This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Possible halo structure of 62,72Ca by forbidden-state-free locally peaked Gaussians

W. Horiuchi [email protected] Department of Physics, Hokkaido University, Sapporo 060-0810, Japan    Y. Suzuki Department of Physics, Niigata University, Niigata 950-2181, Japan RIKEN Nishina Center, Wako 351-0198, Japan    M. A. Shalchi Instituto de Física Teórica, Universidade Estadual Paulista, 01140-070, São Paulo, SP, Brasil    Lauro Tomio Instituto de Física Teórica, Universidade Estadual Paulista, 01140-070, São Paulo, SP, Brasil
Abstract

In order to efficiently describe nucleon orbits around a heavy core nucleus, we propose locally peaked Gaussians orthogonalized to the occupied bound states in the core. We show the advantage of those functions in both numerical stability and fast convergence by taking examples of touchstone calcium isotopes 62,72Ca in Ca60,70+n+n{}^{60,70}{\rm Ca}+n+n three-body models. Both weakly bound configurations and continuum coupling effect are taken into account. We evaluate the neutron radii and the occupation probabilities of two-neutron configurations not only for the ground state but also for some particle-bound excited states by varying the strength of the core-neutron interaction. The emergence of the halo structure in the ground state depends on the energy difference between 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits. Two-neutron [consisting of (s1/2)2(s_{1/2})^{2} configuration] and one-neutron [consisting of (g9/2s1/2)(g_{9/2}s_{1/2}) configuration] halo structures of 62Ca can coexist in narrow energy spacing provided that both of 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits are almost degenerate and barely bound. The ground-state structure of 72Ca is likely to be a two-neutron halo, although its emergence depends on the position of the 2s1/22s_{1/2} level.

I Introduction

The landscape of exotic short-lived nuclei far from the stability line extends to the neutron dripline up to Ne isotopes [1]. In such extreme neutron excess, weakly bound neutrons emerge and often lead to neutron-halo structure. Exploring the halo nuclei has continued since the first discovery of the two-neutron halo nucleus 11Li [2]. Other typical examples of two-neutron halo nuclei include 6He [2], 14Be [3], 19[4], and 22[5, 6, 7]. See also Ref. [8] for more experimental and theoretical references on halo-nuclei studies. Recently, by measuring its interaction cross section, 29F has been identified as the heaviest two-neutron halo nucleus [9].

The two-neutron halo structure has often been described using a core plus two-neutron model. When the core is a light nucleus, the two-neutron motion can be accurately described, e.g., as in 22[5]. Within the search for evidence of Efimov states [10] in two-neutron (2n)(2n) halo nuclei close to the neutron-core (nn-core) unitary limit [11, 12], the emergence of universal properties described by two-body observables and a three-body scaling parameter became quite clear in the case of two neutrons bound to light-core nuclei.

Motivated by the recent measurement reported in Ref. [9], the dripline and near-dripline F isotopes have been studied by three-body models: For 29F, the pp-wave two-neutron halo character of F27+n+n{}^{27}{\rm F}+n+n was verified in Refs. [13, 14, 15], and a competition of halo and antihalo configurations was proposed in Ref. [16] for 31F studied as F29+n+n{}^{29}{\rm F}+n+n. By using the Gamow shell model these F isotopes were also studied in Ref. [17]. Studying the existence of heavier neutron-halo nuclei is interesting as it will help to expose the binding mechanism when increasing the mass number [18, 19], which opens the possibility of verifying universal aspects expected to emerge in low-energy quantum systems, such as ultra-cold atom-molecule heteronuclear systems [12, 21, 20]. However, a microscopic description becomes tougher as the core nucleus becomes heavier because the number of the occupied orbits in the core increases. Therefore, a reliable and efficient method is needed to describe the two-neutron motion around the heavy core nucleus. The complication of such calculation is mainly due to the condition that requires the configuration space available to the valence neutrons to be orthogonal to all the bound states occupied in the core (orthogonality condition model [22, 23]). To eliminate such bound states (called forbidden states) from a three-body solution, a pseudo potential method [24] has often been applied. With increasing the mass number of the core nucleus, the number of the occupied orbits increases and the description of the valence neutron orbits becomes more complicated and often numerically unstable.

To describe the nucleon motion around the core, one can use explicitly correlated Gaussians [25, 26, 27]. Due to the complexity of removing the forbidden states, however, its application is limited only to light nuclei. See, e.g., Refs. [28, 29]. Obviously, a single-particle (sp) basis is advantageous to eliminate the forbidden state. The application of Gaussian sp basis functions to the alpha decay of 212Po in the Pb208+n+n+p+p{}^{208}{\rm Pb}+n+n+p+p model was in fact successfully made in Ref. [30]. It turned out, however, that the description was not perfect for the localized nucleon orbits, i.e., the alpha clustering near the nuclear surface. This is probably because the ordinary Gaussians, rlexp(ar2)r^{l}\exp(-ar^{2}), used there are not always enough to obtain large amplitude near the nuclear surface. In Refs. [31, 32], locally peaked Gaussians (LPGs), r2k+lexp(ar2)r^{2k+l}\exp(-ar^{2}), are proposed to describe the localized configurations. By the additional r2kr^{2k} factor, the LPG basis allows one to describe not only damped short-ranged behavior [25, 26] but also large amplitude far from the center, while keeping the advantage that the matrix elements of the LPG bases can easily be evaluated.

In this paper, we introduce a forbidden-state-free LPG (FFLPG) to efficiently describe neutron orbits around a heavy core nucleus. This opens up the perspectives to study more complicated multinucleon systems around the heavy core. To show its effectiveness, we apply the method to describe 62,72Ca in which the emergence of the two-neutron halo was discussed [18, 19]. The purpose of this paper is twofold: (1) to establish an efficient way to describe both the short-ranged nodal behavior and enhanced amplitude beyond the nuclear surface, which is expected to occur in nucleon orbits around the heavy core, and (2) to clarify the conditions under which halo structure can emerge not only in the ground state but also in some excited states of the Ca60,70+n+n{}^{60,70}{\rm Ca}+n+n three-body systems.

As mentioned above, we focus on the neutron-rich Ca isotopes having large number NN of neutrons. Although the halo structure is expected to appear for N>40N>40, its structure or even its existence is under debate. Very little experimental information is available for the Ca isotopes: The heaviest, 60Ca, was confirmed in Ref. [33] but no information other than its existence is available. Mass measurements have been done up to 57Ca [34], and the charge radii [35] have been determined up to 52Ca, and recently interaction cross sections have been measured up to 51Ca [36]. It was conjectured by the coupled-cluster calculation that the neutron dripline of Ca isotopes is around 60Ca [37] and 62Ca has two-neutron halo structure with dominant ss-wave [18]. On the other hand energy-density-functional [38, 39], shell-model [40], and in-medium similarity renormalization group [41] calculations predicted that the dripline is around 70Ca. The halo structure of 72Ca was also predicted by the relativistic mean-field model [42]. The two-neutron halo structure in the ground state of 72Ca and its relationship to the Efimov physics was studied by a Ca70+n+n{}^{70}{\rm Ca}+n+n model [19]. The dripline of Ca isotopes was predicted to be 72Ca based on a Bayesian analysis of the density functional theory results [43]. Since there is still some ambiguity in determining the dripline, we leave its question open and study possible neutron-halo structure in the spectrum of both cases, 62,72Ca.

The paper is organized as follows. Section II presents the Hamiltonian and some definitions needed to introduce the present approach and explains how to construct the FFLPG. Section III discusses our results. First in Sec. III.1, the Ca+n{\rm Ca}+n potential employed in this paper is investigated. Section III.2 is devoted to test the power of this approach to the Ca60+n+n{}^{60}{\rm Ca}+n+n system. Comparison with a standard projection method is presented. Section III.3 discusses the emergence of various halo structures in the spectrum of 62Ca based on the three-body results. The case of 72Ca is presented in Sec. III.4. Conclusions and future perspectives are given in Sec. IV

II Formulation

II.1 Hamiltonian and variational calculation

The Hamiltonian of a core+2n+2n system consists of the nn-core kinetic (TT) and potential (UU) and the nn-nn potential (vv) terms,

H=i=12(Ti+Ui)+1AmN𝒑1𝒑2+v12,\displaystyle H=\sum_{i=1}^{2}(T_{i}+U_{i})+\frac{1}{Am_{N}}\bm{p}_{1}\cdot\bm{p}_{2}+v_{12}, (1)

where mNm_{N} and AA are the nucleon mass and the mass number of the core nucleus, respectively. We follow the cluster-orbital shell model approach [44], by reducing the three-body problem to a two-body problem using two independent nn-core relative distance vectors, 𝒓1\bm{r}_{1} and 𝒓2\bm{r}_{2} with their respective momentum conjugates 𝒑1\bm{p}_{1} and 𝒑2\bm{p}_{2}. The center-of-mass kinetic energy is subtracted and the corresponding kinetic energies are given by Ti=𝒑i2/(2μ)T_{i}={\bm{p}_{i}^{2}}/{(2\mu)} with μ=mNA/(A+1)\mu=m_{N}A/(A+1). Our choices for the nn-core and nn-nn potentials will be given later.

The total wave function with the angular momentum JJ and its zz component MM is expanded in terms of KK basis functions ΦJM(αi)\Phi_{JM}(\alpha_{i}) (i=1,,K)(i=1,\dots,K):

ΨJM=i=1KciΦJM(αi),\displaystyle\Psi_{JM}=\sum_{i=1}^{K}c_{i}\Phi_{JM}(\alpha_{i}), (2)

where αi\alpha_{i} denotes a set of variational parameters for the iith basis. These basis functions are not restricted to be orthogonal. A set of linear coefficients 𝒄=(c1,,cK)t\bm{c}=(c_{1},\dots,c_{K})^{t} is determined variationally by solving the generalized eigenvalue problem

𝒄=E𝒄,\displaystyle\mathcal{H}\bm{c}=E\mathcal{B}\bm{c}, (3)

where \mathcal{H} and \mathcal{B} are the Hamiltonian and overlap matrices with elements defined by

ij=Φ(αi)|H|Φ(αj),\displaystyle\mathcal{H}_{ij}=\left<\Phi(\alpha_{i})|H|\Phi(\alpha_{j})\right>, (4)

and

ij=Φ(αi)|Φ(αj).\displaystyle\mathcal{B}_{ij}=\left<\Phi(\alpha_{i})|\Phi(\alpha_{j})\right>. (5)

An optimal set of αi\alpha_{i} is selected by the stochastic variational method (SVM) [25, 26]. Its efficiency was demonstrated by a number of examples. See, e.g., Refs. [28, 27, 45]. We increase the basis one by one by selecting the one that gives the lowest energy among randomly generated candidates until the energy convergence is met. This procedure greatly reduces the total number of basis KK, which helps the description of nucleon systems around a heavy core, where the number of possible configurations is quite large.

II.2 Definition of FFLPG

The efficiency of a variational calculation strongly depends on a choice of basis functions. Some basic requirements for a “good” basis include the following in the present case: (i) Weakly bound neutron orbits should be well described, because the system may have large amplitude at and beyond the surface of the core nucleus. (ii) The evaluation of the matrix elements in Eqs. (4) and (5) should be easy and extendable to systems with more neutrons bound to the core. (iii) The removal of many forbidden states should be performed numerically easily and stably.

The sp wave function with total (orbital) angular momentum jj (ll), with their corresponding angular and spin wave functions, Yl(𝒓^)Y_{l}(\hat{\bm{r}}) and χ1/2\chi_{1/2}, is defined by

ϕkljma=ϕkla(r)[Yl(𝒓^)χ1/2]jm,\displaystyle\phi_{kljm}^{a}=\phi_{kl}^{a}(r)\left[Y_{l}(\hat{\bm{r}})\chi_{1/2}\right]_{jm}, (6)

where the LPG function is given by

ϕkla(r)=Nkl(a3π)14(ar)2k+lexp(12ar2),\displaystyle\phi_{kl}^{a}(r)=N_{kl}\left(\frac{a^{3}}{\pi}\right)^{\frac{1}{4}}(\sqrt{a}\,r)^{2k+l}\exp\left(-\frac{1}{2}ar^{2}\right), (7)

Here, NklN_{kl} is the normalization constant

Nkl=22k+l+2(4k+2l+1)!!,\displaystyle N_{kl}=\sqrt{\frac{2^{2k+l+2}}{(4k+2l+1)!!}}, (8)

and aa is a parameter related to the width of the LPG function. We assume kk to be a non-negative integer for the sake of simplicity. The LPG with k=0k=0 is nothing but the ordinary Gaussians. Most of basic matrix elements between the LPG bases can be obtained analytically [31]. Since the LPG reaches a maximum at r=(2k+l)/ar=\sqrt{(2k+l)/a}, a suitable combination of aa and kk can describe such wave packets that are centered at different positions and have different widths. Because of this flexibility, the LPG can describe even linear chain structure [31]. Obviously the tail of a weakly bound orbit can be described well by a superposition of the LPG. This flexibility is vital to describe the sp orbits around a heavy core nucleus.

The LPG basis satisfies the requirements (i) and (ii). Concerning the requirement (iii), we introduce a projection operator onto the forbidden (FF) space,

PF=nljFm=jj|ψnljmψnljm|,\displaystyle P_{F}=\sum_{n^{\prime}l^{\prime}j^{\prime}\in{F}}\sum_{m^{\prime}=-j^{\prime}}^{j^{\prime}}\left|\psi_{n^{\prime}l^{\prime}j^{\prime}m^{\prime}}\right>\left<\psi_{n^{\prime}l^{\prime}j^{\prime}m^{\prime}}\right|, (9)

where the sum extends over all the orbits occupied in the core nucleus. The forbidden states ψnljm\psi_{nljm} are defined by the bound-state solutions of the one-body Hamiltonian, T+UT+U. As will be seen later, they are very well approximated by the harmonic-oscillator (HO) functions

ψnljmν=ψnlν(r)[Yl(𝒓^)χ1/2]jm\displaystyle\psi_{nljm}^{\nu}=\psi_{nl}^{\nu}(r)\left[Y_{l}(\hat{\bm{r}})\chi_{1/2}\right]_{jm} (10)

with the principal quantum number (nn) and an appropriately chosen oscillator parameter ν\nu. We use this approximation in what follows. The FFLPG sp orbit is defined by

ϕ¯kljma\displaystyle\bar{\phi}_{kljm}^{a} =(1PF)ϕkljma\displaystyle=(1-P_{F})\phi_{kljm}^{a}
=ϕkljman;nljFψnlν|ϕklaψnljmν\displaystyle=\phi_{kljm}^{a}-\sum_{n^{\prime};n^{\prime}lj\in F}\left<\psi_{n^{\prime}l}^{\nu}\right.\left|\phi_{kl}^{a}\right>\psi_{n^{\prime}ljm}^{\nu}
ϕ¯kla(r)[Yl(𝒓^)χ1/2]jm.\displaystyle\equiv\bar{\phi}_{kl}^{a}(r)\left[Y_{l}(\hat{\bm{r}})\chi_{1/2}\right]_{jm}. (11)

Because ψnlν(r)\psi_{n^{\prime}l}^{\nu}(r) is a combination of LPG’s, ϕklν(r)(k=0,,n)\phi_{k^{\prime}l}^{\nu}(r)\ (k^{\prime}=0,\ldots,n^{\prime}), ψnlν|ϕkla\left<\psi_{n^{\prime}l}^{\nu}\right.\left|\phi_{kl}^{a}\right> is readily obtained by using

ϕkla|ϕkla\displaystyle\left<\left.\phi_{k^{\prime}l}^{a^{\prime}}\right|\phi_{kl}^{a}\right> =(2k+2k+2l+1)!!(4k+2l+1)!!(4k+2l+1)!!\displaystyle=\frac{(2k+2k^{\prime}+2l+1)!!}{\sqrt{(4k+2l+1)!!(4k^{\prime}+2l+1)!!}}
×a2k+l+32a2k+l+32(a+a2)k+k+l+32,\displaystyle\times\frac{\sqrt{a}^{2k+l+\frac{3}{2}}\sqrt{a^{\prime}}^{2k^{\prime}+l+\frac{3}{2}}}{(\frac{a+a^{\prime}}{2})^{k+k^{\prime}+l+\frac{3}{2}}}, (12)

which leads to easy determination of ϕ¯kljma\bar{\phi}_{kljm}^{a}.

With the use of the sp basis defined above, we construct an antisymmetrized two-neutron basis

ΦJM(a1k1l1j1;a2k2l2j2)\displaystyle\Phi_{JM}(a_{1}k_{1}l_{1}j_{1};a_{2}k_{2}l_{2}j_{2})
=12(1P12){[ϕ¯k1l1j1a1(1)ϕ¯k2l2j2a2(2)]JM},\displaystyle=\frac{1}{\sqrt{2}}(1-P_{12})\left\{\left[\bar{\phi}_{k_{1}l_{1}j_{1}}^{a_{1}}(1)\bar{\phi}_{k_{2}l_{2}j_{2}}^{a_{2}}(2)\right]_{JM}\right\}, (13)

where P12P_{12} exchanges the neutron labels 1 and 2, and [j1j2]JM[j_{1}j_{2}]_{JM} denotes the tensor product. Here, the set of variational parameters of the iith basis αi\alpha_{i} in Eq. (2) stands for αi=(a1,ik1,il1,ij1,i,a2,ik2,il2,ij2,i)\alpha_{i}=(a_{1,i}k_{1,i}l_{1,i}j_{1,i},a_{2,i}k_{2,i}l_{2,i}j_{2,i}). This two-neutron basis can easily be extended to core++few nucleon systems by successively coupling another nucleon one by one.

Note that the variational parameter αi\alpha_{i} comprises eight variables: two continuous ones (a1,ia_{1,i} and a2,ia_{2,i}) and six discrete ones. As will be shown later, both of short- and long-ranged LPG’s have to be superposed to properly describe the asymptotics of the FF halo wave functions and also continuum states with high angular momenta (l1,il_{1,i} and l2,il_{2,i}) have to be included to reach convergence. Therefore, discretizing each component of αi\alpha_{i} on certain grids would lead to enormous basis dimension even for the present three-body system. More crucial is that one has to take into account the possibility of producing two bound states with the same spin and parity. A way to overcome these difficult problems is to reduce the basis dimension by the SVM. An interested reader should refer to Chapter 4 of Ref. [26].

The above basis of Eq. (13) is completely FF. In contrast to this approach, a popular way of eliminating the forbidden components is to add a pseudo-potential to the Hamiltonian as in Ref. [24],

HH+λi=12PF(i),\displaystyle H\to H+\lambda\sum_{i=1}^{2}P_{F}(i), (14)

and to attempt at reaching a stable eigenvalue by taking λ\lambda\to\infty. In practice, λ104\lambda\approx 10^{4} MeV is taken [28]. This pseudo-potential method has the advantage of its simplicity. We compare both approaches in the next section.

III Application to Ca+n+n{\rm Ca}+n+n systems

III.1 Choice of nn-60Ca potential

We take the Woods-Saxon (WS) form for the nn-60Ca potential

U(r)=V0fWS(r)+V1r021rdfWS(r)dr(𝒔),\displaystyle U(r)=-V_{0}f_{\rm WS}(r)+V_{1}r_{0}^{2}\frac{1}{r}\frac{df_{\rm WS}(r)}{dr}(\bm{\ell}\cdot\bm{s}), (15)

where fWS(r)=[1+exp(rRWS)/aWS]1f_{\rm WS}(r)=\left[1+\exp\left(r-R_{\rm WS}\right)/a_{\rm WS}\right]^{-1} with RWS=r0A1/3R_{\rm WS}=r_{0}A^{1/3} and r0=1.27r_{0}=1.27 fm. We use two sets for the diffuseness parameter: One is a standard one, aWS=0.67a_{\rm WS}=0.67 fm (set A), and the other is a larger one, aWS=0.80a_{\rm WS}=0.80 fm (set B).

The 60Ca core nucleus is assumed to have the N=40N=40 closed configuration, that is, the occupied orbits include 0s1/20s_{1/2}, 0p3/20p_{3/2}, 0p1/20p_{1/2}, 1s1/21s_{1/2}, 0d5/20d_{5/2}, 0d3/20d_{3/2}, 1p3/21p_{3/2}, 1p1/21p_{1/2}, 0f7/20f_{7/2}, 0f5/20f_{5/2}. We generate these forbidden states by assuming the WS potential parameters of Eq. (2-182) in Ref. [46]. Those potential parameters are V0=40.00V_{0}=40.00 MeV and V1=17.60V_{1}=17.60 MeV. Note that the potential makes both of 0g9/20g_{9/2} and 2s1/22s_{1/2} orbits unbound. The HO parameter ν\nu is determined so as to maximize the average of the squared overlaps, nljmFψnljm|ψnljmν2/40\sum_{nljm\in{F}}\langle\psi_{nljm}|\psi^{\nu}_{nljm}\rangle^{2}/40 (see Ref. [47]). The maximum average value is 0.988 for set A and the resulting ν\nu value is 0.212 fm-2, while in set B case they are 0.989 and 0.203 fm-2. Approximating the forbidden states with the HO wave functions is quite reasonable.

There is no information about 61Ca. Even its stability is unknown. The valence neutron orbit of 61Ca belongs to the 2n+l=42n+l=4 shell comprising the 0g0g, 1d1d, and 2s2s orbits. The order of these sp orbits is under debate [42, 38, 39, 19, 43, 40, 41]. The standard shell-model filling with the spin-orbit interaction arranges the sp levels in the order of 0g9/20g_{9/2}, 2s1/22s_{1/2}, and 1d5/21d_{5/2} at A60A\approx 60 [46], while Ref. [37] predicted the inverted order of 2s1/22s_{1/2}, 1d5/21d_{5/2}, and 0g9/20g_{9/2}. In the latter case, the halo structure would appear in the ground state of 62Ca [18]. Following this inverted situation, we determine V0V_{0} in Eq. (15) so as to set the 2s1/22s_{1/2} sp energy to be 0.01-0.01 MeV, resulting in V0=44.03V_{0}=44.03 MeV (set A) and 41.89 MeV (set B), respectively. The value turns out to be slightly stronger than the standard value V040V_{0}\approx 40 MeV [46].

Refer to caption
Figure 1: Single-particle energies of 0g9/20g_{9/2} and 1d5/21d_{5/2} orbits of 61Ca as a function of the spin-orbit strength V1V_{1}. The diffuseness parameter, aWSa_{\rm WS}, is 0.67 fm for set A and 0.80 fm for set B, respectively. The 2s1/22s_{1/2} sp energy, independent of V1V_{1}, is set to 0.01-0.01 MeV, as indicated by the thin horizontal line.

We vary the spin-orbit strength V1V_{1} in Eq. (15) to simulate different conditions for the 61Ca structure. Figure 1 plots the sp energies of the 0g9/20g_{9/2} and 1d5/21d_{5/2} orbits. Note that the standard value of V1V_{1} for 60Ca is 0.44V0190.44V_{0}\approx 19 MeV for set A [46], resulting in the sp levels of 0g9/20g_{9/2}, 2s1/22s_{1/2}, and 1d5/21d_{5/2} order. In Fig. 1, both sets exhibit similar sp energy dependence as a function of V1V_{1}, though the value of V1V_{1} for set B tends to be stronger than that for set A to make 0g9/20g_{9/2} and 2s1/22s_{1/2} states degenerate due to more diffused nuclear surface. To fix the range of V1V_{1}, the following two extreme cases are considered:

Vanishing spin-orbit limit. In Ref. [37], the 1d5/21d_{5/2} and 0g9/20g_{9/2} sp energies of 61Ca are predicted to be 1.14 and 2.29 MeV. To realize this situation, we need to take a very small V1V_{1} value, V10V_{1}\approx 0 for set A and 5\approx 5 MeV for set B, respectively. We call this choice of V1V_{1} the vanishing spin-orbit (so) limit.

Degenerate sgsg limit. The sp energies of the 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits are degenerate at V1=11.40V_{1}=11.40 MeV for set A and at V1=21.09V_{1}=21.09 MeV for set B. This choice of V1V_{1} is called the degenerate sgsg limit. Despite both sp energies being the same, their respective root-mean-square (rms) radii are quite different: 36.0 and 5.04 fm for the 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits of set A, and 36.1 and 5.28 fm for these orbits of set B. In both sets, the halo features are noticed in the ss states, while the rms radii shrink significantly for the gg states due to the l=4l=4 centrifugal barrier.

We examine the energy spectrum of 62Ca by varying V1V_{1} between the vanishing and degenerate limits. As the outcomes of both sets A and B are qualitatively the same, we discuss the results obtained with set A, unless otherwise mentioned.

III.2 Tests of FFLPG expansion

Before discussing the structure of 62Ca, we evaluate the power of the FFLPG approach. We use the Minnesota (MN) potential [48] for v12v_{12} of Eq. (1). The MN potential, a soft-core central potential, is designed to reasonably well reproduce the energies and sizes of ss-shell nuclei [28]. Since the two neutrons should be antisymmetric in the spin-orbital space, it is expected that they gain the attraction mostly in the relative ss-wave and spin-singlet state. The spin-orbit and tensor components of v12v_{12} are therefore expected to play an insignificant role in the core+nn+nn model. Refer to Refs. [49, 50, 51] for an example of demonstrating that the MN potential gives similar results as the realistic nn-nn ones.

Let LL denote the relative orbital angular momentum between the two neutrons. The MN potential in the spin-singlet and even LL channel reads 200e1.487r291.85e0.405r2200e^{-1.487r^{2}}-91.85e^{-0.405r^{2}} in MeV, where rr is the two-nucleon distance in fm. For the spin-triplet and odd LL channel it is given by (200e1.487r2178e0.639r2)(u1)(200e^{-1.487r^{2}}-178e^{-0.639r^{2}})(u-1), where uu is a parameter and usually taken to be around 1. In what follows, uu is set to 1 and no interaction acts between the two neutrons in the spin-triplet and odd LL channel.

Refer to caption
Figure 2: Energy convergence of the Ca60+n+n{}^{60}{\rm Ca}+n+n system with Jπ=0+J^{\pi}=0^{+} as a function of the number of basis KK. The calculations are performed only with l1=l2=0l_{1}=l_{2}=0 channel. Thick lines are the results with FFLPG functions, while thin lines are those with the pseudo-potential method. See text for details.

To demonstrate the power of the FFLPG approach, we focus on such a state that is dominated by the weakly bound 2s1/22s_{1/2} orbit because the approach is expected to have the advantage in describing nodal orbits properly. We use the set A potential with V1=0V_{1}=0 and include the LPG bases restricted to l1=l2=0l_{1}=l_{2}=0. Figure 2 plots the energy as a function of KK, for different choices of kmaxk_{\rm max}. The candidates for the basis states are generated randomly in the interval [0.1,40][0.1,40] fm for b=1/ab=1/\sqrt{a} and in the interval [0,kmax][0,k_{\rm max}] for kk, and the best one is selected by the SVM algorithm. The results with the pseudo-potential method are also shown for comparison. The FFLPG approach significantly improves the energy convergence. We confirm that kmax=4k_{\rm max}=4 truncation virtually gives the same curve as the one with kmax=3k_{\rm max}=3. The FFLPG calculation leads to convergence within a few tens of basis functions, while the pseudo-potential method needs more than a hundred bases to reach convergence. As for the pseudo-potential method, it appears that the convergence is fastest when the ordinary Gaussians with kmax=0k_{\rm max}=0 are used, and there is no significant advantage in using the LPG bases with kmax>0k_{\rm max}>0. Probably this occurs because the k=0k=0 bases have the largest overlap with the 0s1/20s_{1/2} and 1s1/21s_{1/2} forbidden states, with the orthogonality requirement being most efficiently met by a superposition of those ordinary Gaussians. However, the energy obtained with K=150K=150 is 0.0537-0.0537 for kmax=0k_{\rm max}=0 and 0.0533-0.0533 MeV for kmax=1k_{\rm max}=1, respectively, which still misses the FFLPG energy of 0.0539-0.0539 MeV. What is more serious in the pseudo-potential method is its numerical instability. Because the k=0,1k=0,1 bases have in general large overlap with the forbidden states, the calculation becomes numerically unstable due to the large prefactor λ\lambda in Eq. (14). It is very hard to extend the basis size without breaking the linear independence of the selected bases. With kmax=2,3k_{\rm max}=2,3 this instability problem is recovered and we get the energy of 0.0539-0.0539 MeV by the pseudo-potential method with K=200K=200.

Here, we should make a comment on the role of kk in the LPG basis. As shown above, the FFLPG basis with kmax>0k_{\rm max}>0 accelerates the energy convergence compared to the ordinary Gaussians with k=0k=0. To understand the reason, we plot in Fig. 3 the ss-wave radial functions with different kk values for some choices of aa. The thick curves are for ϕ¯k0a\bar{\phi}^{a}_{k0} in Eq. (11), while the thin curves are for ϕk0a\phi^{a}_{k0} that in general contains forbidden states. Note that the forbidden states are ψ00ν\psi^{\nu}_{00} and ψ10ν\psi^{\nu}_{10}. In the case of b=1/νb0b=1/\sqrt{\nu}\equiv b_{0} the FFLPG function thus vanishes for k=0k=0 and 1, whereas it has two nodes at short distances for k=2k=2 and 3 because it is orthogonal to the forbidden states. With increasing bb, the amplitude of the FFLPG function becomes smaller in the inner region and finally has no node for k=3k=3 at b=2.5b0b=2.5b_{0} due to small overlap with the forbidden states. The FFLPG basis with k=0k=0 has small amplitude beyond the nuclear surface. With increasing kk, however, it has large amplitude beyond the nuclear surface and damped amplitude at short distances. This property meets the requirement needed for the neutron orbits around the core nucleus and offers the possibility of efficiently describing neutron orbits with large radial extension. The FFLPG basis is also advantageous to gain the energy from the two-neutron interaction because it can have large relative ss-wave components around the nuclear surface.

Refer to caption
Figure 3: ss-wave radial functions of FFLPG ϕ¯k0a{\bar{\phi}}^{a}_{k0} (thick curves) and LPG ϕk0a\phi^{a}_{k0} (thin curves) for different kk values. The width parameter b=1/ab=1/\sqrt{a} is chosen to be (a) b0b_{0}, (b) 1.5b01.5\,b_{0}, (c) 2b02\,b_{0}, and (d) 2.5b02.5\,b_{0}, where b0=1/ν=2.17b_{0}=1/\sqrt{\nu}=2.17 fm with ν\nu being the HO parameter of the forbidden states for Ca60+n{}^{60}{\rm Ca}+n.

III.3 Application to 62Ca

The test example presented in the previous subsection confirms that the FFLPG offer much faster and numerically stabler results than the pseudo-potential method. In this subsection, we study the structure of 62Ca in the Ca60+n+n{}^{60}{\rm Ca}+n+n model using the FFLPG expansion. We take kmax=3k_{\rm max}=3 and lmax=10l_{\rm max}=10 and generate the Gaussian fall-off parameter b=1/ab=1/\sqrt{a} in the interval [0.1, 40] fm. All possible different combinations of (l1,j1)(l_{1},j_{1}) and (l2,j2)(l_{2},j_{2}) are taken into account. For example, the number of combinations are respectively 21, 55, and 69 for Jπ=0+J^{\pi}=0^{+}, 2+2^{+}, and 4+4^{+} states.

We exhibit in Fig. 4 the energies of the states with Jπ=0+,2+J^{\pi}=0^{+},2^{+}, and 4+4^{+} as a function of KK. The spin-orbit strength is V1=11.40V_{1}=11.40 MeV, set A of the degenerate sgsg limit. Note that the energy converges rapidly on a few hundred bases for all the states. This is because all the bases are made free from the forbidden states. The fourth digit of the energy does not change on K=1500K=1500 for these lowest states. The second bound states are found for the 0+0^{+}, 2+2^{+}, and 4+4^{+} states. We further increase the number of basis to lower the second bound states and confirm that they converge very well at K=2000K=2000.

Refer to caption
Figure 4: Energies of Jπ=0+,2+J^{\pi}=0^{+},2^{+}, and 4+4^{+} states of 62Ca as a function of the number of basis KK. The energy drawn is from the two-neutron threshold, and the spin-orbit strength is V1=11.40V_{1}=11.40 MeV, set A of the degenerate sgsg limit.
Table 1: lmaxl_{\rm max}-dependence of the properties of two 0+0^{+} states. EE is the energy from the 60Ca+nn+nn threshold, ϵ2n=i=12(Ti+Ui)\epsilon_{2n}=\langle\sum_{i=1}^{2}(T_{i}+U_{i})\rangle is the sum of the two-neutron sp energies, and r12=(𝒓1𝒓2)2r_{12}=\sqrt{\langle(\bm{r}_{1}-\bm{r}_{2})^{2}\rangle} is the rms relative distance between the two neutrons. Energy is in units of MeV and length is in units of fm.
lmaxl_{\rm max} 4 6 8 10
01+0_{1}^{+} EE 0.99-0.99 1.13-1.13 1.16-1.16 1.19-1.19
ϵ2n\epsilon_{2n} 0.14 0.33 0.38 0.43
v12\left<v_{12}\right> 1.13-1.13 1.39-1.39 1.47-1.47 1.55-1.55
r12r_{12} 7.10 6.85 6.84 6.84
02+0_{2}^{+} EE 0.10-0.10 0.12-0.12 0.14-0.14 0.14-0.14
ϵ2n\epsilon_{2n} 0.17 0.27 0.33 0.35
v12\left<v_{12}\right> 0.27-0.27 0.39-0.39 0.45-0.45 0.48-0.48
r12r_{12} 19.6 18.0 17.3 17.1

The truncation of lmax=10l_{\rm max}=10 appears considerably large. One might question why so large a value is needed. Most of the sp levels are unbound in the present case, that is, they are nonresonant continuum states. The large value of lmaxl_{\rm max} therefore suggests the need to account of the continuum effect to bind the two neutrons. Table 1 lists the lmaxl_{\rm max} dependence of the basic properties of the 01+0_{1}^{+} and 02+0_{2}^{+} states. As the table indicates clearly, ϵ2n\epsilon_{2n} increases, while v12\langle v_{12}\rangle gets more attractive as a function of lmaxl_{\rm max}. The lmaxl_{\rm max} dependence of the two-neutron rms distance, r12r_{12}, shows an apparent correlation with v12\langle v_{12}\rangle. Although we confirm that the truncation with lmax=10l_{\rm max}=10 takes into account most of the continuum effect, it appears that there may be still some room to improve the convergence by increasing lmaxl_{\rm max} further.

In addition to the six states drawn in Fig. 4, we obtain two bound states with 6+6^{+} and 8+8^{+}. Table 2 summarizes the energies, the rms neutron radii, and the occupation probabilities for those bound states. Note that the 0g9/20g_{9/2} and 2s1/22s_{1/2} sp states are set to be degenerate. The ground state exhibits nonhalo structure, occupying the (g9/2)2(g_{9/2})^{2} configuration by 94%. The configuration also produces the 21+,41+,6+2^{+}_{1},4^{+}_{1},6^{+}, and 8+8^{+} bound states that have almost the same structure as the ground state. Halo structure is realized as the 02+0^{+}_{2} and 42+4^{+}_{2} states. Both states have the rms neutron radius, r2nr_{2n}, larger than 10 fm. As shown by the occupation probability, the 02+0^{+}_{2} state may be called an ss-wave two-neutron halo. It should be noted, however, that its r2nr_{2n} value of about 13 fm is by far smaller than the rms radius of the 2s1/22s_{1/2} sp orbit, which is about 36 fm. This dramatic reduction is of course due to the correlated motion of the two neutrons. The one-neutron halo character of the 42+4^{+}_{2} state is brought about by the coupling of the g9/2g_{9/2} neutron with the s1/2s_{1/2} neutron, as revealed by the occupation probabilities. This suggests that both states of two- and one-neutron halo structure can coexist within a narrow energy spacing. Despite the fact that it is barely bound, the 22+2^{+}_{2} state shows no characteristics of halo structure. Its r2nr_{2n} value is considerably large, 6.77 fm, but it is by far smaller than those of the 02+0^{+}_{2} and 42+4^{+}_{2} states. Note that the recoil kinetic energy is not negligible as the binding energy is small. This is because the main configuration of the 22+2^{+}_{2} state is (g9/2d5/2)(g_{9/2}d_{5/2}).

Table 2 also lists the decomposition of the energy EE into the sp energy, the recoil kinetic energy, the two-neutron interaction energy, and the rms two-neutron distance. The positive value of ϵ2n\epsilon_{2n} indicates that the two neutrons move mostly in the continuum states. The recoil kinetic energies are small due to the factor 1/60mN1/60m_{N}. Clearly, the two-neutron interaction v12v_{12} plays a decisive role to make the two neutrons bound. At a closer look, the 22+2_{2}^{+} state shows the largest sp energy and a relatively large energy gain from the two-neutron interaction energy. The state is realized by the coupling of the continuum states such as the d5/2d_{5/2} and other high ll orbits due to the two-neutron correlation.

The (g9/2)2(g_{9/2})^{2} dominance in the ground state can be explained by the pairing antihalo effect pointed out in Ref. [16]. A combination of the higher angular momentum states is more advantageous to gain energy from the pairing. In fact, v12\left<v_{12}\right> values are 1.55-1.55 and 0.48-0.48 MeV for 01+0^{+}_{1} and 02+0_{2}^{+}, respectively. To make the ground state a halo, the energy difference between the 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits should be sufficiently larger than the energy gain of the two-neutron interaction with the (g9/2)2(g_{9/2})^{2} configuration.

As mentioned above, the 42+4^{+}_{2} state is found to have one-neutron halo structure constructed dominantly from the (s1/2g9/2)(s_{1/2}g_{9/2}) configuration. That configuration would suggest a doublet state with 5+5^{+}. However, its dominant configuration is in the spin-triplet and odd LL channel. Possible existence of this unnatural-parity state crucially depends on the choice of uu parameter. Since we set u=1u=1, v12v_{12} vanishes and no doublet state appears. To be more definitive about its existence, we have to test other realistic nn-nn potentials.

Table 2: Properties of the spectrum of 62Ca. The sp energies of 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits are set to be 10-10 keV using the parameters of the degenerate sgsg limit of set A. The sp energy of the 1d5/21d_{5/2} orbit is 0.77 MeV. See Table 1 for the definitions of E,ϵ2nE,\ \epsilon_{2n}, and r12r_{12}. r2n=12(r12+r22)r_{2n}=\sqrt{\langle\frac{1}{2}(r_{1}^{2}+r_{2}^{2})\rangle} is the rms neutron radius, and Trec=160mN𝒑1𝒑2\langle T_{\rm rec}\rangle=\langle\frac{1}{60m_{N}}\bm{p}_{1}\cdot\bm{p}_{2}\rangle is the recoil kinetic energy. PxyP_{xy} denotes the occupation probability of finding x,yx,y sp orbits, where xx and yy stand for s1/2,g9/2,d5/2s_{1/2},\ g_{9/2},\ d_{5/2}, respectively, and ΔP=1xyPxy\Delta P=1-\sum_{xy}P_{xy}.
JπJ^{\pi} EE(MeV) r2nr_{2n}(fm) PssP_{ss} PggP_{gg} PddP_{dd} PsgP_{sg} PsdP_{sd} PgdP_{gd} ΔP\Delta P ϵ2n\epsilon_{2n}(MeV) Trec\left<T_{\rm rec}\right>(MeV) v12\left<v_{12}\right>(MeV) r12r_{12}(fm)
01+0_{1}^{+} 1.19-1.19 5.08 0.01 0.94 0.02 0.02 0.43 -0.07 -1.55 6.84
21+2_{1}^{+} 0.74-0.74 5.12 0.86 0.01 0.01 0.09 0.03 0.40 -0.06 -1.08 6.96
41+4_{1}^{+} 0.36-0.36 5.35 0.87 0.00 0.09 0.03 0.01 0.14 -0.03 -0.47 7.43
6+6^{+} 0.22-0.22 5.03 0.99 0.01 0.01 0.04 -0.02 -0.24 7.05
8+8^{+} 0.21-0.21 5.02 0.99 0.01 0.06 -0.01 -0.26 7.06
02+0_{2}^{+} 0.14-0.14 12.8 0.91 0.02 0.05 0.03 0.35 -0.01 -0.48 17.1
42+4_{2}^{+} 0.11-0.11 10.1 0.10 0.00 0.87 0.02 0.01 0.08 -0.01 -0.18 14.3
22+2_{2}^{+} -0.014 6.77 0.13 0.03 0.14 0.64 0.06 1.04 -0.04 -1.02 9.09

As discussed above, when the spin-orbit strength V1V_{1} of the nn-60Ca potential is taken to be the degenerate sgsg limit, the Jπ=0+J^{\pi}=0^{+} halo structure appears as the excited state. It is interesting to examine how the 0+0^{+} state changes if the spin-orbit strength is weakened towards the vanishing so limit. We introduce a multiplicative factor ff and set the spin-orbit strength as V1=fV1(0f1)V_{1}=fV_{1}^{\prime}\,(0\leq f\leq 1) with V1=11.40V_{1}^{\prime}=11.40 MeV for set A. The limit of f0f\approx 0 is the case used in Ref. [37], where the energy difference between the 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits is large (see Fig. 1) and only one bound state of two-neutron halo structure is predicted.

Refer to caption
Figure 5: (a) Energies, (b) rms two-neutron radii, and (c) occupation probabilities of the 01+0_{1}^{+} and 02+0_{2}^{+} states of 62Ca as a function of ff. The spin-orbit strength is taken to be V1=fV1V_{1}=fV_{1}^{\prime} with V1=11.40V_{1}^{\prime}=11.40 MeV for set A.

Figure 5 (a) displays the energy of the 0+0^{+} state as a function of ff. Only one 0+0^{+} state appears for f<0.7f<0.7 and the existence of the second 0+0^{+} state is possible for f0.7f\geq 0.7. Figure 5 (b) shows the corresponding rms neutron radii. The halo structure having a radius larger than 10 fm emerges in the ground state for f<0.7f<0.7. For f>0.7f>0.7 the 02+0_{2}^{+} state exhibits the two-neutron halo structure, whereas the ground state turns out to be a compact state. At f=0.7f=0.7 the ground state shows intermediate structure between halo and compact states. The rms radius of the 02+0_{2}^{+} state is extremely large due to the small binding energy of -0.03 MeV. This behavior can be understood by showing the occupation probabilities. Figure 5 (c) shows the occupation probabilities of finding (s1/2)2(s_{1/2})^{2} and (g9/2)2(g_{9/2})^{2} components in the 0+0^{+} states. The contributions of (d5/2)2(d_{5/2})^{2} component for the 01+0^{+}_{1} state are at most 0.09 at f=0.7f=0.7 and other contributions are less than 0.05 in total. For the 02+0_{2}^{+} state, those contributions are less about 0.01. As expected, the (s1/2)2(s_{1/2})^{2} component dominates in the ground state at f=0f=0, resulting in a large rms radius of 12 fm, almost the same structure as the 02+0_{2}^{+} state at f=1f=1. By increasing ff, i.e., reducing the energy difference between the 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits, the occupation probability of the 0g9/20g_{9/2} orbit increases gradually and rises suddenly for f>0.6f>0.6. The rms radius decreases simultaneously with the growing occupation of the 0g9/20g_{9/2} orbit that has a much smaller radius. An almost equal mixing of the (s1/2)2(s_{1/2})^{2} and (g9/2)2(g_{9/2})^{2} configurations occurs at f=0.7f=0.7. Here the energy difference between the 2s1/22s_{1/2} and 0g9/20g_{9/2} orbit is approximately 0.5 MeV, which is comparable to the difference of v12/2\left<v_{12}\right>/2 for the two 0+0^{+} states as shown in Table 2. Finally, the ground state becomes g9/2g_{9/2}-dominant at f=1f=1, where the sp energies of the 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits are degenerate, while the 02+0_{2}^{+} state exhibits two-neutron halo structure consisting of the (s1/2)2(s_{1/2})^{2} configuration.

III.4 Probable halo ground state of 72Ca

An extension to the Ca70+n+n{}^{70}{\rm Ca}+n+n model is straightforward. The reader is referred to Ref. [19] for the structure of 72Ca by the hyperspherical method. We follow the case of 62Ca starting from discussing some constraints on the phenomenological WS potential parameters.

The nn-Ca70{}^{70}{\rm Ca} potential should bind the 0g9/20g_{9/2} orbit due to the assumption of the N=50N=50 core. However, no bound 0g9/20g_{9/2} orbit is generated if we assume the parametrization of Ref. [46]. Instead of searching for such a potential that binds the 0g9/20g_{9/2} orbit, we simply assume that all the occupied neutron orbits including the 0g9/20g_{9/2} orbit are described by the HO functions with the oscillator parameter ν\nu^{\prime} that is scaled from the 60Ca parameter ν\nu by ν=(60/70)1/3ν\nu^{\prime}=(60/70)^{1/3}\nu. The sp energy spectrum of 71Ca with respect to V1V_{1} is virtually the same as Fig. 1 except for the absence of the 0g9/20g_{9/2} level.

To bind the 2s1/22s_{1/2} orbit at 0.01-0.01 MeV, the strength V0V_{0} turns out to be 40.26 MeV for set A and 38.32 MeV for set B. Since V0V_{0} gets weaker than the 60Ca-nn case, we have only one bound 0+0^{+} state that has two-neutron halo structure if V1V_{1} is taken to be the same as that of 61Ca: The energy, rms neutron radius, and (s1/2)2(s_{1/2})^{2} and (d5/2)2(d_{5/2})^{2} occupation probabilities are 0.14-0.14 MeV, 12.7 fm, and 0.89 and 0.08 for set A, whereas for set B they are 0.13-0.13 MeV, 13.3 fm, and 0.87 and 0.11, respectively. The nn-70Ca potential used here makes the energy gap between the 2s1/22s_{1/2} and 1d5/21d_{5/2} orbits too large to mix those sp states. To make the 1d5/21d_{5/2} orbit bound is very unlikely because V1V_{1} has to be taken more than two times larger than the standard value. Within the present phenomenological nn-Ca70{}^{70}{\rm Ca} potential, the halo structure emerges if the energy of the 2s1/22s_{1/2} state is close to zero, which is the same conclusion drawn in Ref. [19].

IV Conclusions and prospects

We have developed a forbidden-state-free locally peaked Gaussian expansion method to describe the weakly bound correlated neutron motion around a heavy core. The power of this approach has been tested by taking examples of the weakly bound three-body systems Ca60,70+n+n{}^{60,70}{\rm Ca}+n+n. The expansion proposed here accelerates the energy convergence much faster and offers numerically by far stabler results than the pseudo potential projection method [24]. The present method allows us to predict very weakly bound excited states. The energy spectra of unknown 62,72Ca have been calculated to discuss the possibility of halo-structure emergence by varying the single-particle levels of unknown Ca60,70+n{}^{60,70}{\rm Ca}+n systems.

In the extreme single-particle level ordering of 2s1/2,1d5/22s_{1/2},1d_{5/2}, and 0g9/20g_{9/2} orbits predicted by Ref. [37], only one bound state is found in 62Ca, exhibiting ss-wave two-neutron halo structure. The emergence of the halo structure in the ground state in fact strongly depends on the energy difference between the 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits for 62Ca and between the 2s1/22s_{1/2} and 1d5/21d_{5/2} orbits for 72Ca. Under the very limited condition that the 2s1/22s_{1/2} and 0g9/20g_{9/2} orbits are bound and degenerate, however, the ground state becomes a 0+0^{+} nonhalo state dominated by the 0g9/20g_{9/2} orbit, consistent with the pairing antihalo effect [16]. In that case, a 0+0^{+} two-neutron halo state appears slightly above the ground state. In addition a 4+4^{+} one-neutron halo state appears at almost the same excitation energy as the 0+0^{+} excited state. Apparently experimental information on the single-neutron levels of Ca60,70+n{}^{60,70}{\rm Ca}+n system is crucially important to identify the structure of these neutron-rich Ca isotopes.

The method proposed here can straightforwardly be extended to core plus more-nucleon systems. Since it is advantageous to describe localized orbits near the surface of the core, it is interesting to apply the method for alpha-decay phenomena of heavy nuclei. The result of Ref. [30] for the 212Po problem could be improved with locally peaked Gaussians. A systematic analysis of the degree of the alpha clustering near the nuclear surface is interesting as it has recently been realized in alpha-knockout reactions on Sn isotopes [52].

Acknowledgements.
This work was in part supported by JSPS KAKENHI Grant No. JP18K03635 and the Collaborative Research Program 2021, Information Initiative Center, Hokkaido University. For partial support, L.T. thanks Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP) (Projects No. 2017/05660-0 and No. 2015/11828-5) and Conselho Nacional de Desenvolvimento Científico e Tecnológico [Projects No. 304469/2019-0 and No. 464898/2014-5 (INCT-FNA)]; M.A.S. thanks Coordenação de Aperfeiçoamento de Pessoal de Nível Superior. Y.S. is indebted to L.T. for his generous invitation to São Paulo for September to December 2015 through the FAPESP grant, which made it possible to start this collaboration.

References

  • [1] D. S. Ahn, N. Fukuda, H. Geissel, N. Inabe, N. Iwasa, T. Kubo, K. Kusaka, D. J. Morrissey, D. Murai, T. Nakamura et al., Phys. Rev. Lett. 123, 212501 (2019).
  • [2] I. Tanihata, H. Hamagaki, O. Hashimoto, Y. Shida, N. Yoshikawa, K. Sugimoto, O. Yamakawa, T. Kobayashi, and N. Takahashi, Phys. Rev. Lett. 55, 2676 (1985).
  • [3] I. Tanihata, T. Kobayashi, O. Yamakawa, S. Shimoura, K. Ekuni, K. Sugimoto, N. Takahashi, T. Shimoda, and H. Sato, Phys. Lett. B206, 592 (1988).
  • [4] T. Suzuki, R. Kanungo, O. Bochkarev, L. Chulkov, D. Cortina, M. Fukuda, H. Geissel, M. Hellström, M. Ivanov, R. Janik et al., Nucl. Phys. A658, 313 (1999).
  • [5] W. Horiuchi and Y. Suzuki, Phys. Rev. C 74, 034311 (2006).
  • [6] K. Tanaka, T. Yamaguchi, T. Suzuki, T. Ohtsubo, M. Fukuda, D. Nishimura, M. Takechi, K. Ogata, A. Ozawa, T. Izumikawa et al., Phys. Rev. Lett. 104, 062701 (2010).
  • [7] Y. Togano, T. Nakamura, Y. Kondo, J. A. Tostevin, A. T. Saito, J. Gibelin, N. A.Orr, N. L.Achouri, T. Aumann, H. Baba et al., Phys. Lett. B 761, 412 (2016).
  • [8] I. Tanihata, H. Savajols, and R. Kanungo, Prog. Part. Nucl. Phys. 68, 215 (2013), and references therein.
  • [9] S. Bagchi, R. Kanungo, Y. K. Tanaka, H. Geissel, P. Doornenbal, W. Horiuchi, G. Hagen, T. Suzuki, N. Tsunoda, D. S. Ahn et al., Phys. Rev. Lett. 124, 222504 (2020).
  • [10] V. N. Efimov, Phys. Lett. B 33, 563 (1970).
  • [11] D. V. Fedorov, A. S. Jensen, and K. Riisager, Phys. Rev. Lett. 73, 2817 (1994).
  • [12] T. Frederico, A. Delfino, L. Tomio, and M. T. Yamashita, Prog. Part. Nucl. Phys. 67, 939 (2012).
  • [13] J. Singh, J. Casal, W. Horiuchi, L. Fortunato, and A. Vitturi, Phys. Rev. C 101, 024310 (2020).
  • [14] L. Fortunato, J. Casal, W. Horiuchi, J. Singh, and A. Vitturi, Commun. Phys. 3, 132 (2020).
  • [15] J. Casal, J. Singh, L. Fortunato, W. Horiuchi, and A. Vitturi, Phys. Rev. C 102, 064627 (2020).
  • [16] H. Masui, W. Horiuchi, and M. Kimura, Phys. Rev. C 101, 041303(R) (2020).
  • [17] N. Michel, J. G. Li, F. R. Xu, and W. Zuo, Phys. Rev. C 101, 031301(R) (2020).
  • [18] G. Hagen, P. Hagen, H.-W. Hammer, and L. Platter, Phys. Rev. Lett. 111, 132501 (2013).
  • [19] D. Hove, E. Garrido, P. Sarriguren, D. V. Fedorov, H. O. U. Fynbo, A. S. Jensen, and N. T. Zinner, Phys. Rev. Lett. 120, 052502 (2018).
  • [20] M. A. Shalchi, M. T. Yamashita, M. R. Hadizadeh, E. Garrido, L. Tomio, and T. Frederico, Phys. Rev. A 97, 012701 (2018).
  • [21] J. P. Incao, J. Phys. B: At. Mol. Opt. Phys. 51, 043001 (2018).
  • [22] S. Saito, Prog. Theor. Phys. 40, 893 (1968).
  • [23] S. Saito, Prog. Theor. Phys. 62, 11 (1977).
  • [24] V. I. Kukulin and V. N. Pomerantsev, Ann. Phys. (NY) 111, 330 (1978).
  • [25] K. Varga and Y. Suzuki, Phys. Rev. C 52, 2885 (1995).
  • [26] Y. Suzuki and K. Varga, Stochastic Variational Approach to Quantum-Mechanical Few-Body Problems, Lecture Notes in Physics (Springer, Berlin, 1998), Vol. m54.
  • [27] J. Mitroy, S. Bubin, W. Horiuchi, Y. Suzuki, L. Adamowicz, W. Cencek, K. Szalewicz, J. Komasa, D. Blume, and K. Varga, Rev. Mod. Phys. 85, 693 (2013).
  • [28] Y. Suzuki, W. Horiuchi, M. Orabi, and K. Arai, Few-Body Syst. 42, 33 (2008).
  • [29] W. Horiuchi and Y. Suzuki, Phys. Rev. C 89, 011304(R) (2014).
  • [30] K. Varga and R. J. Liotta, Phys. Rev. C 50, R1292 (1994).
  • [31] Y. Suzuki and W. Horiuchi, Phys. Rev. C 95, 044320 (2017).
  • [32] Y. Suzuki, Phys. Rev. C 101, 014002 (2020).
  • [33] O. B. Tarasov, D. S. Ahn, D. Bazin, N. Fukuda, A. Gade, M. Hausmann, N. Inabe, S. Ishikawa, N. Iwasa, K. Kawata et al., Phys. Rev. Lett. 121, 022501 (2018).
  • [34] S. Michimasa, M. Kobayashi, Y. Kiyokawa, S. Ota, D. S. Ahn, H. Baba, G. P. A. Berg, M. Dozono, N. Fukuda, T. Furuno et al., Phys. Rev. Lett. 121, 022506 (2018).
  • [35] R. F. Garcia Ruiz, M. L. Bissell, K. Blaum, A. Ekström, N. Frömmgen, G. Hagen, M. Hammen, K. Hebeler, J. D. Holt, G. R. Jansen et al., Nat. Phys. 12, 594 (2016).
  • [36] M. Tanaka, M. Takechi, M. Fukuda, D. Nishimura, T. Suzuki, Y. Tanaka, T. Moriguchi, D. S. Ahn, A. Aimaganbetov, M. Amano et al., Phys. Rev. Lett. 124, 102501 (2020).
  • [37] G. Hagen, M. Hjorth-Jensen, G. R. Jansen, R. Machleidt, and T. Papenbrock, Phys. Rev. Lett. 109, 032502 (2012).
  • [38] J. Erler, N. Birge, M. Kortelainen, W. Nazarewicz, E. Olsen, A. M. Perhac, and M. Stoitsov, Nature 486, 509 (2012).
  • [39] C. Forssén, G. Hagen, M. Hjorth-Jensen, W. Nazarewicz, and J. Rotureau, Phys. Scr. T152 014022 (2013).
  • [40] L. Coraggio, G. De Gregorio, A. Gargano ,N. Itaco, T. Fukui, Y. Z. Ma, and F. R. Xu, Phys. Rev. C 102, 054326 (2020).
  • [41] S. R. Stroberg, J. D. Holt, A. Schwenk, and J. Simonis, Phys. Rev. Lett. 126, 022501 (2021).
  • [42] J. Meng, H. Toki, J. Y. Zeng, S. Q. Zhang, and S.-G. Zhou, Phys. Rev. C 65, 041302(R) (2002).
  • [43] L. Neufcourt, Y. Cao, W. Nazarewicz, E. Olsen, and F. Viens, Phys. Rev. Lett. 122, 062502 (2019).
  • [44] Y. Suzuki and K. Ikeda, Phys. Rev. C 38, 410 (1988).
  • [45] Y. Suzuki and W. Horiuchi, Emergent Phenomena in Atomic Nuclei from Large-scale Modeling: A Symmetry-Guided Perspective (World Scientific, Singapore, 2017), Chap. 7, pp. 199-227.
  • [46] A. Bohr and B. R. Mottelson, Nuclear Structure, Vol. I (W. A. Benjamin, New York, 1975).
  • [47] W. Horiuchi, Y. Suzuki, B. Abu-Ibrahim, and A. Kohama, Phys. Rev. C 75, 044607 (2007).
  • [48] D. R. Thompson, M. LeMere, and Y. C. Tang, Nucl. Phys. A 286, 53 (1977).
  • [49] Y. Suzuki, H. Matsumura, and B. Abu-Ibrahim, Phys. Rev. C 70, 051302(R) (2004).
  • [50] W. Horiuchi and Y. Suzuki, Phys. Rev. C 73, 037304 (2006); ibid 74, 019901(E) (2006).
  • [51] W. Horiuchi and Y. Suzuki, Phys. Rev. C 76, 024311 (2007).
  • [52] J. Tanaka, Z. H. Yang, S. Typel, S. Adachi, S. Bai, P. van Beek, D. Beaumel, Y. Fujikawa, J. Han, S. Heil et al., Science 371, 260 (2021).