logo

On shape coexistence and possible shape isomers of nuclei around 172Hg

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

On shape coexistence and possible shape isomers of nuclei around 172Hg

Xin Guan
Jing Guo
Qi-Wen Sun
Bożena Nerlo-Pomorska
Krzysztof Pomorski
Nuclear Science and TechniquesVol.36, No.7Article number 128Published in print Jul 2025Available online 26 May 2025
11300

This study explores the phenomenon of shape coexistence in nuclei around 172Hg, with a focus on the isotopes 170Pt, 172Hg, and 174Pb, as well as the 170Pt to 180Pt isotopic chain. Utilizing a macro-microscopic approach that incorporates the Lublin-Strasbourg Drop model combined with a Yukawa-Folded potential and pairing corrections, we analyze the potential energy surfaces (PESs) to understand the impact of pairing interaction. For170Pt, the PES exhibited a prolate ground-state, with additional triaxial and oblate-shaped isomers. In 172Hg, the ground-state deformation transitions from triaxial to oblate with increasing pairing interaction, demonstrating its nearly γ-unstable nature. Three shape isomers (prolate, triaxial, and oblate) were observed, with increased pairing strength leading to the disappearance of the triaxial isomer. 174Pb exhibited a prolate ground-state that became increasingly spherical with stronger pairing. While shape isomers were present at lower pairing strengths, robust shape coexistence was not observed. For realistic pairing interaction, the ground-state shapes transitioned from prolate in 170Pt to a coexistence of γ-unstable and oblate shapes in 172Hg, ultimately approaching spherical symmetry in 174Pb. A comparison between Exact and Bardeen-Cooper-Schrieffer (BCS) pairing demonstrated that BCS pairing tends to smooth out shape coexistence and reduce the depth of the shape isomer, leading to less pronounced deformation features. The PESs for even-even 170-180Pt isotopes revealed significant shape evolution. 170Pt showed a prolate ground-state, whereas 172Pt exhibited both triaxial and prolate shape coexistence. In 174Pt, the ground-state was triaxial, coexisted with a prolate minimum. For 176Pt, a γ-unstable ground-state coexists with a prolate minimum. By 178Pt and 180Pt, a dominant prolate minimum emerged. These results highlight the role of shape coexistence and γ-instability in the evolution of nuclear structure, especially in the mid-shell region. These findings highlight the importance of pairing interactions in nuclear deformation and shape coexistence, providing insights into the structural evolution of mid-shell nuclei.

Macro-micro modelShape coexistenceShape isomersExact and BCS pairing solutions
1

Introduction

Shape coexistence in atomic nuclei has garnered significant attention in the field of nuclear physics and has become a prominent topic in contemporary research. This phenomenon refers to the presence of multiple distinct shapes within a single nucleus, where states with similar energies exhibit different deformations [1]. Understanding nuclear shapes is crucial for revealing the internal structure and properties of nuclei, providing tools for predicting and explaining nuclear behaviors, and advancing nuclear physics [2-6].

The study of nuclear shapes has a long history, with several foundational studies laying the groundwork for our current understanding. Early theoretical developments included Rainwater's 1950 paper [7], which first proposed the idea of nuclear deformation, and Bohr and Mottelson's collective model [8, 9], which provided a framework for describing rotational spectra in deformed nuclei. Arima and Horie's 1954 study [10] explored the role of configuration mixing in nuclear structure, while Nilsson's work [11] introduced a shell-model approach incorporating deformation effects. Around the same time, Morinaga's 1956 paper [12] specifically addressed the structure of 16O and explained the properties of its first excited state and ground state. He introduced the concept of multi-nucleon cross-shell excitation to describe the deformation characteristics, offering a new perspective on how nuclear shapes evolve. Further developments include Elliott's work in 1958 [13], which further developed the concept of SU(3) symmetry in nuclear deformation and highlighted the interplay between single-particle and collective motion. Over the past five decades, shape coexistence has evolved from a rare phenomenon to a common feature observed in many nuclei, highlighting its significance in nuclear structure research [14]. Recent experimental studies have revealed significant evidence of shape coexistence phenomena in neutron-deficient isotopes of lead and mercury. For instance, one study [15] specifically focuses on the 188Hg isotope, where theoretical predictions suggest the presence of shape coexistence.

These findings have led to increased theoretical investigations into nuclear shape coexistence, utilizing advanced experimental techniques such as tagging techniques at the University of Jyväskylä, Coulomb-excitation experiments at CERN, and relativistic energy-fragmentation experiments at GSI [16]. These experiments underscore the importance of understanding the mechanisms governing the evolution of nuclear shape. Building upon these experimental insights, theoretical investigations have played a pivotal role in elucidating the complexities of shape coexistence [17-19]. Previous studies have employed various theoretical frameworks, including macro-microscopic approaches and self-consistent models, to perform comprehensive calculations of nuclear ground-state masses and deformations across a wide range of nuclei [14]. Ref. [20] highlighted the presence of two distinct coexisting configurations, in platinum isotopes 176-186Pt, oblate and prolate, revealing the intricate shape evolution in this mass region. Therefore, shape coexistence in even-even 172-200Hg isotopes was comprehensively studied using the interacting boson model with configuration mixing [21]. Recently, using the Lublin-Strasbourg Drop (LSD) with Yukawa-Folded single-particle potential and the BCS pairing correction in a macro-microscopic model, Pomorski et al. provided the deformation PESs of nuclei near Z =82. Their study investigated the shape coexistence phenomenon in even-even isotopes of Pt, Hg, and Pb [22]. These studies revealed that nuclei in the vicinity of Hg exhibit a rich variety of shape coexistence phenomena, characterized by the interplay of spherical, oblate, and prolate configurations. Although significant progress has been made in understanding these features of heavier isotopes, lighter isotopes of Hg, Pt, and Pb have been relatively underexplored owing to the scarcity of experimental data [23]. To address this gap, further theoretical investigations are crucial, as they can illuminate the evolution of shape coexistence in these lighter isotopes. Such efforts would not only enhance our theoretical understanding but also provide valuable guidance for future experimental measurements, enabling better interpretation of the limited or ambiguous data that are currently available.

Despite its success, the BCS method [24], as well as the more refined Hartree-Fock-Bogolyubov (HFB) approach face limitations due to the small number of valence nucleons under the pairing correlation's influence [25-31]. These methods often fail to conserve particle numbers, leading to inaccuracies in describing higher-lying excited states [32]. Alternatives such as the shell model provide successful descriptions but are limited by the combinatorial growth of model space sizes, necessitating truncation schemes for heavy nuclei and often being constrained by computational resources [33]. Recent advancements in shell-model truncation techniques, such as the Monte Carlo shell model [34] and angular momentum-projected number-conserved BCS approach [35], have made significant progress in describing deformed nuclei in heavy mass regions, offering improved computational feasibility while maintaining accuracy.

The Exact solution to the standard pairing problem, first obtained by Richardson and now referred to as the Richardson-Gaudin method, offers a promising approach for the microscopic treatment of clustering in heavy nuclei [36-43]. This method is particularly suitable for handling the large model spaces and the pairing and shell effects necessary for accurately describing heavy nuclei [44-48]. In our previous work, the deformed mean-field plus pairing model within the Richardson-Gaudin method was used to explore the quantum phase transition around neutron number N90 in the A150 mass region [49]. The analysis demonstrated the critical behavior of the shape phase transition driven by competition between deformation and pairing interactions. More recently, a new iterative algorithm was developed to find the Exact solution to the standard pairing problem within the Richardson-Gaudin method [50], which has shown excellent agreement with experimental data when applied to actinide fission nuclei isotopes [51-53].

The aim of the current study is to extend this line of inquiry by presenting a systematic study of PESs for even-even Pt, Hg, and Pb isotopes near Z=82. Our investigation leverages recent advancements in shape parametrization and adopted a macro-microscopic approach, integrating the LSD model with a Yukawa-Folded single-particle potential. The analysis focuses on the impact of pairing interactions on the shape coexistence of 170Pt, 172Hg, 174Pb nuclei, as well as 170-180Pt even-even isotopes.

2

Theoretical framework and numerical details

2.1
Deformed mean-field plus standard pairing model

The Hamiltonian of the deformed mean-field plus standard pairing model for either the proton or the neutron sector is given by H^=i=1nεin^iGiiSi+Si, (1) where the sums run over all given i-double degeneracy levels of total number n, G>0 is the overall pairing interaction strength, {εi} are the single-particle energies obtained from mean-field, such as Hartree-Fock, Woods-Saxon potential, Yukawa-Folded single-particle potential, or Nilsson model. ni=aiai+aiai is the fermion number operator for the i-th double degeneracy level, and Si+=aiai [Si=(Si+)=aiai] is the pair creation (annihilation) operator, The up and down arrows in these expressions refer to time-reversed states.

According to the Richardson-Gaudin method [36-43], the exact k-pair eigenstates of (1) with νi=0 for even systems or νi=1 for odd systems, in which i' is the label of the double degeneracy level that is occupied by an unpaired single particle can be written as |k;ξ;νi=S+(x1(ξ))S+(x2(ξ))S+(xk(ξ))|νi, (2) where |νi is the pairing vacuum state with the seniority νi that satisfies Si|νi=0, and n^i|νi=δiiνi|νi for all i. Here, ξ is an additional quantum number for distinguishing different eigenvectors with the same quantum number k and S+(xμ(ξ))=i=1n1xμ(ξ)2εiSi+, (3) in which the spectral parameters xμ(ξ)(μ=1, 2, , k) satisfy the following set of Bethe ansatz equations (BAEs): 1+GiΩixμ(ξ)2εi2G μ=1(μ)k1xμ(ξ)xμ(ξ)=0, (4) where the first sum runs over all i levels and Ωi=1δiiνi. For each solution, the corresponding eigenenergy is given by Ek(ξ)=μ=1kxμ(ξ)+νiεi. (5) In general, according to the polynomial approach in Refs. [45-48], one can find solutions of Eq. (4) by solving the second-order Fuchsian equation [44] as A(x)P(x)+B(x)P(x)V(x)P(x)=0, (6) where A(x)=i=1n(xμ(ξ)2εi) is an n-degree polynomial, B(x)/A(x)=i=1nΩixμ(ξ)2εi1G, (7) V(x) are called Van Vleck polynomials [44] of degree n-1, which are determined according to Eq. (6). They are defined as V(x)=i=0n1bixi. (8) The polynomials P(x) with zeros corresponding to the solutions of Eq. (4) is defined as P(x)=i=1k(xxi(ξ))=i=0kaixi, (9) where k is the number of pairs. bi and ai are the expansion coefficients to be determined instead of the Richardson variables xi. Furthermore, if we set ak=1 in P(x), the coefficient ak1 then equals the negative sum of the P(x) zeros, ak1=i=1kxi(ξ)=Ek(ξ).

If the value of x approaches twice the single-particle energy of a given level δ, i.e., x=2εδ, one can rewrite Eq. (6) in doubly degenerate systems with Ωi=1 as [45, 46, 48] (P(2εδ)P(2εδ))21G(P(2εδ)P(2εδ))=iδ[(P(2εδ)P(2εδ))(P(2εi)P(2εi))]2εδ2εi. (10) In Ref. [50], a new iterative algorithm is established for the exact solution of the standard pairing problem within the Richardson-Gaudin method using the polynomial approach in Eq. (10). It provides efficient and robust solutions for both spherical and deformed systems at a large scale. The key to its success is determining the initial guesses for the large-set nonlinear equations involved in a controllable and physically motivated manner. Moreover, one reduces the large-dimensional problem to a one-dimensional Monte Carlo sampling procedure, which improves the algorithm's efficiency and avoids the nonsolutions and numerical instabilities that persist in most existing approaches. Based on the new iterative algorithm, we applied the model to study the actinide nuclei isotopes, where an excellent agreement with experimental data was obtained [51-53].

2.2
The Fourier shape parametrization

Recent studies demonstrated that the developed Fourier parametrization of deformed nuclear shapes was highly effective in capturing the essential features of nuclear shapes, particularly up to the scission configuration [54, 22]. Current research indicated that combining this innovative Fourier shape parametrization with the LSD + Yukawa-Folded macro-microscopic potential-energy framework was exceptionally efficient [52, 53, 55, 56]. This work primarily adopted the macro-microscopic framework outlined in Refs. [52, 53], where the single-particle energies {ϵi} in the model Hamiltonian (1) were derived from the Yukawa-Folded potential.

The nuclear surface is expanded in terms of a Fourier series of dimensionless coordinates as follows: ρs2(z)R02=n=1[a2ncos((2n1)π2zzshz0)+a2n+1sin(2nπ2zzshz0)], (11) where ρs(z) is the distance from a surface point to the symmetry z-axis, and R0 = 1.2A1/3 fm is the radius of a corresponding spherical shape with the same volume. The shape's extension along the symmetry axis is 2z0, with the left and right ends located at zmin = zsh - z0 and zmax = zsh + z0, respectively. The parameter z0 represents half the shape's extension along the symmetry axis and is determined by volume conservation, while zsh is set such that the center of mass of the nuclear shape is at the origin of the coordinate system. Based on the convergence properties discussed in Ref. [22], the first five terms a2, …, a6 are retained as a starting point, and the parameters an are transformed into deformation parameters qn as follows: q2=a2(0)/a2a2/a2(0),q3=a3,q4=a4+(q2/9)2+(a4(0))2,q5=a5(q22)a3/10,q6=a6(q2/100)2+(a6(0))2, (12) where an(0) are the Fourier coefficients for the spherical shape. Higher-order coordinates q5 and q6 are generally set to zero within the accuracy of the current approach. The set of qi parameters has explicit physical significance in describing the shape of the fissioning nucleus: q2 denotes the elongation, q4 represents the neck parameter, and q3 indicates the left-right asymmetry.

Additionally, the non-axial deformation of nuclear shapes is described as follows, assuming that the surface cross-section at a given z-coordinate is elliptical with semi-axes a(z) and b(z): ϱs2(z, φ)=ρs2(z)1η21+η2+2ηcos(2φ), (13) where η=bab+a characterizes the non-axial deformation. Volume conservation requires that ρs2(z)=a(z)+b(z), with the condition ab=ρs2(z) ensuring volume conservation for non-axial deformations. The semi-axes are then given by: a(z)=ρs(z)1η1+η, b(z)=ρs(z)1+η1η. (14) This description of non-axial shapes using the parameters q2 and η is more general than the commonly used Bohr parametrization (β, γ). For spheroidal shapes, both descriptions are equivalent. However, as shown in Fig. 1, where the two parametrizations are compared, the periodicity of nuclear shapes by a 60° rotation angle is similar in both (q2, η) and (β, γ) planes. It is important to note that this regularity is disrupted when higher multipolarity deformations qn (n>2) are considered, making the (η, q2, q3, q4, q6) shape parametrization substantially more general than the 3-dimensional (ϵ2, ϵ4(γ), γ) parametrization used in Ref. [59, 60]. The two parametrizations coincide only in the special case of spheroidal shapes.

Fig. 1
(Color online) Relationsheep between the elongation parameter q2 and the nonaxiality parameter η [22, 54], and the traditional Bohr deformation parameters β and γ is taken from [57, 58]
pic

It is essential to stress that different points in the (β, γ), and (q2, η) planes can correspond to identical shapes when higher qn (n>2) degrees of freedom are neglected, differing only in the interchange of coordinate system axes. For example, the point (β=0.4, γ=0) corresponds to (q2 = 0.42, η = 0) in the new parametrization, representing the same shape as (β=0.4,γ=120), which corresponds to (q2 = -0.21, η = 0.16) in the new parametrization.

When analyzing potential energy landscapes that include triaxial degrees of freedom, it is crucial to avoid treating as distinct configurations points in the (q2, η) deformation plane that are merely rotational images of each other at γ=60.

In this study, the dynamic process of nuclear fission will be described in the three-dimensional deformation space (η, q2, q4) using the Fourier shape parametrization.

2.3
The potential energy

This study calculated the PESs for the isotopes 170Pt, 172Hg, and 174Pb in a three-dimensional deformation space (η, q2, q4) and analyzed the impact of pairing interactions on the shape coexistence of these isotopes. The results were obtained over the following grid points in the deformation parameter space: η[0.00, 0.20]Δη=0.02q2[0.60, 0.85]Δq2=0.05q4[0.30, 0.30]Δq4=0.03. (15) As indicated in the literature [22], the q3 degree of freedom has no significant impact on the description of shape coexistence for the isotopes discussed in this paper. Therefore, in this study, q3 was set to 0, and for each point on the PES, q4 was minimized to find the energy extremum. The potential energy of the system was calculated within the macro-microscopic approach in this work. The total energy Etotal(N, Z, qn) of a nucleus with a given deformation is calculated as Etotal(N, Z, qn)=ELD(N, Z, qn)+EB(N, Z, qn), (16) where ELD(N, Z, qn) was the macroscopic term obtained by the LSD model with proton number Z and neutron number N [61]. In the current calculation for the potential-energy surface, we just considered the energy EB(N, Z, qn) related to the shape parameter {q2, q4}. EB(N, Z, qn)=Eshell(N, Z, qn)+Epair(N, Z, qn). (17) The microscopic term consisted of the shell correction energy Eshellν(π)(N, Z, {εi}, q2, q4) proposed by Strutinsky [62, 63], and the pairing interaction energy Epairν(π)(N, Z, {εi}, q2, q4) calculated from Eq. (1). Here, ν (π) was the label of the neutron (proton) sector. In the current study, we considered 18 deformed harmonic-oscillator shells in Yukawa-Folded single-particle potential to obtain single-particle levels for the microscopic calculations. For the pairing interaction energy, we performed 29 single-particle levels around the neutron Fermi level and 22 single-particle levels around the proton Fermi level.

To validate our results and further explored the efficacy of the exactly solvable pairing model, we also calculated the PESs for the isotopes considered under the BCS approximation. The pairing interaction was determined as the difference between the BCS energy [24] and the single-particle energy sum and the average pairing energy [64]. Epair=EBCSi=1kεiE˜pair. (18) In the BCS approximation the ground-state energy of a system with an even number of particles and a monopole pairing force was given by EBCS=i=1k2εivk2G(i=1kuivi)2Gi=1kvi4, (19) where the sums run over the pairs of single-particle states contained in the pairing window defined below. The coefficients vi and ui=1vi2 were the BCS occupation amplitudes.

The average projected pairing energy, for a pairing window of width 2Ω, which is symmetric in energy with respect to the Fermi energy, is equal to E˜pair=12g˜Δ˜2+12g˜GΔ˜arctan(ΩΔ˜)log(ΩΔ˜)Δ˜+34GΩ/Δ˜1+(Ω/Δ˜)2/arctan(ΩΔ˜)14G. (20) Here g˜ was the average single-particle level density and Δ˜ the average paring gap corresponding to a pairing strength G Δ˜=2Ωexp(1Gg˜). (21)

2.4
Influence of pairing interactions on the shape coexistence of 170Pt, 172Hg and 174Pb isotopes

Figure 2 shows the PESs of 170Pt projected onto the (q2, η) plane for different pairing interaction strengths (MeV), while the proton pairing interaction strength is fixed at = 0.100 MeV. and represent the neutron and proton pairing interaction strengths (MeV), respectively. The energy is minimized in the q4 direction and q3 is set to 0 and normalized to zero energy at the ground-state value. The choice of varying from 0.03 to 0.145 MeV, and = 0.100 MeV, were based on the fact that our calculations in the next section, when employing = 0.145 MeV, and = 0.100 MeV, closely matched the experimental odd-even mass differences for the 171Pt to 180Pt isotopes. Therefore, this range was selected to study the effects of pairing strength variations on the shape coexistence. The red lines represent the corresponding (β,γ) coordinates, with γ coordinates distributed within 0≤γ≤180°. The β coordinate values are taken as 0.1, 0.2, 0.3 ..., etc.

Fig. 2
(Color online) Potential energy surface of 170Pt projected onto the (q2, η) plane under different pairing interaction strengths (MeV), while the proton pairing interaction strength is fixed at = 0.100 MeV. The energy is minimized in the q4 direction and q3 is set to 0 and normalized to zero energy at the ground-state value. The ground-state deformation is represented by a red dot
pic

In Figs. 2 (a)-(d), the PESs of 170Pt are shown for different values of the neutron pairing interaction strength , while the proton pairing interaction strength is fixed at = 0.100 MeV. The values of are: 0.030, 0.070, 0.105, and 0.145 MeV. It can be seen that the ground-state of the 170Pt isotope is located at (q2 ≈ 0.150,η = 0), indicating a prolate shape for different pairing strengths. The other minimum at (q2 ≈ -0.150,η = 0.04, γ=120°) illustrated in Fig. 2 is simply a reflection of the ground-state minimum.

It is noteworthy to highlight the existence of two distinct shape isomers in 170Pt with different pairing strengths. The first is an oblate shape isomer located at (q2 = -0.400, η = 0), with an energy approximately 3.900 MeV above the ground-state. The second is a triaxial shape isomer at (q2 ≈ 0.600, η ≈ 0.060 (γ ≈ 10°)), positioned around 4.0 MeV above the ground-state. These isomers represent the local minima on the potential energy surface that are separated from the ground-state by energy barriers, highlighting the complex deformation characteristics of the nucleus. With an increase in pairing strength, both shape isomers become shallower. When the pairing strength reaches 0.145 MeV, the oblate isomer disappears (see Fig. 2(d)).

As shown in Figs. 3(a)-(d), the PESs for different pairing interaction strengths demonstrates the evolution of the triaxial minimum at (q2 = 0.150, η = 0.020) to the oblate minimum at (q2 = 0.100, η = 0.040) as the pairing interaction strength increases. The nucleus of 172Hg is nearly γ-unstable, with the energy difference between different points in the ground-state valley not exceeding approximately 0.4 MeV. Additionally, three shape isomers are visible in the (a)-(d) maps: a prolate isomer at (q2 ≈ 0.600, η = 0), E ≈ 5.0 MeV; a triaxial isomer at (q2 ≈ 0.400, η = 0.100), E ≈ 4.0 MeV, and oblate one at (q2 ≈ -0.45, η = 0), E ≈ 4.0 MeV. These local minima are separated by energy barriers of approximately 1 MeV in height. As the pairing strength increases, all shape isomers gradually become shallower. By = 0.145 MeV and = 0.100 MeV (Fig. 3(d)), the triaxial isomer at (q2 ≈ 0.400, η = 0.100) disappeared.

Fig. 3
(Color online) Same as Fig. 2, but for 172Hg
pic

The PESs of 174Pb, as presented in Figs. 4(a)-(d), reveal that a prolate ground-state (q2 ≈ 0.150, η = 0) (in Fig. 4(a)) tend to become spherical (in Fig. 4(d)) as the pairing interaction strength increases. The shape isomers observed here are particularly interesting: a prolate shape at (q2 = 0.600, η = 0, E ≈ 5.0 MeV and a slightly triaxial oblate shape at (q2 = 0.450, η = 0.020, E ≈ 3.9 MeV in Fig. 4(a), and (b), respectively. As the pairing strength increased, both shape isomers gradually became shallower. When = 0.145, MeV, and = 0.100 MeV (Fig. 4(d)), they almost disappeard. Overall, regardless of pairing strength, there was no indication of robust shape coexistence in this nucleus.

Fig. 4
(Color online) Same as Figs. 2 and 3, but for 174Pb
pic

Figures 5 illustrate the PESs projections of 170Pt, 172Hg, and 174Pb under realistic pairing interaction strengths, with = 0.145 MeV, and = 0.100 MeV under both Exact and BCS pairing schemes.

Fig. 5
(Color online) Potential energy surfaces of 170Pt, 172Hg and 174Pb projected on the (q2, η) plane under both Exact and BCS pairing schemes, with the energy minimized in the q4 direction, q3 set to 0 and normalized to zero energy at the ground-state value. The realistic pairing interaction strengths = 0.145 MeV, and = 0.100 MeV are adopted. The ground-state deformation is represented by a red dot, while the coexistence minimum is indicated by a red cross
pic

As shown in Fig. 5, the ground-state of 170Pt is prolate, located at (q2 = 0.15, η = 0) under both the Exact and BCS pairing schemes. However, BCS pairing exhibited a shallower depth for the prolate minimum compared with Exact pairing, indicating a less pronounced prolate ground-state. Furthermore, a triaxial isomer appeared at (q2 ≈ 0.600, η ≈ 0.060 (γ ≈ 10°)) under Exact pairing, whereas it was less distinguishable in the BCS case.

The ground-state of 172Hg (Fig. 5) is found at (q2 = 0.10, η ≈ 0.04) as an oblate minimum, with another minimum at (q2 ≈ -0.100, η ≈ 0.02), which exhibits γ-unstable deformation. The PES of 172Hg provides an excellent example of an almost γ-unstable nucleus. Under Exact pairing, this γ-unstable minimum is more symmetric, with clear reflections around γ=150°, γ=30°, and γ=90°. Under BCS pairing, the γ-unstable features are less prominent, and the oblate minimum becomes more dominant. Additionally, two shape isomers are visible under Exact pairing model: a prolate isomer at (q2 ≈ 0.600, η = 0), E ≈ 4.6 MeV, and an oblate one at (q2 ≈ -0.45, η = 0), E ≈ 4.6 MeV. However, these changes were not distinguishable in the BCS case.

As shown in Figs. 5(c), the ground-state shape of 174Pb tended to be spherical. The PES under Exact pairing revealed a nearly spherical configuration with minor prolate and oblate shape isomers. In contrast, BCS pairing resulted in a more pronounced spherical minimum and diminishes the depth of shape isomers.

In summary, as the number of protons increases, the ground-state transitions from prolate for 170Pt to the coexistence of γ-unstable and oblate for 172Hg, eventually approached a nearly spherical configuration for 174Pb. The comparison between Exact and BCS pairing demonstrates that BCS pairing tends to smooth out shape coexistence and reduce the depth of shape isomer, leading to less pronounced deformation features. The differences in results between Exact and BCS pairing may be attributed to the mean-field approximation in the BCS approach, which likely simplifies the treatment of pairing interactions. This approximation is thought to smooth out shape coexistence phenomena by suppressing pairing fluctuations, energy gaps, and shell effects, potentially leading to less pronounced deformation features.

2.5
Shape coexistence analysis in the Pt isotope chain

In this paper, we investigate the PESs of the even-even 170-180Pt isotopes using the exactly solvable deformed mean-field plus pairing model. Our analysis provides a comprehensive examination of the shape coexistence phenomena across these isotopes.

The pairing interaction strength, denoted as G, serves as the sole adjustable parameter within our model. It is typically determined either through empirical formulas or by fitting to experimental odd-even mass differences [65, 66]. In this study, we determined by fitting the experimental odd-even mass differences for the 171-180Pt isotope chain and by fitting the experimental odd-even mass differences for the 174Pt to 178Pb isotonic chain. The odd-even mass differences are computed using the following expression: P(A)=Etotal(N+1, Z)+Etotal(N1, Z)2Etotal(N, Z). This quantity is highly sensitive to variations in the pairing interaction strength G [67], due to the pairing interaction between nucleons. As shown in Fig. 6, by employing =0.145 MeV and =0.100 MeV, our calculations closely reproduced the experimental odd-even mass differences for the 171-180Pt isotopes, yielding a root mean square deviation of σ=0.465 MeV. Additionally, as display in Fig. 7 for the 174Pt to 178Pb isotonic chain, the calculations closely matched the experimental odd-even mass differences, with a root mean square deviation of σ=1.192 MeV. σ=μ=1N(PμTheor.PμExpt.)2/N, (22) Here, PμTheor. and PμExpt. represent the theoretical and experimental values of the odd-even mass differences, respectively, and 𝒩 denotes the total number of data points.

Fig. 6
Odd-even mass differences (in MeV) for Pt isotopes. "Expt." represents experimental values, and "Theor." represents theoretical values. Experimental data are from [67]
pic
Fig. 7
Odd-even mass differences (in MeV) for the 174Pt to 178Pb isotonic chain. "Expt." represents experimental values, and "Theor." represents theoretical values. Experimental data are from [67]
pic

Next, we examine the PES of the 170-180Pt even-even isotopes under the determined pairing interaction strengths =0.145 MeV and =0.100 MeV. Figure 8 shows the PES projected onto the (q2, η) plane. For 170Pt, the ground-state exhibited a prolate deformation at (q2=0.15, η=0). In contrast, for 172Pt, a more deformed minimum emerged, leading to the coexistence of a triaxial shape (γ≈30°) and a nearly prolate-deformed minimum at (γ≈120°), indicative of γ-instability due to the presence of multiple low-energy configurations at different γ values. The triaxial shape is even more pronounced in 174Pt, where the ground-state is triaxial with deformation parameters (q2=0.020, η=0.10, β≈0.2, γ≈90°) and a coexisting prolate minimum at (q2=0.15, η=0). In 176Pt a γ -unstable ground-state and a prolate minimum coexisted, but by 178Pt and 180Pt, a well-deformed prolate minimum quickly developed, becoming the most pronounced prolate ground-state at the mid-shell.

Fig. 8
(Color online) Potential energy surfaces of the 170-180Pt even-even isotopes chain, projected on the (q2, η) plane using the exact pairing model, where the energy is minimized in the q4 direction with q3 set to 0, with neutron and proton pairing interaction strengths of =0.145 MeV, = 0.100 MeV. The ground-state deformation is represented by a red dot, while the coexistence minimum is indicated by a red cross
pic

The findings of this study are broadly consistent with the results of Ref. [20], which studied the 172-194Pt isotopic chain in the framework of the interacting boson model and self-consistent HFB calculation using the Gogny-D1S interaction. Both studies identified shape coexistence in the 172-176Pt region, with γ-unstable minima and triaxial shapes in 174Pt. Additionally, both studies showed the dominance of prolate deformation in 178Pt, and 180Pt, with the prolate minimum becoming the most pronounced ground state at the mid-shell.

It is noteworthy that a triaxial shape isomer exists for 170-174Pt, characterized by (q2 ≈ 0.600, η ≈ 0.060 (γ ≈ 10°)), and positioned approximately 5.0 MeV above the ground-state. However, this triaxial shape isomer vanishes for 176-180Pt.

3

Conclusion

In this study, we systematically investigated the shape coexistence phenomenon in isotopes near the magic proton number of Z = 82, focusing specifically on the nuclei 170 Pt, 172 Hg, and 174Pb, as well as the Pt isotopic chain from 170Pt to 180Pt. Our analysis, using a macro-microscopic approach that combines the LSD model with a Yukawa-Folded potential and pairing corrections, revealed significant insights into the impact of pairing interactions on nuclear shape evolution.

The PES of 170Pt revealed a prolate ground-state with additional triaxial and oblate shape isomers. Both shape isomers become progressively shallower with increasing neutron pairing strength , and the oblate isomer vanishes at = 0.145 MeV, indicating a significant dependence of shape isomers on pairing strength. The ground-state deformation of 172Hg transitions from triaxial to oblate with increasing , reflecting its nearly γ-unstable nature. Three shape isomers (prolate, triaxial, and oblate) were observed, with energy barriers separating these configurations. As increased, the triaxial isomer disappeared at = 0.145 MeV, demonstrating the impact of pairing interactions on shape stability. 174Pb exhibited a prolate ground-state that became increasingly spherical with stronger pairing interactions. While shape isomers are present at weaker pairing strengths, their prominence diminishes significantly, and robust shape coexistence was not observed in this nucleus.

For realistic pairing interaction, the ground-state shapes transition from prolate in 170Pt to a coexistence of γ-unstable and oblate shapes in 172Hg, ultimately approaching spherical symmetry in 174Pb. This progression highlights the interplay between proton number and pairing interactions in shaping nuclear deformation. The comparison between Exact and BCS pairing for realistic 170Pt, 172Hg, and 174Pb demonstrated that BCS pairing tends to smooth out shape coexistence and reduce the depth of shape isomers, leading to less pronounced deformation features.

These findings emphasize the critical role of pairing interactions in shaping nuclear deformation landscapes and shape coexistence, offering deeper insights into the structural evolution of nuclei near the mid-shell region. This study contributes valuable theoretical perspectives to the understanding of nuclear shape phenomena and the influence of pairing interactions on nuclear dynamics.

Based on the analysis of the PESs for the even-even 170-180Pt isotopes, the results show significant shape evolution across the isotopic chain. For 170Pt, the ground-state exhibited prolate deformation, with deformation parameters. However, for 172Pt, a more deformed minimum appears, leading to the coexistence of a triaxial shape and a nearly prolate-deformed minimum. The triaxial shape becomes even more pronounced in 174Pt, where the ground-state is triaxial with deformation parameters, coexisting with a prolate minimum. For 176Pt, a γ-unstable ground-state coexists with a prolate minimum. By 178Pt, and 180Pt, a well-deformed prolate minimum develops rapidly, becoming the most pronounced prolate ground-state in the mid-shell.

These results highlight the complex shape evolution in the Pt isotopes, with shape coexistence and γ-instability playing significant roles in the nuclear structure evolution, particularly around the mid-shell region where prolate deformation dominates.

References
1.K. Heyde, and J. L. Wood,

Publisher's Note: Shape coexistence in atomic nuclei

. Rev. Mod. Phys. 83, 1467 (2011). https://doi.org/10.1103/RevModPhys.83.1655
Baidu ScholarGoogle Scholar
2.J. Y. Jia, G. Giacalone, B. Bally, et al.,

Imaging the initial condition of heavy-ion collisions and nuclear structure across the nuclide chart

. Nucl. Sci. Tech. 35, 220 (2024). https://doi.org/10.1007/s41365-024-01589-w
Baidu ScholarGoogle Scholar
3.M.Q. Ding, D.Q. Fang, and Y. G. Ma,

Neutron skin and its effects in heavy-ion collisions

. Nucl. Sci. Tech. 35, 211 (2024). https://doi.org/10.1007/s41365-024-01584-1
Baidu ScholarGoogle Scholar
4.B. S. Cai, and C. X. Yuan,

Random forest-based prediction of decay modes and half-lives of superheavy nuclei

. Nucl. Sci. Tech. 34, 204 (2023). https://doi.org/10.1007/s41365-023-01354-5
Baidu ScholarGoogle Scholar
5.J. L. Huang, H. Wang, Y. G. Huang, et al.,

Prediction of (n, 2n) reaction cross-sections of long-lived fission products based on tensor model

. Nucl. Sci. Tech. 35, 184 (2024). https://doi.org/10.1007/s41365-024-01556-5
Baidu ScholarGoogle Scholar
6.Y. Liu, F. R. Xu, and Z. B. Cao,

Shape Coexistence in Selenium Isotopes

. Nucl. Phys. Rev. 27, 2 (2010). https://doi.org/10.11804/NuclPhysRev.27.02.146
Baidu ScholarGoogle Scholar
7.J. Rainwater,

Nuclear energy level argument for a spheroidal nuclear model

. Phys. Rev. 79, 432 (1950). https://doi.org/10.1103/PhysRev.79.432
Baidu ScholarGoogle Scholar
8.A. Bohr and B. R. Mottelson,

Collective and individual-particle aspects of nuclear structure

, Dan. Mat. Fys. Medd. 27, 16 (1953).
Baidu ScholarGoogle Scholar
9.A. Bohr, and B. R. Mottelson,

Moments of Inertia of Rotating Nuclei

. Dan. Mat. Fys. Medd. 30, 1 (1955).
Baidu ScholarGoogle Scholar
10.A. Arima, H. Horie,

Configuration mixing and magnetic moments of odd nuclei

. Prog. Theor. Phys. 12, 623 (1954). https://doi.org/10.1143/PTP.12.623
Baidu ScholarGoogle Scholar
11.S. G. Nilsson,

Binding states of individual nucleons in strongly deformed nuclei

. Dan. Mat. Fys. Medd. 29, 16 (1955).
Baidu ScholarGoogle Scholar
12.H. Morinaga,

Interpretation of some of the excited states of 4n self-conjugate

. Phys. Rev. 101, 254 (1956). https://doi.org/10.1103/PhysRev.101.254
Baidu ScholarGoogle Scholar
13.J. P. Elliott,

Collective motion in the nuclear shell model

. I. Classification schemes for states of mixed configurations. Proc. R. Soc. A 245, 128(1958). https://doi.org/10.1098/rspa.1958.0072
Baidu ScholarGoogle Scholar
14.P. Möller, A. J. Sierk, R. Bengtsson et al.,

Global Calculation of Nuclear Shape Isomers

. Phys. Rev. Lett. 103 212501 (2009). https://doi.org/10.1103/PhysRevLett.103.212501
Baidu ScholarGoogle Scholar
15.M. Siciliano, I. Zanon, A. Goasduff et al.,

Shape coexistence in the neutron-deficient 188Hg investigated via lifetime measurements

. Phys. Rev. C. 102, 014318 (2020). https://doi.org/10.1103/PhysRevC.102.014318
Baidu ScholarGoogle Scholar
16.R. Julin, T. Grahn, J. Pakarinen et al.,

In-beam spectroscopic studies of shape coexistence and collectivity in the neutron-deficient Z ≈ 82 nuclei

. J. Phys. G.43, 024004 (2016). https://doi.org/10.1088/0954-3899/43/2/024004
Baidu ScholarGoogle Scholar
17.M. Bender, P. H. Heenen, and P. G. Reinhard,

Self-consistent mean-field models for nuclear structure

. Rev. Mod. Phys. 75, 121 (2003). https://doi.org/10.1103/RevModPhys.75.121
Baidu ScholarGoogle Scholar
18.T. Nikšić, D. Vretenar, P. Ring et al.,

Shape coexistence in the relativistic Hartree-Bogoliubov approach

. Phys. Rev. C. 65, 054320 (2002). https://doi.org/10.1103/PhysRevC.65.054320
Baidu ScholarGoogle Scholar
19.F. Z. Xing, J. P. Cui, Y. H. Gao et al.,

Structure and α decay of superheavy nucleus 296Og

. Nucl. Phys. Rev. 40, 4 (2023). https://doi.org/10.11804/NuclPhysRev.40.2023059
Baidu ScholarGoogle Scholar
20.J. E. García-Ramos, K. Heyde, L. M. Robledo et al.,

Shape evolution and shape coexistence in Pt isotopes: Comparing interacting boson model configuration mixing and Gogny mean-field energy surfaces

. Phys. Rev. C 98, 034313 (2014). https://doi.org/10.1103/PhysRevC.81.024310
Baidu ScholarGoogle Scholar
21.J. E. García-Ramos, and K. Heyde,

Nuclear shape coexistence: A study of the even-even Hg isotopes using the interacting boson model with configuration mixing

. Phys. Rev. C 89, 014306 (2014). https://doi.org/10.1103/PhysRevC.89.014306
Baidu ScholarGoogle Scholar
22.K. Pomorski, B. Nerlo-Pomorska, A. Dobrowolski, et al.

Shape isomers in Pt, Hg and Pb isotopes with N≤126

. Eur. Phys. J. A, 56, 107 (2020). https://doi.org/10.1140/epja/s10050-020-00115-x
Baidu ScholarGoogle Scholar
23.T. Kibedi, and C. M. Baglin,

ENSDEF

(2010), http://www.nndc.bnl.gov/ensdf
Baidu ScholarGoogle Scholar
24.J. Bardeen, L. N. Cooper, and J. R. Schrieffer,

Theory of Superconductivity

. Phys. Rev. 108, 1175 (1957). https://doi.org/10.1103/PhysRev.108.1175
Baidu ScholarGoogle Scholar
25.A. Bohr, B. R. Mottelson, and D. Pines,

Possible analogy between the excitation spectra of nuclei and those of the superconducting metallic state

. Phys. Rev. 110, 4 (1958). https://doi.org/10.1103/PhysRev.110.936
Baidu ScholarGoogle Scholar
26.S. T. Belyaev,

Effect of pairing correlations on nuclear properties

. Dan. Mat. Fys. Medd. 31, 11 (1959).
Baidu ScholarGoogle Scholar
27.Y. Z. Wang, F. Z. Xing, J. P. Cui et al.,

Roles of tensor force and pairing correlation in two-proton radioactivity of halo nuclei

. Chin. Phys. C 47, 084101 (2023). https://doi.org/10.1088/1674-1137/acd680
Baidu ScholarGoogle Scholar
28.Y. Z. Wang, F. Z. Xing, W. H. Zhang et al.,

Tensor force effect on two-proton radioactivity of 18Mg and 20Si

. Phys. Rev. C 110, 064305 (2024). https://doi.org/10.1103/PhysRevC.110.064305
Baidu ScholarGoogle Scholar
29.Y. Z. Wang, L. Yang, C. Qi et al.,

Pairing effects on bubble nuclei

. Chin. Phys. Lett. 36, 032101 (2019). https://doi.org/10.1088/0256-307X/36/3/032101
Baidu ScholarGoogle Scholar
30.Y. Z. Wang, Z. Y. Hou, Q. L. Zhang et al.,

Effect of a tensor force on the proton bubble structure of 206Hg

. Phys. Rev. C 91, 017302 (2015). https://doi.org/10.1103/PhysRevC.91.017302
Baidu ScholarGoogle Scholar
31.J. Rissanen, R. M. Clark, A. O. Macchiavelli et al.,

Effect of the reduced pairing interaction on α-decay half-lives of multi-quasiparticle isomeric states

. Phys. Rev. C 90, 044324 (2014). https://doi.org/10.1103/PhysRevC.90.044324
Baidu ScholarGoogle Scholar
32.N. Sandulescu, G. F. Bertsch,

Accuracy of BCS-based approximations for pairing in small Fermi systems nuclei

. Phys. Rev. C. 78, 064318 (2008). https://doi.org/10.1103/PhysRevC.78.064318
Baidu ScholarGoogle Scholar
33.I. Talmi, Simple Models of Complex Nuclei. (Harwood Academic Publishers, Switzerland) (1993). https://doi.org/10.1201/9780203739716.
34.T. Otsuka, Y. Tsunoda, T. Abe et al.,

Underlying structure of collective bands and self-organization in quantum systems

. Phys. Rev. Lett. 123, 222502 (2019). https://doi.org/10.1103/PhysRevLett.123.222502
Baidu ScholarGoogle Scholar
35.Y. X. Yu, Y. Lu, G. J. Fu, et al.,

Nucleon-pair truncation of the shell model for medium-heavy nuclei

. Phys. Rev. C 106, 044309 (2022). https://doi.org/10.1103/PhysRevC.106.044309
Baidu ScholarGoogle Scholar
36.R. W. Richardson,

A restricted class of exact eigenstates of the pairing-force Hamiltonian

. Phys. Lett. 3, 3277 (1963). https://doi.org/10.1016/0031-9163(63)90259-2
Baidu ScholarGoogle Scholar
37.R. W. Richardson,

Application to the exact theory of the pairing model to some even isotopes of lead

. Phys. Lett. 5, 82 (1963). https://doi.org/10.1016/S0375-9601(63)80039-0
Baidu ScholarGoogle Scholar
38.R. W. Richardson and N. Sherman,

Exact eigenstates of the pairing-force Hamiltonian

. Nucl. Phys. 52, 221 (1964). https://doi.org/10.1016/0029-5582(64)90687-X
Baidu ScholarGoogle Scholar
39.R. W. Richardson and N. Sherman,

Pairing models of Pb206, Pb204 and Pb202

. Nucl. Phys. 52, 253 (1964). https://doi.org/10.1016/0029-5582(64)90690-X
Baidu ScholarGoogle Scholar
40.M. Gaudin,

Diagonalization of a class of spin Hamiltonians

. Phys. J. 37, 1087 (1976).
Baidu ScholarGoogle Scholar
41.F. Pan, J. P. Draayer, and W. E. Ormand,

A particle-number-conserving solution to the generalized pairing problem

. Phys. Lett. B. 422, 1 (1998). https://doi.org/10.1016/S0370-2693(98)00034-3
Baidu ScholarGoogle Scholar
42.J. Dukelsky, C. Esebbag, and S. Pittel,

Electrostatic mapping of nuclear pairing

. Phys. Rev. Lett. 88, 062501 (2002). https://doi.org/10.1103/PhysRevLett.88.062501.
Baidu ScholarGoogle Scholar
43.J. Dukelsky, S. Pittel, G. Sierra,

Colloquium: Exactly solvable Richardson-Gaudin models for many-body quantum systems

. Rev. Mod. Phys. 76, 643 (2004). https://doi.org/10.1103/RevModPhys.76.643
Baidu ScholarGoogle Scholar
44.X. Guan, K. D. Launey, M. X. Xie et al.,

Heine-Stieltjes correspondence and the polynomial approach to the standard pairing problem

. Phys. Rev. C 86, 024313 (2012). https://doi.org/10.1103/PhysRevC.86.024313
Baidu ScholarGoogle Scholar
45.A. Faribault, O. E. Araby, C. Sträter et al.,

Gaudin models solver based on the correspondence between Bethe ansatz and ordinary differential equations

. Phys. Rev. B 83, 235124 (2011). https://doi.org/10.1103/PhysRevB.83.235124
Baidu ScholarGoogle Scholar
46.O. El Araby, V. Gritsev, and A. Faribault,

Bethe ansatz and ordinary differential equation correspondence for degenerate Gaudin models

. Phys. Rev. B 85, 115130 (2012). https://doi.org/10.1103/PhysRevB.85.115130
Baidu ScholarGoogle Scholar
47.X. Guan, K. D. Launey, M. X. Xie et al.,

Numerical algorithm for the standard pairing problem based on the Heine-Stieltjes correspondence and the polynomial approach

. Comp. Phys. Commun. 185, 2714 (2014). https://doi.org/10.1016/j.cpc.2014.05.023
Baidu ScholarGoogle Scholar
48.C. Qi, and T. Chen,

Exact solution of the pairing problem for spherical and deformed systems

. Phys. Rev. C 92, 051304(R) (2015). https://doi.org/10.1103/PhysRevC.92.051304
Baidu ScholarGoogle Scholar
49.X. Guan, H. C. Xu, F. Pan et al.,

Ground state phase transition in the Nilsson mean-field plus standard pairing model

. Phys. Rev. C 94, 024309 (2016). https://doi.org/10.1103/PhysRevC.94.024309
Baidu ScholarGoogle Scholar
50.X. Guan and C. Qi,

An iterative approach for the exact solution of the pairing Hamiltonian

. Comp. Phys. Comm. 275, 108310 (2022). https://doi.org/10.1016/j.cpc.2022.108310
Baidu ScholarGoogle Scholar
51.X. Guan, Y. Xin, Y. J. Chen et al.,

Impact of pairing interactions on fission in the deformed mean-field plus standard pairing model

. Phys. Rev. C 104, 044329 (2021). https://doi.org/10.1103/PhysRevC.104.044329
Baidu ScholarGoogle Scholar
52.X. Guan, T. C. Wang, W. Q. Jiang et al.,

Impact of the pairing interaction on fission of U isotopes

. Phys. Rev. C 107, 034307 (2023). https://doi.org/10.1103/PhysRevC.107.034307
Baidu ScholarGoogle Scholar
53.X. Guan, J. H. Zheng, and M. Y. Zheng,

Pairing effects on the fragment mass distribution of Th, U, Pu, and Cm isotopes

. Nucl. Sci. Tech. 34, 173 (2023). https://doi.org/10.1007/s41365-023-01316-x
Baidu ScholarGoogle Scholar
54.C. Schmitt, K. Pomorski, B. K. Nerlo-Pomorska et al.,

Performance of the Fourier shape parametrization for the fission process

. Phys. Rev. C 95, 034612 (2017). https://doi.org/10.1103/PhysRevC.95.034612
Baidu ScholarGoogle Scholar
55.K. Pomorski, J. M. Blanco, P. V. Kostryukov et al.,

Fission fragment mass yields of Th to Rf even-even nuclei

. Chin. Phys. C 45, 054109 (2021). https://doi.org/10.1088/1674-1137/abec69
Baidu ScholarGoogle Scholar
56.L. L. Liu, Y. J. Chen, X. Z. Wu et al.,

Analysis of nuclear fission properties with the Langevin approach in Fourier shape parametrization

. Phys. Rev. C 103, 044601 (2021). https://doi.org/10.1103/PhysRevC.103.044601
Baidu ScholarGoogle Scholar
57.A. Bohr,

The coupling of nuclear surface oscillations to the motion of individual nucleons

. Dan. Mat. Fys. Medd. 26, 14 (1952).
Baidu ScholarGoogle Scholar
58.T. Kaniowska, A. Sobiczewski, K. Pomorski et al.,

Microscopic inertial functions for nuclei in the barium region

. Nucl. Phys. A 274, 151 (1976). https://doi.org/10.1016/0375-9474(76)90233-5
Baidu ScholarGoogle Scholar
59.S. G. Rohoziáski, and A. Sobiczewski,

Hexadecapole Nuclear Potential for Non-Axial Shapes

, Acta Phys. Pol. B 12, 1001 (1981). https://doi.org/10.1016/0003-4916(82)90273-1
Baidu ScholarGoogle Scholar
60.P. Möller, A. J. Sierk, R. Bengtsson et al.,

Data tables

. Atom. Data Nucl. Data Tables 98, 149 (2012). https://doi.org/10.1103/PhysRevC.79.064304
Baidu ScholarGoogle Scholar
61.K. Pomorski, and J. Dudek,

Nuclear liquid-drop model and surface-curvature effects

. Phys. Rev. C 67, 044316 (2003). https://doi.org/10.1103/PhysRevC.67044316.
Baidu ScholarGoogle Scholar
62.V. M. Strutinsky,

Shell effects in nuclear masses and deformation energies

. Nucl. Phys. A 95, 420 (1967). https://doi.org/10.1016/0375-9474(67)90510-6
Baidu ScholarGoogle Scholar
63.V. M. Strutinsky,

'Shells' in deformed nuclei

. Nucl. Phys. A 122, 1 (1968). https://doi.org/10.1016/0375-9474(68)90699-4
Baidu ScholarGoogle Scholar
64.S. G. Nilsson, C. F. Tsang, A. Sobiczewski et al.

On the nuclear structure and stability of heavy and superheavy elements

. Nucl. Phys. A. 95, 1 (1969). https://doi.org/10.1016/0375-9474(69)90809-4
Baidu ScholarGoogle Scholar
65.Y. Sun,

Projection techniques to approach the nuclear many-body problem

. Phys. Scr. 91, 043005 (2016). https://doi.org/10.1088/0031-8949/91/4/043005
Baidu ScholarGoogle Scholar
66.M. Bender, K. Rutz, P. G. Reinhard et al.,

Pairing gaps from nuclear mean-field models

. Eur. Phys. J. A 8, 59 (2000). https://doi.org/10.1007/s10050-000-4504-z
Baidu ScholarGoogle Scholar
67.

U. S. National Nuclear Data Center

: http://www.nndc.bnl.gov/
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.