logo

Benchmarking nuclear energy density functionals with new mass data

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Benchmarking nuclear energy density functionals with new mass data

Xiao-Ying Qu
Kang-Min Chen
Cong Pan
Yang-Yang Yu
Kai-Yuan Zhang
Nuclear Science and TechniquesVol.36, No.12Article number 231Published in print Dec 2025Available online 29 Sep 2025
11901

Nuclear masses play a crucial role in both nuclear physics and astrophysics, driving sustained efforts toward precise experimental determination and reliable theoretical predictions. In this study, we compiled the newly measured masses for 296 nuclides from 40 references published between 2021 and 2024, subsequent to the release of the latest Atomic Mass Evaluation. These data were used to benchmark the performance of several relativistic and non-relativistic density functionals, including PC-PK1, TMA, SLy4, SV-min, UNEDF1, and the recently proposed PC-L3R. The results for PC-PK1 and PC-L3R were obtained using the state-of-the-art deformed relativistic Hartree-Bogoliubov theory in continuum (DRHBc), whereas the others were adopted from the existing literature. It was found that the DRHBc calculations with PC-PK1 and PC-L3R achieved an accuracy better than 1.5 MeV, outperforming the other functionals, which all exhibited root-mean-square deviations exceeding 2 MeV. The odd-even effects and isospin dependence in these theoretical descriptions were examined. The PC-PK1 and PC-L3R descriptions were qualitatively similar, exhibiting robust isospin dependence along the isotopic chains. Finally, a quantitative comparison between the PC-PK1 and PC-L3R results is presented, with the largest discrepancies analyzed in terms of the potential energy curves from the constrained DRHBc calculations.

Nuclear massDensity functional theoryDeformed relativistic Hartree-Bogoliubov theory in continuumPC-PK1PC-L3R
1

Introduction

The nuclear mass or binding energy reflects complex nuclear forces that bind protons and neutrons together within a nucleus [1]. This fundamental quantity not only underlies nuclear stability [2] but also critically influences astrophysical phenomena, from nuclear reactions in stellar interiors [3] to the nucleosynthesis processes responsible for elemental production in the universe [4]. Consequently, the precise determination of nuclear masses is indispensable for advancing our understanding of nuclear structures [5] and has significant implications for nuclear astrophysics [6-8]. Thus, improved experimental precision and theoretical accuracy in nuclear mass evaluations not only deepens insights into fundamental research in nuclear physics [9] but also fosters progress in nuclear energy applications via both fusion and fission.

Global investments in rare isotope beam facilities–including the Heavy Ion Research Facility in Lanzhou (HIRFL) [10] and the High Intensity heavy-ion Accelerator Facility (HIAF) at Huizhou [11], China, the Facility for Rare Isotope Beams (FRIB) in the USA [12], the Radioactive Isotope Beam Factory (RIBF) at RIKEN, Japan [13], the Facility for Antiproton and Ion Research (FAIR) in Germany [14], the Rare isotope Accelerator complex for ON-line experiments (RAON) in Korea [15], and Isotope Separator and Accelerator in Canada (ISAC) [16]–have substantially advanced the production, identification, and investigation of nuclides far from the valley of stability. To date, experimental efforts have led to the identification of over 3300 nuclides [17], with mass measurements available for approximately 2500 of these [18-20]. By contrast, theoretical models predict the existence of approximately 7000–10000 nuclides [21, 22]. Given that the proton dripline has been established for isotopes with proton numbers Z90 [23], whereas the neutron dripline has been delineated only up to Z=10 [24], it is anticipated that most unknown neutron-rich nuclei will remain experimentally inaccessible in the near future. Therefore, there is an urgent need for reliable theoretical predictions of the nuclear masses.

Extensive efforts have been devoted to reproducing measured nuclear masses and predicting those that are yet uncharted. Macroscopic-microscopic approaches exemplified by the finite-range droplet model (FRDM) [25] and the Weizsäcker-Skyrme (WS) model [26, 27] have achieved impressive accuracy in describing existing mass data; however, microscopic theories are widely accepted as offering superior predictive capabilities [28, 29]. In this context, density functional theory has emerged as a powerful framework for a unified description of nearly all nuclides across the nuclear chart [30-38]. Its relativistic extension, the covariant density functional theory (CDFT) [39], has been exceptionally successful in describing a variety of nuclear phenomena in both ground and excited states [39-48]. This success is largely attributable to the inherent advantages of CDFT, including the automatic incorporation of spin-orbit coupling [49, 50], natural explanation of pseudospin symmetry in the nucleon spectrum [51-53], spin symmetry in the antinucleon spectrum [53-55], and self-consistent treatment of nuclear magnetism [56, 57].

Within the framework of CDFT, the pairing correlations and continuum effects are taken into account self-consistently in the relativistic continuum Hartree-Bogoliubov (RCHB) theory [58, 59], making it capable of describing both stable and exotic nuclei [58, 60-65]. A pioneering application of RCHB theory is the construction of the first relativistic nuclear mass table incorporating continuum effects, in which the existence of 9035 bound nuclei with 8Z120 is predicted [22]. Notably, the inclusion of continuum effects is essential for extending the neutron dripline to a more neutron-rich region. Nonetheless, the accuracy of the RCHB mass table in reproducing experimental data is limited owing to the assumption of spherical symmetry within the theoretical framework.

Thus, it is natural to propose an upgraded mass table that incorporates not only continuum effects but also nuclear deformation degrees of freedom. This can be realized by employing the deformed extension of the RCHB theory, that is, the deformed relativistic Hartree-Bogoliubov theory in continuum (DRHBc) [66-69]. Axial deformation, pairing correlations, and continuum effects are considered microscopically and self-consistently in the DRHBc theory, which lays an important foundation for its great success [45, 70, 71]. In pursuit of a high-precision mass table [72], a point-coupling version of the DRHBc theory was developed [73, 74] for combination with the density functional PC-PK1 [75], which is probably the most successful density functional for describing nuclear masses [37, 44, 76]. The DRHBc mass table project, now in progress for over six years, has successfully completed the sectors for even-even [77] and even-Z [78] nuclei. Impressively, the root-mean-square (RMS) deviation of the DRHBc calculated masses from the latest Atomic Mass Evaluation (AME2020) data is approximately 1.5 MeV, positioning it among the most accurate density-functional descriptions for nuclear masses. Moreover, lots of relevant studies on halo phenomena [79-89], nuclear charge radii [90-92], shape evolution [93-96], shell structure [97-103], decay properties [104-107], and other topics [108-113] based on the DRHBc mass table underscore its value as a resource that extends far beyond a mere data repository [114].

In this work, inspired by the recent progress in nuclear mass measurements that provide new data beyond AME2020 or reduce the uncertainties of existing data, we further examined the predictive power of the DRHBc mass table using the new mass data. On the theoretical side, a new point-coupling density functional, PC-L3R, has recently been proposed, whose performance is even better than that of PC-PK1 in describing the masses of spherical nuclei [115]. Our second motivation is to test the accuracy of PC-L3R in describing the masses of deformed nuclei when combined with DRHBc theory. The remainder of this paper is organized as follows. The point-coupling DRHBc theory, relativistic density functionals PC-PK1 and PC-L3R, and numerical details are introduced in Sect. 2. DRHBc descriptions with PC-PK1 and PC-L3R for the new masses are presented and compared with those from other density functionals in Sect. 3. Finally, a summary is given in Sect. 4.

2

Theoretical framework

The point-coupling density functional theory starts from the Lagrangian density, L=ψ¯(iγμμM)ψ12αS(ψ¯ψ)(ψ¯ψ)12αV(ψ¯γμψ)(ψ¯γμψ)12αTV(ψ¯τγμψ)(ψ¯τγμψ)12αTS(ψ¯τψ)(ψ¯τψ)13βS(ψ¯ψ)314γS(ψ¯ψ)414γV[ (ψ¯γμψ)(ψ¯γμψ) ]212δSν(ψ¯ψ)ν(ψ¯ψ)12δVν(ψ¯γμψ)ν(ψ¯γμψ)12δTVν(ψ¯τγμψ)ν(ψ¯τγμψ)12δTSν(ψ¯τψ)ν(ψ¯τψ)14FμνFμνeψ¯γμ1τ32Aμψ, (1) where M is the nucleon mass, e the charge unit, and Aμ and Fμν the four-vector potential and field strength tensor of the electromagnetic field, respectively. With the subscripts S, V, and T respectively standing for scalar, vector, and isovector, nine coupling constants, αS, αV, αTV, βS, γS, γV, δS, δV, and δTV, in the Lagrangian density of PC-PK1 and PC-L3R are listed in Table 1. As the isovector-scalar channels involving αTS and δTS terms were found to be less helpful in improving the description of nuclear ground-state properties [116], they are not included in PC-PK1 and PC-L3R.

Table 1
Coupling constants of the relativistic density functionals PC-PK1 [75] and PC-L3R [115]
Coupling constant PC-PK1 PC-L3R
αS (MeV-2) -3.96291×10-4 -3.99289×10-4
βS (MeV-5) 8.6653×10-11 8.65504×10-11
γS (MeV-8) -3.80724×10-17 -3.83950×10-17
δS (MeV-4) -1.09108×10-10 -1.20749×10-10
αV (MeV-2) 2.6904×10-4 2.71991×10-4
γV (MeV-8) -3.64219×10-18 -3.72107×10-18
δV (MeV-4) -4.32619×10-10 -4.26653×10-10
δTV (MeV-2) 2.95018×10-5 2.96688×10-5
δTV (MeV-4) -4.11112×10-10 -4.65682×10-10
Show more

Starting from the Lagrangian density (1), the Hamiltonian can be derived via the quantization of the Dirac spinor field in the Bogoliubov quasiparticle space, and the energy functional can be constructed as its expectation with respect to the Bogoliubov ground state. The relativistic Hartree-Bogoliubov equation obtained by performing the variation of the energy density functional with respect to the generalized density matrix and neglecting the exchange terms reads (h^DλΔ^Δ^*h^D*+λ)(UkVk)=Ek(UkVk), (2) where h^D is the Dirac Hamiltonian, Δ^ is the pairing field, λ is the Fermi surface, Ek is the quasiparticle energy, and Uk and Vk are quasiparticle wave functions. The Dirac Hamiltonian in coordinate space is hD(r)=αp+V(r)+β[M+S(r)], (3) with the scalar S(r) and vector V(r) potentials, S(r)=αSρS+βSρS2+γSρS3+δSΔρS,V(r)=αVρV+γVρV3+δVΔρV+eA0+αTVτ3ρ3 +δTVτ3Δρ3, (4) constructed by various densities, ρS(r)=k>0Vk(r)γ0Vk(r),ρV(r)=k>0Vk(r)Vk(r),ρ3(r)=k>0Vk(r)τ3Vk(r), (5) where the summation index k loops through the positive-energy quasiparticle states in the Fermi sea. Neglecting for simplicity spin and isospin degrees of freedom, the pairing potential reads Δ(r1,r2)=Vpp(r1,r2)κ(r1,r2), (6) where Vpp is the pairing interaction, and κ is the pairing tensor. A density-dependent interaction of zero range is adopted in the present DRHBc theory, Vpp(r1,r2)=V012(1Pσ)δ(r1r2)(1ρ(r1)ρsat), (7) where V0 is the pairing strength, ρsat is the saturation density of nuclear matter, and 12(1Pσ) projects onto the spin-zero S=0 component. Assuming axial symmetry and spatial reflection symmetry, one can expand nuclear densities and potentials in terms of even-order Legendre polynomials [117], f(r)=lfl(r)Pl(cosθ), l=0,2,4,, (8) where the lth-order radial component is calculated by fl(r)=2l+14πdΩf(r)Pl(cosθ). (9) After obtaining the self-consistent solution of the RHB Eq. (2), quantities including the total energy, RMS radius, and deformation parameter can be calculated from nuclear densities. The total energy is [59] ERHB=k>0(λEk)vk2Epaird3r(12αSρS2+12αVρV2+12αTVρ32+23βSρS3+34γSρS4+34γVρV4+12δSρSΔρS+12δVρVΔρV+12δTVρ3Δρ3+12ρpeA0)+Ec.m.+Erot, (10) where vk2=d3rVk(r)Vk(r), (11) and the pairing energy Epair=12d3rκ(r)Δ(r). (12) The center-of-mass (c.m.) correction energy is calculated as Ec.m.=12MAP^2, (13) with A the mass number and P^=iAp^i the total momentum in the c.m. frame. The rotational correction energy for a deformed nucleus from the cranking approximation reads Erot=12JJ^2, (14) where J^=iAj^i is the total angular momentum and J is the moment of inertia that can be estimated using the Inglis-Belyaev formula [118]. The RMS radius is calculated as RRMS=r21/2=d3r[r2ρV(r)]N, (15) where ρV is the vector density and N denotes the corresponding particle number. The quadrupole deformation parameter is calculated as β2=4πr2Y20(Ω)3Nr2, (16) where Y is the spherical harmonic function.

The calculations in this study were performed using the same numerical details as those used to construct the DRHBc mass table [73, 74, 77]. Specifically, the pairing strength V0=-325 MeV fm3, saturation density ρsat = 0.152 fm-3, and pairing window was set to 100 MeV. The Dirac Woods-Saxon basis space was determined by an energy cutoff of Ecut=300 MeV and an angular momentum cutoff of Jcut=232 . The Legendre expansion truncations in Eq. (8) are chosen as lmax=6 and 8 for the nuclei with 8Z70 and 71Z100, respectively. The blocking effects in odd-mass and odd-odd nuclei are included via an equal filling approximation [68, 74, 119].

3

Results and discussion

We have collected the newly measured masses of 296 nuclides from 40 references [120-159], published between 2021 and 2024 (subsequent to the release of AME2020) and summarized in Table 2. The sources and measurement methods for the new experimental data are presented in Table 3. The corresponding mass values in AME2020, which contained 241 experimentally measured values and 55 extrapolated empirical values (labeled #), are listed in Table 2 for comparison. The differences between the new data and the AME2020 values are also given in Table 2 and are scaled by colors in Fig. 1. The nuclides with only empirical values in AME2020 are highlighted by black squares in Fig. 1. Among these 296 mass data, 247 in AME2020 are consistent with new measurements with deviations smaller than 0.15 MeV. The RMS deviation between the new and AME2020 experimental data is σ=0.0984 MeV, and after including the empirical values in AME2020, σ becomes 0.1178 MeV.

Fig. 1
(Color online) The differences between the newly measured masses for 296 nuclides and the corresponding values in AME2020 [20] scaled by colors. The black squares represent nuclides for which AME2020 has only empirical mass values. The dark gray region shows the nuclides observed experimentally, and the light gray region shows the predicted nuclear landscape by the DRHBc mass table [77]
pic

Most of the deviations between the new and AME2020 experimental data lie within experimental uncertainties. Nevertheless, even after considering the experimental uncertainties, disagreement still arises for 73 nuclides, which are labeled in bold in Table 2. Among these 73 data, the smallest difference, 0.00055 MeV, occurs between the upper bound of the new data and the lower bound of the AME2020 data for 111Ag, while the largest one, 0.29516 MeV, occurs between the lower bound of the new data and the upper bound of the AME2020 data for 67Fe. It is important to note that newly measured masses are not necessarily more accurate than those in previous evaluations. Systematic biases or experimental uncertainties may affect the measured values depending on the specific setup and techniques employed. Notably, for 23 nuclides, the central values of the newly reported masses deviated from those in AME2020 by more than 200 keV. Among the 23 nuclides exhibiting discrepancies exceeding 200 keV, 12 cases (23Si, 74Ni, 86Ge, 89As, 91Se, 70Kr, 104Sr, 105Sr, 109Nb, 72Tc, 151La, and 151Yb) show overlapping uncertainty ranges between the new measurements and AME2020 values, indicating potential consistency within experimental uncertainties. For 5 nuclides (69Fe, 60Ga, 88As, 66Se, and 80Zr), while the central value discrepancies also exceed 200 keV and uncertainty ranges do not overlap, the AME2020 masses are extrapolated values. Although such empirical estimates, based on trends in the mass surface and available experimental constraints, are often validated by subsequent measurements, deviations from true mass values may arise, for example, for nuclides exhibiting abrupt changes in shell structure. Finally, for 6 nuclides (28S, 67Fe, 71Kr, 75Sr, 96Ag, and 153Pm), the central values differ by more than 200 keV, the uncertainty ranges do not overlap, and both the new and AME2020 masses are based on experimental data. The cases are compared in Fig. 2, along with earlier measurements referenced in AME2020. The observed discrepancies may arise from differences in the measurement techniques. For 28S, its mass was derived from an indirect measurement in 1982 [160], whereas a recent result was obtained from a direct measurement using Bρ-defined isochronous mass spectrometry (Bρ-IMS) in 2024 [121]. For 67Fe, AME2020 mainly adopted the ion trap data from 2020 [161], which fall within the uncertainty range of contemporaneous Bρ-time-of-flight (Bρ-TOF) measurements [162]. However, the 2022 result obtained using a multiple-reflection TOF mass spectrometer (MR-TOF-MS) [130] supports earlier data from the TOF isochronous spectrometer from 2011 [163] rather than that from 1994 [164]. For 71Kr, the AME2020 value was consistent with storage-ring IMS data [165]. While the new Bρ-IMS result from 2023 [133] shows deviations, it remains largely within the uncertainty range of the results inferred from the electron capture decay energy QEC measurements [166]. For 75Sr and 96Ag, the AME2020 values are consistent with earlier estimates based on QEC [167, 168], but deviate from recent measurements obtained via Bρ-IMS [133] and ion trap techniques [147], respectively. For 153Pm, AME2020 primarily adopted ion trap data from 2012 [169]. Nonetheless, discrepancies were observed among the 2012 results, earlier estimates based on β- decay energy from 1993 [170], and the latest measurement from MR-TOF-MS in 2024 [155]. Such discrepancies necessitate careful data evaluation and/or even further measurements. In this study, we adopt the newly measured masses for convenience in the following examination of the theoretical descriptions.

Fig. 2
The newly measured binding energies of 28S, 67Fe, 71Kr, 75Sr, 96Ag, and 153Pm, in comparison with the AME2020 data and the previous data referenced in the AME
pic

The DRHBc calculations for the 296 nuclides were performed with the density functionals PC-PK1 [75] and PC-L3R [115], and the deviations of the resulting nuclear masses from the experimental data are plotted in Figs. 3(a) and 3(b). For comparison, we also show the mass differences between the new data and the results from the relativistic mean-field plus Bardeen-Cooper-Schrieffer (RMF+BCS) calculations with TMA [32] and the non-relativistic Skyrme Hartree-Fock-Bogoliubov (HFB) calculations [171] with SLy4 [172], SV-min [173], and UNEDF1 [174], respectively, in Figs. 3(c)–(f), respectively. It can be seen that both the DRHBc calculations with PC-PK1 and PC-L3R reproduce the data fairly well within a deviation of 3 MeV, despite a few exceptions. The RMF+BCS calculations with TMA can achieve a similar level of accuracy for nuclides with A<150, but for heavier nuclides, an overestimation of up to 6 MeV arises. In contrast, the Skyrme HFB calculations with SLy4 significantly underestimated the data in the heavy-mass region, with deviations for several nuclides above 10 MeV. The results from the Skyrme HFB calculations with SV-min also exhibited a certain underestimation in the heavy mass region. Although the Skyrme HFB results with UNEDF1 improve the description in A170, they apparently underestimate the masses of a few light and medium-mass nuclei and show significant overestimation at A155. The comparisons in Fig. 3 demonstrate that the descriptions of DRHBc using PC-PK1 and PC-L3R are qualitatively superior.

Fig. 3
(Color online) Mass differences between new data listed in Table 2 and the values from DRHBc calculations with density functionals PC-PK1 (a), PC-L3R (b), RMF+BCS calculations with density functional TMA [32] (c), as well as Skyrme HFB calculations [171] with density functionals SLy4 (d), SV-min (e) and UNEDF1 (f)
pic

For further comparison, we show in Fig. 4 the RMS deviations between the 296 new mass data points and the above theoretical results. The RMS deviations for even-even, odd-mass, and odd-odd nuclei are also computed and presented separately in Fig. 4. It can be found that both the DRHBc descriptions with PC-PK1 and PC-L3R can achieve accuracies better than 1.5 MeV for all datasets, with the exception of DRHBc+PC-L3R for even-even nuclei. In contrast, the accuracies in the other four density functional descriptions were generally worse than 2 MeV. Overall, the odd-even effects on the accuracy are not very significant in the DRHBc results, with a slightly better description for odd-odd nuclei. However, this is not the case for the RMF+BCS description, which obviously deteriorates for odd-odd nuclei. Furthermore, DRHBc descriptions with PC-PK1 and PC-L3R for odd nuclei are expected to be improved by strictly incorporating nuclear magnetism [87]. Instead of self-consistent calculations, Skyrme HFB results for odd nuclei were obtained from interpolations using the masses and average pairing gaps of neighboring even-even nuclei [171]. As expected, the Skyrme HFB descriptions with SV-min and UNEDF1 showed marginal odd-even differences. In contrast, it seems strange that from even-even to odd-mass, and then to odd-odd nuclei, the SLy4 description gradually improves. It should also be noted that the number of mass data points here is not large enough to confirm whether the odd-even features observed in these theoretical results are common across the nuclear chart. Finally, the accuracies in describing the 296 new masses–1.35, 2.04, 3.95, 2.37, and 2.21 MeV for PC-PK1, TMA, SLy4, SV-min, and UNEDF1, respectively–are found to be generally consistent with those obtained for all available masses of even-Z nuclei: 1.43 MeV for PC-PK1, 2.06 MeV for TMA, 5.28 MeV for SLy4, 3.39 MeV for SV-min, and 1.93 MeV for UNEDF1 [78]. Moreover, even within the spherical RHB framework, PC-PK1 and PC-L3R are the only two relativistic density functionals that reproduce the experimental masses with RMS deviations below 8 MeV [22, 175]. Given the superiority of PC-PK1 and PC-L3R, a complete DRHBc mass table including both even-Z and odd-Z nuclei is desirable in the near future, and further large-scale DRHBc+PC-L3R calculations are worth pursuing.

Fig. 4
(Color online) The RMS deviations between the new mass data listed in Table 2 and different theoretical results. The RMS deviations for even-even, odd-mass and odd-odd nuclei are presented separately
pic

One can see from Fig. 3 that the superiority of PC-PK1 and PC-L3R is mainly due to the better description of nuclei with A>150 compared to other density functionals. Therefore, a detailed comparison of the isospin dependence of nuclear masses in this region is necessary. In Fig. 5, the mass differences between the theoretical and experimental values are presented for even-Z nuclei with 70Z80. If the masses of certain nuclei located in the middle of an isotopic chain were absent from the dataset of new measurements, we resorted to AME2020 for completeness. It can be found in Fig. 5 that only the accuracy of the DRHBc description in this region is always better than 2 MeV, whereas other density-functional descriptions show systematic deviations from the data. Furthermore, the DRHBc description along these isotopic chains is almost steady, with a slight, approximately linear isospin dependence, which is in contrast to many obvious staggering behaviors by other descriptions. Another feature, observed in Fig. 5 shows that the nucleus with N=82 and Z=70 is basically described as overbound compared to its neighboring isotopes. This is a well-known weakness of both nonrelativistic and relativistic density functional theories in describing magic nuclei [77]. From the above discussions, it can be concluded that DRHBc theory provides not only an overall high accuracy but also a robust description of isospin dependence for nuclear masses, and the results from PC-PK1 and PC-L3R are qualitatively similar.

Fig. 5
(Color online) Same as Fig. 3 but for the isotopic chains with Z=70, 72, 74, 76, 78, and 80
pic

For a quantitative comparison, the differences between the DRHBc results for PC-PK1 and PC-L3R for binding energies ΔEB=EBPCPK1EBPCL3R, RMS matter radii ΔRm=RmPCPK1RmPCL3R, and quadrupole deformations Δβ2=β2PCPK1β2PCL3R are shown in Fig. 6. For the binding energies in Fig. 6(a), among these 296 nuclei, the ΔEB of 232 nuclei are located within -1.0 < ΔEB < 1.0 MeV, whereas the ΔEB of 64 nuclei are located within 1.0 < ΔEB < 2.0 MeV. Most values of ΔEB are positive, indicating that PC-PK1 generally describes these nuclei as being more bound than PC-L3R. The RMS matter radii shown in Fig. 6(b) reveal a clear trend of decreasing ΔRm with increasing mass number A with only a few exceptions. Notably, the majority of ΔRm values are negative, which is consistent with the general expectation that more strongly bound systems exhibit more compact density distributions. For most nuclei, ΔRm values are confined within ± 0.008 fm, and two outliers emerge: one barely beyond -0.008 fm for 152Pr and the other reaching -0.031 fm for 111Mo. For the quadrupole deformation shown in Fig. 6(c), the Δβ2 values in 291 nuclei are within 0.01, and almost all other nuclei show slightly larger values within 0.02, except for 83Zr and 111Mo. Δβ2 for 83Zr is only -0.021, whereas that for 111Mo reaches 0.193, which corresponds to the large |ΔRm| shown in Fig. 6(b).

Fig. 6
(Color online) Differences between the DRHBc results with PC-PK1 and PC-L3R for binding energies ΔEB=EBPCPK1EBPCL3R (a), RMS matter radii ΔRm=RmPCPK1RmPCL3R (b) and quadrupole deformations Δβ2=β2PCPK1β2PCL3R (c) as functions of mass number A
pic

To understand the large deviations between the PC-PK1 and PC-L3R results for 111Mo, in Fig. 7, the potential energy curves (PECs) of 111Mo from the constrained calculations with PC-PK1 and PC-L3R are shown. For comparison, the corresponding results for the neighboring isotope 112Mo are also shown. The ground state is shown with the filled square (circle) from the calculations with PC-PK1 (PC-L3R). As demonstrated in Refs. [73-75], for PC-PK1, the rotational correction plays an important role in improving the mass description of the deformed nuclei. Therefore, the ground state, after including the rotational correction energy Erot is also presented with an open symbol. Considering that the cranking approximation adopted to calculate Erot in Eqs. (14) is not suitable for spherical and weakly deformed nuclei, we only calculate Erot when |β2|>0.05 and take it as zero when |β2|<0.05. A more appropriate treatment for the correction energies in the nuclei with |β2|<0.05 can be achieved using the collective Hamiltonian method [176] in future work.

Fig. 7
(Color online) Potential energy curves (PECs) of 111Mo (a) and 112Mo (b) in constrained DRHBc calculations with PC-PK1 (red line) and PC-L3R (blue line). The ground state is shown with the filled square (circle) from the calculations with PC-PK1 (PC-L3R). For the case with |β2|>0.05, the ground state after including the rotational correction energy Etot is shown with the open symbol. The green line represents the available data of EBExp
pic

For 111Mo in Fig. 7(a), in both PECs from the calculations with PC-PK1 and PC-L3R, there are three minima, that is, one near-spherical minimum and two well-deformed minima. In each curve, the difference between the oblate and near-spherical minima is within 0.25 MeV, whereas the excitation energy of the prolate minimum is approximately 2 MeV. The ground state from the PC-PK1 calculations is the near-spherical minimum with β2=-0.024, whereas the ground state from the PC-L3R calculations is the oblate minimum with β2=-0.216. This corresponds to the large Δβ2 in Fig. 6(c). In addition, for the near-spherical ground state from PC-PK1, the rotational correction energy is taken as zero. For the well-deformed ground state from PC-L3R, the rotational correction energy is Erot=2.06 MeV, and the result becomes more deeply bound than that from PC-PK1. Finally, the binding energies from the PC-PK1 and PC-L3R calculations are similar, both reproducing the experimental value for 111Mo.

For 112Mo, the shape of the PEC is similar to that for 111Mo, still exhibiting three minima. However, here the ground states from both PC-PK1 and PC-L3R are the spherical minima, while the excitation energies of the oblate and prolate minima are about 0.5 and 3 MeV, respectively. In both results from PC-PK1 and PC-L3R, the significant differences in deformations and the small differences in total energies indicate the shape coexistence in 111,112Mo.

4

Summary

In this study, the newly measured masses for 296 nuclides from 40 references published between 2021 and 2024, subsequent to the release of the latest Atomic Mass Evaluation, were compiled. Although most of the new data are consistent with AME2020, for 73 nuclides, the deviations exceed the uncertainties. The new masses were calculated using the DRHBc theory with the PC-PK1 and PC-L3R density functionals and compared with the results from RMF+BCS calculations with TMA and Skyrme HFB calculations with SLy4, SV-min, and UNEDF1. The DRHBc calculations with both PC-PK1 and PC-L3R reproduce the data fairly well, with an RMS deviation below 1.5 MeV, demonstrating a clear advantage over other models in mass predictions. Taking the even-Z nuclei with 70Z80 as examples, the DRHBc calculations provide not only an overall high accuracy but also a robust description of the isospin dependence for nuclear masses. A quantitative comparison between PC-PK1 and PC-L3R results for the 296 nuclides shows that their differences in binding energies are generally below 1.0 MeV, those in RMS matter radii within 0.008 fm, and those in quadrupole deformations within 0.01. The largest discrepancies were found in 111Mo, attributed to the competing oblate and near-spherical minima with similar energies in the potential energy curve. The significant differences in deformation and the small differences in energies between the two minima indicate the possible shape coexistence in 111Mo and 112Mo.

These results strengthen confidence in the DRHBc predictions of nuclear masses. Continued progress would benefit from additional experimental mass measurements and establishment of a complete DRHBc mass table with PC-PK1. Moreover, large-scale DRHBc calculations using PC-L3R are highly promising. To further improve the mass description, the inclusion of triaxial deformation, rigorous treatment of nuclear magnetism, and beyond-mean-field extensions such as collective Hamiltonian approaches are being pursued within the DRHBc framework.

References
1. F. Wienholtz, D. Beck, K. Blaum, et al.,

Masses of exotic calcium isotopes pin down nuclear forces

. Nature 498, 346-349 (2013). https://doi.org/10.1038/nature12226
Baidu ScholarGoogle Scholar
2. E.M. Ramirez, D. Ackermann, K. Blaum, et al.,

Direct mapping of nuclear shell effects in the heaviest elements

. Science 337, 1207-1210 (2012). https://doi.org/10.1126/science.1225636
Baidu ScholarGoogle Scholar
3. H.A. Bethe,

Energy production in stars

. Phys. Rev. 55, 434-456 (1939). https://doi.org/10.1103/PhysRev.55.434
Baidu ScholarGoogle Scholar
4. E.M. Burbidge, G.R. Burbidge, W.A. Fowler, et al.,

Synthesis of the elements in stars

. Rev. Mod. Phys. 29, 547-650 (1957). https://doi.org/10.1103/RevModPhys.29.547
Baidu ScholarGoogle Scholar
5. D. Lunney, J.M. Pearson, C. Thibault,

Recent trends in the determination of nuclear masses

. Rev. Mod. Phys. 75, 1021-1082 (2003).
Baidu ScholarGoogle Scholar
6. A. Aprahamian, K. Langanke, M. Wiescher,

Nuclear structure aspects in nuclear astrophysics

. Prog. Part. Nucl. Phys. 54, 535-613 (2005). https://doi.org/10.1016/j.ppnp.2004.09.002
Baidu ScholarGoogle Scholar
7. H. Schatz,

Nuclear masses in astrophysics

. Int. J. Mass Spectrom. 349-350, 181-186 (2013). https://doi.org/10.1016/j.ijms.2013.03.016
Baidu ScholarGoogle Scholar
8. M.R. Mumpower, R. Surman, G.C. McLaughlin, et al.,

The impact of individual nuclear properties on r-process nucleosynthesis

. Prog. Part. Nucl. Phys. 86, 86-126 (2016). https://doi.org/10.1016/j.ppnp.2015.09.001
Baidu ScholarGoogle Scholar
9. Y. Yang, H.X. Zhang, Y.B. Wu, et al.,

Physics opportunities of the nuclear excitation by electron capture process

. Nucl. Sci. Tech. 36, 146 (2025). https://doi.org/10.1007/s41365-025-01717-0
Baidu ScholarGoogle Scholar
10. Y.J. Yuan, D.Q. Gao, L.Z. Ma, et al.,

Present status of HIRFL complex in Lanzhou

. J. Phys.: Conf. Ser. 1401, 012003 (2020). https://doi.org/10.1088/1742-6596/1401/1/012003
Baidu ScholarGoogle Scholar
11. X.H. Zhou, J.C. Yang, the HIAF project team,

Status of the high-intensity heavy-ion accelerator facility in China

. AAPPS Bulletin 32, 35 (2022). https://doi.org/10.1007/s43673-022-00064-1
Baidu ScholarGoogle Scholar
12. D. Castelvecchi,

Long-awaited accelerator ready to explore origins of elements

. Nature 605, 201-203 (2022).
Baidu ScholarGoogle Scholar
13. H. Sakurai,

Nuclear physics with RI Beam Factory

. Front. Phys. 13, 132111 (2018). https://doi.org/10.1007/s11467-018-0849-0
Baidu ScholarGoogle Scholar
14. M. Durante, P. Indelicato, B. Jonson, et al.,

All the fun of the FAIR: fundamental physics at the facility for antiproton and ion research

. Phys. Scr. 94, 033001 (2019). https://doi.org/10.1088/1402-4896/aaf93f
Baidu ScholarGoogle Scholar
15. B. Hong,

Overview of the Rare isotope Accelerator complex for ON-line experiments (RAON) project

. J. Phys.: Conf. Ser. 2586, 012143 (2023). https://doi.org/10.1088/1742-6596/2586/1/012143
Baidu ScholarGoogle Scholar
16. G.C. Ball, L. Buchmann, B. Davids, et al.,

Physics with reaccelerated radioactive beams at TRIUMF-ISAC

. J. Phys. G 38, 024003 (2011). https://doi.org/10.1088/0954-3899/38/2/024003
Baidu ScholarGoogle Scholar
17. M. Thoennessen, The discovery of isotopes, (Springer Cham, 2016)
18. F.G. Kondev, M. Wang, W.J. Huang, et al.,

The NUBASE2020 evaluation of nuclear physics properties

. Chin. Phys. C 45, 030001 (2021). https://doi.org/10.1088/1674-1137/abddae
Baidu ScholarGoogle Scholar
19. W.J. Huang, M. Wang, F.G. Kondev, et al.,

The AME 2020 atomic mass evaluation (I). Evaluation of input data, and adjustment procedures

. Chin. Phys. C 45, 030002 (2021). https://doi.org/10.1088/1674-1137/abddb0
Baidu ScholarGoogle Scholar
20. M. Wang, W.J. Huang, F.G. Kondev, et 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
21. J. Erler, N. Birge, M. Kortelainen, et al.,

The limits of the nuclear landscape

. Nature 486, 509 (2012).
Baidu ScholarGoogle Scholar
22. 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, 1-215 (2018). https://doi.org/10.1016/j.adt.2017.09.001
Baidu ScholarGoogle Scholar
23. Z.Y. Zhang, Z.G. Gan, H.B. Yang, et al.,

New isotope 220Np: Probing the robustness of the N=126 shell closure in neptunium

. Phys. Rev. Lett. 122, 192503 (2019). https://doi.org/10.1103/PhysRevLett.122.192503
Baidu ScholarGoogle Scholar
24. D.S. Ahn, N. Fukuda, H. Geissel, et al.,

Location of the neutron dripline at fluorine and neon

. Phys. Rev. Lett. 123, 212501 (2019). https://doi.org/10.1103/PhysRevLett.123.212501
Baidu ScholarGoogle Scholar
25. P. Müller, A.J. Sierk, T. Ichikawa, et 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
26. N. Wang, M. Liu, X.Z. Wu, et al.,

Surface diffuseness correction in global mass formula

. Phys. Lett. B 734, 215-219 (2014). https://doi.org/10.1016/j.physletb.2014.05.049
Baidu ScholarGoogle Scholar
27. M.T. Wan, L. Ou, M. Liu, et al.,

Properties of the drip-line nucleus and mass relation of mirror nuclei

. Nucl. Sci. Tech. 36, 26 (2025). https://doi.org/10.1007/s41365-024-01633-9
Baidu ScholarGoogle Scholar
28. K.Y. Zhang, X.T. He, J. Meng, et al.,

Predictive power for superheavy nuclear mass and possible stability beyond the neutron drip line in deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 104, L021301 (2021). https://doi.org/10.1103/PhysRevC.104.L021301
Baidu ScholarGoogle Scholar
29. X.T. He, J.W. Wu, K.Y. Zhang, et al.,

Odd-even differences in the stability peninsula in the 106≤Z≤112 region with the deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 110, 014301 (2024). https://doi.org/10.1103/PhysRevC.110.014301
Baidu ScholarGoogle Scholar
30. D. Hirata, K. Sumiyoshi, I. Tanihata, et al.,

A systematic study of even-even nuclei up to the drip lines within the relativistic mean field framework

. Nucl. Phys. A 616, 438-445 (1997). https://doi.org/10.1016/S0375-9474(97)00115-2
Baidu ScholarGoogle Scholar
31. G.A. Lalazissis, S. Raman, P. Ring,

Ground-sate properties of even-even nuclei in the relativistic mean-field theory

. At. Data Nucl. Data Tables 71, 1-40 (1999). https://doi.org/10.1006/adnd.1998.0795
Baidu ScholarGoogle Scholar
32. L.S. Geng, H. Toki, J. Meng,

Masses, deformations and charge radii-nuclear ground-state properties in the relativistic mean field model

. Prog. Theor. Phys. 113, 785-800 (2005). https://doi.org/10.1143/PTP.113.785
Baidu ScholarGoogle Scholar
33. S. Goriely, N. Chamel, J.M. Pearson,

Skyrme-Hartree-Fock-Bogoliubov nuclear mass formulas: Crossing the 0.6 MeV accuracy threshold with microscopically deduced pairing

. Phys. Rev. Lett. 102, 152503 (2009). https://doi.org/10.1103/PhysRevLett.102.152503
Baidu ScholarGoogle Scholar
34. S. Goriely, S. Hilaire, M. Girod, et al.,

First Gogny-Hartree-Fock-Bogoliubov nuclear mass model

. Phys. Rev. Lett. 102, 242501 (2009). https://doi.org/10.1103/PhysRevLett.102.242501
Baidu ScholarGoogle Scholar
35. A.V. Afanasjev, S.E. Agbemava, D. Ray, et al.,

Nuclear landscape in covariant density functional theory

. Phys. Lett. B 726, 680-684 (2013). https://doi.org/10.1016/j.physletb.2013.09.017
Baidu ScholarGoogle Scholar
36. S.E. Agbemava, A.V. Afanasjev, D. Ray, et al.,

Global performance of covariant energy density functionals: Ground state observables of even-even nuclei and the estimate of theoretical uncertainties

. Phys. Rev. C 89, 054320 (2014).
Baidu ScholarGoogle Scholar
37. K.Q. Lu, Z.X. Li, Z.P. Li, et al.,

Global study of beyond-mean-field correlation energies in covariant energy density functional theory using a collective Hamiltonian method

. Phys. Rev. C 91, 027304 (2015). https://doi.org/10.1103/PhysRevC.91.027304
Baidu ScholarGoogle Scholar
38. Y.L. Yang, Y.K. Wang, P.W. Zhao, et al.,

Nuclear landscape in a mapped collective Hamiltonian from covariant density functional theory

. Phys. Rev. C 104, 054312 (2021). https://doi.org/10.1103/PhysRevC.104.054312
Baidu ScholarGoogle Scholar
39. J. Meng, Relativistic density functional for nuclear structure, (World Scientific, Singapore, 2016)
40. P. Ring,

Relativistic mean field theory in finite nuclei

. Prog. Part. Nucl. Phys. 37, 193-263 (1996). https://doi.org/10.1016/0146-6410(96)00054-3
Baidu ScholarGoogle Scholar
41. D. Vretenar, A.V. Afanasjev, G.A. Lalazissis, et al.,

Relativistic Hartree-Bogoliubov theory: Static and dynamic aspects of exotic nuclear structure

. Phys. Rep. 409, 101-259 (2005). https://doi.org/10.1016/j.physrep.2004.10.001
Baidu ScholarGoogle Scholar
42. 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, 470-563 (2006). https://doi.org/10.1016/j.ppnp.2005.06.001
Baidu ScholarGoogle Scholar
43. T. Nikšić, D. Vretenar, P. Ring,

Relativistic nuclear energy density functionals: Mean-field and beyond

. Prog. Part. Nucl. Phys. 66, 519-548 (2011). https://doi.org/10.1016/j.ppnp.2011.01.055
Baidu ScholarGoogle Scholar
44. J. Meng, J. Peng, S.Q. Zhang, et al.,

Progress on tilted axis cranking covariant density functional theory for nuclear magnetic and antimagnetic rotation

. Front. Phys. 8, 55-79 (2013). https://doi.org/10.1007/s11467-013-0287-y
Baidu ScholarGoogle Scholar
45. J. Meng, S.G. Zhou,

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

. J. Phys. G 42, 093101 (2015).
Baidu ScholarGoogle Scholar
46. S.G. Zhou,

Multidimensionally constrained covariant density functional theories nuclear shapes and potential energy surfaces

. Phys. Scr. 91, 063008 (2016). https://doi.org/10.1088/0031-8949/91/6/063008
Baidu ScholarGoogle Scholar
47. S.H. Shen, H.Z. Liang, W.H. Long, et al.,

Towards an ab initio covariant density functional theory for nuclear structure

. Prog. Part. Nucl. Phys. 109, 103713 (2019). https://doi.org/10.1016/j.ppnp.2019.103713
Baidu ScholarGoogle Scholar
48. J. Meng, P.W. Zhao,

Relativistic density functional theory in nuclear physics

. AAPPS Bulletin 31, 2 (2021). https://doi.org/10.1007/s43673-021-00001-8
Baidu ScholarGoogle Scholar
49. H. Kucharek, P. Ring,

Relativistic field theory of superfluidity in nuclei

. Z. Phys. A 339, 23-35 (1991). https://doi.org/10.1007/BF01282930
Baidu ScholarGoogle Scholar
50. Z.X. Ren, P.W. Zhao,

Toward a bridge between relativistic and nonrelativistic density functional theories for nuclei

. Phys. Rev. C 102, 021301 (2020). https://doi.org/10.1103/PhysRevC.102.021301
Baidu ScholarGoogle Scholar
51. J.N. Ginocchio,

Pseudospin as a relativistic symmetry

. Phys. Rev. Lett. 78, 436-439 (1997). https://doi.org/10.1103/PhysRevLett.78.436
Baidu ScholarGoogle Scholar
52. J. Meng, K. Sugawara-Tanabe, S. Yamaji, et al.,

Pseudospin symmetry in relativistic mean field theory

. Phys. Rev. C 58, R628-R631 (1998). https://doi.org/10.1103/PhysRevC.58.R628
Baidu ScholarGoogle Scholar
53. H.Z. Liang, J. Meng, S.G. Zhou,

Hidden pseudospin and spin symmetries and their origins in atomic nuclei

. Phys. Rep. 570, 1-84 (2015). https://doi.org/10.1016/j.physrep.2014.12.005
Baidu ScholarGoogle Scholar
54. S.G. Zhou, J. Meng, P. Ring,

Spin symmetry in the antinucleon spectrum

. Phys. Rev. Lett. 91, 262501 (2003). https://doi.org/10.1103/PhysRevLett.91.262501
Baidu ScholarGoogle Scholar
55. X.T. He, S.G. Zhou, J. Meng, et al.,

Test of spin symmetry in anti-nucleon spectra

. Eur. Phys. J. A 28, 265-269 (2006). https://doi.org/10.1140/epja/i2006-10066-0
Baidu ScholarGoogle Scholar
56. W. Koepf, P. Ring,

A relativistic description of rotating nuclei: the yrast line of 20Ne

. Nucl. Phys. A 493, 61-82 (1989). https://doi.org/10.1016/0375-9474(89)90532-0
Baidu ScholarGoogle Scholar
57. W. Koepf, P. Ring,

A relativistic theory of superdeformations in rapidly rotating nuclei

. Nucl. Phys. A 511, 279-300 (1990). https://doi.org/10.1016/0375-9474(90)90160-N
Baidu ScholarGoogle Scholar
58. J. Meng, P. Ring,

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

. Phys. Rev. Lett. 77, 3963-3966 (1996). https://doi.org/10.1103/PhysRevLett.77.3963
Baidu ScholarGoogle Scholar
59. J. Meng,

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

. Nucl. Phys. A 635, 3-42 (1998). https://doi.org/10.1016/S0375-9474(98)00178-X
Baidu ScholarGoogle Scholar
60. J. Meng, P. Ring,

Giant halo at the neutron drip line

. Phys. Rev. Lett. 80, 460-463 (1998). https://doi.org/10.1103/PhysRevLett.80.460
Baidu ScholarGoogle Scholar
61. S.Q. Zhang, J. Meng, S.G. Zhou, et al.,

Giant neutron halo in exotic calcium nuclei

. Chin. Phys. Lett. 19, 312 (2002). https://doi.org/10.1088/0256-307X/19/3/308
Baidu ScholarGoogle Scholar
62. J. Meng, S.G. Zhou, I. Tanihata,

The relativistic continuum Hartree-Bogoliubov description of charge changing cross-section for C,N,O and F isotopes

. Phys. Lett. B 532, 209-214 (2002). https://doi.org/10.1016/S0370-2693(02)01574-5
Baidu ScholarGoogle Scholar
63. W. Zhang, J. Meng, S.Q. Zhang, et al.,

Magic numbers for superheavy nuclei in relativistic continuum Hartree-Bogoliubov theory

. Nucl. Phys. A 753, 106-135 (2005). https://doi.org/10.1016/j.nuclphysa.2005.02.086
Baidu ScholarGoogle Scholar
64. Y. Lim, X.W. Xia, Y. Kim,

Proton radioactivity in relativistic continuum Hartree-Bogoliubov theory

. Phys. Rev. C 93, 014314 (2016). https://doi.org/10.1103/PhysRevC.93.014314
Baidu ScholarGoogle Scholar
65. Y. Kuang, X.L. Tu, J.T. Zhang, et al.,

Systematic study of elastic proton-nucleus scattering using relativistic impulse approximation based on covariant density functional theory

. Eur. Phys. J. A 59, 160 (2023). https://doi.org/10.1140/epja/s10050-023-01072-x
Baidu ScholarGoogle Scholar
66. S.G. Zhou, J. Meng, P. Ring, et al.,

Neutron halo in deformed nuclei

. Phys. Rev. C 82, 011301 (2010). https://doi.org/10.1103/PhysRevC.82.011301
Baidu ScholarGoogle Scholar
67. L.L. Li, J. Meng, P. Ring, et al.,

Deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 85, 024312 (2012). https://doi.org/10.1103/PhysRevC.85.024312
Baidu ScholarGoogle Scholar
68. L.L. Li, J. Meng, P. Ring, et al.,

Odd systems in deformed relativistic Hartree Bogoliubov theory in continuum

. Chin. Phys. Lett. 29, 042101 (2012). https://doi.org/10.1088/0256-307X/29/4/042101
Baidu ScholarGoogle Scholar
69. Y. Chen, P. Ring, J. Meng,

Influence of pairing correlations on the size of the nucleus in relativistic continuum Hartree-Bogoliubov theory

. Phys. Rev. C 89, 014312 (2014). https://doi.org/10.1103/PhysRevC.89.014312
Baidu ScholarGoogle Scholar
70. X.X. Sun, S.G. Zhou,

Shape decoupling effects and rotation of deformed halo nuclei

. Nucl. Phys. Rev. 41, 75-85 (2024). https://doi.org/10.11804/NuclPhysRev.41.2023CNPC56
Baidu ScholarGoogle Scholar
71. K.Y. Zhang, C. Pan, S.Y. Chen, et al.,

Recent progress on halo nuclei in relativistic density functional theory

. Nucl. Phys. Rev. 41, 191-199 (2024). https://doi.org/10.11804/NuclPhysRev.41.2023CNPC28
Baidu ScholarGoogle Scholar
72. K.Y. Zhang, C. Pan, S.Q. Zhang, et al.,

Towards a high-precision nuclear mass table with deformed relativistic Hartree-Bogoliubov theory in continuum

. Chin. Sci. Bull. 66, 3561-3569 (2021). https://doi.org/10.1360/TB-2020-1601
Baidu ScholarGoogle Scholar
73. K.Y. Zhang, M.K. Cheoun, Y.B. Choi, et al.,

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

. Phys. Rev. C 102, 024314 (2020). https://doi.org/10.1103/PhysRevC.102.024314
Baidu ScholarGoogle Scholar
74. C. Pan, M.K. Cheoun, Y.B. Choi, et al.,

Deformed relativistic Hartree-Bogoliubov theory in continuum with a point-coupling functional. II. Examples of odd Nd isotopes

. Phys. Rev. C 106, 014316 (2022). https://doi.org/10.1103/PhysRevC.106.014316
Baidu ScholarGoogle Scholar
75. P.W. Zhao, Z.P. Li, J.M. Yao, et al.,

New parametrization for the nuclear covariant energy density functional with a point-coupling interaction

. Phys. Rev. C 82, 054319 (2010). https://doi.org/10.1103/PhysRevC.82.054319
Baidu ScholarGoogle Scholar
76. P.W. Zhao, L.S. Song, B. Sun, et al.,

Crucial test for covariant density functional theory with new and accurate mass measurements from Sn to Pa

. Phys. Rev. C 86, 064324 (2012). https://doi.org/10.1103/PhysRevC.86.064324
Baidu ScholarGoogle Scholar
77. K.Y. Zhang, M.K. Cheoun, Y.B. Choi, et al.,

Nuclear mass table in deformed relativistic Hartree-Bogoliubov theory in continuum, I: Even-even nuclei

. At. Data Nucl. Data Tables 144, 101488 (2022). https://doi.org/10.1016/j.adt.2022.101488
Baidu ScholarGoogle Scholar
78. P. Guo, X.J. Cao, K.M. Chen, et al.,

Nuclear mass table in deformed relativistic Hartree-Bogoliubov theory in continuum, II: Even-Z nuclei

. At. Data Nucl. Data Tables 158, 101661 (2024). https://doi.org/10.1016/j.adt.2024.101661
Baidu ScholarGoogle Scholar
79. X.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
80. K.Y. Zhang, D.Y. Wang, S.Q. Zhang,

Effects of pairing, continuum, and deformation on particles in the classically forbidden regions for Mg isotopes

. Phys. Rev. C 100, 034312 (2019). https://doi.org/10.1103/PhysRevC.100.034312
Baidu ScholarGoogle Scholar
81. X.X. Sun, J. Zhao, S.G. Zhou,

Study of ground state properties of carbon isotopes with deformed relativistic Hartree-Bogoliubov theory in continuum

. Nucl. Phys. A 1003, 122011 (2020). https://doi.org/10.1016/j.nuclphysa.2020.122011
Baidu ScholarGoogle Scholar
82. Z.H. Yang, Y. Kubota, A. Corsi, et al.,

Quasifree neutron knockout reaction reveals a small s-orbital component in the borromean nucleus 17B

. Phys. Rev. Lett. 126, 082501 (2021). https://doi.org/10.1103/PhysRevLett.126.082501
Baidu ScholarGoogle Scholar
83. X.X. Sun,

Deformed two-neutron halo in 19B

. Phys. Rev. C 103, 054315 (2021). https://doi.org/10.1103/PhysRevC.103.054315
Baidu ScholarGoogle Scholar
84. S.Y. Zhong, S.S. Zhang, X.X. Sun, et al.,

Study of the deformed halo nucleus 31Ne with Glauber model based on microscopic self-consistent structures

. Sci. China Phys. Mech. Astron. 65, 262011 (2022). https://doi.org/10.1007/s11433-022-1894-6
Baidu ScholarGoogle Scholar
85. K.Y. Zhang, S.Q. Yang, J.L. An, et al.,

Missed prediction of the neutron halo in 37Mg

. Phys. Lett. B 844, 138112 (2023). https://doi.org/10.1016/j.physletb.2023.138112
Baidu ScholarGoogle Scholar
86. K.Y. Zhang, P. Papakonstantinou, M.H. Mun, et al.,

Collapse of the N=28 shell closure in the newly discovered 39Na nucleus and the development of deformed halos towards the neutron dripline

. Phys. Rev. C 107, L041303 (2023). https://doi.org/10.1103/PhysRevC.107.L041303
Baidu ScholarGoogle Scholar
87. C. Pan, K.Y. Zhang, S.Q. Zhang,

Nuclear magnetism in the deformed halo nucleus 31Ne

. Phys. Lett. B 855, 138792 (2024). https://doi.org/10.1016/j.physletb.2024.138792
Baidu ScholarGoogle Scholar
88. J.L. An, K.Y. Zhang, Q. Lu, et al.,

A unified description of the halo nucleus 37Mg from microscopic structure to reaction observables

. Phys. Lett. B 849, 138422 (2024). https://doi.org/10.1016/j.physletb.2023.138422
Baidu ScholarGoogle Scholar
89. L.Y. Wang, K.Y. Zhang, J.L. An, et al.,

Toward a unified description of the one-neutron halo nuclei 15C and 19C from structure to reaction

. Eur. Phys. J. A 60, 251 (2024). https://doi.org/10.1140/epja/s10050-024-01464-7
Baidu ScholarGoogle Scholar
90. X.Y. Zhang, Z.M. Niu, W. Sun, et al.,

Nuclear charge radii and shape evolution of Kr and Sr isotopes with the deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 108, 024310 (2023). https://doi.org/10.1103/PhysRevC.108.024310
Baidu ScholarGoogle Scholar
91. M.H. Mun, S. Kim, M.K. Cheoun, et al.,

Odd-even shape staggering and kink structure of charge radii of Hg isotopes by the deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Lett. B 847, 138298 (2023). https://doi.org/10.1016/j.physletb.2023.138298
Baidu ScholarGoogle Scholar
92. C. Pan, J. Meng,

Charge radii and their deformation correlation for even-Z nuclei in deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 112, 024316 (2025). https://doi.org/10.1103/b6j7-6q8z
Baidu ScholarGoogle Scholar
93. P. Guo, C. Pan, Y.C. Zhao, et al.,

Prolate-shape dominance in atomic nuclei within the deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 108, 014319 (2023). https://doi.org/10.1103/PhysRevC.108.014319
Baidu ScholarGoogle Scholar
94. M.H. Mun, E. Ha, Y.B. Choi, et al.,

Nuclear shape evolution of neutron-deficient Au and kink structure of Pb isotopes

. Phys. Rev. C 110, 024310 (2024). https://doi.org/10.1103/PhysRevC.110.024310
Baidu ScholarGoogle Scholar
95. Y.B. Choi, C.H. Lee, M.H. Mun, et al.,

Bubble nuclei with shape coexistence in even-even isotopes of Hf to Hg

. Phys. Rev. C 105, 024306 (2022). https://doi.org/10.1103/PhysRevC.105.024306
Baidu ScholarGoogle Scholar
96. 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). https://doi.org/10.1103/PhysRevC.105.034340
Baidu ScholarGoogle Scholar
97. R.Y. Zheng, X.X. Sun, G.F. Shen, et al.,

Evolution of N=20,28,50 shell closures in the 20≤Z≤30 region in deformed relativistic Hartree-Bogoliubov theory in continuum

. Chin. Phys. C 48, 014107 (2024). https://doi.org/10.1088/1674-1137/ad0bf2
Baidu ScholarGoogle Scholar
98. Y.X. Zhang, B.R. Liu, K.Y. Zhang, et al.,

Shell structure and shape transition in odd-Z superheavy nuclei with proton numbers Z=117, 119: Insights from applying deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 110, 024302 (2024). https://doi.org/10.1103/PhysRevC.110.024302
Baidu ScholarGoogle Scholar
99. W.J. Liu, C.J. Lv, P. Guo, et al.,

Magic number N=350 predicted by the deformed relativistic Hartree-Bogoliubov theory in continuum: Z=136 isotopes as an example

. Particles 7, 1078-1085 (2024).
Baidu ScholarGoogle Scholar
100. P.X. Du, J. Li,

Exploring the neutron magic number in superheavy nuclei: Insights into N=258

. Particles 7, 1086-1094 (2024).
Baidu ScholarGoogle Scholar
101. C. Pan, X.H. Wu,

Examination of possible proton magic number Z = 126 with the deformed relativistic Hartree-Bogoliubov theory in continuum

. Particles 8, 2 (2025).
Baidu ScholarGoogle Scholar
102. L. Wu, W. Zhang, J. Peng, et al.,

Shell structure evolution of U, Pu, and Cm isotopes with deformed relativistic Hartree-Bogoliubov theory in a continuum

. Particles 8, 19 (2025).
Baidu ScholarGoogle Scholar
103. Z.D. Huang, W. Zhang, S.Q. Zhang, et al.,

Ground-state properties and structure evolutions of odd-A transuranium Bk isotopes from deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 111, 034314 (2025). https://doi.org/10.1103/PhysRevC.111.034314
Baidu ScholarGoogle Scholar
104. Y. Xiao, S.Z. Xu, R.Y. Zheng, et al.,

One-proton emission from 148-151Lu in the DRHBc+WKB approach

. Phys. Lett. B 845, 138160 (2023). https://doi.org/10.1016/j.physletb.2023.138160
Baidu ScholarGoogle Scholar
105. Y.B. Choi, C.H. Lee, M.H. Mun, et al.,

α-decay half-lives for even-even isotopes of W to U

. Phys. Rev. C 109, 054310 (2024). https://doi.org/10.1103/PhysRevC.109.054310
Baidu ScholarGoogle Scholar
106. Q. Lu, K.Y. Zhang, S.S. Zhang,

Triaxial shape of the one-proton emitter 149Lu

. Phys. Lett. B 856, 138922 (2024). https://doi.org/10.1016/j.physletb.2024.138922
Baidu ScholarGoogle Scholar
107. M.H. Mun, K. Heo, M.K. Cheoun,

Calculation of α decay half-lives for Tl, Bi, and At isotopes

. Particles 8, 42 (2025).
Baidu ScholarGoogle Scholar
108. C. Pan, K.Y. Zhang, P.S. Chong, et al.,

Possible bound nuclei beyond the two-neutron drip line in the 50≤Z≤70 region

. Phys. Rev. C 104, 024331 (2021). https://doi.org/10.1103/PhysRevC.104.024331
Baidu ScholarGoogle Scholar
109. X.T. He, C. Wang, K.Y. Zhang, et al.,

Possible existence of bound nuclei beyond neutron drip lines driven by deformation

. Chin. Phys. C 45, 101001 (2021). https://doi.org/10.1088/1674-1137/ac1b99
Baidu ScholarGoogle Scholar
110. X.T. He, J.W. Wu, K.Y. Zhang, et al.,

Odd-even differences in the stability “peninsula” in the 106≤Z≤112 region with the deformed relativistic Hartree-Bogoliubov theory in continuum

. Phys. Rev. C 110, 014301 (2024). https://doi.org/10.1103/PhysRevC.110.014301
Baidu ScholarGoogle Scholar
111. M.H. Mun, M.K. Cheoun, E. Ha, et al.,

Symmetry energy from two-nucleon separation energies of Pb and Ca isotopes

. Phys. Rev. C 110, 014314 (2024). https://doi.org/10.1103/PhysRevC.110.014314
Baidu ScholarGoogle Scholar
112. W. Zhang, J.K. Huang, T.T. Sun, et al.,

Inner fission barriers of uranium isotopes in the deformed relativistic Hartree-Bogoliubov theory in continuum

. Chin. Phys. C 48, 1-8 (2024).
Baidu ScholarGoogle Scholar
113. S.B. Wang, P. Guo, C. Pan,

Determining the ground state for superheavy nuclei from the deformed relativistic Hartree-Bogoliubov theory in continuum

. Particles 7, 1139-1149 (2024).
Baidu ScholarGoogle Scholar
114. K.Y. Zhang, C. Pan, X.H. Wu, et al.,

Selected advances in nuclear mass predictions based on covariant density functional theory with continuum effects

. AAPPS Bulletin 35, 13 (2025). https://doi.org/10.1007/s43673-025-00153-x
Baidu ScholarGoogle Scholar
115. Z.X. Liu, Y.H. Lam, N. Lu, et al.,

The optimized point-coupling interaction for the relativistic energy density functional of Hartree-Bogoliubov approach quantifying the nuclear bulk properties

. Phys. Lett. B 842, 137946 (2023). https://doi.org/10.1016/j.physletb.2023.137946
Baidu ScholarGoogle Scholar
116. T. Bürvenich, D.G. Madland, J.A. Maruhn, et al.,

Nuclear ground state observables and QCD scaling in a refined relativistic point coupling model

. Phys. Rev. C 65, 044308 (2002). https://doi.org/10.1103/PhysRevC.65.044308
Baidu ScholarGoogle Scholar
117. C.E. Price, G.E. Walker,

Self-consistent Hartree description of deformed nuclei in a relativistic quantum field theory

. Phys. Rev. C 36, 354-364 (1987). https://doi.org/10.1103/PhysRevC.36.354
Baidu ScholarGoogle Scholar
118. P. Ring, P. Schuck, The nuclear many-body problem, (Springer Science & Business Media, 2004)
119. S. Perez-Martin, L.M. Robledo,

Microscopic justification of the equal filling approximation

. Phys. Rev. C 78, 014304 (2008). https://doi.org/10.1103/PhysRevC.78.014304
Baidu ScholarGoogle Scholar
120. A. Jacobs, C. Andreoiu, J. Bergmann, et al.,

Improved high-precision mass measurements of mid-shell neon isotopes

. Nucl. Phys. A 1033, 122636 (2023). https://doi.org/10.1016/j.nuclphysa.2023.122636
Baidu ScholarGoogle Scholar
121. Y. Yu, Y.M. Xing, Y.H. Zhang, et al.,

Nuclear structure of dripline nuclei elucidated through precision mass measurements of 23Si, 26P, 27,28S, and 31Ar

. Phys. Rev. Lett. 133, 222501 (2024). https://doi.org/10.1103/PhysRevLett.133.222501
Baidu ScholarGoogle Scholar
122. D. Puentes, Z. Meisel, G. Bollen, et al.,

High-precision mass measurement of 24Si and a refined determination of the rp process at the A=22 waiting point

. Phys. Rev. C 106, L012801 (2022). https://doi.org/10.1103/PhysRevC.106.L012801
Baidu ScholarGoogle Scholar
123. I.T. Yandow, A. Abdullah-Smoot, G. Bollen, et al.,

Mass measurement of 27P to constrain type-I x-ray burst models and validate the isobaric multiplet mass equation for the A=27,T=32 isospin quartet

. Phys. Rev. C 108, 065802 (2023). https://doi.org/10.1103/PhysRevC.108.065802
Baidu ScholarGoogle Scholar
124. J. Surbrook, G. Bollen, M. Brodeur, et al.,

First Penning trap mass measurement of 36Ca

. Phys. Rev. C 103, 014323 (2021). https://doi.org/10.1103/PhysRevC.103.014323
Baidu ScholarGoogle Scholar
125. W.S. Porter, E. Dunling, E. Leistenschneider, et al.,

Investigating nuclear structure near N=32 and N=34 Precision mass measurements of neutron-rich Ca, Ti, and V isotopes

. Phys. Rev. C 106, 024312 (2022). https://doi.org/10.1103/PhysRevC.106.024312
Baidu ScholarGoogle Scholar
126. E. , E. Dunling, G. Bollen, et al.,

Precision mass measurements of neutron-rich scandium isotopes refine the evolution of N=32 and N=34 shell closures

. Phys. Rev. Lett. 126, 042501 (2021). https://doi.org/10.1103/PhysRevLett.126.042501
Baidu ScholarGoogle Scholar
127. M. Wang, M. Zhang, X. Zhou, et al.,

Bρ-defined isochronous mass spectrometry: An approach for high-precision mass measurements of short-lived nuclei

. Phys. Rev. C 106, L051301 (2022). https://doi.org/10.1103/PhysRevC.106.L051301
Baidu ScholarGoogle Scholar
128. S. Iimura, M. Rosenbusch, A. Takamine, et al.,

Study of the N=32 and N=34 shell gap for Ti and V by the first high-precision multireflection time-of-flight mass measurements at bigRIPS-SLOWRI

. Phys. Rev. Lett. 130, 012501 (2023). https://doi.org/10.1103/PhysRevLett.130.012501
Baidu ScholarGoogle Scholar
129. R. Silwal, C. Andreoiu, B. Ashrafkhani, et al.,

Summit of the N=40 island of inversion: Precision mass measurements and ab initio calculations of neutron-rich chromium isotopes

. Phys. Lett. B 833, 137288 (2022). https://doi.org/10.1016/j.physletb.2022.137288
Baidu ScholarGoogle Scholar
130. W.S. Porter, B. Ashrafkhani, J. Bergmann, et al.,

Mapping the N=40 island of inversion: Precision mass measurements of neutron-rich Fe isotopes

. Phys. Rev. C 105, L041301 (2022). https://doi.org/10.1103/PhysRevC.105.L041301
Baidu ScholarGoogle Scholar
131. L. Canete, S. Giraud, A. Kankainen, et al.,

Erratum: Precision mass measurements of 67Fe and 69,70Co: Nuclear structure toward N=40 and impact on r-process reaction rates [Phys. Rev. C 101, 041304(R) (2020)]

. Phys. Rev. C 103, 029902 (2021). https://doi.org/10.1103/PhysRevC.103.029902
Baidu ScholarGoogle Scholar
132. S. Giraud, L. Canete, B. Bastin, et al.,

Mass measurements towards doubly magic 78Ni: Hydrodynamics versus nuclear mass contribution in core-collapse supernovae

. Phys. Lett. B 833, 137309 (2022).
Baidu ScholarGoogle Scholar
133. M. Wang, Y.H. Zhang, X. Zhou, et 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
134. S.F. Paul, J. Bergmann, J.D. Cardona, et al.,

Mass measurements of 60−63Ga reduce x-ray burst model uncertainties and extend the evaluated T=1 isobaric multiplet mass equation

. Phys. Rev. C 104, 065803 (2021). https://doi.org/10.1103/PhysRevC.104.065803
Baidu ScholarGoogle Scholar
135. W. Xian, S. Chen, S. Nikas, et al.,

Mass measurements of neutron-rich A≈90 nuclei constrain element abundances

. Phys. Rev. C 109, 035804 (2024). https://doi.org/10.1103/PhysRevC.109.035804
Baidu ScholarGoogle Scholar
136. Y.M. Xing, C.X. Yuan, M. Wang, et al.,

Isochronous mass measurements of neutron-deficient nuclei from 112Sn projectile fragmentation

. Phys. Rev. C 107, 014304 (2023). https://doi.org/10.1103/PhysRevC.107.014304
Baidu ScholarGoogle Scholar
137. M. Horana Gamage, R. Bhandari, G. Bollen, et al.,

Identification of a potential ultralow-Q-value electron-capture decay branch in 75Se via a precise Penning trap measurement of the mass of 75As

. Phys. Rev. C 106, 065503 (2022). https://doi.org/10.1103/PhysRevC.106.065503
Baidu ScholarGoogle Scholar
138. X. Zhou, M. Wang, Y.H. Zhang, et al.,

Mass measurements show slowdown of rapid proton capture process at waiting-point nucleus 64Ge

. Nat. Phys. 28, 1-7 (2023).
Baidu ScholarGoogle Scholar
139. I. Mardor, S.A.S. Andrés, T. Dickel, et 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
140. I. Mukul, C. Andreoiu, J. Bergmann, et al.,

Examining the nuclear mass surface of Rb and Sr isotopes in the A≈104 region via precision mass measurements

. Phys. Rev. C 103, 044320 (2021). https://doi.org/10.1103/PhysRevC.103.044320
Baidu ScholarGoogle Scholar
141. M. Hukkanen, W. Ryssens, P. Ascher, et al.,

Precision mass measurements in the zirconium region pin down the mass surface across the neutron midshell at N=66

. Phys. Lett. B 856, 138916 (2024). https://doi.org/10.1016/j.physletb.2024.138916
Baidu ScholarGoogle Scholar
142. A. Hamaker, E. Leistenschneider, R. Jain, et al.,

Precision mass measurement of lightweight self-conjugate nucleus 80Zr

. Nat. Phys. 17, 1408-1412 (2021). https://doi.org/10.1038/s41567-021-01395-w
Baidu ScholarGoogle Scholar
143. D.S. Hou, A. Takamine, M. Rosenbusch, et al.,

First direct mass measurement for neutron-rich 112Mo with the new ZD-MRTOF mass spectrograph system

. Phys. Rev. C 108, 054312 (2023). https://doi.org/10.1103/PhysRevC.108.054312
Baidu ScholarGoogle Scholar
144. K.L. Wang, A. Estrade, M. Famiano, et al.,

Mass measurements of neutron-rich nuclei near N=70

. Phys. Rev. C 109, 035806 (2024). https://doi.org/10.1103/PhysRevC.109.035806
Baidu ScholarGoogle Scholar
145. W.S. Porter, B. Liu, D. Ray, et al.,

Investigating the effects of precise mass measurements of Ru and Pd isotopes on machine learning mass modeling

. Phys. Rev. C 110, 034321 (2024). https://doi.org/10.1103/PhysRevC.110.034321
Baidu ScholarGoogle Scholar
146. H.F. Li, S. Naimi, T.M. Sprouse, et al.,

First application of mass measurements with the rare-RI ring reveals the solar r-process abundance trend at A=122 and A=123

. Phys. Rev. Lett. 128, 152701 (2022). https://doi.org/10.1103/PhysRevLett.128.152701
Baidu ScholarGoogle Scholar
147. Z. Ge, M. Reponen, T. Eronen, et al.,

High-precision mass measurements of neutron deficient silver isotopes probe the robustness of the N=50 shell closure

. Phys. Rev. Lett. 133, 132503 (2024). https://doi.org/10.1103/PhysRevLett.133.132503
Baidu ScholarGoogle Scholar
148. A. Jaries, M. Stryjczyk, A. Kankainen, et al.,

High-precision Penning-trap mass measurements of Cd and In isotopes at JYFLTRAP remove the fluctuations in the two-neutron separation energies

. Phys. Rev. C 108, 064302 (2023). https://doi.org/10.1103/PhysRevC.108.064302
Baidu ScholarGoogle Scholar
149. M. Mougeot, D. Atanasov, J. Karthein, et al.,

Mass measurements of 99-101In challenge ab initio nuclear theory of the nuclide 100Sn

. Nat. Phys. 17, 1099-1103 (2021).
Baidu ScholarGoogle Scholar
150. D.A. Nesterenko, J. Ruotsalainen, M. Stryjczyk, et al.,

High-precision measurements of low-lying isomeric states in 120-124In with the JYFLTRAP double Penning trap

. Phys. Rev. C 108, 054301 (2023). https://doi.org/10.1103/PhysRevC.108.054301
Baidu ScholarGoogle Scholar
151. C. Izzo, J. Bergmann, K.A. Dietrich, et al.,

Mass measurements of neutron-rich indium isotopes for r-process studies

. Phys. Rev. C 103, 025811 (2021). https://doi.org/10.1103/PhysRevC.103.025811
Baidu ScholarGoogle Scholar
152. D.E.M. Hoff, K. Kolos, G.W. Misch, et al.,

Direct mass measurements to inform the behavior of 128mSb in nucleosynthetic environments

. Phys. Rev. Lett. 131, 262701 (2023). https://doi.org/10.1103/PhysRevLett.131.262701
Baidu ScholarGoogle Scholar
153. A.A. Valverde, F.G. Kondev, B. Liu, et al.,

Precise mass measurements of A=133 isobars with the Canadian Penning trap: Resolving the Qβ- anomaly at 133Te

. Phys. Lett. B 858, 139037 (2024). https://doi.org/10.1016/j.physletb.2024.139037
Baidu ScholarGoogle Scholar
154. O. Beliuskina, D.A. Nesterenko, A. Jaries, et al.,

Mass measurements in the 132Sn region with the JYFLTRAP double Penning trap mass spectrometer

. Phys. Rev. C 110, 034325 (2024). https://doi.org/10.1103/PhysRevC.110.034325
Baidu ScholarGoogle Scholar
155. S. Kimura, M. Wada, H. Haba, et al.,

Comprehensive mass measurement study of 252Cf fission fragments with MRTOF-MS and detailed study of masses of neutron-rich Ce isotopes

. Phys. Rev. C 110, 045810 (2024). https://doi.org/10.1103/PhysRevC.110.045810
Baidu ScholarGoogle Scholar
156. R. Orford, N. Vassh, J.A. Clark, et al.,

Searching for the origin of the rare-earth peak with precision mass measurements across Ce–Eu isotopic chains

. Phys. Rev. C 105, L052802 (2022). https://doi.org/10.1103/PhysRevC.105.L052802
Baidu ScholarGoogle Scholar
157. E.M. Lykiardopoulou, G. Audi, T. Dickel, et al.,

Exploring the limits of existence of proton-rich nuclei in the Z=70−82 region

. Phys. Rev. C 107, 024311 (2023). https://doi.org/10.1103/PhysRevC.107.024311
Baidu ScholarGoogle Scholar
158. S. Beck, B. Kootte, I. Dedes, et al.,

Mass measurements of neutron-deficient Yb isotopes and nuclear structure at the extreme proton-rich side of the N=82 shell

. Phys. Rev. Lett. 127, 112501 (2021). https://doi.org/10.1103/PhysRevLett.127.112501
Baidu ScholarGoogle Scholar
159. T. Niwase, Y.X. Watanabe, Y. Hirayama, et al.,

Discovery of new isotope 241U and systematic high-precision atomic mass measurements of neutron-rich Pa-Pu nuclei produced via multinucleon transfer reactions

. Phys. Rev. Lett. 130, 132502 (2023). https://doi.org/10.1103/PhysRevLett.130.132502
Baidu ScholarGoogle Scholar
160. C.L. Morris, H.T. Fortune, L.C. Bland, et al.,

Target mass dependence of isotensor double charge exchange: Evidence for deltas in nuclei

. Phys. Rev. C 25, 3218-3220 (1982). https://doi.org/10.1103/PhysRevC.25.3218
Baidu ScholarGoogle Scholar
161. L. Canete, S. Giraud, A. Kankainen, et al.,

Precision mass measurements of 67Fe and 69,70Co: Nuclear structure toward N=40 and impact on r-process reaction rates

. Phys. Rev. C 101, 041304 (2020). https://doi.org/10.1103/PhysRevC.101.041304
Baidu ScholarGoogle Scholar
162. Z. Meisel, S. George, S. Ahn, et al.,

Nuclear mass measurements map the structure of atomic nuclei and accreting neutron stars

. Phys. Rev. C 101, 052801 (2020). https://doi.org/10.1103/PhysRevC.101.052801
Baidu ScholarGoogle Scholar
163. A. Estrad��, M. Mato��, H. Schatz, et al.,

Time-of-Flight Mass Measurements for Nuclear Processes in Neutron Star Crusts

. Phys. Rev. Lett. 107, 172503 (2011). https://doi.org/10.1103/PhysRevLett.107.172503
Baidu ScholarGoogle Scholar
164. H.L. Seifert, J.M. Wouters, D.J. Vieira, et al.,

Mass measurement of neutron-rich isotopes from 51Ca to 72Ni

. Z. Phys. A 349, 25-32 (1994). https://doi.org/10.1007/BF01296329
Baidu ScholarGoogle Scholar
165. X.L. Tu, H.S. Xu, M. Wang, et al.,

Direct Mass Measurements of Short-Lived A=2Z−1 Nuclides 63Ge, 65As, 67Se, and 71Kr and Their Impact on Nucleosynthesis in the rp Process

. Phys. Rev. Lett. 106, 112501 (2011). https://doi.org/10.1103/PhysRevLett.106.112501
Baidu ScholarGoogle Scholar
166. M. Oinonen, A. Jokinen, J. Äystö et al.,

βdecay of the proton-rich Tz=−1/2 nucleus, 71Kr

. Phys. Rev. C 56, 745-752 (1997). https://doi.org/10.1103/PhysRevC.56.745
Baidu ScholarGoogle Scholar
167. J. Huikari, M. Oinonen, A. Algora, et al.,

Mirror decay of 75Sr

. Eur. Phys. J. A 16, 359-363 (2003). https://doi.org/10.1140/epja/i2002-10103-0
Baidu ScholarGoogle Scholar
168. L. Batist, J. Döring, I. Mukha, et al.,

Isomerism in 96Ag and non-yrast levels in 96Pd and 95Rh, studied in β decay

. Nucl. Phys. A 720, 245-273 (2003). https://doi.org/10.1016/S0375-9474(03)01093-5
Baidu ScholarGoogle Scholar
169. J. Van Schelt, D. Lascar, G. Savard, et al.,

Mass measurements near the r-process path using the Canadian Penning Trap mass spectrometer

. Phys. Rev. C 85, 045805 (2012). https://doi.org/10.1103/PhysRevC.85.045805
Baidu ScholarGoogle Scholar
170. R.C. Greenwood, M.H. Putnam,

Measurement of β- end-point energies using a Ge detector with Monte Carlo generated response functions

. Nucl. Instrum. Meth. Phys. Res. A 337, 106-115 (1993). https://doi.org/10.1016/0168-9002(93)91142-A
Baidu ScholarGoogle Scholar
171. M. Explorer, https://massexplorer.frib.msu.edu/content/DFTMassTables.html..
172. 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, 231-256 (1998). https://doi.org/10.1016/S0375-9474(98)00180-8
Baidu ScholarGoogle Scholar
173. P. Klüpfel, P.G. Reinhard, T.J. Bürvenich, et al.,

Variations on a theme by Skyrme: A systematic study of adjustments of model parameters

. Phys. Rev. C 79, 034310 (2009). https://doi.org/10.1103/PhysRevC.79.034310
Baidu ScholarGoogle Scholar
174. M. Kortelainen, J. McDonnell, W. Nazarewicz, et al.,

Nuclear energy density optimization: Large deformations

. Phys. Rev. C 85, 024304 (2012). https://doi.org/10.1103/PhysRevC.85.024304
Baidu ScholarGoogle Scholar
175. Z.X. Liu, Y.H. Lam, N. Lu, et al.,

Nuclear ground-state properties probed by the relativistic Hartree-Bogoliubov approach

. At. Data Nucl. Data Tables 156, 101635 (2024). https://doi.org/10.1016/j.adt.2023.101635
Baidu ScholarGoogle Scholar
176. W. Sun, K.Y. Zhang, C. Pan, et al.,

Beyond-mean-field dynamical correlations for nuclear mass table in deformed relativistic Hartree-Bogoliubov theory in continuum

. Chin. Phys. C 46, 064103 (2022). https://doi.org/10.1088/1674-1137/ac53fa
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.