Next Article in Journal
What Is Inside the Double–Double Structure of the Radio Galaxy J0028+0035?
Previous Article in Journal
Triply Heavy Tetraquark States in a Mass-Splitting Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Continuum Effect on Mirror Symmetry Breaking Within the Gamow Frameworks

1
Institute for Advanced Simulation (IAS-4), Forschungszentrum Jülich, D-52425 Jülich, Germany
2
State Key Laboratory of Nuclear Physics and Technology, School of Physics, Peking University, Beijing 100871, China
3
Key Laboratory of Nuclear Physics and Ion-Beam Application (MOE), Institute of Modern Physics, Fudan University, Shanghai 200433, China
4
Shanghai Research Center for Theoretical Nuclear Physics, NSFC and Fudan University, Shanghai 200438, China
*
Authors to whom correspondence should be addressed.
Symmetry 2025, 17(2), 169; https://doi.org/10.3390/sym17020169
Submission received: 30 December 2024 / Revised: 20 January 2025 / Accepted: 21 January 2025 / Published: 23 January 2025
(This article belongs to the Special Issue Isospin Symmetry/Asymmetry in Experimental Nuclear Physics)

Abstract

:
Nuclear physics provides a natural laboratory for studying two kinds of fermions: protons and neutrons. These particles share similarities in mass and strong nuclear interactions, which are often described by isospin symmetry. However, isospin is not a good quantum number due to the differences between protons and neutrons in charge and quark mass. These differences become more pronounced as we approach or move beyond the dripline, affecting the structures and decay properties of mirror nuclei. To explore these intriguing phenomena, researchers have developed novel theoretical frameworks. In this article, we review the results from the Gamow shell model and Gamow coupled-channel, which account for the mirror symmetry breaking influenced by nuclear forces and continuum effects. Specifically, we discuss the recently observed mirror asymmetries in nuclei at the boundaries of the nuclide landscape and their theoretical explanations. We examine the breaking of mirror symmetry in the spectra of N = 8 isotones versus Z = 8 isotopes, as well as the decay properties of the 22Al-22F mirror pair. Such studies enhance our understanding of strong interactions and the behavior of open quantum systems.

1. Introduction

The concept of isospin, first introduced by Heisenberg to explain the relationship between the proton and the newly discovered neutron [1], has become a cornerstone in nuclear physics. Protons and neutrons, regarded as two manifestations of the same fermion under a strong interaction, exhibit analogous behaviors. This symmetry has been effectively utilized in nuclear systems, yielding significant insights into their structure and dynamics. Thus, nuclei with the same nucleon number but varying protons and neutrons exhibit similar properties. However, isospin is not an exact symmetry, as it is affected by a strong interaction (due to the mass difference between up and down quarks) and an electromagnetic interaction (due to the charge difference). These asymmetries become more pronounced in nuclei with extreme proton-to-neutron ratios, particularly near the dripline where they are weakly bound or unbound, with the impact of the isospin asymmetries significantly shaping their structure and decay properties.
Mirror nuclei, pairs of nuclei in which the numbers of protons and neutrons are exchanged, serve as powerful laboratories for probing isospin symmetry and its breaking. Indeed, experiments have demonstrated remarkable similarities in the behavior of most mirror nuclei. However, with advancements in experimental facilities, an increasing number of nuclei located at the boundaries of the nuclear chart have been studied, unveiling the phenomenon of isospin symmetry breaking in properties such as mass [2,3,4,5], energy spectra [6,7,8,9,10,11], and decay characteristics [12,13,14,15,16,17,18,19,20]. These systems are coupled to continuum states due to the proximity of reaction and decay channels, making them open quantum systems [21,22,23]. Studying these systems provides valuable insights into the evolution of nuclear structure and nuclear symmetry at the limits of nuclear landscape boundaries.
In near-threshold and unbound regimes, conventional shell models (SMs) may no longer offer an adequate description of the nuclear system. Instead, more comprehensive theoretical frameworks are required to incorporate the resonant states and coupling to the continuum, which are critical for characterizing weakly bound and unbound nuclei near the dripline. To account for the continuum effect, the continuum shell model [24,25] was developed by projecting the model space onto the subspaces of bound and scattering states within a real-energy basis. By introducing an interaction term to account for continuum coupling, the continuum-coupled shell model [26] was further proposed to incorporate the continuum effect. The Gamow SM (GSM), an extension of the conventional SM, provides another robust framework for tackling these issues [22,27,28,29]. An essential feature of the GSM is the utilization of the Berggren basis [30], an ensemble that enables the incorporation of bound, resonant, and scattering states on the equal footing. Built on this basis, the GSM provides a full description of the interplay between continuum coupling and many-body correlations via configuration mixing. The GSM has been widely used in the study of weakly bound and unbound nuclei [27,28,31,32,33,34,35,36,37,38,39,40,41,42,43,44]. Meanwhile, the Gamow coupled-channel (GCC) approach in Jacobi coordinates with the Berggren basis was also developed to describe the structures and decays of exotic nuclei [45,46,47], which can treat center-of-mass motion and asymptotic behavior precisely [48]. In addition, significant progress in ab initio calculations [49,50,51,52,53,54,55,56] has been achieved through advancements in chiral effective field theory [57,58], the similarity renormalization group [59,60], and various many-body methods [61,62,63,64,65,66,67,68], allowing calculations in the nuclear medium-mass region. The incorporation of chiral three-body forces [69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89] has further enhanced the accuracy and predictive power of ab initio approaches. Hence, several ab initio methods have been successfully applied to open quantum nuclear systems, including the no-core shell model (NCSM) with the resonating group method [90,91], the single-state harmonic oscillator representation of scattering equations [92,93], the NCSM with a continuum [94,95], and the nuclear lattice effective field theory in coordinate space [88]. Additionally, ab initio methods based on the Berggren basis have also undergone significant development in recent years, including the complex coupled cluster [73,96], the complex in-medium similarity renormalization group [97], the no-core GSM [98,99,100], and the ab initio GSM [101,102,103,104].
This article is organized as follows: In Section 2, we introduce the ab initio GSM approach and the GCC approach. We begin with an overview of the Berggren basis and explain the traditional methods used to generate it in Section 2.1. We then focus on the implementation of MBPT based on the Berggren basis and chiral forces in Section 2.2, which forms the foundation of the ab initio GSM. The following is a brief introduction of the GCC framework in Section 2.3. These approaches enable us to provide a comprehensive description of the interplay between continuum coupling and internucleon correlations. In Section 3, we review the recent experimental and theoretical advancements in the study of mirror symmetry breaking (MSB) in nuclei at the edges of the nuclear landscape. We introduce recent experimental observations of MSB along with a theoretical description based on the GCC and GSM approach in Section 3.1. Additionally, we review the application of the ab initio GSM to the excitation spectra of oxygen isotopes and their mirror partners [105] in Section 3.2, and MSB in β -decay processes involving the 22Si-22Al and 22O-22F nuclear pairs [106] in Section 3.3. We highlight that an accurate treatment of the asymptotic behavior of single-particle (s.p.) wavefunctions, along with their coupling to the continuum, is crucial for the precise characterization of isospin asymmetry.

2. Theoretical Framework

In this section, we outline the theoretical framework used in the mirror symmetry studies of exotic nuclei, focusing on the Berggren basis, GSM, and GCC approaches. These frameworks provide a solid foundation for studying the novel phenomena related to open quantum systems.

2.1. Berggren Basis

The Berggren basis is a generalized basis that extends the traditional basis of bound states by incorporating both resonance and continuum states. This basis is particularly useful for describing weakly bound and unbound nuclear systems, offering a comprehensive foundation for many-body frameworks. The Berggren basis was first introduced by Berggren in 1968 [30]; it provides a complete description of s.p. states, including bound states, resonant states, and the associated continuum states, through a completeness relation, as follows:
n u n ( E n , r ) u n ( E n , r ) + L + d E u ( E , r ) u ( E , r ) = δ ( r r ) ,
where u n ( E n , r ) represents the physical s.p. bound and narrow resonant states located in the fourth quadrant of the complex-k plane, with corresponding energies E n . The parameter L + denotes the integral contour surrounding the narrow resonant states, which consists of a series of continuum states characterized by energy E.
There are generally two methods for constructing the Berggren basis. The first involves selecting an auxiliary s.p. potential in coordinate space to generate resonance and continuum states. The second is the Gamow Hartree–Fock (GHF) method [97,101,104], which derives a self-consistent GHF potential from the chiral forces.

2.1.1. Woods–Saxon

The Woods–Saxon central potential, supplemented by the spin–orbit and Coulomb terms, is often used in actual applications. This potential U ( r ) is a finite-depth potential well, and by appropriately setting the radius parameter, it can effectively describe the nuclear density distribution, making it a close approximation to the realistic mean-field potential experienced by nucleons within the atomic nucleus. The parameters of the potential U ( r ) are chosen for the best fit of nuclear s.p. energies and their decay widths. With this potential, we obtain the following radial s.p. Schrödinger equation:
d 2 u ( k , r ) d r 2 = l ( l + 1 ) r 2 + 2 m 2 U ( r ) k 2 u ( k , r ) ,
where l is the orbital angular momentum, and the l ( l + 1 ) r 2 term provides the centrifugal part. The s.p. wavefunction u ( k , r ) represents the Berggren basis, which includes bound states, resonant states, and continuum states. The general solutions of Equation (2) satisfy the boundary condition, as follows:
u ( k , r ) C 0 r l + 1 , r 0 , C + H l , η ( + ) ( k r ) + C H l , η ( ) ( k r ) , r .
Here, C 0 , C + , and C are normalization constants, where C is zero for bound and resonant states. H l ( + ) ( k r ) and H l ( ) ( k r ) are Coulomb wavefunctions with outgoing and incoming behaviors, while η is the Sommerfeld parameter [29]. By using the shooting method (e.g., [29]), the boundary value problem is transformed into an initial value problem so as to solve the corresponding s.p. wavefunction.

2.1.2. Gamow Hartree–Fock

While the Woods–Saxon parameters need to be determined by fitting experimental s.p. energies and their widths, the self-consistent GHF method is proposed to generate the Berggren basis. To derive the GHF basis, we start from the HF calculation. The s.p. HF equation can be written as
β 1 1 A p α 2 2 m δ α β + γ δ ρ γ δ V α γ β δ NN + 1 2 γ μ δ ν ρ γ δ ρ μ ν V α γ μ β δ ν 3 N ψ ( β ) = e α ψ ( α ) ,
where V NN and V 3 N stand for chiral nucleon–nucleon with the incorporated two-body term p i · p j m A and chiral three nucleon force, respectively. α , γ , μ , β , δ , and ν are indexes of a harmonic oscillator basis. ρ γ δ = i ϵ F γ | i i | δ is the one-body density matrix, with the sum of i running over all hole states below the HF Fermi surface ϵ F of the reference state. From the above equation, it is clear that the HF s.p. basis is constructed from the chiral nucleon–nucleon and three-nucleon nuclear forces in a fully microscopic approach. To further include the continuum coupling, we extend the HF equation to the complex-k plane, which leads to the following GHF equation:
2 k 2 2 μ ψ n l j ( k ) + L + d k k 2 U ( l j k k ) ψ n l j ( k ) = e n l j ψ n l j ( k ) ,
where μ = m / ( 1 1 A ) , and k ( k ) is defined on a contour L + in the fourth quadrant of the complex-k plane [104]. The complex GHF s.p. potential U ( l j k k ) is
U ( l j k k ) = α β k | α α | U | β β | k ,
where l and j are the orbital and total angular momenta of the s.p. wavefunction, respectively. Greek letters denote HO states; α | U | β is the HF potential that is obtained by iterating the real-energy HF Equation (4) [97]. In practical calculations, the integral over k along L + is evaluated using the Gauss–Legendre quadrature method [101]. This approach ultimately yields the self-consistent GHF basis, comprising bound, resonant, and continuum states.

2.2. Ab Initio Gamow Shell Model with Many-Body Perturbation Theory

Many-body methods utilizing the Berggren basis, such as the no-core GSM [98,99,100], the complex coupled cluster [73,96], and the complex in-medium similarity renormalization group [97], have been well developed. However, the non-degeneracy of the Berggren basis and the significant computational cost associated with its application severely limit these methods, restricting their use to relatively light nuclei, closed-shell nuclei, and their neighbors. In this article, we introduce the ab initio Gamow shell model, formulated within the framework of MBPT in the complex plane, which extends the capability to calculate properties of open-shell nuclei.

2.2.1. Model Space Effective Hamiltonian

As a perturbative method, MBPT naturally separates the core and valence nucleons, thereby eliminating the need to explicitly compute the core part [107,108]. This results in a relatively lower computational overhead, making it well suited for calculations involving continuum states. Moreover, when combined with the HF basis, it also demonstrates good order-by-order convergence [104,109]. To perform MBPT in the complex plane, we first transform the A-body Hamiltonian, as follows:
H = i = 1 A 1 1 A p i 2 2 m + i < j A v i j NN p i · p j m A + i < j < k A v i j k 3 N ,
in the HO basis to the Berggren basis. Here, p i is the nucleon momentum in the laboratory coordinate, and m is the nucleon mass, while v NN and v 3 N denote the chiral nucleon–nucleon (NN) and three-nucleon (3N) forces, respectively. We then normal-ordered the Hamiltonian with respect to the reference state, which can be the GHF Slater determinant of the closed core. In this step, the normal-ordered two-body approximation is employed for the Hamiltonian [110], as follows:
H ^ = i = 1 A t i i + 1 2 i , j = 1 A W i j i j NN + 1 6 i , j , k = 1 A W i j k i j k 3 N + p q ( t p q + i = 1 A W p i q i NN + 1 2 i , j = 1 A W p i j q i j 3 N ) : a ^ p a ^ q : + 1 4 p q r s ( W p q r s NN + i = 1 A W p q i r s i 3 N ) : a ^ p a ^ q a ^ s a ^ r : ,
where t i j is the matrix element of the nucleon kinetic energy with a CoM-corrected mass μ = m / ( 1 1 A ) , while W p q r s NN and W p q i r s i 3 N are antisymmetric nucleon–nucleon force (2NF) and three-nucleon force (3NF) matrix elements given in the GHF basis, respectively. p , q , r , s and i , j , k represent generic states and hole states, respectively. a ^ and a ^ represent the creation and annihilation operators, respectively. The colons indicate normal ordering with respect to the reference state. The Hamiltonian (8) serves as the initial input of the MBPT [61,62]. The zero-body part of the Hamiltonian (8), combined with Rayleigh–Schrödinger many-body perturbation theory [111] can be used to estimate the core energy. Subsequently, the one-body and two-body parts are employed to derive the model space Hamiltonian based on the one-body S ^ -box and the two-body Q ^ -box folding diagrams [112,113].
The S ^ -box and Q ^ -box diagrams can be calculated by the nondegenerate EKK method [114]. The model space effective Hamiltonian can ultimately be expressed in terms of Q ( ϵ ) and its κ -th derivatives, as follows:
H eff ( κ ) = P H ^ 0 P + Q ^ ( ϵ ) + n = 1 1 n d n Q ^ ( ϵ ) d ϵ n { H eff ( κ 1 ) ϵ } n ,
where ϵ represents the starting energy. The operator P denotes the projection onto the model space, while Q corresponds to its complementary projection operator. The calculation of the Q ^ -box typically involves treating the two-body terms as small perturbations, often expanding to the second order to obtain the model space interaction. In contrast, the corresponding one-body S ^ -box is usually computed up to the third order [104]. Ultimately, the model space Hamiltonian can be expressed as follows:
H ^ eff = p q ϵ p q a ^ p a ^ q + 1 4 p q r s V p q r s eff a ^ p a ^ q a ^ s a ^ r ,
where p, q, r, and s represent valence particles within the valence space. ϵ p q denotes the valence-space effective single-particle (s.p.) matrix elements, obtained through the S ^ -box with the GHF Hamiltonian given in Equation (5). These elements correspond to valence s.p. energies, albeit with small nonzero off-diagonal components arising from coupling within the same partial wave. The effective interaction matrix elements, V p q r s eff , are derived using the Q ^ -box formalism. The complex-symmetric GSM effective Hamiltonian is diagonalized within the valence space using the Jacobi–Davidson method in the m-scheme [115].
In addition, for the GSM using the Woods–Saxon basis, we directly transfer the Hamiltonian (7) to the Woods–Saxon basis and then use Equation (9) to obtain the model space two-body interaction, while the s.p. part usually selects the s.p. energy given by the Woods–Saxon potential [78,103,116,117,118].

2.2.2. Effective Operators

For other observables, such as Gamow–Teller (GT) transitions, electromagnetic transitions, and multipole moments, their bare operators should also be consistently renormalized into the valence space. This renormalization can be achieved through the so-called Θ ^ -box within the same complex MBPT framework, analogous to the Q ^ -box. The valence-space effective operator, denoted as Θ eff , incorporates contributions from the excluded Q space and can be expressed as
Θ eff = α , β | ψ α Ψ ˜ α | Θ | Ψ β ψ ˜ β | ,
where the valence-space wavefunction | ψ α obtained from diagonalizing H eff is the projection of the full-space wavefunction | Ψ α onto the valence space, i.e., ψ α = P Ψ α .
The Θ ^ -box method [106,119,120] is used to derive effective operators in the valence space. The final perturbative expansion of the effective operator Θ eff can be expressed by the Q ^ -box and Θ ^ -box as
Θ eff = ( P + Q ^ 1 + Q ^ 1 Q ^ 1 + Q ^ 2 Q ^ + Q ^ Q ^ 2 + ) Q ^ Q ^ 1 × ( χ 0 + χ 1 + χ 2 + ) = H eff Q ^ 1 ( χ 0 + χ 1 + χ 2 + ) ,
where χ n is related to Θ ^ -box, Q ^ -box, and their derivatives.
In practical calculations, the χ n series is truncated at the χ 2 order, which has been shown to be sufficient to converge [119]. The Θ ^ -box diagrams are computed up to the third order, consistent with the expansions employed in the calculations of the S ^ -box and Q ^ -box. Due to the inclusion of continuum states, the matrix dimension increases significantly as more valence particles are added to the continuum. To manage this, a maximum of two valence particles are allowed in the continuum, a limitation under which converged results are still achieved [103,104,121]. The MBPT effective operators have been successfully applied in both the harmonic oscillator (HO) basis [119,122] and the Berggren basis [106].

2.3. Gamow Coupled-Channel Approach

In this subsection, the Gamow coupled-channel (GCC) approach [48,123,124], a three-body method utilizing the Berggren basis, is introduced. To properly describe the few-body asymptotic behavior, the wavefunctions are expressed in Jacobi coordinates using the Berggren basis [30]. Expressed in hyperspherical harmonics, the wavefunction of the valence nucleons is written as
Ψ J π = ρ 5 / 2 C γ K ( k ) B γ K ( k , ρ ) Y γ K J M ( Ω ) d k ,
where ρ is the hyper-radius, Y γ K J M ( Ω ) is the hyperspherical harmonics for the hyperangle part, and K is the hyperspherical quantum number. γ = { s 1 , s 2 , S 12 , x , y , L } is a set of quantum numbers other than K, where S 12 is the total spin of the valence nucleons and is the orbital angular momentum in the corresponding Jacobi coordinate. The Berggren ensemble B n γ K ( ρ ) defines a complete basis in the complex-momentum k plane, which includes bound, decaying, and scattering states [28,30]. C γ K ( k ) is the corresponding coefficient for each configuration. Equation (13) takes the integral over continuous momenta k (scattering states) and the sum over γ , K, and discretized k (bound and decaying resonant states). By using the Berggren basis, the nuclear wavefunctions in both internal and asymptotic regions can be treated on the same footing, providing the natural connection between nuclear structure and decay aspects of the problem [125,126,127].
The Hamiltonian of the three-body GCC model is
H ^ = i = 1 3 p ^ i 2 2 m i + i = 1 2 V c n r i + V n n r T ^ c . m . ,
where p ^ i 2 / ( 2 m i ) and T ^ c . m . are kinetic operators for each particle and the center of mass of the system, respectively. In Jacobi coordinates, the center of mass is automatically eliminated. V c n and V n n are pairwise interactions for core n and n n , respectively.

3. Results

In this section, we first review recent experimental observations of MSB and their theoretical explanations based on GCC and GSM, including the first observation of unbound 11O [5], the four-proton unbound nucleus 18Mg [10], and the five-proton emitter nucleus 9N [18]. Then, we review the applications of ab initio GSM calculations to weakly bound and unbound nuclei, focusing primarily on the mirror asymmetry in spectra of oxygen isotopes and their mirror nuclei [105], as well as the mirror asymmetry in the β -decay GT transitions [106].

3.1. Mirror Symmetry Breaking in Nuclei at the Edge of the Nuclear Landscape

The extremely proton-rich nucleus 11O has been observed for the first time, characterized by a broad peak with a width of 3.4 MeV [5]. The low-energy structure and decaying properties of 11O were then studied using various approaches [126,128,129,130]. Among them, the GCC approach suggests that the observed peak corresponds to a multiplet of four resonant states with J π = 3 / 2 1 , 5 / 2 1 + , 3 / 2 2 , 5 / 2 2 + [126] (see Figure 1), where J is total angular momentum and π is parity. This multiplet was confirmed later in Ref. [5] using the decay-width analysis.
Meanwhile, a moderate Thomas-Ehrman shift (TES) [6,7] is shown in the spectra of 11O and 11Li. This also impacts the structures of the valence nucleon wavefunctions. As shown in Figure 2, strong correlations between the valence nucleons, manifesting as either diproton or dineutron characteristics, are observed in the two 3 / 2 states of 11O and 11Li. Notably, the dineutron configuration in the bound 11Li is more localized than the diproton configuration in the unbound 12O, even though 11Li is a halo system. Furthermore, the secondary peak strength associated with cigarlike arrangements is significantly diminished in 11O.
The spectra of 12O and its isobaric analog 12Be also indicate an MSB [126,131,132]. In 12O, an excited J π = 1 1 state located between the 0 2 + and the 2 1 + states is predicted. This sequence differs from the level ordering in the mirror system 12Be due to the large TES. The location of the calculated 1 1 state corresponds to the shoulder in the measured invariant mass spectrum of 12O [132]. Because the width of the 1 1 state is similar to that of the 0 2 + state, it might be hidden in the observed peaks attributed to the 0 2 + and 2 1 + states.
In the mid-heavy mass region, extremely proton-rich nuclide 18Mg, a ground-state emitter of four protons, was recently discovered [10]. It exhibits an excited state at an excitation energy of 1.84(14) MeV, likely corresponding to the first excited 2 + state. This nucleus undergoes ground-state decay via two sequential two-proton emissions through the intermediate nucleus 16Ne [10].
To study the structure and two-proton decay mechanism of 18Mg, the GSM and GCC methods were used [10,133,134]. In the latest GCC approach [134], the results show that the ground state of 18Mg has a significant s-wave component due to the continuum coupling. However, the results do not lead to a significant deviation in mirror symmetry in either the structure or the spectroscopy of the 18Mg-18C pair. In GSM [133], the MSBs in carbon isotopes and their mirror partners are also discussed. It is shown that the mixed effects of continuous coupling and Coulomb interactions at the dripline give rise to complex patterns in the isospin multiples. In Figure 3, the calculated spectra exhibit good agreement with experimental data, except for the 3 / 2 + state in 17C, which deviates from the experimental value. MSBs are evident in the states of A = 15 nuclei and in the highest excited states of A = 16 nuclei, characterized by broad decay widths. In contrast, the 0 + and 2 + states of 16Ne and 18Mg exhibit relatively mild MSBs. It is worth noting that the strong continuum coupling in the ground states may lead to a good symmetry of the relative energies in the spectra, as the energy shifts induced by continuum coupling in both the ground and excited states may cancel each other out [86].
On the other hand, a more exotic five-proton emitter, 9N, has been found lately [18], making it one of the most proton-rich isotopes where more than half of its constituent nucleons are unbound [18]. The spectra of 9N, 9He, and 8C are shown in Figure 4. As the mirror partner, the 1 / 2 + ground state of 9He is debated to be a virtual (antibound) state, indicating its status more as a scattering feature rather than as a real state. Due to the presence of the Coulomb force, the 1 / 2 + state in the mirror nucleus 9N exhibits a different physical behavior compared with that in 9He. The GSM interprets the 1 / 2 + state of 9N as a broad resonant state, although the nature of this state remains uncertain.

3.2. Mirror Symmetry Breaking in Excitation Spectra

Although the Coulomb barrier hinders the diffusion of proton wavefunctions, pronounced continuum coupling is evident in the spectra, such as the 3 + excited state in the mirror partner 18Ne-18O [137], and the 1 / 2 + excited state in the mirror partner 19Na-19O. Experimental observations reveal a significant TES in the 1 / 2 + excited state of 19Na [138]. To gain deeper insights into the origins of MSB, we first excluded Coulomb interactions and calculated the excitation spectra of 19Na. In this case, the differences between p p and n n interactions arise solely from the charge-symmetry breaking and charge-independence breaking components in the chiral nuclear force at a next-to-next-to-next-to-leading order [139]. As shown in Figure 5, the calculated excitation spectra of 19Na and 19O exhibit good symmetry. When the Coulomb force is included, the excitation energies of the 5 / 2 + and 3 / 2 + states in 19Na remain nearly unchanged, whereas the 1 / 2 + state shows a significant decrease. Calculations indicate that the 1 / 2 + state of 19Na is primarily composed of a pure 0 d 5 / 2 2 1 s 1 / 2 1 configuration. Due to the absence of a centrifugal barrier for the s-wave component, the wavefunction of the 1 s 1 / 2 orbital is more spatially extended compared with other orbitals. This spatial extension reduces the Coulomb energy contribution for the 1 / 2 + state occupying the 1 s 1 / 2 orbital relative to other excited states, ultimately lowering its position in the energy spectrum. Additionally, the HO basis, due to its localized nature, does not accurately capture the long-range asymptotic behavior of the resonant states. In contrast, the GHF basis, with three-body forces included, can more precisely reproduce the energies of resonant states and effectively describe their long-range asymptotic behavior. In GSM calculations based on the GHF basis, the 1 / 2 + state in 19Na is further reduced in energy due to continuum coupling. The calculations show that the 1 / 2 + state is a resonance with a width of 0.28 MeV, which agrees well with the experimental value of 0.1 MeV. The calculations demonstrate that the inclusion of three-body forces significantly improves the ability of SM calculations to describe experimental results. However, to accurately reproduce the resonant characteristics and TES phenomenon of the 1 / 2 + state in 19Na, it is essential to rigorously account for the asymptotic behavior of the s.p. wavefunction and its coupling to the continuum.
The spectra of 22O and 22Si were further investigated. As shown in Figure 6, the GSM results for the excitation spectrum of 22O show excellent agreement with experimental data. For its mirror nucleus 22Si, however, no experimental data are currently available. GSM calculations predict 22Si to be a proton-rich dripline nucleus, with all its low-lying excited states identified as resonant states. The calculated configuration reveals that in the ground state, the six valence nucleons predominantly occupy the π 0 d 5 / 2 orbital, whereas in the low-lying excited states, one or two nucleons are excited into the π 1 s 1 / 2 orbital. Due to the resonant nature of the π 1 s 1 / 2 s.p. orbital and the absence of a central barrier, the many-body wavefunction containing the π 1 s 1 / 2 orbital component is more spatially extended. This extension contributes to the pronounced TES phenomenon observed in the excited states of 22Si.
The recent experiment observes remarkable β -decay isospin asymmetry in mirror nuclei 22Si-22O [17]. The USDA interaction [140] with isospin-nonconserving (INC) [141] was used to describe the MSB of a low-lying spectrum and β -decay GT transition. INC is related to the loosely bound nature of an s 1 / 2 orbit. As emphasized in Ref. [142], extracting isospin-nonconserving (INC) nuclear effective interactions from spectroscopic data requires careful consideration of the coupling to the many-body continuum in the presence of isospin-conserving nuclear forces. Neglecting or improperly treating these continuum effects can significantly impact the reliability of such analyses. In ref [106], we investigate the spectra of their β -decay mirror daughters 22Al-22F using ab initio GSM with realistic force, which self-consistently incorporates both the continuum coupling and many-body correlations via configuration mixing. Figure 7 compares calculated and experimental MEDs for the 1 + analog states of mirror nuclei 22Al and 22F [106]. The MED of the 1 1 + state is the largest MED in the s d -shell. The SM calculation with USDC [143] cannot reproduce this MED. Including continuum coupling in GSM significantly improves agreement with data compared with standard SM calculations using EM1.8/2.0 [79]. Adjusting INC interactions in USDA can give a good description, which mimic continuum effects, as the s 1 / 2 orbit strongly couples to the continuum. However, standard SM with the USDC interaction underestimates MEDs, emphasizing the importance of continuum coupling.

3.3. Mirror Symmetry Breaking in β -Decay Gamow–Teller Transitions

The β -decay GT transition also exhibits MSB. The MSB was observed in the β -decay GT transition from the ground states of 22Si and 22O to the states 1 1 + of their daughters, 22Al and 22F. Standard SM calculations can reproduce this asymmetry by fine-tuned INC interactions related to the π s 1 / 2 orbit. The ab initio GSM, incorporating resonant and continuum coupling, calculates GT transition matrix elements self-consistently. In the phenomenological calculations of the GT transition, usually a quenching factor is needed to better describe the data [144,145]. As commented in [53,146], in the chiral effective field theory framework, the effect of the coupling of weak interactions to two nucleons can be calculated via two-body currents, which can address the quenching factor in a self-consistent way. In this GSM work [106], the quenching factor from two-body currents is calculated with a Fermi gas approximation [146].
Table 1 presents the calculated | M GT | values for these MSB β -decays. For both nuclei, the | M GT | transition to the 1 1 + state is significantly smaller compared with the 1 2 + state. Additionally, the | M GT | value for 22O →22F is nearly double that for 22Si →22Al when decaying to the 1 1 + state. The SM calculations with the INC USDA interaction successfully capture the observed isospin asymmetry [17]. Similarly, calculations using realistic forces also reproduce the MSB observed in experimental data with high accuracy. Incorporating continuum effects through GSM further improves agreement with the experimental results. While the SM with USDC interaction accurately reproduces the | M GT | values for the 1 2 + states, it underestimates the asymmetry observed in the 1 1 + states.
To investigate the isospin asymmetry in GT transitions at a microscopic level, the configuration of related states have been analyzed. In the SM calculations using EM1.8/2.0, the initial state of 22Si exhibits a dominant proton occupation (5.40 in π 0 d 5 / 2 ). In 22Al, the 1 1 + state shows a larger proton 1 s 1 / 2 occupation but a smaller 0 d 3 / 2 occupation compared with the 1 2 + state. The GT matrix element for transitions to the 1 2 + state is enhanced due to the higher occupation of the d orbital in the 1 2 + state configuration compared with that in the 1 1 + state. GSM calculations further reveal that the proton 1 s 1 / 2 occupation in the 1 1 + state of 22Al is significantly enhanced by continuum coupling, while the occupations in the 1 1 + state of 22F remain unchanged. This enhancement leads to better agreement with the experimental asymmetry.
A systematic investigation of GT transitions for A 20 nuclei, as shown in Figure 8, indicates that GSM and SM yield similar M GT values near β stability, where continuum effects are negligible. For exotic nuclei, however, continuum coupling improves the calculations, as demonstrated in the mirror transitions between 22Si and 22O. The experimental M GT values for decays to the 1 2 + states are comparable, but a significant isospin asymmetry is observed in 1 1 + states, which is better captured by GSM with continuum effects.

4. Conclusions

The intriguing phenomena emerging in nuclei near the dripline challenge our current understanding of nuclear structure. The GCC and GSM, utilizing the Berggren basis, offer robust approaches for describing and supporting experimental observations. Moreover, the ab initio GSM is capable of accounting for continuum coupling and many-body correlations, providing an excellent framework for describing and predicting the behavior of weakly bound and unbound nuclei from chiral forces. In the first part of this paper, we introduce the theoretical framework of the ab initio GSM and GCC developed in recent years. In the second part, we review several extremely proton-rich nuclei discovered in recent years, present the observed MSBs in these nuclei and the corresponding GCC and GSM theoretical explanations, and review the ab initio GSM studies of MSBs in low-excitation spectra and Gamow–Teller transitions of β decays. As powerful approaches for studying weakly bound and unbound nuclei, we aim to further develop these methods and apply them more broadly to the study of dripline nuclei in the future.

Author Contributions

Writing—original draft preparation, S.Z. and Z.X.; writing—review and editing, S.Z., Z.X. and S.W. All authors have read and agreed to the published version of the manuscript.

Funding

The work of Simin Wang was funded by the National Key Research and Development Program (MOST 2022YFA1602303 and MOST 2023YFA1606404) and the National Natural Science Foundation of China under Contract No. 12347106 and No. 12147101. The work of Zhicheng Xu was funded by the National Natural Science Foundation of China under Contract No. 12447122 and the China Postdoctoral Science Foundation under Contract No. 2024M760489. The work of Shuang Zhang was funded in part by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (ERC AdG EXOTIC, grant agreement No. 101018170).

Data Availability Statement

Data are contained within the article.

Acknowledgments

Valuable discussions with Furong Xu and Ulf-G. Meißner are gratefully acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
SMshell model
GSMGamow shell model
GCCGamow coupled-channel
NCSMno-core shell model
MBPTmany-body perturbation theory
MSBmirror symmetry breaking
MEDmirror energy difference
TESThomas-Ehrman shift
GTGamow–Teller
INCisospin-nonconserving
s.p.single particle
GHFGamow Hartree–Fock
2NFnucleon–nucleon force
3NFthree-nucleon force

References

  1. Heisenberg, W. Über den Bau der Atomkerne. I. Z. Für Phys. 1932, 77, 1–11. [Google Scholar] [CrossRef]
  2. Gallant, A.T.; Brodeur, M.; Andreoiu, C.; Bader, A.; Chaudhuri, A.; Chowdhury, U.; Grossheim, A.; Klawitter, R.; Kwiatkowski, A.A.; Leach, K.G.; et al. Breakdown of the Isobaric Multiplet Mass Equation for the A=20 and 21 Multiplets. Phys. Rev. Lett. 2014, 113, 082501. [Google Scholar] [CrossRef] [PubMed]
  3. Stefan, I.; de Oliveira Santos, F.; Sorlin, O.; Davinson, T.; Lewitowicz, M.; Dumitru, G.; Angélique, J.C.; Angélique, M.; Berthoumieux, E.; Borcea, C.; et al. Probing Nuclear forces beyond the drip-line using the mirror nuclei 16N and 16F. Phys. Rev. C 2014, 90, 014307. [Google Scholar] [CrossRef]
  4. Hoff, D.E.M.; Rogers, A.M.; Wang, S.M.; Bender, P.C.; Brandenburg, K.; Childers, K.; Clark, J.A.; Dombos, A.C.; Doucet, E.R.; Jin, S.; et al. Mirror-symmetry violation in bound nuclear ground states. Nature 2020, 580, 52–55. [Google Scholar] [CrossRef]
  5. Webb, T.B.; Wang, S.M.; Brown, K.W.; Charity, R.J.; Elson, J.M.; Barney, J.; Cerizza, G.; Chajecki, Z.; Estee, J.; Hoff, D.E.M.; et al. First observation of unbound 11O, the mirror of the halo nucleus 11Li. Phys. Rev. Lett. 2019, 122, 122501. [Google Scholar] [CrossRef]
  6. Thomas, R.G. An Analysis of the Energy Levels of the Mirror Nuclei, C13 and N13. Phys. Rev. 1952, 88, 1109–1125. [Google Scholar] [CrossRef]
  7. Ehrman, J.B. On the Displacement of Corresponding Energy Levels of C13 and N13. Phys. Rev. 1951, 81, 412–416. [Google Scholar] [CrossRef]
  8. Auerbach, N.; Vinh Mau, N. About Coulomb energy shifts in halo nuclei. Phys. Rev. C 2000, 63, 017301. [Google Scholar] [CrossRef]
  9. Grigorenko, L.V.; Mukha, I.G.; Thompson, I.J.; Zhukov, M.V. Two-Proton Widths of O-12, N-16e, and Three-Body Mechanism of Thomas-Ehrman Shift. Phys. Rev. Lett. 2002, 88, 042502. [Google Scholar] [CrossRef] [PubMed]
  10. Jin, Y.; Niu, C.Y.; Brown, K.W.; Li, Z.H.; Hua, H.; Anthony, A.K.; Barney, J.; Charity, R.J.; Crosby, J.; Dell’Aquila, D.; et al. First Observation of the Four-Proton Unbound Nucleus Mg18. Phys. Rev. Lett. 2021, 127, 262502. [Google Scholar] [CrossRef] [PubMed]
  11. Lalanne, L.; Sorlin, O.; Poves, A.; Assié, M.; Hammache, F.; Koyama, S.; Suzuki, D.; Flavigny, F.; Girard-Alcindor, V.; Lemasson, A.; et al. Structure of Ca36 under the Coulomb Magnifying Glass. Phys. Rev. Lett. 2022, 129, 122501. [Google Scholar] [CrossRef] [PubMed]
  12. Trinder, W.; Adelberger, E.; Janas, Z.; Keller, H.; Krumbholz, K.; Kunze, V.; Magnus, P.; Meissner, F.; Piechaczek, A.; Pfützner, M.; et al. β-decay of37Ca. Phys. Lett. B 1995, 349, 267–271. [Google Scholar] [CrossRef]
  13. Tripathi, V.; Tabor, S.L.; Volya, A.; Liddick, S.N.; Bender, P.C.; Larson, N.; Prokop, C.; Suchyta, S.; Tai, P.L.; VonMoss, J.M. Split Isobaric Analog State in Ni55: Case of Strong Isospin Mixing. Phys. Rev. Lett. 2013, 111, 262501. [Google Scholar] [CrossRef] [PubMed]
  14. Orrigo, S.E.A.; Rubio, B.; Fujita, Y.; Blank, B.; Gelletly, W.; Agramunt, J.; Algora, A.; Ascher, P.; Bilgier, B.; Cáceres, L.; et al. Observation of the β-Delayed γ-Proton Decay of 56Zn and its Impact on the Gamow-Teller Strength Evaluation. Phys. Rev. Lett. 2014, 112, 222501. [Google Scholar] [CrossRef] [PubMed]
  15. Bennett, M.B.; Wrede, C.; Brown, B.A.; Liddick, S.N.; Pérez-Loureiro, D.; Bardayan, D.W.; Chen, A.A.; Chipps, K.A.; Fry, C.; Glassman, B.E.; et al. Isospin Mixing Reveals 30P(p,γ)31S Resonance Influencing Nova Nucleosynthesis. Phys. Rev. Lett. 2016, 116, 102502. [Google Scholar] [CrossRef]
  16. Liu, J.J.; Xu, X.X.; Sun, L.J.; Yuan, C.X.; Kaneko, K.; Sun, Y.; Liang, P.F.; Wu, H.Y.; Shi, G.Z.; Lin, C.J.; et al. Observation of a Strongly Isospin-Mixed Doublet in Si26 via β-Delayed Two-Proton Decay of P26. Phys. Rev. Lett. 2022, 129, 242502. [Google Scholar] [CrossRef] [PubMed]
  17. Lee, J.; Xu, X.X.; Kaneko, K.; Sun, Y.; Lin, C.J.; Sun, L.J.; Liang, P.F.; Li, Z.H.; Li, J.; Wu, H.Y.; et al. Large Isospin Asymmetry in 22Si/22O Mirror Gamow-Teller Transitions Reveals the Halo Structure of 22Al. Phys. Rev. Lett. 2020, 125, 192503. [Google Scholar] [CrossRef]
  18. Charity, R.J.; Wylie, J.; Wang, S.M.; Webb, T.B.; Brown, K.W.; Cerizza, G.; Chajecki, Z.; Elson, J.M.; Estee, J.; Hoff, D.E.M.; et al. Strong Evidence for 9N and the Limits of Existence of Atomic Nuclei. Phys. Rev. Lett. 2023, 131, 172501. [Google Scholar] [CrossRef]
  19. Okołowicz, J.; Michel, N.; Nazarewicz, W.; Płoszajczak, M. Asymptotic normalization coefficients and continuum coupling in mirror nuclei. Phys. Rev. C 2012, 85, 064320. [Google Scholar] [CrossRef]
  20. Wang, S.M.; Nazarewicz, W. Fermion Pair Dynamics in Open Quantum Systems. Phys. Rev. Lett. 2021, 126, 142501. [Google Scholar] [CrossRef]
  21. Dobaczewski, J.; Michel, N.; Nazarewicz, W.; Płoszajczak, M.; Rotureau, J. Shell structure of exotic nuclei. Prog. Part. Nucl. Phys. 2007, 59, 432–445. [Google Scholar] [CrossRef]
  22. Michel, N.; Nazarewicz, W.; Ploszajczak, M.; Vertse, T. Shell Model in the Complex Energy Plane. J. Phys. G 2009, 36, 013101. [Google Scholar] [CrossRef]
  23. Forssén, C.; Hagen, G.; Hjorth-Jensen, M.; Nazarewicz, W.; Rotureau, J. Living on the edge of stability, the limits of the nuclear landscape. Phys. Scr. 2013, 2013, 014022. [Google Scholar] [CrossRef]
  24. Rotureau, J.; Okolowicz, J.; Ploszajczak, M. Microscopic theory of the two-proton radioactivity. Phys. Rev. Lett. 2005, 95, 042503. [Google Scholar] [CrossRef]
  25. Volya, A.; Zelevinsky, V. Discrete and continuum spectra in the unified shell model approach. Phys. Rev. Lett. 2005, 94, 052501. [Google Scholar] [CrossRef] [PubMed]
  26. Tsukiyama, K.; Otsuka, T.; Fujimoto, R. Low-lying continuum states of drip-line Oxygen isotopes. PTEP 2015, 2015, 093D01. [Google Scholar] [CrossRef]
  27. Id Betan, R.; Liotta, R.J.; Sandulescu, N.; Vertse, T. Two particle resonant states in a many body mean field. Phys. Rev. Lett. 2002, 89, 042501. [Google Scholar] [CrossRef]
  28. Michel, N.; Nazarewicz, W.; Ploszajczak, M.; Bennaceur, K. Gamow shell model description of neutron rich nuclei. Phys. Rev. Lett. 2002, 89, 042502. [Google Scholar] [CrossRef] [PubMed]
  29. Michel, N.; Płoszajczak, M. Gamow Shell Model: The Unified Theory of Nuclear Structure and Reactions; Springer: Cham, Switzerland, 2021; Volume 983. [Google Scholar] [CrossRef]
  30. Berggren, T. On the use of resonant states in eigenfunction expansions of scattering and reaction amplitudes. Nucl. Phys. A 1968, 109, 265–287. [Google Scholar] [CrossRef]
  31. Michel, N.; Nazarewicz, W.; Ploszajczak, M.; Okolowicz, J. Gamow shell model description of weakly bound nuclei and unbound nuclear states. Phys. Rev. C 2003, 67, 054311. [Google Scholar] [CrossRef]
  32. Michel, N.; Nazarewicz, W.; Ploszajczak, M.; Rotureau, J. Antibound states and halo formation in the Gamow Shell Model. Phys. Rev. C 2006, 74, 054305. [Google Scholar] [CrossRef]
  33. Fossez, K.; Rotureau, J.; Michel, N.; Nazarewicz, W. Continuum effects in neutron-drip-line oxygen isotopes. Phys. Rev. C 2017, 96, 024308. [Google Scholar] [CrossRef]
  34. Fossez, K.; Rotureau, J.; Nazarewicz, W. Energy Spectrum of Neutron-Rich Helium Isotopes: Complex Made Simple. Phys. Rev. C 2018, 98, 061302. [Google Scholar] [CrossRef]
  35. Mao, X.; Rotureau, J.; Nazarewicz, W.; Michel, N.; Id Betan, R.M.; Jaganathen, Y. Gamow Shell Model description of Li isotopes and their mirror partners. Phys. Rev. C 2020, 102, 024309. [Google Scholar] [CrossRef]
  36. Affranchino, S.; Id Betan, R.M. Neutron-pair structure in the continuum spectrum of 26O. Phys. Rev. C 2020, 102, 044330. [Google Scholar] [CrossRef]
  37. Michel, N.; Li, J.G.; Xu, F.R.; Zuo, W. Two-neutron halo structure of F31. Phys. Rev. C 2020, 101, 031301. [Google Scholar] [CrossRef]
  38. Wylie, J.; Okołowicz, J.; Nazarewicz, W.; Płoszajczak, M.; Wang, S.M.; Mao, X.; Michel, N. Spectroscopic factors in dripline nuclei. Phys. Rev. C 2021, 104, L061301. [Google Scholar] [CrossRef]
  39. Li, J.G.; Michel, N.; Zuo, W.; Xu, F.R. Unbound spectra of neutron-rich oxygen isotopes predicted by the Gamow shell model. Phys. Rev. C 2021, 103, 034305. [Google Scholar] [CrossRef]
  40. Li, H.H.; Li, J.G.; Michel, N.; Zuo, W. Investigation of unbound hydrogen isotopes with the Gamow shell model. Phys. Rev. C 2021, 104, L061306. [Google Scholar] [CrossRef]
  41. Linares Fernandez, J.P.; Michel, N.; Płoszajczak, M.; Mercenne, A. Description of Be7 and Li7 within the Gamow shell model. Phys. Rev. C 2023, 108, 044616. [Google Scholar] [CrossRef]
  42. Michel, N.; Nazarewicz, W.; Płoszajczak, M. Description of the Proton-Decaying 02+ Resonance of the α Particle. Phys. Rev. Lett. 2024, 131, 242502. [Google Scholar] [CrossRef]
  43. Xie, M.R.; Li, J.G.; Michel, N.; Li, H.H.; Wang, S.T.; Ong, H.J.; Zuo, W. Investigation of spectroscopic factors of deeply-bound nucleons in drip-line nuclei with the Gamow shell model. Phys. Lett. B 2023, 839, 137800. [Google Scholar] [CrossRef]
  44. Li, X.; Michel, N.; Li, J.G.; Zhou, X.R. Gamow shell model description of neutron-rich He hyper-isotopes. arXiv 2024, arXiv:2408.16223. [Google Scholar]
  45. Pfützner, M.; Mukha, I.; Wang, S. Two-proton emission and related phenomena. Prog. Part. Nucl. Phys. 2023, 132, 104050. [Google Scholar] [CrossRef]
  46. Zhou, L.; Wang, S.M.; Fang, D.Q.; Ma, Y.G. Recent progress in two-proton radioactivity. Nucl. Sci. Tech. 2022, 33, 105. [Google Scholar] [CrossRef]
  47. Shi, G.Z.; Liu, J.J.; Lin, Z.Y.; Zhu, H.F.; Xu, X.X.; Sun, L.J.; Liang, P.F.; Lin, C.J.; Lee, J.; Yuan, C.X.; et al. β-delayed two-proton decay of 27S at the proton-dripline. Phys. Rev. C 2021, 103, L061301. [Google Scholar] [CrossRef]
  48. Wang, S.M.; Michel, N.; Nazarewicz, W.; Xu, F.R. Structure and decays of nuclear three-body systems: The Gamow coupled-channel method in Jacobi coordinates. Phys. Rev. C 2017, 96, 044307. [Google Scholar] [CrossRef]
  49. Binder, S.; Langhammer, J.; Calci, A.; Roth, R. Ab initio path to heavy nuclei. Phys. Lett. B 2014, 736, 119–123. [Google Scholar] [CrossRef]
  50. Hagen, G.; Ekström, A.; Forssén, C.; Jansen, G.R.; Nazarewicz, W.; Papenbrock, T.; Wendt, K.A.; Bacca, S.; Barnea, N.; Carlsson, B.; et al. Neutron and weak-charge distributions of the 48Ca nucleus. Nat. Phys. 2016, 12, 186–190. [Google Scholar] [CrossRef]
  51. Hagen, G.; Jansen, G.R.; Papenbrock, T. Structure of 78Ni from First-Principles Computations. Phys. Rev. Lett. 2016, 117, 172501. [Google Scholar] [CrossRef]
  52. Morris, T.D.; Simonis, J.; Stroberg, S.R.; Stumpf, C.; Hagen, G.; Holt, J.D.; Jansen, G.R.; Papenbrock, T.; Roth, R.; Schwenk, A. Structure of the Lightest Tin Isotopes. Phys. Rev. Lett. 2018, 120, 152503. [Google Scholar] [CrossRef] [PubMed]
  53. Gysbers, P.; Hagen, G.; Holt, J.D.; Jansen, G.R.; Morris, T.D.; Navrátil, P.; Papenbrock, T.; Quaglioni, S.; Schwenk, A.; Stroberg, S.R.; et al. Discrepancy between experimental and theoretical β-decay rates resolved from first principles. Nat. Phys. 2019, 15, 428–431. [Google Scholar] [CrossRef]
  54. Hu, B.S.; Jiang, W.G.; Miyagi, T.; Sun, Z.H.; Ekström, A.; Forssén, C.; Hagen, G.; Holt, J.D.; Papenbrock, T.; Stroberg, S.R.; et al. Ab initio predictions link the neutron skin of 208Pb to nuclear forces. Nat. Phys. 2022, 18, 1196–1200. [Google Scholar] [CrossRef] [PubMed]
  55. Lu, B.N.; Li, N.; Elhatisari, S.; Lee, D.; Epelbaum, E.; Meißner, U.G. Essential elements for nuclear binding. Phys. Lett. B 2019, 797, 134863. [Google Scholar] [CrossRef]
  56. Elhatisari, S.; Bovermann, L.; Ma, Y.Z.; Epelbaum, E.; Frame, D.; Hildenbr, F.; Kim, M.; Kim, Y.; Krebs, H.; Lähde, T.A.; et al. Wavefunction matching for solving quantum many-body problems. Nature 2024, 630, 59–63. [Google Scholar] [CrossRef]
  57. Epelbaum, E.; Hammer, H.W.; Meißner, U.G. Modern theory of nuclear forces. Rev. Mod. Phys. 2009, 81, 1773–1825. [Google Scholar] [CrossRef]
  58. Machleidt, R.; Entem, D. Chiral effective field theory and nuclear forces. Phys. Rep. 2011, 503, 1–75. [Google Scholar] [CrossRef]
  59. Bogner, S.K.; Furnstahl, R.J.; Perry, R.J. Similarity renormalization group for nucleon-nucleon interactions. Phys. Rev. C 2007, 75, 061001. [Google Scholar] [CrossRef]
  60. Bogner, S.; Furnstahl, R.; Schwenk, A. From low-momentum interactions to nuclear structure. Prog. Part. Nucl. Phys. 2010, 65, 94–147. [Google Scholar] [CrossRef]
  61. Hjorth-Jensen, M.; Kuo, T.T.; Osnes, E. Realistic effective interactions for nuclear systems. Phys. Rep. 1995, 261, 125–270. [Google Scholar] [CrossRef]
  62. Coraggio, L.; Covello, A.; Gargano, A.; Itaco, N.; Kuo, T. Shell-model calculations and realistic effective interactions. Prog. Part. Nucl. Phys. 2009, 62, 135–182. [Google Scholar] [CrossRef]
  63. Hagen, G.; Papenbrock, T.; Hjorth-Jensen, M.; Dean, D.J. Coupled-cluster computations of atomic nuclei. Rep. Prog. Phys. 2014, 77, 096302. [Google Scholar] [CrossRef] [PubMed]
  64. Hergert, H.; Bogner, S.; Morris, T.; Schwenk, A.; Tsukiyama, K. The In-Medium Similarity Renormalization Group: A novel ab initio method for nuclei. Phys. Rep. 2016, 621, 165–222. [Google Scholar] [CrossRef]
  65. Dickhoff, W.H.; Muther, H. Nucleon properties in the nuclear medium. Rep. Prog. Phys. 1992, 55, 1947. [Google Scholar] [CrossRef]
  66. Dickhoff, W.; Barbieri, C. Self-consistent Green’s function method for nuclei and nuclear matter. Prog. Part. Nucl. Phys. 2004, 52, 377–496. [Google Scholar] [CrossRef]
  67. Somà, V. Self-Consistent Green’s Function Theory for Atomic Nuclei. Front. Phys. 2020, 8, 340. [Google Scholar] [CrossRef]
  68. Lähde, T.A.; Meißner, U.G. Nuclear Lattice Effective Field Theory: An introduction; Springer: Cham, Switzerland, 2019; Volume 957. [Google Scholar] [CrossRef]
  69. Navrátil, P.; Gueorguiev, V.G.; Vary, J.P.; Ormand, W.E.; Nogga, A. Structure of A=10-13 Nuclei with Two- Plus Three-Nucleon Interactions from Chiral Effective Field Theory. Phys. Rev. Lett. 2007, 99, 042501. [Google Scholar] [CrossRef] [PubMed]
  70. Otsuka, T.; Suzuki, T.; Holt, J.D.; Schwenk, A.; Akaishi, Y. Three-Body Forces and the Limit of Oxygen Isotopes. Phys. Rev. Lett. 2010, 105, 032501. [Google Scholar] [CrossRef] [PubMed]
  71. Roth, R.; Langhammer, J.; Calci, A.; Binder, S.; Navrátil, P. Similarity-Transformed Chiral NN+3N Interactions for the Ab Initio Description of 12C and 16O. Phys. Rev. Lett. 2011, 107, 072501. [Google Scholar] [CrossRef] [PubMed]
  72. Maris, P.; Vary, J.P.; Navrátil, P.; Ormand, W.E.; Nam, H.; Dean, D.J. Origin of the Anomalous Long Lifetime of 14C. Phys. Rev. Lett. 2011, 106, 202502. [Google Scholar] [CrossRef]
  73. Hagen, G.; Hjorth-Jensen, M.; Jansen, G.R.; Machleidt, R.; Papenbrock, T. Continuum Effects and Three-Nucleon Forces in Neutron-Rich Oxygen Isotopes. Phys. Rev. Lett. 2012, 108, 242501. [Google Scholar] [CrossRef]
  74. Hergert, H.; Binder, S.; Calci, A.; Langhammer, J.; Roth, R. Ab Initio Calculations of Even Oxygen Isotopes with Chiral Two-Plus-Three-Nucleon Interactions. Phys. Rev. Lett. 2013, 110, 242501. [Google Scholar] [CrossRef]
  75. Holt, J.D.; Menéndez, J.; Schwenk, A. Three-Body Forces and Proton-Rich Nuclei. Phys. Rev. Lett. 2013, 110, 022502. [Google Scholar] [CrossRef]
  76. Bogner, S.K.; Hergert, H.; Holt, J.D.; Schwenk, A.; Binder, S.; Calci, A.; Langhammer, J.; Roth, R. Nonperturbative Shell-Model Interactions from the In-Medium Similarity Renormalization Group. Phys. Rev. Lett. 2014, 113, 142501. [Google Scholar] [CrossRef] [PubMed]
  77. Fukui, T.; De Angelis, L.; Ma, Y.Z.; Coraggio, L.; Gargano, A.; Itaco, N.; Xu, F.R. Realistic shell-model calculations for p-shell nuclei including contributions of a chiral three-body force. Phys. Rev. C 2018, 98, 044305. [Google Scholar] [CrossRef]
  78. Ma, Y.Z.; Xu, F.R.; Coraggio, L.; Hu, B.S.; Li, J.G.; Fukui, T.; De Angelis, L.; Itaco, N.; Gargano, A. Chiral three-nucleon force and continuum for dripline nuclei and beyond. Phys. Lett. B 2020, 802, 135257. [Google Scholar] [CrossRef]
  79. Hebeler, K.; Bogner, S.K.; Furnstahl, R.J.; Nogga, A.; Schwenk, A. Improved nuclear matter calculations from chiral low-momentum interactions. Phys. Rev. C 2011, 83, 031301. [Google Scholar] [CrossRef]
  80. Simonis, J.; Stroberg, S.R.; Hebeler, K.; Holt, J.D.; Schwenk, A. Saturation with chiral interactions and consequences for finite nuclei. Phys. Rev. C 2017, 96, 014303. [Google Scholar] [CrossRef]
  81. Jiang, W.G.; Ekström, A.; Forssén, C.; Hagen, G.; Jansen, G.R.; Papenbrock, T. Accurate bulk properties of nuclei from A=2 to from potentials with Δ isobars. Phys. Rev. C 2020, 102, 054301. [Google Scholar] [CrossRef]
  82. Somà, V.; Navrátil, P.; Raimondi, F.; Barbieri, C.; Duguet, T. Novel chiral Hamiltonian and observables in light and medium-mass nuclei. Phys. Rev. C 2020, 101, 014318. [Google Scholar] [CrossRef]
  83. 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]
  84. Hebeler, K. Three-nucleon forces: Implementation and applications to atomic nuclei and dense matter. Phys. Rep. 2021, 890, 1–116. [Google Scholar] [CrossRef]
  85. Zhang, S.; Cheng, Z.H.; Li, J.G.; Xu, Z.C.; Xu, F.R. Ab initio Gamow shell model with chiral three-nucleon force for 14O isotones. Chin. Sci. Bull. 2022, 67, 4101–4107. [Google Scholar] [CrossRef]
  86. Zhang, S.; Xu, F.R.; Li, J.G.; Hu, B.S.; Cheng, Z.H.; Michel, N.; Ma, Y.Z.; Yuan, Q.; Zhang, Y.H. Ab initio descriptions of A=16 mirror nuclei with resonance and continuum coupling. Phys. Rev. C 2023, 108, 064316. [Google Scholar] [CrossRef]
  87. Coraggio, L.; De Gregorio, G.; Fukui, T.; Gargano, A.; Ma, Y.Z.; Cheng, Z.H.; Xu, F.R. The role of three-nucleon potentials within the shell model: Past and present. Prog. Part. Nucl. Phys. 2024, 134, 104079. [Google Scholar] [CrossRef]
  88. Shen, S.; Elhatisari, S.; Lee, D.; Meißner, U.G.; Ren, Z. Ab initio study of the beryllium isotopes 7Be to 12Be. arXiv 2024, arXiv:2411.14935. [Google Scholar]
  89. Zhang, S.; Elhatisari, S.; Meißner, U.G.; Shen, S. Lattice simulation of nucleon distribution and shell closure in the proton-rich nucleus 22Si. arXiv 2024, arXiv:2411.17462. [Google Scholar]
  90. Quaglioni, S.; Navrátil, P. Ab Initio Many-Body Calculations of n-3H, n-4He, p-3,4He, and n-10Be Scattering. Phys. Rev. Lett. 2008, 101, 092501. [Google Scholar] [CrossRef]
  91. Quaglioni, S.; Navrátil, P. Ab initio many-body calculations of nucleon-nucleus scattering. Phys. Rev. C 2009, 79, 044606. [Google Scholar] [CrossRef]
  92. Shirokov, A.M.; Mazur, A.I.; Mazur, I.A.; Vary, J.P. Shell model states in the continuum. Phys. Rev. C 2016, 94, 064320. [Google Scholar] [CrossRef]
  93. Mazur, I.A.; Shirokov, A.M.; Mazur, A.I.; Shin, I.J.; Kim, Y.; Maris, P.; Vary, J.P. Description of Continuum Spectrum States of Light Nuclei in the Shell Model. Phys. Part. Nucl. 2019, 50, 537–543. [Google Scholar] [CrossRef]
  94. Baroni, S.; Navrátil, P.; Quaglioni, S. Ab Initio Description of the Exotic Unbound 7He Nucleus. Phys. Rev. Lett. 2013, 110, 022505. [Google Scholar] [CrossRef]
  95. Baroni, S.; Navrátil, P.; Quaglioni, S. Unified ab initio approach to bound and unbound states: No-core shell model with continuum and its application to 7He. Phys. Rev. C 2013, 87, 034326. [Google Scholar] [CrossRef]
  96. Hagen, G.; Dean, D.J.; Hjorth-Jensen, M.; Papenbrock, T. Complex coupled-cluster approach to an ab-initio description of open quantum systems. Phys. Lett. B 2007, 656, 169–173. [Google Scholar] [CrossRef]
  97. Hu, B.S.; Wu, Q.; Sun, Z.H.; Xu, F.R. Ab initio Gamow in-medium similarity renormalization group with resonance and continuum. Phys. Rev. C 2019, 99, 061302. [Google Scholar] [CrossRef]
  98. Papadimitriou, G.; Rotureau, J.; Michel, N.; Płoszajczak, M.; Barrett, B.R. Ab initio no-core Gamow shell model calculations with realistic interactions. Phys. Rev. C 2013, 88, 044318. [Google Scholar] [CrossRef]
  99. Li, J.G.; Michel, N.; Hu, B.S.; Zuo, W.; Xu, F.R. Ab initio no-core Gamow shell-model calculations of multineutron systems. Phys. Rev. C 2019, 100, 054313. [Google Scholar] [CrossRef]
  100. Li, J.G.; Michel, N.; Zuo, W.; Xu, F.R. Resonances of A=4T=1 isospin triplet states within the ab initio no-core Gamow shell model. Phys. Rev. C 2021, 104, 024319. [Google Scholar] [CrossRef]
  101. Hagen, G.; Hjorth-Jensen, M.; Michel, N. Gamow shell model and realistic nucleon-nucleon interactions. Phys. Rev. C 2006, 73, 064307. [Google Scholar] [CrossRef]
  102. Tsukiyama, K.; Hjorth-Jensen, M.; Hagen, G. Gamow shell-model calculations of drip-line oxygen isotopes. Phys. Rev. C 2009, 80, 051301. [Google Scholar] [CrossRef]
  103. Sun, Z.H.; Wu, Q.; Zhao, Z.H.; Hu, B.S.; Dai, S.J.; Xu, F.R. Resonance and continuum Gamow shell model with realistic nuclear forces. Phys. Lett. B 2017, 769, 227–232. [Google Scholar] [CrossRef]
  104. Hu, B.S.; Wu, Q.; Li, J.G.; Ma, Y.Z.; Sun, Z.H.; Michel, N.; Xu, F.R. An ab-initio Gamow shell model approach with a core. Phys. Lett. B 2020, 802, 135206. [Google Scholar] [CrossRef]
  105. Zhang, S.; Ma, Y.Z.; Li, J.G.; Hu, B.S.; Yuan, Q.; Cheng, Z.H.; Xu, F.R. The roles of three-nucleon force and continuum coupling in mirror symmetry breaking of oxygen mass region. Phys. Lett. B 2022, 827, 136958. [Google Scholar] [CrossRef]
  106. Xu, Z.C.; Zhang, S.; Li, J.G.; Jin, S.L.; Yuan, Q.; Cheng, Z.H.; Michel, N.; Xu, F.R. Complex valence-space effective operators for observables: The Gamow-Teller transition. Phys. Rev. C 2023, 108, L031301. [Google Scholar] [CrossRef]
  107. Ellis, P.J.; Osnes, E. An introductory guide to effective operators in nuclei. Rev. Mod. Phys. 1977, 49, 777–832. [Google Scholar] [CrossRef]
  108. Brandow, B.H. Linked-Cluster Expansions for the Nuclear Many-Body Problem. Rev. Mod. Phys. 1967, 39, 771–828. [Google Scholar] [CrossRef]
  109. Tichai, A.; Langhammer, J.; Binder, S.; Roth, R. Hartree–Fock many-body perturbation theory for nuclear ground-states. Phys. Lett. B 2016, 756, 283–288. [Google Scholar] [CrossRef]
  110. Roth, R.; Binder, S.; Vobig, K.; Calci, A.; Langhammer, J.; Navratil, P. Ab Initio Calculations of Medium-Mass Nuclei with Normal-Ordered Chiral NN+3N Interactions. Phys. Rev. Lett. 2012, 109, 052501. [Google Scholar] [CrossRef]
  111. Hu, B.S.; Xu, F.R.; Sun, Z.H.; Vary, J.P.; Li, T. Ab initio nuclear many-body perturbation calculations in the Hartree-Fock basis. Phys. Rev. C 2016, 94, 014303. [Google Scholar] [CrossRef]
  112. Tsunoda, N.; Takayanagi, K.; Hjorth-Jensen, M.; Otsuka, T. Multi-shell effective interactions. Phys. Rev. C 2014, 89, 024313. [Google Scholar] [CrossRef]
  113. Coraggio, L.; Itaco, N. Self-consistent nuclear shell-model calculation starting from a realistic NN potential. Phys. Lett. B 2005, 616, 43–47. [Google Scholar] [CrossRef]
  114. Takayanagi, K. Effective interaction in non-degenerate model space. Nucl. Phys. A 2011, 852, 61–81. [Google Scholar] [CrossRef]
  115. Michel, N.; Aktulga, H.; Jaganathen, Y. Toward scalable many-body calculations for nuclear open quantum systems using the Gamow Shell Model. Comp. Phys. Comm. 2020, 247, 106978. [Google Scholar] [CrossRef]
  116. Ma, Y.Z.; Xu, F.R.; Michel, N.; Zhang, S.; Li, J.G.; Hu, B.S.; Coraggio, L.; Itaco, N.; Gargano, A. Continuum and three-nucleon force in Borromean system: The 17Ne case. Phys. Lett. B 2020, 808, 135673. [Google Scholar] [CrossRef]
  117. Li, J.G.; Hu, B.S.; Wu, Q.; Gao, Y.; Dai, S.J.; Xu, F.R. Neutron-rich calcium isotopes within realistic Gamow shell model calculations with continuum coupling. Phys. Rev. C 2020, 102, 034302. [Google Scholar] [CrossRef]
  118. Geng, Y.F.; Li, J.G.; Ma, Y.Z.; Hu, B.S.; Wu, Q.; Sun, Z.H.; Zhang, S.; Xu, F.R. Excitation spectra of the heaviest carbon isotopes investigated within the CD-Bonn Gamow shell model. Phys. Rev. C 2022, 106, 024304. [Google Scholar] [CrossRef]
  119. Coraggio, L.; Itaco, N. Perturbative Approach to Effective Shell-Model Hamiltonians and Operators. Front. Phys. 2020, 8, 345. [Google Scholar] [CrossRef]
  120. Suzuki, K.; Okamoto, R. Effective Operators in Time-Independent Approach. Prog. Theor. Phys. 1995, 93, 905–917. [Google Scholar] [CrossRef]
  121. Michel, N.; Nazarewicz, W.; Płoszajczak, M. Proton-neutron coupling in the Gamow shell model: The lithium chain. Phys. Rev. C 2004, 70, 064313. [Google Scholar] [CrossRef]
  122. Xu, Z.C.; Hu, R.Z.; Jin, S.L.; Hou, J.H.; Zhang, S.; Xu, F.R. Collectivity of nuclei near the exotic doubly magic Ni78 by ab initio calculations. Phys. Rev. C 2024, 110, 024308. [Google Scholar] [CrossRef]
  123. Wang, S.M.; Nazarewicz, W.; Charity, R.J.; Sobotka, L.G. Nucleon–nucleon correlations in the extreme oxygen isotopes. J. Phys. G: Nucl. Part. Phys. 2022, 49, 10LT02. [Google Scholar] [CrossRef]
  124. Yang, Y.H.; Ma, Y.G.; Wang, S.M.; Zhou, B.; Fang, D.Q. Structure and decay mechanism of the low-lying states in 9Be and 9B. Phys. Rev. C 2023, 108, 044307. [Google Scholar] [CrossRef]
  125. Wang, Y.T.; Fang, D.Q.; Wang, K.; Xu, X.X.; Sun, L.J.; Bao, P.F.; Bai, Z.; Cao, X.G.; Dai, Z.T.; Ding, B.; et al. Observation of β-delayed 2He emission from the proton-rich nucleus 22Al. Phys. Lett. B 2018, 784, 12–15. [Google Scholar] [CrossRef]
  126. Wang, S.M.; Nazarewicz, W.; Charity, R.J.; Sobotka, L.G. Structure and decay of the extremely proton-rich nuclei 11,12O. Phys. Rev. C 2019, 99, 054302. [Google Scholar] [CrossRef]
  127. Zhang, S.; Geng, Y.F.; Xu, F.R. Ab Initio Gamow Shell-Model Calc. Dripline Nuclei. Nucl. Tech. 2023, 46, 080012. [Google Scholar] [CrossRef]
  128. Fortune, H.T. Energy and width of 11O(g.s.). Phys. Rev. C 2019, 99, 051302. [Google Scholar] [CrossRef]
  129. Garrido, E.; Jensen, A.S. Few-body structures in the mirror nuclei 11O and 11Li. Phys. Rev. C 2020, 101, 034003. [Google Scholar] [CrossRef]
  130. Webb, T.B.; Charity, R.J.; Elson, J.M.; Hoff, D.E.M.; Pruitt, C.D.; Sobotka, L.G.; Brown, K.W.; Barney, J.; Cerizza, G.; Estee, J.; et al. Invariant-mass spectrum of 11O. Phys. Rev. C 2020, 101, 044317. [Google Scholar] [CrossRef]
  131. Suzuki, D.; Iwasaki, H.; Beaumel, D.; Assié, M.; Baba, H.; Blumenfeld, Y.; de Oliveira Santos, F.; de Séréville, N.; Drouart, A.; Franchoo, S.; et al. Second 0+ state of unbound 12O: Scaling of mirror asymmetry. Phys. Rev. C 2016, 93, 024316. [Google Scholar] [CrossRef]
  132. Webb, T.B.; Charity, R.J.; Elson, J.M.; Hoff, D.E.M.; Pruitt, C.D.; Sobotka, L.G.; Brown, K.W.; Barney, J.; Cerizza, G.; Estee, J.; et al. Particle decays of levels in 11,12N and 12O investigated with the invariant-mass method. Phys. Rev. C 2019, 100, 024306. [Google Scholar] [CrossRef]
  133. Michel, N.; Li, J.G.; Xu, F.R.; Zuo, W. Proton decays in 16Ne and 18Mg and isospin-symmetry breaking in carbon isotopes and isotones. Phys. Rev. C 2021, 103, 044319. [Google Scholar] [CrossRef]
  134. Zhou, L.; Fang, D.Q.; Wang, S.M.; Hua, H. Structure and 2p decay mechanism of 18Mg. Nucl. Sci. Tech. 2024, 35, 107. [Google Scholar] [CrossRef]
  135. Brown, K.W.; Charity, R.J.; Sobotka, L.G.; Chajecki, Z.; Grigorenko, L.V.; Egorova, I.A.; Parfenova, Y.L.; Zhukov, M.V.; Bedoor, S.; Buhro, W.W.; et al. Observation of long-range three-body Coloumb effects in the decay of 16Ne. Phys. Rev. Lett. 2014, 113, 232501. [Google Scholar] [CrossRef]
  136. Evaluated Nuclear Structure Data File (ENSDF). Available online: https://www.nndc.bnl.gov/ensdf/ (accessed on 30 December 2024).
  137. Bardayan, D.W.; Blackmon, J.C.; Brune, C.R.; Champagne, A.E.; Chen, A.A.; Cox, J.M.; Davinson, T.; Hansper, V.Y.; Hofstee, M.A.; Johnson, B.A.; et al. Observation of the Astrophysically Important 3+ State in N-18e via Elastic Scattering of a Radioactive F-17 Beam from H-1. Phys. Rev. Lett. 1999, 83, 45–48. [Google Scholar] [CrossRef]
  138. Angulo, C.; Tabacaru, G.; Couder, M.; Gaelens, M.; Leleux, P.; Ninane, A.; Verbist, F.; Davinson, T.; Woods, P.J.; Schweitzer, J.S.; et al. Identification of a new low lying state in the proton dripline nucleus Na-19. Phys. Rev. C 2003, 67, 014308. [Google Scholar] [CrossRef]
  139. Entem, D.R.; Machleidt, R. Accurate charge-dependent nucleon-nucleon potential at fourth order of chiral perturbation theory. Phys. Rev. C 2003, 68, 041001. [Google Scholar] [CrossRef]
  140. Brown, B.A.; Richter, W.A. New “USD” Hamiltonians for the sd shell. Phys. Rev. C 2006, 74, 034315. [Google Scholar] [CrossRef]
  141. Kaneko, K.; Sun, Y.; Mizusaki, T.; Tazaki, S.; Ghorui, S. Isospin-symmetry breaking in superallowed Fermi β-decay due to isospin-nonconserving forces. Phys. Lett. B 2017, 773, 521–526. [Google Scholar] [CrossRef]
  142. Michel, N.; Nazarewicz, W.; Płoszajczak, M. Isospin mixing and the continuum coupling in weakly bound nuclei. Phys. Rev. C 2010, 82, 044315. [Google Scholar] [CrossRef]
  143. Magilligan, A.; Brown, B.A. New isospin-breaking “USD” Hamiltonians for the sd shell. Phys. Rev. C 2020, 101, 064312. [Google Scholar] [CrossRef]
  144. Brown, B.; Wildenthal, B. Experimental and theoretical Gamow-Teller beta-decay observables for the sd-shell nuclei. At. Data Nucl. Data Tables 1985, 33, 347–404. [Google Scholar] [CrossRef]
  145. Towner, I. Quenching of spin matrix elements in nuclei. Phys. Rep. 1987, 155, 263–377. [Google Scholar] [CrossRef]
  146. Hoferichter, M.; Menéndez, J.; Schwenk, A. Coherent elastic neutrino-nucleus scattering: EFT analysis and nuclear responses. Phys. Rev. D 2020, 102, 074018. [Google Scholar] [CrossRef]
Figure 1. Energy spectra (with respect to the core) of 11,12O, their isobaric analogs 11Li and 12Be, and neighboring nuclei 10,11N. The decay widths are marked by gray bars. The GCC results in the top (bottom) panels are for the core+valence potentials whose depths were readjusted to fit the spectra of 11N (10N). This figure is taken from [126].
Figure 1. Energy spectra (with respect to the core) of 11,12O, their isobaric analogs 11Li and 12Be, and neighboring nuclei 10,11N. The decay widths are marked by gray bars. The GCC results in the top (bottom) panels are for the core+valence potentials whose depths were readjusted to fit the spectra of 11N (10N). This figure is taken from [126].
Symmetry 17 00169 g001
Figure 2. Two−nucleon density distributions (in fm 2 ) in Jacobi coordinates predicted in GCC for low-lying resonant states in 11O (ac) and 11Li (df). This figure is taken from [5].
Figure 2. Two−nucleon density distributions (in fm 2 ) in Jacobi coordinates predicted in GCC for low-lying resonant states in 11O (ac) and 11Li (df). This figure is taken from [5].
Symmetry 17 00169 g002
Figure 3. Comparison of the excitation energies ( E x , in MeV) and widths (in keV) of mirroring nuclear states of carbon isotones and isotopes. Excitation energies of a nucleus are given with respect to its ground state energy. Widths are represented by green striped squares, and their explicit values are written above. The GSM calculations are compared with available experimental data. Experimental data of 16Ne are taken from Ref. [135], whereas all other data are taken from Ref. [136]. This figure is taken from [133].
Figure 3. Comparison of the excitation energies ( E x , in MeV) and widths (in keV) of mirroring nuclear states of carbon isotones and isotopes. Excitation energies of a nucleus are given with respect to its ground state energy. Widths are represented by green striped squares, and their explicit values are written above. The GSM calculations are compared with available experimental data. Experimental data of 16Ne are taken from Ref. [135], whereas all other data are taken from Ref. [136]. This figure is taken from [133].
Symmetry 17 00169 g003
Figure 4. Level diagrams of 8C and 9N obtained experimentally and calculated with the GSM. Energies are given relative to the 4He threshold. The level diagram of 9He, the mirror partner of 9N, is shown in the inset. The 1 / 2 + antibound state in 9He is shown with a wavy line to indicate its status more as a scattering feature rather than a real state. The proposed 1 / 2 + state in 9N is shown with both straight and wavy lines to indicate the uncertainty with regard to its nature (a resonance or scattering feature), while the GSM interprets it as a broad resonant state. This figure is taken from [18].
Figure 4. Level diagrams of 8C and 9N obtained experimentally and calculated with the GSM. Energies are given relative to the 4He threshold. The level diagram of 9He, the mirror partner of 9N, is shown in the inset. The 1 / 2 + antibound state in 9He is shown with a wavy line to indicate its status more as a scattering feature rather than a real state. The proposed 1 / 2 + state in 9N is shown with both straight and wavy lines to indicate the uncertainty with regard to its nature (a resonance or scattering feature), while the GSM interprets it as a broad resonant state. This figure is taken from [18].
Symmetry 17 00169 g004
Figure 5. Excitation spectra of A = 19 mirror partners, 19O and 19Na. “NN” and “3N” indicate calculations with NN only and 3NF included, respectively. Dashed levels present SM calculations without continuum included, while solid levels give GSM calculations with continuum included. Shading indicates the resonance of the 1 / 2 + TES state with width (in MeV) given below the level. This figure is taken from [105].
Figure 5. Excitation spectra of A = 19 mirror partners, 19O and 19Na. “NN” and “3N” indicate calculations with NN only and 3NF included, respectively. Dashed levels present SM calculations without continuum included, while solid levels give GSM calculations with continuum included. Shading indicates the resonance of the 1 / 2 + TES state with width (in MeV) given below the level. This figure is taken from [105].
Symmetry 17 00169 g005
Figure 6. GSM calculations of spectra with 3NF included for the mirror nuclei 22O and 22Si. The experimental data of 22O are from [136]. This figure is taken from [105].
Figure 6. GSM calculations of spectra with 3NF included for the mirror nuclei 22O and 22Si. The experimental data of 22O are from [136]. This figure is taken from [105].
Symmetry 17 00169 g006
Figure 7. MEDs for the 1 1 + and 1 2 + analog states between mirror nuclei 22Al and 22F, calculated by standard SM with USDC and EM1.8/2.0 (abbreviated by EM) and by GSM with EM1.8/2.0, compared with data [17,136]. This figure is taken from [106].
Figure 7. MEDs for the 1 1 + and 1 2 + analog states between mirror nuclei 22Al and 22F, calculated by standard SM with USDC and EM1.8/2.0 (abbreviated by EM) and by GSM with EM1.8/2.0, compared with data [17,136]. This figure is taken from [106].
Symmetry 17 00169 g007
Figure 8. A 20 GT transition matrix elements calculated by GSM and SM with EM1.8/2.0 compared with data [17,136]. This figure is taken from [106].
Figure 8. A 20 GT transition matrix elements calculated by GSM and SM with EM1.8/2.0 compared with data [17,136]. This figure is taken from [106].
Symmetry 17 00169 g008
Table 1. GT transition matrix elements | M GT | calculated by standard SM with USDC and EM1.8/2.0 and by GSM with EM1.8/2.0 for mirror nuclei 22Si and 22O decaying into the 1 1 + and 1 2 + states of their mirror daughters 22Al and 22F, compared with data and calculations given in Ref. [17]. This table is taken from [106].
Table 1. GT transition matrix elements | M GT | calculated by standard SM with USDC and EM1.8/2.0 and by GSM with EM1.8/2.0 for mirror nuclei 22Si and 22O decaying into the 1 1 + and 1 2 + states of their mirror daughters 22Al and 22F, compared with data and calculations given in Ref. [17]. This table is taken from [106].
SMGSMRef. [17]
USDCEMEMExpt.Cal.
22Si→22Al 1 1 + 0.2360.3430.2570.176 (16)0.242
1 2 + 0.7211.0421.0120.750 (41)0.863
22O→22F 1 1 + 0.1980.5690.4970.310 (32)0.428
1 2 + 0.7191.0921.0680.775 (77)0.848
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

Zhang, S.; Xu, Z.; Wang, S. Continuum Effect on Mirror Symmetry Breaking Within the Gamow Frameworks. Symmetry 2025, 17, 169. https://doi.org/10.3390/sym17020169

AMA Style

Zhang S, Xu Z, Wang S. Continuum Effect on Mirror Symmetry Breaking Within the Gamow Frameworks. Symmetry. 2025; 17(2):169. https://doi.org/10.3390/sym17020169

Chicago/Turabian Style

Zhang, Shuang, Zhicheng Xu, and Simin Wang. 2025. "Continuum Effect on Mirror Symmetry Breaking Within the Gamow Frameworks" Symmetry 17, no. 2: 169. https://doi.org/10.3390/sym17020169

APA Style

Zhang, S., Xu, Z., & Wang, S. (2025). Continuum Effect on Mirror Symmetry Breaking Within the Gamow Frameworks. Symmetry, 17(2), 169. https://doi.org/10.3390/sym17020169

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