nucl-th0702021/SF5.tex
1: \documentclass[prc,showpacs,aps,twocolumn]{revtex4}
2: 
3: \usepackage{amsmath}
4: 
5: \usepackage{epsfig}
6: 
7: \begin{document}
8: 
9: 
10: \bigskip
11: 
12: \title{Threshold  Effects in Multi-channel Coupling and Spectroscopic
13: Factors in Exotic Nuclei}
14: 
15: 
16: \author
17: {N. Michel$^{\dag \ddag \P}$, W. Nazarewicz$^{\dag \ddag \$}$, 
18: M. P{\l}oszajczak$^{\S}$}
19: 
20: 
21:  
22: 
23: \affiliation{$^{\dag}$ Department of Physics and Astronomy, University of Tennessee, Knoxville, TN 37996, USA\\
24: $^{\ddag}$ Physics Division, Oak Ridge National Laboratory, P.O.~Box 2008, Oak Ridge, TN 37831, USA\\
25: $^{\P}$ Joint Institute for Heavy Ion Research, Oak Ridge, TN 37831, USA\\
26: $^{\$}$ Institute of Theoretical Physics, Warsaw University, ul.~Ho\.{z}a 69, 00-681 Warsaw, Poland\\
27: $^{\S}$ Grand Acc\'el\'erateur National d'Ions Lourds (GANIL), CEA/DSM - CNRS/IN2P3,
28: BP 55027, F-14076 Caen Cedex, France}
29: 
30:  
31: 
32: \date{\today}
33: 
34:  
35: 
36: \begin{abstract}
37: 
38: In the threshold region,  the cross section and the associated overlap
39: integral obey the Wigner threshold law  that results in the Wigner-cusp
40: phenomenon. Due to  flux conservation, a cusp anomaly in one  channel
41: manifests itself  in other  open channels, even if their respective
42: thresholds appear at a different energy. The shape of a threshold cusp
43: depends on the  orbital angular momentum of a scattered particle; hence,
44: studies of Wigner anomalies in weakly bound nuclei with several
45: low-lying thresholds can provide valuable spectroscopic information. In
46: this work,  we investigate  the threshold behavior of  spectroscopic
47: factors in neutron-rich drip-line nuclei using the Gamow Shell Model,
48: which takes into account many-body correlations and  the continuum
49: effects. The presence of threshold anomalies is demonstrated and the
50: implications for spectroscopic factors are discussed.
51: 
52: \end{abstract}
53: 
54: \pacs{21.10.Jx, 03.65.Nk, 21.60.Cs, 24.10.Cn, 24.10.Eq}
55: 
56: 
57: \bigskip 
58: 
59: \maketitle
60: 
61: In 1948, based on general principles, Wigner predicted \cite{Wigner} a
62: characteristic behavior (a cusp) of  scattering and reaction cross
63: sections in the vicinity of  a reaction threshold. This particular
64: behavior (often referred to as the Wigner's threshold law or the
65: Wigner-cusp phenomenon) was given a quantitative explanation a decade
66: later \cite{Breit,Baz,Newton,Meyerhof,Baz1,Lane}. In particular, it has
67: been noted \cite{Baz,Newton,Meyerhof,Baz1,Hategan1} that, due to the
68: unitarity of the scattering matrix and  the resulting flux conservation,
69: the presence of  a threshold anomaly in an opening reaction channel can
70: trigger an appearance of Wigner cusps in other open  channels with lower
71: thresholds. As the shape of the cusp strongly depends on orbital angular
72: momentum (is strongest in $s$ and $p$ waves), it was early realized that
73: the presence of cusp anomaly could provide structural information about
74: reaction products  \cite{Baz,Newton,Adair}.
75: 
76: The Wigner-cusp phenomenon has been studied experimentally and
77: theoretically in various areas of physics: pion-nucleus scattering
78: \cite{Adair,Starostin}, electron-molecule scattering \cite{Domcke},
79: electron-atom scattering  \cite{Scheibner}, and ultracold atom-diatom
80: scattering \cite{Forrey}. In low-energy nuclear physics, threshold
81: effects have been investigated in, e.g.,  charge-exchange reactions
82: \cite{Malmberg}, neutron elastic scattering \cite{Wells}, and deuteron
83: stripping \cite{Moore}. Abramovich {\it et al.} \cite{Abramovich}
84: reviewed threshold phenomena in nuclear reactions, including those with
85: the light neutron-rich  systems such as $^6$He, $^{10}$Be,  and
86: $^{10}$Li (see also Ref.~\cite{Hategan}) that are of particular interest
87: in the context of this work. The influence of threshold effects on 
88: the cross section and the strength function, hence
89: the spectroscopic factor (SF)  of a threshold state was 
90: pointed out in Refs.~\cite{Hategan1,Graw}.
91: 
92: 
93: 
94: The purpose of this study is many-fold. Firstly, we investigate whether
95: the Wigner-cusp phenomenon appears naturally in a microscopic many-body
96: approach rooted in an effective inter-nucleon interaction. The second
97: goal is to investigate the influence of threshold effects and the
98: coupling to the non-resonant continuum on overlap integrals, or
99: spectroscopic factors. We demonstrate  that the energy dependence of SFs
100: caused by an opening of a reaction channel can be described at the
101: shell-model level only if nucleon-nucleon correlations involving
102: scattering states are treated properly. Finally, we emphasize the
103: importance of experimental studies of cusp phenomena in weakly bound,
104: neutron-rich nuclei in which  low-lying one-neutron and two-neutron
105: thresholds appear.
106: 
107: The traditional shell model (SM) of nuclear structure views  the nucleus
108: as a  closed quantum system (CQS) in which nucleons occupy bound orbits.
109: While such an  assumption may be somehow  justified for well-bound
110: nuclei  having  high  particle-emission thresholds, it can no longer
111: be applied to weakly bound or unbound systems in the vicinity of
112: drip lines, where the coupling to the particle continuum (both
113: resonances and the non-resonant scattering states) becomes important.
114: This  coupling  can be considered in the open quantum system (OQS)
115: extension of the SM, the so-called continuum SM (see Ref.
116: \cite{opr} for a recent review). In this work,   we apply the
117: complex-energy implementation of the continuum SM, the so-called Gamow Shell Model
118: (GSM) \cite{Mic02,Bet02} in the version of Refs.~\cite{Mic02,Mic04}. GSM
119: is a multi-configurational SM with a single-particle (s.p.) basis given
120: by the Berggren ensemble \cite{Berggren1} which consists of Gamow (bound
121: and  resonance) states and the  non-resonant continuum. For a given
122: Hamiltonian, the number of particles occupying states of the
123: non-resonant continuum is a result of GSM variational calculations. The
124: resonant states of the GSM are the generalized eigenstates of the
125: time-independent Schr\"{o}dinger equation which are regular at the
126: origin and satisfy purely outgoing boundary conditions. The GSM can thus
127: be viewed as  a quasi-stationary many-body OQS formalism.
128: (An alternative microscopic approach, successfully applied to the structure of
129: weakly bound or unbound nuclei, is the Complex Scaling Method 
130: that  is capable of treating different kinds of reaction channels and continuum
131: states starting from different thresholds \cite{Myo}.)
132: Since the Wigner-cusp phenomenon is most pronounced for  low-$\ell$
133: waves and for neutrons (no Coulomb barrier), as an illustrative example
134: we choose the case of the one-neutron (1n) channel in  the model two-
135: and three-neutron systems outside the inert core:  $^{6}$He and $^7$He.
136: Our aim is not to fit the actual experimental data, but rather to
137: accomplish the physics goals as  stated above. 
138: 
139: \paragraph*{Gamow Shell Model Framework--}
140: The starting point of GSM is the Berggren one-body completeness relation
141: \cite{Berggren1} allowing the expansion of both bound and unbound states.
142: The Berggren ensemble consists of resonant and scattering states generated
143: by a finite-depth potential $V(r)$. 
144: Resonant states are solutions of the Schr\"{o}dinger equation with
145: purely outgoing asymptotics. Their energies and wave functions are in
146: general complex. The resonant states of the GSM have either bound or
147: decaying character; they  form the so-called  {\it pole subspace}.
148: Scattering states entering the Berggren ensemble  are also defined in the complex
149: energy/momentum plane. For a given partial wave $(\ell,j)$, the
150: scattering states are distributed along a contour $L_+^{\ell_j}$ in the
151: complex momentum plane. The set of all bound and decaying states $|u_n
152: \rangle$ enclosed between $L_+^{\ell_j}$ and the real $k$-axis, and
153: scattering states $|u_k \rangle$ in $L_+^{\ell_j}$ is complete
154: \cite{Berggren1}:
155: \begin{equation}
156: \int\hspace{-1.4em}\sum_{\mathcal{B}} |u_\mathcal{B} \rangle 
157: \langle \widetilde{u_\mathcal{B}}| = 1,  \label{one_body_compl_rel}
158: \end{equation}
159: where $|\mathcal{B}\rangle$ is either a discrete resonant state
160: or a scattering continuum state ($L_+^{\ell_j}$ part).
161: The tilde symbol above bra vectors in Eq.~(\ref{one_body_compl_rel})
162: signifies that  the complex conjugation arising in the dual space
163: affects only the angular part and leaves the radial part unchanged
164: \cite{Berggren1}. The continuous part of the completeness relation 
165: (\ref{one_body_compl_rel})  has to be discretized in numerical
166: applications. For that purpose, scattering
167: states are discretized and renormalized in order to obtain a discrete 
168: completeness relation \cite{Berggren1}:
169: \begin{eqnarray}
170: \sum_{\mathcal{B}=1}^{N} |u_{\mathcal{B}} \rangle \langle \widetilde{u_{\mathcal{B}}}| \simeq 1 
171: ~~;~~~~|u_{\mathcal{B}} \rangle = \sqrt{\omega_{\mathcal{B}}} \; 
172: |u_{k_{\mathcal{B}}} \rangle, 
173: \label{one_body_compl_rel_discr}
174: \end{eqnarray}
175: where $\{k_{\mathcal{B}},\omega_{\mathcal{B}}\}$ is the set of
176: discretized momenta and associated weights provided by a Gauss-Legendre
177: quadrature. The many-body GSM basis corresponds to Slater
178: determinants (SD) spanned 
179: by  one-body Berggren states:
180: $\displaystyle |SD_i \rangle = | u_{i_1} \cdots u_{i_A} \rangle$ where
181: $|SD_i \rangle$ is the $i$-th SD 
182: in the $A$-body basis and $u_{i_j}$ is the $j$-th one-body state
183: occupied in $|SD_i \rangle$. The many-body completeness 
184: relation is built from Eq.~(\ref{one_body_compl_rel_discr})
185: by forming all possible SDs generated by the Gamow one-body states:
186: \begin{eqnarray}
187: &&\sum_{i} |SD_i \rangle \langle \widetilde{SD_i}| \simeq 1. \label{Nbody_compl_rel_discr}
188: \end{eqnarray}
189: The completeness in (\ref{Nbody_compl_rel_discr}) is not exact as the
190: one-body completeness relation (\ref{one_body_compl_rel_discr})
191: is approximate due to the discretization. 
192: In the basis (\ref{Nbody_compl_rel_discr}), the GSM Hamiltonian $H$
193: becomes a complex symmetric matrix.  Moreover, many-body bound and
194: resonant states are embedded in the background
195: of non-resonant
196: scattering eigenstates, so that one
197: needs a criterion to isolate them. The
198: overlap method \cite{Mic02} has proven to be very  efficient to solve
199: this problem. For this, one diagonalizes first $H$ in the pole
200: subspace  to
201: generate a zeroth-order vector  $|\Psi_0\rangle$. In the second
202: step, $|\Psi_0\rangle$ is used as a pivot to generate a Lanczos subspace
203: of the full GSM space. Its diagonalization provides
204: eigenvectors of $H$ in the total GSM space, and the requested bound or
205: decaying eigenstate of $H$ is the one which maximizes the overlap $|\langle \Psi_0 |
206: \Psi\rangle|$.
207: 
208: The definition of observables in GSM follows directly from the
209: mathematical setting of quantum mechanics in the rigged Hilbert space
210: rather than  the usual Hilbert space \cite{Bohm,Madrid}. Modified
211: definition of the dual space, embodied by the tilde symbol above bra
212: states, implies that observables in many-body resonances become complex.
213: In this case, the real part of a matrix element corresponds to the
214: expectation value,  and the imaginary part can be interpreted as the
215: uncertainty in the determination of this expectation value due to the
216: possibility of decay of the state during the measuring process
217: \cite{Berggren1,Civ99}.
218: 
219:  
220: 
221: \paragraph*{GSM Hamiltonian--}
222: In this Letter, the s.p.~basis (\ref{one_body_compl_rel}) 
223: is generated by a Woods-Saxon (WS) potential with the radius $R_0$=2 fm,
224: the diffuseness $d$=0.65 fm,  the
225: spin-orbit strength $V_{\rm so}$=7.5 MeV, and  the depth of the central
226: potential $V_0$=47 MeV (the ``$^{5}$He" parameter set). This potential
227: reproduces experimental energies and widths of the s.p.~resonances
228: $3/2_1^-$ and $1/2_1^-$ in $^5$He. The GSM Hamiltonian is a sum of the
229: one-body WS potential, representing the effect of an inert $^4$He core,
230: and of the two-body interaction among valence particles, given by a
231: finite-range surface Gaussian interaction (SGI) \cite{Mic04} with the
232: range $\mu=1$\,fm and the coupling constants depending on the total
233: angular momentum $J$ of the neutron pair: $V_0^{(0)}=-403$ MeV fm$^{3}$
234: and $V_0^{(2)}=-392$ MeV fm$^{3}$. These constants are fitted to
235: reproduce the experimental ground state (g.s.) binding energies of
236: $^6$He and $^7$He with the ``$^{5}$He" WS potential. The valence space
237: for neutrons consists of the $p_{3/2}$ and $p_{1/2}$ partial waves. The
238: $p_{3/2}$ wave functions include a $0p_{3/2}$ resonant state and
239: $p_{3/2}$ non-resonant scattering states along a complex contour
240: enclosing the $0p_{3/2}$ resonance in the complex $k$-plane. For a
241: $p_{1/2}$ part, we take non-resonant scattering states along the
242: real-$k$ axis (the broad $0p_{1/2}$ resonant state  plays a negligible
243: role in the g.s.~wave function of $^{6}$He). For both contours, the
244: maximal momentum value is $k_{\rm{max}}=$3.27 fm$^{-1}$. The contours
245: have been discretized with up to 60 points and the attained precision on
246: energies and widths is better than 0.1 keV. In the subsequent analysis,
247: parameters of the GSM Hamiltonian are varied in order to change
248: positions of 1n thresholds in various isotopes.
249: 
250: In order to illuminate the continuum coupling effects in GSM, we have
251: introduced  a simplified harmonic oscillator SM  (HO-SM) scheme. Here, 
252: the radial wave functions are those of the spherical harmonic oscillator
253:   with the frequency $\hbar \omega=41A^{-1/3}$ MeV. The one-body part of
254: the HO-SM Hamiltonian is given by the real energies of the one-body part
255: of the GSM Hamiltonian.  The HO-SM scheme is supposed to illustrate the
256: ``standard"  CQS SM calculations in which only bound valence  shells are
257: considered in the s.p. basis.
258: 
259: \paragraph*{Spectroscopic Factors in GSM--} SFs are useful indicators of
260: the configuration mixing in the many-body wave function. Extensive
261: attempts have been made to deduce SFs using direct reactions, such as
262: single-nucleon transfer, nucleon knockout,  and elastic break-up
263: reactions, using hadronic and leptonic probes. These analyses often
264: reveal model- and probe-dependence \cite{sf1,sf2,sf3} raising concerns
265: about the possibility of their precise experimental determination. (For
266: an extensive study of spectroscopic factors in exotic nuclei from
267: nucleon-knockout reactions, see Ref.~ \cite{gade}.) Studies of $(e,e'p)$
268: reactions in closed-shell nuclei \cite{eep} demonstrated that the SFs
269: are reduced by $\sim 35\%$ with respect to the standard SM predictions,
270: mainly due to the coupling to high-momentum states reached by the
271: short-range and tensor components of the nucleon-nucleon interaction
272: \cite{dickhoff}. In this work, we point out that additional difficulties
273: in extracting and interpreting SFs from the measured cross sections
274: using the standard SM lie in  the neglect of particle continuum, channel
275: coupling, and strength fragmentation. 
276:  
277: Single-nucleon overlap integrals and the associated spectroscopic
278: factors (SFs) are basic ingredients of the theory of direct reactions
279: (single-nucleon transfer, nucleon knockout, elastic break-up)
280: \cite{satch,sfs}.   Experimentally, SFs  can be  deduced
281: from measured cross sections; they are useful measures of  the configuration mixing
282: in the many-body wave function. The associated reaction-theoretical analysis 
283: often reveals
284: model- and probe-dependence \cite{sf1,sf2,sf3} raising concerns about
285: the accuracy of  experimental determination of SFs. In our study we
286: discuss the  uncertainty in determining SFs due to  the two assumptions commonly used in  the
287: standard SM studies, namely (i) that a nucleon is transferred
288: to/from a specific s.p.~orbit (corresponding to an observed
289: s.p.~state), and  (ii) that the transfer to/from the continuum 
290: of non-resonant scattering states
291: can be disregarded.
292: In this work, we define SFs in a usual way, through
293: the radial
294: overlap  functions $u_{\ell j}(r)$  
295: \cite{sfs,Glenden,Fro_Lipp}:
296: \begin{equation}
297: u_{\ell j}(r) = \langle \Psi^{J_{A}}_{A} | \left[ | \Psi^{J_{A-1}}_{A-1} 
298: \rangle \otimes |\ell,j \rangle \right]^{J_A} \rangle, \label{ovf_basic_def}
299: \end{equation}
300: where $|\Psi^{J_{A}}_{A}\rangle$ and
301: $|\Psi^{J_{A-1}}_{A-1}\rangle$ are wave functions of nuclei $A$ and $A-1$,
302: respectively, and 
303: $|\ell,j\rangle$ is the angular-spin part of the channel function.
304: The angular-spin degrees of freedom
305: are integrated out in Eq.~(\ref{ovf_basic_def})
306: so that  $u_{\ell j}$ depends only on 
307: the relative radial coordinate of the transferred particle
308: $r = |\vec{r}|$.
309: The spectroscopic factor, denoted by $S^2$, is defined as usual
310: through the norm of
311: the overlap function \cite{sfs,Glenden,Fro_Lipp} . 
312: Using a decomposition of the $(\ell,j)$ channel function in the complete
313: Berggren basis,  one obtains:
314: \begin{eqnarray}
315: &&u_{\ell j}(r) = \int\hspace{-1.4em}\sum_{\mathcal{B}} \langle \widetilde{\Psi^{J_{A}}_{A}} || a^+_{\ell j}(\mathcal{B}) || \Psi^{J_{A-1}}_{A-1} \rangle
316: \mbox{ } \langle r \ell j | \mathcal{B} \rangle, \label{ovf_GSM} \\
317: &&S^2 = \int\hspace{-1.4em}\sum_{\mathcal{B}} \langle
318:  \widetilde{\Psi^{J_{A}}_{A}} || a^+_{\ell j}(\mathcal{B}) ||
319:   \Psi^{J_{A-1}}_{A-1} \rangle^2, \label{eq3}
320: \end{eqnarray}
321: where $a^+_{\ell j} (\mathcal{B})$ is a
322: creation operator associated with  a s.p. Berggren state $|\mathcal{B}\rangle$.
323: Since Eqs.~(\ref{ovf_GSM},\ref{eq3}) involve
324: summation  over all discrete Gamow states and integration over all 
325: scattering states along  the contour $L_+^{\ell_j}$, the final result is
326: independent of the s.p.~basis assumed. This is in contrast to standard
327: SF experimental extraction and SM calculations where model-dependence
328: enters through the specific choice of a s.p.~state $a^+_{n \ell j}$,
329: with Eq.~(\ref{eq3}) reducing to the sole matrix element
330: $\displaystyle \langle \Psi^{J_{A}}_{A} || a^+_{n \ell j} || 
331: \Psi^{J_{A-1}}_{A-1} \rangle^2$. 
332: This can lead to sizeable errors if the  states of $A-1$
333: and/or $A$  lie close to a channel threshold.
334: 
335: \begin{figure}[htb]
336: \centerline{\includegraphics[width=8.5cm,angle=0]{fig1.eps}}
337: \caption{The real part of  the 
338: overlap integral as a function of one-neutron
339: separation energy  $S_{1n}$ as indicated.
340: Top and middle: 
341: $\displaystyle \langle ^6{\rm He( g.s.}) | 
342: [^5{\rm He( g.s.}) \otimes p_{3/2}]^{0^+} \rangle$; 
343: bottom:  $\displaystyle \langle ^7{\rm He( g.s.}) | 
344: [^6{\rm He( g.s.}) \otimes p_{3/2}]^{0^+} \rangle$.
345: Solid line:  GSM results; dashed line:  
346: the SM-like approximation (HO-SM) where the SGI matrix elements are calculated in 
347: the HO basis $\{ 0p_{3/2}, 0p_{1/2} \}$.  The 1n thresholds in $^6$He (top and
348: bottom) and $^5$He are marked by arrows.
349: The results displayed in the middle panel are plotted as a function of
350: the energy of the $0p_{3/2}$ resonant state, which, in our model,
351: is negative of $S_{1n}$[$^5$He].
352: For more details, see the discussion in the 
353: text.}
354: \label{fig_SF}
355: \end{figure}
356: 
357: \paragraph*{The Threshold Cusp--}
358: The GSM results for $^6$He g.s.~SF in the channel: $[{^5}{\rm He}({\rm
359: g.s.})\otimes p_{3/2}]^{0^+}$ are shown in Fig.~\ref{fig_SF}a as a
360: function of one-neutron separation energy  $S_{1n}$ in $^6$He. To this
361: end, we have fixed the depth of the  WS potential so that $0p_{3/2}$ and
362: $0p_{1/2}$ are both bound with respective energies -5 and -0.255 MeV,
363: and we have varied the SGI monopole coupling constant $V_0^{(J=0)}$ so
364: that the 1n separation energy of $^6$He goes through zero. At $S_{1n}$=0
365: (1n-emission threshold), the calculated SF exhibits behavior consistent
366: with the Wigner threshold  law: the quickly varying component of SF
367: behaves as $(-S_{1n})^{\ell-1/2}$ ($\ell$=1) {\it below} the 1n
368: threshold and follows the   $(S_{1n})^{\ell+1/2}$ rule   {\it above} the
369:  threshold. It is worth noting that the parameters of the GSM
370: Hamiltonian and the associated S-matrix poles do not show any
371: discontinuities around $S_{1n}$=0. This result constitutes an excellent
372: test of the GSM formalism: the Wigner limit is reached precisely at the
373: 1n threshold obtained from many-body calculations. It is interesting to
374: see that the HO-SM results are depressed by about 25\% as compared to
375: GSM. Indeed, in the HO-SM  s.p. basis, the configurations $[0p_{3/2}
376: \otimes 0p_{3/2}]^{0^+}$ and $[0p_{1/2} \otimes 0p_{1/2}]^{0^+}$ are
377: strongly mixed by the SGI residual interaction; hence, the value of
378: $p_{3/2}$ SF in the g.s.~of $^6$He  is significantly reduced. Moreover,
379: the s.p.~basis in HO-SM calculations comprises HO states whose radial
380: form factors are independent of the depth of the WS potential. Hence, no
381: threshold effect can be seen in HO-SM SFs.
382: 
383: \paragraph*{Threshold Effects Due to Channel Coupling--}
384: 
385: As discussed above, when a new channel opens at energy $E_t$, there
386: appears a  flux redistribution in other open channels with lower
387: thresholds. This flux redistribution may be affected by the presence of
388: the Wigner cusp in the new channel; hence, the reaction cross-sections
389: in {\it all}   open channels may exhibit the threshold anomaly at $E_t$.
390: To illustrate the phenomenon of channel coupling, Fig.~\ref{fig_SF}b
391: shows again the SF for the  $[{^5}{\rm He}({\rm g.s.})\otimes
392: p_{3/2}]^{0^+}$ channel but this time as a function of the $0p_{3/2}$
393: s.p.~state energy (or negative $S_{1n}$ of $^5$He). This calculation was
394: carried out  by varying the depth of the WS potential so that the
395: $p_{3/2}$ pole of the $S$-matrix (which is also the g.s.~of $^5$He in
396: our model space), would change its character from a bound state to an
397: unbound decaying Gamow state. At $S_{1n}$[$^5$He]=0, the 1n and 2n
398: thresholds in $^6$He become degenerate, i.e., 
399: the two channels of $^5$He+n, namely $^6$He and $^4$He+2n couple.
400: As seen in Fig.~\ref{fig_SF}, the  SF in the $[{^5}{\rm He}({\rm
401: g.s.})\otimes p_{3/2}]^{0^+}$ channel exhibits the Wigner cusp at
402: $S_{1n}$[$^5$He]=0, i.e., at a 2n threshold. At this point, the first
403: derivative $\partial S^2/\partial e_{0p_{3/2}}$ of the SF becomes
404: infinite, consistent with the Wigner law for $\ell$=1. The SF in HO-SM
405: varies very little in the whole studied range of $0p_{3/2}$ energies.
406: (Let us mention that  the competition between bound and unbound 
407: states in $^6$He, and the $^5$He+n and $^4$He+2n reaction channels has been 
408: studied within the coupled cluster method \cite{Myo}.)
409: 
410: 
411: In the case shown in  Fig.~\ref{fig_SF}b the GSM Hamiltonian changes at
412: $S_{1n}$[$^5$He]=0 as at this point the $0p_{3/2}$ pole becomes unbound.
413: It is, therefore, instructive to investigate the situation at which the
414: Hamiltonian behaves smoothly around the threshold. Figure~\ref{fig_SF}c
415: illustrates  the case of $^7$He g.s.  in the channel $[{^6}{\rm He}({\rm
416: g.s.})\otimes p_{3/2}]^{3/2^-}$. In order to avoid  secondary
417: open-channel mixing effects, we have adjusted the WS potential depth so
418: that the $^5$He g.s. is bound by 5\,MeV and cannot play any role in the
419: anomalous energy dependence of the studied SF. The $0p_{1/2}$ state is
420: weakly  bound with an energy of $-0.255$ MeV. In order to control the 
421: binding energy of $^6$He, the $V_0^{(J=0)}$ coupling constant is varied
422: so that $^7$He g.s. is always bound while $S_{1n}$  of $^6$He goes
423: through zero. The GSM space is the same as in the previous cases, except
424: that the $0p_{1/2}$ state is now  included in the GSM basis as it is
425: bound. A cusp in the calculated SF for $^7$He is clearly seen at
426: $S_{1n}$[$^6$He]=0. The threshold anomaly shown in Fig.~\ref{fig_SF}c
427: can only result from the cross-channel couplings. Again, the   SF in 
428: HO-SM  smoothly varies  in the whole   energy region considered. 
429: 
430: \paragraph*{Conclusions--} The anomalies in the spectroscopic factors
431: when the total energy of the system varies through the threshold of an
432: opening channel are discussed within the many-body OQS formalism of the
433: GSM. The main conclusions of this work can be summarized as follows: (i)
434: Many-body OQS calculations correctly predict the Wigner-cusp and
435: channel-coupling threshold effects. This constitutes a very strong
436: theoretical check  for the  GSM approach; (ii) The spectroscopic
437: factors  defined in the OQS framework through the norm of the overlap
438: integral, exhibit strong variations around particle thresholds. Such
439: variations cannot be described in a standard CQS SM framework that
440: applies a `one-isolated-state' ansatz and ignores the coupling to the
441: decay and scattering channels. In our model calculations, the
442: contribution to SF from a non-resonant continuum can be as large as 25\%;
443: (iii) Any theoretical model aiming at a meaningful description of SFs of
444: low-$\ell$ states in weakly bound nuclei must meet certain minimal
445: conditions. Namely, it should be able to account for (a)  many-body
446: configuration mixing and the resulting spreading of the spectroscopic
447: strength due to inter-nucleon correlations, (b) coupling to the particle
448: continuum that affects radial properties of wave functions in the
449: neighborhood of reaction thresholds, and (c) coupling between various 
450: reaction channels.
451: 
452: Considering  points (ii) and (iii) above, one should be careful
453: when extracting spectroscopic information from transfer experiments  on
454: drip-line nuclei. The results presented
455: in this work suggest that, similar to other fields, experimental studies
456: of various aspects of   threshold effects  could provide valuable
457: spectroscopic information about the s.p. structure of weakly bound
458: nuclei. Here, of particular interest are systems  with near-lying 1n and
459: 2n thresholds such as $^{6,8}$He or non-Borromean two-neutron halos
460: \cite{Timofeyuk03}, in which cusps in SFs are expected to be
461: particularly strong in low-$\ell$ ($\ell=0,1$) neutron channels. 
462: Similar features were found in the analysis of the continuum-coupling
463: correction to the CQS eigenvalues near the reaction threshold
464: \cite{Op05}. Finally, let us note that a threshold anomaly is also
465: expected  in proton-rich nuclei. While the effect is weaker than in the
466: neutron-rich systems, one still expects
467: anomalous  differences in SFs  and other spectroscopic quantities in
468: mirror nuclei. Work along these lines  is in progress.
469: 
470: 
471: 
472: This work was supported by  the U.S. Department of Energy
473: under Contract Nos. DE-FG02-96ER40963 (University of Tennessee),
474: DE-AC05-00OR22725 with UT-Battelle, LLC (Oak Ridge National
475: Laboratory), and DE-FG05-87ER40361 (Joint Institute for Heavy Ion
476: Research).
477: 
478: 
479: 
480: \begin{thebibliography}{99}
481: 
482: \bibitem{Wigner} E.P.~Wigner, Phys. Rev. {\bf 73}, 1002 (1948).
483: \bibitem{Breit} G.~Breit, Phys. Rev. {\bf 107}, 1612 (1957).
484: \bibitem{Baz} A.I. Baz, Soviet Phys. - JETP {\bf 6}, 709 (1957).
485: \bibitem{Newton} R.G. Newton, Phys. Rev. {\bf 114}, 1611 (1959).
486: \bibitem{Meyerhof}  W.E. Meyerhof, Phys. Rev. {\bf 129}, 692 (1963).
487: \bibitem{Baz1}  A.I.~Baz, Y.B.~Zeldovich,  and A.M.~Peremolov, 
488:  {\it Scattering, Reactions and Decays in Nonrelativistic Quantum Mechanics},  
489: (Jerusalem, 1969, translated from Russian, Nauka, Moscow, 1966).
490: \bibitem{Lane}  A.M. Lane, Phys. Lett. B {\bf 33}, 274 (1970).
491: \bibitem{Hategan1} C.~Hategan, Annals of Physics  {\bf 116} 77 (1978).
492: \bibitem{Adair} R.K. Adair, Phys. Rev. {\bf 111}, 632 (1958).
493: \bibitem{Starostin} A. Starostin {\it et al.},  Phys. Rev. C {\bf 72}, 015205 (2005).
494: \bibitem{Domcke} W. Domcke, J. Phys. B {\bf 14}, 4889 (1981).
495: \bibitem{Scheibner} K.F. Scheibner, A. U. Hazi, and R.J.W. Henry, Phys. Rev. A {\bf 35}, 4869 (1987).
496: \bibitem{Forrey} R.C. Forrey, N. Balakrishnan, V. Kharchenko, and A. Dalgarno, Phys. Rev. A
497:     {\bf 58}, R2645 (1998).
498: \bibitem{Malmberg} P.R. Malmberg, Phys. Rev. {\bf 101}, 114 (1956).
499: \bibitem{Wells} J.T. Wells, A.B. Tucker, and W.E. Meyerhof, Phys. Rev. {\bf 131}, 1644 (1963).
500: \bibitem{Moore} C.F. Moore {\it et al.},  Phys. Rev. Lett. {\bf 17}, 926 (1966).
501: \bibitem{Abramovich} S. Abramovich, B. Guzhovskij, and L. Lasarev,
502: Sov. J. Part. Nucl. {\bf 23}, 129 (1992).
503: \bibitem{Hategan} C.~Hategan, Proc. Rom. Acad. A {\bf 3} 11 (2002). 
504: \bibitem{Graw} G. Graw and C.~Hategan, Phys. Lett. B {\bf 37} 41 (1971). 
505: \bibitem{opr} J. Oko{\l}owicz, M. P{\l}oszajczak,  and I. Rotter, 
506:     Phys. Rep. {\bf 374}, 271 (2003).
507: \bibitem{Mic02} N. Michel, W. Nazarewicz, M. P{\l}oszajczak,
508:    and K. Bennaceur,  Phys. Rev. Lett. {\bf 89}, 042502 (2002);
509:    N. Michel, W. Nazarewicz, M. P{\l}oszajczak, and J. Oko{\l}owicz, 
510:    Phys. Rev. C {\bf 67}, 054311 (2003).
511: \bibitem{Bet02} R. Id Betan, R.J. Liotta, N. Sandulescu, 
512:     and T. Vertse, Phys. Rev. Lett. {\bf 89}, 042501 (2002);
513:     Phys. Rev. C {\bf 67},  014322 (2003).   
514: \bibitem{Mic04} N. Michel, W.~Nazarewicz,  and M. P{\l}oszajczak, 
515:    Phys. Rev. C  {\bf 70}, 064313 (2004).
516: \bibitem{Berggren1} T.~Berggren, Nucl. Phys. A {\bf 109}, 265 (1968).
517: \bibitem{Myo} T. Myo, K. Kat\=o, S. Aoyama, and K. Ikeda, Phys. Rev. C {\bf 63}, 054313
518:    (2001).
519:  \bibitem{Bohm} A. Bohm, {\it The Rigged Hilbert Space and Quantum Mechanics},
520:  Lecture Notes in Physics {\bf 78} (Springer, New York  1978).
521: \bibitem{Madrid} R. de la Madrid, Eur. J. Phys. {\bf 26}, 287 (2005).  
522: \bibitem{Civ99} O. Civitarese, M. Gadella, and R. Id Betan, 
523:     Nucl. Phys. A {\bf 660}, 255 (1999).
524: \bibitem{sf1} P.M. Endt, Atomic Data and Nuclear Data Tables {\bf 19}, 23 (1977);
525:    M.B.~Tsang, Jenny Lee, and W.G.~Lynch, Phys. Rev. Lett. {\bf 95},  222501 (2005).
526: \bibitem{sf2} G.J. Kramer, H.P. Blok, and L. Lapikas, Nucl. Phys. {\bf A679}, 267 (2001).
527: \bibitem{sf3} B.A.~Brown, P.G.~Hansen, B.M.~Sherrill, and J.A.~Tostevin, Phys. Rev. 
528:    C {\bf 65}, 061601(R) (2002);
529:    P.G.~Hansen and J.A.~Tostevin, Ann. Rev. Nucl. Part. Sci. {\bf 53}, 219 (2003).
530: \bibitem{gade} A. Gade {\it et al.}, Phys. Rev. Lett. {\bf 93}, 042501 (2004);
531:    Eur. Phys. J. A {\bf 25}, s1.251 (2005).
532:  \bibitem{eep} A.E.L. Dieperink and P. de Witt Huberts, Ann. Rev. Nucl. Part. Sci. 
533:    {\bf 40}, 239 (1990);
534:    I. Sick and P. de Witt Huberts, Commun. Nucl. Part. Phys. {\bf 20}, 177 (1991);
535:    L. Lapikas, Nucl. Phys. A {\bf 553}, 297c (1993).
536: \bibitem{dickhoff} W.H. Dickhoff and C. Barbieri, Prog. Part. Nucl. Phys. 
537:    {\bf 52}, 377 (2004).
538: \bibitem{satch} G.R. Satchler, {\it Direct Nuclear Reactions} (Clarendon, Oxford, 1983).
539: \bibitem{sfs} M.H. Macfarlane and J.B. French, Rev. Mod. Phys. {\bf 32}, 567 (1960).
540: \bibitem{Glenden} N.K.~Glendenning, {\it Direct Nuclear Reactions} 
541: (Academic Press inc., 1983).
542: \bibitem{Fro_Lipp} P.~Fr{\"o}brich and R.~Lipperheide,
543:  {\it Theory of Nuclear Reactions} (Oxford Science Publications, 
544:  Clarendon Press Oxford, 1996).
545: \bibitem{Timofeyuk03} N.K. Timofeyuk, L.D. Blokhintsev, and J.A. 
546: Tostevin, Phys. Rev. C {\bf 68} 021601(R) (2003).
547: \bibitem{Op05}  N. Michel, W. Nazarewicz, J. Oko{\l}owicz, and M.~P{\l}oszajczak, 
548: Nucl. Phys. A {\bf 752}, 335c (2005);
549: Y. Luo, J. Oko{\l}owicz, M. P{\l}oszajczak, and N.~Michel, arXiv:nucl-th/0201073.
550:               
551: 
552: \end{thebibliography}
553: 
554: \end{document}
555: