nucl-th0411069/fsi.tex
1: % Template article for preprint document class `elsart'
2: % SP 2001/01/05   
3: 
4: \documentclass{elsart}
5: 
6: % Use the option doublespacing or reviewcopy to obtain double line spacing
7: % \documentclass[doublespacing]{elsart}
8: 
9: % if you use PostScript figures in your article
10: % use the graphics package for simple commands
11: % \usepackage{graphics}
12: % or use the graphicx package for more complicated commands
13:  \usepackage{graphicx}
14: % or use the epsfig package if you prefer to use the old commands
15: %\usepackage{epsfig}
16: 
17: % The amssymb package provides various useful mathematical symbols
18: \usepackage{amssymb}
19: 
20: \begin{document}
21: 
22: \begin{frontmatter}
23: 
24: % Title, authors and addresses 
25: 
26: % use the thanksref command within \title, \author or \address for footnotes;
27: % use the corauthref command within \author for corresponding author footnotes;
28: % use the ead command for the email address,
29: % and the form \ead[url] for the home page:
30: % \title{Title\thanksref{label1}}
31: % \thanks[label1]{}
32: % \author{Name\corauthref{cor1}\thanksref{label2}}
33: % \ead{email address}
34: % \ead[url]{home page}
35: % \thanks[label2]{}
36: % \corauth[cor1]{}
37: % \address{Address\thanksref{label3}}
38: % \thanks[label3]{}
39: 
40: %\title{Fragment-fragment final-state interaction in the Coulomb
41: %dissociation of exotic nuclei}
42: \title{
43: Electromagnetic strength of
44: neutron and proton single-particle halo nuclei}
45: 
46: % use optional labels to link authors explicitly to addresses:
47: % \author[label1,label2]{}
48: % \address[label1]{}
49: % \address[label2]{}
50: 
51: \author{S. Typel\thanksref{ST}}
52: \address{Gesellschaft f\"{u}r Schwerionenforschung mbH, Theorie,\\
53: Planckstr.~1, D-64291 Darmstadt, Germany}
54: \thanks[ST]{Corresponding author.}
55: % Fax: +49-6159-712990. Phone +49-6159-712794.}
56: \ead{S.Typel@gsi.de}
57: 
58: \author{G. Baur}
59: \address{Institut f\"{u}r Kernphysik,
60: Forschungszentrum J\"{u}lich,
61: D-52425 J\"{u}lich, Germany}
62: \ead{G.Baur@fz-juelich.de}
63: 
64: \begin{abstract}
65: Electromagnetic strength functions of halo nuclei exhibit 
66: universal features that can be described in terms of
67: characteristic scale parameters.
68: For a nucleus with nucleon+core structure
69: the reduced transition probability, as determined, e.g.,
70: by Coulomb dissociation experiments,
71: shows a typical shape that depends
72: on the nucleon separation energy and the orbital angular
73: momenta in the initial and final states.
74: The sensitivity to the final-state interaction (FSI) between the nucleon
75: and the core
76: can be studied systematically by varying the strength of the interaction
77: in the continuum.
78: In the case of neutron+core nuclei  
79: analytical results for the reduced transition
80: probabilities are obtained by introducing 
81: the effective-range expansion.
82: The scaling with the relevant parameters is found explicitly.
83: General trends are observed by studying
84: several examples of neutron+core and proton+core nuclei
85: in a single-particle model assuming Woods-Saxon potentials.
86: Many important features of the neutron halo case can
87: be obtained from a square-well model. Rather
88: simple analytical formulas are found.
89: The nucleon-core interaction in the continuum
90: affects the determination of astrophysical S factors at zero energy
91: in the method of asymptotic normalisation coefficients (ANC).
92: It is also relevant for the extrapolation of radiative capture
93: cross sections to low energies.
94: \end{abstract}
95: 
96: \begin{keyword}
97: % keywords here, in the form: keyword \sep keyword
98: halo nuclei \sep
99: electromagnetic transitions \sep 
100: effective-range approximation \sep
101: scaling laws \sep
102: final-state interaction \sep  
103: sum rules \sep
104: ANC method  \sep
105: radiative capture
106: % PACS codes here, in the form: \PACS code \sep code
107: \PACS 21.10.Pc \sep 21.10.Ky \sep 25.20.-x \sep 25.40.Lw \sep 27.20.+n
108: \end{keyword}
109: \end{frontmatter}
110: 
111: \newpage 
112: 
113: \section{Introduction}
114: 
115: Light exotic nuclei are available
116: as secondary beams at various 
117: radioactive beam facilities all over the world. These
118: unstable nuclei 
119: are generally  weakly bound with few, if any, bound excited states. 
120: They have been studied extensively
121: in recent years by electromagnetic excitation with the help of the
122: Coulomb breakup method %\cite{Bau86,Ber88,Bau94,Bau03}. 
123: \cite{Bau86,Ber88,Lei01,Bau03,Aum04}. 
124: For low orbital angular momenta 
125: of the lowest bound valence nucleon an extended diffuse 
126: density distribution, a halo, develops 
127: resulting in a large size of the nucleus
128: \cite{Han87,Rii92,Rii94,Han95,Tan96}.
129: Simultaneously, electromagnetic transitions to the continuum 
130: with large strength are observed at low energies.
131: Properties of stable nuclei have successfully been investigated
132: by electromagnetic excitation
133: in photonuclear reactions as well as 
134: in heavy ion collisions for a long time.
135: Their electromagnetic strength functions are dominated by
136: the giant resonances high in the continuum.
137: 
138: Nuclei close to the neutron and proton drip lines
139: often exhibit a pronounced nucleon+core structure
140: that is well described by single-particle models with appropriately
141: chosen potentials.
142: In a microscopic shell-model study \cite{Sag01} 
143: strong low-lying dipole strength in neutron-rich ${}^{14}$Be 
144: and proton-rich ${}^{13}$O
145: was observed to originate from loosely-bound extended 
146: single-particle wave functions. 
147: These extended wave functions were obtained by 
148: adjusting the potential depth in order to reproduce 
149: the empirical binding energies. In order to study low-lying strength
150: theoretically, such a feature has to be added to the ab-initio 
151: microscopic approaches. 
152: A comparison of  strength functions deduced from experiment
153: with theoretical predictions from single-particle models
154: is used to extract spectroscopic factors or asymptotic normalization
155: coefficients (ANC). They can be compared to
156: more elaborated nuclear models. However, the importance of 
157: effects from the interaction between the nuclei in the final state
158: has to be assessed in order to obtain reliable
159: information from experimental data. 
160: Both the shape and the absolute magnitude of the strength function
161: can be affected. This will also have consequences for the application
162: of sum rules that relate the total excitation strength to the
163: properties of the ground state. 
164: 
165: The relevant matrix elements for the electromagnetic transition
166: to the continuum at low energies are esssentially determined
167: by the asymptotics of the bound-state and continuum wave functions
168: of halo nuclei. This allows to study systematically the effects of the 
169: final-state interaction (FSI)
170: without the necessity to introduce sophisticated nuclear structure
171: models. Magnetic contributions to the continuum transitions 
172: are usually much weaker than electric transitions, 
173: except for the deuteron breakup at low
174: energies \cite{Nag97}
175: and for the excitation of resonances. 
176: We will limit ourselves
177: to the discussion of direct electric transitions to the continuum,
178: with emphasis on $E1$ transitions,
179: but the search for low-lying $M1$ strength and 
180: its theoretical description \cite{Bla79} remains an interesting
181: challenge for future studies. 
182: 
183: The nuclear interaction $V_{bc}$
184: between the nucleon $b$ and the core $c$ in the electromagnetic breakup
185: of an exotic nucleus $a$ is responsible
186: for the binding of the nucleus $a$. However, it also affects
187: the structure of the continuum 
188: (usually in partial waves with different $l$ values)
189: even if there are no resonances
190: observed at low excitation energies. 
191: In the experimental analysis of neutron+core
192: breakup reactions it is often neglected and a plane wave is assumed
193: in the final state of the $b+c$ system.
194: The interaction between $b$ and $c$  also appears in the final state of
195: the photo-dissociation reaction $a(\gamma,c)b$ or in the
196: initial state of the radiative capture reaction $b(c,\gamma)a$.
197: Thus, the interaction $V_{bc}$ can affect the 
198: energy dependence of the astrophysical S factor of the radiative
199: capture reaction  that is used to extrapolate
200: experimental data to zero energy. Similarly, the strength of the
201: interaction enters into the
202: calculation of the zero-energy S factor from asymptotic
203: normalization coefficients (ANCs) that are determined experimentally
204: from transfer reactions in the ANC method, see, e.g., 
205: \cite{Xu94,Tra01,Muk01,Cre02}.
206: 
207: Experimentally observed excitation functions of exotic nuclei
208: show an approximately universal shape 
209: when plotted as a function of appropriately scaled variables
210: \cite{Ots94,Men95,Kal96,Typ01a}.
211: They are dominated by direct transitions
212: to the continuum with only few resonances
213: and simple scaling laws apply.
214: The nuclear structure in the initial and final state 
215: depends only on a limited number of relevant quantities
216: that contain all the structural information that is accessible
217: at low energies. Details of the nuclear
218: interaction are not resolved at this low energy scale.
219: The interaction between the fragments leads to
220: a change of the transition strength when compared to the
221: case without nuclear interaction.
222: The actual nuclear potential is often not 
223: %ST - haengt vom Sinn ab
224: well constrained since extrapolations of the corresponding 
225: systematic optical potentials, e.g.\ \cite{Per76},
226: from nuclei in the valley of stability 
227: to unstable nuclei are questionable.
228: The effect of the continuum interaction
229: was studied before only in selected cases. 
230: E.g., it was found that the $s$-wave ground state to $p$-wave continuum
231: $E1$ transition in the case of ${}^{11}$Be 
232: is much less affected by the potential in the final state
233: than the $p \to s$ transition in ${}^{13}$C \cite{Ots94,Men95}.
234: 
235: At low energies the effect of the nuclear potential is 
236: conveniently described by the effective-range expansion
237: \cite{Bet49,New82},
238: with the scattering length and the effective range as the main 
239: parameters.  
240: An effective-range approach for the FSI in electromagnetic
241: excitations was introduced in \cite{Typ04a} and
242: applied to the breakup of ${}^{11}$Be.
243: Recently, the same method was applied to the description of
244: electromagnetic dipole strength in ${}^{23}$O
245: \cite{Noc04}.
246: Here, we will discuss this approach in much more detail.
247: A systematic study for various transitions will shed some additional
248: light on the sensitivity to the interaction in the continuum.
249: We want to expose the dependence on the binding energy of the nucleon
250: and on the angular
251: momentum quantum numbers. Our approach extends the familiar textbook
252: case of the deuteron \cite{Bla79}, that can be considered as the
253: prime example of a halo nucleus, to arbitrary nucleon+core systems.
254: 
255: Our effective-range approach is closely related to effective field theories
256: that are nowadays used for the description of 
257: the nucleon-nucleon system and halo nuclei
258: \cite{Bir99,Kap99,Ber02}. 
259: (The deuteron, the only bound state in the 
260: nucleon-nucleon system, can be considered as
261: a good example of a  halo nucleus).
262: The characteristic low-energy
263: parameters are linked to QCD in systematic expansions.
264: Similar methods are also used in 
265: the study of exotic atoms ($\pi^{-} A$, $\pi^{+} \pi^{-}$, 
266: $\pi^{-}p$, \dots) 
267: in terms of effective-range parameters in Ref.\ \cite{Eri03}.
268: The close relation of effective field theory to the effective-range
269: approach for hadronic atoms was discussed in Ref.\ \cite{Hol99}.
270: In our approach these constants are treated as free parameters.
271: It is not our aim to relate them to the underlying microscopic
272: description. Aspects of the many-body physics are summarized, e.g., 
273: in terms of
274: spectroscopic factors or asymptotic normalization coefficients.
275: 
276: We also investigate in detail a specific model, the square well potential.
277: It has  great merits: it can be solved analytically,
278: it shows the main characteristic features
279: and it leads to rather simple and transparent formulas
280: %hallo
281: where some of them seem to be new.
282: These explicit results can be compared to our general
283: results for low energies (effective-range approach) and also 
284: to Woods-Saxon models. 
285: 
286: This paper is organized as follows.
287: A nucleon+core potential model for halo nuclei is introduced 
288: in section~\ref{sec:ncm}. The relevant scaling parameters are defined 
289: in subsection \ref{subsec:scale} and
290: scaling laws for the 
291: the probabilities to find the nucleon inside the range of the nuclear
292: potential are discussed for bound and scattering states
293: in subsection \ref{subsec:prob}. The
294: scaling of the root-mean-square radius serves as another indication
295: for the halo nature of the bound state.
296: The reduced transition probabilities for the breakup of a nucleon+core
297: nucleus are calculated in
298: subsection \ref{subsec:trans} for electric transitions
299: with multipolarity $\lambda$. 
300: They only depend on the
301: asymptotic normalization of the bound state wave function and
302: on radial integrals with the asymptotic wave functions
303: since the radial integrals are dominated by the contribution from
304: outside the nuclear potential. In the square-well model explicit
305: expressions for the ratio of the interior to the exterior contributions
306: are derived.
307: An alternative calculation of $E1$ transition integrals with the
308: help of a commutator relation is presented in subsection \ref{subsec:comrel}.
309: The strength functions are related 
310: to cross sections of photo-nuclear reactions in
311: subsection \ref{subsec:xs} and the high-energy behaviour is discussed.
312: The effect of the nucleon-core potential is considered
313: in subsection \ref{subsec:scalen}. 
314: The effective-range expansion allows to 
315: parametrize the effects of the continuum interaction 
316: in a suitable way to study systematically its influence on the
317: transition strength.
318: In Section \ref{sec:rrisf} the reduced transition probabilities
319: are expressed in terms of characteristic shape functions that
320: depend on certain reduced integrals. The dependence of the
321: shape functions on the scaling parameters is studied in various
322: limits. Systematic variations of the shape functions are discussed
323: for neutron+core systems in subsections \ref{subsec:n+core}
324: and \ref{subsec:fsinc} without and with FSI, respectively,
325: where analytical results are obtained.
326: Proton+core systems are treated numerically
327: in subsection \ref{subsec:p+core}.
328: The relation of the total excitation strength to the 
329: properties of the bound state
330: with the help of sum rules
331: is considered in section \ref{sec:totsum}.
332: The findings of the model with asymptotic wave functions
333: are corroborated in more realistic
334: calculations using wave function generated from Woods-Saxon potentials
335: in section \ref{sec:examples}.
336: The implication
337: of the nucleon-core interaction in the continuum state
338: on the ANC method is discussed
339: in section \ref{sec:anc}.
340: We close with a summary and conclusions.
341: The appendix contains detailed derivations and explicit expressions
342: of our analytic calculations.
343: 
344: 
345: \section{Nucleon+core model}
346: \label{sec:ncm}
347: 
348: Exotic nuclei close to the driplines often exhibit a pronounced
349: structure with a nucleon $b$ (proton or neutron) 
350: weakly bound to a core $c$. 
351: They are often well
352: described by simple single-particle models
353: that are able to explain 
354: the basic features of low-energy excitations.
355: For small separation energies $S_{b}$ of the nucleon and
356: low orbital angular momenta $l$ the exotic nucleus $a$ develops
357: a proton or neutron halo where the nucleon wave function extends
358: to large radii and there is a large probability of finding
359: the nucleon outside the classically allowed region of the
360: potential $V_{bc}$. Matrix elements for electromagnetic transitions
361: only depend on a small number of characteristic scaling parameters.
362:  
363: \subsection{Scaling parameters of halo nuclei}
364: \label{subsec:scale}
365: 
366: The main scale is set by the nucleon separation energy $S_{b}$
367: ($b=n,p$)
368: that is related to an inverse decay length 
369: \begin{equation}
370:  q = \frac{\sqrt{2\mu S_{b}}}{\hbar} 
371: \end{equation}
372: of the bound state
373: with the reduced mass $\mu = m_{b}m_{c}/(m_{b}+m_{c})$.
374: It becomes very small for typical halo nuclei with small
375: binding energies. The scattering state is characterized by 
376: the momentum $\hbar k$ that is related to the relative energy
377: in the continuum $E$ by
378: \begin{equation}
379:  k = \frac{\sqrt{2\mu E}}{\hbar}  \: .
380: \end{equation}
381: A third parameter is the size of the nuclear core given
382: by the range of the core-nucleon potential $R$.
383: These three parameters allow to define the dimensionless
384: quantities
385: \begin{equation}
386:  \gamma = q R \qquad \mbox{and} \qquad \kappa = k R
387: \end{equation}
388: and the ratio
389: \begin{equation}
390:  x = \frac{\kappa}{\gamma} = \frac{k}{q} = \sqrt{\frac{E}{S_{b}}} 
391: \end{equation}
392: independent of $R$.
393: It can be considered as a definition of halo nuclei that
394: the  parameter $\gamma$ is small. This means that the extension of the 
395: wave function, characterized by $1/q$, is much larger than R. 
396: The parameter $\gamma$ can be used as
397: a convenient expansion parameter in a systematic approach to calculate
398: matrix elements. On the other hand, the
399: relevant range of the parameter $x$ 
400: extends from zero to a value of the order of one.
401: For larger values of $x$ other degrees of freedom, like
402: core excitations, will come into play and 
403: tend to invalidate the simple model.
404: 
405: For proton+core systems an additional scale enters the problem set by
406: the Gamov energy
407: \begin{equation}
408:  E_{G} = \left(\frac{Z_{b}Z_{c}e^{2}}{\hbar}\right)^{2}
409:  \frac{\mu}{2} 
410: \end{equation}
411: or the nuclear Bohr radius 
412: \begin{equation} \label{eq:nbr}
413:  a_{N} = \frac{\hbar}{\sqrt{2\mu E_{G}}}
414:  = \frac{\hbar^{2}}{Z_{b}Z_{c}e^{2}\mu}
415: \end{equation}
416: with charge numbers $Z_{b}$ and $Z_{c}$ of the nucleon
417: $b$ and the core $c$.
418: They serve to define the Sommerfeld parameters
419: \begin{equation}
420:  \eta_{i} = \frac{1}{qa_{N}} 
421:  = \sqrt{\frac{E_{G}}{S_{b}}}
422:  \qquad \mbox{and} \qquad 
423:  \eta_{f} = \frac{1}{ka_{N}} 
424:  = \sqrt{\frac{E_{G}}{E}} 
425: \end{equation}
426: of the bound ($i$) and the scattering ($f$) state
427: with the relation
428: \begin{equation}
429:  x \eta_{f} = \eta_{i} \: .
430: \end{equation}
431: The nuclear Bohr radius (\ref{eq:nbr}) is related to
432: another important parameter: 
433: half the distance of closest approach in a head on collision
434: with energy $E$. It is given by $a_{0}=1/(k^{2} a_{N})$.
435: Typical features of a halo system essentially depend on the three
436: independent parameters
437: $\gamma$, $x$, and $\eta_{i}$. 
438: %%%
439: The characteristic scaling parameters are summarized in Table~\ref{tab:1}.
440: %%%
441: For a neutron+core system one obviously has
442: $\eta_{i}=0$. In case of a proton+core system with
443: a large charge number
444: of the core and/or small binding energy $S_{b}$
445: the parameter $\eta_{i}$ can become quite large.
446: 
447: %%%
448: \begin{table}
449: \caption{\label{tab:1}Characteristic scaling parameters for electromagnetic
450: strength in single-particle halo nuclei.}
451: %\begin{center}
452: \begin{tabular}{ccc}
453:  \hline 
454:  origin & energy scale & dimensionless parameter \\ 
455:  \hline 
456:  bound state & 
457:  $\displaystyle S_{b}= \frac{\hbar^{2}q^{2}}{2\mu}$ & 
458:  $\gamma = q R$ \\
459:  & one-nucleon separation energy & \\
460:  \hline
461:  scattering state & 
462:  $\displaystyle E= \frac{\hbar^{2}k^{2}}{2\mu}$ &  
463:  $\displaystyle x =
464:  \frac{\kappa}{\gamma} = \frac{k}{q} = \sqrt{\frac{E}{S_{b}}}$  \\ 
465:  & nucleon-core relative energy & with $\kappa = kR$\\
466:  \hline
467:  Coulomb field & 
468:  $\displaystyle 
469:  E_{G}= \left( \frac{Z_{b}Z_{c}e^{2}}{\hbar}\right)^{2} \frac{\mu}{2}
470:  = \frac{\hbar^{2}}{2\mu a_{N}^{2}}$ &
471:  $\displaystyle
472:  \eta_{i} = \sqrt{\frac{E_{G}}{S_{b}}} = \frac{1}{qa_{N}} = x \eta_{f}$ \\
473:  & Gamov energy & 
474: % $\displaystyle
475: % \eta_{f} = \sqrt{\frac{E_{G}}{E}} = \frac{1}{ka_{N}}=\frac{\eta_{i}}{x}$ 
476:  \\
477:  \hline 
478: \end{tabular}
479: %\end{center}
480: \end{table}
481: %%%
482: 
483: 
484: \subsection{Probabilities in nucleon+core systems}
485: \label{subsec:prob}
486: 
487: Common to all nucleon+core nuclei with small binding energy
488: is the large probability in the bound state
489: of finding the nucleon outside the range of the nuclear potential.
490: A similar observation is made for the scattering state. For halo
491: nuclei the nucleon does not penetrate strongly into the range
492: of the nuclear potential that describes the halo nature of
493: the bound state.
494: 
495: In a simple neutron+core model assuming a square-well potential
496: of radius $R$
497: the probability $P_{nl}$ of finding the neutron with 
498: principal quantum number $n$ ($=$ number of nodes of the radial
499: wave function including the node at $r=0$) and orbital angular
500: momentum $l$ inside the range of the potential can be
501: calculated analytically (see Appendix \ref{app:A}).
502: It essentially depends on the parameter $\gamma$.
503: One finds that the typical halo structure appears only for
504: low $l$, i.e.\ $s$, $p$  waves, and small neutron 
505: separation energies $S_{n}$, i.e.\ small $\gamma$.
506: Figure~\ref{fig:0a} clearly shows the
507: increase of the probability to find the neutron outside the
508: range of the potential with decreasing $\gamma$.
509: For larger values of $l$ the
510: centrifugal barrier hinders the occurence of a halo structure
511: and the penetration into the classically forbidden region
512: is reduced.
513: A larger number of nodes in the wave function 
514: increases the halo effect again. This effect is most pronounced for 
515: s waves. In the extreme halo limit $\gamma \to 0$, however, there
516: is no dependence on $n$ any more and the probability 
517: approaches finite values of 
518: $0$, $1/3$, and $3/5$
519: for $s$, $p$, and $d$ waves, respectively.
520: In case of a more realistic Woods-Saxon shape of the nuclear potential
521: the probability is even smaller as in a square-well 
522: potential with the same radius \cite{Liu04}. 
523: This is easily understood since the depth of the potential
524: is reduced inside the radius and increased outside the radius. As a
525: consequence the probability is shifted to larger radii.
526: 
527: \begin{figure}
528: \begin{center}
529: \includegraphics[width=135mm]{pnl_bound2.ps}
530: \end{center}
531: \caption{\label{fig:0a} 
532: Probability $P_{nl}$ of finding a neutron with separation energy
533: $S_{n}$ inside the radius $R$ of a square-well potential
534: as a function of the parameter $\gamma^{2} = 2\mu S_{n}R^{2}/\hbar^{2}$
535: for orbital angular momenta $l=0,1,2$ and principal quantum numbers
536: $n=1$ (thick line) and $n=2$ (thin line).
537: }
538: \end{figure}
539: 
540: \begin{figure}
541: \begin{center}
542: \includegraphics[width=135mm]{rms_bound.ps}
543: \end{center}
544: \caption{\label{fig:0c} 
545: Root-mean-square radius $\langle r^{2} \rangle_{nl}^{\frac{1}{2}}$ 
546: of a neutron with separation energy
547: $S_{n}$ in a square-well potential of radius $R$
548: as a function of the parameter $\gamma^{2} = 2\mu S_{n}R^{2}/\hbar^{2}$
549: for orbital angular momenta $l=0,1,2$ and principal quantum numbers
550: $n=1$ (thick line) and $n=2$ (thin line).}
551: \end{figure}
552: 
553: \begin{figure}
554: \begin{center}
555: \includegraphics[width=135mm]{pnl_scatt_4.ps}
556: \end{center}
557: \caption{\label{fig:0b} 
558: Differential probability $dP_{nl}/dx$ 
559: as a function of $x=\kappa/\gamma$ 
560: of finding a neutron with orbital angular momenta $l=0,1,2$ 
561: inside the radius $R$ of a square-well potential
562: that binds a neutron with
563: separation energy $S_{n}$, orbital angular momentum $l$
564: and principal quantum numbers
565: $n=1$ (thick lines) and $n=2$ (thin lines)
566: for various values of the parameter
567: $\gamma^{2} = 2\mu S_{n}R^{2}/\hbar^{2}=0.1$ (solid line),
568: $0.5$ (dashed line), $1.0$ (dotted line), $2.0$ (dot-dashed line).
569: }
570: \end{figure}
571: 
572: 
573: A different way to characterize the halo effect for small orbital angular
574: momenta is the dependence of the root-mean-square (rms) radius 
575: $\sqrt{\langle r^{2} \rangle_{l}}$ on the parameter $\gamma$. Explicit 
576: expressions of the rms radius for $l=0$, $1$, and $2$ in the square-well
577: model were given
578: in Ref.\ \cite{Ham98}, see also \cite{Liu04}. However, simpler
579: expressions for arbitrary values of $l$ can be obtained
580: (see Appendix \ref{app:A}). The scaling behaviour of the
581: rms radius is well known, see, e.g.,
582: \cite{Han87,Rii92,Rii94,Han95,Fed93,Rii00,Jen04} 
583: and Figure \ref{fig:0c}. For $l=0$ and $l=1$ one finds the scaling laws
584: \begin{eqnarray} \label{eq:rms}
585:  \langle r^{2} \rangle_{l} \to
586:  \left\{ \begin{array}{lll}
587:  \frac{R^{2}}{\gamma^{2}} & \mbox{if} & l = 0 \\ 
588:  \frac{5R^{2}}{6\gamma} & \mbox{if} & l = 1
589:  \end{array} \right.
590: \end{eqnarray}
591: for $\gamma \to 0$. For $l\geq 2$ the rms radius approaches
592: a finite value. The divergence of $\langle r^{2} \rangle_{l}$
593: for $l=0$ and $1$ is the typical sign of the halo nature.
594: 
595: Large values for $l$ and small parameters $\gamma$ 
596: prevent the neutron in a continuum state to enter into
597: the nuclear interior
598: as long as there is no resonance. In figure \ref{fig:0b}
599: the corresponding differential probability $dP_{nl}/dx$ 
600: (see Appendix \ref{app:A}) as a function
601: of  the ratio $x = \kappa/\gamma$ is shown for various parameters $\gamma$
602: and $l$. The explicit expression (\ref{eq:pnlcont})
603: shows a scaling proportional to $\gamma$ and inversely
604: proportional to the penetrability. For larger $x$ where resonances
605: in the scattering occur the probability exhibits clear maxima.
606: However, the neutron does not 
607: penetrate deeply into the nuclear interior for small $x$,
608: small $\gamma$ and large $l$.
609: Again, a larger number of nodes in the bound-state wave function
610: that determines the potential depth also reduces the
611: probability to find the neutron in the potential well
612: for continuum states.
613: 
614: In the case of proton+core systems the probabilities of finding
615: the nucleon inside and outside the range of the potential, respectively,
616: cannot be calculated analytically in the model with a
617: square-well potential. Obviously, some qualitative changes are expected.
618: Due to the additional Coulomb barrier the proton penetrates less
619: deeply into the classically forbidden region and the
620: probability for the proton to be inside the nuclear radius is enlarged
621: for bound states. On the other hand, a proton in a continuum state
622: will be found less probable inside the range of the nuclear potential
623: as compared to a corresponding neutron with the same energy.
624: This is described by the Coulomb penetration factor.
625: 
626: 
627: Since the nucleon in halo systems can be found predominantly
628: outside the range of the nuclear potential one can expect that
629: the relevant transition matrix elements mainly depend on the asymptotics
630: of the bound and scattering wave functions. This is especially true
631: for electric multipole transitions that contain an additional
632: $r^{\lambda}$ dependence enhancing contributions from large radii.
633: 
634: \subsection{Reduced transition probabilities and radial integrals}
635: \label{subsec:trans}
636: 
637: The reduced transition probability 
638: $dB(\pi \lambda) / dE$ of multipolarity $\pi \lambda$ ($\pi=E,M$;
639: $\lambda=1,2,\dots$)
640: for the electromagnetic breakup  of the nucleus $a$ 
641: into $b+c$ with relative energy $E$
642: is the basic quantity of our study. It 
643: contains the information
644: on the nuclear structure in the initial ground state and the 
645: interaction in
646: the final continuum state. This strength function 
647: determines the response of the system to a photon and 
648: enters the expressions for the corresponding cross sections.
649: It can be extracted
650: from experimental data in order to compare
651: directly to predictions of nuclear models.
652: 
653: Assuming the nucleon+core picture for the halo nucleus the
654: reduced transition probabilities are easily expressed in terms of
655: certain radial integrals
656: with the wave functions in the 
657: initial (bound) and final (scattering) states denoted
658: by $i$ and $f$, respectively, in the following.
659: In general, the spin $s=1/2$ of the nucleon $b$
660: couples with the orbital angular momentum $l_{i/f}$ 
661: to the total angular momentum $j_{i/f}$.
662: The total angular momentum $J_{i/f}$
663: of the system $a=b+c$ is obtained by coupling 
664: $j_{i/f}$ with the spin of the core $j_{c}$.
665: 
666: Usually there are various combinations
667: of $j_{i/f}$ and $j_{c}$ that are possible to contribute to
668: a given  value of $J_{i/f}$.
669: The reduced transition probability for a specific 
670: electromagnetic transition
671: $\pi \lambda$ 
672: to a final state with momentum $\hbar k$ in the
673: continuum is given by
674: \begin{eqnarray} \label{eq:dbde}
675:   \lefteqn{\frac{dB}{dE} (\pi\lambda, J_{i} s  \to k J_{f} s)
676:   =}
677: \\ \nonumber & &  
678: \frac{2J_{f}+1}{2J_{i}+1}  \sum_{j_{f} l_{f}}
679:  \left| \sum_{j_{i} l_{i} j_{c}} 
680:  \langle k J_{f}j_{f}l_{f}s j_{c} || {\mathcal M}(\pi\lambda) 
681:   || J_{i} j_{i} l_{i} s j_{c} \rangle \right|^{2}
682:  \frac{\mu k}{(2\pi)^{3}\hbar^{2}}
683: \end{eqnarray}
684: depending on reduced multipole matrix elements.
685: In the following we will only consider electric excitations ($\pi = E$)
686: with multipole operator
687: \begin{equation}
688:  {\mathcal M}(E\lambda \mu) = 
689:  Z_{\rm eff}^{(\lambda)}e  r^{\lambda} Y_{\lambda \mu}(\hat{r})
690: \end{equation}
691: that dominate the continuum breakup of exotic nuclei.
692: The effective charge number is given by
693: \begin{equation}
694:  Z_{\rm eff}^{(\lambda)} = 
695:  Z_{b}\left(\frac{m_{c}}{m_{b}+m_{c}}\right)^{\lambda}
696:  +Z_{c}\left(-\frac{m_{b}}{m_{b}+m_{c}}\right)^{\lambda}  \: .
697: \end{equation}
698: For proton+core systems the effective charge numbers for
699: $E1$ and $E2$ transitions are of comparable magnitude and, generally,
700: one has to consider both contributions 
701: in the cross sections for Coulomb breakup,
702: photo dissociation or radiative capture. In the case of a neutron+core
703: system, the $E2$ effective charge number is suppressed by a factor
704: $1/A$ as compared to $E1$ since $Z_{b}=0$. $E1$ transitions
705: dominate the low-lying electromagnetic strength and the $E2$ contribution
706: can be neglected.
707: Neverthess we include the $E2$ case here for completeness.
708: The strong reduction of contributions from higher multipolarities in the
709: neutron+core case was noticed before, e.g.\ in
710: Refs.\ \cite{Typ01a,Ber92}.
711: 
712: In the single-particle model
713: the wave functions of the initial and final state are given by
714: \begin{eqnarray}
715:  \Phi_{i}(\vec{r}) 
716:  & = & \langle \vec{r} | J_{i} j_{i} l_{i} s j_{c} \rangle 
717:  \\ \nonumber 
718:  & = & \frac{1}{r}
719:  \sum_{m_{i}m_{c}} (j_{i} \: m_{i} \: j_{c} \: m_{c} | J_{i} \: M_{i} )
720:  f_{J_{i}j_{i}l_{i}}^{j_{c}}(r) {\mathcal Y}_{j_{i}m_{i}}^{l_{i}s}
721: (\hat{r}) \phi_{j_{c}m_{c}}
722: \end{eqnarray}
723: and
724: \begin{eqnarray}
725:  \Phi_{f}(\vec{r})
726:  & = & \langle \vec{r} | \vec{k} J_{f} j_{f} l_{f} s j_{c} \rangle 
727:  \\ \nonumber 
728:  & = & \frac{4\pi}{kr}
729:  \sum_{m_{f}m_{c}} (j_{f} \: m_{f} \: j_{c} \: m_{c} | J_{f} \: M_{f} )
730:  g_{J_{f}j_{f}l_{f}}^{j_{c}}(r) i^{l_{f}} 
731:  Y_{l_{f}m_{f}}^{\ast}(\hat{k})
732:  {\mathcal Y}_{j_{f}m_{f}}^{l_{f}s}
733: (\hat{r}) \phi_{j_{c}m_{c}} \: ,
734: \end{eqnarray}
735: respectively, 
736: with the radial wave functions $f^{j_{c}}_{J_{i}j_{i}l_{i}}(r)$
737: and $g_{J_{f}j_{f}l_{f}}^{j_{c}}(r)$
738: and with the spinor spherical harmonics
739: \begin{equation}
740: {\mathcal Y}_{jm}^{ls} = \sum_{m_{l} m_{s}}
741:  ( l \: m_{l} \: s \: m_{s} | j \: m) Y_{lm}(\hat{r}) \chi_{s m_{s}} \: .
742: \end{equation}
743: The wave function of the core is denoted by $\phi_{j_{c}m_{c}}$.
744: The reduced matrix element in (\ref{eq:dbde}) can be expressed as
745: \begin{eqnarray} \label{eq:rme}
746:  \lefteqn{\langle k J_{f} j_{f} l_{f} s j_{c} || {\mathcal M}(E\lambda) 
747:  || J_{i} j_{i} l_{i} s j_{c}\rangle =}
748:  \\ \nonumber & & 
749:  \frac{4\pi Z_{\rm eff}^{(\lambda)}e}{k} 
750:  D_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda s j_{c}) \:
751:  (-i)^{l_{f}} I_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
752: \end{eqnarray}
753: with 
754: %the effective charge number
755: the angular momentum coupling coefficient
756: \begin{eqnarray}
757: D_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda s j_{c})  & = & 
758: (-1)^{s+j_{i}+l_{f}+\lambda} 
759: (-1)^{j_{c}+J_{i}+j_{f}+\lambda}
760:  (l_{i} \: 0 \: \lambda \: 0 | l_{f} \: 0 ) 
761:   \\ \nonumber & & 
762: \sqrt{2j_{i}+1} \sqrt{2l_{i}+1}
763:  \sqrt{2J_{i}+1} \sqrt{2j_{f}+1}
764: \\ \nonumber & & 
765:  \sqrt{\frac{2\lambda+1}{4\pi}} 
766: \left\{ \begin{array}{ccc}
767:  l_{i} & s & j_{i} \\ j_{f} & \lambda & l_{f}
768:  \end{array} \right\}
769:  \left\{ \begin{array}{ccc}
770:  j_{i} & j_{c} & J_{i} \\ J_{f} & \lambda & j_{f}
771:  \end{array} \right\} \: ,
772: \end{eqnarray}
773: and the radial integral
774: \begin{equation} \label{eq:radint}
775:  I_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
776:  = \int_{0}^{\infty} dr \: 
777:  g^{j_{c}\ast}_{J_{f}j_{f}l_{f}}(r)
778:  r^{\lambda} f^{j_{c}}_{J_{i}j_{i}l_{i}}(r)
779: \end{equation}
780: that contains the radial wave functions 
781: of the bound state $f^{j_{c}}_{J_{i}j_{i}l_{i}}(r)$
782: and the  continuum state $g^{j_{c}}_{J_{f}j_{f}l_{f}}(r)$,
783: respectively.
784: 
785: The asymptotic of the radial wave functions for the ground state
786: \begin{equation} \label{eq:asymb}
787:  f^{j_{c}}_{J_{i}j_{i}l_{i}}(r) 
788:  \to C^{j_{c}}_{J_{i}j_{i}l_{i}} W_{-\eta_{i}, l_{i}+1/2}
789:  (2qr)
790: \end{equation}
791: is determined by the asymptotic normalization coefficient 
792: $C^{j_{c}}_{J_{i}j_{i}l_{i}}$ of the true many-body wave function and
793: a Whittaker function $W_{-\eta_{i}, l_{i}+1/2}$ \cite{Abr65}.
794: The bound state is characterized by the parameters
795: $q$, $\eta_{i}$, and the orbital angular momentum $l_{i}$.
796: In the pure single-particle model for the nucleon+core system
797: the corresponding 
798: ANC $C^{j_{c}}_{J_{i}j_{i}l_{i}}(\rm sp)$ is determined by the
799: normalization of the wave function $f^{j_{c}}_{J_{i}j_{i}l_{i}}(r)$.
800: Because the radial wave function for small $r$ depends on the
801: nuclear potential of the single-particle model there is a model-dependence
802: of $C^{j_{c}}_{J_{i}j_{i}l_{i}}(\rm sp)$. Similarly, the 
803: spectroscopic factor $S^{j_{c}}_{J_{i}j_{i}l_{i}}$ for the
804: nucleon+core configuration 
805: depends on the single-particle model that is used for the
806: calculation. In contrast, the
807: actual ANC of the true many-body wave function
808: \begin{equation}
809:  C^{j_{c}}_{J_{i}j_{i}l_{i}}
810:  = C^{j_{c}}_{J_{i}j_{i}l_{i}}({\rm sp}) 
811:  \left[ S^{j_{c}}_{J_{i}j_{i}l_{i}} \right]^{\frac{1}{2}}
812: \end{equation}
813: is a model-independent quantity. 
814: It is directly inferred from transfer reactions for example.
815: In the following we always assume
816: a spectroscopic factor of one, i.e.\ the single-particle ANC
817: and the true ANC are identical.
818: 
819: For the scattering state we have the asymptotic form
820: \begin{eqnarray} \label{eq:asyms}
821:  g^{j_{c}}_{J_{f}j_{f}l_{f}}(r) & \to &
822:   \exp\left[i(\sigma_{l_{f}}+\delta^{j_{c}}_{J_{f}j_{f}l_{f}})\right]
823: \\  \nonumber & & \times
824:  \left[ \cos (\delta^{j_{c}}_{J_{f}j_{f}l_{f}}) \: F_{l_{f}}(\eta_{f};kr) 
825:  + \sin ( \delta^{j_{c}}_{J_{f}j_{f}l_{f}}) \:  G_{l_{f}}(\eta_{f};kr) \right]
826: \end{eqnarray}
827: with regular and irregular Coulomb wave functions
828: $F_{l_{f}}$ and $G_{l_{f}}$ \cite{Abr65}, 
829: respectively, and Coulomb phase shifts
830: $\sigma_{l_{f}}$ that depend on 
831: the Sommerfeld parameter $\eta_{f}=\eta_{i}/x$.
832: Effects of the  nuclear interaction in the final-state
833: are contained in the nuclear phase shifts $\delta^{j_{c}}_{J_{f}j_{f}l_{f}}$.
834: 
835: 
836: In general the radial integral (\ref{eq:radint}) has to be calculated
837: with the relevant bound and scattering wave functions
838: with the correct asymptotics (\ref{eq:asymb}) and (\ref{eq:asyms}), 
839: respectively. 
840: They are obtained by
841: solving the Schr\"{o}dinger equation for a given potential, e.g.\
842: of Woods-Saxon form $V(r) = -V_{0}/(1+\exp[(r-R_{0})/a])$
843: with depth $V_{0}$, radius $R_{0}$ and diffuseness parameter $a$.
844: In the following calculations 
845: we will only consider nuclear central potentials and neglect 
846: contributions from the spin-orbit interaction to simplify
847: the discussion. 
848: In case of proton+core systems 
849: the contribution of the Coulomb potential
850: can be assumed to be that of a homogeneously charged sphere
851: of the same radius $R_{0}$ as the nuclear potential. The potential
852: depth $V_{0}$ has to be adjusted to give the correct separation
853: energy $S_{b}$ of the nucleon in the ground state of $a$.
854: However, in the continuum state the parameter $V_{0}$ can be
855: varied freely to investigate the dependence of the transition
856: strength on the final-state interaction. 
857: 
858: \begin{figure}
859: \begin{center}
860: \includegraphics[width=135mm]{rinexb.ps}
861: \end{center}
862: \caption{\label{fig:0d}
863: Modulus of the ratio $R_{l_{i}}^{l_{f}}(1)$ of the interior and
864: exterior radial integrals for dipole transitions $s \to p$ (top) and
865: $p \to s$ (bottom) as a functions of $x^{2}=E/S_{n}$ for various values of
866: the parameter $\gamma^{2}= 2\mu S_{n}R^{2}/\hbar^{2}$. 
867: Thick (thin) lines correspond to the pricipal quantum number $n=1$ ($n=2$).}
868: \end{figure}
869: 
870: A remarkably good approximation 
871: for the nuclear interaction is the
872: square-well potential and many results can be derived 
873: analytically. They already show the main 
874: features for the scaling laws of the matrix elements. 
875: In this case the contributions to the integral
876: (\ref{eq:radint}) from the interior and exterior part can be calculated
877: independently to estimate their relevance in the transition matrix element.
878: In Appendix \ref{app:A4} explicit expressions for the dipole integrals
879: (suppressing irrelevant quantum numbers)
880: \begin{equation} 
881:  I_{l_{i}}^{l_{f}}(1 <)
882:  = \int_{0}^{R} dr \: 
883:  g^{\ast}_{l_{f}}(r) r f_{l_{i}}(r)
884: %\end{equation}
885: \quad \mbox{and} \quad
886: %\begin{equation} 
887:  I_{l_{i}}^{l_{f}}(1 >)
888:  = \int_{R}^{\infty} dr \: 
889:  g^{\ast}_{l_{f}}(r) r f_{l_{i}}(r)
890: \end{equation}
891: in the neutron+core case for $s \to p$ and $p \to s$ transitions
892: are derived. They only depend on the value of the radial wave functions
893: and their logarithmic derivative at the radius $R$ of the square well.
894: The ratio
895: \begin{equation}
896:  R_{l_{i}}^{l_{f}}(1) = \frac{I_{l_{i}}^{l_{f}}(1 <)}{I_{l_{i}}^{l_{f}}(1 >)}
897: \end{equation}
898: as a function of $x^{2}$ for $s \to p$ and $p \to s$ transitions,
899: respectively, is displayed in figure \ref{fig:0d} for various values of
900: $\gamma^{2}$
901: assuming the same depth of the square-well potential for the bound
902: and scattering state. For halo nuclei with small $\gamma^{2}$ the 
903: total radial integral is clearly dominated by the exterior integral
904: even at values $x^{2}=E/S_{n} \gg 1$. In the case of a resonance in the
905: scattering wave (e.g. for the $s \to p$ transition with $\gamma^{2} = 1$)
906: the ratio $R_{l_{i}}^{l_{f}}(1)$ shows a distinct peak since the
907: scattering wave function penetrates into the core of the nucleus.
908: In the limit $x \to 0$ one finds the scaling laws 
909: \begin{equation}
910:  R_{0}^{1}(1)  \to  \frac{\gamma^{4}}{2(2n-1)^{2}\pi^{2}}
911:  \quad \mbox{and} \quad
912:  R_{1}^{0}(1)  \to  -\frac{\gamma^{2}}{4n^{2}\pi^{2}}
913: \end{equation}
914: depending on the principal quantum number $n=1,2,\dots$ 
915: (see appendix \ref{app:A4}).
916: For larger values of $n$ the ratios
917: become smaller but they remain finite for a given $\gamma$ in the limit
918: $x \to 0$. As can be seen from figure \ref{fig:0d} these
919: scaling laws are also well satisfied for larger values of $x^{2}$
920: as long as there are no resonances or accidental zeros in the 
921: interior integral. For halo nuclei one 
922: clearly sees that the total radial integral is well
923: approximated by the exterior part with the asymptotic form of the
924: wave functions that is independent of the details of the potential.
925: This will be even more true for higher multipolarities due to the $r^{\lambda}$
926: factor in the radial integral.
927: Only the parameters $\gamma$, $\kappa$, the ANC $C_{l_{i}}$ 
928: and the phase shift $\delta_{l_{f}}$ are really relevant.
929: 
930: \subsection{Dipole integrals and commutator relations}
931: \label{subsec:comrel}
932: 
933: In the case of electric dipole transitions the relevant reduced matrix element
934: (\ref{eq:rme}) can also be calculated with the help of the
935: commutator relation \cite{Ehr27,Gor29}
936: \begin{equation} \label{eq:commrel}
937:  \langle \Phi_{f} | \left[ H , \left[ H, \vec{r} \right] \right] 
938: | \Phi_{i} \rangle
939:  =  (E_{f}-E_{i})^{2} \langle \Phi_{f} | \vec{r} | \Phi_{i} \rangle \: .
940: \end{equation}
941: Introducing the scaling parameters $\gamma$ and $x$, we
942: find
943: \begin{equation} \label{eq:dme}
944: \langle \Phi_{f} | \vec{r} | \Phi_{i} \rangle 
945:  = \frac{4\mu^{2}R^{4}}{\hbar^{4}\gamma^{4}(1+x^{2})^{2}}
946:  \langle \Phi_{f} | \left[ H , \left[ H, \vec{r} \right] \right] 
947: | \Phi_{i} \rangle
948: \end{equation} 
949: for the dipole matrix element. This equation
950: directly shows the occurence
951: of a pole at $x^{2} = -1$, %\kappa^{2} = -\gamma^{2}$, 
952: i.e., $E=-S_{b}$. This derivation is much more transparent than
953: the corresponding discussion in \cite{Jen98} for $E1$ radiative capture
954: reactions.
955: For halo systems with small nucleon
956: separation energies $S_{b}$ one immediately sees that a series expansion
957: of the dipole matrix element (\ref{eq:dme})
958: in the parameter $x$ (or the energy $E$) 
959: is only of limited value
960: because of the small radius of convergence 
961: %%%
962: $x_{\rm conv} = 1$
963: %%%
964: %$r_{\rm conv} = 1$ 
965: ($E_{\rm conv} = S_{b}$). It would be more advantageous to expand the
966: matrix element with the double commutator 
967: $\left[ H , \left[ H, \vec{r} \right]\right]$ in a power series in $x$.
968: 
969: The double commutator can be calculated assuming different forms
970: of the potential $V$ in the Hamiltonian $H=p^{2}/(2\mu)+V$ of the system.
971: The simplest case is a central potential $V(r)$
972: that commutes with $\vec{r}$, i.e.\ $\left[ V, \vec{r} \right] = 0$. 
973: With $\left[ H , \left[ H, \vec{r} \right] \right]
974: = \hbar^{2} \vec{\nabla} V(r)/\mu$
975: the radial integral (\ref{eq:radint}) can be expressed as
976: \begin{equation} \label{eq:ricom}
977:  I_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 j_{c})
978:  = \frac{4\mu R^{4}}{\hbar^{2}\gamma^{4}(1+x^{2})^{2}} 
979:  \int_{0}^{\infty} dr \: 
980:  g^{j_{c}\ast}_{J_{f}j_{f}l_{f}}(r)
981:  \left( \frac{d}{dr} V(r) \right) f^{j_{c}}_{J_{i}j_{i}l_{i}}(r) \: .
982: \end{equation}
983: 
984: 
985: In the neutron+core case with a square-well potential 
986: \begin{equation}
987:  V(r) = - V_{0} \theta(R-r)
988: \end{equation}
989: we find that the radial integral is given by
990: \begin{equation} \label{eq:radintgf}
991:  I_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 j_{c})
992:  = \frac{2R^{2}v}{(\gamma^{2}+\kappa^{2})^{2}} 
993:  g^{j_{c}\ast}_{J_{f}j_{f}l_{f}}(R)
994:  f^{j_{c}}_{J_{i}j_{i}l_{i}}(R)
995: \end{equation}
996: It depends only on the values of the initial and final radial wave function
997: at the radius $R$
998: and the depth of the potential in the dimensionless parameter
999: \begin{equation}
1000:  v = \frac{2 \mu V_{0} R^{2}}{\hbar^{2}} 
1001: \end{equation}
1002: that characterizes the strength of the interaction.
1003: It is not surprising that this result can also be derived
1004: by evaluating the radial integral (\ref{eq:radint}) directly
1005: (see appendix \ref{app:A4}).
1006: For more realistic potentials the derivative $dV/(dr)$
1007: also peaks close to the nuclear radius and the radial integral is
1008: mainly sensitive to the bound and scattering wave functions near the
1009: nuclear surface.
1010: 
1011: For finite values of $R$ the radial integral (\ref{eq:radintgf})
1012: shows a $x^{-4}$ dependence for large energies independent of the
1013: orbital angular momenta in the initial and final state. 
1014: (The wave function $g_{l_{i}}$ is independent of $x$ and $f_{l_{f}}$
1015: shows an oscillatory behaviour for large $x$.)
1016: %ST added:
1017: This $E^{-2}$ dependence is sufficient for the convergence of
1018: the non energy-weighted and energy-weighted sum rules as discussed in
1019: section \ref{sec:totsum}.
1020: %
1021: %In the special case of
1022: %$l \to l+1$ transitions we can let $R$ go to zero
1023: %in (\ref{eq:radintgf}).
1024: %corresponding to a delta potential
1025: %$\hbar^{2} v \delta(r)/(2\mu)$. 
1026: %We have $f_{l}(R)
1027: %\to (2l-1)!!C_{l}/(qR)^{l}$ and
1028: %$g_{l+1}(R)\propto kR j_{l+1}(kR) \to
1029: %\kappa^{l+2}/(2l+3)!!$ for $R \to 0$ \cite{Abr65}.
1030: %Then, for the radial integral we obtain
1031: %\begin{equation} \label{eq:radintgf0}
1032: % \lim_{R \to 0} I_{l}^{l+1}(1)
1033: % = \frac{2C_{l}v}{(2l+3)q^{2}} \frac{x^{l+2}}{(1+x^{2})^{2}}
1034: %\end{equation}
1035: %supplying an additional factor of $x^{l+2}$
1036: %for large $x$, i.e., the high energy behaviour
1037: %is modified to 
1038: %$I_{l}^{l+1}(1) \propto x^{l-2}$
1039: %and the convergence of the sum rules is not guaranteed
1040: %for large $l$.
1041: 
1042: In the proton+core case the potential is well approximated by
1043: \begin{equation}
1044:  V(r) = -V_{0} \: \theta(R-r)
1045:  + \frac{Z_{c}e^{2}}{r} \: \theta(r-R)
1046: \end{equation}
1047: and the radial integral becomes
1048: \begin{eqnarray} \label{eq:icoul}
1049:  I_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 j_{c})
1050:  & = & \frac{2R^{2}}{(\gamma^{2}+\kappa^{2})^{2}}
1051:  \\ \nonumber & &
1052:  \times
1053:  \left[ (v+2\gamma \eta_{i}) g^{j_{c}\ast}_{J_{f}j_{f}l_{f}}(R)
1054:  f^{j_{c}}_{J_{i}j_{i}l_{i}}(R)
1055:  -  2 \gamma \eta_{i}K_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 j_{c}, R) \right]
1056: \end{eqnarray}
1057: with the 
1058: integral
1059: \begin{equation}
1060:  K_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 j_{c}, R) 
1061:  = R  \int_{R}^{\infty} \frac{dr}{r^{2}} \: 
1062:  g^{j_{c}\ast}_{J_{f}j_{f}l_{f}}(r)
1063:  f^{j_{c}}_{J_{i}j_{i}l_{i}}(r) 
1064: \end{equation} 
1065: that converges much more rapidly than the usual $E1$ transition
1066: integral.
1067: This contribution only depends on the parameters $\gamma$, $\eta_{i}$,
1068: and $x$.
1069: The ratio 
1070: \begin{equation} \label{eq:coulc}
1071:   F_{l_{i}}^{l_{f}}(1) =
1072:  \frac{2 \gamma \eta_{i}}{(v+2\gamma\eta_{i})}
1073:  \frac{K_{l_{i}}^{l_{f}}(1,R)}{g^{\ast}_{l_{f}}(R)f_{l_{i}}(R)} 
1074: \end{equation}
1075: of the two contributions in (\ref{eq:icoul}) is depicted in Fig.\
1076: \ref{fig:0e} as a function of $x^{2}$ for typical values of $\gamma^{2}$
1077: and $\eta_{i}$ for both $s \to p$ and $p \to s$ transitions.
1078: It is clearly seen that the ratio $F_{l_{i}}^{l_{f}}(1)$ of the two
1079: contributions to the total radial integral increases with increasing 
1080: $\eta_{i}$ and decreasing $\gamma$. This is expected since Coulomb
1081: effects become stronger for larger $\eta_{i}$ and the bound state
1082: wave function at larger radii is less steep for smaller $\gamma$. 
1083: The ratio (\ref{eq:coulc}) decreases for larger values of $x^{2}$ with
1084: a larger effect for large values of $\gamma$. Generally, transitions $p \to s$
1085: are more strongly 
1086: affected by the correction than transitions $s \to p$, however,
1087: for large $\gamma$ (less halo effect) the difference becomes smaller.
1088: 
1089: \begin{figure}
1090: \begin{center}
1091: \includegraphics[width=135mm]{coulc_1.ps}
1092: \end{center}
1093: \caption{\label{fig:0e}
1094: Ratio $F_{l_{i}}^{l_{f}}(1)$ of Eq.\ (\ref{eq:coulc}) 
1095: as a function of $x^{2}=E/S_{p}$
1096: for various values of $\gamma^{2}= 2\mu S_{p}R^{2}/\hbar^{2}$ and $\eta_{i}$
1097: for the bound state without a node (principal quantum number $n=1$).
1098: Thick (thin) lines correspond to transitions $s \to p$ ($p \to s$).}
1099: \end{figure}
1100: 
1101: Ususally, one can expect that the Hamiltonian 
1102: $H=p^{2}(2\mu)+V$ contains more general potentials that do not commute
1103: with $\vec{r}$ \cite{Lev50}, e.g.,
1104: \begin{equation} \label{eq:vls}
1105:  V=V_{c}(r)
1106: +V_{ls}(r)\vec{\ell} \cdot \vec{s}
1107: \end{equation}
1108: with a central potential $V_{c}(r)$ and
1109: a spin-orbit contribution $V_{ls}(r)$. In this case the double commutator
1110: in (\ref{eq:commrel}) can be calculated explictly and a rather complicated
1111: radial integral is obtained. Alternatively, we can
1112: introduce $H_{f} = P_{f}HP_{f}$
1113: and $H_{i}=P_{i}HP_{i}$ with projection operators $P_{f}$ and $P_{f}$ on the
1114: final and initial state, respectively. Then the commutator 
1115: relation (\ref{eq:commrel}) can be generalized to
1116: \begin{eqnarray} \label{eq:commrel2}
1117:   (E_{f}-E_{i})^{2} \langle \Phi_{f} | \vec{r} | \Phi_{i} \rangle
1118:  & = & 
1119: \langle \Phi_{f} | \left(
1120:  \left[ \Sigma , \left[ \Sigma, \vec{r} \right] \right] 
1121:  + \left[ \Sigma, \{ \Delta, \vec{r} \} \right]
1122: \right.  \\ \nonumber & &  \left.
1123:  + \{ \Delta, \left[ \Sigma, \vec{r} \right] \}
1124:  + \{ \Delta, \{ \Delta, \vec{r} \} \} \right)
1125: | \Phi_{i} \rangle
1126: \end{eqnarray}
1127: where
1128: $\Sigma=(H_{f}+H_{i})/2$, $\Delta = (H_{f}-H_{i})/2$ and $\{ . ,. \}$ denotes
1129: the anticommutator. For the potential (\ref{eq:vls}) we find 
1130: $H_{k} = p^{2}/(2\mu)+V_{k}(r)$ with $k=i,f$ and
1131: \begin{equation}
1132:  V_{k}(r)  =   V_{c}(r) + \frac{V_{ls}(r)}{2}
1133:  \left[ j_{k}(j_{k}+1)-l_{k}(l_{k}+1)-s(s+1)\right]
1134: \end{equation}
1135: where $j_{k}$, $l_{k}$, and $s$ are the total angular momentum, the orbital
1136: angular momentum, and the spin of the nucleon, respectively. The central
1137: potentials $V_{k}(r)$ with $\left[ V_{k}(r), \vec{r} \right]=0$ are
1138: now different in the bound and scattering states, however, the various
1139: contributions in (\ref{eq:commrel2}) can be calculated easily. Assuming 
1140: a square-well potential for both $V_{c}(R)$ and $V_{ls}(r)$ with the same
1141: radius but different depths the radial integral corresponding
1142: to (\ref{eq:radintgf}) becomes a more complicated 
1143: expression that contains also derivatives of the wave functions at the radius
1144: $R$. Explicit expresssions are obtained from adding the
1145: interior and exterior integrals as derived in appendix \ref{app:A4}.
1146: Also other forms of the potentials, e.g., a surface-peaked spin-orbit
1147: potential $V_{ls} = \mbox{const.} \times \delta(r-R)$, 
1148: can be considered. However, we omit the details of the calculation.
1149: 
1150: 
1151: 
1152: 
1153: \subsection{Cross sections}
1154: \label{subsec:xs}
1155: 
1156: The reduced transition probability determines
1157: the cross sections for photo-nuclear reactions.
1158: The photo dissociation cross section
1159: \begin{equation} \label{eq:sigabs}
1160:  \sigma_{\pi \lambda}(a+\gamma \to b+c) = 
1161:  \frac{\lambda+1}{\lambda} \frac{(2\pi)^{3}}{[(2\lambda+1)!!]^{2}}
1162:  \left(\frac{E_{\gamma}}{\hbar c}\right)^{2\lambda-1}
1163:  \frac{dB(\pi \lambda)}{dE}
1164: \end{equation}
1165: is proportional to $dB(\pi \lambda)/dE$ where
1166: $E_{\gamma}= E+S_{b}$ is the
1167: sum of the binding energy $S_{b}>0$ of the nucleus $a$ with respect
1168: to the breakup into $b+c$ and the relative energy $E$
1169: in the final state. The photo absorption cross section (\ref{eq:sigabs}) also
1170: enters the cross section for
1171: the electromagnetic dissociation reaction $a+X \to b+c+X$
1172: during the scattering of an exotic nucleus $a$ on a target nucleus $X$.
1173: A first-order calculation  gives
1174: \begin{equation} \label{eq:cdxs}
1175:  \frac{d^{2}\sigma}{d\Omega_{aX}dE} =
1176:  \frac{1}{E_{\gamma}} \sum_{\pi\lambda} 
1177:  \sigma_{\pi \lambda}(a+\gamma \to b+c) \frac{dn_{\pi \lambda}}{d\Omega_{aX}}
1178: \end{equation}
1179: with virtual photon numbers $dn_{\pi \lambda} / d\Omega_{aX}$
1180: that can be calculated in the semiclassical theory or in prior-form
1181: DWBA in the quantal approach.
1182: The factorization of the cross section (\ref{eq:cdxs})
1183: into contributions that are related to the  nuclear structure 
1184: of the exotic nucleus $\sigma_{\pi \lambda}$ and to the excitation mechanism
1185: $dn_{\pi \lambda} / d\Omega_{aX}$ is 
1186: no longer valid if higher-order effects from the target-fragment
1187: interactions are significant.
1188: From the photo dissociation reaction it is also possible to
1189: extract information on the
1190: radiative capture reaction $b+c \to a +\gamma$
1191: that for low energies is relevant for nuclear astrophysics.
1192: The capture cross section
1193: \begin{equation} \label{eq:detbal}
1194:  \sigma_{\pi \lambda}(b+c \to a + \gamma)  = 
1195:  \frac{2(2J_{a}+1)}{(2J_{b}+1)(2J_{c}+1)}
1196:  \frac{k_{\gamma}^{2}}{k^{2}}
1197:  \sigma_{\pi \lambda}(a + \gamma \to b + c)
1198: \end{equation}
1199: is obtained by applying the theorem of detailed balance.
1200: The photo absorption cross section (\ref{eq:sigabs})
1201: is multiplied with a factor that contains the spins $J_{i}$ of the particles
1202: $i=a,b,c$, the photon momentum $\hbar k_{\gamma} = E_{\gamma}/c$,
1203: and the final state momentum $\hbar k$.
1204: 
1205: For electric dipole transitions it is easy to find
1206: the high-energy behaviour of the various cross sections.
1207: From equation (\ref{eq:ricom}) the dependence 
1208: $I_{l_{i}}^{l_{f}}(1) \propto x^{-4} \propto E^{-2}$ of the radial transition
1209: integral is extracted. With (\ref{eq:dbde}) and (\ref{eq:rme})
1210: one finds $dB(E1)/dE \propto x^{-1} |I_{l_{i}}^{l_{f}}(1)|^{2}  
1211: \propto x^{-9}$ or $dB(E1)/dE \propto E^{-9/2}$. This corresponds to
1212: a dependence $\sigma_{E1}(a+\gamma \to b +c) \propto E^{-7/2}$
1213: and $\sigma_{E1}(b +c \to a + \gamma) \propto E^{-5/2}$
1214: for the photo dissociation and radiative capture cross sections
1215: at high energiess $E$, respectively.
1216: We note that in the case of the photodissociation of atoms,
1217: with the $1/r$ shape of the Coulomb potential,
1218: one finds the same $E_{\gamma}^{-7/2}$ law, see 
1219: \cite{Bet77}.
1220: 
1221: 
1222: \subsection{Effective-range expansion}
1223: \label{subsec:scalen}
1224: 
1225: The nucleon-core interaction that is responsible for the
1226: binding of the system also generates structures in the continuum.
1227: At low energies in the continuum the phase shifts are insensitive
1228: to the details of the nuclear potential as long as no resonance
1229: appears. They
1230: are determined by only a few parameters that are given by
1231: the effective-range approximation. 
1232: E.g., in the case of charged-particle scattering
1233: the low-energy phase shift $\delta_{l}$ in the partial wave 
1234: with orbital angular momentum $l$ is determined
1235: by the scattering length $a_{l}$ and the effective range
1236: $r_{l}$ in the expansion \cite{Bru96}
1237: \begin{equation} \label{eq:fre}
1238:  C_{l}^{2}(\eta) k^{2l+1} \left[ \cot \delta_{l} 
1239:  + \frac{2\eta h(\eta)}{C_{0}^{2}(\eta)}\right]
1240:  = -\frac{1}{a_{l}} + \frac{r_{l}}{2} k^{2} + \dots
1241: \end{equation}
1242: with the function
1243: \begin{eqnarray}
1244:  h(\eta) & = & %\Re \psi(-i\eta) - \ln \eta =
1245:  \frac{1}{2} \left[ \psi (1+i\eta) + \psi(1-i\eta)\right] - \ln \eta
1246:  \\ \nonumber 
1247:  & = & \eta^{2} \sum_{n=1}^{\infty} \frac{1}{n(n^{2}+\eta^{2})} 
1248:  - \gamma -\ln \eta 
1249: \end{eqnarray}
1250: that depends on the Sommerfeld parameter $\eta$ in the argument
1251: of the Digamma function $\psi$ and
1252: $\gamma=0.5772156649\dots$ is Euler's constant.
1253: The constants $C_{l}^{2}$
1254: for increasing $l$
1255: are obtained by means of a recursion relation
1256: \begin{equation} \label{eq:coulfac}
1257:  C_{l}^{2}(\eta) = C_{l-1}^{2}(\eta) \left( 1 + \frac{\eta^{2}}{l^{2}}\right)
1258:  \qquad \mbox{with} \qquad
1259:  C_{0}^{2}(\eta) = \frac{2\pi\eta}{\exp(2\pi\eta) -1}  \: .
1260: \end{equation}
1261: Note that $a_{l}$ and $r_{l}$ for $l>0$ do not have the dimension
1262: of a length.
1263: The effective-range expansion can be used to express the
1264: phase shifts directly as functions of the momentum $\hbar k$.
1265: 
1266: Taking only the scattering length $a_{l}$ and the effective range
1267: $r_{l}$ into account one does not necessarily find a good approximation
1268: of the phase shift $\delta_{l}$ over a wide range of 
1269: energies in the continuum. 
1270: However, the effective-range expansion motivates to introduce 
1271: an energy-dependent function $b_{l}$ via the equation
1272: \begin{equation}
1273:  C_{l}^{2} x^{2l+1} \left[ \cot \delta_{l} 
1274:  + \frac{2\eta h(\eta)}{C_{0}^{2}}\right]
1275:  = -\frac{1}{b_{l}^{2l+1}} 
1276: \end{equation}
1277: with the dimensionless parameter 
1278: $x=k/q$ where $q$ is constant for a given nucleus.
1279: In general, the function $b_{l}$ 
1280: depends on the momentum $\hbar k$
1281: in the final state. 
1282: It fully describes the effects of the
1283: interaction in the scattering state at all energies.
1284: The quantity $b_{l}$ varies slowly with the parameter $x=k/q$
1285: as long as no resonance is approached.
1286: For small $k$ one can identify $b_{l}$ with a reduced scattering length
1287: since the usual scattering length is obtained in the limit
1288: \begin{equation} \label{eq:abrel}
1289:  a_{l} = \lim_{x\to 0} \left(\frac{b_{l}}{q}\right)^{2l+1}  
1290: \end{equation}
1291: and for $\delta_{l} = 0$ we obviously have $b_{l} = 0$.
1292: 
1293: \section{Reduced radial integrals and shape functions}
1294: \label{sec:rrisf}
1295: 
1296: Considering the asymptotics of the radial wave functions
1297: one can define dimensionless reduced radial integrals 
1298: ${\mathcal I}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})$
1299: by the relation
1300: \begin{equation}
1301:  I_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
1302:  = \frac{C^{j_{c}}_{J_{i}j_{i}l_{i}}}{q^{\lambda+1}} 
1303:  \exp\left[i(\sigma_{l_{f}}+\delta^{j_{c}}_{J_{f}j_{f}l_{f}})\right]
1304:  {\mathcal I}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c}) \: .
1305: \end{equation}
1306: At low energies the main contribution to the radial integral
1307: arises from radii $r$ larger than the radius of the nucleus.
1308: This is especially true for halo nuclei where the probability
1309: of finding the nucleon inside the range of the nuclear potential
1310: is small (see subsection \ref{subsec:prob}).
1311: Neglecting contributions from radii smaller than a cutoff radius $R$
1312: the reduced radial integrals can be approximated by
1313: \begin{eqnarray} \label{eq:idef}
1314:  {\mathcal I}_{l_{i}}^{l_{f}}(\lambda) & = & 
1315:   \left[ \cos (\delta_{l_{f}}) \:  {\mathcal F}_{l_{i}}^{l_{f}}(\lambda)
1316:  + \sin ( \delta_{l_{f}}) \:  {\mathcal G}_{l_{i}}^{l_{f}}(\lambda)\right]
1317: \end{eqnarray}
1318: where
1319: $ {\mathcal F}_{l_{i}}^{l_{f}}(\lambda)$
1320: and
1321: $  {\mathcal G}_{l_{i}}^{l_{f}}(\lambda)$
1322: denote the real and the imaginary part of the function
1323: \begin{eqnarray} \label{eq:hdef}
1324:  {\mathcal H}_{l_{i}}^{l_{f}}(\lambda) 
1325:  & = & q^{\lambda+1} 
1326:  \int_{R}^{\infty} dr \: r^{\lambda} 
1327:   \left[ F_{l_{f}}(\eta_{f};kr)+ i G_{l_{f}}(\eta_{f};kr)   \right]
1328:  W_{-\eta_{i}, l_{i}+\frac{1}{2}} (2qr)  
1329:  \\ \nonumber 
1330:  & = & \gamma^{\lambda+1} 
1331:  \int_{1}^{\infty} dt \: t^{\lambda} 
1332:   \left[ F_{l_{f}}(\eta_{i}/x;x\gamma t)
1333:      + i G_{l_{f}}(\eta_{i}/x;x\gamma t)   \right]
1334:  W_{-\eta_{i}, l_{i}+\frac{1}{2}} (2\gamma t)  
1335:  \: .
1336: \end{eqnarray}
1337: Here, as in the following, the quantum
1338: numbers $J_{i}$, $j_{i}$, $J_{f}$, $j_{f}$ and $j_{c}$
1339: have been suppressed if no confusion arises.
1340: The nuclear phase shifts $\delta_{l_{f}}$ encode the effects of
1341: the final-state interaction. For $\delta_{l_{f}}=0$
1342: one obtains the results without the nuclear interaction between
1343: the nucleon and the core. Note that the function
1344: ${\mathcal H}_{l_{i}}^{l_{f}}(\lambda)$ depends only on the
1345: parameters $\gamma$, $\eta_{i}$ and $x$.
1346: 
1347: If there is only one fixed pair $(j_{i},l_{i})$ in the initial
1348: state and similar $(j_{f},l_{f})$ in the final state the expression 
1349: for the reduced transition probability reduces to
1350: \begin{eqnarray} \label{eq:dbelde}
1351:  \lefteqn{\frac{dB}{dE} (E\lambda, J_{i} s j_{c}  \to k J_{f} s j_{c})=}
1352:  \\ \nonumber & & 
1353:   \left[Z_{\rm eff}^{(\lambda)}e\right]^{2} 
1354:  \frac{2\mu}{\pi\hbar^{2}} 
1355:  \frac{2J_{f}+1}{2J_{i}+1}  
1356:  \left[  D_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda s j_{c}) \right]^{2}
1357:   \frac{\left|C^{j_{c}}_{J_{i}j_{i}l_{i}}\right|^{2}}{q^{2\lambda+3}}
1358:  {\mathcal S}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
1359: \end{eqnarray}
1360: with the dimensionless shape function of the transition strength
1361: \begin{equation} \label{eq:sdef}
1362:  {\mathcal S}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
1363:  =  \frac{1}{x} \left| % \frac{q}{k} \left| 
1364:  {\mathcal I}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
1365: \right|^{2}
1366: \end{equation} 
1367: that completely contains the dependence on the
1368: momentum in the continuum. Similarly, one can define the 
1369: characteristic shape functions
1370: for the photo absorption
1371: \begin{equation} \label{eq:sabs}
1372:   {\mathcal S}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}({\rm abs},\lambda j_{c})
1373:  = \left(1+x^{2}\right)^{2\lambda-1}
1374: % \left(\frac{q^{2}+k^{2}}{q^{2}}\right)^{2\lambda-1}
1375:  {\mathcal S}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
1376: \end{equation}
1377: and for the capture cross section
1378: \begin{equation} \label{eq:scapt}
1379:   {\mathcal S}_{J_{f}j_{f}l_{f}}^{J_{i}j_{i}l_{i}}({\rm capt},\lambda j_{c})
1380:  = \frac{(1+x^{2})^{2\lambda+1}}{x^{2}}
1381: % \frac{(q^{2}+k^{2})^{2\lambda+1}}{k^{2}q^{4\lambda}}
1382:  {\mathcal S}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c}) \: .
1383: \end{equation}
1384: The expression (\ref{eq:dbelde})
1385: directly shows
1386: the dependence of the transition strength on the characteristic parameters
1387: $C^{j_{c}}_{J_{i}j_{i}l_{i}}$ and $q$ of the ground state. For small
1388: separation energies $S_{b}$ of the nucleon the strength becomes very large
1389: since $q$ is a small number. Note that the ANC in general depends on $q$, too.
1390: In a model for neutrons in a square-well potential 
1391: one finds a dependence $C^{j_{c}}_{J_{i}j_{i}l_{i}} \propto \sqrt{q}$ 
1392: for  $l_{i}=0$ 
1393: and $C^{j_{c}}_{J_{i}j_{i}l_{i}} \propto q^{l_{i}}$ for $l_{i}>0$,
1394: respectively,  for small $q$ (see Appendix \ref{app:A}).
1395: 
1396: When the nucleon $b$ is a neutron the reduced radial integrals can be
1397: calculated analytically (see Appendix \ref{app:B}). Then the 
1398: functions (\ref{eq:idef}) only
1399: depend on the dimensionless variables $\gamma = qR$ and 
1400: $\kappa = kR = x \gamma$ 
1401: and the phase shift $\delta_{l_{f}}$ in the final state. 
1402: It is found that the reduced radial integrals have the general form
1403: \begin{eqnarray} \label{eq:iredgen}
1404:   {\mathcal I}_{l_{i}}^{l_{f}}(\lambda) & = &
1405:  \frac{\gamma \exp(-\gamma)}{(\gamma^{2}+\kappa^{2})^{\lambda+1}} 
1406:  \left(\frac{\gamma}{\kappa}\right)^{l_{f}}
1407:  \\ \nonumber & & \times
1408:  \left[ {\mathcal R}_{l_{i}}^{(+)l_{f}}(\lambda) \cos (\kappa+\delta_{l_{f}}) 
1409:  + {\mathcal R}_{l_{i}}^{(-)l_{f}}(\lambda) \sin(\kappa+\delta_{l_{f}})\right]
1410: \end{eqnarray}
1411: with polynomials/rational functions 
1412: ${\mathcal R}_{l_{i}}^{(\pm)l_{f}}(\lambda)$.
1413: Explicit expressions are given in Appendix \ref{app:B}. In general they are 
1414: complicated functions of $\gamma$ and $\kappa$. 
1415: 
1416: From the discussion in subsection \ref{subsec:comrel} a $x^{-4}$
1417: dependence of the reduced radial integrals 
1418: ${\mathcal I}_{l_{i}}^{l_{f}}(1)$ for dipole transitions is expected
1419: for large $x$. 
1420: However, the integrals (\ref{eq:iredgen}) for $\lambda=1$
1421: show a different high-$x$
1422: behaviour and the convergence of the sum rules (see section
1423: \ref{sec:totsum}) is not guaranteed. This is a consequence of neglecting
1424: the interior contribution to the radial integral that becomes relevant
1425: at high relative energies since the nucleon penetrates into the core.
1426: Thus, the reduced radial integrals (\ref{eq:iredgen}) are only a
1427: good approximation for not too high relative energies in the continuum.
1428: %Equation (\ref{eq:iredgen}) can be compared to 
1429: In comparison, equation (\ref{eq:radintgf}) is  valid 
1430: irrespective of the halo nature of the system and for all energies
1431: but only 
1432: for $\lambda =1$ and the square-well case; it contains all contributions 
1433: to the radial integral from zero to infinity. Equation (\ref{eq:iredgen}) 
1434: is a good approximation for halo nuclei at small relative energies, 
1435: since the interior contributions are small 
1436: in this case (cf.\ Fig.\ \ref{fig:0d}). 
1437: It can also be applied easily to cases where
1438: the potential is different in the bound and scattering states.
1439: It is worthwhile to study some limiting cases.
1440: 
1441: \begin{table}
1442: \caption{\label{tab:2}Characteristic shape functions 
1443: ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$ for $\gamma=0$ and no final-state
1444: interaction.}
1445: %\begin{center}
1446: \begin{tabular}{lll}
1447:  \hline 
1448: % monopole & dipole & quadrupole \\
1449:  $\lambda=0$ & $\lambda = 1$ & $\lambda=2$ \\
1450:  \hline 
1451:  $\displaystyle {\mathcal S}_{0}^{0}(0) = \frac{x}{(1+x^{2})^{2}}$ &
1452:  $\displaystyle {\mathcal S}_{1}^{0}(1) = 
1453:  \frac{x(3+x^{2})^{2}}{(1+x^{2})^{4}}$ &
1454:  $\displaystyle {\mathcal S}_{2}^{0}(2) = 
1455:  \frac{x(15+10x^{2}+3x^{4})^{2}}{(1+x^{2})^{6}}$ \\
1456:  $\displaystyle {\mathcal S}_{1}^{1}(0) = \frac{x^{3}}{(1+x^{2})^{2}}$ &
1457:  $\displaystyle {\mathcal S}_{0}^{1}(1) = \frac{4x^{3}}{(1+x^{2})^{4}}$ &
1458:  $\displaystyle {\mathcal S}_{1}^{1}(2) = 
1459:  \frac{4x^{3}(5+x^{2})^{2}}{(1+x^{2})^{6}}$ \\
1460:   &
1461:  $\displaystyle {\mathcal S}_{2}^{1}(1) = 
1462:  \frac{x^{3}(5+3x^{2})^{2}}{(1+x^{2})^{4}}$ &
1463:  \\
1464:  &
1465:  $\displaystyle {\mathcal S}_{1}^{2}(1) = \frac{4x^{5}}{(1+x^{2})^{4}}$ &
1466:  $\displaystyle {\mathcal S}_{0}^{2}(2) = \frac{64x^{5}}{(1+x^{2})^{6}}$ \\
1467:  \hline 
1468: \end{tabular}
1469: %\end{center}
1470: \end{table}
1471: 
1472: \subsection{Shape functions in n+core systems without final-state interaction}
1473: \label{subsec:n+core}
1474: 
1475: \begin{figure}
1476: \begin{center}
1477: \includegraphics[width=135mm]{shape0.ps}
1478: \end{center}
1479: \caption{\label{fig:1} 
1480: Dependence of the generic 
1481: shape functions ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
1482: with cutoff radius $R=0$
1483: on $x^{2}=E/S_{n}$ for various 
1484: transitions $l_{i} \to l_{f}$ of multipolarity
1485: $E1/E2$ (top/bottom). The functions are normalized to 1
1486: at $x^{2}=1$.}
1487: \end{figure}
1488: 
1489: Without nuclear interaction between the neutron and the core 
1490: in the final state the phase shift $\delta_{l_{f}}$ is zero 
1491: and there is no contribution from the integral with the
1492: irregular wave function in  (\ref{eq:idef}). The contribution
1493: for radii $r < R$  in the radial integral is small and it is
1494: possible to take the limit $R \to 0$ keeping the neutron separation
1495: energy $S_{n}$ or equivalently the inverse bound-state decay 
1496: length $q$ constant. In this limit both $\gamma$
1497: and $\kappa$ approach zero but the ratio $x=\kappa/\gamma=
1498: \sqrt{E/S_{n}}$, i.e.\ the relevant variable for the shape
1499: of the strength distribution, is independent of $R$.
1500: In this case the reduced radial integrals (\ref{eq:iredgen})
1501: assume a particular simple form. 
1502: They only depend on this dimensionless variable $x$.
1503: From the reduced radial integrals the shape functions (\ref{eq:sdef})
1504: of the reduced transition probability are easily derived.
1505: The functions ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$ in the limit
1506: $R \to 0$ with $\delta_{l_{f}}=0$ are given in table \ref{tab:2}.
1507: %In the limit $R \to 0$ with $\delta_{0}=0$ they are given by
1508: %(the monopole shape functions are only given for completeness)
1509: %\begin{eqnarray} \label{eq:gens1}
1510: % & & 
1511: % {\mathcal S}_{0}^{0}(0) = \frac{x}{(1+x^{2})^{2}} \qquad
1512: % {\mathcal S}_{1}^{1}(0) = \frac{x^{3}}{(1+x^{2})^{2}} 
1513: % \\ & & \label{eq:sdip_a}
1514: % {\mathcal S}_{1}^{0}(1) = \frac{x(3+x^{2})^{2}}{(1+x^{2})^{4}} \qquad
1515: % {\mathcal S}_{0}^{1}(1) = \frac{4x^{3}}{(1+x^{2})^{4}} \qquad
1516: % \\ & & \label{eq:sdip_b}
1517: % {\mathcal S}_{2}^{1}(1) = \frac{x^{3}(5+3x^{2})^{2}}{(1+x^{2})^{4}} \qquad
1518: % {\mathcal S}_{1}^{2}(1) = \frac{4x^{5}}{(1+x^{2})^{4}}
1519: % \\ & &  
1520: % {\mathcal S}_{2}^{0}(2) = \frac{x(15+10x^{2}+3x^{4})^{2}}{(1+x^{2})^{6}}
1521: % \qquad
1522: % {\mathcal S}_{1}^{1}(2) = \frac{4x^{3}(5+x^{2})^{2}}{(1+x^{2})^{6}} 
1523: % \\ & & 
1524: % {\mathcal S}_{0}^{2}(2) = \frac{64x^{5}}{(1+x^{2})^{6}} 
1525: %  \label{eq:gens4} 
1526: %\end{eqnarray}
1527: Especially the expression ${\mathcal S}_{0}^{1}(1)$ is well known,
1528: see, e.g., \cite{Bla79,Nak94}. It can be found in a different notation
1529: also in Refs.\ \cite{Ber88,Ber92}.
1530: %The $x$ dependence for the shape functions ${\mathcal S}_{l}^{l+1}(1)$
1531: %is consistent with the expectation (\ref{eq:radintgf0}) for the
1532: %square-well potential.
1533: 
1534: In Fig.~\ref{fig:1} the generic form of the shape functions
1535: ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$ is shown as a 
1536: function of $x^{2} = E/S_{n}$ for $E1$ and $E2$ transitions
1537: from a bound state with orbital angular momentum $l_{i}$ to
1538: a scattering state with orbital angular momentum $l_{f}$.
1539: The dependence of the transition shape on the centrifugal barrier
1540: is clearly seen. The peak at small energies is more pronounced
1541: for low orbital angular momenta $l_{f}$. Only for $s$ and $p$ waves
1542: in the continuum a large transition strength is found close to the threshold.
1543: At low $x$ we have the typical dependence
1544: \begin{equation}
1545:  {\mathcal S}_{l_{i}}^{l_{f}}(\lambda) \propto x^{2l_{f}+1} \: .
1546: \end{equation}
1547: The shape function for photo absorption 
1548: ${\mathcal S}_{l_{i}}^{l_{f}}(\mbox{abs},\lambda)$
1549: has the same $x$ dependence as ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
1550: for small final state momenta. 
1551: The shape function
1552: for radiative capture
1553: ${\mathcal S}_{l_{f}}^{l_{i}}(\mbox{capt},\lambda)$,
1554: on the other hand,
1555: shows a $x^{2l_{f}-1}$ dependence for small  momenta in the continuum
1556: due to the additional momentum dependent factor from the theorem
1557: of detailed balance (\ref{eq:detbal}), see equation (\ref{eq:sabs}).
1558: 
1559: \begin{figure}
1560: \begin{center}
1561: \includegraphics[width=135mm]{shape4.ps}
1562: \end{center}
1563: \caption{\label{fig:1b} 
1564: Shape functions ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
1565: as a function of $x^{2}=E_{bc}/S_{n}$ for 
1566: values of the parameter $\gamma$ between $0.0$ and $1.0$ in steps of
1567: $0.2$.}
1568: \end{figure}
1569: 
1570: In case of a finite cutoff radius $R$ in the radial integral
1571: (\ref{eq:radint})
1572: the shape function ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
1573: depends on both $\gamma=qR$ and
1574: $\kappa=kR$. Since $q$ is fixed for a given nucleus is it reasonable again
1575: to use $\gamma$ and $x=\kappa/\gamma=k/q$ as independent parameters.
1576: For halo nuclei with small nucleon separation energy $q$ will be
1577: a small quantity. As long as $R$ does not 
1578: become too large, e.g.\
1579: for heavy nuclei, $\gamma$ will also be small. The variation of
1580: the shape functions with $\gamma$ gives an estimate of the contribution
1581: to the radial integral from the nuclear interior. In Fig.\ \ref{fig:1b}
1582: the change of the shape functions with $\gamma$ is shown for the
1583: transitions of Fig.\ \ref{fig:1}. There is a clear systematic trend.
1584: The shape functions are less sensitive to a change in $\gamma$
1585: for larger final state orbital angular momentum $l_{f}$ and
1586: higher multipolarity $\lambda$ because of the suppression of
1587: the integrand at small $r$ from the spherical Bessel functions
1588: $j_{l_{f}}(kr)$ and $r^{\lambda}$ from the transition operator,
1589: respectively. On the other hand, a larger orbital angular momentum
1590: $l_{i}$ in the bound state increases the sensitivity since
1591: the wave function introduces a $r^{-l_{i}}$ dependence at small radii.
1592: This explains the strong $\gamma$-dependence of the shape function
1593: ${\mathcal S}_{2}^{1}(1)$.
1594: 
1595: \subsection{Shape functions in n+core systems with final-state interaction}
1596: \label{subsec:fsinc}
1597: 
1598: For neutron scattering the finite-range expansion (\ref{eq:fre}) reduces to
1599: \begin{equation} \label{eq:frex}
1600:  k^{2l+1} \cot(\delta_{l}) = - \frac{1}{a_{l}} + \frac{r_{l}}{2} k^{2}
1601:  + \dots
1602: \end{equation}
1603: since $\eta = 0$.
1604: Taking only the contributions with $a_{l}$ and $r_{l}$ into account
1605: the phase shift $\delta_{l}$ crosses the value $\pi/2$ at an energy
1606: $E_{0}=\hbar^{2}/(\mu a_{l} r_{l})$.
1607: This behaviour produces a resonance in the corresponding
1608: partial wave if the phase shift is increasing that, however,
1609: does not have to be physical.
1610: In general, the actual energy-dependence of the phase shift
1611: close to a resonance at energy $E_{R}$ is given by the
1612: Breit-Wigner form
1613: $ \tan \delta_{l} = \Gamma/(2(E_{R}-E))$
1614: with positive width $\Gamma$
1615: where the parameters $E_{R}$ and $\Gamma$ are not related to $a_{l}$
1616: and $r_{l}$ in the effective-range expansion.
1617: %a resonance in the continuum
1618: %is obtained at an energy  Then the width of the resonance is given by 
1619: %$\Gamma = -2 E_{R} a_{l}k_{R}^{2l+1}$. 
1620: Particular values for these parameters that correctly reproduce
1621: the phase shift at low momenta will not necessarily reproduce
1622: the position and the width of an actual 
1623: resonance with orbital angular momentum $l$.
1624: However, the phase shift can be calculated with the help of the relation
1625: \begin{equation} \label{eq:tanbl}
1626:  \tan (\delta_{l}) = - (xb_{l})^{2l+1} 
1627: % \quad \mbox{or} \quad
1628: % \exp(2i\delta_{l}) =  \frac{1-i(xb_{l})^{2l+1}}{1+i(xb_{l})^{2l+1}}
1629: \end{equation}
1630: for general cases assuming an energy dependence of 
1631: the function $b_{l}$.
1632: It replaces the phase shift $\delta_{l}$
1633: in order to take the FSI in the scattering wave function into account.
1634: The dimensionless function $b_{l}$ is related by
1635: \begin{equation}
1636:  D_{l}(E) = - \left(\frac{b_{l}}{q}\right)^{2l+1}
1637: \end{equation}
1638: to the function $D_{l}(E)$ introduced in Ref.\ \cite{Bay04} for
1639: a low-energy expansion of cross sections for radiative capture reactions.
1640: 
1641: The function $b_{l}$ is of the order of $\gamma=qR$ 
1642: unless the logarithmic derivative $L$ of the scattering wave function
1643: at radius $R$ is close to $-l$, see eq.\ (\ref{eqn:scalen_l}).
1644: It is useful in expansions of the
1645: shape functions ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$ if
1646: the limit $R \to 0$ is considered. 
1647: On the other hand, the
1648: scaled function 
1649: \begin{equation} \label{eq:spf}
1650:   c_{l} = \frac{b_{l}}{\gamma}
1651: \end{equation} 
1652: (usually of order one)
1653: is the appropriate quantity
1654: if the limit $\gamma \to 0$ is studied. In this case we have
1655: \begin{equation} 
1656:  a_{l} = \lim_{x\to 0} \left(c_{l}R\right)^{2l+1}  
1657: \end{equation}
1658: with constant radius $R$.
1659: 
1660: \begin{figure}
1661: \begin{center}
1662: \includegraphics[width=135mm]{rsl_Be.ps}
1663: \end{center}
1664: \caption{\label{fig:2} 
1665: Function $b_{l}$ 
1666: in partial waves with orbital angular momentum $l=0,1,2$
1667: as a function of $x=k/q$ for the
1668: scattering of a neutron on ${}^{10}$Be in a single-particle model with
1669: a Woods-Saxon potential of radius $R_{0}=2.78$~fm, diffuseness parameter
1670: $a=0.65$~fm and various potential depths $V_{0}$.
1671: (Note the nonlinear scale on the $y$ axis.)}
1672: \end{figure}
1673: 
1674: 
1675: Typical values of $b_{l}$ and $c_{l}$ can
1676: be estimated from examples employing a single particle model
1677: where a nuclear potential of Woods-Saxon form with typical
1678: parameters is assumed.
1679: In Figure~\ref{fig:2} the function $b_{l}$ 
1680: for orbital angular momenta
1681: $l=0,1,2$ is shown as a function of $x=k/q$ for the
1682: scattering of a neutron on ${}^{10}$Be assuming different depths
1683: $V_{0}$ of the potential. In most cases the function
1684: $b_{l}$ is quite constant
1685: as a function of $x$ except when there is a resonance 
1686: (vertical lines in Fig.~\ref{fig:2}) in the continuum
1687: for a certain fixed depth $V_{0}$. 
1688: For $p$ and $d$ waves $b_{l}$ is
1689: usually in the intervall $[-0.5,0.5]$; for $s$ waves $b_{0}$ covers
1690: a larger range.
1691: Assuming a zero-range potential for the neutron-core interaction
1692: the scattering wave function
1693: is given by $\psi^{(+)}(\vec{r}) = \exp(i\vec{k}\cdot \vec{r})
1694: -\exp(ikr)/[(q+ik)r]$ with the bound state parameter $q$
1695: determined by the binding energy.
1696: The $s$ wave scattering length is just $a_{0} =1/q$ 
1697: corresponding to $b_{0} = 1$ and $c_{0} = 1/\gamma$.
1698: Since the ground state is
1699: close to the threshold for a halo nucleus 
1700: $c_{0}$ becomes unnaturally large and
1701: there will be a large effect
1702: on the continuum in the same partial wave. However, the partial waves in
1703: the final state for an electric excitation have a different orbital angular
1704: momentum 
1705: and parity and usually there will be no resonance in the
1706: energy range of interest.
1707: In general it is reasonable to 
1708: assume $|b_{l}|<0.5$ with a weak momentum dependence.
1709: 
1710: \begin{table}
1711: \caption{\label{tab:3}Characteristic shape functions 
1712: ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$ for $\gamma=0$ and $b_{l_{f}} \neq 0$.}
1713: %\begin{center}
1714: \begin{tabular}{ll}
1715:  \hline 
1716:  $\lambda = 0$ & $\lambda = 1$ \\
1717: % monopole & dipole \\
1718:  \hline 
1719:  $\displaystyle {\mathcal S}_{0}^{0}(0) = \frac{x}{(1+x^{2})^{2}} 
1720:  \frac{(1-b_{0})^{2}}{1+(xb_{0})^{2}}$ &
1721:  $\displaystyle {\mathcal S}_{1}^{0}(1)  = \frac{x}{(1+x^{2})^{4}} 
1722:  \frac{(3-2b_{0}+x^{2})^{2}}{1+(xb_{0})^{2}}$  \\
1723:  &
1724:  $\displaystyle {\mathcal S}_{0}^{1}(1)  = \frac{x^{3}}{(1+x^{2})^{4}}
1725:  \frac{[2-b_{1}^{3}(1+3x^{2})]^{2}}{1+(xb_{1})^{6}}$ \\
1726:  \hline 
1727:  $\lambda = 2$ \\
1728: % quadrupole & \\
1729:  \hline 
1730:  $\displaystyle {\mathcal S}_{2}^{0}(2)  =  \frac{x}{(1+x^{2})^{6}} 
1731:  \frac{(15-8b_{0}+10x^{2}+3x^{4})^{2}}{1+(xb_{0})^{2}}$ & \\
1732:  $\displaystyle {\mathcal S}_{1}^{1}(2) = \frac{x^{3}}{(1+x^{2})^{6}} 
1733:  \frac{4[5+x^{2}-b_{1}^{3}(1+5x^{2})]^{2}}{1+(xb_{1})^{6}}$ & \\
1734:  $\displaystyle {\mathcal S}_{0}^{2}(2) = \frac{x^{5}}{(1+x^{2})^{6}} 
1735:  \frac{[8-b_{2}^{5}(3+10x^{2}+15x^{4})]^{2}}{1+(xb_{2})^{10}}$ & \\
1736:  \hline 
1737: \end{tabular}
1738: %\end{center}
1739: \end{table}
1740: 
1741: The phase shift in the general expression for the
1742: reduced radial integral (\ref{eq:iredgen})
1743: can be replaced by the function $b_{l}$ using the explicit relation
1744: (\ref{eq:tanbl}).
1745: For some of the transitions it is still possible
1746: to take the limit $R\to 0$ in the reduced radial integrals
1747: for $b_{l}\neq 0$. 
1748: The corresponding shape functions are given in table \ref{tab:3}.
1749: %One finds the shape functions
1750: %\begin{eqnarray} \label{eq:sb1}
1751: % {\mathcal S}_{0}^{0}(0) & = & \frac{x}{(1+x^{2})^{2}} \:
1752: % \frac{(1-b_{0})^{2}}{1+(xb_{0})^{2}}
1753: % \\ \label{eq:sb2}
1754: % {\mathcal S}_{1}^{0}(1) & = &\frac{x}{(1+x^{2})^{4}} \:
1755: % \frac{(3-2b_{0}+x^{2})^{2}}{1+(xb_{0})^{2}}
1756: % \\
1757: % {\mathcal S}_{0}^{1}(1) & = &\frac{x^{3}}{(1+x^{2})^{4}} \:  
1758: % \frac{[2-b_{1}^{3}(1+3x^{2})]^{2}}{1+(xb_{1})^{6}}
1759: % \\
1760: % {\mathcal S}_{2}^{0}(2) & = & \frac{x}{(1+x^{2})^{6}} \:
1761: % \frac{(15-8b_{0}+10x^{2}+3x^{4})^{2}}{1+(xb_{0})^{2}} \:
1762: % \\
1763: % {\mathcal S}_{1}^{1}(2) & = & \frac{x^{3}}{(1+x^{2})^{6}} \:
1764: % \frac{4[5+x^{2}-b_{1}^{3}(1+5x^{2})]^{2}}{1+(xb_{1})^{6}}
1765: % \\ \label{eq:sb6}
1766: % {\mathcal S}_{0}^{2}(2) & = &\frac{x^{5}}{(1+x^{2})^{6}} \:
1767: % \frac{[8-b_{2}^{5}(3+10x^{2}+15x^{4})]^{2}}{1+(xb_{2})^{10}}
1768: % \: .
1769: %\end{eqnarray}
1770: For the other transitions the reduced radial integral and the
1771: shape function diverges.
1772: It is obvious that the $x^{2l_{f}+1}$ dependence at small $x$ is
1773: obtained again but the modulus is modified in the limit $x\to 0$
1774: as compared to the generic shape functions in table \ref{tab:2}
1775: %(\ref{eq:gens1}) - (\ref{eq:gens4}) 
1776: because of the FSI.
1777: 
1778: \begin{figure}
1779: \begin{center}
1780: \includegraphics[width=135mm]{shape1.ps}
1781: \end{center}
1782: \caption{\label{fig:3} 
1783: Shape functions ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
1784: for $\gamma=0$
1785: as a function of $x^{2}=E/S_{n}$ for various 
1786: values of the function $b_{l_{f}}$ in steps of
1787: $0.1$.  The thick dashed, solid, and dot-dashed lines correspond to
1788: $b_{l_{f}}=+0.5$, $0.0$, $-0.5$, respectively.}
1789: \end{figure}
1790: 
1791: In Fig.~\ref{fig:3} the dependence of the shape function 
1792: in the limit $R \to 0$ on $b_{l_{f}}$ is shown
1793: for the cases of table \ref{tab:3}. %(\ref{eq:sb2}) - (\ref{eq:sb6}). 
1794: (Transitions with $\lambda=0$
1795: are not relevant.) The function $b_{l_{f}}$
1796: varies in the interval $[-0.5,0.5]$. This corresponds
1797: to reasonable values that are far from a resonance in the particular
1798: continuum channel. Depending on the magnitude of $b_{l_{f}}$ 
1799: the shape functions show a pronounced variation around
1800: the generic form that reflects the effect
1801: of the potential in the final state. Small changes in the reduced
1802: scattering length lead to a smaller effect for higher orbital
1803: angular momenta $l_{f}$ due to the $b_{l_{f}}^{2l_{f}+1}$ dependence
1804: in the analytical expressions. For positive values of $b_{l_{f}}$
1805: one finds an increase of 
1806: ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$ whereas a negative $b_{l_{f}}$ leads
1807: to a decrease in absolute value. There is also a change in the shape
1808: observed and a shift of the maximum. This shift 
1809: becomes noticable only for large values of $b_{l}$.
1810: The position of the maximum moves to larger $x^{2}$ with
1811: decreasing $b_{l}$ for $p$ and
1812: $d$ waves in the final state, whereas the trend is opposite
1813: for $s$-wave final states. There is also a clear hierachy observed.
1814: The maximum of the shape function appears at higher $x^{2}$ with
1815: larger orbital angular momentum $l_{f}$ in the continuum.
1816: 
1817: \begin{table}
1818: \caption{\label{tab:4}Expansion of the characteristic shape functions 
1819: ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$ in the parameter $x$
1820: for finite $\gamma$ and $b_{l_{f}} \neq 0$.}
1821: %\begin{center}
1822: \begin{tabular}{l}
1823:  \hline 
1824:  $\lambda = 0$ \\
1825: % monopole  \\
1826:  \hline 
1827:  $\displaystyle {\mathcal S}_{0}^{0}(0) =
1828:  \exp(-2\gamma) \left( 1+\gamma-b_{0}\right)^{2} x$  \\
1829:  $\displaystyle {\mathcal S}_{1}^{1}(0) =
1830:  \frac{\exp(-2\gamma)}{9\gamma^{2}}
1831:  \left[\gamma(3+3\gamma+\gamma^{2})-3b_{1}^{3}\right]^{2} x^{3}$ \\
1832:  \hline 
1833:  $\lambda = 1$ \\
1834: % dipole \\
1835:  \hline 
1836:  $\displaystyle {\mathcal S}_{1}^{0}(1) =
1837:  \exp(-2\gamma) \left[ 3+3\gamma+\gamma^{2}-b_{0}(2+\gamma)\right]^{2} x$ \\
1838:  $\displaystyle {\mathcal S}_{0}^{1}(1) =
1839:  \frac{\exp(-2\gamma)}{9} 
1840:  \left(6+6\gamma+3\gamma^{2}+\gamma^{3}-3b_{1}^{3} \right)^{2} x^{3}$ \\
1841:  $\displaystyle {\mathcal S}_{2}^{1}(1) =
1842:  \frac{\exp(-2\gamma)}{9\gamma^{2}} 
1843:  \left[\gamma(15+15\gamma+6\gamma^{2}+\gamma^{3})
1844:    -3b_{1}^{3}(3+\gamma)\right]^{2} x^{3}$ \\
1845:  $\displaystyle {\mathcal S}_{1}^{2}(1) =
1846:  \frac{\exp(-2\gamma)}{225\gamma^{2}} 
1847:  \left[\gamma(30+30\gamma+45\gamma^{2}+5\gamma^{3}+\gamma^{4})
1848:    -45b_{2}^{5} \right]^{2} x^{5}$\\
1849:  \hline 
1850:  $\lambda = 2$ \\
1851: % quadrupole \\
1852:  \hline 
1853:  $\displaystyle {\mathcal S}_{2}^{0}(2) =
1854:  \exp(-2\gamma) \left[15+15\gamma+6\gamma^{2}+\gamma^{3}
1855:    -b_{0}(8+5\gamma+\gamma^{2})\right]^{2} x$ \\
1856:  $\displaystyle {\mathcal S}_{1}^{1}(2) =
1857:  \frac{\exp(-2\gamma)}{9} 
1858:  \left[30+30\gamma+15\gamma^{2}+5\gamma^{3}+\gamma^{4}
1859:    -3b_{1}^{3}(2+\gamma) \right]^{2} x^{3}$ \\
1860:  $\displaystyle {\mathcal S}_{0}^{2}(2) =
1861:  \frac{\exp(-2\gamma)}{225} 
1862:  \left(120+120\gamma+60\gamma^{2}+20\gamma^{3}+5\gamma^{4}+\gamma^{5}
1863:    -45b_{2}^{5} \right)^{2} x^{5}$ \\
1864:  \hline 
1865: \end{tabular}
1866: %\end{center}
1867: \end{table}
1868: 
1869: The analytic expressions of the shape functions in table \ref{tab:3}
1870: %(\ref{eq:sb1}) - (\ref{eq:sb6}) 
1871: were obtained by extending the asymptotic
1872: form of the bound and scattering wave functions to zero radius
1873: in the radial integral. The integrand with the regular scattering wave
1874: function shows a $r^{\lambda-l_{i}+l_{f}+1}$ dependence; in contrast to that
1875: one finds a $r^{\lambda-l_{i}-l_{f}}$ 
1876: behaviour at small $r$ for the integrand with the
1877: irregular scattering wave function.
1878: One might expect that the irregular contribution 
1879: is overestimated in this limit since it diverges for $r \to 0$. 
1880: Expanding the general
1881: shape functions in powers of $x$ 
1882: with constant $\gamma = qR$
1883: %for finite cutoff radius $R$
1884: one finds the results as given in table \ref{tab:4}.
1885: %\begin{eqnarray}
1886: % {\mathcal S}_{0}^{0}(0) & = & 
1887: % \exp(-2\gamma) \left( 1+\gamma-b_{0}\right)^{2} x
1888: % \\
1889: % {\mathcal S}_{1}^{1}(0) & = &
1890: % \frac{\exp(-2\gamma)}{9\gamma^{2}}
1891: % \left[\gamma(3+3\gamma+\gamma^{2})-3b_{1}^{3}\right]^{2} x^{3}
1892: % \\ \label{eq:s110x}
1893: % {\mathcal S}_{1}^{0}(1) & = &
1894: % \exp(-2\gamma) \left[ 3+3\gamma+\gamma^{2}-b_{0}(2+\gamma)\right]^{2} x
1895: % \\
1896: % {\mathcal S}_{0}^{1}(1) & = &
1897: % \frac{\exp(-2\gamma)}{9} 
1898: % \left(6+6\gamma+3\gamma^{2}+\gamma^{3}-3b_{1}^{3} \right)^{2} x^{3}
1899: % \\
1900: % {\mathcal S}_{2}^{1}(1) & = &
1901: % \frac{\exp(-2\gamma)}{9\gamma^{2}} 
1902: % \left[\gamma(15+15\gamma+6\gamma^{2}+\gamma^{3})
1903: %   -3b_{1}^{3}(3+\gamma)\right]^{2} x^{3}
1904: % \\ \label{eq:s121x}
1905: % {\mathcal S}_{1}^{2}(1) & = &
1906: % \frac{\exp(-2\gamma)}{225\gamma^{2}} 
1907: % \left[\gamma(30+30\gamma+45\gamma^{2}+5\gamma^{3}+\gamma^{4})
1908: %   -45b_{2}^{5} \right]^{2} x^{5}
1909: % \\
1910: % {\mathcal S}_{2}^{0}(2) & = &
1911: % \exp(-2\gamma) \left[15+15\gamma+6\gamma^{2}+\gamma^{3}
1912: %   -b_{0}(8+5\gamma+\gamma^{2})\right]^{2} x
1913: % \\
1914: % {\mathcal S}_{1}^{1}(2) & = &
1915: % \frac{\exp(-2\gamma)}{9} 
1916: % \left[30+30\gamma+15\gamma^{2}+5\gamma^{3}+\gamma^{4}
1917: %   -3b_{1}^{3}(2+\gamma) \right]^{2} x^{3}
1918: % \\
1919: % {\mathcal S}_{0}^{2}(2) & = &
1920: % \frac{\exp(-2\gamma)}{225} 
1921: % \left(120+120\gamma+60\gamma^{2}+20\gamma^{3}+5\gamma^{4}+\gamma^{5}
1922: %   -45b_{2}^{5} \right)^{2} x^{5}
1923: %\end{eqnarray}
1924: The functions again show the typical
1925: $x^{2l_{f}+1}$ dependence at low $x=k/q$ but the slope depends
1926: less strongly on $b_{l_{f}}$ 
1927: for finite $\gamma$ than for the functions given in table \ref{tab:3}.
1928: %in Eqs.\ (\ref{eq:sb1}) - (\ref{eq:sb6}). 
1929: The shape functions of the transitions $d\to p$ and $p \to d$ 
1930: ($p\to p$) for $\lambda=1$ ($\lambda=0$) diverge in the limit $\gamma \to 0$
1931: or $R \to 0$.
1932: 
1933: \begin{figure}
1934: \begin{center}
1935: \includegraphics[width=135mm]{shape2.ps}
1936: \end{center}
1937: \caption{\label{fig:4} 
1938: Shape functions ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
1939: for $\gamma=0.5$
1940: as a function of $x^{2}=E/S_{n}$ for various 
1941: values of $b_{l_{f}}$ in steps of
1942: $0.1$.  The thick dashed, solid, and dot-dashed lines correspond to
1943: $b_{l_{f}}=+0.5$, $0.0$, and $-0.5$, respectively.}
1944: \end{figure}
1945: 
1946: \begin{figure}
1947: \begin{center}
1948: \includegraphics[width=135mm]{shape3.ps}
1949: \end{center}
1950: \caption{\label{fig:5} 
1951: Shape functions ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
1952: for $\gamma=1.0$
1953: as a function of $x^{2}=E/S_{n}$ for various 
1954: values of $b_{l_{f}}$ in steps of
1955: $0.2$.  The thick dashed, solid, and dot-dashed lines correspond to
1956: $b_{l_{f}}= +1.0$, $0.0$, and $-1.0$, respectively. 
1957: The scale of the $y$ axis is chosen to
1958: be the same as in Fig.~\ref{fig:4} for a better comparison.}
1959: \end{figure}
1960: 
1961: Fig.~\ref{fig:4} shows the dependence of the shape functions
1962: on $b_{l_{f}}$ for constant $\gamma=0.5$, i.e.\ a constant 
1963: cutoff radius $R$. 
1964: This value of $\gamma$ 
1965: corresponds to $R=3.36$~fm in case of neutron
1966: scattering on ${}^{10}$Be with a bound state parameter 
1967: $q=0.1487$~fm$^{-1}$ for a neutron separation energy of $S_{n}=0.504$~MeV.
1968: The shape functions are very similar
1969: to the case with $R\to 0$, cf.\ Fig.\ \ref{fig:3}, except for
1970: two cases where ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$ diverges
1971: in the limit $R \to 0$ for finite $b_{l_{f}}$. In general the shape functions
1972: for finite $R$ are slightly smaller than in the case $R=0$.
1973: Final state effects
1974: for a finite value of $\gamma$ are less pronounced than in the
1975: case with $\gamma = 0$ but the differences is small.
1976: Therefore the dependence of the shape functions 
1977: in table \ref{tab:3}
1978: %(\ref{eq:sb1}) - (\ref{eq:sb6})
1979: on $b_{l_{f}}$ gives a reasonable impression about the importance
1980: of the final state interaction.
1981: For $E1$ transitions $p \to d$ or $d \to p$ a strong dependence
1982: of the shape function on $b_{l_{f}}$
1983: is found. The shape and magnitude of the reduced transition
1984: probability depends sensitively on the choice of $b_{l_{f}}$ and
1985: $R$.
1986: 
1987: The sensitivity of the shape functions to final-state effects
1988: on the neutron separation
1989: energy can be estimated from an expansion of 
1990: ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
1991: in terms of $\gamma$ by replacing the function $b_{l}$
1992: with the quantity $c_{l}$ as defined in equation (\ref{eq:spf}).
1993: This approach was introduced in \cite{Typ04a}.
1994: Analytical expressions for 
1995: dipole transitions are given by
1996: \begin{eqnarray}
1997:  {\mathcal S}_{1}^{0}(1) & = & 
1998:  \frac{x(3+x^{2})^{2}}{(1+x^{2})^{4}} 
1999:  \left[ 1 - \frac{4c_{0}}{3+x^{2}} \gamma + \dots \right] \: ,
2000:  \\
2001:  {\mathcal S}_{0}^{1}(1) & = &
2002:  \frac{4x^{3}}{(1+x^{2})^{4}}
2003:  \left[1 - c_{1}^{3}(1+3x^{2})\gamma^{3} + \dots \right] \: ,
2004:  \\
2005:  {\mathcal S}_{2}^{1}(1) & = &
2006:  \frac{x^{3}(5+3x^{2})^{2}}{(1+x^{2})^{4}} 
2007:  \left[ 1 - \frac{(1+6c_{1}^{3})(1+x^{2})^{2}}{5+3x^{2}}\gamma^{2}
2008:  + \dots \right] \: ,
2009:  \\
2010:  {\mathcal S}_{1}^{2}(1) & = &
2011:  \frac{4x^{5}}{(1+x^{2})^{4}} 
2012:  \left[ 1 - \frac{1+180c_{2}^{5}}{60}(1+x^{2})^{2}\gamma^{4} + \dots \right] 
2013:  \: .
2014: \end{eqnarray}
2015: These expansions are consistent with the results 
2016: in table \ref{tab:4}
2017: %(\ref{eq:s110x}) - (\ref{eq:s121x})
2018: in the limit $\gamma \to 0$ 
2019: ($c_{l} = b_{l}/\gamma$, see eq.\ (\ref{eq:spf})).
2020: There are corrections to the $E1$ shape functions 
2021: in table \ref{tab:3} %(\ref{eq:sdip_a})
2022: linear in $\gamma$ for $p \to s$ transitions and
2023: proportional to $\gamma^{3}$ for $s \to p$ transitions
2024: for finite values of $c_{l_{f}}$. For $d \to p$ and
2025: $p \to d$ transitions one finds corrections to the
2026: forms in table \ref{tab:3} %(\ref{eq:sdip_b}) 
2027: of order $\gamma^{2}$ and $\gamma^{4}$,
2028: respectively.
2029: They contain a contribution depending on the interaction
2030: ($c_{l_{f}}$) and a correction from the finite size of the system.
2031: Transitions $l \to l-1$ are affected 
2032: more strongly by the interaction
2033: than transitions $l \to l+1$.
2034: 
2035: Another approach to estimate the sensitivity 
2036: to the FSI is a comparison of the shape functions by scaling 
2037: $\gamma$ and $b_{l}$ appropriately.
2038: In Fig.~\ref{fig:4} the shape functions were shown for $\gamma=0.5$
2039: and constant $b_{l_{f}} \in [-0.5,0.5]$. Multiplying $\gamma$ by a factor of
2040: 2 corresponds to an increase of $S_{n}$ by a factor of 4. In order
2041: to cover the same range of scattering lengths $a_{l}$ as in Fig.~\ref{fig:4}
2042: also the interval for the function $b_{l}$ 
2043: has to be increased to $[-1.0,1.0]$ because the ratio $b_{l_{f}}/q$ has
2044: to be kept constant, cf.\ Eq.~(\ref{eq:abrel}). 
2045: The corresponding shape functions for $\gamma = 1.0$
2046: are depicted in Fig.~\ref{fig:5}. It is obvious that a higher 
2047: neutron separation energy leads to a much stronger dependence of
2048: the shape functions on the strength of the final-state interaction.
2049: This behaviour is easily understood from the radial dependence of the
2050: bound state wave function. For larger separation energy 
2051: the slope of the asymptotic wave function is much steeper since $q$ 
2052: increases. The radial integral is more sensitive to a small shift
2053: of the continuum wave function from a finite nuclear phase shift
2054: because the main contribution to the integrals arises from a
2055: smaller intervall in the radius on the surface of the nucleus.
2056: The effect is less pronounced for $s$ waves in the final state
2057: but increases dramatically for higher partial waves in the continuum.
2058: 
2059: \subsection{Shape functions in p+core systems}
2060: \label{subsec:p+core}
2061: 
2062: \begin{figure}
2063: \begin{center}
2064: \includegraphics[width=135mm]{shape5.ps}
2065: \end{center}
2066: \caption{\label{fig:1c} 
2067: Scaled shape functions $e^{4\eta_{i}} {\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
2068: as a function of $x^{2}=E_{bc}/S_{p}$ for $\gamma=0$ and
2069: values of the parameter $\eta_{i}$ between $0.0$ and $1.0$ in steps of
2070: $0.2$ without nuclear final-state interaction.}
2071: \end{figure}
2072: 
2073: \begin{figure}
2074: \begin{center}
2075: \includegraphics[width=135mm]{shape5e.ps}
2076: \end{center}
2077: \caption{\label{fig:1d} 
2078: Ratios $R_{l_{i}}^{l_{f}}(\lambda,\eta_{i})$
2079: as a function of $x^{2}=E_{bc}/S_{p}$ for $\gamma=0$ and
2080: values of the parameter $\eta_{i}$ between $0.0$ and $1.0$ in steps of
2081: $0.2$ without nuclear final-state interaction.}
2082: \end{figure}
2083: 
2084: The Coulomb interaction in p+core systems leads to a systematic 
2085: modification of the characteristic shape functions as compared to
2086: n+core systems. Since analytical results are not available in general one
2087: has to resort to a numerical integration of the radial integral
2088: (\ref{eq:hdef})
2089: if the dependence on arbitrary values of $x$ is studied.
2090: In Ref. \cite{Muk02} an analytic result for the radial integral was 
2091: obtained in the special case of scattering on a solid sphere 
2092: with given radius.
2093: For $x \ll 1$ it is possible to obtain analytical results
2094: in an approach with an expansion for small energies as
2095: presented in Refs.\ \cite{Jen98,Bay00}.
2096: For small $x$ one finds a suppression of the shape function
2097: approximately
2098: proportional to $\exp(-2\pi \eta_{f})$ with the
2099: Sommerfeld parameter $\eta_{f}$ of the scattering state.
2100: This scaling is 
2101: characteristic for a case with a Coulomb barrier.
2102: However, at larger energies the absolute value of the shape functions
2103: is determined by the Sommerfeld parameter $\eta_{i}$ of the 
2104: bound state since it defines the range of radii with the largest
2105: contribution to the radial integral. 
2106: In figure~\ref{fig:1c} the variation of the 
2107: shape functions with $\eta_{i}$ for $\gamma=0$
2108: and for various transitions is shown. The functions are scaled 
2109: with $\exp(4\eta_{i})$.
2110: Effects from the nuclear
2111: interaction in the final state are not taken into account.
2112: For $\eta_{i}=0$ the results for neutrons are recovered.
2113: With increasing $\eta_{i}$ the maximum of the shape function
2114: shifts to higher $x^{2}$, i.e.\ to higher relative energies.
2115: At the same time the width of the strength distribution increases
2116: and the absolute value of ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
2117: reduces considerably (note the scaling).
2118: 
2119: At low $x^{2}$ the suppression of the shape functions with
2120: finite $\eta_{i}$ as compared to the case $\eta_{i}=0$
2121: is quantitatively
2122: given by the factor $C_{l_{f}}^{2}(\eta_{f})$ as defined in
2123: equation (\ref{eq:coulfac}). For $x\to \infty$ corresponding
2124: to $\eta_{f}=\eta_{i}/x \to 0$
2125: we have $\lim_{\eta_{f} \to 0} C_{l_{f}}(\eta_{f})= 1$. 
2126: This suggests to scale the shape functions
2127: for the proton+core case with the penetrability factor
2128: $C_{l_{f}}^{-2}(\eta_{f})$.
2129: In figure \ref{fig:1d} the ratio
2130: \begin{equation}
2131:  R_{l_{i}}^{l_{f}}(\lambda,\eta_{i})
2132:  = \left[ C_{l_{f}}(\eta_{f}) \right]^{-2} 
2133:  \frac{{\mathcal S}_{l_{i}}^{l_{f}}(\lambda,\eta_{i})}{
2134:  {\mathcal S}_{l_{i}}^{l_{f}}(\lambda,0)}
2135: \end{equation}
2136: is depicted for various transitions and parameters $\eta_{i}$.
2137: The ratio only weakly depends on $x^{2}$. Therefore, the main
2138: difference between the proton+core and neutron+core cases
2139: is well described by the penetrability factor $C_{l_{f}}^{-2}(\eta_{f})$.
2140: 
2141: 
2142: There is a major difference between n+core and p+core nuclei.
2143: Due to the increasing Coulomb barrier the
2144: observation of a large transition strength at low relative
2145: energies will be lost for heavier p+core nuclei and the difference
2146: between transitions of different orbital angular momenta in the
2147: initial and final states will be become smaller. In contrast,
2148: even for heavy n+core nuclei the strong transition strength 
2149: at small energies will persist and the shape will still be
2150: characteristic of the particular transition.
2151: In principle, variations of the generic shape functions as shown
2152: in figure~\ref{fig:1c} with finite parameters $\gamma=qR$ and
2153: $b_{l}$ can be studied along the lines
2154: as for the n+core case. 
2155: Very similar trends are observed
2156: and we omit a detailed discussion.
2157: 
2158: 
2159: \section{Total transition strength and sum rules}
2160: \label{sec:totsum}
2161: 
2162: The total strength for an $E\lambda$ transition
2163: from a bound state with quantum numbers $J_{i}, s, j_{c}, l_{i}$
2164: to all possible final states
2165: is related by the non energy-weighted sum rule 
2166: \begin{eqnarray} \label{eq:newsr}
2167:  B(E\lambda, J_{i} s j_{c} l_{i}) & = &
2168:  \sum_{J_{f} l_{f}} \int_{0}^{\infty} dE \: 
2169:  \frac{dB}{dE}(E\lambda, J_{i} s j_{c} l_{i} \to 
2170:   k J_{f} s j_{c} l_{f})
2171:  \\ \nonumber & = &
2172:  \left[ Z_{\rm eff}^{(\lambda)}e \right]^{2}
2173:  \frac{2\lambda+1}{4\pi} \langle r^{2\lambda} \rangle_{l_{i}}
2174: \end{eqnarray}
2175: to the expectation value $\langle r^{2\lambda} \rangle_{l_{i}}$
2176: of the initial bound state with orbital angular momentum $l_{i}$,
2177: see, e.g., \cite{Jon04}.
2178: Note that for the integration over the energy $E$ not only 
2179: continuum states but also bound states
2180: have to be taken into account if they can be reached by an $E\lambda$
2181: transition from the initial bound state.
2182: In the special case of dipole transitions 
2183: the root-mean-square radius of the bound state wave function
2184: determines the total $E1$ strength. 
2185: From the scaling laws of $\langle r^{2} \rangle_{l_{i}}$
2186: as discussed
2187: in subsection \ref{subsec:prob} 
2188: we expect a divergence of the total transition
2189: strengths $B(E1, l_{i})\propto \gamma^{-2}$ and
2190: $B(E1, l_{i})\propto \gamma^{-1}$ for $l_{i}=0$ and
2191: $1$, respectively, in the limit $\gamma \to 0$. 
2192: In contrast, the total $E1$ strength should remain
2193: finite for larger orbital angular momenta of the bound  state
2194: wave function.
2195: 
2196: The total reduced transition probability to the 
2197: continuum is obtained in our approach with asymptotic wave functions
2198: from
2199: \begin{eqnarray} \label{eq:bel}
2200:  \lefteqn{B_{\rm cont} (E\lambda, J_{i} s j_{c}l_{i}  
2201:  \to J_{f} s j_{c} l_{f})=}
2202:  \\ \nonumber & & 
2203:    \left[Z_{\rm eff}^{(\lambda)}e\right]^{2} \sum_{J_{f} l_{f}}
2204:  \frac{2J_{f}+1}{2J_{i}+1} 
2205:  \left[  D_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda s j_{c}) \right]^{2}
2206:   \frac{\left|C^{j_{c}}_{J_{i}j_{i}l_{i}}\right|^{2}}{q^{2\lambda+1}}
2207:  {\mathcal T}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
2208: \end{eqnarray}
2209: with the dimensionless integral
2210: \begin{equation} \label{eq:tdef}
2211: {\mathcal T}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
2212:  = \frac{2}{\pi q^{2}} \int_{0}^{\infty} dk \: k \:
2213: {\mathcal S}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c})
2214:  = \frac{2}{\pi} \int_{0}^{\infty} dx \: \left|
2215: {\mathcal I}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(\lambda j_{c}) 
2216:  \right|^{2} \: .
2217: \end{equation}
2218: Evidently, the continuum contribution is only a fraction of the total strength
2219: if the FSI allows the occurrence of bound
2220: states in the corresponding channels.
2221: Since we use only the exterior contribution to the radial integral
2222: it is not guaranteed that the high energy behaviour of the
2223: reduced radial integrals (\ref{eq:iredgen}) is correct. Indeed,
2224: generally we do not find the $x^{-4}$ dependence for $E1$ transitions as
2225: expected from the general considerations in section \ref{subsec:xs}.
2226: For large $x$ there is a sizable contribution from the interior
2227: integral that is necessary to generate the correct $x$ dependence
2228: at high energies. However, the main contribution to the integral
2229: (\ref{eq:tdef}) arises at small $x$ close to the threshold
2230: and as long as the integral converges it will give a reasonable
2231: approximation of the total transition strength.
2232: 
2233: 
2234: \begin{table}
2235: \caption{\label{tab:5}Functions 
2236: ${\mathcal T}_{l_{i}}^{l_{f}}(\lambda)$ 
2237: for finite $\gamma$ and $b_{l_{f}} = 0$.}
2238: %\begin{center}
2239: \begin{tabular}{l}
2240:  \hline 
2241:  $\lambda = 0$ \\
2242: % monopole  \\
2243:  \hline 
2244:  $\displaystyle {\mathcal T}_{0}^{0}(0) = 
2245:  \exp(-2\gamma)/2$ \\
2246:  $\displaystyle {\mathcal T}_{1}^{1}(0) = 
2247:  (2+\gamma)\exp(-2\gamma)/(2\gamma)$ \\
2248:  \hline 
2249:  $\lambda = 1$ \\
2250: % dipole \\
2251:  \hline 
2252:  $\displaystyle {\mathcal T}_{1}^{0}(1) = 
2253:  (5+6\gamma+2\gamma^{2}) \exp(-2\gamma)/4$ \\
2254:  $\displaystyle  {\mathcal T}_{0}^{1}(1) = 
2255:  (1+2\gamma+\gamma^{2}) \exp(-2\gamma)/4 $\\
2256:  $\displaystyle {\mathcal T}_{2}^{1}(1) = 
2257:  (36+37\gamma+14\gamma^{2} +2\gamma^{3}) \exp(-2\gamma)/(4\gamma)$\\
2258:  $\displaystyle {\mathcal T}_{1}^{2}(1) = 
2259:  (5+6\gamma+2\gamma^{2}) \exp(-2\gamma)/4$\\
2260:  \hline 
2261:  $\lambda = 2$ \\
2262: % quadrupole \\
2263:  \hline 
2264:  $\displaystyle {\mathcal T}_{2}^{0}(2) = 
2265:  (63+90\gamma+54\gamma^{2}+16\gamma^{3}+2\gamma^{4}) 
2266:  \exp(-2\gamma)/4$ \\
2267:  $\displaystyle {\mathcal T}_{1}^{1}(2) = 
2268:  (7+14\gamma+14\gamma^{2}+8\gamma^{3}+2\gamma^{4}) \exp(-2\gamma)/4$ \\
2269:  $\displaystyle {\mathcal T}_{0}^{2}(2) = 
2270:  (3+6\gamma+6\gamma^{2}+4\gamma^{3}+2\gamma^{4}) \exp(-2\gamma)/4$ \\
2271:  \hline
2272: \end{tabular}
2273: %\end{center}
2274: \end{table}
2275: 
2276: 
2277: If one assumes that there is no nucleon-core interaction in the final states
2278: (i.e.\ plane waves)
2279: the full strength lies in the continuum.
2280: In this case the functions 
2281: ${\mathcal T}_{l_{i}}^{l_{f}}(\lambda)$ 
2282: can be calculated analytically for neutron+core systems in our approach.
2283: These functions are given in table \ref{tab:5}
2284: %by
2285: %\begin{eqnarray} \label{eq:gent1}
2286: % & & 
2287: % {\mathcal T}_{0}^{0}(0) = \frac{1}{2} e^{-2\gamma}\qquad
2288: % {\mathcal T}_{1}^{1}(0) = \frac{1}{2\gamma} (2+\gamma)e^{-2\gamma}
2289: % \\ & & 
2290: % {\mathcal T}_{1}^{0}(1) = \frac{1}{4}(5+6\gamma+2\gamma^{2})
2291: %  e^{-2\gamma} \qquad
2292: % {\mathcal T}_{0}^{1}(1) = \frac{1}{4}(1+2\gamma+\gamma^{2}) 
2293: %  e^{-2\gamma} 
2294: % \\ & & 
2295: % {\mathcal T}_{2}^{1}(1) = \frac{1}{4\gamma}(36+37\gamma+14\gamma^{2}
2296: % +2\gamma^{3}) e^{-2\gamma} 
2297: % \\ & & 
2298: % {\mathcal T}_{1}^{2}(1) = \frac{1}{4} (5+6\gamma+2\gamma^{2})
2299: % e^{-2\gamma}
2300: % \\ & & 
2301: % {\mathcal T}_{2}^{0}(2) = \frac{1}{4}
2302: % (63+90\gamma+54\gamma^{2}+16\gamma^{3}+2\gamma^{4}) e^{-2\gamma}
2303: % \\ & & 
2304: % {\mathcal T}_{1}^{1}(2) = \frac{1}{4} 
2305: % (7+14\gamma+14\gamma^{2}+8\gamma^{3}+2\gamma^{4}) e^{-2\gamma}
2306: % \\ & & 
2307: % {\mathcal T}_{0}^{2}(2) = \frac{1}{4} 
2308: % (3+6\gamma+6\gamma^{2}+4\gamma^{3}+2\gamma^{4}) e^{-2\gamma}
2309: %  \label{eq:gent4} 
2310: %\end{eqnarray}
2311: where we suppressed unimportant quantum numbers for simplification.
2312: The integrals ${\mathcal T}_{1}^{1}(0)$ and 
2313: ${\mathcal T}_{2}^{1}(1)$ diverge in the limit  $\gamma\to 0$
2314: but remain finite for $\gamma > 0$.  In the limit $\gamma \to \infty$, i.e.\
2315: if the bound state does not develop a halo, the functions become very small.
2316: 
2317: The total transition strength (\ref{eq:bel})
2318: is proportional to 
2319: $\left| C_{l_{i}} \right|^{2} {\mathcal T}_{l_{i}}^{l_{f}}(\lambda)
2320: /q^{2\lambda+1}$.
2321: The scaling of the ANC $C_{l_{i}}$ in the limit $\gamma \to 0$ can be
2322: obtained by considering the square-well model 
2323: (see appendix \ref{app:A}).
2324: We find 
2325: \begin{equation} \label{eq:slanc}
2326:  |C_{l_{i}}|^{2} \propto \left\{ 
2327: \begin{array}{lll}
2328:  q & \mbox{if} & l_{i} = 0 \\
2329:  q \gamma^{2l_{i}-1} & \mbox{if} & l_{i} \geq 1
2330: \end{array}
2331:  \right.
2332: \end{equation}
2333: for $\gamma \to 0$.
2334: Taking this dependence into consideration
2335: we find the scaling of $B(E1,l_{i})$
2336: consistent with the expectation from the non energy-weighted sum rule.
2337: 
2338: Now, let us consider the case with non-vanishing nuclear interaction
2339: in the continuum states. The FSI is parametrized in our approach by the
2340: function $b_{l_{f}}$ that in general depends on the energy $E$.
2341: If we assume that $b_{l_{f}}$ is constant for all energies
2342: the functions ${\mathcal T}_{l_{i}}^{l_{f}}(\lambda)$ can be
2343: calculated analytically again.
2344: One obtains, e.g.\, for the $E1$ $p \to s$ transition
2345: \begin{eqnarray} \label{eq:t101}
2346:  {\mathcal T}_{1}^{0}(1) & = &  e^{-2\gamma} 
2347:  \left\{ \begin{array}{lll}
2348:   \left[ \frac{5+6\gamma+2\gamma^{2}}{4} 
2349:  - 2b_{0}\frac{[1+2b_{0}+\gamma(1+b_{0})]^{2}}{(1+b_{0})^{4}}
2350: \exp\left( -\frac{2\gamma}{b_{0}}\right)\right]
2351:  & \mbox{for} & b_{0} \geq 0 \\
2352:  \frac{5+6\gamma+2\gamma^{2}}{4}  & \mbox{for} & b_{0} \leq 0 
2353:  \end{array} \right. 
2354:  \: .
2355: \end{eqnarray}
2356: For other transitions the expressions become quite complicated and
2357: we abstain from giving them here.
2358: Since the function
2359: $b_{l_{f}}$ cannot be assumed constant over the whole range of
2360: continuum energies one cannot expect that the calculated
2361: ${\mathcal T}_{l_{i}}^{l_{f}}(\lambda)$ corresponds to
2362: a reasonable result in general.
2363: However, it is sufficient to assume 
2364: that the function $b_{l_{f}}$ does not vary too much
2365: over the peak of the transition strength
2366: that is described by the shape
2367: function ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$.
2368: For large $|b_{0}|$ the excitation spectrum is distorted significantly
2369: in the relevant low-energy region.
2370: 
2371: Equation (\ref{eq:t101}) is quite instructive.
2372: It gives a good idea
2373: how the $B(E1)$ strength is modified when the depth $V_{0}$
2374: of the nuclear potential (with reasonable geometry) changes. 
2375: For $b_{0} > 0$ there is a reduction of the total strength
2376: in the continuum whereas in the case $b_{0} < 0$ no reduction
2377: is found. This behaviour is related to the occurrence of bound states
2378: in the $s$-wave continuum.
2379: For $V_{0}=0$ the nuclear phase shift $\delta_{0}$
2380: and the function $b_{0}$ is zero. 
2381: Increasing the potential depth 
2382: the phase shift at low energies becomes positive, corresponding to
2383: a negative scattering length. Increasing the depth $V_{0}$ further
2384: will lead to a more negative $b_{0}$ and a resonance will 
2385: appear at low energies. Then there comes the point
2386: where the potential will be able to support a bound state and
2387: $b_{0}$ suddenly jumps to a large positive value. Increasing the
2388: depth $V_{0}$ further will reduce the scattering length again.
2389: Finally, $b_{0}$ becomes zero and the cycle starts again.
2390: Considering this relation between strength of the potential
2391: and the scattering length it is reasonable to plot
2392: ${\mathcal T}_{1}^{0}(1)$ as a function of $1/b_{0}$ as depicted 
2393: in figure~\ref{fig:3c}. As long as the potential is not able to bind
2394: a state the continuum $B(E1)$ strength does not change. As soon
2395: as a resonance in the continuum becomes bound a sudden drop
2396: of the transition strength is observed that recovers to its
2397: initial value when $b_{0}$ approaches zero
2398: from positive values. 
2399: %ST added:
2400: An analogous effect was found for the $s \to p$ transition
2401: for the $E1$ excitation of ${}^{11}$Be in Ref.\ \cite{Typ04a}.
2402: %
2403: Similar observations are expected for other transitions.
2404: 
2405: \begin{figure}
2406: \begin{center}
2407: \includegraphics[width=135mm]{t101_g2.ps}
2408: \end{center}
2409: \caption{\label{fig:3c} 
2410: ${\mathcal T}_{1}^{0}(1)$
2411: as a function 
2412: of  $1/b_{0}$ for values of
2413: the parameter $\gamma$ between $0.0$ and $1.0$ in steps of $0.2$.}
2414: \end{figure}
2415: 
2416: This effect of the FSI has consequences for the extraction of the
2417: ANC or the spectroscopic factor for an assumed ground-state
2418: single-particle configuration from experimentally obtained
2419: total reduced transition probabilities. When the value of $B(E1,l_{i})$ 
2420: from the experiment is compared to the theoretical result assuming
2421: a single-particle model with
2422: plane waves in the continuum the extracted ANC or spectroscopic
2423: factor will be underestimated if the actual nuclear potential
2424: supports also bound states. 
2425: In the latter case, a part of the total transition strength 
2426: goes to these
2427: bound states, even if they cannot be reached by a transition of the
2428: halo nucleon because they are occupied by nucleons of the core.
2429: 
2430: Using the relation
2431: \begin{equation}
2432:  \sum_{f} (E_{f}-E_{i})\left| \langle f | r_{i} | i \rangle \right|^{2}
2433:  = \frac{1}{2}
2434:  \langle i | \left[ r_{i}, \left[ H, r_{i} \right] \right] | i \rangle
2435: \end{equation}
2436: the energy-weighted (or Thomas-Reiche-Kuhn) sum rule
2437: \begin{eqnarray} \label{eq:ewsr}
2438:  S(E1, J_{i} s j_{c} l_{i})
2439:  & = &  \frac{2\mu}{\hbar^{2}}
2440:  \sum_{J_{f} l_{f}} \int_{0}^{\infty} dE \: (E+S_{b}) \:
2441:  \frac{dB}{dE}(E1, J_{i} s j_{c} l_{i} \to 
2442:   k J_{f} s j_{c} l_{f})
2443:  \\ \nonumber & = & \frac{9}{4\pi} \left[ Z_{\rm eff}^{(1)} e \right]^{2}
2444: \end{eqnarray}
2445: for dipole transitions is obtained. 
2446: In constrast to the non energy-weighted sum rule (\ref{eq:newsr})
2447: it gives a result independent of the orbital angular
2448: momentum $l_{i}$ of the bound state. In principle, the sum contains
2449: all bound and unbound states that can be reached 
2450: from the initial state by an $E1$ transition.
2451: Since the photon energy is $E_{\gamma} = E_{f}-E_{i} = E+S$
2452: the sum rule (\ref{eq:ewsr}) can be expressed in terms of 
2453: the total $E1$ photo absorption cross section
2454: (\ref{eq:sigabs}) as
2455: \begin{eqnarray}
2456:  \lefteqn{\int_{0}^{\infty} dE_{\gamma} \: \sigma_{E1}(a+\gamma \to b+c)}
2457:  \\ \nonumber 
2458:  & = & \frac{2\pi^{2}\hbar}{\mu c} \left[ Z_{\rm eff}^{(1)} e \right]^{2} 
2459:  =  \frac{2\pi^{2}\hbar e^{2}}{m c} 
2460: \left( \frac{N_{a}Z_{a}}{A_{a}} - \frac{N_{b}Z_{b}}{A_{b}} 
2461:  - \frac{N_{c}Z_{c}}{A_{c}} \right)
2462: \end{eqnarray}
2463: for the disintegration into core+nucleon.
2464: It has the form of a cluster
2465: sum rule since we only consider the excitation of the halo nucleon but not
2466: of the core \cite{Alh82,Hen04}.
2467: In our approach we can write
2468: for the continuum contribution
2469: \begin{eqnarray}  \label{eq:ewsr2}
2470:  \lefteqn{S_{\rm cont} (E1, J_{i} s j_{c}l_{i})}
2471:  \\ \nonumber & = & 
2472:    \left[Z_{\rm eff}^{(1)}e\right]^{2} \sum_{J_{f} l_{f}}
2473:  \frac{2J_{f}+1}{2J_{i}+1} 
2474:  \left[  D_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 s j_{c}) \right]^{2}
2475:   \frac{\left|C^{j_{c}}_{J_{i}j_{i}l_{i}}\right|^{2}}{q}
2476:  {\mathcal U}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 j_{c})
2477: \end{eqnarray}
2478: with the dimensionless integral
2479: \begin{eqnarray} 
2480: {\mathcal U}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 j_{c})
2481:  & = & \frac{2}{\pi q^{4}} \int_{0}^{\infty} dk \: k \: 
2482:  (q^{2}+k^{2})\:
2483: {\mathcal S}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 j_{c})
2484:  \\ \nonumber 
2485:  & = & \frac{2}{\pi} \int_{0}^{\infty} dx \: (1+x^{2}) \: \left|
2486: {\mathcal I}_{J_{i}j_{i}l_{i}}^{J_{f}j_{f}l_{f}}(1 j_{c}) 
2487:  \right|^{2} \: .
2488: \end{eqnarray}
2489: If there are bound states in the final channel we expect
2490: $S_{\rm cont} (E1, J_{i} s j_{c}l_{i}) < S (E1, J_{i} s j_{c}l_{i})$.
2491: Here again, the remarks on the high energy behaviour of the integrand
2492: apply as for the functions (\ref{eq:tdef}). For the most important
2493: transition $s \to p$ 
2494: without interaction in the continuum states we obtain the finite result
2495: \begin{equation}
2496:   {\mathcal U}_{0}^{1}(1) = %{\mathcal U}_{1}^{0}(1) = 
2497:   e^{-2\gamma} \frac{3+2\gamma}{2}
2498: \end{equation}
2499: where we suppressed irrelevant quantum numbers.
2500: Considering the scaling law (\ref{eq:slanc}) of the ANC for $l_{i}=0$
2501: we find the correct result (\ref{eq:ewsr})
2502: for the  energy-weighted sum rule
2503: in the limit $\gamma \to 0$. For the ground state with $l_{i}=1$ there
2504: are two contributions from $l_{f}=0$ and $l_{f}=2$ in the final state.
2505: In this case the integral ${\mathcal U}_{1}^{2}(1)$ diverges with
2506: $\gamma^{-1}$ but this dependence is compensated by the scaling of the
2507: ANC. However, since the radial integrals 
2508: in the external approximation do not vanish fast enough with increasing $x$,
2509: the correct value for the energy-weighted sum rule 
2510: is not recovered in the limit $\gamma \to 0$.
2511:  
2512: \section{Examples for transition strengths of nucleon+core nuclei}
2513: \label{sec:examples}
2514: 
2515: Let us now look at some specific 
2516: examples of exotic nuclei with nucleon+core structure.
2517: The dependence of the reduced transition probability and of the 
2518: cross sections for electromagnetic transitions
2519: on the strength of the fragment-fragment
2520: interaction in the final state
2521: can be studied in the simple but realistic single-particle model of
2522: subsection~\ref{subsec:trans}. 
2523: For the nuclear potential a
2524: Woods-Saxon shape with
2525: the representative parameters $a=0.65$~fm and 
2526: $R_{0}=r_{0}A^{1/3}$ where $r_{0}=1.25$~fm is assumed.
2527: The depth of the
2528: potential in the bound state is adjusted to reproduce the experimental
2529: neutron and proton separation energies, respectively. In order to
2530: simulate a varying interaction strength in the final states 
2531: the transition strength is studied for 
2532: the depth of the Woods-Saxon potential in the continuum 
2533: in the range from 0 to
2534: 80~MeV. In Tables \ref{tab:B1} and \ref{tab:B2} the separation energy $S_{b}$, 
2535: potential radius $R_{0}$, potential depth $V_{0}$, and orbital angular
2536: momentum $l_{i}$ 
2537: for the ground state are given for the neutron+core nuclei 
2538: and proton+core nuclei, respectively. For simplicity, no spin-orbit
2539: potential is considered in the calculations.
2540: Also the characteristic parameter $\gamma$ is given in Tables
2541: \ref{tab:B1} and \ref{tab:B2}. Only nuclei with $\gamma <1$ can be
2542: considered to be halo nuclei, however, the boundary to nuclei without
2543: an extended nucleon distribution is smooth. For p+core nuclei
2544: the parameter $\eta_{i}$ is given in addition. A larger value
2545: indicates a stronger dominance of Coulomb distortions of the 
2546: $dB(E1)/dE$ strength.
2547: In the case of neutron+core nuclei $E1$ $s\to p$ and $d\to p$ transition
2548: will be considered. In the case of proton+core nuclei we will study
2549: %ST added: 
2550: especially
2551: %
2552: $E1$ $p \to s$ excitations.
2553: For many observables, a square-well model would
2554: also be a good approximation, especially for low
2555: energy processes where the ``shape independence'' 
2556: is valid to a large extent.  
2557: 
2558: \subsection{Neutron+core nuclei}
2559: 
2560: \begin{table}
2561: \caption{\label{tab:B1}Neutron 
2562: separation energy $S_{n}$, radius $R_{0}$, depth $V_{0}$
2563: of the Woods-Saxon potential and parameter $\gamma=qR_{0}$
2564: for the ground state with orbital angular momentum $l_{i}$
2565: of the  neutron+core nuclei in the single-particle model.
2566: The diffuseness parameter is $a=0.65$~fm.}
2567: \begin{center}
2568: \begin{tabular}{ccccccc}
2569:   \hline
2570:   &       & ${}^{11}$Be & ${}^{15}$C & ${}^{17}$O &
2571:             ${}^{23}$O  & ${}^{23}$O \\
2572:   \hline
2573:   $S_{n}$ & [MeV] & 0.504   & 1.218   & 4.140   & 
2574:                     2.740   & 2.740   \\
2575:   $R_{0}$ & [fm]  & 2.78    & 3.08    & 3.21    & 
2576:                     3.55    & 3.55    \\
2577:   $V_{0}$ & [MeV] & 54.3765 & 48.562  & 55.3726 & 
2578:                     42.416  & 43.1932 \\
2579:   $l_{i}$ &       & 0       & 0       & 2       & 
2580:                     0       & 2       \\
2581:   $\gamma$ &      & 0.413   & 0.722   & 1.393   &
2582:                     1.264   & 1.264   \\
2583:   \hline 
2584: \end{tabular}
2585: \end{center}
2586: \end{table}
2587: 
2588: The $\frac{1}{2}^{+}$ 
2589: ground state of ${}^{11}$Be is considered as a neutron in a
2590: $2s_{1/2}$ state coupled to the $0^{+}$ ground state of ${}^{10}$Be.
2591: The experimental neutron separation energy of $S_{n} = 0.504$ 
2592: MeV is
2593: quite small so that ${}^{11}$Be is a typical example of a halo nucleus.
2594: It was studied by Coulomb dissociation in the experiments 
2595: \cite{Nak94,Ann94,Pal03,Fuk04}.
2596: In Fig.~\ref{fig:6} the reduced transition probability 
2597: $dB(E1)/dE$ for a transition into $p$ waves is shown for 
2598: varying depths $V_{0}$ in the continuum between 0 and 80~MeV
2599: in steps of 5~MeV. The E1 strength shows a large peak at relative
2600: energies well below 1~MeV typical for a halo nucleus. The shape
2601: changes smoothly with increasing potential depth except for $V_{0}$
2602: close to 30~MeV where a $p$-wave resonance in the continuum becomes
2603: bound. For a potential depth similar to the one required for
2604: the correct separation energy in the ground state (cf. Tab.~\ref{tab:B1})
2605: the shape of the E1 transition strength is very similar to the
2606: result for a plane wave, i.e.\ $V_{0}=0$~MeV.
2607: Experimental data for the $dB(E1)/dE$ distribution in ${}^{11}$Be 
2608: were described in \cite{Typ04a} in our approach by fitting the
2609: asymptotic normalization coefficient and the constants $c_{1}^{3/2}$
2610: and $c_{1}^{1/2}$ in the effective-range expansion for the phase shifts
2611: in the $p$ waves with total angular momentum $j=3/2$ and $j=1/2$.
2612: An unnaturally large value for $c_{1}^{1/2}$ was found that is related 
2613: to the low separation energy of the first bound excited state in ${}^{11}$Be.
2614: The shape of the dipole strength function in the effective-range approach 
2615: was found to agree very well to a calculation in the Woods-Saxon model.
2616: A corresponding figure and more details can be found in Ref.\ \cite{Typ04a}.
2617: 
2618: 
2619: 
2620: \begin{figure}
2621: \begin{center}
2622: \includegraphics[width=120mm]{fig_Be11.ps}
2623: \end{center}
2624: \caption{\label{fig:6} 
2625: Reduced transition probability
2626: $dB(E1)/dE$ for the breakup of ${}^{11}$Be into a neutron and ${}^{10}$Be
2627: as a function of the c.m.\ energy $E$ 
2628: for various depths $V_{0}$ of the potential in the contiuum.
2629: The quantum numbers for the single particle bound state and the
2630: neutron separation energy in MeV are given in parenthesis.}
2631: \end{figure}
2632: 
2633: Another example of a neutron+core nucleus with the neutron in
2634: a $s$ wave ground state is ${}^{15}$C but with a larger neutron separation
2635: energy of $S_{n}=1.218$~MeV \cite{Dat03}. Comparing the dependence of 
2636: the dipole transition strength $dB(E1)/dE$ on the potential depth $V_{0}$
2637: as shown in Fig.~\ref{fig:7} with the case of ${}^{11}$Be one observes
2638: a stronger variation of the shape and absolute value. 
2639: This clearly shows the increased sensitivity of the transition
2640: strength on the final-state interaction when the neutron separation
2641: energy increases as expected from the analytical model in 
2642: subsection~\ref{subsec:fsinc}. Again,
2643: the reduced transition probability changes quite abruptly
2644: when a $p$ wave resonance in the continuum becomes a bound state.
2645: 
2646: \begin{figure}
2647: \begin{center}
2648: \includegraphics[width=120mm]{fig_C15.ps}
2649: \end{center}
2650: \caption{\label{fig:7}
2651: Same as Fig.~\ref{fig:6} but for the 
2652: breakup of ${}^{15}$C into a neutron and ${}^{14}$C.}
2653: \end{figure}
2654: 
2655: An even more drastic change of the transition strength is observed
2656: in the case of ${}^{23}$O when the neutron is assumed to be in a
2657: $s$-wave ground state, see Fig.~\ref{fig:8}. 
2658: This nucleus with a neutron separation 
2659: energy of $S_{n}=2.74$~MeV cannot be considered as a real halo nucleus.
2660: Yet, it can be said that there is substantial low-energy 
2661: $E1$ strength.
2662: 
2663: \begin{figure}
2664: \begin{center}
2665: \includegraphics[width=120mm]{fig_O23s.ps}
2666: \end{center}
2667: \caption{\label{fig:8}
2668: Same as Fig.~\ref{fig:6} but for the 
2669: breakup of ${}^{23}$O in a $s$-wave ground state
2670: into a neutron and ${}^{22}$O.}
2671: \end{figure}
2672: 
2673: \begin{figure}
2674: \begin{center}
2675: \includegraphics[width=120mm]{fig_O23d.ps}
2676: \end{center}
2677: \caption{\label{fig:9}
2678: Same as Fig.~\ref{fig:6} but for the 
2679: breakup of ${}^{23}$O in a $d$-wave ground state
2680: into a neutron and ${}^{22}$O.}
2681: \end{figure}
2682: 
2683: Comparing the three cases ${}^{11}$Be, ${}^{15}$C, and ${}^{23}$O
2684: one observes a shift of the 
2685: peak in the transition strength to higher relative energies in the
2686: continuum. The position of the peak scales with the separation energy
2687: as expected from the dependence of the shape functions on the
2688: ratio $x^{2}=S_{n}/E$ in Section~\ref{sec:rrisf}. Additionally, one finds
2689: a reduction of the overall strength as suggested from
2690: Eqs.~(\ref{eq:dbelde}) and (\ref{eq:bel}) 
2691: with an increasing value of the ground state 
2692: parameter $q$.
2693: 
2694: Since the ground state angular momentum of ${}^{23}$O is not known
2695: uniquely from experiments \cite{Noc04,Kan01,Kan02,Bro03,Tho03,Sau04,Cor04} 
2696: it is also possible that the
2697: neutron in the ground state occupies a $d_{5/2}$ single-particle state.
2698: The strength function for an $E1$ transition to $p$-wave final
2699: states under this assumption is shown in Fig.~\ref{fig:9}.
2700: The absolute magnitude is about a factor 4 smaller than in the case
2701: of a $s$-wave neutron in the ground state and the maximum is
2702: shifted to larger relative energy. This is explained by 
2703: the additional centrifugal barrier in the $d$ wave that reduces
2704: the probability of finding the neutron at large distances from the
2705: core in the classically forbidden region.
2706: 
2707: \begin{figure}
2708: \begin{center}
2709: \includegraphics[width=120mm]{fig_O17.ps}
2710: \end{center}
2711: \caption{\label{fig:10}
2712: Same as Fig.~\ref{fig:6} but for the 
2713: breakup of ${}^{17}$O into a neutron and ${}^{16}$O.}
2714: \end{figure}
2715: 
2716: A further example for a neutron+core nucleus with a $d$-wave 
2717: ground-state configuration is ${}^{17}$O, a nucleus close to the valley of
2718: stability. Here the separation energy
2719: of $S_{n}=4.14$~MeV is even larger than in the case of ${}^{23}$O
2720: and the shape and magnitude of the reduced $E1$ transition probability
2721: to $p$ waves in the final state vary drastically with an increasing
2722: strength of the final-state interaction.
2723: 
2724: \begin{figure}
2725: \begin{center}
2726: \includegraphics[width=135mm]{BE1.ps}
2727: \end{center}
2728: \caption{\label{fig:11} Total reduced transition probability $B(E1)$
2729: for the breakup into a neutron+core with c.m. energies $E$
2730: between 0 and 10~MeV as a function of the potential
2731: depth $V_{0}$ for the continuum states. The filled circle gives
2732: the result for the same potential depth as for the single particle
2733: bound state.
2734: The quantum numbers for the single particle bound state and the
2735: neutron separation energy for each system are given in parenthesis.}
2736: \end{figure}
2737: 
2738: In Fig.~\ref{fig:11} the total $E1$ transition strength integrated
2739: from 0 to 10~MeV is shown for the nuclei discussed above as a function
2740: of the potential depth $V_{0}$ in the continuum. There is
2741: a distinctive difference between nuclei with a $s$-wave and $d$-wave
2742: neutron in the ground state. In the former case the integrated strength
2743: stays almost constant when $V_{0}$ is increased starting at 0~MeV,
2744: i.e.\ the plane-wave result. Then there is a sudden drop of the total
2745: $E1$ strength when the $p$-wave resonance from the continuum becomes
2746: a bound state. However, the total strength recovers to its value
2747: for $V_{0}=0$ when the potential depth is increased further
2748: until the next higher $p$-wave resonance becomes bound. 
2749: The dependence of the total transition probability on the strength
2750: of the FSI is
2751: in qualitative agreement with the expectations from the analytical
2752: neutron+core model as shown in figure \ref{fig:3c}.
2753: The rise
2754: of the total $E1$ strength after the continuum to bound state
2755: crossing is faster when the neutron separation energy is smaller.
2756: For halo nuclei there is only a small variation of the total
2757: reduced transition probability if one is not too close to a potential
2758: strength where the crossing appears. When the neutron is in the
2759: $d$-wave ground state there is already a considerable
2760: dependence of the total $E1$ strength on the depth of the 
2761: potential close to the plane-wave limit. Beyond the sudden drop of the 
2762: strength at the continuum-to-bound-state crossing the
2763: absolute value increases again but it does not reach the plane-wave
2764: limit again. Comparing the total $E1$ strength calculated
2765: for a continuum potential that is identical to the bound state
2766: potential (filled circle in Fig.~\ref{fig:11}) one immediately
2767: finds that the integrated $E1$ strength is almost the same as
2768: in the plane-wave case for a $s$-wave neutron in the ground state
2769: but it is significantly smaller for a $d$-wave neutron.
2770: 
2771: \subsection{Proton+core nuclei}
2772: 
2773: \begin{table}
2774: \caption{\label{tab:B2}Proton 
2775: separation energy $S_{p}$, radius $R_{0}$, depth $V_{0}$
2776: of the Woods-Saxon potential, parameters $\gamma = qR_{0}$
2777: and $\eta_{i}$
2778: for the ground state with orbital angular momentum $l_{i}$
2779: of the  proton+core nuclei in the single-particle model.
2780: The diffuseness parameter is $a=0.65$~fm.}
2781: \begin{center}
2782: \begin{tabular}{ccccc}
2783:   \hline
2784:   &       & ${}^{8}$B  & ${}^{9}$C   & ${}^{12}$N \\
2785:   \hline
2786:   $S_{p}$ & [MeV] & 0.1375  & 1.296   & 0.601  \\
2787:   $R_{0}$ & [fm]  & 2.50    & 2.60    & 2.86   \\
2788:   $V_{0}$ & [MeV] & 43.183  & 44.4456 & 36.474 \\
2789:   $l_{i}$ &       & 1       & 1       & 1      \\
2790:   $\gamma$ &      & 0.190   & 0.613   & 0.466  \\
2791:   $\eta_{i}$ &    & 1.595   & 0.655   & 1.171  \\
2792:   \hline 
2793: \end{tabular}
2794: \end{center}
2795: \end{table}
2796: 
2797: Systems with proton+core structure show similar features as compared
2798: to neutron+core systems, however, with modifications due to the
2799: appearence of the 
2800: Coulomb barrier. Since analytical results for the
2801: shape functions are not available one has to resort to the numerical
2802: calculations, e.g.\ in the simple single-particle model.
2803: 
2804: 
2805: As examples we consider the three unstable nuclei 
2806: ${}^{8}$B \cite{Xu94,Tra01,Azh01,Oga03b,Tra03,Jun02,Sch03}, 
2807: ${}^{9}$C \cite{Bea01,Tra02,End03} and
2808: ${}^{12}$N \cite{Lef95,Tan03,Tim03}. 
2809: Their ground state is well described by a proton in
2810: a $p$-wave bound state. The proton wave function is calculated in
2811: a single-particle model.  
2812: %ST modified:
2813: The depth of the potentials was adjusted to reproduce the experimental 
2814: binding energies as in the case of the neutron+core nuclei. 
2815: The corresponding parameters of
2816: the Woods-Saxon potential are given in table~\ref{tab:B2}.
2817: %
2818: As a consequence the ANC of the bound state 
2819: wave functions is uniquely determined in this model.
2820: 
2821: \begin{figure}
2822: \begin{center}
2823: \includegraphics[width=120mm]{fig_B8.ps}
2824: \end{center}
2825: \caption{\label{fig:13} 
2826: Reduced transition probability
2827: $dB(E1)/dE$ for the breakup of ${}^{8}$B into a proton and ${}^{7}$Be
2828: into a $s$ wave continuum state
2829: as a function of the c.m.\ energy $E$ 
2830: for various depths $V_{0}$ of the potential in the contiuum.
2831: The proton separation energy in MeV is given in parenthesis.}
2832: \end{figure}
2833: 
2834: \begin{figure}
2835: \begin{center}
2836: \includegraphics[width=120mm]{fig_N12.ps}
2837: \end{center}
2838: \caption{\label{fig:14} 
2839: Same as Fig.~\ref{fig:13} but for the 
2840: breakup of ${}^{12}$N into a proton and ${}^{11}$C.}
2841: \end{figure}
2842: 
2843: \begin{figure}
2844: \begin{center}
2845: \includegraphics[width=120mm]{fig_C9.ps}
2846: \end{center}
2847: \caption{\label{fig:15} 
2848: Same as Fig.~\ref{fig:13} but for the 
2849: breakup of ${}^{9}$C into a proton and ${}^{8}$B.}
2850: \end{figure}
2851: 
2852: The reduced transition probability $dB(E1)/dE$ as a function of
2853: the relative energy $E$ is shown
2854: in figures \ref{fig:13}, \ref{fig:14}, and \ref{fig:15} for these
2855: nuclei for various
2856: depths of the continuum potential. In the calculation
2857: it was assumed that the channel spin $S$ of the ground state is
2858: given by $S=j_{c}+1/2$ with the core spin $j_{c}$.
2859: The  additional Coulomb interaction between nucleon and core leads to a
2860: substantial modification of the shape functions.
2861: The transition strength still peaks at low relative energies,
2862: however, the maximum is shifted to higher energies 
2863: as compared to the neutron+core case.
2864: For a pure Coulomb wave in the continuum state, i.e.\ $V_{0}=0$,
2865: the maximum is reached at an energy of $E/S_{p}= 4.28$, $2.57$ and
2866: $1.26$, for ${}^{8}$B, ${}^{12}$N, and ${}^{9}$C,
2867: respectively. The larger shift of the maximum to higher energies 
2868: corresponds to an increase of the parameter $\eta_{i}$,
2869: cf.\ Table \ref{tab:B2}.
2870: In the neutron+core case the maximum is expected
2871: at a value of only $0.18$. 
2872: The overall shape of the transition strength
2873: changes similarly with the depth of the potential
2874: as for neutron+core nuclei and the occurence of resonances in
2875: the $p$-wave continuum is observed again. 
2876: Also the integrated reduced transition probability, as depicted in figure
2877: \ref{fig:16}, resembles in its dependence on the potential depth the 
2878: results for neutron+core systems between the two cases of
2879: $s \to p$ and $d \to p$ transitions. The $B(E1)$ value assuming
2880: the same potential in the continuum as in the bound state is clearly
2881: smaller than the value for pure Coulomb waves in the continuum.
2882: 
2883: \begin{figure}
2884: \begin{center}
2885: \includegraphics[width=135mm]{BE1b.ps}
2886: \end{center}
2887: \caption{\label{fig:16} Reduced transition probability $B(E1)$
2888: for the breakup into a proton+core with c.m. energies $E$ between
2889: 0 and 10~MeV as a function of the potential
2890: depth $V_{0}$ for the continuum states. The filled circle gives
2891: the result for the same potential depth as for the single particle
2892: bound state.
2893: The proton separation energy for each system is given in parenthesis.}
2894: \end{figure}
2895: 
2896: 
2897: 
2898: \section{Low-energy behaviour and ANC method}
2899: \label{sec:anc}
2900: 
2901: At small relative energies the reduced transition probability
2902: $dB(E\lambda)/dE \propto {\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$
2903: is strongly suppressed 
2904: due to penetration effects of the centrifugal barrier. It is useful to
2905: change to a different quantity to study the low-energy
2906: behaviour and the effects of the
2907: nucleon-core interaction in the continuum. 
2908: Considering the approximation of the shape functions
2909: for small $x$ in table \ref{tab:4}
2910: %equations (\ref{eq:sb1}) - (\ref{eq:sb6})  
2911: one can use
2912: \begin{equation}
2913:  \tilde{\mathcal S}_{l_{i}}^{l_{f}}(\lambda) 
2914:  = x^{-2l_{f}-1} {\mathcal S}_{l_{i}}^{l_{f}}(\lambda)
2915: \end{equation}
2916: with a finite limit for $x \to 0$
2917: in the case of neutron+core systems.
2918: This quantitiy 
2919: takes the angular momentum barrier into account. 
2920: Correspondingly the quantities 
2921: \begin{equation}
2922:  \tilde{\mathcal S}_{l_{i}}^{l_{f}}({\rm abs},\lambda)
2923:  =x^{-2l_{f}-1}
2924:  {\mathcal S}_{l_{i}}^{l_{f}}({\rm abs},\lambda)
2925:  = x^{-2l_{f}-1} (1+x^{2})^{2\lambda-1}{\mathcal S}_{l_{i}}^{l_{f}}(\lambda)
2926: \end{equation}
2927: and 
2928: \begin{equation} \label{eq:scaptmod}
2929:  \tilde{\mathcal S}_{l_{i}}^{l_{f}}({\rm capt},\lambda) 
2930:  = x^{-2l_{i}+1}
2931:  {\mathcal S}_{l_{i}}^{l_{f}}({\rm capt},\lambda)
2932:  = x^{-2l_{i}-1}(1+x^{2})^{2\lambda+1}
2933:   {\mathcal S}_{l_{f}}^{l_{i}}(\lambda)
2934: \end{equation} 
2935: are the relevant forms with a finite value for $x \to 0$
2936: for photon absorption and radiative capture,
2937: respectively,
2938: cf.\ equations (\ref{eq:sabs}) and (\ref{eq:scapt}).
2939: (Note that for the capture reaction the initial and final state 
2940: are interchanged.)
2941: The low-energy behaviour is clearly dominated by the centrifugal barrier
2942: in the continuum state. From equation (\ref{eq:scaptmod})
2943: one finds the famous $1/v \propto 1/x$ law of the cross section
2944: for the neutron capture from a $s$ wave
2945: in the continuum $l_{i}=0$.
2946: The above quantities approach a finite value in the limit $E\to 0$
2947: that depends on the function $b_{l_{f}}$.
2948: Thus, the interaction in the continuum 
2949: has an effect on the absolute normalization of the low-energy cross
2950: sections.
2951: 
2952: Apart from the typical $x^{2l_{f}+1}$ dependence the expressions
2953: for ${\mathcal S}_{l_{i}}^{l_{f}}(\lambda)$ contain a factor
2954: $(1+x^{2})^{2\lambda+2}$ 
2955: in the denominator that corresponds to a pole at 
2956: $x^{2}=-1$. Considering equation (\ref{eq:scaptmod}) one sees
2957: that the shape function for capture reactions 
2958: has a simple pole at the separation energy $E=-S_{n}$.
2959: This pole limits the range of convergence
2960: of an expansion of the shape functions or cross sections
2961: in terms of the parameter $x$ or the energy $E$ \cite{Bay04}.
2962: Since the separation energy is very small for halo systems a
2963: corresponding expansion has only limited applicability in this case.
2964: 
2965: A similar effect is observed for proton+core systems \cite{Jen98,Bay00}.
2966: Here, the low-energy behaviour is dominated by the Coulomb barrier
2967: in the continuum state. In nuclear astrophysics one needs, e.g., 
2968: cross sections for radiative proton capture at very low energies. 
2969: Instead of extrapolating the strongly energy dependent capture
2970: cross section $\sigma_{\pi \lambda}(p+c \to a+\gamma)$ to low energies 
2971: the astrophysical S factor
2972: \begin{equation}
2973:  S(E) = \sigma_{\pi \lambda}(p+c \to a+\gamma) E \exp(2\pi\eta_{i})
2974: \end{equation}
2975: is employed. It is weakly dependent on energy and approaches a finite
2976: value for $E \to 0$.
2977: The exponential term depending on the Sommerfeld
2978: parameter $\eta_{i}$ in the continuum state cancels in leading order
2979: the suppression of the capture cross section due to the Coulomb barrier.
2980: The factor $E$ is proportional to the $k^{2}$ factor in the theorem
2981: of detailed balance (\ref{eq:detbal}) from the phase space in the continuum.
2982: Considering the scaling of the shape functions for
2983: proton+core systems in section \ref{subsec:p+core}
2984: the generalitzation of (\ref{eq:scaptmod}) is found.
2985: The quantity
2986: \begin{equation} 
2987:  \tilde{\mathcal S}_{l_{i}}^{l_{f}}({\rm capt},\lambda) 
2988:  = C_{l_{i}}^{2}(\eta_{i}) x^{2}
2989:  {\mathcal S}_{l_{i}}^{l_{f}}({\rm capt},\lambda)
2990:  =  C_{l_{i}}^{2}(\eta_{i}) (1+x^{2})^{2\lambda+1}
2991:   {\mathcal S}_{l_{f}}^{l_{i}}(\lambda)
2992: \end{equation} 
2993: approaches a finite value for $x \to 0$ because
2994: $C_{l_{i}}^{2}(\eta_{i}) \to \eta_{i}^{2l_{i}+1} \exp(-2\pi\eta_{i})$ 
2995: for $\eta_{i} = \eta_{f}/x \to \infty$ with constant $\eta_{f}$.
2996: It follows that
2997: \begin{eqnarray}
2998:  S(E) & \propto & \tilde{\mathcal S}_{l_{i}}^{l_{f}}({\rm capt},\lambda) 
2999:  \propto \frac{{\rm const.}}{1+x^{2}}
3000: \end{eqnarray}
3001: for $x\to 0$.
3002: For small separation energies of the proton in the bound state
3003: a strong increase of the S factor at small energies is observed due to
3004: the closeness of the pole at $E=-S_{p}$.
3005: In general, the absolute value of $S(0)$ depends on the ANC of the ground state
3006: and the scattering length, i.e.\ the strength of the continuum interaction.
3007: 
3008: An interesting case is the direct $E1$ radiative capture from the 
3009: p + ${}^{16}$O
3010: continuum to the first excited ($1/2^{+}$) state in ${}^{17}$F
3011: that is relevant to nuclear astrophysics \cite{Bru96,Mor97}.
3012: This state with $l=0$ and a small proton
3013: separation energy of 105~keV can be considered as a typical halo state.
3014: %ST added:
3015: Assuming a radius of $R=3.15$~fm we find $\gamma=0.217$ and
3016: $\eta_{i}=3.787$.
3017: %
3018: At low energies the $p \to s$ transition gives the main contribution 
3019: to the astrophysical S factor (apart from the small contribution of the
3020: $p \to d$ transition to the $5/2^{+}$ ground state with $S_{p} = 600$~keV).
3021: Due to the closeness of the pole at $E=-S_{p}$ a strong rise of the
3022: astrophysical S factor is observed for $E \to 0$ that is practically
3023: independent of the potential in the continuum.
3024: 
3025: 
3026: \begin{figure}
3027: \begin{center}
3028: \includegraphics[width=135mm]{sfac.ps}
3029: \end{center}
3030: \caption{\label{fig:12} S factor $S(E,V_{0})$
3031: at energy $E=0$ as a function of the
3032: depth $V_{0}$ of a Woods-Saxon potential in the continuum 
3033: scaled to the S factor calculated with pure Coulomb waves for three
3034: radiative capture reactions. The numbers in parenthesis are the 
3035: separation energies of the proton. The filled circle shows
3036: the result assuming the same potential depth in the continuum
3037: as for the bound state.}
3038: \end{figure}
3039: 
3040: The sensitivity of the low-energy S factor to the interaction in the
3041: continuum can be
3042: estimated by comparing $S(0)$ for various p+core systems
3043: as a function of the interaction strength.
3044: In figure~\ref{fig:12} the
3045: dependence of the S factor at zero energy is shown for the
3046: reactions (a) ${}^{7}$Be(p,$\gamma$)${}^{8}$B,
3047: (b) ${}^{11}$C(p,$\gamma$)${}^{12}$N, and
3048: (c) ${}^{8}$B(p,$\gamma$)${}^{9}$C.
3049: In all cases we have an $E1$ $s \to p$ transition.
3050: The S factor
3051: is calculated with varying depths of the Woods-Saxon potential
3052: in the continuum $s$ waves. Spin-orbit contribution to the potential
3053: are neglected for simplicity. 
3054: For an easier comparision all zero-energy S factors are divided by
3055: the zero-energy S factor for pure Coulomb waves, i.e.\ $V_{0}=0$
3056: for the nuclear potential. These are given by
3057: (a) $22.07$~eV~b, (c) $129.4$~eV~b, and (b) $68.41$~eV~b
3058: in the present model.
3059: As is clearly seen in figure~\ref{fig:12}
3060: the nuclear proton-core interaction
3061: directly has an effect on the zero-energy S factor. There
3062: is a considerable variation, especially if the depth $V_{0}$ approaches
3063: a value where an $s$-wave continuum state becomes a bound state, e.g.\ at 
3064: $V_{0}=16.3$~MeV and $71.6$~MeV in case of the
3065: ${}^{7}$Be(p,$\gamma$)${}^{8}$B reaction. Furthermore, we find that
3066: the sensitivity to the potential in the continuum state increases
3067: with the separation energy of the proton. Assuming the same depth
3068: for the continuum potential as for the bound state a S factor is 
3069: obtained that is somehow smaller, but close the value from a 
3070: calculation with pure Coulomb waves.
3071: 
3072: Our results indicate that the zero-energy S factor
3073: is not uniquely determined by the ground state ANC 
3074: that can be extracted from
3075: experiments.
3076: In general it is necessary to
3077: consider the nuclear interaction in the continuum that can be 
3078: parametrized at small energies by the scattering lengths in the
3079: relevant partial waves \cite{Bay04,Bay00}. 
3080: One example of particular interest is
3081: the ${}^{7}$Be(p,$\gamma$)${}^{8}$B reaction where experimental
3082: data from direct and indirect experiments for the S factor
3083: have to be extrapolated to zero energy assuming a certain
3084: energy dependence from theoretical models \cite{Dav03}. 
3085: For the ${}^{7}$Be+p system $s$-wave scattering
3086: lengths $a_{S}$ for channel spin S=2 and S=1 (from coupling the spins of
3087: the proton and ${}^{7}$Be) have been determined experimentally,
3088: however with large uncertainties \cite{Ang03}.
3089: More precise values are available
3090: for the mirror system ${}^{7}$Li+n \cite{Koe83}. 
3091: The scattering lengths are easily
3092: reproduced in the single-particle model by adjusting the depth of the
3093: Woods-Saxon potential in the continuum. As is seen clearly in figure
3094: 3 of Ref.\ \cite{Dav03} the energy dependence of the S factor shows
3095: a considerable variation that affects the extrapolation of the
3096: experimental data to zero energy.
3097: 
3098: 
3099: 
3100: \section{Conclusions and Outlook}
3101: 
3102: In this paper we study single-particle aspects
3103: of neutron- and proton-halo nuclei,
3104: i.e.\ the effects which appear when the separation energy 
3105: of the least-bound nucleon tends to zero. 
3106: These nuclear systems
3107: display universal features, 
3108: e.g.\ for the electromagnetic transition strength,
3109: that can be described in simple models.
3110: 
3111: We use different approaches in our theoretical studies:
3112: one is a square-well model, for which we obtain remarkably simple
3113: analytical results. A more realistic and conventional 
3114: approach is the model with Woods-Saxon potentials.
3115: For our applications to halo nuclei
3116: we find that it is, even quantitatively,
3117: quite similar to the square-well model. 
3118: The underlying reason is quite simple:
3119: For the low energies relevant for the halo systems
3120: there is shape independence: the properties of the 
3121: potential are encoded in some low-energy parameters
3122: which do not depend on the details of the chosen specific 
3123: potential. Halo nuclei are low-energy phenomena, related to large 
3124: wavelengths. It is well known that
3125: a probe of a given wavelength is insensitive to details of 
3126: structure at distances much smaller than this
3127: wavelength \cite{Lep97}. This means that we can mimic the 
3128: real short-distance structure (e.g.\ determined from
3129: the nuclear many-body problem or a Woods-Saxon
3130: single-particle model) by a simple short distant structure,
3131: e.g.\ a square well model with its agreable analytical properties. 
3132: The depth and radius parameters serve as adjustable 
3133: parameters that, e.g., reproduce the binding
3134: energy and the scattering length.
3135: 
3136: We mention also that there is a need to 
3137: supplement full scale ab-initio microscopic 
3138: approaches to nuclear structure by some empirical
3139: fine tuning of parameters to reproduce the actual
3140: position of loosely bound single particle structure,
3141: see e.g.\ \cite{Sag01}. This is indispensible
3142: in order to obtain a satisfactory description of
3143: low-lying electromagnetic strength.
3144: 
3145: Following a study of some general single-particle properties 
3146: we consider the radial matrix elements for electric excitations to the
3147: continuum.
3148: They essentially depend on the asymptotic wave functions of the
3149: bound and scattering states.
3150: We identify the important parameters and find 
3151: rather simple analytical formulae for   
3152: the most relevant cases for neutron
3153: halo nuclei. The most important parameter is 
3154: $\gamma=qR$ depending on the separation energy $S = \hbar^{2}q^{2}/(2\mu)$
3155: and the size $R$ of the system.  It is shown that $\gamma$
3156: serves as a convenient expansion parameter.
3157: Proton halo nuclei are studied numerically.
3158: We find simple scaling laws for the transition
3159: strength. We also discuss typical
3160: examples for light halo nuclei in the Woods-Saxon model.
3161: 
3162: We quantitatively analyze the effects of the 
3163: final-state interaction in the continuum on the transition strength.
3164: We find that it is determined mainly by the 
3165: scattering length. In general, this nucleon-core
3166: interaction has an influence on the radiative capture
3167: process down to very small energies.
3168: This can serve as a warning for the simple
3169: application of the ANC method: the results are rather insensitive to the
3170: potential in the continuum only for true halo systems.
3171: 
3172: The power of the present approach is 
3173: demonstrated by recent applications to actual nuclei \cite{Typ04a,Noc04}.
3174: It will serve also as a framework for further applications,
3175: notably for the upcoming radioactive beam facilities RIKEN,
3176: FAIR/GSI, RIA. One important conclusion is that we quantitatively 
3177: understand halo effects. It is expected from our
3178: studies that neutron-halo effects will show up
3179: irrespective of the mass number $A$, whereas 
3180: proton halo effects will tend to disappear
3181: with increasing charge number $Z$.
3182: 
3183: We hope that methods similar to ones developed in this paper
3184: will also be useful for the (much) more complicated
3185: problems of two-, three, and more nucleon halo nuclei
3186: \cite{Rii00,Jen04}.
3187: %ST: modified
3188: One early discussion of that problem is 
3189: \cite{Mig73}. A review on modern approaches in context of effective
3190: field theories and references are found in \cite{Bra04}.
3191: %
3192: 
3193: 
3194: %(Perhaps it would be rewarding to compare
3195: %these simple models to more sophisticated
3196: %direct capture calculations, see e.g. 
3197: %Thomas Rauscher and Klaus Gruber Phys.Rev. C66
3198: %(2002)028802. the problem: how to determine
3199: %the parameters V and a....)
3200: %It is applicable to low as well as high energies.
3201: 
3202: 
3203: \section{Acknowledgments}
3204: 
3205: We greatfully acknowledge discussions with
3206: T. Aumann, U. Datta Pramanik, D. Baye, H. Emling, G. Hansen, and C. Nociforo.
3207: 
3208: \section*{Note added in proof:}
3209: In the meantime a paper ``Low-lying dipole strength for weakly bound
3210: systems: A simple analytic estimate'' \cite{Nag05} appeared that contains 
3211: results which are related to this paper.
3212: 
3213: \appendix 
3214: \section{Neutron in a square-well potential}
3215: \label{app:A}
3216: 
3217: The problem of a neutron in a square-well potential is well studied
3218: for bound and scattering states
3219: in the literature, see e.g.\ \cite{Mes61}, and many analytic results
3220: have been obtained. However, often only results for 
3221: a few selected cases are presented.
3222: Here, formulas for arbitrary values of the orbital angular
3223: momentum are given by generalizing earlier calculations
3224: and simplifying the expressions. 
3225: %ST added:
3226: We also derive results that, to our knowledge, were not quoted in the
3227: literature before.
3228: %
3229: 
3230: 
3231: \subsection{Wave functions and probabilities}
3232: \label{app:A1}
3233: 
3234: The bound-state wave function
3235: \begin{equation}
3236:  \Phi_{nlm}^{i}(\vec{r}) = \frac{f_{nl}(r)}{r} Y_{lm} (\hat{r})
3237: \end{equation}
3238: for a neutron with orbital angular momentum $l$ and projection $m$
3239: in a square-well potential of Radius $R$ and
3240: depth $V_{n}$ is completely characterized by the separation
3241: energy $S>0$ and the principal quantum number $n=1,2,\dots$ that counts the
3242: number of nodes (including the node at $r=0$). 
3243: The radial wave function
3244: \begin{equation}
3245:  f_{nl}(r) = \left\{ \begin{array}{ll}
3246:  A_{nl} r j_{l}(\bar{q}_{nl} r) & \mbox{for} \quad r \leq R \\
3247:  B_{nl} r i^{l} h^{(1)}_{l}(iq r) & \mbox{for} \quad r \geq R
3248:  \end{array} \right.
3249: \end{equation}
3250: is expressed in terms of spherical Bessel and Hankel functions
3251: $j_{l}$ and $h^{(1)}_{l}$ \cite{Abr65}. The quantities
3252: $q = \sqrt{2\mu S}/\hbar$
3253: and
3254: $\bar{q}_{nl} = \sqrt{2\mu(V_{nl}-S)}/\hbar$
3255: in the arguments are related by
3256: \begin{equation}
3257:  q^{2} + \bar{q}_{nl}^{2} = \frac{2\mu V_{nl}^{i}}{\hbar^{2}}
3258: \end{equation}
3259: where an increasing number of radial nodes $n$ requires
3260: larger values of  $\bar{q}_{nl}$ and of the depth $V_{nl}^{i}$ 
3261: of the potential in the ground state.
3262: The relation between $q$ and $\bar{q}_{nl}$ for given $n$
3263: is determined by the continuity of the logarithmic derivative
3264: \begin{equation}
3265:  L_{nl}^{i} = \left.
3266:  \frac{r\frac{d}{dr}f_{nl}(r)}{f_{nl}(r)} \right|_{r=R}
3267: \end{equation}
3268: of the radial wave function $\phi_{nl}(r)$ at $r=R$. This condition
3269: can be written as
3270: \begin{equation}
3271:  L_{nl}^{i} = l+1 - Y_{l}^{(+)} = Y_{l}^{(-)} - l
3272: \end{equation}
3273: with
3274: \begin{equation}
3275:  Y_{l}^{(\pm)} = 
3276:  \bar{\gamma}_{nl} 
3277:  \frac{j_{l\pm 1}(\bar{\gamma}_{nl})}{j_{l}(\bar{\gamma}_{nl})} = 
3278:  i \gamma \frac{h^{(1)}_{l\pm 1}(i\gamma)}{h^{(1)}_{l}(i\gamma)} 
3279: \end{equation}
3280: where
3281: \begin{equation} \label{eq:vi}
3282:  \gamma = q R \: ,\quad \bar{\gamma}_{nl} = \bar{q}_{nl} R \quad
3283:  \mbox{and} \quad
3284:  \gamma^{2} + \bar{\gamma}_{nl}^{2} 
3285:  = \frac{2\mu R^{2} V_{nl}^{i}}{\hbar^{2}}  = v_{i} \: .
3286: \end{equation}
3287: The (complex) constants $A_{nl}$ and $B_{nl}$ 
3288: are determined by the continuity
3289: condition for the wave function
3290: \begin{equation} \label{eq:cont}
3291:   A_{nl} j_{l}(\bar{\gamma}_{nl}) = B_{nl} i^{l} h^{(1)}_{l}(i\gamma)
3292: \end{equation}
3293: and the normalization condition
3294: \begin{equation}
3295:  1 = P^{<}_{nl} + P^{>}_{nl} 
3296: \end{equation}
3297: with
3298: \begin{equation}
3299:  P^{<}_{nl} = 
3300:  \left|A_{nl}\right|^{2} 
3301:  \int_{0}^{R} dr \: r^{2} \: \left[j_{l}(\bar{q}_{nl} r) \right]^{2}
3302: \end{equation}
3303: and
3304: \begin{equation}
3305:  P^{>}_{nl} = 
3306:  \left|B_{nl}\right|^{2} \int_{R}^{\infty} 
3307:  dr \: r^{2} \: \left|h^{(1)}_{l}(iq r) \right|^{2} \: .
3308: \end{equation}
3309: These radial integrals are easily evaluated (see Appendix \ref{app:E})
3310: with the results
3311: \begin{eqnarray}
3312:  P^{<}_{nl} & = & 
3313:   \frac{|B_{nl}|^{2}}{2}R^{3}  \left| h^{(1)}_{l}(i\gamma) \right|^{2}
3314:    \left( 
3315:  \frac{\gamma^{2}}{\bar{\gamma}_{nl}^{2}} X_{l} + 1 \right)
3316: \end{eqnarray}
3317: and
3318: \begin{eqnarray}
3319:  P^{>}_{nl} & = & 
3320:    \frac{|B_{nl}|^{2}}{2}R^{3}  \left| h^{(1)}_{l}(i\gamma) \right|^{2}
3321:    \left( X_{l} - 1  \right)
3322: \end{eqnarray}
3323: where the continuity equation (\ref{eq:cont}) was used.
3324: The quantity
3325: \begin{equation}
3326:  X_{l} = -\frac{Y_{l}^{(-)}Y_{l}^{(+)}}{\gamma^{2}}
3327:  = \frac{h^{(1)}_{l-1}(i\gamma)
3328:  h^{(1)}_{l+1}(i\gamma)}{\left[ h^{(1)}_{l}(i\gamma) \right]^{2}} 
3329:  = \frac{K_{l-1/2}(\gamma)
3330:    K_{l+3/2}(\gamma)}{\left[K_{l+1/2}(\gamma)\right]^{2}}
3331: \end{equation}
3332: is a rational function in the variable $\gamma$.
3333: From the ratio 
3334: \begin{equation}
3335:  \frac{P^{<}_{nl}}{P^{>}_{nl}} =  
3336:  \frac{\gamma^{2}X_{l}+\bar{\gamma}_{nl}^{2}}{\bar{\gamma}_{nl}^{2}
3337:  \left[X_{l}-1\right]} 
3338: \end{equation}
3339: the probability of finding the neutron at radii $r \leq R$ 
3340: \begin{equation}
3341:  P_{nl} = \frac{P^{<}_{nl}}{P^{<}_{nl}+P^{>}_{nl}} 
3342:  =  \frac{\gamma^{2} X_{l} + \bar{\gamma}_{nl}^{2}}{\left(\bar{\gamma}_{nl}^{2}
3343: + \gamma^{2} \right) X_{l}}
3344: \end{equation}
3345: is obtained.
3346: In the limit $\gamma \to 0$ we have
3347: \begin{equation}
3348:  \lim_{\gamma \to 0} X_{l} = \left\{ 
3349: \begin{array}{ll}
3350:  \infty & \mbox{for} \quad l=0 \\
3351:  \frac{2l+1}{2l-1}& \mbox{for} \quad l>0
3352: \end{array}\right.
3353: \end{equation}
3354: and 
3355: \begin{equation}
3356:  \lim_{\gamma \to 0} P_{nl} = \left\{ 
3357: \begin{array}{ll}
3358:  0 & \mbox{for} \quad l=0 \\
3359:  \frac{2l-1}{2l+1}& \mbox{for} \quad l>0
3360: \end{array}\right. 
3361: \end{equation}
3362: independent of the principal quantum number $n$.
3363: Equivalent expressions for neutrons 
3364: were given recently for $l=0,1,2$ 
3365: in Ref.~\cite{Liu04}.
3366:  
3367: In case of a scattering state the wave function has the form
3368: \begin{equation}
3369:  \Phi_{lm}^{f}(\vec{r})
3370:   = 4\pi i^{l} \frac{g_{l}(r)}{kr} Y_{lm}(\hat{r}) Y_{lm}^{\ast}(\hat{k})
3371: \end{equation}
3372: for energy $E=\hbar^{2}k^{2}/(2\mu)$
3373: with the radial wave function  
3374: \begin{equation}
3375:  g_{l}(r) = \left\{ \begin{array}{ll} 
3376:  \bar{A}_{nl} \bar{k}_{nl} r j_{l}(\bar{k}_{nl}r) & \mbox{for} \quad r\leq R
3377:  \\
3378:  \frac{1}{2i} 
3379:  \left[ \exp(2i\delta_{l}) u_{l}^{(+)}(kr) - u_{l}^{(-)}(kr) \right]
3380:  & \mbox{for} \quad r\geq R
3381:  \end{array} \right.
3382: \end{equation}
3383: where 
3384: \begin{equation}
3385:  u_{l}^{(\pm)} (x) = x \left[-y_{l}(x) \pm i j_{l}(x) \right]
3386: \end{equation}
3387: and $j_{l}$ and $y_{l}$ are
3388: spherical Bessel and Neumann functions, respectively \cite{Abr65}.
3389: The phase shift is denoted by $\delta_{l}$ and 
3390: $\bar{k}_{nl} = \sqrt{2\mu(E+V_{nl})}/\hbar$
3391: where $V_{nl}$ is the depth of the square-well potential
3392: that gives the correct separation energy $S$ of the neutron.
3393: The continuity condition for the logarithmic derivative
3394: \begin{eqnarray}
3395:  L_{nl}^{f}  & = &
3396:  \left. \frac{r \frac{d}{dr} g_{l}(r)}{g_{l}(r)} \right|_{r=R}
3397:  = 1 + \frac{\bar{\kappa}_{nl}
3398:  j_{l}^{\prime}(\bar{\kappa}_{nl})}{j_{l}(\bar{\kappa}_{nl})}
3399:  \\ \nonumber & = & 
3400:  \frac{\kappa\left[ \exp(2i\delta_{l}) u_{l}^{(+)\prime}(\kappa)
3401:  - u_{l}^{(-)\prime}(\kappa)\right]}{\exp(2i\delta_{l})u_{l}^{(+)}(\kappa)
3402:  - u_{l}^{(-)}(\kappa)}
3403: \end{eqnarray}
3404: at $r=R$ with 
3405: \begin{equation}
3406:  \kappa = kR \qquad \bar{\kappa}_{nl} = \bar{k}_{nl} R
3407: \end{equation}
3408: determines the phase shift $\delta_{l}$ from
3409: \begin{equation}
3410:  \exp(2i\delta_{l}) = 
3411:  \frac{L_{nl}^{f}u_{l}^{(-)}(\kappa)-\kappa u_{l}^{(-)\prime}
3412:  (\kappa)}{L_{nl}^{f}u_{l}^{(+)}(\kappa)-\kappa 
3413:  u_{l}^{(+)\prime}(\kappa)} \: .
3414: \end{equation}
3415: %or
3416: %\begin{equation}
3417: % \tan(\delta_{nl}) = 
3418: % \frac{(L_{nl}^{f}-1)j_{l}(\kappa)
3419: %   -\kappa j_{l}^{\prime}(\kappa)}{(L_{nl}^{f}-1)
3420: %   y_{l}(\kappa)-\kappa y_{l}^{\prime}(\kappa)} 
3421: % = \frac{j_{l}(\kappa)}{y_{l}(\kappa)}
3422: %  \frac{L_{nl}^{f}-l-1+\kappa 
3423: % \frac{j_{l+1}(\kappa)}{j_{l}(\kappa)}}{L_{nl}^{f}
3424: %  +l - \kappa \frac{y_{l-1}(\kappa)}{y_{l}(\kappa)}}  
3425: % \: .
3426: %\end{equation}
3427: The phase shift is a sum
3428: \begin{equation}
3429:  \delta_{l} = \tau_{l} + \rho_{l}
3430: \end{equation}
3431: of the hard-sphere phase shift $\tau_{l}$ with
3432: \begin{equation}
3433:  \exp(2i\tau_{l}) = \frac{u_{l}^{(-)}(\kappa)}{u_{l}^{(+)}(\kappa)}
3434: % \quad \mbox{or} \quad
3435: % \tan \left( \tau_{l} \right)
3436: % = \frac{j_{l}(\kappa)}{y_{l}(\kappa)}
3437: \end{equation}
3438: and the additional phase shift $\rho_{l}$ with
3439: \begin{equation}
3440:  \exp(2i\rho_{l}) = 
3441:  \frac{L_{nl}^{f}-q_{l}^{(-)}(\kappa)}{L_{nl}^{f}
3442:  -q_{l}^{(+)}(\kappa)}
3443: % \quad \mbox{or} \quad
3444: % \tan \left( \rho_{l} \right) 
3445: % = \frac{{\rm Im}\left[q_{l}^{(+)}(\kappa)\right]}{{\mathcal L}_{nl}
3446: % -{\rm Re}\left[q_{l}^{(+)}(\kappa)\right]}
3447: \end{equation}
3448: where
3449: \begin{equation}
3450:  q_{l}^{(\pm)}(\kappa) = 
3451:  \frac{\kappa u_{l}^{(\pm)\prime}(\kappa)}{u_{l}^{(\pm)}(\kappa)}
3452:  \: .
3453: \end{equation}
3454: The relation
3455: \begin{equation} \label{eq:vf}
3456:   \bar{\kappa}_{nl}^{2} - \kappa^{2} =
3457:  \frac{2\mu V_{nl}^{f}R^{2}}{\hbar^{2}} = 
3458:   v_{f}
3459: \end{equation}
3460: with the depth $V_{nl}^{f}$ of the potential in the
3461: scattering state
3462: determines the quantity $\bar{\kappa}_{nl}$. 
3463: Generally, $V_{nl}^{i}$ in (\ref{eq:vi}) can be different from
3464: $V_{nl}^{f}$.
3465: In the limit $k \to 0$ we have
3466: \begin{equation}
3467:  \tan(\delta_{nl}) \to - a_{l} k^{2l+1}
3468: \end{equation}
3469: with the scattering length
3470: \begin{equation} \label{eqn:scalen_l}
3471:  a_{l} = a_{l}^{hs} 
3472:  \left( 1 - \frac{2l+1}{L_{nl}^{f}(0)+l}\right)
3473: \end{equation}
3474: where
3475: \begin{equation}
3476:  a_{l}^{hs} = \left\{ \begin{array}{lll}
3477:  R & \mbox{if} & l=0 \\
3478:  \frac{R^{2l+1}}{(2l+1)!!(2l-1)!!} & \mbox{if} & l>0
3479: \end{array} \right. 
3480: \end{equation}
3481: is the scattering length of a hard sphere of radius $R$, 
3482: i.e.\ $g_{l}(R)=0$ corresponding to
3483: $L_{nl}^{f} \to \infty$.
3484: 
3485: Defining the penetrability
3486: \begin{equation}
3487:  P_{l}(x) = \left| u_{l}^{(\pm)}(x) \right|^{-2}
3488:  = x^{-2} \left[y_{l}^{2}(x) + j_{l}^{2}(x)\right]^{-1}
3489: \end{equation}
3490: we can write
3491: \begin{equation}
3492:  u_{l}^{(\pm)} (x) = P_{l}^{-\frac{1}{2}}(x) 
3493:  \exp\left[ \mp i \tau_{l}(x)\right]
3494: \end{equation}
3495: with the hard-sphere phase shift $\tau_{l}$.
3496: Then the continuity of the wave function at $r=R$
3497: \begin{eqnarray} \label{eq:contf}
3498:  \bar{A}_{nl} \bar{\kappa}_{nl} j_{l}(\bar{\kappa}_{nl}) & = &
3499: \frac{1}{2i} 
3500:  \left[ \exp(2i\delta_{l}) u_{l}^{(+)}(\kappa) - u_{l}^{(-)}(\kappa) \right] 
3501:  \\ \nonumber & = & 
3502:   P_{l}^{-\frac{1}{2}}(\kappa) \:
3503:   \exp(i\delta_{l}) \:
3504:   \sin \left( \delta_{l}-\tau_{l}\right)
3505: \end{eqnarray}
3506: fixes the constant $\bar{A}_{nl}$. 
3507: The differential probability of finding the neutron
3508: inside the square-well potential of radius $R$
3509: is given by
3510: \begin{eqnarray}
3511:  \frac{dP_{nl}}{dk} & = & 
3512:  k^{2} \int \frac{d\Omega_{k}}{(2\pi)^{3}} 
3513:  \int d\Omega_{r} \int_{0}^{R} dr \: r^{2} \: \left| \Phi_{lm}^{f} \right|^{2}
3514:  \\ \nonumber & = & 
3515:  \frac{R}{\pi}
3516:   \left| \bar{A}_{nl} \right|^{2} \bar{\kappa}_{nl}^{2}
3517:  \left(\left[j_{l}(\bar{\kappa}) \right]^{2} 
3518:  - j_{l-1}(\bar{\kappa}) j_{l+1}(\bar{\kappa})\right)
3519:  \\ \nonumber & = & 
3520:  R
3521:  \frac{\sin^{2}\left( \delta_{l}
3522:  -\tau_{l}\right)}{\pi P_{l}(\kappa)}
3523:  \left( 1 - \frac{(L_{nl}^{f}+l)}{\bar{\kappa}_{nl}} 
3524:  \frac{(l+1-L_{nl}^{f})}{\bar{\kappa}_{nl}} \right)
3525:  \: .
3526: \end{eqnarray}
3527: Using the continuity relations for the logarithmic derivative
3528: one obtains
3529: \begin{eqnarray} \label{eq:pnlcont}
3530:  \frac{dP_{nl}}{dx} & = & 
3531:  \frac{\gamma}{4\pi}
3532:  \left(  \left| \exp(i\delta_{l}) u_{l}^{(+)}(\kappa) 
3533: - \exp(-i\delta_{l}) u_{l}^{(-)}(\kappa) \right|^{2} 
3534:  \right. \\ \nonumber & & \left.
3535:  - \frac{\kappa^{2}}{\bar{\kappa}_{nl}^{2}} 
3536:  [\exp(i\delta_{l}) u_{l-1}^{(+)}(\kappa) 
3537: - \exp(-i\delta_{l}) u_{l-1}^{(-)}(\kappa)]
3538:  \right. \\ \nonumber & & \left.
3539:  [\exp(i\delta_{l}) u_{l+1}^{(+)}(\kappa) 
3540: - \exp(-i\delta_{l}) u_{l+1}^{(-)}(\kappa)]^{\ast}
3541:  \right)
3542:  \\ \nonumber & = & 
3543:  \frac{\gamma}{\pi}
3544:  \left( \frac{\sin^{2}\left( \delta_{l}
3545:  -\tau_{l}\right)}{P_{l}(\kappa)}
3546:  - \frac{\kappa^{2}}{\bar{\kappa}_{nl}^{2}} 
3547:  \frac{\sin\left( \delta_{l}
3548:  -\tau_{l-1}\right)\sin\left( \delta_{l}
3549:  -\tau_{l+1}\right)}{\sqrt{P_{l-1}(\kappa)
3550:  P_{l+1}(\kappa)}}
3551:  \right)
3552:  \: .
3553: \end{eqnarray}
3554: The probability is determined mainly by 
3555: the penetrabilities $P_{l,l\pm 1}$ and the differences of the
3556: phase shift $\delta_{l}$ from the hard-sphere values $\tau_{l,l\pm 1}$.
3557: 
3558: \subsection{Root-mean-square radius}
3559: \label{app:A2}
3560: 
3561: The root-mean-square radius of the bound state wave function
3562: is obtained from
3563: \begin{equation}
3564:  \langle r^{2} \rangle = \frac{R_{2}}{R_{0}}
3565: \end{equation}
3566: with
3567: \begin{equation}
3568:   R_{n} = \int_{0}^{\infty} dr \: r^{n+2} \: \left| \phi_{nl}(r) \right|^{2}
3569: \end{equation}
3570: With the help of the integrals in appendix \ref{app:E} and the continuity
3571: equation for the wave function and the logarithmic derivative we find
3572: \begin{eqnarray}
3573:  R_{0} & = & 
3574:  \left| A_{nl} \right|^{2} \int_{0}^{R} dr \: r^{2} \: 
3575:  \left[ j_{l}(\bar{q}_{nl}r)\right]^{2} 
3576:  + \left| B_{nl} \right|^{2} \int_{R}^{\infty} dr \: r^{2} \: 
3577:  \left| h^{(1)}_{l}(iqr)\right|^{2} 
3578:  \\ \nonumber & = & 
3579:  \frac{\left| A_{nl} \right|^{2}}{\bar{q}_{nl}^{3}} 
3580: \frac{\bar{\gamma}_{nl}}{2} \left[ j_{l}(\bar{\gamma}_{nl}) \right]^{2} 
3581:  \left( \bar{\gamma}_{nl}^{2}
3582:  -  Y_{l}^{(-)} Y_{l}^{(+)}
3583:  \right)
3584:  \\ \nonumber & & 
3585:  - (-1)^{l} \frac{\left| B_{nl} \right|^{2}}{(iq)^{3}} 
3586: \frac{i\gamma}{2} \left[ h^{(1)}_{l}(i\gamma) \right]^{2} 
3587:  \left( [i\gamma]^{2}
3588:    -   Y_{l}^{(-)} Y_{l}^{(+)}\right)
3589:  \\ \nonumber & = & 
3590:  - \frac{R^{3}}{2} \left| B_{nl} \right|^{2}
3591:  \left| h^{(1)}_{l}(i\gamma) \right|^{2}\left(
3592:  \frac{1}{\bar{\gamma}_{nl}^{2}} + \frac{1}{\gamma^{2}}
3593:  \right) Y_{l}^{(-)} Y_{l}^{(+)} 
3594:  \: .
3595: \end{eqnarray}
3596: Similarly
3597: we obtain
3598: \begin{eqnarray}
3599:  R_{2} & = & 
3600:  \left| A_{nl} \right|^{2} \int_{0}^{R} dr \: r^{4} \: 
3601:  \left[ j_{l}(\bar{q}_{nl}r)\right]^{2} 
3602:  + \left| B_{nl} \right|^{2} \int_{R}^{\infty} dr \: r^{4} \: 
3603:  \left| h^{(1)}_{l}(iqr)\right|^{2} 
3604:  \\ \nonumber & = & 
3605:  \frac{\left| A_{nl} \right|^{2}}{\bar{q}_{nl}^{5}} 
3606: \frac{\bar{\gamma}_{nl}}{12} [j_{l}(\bar{\gamma}_{nl})]^{2} 
3607:  \left( 3 \bar{\gamma}_{nl}^{4}
3608:  -2 \bar{\gamma}_{nl}^{2}
3609:  Y_{l}^{(-)} Y_{l}^{(+)}
3610:  \right.
3611:  \\ \nonumber & & \left.
3612:  -  \left\{(2l-1)Y_{l}^{(-)} 
3613:  -\bar{\gamma}_{nl}^{2} \right\}
3614:  \left\{(2l+3)Y_{l}^{(+)} 
3615:  -\bar{\gamma}_{nl}^{2} \right\}
3616:  \right) 
3617:  \\ \nonumber & & 
3618:  - (-1)^{l} \frac{\left| B_{nl} \right|^{2}}{(iq)^{5}} 
3619: \frac{i\gamma}{12} [h^{(1)}_{l}(i\gamma)]^{2} \left( 3 (i\gamma)^{4}
3620:  -2 (i\gamma)^{2} Y_{l}^{(-)} Y_{l}^{(+)}
3621:  \right.
3622:  \\ \nonumber & & \left.
3623:  -  \left\{(2l-1)Y_{l}^{(-)} 
3624:  -(i\gamma)^{2} \right\}
3625:  \left\{(2l+3) Y_{l}^{(+)} 
3626:  -(i\gamma)^{2} \right\}
3627:  \right) 
3628:  \\ \nonumber & = & 
3629:  \left| B_{nl} \right|^{2} \left|h^{(1)}_{l}(i\gamma)\right|^{2} 
3630:  \frac{R^{5}}{12}
3631:  \left(\frac{1}{\bar{\gamma}_{nl}^{2}}+\frac{1}{\gamma^{2}} \right) 
3632:  \left[-2 Y_{l}^{(-)} Y_{l}^{(+)}
3633: \right. 
3634:  \\ \nonumber & & 
3635:  + \left\{ (2l+3) Y_{l}^{(+)} 
3636:  + (2l-1) Y_{l}^{(-)}  
3637:  \right\}
3638:  \\ \nonumber & & \left.
3639:  - \left(\frac{1}{\bar{\gamma}_{nl}^{2}} 
3640:  - \frac{1}{\gamma^{2}} \right) \left\{
3641:  (2l-1)(2l+3) Y_{l}^{(-)} Y_{l}^{(+)}
3642:  \right\}  \right] \: .
3643: \end{eqnarray}
3644: The ratio $R_{2}/R_{0}$ gives
3645: \begin{eqnarray}
3646:  \langle r^{2} \rangle_{l} & = & 
3647:     \frac{R^{2}}{6} \left[2
3648:  -  \frac{2l+3}{Y_{l}^{(-)}}
3649:  - \frac{2l-1}{Y_{l}^{(+)}} 
3650:  + \left(\frac{1}{\bar{\gamma}_{nl}^{2}} - \frac{1}{\gamma^{2}} \right) 
3651:  (2l-1)(2l+3) \right] \: .
3652: \end{eqnarray}
3653: We note
3654: \begin{equation} \label{eq:recur}
3655:  Y_{l}^{(-)} Y_{l-1}^{(+)} = - \gamma^{2}
3656:  \quad \mbox{and} \quad
3657:  Y_{l}^{(-)} + Y_{l}^{(+)} = 2l+1
3658: \end{equation}
3659: and find
3660: \begin{eqnarray}
3661:  \langle r^{2} \rangle_{l} & = & 
3662:     \frac{R^{2}}{6} \left[2
3663:  + \frac{2l+3}{Y_{l-2}^{(+)}}
3664:  - \frac{2l-1}{Y_{l}^{(+)}} 
3665:  + \frac{(2l-1)(2l+3)}{\bar{\gamma}_{nl}^{2}}
3666:   \right]
3667: \end{eqnarray}
3668: Explicit expressions for $Y_{l}^{(\pm)}$ are found from the recursion
3669: relations (\ref{eq:recur})
3670: with 
3671: \begin{equation}
3672:  Y_{0}^{(+)} = 1+\gamma \qquad
3673:  Y_{0}^{(-)} = - \gamma
3674: \end{equation}
3675: for $l=0$.
3676: We have
3677: \begin{equation}
3678:  \langle r^{2} \rangle_{0} = 
3679:  \frac{R^{2}}{6} \left[ \frac{3+2\gamma}{1+\gamma}
3680:  + \frac{3}{\gamma} + \frac{3}{\gamma^{2}} 
3681:  - \frac{3}{\bar{\gamma}_{nl}^{2}}\right]
3682: \end{equation}
3683: and
3684: \begin{equation}
3685:  \langle r^{2} \rangle_{1} = 
3686:  \frac{R^{2}}{6} \left[ \frac{5+5\gamma+2\gamma^{2}}{3+3\gamma+\gamma^{2}}
3687:  + \frac{5}{\gamma} 
3688:  + \frac{5}{\bar{\gamma}_{nl}^{2}}\right]
3689: \end{equation}
3690: for the two lowest orbital angular momenta.
3691: This leads to the divergences
3692: \begin{equation}
3693:  \langle r^{2} \rangle_{0} \to 
3694:  \frac{R^{2}}{2\gamma^{2}} 
3695: \quad \mbox{and} \quad
3696:   \langle r^{2} \rangle_{1} \to
3697:  \frac{5R^{2}}{6\gamma} 
3698: \end{equation}
3699: in the limit $\gamma \to 0$.
3700: With the asymptotic behaviour
3701: \begin{equation}
3702:  Y_{l}^{(+)} \to 2l+1
3703:  \quad \mbox{for} \quad l \geq 0
3704: \end{equation}
3705: for $\gamma \to 0$
3706: we obtain
3707: \begin{eqnarray}
3708:   \langle r^{2} \rangle_{l} & \to & 
3709:      \frac{(2l-1)(2l+3)}{6} R^{2} \left[  
3710:  \frac{2}{(2l-3)(2l+1)} + \frac{1}{\bar{\gamma}_{nl}^{2}} 
3711:   \right] 
3712: \end{eqnarray}
3713: for higher orbital angular momenta. There is no divergence
3714: of the root-mean-square radius in the limit $\gamma \to 0$.
3715: It approaches a finite value.
3716: 
3717: \subsection{Asymptotic normalization coefficient}
3718: \label{app:A3}
3719: 
3720: The modulus of the quantity $B_{nl}$ 
3721: is obtained from the normalization condition
3722: \begin{equation}
3723:  1 = P_{<}(nl) + P_{>}(nl) = 
3724:  (-1)^{l} \frac{\left|B_{nl}\right|^{2}}{2}R^{3}  
3725:    \left(1 + \frac{\gamma^{2}}{\bar{\gamma}_{nl}^{2}} \right)
3726:  h^{(1)}_{l-1}(i\gamma)h^{(1)}_{l+1}(i\gamma) 
3727: \end{equation}
3728: It is related via $B_{nl}=q C_{nl} = \gamma C_{nl}/R$ to the ANC $C_{nl}$.
3729: Thus we find for the ANC
3730: \begin{equation}
3731:  C_{nl} = 
3732:  \sqrt{2q} \left[ (-1)^{l} \gamma^{3}
3733:    \left(1 + \frac{\gamma^{2}}{\bar{\gamma}_{nl}^{2}} \right)
3734:  h^{(1)}_{l-1}(i\gamma)h^{(1)}_{l+1}(i\gamma) \right]^{-\frac{1}{2}}
3735: \end{equation}
3736: In the case $l=0$ one has explicitly
3737: \begin{equation}
3738:  C_{n0} =
3739:  \sqrt{2q} \exp(\gamma) \left[
3740:    \left(1 + \frac{\gamma^{2}}{\bar{\gamma}_{n0}^{2}} \right)
3741:  (1+\gamma) \right]^{-\frac{1}{2}}
3742: \end{equation}
3743: and
3744: \begin{equation}
3745:  \lim_{\gamma \to 0} C_{n0} = \sqrt{2q} \: .
3746: \end{equation}
3747: For $l=1$ we find
3748: \begin{equation}
3749:  C_{n1} \to \sqrt{\frac{2q}{3}} \: \gamma^{1/2} 
3750: \end{equation}
3751: and for $l>1$
3752: \begin{equation}
3753:  C_{nl} \to \sqrt{\frac{2q}{(2l-3)!!(2l+1)!!}}  \:
3754:  \gamma^{l-1/2} 
3755: \end{equation}
3756: if $\gamma$ approaches zero.
3757: 
3758: \subsection{Transition integrals}
3759: \label{app:A4}
3760: 
3761: Using the notation for the radial wave functions of appendix \ref{app:A1},
3762: the interior and exterior contribution to the radial transition integral
3763: (\ref{eq:radint}) are given by
3764: \begin{eqnarray}
3765:  I_{l_{i}}^{l_{f}}(\lambda,<) & = &
3766: % \int_{0}^{R} dr \: r^{2} \:
3767: % A_{nl_{i}} j_{l_{i}}(\bar{q}_{nl_{i}} r) \: r^{\lambda} \:
3768: % \bar{A}_{nl_{f}}  \frac{\bar{k}_{nl_{f}}r}{kr}
3769: % j_{l_{f}}(\bar{k}_{nl_{f}}r)
3770: % \\ \nonumber & = & 
3771:   A_{nl_{i}} \bar{A}_{nl_{f}}^{\ast}
3772:   \bar{q}_{nl_{i}}^{-\lambda-2} 
3773:  M_{l_{i}}^{l_{f}}(\lambda) 
3774: %,\bar{k}_{n_{f}l_{f}}, \bar{q}_{n_{i}l_{i}})
3775: \end{eqnarray}
3776: and
3777: \begin{eqnarray}
3778:  I_{l_{i}}^{l_{f}}(\lambda,>) & = &
3779: % \int_{R}^{\infty} dr \: r^{2} \:
3780: % B_{nl_{i}} i^{l_{i}} h^{(1)}_{l_{i}}(iq r) \: r^{\lambda} \:
3781: % \frac{\left[ \exp(2i\delta_{l}) u_{l}^{(+)}(kr) 
3782: %     - u_{l}^{(-)}(kr) \right]}{2ikr}
3783: % \\ \nonumber & = & 
3784: % \frac{B_{nl_{i}}}{2ik}  \int_{R}^{\infty} dr \: i^{l_{i}}
3785: % h^{(1)}_{l_{i}}(iq r) \: r^{\lambda+1} \:
3786: % \left[ \exp(2i\delta_{l_{f}}) u_{l_{f}}^{(+)}(kr) 
3787: %   - u_{l_{f}}^{(-)}(kr) \right]
3788: % \\ \nonumber & = & 
3789:  B_{nl_{i}} q^{-\lambda-2}
3790:  \left[ \frac{\exp(2i\delta_{l_{f}}) N_{l_{i}}^{(+)l_{f}}(\lambda) %, k, q) 
3791:  - N_{l_{i}}^{(-)l_{f}}(\lambda) %, k, q)
3792:  }{2i} 
3793:  \right]^{\ast}
3794: \end{eqnarray}
3795: with the dimensionless integrals
3796: \begin{equation}
3797:  M_{l_{i}}^{l_{f}}(\lambda) %,\bar{k}, \bar{q})  
3798:  =
3799:    \bar{k} \bar{q}^{\lambda+2}  \int_{0}^{R} dr \: 
3800:   j_{l_{i}}(\bar{q}r) \: r^{\lambda+2} \:   j_{l_{f}}(\bar{k}r)
3801: % \\ \nonumber & = & 
3802: %  \bar{x}  \int_{0}^{\bar{\gamma}} dt \: 
3803: %  j_{l_{i}}(t) \: t^{\lambda+2} \:   j_{l_{f}}(\bar{x}t)
3804: \end{equation}
3805: and
3806: \begin{equation}
3807:  N_{l_{i}}^{(\pm)l_{f}}(\lambda) %, k , q) 
3808:  =
3809:   q^{\lambda+2} \int_{R}^{\infty} dr \: i^{l_{i}}
3810:  h^{(1)}_{l_{i}}(iq r) \: r^{\lambda+1} \:
3811:  u_{l_{f}}^{(\pm)}(kr)
3812: % \\ \nonumber & = & 
3813: % \int_{\gamma}^{\infty} dt \: i^{l_{i}}
3814: % h^{(1)}_{l_{i}}(it) \: t^{\lambda+1} \:
3815: % u_{l_{f}}^{(\pm)}(xt)
3816:  \: .
3817: \end{equation}
3818: These integrals obey the recursion relations
3819: \begin{eqnarray}
3820:  \label{eq:recmf}
3821:  M_{l_{i}}^{l_{f}+1}(\lambda+1) %, \bar{k}, \bar{q}) 
3822:  & = & 
3823:  \bar{q}  \left[ \frac{l_{f}+1}{\bar{k}} - \frac{d}{d\bar{k}} \right]
3824:  M_{l_{i}}^{l_{f}}(\lambda) %,\bar{k}, \bar{q}) 
3825:  \\
3826:  \label{eq:recmi}
3827:  M_{l_{i}+1}^{l_{f}}(\lambda+1) %,\bar{k}, \bar{q}) 
3828:  & = & 
3829:  \left[ l_{i}+\lambda+2 - \bar{q}\frac{d}{d\bar{q}} \right]
3830:  M_{l_{i}}^{l_{f}}(\lambda) %,\bar{k}, \bar{q}) 
3831:  \\
3832:  \label{eq:recnf}
3833:  N_{l_{i}}^{(\pm)l_{f}+1}(\lambda+1) %,k,q) 
3834:  & = & 
3835:  q  \left[ \frac{l_{f}+1}{k} - \frac{d}{dk} \right]
3836:  N_{l_{i}}^{(\pm)l_{f}}(\lambda) %,k,q) 
3837:  \\
3838:  \label{eq:recni}
3839:  N_{l_{i}+1}^{(\pm)l_{f}}(\lambda+1) %,k,q) 
3840:  & = & 
3841:   \left[ l_{i}+\lambda+2 - q\frac{d}{dq} \right]
3842:  N_{l_{i}}^{(\pm)l_{f}}(\lambda) %,k,q) 
3843:  \: .
3844: \end{eqnarray}
3845: Using the continuity equations (\ref{eq:cont}) and (\ref{eq:contf})
3846: we obtain
3847: \begin{eqnarray}
3848:  I_{l_{i}}^{l_{f}}(\lambda,<) & = &
3849:   A_{nl_{i}} \bar{A}_{nl_{f}}^{\ast} 
3850:   \bar{q}_{nl_{i}}^{-\lambda-2} 
3851:  M_{l_{i}}^{l_{f}}(\lambda) 
3852:  \\ \nonumber & = & 
3853:   B_{nl_{i}}  \bar{q}_{nl_{i}}^{-\lambda-2}
3854:  \frac{i^{l_{i}} h^{(1)}_{l_{i}}(i\gamma)M_{l_{i}}^{l_{f}}(\lambda)}{
3855:  j_{l_{i}}(\bar{\gamma}_{nl_{i}})\bar{\kappa}_{nl_{f}}  
3856:  j_{l_{f}}(\bar{\kappa}_{nl_{f}})}
3857:  \left[\frac{\exp(2i\delta_{l_{f}}) u_{l_{f}}^{(+)}(\kappa) 
3858:  - u_{l_{f}}^{(-)}(\kappa)}{2i}  \right]^{\ast}
3859: % \frac{M_{l_{i}}^{l_{f}}(\lambda)}{\bar{\kappa}_{nl_{f}}  
3860: % j_{l_{f}}(\bar{\kappa}_{nl_{f}})}
3861: %,\bar{k}_{n_{f}l_{f}}, \bar{q}_{n_{i}l_{i}})
3862: \end{eqnarray}
3863: and the ratio
3864: \begin{eqnarray}
3865:  R_{l_{i}}^{l_{f}}(\lambda) & = & 
3866:  \frac{I_{l_{i}}^{l_{f}}(\lambda,<)}{I_{l_{i}}^{l_{f}}(\lambda,>)}
3867:  \\ \nonumber & = & 
3868:  \left(\frac{\gamma}{\bar{\gamma}_{nl_{i}}}\right)^{\lambda+2}
3869:  \frac{i^{l_{i}} h^{(1)}_{l_{i}}(i\gamma)
3870:  M_{l_{i}}^{l_{f}}(\lambda)}{j_{l_{i}}(\bar{\gamma}_{nl_{i}})
3871:  \bar{\kappa}_{nl_{f}}  
3872:  j_{l_{f}}(\bar{\kappa}_{nl_{f}})}
3873: % \\ \nonumber &  & 
3874: \frac{\left[ \exp(2i\delta_{l_{f}}) u_{l_{f}}^{(+)}(\kappa) 
3875:  - u_{l_{f}}^{(-)}(\kappa) \right]^{\ast}}{\left[ 
3876:  \exp(2i\delta_{l_{f}}) N_{l_{i}}^{(+)l_{f}}(\lambda) %, k, q) 
3877:  - N_{l_{i}}^{(-)l_{f}}(\lambda) %, k, q) 
3878:  \right]^{\ast}} 
3879: \end{eqnarray}
3880: With the integral formula \cite{Abr65}
3881: \begin{equation}
3882:  \int dt \: f_{l}(at) \: t^{2} \: g_{l}(bt)
3883:  = \frac{t^{2}}{a^{2}-b^{2}} 
3884:  \left[ a f_{l+1}(at)g_{l}(bt)-bf_{l}(at)g_{l+1}(bt)\right]
3885: \end{equation}
3886: for spherical Bessel/Neumann/Hankel functions $f_{l}$ and $g_{l}$
3887: we find the monopole functions
3888: \begin{equation}
3889:  M_{l}^{l}(0) =
3890: % & = & 
3891: %    \bar{x}  \int_{0}^{\bar{\gamma}} dt \: 
3892: %  j_{l}(t) \: t^{2} \:   j_{l}(\bar{x}t)
3893: % \\ \nonumber & = & 
3894: %  \frac{\bar{x}\bar{\gamma}^{2}}{1-\bar{x}^{2}}
3895: % \left[ j_{l+1}(\bar{\gamma})j_{l}(\bar{x}\bar{\gamma})
3896: % - \bar{x} j_{l}(\bar{\gamma})j_{l+1}(\bar{x}\bar{\gamma})\right]
3897: % \\ \nonumber & = & 
3898:   \frac{\bar{\kappa}\bar{\gamma}^{2}}{\bar{\gamma}^{2}-\bar{\kappa}^{2}}
3899:  \left[\bar{\gamma} j_{l+1}(\bar{\gamma})j_{l}(\bar{\kappa})
3900:  - \bar{\kappa} j_{l}(\bar{\gamma})j_{l+1}(\bar{\kappa})\right]
3901: \end{equation}
3902: and
3903: \begin{equation}
3904:  N_{l}^{(\pm)l}(0)   =
3905: % & = &
3906: %   \int_{\gamma}^{\infty} dt \: i^{l}
3907: % h^{(1)}_{l}(it) \: t^{1} \: u_{l}^{(\pm)}(xt) 
3908: % \\ \nonumber & = &
3909: %  x  \int_{\gamma}^{\infty} dt \: i^{l}
3910: % h^{(1)}_{l}(it) \: t^{2} \: \left[-y_{l}(xt)\pm i j_{l}(xt) \right]
3911: % \\ \nonumber & = & 
3912: % \frac{i^{l} x\gamma^{2}}{1+x^{2}}
3913: % \left\{ ih_{l+1}^{(1)}(i\gamma)
3914: % \left[-y_{l}(x\gamma)\pm i j_{l}(x\gamma)\right]
3915: % \right. \\ \nonumber & &  \left.
3916: % - x h_{l}^{(1)}(i\gamma)\left[ -y_{l+1}(x\gamma) 
3917: % \pm i j_{l+1}(x\gamma)\right]
3918: %\right\}
3919: % \\ \nonumber & = & 
3920:  \frac{\gamma^{2}}{\gamma^{2}+\kappa^{2}}
3921:  \left[ i^{l+1}\gamma h_{l+1}^{(1)}(i\gamma) u_{l}^{(\pm)}(\kappa)
3922:  -  i^{l} \kappa h_{l}^{(1)}(i\gamma)
3923:  u_{l+1}^{(\pm)}(\kappa) \right]
3924: \end{equation}
3925: for general orbital angular momenta $l$.
3926: The logarithmic derivative of the scattering wave function is given by
3927: \begin{equation}
3928:   L_{l_{f}}^{f}
3929:  = \left. \frac{r\frac{d}{dr}g_{nl_{f}}}{g_{nl_{f}}} \right|_{r=R}
3930:  = l_{f} + 1 - Z_{l_{f}}^{(+)} = Z_{l_{f}}^{(-)} - l_{f} 
3931: \end{equation}
3932: with
3933: \begin{equation}
3934:  Z_{l_{f}}^{(\pm)} =
3935:   \bar{\kappa}_{nl_{f}} 
3936:  \frac{j_{l_{f}\pm 1}(\bar{\kappa}_{nl_{f}})}{j_{l_{f}}
3937:  (\bar{\kappa}_{nl_{f}})} 
3938:  = \kappa \frac{\left[\exp(2i\delta_{l_{f}})u_{l_{f}\pm1}^{(+)}(\kappa)
3939:  -u_{l_{f}\pm1}^{(-)}(\kappa)\right]}{\left[ 
3940:  \exp(2i\delta_{l_{f}}) u_{l_{f}}^{(+)}(\kappa) 
3941:      - u_{l_{f}}^{(-)}(\kappa) \right]} \: .
3942: \end{equation}
3943: We note
3944: \begin{equation}
3945:  Z_{l}^{(-)} Z_{l-1}^{(+)} = \kappa^{2}
3946:  \quad \mbox{and} \quad
3947:  Z_{l}^{(+)} + Z_{l}^{(-)} = 2l+1
3948: \end{equation}
3949: similar as in the case for the functions $Y_{l}^{(\pm)}$.
3950: Introducing $Y_{l}^{(\pm)}$ and $Z_{l}^{(\pm)}$ we find
3951: \begin{eqnarray}
3952:  M_{l}^{l}(0)   
3953:  & = & 
3954: %  \frac{\bar{\kappa}\bar{\gamma}^{2}}{\bar{\gamma}^{2}-\bar{\kappa}^{2}}
3955: % j_{l}(\bar{\kappa}) j_{l}(\bar{\gamma})
3956: % \left[\bar{\gamma} \frac{j_{l+1}(\bar{\gamma})}{j_{l}(\bar{\gamma})}
3957: % - \bar{\kappa} 
3958: % \frac{j_{l+1}(\bar{\kappa})}{j_{l}(\bar{\kappa})}\right]
3959: % \\ \nonumber & = & 
3960:   \frac{\bar{\kappa}\bar{\gamma}^{2}}{\bar{\gamma}^{2}-\bar{\kappa}^{2}}
3961:  j_{l}(\bar{\kappa}) j_{l}(\bar{\gamma})
3962:  \left[ Y_{l}^{(+)} - Z_{l}^{(+)} \right]
3963: % \\
3964: % N_{l}^{(\pm)l}(0)  
3965: % & = &
3966: % \frac{\gamma^{2}}{\gamma^{2}+\kappa^{2}}
3967: % i^{l} h_{l}^{(1)}(i\gamma) 
3968: % \left[ i\gamma \frac{h_{l+1}^{(1)}(i\gamma)}{h_{l}^{(1)}(i\gamma)} 
3969: % u_{l}^{(\pm)}(\kappa)
3970: % -  \kappa u_{l+1}^{(\pm)}(\kappa) \right]
3971: \end{eqnarray}
3972: and
3973: \begin{eqnarray}
3974:  \lefteqn{\exp(2i \delta_{l}) N_{l}^{(+)l}(0) - N_{l}^{(-)l}(0)}
3975:  \\ \nonumber  & = &
3976:  \frac{\gamma^{2}}{\gamma^{2}+\kappa^{2}}
3977:  i^{l} h_{l}^{(1)}(i\gamma) 
3978:  \left[ \exp(2i \delta_{l})  u_{l}^{(+)}(\kappa) 
3979:  -  u_{l}^{(-)}(\kappa)\right] 
3980:  \left[ Y_{l}^{(+)} - Z_{l}^{(+)} \right] \: .
3981: \end{eqnarray}
3982: The ratio of the monopole integrals is given by
3983: \begin{equation}
3984:  R_{l}^{l}(0) 
3985:   = 
3986:  \left(\frac{\gamma}{\bar{\gamma}_{nl}}\right)^{2}
3987:   \frac{\bar{\kappa}_{nl}
3988:  \bar{\gamma}_{nl}^{2}}{\bar{\gamma}_{nl}^{2}
3989:  -\bar{\kappa}^{2}}
3990:  \frac{\gamma^{2}+\kappa^{2}}{\gamma^{2}} \frac{1}{\bar{\kappa}_{nl}}
3991: % = % \\ \nonumber & = & 
3992: %  \frac{\gamma^{2}+\kappa^{2}}{\bar{\gamma}_{nl}^{2}-\bar{\kappa}_{nl}^{2}}
3993:  = \frac{\gamma^{2}+\kappa^{2}}{v_{i}-v_{f}-\gamma^{2}-\kappa^{2}}
3994: % = -1
3995: \end{equation}
3996: where the relations (\ref{eq:vi}) and (\ref{eq:vf}) were used.
3997: If the depths of the potential in the bound and scattering state are
3998: identical, i.e.\ $v_{i} = v_{f}$, we have
3999: \begin{equation}
4000:  R_{l}^{l}(0) = -1
4001:  \quad \mbox{i.e.} \quad
4002:  I_{l}^{l}(0) = 0
4003: \end{equation}
4004: and the bound and scattering wave functions are orthogonal.
4005: 
4006: Applying the recursion relations (\ref{eq:recmf}) and
4007: (\ref{eq:recnf}) the relevant integrals for
4008: $E1$ transitions $l \to l+1$ are found to be
4009: \begin{eqnarray}
4010:  \lefteqn{M_{l}^{l+1}(1) =
4011:  \frac{\bar{\kappa}\bar{\gamma}^{3}}{(\bar{\gamma}^{2}-\bar{\kappa}^{2})^{2}}
4012:  j_{l}(\bar{\gamma}) j_{l+1}(\bar{\kappa})}
4013:  \\ \nonumber & & \times
4014:  \left[ (\bar{\gamma}^{2} - \bar{\kappa}^{2})
4015:  \left( Y_{l}^{(+)} +Z_{l+1}^{(-)} \right) -2 Y_{l}^{(+)} Z_{l+1}^{(-)}
4016:  + (2l+3) \bar{\kappa}^{2} - (2l+1) \bar{\gamma}^{2} \right]
4017: \end{eqnarray}
4018: and
4019: \begin{eqnarray}
4020:  \lefteqn{\exp(2i\delta_{l+1})N_{l}^{(+)l+1}(1) - N_{l}^{(-)l+1}(1) =}
4021:  \\ \nonumber &  & 
4022:  \frac{\gamma^{3}}{(\gamma^{2}+\kappa^{2})^{2}}
4023:   i^{l} h_{l}^{(1)}(i\gamma)
4024:  \left[ \exp(2i\delta_{l+1})u_{l+1}^{(+)}(\kappa)-u_{l+1}^{(-)}(\kappa)\right]
4025:  \\ \nonumber & &  \times
4026:  \left[ (\gamma^{2}+\kappa^{2}) 
4027:  \left(  Y_{l}^{(+)} + Z_{l+1}^{(-)}\right)
4028:  +  2  Y_{l}^{(+)} Z_{l+1}^{(-)}
4029:  - (2l+1) \gamma^{2} - (2l+3) \kappa^{2} \right] \: .
4030: \end{eqnarray}
4031: Then the radial integrals are
4032: \begin{eqnarray} \label{eq:intfk}
4033:  \lefteqn{I_{l}^{l+1}(1,<) =
4034:   \frac{R^{2}}{(\bar{\gamma}^{2}-\bar{\kappa}^{2})^{2}}
4035:  f_{nl}(R) g_{nl+1}^{\ast}(R)}
4036:  \\ \nonumber & & \times
4037:  \left[ (\bar{\gamma}^{2} - \bar{\kappa}^{2})
4038:  \left( Y_{l}^{(+)} +Z_{l+1}^{(-)} \right) -2 Y_{l}^{(+)} Z_{l+1}^{(-)}
4039:  + (2l+3) \bar{\kappa}^{2} - (2l+1) \bar{\gamma}^{2} \right]
4040: \end{eqnarray}
4041: and
4042: \begin{eqnarray} \label{eq:intfg}
4043:  \lefteqn{I_{l}^{l+1}(1,>) =
4044:   \frac{R^{2}}{(\gamma^{2}+\kappa^{2})^{2}}
4045:   f_{nl}(R) g_{nl+1}^{\ast}(R)}
4046:  \\ \nonumber & & 
4047:  \times \left[ (\gamma^{2}+\kappa^{2}) 
4048:  \left(  Y_{l}^{(+)} + Z_{l+1}^{(-)}\right)
4049:  +  2  Y_{l}^{(+)} Z_{l+1}^{(-)}
4050:  - (2l+1) \gamma^{2} - (2l+3) \kappa^{2} \right] \: .
4051: \end{eqnarray}
4052: Similarly, the relevant integrals for $E1$ transitions $l+1 \to l$
4053: \begin{eqnarray} 
4054:  \lefteqn{M_{l+1}^{l}(1) =
4055:  \frac{\bar{\kappa}\bar{\gamma}^{3}}{(\bar{\gamma}^{2}-\bar{\kappa}^{2})^{2}}
4056:   j_{l+1}(\bar{\gamma}) j_{l}(\bar{\kappa})}
4057:  \\ \nonumber & & 
4058:  \times \left[ (2l+3) \bar{\gamma}^{2} - (2l+1) \bar{\kappa}^{2}
4059:  - \left( \bar{\gamma}^{2}-\bar{\kappa}^{2}\right)
4060:  \left( Y_{l+1}^{(-)} + Z_{l}^{(+)} \right)
4061:  - 2  Y_{l+1}^{(-)} Z_{l}^{(+)} \right]
4062: \end{eqnarray}
4063: and
4064: \begin{eqnarray} 
4065:  \lefteqn{ \exp(2i\delta_{l}) N_{l+1}^{(+)l}(1)
4066:  - N_{l+1}^{(-)l}(1) = }
4067:  \\ \nonumber &  & 
4068:  \frac{\gamma^{3}}{(\gamma^{2}+\kappa^{2})^{2}}
4069:  i^{l+1} h_{l+1}^{(1)}(i\gamma)
4070:  \left[ \exp(2i\delta_{l}) u_{l}^{(+)}(\kappa) - u_{l}^{(-)}(\kappa) \right]
4071:  \\ \nonumber & & \times
4072:  \left[ (2l+3) \gamma^{2} + (2l+1) \kappa^{2}
4073:  + 2 Y_{l+1}^{(-)} Z_{l}^{(+)} - (\gamma^{2}+\kappa^{2}) 
4074:  \left( Y_{l+1}^{(-)}+ Z_{l}^{(+)}\right) \right]
4075: \end{eqnarray}
4076: are found with the recursion relations
4077: (\ref{eq:recmi}) and (\ref{eq:recni}).
4078: Correspondingly, the radial integrals
4079: \begin{eqnarray} \label{eq:intik}
4080:  \lefteqn{I_{l+1}^{l}(1,<) =
4081:  \frac{R^{2}}{(\bar{\gamma}^{2}-\bar{\kappa}^{2})^{2}} 
4082:  f_{nl+1}(R) g_{nl}^{\ast}(R)}
4083:  \\ \nonumber & & \times
4084:  \left[ (2l+3) \bar{\gamma}^{2} - (2l+1) \bar{\kappa}^{2}
4085:  - 2  Y_{l+1}^{(-)} Z_{l}^{(+)}
4086:  - \left( \bar{\gamma}^{2}-\bar{\kappa}^{2}\right)
4087:  \left( Y_{l+1}^{(-)} + Z_{l}^{(+)} \right)
4088:   \right]
4089: \end{eqnarray}
4090: and
4091: \begin{eqnarray} \label{eq:intig}
4092:  \lefteqn{I_{l+1}^{l}(1,>) =
4093:  \frac{R^{2}}{(\gamma^{2}+\kappa^{2})^{2}}
4094:  f_{nl+1}(R) g_{nl}^{\ast}(R)}
4095:  \\ \nonumber & & \times
4096:  \left[ (2l+3) \gamma^{2} + (2l+1) \kappa^{2}
4097:  + 2 Y_{l+1}^{(-)} Z_{l}^{(+)} - (\gamma^{2}+\kappa^{2}) 
4098:  \left( Y_{l+1}^{(-)}+ Z_{l}^{(+)}\right) \right]
4099: \end{eqnarray}
4100: are obtained. 
4101: 
4102: Assuming that the depths of the potential are the same in
4103: the bound and the scattering state, i.e.
4104: \begin{equation}
4105:  v = \bar{\gamma}^{2}+\gamma^{2} = v_{i} = 
4106:  v_{f} = \bar{\kappa}^{2}-\kappa^{2}
4107: \end{equation}
4108: we find for the total integrals
4109: \begin{equation}
4110:  I_{l}^{l+1}(1) =
4111:   \frac{2R^{2}v}{(\gamma^{2}+\kappa^{2})^{2}}
4112:  f_{nl}(R) g_{nl+1}^{\ast}(R)
4113: \end{equation}
4114: and
4115: \begin{equation}
4116:  I_{l+1}^{l}(1) =
4117:  \frac{2R^{2}v}{(\gamma^{2}+\kappa^{2})^{2}}
4118:  f_{nl+1}(R) g_{nl}^{\ast}(R)
4119: \end{equation}
4120: consistent with the result (\ref{eq:radintgf}) of the commutator relation.
4121: The ratios of the interior to the exterior dipole integral are given by
4122: \begin{eqnarray}
4123:  \lefteqn{R_{l}^{l+1}(1) = \frac{I_{l}^{l+1}(1,<)}{I_{l}^{l+1}(1,>)}}
4124:  \\ \nonumber & = & 
4125:   \frac{(\bar{\gamma}^{2} - \bar{\kappa}^{2})
4126:  \left( Y_{l}^{(+)} +Z_{l+1}^{(-)} \right) -2 Y_{l}^{(+)} Z_{l+1}^{(-)}
4127:  + (2l+3) \bar{\kappa}^{2} - (2l+1) \bar{\gamma}^{2}}{
4128:  (\gamma^{2}+\kappa^{2}) 
4129:  \left(  Y_{l}^{(+)} + Z_{l+1}^{(-)}\right)
4130:  +  2  Y_{l}^{(+)} Z_{l+1}^{(-)}
4131:  - (2l+1) \gamma^{2} - (2l+3) \kappa^{2}}
4132: \end{eqnarray}
4133: and
4134: \begin{eqnarray}
4135:  \lefteqn{R_{l+1}^{l}(1) = \frac{I_{l+1}^{l}(1,<)}{I_{l+1}^{l}(1,>)}}
4136:  \\ \nonumber & = & 
4137:  \frac{(2l+3) \bar{\gamma}^{2} - (2l+1) \bar{\kappa}^{2}
4138:  - 2  Y_{l+1}^{(-)} Z_{l}^{(+)}
4139:  - \left( \bar{\gamma}^{2}-\bar{\kappa}^{2}\right)
4140:  \left( Y_{l+1}^{(-)} + Z_{l}^{(+)} \right)}{
4141:  (2l+3) \gamma^{2} + (2l+1) \kappa^{2}
4142:  + 2 Y_{l+1}^{(-)} Z_{l}^{(+)} - (\gamma^{2}+\kappa^{2}) 
4143:  \left( Y_{l+1}^{(-)}+ Z_{l}^{(+)}\right)}
4144: \end{eqnarray}
4145: for the transitions $l \to l+1$ and $l+1 \to l$, respectively.
4146: The scaling behaviour of these ratios is obtained by an
4147: expansion
4148: in terms of small
4149: $\gamma$ with $\kappa \to 0$. 
4150: %ST modified:
4151: In the case with a $s$ wave bound state we have the relation
4152: \begin{equation}
4153:  Y_{0}^{(-)} = -\gamma = \bar{\gamma} \cot \bar{\gamma} \: .
4154: \end{equation}
4155: Then we find the expansion
4156: %
4157: \begin{equation}
4158:  \bar{\gamma}_{n} = 
4159:  s_{n} \left( 1 + \frac{\gamma}{s_{n}^{2}} 
4160:  - \frac{\gamma^{2}}{s_{n}^{4}}
4161:  + \left(2-\frac{s_{n}^{2}}{3} \right) \frac{\gamma^{3}}{s_{n}^{6}}
4162:  - \left(5-\frac{4s_{n}^{2}}{3} \right) \frac{\gamma^{4}}{s_{n}^{8}} + \dots
4163:  \right)
4164: \end{equation}
4165: with $s_{n} = (2n-1)\pi/2$
4166: for $\gamma \to 0$
4167: depending on the principal quantum number $n=1,2,\dots $.
4168: Similarly, for the case of a $p$ wave bound state the relation
4169: \begin{equation}
4170:  Y_{1}^{(-)} = -\frac{\gamma^{2}}{1+\gamma} 
4171:  = \frac{\bar{\gamma}^{2}}{1- \bar{\gamma} \cot \bar{\gamma}}
4172: \end{equation}
4173: leads to the expansion
4174: \begin{equation}
4175:  \bar{\gamma}_{n} = 
4176:  s_{n} \left( 1 + \frac{\gamma^{2}}{s_{n}^{2}} 
4177:  - \frac{\gamma^{3}}{s_{n}^{2}}
4178:  + \left(s_{n}^{2}-2 \right) \frac{\gamma^{4}}{s_{n}^{4}}
4179:  + \dots
4180:  \right)
4181: \end{equation}
4182: with $s_{n} = n\pi$.
4183: Using
4184: \begin{equation}
4185:  Z_{1}^{(-)} = \frac{\bar{\kappa}^{2}}{1-\bar{\kappa} \cot \bar{\kappa}}
4186:  \quad \mbox{and} \quad
4187:  Z_{0}^{(+)} = 1-\bar{\kappa} \cot \bar{\kappa}
4188: \end{equation}
4189: we obtain
4190: \begin{equation}
4191:  \lim_{\kappa \to 0} R_{0}^{1}(1) =  \frac{\gamma^{4}}{8s_{n}^{2}}
4192:  + \dots
4193:  \quad \mbox{\and} \quad
4194:  \lim_{\kappa \to 0} R_{1}^{0}(1)  =  - \frac{\gamma^{2}}{4s_{n}^{2}}
4195:  + \dots
4196: \end{equation}
4197: for small $\gamma$. 
4198: 
4199: 
4200: 
4201: 
4202: \section{Integrals of spherical cylinder functions}
4203: \label{app:E}
4204: 
4205: With the recursion relations \cite{Abr65}
4206: \begin{eqnarray}
4207:  & & f_{l}^{\prime}(z) = f_{l-1}(z) - \frac{l+1}{z} f_{l}(z)
4208:  = \frac{l}{z} f_{l}(z)-f_{l+1}(z)
4209:  \\
4210:  & & 
4211:  f_{l+1}(z) + f_{l-1}(z) = \frac{2l+1}{z}f_{l}(z)
4212: \end{eqnarray}
4213: for any  spherical Bessel, Neumann
4214: or Hankel function $f_{l}(z)$  the well-known integral formula
4215: \begin{equation}
4216:  \int dz \: z^{2} \left[ f_{l}(z) \right]^{2}
4217:  = \frac{z^{3}}{2} \left( \left[ f_{l}(z) \right]^{2}
4218:  -  f_{l-1}(z)f_{l+1}(z) \right)
4219: \end{equation}
4220: is easily proven. In a similar fashion the new relation
4221: \begin{equation}
4222:  \int dz \: z^{4} \: \left[ f_{l}(z)\right]^{2}
4223:  = \frac{z^{5}}{12} \left( 3 [f_{l}(z)]^{2}
4224:  -2 f_{l-1}(z) f_{l+1}(z) - f_{l-2}(z) f_{l+2}(z)
4225:  \right) 
4226: \end{equation}
4227: is obtained. Similarly, integrals with higher powers in $z$ can be treated.
4228: Introducing the logarithmic derivative
4229: \begin{equation}
4230:  L_{l}(z) = z\frac{f_{l}^{\prime}(z)}{f_{l}(z)} 
4231:  = l-Y_{l}^{(+)}(z)
4232:  =  Y_{l}^{(-)}(z)-l-1 
4233: \end{equation}
4234: with
4235: \begin{equation}
4236:  Y_{l}^{(\pm)}(z) = z\frac{f_{l\pm1}(z)}{f_{l}(z)} \: .
4237: \end{equation}
4238: one finds
4239: \begin{eqnarray}
4240:  f_{l-1}(z)f_{l+1}(z) & = & 
4241: \frac{[f_{l}(z)]^{2}}{z^{2}} 
4242: \left[ l+1+L_{l}(z)\right]\left[ l - L_{l}(z)\right] 
4243: \end{eqnarray}
4244: and
4245: \begin{eqnarray}
4246:  f_{l-2}(z)f_{l+2}(z) & = & 
4247:  \frac{[f_{l}(z)]^{2}}{z^{4}} 
4248:  \left\{ (2l-1)[L_{l}(z)+l+1]-z^{2}\right\}
4249:  \\ \nonumber &  &
4250:  \left\{ (2l+3)[l-L_{l}(z)]-z^{2}\right\}
4251: \end{eqnarray}
4252: With these relations we obtain
4253: \begin{eqnarray}
4254:  \int dz \: z^{2} \left[ f_{l}(z) \right]^{2}
4255:  & = & 
4256: \frac{z}{2} \left[ f_{l}(z) \right]^{2} \left( z^{2}
4257: - \left[ l+1+L_{l}(z)\right]\left[ l - L_{l}(z)\right] \right)
4258: \end{eqnarray}
4259: and
4260: \begin{eqnarray}
4261:  \lefteqn{\int dz \: z^{4} \: \left[ f_{l}(z)\right]^{2}}
4262:  \\ \nonumber 
4263:  & = & \frac{z}{12} [f_{l}(z)]^{2} \left( 3 z^{4}
4264:  -2 z^{2}
4265: \left[ l+1+L_{l}(z)\right]\left[ l - L_{l}(z)\right]  \right.
4266:  \\ \nonumber & & \left.
4267:  -  \left\{(2l-1)\left[ l+1+L_{l}(z)\right] - z^{2}\right\}
4268:  \left\{(2l+3)\left[ l - L_{l}(z)\right] - z^{2} \right\}
4269:  \right) \: .
4270: \end{eqnarray}
4271: 
4272: \section{Radial transition integrals for neutron+core systems}
4273: \label{app:B}
4274: 
4275: In the case when $b$ is a neutron analytical expressions of the functions
4276: ${\mathcal H}_{l_{i}}^{l_{f}}(\lambda)$
4277: are found. With $\eta = 0$
4278: the Coulomb wave functions 
4279: \begin{eqnarray}
4280:  F_{l}(\eta;z) % & \to & \sqrt{\frac{\pi z}{2}} J_{l+\frac{1}{2}}(z)
4281:  \to z j_{l}(z)
4282:  \qquad
4283:  G_{l}(\eta;z) %& \to & (-1)^{l} \sqrt{\frac{\pi z}{2}} J_{-(l+\frac{1}{2})}(z)
4284: % (-1)^{l} z j_{-l-1}(z) = 
4285:  \to - z y_{l}(z) 
4286: \end{eqnarray}
4287: in the integral (\ref{eq:hdef})
4288: reduce to spherical Bessel and Neumann functions \cite{Abr65}. 
4289: Similarly, the Whittaker function 
4290: \begin{equation}
4291:  W_{-\eta,l+\frac{1}{2}}(2qr) \to -qr \: i^{l} \: h_{l}^{(1)}(iqr)
4292: \end{equation}
4293: reduces to a spherical Hankel function.
4294: This leads to
4295: \begin{eqnarray}
4296:  {\mathcal H}_{l_{i}}^{l_{f}}(\lambda) & = &
4297:  - i^{l_{i}} \kappa \gamma^{\lambda+2}  \int_{1}^{\infty} dt  \: 
4298:  t^{\lambda+2} \: 
4299:   h^{(2)}_{l_{f}}(\kappa t) \:
4300:    \:  h^{(1)}_{l_{i}}(i\gamma t)
4301: \end{eqnarray}
4302: with the quantities $\kappa = kR$ and $\gamma = qR$.
4303: From  the recursion relation 
4304: \begin{eqnarray}
4305:  g_{l+1}(z) & = & \left[ \frac{l}{z}  - \frac{d}{dz} \right] g_{l}(z)
4306: \end{eqnarray}
4307: for any spherical Bessel function $g_{l}$
4308: the recursion relations
4309: \begin{eqnarray} \label{eq:recf}
4310:  {\mathcal H}_{l_{i}}^{l_{f}+1}(\lambda+1) & = &
4311:  \gamma \left[ \frac{l_{f}+1}{\kappa} 
4312:  - \frac{d}{d\kappa} \right] {\mathcal H}_{l_{i}}^{l_{f}}(\lambda)
4313:  \\ \label{eq:reci}
4314:  {\mathcal H}_{l_{i}+1}^{l_{f}}(\lambda+1) & = &
4315:  \left[ l_{i}+\lambda+2 
4316:  - \gamma \frac{d}{d\gamma} \right] {\mathcal H}_{l_{i}}^{l_{f}}(\lambda)
4317: \end{eqnarray}
4318: for the functions ${\mathcal H}_{l_{i}}^{l_{f}}(\lambda)$ are obtained
4319: that allow the calculation of all relevant functions for increasing 
4320: values of $l_{i}$, $l_{f}$, and $\lambda$ from basic integrals with
4321: $\lambda = 0$.
4322: Explicitly one finds
4323: \begin{eqnarray}
4324:  {\mathcal H}_{0}^{0}(0) & = &
4325:  i \gamma \int_{1}^{\infty} dt \: \exp[-(\gamma+i \kappa)t]
4326:  = \frac{i\gamma}{\gamma+i\kappa} \exp(-\gamma-i\kappa)
4327: \end{eqnarray}
4328: and
4329: \begin{eqnarray}
4330:  {\mathcal H}_{1}^{1}(0) & = &
4331:   -   \gamma \int_{1}^{\infty} dt  \:   \exp[-(\gamma+i \kappa)t]
4332:   \left( 1 + \frac{1}{i\kappa t} 
4333:   + \frac{1}{\gamma t} + \frac{1}{i\kappa\gamma t^{2}}\right)
4334:  \\ \nonumber & = &
4335:    \left( \frac{i}{\kappa}  - \frac{\gamma}{\gamma+i\kappa}
4336:  \right) \exp(-\gamma-i\kappa)
4337: \end{eqnarray}
4338: with the help of the exponential integral and its recursion relations
4339: \cite{Abr65}.
4340: Separating real and imaginary parts one obtains the reduced
4341: monopole integrals 
4342: \begin{eqnarray} \label{eq:i000}
4343:  {\mathcal I}_{0}^{0}(0) & = & 
4344:  \frac{\gamma \exp(-\gamma)}{\gamma^{2}+\kappa^{2}} 
4345:  \left[  \kappa \cos (\kappa+\delta_{0}) 
4346:  + \gamma \sin(\kappa+\delta_{0})\right]
4347:  \\ \label{eq:i110}
4348:  {\mathcal I}_{1}^{1}(0) & = & 
4349:  \frac{\gamma \exp(-\gamma)}{\gamma^{2}+\kappa^{2}} 
4350:  \left[ -\gamma \cos (\kappa+\delta_{1}) 
4351:  + \left( \kappa +\frac{\gamma^{2}+\kappa^{2}}{\gamma \kappa}\right) 
4352:  \sin(\kappa+\delta_{1})\right]
4353: \end{eqnarray}
4354: where the addition theorems of the sine and cosine
4355: functions have been used. 
4356: The reduced integrals ${\mathcal I}_{l_{i}}^{l_{f}}(\lambda)$
4357: for fixed phase shift $\delta$
4358: obey the same recursion relations (\ref{eq:recf},\ref{eq:reci}) 
4359: as the functions ${\mathcal H}_{l_{i}}^{l_{f}}(\lambda)$.
4360: 
4361: Considering the general form (\ref{eq:iredgen}) 
4362: for the reduced radial integrals
4363: the monopole functions 
4364: \begin{equation}
4365:  {\mathcal R}_{0}^{(+)0}(0) = \kappa  \qquad 
4366:  {\mathcal R}_{0}^{(-)0}(0) = \gamma 
4367: \end{equation}
4368: and
4369: \begin{equation}
4370:  {\mathcal R}_{1}^{(+)1}(0) = - \kappa  \qquad 
4371:  {\mathcal R}_{1}^{(-)1}(0) = 
4372:  [\kappa^{2}(1+\gamma)+\gamma^{2}]/\gamma^{2}
4373: \end{equation}
4374: are extracted from Eqs.\ (\ref{eq:i000}) and (\ref{eq:i110}).
4375: General expression for  ${\mathcal R}_{l_{i}}^{(\pm)l_{f}}(\lambda)$
4376: in Eq.~(\ref{eq:iredgen}) 
4377: for larger values of $l_{f}$, $l_{i}$, and
4378: $\lambda$ can be obtained 
4379: by applying the recursion relations
4380: \begin{eqnarray}
4381:  {\mathcal R}_{l_{i}}^{(\pm)l_{f}+1}(\lambda+1)
4382:  & = & \left[ 2\kappa^{2}(\lambda+1)+
4383:  (\gamma^{2}+\kappa^{2})(2l_{f}+1)\right]
4384:  {\mathcal R}_{l_{i}}^{(\pm)l_{f}}(\lambda)
4385:  \\ \nonumber & & 
4386:  - \kappa (\gamma^{2}+\kappa^{2}) \left[
4387:  \frac{d}{d\kappa} {\mathcal R}_{l_{i}}^{(\pm)l_{f}}(\lambda)
4388:  \pm {\mathcal R}_{l_{i}}^{(\mp)l_{f}}(\lambda)\right]
4389: \end{eqnarray}
4390: \begin{eqnarray}
4391:  {\mathcal R}_{l_{i}+1}^{(\pm)l_{f}}(\lambda+1)
4392:  & = & \left[ 2\gamma^{2}(\lambda+1)+
4393:  (\gamma^{2}+\kappa^{2})(l_{i}+\lambda+1-l_{f}+\gamma)\right]
4394:  \\ \nonumber & & \times
4395:  {\mathcal R}_{l_{i}}^{(\pm)l_{f}}(\lambda)
4396: % \\ \nonumber & & 
4397:  - \gamma (\gamma^{2}+\kappa^{2}) 
4398:  \frac{d}{d\gamma} {\mathcal R}_{l_{i}}^{(\pm)l_{f}}(\lambda)
4399: \end{eqnarray}
4400: Explicitly one finds
4401: \begin{eqnarray}
4402:  {\mathcal R}_{0}^{(+)1}(1) & = & -\kappa[\kappa^{2}(\gamma-2)+\gamma^{3}]
4403:  \\
4404:  {\mathcal R}_{0}^{(-)1}(1) & = & 
4405:  \kappa^{4}+\kappa^{2}\gamma(3+\gamma)+\gamma^{3}
4406:  \\
4407:  {\mathcal R}_{1}^{(+)0}(1) & = & 
4408:  \kappa[\kappa^{2}(1+\gamma)+\gamma^{2}(3+\gamma)]
4409:  \\
4410:  {\mathcal R}_{1}^{(-)0}(1) & = & \gamma^{2}[\kappa^{2}+\gamma(2+\gamma)]
4411:  \\
4412:  {\mathcal R}_{1}^{(+)2}(1) & = & -
4413:  \kappa[\kappa^{4}(1+\gamma)+\kappa^{2}\gamma^{2}(6+\gamma)
4414:  +3\gamma^{4}]/\gamma^{2}
4415:  \\
4416:  {\mathcal R}_{1}^{(-)2}(1) & = & 
4417:  [\kappa^{4}(3+3\gamma-\gamma^{2})+\kappa^{2}\gamma^{2}(6+\gamma-\gamma^{2})
4418:  +3\gamma^{4}]/\gamma^{2}
4419:  \\
4420:  {\mathcal R}_{2}^{(+)1}(1) & = & -
4421:  \kappa[\kappa^{2}(1+\gamma)+\gamma^{2}(3+\gamma)]
4422:  \\
4423:  {\mathcal R}_{2}^{(-)1}(1) & = & 
4424:  [\kappa^{4}(3+3\gamma+\gamma^{2})+\kappa^{2}\gamma^{2}(6+6\gamma+\gamma^{2})
4425:  +\gamma^{4}(3+\gamma)]/\gamma^{2}
4426: \end{eqnarray}
4427: for $\lambda=1$ and
4428: \begin{eqnarray}
4429:  {\mathcal R}_{0}^{(+)2}(2) & = &
4430:  -\kappa [\kappa^{6}+\kappa^{4}(-8+7\gamma+2\gamma^{2})
4431:  +\kappa^{2}\gamma^{3}(10+\gamma)+3\gamma^{5}]
4432:  \\
4433:  {\mathcal R}_{0}^{(-)2}(2) & = & 
4434:  \kappa^{6} (5-\gamma) +\kappa^{4}\gamma(15+6\gamma-2\gamma^{2}) 
4435:  + \kappa^{2}\gamma^{3}(10+\gamma-\gamma^{2}) 
4436:  \\ \nonumber & &
4437:  + 3\gamma^{5}
4438:  \\
4439:  {\mathcal R}_{1}^{(+)1}(2) & = &
4440:  \kappa[\kappa^{4}(2+2\gamma-\gamma^{2})+2\kappa^{2}\gamma^{2}(5-\gamma^{2})
4441:  -\gamma^{5}(2+\gamma)]
4442:  \\
4443:  {\mathcal R}_{1}^{(-)1}(2) & = & 
4444:  \kappa^{6}(1+\gamma) +\kappa^{4}\gamma^{2}(7+2\gamma) 
4445:  +\kappa^{2}\gamma^{3}(10+7\gamma+\gamma^{2})
4446:  \\ \nonumber & & 
4447:  +\gamma^{5}(2+\gamma)
4448:  \\
4449:  {\mathcal R}_{2}^{(+)0}(2) & = &
4450:  \kappa [\kappa^{4}(3+3\gamma+\gamma^{2})
4451:  +2\kappa^{2}\gamma^{2}(5+5\gamma+\gamma^{2})
4452:  \\ \nonumber & & 
4453:  +\gamma^{4}(15+7\gamma+\gamma^{2})]
4454:  \\
4455:  {\mathcal R}_{2}^{(-)0}(2) & = & 
4456:  \gamma^{2}[\kappa^{4}(1+\gamma)+2\kappa^{2}\gamma^{2}(3+\gamma)
4457:  +\gamma^{3}(8+5\gamma+\gamma^{2})]
4458: \end{eqnarray}
4459: for $\lambda=2$. They cover most of the relevant $E1$ and $E2$ 
4460: transitions between $s$, $p$, and $d$ waves in the initial and final
4461: states.
4462: 
4463: 
4464: 
4465: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4466: 
4467: \begin{thebibliography}{00}
4468: 
4469: \bibitem{Bau86}
4470:  G. Baur, C. A. Bertulani, and H. Rebel,
4471:  Nucl. Phys. A 459 (1986) 188.
4472: 
4473: \bibitem{Ber88}
4474:  C. A. Bertulani and G. Baur,
4475:  Phys. Rep. 163 (1988) 299.
4476: 
4477: \bibitem{Lei01}
4478:  A. Leistenschneider et al.,
4479:  Phys. Rev. Lett. 86 (2001) 5442.
4480: 
4481: \bibitem{Bau03}
4482:  G. Baur, K. Hencken, and D. Trautmann,
4483:  Prog. in Part. Nucl. Phys. 51 (2003) 487.
4484: 
4485: \bibitem{Aum04}
4486:  T. Aumann, in: 
4487:  Proceedings of the International Workshop XXXII on Gross Properties
4488:  of Nuclei and Nuclear Excitations, edited by M. Buballa, J. Knoll,
4489:  W. N\"{o}renberg, B.-J. Sch\"{a}fer and J. Wambach, Hirschegg, Austria, 
4490:  January 11 - 17, 2004, GSI, Darmstadt (2004), p.\ 243 ff.
4491: 
4492: \bibitem{Han87}
4493:  P. G. Hansen and B. Jonson,
4494:  Europhys. Lett. 4 (1987) 409.
4495: 
4496: \bibitem{Rii92}
4497:  K. Riisager, A. S. Jensen, P. M{\o}ller,
4498:  Nucl. Phys. A 548 (1992) 393.
4499: 
4500: \bibitem{Rii94}
4501:  K. Riisager,
4502:  Rev. Mod. Phys. 66 (1994) 1105.
4503: 
4504: \bibitem{Han95}
4505:  P. G. Hansen, A. S. Jensen, and B. Jonson,
4506:  Annu. Rev. Nucl. Part. Sci. 45 (1995) 591.
4507: 
4508: \bibitem{Tan96}
4509:  I. Tanihata,
4510:  J. Phys. G 22 (1996) 157.
4511: 
4512: \bibitem{Sag01}
4513:  H. Sagawa, T. Suzuki, H. Iwasaki, and M. Ishihara,
4514:  Phys. Rev. C 63 (2001) 034310.
4515: 
4516: \bibitem{Nag97}
4517:  Y. Nagai et al.,
4518:  Phys. Rev. C 56 (1997) 3173.
4519: 
4520: \bibitem{Bla79}
4521:  J. M. Blatt and V. E. Weisskopf,
4522:  Theoretical Nuclear Physics,
4523:  Springer-Verlag, New York (1979), p.\ 48 ff.
4524: 
4525: \bibitem{Xu94}
4526:  H.M. Xu et al.,
4527:  Phys. Rev. Lett. 73 (1994) 2027.
4528: 
4529: \bibitem{Tra01}
4530:  L. Trache, F. Carstoiu, C.A. Gagliardi, and R.E. Tribble,
4531:  Phys. Rev. Lett. 87 (2001) 271102.
4532: 
4533: \bibitem{Muk01}
4534:  A.M. Mukhamedzhanov, C.A. Gagliardi, and R.E. Tribble,
4535:  Phys. Rev. C 63 (2001) 024612.
4536: 
4537: \bibitem{Cre02} 
4538:  R. Crespo and F.M. Nunes,
4539:  Nucl. Phys. A 701 (2002) 637.
4540: 
4541: \bibitem{Ots94}
4542:  T. Otsuka, M. Ishihara, N. Fukunishi, T. Nakamura, and M. Yokoyama,
4543:  Phys. Rev. C 49 (1994) R2289.
4544: 
4545: \bibitem{Men95}
4546:  A. Mengoni, T. Otsuka, and M. Ishihara,
4547:  Phys. Rev. C 52 (1995), R2334.
4548: 
4549: \bibitem{Kal96}
4550:  D. M. Kalassa and G. Baur,
4551:  J. Phys. G 22 (1996) 115.
4552: 
4553: \bibitem{Typ01a}
4554:  S. Typel and G. Baur, 
4555:  Phys. Rev. C 64 (2001) 024601.
4556: 
4557: \bibitem{Per76}
4558:  C. M. Perey and F. G. Perey,
4559:  Atomic Data and Nuclear Data Tables 17 (1976) 1.
4560: 
4561: \bibitem{Bet49}
4562:  H. A. Bethe.
4563:  Phys. Rev. 76 (1949) 38.
4564: 
4565: \bibitem{New82}
4566:   R. G. Newton,
4567:   Scattering Theory of Waves and Particles,
4568:   second edition, Springer-Verlag, New York (1982), p.\ 306 ff.
4569: 
4570: \bibitem{Typ04a}
4571:  S. Typel and G. Baur,
4572:  Phys. Rev. Lett. 93 (2004) 142502.
4573: 
4574: \bibitem{Noc04}
4575:  C. Nociforo et al.,
4576:  Electromagnetic excitation of ${}^{23}$O,
4577:  Phys. Lett. B 605 (2005) 79.
4578: 
4579: \bibitem{Bir99}
4580:  M. C. Birse, J. A. McGovern, and K. G. Richardson,
4581:  Phys. Lett. B 464 (1999) 169. 
4582: 
4583: \bibitem{Kap99}
4584:  D. B. Kaplan, M. J. Savage, and M. B. Wise,
4585:  Phys. Rev. C 59 (1999) 617. 
4586: 
4587: \bibitem{Ber02}
4588:  C. A. Bertulani, H.-W. Hammer, and U. van Kolck,
4589:  Nucl. Phys. A 712 (2002) 37. 
4590: 
4591: \bibitem{Eri03}
4592:  T. E. O. Ericsson, B. Loiseau, and S. Wycech,
4593:  hep-ph/0310134 (2003). 
4594: 
4595: \bibitem{Hol99}
4596:  B. R. Holstein,
4597:  Phys. Rev. D 60 (1999) 114030.
4598: 
4599: \bibitem{Liu04}
4600:  Z. H. Liu, X. Z. Zhang, and H. Q. Zhang,
4601:  Phys. Rev. C 68 (2003) 024305.
4602: 
4603: \bibitem{Ham98}
4604:  I. Hamamoto and X. Z. Zhang,
4605:  Phys. Rev. C 58 (1998) 3388.
4606: 
4607: \bibitem{Fed93}
4608:  D. Fedorov, A.S. Jensen, K. Riisager,
4609:  Phys. Lett. B 312 (1993) 1.
4610: 
4611: \bibitem{Rii00}
4612:  K. Riisager, D.V. Fedorov, and A.S. Jensen,
4613:  Europhys. Lett. 49 (2000) 547.
4614: 
4615: \bibitem{Jen04}
4616:  A. Jensen, K. Riisager, D.V. Fedorov, and E. Garrido,
4617:  Rev. Mod. Phys. 76 (2004) 215.
4618: 
4619: \bibitem{Ber92}
4620:  C. A. Bertulani and A. Sustich,
4621:  Phys. Rev. C 46 (1992) 2340.
4622: 
4623: \bibitem{Abr65}
4624:  M. Abramowitz and I. S. Stegun,
4625:  Handbook of Mathematical Functions,
4626:  Dover Publications, New York (1965).
4627: 
4628: \bibitem{Ehr27}
4629:  P. Ehrenfest,
4630:  Z. Physik 45 (1927) 455.
4631: 
4632: \bibitem{Gor29}
4633:  W. Gordon,
4634:  Ann. Phys. 2 (1929) 1031.
4635: 
4636: \bibitem{Jen98}
4637:  B. K. Jennings, S. Karataglidis, and T. D. Shoppa,
4638:  Phys. Rev. C 58 (1998) 3711.
4639: 
4640: \bibitem{Lev50}
4641:  J.S. Levinger and H.A. Bethe,
4642:  Phys. Rev. 78 (1950) 115.
4643: 
4644: \bibitem{Bet77}
4645:  H. A. Bethe and E. E. Salpeter,
4646:  Quantum Mechanics of One- and Two-Electron Atoms,
4647:  Plenum, New York (1977). 
4648: 
4649: \bibitem{Bru96}
4650:  C. R. Brune,
4651:  Nucl. Phys. A 596 (1996) 122.
4652: 
4653: \bibitem{Nak94}
4654:  T. Nakamura et al.,
4655:  Phys. Lett. B 331 (1994) 296.
4656: 
4657: \bibitem{Bay04}
4658:  D. Baye,
4659:  Phys. Rev. C 70 (2004) 015801.
4660: 
4661: \bibitem{Muk02}
4662:  A. M. Mukhamedzhanov and F. M. Nunes,
4663:  Nucl. Phys. A 708, (2002) 437.
4664: 
4665: \bibitem{Bay00}
4666:  D. Baye and E. Brainis,
4667:  Phys. Rev. C 61 (2000) 025801.
4668: 
4669: \bibitem{Jon04}
4670:  B. Jonson,
4671:  Phys. Rep. 389 (2004) 1.
4672: 
4673: \bibitem{Alh82}
4674:  Y. Alhassid, M. Gai, and G. F. Bertsch,
4675:  Phys. Rev. Lett. 49 (1982) 1482.
4676: 
4677: \bibitem{Hen04}
4678:  K. Hencken, G. Baur, and D. Trautmann,
4679:  Nucl. Phys. A 733 (2004) 200.
4680: 
4681: \bibitem{Ann94}
4682:  R. Anne et al.,
4683:  Nucl. Phys. A 575 (1994) 125.
4684: 
4685: \bibitem{Pal03}
4686:  R. Palit et al.,
4687:  Phys. Rev. C 68 (2003) 034318.
4688: 
4689: \bibitem{Fuk04}
4690:  N. Fukuda et al.,
4691:  Phys. Rev. C 70 (2004) 054606.
4692: 
4693: \bibitem{Dat03}
4694:  U. Datta Pramanik et al.,
4695:  Phys. Lett. B 551 (2003) 67.
4696: 
4697: \bibitem{Kan01}
4698:  R. Kanungo, I. Tanihata, and A. Ozawa,
4699:  Phys. Lett. B 512 (2001) 261.
4700: 
4701: \bibitem{Kan02}
4702:  R. Kanungo et al.,
4703:  Phys. Rev. Lett. 88 (2002) 142502.
4704: 
4705: \bibitem{Bro03}
4706:   B. A. Brown, P. G. Hansen, and J. A. Tostevin,
4707:   Phys. Rev. Lett. 90 (2003) 159201.
4708: 
4709: \bibitem{Tho03}
4710:  M. Thoennessen et al.,
4711:  Phys. Rev. C 68 (2003) 044318.
4712: 
4713: \bibitem{Sau04}
4714:  E. Sauvan et al.,
4715:  Phys. Rev. C 69 (2004) 044603.
4716: 
4717: \bibitem{Cor04}
4718:  D. Cortina-Gil et al.,
4719:  Phys. Rev. Lett. 93 (2004) 062501.
4720: 
4721: \bibitem{Azh01}
4722:  A. Azhari et al,
4723:  Phys. Rev. C 63 (2001) 055803;
4724:  Phys. Rev. C 60 (1999) 055803; 
4725:  Phys. Rev. Lett. 82 (1999) 3960.
4726: 
4727: \bibitem{Oga03b}
4728:  K. Ogata et al.,
4729:  Phys. Rev. C 67 (2003) 011602(R), 019902.
4730: 
4731: \bibitem{Tra03}
4732:  L. Trache et al.,
4733:  Phys. Rev. C 67 (2003) 062801(R).
4734: 
4735: \bibitem{Jun02}
4736:  A. R. Junghans et al.,
4737:  Phys. Rev. Lett. 88 (2002) 041101;
4738:  Phys. Rev. C 68 (2003) 065803.
4739: 
4740: \bibitem{Sch03}
4741:  F.~Sch\"{u}mann et al.,
4742:  Phys. Rev. Lett. 90 (2003) 232501.
4743: 
4744: \bibitem{Bea01}
4745:  D. Beaumel et al.,
4746:  Phys. Lett. B 514 (2001) 226.
4747: 
4748: \bibitem{Tra02}
4749:  L. Trache et al.,
4750:  Phys. Rev. C 66 (2002) 035801.
4751: 
4752: \bibitem{End03}
4753:  J.  Enders et al.,
4754:  Phys. Rev. C 67 (2003) 064301.
4755: 
4756: \bibitem{Lef95}
4757:  A. Lefebvre et al.,
4758:  Nucl. Phys. A 592 (1995) 69.
4759: 
4760: \bibitem{Tan03}
4761:  Xiaodong Tang et al.,
4762:  Phys. Rev. C 67 (2003) 015804.
4763: 
4764: \bibitem{Tim03} 
4765:  N.K. Timofeyuk and S.B. Igamov,
4766:  Nucl. Phys A 713 (2003) 217.
4767: 
4768: \bibitem{Mor97}
4769:  R. Morlock et al.,
4770:  Phys. Rev. Lett. 79 (1997) 3837.
4771: 
4772: \bibitem{Dav03}
4773:  B. Davids and S. Typel, 
4774:  Phys. Rev. C 68 (2003) 045802.
4775: 
4776: \bibitem{Ang03}
4777:  C. Angulo et al.,
4778:  Nucl. Phys. A 716 (2003) 211.
4779: 
4780: \bibitem{Koe83}
4781:  L. Koester, K. Knopf, and W. Waschkowski, 
4782:  Z. Phys. A 312 (1983) 81. 
4783: 
4784: \bibitem{Mes61}
4785:  A. Messiah,
4786:  {\it Quantum Mechanics}, Vol. I,
4787:  North Holland, Amsterdam (1961).
4788: 
4789: \bibitem{Lep97}
4790:  G. P. Lepage,
4791:  How to Renormalize the Schr\"{o}dinger Equation, 
4792:  Lectures given at the VIII Jorge Andre Swieca Summer School, Brazil (1997),
4793:  nucl-th/9706029.
4794: 
4795: \bibitem{Mig73}
4796:  A. B. Migdal, 
4797:  Sov. J. of Nucl. Phys. 16 (1973) 238.
4798: 
4799: \bibitem{Bra04}
4800:  E. Braaten and H.-W. Hammer,
4801:  Universality in Few-body Systems with Large Scattering Length,
4802:  cond-mat/0410417.
4803: 
4804: \bibitem{Nag05}
4805:  M. A. Nagarajan, S. M. Lenzi, and A. Vitturi,
4806:  Eur. Phys. J. A 24 (2005) 63.
4807: 
4808: \end{thebibliography}
4809: 
4810: \end{document}
4811: 
4812: 
4813: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4814: 
4815: 
4816: 
4817: