Continuum Skyrme Hartree–Fock–Bogoliubov theory with Green's function method for neutron-rich Ca, Ni, Zr, and Sn isotopes

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Continuum Skyrme Hartree–Fock–Bogoliubov theory with Green's function method for neutron-rich Ca, Ni, Zr, and Sn isotopes

En-Bo Huo
Ke-Ran Li
Xiao-Ying Qu
Ying Zhang
Ting-Ting Sun
Nuclear Science and TechniquesVol.34, No.7Article number 105Published in print Jul 2023Available online 20 Jul 2023
10202

The possible exotic nuclear properties in the neutron-rich Ca, Ni, Zr, and Sn isotopes are examined with the continuum Skyrme Hartree–Fock–Bogoliubov theory in the framework of the Green's function method. The pairing correlation, the couplings with the continuum, and the blocking effects for the unpaired nucleon in odd-A nuclei are properly treated. The Skyrme interaction SLy4 is adopted for the ph channel and the density-dependent δ interaction is adopted for the pp channel, which well reproduce the experimental two-neutron separation energies S2n and one-neutron separation energies Sn. It is found that the criterion Sn>0 predicts a neutron drip line with neutron numbers much smaller than those for S2n>0. Owing to the unpaired odd neutron, the neutron pairing energies -Epair in odd-A nuclei are much lower than those in the neighboring even-even nuclei. By investigating the single-particle structures, the possible halo structures in the neutron-rich Ca, Ni, and Sn isotopes are predicted, where sharp increases in the root-mean-square (rms) radii with significant deviations from the traditional rA1/3 rule and diffuse spatial density distributions are observed. Analyzing the contributions of various partial waves to the total neutron density ρlj(r)/ρ(r) reveals that the orbitals located around the Fermi surface—particularly those with small angular momenta—significantly affect the extended nuclear density and large rms radii. The number of neutrons Nλ (N0) occupying above the Fermi surface λn (continuum threshold) is discussed, whose evolution as a function of the mass number A in each isotope is consistent with that of the pairing energy, supporting the key role of the pairing correlation in halo phenomena.

Neutron-rich nucleineutron haloSkyrme Hartree–Fock–Bogoliubov theoryGreen's function method
1

Introduction

The study of exotic nuclei far from the β stability line is a challenging frontier in experimental [1-4] and theoretical [5-10] nuclear physics. The unstable nuclei with extreme N/Z ratios, which are weakly bound systems, have exhibited many exotic properties that differ from those of the stable nuclei, such as halo structures [11-17], changes in the traditional magic numbers [18-23], and new nuclear excitation modes [24, 25], which may herald new physics. The study of exotic nuclei is crucial for not only comprehensively understanding the rich nuclear structures and properties but also investigating element synthesis and nuclear astrophysics [26]. However, because of the short lifespan, the cross sections for synthesizing exotic nuclei are small, which makes it difficult to create them experimentally. Thus, advanced large-scale radioactive beam facilities and updated detector techniques have been developed, upgraded, or planned worldwide [27-34]. Meanwhile, abundant theoretical studies on exotic nuclei provide valuable guidance for the design of experiments and analysis of experimental results [35-37].

In the exotic nuclei—particularly the drip line nuclei—the neutron or proton Fermi surfaces are typically close to the continuum threshold. With the effects of the pairing correlation, the valence nucleons have a certain probability to be scattered into the continuum and occupy the resonant states therein, making the nuclear density distributions very diffuse and extended. It is therefore essential to treat the pairing correlations and the couplings to the continuum properly in the theoretical descriptions of exotic nuclei [38-42]. Additionally, for one-neutron halo nuclei, e.g., 31Ne [43] and 37Mg [44], the blocking effect [45] should be considered to treat the unpaired odd nucleon.

The Hartree–Fock–Bogliubov (HFB) theory has achieved great success in describing exotic nuclei with a unified description of the mean field and pairing correlation via Bogoliubov transformation [45]. Different models based on the HFB theory have been used to study exotic nuclei, such as the Gogny–HFB theory [46], the Skyrme–HFB theory [47], the relativistic continuum Hartree–Bogoliubov (RCHB) theory [39], and the density-dependent relativistic Hartree–Fock–Bogoliubov (RHFB) theory [48]. To explore the halo phenomena in deformed nuclei, these models have been extended to the deformed framework, e.g., the deformed relativistic Hartree–Bogoliubov (DRHB) theory [16, 49, 50] and the coordinate-space Skyrme–HFB approach [51-54].

Traditionally, these H(F)B equations are often solved in configuration spaces, i.e., via the basis expansion method [55]. However, the calculations are closely related to the space size and the shape of the expanded basis. Although the harmonic oscillator basis, which has a tail in the shape of a Gaussian function, can efficiently describe stable nuclei, significant difficulties were encountered when it was applied to exotic nuclei. Bases with proper shapes, such as the Woods–Saxon basis [56] and the transformed harmonic oscillator basis [57, 58], are often used for the exotic nuclei. For example, to explore deformed halos [59, 60], the DRHB theory based on a Woods–Saxon basis [16, 49] was developed. In contrast to the basis expansion method, solving the HFB equation in the coordinate space is believed to be more effective. In the coordinate space, the discretized method with the box boundary condition has been widely used, whereby a series of discrete quasiparticle levels can be easily obtained. However, flaws in this method have been identified, such as the nonphysical drops in the nuclear densities at the box boundary, the discretization of continua and resonant states, and the inclusion of unphysical states. In contrast, the Green's function (GF) method [61-63] in the coordinate space can avoid these problems and has significant advantages, i.e., it can describe the asymptotic behaviors of wave functions properly, provide the energies and widths of the resonant states directly, and treat the bound states and the continua on the same footing.

Owing to these advantages, the GF method has been applied extensively in nuclear physics to study the contribution of the continuum to the nuclear structures and excitations. In 1987, Belyaev et al. formulated the GF for the HFB equation [64]. Subsequently, this HFB GF was applied to the quasiparticle random-phase approximation [65], which was further used to describe the collective excitations coupled to the continuum [66-71]. In 2009, the continuum HFB theory in a coupled channel representation was developed to explore the effects of the continuum and pairing correlation in deformed neutron-rich Mg isotopes [72]. In 2011, Zhang et al. developed the fully self-consistent continuum Skyrme–HFB theory with the GF method [73], which was applied to investigate the giant halos [74] and the effects of the pairing correlation on the quasiparticle resonances [75, 76]. In 2019, to explore the halo phenomena in neutron-rich odd-A nuclei, the self-consistent continuum Skyrme–HFB theory was extended by including the blocking effect [77]. In recent years, the GF method has also been adopted for the covariant density functional theory [78-86] in studies on nuclear structure. For example, by introducing the GF method to the relativistic mean field theory (GF-RMF), the single-particle level structures, including the bound states and resonant states and the pseudospin symmetries therein, were investigated for neutrons [87, 88], protons [89], and Λ hyperons [90]. Additionally, it was confirmed that exact values of the energies and widths could be obtained by searching for the poles of the GF or the extremes of the density of states in disregard of the widths of resonant states [91, 92]. By combining the GF method with the RCHB theory, the pairing correlation and continuum are well described in the giant halos of the Zr isotopes [93]. By extending the GF-RMF model to the coupled channel representation, the halo candidate nucleus 37Mg reported experimentally was analyzed and confirmed to be a p-wave one-neutron halo according to the Nilsson levels [94]. In addition, the complex-scaled GF method [95] has been established as a powerful tool for the exploration of resonant states, which was extended to the framework of the relativistic mean field [96] and deformed nuclei [97]. Additionally, the RMF-CMR-GF approach was developed by combining the complex momentum representation method with the GF method in the relativistic mean-field framework to study the halo structures in neutron-rich nuclei [98]. These studies proved the effectiveness of the GF method for the description of the continuum.

In this study, the neutron-rich Ca, Ni, Zr, and Sn isotopes are investigated systematically using the continuum Skyrme–HFB theory formulated with the GF method in the coordinate space, in which the pairing correlations, the couplings with the continuum, and the blocking effect for the odd unpaired nucleon are treated properly. The remainder of the paper is organized as follows. In Sec. 2, we briefly introduce the continuum Skyrme–HFB theory. Numerical details are presented in Sec. 3. The results and discussions are presented in Sec. 4, and conclusions are drawn in Sec. 5.

2

Theoretical framework

In the Hartree–Fock–Bogoliubov (HFB) theory [45], the pair correlated nuclear system is described in terms of independent quasiparticles by the Bogoliubov transformation. The HFB equation in the coordinate space [38] is (hλh˜h˜*h*+λ)ϕi(rσ)=Eiϕi(rσ), (1) where Ei represents the quasiparticle energy, ϕi(rσ) represents the quasiparticle wave function, and λ represents the Fermi energy, which is determined by constraining the expectation value of the nucleon number. The HF Hamiltonian h(rσ,r'σ) and the pair Hamiltonian h˜(rσ,r'σ) are obtained by the variation of the total energy functional with respect to the particle density ρ(rσ,r'σ) and the pair density ρ˜(rσ,r'σ), respectively. The solutions of the HFB equations have two symmetric branches: one is positive (Ei>0) with the quasiparticle wave function ϕi(rσ), and the other is negative (-Ei<0) with the conjugate wave function ϕ¯i˜(rσ). The wave functions ϕi(rσ) and ϕ¯i˜(rσ) are in two components and are related as follows: ϕi(rσ)=(φ1,i(rσ)φ2,i(rσ)),  ϕ¯i˜(rσ)=(φ2,i*(rσ˜)φ1,i*(rσ˜)), (2) with φ(rσ˜)=2(r,σ). In this paper, we follow the notation used in Ref. [65] for convenience.

For an even-even nucleus, the ground state |Φ0 is a quasiparticle vacuum with the particle density ρ(rσ,r'σ) and pairing density ρ˜(rσ,r'σ) determined as follows: ρ(rσ,r'σ)Φ0|cr'σcrσ|Φ0, (3a) ρ˜(rσ,r'σ)Φ0|cr'σ˜'crσ|Φ0, (3b) where crσ and crσ are the particle creation and annihilation operators, respectively. The densities can be unified as a generalized density matrix R(rσ,r'σ), with ρ(rσ,r'σ) and ρ˜(rσ,r'σ) being the “11” and “22” elements, respectively. With the quasiparticle wave functions, R(rσ,r'σ) can be written in a simple form: R(rσ,r'σ)=iϕ¯i˜(rσ)ϕ¯i˜(rσ). (4) For an odd-A nucleus, the last odd nucleon is unpaired, for which the blocking effect should be considered. The nuclear ground state in this case is a one-quasiparticle state |Φ1, which can be constructed on the basis of a HFB vacuum |Φ0 as |Φ1=βib|Φ0, (5) where βib is the quasiparticle creation operator and ib denotes the blocked quasiparticle level occupied by the odd nucleon. Accordingly, the particle density ρ(rσ,r'σ) and the pairing density ρ˜(rσ,r'σ) are ρ(rσ,r'σ)Φ1|cr'σcrσ|Φ1, (6a) ρ˜(rσ,r'σ)Φ1|cr'σ˜'crσ|Φ1, (6b) and the generalized density matrix R(rσ,r'σ) becomes R(rσ,r'σ)=i:allϕ¯i˜(rσ)ϕ¯i˜(rσ)ϕ¯i˜b(rσ)ϕ¯i˜b(rσ)+ϕib(rσ)ϕib(rσ), (7) where two more terms are introduced compared with those for the even-even nuclei.

In the conventional Skyrme–HFB theory, the HFB equation (1) in the coordinate space is often solved with the box boundary condition, and a series of discretized eigensolutions including the quasiparticle energy Ei and the corresponding wave functions ϕi(rσ) can be obtained. Then, the generalized density matrix R(rσ,r'σ) can be calculated by summing these discretized quasiparticle states, in accordance with Eqs. (5) and (9). We call this method the box-discretized approach. However, the applicability of the box boundary condition in the description of exotic nuclei—particularly those close to the drip line—is poor. A sufficiently large coordinate space (or box size) should be used to describe the extended density distribution.

The GF method can avoid these problems of the box-discretized approach, as it imposes the correct asymptotic behaviors on the wave functions—particularly for the weakly bound sates and the continuum. The GF G(rσ,r'σ;E) with an arbitrary quasiparticle energy E defined for the coordinate-space HFB equation obeys [E(hλh˜h˜*h*+λ)]G(rσ,r'σ;E)    =δ(rr)δσσ, (8) which is a 2× 2 matrix. The generalized density matrix R(rσ,r'σ) in Eq. (9) can be calculated by taking the integrals of the GFs on the complex quasiparticle energy plane, as follows: R(rσ,r'σ)=12πi[CE<0dEG(rσ,r'σ;E)CbdEG(rσ,r'σ;E)+Cb+dEG(rσ,r'σ;E)], (9) where the contour path CE<0 encloses all the negative quasiparticle energies -Ei<0, Cb encloses only the pole of Eib, and Cb+ encloses only the pole of Eib.

In the spherical case, the quasiparticle wave functions ϕi(rσ) and ϕ¯i˜(rσ) are only dependent on the radial parts, and they can be expanded as follows: ϕi(rσ)=1rϕnlj(r)Yjml(r^σ), (10a) ϕ¯i˜(rσ)=1rϕ¯nlj(r)Yjml*(r^σ˜), (10b) ϕnlj(r)=(φ1,nlj(r)φ2,nlj(r)), (10c) ϕ¯nlj(r)=(φ2,nlj*(r)φ1,nlj*(r)), (10d) where Yjml(r^σ) is the spin spherical harmonic, and Yjml(r^σ˜)=2σYjml(r^σ). Similarly, the generalized density matrix R(rσ,r'σ) and the GF G(rσ,r'σ;E) can be expanded as R(rσ,rσ)=ljmYjml(r^σ)Rlj(r,r)Yjml*(r^'σ), (11a) G(rσ,rσ;E)=ljmYjml(r^σ)Glj(r,r;E)rrYjml*(r^'σ), (11b) where Rlj(r,r) and Glj(r,r;E) are the radial parts of the generalized density matrix and GF, respectively.

Thus, the radial local generalized density matrix R(r)=R(r,r) can be expressed by the radial HFB GF Glj(r,r;E) as follows: R(r)=ljRlj(r,r)=14πr2[lj:all(2j+1)n:allϕ¯nlj2(r)ϕ¯nblbjb2(r)+ϕnblbjb2(r)]=14πr212πi[lj:all(2j+1)CE<0dEGlj(r,r;E)CbdEGlbjb(r,r;E)+Cb+dEGlbjb(r,r;E)]. (12) From the radial generalized matrix R(r), one can easily obtain the radial local particle density ρ(r) and pair density ρ˜(r), which are the “11” and “12” components of R(r), respectively. In the same way, one can express other radial local densities needed in the functional of the Skyrme interaction, such as the kinetic-energy density τ(r) and the spin–orbit density J(r), in terms of the radial GF. For the construction of the GF, see Refs. [73, 77].

3

Numerical details

For the Skyrme interaction in the ph channel, the SLy4 parameter set [100] is adopted. For the pairing interaction in the pp channel, a density-dependent δ interaction (DDDI) is used: vpair(r,r)=12(1Pσ)V0[1η(ρ(r)ρ0)α]δ(rr). (13) The pair Hamiltonian h˜(rσ,r'σ) is then reduced to a local pair potential [47]: Δ(r)=12V0[1η(ρ(r)ρ0)α]ρ˜(r). (14) The strength of the pairing force V0=-458.4 MeVfm3, density ρ0=0.08 fm-3, and parameters η=0.71 and α=0.59 are constrained by reproducing the experimental neutron pairing gaps for the Sn isotopes [101, 102, 68]. With these parameters, the DDDI can reproduce the scattering length a=-18.5 fm in the 1S channel of the bare nuclear force in the low density limit [101]. The cutoff of the quasiparticle states is set to a maximal angular momentum of jmax=25/2 and a maximal quasiparticle energy of Ecut=60 MeV.

The HFB equation is solved in the coordinate space with the space size R=20 fm and mesh size dr=0.1 fm. To calculate the densities with the GF, the integrals of the GFs are performed along a contour path CE<0, which is set as a rectangle with height γ = 0.1 MeV and length Ecut = 60 MeV to enclose all the quasiparticle states with negative energies. For the odd-A nuclei, two more contour paths Cb+ and Cb, which only enclose the blocked quasiparticle states at energies Eib and Eib, are introduced owing to the blocking effect of the odd unpaired nucleon. Details are presented in Ref. [77]. To perform the contour integration, an energy step of Δ E = 0.01 MeV on the contour path is adopted.

4

Results and discussion

In Fig. 1, the two-neutron separation energies S2n(N,Z) =E(N-2,Z) - E(N, Z) are plotted for the even-even and odd-even Ca, Ni, Zr, and Sn isotopes. Red circles indicate those calculated according to the continuum Skyrme–HFB theory with the SLy4 parameter set, blue triangles indicate the results of the discretized method, and black squares indicate the available experimental data [99]. The differences between the S2n values obtained via the GF method and discretized method are small. Good agreement with the experimental data is observed, indicating the reliability of the continuum Skyrme–HFB theory for the prediction of neutron drip line. The traditional shell closures, i.e., N=28 in Ca isotopes, N=50 in Ni isotopes, N=50, 82 in Zr isotopes, and N=82 in Sn isotopes, can be identified, where the S2n decreases sharply. For example, in the Sn chain, S2n decreases from 13.25 MeV at 132Sn to 4.94 MeV at 134Sn with the neutron number exceeding the magic number N=82. In the Ca, Ni, and Zr chains, the two-neutron separation energies quickly reach 0 as mass increasing, resulting in relatively short neutron drip lines, which are 67Ca, 89Ni, and 123Zr, respectively. In contrast, in the Sn chain, S2n remains below 1.0 MeV in a wide mass region after the gap of N=82 and finally becomes negative until A=178, suggesting that 177Sn is a neutron drip line nucleus. These weakly bound nuclei are interesting owing to the possible appearance of neutron halos, although this is experimentally difficult to achieve. In addition, the exploration of the neutron drip line and the determination of the limit of the nuclear landscape are important in nuclear physics. However, various theoretical studies indicate that the predicted neutron drip line is very model-dependent [103]. Moreover, different physical quantities and criteria yield different neutron drip line predictions.

Fig. 1
(Color online) Two-neutron separation energies S2n in the Ca, Ni, Zr, and Sn isotopes as a function of the mass number A calculated according to the continuum Skyrme–HFB theory with the SLy4 parameter set (filled red circles), in comparison with the results of the discretized method (open blue triangles) and experimental data (filled black squares) [99].
pic

To explore the neutron drip lines in the Ca, Ni, Zr, and Sn isotopes, in Fig. 2 the single-neutron separation energies Sn(N,Z)=E(N1,Z)E(N,Z) are plotted. The results obtained using the continuum Skyrme–HFB theory with the SLy4 parameter set are indicated by red circles, which are consistent with the experimental data [99] indicated by the black squares. Strong odd-even staggering is observed in all isotopes. In general, the Sn in the even-even nucleus is approximately 2~3 MeV larger than those in the neighboring odd-A nuclei, which is attributed to the unpaired odd neutron with vanishing pairing energy. Consequently, compared with those in Fig. 1, the neutron drip lines determined via the one-neutron separation energy are significantly shortened. In the Ca, Ni, Zr, and Sn isotopes, the drip line nuclei are 60Ca, 86Ni, 122Zr, and 148Sn, respectively, whose positions are indicated by the black arrows. Outside the neutron drip line determined by Sn, the bound even-even nuclei behave as interesting Borromean systems. For example, considering the bound nucleus 60Ca, 60Ca+n is unbound, while 60Ca+n+n is bound.

Fig. 2
(Color online) Single-neutron separation energies Sn as a function of the mass number A in the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes calculated according to the continuum Skyrme–HFB theory with the SLy4 parameter set (filled red circles), in comparison with the results of the discretized method (open blue triangles) and experimental data (filled black squares) [99].
pic

In Fig. 3, we plot the neutron pairing energy Epair, which is expressed as Epair=12drΔ(r)ρ˜(r). (15) The red solid symbols correspond to the even-even nuclei, and the open symbols correspond to the odd-A nuclei. The neutron pairing energies Epair of the odd-A nuclei are obviously lower than those of the neighboring even-even nuclei, owing to the absent contribution of pairing energy from the unpaired neutron. This also explains why the drip line determined via the single-neutron separation energy Sn is much shorter than that obtained via the two-neutron separation energy S2n. In addition, obvious shell effects are observed in the pairing energy. For example, in Sn isotopes, the pairing energy is 0 at N=82 and N=126, and it is maximized at the half-shell N=102. As a result, the traditional shell closures, i.e., N=28, 40 in Ca isotopes, N=40, 50 in Ni isotopes, N=50, 82 in Zr isotopes, and N=82,126 in Sn isotopes, can also be clearly observed, which are consistent with those shown in Fig. 1. In addition, a sub-shell N=32 is observed in Ca isotopes.

Fig. 3
(Color online) Neutron pairing energy -Epair as a function of the mass number A in the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes calculated according to the continuum Skyrme–HFB theory with the SLy4 parameter set. The filled and open squares indicate the results for the even-even and odd-even nuclei, respectively.
pic

In the following, the possible neutron halos in the Ca, Ni, Zr, and Sn isotopes are examined—particularly those in the weakly bound nuclei close to the neutron drip line, where the Fermi surfaces are very close to the continuum threshold and the valence neutrons can be easily scattered to the continuum by the pairing correlation.

In Fig. 4, the neutron canonical single-particle structure as a function of the mass number is plotted for the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes. Details on obtaining the canonical single-particle levels are presented in Refs. [104, 105]. The neutron Fermi energy λn as well as the canonical single-particle energies ε are shown. As the neutron number increases, the Fermi energy λn in each chain increases, finally reaching the continuum threshold, while all the HF single-particle levels decrease. The traditional shell closures, i.e., N=28, 40 in Ca isotopes, N=40, 50 in Ni isotopes, N=82 in Zr isotopes, and N=82 in Sn isotopes, can be observed, exhibiting large gaps. Different single-particle structures are revealed in the Ca, Ni, Zr, and Sn chains, whereby halos may be formed. In the Ni chain, above the shell closure of N=50, there are several weakly bound states and low-lying positive canonical states in the continuum with small angular momenta, which favor the formation of halos. For example, in 86Ni, around the Fermi surface, there are two weakly bound states, i.e., 2d5/2 and 3s1/2, and one low-lying positive canonical state 2d3/2. The Sn chain is similar to the Ni chain but has more advantages for halo formation, where weakly bound states and the low-lying positive states in the continuum exist above the N=82 shell closure. Additionally, the Fermi surface λn gradually approaches 0, and the Sn isotopes in a large mass region are weakly bound. In the Ca chain, above the shell closure of N=40, the main state is 1g9/2, which evolves from a canonical positive state in the continuum (A62) to a weakly bound level (A64). Although there is a possibility of valence neutrons occupying the 1g9/2 orbital, the contributed density is very localized owing to the large central barrier. In the neutron-rich Ca isotopes, the low-lying positive canonical states in the continuum 3s1/2 and 2d5/2 also play important roles. The formation of halos is most unlikely for the Zr chain, where the shell closure of N=82 is located around the threshold of the continuum and it is difficult for the valence neutrons to overcome the large gap and occupy the continuum.

Fig. 4
(Color online) Single-particle levels for neutrons around the Fermi surface in the canonical basis for the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes calculated according to the continuum Skyrme–HFB theory. The short dashed lines represent the Fermi surface.
pic

In Fig. 5, the neutron root-mean-square (rms) radii rrms=dr4πr4ρ(r)dr4πr2ρ(r), (16) are plotted for the Ca, Ni, Zr, and Sn isotopes, which are based on Skyrme–HFB calculations with the GF method (filled circles and solid lines) and box-discretized method (open triangles and dashed lines) employing the SLy4 parameter set. The radii r=b0 A1/3 in the traditional liquid-drop model (black lines) are shown as well, with the coefficient b0 determined via the radii of deeply bound nuclei. In the Ca, Ni, Zr, and Sn chains, they are r0.991A1/3, 0.984A1/3, 0.957A1/3, and 0.961 A1/3, respectively. In each chain, with the addition of neutrons, the nuclear rms radii rrms increase steeply and deviate from the radii A1/3 rule. For example, in the Ca chain, compared with the isotopic trend in N20 with r0.991A1/3, the neutron rms radii in 50Ca and the heavier isotopes exhibit steep increases with an increase in N. In this mass range, possible neutron halos may occur. Additionally, odd-even staggering of rms radii can be clearly observed in the Sn chain, where the odd-A nuclei 151-165Sn have larger rms radii than the neighboring two even-even nuclei. Details are presented in Ref. [77]. The odd-even staggering phenomena in nuclear radii and nuclear mass have attracted considerable research interest in recent years [106-112].

Fig. 5
(Color online) Neutron rms radii rrms for the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes calculated according to the Skyrme–HFB theory with the GF method (filled circles and solid lines), in comparison with the results of the discretized method (open triangles and dashed lines).
pic

In exotic nuclei, diffuse density distributions in the coordinate space are often observed. Thus, in Fig. 6, to explore the exotic structures in the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes, we also plot the neutron density distributions ρ(r), where the solid and dashed lines indicate those obtained via the GF method and the box-discretized method, respectively. As a global trend, the neutron density distributions are extended with an increase in the neutron number. The shell structures significantly affect the density distribution; i.e., compared with the bound nuclei, the density distributions of the neutron-rich nuclei in the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes with the neutron number exceeding the neutron closure N=28, 50, 50, 82 are far more extended, which is consistent with the behaviors of the rms radii plotted in Fig. 5. In addition, compared with the Ca and Zr chains, the Ni and Sn chains exhibit more diffuse density distributions, which can be explained by their small two-neutron separation energies S2n in a large mass range, as shown in Fig. 1. For the Zr isotopes, the density distributions are relatively localized. The large S2n shown in Fig. 1 suggests the absence of halos in Zr isotopes. However, according to the RCHB theory with the NLSH parameter set [113] and the continuum Skyrme–HFB theory with the SKI4 parameter set [114, 74], giant halos in Zr isotopes have been predicted. In all isotopes, compared with the box-discretized method predicting nonphysical sharp reductions in density at the space boundary, the GF method can better describe the extended density distributions—particularly for very neutron-rich isotopes. Additionally, the densities obtained via the GF method can be independent of the space sizes, as discussed in Ref. [77], which are mainly determined by the proper boundary conditions of the bound states, weakly bound sates, and the continuum employed when constructing the GFs.

Fig. 6
(Color online) Neutron density ρ(r) for the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes calculated according to the Skyrme–HFB theory with the GF method (solid lines) and the discretized method (dashed lines).
pic

To explore the contributions of different partial waves to the extended density distributions in Fig. 6, taking the neutron-rich (a) 64Ca, (b) 86Ni, (c) 120Zr, and (d) 174Sn as examples, we plot in Fig. 7 the compositions ρlj(r)/ρ(r) as functions of the radial coordinate r. As shown, outside the nuclear surface referring to the right boundary of the shallow regions of the total nuclear density distributions, the orbitals located around the Fermi surface have the most significant effect on the density distributions. For example, in neutron-rich 64Ca, the partial waves p1/2, f5/2, g9/2, s1/2, and d5/2 contribute significantly to the total density in the area of 5 fm<r<15 . These levels are located within ~5 MeV around the Fermi surface, as shown in Fig. 8, where the neutron canonical single-particle levels as well as the occupation probabilities are presented. As we move further in the coordinate space with r>15 fm, the contributions of the partial waves s1/2 and d5/2 with small angular momenta increase significantly, while the contributions of other partial waves decrease. In the case of 86Ni, the single-particle levels 2d5/2, 3s1/2, and 2d3/2 are located above the neutron shell of N=50 and close to the Fermi surface, playing the main role in the neutron density distribution. Although the positive state 1g7/2 is also very close to the Fermi surface, it contributes little to the density in the large coordinate space owing to the large centrifugal barrier. For the nucleus 120Zr with the neutron number very close to the closure of N=82, the single-particle levels between the closures N=50 and N=82, including 2d5/2, 1g7/2, 3s1/2, 2d3/2, and 1h11/2, contribute significantly to the densities in the large coordinate space. The insignificant occupation of the positive state 2f7/2 leads to a small contribution to the density. Regarding the neutron-rich 174Sn with the neutron number exceeding the closure of N=82, the weakly bound single-particle levels 2f7/2, 2f5/2, 3p3/2, and 3p1/2 with small angular momenta play the key role in the extended density distributions in the large coordinate space. From this analysis, we can conclude that the single-particle levels around the Fermi surface—particularly the waves with small angular momenta—are the main cause of the extended nuclear density distributions.

Fig. 7
(Color online) Contributions of different partial waves to the total neutron density ρlj(r)/ρ(r) as a function of the radial coordinate r for the neutron-rich nuclei (a) 64Ca, (b) 86Ni, (c) 120Zr, and (d) 174Sn. The shallow regions are for the nuclear density distributions ρ(r), which are rescaled by multiplying by a factor of 5.
pic
Fig. 8
(Color online) Neutron canonical single-particle levels in (a) 64Ca, (b) 86Ni, (c) 120Zr, and (d) 174Sn calculated according to the Skyrme–HFB theory with the GF method. The blue dashed lines represent the Fermi surface. The occupation probabilities (v2) for the canonical orbitals are proportional to the lengths of the lines (logarithmic scale).
pic

In Fig. 8, the particle occupation probabilities v2 on different canonical levels εNcan are indicated by the lengths of the lines. Without pairing, the values of the occupation probabilities v2 should be either 1 or 0, separated by the Fermi surface. With the effect of pairing, the nucleons occupying the levels below the Fermi surface can be scattered to higher levels, which results in the occupation of the weakly bound states above the Fermi surface and even levels in the continuum. In the neutron-rich nuclei 64Ca, 86Ni, 120Zr, and 174Sn, the numbers of neutrons Nλ=εk>λn(2j+1)vk2 scattered above the Fermi surface λn are 4.33, 2.51, 0.159, and 0.173, respectively. In the case of 64Ca, the weakly bound single-particle 1g9/2 contributes approximately 3.7 neutrons.

To explore the effects of pairing, we plot in Fig. 9 the numbers of neutrons Nλ scattered above the Fermi surface for the (a) Ca, (b) Ni, (c) Zr, and (d) Sn chains obtained via the Skyrme–HFB theory with the GF method. As shown, large numbers of neutrons are scattered from the single-particle levels below the Fermi surface to the weakly bound states above the Fermi surface and even levels in the continuum owing to the pairing—particularly in the nuclei with the neutron number filling the half-full shells. Additionally, an obvious shell structure is observed. When the number of neutrons reaches a magic number, i.e., N=28, 40 in Ca isotopes, N=40, 50 in Ni isotopes, N=50, 82 in Zr isotopes, and N=82,126 in Sn isotopes, Nλ is almost 0 owing to the absence of the pairing for the closed-shell nuclei. Additionally, at the points of N=32 in the Ca isotopes, N=54,68 in the Zr isotopes, and N=88 in the Sn isotopes, very small numbers of neutrons (Nλ) are obtained, indicating weak pairing in these nuclei, along with the possible existence of sub-shells and new magic numbers. Furthermore, the evolution of Nλ is consistent with the trend of the pair energy -Epair in Fig. 3.

Fig. 9
(Color online) Number of neutrons Nλ occupying the single-particle levels above the Fermi surface λn as a function of the mass number A for the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes calculated according to the Skyrme–HFB theory with the GF method.
pic

In Fig. 10, we further investigate the number of neutrons occupying the continuum with single-particle energies of ε>0 MeV, i.e., N0=εk>0(2j+1)vk2. Compared with the number of neutrons Nλ occupying the levels above the Fermi surface, the number of neutrons occupying the continuum N0 is significantly smaller. For example, in the Ca chain, N0 is less than 1 in all isotopes except 62Ca. In 62Ca, the single-particle level 1gg/2 appears as a low-lying canonical positive state in the continuum with energy of ε=0.198 MeV and a high occupation probability of v2 = 0.197, resulting in almost 1.979 neutrons occupying it. However, in the neighboring 60Ca, a very low occupation probability v2=0.015 of 1g9/2 is obtained, and in 64Ca, the 1g9/2 state drops to a weakly bound level with energy of ε=-0.072 MeV. After the neutron contribution is removed from 1gg/2 in 62Ca, only 0.38 neutrons are in the continuum, which is denoted by an empty circle in panel (a). Except for the Sn chain, the shape of the evolution of N0 is very close to those of the pairing energy and pairing gap. Although the case of the Sn chain becomes very complex, we can observe the shell structure at N=82 and N=126, where N0 is almost 0.

Fig. 10
(Color online) Number of neutrons N0 occupying in the continuum (above the threshold ε=0 MeV) as a function of the mass number A for the (a) Ca, (b) Ni, (c) Zr, and (d) Sn isotopes calculated according to the Skyrme–HFB theory with the GF method.
pic
5

Summary

In this study, the exotic nuclear properties of neutron-rich Ca, Ni, Zr, and Sn isotopes were examined systematically according to the continuum Skyrme–HFB theory in the coordinate space formulated with the GF method, in which the pairing correlations, the couplings to the continuum, and the blocking effects for the unpaired nucleon in odd-A nuclei are treated properly.

First, the two-neutron separation energies S2n and one-neutron separation energies Sn were calculated, which were consistent with experimental data. Significant differences exist for the drip lines determined by S2n and Sn. In the Ca, Ni, Zr, and Sn isotopes, the drip line nuclei are 67Ca, 89Ni, 123Zr, and 177Sn according to S2n and 60Ca, 86Ni, 122Zr, and 148Sn according to Sn. Owing to the absent contribution of pairing energy from the single unpaired odd neutron, the neutron pairing energies (Epair) of the odd-A nuclei are approximately 2 MeV lower than those of the neighboring even-even nuclei. This explains why the drip lines determined via Sn are much shorter than those determined via S2n. In addition, from the fluctuation trends of the pairing energy, the traditional neutron magic numbers are clearly displayed, i.e., N=28, 40 in Ca isotopes, N=40, 50 in Ni isotopes, N=50, 82 in Zr isotopes, and N=82,126 in Sn isotopes.

Second, to explore the possible halo structures in the neutron-rich Ca, Ni, Zr, and Sn isotopes, the neutron single-particle structures, the rms radii, and the density distributions were investigated. In the neutron-rich Ca, Ni, Sn nuclei—particularly the weakly bound nuclei close to neutron drip line—the rms radii increase sharply, with significant deviations from the traditional rA1/3 rule. Additionally, very diffuse spatial density distributions are observed in these nuclei, possibly indicating a halo phenomenon therein. By analyzing the contributions of different partial waves to the total density, we found that the orbitals located around the Fermi surface—particularly those with small angular momenta—are the main cause of the extended nuclear density and large rms radii.

Finally, the numbers of halo nucleons that can reflect the effects of pairing were examined. Two different numbers of neutrons were defined: Nλ neutrons occupying the single-particle levels above the Fermi surface λn and N0 neutrons occupying the continuum. We found that the evolutions of Nλ and N0 with respect to the mass number A are consistent with the trend of the pairing energy -Epair, which supports the key role of the pairing correlations in the halo phenomena.

The theoretical calculations were supported by the nuclear data storage system at Zhengzhou University.

References
1. I. Tanihata,

Nuclear structure studies from reaction induced by radioactive nuclear beams

. Prog. Part. Nucl. Phys. 35, 505 (1995). doi: 10.1016/0146-6410(95)00046-L
Baidu ScholarGoogle Scholar
2. B. Jonson,

Light drip line nuclei

. Phys. Rep. 389, 1 (2004). doi: 10.1016/j.physrep.2003.07.004
Baidu ScholarGoogle Scholar
3. I. Tanihata, H. Savajols, and R. Kanungo,

Recent experimental progress in nuclear halo structure studies

. Prog. Part. Nucl. Phys. 68, 215 (2013). doi: 10.1016/j.ppnp.2012.07.001
Baidu ScholarGoogle Scholar
4. T. Nakamura,

Neutron halo - recent experimental progress at RIBF

. AAPPS Bull. 29, 19 (2019). doi: 10.22661/AAPPSBL.2019.29.5.19
Baidu ScholarGoogle Scholar
5. A.C. Mueller and B.M. Sherrill,

Nuclei at the limits of particle stability

. Annu. Rev. Nucl. Part. Sci. 43, 529 (1993). doi: 10.1146/annurev.ns.43.120193.002525
Baidu ScholarGoogle Scholar
6. P. Hansen,

Nuclear halos: Structure and reactions

. Nucl. Phys. A 588, c1 (1995). doi: 10.1016/0375-9474(95)00091-E
Baidu ScholarGoogle Scholar
7. R. Casten and B. Sherrill,

The study of exotic nuclei

. Prog. Part. Nucl. Phys. 45, S171 (2000). doi: 10.1016/S0146-6410(00)90013-9
Baidu ScholarGoogle Scholar
8. A.S. Jensen, K. Riisager, D.V. Fedorov et al.,

Structure and reactions of quantum halos

. Rev. Mod. Phys. 76, 215 (2004). doi: 10.1103/RevModPhys.76.215
Baidu ScholarGoogle Scholar
9. J. Meng, H. Toki, S.-G. Zhou et al.,

Relativistic continuum Hartree Bogoliubov theory for ground-state properties of exotic nuclei

. Prog. Part. Nucl. Phys. 57, 470 (2006). doi: 10.1016/j.ppnp.2005.06.001
Baidu ScholarGoogle Scholar
10. S.N. Ershov, L.V. Grigorenko, J.S. Vaagen et al.,

Halo formation and breakup: lessons and open questions

. J. Phys. G Nucl. Phys. 37, 064026 (2010). doi: 10.1088/0954-3899/37/6/064026
Baidu ScholarGoogle Scholar
11. I. Tanihata, H. Hamagaki, O. Hashimoto, et al.,

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

. Phys. Rev. Lett. 55, 2676 (1985). doi: 10.1103/PhysRevLett.55.2676
Baidu ScholarGoogle Scholar
12. T. Minamisono, T. Ohtsubo, I. Minami et al.,

Proton halo of 8B disclosed by its giant quadrupole moment

. Phys. Rev. Lett. 69, 2058-2061 (1992). doi: 10.1103/PhysRevLett.69.2058
Baidu ScholarGoogle Scholar
13. W. Schwab, H. Geissel, H. Lenske et al.,

Observation of a proton halo in 8B

. Z. Phys. A 350, 283 (1995). doi:https://link.springer.com/content/pdf/10.1007/bf01291183
Baidu ScholarGoogle Scholar
14. J. Meng and P. Ring,

Relativistic Hartree-Bogoliubov description of the neutron halo in 11Li

. Phys. Rev. Lett. 77, 3963 (1996). doi: 10.1103/PhysRevLett.77.3963
Baidu ScholarGoogle Scholar
15. J. Meng and P. Ring,

Giant halo at the neutron drip line

. Phys. Rev. Lett. 80, 460 (1998). doi: 10.1103/PhysRevLett.80.460
Baidu ScholarGoogle Scholar
16. S.-G. Zhou, J. Meng, P. Ring et al.,

Neutron halo in deformed nuclei

. Phys. Rev. C 82, 011301 (2010). doi: 10.1103/PhysRevC.82.011301
Baidu ScholarGoogle Scholar
17. X.-X. Sun and S.-G. Zhou,

Rotating deformed halo nuclei and shape decoupling effects

. Sci. Bull. 66, 2072 (2021). doi: 10.1016/j.scib.2021.07.005
Baidu ScholarGoogle Scholar
18. A. Ozawa, T. Kobayashi, T. Suzuki, K. Yoshida, and I. Tanihata,

New magic number, N=16, near the neutron drip line

. Phys. Rev. Lett. 84, 5493 (2000). doi: 10.1103/PhysRevLett.84.5493
Baidu ScholarGoogle Scholar
19. T. Otsuka, R. Fujimoto, Y. Utsuno, B. A. Brown, M. Honma, and T. Mizusaki,

Magic numbers in exotic nuclei and spin-isospin properties of the NN Interaction

. Phys. Rev. Lett. 87, 082502 (2001).doi: 10.1103/PhysRevLett.87.082502
Baidu ScholarGoogle Scholar
20. M. Rejmund, S. Bhattacharyya, A. Navin et al.,

Shell evolution and the N=34 ”magic number”

. Phys. Rev. C 76, 021304 (2007). doi: 10.1103/PhysRevC.76.021304
Baidu ScholarGoogle Scholar
21. M. Rosenbusch, P. Ascher, D. Atanasov et al.,

Probing the N=32 shell closure below the magic proton number Z=20: Mass measurements of the exotic isotopes 52,53K

. Phys. Rev. Lett. 114, 202501 (2015). doi: 10.1103/PhysRevLett.114.202501
Baidu ScholarGoogle Scholar
22. S. Chen, J. Lee, P. Doornenbal, et al.,

Quasifree neutron knockout from 54Ca corroborates arising N=34 neutron magic number

. Phys. Rev. Lett. 123, 142501 (2019). doi: 10.1103/PhysRevLett.123.142501
Baidu ScholarGoogle Scholar
23. X.-X. Sun, J. Zhao, and 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 (2018). doi: 10.1016/j.physletb.2018.08.071
Baidu ScholarGoogle Scholar
24. A. Zilges, M. Babilon, T. Hartmann et al.,

Collective excitations close to the particle threshold

. Prog. Part. Nucl. Phys. 55, 408 (2005). doi: 10.1016/j.ppnp.2005.01.018
Baidu ScholarGoogle Scholar
25. P. Adrich, A. Klimkiewicz, M. Fallot, et al.,

Evidence for pygmy and giant dipole resonances in 130Sn and 132Sn

. Phys. Rev. Lett. 95, 132501 (2005). doi: 10.1103/PhysRevLett.95.132501
Baidu ScholarGoogle Scholar
26. M. Arnould, S. Goriely, and K. Takahashi,

The r-process of stellar nucleosynthesis: Astrophysics and nuclear physics achievements and mysteries

. Phys. Rep. 450, 97 (2007). doi: 10.1016/j.physrep.2007.06.002
Baidu ScholarGoogle Scholar
27. J.-W. Xia, W.-L. Zhan, B.-W. Wei et al.,

The heavy ion cooler-storage-ring project (HIRFL-CSR) at Lanzhou

. Nucl. Instrum. Meth. A 488, 11 (2002). doi: 10.1016/S0168-9002(02)00475-8
Baidu ScholarGoogle Scholar
28. W.-L. Zhan, H.-S. Xu, G.-Q. Xiao et al.,

Progress in HIRFL-CSR

. Nucl. Phys. A 834, 694c (2010). doi: 10.1016/j.nuclphysa.2010.01.126
Baidu ScholarGoogle Scholar
29. C. Sturm, B. Sharkov, and H. Stocker,

1, 2, 3 … FAIR!

. Nucl. Phys. A 834, 682c (2010). doi: 10.1016/j.nuclphysa.2010.01.124
Baidu ScholarGoogle Scholar
30. S. Gales,

SPIRAL2 at GANIL: next generation of ISOL facility for intense secondary radioactive ion beams

. Nucl. Phys. A 834, 717c (2010). doi: 10.1016/j.nuclphysa.2010.01.130
Baidu ScholarGoogle Scholar
31. T. Motobayashi,

RIKEN RI beam factory-recent results and perspectives

. Nucl. Phys. A 834, 707c (2010). doi: 10.1016/j.nuclphysa.2010.01.128
Baidu ScholarGoogle Scholar
32. M. Thoennessen,

Plans for the facility for rare isotope beams

. Nucl. Phys. A 834, 688c (2010). doi: 10.1016/j.nuclphysa.2010.01.125
Baidu ScholarGoogle Scholar
33. X.H. Zhou,

Physics opportunities at the new facility HIAF

. Nucl. Phys. Rev. 35, 339 (2018). doi: 10.11804/NuclPhysRev.35.04.339
Baidu ScholarGoogle Scholar
34. X. Zhou, M. Wang, Y.-H. Zhang et al.,

Charge resolution in the isochronous mass spectrometry and the mass of 51Co

. Nucl. Sci. Tech. 32, 37 (2021). doi: 10.1007/s41365-021-00876-0
Baidu ScholarGoogle Scholar
35. Relativistic Density Functional for Nuclear Structure, edited by J. Meng, International Review of Nuclear Physics Vol. 10, (World Scientific, Singapore, 2016).
36. X.-B. Wei, H.-L. Wei, Y.-T. Wang et al.,

Multiple-models predictions for drip line nuclides in projectile fragmentation of 40,48Ca, 58,64Ni, and 78,86Kr at 140 MeV/u

. Nucl. Sci. Tech. 33, 155 (2022). doi: 10.1007/s41365-022-01137-4
Baidu ScholarGoogle Scholar
37. Y.-F. Gao, B.-S. Cai, and C.-X. Yuan,

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

. Nucl. Sci. Tech. 34, 9 (2023). doi: 10.1007/s41365-022-01153-4
Baidu ScholarGoogle Scholar
38. J. Dobaczewski, W. Nazarewicz, T. R. Werner et al.,

Mean-field description of ground-state properties of drip-line nuclei: Pairing and continuum effects

. Phys. Rev. C 53, 2809 (1996). doi: 10.1103/PhysRevC.53.2809
Baidu ScholarGoogle Scholar
39. J. Meng,

Relativistic continuum Hartree-Bogoliubov theory with both zero range and finite range Gogny force and their application

. Nucl. Phys. A 635, 3 (1998). doi: 10.1016/S0375-9474(98)00178-X
Baidu ScholarGoogle Scholar
40. M. Grasso, N. Sandulescu, N. Van Giai et al.,

Pairing and continuum effects in nuclei close to the drip line

. Phys. Rev. C 64, 064321 (2001). doi: 10.1103/PhysRevC.64.064321
Baidu ScholarGoogle Scholar
41. N. Sandulescu, L.S. Geng, H. Toki et al.,

Pairing correlations and resonant states in the relativistic mean field theory

. Phys. Rev. C 68, 054323 (2003). doi: 10.1103/PhysRevC.68.054323
Baidu ScholarGoogle Scholar
42. L.-G. Cao, and Z.-Y. Ma,

Effect of resonant continuum on pairing correlations in the relativistic approach

. Eur. Phys. J. A 22, 189 (2004). doi: 10.1140/epja/i2004-10029-5
Baidu ScholarGoogle Scholar
43. T. Nakamura, N. Kobayashi, Y. Kondo et al.,

Halo structure of the island of inversion nucleus 31Ne

. Phys. Rev. Lett. 103, 262501 (2009). doi: 10.1103/PhysRevLett.103.262501
Baidu ScholarGoogle Scholar
44. N. Kobayashi, T. Nakamura, Y. Kondo et al.,

Observation of a p-wave one-neutron halo configuration in 37Mg

. Phys. Rev. Lett. 112, 242501 (2014). doi: 10.1103/PhysRevLett.112.242501
Baidu ScholarGoogle Scholar
45. P. Ring and P. Schuck, The nuclear many-body problem (Springer Science & Business Media, 2004).
46. J. Dechargé and D. Gogny,

Hartree-Fock-Bogolyubov calculations with the D1 effective interaction on spherical nuclei

. Phys. Rev. C 21, 1568 (1980). doi: 10.1103/PhysRevC.21.1568
Baidu ScholarGoogle Scholar
47. J. Dobaczewski, H. Flocard, and J. Treiner,

Hartree-Fock-Bogolyubov description of nuclei near the neutron-drip line

. Nucl. Phys. A 422, 103 (1984). doi: 10.1016/0375-9474(84)90433-0
Baidu ScholarGoogle Scholar
48. W.-H. Long, P. Ring, N. V. Giai et al.,

Relativistic Hartree-Fock-Bogoliubov theory with density dependent meson-nucleon couplings

. Phys. Rev. C 81, 024308 (2010). doi: 10.1103/PhysRevC.81.024308
Baidu ScholarGoogle Scholar
49. L.-L. Li, J. Meng, P. Ring et al.,

Deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 85, 024312 (2012). doi: 10.1103/PhysRevC.85.024312
Baidu ScholarGoogle Scholar
50. Y. Chen, L.-L. Li, H.-Z. Liang et al.,

Density-dependent deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 85, 067301 (2012). doi: 10.1103/PhysRevC.85.067301
Baidu ScholarGoogle Scholar
51. J.-C. Pei, Y.-N. Zhang, and F.-R. Xu,

Evolution of surface deformations of weakly bound nuclei in the continuum

. Phys. Rev. C 87, 051302 (2013). doi: 10.1103/PhysRevC.87.051302
Baidu ScholarGoogle Scholar
52. Y.-N. Zhang, J.-C. Pei, and F.-R. Xu,

Hartree-Fock-Bogoliubov descriptions of deformed weakly bound nuclei in large coordinate spaces

. Phys. Rev. C 88, 054305 (2013). doi: 10.1103/PhysRevC.88.054305
Baidu ScholarGoogle Scholar
53. J.C. Pei, G.I. Fann, R.J. Harrison et al.,

Adaptive multi-resolution 3D Hartree-Fock-Bogoliubov solver for nuclear structure

. Phys. Rev. C 90, 024317 (2014). doi: 10.1103/PhysRevC.90.024317
Baidu ScholarGoogle Scholar
54. Y. Shi,

Precision of finite-difference representation in 3D coordinate-space Hartree-Fock-Bogoliubov calculations

. Phys. Rev. C 98, 014329 (2018). doi: 10.1103/PhysRevC.98.014329
Baidu ScholarGoogle Scholar
55. Y. Gambhir, P. Ring, and A. Thimet,

Relativistic mean field theory for finite nuclei

. Ann. Phys. 198, 132 (1990). doi: 10.1016/0003-4916(90)90330-Q
Baidu ScholarGoogle Scholar
56. S.-G. Zhou, J. Meng, and P. Ring,

Spherical relativistic Hartree theory in a Woods-Saxon basis

. Phys. Rev. C 68, 034323 (2003). doi: 10.1103/PhysRevC.68.034323
Baidu ScholarGoogle Scholar
57. M. V. Stoitsov, W. Nazarewicz, and S. Pittel,

New discrete basis for nuclear structure studies

. Phys. Rev. C 58, 2092 (1998). doi: 10.1103/PhysRevC.58.2092
Baidu ScholarGoogle Scholar
58. M. Stoitsov, J. Dobaczewski, W. Nazarewicz et al.,

Axially deformed solution of the Skyrme-Hartree-Fock-Bogolyubov equations using the transformed harmonic oscillator basis

. The program HFBTHO. Comput. Phys. Commun. 167, 43 (2005). doi: 10.1016/j.cpc.2005.01.001
Baidu ScholarGoogle Scholar
59. K.-Y. Zhang, M.-K. Cheoun, Y.-B. Choi, et al. (DRHBc collaboration),

Deformed relativistic Hartree-Bogoliubov theory in continuum with a point-coupling functional: Examples of even-even Nd isotopes

. Phys. Rev. C 102, 024314 (2020). doi: 10.1103/PhysRevC.102.024314
Baidu ScholarGoogle Scholar
60. S. Kim, M.-H. Mun, M.-K. Cheoun et al.,

Shape coexistence and neutron skin thickness of Pb isotopes by the deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 105, 034340 (2022). doi: 10.1103/PhysRevC.105.034340
Baidu ScholarGoogle Scholar
61. E. Tamura,

Relativistic single-site Green function for general potentials

. Phys. Rev. B 45, 3271 (1992). doi: 10.1103/PhysRevB.45.3271
Baidu ScholarGoogle Scholar
62. D. L. Foulis,

Partial-wave Green-function expansions for general potentials

. Phys. Rev. A 70, 022706 (2004). doi: 10.1103/PhysRevA.70.022706
Baidu ScholarGoogle Scholar
63. E.N. Economou, Green's Fucntion in Quantum Physics (Springer-Verlag, Berlin, 2006).
64. S.T. Belyaev, A.V. Smirnov, S.V. Tolokonnikov et al.,

Pairing in nuclei in the coordinate representation

. Sov. J. Nucl. Phys. 45, 783 (1987).
Baidu ScholarGoogle Scholar
65. M. Matsuo,

Continuum linear response in coordinate space Hartree–Fock–Bogoliubov formalism for collective excitations in drip-line nuclei

. Nucl. Phys. A 696, 371 (2001). doi: 10.1016/S0375-9474(01)01133-2
Baidu ScholarGoogle Scholar
66. M. Matsuo,

Collective excitations and pairing effects in drip-line nuclei: continuum RPA in coordinate-space HFB

. Prog. Theor. Phys. Suppl. 146, 110 (2002). doi: 10.1143/PTPS.146.110
Baidu ScholarGoogle Scholar
67. M. Matsuo, K. Mizuyama, and Y. Serizawa,

Di-neutron correlation and soft dipole excitation in medium mass neutron-rich nuclei near drip line

. Phys. Rev. C 71, 064326 (2005). doi: 10.1103/PhysRevC.71.064326
Baidu ScholarGoogle Scholar
68. M. Matsuo and Y. Serizawa,

Surface-enhanced pair transfer amplitude in quadrupole states of neutron-rich Sn isotopes

. Phys. Rev. C 82, 024318 (2010). doi: 10.1103/PhysRevC.82.024318
Baidu ScholarGoogle Scholar
69. H. Shimoyama and M. Matsuo,

Anomalous pairing vibration in neutron-rich Sn isotopes beyond the N=82 magic number

. Phys. Rev. C 84, 044317 (2011). doi: 10.1103/PhysRevC.84.044317
Baidu ScholarGoogle Scholar
70. H. Shimoyama and M. Matsuo,

Di-neutron correlation in monopole two-neutron transfer modes in the Sn isotope chain

. Phys. Rev. C 88, 054308 (2013). doi: 10.1103/PhysRevC.88.054308
Baidu ScholarGoogle Scholar
71. M. Matsuo,

Continuum quasiparticle random-phase approximation for astrophysical direct neutron capture reactions on neutron-rich nuclei

. Phys. Rev. C 91, 034604 (2015). doi: 10.1103/PhysRevC.91.034604
Baidu ScholarGoogle Scholar
72. H. Oba and M. Matsuo,

Continuum Hartree-Fock-Bogoliubov theory for weakly bound deformed nuclei using the coordinate-space Green's function method

. Phys. Rev. C 80, 024301 (2009). doi: 10.1103/PhysRevC.80.024301
Baidu ScholarGoogle Scholar
73. Y. Zhang, M. Matsuo, and J. Meng,

Persistent contribution of unbound quasiparticles to the pair correlation in the continuum Skyrme-Hartree-Fock-Bogoliubov approach

. Phys. Rev. C 83, 054301 (2011). doi: 10.1103/PhysRevC.83.054301
Baidu ScholarGoogle Scholar
74. Y. Zhang, M. Matsuo, and J. Meng,

Pair correlation of giant halo nuclei in continuum Skyrme-Hartree-Fock-Bogoliubov theory

. Phys. Rev. C 86, 054318 (2012). doi: 10.1103/PhysRevC.86.054318
Baidu ScholarGoogle Scholar
75. X. Qu and Y. Zhang,

Effects of mean-field and pairing correlations on the Bogoliubov quasiparticle resonance

. Sci. China-Phys. Mech. Astron. 62, 112012 (2019). doi: 10.1007/s11433-019-9409-y
Baidu ScholarGoogle Scholar
76. Y. Zhang and X.-Y. Qu,

Effects of pairing correlation on the quasiparticle resonance in neutron-rich Ca isotopes

. Phys. Rev. C 102, 054312 (2020). doi: 10.1103/PhysRevC.102.054312
Baidu ScholarGoogle Scholar
77. T.-T. Sun, Z.-X. Liu, L. Qian et al.,

Continuum Skyrme-Hartree-Fock-Bogoliubov theory with Green's function method for odd-A nuclei

. Phys. Rev. C 99, 054316 (2019). doi: 10.1103/PhysRevC.99.054316
Baidu ScholarGoogle Scholar
78. J. Meng and S.-G. Zhou,

Halos in medium-heavy and heavy nuclei with covariant density functional theory in continuum

. J. Phys. G Nucl. Phys. 42, 093101 (2015). doi: 10.1088/0954-3899/42/9/093101
Baidu ScholarGoogle Scholar
79. T.-T. Sun, E. Hiyama, H. Sagawa et al.,

Mean-field approaches for Ξ− hypernuclei and current experimental data

. Phys. Rev. C 94, 064319 (2016). doi: 10.1103/PhysRevC.94.064319
Baidu ScholarGoogle Scholar
80. W.-L. Lu, Z.-X. Liu, S.-H. Ren et al.,

(Pseudo)spin symmetry in the single-neutron spectrum of hypernuclei

. J. Phys. G Nucl. Phys. 44, 125104 (2017). doi: 10.1088/1361-6471/aa8e2d
Baidu ScholarGoogle Scholar
81. T.-T. Sun, W.-L. Lu, and S.-S. Zhang,

Spin and pseudospin symmetries in the single-Λ spectrum

. Phys. Rev. C 96, 044312 (2017). doi: 10.1103/PhysRevC.96.044312
Baidu ScholarGoogle Scholar
82. T.-T. Sun, C.-J. Xia, S.-S. Zhang et al.,

Massive neutron stars and -hypernuclei in relativistic mean field models

. Chin. Phys. C 42, 025101 (2018). doi: 10.1088/1674-1137/42/2/025101
Baidu ScholarGoogle Scholar
83. Z.-X. Liu, C.-J. Xia, W.-L. Lu et al.,

Relativistic mean-field approach for Λ,Ξ, and ∑ hypernuclei

. Phys. Rev. C 98, 024316 (2018). doi: 10.1103/PhysRevC.98.024316
Baidu ScholarGoogle Scholar
84. T.-T. Sun, S.-S. Zhang, Q.-L. Zhang et al.,

Strangeness and resonance in compact stars with relativistic-mean-field models

. Phys. Rev. D 99, 023004 (2019). doi: 10.1103/PhysRevD.99.023004
Baidu ScholarGoogle Scholar
85. C. Chen, Q.-K. Sun, Y.-X. Li et al.,

Possible shape coexistence in Ne isotopes and the impurity effect of Λ hyperon

. Sci. China-Phys. Mech. Astron. 64, 282011 (2021). doi: 10.1007/s11433-021-1721-1
Baidu ScholarGoogle Scholar
86. Y. Tanimura, H. Sagawa, T.-T. Sun et al.,

Ξ hypernuclei Ξ15C and Ξ12Be, and the ΞN two-body interaction

. Phys. Rev. C 105, 044324 (2022). doi: 10.1103/PhysRevC.105.044324
Baidu ScholarGoogle Scholar
87. T.-T. Sun, S.-Q. Zhang, Y. Zhang et al.,

Green's function method for single-particle resonant states in relativistic mean field theory

. Phys. Rev. C 90, 054321 (2014). doi: 10.1103/PhysRevC.90.054321
Baidu ScholarGoogle Scholar
88. T.-T. Sun, W.-L. Lu, L. Qian et al.,

Green's function method for the spin and pseudospin symmetries in the single-particle resonant states

. Phys. Rev. C 99, 034310 (2019). doi: 10.1103/PhysRevC.99.034310
Baidu ScholarGoogle Scholar
89. T.-T. Sun, Z.-M. Niu, and S.-Q. Zhang,

Single-proton resonant states and the isospin dependence investigated by Green's function relativistic mean field theory

. J. Phys. G Nucl. Phys. 43, 045107 (2016). doi: 10.1088/0954-3899/43/4/045107
Baidu ScholarGoogle Scholar
90. S.-H. Ren, T.-T. Sun, and W. Zhang,

Green's function relativistic mean field theory for Λ hypernuclei

. Phys. Rev. C 95, 054318 (2017). doi: 10.1103/PhysRevC.95.054318
Baidu ScholarGoogle Scholar
91. C. Chen, Z. P. Li, Y. X. Li et al.,

Green's function relativistic mean field theory for Λ hypernuclei

. Chin. Phys. C 44, 084105 (2020). doi: 10.1088/1674-1137/44/8/084105
Baidu ScholarGoogle Scholar
92. Y.-T. Wang and T.-T. Sun,

Searching for single-particle resonances with the Green's function method

. Nucl. Sci. Tech. 32, 46 (2021). doi: 10.1007/s41365-021-00884-0
Baidu ScholarGoogle Scholar
93. T.-T. Sun,

Green's function method in covariant density functional theory

. Sci. Sin.-Phys. Mech. Astron. 46, 12006 (2016). doi: 10.1360/SSPMA2015-00371
Baidu ScholarGoogle Scholar
94. T.-T. Sun, L. Qian, C. Chen et al.,

Green's function method for the single-particle resonances in a deformed Dirac equation

. Phys. Rev.C 101, 014321 (2020). doi: 10.1103/PhysRevC.101.014321
Baidu ScholarGoogle Scholar
95. J. Carbonell, A. Deltuva, A.C. Fonseca et al.,

Bound state techniques to solve the multiparticle scattering problem

. Prog. Part. Nucl. Phys. 74, 55 (2014). doi: 10.1016/j.ppnp.2013.10.003
Baidu ScholarGoogle Scholar
96. M. Shi, J.-G. Guo, Q. Liu et al.,

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

. Phys. Rev. C 92, 054313 (2015). doi: 10.1103/PhysRevLett.55.2676
Baidu ScholarGoogle Scholar
97. X.-X. Shi, M. Shi, Z.-M. Niu et al.,

Probing resonances in deformed nuclei by using the complex-scaled Green's function method

. Phys. Rev. C 94, 024302 (2016). doi: 10.1103/PhysRevC.94.024302
Baidu ScholarGoogle Scholar
98. M. Shi, Z.-M. Niu, and H.-Z. Liang,

Combination of complex momentum representation and Green's function methods in relativistic mean-field theory

. Phys. Rev. C 97, 064301 (2018). doi: 10.1103/PhysRevC.97.064301
Baidu ScholarGoogle Scholar
99. M. Wang, W.-J. Huang, F. Kondev et al.,

The AME 2020 atomic mass evaluation (II). Tables, graphs and references*

. Chin. Phys. C 45, 030003 (2021). doi: 10.1088/1674-1137/abddaf
Baidu ScholarGoogle Scholar
100. E. Chabanat, P. Bonche, P. Haensel et al.,

A Skyrme parametrization from subnuclear to neutron star densities Part II

. Nuclei far from stabilities. Nucl. Phys. A 635, 231 (1998). doi: 10.1016/S0375-9474(98)00180-8
Baidu ScholarGoogle Scholar
101. M. Matsuo,

Spatial structure of neutron cooper pair in low density uniform matter

. Phys. Rev. C 73, 044309 (2006). doi: 10.1103/PhysRevC.73.044309
Baidu ScholarGoogle Scholar
102. M. Matsuo, Y. Serizawa, and K. Mizuyama,

Pairing collectivity in medium-mass neutron-rich nuclei near drip-line

. Nucl. Phys. A 788, 307 (2007). doi: 10.1016/j.nuclphysa.2007.01.017
Baidu ScholarGoogle Scholar
103. X.-W. Xia, Y. Lim, P.-W. Zhao et al.,

The limits of the nuclear landscape explored by the relativistic continuum Hartree-Bogoliubov theory

. At. Data Nucl. Data Tables 121-122, 1 (2018). doi: 10.1016/j.adt.2017.09.001
Baidu ScholarGoogle Scholar
104. X.-Y. Qu and Y. Zhang,

Canonical states in continuum Skyrme Hartree-Fock-Bogoliubov theory with Green's function method

. Phys. Rev. C 99, 014314 (2019). doi: 10.1103/PhysRevC.99.014314
Baidu ScholarGoogle Scholar
105. X.Y. Qu, H. Tong, and S.Q. Zhang,

Canonical states in relativistic continuum theory with the Green's function method: Neutrons in continuum of Zirconium giant-halo nuclei

. Phys. Rev. C 105, 014326 (2022). doi: 10.1103/PhysRevC.105.014326
Baidu ScholarGoogle Scholar
106. W. Satuła, J. Dobaczewski, and W. Nazarewicz,

Odd-even staggering of nuclear masses: Pairing or shape effect

. Phys. Rev. Lett. 81, 3599 (1998). doi: 10.1103/PhysRevLett.81.3599
Baidu ScholarGoogle Scholar
107. J. Dobaczewski, P. Magierski, W. Nazarewicz et al.,

Odd-even staggering of binding energies as a consequence of pairing and mean-field effects

. Phys. Rev. C 63, 024308 (2001). doi: 10.1103/PhysRevC.63.024308
Baidu ScholarGoogle Scholar
108. Y. A. Litvinov, T. J. Bürvenich, H. Geissel et al.,

Isospin dependence in the odd-even staggering of nuclear binding energies

. Phys. Rev. Lett. 95, 042501 (2005). doi: 10.1103/PhysRevLett.95.042501
Baidu ScholarGoogle Scholar
109. K. Hagino and H. Sagawa,

Pairing correlations and odd-even staggering in reaction cross sections of weakly bound nuclei

. Phys. Rev. C 85, 014303 (2012). doi: 10.1103/PhysRevC.85.014303
Baidu ScholarGoogle Scholar
110. L. J. Wang, B. Y. Sun, J. M. Dong et al.,

Odd-even staggering of the nuclear binding energy described by covariant density functional theory with calculations for spherical nuclei

. Phys. Rev. C 87, 054331 (2013). doi: 10.1103/PhysRevC.87.054331
Baidu ScholarGoogle Scholar
111. L. Coraggio, A. Covello, A. Gargano et al.,

Behavior of odd-even mass staggering around 132Sn

. Phys. Rev. C 88, 041304 (2013). doi: 10.1103/PhysRevC.88.041304
Baidu ScholarGoogle Scholar
112. W.J. Chen, C.A. Bertulani, F.R. Xu et al.,

Odd-even mass staggering with Skyrme-Hartree-Fock-Bogoliubov theory

. Phys. Rev. C 91, 047303 (2015). doi: 10.1103/PhysRevC.91.047303
Baidu ScholarGoogle Scholar
113. J. Meng and P. Ring,

Giant halo at the neutron drip line

. Phys. Rev. Lett. 80, 460 (1998). doi: 10.1103/PhysRevLett.80.460
Baidu ScholarGoogle Scholar
114. M. Grasso, S. Yoshida, N. Sandulescu, and N. Van Giai,

Giant neutron halos in the non-relativistic mean field approach

. Phys. Rev. C 74, 064317 (2006). doi: 10.1103/PhysRevC.74.064317
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.