Mechanisms of mirror energy difference for states exhibiting Thomas-Ehrman shift: Gamow shell model case studies of 18Ne/18O and 19Na/19O

J.G. Li K. H. Li N. Michel H. H. Li W. Zuo CAS Key Laboratory of High Precision Nuclear Spectroscopy, Institute of Modern Physics, Chinese Academy of Sciences, Lanzhou 730000, China School of Nuclear Science and Technology, University of Chinese Academy of Sciences, Bei**g 100049, China Southern Center for Nuclear-Science Theory (SCNT), Institute of Modern Physics, Chinese Academy of Sciences, Huizhou 516000, Guangdong Province, China Institute of Particle and Nuclear Physics, Henan Normal University, Xinxiang 453007, China
Abstract

The mirror energy difference (MED) of the mirror state, especially for states bearing the Thomas-Erhman shift, serves as a sensitive probe of isospin symmetry breaking. We employ the Gamow shell model, which includes the inter-nucleon correlation and continuum coupling, to investigate the MED for sd𝑠𝑑sditalic_s italic_d-shell nuclei by taking the 18Ne/18O and 19Na/19O as examples. Our GSM provides good descriptions for the excitation energies and MEDs for the 18Ne/18O and 19Na/19O. Moreover, our calculations also reveal that the large MED of the mirror states is caused by the significant occupation of the weakly bound or unbound s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT waves, giving the radial density distribution of the state in the proton-rich nucleus more extended than that of mirror states in deeply-bound neutron-rich nuclei. Furthermore, our GSM calculation shows that the contribution of Coulomb is different for the low-lying states in proton-rich nuclei, which significantly contributes to MEDs of mirror states. Moreover, the contributions of the nucleon-nucleon interaction are different for the mirror state, especially for the state of proton-rich nuclei bearing the Thomas-Erhman shift, which also contributes to the significant isospin symmetry breaking with large MED.

keywords:
isospin symmetry breaking , Thomas-Ehrman shift , mirror energy difference , continuum coupling , Gamow shell model
journal: Physics Letters B

Introduction.  Exotic nuclear structures in drip-line nuclei have become a subject of great interest in recent years, as they are characterized by unique properties that exhibit significant differences compared to those of stable nuclei. One of the most significant phenomena observed in these systems is the Thomas-Ehrman shift (TES) [1, 2]. This effect is most pronounced in nuclei close to the proton drip lines, where the balance between the strong force and the Coulomb force is most delicate. States exhibiting TES effects are often weakly bound or unbound, characteristic of open quantum systems, while their neutron-rich mirror counterparts remain deeply bound, resulting in a large mirror energy difference (MED) in their isobaric states [1, 2, 3, 4, 5, 6]. The large MEDs are attributed to their proximity to near-threshold effects, in which the continuum effects need to be well treated. A thorough comprehension of the Thomas-Ehrman shift is pivotal for elucidating the dynamics of weakly bound and unbound nuclear systems and understanding the mechanisms underlying isospin symmetry breaking in mirror nuclei.

Two possible reasons exist for the states with large MED, of external or internal character. If extended single-particle wave functions of weakly- or unbound s𝑠sitalic_s- or p𝑝pitalic_p-waves are significantly occupied in the considered states, the large MED is of external nature, as in the TES states [7, 4, 8]. The second possibility, related to configuration mixing (see Refs.[9, 10]), is of internal nature. In this case, the extended wave function is given via the strong configuration mixing in which a few nodal states of s𝑠sitalic_s or p𝑝pitalic_p waves are included in the calculations. These two external and internal effects are different but can be intertwined in a complex manner. For instance, the inversion of ground states in the 16F and 16N mirror nuclei is primarily due to the unbound proton 1s1/21subscript𝑠121s_{1/2}1 italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT orbital, which can also be well-described in GSM calculation within the configuration mixing framework [8, 11].

The sd𝑠𝑑sditalic_s italic_d-shell nuclei, situated at the boundary between the light and heavy nuclei, exhibit a wide range of nuclear structure phenomena that remain somewhat mysterious [12]. In recent years, these nuclei have been extensively studied using a variety of experimental techniques [13, 14, 15]. A wealth of information on the Thomas-Ehrman shift has been gleaned from sd𝑠𝑑sditalic_s italic_d-shell proton drip-line nuclei, where numerous states exhibiting significant TES effects have been identified [3, 16, 4, 5]. For instance, the mirror pairs 18Ne/18[3, 17] and 19Na/19[16] serve as notable examples. For the sd𝑠𝑑sditalic_s italic_d-shell nuclei, TES is mainly driven by s𝑠sitalic_s-waves. Indeed, the proton 1s1/21subscript𝑠121s_{1/2}1 italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT orbital is weakly bound or unbound in proton drip-line nuclei, whereas the neutron 1s1/21subscript𝑠121s_{1/2}1 italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT is well-bound in their mirror neutron-rich nuclei.

Several theoretical models have been developed to probe the isospin asymmetry for mirror nuclei, especially the MEDs of sd𝑠𝑑sditalic_s italic_d-shell nuclei, such as the standard shell model (SM) [18, 19, 20, 21], mean-field calculations [22, 23], and ab initio approaches [24, 25, 26, 17, 27].Within the standard SM calculations, weakly bound and unbound wave functions on eigenenergies are indirectly considered via phenomenologically adjusting the matrix element related to 1s1/21subscript𝑠121s_{1/2}1 italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT orbit [7, 4]. Mean-field calculations, such as the Skyrme-Hartree-Fock, have also been extensively employed in MED studies [22, 23]. However, these models involve parameters that are constrained by data [18, 22, 28]. In recent years, ab initio approaches, such as the ab initio valence-space in-medium similarity renormalization group have also been applied to study MEDs of sd𝑠𝑑sditalic_s italic_d-shell nuclei [24, 25, 26, 17, 6, 5, 27, 29], in which, the extended many-body wavefunctions are partially described by using a large number of HO spaces. Moreover, current theoretical calculations have pointed out that the TES is caused by the repulsive Coulomb interaction and the occupations of weakly bound or unbound s𝑠sitalic_s- or p𝑝pitalic_p-waves for valence protons. However, detailed studies on the mechanism of the TES are also lacking.

One of the major challenges in studying drip line nuclei is accounting for the interplay between configuration-mixing and continuum effects. The Gamow shell model (GSM) [30, 31, 32, 33, 34, 17, 35, 9, 24, 10] has emerged as a powerful tool in this regard, as it provides a unified framework to describe the structure of nuclei close to the particle emission threshold and allows for a accurate understanding of the exotic properties in drip line nuclei. Based on the above situation, we employ the GSM to investigate the significant isospin symmetry breaking with large MED values and behind mechanism for the sd𝑠𝑑sditalic_s italic_d-shell nuclei, taking the 18Ne/18O and 19Na/19O mirror partner as examples.

Method.—  GSM is a multiconfiguration shell model framework, which works in the picture of a core plus valence nucleons [32, 36, 37, 38, 39]. At the heart of GSM lies the utilization of the one-body Berggren basis [40], which possesses bound, resonance, and scattering states, generated by a finite-range potential, typically of Woods-Saxon (WS) types (see details in Ref. [40, 32, 36]). Contrary to the traditional SM, the GSM Hamiltonian matrix is characterized as complex symmetric [32, 36]. Moreover, the GSM Hamiltonian possesses numerous many-body scattering states, so that the bound or resonance eigenstates are embedded among scattering eigenstates. Thus, one developed the overlap method along with the Jacobi-Davidson method extended to complex-symmetric matrices to diagonalize and identify many-body resonance eigenstates [32, 41, 36]. In the case of well-bound states in traditional SM calculations, the use of the Lanczos method is optimal [13]. As a consequence, the GSM calculation includes both the inter-nucleon correlations and continuum coupling [32, 36, 37].

The many-body Schrödinger equation of GSM Hamiltonian can be solved within the so-called cluster orbital shell model (COSM) formalism [42] (see Refs. [43, 44, 36]). The GSM Hamiltonian in COSM coordinates reads [43, 44, 36]:

H^GSM=i=1Aval(𝐩i22μi+U^i(c))+i<jAval(V^ij+𝐩i𝐩jMc),subscript^𝐻GSMsuperscriptsubscript𝑖1subscript𝐴𝑣𝑎𝑙superscriptsubscript𝐩𝑖22subscript𝜇𝑖superscriptsubscript^𝑈𝑖𝑐superscriptsubscript𝑖𝑗subscript𝐴𝑣𝑎𝑙subscript^𝑉𝑖𝑗subscript𝐩𝑖subscript𝐩𝑗subscript𝑀𝑐\!\!\!\hat{H}_{\rm GSM}\!=\!\!\sum_{i=1}^{A_{val}}\left(\frac{\mathbf{p}_{i}^{% 2}}{2\mu_{i}}+\hat{U}_{i}^{(c)}\right)+\sum_{i<j}^{A_{val}}\left(\hat{V}_{ij}+% \frac{\mathbf{p}_{i}\cdot\mathbf{p}_{j}}{M_{c}}\right),over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_GSM end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_v italic_a italic_l end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( divide start_ARG bold_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG + over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_c ) end_POSTSUPERSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_i < italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_v italic_a italic_l end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( over^ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT + divide start_ARG bold_p start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ bold_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ) , (1)

where Avalsubscript𝐴𝑣𝑎𝑙A_{val}italic_A start_POSTSUBSCRIPT italic_v italic_a italic_l end_POSTSUBSCRIPT is the number of valence nucleons, μisubscript𝜇𝑖\mu_{i}italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the effective mass of the nucleon, U^i(c)superscriptsubscript^𝑈𝑖𝑐\hat{U}_{i}^{(c)}over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_c ) end_POSTSUPERSCRIPT is represented by a one-body WS potential mimicking the inert core. V^ijsubscript^𝑉𝑖𝑗\hat{V}_{ij}over^ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the residual inter-nucleon interaction, which is modeled by a pionless effective field theory (EFT) interaction [45, 46], and the last term embodies the recoil effects induced by the finite mass of the core Mcsubscript𝑀𝑐M_{c}italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Concerning the used pionless EFT interaction [45, 46], only two-body contact terms up to next-to-next leading-order are considered in the present calculations, and EFT interaction is optimized to reproduce the low-lying states of selected nuclei. The harmonic oscillator (HO) basis used for the representation of the EFT interaction is limited to a few shells. This regularization approach has been recently utilized in Refs. [47, 48, 49, 50], and in particular in GSM calculation of unbound neutron-rich oxygen isotopes [38].

In the present work, the 18Ne/18O and 19Na/19O mirror partners are taken as examples. The doubly magic nucleus 16O is chosen as the inert core, and the s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT, p1/2,3/2subscript𝑝1232p_{1/2,3/2}italic_p start_POSTSUBSCRIPT 1 / 2 , 3 / 2 end_POSTSUBSCRIPT and d3/2,5/2subscript𝑑3252d_{3/2,5/2}italic_d start_POSTSUBSCRIPT 3 / 2 , 5 / 2 end_POSTSUBSCRIPT partial waves are represented by the Berggren basis, in which 40 discretization points are used in total for continuum states in each partial wave, and the f5/2,7/2subscript𝑓5272f_{5/2,7/2}italic_f start_POSTSUBSCRIPT 5 / 2 , 7 / 2 end_POSTSUBSCRIPT partial waves are treated using the HO basis, in which 6 HO states are adopted for each partial wave. Only the Coulomb force is considered for the isospin non-conserving part of the GSM Hamiltonian. The contribution of the isospin-dependent part of nuclear interaction to TES is small, which is neglected in the present GSM calculations. The adopted Hamiltonian is optimized to produce the selected experimental data of sd𝑠𝑑sditalic_s italic_d-shell nuclei. The calculated excitation energies of 18Ne/18O and 19Na/19O mirror partners are presented in Tables. 1 and 2, which show good agreements with experimental data [51]. In the following section, the mechanics of mirror energies difference for the 18Ne/18O and 19Na/19O mirror partners are detail investigated.

Jπsuperscript𝐽𝜋J^{\pi}italic_J start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT 18Ne 18O
Eexpsubscript𝐸expE_{\rm exp}italic_E start_POSTSUBSCRIPT roman_exp end_POSTSUBSCRIPT EGSMsubscript𝐸GSME_{\rm GSM}italic_E start_POSTSUBSCRIPT roman_GSM end_POSTSUBSCRIPT ΓexpsubscriptΓexp\Gamma_{\rm exp}roman_Γ start_POSTSUBSCRIPT roman_exp end_POSTSUBSCRIPT ΓGSMsubscriptΓGSM\Gamma_{\rm GSM}roman_Γ start_POSTSUBSCRIPT roman_GSM end_POSTSUBSCRIPT Eexpsubscript𝐸expE_{\rm exp}italic_E start_POSTSUBSCRIPT roman_exp end_POSTSUBSCRIPT EGSMsubscript𝐸GSME_{\rm GSM}italic_E start_POSTSUBSCRIPT roman_GSM end_POSTSUBSCRIPT MEDexp MEDGSM
01+superscriptsubscript010_{1}^{+}0 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 0 0 0 0 0 0 0
21+superscriptsubscript212_{1}^{+}2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 1.89 1.83 0 1.98 1.93 9595-95- 95 106106-106- 106
41+superscriptsubscript414_{1}^{+}4 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 3.38 2.72 0 3.56 2.72 179179-179- 179 33-3- 3
22+superscriptsubscript222_{2}^{+}2 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 3.62 3.98 0 3.92 4.40 304304-304- 304 420420-420- 420
02+superscriptsubscript020_{2}^{+}0 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 3.58 4.58 0 3.63 5.42 5858-58- 58 834834-834- 834
31+superscriptsubscript313_{1}^{+}3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 4.56 4.63 18 80 5.38 5.53 817817-817- 817 892892-892- 892
Table 1: The calculated excitation energies of 18Ne/18O with GSM calculations, along with experimental data [51]. The unit of excitation energy is given in MeV, and the units of particle decay width and MED are given in keV.
Jπsuperscript𝐽𝜋J^{\pi}italic_J start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT 19Na 19O
Eexpsubscript𝐸expE_{\rm exp}italic_E start_POSTSUBSCRIPT roman_exp end_POSTSUBSCRIPT EGSMsubscript𝐸GSME_{\rm GSM}italic_E start_POSTSUBSCRIPT roman_GSM end_POSTSUBSCRIPT Eexpsubscript𝐸expE_{\rm exp}italic_E start_POSTSUBSCRIPT roman_exp end_POSTSUBSCRIPT EGSMsubscript𝐸GSME_{\rm GSM}italic_E start_POSTSUBSCRIPT roman_GSM end_POSTSUBSCRIPT MEDexp MEDGSM
5/21+5superscriptsubscript215/2_{1}^{+}5 / 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 0 0 0 0 0 0
3/21+3superscriptsubscript213/2_{1}^{+}3 / 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 0.12 0.64 0.10 0.71 24 6464-64- 64
1/21+1superscriptsubscript211/2_{1}^{+}1 / 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT 0.75 0.69 1.47 1.30 727727-727- 727 607607-607- 607
Table 2: Similar to Table 1, but for 19Na/19O.

Results.— Our GSM calculations accurately describe the excitation energies of low-lying states for the mirror partners 18Ne/18O and 19Na/19O. To delve deeper into the significant isospin symmetry breaking observed in these mirror nuclei, we define the MED for a given state Jπsuperscript𝐽𝜋J^{\pi}italic_J start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT as MED(Jπ)superscript𝐽𝜋(J^{\pi})( italic_J start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT ) = Ex(Tz<,Jπ)Ex(Tz>,Jπ)subscript𝐸𝑥superscriptsubscript𝑇𝑧superscript𝐽𝜋subscript𝐸𝑥superscriptsubscript𝑇𝑧superscript𝐽𝜋E_{x}(T_{z}^{<},J^{\pi})-E_{x}(T_{z}^{>},J^{\pi})italic_E start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_T start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT < end_POSTSUPERSCRIPT , italic_J start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT ) - italic_E start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_T start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT > end_POSTSUPERSCRIPT , italic_J start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT ), where Tz<superscriptsubscript𝑇𝑧T_{z}^{<}italic_T start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT < end_POSTSUPERSCRIPT and Tz>superscriptsubscript𝑇𝑧T_{z}^{>}italic_T start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT > end_POSTSUPERSCRIPT denote the negative and positive isospin projection Tz=(NZ)/2subscript𝑇𝑧𝑁𝑍2T_{z}=(N-Z)/2italic_T start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ( italic_N - italic_Z ) / 2, respectively, for a mirror pair. We have calculated MED values for mirror states in 18Ne/18O and 19Na/19O, as presented in Tables 1 and 2, along with experimental data. It is observed that the calculated MED values for low-lying states in the mirror partners 18Ne/18O and 19Na/19O, align well with the experimental data. However, an exception is noted for the 02+superscriptsubscript020_{2}^{+}0 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state in the 18Ne/18O mirror nuclei, where our GSM calculations yield larger values than the experimental data. Both our GSM calculations and the experimental data highlight significant isospin symmetry breaking in the 31+superscriptsubscript313_{1}^{+}3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state of the 18Ne/18O mirror nuclei and the 1/21+1superscriptsubscript211/2_{1}^{+}1 / 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state of the 19Na/19O mirror nuclei, evidenced by their large MED values.

Refer to caption
Figure 1: The average occupation numbers for the s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT and d5/2subscript𝑑52d_{5/2}italic_d start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT partial waves in the low-lying states of the 18Ne/18O mirror pair, calculated using the GSM above the 16O core.
Refer to caption
Figure 2: Similar to Fig. 1, but for low-lying states in 19Na/19O.

To investigate the significant isospin symmetry breaking and the associated large MEDs, we begin with calculating the average occupations of low-lying states through the GSM. The focus is particularly on the s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT and d5/2subscript𝑑52d_{5/2}italic_d start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT partial waves above the 16O core for the 18Ne/18O and 19Na/19O mirror nuclei, as illustrated in Figs. 1 and 2. Notably, other partial waves like d3/2subscript𝑑32d_{3/2}italic_d start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT, f5/2,7/2subscript𝑓5272f_{5/2,7/2}italic_f start_POSTSUBSCRIPT 5 / 2 , 7 / 2 end_POSTSUBSCRIPT exhibit negligible occupations and are, therefore, excluded from these figures. The calculated average occupations reveal almost identical patterns for mirror states within the 18Ne/18O and 19Na/19O pairs. Our GSM calculations further indicate that states exhibiting significant isospin symmetry breaking with large MED values also show significant occupancy in the s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT partial waves—markedly higher than in their respective ground states. For instance, the occupations of the s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT partial wave for the 31+superscriptsubscript313_{1}^{+}3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and 1/21+1superscriptsubscript211/2_{1}^{+}1 / 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT states in the 18Ne/18O and 19Na/19O mirror pairs, respectively, are substantially greater than those of the ground states. Additionally, our calculations give that the 02+superscriptsubscript020_{2}^{+}0 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT of 18Ne/18O demonstrate a notable s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT partial wave occupations compared to the ground states, resulting in a large MED. Contrastingly, experimental data give a smaller MED value, hinting at a complex structure of the 02+superscriptsubscript020_{2}^{+}0 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT states in 18Ne/18O that might not be fully captured by 16O plus valence particles picture.

Refer to caption
Figure 3: The calculated radial density distribution of valence protons and valence neutrons for low-lying mirror states in 18Ne/18O, respectively, above 16O inner core, using GSM.
Refer to caption
Figure 4: Similar to Fig. 3, but for low-lying states in 19Na/19O.

Aligned with results from other theoretical frameworks, such as the standard shell model and the ab initio VS-IMSRG approach, our results indicate that the significant isospin symmetry breaking with large MEDs observed in mirror states stem primarily from the extensive occupation of the s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT partial waves, which is weakly bound or unbound in the proton-rich nucleus but deeply bound in its mirror neutron-rich nucleus, called TES. However, a further deep understanding of the mirror state bearing significant isospin symmetry breaking with large MED value is lacking. The GSM is a very suitable model, which properly treats both the inter-nucleon correlations and continuum coupling, to describe the properties of dripline nuclei, including a precious description of the many-body wave function in the asymptotic regions [38, 52, 53, 54].

To elucidate the underlying mechanism of large MEDs, we conduct a detailed analysis of the radial density distributions of mirror states in 18Ne/18O and 19Na/19O pairs using the GSM. The results allow us to systematically compare the radial distributions of valence protons in proton-rich nuclei and valence neutrons in their neutron-rich mirror counterparts. The results are presented in Figs. 3 and 4 for 18Ne/18O and 19Na/19O mirror partners, respectively. Our GSM results reveal that the states characterized by minor isospin symmetry breaking with small MEDs exhibit almost identical radial density distributions, which decline sharply in the asymptotic regions, such as the ground states of both 18Ne/18O and 19Na/19O. This phenomenon is largely attributed to the dominance of d5/2subscript𝑑52d_{5/2}italic_d start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT partial waves, which are constrained within the nuclear region by high centrifugal and Coulomb barriers, despite the state being unbound. Conversely, GSM calculations depict the radial density distributions of the 31+superscriptsubscript313_{1}^{+}3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state of 18Ne and the 1/21+1superscriptsubscript211/2_{1}^{+}1 / 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state of 19Na as more extended in the asymptotic region than their neutron-rich counterparts, 18O and 19O, respectively. This disparity stems from the non-existent centrifugal barrier for the s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT partial wave, leading to a more pronounced distribution in the proton-rich nucleus due to the weakly bound or unbound nature of the s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT partial wave. A similar mechanism underlies the formation of halo nuclei, where the valence nucleons occupy weakly bound s𝑠sitalic_s- or p𝑝pitalic_p- partial waves, resulting in an extended density distribution due to the minimal or absent centrifugal barrier [43, 9, 55]. Consequently, our GSM calculations provide direct calculations for radial density distribution and unveil that the mirror states demonstrating significant isospin symmetry breaking with large MEDs possess many-body wave functions in proton-rich nuclei that are more extended than those in their neutron-rich mirror states, which add a new dimension to our understanding of the role of isospin symmetry breaking in sha** their properties.

Refer to caption
Figure 5: Upper panel: calculated energies (GSM), energies minus the two-body Coulomb contribution (GSM-2BC), and energies minus one- and two-body Coulomb total contributions (GSM-1BC-2BC) of low-lying states of 18Ne using GSM, along with the energies of mirror states in 18O, with respect to the 16O inner core. The GSM results are also compared with experimental data. Lower panel: the contribution for the calculated energy difference between the mirror states.
Refer to caption
Figure 6: Similar to Fig. 5, but for low-lying states in 19Na/19O.

Within the present GSM calculations, the GSM Hamiltonian, as shown in Eq. (1), can be divided into nuclear interaction (encompassing core-nucleons and nucleon-nucleon interaction) and Coulomb interaction (including one-body Coulomb (1BC) interaction between the inner core and valence protons, two-body Coulomb (2BC) interaction between valence protons). We delve into further calculations to dissect the contributions from different parts of the Hamiltonian, aiming to shed light on the underlying mechanisms in mirror states exhibiting significant isospin symmetry breaking with large MED.

The computed energies for the low-lying mirror states in the pairs 18Ne/18O and 19Na/19O, along with experimental data [51], are showcased in Figs. 5 and 6, respectively. To gain deeper insights, we also present the energy minus 2BC contribution (GSM-2BC) and energy minus 1BC and 2BC contribution (GSM-1BC-2BC) in proton-rich nuclei 18Ne and 19Na. Indeed, the GSM-1BC-2BC also corresponds to the contribution of nuclear interaction. Within the isospin symmetry picture, the difference in mirror state energies should solely stem from Coulomb interactions, implying that the GSM-1BC-2BC values for a state in a proton-rich nucleus would be the same as its mirror state in the neutron-rich nucleus.

Our GSM calculation gives that the GSM-1BC-2BC values for the ground states of 18Ne and 19Na closely align with the computed ground-state energies of their neutron-rich counterparts, 18O and 19O, respectively. The results indicate the preservation of isospin symmetry for these ground states. Conversely, for the excited 31+superscriptsubscript313_{1}^{+}3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state in 18Ne/18O and the 1/21+1superscriptsubscript211/2_{1}^{+}1 / 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state in 19Na/19O, our GSM calculations showcase a deviation from this symmetry. Specifically, the GSM-1BC-2BC values for the 31+superscriptsubscript313_{1}^{+}3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state of 18Ne and the 1/21+1superscriptsubscript211/2_{1}^{+}1 / 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT state of 19Na are more unbound compared to their computed energies of mirror states. To quantitatively examine this discrepancy, we introduce ΔEΔ𝐸\Delta Eroman_Δ italic_E as the differential metric. ΔEΔ𝐸\Delta Eroman_Δ italic_E encapsulates the disparity between the GSM-1BC-2BC values in the state of the proton-rich nucleus and the energy calculated for the corresponding state in the neutron-rich mirror nucleus, which read as ΔE=Ψproton|HNN|ΨprotonΨneutron|HNN|ΨneutronΔ𝐸quantum-operator-productsubscriptΨprotonsubscript𝐻𝑁𝑁subscriptΨprotonquantum-operator-productsubscriptΨneutronsubscript𝐻𝑁𝑁subscriptΨneutron\Delta E=\langle\Psi_{\rm proton}|H_{NN}|\Psi_{\rm proton}\rangle-\langle\Psi_% {\rm neutron}|H_{NN}|\Psi_{\rm neutron}\rangleroman_Δ italic_E = ⟨ roman_Ψ start_POSTSUBSCRIPT roman_proton end_POSTSUBSCRIPT | italic_H start_POSTSUBSCRIPT italic_N italic_N end_POSTSUBSCRIPT | roman_Ψ start_POSTSUBSCRIPT roman_proton end_POSTSUBSCRIPT ⟩ - ⟨ roman_Ψ start_POSTSUBSCRIPT roman_neutron end_POSTSUBSCRIPT | italic_H start_POSTSUBSCRIPT italic_N italic_N end_POSTSUBSCRIPT | roman_Ψ start_POSTSUBSCRIPT roman_neutron end_POSTSUBSCRIPT ⟩, the ΨprotonsubscriptΨproton\Psi_{\rm proton}roman_Ψ start_POSTSUBSCRIPT roman_proton end_POSTSUBSCRIPT and ΨneutronsubscriptΨneutron\Psi_{\rm neutron}roman_Ψ start_POSTSUBSCRIPT roman_neutron end_POSTSUBSCRIPT correspond to the many-body wave function of proton-rich and neutron-rich nuclei, respectively. The ΔEΔ𝐸\Delta Eroman_Δ italic_E corresponds to the difference in the contribution of nuclear interactions in the mirror state.

Our GSM calculations show that both ΔEΔ𝐸\Delta Eroman_Δ italic_E and Coulomb interactions, including 1BC and 2BC , significantly influence the energy discrepancies observed in mirror states. Predominantly, the Coulomb interaction emerges as the dominant factor contributing to these differences. Illustrated in the lower panels of Figs. 5 and 6, we detail the ΔEΔ𝐸\Delta Eroman_Δ italic_E, 1BC, and 2BC contributions to the energy differences in the low-lying mirror states of 18Ne/18O and 19Na/19O mirror pairs. Our GSM results indicate that the energy differences in the ground states of 18Ne/18O primarily stem from Coulomb interactions, with ΔEΔ𝐸\Delta Eroman_Δ italic_E making a minimal contribution. Furthermore, for 19Na/19O, the ΔEΔ𝐸\Delta Eroman_Δ italic_E contribution is noted to be around 100 keV. Interestingly, we find varying contributions of ΔEΔ𝐸\Delta Eroman_Δ italic_E, 1BC, and 2BC across different mirror states within each state. For instance, the 31+superscriptsubscript313_{1}^{+}3 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT states in 18Ne/18O exhibit a higher ΔEΔ𝐸\Delta Eroman_Δ italic_E contribution and lower Coulomb interactions, relative to their ground states. The heightened ΔEΔ𝐸\Delta Eroman_Δ italic_E values underscore the distinct nuclear interaction contributions to isospin symmetry breaking in these systems, showcasing the complex interplay of forces that shape the energy landscapes of mirror nuclei.

In evaluating the MED, the ground state energy of mirror nuclei serves as the baseline, with MED being determined by the discrepancy in excitation energies of corresponding mirror states. Adopting the energy difference of the ground states of mirror nuclei as a reference—illustrated by red dashed lines in Figs. 5 and 6. The difference between the values of ground and excited mirror states corresponds to the MED, highlighted by the red arrows in these figures. The results reveal that both the ΔEΔ𝐸\Delta Eroman_Δ italic_E values and the Coulomb interaction exhibit significant variations across different mirror states, both contributing to the MED.

Summary.—Based on the GSM calculations, in which both the inter-nucleon correlation and continuum coupling are properly treated, we deduce that significant isospin symmetry breaking in mirror states, leading to large MED values, arises from the occupation of weakly-bound or unbound s1/2subscript𝑠12s_{1/2}italic_s start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT partial waves in the proton-rich nucleus, while its counterpart in the neutron-rich nucleus remains deeply bound. This dichotomy culminates in a more expansive radial density distribution for states within the proton-rich nucleus, as opposed to their mirror counterparts. Additionally, the difference in radial density distributions between mirror states implies disparate contributions from nuclear interactions, underscored by significant ΔEΔ𝐸\Delta Eroman_Δ italic_E values, which further highlight the presence of isospin symmetry breaking. Moreover, states with an extended radial density distribution tend to yield smaller Coulomb contributions compared to ground states characterized by more localized distributions. This factor chiefly accounts for the reduced excitation energies in states influenced by the Thomas-Ehrman shift effect, thereby engendering substantial negative MED values in mirror states. Our GSM calculations corroborate that both nuclear and Coulomb interactions play crucial roles in manifesting the significant isospin symmetry breaking associated with significant MED values.

Acknowledgments — This work has been supported by the National Key R&D Program of China under Grant No. 2023YFA1606403; the National Natural Science Foundation of China under Grant Nos. 12205340, 12347106, and 12121005; the Gansu Natural Science Foundation under Grant No. 22JR5RA123 and 23JRRA614; the Key Research Program of the Chinese Academy of Sciences under Grant No. XDPB15; the State Key Laboratory of Nuclear Physics and Technology, Peking University under Grant No. NPT2020KFY13. The numerical calculations in this paper have been done on Hefei advanced computing center.

References

References

  • [1] R. G. Thomas, Phys. Rev. 88 (1952) 1109–1125.
  • [2] J. B. Ehrman, Phys. Rev. 81 (1951) 412–416.
  • [3] K. A. Chipps, D. W. Bardayan, J. C. Blackmon, K. Y. Chae, U. Greife, R. Hatarik, R. L. Kozub, C. Matei, B. H. Moazen, C. D. Nesaraja, S. D. Pain, W. A. Peters, S. T. Pittman, J. F. Shriner, M. S. Smith, Phys. Rev. Lett. 102 (2009) 152502.
  • [4] J. Lee, X. X. Xu, K. Kaneko, Y. Sun, C. J. Lin, L. J. Sun, P. F. Liang, Z. H. Li, J. Li, H. Y. Wu, D. Q. Fang, J. S. Wang, Y. Y. Yang, C. X. Yuan, Y. H. Lam, Y. T. Wang, K. Wang, J. G. Wang, J. B. Ma, J. J. Liu, P. J. Li, Q. Q. Zhao, L. Yang, N. R. Ma, D. X. Wang, F. P. Zhong, S. H. Zhong, F. Yang, H. M. Jia, P. W. Wen, M. Pan, H. L. Zang, X. Wang, C. G. Wu, D. W. Luo, H. W. Wang, C. Li, C. Z. Shi, M. W. Nie, X. F. Li, H. Li, P. Ma, Q. Hu, G. Z. Shi, S. L. **, M. R. Huang, Z. Bai, Y. J. Zhou, W. H. Ma, F. F. Duan, S. Y. **, Q. R. Gao, X. H. Zhou, Z. G. Hu, M. Wang, M. L. Liu, R. F. Chen, X. W. Ma, Phys. Rev. Lett. 125 (2020) 192503.
  • [5] M. Z. Sun, Y. Yu, X. P. Wang, M. Wang, J. G. Li, Y. H. Zhang, K. Blaum, Z. Y. Chen, R. J. Chen, H. Y. Deng, C. Y. Fu, W. W. Ge, W. J. Huang, H. Y. Jiao, H. H. Li, H. F. Li, Y. F. Luo, T. Liao, Y. A. Litvinov, M. Si, P. Shuai, J. Y. Shi, Q. Wang, Y. M. Xing, X. Xu, H. S. Xu, F. R. Xu, Q. Yuan, T. Yamaguchi, X. L. Yan, J. C. Yang, Y. J. Yuan, X. H. Zhou, X. Zhou, M. Zhang, Q. Zeng, Chinese Physics C 48 (3) (2024) 034002.
  • [6] H. H. Li, Q. Yuan, J. G. Li, M. R. Xie, S. Zhang, Y. H. Zhang, X. X. Xu, N. Michel, F. R. Xu, W. Zuo, Phys. Rev. C 107 (2023) 014302.
  • [7] C. Yuan, C. Qi, F. Xu, T. Suzuki, T. Otsuka, Phys. Rev. C 89 (2014) 044327.
  • [8] I. Stefan, F. de Oliveira Santos, O. Sorlin, T. Davinson, M. Lewitowicz, G. Dumitru, J. C. Angélique, M. Angélique, E. Berthoumieux, C. Borcea, R. Borcea, A. Buta, J. M. Daugas, F. de Grancey, M. Fadil, S. Grévy, J. Kiener, A. Lefebvre-Schuhl, M. Lenhardt, J. Mrazek, F. Negoita, D. Pantelica, M. G. Pellegriti, L. Perrot, M. Ploszajczak, O. Roig, M. G. Saint Laurent, I. Ray, M. Stanoiu, C. Stodel, V. Tatischeff, J. C. Thomas, Phys. Rev. C 90 (2014) 014307.
  • [9] J. G. Li, N. Michel, H. H. Li, W. Zuo, Phys. Lett. B 832 (2022) 137225.
  • [10] X. Mao, J. Rotureau, W. Nazarewicz, N. Michel, R. M. Id Betan, Y. Jaganathen, Phys. Rev. C 102 (2020) 024309.
  • [11] N. Michel, J. G. Li, L. H. Ru, W. Zuo, Phys. Rev. C 106 (2022) L011301.
  • [12] K. Way, Science 122 (3170) (1955) 603–603.
  • [13] E. Caurier, G. Martínez-Pinedo, F. Nowacki, A. Poves, A. P. Zuker, Rev. Mod. Phys. 77 (2005) 427–488.
  • [14] F. Ajzenberg-Selove, Nuclear Physics A 490 (1) (1988) 1–225.
  • [15] P. Campbell, I. Moore, M. Pearson, Progress in Particle and Nuclear Physics 86 (2016) 127–180.
  • [16] C. Angulo, G. Tabacaru, M. Couder, M. Gaelens, P. Leleux, A. Ninane, F. Vanderbist, T. Davinson, P. J. Woods, J. S. Schweitzer, N. L. Achouri, J. C. Angélique, E. Berthoumieux, F. de Oliveira Santos, P. Himpe, P. Descouvemont, Phys. Rev. C 67 (2003) 014308.
  • [17] S. Zhang, Y. Ma, J. Li, B. Hu, Q. Yuan, Z. Cheng, F. Xu, Phys. Lett. B 827 (2022) 136958.
  • [18] M. Bentley, S. Lenzi, Prog. Part. Nucl. Phys 59 (2) (2007) 497–561.
  • [19] A. P. Zuker, S. M. Lenzi, G. Martínez-Pinedo, A. Poves, Phys. Rev. Lett. 89 (2002) 142502.
  • [20] K. Kaneko, Y. Sun, T. Mizusaki, S. Tazaki, Phys. Rev. Lett. 110 (2013) 172505.
  • [21] Y. H. Lam, N. A. Smirnova, E. Caurier, Phys. Rev. C 87 (2013) 054304.
  • [22] P. Ba̧czyk, J. Dobaczewski, M. Konieczka, W. Satuła, T. Nakatsukasa, K. Sato, Phys. Lett. B 778 (2018) 178–183.
  • [23] R. D. O. Llewellyn, M. A. Bentley, R. Wadsworth, H. Iwasaki, J. Dobaczewski, G. de Angelis, J. Ash, D. Bazin, P. C. Bender, B. Cederwall, B. P. Crider, M. Doncel, R. Elder, B. Elman, A. Gade, M. Grinder, T. Haylett, D. G. Jenkins, I. Y. Lee, B. Longfellow, E. Lunderberg, T. Mijatović, S. A. Milne, D. Muir, A. Pastore, D. Rhodes, D. Weisshaar, Phys. Rev. Lett. 124 (2020) 152501.
  • [24] J. G. Li, N. Michel, W. Zuo, F. R. Xu, Phys. Rev. C 104 (2021) 024319.
  • [25] M. S. Martin, S. R. Stroberg, J. D. Holt, K. G. Leach, Phys. Rev. C 104 (2021) 014324.
  • [26] E. Caurier, P. Navrátil, W. E. Ormand, J. P. Vary, Phys. Rev. C 66 (2002) 024314.
  • [27] J. G. Li, H. H. Li, S. Zhang, Y. M. Xing, W. Zuo, Physics Letters B 846 (2023) 138197.
  • [28] S. Uthayakumaar, M. A. Bentley, E. C. Simpson, T. Haylett, R. Yajzey, S. M. Lenzi, W. Satuła, D. Bazin, J. Belarge, P. C. Bender, P. J. Davies, B. Elman, A. Gade, H. Iwasaki, D. Kahl, N. Kobayashi, B. Longfellow, S. J. Lonsdale, E. Lunderberg, L. Morris, D. R. Napoli, T. G. Parry, X. Pereira-Lopez, F. Recchia, J. A. Tostevin, R. Wadsworth, D. Weisshaar, Phys. Rev. C 106 (2022) 024327.
  • [29] H. H. Li, J. G. Li, M. R. Xie, W. Zuo, Chin. Phys. C 47 (2023) 124101.
  • [30] R. Id Betan, R. J. Liotta, N. Sandulescu, T. Vertse, Phys. Rev. Lett. 89 (2002) 042501.
  • [31] N. Michel, W. Nazarewicz, M. Płoszajczak, K. Bennaceur, Phys. Rev. Lett. 89 (2002) 042502.
  • [32] N. Michel, W. Nazarewicz, M. Płoszajczak, T. Vertse, Jour. Phys. G. Nucl. Part. Phys. 36 (1) (2009) 013101.
  • [33] N. Michel, M. Płoszajczak, Gamow Shell Model, Springer Berlin Heidelberg, 2021.
  • [34] J. G. Li, B. S. Hu, Q. Wu, Y. Gao, S. J. Dai, F. R. Xu, Phys. Rev. C 102 (2020) 034302.
  • [35] H. H. Li, J. G. Li, N. Michel, W. Zuo, Phys. Rev. C 104 (2021) L061306.
  • [36] N. Michel, M. Płoszajczak, Gamow Shell Model, The Unified Theory of Nuclear Structure and Reactions, Lecture Notes in Physics, Vol. 983, Springer, Berlin, 2021.
  • [37] J. G. Li, Y. Z. Ma, N. Michel, B. S. Hu, Z. H. Sun, W. Zuo, F. R. Xu, Physics 3 (4) (2021) 977–997.
  • [38] J. G. Li, N. Michel, W. Zuo, F. R. Xu, Phys. Rev. C 103 (2021) 034305.
  • [39] S. Zhang, F. R. Xu, J. G. Li, B. S. Hu, Z. H. Cheng, N. Michel, Y. Z. Ma, Q. Yuan, Y. H. Zhang, Phys. Rev. C 108 (2023) 064316.
  • [40] T. Berggren, Nucl. Phys. A 109 (2) (1968) 265 – 287.
  • [41] N. Michel, H. Aktulga, Y. Jaganathen, Comput. Phys. Commun. 247 (2020) 106978.
  • [42] Y. Suzuki, K. Ikeda, Phys. Rev. C 38 (1988) 410–413.
  • [43] G. Papadimitriou, A. T. Kruppa, N. Michel, W. Nazarewicz, M. Płoszajczak, J. Rotureau, Phys. Rev. C 84 (2011) 051304(R).
  • [44] Y. Jaganathen, R. M. Id Betan, N. Michel, W. Nazarewicz, M. Płoszajczak, Phys. Rev. C 96 (2017) 054316.
  • [45] H.-W. Hammer, A. Nogga, A. Schwenk, Rev. Mod. Phys. 85 (2013) 197–217.
  • [46] H.-W. Hammer, S. König, U. van Kolck, Rev. Mod. Phys. 92 (2020) 025004.
  • [47] S. Binder, A. Ekström, G. Hagen, T. Papenbrock, K. A. Wendt, Phys. Rev. C 93 (2016) 044332.
  • [48] R. J. Furnstahl, G. Hagen, T. Papenbrock, Phys. Rev. C 86 (2012) 031301(R).
  • [49] A. Bansal, S. Binder, A. Ekström, G. Hagen, G. R. Jansen, T. Papenbrock, Phys. Rev. C 98 (2018) 054301.
  • [50] L. Huth, V. Durant, J. Simonis, A. Schwenk, Phys. Rev. C 98 (2018) 044301.
  • [51] http://www.nndc.bnl.gov/ensdf.
  • [52] N. Michel, J. G. Li, F. R. Xu, W. Zuo, Phys. Rev. C 100 (2019) 064303.
  • [53] M. R. Xie, J. G. Li, N. Michel, H. H. Li, S. T. Wang, H. J. Ong, W. Zuo, Phys. Lett. B (2023) 137800.
  • [54] M. Xie, J. Li, N. Michel, H. Li, W. Zuo, Science China Physics, Mechanics & Astronomy 67 (1) (2023) 212011.
  • [55] N. Michel, J. G. Li, F. R. Xu, W. Zuo, Phys. Rev. C 101 (2020) 031301(R).