An improved nuclear mass formula: WS3

Similar documents
Study of alpha decay with WS-type mass formula and RBF Correction

Joint ICTP-IAEA Workshop on Nuclear Structure Decay Data: Theory and Evaluation August Introduction to Nuclear Physics - 1

arxiv: v1 [nucl-th] 16 Sep 2017

Fusion Barrier of Super-heavy Elements in a Generalized Liquid Drop Model

Systematic Evaluation of the Nuclear Binding Energies as Functions of F -spin

Magic Numbers of Ultraheavy Nuclei

Microscopic description of fission properties for r-process nuclei

Systematic Study of Survival Probability of Excited Superheavy Nucleus

One and two proton separation energies from nuclear mass systematics using neural networks

arxiv: v1 [nucl-th] 21 Aug 2007

Lisheng Geng. Ground state properties of finite nuclei in the relativistic mean field model

FAVORABLE HOT FUSION REACTION FOR SYNTHESIS OF NEW SUPERHEAVY NUCLIDE 272 Ds

Nuclear Mass Models. J. M. Pearson

Coefficients and terms of the liquid drop model and mass formula

An extended liquid drop approach

Theoretical Study on Alpha-Decay Chains of

The limits of the nuclear chart set by fission and alpha decay

The role of isospin symmetry in collective nuclear structure. Symposium in honour of David Warner

arxiv:nucl-th/ v1 27 Apr 2004

Evolution Of Shell Structure, Shapes & Collective Modes. Dario Vretenar

arxiv:nucl-th/ v4 30 Jun 2005

Theoretical description of decay chains of SHN

Correlation between alpha-decay energies of superheavy nuclei

Programs for experimental nuclear physics with relativistic heavy ion beams at Beihang

Projected shell model for nuclear structure and weak interaction rates

Theoretical Nuclear Physics

EXTRACTION OF A "LIQUID-DROP" PART OP THE HF-ENERGY. M.Brack

Symmetry energy, masses and T=0 np-pairing

arxiv: v1 [nucl-th] 22 Mar 2012

Alpha Decay of Superheavy Nuclei

Ground-state properties of some N=Z medium mass heavy nuclei. Keywords: Nuclear properties, neutron skin thickness, HFB method, RMF model, N=Z nuclei

Structure of Atomic Nuclei. Anthony W. Thomas

Theory for nuclear processes in stars and nucleosynthesis

ALPHA-DECAY AND SPONTANEOUS FISSION HALF-LIVES OF SUPER-HEAVY NUCLEI AROUND 270Hs

Mass measurements of n-rich nuclei with A~70-150

New T=1 effective interactions for the f 5/2 p 3/2 p 1/2 g 9/2 model space: Implications for valence-mirror symmetry and seniority isomers

IAS and Skin Constraints. Symmetry Energy

Investigation of Nuclear Ground State Properties of Fuel Materials of 232 Th and 238 U Using Skyrme- Extended-Thomas-Fermi Approach Method

Charge density distributions and charge form factors of some even-a p-shell nuclei

Large-scale Gogny-HFB Calculation for r-process Nucleosynthesis: Towards a fully-converged microscopic mass table

Impact of fission on r-process nucleosynthesis within the energy density functional theory

WEAKLY BOUND NEUTRON RICH C ISOTOPES WITHIN RMF+BCS APPROACH

Microscopic Fusion Dynamics Based on TDHF

Nuclear Symmetry Energy Constrained by Cluster Radioactivity. Chang Xu ( 许昌 ) Department of Physics, Nanjing University

The isospin dependence of the nuclear force and its impact on the many-body system

HALF-LIVES OF NUCLEI AROUND THE SUPERHEAVY NUCLEUS

Nuclear structure theory

Compressibility of Nuclear Matter from Shell Effects in Nuclei

Interaction cross sections for light neutron-rich nuclei

Lecture 4: Nuclear Energy Generation

Nuclear Astrophysics

The Nuclear Many-Body Problem

ROLE OF THE SHELL AND PAIRING EFFECTS IN NUCLEAR FISSION 1

c E If photon Mass particle 8-1

Theoretical and experimental α decay half-lives of the heaviest odd-z elements and general predictions

On the robustness of the r-process in neutron-star mergers against variations of nuclear masses

Nuclear structure and stellar weak interaction rates of nuclei below the 132 Sn core

Sunday. Monday Thursday Friday

The role of fission in the r-r process nucleosynthesis

COLD FUSION SYNTHESIS OF A Z=116 SUPERHEAVY ELEMENT

Microscopic Theories of Nuclear Masses

Observables predicted by HF theory

APPLICATIONS OF A HARD-CORE BOSE HUBBARD MODEL TO WELL-DEFORMED NUCLEI

Functional Orsay

Nuclear structure Anatoli Afanasjev Mississippi State University

8 Nuclei. introduc)on to Astrophysics, C. Bertulani, Texas A&M-Commerce 1

Influence of entrance channels on formation of superheavy nuclei in massive fusion reactions

Shape Coexistence and Band Termination in Doubly Magic Nucleus 40 Ca

arxiv: v1 [nucl-th] 19 Feb 2018

RFSS: Lecture 8 Nuclear Force, Structure and Models Part 1 Readings: Nuclear Force Nuclear and Radiochemistry:

Nucleon Pair Approximation to the nuclear Shell Model

Single universal curve for decay derived from semi-microscopic calculations

Hartree-Fock-Bogoliubov atomic mass models and the description of the neutron-star crust

arxiv: v1 [nucl-th] 24 Oct 2007

Nuclear Science Seminar (NSS)

Nuclear liquid-drop model and surface-curvature effects

Modern nuclear mass models

INTRODUCTION. The present work is mainly concerned with the comprehensive analysis of nuclear binding energies and the

What did you learn in the last lecture?

arxiv: v1 [nucl-th] 16 Jan 2014

LOW-LYING STATES OF ODD NUCLEI IN THE SOUTH WEST 208 Pb REGION

QRPA Calculations of Charge Exchange Reactions and Weak Interaction Rates. N. Paar

Particle-number projection in finite-temperature mean-field approximations to level densities

Symmetry Energy. in Structure and in Central and Direct Reactions

Quantum Theory of Many-Particle Systems, Phys. 540

Lesson 5 The Shell Model

Nuclear Symmetry Energy and its Density Dependence. Chang Xu Department of Physics, Nanjing University. Wako, Japan

Journal of Nuclear Sciences

Nuclear matter inspired Energy density functional for finite nuc

Study of Fission Barrier Heights of Uranium Isotopes by the Macroscopic-Microscopic Method

The effect of the tensor force on the predicted stability of superheavy

Nuclear physics input for the r-process

Role of fission in the r-process nucleosynthesis

Symmetry Energy in Surface

Relativistic Hartree-Bogoliubov description of sizes and shapes of A = 20 isobars

Clusters in Dense Matter and the Equation of State

PENINSULA OF NEUTRON STABILITY OF NUCLEI IN THE NEIGHBORHOOD OF N=258

arxiv: v1 [nucl-th] 12 Jul 2012

Structure properties of medium and heavy exotic nuclei

Symmetry energy of dilute warm nuclear matter

Transcription:

Journal of Physics: Conference Series An improved nuclear mass formula: WS3 To cite this article: Ning Wang and Min Liu 2013 J. Phys.: Conf. Ser. 420 012057 View the article online for updates and enhancements. Related content - Statistical errors in Weizsäcker-Skyrme mass model * Min Liu, Yu Gao and Ning Wang - Nuclear masses, deformations and shell effects Jorge G Hirsch, César A Barbero and Alejandro E Mariano - Weizsäcker Skyrme-type mass formula by considering radial basis function correction Na Na Ma, Hai Fei Zhang, Xiao Jun Bao et al. Recent citations - Neutron Star Mergers and Nucleosynthesis of Heavy Elements F.-K. Thielemann et al - Improvement in a phenomenological formula for ground state binding energies G. Gangopadhyay - Modification of magicity toward the dripline and its impact on electron-capture rates for stellar core collapse Ad. R. Raduta et al This content was downloaded from IP address 148.251.232.83 on 06/11/2018 at 12:52

An improved nuclear mass formula: WS3 Ning Wang and Min Liu Department of Physics, Guangxi Normal University, Guilin 541004, P. R. China E-mail: wangning@gxnu.edu.cn Abstract. We introduce a global nuclear mass formula which is based on the macroscopicmicroscopic method, the Skyrme energy-density functional and the isospin symmetry in nuclear physics. The rms deviation with respect to 2149 known nuclear masses falls to 336 kev, and the rms deviations from 1988 neutron separation energies and α-decay energies of 46 superheavy nuclei are significantly reduced to 286 and 248 kev, respectively. The predictive power of the mass formula for describing new measured masses in GSI and those in AME2011 is excellent. In addition, we introduce an efficient and powerful systematic method, radial basis function approach, for further improving the accuracy and predictive power of global nuclear mass models. 1. Introduction It is known that nuclear masses are of great importance for the study of super-heavy nuclei [1, 2, 3, 4], nuclear symmetry energy [5] and nuclear astrophysics [6, 7, 8]. For the synthesis of super-heavy nuclei (SHN), a reliable nuclear mass formula is required for providing some valuable information on the central position of the island of stability, the position of proton drip line and the properties of SHN at their ground states. In addition, most of masses for nuclei along the r-process path have not yet been measured. Accurate predictions for the masses of these extremely neutron-rich nuclei play a key role for the study of nuclear astrophysics. Available nuclear mass formulas include some global and local formulas. For the global formulas, the model parameters are usually determined by all measured masses and the masses of almost all bound nuclei can be calculated. Some global nuclear mass models have been successfully established, such as the finite range droplet model (FRDM) [9] which is based on the macroscopic-microscopic method, the Hartree-Fock-Bogliubov (HFB) model [10], the Duflo-Zuker formula [11, 12] and the Weizsäcker-Skyrme formula [13, 14, 15]. These models can reproduce the measured masses with accuracy at the level of 600 to 300 kev. The local mass formulas are generally based on algebraic or systematic approaches. They predict the masses of unknown nuclei from the masses of known neighboring nuclei, such as the Garvey-Kelson relations [16], the isobaric multiplet mass equation, the residual proton-neutron interactions [17, 18] and the image reconstruction technique (like the CLEAN algorithm [19] and the radial basis function method [20]). The main difficult of the local mass formulas is that the model errors rapidly increase for nuclei far from the measured nuclei. For neutron drip line nuclei and super-heavy nuclei, the differences between the calculated masses from these different models are quite large. It is therefore necessary to check the reliability and predictive power of the nuclear mass formulas. Published under licence by Ltd 1

2. Weizsäcker-Skyrme mass formula In Ref. [15] an improved nuclear mass formula, Weizsäcker-Skyrme mass formula, was proposed. The total energy of a nucleus is written as a sum of the liquid-drop energy, the Strutinsky shell correction E and the residual correction res, E(A, Z, β) = E LD (A, Z) k 2 ( ) 1 + b k βk 2 + E(A, Z, β) + res. (1) The liquid-drop energy of a spherical nucleus E LD (A, Z) is described by a modified Bethe- Weizsäcker mass formula, E LD (A, Z) = a v A + a s A 2/3 + E C + a sym I 2 A + a pair A 1/3 δ np + W (2) with the isospin asymmetry I = (N Z)/A and the Coulomb energy, E C = a c Z 2 A 1/3 ( 1 0.76Z 2/3). (3) In the original Bethe-Weizsäcker formula, the symmetry energy coefficient a sym is simply written as a constant. To consider the surface-symmetry energy of finite nuclei, the massdependence of the a sym was proposed in Refs. [21, 22]. To check the mass-dependence of a sym, we have studied the symmetry energy coefficients of finite nuclei based on the experimental data for masses [23]. The liquid-drop energy which is a function of mass number A and charge number Z can also be expressed as a function of A and the isospin asymmetry I. By performing a partial derivative of the liquid-drop energy per particle with respect to I, the symmetry energy coefficients of finite nuclei can be extracted. In addition to the mass-dependence of a sym, the isospin dependence of the symmetry energy coefficients of finite nuclei can also be clearly observed. To consider the isospin dependence of a sym, we proposed a new form for the symmetry energy coefficient [ a sym = c sym 1 κ ] 2 I + ξ, (4) A1/3 2 + I A with which the mass- and isospin-dependence of a sym can be reproduced reasonably well. The I term in a sym represents the traditional Wigner effect [23]. With the proposed form for a sym, we found that the accuracy of the liquid-drop formula can be improved by 5% and the value of c sym which represents the nuclear symmetry energy at normal density goes up to about 30 MeV (very close to the extracted value from other approaches). Furthermore, it was found that the model parameters in the liquid-drop formula with the proposed form of a sym are quite stable for different mass region [24]. The a pair term empirically describes the pairing effect with δ np = 2 I : N and Z even I : N and Z odd 1 I : N even, Z odd, and N > Z 1 I : N odd, Z even, and N < Z 1 : N even, Z odd, and N < Z 1 : N odd, Z even, and N > Z The influence of nuclear deformations on the liquid-drop energy is considered based on a parabolic approximation. We studied the variation of the energy of a nucleus as a function of β 2 and β 4 by using the Skyrme energy-density functional plus the extended Thomas-Fermi (5) 2

approach [25], and found that the parabolic approximation is applicable for small deformations. The terms with b k in Eq.(1) describe the contribution of nuclear deformation (including β 2, β 4 and β 6 ) to the macroscopic energy. Mass dependence of the curvatures b k is written as [13], b k = ( ) ( ) k k 2 g 1 A 1/3 + g 2 A 1/3, (6) 2 2 according to the Skyrme energy-density functional, which greatly reduces the computation time for the calculation of deformed nuclei. The microscopic shell correction E = c 1 E sh + I E sh (7) is obtained with the traditional Strutinsky procedure by setting the order p = 6 of the Gauss- Hermite polynomials and the smoothing parameter γ = 1.2 hω 0 with hω 0 = 41A 1/3 MeV. E sh and E sh denote the shell energy of a nucleus and of its mirror nucleus, respectively. The I term in E is to take into account the mirror nuclei constraint [14] from the isospin symmetry, with which the accuracy of the mass model can be significantly improved by 10% without introducing any new model parameters. The charge-symmetry and charge-independence of nuclear force implies the energies of a pair of mirror nuclei should be close to each other if removing the Coulomb term. Combining the macro-micro method, one finds that the difference of the shell correction for mirror nuclei should be close to each other. The experimental data also support this point. From the traditional Strutinsky shell correction calculations, we note that the difference of the shell correction reaches a few MeV for some nuclei. The new expression for the shell correction in which the shell energy of the mirror nucleus is also involved, is proposed to consider the mirror effect. In the calculations of the shell corrections, the single-particle levels are obtained under an axially deformed Woods-Saxon potential [26]. Simultaneously, the isospin-dependent spin-orbit strength is adopted based on the Skyrme energy-density functional, ( λ = λ 0 1 + N ) i A with N i = Z for protons and N i = N for neutrons, which strongly affects the shell structure of neutron-rich nuclei and super-heavy nuclei. In this formula, we also consider the Wigner effect of heavy nuclei coming from the approximate symmetry between valence protons and neutrons. We found that some heavy doubly magic nuclei, such as 132 Sn, 208 Pb and 270 Hs, lie along a straight line N = 1.37Z + 13.5. The shell corrections of nuclei are approximately symmetric along this straight line. The Wigner effect of heavy nuclei causes the nuclei along this line are more bound. The term W is to consider this effect and reduces the rms error by 5% (see Ref.[15] for details). Finally, we would like to introduce a very efficient and powerful systematic correction to the global nuclear mass formulas - radial basis function (RBF) correction [20]. The RBF approach is a prominent global interpolation and extrapolation scheme for scattered data fitting. It is widely used in surface reconstruction. Based on a global nuclear mass formula and known masses, the differences between model calculation and experimental data are analyzed. The aim of the RBF approach is to find a smooth function S(N, Z) to describe the differences M exp M th. Once the reconstructed function S is obtained, the revised masses for unmeasured nuclei are given by = M th + S. Here, M th denotes the calculated mass from a global nuclear mass model. We find that the RBF correction can improve the accuracy of a global mass formula by 10% to 40% without introducing any new model parameters. Mth RBF (8) 3

Table 1. rms deviations between data AME2003 [27] and predictions of some models (in kev). The line σ(m) refers to all the 2149 measured masses, the line σ(s n ) to the 1988 measured neutron separation energies S n, the line σ(q α ) to the α-decay energies of 46 super-heavy nuclei. FRDM HFB-17 DZ10 WS WS* DZ31 DZ28 WS3 σ(m) 656 581 561 516 441 362 360 336 σ(s n ) 399 506 342 346 332 299 306 286 σ(q α ) 566 916 284 263 1052 936 248 Reference [9] [10] [11] [13] [14] [11] [12] [15] 100 WS3 (a) (b) Proton number 80 60 40 20-1.2-1.0-0.80-0.60-0.40-0.20 0 0.20 0.40 0.60 0.80 1.0 1.2 20 40 60 80 100 120 140 160 Neutron number AME2003 New masses in AME2011 New measured data in GSI 20 40 60 80 100 120 140 160 Figure 1. (Color online) (a) Difference between the calculated masses with the WS3 formula [15] and the experimental data AME2003. (b) Positions of new measured masses. The blue and red squares denote the new measured nuclei in AME2011 [28] and those in GSI [29], respectively. 3. Results The rms deviations between the 2149 experimental masses AME2003 [27] and predictions of some models are calculated, σ(m) = [ 1 ( M exp (i) M (i) ) ] 2 1/2 th. (9) m We show in Table 1 the calculated rms deviations σ(m) (in kev). σ(s n ) denotes the rms deviation to the 1988 measured neutron separation energies S n. σ(q α ) denotes the rms deviation with respect to the α-decay energies of 46 super-heavy nuclei (Z 106) [14]. FRDM and HFB-17 denote the finite-range droplet model [9] and the latest Hartree-Fock-Bogoliubov (HFB) model with the improved Skyrme energy-density functional [10], respectively. DZ10, DZ28 and DZ31 4

denote the Duflo-Zuker mass models with 10, 28 and 31 parameters, respectively [11, 12]. The rms deviation from the 2149 masses with WS3 is remarkably reduced to 336 kev, much smaller than the results from the FRDM and the latest HFB-17 calculations, even lower than that achieved with the best of the Duflo-Zuker models. The rms deviation with respect to the 1988 neutron separation energies is reduced to 286 kev, and the rms deviation to the α-decay energies of 46 super-heavy nuclei is reduced to 248 kev, much smaller than the result of Duflo-Zuker formula (which is about one MeV). Fig. 1 (a) shows the difference between the experimental data AME2003 and the calculated masses with the WS3 formula [15] (M exp M th ). One sees that most of nuclei can be described remarkably well and the deviations for some semi-magic nuclei are slightly large. Table 2. rms deviations with respect to 154 new masses in AME2011 [28] and those to 53 new measured masses in GSI [29] from four mass models (in kev). FRDM HFB17 DZ28 WS3 154 new masses in AME2011 723 693 622 433 53 new masses in GSI 600 582 504 360 Table 3. The same as Table 2, but the radial basis function (RBF) corrections are involved. FRDM HFB17 DZ28 WS3 154 new masses in AME2011 475 596 442 397 53 new masses in GSI 193 250 287 229 Here, we also check the predictive power of different mass formulas for description of new masses. The crosses in Fig. 1(b) denote the available masses in AME2003. After 2003, 154 new measured data have been collected by Audi and Wang (blue squares) in the interim AME2011 [28] and 53 new masses are measured in GSI [29] (red squares). The rms deviations with respect to these new data are listed in Table 2. For the new masses in AME2011, the corresponding results of FRDM and HFB are about 700 kev and the result of DZ28 model is more than 600 kev. The result of WS3 is only 433 kev. For the 53 new data in GSI, the result from WS3 is also the smallest one (only 360 kev). Combining the radial basis function correction, we find that the accuracy of global mass formulas can be significantly improved. With the RBF corrections (see Table 3), the rms error to the new data in AME2011 is reduced from 723 kev to 475 kev and the rms error to the new data in GSI is reduced from 600 kev to about 200 kev based on the FRDM calculations. For other models, the improvement is also remarkable. Fig. 2 shows the difference between different model predictions and experimental data for the new masses measured in GSI [29]. The solid circles denote the results when the RBF corrections are involved. For these global nuclear mass models, especially the FRDM and HFB17 models, the RBF approach plays an important role for improving the systematic errors. Here, the measured masses in AME2003 are adopted for training the RBF [to obtain the function S(N, Z)] based on the leave-one-out cross-validation 5

M GSI - M (MeV) exp th 1.2 0.8 0.4 0.0-0.4-0.8-1.2 0.8 0.4 0.0-0.4-0.8-1.2 (a) WS3 WS3+RBF DZ28 DZ28+RBF rms = 360 229 kev FRDM FRDM+RBF 120 130 140 150 (b) HFB17 HFB17+RBF rms = 600 193 kev (c) rms = 504 287 kev (d) rms = 582 250 kev Neutron number 130 140 150 Figure 2. (Color online) Difference between model predictions and the experimental data for the new measured masses in GSI. Proton number 120 WS3 FRDM 100 80 60 40 20 120 100 DZ28 80 60 40 20 0 0 40 80 120 160 200 HFB17 0 40 80 120 160 200 Neutron number -2.00-1.75-1.50-1.25-1.00-0.750-0.500-0.250 0 0.250 0.500 0.750 1.00 1.25 1.50 1.75 2.00 Figure 3. (Color online) Radial basis function corrections for different global nuclear mass models. The dashed lines show the magic numbers. 6

Q th - Q exp (MeV) 1.0 0.5 0.0-0.5-1.0 FRDM WS3 106 108 110 112 114 116 118 Charge number Figure 4. (Color online) Difference between model predictions and the experimental data for the α-decay energies of super-heavy nuclei. The squares and circles denote the results of FRDM and WS3, respectively. method. Simultaneously, the Garvey-Kelson relation [16], which contains 12 estimates for a nucleus with the corresponding values of its 21 neighbors, is also adopted for further improving the smoothness of the function S(N, Z). For unmeasured nuclei, the masses are predicted by using Mth RBF = M th + S with the calculated mass M th from a global nuclear mass model. Fig. 3 shows the RBF corrections for different global nuclear mass models. For the super-heavy nuclei region, the RBF corrections are large for the FRDM and DZ28 model. For the DZ28 model, one can see from Table 1 that the rms deviation with respect to the known masses is only 360 kev, however, the rms deviation σ(q α ) with respect to the α-decay energies of 46 super-heavy nuclei goes up to 936 kev, which also implies that the systematic errors for the masses of super-heavy nuclei could be large in this model. It is known that the α-decay energies of super-heavy nuclei have been measured with a high precision, which provides us with useful data for testing mass models. Fig. 4 shows the difference between model predictions and the experimental data for the α-decay energies of SHN. Comparing with the FRDM, the α-decay energies of the super-heavy nuclei are much better reproduced with the WS3 model. The difference is within Q α = ±0.5 MeV for SHN with the WS3 formula. The corresponding result from the FRDM is ±1.0 MeV. Here, we would like to emphasize that the description for the α-decay energies of SHN is also a useful tool to test the predictive power of the mass models because the measured α-decay energies are not involved in the fit for the model parameters. Based on the WS3 formula, the shell corrections and the surface of the α-decay energies of super-heavy nuclei have been studied simultaneously. In Fig. 5, we show the calculated shell corrections E of nuclei. There are two islands in the super-heavy mass region. One is located at N = 162, Z = 108 and the other is located around N = 178, Z = 118. The triangles denote the synthesized SHN in Dubna through fusion reactions by using 48 Ca bombarding on actinide targets. The calculated deformations of nuclei demonstrate that the nuclei with N = 184 are (nearly) spherical in shape. However, the maximum of the shell corrections occurs at around N = 178 instead of N = 184. The shell corrections (in absolute value) for nuclei with Z = 126 are smaller than that for nucleus (N = 178, Z = 118) by more than one MeV. Furthermore, 7

Proton number 130 126 WS3-7.3 120 120-7.0 proton drip line -6.7-6.4 114-6.1 110-5.8 108-5.5-5.2 100-4.9-4.6 162 178 184 140 160 180 200 Neutron number Q (MeV) 14 12 10 8 Cn Ds Hs Sg WS3 120 114 116118 6 144 152 160 168 176 184 192 Neutron number Figure 5. Shell corrections of nuclei in super-heavy region from WS3 calculations. The black squares denote the nuclei with spherical shapes and the triangles denote the synthesized SHN in Dubna. The dark gray zigzag line denotes the calculated proton drip line. The dashed lines show the possible magic numbers. Figure 6. Predicted (ground state to ground state) α-decay energies of super-heavy nuclei. The thick and thin curves denote the results for even-z and odd-z nuclei, respectively. The two dashed lines show neutron numbers N = 162 and 184, and the solid line show the position of N = 178. nuclei with Z = 126 and N 184 locate beyond the calculated proton drip line, which indicates that the probability to synthesize nuclei with Z = 126 would be much smaller than that of produced super-heavy nuclei already. Fig. 6 shows the predicted α-decay energies of superheavy nuclei with the WS3 model. The two valleys at N = 162 and 184 can be clearly observed. One can also see the valley at N = 178. For SHN with Z = 116 118, the α-decay energies for nuclei with N = 178 are smaller than those for nuclei with N = 184. 4. Summary In this talk, we introduced a global nuclear mass formula, Weizsäcker-Skyrme mass formula, which is based on the Skyrme energy-density functional, the macroscopic-microscopic method and the isospin symmetry in nuclear physics. The rms deviations from 2149 measured masses and 1988 neutron separation energies are significantly reduced to 336 and 286 kev, respectively. As a test of the extrapolation of the mass model, the α-decay energies of 46 super-heavy nuclei have been systematically studied. The rms deviation with the proposed model falls to 248 kev, much smaller than the result 936 kev from the DZ28 model. Shell effect and isospin effect, especially isospin symmetry play a crucial role for improving the nuclear mass formula. For further testing the predictive power of the nuclear mass models, 154 new measured data in AME2011 and 53 new masses measured in GSI have been studied. The Weizsäcker-Skyrme mass formula can reproduce these new data remarkably well. We also introduced an efficient and powerful systematic method, radial basis function (RBF) approach, for further improving the accuracy and predictive power of global nuclear mass models. With the proposed mass formula, the shell corrections and α-decay energies of super-heavy nuclei have been studied simultaneously. The calculated shell corrections and α-decay energies of super-heavy nuclei imply that the shell gap at N = 178 also influences the central position of the island of stability for SHN. 8

5. Acknowledgments This work was supported by National Natural Science Foundation of China, Nos 11275052, 11005022, 10875031 and 10847004. The obtained mass table with the formula WS3 is available at http://www.imqmd.com/wangning/ws3.6.txt. The nuclear mass tables with the radial basis function corrections are available at http://www.imqmd.com/wangning/rbfmasses.txt. References [1] Wang N, Tian J, Scheid W, 2011 Phys. Rev. C 84 061601(R) [2] Zhang H, et al. 2012 Phys. Rev. C 85 014325 [3] Lu B, Zhao E, and Zhou S, 2012 Phys. Rev. C 85 011301(R) [4] Siwek-Wilczynska K et al. 2007 Int. J. Mod. Phys. E 16 493 [5] Liu M, Wang N, Li Z, Zhang F 2010 Phys. Rev. C 82 064306 [6] Arcones A and Martínez-Pinedo G 2011 Phys. Rev. C 83 045809 [7] Sun B, Montes F, Geng L. et al. 2008 Phys. Rev. C 78 025806 [8] Li Z, Niu Z, Sun B, Wang N, Meng J, 2012 Acta Phys. Sin. 61 017601 [9] Möller P, Nix J et al. 1995 At. Data and Nucl. Data Tables 59 185 [10] Goriely S, Chamel N and Pearson J 2009 Phys. Rev. Lett. 102 152503 [11] Duflo J and Zuker A 1996 at http://amdc.in2p3.fr/web/dz.html [12] Duflo J and Zuker A 1995 Phys. Rev. C 52 23 [13] Wang N, Liu M and Wu X 2010 Phys. Rev. C 81 044322 [14] Wang N, Liang Z, Liu M and Wu X 2010 Phys. Rev. C 82 044304 [15] Liu M, Wang N, Deng Y and Wu X 2011 Phys. Rev. C 84 014333 [16] Barea J, Frank A, Hirsch J et al. 2008 Phys. Rev. C 77, 041304 [17] Fu G, Jiang H, Zhao Y et al. 2010 Phys. Rev. C 82 034304 [18] Jiang H, Fu G, Sun B et al. 2012 Phys. Rev. C 85 054303 [19] Morales I, Isacker P et al. 2010 Phys. Rev. C 81, 024304 [20] Wang N and Liu M 2011 Phys. Rev. C 84 051303(R) [21] Myers W and Swiatecki W 1966 Nucl. Phys. 81 1 [22] Danielewicz P et al. 2009 Nucl. Phys. A 818 36 [23] Wang N and Liu M 2010 Phys. Rev. C 81 067302 [24] Hirsch J, Barbero C and Mariano A 2011 J. Phys.: Conf. Ser. 322 012017 [25] Liu M, Wang N et al. 2006 Nucl. Phys. A 768 80 [26] Cwoik S, Dudek J et al. 1987 Comp. Phys. Comm. 46 379 [27] Audi G, Wapstra A and Thibault C 2003 Nucl. Phys. A 729 337 [28] Audi G and Wang M (private communication). [29] Chen L, Plaß WR, Geissel H et al. 2012 Nucl. Phys. A 882 71 9