nucl-th0412049/be.tex
1: %\documentclass[aps,prc,twocolumn,floatfix,showpacs,amsmath,amssymb]{revtex4}
2: %\documentclass[aps,prc,preprint,endfloats,floatfix,showpacs,amsmath,amssymb]{revtex4}
3: \documentclass[aps,prc,twocolumn,floatfix,showpacs,amsmath,amssymb]{revtex4}
4: \usepackage{graphicx}
5: \usepackage{dcolumn}% Align table columns on decimal point
6: \usepackage{color} % To use colored text to denote comments
7: \bibliographystyle{apsrev}
8: %-------------------------
9: % Personal new commands and shorthands
10: % Define \nuc{A}{Z} command
11:   \def\nuc#1#2{\relax\ifmmode{}^{#1}{\protect\text{#2}}\else${}^{#1}$#2\fi}
12:   \def\itnuc#1#2{\setbox\@tempboxa=\hbox{\scriptsize\it #1}
13:     \def\@tempa{{}^{\box\@tempboxa}\!\protect\text{\it #2}}\relax
14:     \ifmmode \@tempa \else $\@tempa$\fi}
15: % Shorthand for 9,11Be and other common expressions
16:   \newcommand{\beel}{\nuc{11}{Be}}
17:   \newcommand{\beni}{\nuc{9}{Be}}
18:   \newcommand{\nm}{\ensuremath{N_\mathrm{max}}}
19:   \newcommand{\ho}{\ensuremath{\hbar \Omega}}
20:   \newcommand{\nn}{\ensuremath{N\!N}}
21:   \newcommand{\co}{(Color online)}
22: % Define command to print spin/parity/isospin for odd isotopes
23:   \newcommand{\jpt}[4]{\ensuremath{\left( \frac{#1}{2}_{_#2}^{#3} \:
24:       \frac{#4}{2} \right)}}
25: % ...and for even isotopes
26:   \newcommand{\jpte}[4]{\ensuremath{\left( #1_{#2}^{#3} \: #4 \right)}}
27: % Short-hands for the different interactions
28:  \newcommand{\avp}{AV8$^\prime$}
29:  \newcommand{\cdb}{CDB2k}
30:  \newcommand{\nlo}{N$^3$LO}
31:  \newcommand{\inoy}{INOY}
32: % Comments/notes  (to be changed)
33:  \newcommand{\nb}[1]{\textcolor{red}{\emph{[#1]}}}
34: %-------------------------
35: %**************************************************************************
36: \begin{document}
37: \preprint{UCRL-JRNL-208555}
38: \title{Large basis \emph{ab initio} shell model investigation of
39:   $^{9}${Be} and $^{11}${Be}}
40: \author{C. Forss\'en}
41: \email[]{c.forssen@llnl.gov}
42: \author{P. Navr\'atil}
43: \author{W.E. Ormand}
44: \affiliation{Lawrence Livermore National Laboratory, P.O. Box 808, L-414, 
45: Livermore, CA  94551}
46: \author{E. Caurier}
47: \affiliation{Institut de Recherches Subatomiques
48:             (IN2P3-CNRS-Universit\'e Louis Pasteur)\\
49:             Batiment 27/1,
50:             67037 Strasbourg Cedex 2, France}
51: 
52: \date{\today}
53: 
54: \begin{abstract}
55: We are presenting the first \emph{ab initio} structure investigation of
56: the loosely bound $^{11}${Be} nucleus, together with a study of the
57: lighter isotope $^{9}${Be}. The nuclear structure of these isotopes is
58: particularly interesting due to the appearance of a parity-inverted
59: ground state in $^{11}${Be}. Our study is performed in the framework of
60: the \emph{ab initio} no-core shell model. Results obtained using four
61: different, high-precision two-nucleon interactions, in model spaces up
62: to 9$\hbar\Omega$, are shown. For both nuclei, and all potentials, we
63: reach convergence in the level ordering of positive- and negative-parity
64: spectra separately. Concerning their relative position, the
65: positive-parity states are always too high in excitation energy, but a
66: fast drop with respect to the negative-parity spectrum is observed when
67: the model space is increased. This behavior is most dramatic for
68: $^{11}${Be}. In the largest model space we were able to reach, the $1/2^+$
69: level has dropped down to become either the first or the second excited
70: state, depending on which interaction we use. We also observe a
71: contrasting behavior in the convergence patterns for different
72: two-nucleon potentials, and argue that a three-nucleon interaction is
73: needed to explain the parity inversion. Furthermore, large-basis
74: calculations of $^{13}${C} and $^{11}${B} are performed. This allows
75: us to study the systematics of the position of the first
76: unnatural-parity state in the $N=7$ isotone and the $A=11$ isobar. The
77: $^{11}${B} run in the $9\hbar\Omega$ model space involves a matrix with
78: dimension exceeding $1.1 \times 10^9$, and is our largest calculation so
79: far. We present results on binding energies, excitation spectra, level
80: configurations, radii, electromagnetic observables, and $^{10}\mathrm{Be} +
81: n$ overlap functions.
82: \end{abstract}
83: % insert suggested PACS numbers in braces on next line
84: \pacs{21.60.Cs, 21.45.+v, 21.30.-x, 21.30.Fe, 27.20.+n}
85: \maketitle
86: %
87: %**************************************************************************
88: \section{\label{sec:intro}Introduction}
89: %
90: Studies of how nuclear structure evolves when varying the $N/Z$ ratio
91: are important in order to improve our fundamental understanding of
92: nuclear forces. For this reason, research on light neutron-rich nuclei
93: has attracted an increasing amount of theoretical and experimental
94: effort since the advent of radioactive nuclear beams. The application of
95: a standard mean-field picture to describe these few-body systems is
96: questionable, and it is not surprising that substantial deviations from
97: regular shell structure has been observed. The $A=11$ isobar is of
98: particular interest in this respect since it exhibits some anomalous
99: features that are not easily explained in a simple shell-model
100: framework. Most importantly, the parity-inverted $1/2^+$ ground state of
101: \beel\ was noticed by Talmi and Unna~\cite{tal60:4} already in the early
102: 1960s, and it still remains one of the best examples of the
103: disappearance of the $N=8$ magic number.
104: 
105: Many theoretical studies of odd-$A$ beryllium isotopes have already
106: been performed using various models. A thourough review of the
107: structure of unstable light nuclei in terms of the shell model can be
108: found in Ref.~\cite{mil01:693}. Of particular interest is the study on
109: unnatural-parity states of the $A=11$ isobar by Teeters and
110: Kurath~\cite{tee77:275} using a 1\ho\ model space and the
111: Millener-Kurath interaction with modified $0s$ and $sd$
112: single-particle energies. The halo structure of the \beel\ ground
113: state was reproduced with the variational shell model by Otsuka
114: \emph{et al}~\cite{ots93:70}. They used Skyrme interactions and
115: constructed multi-nucleon wave functions from a variational
116: single-particle basis in a (0--1)\ho\ model space. Alternatively, the
117: loosely-bound nature of the valence neutron in \beel\ can be treated
118: explicitly in a $\nuc{10}{Be} + n$ picture with a Woods-Saxon
119: potential, see e.g.  Refs.~\cite{esb95:51,nun96:596}. Using a
120: coupled-channels treatment, the authors of these papers found a
121: significant overlap with excited-core states. Possible explanations
122: for the parity inversion of the \beel\ ground state has also been
123: investigated using the AMD+HF model~\cite{dot00:103}, which is a
124: combination of anti-symmetrized molecular dynamics with the concept of
125: single-particle motion. An extended version of the AMD framework was
126: later used to study excited states of \beel, and the existence of
127: three negative-parity rotational bands was proposed~\cite{kan02:66}.
128: 
129: There are also several calculations involving different cluster
130: models. In particular, $\alpha$-clustering has been considered to play
131: an important role in these systems. With this assumption as a starting
132: point, K. Arai \emph{et al} used an $\alpha + \alpha + n$ model and
133: obtained the ground state of \beni\ using the stochastical variational
134: method, while several particle-unbound excited states were studied
135: simultaneously with the complex scaling method~\cite{ara96:54}. A
136: similar $\alpha + \alpha + Xn$ description was employed by
137: P. Descouvemont~\cite{des02:699} in his study of possible rotational
138: bands in \nuc{9-11}{Be} using the Generator Coordinate Method. His
139: conclusion, however, was that the degree of $\alpha$-clustering
140: decreased with increasing mass, and consequently his model was not able
141: to reproduce some of the anomalous properties of \beel.
142: 
143: Unfortunately, due to the complexity of the problem, there has been no
144: genuine \emph{ab initio} investigation of \beel\ starting from realistic
145: inter-nucleon interactions. There is no doubt that the cluster and
146: potential models are very successful, and can provide reasonable
147: explanations for many observations. Still, one has to remember that they
148: rely upon the fundamental approximation that the total wave function can
149: be separated into cluster components. Furthermore, the effective
150: interactions used in all models must be fitted to some observables for
151: each individual case. On the contrary, a truly microscopic theory such
152: as the Green's Function Monte Carlo (GFMC) method~\cite{pud97:56}, or
153: the \emph{ab initio} no-core shell model
154: (NCSM)~\cite{nav00:84,nav00:62}, starts from a realistic inter-nucleon
155: potential and solves the $A$-body problem, producing an antisymmetrized
156: total wave function. It is a true challenge of our understanding of
157: atomic nuclei to investigate nuclear many-body systems ($A > 4$) using
158: such \emph{ab initio} approaches.
159: 
160: This paper represents an effort to fill this gap. Our study is performed
161: in the framework of the \emph{ab initio} NCSM, in which the $A$-body
162: Schr\"odinger equation is solved using a large Slater determinant
163: harmonic oscillator (HO) basis. However, it is well known that the HO
164: basis functions have incorrect asymptotics which might be a problem when
165: trying to describe loosely bound systems. Therefore it is desirable to
166: include as many terms as possible in the expansion of the total wave
167: function. By restricting our study to two-nucleon (\nn) interactions,
168: even though the NCSM allows for the inclusion of three-body forces, we
169: are able to maximize the model space and to better observe the
170: convergence of our results.
171: 
172: In Sec.~\ref{sec:theory} the framework for the NCSM will be briefly
173: outlined, and the four different high-precision \nn\ interactions that
174: are used in this work will be introduced. Sec.~\ref{sec:results} is
175: devoted to a presentation and discussion of our complete set of results
176: for \nuc{9,11}{Be}, with a particular focus on the position of the first
177: unnatural-parity state. Concluding remarks are presented in
178: Sec.~\ref{sec:conc}.
179: %
180: %************************************************************************** 
181: \section{\label{sec:theory}\emph{Ab initio} no-core shell model}
182: %
183: Applying the \emph{ab initio} NCSM is a multi-step process. The first
184: step is to derive the effective interaction from the underlying
185: inter-nucleon forces, and to transform it from relative coordinates into
186: the single-particle $M$-scheme basis. The second step is to evaluate and
187: diagonalize the effective Hamiltonian in an $A$-nucleon ($Z$ protons and
188: $N$ neutrons) Slater determinant HO basis that spans the complete $\nm
189: \ho$ model space. Finally, we can use the resulting wave functions for
190: further processing. This section contains a short discussion on each of
191: these steps. We stress that an important strength of the method is the
192: possibility to include virtually any type of inter-nucleon
193: potential. The four high-precision \nn\ interactions that have been used
194: in this study will be introduced in Sec.~\ref{sec:int}. A more detailed
195: description of the NCSM approach, as it is implemented in this study,
196: can be found in, e.g., Ref.~\cite{nav00:62}.
197: %
198: \subsection{Hamiltonian and effective interactions}
199: %
200: The goal is to solve the $A$-body Schr\"odinger equation with an
201: intrinsic Hamiltonian of the form
202: %
203: \begin{equation}
204:     H_A= %T_{\rm rel} + {\cal V} =
205:     \frac{1}{A}\sum_{i<j}^{A}\frac{(\vec{p}_i-\vec{p}_j)^2}{2m}
206:     + \sum_{i<j}^A V_{\nn, ij}  \; ,
207:     \label{eq:H}
208: \end{equation}
209: %
210: where $m$ is the nucleon mass and $V_{\nn, ij}$ is the \nn\ interaction
211: including both strong and electromagnetic components. As mentioned
212: earlier, we will not use three-body forces in this study since we strive
213: to maximize the size of the model space. By adding a center-of-mass (CM)
214: HO Hamiltonian $H_\mathrm{CM}^\Omega = T_\mathrm{CM} +
215: U_\mathrm{CM}^\Omega$ (where $U_\mathrm{CM}^\Omega = A m \Omega^2
216: \vec{R}^2 / 2, \; \vec{R} = \sum_{i=1}^A \vec{r} / A$), we facilitate
217: the use of a convenient HO basis. The modified Hamiltonian can be
218: separated into one- and two-body terms
219: %
220: %\nb{Split equation into three lines}
221: %
222: \begin{equation}
223:   \begin{split}
224:     H_A^\Omega = & H_A + H_\mathrm{CM}^\Omega = \sum_{i=1}^A h_i +
225:     \sum_{i<j}^A V_{ij}^{\Omega,A} 
226:     \\ 
227:     = &\sum_{i=1}^A \left[ \frac{\vec{p}_i^2}{2m}
228:     +\frac{1}{2}m\Omega^2 \vec{r}^2_i
229:     \right] 
230:     \\
231:     &+ \sum_{i<j}^A \left[ V_{{\nn}, ij}
232:     -\frac{m\Omega^2}{2A} (\vec{r}_i-\vec{r}_j)^2 \right] \; .
233:     \label{eq:Homega}
234:   \end{split}
235: \end{equation}
236: %
237: The next step is to divide the $A$-nucleon infinite HO basis space into
238: an active, finite model space ($P$) and an excluded space ($Q = 1-
239: P$). The model space consists of all configurations with $\leq \nm \ho$
240: excitations above the unperturbed ground state. In this approach there
241: is no closed-shell core; meaning that all nucleons are active.
242: 
243: Since we solve the many-body problem in a finite model space, the
244: realistic \nn\ interaction will yield pathological results because of
245: the short-range repulsion. Consequently, we employ effective interaction
246: theory. In the \emph{ab initio} NCSM approach, the model-space dependent
247: effective interaction is constructed by performing a unitary
248: transformation of the Hamiltonian~\eqref{eq:Homega}, $e^{-S} H_A^\Omega
249: e^S$, such that the model space and the excluded space are decoupled $Q
250: e^{-S} H_A^\Omega e^S P = 0$. This procedure has been discussed by Lee
251: and Suzuki~\cite{suz80:64,suz82:68}, and yields a Hermitian effective
252: interaction $H_\mathrm{eff} = Pe^{-S} H_A^\Omega e^S P$ which acts in
253: the model space and reproduces exactly a subset of the eigenspectrum to
254: the full-space Hamiltonian. In general, this effective Hamiltonian will
255: be an $A$-body operator which is essentially as difficult to construct
256: as to solve the full $A$-body problem. In this study we approximate the
257: effective Hamiltonian at the two-body cluster level. The basic idea is
258: to derive it from high-precision solutions to the two-body problem with
259: $\mathcal{H}_2 = h_1 + h_2 + V_{12}^{\Omega,A}$, where the two-body term
260: is the same as in Eq.~\eqref{eq:Homega}. The final result will be a
261: two-body effective interaction $V_{12,\mathrm{eff}}^{\Omega,A}$. See
262: Ref.~\cite{nav00:62,cau02:66} for details.
263: 
264: We note that our approximated effective interaction will depend on the
265: nucleon number $A$, the HO frequency $\Omega$, and the size of the model
266: space which is defined by \nm. However, by construction, it will
267: approach the starting bare interaction $V_{ij,\mathrm{eff}}^{\Omega,A}
268: \to V_{ij}^{\Omega,A}$, as $\nm \to \infty$. Consequently, the dependence on
269: $\Omega$ will decrease with increasing model space, and the NCSM results
270: will converge to the exact solution. A very important feature of the
271: NCSM is the fact that the effective interaction is translationally
272: invariant so that the solutions can be factorized into a CM component
273: times a wave function corresponding to the internal degrees of
274: freedom. Due to this property it is straightforward to remove CM effects
275: completely from all observables.
276: %
277: \subsection{Solution of the many-body {S}chr\"odinger equation}
278: %
279: Once the effective interaction has been derived, we can construct the
280: effective Hamiltonian in the $A$-body space. In this process we
281: subtract the CM Hamiltonian $H_\mathrm{CM}^\Omega$, and add
282: the Lawson projection term $\beta(H_\mathrm{CM} - \frac{3}{2}\ho )$ to
283: shift eigenstates with excited CM motion up to high energies. States
284: with $0S$ CM motion are not affected by this term and, consequently,
285: their eigenenergies will be independent of the particular choice of
286: $\beta$. We are now left with a Hamiltonian of the form
287: %
288: %\nb{Split equation using multline}
289: %
290: %\begin{equation}
291: \begin{multline}
292:   H_{A,\mathrm{eff}}^\Omega = P \Bigg\{ \sum_{i<j}^A 
293:   \bigg[ \frac{(\vec{p}_i-\vec{p}_j)^2}{2Am}
294:   +\frac{m\Omega^2}{2A} (\vec{r}_i-\vec{r}_j)^2 
295:   \\
296:   + V_{ij,\mathrm{eff}}^{\Omega,A} \bigg]  + 
297:   \beta \left( H_\mathrm{CM} - \frac{3}{2}\ho \right) \Bigg\} P \; .
298:   \label{eq:Homegaeff}
299: \end{multline}
300: %\end{equation}
301: %
302: 
303: The computational problem of obtaining the many-body eigenvalues is
304: non-trivial due to the very large matrix dimensions involved. The
305: largest model space that we encountered in this study was the
306: \nuc{11}{B} 9\ho\ (positive parity) space, for which the dimension
307: exceeds $d_P = 1.1 \times 10^9$. For \beel(\beni), the 9\ho\ space gives
308: $d_P = 7.1 \times 10^8(2.0 \times 10^8)$. To solve this problem we have
309: used a specialized version of the shell model code
310: \textsc{antoine}~\cite{cau99:30,cau99:59}, recently adapted to the
311: NCSM~\cite{cau01:64}. This code works in the $M$ scheme for basis
312: states, and uses the Lanczos algorithm for diagonalization. The number
313: of iterations needed to converge the first eigenstates is significantly
314: reduced by the implementation of a sophisticated strategy for selecting
315: the pivot vectors. This feature of the code is absolutely crucial when
316: using it to perform calculations in very large model spaces.
317: 
318: Furthermore, the code takes advantage of the fact that the dimensions of
319: the neutron and proton spaces are small with respect to the full
320: dimension. Therefore, before the diagonalization, all the matrix
321: elements involving one- and two-body operators acting in a single
322: subspace (proton or neutron) are calculated and stored. As a
323: consequence, during the Lanczos procedure, all non-zero proton-proton
324: and neutron-neutron matrix elements can be generated with a simple
325: loop. Furthermore, the proton-neutron matrix elements are obtained with
326: three integer additions~\cite{cau99:30}. However, for no-core
327: calculations (in which all nucleons are active) the number of shells
328: and, consequently, the number of matrix elements that are precalculated,
329: becomes very large. Consider, e.g., the \nuc{11}{B} calculation in the
330: 9\ho\ space. The full dimension is $d_P = 1.1 \times 10^9$, while the
331: number of active shells is 66 and the total number of neutron plus
332: proton Slater determinants is $N(n) + N(p) = 1.0 \times 10^7$. This
333: corresponds to 80~Gb of precalculated and stored matrix elements. In
334: contrast, consider a shell-model calculation of \nuc{57}{Ni} in the full
335: $fp$ space. The total dimension is larger, $d_P = 1.4 \times 10^9$,
336: but there are only four active shells which gives $N(n) + N(p) = 1.8
337: \times 10^6$ and requires merely 1~Gb of precalculated data.
338: 
339: A recent development of the NCSM is the ability to further process the
340: wave functions, resulting from the shell-model calculation, to obtain
341: translationally invariant densities~\cite{nav04:70} and cluster form
342: factors~\cite{nav04:70_2}. The latter can be used to obtain
343: spectroscopic factors, but can also serve as a starting point for an
344: \emph{ab initio} description of low-energy nuclear reactions. We have
345: employed these new capabilities to study the overlap of \beel\ with
346: different $\nuc{10}{Be}+n$ channels.
347: %
348: \subsection{\label{sec:int}Realistic \nn\ interactions}
349: %
350: Four different, high-precision \nn\ interactions have been used in this
351: study. These are: the Argonne V8$^\prime$
352: (\avp)~\cite{wir95:51,pud97:56}, the CD-Bonn 2000
353: (\cdb)~\cite{mac01:63}, the N$^3$LO~\cite{ent03:68}, and the
354: INOY~\cite{dol04:69,dol03:67} potentials. We can divide these
355: interactions into three different types:
356: 
357: \emph{1. Local in coordinate space:} The \avp\ interaction is
358: an isospin-invariant subset of the phenomenological Argonne $v_{18}$
359: potential~\cite{wir95:51} plus a screened Coulomb potential. This
360: interaction is local in coordinate space and it is also employed in
361: the Green's Function Monte Carlo (GFMC)
362: approach~\cite{pud97:56}. Consequently, the use of this potential
363: allows for a direct comparison of results from the NCSM and the GFMC
364: methods.
365: 
366: \emph{2. Non-Local in momentum space:} The \cdb\
367: interaction~\cite{mac01:63} is a charge-dependent \nn\ interaction based
368: on one-boson exchange. It is described in terms of covariant Feynman
369: amplitudes, which are non-local. Consequently, the off-shell behavior of
370: the CD-Bonn interaction differs from commonly used local potentials
371: which leads to larger binding energies in nuclear few-body systems. The
372: newly developed \nlo\ interaction~\cite{ent03:68} is also represented by
373: a finite number of Feynman diagrams. This interaction, however, is based
374: on chiral perturbation theory at the fourth order, which means that it
375: is derived from a Lagrangian that is consistent with the symmetries of
376: QCD. It represents a novel development in the theory of nuclear
377: forces. It is particularly interesting to note that many-body forces
378: appear naturally already at the next-to-next-to-leading order (NNLO) of
379: this low-energy expansion. However, in this study we use solely the \nn\
380: part of the \nlo\ interaction. This \nn\ potential has previously been
381: applied in the NCSM approach to study the $p$-shell nuclei \nuc{6}{Li}
382: and \nuc{10}{B}~\cite{nav04:69}.
383: 
384: \emph{3. Non-Local in coordinate space:} A new type of interaction,
385: which respects the local behavior of traditional \nn\ interactions at
386: longer ranges but exhibits a non-locality at shorter distances, was
387: recently proposed by Doleschall \emph{et al}~\cite{dol04:69,dol03:67}.
388: The authors are exploring the extent to which effects of multi-nucleon
389: forces can be absorbed by non-local terms in the \nn\ interaction. Their
390: goal was to investigate if it is possible to introduce non-locality in
391: the \nn\ interaction so that it correctly describes the three-nucleon
392: bound states \nuc{3}{H} and \nuc{3}{He}, while still reproducing \nn\
393: scattering data with high precision. Note that all other \nn\
394: interactions give a large underbinding of $A \geq 3$ systems. In
395: practice, the \inoy\ interaction was constructed by combining an inner
396: ($<3$~fm) phenomenological non-local part with a local Yukawa
397: tail. Hence the name INOY (Inside Nonlocal Outside Yukawa). The so
398: called IS version of this interaction, introduced in
399: Ref.~\cite{dol03:67}, contains short-range non-local potentials in
400: $^1S_0$ and $^3S_1-^{3\!\!}D_1$ partial waves while higher partial waves are
401: taken from Argonne $v_{18}$. In this study we are using the IS-M
402: version, which includes non-local potentials also in the $P$ and $D$
403: waves~\cite{dol04:69}. It is important to note that, for this particular
404: version, the on-shell properties of the triplet $P$-wave interactions
405: have been modified in order to improve the description of $3N$ analyzing
406: powers. The $^{3\!}P_0$ interaction was adjusted to become less attractive,
407: the $^3P_1$ became more repulsive, and the $^{3\!}P_2$ more
408: attractive. Unfortunately, this gives a slightly worse fit to the
409: Nijmegen $^{3\!}P$ phase shifts.
410: %
411: %**************************************************************************
412: \section{\label{sec:results}Results}
413: %
414: By construction, the \emph{ab initio} NCSM method is guaranteed to
415: converge either by calculating the effective interaction using a fixed
416: cluster approximation (e.g., two-body) and then solving the eigenvalue
417: problem in increasing model spaces ($\nm \to \infty$), or by working in
418: a limited model space but increasing the clustering of the effective
419: interaction towards the full $A$-body one. Our codes are currently
420: constructed to derive effective interactions up to the level of
421: three-body clustering (with or without three-body forces). However, in
422: this study we have chosen to approach convergence by trying to maximize
423: our model space and, therefore, we limit ourselves to the use of
424: two-body effective interactions. Thus we are able to reach the 9\ho\
425: model space for nuclei with $A=11$. This maximal space corresponds to
426: basis dimensions of $d_P =$ $2.0 \times 10^8$ (\beni), $7.0 \times 10^8$
427: (\beel), and $1.1 \times 10^9$ (\nuc{11}{B}). For \nuc{13}{C}, which is
428: briefly discussed in connection to the parity-inversion problem, the
429: largest space that we were able to reach was $8\ho$ ($d_P = 8.2 \times
430: 10^8$).
431: 
432: Note that model spaces with an even(odd) number of HO excitations give
433: negative-(positive-)parity states for the nuclei under study. When
434: constructing a full spectrum we combine the $\nm\ho$ and $(\nm+1)\ho$
435: results, with \nm\ being an even number. In connection to this, we should
436: also point out that very few states in \beni\ and \beel\ are particle
437: stable. However, in the NCSM approach, all states are artificially bound
438: due to the truncated model space and the use of HO basis functions.
439: 
440: In addition to a careful study of the level ordering in \beni\ and
441: \beel, with a particular focus on the position of the positive-parity
442: states, we also calculate electromagnetic moments and transition
443: strengths. For this we use traditional one-body transition operators
444: with free-nucleon charges. Note that, due to the factorization of our
445: wave function into CM and intrinsic components, we obtain
446: translationally invariant matrix elements for all observables that we
447: investigate in this work. However, we have not renormalized the
448: operators, which means that the results are not corrected for the fact
449: that we work in a truncated model space. The theoretical framework for
450: performing this renormalization is in place, and the process is
451: underway~\cite{ste04:nucl-th}. Until we are ready to implement the use
452: of effective operators, we can get an indication on the need for
453: renormalization by studying the basis-size dependence of our calculated
454: observables .
455: 
456: The cluster decomposition of the \beel\ ground state into
457: $\nuc{10}{Be}+n$ is of particular interest due to the small neutron
458: separation energy. We have employed the formalism recently developed in
459: Refs.~\cite{nav04:70,nav04:70_2} to calculate cluster overlap functions
460: using our NCSM wave functions. These results are presented in
461: Sec.~\ref{sec:11be}.
462: %
463: %**************************************************************************
464: \subsection{\label{sec:hodep}Dependence on HO frequency}
465: %-------------------------
466: %
467: The first step in our study is a search for the optimal HO frequency. In
468: principle, the intrinsic properties of the nucleus should not depend on
469: the particular value of \ho\ since it only enters the
470: Hamiltonian~\eqref{eq:Homega} through a CM-dependent term. In practice
471: however, due to the cluster approximation of the effective interaction,
472: our results will be sensitive to the choice of \ho. Furthermore, by
473: construction, the effective interactions depend on the size of the model
474: space, \nm, and on the number of nucleons, $A$. In order to investigate
475: these dependences we have performed a large series of calculations for a
476: sequence of frequencies. The results from this study are presented in
477: Figs.~\ref{fig:9freqdep} and \ref{fig:11freqdep} (for \beni\ and \beel,
478: respectively) as curves showing the frequency dependence of the binding
479: energy in different model spaces.  We have studied this dependence for
480: the lowest state of each parity.  We are looking for the region in which
481: the dependence on $\Omega$ is the smallest; and we select this frequency
482: (from the calculation in the largest model space) to use in our detailed
483: investigation of excited states. In our present case, this optimal
484: frequency always corresponds to an energy minimum. Note, however, that
485: the NCSM is not a variational method and the convergence of the binding
486: energy with increasing model space is not always from above.
487: %
488: \begin{figure*}[hbtp]
489:   \begin{minipage}{0.95\textwidth}
490:     \begin{minipage}[t]{0.47\textwidth}
491:       \centering
492:       \mbox{}\\
493:       \includegraphics*[width=\textwidth]{fig1a_a9v8p.eps}\\
494:       \includegraphics*[width=\textwidth]{fig1c_a9n3lo.eps}\\ 
495:       \mbox{}
496:     \end{minipage}
497:   \hfill
498:     \begin{minipage}[t]{0.47\textwidth}
499:       \centering
500:       \mbox{}\\
501:       \includegraphics*[width=\textwidth]{fig1b_a9cdb.eps}\\
502:       \includegraphics*[width=\textwidth]{fig1d_a9inoy.eps}\\
503:       \mbox{}
504:     \end{minipage}
505:   \end{minipage}
506:   \caption{\co\ The dependence on HO frequency for the calculated
507:   \beni\jpt{3}{1}{-}{1} (solid lines) and \beni\jpt{1}{1}{+}{1} (dashed
508:   lines) binding energies. Each panel correspond to a particular \nn\
509:   interaction: (a) \avp, (b) \cdb, (c) \nlo, (d) \inoy; and each
510:   separate line corresponds to a specific model space. The insets
511:   demonstrate how the minima of the curves converge as the model space
512:   is increased. The horizontal lines are the experimental values.%
513:   \label{fig:9freqdep}}
514: \end{figure*}
515: %
516: %
517: \begin{figure*}[hbtp]
518:   \begin{minipage}{0.95\textwidth}
519:     \begin{minipage}[t]{0.47\textwidth}
520:       \centering
521:       \mbox{}\\
522:       \includegraphics*[width=\textwidth]{fig2a_a11v8p.eps}\\
523:       \includegraphics*[width=\textwidth]{fig2c_a11n3lo.eps}\\ 
524:       \mbox{}
525:     \end{minipage}
526:   \hfill
527:     \begin{minipage}[t]{0.47\textwidth}
528:       \centering
529:       \mbox{}\\
530:       \includegraphics*[width=\textwidth]{fig2b_a11cdb.eps}\\ 
531:       \includegraphics*[width=\textwidth]{fig2d_a11inoy.eps}\\
532:       \mbox{}
533:     \end{minipage}
534:   \end{minipage}
535:   \caption{\co\ The dependence on HO frequency for the calculated
536:   \beel\jpt{1}{1}{-}{3} (solid lines) and \beel\jpt{1}{1}{+}{3} (dashed
537:   lines) binding energies. Each panel correspond to a particular \nn\
538:   interaction: (a) \avp, (b) \cdb, (c) \nlo, (d) \inoy; and each
539:   separate line corresponds to a specific model space. The insets
540:   demonstrate how the minima of the curves converge as the model space
541:   is increased. The horizontal lines are the experimental values.%
542:   \label{fig:11freqdep}}
543: \end{figure*}
544: %
545: 
546: Following this procedure for each nucleus and interaction, we obtain the
547: optimal HO frequencies that are listed in Table~\ref{tab:optfreq}.  A
548: few general remarks regarding the HO dependence, observed for the
549: different interactions in Figs.~\ref{fig:9freqdep} and
550: \ref{fig:11freqdep}, can be made: (1) Clear signs of convergence is
551: observed. The dependence on $\Omega$ becomes weaker with increasing size
552: of the model space, and the relative difference between the calculated
553: ground-state energies is in general decreasing. Furthermore, the optimal
554: frequencies for the largest model spaces of each parity (8\ho\ and 9\ho)
555: are approximately the same. This motivates our use of a single frequency
556: to compute both positive- and negative-parity states; (2) This single
557: frequency is found to be in the range of about \ho = 11--13~MeV for all
558: interactions except for \inoy, which seems to prefer a significantly
559: larger HO frequency (\ho = 16--17~MeV); (3) In general, the behavior of
560: the \avp, \cdb, and \nlo\ interactions are very similar, but with \nlo\
561: having the largest dependence on \ho; (4) As could be expected, since it
562: is the only \nn\ potential which is capable of reproducing $3N$ binding
563: energies, the \inoy\ interaction exhibits a distinctively different
564: behavior compared to the three others. The dependence on $\Omega$ is
565: encouragingly small, but the ground-state energy is still changing with
566: increasing basis size. This is particularly true for the positive-parity
567: state. We also note, from the insets of Figs.~\ref{fig:9freqdep} and
568: \ref{fig:11freqdep}, that it is the only interaction for which the
569: resulting binding energies are approaching the experimental values.
570: %
571: \begin{table}[hbtp]
572:   \caption{Selected optimal HO frequencies (in [MeV]). These choices are
573:     based on the frequency variation studies presented in
574:     Figs.~\ref{fig:9freqdep} and \ref{fig:11freqdep}.\vspace*{1ex}%
575:   \label{tab:optfreq}}
576:   \begin{ruledtabular}
577:     \begin{tabular}{ccccc}
578:       Nucleus &  \multicolumn{4}{c}{Interaction}                      \\   
579:                        & \inoy\    &  \cdb\    & \nlo\     & \avp\     \\
580:       \hline
581:       \beni            & 16        & 12        & 11        & 12        \\
582:       \beel            & 17        & 13        & 12        & 12        \\
583:     \end{tabular}
584:   \end{ruledtabular}
585: \end{table}
586: %
587: %
588: %**************************************************************************
589: \subsection{\label{sec:9be}\beni}
590: %-------------------------
591: %
592: By studying the HO frequency dependence of the \beni\ binding energy
593: obtained with different \nn\ interactions (see Fig.~\ref{fig:9freqdep})
594: it is clear that the \cdb\ results have a slightly better convergence
595: rate and a weaker HO frequency dependence than \avp\ and \nlo. The
596: \inoy\ results display an even weaker frequency dependence, but the
597: binding energy is still moving with increasing \nm. It is clear from
598: Table~\ref{tab:9energies} that all interactions, with the possible
599: exception of \inoy, underbind the system. Actually, by studying the
600: convergence rate of the \inoy\ results in Fig.~\ref{fig:9freqdep}d, it seems
601: as if this interaction will eventually overbind \beni. This observation
602: confirms that the additional binding, usually provided by $3N$ forces,
603: can be produced by the \inoy\ interaction. The other three \nn\
604: interactions underbind by 12--14\%. The local \avp\ potential was also
605: used in a recent GFMC study of negative-parity states in
606: \beni~\cite{pie02:66}, and we find an excellent agreement with their
607: ground-state binding energy (see Table~\ref{tab:9energies}).
608: 
609: In principle, the frequency dependence for each excited state should be
610: studied in order to compute its energy. This is particularly true in our
611: case where we want to compare negative- and positive-parity states. It
612: is therefore very encouraging that we find the same optimal frequency
613: for the first positive-parity state as for the negative-parity ground
614: state; and we select this frequency to use in our detailed investigation
615: of excited states. In Figs.~\ref{fig:9v8pspec},~\ref{fig:9inoyspec} and
616: Table~\ref{tab:9energies} we present our NCSM low-energy spectra for
617: different \nn\ interactions and compare the results to known
618: experimental levels.
619: %
620: \begin{table*}[hbtp]
621:   \caption{Experimental and calculated energies (in [MeV]) of the lowest
622:     negative- and positive-parity states in \beni. Quadrupole and
623:     magnetic moments (in [$e$fm$^2$] and [$\mu_N$]) for the ground
624:     state, as well as E2 and M1 strengths for selected transitions (in
625:     [$e^2$fm$^4$] and [$\mu_N^2$]). Results for the \avp, \cdb, \nlo\,
626:     and \inoy\ \nn\ interactions are presented. These calculations were
627:     performed in the 8(9)$\ho$ model space for
628:     negative-(positive-)parity states, using the HO frequencies listed
629:     in Table~\ref{tab:optfreq}. The GFMC results~\cite{pie02:66} are
630:     shown for comparison. Experimental values are
631:     from~\cite{til04:745}. $E_{x^+}$ denotes the excitation energy
632:     relative to the lowest positive-parity state.\vspace*{1ex}%
633:   \label{tab:9energies}}
634:   \begin{ruledtabular}
635:     \begin{tabular}{c@{\extracolsep\fill}cc@{\extracolsep{8ex}}ccc@{\extracolsep\fill}c}
636:       \beni\  &        & \multicolumn{4}{c}{NCSM}                      & GFMC  \\   
637:               & Exp    & \inoy\    &  \cdb\    & \nlo\     & \avp\     & \avp\ \\
638:       \hline
639:       $E_\mathrm{gs}\jpt{3}{1}{-}{1}$ 
640:               & -58.16 & -56.05    & -51.16    & -50.47    & -50.20    & -49.9(2) \\ 
641:       $E_x \jpt{3}{1}{-}{1}$
642:               & 0      & 0         & 0         & 0         & 0         & 0      \\
643:       $E_x \jpt{5}{1}{-}{1}$
644:               & 2.43   & 2.96      & 2.78      & 2.64      & 2.70      & 2.1      \\
645:       $E_x \jpt{1}{1}{-}{1}$
646:               & 2.78   & 4.57      & 2.68      & 2.33      & 2.50      & 1.7      \\
647:       $E_x \jpt{3}{2}{-}{1}$
648:               & 5.59\footnote[1]{The experimental spin-parity
649:       assignment of this level is ``less certain'' according to
650:       the TUNL Nuclear Data Evaluation~\cite{til04:745}.}   
651:                        & 7.02      & 4.98      & 4.53      & 4.74      & ---       \\
652:       $E_x \jpt{7}{1}{-}{1}$
653:               & 6.38   & 8.09      & 7.80      & 7.40      & 7.56      & 6.4      \\
654:       \hline
655:       $E\jpt{1}{1}{+}{1}$ 
656:               & -56.48 & -50.95    & -47.81    & -47.57    & -46.84    & ---       \\ 
657:       $E\jpt{1}{1}{+}{1} - E_\mathrm{gs}$
658:               & 1.68   & 5.10      & 3.35      & 2.90      & 3.35      & ---       \\
659:       $E_{x^+} \jpt{1}{1}{+}{1}$
660:               & 0    & 0       & 0       & 0       & 0       & ---       \\
661:       $E_{x^+} \jpt{5}{1}{+}{1}$
662:               & 1.37   & 1.39      & 1.68      & 1.68      & 1.66      & ---       \\
663:       $E_{x^+} \jpt{3}{1}{+}{1}$
664:               & 3.02\footnotemark[1]   
665:                        & 4.06      & 3.60      & 3.37      & 3.49      & ---       \\
666:       $E_{x^+} \jpt{9}{1}{+}{1}$
667:               & 5.08   & 6.22      & 6.36      & 6.21      & 6.24      & ---       \\
668:       \hline
669:       $Q_\mathrm{gs}$ 
670:               & 5.288(38)& 3.52      & 4.01      & 4.21      & 4.01      & 5.0(3) \\
671:       $\mu_\mathrm{gs}$ 
672:               & -1.1778(9)& -1.06    & -1.22     & -1.24     & -1.22     & -1.35(2) \\
673:       \hline
674:       $B \left( \mathrm{E}2; \frac{5}{2}_{_1}^- \to \frac{3}{2}_{_1}^- \right)$ 
675:               & 27.1(2) & 10.9     &  14.9     & 16.7      & 15.0      & --- \\
676:       $B \left( \mathrm{M}1; \frac{5}{2}_{_1}^- \to \frac{3}{2}_{_1}^- \right)$ 
677:               & 0.54(6)  & 0.46    &  0.37     & 0.36      & 0.37      & --- \\
678:     \end{tabular}
679:   \end{ruledtabular}
680: \end{table*}
681: %
682: As can be seen, the \avp, \cdb\ and \nlo\
683: interactions give the same level ordering and very similar excitation
684: energies. It is noteworthy that all these high-precision \nn\
685: interactions perform equally when applied to the $A=9$ system. We let
686: the \avp\ spectrum shown in Fig.~\ref{fig:9v8pspec} be the graphical
687: representation of all of them. Using the \avp\ we can also make a
688: comparison to the recent GFMC calculation~\cite{pie02:66}. 
689: 
690: %
691: \begin{figure}[hbtp]
692:   \includegraphics*[width=0.9\columnwidth]{fig3_Be9v8p_spectra.eps}
693:   \caption{\co\ Excitation spectrum for \beni\ calculated using the
694:   \avp\ interaction in 0\ho--9\ho\ model spaces with a fixed HO
695:   frequency of $\hbar\Omega = 12$~MeV. The experimental values are from
696:   Ref.~\cite{til04:745}. The \avp\ results obtained by the GFMC
697:   method~\cite{pie02:66} are shown for comparison (note that only
698:   negative-parity states were computed). The two lower graphs show
699:   separately the negative- and positive-parity spectra, while the upper
700:   graph shows the combined spectrum with selected states.%
701:   \label{fig:9v8pspec}}
702: \end{figure}
703: %
704: %
705: \begin{figure}[hbtp]
706:   \includegraphics*[width=0.9\columnwidth]{fig4_Be9inoy_spectra.eps}
707:   \caption{\co\ Excitation spectrum for \beni\ calculated using the
708:   \inoy\ interaction in 0\ho--9\ho\ model spaces with
709:   a fixed HO frequency of $\hbar\Omega = 16$~MeV. The experimental
710:   values are from Ref.~\cite{til04:745}. The two lower graphs show
711:   separately the negative- and positive-parity spectra, while the upper
712:   graph shows the combined spectrum with selected states.%
713:   \label{fig:9inoyspec}}
714: \end{figure}
715: %
716: 
717: In general, we observe a very reasonable agreement with experimental
718: levels of natural parity, while the unnatural-parity states are
719: consistently high in excitation energy. For both parities, there is a
720: general trend of convergence with increasing model space. When plotting
721: the negative- and positive-parity spectra separately, it is evident that
722: the relative level spacings are almost independent on the model space,
723: so that the level ordering within each parity projection is remarkably
724: stable. It is clear, however, that the relative position of the
725: negative- versus positive-parity states is still not
726: converged. Furthermore, when studying the \avp\ convergence pattern in
727: the upper panel of Fig.~\ref{fig:9v8pspec}, it seems as if this
728: interaction will predict the positive-parity states at too high
729: excitation energies even when the calculations will be converged. This
730: finding is consistent with an overall trend observed in other NCSM
731: calculations, and it has been speculated whether a three-body force will
732: correct this behavior~\cite{cau02:66}. Although we are still not able to
733: apply a true three-body force in a large enough model space, we get some
734: indications from the performance of the \inoy\ interaction. In
735: Fig.~\ref{fig:9inoyspec} we see that, for this interaction, the
736: positive-parity states are even higher in small model spaces, but that
737: they are also dropping much faster with increasing \nm. This issue is
738: investigated in further detail in Sec.~\ref{sec:pospar} where we discuss
739: the important question of parity inversion, and the general trend of the
740: position of natural- versus unnatural-parity states.
741: 
742: In addition to an increase in binding energy, it has been found that the
743: level ordering for many nuclei can be sensitive to the presence of
744: multi-nucleon forces~\cite{nav02:88,nav03:68,pie02:66}. This sensitivity
745: is the largest for those states where the spin-orbit interaction
746: strength is known to play a role. For \beni\ we find that our
747: calculations with the \avp, \cdb\ and \nlo\ interactions predict the
748: first-excited negative-parity state to be a $1/2^-$, while experiments
749: show that it is a $5/2^-$ (Note, however, that the \cdb\ interaction
750: predicts these two states to be almost degenerate, and to exhibit a
751: convergence trend indicating a possible level crossing at larger model
752: spaces.). This level reversal was also found in the GFMC calculations
753: using \avp. The \inoy\ interaction, on the other hand, gives the correct
754: level ordering, but instead overpredicts the spin-orbit splitting. By
755: performing a calculation in a smaller model space using the \avp\ plus
756: the Tucson-Melbourne TM$^\prime$(99)~\cite{coo01:30} three-nucleon
757: interaction, we found a similar result as with \inoy.
758: 
759: Our discussion up to this point has been concentrated on the low-lying
760: levels in \beni. However, in response to the recent evaluation published
761: by the TUNL Nuclear Data Evaluation Project~\cite{til04:745}, we have
762: also decided to summarize our results for higher excited states. It is
763: important to note that the experimental widths of these states are
764: generally quite large, and to compute them correctly with the NCSM
765: method requires a very large model space. Furthermore, at high
766: excitation energies, it is very probable that there will be some
767: admixture of 2\ho\ intruders, and these are usually predicted too high
768: in the NCSM. In any case, our results can serve as an important
769: guideline as to which $p$-shell states that can be expected to appear in
770: the spectrum, and consequently should be looked for in experiments. In
771: Table~\ref{tab:9energiescdb}, we present all levels that we have
772: calculated using the \cdb\ interaction in the 8\ho\ and 9\ho\ model
773: spaces. We also show the tabulated experimental levels below $E_x =
774: 13$~MeV, taken from the most recent evaluation~\cite{til04:745}. A quick
775: comparison with the previous, published evaluation~\cite{ajz88:490}
776: (from 1988), reveals that several new levels have been discovered and
777: many spin-parity assignments have been changed. Altogether, these
778: changes lead to a much better agreement with our results. In the
779: negative-parity spectrum, our calculations give the correct level
780: ordering for the first six states. In particular, we correctly reproduce
781: the second $3/2^-$ and $5/2^-$ states that were introduced in the new
782: evaluation. On the other hand, we find a third $3/2^-$ state and a
783: second $1/2^-$ state that have not been observed in
784: experiments. However, these states are not fully converged in our 8\ho\
785: calculation, and they are still moving towards higher excitation
786: energies. We also find a $9/2^-$ state, which is fairly stable, and that
787: has not been experimentally identified. In the positive-parity spectrum,
788: the 6.76~MeV level has now been changed to being a $9/2^+$ which agrees
789: well with our level ordering. Finally, it is interesting to note that
790: our second $1/2^+$, $3/2^+$, and $5/2^+$ levels all appear in between
791: the first $9/2^+$ and $7/2^+$. None of these three states have, however,
792: been experimentally confirmed.
793: %
794: \begin{table}[hbtp]
795:   \caption{Experimental and calculated energies (in [MeV]) of the lowest
796:     negative- and positive-parity states in \beni. The calculations were
797:     performed in the 8(9)$\ho$ model space for
798:     negative-(positive-)parity states, using the \cdb\ \nn\ interaction
799:     with $\ho=12$~MeV. This table represents a more complete compilation
800:     of our computed levels (albeit for only one interaction) as compared
801:     to Table~\ref{tab:9energies}. Experimental values are
802:     from~\cite{til04:745}. $E_{x^+}$ denotes the excitation energy
803:     relative to the lowest positive-parity state.\vspace*{1ex}%
804:   \label{tab:9energiescdb}}
805:   \begin{ruledtabular}
806:     \begin{tabular}{ccc@{\hspace\fill}ccc}
807:       \multicolumn{3}{c}{Negative parity states} & 
808:       \multicolumn{3}{c}{Positive parity states} \\
809:       \beni\          & Exp    &  \cdb     & \beni          & Exp    &  \cdb    \\
810:       \hline
811:       $E_\mathrm{gs}\jpt{3}{1}{-}{1}$ 
812:                       & -58.16 & -51.16    & $E\jpt{1}{1}{+}{1}$ 
813:                                                             & -56.48 & -47.81    \\ 
814:       $E_x \jpt{3}{1}{-}{1}$
815:                       & 0      & 0         & $E\jpt{1}{1}{+}{1} - E_\mathrm{gs}$
816:                                                             & 1.68   & 3.35      \\
817:       $E_x \jpt{5}{1}{-}{1}$
818:                       & 2.43   & 2.78      & $E_{x^+} \jpt{1}{1}{+}{1}$
819:                                                             & 0    & 0           \\
820:       $E_x \jpt{1}{1}{-}{1}$
821:                       & 2.78   & 2.68      & $E_{x^+} \jpt{5}{1}{+}{1}$
822:                                                             & 1.37   & 1.68      \\
823:       $E_x \jpt{3}{2}{-}{1}$
824:                       & 5.59\footnote[1]{The experimental spin-parity
825:       assignment of this level is ``less certain'' according to
826:       the TUNL Nuclear Data Evaluation~\cite{til04:745}.}   
827:                                & 4.98      & $E_{x^+} \jpt{3}{1}{+}{1}$
828:                                                  & 3.02\footnotemark[1]   & 3.60      \\
829:       $E_x \jpt{7}{1}{-}{1}$
830:                       & 6.38   & 7.80      & $E_{x^+} \jpt{9}{1}{+}{1}$
831:                                                             & 5.08   & 6.36      \\
832: 
833:       $E_x \jpt{5}{2}{-}{1}$
834:                       & 7.94\footnotemark[1]   & 7.96      & $E_{x^+} \jpt{5}{2}{+}{1}$
835:                                                             &        & 7.66\footnote[2]
836: 		      {Calculated in a smaller, 7$\ho$, model space.}         \\
837:       $E_x \jpt{3}{3}{-}{1}$
838:                       &        & 11.26     & $E_{x^+} \jpt{3}{2}{+}{1}$
839:                                                             &        & 7.91\footnotemark[2]      \\
840:       $E_x \jpt{1}{2}{-}{1}$
841:                       &        & 11.86     & $E_{x^+} \jpt{1}{2}{+}{1}$
842:                                                             &        & 8.13\footnotemark[2]      \\
843:       $E_x \jpt{9}{1}{-}{1}$
844:                       &        & 12.45     & $E_{x^+} \jpt{7}{1}{+}{1}$
845:                                                             &        & 8.48      \\
846:       $E_x \jpt{7}{2}{-}{1}$
847:                       & 11.28\footnotemark[1]  & 12.61     \\
848:       $E_x \jpt{5}{3}{-}{1}$
849:                       & 11.81  & 13.02     \\
850:     \end{tabular}
851:   \end{ruledtabular}
852: \end{table}
853: %
854: 
855: In Table~\ref{tab:9energies}, we also present our results for the
856: ground-state quadrupole and magnetic moments, as well as for selected
857: electromagnetic transition strengths. We should stress that free-nucleon
858: charges have been used in these calculations, and that the operators
859: have not been renormalized. On the other hand, the stability of our
860: results can be judged by investigating the dependence on the model
861: space. We find that the calculated ground-state magnetic moment and the
862: $B(\mathrm{M1};\frac{5}{2}_{_1}^- \to \frac{3}{2}_{_1}^-)$ transition
863: strength are almost converged, and in fair agreement with the
864: experimental values. The results for electromagnetic quadrupole
865: observables are, however, steadily increasing with basis size
866: enlargement and should clearly benefit from the use of effective
867: operators. As an example we can consider the evolution of the \nlo\
868: results: For the $\left\{4\:-\:6\:-\:8\right\}\ho$ sequence of model
869: spaces these observables increase as $Q_\mathrm{gs} =
870: \left\{+3.96\:-\:+4.10\:-\:+4.21\right\}$~[$e$fm$^2$], and $B \left(
871: \mathrm{E}2; \frac{5}{2}_{_1}^- \to \frac{3}{2}_{_1}^- \right) =
872: \left\{14.9\:-\:15.7\:-\:16.7\right\}$~[$e^2$fm$^4$]. We would also like
873: to highlight the fact that \inoy\ gives much smaller values for
874: $Q_\mathrm{gs}$ and $B(\mathrm{E}2)$ than the other interactions. This
875: is partly due to the fact that, for \inoy, our selected frequency is
876: much larger than for the other potentials which, in our limited model
877: space, corresponds to a smaller radial scale. In principle, a HO
878: frequency dependence study should be made for each operator. However, we
879: have also applied the \inoy\ interaction in studies of $A=3,4$ systems,
880: for which convergence can be easily reached. It was found that, in
881: particular the rms proton radius is always underpredicted. The same
882: result was obtained in Ref.~\cite{laz04:70} through exact solutions of
883: the Faddeev-Yakubovski equations, and it demonstrates that the
884: interaction is too soft, resulting in a faster condensation of nuclear
885: matter. In this work, we have studied the \beni\ point-nucleon radii as
886: well as the strong E1 transition from the first-excited to the ground
887: state using the \avp\ interaction. However, since these results will be
888: compared to \beel\ data, we postpone the discussion to
889: Sec.~\ref{sec:11be}.
890: 
891: In Tables~\ref{tab:9conf} and~\ref{tab:9spocc}, we show the resulting
892: $N\ho$-configurations and the single-particle occupancies of the \beni\
893: wave function, obtained with the four different interactions. Although
894: these quantities are not physical observables, they can still give
895: interesting information. We see that the wave functions obtained with
896: the \avp, \cdb\ and \nlo\ interactions are almost identical, while the
897: \inoy\ wave function has a considerably larger fraction of low-\ho\
898: excitations. This fact is in part due to the higher HO frequency being
899: used in the \inoy\ calculations. Furthermore, from Table~\ref{tab:9conf}
900: we see that this particular interaction gives a different distribution
901: of \ho\ excitations for the \jpt{3}{1}{-}{1} and \jpt{1}{1}{+}{1}
902: states; which would indicate that the latter is slightly more
903: deformed. However, this behavior is not observed for the other
904: interactions. The differences in occupations of single-particle levels
905: reflect some properties of the interactions. The fact that the
906: $0p_{3/2}$ and $0d_{5/2}$ levels have a larger occupation in the \inoy\
907: eigenstates is direct evidence for a stronger spin-orbit interaction.
908: %
909: \begin{table}[hbtp]
910:   \caption{Calculated configurations of the first negative- and
911:     positive-parity states in \beni. Results obtained in our largest
912:     model spaces (8$\hbar\Omega$ and 9$\hbar\Omega$, respectively) are
913:     presented. The calculations were performed with the HO frequencies
914:     listed in Table~\ref{tab:optfreq}.\vspace*{1ex}%
915:   \label{tab:9conf}}
916:   \begin{ruledtabular}
917:     \begin{tabular}{lccccc}
918:       & \multicolumn{5}{c}{\beni\jpt{3}{1}{-}{1} (8$\hbar\Omega$ model space)}\\
919:       \nn\ interaction & 0$\hbar\Omega$ & 2$\hbar\Omega$ & 4$\hbar\Omega$
920:         & 6$\hbar\Omega$ & 8$\hbar\Omega$ \\
921:       \hline
922:       \inoy\ & 0.58 & 0.19 & 0.13 & 0.06 & 0.04 \\
923:       \cdb\  & 0.49 & 0.22 & 0.15 & 0.08 & 0.06 \\
924:       \nlo\  & 0.47 & 0.24 & 0.16 & 0.08 & 0.06 \\
925:       \avp\  & 0.49 & 0.22 & 0.15 & 0.08 & 0.06 \\[2ex]
926:       \hline
927:       & \multicolumn{5}{c}{\beni\jpt{1}{1}{+}{1} (9$\hbar\Omega$ model space)}\\
928:       \nn\ interaction & 1$\hbar\Omega$ & 3$\hbar\Omega$ & 5$\hbar\Omega$
929:         & 7$\hbar\Omega$ & 9$\hbar\Omega$ \\
930:       \hline
931:       \inoy\ & 0.53 & 0.22 & 0.14 & 0.07 & 0.04 \\
932:       \cdb\  & 0.48 & 0.23 & 0.15 & 0.08 & 0.06 \\
933:       \nlo\  & 0.46 & 0.24 & 0.16 & 0.08 & 0.05 \\
934:       \avp\  & 0.49 & 0.23 & 0.15 & 0.08 & 0.06 \\[2ex]
935:     \end{tabular}
936:   \end{ruledtabular}
937: \end{table}
938: %
939: %
940: \begin{table}[hbtp]
941:   \caption{Calculated occupations of neutron single-particle levels for
942:     the first negative- and positive-parity states in \beni. Results
943:     obtained in our largest model spaces (8$\hbar\Omega$ and
944:     9$\hbar\Omega$, respectively) are presented. The calculations were
945:     performed with the HO frequencies listed in
946:     Table~\ref{tab:optfreq}.\vspace*{1ex}%
947:   \label{tab:9spocc}}
948:   \begin{ruledtabular}
949:     \begin{tabular}{lcccccc}
950:       & \multicolumn{5}{c}{\beni\jpt{3}{1}{-}{1} (8$\hbar\Omega$ model space)}\\
951:       \nn\ interaction & $0s_{1/2}$ & $0p_{1/2}$ & $0p_{3/2}$
952:         & $1s_{1/2}$ & $0d_{3/2}$ & $0d_{5/2}$ \\
953:       \hline
954:       \inoy\ & 1.804& 0.454& 2.380& 0.043& 0.064& 0.069\\
955:       \cdb\  & 1.773& 0.511& 2.277& 0.067& 0.060& 0.072\\
956:       \nlo\  & 1.768& 0.521& 2.256& 0.077& 0.060& 0.072\\
957:       \avp\  & 1.778& 0.519& 2.273& 0.064& 0.061& 0.071\\[2ex]
958:       \hline
959:       & \multicolumn{5}{c}{\beni\jpt{1}{1}{+}{1} (9$\hbar\Omega$ model space)}\\
960:       \nn\ interaction & $0s_{1/2}$ & $0p_{1/2}$ & $0p_{3/2}$
961:         & $1s_{1/2}$ & $0d_{3/2}$ & $0d_{5/2}$ \\
962:       \hline
963:       \inoy\ & 1.792 & 0.498 & 1.428 & 0.573 &
964:       0.111 & 0.379 \\
965:       \cdb\  & 1.761& 0.530& 1.377& 0.666& 0.113& 0.317\\
966:       \nlo\  & 1.757& 0.536& 1.365& 0.676& 0.117& 0.312\\
967:       \avp\  & 1.765& 0.535& 1.376& 0.664& 0.113& 0.313 \\[2ex]
968:     \end{tabular}
969:   \end{ruledtabular}
970: \end{table}
971: %**************************************************************************
972: \subsection{\label{sec:11be}\beel}
973: %
974: Most of the observations made for \beni\ concerning the frequency
975: dependence of the calculated binding energies also hold true for
976: \beel. In general, however, the sensitivity to \ho\ is stronger in the
977: \beel\ case. Another important remark, that can be made from studying
978: Fig.~\ref{fig:11freqdep}, is that the binding energy of the first
979: positive-parity state calculated with the \inoy\ interaction is clearly not
980: converged. The relative shift in energy is actually slowly increasing
981: with model-space enlargement.
982: 
983: The experimental ground state of \beel\ is an intruder $1/2^+$ level,
984: while the first $p$-shell state is a $1/2^-$ situated at
985: $E_x=320$~keV. The neutron separation energy is only 503~keV, and there
986: are no additional bound states. This level-ordering anomaly constitutes
987: the famous parity-inversion problem. A number of excited states have
988: been observed in different reactions and beta-decay studies. However, as
989: can be seen from the summary presented in Table~\ref{tab:explevels},
990: there are considerable ambiguities in the spin-parity assignments.
991: %
992: \begin{table*}[hbtp]
993:   \caption{Present situation of the spin-parity assignments for the
994:     lowest states in \beel. The table contains published results from
995:     the FAS evaluation of 1990~\cite{ajz90:506rev} and from more recent
996:     experimental studies. These studies include direct reactions such as
997:     $(t,p)$ (Liu-Fortune) and $\nuc{12}{C}(\beel,\beel')$ (Fukuda) in
998:     which the extracted angular distributions were analyzed using DWBA
999:     theory. The remaining references are measurements of $\beta$-delayed
1000:     neutrons in coincidence with $\gamma$-rays. All decays that were
1001:     observed in these experiments had $\log(ft)$ values that were
1002:     consistent with allowed transitions, indicating that the
1003:     corresponding final states have negative parity and $J \leq
1004:     \frac{5}{2}$. \vspace*{1ex}%
1005:   \label{tab:explevels}}
1006:   \begin{ruledtabular}
1007:     \begin{tabular}{lccccccccc}
1008:          & \multicolumn{9}{c}{States (MeV)} \\
1009:     Ref. & 0.0 & 0.32 & 1.78 & 2.69 & 3.41 & 3.89 & 3.96 & 5.24 & 5.86
1010:     \\
1011:     \hline
1012:     Ajz.-Sel.~\cite{ajz90:506rev} & $\frac{1}{2}^+$ & $\frac{1}{2}^-$ &
1013:     $\left( \frac{5}{2} , \frac{3}{2} \right)^+$ & $\left(
1014:     \frac{1}{2},\frac{3}{2},\frac{5}{2}^+\right)$ & $\left(
1015:     \frac{1}{2},\frac{3}{2},\frac{5}{2}^+\right)$ & $\geq \frac{7}{2}$ &
1016:     $\frac{3}{2}^-$ &     &     \\
1017:     Liu~\cite{liu90:42} & $\frac{1}{2}^+$ & $\frac{1}{2}^-$ & $\frac{5}{2}^+$ &
1018:     $\frac{3}{2}^-$ & $\frac{3}{2}^-$ & $\frac{3}{2}^+$ &
1019:     $\frac{3}{2}^-$ & $\frac{5}{2}^-$ &
1020:     $\left(\frac{1}{2}^+,\frac{1}{2}^- \right)$ \\
1021:     Morrisey~\cite{mor97:627} & $\frac{1}{2}^+$ & $\frac{1}{2}^-$ & $(+)$ & $(-)$ & $(+)$
1022:     & $(-)$ & $(-)$ & $(-)$ & $(-)$ \\
1023:     Aoi~\cite{aoi97:616} & $\frac{1}{2}^+$ & $\frac{1}{2}^-$ & $\frac{5}{2}^+$ &
1024:     $\frac{3}{2}^-$ & $\frac{3}{2}^-$ & $\frac{3}{2}^+$ &
1025:     $\frac{3}{2}^-$ & $\frac{5}{2}^-$ &     \\
1026:     Hirayama~\cite{hir04:738} & $\frac{1}{2}^+$ & $\frac{1}{2}^-$ & $(+)$ &
1027:     $\frac{3}{2}^-$ & $\frac{3}{2}^-$ & $\frac{5}{2}^-$ &
1028:     $\frac{3}{2}^-$ & $\frac{5}{2}^-$ &     \\
1029:     Fukuda~\cite{fuk04:70} & $\frac{1}{2}^+$ & $\frac{1}{2}^-$ & $\left(
1030:       \frac{3}{2} , \frac{5}{2} \right)^+$ &     & $\left(
1031:       \frac{3}{2} , \frac{5}{2} \right)^+$ &     &     &     &     \\
1032:    \end{tabular}
1033:   \end{ruledtabular}
1034: \end{table*}
1035: %
1036: %
1037: \begin{figure}[hbtp]
1038:   \includegraphics*[width=0.9\columnwidth]{fig5_be11v8p_hW12.eps}
1039:   \caption{\co\ Excitation spectrum for \beel\ calculated using the
1040:   \avp\ interaction in 0\ho--9\ho\ model spaces with
1041:   a fixed HO frequency of $\hbar\Omega = 12$~MeV. The experimental
1042:   values are from Ref.~\cite{ajz90:506rev}. The two lower graphs show
1043:   separately the negative- and positive-parity spectra, while the upper
1044:   graph shows the combined spectrum with selected states.%
1045:   \label{fig:11v8pspec}}
1046: \end{figure}
1047: %
1048: %
1049: \begin{figure}[hbtp]
1050:   \includegraphics*[width=0.9\columnwidth]{fig6_be11cdb_hW13.eps}
1051:   \caption{\co\ Excitation spectrum for \beel\ calculated using the
1052:   \cdb\ interaction in 0\ho--9\ho\ model spaces with
1053:   a fixed HO frequency of $\hbar\Omega = 13$~MeV. The experimental
1054:   values are from Ref.~\cite{ajz90:506rev}. The two lower graphs show
1055:   separately the negative- and positive-parity spectra, while the upper
1056:   graph shows the combined spectrum with selected states.%
1057:   \label{fig:11cdbspec}}
1058: \end{figure}
1059: %
1060: %
1061: \begin{figure}[hbtp]
1062:   \includegraphics*[width=0.9\columnwidth]{fig7_be11n3lo_hW12.eps}
1063:   \caption{\co\ Excitation spectrum for \beel\ calculated using the
1064:   \nlo\ interaction in 0\ho--9\ho\ model spaces with
1065:   a fixed HO frequency of $\hbar\Omega = 12$~MeV. The experimental
1066:   values are from Ref.~\cite{ajz90:506rev}. The two lower graphs show
1067:   separately the negative- and positive-parity spectra, while the upper
1068:   graph shows the combined spectrum with selected states.%
1069:   \label{fig:11n3lospec}}
1070: \end{figure}
1071: %
1072: %
1073: \begin{figure}[hbtp]
1074:   \includegraphics*[width=0.9\columnwidth]{fig8_be11inoy_hW17.eps}
1075:   \caption{\co\ Excitation spectrum for \beel\ calculated using the
1076:   \inoy\ interaction in 0\ho--9\ho\ model spaces with
1077:   a fixed HO frequency of $\hbar\Omega = 17$~MeV. The experimental
1078:   values are from Ref.~\cite{ajz90:506rev}. The two lower graphs show
1079:   separately the negative- and positive-parity spectra, while the upper
1080:   graph shows the combined spectrum with selected states.%
1081:   \label{fig:11inoyspec}}
1082: \end{figure}
1083: %
1084: %
1085: \begin{table}[hbtp]
1086:   \caption{Experimental and calculated energies (in [MeV]) of the lowest
1087:     negative- and positive-parity states in \beel, as well as the
1088:     magnetic moment (in [$\mu_N$]) of the ground state. Results for the
1089:     \avp, \cdb, \nlo\, and \inoy\ \nn\ interactions are presented. These
1090:     calculations were performed in the 8(9)$\ho$ model space for
1091:     negative-(positive-)parity states, using the HO frequencies listed in
1092:     Table~\ref{tab:optfreq}. $E_{x^-}$ denotes the excitation
1093:     energy relative to the lowest negative-parity state. Experimental
1094:     values are from~\cite{gei99:83,ajz90:506rev}. \vspace*{1ex}%
1095:   \label{tab:11energies}}
1096:   \begin{ruledtabular}
1097:     \begin{tabular}{cccccc}
1098:       \beel\  &        & \multicolumn{4}{c}{NCSM}                      \\   
1099:               & Exp    & \inoy\    &  \cdb\    & \nlo\     & \avp\     \\
1100:       \hline
1101:       $E \jpt{1}{1}{-}{3}$
1102:               & -65.16 & -62.40    & -56.95    & -56.57    & -55.52    \\ 
1103:       $E\jpt{1}{1}{-}{3} - E_\mathrm{gs}$
1104:               & 0.32   & -2.89     & -2.69     & -2.44     & -2.54     \\
1105:       $E_{x^-} \jpt{1}{1}{-}{3}$
1106:               & 0      & 0         & 0         & 0         & 0         \\
1107:       $E_{x^-} \jpt{3}{1}{-}{3}$
1108:               & ?\footnote[1]{There are large ambiguities in the
1109:               experimental spin-parity assignments,
1110:               cf. Table~\ref{tab:explevels}} 
1111:                        & 2.99      & 2.27      & 2.08      & 2.04      \\
1112:       $E_{x^-} \jpt{5}{1}{-}{3}$
1113: 	      & ?\footnotemark[1]
1114:                        & 3.82      & 3.93      & 3.70      & 3.81      \\
1115:       $E_{x^-} \jpt{3}{2}{-}{3}$ 
1116:               & ?\footnotemark[1]
1117:                        & 6.93      & 4.91      & 4.43      & 4.38      \\
1118:       \hline
1119:       $E_\mathrm{gs}\jpt{1}{1}{+}{3}$
1120:               & -65.48 & -59.51    & -54.26    & -54.13    & -52.98    \\
1121:       $E_x \jpt{1}{1}{+}{3}$
1122:               & 0      & 0         & 0         & 0         & 0       \\
1123:       $E_x \jpt{5}{1}{+}{3}$
1124:               & ?\footnotemark[1]
1125:                        & 1.85      & 2.01      & 1.98      & 2.03    \\
1126:       \hline
1127:       $\mu_\mathrm{gs}$
1128:               & -1.6816(8)
1129:                        & -1.47     & -1.55     & -1.58     & -1.58     \\
1130:     \end{tabular}
1131:   \end{ruledtabular}
1132: \end{table}
1133: %
1134: 
1135: The low-lying experimental spectrum is compared to our NCSM calculated
1136: levels (obtained using the four different \nn\ interactions) in
1137: Figs.~\ref{fig:11v8pspec}--\ref{fig:11inoyspec}
1138: %
1139: %\nb{make sure that this range is correct} 
1140: %
1141: and in Table~\ref{tab:11energies}. In the two lower panels of these
1142: figures we show, separately, the negative- and positive-parity spectra,
1143: while the upper panel shows the combined spectrum with selected
1144: states. Note that, those experimental levels for which there is an
1145: uncertainty in the parity assignment, are included in all three
1146: panels. There are clear signs of convergence with increasing model
1147: space. However, as was also observed for \beni, the relative position of
1148: the negative- and positive-parity spectra has clearly not converged, and
1149: the latter is still moving down. The most dramatic drop is observed in
1150: the \inoy\ spectrum, thus indicating the importance of a $3N$
1151: force. With this particular interaction, the $1/2^+$ level actually ends
1152: up below all, but one, of the negative-parity states in the (8--9)\ho\
1153: calculation. We refer to Sec.~\ref{sec:pospar} for further discussions
1154: on this topic.
1155: 
1156: We stress again that the relative level spacings, observed when plotting
1157: negative- and positive-parity states separately, is remarkably
1158: stable. Furthermore, the ordering of the first six(four) levels of
1159: negative(positive) parity, is the same for all four potentials. This
1160: calculated level ordering is summarized in
1161: Table~\ref{tab:theorordering}.
1162: %
1163: \begin{table}[hbtp]
1164:   \caption{NCSM observed ordering (from left to right) of \beel\
1165:   negative- and positive-parity states (separately). Note that all four
1166:   \nn\ interactions used in this study give the same ordering for the
1167:   first six(four) negative-(positive-)parity states.
1168:   %
1169:   %\nb{Maybe we should make sure that this table is adjacent to the review of
1170:   %experimental assignments?}
1171:   %
1172:   \vspace*{1ex}%
1173:   \label{tab:theorordering}}
1174:   \begin{ruledtabular}
1175:     \begin{tabular}{cccccc}
1176:       \multicolumn{6}{c}{Negative parity}\\
1177:       $\frac{1}{2}^-$ & $\frac{3}{2}^-$ & $\frac{5}{2}^-$ &
1178:       $\frac{3}{2}^-$ & $\frac{3}{2}^-$ & $\frac{5}{2}^-$ \\[1ex]
1179:     \hline
1180:       \multicolumn{6}{c}{Positive parity}\\
1181:       $\frac{1}{2}^+$ & $\frac{5}{2}^+$ & $\frac{3}{2}^+$ &
1182:       $\frac{5}{2}^+$ & \multicolumn{2}{c}{} \\ 
1183:    \end{tabular}
1184:   \end{ruledtabular}
1185: \end{table}
1186: %
1187: Our results can, therefore, provide input to help resolve the
1188: uncertainties of the experimental spin-parity assignments
1189: (cf.~Table~\ref{tab:explevels}). Note in particular that some
1190: experiments suggest that there are three low-lying $3/2^-$ states. The
1191: task to compute three levels with the same spin quickly becomes very
1192: time consuming with increasing dimension, since it requires many Lanczos
1193: iterations. Therefore, this third state was studied in two separate
1194: runs, using only the \avp\ and \inoy\ potentials, and is included in
1195: Figs.~\ref{fig:11v8pspec} and \ref{fig:11inoyspec} up to the 6\ho\ model
1196: space. These calculations confirm the existence of three low-lying
1197: $3/2^-$ levels, but they also stress the presence of a $5/2^-$ state
1198: which is not completely consistent with
1199: Refs.~\cite{liu90:42,aoi97:616,hir04:738}. However, we can not rule out
1200: the possibility of a low-lying intruder 2\ho-dominated state, which
1201: would avoid detection in our study. These states have a different
1202: convergence pattern than 0\ho\ states and generally appear at too high
1203: an excitation energy in the smaller model spaces, see
1204: e.g. Ref.~\cite{cau01:64}.
1205: 
1206: In summary, our results suggest that there are two excited
1207: positive-parity states below 4~MeV (rather than three as stated in
1208: Ref.~\cite{ajz90:506rev}). The 1.78~MeV level should be a $5/2^+$ state,
1209: while either the 3.41 or the 3.89~MeV level is a $3/2^+$. Our results do
1210: not support the presence of a high-spin ($J \geq 7/2$) state, which one
1211: can find in Ref.~\cite{ajz90:506rev}. We do observe three low-lying
1212: $3/2^-$ states although they are accompanied by a $5/2^-$ state.
1213: 
1214: The strength of the electric dipole transition between the two bound
1215: states in \beel\ is of fundamental importance. This is an observable
1216: which has attracted much attention since it was first measured in
1217: 1971~\cite{han71:3}, and again in 1983~\cite{mil83:28}. The cited value
1218: of 0.36 W.u. is still the strongest known transition between low-lying
1219: states, and it has been attributed to the halo character of the
1220: bound-state wave functions. Unfortunately, by working in a HO basis, we
1221: suffer from an incorrect description of the long-range asymptotics, and
1222: we would need an extremely large number of basis states in order to
1223: reproduce the correct form. This shortcoming of the HO basis is
1224: illustrated by the fact that we obtain a value for the E1 strength which
1225: is 20 times too small (see Table~\ref{tab:radius}). When studying the
1226: dependence of this value on the size of the model space, we observe an
1227: almost linear increase, indicating that our result is far from
1228: converged.  For the $\left\{(4-5)\:-\:(6-7)\:-\:(8-9)\right\}\ho$
1229: sequence of model spaces, the \beel\ E1 strength, $B\left( \mathrm{E}1;
1230: \frac{1}{2}_{_1}^- \to \frac{1}{2}_{_1}^+ \right)$, calculated with the
1231: \avp\ interaction increases as:
1232: $\left\{0.0054\:-\:0.0059\:-\:0.0065\right\}$~[$e^2$fm$^2$].  The
1233: corresponding sequence of results for \beni\ is: $B \left( \mathrm{E}1;
1234: \frac{1}{2}_{_1}^+ \to \frac{3}{2}_{_1}^- \right) =
1235: \left\{0.029\:-\:0.031\:-\:0.033\right\}$~[$e^2$fm$^2$], which
1236: demonstrates a similar increase. However, for this nucleus we note that,
1237: in the largest model space, our calculated E1 strength is only off by a
1238: factor of two compared to experiment. In addition, a consistent result
1239: is found for the much weaker $\frac{5}{2}_{_1}^+ \to \frac{3}{2}_{_1}^-
1240: $ E1 transition in \beni, where we also obtain a factor of two smaller
1241: $B(\mathrm{E}1)$ than experiment. These results accentuates the
1242: anomalous strength observed for \beel. A simple explanation for the
1243: failure of HO calculations in the \beel\ case was given by Millener
1244: \emph{et al}~\cite{mil83:28}. It was shown that there is a strong
1245: cancellation in the calculated E1 transition amplitude due to the
1246: insufficient description of the long-range asymptotics (see in
1247: particular Tables IV and V in Ref.~\cite{mil83:28}). By simply replacing
1248: their HO single-particle wave functions with solutions to the
1249: Schr\"odinger equation with a Woods-Saxon potential, they found that the
1250: magnitude of the neutron $1s_{1/2}0p_{1/2}$ single-particle matrix
1251: element increased significantly so that the cancellation was removed.
1252: Even though our multi-\ho\ calculations give a significant improvement
1253: of the calculated E1 strengths as compared to their simple
1254: (0--1)\ho\ model, the underlying problem is still present.
1255: 
1256: Another operator which is sensitive to the long-range behavior of the
1257: wave function is the point-nucleon radius. However, even though no
1258: operator renormalization has been applied, our results show a fair
1259: stability with increasing model space, and they are in rather good
1260: agreement with experimental findings for both \beni\ and \beel\ (see
1261: Table~\ref{tab:radius}). It is probably safe to assume that the missing
1262: part of the \beel\ matter radius originates mainly in an underestimation
1263: of the point-neutron radius. One should also remember that the
1264: experimental results for matter radii, in these light systems, are
1265: highly model-dependent and are usually theoretically extracted from
1266: measurements of the interaction cross section. In addition, we have also
1267: calculated the radii of the first excited state. For both isotopes it is
1268: found that the unnatural-parity state has a $~10$\% larger neutron
1269: radius than the natural-parity one, probably due to a larger admixture
1270: of $sd$-shell neutrons. Finally, the ground-state magnetic
1271: moment of \beel\ has been measured~\cite{gei99:83} and we find a
1272: reasonable agreement with our calculated value, see
1273: Table~\ref{tab:11energies}.
1274: %
1275: \begin{table}[hbtp]
1276:   \caption{Nuclear ground-state radii (in [fm]) and the E1 strengths (in
1277:     [$e^2$fm$^2$]) for the strong ground-state transitions in \beni\ and
1278:     \beel. The NCSM calculations were performed in the 8(9)$\ho$ model
1279:     space for negative-(positive-)parity states using the \avp\
1280:     interaction. The GFMC result for \beni, with the same
1281:     interaction~\cite{pie02:66}, is shown for comparison. Experimental
1282:     values are
1283:     from~\cite{til04:745,ajz90:506rev,tan88:206,fri95:60}. \vspace*{1ex}%
1284:   \label{tab:radius}}
1285:   \begin{ruledtabular}
1286:     \begin{tabular}{ccccccc}
1287:               & \multicolumn{5}{c}{\beni\jpt{3}{1}{-}{1} } \\
1288:               & & & & \multicolumn{2}{c}{$B ( \mathrm{E}1 )$} \\
1289:               & $R_n$      & $R_p$        & $R_\mathrm{mat}$    & 
1290:       $\frac{1}{2}_{_1}^+ \to \frac{3}{2}_{_1}^-$  &
1291:       $\frac{5}{2}_{_1}^+ \to \frac{3}{2}_{_1}^-$ \\
1292:       \hline							       
1293:       Exp     &            & 2.39         & 2.45(1)\footnote{Interaction
1294:         radius}                                                 
1295:       & 0.061(25) & 0.0100(84) \\
1296:       NCSM    & 2.40       & 2.27         & 2.34                
1297:       & 0.033      & 0.0057 \\
1298:       GFMC    & ---        & 2.41(1)      & ---                 & --- & --- \\
1299:       \hline
1300:       \hline
1301:               & \multicolumn{5}{c}{\beel\jpt{1}{1}{+}{3} } \\
1302:                & & & & \multicolumn{2}{c}{$B ( \mathrm{E}1 )$} \\
1303:              & $R_n$      & $R_p$        & $R_\mathrm{mat}$   & 
1304:       \multicolumn{2}{c}{$\frac{1}{2}_{_1}^- \to \frac{1}{2}_{_1}^+$} \\
1305:       \hline							       
1306:       Exp     &            &              & 2.86(4)            & 
1307:       \multicolumn{2}{c}{0.116(12)} \\
1308:       NCSM    & 2.66       & 2.30         & 2.54               & 
1309:       \multicolumn{2}{c}{0.0065} \\
1310:     \end{tabular}
1311:   \end{ruledtabular}
1312: \end{table}
1313: %
1314: 
1315: The standard halo picture of the \beel\ ground state is a simple
1316: two-body configuration consisting of an inert \nuc{10}{Be} core coupled
1317: to an $s_{1/2}$ valence neutron. Theoretical estimates of the
1318: spectroscopic factor for this component range from 0.55 to 0.92, see
1319: e.g. Table~1 in Ref.~\cite{win01:683}. The experimental situation is
1320: also unclear since the extracted results are generally
1321: model-dependent. In the literature one can find values from 0.36 to 0.8,
1322: see e.g. Fig.~8 in Ref.~\cite{pal03:68}. An important question is to
1323: which extent the first-excited \nuc{10}{Be}$\left( 2_1^+ \right)$ state
1324: contributes to the simple two-body configuration. The formalism for
1325: investigating cluster structures of NCSM eigenstates was recently
1326: developed in Ref.~\cite{nav04:70_2}. We have calculated the overlap of
1327: the \beel\jpt{1}{1}{+}{3} state with different $\nuc{10}{Be} + n$
1328: channels. To this aim, the \beel(\nuc{10}{Be}) wave functions were
1329: calculated using the \cdb\ interaction in a 7(6)\ho\ model space. We
1330: used a HO frequency of $\ho = 14$~MeV, which corresponds to the optimal
1331: value for calculating binding energies in these two model spaces. The
1332: largest overlap functions (in $jj$ coupling) are presented in
1333: Fig.~\ref{fig:11overlap}, while the corresponding spectroscopic factors
1334: (the overlap function squared and integrated over all $r$) are
1335: summarized in Table~\ref{tab:11specfac}. Several additional channels,
1336: such as the overlap with the second excited $2_2^+$ state in
1337: \nuc{10}{Be}, were also computed but their spectroscopic factors were
1338: found to be very small ($\lesssim 0.001$).
1339: %
1340: \begin{figure}[hbtp]
1341:   \includegraphics*[width=0.9\columnwidth]{fig9_be11be10cdb98_76_14_rxf.eps}
1342:   \caption{\co\ The largest radial overlap functions for the
1343:     \beel\jpt{1}{1}{+}{3} state decomposed as $\nuc{10}{Be} + n$ in
1344:     $jj$-coupling. The results presented here were obtained with the
1345:     \cdb\ interaction ($\ho = 14$~MeV) with the \beel(\nuc{10}{Be}) wave
1346:     function calculated in a $7(6)\hbar\Omega$ model space. The thin,
1347:     dotted line shows the dominant overlap function calculated in a
1348:     smaller $5(4)\hbar\Omega$ model space.
1349:   \label{fig:11overlap}}
1350: \end{figure}
1351: %
1352: %
1353: \begin{table}[hbtp]
1354:   \caption{Spectroscopic factors for the \beel\ \jpt{1}{1}{+}{3} ground
1355:     state decomposed as $\nuc{10}{Be} + n$ in $jj$
1356:     coupling. The results presented here were obtained with the \cdb\
1357:     interaction ($\ho = 14$~MeV) with the \beel(\nuc{10}{Be}) wave
1358:     function calculated in a $7(6)\hbar\Omega$ model
1359:     space. For comparison, we list spectroscopic factors extracted from
1360:     three recent experiments utilizing different reactions.\vspace*{1ex}%
1361:   \label{tab:11specfac}}
1362:   \begin{ruledtabular}
1363:     \begin{tabular}{cccccc}
1364: %      \multicolumn{6}{c}{$\beel\jpt{1}{1}{+}{3} \to \nuc{10}{Be}+n$}\\
1365: %      $\nuc{10}{Be} \left( J^\pi \right)$ & $n(l,j)$ & NCSM
1366: %        & Transfer~\cite{win01:683} & Knockout~\cite{aum00:84} &
1367: %        Breakup~\cite{pal03:68} \\
1368:       \multicolumn{2}{c}{$\nuc{10}{Be} \otimes n$} &
1369:         & Transfer & Knockout &
1370:         Breakup \\
1371: 	$J^\pi$ &  $(l,j)$ & NCSM
1372: 	& \cite{win01:683}\footnote{DWBA analysis of $\beel(p,d)$.}
1373:         & \cite{aum00:84}\footnote{From $\beni \left(
1374:         \beel,\nuc{10}{Be}+\gamma \right)$} 
1375: 	& \cite{pal03:68}\footnote{Spectroscopic factors extracted from
1376:         \beel\ breakup on lead and carbon targets respectively.} \\
1377:       \hline							
1378:       $0_1^+$ & $\left( 0, \frac{1}{2} \right) $     & 0.818 
1379:         & 0.67-0.80 & 0.78 & 0.61(5), 0.77(4) \\
1380:       $2_1^+$ & $\left( 2, \frac{5}{2} \right) $     & 0.263 
1381:         & 0.09-0.16 \\
1382:       $2_1^+$ & $\left( 2, \frac{3}{2} \right) $     & 0.022 \\
1383:       $1_1^+$ & $\left( 0, \frac{1}{2} \right) $     & 0.032 \\
1384:       $0_4^+$ & $\left( 0, \frac{1}{2} \right) $     & 0.005 \\
1385:       $0_8^+$ & $\left( 0, \frac{1}{2} \right) $     & 0.037 \\
1386:     \end{tabular}
1387:   \end{ruledtabular}
1388: \end{table}
1389: %
1390: Several observations can be made when studying these results: (1) The
1391: \beel\ ground state has a large overlap with $\left[ \nuc{10}{Be}(0_1^+)
1392: \otimes n\left(s_{1/2}\right) \right]$ ($S= 0.82$), but also with the
1393: core-excited $\left[ \nuc{10}{Be}(2_1^+) \otimes n\left(d_{5/2}\right)
1394: \right]$ channel ($S=0.26$). These results are in good agreement with
1395: the consensus of recent experimental studies, see
1396: e.g. Refs.~\cite{aum00:84,win01:683}.  (2) The thin dotted line in
1397: Fig.~\ref{fig:11overlap} shows the $\left[ \nuc{10}{Be}(0_1^+) \otimes
1398: n\left(s_{1/2}\right) \right]$ overlap function calculated in a smaller
1399: model space. From this comparison it is clear that the results are quite
1400: stable with regards to a change in \nm. The interior part does hardly
1401: change at all, while the tail is slowly extending towards larger
1402: inter-cluster distances. This statement is true for all channels shown
1403: in the figure except for those involving the two high-lying $0^+$ states
1404: [see bullet (4) below]. (3) The inset shows the main component plotted
1405: on a logarithmic scale. This graph clearly demonstrates the fact that
1406: our HO basis is not large enough to reproduce the correct asymptotic
1407: behavior. Even though the tail is extending further with increasing \nm,
1408: it still does not reach the expected exponential decay. Instead it dies
1409: of too fast. (4) Our calculated \nuc{10}{Be} $0_4^+$ and $0_8^+$ states
1410: are found to be 2\ho\ dominated, and their binding energies have not
1411: converged in the NCSM calculation. The cluster overlaps with these
1412: states do not display the same stability as observed for the other
1413: channels. Instead, there is a large dependence on \nm. A similar result
1414: was found in Ref.~\cite{nav04:70_2} and it is just another manifestation
1415: of the slower convergence of the 2\ho\ states in the NCSM.
1416: 
1417: Finally, we compare, in Tables~\ref{tab:11conf} and~\ref{tab:11spocc},
1418: the resulting configurations and the occupancies of single-particles
1419: states obtained with different interactions. Again, it is clear that the
1420: \inoy\ eigenstates have a larger fraction of low-\ho\ excitations and
1421: that this interaction results in a different single-particle spectrum
1422: due, most likely, to a stronger spin-orbit interaction.
1423: %
1424: \begin{table}[hbtp]
1425:   \caption{Calculated configurations of the first negative- and
1426:     positive-parity states in \beel. Results obtained in our largest
1427:     model spaces (8$\hbar\Omega$ and 9$\hbar\Omega$, respectively) are
1428:     presented. The calculations were performed with the HO frequencies
1429:     listed in Table~\ref{tab:optfreq}.\vspace*{1ex}%
1430:   \label{tab:11conf}}
1431:   \begin{ruledtabular}
1432:     \begin{tabular}{lccccc}
1433:       & \multicolumn{5}{c}{\beel\jpt{1}{1}{-}{3} (8$\hbar\Omega$ model space)}\\
1434:       \nn\ interaction & 0$\hbar\Omega$ & 2$\hbar\Omega$ & 4$\hbar\Omega$
1435:         & 6$\hbar\Omega$ & 8$\hbar\Omega$ \\
1436:       \hline
1437:       \inoy\ & 0.59 & 0.17 & 0.14 & 0.06 & 0.04 \\
1438:       \cdb\  & 0.51 & 0.20 & 0.15 & 0.08 & 0.06 \\
1439:       \nlo\  & 0.49 & 0.22 & 0.15 & 0.08 & 0.06 \\
1440:       \avp\  & 0.48 & 0.21 & 0.16 & 0.08 & 0.07 \\[2ex]
1441:       \hline
1442:       & \multicolumn{5}{c}{\beel\jpt{1}{1}{+}{3} (9$\hbar\Omega$ model space)}\\
1443:       \nn\ interaction & 1$\hbar\Omega$ & 3$\hbar\Omega$ & 5$\hbar\Omega$
1444:         & 7$\hbar\Omega$ & 9$\hbar\Omega$ \\
1445:       \hline
1446:       \inoy\ & 0.56 & 0.20 & 0.14 & 0.06 & 0.04 \\
1447:       \cdb\  & 0.50 & 0.21 & 0.15 & 0.08 & 0.06 \\
1448:       \nlo\  & 0.49 & 0.22 & 0.16 & 0.08 & 0.06 \\
1449:       \avp\  & 0.48 & 0.22 & 0.16 & 0.08 & 0.07 \\
1450:     \end{tabular}
1451:   \end{ruledtabular}
1452: \end{table}
1453: %
1454: %
1455: \begin{table}[hbtp]
1456:   \caption{Calculated occupations of neutron single-particle levels for
1457:     the first negative- and positive-parity states in \beel. Results
1458:     obtained in our largest model spaces (8$\hbar\Omega$ and
1459:     9$\hbar\Omega$, respectively) are presented. The calculations were
1460:     performed with the HO frequencies listed in
1461:     Table~\ref{tab:optfreq}.\vspace*{1ex}%
1462:   \label{tab:11spocc}}
1463:   \begin{ruledtabular}
1464:     \begin{tabular}{lcccccc}
1465:       & \multicolumn{5}{c}{\beel\jpt{1}{1}{-}{3} (8$\hbar\Omega$ model
1466:       space)}\\
1467:       \nn\ interaction & $0s_{1/2}$ & $0p_{1/2}$ & $0p_{3/2}$
1468:         & $1s_{1/2}$ & $0d_{3/2}$ & $0d_{5/2}$ \\
1469:       \hline
1470:       \inoy\ & 1.862& 1.078& 3.643& 0.046& 0.065& 0.075\\
1471:       \cdb\  & 1.835& 1.093& 3.597& 0.066& 0.062& 0.072\\
1472:       \nlo\  & 1.832& 1.095& 3.586& 0.073& 0.062& 0.072\\
1473:       \avp\  & 1.828& 1.094& 3.579& 0.073& 0.061& 0.072\\[2ex]
1474:       \hline
1475:       & \multicolumn{5}{c}{\beel\jpt{1}{1}{+}{3} (9$\hbar\Omega$ model
1476:       space)}\\
1477:       \nn\ interaction & $0s_{1/2}$ & $0p_{1/2}$ & $0p_{3/2}$
1478:         & $1s_{1/2}$ & $0d_{3/2}$ & $0d_{5/2}$ \\
1479:       \hline
1480:       \inoy\ & 1.845 & 0.504& 3.300& 0.658& 0.086& 0.345\\
1481:       \cdb\  & 1.824 & 0.600& 3.181& 0.742& 0.088& 0.285\\
1482:       \nlo\  & 1.823 & 0.616& 3.153& 0.752& 0.091& 0.281 \\
1483:       \avp\  & 1.820 & 0.630 & 3.135 & 0.768 & 0.088 & 0.265
1484:       \\[2ex] 
1485:     \end{tabular}
1486:   \end{ruledtabular}
1487: \end{table}
1488: %
1489: %**************************************************************************
1490: \subsection{\label{sec:pospar}Parity inversion}
1491: %
1492: One of the main objectives of this study has been to investigate the
1493: relative position of negative- and positive-parity states in the region
1494: around \beel. As we have shown, none of our calculations reproduce the
1495: parity inversion that is observed for this nucleus. However, considering
1496: the slower convergence rate for 1\ho-dominated states in the NCSM, and
1497: the large, but still finite, model spaces that we were able to use, our
1498: results are actually very promising. In all nuclei, we found a fast drop
1499: of the unnatural-parity states with respect to the natural ones. This
1500: behavior has already been demonstrated in earlier NCSM studies, but the
1501: drop that we observe in \beel\ is the most dramatic so far. Furthermore,
1502: the results obtained with the \inoy\ interaction are clearly different
1503: from the others, which indicates the significance that a realistic $3N$
1504: force should have in a fundamental explanation of the parity
1505: inversion. Note that \inoy\ is a two-body interaction, but that it
1506: simulates the main effects of $3N$ forces by short-range, non-local
1507: terms. Furthermore, the $^{3\!}P$ \nn\ interactions are slightly modified in
1508: order to improve the description of $3N$ analyzing
1509: powers. Fig.~\ref{fig:beeminextrapol} shows the calculated excitation
1510: energy of the first positive-parity states in \beni\ and \beel\ as a
1511: function of the basis size, \nm. For illustrative purposes we have
1512: extrapolated our results to larger model spaces assuming an exponential
1513: dependence on \nm, i.e.,~$E_x = E_{x,\,\infty} + a \exp\left( -b \nm
1514: \right)$. Note that the (0--1)\ho\ points are excluded from the fits.
1515: %
1516: \begin{figure}[hbtp]
1517:   \includegraphics*[width=0.9\columnwidth]{fig10a_a9energyminfixfreq_paper.eps}\\
1518:   \includegraphics*[width=0.9\columnwidth]{fig10b_be11energyminfixfreq_paper.eps}
1519:   \caption{\co\ Basis size dependence of the calculated $E_x\left(
1520:     \frac{1}{2}_{_1}^{+} \right)$ excitation energy relative to the
1521:     lowest negative-parity state in (a) \beni\ and (b) \beel. The results for
1522:     four different \nn\ interactions are compared. For each potential, a
1523:     single, fixed HO frequency was used (see
1524:     Table~\ref{tab:optfreq}). The dashed lines correspond to exponential
1525:     fits of the calculated data and, for illustration, these curves are
1526:     extrapolated to larger model spaces.%
1527:   \label{fig:beeminextrapol}}
1528: \end{figure}
1529: %
1530: The extrapolated \inoy\ results end up below the other interactions; and
1531: for \beni\ the curve is actually approaching the experimental
1532: value. With all other interactions, the extrapolated excitation energy
1533: is $\approx$1--2~MeV too high.
1534: 
1535: When discussing the position of the first unnatural-parity state, it is
1536: very interesting to study the systematics within the $A=11$ isobar and
1537: the $N=7$ isotone. To this aim, we have performed large-basis
1538: calculations for \nuc{11}{B} and \nuc{13}{C}. The diagonalization of the
1539: \nuc{11}{B} Hamiltonian in the 9\ho\ space proved to be our largest
1540: calculation so far. For \nuc{13}{C} we were only able to reach the 8\ho\
1541: space. Both studies were performed using the \cdb\ interaction and an HO
1542: frequency of $\ho = 13$~MeV. The ground-state binding energies (obtained
1543: in the 8\ho\ space) are: $E \left( \nuc{11}{B}; \frac{3}{2}_{_1}^{-} \:
1544: \frac{1}{2} \right) = 66.25$~MeV, and $E \left( \nuc{13}{C};
1545: \frac{1}{2}_{_1}^{-} \: \frac{1}{2} \right) = 86.53$~MeV. Our calculated
1546: \nuc{11}{B} spectrum, including the first negative-parity state for each
1547: spin up to $J=9/2$, plus the lowest positive-parity state, is compared
1548: to known experimental levels in Fig.~\ref{fig:b11cdbspec}. Note that we
1549: obtain an incorrect $1/2^-$ ground-state spin in our largest model
1550: space. However, the first $3/2^-$ and $1/2^-$ states are found to be
1551: almost degenerate, and there is a trend indicating that the position of
1552: these levels may eventually intersect as the basis size is increased. In
1553: principle, a thorough frequency variation study should be performed in
1554: order to clarify the fine details of the predictions. In any case, it is
1555: clear that the level splitting is described incorrectly with this
1556: interaction. Basically the same result was found in an earlier NCSM
1557: study~\cite{nav03:68} using a three-body effective interaction derived
1558: from \avp. In that paper, it was also shown that the correct level
1559: ordering can be reproduced, and the splitting greatly improved, by
1560: adding a realistic $3N$ force. For \nuc{13}{C} we have only computed the
1561: lowest state for each parity. However, this nucleus has also been
1562: studied previously using the NCSM. A spectrum obtained with the \cdb\
1563: interaction was presented in Ref.~\cite{thi02:697}, while calculations
1564: with a genuine $3N$ force were reported in Ref.~\cite{nav03:68}. In both
1565: papers, the study was limited to negative-parity states and a smaller
1566: model space (4\ho) was used.
1567: %
1568: \begin{figure}[hbtp]
1569:   \includegraphics*[width=0.9\columnwidth]{fig11_b11cdb_hW13.eps}
1570:   \caption{\co\ Excitation spectrum for \nuc{11}{B} calculated using the
1571:   \cdb\ interaction in 0\ho--9\ho\ model spaces with a fixed HO
1572:   frequency of $\hbar\Omega = 13$~MeV. The experimental levels are from
1573:   Ref.~\cite{ajz90:506rev}. Note that there are many additional levels
1574:   between the experimental $1/2^+$ and $9/2^-$ shown in the
1575:   figure. However, we have only computed the first level for a given
1576:   spin, and the $1/2^+$ was the only positive-parity state that was
1577:   considered.%
1578:   \label{fig:b11cdbspec}}
1579: \end{figure}
1580: %  
1581: 
1582: Let us now comment on our \nuc{11}{B} and \nuc{13}{C} results and return
1583: to the important question of the position of the first positive-parity
1584: state. The calculated $1/2^+$ excitation energy for these two nuclei, as
1585: a function of \nm, is shown in Fig.~\ref{fig:bceminextrapol}. It is a
1586: fascinating empirical fact that, by simply going from $Z=4 \to Z=6$, the
1587: first $1/2^+$ state moves from being the ground state in \beel\ to
1588: become an excited state at 3.1~MeV in \nuc{13}{C}. In the odd-$Z$
1589: nucleus \nuc{11}{B}, the first positive-parity state is found quite high
1590: in the excitation spectrum, namely at 6.8~MeV. It is a significant
1591: success of the NCSM method, and of the \nn\ interactions being employed,
1592: that these huge shifts are accurately reproduced in our
1593: calculations. However, as can be seen from
1594: Figs.~\ref{fig:beeminextrapol} and~\ref{fig:bceminextrapol}, the
1595: calculated excitation energy always turns out to be too large. A
1596: comparison of our extrapolated \cdb\ results shows that they exceed the
1597: experimental values by $\approx$1--2~MeV for all four isotopes. As a
1598: final remark, our \inoy\ results for \beni\ and \beel\ indicate that the
1599: use of a realistic $3N$ force in a large basis space might correct this
1600: discrepancy.
1601: %
1602: \begin{figure}[hbtp]
1603:   \includegraphics*[width=0.9\columnwidth]
1604: 		   {fig12a_b11energyminfixfreq_paper.eps}\\
1605:   \includegraphics*[width=0.9\columnwidth]
1606: 		   {fig12b_c13energyminfixfreq_paper.eps}
1607:   \caption{\co\ Basis size dependence of the calculated $E_x\left(
1608:     \frac{1}{2}_{_1}^{+} \right)$ excitation energy relative to the
1609:     lowest negative-parity state in (a) \nuc{11}{B} and (b)
1610:     \nuc{13}{C}. These results are obtained with the \cdb\ interaction
1611:     and $\ho = 13$~MeV. The dashed lines correspond to exponential fits
1612:     of the calculated data and, for illustration, these curves are
1613:     extrapolated to larger model spaces.%
1614:   \label{fig:bceminextrapol}}
1615: \end{figure}
1616: %
1617: %**************************************************************************
1618: \section{\label{sec:conc}Conclusions}
1619: %
1620: We have performed large-basis \emph{ab initio} no-core shell model
1621: calculations for \beni\ and \beel\ using four different realistic \nn\
1622: interactions. One of these, the non-local \inoy\ interaction, has never
1623: before been used in nuclear structure calculations. Although it is
1624: formally a two-body potential, it reproduces not only \nn\ data (beware
1625: of the fact that the $^{3\!}P$ interactions are slightly modified in the
1626: IS-M version that we are using) but also the binding energies of
1627: \nuc{3}{H} and \nuc{3}{He}. Therefore it has been of particular interest
1628: for our current application, where we have striven to maximize the model
1629: space by limiting ourselves to \nn\ interactions, but have still been
1630: very much interested in the effects of three-body forces. We have
1631: computed: binding energies, excited states of both parities,
1632: electromagnetic moments and transition strengths, point-nucleon radii,
1633: and also the core-plus-neutron cluster decomposition of the \beel\
1634: ground state.
1635: 
1636: In summary, for the calculated spectra we found clear signs of
1637: convergence, and a remarkable agreement between the predictions of
1638: different \nn\ interactions. In particular, the relative level spacings
1639: observed when plotting positive- and negative-parity states separately,
1640: were found to be very stable and to agree well with experimental
1641: spectra. This has allowed us to make some conclusions regarding the
1642: largely unknown spin-parities of unbound, excited states in \beel. An
1643: overall observation is that the \avp\ and \nlo\ potentials produce very
1644: similar results, while \cdb\ gives slightly more binding. The \inoy\
1645: interaction is clearly different; giving a much larger binding energy
1646: and a stronger spin-orbit splitting. Both these effects would be
1647: expected from a true $3N$ force, but are here achieved by the use of
1648: short-range, non-local terms in the \nn\ interaction.
1649: 
1650: Furthermore, it was also clear from our study that our results for
1651: observables connected to long-range operators, have not converged. These
1652: calculations would clearly benefit from operator renormalization, in
1653: order to correct for the limited model space being used. In particular,
1654: the extremely strong E1 transition between the two bound states in
1655: \beel, was underestimated by a factor of 20. We have discussed how this
1656: illustrates the fact that the anomalous strength is due to the halo
1657: character, and hence large overlap, of the initial and final state wave
1658: functions; a property which is extremely hard to reproduce using a HO
1659: basis. In the NCSM approach, there is no fitting to single-particle
1660: properties, e.g., by the use of empirical interactions. Instead, the
1661: effective interactions are derived from the underlying
1662: inter-nucleon forces. Therefore, it is likely that a good
1663: description of loosely bound, and unbound, single-particle states
1664: might require a very large number of HO basis functions.
1665: 
1666: An important topic of this work has been the investigation of the parity
1667: inversion found in \beel. We did not reproduce the anomalous $1/2^+$
1668: ground state in our \emph{ab initio} approach, but did observe a
1669: dramatic drop of the positive-parity excitation energies with increasing
1670: model space. Furthermore, the behavior of our \inoy\ results suggests
1671: that a realistic $3N$ force will have an important influence on the
1672: parity inversion. However, in order to pursue this question further, an
1673: improved computational capacity is needed. We have also performed
1674: large-basis calculations for \nuc{11}{B} and \nuc{13}{C}. In this way,
1675: we were able to put our \beel\ positive-parity results into a wider
1676: context by studying the systematics within the $Z=4$ isotopes (\beni),
1677: the $N=7$ isotone (\nuc{13}{C}), and the $A=11$ isobar
1678: (\nuc{11}{B}). Although we found that the NCSM always overestimates the
1679: excitation energy of the first unnatural-parity state, we did reproduce
1680: the very large shifts observed for these different nuclei. This is an
1681: important finding which leads us to the optimistic conclusion that the
1682: parity-inversion problem should be possible to reproduce in the NCSM
1683: starting from realistic inter-nucleon interactions.
1684: %
1685: %**************************************************************************
1686: \begin{acknowledgments}
1687: This work was partly performed under the auspices of the
1688: U. S. Department of Energy by the University of California, Lawrence
1689: Livermore National Laboratory under contract No. W-7405-Eng-48. Support
1690: from the LDRD contract No.~04--ERD--058, and from U.S.~Department of
1691: Energy, Office of Science, (Work Proposal Number SCW0498) is
1692: acknowledged.
1693: \end{acknowledgments}
1694: %**************************************************************************
1695: \bibliography{ncsm,f2b_refs,cf}
1696: %**************************************************************************
1697: %\newpage
1698: %\printfigures
1699: %**************************************************************************
1700: \end{document}
1701: