Next Article in Journal
Brain Immunoinformatics: A Symmetrical Link between Informatics, Wet Lab and the Clinic
Next Article in Special Issue
Quantifying the Effect of Initial Fluctuations on Isospin-Sensitive Observables from Heavy-Ion Collisions at Intermediate Energies
Previous Article in Journal
Software Defect Prediction Using Wrapper Feature Selection Based on Dynamic Re-Ranking Strategy
Previous Article in Special Issue
np-Pair Correlations in the Isovector Pairing Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cross-Shell Excitation in F and Ne Isotopes around N = 20

Sino-French Institute of Nuclear Engineering and Technology, Sun Yat-sen University, Zhuhai 519082, China
*
Author to whom correspondence should be addressed.
Submission received: 29 September 2021 / Revised: 2 November 2021 / Accepted: 8 November 2021 / Published: 12 November 2021
(This article belongs to the Special Issue Experiments and Theories of Radioactive Nuclear Beam Physics)

Abstract

:
Within the framework of the configuration–interaction shell model, the present work applies three effective interactions to investigate the effects of the cross-shell excitation on F and Ne isotopes around N = 20, which are significantly proton–neutron asymmetric, and have different properties compared with the proton–neutron symmetric nuclei. It is shown that cross-shell excitation is necessary in order to reproduce separation energies, neutron drip lines, and low-energy levels of these isotopes. Furthermore, the cross-shell excitation of (0–5)ħω is suggested to be important in the description of 29F and 30Ne. However, the three interactions are insufficient in describing the bound structure of 29,31Ne, and provide inconsistent shell structures and evolutions in the target nuclei. Their cross-shell interactions are suggested to be improved.

1. Introduction

Nuclei with neutron numbers close to 20 are of interest. According to the conventional shell model, the N = 20 shell is closed, defining the boundary between the sd shell and the pf shell [1,2]. As a result, nuclei with N = 20 should be magic, which is valid for nuclei near the β-stability line, where protons and neutrons are almost symmetrical in the light nuclear region. However, for lighter elements, such as F [3], Ne [4,5,6], Na [7], and Mg [8], it is known that the N = 20 shell gap vanishes in extremely proton–neutron asymmetric nuclei, leading to the existence of the “island of inversion” [7]. Moreover, the neutron drip lines of C, N, and O have 16 neutrons [9,10,11], while 31F and 34Ne are the heaviest bound isotopes of F and Ne, respectively [12]. The abrupt change of the neutron drip line also manifests the evolution of the N = 20 shell. Since 29F and 30Ne are the two lightest bound nuclei with N = 20, a systematic study on the F and Ne isotopes around 29F and 30Ne is significant.
F and Ne isotopes with N~20 are challenging to study due to their extreme proton–neutron asymmetry. For instance, 31F has been known to be bound since 1999 [13], while the unbound character of 32,33F was not determined until 2019 [3]. In fact, to date, only the masses of the F (Ne) isotopes with mass numbers less than 30 (32) have been experimentally determined [14]. On the other hand, some interesting phenomena in such nuclei have been experimentally discovered in recent years. For example, the “island of inversion” has been extended with several F and Ne isotopes [3,15], and exotic properties of F and Ne isotopes in the N~20 region have been discovered [16,17]. In 2020, the two-neutron halo structure in 29F [18] was experimentally identified. In 2021, neutron-unbound states were experimentally explored in 31Ne [19]. Therefore, theoretical models can be helpful to investigate the extremely neutron-rich F and Ne nuclei.
There usually are three types of theoretical models of nuclear structure: the ab initio theories [20,21], mean-field approaches [22], and shell models [1,2,23]. All three have been employed for neutron-rich F and Ne nuclei [5,24,25,26]. In the past two years, various theoretical models considering the continuum coupling for nuclei near the drip line—including the coupled-cluster method [18], Gamow shell model [27], and Green’s function method in the complex-momentum representation [25]—have been used to investigate the halo structure in 29,31F. Three-body systems have been built to analyze such exotic structures [28,29]. Nevertheless, a systematic study can be helpful to understand such unexpected phenomena.
The shell model is an essential tool for the systematic understanding of the nuclear structure. Recently, it has been employed in describing light, proton-rich, exotic nuclei and the related isospin symmetry breaking [30,31,32,33], interpreting the shell evolution in the medium mass region [34,35], reproducing isomeric states in the medium [36,37] and heavy nuclei [38,39,40], and investigating the α-decay properties of heavy-mass nuclei [41,42].
Within the shell model framework, it should be noticed that pure sd- or pf-shell model space is insufficient to study F and Ne isotopes around N = 20, due to the vanishing of the N = 20 shell gap. On this basis, the neutron excitation across the N = 20 shell is rather crucial. Fukunishi et al. employed an interaction in the model space, including the whole sd shell and the 0f7/2 and 1p3/2 orbits, and calculated some properties of 30Ne, ensuring the necessity of cross-shell excitation [43]. The SDPF-M interaction, proposed by Utsuno et al., was presented with its results on the average neutron occupancy in the pf shell of even-mass 26–34Ne [44]. Caurier et al. proposed the SDPF-U-MIX interaction, and showed the results of 29,31F and 30–32Ne, considering the contribution of the particle-hole states in [45].
In fact, cross-shell excitation is not a fresh idea; just as near the N = 20 shell, it can also be seen in the N = 8 and N = 28 shells. For example, shell model calculations with the YSOX interaction have succeeded in reproducing the properties of neutron-rich Be [46,47], B [48], C [49,50], and O [51] isotopes with the help of the cross-shell excitation beyond N = 8. In the sd region, the interpretation of the exotic β-γ-α decay mode in 20Na [52] also requires the neutron cross-shell excitation. Moreover, shell model studies with the SDPF-MU [53] and SDPF-U-SI [54] interactions were performed to describe the exotic Mg [55], Si [56], and S [53,54] isotopes, considering the neutron excitation across the N = 28 shell.
Despite the previous studies, a systematic investigation of the influences of cross-shell excitation on F and Ne isotopes around N = 20 is lacking. It would be interesting to know, for example, how many neutrons of cross-shell excitation should be considered if we wish to reproduce the neutron drip line of F and Ne. For this purpose, based on the widely used configuration–interaction shell model (CISM), this paper applied three different interactions—SDPF-M, SDPF-MU, and SDPF-U-SI—to investigate the role of the cross-shell excitation in the F and Ne isotopes with N~20.
The paper is organized as follows: in the next section, CISM is briefly introduced, and the details of the three interactions are given; then, the results from the shell model calculations are shown to analyze the effects of the cross-shell excitation on the F and Ne neutron-rich isotopes; finally, a summary is presented.

2. Theories

The nuclear structure problem is essentially a many-body quantum problem, where protons and neutrons interact in a complex manner. To this end, CISM assumes an inert core, and defines a model space where the remaining nucleons can occupy the orbits. In the truncated model space, the effective Hamiltonian—also called the effective interaction—can be built from the nuclear force. Usually, the nuclear force can be derived from either the nucleon–nucleon scattering experimental data [57] or the observed nuclear structure data [58], called realistic nuclear force and phenomenological nuclear force, respectively. The former starts in a more ab initio manner, which needs to deal with hard-core and in-medium effects before the utilization. On the other hand, the phenomenological nuclear force depends a lot on experiments. The three effective interactions applied in this work are all constructed in a phenomenological manner. They can partially include continuum-coupling effects and proton–neutron asymmetry via fitting nuclear structure data of nuclei near the drip line. Then, through diagonalizing the Hamiltonian matrix, which consists of single-particle energies (SPEs) and two-body matrix elements (TBMEs), the wavefunctions can be derived as eigenstates of the many-body Schrodinger equation. As a result, configuration-mixing is used to express each state.
There are normally three steps to perform CISM calculations in practice: The first step of CISM is to choose an appropriate model space. The second step, which is also the most critical step, is to construct the effective interaction in the model space. Finally, the shell model codes are used to diagonalize the Hamiltonian matrix. In this paper, the size of the model space is discussed by analyzing the effects of the cross-shell excitation. The input of the effective interaction is varied with three choices, while all of the calculations are performed with the code KSHELL [59,60,61].
To reproduce the structure of F and Ne isotopes around N = 20, the model space includes not only the sd-shell orbits, but also two or all pf-shell orbits. On this basis, the size of the model space can be further determined by restricting the number of neutrons of cross-shell excitation. Then, the interactions involving two major shells are divided into three parts: the sd-shell part, the pf-shell part, and the cross-shell part. The two former terms are based on well-known interactions in the sd- and pf-shell model space. Thus, the interactions constructed in the larger space can immediately describe the N~Z nuclei. As for F and Ne nuclei around N = 20, the cross-shell interactions can be crucial, since both the sd- and pf-shell configurations contribute significantly to these nuclei. In the following subsections, the details of the three interactions will be described.

2.1. SDPF-M

The SDPF-M interaction was proposed by Utsuno et al. for studying neutron-rich nuclei with nearly 20 neutrons [44]. The interaction takes 16O as the inert core, and considers the model space consisting of the full sd shell and the 0f7/2 and 1p3/2 orbits; it uses the USD interaction [62] and the KB interaction [63] for the sd-shell and pf-shell parts, respectively. More importantly, its cross-shell part starts from the MK interaction [64].
In detail, the MK interaction is close to a G-matrix interaction in character, except that MK has a repulsive triplet-odd central force. Warburton et al. [65] modified the eight TBMEs <0f7/20d3/2|V|0f7/20d3/2> J = 25,T = 0,1 of the MK interaction to build the cross-shell part of the SDPF interaction. The modified MK interaction is exactly the adopted cross-shell interaction of the SDPF-M interaction.
On the basis of the interactions—including the USD, KB, and modified MK interactions—two important modifications are performed: On the one hand, in order to reproduce the unbound property of 26O, modifications are added to the monopole interactions [66]. On the other hand, the pairing matrix elements of the USD interaction are reduced, as it implicitly includes the effects of the pf shell.
The finally obtained SDPF-M was successfully applied to study the exotic properties of the O [44], F [67], Ne [4,16,44], Na [68], Mg [8,44], and Si [44,69] isotopes, such as the unbound structure of 26,28O and the bound structure of 32,34Ne. This interaction did well in describing the nuclei with N ~ 20, and was used to study the evolution of the N = 20 shell gap [44].

2.2. SDPF-MU

The SDPF-MU interaction [53] is based on the USD interaction, the GXPF1B interaction [70], and the VMU interaction [71]. More precisely, its sd-shell part starts from the USD interaction, with the monopole part for the 0d3/2 and 0d5/2 orbits modified in the same way as that in the SDPF-M interaction. The GXPF1B interaction is used for the pf-shell part, whereas the TBMEs <0f7/20f7/2|V|0f7/20f7/2> J = 0,2 are those from the interaction KB3 [72], so as to better describe the nuclei around N = 22. Finally, the cross-shell part is composed of the VMU and the M3Y spin-orbit force [73]. To be clearer, VMU is the monopole-based universal force, including the Gaussian central force and the π + ρ tensor force. SDPF-MU has been used in the exotic F [3], Mg [55], Si [53,56], and S [53] isotopes [53,56].

2.3. SDPF-U-SI

The SDPF-U/SDPF-U-SI interaction [54] proposed by Nowacki et al. in 2009 is mainly independent of the two interactions mentioned previously. Its proposition aims firstly at nuclei with 8 ≤ Z ≤ 20 and 20 ≤ N ≤ 40. The model space comprises the full sd shell for the protons and the full pf shell for the neutrons. Thus, nuclei with around 28 neutrons, along with the neutron excitation across the N = 28 shell gap, are the main focus. This is actually based on the SDPF-NR interaction [74,75], which takes the USD interaction, KB’ interaction [72], and G-matrix of Kahana, Lee, and Scott [76] as the sd-shell interaction, pf-shell interaction, and cross-shell interaction, respectively. Iterations are realized to derive the proper monopole changes with the help of the newly discovered experimental data. For instance, after iterations, SDPF-U-SI is able to reproduce the level of the 3/21 state of 35Si and the newly discovered spectrum of 41Ca [54].
When applying the three interactions involving the orbits of two major shells, the center-of-mass (c.m.) correction is also considered. This work uses the model of Gloeckner and Lawson [77], which writes the Hamiltonian as H = HSM + βc.m.Hc.m. in order to decouple the original Hamiltonian (HSM) and the c.m. Hamiltonian (Hc.m.). According to [51], βc.m. = 10 is large enough in the calculations of low-lying states, and is used in this work.
In sum, the three interactions are proposed in different ways. The model space of SDPF-MU is the same as that of SDPF-U-SI, while it is larger than that of SDPF-M. Moreover, SDPF-MU and SDPF-U-SI are proposed primarily for nuclei with N ~ 28. The two interactions have been compared in the calculations of 42Si. Results showed that CISM calculations with SDPF-MU agreed better with existing data [56]. Nevertheless, in the calculations of nuclei around N = 20, the best interaction is not evident. Despite the differences, all of the interactions showed good performance for nuclei in the sdpf region. With the three interactions, this paper focuses on the effects of the cross-shell excitation on describing F and Ne nuclei around N = 20.

3. Results and Discussions

In this work, we performed CISM calculations to derive the binding energies and energy spectra of F and Ne neutron-rich isotopes. We started with three interactions—SDPF-M, SDPF-MU, and SDPF-U-SI—separately, in order to examine the effects of the cross-shell excitation on the ground-state properties and excitation energies of F and Ne isotopes near N = 20. The CISM calculations were carried out in three cases: with cross-shell excitation of the least neutrons (no for positive states and one for negative states, labeled as (0–1)ħω), with cross-shell excitation of at most two neutrons beyond the first case (labeled as (0–3)ħω), and with cross-shell excitation of at most four neutrons beyond the first case (labeled as (0–5)ħω). There are a total of nine groups of theoretical results for all of the involved nuclei.
In this section, the nine groups of ground-state properties and excitation energies are compared with experimental results in order to analyze the contributions of the cross-shell excitation, and then the results of configuration-mixing and occupancy are given to illustrate the analysis; furthermore, the shell structure and evolution described by the three interactions are discussed in order to complete the analysis.

3.1. Ground-State Properties

The foremost task of the theoretical study on F and Ne isotopes around N = 20 is to reproduce and explain the location of neutron drip lines by analyzing the separation energies of one neutron (Sn) and two neutrons (S2n). Normally, if the Sn (S2n) of a nucleus is negative, it will be single (double)-neutron(s) unbound. Thus, both the Sn and S2n of a bound nucleus should be positive. In addition, if adding one or two neutrons to a bound nucleus results in unbound nuclei, it can be speculated that such bound nuclei are located in the neutron drip line.
Table 1 shows the theoretical and experimental [14] separation energies of F and Ne nuclei around N = 20. For the nuclei with experimental separation energies listed in Table 1, the inclusion of the (2–3)ħω cross-shell excitation can globally improve the calculation accuracy. Figure 1 illustrates the root-mean-square errors (RMSEs) of the calculated Sn and S2n values of nuclei 2729F and 2831Ne, compared to experimental data. It can be seen that the RMSEs are smaller after considering the (2–3)ħω cross-shell excitation. However, the cross-shell excitation part of the SDPF-MU interaction should be improved, as the RMSEs of Sn in the nuclei of 2729F and 2831Ne are larger in the (0–5)ħω case than those in the (0–3)ħω case.
It should be noted that the description improvement of the ground-state properties caused by the cross-shell excitation is rather evident in 29F and 30Ne. The Sn and S2n values of the two nuclei with N = 20 are shown in Figure 2. In detail, CISM with the least cross-shell excitation underestimates the energies necessary to separate one or two neutrons from 29F and 30Ne. Then, the inclusion of higher ħω states leads to the increase in the Sn and S2n values of 29F and 30Ne, typically making the theoretical values closer to the observation results. For instance, in the calculation of Sn (29F) with SDPF-M, the further inclusion of (2–3)ħω states can take it from negative value to positive value, while that of (4–5)ħω states allows it to approach the experimental results further. In total, a model space including (0–5)ħω cross-shell excitation is suggested to sufficiently describe the ground-state properties of 29F and 30Ne within the framework of CISM.
The qualitative description of the ground states in the nuclei near the F and Ne drip line requires the consideration of the neutron cross-shell excitation beyond the (0–1)ħω case. Based on the Sn and S2n values, the neutron drip lines are finally determined. In the (0–1)ħω case, CISM calculations using SDPF-M, SDPF-MU, and SDPF-U-SI predict 28, 31, and 33, respectively, as the largest atomic number of bound F isotopes. However, according to recent experiments, the heaviest bound F isotope is 31F [12]. In the (0–3)ħω case, both SDPF-M and SDPF-U-SI succeed in reproducing the drip line of F; the results of Ne are rather similar. Only with (0–1)ħω do CISM calculations with SDPF-M predict an inner drip line of Ne, while those with SDPF-U-SI predict outer. Then, in the case of (0–3)ħω, the results of the two interactions are consistent with the observations. Though SDPF-MU with (0–1)ħω can reproduce the location of the drip line of F and Ne, it wrongly gives unbound results of 29F and 30Ne, while the neutron cross-shell excitation corrects the descriptions.
Nevertheless, whether or not CISM with (0–5)ħω can give a significantly better theoretical description of F and Ne chains than that with (0–3)ħω requires further study. For instance, the three interactions with (0–5)ħω all incorrectly result in the unbound properties of 29,31Ne, while those with (0–3)ħω do not, which should not happen in principle. Improvements are needed in the cross-shell excitation parts of the three interactions.
The cross-shell excitation is also fundamental to reproducing the ground-state spin of F and Ne nuclei around N = 20. Table 2 shows the ground-state spin of 27–31F and 28–34Ne, given by experiments or CISM calculations. In the (0–1)ħω case, the CISM calculation with the SDPF-M interaction cannot reproduce the ground-state spin of 28F and 29,31,32Ne, while those with SDPF-MU or with SDPF-U-SI fail, in 31Ne. In the (0–3)ħω case, the CISM calculation with SDPF-M corrects the ground-state spin of 28F and 31,32Ne. Then, in the (0–5)ħω case, the CISM calculations using the SDPF-MU interaction correct the ground-state spin of 31Ne calculated with (0–1)ħω and (0–3)ħω. Therefore, the cross-shell excitation is necessary in order to calculate the ground-state spin in F and Ne nuclei around N = 20, and more than (0–3)ħω should be considered. Furthermore, it is suggested that 4 and 3/2 should be the ground-state spin of 30F and 33Ne, respectively.
In total, the cross-shell excitation is crucial to describing the ground states of F and Ne isotopes around N = 20. Furthermore, (0–5)ħω excitation could be necessary to the 29F and 30Ne nuclei. The cross-shell excitation parts of the three interactions may not be well fixed yet.

3.2. Excitation Energies

Since many of the F and Ne nuclei around N = 20 are weakly bound, the experimental results of the excited states are rather rare. To date, there is no information about the excited states of the F (Ne) isotopes heavier than 29 (32) in the Nudat2 database [79]. Therefore, this paper focused on selected low-lying states. Table 3 shows the experimental and theoretical energy levels of the excited states in F and Ne nuclei around N = 20.
For F isotopes, the odd-A nuclei are our first focused. According to the shell model, such nuclei should have a 5/2+ ground state caused by a valence proton on the π0d5/2 orbit. When the valence proton moves to the π1s1/2 (π0d3/2) orbit, the spin of the excited states should be 1/2+ (3/2+). Although a state is rarely a pure, single-particle state in the view of CISM, the order 5/2+, 1/2+, and 3/2+ holds. This is because the π0d5/2 configuration dominates on the 5/2+ ground state, and similar explanations work for the other cases.
In all cases discussed in this paper, the theoretical ground-state spins of odd-A F isotopes are 5/2+. For the first 1/2+ state, the cross-shell excitation is shown to be necessary to the CISM calculation. It can be seen from Table 3 that the theoretical energy levels of 1/21+ of 27,29F are much higher than the existing experimental levels. From (0–1)ħω to (0–3)ħω, those energy levels become lower; then, from (0–3)ħω to (0–5)ħω, they decrease again. It can be noted that the latter decreases are less significant than the former. Overall, the inclusion of cross-shell excitation beyond the (0–1)ħω states makes the theoretical levels of the 1/21+ of odd-A 27,29F approach the experimental results significantly. Considering the experimental errors and the possible deficiency of the interactions involved, it is suggested that (0–3)ħω is insufficient to reproduce those 1/21+ states. In particular, the results of the 1/21+ state in 29F are illustrated in Figure 3; according to this figure, the necessity of (0–5)ħω is indicated in the interpretation of the energy spectrum of 29F.
Theoretically, the three lowest lying states of even-mass Ne isotopes should be 0+, 2+, and 4+ by orders. The CISM with SDPF-M wrongly gives 3 as the ground-state spin in the (0–1)ħω case, and corrects the result with (0–3)ħω. The levels of the 21+ and 41+ states in even-A 28–32Ne can be seen in Table 3. CISM calculations of these levels with the three interactions are consistent with experimental results. It can be seen that the SDPF-M interaction reasonably reproduces these levels, with RMSE values less than 0.4 MeV. The levels of the 21+ state in 30Ne are also presented in Figure 3. Including higher ħω states results in slight variations in the energy spectra of 28Ne, while the inclusion of (2–3)ħω states helps SDPF-M to better describe the 21+ state in 30Ne. For SDPF-U-SI, theoretical results with (0–5)ħω are the most consistent with the experimental results, both for the 21+ state in 30Ne, and for the overall levels in even-A 28–32Ne. However, it should also be noted that the global description accuracy of levels of SDPU-MU deteriorates with the enlargement of the model space. In sum, it is believed that the cross-shell excitation plays an important role in CISM calculations of Ne isotopes around N = 20.
The level of the first intruder state is also of interest. It was found that the levels of the first negative state of odd-A 27–31F and even-A 28–34Ne increase with the addition of ħω. Moreover, the level variation is more significant from the (0–1)ħω case to the (0–3)ħω case than from the (0–3)ħω case to the (0–5)ħω case.
Overall, the cross-shell excitation is necessary to well describe the excited states of F and Ne nuclei around N = 20, and it is suggested to consider (0–5)ħω in the calculations of 29F and 30Ne.

3.3. Configuration Occupancies

From Section 3.1 and Section 3.2, it is known that the cross-shell excitation is crucial to describe F and Ne nuclei around N = 20 within the CISM framework. This means that the N = 20 shell gap is reduced or vanished, and neutrons occupy the pf-shell orbits. This phenomenon can be illustrated with the configuration occupancy.
Table 4 shows the average neutrons that occupy the pf-shell orbits in the odd-F and even-Ne isotopes. For the N = 18 nuclei, there are large possibilities of neutron excitation across the N = 20 shell gap; thus, the interactions without cross-shell excitation are insufficient to reproduce the ground-state properties of 27F and 28Ne. Similar explanations can be found in the nuclei with N = 22 or 24. Thus, the overall description of the ground states in F and Ne nuclei near the neutron drip lines—including the determination of the neutron drip lines—requires the cross-shell excitation.
Among the nuclei investigated in this paper, 29F and 30Ne are the most special, as the sd shell is fully occupied, without cross-shell excitation. However, results in a larger model space show that neutrons in 29F and 30Ne occupy the pf-shell orbits. CISM calculations using the SDPF-M interaction with the (0–5)ħω case indicate that more than two neutrons in 29F and 30Ne excite across the N = 20 shell gap on average. This explains the necessity of a model space including (0–5)ħω excitation in describing 29F and 30Ne.
In detail, Table 5 gives the configuration occupancies of the ground states in 29F and 30Ne. In the (0–1)ħω case, 29F and 30Ne are pure π0d5/2 ν(0d3/2)4 and π(0d5/2)2 ν(0d3/2)4 configurations, respectively. However, results considering the cross-shell excitation show that the π0d5/2 ν(0d3/2)2 (0f7/2)2 configuration is strongly mixed with π0d5/2 ν(0d3/2)4 in 29F. CISM calculations with SDPF-M indicate that the π0d5/2 ν(0d3/2)2 (0f7/2)2 configuration is actually dominant in the ground state of 29F. In such a configuration, two neutrons of 29F (30Ne also) excite from the ν0d3/2 orbit to the ν0f7/2 orbit. The configurations π0d5/2 ν(0d3/2)2 (1p3/2)2 and π(0d5/2)2ν(0d3/2)2 (1p3/2)2, which represent the excitation of a pair of neutrons to the ν1p3/2 orbit, are also important in 29F and 30Ne, respectively. Therefore, it is necessary to consider the three kinds of configurations in 29F and 30Ne.
In conclusion, neutrons of F and Ne nuclei excite from sd-shell orbits to pf-shell orbits when the neutron number approaches 20. This phenomenon makes it necessary for F and Ne nuclei around N = 20 to include the cross-shell excitation. Taking 29F and 30Ne, for example, in both the ground state and excited states, the neutrons exciting to the pf-shell are in significant numbers, which explains why CISM calculations with (0–1)ħω or (0–3)ħω fail in describing 29F and 30Ne.

3.4. Shell Structure and Evolution

This section tries to explain why some CISM calculations, including (4–5)ħω states, are less consistent with experimental results. The shell structure and its evolution are compared among the three interactions via their effective single-particle energies (ESPEs) [80,81] and a monopole-frozen analysis [82].
For an orbit j, the ESPE can be expressed as [80,81]:
ε j = ε j c o r e + j V j j ψ N ^ j ψ
where ε j c o r e represents the SPE of the j orbit regarding the inert core, V j j represents the monopole interaction [66] between the j and j orbits, and ψ N ^ j ψ is the occupancy number of particles in the j orbit. The monopole interactions are derived from the average of the diagonal TBMEs. In detail, V j j for a given isospin T is defined as:
V j j , T = J 1 1 2 j J T + 1 δ j j 2 J + 1 2 j + 1 ) ( 2 j + 1 j j V j j J T
with J as the angular momentum and δ as Kronecker’s delta. When valence particles are added to orbits, the monopole interactions will change the SPEs.
To illustrate the shell evolution due to the addition of neutrons, Figure 4 shows the ESPEs in the ground states of Ne isotopes (the results of F isotopes are fairly similar), derived using the three interactions in the (0–5)ħω model space. It can be deduced that the three interactions give different interpretations of the shell structure and evolution. Firstly, the SDPF-M interaction indicates a rather strong mixing of the three orbits ν0d3/2, ν1p3/2, and ν0f7/2, while ν1p3/2, ν1p1/2, and ν0f7/2 orbits are strongly mixed in the other two interactions. With the increase in mass number, the relative position of ν0f7/2 (ν1p3/2) to ν1p1/2 varies more slowly in SDPF-MU than in SDPF-U-SI. Such differences in shell structure suggest the necessity of improvements in the cross-shell interactions of at least two among the three interactions.
For a given nucleus, the shell evolution can also result from the excitation of valence particles [35,83], which is called the type Ⅱ shell evolution [83]. We employed monopole-frozen analysis to examine the effects of such shell evolution, as described by the three interactions. In detail, we substituted ESPEs for SPEs, removed the monopole interaction from the diagonal TBMEs, and performed CISM calculations for 29F and 30Ne in the (0–5)ħω model space. If the type Ⅱ shell evolution is not significant in the two nuclei, their excitation energies should not be significantly changed by the monopole-frozen calculations.
Figure 5 presents the energy spectra of the two nuclei derived before and after the monopole-frozen modification. Despite the distinct energy spectra, it can be seen that the monopole-frozen modification generates different effects on the selected states for different interactions. For instance, positive yrast states of 29F and 30Ne from all three interactions change little with monopole-frozen calculations, which means that they share similar shell structures to the ground states. After monopole-frozen modification, the negative states of 30Ne from SDPF-MU and SDPF-U-SI interactions and the second 0+ state of 30Ne from SDPF-M have changed a lot. Therefore, the three interactions provide different illustrations of the excitation of the ground state for both 29F and 30Ne. Furthermore, a more reliable and effective interaction to describe N~20 nuclei with extreme proton–neutron asymmetry is required.
The ESPEs and monopole-frozen effects of the three interactions provide various shell structures and evolutions for F and Ne nuclei near N = 20, indicating that they can hardly be consistent with one another. Therefore, the cross-shell parts of the three interactions are suggested to be improved in order to describe the neutron-rich nuclei with N~20 better.

4. Conclusions

In this work, three interactions—SDPF-M, SDPF-MU, and SDPF-U-SI—were applied to calculate the structure of neutron-rich F and Ne isotopes within the framework of the configuration–interaction shell model. This paper focused on the cross-shell excitation of F and Ne nuclei near N = 20, which are among the systems with the most proton–neutron asymmetry. As a result, the cross-shell excitation was shown to be necessary for the overall description of F and Ne isotopes around N = 20. For 29F and 30Ne, it is important to include the cross-shell excitation of (0–5)ħω. Considering the insufficiency of the three interactions in the (0–5)ħω model space to reproduce properties such as the bound structure of 29,31Ne, it is also suggested to optimize the three interactions in order to better describe the cross-shell excitation near 29F and 30Ne in future works. This conclusion is supported by the investigation of ESPEs and a monopole-frozen analysis, which indicate the different illustrations of the three interactions on the shell structure and evolution in the target nuclear region.

Author Contributions

Conceptualization, C.Y.; methodology, C.Y. and M.L.; investigation, C.Y. and M.L.; data curation, C.Y. and M.L.; writing—original draft preparation, M.L.; writing—review and editing, C.Y. and M.L.; supervision, C.Y.; project administration, C.Y.; funding acquisition, C.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China under Grant No. 11775316 and 11305272, the computational resources from SYSU, the National Supercomputer Center in Guangzhou, and the Key Laboratory of High-Precision Nuclear Spectroscopy, Institute of Modern Physics, Chinese Academy of Sciences.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors appreciate fruitful discussions from Takaharu Otsuka, Noritaka Shimizu, and Xinxing Xu.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Haxel, O.; Jensen, J.H.D.; Suess, H.E. On the “magic numbers” in nuclear structure. Phys. Rev. 1949, 75, 1766. [Google Scholar] [CrossRef]
  2. Mayer, M.G. On closed shells in nuclei. II. Phys. Rev. 1949, 75, 1969–1970. [Google Scholar] [CrossRef]
  3. Revel, A.; Sorlin, O.; Marques, F.M.; Kondo, Y.; Kahlbow, J.; Nakamura, T.; Orr, N.A.; Nowacki, F.; Tostevin, J.A.; Yuan, C.X.; et al. Extending the southern shore of the island of inversion to 28F. Phys. Rev. Lett. 2020, 124, 152502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Terry, J.R.; Bazin, D.; Brown, B.A.; Campbell, C.M.; Church, J.A.; Cook, J.M.; Davies, A.D.; Dinca, D.C.; Enders, J.; Gade, A.; et al. Direct evidence for the onset of intruder configurations in neutron-rich Ne isotopes. Phys. Lett. B 2006, 640, 86–90. [Google Scholar] [CrossRef] [Green Version]
  5. Dombradi, Z.; Elekes, Z.; Saito, A.; Aoi, N.; Baba, H.; Demichi, K.; Fulop, Z.; Gibelin, J.; Gomi, T.; Hasegawa, H.; et al. Vanishing N = 20 shell gap: Study of excited states in 27,28Ne. Phys. Rev. Lett. 2006, 96, 182501. [Google Scholar] [CrossRef]
  6. Liu, H.; Lee, J.; Doornenbal, P.; Scheit, H.; Takeuchi, S.; Aoi, N.; Li, K.; Matsushita, M.; Steppenbeck, D.; Wang, H.; et al. Single-neutron knockout reaction from 30Ne. JPS Conf. Proc. 2015, 6, 030003. [Google Scholar]
  7. Thibault, C.; Klapisch, R.; Rigaud, C.; Poskanzer, A.M.; Prieels, R.; Lessard, L.; Reisdorf, W. Direct measurement of the masses of 11Li and 26-32Na with an on-line mass spectrometer. Phys. Rev. C 1975, 12, 644–657. [Google Scholar] [CrossRef] [Green Version]
  8. Terry, J.R.; Brown, B.A.; Campbell, C.M.; Cook, J.M.; Davies, A.D.; Dinca, D.C.; Gade, A.; Glasmacher, T.; Hansen, P.G.; Sherrill, B.M.; et al. Single-neutron knockout from intermediate energy beams of 30,32Mg: Mapping the transition into the “island of inversion”. Phys. Rev. C 2008, 77, 014316. [Google Scholar] [CrossRef] [Green Version]
  9. Guillemaud-Mueller, D.; Jacmart, J.C.; Kashy, E.; Latimier, A.; Mueller, A.C.; Pougheon, F.; Richard, A.; Penionzhkevich, Y.E.; Artuhk, A.G.; Belozyorov, A.V.; et al. Particle stability of the isotopes 26O and 32Ne in the reaction 44 MeV/nucleon 48Ca+Ta. Phys. Rev. C Nucl. Phys. 1990, 41, 937–941. [Google Scholar] [CrossRef]
  10. Tarasov, O.; Allatt, R.; Angélique, J.C.; Anne, R.; Borcea, C.; Dlouhy, Z.; Donzaud, C.; Grévy, S.; Guillemaud-Mueller, D.; Lewitowicz, M.; et al. Search for 28O and study of neutron-rich nuclei near the N = 20 shell closure. Phys. Lett. B 1997, 409, 64–70. [Google Scholar] [CrossRef] [Green Version]
  11. Thoennessen, M. Reaching the limits of nuclear stability. Rep. Prog. Phys. 2004, 67, 1187–1232. [Google Scholar] [CrossRef] [Green Version]
  12. Ahn, D.S.; Fukuda, N.; Geissel, H.; Inabe, N.; Iwasa, N.; Kubo, T.; Kusaka, K.; Morrissey, D.J.; Murai, D.; Nakamura, T.; et al. Location of the neutron dripline at fluorine and neon. Phys. Rev. Lett. 2019, 123, 212501. [Google Scholar] [CrossRef]
  13. Sakurai, H.; Lukyanov, S.M.; Notani, M.; Aoi, N.; Beaumel, D.; Fukuda, N.; Hirai, M.; Ideguchi, E.; Imai, N.; Ishihara, M.; et al. Evidence for particle stability of F and particle instability of N and O. Phys. Lett. B 1999, 448, 180–184. [Google Scholar] [CrossRef]
  14. Wang, M.; Huang, W.J.; Kondev, F.G.; Audi, G.; Naimi, S. The AME 2020 atomic mass evaluation (II). Tables, graphs and references*. Chin. Phys. C 2021, 45, 030003. [Google Scholar] [CrossRef]
  15. Kobayashi, N.; Nakamura, T.; Kondo, Y.; Tostevin, J.A.; Aoi, N.; Baba, H.; Barthelemy, R.; Famiano, M.A.; Fukuda, N.; Inabe, N.; et al. One-neutron removal from 29Ne: Defining the lower limits of the island of inversion. Phys. Rev. C 2016, 93, 014613. [Google Scholar] [CrossRef] [Green Version]
  16. Nakamura, T.; Kobayashi, N.; Kondo, Y.; Satou, Y.; Aoi, N.; Baba, H.; Deguchi, S.; Fukuda, N.; Gibelin, J.; Inabe, N.; et al. Halo structure of the island of inversion nucleus 31Ne. Phys. Rev. Lett. 2009, 103, 262501. [Google Scholar] [CrossRef]
  17. Gaudefroy, L.; Mittig, W.; Orr, N.A.; Varet, S.; Chartier, M.; Roussel-Chomaz, P.; Ebran, J.P.; Fernandez-Dominguez, B.; Fremont, G.; Gangnant, P.; et al. Direct mass measurements of 19B, 22C, 29F, 31Ne, 34Na and other light exotic nuclei. Phys. Rev. Lett. 2012, 109, 202503. [Google Scholar] [CrossRef] [Green Version]
  18. Bagchi, S.; Kanungo, R.; Tanaka, Y.K.; Geissel, H.; Doornenbal, P.; Horiuchi, W.; Hagen, G.; Suzuki, T.; Tsunoda, N.; Ahn, D.S.; et al. Two-Neutron Halo is Unveiled in 29F. Phys. Rev. Lett. 2020, 124, 222504. [Google Scholar] [CrossRef]
  19. Chrisman, D.; Kuchera, A.N.; Baumann, T.; Blake, A.; Brown, B.A.; Brown, J.; Cochran, C.; DeYoung, P.A.; Finck, J.E.; Frank, N.; et al. Neutron-unbound states in 31Ne. Phys. Rev. C 2021, 104, 034313. [Google Scholar] [CrossRef]
  20. Jansen, G.R.; Engel, J.; Hagen, G.; Navratil, P.; Signoracci, A. Ab initio coupled-cluster effective interactions for the shell model: Application to neutron-rich oxygen and carbon isotopes. Phys. Rev. Lett. 2014, 113, 142502. [Google Scholar] [CrossRef] [Green Version]
  21. Sun, Z.; Wu, Q.; Xu, F. Green’s function calculations of light nuclei. Sci. China Phys. Mech. 2016, 59, 692013. [Google Scholar] [CrossRef]
  22. Bender, M.; Heenen, P.-H.; Reinhard, P.-G. Self-consistent mean-field models for nuclear structure. Rev. Mod. Phys. 2003, 75, 121–180. [Google Scholar] [CrossRef]
  23. Otsuka, T.; Honma, M.; Mizusaki, T.; Shimizu, N.; Utsuno, Y. Monte Carlo shell model for atomic nuclei. Prog. Part. Nucl. Phys. 2001, 47, 319–400. [Google Scholar] [CrossRef]
  24. Miyagi, T.; Stroberg, S.R.; Holt, J.D.; Shimizu, N. Ab initio multishell valence-space Hamiltonians and the island of inversion. Phys. Rev. C 2020, 102, 034320. [Google Scholar] [CrossRef]
  25. Luo, Y.-X.; Fossez, K.; Liu, Q.; Guo, J.-Y. Role of quadrupole deformation and continuum effects in the “island of inversion” nuclei F28,29,31. Phys. Rev. C 2021, 104, 014307. [Google Scholar] [CrossRef]
  26. Tripathi, V.; Tabor, S.L.; Mantica, P.F.; Utsuno, Y.; Bender, P.; Cook, J.; Hoffman, C.R.; Lee, S.; Otsuka, T.; Pereira, J.; et al. Competition between normal and intruder states inside the “island of inversion”. Phys. Rev. C 2007, 76, 021301. [Google Scholar] [CrossRef] [Green Version]
  27. Michel, N.; Li, J.G.; Xu, F.R.; Zuo, W. Two-neutron halo structure of 31F. Phys. Rev. C 2020, 101, 031301. [Google Scholar] [CrossRef] [Green Version]
  28. Masui, H.; Horiuchi, W.; Kimura, M. Two-neutron halo structure of 31F and a novel pairing antihalo effect. Phys. Rev. C 2020, 101, 041303. [Google Scholar] [CrossRef] [Green Version]
  29. Singh, J.; Casal, J.; Horiuchi, W.; Fortunato, L.; Vitturi, A. Exploring two-neutron halo formation in the ground state of 29F within a three-body model. Phys. Rev. C 2020, 101, 024310. [Google Scholar] [CrossRef] [Green Version]
  30. Lee, J.; Xu, X.X.; Kaneko, K.; Sun, Y.; Lin, C.J.; Sun, L.J.; Liang, P.F.; Li, Z.H.; Li, J.; Wu, H.Y.; et al. Large isospin asymmetry in 22Si/22O mirror Gamow-Teller transitions reveals the halo structure of 22Al. Phys. Rev. Lett. 2020, 125, 192503. [Google Scholar] [CrossRef]
  31. Liang, P.F.; Sun, L.J.; Lee, J.; Hou, S.Q.; Xu, X.X.; Lin, C.J.; Yuan, C.X.; He, J.J.; Li, Z.H.; Wang, J.S.; et al. Simultaneous measurement of β-delayed proton and γ emission of 26P for the 25Al(p,γ)26Si reaction rate. Phys. Rev. C 2020, 101, 024305. [Google Scholar] [CrossRef] [Green Version]
  32. Sun, L.J.; Xu, X.X.; Hou, S.Q.; Lin, C.J.; José, J.; Lee, J.; He, J.J.; Li, Z.H.; Wang, J.S.; Yuan, C.X.; et al. Experimentally well-constrained masses of 27P and 27S: Implications for studies of explosive binary systems. Phys. Lett. B 2020, 802, 135213. [Google Scholar] [CrossRef]
  33. Shi, G.Z.; Liu, J.J.; Lin, Z.Y.; Zhu, H.F.; Xu, X.X.; Sun, L.J.; Liang, P.F.; Lin, C.J.; Lee, J.; Yuan, C.X.; et al. β-delayed two-proton decay of 27S at the proton-drip line. Phys. Rev. C 2021, 103, L061301. [Google Scholar] [CrossRef]
  34. Chen, Z.Q.; Li, Z.H.; Hua, H.; Watanabe, H.; Yuan, C.X.; Zhang, S.Q.; Lorusso, G.; Nishimura, S.; Baba, H.; Browne, F.; et al. Proton shell evolution below 132Sn: First measurement of low-lying beta-emitting isomers in 123,125Ag. Phys. Rev. Lett. 2019, 122, 212502. [Google Scholar] [CrossRef] [Green Version]
  35. Xu, X.; Liu, J.H.; Yuan, C.X.; Xing, Y.M.; Wang, M.; Zhang, Y.H.; Zhou, X.H.; Litvinov, Y.A.; Blaum, K.; Chen, R.J.; et al. Masses of ground and isomeric states of 101In and configuration-dependent shell evolution in odd-A indium isotopes. Phys. Rev. C 2019, 100, 051303. [Google Scholar] [CrossRef] [Green Version]
  36. Yuan, C.X.; Liu, Z.; Xu, F.R.; Walker, P.M.; Podolyák, Z.; Xu, C.; Ren, Z.Z.; Ding, B.; Liu, M.L.; Liu, X.Y.; et al. Isomerism in the “south-east” of 132Sn and a predicted neutron-decaying isomer in 129Pd. Phys. Lett. B 2016, 762, 237–242. [Google Scholar] [CrossRef]
  37. Phong, V.H.; Lorusso, G.; Davinson, T.; Estrade, A.; Hall, O.; Liu, J.; Matsui, K.; Montes, F.; Nishimura, S.; Boso, A.; et al. Observation of a μs isomer in 134In: Proton-neutron coupling “southeast” of 132Sn. Phys. Rev. C 2019, 100, 011302. [Google Scholar] [CrossRef]
  38. Li, C.B.; Zhang, G.L.; Yuan, C.X.; Zhang, G.X.; Hu, S.P.; Qu, W.W.; Zheng, Y.; Zhang, H.Q.; Mengoni, D.; Testov, D.; et al. New level scheme and shell model description of 212Rn. Phys. Rev. C 2020, 101, 044313. [Google Scholar] [CrossRef]
  39. Zhang, M.M.; Yang, H.B.; Gan, Z.G.; Zhang, Z.Y.; Huang, M.H.; Ma, L.; Yang, C.L.; Yuan, C.X.; Wang, Y.S.; Tian, Y.L.; et al. A new isomeric state in 218Pa. Phys. Lett. B 2020, 800, 135102. [Google Scholar] [CrossRef]
  40. Zhou, H.B.; Gan, Z.G.; Wang, N.; Yang, H.B.; Ma, L.; Huang, M.H.; Yang, C.L.; Zhang, M.M.; Tian, Y.L.; Wang, Y.S.; et al. Lifetime measurement for the isomeric state in 213Th. Phys. Rev. C 2021, 103, 044314. [Google Scholar] [CrossRef]
  41. Cai, B.; Chen, G.; Xu, J.; Yuan, C.; Qi, C.; Yao, Y. α decay half-life estimation and uncertainty analysis. Phys. Rev. C 2020, 101, 054304. [Google Scholar] [CrossRef]
  42. Zhang, Z.Y.; Yang, H.B.; Huang, M.H.; Gan, Z.G.; Yuan, C.X.; Qi, C.; Andreyev, A.N.; Liu, M.L.; Ma, L.; Zhang, M.M.; et al. New alpha-emitting isotope 214U and abnormal enhancement of alpha-particle clustering in lightest uranium isotopes. Phys. Rev. Lett. 2021, 126, 152502. [Google Scholar] [CrossRef]
  43. Fukunishi, N.; Otsuka, T.; Sebe, T. Vanishing of the shell gap in N = 20 neutron-rich nuclei. Phys. Lett. B 1992, 296, 279–284. [Google Scholar] [CrossRef]
  44. Utsuno, Y.; Otsuka, T.; Mizusaki, T.; Honma, M. Varying shell gap and deformation in N~20 unstable nuclei studied by the Monte Carlo shell model. Phys. Rev. C 1999, 60, 054315. [Google Scholar] [CrossRef]
  45. Caurier, E.; Nowacki, F.; Poves, A. Merging of the islands of inversion at N = 20 and N = 28. Phys. Rev. C 2014, 90, 014302. [Google Scholar] [CrossRef] [Green Version]
  46. Chen, J.; Auranen, K.; Avila, M.L.; Back, B.B.; Caprio, M.A.; Hoffman, C.R.; Gorelov, D.; Kay, B.P.; Kuvin, S.A.; Liu, Q.; et al. Experimental study of the low-lying negative-parity states in 11Be using the 12B(d,3He)11Be reaction. Phys. Rev. C 2019, 100, 064314. [Google Scholar] [CrossRef]
  47. Chen, J.; Wang, S.M.; Fortune, H.T.; Lou, J.L.; Ye, Y.L.; Li, Z.H.; Michel, N.; Li, J.G.; Yuan, C.X.; Ge, Y.C.; et al. Observation of the near-threshold intruder 0 resonance in 12Be. Phys. Rev. C 2021, 103, L031302. [Google Scholar] [CrossRef]
  48. Yang, Z.H.; Kubota, Y.; Corsi, A.; Yoshida, K.; Sun, X.X.; Li, J.G.; Kimura, M.; Michel, N.; Ogata, K.; Yuan, C.X.; et al. Quasifree neutron knockout reaction reveals a small s-orbital component in the Borromean nucleus 17B. Phys. Rev. Lett. 2021, 126, 082501. [Google Scholar] [CrossRef]
  49. Yuan, C.X. Impact of off-diagonal cross-shell interaction on 14C. Chin. Phys. C 2017, 41, 104102. [Google Scholar] [CrossRef] [Green Version]
  50. Jiang, Y.; Lou, J.L.; Ye, Y.L.; Liu, Y.; Tan, Z.W.; Liu, W.; Yang, B.; Tao, L.C.; Ma, K.; Li, Z.H.; et al. Quadrupole deformation of 16C studied by proton and deuteron inelastic scattering. Phys. Rev. C 2020, 101, 024601. [Google Scholar] [CrossRef]
  51. Yuan, C.X.; Suzuki, T.; Otsuka, T.; Xu, F.; Tsunoda, N. Shell-model study of boron, carbon, nitrogen, and oxygen isotopes with a monopole-based universal interaction. Phys. Rev. C 2012, 85, 064324. [Google Scholar] [CrossRef] [Green Version]
  52. Wang, Y.B.; Su, J.; Han, Z.Y.; Tang, B.; Cui, B.Q.; Ge, T.; Lyu, Y.L.; Brown, B.A.; Yuan, C.X.; Chen, L.H. Direct observation of the exotic β−γ−α decay mode in the Tz = −1 nucleus 20Na. Phys. Rev. C 2021, 103, L011301. [Google Scholar] [CrossRef]
  53. Utsuno, Y.; Otsuka, T.; Brown, B.A.; Honma, M.; Mizusaki, T.; Shimizu, N. Shape transitions in exotic Si and S isotopes and tensor-force-driven Jahn-Teller effect. Phys. Rev. C 2012, 86, 051301. [Google Scholar] [CrossRef] [Green Version]
  54. Nowacki, F.; Poves, A. New effective interaction for 0ℏω shell-model calculations in the sd−pf valence space. Phys. Rev. C 2009, 79, 014310. [Google Scholar] [CrossRef] [Green Version]
  55. Bazin, D.; Aoi, N.; Baba, H.; Chen, J.; Crawford, H.; Doornenbal, P.; Fallon, P.; Li, K.; Lee, J.; Matsushita, M.; et al. Spectroscopy of 33Mg with knockout reactions. Phys. Rev. C 2021, 103, 064318. [Google Scholar] [CrossRef]
  56. Gade, A.; Brown, B.A.; Tostevin, J.A.; Bazin, D.; Bender, P.C.; Campbell, C.M.; Crawford, H.L.; Elman, B.; Kemper, K.W.; Longfellow, B.; et al. Is the structure of 42Si understood? Phys. Rev. Lett. 2019, 122, 222501. [Google Scholar] [CrossRef] [Green Version]
  57. Hjorth-Jensen, M.; Kuo, T.T.S.; Osnes, E. Realistic effective interactions for nuclear systems. Phys. Rep. 1995, 261, 125–270. [Google Scholar] [CrossRef]
  58. Brown, B.A. The nuclear shell model towards the drip lines. Prog. Part. Nucl. Phys. 2001, 47, 517–599. [Google Scholar] [CrossRef]
  59. Noritaka, S. Nuclear shell-model code for massive parallel computation, “KSHELL”. arXiv 2013, arXiv:1310.5431. [Google Scholar]
  60. Noritaka, S.; Takahiro, M.; Yutaka, U.; Yusuke, T. Thick-restart block Lanczos method for large-scale shell-model calculations. Comput. Phys. Commun. 2019, 244, 372–384. [Google Scholar]
  61. Available online: https://sites.google.com/a/cns.s.u-tokyo.ac.jp/kshell/ (accessed on 1 November 2021).
  62. Brown, B.A.; Richter, W.A.; Julies, R.E.; Wildenthal, B.H. Semi-empirical effective interactions for the 1s-0d shell. Ann. Phys. 1988, 182, 191–236. [Google Scholar] [CrossRef]
  63. Kuo, T.T.S.; Brown, G.E. Reaction matrix elements for the 0f-1p shell nuclei. Nucl. Phys. A 1968, 114, 241–279. [Google Scholar] [CrossRef]
  64. Millener, D.J.; Kurath, D. The particle-hole interaction and the beta decay of 14B. Nucl. Phys. A 1975, 255, 315–338. [Google Scholar] [CrossRef]
  65. Warburton, E.K.; Alburger, D.E.; Becker, J.A.; Brown, B.A.; Raman, S. Probe of the shell crossing at A=40 via beta decay: Experiment and theory. Phys. Rev. C 1986, 34, 1031–1051. [Google Scholar] [CrossRef] [Green Version]
  66. Bansal, R.K.; French, J.B. Even-parity-hole states in f7/2 -shell nuclei. Phys. Lett. 1964, 11, 145–148. [Google Scholar] [CrossRef]
  67. Doornenbal, P.; Scheit, H.; Takeuchi, S.; Utsuno, Y.; Aoi, N.; Li, K.; Matsushita, M.; Steppenbeck, D.; Wang, H.; Baba, H.; et al. Low-Z shore of the “island of inversion” and the reduced neutron magicity toward 28O. Phys. Rev. C 2017, 95, 041301. [Google Scholar] [CrossRef] [Green Version]
  68. Utsuno, Y.; Otsuka, T.; Glasmacher, T.; Mizusaki, T.; Honma, M. Onset of intruder ground state in exotic Na isotopes and evolution of the N = 20 shell gap. Phys. Rev. C 2004, 70, 044307. [Google Scholar] [CrossRef] [Green Version]
  69. Han, R.; Li, X.Q.; Jiang, W.G.; Li, Z.H.; Hua, H.; Zhang, S.Q.; Yuan, C.X.; Jiang, D.X.; Ye, Y.L.; Li, J.; et al. Northern boundary of the “island of inversion” and triaxiality in 34Si. Phys. Lett. B 2017, 772, 529–533. [Google Scholar] [CrossRef]
  70. Honma, M.; Otsuka, T.; Mizusaki, T.; Hjorthjensen, M. Shell-model description of beta-decays for pfg-shell nuclei. Riken Acce. Prog. Rep. 2009, 42, 43. [Google Scholar]
  71. Otsuka, T.; Suzuki, T.; Honma, M.; Utsuno, Y.; Tsunoda, N.; Tsukiyama, K.; Hjorth-Jensen, M. Novel features of nuclear forces and shell evolution in exotic nuclei. Phys. Rev. Lett. 2010, 104, 012501. [Google Scholar] [CrossRef] [Green Version]
  72. Poves, A.; Zuker, A. Theoretical spectroscopy and the fp shell. Phys. Rep. 1981, 70, 235–314. [Google Scholar] [CrossRef]
  73. Bertsch, G.; Borysowicz, J.; McManus, H.; Love, W.G. Interactions for inelastic scattering derived from realistic potentials. Nucl. Phys. A 1977, 284, 399–419. [Google Scholar] [CrossRef]
  74. Nummela, S.; Baumann, P.; Caurier, E.; Dessagne, P.; Jokinen, A.; Knipper, A.; Le Scornet, G.; Miehé, C.; Nowacki, F.; Oinonen, M.; et al. Spectroscopy of 34,35Si by β decay: Sd−fp shell gap and single-particle states. Phys. Rev. C 2001, 63, 044316. [Google Scholar] [CrossRef] [Green Version]
  75. Retamosa, J.; Caurier, E.; Nowacki, F.; Poves, A. Shell model study of the neutron-rich nuclei around N = 28. Phys. Rev. C 1997, 55, 1266–1274. [Google Scholar] [CrossRef] [Green Version]
  76. Kahana, S.; Lee, H.C.; Scott, C.K. Effect of Woods-Saxon wave functions on the calculation of A=18, 206, 210 spectra with a realistic interaction. Phys. Rev. 1969, 180, 956–966. [Google Scholar] [CrossRef]
  77. Gloeckner, D.H.; Lawson, R.D. Spurious center-of-mass motion. Phys. Lett. B 1974, 53, 313–318. [Google Scholar] [CrossRef]
  78. Kondev, F.G.; Wang, M.; Huang, W.J.; Naimi, S.; Audi, G. The NUBASE2020 evaluation of nuclear physics properties. Chin. Phys. C 2021, 45, 030001. [Google Scholar] [CrossRef]
  79. Available online: https://www.nndc.bnl.gov/nudat2/ (accessed on 1 November 2021).
  80. Umeya, A.; Muto, K. Single-particle energies in neutron-rich nuclei by shell model sum rule. Phys. Rev. C 2006, 74, 034330. [Google Scholar] [CrossRef]
  81. Yuan, C.X.; Qi, C.; Xu, F.R. Shell evolution in neutron-rich carbon isotopes: Unexpected enhanced role of neutron–neutron correlation. Nucl. Phys. A 2012, 883, 25–34. [Google Scholar] [CrossRef] [Green Version]
  82. Otsuka, T.; Tsunoda, Y.; Abe, T.; Shimizu, N.; Van Duppen, P. Underlying Structure of Collective Bands and Self-Organization in Quantum Systems. Phys. Rev. Lett. 2019, 123, 222502. [Google Scholar] [CrossRef] [Green Version]
  83. Yuan, C.X.; Liu, M.L.; Ge, Y.L. Shell-model explanation on some newly discovered isomers. Nucl. Phys. R 2020, 37, 447–454. [Google Scholar]
Figure 1. RMSEs of theoretical Sn (a) and S2n (b) values of 2729F and 2831Ne.
Figure 1. RMSEs of theoretical Sn (a) and S2n (b) values of 2729F and 2831Ne.
Symmetry 13 02167 g001
Figure 2. (a) Sn values of 29F, (b) S2n values of 29F, (c) Sn values of 30Ne and (d) S2n values of 30Ne, as calculated by CISM (experimental results are from [14]).
Figure 2. (a) Sn values of 29F, (b) S2n values of 29F, (c) Sn values of 30Ne and (d) S2n values of 30Ne, as calculated by CISM (experimental results are from [14]).
Symmetry 13 02167 g002
Figure 3. The energy levels of (a) the 1/21+ state in 29F and (b) the 21+ state in 30Ne, calculated by the CISM (experimental data are from [67] (a) and Nudat2 (b)).
Figure 3. The energy levels of (a) the 1/21+ state in 29F and (b) the 21+ state in 30Ne, calculated by the CISM (experimental data are from [67] (a) and Nudat2 (b)).
Symmetry 13 02167 g003
Figure 4. ESPEs in the ground state of Ne isotopes calculated in the (0–5)ħω model space, with SDPF-M (a), SDPF-MU (b), and SDPF-U-SI (c).
Figure 4. ESPEs in the ground state of Ne isotopes calculated in the (0–5)ħω model space, with SDPF-M (a), SDPF-MU (b), and SDPF-U-SI (c).
Symmetry 13 02167 g004
Figure 5. Comparison of energy spectra of (a) 29F and (b) 30Ne. All CISM calculations were performed in the (0–5)ħω model space. The symbol * represents where the interaction is modified via frozen monopole.
Figure 5. Comparison of energy spectra of (a) 29F and (b) 30Ne. All CISM calculations were performed in the (0–5)ħω model space. The symbol * represents where the interaction is modified via frozen monopole.
Symmetry 13 02167 g005
Table 1. The Sn and S2n values (in MeV) of 27–33F and 28–36Ne nuclei calculated by CISM, in comparison with experimental data from the Atomic Mass Evaluation Table 2020 [14].
Table 1. The Sn and S2n values (in MeV) of 27–33F and 28–36Ne nuclei calculated by CISM, in comparison with experimental data from the Atomic Mass Evaluation Table 2020 [14].
Exp. (Error)SDPF-M
(0–1)ħω
SDPF-M
(0–3)ħω
SDPF-M
(0–5)ħω
SDPF-MU
(0–1)ħω
SDPF-MU
(0–3)ħω
SDPF-MU
(0–5)ħω
SDPF-U-SI
(0–1)ħω
SDPF-U-SI
(0–3)ħω
SDPF-U-SI
(0–5)ħω
Sn (27F)1.610 (0.060)0.8862.4372.6591.1143.2003.9342.4023.0273.154
Sn (28F)−0.199 (0.006)0.157−0.479−0.5840.003−0.796−1.113−0.090−0.541−0.622
Sn (29F)1.320 (0.540)−1.7620.1161.498−0.9800.9082.6190.1521.4652.016
Sn (30F)/0.830−0.799−1.2171.075−1.005−1.7740.240−0.362−0.652
Sn (31F)/−0.5020.9991.0050.1171.4911.8451.3251.2321.200
Sn (32F)/0.178−1.423−1.787−0.222−2.015−2.689−0.095−0.860−1.077
Sn (33F)/−0.8510.5350.564−0.1400.8951.0670.5370.6360.675
Sn (28Ne)3.820 (0.160)2.6074.4774.8242.5005.3286.3594.0864.8615.060
Sn (29Ne)0.970 (0.200)0.9080.005−0.0911.1110.137−0.4100.5920.059−0.081
Sn (30Ne)3.190 (0.290)−0.7712.2253.562−0.2741.7573.9871.0322.5323.663
Sn (31Ne)0.170 (0.130)2.8640.304−0.4743.6570.998−1.1792.4410.681−0.478
Sn (32Ne)/0.5202.2472.3521.2472.6163.7632.2392.9903.319
Sn (33Ne)/1.495−0.969−1.3341.598−1.216−2.2240.723−0.127−0.446
Sn (34Ne)/−0.0851.4831.5400.9702.3152.4682.2642.1732.186
Sn (35Ne)/0.067−1.629−1.902−1.514−2.470−3.016−0.512−0.943−1.032
Sn (36Ne)/−0.1101.1561.1430.7211.2111.3270.7980.8220.830
S2n (27F)2.340 (0.150)1.2762.7462.9752.0384.3075.0793.3234.2174.353
S2n (28F)1.410 (0.060)1.0431.9582.0751.1172.4042.8212.3122.4862.532
S2n (29F)1.130 (0.540)−1.605−0.3630.914−0.9770.1121.5060.0620.9241.394
S2n (30F)/−0.932−0.6830.2810.095−0.0970.8450.3921.1031.364
S2n (31F)/0.3280.200−0.2121.1920.4860.0711.5650.8700.548
S2n (32F)/−0.324−0.424−0.782−0.105−0.524−0.8441.2300.3720.123
S2n (33F)/−0.673−0.888−1.223−0.362−1.120−1.6220.442−0.224−0.402
S2n (28Ne)5.320 (0.130)3.7665.3675.7194.4156.4047.3785.3826.1836.375
S2n (29Ne)4.790 (0.170)3.5154.4824.7333.6115.4655.9494.6784.9204.979
S2n (30Ne)4.160 (0.280)0.1372.2303.4710.8371.8943.5771.6242.5913.582
S2n (31Ne)3.360 (0.310)2.0932.5293.0883.3832.7552.8083.4733.2133.185
S2n (32Ne)/3.3842.5511.8784.9043.6142.5844.6803.6712.841
S2n (33Ne)/2.0151.2781.0182.8451.4001.5392.9622.8632.873
S2n (34Ne)/1.4100.5140.2062.5681.0990.2442.9872.0461.740
S2n (35Ne)/−0.018−0.146−0.362−0.544−0.155−0.5481.7521.2301.154
S2n (36Ne)/−0.043−0.473−0.759−0.793−1.259−1.6890.286−0.121−0.202
Table 2. The experimental and theoretical ground-state spin of 27–31F and 28–34Ne (experimental results are from [78]).
Table 2. The experimental and theoretical ground-state spin of 27–31F and 28–34Ne (experimental results are from [78]).
NucleusExp.SDPF-M
(0–1)ħω
SDPF-M
(0–3)ħω
SDPF-M
(0–5)ħω
SDPF-MU
(0–1)ħω
SDPF-MU
(0–3)ħω
SDPF-MU
(0–5)ħω
SDPF-U-SI
(0–1)ħω
SDPF-U-SI
(0–3)ħω
SDPF-U-SI
(0–5)ħω
27F5/2+5/2+5/2+5/2+5/2+5/2+5/2+5/2+5/2+5/2+
28F4644444444
29F5/2+5/2+5/2+5/2+5/2+5/2+5/2+5/2+5/2+5/2+
30F/3+443+3+43+44
31F/5/2+5/2+5/2+5/2+5/2+5/2+5/2+5/2+5/2+
28Ne0+0+0+0+0+0+0+0+0+0+
29Ne3/27/27/23/2+3/23/23/23/23/23/2
30Ne0+0+0+0+0+0+0+0+0+0+
31Ne3/23/2+3/23/23/2+3/2+3/23/2+3/2+3/2+
32Ne0+30+0+0+0+0+0+0+0+
33Ne/3/2+3/23/2-3/2+3/2+3/23/23/23/2
34Ne/0+0+0+0+0+0+0+0+0+
Table 3. The experimental and theoretical energy levels (in MeV) of the excited state in F and Ne nuclei (the experimental results are from [79], except those for 27,29F, which are from [67]).
Table 3. The experimental and theoretical energy levels (in MeV) of the excited state in F and Ne nuclei (the experimental results are from [79], except those for 27,29F, which are from [67]).
NucleusStateExp.SDPF-M
(0–1)ħω
SDPF-M
(0–3)ħω
SDPF-M
(0–5)ħω
SDPF-MU
(0–1)ħω
SDPF-MU
(0–3)ħω
SDPF-MU
(0–5)ħω
SDPF-U-SI
(0–1)ħω
SDPF-U-SI
(0–3)ħω
SDPF-U-SI
(0–5)ħω
27F1/21+0.915 (0.012)1.9801.2971.1842.1761.9521.4821.9231.7701.715
3/21+/2.6262.7322.6752.8473.5713.4522.8723.2183.293
29F1/21+1.080 (0.018)3.5530.9520.7973.6692.8331.5152.7362.1161.294
3/21+/6.4072.2462.2946.3495.0473.4426.6603.8292.816
31F1/21+/0.6480.5420.6280.3110.6800.7960.1760.3400.390
3/21+/1.8681.7012.0191.8212.3422.6181.4471.6811.764
28Ne21+1.304 (0.003)1.5531.4011.3621.8302.1102.1961.8041.9471.983
41+3.010 (0.006)2.9522.7702.7573.3023.5123.7793.3423.4583.510
30Ne21+0.792 (0.004)1.5310.8731.0281.9121.9971.8081.9011.7601.201
41+2.235 (0.012)2.4441.9452.3642.8172.9543.3592.8792.9352.535
32Ne21+0.722 (0.009)0.7680.6820.8410.8571.2621.4820.6590.8580.932
41+/1.7631.7782.0862.1202.7843.2531.7622.0922.241
34Ne21+/0.5210.8380.9690.6901.1101.4030.4760.6160.655
41+/1.3872.0282.2671.9222.6703.1581.4241.6791.747
Table 4. The average neutron occupancies of F and Ne nuclei in the pf shell with N~20.
Table 4. The average neutron occupancies of F and Ne nuclei in the pf shell with N~20.
NucleusState(0–1)ħω(0–3)ħω
SDPF-M
(0–3)ħω
SDPF-MU
(0–3)ħω
SDPF-U-SI
(0–5)ħω
SDPF-M
(0–5)ħω
SDPF-MU
(0–5)ħω
SDPF-U-SI
27F5/21+00.7950.5430.2360.9510.8770.300
29F5/21+01.4970.6460.4572.3541.6600.930
29F1/21+01.9721.0031.2262.7162.3021.967
29F3/21+01.9941.3561.9712.7302.3302.127
31F5/21+23.4132.6892.3464.0303.4522.519
33F5/21+45.3374.6694.2345.7455.2144.298
28Ne01+01.0900.5700.2611.3071.0890.377
30Ne01+01.8570.7000.6692.4791.9581.619
30Ne21+01.9650.6901.3302.5332.1922.053
30Ne41+01.9810.5891.0162.4041.9532.050
32Ne01+23.7052.6932.3854.0383.5152.660
34Ne01+45.3444.5664.2045.6545.1014.268
Table 5. Percentages (in %) of the configurations in the ground states.
Table 5. Percentages (in %) of the configurations in the ground states.
NucleusConfiguration(0–1)ħω(0–3)ħω
SDPF-M
(0–3)ħω
SDPF-MU
(0–3)ħω
SDPF-U-SI
(0–5)ħω
SDPF-M
(0–5)ħω
SDPF-MU
(0–5)ħω
SDPF-U-SI
29Fπ0d5/2ν(0d3/2)4124.9865.8176.248.5330.6556.14
π0d5/2ν(0d3/2)2 (0f7/2)2043.6513.839.5336.1620.5615.91
π0d5/2ν(0d3/2)2 (1p3/2)2015.414.422.9315.1414.287.92
30Neπ(0d5/2)2 ν(0d3/2)416.5557.8859.732.9817.4522.58
π(0d5/2)2ν(0d3/2)2 (0f7/2)2045.412.8811.0832.5317.9417.49
π(0d5/2)2ν(0d3/2)2 (1p3/2)2010.273.413.748.5610.3310.21
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, M.; Yuan, C. Cross-Shell Excitation in F and Ne Isotopes around N = 20. Symmetry 2021, 13, 2167. https://0-doi-org.brum.beds.ac.uk/10.3390/sym13112167

AMA Style

Liu M, Yuan C. Cross-Shell Excitation in F and Ne Isotopes around N = 20. Symmetry. 2021; 13(11):2167. https://0-doi-org.brum.beds.ac.uk/10.3390/sym13112167

Chicago/Turabian Style

Liu, Menglan, and Cenxi Yuan. 2021. "Cross-Shell Excitation in F and Ne Isotopes around N = 20" Symmetry 13, no. 11: 2167. https://0-doi-org.brum.beds.ac.uk/10.3390/sym13112167

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop