Introduction
The nuclear mass is one of the most fundamental properties of the nucleus and is important in many fields of nuclear physics [1]. On the one hand, the nuclear mass can provide important information on the nuclear structure and reactions, such as the nuclear deformation [2], properties of neutron-rich nuclei [3, 4], structure and decay of superheavy nuclei [5-8], shell effect [9-12], nuclear drip line [13], halo nuclei [14], and synthesis of superheavy nuclei [15-18]. On the other hand, the nuclear mass is a key input to many problems in nuclear astrophysics. It has a vital influence on the composition[19] and cooling rate [20] of a neutron star and directly determines the evolution path of the rapid neutron capture process in stellar nucleosynthesis [21-23].
The aforementioned research on nuclear physics requires a high-precision nuclear mass table. With the rapid development of radioactive nuclear beam devices and detection technology, nuclei with measured masses continue to move toward the drip lines. Recently, significant progress has been made in the measurement of short-lived isotope nuclear masses far from stable regions. Approximately 2500 nuclear masses have been measured with an accuracy of less than hundreds of keV [24-27]. However, owing to a series of difficulties in synthesis, separation, and detection in experiments, the nuclei involved in the studies of superheavy islands, drip-line nuclei, nuclear astrophysics, etc., are still significantly beyond the current scope of nuclei with measured masses. No significant breakthroughs may occur in the foreseeable future, particularly on the neutron-rich side. Therefore, a model that can accurately describe known masses and accurately predict unknown masses is crucial.
Since Weizsäcker first proposed the liquid-drop model nuclear mass formula in 1935 [28], various types of nuclear mass formulas have been proposed, including the macroscopic model (e.g., Bethe–Weizsäcker (BW) models [28-30]), microscopic model (e.g., Skyrme Hartree–Fock–Bogoliubov (HFB) models [31-33], relativistic-mean-field (RMF) model [34]), macroscopic–microscopic models (e.g., finite-range droplet model (FRDM) [35, 36], Weizsäcker–Skyrme (WS) model [37-41], Koura–Tachibana–Uno–Yamada (KTUY) [42], Duflo-Zuker model (DZ) [43] etc.), systematic local-mass Garvey–Kelson relation [44], and Audi–Wapstra extrapolation method [26, 27, 45-47]. These nuclear mass formulas can reproduce the known nuclear masses with a certain accuracy, but their predictions appear to diverge as the isospin asymmetry increases, particularly for unknown masses. For details on the introduction and comparison of these models, refer to [48] and the references therein. Therefore, a more accurate nuclear mass formula requires more accurate experimental data, particularly for extremely neutron-rich nuclei, and a deeper understanding of the nuclear force, particularly the isospin symmetry breaking from theory. However, efforts have been made to improve the precision and extrapolation ability of nuclear mass models using machine-learning algorithms [49-53].
The macroscopic-microscopic mass model can systematically and quickly calculate the mass of nuclei on the entire nuclide chart with a high accuracy and good prediction ability. In this paper, the properties of drip-line nuclei are studied using the WS nuclear mass formula (WS3.3)[38], and the nuclear Coulomb energy is studied using the mass relations of mirror nuclei. The remainder of this paper is structured as follows: Sect. 2 briefly introduces the WS3.3 nuclear mass formula. The properties of drip-line nuclei based on the WS3.3 formula and the Coulomb energy based on the mass relations of mirror nuclei are presented in Sect. 3 and 4, respectively. Finally, a summary is given in Sect. 5.
WS3.3 nuclear mass formula
The WS nuclear mass formula is based on the Bathe–Weizsäcker liquid-drop model [30] and Skyrme energy density functional theory. The evolution of the model and a detailed introduction are presented in [37-41]. The WS3.3 version is adopted in this paper. In the WS nuclear mass formula, the macroscopic part of the binding energy of the nucleus considers the correction of the deformation based on the liquid-drop model, and the microscopic part is the shell-correction energy. Considering the deformed liquid-drop and shell-correction energies, the total energy of the nucleus can be expressed as
The microscopic shell correction of the binding energy is obtained using the traditional Strutinsky shell-correction method [56]:
The calculation of the single-particle energy levels of a nucleus based on the deformed Woods–Saxon potential, combined with microscopic shell correction, requires four parameters: the potential well depth parameter V0, radius parameter r0, surface dispersion parameter a, and spin-orbit coupling parameter λ0. With nine parameters in the macroscopic part, the WS3.3 model has a total of 13 parameters. The parameters in the improved model shown in Table 1 are determined by fitting the experimental data from 2149 nuclei with N and
Parameter | Value | Parameter | Value |
---|---|---|---|
av (MeV) | -15.6223 | V0 (MeV) | -46.8784 |
as (MeV) | 18.0571 | r0 (fm) | 1.3840 |
ac (MeV) | 0.7194 | a (fm) | 0.7842 |
csym (MeV) | 29.1563 | λ0 | 26.3163 |
κ | 1.3484 | g1 | 0.00895 |
apair (MeV) | -5.4423 | g2 | -0.46324 |
c1 | 0.62966 |
Properties of the drip-line nucleus
The drip-line nucleus is a nucleus whose separation energy of the last neutron or proton is less than zero. The drip line can be obtained by marking these nuclei in the nuclide chart, and it serves as the boundary for the existence of nuclei in the nuclide chart. When the drip line is crossed, the binding energy of the nucleus is insufficient to bind the last nucleon. Consequently, the near-drip-line nucleus exhibits extremely weak binding, resulting in a significantly more important proportion of coupling between nucleus-bound states and the continuum spectrum. This phenomenon causes many peculiarities among the nuclei near the drip line, such as the neutron skin and neutron halo [57, 58], cluster structure [59], and the emergence and disappearance of traditional magic numbers [60, 61]. Owing to their very low separation energies, the accurate prediction of drip-line positions requires high accuracy in nuclear mass formulas. Note that the differences in the nuclear masses given by the various models increase as the deviation from the β-stable valley increases. Therefore, the properties related to the drip line and drip-line nuclei serve as important standards for testing nuclear mass formulas.
First, the masses of the nuclei are calculated, and the drip, β-stable, and most-bound-nucleus (MBN) lines are determined using the WS3.3 nuclear mass formula, as shown in Fig. 1. The MBN is identified as the nucleus with the maximum specific binding energy in each isotope chain, whereas the β-stable nucleus has the maximum specific binding energy in each isobar chain. Nuclei with known experimental masses and mirror nuclei are obtained from the AME2020 mass table.
-202502/1001-8042-36-02-007/alternativeImage/1001-8042-36-02-007-F001.jpg)
The isospin asymmetry of the β-stable nucleus, drip-line nucleus, and MBN as functions of the neutron number given by the WS3.3 model are shown in Fig. 2. Separate representations of odd and even charges are provided for the drip-line nuclei to mitigate the influence of the symmetry effect. These results show that in the light nucleus region, both β-stable nuclei and MBN exhibit essentially similar isospin asymmetry, which increase as the neutron number increases because the Coulomb repulsion must be offset by excess neutrons. Furthermore, the isospin asymmetry of most β-stable nuclei progressively exceed those of the MBN. In light nuclear regions, the isospin asymmetry of the proton drip-line nucleus tends to be smaller than zero, indicating a proton-rich state. However, this tendency shifts towards increasing values, aligning more closely with those exhibited by the MBN as the neutron number increases. Conversely, as a natural consequence, the isospin asymmetry displayed by neutron-drip-line nuclei significantly surpass those observed in β-stable nuclei, MBNs, and proton drip-line nuclei. As the nucleus size increases, the isospin asymmetry of the neutron drip-line nucleus tends to decrease, eventually saturating at approximately 0.38 in the heavy nuclei region. Some nuclear mass models, such as HFB27 [33], RMF [34], FRDM12 [36], KTUY [42], DZ31 [43], and Bhagwat [62] also predict the same tendency, as shown by the pink block in Fig. 2. This suggests that the increased Coulomb energy of the heavy drip-line nucleus, resulting from the increased number of protons, is not simply offset by a further increase in the neutron number.
-202502/1001-8042-36-02-007/alternativeImage/1001-8042-36-02-007-F002.jpg)
To elucidate the mechanism for equilibrating the binding energy of heavy drip-line nuclei, we analyze the quadrupole deformation and deformation energy of the nucleus as a function of the neutron number, as presented in Fig. 3. The deformation energy is defined as the difference between the energies of the deformed and conventional spherical nuclei [6]:
-202502/1001-8042-36-02-007/alternativeImage/1001-8042-36-02-007-F003.jpg)
We can observe that the β2 distribution of the proton drip-line nucleus closely resembles that of β-stable nuclei throughout the nuclear region, ranging from -0.3 to 0.3. Similarly, the β2 distribution of the neutron drip-line nucleus is relatively uniform in the regions where N<200. This implies that these nuclei can counterbalance the increased Coulomb repulsion owing to charge enhancement through symmetry energy by enriched neutrons and deformation energy by suitable deformation. Conversely, in the region where N >200 for the neutron drip-line nucleus, no nucleus has a small deformation (
Furthermore, a clear correlation between isospin asymmetry of nucleus at the neutron drip line and magic number can be observed in Fig. 2. In Fig. 2, the blue dotted lines denote the positions of traditional magic numbers: 28, 34, 50, 82, and 126, which precisely correspond to locations where saltation occurs simultaneously in isospin asymmetry of the MBN and that of neutron drip-line nucleus with odd Z and even Z. This correlation is not apparent in the proton drip-line nucleus. Nevertheless, no confirmed magic number corresponds to some saltation positions in the heavy-nuclei region. This implies that the neutron numbers corresponding to these positions, i.e., 162 (164), 184, 212, and 236 (238), denoted by magenta lines, might also be neutron magic numbers. To obtain a more distinct perspective on the correlation between the isospin asymmetry of the neutron drip-line nucleus and the magic number, Fig. 4 presents the differences in proton numbers between two neighboring nuclei at the neutron drip line ΔZ as a function of the neutron number. We can imagine that, without the shell effect, the proton number should change smoothly with the neutron number. However, Fig. 4 shows that for some nuclei, with the addition of two neutrons, the proton number increases by -2, 4, 6, even 8. This indicates a shell effect.
-202502/1001-8042-36-02-007/alternativeImage/1001-8042-36-02-007-F004.jpg)
The positions where ΔZ of the neutron drip-line nucleus with odd Z and even Z simultaneously undergo saltation nearby are 28, 34, 50, 82, 126, 184, and 236 (238). They reproduce known magic numbers (28, 34, 50, 82, and 126) quite well. The neutron numbers 184, 236 (238) may be new magic numbers. Using shell model calculations, 184 [63-74] and 238 [63, 66] are also predicted as new neutron magic numbers. In addition, 162 [63, 69, 70, 73] and 164 [63, 69] are predicted as new neutron numbers, which are predicted using isospin asymmetry saltation but not by the charge number saltation in this paper. This indicates that the saltations in the isospin asymmetry of the neutron drip-line nucleus and in the difference in proton numbers between two neighboring nuclei at the neutron drip line can be used as criteria to verify the prospective magic number.
Mass relation of mirror nuclei
A pair of nuclei with the same mass number A and, exchanged proton number Z and neutron number N are called mirror nuclei. With symmetric properties, mirror nuclei are widely used to study nuclear structures [75-78] and reactions [79-82].
Based on the isospin symmetry of nuclear interactions, the binding energy difference between mirror nuclei (BEDbMN) results from the Coulomb energy and small mass difference between the neutron and proton [83, 84]. Precise measurement of the masses of mirror nuclei enables the study of isospin symmetry and the charge independence of the nuclear force and can be employed to test nuclear structure models. It is an effective approach for enhancing the precision of nuclear mass formulas by utilizing the mass relations of mirror nuclei to investigate the nuclear mass formula, particularly the reasonable correction of the Coulomb term [38, 88-93].
By subtracting the Coulomb energy, the nuclear-force-dependent part in the BEDbMN can be considered approximately equal, i.e.,
Because the nuclear force is independent of the charge of the nucleus, the BEDbMN is only reflected in the Coulomb energy terms. Therefore, the Coulomb term coefficient can be determined using the BEDbMN. Considering the Coulomb exchange effect, the Coulomb term in (3) can be expressed as
ac (MeV) | c | σ (MeV) (AME2003) | σ (MeV) (AME2020) | |
---|---|---|---|---|
This work | 0.73 | 1.49 | 0.312 | 0.334 |
Tian | 0.69 | 1.19 | 0.418 | 0.453 |
WS3.3 | 0.71 | 1.0 | 0.684 | 0.816 |
-202502/1001-8042-36-02-007/alternativeImage/1001-8042-36-02-007-F005.jpg)
Because only one pair of mirror nuclei has N-Z=5 (19Mg and 19N) in the AME2003 mass table, a linear fit is not feasible. The empirical curve for N-Z=5 shown in Fig. 6 is an extrapolation of expression (17). An apparent variation in the experimental value for mirror nulei pair of 19Mg and 19N between AME2003 and AME2020 can be observed. This is because a correction of 1.2 MeV is applied to the binding energy of 19Mg. With this correction and one new mirror nuclei pair (17C and 17Na) added to AME2020, the extrapolated result for N-Z=5 corresponds with the data of AME2020.
-202502/1001-8042-36-02-007/alternativeImage/1001-8042-36-02-007-F006.jpg)
Compared with AME2003, the masses of 21 new mirror nuclei pairs are included in the AME2020 mass table, which can be utilized to verify the accuracy of the mirror nuclei binding energy difference formula. The RMS deviations between the experimental and theoretical masses for the 21 newly added mirror nuclei pairs in this work and the WS3.3 model are shown in Table 3. Considering the mass relations of the mirror nuclei, the mass formula generates a lower RMS deviation. Currently, proton-rich nuclei whose masses have been measured are considerably close to the proton drip line; however, some nuclei remain to be synthesized and measured. More accurate predictions of the experimentally unmeasured masses of mirror nuclei toward the proton drip line by employing the mass relation of the mirror nuclei would be beneficial.
The study of the BEDbMN reveals that a more precise Coulomb term, particularly the Coulomb exchange term coefficient, can be acquired from a BEDbMN. In addition, the nuclear mass formula have a better physical basis. Therefore, we attempt to utilize three schemes in the nuclear mass formula to refit the AME2003 mass table and investigate the effect of the Coulomb exchange term on the accuracy of the nuclear mass formula.
A. The Coulomb term coefficients ac and c are fixed by taking the values determined from the mass relations of mirror nuclei. The other 12 parameters are determined via the optimization method.
B. The Coulomb exchange term coefficient c assumes different values, the other 13 parameters, including ac, are determined using the optimization method. Specifically, the WS3.3 parameter is obtained when c=1.0.
C. The Coulomb term coefficients ac and c and the other 14 parameters are determined using the optimization method.
The values of the Coulomb term coefficients for the three schemes and their mass RMS deviations are listed in Table 4. Although Scheme A provided the most accurate description of the binding energy difference for 71 pairs of mirror nuclei, the RMS deviation of the mass of 2149 nuclei is larger than that provided by the WS3.3 parameter set. As shown in Scheme B, the Coulomb exchange term coefficient ranges from 0 to 2.0, and the Coulomb direct term coefficient ac varies only slightly. This result indicates that the Coulomb direct-term coefficient is robust against the Coulomb exchange-term coefficient. The mass RMS deviation as a function of the Coulombic exchange coefficient is shown in Fig. 7. The RMS deviation is the smallest when
-202502/1001-8042-36-02-007/alternativeImage/1001-8042-36-02-007-F007.jpg)
The preceding discussion reveals that the accuracy of the nuclear mass formula is not enhanced but rather diminishes when directly using the Coulomb term coefficients determined by the mass relations of mirror nuclei pairs. A possible reason for this is that 71 pairs of mirror nuclei constitute a relatively small proportion of the 2149 nuclei, and no mirror nuclei exist in the heavy nuclei region. The Coulomb term coefficient determined by the mirror nuclei in the light nuclei region based on the local mass relation is less precise when extrapolated to the heavy nuclei region. As a test, the above three schemes are employed to calculate the masses of 71 pairs of mirror nuclei, and the obtained RMS deviations are presented in Table 5.
Similar to the results based on the 2149 nuclei optimization parameters, the RMS deviation obtained by the method in this paper remains higher than that of WS3.3 parameters, although the disparity in RMS deviations between the two methods is not significant. The accuracy of the result for c=1.0 is still higher than that for c=1.49. Although the accuracy of Scheme C is slightly improved, the Coulomb coefficient determined by mirror nuclei masses differs significantly from that determined by the 2149 nuclear masses, and the experimental data of the 2149 nuclear masses cannot be well represented using the coefficients in Scheme C. These observations suggest a systematic correlation among the terms of the nuclear mass formula. An improvement in the accuracy of a single term in the formula, such as the Coulomb term, may not enhance the accuracy of the entire nuclear mass formula.
Finaly, let us consider the Coulomb exchange coefficient c=1.49 obtained by fitting the BEDbMN in this study. The well-known Fermi gas model yields a considerably smaller result of 0.76 (that is,
The Coulomb energies of the β-stable nuclei, calculated using various Coulomb energy forms, are presented in Fig. 8. As shown in the figure, Ec with c=0.76 can reproduce
-202502/1001-8042-36-02-007/alternativeImage/1001-8042-36-02-007-F008.jpg)
summary
The properties of the drip-line nucleus and mass relation of mirror nuclei are studied based on the WS3.3 nuclear mass formula. We observe that the isospin asymmetry of heavy nuclei on the neutron drip line tends toward a saturation value of 0.38. An analysis of the nuclear quadrupole deformation value β2 and the deformation energy as a function of the neutron number suggests that in the heavy nuclei region, the strong Coulomb energy caused by the augmented proton is resisted by enormous deformation but not by the further addition of neutrons because of isospin asymmetry saturation. Additionally, the conventional magic number has a distinct correspondence with the saltation position of the isospin asymmetries of the MBN, the difference in proton numbers between two neighboring nuclei at the neutron drip line, and isospin asymmetry degree at the neutron drip-line nucleus as a function of the neutron number. Therefore, saltation without a known corresponding magic number can serve as a reference to verify the undetermined neutron magic number.
By considering the mass relation of mirror nuclei, i.e., BEDbMN, a Coulomb term with a stronger physical foundation and greater accuracy is obtained. However, the accuracy of the mass formula with the BEDbMN considered is lower than that without the BEDbMN considered. A reason for this is that the Coulomb term coefficients are determined based on insufficient experimental data of mirror nuclei with A <75, which are insufficient to describe the entire mass table globally. The other reason is that a systematic relationship exists between the coefficients in the mass formula; we should not expect it to improve accuracy by improving some of the coefficients independently. However, the relation of the BEDbMN can be used to accurately determine the mass of one of the mirror nuclei, which is experimentally unknown, by another whose mass is known. This is particularly useful for predicting the mass of isotopes that are difficult to synthesize or measure experimentally toward a proton drip line for N <50 where a mirror nuclei pair exists.
Recent trends in the determination of nuclear masses
. Rev. Mod. Phys. 75, 1021 (2003). https://doi.org/10.1103/RevModPhys.75.1021Nuclear deformation in the A≈100100 region: Comparison between new masses and mean-field predictions
. Phys. Rev. C 96,Effect of time-odd fields on odd-even mass differences of semi-magic nuclei
. Sci. China- Phys. Mech. Astron. 59,Masses of exotic calcium isotopes pin down nuclear forces
. Nature 498, 346 (2013). https://doi.org/10.1038/nature12226Nazarewicz, Shape coexistence and triaxiality in the superheavy nuclei
. Nature 433, 705 (2005). https://doi.org/10.1038/nature03336Description of structure and properties of superheavy nuclei
. Prog. Part. Nucl. Phys. 58, 292 (2007). https://doi.org/10.1016/j.ppnp.2006.05.001Theoretical description of the decay chain of the nucleus 294 118
. J. Phys. G: Nucl. Part. Phys. 43,Superheavy nuclei from 48Ca-induced reactions
. Nucl. Phys. A 944, 62 (2015). https://doi.org/10.1088/0954-3899/43/9/095106Systematic study of shell gaps in nuclei
. Phys. Rev. C 90,Magic Nature of Neutrons in 54Ca: First Mass Measurements of 55-57Ca
. Phys. Rev. Lett. 121,Probing the N=32 Shell Closure below the Magic Proton Number Z=20: Mass Measurements of the Exotic Isotopes 52,53K
. Phys. Rev. Lett. 121,Dawning of the N = 32 Shell Closure Seen through Precision Mass Measurements of Neutron-Rich Titanium Isotopes
. Phys. Rev. Lett. 121,Mean-field description of ground-state properties of drip-line nuclei: Pairing and continuum effects
. Phys. Rev. C 53, 2809 (1996). https://doi.org/10.1103/PhysRevC.53.2809Direct mass measurements of 19B, 22C, 29F, 31Ne, 34Na and other light exotic nuclei
. Phys. Rev. Lett. 109,The limits of the nuclear landscape
. Nature 486, 509 (2012). https://doi.org/10.1038/nature11188Study on superheavy nuclei and superheavy elements
. Nucl. Phys. Rev. 34, 318 (2017). https://doi.org/10.11804/NuclPhysRev.34.03.318The limits of nuclear mass and charge
. Nature Physics 14, 5 (2018). https://doi.org/10.1038/s41567-018-0163-3Direct mapping of nuclear shell effects in the heaviest elements
. Science 337, 1207 (2012). https://doi.org/10.1126/science.1225636Plumbing neutron stars to new depths with the binding energy of the exotic nuclide Zn 82
. Phys. Rev. Lett. 110,Strong neutrino cooling by cycles of electron capture and β-decay in neutron star crusts
. Nature 505, 62 (2014). https://doi.org/10.1038/nature12757Where, oh where has the r-process gone
? Physics Reports 442, 237 (2007). https://doi.org/10.1016/j.physrep.2007.02.006The r-process of stellar nucleosynthesis: Astrophysics and nuclear physics achievements and mysteries
. Physics Reports 450, 97 (2007). https://doi.org/10.1016/j.physrep.2007.06.002WLW mass model in nuclear r-process calculations
. Acta Phys. Sin. 61,Mass and lifetime measurements of exotic nuclei in storage rings
. Mass Spectrometry Reviews 27, 428 (2008). https://doi.org/10.1002/mas.20173Toward precision mass measurements of neutron-rich nuclei relevant to r-process nucleosynthesis
. Front. Phys. 10,The AME 2020 atomic mass evaluation (II). Tables, graphs and references
. Chin. Phys. C. 45,The AME 2020 atomic mass evaluation (I). Evaluation of input data, and adjustment procedures
. Chin. Phys. C. 45,Zur Theorie der Kernmassen
. Z. Physik 96, 431 (1935). https://doi.org/10.1007/BF01337700Mutual influence of terms in a semi-empirical mass formula
. Nucl. Phys. A 798, 29 (2008). https://doi.org/10.1016/j.nuclphysa.2007.10.011Nuclear physics A
. Stationary states of nuclei. Rev. Mod. Phys. 8, 82 (1936). https://doi.org/10.1103/RevModPhys.8.82Further explorations of Skyrme-Hartree-Fock-Bogoliubov mass formulas.VII.Simultaneous fits to masses and fission barriers
. Phys. Rev. C 75,Skyrme-Hartree-Fock-Bogoliubov nuclear mass formulas: Crossing the 0.6 MeV accuracy threshold with microscopically deduced pairing
. Phys. Rev. Lett. 102,Hartree-Fock-Bogoliubov nuclear mass model with 0.50 MeV accuracy based on standard forms of Skyrme and pairing functionals
. Phys. Rev. C 88,Masses, deformations and charge radii–nuclear ground-state properties in the relativistic mean field model
. Prog. Theo. Phys. 113, 785 (2005). https://doi.org/10.1143/PTP.113.785Nuclear ground-state masses and deformations
. At. Data Nucl. Data Tables 59, 185 (1995). https://doi.org/10.1006/adnd.1995.1002Nuclear ground-state masses and deformations: FRDM(2012)
. At. Data Nucl. Data Tables 109-110, 1 (1996). https://doi.org/10.1016/j.adt.2015.10.002Modification of nuclear mass formula by considering isospin effects
. Phys. Rev. C 81,Mirror nuclei constraint in nuclear mass formula
. Phys. Rev. C 82,Nuclear mass predictions with a radial basis function approach
. Phys. Rev. C 84,Further improvements on a global nuclear mass model
. Phys. Rev. C 84,Surface diffuseness correction in global mass formula
. Phys. Lett. B 734, 215 (2014). https://doi.org/10.1016/j.physletb.2014.05.049Nuclidic mass formula on a spherical basis with an improved even-odd term
. Prog. Theor. Phys. 113, 305 (2005). https://doi.org/10.1143/PTP.113.305Microscopic mass formulas
. Phys. Rev. C 52,Set of nuclear-mass relations and a resultant mass table
. Rev. Mod. Phys. 41, S1 (1969). https://doi.org/10.1103/RevModPhys.41.S1The Ame2003 atomic mass evaluation
. Nucl. Phys. A 729, 337 (2003). https://doi.org/10.1016/j.nuclphysa.2003.11.003The AME2016 atomic mass evaluation (I). Evaluation of input data; and adjustment procedures
. Chin. Phys. C 41,The AME2016 atomic mass evaluation (II). Tables, graphs and references
. Chin. Phys. C 41,Masses of exotic nuclei
. Prog. Part. Nucl. Phys. 120,Nuclear mass predictions based on Bayesian neural network approach with pairing and shell effects
. Phys. Lett. B 778, 48 (2018). https://doi.org/10.1016/j.physletb.2018.01.002Nuclear mass predictions with machine learning reaching the accuracy required by r-process studies
. Phys. Rev. C 106,Multi-task learning on nuclear masses and separation energies with the kernel ridge regression
. Phys. Lett. B 834,Machine learning the nuclear mass
. Nucl. Sci. Tech. 32, 109 (2021). https://doi.org/10.1007/s41365-021-00956-1Studies on several problems in nuclear physics by using machine learning
. Nucl. Tech. 46, 95 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080009 (in Chinese)On the origin of the Wigner energy
. Phys. Lett. B 407, 103 (1997). https://doi.org/10.1016/S0370-2693(97)00711-9The anatomy of the simplest Duflo–Zuker mass formula
. Nucl. Phys. A 843, 14 (2010). https://doi.org/10.1016/j.nuclphysa.2010.05.055A new definition of shell corrections to the liquid drop energy
. Nucl. Phys. A 225, 405 (1975). https://doi.org/10.1016/0375-9474(75)90688-0Recent experimental progress in nuclear halo structure studies
. Prog. Part. Nucl. Phys. 68, 215 (2013). https://doi.org/10.1016/j.ppnp.2012.07.001Recent experimental progress in nuclear halo structure studies
. Phys. Rev. Lett. 87,Nuclear clusters and nuclear molecules
. Phys. Rep. 432, 43 (2006). https://doi.org/10.1016/j.physrep.2006.07.001Nuclear structure studies from reaction induced by radioactive nuclear beams
. Prog. Part. Nucl. Phys. 35, 505 (1995). https://doi.org/10.1016/0146-6410(95)00046-LNuclear magic numbers: New features far from stability
. Prog. Part. Nucl. Phys. Supp. 61, 602 (2008). https://doi.org/10.1016/j.ppnp.2008.05.001Simple nuclear mass formula
. Phys. Rev. C 90,Relativistic continuum Hartree Bogoliubov theory for ground-state properties of exotic nuclei
. Prog. Part. Nucl. Phys. 57, 470 (2006). https://doi.org/10.1016/j.ppnp.2005.06.001Superheavy magic structures in the relativistic Hartree–Fock–Bogoliubov approach
. Phys. Lett. B 732, 169 (2014). https://doi.org/10.1016/j.physletb.2014.03.031Superheavy nuclei with the vector self-coupling of the ω-meson in relativistic mean-field theory
. J. Phys. G: Nucl. Part. Phys. 36,Magic numbers for superheavy nuclei in relativistic continuum Hartree–Bogoliubov theory
. Nucl. Phys. A 753, 106 (2005). https://doi.org/10.1016/j.nuclphysa.2005.02.086On the nuclear structure and stability of heavy and superheavy elements
. Nucl. Phys. A 131, 1 (1969). https://doi.org/10.1016/0375-9474(69)90809-4On the stability of superheavy nuclei against fission
. Z. Physik 222, 261 (1969). https://doi.org/10.1007/BF01392125Stability of heavy and superheavy elements
. J. Phys. G Nucl. Part. Phys. 20, 1681 (1994). https://doi.org/10.1088/0954-3899/20/11/003Ground-state properties of the heaviest nuclei analyzed in a multidimensional deformation space
. Nucl. Phys. A 533, 132 (1991). https://doi.org/10.1016/0375-9474(91)90823-OSuperheavy nuclei in self-consistent nuclear calculations
. Phys. Rev. C 56, 238 (1997). https://doi.org/10.1103/PhysRevC.56.238Shell structure of superheavy nuclei in self-consistent mean-field models
. Phys. Rev. C 60,Shell structure of the superheavy elements
. Nucl. Phys. A 611, 211 (1996). https://doi.org/10.1016/S0375-9474(96)00337-5Superheavy nuclei in the relativistic mean-field theory
. Nucl. Phys. A 608, 202 (1996). https://doi.org/10.1016/0375-9474(96)00273-4Binding energies of near proton-drip line Z = 22-28 isotopes determined from measured isotopic cross section distributions
. Sci. China Phys. Mech. Astron. 62,Shell and isospin effects in nuclear charge radii
. Phys. Rev. C 88, 011301(R) (2013). https://doi.org/10.1103/PhysRevC.88.011301Constraining nuclear symmetry energy with the charge radii of mirror pair nuclei
. Nucl. Sci. Tech. 34, 119 (2023). https://doi.org/10.1007/s41365-023-01269-1Ab initio Gamow shell-model calculations for dripline nuclei
. (in Chinese) Nucl. Tech. 46, 121 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080012Systematic behavior of fragments in Bayesian neural network models for projectile fragmentation reactions
. Phys. Rev. C 108,Determination of the astrophysical 26Si(p,γ)27P reaction rate from the asymptotic normalization coefficients of 27Mg→26Mg+n
. Phys. Rev. C 73,Large Isospin Asymmetry in 22Si/22O Mirror Gamow-Teller Transitions Reveals the Halo Structure of 22Al
. Phys. Rev. Lett. 125,Progress on nuclear reactions and related nuclear structure at low energies
. Nucl. Tech. 46,Nuclear Coulomb energies
. Rep. Prog. Phys. 41, 957 (1978). https://doi.org/10.1088/0034-4885/41/7/001Mapping the proton drip line up to A = 70
. Phys. Rev. C 55, 2407 (1997). https://doi.org/10.1103/PhysRevC.55.2407Simple relations between masses of mirror nuclei
. Phys. Rev. C 94,Mass relations of mirror nuclei for both bound and unbound systems
. Phys. Rev. C 152,Improved mass relations of mirror nuclei
. Nucl. Sci. Tech. 35, 157 (2024). https://doi.org/10.1007/s41365-024-01501-6Improved Kelson-Garvey mass relations for proton-rich nuclei
. Phys. Rev. C 87,Mass relations of corresponding mirror nuclei
. Phys. Rev. C 100,Mass relations of mirror nuclei
. Phys. Rev. C 102,Mass relations of mirror nuclei
. Phys. Rev. C 102,Isochronous mass measurements of Tz=-1 fp-shell nuclei from projectile fragmentation of 58Ni
. Phys. Rev. C 98,Nuclear mass parabola and its applications
. Chin. Phys. C 43,Coulomb Energies
. Annu. Rev. Nucl. Sci. 19, 471 (1969). https://doi.org/10.1146/annurev.ns.19.120169.002351The authors declare that they have no competing interests.