. Villari, H.~G. Bohlen,
558: S. Kubono, E. Stiliaris, W. von Oertzen, Nucl. Phys. {\bf A555} (1993) 455.
559:
560: \bibitem{annakkage} T. Annakkage, J. J\"{a}necke, J.~S. Winfield,
561: G.~P.~A.Berg, J.~A. Brown, G. Crawley, S. Danczyk, M. Fujiwara,
562: D.~J. Mercer, K. Pham, D.~A. Roberts, J. Stasko, G.~H. Yoo, Nucl. Phys. {\bf A648} (1999) 3.
563:
564:
565:
566: \bibitem{nolen} J.~A. Nolen and J.~P. Schiffer, Annu. Rev. Nucl. Sci.
567: {\bf 19} (1969) 471.
568:
569: \bibitem{ekman04} J. Ekman, D. Rudolph, C. Fahlander, A.~P. Zuker,
570: M.~A. Bentley, S.~M. Lenzi, C. Andreoiu, M. Axiotis, G. de Angelis,
571: E. Farnea, A. Gadea, Th Kr\"{o}ll, N. M\v{a}rginean, T. Martinez, M.~N. Mineva,
572: C. Rossi Alvarez, C.~A. Ur, Phys. Rev. Lett. {\bf 92} (2004) 132502.
573:
574: \bibitem{beiner} M. Beiner, H. Flocard, N. Van Giai and Ph. Quentin,
575: Nucl. Phys. {\bf A238} (1975) 29.
576:
577: \bibitem{sgii} N. Van Giai and H. Sagawa, Phys. Lett. {\bf B106} (1981) 379.
578:
579: \bibitem{skop} P.-G. Reinhard, D.J. Dean, W. Nazarewicz, J. Dobaczewski, J.A. Maruhn
580: and M.R. Strayer, Phys. Rev. {\bf C60} (1999) 014316.
581:
582: \bibitem{note_sc} However, we must stress that this is the only
583: calculation of our work which is characterized by some (arguably
584: small) self-consistency breaking. In fact, the calculation of the ground
585: state includes the spin-gradient terms (the so-called
586: terms in $J^2$, $\vec J$ being the spin density) while the corresponding {\itshape p-h} residual
587: force is omitted.
588:
589: \end{thebibliography}
590:
591: \newpage
592:
593: % 1
594: \begin{figure}
595: \caption{The values of the pairing gaps $\Delta$ in the Sn isotopes. The open squares
596: correspond to the empirical values, extracted by attributing to the isotope with $N$ neutrons
597: the value which results from the three-point formula centered in $N+1$.
598: The black squares correspond to the theoretical results: in this case, the
599: values of the state-dependent HF-BCS pairing gaps $\Delta_i$ are averaged in an energy
600: interval centered at the neutron Fermi energy and having a width of $\pm$5 MeV.}
601: \label{average_delta}
602: \end{figure}
603:
604: % 2
605: \begin{figure}
606: \caption{Systematic trend of the IAR energies in the stable and unstable Sn isotopes. The
607: theoretical results, displayed by means of black circles, are compared with the experimental
608: data (open squares) whenever these are available. It must be noticed that the energies
609: are referred to the daughter nuclei.}
610: \label{syst_trend}
611: \end{figure}
612:
613: % 3
614: \begin{figure}
615: \caption{Strength function associated with the IVGMR operator (\ref{op_ivgmr}) in
616: $^{120}$In. The discrete QRPA peaks have been smoothed by using a Lorentzian averaging
617: (the Lorentzian width is 1 MeV). See the text for a comparison with the available
618: experimental results.}
619: \label{In_channel}
620: \end{figure}
621:
622: % 4
623: \begin{figure}
624: \caption{Importance of the residual proton-neutron {\itshape p-p} interaction for the collectivity
625: of the IAR. The left, central and right panels refer respectively to RPA, QRPA without
626: that term in the residual force, and complete QRPA. The result is analogous to the one
627: shown in Fig.\ 5 of Ref.~\cite{paar0304}.}
628: \label{pn}
629: \end{figure}
630:
631: % 5
632: \begin{figure}
633: \caption{Result for the IAR energy in $^{120}$Sn obtained using different approximations.
634: The values labelled by $\Delta$ represent the energy shifts of the IAR (in keV) at each step.
635: See the text for a detailed discussion.}
636: \label{contrib}
637: \end{figure}
638:
639: % 6
640: \begin{figure}
641: \caption{Effect of the overall pairing strength $V_0$ which defines the effective
642: force (\ref{ppforce}) on the pairing gap (upper panel) and the IAR energy (lower
643: panel) in $^{116}$Sn. The experimental values are marked by horizontal full lines,
644: whereas the vertical dashed line indicates the adopted value of $V_0$.}
645: \label{pairing_effect}
646: \end{figure}
647:
648: \end{document}
649: