Next Article in Journal
Symmetrical Hybrid Coupled Fuzzy Fixed-Point Results on Closed Ball in Fuzzy Metric Space with Applications
Next Article in Special Issue
A Symmetry In-between the Shapes, Shells, and Clusters of Nuclei
Previous Article in Journal
Breaking Material Symmetry to Control Mechanical Performance in 3D Printed Objects
Previous Article in Special Issue
Overview of Seniority Isomers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Islands of Shape Coexistence: Theoretical Predictions and Experimental Evidence

by
Andriana Martinou
1,
Dennis Bonatsos
1,*,
Spyridon Kosmas Peroulis
1,
Konstantinos Eleftherios Karakatsanis
2,3,
Theodoros John Mertzimekis
4 and
Nikolay Minkov
5
1
Institute of Nuclear and Particle Physics, National Centre for Scientific Research “Demokritos”, GR-15310 Aghia Paraskevi, Attiki, Greece
2
Department of Physics, Faculty of Science, University of Zagreb, HR-10000 Zagreb, Croatia
3
Physics Department, Aristotle University of Thessaloniki, GR-54124 Thessaloniki, Greece
4
Department of Physics, National and Kapodistrian University of Athens, Zografou Campus, GR-15784 Athens, Greece
5
Institute of Nuclear Research and Nuclear Energy, Bulgarian Academy of Sciences, 72 Tzarigrad Road, 1784 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Symmetry 2023, 15(1), 29; https://doi.org/10.3390/sym15010029
Submission received: 13 November 2022 / Revised: 4 December 2022 / Accepted: 6 December 2022 / Published: 22 December 2022

Abstract

:
Parameter-free theoretical predictions based on a dual shell mechanism within the proxy-SU(3) symmetry of atomic nuclei, as well as covariant density functional theory calculations using the DDME2 functional indicate that shape coexistence (SC) based on the particle-hole excitation mechanism cannot occur everywhere on the nuclear chart but is restricted on islands lying within regions of 7–8, 17–20, 34–40, 59–70, 96–112, 146–168 protons or neutrons. Systematics of data for even-even nuclei possessing K = 0 (beta) and K = 2 (gamma) bands support the existence of these islands, on which shape coexistence appears whenever the K = 0 bandhead 0 2 + and the first excited state of the ground state band 2 1 + lie close in energy, with nuclei characterized by 0 2 + lying below the 2 1 + found in the center of these islands. In addition, a simple theoretical mechanism leading to multiple-shape coexistence is briefly discussed.

1. Introduction

Shape coexistence (SC) is the term used to describe the situation in which in an atomic nucleus the ground state band and another band lying nearby in energy possess similar energy levels, but radically different structures, for example, one of them being spherical and the other one deformed, or both of them being deformed, but one of them exhibiting prolate (rugby ball-like) deformation and the other one showing oblate (pancake-like) deformation. First proposed in 1956 by Morinaga [1], in relation to the spectrum of 16 O, SC has been observed in many odd [2] and even-even [3,4] nuclei, recently receiving intense experimental attention [5] since more cases become accessible through novel radioactive ion beam facilities.
It has been believed over the years that shape coexistence can occur practically everywhere across the nuclear chart, despite the fact that experimental examples have been appearing to cluster into islands on the nuclear chart, as seen, for example, in Figure 8 of the review article [4]. However, it has been recently proposed, within the framework of the proxy-SU(3) symmetry [6,7], that SC can occur only within the proton or neutron number intervals 7–8, 17–20, 34–40, 59–70, 96–112, 146–168 [8,9], which form certain horizontal and vertical stripes on the nuclear chart. Furthermore, recent covariant density functional theory (CDFT) calculations using the DDME2 functional [10,11] suggest that SC arising from nucleon particle-hole excitations occurs only within certain islands on the nuclear chart, the islands themselves lying within the stripes predicted by the proxy-SU(3) symmetry. In particular, islands of neutron-induced SC, in which SC is due to particle-hole excitations of protons, caused by the neutrons, as well as islands of proton-induced SC, in which SC is due to particle-hole excitations of neutrons, caused by the protons, have been found, in addition to islands in which both particle-hole mechanisms are present simultaneously [10,11].
In the present work, after a brief review of the proxy-SU(3) and CDFT predictions, we consider new systematics of data that support these findings. In particular, it is shown that even-even nuclei with experimentally known K = 0 bands [12], as well as nuclei with experimentally known K = 2 bands [12], form islands on the nuclear chart, within the stripes [8] predicted by the proxy-SU(3) symmetry. Furthermore, it is shown that even-even nuclei in which the first excited state of the ground state band, 2 1 + , and the first K = 0 bandhead, 0 2 + , lie close in energy, making these nuclei good SC candidates, also form islands on the nuclear chart, in agreement with the stripes predicted by the proxy-SU(3) symmetry. In addition, a qualitative mechanism giving rise to multiple SC [13,14] is briefly described.

2. Harmonic Oscillator (HO) and Spin-Orbit (SO) Magic Numbers

Magic numbers have played a major role since the infancy of nuclear structure. The first set of magic numbers introduced in nuclear structure was the set of the magic numbers 2, 8, 20, 28, 50, 82, …, which gave rise to the shell model [15,16]. It has been understood that these magic numbers rise from the three-dimensional isotropic harmonic oscillator (3D-HO) magic numbers 2, 8, 20, 40, 70, 112, 168, … [17,18,19] with the addition of the spin–orbit interaction [15,16], which is small up to the s d shell, but modifies the composition of the HO shells beyond this point. More recent experimental evidence [20] led to the distinction between the HO magic numbers 2, 8, 20, 40, 70, 112, 168, …, which are valid everywhere in the absence of any spin–orbit interaction, and spin-orbit (SO) magic numbers 6, 14, 28, 50, 82, 126, 184, …, [20,21], which are valid in the case of strong spin–orbit interaction everywhere. It is clear that the shell model magic numbers are a mixture of these two sets, starting with HO magic numbers in light nuclei and ending up in medium-mass and heavy nuclei with SO magic numbers.

3. The Dual-Shell Mechanism

A dual-shell mechanism [8], based on the interplay of the HO and SO shells have been recently introduced, predicting the existence of specific regions in the nuclear chart, in which shape coexistence (SC) can occur.
From the conceptual point of view, the dual-shell mechanism is based on the fact that the shell model magic numbers are known to be valid only at zero nuclear deformation, or close to it. When nuclear deformation comes in, the shell model magic numbers are not valid anymore, as one can see in the standard Nilsson diagrams [22,23,24], in which the change of the nucleon (proton or neutron) single-particle energies as a function of the nuclear deformation is depicted. The Nilsson model [22,23] consists of a three-dimensional HO with cylindrical symmetry, to which a spin–orbit interaction is added. Despite its simplicity, it has been extremely helpful for decades in analyzing experimental data. Away from zero deformation, one can see in the Nilsson diagrams that the large gaps appearing above shell model magic numbers at zero deformation deteriorate quickly, while new magic numbers appear and disappear with increasing deformation. The appearance of new magic numbers has recently attracted considerable attention, both from the experimental [20] and theoretical [25] viewpoints.
From the algebraic point of view, the dual-shell mechanism is based on the SU(3) symmetry of the Elliott model [26,27,28,29], prevailing in light nuclei up to the s d shell, and the proxy-SU(3) symmetry [6,7], proved to be a very good approximation valid in medium mass and heavy nuclei [21,30]. The dual-shell mechanism predicts that SC can be obtained only when the proton and/or the neutron numbers, Z and N, respectively, of a given nucleus, are within the intervals 7–8, 17–20, 34–40, 59–70, 96–112, 146–168. While the right borders of these regions are the well-known HO magic numbers [17,18], the left borders, 7, 17, 34, 59, 96, 146 form a new set of magic numbers (called the SC magic numbers in this work), determined by the condition that the quadrupole-quadrupole interaction within the HO and SO schemes become equal [8,9]. Actually, the intervals just mentioned represent regions in which the magnitude of the quadrupole–quadrupole interaction (expressed through the second-order Casimir operator of the relevant SU(3) irreducible representation (irrep) [7,31,32]) coming from the SO shell exceeds the one coming from the HO shell.
In more detail, in Ref. [8] the third set of magic numbers (the SC magic numbers) has been determined using the eigenvalue of the second-order Casimir operator of SU(3) [19,31]
C 2 ( λ , μ ) = λ 2 + μ 2 + λ μ + 3 ( λ + μ ) ,
where λ and μ are the Elliott quantum numbers [26,27,28,29] characterizing the ( λ , μ ) irreducible representation of SU(3). The Casimir C 2 is known to be proportional to the quadrupole–quadrupole operator Q Q [26,27,28,29], as well as to the square of the collective deformation variable β [7,33]. For the protons (or the neutrons) of each nucleus, we have two possibilities: The ( λ , μ ) irrep coming from the HO picture, or the irrep coming from the SO picture (see Table 1 of Ref. [8] for ready-to-use irreps). The need to use the highest weight irreps has been emphasized in Ref. [7] and fully justified mathematically in Ref. [32]. The SC magic number occurs at the point at which the SO Casimir, or equivalently the QQ interaction or the β deformation variable, exceeds the corresponding HO one. Apparently, this point is related to the onset of deformation [34,35,36].
If the neutron number N of a given nucleus belongs to one of the above-mentioned intervals, we say that this nucleus exhibits neutron-induced SC. As we shall see below, this is the kind of SC occurring in the Pb and Hg isotopes around the Z = 82 shell closure [4], as well as in the Sn and Te isotopes around the Z = 50 shell closure [3]. In the first case, SC occurs in the interval N = 96 –112, centered around N = 104 , while in the second case, SC occurs in the interval N = 59 –70, centered around N = 64 .
If the proton number Z of a given nucleus belongs to one of the above-mentioned intervals, we say that this nucleus exhibits proton-induced SC. As we shall see below, such a case occurs, for example, in the light (with N < 92 ) Sm, Gd, and Dy isotopes. SC is expected in this case just below N = 92 and within the interval Z = 59 –72, centered around Z = 64 .
If both the proton and the neutron numbers of a given nucleus lie within the above-mentioned intervals, we say that in this nucleus SC is both proton-induced and neutron-induced. As we shall see below, this happens, for example, in the Sr and Zr isotopes in the region Z N 40 .
The regions of the nuclear chart in which SC is expected according to the dual shell mechanism are shown in Figure 1. The colored regions possess proton or neutron numbers in one of the intervals 7–8, 17–20, 34–40, 59–70, 96–112, 145–168. The horizontal stripes correspond to the proton-induced SC, while the vertical stripes correspond to the neutron-induced SC.

4. Experimental Evidence for the Shape Coexistence (SC) Magic Numbers

This third set of magic numbers is not entirely unheard of. In the authoritative review on nuclear magic numbers of Ref. [20] based on experimental evidence, new gaps corresponding to magic numbers have been identified in N = 6 , 16, 32, 34 (see [20], p. 667), which are very close to 7, 17, 34 of the third set of magic numbers mentioned above. In the very recent review of the evolution of shell structure in exotic nuclei [25], the new magic numbers N = 16 and N = 34 emerged when using central and tensor forces in actual nuclei. We see, therefore, that the first three members of this new set of magic numbers have already been well established, from both the experimental [20] and the theoretical [25] viewpoint. For easy reference we shall call them the SC magic numbers, taking advantage of the fact that they correspond to the left borders of the regions in which SC can appear [8], while the right borders are the HO magic numbers.
Empirical evidence for the next two members (59, 96) of the SC magic numbers comes from the parabolic behavior observed when plotting certain energy levels vs. N. Parabolas in plots of specific nuclear states in even nuclei, such as the 2 1 + and 4 1 + states of the ground state band (gsb) [25,37], or the 2 γ + band-head of the γ 1 collective band [38], vs. the neutron number N have been known for a long time. The parabolas are bordered by the usual shell model magic numbers.
Parabolas also appear when plotting the 0 2 + states of even nuclei, usually considered as the band-heads of the relevant β 1 collective bands [3,4], vs. N. However, in these cases, the parabolas are bordered by new sets of magic numbers. 0 2 + states in the Sn isotopes are bordered by N = 60 and 70 [3], while 0 2 + states in the Hg and Pb isotopes are bordered by N = 96 and 112 [4]. In Ref. [8], it has been shown that these are regions in which SC is expected to appear, bordered on the left side by SC magic numbers and on the right side by HO magic numbers, according to the terminology introduced above.
Over the years some subshell closures, such as Z = 34 , 40, 64 have received considerable attention [3,4]. Within the framework described above, 34 is a member of the third (SC) set of magic numbers and 40 is a member of the set of HO magic numbers, while 64 is the midshell of the SC candidate region 59–70.
The empirical support for the members of the SC magic numbers discussed above, invites further consideration of the physics underlying them. This task will be undertaken in Section 6.

5. Islands of Shape Coexistence in Covariant Density Functional Theory

Before considering further experimental evidence in favor of the formation of islands of SC, it is worth examining what microscopic methods could say in comparison to the parameter-free symmetry-based predictions of the dual shell mechanism. In this direction, covariant density functional theory [39,40,41,42,43,44,45] calculations using the DDME2 functional [46] have been performed [10,11] using the code of Ref. [47], in which a finite range paring force [48,49] has been taken into account in its default value. The main results are shown in Figure 2. Two islands of neutron-induced SC are seen close to the magic numbers Z = 82 and Z = 50 , within the neutron regions in which the parabolas mentioned in Section 4 are observed. In addition, two islands of proton-induced SC are seen close to Z = 64 and Z = 40 discussed in Section 4 while an island located near N = Z = 40 corresponds to nuclei in which both neutron-induced and proton-induced SC are active. All islands lie entirely within the horizontal and vertical stripes predicted by the dual shell mechanism. Further discussion is deferred to Section 9.

6. Islands of K = 0 and K = 2 Bands

A systematic collection of experimental data from the ENSDF database [12] for the collective K = 0 and K = 2 bands in even-even nuclei throughout the nuclear chart has been performed.
In Figure 3 are shown the nuclei in which well-developed excited K = 0 bands appear, also listed in Table 1. All even-even nuclei up to Z = 84 and N = 126 have been considered. 104 nuclei are included, fulfilling the following requirements: (a) In Ref. [12] these bands are denoted as β bands, or quasi- β bands, or K = 0 bands. (b) At least two levels of these bands are known experimentally.
In Figure 3, the regions in which shape coexistence is expected according to the dual shell mechanism, i.e., 7–8, 17–20, 34–40, 59–70, 96–112, are also shown for both protons and neutrons.
We see that almost all nuclei with known K = 0 bands fall within a horizontal (proton) or vertical (neutron) zone, while the few exceptions touch the borders of these zones.
This fact raises several comments and questions:
(1) The appearance of experimental K = 0 bands within certain zones cannot be accidental. It is understood that searches for K = 0 bands [50] have been performed over the years in other regions as well, with no K = 0 bands been identified.
(2) K = 0 bands are seen within proton-induced coexistence zones (horizontal zones), as well as within neutron-induced coexistence zones (vertical zones).
(3) The nature of K = 0 bands has been disputed in recent years by Sharpey-Schafer [51,52,53] and others [50,54]. The present study might lead to additional insights in relation to their nature.
In Figure 4 are shown the nuclei in which well-developed excited K = 2 bands appear, also listed in Table 2. All even-even nuclei up to Z = 84 and N = 126 have been considered. 164 nuclei are included, fulfilling the following requirements: (a) In Ref. [12] these bands are denoted as γ bands, or quasi- γ bands, or K = 2 bands. (b) At least two levels of these bands are known experimentally.
The regions in which shape coexistence is expected are shown in the same way as in Figure 3.
We see that most of the nuclei with known K = 2 bands fall within a horizontal zone and/or a vertical zone, as in Figure 3. However, in the present case, there are a few exceptions not touching the borders of these zones.
Comments (1) and (2) made on K = 0 bands apply also to the K = 2 bands. However, in contrast to the K = 0 bands, the nature of the K = 2 bands as being collective gamma vibrations has not been disputed [55].
It seems, therefore, that almost all collective K = 0 bands, as well as most of the K = 2 bands, fall within the borders of the regions proposed through the dual shell mechanism of Ref. [8]. The remaining task is then to distinguish among these nuclei the ones in which SC is observed from those in which SC is not seen. In this direction, in the next section, we apply the criterion that the gsb and the excited K = 0 band should be close in energy, in order to narrow down the islands seen in Figure 3.

7. Energy Differences of 0 2 + and 2 1 +

For all nuclei exhibited in Figure 3, in which excited K = 0 bands known experimentally are shown, the energy difference of the 0 2 + and 2 1 + states has been considered. The energies of the 2 + state of the ground state band, the 0 + bandhead of the K = 0 ( β ) band, and the 2 + bandhead of the K = 2 ( γ ) band, wherever known [12], are shown in Table 3. Nuclei for which the energy difference E ( 0 2 + ) E ( 2 1 + ) remains lower than 800 keV are shown in boldface in Table 1 and depicted in Figure 5 (see the last paragraph of this session for further discussion of this arbitrary choice). These are nuclei for which SC is expected to appear since the ground state band and the first excited K = 0 band remain close in energy. All nuclei in which the 0 2 + and 2 1 + states come close in energy lie within the stripes predicted by the dual shell mechanism.
The emerging picture is simple and clear. The nuclei in which the 0 2 + and 2 1 + states are close in energy can be divided into three classes:
(a) Nuclei in which only the neutrons belong to the intervals 7–8, 17–20, 34–40, 59–70, 96–112. These are nuclei in which neutron-induced SC is expected.
(b) Nuclei in which only the protons belong to the intervals 7–8, 17–20, 34–40, and 59–70. These are nuclei in which proton-induced SC is expected.
(c) Nuclei in which both protons and neutrons belong to the intervals 7–8, 17–20, 34–40, and 59–70. These are nuclei in which SC is expected to be both proton-induced and neutron-induced.
As an example, we assign to these classes all nuclei for which the condition E ( 0 2 + ) < E ( 2 1 + ) is fulfilled, which appear to be the “core nuclei” around which islands are formed.
(a) Nuclei with neutron-induced SC: Pb 102 , 104 , 106 , 112 82 184 , 186 , 188 , 194 , Ge 40 32 72 .
(b) Nuclei with proton-induced SC: Zr 56 , 58 40 96 , 98 .
(c) Nuclei with both neutron-induced and proton-induced SC: Ca 20 20 40 .
As a result, some nuclei which appear as border cases of SC from one viewpoint can be core cases from the other viewpoint. For example, Zr 56 , 58 40 96 , 98 might appear as border cases from the viewpoint of neutron-induced SC, since their neutron numbers lie just outside the N = 59 –70 region, but they are core cases from the viewpoint of proton-induced SC since their proton number does lie within the interval Z = 34 –40. On the contrary, Ge 40 32 72 might appear as a border case from the viewpoint of proton-induced SC, since its proton number lies just outside the Z = 34 –40 region, but it is a core case from the viewpoint of neutron-induced SC since its neutron number does lie within the interval N = 34 –40.
It should be noticed that the arbitrary choice of the 800 keV cut-off for the energy difference E ( 0 2 + ) E ( 2 1 + ) , although arbitrary, does not influence the conclusions drawn in this section. Lowering the cut-off will make the islands formed around the “core” nuclei having E ( 0 2 + ) < E ( 2 1 + ) narrower while raising the cut-off by 200 keV, for example, will make the islands formed around the “core” nuclei wider, but still lying within the stripes predicted by the dual shell mechanism.

8. A Mechanism for Multiple Shape Coexistence

While in most cases reported in the relevant review articles [3,4,5], SC of two bands is seen, multiple SC has been recently observed experimentally in 110 , 112 Cd, in which four coexisting bands have been located [13,14]. We shall see that multiple-shape coexistence can occur within the dual shell mechanism mentioned earlier in a simple way.
From the considerations made so far, it becomes plausible that in regions of moderate or large deformation, both sets of magic numbers, the SO magic numbers (6, 14, 20, 28, 50, 82, 126, … [20]) and the 3D-HO magic numbers (2, 8, 20, 40, 70, 112, 168, … [23]) should play roles of comparable importance. This implies that in a given nucleus, the valence protons or neutrons could be described in SU(3) either by the relevant SO irrep or by the HO irrep. This leads to four possible combinations for protons-neutrons: SO-SO, SO-HO, HO-SO, and HO-HO. In all cases, the stretched irreps ( λ , μ ) = ( λ p + λ n , μ p + μ n ) [56] can be used as a lowest order approximation. Given the success of the proxy-SU(3) symmetry in predicting the β and γ deformation variables in much medium mass and heavy nuclei [7], it is expected that the SO-SO combination would correspond to the ground state band, while other irreps would correspond to other low-lying configurations. Since in most nuclei four very different irreps result from these combinations (or three different irreps, if equal numbers of valence protons and neutrons occupy the same shell), the structures of the corresponding bands are expected to be quite different. Therefore, if some of them lie close in energy, multiple SC would arise. It is clear that this simple mechanism can produce up to four coexisting bands. Further work is called for in this direction.

9. Discussion

The present work accumulates evidence that shape coexistence based on particle-hole excitations can occur only on certain islands of the nuclear chart, the main relevant points listed here.
(a) Parameter-free predictions based on a dual shell mechanism within the proxy-SU(3) symmetry indicate that SC can only occur within the nucleon regions 7–8, 17–20, 34–40, 59–70, 96–112, and 146–168, which form horizontal and vertical stripes on the nuclear chart.
(b) Covariant density functional theory calculations using the DDME2 functional suggest that SC based on particle-hole excitations can only occur on islands of the nuclear chart lying entirely within the proxy-SU(3) stripes mentioned above.
(c) In parallel, nuclei with experimentally known collective K = 0 and/or K = 2 bands also lie on similar islands of the nuclear chart located mostly within the proxy-SU(3) stripes, indicating that a certain amount of collectivity is a prerequisite for the occurrence of SC.
(d) Choosing from the islands of c) only the nuclei in which the 0 2 + and 2 1 + states are energetically close to each other, islands lying entirely within the proxy-SU(3) stripes are obtained.
In other words, the above findings indicate that SC based on particle-hole excitations can occur only on certain islands of the nuclear chart. These islands are characterized by the presence of collective K = 0 and/or K = 2 bands, a fact indicating that a certain minimal amount of collectivity is needed for the appearance of SC. The islands get narrower by insisting on the proximity of the energies of the 0 2 + and 2 1 + states. The islands determined in this way are in agreement with the islands on which particle-hole excitations can occur according to covariant density functional theory calculations. The islands also lie entirely within the SC stripes determined by parameter-free arguments within the proxy-SU(3) symmetry, using a dual shell mechanism.
Looking the other way around, parameter-free arguments based on the proxy-SU(3) symmetry suggest that SC can occur only within horizontal and vertical stripes of the nuclear chart, bordered by the nucleon numbers 7–8, 17–20, 34–40, 59–70, 96–112, and 146–168. Systematics of the appearance of K = 0 and K = 2 bands, as well as of the proximity of the 0 2 + and 2 1 + states, suggest that SC can be expected on islands lying entirely within the proxy-SU(3) stripes. This conclusion is corroborated by covariant density functional theory calculations using the DDME2 functional, indicating that particle-hole excitations can occur only within the same islands, lying within the proxy-SU(3) stripes.
The possible role of the pairing force should be mentioned at this point. Within the covariant density functional theory calculations, a finite range pairing force [48,49] has been taken into account in its default value, while no pairing force has been included so far in the proxy-SU(3) scheme. The pairing force is known [48,49] to improve the agreement between relativistic mean field predictions and the experimental values for the collective deformation variable β (see Figure 4 of Ref. [49]). Although the parameter-free predictions of proxy-SU(3) for β lie in general closer to the experimental values than RMF results (see Figure 3 of Ref. [7], for example), it is clear that room for improvement through inclusion of the pairing force in the proxy-SU(3) scheme does exist.
In view of the above, the proxy-SU(3) stripes represent a necessary but not sufficient condition for the existence of SC. The stripes are narrowed down into islands by the empirically seen necessary condition of having adequate collectivity and keeping the 0 2 + and 2 1 + states close to each other. These narrower islands are corroborated by covariant density functional theory calculations using the DDME2 functional, which indicate that particle-hole excitations are found only within these islands.
The above findings do not exclude the existence of SC in other regions, based on a different mechanism. The present results regard SC based on particle-hole excitations, for which a minimal degree of collectivity seems to be required.
The parameter-free proxy-SU(3) predictions mentioned above are obtained through a dual shell mechanism, based on the collapse of the usual shell model magic numbers away from zero deformation, where two sets of magic numbers can be considered, the HO magic numbers occurring in the absence of any spin–orbit interaction, and the SO magic numbers, prevailing when the spin–orbit interaction is strong. For a given number of protons, i.e., within a series of isotopes, the competition between the SO and HO neutron magic numbers leads to neutron-induced SC. Similarly, for a given number of neutrons, i.e., within a series of isotones, the competition between the SO and HO proton magic numbers leads to proton-induced SC. Within a more general scheme, competition between the SO and HO magic numbers can occur for both protons and neutrons, leading to multiple shape coexistence of up to four different shapes.

10. Conclusions

In conclusion, both the dual shell mechanism based on the proxy-SU(3) symmetry and covariant density functional theory calculations using the DDME2 functional suggest that SC cannot occur everywhere on the nuclear chart, but is restricted within certain islands. The existence of these islands is corroborated by empirical evidence for nuclei exhibiting K = 0 and K = 2 bands and having closely lying 0 2 + and 0 1 + states.
The predictions of a dual shell mechanism extended to both protons and neutrons for the regions in which multiple shape coexistence can occur is a project worth pursuing, in view of the recent experimental findings [13,14].

Author Contributions

Conceptualization, A.M., D.B., S.K.P., K.E.K., T.J.M., N.M.; Methodology, A.M., D.B., S.K.P., K.E.K., T.J.M., N.M.; Investigation, A.M., D.B., S.K.P., K.E.K., T.J.M., N.M.; Writing—original draft, D.B.; Writing—review & editing, A.M., S.K.P., K.E.K., T.J.M., 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 Tenure Track Pilot Programme of the Croatian Science Foundation and the Ecole Polytechnique Fédérale de Lausanne, the Project TTP-2018-07-3554 Exotic Nuclear Structure and Dynamics with funds of the Croatian-Swiss Research Programme, as well as by the Bulgarian National Science Fund (BNSF) under Contract No. KP-06-N48/1.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

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. 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]
  6. 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] [Green Version]
  7. 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] [Green Version]
  8. 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]
  9. Martinou, A. A mechanism for shape coexistence. EPJ Web Conf. 2021, 252, 02005. [Google Scholar] [CrossRef]
  10. 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]
  11. 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]
  12. ENSDF Database. Available online: https://www.nndc.bnl.gov/ensdf (accessed on 30 December 2020).
  13. 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. Phys. Rev. Lett. 2019, 123, 142502. [Google Scholar] [CrossRef] [PubMed]
  14. 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]
  15. Mayer, M.G. On closed shells in nuclei. II. Phys. Rev. 1949, 75, 1969. [Google Scholar] [CrossRef]
  16. Mayer, M.G.; Jensen, J.H.D. Elementary Theory of Nuclear Shell Structure; Wiley: New York, NY, USA, 1955. [Google Scholar]
  17. Wybourne, B.G. Classical Groups for Physicists; Wiley: New York, NY, USA, 1974. [Google Scholar]
  18. Moshinsky, M.; Smirnov, Y.F. The Harmonic Oscillator in Modern Physics; Harwood: Amsterdam, The Netherlands, 1996. [Google Scholar]
  19. Iachello, F. Lie Algebras and Applications; Springer: Berlin, Germany, 2006. [Google Scholar]
  20. Sorlin, O.; Porquet, M.-G. Nuclear magic numbers: New features far from stability. Prog. Part. Nucl. Phys. 2008, 61, 602. [Google Scholar] [CrossRef] [Green Version]
  21. Martinou, A.; Bonatsos, D.; Minkov, N.; Assimakis, I.E.; Peroulis, S.K.; Sarantopoulou, S. Proxy-SU(3) symmetry in the shell model basis. J. Cseh. Eur. Phys. J. A 2020, 56, 239. [Google Scholar] [CrossRef]
  22. Nilsson, S.G. Binding states of individual nucleons in strongly deformed nuclei. Mat. Fys. Medd. K. Dan. Vidensk. Selsk. 1955, 29, 16. [Google Scholar]
  23. Nilsson, S.G.; Ragnarsson, I. Shapes and Shells in Nuclear Structure; Cambridge University Press: Cambridge, MA, USA, 1995. [Google Scholar]
  24. Lederer, C.M.; Shirley, V.S. (Eds.) Table of Isotopes, 7th ed.; Wiley: New York, NY, USA, 1978. [Google Scholar]
  25. 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] [Green Version]
  26. Elliott, J.P. Collective motion in the nuclear shell model. I. Classification schemes for states of mixed configurations. Proc. Roy. Soc. Lond. Ser. A 1958, 245, 128. [Google Scholar]
  27. Elliott, J.P. Collective motion in the nuclear shell model II. The introduction of intrinsic wave-functions. Proc. Roy. Soc. Lond. Ser. A 1958, 245, 562. [Google Scholar]
  28. Elliott, J.P.; Harvey, M. Collective motion in the nuclear shell model III. The calculation of spectra. Proc. Roy. Soc. Lond. Ser. A 1963, 272, 557. [Google Scholar]
  29. Elliott, J.P.; Wilsdon, C.E. Collective motion in the nuclear shell model IV. Odd-mass nuclei in the sd shell. Proc. Roy. Soc. Lond. Ser. A 1968, 302, 509. [Google Scholar]
  30. 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]
  31. Kota, V.K.B. SU(3) Symmetry in Atomic Nuclei; Springer: Singapore, 2020. [Google Scholar]
  32. 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]
  33. Castaños, O.; Draayer, J.P.; Leschber, Y. Shape variables and the shell model. Z. Phys. A 1988, 329, 33. [Google Scholar]
  34. 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]
  35. 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(R). [Google Scholar] [CrossRef]
  36. 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]
  37. Sheline, R.K. Vibrational states in deformed even-even nuclei. Rev. Mod. Phys. 1960, 32, 1. [Google Scholar] [CrossRef]
  38. Zamfir, N.V.; Bucurescu, D.; Casten, R.F.; Ivasvu, M. A simple phenomenology for 2 γ + states. Phys. Lett. B 1990, 241, 463. [Google Scholar] [CrossRef]
  39. Ring, P. Relativistic mean field theory in finite nuclei. Prog. Part. Nucl. Phys. 1996, 37, 193. [Google Scholar] [CrossRef]
  40. 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]
  41. 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] [CrossRef]
  42. Meng, J.; Toki, H.; Zhou, S.G.; Zhang, S.Q.; Long, W.H.; Geng, L.S. Relativistic continuum Hartree Bogoliubov theory for ground-state properties of exotic nuclei. Prog. Part. Nucl. Phys. 2006, 57, 470. [Google Scholar] [CrossRef] [Green Version]
  43. Nikšić, T.; Vretenar, D.; Ring, P. Relativistic nuclear energy density functionals: Mean-field and beyond. Prog. Part. Nucl. Phys. 2011, 66, 519. [Google Scholar] [CrossRef] [Green Version]
  44. Meng, J.; Zhou, S.G. Halos in medium-heavy and heavy nuclei with covariant density functional theory in continuum. J. Phys. G Nucl. Part. Phys. 2015, 42, 093101. [Google Scholar] [CrossRef]
  45. Liang, H.; Meng, J.; Zhou, S.-G. Hidden pseudospin and spin symmetries and their origins in atomic nuclei. Phys. Rep. 2015, 570, 1. [Google Scholar] [CrossRef] [Green Version]
  46. Lalazissis, G.A.; Nikšić, T.; Paar, N.; Vretenar, D.; Ring, P. New relativistic mean-field interaction with density-dependent meson-nucleon couplings. Phys. Rev. C 2005, 71, 024312. [Google Scholar] [CrossRef] [Green Version]
  47. Nikšić, 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] [Green Version]
  48. 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] [Green Version]
  49. 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] [Green Version]
  50. 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] [CrossRef]
  51. 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] [Green Version]
  52. 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, G. Blocking of coupling to the 0 2 + excitation in 154Gd by the [505]11/2- neutron in 155Gd. J. Eur. Phys. J. A 2011, 47, 6. [Google Scholar] [CrossRef] [Green Version]
  53. 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]
  54. Garrett, P.E.; Green, K.L.; Austin, R.A.E.; Ball, G.C.; Bandyopadhyay, D.S.; Colosimo, S.; Cross, D.S.; Demand, G.A.; Grinyer, G.F.; Hackman, G.; et al. 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]
  55. McCutchan, E.A.; Bonatsos, D.; Zamfir, N.V.; Casten, R.F. Staggering in γ-band energies and the transition between different structural symmetries in nuclei. Phys. Rev. C 2007, 76, 024306. [Google Scholar] [CrossRef]
  56. 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. (N. Y.) 1984, 156, 41. [Google Scholar] [CrossRef]
Figure 1. Regions in which shape coexistence can appear, according to the dual shell mechanism of Section 3. The colored regions possess proton or neutron numbers within the intervals 7–8, 17–20, 34–40, 59–70, 96–112, and 145–168. The horizontal stripes correspond to the proton-induced SC, while the vertical stripes correspond to the neutron-induced SC. Adapted from Ref. [8]. See Section 3 for further discussion.
Figure 1. Regions in which shape coexistence can appear, according to the dual shell mechanism of Section 3. The colored regions possess proton or neutron numbers within the intervals 7–8, 17–20, 34–40, 59–70, 96–112, and 145–168. The horizontal stripes correspond to the proton-induced SC, while the vertical stripes correspond to the neutron-induced SC. Adapted from Ref. [8]. See Section 3 for further discussion.
Symmetry 15 00029 g001
Figure 2. Islands of shape coexistence found through covariant density functional theory calculations, as described in Section 5. Islands corresponding to neutron-induced SC are shown in yellow, islands due to proton-induced SC are exhibited in azure, while islands due to both mechanisms are shown in green. The stripes in which SC is allowed according to the dual shell mechanism are shown in grey. Adapted from Ref. [11]. See Section 5 for further discussion.
Figure 2. Islands of shape coexistence found through covariant density functional theory calculations, as described in Section 5. Islands corresponding to neutron-induced SC are shown in yellow, islands due to proton-induced SC are exhibited in azure, while islands due to both mechanisms are shown in green. The stripes in which SC is allowed according to the dual shell mechanism are shown in grey. Adapted from Ref. [11]. See Section 5 for further discussion.
Symmetry 15 00029 g002
Figure 3. Nuclei with well-developed K = 0 bands (blue boxes), also listed in Table 1, are given on the proton-neutron map including the stripes within which proton-induced (yellow) or neutron-induced (azure) SC can be expected. Based on data taken from Ref. [12]. See Section 6 for further discussion.
Figure 3. Nuclei with well-developed K = 0 bands (blue boxes), also listed in Table 1, are given on the proton-neutron map including the stripes within which proton-induced (yellow) or neutron-induced (azure) SC can be expected. Based on data taken from Ref. [12]. See Section 6 for further discussion.
Symmetry 15 00029 g003
Figure 4. Nuclei with well-developed K = 2 bands (blue boxes), also listed in Table 2, are given on the proton-neutron map including the stripes within which proton-induced (yellow) or neutron-induced (azure) SC can be expected. Based on data taken from Ref. [12]. See Section 6 for further discussion.
Figure 4. Nuclei with well-developed K = 2 bands (blue boxes), also listed in Table 2, are given on the proton-neutron map including the stripes within which proton-induced (yellow) or neutron-induced (azure) SC can be expected. Based on data taken from Ref. [12]. See Section 6 for further discussion.
Symmetry 15 00029 g004
Figure 5. Nuclei with energy differences E ( 0 2 + ) E ( 2 1 + ) less than 800 keV (blue boxes), are given on the proton-neutron map including the stripes within which proton-induced (yellow) or neutron-induced (azure) SC can be expected. Based on data taken from Ref. [12]. See Section 7 for further discussion.
Figure 5. Nuclei with energy differences E ( 0 2 + ) E ( 2 1 + ) less than 800 keV (blue boxes), are given on the proton-neutron map including the stripes within which proton-induced (yellow) or neutron-induced (azure) SC can be expected. Based on data taken from Ref. [12]. See Section 7 for further discussion.
Symmetry 15 00029 g005
Table 1. Nuclei with experimentally known [12] well-developed K = 0 bands, also shown in Figure 3. For each nucleus, the angular momentum L of the highest state known in the K = 0 band is shown. Based on data taken from Ref. [12]. See Section 6 for further discussion. The nuclei in which the energy difference E ( 0 2 + ) E ( 2 1 + ) is less than 800 keV are shown in boldface and are depicted in Figure 5. See Section 7 for further discussion of them.
Table 1. Nuclei with experimentally known [12] well-developed K = 0 bands, also shown in Figure 3. For each nucleus, the angular momentum L of the highest state known in the K = 0 band is shown. Based on data taken from Ref. [12]. See Section 6 for further discussion. The nuclei in which the energy difference E ( 0 2 + ) E ( 2 1 + ) is less than 800 keV are shown in boldface and are depicted in Figure 5. See Section 7 for further discussion of them.
nuc.Lnuc.Lnuc.Lnuc.Lnuc.Lnuc.Lnuc.L
36 Ar16 100 Zr12 124 Xe4 162 Gd2 166 Yb10 176 W12 180 Pt6
40 Ar12 100 Mo2 126 Xe4 154 Dy12 168 Yb4 178 W22 182 Pt8
40 Ca8 102 Mo2 132 Ce4 156 Dy10 170 Yb18 182 W4 184 Pt6
42 Ca12 104 Ru4 134 Ce2 158 Dy8 172 Yb14 184 W4 186 Pt6
44 Ti4 110 Pd6 146 Ce6 160 Dy4 174 Yb4 186 W4 196 Pt2
68 Ge10 112 Pd4 148 Nd8 162 Dy28 176 Yb2 172 Os8 184 Hg20
70 Ge8 110 Cd6 150 Nd6 166 Dy2 178 Yb4 174 Os6 186 Hg28
72 Ge6 112 Cd8 152 Nd4 156 Er22 168 Hf4 176 Os6 188 Hg22
82 Ge6 116 Cd2 152 Sm16 158 Er4 170 Hf4 178 Os6 190 Hg6
72 Se2 114 Sn30 154 Sm6 160 Er2 172 Hf4 180 Os6 184 Pb8
76 Kr2 116 Sn14 152 Gd10 162 Er2 174 Hf26 184 Os6 186 Pb14
96 Sr18 118 Sn12 154 Gd10 164 Er10 176 Hf10 186 Os10 188 Pb14
98 Sr6 122 Te18 156 Gd14 166 Er12 178 Hf6 190 Os2 194 Pb6
96 Zr6 118 Xe4 158 Gd6 168 Er6 180 Hf14 192 Os2 196 Pb8
98 Zr12 120 Xe4 160 Gd4 170 Er22 170 W6 178 Pt8
Table 2. Nuclei with experimentally known [12] well-developed K = 2 bands, also shown in Figure 4. For each nucleus, the angular momenta L e and L o of the highest even and odd states known in the K = 2 band are shown. Based on data taken from Ref. [12]. See Section 6 for further discussion.
Table 2. Nuclei with experimentally known [12] well-developed K = 2 bands, also shown in Figure 4. For each nucleus, the angular momenta L e and L o of the highest even and odd states known in the K = 2 band are shown. Based on data taken from Ref. [12]. See Section 6 for further discussion.
nuc. L e L o nuc. L e L o nuc. L e L o nuc. L e L o nuc. L e L o nuc. L e L o nuc. L e L o
38 Ar8 84 Zr621 116 Pd813 134 Ce87 158 Dy87 166 Hf25 188 Os67
40 Ca 13 86 Mo25 118 Pd45 132 Nd43 160 Dy2225 168 Hf65 190 Os105
58 Fe43 88 Mo25 112 Cd1011 134 Nd443 162 Dy1817 170 Hf43 192 Os107
62 Zn67 102 Zr43 110 Te6 136 Nd45 164 Dy149 172 Hf65 180 Pt49
64 Zn8 100 Mo43 116 Te65 138 Nd65 166 Dy45 174 Hf45 182 Pt67
64 Ge8 102 Mo6 114 Xe20 136 Sm14 156 Er1415 176 Hf87 184 Pt67
66 Ge10 104 Mo1817 116 Xe6 138 Sm87 158 Er107 178 Hf1415 186 Pt109
68 Ge47 106 Mo1817 118 Xe1411 138 Gd22 160 Er213 180 Hf2015 188 Pt243
70 Ge85 108 Mo1211 120 Xe285 140 Gd87 162 Er1211 170 W43 190 Pt65
68 Se14 110 Mo45 122 Xe1415 140 Xe 13 164 Er1821 176 W45 192 Pt87
70 Se14 100 Ru89 124 Xe1021 146 Ce45 166 Er1413 178 W45 194 Pt45
74 Se619 102 Ru1011 126 Xe87 148 Nd63 168 Er127 180 W67 196 Pt85
76 Se89 104 Ru87 128 Xe65 150 Nd43 170 Er1819 182 W85 198 Pt8
78 Se109 106 Ru87 122 Ba411 152 Sm1211 172 Er25 184 W125 182 Hg20
74 Kr 7 108 Ru1013 124 Ba1011 154 Sm67 160 Yb23 186 W125 186 Hg27
76 Kr89 110 Ru1415 126 Ba169 152 Gd813 162 Yb23 188 W4 180 Pb8
78 Kr109 112 Ru1619 128 Ba149 154 Gd67 164 Yb413 172 Os65 182 Pb12
80 Kr67 114 Ru69 130 Ba105 156 Gd1215 166 Yb1213 174 Os45 190 Pb8
82 Kr45 116 Ru65 132 Ba127 158 Gd65 168 Yb629 176 Os45 194 Po6
84 Kr10 118 Ru43 124 Ce32 160 Gd127 170 Yb1617 178 Os45 198 Po4
80 Sr1021 108 Pd107 126 Ce43 162 Gd43 172 Yb45 180 Os69
82 Sr209 110 Pd85 128 Ce189 166 Gd45 174 Yb45 182 Os67
84 Sr49 112 Pd1013 130 Ce26 154 Dy67 176 Yb45 184 Os65
82 Zr 13 114 Pd1013 132 Ce145 156 Dy1215 156 Hf6 186 Os1213
Table 3. Nuclei with experimentally known [12] well-developed K = 0 bands, also shown in Table 1 and Figure 3. For each nucleus, the energies of the 2 + state of the ground state band ( 2 g + ), the 0 + state of the K = 0 ( β ) band ( 0 β + ), and the 2 + state of the K = 2 ( γ ) band ( 2 γ + ), taken from Ref. [12], are shown in keV. The nuclei in which the energy difference E ( 0 2 + ) E ( 2 1 + ) is less than 800 keV are shown in boldface and are depicted in Figure 5. See Section 7 for further discussion.
Table 3. Nuclei with experimentally known [12] well-developed K = 0 bands, also shown in Table 1 and Figure 3. For each nucleus, the energies of the 2 + state of the ground state band ( 2 g + ), the 0 + state of the K = 0 ( β ) band ( 0 β + ), and the 2 + state of the K = 2 ( γ ) band ( 2 γ + ), taken from Ref. [12], are shown in keV. The nuclei in which the energy difference E ( 0 2 + ) E ( 2 1 + ) is less than 800 keV are shown in boldface and are depicted in Figure 5. See Section 7 for further discussion.
nuc. 2 g + 0 β + 2 γ + nuc. 2 g + 0 β + 2 γ + nuc. 2 g + 0 β + 2 γ + nuc. 2 g + 0 β + 2 γ +
36 Ar1970.44329.1 118 Sn1229.71758.3 156 Er344.5930.1930.5 184 W111.21002.5903.3
40 Ar1460.82120.9 122 Te564.11357.4 158 Er192.2806.4820.1 186 W122.6883.6738.0
40 Ca5248.83352.6 118 Xe337.3830.4928.1 160 Er125.8893.6854.4 172 Os227.8758.3918.8
42 Ca1524.71837.3 120 Xe322.6908.7876.1 162 Er102.01087.2900.7 174 Os158.6545.3846.2
44 Ti1083.11904.3 124 Xe354.01268.9846.5 164 Er91.41246.0860.3 176 Os135.1601.2863.6
68 Ge1015.81754.51777.4 126 Xe388.61313.9879.9 166 Er80.61460.0785.9 178 Os132.2650.5864.4
70 Ge1039.51215.61707.7 132 Ce325.31158.4822.2 168 Er79.81217.2821.2 180 Os132.1736.4870.4
72 Ge834.0691.4 134 Ce409.21533.5965.7 170 Er78.6890.9934.0 184 Os119.81042.0942.9
82 Ge1348.32333.6 146 Ce258.51043.21381.9 166 Yb102.41043.0932.4 186 Os137.21061.0767.5
72 Se862.1937.2 148 Nd301.7916.91248.9 168 Yb87.71155.2984.0 190 Os186.7911.8558.0
76 Kr424.0769.91221.7 150 Nd130.2675.91062.1 170 Yb84.31069.41145.7 192 Os205.8956.5489.1
96 Sr814.91229.3 152 Nd72.41139.0 172 Yb78.71042.91465.9 178 Pt170.3421.0
98 Sr144.2215.6 152 Sm121.8684.81085.8 174 Yb76.51487.11634.0 180 Pt153.2478.1677.5
96 Zr1750.51581.6 154 Sm82.01099.31440.0 176 Yb82.11139.01260.9 182 Pt155.0499.7667.8
98 Zr1222.9854.0 152 Gd344.3615.41109.2 178 Yb84.01315.0 184 Pt163.0491.8648.7
100 Zr212.5331.1 154 Gd123.1680.7996.3 168 Hf124.1942.1875.9 186 Pt191.5471.5607.2
100 Mo535.6695.11063.8 156 Gd89.01049.51154.2 170 Hf100.8879.6961.3 196 Pt355.71135.3688.7
102 Mo296.6698.3847.9 158 Gd79.51196.21187.1 172 Hf95.2871.31075.3 184 Hg366.8375.1
104 Ru358.0988.3893.1 160 Gd75.31325.7988.4 174 Hf91.0828.11226.8 186 Hg405.3523.01096.6
110 Pd373.8946.7931.2 162 Gd71.61427.0864.0 176 Hf88.31149.91341.3 188 Hg412.8824.5
112 Pd348.71125.5813.6 154 Dy334.3660.61027.0 178 Hf93.21199.41174.6 190 Hg416.31278.6
110 Cd657.81473.1 156 Dy137.8675.6890.5 180 Hf93.31101.91199.7 184 Pb 570.0
112 Cd617.51224.31312.4 158 Dy98.9990.5946.3 170 W156.7 937.1 186 Pb 655.0
116 Cd513.51282.6 160 Dy86.81279.9966.2 176 W108.3843.31040.2 188 Pb723.6591.0
114 Sn1299.91953.3 162 Dy80.71400.3888.2 178 W105.9997.01110.4 194 Pb965.1930.7
116 Sn1293.61756.9 166 Dy76.61149.0857.2 182 W100.11135.81221.4 196 Pb1049.21142.9
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

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. https://doi.org/10.3390/sym15010029

AMA Style

Martinou A, Bonatsos D, Peroulis SK, Karakatsanis KE, Mertzimekis TJ, Minkov N. Islands of Shape Coexistence: Theoretical Predictions and Experimental Evidence. Symmetry. 2023; 15(1):29. https://doi.org/10.3390/sym15010029

Chicago/Turabian Style

Martinou, Andriana, Dennis Bonatsos, Spyridon Kosmas Peroulis, Konstantinos Eleftherios Karakatsanis, Theodoros John Mertzimekis, and Nikolay Minkov. 2023. "Islands of Shape Coexistence: Theoretical Predictions and Experimental Evidence" Symmetry 15, no. 1: 29. https://doi.org/10.3390/sym15010029

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