logo

Progress in ab initio in-medium similarity renormalization group and coupled-channel method with coupling to the continuum

Special Issue: Dedicated to Professor Wenqing Shen in Honour of his 80th Birthday

Progress in ab initio in-medium similarity renormalization group and coupled-channel method with coupling to the continuum

Xin-Yu Xu
Si-Qin Fan
Qi Yuan
Bai-Shan Hu
Jian-Guo Li
Si-Min Wang
Fu-Rong Xu
Nuclear Science and TechniquesVol.35, No.12Article number 215Published in print Dec 2024Available online 26 Nov 2024
18705

Over the last decade, nuclear theory has made dramatic progress in few-body and ab initio many-body calculations. These great advances stem from chiral effective field theory (χEFT), which provides an efficient expansion and consistent treatment of nuclear forces as inputs of modern many-body calculations, among which the in-medium similarity renormalization group (IMSRG) and its variants play a vital role. On the other hand, significant efforts have been made to provide a unified description of the structure, decay, and reactions of the nuclei as open quantum systems. While a fully comprehensive and microscopic model has yet to be realized, substantial progress over recent decades has enhanced our understanding of open quantum systems around the dripline, which are often characterized by exotic structures and decay modes. To study these interesting phenomena, Gamow coupled-channel (GCC) method, in which the open quantum nature of few-body valence nucleons coupled to a deformed core, has been developed. This review focuses on the developments of the advanced IMSRG and GCC, and their applications to nuclear structure and reactions.

Ab initio calculationsChiral effective field theoryIn-medium similarity renormalization groupGamow-coupled channelResonance and continuumOpen quantum systems
1

Introduction

In the past decades, great progress in nuclear forces [1, 2], and ab initio many-body theories [3-10] has been made. Using low-energy expansion with nucleons and pions as explicit degrees of freedom, the chiral effective field theory (χEFT) [1, 2] with Weinberg’s power counting (WPC) [11-13] provides a powerful framework in which two- and many-nucleon interactions, and electroweak currents can be naturally derived with the uncertainties associated with each expansion order. The three-nucleon force (3NF) has been shown to be crucial in the quantitative predictions of nuclear structure [14-30].

However, it still challenges current calculations to extend the ab initio frontier to heavier nuclei. In ab initio calculations, one needs to handle the coupling between low and high momenta of nuclear forces. In recent years, new approaches to nuclear forces have been developed based on the ideas of the renormalization group (RG), whereby high-momentum degrees of freedom are decoupled by lowering the resolution (or cutoff) scale in nuclear forces to typical nuclear structure momentum scales, which greatly accelerates the convergence of the nuclear structure calculations [31-33]. The similarity renormalization group (SRG) [34, 35] provides a powerful method to decouple the high-momentum degrees of freedom, using continuous unitary transformations that suppress off-diagonal matrix elements and drive the Hamiltonian towards a band-diagonal form. The SRG softened nuclear forces can accelerate many-body calculations without compromising the nature of realistic nuclear forces or the accuracy of calculations.

The ab initio SRG method has been further developed to treat the nuclear many-body problems, which are in-medium similarity renormalization group (IMSRG) [36-38] for the ground states of closed-shell nuclei and valence-space IMSRG (VS-IMSRG) [39, 21, 28] for open-shell nuclei. The IMSRG employs a continuous unitary transformation of the many-body Hamiltonian to decouple the ground state from all excitations, thereby solving the many-body problems [39, 21, 37, 40, 28, 10]. Other advanced ab initio many-body methods include coupled-cluster (CC) theory [41-43], many-body perturbation theory (MBPT) [44, 6, 45], and self-consistent Green’s function (SCGF) [46, 27]. Current ab initio many-body approaches have become possible to accurately describe more than one hundred fully interacting nucleons in a controlled way [7, 47].

The traditional spherical symmetry-conserving single-reference scheme of IMSRG is applicable only to closed-shell nuclei. To calculate open-shell nuclei, symmetry-breaking schemes have been developed, which include single- and multi-reference approaches [48-52]. To capture the strong collective correlations, Yuan et al. developed an ab initio deformed single-reference IMSRG approach for open-shell nuclei in the m-scheme Hartree-Fock (HF) basis, referred to as D-IMSRG [53]. Using the m-scheme, a single HF reference state can be constructed for any even-even nuclei. The deformed reference state efficiently includes the important configurations of the deformed nucleus and captures more correlations through symmetry restoration, which would be many-particle-many-hole excitations in the spherical scheme. The calculations under the axially deformed HF basis break the SU(2) rotational symmetry associated with angular momentum conservation. The broken rotation symmetry can be restored by angular momentum projection.

Next-generation rare isotope beam (RIB) facilities have the ability to produce most of the rare isotopes located at the edge of the nuclear landscape, thereby shedding light on the origin of elements, the fundamental problems of nuclear structure, and nuclear forces. However, providing theoretical descriptions of proton- or neutron-rich nuclei in these regions is challenging due to the complexity of theoretical methods and computational demands. As nuclei approach the dripline, the effects of single-particle long-distance asymptotic behavior and coupling to the continuum become crucial for understanding the open quantum systems [54]. The complex-energy Berggren basis provides an efficient framework to treat bound, resonant, and scattering continuum states on an equal footing [55, 56]. To include the coupling to the continuum, Hu et al. developed a Gamow IMSRG (G-IMSRG) [57] in the complex-energy Berggren basis. The advanced G-IMSRG is capable of describing the resonance and non-resonant continuum properties of weakly bound and unbound nuclei. The known heaviest Borromean halo 22C is a challenging nucleus for many theoretical calculations [58-60]. The halo structure of 22C can be clearly visualized by calculating the density distribution in which the continuum s channel plays a crucial role, and the low-lying resonant excited states in 22C are predicted via the G-IMSRG [57].

Significant efforts have been made to develop theoretical frameworks in an alternative direction to provide a comprehensive description of dripline systems, which often exhibit exotic structures and decay modes. These approaches aim to unify the treatment of structure, decay, and reactions within a single framework. Although a fully comprehensive and microscopic model achieving this goal does not yet exist, substantial advances have been made over the past few decades [61-63]. Notably, in [64, 65], a method was demonstrated for integrating structural and reaction aspects starting from an ab initio framework. I. In this framework, each component of the three-body system is calculated using the no-core shell model (NCSM) in Jacobi coordinates. The inter-cluster motion is described using the resonating group method (RGM), which has been widely applied in nuclear reactions. Recent developments have also incorporated continuum effects, exemplified by the Gamow shell model coupled channel (GSM-CC) [54, 66] and Gamow coupled-channel (GCC) method [67, 68]. The former focuses on configuration mixing with self-consistent continuum effects [54, 69], whereas the latter emphasizes the open quantum nature of few-body valence nucleons coupled to a deformed core [68, 70, 71]. This review primarily focuses on the recent advancements in the GCC method and its applications to exotic decays in the dripline region.

This review is organized as follows. The basic IMSRG approach and its extensions are expounded in Sec. 2.1-2.4. The theory of GCC method is formulated in Sec. 2.5. Sec. 3 describes the main results of the corresponding IMSRG and GCC computations of atomic nuclei. Finally, a summary is presented in Sec. 4.

2

Outline of developed methods

In this section, the basic formulae for the developed IMSRG and GCC approaches are presented. Sec. 2.1 is dedicated to the main ideas and previous developments of IMSRG itself. The symmetry-breaking m-scheme D-IMSRG is introduced in Sec. 2.2 for open-shell nuclei applications. Another approach to treat the open-shell nuclei while preserving the spherical symmetry is the VS-IMSRG, which combines the shell model and IMSRG, as presented in Sec. 2.3. The G-IMSRG with the Berggren basis for the description of weakly bound and unbound nuclei is formulated in Sec. 2.4. Finally, the reaction-related GCC method and its extensions to deformed systems and time-dependent approaches are briefly introduced in Sec. 2.5 and 2.6.

2.1
The in-medium similarity renormalization group

The SRG is to evolve the Hamiltonian H(s) to be band-diagonal by using the continuous unitary transformation as [35, 34] H(s)=U(s)HU(s)Hd(s)+Hod(s)Hd(), (1) where s denotes the so-called flow parameter, and Hd(s) and Hod(s) are appropriately defined as the “diagonal” and “off-diagonal” parts of the Hamiltonian, respectively. Although the evolution should continue up to s→∞, a finite number of evolution steps is usually sufficient to make H(s) approach the band-diagonal form of Hd(∞).

Equation (1) expresses a general ideal. In practice, taking the derivative of Eq. (1), a flow equation is defined to evolve the Hamiltonian H(0), ddsH(s)=[η(s),H(s)], (2) where the anti-Hermitian generator η(s) is related to the unitary transformation U(s) by η(s)=dU(s)dsU(s)=η(s). (3) A commonly used generator is defined as η(s)=[Hd(s),H(s)]=[Hd(s),Hod(s)], (4) which guarantees that the off-diagonal coupling of Hod is driven exponentially to zero with increasing in the value of the flow parameter s [35]. In practice, the demand for strict diagonality is usually relaxed to band diagonality of the Hamiltonian matrix in a chosen basis, such as in relative momentum or harmonic oscillator (HO) spaces. In nuclear physics, the SRG is used to decouple the momentum or energy scales in free space to construct “soft” NN and 3N interactions, thereby rendering the nuclear Hamiltonian more suitable for ab initio many-body calculations [31, 33, 72-74].

The SRG is used to soften the nuclear force which has a hard core in free space. This renormalization can significantly accelerate ab initio calculations of nuclei. Another development of the SRG theory is the in-medium SRG (IMSRG) [36-38] which evolves the many-body Hamiltonian to block diagonal form. The decoupling between the lowest-energy ground state and excited states of the Hamiltonian directly provides the energy of the ground state of the nucleus. A distinct advantage of IMSRG, compared to the SRG free-space evolution, is its ability to approximately evolve 3, …, A-body operators using only two-body machinery. This simplification is primarily achieved through the use of normal ordering with respect to a reference state |Φ, usually the Hartree-Fock (HF) state.

The intrinsic Hamiltonian of A-body nuclear system is expressed as H=i=1A(11A)pi22m+i<jA(vijNNpipjmA)+i<j<kAvijk3N, (5) where pi is the nucleon momentum in laboratory coordinates, and m is the nucleon mass, with vNN and v3N denoting the NN and 3N interactions, respectively. In order to generate the reference state in the IMSRG calculation, the HF equation for the intrinsic Hamiltonian Eq. (5) is first solved. The Wick’s theorem is applied to normal order all operators starting from a general second-quantized Hamiltonian with two- and three-body interactions, with respect to the HF ground state. H=E0+ijfij:aiaj:+12!2ijklΓijkl:aiajalak:       +13!2ijklmnWijklmn:aiajakanamal:, (6) where E0, f, Γ, and W correspond to the normal-ordered zero-, one-, two-, and three-body terms, respectively, given by E0=iTiini+12ij(Tijij+vijijNN)ninj+16ijkvijkijk3Nninjnk, (7) fij=Tij+k(Tikjk+vikjkNN)nk+12klvikljkl3Nnknl, (8) Γijkl=Tijkl+vijklNN+14mvijmklm3Nnm, (9) Wijklmn=vijklmn3N, (10) where ni=θ(εFεi) represents the occupation numbers in the reference state |Φ, with εF denoting the Fermi energy of the reference state and T representing the kinetic part of the Hamiltonian. From Eqs. (7)-(10), it is evident that all the zero-, one- and two-body parts of the Hamiltonian contain the in-medium effects from the free-space 3N interactions.

The exact treatment of the 3NF is computationally expensive. Therefore, the residual 3NF is usually neglected, which provides a reasonably good approximation in nuclear structure calculations. The omission of the residual normal-ordered three-body component of the Hamiltonian has been shown to result in only 1%-2% discrepancy in ground-state and excited-states energies for light and medium-mass nuclei [16, 75]. The approximation of normal-ordered two-body (NO2B) for the Hamiltonian has been proved to be useful and beneficial in practical calculations, offering an efficient method to include 3NF effects in nuclear many-body calculations, thereby avoiding the computational burden of directly dealing with three-body operators.

Similar to the evolution of the Hamiltonian, the operators of other observables can also be evolved using the flow equation ddsO(s)=[η(s),O(s)]. (11) The Magnus expansion was usually used in matrix differential equations [76], and was applied to reformulate the IMSRG [77] for more efficient calculations. In the Magnus approach, the IMSRG transformation can be written as an exponential expression [76], U(s)eΩ(s). (12) The Magnus evolution operator Ω(s) works for both the Hamiltonian and other observable operators, which allows the derivation of the flow equation for the anti-Hermitian Magnus operator Ω(s), dΩds=k=0Bkk!adΩk(η), (13) where Bk denote the Bernoulli numbers and adΩ0(η)=η (14) adΩk(η)=[Ω,adΩk1(η)]. (15) In practical calculations, η and Ω are truncated along with their commutators at the two-body level, called the Magnus(2) approximation. The series of nested commutators generated by adΩk are recursively evaluated until a satisfactory convergence of the right-hand side of Eq. (13) is reached [77]. At each integration step, U(s) is used to construct the Hamiltonian H(s) via the Baker-Campbell-Hausdorff (BCH) formula H(s)eΩ(s)H(0)eΩ(s)=k=01k!adΩ(s)k(H(0)). (16) The Magnus formulation offers a significant advantage, as it enables the evaluation of arbitrary observables by utilizing the final Magnus operator Ω(), O()eΩ()O(0)eΩ(). (17) The computational effort for solving the IMSRG(2) flow equations is primarily dictated by the two-body flow equation, which exhibits polynomial complexity of O(N6) based on the single-particle size N.

2.2
The deformed IMSRG

The use of deformations as degrees of freedom in nuclear many-body problems can make the calculations more efficient [78, 79]. The standard IMSRG conserves spherical symmetry with a single reference, which works for closed-shell nuclei. To calculate open-shell nuclei, symmetry-breaking schemes have been developed, including both single- and multi-reference approaches. The single-reference Hartree-Fock-Bogoliubov (HFB) IMSRG, which selects a single HFB state as the reference state, has been proposed [48]. The HFB quasiparticle state breaks the particle number conservation, necessitating that particle number projection be performed. To choose a reference state closer to the true solution, the multi-reference IMSRG with particle-number-projected spherical HFB [49, 50] has been suggested. Calculations based on the Bogoliubov quasiparticle states significantly complicate the formalism and increase computational costs. Using the m scheme, a single HF reference state can be constructed for any even-even nuclei, with the particle number conserved but rotational symmetry broken. This deformed reference state may better reflect the intrinsic structure of a deformed nucleus and capture more correlations through symmetry restoration, which would otherwise be many-particle-many-hole excitations in the spherical scheme. The expected symmetry preconsiderations, e.g., as in the symmetry-adapted approach [80, 81], provide an efficient way to capture the expected features of nuclear states of interest while simultaneously reducing the computational cost.

As indicated in Sec. 2.1, the standard IMSRG is limited to extracting the ground-state energy of a closed-shell nucleus. An extension of the IMSRG to the deformed scheme would be useful for the description of open-shell nuclei. Therefore, we developed the D-IMSRG method [53] within the deformed HF basis, i.e., the m-scheme HF basis. First, the axially deformed HF equation of the even-even nucleus is solved within the spherical HO basis. Under the j-scheme, the initial Hamiltonian is typically expressed in the spherical HO basis. Using the Wigner-Eckart theorem, the matrix elements of operators, including the Hamiltonian in the j-scheme, can be converted to the matrix elements in the m-scheme, α'j'm'|Okq|αjm=jmkq|j'm'×12j'+1α'j'Okαj, (18) where |αjm are quantum states with good angular momentum j and projection m. The quantity α contains all other quantum numbers needed to completely specify the quantum state. Okq is a spherical tensor of rank k; jmkq|j'm' is the Clebsch-Gordan coefficient; and α'j'Okαj denotes the reduced matrix elements. The one-, two-, and three-body matrix elements can thus be converted from j-scheme to m-scheme through Eq. (18). The m-scheme HF single-particle levels obtained are twofold degenerate with respect to the angular momentum projection quantum number m of the orbital (i.e., the energies are the same with respect to ±m). Filling the deformed HF single-particle levels up to the Fermi surfaces of neutrons and protons in ±m pairing from low to high |m|, keeps the axial, parity, and time-reversal symmetries of the even-even ground state, thereby creating an oblate or prolate deformed HF reference state [82]. Subsequently, the intrinsic A-body Hamiltonian in Eq. (5) is normal ordered with respect to the deformed A-dependent reference state |Φ (i.e., the m-scheme HF ground state of the target nucleus). The off-diagonal parts of the Hamiltonian are consistent with the standard IMSRG, and the flow equations are evolved using the Magnus expansion Eqs. (16)-(17).

Subsequently, the ground-state energy and other observables can be calculated using the D-IMSRG ground-state wave function |Ψ=eΩ|Φ (here |Φ is the deformed HF reference state of the nucleus) expressed as E=Ψ|H|Ψ=Φ|eΩHeΩ|Φ=Φ|H˜|Φ, (19) O=Ψ|O|Ψ=Φ|eΩOeΩ|Φ=Φ|O˜|Φ. (20) In the D-IMSRG, the reference state is just the ground state of the deformed even-even nucleus. However, performing exact symmetry restoration of the D-IMSRG wave function is mathematically cumbersome and computationally expensive due to the exponential increase of configurations in projecting the wave function |Ψ=eΩ|Φ.

Therefore, an HF projection correction is introduced as a first approximation, to account for the angular momentum projection effect. The projection correction to the ground-state energy is estimated by ΔEproj=Φ|HP|ΦΦ|P|ΦΦ|H|ΦΦ|Φ, (21) where PMMJ=2J+18π2dωDMMJ*(ω)R(ω) is the angular momentum projection operator. This provides a D-IMSRG ground-state energy given by E+ΔEproj with the projection correction estimated by the HF wave function (here E is obtained by Eq. (19), that is, the ground-state energy without the projection).

A deformed coupled-cluster calculation has estimated that the angular-momentum projection of the HF state reduced the HF energy by approximately 3-5 MeV in the sd shell [82], corresponding to the static correlation, which is not size extensive. Since modern ab initio calculations already include some of the correlations associated with the projection, the energy correction obtained by projecting the ab initio wave function would be slightly smaller than the HF projection correction [82, 83].

In the spherical j-scheme, single-particle levels within the same j shell are degenerate. However, this degeneracy is broken with the onset of deformation although a twofold degeneracy with respect to ±m persists in axially symmetric shapes, significantly increasing the model-space dimension. The dimension of D-IMSRG calculation depends on the nucleon number A and basis-space size Nshell (the number of spherical HO major shells considered in solving the deformed HF). For 40Mg, the number of D-IMSRG Hamiltonian matrix elements already exceeds 109 at Nshell=10. Nonetheless, such a large model space may still not be sufficient to make the calculation converged. To estimate the converged ground-state energy, a simple exponential fitting method was applied with respect to Nshell to extrapolate the D-IMSRG result to an infinite basis space, similar to the ones used in NCSM [16, 84-87] and multi-reference IMSRG [19] calculations, E(Nshell)=b0+b1exp(b2Nshell), (22) where b0, b1, and b2 are the fitting parameters. The value of b0E(Nshell) provides the estimate for the fully converged energy.

2.3
The valence-space IMSRG

The spherical j-scheme IMSRG can only treat closed-shell nuclei. The m-scheme D-IMSRG was designed to calculate open-shell nuclei. Unfortunately, the D-IMSRG does not conserve the angular momentum, and the exact angular-momentum projection D-IMSRG has not been available.

The nuclear shell model (SM) has served as one of the most powerful theoretical and computational tools for nuclear structure calculations [38, 88-91]. In the SM, valence nucleons move in the mean field generated by the inert core and interact via residual effective interactions. While the SM has been used predominantly in a phenomenological context [92, 91], there have been efforts dating back to decades ago to derive shell-model parameters based on a realistic interaction between nucleons [93, 94], the ab initio effective shell-model interactions. For open-shell systems, in addition to solving the full A-body problem, such as the D-IMSRG mentioned in Sec. 2.2, it is beneficial to follow the shell-model paradigm to construct and diagonalize the effective Hamiltonian in which the active degrees of freedom are Av valence nucleons confined to a few orbitals near the Fermi level. As for the IMSRG, a valence-space effective interaction can be derived using the spherical symmetry-conserving single-reference IMSRG at a shell closure to perform ab initio shell-model calculations for open-shell nuclei. This method, which combines the IMSRG and SM, is referred to as the VS-IMSRG [39].

The utility of the IMSRG lies in its flexibility to customize the definition of Hod to address specific problems. For the ground state of closed-shell nuclei, all terms that couple the reference state |Φ to the rest of the Hilbert space can be eliminated, as in the standard IMSRG. For open-shell nuclei, |Φv is decoupled from states containing non-valence states. This can be achieved by defining the Hod using the following matrix elements, Hod=fph,fpp',fhh',Γpp'hh',Γpp'vh,Γpqvv'+H.c., (23) where p=v,q. These off-diagonal parts of the generators evolve the Hamiltonian to diagonal Hd form, where states outside the valence space are decoupled using the flow equation, as illustrated in Fig. 1, nonperturbatively satisfying the decoupling equation: PHd()Q=QHd()P=0, (24) with P=v|ΦvΦv| and Q=1P. After the evolution is complete, the effective shell-model Hamiltonian is obtained. The last step is to use the SM code, such as KSHELL [95], to diagonalize the effective Hamiltonian and express it as a reduced eigenvalue problem in the valence-particle space.

Fig. 1
(Color online) Schematic illustration of the VS-IMSRG decoupling from the initial Hamiltonian H(0) to obtain the final Hamiltonian H() for the two valence nucleons
pic

The current VS-IMSRG is at two-body approximation without explicitly considering 3NF or three-body correlations. To reduce the residual 3NF effect, a fractional filling of open-shell orbitals in an open-shell nucleus, named ensemble normal ordering (ENO), has been suggested [40]. Using the ENO approximation of the VS-IMSRG, nucleus-dependent valence-space effective Hamiltonian and effective operators of other observables can be obtained.

2.4
The Gamow IMSRG with coupling to continuum

Weakly bound and unbound nuclei belong to the category of open quantum systems, where coupling to the particle continuum profoundly affects the system behavior [96, 97]. Many novel phenomena, including halos [98, 99], genuine intrinsic resonances [100, 101], and new collective modes [102-104], have been observed or predicted in exotic nuclei. However, the majority of IMSRG calculations are performed within the HO or real-energy HF basis. Here the real-energy HF means that the HF approach is performed under the HO basis, which is bound and localized and hence isolated from the environment of unbound scattering states because of the Gaussian falloff of the HO functions. Similarly, the real-energy HF basis cannot include the continuum effect in IMSRG calculations.

The complex-energy Berggren basis offers an elegant framework for treating bound, resonant, and scattering continuum states on an equal footing [55, 56]. This basis is a generalization of the standard completeness relation from the real-energy axis to the complex-energy plane. Completeness encompasses a finite set of bound and resonance states together with a complex-energy scattering continuum: nun(En,r)un(En,r')+L dEu(E,r)u(E,r')=δ(rr'), (25) where un(En,r)Ol(knr)eiknr, (26) Here, kn=iκn(κn>0) for bound states and kn=γniκn(κn,γn>0) for decaying resonances located in the fourth quadrant of the complex-momentum (complex-k) plane.

In practical applications, it is more convenient to express Eq. (25) in momentum space, n(b,d)|unun|+L+|u(k)u(k)|dk=1. (27)

As depicted in Fig. 2, bound and resonant states appear as poles of the scattering matrix within the complex-k plane, and the scattering continuum is represented by the blue contour. Within the Berggren basis, the GSM [23, 100, 101, 105-115] and complex coupled cluster [116, 18] methods have been well developed and widely applied to the calculations of weakly bound and unbound nuclei.

Fig. 2
(Color online) Location of one-body states in the complex-k plane. The Berggren completeness relation in Eq. (25) involves the bound states (brown-filled circles) lying on the imaginary k-axis, scattering states lying on the contour (solid blue line), and decaying resonant states (blue-filled circles) in the fourth quarter of the complex-k plane lying between the real axis and scattering contour. The capturing states (purple-filled circles) and antibound states (cyan-filled circles) are not included in the present completeness relation
pic

For calculations within the Gamow-Berggren framework, selecting an appropriate one-body potential is essential for generating resonance and the continuum Berggren basis, frequently using the phenomenological Woods-Saxon potential [100, 101, 105, 117, 111]. In our approach, we used the Gamow Hartree-Fock (GHF) method with chiral potentials to produce an ab initio Berggren single-particle basis, which is convenient for computing the Berggren basis using an analytical continuation of Schrödinger’s equation in complex-k space. The complex-k single-particle GHF equation is formulated as follows: 2k22μψi(k)+L+dk'k'2U(ljk'k)ψi(k')=eiψi(k), (28) where μ=m/(11A), and k(k) is defined on the scattering contour. U(ljk'k) is the GHF single-particle potential, U(ljk'k)=k|U|k'=αβk'|αα|U|ββ|k, (29) where l, j are the orbital and total angular momenta of a single-particle orbital, respectively. Greek letters denote HO states, indicating that 〈β/k〉 is the HO basis wave functions |β expressed in the complex-k plane β|k=(i)2n+lel/2b2k2(bk)l×2n!b3Γ(n+l+3/2)Lnl+1/2(b2k2). (30) Here Lnl+1/2 denotes the generalized Laguerre polynomial; b=/mω; and ω is the frequency of the oscillator basis. Since Lnl+1/2 is analytic, it can be extended to the complex momentum space, expressing its real part as Re[Ln(x+iy)]=j=0|n2|(1)jy2j(2j)!Ln22j(x), (31) and its imaginary part as Im[Ln(x+iy)]=j=0|n12|(1)j1y2j+1(2j+1)!Ln2j12j+1(x), (32) where x and y represent the real and imaginary parts of b2k2, respectively; α|U|β is the HF single-particle potential which can be obtained by solving the real-energy HF equation in the HO basis α|U|β=i=1Aγδαγ|vNN|βδDγi*Dδi+12i,j=1Aγδϵζαγϵ|v3N|βδζDγi*DδiDϵj*Dζj, (33) where D is the coefficient of the HF single-particle state. In numerical calculations, the GHF equation is solved using the Gauss-Legendre quadrature scheme [118, 107, 112] with discrete points on the contour L+, 2kα22μψi(kα)+βωβkβ2kα|U|kβψi(kβ)=eiψi(kα). (34) Here () are the discrete momentum points, and ωα (ωβ) are the corresponding Gauss-Legendre quadrature weights. We define ψi'(kα)=ψi(kα)kαωα. (35) Then, Eq. (34) can be written as βhαβψi'(kβ)=eiψi'(kα), (36) with hαβ=22μkα2δαβ+ωαωβkα|U|kβ. (37) The bound, resonant, and continuum GHF basis can be obtained by diagonalizing the complex-energy in Eq. (36).

Within the GHF basis, the G-IMSRG calculations can be conducted. Notably, the Hamiltonian in real-energy space is Hermitian, H=H. Therefore, in practice, the similarity transformation U(s) is a unitary transformation, satisfying U(s)U(s)=U(s)U1(s)=1, and η(s)=dU(s)dsU(s)=η(s) is the anti-Hermitian generator. However, in the G-IMSRG framework, the Hamiltonian is complex symmetric, H=HT (here T indicates the transpose). Therefore, employing a continuous orthogonal transformation, U(s)UT(s)=U(s)U1(s)=1, and the Hamiltonian H(s) can be transformed into a band or block diagonal form, H(s)=U(s)H(0)UT(s). (38) Correspondingly, the generator η(s) becomes η(s)=dU(s)dsUT(s)=ηT(s). (39) The G-IMSRG method can directly compute the ground state of a closed-shell nucleus by decoupling the Hamiltonian from the excitations above the closed-shell Fermi surface. To handle open-shell nuclei or excited states, we employed G-IMSRG using the equation-of-motion (EOM) approach [119]. This approach offers a useful alternative to the shell model strategy for calculating excited states, especially when extended valence spaces lead to prohibitively large shell-model basis dimensions. Within the EOM framework, the Schrödinger equation is rewritten using ladder operators, which create excited eigenstates from the exact ground state, H|Ψn=En|ΨHXn|Ψ0=EnXn|Ψ0, (40) where Xn is given by the dyadic product |ΨnΨ0|. Further rewriting Eq. (40) as EOM [H,Xn]|Ψ0=(EnE0)Xn|Ψ0ωnXn|Ψ0. (41) The amplitudes of Xn are determined by solving a generalized eigenvalue problem [120].

Coupling EOM methods with G-IMSRG is natural, as the reference state |Φ0 corresponds to the ground state of H¯U()HUT(). Multiplying Eq. (41) by U() and recalling that U()|Ψ0=|Φ0, (42) we obtain the similarity transformed EOM [H¯,X¯n]|Φ0=ωnX¯n|Φ0, (43) where X¯nU()XnUT(). The solutions X¯n can then be used to obtain the eigenstates of the unevolved Hamiltonian via |Ψn=UT()X¯n|Φ0. (44) Currently, in our applications, we include up to 2p2h excitations in the ladder operator [120] X¯n=phX¯ph(n)apah+14pp'hh'X¯pp'hh'(n)apap'ah'ah. (45) In principle, the EOM ladder operator can include any excitation rank up to ApAh, which would constitute an exact diagonalization of H¯ and can be computationally expensive. In practical applications, the EOM-G-IMSRG method is commonly employed in an approximative systematically improbable form, referred to as EOM-G-IMSRG(m,n), where m and n denote the truncation level in EOM and G-IMSRG, respectively. The calculations in the present work were performed using the EOM-G-IMSRG(2,2) approximation.

2.5
The Gamow coupled-channel method

To provide a thorough description of open quantum systems, the GCC method [67, 68, 70] has been advanced as an alternative approach, focusing on the few-body decay processes influenced by continuum and structural factors [121, 122]. The methodology involves constructing a robust three-body framework, utilizing the Berggren basis [55]. As elucidated in Sec. 2.4, the Berggren basis is specifically designed to incorporate continuum effects, thereby facilitating the analysis of weakly bound and unbound nuclear systems.

2.5.1
Spherical system with an inert core

In the context of the three-body GCC model, the nucleus comprises a core and two valence nucleons or clusters. The GCC Hamiltonian is formulated as follows: H=i=13p¯^i22mi+i>j=13Vij(rij)T^c.m., (46) where Vij denotes the interaction between clusters i and j, and T^c.m. represents the kinetic energy of the center-of-mass. Each i-th cluster (i=c,n1,n2) is characterized by its position vector ri and linear momentum ki. To accurately describe three-body asymptotics and eliminate the spurious center-of-mass motion, it is advantageous to utilize the relative (Jacobi) coordinates: x=μx(ri1ri2),y=μy(Ai1ri1+Ai2ri2Ai1+Ai2ri3), (47) where i1=n1, i2=n2 and i3=c for T-coordinates, whereas i1=n2, i2=c and i3=n1 for Y-coordinates, as depicted in Fig. 3. Here, Ai represents the mass number of the i-th cluster; μx=Ai1Ai2Ai1+Ai2 and μy=(Ai1+Ai2)Ai3Ai1+Ai2+Ai3 are the reduced masses associated with x and y, respectively. For analytical convenience, the hyperradius ρ=x2+y2, which remains invariant across different Jacobi coordinate transformations, is introduced.

Fig. 3
(Color online) Illustration of the coordinate and momentum configurations in a core + nucleon + nucleon system: (a) Jacobi T (solid lines) and Y (dashed lines) coordinates, where the former is used to describe the interactions between the nucleons (n1 and n2) and the latter for interactions involving the core (c). (b) Corresponding momentum scheme within the c.m. frame. Here, A denotes the mass number; μij represents the reduced mass between clusters i and j; and k1, k2, and kc indicate the momenta of nucleons n1, n2, and core c, respectively. The figure is taken from [70]
pic

Experimental measurements in the momentum space necessitate the definition of relative momenta as follows: kx=μx(ki1Ai1ki2Ai2),ky=μy(ki1+ki2Ai1+Ai2ki3Ai3). (48) In the absence of c.m. motion, it is evident that iki=0, and notably, ky is oriented in the opposite direction to θi3. The angles θk and θk represent the opening angles of the vectors (kx, ky) in T- and Y-Jacobi coordinates, respectively (refer to Fig. 3). For example, in the scenario of two-nucleon decay, the kinetic energy associated with the relative motion of the emitted nucleons is expressed as Epp/nn=2kx22μx, with Ecore-p/n pertaining to the kinetic energy of the core-nucleon pair. The distribution types T (θk, Epp/nn) and Y (θk, Ecore-p/n) elucidate the nucleon-nucleon correlations and provide insights into the structural dynamics of the progenitor nucleus. The total momentum k is defined as kx2μx+ky2μy, which asymptotically approaches 2mQ2p/2n as time progresses, where Q2p/2n is the two-nucleon decay energy derived from the binding energy difference between parent and daughter nuclei.

The presence of Pauli forbidden states in three-body models represents a challenge arising from the lack of antisymmetrization between core and valence particles. To address this issue, the orthogonal projection method [123-125], which entails the inclusion of a Pauli operator in the GCC Hamiltonian, was adopted and formulated as: Q^=Λc|φjcmcφjcmc|, (49) where Λ is a large constant, and |φjcmc represents a two-body state comprising forbidden single-particle (s.p.) states of core nucleons. By setting Λ to high values, Pauli forbidden states are elevated to higher energies, which effectively suppresses their influence within the physical spectrum of the system.

This standard projection technique [67] may introduce minor numerical errors in the asymptotic region because of coordinate transformations. The supersymmetric transformation method [125-127] offers an alternative solution for the exclusion of Pauli-forbidden states. This method employs an auxiliary repulsive “Pauli core” within the original core-proton interaction, thereby effectively eliminating the influence of Pauli-forbidden states from the system.

The total wave function is expressed using hyperspherical harmonics as: ΨJMπ(ρ,Ω5)=ρ5/2γKψγKJπ(ρ)YγKJM(Ω5), (50) where K denotes the hyperspherical quantum number. The set γ=s1,s2,s3,S12,S,lx,ly,L encapsulates the quantum numbers other than K. Here, s represents spin, and l denotes orbital angular momentum. The function ψγKJπ(ρ) specifies the hyperradial wave function, and YγKJM(Ω5) represents the corresponding hyperspherical harmonic [125].

The resulting Schrödinger equation for the hyperradial wave functions can be written as a set of coupled-channel equations: [22m(d2dρ2(K+3/2)(K+5/2)ρ2)E˜]ψγKJπ(ρ)  +KγVKγ,KγJπ(ρ)ψγKJπ(ρ)  +Kγ0WKγ,Kγ(ρ,ρ)ψγKLπ(ρ)dρ=0, (51) where VKγ,KγLπ(ρ)=YγKJM|i>j=13Vij(rij)|YγKJM (52) and WKγ,Kγ(ρ,ρ)=YγKJM|Q^|YγKJM (53) is the non-local potential generated by the Pauli projection operator, as defined in Eq. 49.

To properly treat the positive-energy continuum space, the Berggren expansion technique is utilized for the hyperradial wave function: ψγKJπ(ρ)=nCγnKJπMBγnJπ(ρ), (54) where BγnJπ(ρ) denotes an s.p. state within the Berggren ensemble [55] (detailed in Sec. 2.4). To compute radial matrix elements using the Berggren basis, exterior complex scaling [128] is employed, whereby integrals are evaluated along a complex radial trajectory: Bn|V(ρ)|Bm=0RBn(ρ)V(ρ)Bm(ρ)dρ      +0+Bn(R+ρeiθ)V(R+ρeiθ)Bm(R+ρeiθ)dρ. (55) For potentials that decay as O(1/ρ2) (such as the centrifugal potential) or more rapidly (such as the nuclear potential), R should be large enough to circumvent all singularities, with the scaling angle θ selected to ensure that the integrals converge (see [129] for further details). Since the Coulomb potential is not square-integrable, its matrix elements diverge when the complex momenta kn=km. To address this, the “off-diagonal method” introduced in [130], where a slight offset ±δk is added to the linear momenta of the involved scattering wave-functions, was applied to facilitate the convergence of the resulting diagonal Coulomb matrix elements. The complex-momentum representation has also been adopted in other methods, e.g., in mean-field calculations [131, 132], to describe the continuum effect.

2.5.2
Deformed core

The fact that the open-shell nuclei are often accompanied by deformation, particularly around the dripline region, substantially changes the corresponding nuclear structure and affects the decay process. To this end, GCC method was extended to the deformed system by allowing the pair of nucleons to couple to the collective states of the core. Consequently, the wave function of the parent nucleus can be written as ΨJπ=Jpπpjcπc[ΦJpπpϕjcπc]Jπ, where ΦJpπp and ϕjcπc are the wave functions of the two valence protons and the core, respectively. Similar to the spherical case, the wave function ΦJpπp for the valence nucleons is constructed in Jacobi coordinates using the hyperspherical harmonics YγKJpM(Ω) for the hyperangle part, and the hyperradial part ψγK(ρ) is expanded in the Berggren ensemble [133, 67].

In the deformed case, the core+p+p Hamiltonian of GCC is H=i=c,p1,p23p^i22mi+i>j=13Vij(rij)+H^cT^c.m.. (56) This definition is similar to Eq. 46, except that H^c is the core Hamiltonian represented by excitation energies of the core Ejcπc. For nuclei exhibiting small shape deformations, the vibrational coupling model is utilized, following the methodologies outlined in [134, 135]. Conversely, for large quadrupole deformations, rotational coupling is employed, consistent with the non-adiabatic approach to deform proton emitters as in [136, 137]. This approach allows for the differentiated treatment of nuclear dynamics depending on the extent of deformation, thereby enhancing the accuracy of theoretical predictions in nuclear structure analysis.

By employing hyperspherical harmonics and the Berggren basis, the Schrödinger equation can be formulated as a coupled-channel equation. This formulation incorporates couplings not only among the hyperspherical basis states but also among the collective states of the core. The resulting complex eigenvalues provide information about the resonance energies and decay widths. However, in the case of medium-mass nuclei, proton decay widths typically fall below the numerical precision of calculations. Nevertheless, decay widths can be estimated using the current expression presented in [138], as demonstrated in previous works [139, 140, 67]. According to R-matrix theory, if the contribution from the off-diagonal part of the Coulomb interaction in the asymptotic region is neglected, the hyperradial wave function of the resonance, ψγK(ρ), is proportional to the outgoing Coulomb function HK+3/2+(ηγK,kpρ) [141-143]. By assuming a small decay width and adopting the expression ψ'/ψ=kpH+'/H+ [136, 137], the numerical derivative of the small-wave function in the asymptotic region that appears in the original current expression can be avoided, thereby significantly enhancing numerical precision [144].

2.6
Time-dependent approach

To tackle the decay process, a time-dependent formalism was developed, to allow precise, numerically stable, and transparent investigations of a broad range of phenomena, such as configuration evolution [145], decay rates [146], and fission [147]. For two-nucleon decay, the measured inter-particle correlations can be interpreted by propagating the solutions over long times. Despite previous efforts in this direction [148-150], capturing the asymptotic correlation of emitted particles still requires a precise description of the resonance wave function at large distances. To this end, we utilized the complex-momentum state Ψ0Jπ obtained using the GCC method. This state can be decomposed into real-momentum scattering states using the Fourier-Bessel series expansion in the real-energy Hilbert space [151]. The resulting wave packet is propagated by the time evolution operator through the Chebyshev expansion [152, 153]: eiH^t=n=0(i)n(2δn0)Jn(t)Tn(H^/), (57) where Jn are the Bessel functions of the first kind and Tn are the Chebyshev polynomials.

The time evolution was limited to the real momentum space, to restore the Hermitian property of the Hamiltonian matrix and ensure conservation of total density. Our implementation of the time-dependent approach is based on the integral equation, which allows maintaining high numerical precision by utilizing the Chebyshev expansion’s good convergence rate [154, 153]. Furthermore, the evolving wave packet has an implicit cutoff at large distances, preventing the divergence of the Coulomb interaction in momentum space. In practice, we considered interactions within a sphere of radius of approximately 500 fm, and the wave function remained defined in momentum space beyond this cutoff, preventing unwanted reflections at the boundary.

3

The calculations and Discussions

In this section, we primarily review the calculations by our developed D-IMSRG, G-IMSRG, and GCC. Sec. 3.1 presents the ground-state energies of 8,10Be isotopes as benchmark, along with the ground-state energies and charge radii of even-even nuclei from light beryllium to medium-mass magnesium isotopes using D-IMSRG. In Sec. 3.2, using VS-IMSRG, the residual proton-neutron interaction δVpn values in the upper fp shell were investigated, indicating the important role played by 3NF in explaining the experimental observations. Resonant states observed in the neutron-dripline 24O and the halo structure of the known heaviest Borromean nucleus 22C are presented in Sec. 3.3 for G-IMSRG. Furthermore, the low-lying resonant excited states in 22 are also predicted. Sec. 3.4 presents the applications of the GCC method, focusing on the exotic few-body decay beyond the dripline and the intriguing phenomena in open quantum systems. Specifically, the decay dynamics and properties of exotic two-proton (2p) emissions are discussed, including the impact of structure, deformation, and continuum effects.

3.1
The D-IMSRG calculations of deformed light nuclei

In [53], Yuan and his collaborators developed D-IMSRG, as formulated in Sec. 2.2, which better reflects the intrinsic structure of the deformed nucleus and captures more correlations through symmetry restoration. As a test ground, the D-IMSRG was first performed to calculate the ground-state energies of 8Be and 10Be, which are exotic nuclei with 2α cluster structure or elongated shapes, benchmarked against NCSM and VS-IMSRG. Subsequently, D-IMSRG was applied to describe the ground-state properties of even-even nuclei ranging from light beryllium to medium-mass magnesium isotopes. The optimized chiral NN interaction NNLOopt [156, 157], which gives good descriptions of nuclear binding energies, excitation spectra and neutron matter equation of state without the inclusion of the 3N force, was used during the calculation in [53] with ω=24 MeV.

The ground-state energies of 8Be and 10Be were first studied though D-IMSRG, as shown in Fig. 4, with and without the approximate angular momentum projection. It was found out that the trend of calculated energy by D-IMSRG is similar to those of NCSM [16, 84-87] and multi-reference IMSRG [19] calculations, exhibiting an exponential convergence with respect to the basis-space size Nshell. The energies extrapolated to infinite model space using an exponential fit based on Eq. (22) are depicted in Fig. 4. The extrapolated “Extrap” results fitted with different data points are provided in Fig. 4 along with their uncertainties, verifying that the calculation results of D-IMSRG converge exponentially with an increase in Nshell, thereby demonstrating the reliability of the calculations.

Fig. 4
(Color online) Ground-state energies for 8Be and 10Be calculated by D-IMSRG, with and without projection correction, are shown as a function of basis-space size Nshell. Symbols below “Extrap” represent the energies extrapolated to the infinite basis space using an exponential fit, based on different data points: Nshell=3-7, 3-10, and 6-10. The fitting uncertainties are indicated by error bars. Extrapolation uncertainties in NCSM and VS-IMSRG calculations are also represented by error bars. Experimental data were taken from AME2020 [155], and the figure was taken from [53]
pic

In Fig. 4, the angular momentum projection corrections are -7.2 MeV and -5.9 MeV for 8Be and 10Be, respectively. These significant corrections are caused by the large deformations of these two nuclei. The results of NCSM and VS-IMSRG and the experimental data are also shown in Fig. 4.

The results of VS-IMSRG underestimate the ground-state energy in 8,10Be, which may be attributed to the omission of higher-order collective excitations that are not handled well in VS-IMSRG at the IMSRG(2) level, as discussed in [158, 159]. However, this omission can be compensated by the angular momentum projection correction through D-IMSRG, as illustrated in Fig. 4.

The ground-state energies and two-neutron separation energies calculated by D-IMSRG for 6-16Be are shown in Fig. 5 (top panel and bottom panel, respectively), along with VS-IMSRG calculations and experimental data. The angular momentum projection lowered the ground-state energies of 8-16Be by about 5-6 MeV, making the calculated energies closer to the data. Both D-IMSRG and VS-IMSRG calculations indicated that the neutron dripline of beryllium isotopes was at 12Be, contrary to the experimental position of 14Be, which may be due to the absence of a continuum effect [105, 107, 57, 160].

Fig. 5
(Color online) Ground-state energies (upper panel) and two-neutron separation energies S2n (lower panel) of 6-16Be calculated by D-IMSRG with and without projection correction. The VS-IMSRG results and experimental data [155] are also presented for comparison. The figure is taken from [53]
pic

In [53], the heavier nuclei of C, O, Ne, and Mg isotopes were also calculated by D-IMSRG, as shown in Fig. 6 along with VS-IMSRG calculations and experimental data. The D-IMSRG calculations with the projection correction agreed well with VS-IMSRG results and experimental data. The angular momentum projection corrections were zero for the closed-shell nuclei 14C and 14,16,22,24,28O, indicating the spherical characteristics of these nuclei. However, for the expected closed-shell nuclei of 12,22C, the angular momentum projection corrections were not zero but -5.5 MeV and -2.7 MeV, respectively, indicating their deformation. For Ne and Mg isotopes, the projection results provided energy gains of about 3-6 MeV near the neutron number N=20 island of inversion [161, 162]. There is strong configuration mixing between sd and pf shells in nuclei located in the region of the island of inversion. This cross-shell mixing is missing in the VS-IMSRG calculations although a multi-shell VS-IMSRG has been proposed [163]. Therefore, it can be concluded that deformation effectively brings the deformation orbitals into the wave function of the state in D-IMSRG calculations.

Fig. 6
(Color online) Ground-state energies of C, O, Ne, and Mg isotopes. D-IMSRG results are extrapolated to the infinite basis space using the Nshell=6-10 data points, whereas the VS-IMSRG results are extrapolated based on Nshell=8-13. In VS-IMSRG calculations, the model space includes both protons and neutrons in 0p3/2,1/2 for 6-14C; protons in 0p3/2,1/2 and neutrons in 1s1/20d5/2,3/2 for 14-22C; and both protons and neutrons in 1s1/20d5/2,3/2 for O, Ne, and Mg isotopes. The experimental data were taken from AME2020 [155]. The figure was taken from [53]
pic

For nuclei, another important observable is the charge radius. The radii of the studied isotopes were also calculated in [53]. Compared with the calculation of the ground-state energy, the convergence of the radius calculation showed a different trend with increasing basis-space size, as discussed in [32, 168, 87] and as also observed in the D-IMSRG calculations [53], indicating that the exponential fit was not applicable to the radius. Therefore, the authors in [53] did not use extrapolation to fit the radius in the calculations. The angular momentum projection correction was also estimated by the HF wave function in the study. As shown in Fig. 7, the charge radii of Be, C, O, Ne, and Mg isotopes were investigated, along with the VS-IMSRG calculations and experimental data. The projection correction to the charge radius is small, making the D-IMSRG radii with and without the projection correction close to each other and in good agreement with the VS-IMSRG calculations except for 8Be, where the D-IMSRG radius is larger. This difference can be attributed to the large deformation of 8Be. Therefore, it can be concluded that the calculated charge radii by D-IMSRG and VS-IMSRG are reasonable compared with experiment data, as shown in Fig. 7 although the NNLOopt interaction tends to underestimate the nuclear radii, as noted in [157].

Fig. 7
(Color online) Charge radii calculated by D-IMSRG in a basis space with Nshell=10 and using VS-IMSRG with Nshell=13 for Be, C, O, Ne, and Mg isotopes. Experimental data were taken from [157, 164-167]. The figure was taken from [53]
pic
3.2
δVpn bifurcation by the VS-IMSRG

The VS-IMSRG was first introduced by Tsukiyama et al. in 2012 [39]. This method combines the SM and IMSRG to nonperturbatively derive effective valence-space Hamiltonians and operators, as detailed in Sec. 3.2. Recently, it has been applied to describe the δVpn values in the upper fp shell, incorporating a chiral three-nucleon force (3NF), as reported in [169].

Nuclear binding energy B(Z,N) represents the total interaction energy of interacting nucleons in the nucleus with Z protons and N neutrons. Differences in binding energy can isolate specific types of interactions and provide insights into modifications in nuclear structure [170, 171]. The double binding energy difference denoted as δVpn, has been used as an important mass filter to extract the residual proton-neutron (pn) interaction [172-174], particularly for the N=Z nuclei. The residual proton-neutron interaction δVpn can be extracted using δVpnee(Z,N)=14[B(Z,N)B(Z,N2)B(Z2,N)+B(Z2,N2)], (58) for the nucleus with N=Z= an even number, and δVpnoo(Z,N)=[B(Z,N)B(Z,N1)B(Z1,N)+B(Z1,N1)], (59) for the nucleus with N=Z= an odd number.

Weakly bound proton-rich nuclei are attracting interest in novel structure [175]. In [169], the masses of 62Ge, 64As, 66Se, and 70Kr were measured for the first time. Additionally, the masses of six N=Z-1 nuclides 61Ga, 63Ge, 65As, 67Se, 71Kr, and 75Sr were redetermined with improved accuracy, using a novel method of isochronous mass spectrometry conducted at the Heavy Ion Research Facility in Lanzhou (HIRFL). These newly measured masses provide updated δVpn values, which offer great test ground for state-of-the-art theoretical calculations. The updated δVpn values show a clear increasing trend in δVpnoo beyond Z=28, which is interpreted as an indication of the restoration of pseudo-SU(4) symmetry in the fp shell, as suggested in [176, 177]. In contrast, δVpnee shows a decreasing trend that was previously observed in the lower mass region [173, 178]. These δVpn values were extracted using predicted masses from frequently used mass models; however, none of these models successfully reproduce the bifurcation in δVpn values [169], except for the ab initio VS-IMSRG calculations. Within the ab initio VS-IMSRG calculation, a chiral 2NF plus 3NF, labeled by EM1.8/2.0 [24], is adopted, which can reproduce well the ground-state energies up to A≈100 region nuclei [179, 7, 28]. The effective Hamiltonian in the full fp-shell above the 40Ca core was derived using the VS-IMSRG, and the final diagonalization of the valence-space Hamiltonian was realized using KSHELL [95].

As shown in the lower panel of Fig. 8, VS-IMSRG calculations with chiral 2NF plus 3NF reproduce the experimental δVpn for the N=Z+2 nuclei exceptionally well. For the odd-odd N=Z nuclei, 62Ga,66As,70Br and 74Rb, the ground states have been identified as (T,Jπ)= (1,0+) [155]. VS-IMSRG calculations, which inherently incorporate both T=0 and T=1 pn correlations, achieve a commendable match with the experimental δVpnoo values for nuclei ranging from 58Cu to 70Br. Particularly noteworthy is that the calculation successfully reproduces the observed increasing trend in δVpnoo with an increase in nucleon number A. Our calculations consistently attributed an isospin of T=1 to the ground states of these odd-odd nuclei, aligning with experimental results, with the exception of 58Cu. Moreover, the decreased trend in the even-even δVpnee was also well reproduced by our VS-IMSRG calculations.

Fig. 8
(Color online) Experimental δVpn for (a) N=Z and (b) N=Z+2 nuclei beyond A=56, compared with the ab initio VS-IMSRG calculations. Data uncertainties are indicated by the size of symbols. δVpn values from ab initio calculations using 2NF + 3NF and only 2NF plotted as red and blue lines (solid lines for even-even and dashed lines for odd-odd), respectively. The figure is taken from [169]
pic

In mass regions with extremely asymmetric N/Z ratios, 3NF usually provides a repulsive effect on the neutron-neutron (nn) and proton-proton (pp) interactions [180, 28], which is essential for the emergence of new magic numbers [180] and also for reproducing the neutron or proton dripline [28]. To understand the role played by 3NF in the δVpn of upper fp-shell nuclei, we performed calculations using only a chiral 2NF at N3LO. The results using only 2NF show significant deviations from those calculated with 3NF included, as demonstrated in Fig. 8. Specifically, the agreement with experimental δVpn values was markedly poor in the calculations without the 3NF included. Additionally, the predicted isospins of ground states for odd-odd nuclei ranging from 62Ga to 74Rb were all erroneously identified as T=0 in the calculations with the 3NF included, contradicting experimental data. Furthermore, without the inclusion of 3NF, the calculated δVpnoo values of N=Z nuclei were lower than δVpnee calculated with 3NF included. The 3NF enhances the pn correlations in N=Z nuclei favoring a T=1 isospin coupling, which changes the δVpn behavior.

3.3
The G-IMSRG with the coupling to the continuum

A novel G-IMSRG [57] was developed by Hu et al., using the complex-energy Berggren representation, as introduced in Sec. 2.4. This advanced G-IMSRG is capable of describing the weakly bound and unbound nature of nuclei in the vicinity of nuclear driplines. We applied G-IMSRG to oxygen and carbon isotopes. Recent experiments [181-184] highlight that 22C is a Borromean halo nucleus, with an experimentally deduced root-mean-squared matter radius of 3.44±0.08 fm [184]. The continuum coupling plays a vital role in generating the extended density distribution. Notably, experimental information about the excited states of 22C, which can offer additional insights into its halo structure, is lacking. In this study, we performed an ab initio G-IMSRG calculation of the halo 22C, using both chiral 2NF NNLOopt and 2NF plus 3NF NNLOsat interactions. The NNLOopt interaction matrix elements were expanded within 12 major HO shells at a frequency of ω=20 MeV, whereas the NNLOsat interaction was truncated at 13 major HO shells with ω=22 MeV [185, 158]. The NNLOopt potential provides a good description of nuclear structure, including binding energies, excitation spectrum, and dripline position without the need for 3NF [156]. The NNLOsat interaction can provide accurate descriptions of charge radii in light- and mid-mass isotopes [186].

For the sd shell, the neutron 0d3/2 is a narrow-resonance orbital. With no centrifugal barrier of the l=0 s partial wave, a weakly bound 1s1/2 orbital can significantly affect the spatial distribution of the wave functions of weakly bound nuclei. Therefore, the 0d3/2 and 1s1/2 orbitals should be treated in the Berggren basis, which includes coupling to the continuum, whereas other orbitals can be treated in the real-energy HF basis to reduce the computational cost, as in [57].

Although the Hamiltonian (5) is intrinsic, the IMSRG wave functions are expressed in laboratory coordinates. Therefore, considering center-of-mass (CoM) motion corrections may be necessary. Our previous work has indicated that the CoM effect on an intrinsic Hamiltonian is small for low-lying states [105]. Thus, the approximation method suggested in [187, 37] can be adopted to estimate the CoM effect in IMSRG calculations. Fig. 9 presents real-energy EOM-IMSRG calculations without and with the CoM multiplication term βHCoM=β(P22mA+12mAω˜2R232ω˜). Note that the value of the CoM vibration frequency ω˜ can differ from the ω frequency of the HO basis [187]. As illustrated in Fig. 9, the CoM effect remains negligible for these low-lying states.

Fig. 9
(Color online) 24O spectra calculated using NNLOopt and NNLOsat interactions. The first two columns display the results from real-energy EOM-IMSRG calculations (denoted as R-IMSRG) without and with CoM treatment βHCoM, using the multiplier β = 5. The subsequent three columns present the EOM-G-IMSRG calculation (denoted as G-IMSRG), which are compared with the data from [188, 189]. Resonant states are highlighted by shading, and their widths (in MeV) are annotated nearby. The gray shading indicates the continuum region above the particle emission threshold. The figure is taken from [57]
pic

As described in Sec. 2.4, the equation-of-motion approach can be used to calculate nuclear excited states. Fig. 9 presents the calculated spectrum of 24O, showing resonant excited states. The EOM Gamow IMSRG (indicated by EOM-G-IMSRG) calculations reproduced the experimental excitation energies and resonance widths of the observed states well. A high excitation energy of the first 2+ state supports the shell closure at N = 16 in the oxygen chain. Additionally, the calculation predicted three resonant states around the excitation energies of 8 MeV with Jπ=2+4+, aligning with the experimentally ambiguous states observed around 7.6 MeV [189]. This finding is consistent with the complex coupled cluster calculation which uses a schematic 3NF [18].

The Borromean halo nucleus 22C poses significant challenges for many theoretical models [58, 59, 60]. Our GHF calculations suggested that the neutron ν1s1/2 orbital is weakly bound. Therefore, the two-neutron configuration (ν1s1/2)2 is responsible for the formation of the halo structure [181-184]. Figure 10 presents the ground-state densities of 22C calculated by the real-energy R-IMSRG and complex-energy G-IMSRG using two different chiral interactions. The density was computed by an effective density operator evolving within the Magnus framework (16, 17). The G-IMSRG calculation revealed a long tail in the density distribution, supporting the halo structure of 22C.

Fig. 10
22C ground-state densities calculated by real-energy IMSRG (R-IMSRG) and Gamow IMSRG (G-IMSRG), displayed on a logarithmic scale. The inset provides a detailed view of the densities in the central region of the nucleus on a standard scale. The figure is taken from [57]
pic

To assess the continuum effect of the s channel on the properties of 22C, we performed two types of G-IMSRG calculations using: (i) discrete s states obtained from the real-energy HF calculation, and (ii) Berggren s states obtained from the complex-energy GHF calculation. In both calculations, the neutron d3/2 channel remained in the GHF basis. Calculations using NNLOopt with discrete real-energy HF s states yielded a radius of 2.798 fm for the 22C ground state, which increased to 2.928 fm upon incorporating the continuum s wave. Similarly, the calculations using NNLOsat yielded a radius of 2.983 fm for the real-energy discrete s states and 3.139 fm for the continuum GHF s wave. The experimentally estimated radius was reported to be 5.4±0.9 fm in an earlier work [181], and later works found it to be 3.44±0.08 fm [184] and 3.38±0.10 fm [190]. These findings highlight the crucial role of the continuum s wave in predicting the radius and understanding the halo structure.

Currently, no experimental data are available for the excited states in 22C. Figure 11 displays the EOM-G-IMSRG predictions of low-lying states, benchmarked against results from complex CC calculations. Both methods yielded consistent results. The first 2+ excited state was bound in both G-IMSRG and coupled cluster calculations. We found that the 21+ state was dominated by the proton 1p1h excitation from the proton 0p3/2 hole to proton 0p1/2 particle orbits. The proton 21+ excited state was lower in energy than the neutron 2+ state calculated by the real-energy SM with the 14C core [191, 192]. The real-energy R-IMSRG results in Fig. 11 show a neutron 2+ energy similar to that in [191, 192]. Additionally, there were superposed resonant states with Jπ=1+4+ at energies of 3.5-4.0 MeV and widths of 0.15-0.25 MeV. The NNLOsat calculations yielded slightly higher excitation energies and broader resonant widths for these states, as illustrated in Fig. 11. The resonances were primarily characterized by neutron 1p1h excitations from the v0d5/2 hole to v0d3/2 particle orbitals. Structure and decay modes of loosely-bound nuclei are of interest in many respects [193, 194].

Fig. 11
Excited states in 22C predicted by R-IMSRG and G-IMSRG using two different chiral interactions compared with complex coupled cluster results. The channels listed at the top of the panel indicate that the partial waves are treated in the resonance and continuum Berggren representation. The other labels are the same as those in Fig. 9. The figure is taken from [57]
pic
3.4
Few-body decay by GCC

In this section, the discussion centers on the decay properties of open quantum systems, with a particular focus on two-proton (2p) emitters, exploring how deformation and continuum effects influence the decay dynamics of these systems. Previous research highlighted the significant role that these factors play in shaping the decay characteristics of 2p emitters. A critical aspect of our analysis involves extracting structural information from these systems through the measurement of asymptotic nucleon-nucleon correlations, which are experimentally accessible.

In analyzing these correlations, the objective is to deepen our understanding of the universal properties inherent to open quantum systems. This approach not only elucidates the fundamental interactions within these systems but also provides a framework for interpreting experimental results in terms of underlying nuclear structure and dynamics. Such insights are invaluable for advancing our comprehension of the complex behaviors exhibited by open quantum systems under various conditions.

3.4.1
Impact of structure on the decay process

As the heaviest 2p emitter identified to date, 67Kr serves as a pivotal case study for examining the influence of structure on decay properties. Notably, shape coexistence is often observed with Kr isotopes. In addition, the deformation effects can significantly impact the lifetime of a decaying system, as evidenced by prior research on one-proton (1p) emitters [135-137, 144, 195-199]. This offers a good opportunity to study how the 2p decay properties change as a function of deformation.

As discussed in [68], Fig. 12a illustrates the proton Nilsson levels (labeled by asymptotic quantum numbers Ω[NnzΛ]) within the core-p potential. At modest deformations, specifically |β2|0.1, the valence protons predominantly occupy the f5/2 shell. The half-life predicted under vibrational conditions, calculated as T1/2>218 ms, significantly surpasses the experimentally observed value by more than an order of magnitude, as depicted in Fig. 12b. This discrepancy underscores the need for further theoretical refinement and is consistent with prior theoretical estimates [200, 201], suggesting an intricate interplay between nuclear structure and decay dynamics in 67Kr.

Fig. 12
Top: Nilsson levels Ω[NnzΛ] of the deformed core-p potential as a function of the oblate quadrupole deformation β2 of the core. The dotted line indicates the valence level primarily occupied by the two valence protons. Bottom: Decay width (half-life) of the 2p ground state radioactivity of 67Kr. The solid and dashed lines mark the results for the rotational and vibrational couplings, respectively. The rotational-coupling calculations were carried out by assuming that the 1/2[321] orbital is either occupied by the core (9/2[404]-valence) or valence (1/2[321]-valence) protons. The figure is taken from [68]
pic

As the core deformation increases, a notable oblate gap at Z=36 emerges due to the descending 9/2[404] Nilsson level, which stems from the 0g9/2 shell. This gap plays a crucial role in shaping the oblate ground state (g.s.) configurations of proton-deficient Kr isotopes [202-204]. As the oblate deformation intensifies, the structure of the valence proton orbital transitions from the 9/2[404] (l=4) state to the 1/2[321] orbital, featuring a significant l=1 component. The precise level crossing between 1/2[321] and 9/2[404] is contingent on the specifics of the core-proton parametrization, yet the overarching pattern remains consistent as depicted in Fig. 12a: a shift from the 2p wave function dominated by l=4 components towards l=1 components as the oblate deformation progresses.

Figure 12b illustrates the predicted 2p decay width within the rotational model for two scenarios: (i) the 1/2[321] level is integrated into the core with the valence protons predominantly residing in the 9/2[404] level, and (ii) the valence protons primarily occupy the 1/2[321] level. Consequently, the reduction in l content within the 2p wave function markedly enhances the decay width.

At a deformation of β20.3, aligned with estimates from mirror nuclei and various theoretical models [205-208], the calculated 2p ground state half-life of 67Kr is 247+10 ms. This estimation not only concurs with experimental findings [209] but also underscores the significant impact of nuclear shape and structural dynamics on the decay properties of 67Kr.

3.4.2
Dynamics of two-proton decay

For the light 2p emitters, both direct and sequential decays are possible [210, 211], providing a good opportunity to study the decay dynamics as well as the interplay between nucleon-nucleon correlation and single-particle emission.

An illustrative example is the decay of 6Be, where the neighboring g.s. of 5Li exists as a broad resonance, characterized by a proton decay width of Γ=1.23 MeV [212]. The complex three-body decay dynamics of 6Be remain an area of active research and debate, with existing studies indicating unresolved aspects of the decay [141, 211, 213-219]. Theoretical predictions of a diproton structure and experimental observations of broad angular correlations between emitted protons lead to conflicting interpretations.

Based on the time-dependent approach, Fig. 13a,b depicts the temporal evolution of the 2p density and momentum distribution in the ground state of 6Be across an extensive time frame. Initially, at t=0, the wave function inside the nucleus shows a localized character with a density distribution exhibiting two maxima. These maxima represent diproton (compact) and cigar-like (extended) configurations based on the relative distances between the valence protons [220].

Fig. 13
The density and momentum distributions of two-nucleon decays from the g.s. of 6Be (left) and 6He′ (right) for four different time slices. The density distributions are shown in the Jacobi-T coordinates (see Fig. 3). The momentum distribution of the second nucleon is depicted with respect to the momentum of the first nucleon. To clearly show the asymptotic wave function, all the particle densities (in fm-1) are multiplied by the polar Jacobi coordinate ρ. The dimensionless momentum (angular) distributions are divided by the total momentum k. The figure is taken from [70]
pic

As the decay progresses, Fig. 13b highlights two predominant flux branches. The primary branch shows protons emitted at narrow angles, indicating the presence of a diproton structure during the tunneling phase. This phenomenon is interpreted through nucleonic pairing, which favors low angular momentum states, reducing the centrifugal barrier and enhancing the 2p tunneling likelihood [216, 149, 150, 220]. The secondary branch illustrates protons emitted in nearly opposite directions. Despite their spatial separation, these protons exhibit simultaneous decay, suggesting three-body decay dynamics characterized by correlated decay pathways of the protons with respect to the core. This configuration reveals intricate interplays within the decay process, shedding light on the multifaceted nature of three-body decays in light 2p emitters.

After tunneling through the Coulomb barrier, the two emitted protons tend to gradually separate due to Coulomb repulsion. This is reflected in the bent trajectory of the diproton decay branch and gradual reduction of the momentum alignment observed in Fig. 13a,b. Eventually, the 2p density becomes spatially diffuse, which is consistent with the broad angular distribution measured in [217]. One may notice that even beyond 100 fm (at t≈2 pm/c), the Coulomb repulsion tends to reduce the inter-proton correlation. According to our calculations, the angular correlation starts to stabilize only after very long times greater than 9 pm/c. Therefore, to obtain meaningful estimates of asymptotic observables, very long propagation times are required.

After the protons tunnel through the Coulomb barrier, the Coulomb repulsion between them leads to an increasing spatial separation, impacting their trajectories and momentum alignment. This dynamic is depicted in the bending trajectory of the diproton decay branch, and a reduction in momentum alignment is observed in Fig. 13a,b. Over time, the 2p density distribution becomes increasingly diffuse, aligning with the broad angular distributions reported in experimental studies such as [217].

To provide insights into the nuclear-Coulomb interplay in two-nucleon decay processes, an artificially unbound variant of 6He, denoted as 6He′, was built to study its two-neutron decay. Initially, the density distributions of 6He′ and 6Be display similarities, as depicted in Fig. 13, a consequence of the isospin symmetry inherent in the nuclear force.

However, the absence of Coulomb repulsion in the 6He′ scenario significantly influences the decay dynamics. In the case of 6He′, the dineutron decay branch is more pronounced, with the emitted neutrons maintaining their spatial correlations over time more robustly than their proton counterparts in 6Be. This differential behavior leads to distinct nucleon-nucleon correlation patterns in the asymptotic regime. As a result, the nucleon-nucleon correlations observed in 6He′, as illustrated in Fig. 14, diverge markedly from those in 6Be, underscoring the critical role of Coulomb forces in shaping the decay pathways and final-state interactions in these mirror nuclei.

Fig. 14
Asymptotic energy (left) and angular (right) correlations of emitted nucleons from the g.s. of 6Be (top) and 6He′ (bottom) calculated at t=15 pm/c with different strengths of the Minnesota interaction [221]: standard (solid line), strong (increased by 50%; dashed line), and weak (decreased by 50%; dash-dotted line). Also shown are the benchmarking results obtained using the Green’s function method (GF; dotted line) with standard interaction strength; θk is the opening angle between kx and k1 in the Jacobi-Y coordinate system, and Epp/nn is the kinetic energy of the relative motion of the emitted nucleons (see Fig. 3 for definitions). The figure is taken from [70]
pic
3.4.3
Nucleon-nucleon correlations

The experimental measurements of nucleon-nucleon correlations among emitted nucleons provide data pertinent to the nuclear structure. This methodology serves as a distinctive avenue for investigating the internal structure and decay mechanisms of 2p emitters. However, the nucleon-nucleon correlations recorded by detectors are influenced by final-state interactions and distort the original nuclear correlations. Consequently, self-consistent theoretical frameworks are urgently required to establish a linkage between the nuclear structure and observed asymptotic correlations. In this section, we elucidate how the GCC method and time-dependent (TD) approach have been effectively employed to address such challenges.

Compared to the neutron dripline, the proton dripline is relatively closer to the line of β-stability, facilitating the acquisition of 2p correlation data in several instances, as evidenced by the findings reported in studies [222, 223]. Remarkably, the energy correlation of protons emitted from the ground state of 12O closely resembles those observed in other sd-shell 2p emitters, such as 16Ne and 19Mg. This resemblance suggests potential structural similarities in the configurations of valence protons across these isotopes despite their differing proton numbers. This observation underscores the intricate interplay between nuclear structure and decay processes, hinting at underlying uniformities in the spatial and energetic distributions of valence protons within this specific shell.

To elucidate the 2p correlation patterns observed in 12O and its isotonic neighbor 11O, a time-dependent approach was employed, as detailed in recent research [71]. The initial 2p density configurations in 11,12O exhibit a prominent diproton arrangement alongside a secondary, cigar-like structure [224, 220], bearing resemblance to configurations typical of p-shell nuclei. However, in the case of 12O, the protons emerging from these configurations coalesce, leading to a broad distribution as reported in [71]. This pattern contrasts starkly with the decay dynamics of 6Be, as shown in Fig. 13 and [70]. These observations align with the flux current calculations presented in [220], which indicate a competitive interplay between diproton and cigar-like decay modes, culminating in a so-called democratic decay process. This comparative analysis underscores the complex interdependencies within nuclear decay pathways and highlights the distinctive decay characteristics of 12O relative to its nuclear peers.

Consequently, the asymptotic 2p correlations for 12O, as depicted in Fig. 15, demonstrate robust alignment with the empirical data [223] and corroborate prior theoretical investigations [225]. Notably, the subtle discrepancies between the experimental results and theoretical predictions, enhanced via Monte Carlo simulations incorporating experimental acceptance and resolution, might be mitigated by refining the Minnesota force initially utilized for modeling the nucleon-nucleon interactions. More importantly, the absence of distinct diproton emissions during the temporal evolution suggests that a low-Epp correlation does not inherently imply a diproton decay. Moreover, recognizing that the 2p system may manifest as a subthreshold resonance characterized by a broad decay width around 1 MeV [226], is critical. This continuum characteristic potentially influences the energy correlation observed in 2p emitters possessing minimal decay energies, underscoring the complex interplay between nuclear structure and decay dynamics.

Fig. 15
Asymptotic (a) energy and (b) angular correlations of the protons emitted from the two-proton unbound 12O isotope. Also shown is (c) the momentum scheme for three-body system. Theoretical distributions were obtained within the time-dependent approach (TD, red line) at t=15 pm/c. MC (green line chart) labels the Monte Carlo simulation of TD results which include experimental resolution and efficiency [223]. The calculated 2p correlations (TD and MC) are compared with experimental data (Exp, blue histogram) of [223]. A is the mass number and k1, k2, and kc are the momenta of the nucleons n1 and n2, and the core c, respectively in the c.m. coordinate frame. The figure is taken from [71]
pic

To further illustrate, the lightest oxygen isotope, 11O, has been characterized as a broad structure encompassing multiple resonances [224, 227-230]. Utilizing GCC calculations, four low-lying states with quantum numbers Jπ=3/21, 5/21+, 3/22, and 5/22+ were predicted within the experimental energy range [220]. Each state exhibits a substantial decay width ranging from 1 MeV to 2 MeV. In the time-dependent calculations [71], these states were propagated individually, effectively disregarding potential interference effects. This approach underscores their significant decay widths, which foster robust continuum coupling, resulting in a more homogeneous density distribution throughout the decay process compared to 12O, as noted in [71].

The emergent Y-type correlations exhibit a pronounced dependency on angular momentum, which can be instrumental in experimentally determining spin assignments. To simulate the asymptotic correlations of the emitted valence protons, the correlations of these four states were amalgamated, utilizing the weights derived from resonance-shape fitting [224]. This composite correlation, depicted in Fig. 16(a), closely aligns with the experimental observations shown in Fig. 16(b), lending credence to the hypothesis that the observed broad structure integrates states with Jπ=3/2 and 5/2+ [224, 227]. Moreover, Figs. 16(g) and (h) illustrate the energy and angular correlations, respectively, adjusted for experimental resolution and efficiency. The qualitative agreement between these corrected simulations and experimental data provides additional support for the multi-resonance composition within the observed peak, further validating the complex resonance structure of 11O.

Fig. 16
(a) Theoretical and (b) experimental Jacobi-Y correlations of the two protons emitted from the broad low-energy structure in 11O; (c)-(f) the corresponding contributions predicted from each low-lying state. The experimental resolution and efficiency have been considered in (a) Monte Carlo simulations. Also shown are the corresponding (g) energy and (h) angular correlations obtained in the time-dependent (TD) calculations (red line), Monte Carlo (MC) simulations (green step chart), and experiments (Exp, blue histogram). The figure is taken from [71]
pic
3.4.4
Exotic decays in open quantum systems

In addition to 2p decay, there exist several other exotic decay modes in nuclear physics, such as two-neutron decay, two-alpha decay, and multi-particle emission. These decay processes exhibit intriguing behaviors that are unique to the quantum realm. The classical understanding of radioactive decay is based on the principle that the rate of decay is directly proportional to the quantity of the radioactive substance present. This model fundamentally assumes that the decay is a stochastic process at the level of individual particles, meaning that the probability of decay is independent of the system’s previous history.

However, this classical perspective is challenged within the quantum framework due to phenomena such as memory effects and quantum entanglement. These effects introduce deviations from the exponential decay law, traditionally used to describe radioactive decay. Notably, quantum mechanics reveals that decay probabilities can exhibit non-exponential behavior at very short and very long timescales. Theoretical and experimental studies have demonstrated that quantum systems do not always follow the expected exponential decay pattern [151, 231-241].

These findings underscore the complex nature of decay processes in quantum systems, where the inherent properties of the particles and their interactions can lead to observable departures from classical predictions. The implications of these quantum behaviors are profound, impacting our understanding of fundamental decay processes and the predictive models used in nuclear physics.

In analyzing the non-exponential decay in quantum systems, one crucial concept is the survival amplitude A(t). This amplitude is defined as the overlap between the initial quantum state Ψ(0) and the state at a later time Ψ(t). Mathematically, the survival amplitude can be expressed and computed using the Fourier transform of the spectral function ρ(E), as illustrated in [242]: A(t)=Ψ(0)|Ψ(t)=0+ρ(E)eiEtdE. (60) In this formulation, ρ(E), which is the probability density of finding a system with energy E, is derived from the squared modulus of the projection of the state vector Ψ onto the real-energy eigenstates E|Ψ. Thus, ρ(E)=|E|Ψ|2 represents the distribution of the initial state over the various energy eigenstates.

The survival probability S(t), which is a measure of the likelihood that the system remains in its initial state at time t, is then calculated in a straightforward manner from the survival amplitude: S(t)=|A(t)|2. (61) This probabilistic measure reflects how the state evolves over time, deviating from its initial configuration. This deviation is a key indicator of quantum mechanical effects in decay processes and provides insight into the complex nature of quantum dynamics.

Non-exponential decay of a threshold resonance. The survival probability of a quantum state is intrinsically linked to the energy distribution described by the spectral function of the system. Typically, for a system exhibiting exponential decay, the spectral function is expected to follow a Breit-Wigner distribution, characterized by a Lorentzian shape centered around the resonance energy with a width corresponding to the decay rate. However, this does not hold for near-threshold states, particularly those with large decay widths [233, 151, 240, 241, 243].

These near-threshold states often display significant deviations from the exponential decay law. The temporal evolution of resonance states has been shown to involve both exponential and non-exponential components [244, 245]. The exponential components, characterized by a rapid decrease in probability, dominate the early time behavior of the decay process. However, as these components decay, the non-exponential elements become more prominent.

Over time, as the influence of the exponential decay wanes, a transition to a power-law regime becomes inevitable. This regime is indicative of the long-time tail behavior common to quantum systems with broad spectral distributions. Such transitions are crucial for understanding the full dynamics of decay processes, particularly in scenarios where classical exponential decay laws fail to capture the complexities introduced by quantum mechanical principles.

Figure 17a,c illustrates this transition, highlighting how the decay initially follows an exponential decrease before transitioning to a power-law decay. This behavior underscores the complex nature of quantum decay processes and the need for a deeper exploration into the underlying physics, particularly for states near the energy threshold or those with significant quantum mechanical interactions.

Fig. 17
Survival probability S(t) as a function of time (relative to T1/2) for (a) the 1/2- state of 9He for different depths V0 of the Woods-Saxon (WS) potential, and (c) the low-lying states of 9N. The near-threshold behavior of the spectral function ρ (relative to the Breit-Wigner distribution) is shown for (b) neutron and (d) proton s,p, d partial waves. The polar angle φ indicates the location of the resonant state in the complex-k plane. Also shown is the survival probability for the virtual 1/2+ state in 9He. For this state, T1/2 was assumed to be 20 fm/c. The figure is taken from [246]
pic

The universality of the transition from exponential to non-exponential decay is a key characteristic of quantum decay processes, the specific dynamics of which are deeply influenced by several factors. These include the structure of the initial state, the chosen decay channel, and notably, the nature of the scattering continuum which drives the post-exponential decay behavior [246].

To provide a concrete example of these dynamics, the survival probability of the 1/2- resonant state in 9He has been analyzed by adjusting the depth of core-nucleon potential used in the calculations [246]. This adjustment affects the resonant states, which can be characterized in the complex-k plane. The positioning of these states is determined using the polar angle φ=cot1(2E/Γ)/2, which offers insights into the relative contributions of the exponential and non-exponential decay components in state evolution.

In this analysis, as depicted in Fig. 17a, the deviations from exponential decay become increasingly pronounced as φ shifts toward -45°. This angular movement suggests a strengthening of the non-exponential decay component, particularly in threshold resonances where the resonant energy Er is approximately equal to the decay width Γ. This result is consistent with [240, 247, 248], which indicate that post-exponential decay features tend to dominate more rapidly in systems where the resonance lies near the decay threshold. Such resonances provide a clearer and more readily observable transition to non-exponential decay, making them ideal subjects for experimental and theoretical studies aiming to explore quantum decay dynamics beyond the conventional exponential model.

Interference between near-lying states. Besides the threshold effect, the decay dynamics of quantum systems can also be significantly influenced by the interaction between closely lying resonances, particularly when these states share the same spin-parity configuration. This scenario leads to an intricate interplay due to the interference between overlapping resonances, which in turn modify the decay characteristics [249-251].

The mechanism underlying this behavior is related to the quantum interference effects, whereby the wavefunctions of the resonant states overlap and coherently interact with the continuum states. This interplay can lead to a redistribution of decay widths among the resonances, with one or more states experiencing an enhancement in decay widths due to the increased coupling [252-255]. This enhanced coupling is a critical factor in the non-exponential decay characteristics observed in such systems, as it directly impacts the decay pathways and probabilities.

To provide a detailed examination of how continuum coupling affects the spectral functions of overlapping resonances, a hypothetical study was conducted on a two-level 0+ system in an artificial nucleus labeled as 6He′ [246]. This recent study [246] explored how two 0+ states, lying close in energy, interact with each other and the continuum, highlighting that the interference between these states not only affects their individual decay rates but also alters the overall spectral shape of the system. This interaction leads to one of the resonances showing a collective enhancement in decay width, which is a direct manifestation of the increased coupling with the continuum.

In this scenario, the excited state |1 predominantly features a d2 configuration, whereas the ground state |2 mainly consists of a p2 configuration. Figure 18 illustrates the evolution of the spectral functions and the corresponding survival probabilities for different energy splittings, ΔE=|Er(1)Er(2)|, of the doublet states.

Fig. 18
Interference between two close-lying 0+ resonances in 6He′ for the three values of the doublet Δ E (in MeV) energy splitting. Left: Spectral functions versus decay energy. The arrow indicates the suppression of the spectral function of |2. Right: Time dependence of the corresponding survival probabilities. The decay widths (in keV) of the doublet (Γ1, Γ2) are (34, 60), (30, 52), and (6, 68) for large, moderate, and small values of Δ E, respectively. The figure is taken from [246]
pic

When the energy splitting Δ E is large, only a minor suppression occurs at the tail of the spectral function for state |2, and both states exhibit comparable decay widths. However, as the states begin to overlap, significant interference effects arise, which dramatically impact the spectral functions of the doublet (Fig. 18e, f). This interference leads to pronounced deviations from the exponential decay regime in survival probabilities.

Specifically, state |1 decays much more rapidly than its intrinsic decay width would suggest, whereas state |2 exhibits a remarkably slow decay. These observations are consistent with the findings of previous studies [253, 255], which discuss how such exponential deviations during the decay process can occur between any near-lying resonances of the same spin-parity. This phenomenon is driven by virtual transitions governed by the scattering continuum and differing orbital angular momentum structures of the doublet states.

This behavior highlights the complex dynamics that can arise from interference effects in quantum decay processes, particularly when closely spaced resonances are involved. The interplay between the initial state configurations, decay channels, and the nature of the continuum coupling leads to non-trivial modifications in decay rates and survival probabilities, underscoring the intricate nature of quantum mechanical decay processes.

4

Summary

In this paper, recent developments in IMSRG were reviewed, focusing on our work on deformed IMSRG and Gamow IMSRG. The developed IMSRG approaches were successfully applied to nuclei which are elongated in shape or exhibit weakly bound or even unbound resonance. The reaction-related GCC method and its extensions to deformed systems and time-dependent approaches are also summarized. Starting with the axially deformed HF reference state, the D-IMSRG enables the IMSRG to compute open-shell nuclei and includes important deformed configurations. The valence-space IMSRG was first developed by Tsukiyama et al. to derive an ab initio shell-model effective interaction in a non-perturbative way. We used this method to investigate the residual neutron-proton interaction δVpn in the upper fp shell with chiral 3NF included. The bifurcation of even-even and odd-odd δVpn values were found experimentally in this region.

Without 3NF, we could not reproduce the bifurcation in this region, which in turn means that 3NF plays a vital role on the behavior of δVpn by enhancing the pn correlations with a stronger T=1 isospin coupling. The G-IMSRG uses the Berggren basis to include effects from continuum coupling and describes the resonance and non-resonant continuum properties of weakly bound and unbound nuclei.

There are still many challenges on the path to developing IMSRG. The IMSRG(2) approximation is computationally efficient and capable of accurately capturing dynamic correlations. However, when treating observables characterized by strong static or collective correlations, such as E2 transition probabilities, IMSRG(2) usually fails to precisely reproduce experimental values. This is typically attributed to the lack of contributions from many-particle many-hole excitations [256] and the large uncertainty of the nuclear force [165]. IMSRG(3) was developed by [257]; however, its computational demands are so immense that it cannot be applied to medium- and heavy-mass nuclei. Several new approximations have been proposed that aim to capture as many of the essential IMSRG(3) correlations as possible while minimizing the computational cost [258, 259] and it is still an open problem. G-IMSRG is a powerful tool to treat weakly bound and unbound nuclei but is currently limited to closed-shell nuclei. New methods that combine G-IMSRG and VS-IMSRG are on the way to broaden the applicability of G-IMSRG to open-shell nuclei.

Notably, methods such as GSM-CC and GCC are dedicated to providing a unified description of the structure, decay, and reactions within open quantum systems. These developments enable a meticulous study of exotic decays in the dripline region. Deformation and continuum effects have been demonstrated to significantly influence the 2p decay process. Additionally, the observed nucleon-nucleon correlations serve as a valuable tool for probing the internal structure of dripline nuclei. These studies enhance our understanding of the complex dynamics and universal properties within open quantum systems.

Meanwhile, the existing framework of GCC remains incomplete at a microscopic level; the description of the core wave function is relatively simplistic, capturing only the collective motions. Thus, advancements towards a more detailed, microscopic framework are anticipated in future developments. Another avenue can involve extending the model to multi-particle decay studies near or beyond the dripline, which have garnered significant interest recently. Their applications to reaction-related problems, such as analyzing cross sections and reaction mechanisms, would be invaluable in providing structural and reactive insights in both theoretical and experimental studies.

References
1E. Epelbaum, H.W. Hammer, U.G. Meißner,

Modern theory of nuclear forces

. Rev. Mod. Phys. 81, 1773-1825 (2009). https://doi.org/10.1103/RevModPhys.81.1773
Baidu ScholarGoogle Scholar
2R. Machleidt, D. Entem,

Chiral effective field theory and nuclear forces

. Phys. Rep. 503, 1-75 (2011). https://doi.org/10.1016/j.physrep.2011.02.001
Baidu ScholarGoogle Scholar
3S. Binder, J. Langhammer, A. Calciet al.,

Ab initio path to heavy nuclei

. Phys. Lett. B 736, 119-123 (2014). https://doi.org/10.1016/j.physletb.2014.07.010
Baidu ScholarGoogle Scholar
4G. Hagen, A. Ekström, C. Forssénet al.,

Neutron and weak-charge distributions of the 48Ca nucleus

. Nat. Phys. 12, 186 (2016). https://doi.org/10.1038/nphys3529
Baidu ScholarGoogle Scholar
5G. Hagen, G.R. Jansen, T. Papenbrock,

Structure of 78Ni from first-principles computations

. Phys. Rev. Lett. 117, 172501 (2016). https://doi.org/10.1103/PhysRevLett.117.172501
Baidu ScholarGoogle Scholar
6B.S. Hu, F.R. Xu, Z.H. Sunet al.,

Ab initio nuclear many-body perturbation calculations in the hartree-fock basis

. Phys. Rev. C 94, 014303 (2016). https://doi.org/10.1103/PhysRevC.94.014303
Baidu ScholarGoogle Scholar
7T.D. Morris, J. Simonis, S.R. Stroberget al.,

Structure of the lightest tin isotopes

. Phys. Rev. Lett. 120, 152503 (2018). https://doi.org/10.1103/PhysRevLett.120.152503
Baidu ScholarGoogle Scholar
8P. Gysbers, G. Hagen, J.D. Holtet al.,

Discrepancy between experimental and theoretical β-decay rates resolved from first principles

. Nat. Phys. 15, 428-431 (2019). https://doi.org/10.1038/s41567-019-0450-7
Baidu ScholarGoogle Scholar
9B.N. Lu, N. Li, S. Elhatisariet al.,

Perturbative quantum monte carlo method for nuclear physics

. Phys. Rev. Lett. 128, 242501 (2022). https://doi.org/10.1103/PhysRevLett.128.242501
Baidu ScholarGoogle Scholar
10B.S. Hu, W.G. Jiang, T. Miyagiet al.,

Ab initio predictions link the neutron skin of 208Pb to nuclear forces

. Nat. Phys. 18, 1196 (2022). https://doi.org/10.1038/s41567-022-01715-8
Baidu ScholarGoogle Scholar
11S. Weinberg,

Nuclear forces from chiral lagrangians

. Phys. Lett. B 251, 288-292 (1990). https://doi.org/10.1016/0370-2693(90)90938-3
Baidu ScholarGoogle Scholar
12S. Weinberg,

Effective chiral lagrangians for nucleon-pion interactions and nuclear forces

. Nucl. Phys. B 363, 3-18 (1991). https://doi.org/10.1016/0550-3213(91)90231-L
Baidu ScholarGoogle Scholar
13C. Ordóñez, U. van Kolck,

Chiral lagrangians and nuclear forces

. Phys. Lett. B 291, 459-464 (1992). https://doi.org/10.1016/0370-2693(92)91404-W
Baidu ScholarGoogle Scholar
14P. Navrátil, V.G. Gueorguiev, J.P. Varyet al.,

Structure of A=10-13 nuclei with two- plus three-nucleon interactions from chiral effective field theory

. Phys. Rev. Lett. 99, 042501 (2007). https://doi.org/10.1103/PhysRevLett.99.042501
Baidu ScholarGoogle Scholar
15T. Otsuka, T. Suzuki, J.D. Holtet al.,

Three-body forces and the limit of oxygen isotopes

. Phys. Rev. Lett. 105, 032501 (2010). https://doi.org/10.1103/PhysRevLett.105.032501
Baidu ScholarGoogle Scholar
16R. Roth, J. Langhammer, A. Calciet al.,

Similarity-transformed chiral NN+3N interactions for the ab initio description of 12C and 16O

. Phys. Rev. Lett. 107, 072501 (2011). https://doi.org/10.1103/PhysRevLett.107.072501
Baidu ScholarGoogle Scholar
17P. Maris, J.P. Vary, P. Navrátilet al.,

Origin of the anomalous long lifetime of 14C

. Phys. Rev. Lett. 106, 202502 (2011). https://doi.org/10.1103/PhysRevLett.106.202502
Baidu ScholarGoogle Scholar
18G. Hagen, M. Hjorth-Jensen, G.R. Jansenet al.,

Continuum effects and three-nucleon forces in neutron-rich oxygen isotopes

. Phys. Rev. Lett. 108, 242501 (2012). https://doi.org/10.1103/PhysRevLett.108.242501
Baidu ScholarGoogle Scholar
19H. Hergert, S. Binder, A. Calciet al.,

Ab initio calculations of even oxygen isotopes with chiral two-plus-three-nucleon interactions

. Phys. Rev. Lett. 110, 242501 (2013). https://doi.org/10.1103/PhysRevLett.110.242501
Baidu ScholarGoogle Scholar
20J.D. Holt, J. Menéndez, A. Schwenk,

Three-body forces and proton-rich nuclei

. Phys. Rev. Lett. 110, 022502 (2013). https://doi.org/10.1103/PhysRevLett.110.022502
Baidu ScholarGoogle Scholar
21S.K. Bogner, H. Hergert, J.D. Holtet al.,

Nonperturbative Shell-Model Interactions from the In-Medium Similarity Renormalization Group

. Phys. Rev. Lett. 113, 142501 (2014). https://doi.org/10.1103/PhysRevLett.113.142501
Baidu ScholarGoogle Scholar
22T. Fukui, L. De Angelis, Y.Z. Maet al.,

Realistic shell-model calculations for p-shell nuclei including contributions of a chiral three-body force

. Phys. Rev. C 98, 044305 (2018). https://doi.org/10.1103/PhysRevC.98.044305
Baidu ScholarGoogle Scholar
23Y. Ma, F. Xu, L. Coraggioet al.,

Chiral three-nucleon force and continuum for dripline nuclei and beyond

. Phys. Lett. B 802, 135257 (2020). https://doi.org/10.1016/j.physletb.2020.135257
Baidu ScholarGoogle Scholar
24K. Hebeler, S.K. Bogner, R.J. Furnstahlet al.,

Improved nuclear matter calculations from chiral low-momentum interactions

. Phys. Rev. C 83, 031301(R) (2011). https://doi.org/10.1103/PhysRevC.83.031301
Baidu ScholarGoogle Scholar
25J. Simonis, S.R. Stroberg, K. Hebeleret al.,

Saturation with chiral interactions and consequences for finite nuclei

. Phys. Rev. C 96, 014303 (2017). https://doi.org/10.1103/PhysRevC.96.014303
Baidu ScholarGoogle Scholar
26W.G. Jiang, A. Ekström, C. Forssénet al.,

Accurate bulk properties of nuclei from A=2 to ∞ from potentials with Δ isobars

. Phys. Rev. C 102, 054301 (2020). https://doi.org/10.1103/PhysRevC.102.054301
Baidu ScholarGoogle Scholar
27V. Somà, P. Navrátil, F. Raimondiet al.,

Novel chiral Hamiltonian and observables in light and medium-mass nuclei

. Phys. Rev. C 101, 014318 (2020). https://doi.org/10.1103/PhysRevC.101.014318
Baidu ScholarGoogle Scholar
28S.R. Stroberg, J.D. Holt, A. Schwenket al.,

Ab initio limits of atomic nuclei

. Phys. Rev. Lett. 126, 022501 (2021). https://doi.org/10.1103/PhysRevLett.126.022501
Baidu ScholarGoogle Scholar
29K. Hebeler,

Three-nucleon forces: Implementation and applications to atomic nuclei and dense matter

. Phys. Rep. 890, 1-116 (2021). https://doi.org/10.1016/j.physrep.2020.08.009
Baidu ScholarGoogle Scholar
30S. Zhang, Z. Cheng, J. Liet al.,

Ab initio gamow shell model with chiral three-nucleon force for 14O isotones

. Chin. Sci. Bull. 67, 4101-4107 (2022). https://doi.org/10.1360/TB-2022-0432
Baidu ScholarGoogle Scholar
31S. Bogner, R. Furnstahl, S. Ramananet al.,

Low-momentum interactions with smooth cutoffs

. Nucl. Phys. A 784, 79-103 (2007). https://doi.org/10.1016/j.nuclphysa.2006.11.123
Baidu ScholarGoogle Scholar
32S. Bogner, R. Furnstahl, P. Mariset al.,

Convergence in the no-core shell model with low-momentum two-nucleon interactions

. Nucl. Phys. A 801, 21-42 (2008). https://doi.org/10.1016/j.nuclphysa.2007.12.008
Baidu ScholarGoogle Scholar
33E.D. Jurgenson, P. Navrátil, R.J. Furnstahl,

Evolution of nuclear many-body forces with the similarity renormalization group

. Phys. Rev. Lett. 103, 082501 (2009). https://doi.org/10.1103/PhysRevLett.103.082501
Baidu ScholarGoogle Scholar
34S.D. Glazek, K.G. Wilson,

Renormalization of hamiltonians

. Phys. Rev. D 48, 5863-5872 (1993). https://doi.org/10.1103/PhysRevD.48.5863
Baidu ScholarGoogle Scholar
35F. Wegner,

Flow-equations for hamiltonians

. Ann. Phys. Lpz 3, 77 (1994). https://doi.org/10.1002/andp.19945060203
Baidu ScholarGoogle Scholar
36K. Tsukiyama, S.K. Bogner, A. Schwenk,

In-Medium Similarity Renormalization Group For Nuclei

. Phys. Rev. Lett. 106, 222502 (2011). https://doi.org/10.1103/PhysRevLett.106.222502
Baidu ScholarGoogle Scholar
37H. Hergert, S. Bogner, T. Morriset al.,

The in-medium similarity renormalization group: A novel ab initio method for nuclei

. Phys. Rep. 621, 165-222 (2016). https://doi.org/10.1016/j.physrep.2015.12.007
Baidu ScholarGoogle Scholar
38S.R. Stroberg, H. Hergert, S.K. Bogneret al.,

Nonempirical Interactions for the Nuclear Shell Model: An Update

. Annu. Rev. Nucl. Part. Sci. 69, 307-362 (2019). https://doi.org/10.1146/annurev-nucl-101917-021120
Baidu ScholarGoogle Scholar
39K. Tsukiyama, S.K. Bogner, A. Schwenk,

In-medium similarity renormalization group for open-shell nuclei

. Phys. Rev. C 85, 061304(R) (2012). https://doi.org/10.1103/PhysRevC.85.061304
Baidu ScholarGoogle Scholar
40S.R. Stroberg, A. Calci, H. Hergertet al.,

Nucleus-Dependent Valence-Space Approach to Nuclear Structure

. Phys. Rev. Lett. 118, 032502 (2017). https://doi.org/10.1103/PhysRevLett.118.032502
Baidu ScholarGoogle Scholar
41G. Hagen, T. Papenbrock, D.J. Deanet al.,

Ab initio coupled-cluster approach to nuclear structure with modern nucleon-nucleon interactions

. Phys. Rev. C 82, 034330 (2010). https://doi.org/10.1103/PhysRevC.82.034330
Baidu ScholarGoogle Scholar
42G. Hagen, T. Papenbrock, M. Hjorth-Jensenet al.,

Coupled-cluster computations of atomic nuclei

. Rep. Prog. Phys. 77, 096302 (2014). https://doi.org/10.1088/0034-4885/77/9/096302
Baidu ScholarGoogle Scholar
43S. Novario, P. Gysbers, J. Engelet al.,

Coupled-cluster calculations of neutrinoless double-β decay in 48Ca

. Phys. Rev. Lett. 126, 182502 (2021). https://doi.org/10.1103/PhysRevLett.126.182502
Baidu ScholarGoogle Scholar
44J.D. Holt, J. Menéndez, J. Simoniset al.,

Three-nucleon forces and spectroscopy of neutron-rich calcium isotopes

. Phys. Rev. C 90, 024312 (2014). https://doi.org/10.1103/PhysRevC.90.024312
Baidu ScholarGoogle Scholar
45A. Tichai, J. Langhammer, S. Binderet al.,

Hartree–Fock many-body perturbation theory for nuclear ground-states

. Phys. Lett. B 756, 283-288 (2016). https://doi.org/10.1016/j.physletb.2016.03.029
Baidu ScholarGoogle Scholar
46W.H. Dickhoff, C. Barbieri,

Self-consistent Green’s function method for nuclei and nuclear matter

. Prog. Part. Nucl. Phys. 52, 377-496 (2004). https://doi.org/10.1016/j.ppnp.2004.02.038
Baidu ScholarGoogle Scholar
47T. Miyagi, S.R. Stroberg, P. Navrátilet al.,

Converged ab initio calculations of heavy nuclei

. Phys. Rev. C 105, 014302 (2022). https://doi.org/10.1103/PhysRevC.105.014302
Baidu ScholarGoogle Scholar
48A. Tichai, P. Arthuis, H. Hergertet al.,

Adg: automated generation and evaluation of many-body diagrams

. Eur. Phys. J. A 58, 2 (2022). https://doi.org/10.1140/epja/s10050-021-00621-6
Baidu ScholarGoogle Scholar
49H. Hergert, S. Binder, A. Calciet al.,

Ab Initio Calculations of Even Oxygen Isotopes with Chiral Two-Plus-Three-Nucleon Interactions

. Phys. Rev. Lett. 110, 242501 (2013). https://doi.org/10.1103/PhysRevLett.110.242501
Baidu ScholarGoogle Scholar
50H. Hergert, S.K. Bogner, T.D. Morriset al.,

Ab initio multireference in-medium similarity renormalization group calculations of even calcium and nickel isotopes

. Phys. Rev. C 90, 041302 (2014). https://doi.org/10.1103/PhysRevC.90.041302
Baidu ScholarGoogle Scholar
51T. Duguet,

Symmetry broken and restored coupled-cluster theory: I. rotational symmetry and angular momentum

. J. Phys. G, 42, 025107 (2014). https://doi.org/10.1088/0954-3899/42/2/025107
Baidu ScholarGoogle Scholar
52J.M. Yao, B. Bally, J. Engelet al.,

Ab initio treatment of collective correlations and the neutrinoless double beta decay of 48Ca

. Phys. Rev. Lett. 124, 232501 (2020). https://doi.org/10.1103/PhysRevLett.124.232501
Baidu ScholarGoogle Scholar
53Q. Yuan, S.Q. Fan, B.S. Huet al.,

Deformed in-medium similarity renormalization group

. Phys. Rev. C 105, L061303 (2022). https://doi.org/10.1103/PhysRevC.105.L061303
Baidu ScholarGoogle Scholar
54N. Michel, W. Nazarewicz, M. Płoszajczaket al.,

Shell model in the complex energy plane

. J. Phys. G 36, 013101 (2009). https://doi.org/10.1088/0954-3899/36/1/013101
Baidu ScholarGoogle Scholar
55T. Berggren,

On the use of resonant states in eigenfunction expansions of scattering and reaction amplitudes

. Nucl. Phys. A 109, 265-287 (1968). https://doi.org/10.1016/0375-9474(68)90593-9
Baidu ScholarGoogle Scholar
56R. Liotta, E. Maglione, N. Sandulescuet al.,

A representation to describe nuclear processes in the continuum

. Phys. Lett. B 367, 1-4 (1996). https://doi.org/10.1016/0370-2693(95)01415-2
Baidu ScholarGoogle Scholar
57B.S. Hu, Q. Wu, Z.H. Sunet al.,

Ab initio gamow in-medium similarity renormalization group with resonance and continuum

. Phys. Rev. C 99, 061302(R) (2019). https://doi.org/10.1103/PhysRevC.99.061302
Baidu ScholarGoogle Scholar
58B. Acharya, C. Ji, D. Phillips,

Implications of a matter-radius measurement for the structure of carbon-22

. Phys. Lett. B 723, 196-200 (2013). https://doi.org/10.1016/j.physletb.2013.04.055
Baidu ScholarGoogle Scholar
59T. Suzuki, T. Otsuka, C. Yuanet al.,

Two-neutron “halo” from the low-energy limit of neutron-neutron interaction: Applications to drip-line nuclei 22C and 24O

. Phys. Lett. B 753, 199-203 (2016). https://doi.org/10.1016/j.physletb.2015.12.001
Baidu ScholarGoogle Scholar
60X.X. Sun, J. Zhao, S.G. Zhou,

Shrunk halo and quenched shell gap at N=16 in 22C: Inversion of sd states and deformation effects

. Phys. Lett. B 785, 530-535 (2018). https://doi.org/10.1016/j.physletb.2018.08.071
Baidu ScholarGoogle Scholar
61S. Elhatisari, D. Lee, G. Rupaket al.,

Ab initio alpha–alpha scattering

. Nature 528, 111-114 (2015).
Baidu ScholarGoogle Scholar
62P. Navrátil, S. Quaglioni, G. Hupinet al.,

Unified ab initio approaches to nuclear structure and reactions

. Phys. Scripta 91, 053002 (2016).
Baidu ScholarGoogle Scholar
63A. Kumar, R. Kanungo, A. Calciet al.,

Nuclear force imprints revealed on the elastic scattering of protons with 10C

. Phys. Rev. Lett. 118, 262502 (2017). https://doi.org/10.1103/PhysRevLett.118.262502
Baidu ScholarGoogle Scholar
64S. Quaglioni, C. Romero-Redondo, P. Navrátil,

Three-cluster dynamics within an ab initio framework

. Phys. Rev. C 88, 034320 (2013). https://doi.org/10.1103/PhysRevC.88.034320
Baidu ScholarGoogle Scholar
65S. Quaglioni, C. Romero-Redondo, P. Navrátil,

Erratum: Three-cluster dynamics within an ab initio framework [phys. rev. c 88, 034320 (2013)]

. Phys. Rev. C 94, 019902 (2016). https://doi.org/10.1103/PhysRevC.94.019902
Baidu ScholarGoogle Scholar
66N. Michel, M. Płoszajczak, Gamow Shell Model, The Unified Theory of Nuclear Structure and Reactions, Vol. 983 of Lecture Notes in Physics, (Springer, Cham, 2021). https://doi.org/10.1007/978-3-030-69356-5
67S.M. Wang, N. Michel, W. Nazarewiczet al.,

Structure and decays of nuclear three-body systems: The Gamow coupled-channel method in Jacobi coordinates

. Phys. Rev. C 96, 044307 (2017). https://doi.org/10.1103/PhysRevC.96.044307
Baidu ScholarGoogle Scholar
68S.M. Wang, W. Nazarewicz,

Puzzling two-proton decay of 67Kr

. Phys. Rev. Lett. 120, 212502 (2018). https://doi.org/10.1103/PhysRevLett.120.212502
Baidu ScholarGoogle Scholar
69N. Michel, W. Nazarewicz, M. Płoszajczak,

Description of the proton-decaying 02+ resonance of the α particle

. Phys. Rev. Lett. 131, 242502 (2023). https://doi.org/10.1103/PhysRevLett.131.242502
Baidu ScholarGoogle Scholar
70S.M. Wang, W. Nazarewicz,

Fermion pair dynamics in open quantum systems

. Phys. Rev. Lett. 126, 142501 (2021). https://doi.org/10.1103/PhysRevLett.126.142501
Baidu ScholarGoogle Scholar
71S.M. Wang, W. Nazarewicz, R.J. Charityet al.,

Nucleon-nucleon correlations in the extreme oxygen isotopes

. J. Phys. G 49, 10LT02 (2022). https://doi.org/10.1088/1361-6471/ac888f
Baidu ScholarGoogle Scholar
72S. Bogner, R. Furnstahl, A. Schwenk,

From low-momentum interactions to nuclear structure

. Prog. Part. Nucl. Phys. 65, 94-147 (2010). https://doi.org/10.1016/j.ppnp.2010.03.001
Baidu ScholarGoogle Scholar
73S.K. Bogner, R.J. Furnstahl, R.J. Perry,

Similarity renormalization group for nucleon-nucleon interactions

. Phys. Rev. C 75, 061001(R) (2007). https://doi.org/10.1103/PhysRevC.75.061001
Baidu ScholarGoogle Scholar
74K. Hebeler,

Momentum-space evolution of chiral three-nucleon forces

. Phys. Rev. C 85, 021002 (2012). https://doi.org/10.1103/PhysRevC.85.021002
Baidu ScholarGoogle Scholar
75R. Roth, S. Binder, K. Vobiget al.,

Medium-mass nuclei with normal-ordered chiral NN+3N interactions

. Phys. Rev. Lett. 109, 052501 (2012). https://doi.org/10.1103/PhysRevLett.109.052501
Baidu ScholarGoogle Scholar
76W. Magnus,

On the exponential solution of differential equations for a linear operator

. Pure Appl. Math 7, 649-673 (1954). https://doi.org/10.1002/cpa.3160070404
Baidu ScholarGoogle Scholar
77T.D. Morris, N.M. Parzuchowski, S.K. Bogner,

Magnus expansion and in-medium similarity renormalization group

. Phys. Rev. C 92, 034331 (2015). https://doi.org/10.1103/PhysRevC.92.034331
Baidu ScholarGoogle Scholar
78Y. Sun, P.M. Walker, F.R. Xuet al.,

Rotation-driven prolate-to-oblate shape phase transition in 190W: A projected shell model study

. Phys. Lett. B 659, 165-169 (2008). https://doi.org/10.1016/j.physletb.2007.10.067
Baidu ScholarGoogle Scholar
79H.L. Liu, F.R. Xu, P.M. Walkeret al.,

Effects of high-order deformation on high-k isomers in superheavy nuclei

. Phys. Rev. C 83, 011303 (2011). https://doi.org/10.1103/PhysRevC.83.011303
Baidu ScholarGoogle Scholar
80T. Dytrych, K.D. Launey, J.P. Draayeret al.,

Collective modes in light nuclei from first principles

. Phys. Rev. Lett. 111, 252501 (2013). https://doi.org/10.1103/PhysRevLett.111.252501
Baidu ScholarGoogle Scholar
81T. Dytrych, K.D. Launey, J.P. Draayeret al.,

Physics of nuclei: Key role of an emergent symmetry

. Phys. Rev. Lett. 124, 042501 (2020). https://doi.org/10.1103/PhysRevLett.124.042501
Baidu ScholarGoogle Scholar
82S.J. Novario, G. Hagen, G.R. Jansenet al.,

Charge radii of exotic neon and magnesium isotopes

. Phys. Rev. C 102, 051303 (2020). https://doi.org/10.1103/PhysRevC.102.051303
Baidu ScholarGoogle Scholar
83G. Hagen, S.J. Novario, Z.H. Sunet al.,

Angular-momentum projection in coupled-cluster theory: Structure of 34Mg

. Phys. Rev. C 105, 064311 (2022). https://doi.org/10.1103/PhysRevC.105.064311
Baidu ScholarGoogle Scholar
84R. Roth, P. Navrátil,

Ab initio study of 40Ca with an importance-truncated no-core shell model

. Phys. Rev. Lett. 99, 092501 (2007). https://doi.org/10.1103/PhysRevLett.99.092501
Baidu ScholarGoogle Scholar
85R. Roth,

Importance truncation for large-scale configuration interaction approaches

. Phys. Rev. C 79, 064324 (2009). https://doi.org/10.1103/PhysRevC.79.064324
Baidu ScholarGoogle Scholar
86M.A. Caprio, P. Maris, J.P. Varyet al.,

Collective rotation from ab initio theory

. Int. J. Mod. Phys. E 24, 1541002 (2015). https://doi.org/10.1142/s0218301315410025
Baidu ScholarGoogle Scholar
87T. Abe, P. Maris, T. Otsukaet al.,

Ground-state properties of light 4n self-conjugate nuclei in ab initio no-core monte carlo shell model calculations with nonlocal NN interactions

. Phys. Rev. C 104, 054315 (2021). https://doi.org/10.1103/PhysRevC.104.054315
Baidu ScholarGoogle Scholar
88M.G. Mayer, J.H.D. Jensen, Elementary Theory of Nuclear Shell Structure, (John Wiley & Sons, New York, 1955)
89B.A. Brown, B.H. Wildenthal,

Status of the nuclear shell model

. Annu. Rev. Nucl. Part. Sci. 38, 29-66 (1988). https://doi.org/10.1146/annurev.ns.38.120188.000333
Baidu ScholarGoogle Scholar
90T. Otsuka, M. Honma, T. Mizusakiet al.,

Monte Carlo shell model for atomic nuclei

. Prog. Part. Nucl. Phys. 47, 319-400 (2001). https://doi.org/10.1016/S0146-6410(01)00157-0
Baidu ScholarGoogle Scholar
91E. Caurier, G. Martínez-Pinedo, F. Nowackiet al.,

The shell model as a unified view of nuclear structure

. Rev. Mod. Phys. 77, 427-488 (2005). https://doi.org/10.1103/RevModPhys.77.427
Baidu ScholarGoogle Scholar
92B.A. Brown,

The nuclear shell model towards the drip lines

. Prog. Part. Nucl. Phys. 47, 517-599 (2001). https://doi.org/10.1016/S0146-6410(01)00159-4
Baidu ScholarGoogle Scholar
93P. Navrátil, J.P. Vary, B.R. Barrett,

Properties of 12C in the ab initio nuclear shell model

. Phys. Rev. Lett. 84, 5728-5731 (2000). https://doi.org/10.1103/PhysRevLett.84.5728
Baidu ScholarGoogle Scholar
94P. Navrátil, J.P. Vary, B.R. Barrett,

Large-basis ab initio no-core shell model and its application to 12C

. Phys. Rev. C 62, 054311 (2000). https://doi.org/10.1103/PhysRevC.62.054311
Baidu ScholarGoogle Scholar
95N. Shimizu, T. Mizusaki, Y. Utsunoet al.,

Thick-restart block lanczos method for large-scale shell-model calculations

. Comput. Phys. Commun. 244, 372-384 (2019). https://doi.org/10.1016/j.cpc.2019.06.011
Baidu ScholarGoogle Scholar
96J. Okołowicz, M. Płoszajczak, I. Rotter,

Dynamics of quantum systems embedded in a continuum

. Physics Reports 374, 271-383 (2003). https://doi.org/10.1016/S0370-1573(02)00366-6
Baidu ScholarGoogle Scholar
97N. Michel, W. Nazarewicz, M. Płoszajczaket al.,

Shell model in the complex energy plane

. J. Phys. G: Nucl. Part. Phys. 36, 013101 (2008). https://doi.org/10.1088/0954-3899/36/1/013101
Baidu ScholarGoogle Scholar
98I. Tanihata, H. Hamagaki, O. Hashimotoet al.,

Measurements of interaction cross sections and nuclear radii in the light p-shell region

. Phys. Rev. Lett. 55, 2676-2679 (1985). https://doi.org/10.1103/PhysRevLett.55.2676
Baidu ScholarGoogle Scholar
99A.S. Jensen, K. Riisager, D.V. Fedorovet al.,

Structure and reactions of quantum halos

. Rev. Mod. Phys. 76, 215-261 (2004). https://doi.org/10.1103/RevModPhys.76.215
Baidu ScholarGoogle Scholar
100R. Id Betan, R.J. Liotta, N. Sandulescuet al.,

Two-particle resonant states in a many-body mean field

. Phys. Rev. Lett. 89, 042501 (2002). https://doi.org/10.1103/PhysRevLett.89.042501
Baidu ScholarGoogle Scholar
101N. Michel, W. Nazarewicz, M. Płoszajczaket al.,

Gamow shell model description of neutron-rich nuclei

. Phys. Rev. Lett. 89, 042502 (2002). https://doi.org/10.1103/PhysRevLett.89.042502
Baidu ScholarGoogle Scholar
102R. Kanungo, A. Sanetullaev, J. Tanakaet al.,

Evidence of soft dipole resonance in 11Li with isoscalar character

. Phys. Rev. Lett. 114, 192502 (2015). https://doi.org/10.1103/PhysRevLett.114.192502
Baidu ScholarGoogle Scholar
103K. Fossez, J. Rotureau, N. Michelet al.,

Single-particle and collective motion in unbound deformed 39Mg

. Phys. Rev. C 94, 054302 (2016). https://doi.org/10.1103/PhysRevC.94.054302
Baidu ScholarGoogle Scholar
104G. Colò,

A novel way to study the nuclear collective excitations

. Nucl. Sci. Tech. 34, 189 (2023). https://doi.org/10.1007/s41365-023-01343-8
Baidu ScholarGoogle Scholar
105Z. Sun, Q. Wu, Z. Zhaoet al.,

Resonance and continuum gamow shell model with realistic nuclear forces

. Phys. Lett. B 769, 227-232 (2017). https://doi.org/10.1016/j.physletb.2017.03.054
Baidu ScholarGoogle Scholar
106J.G. Li, N. Michel, B.S. Huet al.,

Ab initio no-core gamow shell-model calculations of multineutron systems

. Phys. Rev. C 100, 054313 (2019). https://doi.org/10.1103/PhysRevC.100.054313
Baidu ScholarGoogle Scholar
107B. Hu, Q. Wu, J. Liet al.,

An ab-initio gamow shell model approach with a core

. Phys. Lett. B 802, 135206 (2020). https://doi.org/10.1016/j.physletb.2020.135206
Baidu ScholarGoogle Scholar
108Y. Ma, F. Xu, N. Michelet al.,

Continuum and three-nucleon force in borromean system: The 17ne case

. Phys. Lett. B 808, 135673 (2020). https://doi.org/10.1016/j.physletb.2020.135673
Baidu ScholarGoogle Scholar
109J.G. Li, N. Michel, W. Zuoet al.,

Resonances of A=4 T=1 isospin triplet states within the ab initio no-core gamow shell model

. Phys. Rev. C 104, 024319 (2021). https://doi.org/10.1103/PhysRevC.104.024319
Baidu ScholarGoogle Scholar
110J.G. Li, N. Michel, W. Zuoet al.,

Unbound spectra of neutron-rich oxygen isotopes predicted by the gamow shell model

. Phys. Rev. C 103, 034305 (2021). https://doi.org/10.1103/PhysRevC.103.034305
Baidu ScholarGoogle Scholar
111Y.F. Geng, J.G. Li, Y.Z. Maet al.,

Excitation spectra of the heaviest carbon isotopes investigated within the cd-bonn gamow shell model

. Phys. Rev. C 106, 024304 (2022). https://doi.org/10.1103/PhysRevC.106.024304
Baidu ScholarGoogle Scholar
112S. Zhang, Y. Ma, J. Liet al.,

The roles of three-nucleon force and continuum coupling in mirror symmetry breaking of oxygen mass region

. Phys. Lett. B 827, 136958 (2022). https://doi.org/10.1016/j.physletb.2022.136958
Baidu ScholarGoogle Scholar
113S. Zhang, F.R. Xu, J.G. Liet al.,

Ab initio descriptions of A=16 mirror nuclei with resonance and continuum coupling

. Phys. Rev. C 108, 064316 (2023). https://doi.org/10.1103/PhysRevC.108.064316
Baidu ScholarGoogle Scholar
114Z.C. Xu, S. Zhang, J.G. Liet al.,

Complex valence-space effective operators for observables: The gamow-teller transition

. Phys. Rev. C 108, L031301 (2023). https://doi.org/10.1103/PhysRevC.108.L031301
Baidu ScholarGoogle Scholar
115S. Zhang, Y.F. Geng, F.R. Xu,

Ab initio gamow shell-model calculations for dripline nuclei

. Nucl. Tech. 46, 121-128 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080012
Baidu ScholarGoogle Scholar
116G. Hagen, D. Dean, M. Hjorth-Jensenet al.,

Complex coupled-cluster approach to an ab-initio description of open quantum systems

. Phys. Lett. B 656, 169-173 (2007). https://doi.org/10.1016/j.physletb.2007.07.072
Baidu ScholarGoogle Scholar
117S.M. Wang, W. Nazarewicz,

Puzzling two-proton decay of 67Kr

. Phys. Rev. Lett. 120, 212502 (2018). https://doi.org/10.1103/PhysRevLett.120.212502
Baidu ScholarGoogle Scholar
118G. Hagen, M. Hjorth-Jensen, N. Michel,

Gamow shell model and realistic nucleon-nucleon interactions

. Phys. Rev. C 73, 064307 (2006). https://doi.org/10.1103/PhysRevC.73.064307
Baidu ScholarGoogle Scholar
119D.J. Rowe,

Equations-of-motion method and the extended shell model

. Rev. Mod. Phys. 40, 153-166 (1968). https://doi.org/10.1103/RevModPhys.40.153
Baidu ScholarGoogle Scholar
120N.M. Parzuchowski, T.D. Morris, S.K. Bogner,

Ab initio excited states from the in-medium similarity renormalization group

. Phys. Rev. C 95, 044304 (2017). https://doi.org/10.1103/PhysRevC.95.044304
Baidu ScholarGoogle Scholar
121M. Pfützner, I. Mukha, S.M. Wang,

Two-proton emission and related phenomena

. Prog. Part. Nucl. Phys. 123, 104050 (2023). https://doi.org/10.1016/j.ppnp.2023.104050
Baidu ScholarGoogle Scholar
122L. Zhou, D.Q. Fang, S.M. Wanget al.,

Recent progress in two-proton radioactivity

. Nucl. Sci. Tech. 33, 105 (2022). https://doi.org/10.1007/s41365-022-01091-1
Baidu ScholarGoogle Scholar
123S. Saito,

Interaction between clusters and pauli principle

. Prog. Theor. Phys. 41, 705 (1969).
Baidu ScholarGoogle Scholar
124V. Kukulin, V. Pomerantsev,

The orthogonal projection method in the scattering theory

. Ann. Phys. (NY) 111, 330 (1978).
Baidu ScholarGoogle Scholar
125P. Descouvemont, C. Daniel, D. Baye,

Three-body systems with lagrange-mesh techniques in hyperspherical coordinates

. Phys. Rev. C 67, 044309 (2003). https://doi.org/10.1103/PhysRevC.67.044309
Baidu ScholarGoogle Scholar
126I.J. Thompson, B.V. Danilin, V.D. Efroset al.,

Pauli blocking in three-body models of halo nuclei

. Phys. Rev. C 61, 024318 (2000). https://doi.org/10.1103/PhysRevC.61.024318
Baidu ScholarGoogle Scholar
127I.J. Thompson, F.M. Nunes, B.V. Danilin,

FaCE: a tool for three body Faddeev calculations with core excitation

. Comput. Phys. Commun. 161, 87-107 (2004). https://doi.org/10.1016/j.cpc.2004.03.007
Baidu ScholarGoogle Scholar
128B. Gyarmati, T. Vertse,

On the normalization of Gamow functions

. Nucl. Phys. A 160, 523-528 (1971). https://doi.org/10.1016/0375-9474(71)90095-9
Baidu ScholarGoogle Scholar
129N. Michel, W. Nazarewicz, M. Płoszajczaket al.,

Gamow shell model description of weakly bound nuclei and unbound nuclear states

. Phys. Rev. C 67, 054311 (2003). https://doi.org/10.1103/PhysRevC.67.054311
Baidu ScholarGoogle Scholar
130N. Michel,

Numerical treatment of the long-range Coulomb potential with Berggren bases

. Phys. Rev. C 83, 034325 (2011). https://doi.org/10.1103/PhysRevC.83.034325
Baidu ScholarGoogle Scholar
131E.B. Huo, K.R. Li, X.Y. Quet al.,

Continuum skyrme hartree–fock–bogoliubov theory with green’s function method for neutron-rich ca, ni, zr, and sn isotopes

. Nucl. Sci. Tech. 34, 105 (2023). https://doi.org/10.1007/s41365-023-01261-9
Baidu ScholarGoogle Scholar
132S.Z. Xu, S.S. Zhang, X.Q. Jianget al.,

The complex momentum representation approach and its application to low-lying resonances in 17O and 29,31F

. Nucl. Sci. Tech. 34, 5 (2023). https://doi.org/10.1007/s41365-022-01159-y
Baidu ScholarGoogle Scholar
133T. Berggren,

On the use of resonant states in eigenfunction expansions of scattering and reaction amplitudes

. Nucl. Phys. A 109, 265-287 (1968). https://doi.org/10.1016/0375-9474(68)90593-9
Baidu ScholarGoogle Scholar
134K. Hagino, N. Rowley, A. Kruppa,

A program for coupled-channel calculations with all order couplings for heavy-ion fusion reactions

. Comput. Phys. Commun. 123, 143-152 (1999). https://doi.org/10.1016/S0010-4655(99)00243-X
Baidu ScholarGoogle Scholar
135K. Hagino,

Role of dynamical particle-vibration coupling in reconciliation of the d3/2 puzzle for spherical proton emitters

. Phys. Rev. C 64, 041304 (2001). https://doi.org/10.1103/PhysRevC.64.041304
Baidu ScholarGoogle Scholar
136B. Barmore, A.T. Kruppa, W. Nazarewiczet al.,

Theoretical description of deformed proton emitters: Nonadiabatic coupled-channel method

. Phys. Rev. C 62, 054315 (2000). https://doi.org/10.1103/PhysRevC.62.054315
Baidu ScholarGoogle Scholar
137A.T. Kruppa, W. Nazarewicz,

Gamow and r-matrix approach to proton emitting nuclei

. Phys. Rev. C 69, 054311 (2004). https://doi.org/10.1103/PhysRevC.69.054311
Baidu ScholarGoogle Scholar
138J. Humblet, L. Rosenfeld,

Theory of nuclear reactions I

. Resonant states and collision matrix. Nucl. Phys. 26, 529 (1961).
Baidu ScholarGoogle Scholar
139L.V. Grigorenko, R.C. Johnson, I.G. Mukhaet al.,

Theory of two-proton radioactivity with application to 19Mg and 48Ni

. Phys. Rev. Lett. 85, 22-25 (2000). https://doi.org/10.1103/PhysRevLett.85.22
Baidu ScholarGoogle Scholar
140L.V. Grigorenko, M.V. Zhukov,

Two-proton radioactivity and three-body decay. iii. integral formulas for decay widths in a simplified semianalytical approach

. Phys. Rev. C 76, 014008 (2007). https://doi.org/10.1103/PhysRevC.76.014008
Baidu ScholarGoogle Scholar
141L.V. Grigorenko, T.D. Wiser, K. Mierniket al.,

Complete correlation studies of two-proton decays: 6Be and 45Fe

. Phys. Lett. B 677, 30-35 (2009). https://doi.org/10.1016/j.physletb.2009.04.085
Baidu ScholarGoogle Scholar
142P. Descouvemont, E. Tursunov, D. Baye,

Three-body continuum states on a lagrange mesh

. Nucl. Phys. A 765, 370-389 (2006). https://doi.org/10.1016/j.nuclphysa.2005.11.010
Baidu ScholarGoogle Scholar
143V. Vasilevsky, A.V. Nesterov, F. Arickxet al.,

Algebraic model for scattering in three-s-cluster systems. i. theoretical background

. Phys. Rev. C 63, 034606 (2001). https://doi.org/10.1103/PhysRevC.63.034606
Baidu ScholarGoogle Scholar
144H. Esbensen, C.N. Davids,

Coupled-channels treatment of deformed proton emitters

. Phys. Rev. C 63, 014315 (2000). https://doi.org/10.1103/PhysRevC.63.014315
Baidu ScholarGoogle Scholar
145A. Volya,

Computational approaches to many-body dynamics of unstable nuclear systems.

, in Proceedings of the International Conference ‘Nuclear Theory in the Supercomputing Era’, Khabarovsk, Russia (2014). arXiv:1412.6335
Baidu ScholarGoogle Scholar
146M. Peshkin, A. Volya, V. Zelevinsky,

Non-exponential and oscillatory decays in quantum mechanics

. Europhys. Lett. 107, 40001 (2014). https://doi.org/10.1209/0295-5075/107/40001
Baidu ScholarGoogle Scholar
147M. Bender, R. Bernard, G. Bertschet al.,

Future of nuclear fission theory

. J. Phys. G 47, 113002 (2020). https://doi.org/10.1088/1361-6471/abab4f
Baidu ScholarGoogle Scholar
148C.A. Bertulani, M.S. Hussein, G. Verde,

Blurred femtoscopy in two-proton decay

. Phys. Lett. B 666, 86-90 (2008). https://doi.org/10.1016/j.physletb.2008.06.062
Baidu ScholarGoogle Scholar
149T. Oishi, K. Hagino, H. Sagawa,

Role of diproton correlation in two-proton-emission decay of the 6Be nucleus

. Phys. Rev. C 90, 034303 (2014). https://doi.org/10.1103/PhysRevC.90.034303
Baidu ScholarGoogle Scholar
150T. Oishi, M. Kortelainen, A. Pastore,

Dependence of two-proton radioactivity on nuclear pairing models

. Phys. Rev. C 96, 044327 (2017). https://doi.org/10.1103/PhysRevC.96.044327
Baidu ScholarGoogle Scholar
151A.I. Baz’, Y.B. Zel’dovich, A.M. Perelomov, Scattering, reactions and decay in nonrelativistic quantum mechanics, (Israel Program for Scientific Translation, Jerusalem, 1969)
152T. Ikegami, S. Iwata,

Spectral density calculation by using the Chebyshev expansion

. J. Comput. Chem. 23, 310-318 (2002). https://doi.org/10.1002/jcc.10010
Baidu ScholarGoogle Scholar
153A. Volya,

Time-dependent approach to the continuum shell model

. Phys. Rev. C 79, 044308 (2009). https://doi.org/10.1103/PhysRevC.79.044308
Baidu ScholarGoogle Scholar
154Y.L. Loh, S.N. Taraskin, S.R. Elliott,

Fast Chebyshev-polynomial method for simulating the time evolution of linear dynamical systems

. Phys. Rev. E 63, 056706 (2001). https://doi.org/10.1103/PhysRevE.63.056706
Baidu ScholarGoogle Scholar
155M. Wang, W. Huang, F. Kondevet al.,

The AME 2020 atomic mass evaluation (II)

. tables, graphs and references*. Chin. Phys. C 45, 030003 (2021). https://doi.org/10.1088/1674-1137/abddaf
Baidu ScholarGoogle Scholar
156A. Ekström, G. Baardsen, C. Forssénet al.,

Optimized chiral nucleon-nucleon interaction at next-to-next-to-leading order

. Phys. Rev. Lett. 110, 192502 (2013). https://doi.org/10.1103/PhysRevLett.110.192502
Baidu ScholarGoogle Scholar
157R. Kanungo, W. Horiuchi, G. Hagenet al.,

Proton distribution radii of 12-19C illuminate features of neutron halos

. Phys. Rev. Lett. 117, 102501 (2016). https://doi.org/10.1103/PhysRevLett.117.102501
Baidu ScholarGoogle Scholar
158J. Henderson, G. Hackman, P. Ruotsalainenet al.,

Testing microscopically derived descriptions of nuclear collectivity: Coulomb excitation of 22Mg

. Phys. Lett. B 782, 468-473 (2018). https://doi.org/10.1016/j.physletb.2018.05.064
Baidu ScholarGoogle Scholar
159J. Henderson, G. Hackman, P. Ruotsalainenet al.,

Coulomb excitation of the |Tz|=12, A=23 mirror pair

. Phys. Rev. C 105, 034332 (2022). https://doi.org/10.1103/PhysRevC.105.034332
Baidu ScholarGoogle Scholar
160G. Hagen, M. Hjorth-Jensen, G.R. Jansenet al.,

Emergent properties of nuclei from ab initio coupled-cluster calculations*

. Physica Scripta 91, 063006 (2016). https://doi.org/10.1088/0031-8949/91/6/063006
Baidu ScholarGoogle Scholar
161A. Poves, J. Retamosa,

The onset of deformation at the N = 20 neutron shell closure far from stability

. Phys. Lett. B 184, 311-315 (1987). https://doi.org/10.1016/0370-2693(87)90171-7
Baidu ScholarGoogle Scholar
162E.K. Warburton, J.A. Becker, B.A. Brown,

Mass systematics for A=29-44 nuclei: The deformed A~32 region

. Phys. Rev. C 41, 1147-1166 (1990). https://doi.org/10.1103/PhysRevC.41.1147
Baidu ScholarGoogle Scholar
163T. Miyagi, S.R. Stroberg, J.D. Holtet al.,

Ab initio multishell valence-space hamiltonians and the island of inversion

. Phys. Rev. C 102, 034320 (2020). https://doi.org/10.1103/PhysRevC.102.034320
Baidu ScholarGoogle Scholar
164A. Krieger, W. Nörtershäuser, C. Geppertet al.,

Frequency-comb referenced collinear laser spectroscopy of Be+ for nuclear structure investigations and many-body qed tests

. Appl. Phys. B 123, 15 (2016). https://doi.org/10.1007/s00340-016-6579-5
Baidu ScholarGoogle Scholar
165V. Lapoux, V. Somà, C. Barbieriet al.,

Radii and binding energies in oxygen isotopes: A challenge for nuclear forces

. Phys. Rev. Lett. 117, 052501 (2016). https://doi.org/10.1103/PhysRevLett.117.052501
Baidu ScholarGoogle Scholar
166B. Ohayon, H. Rahangdale, A.J. Geddeset al.,

Isotope shifts in 20,22Ne: Precision measurements and global analysis in the framework of intermediate coupling

. Phys. Rev. A 99, 042503 (2019). https://doi.org/10.1103/PhysRevA.99.042503
Baidu ScholarGoogle Scholar
167D.T. Yordanov, M.L. Bissell, K. Blaumet al.,

Nuclear charge radii of 21-32Mg

. Phys. Rev. Lett. 108, 042504 (2012). https://doi.org/10.1103/PhysRevLett.108.042504
Baidu ScholarGoogle Scholar
168J. Hoppe, C. Drischler, K. Hebeleret al.,

Probing chiral interactions up to next-to-next-to-next-to-leading order in medium-mass nuclei

. Phys. Rev. C 100, 024318 (2019). https://doi.org/10.1103/PhysRevC.100.024318
Baidu ScholarGoogle Scholar
169M. Wang, Y.H. Zhang, X. Zhouet al.,

Mass measurement of upper fp-shell N=Z-2 and N=Z-1 nuclei and the importance of three-nucleon force along the N=Z line

. Phys. Rev. Lett. 130, 192501 (2023). https://doi.org/10.1103/PhysRevLett.130.192501
Baidu ScholarGoogle Scholar
170K. Blaum,

High-accuracy mass spectrometry with stored ions

. Phys. Rep. 425, 1-78 (2006). https://doi.org/10.1016/j.physrep.2005.10.011
Baidu ScholarGoogle Scholar
171T. Yamaguchi, H. Koura, Y. Litvinovet al.,

Masses of exotic nuclei

. Prog. Part. Nucl. Phys. 120, 103882 (2021). https://doi.org/10.1016/j.ppnp.2021.103882
Baidu ScholarGoogle Scholar
172J.Y. Zhang, R. Casten, D. Brenner,

Empirical proton-neutron interaction energies. linearity and saturation phenomena

. Phys. Lett. B 227, 1-5 (1989). https://doi.org/10.1016/0370-2693(89)91273-2
Baidu ScholarGoogle Scholar
173D. Brenner, C. Wesselborg, R. Castenet al.,

Empirical p-n interactions: global trends, configuration sensitivity and N=Z enhancements

. Phys. Lett. B 243, 1-6 (1990). https://doi.org/10.1016/0370-2693(90)90945-3
Baidu ScholarGoogle Scholar
174P. Van Isacker, D.D. Warner, D.S. Brenner,

Test of wigner’s spin-isospin symmetry from double binding energy differences

. Phys. Rev. Lett. 74, 4607-4610 (1995). https://doi.org/10.1103/PhysRevLett.74.4607
Baidu ScholarGoogle Scholar
175C.X. Yuan, C. Qi, F.R. Xuet al.,

Mirror energy difference and the structure of loosely bound proton-rich nuclei around A=20

. Phys. Rev. C 89, 044327 (2014). https://doi.org/10.1103/PhysRevC.89.044327
Baidu ScholarGoogle Scholar
176P. Schury, C. Bachelet, M. Blocket al.,

Precision mass measurements of rare isotopes near N=Z=33 produced by fast beam fragmentation

. Phys. Rev. C 75, 055801 (2007). https://doi.org/10.1103/PhysRevC.75.055801
Baidu ScholarGoogle Scholar
177I. Mardor, S.A.S. Andrés, T. Dickelet al.,

Mass measurements of as, se, and br nuclei, and their implication on the proton-neutron interaction strength toward the N=Z line

. Phys. Rev. C 103, 034319 (2021). https://doi.org/10.1103/PhysRevC.103.034319
Baidu ScholarGoogle Scholar
178D.S. Brenner, R.B. Cakirli, R.F. Casten,

Valence proton-neutron interactions throughout the mass surface

. Phys. Rev. C 73, 034315 (2006). https://doi.org/10.1103/PhysRevC.73.034315
Baidu ScholarGoogle Scholar
179J. Simonis, S.R. Stroberg, K. Hebeleret al.,

Saturation with chiral interactions and consequences for finite nuclei

. Phys. Rev. C 96, 014303 (2017). https://doi.org/10.1103/PhysRevC.96.014303
Baidu ScholarGoogle Scholar
180A.T. Gallant, J.C. Bale, T. Brunneret al.,

New precision mass measurements of neutron-rich calcium and potassium isotopes and three-nucleon forces

. Phys. Rev. Lett. 109, 032506 (2012). https://doi.org/10.1103/PhysRevLett.109.032506
Baidu ScholarGoogle Scholar
181K. Tanaka, T. Yamaguchi, T. Suzukiet al.,

Observation of a large reaction cross section in the drip-line nucleus 22C

. Phys. Rev. Lett. 104, 062701 (2010). https://doi.org/10.1103/PhysRevLett.104.062701
Baidu ScholarGoogle Scholar
182N. Kobayashi, T. Nakamura, J.A. Tostevinet al.,

One- and two-neutron removal reactions from the most neutron-rich carbon isotopes

. Phys. Rev. C 86, 054604 (2012). https://doi.org/10.1103/PhysRevC.86.054604
Baidu ScholarGoogle Scholar
183L. Gaudefroy, W. Mittig, N.A. Orret al.,

Direct mass measurements of 19B, 22C, 29F, 31Ne, 34Na and other light exotic nuclei

. Phys. Rev. Lett. 109, 202503 (2012). https://doi.org/10.1103/PhysRevLett.109.202503
Baidu ScholarGoogle Scholar
184Y. Togano, T. Nakamura, Y. Kondoet al.,

Interaction cross section study of the two-neutron halo nucleus 22C

. Phys. Lett. B 761, 412-418 (2016). https://doi.org/10.1016/j.physletb.2016.08.062
Baidu ScholarGoogle Scholar
185T. Heng, J.P. Vary, P. Maris,

Ab initio no-core properties of 7Li and 7Be with the jisp16 and chiral NNLOopt interactions

. Phys. Rev. C 95, 014306 (2017). https://doi.org/10.1103/PhysRevC.95.014306
Baidu ScholarGoogle Scholar
186A. Ekström, G.R. Jansen, K.A. Wendtet al.,

Accurate nuclear radii and binding energies from a chiral interaction

. Phys. Rev. C 91, 051301 (2015). https://doi.org/10.1103/PhysRevC.91.051301
Baidu ScholarGoogle Scholar
187G. Hagen, T. Papenbrock, D.J. Dean,

Solution of the center-of-mass problem in nuclear structure calculations

. Phys. Rev. Lett. 103, 062503 (2009). https://doi.org/10.1103/PhysRevLett.103.062503
Baidu ScholarGoogle Scholar
188C. Hoffman, T. Baumann, D. Bazinet al.,

Evidence for a doubly magic 24O

. Phys. Lett. B 672, 17-21 (2009). https://doi.org/10.1016/j.physletb.2008.12.066
Baidu ScholarGoogle Scholar
189C.R. Hoffman, T. Baumann, J. Brownet al.,

Observation of a two-neutron cascade from a resonance in 24O

. Phys. Rev. C 83, 031303 (2011). https://doi.org/10.1103/PhysRevC.83.031303
Baidu ScholarGoogle Scholar
190T. Nagahisa, W. Horiuchi,

Examination of the 22C radius determination with interaction cross sections

. Phys. Rev. C 97, 054614 (2018). https://doi.org/10.1103/PhysRevC.97.054614
Baidu ScholarGoogle Scholar
191L. Coraggio, A. Covello, A. Garganoet al.,

Shell-model calculations for neutron-rich carbon isotopes with a chiral nucleon-nucleon potential

. Phys. Rev. C 81, 064303 (2010). https://doi.org/10.1103/PhysRevC.81.064303
Baidu ScholarGoogle Scholar
192G.R. Jansen, J. Engel, G. Hagenet al.,

Ab initio coupled-cluster effective interactions for the shell model: Application to neutron-rich oxygen and carbon isotopes

. Phys. Rev. Lett. 113, 142502 (2014). https://doi.org/10.1103/PhysRevLett.113.142502
Baidu ScholarGoogle Scholar
193Y.F. Gao, B.S. Cai, C.X. Yuan,

Investigation of β--decay half-life and delayed neutron emission with uncertainty analysis

. Nucl. Sci. Tech. 34, 9 (2023). https://doi.org/10.1007/s41365-022-01153-4
Baidu ScholarGoogle Scholar
194J.Z. Han, S. Xu, A. Jaliliet al.,

Investigation of the level spectra of nuclei in the northeast region of doubly magic 40Ca with intruder orbit g9/2

. Nucl. Sci. Tech. 34, 85 (2023). https://doi.org/10.1007/s41365-023-01243-x
Baidu ScholarGoogle Scholar
195A.T. Kruppa, B. Barmore, W. Nazarewiczet al.,

Fine structure in the decay of deformed proton emitters: Nonadiabatic approach

. Phys. Rev. Lett. 84, 4549-4552 (2000). https://doi.org/10.1103/PhysRevLett.84.4549
Baidu ScholarGoogle Scholar
196C.N. Davids, H. Esbensen,

Particle-vibration coupling in proton decay of near-spherical nuclei

. Phys. Rev. C 64, 034317 (2001). https://doi.org/10.1103/PhysRevC.64.034317
Baidu ScholarGoogle Scholar
197C.N. Davids, H. Esbensen,

Decay rate of triaxially deformed proton emitters

. Phys. Rev. C 69, 034314 (2004). https://doi.org/10.1103/PhysRevC.69.034314
Baidu ScholarGoogle Scholar
198G. Fiorin, E. Maglione, L.S. Ferreira,

Theoretical description of deformed proton emitters: Nonadiabatic quasiparticle method

. Phys. Rev. C 67, 054302 (2003). https://doi.org/10.1103/PhysRevC.67.054302
Baidu ScholarGoogle Scholar
199P. Arumugam, E. Maglione, L.S. Ferreira,

Nonadiabatic quasiparticle description of triaxially deformed proton emitters

. Phys. Rev. C 76, 044311 (2007). https://doi.org/10.1103/PhysRevC.76.044311
Baidu ScholarGoogle Scholar
200L.V. Grigorenko, M.V. Zhukov,

Two-proton radioactivity and three-body decay. II. Exploratory studies of lifetimes and correlations

. Phys. Rev. C 68, 054005 (2003). https://doi.org/10.1103/PhysRevC.68.054005
Baidu ScholarGoogle Scholar
201M. Gonçalves, N. Teruya, O. Tavareset al.,

Two-proton emission half-lives in the effective liquid drop model

. Phys. Lett. B 774, 14-19 (2017). https://doi.org/10.1016/j.physletb.2017.09.032
Baidu ScholarGoogle Scholar
202W. Nazarewicz, J. Dudek, R. Bengtssonet al.,

Microscopic study of the high-spin behaviour in selected A=80 nuclei

. Nucl. Phys. A 435, 397-447 (1985). https://doi.org/10.1016/0375-9474(85)90471-3
Baidu ScholarGoogle Scholar
203M. Yamagami, K. Matsuyanagi, M. Matsuo,

Symmetry-unrestricted skyrme-hartree-fock-bogoliubov calculations for exotic shapes in N=Z nuclei from 64Ge to 84Mo

. Nucl. Phys. A 693, 579-602 (2001). https://doi.org/10.1016/S0375-9474(01)00918-6
Baidu ScholarGoogle Scholar
204K. Kaneko, M. Hasegawa, T. Mizusaki,

Shape transition and oblate-prolate coexistence in N=Z fpg-shell nuclei

. Phys. Rev. C 70, 051301 (2004). https://doi.org/10.1103/PhysRevC.70.051301
Baidu ScholarGoogle Scholar
205B. Pritychenko, M. Birch, B. Singhet al.,

Tables of E2 transition probabilities from the first 2+ states in even-even nuclei

. At. Data Nucl. Data Tables 107, 1-139 (2016). https://doi.org/10.1016/j.adt.2015.10.001
Baidu ScholarGoogle Scholar
206Y. Aboussir, J. Pearson, A. Duttaet al.,

Nuclear mass formula via an approximation to the hartree-fock method

. At. Data Nucl. Data Tables 61, 127-176 (1995). https://doi.org/10.1016/S0092-640X(95)90014-4
Baidu ScholarGoogle Scholar
207

Mass Explorer

, http://massexplorer.frib.msu.edu/
Baidu ScholarGoogle Scholar
208P. Möller, A. Sierk, T. Ichikawaet al.,

Nuclear ground-state masses and deformations: Frdm(2012)

. At. Data Nucl. Data Tables 109-110, 1-204 (2016). https://doi.org/10.1016/j.adt.2015.10.002
Baidu ScholarGoogle Scholar
209T. Goigoux, P. Ascher, B. Blanket al.,

Two-proton radioactivity of 67Kr

. Phys. Rev. Lett. 117, 162501 (2016). https://doi.org/10.1103/PhysRevLett.117.162501
Baidu ScholarGoogle Scholar
210O.V. Bochkarev, L.V. Chulkov, A.A. Korsheninniicovet al.,

Democratic decay of 6Be states

. Nucl. Phys. A 505, 215-240 (1989). https://doi.org/10.1016/0375-9474(89)90371-0
Baidu ScholarGoogle Scholar
211M. Pfützner, M. Karny, L.V. Grigorenkoet al.,

Radioactive decays at limits of nuclear stability

. Rev. Mod. Phys. 84, 567-619 (2012). https://doi.org/10.1103/RevModPhys.84.567
Baidu ScholarGoogle Scholar
212P. Papka, R. Álvarez-Rodríguez, F. Nemulodiet al.,

Decay of 6Be populated in the 6Li(3He, 3H) charge-exchange reaction

. Phys. Rev. C 81, 054308 (2010). https://doi.org/10.1103/PhysRevC.81.054308
Baidu ScholarGoogle Scholar
213F.C. Barker,

R-matrix formulas for three-body decay widths

. Phys. Rev. C 68, 054602 (2003). https://doi.org/10.1103/PhysRevC.68.054602
Baidu ScholarGoogle Scholar
214B. Blank, M. Płoszajczak,

Two-proton radioactivity

. Rep. Prog. Phys. 71, 046301 (2008). https://doi.org/10.1088/0034-4885/71/4/046301
Baidu ScholarGoogle Scholar
215R. Álvarez-Rodríguez, H.O.U. Fynbo, A.S. Jensenet al.,

Distinction between sequential and direct three-body decays

. Phys. Rev. Lett. 100, 192501 (2008). https://doi.org/10.1103/PhysRevLett.100.192501
Baidu ScholarGoogle Scholar
216L.V. Grigorenko, T.D. Wiser, K. Mercurioet al.,

Three-body decay of 6Be

. Phys. Rev. C 80, 034602 (2009). https://doi.org/10.1103/PhysRevC.80.034602
Baidu ScholarGoogle Scholar
217I.A. Egorova, R.J. Charity, L.V. Grigorenkoet al.,

Democratic decay of 6Be exposed by correlations

. Phys. Rev. Lett. 109, 202502 (2012). https://doi.org/10.1103/PhysRevLett.109.202502
Baidu ScholarGoogle Scholar
218V. Chudoba, L.V. Grigorenko, A.S. Fomichevet al.,

Three-body correlations in direct reactions: Example of 6Be populated in the (p,n) reaction

. Phys. Rev. C 98, 054612 (2018). https://doi.org/10.1103/PhysRevC.98.054612
Baidu ScholarGoogle Scholar
219D.X. Zhu, Y.Y. Xu, H.M. Liuet al.,

Two-proton radioactivity of the excited state within the gamow-like and modified gamow-like models

. Nucl. Sci. Tech. 33, 122 (2022). https://doi.org/10.1007/s41365-022-01116-9
Baidu ScholarGoogle Scholar
220S.M. Wang, W. Nazarewicz, R.J. Charityet al.,

Structure and decay of the extremely proton-rich nuclei 11,12O

. Phys. Rev. C 99, 054302 (2019). https://doi.org/10.1103/PhysRevC.99.054302
Baidu ScholarGoogle Scholar
221D.R. Thompson, M. Lemere, Y.C. Tang,

Systematic investigation of scattering problems with the resonating-group method

. Nucl. Phys. A 286, 53-66 (1977). https://doi.org/10.1016/0375-9474(77)90007-0
Baidu ScholarGoogle Scholar
222K. Miernik, W. Dominik, Z. Janaset al.,

Two-proton correlations in the decay of 45Fe

. Phys. Rev. Lett. 99, 192501 (2007). https://doi.org/10.1103/PhysRevLett.99.192501
Baidu ScholarGoogle Scholar
223T.B. Webb, R.J. Charity, J.M. Elsonet al.,

Particle decays of levels in 11,12N and 12O investigated with the invariant-mass method

. Phys. Rev. C 100, 024306 (2019). https://doi.org/10.1103/PhysRevC.100.024306
Baidu ScholarGoogle Scholar
224T.B. Webb, S.M. Wang, K.W. Brownet al.,

First observation of unbound 11O, the mirror of the halo nucleus 11Li

. Phys. Rev. Lett. 122, 122501 (2019). https://doi.org/10.1103/PhysRevLett.122.122501
Baidu ScholarGoogle Scholar
225L.V. Grigorenko, I.G. Mukha, I.J. Thompsonet al.,

Two-proton widths of 12O, 16Ne, and three-body mechanism of thomas-ehrman shift

. Phys. Rev. Lett. 88, 042502 (2002). https://doi.org/10.1103/PhysRevLett.88.042502
Baidu ScholarGoogle Scholar
226L.P. Kok,

Accurate determination of the ground-state level of the 2He nucleus

. Phys. Rev. Lett. 45, 427-430 (1980). https://doi.org/10.1103/PhysRevLett.45.427
Baidu ScholarGoogle Scholar
227T.B. Webb, R.J. Charity, J.M. Elsonet al.,

Invariant-mass spectrum of 11O

. Phys. Rev. C 101, 044317 (2020). https://doi.org/10.1103/PhysRevC.101.044317
Baidu ScholarGoogle Scholar
228H.T. Fortune,

Energy and width of 11O(g.s.)

. Phys. Rev. C 99, 051302 (2019). https://doi.org/10.1103/PhysRevC.99.051302
Baidu ScholarGoogle Scholar
229E. Garrido, A.S. Jensen,

Few-body structures in the mirror nuclei 11O and 11Li

. Phys. Rev. C 101, 034003 (2020). https://doi.org/10.1103/PhysRevC.101.034003
Baidu ScholarGoogle Scholar
230X. Mao, J. Rotureau, W. Nazarewiczet al.,

Gamow-shell-model description of Li isotopes and their mirror partners

. Phys. Rev. C 102, 024309 (2020). https://doi.org/10.1103/PhysRevC.102.024309
Baidu ScholarGoogle Scholar
231E. Merzbacher, Quantum Mechanics, (John Wiley and Sons, inc, 1998)
232E.J. Hellund,

The decay of resonance radiation by spontaneous emission

. Phys. Rev. 89, 919-922 (1953). https://doi.org/10.1103/PhysRev.89.919
Baidu ScholarGoogle Scholar
233L.A. Khalfin,

Contribution to the decay theory of a quasi-stationary state

. Sov. Phys. JETP 6, 1053 (1958).
Baidu ScholarGoogle Scholar
234M. Lévy,

On the validity of the exponential law for the decay of an unstable particle

. Il Nuovo Cimento (1955-1965) 14, 612-624 (1959). https://doi.org/10.1007/BF02726390
Baidu ScholarGoogle Scholar
235J. Schwinger,

Field theory of unstable particles

. Ann. Phys. 9, 169-193 (1960). https://doi.org/10.1016/0003-4916(60)90027-0
Baidu ScholarGoogle Scholar
236R.G. Winter,

Evolution of a quasi-stationary state

. Phys. Rev. 123, 1503-1507 (1961). https://doi.org/10.1103/PhysRev.123.1503
Baidu ScholarGoogle Scholar
237R.G. Newton,

The exponential decay law of unstable systems

. Ann. Phys. (N.Y.) 14, 333-345 (1961). https://doi.org/10.1016/0003-4916(61)90060-4
Baidu ScholarGoogle Scholar
238M.L. Goldberger, K.M. Watson,

Lifetime and decay of unstable particles in S-matrix theory

. Phys. Rev. 136, B1472-B1480 (1964). https://doi.org/10.1103/PhysRev.136.B1472
Baidu ScholarGoogle Scholar
239L. Fonda, G.C. Ghirardi, A. Rimini,

Decay theory of unstable quantum systems

. Rep. Prog. Phys. 41, 587-631 (1978). https://doi.org/10.1088/0034-4885/41/4/003
Baidu ScholarGoogle Scholar
240P.T. Greenland,

Seeking non-exponential decay

. Nature 335, 298-298 (1988). https://doi.org/10.1038/335298a0
Baidu ScholarGoogle Scholar
241G. Esposito, G. Marmo, G. Sudarshan, From Classical to Quantum Mechanics: An Introduction to the Formalism, Foundations and Applications, (Cambridge University Press, 2004). https://doi.org/10.1017/CBO9780511610929
242V. Fock, N. Krylov,

On two main interpretations of energy-time uncertainty

. J. Exp. Theor. Phys. 17, 93 (1947).
Baidu ScholarGoogle Scholar
243M. Miyamoto,

Zero energy resonance and the logarithmically slow decay of unstable multilevel systems

. J. Math. Phys. 47, 082103 (2006). https://doi.org/10.1063/1.2227260
Baidu ScholarGoogle Scholar
244D.F. Ramírez Jiménez, N.G. Kelkar,

Quantum decay law: critical times and the equivalence of approaches

. J. Phys. A 52, 055201 (2019). https://doi.org/10.1088/1751-8121/aaf9f3
Baidu ScholarGoogle Scholar
245D.F. Ramírez Jiménez, N.G. Kelkar,

Formal aspects of quantum decay

. Phys. Rev. A 104, 022214 (2021). https://doi.org/10.1103/PhysRevA.104.022214
Baidu ScholarGoogle Scholar
246S.M. Wang, W. Nazarewicz, A. Volyaet al.,

Probing the nonexponential decay regime in open quantum systems

. Phys. Rev. Res. 5, 023183 (2023). https://doi.org/10.1103/PhysRevResearch.5.023183
Baidu ScholarGoogle Scholar
247C. Rothe, S.I. Hintschich, A.P. Monkman,

Violation of the exponential-decay law at long times

. Phys. Rev. Lett. 96, 163601 (2006). https://doi.org/10.1103/PhysRevLett.96.163601
Baidu ScholarGoogle Scholar
248T. Mercouris, C.A. Nicolaides,

Time dependence and properties of nonstationary states in the continuous spectrum of atoms

. J. Phys. B 30, 811 (1997). https://doi.org/10.1088/0953-4075/30/4/006
Baidu ScholarGoogle Scholar
249D.S. Onley, A. Kumar,

Time dependence in quantum mechanics-study of a simple decaying system

. Am. J. Phys. 60, 432-439 (1992). https://doi.org/10.1119/1.16897
Baidu ScholarGoogle Scholar
250U. Peskin, H. Reisler, W.H. Miller,

On the relation between unimolecular reaction rates and overlapping resonances

. J. Chem. Phys. 101, 9672-9680 (1994). https://doi.org/10.1063/1.467932
Baidu ScholarGoogle Scholar
251R. de la Madrid,

Numerical calculation of the decay widths, the decay constants, and the decay energy spectra of the resonances of the delta-shell potential

. Nucl. Phys. A 962, 24-45 (2017). https://doi.org/10.1016/j.nuclphysa.2017.03.006
Baidu ScholarGoogle Scholar
252V.V. Sokolov, V.G. Zelevinsky,

Dynamics and statistics of unstable quantum states

. Nucl. Phys. A 504, 562-588 (1989). https://doi.org/10.1016/0375-9474(89)90558-7
Baidu ScholarGoogle Scholar
253A. Volya, V. Zelevinsky,

Non-hermitian effective hamiltonian and continuum shell model

. Phys. Rev. C 67, 054322 (2003). https://doi.org/10.1103/PhysRevC.67.054322
Baidu ScholarGoogle Scholar
254A.I. Magunov, I. Rotter, S.I. Strakhova,

Fano resonances in the overlapping regime

. Phys. Rev. B 68, 245305 (2003). https://doi.org/10.1103/PhysRevB.68.245305
Baidu ScholarGoogle Scholar
255K. Kravvaris, A. Volya,

Quest for superradiance in atomic nuclei

. AIP Conf. Proc. 1912, 020010 (2017). https://doi.org/10.1063/1.5016135
Baidu ScholarGoogle Scholar
256S.R. Stroberg, J. Henderson, G. Hackmanet al.,

Systematics of e2 strength in the sd shell with the valence-space in-medium similarity renormalization group

. Phys. Rev. C 105, 034333 (2022). https://doi.org/10.1103/PhysRevC.105.034333
Baidu ScholarGoogle Scholar
257M. Heinz, A. Tichai, J. Hoppeet al.,

In-medium similarity renormalization group with three-body operators

. Phys. Rev. C 103, 044318 (2021). https://doi.org/10.1103/PhysRevC.103.044318
Baidu ScholarGoogle Scholar
258B.C. He, S.R. Stroberg,

Factorized approximation to the imsrg(3)

. (2024). arXiv:2405.19594
Baidu ScholarGoogle Scholar
259S.R. Stroberg, T.D. Morris, B.C. He,

Imsrg with flowing 3 body operators, and approximations thereof

. (2024). arXiv:2406.13010
Baidu ScholarGoogle Scholar
Footnote

Dedicated to Professor Wenqing Shen in honour of his 80th birthday