Next Article in Journal
Extreme Ultraviolet Radiation Sources from Dense Plasmas
Previous Article in Journal
Search for Weak Side Branches in the Electromagnetic Decay Paths of the 6526-keV 10+ Isomer in 54Fe
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Shape Coexistence in Even–Even Nuclei: A Theoretical Overview

by
Dennis Bonatsos
1,*,
Andriana Martinou
1,
Spyridon K. Peroulis
1,
Theodoros J. Mertzimekis
2 and
Nikolay Minkov
3
1
Institute of Nuclear and Particle Physics, National Centre for Scientific Research “Demokritos”, 15310 Aghia Paraskevi, Attiki, Greece
2
Department of Physics, Zografou Campus, National and Kapodistrian University of Athens, 15784 Athens, Greece
3
Institute of Nuclear Research and Nuclear Energy, Bulgarian Academy of Sciences, 72 Tzarigrad Road, 1784 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Atoms 2023, 11(9), 117; https://doi.org/10.3390/atoms11090117
Submission received: 10 July 2023 / Revised: 18 August 2023 / Accepted: 21 August 2023 / Published: 30 August 2023
(This article belongs to the Section Nuclear Theory and Experiments)

Abstract

:
The last decade has seen a rapid growth in our understanding of the microscopic origins of shape coexistence, assisted by the new data provided by the modern radioactive ion beam facilities built worldwide. Islands of the nuclear chart in which shape coexistence can occur have been identified, and the different microscopic particle–hole excitation mechanisms leading to neutron-induced or proton-induced shape coexistence have been clarified. The relation of shape coexistence to the islands of inversion, appearing in light nuclei, to the new spin-aligned phase appearing in N = Z nuclei, as well as to shape/phase transitions occurring in medium mass and heavy nuclei, has been understood. In the present review, these developments are considered within the shell-model and mean-field approaches, as well as by symmetry methods. In addition, based on systematics of data, as well as on symmetry considerations, quantitative rules are developed, predicting regions in which shape coexistence can appear, as a possible guide for further experimental efforts that can help in improving our understanding of the details of the nucleon–nucleon interaction, as well as of its modifications occurring far from stability.

1. Introduction

Shape coexistence (SC) in atomic nuclei refers to the special case in which the ground state band of a nucleus lies close in energy with another band possessing a radically different structure. SC has been suggested in 1956 by Morinaga [1], in order to interpret 0 + states corresponding to deformed shapes lying close in energy with the spherical ground state bands of 16 O, 20 Ne, and 24 Mg. Many more examples of SC have been suggested since then. Manifestations of SC in odd-mass nuclei have been reviewed in 1983 by Heyde et al. [2], while cases observed in even–even nuclei have been reviewed in 1992 by Wood et al. [3]. A review of both experimental observations and relevant theoretical developments has been given in 2011 by Heyde and Wood [4]; see also [5] for an account of the historical development of the subject, as well as the volume [6,7] for a compilation of recent work. The advent of radioactive ion beam facilities in recent years largely increased the number of nuclei available for experimental investigation. Experimental manifestations of SC have been recently (2022) reviewed by Garrett et al. [8]. In parallel, the progress made in theoretical calculations, both in the mean-field and the shell-model frameworks, has accumulated many new results over the last decade, including massive calculations over entire series of isotopes, or even large regions of the nuclear chart involving several series of isotopes. It is the scope of the present review to summarize these recent theoretical developments, draw conclusions from the existing results and pave the way towards future developments.
Nobel prizes in nuclear structure have been awarded to Wigner, Goeppert-Mayer, and Jensen in 1963 [9], as well as to Bohr, Mottelson, and Rainwater in 1975 [10], corresponding to breakthrough developments of our understanding of atomic nuclei. Wigner was rewarded for the use of symmetries in nuclear structure, Goeppert-Mayer and Jensen for the introduction of the spherical shell model, and Bohr, Mottelson, and Rainwater for the introduction of the collective model describing nuclear deformation. Recent developments along all these three lines will be considered in the present review, which will be limited to even–even nuclei, for the sake of size.

1.1. Statement of Purpose

The present review considers SC of the ground state band with K π = 0 + bands lying close in energy and having radically different structures. The term SC has also been used in the literature to refer to pairs and/or bunches of high-lying bands (see, for example, [11,12,13,14,15,16,17,18]), isomeric states (see, for example, [19,20,21,22,23,24]), superdeformed bands (see, for example, [25,26,27]), or bands of negative parity, related to the octupole degree of freedom (see, for example, [28,29,30,31]). Such bands are not considered here.
The main purpose of the present review is to focus attention on the many theoretical advances made in the description of SC since the last authoritative review, by Heyde and Wood in 2011 [4]. Since the experimental work on SC has been recently reviewed by Garrett et al. in 2022 [8], our emphasis has been placed on the theoretical developments, with a number of experimental papers cited either for historical purposes or thanks to the theoretical calculations contained in them.
The progress made on the theoretical front of SC over the last decade has been tremendous. The advent of both theoretical methods and computational facilities already allows for systematic calculations over extended regions of the nuclear chart using the same set of model parameters for all nuclei, in contrast to old calculations in which individual fits of the parameters had to be made for each nucleus separately. This extends the predictive power of the models beyond the experimentally known nuclei, thus providing predictions which can guide further experimental efforts.
The present work has been focused on spectra and B(E2) transition rates, as well as on potential energy surfaces (PESs). A review of additional experimental quantities (transfer cross sections, decay properties, charge radii) related to the presence of SC has been given recently by Garrett et al. [8]. It should be emphasized that PESs alone do not suffice for establishing SC, since they do not contain all the necessary information, which has to be supplied by spectra, B(E2)s, and/or other experimental quantities.
For the sake of simplicity and space, only even–even nuclei are considered in this review. Relevant literature has been considered up to December 2022, with few exceptions. A similar review on odd-A nuclei is called for.

1.2. Outline

In the first part (Section 2) of this article, the theoretical methods used for the description of even–even nuclei are briefly reviewed. The purpose of this part is to introduce the language needed and the basic concepts used, practically using no equations. The readers are directed to the relevant authoritative reviews for technical details. The acronyms used throughout this review are listed in Abbreviations, while the theoretical methods used are listed in Appendix A, along with the section in which they are defined. Some additional methods are also briefly described in Appendix A.
In the second part of this article (Section 3, Section 4, Section 5, Section 6, Section 7, Section 8, Section 9, Section 10, Section 11 and Section 12), the series of isotopes from Rn ( Z = 86 ) to Be ( Z = 4 ) are studied separately. An effort has been made to keep the study of each series of isotopes self-contained, at the expense of some repetition. Important figures from the articles considered are not reproduced in the body of this work, in order to keep its size finite. Detailed citations are given instead, so that the interested reader can easily reach them in this era of easy access to articles in an additional window on the computer in use. A brief summary is presented at the end of each subsection, regarding the relevant series of isotopes. One can choose to consider only the atomic numbers of their own interest, without any need to go through the rest. The only sections in which the properties of several series of isotopes are discussed together are Section 5.5, Section 8.5, and Section 9.5, in which the relation between SC and SPT is discussed in the regions ( N = 90 , Z = 60 ) , ( N = 60 , Z = 40 ) , and ( N = 40 , Z = 34 ) , respectively, as well as the Section 11 and Section 12, in which N = Z nuclei and the islands of inversion are discussed, respectively.
In the third part (Section 13, Section 14 and Section 15) of this article, the information presented in the summaries of the second part is gathered together and discussed in Section 13, in order to reach the final conclusions of the present study in Section 14, as well as to make proposals for further work in Section 15.

2. Theoretical Approaches to Shape Coexistence

2.1. The Nuclear Shell Model

The nuclear shell model [32,33,34,35,36,37] was introduced in 1949 in order to explain the appearance of the magic numbers 2, 8, 20, 28, 50, 82, 126, …, associated with nuclei exhibiting extreme stability. It is based on the three-dimensional isotropic harmonic oscillator (3D-HO), which possesses the SU(3) symmetry and is characterized by closed shells at 2, 8, 20, 40, 70, 112, 168, … protons or neutrons [38,39,40]. To the 3D-HO potential the spin–orbit interaction is added, which destroys the SU(3) symmetry beyond the sd nuclear shell, leading from the 3D-HO magic numbers 2, 8, 20, 40, 70, 112, 168, … to the spherical shell-model magic numbers 2, 8, 20, 28, 50, 82, 126, …. The spectroscopic notation, labeling the shells with 0, 1, 2, 3, 4, 5, 6, 7, … oscillator quanta by the letters s, p, d, f, g, h, i, j, … is used. The Woods–Saxon potential [41] can be used alternatively. It may be noticed that the idea of a shell model for atomic nuclei based on the 3D-HO has already existed since the 1930s [42,43,44], but it was the crucial addition of the spin–orbit interaction [32,34,35,36] to it, which led to the successful prediction of the magic energy gaps.
In this primitive version of the shell model, also called the independent particle model, degenerate energy levels within each shell occur. In order to break the degeneracies, one has to consider that in addition to the two-body nucleon–nucleon interaction creating the spherical mean field, a “residual” two-body interaction, appearing when two or more nucleons are present in the same shell, has to be taken into account. As a consequence, the computational effort needed for diagonalizing the relevant Hamiltonian increases dramatically with the size of the shell.
A first compromise in order to reduce the size of the calculation is to limit it to the valence protons and neutrons outside closed shells, with the latter being treated as an inert core. Thus, calculations have to use a limited valence space and an appropriate effective interaction adapted to it. The interaction of the valence nucleons with the core is included in the “residual” interaction. A review of these developments has been given by Caurier et al. [45], while a review of the shell-model applications on SC can be found in [6,46]. Applications of the shell-model approach to specific nuclei [47,48,49,50,51] will be discussed in the relevant sections regarding the corresponding series of isotopes.
Computational limitations have restricted the application of the shell model to light- and medium-mass nuclei near to closed shells. A major step towards large-scale shell-model calculations has been taken by the use of the quantum Monte Carlo technique [52,53,54], which led to the formulation of the Monte Carlo shell model (MCSM) [55,56,57] scheme. Within this scheme, the size of the calculation is radically reduced by selecting out of the huge number of many-body basis states a much smaller number of highly selected many-body basis states, in what can be considered as an “importance truncation” [58]. The question of uncertainty quantification, which is of obvious importance for predictions in experimentally unknown regions, has also been considered [59] within this approach. Applications of this technique to specific nuclei, including state-of-the-art calculations over extended series of isotopes, can be found in [58,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75]. These will be discussed in the relevant sections regarding the corresponding series of isotopes.
An alternative path has been taken through no-core shell-model calculations [76,77]. Within this scheme, one is trying to reduce the size of the calculation by taking advantage of the underlying SU(3) symmetry of the 3D-HO, thus ending up with a symmetry-adapted no-core shell model (SANCSM) [78,79,80,81,82], which will be further discussed in the SU(3) Section 2.4.
Within the shell-model framework, SC can be accounted for by two-particle–two-hole (2p-2h) excitations across major shells [83,84,85,86,87]. In a nucleus lying near to closed shells, the excitation of two protons or two neutrons above the Fermi level effectively adds two more valence pairs (a pair of particles above the Fermi surface and a pair of holes below the Fermi surface), thus largely increasing the collectivity of the nucleus, which is based on the numbers of valence protons and valence neutrons outside closed shells. In even–even nuclei, 2p-2h and 4p-4h excitations are the most common ones. Mixing [88,89,90] of the ground state configuration (which has no p-h excitations) and the excited configuration leads then to the two coexisting bands.
The last statement is the basis of the two-state mixing model (see Section 2.2 of [8]), widely used [88,89,90,91,92,93,94,95,96,97,98,99,100,101] by experimentalists in order to directly detect SC from the data. The two-state mixing model is based on the assumption that two observed states bearing the same angular momentum and parity can be understood as linear combinations of two unperturbed states, with coefficients (mixing amplitudes) depending on a mixing angle, which is a measure of the occurring mixing and can be determined from the spectroscopic data.

2.1.1. The Tensor Force

A major step forward in understanding shell evolution in exotic nuclei was made by Otsuka et al. in 2005 [102] with the introduction of the tensor force, based on the exchange of π and ρ mesons, in the framework of the meson-exchange process introduced by Yukawa [103] in 1935. The main effect caused by the tensor force is the influence of protons on the neutron gaps, as well as the influence of neutrons on the proton gaps. Using for the nuclear orbitals the symbols j > = l + 1 / 2 and j < = l 1 / 2 , it turns out that the interaction between orbitals with different >, < subscripts is attractive, while the interaction between orbitals with the same subscript is repulsive.
Let us see how the tensor force influences the shell gaps through an example [102]. Below the Z = 28 gap one finds the proton orbital 1 f 7 / 2 > , while above it the proton orbital 1 f 5 / 2 < is lying. Lower than both of them in energy, the neutron orbital 1 g 9 / 2 > lies, since the neutron potential is deeper than the proton potential, see, for example, Figure 6.4 of [104]. The tensor force between 1 g 9 / 2 > and 1 f 7 / 2 > is repulsive, thus 1 f 7 / 2 > is pushed up in energy by the tensor component of the proton–neutron interaction. On the contrary, the tensor force between 1 g 9 / 2 > and 1 f 5 / 2 < is attractive, thus 1 f 5 / 2 < is pushed down in energy by the tensor component of the proton–neutron interaction. As a result, the two proton orbitals are approaching each other, and therefore the Z = 28 gap gets reduced, thus making the creation of proton particle–hole excitations across the Z = 28 gap easier and therefore facilitating the appearance of SC [105].
Recent reviews on the properties of the tensor force and its role in the occurrence of SC and in shell evolution can be found in [105,106,107], respectively. The terms type I shell evolution, corresponding to shell evolution along a series of isotopes or isotones, and type II shell evolution, depending on the balance of the central and tensor components of the effective nucleon–nucleon interaction, and influencing the appearance of SC, have been coined [105]. The shape of the tensor force as a function of the nuclear radius is given for various interaction models in Figure 3 of [102], as well as in Figure 15 of [107]. It is interesting to compare the shape of the tensor force due to the π and ρ meson exchange to the shape of the central force, see Figure 1 of [64]. While the central force has a Gaussian shape with minimum at r = 0 , the tensor force has a hard core and a minimum at r > 0 , roughly resembling the Kratzer potential [108].

2.2. Self-Consistent Mean-Field Models

In the shell-model approach one assumes that the nuclear mean field has the form of a specific single-particle potential and then performs configuration mixing involving all possible states appearing in the area of its application, called the shell-model space. An alternative approach is to start with the self-consistent determination of a nuclear mean field fitted to nuclear structure data [109]. The simplest variational method for the determination of the wave functions of a many-body system is the Hartree–Fock (HF) method [110,111]. In the case of nuclei, the pairing interaction has to be taken into account, leading into the Hartree–Fock–Bogoliubov (HFB)approach [110,111]. The inclusion of pairing is simplified by the BCS approximation, which allows the formation of pairs of degenerate states connected through time reversal, as in the BCS theory of superconductivity in solids [112]. Three different choices have been made for the effective nucleon–nucleon interaction: the Gogny interaction [113,114,115] and the Skyrme interaction [116,117,118], used in a non-relativistic framework, and the relativistic mean-field (RMF) models [119,120,121,122,123,124,125,126,127,128], the latter having the advantage of providing a natural explanation of the spin–orbit force. A compilation of various parametrizations of the effective interactions in use can be found in Table I of the review article [109]. Extensive codes have been published [129], facilitating the use of RMF all over the nuclear chart.
Applications of the Gogny interaction to specific nuclei can be found in [130,131,132,133,134,135,136,137,138,139,140], while applications of the Skyrme interaction to specific nuclei can be found in [69,133,141,142,143,144,145,146,147,148]. Applications of the RMF method to specific nuclei can be found in [149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171]. These will be discussed in the relevant sections regarding the corresponding series of isotopes.
The self-consistent mean-field approach used in nuclear structure [172,173] strongly resembles the density functional theory used for the description of many-electron systems [174,175]. The main difference is that in electronic systems one can derive very accurately the energy density functionals by ab initio methods from the theory of electron gas, while in nuclear structure the effective energy density functionals are motivated by ab initio theory [176], but they contain free parameters fitted to the nuclear structure data.
The simplest case of mean-field methods is the static mean-field approach, in which the lowest order nuclear binding energy is obtained [109]. In order to describe a larger set of data, correlations have to be included, in what is called the beyond-mean-field approach. SC in the beyond-mean-field method has been considered, for example, in [131,136,137,138,148,177,178]. The most important correlations are due to large amplitude collective motion, and are dealt by the generator coordinate method, which is a variational method in which the single-particle and collective nuclear dynamics are considered within a single coherent quantum mechanical formulation (see [109] and references therein). An alternative approach is to consider a much larger time-dependent (HFB) (TDHFB) phase space and extract from it a collective submanifold. This TDHFB approach [179] has been used in several light and medium-mass nuclei exhibiting SC [180,181].
A non-relativistic approach widely used [22,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196] for the study of nuclei exhibiting SC is the excited VAMPIR [197]. This is an (HFB) approach with spin and number projection before variation, extended to non-yrast states.
In addition, a five-dimensional quadrupole collective Hamiltonian (5DQCH) has been derived [198] microscopically from large amplitude collective motion and applied to several nuclei exhibiting SC [199,200,201,202,203,204,205].
The relativistic self-consistent mean-field approaches determine the ground state properties of nuclei. A method allowing the calculation of excitation spectra and B(E2) transition rates has been introduced by Otsuka and Nomura [206,207,208,209,210]. The parameters of the Interacting Boson Model (IBM) [211,212,213,214,215] are determined by mapping the total energy surface obtained from the RMF approach onto the total energy surface determined by IBM. One is then free to use the IBM Hamiltonian with these microscopically derived parameters for the calculation of spectra and electromagnetic transition rates. This method has already been applied to wide regions of the nuclear chart exhibiting SC [216,217,218,219,220,221,222,223,224,225,226,227,228,229,230].
In parallel, a 5DQCH with parameters microscopically determined from relativistic density functional theory [124,231,232,233,234,235,236,237] has also been used for the study of extended regions of the nuclear chart in which SC appears [238,239,240,241,242,243,244].

2.3. Nuclear Deformation and the Nuclear Collective Model

The primordial shell model introduced by Mayer and Jensen in 1949 [32,33,34,35,36,37] did explain the nuclear magic numbers, but it failed to explain the large nuclear quadrupole moments measured away from closed shells. It was soon realized by Rainwater [245] in 1950 that a spheroidal nuclear shape was energetically favored over the spherical one, offering greater stability to the nucleus. The collective model of Bohr and Mottelson [110,246,247,248,249] was then introduced in 1952, allowing the description of deformed nuclear shapes in terms of the collective variables β (describing the departure from sphericity) and γ (an angle representing the departure from axial symmetry).
The realization of the importance of nuclear deformation immediately led to the introduction of a deformed shell model by Nilsson [104,250,251] in 1955, in which an axially deformed three-dimensional harmonic oscillator is employed, characterized by a deformation parameter ϵ , expressing the departure from isotropy. The plots of the single-particle energies of the protons and of the neutrons as a function of the deformation parameter ϵ , known as the Nilsson diagrams [252], have accompanied experimental efforts for understanding the microscopic structure of nuclei for several decades, becoming an integral part of the nuclear terminology.
Soon thereafter the pairing-plus-quadrupole (PPQ) model of Kumar and Baranger was formulated [253,254,255] in 1965, incorporating the two most important features of the “residual” interaction, namely the pairing effects due to nuclear superfluidity [256] and the deformation effects due to the quadrupole–quadrupole interaction. Despite the fact that the true “residual” interaction might be much more complicated in form, the success of the PPQ model suggests that these two terms offer a very good approximate manifestation of the real interaction.
The importance of nuclear pairing has been demonstrated by Talmi [37,257,258,259,260,261] in 1962, through the introduction of the seniority quantum number v, which counts nucleon pairs in a nuclear shell not coupled to zero angular momentum. The seniority scheme is best manifested in semi-magic nuclei, having a magic number of protons and therefore only neutrons as valence particles. It provided clear explanations of the linear dependence of neutron separation energies on the mass number in the semi-magic Ca ( Z = 20 ) , Ni ( Z = 28 ) , and Sn ( Z = 50 ) series of isotopes, as well as of the nearly constant excitation energies of the 2 + states in the even–even Sn isotopes.
The importance of the quadrupole–quadrupole interaction has been most dramatically demonstrated by Casten in 1985 through the N p N n scheme [262,263,264,265], in which N p ( N n ) is the number of valence protons (neutrons) measured from the nearest closed shell. The N p N n scheme demonstrates that each nuclear observable exhibits a smooth behavior as a function of N p N n , independent of the region in which the nucleus sits in the nuclear chart.
A measure of nuclear collectivity, expressing the relative strength of the quadrupole–quadrupole interaction over the pairing interaction is the P-factor [265,266]
P = N p N n N p + N n .
Deformed nuclei possess P > 4 –5, while nuclei with P < 4 cannot be deformed. In relation to SC, if we consider a nucleus with N p = 4 , N n = 6 , we obtain P = 2.4 . Assuming 2p-2h excitations in the protons, we are going to effectively have N p = 8 , N n = 6 , yielding P = 3.43 , which is much higher than 2.4.
Another, easy to use, measure of nuclear collectivity is the ratio
R 4 / 2 = E ( 4 1 + ) E ( 2 1 + )
of the first two excited states of the ground state band of an even–even nucleus, which obtains values 2.0–2.4 for vibrational (near-spherical) nuclei, 2.4–3.0 for intermediate γ -unstable nuclei, and 3.0–3.33 for rotational (deformed) nuclei. For excited K = 0 bands, the ratio takes the form R 4 / 2 = ( E ( 4 ) E ( 0 ) ) / ( E ( 2 ) E ( 0 ) ) .
The proton-neutron interaction of the last proton and neutron in an even–even nucleus can be obtained from double differences of binding energies [267,268,269,270,271,272,273,274,275,276,277]
δ V p n ( Z , N ) = 1 4 [ { B ( Z , N ) B ( Z , N 2 ) } { B ( Z 2 , N ) B ( Z 2 , N 2 ) } ] ,
where B ( Z , N ) stands for the binding energy of the nucleus consisting of Z protons and N neutrons. In the notation of the Nilsson model, in which orbitals are labeled by K [ N N z Λ ] , where N is the total number of oscillator quanta, N z is the number of quanta along the axis z of cylindrical symmetry, and K ( Λ ) is the projection of the total (orbital) angular momentum on the symmetry axis, it has been found [268,273,274,278] that the proton–neutron pairs exhibiting maximum interaction are characterized by Nilsson labels differing by Δ K [ Δ N Δ N z Δ Λ ] = 0 [ 110 ] , which are called 0[110] Nilsson pairs.
The Bohr collective Hamiltonian with a sextic potential possessing two minima has been used [279,280,281,282,283,284] for the description of SC of spherical and prolate deformed shapes. A similar effort regarding SC of spherical and γ -unsable shapes has also been made [285]. More general potentials possessing two minima have been considered within the Bohr collective Hamiltonian in [286,287,288], while oblate–prolate SC has been considered in [289].

2.4. The SU(3) Symmetry

2.4.1. The SU(3) Symmetry of Elliott

A bridge between the microscopic picture of the shell model and the macroscopic picture of the collective model of Bohr and Mottelson has been provided in 1958 by Elliott [290,291,292,293,294], who proved that nuclear deformation naturally occurs within the sd shell of the shell model, which is characterized by a U(6) algebra possessing an SU(3) subalgebra, generated by the angular momentum and quadrupole operators [38,40,295]. Elliott’s papers stimulated extensive work on the development of SU(3) techniques appropriate for the description of deformed nuclei, recently reviewed by Kota [296], as well as on the use of algebraic techniques for the description of nuclear structure in general [297]. In what follows we are going to need Elliott’s notation for the irreducible representations of SU(3), which is ( λ , μ ) , where λ = f 1 f 2 and μ = f 2 , with f 1 ( f 2 ) being the number of boxes in the first (second) line of the relevant Young diagram.
The p, s d , p f , s d g , p f h , s d g i , p f h j shells of the isotropic 3D-HO are known to possess the U(3), U(6), U(10), U(15), U(21), U(28), U(36) symmetry, respectively, each of them having an SU(3) subalgebra [39,298]. In short, the shell with n oscillator quanta is characterized by a U((n + 1)(n + 2)/2) algebra. However, the SU(3) symmetry beyond the s d shell is destroyed by the spin–orbit interaction, which within each major shell acts most strongly on the orbital with the highest total angular momentum j, pushing it down into the shell below. Since the parity of the n shell is given by ( 1 ) n , each shell modified by the spin–orbit interaction ends up consisting of its original orbitals (except the one which escaped to the shell below), called the normal parity orbitals, plus the orbital invading from the shell above, which has the opposite parity and is called the abnormal parity or the intruder orbital.

2.4.2. The Pseudo-SU(3) Symmetry

A first attempt to restore the SU(3) symmetry beyond the sd shell was the development of the pseudo-SU(3) scheme [299,300,301,302,303,304,305,306], in which the normal parity orbitals remaining in each shell are mapped onto the full set of orbitals of the shell below, through a unitary transformation [307,308,309] introducing the quantum numbers of pseudo-angular momentum and pseudo-spin for each orbital. For example, in the 28-50 shell the Nilsson orbitals 1/2[301], 1/2[321] are mapped onto the pseudo-orbitals 1/2[200], 1/2[220] respectively, the orbitals 3/2[312], 1/2[310] are mapped onto 3/2[211], 1/2[211], and the orbitals 3/2[301], 5/2[303] are mapped onto 3/2[202], 5/2[202]. Remembering the Nilsson notation K [ N N z Λ ] and taking into account that the projection of the spin on the symmetry axis is Σ = K Λ , we remark that, when making the pseudo-SU(3) transformation, N is reduced by one unit (and therefore parity is changing sign), n z and K remain intact, while Σ is changing sign and Λ is modified accordingly in order to keep the rule Σ = K Λ satisfied. Tables for the correspondence of the normal and pseudo-SU(3) levels can be found in [310,311]. In this way, the SU(3) symmetry is restored for the normal parity orbitals, thus SU(3) techniques can be applied to them, while the intruder orbitals still have to be treated by shell-model methods. The pseudo-SU(3) scheme has been very successful in describing many aspects of medium-mass and heavy nuclei, like spectra and B(E2) transition rates in even–even [303,312] and odd [313,314] nuclei, the pairing interaction [315,316], double β decay [317,318] and superdeformation [26], in which a full-shell-model calculation would have been intractable.

2.4.3. The Quasi-SU(3) Symmetry

A second attempt to extend the SU(3) symmetry beyond the sd shell has been made in 1995 by Zuker et al. [319,320] with the introduction of the quasi-SU(3) symmetry. The basic idea of this symmetry is that in the p f shell the intruder orbital 1 g 9 / 2 , coming down from the sdg shell, does not suffice to create the necessary collectivity when combined with the normal parity orbitals 2 p 1 / 2 , 2 p 3 / 2 , 1 f 5 / 2 , remaining in the p f shell after the sinking of the 1 f 7 / 2 orbital to the shell below. In order to enhance collectivity, one has to include in the calculation in addition the orbitals 2 d 5 / 2 and 3 s 1 / 2 of the s d g shell, thus including in the collection all orbitals of the s d g shell successively differing by Δ j = 2 from the 1 g 9 / 2 orbital, where j stands for the total angular momentum. The Δ j = 2 rule is justified by looking at the matrix elements of the quadrupole operator in L-S and j-j coupling [319]. The quasi-SU(3) symmetry has been recently extended to the s d g shell [321,322], the orbitals of which ( 3 s 1 / 2 , 2 d 3 / 2 , 2 d 5 / 2 , 1 g 7 / 2 , 1 g 9 / 2 ) should be combined not only to the 1 h 11 / 2 intruder orbital, but in addition to its Δ j = 2 partner orbital 2 f 7 / 2 in order to create the desired collectivity.

2.4.4. The Proxy-SU(3) Symmetry

A third attempt to restore the SU(3) symmetry beyond the sd shell has been made in 2017 with the introduction of the proxy-SU(3) symmetry [323,324,325,326,327,328]. The proxy-SU(3) symmetry takes advantage of the Nilsson 0[110] pairs mentioned above, for which it is known that they correspond to orbitals having maximum spatial overlap [268]. One can then observe that in each shell the intruder orbitals (except the one with the highest K) are 0[110] partners of the orbitals which deserted the shell by going into the shell below. For example, in the 28–50 shell, the intruding 1g9/2 orbitals 1/2[440], 3/2[431], 5/2[422], 7/2[413] are 0[110] partners of the deserting 1 f 7 / 2 orbitals 1/2[330], 3/2[321], 5/2[312], 7/2[303], while the invading orbital 9/2[404] has no partner. 0[110] partner orbitals possess the same angular momentum and spin projections K, Λ , Σ , while they differ only by one quantum in the z-direction (and as a consequence, also by one quantum in N, thus having opposite parity). Because of this similarity, the intruder orbitals can act as “proxies” of the deserting orbitals, thus restoring the U(10) symmetry of the p f shell and its SU(3) subalgebra. The intruder orbital with the highest value of K remains alone and has to be treated separately, but, luckily enough, as one can see from the standard Nilsson diagrams [252] it is the orbital lying highest within the 28–50 shell, thus it should be empty for most nuclei within this shell. The same process, which is equivalent to a unitary transformation [329], applies to all higher shells. The accuracy of the approximation has been tested [323] by comparing the results of Nilsson calculations before and after the replacement, the outcome being that the accuracy of the approximation is improved with increasing size of the shell. The connection of the proxy-SU(3) scheme to the shell model [329,330] has proved that Nilsson 0[110] pairs correspond to | Δ n Δ l Δ j Δ m j = | 0110 pairs in the shell-model language, where n is the number of oscillator quanta, l (j) is the orbital (total) angular momentum, and m j is the projection of j on the symmetry axis z. The | 0110 pairs of orbitals are identical to the proton–neutron pairs identified by de Shalit and Goldhaber [331] in 1953 during studies of β transition probabilities. In particular, they observed that nucleons of one kind (neutrons, for example) have a stabilizing effect on pairs of nucleons of the other kind (neutrons in the example), thus favoring the development of nuclear deformation.
The proxy-SU(3) symmetry has been able to predict quite accurately the dependence of the deformation variable β along several series of isotopes of medium-mass and heavy nuclei, without using any free parameters [324,326]. This has been achieved by using the relations [332,333] between the Elliott quantum numbers λ , μ , characterizing the SU(3) irreps ( λ , μ ) , and the collective variables β , γ describing the nuclear shape. These relations have been obtained by establishing a mapping [332] between the invariant quantities of the rigid rotor, which are the second- and third-order Casimir operators of SU(3), C 2 ( λ , μ ) and C 3 ( λ , μ ) [38,40], and the invariants of the Bohr Hamiltonian, which are β 2 and β 3 cos 3 γ [246,247,249]. In addition, it provides a natural explanation of the dominance of prolate over oblate shapes in the ground states of even–even nuclei [324,328], and it successfully predicts [324] the transition from prolate to oblate shapes in the rare earths around N = 114 , again in a parameter-free way. As we shall see below in Section 2.4.6, it also provides parameter-free predictions for regions of the nuclear chart in which SC can be expected [334,335].

2.4.5. The SU(3) Symmetry in the Symplectic Model

An SU(3) subalgebra also exists within the symplectic model [336,337,338], characterized by the overall Sp(6,R) non-compact algebra (called Sp(3,R) in the papers in which it was introduced). The symplectic model is a generalization of Elliott’s SU(3) model, in which many oscillator major shells are taken into account. The advantage of the model is that it allows the determination of the shell-model configurations necessary for building collective rotations, as well as quadrupole and monopole vibrations in nuclei, thus providing a bridge between the shell model and the collective model of Bohr and Mottelson.
The SU(3) symmetry plays a crucial role within the symmetry-adapted no-core shell model (SANCSM) [78,79,80,81,82], which takes advantage of the appearance [339,340] of the symplectic symmetry in no-core shell-model calculations [76,77], thus allowing the use of SU(3) techniques within shell-model calculations. Such calculations have been carried out up to now in light nuclei, including 6 He, 6 Li, and 8 Be [341], 12 C and 16 O [339,340,342,343,344], 20 O, 20 , 22 , 24 Ne, and 20 , 22 Mg [345]. Recent SANCSM calculations in 6 Li and 20 Ne have revealed the importance of a few basic equilibrium shapes (deformed or not) in producing the structure of each nucleus [346].

2.4.6. The Dual-Shell Mechanism

Within the framework of the proxy-SU(3) symmetry a dual-shell mechanism [334,335] has been developed, allowing for the prediction of regions in the nuclear chart, in which the appearance of SC can be expected.
The three-dimensional isotropic harmonic oscillator (3D-HO) is known [38,39,40] to exhibit shell structure, with closed shells appearing at the harmonic oscillator (HO) magic numbers 2, 8, 20, 40, 70, 112, 168, ….
On the other hand, the connection [329] of the proxy-SU(3) symmetry to the shell model shows that the addition of the spin–orbit interaction to the 3D-HO modifies the shell structure, leading to the magic numbers 6, 14, 28, 50, 82, 126, 184, … (see Table 7 of [329]), which for the sake of brevity we call the spin–orbit (SO) magic numbers, although in reality they are 3D-HO magic numbers modified by the spin–orbit interaction.
In spherical nuclei (i.e., in nuclei with zero deformation), the HO magic numbers appear up to the s d shell, while the SO magic numbers take over above it, resulting in the well known set of shell-model magic numbers 2, 8, 20, 28, 50, 82, 126, … [32,34,35,36].
As soon as deformation sets in, the magic gaps in the single-particle energy diagrams, known as the Nilsson diagrams [252], soon disappear. However, this does not mean that the HO and/or the SO magic numbers themselves disappear. The HO and SO magic numbers are still valid: just the single particle energy levels become distorted by the interactions, leading to nuclear deformation. This is how the dual-shell mechanism emerges.
Looking into the details of the relevant Hamiltonian [334], one ends up with the conclusion that SC can appear only within certain regions of neutrons or protons, namely 7–8, 17–20, 34–40, 59–70, 96–112, 146–168, …. These regions appear as horizontal and vertical stripes on the nuclear chart (see Figure 25 of [334]). It should be emphasized that these stripes, which can be seen in Figure 1, Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6, express a necessary condition for the appearance of SC, and not a sufficient one.

2.5. Algebraic Models Using Bosons

A different path for reducing the size of shell-model calculations has been taken through representing correlated fermion pairs by bosons. This idea led to the introduction of the Interacting Boson Model (IBM) by Arima and Iachello in 1975 [211,212,213,214,215], the basic ingredients of which are s and d bosons with angular momentums of 0 and 2, respectively, a reasonable assumption since the quadrupole degree of freedom is dominant in nuclear collectivity. The bosons composing a nucleus correspond to valence fermion pairs counted from the nearest closed shell. In the simplest version of the model, labeled as IBM-1, no distinction is made between protons and neutrons. This distinction is taken into account in the version called IBM-2 or pn-IBM. Odd nuclei (not included in the present review) are described by considering a single fermion in addition to the bosons representing the even–even part of the nucleus, thus forming the Interacting Boson–Fermion Model (IBFM) [215]. Nuclear supersymmetries, describing quadruplets of even–even, even–odd, odd–even, and odd–odd nuclei by the same Hamiltonian have also been developed [215].
IBM-1 is characterized by a U(6) overall symmetry, having three different chains of subalgebras containing the angular momentum algebra SO(3), so that angular momentum can be used for labeling the nuclear energy states. Each chain of subalgebras corresponds to a different dynamical symmetry, representing different physics. The U(5) dynamical symmetry [408] is appropriate for vibrational (near spherical) nuclei appearing near to closed shells, while the SU(3) dynamical symmetry [409] corresponds to axially deformed prolate (rugby ball like) nuclei appearing away from closed shells. A variant called SU ( 3 ) ¯ corresponds to axially symmetric oblate (pancake like) nuclei. The O(6) dynamical symmetry [410] corresponds to nuclei with intermediate values of deformation, characterized as γ -unstable, since they are soft with respect to the value of the collective variable γ , thus exhibiting unstable triaxial shapes.
A convenient way to visualize the three dynamical symmetries of IBM is to place them on the vertices of a triangle [265,411,412], with each point within the triangle corresponding to a different set of parameter values of the IBM Hamiltonian.
Partial dynamical symmetries [413,414] have also been developed in the IBM framework, regarding situations in which a symmetry is obeyed by a subset of soluble eigenstates, but is not shared by the Hamiltonian. SC of spherical and deformed shapes, as well as of prolate and oblate shapes, have been predicted [415,416] within the PDS framework.
IBM-1 can describe only states with positive parity. Negative parity states can be accommodated with the addition of p and f bosons with negative parity and angular momenta of 1 and 3, respectively [417,418]. The resulting spdf-IBM can be used for the description of octupole bands appearing in the rare earth [419,420] and in the actinide [421,422] region.
SC can be described within the extended IBM framework [423,424,425,426,427,428,429,430] by allowing two-particle–two-hole (2p-2h) excitations to occur, which obviously increases the number of bosons by 2, since two additional pairs are taken into account. Mixing [431,432] of the ground state configuration of N bosons with the excited configuration of N + 2 bosons leads then to the two coexisting bands. This scheme is called the IBM with configuration mixing (IBM-CM). It has already been applied for the description of SC in several regions of the nuclear chart [433,434,435,436,437,438,439,440,441,442,443,444,445,446].

2.6. Shape/Phase Transitions and Critical Point Symmetries

Shape/phase transitions (SPTs) [347,348,447,448,449,450,451] in chains of isotopes occur when the shape of the nucleus is undergoing a radical change by the addition of two neutrons. A spectacular example is seen in the Sm series of isotopes, in which the R 4 / 2 ratio jumps from the near-vibrational value of 2.316 in 150 Sm to the near-rotational value of 3.009 in 152 Sm [406].
In the IBM a first order SPT (according to the Ehrenfest classification) is known since 1981 [411] to occur between the dynamical symmetries U(5) and SU(3) [452], while a second-order SPT is found between the dynamical symmetries U(5) and O(6) [411]. No SPT occurs between SU(3) and O(6), but O(6) has been suggested as the critical point of the transition from prolate SU(3) shapes to oblate SU ( 3 ) ¯ shapes [453,454,455]. A convenient way to form a picture of these SPTs is to place them on a triangle [265,411,412], with the vertices of the triangle corresponding to the U(5), SU(3), and O(6) dynamical symmetries of IBM, and with each point within the triangle corresponding to a different set of parameter values of the IBM Hamiltonian. This triangle is doubled [453,454,455] by the addition of the SU ( 3 ) ¯ symmetry.
In Bohr’s collective Hamiltonian, a critical point symmetry (CPS) called E(5) has been suggested by Iachello [450] in 2000 between spherical and γ -unstable nuclei, while a CPS called X(5) has been suggested by Iachello [451] in 2001 between spherical and prolate deformed nuclei. A convenient way to demonstrate these critical point symmetries is to place them on a triangle [456], the vertices of which correspond to spherical, deformed, and γ -unstable nuclei, with each point within the triangle corresponding to a different set of parameter values of the Bohr Hamiltonian. In addition, a Y(5) CPS between axial and triaxial nuclei [457] and a Z(5) CPS between prolate and oblate shapes [458] have been suggested. Fixing γ to the value of maximum triaxiality, π / 6 , the CPSs X(5) and Z(5) reduce to the CPSs X(3) [459] and Z(4) [460], respectively. Equal shape/phase space mixing of the γ -stable X(5) and the γ -rigid X(3) solutions leads to the X(4) [461] CPS. A link between X(4) and Z(4) is provided by T(4) [462], while a link between E(5) and X(5) is given by F(5) [463].
The best experimental manifestations of X(5) have been found in the N = 90 isotones 150 Nd [361,362], 152 Sm [366,372], and 154 Gd [375,376]. Several other candidates have been suggested, reviewed in Refs. [347,356,464], with relevant references given in the caption of Figure 2.
The only fully confirmed experimental manifestation of E(5) is found in 134 Ba [402,403,404,405]. Several candidates have been suggested, reviewed in Refs. [347,391,464], with relevant references given in the caption of Figure 2.
The way the energy levels are rearranged by SPTs has been clarified in terms of quasidynamical symmetries in a series of papers by Rowe et al. [465,466,467,468].
The close connection between SPTs and SC [469] has been studied in the region around N = 60 [439]. Further discussion is deferred to the relevant Section 5.5, Section 8.5, and Section 9.5.
In relation to the microscopic origin of SPTs, the crucial role played by the proton–neutron interaction in the creation of deformation in medium-mass and heavy nuclei, clarified by Federman and Pittel in 1977 [470,471,472,473], offers an answer. In light, intermediate-mass, rare earth, and actinide nuclei the ( 1 d 5 / 2 , 1 d 3 / 2 ), ( 1 g 9 / 2 , 1 g 7 / 2 ), ( 1 h 11 / 2 , 1 h 9 / 2 ), and ( 1 i 13 / 2 , 1 i 11 / 2 ) proton–neutron pairs of orbitals, respectively, play a major role at the onset of deformation, while the pairs ( 1 d 5 / 2 , 1 f 7 / 2 ), ( 1 g 9 / 2 , 1 h 11 / 2 ), ( 1 h 11 / 2 , 1 i 13 / 2 ), and ( 1 i 13 / 2 , 1 j 15 / 2 ) play a major role in the increase of deformation after its onset, respectively. Notice that the first four couples of orbitals are | Δ n Δ l Δ j Δ m j = | 0010 shell-model pairs, while the last 4 couples are | 0110 shell-model pairs. Notice also that the same proton orbitals appear in both sets. In the first set they are coupled to neutron orbitals of the same parity, while in the second set they are coupled to neutron orbitals of the opposite parity, which are the orbitals of maximum j in their shell, pushed down to the shell below by the spin–orbit interaction and thus becoming the intruder orbitals. In other words, the intruder neutron orbitals play a major role in the development of deformation across the nuclear chart. Therefore they are expected to be the protagonists in the regions of the sharp SPT from spherical to deformed nuclei occurring at N = 90 , as well as at N = 60 and N = 40 , and in the appearance of SC there, as we shall see in the relevant Section 5.5, Section 8.5 and Section 9.5.
A recent major step forward in the development of CPSs in atomic nuclei has been the introduction in their study of the conformable fractional derivatives by Hammad [474] in 2021. While in the standard fractional calculus [475,476,477], fractional derivatives do not obey the familiar basic properties of derivatives, the conformable fractional derivatives [478] do obey them, thus offering obvious analytical advantages. Analytical models using conformable fractional derivatives contain an extra parameter, the order of the derivative, which allows closer approach to the data of critical nuclei [474,479,480,481].

2.7. The O(6) Symmetry

Since the first article by Elliott [290] in 1958, in which the description of nuclear deformation in the s d shell of the shell model that has an overall U(6) symmetry, in terms of the SU(3) symmetry has been introduced, it is known that an alternative formulation of the s d shell in terms of the O(6) symmetry is possible (see the Appendix of [290]). However, while the SU(3) subalgebra is present in all algebras U( ( n + 1 ) ( n + 2 ) / 2 ) characterizing the n-th shell of the isotropic 3D-HO [39,40,298], this is not the case for the O(6). As a consequence, while SU(3) can be applied to heavier shells of the shell model through the use of pseudo-SU(3) and/or quasi-SU(3) and/or proxy-SU(3), no similar generalization regarding O(6) exists so far.
However, the use of O(6) in heavier nuclei is possible in the framework of the IBM [214] (see Figure 2.4 of [214] for the regions of the nuclear chart in which experimental manifestations of the O(6) dynamical symmetry have been seen).
O(6) is related to two SPTs. One of them is the second-order SPT in IBM between U(5) and O(6) [411], characterized by the common O(5) subalgebra of these two dynamical symmetries [482], and labeled by E(5) [450] in the framework of the Bohr Hamiltonian, as mentioned in the preceding section.
The other one is the SPT in IBM between prolate (rugby-ball like) and oblate (pancake-like) deformed nuclei, characterized by the dynamical symmetries SU(3) and SU ( 3 ) ¯ [453] respectively. It has been argued [453,454,455] that O(6) is the critical point of the SPT from prolate to oblate shapes, with relevant experimental support provided in [483]. Within the framework of the Bohr Hamiltonian the prolate-to-oblate SPT has been described by the Z(5) [458] CPS. The prolate-to-oblate transition is predicted [324] in a parameter-free way within the proxy-SU(3) symmetry.
An analytic approach to SC within the U(5) to O(6) transition has been given in [484,485], with the role played by their common O(5) subalgebra being emphasized.
It may be mentioned that SC has also been considered [486] within a simplified shell model with O(4) symmetry, a multi-shell version of which has been studied through TDHFB theory. In addition, SC has been considered [487] within a simplified model with protons in two subshells and neutrons in one subshell, with each subshell possessing the SO(8) symmetry of the Ginocchio model [488,489], which guarantees that a space of s and d pairs is closed, i.e., it does not bring in pairs of higher angular momentum.

2.8. The Nature of 0 + States

The most common manifestation of SC in even–even nuclei regards the occurrence of a 0 + band lying close in energy with the ground state band and having radically different structure from it, for example with one of the bands being spherical and the other one deformed. It is therefore of interest to clarify the nature of low-lying 0 + bands in even–even nuclei.
At least since 2001 [490] it has been clear that most of the first excited 0 2 + states are not β vibrations of the Bohr–Mottelson collective model. In order to characterize a 0 + band as a β vibrational band, one should have appreciable B(E2) transition rates connecting this band to the ground state band (appreciable B ( E 2 ; 0 β + 2 g s + ) or B ( E 2 ; 2 β + 0 g s + ) . Furthermore, the E 0 transition connecting the two bands should also be strong [491]. However, in order to unambiguously characterize a 0 + state as a β vibrational band-head, the presence of anharmonicity contibutions should be taken into account.
The number of 0 + states in an even–even nucleus can be large. The first example found experimentally in 2002 was 158 Gd, in which 13 excited 0 + states have been found below 3.1 MeV [492,493,494,495]. Many theoretical interpretations have been suggested for these 0 + states, including two-phonon octupole character [496], the pseudo-SU(3) model [497,498,499], the quasiparticle–phonon model [500], and the pairing plus quadrupole model [501].
Further experimental investigations found many 0 + states in several rare earth nuclei [502,503,504,505,506,507], and theoretical interpretations have been provided using the pseudo-SU(3) symmetry [508,509], the quasiparticle–phonon model [504], the projected shell model [504], the pairing plus quadrupole model [510,511], as well as simple band mixing calculations [512]. Special attention has also been given to the Cd isotopes [513,514,515], as well as to the Sn and Te isotopes [516,517], since SC appears there, as well as to 152 Sm [518] (using the multiphonon approach), since it is a textbook example of the X(5) CPS. Examples have also been found in the actinides [519], interpreted in terms of the quasiparticle–phonon model [520] and the s p d f version of the IBM [519].
The vibrational character of the 0 2 + state in 154 Gd was disputed in 2008 by Sharpey-Schafer [521,522,523,524,525], by suggesting that this state is not a β vibration based on the ground state, but a second vacuum, each with its own γ and octupole vibrations. Similar arguments have been presented for the neighboring N = 90 isotones, 150 Nd, 152 Sm, and 156 Dy, which, along with 154 Gd, happen to be the best manifestations of the X(5) CPS [347,464]. Further discussion on the relation between the X(5) CPS and SC is deferred to the relevant Section 5.5.
Strong electric monopole transitions, characterized by the monopole strength ρ 2 ( E 0 ) , connecting excited 0 + states to the ground state, are an important signature of SC [491], since they provide a clear sign of shape mixing. The relevant data have been reviewed in [526,527], while novel experimental approaches for their identification, based on conversion electron spectroscopy, have been reviewed by Zganjar [528].

2.9. Multiple Shape Coexistence

In some nuclei more than one excited low-lying 0 + states have been observed, leading to the suggestion of multiple SC.
The first example has been found in 2000 in 82 186 Pb 104 [529], in which three coexisting bands with spherical, oblate, and prolate shapes have been determined. Similar observations have been made in other Pb isotopes [530].
Three different configurations have been found in 2018 in 38 96 Sr 58 [531], in accordance to earlier (2012) predictions by Petrovici within the complex excited VAMPIR approach [190]. In the same paper [190], triple SC has been predicted also in the neighboring nucleus 40 98 Zr 58 , in which the influence of multiple SC in 40 98 Zr 58 on the β decay from 39 96 Y 57 to it is discussed [196].
Four coexisting bands have been found in 2019 in 48 110 , 112 Cd 62 , 64 [177,178], interpreted in terms of beyond-mean-field approach with the Gogny D1S interaction [177]. Energy systematics suggest that they are built on multiparticle–multihole proton excitations across the Z = 50 gap [178].
Triple coexistence of spherical, prolate, and oblate shapes has been found in 2022 in 28 64 Ni 36 [532], interpreted in terms of MCSM calculations. Triple coexistence of spherical, prolate, and oblate shapes has also been predicted [415,416] in the framework of partial dynamical symmetries [413,414] of the IBM.
Multiple SC has also been predicted in 2011 in 40 80 Zr 40 [138], using the beyond-mean-field approach with the Gogny D1S interaction, as well as in 2017 in 36 76 , 78 Kr 40 , 42 [224], using an IBM Hamiltonian with parameters derived microscopically from a Gogny D1M energy density functional.
A mechanism for multiple SC, being an extension of the dual-shell mechanism described in Section 2.4.6 has been recently suggested [335]. In short, the simultaneous presence of the harmonic oscillator (HO) magic numbers and the spin–orbit (SO) magic numbers, described in Section 2.4.6, can lead to four different combinations of magic numbers of protons and neutrons, namely SO-SO, SO-HO, HO-SO, and HO-HO, leading to four different shapes for the same nucleus [335].

2.10. Islands of Inversion

In light nuclei the term islands of inversion (IoIs) has been coined [533] in the N = 20 region, in order to describe the situation in which in a group of semimagic or even doubly magic nuclei, expected to have a spherical ground state, turn out to have a deformed ground state. This effect is attributed [534] to the occurrence of highly correlated many-particle–many-hole configurations, called intruder configurations, pushed low in energy by the proton–neutron interaction, so that they go lower in energy than the spherical configuration that otherwise would have been the ground state. This mechanism is identical [236] to the particle–hole excitation mechanism leading to SC. It also clarifies the relation between SC and SPTs from spherical to deformed ground states, with the SPT expected to take place on the shores of the island of inversion.
Well-documented islands of inversion are known to occur [535,536] at N = 8 [537], N = 20 [533,537,538,539,540,541,542,543], N = 28 [49,544,545], N = 40 [50,546,547,548,549,550,551], and N = 50 [534]. Merging of the IoIs occurring at N = 20 and N = 28 has been suggested in [49,545], while merging of the IoIs occurring at N = 40 and N = 50 has been proposed in [534,551].
Islands of inversion will be further discussed in Section 12.

3. The Z ≈ 82 Region

3.1. The Hg (Z = 80) Isotopes

The Hg isotopes are the ones in which SC has been first observed.
The plot of experimental energy-level systematics of the Hg isotopes, appearing in similar form in many papers (see Figure 6 of [552] , Figure 10 of [4], Figure 1 of [435], Figure 1 of [553]), and shown here in Figure 7a, is the hallmark of SC. Excited K π = 0 + bands invade the practically vibrational spectra of the ground state bands of Hg isotopes, forming parabolas with minima at N = 102 , which is close to the midshell N = 104 . The parabolas raise towards infinity at N = 96 on the left and at N = 110 on the right, providing strong evidence for SC of a deformed excited 0 + band with a spherical ground state band.
Indeed 184 Hg 104 and 186 Hg 106 were the first Hg isotopes in which SC has been observed in 1973 [554,555,556,557], confirmed through both spectra and lifetimes. Further experimental studies confirmed the occurrence of SC in these isotopes [558,559,560,561].
SC in the neighbors of the previous isotopes, 182 Hg 102 and 188 Hg 108 , has been seen in [28,553,554,561,562,563,564,565]. Further to the left, SC has been seen in the isotopes 178 , 180 Hg 98 , 100 in [28,566,567,568,569,570,571].
The lower border has been reached in [566], in which SC is seen in 178 Hg 98 but not in 176 Hg 96 . The upper border has been reached in [552,564]. In [564] a systematic study has been carried out in 188 200 Hg 108 120 , making clear that the onset of state mixing, and therefore SC, occurs at 190 Hg 110 , while no state mixing and SC is seen in 192 200 Hg 112 120 .
In conclusion, experimental findings support the occurrence of SC within the region 176 190 Hg 96 110 , while no SC is seen outside it.
From the theoretical point of view, SC in 178 188 Hg 96 110 has been predicted and/or confirmed by many calculations using different methods, including mean field [148,170,572,573,574,575], particle plus rotor [576], and extended IBM [424] calculations. 190 Hg 110 has been suggested as the upper limit of SC in mean-field calculations by Zhang and Hamilton in 1991 [577] and Nazarewicz in 1993 [578]. Recent calculations using covariant density functional theory (CDFT) with the DDME2 functional with standard pairing included, have found that proton particle–hole excitations induced by the neutrons occur in 176 190 Hg 96 110 [579,580].
The advent of computational facilities made possible in 2013 extended IBM calculations by Nomura et al. [220,222] with parameters microscopically derived from RMF theory with the Gogny–D1M interaction. Detailed calculations all over the region 172 204 Hg 92 124 have revealed the onset of a second minimum in the potential energy surface at 176 Hg 96 , coexistence of prolate and oblate minima in 178 190 Hg 98 110 , and the disappearance of the prolate minimum at 192 Hg 112 .
Extended calculations have also been performed in 2014 in the IBM-CM by García-Ramos and Heyde [435]. The isotopes 172 200 Hg 92 120 have been considered. Clear cases of SC have been found in 180 188 Hg 100 108 . The prolate minimum is found to disappear at 190 Hg 110 , in remarkably close agreement with the RMF results of [220].
Earlier extended RMF calculations have been carried out in 1994 by Patra et al. [160] in 170 200 Hg 90 120 . An oblate ground state has been found in most cases, but a transition from oblate to prolate shape has been seen at 178 Hg 98 and a transition from prolate back to the oblate shape has been seen at 188 Hg 108 . These results are in agreement with the more recent calculations predicting SC of prolate and oblate shapes in the 178 188 Hg 98 108 region, but they differ in relation to the nature of the ground state in this region, which appears to be oblate and not prolate, as pointed out by Heyde et al. [152], based on the data [566,567,568], and in agreement with earlier calculations [572,578]. As pointed out in Ref. [165], the energy difference between the prolate and oblate configurations is very sensitive to the choice of the input parameters, as well as to the amount of pairing involved.
In conclusion, the theoretical predictions closely agree with the experimental findings, indicating SC in the region starting at N = 96 or 98 and ending up at N = 108 or 110.

3.2. The Pb (Z = 82) Isotopes

Following the observation of proton particle–hole excitations across the Z = 82 magic number in the Hg isotopes, it was expected to search for similar excitations in the Pb isotopes, lying exactly at the Z = 82 energy gap. Some findings are summarized in Figure 7b, in which again intruder states forming parabolas with minima at N = 104 are seen.
Low-lying 0 + states in the Pb isotopes have been first identified in 1984 in 192 200 Pb 110 118 [581,582,583], and the right branch of a parabola formed by them, going down in energy from the upper border of the 82–126 neutron shell towards the midshell ( N = 104 ) has been formed [583]. It has been realized that the 0 + state becomes the first excited state in 192 , 194 Pb 110 , 112 , sinking lower than the 2 + member of the ground state band [582]. The microscopic nature of these 0 + states, based on proton 2p-2h excitations, has also been clarified [581,582,583]. It was even realized that “the general idea of obtaining low-lying intruder states near or in single-closed-shell nuclei with a maximal number of valence nucleons of the other type is shown to be correct” [582]. The isotopes 188 198 Pb 106 116 , forming the right branch of the parabola, have been reviewed recently in [565].
The next important step was the identification in 1998 of coexisting spherical, prolate, and oblate configurations in 188 Pb 106 [530,584,585,586], offering the first example of triple SC. Triple SC has also been seen in 2000 at the neutron midshell, i.e., in 186 Pb 104 [529], by realizing that the three lowest in energy states are spherical, oblate, and prolate. An oblate band has been identified [587] in 186 Pb 104 , in addition to the prolate ground state band.
The left branch of the parabola of intruder 0 + states has been completed by searches in 184 Pb 102 [588], 182 Pb 100 [589], and 180 Pb 98 [590], reviewed in [591], in which an updated figure of the parabola is given (Figure 1 in [591], Figure 3 in [590]). It is seen that the prolate intruder 0 + band reaches a minimum in energy at midshell ( N = 104 ) and keeps rising sharply as the neutron number is reduced.
It should be noticed that 188 Pb 106 and 194 Pb 112 represent two rare cases in which the 0 2 + band-head of a well-developed K π = 0 + band lies below the 2 1 + state of the ground state band [335]. It has been found [335] that such nuclei tend to be surrounded by nuclei exhibiting SC.
Early theoretical considerations over extended series of Pb isotopes included calculations [84] in the 186 204 Pb 104 122 region, confirming the equivalence of the spherical and deformed shell-model approaches to the interpretation of intruder 0 + states through the particle–hole excitation mechanism, as well as RMF calculations in the 178 208 Pb 96 126 region, predicting oblate ground states in the 188 196 Pb 106 114 and prolate ones outside it [170]. This prediction has been questioned by Ref. [152], leading to the realization [165] that the ordering between prolate and oblate configurations is very sensitive to the input parameters, the pairing interaction included. In addition, calculations have been performed in the 190 196 Pb 108 114 region, studying particle–hole excitations in the extended IBM formalism [424], as well as in the particle-plus-rotor model, in which configuration mixing in relation to SC has been considered [576].
Theoretical investigations focused on predicting the triple SC of spherical, prolate and oblate shapes in 186 Pb 104 have been performed in 2003 in the mean-field framework using the Skyrme SLy6 interaction [142], as well as in 2004 in the IBM-CM framework [433]. Soon thereafter, calculations covering the region 182 194 Pb 100 112 have been performed in the mean-field framework using the Gogny D1S force [130,140], the Skyrme interaction [148], the density-dependent cluster model [592], the total-Routhian-surface (TRS) method [574], the configuration-constrained potential-energy-surface method [575], the IBM-CM [444] with parameters fitted to the data, and the IBM with parameters determined through the Gogny D1M energy density functional [219], consistently describing SC in these nuclei and reproducing the parabolic behavior of the intruder 0 + states (see Figure 7 of [140], Figure 3 of [219], Figure 7 of [444]).
It was suggested in Ref. [130] that the low-lying prolate and oblate 0 + states are predominantly due to neutron correlations, while the protons behave rather as spectators, in contrast to the traditional interpretation of proton 2p-2h excitations [581,582,583]. The recent findings in the framework of covariant DFT calculations [579,580] seem to bridge this gap, suggesting that SC in this region is indeed due to proton particle–hole excitations, but the cause of these particle–hole excitations is based on the increased role played by the neutrons near midshell.
Recent (2022) extended calculations [153] using the deformed relativistic Hartree–Bogoliubov (RHB) theory in the 172 302 Pb 90 220 region, have reconfirmed SC in 184 188 Pb 102 106 , in addition to finding a peninsula near the neutron drip line, consisting of the 278 296 , 300 Pb 196 214 , 218 isotopes. Furthermore, recent calculations using CDFT with the DDME2 functional with standard pairing included have found that proton particle–hole excitations induced by the neutrons occur in 180 190 Pb 98 108 [579,580].
In summary, the theoretical predictions closely agree with the experimental findings, indicating SC in the region starting at N = 98 and ending up at N = 112.

3.3. The Po (Z = 84) Isotopes

Following the observation of proton particle–hole excitations across the Z = 82 magic number in the Hg isotopes, lying two protons below the Z = 82 magic number, it was reasonable to search for similar excitations in the Po isotopes, lying two protons above the Z = 82 energy gap. Some findings are summarized in Figure 7c, in which intruder states forming the right branch of a parabola are seen.
Since 1991, experimental examples of SC due to proton 4p-2h excitations and configuration mixing with the proton 2p ground state have been found in 192 196 Po 108 112 in several works [584,593,594,595,596,597]. In addition, experimental examples of SC have been seen in 198 , 200 Po 114 , 116 [593,597,598,599]. In Ref. [598], a transition was suggested to occur between 202 Po 118 and 200 Po 116 , corresponding to a transition between heavier Po isotopes, in which the positive parity states can be described as two-phonon multiplets within a quadrupole vibrational model, and lighter Po isotopes, in which proton 4p-2h excitations appear, mixing with the proton 2p ground state band and leading to SC, see also [565] in relation to this transition. The intruder 0 2 + states are seen to form the right branch of a parabola reaching 202 Po 118 (see Figure 4 of [597], Figure 9 of [598], Figure 7 of [596], and Figure 1 of [599]).
From the theoretical point of view, several calculations revealing SC have been performed in the region 188 200 Po 104 116 , including extended-IBM [424,425] and particle-core model [600] calculations, based on the proton 4p-2h excitations picture, as well as mean-field calculations in the standard deformed Woods-Saxon plus pairing approach [573,575] and calculations in the density-dependent cluster model [592]. Efforts extending beyond 200 Po 116 include IBM-CM calculations [437], in which the change from the SC picture involving 4p-2h configurations to regular rotational behavior when passing from 200 Po 116 to 202 Po 118 has been corroborated, as well as beyond-mean-field calculations with a Skyrme energy density functional [148], which indicate that the ground state is oblate above N = 106 and prolate at and below it, in agreement with the findings within the particle-core model [600], in which it has been found that in 192 Po 108 the intruder band becomes the ground-state configuration. Recent calculations using CDFT with the DDME2 functional with standard pairing included have found that proton particle–hole excitations induced by the neutrons occur in 182 192 Po 98 108 [579,580].
In conclusion, both experimental and theoretical pieces of evidence converge to the conclusion that SC based on proton 4p–2h excitations is seen from N = 98 up to N = 116 , while the regular rotational character dominates above it.

3.4. The Rn (Z = 86) Isotopes

Having seen signs of SC in the Po isotopes, lying two protons above the Z = 82 energy gap, it is reasonable to search for similar effects in the Rn isotopes, which lie four protons above the Z = 82 gap.
Beyond-mean-field calculations with a Skyrme energy density functional [148] in the region 194 204 Rn 108 118 , indicate oblate ground states in these nuclei (see Figures 3 and 6 of [148] for the relevant potential energy curves). Mixing of the ground state and the 0 2 + state has been seen in 202 Rn 116 using a standard deformed Woods–Saxon plus pairing approach [573]. Therefore, theoretical predictions appear to be similar as in the Po isotopes, suggesting the occurrence of SC up to N = 116 . However, further experimental efforts are needed to confirm this suggestion, since the 0 2 + state in 202 Rn 116 remains elusive [565].
In summary, some hints for SC exists in the N = 108 -116 region.

3.5. Heavy Nuclei Beyond Z = 86

Data related to SC become scarce in the actinides beyond Z = 86 . For example, sets of 0 + states have been observed [519] in 90 228 , 230 Th 138 , 140 and 92 232 U 140 , and their microscopic interpretation in terms of the quasiparticle–phonon model has been given [520]. In addition, the spectrum of 94 238 Pu 144 has been described [282] within the Bohr Hamiltonian with a sextic potential possessing two minima, thus possibly related to a first-order SPT from spherical to prolate deformed shapes and/or SC related to it.
SC in superheavy nuclei with Z = 100 –114 has been considered [161] by self-consistent RMF theory, finding that in some cases near Z = 114 and N = 174 the superdeformed band may become the ground state. More recent calculations on superheavy nuclei using CDFT with various functionals [601,602,603] indicate that the nature of the ground state depends on the functional used. For example, in the Z = 120 , N = 184 region, some functionals (NL3*, DD-ME2, and PC-PK1) predict spherical ground states, while other functionals (DD-PC1,DD-ME δ ) predict oblate shapes [601].
Recent beyond-mean-field calculations [131] using the Gogny force for the 114 288 298 Fl 174 184 superheavy nuclei find that triaxial shapes dominate in this region. A prolate ground state is predicted in 114 288 Fl 174 , while an oblate ground state is predicted in 114 296 Fl 182 , with a prolate-to-oblate transition predicted to occur between 114 290 Fl 176 and 114 292 Fl 178 , with a novel type of SC of two different triaxial shapes occurring in 114 290 Fl 176 .
It seems that superheavy nuclei present a special interest for further studies, since a rich variety of novel phenomena appears to be predicted there.

3.6. The Pt (Z = 78) Isotopes

Following the observation of proton particle–hole excitations across the Z = 82 magic number in the Hg isotopes, lying two protons below the Z = 82 magic number, it was reasonable to examine to which extent similar excitations appear in the Pt isotopes, lying four protons below the Z = 82 energy gap.
SC in the Pt isotopes was first seen experimentally in 1986 in 184 Pt 106 [604,605,606]. It was later found in lighter isotopes, down to 174 Pt 96 [607,608,609,610,611], as well as in heavier isotopes, up to 188 Pt 110 [612,613]. In summary, experimental evidence for SC has been found in the region 174 188 Pt 96 110 , in agreement with Figure 3.26 of the review article [3] (see also Figure 3 of [434]), in which the weakly deformed and the strongly deformed states in the even–even Pt isotopes are depicted, showing that the strongly deformed states nearly form a well with walls sharply rising at N = 96 on the left and at N = 110 on the right.
Several theoretical calculations have been carried out using the IBM-CM [434,443,446], in which the model parameters have been fitted to the data. In addition, IBM calculations have been carried out with parameters determined from energy density functional theory with Gogny-D1S [216] and Gogny-D1M [218] interactions, as well as with the DD-PC1 functional [217]. A comparison between these two approaches has been carried out in [436], reaching the conclusion that SC is seen in the region 176 186 Pt 98 108 , in agreement with the data.
An alternative point of view had been given in [614], in which IBM calculations have been carried out in the region 172 194 Pt 94 116 , claiming that a good description of these isotopes can be obtained without configuration mixing, i.e., without involving particle–hole excitations. It seems that p-h effects are stronger for nuclei lying at Z = 82 ± 2 (namely Pb, Po, Hg), while they are “fading away” when moving further away from the magic number 82. As a consequence, the SC effects in Pt, which appears to be a “border case”, can also be described in a theoretical framework involving free parameters without explicitly including the p-h excitations.
A detailed analysis [101] of 194 Pt 116 in terms of IBM with configuration mixing has shown that the mixing between the ground-state band and the excited band is very small, while the lower band is only slightly more collective than the upper band, showing that the criteria for SC are not satisfied in this nucleus.
Configuration mixing and/or SC of prolate and oblate shapes within the 176 188 Pt 98 110 region have been corroborated by several calculations in the mean-field framework [572,573,575], as well as within the particle plus rotor model [576] and the Bohr collective model [615]. Deformed RMF calculations [170] have confirmed the absence of spherical shapes in 176 180 Pt 98 102 , in contrast to the Pb isotopes, in which triple SC of spherical, prolate, and oblate structures has been seen.
Several calculations have focused on the region 184 200 Pt 106 122 . Both mean-field calculations [145,616], as well as 5DQCH calculations with parameters determined through CDFT [244] have found that an oblate-to-prolate transition occurs at N = 116 , where the prolate and oblate minima have approximately the same depth. In addition, RMF calculations have shown that deformation and SC are favored within the 176 190 Pt 98 112 region [168], in which oblate shapes are favored [162].
Recent calculations [151] within the deformed RHB theory have indicated that bubble structures with SC can be obtained in the extremely neutron-rich nucleus 210 Pt 132 .
In summary, both experimental findings and theoretical predictions suggest SC in the region bordered roughly by 174 Pt 96 and 188 Pt 110 .

4. The Z = 68–76 Desert

4.1. The Os (Z = 76) Isotopes

Experimental results on 176 180 Os 100 104 [383] have pointed to 178 Os 102 as a good candidate for the X(5) CPS. SC in 178 Os 102 of a prolate deformed ground state with γ -soft excited states has been supported by experimental results in [617], backed up by projected angular momentum deformed HF calculations, as well as by cranked Woods–Saxon model calculations [617]. Shape/phase mixing in the 174 180 Os 98 104 region has been supported by calculations in a hybrid collective model [615].
The 0 + states in 190 Os 114 have attracted experimental attention [505,506], since the enhanced density of low-lying 0 + states has been considered as evidence for shape/phase transitional behavior [505,506]. However, such an enhanced density has been located in 154 Gd 90 and not in 190 Os 114 [506].
The 186 198 Os 110 122 region has attracted much theoretical attention, since SC is expected to occur in relation to the SPT from prolate to oblate shapes seen around N = 116 [324], which implies the occurrence of SC around it. Potential energy surfaces revealing this transition have been calculated within a self-consistent HFB approach [616], a Skyrme HF plus BCS framework [145], as well as with an IBM Hamiltonian with parameters determined from the microscopic energy density functional Gogny D1M [218] and Gogny D1S [209] (see Figure 2 of [616], Figure 2 of [145], Figure 1 of [218], and Figure 1 of [209] for the corresponding PES, as well as Figure 5 of [209] for relevant level schemes).
More recently, extended calculations with a 5DQCH with parameters determined from CDFT have been performed in 178 200 Os 102 124 [244] (see Figure 5 of [244] for the relevant potential energy surfaces and Figure 9 for level schemes). An evolution is seen from well-deformed prolate shapes at N = 102 (near to the midshell N = 104 ) to prolate–oblate SC in 190 , 192 Os 114 , 116 , and then to oblate shapes at N = 118 , 120 and spherical shapes at N = 122 , 124, where the shell closure at N = 126 is approached. While SC around the critical point of the prolate-to-oblate SPT has been seen in 182 , 184 Er 114 , 116 and 184 , 186 Yb 114 , 116 , this is not the case in 190 , 192 Os 114 , 116 [244], since the transition from prolate to oblate shapes becomes smooth at higher Z.
Potential energy surfaces in the region 206 216 Os 130 140 have been calculated [207] using an IBM Hamiltonian with parameters determined from mean-field calculations with the Skyrme force (see Figure 21 of [207] for the relevant PES). 210 Os 134 has been suggested [207] as a possible candidate for the E(5) CPS, its level scheme given in Figure 25 of [207].
Recent calculations [151] within the deformed RHB theory have indicated that bubble structures with SC can be obtained in the extremely neutron-rich nuclei 196 , 206 , 208 , 256 Os 120 , 130 , 132 , 180 .
In summary, some experimental and theoretical support exists for SC and shape/phase mixing in the 174 180 Os 98 104 region.

4.2. The W (Z = 74) Isotopes

The 0 + states in 180 184 W 106 110 have attracted considerable experimental attention [490,505,506], since the enhanced density of low-lying 0 + states has been considered as evidence for shape/phase transitional behavior [505,506]. However, such an enhanced density has been located in 154 Gd 90 and not in 180 , 184 W 106 , 110 [506].
The 184 196 W 110 122 region has attracted much theoretical attention, since SC is expected to occur in relation to the SPT from prolate to oblate shapes seen around N = 116 [324], which implies the occurrence of SC around it. Potential energy surfaces revealing this transition have been calculated within a self-consistent HFB approach [616], a Skyrme HF plus BCS framework [145], as well as with an IBM Hamiltonian with parameters determined from the microscopic energy density functional Gogny D1M [218] and Gogny D1S [209] (see Figure 2 of [616], Figure 2 of [145], Figure 1 of [218], and Figure1 of [209] for the corresponding PES, as well as Figure 5 of [209] for relevant level schemes).
More recently, extended calculations with a 5DQCH with parameters determined from CDFT have been performed in 176 198 W 102 124 [244] (see Figure 4 of [244] for the relevant potential energy surfaces). An evolution is seen from well-deformed prolate shapes at N = 102 (near to the midshell N = 104 ) to prolate–oblate SC in 188 , 190 W 114 , 116 , and then to oblate shapes at N = 118 , 120 and spherical shapes at N = 122 , 124, where the shell closure at N = 126 is approached. While SC around the critical point of the prolate-to-oblate SPT has been seen in 182 , 184 Er 114 , 116 and 184 , 186 Yb 114 , 116 , this is not the case in 188 , 190 W 114 , 116 [244], since the transition from prolate to oblate shapes becomes smooth at higher Z.
Potential energy surfaces in the region 204 214 W 130 140 have been calculated [207] using an IBM Hamiltonian with parameters determined from mean-field calculations with the Skyrme force (see Figure 21 of [207] for the relevant PES). 208 W 134 has been suggested [207] as a possible candidate for the E(5) CPS, its level scheme given in Figure 25 of [207].
Recent calculations [151] within the deformed RHB theory have indicated that bubble structures with SC can be obtained in the extremely neutron rich nuclei 194 , 204 , 254 W 120 , 130 , 180 .
In summary, SC in the W isotopes is not predicted anywhere, not even in the N = 114 , 116 region, in which an SPT from prolate to oblate shapes is predicted in the Er and Yb series of isotopes.

4.3. The Hf (Z = 72) Isotopes

The 166 170 Hf 94 98 isotopes have attracted considerable attention, since they lie close to the Nd, Sm, Gd, and Dy N = 90 isotones, which are considered as manifestations of the X(5) CPS. Experimental investigations in 166 Hf 94 [379,380] have shown reasonable agreement of spectra and B(E2)s to the X(5) CPS, which is using an infinite square well potential in the β collective variable, while in 168 Hf 96 [381] a Davidson potential appears to provide better results, and in 170 Hf 98 [282,382] a sextic potential possessing two minima appears to be the most appropriate one.
The 0 + states in 174 178 Hf 102 106 have attracted much experimental attention [490,492,505,506,512], since the enhanced density of low-lying 0 + states has been considered as evidence for shape/phase transitional behavior [505,506]. However, such an enhanced density has been located in 154 Gd 90 and not in 176 Hf 104 [506]. A theoretical description for 176 Hf 104 has been provided within the pseudo-SU(3) symmetry [508,509].
The 182 194 Hf 110 122 region has attracted much theoretical attention, since SC is expected to occur in relation to the SPT from prolate to oblate shapes seen around N = 116 [324], which implies the occurrence of SC around it. Potential energy surfaces revealing this transition have been calculated within a self-consistent HFB approach [616], a Skyrme HF plus BCS framework [145], as well as with an IBM Hamiltonian with parameters determined from the microscopic energy density functional Gogny D1M [218] (see Figure 2 of [616], Figure 2 of [145], and Figure 1 of [218] for the corresponding PES).
More recently, extended calculations with a 5DQCH with parameters determined from CDFT have been performed in 174 196 Hf 102 124 [244] (see Figure 3 of [244] for the relevant potential energy surfaces). An evolution is seen from well-deformed prolate shapes at N = 102 (near to the midshell N = 104 ) to prolate–oblate SC in 186 , 188 Hf 114 , 116 , and then to oblate shapes at N = 118 , 120 and spherical shapes at N = 122 , 124, where the shell closure at N = 126 is approached. While SC around the critical point of the prolate-to-oblate SPT has been seen in 182 , 184 Er 114 , 116 and 184 , 186 Yb 114 , 116 , this is not the case in 186 , 188 Hf 114 , 116 [244], since the transition from prolate to oblate shapes becomes smooth at higher Z.
Recent calculations [151] within the deformed RHB theory have indicated that bubble structures with SC can be obtained in the extremely neutron rich nuclei 202 , 234 , 236 , 240 , 250 Hf 130 , 162 , 164 , 168 , 178 .
In summary, SC in the Hf isotopes is not predicted anywhere, not even in the N = 114 , 116 region, in which an SPT from prolate to oblate shapes is predicted in the Er and Yb series of isotopes.

4.4. The Yb (Z = 70) Isotopes

The 0 + states in 168 172 Yb 98 102 have attracted much experimental attention [490,506,512], since the enhanced density of low-lying 0 + states has been considered as evidence for shape/phase transitional behavior [505,506]. However, such an enhanced density has been located in 154 Gd 90 and not in 170 Yb 100 [506]. A theoretical description for the many 0 + states in 170 Yb 100 has been provided within a monopole pairing plus quadrupole–quadrupole model with a spin quadrupole force [511], while for 172 Yb 102 the pseudo-SU(3) symmetry [508,509] has been used.
The 158 162 Yb 88 92 isotopes [241,379,618] have attracted considerable attention, since 160 Yb 90 is in line with the Nd, Sm, Gd, and Dy N = 90 isotones, which are considered as manifestations of the X(5) CPS. Calculations have been performed within the cranked HFB approximation [618], and a 5DQCH with parameters obtained from CDFT [241]. Marginal relation to the X(5)CPS has been found [379]. Extensive calculations [143] in 152 170 Yb 82 100 , performed with a 5DQCH with parameters determined from three different mean-field solutions have demonstrated the significant influence of the pairing correlations on the speed of the development of collectivity with increasing neutron number.
The 180 192 Yb 110 122 region has attracted much theoretical attention, since SC is expected to occur in relation to the SPT from prolate to oblate shapes seen around N = 116 [324], which implies the occurrence of SC around it. Potential energy surfaces revealing this transition have been calculated within a self-consistent HFB approach [616], a Skyrme HF plus BCS framework [145], as well as with an IBM Hamiltonian with parameters determined from the microscopic energy density functional Gogny D1M [218] (see Figure 2 of [616], Figure 2 of [145], and Figure 1 of [218] for the corresponding PES).
More recently, extended calculations with a 5DQCH with parameters determined from CDFT have been performed in 172 194 Yb 102 124 [244] (see Figure 2 of [244] for the relevant potential energy surfaces). An evolution is seen from well-deformed prolate shapes at N = 102 (near to the midshell N = 104 ) to prolate–oblate SC in 184 , 186 Yb 114 , 116 , and then to oblate shapes at N = 118 , 120 and spherical shapes at N = 122 , 124, where the shell closure at N = 126 is approached. The level scheme of 186 Yb 116 can be seen in Figure 11 of [244]. The occurrence of SC around the critical point of the prolate-to-oblate SPT seen in this exotic region bears great similarity to the SC seen in Section 5.5 and Section 8.5.
In summary, SC in the Yb isotopes is predicted only in the N = 114 , 116 region, in which an SPT from prolate to oblate shapes is predicted. Once more, the close relation between SC and SPT is pointed out.

4.5. The Er (Z = 68) Isotopes

The 0 + states in 168 Er 100 have attracted much experimental attention [504,505,506,512], since the enhanced density of low-lying 0 + states has been considered as evidence for shape/phase transitional behavior [505,506]. However, such an enhanced density has been located in 154 Gd 90 and not in 168 Er 100 [505,506]. Theoretical descriptions for the many 0 + states in 168 Er 100 have been provided within the quasiparticle–phonon model [504,520] and the projected shell model [504], as well as within the pseudo-SU(3) symmetry [508,509].
The development of nuclear deformation, leading to the emergence of simplicity and SC has been theoretically studied in 168 Er 100 in the framework of the symplectic shell model [619], and in 166 Er 98 within an algebraic many-nucleon version of the Bohr–Mottelson unified model [620]. Detailed level schemes for 166 Er 98 [239] have been provided by a 5DQCH with parameters determined through the relativistic energy density functionals DD-PC1 and PC-F1 (see Figures 5 and 8 of [239] for the corresponding level schemes).
The 156 160 Er 88 92 isotopes have attracted considerable attention, since 158 Er 90 is a neighbor of the Nd, Sm, Gd, and Dy N = 90 isotones, which are considered as manifestations of the X(5) CPS. Calculations have been performed within a hybrid collective model [615], a quartic potential [503], the cranked HFB approximation [618], and a 5DQCH with parameters obtained from CDFT [241]. Marginal relation to the X(5) CPS has been found.
Extended calculations with a 5DQCH with parameters determined from CDFT have been performed in 170 192 Er 102 124 [244] (see Figure 1 of [244] for the relevant potential energy surfaces). An evolution is seen from well-deformed prolate shapes at N = 102 (near to the midshell N = 104 ) to prolate–oblate SC in 182 , 184 Er 114 , 116 , and then to oblate shapes at N = 118 , 120 and spherical shapes at N = 122 , 124, where the shell closure at N = 126 is approached. The level scheme of 184 Er 116 can be seen in Figure 11 of [244]. The occurrence of SC around the critical point of the prolate-to-oblate SPT seen in this exotic region bears great similarity to the SC seen in Setions Section 5.5 and Section 8.5.
In summary, SC in the Er isotopes is predicted only in the N = 114 , 116 region, in which an SPT from prolate to oblate shapes is predicted. Once more, the close relation between SC and SPT is pointed out.

5. The Z ≈  60 , N 90 Region

5.1. The Sm (Z = 62) Isotopes

5.1.1. Sm (Z = 62) Isotopes above N = 82

The change of equilibrium (ground state) deformation occurring between 150 Sm 88 and 152 Sm 90 has been known since 1966 [621], while the question of SC in 152 Sm 90 has been raised since 1969 [622], albeit in relation to the 0 3 + state.
The idea of two coexisting phases in 152 Sm 90 , related to the transition from spherical U(5) to deformed SU(3) shapes, and corresponding in the IBM framework to a potential energy surface with two minima, was proposed in 1998 [412], followed by extensive experimental work [368,371] supporting this idea. Further evidence in favor of this description has been provided both in the IBM [367] and the geometrical collective model [373] frameworks, while the question regarding the occurrence of a phase transition in a system characterized by discrete integer nucleon numbers has been addressed in [623].
In relation to this SPT, the concept of the X(5) CPS was introduced by Iachello [451] in 2001, using an approximate solution of the Bohr Hamiltonian with an infinite square well potential in the β variable (which can be considered as a crude approximation to a double-well potential with two wells of equal depth), and a steep harmonic oscillator potential in the γ variable, centered at γ = 0 and therefore corresponding to prolate deformed shapes. Characteristic parameter-independent predictions of the X(5) CPS is the R 4 / 2 ratio for the ground state band, which is 2.91, and the relative excitation energy of the first excited 0 + state, E ( 0 2 + ) / E ( 2 1 + ) , which should be 5.67, both of them being very close to the experimental values found for 152 Sm 90 . The parameter-independent (up to an overall scale factor) predictions of the X(5) CPS for spectra and B(E2) transition rates involving the ground state and other K = 0 bands have been found to be in good agreement with the data for 152 Sm 90 [356,358,360,366,372]. Good agreement has also been found for the γ -band, in the case of which a free parameter enters [364].
The nature of the 0 2 + state of 152 Sm 90 has been thoroughly discussed, since it plays a central role in the X(5) CPS. While initially this state was thought as being a clear example of a β -vibration in a deformed nucleus [365,490], within the X(5) CPS it appears as a spherical state coexisting with the deformed ground state band (gsb). (It is worth noticing at this point that the gsb and the K = 0 bands based on 0 2 + and 0 3 + in 152 Sm 90 have R 4 / 2 ratios of 3.01, 2.69, and 2.53 respectively [406]). The characterization of the bands based on the 0 2 + states in the N = 90 region as β -vibrations has been questioned [525], starting with the 0 2 + state in 154 Gd 90 [522,523,524]. Furthermore, no experimental evidence for multiphonon structures has been found in 152 Sm 90 [369,370,518].
As mentioned above, before the introduction of the X(5) CPS in 2001 [451], the term phase coexistence was used [367,368,371,412,623], in order to distinguish this effect from SC, which was already identified as connected to intruder bands arising from particle–hole excitations [490]. The term phase/shape coexistence started also being used [373]. It was also suggested that 152 Sm 90 indicates a complex example of SC instead of a critical point nucleus [624]. Given the recent work on the close relation between SC and SPTs [439,469], this question appears to have been answered. The study in Ref. [439] is particularly relevant, since it regards the Z 40 , N 60 region, which bears great similarity to the present Z 60 , N 90 region, as we will further see below in Section 8.5.
150 Sm 88 has been found [363] to lie very close to the SPT from spherical to prolate deformed shapes. Recent lifetime measurements [625] support the appearance of SC in it. See also Ref. [626] for an early comparison between 150 Sm 88 and 70 Ge 38 , supporting the occurrence of SC in both.
On the theoretical side, spectra and B(E2)s of 152 Sm 90 have been successfully described in the geometric collective model [373] and in the dynamic pairing plus quadrupole (DPPQ) model [627].
More recently, extended calculations have been performed in the region 144 158 Sm 82 96 within several schemes. Early RMF calculations [157] found relatively flat potential surfaces for 148 152 Sm 86 90 , in agreement with the idea of an SPT. Subsequent calculations using the self-consistent Skyrme-HF+BCS approach [144], the self-consistent HFB approximation with the D1S parametrization of the Gogny interaction [135,137], as well as relativistic HB calculations using the DD-ME2 force [158] have found potential energy curves with two minima for 152 Sm 90 and its neighbors 150 , 154 Sm 88 , 92 , approximating the flat potential needed for an SPT (see Figure 9 of [144], Figure 10 of [135], Figure 4 of [137], Figure 2 of [158]). Recent calculations using CDFT with the DDME2 functional with standard pairing included, have found that neutron particle–hole excitations induced by the protons occur in 152 , 154 Sm 90 , 92 [579,580].
More detailed calculations in the region 144 158 Sm 82 96 , or parts of it, in which spectra and B(E2)s are obtained, have been performed with a 5DQCH with parameters calculated from CDFT [231,233,235,241] (for level schemes see Figure 12 of [231] and Figure 5 of [233]). In addition, IBM calculations have also been performed, with parameters determined from the Skyrme interaction [207,226] and from the PC-PK1 energy density functional [228]. In the latter calculation, one can see that the addition of pairing significantly lowers the energy of the excited K = 0 bands, thus improving agreement to the data (see Figure 10 of [228]). Finally, the 0 + states of 156 Sm 94 have been successfully described [508,509] in the framework of the pseudo-SU(3) symmetry.
In summary, signatures of SC have been seen in the critical nucleus 152 Sm 90 and its neighbors 150 , 154 Sm 88 , 92 .

5.1.2. Sm (Z = 62) Isotopes below N = 82

130 144 Sm 68 82 have been described [321] in the quasi-SU(3) framework. It is found that for the neutrons, the orbitals of the s d g shell, plus the intruder orbital 1 h 11 / 2 , coming down from the p f h shell, do not suffice to explain the large B(E2)s observed in 134 Sm 72 , and that the neutron 2 f 7 / 2 orbital, coming from the 82–126 shell, has to be added to the model space for large enough collectivity to be obtained. The simultaneous inclusion of the 1 h 11 / 2 and 2 f 7 / 2 orbitals, differing by Δ j = 2 , is the hallmark of the quasi-SU(3) symmetry. In other words, neutron particle–hole excitations from the 1 h 11 / 2 to the 2 f 7 / 2 orbital are needed in order to account for the increased collectivity, thus supporting microscopically the possible occurrence of SC in 134 Sm 72 .
Within a completely different approach, the isotopes 128 142 Sm 66 80 have been described [243] by a 5DQCH with parameters determined from the relativistic energy density functional PC-PK1. In this case, the first two 0 + states in 136 Sm 74 are found to possess significantly different deformation, again signifying the presence of SC.
In addition, experimental signs of SC have been found in the near spherical nucleus 142 Sm 80 [628], as well as in 140 Sm 78 [629], where the onset of deformation below N = 82 occurs. The role of the intruder neutron orbitals in the onset of deformation has been emphasized in [628].
In summary, evidence for SC starts appearing in the neutron-deficient Sm isotopes with N < 82 , the isotopes with N = 72 , 74, 78, 80 obtained as the first candidates.

5.2. The Gd (Z = 64) Isotopes

5.2.1. Gd (Z = 64) Isotopes above N = 82

The analogy between the 0 2 + states of 154 Gd 90 and 152 Sm 90 was realized early [490], as well as the similarity of the spectrum and B(E2)s of 154 Gd 90 to the predictions of the X(5) CPS [356,375,376]. The lack of β -vibrations in 152 156 Gd 88 92 has also been realized [521,522,523,524,525], as in the case of the Sm isotopes. The 0 2 + state of 152 Gd 88 has been found [374] not to be of intruder nature.
A large number of excited 0 + states has been found in 158 Gd 94 and 156 Gd 92 [492,493,494,495], 13 and 6 respectively. The 0 6 + state (2276.66 keV) of 158 Gd 94 [495], as well as the 0 4 + state (1715.2 keV) of 156 Gd 92 [492] have been found to correspond to two-phonon γ γ vibrations. The enhanced density of low-lying 0 + states in 154 Gd 90 has been suggested [505,506] as a new signature of shape/phase transitional behavior near the critical point (see Figure 4 of [505] and Figure 11 of [506]).
The large number of excited 0 + states appearing in 156 , 158 Gd 92 , 94 has been interpreted in terms of the geometric collective model [496], the IBM [496], the quasiparticle–phonon model [500], the pairing plus quadrupole model [501], and the pseudo-SU(3) symmetry [497,498,499,508,509].
Calculations extended at least over the 152 156 Gd 88 92 region, focusing on the SPT from spherical to prolate deformed shapes, have been performed using the RMF theory with the NL3 interaction [164], the Skyrme-HF plus BCS approach [144], the beyond-mean-field method with the Gogny D1S interaction [135,137], and the relativistic HFB approach using the DD-ME2 parametrization [158]. In all cases a double-well potential energy surface has been obtained in the 152 156 Gd 88 92 region, resembling the flat potential needed for a SPT (see Figure 2 of [164], Figure 11 of [144], Figure 11 of [135], Figure 4 of [137], Figure 2 of [158]). In addition, adiabatic TDHFB calculations [181] have been able to reproduce the X(5)-like behavior of 154 Gd 90 . Recent calculations using CDFT with the DDME2 functional with standard pairing included, have found that neutron particle–hole excitations induced by the protons occur in 154 , 156 Gd 90 , 92 [579,580].
Recently, more detailed calculations, allowing for predictions of the energy levels and the B(E2)s among them, have been performed in the 152 156 Gd 88 92 region and around it, using a 5DQCH with parameters determined through RMF theory [231,233,241]; see Figure 13 of [231] and Figure 6 of [233] for relevant level schemes. The potential energy surface determined for 154 Gd 90 in [241] exhibits a soft potential in β without two coexisting minima, in agreement with the flat potential expected for an SPT in the framework of the X(5) CPS.
Furthermore, detailed calculations allowing for predictions of the energy levels and the B(E2)s among them have been performed in the 152 156 Gd 88 92 region and around it, using the IBM Hamiltonian with parameters determined from CDFT [226,228]. The addition of pairing in these calculations [228] lowers the energy of the excited 0 + states, improving agreement to the data (see Figure 11 of [228]).
In conclusion, two competing pictures arise from the theoretical calculations in relation to the critical nucleus 154 Gd 90 . On one hand, most calculations predict potential energy surfaces with two minima of comparable depth in the β direction, which support the presence of SC, and can be considered as an approximation to the flat-bottom potential needed in order to achieve an SPT. On the other hand, some calculations predict flat potential energy surfaces, which are in perfect agreement with the flat-bottom potential needed in order to achieve an SPT, but do not support SC. The sensitivity of these predictions to the parameter sets used in each case has been pointed out by many authors. Theoretical approaches based on first principles, such as symmetries, can be helpful in resolving this issue.
In summary, signatures of SC have been seen in the 152 156 Gd 88 92 region, surrounding the critical point of the SPT from spherical to prolate deformed shapes.

5.2.2. Gd (Z = 64) Isotopes below N = 82

Four-particle four-hole excitations have been suggested in 146 Gd 82 [85], corresponding to proton 2p-2h excitations across the Z = 64 gap and neutron 2p-2h excitations across the N = 82 gap.

5.3. The Dy (Z = 66) Isotopes

Experimental attention has been focused on 156 Dy 90 , which has been found [356,375,377,378,503] to be a reasonable candidate for the X(5) CPS, albeit of lower quality [375,378] in comparison to the Nd, Sm, and Gd N = 90 isotones, since the γ degree of freedom appears to play a more important role here. Twelve 0 + states have been observed in 162 Dy 96 [505,506].
The seven excited 0 + states appearing in 164 Dy 98 have been interpreted in terms of the pseudo-SU(3) symmetry [508,509].
Calculations extended over the 148 162 Dy 82 96 region, or parts of it, focusing on the SPT from spherical to prolate deformed shapes, have been performed using the RMF theory with the NL3 interaction [164], the Skyrme-HF plus BCS approach [144], the self-consistent HFB method with the Gogny D1S interaction [135], and the relativistic HFB approach using the DD-ME2 parametrization [158]. In all cases a double-well potential energy surface has been obtained in the 154 158 Dy 88 92 region, resembling the flat potential needed for an SPT(see Figure 3 of [164], Figure 12 of [144], Figure 11 of [135], Figure 2 of [158]). The fact that 156 Dy 90 represents a poor candidate for the X(5) CPS, in comparison to its N = 90 neighbors in the Nd, Sm, and Gd chains of isotopes, has been emphasized [158].
Calculations allowing for predictions of the energy levels and the B(E2)s among them, have been recently performed for 156 Dy 90 and its neighbors, using a 5DQCH with parameters determined through RMF theory [233,241], see Figure 7 of [233] for a relevant level scheme. The potential energy surface determined for 156 Dy 90 in [241] exhibits a soft potential in β without two coexisting minima, in agreement with the flat potential expected for an SPT in the framework of the X(5) CPS.
Furthermore, detailed calculations allowing for predictions of the energy levels and the B(E2)s among them, have been performed in the 150 162 Dy 84 96 region, using the IBM Hamiltonian with parameters determined from CDFT [226,228]. The addition of pairing in these calculations [228] lowers the energy of the excited 0 + states, improving agreement to the data (see Figure 12 of [228]).
The critical point nucleus 156 Dy 90 has been adequately described in a hybrid collective model [615], revealing a sizeable shape/phase mixing.
In summary, signatures of SC have been seen in the critical nucleus 156 Dy 90 and its neighbors 154 , 158 Dy 88 , 92 .

5.4. The Nd (Z = 60) Isotopes

5.4.1. Nd (Z = 60) Isotopes above N = 82

Experimental work has been focused on 150 Nd 90 , which has been found to agree very well [356,358,360,361,362] with the parameter-independent predictions of the X(5) CPS.
Calculations extended over the 142 156 Nd 82 96 region, or parts of it, focusing on the SPT from spherical to prolate deformed shapes, have been performed using the RMF theory with the NL3 interaction [164] and the PC-F1 interaction [159], the Skyrme-HF plus BCS approach [144], the self-consistent HFB method with the Gogny D1S interaction [135], the beyond-mean-field method with the Gogny interaction [136], and the relativistic HFB approach using the DD-ME2 parametrization [158]. In all cases a double-well potential energy surface has been obtained in the 148 152 Nd 88 92 region, resembling the flat potential needed for an SPT (see Figure 1 of [164], Figure 1 of [159], Figure 8 of [144], Figure 10 of [135], Figure 4 of [136], Figure 2 of [158]). Recent calculations using CDFT with the DDME2 functional with standard pairing included have found that neutron particle–hole excitations induced by the protons occur in 150 , 152 Nd 90 , 92 [579,580].
A step further has been taken in Ref. [231], in which a 5DQCH with parameters determined through RMF theory has been used, allowing for the calculation of spectra and B(E2)s, see Figure 6 of [231] for the relevant level scheme of 150 Nd 90 . In Ref. [231] both the β and γ degrees of freedom have been included, in contrast to Ref. [159], in which only β has been considered. However, the potential energy surfaces obtained from the two calculations are very similar, as one can see by comparing Figure 2 of [231] to Figure 1 of [159].
Spectra and B(E2)s have also been calculated in Ref. [228], using an IBM Hamiltonian with parameters determined from CDFT [228]. The addition of pairing in these calculations lowers the energy of the excited 0 + states, improving agreement to the data, as one can see in Figure 9 of [228] for the level scheme of 152 Nd 92 .
Spectra and B(E2)s have also been successfully calculated for 150 Nd 90 , using a hybrid collective model with shape/phase mixing [615], as well as for 152 Nd 92 with the Bohr Hamiltonian with a sextic potential possessing two minima of equal depth [281,282]. The excited 0 + states in 152 Nd 92 have been considered within the pseudo-SU(3) approach in [508,509].
In summary, signatures of SC have been seen in the 148 152 Nd 88 92 region, surrounding the critical point of the SPT from spherical to prolate deformed shapes.

5.4.2. Nd (Z = 60) Isotopes below N = 82

128 142 Nd 68 82 have been described [321] in the quasi-SU(3) framework. It is found that for the neutrons, the orbitals of the s d g shell, plus the intruder orbital 1 h 11 / 2 , coming down from the p f h shell, do not suffice to explain the large B(E2)s observed in 130 , 132 Nd 70 , 72 , and that the neutron 2 f 7 / 2 orbital, coming from the 82–126 shell, has to be added to the model space for large enough collectivity to be obtained. The simultaneous inclusion of the 1 h 11 / 2 and 2 f 7 / 2 orbitals, differing by Δ j = 2 , is the hallmark of the quasi-SU(3) symmetry. In other words, neutron particle–hole excitations from the 1 h 11 / 2 to the 2 f 7 / 2 orbital are needed in order to account for the increased collectivity, thus supporting microscopically the possible occurrence of SC in 130 , 132 Nd 70 , 72 .
Within a completely different approach, the isotopes 126 140 Nd 66 80 have been described [243] by a 5DQCH with parameters determined from the relativistic energy density functional PC-PK1. In this case, the first two 0 + states in 134 Nd 74 are found to possess significantly different deformation, again signifying the presence of SC.
In summary, theoretical evidence for SC starts appearing in the neutron-deficient Nd isotopes with N < 82 , the isotopes with N = 70 –74 obtained as the first candidates.

5.5. Shape Coexistence and Shape/Phase Transition at N 90

In Figure 8 the ground state bands and the K π = 0 + bands invading them are plotted for N = 88 , 90, 92 in the Z 60 region.
In Figure 8a we see that for N = 88 the intruder levels form a parabola, presenting a minimum at Z = 64 , i.e., near the proton midshell ( Z = 66 ). In Figure 8b the same happens for N = 90 , while for N = 92 the minimum seems to be shifted towards Z = 68 , as seen in Figure 8c.
In the case of N = 88 the intruder 0 2 + states around Z = 64 fall below the 4 1 + members of the ground state band (gsb), while in the case of N = 90 the intruder 0 2 + states at Z = 60 –66 are nearly degenerate with the 6 1 + members of the gsb. In the case of N = 92 the intruder 0 2 + states at Z = 60 –66 are already lying far above the 6 1 + members of the gsb, but they are almost degenerate to them at Z = 68 , 70.
Several comments are in place.
The parabolas appearing in Figure 8 vs. Z for the N 90 isotones are similar to the parabolas appearing in Figure 7 and Figure 9 vs. N, for the Z 82 and Z 50 isotopes, respectively. In all cases, the minima of the parabolas appear near the relevant midshell. The similarity of the data suggests similarity in the microscopic mechanism leading to the appearance of these effects. In the cases of the Z 82 and Z 50 isotopes it is known [83,84,85,86,87] that the microscopic origin of the intruder states is lying in proton particle–hole excitations across the Z = 82 and Z = 50 shell closures, respectively. Therefore it is plausible that in the present case of the N 90 isotones the underlying microscopic origin should be related to neutron particle–hole excitations. Indeed, in [579,580] it has been suggested that, in this case, neutron particle–hole excitations occur from normal-parity orbitals to intruder orbitals, which of course bear the opposite parity. It should be noticed that the particle–hole excitations occurring across the Z = 82 and Z = 50 major shells also correspond to particle–hole excitations between orbitals of opposite parity, since the orbitals involved belong to different major shells.
The appearance of particle–hole excitations across Z = 82 and Z = 50 for relevant series of isotopes can be attributed to the influence of the changing number of neutrons on the constant set of protons. Thus, SC appearing in these cases has been called neutron-induced SC [579,580]. In a similar manner, particle–hole excitations across N = 90 for relevant series of isotones can be attributed to the influence of the changing number of protons on the constant set of neutrons. Thus, SC appearing in these cases has been called proton-induced SC [579,580]. This interpretation is consistent with the role played in the shell-model framework by the tensor force [102,105,106,107], which is known to be responsible for the influence of the protons on the neutron gaps, as well as the influence of the neutrons on the proton gaps, as described in Section 2.1.1.
The fact that for N = 88 and Z = 60 –64 the 0 2 + intruder levels remain close to the 4 1 + members of the ground state bands means that the K π = 0 + intruder band remains close to the ground state band, thus the appearance of SC is possible. SC remains plausible also for N = 90 , since the 0 2 + intruder levels remain close to the 6 1 + members of the ground state band, implying again that the K π = 0 + intruder band still remains quite close to the ground state band. In different words, 148 , 150 Nd 88 , 90 , 150 , 152 Sm 88 , 90 , 152 , 154 Gd 88 , 90 , and 154 , 156 Dy 88 , 90 are good candidates for the appearance of SC. This closeness deteriorates for Z = 60 –64 at N = 92 , which means that 152 Nd 92 , 154 Sm 92 , 156 Gd 92 , and 158 Dy 92 are not candidates for SC.
The fact that for N = 90 and Z = 60 –64 the 0 2 + intruder levels remain close to the 6 1 + members of the ground state bands implies that the isotones 150 Nd 90 , 152 Sm 90 , 154 Gd 90 , and 156 Dy 90 are good candidates for the X(5) CPS [451], of which the degeneracy of the 0 2 + and 6 1 + states is a hallmark. Indeed these isotones have been suggested as being the best manifestations of the X(5) CPS, as discussed in Section 2.6, with detailed relevant references given in the caption of Figure 2. On the other hand, the proximity of the 0 2 + states to the 6 1 + states seen in N = 92 for the Z = 68 , 70 cases, suggests that 160 Er 92 and 162 Yb 92 might be reasonable candidates for X(5). Indeed, such a suggestion exists [379] for the latter.
It should be noticed that the strong similarity between the N 90 , Z 64 and the N 60 , Z 40 regions has been pointed out by Casten et al. already in 1981 [630], suggesting a common microscopic interpretation of the onset of deformation in these two regions.

6. The Z = 54–58 Desert

6.1. The Ce (Z = 58) Isotopes

On the experimental side, 130 Ce 72 [356,359] and 128 Ce 70 [357] have been suggested as examples of the X(5) CPS.
Moving to lighter Ce isotopes, one sees in Figure 9 of [502] that the 0 2 + states in 128 134 Ce 70 76 exhibit a parabolic behavior with a minimum at 130 Ce 72 , while the 0 2 + state in 128 Ce 70 remains close to this minimum, suggesting a region of SC starting at 130 Ce 72 and extending to the lighter Ce isotopes.
It should be recalled that 128 134 Ce 70 76 are considered as good examples of the O(6) dynamical symmetry of IBM, as seen in Figure 2.4 of the book [214].
Extended RMF calculations have been performed [171] in the 120 142 Ce 62 84 region, using different interactions. Potential energy surfaces with two minima of similar depth and different deformation have been seen in 126 130 Ce 68 72 with the PK1 and NL3 interactions [171], as well as in 124 128 Ce 66 70 with the TM1 interaction [171].
Recently, 126 140 Ce 68 82 have been described [321] in the quasi-SU(3) framework. It is found that for the neutrons, the orbitals of the s d g shell, plus the intruder orbital 1 h 11 / 2 , coming down from the p f h shell, do not suffice to explain the large B(E2)s observed in 124 , 126 Ce 66 , 68 , and that the neutron 2 f 7 / 2 orbital, coming from the 82–126 shell, has to be added to the model space for large enough collectivity to be obtained. The simultaneous inclusion of the 1 h 11 / 2 and 2 f 7 / 2 orbitals, differing by Δ j = 2 , is the hallmark of the quasi-SU(3) symmetry. In other words, neutron particle–hole excitations from the 1 h 11 / 2 to the 2 f 7 / 2 orbital are needed in order to account for the increased collectivity, thus supporting, microscopically, the possible occurrence of SC in 124 , 126 Ce 66 , 68 .
148 Ce 90 has been considered within a wider study of N = 90 isotones within a hybrid collective model [615], turning out to be a marginal case neighboring 150 Nd 90 , which is a clear example of the X(5) CPS.
In summary, no clear evidence for SC exists so far in the Ce isotopes, except for signs of neutron particle–hole excitations supporting the occurrence of SC in 124 , 126 Ce 66 , 68 , as well as some theoretical indications for a double-minimum potential in the 124 130 Ce 66 72 region, and/or the occurrence of the X(5) CPS in 128 , 130 Ce 70 , 72 .

6.2. The Ba (Z = 56) Isotopes

On the experimental side, 134 Ba 78 [366,391,394,403,404,405] is considered as the textbook example of the E(5) CPS, being the only unquestionable example of this symmetry identified so far. Further measurements [402] in 132 , 134 Ba 76 , 78 corroborate the transitional character from O(6) to U(5) symmetry in these nuclei, while 130 Ba 74 [507,631] is found to be close to the O(6) dynamical symmetry [507], in agreement to Figure 2.4 of the book [214], in which 126 132 Ba 70 76 appear as good examples of the O(6) dynamical symmetry of IBM.
Moving to lighter Ba isotopes, one sees in Figure 8 of [502] that the 0 2 + states in 122 134 Ba 66 78 exhibit a parabolic behavior with a minimum at 128 Ba 72 , while the 0 2 + states in 124 , 126 Ba 68 , 70 remain close to this minimum, suggesting a region of SC starting at 128 Ba 72 and extending to the lighter isotopes 124 , 126 Ba 68 , 70 .
Independently, 126 Ba 70 [356] and 122 Ba 66 [354,355] have been suggested as examples of the X(5) CPS, which is known to be related to the first-order SPT between spherical and deformed shapes and related to SC in the N = 90 region.
In summary, when moving down from the N = 82 shell closure to the N = 66 midshell one first encounters the E(5) CPS at N = 78 and the transition from U(5) shapes above N = 78 to O(6) shapes below N = 78 . Moving further down towards the midshell, one finds the border of a region of SC at N = 72 and nuclei exhibiting SC and/or the X(5) CPS below it.
On the theoretical side, potential energy surfaces (PES) for 120 138 Ba 64 82 have been calculated in the Skyrme-HF plus BCS approach [144], as well as in 130 134 Ba 74 78 in the self-consistent HFB approximation with the Gogny D1S interaction [135]. Relatively flat PES have been found for 132 , 134 Ba 76 , 78 , as seen in Figure 6 of [144] and in Figure 6 of [135], in agreement with the expectations for an SPT. On the other hand, two minima of different deformations appear in 120 126 Ba 64 70 , which could give rise to SC. Flat PES in 132 , 134 Ba 76 , 78 have also been obtained [238] by a 5DQCH with parameters determined from self-consistent RMF calculations, as seen in Figure 1 of [238], with the corresponding level scheme for 134 Ba 78 given in Figure 9 of [238]. Flat PES in 132 , 134 Ba 76 , 78 have also been found in IBM calculations with parameters determined by mean-field calculations with the Skyrme interaction [207], as well as in RHB calculations with the DD-ME2 interaction [158], as seen in Figure 10 of [207] and Figure 1 of [158].
IBM calculations in 124 132 Ba 68 76 with parameters determined from the Gogny D1M energy density functional [227] tend to overestimate the energy of the 0 2 + state, as seen in Figure 1 of [227]. The same problem is appearing in 124 132 Xe 70 78 [227], but it has been resolved in IBM calculations with parameters determined from the PC-PK1 energy density functional [228] in 122 Xe 68 by taking into account configuration mixing, i.e., by inclusion of the pairing vibrations, as seen in Figure 4 of [228]. Presumably the same remedy would work in the case of the Ba isotopes. The excited 0 + states in 130 134 Ba 74 78 have been considered within a Hamiltonian with pairing, quadrupole–quadrupole, and spin–quadrupole interactions in [510].
Extensive shell-model calculations [321] in the N = 56 –82 region have shown that the quasi-SU(3) symmetry, implying the inclusion of the 2 f 7 / 2 neutron orbital in addition to the 1 h 11 / 2 one, is needed in order to account for large B(E2) values in some Ce, Nd, and Sm isotopes, as described in the relevant subsections, but such need does not arise for Ba isotopes. However, more recent shell-model calculations [322] in the 126 136 Ba 70 80 region indicate that the inclusion of both the 2 f 7 / 2 and the 1 h 11 / 2 neutron orbitals is necessary for explaining the oblate to prolate SPT with N = 76 as the critical point. It should be remembered that the O(6) dynamical symmetry of IBM has been suggested as lying at the critical point of the oblate-to-prolate transition [453,454,455]. Therefore the inclusion of quasi-SU(3) in shell-model calculations appears to be connected to the creation of collectivity leading to the O(6) dynamical symmetry of IBM.
This finding points towards an underlying relation between the E(5) CPS, suggested to be present in 132 , 134 Ba 76 , 78 , as mentioned in the beginning of this subsection, the O(6) dynamical symmetry of the IBM, and the critical point of the oblate-to-prolate transition. The existence of the subalgebra O(5) in both the E(5) and O(6) symmetries, which guarantees seniority as a good quantum number, appears to play a leading role in this connection, further discussed in Section 2.7. It should be remembered that O(5) not only is a subalebgra of the O(6) and U(5) dynamical symmetries of the IBM, but in addition underlies the whole O(6) to U(5) transition region [482]. Therefore it also underlies the critical point of the relevant second-order SPT [411] from γ -unstable to spherical shapes, which has been called E(5) [450] in the framework of the Bohr Hamiltonian.
In summary, no clear evidence for SC exists so far in the Ba isotopes, except for some theoretical indications for a double-minimum potential in the 120 126 Ba 64 70 region and/or the occurrence of the X(5) CPS in this area. On the other hand, strong evidence is accumulated for an SPT from U(5) shapes above N = 78 to O(6) shapes below N = 78 at 132 Ba 76 and 134 Ba 78 and for the occurrence of the E(5) CPS in this area.

6.3. The Xe (Z = 54) Isotopes

On the experimental side, 128 Xe 74 [391,395,399], 130 Xe 76 [400,401], and 132 Xe 78 [400], have been considered as candidates for the E(5) CPS.
On the theoretical side, potential energy surfaces (PES) for 118 136 Xe 64 82 have been calculated both in the Skyrme-HF plus BCS approach [144] and in the self-consistent HFB approximation with the Gogny D1S interaction [135]. Relatively flat PES have been found for 128 132 Xe 74 78 , as seen in Figure 5 of [144] and in Figure 2 of [135], in agreement with the expectations for a SPT. On the other hand, two minima of different deformations appear in 118 124 Xe 64 70 , which could give rise to SC. Flat PES in 128 132 Xe 74 78 have also been obtained [238] by a 5DQCH with parameters determined from self-consistent RMF calculations, as seen in Figure 2 of [238], with the corresponding level schemes given in Figures 11–13 of [238]. Flat PES in 128 132 Xe 74 78 have also been found in IBM calculations with parameters determined by mean-field calculations with the Skyrme interaction [207], as well as in RHB calculations with the DD-ME2 interaction [158], as seen in Figure 11 of [207] and Figure 1 of [158].
IBM calculations in 124 132 Xe 70 78 with parameters determined from the Gogny D1M energy density functional [227] tend to overestimate the energy of the 0 2 + state, as seen in Figure 1 of [227]. This problem has been resolved in IBM calculations with parameters determined from the PC-PK1 energy density functional [228] in 122 Xe 68 by taking into account configuration mixing, i.e., by inclusion of the pairing vibrations, as seen in Figure 4 of [228].
Extensive shell-model calculations [321] in the N = 56 –82 region have shown that the quasi-SU(3) symmetry, implying the inclusion of the 2 f 7 / 2 neutron orbital in addition to the 1 h 11 / 2 one, is needed in order to account for large B(E2) values in some Ce, Nd, and Sm isotopes, as described in the relevant subsections, but such a need does not arise for Xe isotopes. However, more recent shell-model calculations [322] in the 124 134 Xe 70 80 region indicate that the inclusion of both the 2 f 7 / 2 and the 1 h 11 / 2 neutron orbitals is necessary for explaining the oblate to prolate SPT with N = 76 as the critical point. It should be remembered that the O(6) dynamical symmetry of IBM has been suggested as lying at the critical point of the oblate-to-prolate transition [453,454,455]. Therefore, the inclusion of quasi-SU(3) in shell-model calculations appears to be connected to the creation of collectivity leading to the O(6) dynamical symmetry of IBM.
This finding points towards an underlying relation between the E(5) CPS, suggested to be present in 128 132 Xe 74 78 , as mentioned in the beginning of this subsection, the O(6) dynamical symmetry of the IBM, and the critical point of the oblate-to-prolate transition. The existence of the subalgebra O(5) in both the E(5) and O(6) symmetries, which guarantees seniority as a good quantum number, appears to play a leading role in this connection, further discussed in Section 2.7. It should be remembered that O(5) is not only a subalebgra of the O(6) and U(5) dynamical symmetries of the IBM, but in addition underlies the whole O(6) to U(5) transition region [482]. Therefore it also underlies the critical point of the relevant second-order SPT [411] from γ -unstable to spherical shapes, which has been called E(5) [450] in the framework of the Bohr Hamiltonian.
In summary, no clear evidence for SC exists so far in the Xe isotopes, except for some theoretical indications for a double-minimum potential in the 118 124 Xe 64 70 region. On the other hand, strong evidence is accumulated for a SPT from oblate to prolate shapes between 130 Xe 76 and 132 Xe 78 and/or the occurrence of the E(5) CPS in this area.

7. The Z 50 Region

7.1. The Sn (Z = 50) Isotopes

The observation of SC due to proton particle–hole excitations across the magic number Z = 82 led to the question if similar excitations also occur across the magic number Z = 50 . Some findings are summarized in Figure 9b, in which again intruder states forming parabolas with minima at the midshell, here lying at N = 66 , are seen.
Since 1979, when collective intruder bands based on low-lying 0 2 + states corresponding to proton 2p-2h excitations were first observed in 112 118 Sn 62 68 [632], SC in the 110 120 Sn 60 70 region has been observed in several experiments [633,634,635,636,637,638,639,640,641], recently reviewed in [297,516] (see in particular Figure 1.32 of [297], reproduced as Figure 23 in [516]). The neighborhood below this region, namely 106 , 108 Sn 56 , 58 , has been investigated in [642,643,644,645], in which rotational bands have been found that are much higher in energy and starting at much higher angular momenta, thus providing no evidence of SC with the ground state band.
The low-lying 0 2 + collective states in the 112 118 Sn 62 68 isotopes have been interpreted in terms of proton 2p-2h excitations, mixing with the ground state band, both in a microscopic framework [86,87,646], as well as in extended versions of the IBM [430,647]. Calculations extending over much wider regions have been recently performed using the MCSM [72] and the shell model taking advantage of the quasi-SU(3) symmetry [321].
In the MCSM calculation [72], covering the 100 138 Sn 50 88 region, for both protons and neutrons the whole s d g shell, consisting of the 1 g 9 / 2 , 1 g 7 / 2 , 2 d 5 / 2 , 2 d 3 / 2 , 3 s 1 / 2 orbitals has been taken into account, along with the lower part of the p f h shell, consisting of the 1 h 11 / 2 , 2 f 7 / 2 , 3 p 3 / 2 orbitals. The major finding of this work was the explanation of the sudden increase of the B ( E 2 ; 0 1 + 2 1 + ) value, seen at 110 Sn 60 , in terms of proton excitations from the proton 1 g 9 / 2 orbital, which is in accordance with the earlier explanation of SC in terms of 2p–2h proton excitations in this region, having in mind that the 1 g 9 / 2 orbital is the one lying immediately below the Z = 50 magic gap, and further corroborating the experimental fact [642,643,644,645] that no SC appears at N = 58 and below it.
The shell-model calculations of Ref. [321] corroborate the above findings. Taking into account the full s d g shell plus the 1 h 11 / 2 orbital, one sees that the B ( E 2 ; 0 1 + 2 1 + ) values are underestimated near the neutron midshell, while this discrepancy goes away as soon as the 2 f 7 / 2 orbital is included in the calculation. This finding also clarifies the essence of the quasi-SU(3) symmetry, which is the inclusion of orbitals differing by Δ j = 2 , the 1 h 11 / 2 and 2 f 7 / 2 orbitals in the present case. It is remarkable that in the MCSM calculation of Ref. [72], the full set of orbitals differing by Δ j = 2 from the p f h shell, namely the 1 h 11 / 2 , 2 f 7 / 2 , 3 p 3 / 2 orbitals, have been taken into account.
The MCSM calculation [72] also predicts a second-order SPT around 116 Sn 66 from a moderately deformed phase to the pairing (seniority) phase. As we shall see below, this prediction is in accordance with the earlier suggestion [396] of 114 Cd 66 (also having N = 66 ) as a candidate for the E(5) CPS.
In summary, the experimental findings and theoretical predictions agree that SC is seen in the region bordered by 110 Sn 60 and 120 Sn 70 .

7.2. The Cd (Z = 48) Isotopes

The Cd isotopes, lying two protons below the magic number Z = 50 , are the analog of the Hg isotopes, which lie two protons below the magic number Z = 82 . Therefore, strong evidence for SC is expected to be seen in them. Some findings are summarized in Figure 9a, in which again intruder states forming parabolas with minima at the midshell ( N = 66 ), are seen.
Starting in 1977 [648], the first examples of SC in the Cd isotopes, interpreted in terms of proton 2p-4h excitations, were seen experimentally in 110 114 Cd 62 66 [648,649,650,651]. It was also in 110 , 112 Cd 62 , 64 that multiple SC, based on multiparticle–multihole proton excitations, were seen [177,178,652].
Systematics of energy levels revealed a parabola formed by the 0 2 + states, related to 2p-4h excitations, having its minimum at the neutron midshell N = 66 and extending from 106 Cd 58 to 120 Cd 72 (see Figure 1 of [513], Figure 1 of [653], Figure 1 of [654]). B(E2)s connecting the intruder band to the ground-state band corroborate this picture (see Figure 15 of [516] and Figure 5 of [517]).
The lower border of the SC region has been considered as a place of possible manifestation of the E(5) CPS. The 106 , 108 Cd 58 , 60 isotopes have been suggested as possible E(5) candidates [391,395]. Recent investigations [655] suggest that at 106 Cd 58 the onset of collective rotation occurs (see also Figure 4 of [655], in which the energy systematics suggest that intruder 0 2 + and 0 3 + bands start from 106 Cd 58 towards heavier Cd isotopes).
The upper border of the SC region appears to be reached at 118 Cd 70 , which is considered as a textbook example of a near-harmonic vibrational nucleus [656], since, in addition to the two-phonon triplet, the three-phonon quintuplet is also seen. The arguments for and against vibrational motion in the region 110 116 Cd 62 68 , i.e., at the center of the region exhibiting SC, have been considered in detail in [514,515].
Already in 1990 it was seen [653,654] that two sets of low-lying 0 + states cross in energy between 114 Cd 66 and 116 Cd 68 in what can be interpreted as an early sign of multiple SC, confirmed recently in 110 , 112 Cd 62 , 64 [177,178,652].
114 Cd 66 has been suggested as a candidate for the E(5) CPS in [396], in accordance with the recent suggestion [72] in the framework of MCSM calculations of a second order SPT around 116 Sn 66 (also having N = 66 ) from a moderately deformed phase to the pairing (seniority) phase.
From the theoretical point of view, SC in 112 , 114 Cd 64 , 66 was attributed as early as in 1985 within the shell-model framework to 2p-4h proton excitations and configuration mixing [83,646]. It was within this framework that the close connection between SPTs and SC has been first considered in 2004 [469] in the N = 66 isotones, in accordance with the experimental observations mentioned above.
Within the IBM framework, the influence on SC of the O(5) as a common subalgebra in the U(5) and O(6) dynamical symmetries of the model has been considered [484,485], and detailed predictions for 108 114 Cd 60 66 have been given within the extended IBM framework, involving p-h excitations and configuration mixing.
More recently, calculations discussing SC in the 108 116 Cd 60 68 region have been performed using parameters determined microscopically through CDFT both in the IBM framework [225] and with a quadrupole collective Hamiltonian [657]. Furthermore, an extended calculation covering the region 96 136 Cd 48 88 [154] has been performed using CDFT, corroborating the oblate–prolate SC in 110 114 Cd 62 66 .
In summary, the experimental findings and theoretical predictions agree that SC is seen in the region starting at N = 58 or 60 and ending at N = 70 , with special attention paid to the relation between SC and an SPT from moderately deformed to nearly spherical shapes seen at N = 66 .

7.3. The Te (Z = 52) Isotopes

The Te isotopes, lying two protons above the magic number Z = 50 , are the analog of the Po isotopes, which lie two protons above the magic number Z = 82 . Therefore, strong evidence for SC is expected to be seen in them. Some findings are summarized in Figure 9c, in which again intruder states forming a parabola with a minimum at the midshell ( N = 66 ), are seen.
Extended experimental investigations of the Te isotopes since 1985 by Rikovska et al. [658,659,660,661] concluded that SC appears in the region 116 124 Te 64 72 , corresponding to 4p-2h proton excitations, analogous to the 2p-4h proton excitations seen in the Cd isotopes. The upper limit of this region, 124 Te 72 , has been suggested as a manifestation of the E(5) CPS [391,397,398], while behavior similar to E(5), especially in relation to the decays from the low-lying 0 + levels, has been seen in the neighboring isotopes 122 Te 70 and 126 Te 74 [398].
In recent reviews [516,517], the 0 2 + states have been seen to form a parabola having its minimum at the midshell isotope 118 Te 66 (see Figure 28 of [516] and Figure 17 of [517]). SC has been confirmed in 118 124 Te 66 72 [516], while a 4p-2h proton excitations character for the 0 2 + states has been suggested in [517].
Theoretical investigations within IBM-2 [658,660,662] and the extended IBM [423,430], in which configuration mixing of the 4p-2h configuration with the ground-state 2p configuration has been used, have corroborated SC in 116 124 Te 64 72 , while the relation between SC and SPTs has been considered by [469] in 118 Te 66 , along with other N = 66 isotones. The upper limit of the SC region has been discussed [663] within a transitional IBM Hamiltonian taking advantage of the affine SU ( 1 , 1 ) ^ Lie algebra in the transitional region between U(5) and O(6), concluding [663] that SC is seen in 118 122 Te 66 70 , while 124 , 126 Te 72 , 74 are exhibiting E(5) features, in accordance with [391,397,398] mentioned above. Recently the spectral features of 118 124 Te 66 72 have been successfully reproduced [664] both in the IBM-1 and the dynamic pairing plus quadrupole model, while spectra and B(E2)s of 114 124 Te 62 72 have been calculated [286] using the Bohr collective Hamiltonian with an infinite square well potential possessing a step.
Early shell-model calculations [646] with configuration mixing corroborated SC in 116 120 Te 64 68 , while recently [321,322] extended shell-model calculations taking advantage of the quasi-SU(3) symmetry, as described above (see the second half of Section 7.1) for the Sn isotopes, have been performed in 108 134 Te 56 82 . In addition, extended RHB calculations [163] in the 104 144 Te 52 92 region have corroborated SC in 116 120 Te 64 68 . Recent calculations using CDFT with the DDME2 functional with standard pairing included, have found that proton particle–hole excitations induced by the neutrons occur in 116 120 Te 64 68 [579,580].
In summary, the experimental findings and theoretical predictions agree that SC is seen in the region bordered by 116 Te 64 and 124 Te 72 , with possible occurrence of the E(5) CPS at the upper limit.

7.4. The Pd (Z = 46) Isotopes

The Pd isotopes, lying four protons below the magic number Z = 50 , are the analog of the Pt isotopes, which lie four protons below the magic number Z = 82 . Therefore some evidence for SC is expected to be seen in them.
Low-lying 0 + states with intruder bands built on them have been seen in 106 112 Pd 60 66 [665,666,667]. Systematics of 0 + states in the 102 116 Pd 56 70 region [517] show a parabola with a minimum at N = 64 , next to the N = 66 midshell. The observation of a strong ρ 2 ( E 0 ) in 106 Pd 60 (see Figure 21 in [668]) supports the SC case around this minimum, as also suggested by B(E2) transition rates (see Figure 22 in [517]).
102 Pd 56 [391,392,669] and 108 Pd 62 [393,394,395] have been suggested as candidates for the E(5) CPS, while in 100 Pd 54 spectra appear to be closer to the U(5) limit of IBM, but B(E2)s appear to be closer to the O(6) limit, indicating the absence of any IBM dynamical symmetry in this region.
Early theoretical considerations interpreted SC in the 108 112 Pd 62 66 region in terms of proton p-h excitations in the shell model [86], as well as in the extended IBM framework [430], while the appearance of the E(5) CPS in 102 Pd 56 has been supported by adiabatic TDHFB calculations [181].
More recently, extended calculations of potential energy surfaces covering the whole 96 128 Pd 50 82 region, or a large part of it, have been performed using mean-field methods with the Skyrme [144] and Gogny [135] interactions, as well as RMF methods using the DD-ME2 energy density functional [158]. In addition, detailed spectroscopic predictions have been obtained in the IBM framework [670]. These studies were focused on locating candidates for the E(5) CPS, with several suggestions ( 106 Pd 62 [670], 108 Pd 62 [135,144,158], 110 Pd 64 [144,158], 112 118 Pd 66 72 [135]) made, while in [158] it has been suggested that 102 Pd 56 is not a suitable candidate for E(5).
On the other hand, recent RMF calculations using the DD-PC1 and DD-PCX parametrizations [167] have suggested 108 Pd 62 as a clear example of prolate–oblate coexistence, locating in addition a prolate-to-oblate SPT at 110 Pd 64 and an oblate-to-prolate SPT at 120 Pd 74 . In addition, extended IBM calculations [207] with parameters determined microscopically from the Skyrme interaction have found weakly prolate structures for N = 60 –72, in rough agreement with [167].
In summary, evidence in favor of SC has been presented for 106 Pd 60 [666,667,668], 108 Pd 62 [167,430,667], 110 Pd 64 [430,665], 112 Pd 66 [86,665], i.e., in the region N = 60 –66, while 108 Pd 62 appears to be the most common suggestion for the E(5) CPS, thus offering a further argument in favor of the connection between SC and SPTs.

8. The Z 40 , N 60 Region

As pointed out in [630], the Z 40 , N 60 region is expected to be very similar to the Z 60 , N 90 region. This point will be further discussed in Section 8.5.

8.1. The Zr (Z = 40) Isotopes

8.1.1. Zr (Z = 40) Isotopes above N = 50

Strong experimental evidence for SC has been accumulated in 100 Zr 60 [671,672,673,674,675,676] and 98 Zr 58 [388,671,677,678], gradually extended to 96 Zr 56 [88,679,680] and 94 Zr 54 [681]. The sharp SPT taking place between 98 Zr 58 and 100 Zr 60 has been pointed out. In addition, a detailed study of 110 Zr 70 , lying at the 3D-HO magic numbers Z = 40 and N = 70 , showed no sign of stabilization of harmonic oscillator shell structure there [682].
It should be noticed that 96 Zr 56 and 98 Zr 58 represent two rare cases in which the 0 2 + band-head of a well-developed K π = 0 + band lies below the 2 1 + state of the ground state band [335]. It has been found [335] that such nuclei tend to be surrounded by nuclei exhibiting SC.
The 88 102 Zr 48 62 isotopes have been (along with the Mo isotopes in this region) the textbook example used by Federman and Pittel [470,471,472,473] for connecting nuclear deformation to the enhanced proton–neutron interaction among certain Nilsson orbitals, as described in Section 2.6. The abrupt change in the spectra of the Zr isotopes can be seen in Figure 1 of [470], Figure 1 of [472], and Figure 2 of [473], in which the large energy drop of the 2 1 + and 0 2 + levels when passing from 98 Zr 58 to 100 Zr 60 is clearly seen, signifying both the sudden increase of deformation (indicated by the drop of 2 1 + ) and the onset of SC (indicated by the drop of 0 2 + ), offering a dramatic manifestation of the close connection between SPTs and SC, discussed later in detail by García-Ramos and Heyde [439,683]. It is then not surprising that 100 Zr 60 has been suggested [350] as a candidate for the X(5) CPS.
Calculations focused on 98 , 100 Zr 58 , 60 have been carried out both in the mean-field framework using the Skyrme interaction [69,146] and the complex excited VAMPIR approach [190], as well as in the MCSM framework [66,69] and within a simple two-state-mixing model [684], confirming both the rapid change of structure occurring between 98 Zr 58 and 100 Zr 60 , as well as configuration mixing and SC in the latter. Furthermore, triple SC has been predicted in 98 Zr 58 [190] and 96 Zr 56 [196] within the complex excited VAMPIR approach, while SC in 96 Zr 56 has been corroborated through calculations [287,685] using a quadrupole collective Bohr Hamiltonian, later extended to the 92 102 Zr 52 62 region [288]. Recent calculations using CDFT with the DDME2 functional with standard pairing included, have found that neutron particle–hole excitations induced by the protons occur in 98 , 100 Zr 58 , 60 [579,580].
Within the last decade, calculations covering the whole N = 50 –70 region, or most of it, have been performed within several different theoretical frameworks.
MCSM calculations [71,679] have corroborated the sharp SPT taking place at N = 60 , and in addition have exhibited the difference between the SC occurring in 100 Zr 60 , in which a prolate band coexists with an oblate one, and in 110 Zr 70 , in which a prolate band coexists with a triaxial one. The crucial role played by the tensor force, discussed in Section 2.1, has been emphasized.
Extended CDFT calculations [150,155,166] have found SC in the 98 110 Zr 58 70 region, while in contrast the absence of SC in 88 92 Zr 48 52 has been demonstrated [150]. Similar results have been obtained using a 5DQCH [242] or an IBM Hamiltonian [221], with the relevant parameters in both cases obtained from CDFT.
Extended calculations [438,439] have also been performed within the IBM-CM framework, confirming the abrupt structural change from spherical to deformed shapes occurring at 100 Zr 60 and the close connection between SPT and SC. SC in 100 110 Zr 60 70 is confirmed, with the evolution of the intruder configuration across this region pointed out [439]. The latter point has been studied in detail in [440,441,442], suggesting that in the Zr isotopes two intertwined SPTs take place, sharing the same critical point at 100 Zr 60 . On one hand, the transition from normal (without particle–hole excitations) to intruder (involving particle–hole excitations) configurations takes place at N = 60 , signaling also the beginning of the SC region extending up to N = 70. On the other hand, the nature of the intruder band itself is varying from nearly spherical in 92 98 Zr 52 58 to prolate deformed in 102 104 Zr 62 64 and finally to γ -unstable shapes in 106 110 Zr 66 70 . It is worth pointing out that the last point is in accordance to the results of MCSM calculations, finding SC of prolate and triaxial shapes in 110 Zr 70 [71].
In summary, SC is well established both experimentally and theoretically in 94 100 Zr 54 60 , while theoretical predictions are extending it up to 110 Zr 70 .

8.1.2. Zr (Z = 40) Isotopes below N = 50

Early RMF calculations in the 76 90 Zr 36 50 region have shown SC of a spherical and a deformed minimum in 76 82 Zr 36 42 [156]. SC has been considered in 80 Zr 40 within the complex-excited VAMPIR framework [185,188], as well as within beyond-mean-field calculations involving the Gogny D1S interaction [138], revealing in it multiple SC of five 0 + states exhibiting different shapes. 80 Zr 40 has also been suggested as a candidate for the X(5) CPS [350]. Total-Routhian-surface calculations [686] have revealed SC in the region 80 84 Zr 40 44 . The same holds for Skyrme HF calculations [69] in 80 84 Zr 40 44 (see Figure 9 of [69]). Recent calculations using CDFT with the DDME2 functional with standard pairing included, have found that neutron particle–hole excitations induced by the protons, as well as proton particle–hole excitations induced by the neutrons, occur in 78 , 80 Zr 38 , 40 [579,580].
In summary, several theoretical predictions support the occurrence of SC in 76 84 Zr 36 44 .

8.2. The Sr (Z = 38) Isotopes

8.2.1. Sr (Z = 38) Isotopes above N = 50

Strong experimental evidence for SC in 96 , 98 Sr 58 , 60 has been accumulated [89,531,673,687,688,689,690,691,692], while signs of an abrupt shape transition from N = 58 to 60 have been identified [687,688,689,691].
Early theoretical calculations using a two-state mixing model [684] and the VAMPIR mean-field approach [190] have been focused on SC in 98 Sr 60 and 96 Sr 58 , respectively, finding triple SC in the latter. More recently, extended calculations have been carried out, covering large parts of the region 88 108 Sr 50 70 , using the RHB formalism [150], a 5DQCH with parameters determined from CDFT [242], the IBM with parameters determined from a Gogny energy density functional [221], and the IBM-CM [445]. In all cases an abrupt change of structure is seen between 96 Sr 58 and 98 Sr 60 , with SC predicted then up to 108 Sr 70 .
In MCSM calculations [691], in which the SPT occurring at N = 60 is attributed to type II shell evolution, proton particle–hole excitations from the p f shell into the 1 g 9 / 2 orbital are seen as the mechanism leading to SC. This means that p-h excitations across the Z = 40 subshell closure are supposed to play the major role in creating SC, in contrast to the opinion expressed in the review article [4] (see p. 1486 of [4]), that subshell structure does not play a major role in nuclear collectivity, since it is “smeared out” by the pairing interaction into a uniform distribution. The same reservation has been expressed in the review article [8] (see p. 71 of [8]) in relation to the similar region of SC occurring around N = 90 , Z = 64 .
In conclusion, SC is well established both experimentally and theoretically in 96 , 98 Sr 58 , 60 , while theoretical predictions are extending it up to 108 Sr 70 .

8.2.2. Sr (Z = 38) Isotopes below N = 50

Experimental evidence for SC has been seen in 82 Sr 44 [693], while 76 , 78 Sr 38 , 40 have been considered as possible X(5) candidates [350]. Along the N = Z line collectivity has been found [689] to increase strongly between 64 Ge 32 and 76 Sr 38 .
Early RMF calculations in the 74 88 Sr 36 50 region have shown SC of a spherical and a deformed minimum in 74 80 Sr 36 42 [156]. SC has been considered in the complex excited VAMPIR framerwork in 74 Sr 36 [192,195] and in 76 , 78 Sr 38 , 40 [185,188]. SC of spherical, prolate, and oblate shapes has been found by total-Routhian-surface calculations in 76 80 Sr 38 42 [686], while shell-model calculations along the N = Z line have shown [60] that a sudden jump of nucleons into the 1 g 9 / 2 2 d 5 / 2 orbitals occurs at N = Z = 36 , thus supporting SC in 76 Sr 38 . Recent calculations using CDFT with the DDME2 functional with standard pairing included, have found that neutron particle–hole excitations induced by the protons, as well as proton particle–hole excitations induced by the neutrons, occur in 78 Sr 40 [579,580].
In summary, SC is supported experimentally and/or theoretically in 74 82 Sr 36 44 .

8.3. The Mo (Z = 42) Isotopes

8.3.1. Mo (Z = 42) Isotopes above N = 50

Clear experimental evidence for SC has been reported in 98 Mo 56 [229,230], 100 Mo 58 [672,674,675], 102 Mo 60 [694]. 104 Mo 62 has been suggested [350,353] as a candidate for the X(5) CPS, based on energy spectra, but B(E2)s were found [351] to support a rather rotational behavior. The recent observation of low-lying 0 2 + states in 108 , 110 Mo 66 , 68 [695] appears to extend the SC region up to N = 68 . The lower border of the SC region appears to be 98 Mo 56 , since no evidence of SC has been seen in 96 Mo 54 [230]. Experimental information reviewed in [8,517,689] supports these borders. The energy of the 0 2 + states appears to be dropping fast up to N = 56 , being nearly stabilized up to N = 60 and then increasing very slowly up to N = 68 (see Figure 7 of [230] and Figure 29 of [8]).
The 92 106 Mo 50 64 isotopes have been (along with the Zr isotopes in this region) the textbook example used by Federman and Pittel [470,471,472,473] for connecting nuclear deformation to the enhanced proton–neutron interaction among certain Nilsson orbitals, as described in Section 2.6. The abrupt change in the spectra of the Mo isotopes can be visualized in Figure 5 of [473], in which the large energy drop of 0 2 + when passing from 96 Mo 54 to 98 Mo 56 is clearly seen, signifying the onset of SC at N = 56 . The drop of the 2 1 + state along the Mo chain of isotopes is gradual, in contrast to the Zr chain of isotopes, in which a dramatic drop is seen between N = 56 and 58.
Extended theoretical calculations covering the 92 112 Mo 50 70 region, or even going to higher N, have been performed recently using CDFT [149,155,166], predicting SC in the 98 108 Mo 56 66 region. Similar results have been obtained using a 5DQCH with parameters determined through CDFT using the PC-PK1 parametrization [242], and with an IBM Hamiltonian with parameters determined through CDFT using the Gogny-D1M energy density functional [221]. Recent calculations using CDFT with the DDME2 functional with standard pairing included have found that neutron particle–hole excitations induced by the protons occur in 102 Mo 60 [579,580].
The shape mixing evolution in the 96 100 Mo 54 58 has also been described [284] in terms of a Bohr Hamiltonian with a double-well sextic potential. The need for a double-well potential is related to a first order SPT (see, for example, Figure 10 of [449]).
In summary, SC is well established both experimentally and theoretically in 98 104 Mo 56 62 , while theoretical predictions are extending it up to 110 Mo 68 .

8.3.2. Mo (Z = 42) Isotopes below N = 50

Theoretical predictions for the occurrence of SC have been provided by complex excited VAMPIR calculations [185] in 84 Mo 42 , as well as by total-Routhian-surface calculations [686] in 84 , 86 Mo 42 , 44 .

8.4. The Ru (Z = 44) Isotopes

Despite the proximity of the Ru (Z = 44) isotopes to the Mo (Z = 42) ones, in which SC has been observed, there is very little evidence for SC in the Ru isotopes. The evolution of the 0 2 + states in the Ru isotopes (see Figure 32 of [8]) looks very similar to the corresponding plot for the Mo isotopes (see Figure 29 of [8]), dropping rapidly up to 98 Ru 54 , then dropping slowly up to 102 Ru 58 , then raising slowly up to 110 Ru 66 . The Ru isotopes beyond N = 60 appear to be certainly deformed [517], while signs of triaxiality have been seen in 110 Ru 66 [689]. In parallel, the presence of multiphonon spherical vibrations in the Ru isotopes has been recently completely ruled out [696].
On the theoretical side, early investigations [423,469] led to the conclusion that 110 Ru 66 presents a regular behavior, and therefore shows no signs of SC. 104 Ru 60 has been tested as a candidate for the E(5) CPS by IBM [389] and adiabatic TDHFB [181] methods, but no firm evidence supporting this suggestion has been found.
Recently, extended calculations have been performed, covering the region 96 124 Ru 52 80 , or most of it, using an IBM Hamiltonian with parameters fixed through microscopic calculations involving the Skyrme force [207] and the Gogny energy density functional [221], with no signs of SC found in the Ru isotopes, in contrast to the Sr, Zr, and Mo isotopes considered within the same framework [221]. The difference can be clearly seen from the configurations used in order to obtain satisfactory agreement with the data, reported on page 5 of [221]. While the mixing of two configurations was required in the case of the Mo isotopes, and mixing of three configurations was found necessary in the case of the Zr and Sr isotopes, a single configuration sufficed in the case of the Ru isotopes.
Extended calculations within the same wide region have been recently performed also using CDFT [149,166]. In both cases, only 104 Ru 60 has been singled out as a possible candidate for SC. It is of interest to compare Tables I and II of Ref. [149]. While for most of the Mo isotopes considered in that study, two configurations have been involved (as seen in Table II of [149]), in the case of the Ru isotopes only one configuration was found, as seen in Table I of [149], with the sole exception of 104 Ru 60 , in which two configurations have been found.
In summary, only 104 Ru 60 appears as a possible candidate for SC, in contrast to the Sr, Zr, and Mo series of isotopes, in which several candidates have been found.

8.5. Shape Coexistence and Shape/Phase Transition at N 60

In Figure 10 the ground-state bands and the intruder levels are plotted for N = 58 , 60, 62 in the Z 40 region.
In Figure 10a we see that for N = 58 the intruder levels form a parabola, presenting a minimum at Z = 42 , i.e., near the proton midshell ( Z = 39 ). In Figure 10b the same happens for N = 60 at Z = 38 , while for N = 62 the minimum seems to be flat, as seen in Figure 10c.
In the case of N = 58 , the intruder 0 2 + states in the region Z = 38 –44 fall below the 4 1 + members of the ground state band (gsb) and approach the 2 1 + levels, with 98 Zr 58 representing a rare case in which the 0 2 + state lies below the 2 1 + state, while for Z = 46 –50 they stay close to the 4 1 + state. In the case of N = 60 , the intruder 0 2 + states at Z = 38 –40 are nearly degenerate with the 2 1 + members of the gsb, while at Z = 42 -50 are nearly degenerate with the 4 1 + members of the gsb. In the case of N = 62 , the intruder 0 2 + states at Z = 38 –44 are already lying far above the 4 1 + members of the gsb at Z = 38 –44, still remaining below the 6 1 + members of the gsb though, but they are almost degenerate to the 4 1 + members of the gsb at Z = 46 –50.
The evolution seen in Figure 10 from N = 58 to 62 is very similar to the evolution seen in Figure 8 from N = 88 to 92, the main difference being that the intruder levels in Figure 10 appear to have sank lower within the gsb in comparison to what is seen in Figure 8. As a consequence, while for N = 90 near degeneracy is seen between the 0 2 + and 6 1 + states in the region Z = 60 –66, in the present case for N = 60 one sees degeneracy of the 0 2 + state to the 2 1 + state for Z = 38 , 40 and to the 4 1 + state for Z = 42 –50.
Several comments are in place.
The parabolas appearing in Figure 10 vs. Z for the N 60 isotones are similar to the parabolas appearing in Figure 7 and Figure 9 vs. N, for the Z 82 and Z 50 isotopes, respectively. In all cases the minima of the parabolas appear near the relevant midshell. The similarity of the data suggests similarity in the microscopic mechanism leading to the appearance of these effects. In the cases of the Z 82 and Z 50 isotopes, it is known [83,84,85,86,87] that the microscopic origin of the intruder states is lying in proton particle–hole excitations across the Z = 82 and Z = 50 shell closures, respectively. Therefore, it is plausible that in the present case of the N 60 isotones the underlying microscopic origin should be related to neutron particle–hole excitations. Indeed, in [579,580] it has been suggested that in this case neutron particle–hole excitations occur from normal parity orbitals to intruder orbitals, which of course bear the opposite parity. It should be noticed that the particle–hole excitations occurring across the Z = 82 and Z = 50 major shells, also correspond to particle–hole excitations between orbitals of opposite parity, since the orbitals involved belong to different major shells.
The appearance of particle–hole excitations across Z = 82 and Z = 50 for relevant series of isotopes can be attributed to the influence of the changing number of neutrons on the constant set of protons. Thus SC appearing in these cases has been called neutron-induced SC [579,580]. In a similar manner, particle–hole excitations across N = 60 for relevant series of isotones can be attributed to the influence of the changing number of protons on the constant set of neutrons. Thus SC appearing in these cases has been called proton-induced SC [579,580]. This interpretation is consistent with the role played in the shell-model framework by the tensor force [102,105,106,107], which is known to be responsible for the influence of the protons on the neutron gaps, as well as the influence of the neutrons on the proton gaps, as described in Section 2.1.1.
The fact that in practically all cases for N = 58 –62 and Z = 38 –50 the intruder 0 2 + levels remain below the 6 1 + members of the gsb, or are almost degenerate to them, suggests that the appearance of SC in all these nuclei is plausible. In different words, 96 100 Sr 58 62 , 98 102 Zr 58 62 , 100 104 Mo 58 62 , 102 106 Ru 58 62 , 104 108 Pd 58 62 , 106 110 Cd 58 62 , and 108 112 Sn 58 62 , are good candidates for the appearance of SC.
The strong similarity between the Z 40 , N 60 region and the Z 60 , N 90 region has a further consequence. The N = 90 isotones 150 Nd 90 , 152 Sm 90 , 154 Gd 90 , and 156 Dy 90 are known to be very good examples of the X(5) CPS, occurring at the critical point of a first order SPT from spherical to prolate deformed shapes. Because of the similarities mentioned above, it should be expected that in the region with Z = 38 –42 and N = 58 –62, i.e., among 96 100 Sr 58 62 , 98 102 Zr 58 62 , and 100 104 Mo 58 62 , some good examples of this SPT should occur, thus providing further connection between the SC and SPT concepts, as proposed by Heyde et al. [438,439,445,469,683]. However, it should be pointed out that the numerical hallmarks of this critical point symmetry might be different from those of X(5), since in the Z 40 , N 60 region the 0 2 + state appears to be closer to the 2 1 + and/or 4 1 + states and not nearly degenerate to the 6 1 + state. In other words, some formulation of the CPS more flexible than X(5) would be required in the Z 40 , N 60 region.
The above observations are compatible with past findings regarding the 104 Mo 62 nucleus, which had been suggested [353] as an example of X(5) based on spectral characteristics, but later measurements of B(E2) transition rates [351] have shown that the 104 , 106 Mo 62 , 64 isotopes are of rotational and not of critical nature.
It should be recalled that the strong similarity between the N 90 , Z 64 and the N 60 , Z 40 regions has been pointed out by Casten et al. already in 1981 [670], suggesting a common microscopic interpretation of the onset of deformation in these two regions.

9. The Z 34 , N 40 Region

9.1. The Kr (Z = 36) Isotopes

9.1.1. Kr (Z = 36) Isotopes below N = 50

Coexistence of a prolate and an oblate band has been seen experimentally in the region 72 76 Kr 36 40 [689,697,698,699,700,701,702,703,704], recently extended to 70 Kr 34 [705,706]. Plotting in Figure 11d the 0 2 + states [406] in the region 72 86 Kr 36 50 , one sees a rapid fall with decreasing N, reaching a plateau at 72 76 Kr 36 40 , which looks like the bottom of a parabola, of which the left branch is still missing. It should be noticed that 72 Kr 36 is one of the very few nuclei in which the 0 2 + state falls below the 2 1 + state.
Theoretical calculations have been focused on the 70 76 Kr 34 40 region. Shell-model calculations have shown the oblate shape of the ground state and the prolate–oblate SC in 72 Kr 36 [61], as well as the importance of 4p-4h excitations from ( 2 p 3 / 2 1 f 5 / 2 ) to ( 2 p 1 / 2 1 g 9 / 2 ) for the oblate configurations. The sudden structural change occurring along the N = Z line when passing from N = Z 35 to N = Z 36 is attributed to a sudden jump of nucleons into the 1 g 9 / 2 2 d 5 / 2 orbitals [60]. It has been shown [65] that this jump favors the creation of configurations with large prolate deformation, while configurations within the p f shell favor oblate deformation. The role of the tensor force in keeping these two different configurations close in energy, which is a prerequisite for obtaining SC, has been clarified [65]. The mixing between the oblate and prolate configurations in 72 Kr 36 has also been discussed in [46].
SC in 70 74 Kr 34 38 has been extensively discussed in the framework of the excited VAMPIR [184,185] and the complex excited VAMPIR [186,188,191,192,193,194,195]. A microscopic quadrupole collective Hamiltonian has been derived considering large amplitude collective motion [198,203,707], and SC and prolate–oblate mixing in 72 76 Kr 36 40 have been considered in it (see Figures 6, 8 and 11 of [203] and Figure 2 of [198] for the relevant level schemes). Mean-field calculations for 70 78 Kr 34 42 have been performed with the Skyrme interaction [141] and the Gogny D1S interaction [132,134] (for relevant level schemes see Figures 5, 9, and 13 of [141] and Figure 2 of [132]), as well as with a Bohr Hamiltonian with a sextic potential possessing two minima [279,708] (see Figure 5 of [279] and Figure 3 of [708] for relevant level schemes).
In addition, extensive mean-field calculations covering the whole 70 98 Kr 34 62 region have been performed [139] with the Gogny D1S interaction (see Figures 6–9 of [139] for relevant level schemes). SC is seen in 70 78 Kr 34 42 , and then again in 94 98 Kr 58 62 , with a gap with no SC appearing in between ( N = 44 –56).
Extensive calculations covering the whole 70 100 Kr 34 64 region have been performed [150] in the relativistic HF framework with the DD-PC1 parametrization, leading to two coexisting minima in 72 78 Kr 36 42 and then again in 92 , 96 100 Kr 56 , 60 64 (see Table I of [150]), as well as with the DD-ME2 parametrization, leading to two coexisting minima in 70 78 Kr 34 42 and then again in 92 100 Kr 56 64 (see Table II of [150]). Once more, there is some dependence on the details of the interactions, but they both agree that SC is observed roughly in the regions N = 34 –42 and N = 56 –64, with a large gap with no SC in between.
Extensive calculations covering the whole 70 100 Kr 34 64 region have also been performed [224] using an IBM Hamiltonian with parameters determined from a Gogny energy density functional. Two coexisting oblate and prolate minima have been found in 70 , 78 Kr 34 , 42 , and then again in 92 100 Kr 56 64 , while triple SC of spherical, prolate and oblate minima has been seen in 72 76 Kr 36 40 (see Table I of [224]). The same gap without SC in the region N = 44 –54 is seen, as in [150]. Level schemes can be seen in Figures 7–9 of [224].

9.1.2. Kr (Z = 36) Isotopes above N = 50

Experimental evidence for a smooth onset of deformation has been seen in 94 96 Kr 58 60 [709], with the low-Z boundary of the island of deformation seen at 96 Kr 60 [710]. Oblate-prolate SC has been in 98 Kr 62 [711]. A relevant review has been given in [689].
On the theoretical side, extensive RMF calculations have been performed [242] in the 86 104 Kr 50 68 region, using a 5DQCH with parameters determined from the PC-PK1 energy density functional. In addition, SC in 94 Kr 58 has been considered in the complex excited VAMPIR approach [194].
As mentioned above, extensive mean-field calculations covering the whole 70 98 Kr 34 62 region have been performed [139] with the Gogny D1S interaction, showing SC in 94 98 Kr 58 62 (see Figure 9 of [139] for relevant level schemes).
Again, as mentioned above, extensive calculations covering the whole 70 100 Kr 34 64 region have been performed [150] in the relativistic HF framework with the DD-PC1 parametrization, leading to SC in 92 , 96 100 Kr 56 , 60 64 (see Table I of [150]), as well as with the DD-ME2 parametrization, leading to SC in 92 100 Kr 56 64 (see Table II of [150]). Once more, there is some dependence on the details of the interactions, but they both agree that SC is observed roughly in the region N = 56 –64.
Again, as mentioned above, extensive calculations covering the whole 70 100 Kr 34 64 region have also been performed [224] using an IBM Hamiltonian with parameters determined from a Gogny energy density functional. SC is seen in 92 100 Kr 56 64 (see Table I of [224]). Relevant level schemes can be seen in Figures 8, 9 of [224].
Experimental evidence for SC has also been seen in 88 Kr 52 [712], as well as in 86 Se 52 and 84 Ge 52 . These observations suggest that the N = 50 island of inversion [534] is extended beyond 78 Ni 50 .
In conclusion, SC is well established both experimentally and theoretically in the region 70 76 Kr 34 40 , while in addition SC related to the N = 50 island of inversion is seen in 88 Kr 52 and both experimental and theoretical signs for the beginning of a new region of SC starting at N = 60 have been seen in 94 100 Kr 58 64 .

9.2. The Se (Z = 34) Isotopes

9.2.1. Se (Z = 34) Isotopes below N = 50

Coexistence of a nearly spherical ground state band with a prolate deformed band built on the 0 2 + state (at 937 keV) has been first observed in 1974 by Hamilton et al. [713,714] in 72 Se 38 . Later experiments pointed to a slightly oblate shape for the ground state [715,716], while negative parity bands have also been observed [717]. An up-to-date level scheme can be seen in Figure 1 of [718]. A very similar situation, with a prolate deformed band built on the 0 2 + state (at 854 keV) and a near-spherical ground state has been seen also in 74 Se 40 [719,720], while in 68 Se 34 an excited prolate deformed band built on 0 2 + and a slightly oblate ground state band have been seen [715,721,722]. A similar picture is seen in 70 Se 36 [706,715]. A short review of the experimental situation in 68 72 Se 34 38 has been given in [689]. An oblate ground state has also been seen in 66 Se 32 [723].
The 0 2 + states in the region 68 82 Se 34 48 , plotted vs. N, form a parabola with its bottom located at N = 38 , 40 (see Figure 17 of [724] and Figure 11c), while the 0 2 + states of the N = 40 isotones in the Z = 28 –36 region also appear to form a parabola with its bottom lying at Z = 32 –36 (see Figure 18 of [724] and Figure 12b). These curves support the occurrence of SC in the region around Z = 34 , N = 40 .
Theoretical calculations have been focused on the 68 74 Se 34 40 region. Shell-model calculations have shown the oblate shape of the ground state and the prolate–oblate SC in 68 Se 34 [61], as well as the importance of 4p-4h excitations from ( 2 p 3 / 2 1 f 5 / 2 ) to ( 2 p 1 / 2 1 g 9 / 2 ) for the oblate configurations. The sudden structural change occurring along the N = Z line when passing from N = Z 35 to N = Z 36 is attributed to a sudden jump of nucleons into the 1 g 9 / 2 2 d 5 / 2 orbitals [60]. A systematic study of 68 74 Se 34 40 has been performed in the shell-model framework using the 2 p 1 f 5 / 2 1 g 9 / 2 model space [64], taking into account a Gaussian central force and the tensor force. A simple two-state mixing model [90] has shown that the excited 0 2 + band is significantly more collective than the ground state band in 76 Se 42 .
SC in 68 72 Se 34 38 has been accounted for in the excited VAMPIR [182,183] and the complex excited VAMPIR [189,191,193,195] mean-field approaches. The oblate–prolate shape mixing in these nuclei has also been described within the adiabatic self-consistent collective coordinate (ASCC) method [199]. Spectra and B(E2)s have been calculated using a 5DQCH (see Figures 10, 12, and 14 of [200] for the relevant level schemes and [134] for a relevant review).
Spectra and B(E2)s in the extensive region 68 96 Se 34 62 have been calculated [223] using an IBM Hamiltonian with parameters determined from the Gogny-D1M energy density functional. SC has been seen in 72 , 74 Se 38 , 40 , and also in the neutron-rich 94 Se 60 , which will be discussed in Section 9.2.2 (see Figures 15, 17 and 18 of [223] for the relevant level schemes).
Finally, SC in 72 76 Se 38 42 has been described [283] using a Bohr Hamiltonian with a sextic potential possessing two unequal, competing minima (see Figure 3 of [283] for the potentials and Figure 1 of [283] for the relevant level schemes).

9.2.2. Se (Z = 34) Isotopes above N = 50

Experimental signs of SC have been seen in 86 Se 52 [712], while in 88 Se 54 excited states have been detected [725], indicating increased collectivity in relation to 86 Se 52 . A transition from γ -soft to rigid oblate ground state is seen between 92 Se 58 and 94 Se 60 [726]. Low-lying structures and shape evolution in the region 88 94 Se 54 60 have been considered in [727].
As mentioned above, spectra and B(E2)s in the extensive region 68 96 Se 34 62 have been calculated [223] using an IBM Hamiltonian with parameters determined from the Gogny-D1M energy density functional, with SC seen in 94 Se 60 (see Figure 18 of [223] for the relevant level scheme).
In conclusion, SC is well established both experimentally and theoretically in the region 68 74 Se 34 40 , while signs of SC have been seen in 86 Se 52 , related to the N = 50 island of inversion, and in 94 Se 60 , signing the beginning of a new region of SC starting at N = 60 .

9.3. The Ge (Z = 32) Isotopes

9.3.1. Ge (Z = 32) Isotopes below N = 50

Experimental evidence for SC in 68 74 Ge 36 42 has been reported in [728], in which a transition from spherical to deformed shapes is also seen between 72 Ge 40 and 74 Ge 42 . Further evidence for SC in 70 Ge 38 and 72 Ge 40 has been reported in [626] and [729] respectively. Experimental evidence is reviewed in [724], in Figure 10 of which the 0 2 + states of 68 78 Ge 36 46 are seen to exhibit a parabolic behavior with a sharp minimum at N = 40 (see also Figure 11b).
It should be noticed (see Figure 11b) that 72 Ge 40 represents a rare example in which the 0 2 + band-head of a well-developed K π = 0 + band lies below the 2 1 + state of the ground state band [335]. It has been found [335] that such nuclei tend to be surrounded by nuclei exhibiting SC.
Theoretical corroboration for SC in 68 Ge 36 and 72 Ge 40 has been given by mean-field calculations in the excited VAMPIR approach [182,183,184]. Earlier shell-model calculations [63,70], corroborating SC in 72 Ge 40 , have been recently extended [64] to the whole 64 76 Ge 32 44 region, emphasizing the role played by the tensor force and supporting SC in 68 76 Ge 36 44 , as seen in Figures 15–19 of [64]. SC of spherical and oblate shapes in 72 Ge 40 has also been found in extended IBM calculations with parameters determined through CDFT using the Gogny-D1M energy density functional [223]. IBM calculations in 64 80 Ge 32 48 have also been carried out and compared to solutions of a γ -unstable Bohr Hamiltonian with a Davidson potential, with both approaches giving satisfactory results [730]. SC in 74 Ge 42 has also been studied [279] by a Bohr Hamiltonian with a sextic potential in the β variable, exhibiting two unequal minima at different deformations, as seen in Figure 6 of [279]. A similar Bohr calculation [280] in 80 Ge 48 has shown absence of SC in this isotope.
Shell-model calculations [61] for N = Z nuclei in the Z = 30–36 region have found coexistence of prolate and oblate shapes in 68 Se 34 and 72 Kr 36 , but not in 60 Zn 30 and 64 Ge 32 .

9.3.2. Ge (Z = 32) Isot4opes above N = 50

Experimental signs of SC in 82 Ge 50 and 84 Ge 52 have been given in [731] and [712], respectively. Signs of SC in 80 Ge 48 given in [732,733] have been questioned by [734].
Gogny HFB, as well as Skyrme HF plus BCS calculations, have been carried out in 58 108 Ge 26 76 [133], along with IBM calculations in 66 94 Ge 34 62 with parameters determined through CDFT using the Gogny-D1M energy density functional [223]. In the latter calculation, nuclei exhibiting γ -soft shapes and coexistence of prolate and oblate shapes are seen in 52 N 62 [224].
In summary, SC is experimentally and theoretically established in 68 76 Ge 36 44 , while evidence for SC is also seen in 80 94 Ge 48 62 .

9.4. The Zn (Z = 30) Isotopes

Experimental evidence for SC in 68 , 70 Zn 38 , 40 has been seen in [724,735,736], backed by Nilsson–Strutinsky [735] and large-scale shell-model [736] calculations. The 0 2 + states in 66 72 Zn 36 42 form a parabola with a minimum at N = 40 (see Figure 1 of [735], Figure 4 of [724], Figure 1 of [736], as well as Figure 11a).
Evidence for the robustness of the Z = 50 gap and the possibility of inversion of proton orbitals has been seen in 78 , 80 Zn 48 , 50 [737,738], supported by evidence for SC in 79 Zn 49 [739,740]. Large-scale shell-model [738,739,740] and MCSM [737] calculations support these findings.
Shell-model calculations [61] for N = Z nuclei in the Z = 30 –36 region have found coexistence of prolate and oblate shapes in 68 Se 34 and 72 Kr 36 , but not in 60 Zn 30 and 64 Ge 32 .
In summary, SC is established in 68 , 70 Zn 38 , 40 , related to the N = 40 island of inversion, while evidence for SC is also seen in 78 80 Zn 48 50 , related to the N = 50 IoI.

9.5. Shape Coexistence and Shape/Phase Transition at N 40

In Figure 12 the ground state bands and the intruder levels are plotted for N = 38 , 40, 42 in the Z 34 region.
In Figure 12a we see that for N = 38 , the intruder levels seem to form the left branch of a parabola, reaching a minimum at Z = 36 , i.e., near the proton midshell ( Z = 34 ) between the major shell closure at N = 28 and the subshell closure at Z = 40 . The same happens in Figure 12b for N = 40 and in Figure 12c for N = 42 . The difference among the three cases is that at N = 38 the intruder 0 2 + state starts between the 2 1 + and 4 1 + states at N = 28 and gradually approaches 2 1 + , with which it becomes practically degenerate for Z = 32 –36, while in the N = 40 case this degeneracy occurs in the whole Z = 28 –34 region. In contrast, for N = 42 , the 0 2 + state becomes practically degenerate with the 4 1 + state for Z = 30 –38.
The evolution from lower N to higher N seen in Figure 12 for the N 40 region strongly resembles the evolution seen in Figure 8 and Figure 10 for the N 60 and N 90 regions, respectively, the main difference being that approximate degeneracy of the 0 2 + state is seen in N = 90 with the 6 1 + state, in N = 60 with the the 4 1 + state, and in N = 40 with the the 2 1 + state.
Several comments are in place.
In Section 5.5 and Section 8.5 we have seen that the parabolas appearing at N 90 and N 60 have been attributed to neutron particle–hole excitations caused by the influence of the protons, leading to proton-induced SC. In addition, in Section 3 and Section 7 we have seen that the parabolas appearing at Z 82 and Z 50 have been attributed to proton particle–hole excitations caused by the influence of the neutrons, leading to neutron-induced SC. It has been suggested [579,580] that in the region Z = 38 –40, N = 38 –40, both mechanisms are simultaneously active. This interpretation is consistent with the role played in the shell-model framework by the tensor force [102,105,106,107], which is known to be responsible for the influence of the protons on the neutron gaps, as well as the influence of the neutrons on the proton gaps, as described in Section 2.1.1.
The fact that in practically all cases for N = 38 –42 and Z = 28 –38 the intruder 0 2 + levels remain below the 4 1 + members of the gsb, or are almost degenerate to them, suggests that the appearance of SC in all these nuclei is plausible. In different words, 56 66 Ni 28 38 , 58 68 Zn 28 38 , 60 70 Ge 28 38 , 62 72 Se 28 38 , 64 74 Kr 28 38 , and 66 76 Sr 28 38 , are good candidates for the appearance of SC.
The strong similarity between the Z 34 , N 40 region and the Z 40 , N 60 and Z 60 , N 90 regions has a further consequence. The N = 90 isotones 150 Nd 90 , 152 Sm 90 , 154 Gd 90 , and 156 Dy 90 are known to be very good examples of the X(5) CPS, occurring at the critical point of a first-order SPT from spherical to prolate deformed shapes. Because of the similarities mentioned above, it should be expected that the N = 40 isotones could also be good candidates for the X(5) CPS, thus providing further connection between the SC and SPT concepts, as proposed by Heyde et al. [438,439,445,469,683]. However, it should be pointed out that the numerical hallmarks of this critical point symmetry might be different from those of X(5), since in the Z 34 , N 40 region the 0 2 + states appears to be nearly degenerate with the 2 1 + states and not nearly degenerate to the 6 1 + states. In other words, the N = 40 isotones 68 Ni 40 , 70 Zn 40 , 72 Ge 40 , 74 Se 40 , and 76 Kr 40 may be good candidates for the first order SPT, but some formulation of the CPS more flexible than X(5) would be required in the Z 34 , N 40 region.

10. The Light Nuclei at and below Z28

10.1. The Ni (Z = 28) Isotopes

SC in this region has first been observed in 68 Ni 40 [741,742,743,744,745,746,747] and 70 Ni 42 [742,748,749], attributed to proton 2p-2h excitations across the Z = 28 shell. The early implications [750] that configuration mixing is needed in order to interpret the spectra of 64 , 66 Ni 36 , 38 was followed by the observation of SC in these two isotopes, see [751,752] for 66 Ni 38 and [532,753] for 64 Ni 36 . Existing evidence for SC in these isotopes is reviewed in [754], in which it is seen (in Figure 7 of [754]) that the 0 2 + states in 64 70 Ni 36 42 resemble the left branch of a parabola reaching a minimum/plateau at N = 40 , 42 (see Figure 13c).
The above observations indicate the occurrence of SC in the neighborhood of Z = 28 near the neutron 28–50 midshell ( N = 39 ), in direct analogy to what has been seen in the neighborhood of Z = 50 near the neutron 50–82 mideshell ( N = 66 ), as well as in the neighborhood of Z = 82 near the neutron 82–126 midshell ( N = 104 ).
Recent experiments [755,756] suggest that SC might also appear in the doubly magic nucleus 78 Ni 50 , signifying a new island of inversion for Z 28 .
68 Ni 40 has been considered as a test ground for examining the roles played by the various components of the nucleon–nucleon interaction in the shell-model framework in the appearance of SC. The balance of the tensor and the central force has been considered in the MCSM framework [105], while the competition between the spherical mean field and the pairing and quadrupole–quadrupole interactions has been studied in large-scale shell-model calculations taking advantage of the pseudo-SU(3) and quasi-SU(3) symmetries [46]. Calculations within these two frameworks extending to several Ni isotopes can be found in [58,73] and [47], respectively.
SC in 78 Ni 50 has been shown to occur within large-scale shell-model calculations taking advantage of the pseudo-SU(3) and quasi-SU(3) symmetries [534], signifying a fifth island of inversion at N = 50 , in addition to the ones known at N = 8 , 20, 28, 40. Earlier considerations on the evolution of the N = 50 gap can be found in [757].
SC in the doubly magic 56 Ni 28 has been studied in the MCSM framework [56,58,67,73], as well as with a pairing plus multipole Hamiltonian with a monopole interaction [63] and within the complex excited VAMPIR mean-field approach [187].
Extended MCSM calculations incorporating SC at the doubly magic nuclei 56 , 68 , 78 Ni 28 , 40 , 50 can be found in [58,73].
In summary, much experimental and theoretical evidence for SC is accumulated in the region 64 70 Ni 36 42 , i.e., around N = 40 , while some experimental and theoretical evidence for SC exists in 78 Ni 50 , signifying an island of inversion at N = 50 . Theoretical predictions for SC in the N = Z nucleus 56 Ni 28 also exist.

10.2. The Fe (Z = 26) Isotopes

Experimental evidence for particle–hole excitations across the N = 40 subshell closure and enhanced collectivity, related to the N = 40 island of inversion, have been observed in 62 Fe 36 [548,758,759,760], 64 Fe 38 [548,758,759,760], 66 Fe 40 [546,759,760,761,762], 68 Fe 42 [546,762]. Strong experimental evidence for particle–hole cross-shell excitations has been seen [763] in the N = Z isotope 52 Fe 26 .
Large-scale shell-model [47,50] as well as two-band mixing [764] calculations in the 60 66 Fe 34 40 region have indicated the existence of strongly mixed spherical and deformed configurations, as well as the occupation of neutron intruder orbitals as N = 40 is approached, related to the N = 40 island of inversion. Shell-model calculations show evidence for particle–hole excitations and SC in 54 Fe 28 [68] and 76 Fe 50 [534], in relation to the N = 28 and N = 50 islands of inversion, while the excited 0 + states in the N = Z isotope 52 Fe 26 have been considered in [62].
In summary, experimental and theoretical evidence for SC is accumulated for 62 66 Fe 36 40 , especially for 66 Fe 40 , while theoretical evidence also exists for 54 Fe 28 and 76 Fe 50 , in relation to the N = 28 and N = 50 islands of inversion, as well as for the N = Z isotope 52 Fe 26 .

10.3. The Cr (Z = 24) Isotopes

Proton inelastic scattering experiments in inverse kinematics on 58 Ti 36 , 60 Cr 36 and 62 Cr 38 , backed by shell-model calculations, have demonstrated [384] that the contributions of both protons and neutrons are crucial for the building up of collectivity through the admixture of the p f and d g shells across the N = 40 gap. 58 Cr 34 has been suggested as a candidate for the E(5) CPS in several works [384,385,386,387]. SC in 64 , 66 Mn 39 , 41 has been found through the population of levels following the β decay of 64 Cr 40 and 66 Cr 42 [765]. These findings cover the whole region N = 34 –40, in which SC is expected to occur according to the dual-shell mechanism of the proxy-SU(3) symmetry [334]. Furthermore, the appearance of the CPS E(5) has been suggested to occur at the left border of this region, while the N = 40 island of inversion appears at its right border [536,546].
Furthermore, experimental evidence for SC has been seen in 52 Cr 28 [766], related to the N = 28 island of inversion, as well as in the N = Z isotope 48 Cr 24 [763].
Shell-model calculations focused on the IoI around 64 Cr 40 have been performed in [46,50]. Extended large-scale shell-model [47] and MCSM [58] calculations covering the whole 48 64 Cr 24 40 region, as well as calculations using a 5DQCH [204,205] covering the 58 68 Cr 34 44 region, corroborate the evolution of shape changes and the occurrence of shape mixing in the 58 64 Cr 34 40 region, in agreement with a two-band mixing calculation [764] in this region.
The relation between SC and isobaric analog states has been considered in 44 Cr 20 within an algebraic framework in [767]. The N = Z isotope 48 Cr 24 has been considered by shell-model calculations in [64], while 74 Cr 50 has been considered by large-scale shell-model calculations in relation to the N = 50 IoI in [534].
In summary, much experimental and theoretical evidence for SC is accumulated for 58 64 Cr 34 40 , while experimental and/or theoretical evidence also exists for 44 Cr 20 , 52 Cr 28 ,and 74 Cr 50 , in relation to the N = 20 , 28, 50 islands of inversion, as well as for the N = Z isotope 48 Cr 24 .

10.4. The Ti (Z = 22) Isotopes

Experimental evidence [763,768,769,770] for SC and particle–hole excitations has been found in 44 Ti 22 , corroborated by d f shell-model [768] and p f shell-model [763,771] calculations, while theoretical considerations [767] in an algebraic framework have revealed the relation between isobaric analog states and SC.
The lowest four 0 + states of 50 Ti 28 have been considered [197] within the excited VAMPIR framework, with considerable mixing among them found. Shell-model calculations [50,546] have considered 60 , 62 Ti 38 , 40 as members of the N = 40 island of inversion, while experiments on 58 Ti 36 , 60 Cr 36 and 62 Cr 38 have demonstrated [384] that the contributions of both protons and neutrons are crucial for the building up of collectivity through neutron excitations across the N = 40 gap. Large-scale shell-model calculations [534] have been performed for 72 Ti 50 in relation to the N = 50 island of inversion.
In summary, much experimental and theoretical evidence for SC is accumulated for 44 Ti 22 , i.e., at N = Z , while theoretical studies also exist for 50 Ti 28 , 60 , 62 Ti 38 , 40 , and 72 Ti 50 , in relation to the N = 28 , 40, 50 islands of inversion.

10.5. The Ca (Z = 20) Isotopes

SC in 40 Ca 20 [772] and 42 Ca 22 [773] have been known experimentally since 1972, and have been interpreted as multi-paricle–multi-hole excitations in the framework of large-scale shell-model calculations [46,48]. Recent experimental evidence [769,774,775] corroborated SC in 42 Ca 22 and extended it to 44 Ca 24 , the latter having been theoretically considered [767] in an algebraic framework revealing the relation between isobaric analog states and SC.
It should be noticed that 40 Ca 20 represents a rare example in which the 0 2 + band-head of a well-developed K π = 0 + band lies below the 2 1 + state of the ground state band [335] (see Figure 13b). It has been found [335] that such nuclei tend to be surrounded by nuclei exhibiting SC. Indeed the intruder states in Figure 13b seem to form the bottom of a parabola around the N = 24 midshell.
Theoretical studies in relation to SC and islands of inversion have been performed for 48 Ca 28 using a 5DQCH with parameters determined from the relativistic energy density funcional DD-PC1 [240], as well as within TDHFB theory [180]. Large-scale shell-model calculations for 36 Ca 16 [51], 60 Ca 40 [50] and 70 Ca 50 , related to the IoIs at N = 20 , N = 40 and N = 50 , have also been carried out.
In summary, much experimental and theoretical evidence for SC is accumulated for 36 , 40 44 Ca 16 , 20 24 , i.e., around N = 20 , while theoretical studies also exist for 48 Ca 28 , 60 Ca 40 , and 70 Ca 50 , in relation to the N = 20 , 28, 40, 50 islands of inversion.

10.6. The Ar (Z = 18) Isotopes

Experimental evidence [776,777,778,779,780] for SC and particle–hole excitations has been found in 36 40 Ar 18 22 , corroborated by shell-model [778,780], cranked Nilsson–Strutinsky [780], and cranked HFB [777] calculations. SC in 36 Ar 18 has also been considered [169] within RMF theory, in conjunction with the hypernucleus Λ 37 Ar.
Experimental evidence for the existence of a strong N = 28 shell gap at 46 Ar 28 has been provided by mass measurements [781]. The importance of the quadrupole interaction between protons and neutrons in coupling the proton hole states with neutron excitations across the N = 28 shell gap in 46 Ar 28 has been demonstrated through TDHFB calculations [180]. In addition, 46 Ar 28 has been considered within a 5DQCH with parameters determined through the relativistic energy density functional DD-PC1 [233,240].
In summary, much experimental and theoretical evidence for SC is accumulated for 36 40 Ar 18 22 , i.e., around N = 20 , while experimental evidence and theoretical studies also exist for 46 Ar 28 , in relation to the N = 28 island of inversion.

10.7. The S (Z = 16) Isotopes

Experimental evidence for SC has been seen in 44 S 28 [754,782,783] and 43 S 27 [784,785], i.e., close to the N = 28 shell closure.
SC in 44 S 28 has been seen in theoretical calculations within self-consistent mean-field theory with Skyrme interaction [147], TDHFB theory [180], shell model with variation after angular-momentum projection [75], as well as within a 5DQCH with parameters determined from the relativistic energy density functional DD-PC1 [233,240] (see Figure 1 of [75], Figure 7 of [240], and Figure 3 of [233] for relevant level schemes).
Calculations covering the 36 44 S 20 28 region have been performed in the self-consistent HF framework with Skyrme interaction [69], as well as within various shell-model approaches [47,69,74]. A deformed ground state is predicted [69] for 44 S 28 , with strong mixing of the closed-shell and excited configurations [47]. The crucial role played by the tensor force in creating SC in 44 S 28 is clearly seen in Figure 5 of [74], where it is made clear that the tensor force is crucial for developing a second minimum in the potential energy surface.
Shell-model calculations with configuration interaction [51] point out the different structures of the intruder and normal states in 36 S 20 , thus giving a signal related to the N = 20 island of inversion.
In summary, much experimental and theoretical evidence for SC is accumulated for 44 S 28 , i.e., at the N = 28 island of inversion, while some theoretical evidence also exists for 36 S 20 , in relation to the N = 20 island of inversion.

10.8. The Si (Z = 14) Isotopes

SC has been observed in 34 Si 20 [754,786,787] and 32 Si 18 [788]. Large-scale shell-model calculations have also been performed for these nuclei [46,47,49].
SC in 42 Si 28 has been considered in shell-model calculations [49,74], as well as in a 5DQCH with parameters determined from RMF calculations using the DD-PC1 energy density functional [233,240] and in a TDHFB framework [180].
SC in the N = Z nucleus 28 Si 14 has been considered by the Nilsson–Strutinsky approach [789,790], as well as by constrained HFB plus local quasiparticle RPA calculations [201] and large-scale shell-model calculations [49].
In summary, much experimental and theoretical evidence for SC exists for 32 , 34 Si 18 , 20 , i.e., around N = 20 , while theoretical evidence for SC also exists both for 42 Si 28 and 28 Si 14 , i.e., at the N = 28 island of inversion and along the N = Z line.

10.9. The Mg (Z = 12) Isotopes

SC has been observed in 30 Mg 18 [754,791,792,793] and 32 Mg 20 [543,754], as well as in 40 Mg 28 [754]. SC is also seen in the N = Z nucleus 24 Mg 12 [794]. Intruder states shown in Figure 13a seem to form the left branch of a parabola reaching a minimum at N = 20 .
Mass systematics [533] have suggested 32 , 34 Mg 20 , 22 as candidates for the island of inversion at N = 20 (see also [542]). SC in 30 34 Mg 18 22 has been considered in self-consistent HF calculations with a Skyrme interaction [69], as well as in large-scale shell-model calculations [46,49], and within a 5DQCH [198,202].
SC in the N = 28 isotones with Z = 12 –20, including 40 Mg 28 , has been studied in a 5DQCH with parameters determined from RMF calculations using the DD-PC1 energy density functional [233,240]. A bridge between the N = 20 and N = 28 islands of inversion has been provided for the Mg isotopes by large-scale shell-model calculations [49], supported by recent experimental evidence [545].
SC in the N = Z nucleus 24 Mg 12 has been considered within the extended IBM formalism [427,428,429], as well as within shell-model [62] and constrained HFB plus local quasiparticle RPA calculations [201]. In contrast, hindrance of shape mixing in 26 Mg 14 has been found in the last calculation [201], based on the softness of the collective potential against the β and γ deformations.
In summary, much experimental and theoretical evidence for SC is accumulated for 30 , 32 Mg 18 , 20 , i.e., around N = 20 , while experimental and theoretical evidence for SC also exists both for 40 Mg 28 and 24 Mg 12 , i.e., at the N = 28 island of inversion and along the N = Z line.

10.10. The Ne (Z = 10) Isotopes

Mass systematics [533] have suggested 30 , 32 Ne 20 , 22 as candidates for the island of inversion at N = 20 (see also [542]). Further experimental evidence supporting that 32 Ne 22 does belong to the N = 20 IoI has been provided in [795], while experimental evidence on 28 Ne 18 [796] suggests that this nucleus does not belong to the N = 20 IoI. Self-consistent HF calculations with a Skyrme interaction [69] suggested that shape mixing occurs in 30 Ne 20 , but to a much smaller degree in comparison to 32 Mg 20 (compare Figures 2 and 3 of [69]). Large-scale shell-model calculations [49] corroborated that 30 , 32 Ne 20 , 22 lie in the N = 20 island of inversion, and suggested that the Ne isotopes might also provide a bridge between the N = 20 IoI and the N = 28 IoI, if the neutron drip line does reach 38 Ne 28 .
In contrast, hindrance of shape mixing in 24 Ne 14 has been found in constrained HFB plus local quasiparticle RPA calculations [201], based on the softness of the collective potential against the β and γ deformations.
20 Ne 10 has been known [473] to be a textbook example of a deformed nucleus in the very light mass region, having a high-lying 0 2 + state [62]. Recent calculations within the symmetry-adapted no-core shell-model have revealed [346] that 20 Ne 10 is made out of a few equilibrium shapes in the framework of the symplectic symmetry.
In summary, both experimental and theoretical evidence for SC is concentrated on 30 , 32 Ne 20 , 22 , lying within the N = 20 IoI.

10.11. The O (Z = 8) Isotopes

16 O 8 is the nucleus in which SC was first suggested by Morinaga in 1956 [1]. Soon thereafter, the two-particle–two-hole excitation mechanism was discussed in the shell-model framework for 16 O 8 [797,798,799] and 18 O 10 [798].
Proton ( π ) and neutron ( ν ) particle–hole (p-h) excitations in 16 O 8 have been considered in the extended IBM-2 model within the Wigner SU(4) symmetry [800,801,802,803], further discussed in Section 11. In particular, the π (2p-2h) ν (2p-2h), π (4p-4h), and ν (4p-4h) configurations of 16 O 8 have been placed in the same multiplet with the π (4p) ν (4p) configuration of 24 Mg 12 [427,428,429].
The emergence of deformation and SC in 16 O 8 has been discussed by Rowe et al. in the framework of the symplectic model [619,620,804].
The spectrum of 20 O 12 has been used, in comparison to the spectrum of 20 Ne 10 , in order to demonstrate the importance of the proton–neutron interaction for the development of nuclear deformation in the latter [473].
Recent experimental evidence [539] suggests 28 O 20 as the low-Z shore of the N = 20 IoI.
In summary, SC appears in 16 O 8 and 28 O 20 .

10.12. The C (Z = 6) Isotopes

The 0 2 + state of 12 C 6 at 7.65 MeV is of special interest for stellar nucleosynthesis, since it plays a key role in allowing it to proceed beyond 12 C 6 to heavier elements. First suggested by the astrophysicist Hoyle in 1954 [805], it was considered from the SC point of view in 1956 by Morinaga, his work being the first study of SC in atomic nuclei [1]. An important step towards establishing a band based on the Hoyle state was the observation of a 2 + state around 10 MeV [806,807]. The structure of 12 C 6 has also been discussed by Rowe et al. [619,620] in the framework of the symplectic model. Extended work has been done over the years on the structure of the Hoyle state and its description as an alpha-particle condensate, recently reviewed in [808,809], respectively, while a relevant review on microscopic clustering in light nuclei has been given in [810]. The Hoyle state is still an active field of research (see, for example, [811] and references therein).

10.13. The Be (Z = 4) Isotopes

The breakdown of the N = 8 shell closure in 12 Be 8 has been seen experimentally in 2000 in [812,813], see [814] for recent references. This leads to the lowest island of inversion at N = 8 [536].

11. The N = Z Nuclei

Summarizing the evidence for SC in N = Z nuclei cited in the subsections of Section 3, Section 4, Section 5, Section 6, Section 7, Section 8, Section 9 and Section 10 on individual series of isotopes and mentioned in their conclusions, the following picture emerges.
SC is seen in the isolated nuclei 6 12 C 6 , 8 16 O 8 , 12 24 Mg 12 , 14 28 Si 14 , 22 44 Ti 22 , 24 48 Cr 24 , 26 52 Fe 26 , 28 56 Ni 28 .
In contrast, SC in N = Z nuclei surrounded by more isotopes exhibiting SC is seen in the regions 18 36 40 Ar 18 22 , 20 40 44 Ca 20 24 , 34 68 74 Se 34 40 , 36 70 76 Kr 34 40 , 38 74 82 Sr 36 44 , 40 76 84 Zr 36 44 .
The difference between these two groups of N = Z nuclei can be explained through the dual-shell mechanism [334,335] developed in the framework of the proxy-SU(3) symmetry [323,327]. The Z number of the nuclei belonging to the second group lies within the regions 7–8, 17–20, 34–40, 59–70, 96–112 …, in which SC can appear according to the dual-shell mechanism; thus, the N = Z nucleus can be accompanied by additional isotopes with N lying within these regions. For example, 34 68 Se 34 can be accompanied by its isotopes lying within the N = 34 –40 region, i.e., the whole group 34 68 74 Se 34 40 can exhibit SC.
N = Z nuclei are the ideal test ground for the manifestation of Wigner’s spin-isospin symmetry, introduced in 1937 [803] and called the Wigner SU(4) supermultiplet model [800,801,802]. Wigner suggested that, at first approximation, the nuclear forces should be independent of the orientation of the spins and the isospins of the nuclei constituting the nucleus. The double binding energy differences of Equation (3) have provided a test [815] of the Wigner SU(4) supermultiplet, while further details on its implications for heavy N = Z nuclei can be found in chapter 6 of [816].

11.1. T = 0 Pairs in N = Z Nuclei

The fact that neutrons and protons in N = Z nuclei occupy the same orbitals gives room to the development of unique proton–neutron correlations which cannot be seen away from the N = Z line. Much attention has been attracted by the experimental observation in 2011 [817] of an isoscalar ( T = 0 ) spin-aligned ( S = 1 ) neutron-proton phase in the spectrum of 46 92 Pd 46 . This phase is expected to replace the standard seniority [37,258,259,260,261] coupling in the ground and low-lying excited states of the heaviest N = Z nuclei. The existence of this phase in 46 92 Pd 46 and 48 96 Cd 48 has been corroborated by shell-model calculations [818,819,820,821,822], within a model using quartets [823,824], as well as by mapping the shell-model wave functions onto bosons [825], see also [816,826] for relevant reviews. This phase has been recently seen also in 44 88 Ru 44 [827].
The quartet formalism has also been successfully applied in the heavier N = Z nuclei 52 104 Te 52 , 54 108 Xe 54 , and 56 112 Ba 56 [828,829,830], as well as in the light N = Z nuclei 10 20 Ne 10 , 12 24 Mg 12 , 14 28 Si 14 , and 16 32 S 16 [823,828,829,830]. Shape isomers and clusterization, taking advantage of a quasidynamical SU(3) symmetry, have been studied in 14 28 Si 14 [831] and 18 36 Ar 18 [832,833].
The quartet formalism has also been successfully applied in the N = Z nuclei 22 44 Ti 22 , 24 48 Cr 24 , and 26 52 Fe 26 [828,829,830]. These nuclei have also been considered within cranked mean-field [834,835] and shell-model [821,836,837] calculations.
Experimental investigations in the N = Z nuclei 28 56 Ni 28 , 30 60 Zn 30 , and 32 64 Ge 32 [838,839,840,841] have been followed by mean-field [842,843,844] and shell-model [845] calculations focusing on the T = 0 neutron–proton pairing correlations and their competition with the T = 1 interactions, as well as by clusterization [846] investigations. The importance of the T = 0 channel has been questioned by [847]. A review of these efforts can be found in [848].
Experimental investigations at higher masses have been carried out in 34 68 Se 34 [849], 36 72 Kr 36 [838], and 38 76 Sr 38 [850]. An overview of the N = Z = 32 –46 region has been given in Section 11.3 of [816], while for the N = Z = 44 –48 region one can see [826].
In summary, while SC in the N = Z nuclei has been seen up to N = Z = 42 , theoretical investigations are extended up to N = Z = 56 , recently motivated by the observation of the isoscalar ( T = 0 ) spin-aligned ( S = 1 ) neutron–proton phase in the region of 46 92 Pd 46 . No SC is expected in the N = Z = 44 –56 region according to the dual-shell mechanism [334,335] developed in the framework of the proxy-SU(3) symmetry [323,327], since these nuclei lie outside the 7–8, 17–20, 34–40, 59–70, 96–112 …regions, in which SC can appear.

12. Islands of Inversion

Summarizing the evidence for IoIs cited in the subsections of Section 3, Section 4, Section 5, Section 6, Section 7, Section 8, Section 9 and Section 10 on individual series of isotopes and mentioned in their conclusions, the following picture emerges.
Signs for the IoI at N = 8 appear at Z = 4 , in 4 12 Be 8 .
Evidence for the IoI at N = 20 appears at Z = 8 –20, 24, namely in 8 28 O 20 , 10 30 , 32 Ne 20 , 22 , 12 30 , 32 Mg 18 , 20 , 14 32 , 34 Ne 18 , 20 , 16 36 S 20 , 18 36 40 Ar 18 22 , 20 40 44 Ca 20 24 , and 24 44 Cr 20 . In other words, a robust N = 20 IoI appears at Z = 8 -20, with an additional point at Z = 24 . In some cases ( Z = 12 , 14 , 18 ), SC is also seen at N = 18 , while in other cases ( Z = 10 , 18 , 20 ) it is also seen at N = 22 .
Evidence for the IoI at N = 28 appears at Z = 12 –28, namely in 12 40 Mg 28 , 14 42 Si 28 , 16 44 S 28 , 18 46 Ar 28 , 20 48 Ca 28 , 22 50 Ti 28 , 24 52 Cr 28 , 26 54 Fe 28 , and 28 56 Ni 28 . In other words, a robust N = 28 IoI appears at Z = 12 –28, with no additional cases seen at N = 26 or N = 30 .
Evidence for the IoI at N = 40 appears at Z = 20 –40, namely in 20 60 Ca 40 , 22 60 , 62 Ti 38 , 40 , 24 58 64 Cr 34 40 , 26 62 66 Fe 36 40 , 28 64 70 Ni 36 42 , 30 68 70 Zn 38 , 40 , 32 68 76 Ge 36 44 , 34 68 72 Se 34 40 , 36 70 76 Kr 34 40 , 38 74 82 Sr 36 44 , and 40 76 84 Zr 36 44 . In other words, a robust N = 40 IoI appears at Z = 20 –40, which in most cases ( Z = 22 –40) is extended to lower N as well.
Evidence for the IoI at N = 50 appears at Z = 20 –32, namely in 20 70 Ca 50 , 22 72 Ti 50 , 24 74 Cr 50 , 26 76 Fe 50 , 28 78 Ni 50 , 30 78 , 80 Zn 48 , 50 , 32 80 94 Ge 48 62 . In addition, signs of SC appear further up in Z = 34 , 36 in the N = 52 isotones 34 86 Se 52 , and 36 88 Kr 52 . In other words, a robust N = 50 IoI appears at Z = 30 -32, with no additional cases seen up to Z = 28 , while above Z = 28 SC starts to be seen in N = 48 or N = 52 up to Z = 36 .
It is seen that the N = 28 and N = 50 IoIs remain “narrow”, while the N = 20 and N = 40 IoIs get “wide”. A way to understand this difference is again offered by the dual-shell mechanism [334,335] developed in the framework of the proxy-SU(3) symmetry [323,327]. According to the dual-shell mechanism, SC can appear within the neutron intervals 7–8, 17–20, 34–40, 59–70, 96–112 …. N = 28 and N = 50 lie outside these intervals, therefore signs of inversion of single particle orbitals and/or signs of SC can appear only at the shell-model magic numbers 28 and 50, presumably due to particle–hole excitations across these gaps. On the contrary, N = 20 and N = 40 are the upper borders of two of these intervals; thus, SC can be expected also at lower N below them. It should be noticed that “narrow” IoIs correspond to shell-model magic numbers, while “wide” IoIs are related to isotropic 3D-HO magic numbers.
As mentioned in Section 2.10, islands of inversion [534] are formed by nuclei appearing close to semimagic or even doubly magic nuclei, expected to be spherical in their ground state, but found to be deformed. How does this happen? The proton–neutron interaction produces a shape transition, in which highly correlated many-particles–many-holes configurations, called intruders, become more bound than the spherical ones [534]. This mechanism is the same as the one widely accepted to be the microscopic cause of SC. Therefore, SC and SPT from spherical to deformed nuclei are one and the same effect, caused by the proton–neutron interaction. This is why the dual-shell mechanism, based on the proxy-SU(3) symmetry, works well in predicting the regions in which SC can appear. The present mechanism is also consistent with the argument that proton p-h excitations are caused by the neutrons and vice versa, in agreement to the role played in the shell-model framework by the tensor force.
Large-scale shell-model calculations [49] have suggested the merging of the N = 20 and N = 28 IoIs. Indeed the intruder states shown in Figure 13a reveal such a tendency, since the 0 2 + states keep falling from N = 12 to N = 20 . In order to clarify this issue, the 0 2 + states in 34 38 Mg 22 26 should be measured. Experimental findings [545] for the R 4 / 2 ratios within the ground state bands of these nuclei seem to support the idea of merging.
A similar picture appears in the case of the intruder states shown in Figure 13c, in which the 0 2 + states keep falling from N = 36 to N = 42 . In order to clarify the question of merging of the N = 40 and N = 50 [534] IoIs, the 0 2 + states in 72 76 Ni 44 48 should be measured.

13. Unified Perspectives for Shape Coexistence

13.1. Islands of Shape Coexistence and Shape/Phase Transitions

The conclusions of all subsections of Section 3, Section 4, Section 5, Section 6, Section 7, Section 8, Section 9 and Section 10 are depicted in Figure 1, in which the stripes among the nucleon numbers 7–8, 17–20, 34–40, 59–70, 96–112, 146–168, predicted by the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327] are also shown in azure. According to the dual-shell mechanism [334], SC can occur only within these stripes. We remark that all nuclei in which SC has been found fall within these stripes, with two notable exceptions. The first exception regards some N = Z nuclei, in which additional components of the nucleon–nucleon interaction are activated, as discussed in Section 11 and Section 11.1. The second exception consists of some nuclei within the N = 28 and N = 50 islands of inversion, discussed in Section 12. Notice that these are the IoIs which correspond to shell-model magic numbers created by the spin–orbit interaction, while the IoIs at N = 8 , N = 20 , and N = 40 , which correspond to magic numbers of the isotropic 3D-HO, fall within the relevant stripes of the dual-shell mechanism. A few other nuclei appear just outside the borders of the stripes, simply emphasizing the approximate nature of the arguments which led to their introduction (see Section 8 of [334] for relevant details).
The prediction of the proton and the neutron driplines has been recently receiving attention within several theoretical approaches, including large-scale shell-model [851], energy density functional theory [852], and ab initio calculations starting from chiral two- and three-nucleon interactions and taking advantage of the renormalization group [853]. The proton (neutron) driplines shown in Figure 1 and subsequent figures have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349].
In relation to the microscopic mechanisms underlying the existence of these islands, Figure 14 can be helpful. According to the results [579,580] obtained through CDFT with standard amount of pairing included, the “horizontal” islands along Z = 82 and Z = 50 correspond to proton particle–hole excitations across these proton magic numbers, created by the influence of the neutrons. The term neutron-induced shape coexistence has been used for them [579,580]. In contrast, the “vertical” islands around N = 90 and N = 60 correspond to neutron particle–hole excitations from normal parity orbitals to intruder parity orbitals, created by the influence of the protons. The term proton-induced shape coexistence has been used for them [579,580]. In the island formed around N = Z = 40 , both mechanisms are simultaneously present. As seen in Section 2.1, the influence of protons on the neutron gaps, as well as the influence of neutrons on the proton gaps, is the main effect caused by the tensor force [102,105,106,107] in the shell-model framework. Furthermore, it has been recently realized [445] within the IBM framework with configuration mixing that SC in the N = 60 region is due to the crossing of regular and intruder configurations.
While the nature of SC along the Z = 82 and Z = 50 lines has been understood since the early days of SC [2,3,83,84,85,86,87] as due to proton particle–hole excitations across these proton gaps, the nature of SC in the N = 90 and N = 60 regions has been mentioned as an open question (see Sections III.A.3 and III.A.4 of [4]) until very recently (see Section 4 of [8]). It seems that the recent findings clarify the nature of SC in the N = 90 and N = 60 regions in three different frameworks, namely the shell model [102,105,106,107], the CDFT with pairing interaction included [579,580], and the IBM-CM [445].
Within the shell-model [102,105,106,107] framework, the occurrence of SC is explained in terms of the action of the tensor force, which reduces the widths of certain energy gaps. Within the CDFT framework with pairing interaction included [579,580], SC is explained in terms of particle–hole excitations among shells of different parity. Within the IBM-CM [445] framework, SC is explained in terms of mixing of a configuration with particle–hole excitations with the ground state configuration. The interconnections between these three approaches is evident, reinforcing their validity. The reduced gaps predicted by the tensor force in the shell-model framework facilitate the occurrence of particle–hole excitations in the CDFT approach, which is equivalent to the appearance of a configuration with increased number of bosons in the IBM description.
The close connection between SPTs and SC, first suggested by Heyde et al. in 2004 [469], has recently been clarified through detailed studies in the Zr ( Z = 40 ) [438,439,683] and Sr ( Z = 38 ) [445] series of isotopes around N = 60 , in the framework of the IBM-CM. It turns out that the Z 40 , N 60 region offers a clear example of an SPT from spherical to prolate deformed shapes, similar to the SPT taking place in the Z 64 , N 90 region, which is known to represent the best manifestation of the first order SPT from spherical to prolate deformed shapes in the IBM [411], described in the framework of the Bohr Hamiltonian by the X(5)CPS [451].
The similarity between these two regions is further emphasized by Figure 2, in which all nuclei reported in the literature as candidates for the X(5) CPS are shown, along with the P 5 contours (see Equation (1)) indicating the borders of regions of deformed nuclei and therefore the regions of the onset of deformation. We see that all X(5) candidates in the Z 40 , N 60 and Z 64 , N 90 regions lie on the P 5 contours or they are touching them. The same happens in the Z 40 , N 40 region, as well as for all other X(5) candidates reported in the literature. For the sake of comparison, all nuclei reported in the literature as candidates for the E(5) CPS are also shown.
It should be emphasized that the islands of SC appearing in Figure 1 are based on the mechanism of particle–hole excitations, which can also be described as a change of the order of single-particle orbitals in the case of the islands of inversion. It might be possible that some other mechanism might exist, leading to SC far from the shores of these islands, but up to now signs of SC far from these islands is practically missing.

13.2. Systematics of Data

In order to observe SC, a K π = 0 + band lying close to the ground state band is needed. In Figure 3, the nuclei in which well-developed K π = 0 + bands are experimentally known are shown. It is seen that they form islands lying within the stripes predicted by the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327].
For SC to occur, the K π = 0 + band should not be too far away from the ground state band (gsb). In Figure 4, the nuclei in which the energy difference E ( 0 2 + ) E ( 2 1 + ) is less that 800 keV [406] are shown. Again they are seen to cluster together into islands, lying within the stripes predicted by the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327].
In Ref. [407] it has been shown that the proximity of the K π = 0 + band to the gsb can be expressed quantitatively through the energy ratio
R 0 / 2 = E ( 0 2 + ) E ( 2 1 + ) ,
by requiring R 0 / 2 < 5.7 , which is the parameter-independent value predicted by the X(5) CPS [451]. In parallel, it has been found [407] that for SC to occur, the energy ratio R 4 / 2 = E ( 4 1 + ) / E ( 2 1 + ) should be less than 3.05, which guarantees that the nucleus stays below the well-deformed region (notice that R 4 / 2 = 2.9 is the parameter-independent value predicted by the X(5) CPS) [451]). In other words, these two conditions guarantee that the nucleus is not getting well deformed, since in that case the K π = 0 + band goes to very high-energy values.
In addition, for SC to occur, the K π = 0 + band should be strongly connected to the ground-state band. A measure of this connection is the ratio
B 02 = B ( E 2 ; 0 2 + 2 1 + ) B ( E 2 ; 2 1 + 0 1 + ) ,
connecting the band-head of the excited K π = 0 + band to the first excited state of the ground state band. In Ref. [407] it has been found that for SC to occur one should have B 02 > 0.1 .
In Figure 5, all nuclei with experimentally known B 02 > 0.1 , R 0 / 2 < 5.7 , and R 4 / 2 < 3.05 , known to exhibit SC [407], are shown. Again they are seen to cluster together into islands, lying within the stripes predicted by the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327]. The only notable exception is seen in the Pt region ( 198 Pt 120 ), in which it is known that the level schemes can admit an alternative description not involving particle–hole excitations [614].
The B ( E 0 ; 0 2 + 0 1 + ) transition rate is expected [491,527] to be a very reliable indicator of SC, since strong B ( E 0 ) s would indicate bands with very different structures lying close in energy. However, the existing data [406,526,527] do not suffice for making meaningful systematics.
A dimensionless ratio of B(E2)s involving the 0 1 + , 2 1 + , as well as the 2 2 + or 2 3 + states has been found [854] to obtain values lying within two narrow bands in the case of nuclei exhibiting SC.
As shown in the recent (2022) review article by Garrett et al. [8], several additional experimental quantities (transfer cross sections, decay properties, charge radii) can serve as indicators of the presence of SC. It would be interesting to gather existing data for each of these quantities on the nuclear chart and see to which extend they are consistent with the predictions of the dual-shell mechanism. The impression formed from Figures 7, 13, 14, 15, 21, 33, and 52 of Ref. [8], in which the information from many different experimental quantities is simultaneously shown in various regions of the nuclear chart, is that these extra quantities also form islands consistent with the predictions of the dual-shell mechanism, but a detailed study of each quantity separately over the whole nuclear chart is necessary before reaching any final conclusions.

13.3. Comparison to Earlier Compilations

In Figure 8 of the authoritative review article by Heyde and Wood [4], the main regions of SC are shown. In Figure 6 a comparison between the regions of SC shown in that figure and the present findings is attempted.
The regions L and J of Figure 8 of [4] correspond to Z 82 and Z 50 respectively. They are elongated along the Z = 82 and Z = 50 lines and fall almost entirely within the N = 96 –112 and N = 59 –70 stripes predicted by the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327]. Furthermore, they agree with the findings of the present work, depicted in Figure 1 of the present work, in which it is seen that these islands are actually wider in the Z direction in comparison to Figure 8 of [4]. Their microscopic origin is attributed [579,580] to proton particle–hole excitations across the Z = 82 and Z = 50 energy gaps, corresponding to neutron-induced SC, in accordance to the early suggestions of Heyde et al. [2,3,83,84,85,86,87].
The regions K and I of Figure 8 of [4] correspond to Z 64 , N 90 and Z 40 , N 60 respectively. The first one falls entirely within the Z = 59 –70 stripe of the dual-shell mechanism, while the second one falls partly within the Z = 34 –40 stripe and partly within the N = 59 –70 stripe of the dual-shell mechanism. Their microscopic origin is attributed [579,580] to neutron particle–hole excitations from normal parity orbitals to intruder orbitals, corresponding to proton-induced SC, thus answering the long-standing question [4,8] about their nature. The close relation of SC in these two very similar in nature regions to the shape/phase transition from spherical to deformed shapes has also been understood, as mentioned in Section 13.1. The regions K and I agree with the findings of the present work, depicted in Figure 1 of the present work, except for the upper left part of region I, for which minimal evidence has been found in the present work.
It is interesting that in Figure 6 the analogs of regions L and K remain disconnected, while the analogs of regions J and I almost join within the N = 59 –70 stripe of the dual-shell mechanism, practically separated only by the Ru ( Z = 44 ) isotopes.
The region H of Figure 8 of [4] corresponds to Z 34 , N 40 . It falls partly within the Z = 34 –40 stripe and partly within the N = 34 –40 stripe of the dual-shell mechanism, in agreement with the findings of the present work, depicted in Figure 1 of the present work, except for the lower right part of region H, for which minimal evidence has been found in the present work. The latter fact reinforces the interpretation of this region as due to the simultaneous presence of neutron-induced particle–hole excitations of protons and proton-induced particle–hole excitations of neutrons [579,580].
Region G of Figure 8 of [4] corresponds to superdeformed nuclei, which have not been considered in the present work, in view of the limitations set in Section 1.1.
Region A of Figure 8 of [4] corresponds to N = Z nuclei, discussed in Section 11 and Section 11.1. Region A extends up to N = Z = 28 , while in the present work it goes further up to N = Z = 42 , as seen in Figure 1 of the present work.
Regions B and C of Figure 8 of [4] correspond to the N = 8 and N = 20 islands of inversion, in agreement with what is seen in Figure 1 of the present work, in which the N = 20 IoI joins the N = Z line and probably starts to be extended above it.
Regions D and F of Figure 8 of [4] correspond to the N = 28 island of inversion, in agreement to what is seen in Figure 1 of the present work, in which these two regions are merged and the N = 28 IoI reaches the N = Z line.
Finally, the lower part of region E of Figure 8 of [4] falls within the Z = 17 –20 stripe of the dual-shell mechanism, while the upper part reaches the N = Z line, in agreement with what is seen in Figure 1 of the present work.
What is missing in Figure 8 of [4], in comparison to Figure 1 of the present work, is the N = 50 island of inversion, which had not been discovered yet at that time [534].
In conclusion, very close agreement is seen between the pioneering, early Figure 8 of [4] and Figure 1 of the present work, the main differences being the discovery of new examples of SC over the last 12 years, as well as the improved theoretical understanding of the nature of SC in the Z 64 , N 90 and Z 40 , N 60 regions, connected to the appearance of shape/phase transitions [438,439,445,683]. These regions have been found to correspond to neutron-induced particle–hole excitations caused by the protons [579,580], in accordance with the mechanism supported by the tensor force [102,105,106,107] of the nucleon–nucleon interaction.
It should be noticed that global calculations for low-lying levels and B(E2)s among them have been carried out [236] in the framework of CDFT. In addition to the regions of SC and the IoIs discussed above, evidence for SC is seen in the regions Z 64 , N 76 and Z 40 , N 70 . Both regions lie within the stripes predicted by the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327]. In addition, in Figure 1 of the present work, one sees that evidence corroborating these predictions already exists.

14. Conclusions

The last decade has seen a rapid growth in experimental searches for SC, assisted by the new possibilities provided by the modern radioactive ion beam facilities built worldwide. In parallel, our understanding of its microscopic origins has been greatly improved, both through large-scale shell-model and density functional theory calculations, as well as through the use of symmetries.
It is by now becoming evident in even–even nuclei that SC of the ground-state band with another K π = 0 + band, lying close in energy but possessing radically different structure, is not occurring all over the nuclear chart, but within certain regions of it, called islands of SC (Section 13.1 and Section 13.3).
Within these islands, SC is created either by proton particle–hole excitations caused by the influence of neutrons (neutron-induced shape coexistence), or by neutron particle–hole excitations caused by the influence of protons (proton-induced shape coexistence), while in some regions both mechanisms can act simultaneously (Section 13.1). These mechanisms, corroborated by CDFT (Section 13.1), can be traced back to the properties of the tensor force within the nuclear shell model (Section 2.1.1).
In light- and medium-mass nuclei the shell-model theoretical approach (including symmetry-based no-core shell-model schemes) prevails, while in heavy nuclei the mean-field approaches, including both non-relativistic schemes and CDFT, as well as symmetry-based methods are used.
In light- and medium-mass nuclei, a rearrangement of nuclear orbitals, similar in nature to the particle–hole mechanism leading to SC, appears along the N = 8 , 20, 28, 40, 50 lines of the nuclear chart, forming the N = 8 , 20, 28, 40, 50 islands of inversion (Section 12). In addition, SC appears along the N = Z line of the nuclear chart, in which the contribution of spin-aligned proton–neutron pairs becomes important (Section 11 and Section 11.1).
In medium-mass and heavy nuclei, the connection between SC and critical point symmetries corresponding to a first-order shape/phase transition from spherical to deformed nuclear shapes has been clarified (Section 13.1).
Based on the systematics of data, as well as on symmetry considerations, quantitative rules have been developed, predicting regions in which SC can appear, as a possible guide for the experimental effort (Section 13.2).

15. Outlook

Further experimental results on SC can help in improving our understanding of the details of the nucleon–nucleon interaction, as well as of its modifications occurring far from stability.
The separate study of existing data [8] for individual structure indicators over the nuclear chart can provide further evidence for the regions in which SC can be expected (Section 13.2).
Growing evidence for multiple SC (Section 2.9) will provide a further test for the SC islands predicted by the dual-shell mechanism developed within the proxy-SU(3) symmetry. If the mechanism holds, multiple SC should be confined within the areas of these islands [335].
Superheavy nuclei present a special interest, since a rich variety of novel phenomena seems to be predicted there (Section 3.5).
Theoretical predictions have been found in some cases to be sensitive to the parameter sets used (Section 5.2). Parameter-independent symmetry predictions can be useful in some cases in resolving any disagreements. The development of methods allowing the estimation of the accuracy [59] of numerical theoretical predictions can also be helpful in the long run.
It would be interesting to examine if the particle–hole mechanism found in CDFT calculations using the DDME2 functional and including standard pairing [579,580] in the N = 90 , N = 60 and N = 40 regions is corroborated by calculations in other relativistic or non-relativistic calculations using different forces and/or parametrizations.
While the CPS X(5) is well established for the description of the first-order shape/phase transition seen in the N 90 , Z 60 region (Section 5.5) and the SC occurring in relation to it, it seems that a more flexible scheme is needed for the description of the very similar SPTs seen in the regions N 60 , Z 40 (Section 8.5) and N 40 , Z 34 (Section 9.5), and the SC occurring in relation to them.
The relation between the upper limits of SC and the prolate–oblate shape/phase transition should also be explored. The upper limit of the N = 96 –112 region, in which SC can be expected according the dual-shell mechanism developed within the proxy-SU(3) symmetry in the heavy rare earths, touches N = 114 , at which the shape/phase transition from prolate to oblate shapes is expected to occur in this region (see Table II of [324]). It should be noticed that the textbook example [214] of the O(6) dynamical symmetry of IBM, 196 Pt 118 , lies next to this point. The relation among the upper border of SC, the prolate-to-oblate shape/phase transition, and the O(6) dynamical symmetry of IBM needs to be clarified.
A similar situation occurs theoretically for the rare earths with neutrons in the 50–82 shell. The upper limit of the N = 59 –70 region, in which SC can be expected according the dual-shell mechanism developed within the proxy-SU(3) symmetry in the heavy rare earths, touches N = 72 , at which the shape/phase transition from prolate to oblate shapes is expected to occur in this region (see Table III of [324]).
The problem of clarifying the relation among SC, the E(5) CPS related to the second order shape/phase transition between spherical and γ -unstable shapes, and the dynamical symmetry O(6) of IBM has been described in detail in Section 6.2, to which the reader is referred.
Negative parity bands related to the octupole degree of freedom [855,856,857], not considered in the present review, might offer a fertile ground for the search for SC of positive- and negative-parity bands, since it is known [858,859] that octupole deformation occurs when a negative-parity band with levels characterized by J π = 1 , 3 , 5 , 7 , 9 , …joins the ground-state band, which has levels with J π = 0 + , 2 + , 4 + , 6 + , 8 + , …, thus forming a single band with J π = 0 + , 1 , 2 + , 3 , 4 + , 5 , 6 + , 7 , 8 + , …. The fact that the bands are lying close in energy and that a double-well potential [860] in the octupole deformation collective variable β 3 is needed in order to describe the octupole deformed shape may favor the appearance of SC. The regions to be investigated are the ones lying close to the octupole magic numbers 34, 56, 88, 134 (see Figure 1 of [857]), where the manifestation of octupole effects is expected. For some examples, see [29,30] for the N 88 region, and [31] in the Z 56 region.

Author Contributions

Conceptualization, D.B., A.M., S.K.P., T.J.M. and N.M.; methodology, D.B., A.M., S.K.P., T.J.M. and N.M.; writing–original draft preparation, D.B.; writing–review and editing, D.B., A.M., S.K.P., T.J.M. and N.M.; supervision, D.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Bulgarian National Science Foundation (BNSF) under contract no. KP-06-N48/1.

Data Availability Statement

No new data have been reported in this review article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

5DQCHfive-dimensional quadrupole collective Hamiltonian
CDFTcovariant density functional theory
CPScritical point symmetry
GCMgenerator coordinate method
HFHartree–Fock
HFBHartree–Fock–Bogoliubov
IBMinteracting boson model
IBM-CMinteracting boson model with configuration mixing
MCSMMonte Carlo shell model
PDSpartial dynamical symmetry
PPQpairing plus quadrupole
RHBrelativistic Hartree–Bogoliubov
RMFrelativistic mean field
SANCSMsymmetry-adapted no-core shell model
SCshape coexistence
SPTshape/phase transition
TDHFBtime-dependent Hartree–Fock–Bogoliubov
VAMPIRvariation after mean-field projection in realistic model spaces

Appendix A. Theoretical Methods

In this Appendix, a list of theoretical terms and models used in the text of this review are listed (in alphabetical order) first, along with the section in which their definition and/or brief description appears. Below this list, additional theoretical terms and models used in the text of this review without definition, mostly in Section 3, Section 4, Section 5, Section 6, Section 7, Section 8, Section 9 and Section 10, are briefly discussed, again in alphabetical order.

Appendix A.1. Theoretical Approaches

BCS approximationSection 2.2
beyond-mean-field approachSection 2.2
collective model of Bohr and MottelsonSection 2.3
covariant density functional theory (CDFT)Section 3.1
critical point symmetry (CPS)Section 2.6
density functional theory (DFT)Section 2.2
dual shell mechanismSection 2.4.6
E(5) CPSSection 2.6
Elliott SU(3) modelSection 2.4.1
extended IBMSection 2.5
five-dimensional quadrupole collective Hamiltonian (5DQCH)Section 2.2
generator coordinate method (GCM)Section 2.2
Gogny interactionSection 2.2
Hartree–Fock (HF)Section 2.2
Hartree–Fock–Bogoliubov (HFB)Section 2.2
IBM with configuration mixing (IBM-CM)Section 2.5
interacting boson model (IBM)Section 2.5
Monte Carlo shell model (MCSM)Section 2.1
Nilsson modelSection 2.3
no-core shell modelSection 2.1
O(6) symmetrySection 2.7
pairing interactionSection 2.3
pairing plus quadrupole (PPQ) modelSection 2.3
partial dynamical symmetry (PDS)Section 2.5
proxy-SU(3) symmetrySection 2.4.4
quasi-SU(3) symmetrySection 2.4.3
pseudo-SU(3) symmetrySection 2.4.2
relativistic mean field (RMF)Section 2.2
shape/phase transition (SPT)Section 2.6
Skyrme interactionSection 2.2
spdf-IBMSection 2.5
symmetry-adapted no-core shell model (SANCSM)Section 2.1
symplectic modelSection 2.4.5
tensor forceSection 2.1.1
time-dependent Hartree–Fock–Bogoliubov (TDHFB)Section 2.2
VAMPIRSection 2.2
X(5) CPSSection 2.6
Z(5) CPSSection 2.6

Appendix A.2. The Constrained HFB Theory

The (HF) and (HFB) variational methods, mentioned in Section 2.2, determine only one point on the potential energy surface, namely the minimum. If somebody wishes to locate additional local minima, one has to introduce relevant constraints (i.e., subsidiary conditions) and look for a minimum obeying these additional restrictions (see Section 7.6 of [111]). For an application of configuration-constrained calculations to SC, see [575].

Appendix A.3. The Cranking Model and the Routhians

The cranking model, introduced by Inglis in 1954 [861,862] (see also Section 3.4 of [111]), in its initial form describes nuclear motion in a rotating potential well. However, a similar formulation can be applied to more general collective motions [863]; see, for example, [24,25,618,789] for applications related to SC.
The single particle energies in the rotating system are called Routhians (see Section 12.1 of [104]). Total-Routhian-surface (TRS) calculations related to SC have been performed, for example, in [574,686].

Appendix A.4. The Deformed Woods–Saxon Potential

The deformed Woods–Saxon potential [864] is a generalization of the spherical Woods–Saxon potential, in which the distance of a point from the nuclear surface is specified by a set of shape parameters and is generated numerically [865]. Applications of the deformed Woods–Saxon potential related to SC can be found in [25,572,573,575,578].

Appendix A.5. The Nilsson–Strutinsky Model

In phenomenological models, like the liquid drop model of vibrations and rotations (see Appendix 6A of [249]), the distribution of nucleons in phase space is supposed to be homogeneous, which is not the case in the nuclear shell model, in which the distribution becomes inhomogeneous. Strutinsky’s method ([866], see also Section 2.9 of [111]) allows for calculations of the shell-model corrections to the liquid drop energy of the nucleus, starting from Nilsson’s level scheme [104,250,251], which is a simplified deformed shell model. The corrections depend on the occupation number and on the deformation. The Nilsson–Strutinsky model has been used for the description of SC, for example in [572,577,789,790].

Appendix A.6. The Particle-Plus-Rotor Model

In order to describe the interplay between the independent motion of individual particles and the collective motion of the nucleus as a whole, Bohr and Mottelson in 1953 ([247], see also Section 3.3 of [111]) suggested to describe the nucleus in terms of a few valence particles, which move more or less independently in the deformed well of the core, and couple them to the collective rotor, which represents the rest of the particles. This particle-plus-rotor or particle-core model has been used for the description of SC for example in [576,600].

Appendix A.7. The Projection Method

The projection method is a method of restoring broken symmetries. It is used in cases in which one has a Hamiltonian possessing a certain symmetry and a wave function of this Hamiltonian violating this symmetry. It is a variational procedure which minimizes the energy and at the same time produces a wave function satisfying the symmetry of the Hamiltonian [111]. For details on particle number projection and angular momentum projection one may see Section 11.4 of the book [111]. For example, spin and number projection has been used in the excited VAMPIR [197] approach.

References

  1. Morinaga, H. Interpretation of Some of the Excited States of 4n Self-Conjugate Nuclei. Phys. Rev. 1956, 101, 254. [Google Scholar] [CrossRef]
  2. Heyde, K.; Van Isacker, P.; Waroquier, M.; Wood, J.L.; Meyer, R.A. Coexistence in odd-mass nuclei. Phys. Rep. 1983, 102, 291. [Google Scholar] [CrossRef]
  3. Wood, J.L.; Heyde, K.; Nazarewicz, W.; Huyse, M.; Van Duppen, P. Coexistence in even-mass nuclei. Phys. Rep. 1992, 215, 101. [Google Scholar] [CrossRef]
  4. Heyde, K.; Wood, J.L. Shape coexistence in atomic nuclei. Rev. Mod. Phys. 2011, 83, 1467. [Google Scholar] [CrossRef]
  5. Heyde, K.; Wood, J.L. Nuclear shapes: From earliest ideas to multiple shape coexisting structures. Phys. Scr. 2016, 91, 083008. [Google Scholar] [CrossRef]
  6. Poves, A. Shape coexistence in nuclei. J. Phys. G Nucl. Part. Phys. 2016, 43, 020401. [Google Scholar] [CrossRef]
  7. Wood, J.L.; Heyde, K. A focus on shape coexistence in nuclei. J. Phys. G Nucl. Part. Phys. 2016, 43, 020402. [Google Scholar] [CrossRef]
  8. Garrett, P.E.; Zielińska, M.; Clément, E. An experimental view on shape coexistence in nuclei. Prog. Part. Nucl. Phys. 2022, 124, 103931. [Google Scholar] [CrossRef]
  9. Nobel Foundation. Nobel Lectures, Physics 1963–1970; Elsevier: Amsterdam, The Netherlands, 1972. [Google Scholar]
  10. Nobel Foundation. Nobel Lectures, Physics 1971–1980; Lundqvist, S., Ed.; World Scientific: Singapore, 1992. [Google Scholar]
  11. Bark, R.A.; Dracoulis, G.D.; Stuchbery, A.E. Shape coexistence or particle alignment in the light osmium isotopes 171Os, 172Os and 173Os. Nucl. Phys. A 1990, 514, 503. [Google Scholar] [CrossRef]
  12. Davidson, P.M.; Dracoulis, G.D.; Kibédi, T.; Byrne, A.P.; Anderssen, S.S.; Baxter, A.M.; Fabricius, B.; Lane, G.J.; Stuchbery, A.E. Non-yrast states and shape co-existence in 172Os. Nucl. Phys. A 1994, 568, 90. [Google Scholar] [CrossRef]
  13. Joss, D.T.; King, S.L.; Page, R.D.; Simpson, J.; Keenan, A.; Amzal, N.; Bäck, T.; Bentley, M.A.; Cederwall, B.; Cocks, J.F.; et al. Identification of excited states in 167Os and 168Os: Shape coexistence at extreme neutron deficiency. Nucl. Phys. A 2001, 689, 631. [Google Scholar] [CrossRef]
  14. Kibédi, T.; Dracoulis, G.D.; Byrne, A.P.; Davidson, P.M. Low-spin non-yrast states in light tungsten isotopes and the evolution of shape coexistence. Nucl. Phys. A 2001, 688, 669. [Google Scholar] [CrossRef]
  15. Paul, E.S.; Fossan, D.B.; Liang, Y.; Ma, R.; Xu, N. Shape coexistence in 132Ba. Phys. Rev. C 1989, 40, 1255. [Google Scholar] [CrossRef]
  16. Procter, M.G.; Cullen, D.M.; Scholey, C.; Niclasen, B.; Mason, P.J.R.; Rigby, S.V.; Dare, J.A.; Dewald, A.; Greenlees, P.T.; Iwasaki, H.; et al. Lifetime measurements and shape coexistence in 144Dy. Phys. Rev. C 2010, 81, 054320. [Google Scholar] [CrossRef]
  17. Revill, J.P.; Paul, E.S.; Wang, X.; Riley, M.A.; Simpson, J.; Janssens, R.V.F.; Ollier, J.; Boston, A.J.; Carpenter, M.P.; Chiara, C.J.; et al. Quadrupole moments of coexisting collective shapes at high spin in 154Er. Phys. Rev. C 2013, 88, 031304. [Google Scholar] [CrossRef]
  18. Smith, M.B.; Appelbe, D.E.; Twin, P.J.; Beausang, C.W.; Beck, F.A.; Bentley, M.A.; Cullen, D.M.; Curien, D.; Dagnall, P.J.; de France, G.; et al. Deformed rotational cascades in 152Dy: Further evidence for shape coexistence at high spin. Phys. Rev. C 2000, 61, 034314. [Google Scholar] [CrossRef]
  19. Fant, B.; Tanner, R.J.; Butler, P.A.; James, A.N.; Jones, G.D.; Poynter, R.J.; White, C.A.; Ying, K.L.; Love, D.J.G.; Simpson, J.; et al. Search for shape coexistence in 194Pb. J. Phys. G Nucl. Part. Phys. 1991, 17, 319. [Google Scholar]
  20. Ionescu-Bujor, M.; Iordachescu, A.; Mǎrginean, N.; UR, C.A.; Bucurescu, D.; Suliman, G.; Balabanski, D.L.; Brandolini, F.; Chmel, S.; Detistov, P.; et al. Shape coexistence in neutron-deficient Pb nuclei probed by quadrupole moment measurements. Phys. Lett. B 2007, 650, 141. [Google Scholar] [CrossRef]
  21. Ionescu-Bujor, M.; Iordachescu, A.; Ur, C.A.; Marginean, N.; Suliman, G.; Bucurescu, D.; Brandolini, F.; Della Vedova, F.; Chmel, S.; Lenzi, S.M.; et al. g factors of coexisting isomeric states in 188Pb. Phys. Rev. C 2010, 81, 024323. [Google Scholar] [CrossRef]
  22. Mare, A.S.; Petrovici, A. Shape coexistence and isomeric states in 94Pd within a beyond-mean-field approach. Phys. Rev. C 2022, 106, 054306. [Google Scholar] [CrossRef]
  23. Nara Singh, B.S.; Liu, Z.; Wadsworth, R.; Grawe, H.; Brock, T.S.; Boutachkov, P.; Braun, N.; Blazhev, A.; Górska, M.; Pietri, S.; et al. 16+ spin-gap isomer in 96Cd. Phys. Rev. Lett. 2011, 107, 172502. [Google Scholar] [CrossRef] [PubMed]
  24. Oi, M.; Walker, P.M.; Ansari, A. Shape coexistence and tilted-axis rotation in neutron-rich hafnium isotopes. Phys. Lett. B 2001, 505, 75. [Google Scholar] [CrossRef]
  25. Dudek, J.; Nazarewicz, W.; Rowley, N. Shape coexistence effects and superdeformation in 84Zr. Phys. Rev. C 1987, 35, 1489. [Google Scholar] [CrossRef] [PubMed]
  26. Dudek, J.; Nazarewicz, W.; Szymanski, Z.; Leander, G.A. Abundance and systematics of nuclear superdeformed states; relation to the pseudospin and pseudo-SU(3) symmetries. Phys. Rev. Lett. 1987, 59, 1405. [Google Scholar] [CrossRef] [PubMed]
  27. Lagergren, K.; Cederwall, B.; Bäck, T.; Wyss, R.; Ideguchi, E.; Johnson, A.; Ataç, A.; Axelsson, A.; Azaiez, F.; Bracco, A.; et al. Coexistence of Superdeformed Shapes in 154Er. Phys. Rev. Lett. 2001, 87, 022502. [Google Scholar] [CrossRef]
  28. Grahn, T.; Petts, A.; Scheck, M.; Butler, P.; Page, R.; Dewald, A.; Jolie, J.; Melon, B.; Pissulla, T.; Hornillos, M.; et al. Evolution of collectivity in 180Hg and 182Hg. Phys. Rev. C 2009, 80, 014324. [Google Scholar] [CrossRef]
  29. Leandri, J.; Piepenbring, R. Coupled Kπ=0+ and Kπ=0 vibrations in 152Sm. Phys. Lett. B 1989, 232, 437. [Google Scholar] [CrossRef]
  30. Urban, W.; Lieder, R.M.; Gast, W.; Hebbinghaus, G.; Krämer-Flecken, A.; Blume, K.P.; Hübel, H. Evidence for coexistence of reflection asymmetric and symmetric shapes in 150Sm. Phys. Lett. B 1987, 185, 331. [Google Scholar] [CrossRef]
  31. Zhu, S.J.; Wang, E.H.; Hamilton, J.H.; Ramayya, A.V.; Liu, Y.X.; Brewer, N.T.; Luo, Y.X.; Rasmussen, J.O.; Xiao, Z.G.; Huang, Y.; et al. Coexistence of Reflection Asymmetric and Symmetric Shapes in 144Ba. Phys. Rev. Lett. 2020, 124, 032501. [Google Scholar] [CrossRef]
  32. Haxel, O.; Jensen, J.H.D.; Suess, H.E. On the “Magic Numbers” in Nuclear Structure. Phys. Rev. 1949, 75, 1766. [Google Scholar] [CrossRef]
  33. Heyde, K.L.G. The Nuclear Shell Model; Springer: Berlin/Heidelberg, Germany, 1990. [Google Scholar]
  34. Mayer, M.G. On Closed Shells in Nuclei. Phys. Rev. 1948, 74, 235. [Google Scholar] [CrossRef]
  35. Mayer, M.G. On Closed Shells in Nuclei. II. Phys. Rev. 1949, 75, 1969. [Google Scholar] [CrossRef]
  36. Mayer, M.G.; Jensen, J.H.D. Elementary Theory of Nuclear Shell Structure; Wiley: New York, NY, USA, 1955. [Google Scholar]
  37. Talmi, I. Simple Models of Complex Nuclei: The Shell Model and the Interacting Boson Model; Harwood: Chur, Switzerland, 1993. [Google Scholar]
  38. Iachello, F. Lie Algebras and Applications; Springer: Berlin/Heidelberg, Germany, 2006. [Google Scholar]
  39. Moshinsky, M.; Smirnov, Y.F. The Harmonic Oscillator in Modern Physics; Harwood: Amsterdam, The Netherlands, 1996. [Google Scholar]
  40. Wybourne, B.G. Classical Groups for Physicists; Wiley: New York, NY, USA, 1974. [Google Scholar]
  41. Woods, R.D.; Saxon, D.S. Diffuse Surface Optical Model for Nucleon-Nuclei Scattering. Phys. Rev. 1954, 95, 577. [Google Scholar] [CrossRef]
  42. Bethe, H.A.; Bacher, R.F. Nuclear Physics A. Stationary States of Nuclei. Rev. Mod. Phys. 1936, 8, 82. [Google Scholar] [CrossRef]
  43. Elsasser, W.M. Sur le principe de Pauli dans les noyaux-II. J. Phys. Radium 1934, 5, 389–397. [Google Scholar] [CrossRef]
  44. Elsasser, W.M. Sur le principe de Pauli dans les noyaux-III. J. Phys. Radium 1934, 5, 635–639. [Google Scholar] [CrossRef]
  45. Caurier, E.; Martínez-Pinedo, G.; Nowacki, F.; Poves, A.; Zuker, A.P. The shell model as a unified view of nuclear structure. Rev. Mod. Phys. 2005, 77, 427. [Google Scholar] [CrossRef]
  46. Poves, A. Shape coexistence: The shell model view. J. Phys. G Nucl. Part. Phys. 2016, 43, 024010. [Google Scholar] [CrossRef]
  47. Caurier, E.; Nowacki, F.; Poves, A. Large-scale shell model calculations for exotic nuclei. Eur. Phys. J. A 2002, 15, 145. [Google Scholar] [CrossRef]
  48. Caurier, E.; Menéndez, J.; Nowacki, F.; Poves, A. Coexistence of spherical states with deformed and superdeformed bands in doubly magic 40Ca: A shell-model challenge. Phys. Rev. C 2007, 75, 054317. [Google Scholar] [CrossRef]
  49. Caurier, E.; Nowacki, F.; Poves, P. Merging of the islands of inversion at N = 20 and N = 28. Phys. Rev. C 2014, 90, 014302. [Google Scholar] [CrossRef]
  50. Lenzi, S.M.; Nowacki, F.; Poves, A.; Sieja, K. Island of inversion around 64Cr. Phys. Rev. C 2010, 82, 054301. [Google Scholar] [CrossRef]
  51. Valiente-Dobón, J.J.; Poves, A.; Gadea, A.; Fernández-Domínguez, B. Broken mirror symmetry in 36S and 36Ca. Phys. Rev. C 2018, 98, 011302. [Google Scholar] [CrossRef]
  52. Honma, M.; Mizusaki, T.; Otsuka, T. Diagonalization of Hamiltonians for Many-Body Systems by Auxiliary Field Quantum Monte Carlo Technique. Phys. Rev. Lett. 1995, 75, 1284. [Google Scholar] [CrossRef]
  53. Honma, M.; Mizusaki, T.; Otsuka, T. Nuclear Shell Model by the Quantum Monte Carlo Diagonalization Method. Phys. Rev. Lett. 1996, 77, 3315. [Google Scholar] [CrossRef] [PubMed]
  54. Mizusaki, T.; Honma, M.; Otsuka, T. Quantum Monte Carlo diagonalization with angular momentum projection. Phys. Rev. C 1996, 53, 2786. [Google Scholar] [CrossRef] [PubMed]
  55. Otsuka, T.; Honma, M.; Mizusaki, T. Structure of the N=Z=28 Closed Shell Studied by Monte Carlo Shell Model Calculation. Phys. Rev. Lett. 1998, 81, 1588. [Google Scholar] [CrossRef]
  56. Otsuka, T.; Honma, M.; Mizusaki, T.; Shimizu, N.; Utsuno, Y. Monte Carlo shell model for atomic nuclei. Prog. Part. Nucl. Phys. 2001, 47, 319. [Google Scholar] [CrossRef]
  57. Shimizu, N.; Otsuka, T.; Mizusaki, T.; Honma, M. Transition from Spherical to Deformed Shapes of Nuclei in the Monte Carlo Shell Model. Phys. Rev. Lett. 2001, 86, 1171. [Google Scholar] [CrossRef]
  58. Shimizu, N.; Abe, T.; Tsunoda, Y.; Utsuno, Y.; Yoshida, T.; Mizusaki, T.; Honma, M.; Otsuka, T. New-generation Monte Carlo shell model for the K computer era. Prog. Theor. Exp. Phys. 2012, 2012, 01A205. [Google Scholar] [CrossRef]
  59. Yoshida, S.; Shimizu, N.; Togashi, T.; Otsuka, T. Uncertainty quantification in the nuclear shell model. Phys. Rev. C 2018, 98, 061301. [Google Scholar] [CrossRef]
  60. Hasegawa, M.; Kaneko, K.; Mizusaki, T.; Sun, Y. Phase transition in exotic nuclei along the N = Z line. Phys. Lett. B 2007, 656, 51. [Google Scholar] [CrossRef]
  61. Kaneko, K.; Hasegawa, H.; Mizusaki, T. Shape transition and oblate-prolate coexistence in N = Z fpg-shell nuclei. Phys. Rev. C 2004, 70, 051301. [Google Scholar] [CrossRef]
  62. Kaneko, K.; Casten, R.F.; Hasegawa, M.; Mizusaki, T.; Zhang, J.-Y.; McCutchan, E.A.; Zamfir, N.V.; Krücken, R. Anomalous behavior of the first excited 0+ state in NZ nuclei. Phys. Rev. C 2005, 71, 014319. [Google Scholar] [CrossRef]
  63. Kaneko, K.; Mizusaki, T.; Sun, Y.; Tazaki, S. Toward a unified realistic shell-model Hamiltonian with the monopole-based universal force. Phys. Rev. C 2014, 89, 011302. [Google Scholar] [CrossRef]
  64. Kaneko, K.; Mizusaki, T.; Sun, Y.; Tazaki, S. Systematical shell-model calculation in the pairing-plus-multipole Hamiltonian with a monopole interaction for the pf5/2g9/2 shell. Phys. Rev. C 2015, 92, 044331. [Google Scholar] [CrossRef]
  65. Kaneko, K.; Sun, Y.; Wadsworth, R. Shape coexistence and shape transition in self-conjugate nucleus 72Kr and the tensor force. Phys. Scr. 2017, 92, 114008. [Google Scholar] [CrossRef]
  66. Lay, J.A.; Vitturi, A.; Fortunato, L.; Tsunoda, Y.; Togashi, T.; Otsuka, T. Two-particle transfer processes as a signature of shape phase transition in Zirconium isotopes. Phys. Lett. B 2023, 838, 137719. [Google Scholar] [CrossRef]
  67. Mizusaki, T.; Otsuka, T.; Utsuno, Y.; Honma, M.; Sebe, T. Shape coexistence in doubly-magic 56Ni by the Monte Carlo shell model. Phys. Rev. C 1999, 59, R1846. [Google Scholar] [CrossRef]
  68. Mizusaki, T.; Otsuka, T.; Honma, M.; Brown, B.A. Spherical-deformed shape coexistence for the pf shell in the nuclear shell model. Phys. Rev. C 2001, 63, 044306. [Google Scholar] [CrossRef]
  69. Reinhard, P.-G.; Dean, D.J.; Nazarewicz, W.; Dobaczewski, J.; Maruhn, J.A.; Strayer, M.R. Shape coexistence and the effective nucleon-nucleon interaction. Phys. Rev. C 1999, 60, 014316. [Google Scholar] [CrossRef]
  70. Shimizu, N.; Utsuno, Y.; Mizusaki, T.; Honma, M.; Tsunoda, Y.; Otsuka, T. Variational procedure for nuclear shell-model calculations and energy-variance extrapolation. Phys. Rev. C 2012, 85, 054301. [Google Scholar] [CrossRef]
  71. Togashi, T.; Tsunoda, Y.; Otsuka, T.; Shimizu, N. Quantum Phase Transition in the Shape of Zr isotopes. Phys. Rev. Lett. 2016, 117, 172502. [Google Scholar] [CrossRef]
  72. Togashi, T.; Tsunoda, Y.; Otsuka, T.; Shimizu, N.; Honma, M. Novel Shape Evolution in Sn Isotopes from Magic Numbers 50 to 82. Phys. Rev. Lett. 2018, 121, 062501. [Google Scholar] [CrossRef] [PubMed]
  73. Tsunoda, Y.; Otsuka, T.; Shimizu, N.; Honma, M.; Utsuno, Y. Novel shape evolution in exotic Ni isotopes and configuration-dependent shell structure. Phys. Rev. C 2014, 89, 031301. [Google Scholar] [CrossRef]
  74. 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]
  75. Utsuno, Y.; Shimizu, N.; Otsuka, T.; Yoshida, T.; Tsunoda, Y. Nature of Isomerism in Exotic Sulfur Isotopes. Phys. Rev. Lett. 2015, 114, 032501. [Google Scholar] [CrossRef]
  76. Navrátil, P.; Vary, J.P.; Barrett, B.R. Properties of 12C in the Ab Initio Nuclear Shell Model. Phys. Rev. Lett. 2000, 84, 5728. [Google Scholar] [CrossRef]
  77. Navrátil, P.; Vary, J.P.; Barrett, B.R. Large-basis ab initio no-core shell model and its application to 12C. Phys. Rev. C 2000, 62, 054311. [Google Scholar] [CrossRef]
  78. Dytrych, T.; Sviratcheva, K.D.; Draayer, J.P.; Bahri, C.; Vary, J.P. Ab initio symplectic no-core shell model. J. Phys. G Nucl. Part. Phys. 2008, 35, 123101. [Google Scholar] [CrossRef]
  79. Launey, K.D.; Draayer, J.P.; Dytrych, T.; Sun, G.-H.; Dong, S.-H. Approximate symmetries in atomic nuclei from a large-scale shell-model perspective. Int. J. Mod. Phys. E 2015, 24, 1530005. [Google Scholar] [CrossRef]
  80. Launey, K.D.; Dytrych, T.; Draayer, J.P. Symmetry-guided large-scale shell-model theory. Prog. Part. Nucl. Phys. 2016, 89, 101. [Google Scholar] [CrossRef]
  81. Launey, K.D.; Dytrych, T.; Sargsyan, G.H.; Baker, R.B.; Draayer, J.P. Emergent symplectic symmetry in atomic nuclei: Ab initio symmetry-adapted no-core shell model. Eur. Phys. J. Spec. Top. 2020, 229, 2429. [Google Scholar] [CrossRef]
  82. Launey, K.D.; Mercenne, A.; Dytrych, T. Nuclear Dynamics and Reactions in the Ab Initio Symmetry-Adapted Framework. Annu. Rev. Nucl. Part. Sci. 2021, 71, 253. [Google Scholar] [CrossRef]
  83. Heyde, K.; Van Isacker, P.; Casten, R.F.; Wood, J.L. A shell-model interpretation of intruder states and the onset of deformation in even-even nuclei. Phys. Lett. B 1985, 155, 303. [Google Scholar] [CrossRef]
  84. Heyde, K.; De Coster, C.; Ryckebusch, J.; Waroquier, M. Equivalence of the spherical and deformed shell-model approach to intruder states. Phys. Lett. B 1989, 218, 287. [Google Scholar] [CrossRef]
  85. Heyde, K.; Meyer, R.A. Possible evidence for four-particle, four-hole excitations in 146Gd. Phys. Rev. C 1990, 41, 280. [Google Scholar] [CrossRef]
  86. Heyde, K. A shell-model description of intruder states and shape coexistence. Nucl. Phys. A 1990, 507, 149c. [Google Scholar] [CrossRef]
  87. Wenes, G.; Van Isacker, P.; Waroquier, M.; Heyde, K.; Van Maldeghem, J. Collective bands in doubly-even Sn nuclei: Energy spectra and electromagnetic decay properties. Phys. Rev. C 1981, 23, 2291. [Google Scholar] [CrossRef]
  88. Fortune, H.T. Shape coexistence and mixing in 96Zr. Phys. Rev. C 2017, 95, 054313. [Google Scholar] [CrossRef]
  89. Fortune, H.T. Coexistence and B(E2)’s in 98Sr. Nucl. Phys. A 2017, 957, 184–187. [Google Scholar] [CrossRef]
  90. Fortune, H.T. Coexistence and mixing in 76Se. Phys. Rev. C 2019, 99, 054320. [Google Scholar] [CrossRef]
  91. Carchidi, M.; Fortune, H.T. Coexistence in the even zinc isotopes. Phys. Rev. C 1988, 37, 556–570. [Google Scholar] [CrossRef]
  92. Fortune, H.T. Nature of collectivity in Pd isotopes. J. Phys. G Nucl. Phys. 1985, 11, 1305–1308. [Google Scholar] [CrossRef]
  93. Fortune, H.T.; Carchidi, M. Coexistence and B(E2)’s in even Ge nuclei. Phys. Rev. C 1987, 36, 2584–2589. [Google Scholar] [CrossRef]
  94. Fortune, H.T. Coexistence and B(E2) values in 72Ge. Phys. Rev. C 2016, 94, 024318. [Google Scholar] [CrossRef]
  95. Fortune, H.T. Mixing of higher-J states in 72Ge. Phys. Rev. C 2017, 95, 044317. [Google Scholar] [CrossRef]
  96. Fortune, H.T. Nature of first two rotational bands in 152Sm. Nucl. Phys. A 2017, 966, 47–53. [Google Scholar] [CrossRef]
  97. Fortune, H.T. Band mixing and structure of 106,108Pd. Phys. Rev. C 2018, 98, 064303. [Google Scholar] [CrossRef]
  98. Fortune, H.T. Band mixing in 154Gd. Eur. Phys. J. A 2018, 54, 178. [Google Scholar] [CrossRef]
  99. Fortune, H.T. Band mixing in 74,76,78Kr. Eur. Phys. J. A 2018, 54, 229. [Google Scholar] [CrossRef]
  100. Fortune, H.T. Coexistence and mixing in 182,184Hg. Phys. Rev. C 2019, 100, 044303. [Google Scholar] [CrossRef]
  101. Majarshin, A.J.; Luo, Y.-A.; Pan, F.; Fortune, H.T.; Draayer, J.P. Nuclear structure and band mixing in 194Pt. Phys. Rev. C 2021, 103, 024317. [Google Scholar] [CrossRef]
  102. Otsuka, T.; Suzuki, T.; Fujimoto, R.; Grawe, H.; Akaishi, Y. Evolution of Nuclear Shells due to the Tensor Force. Phys. Rev. Lett. 2005, 95, 232502. [Google Scholar] [CrossRef] [PubMed]
  103. Yukawa, H. On the Interaction of Elementary Particles. I. Proc. Phys. Math. Soc. Jpn. 1935, 17, 48. [Google Scholar]
  104. Nilsson, S.G.; Ragnarsson, I. Shapes and Shells in Nuclear Structure; Cambridge U. Press: Cambridge, UK, 1995. [Google Scholar]
  105. Otsuka, T.; Tsunoda, Y. The role of shell evolution in shape coexistence. J. Phys. G Nucl. Part. Phys. 2016, 43, 024009. [Google Scholar] [CrossRef]
  106. Otsuka, T. Exotic nuclei and nuclear forces. Phys. Scr. 2013, T152, 014007. [Google Scholar] [CrossRef]
  107. Otsuka, T.; Gade, A.; Sorlin, O.; Suzuki, T.; Utsuno, Y. Evolution of shell structure in exotic nuclei. Rev. Mod. Phys. 2020, 92, 015002. [Google Scholar] [CrossRef]
  108. Kratzer, A. Die ultraroten Rotationsspektren der Halogenwasserstoffe. Z. Phys. 1920, 3, 289. [Google Scholar] [CrossRef]
  109. Bender, M.; Heenen, P.-H.; Reinhard, P.-G. Self-consistent mean-field models for nuclear structure. Rev. Mod. Phys. 2003, 75, 121. [Google Scholar] [CrossRef]
  110. Greiner, W.; Maruhn, J.A. Nuclear Models; Springer: Berlin/Heidelberg, Germany, 1996. [Google Scholar]
  111. Ring, P.; Schuck, P. The Nuclear Many-Body Problem; Springer: Berlin/Heidelberg, Germany, 1980. [Google Scholar]
  112. Bardeen, J.; Cooper, L.N.; Schrieffer, J.R. Theory of Superconductivity. Phys. Rev. 1957, 108, 1175. [Google Scholar] [CrossRef]
  113. Delaroche, J.-P.; Girod, M.; Libert, J.; Goutte, H.; Hilaire, S.; Péru, S.; Pillet, N.; Bertsch, G.F. Structure of even-even nuclei using a mapped collective Hamiltonian and the D1S Gogny interaction. Phys. Rev. C 2010, 81, 014303. [Google Scholar] [CrossRef]
  114. Gogny, D. Hartree-Fock Bogolyubov method with density-dependent interaction. In Proceedings of the International Conference on Nuclear Physics, Munich, Germany, 27 August–1 September 1973; de Boer, J., Mang, H.J., Eds.; North Holland: Amsterdam, The Netherlands, 1973; p. 48. [Google Scholar]
  115. Gogny, D. Perturbation theory with a soft core two nucleon interaction. In Proceedings of the International Conference on Nuclear Self-Consistent Fields, Trieste, Italy, 24–28 February 1975; Ripka, G., Porneuf, M., Eds.; North Holland: Amsterdam, The Netherlands, 1975; p. 149. [Google Scholar]
  116. Erler, J.; Klüpfel, P.; Reinhard, P.-G. Self-consistent nuclear mean-field models: Example Skyrme–Hartree–Fock. J. Phys. G Nucl. Part. Phys. 2011, 38, 033101. [Google Scholar] [CrossRef]
  117. Skyrme, T.H.R. CVII. The nuclear surface. Phil. Mag. 1956, 1, 1043. [Google Scholar] [CrossRef]
  118. Skyrme, T.H.R. The spin-orbit interaction in nuclei. Nucl. Phys. 1959, 9, 615. [Google Scholar] [CrossRef]
  119. Lalazissis, G.A.; König, J.; Ring, P. New parametrization for the Lagrangian density of relativistic mean field theory. Phys. Rev. C 1997, 55, 540. [Google Scholar] [CrossRef]
  120. Lalazissis, G.A.; Niksić, T.; Vretenar, D.; Ring, P. New relativistic mean-field interaction with density-dependent meson-nucleon couplings. Phys. Rev. C 2005, 71, 024312. [Google Scholar] [CrossRef]
  121. Niksić, T.; Vretenar, D.; Ring, P.; Lalazissis, G.A. Shape coexistence in the relativistic Hartree-Bogoliubov approach. Phys. Rev. C 2002, 65, 054320. [Google Scholar] [CrossRef]
  122. Niksić, T.; Vretenar, D.; Ring, P. Beyond the relativistic mean-field approximation: Configuration mixing of angular-momentum-projected wave functions. Phys. Rev. C 2006, 73, 034308. [Google Scholar] [CrossRef]
  123. Niksić, T.; Vretenar, D.; Ring, P. Beyond the relativistic mean-field approximation. II. Configuration mixing of mean-field wave functions projected on angular momentum and particle number. Phys. Rev. C 2006, 74, 064309. [Google Scholar] [CrossRef]
  124. Niksić, T.; Li, Z.P.; Vretenar, D.; Próchniak, L.; Meng, J.; Ring, P. Beyond the relativistic mean-field approximation. III. Collective Hamiltonian in five dimensions. Phys. Rev. C 2009, 79, 034303. [Google Scholar] [CrossRef]
  125. Ring, P.; Gambhir, Y.K.; Lalazissis, G.A. Computer program for the relativistic mean field description of the ground state properties of even-even axially deformed nuclei. Comp. Phys. Commun. 1997, 105, 77. [Google Scholar] [CrossRef]
  126. Tian, Y.; Ma, Z.Y.; Ring, P. A finite range pairing force for density functional theory in superfluid nuclei. Phys. Lett. B 2009, 676, 44. [Google Scholar] [CrossRef]
  127. Tian, Y.; Ma, Z.-Y.; Ring, P. Axially deformed relativistic Hartree Bogoliubov theory with a separable pairing force. Phys. Rev. C 2009, 80, 024313. [Google Scholar] [CrossRef]
  128. Vretenar, D.; Afanasjev, A.V.; Lalazissis, G.A.; Ring, P. Relativistic Hartree–Bogoliubov theory: Static and dynamic aspects of exotic nuclear structure. Phys. Rep. 2005, 409, 101. [Google Scholar]
  129. Niksić, T.; Paar, N.; Vretenar, D.; Ring, P. DIRHB—A relativistic self-consistent mean-field framework for atomic nuclei. Comp. Phys. Commun. 2014, 185, 1808. [Google Scholar] [CrossRef]
  130. Egido, J.L.; Robledo, L.M.; Rodríguez-Guzmán, R.R. Unveiling the Origin of Shape Coexistence in Lead Isotopes. Phys. Rev. Lett. 2004, 93, 082502. [Google Scholar] [CrossRef]
  131. Egido, J.L.; Jungclaus, A. Predominance of Triaxial Shapes in Transitional Super-Heavy Nuclei: Ground-State Deformation and Shape Coexistence along the Flerovium (Z = 114) Chain of Isotopes. Phys. Rev. Lett. 2020, 125, 192504. [Google Scholar] [CrossRef] [PubMed]
  132. Girod, M.; Delaroche, J.-P.; Görgen, A.; Obertelli, A. The role of triaxiality for the coexistence and evolution of shapes in light krypton isotopes. Phys. Lett. B 2009, 676, 39. [Google Scholar] [CrossRef]
  133. Guo, L.; Maruhn, J.A.; Reinhard, P.-G. Triaxiality and shape coexistence in germanium isotopes. Phys. Rev. C 2007, 76, 034317. [Google Scholar] [CrossRef]
  134. Péru, S.; Martini, M. Mean field based calculations with the Gogny force: Some theoretical tools to explore the nuclear structure. Eur. Phys. J. A 2014, 50, 88. [Google Scholar] [CrossRef]
  135. Robledo, L.M.; Rodríguez-Guzmán, R.R.; Sarriguren, P. Evolution of nuclear shapes in medium mass isotopes from a microscopic perspective. Phys. Rev. C 2008, 78, 034314. [Google Scholar] [CrossRef]
  136. Rodríguez, T.R.; Egido, J.L. A beyond mean field analysis of the shape transition in the Neodymium isotopes. Phys. Lett. B 2008, 663, 49. [Google Scholar] [CrossRef]
  137. Rodríguez, T.R.; Egido, J.L. Study of shape transitions in N∼90 isotopes with beyond mean field calculations. AIP Conf. Proc. 2009, 1090, 419. [Google Scholar]
  138. Rodríguez, T.R.; Egido, J.L. Multiple shape coexistence in the nucleus 80Zr. Phys. Lett. B 2011, 705, 255. [Google Scholar] [CrossRef]
  139. Rodríguez, T.R. Structure of krypton isotopes calculated with symmetry-conserving configuration-mixing methods. Phys. Rev. C 2014, 90, 034306. [Google Scholar] [CrossRef]
  140. Rodríguez-Guzmán, R.R.; Egido, J.L.; Robledo, L.M. Beyond mean field description of shape coexistence in neutron-deficient Pb isotopes. Phys. Rev. C 2004, 69, 054319. [Google Scholar] [CrossRef]
  141. Bender, M.; Bonche, P.; Heenen, P.-H. Shape coexistence in neutron-deficient Kr isotopes: Constraints on the single-particle spectrum of self-consistent mean-field models from collective excitations. Phys. Rev. C 2006, 74, 024312. [Google Scholar] [CrossRef]
  142. Duguet, T.; Bender, M.; Bonche, P.; Heenen, P.-H. Shape coexistence in 186Pb: Beyond-mean-field description by configuration mixing of symmetry restored wave functions. Phys. Lett. B 2003, 559, 201. [Google Scholar] [CrossRef]
  143. Fu, Y.; Tong, H.; Wang, X.F.; Wang, H.; Wang, D.Q.; Wang, X.Y.; Yao, J.M. Microscopic analysis of shape transition in neutron-deficient Yb isotopes. Phys. Rev. C 2018, 97, 014311. [Google Scholar] [CrossRef]
  144. Rodríguez-Guzmán, R.; Sarriguren, P. E(5) and X(5) shape phase transitions within a Skyrme-Hartree-Fock + BCS approach. Phys. Rev. C 2007, 76, 064303. [Google Scholar] [CrossRef]
  145. Sarriguren, P.; Rodríguez-Guzmán, R.; Robledo, L.M. Shape transitions in neutron-rich Yb, Hf, W, Os, and Pt isotopes within a Skyrme Hartree-Fock + BCS approach. Phys. Rev. C 2008, 77, 064322. [Google Scholar] [CrossRef]
  146. Skalski, J.; Heenen, P.-H.; Bonche, P. Shape coexistence and low-lying collective states in A≈100 Zr nuclei. Nucl. Phys. A 1993, 559, 221. [Google Scholar] [CrossRef]
  147. Werner, T.R.; Sheikh, J.A.; Nazarewicz, W.; Strayer, M.R.; Umar, A.S.; Misu, M. Shape coexistence around S 28 16 44 : The deformed N = 28 region. Phys. Lett. B 1994, 335, 259. [Google Scholar] [CrossRef]
  148. Yao, J.M.; Bender, M.; Heenen, P.-H. Systematics of low-lying states of even-even nuclei in the neutron-deficient lead region from a beyond-mean-field calculation. Phys. Rev. C 2013, 87, 034322. [Google Scholar] [CrossRef]
  149. Abusara, H.; Ahmad, S.; Othman, S. Triaxiality softness and shape coexistence in Mo and Ru isotopes. Phys. Rev. C 2017, 95, 054302. [Google Scholar] [CrossRef]
  150. Abusara, H.; Ahmad, S. Shape evolution in Kr, Zr, and Sr isotopic chains in covariant density functional theory. Phys. Rev. C 2017, 96, 064303. [Google Scholar] [CrossRef]
  151. Choi, Y.-B.; Lee, C.-H.; Mun, M.-H.; Kim, Y. Bubble nuclei with shape coexistence in even-even isotopes of Hf to Hg. Phys. Rev. C 2022, 105, 024306. [Google Scholar] [CrossRef]
  152. Heyde, K.; De Coster, C.; Van Duppen, P.; Huyse, M.; Wood, J.L.; Nazarewicz, W. Comment on “Shape and superdeformed structure in Hg isotopes in relativistic mean field model” and “Structure of neutron-deficient Pt, Hg, and Pb isotopes”. Phys. Rev. C 1996, 53, 1035. [Google Scholar] [CrossRef]
  153. Kim, S.; Mun, M.-H.; Cheoun, M.-K.; Ha, E. Shape coexistence and neutron skin thickness of Pb isotopes by the deformed relativistic Hartree-Bogoliubov theory in continuum. Phys. Rev. C 2022, 105, 034340. [Google Scholar] [CrossRef]
  154. Kumar, P.; Dhiman, S.K. Microscopic study of shape evolution and ground state properties in even-even Cd isotopes using covariant density functional theory. Nucl. Phys. A 2020, 1001, 121935. [Google Scholar] [CrossRef]
  155. Kumar, P.; Thakur, V.; Thakur, S.; Kumar, V.; Dhiman, S.K. Nuclear shape evolution and shape coexistence in Zr and Mo isotopes. Eur. Phys. J. A 2021, 57, 36. [Google Scholar] [CrossRef]
  156. Maharana, J.P.; Gambhir, Y.K.; Sheikh, J.A.; Ring, P. Shape coexistence and extreme deformations near A = 80. Phys. Rev. C 1992, 46, R1163. [Google Scholar] [CrossRef] [PubMed]
  157. Meng, J.; Zhang, W.; Zhou, S.G.; Toki, H.; Geng, L.S. Shape evolution for Sm isotopes in relativistic mean-field theory. Eur. Phys. J. A 2005, 25, 23. [Google Scholar] [CrossRef]
  158. Naz, T.; Bhat, G.H.; Jehangir, S.; Ahmad, S.; Sheikh, J.A. Microscopic description of structural evolution in Pd, Xe, Ba, Nd, Sm, Gd and Dy isotopes. Nucl. Phys. A 2018, 979, 1. [Google Scholar] [CrossRef]
  159. Niksić, T.; Vretenar, D.; Lalazissis, G.A.; Ring, P. Microscopic Description of Nuclear Quantum Phase Transitions. Phys. Rev. Lett. 2007, 99, 092502. [Google Scholar] [CrossRef]
  160. Patra, S.K.; Yoshida, S.; Takigawa, N.; Praharaj, C.R. Shape and superdeformed structure in Hg isotopes in relativistic mean field model. Phys. Rev. C 1994, 50, 1924. [Google Scholar] [CrossRef]
  161. Ren, Z. Shape coexistence in even-even superheavy nuclei. Phys. Rev. C 2002, 65, 051304. [Google Scholar] [CrossRef]
  162. Sharma, M.M.; Ring, P. Relativistic mean-field description of neutron-deficient platinum isotopes. Phys. Rev. C 1992, 46, 1715. [Google Scholar] [CrossRef]
  163. Sharma, S.; Devi, R.; Khosa, S.K. Microscopic study of evolution of shape change across even-even mass chain of tellurium isotopes using relativistic Hartree-Bogoliubov model. Nucl. Phys. A 2019, 988, 9. [Google Scholar] [CrossRef]
  164. Sheng, Z.-Q.; Guo, J.-Y. Systematic analysis of critical point nuclei in the rare-earth region with relativistic mean field theory. Mod. Phys. Lett. A 2005, 20, 2711. [Google Scholar] [CrossRef]
  165. Takigawa, N.; Yoshida, S.; Hagino, K.; Patra, S.K.; Praharaj, C.R. Reply to “Comment on `Shape and superdeformed structure in Hg isotopes in relativistic mean field model’ and `Structure of neutron-deficient Pt, Hg, and Pb isotopes”’. Phys. Rev. C 1996, 53, 1038. [Google Scholar] [CrossRef] [PubMed]
  166. Thakur, S.; Kumar, P.; Thakur, V.; Kumar, V.; Dhiman, S.K. Shape transitions and shell structure study in zirconium, molybdenum and ruthenium. Nucl. Phys. A 2021, 1014, 122254. [Google Scholar] [CrossRef]
  167. Thakur, S.; Kumar, P.; Thakur, V.; Kumar, V.; Dhiman, S.K. Nuclear Shape Evolution in Palladium Isotopes. Acta Phys. Pol. B 2021, 52, 1433. [Google Scholar] [CrossRef]
  168. Wang, G.; Fang, X.-Z.; Guo, J.-Y. Analysis of shape evolution for Pt isotopes with relativistic mean field theory. Acta Phys. Sin. 2012, 61, 102101. [Google Scholar] [CrossRef]
  169. Wu, X.Y.; Mei, H.; Yao, J.M.; Zhou, X.-R. Beyond-mean-field study of the hyperon impurity effect in hypernuclei with shape coexistence. Phys. Rev. C 2017, 95, 034309. [Google Scholar] [CrossRef]
  170. Yoshida, S.; Patra, S.K.; Takigawa, N.; Praharaj, C.R. Structure of neutron-deficient Pt, Hg, and Pb isotopes. Phys. Rev. C 1994, 50, 1398. [Google Scholar] [CrossRef]
  171. Yu, M.; Zhang, P.-F.; Ruan, T.-N.; Guo, J.-Y. Shape evolution for Ce isotopes in relativistic mean-field theory. Int. J. Mod. Phys. E 2006, 15, 939. [Google Scholar] [CrossRef]
  172. Dobaczewski, J. Current Developments in Nuclear Density Functional Methods. J. Phys. Conf. Ser. 2011, 312, 092002. [Google Scholar] [CrossRef]
  173. Dobaczewski, J.; Bennaceur, K.; Raimondi, F. Effective theory for low-energy nuclear energy density functionals. J. Phys. G Nucl. Part. Phys. 2012, 39, 125103. [Google Scholar] [CrossRef]
  174. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864. [Google Scholar] [CrossRef]
  175. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133. [Google Scholar] [CrossRef]
  176. Drut, J.E.; Furnstahl, R.J.; Platter, L. Toward ab initio density functional theory for nuclei. Prog. Part. Nucl. Phys. 2010, 64, 120. [Google Scholar] [CrossRef]
  177. Garrett, P.E.; Rodríguez, T.R.; Varela, A.D.; Green, K.L.; Bangay, J.; Finlay, A.; Austin, R.A.E.; Ball, G.C.; Garnsworthy, A.B.; Grinyer, G.F.; et al. Multiple Shape Coexistence in 110,112Cd. Phys. Rev. Lett. 2019, 123, 142502. [Google Scholar] [CrossRef]
  178. Garrett, P.E.; Rodríguez, T.R.; Diaz Varela, A.; Green, K.L.; Bangay, J.; Finlay, A.; Austin, R.A.E.; Ball, G.C.; Bandyopadhyay, D.S.; Bildstein, V.; et al. Shape coexistence and multiparticle-multihole structures in 110,112Cd. Phys. Rev. C 2020, 101, 044302. [Google Scholar] [CrossRef]
  179. Dobaczewski, J.; Skalski, J. The quadrupole vibrational inertial function in the adiabatic time-dependent Hartree-Fock-Bogolyubov approximation. Nucl. Phys. A 1981, 369, 123. [Google Scholar] [CrossRef]
  180. Ebata, S.; Kimura, M. Quenching of N = 28 shell gap and a low-lying quadrupole mode in the vicinity of neutron-rich N = 28 isotones. JPS Conf. Proc. 2015, 6, 030093. [Google Scholar]
  181. Próchniak, L. E(5) and X(5) Dynamical Symmetries from a Microscopic Perspective. Acta Phys. Pol. B 2007, 38, 1605. [Google Scholar]
  182. Petrovici, A.; Schmid, K.W.; Grümmer, F.; Faessler, A. Shape coexistence at high spins in the nuclei 68Ge and 72Se. Nucl. Phys. A 1989, 504, 277. [Google Scholar] [CrossRef]
  183. Petrovici, A.; Schmid, K.W.; Grümmer, F.; Faessler, A. Some new aspects of the shape coexistence in the A = 70 mass region. Nucl. Phys. A 1990, 517, 108. [Google Scholar] [CrossRef]
  184. Petrovici, A.; Hammarén, E.; Schmid, K.W.; Grümmer, F.; Faessler, A. Shape coexistence in the A∼70 region including neutron-proton interaction and unnatural-parity correlations in the mean field. Nucl. Phys. A 1992, 549, 352. [Google Scholar] [CrossRef]
  185. Petrovici, A.; Schmid, K.W.; Faessler, A. Shape coexistence and shape transition in NZ nuclei from krypton to molybdenum. Nucl. Phys. A 1996, 605, 290. [Google Scholar] [CrossRef]
  186. Petrovici, A.; Schmid, K.W.; Faessler, A. Microscopic aspects of shape coexistence in 72Kr and 74Kr. Nucl. Phys. A 2000, 665, 333. [Google Scholar] [CrossRef]
  187. Petrovici, A.; Schmid, K.W.; Faessler, A. Variational approach to shape coexistence in 56Ni. Nucl. Phys. A 2001, 689, 707. [Google Scholar] [CrossRef]
  188. Petrovici, A.; Schmid, K.W.; Faessler, A. Shape coexistence and center-of-mass effects in N=Z medium mass nuclei. Nucl. Phys. A 2002, 708, 190. [Google Scholar] [CrossRef]
  189. Petrovici, A.; Schmid, K.W.; Faessler, A. Variational approach to shape coexistence in 68Se. Nucl. Phys. A 2002, 710, 246. [Google Scholar] [CrossRef]
  190. Petrovici, A. Triple shape coexistence and shape evolution in the N = 58 Sr and Zr isotopes. Phys. Rev. C 2012, 85, 034337. [Google Scholar] [CrossRef]
  191. Petrovici, A. Isospin-symmetry breaking and shape coexistence in A≈70 analogs. Phys. Rev. C 2015, 91, 014302. [Google Scholar] [CrossRef]
  192. Petrovici, A.; Andrei, O. Weak interaction rates and shape coexistence for the Z=N+2 isotopes 70Kr and 74Sr. Phys. Rev. C 2015, 92, 064305. [Google Scholar] [CrossRef]
  193. Petrovici, A.; Andrei, O. Stellar weak interaction rates and shape coexistence for 68Se and 72Kr waiting points. Eur. Phys. J. A 2015, 51, 133. [Google Scholar] [CrossRef]
  194. Petrovici, A. Shape evolution in proton-rich and neutron-rich Kr isotopes within the beyond-mean-field approach. Phys. Scr. 2017, 92, 064003. [Google Scholar] [CrossRef]
  195. Petrovici, A.; Andrei, O.; Chilug, A. Exotic phenomena in medium mass NZ nuclei within the beyond-mean-field approach. Phys. Scr. 2018, 93, 114001. [Google Scholar] [CrossRef]
  196. Petrovici, A.; Mare, A.S. Triple shape coexistence and β decay of 96Y to 96Zr. Phys. Rev. C 2020, 101, 024307. [Google Scholar] [CrossRef]
  197. Schmid, K.W.; Grümmer, F.; Kyotoku, M.; Faessler, A. Selfconsistent description of non-yrast states in nuclei: The excited VAMPIR approach. Nucl. Phys. A 1986, 452, 493. [Google Scholar] [CrossRef]
  198. Matsuyanagi, K.; Matsuo, M.; Nakatsukasa, T.; Yoshida, K.; Hinohara, N.; Sato, K. Microscopic derivation of the quadrupole collective Hamiltonian for shape coexistence/mixing dynamics. J. Phys. G Nucl. Part. Phys. 2016, 43, 024006. [Google Scholar] [CrossRef]
  199. Hinohara, N.; Nakatsukasa, T.; Matsuo, M.; Matsuyanagi, K. Microscopic description of oblate-prolate shape mixing in proton-rich Se isotopes. Phys. Rev. C 2009, 80, 014305. [Google Scholar] [CrossRef]
  200. Hinohara, N.; Sato, K.; Nakatsukasa, T.; Matsuo, M.; Matsuyanagi, K. Microscopic description of large-amplitude shape-mixing dynamics with inertial functions derived in local quasiparticle random-phase approximation. Phys. Rev. C 2010, 82, 064313. [Google Scholar] [CrossRef]
  201. Hinohara, N.; Kanada-En’yo, Y. Triaxial quadrupole deformation dynamics in sd-shell nuclei around 26Mg. Phys. Rev. C 2011, 83, 014321. [Google Scholar] [CrossRef]
  202. Hinohara, N.; Sato, K.; Yoshida, K.; Nakatsukasa, T.; Matsuo, M.; Matsuyanagi, K. Shape fluctuations in the ground and excited 0+ states of 30,32,34 Mg. Phys. Rev. C 2011, 84, 061302. [Google Scholar] [CrossRef]
  203. Sato, K.; Hinohara, N. Shape mixing dynamics in the low-lying states of proton-rich Kr isotopes. Nucl. Phys. A 2011, 849, 53. [Google Scholar] [CrossRef]
  204. Sato, K.; Hinohara, N.; Yoshida, K.; Nakatsukasa, T.; Matsuo, M.; Matsuyanagi, K. Shape transition and fluctuations in neutron-rich Cr isotopes around N = 40. Phys. Rev. C 2012, 86, 024316. [Google Scholar] [CrossRef]
  205. Yoshida, K.; Hinohara, N. Shape changes and large-amplitude collective dynamics in neutron-rich Cr isotopes. Phys. Rev. C 2011, 83, 061302. [Google Scholar] [CrossRef]
  206. Nomura, K.; Shimizu, N.; Otsuka, T. Mean-Field Derivation of the Interacting Boson Model Hamiltonian and Exotic Nuclei. Phys. Rev. Lett. 2008, 101, 142501. [Google Scholar] [CrossRef] [PubMed]
  207. Nomura, K.; Shimizu, N.; Otsuka, T. Formulating the interacting boson model by mean-field methods. Phys. Rev. C 2010, 81, 044307. [Google Scholar] [CrossRef]
  208. Nomura, K.; Otsuka, T.; Shimizu, N.; Guo, L. Microscopic formulation of the interacting boson model for rotational nuclei. Phys. Rev. C 2011, 83, 041302. [Google Scholar] [CrossRef]
  209. Nomura, K.; Otsuka, T.; Rodríguez-Guzmán, R.; Robledo, L.M.; Sarriguren, P.; Regan, P.H.; Stevenson, P.D.; Podolyák, Z. Spectroscopic calculations of the low-lying structure in exotic Os and W isotopes. Phys. Rev. C 2011, 83, 054303. [Google Scholar] [CrossRef]
  210. Nomura, K.; Shimizu, N.; Vretenar, D.; Niksić, T.; Otsuka, T. Robust Regularity in γ-Soft Nuclei and Its Microscopic Realization. Phys. Rev. Lett. 2012, 108, 132501. [Google Scholar] [CrossRef]
  211. Arima, A.; Iachello, F. Collective Nuclear States as Representations of a SU(6) Group. Phys. Rev. Lett. 1975, 35, 1069. [Google Scholar] [CrossRef]
  212. Casten, R.F. (Ed.) Algebraic Approaches to Nuclear Structure: Interacting Boson and Fermion Models; Harwood: Chur, Switzerland, 1993. [Google Scholar]
  213. Frank, A.; Van Isacker, P. Symmetry Methods in Molecules and Nuclei; S y G Editores: Ciudad de Mexico, Mexico, 2005. [Google Scholar]
  214. Iachello, F.; Arima, A. The Interacting Boson Model; Cambridge U. Press: Cambridge, UK, 1987. [Google Scholar]
  215. Iachello, F.; Van Isacker, P. The Interacting Boson-Fermion Model; Cambridge U. Press: Cambridge, UK, 1991. [Google Scholar]
  216. Nomura, K.; Otsuka, T.; Rodríguez-Guzmán, R.; Robledo, L.M.; Sarriguren, P. Structural evolution in Pt isotopes with the interacting boson model Hamiltonian derived from the Gogny energy density functional. Phys. Rev. C 2011, 83, 014309. [Google Scholar] [CrossRef]
  217. Nomura, K.; Niksić, T.; Otsuka, T.; Shimizu, N.; Vretenar, D. Quadrupole collective dynamics from energy density functionals: Collective Hamiltonian and the interacting boson model. Phys. Rev. C 2011, 84, 014302. [Google Scholar] [CrossRef]
  218. Nomura, K.; Otsuka, T.; Rodríguez-Guzmán, R.; Robledo, L.M.; Sarriguren, P. Collective structural evolution in neutron-rich Yb, Hf, W, Os, and Pt isotopes. Phys. Rev. C 2011, 84, 054316. [Google Scholar] [CrossRef]
  219. Nomura, K.; Rodríguez-Guzmán, R.; Robledo, L.M.; Shimizu, N. Shape coexistence in lead isotopes in the interacting boson model with a Gogny energy density functional. Phys. Rev. C 2012, 86, 034322. [Google Scholar] [CrossRef]
  220. Nomura, K.; Rodríguez-Guzmán, R.; Robledo, L.M. Shape evolution and the role of intruder configurations in Hg isotopes within the interacting boson model based on a Gogny energy density functional. Phys. Rev. C 2013, 87, 064313. [Google Scholar] [CrossRef]
  221. Nomura, K.; Rodríguez-Guzmán, R.; Robledo, L.M. Structural evolution in A≈100 nuclei within the mapped interacting boson model based on the Gogny energy density functional. Phys. Rev. C 2016, 94, 044314. [Google Scholar] [CrossRef]
  222. Nomura, K.; Otsuka, T.; Van Isacker, P. Shape coexistence in the microscopically guided interacting boson model. J. Phys. G Nucl. Part. Phys. 2016, 43, 024008. [Google Scholar] [CrossRef]
  223. Nomura, K.; Rodríguez-Guzmán, R.; Robledo, L.M. Structural evolution in germanium and selenium nuclei within the mapped interacting boson model based on the Gogny energy density functional. Phys. Rev. C 2017, 95, 064310. [Google Scholar] [CrossRef]
  224. Nomura, K.; Rodríguez-Guzmán, R.; Humadi, Y.M.; Robledo, L.M.; Abusara, H. Structure of krypton isotopes within the interacting boson model derived from the Gogny energy density functional. Phys. Rev. C 2017, 96, 034310. [Google Scholar] [CrossRef]
  225. Nomura, K.; Jolie, J. Structure of even-even cadmium isotopes from the beyond-mean-field interacting boson model. Phys. Rev. C 2018, 98, 024303. [Google Scholar] [CrossRef]
  226. Nomura, K.; Zhang, Y. Two-neutron transfer reactions and shape phase transitions in the microscopically formulated interacting boson model. Phys. Rev. C 2019, 99, 024324. [Google Scholar] [CrossRef]
  227. Nomura, K.; Rodríguez-Guzmán, R.; Robledo, L. M > β decay of even-A nuclei within the interacting boson model with input based on nuclear density functional theory. Phys. Rev. C 2020, 101, 044318. [Google Scholar] [CrossRef]
  228. Nomura, K.; Vretenar, D.; Li, Z.P.; Xiang, J. Pairing vibrations in the interacting boson model based on density functional theory. Phys. Rev. C 2020, 102, 054313. [Google Scholar] [CrossRef]
  229. Thomas, T.; Nomura, K.; Werner, V.; Ahn, T.; Cooper, N.; Duckwitz, H.; Hinton, M.; Ilie, G.; Jolie, J.; Petkov, P.; et al. Evidence for shape coexistence in 98Mo. Phys. Rev. C 2013, 88, 044305. [Google Scholar] [CrossRef]
  230. Thomas, T.; Werner, V.; Jolie, J.; Nomura, K.; Ahn, T.; Cooper, N.; Duckwitz, H.; Fitzler, A.; Fransen, C.; Gade, A.; et al. Nuclear structure of 96,98Mo: Shape coexistence and mixed-symmetry states. Nucl. Phys. A 2016, 947, 203–233. [Google Scholar] [CrossRef]
  231. Li, Z.P.; Niksić, T.; Vretenar, D.; Meng, J.; Lalazissis, G.A.; Ring, P. Microscopic analysis of nuclear quantum phase transitions in the N≈90 region. Phys. Rev. C 2009, 79, 054301. [Google Scholar] [CrossRef]
  232. Li, Z.P.; Niksić, T.; Vretenar, D.; Meng, J. Microscopic analysis of order parameters in nuclear quantum phase transitions. Phys. Rev. C 2009, 80, 061301. [Google Scholar] [CrossRef]
  233. Li, Z.P.; Niksić, T.; Vretenar, D. Coexistence of nuclear shapes: Self-consistent mean-field and beyond. J. Phys. G Nucl. Part. Phys. 2016, 43, 024005. [Google Scholar] [CrossRef]
  234. Lu, K.Q.; Li, Z.X.; Li, Z.P.; Yao, J.M.; Meng, J. Global study of beyond-mean-field correlation energies in covariant energy density functional theory using a collective Hamiltonian method. Phys. Rev. C 2015, 91, 027304. [Google Scholar] [CrossRef]
  235. Niksić, T.; Vretenar, D.; Ring, P. Relativistic nuclear energy density functionals: Mean-field and beyond. Prog. Part. Nucl. Phys. 2011, 66, 519. [Google Scholar] [CrossRef]
  236. Quan, S.; Chen, Q.; Li, Z.P.; Niksić, T.; Vretenar, D. Global analysis of quadrupole shape invariants based on covariant energy density functionals. Phys. Rev. C 2017, 95, 054321. [Google Scholar] [CrossRef]
  237. Yang, Y.L.; Wang, Y.K.; Zhao, P.W.; Li, Z.P. Nuclear landscape in a mapped collective Hamiltonian from covariant density functional theory. Phys. Rev. C 2021, 104, 054312. [Google Scholar] [CrossRef]
  238. Li, Z.P.; Niksić, T.; Vretenar, D.; Meng, J. Microscopic description of spherical to γ–soft shape transitions in Ba and Xe nuclei. Phys. Rev. C 2010, 81, 034316. [Google Scholar] [CrossRef]
  239. Li, Z.P.; Niksić, T.; Vretenar, D.; Ring, P.; Meng, J. Relativistic energy density functionals: Low-energy collective states of 240Pu and 166Er. Phys. Rev. C 2010, 81, 064321. [Google Scholar] [CrossRef]
  240. Li, Z.P.; Yao, J.M.; Vretenar, D.; Niksić, T.; Chen, H.; Meng, J. Energy density functional analysis of shape evolution in N = 28 isotones. Phys. Rev. C 2011, 84, 054304. [Google Scholar] [CrossRef]
  241. Majola, S.N.T.; Shi, Z.; Song, B.Y.; Li, Z.P.; Zhang, S.Q.; Bark, R.A.; Sharpey-Schafer, J.F.; Aschman, D.G.; Bvumbi, S.P.; Bucher, T.D.; et al. β and γ bands in N = 88, 90, and 92 isotones investigated with a five-dimensional collective Hamiltonian based on covariant density functional theory: Vibrations, shape coexistence, and superdeformation. Phys. Rev. C 2019, 100, 044324. [Google Scholar] [CrossRef]
  242. Xiang, J.; Li, Z.P.; Li, Z.X.; Yao, J.M.; Meng, J. Covariant description of shape evolution and shape coexistence in neutron-rich nuclei at N≈60. Nucl. Phys. A 2012, 873, 1. [Google Scholar] [CrossRef]
  243. Xiang, J.; Li, Z.P.; Long, W.H.; Niksić, T.; Vretenar, D. Shape evolution and coexistence in neutron-deficient Nd and Sm nuclei. Phys. Rev. C 2018, 98, 054308. [Google Scholar] [CrossRef]
  244. Yang, X.Q.; Wang, L.J.; Xiang, J.; Wu, X.Y.; Li, Z.P. Microscopic analysis of prolate-oblate shape phase transition and shape coexistence in the Er-Pt region. Phys. Rev. C 2021, 103, 054321. [Google Scholar] [CrossRef]
  245. Rainwater, J. Nuclear Energy Level Argument for a Spheroidal Nuclear Model. Phys. Rev. 1950, 79, 432. [Google Scholar] [CrossRef]
  246. Bohr, A. The coupling of nuclear surface oscillations to the motion of individual nucleons. Dan. Mat. Fys. Medd. 1952, 26, 14. [Google Scholar]
  247. Bohr, A.; Mottelson, B.R. Collective and individual-particle aspects of nuclear structure. Dan. Mat. Fys. Medd. 1953, 27, 16. [Google Scholar]
  248. Bohr, A.; Mottelson, B.R. Nuclear Structure Vol. I: Single-Particle Motion; World Scientific: Singapore, 1998. [Google Scholar]
  249. Bohr, A.; Mottelson, B.R. Nuclear Structure Vol. II: Nuclear Deformations; World Scientific: Singapore, 1998. [Google Scholar]
  250. Nilsson, S.G. Binding states of individual nucleons in strongly deformed nuclei. Dan. Mat. Fys. Medd. 1955, 29, 16. [Google Scholar]
  251. Ragnarsson, I.; Nilsson, S.G. Shell structure in nuclei. Phys. Rep. 1978, 45, 1. [Google Scholar] [CrossRef]
  252. Lederer, C.M.; Shirley, V.S. (Eds.) Table of Isotopes, 7th ed.; Wiley: New York, NY, USA, 1978. [Google Scholar]
  253. Baranger, M.; Kumar, K. Nuclear deformations in the pairing-plus-quadrupole model: (I). The single-j shell. Nucl. Phys. 1965, 62, 113. [Google Scholar] [CrossRef]
  254. Baranger, M.; Kumar, K. Nuclear deformations in the pairing-plus-quadrupole model: (II). Discussion of validity of the model. Nucl. Phys. A 1968, 110, 490. [Google Scholar] [CrossRef]
  255. Kumar, K.; Baranger, M. Nuclear deformations in the pairing-plus-quadrupole model: (III). Static nuclear shapes in the rare-earth region. Nucl. Phys. A 1968, 110, 529. [Google Scholar] [CrossRef]
  256. Brink, D.M.; Broglia, R.A. Nuclear Superfluidity: Pairing in Finite Systems; Cambridge U. Press: Cambridge, UK, 2005. [Google Scholar]
  257. Shlomo, S.; Talmi, I. Shell-model hamiltonians with generalized seniority eigenstates. Nucl. Phys. A 1972, 198, 81. [Google Scholar] [CrossRef]
  258. Talmi, I. Effective Interactions and Coupling Schemes in Nuclei. Rev. Mod. Phys. 1962, 34, 704. [Google Scholar] [CrossRef]
  259. Talmi, I. Generalized seniority and structure of semi-magic nuclei. Nucl. Phys. A 1971, 172, 1. [Google Scholar] [CrossRef]
  260. Talmi, I. Coupling schemes in nuclei. Riv. Nuovo C. 1973, 3, 85. [Google Scholar] [CrossRef]
  261. Talmi, I. Effective interactions and coupling schemes in nuclei. Nucl. Phys. A 1994, 570, 319c. [Google Scholar] [CrossRef]
  262. Casten, R.F. Possible Unified Interpretation of Heavy Nuclei. Phys. Rev. Lett. 1985, 54, 1991. [Google Scholar] [CrossRef]
  263. Casten, R.F. NpNn systematics in heavy nuclei. Nucl. Phys. A 1985, 443, 1. [Google Scholar] [CrossRef]
  264. Casten, R.F. Nuclei far off stability in the NpNn scheme. Phys. Rev. C 1986, 33, 1819. [Google Scholar] [CrossRef] [PubMed]
  265. Casten, R.F. Nuclear Structure from a Simple Perspective; Oxford University Press: Oxford, UK, 1990. [Google Scholar]
  266. Casten, R.F.; Brenner, D.S.; Haustein, P.E. Valence p-n interactions and the development of collectivity in heavy nuclei. Phys. Rev. Lett. 1987, 58, 658. [Google Scholar] [CrossRef] [PubMed]
  267. Bender, M.; Heenen, P.-H. What can be learned from binding energy differences about nuclear structure: The example of δVpn. Phys. Rev. C 2011, 83, 064319. [Google Scholar] [CrossRef]
  268. Bonatsos, D.; Karampagia, S.; Cakirli, R.B.; Casten, R.F.; Blaum, K.; Amon Susam, L. Emergent collectivity in nuclei and enhanced proton-neutron interactions. Phys. Rev. C 2013, 88, 054309. [Google Scholar] [CrossRef]
  269. Brenner, D.S.; Wesselborg, C.; Casten, R.F.; Warner, D.D.; Zhang, J.-Y. Empirical p-n interactions: Global trends, configuration sensitivity and N = Z enhancements. Phys. Lett. B 1990, 243, 1–6. [Google Scholar] [CrossRef]
  270. Brenner, D.S.; Cakirli, R.B.; Casten, R.F. Valence proton-neutron interactions throughout the mass surface. Phys. Rev. C 2006, 73, 034315. [Google Scholar] [CrossRef]
  271. Cakirli, R.B.; Brenner, D.S.; Casten, R.F.; Millman, E.A. Proton-Neutron Interactions and the New Atomic Masses. Phys. Rev. Lett. 2005, 94, 092501, Erratum in Phys. Rev. Lett. 2005, 95, 119903. [Google Scholar] [CrossRef]
  272. Cakirli, R.B.; Casten, R.F. Direct Empirical Correlation between Proton-Neutron Interaction Strengths and the Growth of Collectivity in Nuclei. Phys. Rev. Lett. 2006, 96, 132501. [Google Scholar] [CrossRef]
  273. Cakirli, R.B.; Blaum, K.; Casten, R.F. Indication of a mini-valence Wigner-like energy in heavy nuclei. Phys. Rev. C 2010, 82, 061304. [Google Scholar] [CrossRef]
  274. Cakirli, R.B.; Casten, R.F. Nuclear binding and nuclear structure. Int. J. Mass Spectrom. 2013, 187, 349–350. [Google Scholar] [CrossRef]
  275. Oktem, Y.; Cakirli, R.B.; Casten, R.F.; Casperson, R.J.; Brenner, D.S. Simple interpretation of proton-neutron interactions in rare earth nuclei. Phys. Rev. C 2006, 74, 027304. [Google Scholar] [CrossRef]
  276. Stoitsov, M.; Cakirli, R.B.; Casten, R.F.; Nazarewicz, W.; Satuła, W. Empirical Proton-Neutron Interactions and Nuclear Density Functional Theory: Global, Regional, and Local Comparisons. Phys. Rev. Lett. 2007, 98, 132502. [Google Scholar] [CrossRef] [PubMed]
  277. Zhang, J.-Y.; Casten, R.F.; Brenner, D.S. Empirical proton-neutron interaction energies. Linearity and saturation phenomena. Phys. Lett. B 1989, 227, 1. [Google Scholar] [CrossRef]
  278. Casten, R.F.; Cakirli, R.B. The evolution of collectivity in nuclei and the proton–neutron interaction. Phys. Scr. 2016, 91, 033004. [Google Scholar] [CrossRef]
  279. Ait Ben Mennana, A.; Benjedi, R.; Budaca, R.; Buganu, P.; El Bassem, Y.; Lahbas, A.; Oulne, M. Mixing of the coexisting shapes in the ground states of 74Ge and 74Kr. Phys. Scr. 2021, 96, 125306. [Google Scholar] [CrossRef]
  280. Ait Ben Mennana, A.; Benjedi, R.; Budaca, R.; Buganu, P.; El Bassem, Y.; Lahbas, A.; Oulne, M. Shape and structure for the low-lying states of the 80Ge nucleus. Phys. Rev. C 2022, 105, 034347. [Google Scholar] [CrossRef]
  281. Budaca, R.; Buganu, P.; Budaca, A.I. Bohr Model Solution for a Shape Coexisting Potential. Bulg. J. Phys. 2017, 44, 319. [Google Scholar]
  282. Budaca, R.; Buganu, P.; Budaca, A.I. Bohr model description of the critical point for the first order shape phase transition. Phys. Lett. B 2018, 776, 26. [Google Scholar] [CrossRef]
  283. Budaca, R.; Buganu, P.; Budaca, A.I. Geometrical model description of shape coexistence in Se isotopes. Nucl. Phys. A 2019, 990, 137. [Google Scholar] [CrossRef]
  284. Budaca, R.; Budaca, A.I.; Buganu, P. Application of the Bohr Hamiltonian with a double-well sextic potential to collective states in Mo isotopes. J. Phys. G Nucl. Part. Phys. 2019, 46, 125102. [Google Scholar] [CrossRef]
  285. Georgoudis, P.E.; Leviatan, A. Aspects of Shape Coexistence in the Geometric Collective Model of Nuclei. J. Phys. Conf. Ser. 2018, 966, 012043. [Google Scholar] [CrossRef]
  286. Budaca, R.; Budaca, A.I. Stepped infinite square well potential for collective excitations in even–even nuclei. Eur. Phys. J. Plus 2021, 136, 983. [Google Scholar] [CrossRef]
  287. Mardyban, E.V.; Kolganova, E.A.; Shneidman, T.M.; Jolos, R.V.; Pietralla, N. Description of the low-lying collective states of 96Zr based on the collective Bohr Hamiltonian including the triaxiality degree of freedom. Phys. Rev. C 2020, 102, 034308. [Google Scholar] [CrossRef]
  288. Mardyban, E.V.; Kolganova, E.A.; Shneidman, T.M.; Jolos, R.V. Evolution of the phenomenologically determined collective potential along the chain of Zr isotopes. Phys. Rev. C 2022, 105, 024321. [Google Scholar] [CrossRef]
  289. Sato, K.; Hinohara, N.; Nakatsukasa, T.; Matsuo, M.; Matsuyanagi, K. A model analysis of triaxial deformation dynamics in oblate-prolate shape coexistence phenomena. Prog. Theor. Phys. 2010, 123, 129–155. [Google Scholar] [CrossRef]
  290. Elliott, J.P. Collective motion in the nuclear shell model. I. Classification schemes for states of mixed configurations. Proc. R. Soc. A Ser. A 1958, 245, 128. [Google Scholar]
  291. Elliott, J.P. Collective motion in the nuclear shell model II. The introduction of intrinsic wave-functions. Proc. R. Soc. A Ser. A 1958, 245, 562. [Google Scholar]
  292. Elliott, J.P.; Harvey, M. Collective motion in the nuclear shell model III. The calculation of spectra. Proc. R. Soc. A Ser. A 1963, 272, 557. [Google Scholar]
  293. Elliott, J.P.; Wilsdon, C.E. Collective motion in the nuclear shell model IV. Odd-mass nuclei in the sd shell. Proc. R. Soc. A Ser. A 1968, 302, 509. [Google Scholar]
  294. Harvey, M. The nuclear SU3 model. Adv. Nucl. Phys. 1968, 1, 67. [Google Scholar]
  295. Gilmore, R. Lie Groups, Lie Algebras, and Some of Their Applications; Wiley: New York, NY, USA, 1974. [Google Scholar]
  296. Kota, V.K.B. SU(3) Symmetry in Atomic Nuclei; Springer Nature: Singapore, 2020. [Google Scholar]
  297. Rowe, D.J.; Wood, J.L. Fundamentals of Nuclear Models: Foundation Models; World Scientific: Singapore, 2010. [Google Scholar]
  298. Bonatsos, D.; Klein, A. Exact boson mappings for nuclear neutron (proton) shell-model algebras having SU(3) subalgebras. Ann. Phys. 1986, 169, 61. [Google Scholar] [CrossRef]
  299. Arima, A.; Harvey, M.; Shimizu, K. Pseudo LS coupling and pseudo SU3 coupling schemes. Phys. Lett. B 1969, 30, 517. [Google Scholar] [CrossRef]
  300. Bahri, C.; Draayer, J.P.; Moszkowski, S.A. Pseudospin symmetry in nuclear physics. Phys. Rev. Lett. 1992, 68, 2133. [Google Scholar] [CrossRef]
  301. Draayer, J.P.; Weeks, K.J.; Hecht, K.T. Strength of the Qπ·Qν interaction and the strong-coupled pseudo-SU(3) limit. Nucl. Phys. A 1982, 381, 1. [Google Scholar] [CrossRef]
  302. Draayer, J.P.; Weeks, K.J. Shell-Model Description of the Low-Energy Structure of Strongly Deformed Nuclei. Phys. Rev. Lett. 1983, 51, 1422. [Google Scholar] [CrossRef]
  303. Draayer, J.P.; Weeks, K.J. Towards a shell model description of the low-energy structure of deformed nuclei I. Even-even systems. Ann. Phys. 1984, 156, 41. [Google Scholar] [CrossRef]
  304. Ginocchio, J.N. Pseudospin as a Relativistic Symmetry. Phys. Rev. Lett. 1997, 78, 436. [Google Scholar] [CrossRef]
  305. Hecht, K.T.; Adler, A. Generalized seniority for favored J≠0 pairs in mixed configurations. Nucl. Phys. A 1969, 137, 129. [Google Scholar] [CrossRef]
  306. Ratna Raju, R.D.; Draayer, J.P.; Hecht, K.T. Search for a coupling scheme in heavy deformed nuclei: The pseudo SU(3) model. Nucl. Phys. A 1973, 202, 433. [Google Scholar] [CrossRef]
  307. Castaños, O.; Moshinsky, M.; Quesne, C. Transformation from U(3) to pseudo U(3) basis. In Group Theory and Special Symmetries in Nuclear Physics (AnnArbor, 1991); Draayer, J.P., Jänecke, J., Eds.; World Scientific: Singapore, 1992; p. 80. [Google Scholar]
  308. Castaños, O.; Moshinsky, M.; Quesne, C. Transformation to pseudo-SU(3) in heavy deformed nuclei. Phys. Lett. B 1992, 277, 238. [Google Scholar] [CrossRef]
  309. Castaños, O.; Velázquez, A.V.; Hess, P.O.; Hirsch, J.G. Transformation to pseudo-spin-symmetry of a deformed Nilsson hamiltonian. Phys. Lett. B 1994, 321, 303. [Google Scholar] [CrossRef]
  310. Bonatsos, D.; Assimakis, I.E.; Martinou, A. Proton-Neutron Pairs in Heavy Deformed Nuclei. Bulg. J. Phys. 2015, 42, 439. [Google Scholar]
  311. Martinou, A.; Assimakis, I.E.; Minkov, N.; Bonatsos, D. Emergence of SU(3) symmetry in heavy deformed nuclei. Nucl. Theory 2016, 35, 224. [Google Scholar]
  312. Castaños, O.; Draayer, J.P.; Leschber, L. Towards a shell-model description of the low-energy structure of deformed nuclei II. Electromagnetic properties of collective M1 bands. Ann. Phys. 1987, 180, 290. [Google Scholar] [CrossRef]
  313. Vargas, C.E.; Hirsch, J.G.; Draayer, J.P. Pseudo SU(3) shell model: Normal parity bands in odd-mass nuclei. Nucl. Phys. A 2000, 673, 219. [Google Scholar] [CrossRef]
  314. Weeks, K.J.; Draayer, J.P. Shell-model predictions for unique parity yrast configurations of odd-mass deformed nuclei. Nucl. Phys. A 1983, 393, 69. [Google Scholar] [CrossRef]
  315. Troltenier, D.; Bahri, C.; Draayer, J.P. Generalized pseudo-SU(3) model and pairing. Nucl. Phys. A 1995, 586, 53. [Google Scholar] [CrossRef]
  316. Troltenier, D.; Bahri, C.; Draayer, J.P. Effects of pairing in the pseudo-SU(3) model. Nucl. Phys. A 1995, 589, 75. [Google Scholar] [CrossRef]
  317. Castaños, O.; Hirsch, J.G.; Civitarese, O.; Hess, P.O. Double-beta decay in the pseudo SU(3) scheme. Nucl. Phys. A 1994, 571, 276. [Google Scholar] [CrossRef]
  318. Hirsch, J.G.; Castaños, O.; Hess, P.O. Neutrinoless double beta decay in heavy deformed nuclei. Nucl. Phys. A 1995, 582, 124. [Google Scholar] [CrossRef]
  319. Zuker, A.P.; Retamosa, J.; Poves, A.; Caurier, E. Spherical shell model description of rotational motion. Phys. Rev. C 1995, 52, R1741. [Google Scholar] [CrossRef]
  320. Zuker, A.P.; Poves, A.; Nowacki, F.; Lenzi, S.M. Nilsson-SU3 self-consistency in heavy N = Z nuclei. Phys. Rev. C 2015, 92, 024320. [Google Scholar] [CrossRef]
  321. Kaneko, K.; Shimizu, N.; Mizusaki, T.; Sun, Y. Quasi-SU(3) coupling of (1h11/2, 2f7/2) across the N = 82 shell gap: Enhanced E2 collectivity and shape evolution in Nd isotopes. Phys. Rev. C 2021, 103, L021301. [Google Scholar] [CrossRef]
  322. Kaneko, K.; Sun, Y.; Shimizu, N.; Mizusaki, T. Quasi-SU(3) Coupling Induced Oblate-Prolate Shape Phase Transition in the Casten Triangle. Phys. Rev. Lett. 2023, 130, 052501. [Google Scholar] [CrossRef]
  323. Bonatsos, D.; Assimakis, I.E.; Minkov, N.; Martinou, A.; Cakirli, R.B.; Casten, R.F.; Blaum, K. Proxy-SU(3) symmetry in heavy deformed nuclei. Phys. Rev. C 2017, 95, 064325. [Google Scholar] [CrossRef]
  324. Bonatsos, D.; Assimakis, I.E.; Minkov, N.; Martinou, A.; Sarantopoulou, S.; Cakirli, R.B.; Casten, R.F.; Blaum, K. Analytic predictions for nuclear shapes, prolate dominance, and the prolate-oblate shape transition in the proxy-SU(3) model. Phys. Rev. C 2017, 95, 064326. [Google Scholar] [CrossRef]
  325. Bonatsos, D. Prolate over oblate dominance in deformed nuclei as a consequence of the SU(3) symmetry and the Pauli principle. Eur. Phys. J. A 2017, 53, 148. [Google Scholar] [CrossRef]
  326. Bonatsos, D.; Martinou, A.; Sarantopoulou, S.; Assimakis, I.E.; Peroulis, S.; Minkov, N. Parameter-free predictions for the collective deformation variables β and γ within the pseudo-SU(3) scheme. Eur. Phys. J. Spec. Top. 2020, 229, 2367. [Google Scholar] [CrossRef]
  327. Bonatsos, D.; Martinou, A.; Peroulis, S.K.; Mertzimekis, T.J.; Minkov, N. The Proxy-SU(3) Symmetry in Atomic Nuclei. Symmetry 2023, 15, 169. [Google Scholar] [CrossRef]
  328. Martinou, A.; Bonatsos, D.; Karakatsanis, K.E.; Sarantopoulou, S.; Assimakis, I.E.; Peroulis, S.K.; Minkov, N. Why nuclear forces favor the highest weight irreducible representations of the fermionic SU(3) symmetry. Eur. Phys. J. A 2021, 57, 83. [Google Scholar] [CrossRef]
  329. Martinou, A.; Bonatsos, D.; Minkov, N.; Assimakis, I.E.; Peroulis, S.K.; Sarantopoulou, S.; Cseh, J. Proxy-SU(3) symmetry in the shell model basis. Eur. Phys. J. A 2020, 56, 239. [Google Scholar] [CrossRef]
  330. Bonatsos, D.; Sobhani, H.; Hassanabadi, H. Shell model structure of proxy-SU(3) pairs of orbitals. Eur. Phys. J. Plus 2020, 135, 710. [Google Scholar] [CrossRef]
  331. de-Shalit, A.; Goldhaber, M. Mixed Configurations in Nuclei. Phys. Rev. 1953, 92, 1211. [Google Scholar] [CrossRef]
  332. Castaños, O.; Draayer, J.P.; Leschber, Y. Shape variables and the shell model. Z. Phys. A 1988, 329, 33. [Google Scholar]
  333. Draayer, J.P.; Park, S.C.; Castaños, O. Shell-Model Interpretation of the Collective-Model Potential-Energy Surface. Phys. Rev. Lett. 1989, 62, 20. [Google Scholar] [CrossRef]
  334. Martinou, A.; Bonatsos, D.; Metzimekis, T.J.; Karakatsanis, K.E.; Assimakis, I.E.; Peroulis, S.K.; Sarantopoulou, S.; Minkov, N. The islands of shape coexistence within the Elliott and the proxy-SU(3) Models. Eur. Phys. J. A 2021, 57, 84. [Google Scholar] [CrossRef]
  335. Martinou, A.; Bonatsos, D.; Peroulis, S.K.; Karakatsanis, K.E.; Mertzimekis, T.J.; Minkov, N. Islands of Shape Coexistence: Theoretical Predictions and Experimental Evidence. Symmetry 2023, 15, 29. [Google Scholar] [CrossRef]
  336. Rosensteel, G.; Rowe, D.J. Nuclear Sp(3,R) Model. Phys. Rev. Lett. 1977, 38, 10. [Google Scholar] [CrossRef]
  337. Rosensteel, G.; Rowe, D.J. On the algebraic formulation of collective models III. The symplectic shell model of collective motion. Ann. Phys. 1980, 126, 343. [Google Scholar] [CrossRef]
  338. Rowe, D.J. Microscopic theory of the nuclear collective model. Rep. Prog. Phys. 1985, 48, 1419. [Google Scholar] [CrossRef]
  339. Dytrych, T.; Sviratcheva, K.D.; Bahri, C.; Draayer, J.P.; Vary, J.P. Evidence for Symplectic Symmetry in Ab Initio No-Core Shell Model Results for Light Nuclei. Phys. Rev. Lett. 2007, 98, 162503. [Google Scholar] [CrossRef] [PubMed]
  340. Dytrych, T.; Sviratcheva, K.D.; Bahri, C.; Draayer, J.P.; Vary, J.P. Dominant role of symplectic symmetry in ab initio no-core shell model results for light nuclei. Phys. Rev. C 2007, 76, 014315. [Google Scholar] [CrossRef]
  341. Dytrych, T.; Launey, K.D.; Draayer, J.P.; Maris, P.; Vary, J.P.; Saule, E.; Catalyurek, U.; Sosonkina, M.; Langr, D.; Caprio, M.A. Collective Modes in Light Nuclei from First Principles. Phys. Rev. Lett. 2013, 111, 252501. [Google Scholar] [CrossRef] [PubMed]
  342. Dreyfuss, A.C.; Launey, K.D.; Dytrych, T.; Draayer, J.P.; Bahri, C. Hoyle state and rotational features in Carbon-12 within a no-core shell-model framework. Phys. Lett. B 2013, 727, 511. [Google Scholar] [CrossRef]
  343. Dreyfuss, A.C.; Launey, K.D.; Dytrych, T.; Draayer, J.P.; Baker, R.B.; Deibel, C.M.; Bahri, C. Understanding emergent collectivity and clustering in nuclei from a symmetry-based no-core shell-model perspective. Phys. Rev. C 2017, 95, 044312. [Google Scholar] [CrossRef]
  344. Dytrych, T.; Sviratcheva, K.D.; Bahri, C.; Draayer, J.P.; Vary, J.P. Highly deformed modes in the ab initio symplectic no-core shell model. J. Phys. G Nucl. Part. Phys. 2008, 35, 095101. [Google Scholar] [CrossRef]
  345. Tobin, G.K.; Ferriss, M.C.; Launey, K.D.; Dytrych, T.; Draayer, J.P.; Dreyfuss, A.C.; Bahri, C. Symplectic no-core shell-model approach to intermediate-mass nuclei. Phys. Rev. C 2014, 89, 034312. [Google Scholar] [CrossRef]
  346. Dytrych, T.; Launey, K.D.; Draayer, J.P.; Rowe, D.J.; Wood, J.L.; Rosensteel, G.; Bahri, C.; Langr, D.; Baker, R.B. Physics of Nuclei: Key Role of an Emergent Symmetry. Phys. Rev. Lett. 2020, 124, 042501. [Google Scholar] [CrossRef]
  347. Casten, R.F.; McCutchan, E.A. Quantum phase transitions and structural evolution in nuclei. J. Phys. G Nucl. Part. Phys. 2007, 34, R285. [Google Scholar] [CrossRef]
  348. Casten, R.F. Quantum phase transitions and structural evolution in nuclei. Prog. Part. Nucl. Phys. 2009, 62, 183. [Google Scholar] [CrossRef]
  349. 0 Database. Available online: https://www.nndc.bnl.gov/nudat3 (accessed on 31 January 2023).
  350. Brenner, D.S. Are There X(5) Nuclei In The A∼80 and A∼100 Regions? AIP Conf. Proc. 2002, 638, 223. [Google Scholar]
  351. Hutter, C.; Krücken, R.; Aprahamian, A.; Barton, C.J.; Beausang, C.W.; Caprio, M.A.; Casten, R.F.; Chou, W.-T.; Clark, R.M.; Cline, D.; et al. B(E2) values and the search for the critical point symmetry X(5) in 104Mo and 106Mo. Phys. Rev. C 2003, 67, 054315. [Google Scholar] [CrossRef]
  352. Regan, P.H.; Yamamoto, A.D.; Beausang, C.W.; Zamfir, N.V.; Casten, R.F.; Zhang, J.-Y.; Caprio, M.A.; Gúrdal, G.; Hecht, A.A.; Hutter, C.; et al. The Highs And Lows Of The A=100 Region: Vibration-To-Rotation Evolution In Mo And Ru Isotopes. AIP Conf. Proc. 2003, 656, 422. [Google Scholar]
  353. Bizzeti, P.G.; Bizzeti-Sona, A.M. Evidence of X(5) symmetry for nγ=0, 1, 2 bands in 104Mo. Phys. Rev. C 2002, 66, 031301. [Google Scholar] [CrossRef]
  354. Bizzeti, P.G.; Bizzeti-Sona, A.M.; Tonev, D.; Ur, C.A.; Dewald, A.; Giannatiempo, A.; Melon, B.; Bazzacco, D.; Costin, A.; de Angelis, G.; et al. Transition probabilities in the X(5) candidate 122Ba. Phys. Rev. C 2010, 82, 054311. [Google Scholar] [CrossRef]
  355. Fransen, C.; Pietralla, N.; Linnemann, A.; Werner, V.; Bijker, R. Low-spin γ-ray spectroscopy of the (critical-point?) nucleus 122Ba. Phys. Rev. C 2004, 69, 014313. [Google Scholar] [CrossRef]
  356. Clark, R.M.; Cromaz, M.; Deleplanque, M.A.; Descovich, M.; Diamond, R.M.; Fallon, P.; Firestone, R.B.; Lee, I.Y.; Macchiavelli, A.O.; Mahmud, H.; et al. Searching for X(5) behavior in nuclei. Phys. Rev. C 2003, 68, 037301. [Google Scholar] [CrossRef]
  357. Balabanski, D.L.; Gladnishki, K.A.; Lo Bianco, G.; Saltarelli, A.; Zamfir, N.V.; McCutchan, E.A.; Ai, H.; Casten, R.F.; Heinz, A.; Meyer, D.A.; et al. Evidence for X(5) critical point symmetry in 128Ce. Int. J. Mod. Phys. E 2006, 15, 1735. [Google Scholar] [CrossRef]
  358. Casten, R.F.; Zamfir, N.V.; Krücken, R. Comment on “Reexamination of the N = 90 transitional nuclei 150Nd and 152Sm”. Phys. Rev. C 2003, 68, 059801. [Google Scholar] [CrossRef]
  359. Mertz, A.F.; McCutchan, E.A.; Casten, R.F.; Casperson, R.J.; Heinz, A.; Huber, B.; Lüttke, R.; Qian, J.; Shoraka, B.; Terry, J.R.; et al. First experimental test of X(5) critical-point symmetry in the A∼130 mass region: Low-spin states and the collective structure of 130Ce. Phys. Rev. C 2008, 77, 014307. [Google Scholar] [CrossRef]
  360. Clark, R.M.; Cromaz, M.; Deleplanque, M.A.; Diamond, R.M.; Fallon, P.; Görgen, A.; Lee, I.Y.; Macchiavelli, A.O.; Stephens, F.S.; Ward, D. Reexamination of the N = 90 transitional nuclei 150Nd and 152Sm. Phys. Rev. C 2003, 67, 041302. [Google Scholar] [CrossRef]
  361. Krücken, R.; Albanna, B.; Bialik, C.; Casten, R.F.; Cooper, J.R.; Dewald, A.; Zamfir, N.V.; Barton, C.J.; Beausang, C.W.; Caprio, M.A.; et al. B(E2) values in 150Nd and the critical point symmetry X(5). Phys. Rev. Lett. 2002, 88, 232501. [Google Scholar] [CrossRef]
  362. Zhang, D.-L.; Zhao, H.-Y. Empirical example of nucleus with transitional dynamical symmetry X(5). Chin. Phys. Lett. 2002, 19, 779. [Google Scholar]
  363. Börner, H.G.; Mutti, P.; Jentschel, M.; Zamfir, N.V.; Casten, R.F.; McCutchan, E.A.; Krücken, R. Low-energy phonon structure of 150Sm. Phys. Rev. C 2006, 73, 034314. [Google Scholar] [CrossRef]
  364. Bijker, R.; Casten, R.F.; Zamfir, N.V.; McCutchan, E.A. Test of X(5) for the γ degree of freedom. Phys. Rev. C 2003, 68, 064304, Erratum in Phys. Rev. C 2004, 69, 059901. [Google Scholar] [CrossRef]
  365. Casten, R.F.; Wilhelm, M.; Radermacher, E.; Zamfir, N.V.; von Brentano, P. The first excited 0+ state in 152Sm. Phys. Rev. C 1998, 57, R1553. [Google Scholar] [CrossRef]
  366. Casten, R.F.; Zamfir, N.V. Empirical Realization of a Critical Point Description in Atomic Nuclei. Phys. Rev. Lett. 2001, 87, 052503. [Google Scholar] [CrossRef]
  367. Jolie, J.; Cejnar, P.; Dobes, J. Phase coexistence in the interacting boson model and 152Sm. Phys. Rev. C 1999, 60, 061303. [Google Scholar] [CrossRef]
  368. Klug, T.; Dewald, A.; Werner, V.; von Brentano, P.; Casten, R.F. The B(E2: 4 2 + 2 2 + ) value in 152Sm and β-softness in phase coexisting structures. Phys. Lett. B 2000, 495, 55. [Google Scholar] [CrossRef]
  369. Kulp, W.D.; Wood, J.L.; Allmond, J.M.; Eimer, J.; Furse, D.; Krane, K.S.; Loats, J.; Schmelzenbach, P.; Stapels, C.J.; Larimer, R.-M.; et al. N = 90 region: The decays of 152Eum,g to 152Sm. Phys. Rev. C 2007, 76, 034319. [Google Scholar] [CrossRef]
  370. Kulp, W.D.; Wood, J.L.; Garrett, P.E.; Wu, C.Y.; Cline, D.; Allmond, J.M.; Bandyopadhyay, D.; Dashdorj, D.; Choudry, S.N.; Hayes, A.B.; et al. Search for intrinsic collective excitations in 152Sm. Phys. Rev. C 2008, 77, 061301. [Google Scholar] [CrossRef]
  371. Zamfir, N.V.; Casten, R.F.; Caprio, M.A.; Beausang, C.W.; Krücken, R.; Novak, J.R.; Cooper, J.R.; Cata-Danil, G.; Barton, C.J. B(E2) values and phase coexistence in 152Sm. Phys. Rev. C 1999, 60, 054312. [Google Scholar] [CrossRef]
  372. Zamfir, N.V.; Börner, H.G.; Pietralla, N.; Casten, R.F.; Berant, Z.; Barton, C.J.; Beausang, C.W.; Brenner, D.S.; Caprio, M.A.; Cooper, J.R.; et al. Lifetime and B(E2) values for the 3 1 + level in 152Sm. Phys. Rev. C 2002, 65, 067305. [Google Scholar] [CrossRef]
  373. Zhang, J.-Y.; Caprio, M.A.; Zamfir, N.V.; Casten, R.F. Phase/shape coexistence in 152Sm in the geometric collective model. Phys. Rev. C 1999, 60, 061304. [Google Scholar] [CrossRef]
  374. Adam, J.; Dobes, J.; Honusek, M.; Kalinnikov, V.G.; Mrázek, J.; Pronskikh, V.S.; Caloun, P.; Lebedev, N.A.; Stegailov, V.I.; Tsoupko-Sitnikov, V.M. Properties of 152Gd collective states. Eur. Phys. J. A 2003, 18, 605. [Google Scholar] [CrossRef]
  375. Dewald, A.; Möller, O.; Tonev, D.; Fitzler, A.; Saha, B.; Jessen, K.; Heinze, S.; Linnemann, A.; Jolie, J.; Zell, K.O.; et al. Shape changes and test of the critical-point symmetry X(5) in N = 90 nuclei. Eur. Phys. J. A 2003, 20, 173. [Google Scholar] [CrossRef]
  376. Tonev, D.; Dewald, A.; Klug, T.; Petkov, P.; Jolie, J.; Fitzler, A.; Möller, O.; Heinze, S.; von Brentano, P.; Casten, R.F. Transition probabilities in 154Gd: Evidence for X(5) critical point symmetry. Phys. Rev. C 2004, 69, 034334. [Google Scholar] [CrossRef]
  377. Caprio, M.A.; Zamfir, N.V.; Casten, R.F.; Barton, C.J.; Beausang, C.W.; Cooper, J.R.; Hecht, A.A.; Krücken, R.; Newman, H.; Novak, J.R.; et al. Low-spin structure of 156Dy through γ-ray spectroscopy. Phys. Rev. C 2002, 66, 054310. [Google Scholar] [CrossRef]
  378. Möller, O.; Dewald, A.; Petkov, P.; Saha, B.; Fitzler, A.; Jessen, K.; Tonev, D.; Klug, T.; Heinze, S.; Jolie, J.; et al. Electromagnetic transition strengths in 156Dy. Phys. Rev. C 2006, 74, 024313. [Google Scholar] [CrossRef]
  379. McCutchan, E.A.; Zamfir, N.V.; Casten, R.F.; Ai, H.; Amro, H.; Babilon, M.; Brenner, D.S.; Gürdal, G.; Heinz, A.; Hughes, R.O.; et al. Lifetime measurements of yrast states in 162Yb and 166Hf. Phys. Rev. C 2006, 73, 034303. [Google Scholar] [CrossRef]
  380. McCutchan, E.A.; Zamfir, N.V.; Casten, R.F.; Caprio, M.A.; Ai, H.; Amro, H.; Beausang, C.W.; Hecht, A.A.; Meyer, D.A.; Ressler, J.J. γ-ray spectroscopy of 166Hf: X(5) in N>90? Phys. Rev. C 2005, 71, 024309. [Google Scholar] [CrossRef]
  381. McCutchan, E.A.; Casten, R.F.; Werner, V.; Wolf, A.; Berant, Z.; Casperson, R.J.; Heinz, A.; Lüttke, R.; Shoraka, B.; Terry, J.R.; et al. β decay study of 168Hf and a test of new geometrical models. Phys. Rev. C 2007, 76, 064307. [Google Scholar] [CrossRef]
  382. Costin, A.; Ahn, T.; Bochev, B.; Dusling, K.; Li, T.C.; Pietralla, N.; Rainovski, G.; Rother, W. Lifetime measurement for the 2 1 + state of 170Hf. Phys. Rev. C 2006, 74, 067301. [Google Scholar] [CrossRef]
  383. Dewald, A.; Möller, O.; Saha, B.; Jessen, K.; Fitzler, A.; Melon, B.; Pissulla, T.; Heinze, S.; Jolie, J.; Zell, K.O.; et al. Test of the critical point symmetry X(5) in the mass A = 180 region. J. Phys. G Nucl. Part. Phys. 2005, 31, S1427. [Google Scholar] [CrossRef]
  384. Aoi, N.; Suzuki, H.; Takeshita, E.; Takeuchi, S.; Ota, S.; Baba, H.; Bishop, S.; Fukui, T.; Hashimoto, Y.; Ong, H.J.; et al. Shape transition observed in neutron-rich pf-shell isotopes studied via proton inelastic scattering. Nucl. Phys. A 2008, 805, 400c. [Google Scholar] [CrossRef]
  385. Gadea, A. First results of the CLARA-PRISMA setup installed at LNL. Eur. Phys. J. A 2005, 25, 421. [Google Scholar] [CrossRef]
  386. Gadea, A. The CLARA-PRISMA setup installed at LNL: First results. J. Phys. G Nucl. Part. Phys. 2005, 31, S1443. [Google Scholar]
  387. Mărginean, N.; Lenzi, S.M.; Gadea, A.; Farnea, E.; Freeman, S.J.; Napoli, D.R.; Bazzacco, D.; Beghini, S.; Behera, B.R.; Bizzeti, P.G.; et al. Shape transitions far from stability: The nucleus 58Cr. Phys. Lett. B 2006, 633, 696. [Google Scholar] [CrossRef]
  388. Bettermann, L.; Régis, J.-M.; Materna, T.; Jolie, J.; Köster, U.; Moschner, K.; Radeck, D. Lifetime measurement of excited states in the shape-phase-transitional nucleus 98Zr. Phys. Rev. C 2010, 82, 044310. [Google Scholar] [CrossRef]
  389. Frank, A.; Alonso, C.E.; Arias, J.M. Search for E(5) symmetry in nuclei: The Ru isotopes. Phys. Rev. C 2001, 65, 014301. [Google Scholar] [CrossRef]
  390. Radeck, D.; Blazhev, A.; Albers, M.; Bernards, C.; Dewald, A.; Fransen, C.; Heidemann, M.; Jolie, J.; Melon, B.; Mücher, D.; et al. First measurement of lifetimes in the yrast band of 100Pd. Phys. Rev. C 2009, 80, 044331. [Google Scholar] [CrossRef]
  391. Clark, R.M.; Cromaz, M.; Deleplanque, M.A.; Descovich, M.; Diamond, R.M.; Fallon, P.; Lee, I.Y.; Macchiavelli, A.O.; Mahmud, H.; Rodriguez-Vieitez, R.; et al. Searching for E(5) behavior in nuclei. Phys. Rev. C 2004, 69, 064322. [Google Scholar] [CrossRef]
  392. Konstantinopoulos, T.; Ashley, S.F.; Axiotis, M.; Spyrou, A.; Harissopulos, S.; Dewald, A.; Litzinger, J.; Möller, O.; Müller-Gatterman, C.; Petkov, P.; et al. Lifetime measurements in 102Pd: Searching for empirical proof of the E(5) critical-point symmetry in nuclear structure. Phys. Rev. C 2016, 93, 014320. [Google Scholar] [CrossRef]
  393. Zhang, D.-L.; Liu, Y.-X. Empirical example of possible E(5) symmetry nucleus 108Pd. Phys. Rev. C 2002, 65, 057301. [Google Scholar] [CrossRef]
  394. Zhang, D.-L.; Wang, H.-L. An IBM2 Description of the E(5) Symmetry in 134Ba and 108Pd. Commun. Theor. Phys. 2002, 38, 71. [Google Scholar]
  395. Kirson, M.W. Comment on “Searching for E(5) behavior in nuclei”. Phys. Rev. C 2004, 70, 049801. [Google Scholar] [CrossRef]
  396. Zhang, J.-F.; Long, G.-L.; Sun, Y.; Zhu, S.-J.; Liu, F.-Y.; Jia, Y. Description of 114Cd in the E(5) Symmetry. Chin. Phys. Lett. 2003, 20, 1231. [Google Scholar]
  397. Ghiţă, D.G.; Căta-Danil, G.; Bucurescu, D.; Căta-Danil, I.; Ivascu, M.; Mihai, C.; Suliman, G.; Stroe, L.; Sava, T.; Zamfir, N.V. 124Te and the E(5) critical point symmetry. Int. J. Mod. Phys. A 2008, 17, 1453. [Google Scholar] [CrossRef]
  398. Hicks, S.F.; Vanhoy, J.R.; Burkett, P.G.; Champine, B.R.; Etzkorn, S.J.; Garrett, P.E.; Yates, S.W.; Yeh, M. Lifetimes in 124Te: Examining critical-point symmetry in the Te nuclei. Phys. Rev. C 2017, 95, 034322. [Google Scholar] [CrossRef]
  399. Coquard, L.; Pietralla, N.; Ahn, T.; Rainovski, G.; Bettermann, L.; Carpenter, M.P.; Janssens, R.V.F.; Leske, J.; Lister, C.J.; Möller, O.; et al. Robust test of E(5) symmetry in 128Xe. Phys. Rev. C 2009, 80, 061304. [Google Scholar] [CrossRef]
  400. Peters, E.E.; Ross, T.J.; Ashley, S.F.; Chakraborty, A.; Crider, B.P.; Hennek, M.D.; Liu, S.H.; McEllistrem, M.T.; Mukhopadhyay, S.; Prados-Estévez, F.M.; et al. 0+ states in 130,132Xe: A search for E(5) behavior. Phys. Rev. C 2016, 94, 024313. [Google Scholar] [CrossRef]
  401. Zhang, D.-L.; Liu, Y.-X. Evidence for a Possible E(5) Symmetry in 130Xe. Chin. Phys. Lett. 2003, 20, 1028–1030. [Google Scholar]
  402. Pascu, S.; Căta-Danil, G.; Bucurescu, D.; Mărginean, N.; Müller, C.; Zamfir, N.V.; Graw, G.; Gollwitzer, A.; Hofer, D.; Valnion, B.D. Structure investigation with the (p, t) reaction on 132,134Ba nuclei. Phys. Rev. C 2010, 81, 014304. [Google Scholar] [CrossRef]
  403. Arias, J.M. E2 transitions and quadrupole moments in the E(5) symmetry. Phys. Rev. C 2001, 63, 034308. [Google Scholar] [CrossRef]
  404. Casten, R.F.; Zamfir, N.V. Evidence for a Possible E(5) Symmetry in 134Ba. Phys. Rev. Lett. 2000, 85, 3584. [Google Scholar] [CrossRef]
  405. Luo, Y.-A.; Zhang, X.-B.; Pan, F.; Ning, P.-Z.; Draayer, J.P. Magnetic excitations in the nucleon-pair shell model. Phys. Rev. C 2001, 64, 047302. [Google Scholar] [CrossRef]
  406. ENSDF Database. Available online: https://www.nndc.bnl.gov/ensdf (accessed on 31 January 2023).
  407. Bonatsos, D.; Martinou, A.; Peroulis, S.; Mertzimekis, T.; Minkov, N. Signatures for shape coexistence and shape/phase transitions in even-even nuclei. J. Phys. G Nucl. Part. Phys. 2023, 50, 075105. [Google Scholar] [CrossRef]
  408. Arima, A.; Iachello, F. Interacting boson model of collective states I. The vibrational limit. Ann. Phys. 1976, 99, 253. [Google Scholar] [CrossRef]
  409. Arima, A.; Iachello, F. Interacting boson model of collective nuclear states II. The rotational limit. Ann. Phys. 1978, 111, 201. [Google Scholar] [CrossRef]
  410. Arima, A.; Iachello, F. Interacting boson model of collective nuclear states IV. The O(6) limit. Ann. Phys. 1979, 123, 468. [Google Scholar] [CrossRef]
  411. Feng, D.H.; Gilmore, R.; Deans, S.R. Phase transitions and the geometric properties of the interacting boson model. Phys. Rev. C 1981, 23, 1254. [Google Scholar] [CrossRef]
  412. Iachello, F.; Zamfir, N.V.; Casten, R.F. Phase Coexistence in Transitional Nuclei and the Interacting-Boson Model. Phys. Rev. Lett. 1998, 81, 1191. [Google Scholar] [CrossRef]
  413. Alhassid, Y.; Leviatan, A. Partial dynamical symmetry. J. Phys. A Math. Gen. 1992, 25, L1265–L1271. [Google Scholar] [CrossRef]
  414. Leviatan, A. Partial dynamical symmetries. Prog. Part. Nucl. Phys. 2011, 66, 93–143. [Google Scholar] [CrossRef]
  415. Leviatan, A.; Shapira, D. Algebraic benchmark for prolate-oblate coexistence in nuclei. Phys. Rev. C 2016, 93, 051302. [Google Scholar] [CrossRef]
  416. Leviatan, A.; Gavrielov, N. Partial dynamical symmetries and shape coexistence in nuclei. Phys. Scr. 2017, 92, 114005. [Google Scholar] [CrossRef]
  417. Engel, J.; Iachello, F. Quantization of Asymmetric Shapes in Nuclei. Phys. Rev. Lett. 1985, 54, 1126. [Google Scholar] [CrossRef]
  418. Engel, J.; Iachello, F. Interacting boson model of collective octupole states: (I). The rotational limit. Nucl. Phys. A 1987, 472, 61. [Google Scholar] [CrossRef]
  419. Kusnezov, D.; Iachello, F. A study of collective octupole states in barium in the interacting boson model. Phys. Lett. B 1988, 209, 420. [Google Scholar] [CrossRef]
  420. Kusnezov, D.F.; Henry, E.A.; Meyer, R.A. An octupole two-phonon multiplet in 96Zr and its description within the spdf interacting boson model. Phys. Lett. B 1989, 228, 11. [Google Scholar] [CrossRef]
  421. Zamfir, N.V.; Kusnezov, D. Octupole correlations in the transitional actinides and the spdf interacting boson model. Phys. Rev. C 2001, 63, 054306. [Google Scholar] [CrossRef]
  422. Zamfir, N.V.; Kusnezov, D. Octupole correlations in U and Pu nuclei. Phys. Rev. C 2003, 67, 014305. [Google Scholar] [CrossRef]
  423. De Coster, C.; Heyde, K.; Decroix, B.; Van Isacker, P.; Jolie, J.; Lehmann, H.; Wood, J.L. Particle-hole excitations in the interacting boson model (I) General structure and symmetries. Nucl. Phys. A 1996, 600, 251. [Google Scholar] [CrossRef]
  424. De Coster, C.; Decroix, B.; Heyde, K.; Wood, J.L.; Jolie, J.; Lehmann, H. Particle-hole excitations in the interacting boson model (III): The O(6)-SU(3) coupling. Nucl. Phys. A 1997, 621, 802. [Google Scholar] [CrossRef]
  425. De Coster, C.; Decroix, B.; Heyde, K.; Jolie, J.; Lehmann, H.; Wood, J.L. Particle-hole excitations in the interacting boson model (IV). The U(5)-SU(3) coupling. Nucl. Phys. A 1999, 651, 31. [Google Scholar] [CrossRef]
  426. De Coster, C.; De Beule, J.; Decroix, B.; Heyde, K.; Oros, A.M.; Van Isacker, P. An algebraic approach to shape coexistence. J. Phys. G Nucl. Part. Phys. 1999, 25, 671. [Google Scholar] [CrossRef]
  427. Decroix, B.; De Beule, J.; De Coster, C.; Heyde, K. SU(4) symmetry in the extended proton-neutron interacting boson model: Multiplets and symmetry breaking. Phys. Lett. B 1998, 439, 237. [Google Scholar] [CrossRef]
  428. Decroix, B.; De Beule, J.; De Coster, C.; Heyde, K.; Oros, A.M.; Van Isacker, P. Particle-hole excitations in the interacting boson models IBM-2 and IBM-3. Phys. Rev. C 1998, 57, 2329. [Google Scholar] [CrossRef]
  429. Decroix, B.; De Coster, C.; Heyde, K.; Oros, A.M.; De Beule, J.; Van Isacker, P. The extended proton-neutron interacting boson model and possible applications. J. Phys. G Nucl. Part. Phys. 1999, 25, 855. [Google Scholar] [CrossRef]
  430. Lehmann, H.; Jolie, J.; De Coster, C.; Decroix, B.; Heyde, K.; Wood, J.L. Particle-hole excitations in the interacting boson model (II): The U(5)-O(6) coupling. Nucl. Phys. A 1997, 621, 767. [Google Scholar] [CrossRef]
  431. Duval, P.D.; Barrett, B.R. Configuration mixing in the interacting boson model. Phys. Lett. B 1981, 100, 223. [Google Scholar] [CrossRef]
  432. Duval, P.D.; Barrett, B.R. Quantitative description of configuration mixing in the interacting boson model. Nucl. Phys. A 1982, 376, 213. [Google Scholar] [CrossRef]
  433. Frank, A.; Van Isacker, P.; Vargas, C.E. Evolving shape coexistence in the lead isotopes: The geometry of configuration mixing in nuclei. Phys. Rev. C 2004, 69, 034323. [Google Scholar] [CrossRef]
  434. García-Ramos, J.E.; Hellemans, V.; Heyde, K. Platinum nuclei: Concealed configuration mixing and shape coexistence. Phys. Rev. C 2011, 84, 014331. [Google Scholar] [CrossRef]
  435. García-Ramos, J.E.; Heyde, K. Nuclear shape coexistence: A study of the even-even Hg isotopes using the interacting boson model with configuration mixing. Phys. Rev. C 2014, 89, 014306. [Google Scholar] [CrossRef]
  436. García-Ramos, J.E.; Heyde, K.; Robledo, L.M.; Rodríguez-Guzmán, R. Shape evolution and shape coexistence in Pt isotopes: Comparing interacting boson model configuration mixing and Gogny mean-field energy surfaces. Phys. Rev. C 2014, 89, 034313. [Google Scholar] [CrossRef]
  437. García-Ramos, J.E.; Heyde, K. Nuclear shape coexistence in Po isotopes: An interacting boson model study. Phys. Rev. C 2015, 92, 034309. [Google Scholar] [CrossRef]
  438. García-Ramos, J.E.; Heyde, K. Quest of shape coexistence in Zr isotopes. Phys. Rev. C 2019, 100, 044315. [Google Scholar] [CrossRef]
  439. García-Ramos, J.E.; Heyde, K. Subtle connection between shape coexistence and quantum phase transition: The Zr case. Phys. Rev. C 2020, 102, 054333. [Google Scholar] [CrossRef]
  440. Gavrielov, N.; Leviatan, A.; Iachello, F. Intertwined quantum phase transitions in the Zr isotopes. Phys. Rev. C 2019, 99, 064324. [Google Scholar] [CrossRef]
  441. Gavrielov, N.; Leviatan, A.; Iachello, F. Interplay between shape-phase transitions and shape coexistence in the Zr isotopes. Phys. Scr. 2020, 95, 024001. [Google Scholar] [CrossRef]
  442. Gavrielov, N.; Leviatan, A.; Iachello, F. Zr isotopes as a region of intertwined quantum phase transitions. Phys. Rev. C 2022, 105, 014305. [Google Scholar] [CrossRef]
  443. Harder, M.K.; Tang, K.T.; Van Isacker, P. An IBM description of coexistence in the platinum isotopes. Phys. Lett. B 1997, 405, 25. [Google Scholar] [CrossRef]
  444. Hellemans, V.; De Baerdemacker, S.; Heyde, K. Configuration mixing in the neutron-deficient 186-196Pb isotopes. Phys. Rev. C 2008, 77, 064324. [Google Scholar] [CrossRef]
  445. Maya-Barbecho, E.; García-Ramos, J.E. Shape coexistence in Sr isotopes. Phys. Rev. C 2022, 105, 034341. [Google Scholar] [CrossRef]
  446. Morales, I.O.; Frank, A.; Vargas, C.E.; Van Isacker, P. Shape coexistence and phase transitions in the platinum isotopes. Phys. Rev. C 2008, 78, 024303. [Google Scholar] [CrossRef]
  447. Casten, R.F. Shape phase transitions and critical-point phenomena in atomic nuclei. Nat. Phys. 2006, 2, 811. [Google Scholar] [CrossRef]
  448. Cejnar, P.; Jolie, J. Quantum phase transitions in the interacting boson model. Prog. Part. Nucl. Phys. 2009, 62, 210. [Google Scholar] [CrossRef]
  449. Cejnar, P.; Jolie, J.; Casten, R.F. Quantum phase transitions in the shapes of atomic nuclei. Rev. Mod. Phys. 2010, 82, 2155. [Google Scholar] [CrossRef]
  450. Iachello, F. Dynamic Symmetries at the Critical Point. Phys. Rev. Lett. 2000, 85, 3580. [Google Scholar] [CrossRef]
  451. Iachello, F. Analytic Description of Critical Point Nuclei in a Spherical-Axially Deformed Shape Phase Transition. Phys. Rev. Lett. 2001, 87, 052502. [Google Scholar] [CrossRef]
  452. Scholten, O.; Iachello, F.; Arima, A. Interacting boson model of collective nuclear states III. The transition from SU(5) to SU(3). Ann. Phys. 1978, 115, 325. [Google Scholar] [CrossRef]
  453. Jolie, J.; Casten, R.F.; von Brentano, P.; Werner, V. Quantum Phase Transition for γ-Soft Nuclei. Phys. Rev. Lett. 2001, 87, 162501. [Google Scholar] [CrossRef] [PubMed]
  454. Jolie, J.; Cejnar, P.; Casten, R.F.; Heinze, S.; Linnemann, A.; Werner, V. Triple Point of Nuclear Deformations. Phys. Rev. Lett. 2002, 89, 182502. [Google Scholar] [CrossRef] [PubMed]
  455. Warner, D. A triple point in nuclei. Nature 2002, 420, 614. [Google Scholar] [CrossRef]
  456. Zhang, J.Y.; Casten, R.F.; Zamfir, N.V. A structural triangle for the geometric collective model. Phys. Lett. B 1997, 407, 201. [Google Scholar] [CrossRef]
  457. Iachello, F. Phase Transitions in Angle Variables. Phys. Rev. Lett. 2003, 91, 132502. [Google Scholar] [CrossRef] [PubMed]
  458. Bonatsos, D.; Lenis, D.; Petrellis, D.; Terziev, P.A. Z(5): Critical point symmetry for the prolate to oblate nuclear shape phase transition. Phys. Lett. B 2004, 588, 172. [Google Scholar] [CrossRef]
  459. Bonatsos, D.; Lenis, D.; Petrellis, D.; Terziev, P.A.; Yigitoglu, I. X(3): An exactly separable γ-rigid version of the X(5) critical point symmetry. Phys. Lett. B 2006, 632, 238–242. [Google Scholar] [CrossRef]
  460. Bonatsos, D.; Lenis, D.; Petrellis, D.; Terziev, P.A.; Yigitoglu, I. γ-rigid solution of the Bohr Hamiltonian for γ=30o compared to the E(5) critical point symmetry. Phys. Lett. B 2005, 621, 102–108. [Google Scholar] [CrossRef]
  461. Budaca, R.; Budaca, A.I. Emergence of Euclidean dynamical symmetry as a consequence of shape phase mixing. Phys. Lett. B 2016, 759, 349–353. [Google Scholar] [CrossRef]
  462. Zhang, Y.; Pan, F.; Liu, Y.-X.; Luo, Y.-A.; Draayer, J.P. γ-rigid solution of the Bohr Hamiltonian for the critical point description of the spherical to γ-rigidly deformed shape phase transition. Phys. Rev. C 2017, 96, 034323. [Google Scholar] [CrossRef]
  463. Zhang, Y.; Liu, Y.-X.; Pan, F.; Sun, Y.; Draayer, J.P. Euclidean dynamical symmetry in nuclear shape phase transitions. Phys. Lett. B 2014, 732, 55–58. [Google Scholar] [CrossRef]
  464. Bonatsos, D.; Lenis, D.; Petrellis, D. Special solutions of the Bohr hamiltonian related to shape phase transitions in nuclei. Rom. Rep. Phys. 2007, 59, 273. [Google Scholar]
  465. Rosensteel, G.; Rowe, D.J. Phase transitions and quasi-dynamical symmetry in nuclear collective models, III: The U(5) to SU(3) phase transition in the IBM. Nucl. Phys. A 2005, 759, 92. [Google Scholar] [CrossRef]
  466. Rowe, D.J. Quasidynamical Symmetry in an Interacting Boson Model Phase Transition. Phys. Rev. Lett. 2004, 93, 122502. [Google Scholar] [CrossRef]
  467. Rowe, D.J. Phase transitions and quasidynamical symmetry in nuclear collective models: I. The U(5) to O(6) phase transition in the IBM. Nucl. Phys. A 2004, 745, 47. [Google Scholar] [CrossRef]
  468. Turner, P.S.; Rowe, D.J. Phase transitions and quasidynamical symmetry in nuclear collective models. II. The spherical vibrator to gamma-soft rotor transition in an SO(5)-invariant Bohr model. Nucl. Phys. A 2005, 756, 333–355. [Google Scholar] [CrossRef]
  469. Heyde, K.; Jolie, J.; Fossion, R.; De Baerdemacker, S.; Hellemans, V. Phase transitions versus shape coexistence. Phys. Rev. C 2004, 69, 054304. [Google Scholar] [CrossRef]
  470. Federman, P.; Pittel, S. Towards a unified microscopic description of nuclear deformation. Phys. Lett. B 1977, 69, 385. [Google Scholar] [CrossRef]
  471. Federman, P.; Pittel, S. Hartree-Fock-Bogolyubov study of deformation in the Zr-Mo region. Phys. Lett. B 1978, 77, 29. [Google Scholar] [CrossRef]
  472. Federman, P.; Pittel, S.; Campos, R. Microscopic study of the shape transition in the zirconium isotopes. Phys. Lett. B 1979, 82, 9. [Google Scholar] [CrossRef]
  473. Federman, P.; Pittel, S. Unified shell-model description of nuclear deformation. Phys. Rev. C 1979, 20, 820. [Google Scholar] [CrossRef]
  474. Hammad, M.M. On the conformable fractional E(5) critical point symmetry. Nucl. Phys. A 2021, 1011, 122203. [Google Scholar] [CrossRef]
  475. Herrmann, R. Fractional Calculus—An Introduction for Physicists; World Scientific: Singapore, 2011. [Google Scholar]
  476. Miller, K.S.; Ross, B. An Introduction to the Fractional Calculus and Fractional Differential Equations; Wiley: New York, NY, USA, 1993. [Google Scholar]
  477. Podlubny, I. Fractional Differential Equations; Academic Press: San Diego, CA, USA, 1999. [Google Scholar]
  478. Khalil, R.; Al Horani, M.; Yousef, A.; Sababheh, M. A new definition of fractional derivative. J. Comput. Appl. Math. 2014, 264, 65–70. [Google Scholar] [CrossRef]
  479. Hammad, M.M. Conformable fractional Bohr Hamiltonian with Bonatsos and double-well sextic potentials. Phys. Scr. 2021, 96, 115304. [Google Scholar] [CrossRef]
  480. Hammad, M.M.; Yaqut, A.S.H.; Abdel-Khalek, M.A.; Doma, S.B. Analytical study of conformable fractional Bohr Hamiltonian with Kratzer potential. Nucl. Phys. A 2021, 1015, 122307. [Google Scholar] [CrossRef]
  481. Hammad, M.M.; Yahia, M.M.; Bonatsos, D. Triaxial nuclei and analytical solutions of the conformable fractional Bohr Hamiltonian with some exponential-type potentials. Nucl. Phys. A 2023, 1030, 122576. [Google Scholar] [CrossRef]
  482. Leviatan, A.; Novoselsky, A.; Talmi, I. O(5) symmetry in IBA-1—The O(6)—U(5) transition region. Phys. Lett. B 1986, 172, 144. [Google Scholar]
  483. Jolie, J.; Linnemann, A. Prolate-oblate phase transition in the Hf-Hg mass region. Phys. Rev. C 2003, 68, 031301. [Google Scholar] [CrossRef]
  484. Jolie, J.; Lehmann, H. On the influence of the O(5) symmetry on shape coexistence in atomic nuclei. Phys. Lett. B 1995, 342, 1. [Google Scholar]
  485. Lehmann, H.; Jolie, J. The U(5)-O(6) model: An analytical approach to shape coexistence. Nucl. Phys. A 1995, 588, 623. [Google Scholar] [CrossRef]
  486. Nakatsukasa, T.; Walet, N.R. Collective coordinates, shape transitions, and shape coexistence: A microscopic approach. Phys. Rev. C 1998, 58, 3397. [Google Scholar]
  487. Kaup, U.; Barrett, B.R. Nuclear shape coexistence in a schematic model. Phys. Rev. C 1990, 42, 981. [Google Scholar] [CrossRef]
  488. Ginocchio, J.N. On a generalization of quasispin to monopole and quadrupole pairing. Phys. Lett. B 1979, 85, 9–13. [Google Scholar] [CrossRef]
  489. Ginocchio, J.N. A schematic model for monopole and quadrupole pairing in nuclei. Ann. Phys. 1980, 126, 234–276. [Google Scholar] [CrossRef]
  490. Garrett, P.E. Characterization of the β vibration and 0 2 + states in deformed nuclei. J. Phys. G Nucl. Part. Phys. 2001, 27, R1. [Google Scholar]
  491. Zganjar, E.F.; Wood, J.L. Electric monopole transitions and shape coexistence in nuclei. Nucl. Phys. A 1990, 520, 427c. [Google Scholar] [CrossRef]
  492. Aprahamian, A. What is the nature of Kπ=0+ bands in deformed nuclei? AIP Conf. Proc. 2002, 638, 77. [Google Scholar]
  493. Aprahamian, A.; Lesher, S.R.; Stratman, A. First Excited 0+ States in Deformed Nuclei. Bulg. J. Phys. 2017, 44, 372. [Google Scholar]
  494. Lesher, S.R.; Aprahamian, A.; Trache, L.; Oros-Peusquens, O.; Deyliz, S.; Gollwitzer, A.; Hertenberger, R.; Valnion, B.D.; Graw, G. New 0+ states in 158Gd. Phys. Rev. C 2002, 66, 051305. [Google Scholar] [CrossRef]
  495. Lesher, S.R.; Orce, J.N.; Ammar, Z.; Hannant, C.D.; Merrick, M.; Warr, N.; Brown, T.B.; Boukharouba, N.; Fransen, C.; Scheck, M.; et al. Study of 0+ excitations in 158Gd with the (n,n’γ) reaction. Phys. Rev. C 2007, 76, 034318. [Google Scholar] [CrossRef]
  496. Zamfir, N.V.; Zhang, J.-Y.; Casten, R.F. Interpreting recent measurements of 0+ states in 158Gd. Phys. Rev. C 2002, 66, 057303. [Google Scholar] [CrossRef]
  497. Hirsch, J.G.; Popa, G.; Lesher, S.R.; Aprahamian, A.; Vargas, C.E.; Draayer, J.P. Low energy 0+ excitations in 158Gd. Rev. Mex. Fis. 2006, S52, 69. [Google Scholar]
  498. Popa, G.; Hirsch, J.G.; Draayer, J.P. Shell model description of normal parity bands in even-even heavy deformed nuclei. Phys. Rev. C 2000, 62, 064313. [Google Scholar] [CrossRef]
  499. Popa, G.; Georgieva, A. Description of heavy deformed nuclei within the pseudo-SU(3) shell model. J. Phys. G Conf. Ser. 2012, 403, 012009. [Google Scholar] [CrossRef]
  500. Lo Iudice, N.; Sushkov, A.V.; Shirikova, N.Y. Microscopic structure of low-lying 0+ states in the deformed 158Gd. Phys. Rev. C 2004, 70, 064316. [Google Scholar] [CrossRef]
  501. Gerçceklioğlu, M. Excited 0+ states in 156,158,160,162Gd. Ann. Phys. 2005, 14, 312. [Google Scholar] [CrossRef]
  502. Asai, M.; Sekine, T.; Osa, A.; Koizumi, M.; Kojima, Y.; Shibata, M.; Yamamoto, H.; Kawade, K. Energy systematics of low-lying 0+ states in neutron-deficient Ba nuclei. Phys. Rev. C 1997, 56, 3045. [Google Scholar] [CrossRef]
  503. Balabanski, D.L.; Lo Bianco, G.; Atanasova, L.; Blasi, N.; Das Gupta, S.; Detistov, P.; Gladnishki, K.; Fortunato, L.; Kusoglu, A.; Nardelli, S.; et al. E0 decay and lifetimes of 0 2 + states in the rare-earth region: The case of 156Dy and 160Er. In Proceedings of the 30th International Workshop on Nuclear Theory, Rila Mountains, Bulgaria, 27 June–1 July 2011; Georgieva, A., Minkov, N., Eds.; Heron Press: Sofia, Bulgaria, 2011. [Google Scholar]
  504. Bucurescu, D.; Graw, G.; Hertenberger, R.; Wirth, H.-F.; Lo Iudice, N.; Sushkov, A.V.; Shirikova, N.Y.; Sun, Y.; Faestermann, T.; Krücken, R.; et al. High-resolution study of 0+ and 2+ excitations in 168Er with the (p, t) reaction. Phys. Rev. C 2006, 73, 064309. [Google Scholar] [CrossRef]
  505. Meyer, D.A.; Wood, V.; Casten, R.F.; Fitzpatrick, C.R.; Graw, G.; Bucurescu, D.; Jolie, J.; von Brentano, P.; Hertenberger, R.; Wirth, H.-F.; et al. Enhanced density of low-lying 0+ states: A corroboration of shape phase transitional behavior. Phys. Lett. B 2006, 638, 44. [Google Scholar] [CrossRef]
  506. Meyer, D.A.; Wood, V.; Casten, R.F.; Fitzpatrick, C.R.; Graw, G.; Bucurescu, D.; Jolie, J.; von Brentano, P.; Hertenberger, R.; Wirth, H.-F.; et al. Extensive investigation of 0+ states in rare earth region nuclei. Phys. Rev. C 2006, 74, 044309. [Google Scholar] [CrossRef]
  507. Suliman, G.; Bucurescu, D.; Hertenberger, R.; Wirth, H.-F.; Faestermann, T.; Krücken, R.; Behrens, T.; Bildstein, V.; Eppinger, K.; Hinke, C.; et al. Study of the 130Ba nucleus with the (p, t) reaction. Eur. Phys. J. A 2008, 36, 243. [Google Scholar] [CrossRef]
  508. Popa, G.; Georgieva, A.; Draayer, J.P. Systematics in the structure of low-lying, nonyrast bandhead configurations of strongly deformed nuclei. Phys. Rev. C 2004, 69, 064307. [Google Scholar] [CrossRef]
  509. Popa, G.; Aprahamian, A.; Georgieva, A.; Draayer, J.P. Systematics in the structure of low-lying, non-yrast band-head configurations of strongly deformed nuclei. Eur. Phys. J. A 2005, 25, s01, 451. [Google Scholar] [CrossRef]
  510. Gerçeklioğlu, M. Transfer strengths to the 0+ states excited by (p, t) reactions in 130,132,134Ba. Phys. Rev. C 2010, 82, 024306. [Google Scholar] [CrossRef]
  511. Gerçeklioğlu, M. The 0+ states excited by (p, t) reaction in 170Yb. Eur. Phys. J. A 2012, 48, 141. [Google Scholar] [CrossRef]
  512. Cakirli, R.B.; Casten, R.F. Anomalies in the Behavior of the First Excited K = 0 Band in Deformed Nuclei. Bulg. J. Phys. 2017, 44, 380. [Google Scholar]
  513. Aprahamian, A.; Brenner, D.S.; Casten, R.F.; Gill, R.L.; Piotrowski, A.; Heyde, K. Observation of 0+ states in 118Cd and the systematics of intruder states. Phys. Lett. B 1984, 140, 22. [Google Scholar] [CrossRef]
  514. Garrett, P.E.; Wood, J.L. On the robustness of surface vibrational modes: Case studies in the Cd region. J. Phys. G Nucl. Part. Phys. 2010, 37, 064028, Erratum in J. Phys. G Nucl. Part. Phys. 2010, 37, 069701. [Google Scholar] [CrossRef]
  515. Garrett, P.E.; Green, K.L.; Austin, R.A.E.; Ball, G.C.; Byopadhyay, D.S.; Colosimo, S.; Cross, D.S.; Demand, G.A.; Grinyer, G.F.; Hackman, G. Using β-decay to Map the E2 Strength in the Cd Isotopes and the Downfall of Vibrational Motion. Acta Phys. Pol. B 2011, 42, 799. [Google Scholar] [CrossRef]
  516. Garrett, P.E. Shape coexistence at low spin in the Z = 50 region and its spectroscopic signatures. J. Phys. G Nucl. Part. Phys. 2016, 43, 084002. [Google Scholar] [CrossRef]
  517. Garrett, P.E.; Wood, J.L.; Yates, S.W. Critical insights into nuclear collectivity from complementary nuclear spectroscopic methods. Phys. Scr. 2018, 93, 063001. [Google Scholar] [CrossRef]
  518. Burke, D.G. Search for experimental evidence supporting the multiphonon description of excited states in 152Sm. Phys. Rev. C 2002, 66, 024312. [Google Scholar] [CrossRef]
  519. Wirth, H.-F.; Graw, G.; Christen, S.; Cutoiu, D.; Eisermann, Y.; Günther, C.; Hertenberger, R.; Jolie, J.; Levon, A.I.; Möller, O.; et al. 0+ states in deformed actinide nuclei by the (p, t) reaction. Phys. Rev. C 2004, 69, 044310. [Google Scholar] [CrossRef]
  520. Lo Iudice, N.; Sushkov, A.V.; Shirikova, N.Y. Microscopic structure of low-lying 0+ states in deformed nuclei. Phys. Rev. C 2005, 72, 034303. [Google Scholar] [CrossRef]
  521. Sharpey-Schafer, J.F.; Mullins, S.M.; Bark, R.A.; Gueorguieva, E.; Kau, J.; Komati, F.; Lawrie, J.J.; Maine, P.; Minkova, A.; Murray, S.H.T.; et al. Shape Transitional Nuclei: What can we learn from the Yrare States? or Hello the Double Vacuum; Goodbye β-vibrations! AIP Conf. Proc. 2008, 1012, 19. [Google Scholar]
  522. Sharpey-Schafer, J.F.; Bark, R.A.; Bvumbi, S.P.; Lawrie, E.A.; Lawrie, J.J.; Madiba, T.E.; Majola, S.N.T.; Minkova, A.; Mullins, S.M.; Papka, P.; et al. A double vacuum, configuration dependent pairing and lack of β-vibrations in transitional nuclei: Band structure of N = 88 to N = 91 Nuclei. Nucl. Phys. A 2010, 834, 45c. [Google Scholar] [CrossRef]
  523. Sharpey-Schafer, J.F.; Mullins, S.M.; Bark, R.A.; Kau, J.; Komati, F.; Lawrie, E.A.; Lawrie, J.J.; Madiba, T.E.; Maine, P.; Minkova, A.; et al. Congruent band structures in 154Gd: Configuration-dependent pairing, a double vacuum and lack of β-vibrations. Eur. Phys. J. A 2011, 47, 5. [Google Scholar] [CrossRef]
  524. Sharpey-Schafer, J.F.; Madiba, T.E.; Bvumbi, S.P.; Lawrie, E.A.; Lawrie, J.J.; Minkova, A.; Mullins, S.M.; Papka, P.; Roux, D.G.; Timár, J. Blocking of coupling to the 0 2 + excitation in 154Gd by the [505]11/2- neutron in 155Gd. Eur. Phys. J. A 2011, 47, 6. [Google Scholar] [CrossRef]
  525. Sharpey-Schafer, J.F.; Bark, R.A.; Bvumbi, S.P.; Dinoko, T.R.S.; Majola, S.N.T. “Stiff” deformed nuclei, configuration dependent pairing and the β and γ degrees of freedom. Eur. Phys. J. A 2019, 55, 15. [Google Scholar] [CrossRef]
  526. Kibédi, T.; Spear, R.H. Electric monopole transitions between 0+ states for nuclei throughout the periodic table. At. Data Nucl. Data Tables 2005, 89, 77. [Google Scholar] [CrossRef]
  527. Wood, J.L.; Zganjar, E.F.; De Coster, C.; Heyde, K. Electric monopole transitions from low energy excitations in nuclei. Nucl. Phys. A 1999, 651, 323. [Google Scholar] [CrossRef]
  528. Zganjar, E.F. Conversion electron spectroscopy and its role in identifying shape coexisting structures in nuclei via E0 transitions. J. Phys. G Nucl. Part. Phys. 2016, 43, 024013. [Google Scholar] [CrossRef]
  529. Andreyev, A.N.; Huyse, M.; Van Duppen, P.; Weissman, L.; Ackermann, D.; Gerl, J.; Hessberger, F.P.; Hofmann, S.; Kleinböhl, A.; Münzenberg, G.; et al. A triplet of differently shaped spin-zero states in the atomic nucleus 186Pb. Nature 2000, 405, 430. [Google Scholar] [CrossRef] [PubMed]
  530. Le Coz, Y.; Becker, F.; Kankaanpää, H.; Korten, W.; Mergel, E.; Butler, P.A.; Cocks, J.F.C.; Dorvaux, O.; Hawcroft, D.; Helariutta, K.; et al. Evidence of multiple shape-coexistence in 188Pb. EPJdirect 1999, A3, 1. [Google Scholar]
  531. Cruz, S.; Bender, P.C.; Krucken, R.; Wimmer, K.; Ames, F.; Andreoiu, C.; Austin, R.A.E.; Bancroft, C.S.; Braid, R.; Bruhn, T.; et al. Shape coexistence and mixing of low-lying 0+ states in 96Sr. Phys. Lett. B 2018, 786, 94. [Google Scholar] [CrossRef]
  532. Little, D.; Ayangeakaa, A.D.; Janssens, R.V.F.; Zhu, S.; Tsunoda, Y.; Otsuka, T.; Brown, B.A.; Carpenter, M.P.; Gade, A.; Rhodes, D.; et al. Multistep Coulomb excitation of 64Ni: Shape coexistence and nature of low-spin excitations. Phys. Rev. C 2022, 106, 044313. [Google Scholar] [CrossRef]
  533. Warburton, E.K.; Becker, J.A.; Brown, B.A. Mass systematics for A = 29–44 nuclei: The deformed A ∼ 32 region. Phys. Rev. C 1990, 41, 1147. [Google Scholar] [CrossRef]
  534. Nowacki, F.; Poves, A.; Caurier, E.; Bounthong, B. Shape Coexistence in 78Ni as the Portal to the Fifth Island of Inversion. Phys. Rev. Lett. 2016, 117, 272501. [Google Scholar] [CrossRef] [PubMed]
  535. Brown, B.A. Islands of insight in the nuclear chart. Physics 2010, 3, 104. [Google Scholar] [CrossRef]
  536. Sorlin, O. Shell Evolutions and Nuclear Forces. EPJ Web Conf. 2014, 66, 01016. [Google Scholar] [CrossRef]
  537. 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]
  538. Doornenbal, P.; Scheit, H.; Takeuchi, S.; Aoi, N.; Li, K.; Matsushita, M.; Steppenbeck, D.; Wang, H.; Baba, H.; Ideguchi, E.; et al. Mapping the deformation in the “island of inversion”: Inelastic scattering of 30Ne and 36Mg at intermediate energies. Phys. Rev. C 2016, 93, 044306. [Google Scholar] [CrossRef]
  539. 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]
  540. MacGregor, P.T.; Sharp, D.K.; Freeman, S.J.; Hoffman, C.R.; Kay, B.P.; Tang, T.L.; Gaffney, L.P.; Baader, E.F.; Borge, M.J.G.; Butler, P.A.; et al. Evolution of single-particle structure near the N = 20 island of inversion. Phys. Rev. C 2021, 104, L051301. [Google Scholar] [CrossRef]
  541. Pritychenko, B.V.; Glasmacher, T.; Cottle, P.D.; Ibbotson, R.W.; Kemper, K.W.; Miller, K.L.; Riley, L.A.; Scheit, H. Transition to the “island of inversion”: Fast-beam γ-ray spectroscopy of 28, 30Na. Phys. Rev. C 2002, 66, 024325. [Google Scholar] [CrossRef]
  542. Scheit, H. Spectroscopy in and around the Island of Inversion. J. Phys. Conf. Ser. 2011, 312, 092010. [Google Scholar] [CrossRef]
  543. Wimmer, K.; Kröll, T.; Krücken, R.; Bildstein, V.; Gernhäuser, R.; Bastin, B.; Bree, N.; Diriken, J.; Duppen, P.V.; Huyse, M.; et al. Discovery of the Shape Coexisting 0+ State in 32Mg by a Two Neutron Transfer Reaction. Phys. Rev. Lett. 2010, 105, 252501. [Google Scholar] [CrossRef] [PubMed]
  544. Choudhary, V.; Horiuchi, W.; Kimura, M.; Chatterjee, R. Enormous nuclear surface diffuseness of Ne and Mg isotopes in the island of inversion. Phys. Rev. C 2021, 104, 054313. [Google Scholar] [CrossRef]
  545. Doornenbal, P.; Scheit, H.; Takeuchi, S.; Aoi, N.; Li, K.; Matsushita, M.; Steppenbeck, D.; Wang, H.; Baba, H.; Crawford, H.; et al. In-Beam γ-Ray Spectroscopy of 34,36,38Mg: Merging the N = 20 and N = 28 Shell Quenching. Phys. Rev. Lett. 2013, 111, 212502. [Google Scholar] [CrossRef]
  546. Gade, A. Reaching into the N = 40 Island of Inversion with Nucleon Removal Reactions. Physics 2021, 3, 1226. [Google Scholar] [CrossRef]
  547. Horiuchi, W.; Inakura, T.; Michimasa, S. Large enhancement of total reaction cross sections at the edge of the island of inversion in Ti, Cr, and Fe isotopes. Phys. Rev. C 2022, 105, 014316. [Google Scholar] [CrossRef]
  548. Ljungvall, J.; Görgen, A.; Obertelli, A.; Korten, W.; Clément, E.; de France, G.; Bürger, A.; Delaroche, J.-P.; Dewald, A.; Gadea, A.; et al. Onset of collectivity in neutron-rich Fe isotopes: Toward a new island of inversion? Phys. Rev. C 2010, 81, 061301. [Google Scholar] [CrossRef]
  549. Porter, W.S.; Ashrafkhani, B.; Bergmann, J.; Brown, C.; Brunner, T.; Cardona, J.D.; Curien, D.; Dedes, I.; Dickel, T.; Dudek, J.; et al. Mapping the N = 40 island of inversion: Precision mass measurements of neutron-rich Fe isotopes. Phys. Rev. C 2022, 105, L041301. [Google Scholar] [CrossRef]
  550. Rocchini, M.; Garrett, P.; Zielińska, M.; Lenzi, S.; Dao, D.; Nowacki, F.; Bildstein, V.; MacLean, A.; Olaizola, B.; Ahmed, Z.; et al. First Evidence of Axial Shape Asymmetry and Configuration Coexistence in 74Zn : Suggestion for a Northern Extension of the N = 40 Island of Inversion. Phys. Rev. Lett. 2023, 130, 122502. [Google Scholar] [CrossRef]
  551. Louchart, C.S.C.; Obertelli, A.; Werner, V.; Doornenbal, P.; Nowacki, F.; Authelet, G.; Baba, H.; Calvet, D.; Château, F.; Corsi, A. Extension of the N = 40 Island of Inversion towards N = 50 : Spectroscopy of 66Cr, 70, 72Fe. Phys. Rev. Lett. 2015, 115, 192501. [Google Scholar]
  552. Kortelahti, M.O.; Zganjar, E.F.; Wood, J.L.; Bingham, C.R.; Carter, H.K.; Toth, K.S.; Hamilton, J.H.; Kormicki, J.; Chaturvedi, L.; Newbolt, W.B. Shape coexistence in 190Hg. Phys. Rev. C 1991, 43, 484. [Google Scholar] [CrossRef]
  553. Siciliano, M.; Zanon, I.; Goasduff, A.; John, P.R.; Rodríguez, T.R.; Péru, S.; Deloncle, I.; Libert, J.; Zielińska, M.; Ashad, D.; et al. Shape coexistence in neutron-deficient 188Hg investigated via lifetime measurements. Phys. Rev. C 2020, 102, 014318. [Google Scholar] [CrossRef]
  554. Hamilton, J.H.; Ramayya, A.V.; Bosworth, E.L.; Lourens, W.; Cole, J.D.; van Nooijen, B.; Garcia-Bermudez, G.; Martin, B.; Rao, B.N.; Kawakami, H.; et al. Crossing of Near-Spherical and Deformed Bands in 186,188Hg and New Isotopes 186,188Tl. Phys. Rev. Lett. 1975, 35, 562. [Google Scholar] [CrossRef]
  555. Proetel, D.; Diamond, R.M.; Kienle, P.; Leigh, J.R.; Maier, K.H.; Stephens, F.S. Evidence for Strongly Deformed Shapes in 186Hg. Phys. Rev. Lett. 1973, 31, 896. [Google Scholar] [CrossRef]
  556. Proetel, D.; Diamond, R.M.; Stephens, F.S. Nuclear deformations in 186Hg from lifetime measurements. Phys. Lett. B 1974, 48, 102. [Google Scholar] [CrossRef]
  557. Rud, N.; Ward, D.; Andrews, H.R.; Graham, R.L.; Geiger, J.S. Lifetimes in the Ground-State Band of 184Hg. Phys. Rev. Lett. 1973, 31, 1421. [Google Scholar] [CrossRef]
  558. Cole, J.D.; Ramayya, A.V.; Hamilton, J.H.; Kawakami, H.; van Nooijen, B.; Nettles, W.G.; Riedinger, L.L.; Turner, F.E.; Bingham, C.R.; Carter, H.K.; et al. Shape coexistence in 186Hg and the decay of 186Tl. Phys. Rev. C 1977, 16, 2010. [Google Scholar] [CrossRef]
  559. Gaffney, L.P.; Hackstein, M.; Page, R.D.; Grahn, T.; Scheck, M.; Butler, P.A.; Bertone, P.F.; Bree, N.; Carroll, R.J.; Carpenter, M.P.; et al. Shape coexistence in neutron-deficient Hg isotopes studied via lifetime measurements in 184,186Hg and two-state mixing calculations. Phys. Rev. C 2014, 89, 024307, Erratum in Phys. Rev. C 2014, 89, 059905(E). [Google Scholar] [CrossRef]
  560. Ma, W.C.; Ramayya, A.V.; Hamilton, J.H.; Robinson, S.J.; Cole, J.D.; Zganjar, E.F.; Spejewski, E.H.; Bengtsson, R.; Nazarewicz, W.; Zhang, J.-Y. The structure of high spin states in 184 Hg and 186 Hg. Phys. Lett. B 1986, 167, 277. [Google Scholar] [CrossRef]
  561. Bree, N.; Wrzosek-Lipska, K.; Petts, A.; Andreyev, A.; Bastin, B.; Bender, M.; Blazhev, A.; Bruyneel, B.; Butler, P.A.; Butterworth, J.; et al. Shape Coexistence in the Neutron-Deficient Even-Even 182–188Hg Isotopes Studied via Coulomb Excitation. Phys. Rev. Lett. 2014, 112, 162701. [Google Scholar] [CrossRef]
  562. Cole, J.D.; Hamilton, J.H.; Ramayya, A.V.; Lourens, W.; van Nooijen, B.; Kawakami, H.; Mink, L.A.; Spejewski, E.H.; Carter, H.K.; Mlekodaj, R.L.; et al. Decay of 188TI and observed shape coexistence in the bands of 188Hg. Phys. Rev. C 1984, 30, 1267. [Google Scholar] [CrossRef]
  563. Ma, W.C.; Ramayya, A.V.; Hamilton, J.H.; Robinson, S.J.; Barclay, M.E.; Zhao, K.; Cole, J.D.; Zganjar, E.F.; Spejewski, E.H. Ground state shaoe and crossing of near spherical and deformed bands in 182 Hg. Phys. Lett. B 1984, 139, 276. [Google Scholar] [CrossRef]
  564. Olaizola, B.; Garnsworthy, A.B.; Ali, F.A.; Andreoiu, C.; Ball, G.C.; Bernier, N.; Bidaman, H.; Bildstein, V.; Bowry, M.; Caballero-Folch, R.; et al. Shape coexistence in the neutron-deficient lead region: A systematic study of lifetimes in the even-even 188--200Hg with the GRIFFIN spectrometer at TRIUMF. Phys. Rev. C 2019, 100, 024301. [Google Scholar] [CrossRef]
  565. Wrzosek-Lipska, K.; Gaffney, L.P. Unique and complementary information on shape coexistence in the neutron-deficient Pb region derived from Coulomb excitation. J. Phys. G Nucl. Part Phys. 2016, 43, 024012. [Google Scholar] [CrossRef]
  566. Carpenter, M.P.; Janssens, R.V.F.; Amro, H.; Blumenthal, D.J.; Brown, L.T.; Seweryniak, D.; Woods, P.J.; Ackermann, D.; Ahmad, I.; Davids, C.; et al. Excited States in 176,178Hg and Shape Coexistence in Very Neutron-Deficient Hg Isotopes. Phys. Rev. Lett. 1997, 78, 3650. [Google Scholar] [CrossRef]
  567. Dracoulis, G.D.; Stuchbery, A.E.; Macchiavelli, A.O.; Beausang, C.W.; Burde, J.; Deleplanque, M.A.; Diamond, R.M.; Stephens, F.S. Shape co-existence in 180Hg and delineation of the midshell minimum. Phys. Lett. B 1988, 208, 365. [Google Scholar] [CrossRef]
  568. Elseviers, J.; Andreyev, A.; Antalic, S.; Barzakh, A.; Bree, N.; Cocolios, T.; Comas, V.; Diriken, J.; Fedorov, D.; Fedosseyev, V.N.; et al. Shape coexistence in 180Hg studied through the β decay of 180Tl. Phys. Rev. C 2011, 84, 034307. [Google Scholar] [CrossRef]
  569. Kondev, F.G.; Carpenter, M.P.; Janssens, R.V.F.; Wiedenhöver, I.; Alcorta, M.; Bhattacharyya, P.; Brown, L.T.; Davids, C.N.; Fischer, S.M.; Khoo, T.L.; et al. Complex band structure in neutron-deficient 178Hg. Phys. Rev. C 1999, 61, 011303. [Google Scholar] [CrossRef]
  570. Kondev, F.G.; Janssens, R.V.F.; Carpenter, M.P.; Abu Saleem, K.; Ahmad, I.; Alcorta, M.; Amro, H.; Bhattacharyya, P.; Brown, L.T.; Caggiano, T.; et al. Interplay between octupole and quasiparticle excitations in 178Hg and 180Hg. Phys. Rev. C 2000, 62, 044305. [Google Scholar] [CrossRef]
  571. Müller-Gatermann, C.; Dewald, A.; Fransen, C.; Auranen, K.; Badran, H.; Beckers, M.; Blazhev, A.; Blazhev, A.; Cullen, D.M.; Fruet, G.; et al. Shape coexistence in 178Hg. Phys. Rev. C 2019, 99, 054325. [Google Scholar] [CrossRef]
  572. Bengtsson, R.; Bengtsson, T.; Dudek, J.; Leander, G.; Nazarewicz, W.; Zhang, J.-Y. Shape coexistence and shape transitions in even-even Pt and Hg isotopes. Phys. Lett. B 1987, 183, 1. [Google Scholar] [CrossRef]
  573. Delion, D.S.; Liotta, R.J.; Qi, C.; Wyss, R. Probing shape coexistence by α decays to 0+ states. Phys. Rev. C 2014, 90, 061303. [Google Scholar]
  574. Jiao, C.F.; Shi, Y.; Liu, H.L.; Xu, F.R.; Walker, P.M. Shape-coexisting rotation in neutron-deficient Hg and Pb nuclei. Phys. Rev. C 2015, 91, 034309. [Google Scholar] [CrossRef]
  575. Shi, Y.; Xu, F.R.; Liu, H.L.; Walker, P.M. Multi-quasiparticle excitation: Extending shape coexistence in A∼190 neutron-deficient nuclei. Phys. Rev. C 2010, 82, 044314. [Google Scholar] [CrossRef]
  576. Richards, J.D.; Berggren, T.; Bingham, C.R.; Nazarewicz, W.; Wauters, J. α decay and shape coexistence in the α-rotor model. Phys. Rev. C 1997, 56, 1389. [Google Scholar] [CrossRef]
  577. Zhang, J.-Y.; Hamilton, J.H. Theoretical calculations of the different shape coexistence observed in 190Hg. Phys. Lett. B 1991, 260, 11. [Google Scholar] [CrossRef]
  578. Nazarewicz, W. Variety of shapes in the mercury and lead isotopes. Phys. Lett. B 1993, 305, 195. [Google Scholar] [CrossRef]
  579. Bonatsos, D.; Karakatsanis, K.E.; Martinou, A.; Mertzimekis, T.J.; Minkov, N. Microscopic origin of shape coexistence in the N = 90, Z = 64 region. Phys. Lett. B 2022, 829, 137099. [Google Scholar] [CrossRef]
  580. Bonatsos, D.; Karakatsanis, K.E.; Martinou, A.; Mertzimekis, T.J.; Minkov, N. Islands of shape coexistence from single-particle spectra in covariant density functional theory. Phys. Rev. C 2022, 106, 044323. [Google Scholar] [CrossRef]
  581. Penninga, J.; Hesselink, W.H.A.; Balanda, A.; Stolk, A.; Verheul, H.; van Klinken, J.; Riezebos, H.J.; de Voigt, M.J.A. Proton particle-hole states and collective excitations in 196Pb. Nucl. Phys. A 1987, 471, 535. [Google Scholar] [CrossRef]
  582. Van Duppen, P.; Coenen, E.; Deneffe, K.; Huyse, M.; Heyde, K.; Van Isacker, P. Observation of Low-Lying Jπ=0+ States in the Single-Closed-Shell Nuclei 192-198Pb. Phys. Rev. Lett. 1984, 52, 1974. [Google Scholar] [CrossRef]
  583. Van Duppen, P.; Coenen, E.; Deneffe, K.; Huyse, M.; Wood, J.L. β+/electron-capture decay of 192,194,196,198,200Bi: Experimental evidence for low lying 0+ states. Phys. Rev. C 1987, 35, 1861. [Google Scholar] [CrossRef] [PubMed]
  584. Allatt, R.G.; Page, R.D.; Leino, M.; Enqvist, T.; Eskola, K.; Greenlees, P.T.; Jones, P.; Julin, R.; Kuusiniemi, P.; Trzaska, W.H.; et al. Fine structure in 192Po α-decay and shape coexistence in 188Pb. Phys. Lett. B 1998, 437, 29. [Google Scholar] [CrossRef]
  585. Dracoulis, G.D.; Lane, G.J.; Byrne, A.P.; Baxter, A.M.; Kibédi, T.; Macchiavelli, A.O.; Fallon, P.; Clark, R.M. Isomer bands, E0 transitions, and mixing due to shape coexistence in Pb 106 82 188 . Phys. Rev. C 2003, 67, 051301. [Google Scholar] [CrossRef]
  586. Dracoulis, G.D.; Lane, G.J.; Byrne, A.P.; Kibédi, T.; Baxter, A.M.; Macchiavèlli, A.O.; Fallon, P.; Clark, R.M. Spectroscopy of 82188Pb106: Evidence for shape coexistence. Phys. Rev. C 2004, 69, 054318. [Google Scholar] [CrossRef]
  587. Pakarinen, J.; Darby, I.G.; Eeckhaudt, S.; Enqvist, T.; Grahn, T.; Greenlees, P.T.; Hellemans, V.; Heyde, K.; Johnston-Theasby, F.; Jones, P.; et al. Evidence for oblate structure in 186Pb. Phys. Rev. C 2005, 72, 011304. [Google Scholar] [CrossRef]
  588. Cocks, J.F.C.; Muikku, M.; Korten, W.; Wadsworth, R.; Chmel, S.; Domscheit, J.; Greenlees, P.T.; Helariutta, K.; Hibbert, I.; Houry, M.; et al. First observation of excited states in 184Pb: Spectroscopy beyond the neutron mid-shell. Eur. Phys. J. A 1998, 3, 17. [Google Scholar] [CrossRef]
  589. Jenkins, D.G.; Muikku, M.; Greenlees, P.T.; Hauschild, K.; Helariutta, K.; Jones, P.M.; Julin, R.; Juutinen, S.; Kankaanpää, H.; Kelsall, N.S.; et al. First observation of excited states in 182Pb. Phys. Rev. C 2000, 62, 021302. [Google Scholar] [CrossRef]
  590. Rahkila, P.; Jenkins, D.G.; Pakarinen, J.; Gray-Jones, C.; Greenlees, P.T.; Jakobsson, U.; Jones, P.; Julin, R.; Juutinen, S.; Ketelhut, S.; et al. Shape coexistence at the proton drip-line: First identification of excited states in 180Pb. Phys. Rev. C 2010, 82, 011303. [Google Scholar] [CrossRef]
  591. Julin, R.; Grahn, T.; Pakarinen, J.; Rahkila, P. In-beam spectroscopic studies of shape coexistence and collectivity in the neutron-deficient Zapprox82 nuclei. J. Phys. G Nucl. Part. Phys. 2016, 43, 024004. [Google Scholar] [CrossRef]
  592. Xu, C.; Ren, Z. α transitions to coexisting 0+ states in Pb and Po isotopes. Phys. Rev. C 2007, 75, 044301. [Google Scholar] [CrossRef]
  593. Alber, D.; Alfier, R.; Bach, C.E.; Fossan, D.B.; Grawe, H.; Kluge, H.; Lach, M.; Maier, K.H.; Schramm, M.; Schubart, R.; et al. Quadrupole and octupole collectivity in light Po isotopes. Z. Phys. A 1991, 339, 225. [Google Scholar] [CrossRef]
  594. Grahn, T.; Dewald, A.; Moller, O.; Julin, R.; Beausang, C.W. Collectivity and Configuration Mixing in 186,188Pb and 194Po. Phys. Rev. Lett. 2006, 97, 062501. [Google Scholar] [CrossRef]
  595. Grahn, T.; Dewald, A.; Möller, O.; Julin, R.; Beausang, C.W.; Christen, S.; Darby, I.G.; Eeckhaudt, S.; Greenlees, P.T.; Görgen, A. Lifetimes of intruder states in 186,188Pb and 194Po. Nucl. Phys. A 2008, 801, 83. [Google Scholar] [CrossRef]
  596. Helariutta, K.; Cocks, J.F.C.; Enqvist, T.; Greenlees, P.T.; Jones, P.; Julin, R.; Juutinen, S.; Jämsen, P.; Kankaanpää, H.; Kettunen, H.; et al. Gamma-ray spectroscopy of 192-195Po. Eur. Phys. J. A 1999, 6, 289. [Google Scholar] [CrossRef]
  597. Younes, W.; Cizewski, J.A. Systematical behavior of even—A polonium isotopes. Phys. Rev. C 1997, 55, 1218. [Google Scholar] [CrossRef]
  598. Bijnens, N.; Decrock, P.; Franchoo, S.; Gaelens, M.; Huyse, M.; Hwang, H.-Y.; Reusen, I.; Szerypo, J.; von Schwarzenberg, J.; Vancraeynest, G.; et al. Study of 200,202Po through β+ and electron-capture decay and the manifestation of shape coexistence in the lighter Po isotopes. Phys. Rev. C 1998, 58, 754. [Google Scholar] [CrossRef]
  599. Kesteloot, N.; Bastin, B.; Gaffney, L.P.; Wrzosek-Lipska, K.; Auranen, K.; Bauer, C.; Bender, M.; Bildstein, V.; Blazhev, A.; Bonig, S.; et al. Deformation and mixing of coexisting shapes in neutron-deficient polonium isotopes. Phys. Rev. C 2015, 92, 054301. [Google Scholar] [CrossRef]
  600. Oros, A.M.; Heyde, K.; De Coster, C.; Decroix, B.; Wyss, R.; Barrett, B.R.; Navratil, P. Shape coexistence in the light Po isotopes. Nucl. Phys. A 1999, 645, 107. [Google Scholar] [CrossRef]
  601. Agbemava, S.E.; Afanasjev, A.V.; Nakatsukasa, T.; Ring, P. Covariant density functional theory: Reexamining the structure of superheavy nuclei. Phys. Rev. C 2015, 92, 054310. [Google Scholar] [CrossRef]
  602. Prassa, V.; Nikšić, T.; Lalazissis, G.A.; Vretenar, D. Relativistic energy density functional description of shape transitions in superheavy nuclei. Phys. Rev. C 2012, 86, 024317. [Google Scholar] [CrossRef]
  603. Prassa, V.; Nikšić, T.; Vretenar, D. Structure of transactinide nuclei with relativistic energy density functionals. Phys. Rev. C 2013, 88, 044324. [Google Scholar] [CrossRef]
  604. Garg, U.; Chaudhury, A.; Drigert, M.W.; Funk, E.G.; Mihelich, J.W.; Radford, D.C.; Helppi, H.; Holzmann, R.; Janssens, R.V.F.; Khoo, T.L.; et al. Lifetime measurements in 184Pt and the shape coexistence picture. Phys. Lett. B 1986, 180, 319. [Google Scholar] [CrossRef]
  605. Larabee, A.J.; Carpenter, M.P.; Riedinger, L.L.; Courtney, L.H.; Waddington, J.C.; Janzen, V.P.; Nazarewicz, W.; Zhang, J.-Y.; Bengtsson, R.; Lèander, G.A. Shape coexistence and alignment processes in the light Pt and Au region. Phys. Lett. B 1986, 169, 21. [Google Scholar] [CrossRef]
  606. Xu, Y.; Krane, K.S.; Gummin, M.A.; Jarrio, M.; Wood, J.L.; Zganjar, E.F.; Carter, H.K. Shape coexistence and electric monopole transitions in 184Pt. Phys. Rev. Lett. 1992, 68, 3853. [Google Scholar] [CrossRef] [PubMed]
  607. Davidson, P.M.; Dracoulis, G.D.; Kibédi, T.; Byrne, A.P.; Anderssen, S.S.; Baxter, A.M.; Fabricius, B.; Lane, G.J.; Stuchbery, A.E. Non-yrast states and shape co-existence in light Pt isotopes. Nucl. Phys. A 1999, 657, 219. [Google Scholar] [CrossRef]
  608. Dracoulis, G.D.; Stuchbery, A.E.; Byrne, A.P.; Poletti, A.R.; Poletti, S.J.; Gerl, J.; Bark, R.A. Shape coexistence in very neutron-deficient Pt isotopes. J. Phys. G Nucl. Phys. 1986, 12, L97. [Google Scholar] [CrossRef]
  609. Dracoulis, G.D.; Fabricius, B.; Stuchbery, A.E.; Macchiavelli, A.O.; Korten, W.; Azaiez, F.; Rubel, E.; Deleplanque, M.A.; Diamond, R.M.; Stephens, F.S. Shape coexistence from the structure of the yrast band in 174Pt. Phys. Rev. C 1991, 44, R1246. [Google Scholar] [CrossRef]
  610. Gladnishki, K.A.; Petkov, P.; Dewald, A.; Fransen, C.; Hackstein, M.; Jolie, J.; Pissulla, T.; Rother, W.; Zell, K.O. Yrast electromagnetic transition strengths and shape coexistence in 182Pt. Nucl. Phys. A 2012, 877, 19. [Google Scholar] [CrossRef]
  611. Goon, J.T.M.; Hartley, D.J.; Riedinger, L.L.; Carpenter, M.P.; Kondev, F.G.; Janssens, R.V.F.; Abu Saleem, K.H.; Ahmad, I.; Amro, H.; Cizewski, J.A.; et al. Shape coexistence and band crossings in 174Pt. Phys. Rev. C 2004, 70, 014309. [Google Scholar] [CrossRef]
  612. Mukhopadhyay, S.; Biswas, D.C.; Tanel, S.K.; Danu, L.S.; Joshi, B.N.; Prajapati, G.K.; Nag, S.; Trivedi, T.; Saha, S.; Sethi, J.; et al. Coexisting shape- and high-K isomers in the shape transitional nucleus 188Pt. Phys. Lett. B 2014, 739, 462. [Google Scholar] [CrossRef]
  613. Stuchbery, A.E.; Anderssen, S.S.; Byrne, A.P.; Davidson, P.M.; Dracoulis, G.D.; Lane, G.J. Measured Magnetic Moments and Shape Coexistence in the Neutron-Deficient Nuclei 184,186,188Pt. Phys. Rev. Lett. 1996, 76, 2246. [Google Scholar] [CrossRef] [PubMed]
  614. McCutchan, E.A.; Casten, R.F.; Zamfir, N.V. Simple interpretation of shape evolution in Pt isotopes without intruder states. Phys. Rev. C 2005, 71, 061301. [Google Scholar] [CrossRef]
  615. Budaca, R.; Budaca, A.I. Shape phase mixing in critical point nuclei. Phys. Rev. C 2016, 94, 054306. [Google Scholar] [CrossRef]
  616. Stevenson, P.D.; Brine, M.P.; Podolyak, Z.; Regan, P.H.; Walker, P.M.; Rikovska Stone, J. Shape evolution in the neutron-rich tungsten region. Phys. Rev. C 2005, 72, 047303. [Google Scholar] [CrossRef]
  617. Kumar, R.; Govil, I.M.; Dhal, A.; Chaturvedi, L.; Praharaj, C.R.; Rath, A.K.; Kiran Kumar, G.; Basu, S.K.; Chakraborty, A.; Krishichayan; et al. Triaxial shape coexistence and new aligned band in 178Os. Phys. Rev. C 2009, 80, 054319. [Google Scholar] [CrossRef]
  618. Tanabe, K.; Sugawara-Tanabe, K. Shape coexistence at high spins in 158Er and 160Yb predicted by the self-consistent calculation. Phys. Lett. B 1984, 135, 353. [Google Scholar] [CrossRef]
  619. Rowe, D.J. The emergence of deformation and rotational states in the many-nucleon quantum theory of nuclei. J. Phys. G Nucl. Part. Phys. 2016, 43, 024011. [Google Scholar] [CrossRef]
  620. Rowe, D.J. Nuclear shape coexistence from the perspective of an algebraic many-nucleon version of the Bohr-Mottelson unified model. Phys. Rev. C 2020, 101, 054301. [Google Scholar] [CrossRef]
  621. Bjerregaard, J.H.; Hansen, O.; Nathan, O.; Hinds, S. The (t, p) reaction with the even isotopes of Sm. Nucl. Phys. 1966, 86, 145. [Google Scholar] [CrossRef]
  622. McLatchie, W.; Kitching, J.E.; Darcey, W. The reaction 154Sm (p, t) 152Sm and further evidence for shape coexistence in 152Sm. Phys. Lett. B 1969, 30, 529. [Google Scholar] [CrossRef]
  623. Casten, R.F.; Kusnezov, D.; Zamfir, N.V. Phase Transitions in Finite Nuclei and the Integer Nucleon Number Problem. Phys. Rev. Lett. 1999, 82, 5000. [Google Scholar] [CrossRef]
  624. Garrett, P.E.; Kulp, W.D.; Wood, J.L.; Byopadhyay, D.; Choudry, S.; Dashdorj, D.; Lesher, S.R.; McEllistrem, M.T.; Mynk, M.; Orce, J.N.; et al. New Features of Shape Coexistence in 152Sm. Phys. Rev. Lett. 2009, 103, 062501. [Google Scholar] [CrossRef] [PubMed]
  625. Basak, S.; Alam, S.S.; Kumar, D.; Saha, A.; Bhattacharjee, T. Shape coexistence scenario in 150Sm from a γ-γ fast-timing measurement. Phys. Rev. C 2021, 104, 024320. [Google Scholar] [CrossRef]
  626. Passoja, A.; Julin, R.; Kantele, J.; Luontama, M.; Vergnes, M. E0 transitions in 70Ge and shape-coexistence interpretation of even-mass Ge isotopes. Nucl. Phys. A 1985, 441, 261. [Google Scholar] [CrossRef]
  627. Gupta, J.B.; Hamilton, J.H. Outstanding problems in the band structures of 152Sm. Phys. Rev. C 2017, 96, 034321. [Google Scholar] [CrossRef]
  628. Rajbanshi, S.; Bisoi, A.; Nag, S.; Saha, S.; Sethi, J.; Trivedi, T.; Bhattacharjee, T.; Bhattacharyya, S.; Chattopadhyay, S.; Gangopadhyay, G.; et al. Shape coexistence in the near-spherical 142Sm nucleus. Phys. Rev. C 2014, 89, 014315. [Google Scholar] [CrossRef]
  629. Cardona, M.A.; Lunardi, S.; Bazzacco, D.; de Angelis, G.; Roca, V. Shape coexistence in 140Sm and the onset of deformation below N = 82 from lifetime measurements. Phys. Rev. C 1991, 44, 891. [Google Scholar] [CrossRef]
  630. Casten, R.F.; Warner, D.D.; Brenner, D.S.; Gill, R.L. Relation between the Z = 64 Shell Closure and the Onset of Deformation at N = 88–90. Phys. Rev. Lett. 1981, 47, 1433. [Google Scholar] [CrossRef]
  631. Petrache, C.M.; Walker, P.M.; Guo, S.; Chen, Q.B.; Frauendorf, S.; Liu, Y.X.; Wyss, R.A.; Mengoni, D.; Qiang, Y.H.; Astier, A.; et al. Diversity of shapes and rotations in the γ-soft 130Ba nucleus: First observation of a t-band in the A = 130 mass region. Phys. Lett. B 2019, 795, 241. [Google Scholar] [CrossRef]
  632. Bron, J.; Hesselink, W.H.A.; Van Poelgeest, A.; Zalmstra, J.J.A.; Uitzinger, M.J.; Verheul, H.; Heyde, K.; Waroquier, M.; Vincx, H.; Van Isacker, P. Collective bands in even mass Sn isotopes. Nucl. Phys. A 1979, 318, 335. [Google Scholar] [CrossRef]
  633. Cross, D.S.; Pore, J.L.; Andreoiu, C.; Ball, G.C.; Bender, P.C.; Chester, A.S.; Churchman, R.; Demand, G.A.; Diaz Varela, A.; Dunlop, R.; et al. Conversion-electron spectroscopy and gamma-gamma angular correlation measurements in 116Sn. Eur. Phys. J. A 2017, 53, 216. [Google Scholar] [CrossRef]
  634. Harada, H.; Murakami, T.; Yoshida, K.; Kasagi, J.; Inamura, T.; Kubo, T. Intruder deformed bands in 110Sn and 112Sn. Phys. Lett. B 1988, 207, 17. [Google Scholar] [CrossRef]
  635. Harada, H.; Sugawara, M.; Kusakari, H.; Shinohara, H.; Ono, Y.; Furuno, K.; Hosoda, T.; Adachi, M.; Matsuki, S.; Kawamura, N. High-spin states in 114Sn. Phys. Rev. C 1989, 39, 132. [Google Scholar] [CrossRef]
  636. Petrache, C.M.; Régis, J.-M.; Andreoiu, C.; Spieker, M.; Michelagnoli, C.; Garrett, P.E.; Astier, A.; Dupont, E.; Garcia, F.; Guo, S.; et al. Collectivity of the 2p-2h proton intruder band of 116Sn. Phys. Rev. C 2019, 99, 024303. [Google Scholar] [CrossRef]
  637. Pore, J.L.; Cross, D.S.; Andreoiu, C.; Ashley, R.; Ball, G.C.; Bender, P.C.; Chester, A.S.; Diaz Varela, A.; Demand, G.A.; Dunlop, R.; et al. Study of the β-decay of 116m1In: A new interpretation of low-lying 0+ states in 116Sn. Eur. Phys. J. A 2017, 53, 27. [Google Scholar] [CrossRef]
  638. Schimmer, M.; Wirowski, R.; Albers, S.; Böhm, G.; Dewald, A.; Gelberg, A.; von Brentano, P. Observation of a new deformed structure in 114Sn. Z. Phys. A 1991, 338, 117. [Google Scholar] [CrossRef]
  639. Spieker, M.; Petkov, P.; Litvinova, E.; Müller-Gatermann, C.; Pickstone, S.G.; Prill, S.; Scholz, P.; Zilges, A. Shape coexistence and collective low-spin states in 112,114Sn studied with the (p, p’ γ) Doppler-shift attenuation coincidence technique. Phys. Rev. C 2018, 97, 054319. [Google Scholar] [CrossRef]
  640. Viggars, D.A.; Taylor, H.W.; Singh, B.; Waddington, J.C. Observation of collective behavior in 110Sn. Phys. Rev. C 1987, 36, 1006. [Google Scholar] [CrossRef]
  641. Wolińska-Cichocka, M.; Kownacki, J.; Urban, W.; Ruchowska, E.; Płóciennik, W.A.; Bekman, B.; Droste, C.; Gast, W.; Lieder, R.; Mȩczyński, W.; et al. Gamma-ray spectroscopy in 110Sn and 111Sn. Eur. Phys. J. A 2005, 24, 259. [Google Scholar] [CrossRef]
  642. Azaiez, F.; Andriamonje, S.; Chemin, J.F.; Fidah, M.; Scheurer, J.N.; Bastin, G.; Thibaud, J.P.; Beck, F.; Costa, G.; Bruandet, J.F.; et al. High-spin states in neutron deficient 106Sn and 108Sn isotopes. Nucl. Phys. A 1989, 501, 401. [Google Scholar] [CrossRef]
  643. Juutinen, S.; Mäkelä, E.; Julin, R.; Piiparinen, M.; Törmänen, S.; Virtanen, A.; Adamides, E.; Ataç, A.; Blomqvist, J.; Cederwall, B.; et al. Coexistence of collective and quasiparticle structures in 106Sn and 108Sn. Nucl. Phys. A 1997, 617, 74. [Google Scholar] [CrossRef]
  644. Wadsworth, R.; Andrews, H.R.; Clark, R.M.; Fossan, D.B.; Galindo-Uribarri, A.; Hughes, J.R.; Janzen, V.P.; LaFosse, D.R.; Mullins, S.M.; Paul, E.S.; et al. Intruder bands in 108Sn. Nucl. Phys. A 1993, 559, 461. [Google Scholar] [CrossRef]
  645. Wadsworth, R.; Beausang, C.W.; Cromaz, M.; DeGraaf, J.; Drake, T.E.; Fossan, D.B.; Flibotte, S.; Galindo-Uribarri, A.; Hauschild, K.; Hibbert, I.M.; et al. Smooth band termination in 108Sn. Phys. Rev. C 1996, 53, 2763. [Google Scholar] [CrossRef] [PubMed]
  646. Song, T.; Yang, L.M. Microscopic investigation of pair excitation and shape coexistence in even-even nuclei. Phys. Rev. C 1989, 40, 1782. [Google Scholar] [CrossRef] [PubMed]
  647. Yang, L.M.; Song, T.; Wang, X.H. Microscopic treatment of shape coexistence in even-even nuclei and related extension of IBM. Phys. Lett. B 1986, 175, 6. [Google Scholar] [CrossRef]
  648. Meyer, R.A.; Peker, L. Evidence for the coexistence of shapes in even-mass Cd nuclei. Z. Phys. A 1977, 283, 379. [Google Scholar] [CrossRef]
  649. Fahlander, C.; Bäcklin, A.; Hasselgren, L.; Kavka, A.; Mittal, V.; Svensson, L.E.; Varnestig, B.; Cline, D.; Kotlinski, B.; Grein, H.; et al. Quadrupole collective properties of 114Cd. Nucl. Phys. A 1988, 485, 327. [Google Scholar] [CrossRef]
  650. Kern, J.; Bruder, A.; Drissi, S.; Ionescu, V.A.; Kusnezov, D. Study of 110Cd by the 108Pd (α,2) reaction. Nucl. Phys. A 1990, 512, 1. [Google Scholar] [CrossRef]
  651. Kusnezov, D.; Bruder, A.; Ionescu, V.; Kern, J.; Rast, M.; Heyde, K.; Van Isacker, P.; Moreau, J.; Waroquier, M.; Meyer, R.A. Mixing of the ground band and two-particle four-hole intruder band in 110Cd. Helv. Phys. Acta 1987, 60, 456. [Google Scholar]
  652. Garrett, P.E.; Rodríguez, T.R.; Diaz Varela, A.; Green, K.L.; Bangay, J.; Finlay, A.; Austin, R.A.E.; Ball, G.C.; Bandyopadhyay, D.S.; Bildstein, V.; et al. Multiple Shape Coexistence in 110,112Cd and Beyond Mean Field Calculations. J. Phys. Conf. Ser. 2020, 1643, 012131. [Google Scholar] [CrossRef]
  653. Kumpulainen, J.; Julin, R.; Kantele, J.; Passoja, A.; Trzaska, W.H.; Verho, E.; Väärämäki, J. New features in systematics of low-spin states in even 106-120Cd. Z. Phys. A At. Nucl. 1990, 335, 109. [Google Scholar] [CrossRef]
  654. Kumpulainen, J.; Julin, R.; Kantele, J.; Passoja, A.; Trzaska, W.H.; Verho, E.; Väärämäki, J.; Cutoiu, D.; Ivascu, M. Systematic study of low-spin states in even Cd nuclei. Phys. Rev. C 1992, 45, 640. [Google Scholar] [CrossRef] [PubMed]
  655. Gray, T.J.; Allmond, J.M.; Janssens, R.V.F.; Korten, W.; Stuchbery, A.E.; Wood, J.L.; Ayangeakaa, A.D.; Bottoni, S.; Bucher, B.M.; Campbell, C.M.; et al. E2 rotational invariants of 0 1 + and 2 1 + states for 106Cd: The emergence of collective rotation. Phys. Lett. B 2022, 834, 137446. [Google Scholar] [CrossRef]
  656. Aprahamian, A.; Brenner, D.S.; Casten, R.F.; Gill, R.L.; Piotrowski, A. First observation of a near-harmonic vibrational nucleus. Phys. Rev. Lett. 1987, 59, 535. [Google Scholar] [CrossRef]
  657. Nomura, K.; Karakatsanis, K.E. Collective-model description of shape coexistence and intruder states in cadmium isotopes based on a relativistic energy density functional. Phys. Rev. C 2022, 106, 064317. [Google Scholar] [CrossRef]
  658. Rikovska, J.; Stone, N.J.; Green, V.R.; Walker, P.M. IBA-2 model calculation on even mass tellurium isotopes compared with results of recent nuclear orientation experiments. Hyperfine Interact. 1985, 22, 405. [Google Scholar] [CrossRef]
  659. Rikovska, J.; Stone, N.J.; Walters, W.B. Dynamical symmetries in even-even Te nuclides. Phys. Rev. C 1987, 36, 2162. [Google Scholar] [CrossRef]
  660. Rikovska, J.; Stone, N.J.; Walker, P.M.; Walters, W.B. Intruder states in even-even Te nuclei. Nucl. Phys. A 1989, 505, 145. [Google Scholar] [CrossRef]
  661. Walker, P.M.; Ashworth, C.J.; Grant, I.S.; Green, V.R.; Rikovska, J.; Shaw, T.L.; Stone, N.J. E0 transitions in the light tellurium isotopes: Evidence for intruder states. J. Phys. G Nucl. Phys. 1987, 13, L195. [Google Scholar] [CrossRef]
  662. Abood, S.N.; Najim, L.A. Interacting boson model (IBM-2) calculations of selected even-even Te nuclei. Adv. Appl. Sci. Res. 2013, 4, 444. [Google Scholar]
  663. Sabri, H.; Jahangiri, Z.; Mohammadi, M.A. Investigation of shape coexistence in 118-128Te isotopes. Nucl. Phys. A 2016, 946, 11. [Google Scholar] [CrossRef]
  664. Gupta, J.B. Spectral features of vibrational Te isotopes. Phys. Rev. C 2023, 107, 034315. [Google Scholar] [CrossRef]
  665. Lhersonneau, G.; Wang, J.C.; Hankonen, S.; Dendooven, P.; Jones, P.; Julin, R.; Äystö, J. Decays of 110Rh and 112Rh to the near neutron midshell isotopes 110Pd and 112Pd. Phys. Rev. C 1999, 60, 014315. [Google Scholar] [CrossRef]
  666. Prados-Estévez, F.M.; Peters, E.E.; Chakraborty, A.; Mynk, M.G.; Bandyopadhyay, D.; Boukharouba, N.; Choudry, S.N.; Crider, B.P.; Garrett, P.E.; Hicks, S.F.; et al. Collective quadrupole behavior in 106Pd. Phys. Rev. C 2017, 95, 034328. [Google Scholar] [CrossRef]
  667. Svensson, L.E.; Fahlander, C.; Hasselgren, L.; Bäcklin, A.; Westerberg, L.; Cline, D.; Czosnyka, T.; Wu, C.Y.; Diamond, R.M.; Kluge, H. Multiphonon vibrational states in 106,108Pd. Nucl. Phys. A 1995, 584, 547. [Google Scholar] [CrossRef]
  668. Peters, E.E.; Prados-Estévez, F.M.; Chakraborty, A.; Mynk, M.G.; Bandyopadhyay, D.; Choudry, S.N.; Crider, B.P.; Garrett, P.E.; Hicks, S.F.; Kumar, A.; et al. E0 transitions in 106Pd: Implications for shape coexistence. Eur. Phys. J. A 2016, 52, 96. [Google Scholar] [CrossRef]
  669. Zamfir, N.V.; Caprio, M.A.; Casten, R.F.; Barton, C.J.; Beausang, C.W.; Berant, Z.; Brenner, D.S.; Chou, W.T.; Cooper, J.R.; Hecht, A.A.; et al. 102Pd: An E(5) nucleus? Phys. Rev. C 2002, 65, 044325. [Google Scholar] [CrossRef]
  670. Hosseinnezhad, A.; Jalili Majarshin, A.; Luo, Y.A.; Ahmadian, D.; Sabri, H. Deformation in 92–128 Pd isotopes. Nucl. Phys. A 2022, 1028, 122523. [Google Scholar] [CrossRef]
  671. Ansari, S.; Régis, J.; Jolie, J.; Saed-Samii, N.; Warr, N.; Korten, W.; Zielińska, M.; Salsac, M.; Blanc, A.; Jentschel, M.; et al. Experimental study of the lifetime and phase transition in neutron-rich 98,100,102Zr. Phys. Rev. C 2017, 96, 054323. [Google Scholar] [CrossRef]
  672. Heyde, K.; Meyer, R.A. Comment on “Monopole strength and shape coexistence in the A≃100 mass region”. Phys. Rev. C 1990, 42, 790. [Google Scholar] [CrossRef]
  673. Mach, H.; Moszyǹski, M.; Gill, R.L.; Wohn, F.K.; Winger, J.A.; Hill, J.C.; Molnár, G.; Sistemich, K. Deformation and shape coexistence of 0+ states in 98Sr and 100Zr. Phys. Lett. B 1989, 230, 21. [Google Scholar] [CrossRef]
  674. Mach, H.; Moszyǹski, M.; Gill, R.L.; Molnár, G.; Wohn, F.K.; Winger, J.A.; Hill, J.C. Monopole strength and shape coexistence in the A ≃ 100 mass region. Phys. Rev. C 1990, 41, 350. [Google Scholar] [CrossRef] [PubMed]
  675. Mach, H.; Moszyǹski, M.; Gill, R.L.; Molnár, G.; Wohn, F.K.; Winger, J.A.; Hill, J.C. Reply to “Comment on `Monopole strength and shape coexistence in the A ≃ 100 mass region”’. Phys. Rev. C 1990, 42, 793. [Google Scholar] [CrossRef] [PubMed]
  676. Wohn, F.K.; Hill, J.C.; Howard, C.B.; Sistemich, K.; Petry, R.F.; Gill, R.L.; Mach, H.; Piotrowski, A. Shape coexistence and level structure of 100Zr from decay of the low-spin isomer of 100Y. Phys. Rev. C 1986, 33, 677. [Google Scholar] [CrossRef]
  677. Singh, P.; Korten, W.; Hagen, T.; Görgen, A.; Grente, L.; Salsac, M.-D.; Farget, F.; Clément, E.; de France, G.; Braunroth, T.; et al. Evidence for Coexisting Shapes through Lifetime Measurements in 98Zr. Phys. Rev. Lett. 2018, 121, 192501. [Google Scholar] [CrossRef]
  678. Witt, W.; Werner, V.; Pietralla, N.; Albers, M.; Ayangeakaa, A.D.; Bucher, B.; Carpenter, M.P.; Cline, D.; David, H.M.; Hayes, A.; et al. Sub-shell closure and shape coexistence in the transitional nucleus 98Zr. Phys. Rev. C 2018, 98, 041302. [Google Scholar] [CrossRef]
  679. Kremer, C.; Aslanidou, S.; Bassauer, S.; Hilcker, M.; Krugmann, A.; von Neumann-Cosel, P.; Otsuka, T.; Pietralla, N.; Ponomarev, V.Y.; Shimizu, N.; et al. First Measurement of Collectivity of Coexisting Shapes Based on Type II Shell Evolution: The Case of 96Zr. Phys. Rev. Lett. 2016, 117, 172503. [Google Scholar] [CrossRef]
  680. Witt, W.; Pietralla, N.; Werner, V.; Beck, T. Data on the structural coexistence in the 96Zr nucleus. Eur. Phys. J. A 2019, 55, 79. [Google Scholar] [CrossRef]
  681. Chakraborty, A.; Peters, E.E.; Crider, B.P.; Andreoiu, C.; Bender, P.C.; Cross, D.S.; Demand, G.A.; Garnsworthy, A.B.; Garrett, P.E.; Hackman, G.; et al. Collective Structure in 94Zr and Subshell Effects in Shape Coexistence. Phys. Rev. Lett. 2013, 110, 022504. [Google Scholar] [CrossRef]
  682. Paul, N.; Corsi, A.; Obertelli, A.; Doornenbal, P.; Authelet, G.; Baba, H.; Bally, B.; Bender, M.; Calvet, D.; Château, F.; et al. Are There Signatures of Harmonic Oscillator Shells Far from Stability? First Spectroscopy of 110Zr. Phys. Rev. Lett. 2017, 118, 032501. [Google Scholar]
  683. García-Ramos, J.E.; Heyde, K. On the nature of the shape coexistence and the quantum phase transition phenomena: Lead region and Zr isotopes. EPJ Web Conf. 2018, 178, 05005. [Google Scholar] [CrossRef]
  684. Wu, C.Y.; Hua, H.; Cline, D. Shape coexistence and their configuration mixing in 98Sr and 100Zr. Phys. Rev. C 2003, 68, 034322. [Google Scholar] [CrossRef]
  685. Sazonov, D.A.; Kolganova, E.A.; Shneidman, T.M.; Jolos, R.V.; Pietralla, N.; Witt, W. Description of shape coexistence in 96Zr based on the quadrupole-collective Bohr Hamiltonian. Phys. Rev. C 2019, 99, 031304. [Google Scholar] [CrossRef]
  686. Zheng, S.J.; Xu, F.R.; Shen, S.F.; Liu, H.L.; Wyss, R.; Yan, Y.P. Shape coexistence and triaxiality in nuclei near 80Zr. Phys. Rev. C 2014, 90, 064309. [Google Scholar] [CrossRef]
  687. Clément, E.; Zielińska, M.; Görgen, A.; Korten, W.; Péru, S.; Libert, J.; Goutte, H.; Hilaire, S.; Bastin, B.; Bauer, C.; et al. Spectroscopic Quadrupole Moments in 96,98Sr: Evidence for Shape Coexistence in Neutron-Rich Strontium Isotopes at N = 60. Phys. Rev. Lett. 2016, 116, 022701, Erratum in Phys. Rev. Lett. 2016 , 117, 099902. [Google Scholar] [CrossRef]
  688. Clément, E.; Zielińska, M. Shape coexistence in neutron-rich strontium isotopes at N = 60. Phys. Scr. 2017, 92, 084002. [Google Scholar] [CrossRef]
  689. Görgen, A.; Korten, W. Coulomb excitation studies of shape coexistence in atomic nuclei. J. Phys. G Nucl. Part Phys. 2016, 43, 024002. [Google Scholar] [CrossRef]
  690. Park, J.; Garnsworthy, A.B.; Krücken, R.; Andreoiu, C.; Ball, G.C.; Bender, P.C.; Chester, A.; Close, A.; Finlay, P.; Garrett, P.E.; et al. Shape coexistence and evolution in 98Sr. Phys. Rev. C 2016, 93, 014315. [Google Scholar] [CrossRef]
  691. Régis, J.-M.; Jolie, J.; Saed-Samii, N.; Warr, N.; Pfeiffer, M.; Blanc, A.; Jentschel, M.; Köster, U.; Mutti, P.; Soldner, T.; et al. Abrupt shape transition at neutron number N = 60: B(E2) values in 94,96,98Sr from fast γ-γ timing. Phys. Rev. C 2017, 95, 054319, Erratum in Phys. Rev. C 2017 , 95, 069902. [Google Scholar] [CrossRef]
  692. Schussler, F.; Pinston, J.A.; Monnand, E.; Moussa, A.; Jung, G.; Koglin, E.; Pfeiffer, B.; Janssens, R.V.F.; van Klinken, J. Discovery of a very low-lying 0+ state in 98Sr and shape coexistence implication in 98Sr. Nucl. Phys. A 1980, 339, 415. [Google Scholar] [CrossRef]
  693. Baktash, C.; Garcia-Bermudez, G.; Sarantites, D.G.; Nazarewicz, W.; Abenante, V.; Beene, J.R.; Griffin, H.C.; Halbert, M.L.; Hensley, D.C.; Johnson, N.R.; et al. Shape coexistence and disappearance of pairing correlations in 82Sr. Phys. Lett. B 1991, 255, 174. [Google Scholar] [CrossRef]
  694. Esmaylzadeh, A.; Karayonchev, V.; Nomura, K.; Jolie, J.; Beckers, M.; Blazhev, A.; Dewald, A.; Fransen, C.; Gerst, R.-B.; Häfner, G.; et al. Lifetime measurements to investigate γ softness and shape coexistence in 102Mo. Phys. Rev. C 2021, 104, 064314. [Google Scholar] [CrossRef]
  695. Ha, J.; Sumikama, T.; Browne, F.; Hinohara, N.; Bruce, A.M.; Choi, S.; Nishizuka, I.; Nishimura, S.; Doornenbal, P.; Lorusso, G.; et al. Shape evolution of neutron-rich 106,108,110Mo isotopes in the triaxial degree of freedom. Phys. Rev. C 2020, 101, 044311. [Google Scholar] [CrossRef]
  696. Garrett, P.E.; Makhathini, L.; Bark, R.A.; Rodríguez, T.R.; Valbuena, S.; Bildstein, V.; Bucher, T.D.; Burbadge, C.; Dubey, R.; Faestermann, T.; et al. Observation of the 0 2 + and γ bands in 98Ru, and shape coexistence in the Ru isotopes. Phys. Lett. B 2020, 809, 135762. [Google Scholar] [CrossRef]
  697. Becker, F.; Korten, W.; Hannachi, F.; Paris, P.; Buforn, N.; Chandler, C.; Houry, M.; Hübel, H.; Jansen, A.; Le Coz, Y.; et al. Investigation of prolate-oblate shape-coexistence in 74Kr. Eur. Phys. J. A 1999, 4, 103. [Google Scholar] [CrossRef]
  698. Briz, J.A.; Nácher, E.; Borge, M.J.G.; Algora, A.; Rubio, B.; Dessagne, P.; Maira, A.; Cano-Ott, D.; Courtin, S.; Escrig, D.; et al. Shape study of the N = Z nucleus 72Kr via β decay. Phys. Rev. C 2015, 92, 054326. [Google Scholar] [CrossRef]
  699. Clément, E.; Görgen, A.; Korten, W.; Bouchez, E.; Chatillon, A.; Delaroche, J.; Girod, M.; Goutte, H.; Huerstel, A.; Coz, Y.; et al. Shape coexistence in neutron-deficient krypton isotopes. Phys. Rev. C 2007, 75, 054313. [Google Scholar] [CrossRef]
  700. Dejbakhsh, H.; Cormier, T.M.; Zhao, X.; Ramayya, A.V.; Chaturvedi, L.; Zhu, S.; Kormicki, J.; Hamilton, J.H.; Satteson, M.; Lee, I.Y.; et al. Signature of the shape coexistence in 72Kr: Discontinuities of the moment of inertia at low spin. Phys. Lett. B 1990, 249, 195. [Google Scholar] [CrossRef]
  701. Korten, W.; Bouchez, E.; Clément, E.; Chatillon, A.; Görgen, A.; Le Coz, Y.; Theisen, C.; Wilson, J.; Casandjian, J.M.; de France, G.; et al. Shape coexistence in Krypton isotopes studied through Coulomb excitation of radioactive Krypton ion beams. Nucl. Phys. A 2004, 746, 90c. [Google Scholar] [CrossRef]
  702. Piercey, R.B.; Ramayya, A.V.; Hamilton, J.H.; Sun, X.J.; Zhao, Z.Z.; Robinson, R.L.; Kim, H.J.; Wells, J.C. Collective band structure to high spin and shape coexistence in 76Kr. Phys. Rev. C 1982, 25, 1941. [Google Scholar] [CrossRef]
  703. Varley, B.J.; Campbell, M.; Chishti, A.A.; Gelletly, W.; Goettig, L.; Lister, C.J.; James, A.N.; Skeppstedt, O. Evidence for shape coexistence in the N = Z nucleus Kr 36 36 72 . Phys. Lett. B 1987, 194, 463. [Google Scholar] [CrossRef]
  704. Wimmer, K.; Arici, T.; Korten, W.; Doornenbal, P.; Delaroche, J.-P.; Girod, M.; Libert, J.; Rodríguez, T.R.; Aguilera, P.; Algora, A.; et al. Shape coexistence revealed in the N = Z isotope 72Kr through inelastic scattering. Eur. Phys. J. A 2020, 56, 159. [Google Scholar] [CrossRef]
  705. Wimmer, K.; Korten, W.; Arici, T.; Doornenbal, P.; Aguilera, P.; Algora, A.; Ando, T.; Baba, H.; Blank, B.; Boso, A.; et al. Shape coexistence and isospin symmetry in A = 70 nuclei: Spectroscopy of the Tz=-1 nucleus 70Kr. Phys. Lett. B 2018, 785, 441. [Google Scholar] [CrossRef]
  706. Wimmer, K.; Korten, W.; Doornenbal, P.; Arici, T.; Aguilera, P.; Algora, A.; Ando, T.; Baba, H.; Blank, B.; Boso, A.; et al. Shape Changes in the Mirror Nuclei 70Kr and 70Se. Phys. Rev. Lett. 2021, 126, 072501. [Google Scholar] [CrossRef]
  707. Almehed, D.; Walet, N.R. Shape coexistence in 72Kr at finite angular momentum. Phys. Lett. B 2004, 604, 163. [Google Scholar] [CrossRef]
  708. Budaca, R.; Budaca, A.I. Coexistence, mixing and fluctuation of nuclear shapes. EPL 2018, 123, 42001. [Google Scholar] [CrossRef]
  709. Albers, M.; Warr, N.; Nomura, K.; Blazhev, A.; Jolie, J.; Mücher, D.; Bastin, B.; Bauer, C.; Bernards, C.; Bettermann, L.; et al. Evidence for a Smooth Onset of Deformation in the Neutron-Rich Kr Isotopes. Phys. Rev. Lett. 2012, 108, 062701, Erratum in Phys. Rev. Lett. 2012, 109, 209904; Phys. Rev. Lett. 2015, 114, 189902. [Google Scholar] [CrossRef]
  710. Dudouet, J.; Lemasson, A.; Duchêne, G.; Rejmund, M.; Clément, E.; Michelagnoli, C.; Didierjean, D.; Korichi, A.; Maquart, G.; Stezowski, O.; et al. Kr 60 36 96 –Low-Z Boundary of the Island of Deformation at N = 60. Phys. Rev. Lett. 2017, 118, 162501. [Google Scholar] [CrossRef]
  711. Flavigny, F.; Doornenbal, P.; Obertelli, A.; Delaroche, J.-P.; Girod, M.; Libert, J.; Rodriguez, T.R.; Authelet, G.; Baba, H.; Calvet, D.; et al. Shape Evolution in Neutron-Rich Krypton Isotopes Beyond N = 60: First Spectroscopy of 98,100Kr. Phys. Rev. Lett. 2017, 118, 242501. [Google Scholar] [CrossRef]
  712. Delafosse, C.; Verney, D.; Marević, P.; Gottardo, A.; Michelagnoli, C.; Lemasson, A.; Goasduff, A.; Ljungvall, J.; Clément, E.; Korichi, A.; et al. Pseudospin Symmetry and Microscopic Origin of Shape Coexistence in the 78Ni Region: A Hint from Lifetime Measurements. Phys. Rev. Lett. 2018, 121, 192502. [Google Scholar] [CrossRef]
  713. Hamilton, J.H.; Ramayya, A.V.; Pinkston, W.T.; Ronningen, R.M.; Garcia-Bermudez, G.; Carter, H.K.; Robinson, R.L.; Kim, H.J.; Sayer, R.O. Evidence for Coexistence of Spherical and Deformed Shapes in 72Se. Phys. Rev. Lett. 1974, 32, 239. [Google Scholar] [CrossRef]
  714. Hamilton, J.H.; Crowell, H.L.; Robinson, R.L.; Ramayya, A.V.; Collins, W.E.; Ronningen, R.M.; Maruhn-Rezwani, V.; Maruhn, J.A.; Singhal, N.C.; Kim, H.J.; et al. Lifetime Measurements to Test the Coexistence of Spherical and Deformed Shapes in 72Se. Phys. Rev. Lett. 1976, 36, 340. [Google Scholar] [CrossRef]
  715. Ljungvall, J.; Görgen, A.; Girod, M.; Delaroche, J.-P.; Dewald, A.; Dossat, C.; Farnea, E.; Korten, W.; Melon, B.; Menegazzo, R.; et al. Shape Coexistence in Light Se Isotopes: Evidence for Oblate Shapes. Phys. Rev. Lett. 2008, 100, 102502. [Google Scholar] [CrossRef] [PubMed]
  716. McCutchan, E.A.; Lister, C.J.; Ahn, T.; Casperson, R.J.; Heinz, A.; Ilie, G.; Qian, J.; Williams, E.; Winkler, R.; Werner, V. Shape coexistence in 72Se investigated following the β decay of 72Br. Phys. Rev. C 2011, 83, 024310. [Google Scholar] [CrossRef]
  717. Palit, R.; Jain, H.C.; Joshi, P.K.; Sheikh, J.A.; Sun, Y. Shape coexistence in 72Se. Phys. Rev. C 2001, 63, 024313. [Google Scholar] [CrossRef]
  718. Mukherjee, A.; Bhattacharya, S.; Trivedi, T.; Singh, R.P.; Muralithar, S.; Negi, D.; Palit, R.; Nag, S.; Rajbanshi, S.; Kumar Raju, M.; et al. Shape coexistence and octupole correlations in 72Se. Phys. Rev. C 2022, 105, 014322. [Google Scholar] [CrossRef]
  719. Cottle, P.D.; Holcomb, J.W.; Johnson, T.D.; Stuckey, K.A.; Tabor, S.L.; Womble, P.C.; Buccino, S.G.; Durham, F.E. Shape coexistence and octupole vibrations in 74Se. Phys. Rev. C 1990, 42, 1254. [Google Scholar] [CrossRef]
  720. McCutchan, E.A.; Lister, C.J.; Ahn, T.; Anagnostatou, V.; Cooper, N.; Elvers, M.; Goddard, P.; Heinz, A.; Ilie, G.; Radeck, D.; et al. Shape coexistence and high-K states in 74Se populated following the β decay of 74Br. Phys. Rev. C 2013, 87, 014307. [Google Scholar] [CrossRef]
  721. Fischer, S.M.; Balamuth, D.P.; Hausladen, P.A.; Lister, C.J.; Carpenter, M.P.; Seweryniak, D.; Schwartz, J. Evidence for Collective Oblate Rotation in N = Z 68Se. Phys. Rev. Lett. 2000, 84, 4064. [Google Scholar] [CrossRef]
  722. Jenkins, D.G.; Balamuth, D.P.; Carpenter, M.P.; Lister, C.J.; Fischer, S.M.; Clark, R.M.; Macchiavelli, A.O.; Fallon, P.; Svensson, C.E.; Kelsall, N.S.; et al. Stability of oblate shapes in the vicinity of N=Z=34 68Se: Bands in 69Se and 67As. Phys. Rev. C 2001, 64, 064311. [Google Scholar] [CrossRef]
  723. Obertelli, A.; Baugher, T.; Bazin, D.; Boissinot, S.; Delaroche, J.-P.; Dijon, A.; Flavigny, F.; Gade, A.; Girod, M.; Glasmacher, T.; et al. First spectroscopy of 66Se and 65As: Investigating shape coexistence beyond the N=Z line. Phys. Lett. B 2011, 701, 417. [Google Scholar] [CrossRef]
  724. Gupta, J.B.; Hamilton, J.H. Empirical study of the shape evolution and shape coexistence in Zn, Ge and Se isotopes. Nucl. Phys. A 2019, 983, 20. [Google Scholar] [CrossRef]
  725. Gratchev, I.N.; Simpson, G.S.; Thiamova, G.; Ramdhane, M.; Sieja, K.; Blanc, A.; Jentschel, M.; Köster, U.; Mutti, P.; Soldner, T.; et al. Identification of excited states and collectivity in 88Se. Phys. Rev. C 2017, 95, 051302. [Google Scholar] [CrossRef]
  726. Lizarazo, C.; Söderström, P.-A.; Werner, V.; Pietralla, N.; Walker, P.; Dong, G.; Xu, F.; Rodríguez, T.; Browne, F.; Doornenbal, P.; et al. Metastable States of 92,94Se: Identification of an Oblate K Isomer of 94Se and the Ground-State Shape Transition between N = 58 and 60. Phys. Rev. Lett. 2020, 124, 222501. [Google Scholar] [CrossRef] [PubMed]
  727. Chen, S.; Doornenbal, P.; Obertelli, A.; Rodríguez, T.R.; Authelet, G.; Baba, H.; Calvet, D.; Browne, F.; Bruce, A.; Nobs, C. Low-lying structure and shape evolution in neutron-rich Se isotopes. Phys. Rev. C 2017, 95, 041302. [Google Scholar] [CrossRef]
  728. Lecomte, R.; Kajrys, G.; Landsberger, S.; Paradis, P.; Monaro, S. Shape coexistence and shape transitions in the even-A Ge nuclei. Phys. Rev. C 1982, 25, 2812, Erratum in Phys. Rev. C 1982, 26, 1334. [Google Scholar] [CrossRef]
  729. Ayangeakaa, A.D.; Janssens, R.V.F.; Wu, C.Y.; Allmond, J.M.; Wood, J.L.; Zhu, S.; Albers, M.; Almaraz-Calderon, S.; Bucher, B.; Carpenter, M.P.; et al. Shape coexistence and the role of axial asymmetry in 72Ge. Phys. Lett. B 2016, 754, 254. [Google Scholar] [CrossRef]
  730. Turkan, N.; Maras, I. E(5) behaviour of the Ge isotopes. Math. Comput. Appl. 2010, 15, 428. [Google Scholar] [CrossRef]
  731. Hwang, J.K.; Hamilton, J.H.; Ramayya, A.V.; Brewer, N.T.; Luo, Y.X.; Rasmussen, J.O.; Zhu, S.J. Possible excited deformed rotational bands in 82Ge. Phys. Rev. C 2011, 84, 024305. [Google Scholar] [CrossRef]
  732. Ahn, S.; Bardayan, D.W.; Jones, K.L.; Adekola, A.S.; Arbanas, G.; Blackmon, J.C.; Chae, K.Y.; Chipps, K.A.; Cizewski, J.A.; Hardy, S.; et al. Direct neutron capture cross section on 80Ge and probing shape coexistence in neutron-rich nuclei. Phys. Rev. C 2019, 100, 044613. [Google Scholar] [CrossRef]
  733. Gottardo, A.; Verney, D.; Delafosse, C.; Ibrahim, F.; Roussière, B.; Sotty, C.; Roccia, S.; Andreoiu, C.; Costache, C.; Delattre, M.C.; et al. First Evidence of Shape Coexistence in the 78Ni Region: Intruder 0 2 + State in 80Ge. Phys. Rev. Lett. 2016, 116, 182501. [Google Scholar] [CrossRef]
  734. Garcia, F.H.; Andreoiu, C.; Ball, G.C.; Bell, A.; Garnsworthy, A.B.; Nowacki, F.; Petrache, C.M.; Poves, A.; Whitmore, K.; Ali, F.A.; et al. Absence of Low-Energy Shape Coexistence in 80Ge: The Nonobservation of a Proposed Excited 0 2 + Level at 639 keV. Phys. Rev. Lett. 2020, 125, 172501. [Google Scholar] [CrossRef]
  735. Koizumi, M.; Seki, A.; Toh, Y.; Osa, A.; Utsuno, Y.; Kimura, A.; Oshima, M.; Hayakawa, T.; Hatsukawa, Y.; Katakura, J.; et al. Multiple Coulomb excitation experiment of 68Zn. Nucl. Phys. A 2004, 730, 46. [Google Scholar] [CrossRef]
  736. Rocchini, M.; Hadyńska-Klȩk, K.; Nannini, A.; Goasduff, A.; Zielińska, M.; Testov, D.; Rodríguez, T.R.; Gargano, A.; Nowacki, F.; Gregorio, G.D.; et al. Onset of triaxial deformation in 66Zn and properties of its first excited 0+ state studied by means of Coulomb excitation. Phys. Rev. C 2021, 103, 014311. [Google Scholar] [CrossRef]
  737. Shiga, Y.; Yoneda, K.; Steppenbeck, D.; Aoi, N.; Doornenbal, P.; Lee, J.; Liu, H.; Matsushita, M.; Takeuchi, S.; Wang, H.; et al. Investigating nuclear shell structure in the vicinity of 78Ni: Low-lying excited states in the neutron-rich isotopes 80,82Zn. Phys. Rev. C 2016, 93, 024320, Erratum in Phys. Rev. C 2016, 93, 049904. [Google Scholar] [CrossRef]
  738. Van de Walle, J.; Aksouh, F.; Behrens, T.; Bildstein, V.; Blazhev, A.; Cederkäll, J.; Clément, E.; Cocolios, T.E.; Davinson, T.; Delahaye, P.; et al. Low-energy Coulomb excitation of neutron-rich zinc isotopes. Phys. Rev. C 2009, 79, 014309. [Google Scholar] [CrossRef]
  739. Orlandi, R.; Mücher, D.; Raabe, R.; Jungclaus, A.; Pain, S.D.; Bildstein, V.; Chapman, R.; de Angelis, G.; Johansen, J.G.; Van Duppen, P.; et al. Single-neutron orbits near 78Ni: Spectroscopy of the isotope 79Zn. Phys. Lett. B 2015, 740, 298. [Google Scholar] [CrossRef]
  740. Yang, X.F.; Wraith, C.; Xie, L.; Babcock, C.; Billowes, J.; Bissell, M.L.; Blaum, K.; Cheal, B.; Flanagan, K.T.; Garcia Ruiz, R.F.; et al. Isomer Shift and Magnetic Moment of the Long-Lived 1/2+ Isomer in Zn 49 30 79 : Signature of Shape Coexistence near 78Ni. Phys. Rev. Lett. 2016, 116, 182502. [Google Scholar] [CrossRef] [PubMed]
  741. Chiara, C.J.; Broda, R.; Walters, W.B.; Janssens, R.V.F.; Albers, M.; Alcorta, M.; Bertone, P.F.; Carpenter, M.P.; Hoffman, C.R.; Lauritsen, T.; et al. Low-spin states and the non-observation of a proposed 2202-keV, 0+ isomer in 68Ni. Phys. Rev. C 2012, 86, 041304. [Google Scholar] [CrossRef]
  742. Crider, B.P.; Prokop, C.J.; Liddick, S.N.; Al-Shudifat, M.; Ayangeakaa, A.D.; Carpenter, M.P.; Carroll, J.J.; Chen, J.; Chiara, C.J.; David, H.M.; et al. Shape coexistence from lifetime and branching-ratio measurements in 68,70Ni. Phys. Lett. B 2016, 763, 108. [Google Scholar] [CrossRef]
  743. Dijon, A.; Clément, E.; France, G.D.; Angelis, G.D.; Duchêne, G.; Dudouet, J.; Franchoo, S.; Gadea, A.; Gottardo, A.; Hüyük, T.; et al. Discovery of a new isomeric state in 68Ni: Evidence for a highly deformed proton intruder state. Phys. Rev. C 2012, 85, 031301. [Google Scholar] [CrossRef]
  744. Flavigny, F.; Pauwels, D.; Radulov, D.; Darby, I.J.; de Witte, H.; Diriken, J.; Fedorov, D.V.; Fedosseev, V.N.; Prieto, F.; Mario, L.; et al. Characterization of the low-lying 0+ and 2+ states in 68Ni via β decay of the low-spin 68Co isomer. Phys. Rev. C 2015, 91, 034310. [Google Scholar] [CrossRef]
  745. Flavigny, F.; Elseviers, J.; Andreyev, A.; Bauer, C.; Bildstein, V.; Blazhev, A.; Brown, B.A.; Witte, H.; Diriken, J.; Fedosseev, V.; et al. Microscopic structure of coexisting 0+ states in 68Ni probed via two-neutron transfer. Phys. Rev. C 2019, 99, 054332. [Google Scholar] [CrossRef]
  746. Recchia, F.; Chiara, C.J.; Janssens, R.V.F.; Weisshaar, D.; Gade, A.; Walters, W.B.; Albers, M.; Alcorta, M.; Bader, V.M.; Baugher, T.; et al. Configuration mixing and relative transition rates between low-spin states in 68Ni. Phys. Rev. C 2013, 88, 041302. [Google Scholar] [CrossRef]
  747. Suchyta, S.; Liddick, S.N.; Tsunoda, Y.; Otsuka, T.; Bennett, M.B.; Chemey, A.; Honma, M.; Larson, N.; Prokop, C.J.; Quinn, S.J.; et al. Shape coexistence in 68Ni. Phys. Rev. C 2014, 89, 021301. [Google Scholar] [CrossRef]
  748. Chiara, C.; Weisshaar, D.; Janssens, R.; Tsunoda, Y.; Otsuka, T.; Harker, J.; Walters, W.; Recchia, F.; Albers, M.; Alcorta, M.; et al. Identification of deformed intruder states in semi-magic 70Ni. Phys. Rev. C 2015, 91, 044309. [Google Scholar] [CrossRef]
  749. Prokop, C.J.; Crider, B.P.; Liddick, S.N.; Ayangeakaa, A.D.; Carpenter, M.P.; Carroll, J.J.; Chen, J.; Chiara, C.J.; David, H.M.; Dombos, A.C.; et al. New low-energy 0+ state and shape coexistence in 70Ni. Phys. Rev. C 2015, 92, 061302, Erratum in Phys. Rev. C 2016, 93, 019901(E). [Google Scholar] [CrossRef]
  750. Broda, R.; Pawłat, T.; Królas, W.; Janssens, R.V.F.; Zhu, S.; Walters, W.B.; Fornal, B.; Chiara, C.J.; Carpenter, M.P.; Hoteling, N.; et al. Spectroscopic study of the 64,66,68Ni isotopes populated in 64Ni + 238U collisions. Phys. Rev. C 2012, 86, 064312. [Google Scholar] [CrossRef]
  751. Leoni, S.; Fornal, B.; Mărginean, N.; Sferrazza, M.; Tsunoda, Y.; Otsuka, T.; Bocchi, G.; Crespi, F.C.L.; Bracco, A.; Aydin, S.; et al. Multifaceted Quadruplet of Low-Lying Spin-Zero States in 66Ni: Emergence of Shape Isomerism in Light Nuclei. Phys. Rev. Lett. 2017, 118, 162502. [Google Scholar] [CrossRef]
  752. Olaizola, B.; Fraile, L.M.; Mach, H.; Poves, A.; Nowacki, F.; Aprahamian, A.; Briz, J.A.; Cal-González, J.; Ghiţa, D.; Köster, U.; et al. Search for shape-coexisting 0+ states in 66Ni from lifetime measurements. Phys. Rev. C 2017, 95, 061303. [Google Scholar] [CrossRef]
  753. Mărginean, N.; Little, D.; Tsunoda, Y.; Leoni, S.; Janssens, R.V.F.; Fornal, B.; Otsuka, T.; Michelagnoli, C.; Stan, L.; Crespi, F.C.L.; et al. Shape Coexistence at Zero Spin in 64Ni Driven by the Monopole Tensor Interaction. Phys. Rev. Lett. 2020, 125, 102502. [Google Scholar] [CrossRef] [PubMed]
  754. Gade, A.; Liddick, S.N. Shape coexistence in neutron-rich nuclei. J. Phys. G Nucl. Part. Phys. 2016, 43, 024001. [Google Scholar] [CrossRef]
  755. Olivier, L.; Franchoo, S.; Niikura, M.; Vajta, Z.; Sohler, D.; Doornenbal, P.; Obertelli, A.; Tsunoda, Y.; Otsuka, T.; Authelet, G.; et al. Persistence of the Z = 28 Shell Gap Around 78Ni: First Spectroscopy of 79Cu. Phys. Rev. Lett. 2017, 119, 192501. [Google Scholar] [CrossRef] [PubMed]
  756. Welker, A.; Althubiti, N.A.S.; Atanasov, D.; Blaum, K.; Cocolios, T.E.; Herfurth, F.; Kreim, S.; Lunney, D.; Manea, V.; Mougeot, M.; et al. Binding Energy of 79Cu: Probing the Structure of the Doubly Magic 78Ni from Only One Proton Away. Phys. Rev. Lett. 2017, 119, 192502. [Google Scholar] [CrossRef] [PubMed]
  757. Porquet, M.-G.; Sorlin, O. Evolution of the N = 50 gap from Z = 30 to Z = 38 and extrapolation toward 78Ni. Phys. Rev. C 2012, 85, 014307. [Google Scholar] [CrossRef]
  758. Klintefjord, M.; Ljungvall, J.; Görgen, A.; Lenzi, S.M. Measurement of lifetimes in 62,64Fe, 61,63Co, and 59Mn. Phys. Rev. C 2017, 95, 024312. [Google Scholar] [CrossRef]
  759. Lunardi, S.; Lenzi, S.M.; Della Vedova, F.; Farnea, E.; Gadea, A.; Mărginean, N.; Bazzacco, D.; Beghini, S.; Bizzeti, P.G.; Bizzeti-Sona, A.M.; et al. Spectroscopy of neutron-rich Fe isotopes populated in the 64Ni+238U reaction. Phys. Rev. C 2007, 76, 034303. [Google Scholar] [CrossRef]
  760. Rother, W.; Dewald, A.; Iwasaki, H.; Lenzi, S.M.; Starosta, K.; Bazin, D.; Baugher, T.; Brown, B.A.; Crawford, H.L.; Fransen, C.; et al. Enhanced Quadrupole Collectivity at N = 40: The Case of Neutron-Rich Fe Isotopes. Phys. Rev. Lett. 2011, 106, 022502. [Google Scholar] [CrossRef]
  761. Olaizola, B.; Fraile, L.M.; Mach, H.; Poves, A.; Aprahamian, A.; Briz1, J.A.; Cal-González, J.; Ghiţa, D.; Köster, U.; Kurcewicz, W.; et al. Beta decay of 66Mn to the N = 40 nucleus 66Fe. J. Phys. G Nucl. Part. Phys. 2017, 44, 125103. [Google Scholar] [CrossRef]
  762. Liddick, S.N.; Abromeit, B.; Ayres, A.; Bey, A.; Bingham, C.R.; Brown, B.A.; Cartegni, L.; Crawford, H.L.; Darby, I.G.; Grzywacz, R.; et al. Low-energy level schemes of 66,68Fe and inferred proton and neutron excitations across Z = 28 and N = 40. Phys. Rev. C 2013, 87, 014325. [Google Scholar] [CrossRef]
  763. Arnswald, K.; Braunroth, T.; Seidlitz, M.; Coraggio, L.; Reiter, P.; Birkenbach, B.; Blazhev, A.; Dewald, A.; Fransen, C.; Fu, B.; et al. Enhanced collectivity along the N = Z line: Lifetime measurements in 44Ti, 48Cr, and 52Fe. Phys. Lett. B 2017, 772, 599. [Google Scholar] [CrossRef]
  764. Carpenter, M.P.; Janssens, R.V.F.; Zhu, S. Shape coexistence in neutron-rich nuclei near N = 40. Phys. Rev. C 2013, 87, 041305. [Google Scholar] [CrossRef]
  765. Liddick, S.N.; Suchyta, S.; Abromeit, B.; Ayres, A.; Bey, A.; Bingham, C.R.; Bolla1, M.; Carpenter, M.P.; Cartegni, L.; Chiara, C.J.; et al. Shape coexistence along N = 40. Phys. Rev. C 2011, 84, 061305. [Google Scholar] [CrossRef]
  766. Kumar, R.; Chamoli, S.K.; Govil, I.M.; Dhal, A.; Sinha, R.K.; Chaturvedi, L.; Naik, Z.; Praharaj, C.R.; Muralithar, S.; Singh, R.P.; et al. Shape coexistence and high spin states in 52Cr. Phys. Rev. C 2007, 76, 034301. [Google Scholar] [CrossRef]
  767. Rowe, D.J.; Wood, J.L. A relationship between isobaric analog states and shape coexistence in nuclei. J. Phys. G Nucl. Part. Phys. 2018, 45, 06LT01. [Google Scholar] [CrossRef]
  768. O’Leary, C.D.; Bentley, M.A.; Brown, B.A.; Appelbe, D.E.; Bark, R.A.; Cullen, D.M.; Ertürk, S.; Maj, A.; Merchant, A.C. Nonyrast high-spin states in N = Z 44Ti. Phys. Rev. C 2000, 61, 064314. [Google Scholar] [CrossRef]
  769. Schielke, S.; Hohn, D.; Speidel, K.-H.; Kenn, O.; Leske, J.; Gemein, N.; Offer, M.; Gerber, J.; Maier-Komor, P.; Zell, O.; et al. Evidence for 40Ca core excitation from g factor and B(E2) measurements on the 2 1 + states of 42,44Ca. Phys. Lett. B 2003, 571, 29. [Google Scholar] [CrossRef]
  770. Simpson, J.J.; Dixon, W.R.; Storey, R.S. Evidence for Rotational Bands in 44Ti. Phys. Rev. Lett. 1973, 31, 946. [Google Scholar] [CrossRef]
  771. Schielke, S.; Speidel, K.-H.; Kenn, O.; Leske, J.; Gemein, N.; Offer, M.; Sharon, Y.Y.; Zamick, L.; Gerber, J.; Maier-Komor, P. First measurement and shell model interpretation of the g factor of the 2 1 + state in self-conjugate radioactive 44Ti. Phys. Lett. B 2003, 567, 153. [Google Scholar] [CrossRef]
  772. Middleton, R.; Garrett, J.D.; Fortune, H.T. Search for multiparticle-multihole states of 40Ca with the 32S(12C,α) reaction. Phys. Lett. B 1972, 39, 339. [Google Scholar] [CrossRef]
  773. Ellegaard, C.; Lien, J.R.; Nathan, O.; Sletten, G.; Ingebretsen, F.; Osnes, E.; Tjøm, P.O.; Hansen, O.; Stock, R. The (1f)2 multiplet in 42Ca. Phys. Lett. B 1972, 40, 641. [Google Scholar] [CrossRef]
  774. Hadyńska-Klȩk, K.; Napiorkowski, P.J.; Zielinska, M.; Srebrny, J.; Maj, A.; Azaiez, F.; aliente Dobon, J.J.; Kicinska-Habior, M.; Nowacki, F.; Naïdja, H.; et al. Superdeformed and Triaxial States in 42Ca. Phys. Rev. Lett. 2016, 117, 062501. [Google Scholar] [CrossRef]
  775. Hadyńska-Klȩk, K.; Napiorkowski, P.J.; Zielinska, M.; Srebrny, J.; Maj, A.; Azaiez, F.; aliente Dobon, J.J.; Kicinska-Habior, M.; Nowacki, F.; Naïdja, H.; et al. Quadrupole collectivity in 42Ca from low-energy Coulomb excitation with AGATA. Phys. Rev. C 2018, 97, 024326. [Google Scholar] [CrossRef]
  776. Flynn, E.R.; Hansen, O.; Casten, R.F.; Garrett, J.D.; Ajzenberg-Selove, F. The (t, p) reaction on 36, 38, 40Ar. Nucl. Phys. A 1975, 246, 117. [Google Scholar] [CrossRef]
  777. Ideguchi, E.; Ota, S.; Morikawa, T.; Oshima, M.; Koizumi, M.; Toh, Y.; Kimura, A.; Harada, H.; Furutaka, K.; Nakamura, S.; et al. Superdeformation in asymmetric N>Z nucleus 40Ar. Phys. Lett. B 2010, 686, 18. [Google Scholar] [CrossRef]
  778. Speidel, K.-H.; Schielke, S.; Leske, J.; Gerber, J.; Maier-Komor, P.; Robisnon, S.J.O.; Sharon, Y.Y.; Zamick, L. Experimental g factors and B(E2) values in Ar isotopes: Crossing the N = 20 semi-magic divide. Phys. Lett. B 2006, 632, 207. [Google Scholar] [CrossRef]
  779. Stefanova, E.A.; Benczer-Koller, N.; Kumbartzki, G.J.; Sharon, Y.Y.; Zamick, L.; Robinson, S.J.Q.; Bernstein, L.; Cooper, J.R.; Judson, D.; Taylor, M.J.; et al. Near spherical shell-model structure of the 2 + 1 state in 40Ar from g-factor measurements. Phys. Rev. C 2005, 72, 014309. [Google Scholar] [CrossRef]
  780. Svensson, C.E.; Macchiavelli, A.O.; Juodagalvis, A.; Poves, A.; Ragnarsson, I.; Åberg, S.; Appelbe, D.E.; Austin, R.A.E.; Baktash, C.; Ball, G.C.; et al. Superdeformation in the N = Z Nucleus 36Ar: Experimental, Deformed Mean Field, and Spherical Shell Model Descriptions. Phys. Rev. Lett. 2000, 85, 2693. [Google Scholar] [CrossRef]
  781. Meisel, Z.; George, S.; Ahn, S.; Browne, J.; Bazin, D.; Brown, B.A.; Carpino, J.F.; Chung, H.; Cyburt, R.H.; Estradé, A.; et al. Mass Measurements Demonstrate a Strong N = 28 Shell Gap in Argon. Phys. Rev. Lett. 2015, 114, 022501. [Google Scholar] [CrossRef]
  782. Cottle, P.D.; Kemper, K.W. Persistence of the N = 28 shell closure in neutron-rich nuclei. Phys. Rev. C 1998, 58, 3761. [Google Scholar] [CrossRef]
  783. Force, C.; Grévy, S.; Gaudefroy, L.; Sorlin, O.; Cáceres, L.; Rotaru, F.; Mrazek, J.; Achouri, N.L.; Angélique, J.C.; Azaiez, F.; et al. Prolate-Spherical Shape Coexistence at N = 28 in 44S. Phys. Rev. Lett. 2010, 105, 102501. [Google Scholar] [CrossRef]
  784. Longfellow, B.; Weisshaar, D.; Gade, A.; Brown, B.A.; Bazin, D.; Brown, K.W.; Elman, B.; Pereira, J.; Rhodes, D.; Spieker, M.; et al. Shape Changes in the N = 28 Island of Inversion: Collective Structures Built on Configuration-Coexisting States in 43S. Phys. Rev. Lett. 2020, 125, 232501. [Google Scholar] [CrossRef] [PubMed]
  785. Sarazin, F.; Savajols, H.; Mittig, W.; Nowacki, F.; Orr, N.A.; Ren, Z.; Roussel-Chomaz, P.; Auger, G.; Baiborodin, D.; Belozyorov, A.V.; et al. Shape Coexistence and the N = 28 Shell Closure Far from Stability. Phys. Rev. Lett. 2000, 84, 5062. [Google Scholar] [CrossRef] [PubMed]
  786. Mittig, W.; Savajols, H.; Baiborodin, D.; Casandjian, J.M.; Demonchy, C.E.; Roussel-Chomaz, P.; Sarazin, F.; Dlouhý, Z.; Mrazek, J.; Belozyorov, A.V.; et al. Shape coexistence and the N = 20 shell closure far from stability by inelastic scattering. Eur. Phys. J. A 2002, 15, 157. [Google Scholar] [CrossRef]
  787. Rotaru, F.; Negoita, F.; Grévy, S.; Mrazek, J.; Lukyanov, S.; Owacki, F.; Poves, A.; Sorlin, O.; Borcea, C.; Borcea, R.; et al. Unveiling the Intruder Deformed 0 2 + State in 34Si. Phys. Rev. Lett. 2012, 109, 092503. [Google Scholar] [CrossRef] [PubMed]
  788. Macchiavelli, A.O.; Crawford, H.L.; Campbell, C.M.; Clark, R.M.; Cromaz, M.; Fallon, P.; Lee, I.Y.; Gade, A.; Poves, A.; Rice, E.; et al. Structure of 43P and 42Si in a two-level shape-coexistence model. Phys. Rev. C 2022, 105, 014309. [Google Scholar] [CrossRef]
  789. Ragnarsson, I.; Åberg, S. Shape coexistence and high-spin states in 28Si. Phys. Lett. B 1982, 114, 387. [Google Scholar] [CrossRef]
  790. Sheline, R.K.; Kubono, S.; Morita, K.; Tanaka, M.H. Coexistence of three different shapes in 28Si. Phys. Lett. B 1982, 119, 263. [Google Scholar] [CrossRef]
  791. Kitamura, N.; Wimmer, K.; Shimizu, N.; Bader, V.M.; Bancroft, C.; Barofsky, D.; Baugher, T.; Bazin, D.; Berryman, J.S.; Bildstein, V.; et al. Structure of 30Mg explored via in-beam γ-ray spectroscopy. Phys. Rev. C 2020, 102, 054318. [Google Scholar] [CrossRef]
  792. Nishibata, H.; Tajiri, K.; Shimoda, T.; Odahara, A.; Morimoto, S.; Kanaya, S.; Yagi, A.; Kanaoka, H.; Pearson, M.R.; Levy, C.D.P.; et al. Structure of the neutron-rich nucleus 30Mg. Phys. Rev. C 2020, 102, 054327. [Google Scholar] [CrossRef]
  793. Schwerdtfeger, W.; Thirolf, P.G.; Wimmer, K.; Habs, D.; Mach, H.; Rodriguez, T.R.; Bildstein, V.; Egido, J.L.; Fraile, L.M.; Gernhäuser, R.; et al. Shape Coexistence Near Neutron Number N = 20: First Identification of the E0 Decay from the Deformed First Excited Jπ=0+ State in 30Mg. Phys. Rev. Lett. 2009, 103, 012501. [Google Scholar] [CrossRef] [PubMed]
  794. Dowie, J.T.H.; Kibedi, T.; Jenkins, D.J.; Stuchbery, A.E.; Akber, A.; Alshammari, H.; Aoi, N.; Abaraham, A.; Bignell, L.; Vernon Chisapi, M.V.; et al. Evidence for shape coexistence and superdeformation in 24Mg. Phys. Lett. B 2020, 811, 135855. [Google Scholar] [CrossRef]
  795. Murray, I.; MacCormick, M.; Bazin, D.; Doornenbal, P.; Aoi, N.; Baba, H.; Crawford, H.; Fallon, P.; Li, K.; Lee, J.; et al. Spectroscopy of strongly deformed 32Ne by proton knockout reactions. Phys. Rev. C 2019, 99, 011302. [Google Scholar] [CrossRef]
  796. Dombrádi, Z.; Elekes, Z.; Saito, A.; Aoi, N.; Baba, H.; Demichi, K.; Fülöp, 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]
  797. Brown, G.E.; Green, A.M. Even parity states of 16O and 17O. Nucl. Phys. 1966, 75, 401. [Google Scholar] [CrossRef]
  798. Brown, G.E.; Green, A.M. Nuclear coexistence in the oxygen region and realistic nucleon-nucleon forces. Nucl. Phys. 1966, 85, 87. [Google Scholar] [CrossRef]
  799. Unna, I.; Talmi, I. Energies of Ground and Excited Nuclear Configurations in the First p1/2 Region. Phys. Rev. 1958, 112, 452. [Google Scholar] [CrossRef]
  800. Franzini, P.; Radicati, L.A. On the validity of the supermultiplet model. Phys. Lett. 1963, 6, 322–324. [Google Scholar] [CrossRef]
  801. Hecht, K.T.; Pang, S.C. On the Wigner Supermultiplet Scheme. J. Math. Phys. 1969, 10, 1571. [Google Scholar] [CrossRef]
  802. Van Isacker, P.; Pittel, S. Symmetries and deformations in the spherical shell model. Phys. Scr. 2016, 91, 023009. [Google Scholar] [CrossRef]
  803. Wigner, E. On the Consequences of the Symmetry of the Nuclear Hamiltonian on the Spectroscopy of Nuclei. Phys. Rev. 1937, 51, 106. [Google Scholar] [CrossRef]
  804. Rowe, D.J.; Thiamova, G.; Wood, J.L. Implications of Deformation and Shape Coexistence for the Nuclear Shell Model. Phys. Rev. Lett. 2006, 97, 202501. [Google Scholar] [CrossRef] [PubMed]
  805. Hoyle, F. On Nuclear Reactions Occurring in Very Hot STARS. I. the Synthesis of Elements from Carbon to Nickel. Astrophys. J. Suppl. 1954, 1, 121. [Google Scholar]
  806. Fynbo, H.O.U.; Freer, M. Viewpoint: Rotations of the Hoyle state in 12C. Physics 2011, 4, 94. [Google Scholar] [CrossRef]
  807. Itoh, M.; Akimune, H.; Fujiwara, M.; Garg, U.; Hashimoto, N.; Kawabata, T.; Kawase, K.; Kishi, S.; Murakami, T.; Nakanishi, K.; et al. Candidate for the 2+ excited Hoyle state at Ex∼10 MeV in 12C. Phys. Rev. C 2011, 84, 054308. [Google Scholar] [CrossRef]
  808. Freer, M.; Fynbo, H.O.U. The Hoyle state in 12C. Prog. Part. Nucl. Phys. 2014, 78, 1. [Google Scholar] [CrossRef]
  809. Tohsaki, A.; Horiuchi, H.; Schuck, P.; Röpke, G. Colloquium: Status of α-particle condensate structure of the Hoyle state. Rev. Mod. Phys. 2017, 89, 011002. [Google Scholar] [CrossRef]
  810. Freer, M.; Horiuchi, H.; Kanada-En’yo, Y.; Lee, D.; Meissner, U.-G. Microscopic clustering in light nuclei. Rev. Mod. Phys. 2018, 90, 035004. [Google Scholar] [CrossRef]
  811. Takemoto, H.; Myo, T.; Horiuchi, H.; Toki, H.; Isaka, M.; Lyu, M.; Zhao, Q.; Wan, N. Appearance of the Hoyle state and its breathing mode in 12C despite strong short-range repulsion of the nucleon-nucleon potential. Phys. Rev. C 2023, 107, 044304. [Google Scholar] [CrossRef]
  812. Iwasaki, H.; Motobayashi, T.; Akiyoshi, H.; Ando, Y.; Fukuda, N.; Fujiwara, H.; Fülöp, Z.; Hahn, K.I.; Higurashi, Y.; Hirai, M.; et al. Low-lying intruder 1 state in 12Be and the melting of the N = 8 shell closure. Phys. Lett. B 2000, 491, 8. [Google Scholar] [CrossRef]
  813. Navin, A.; Anthony, D.W.; Aumann, T.; Baumann, T.; Bazin, D.; Blumenfeld, Y.; Brown, B.A.; Glasmacher, T.; Hansen, P.G.; Ibbotson, R.W.; et al. Direct Evidence for the Breakdown of the N = 8 Shell Closure in 12Be. Phys. Rev. Lett. 2000, 85, 266. [Google Scholar] [CrossRef]
  814. Lyu, M.; Yoshida, K.; Kanada-En’yo, Y.; Ogata, K. Direct probing of the cluster structure in 12Be via the α-knockout reaction. Phys. Rev. C 2019, 99, 064610. [Google Scholar] [CrossRef]
  815. Van Isacker, P.; Warner, D.D.; Brenner, D.S. Test of Wigner’s Spin-Isospin Symmetry from Double Binding Energy Differences. Phys. Rev. Lett. 1995, 74, 4607. [Google Scholar] [CrossRef]
  816. Kota, V.K.B.; Sahu, R. Structure of Medium Mass Nuclei: Deformed Shell Model and Spin-Isospin Interacting Boson Model; CRC Press: Boca Raton, FL, USA, 2017. [Google Scholar]
  817. Cederwall, B.; Moradi, F.G.; Bäck, T.; Johnson, A.; Blomqvist, J.; Clément, E.; de France, G.; Wadsworth, R.; Andgren, K.; Lagergren, K.; et al. Evidence for a spin-aligned neutron–proton paired phase from the level structure of 92Pd. Nature 2011, 469, 68. [Google Scholar] [CrossRef] [PubMed]
  818. Qi, C.; Blomqvist, J.; Bäck, T.; Cederwall, B.; Johnson, A.; Liotta, R.J.; Wyss, R. Spin-aligned neutron-proton pair mode in atomic nuclei. Phys. Rev. C 2011, 84, 021301. [Google Scholar] [CrossRef]
  819. Qi, C.; Blomqvist, J.; Bäck, T.; Cederwall, B.; Johnson, A.; Liotta, R.J.; Wyss, R. Coherence features of the spin-aligned neutron–proton pair coupling scheme. Phys. Scr. 2012, T150, 014031. [Google Scholar] [CrossRef]
  820. Qi, C.; Wyss, R. N = Z nuclei: A laboratory for neutron–proton collective mode. Phys. Scr. 2016, 91, 013009. [Google Scholar] [CrossRef]
  821. Robinson, S.J.Q.; Fan, C.; Harper, M.; Zamick, L. On the vibrational model of 92Pd and comparison with 48Cr. Int. J. Mod. Phys. E 2021, 30, 2150047. [Google Scholar] [CrossRef]
  822. Xu, Z.X.; Qi, C.; Blomqvist, J.; Liotta, R.J.; Wyss, R. Multistep shell model description of spin-aligned neutron–proton pair coupling. Nucl. Phys. A 2012, 877, 51. [Google Scholar] [CrossRef]
  823. Sambataro, M.; Sandulescu, N. Four-Body Correlations in Nuclei. Phys. Rev. Lett. 2015, 115, 112501. [Google Scholar] [CrossRef]
  824. Sambataro, M.; Sandulescu, N. Quarteting and spin-aligned proton-neutron pairs in heavy N = Z nuclei. Phys. Rev. C 2015, 91, 064318. [Google Scholar] [CrossRef]
  825. Zerguine, S.; Van Isacker, P. Spin-aligned neutron-proton pairs in N = Z nuclei. Phys. Rev. C 2011, 83, 064314. [Google Scholar] [CrossRef]
  826. Frauendorf, S.; Macchiavelli, A.O. Overview of neutron–proton pairing. Prog. Part. Nucl. Phys. 2014, 78, 24. [Google Scholar] [CrossRef]
  827. Cederwall, B.; Liu, X.; Aktas, O.; Ertoprak, A.; Zhang, W.; Qi, C.; Clement, E.; De France, G.; Ralet, D.; Gadea, A.; et al. Isospin Properties of Nuclear Pair Correlations from the Level Structure of the Self-Conjugate Nucleus 88Ru. Phys. Rev. Lett. 2020, 124, 062501. [Google Scholar] [CrossRef] [PubMed]
  828. Sambataro, M.; Sandulescu, N.; Johnson, C.W. Isoscalar and isovector pairing in a formalism of quartets. Phys. Lett. B 2015, 740, 137. [Google Scholar] [CrossRef]
  829. Sambataro, M.; Sandulescu, N. Isoscalar-isovector proton-neutron pairing and quartet condensation in N = Z nuclei. Phys. Rev. C 2016, 93, 054320. [Google Scholar] [CrossRef]
  830. Sandulescu, N.; Negrea, D.; Gambacurta, D. Proton-neutron pairing in N = Z nuclei: Quartetting versus pair condensation. Phys. Lett. B 2015, 751, 348–351. [Google Scholar] [CrossRef]
  831. Darai, J.; Cseh, J.; Jenkins, D.G. Shape isomers and clusterization in the 28Si nucleus. Phys. Rev. C 2012, 86, 064309. [Google Scholar] [CrossRef]
  832. Cseh, J.; Algora, A.; Darai, J.; Hess, P.O. Deformation dependence of nuclear clusterization. Phys. Rev. C 2004, 70, 034311. [Google Scholar] [CrossRef]
  833. Cseh, J.; Darai, J.; Sciani, W.; Otani, Y.; Lépine-Szily, A.; Benjamim, E.A.; Chamon, L.C.; Lichtenthäler Filho, R. Elongated shape isomers in the 36Ar nucleus. Phys. Rev. C 2009, 80, 034320. [Google Scholar] [CrossRef]
  834. Satuła, W.; Wyss, R. Competition between T = 0 and T = 1 pairing in proton-rich nuclei. Phys. Lett. B 1997, 393, 1. [Google Scholar] [CrossRef]
  835. Terasaki, J.; Wyss, R.; Heenen, P.-H. Onset of T = 0 pairing and deformations in high spin states of the N = Z nucleus 48Cr. Phys. Lett. B 1998, 437, 1. [Google Scholar] [CrossRef]
  836. Lei, Y.; Pittel, S.; Sandulescu, N.; Poves, A.; Thakur, B.; Zhao, Y.M. Systematic study of proton-neutron pairing correlations in the nuclear shell model. Phys. Rev. C 2011, 84, 044318. [Google Scholar] [CrossRef]
  837. Poves, A.; Martinez-Pinedo, G. Pairing and the structure of the pf-shell NZ nuclei. Phys. Lett. B 1998, 430, 203. [Google Scholar] [CrossRef]
  838. de Angelis, G.; Fahlander, C.; Gadea, A.; Farnea, E.; Gelletly, W.; Aprahamian, A.; Axelsson, A.; Bazzacco, D.; Becker, F.; Bizzeti, P.G.; et al. T = 0 pairing correlations and band crossing phenomena in N = Z nuclei. Nucl. Phys. A 1998, 630, 426c. [Google Scholar] [CrossRef]
  839. Rudolph, D.; Baktash, C.; Satuła, W.; Dobaczewski, J.; Nazarewicz, W.; Brinkman, M.J.; Devlin, M.; Jin, H.-Q.; LaFosse, D.R.; Riedinger, L.L.; et al. High-spin γ-ray spectroscopy in the vicinity of 56Ni. Nucl. Phys. A 1998, 630, 417c. [Google Scholar] [CrossRef]
  840. Rudolph, D.; Baktash, C.; Brinkman, M.J.; Caurier, E.; Dean, D.J.; Devlin, M.; Dobaczewski, J.; Heenen, P.-H.; Jin, H.-Q.; LaFosse, D.R.; et al. Rotational Bands in the Doubly Magic Nucleus 56Ni. Phys. Rev. Lett. 1999, 82, 3763. [Google Scholar] [CrossRef]
  841. Svensson, C.E.; Rudolph, D.; Baktash, C.; Bentley, M.A.; Cameron, J.A.; Carpenter, M.P.; Devlin, M.; Eberth, J.; Flibotte, S.; Galindo-Uribarri, A.; et al. Decay Out of the Doubly Magic Superdeformed Band in the N = Z Nucleus 60Zn. Phys. Rev. Lett. 1999, 82, 3400. [Google Scholar] [CrossRef]
  842. Dobaczewski, J.; Dudek, J.; Wyss, R. T = 0 neutron-proton pairing correlations in the superdeformed rotational bands around 60Zn. Phys. Rev. C 2003, 67, 034308. [Google Scholar] [CrossRef]
  843. Šimkovic, F.; Moustakidis, C.C.; Pacearescu, L.; Faessler, A. Proton-neutron pairing in the deformed BCS approach. Phys. Rev. C 2003, 68, 054319. [Google Scholar] [CrossRef]
  844. Yoshida, K. Proton-neutron pairing vibrations in N = Z nuclei: Precursory soft mode of isoscalar pairing condensation. Phys. Rev. C 2014, 90, 031303. [Google Scholar] [CrossRef]
  845. Sagawa, H.; Tanimura, Y.; Hagino, K. Competition between T = 1 and T = 0 pairing in pf-shell nuclei with N = Z. Phys. Rev. C 2013, 87, 034310. [Google Scholar] [CrossRef]
  846. Darai, J.; Cseh, J.; Antonenko, N.V.; Royer, G.; Algora, A.; Hess, P.O.; Jolos, R.V.; Scheid, W. Clusterization in the shape isomers of the 56Ni nucleus. Phys. Rev. C 2011, 84, 024302. [Google Scholar] [CrossRef]
  847. Macchiavelli, A.O.; Fallon, P.; Clark, R.M.; Cromaz, M.; Deleplanque, M.A.; Diamond, R.M.; Lane, G.J.; Lee, I.Y.; Stephens, F.S.; Svensson, C.E.; et al. Collective T = 0 pairing in N = Z nuclei? Pairing vibrations around 56Ni revisited. Phys. Lett. B 2000, 480, 1. [Google Scholar]
  848. Sagawa, H.; Bai, C.L.; Colò, G. Isovector spin-singlet (T = 1, S = 0) and isoscalar spin-triplet (T = 0, S = 1) pairing interactions and spin-isospin response. Phys. Scr. 2016, 91, 083011. [Google Scholar] [CrossRef]
  849. Skoda, S.; Fiedler, B.; Becker, F.; Eberth, J.; Freund, S.; Steinhardt, T.; Stuch, O.; Thelen, O.; Thomas, H.G.; Käubler, L.; et al. Identification of excited states in the N = Z nucleus 68Se with cluster detectors. Phys. Rev. C 1998, 58, R5. [Google Scholar] [CrossRef]
  850. Lieb, K.P.; Harder, A.; Rudolph, D.; Dönau, F.; Gelletly, W.; Gross, C.J.; Jungclaus, A.; Sheikh, J.A.; EUROGRAM 1 Collaboration; NORDBALL Collaboration. High-spin spectroscopy near the deformed N≈Z≈38 shell gap: The light Rb isotopes. Prog. Part. Nucl. Phys. 1997, 38, 101. [Google Scholar] [CrossRef]
  851. Tsunoda, N.; Otsuka, T.; Takayanagi, K.; Shimizu, N.; Suzuki, T.; Utsuno, Y.; Yoshida, S.; Ueno, H. The impact of nuclear shape on the emergence of the neutron dripline. Nature 2020, 587, 66. [Google Scholar] [CrossRef]
  852. Neufcourt, L.; Cao, Y.; Giuliani, S.A.; Nazarewicz, W.; Olsen, E.; Tarasov, O.B. Quantified limits of the nuclear landscape. Phys. Rev. C 2020, 101, 044307. [Google Scholar] [CrossRef]
  853. Stroberg, S.R.; Holt, J.D.; Schwenk, A.; Simonis, J. Ab Initio Limits of Atomic Nuclei. Phys. Rev. Lett. 2021, 126, 022501. [Google Scholar] [CrossRef] [PubMed]
  854. Fortune, H.T. Proposed new classification among coexistence nuclei. Nucl. Phys. A 2020, 1004, 122063. [Google Scholar] [CrossRef]
  855. Ahmad, I.; Butler, P.A. Octupole Shapes in Nuclei. Annu. Rev. Nucl. Part. Sci. 1993, 43, 71–116. [Google Scholar] [CrossRef]
  856. Butler, P.A.; Nazarewicz, W. Intrinsic reflection asymmetry in atomic nuclei. Rev. Mod. Phys. 1996, 68, 349. [Google Scholar] [CrossRef]
  857. Butler, P.A. Octupole collectivity in nuclei. J. Phys. G Nucl. Part. Phys. 2016, 43, 073002. [Google Scholar] [CrossRef]
  858. Phillips, W.R.; Ahmad, I.; Emling, H.; Holzmann, R.; Janssens, R.V.F.; Khoo, T.-L.; Drigert, M.W. Octupole deformation in neutron-rich barium isotopes. Phys. Rev. Lett. 1986, 57, 3257. [Google Scholar] [CrossRef]
  859. Schüler, P.; Lauterbach, C.; Agarwal, Y.K.; De Boer, J.; Blume, K.P.; Butler, P.A.; Euler, K.; Fleischmann, C.; Günther, C.; Hauber, E.; et al. High-spin states in 224,226,228Th and the systematics of octupole effects in even Th isotopes. Phys. Lett. B 1986, 174, 241. [Google Scholar] [CrossRef]
  860. Leander, G.A.; Sheline, R.K.; Möller, P.; Olanders, P.; Ragnarsson, I.; Sierk, A.J. The breaking of intrinsic reflection symmetry in nuclear ground states. Nucl. Phys. A 1982, 388, 452–476. [Google Scholar] [CrossRef]
  861. Inglis, D.R. Particle Derivation of Nuclear Rotation Properties Associated with a Surface Wave. Phys. Rev. 1954, 96, 1059–1065. [Google Scholar] [CrossRef]
  862. Inglis, D.R. Nuclear Moments of Inertia due to Nucleon Motion in a Rotating Well. Phys. Rev. 1956, 103, 1786–1795. [Google Scholar] [CrossRef]
  863. Kerman, A.K. Pairing forces and nuclear collective motion. Ann. Phys. 1961, 12, 300–329. [Google Scholar] [CrossRef]
  864. Nazarewicz, W.; Dudek, J.; Bengtsson, R.; Bengtsson, T.; Ragnarsson, I. Microscopic study of the high-spin behaviour in selected A≈80 nuclei. Nucl. Phys. A 1985, 435, 397–447. [Google Scholar] [CrossRef]
  865. Cwiok, S.; Dudek, J.; Nazarewicz, W.; Skalski, J.; Werner, T. Single-particle energies, wave functions, quadrupole moments and g-factors in an axially deformed woods-saxon potential with applications to the two-centre-type nuclear problems. Comp. Phys. Commun. 1987, 46, 379–399. [Google Scholar] [CrossRef]
  866. Strutinsky, V.M. Shell effects in nuclear masses and deformation energies. Nucl. Phys. A 1967, 95, 420. [Google Scholar] [CrossRef]
Figure 1. Nuclei exhibiting SC according to the evidence discussed in Section 13.1. The nuclei mentioned in the conclusion or summary of each subsection in Section 3, Section 4, Section 5, Section 6, Section 7, Section 8, Section 9 and Section 10 are included. Most nuclei fall within the stripes predicted by the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327], shown in azure, with notable exceptions corresponding to the islands of inversion (see Section 12) and the N = Z line (see Section 11 and Section 11.1). Contours corresponding to P 5 [266,347,348] are also shown in orange, with deformed nuclei lying within them. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. See Section 13.1 for further discussion.
Figure 1. Nuclei exhibiting SC according to the evidence discussed in Section 13.1. The nuclei mentioned in the conclusion or summary of each subsection in Section 3, Section 4, Section 5, Section 6, Section 7, Section 8, Section 9 and Section 10 are included. Most nuclei fall within the stripes predicted by the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327], shown in azure, with notable exceptions corresponding to the islands of inversion (see Section 12) and the N = Z line (see Section 11 and Section 11.1). Contours corresponding to P 5 [266,347,348] are also shown in orange, with deformed nuclei lying within them. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. See Section 13.1 for further discussion.
Atoms 11 00117 g001
Figure 2. Nuclei suggested in the literature as candidates for the X(5) CPS, related to the SPT from spherical to deformed nuclear shape are shown as blue circles. The nuclei shown are 38 76 Sr 38 [350], 38 78 Sr 40 [350], 40 80 Zr 40 [350], 40 100 Zr 60 [350], 42 98 Mo 56 [351,352], 42 104 Mo 62 [350,351,353], 42 106 Mo 64 [351], 44 102 Ru 58 [352], 56 122 Ba 66 [354,355], 56 126 Ba 70 [356], 58 128 Ce 70 [357], 58 130 Ce 72 [356,358,359], 60 150 Nd 90 [360,361,362], 62 150 Sm 88 [363], 62 152 Sm 90 [358,360,364,365,366,367,368,369,370,371,372,373], 64 152 Gd 88 [374], 64 154 Gd 90 [375,376], 66 156 Dy 90 [375,377,378], 70 162 Yb 92 [379], 72 166 Hf 94 [379,380], 72 168 Hf 96 [381], 72 170 Hf 98 [382], 76 176 Os 100 [383], 76 178 Os 102 [383], 76 180 Os 104 [383]. Contours corresponding to P 5 [266,347,348] are also shown in orange, with deformed nuclei lying within them. We remark that candidates for X(5) lie on or next to the P 5 contours. The stripes in which SC can be expected according to the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327] are shown in azure. In addition, nuclei suggested in the literature as candidates for the E(5) CPS, related to the SPT from spherical to γ -unstable shape are shown as red X’s, for comparison. The nuclei shown are 22 58 Ti 36 [384], 24 58 Cr 34 [384,385,386,387], 24 60 Cr 36 [384], 24 62 Cr 38 [384], 40 98 Zr 58 [388], 44 104 Ru 60 [389], 46 100 Pd 54 [390], 46 102 Pd 56 [372,391,392], 46 108 Pd 62 [393,394], 48 106 Cd 58 [391], 48 108 Cd 60 [391,395], 48 114 Cd 66 [396], 52 124 Te 72 [391,397,398], 54 128 Xe 74 [391,399], 54 130 Xe 76 [400,401], 54 132 Xe 78 [400], 56 132 Ba 76 [402], 56 134 Ba 78 [366,391,394,402,403,404,405]. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. See Section 13.1 for further discussion.
Figure 2. Nuclei suggested in the literature as candidates for the X(5) CPS, related to the SPT from spherical to deformed nuclear shape are shown as blue circles. The nuclei shown are 38 76 Sr 38 [350], 38 78 Sr 40 [350], 40 80 Zr 40 [350], 40 100 Zr 60 [350], 42 98 Mo 56 [351,352], 42 104 Mo 62 [350,351,353], 42 106 Mo 64 [351], 44 102 Ru 58 [352], 56 122 Ba 66 [354,355], 56 126 Ba 70 [356], 58 128 Ce 70 [357], 58 130 Ce 72 [356,358,359], 60 150 Nd 90 [360,361,362], 62 150 Sm 88 [363], 62 152 Sm 90 [358,360,364,365,366,367,368,369,370,371,372,373], 64 152 Gd 88 [374], 64 154 Gd 90 [375,376], 66 156 Dy 90 [375,377,378], 70 162 Yb 92 [379], 72 166 Hf 94 [379,380], 72 168 Hf 96 [381], 72 170 Hf 98 [382], 76 176 Os 100 [383], 76 178 Os 102 [383], 76 180 Os 104 [383]. Contours corresponding to P 5 [266,347,348] are also shown in orange, with deformed nuclei lying within them. We remark that candidates for X(5) lie on or next to the P 5 contours. The stripes in which SC can be expected according to the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327] are shown in azure. In addition, nuclei suggested in the literature as candidates for the E(5) CPS, related to the SPT from spherical to γ -unstable shape are shown as red X’s, for comparison. The nuclei shown are 22 58 Ti 36 [384], 24 58 Cr 34 [384,385,386,387], 24 60 Cr 36 [384], 24 62 Cr 38 [384], 40 98 Zr 58 [388], 44 104 Ru 60 [389], 46 100 Pd 54 [390], 46 102 Pd 56 [372,391,392], 46 108 Pd 62 [393,394], 48 106 Cd 58 [391], 48 108 Cd 60 [391,395], 48 114 Cd 66 [396], 52 124 Te 72 [391,397,398], 54 128 Xe 74 [391,399], 54 130 Xe 76 [400,401], 54 132 Xe 78 [400], 56 132 Ba 76 [402], 56 134 Ba 78 [366,391,394,402,403,404,405]. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. See Section 13.1 for further discussion.
Atoms 11 00117 g002
Figure 3. Nuclei with well-developed K = 0 bands (blue boxes) (listed in Table I of [335], based on data taken from Ref. [406]). The stripes within which SC can be expected according to the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327] are also shown in light green. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. Adapted from [335]. See Section 13.2 for further discussion.
Figure 3. Nuclei with well-developed K = 0 bands (blue boxes) (listed in Table I of [335], based on data taken from Ref. [406]). The stripes within which SC can be expected according to the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327] are also shown in light green. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. Adapted from [335]. See Section 13.2 for further discussion.
Atoms 11 00117 g003
Figure 4. Nuclei with energy differences E ( 0 2 + ) E ( 2 1 + ) less that 800 keV (blue boxes), based on data taken from Ref. [406]. The stripes within which SC can be expected according to the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327] are also shown in light green. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. Adapted from [335]. See Section 13.2 for further discussion.
Figure 4. Nuclei with energy differences E ( 0 2 + ) E ( 2 1 + ) less that 800 keV (blue boxes), based on data taken from Ref. [406]. The stripes within which SC can be expected according to the dual-shell mechanism [334] developed in the framework of the proxy-SU(3) symmetry [323,327] are also shown in light green. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. Adapted from [335]. See Section 13.2 for further discussion.
Atoms 11 00117 g004
Figure 5. Nuclei with experimentally known [406] B ( E 2 ; 0 2 + 2 1 + ) transition rates and K π = 0 + bands, satisfying the conditions R 4 / 2 < 3.05 (Equation (2)), R 0 / 2 < 5.7 (Equation (4)), and B 02 > 0.1 (Equation (5)), and known to exhibit SC according to the references listed in the relevant to each series of isotopes subsection of Section 3, Section 4, Section 5, Section 6, Section 7, Section 8, Section 9 and Section 10, are shown as blue circles. The azure stripes within which SC can be expected to occur according to the dual-shell mechanism [335] developed within the framework of the proxy-SU(3) symmetry [323,327] are also shown, along with the contours (indicated by orange boxes) corresponding to P 5 [266,347,348]. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. Adapted from [407]. See Section 13.2 for further discussion.
Figure 5. Nuclei with experimentally known [406] B ( E 2 ; 0 2 + 2 1 + ) transition rates and K π = 0 + bands, satisfying the conditions R 4 / 2 < 3.05 (Equation (2)), R 0 / 2 < 5.7 (Equation (4)), and B 02 > 0.1 (Equation (5)), and known to exhibit SC according to the references listed in the relevant to each series of isotopes subsection of Section 3, Section 4, Section 5, Section 6, Section 7, Section 8, Section 9 and Section 10, are shown as blue circles. The azure stripes within which SC can be expected to occur according to the dual-shell mechanism [335] developed within the framework of the proxy-SU(3) symmetry [323,327] are also shown, along with the contours (indicated by orange boxes) corresponding to P 5 [266,347,348]. The proton (neutron) driplines, shown as thick black lines, have been extracted from the two-proton (two-neutron) separation energies appearing in the database NuDat3.0 [349], involving standard extrapolations. Adapted from [407]. See Section 13.2 for further discussion.
Atoms 11 00117 g005
Figure 6. The regions of SC reported in Figure 8 of the review article [4], shown as blue ovals, are compared to light green stripes predicted by the dual-shell mechanism [335] developed within the framework of the proxy-SU(3) symmetry [323,327]. See Section 13.3 for further discussion.
Figure 6. The regions of SC reported in Figure 8 of the review article [4], shown as blue ovals, are compared to light green stripes predicted by the dual-shell mechanism [335] developed within the framework of the proxy-SU(3) symmetry [323,327]. See Section 13.3 for further discussion.
Atoms 11 00117 g006
Figure 7. Experimental energy levels in keV [406] for the Hg ( Z = 80 ) (a), Pb ( Z = 82 ) (b), and Po ( Z = 84 ) (c) series of isotopes. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 3.1, Section 3.2 and Section 3.3 for further discussion.
Figure 7. Experimental energy levels in keV [406] for the Hg ( Z = 80 ) (a), Pb ( Z = 82 ) (b), and Po ( Z = 84 ) (c) series of isotopes. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 3.1, Section 3.2 and Section 3.3 for further discussion.
Atoms 11 00117 g007
Figure 8. Experimental energy levels in keV [406] for the N = 88 (a), N = 90 (b), and N = 92 (c) series of isotones. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 5.5 for further discussion.
Figure 8. Experimental energy levels in keV [406] for the N = 88 (a), N = 90 (b), and N = 92 (c) series of isotones. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 5.5 for further discussion.
Atoms 11 00117 g008
Figure 9. Experimental energy levels in keV [406] for the Cd ( Z = 48 ) (a), Sn ( Z = 50 ) (b), and Te ( Z = 52 ) (c) series of isotopes. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 7.1, Section 7.2 and Section 7.3 for further discussion.
Figure 9. Experimental energy levels in keV [406] for the Cd ( Z = 48 ) (a), Sn ( Z = 50 ) (b), and Te ( Z = 52 ) (c) series of isotopes. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 7.1, Section 7.2 and Section 7.3 for further discussion.
Atoms 11 00117 g009
Figure 10. Experimental energy levels in keV [406] for the N = 58 (a), N = 60 (b), and N = 62 (c) series of isotones. The levels of the ground-state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 8.5 for further discussion.
Figure 10. Experimental energy levels in keV [406] for the N = 58 (a), N = 60 (b), and N = 62 (c) series of isotones. The levels of the ground-state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 8.5 for further discussion.
Atoms 11 00117 g010
Figure 11. Experimental energy levels in keV [406] for the Zn ( Z = 30 ) (a), Ge ( Z = 32 ) (b), Se ( Z = 34 ) (c), and Kr ( Z = 36 ) (d) series of isotopes. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 9.1, Section 9.2, Section 9.3 and Section 9.4 for further discussion.
Figure 11. Experimental energy levels in keV [406] for the Zn ( Z = 30 ) (a), Ge ( Z = 32 ) (b), Se ( Z = 34 ) (c), and Kr ( Z = 36 ) (d) series of isotopes. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 9.1, Section 9.2, Section 9.3 and Section 9.4 for further discussion.
Atoms 11 00117 g011
Figure 12. Experimental energy levels in keV [406] for the N = 38 (a), N = 40 (b), and N = 42 (c) series of isotones. The levels of the ground-state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 9.5 for further discussion.
Figure 12. Experimental energy levels in keV [406] for the N = 38 (a), N = 40 (b), and N = 42 (c) series of isotones. The levels of the ground-state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 9.5 for further discussion.
Atoms 11 00117 g012
Figure 13. Experimental energy levels in keV [406] for the Mg ( Z = 12 ) (a), Ca ( Z = 20 ) (b), and Ni ( Z = 28 ) (c) series of isotopes. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 10.1, Section 10.5 and Section 10.9 and for further discussion.
Figure 13. Experimental energy levels in keV [406] for the Mg ( Z = 12 ) (a), Ca ( Z = 20 ) (b), and Ni ( Z = 28 ) (c) series of isotopes. The levels of the ground state band bear the subscript g, while the levels of the excited K π = 0 + band bear no subscript. See Section 10.1, Section 10.5 and Section 10.9 and for further discussion.
Atoms 11 00117 g013
Figure 14. Nuclei found to exhibit particle–hole excitations, and therefore showing SC, in CDFT calculations with standard pairing included [579,580]. Islands of SC corresponding to neutron-induced SC are shown in yellow, islands due to proton-induced SC are exhibited in cyan, while islands due to both mechanisms are shown in green. Adapted from [580]. See Section 13.1 for further discussion.
Figure 14. Nuclei found to exhibit particle–hole excitations, and therefore showing SC, in CDFT calculations with standard pairing included [579,580]. Islands of SC corresponding to neutron-induced SC are shown in yellow, islands due to proton-induced SC are exhibited in cyan, while islands due to both mechanisms are shown in green. Adapted from [580]. See Section 13.1 for further discussion.
Atoms 11 00117 g014
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bonatsos, D.; Martinou, A.; Peroulis, S.K.; Mertzimekis, T.J.; Minkov, N. Shape Coexistence in Even–Even Nuclei: A Theoretical Overview. Atoms 2023, 11, 117. https://doi.org/10.3390/atoms11090117

AMA Style

Bonatsos D, Martinou A, Peroulis SK, Mertzimekis TJ, Minkov N. Shape Coexistence in Even–Even Nuclei: A Theoretical Overview. Atoms. 2023; 11(9):117. https://doi.org/10.3390/atoms11090117

Chicago/Turabian Style

Bonatsos, Dennis, Andriana Martinou, Spyridon K. Peroulis, Theodoros J. Mertzimekis, and Nikolay Minkov. 2023. "Shape Coexistence in Even–Even Nuclei: A Theoretical Overview" Atoms 11, no. 9: 117. https://doi.org/10.3390/atoms11090117

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