logo

Production of neutron-rich actinide isotopes in isobaric collisions via multinucleon transfer reactions

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Production of neutron-rich actinide isotopes in isobaric collisions via multinucleon transfer reactions

Peng‑Hui Chen
Chang Geng
Zu‑Xing Yang
Xiang‑Hua Zeng
Zhao‑Qing Feng
Nuclear Science and TechniquesVol.34, No.10Article number 160Published in print Oct 2023Available online 31 Oct 2023
43902

We systematically calculated the multinucleon transfer reactions of 208Os, 208Pt, 208Hg, 208Pb,208Po, 208Rn, 208Ra, and 132,136Xe when bombarded on 232Th and 248Cm at Coulomb barrier energies within the dinuclear system model. These results are in good agreement with the available experimental data. The influence of Coulomb and shell effects on actinide production in these reactions has been rigorously studied. We calculated and analyzed the potential energy surface (PES) and total kinetic energy (TKE) mass distributions for the reactions involving 208Hg, 208Pb, and 208Po with 248Cm and 232Th. The PES and TKE spectra shed light on the fragment formation mechanisms in multinucleon transfer reactions, with clear indications of isospin and shell effects. The production cross-sections for multinucleon transfer products show a strong dependence on isobar projectiles with a mass number A=208. Isobar projectiles with high N/Z ratios are advantageous for generating neutron-rich target-like fragments. Conversely, products induced by isobar projectiles with larger charge numbers tend to shift toward proton-rich regions. The intertwining of the Coulomb potential and shell effect is evident in the production cross-sections of actinide isotopes. Drawing from reactions induced by radioactive projectiles, we anticipate the discovery of several new actinide isotopes near the nuclear drip lines, extending our reach into the superheavy nuclei domain.

Dinuclear system modelIsobaric collisionsMultinucleon transfer reactionsNeutron-rich actinides
1

Introduction

To date, the synthesis of 11 isotopes – 149Lu [1], 207Th [2], 251,264Lr [3, 4], 166Pm, 168Sm, 170Eu, 172Gd [5], 204Ac [6], 39Na [7], and 286Mc [8] – in the last year has brought the total count to 3,327 known nuclei in the nuclide chart. This comprises 288 natural nuclides (254 stable isotopes with lifespans longer than the Earth’s age and 34 unstable nuclides) and 3,039 species of nuclei synthesized in global laboratories. These syntheses utilize techniques such as fusion-evaporation (FE), multinucleon transfer (MNT), deep inelastic reactions (DIR), projectile fragmentation (PF), spallation, fission (SF), neutron capture (NC), and thermonuclear tests (TT) [9]. However, various theoretical models predict the existence of an additional 8,000 to 10,000 unknown bound isotopes in the nuclei chart [10-12]. This suggests that at least over 5,000 nuclides remain to be discovered by nuclear experimentalists, particularly in the realms of nuclear drip lines and the stability islands of superheavy nuclei.

In recent years, laboratories across the globe have made significant advancements in nuclear synthesis. From an experimental standpoint, several new nuclear species have been produced: 207Th, 235Cm, 214U, 222Np, 211Pa, 280Ds[2, 13-15], and others via fusion-evaporation (FE) reactions; 110Zr, 121Tc, 129Pd, and more through projectile fragmentation (PF)[16]; and 223,229Am, 233Bk[17], among others, using multinucleon transfer (MNT) reactions. This study has garnered significant attention from leading research facilities, including the Lanzhou Heavy Ion Research Facility (HIRFL) in China, Joint Institute for Nuclear Research (JINR) in Russia, Helmholtz Centre for Heavy Ion Research (GSI) in Germany, Grand Accëlërateur National d’Ions Lourds (GANIL) in France, and Argonne National Laboratory (ANL) in the USA, actively working towards the synthesis of new nuclides in regions near the drip lines and within the superheavy region.

In efforts to elucidate the damped collision mechanism and predict the synthesis cross-sections of target nuclides, theorists have crafted intricate yet pragmatic models to represent the multinucleon transfer reactions occurring at incident energies close to the Coulomb barrier. Notable among these are the GRAZING model[18-20], the dinuclear system (DNS) model[21-30], and a dynamical model anchored in the Langevin equations[31, 32]. There are also microscopic methods that account for the degrees of freedom inherent to nucleons. Examples include the time-dependent Hartree (TDHF) approach[33-35] and the improved quantum molecular dynamics model (ImQMD)[36, 37]. Although each model possesses its distinctive attributes, they are all capable of faithfully replicating available experimental data. The DNS model, in particular, offers a comprehensive view, considering factors, such as the shell effect, dynamic deformations, fission, quasi-fission, deep-inelastic mechanisms, and the odd-even effect. Furthermore, it stands out for its exceptional computational efficiency.

In this study, we compared the calculated cross-sections of target-like fragments resulting from MNT reactions of 132,136Xe + 248Cm at incident energies proximate to the Coulomb barriers with the available experimental data, using the DNS model as a basis. To delve into the interplay between the Coulomb force and shell effect in the MNT process, we selected isobaric projectiles with a mass number A=208A=208 in proximity to the double magic nucleus 208Pb. These projectiles were directed to bombard targets 232Th and 248Cm at energies consistent with the Coulomb barrier. Our analysis further explores the production cross-sections of yet-to-be-identified actinide isotopes in isobaric collisions. The structure of this study is as follows: In Sec. 2, we provide a concise overview of the DNS model. Sec. 3 presents the calculation results and subsequent discussion. Finally, a summary and conclusions drawn from our study are detailed in Sec. 4.

2

Model Description

Initially, Volkov introduced a concept to depict deep inelastic heavy-ion collisions[38]. Later on, Adamian incorporated a quasi-fission component into the massive fusion process[39, 40]. Subsequent enhancements, including modifications to the relative motion energy and angular momentum of the colliding nuclei, in tandem with nucleon transfer, were developed within the DNS framework by the Lanzhou Group[41]. The production cross-sections of superheavy nuclei (SHN), quasi-fission, and fusion–fission dynamics have been comprehensively examined within the dynamical DNS model. The dynamic evolution of the colliding system involves sequential processes: overcoming the Coulomb barrier to form the DNS, adjusting the relative motion energy, angular momentum, mass, charge asymmetry, among other factors, within the potential energy surface, and finally, the de-excitation of primary fragments[42]. The production cross-section of the MNT fragments was determined as follows: σtr(Z1,N1,Ec.m.)=J=0Jmaxσcap(Ec.m.,J)f(B)×P(Z1,N1,J1,B)×Wsur(E1,J1,s)dB. (1) σcap(Ec.m.,J) denotes the cross-section of the DNS formation derived by the Hill–Wheeler formula with a barrier distribution function[43]. Wsur(E1,J1,s) denotes the survival probability of the fragment formation during the MNT process. Furthermore, s denotes the decay channels for fragments (Z1,N1) such as neutrons, protons, deuteron, alpha rays, and gamma rays. Ec.m. denotes the incident energy at the center of the mass frame. The highest angular momentum Jmax was calculated for the colliding system in the grazing configuration. The angular momentum J was considered at the initial collision configuration before dissipation. E1 and J1 denote the excitation energy and angular momentum of the fragment with proton number Z1 and neutron number N1, respectively, in DNS, respectively. Additionally, P(Z1,N1,J1,B) is the formation probability of fragments (Z1,N1). For the barrier distribution function, we adopted an asymmetric Gaussian[44] form. f(B)=1Nexp[(BBmΔ)2] (2) The quantities Δ and Bm were evaluated using Δ =(Bt+Bs)/2 and Bm=(Bt+Bs)/2. Furthermore, Bt and Bs represent the height of the Coulomb barrier and minimum point of deformation under tip-tip collision, respectively. The normalization constant satisfies f(B)dB=1.

In the DNS model, the solution for nucleon transfer and relative motion involves a set of microscopic derivations, and the master equations distinguish between protons and neutrons. The fragment distribution probability, P(Z1,N1,E1), represents the proton number Z1, neutron number N1, and excitation energy E1 for DNS fragment 1, which is described by the following master equation: dP(Z1,N1,E1,β,t)dt=Z1'WZ1,N1,β;Z1,N1,β'(t)[dZ1,N1P(Z1,N1,E1,β',t)dZ1,N1P(Z1,N1,E1,β,t)]+N1WZ1,N1,β;Z1,N1,β'(t)[dZ1,N1P(Z1,N1,E1,β',t)dZ1,N1P(Z1,N1,E1,β,t)][ΛA1,E1,tqf(Θ)+ΛA1,E1,tfis(Θ)]P(Z1,N1,E1,t) (3) Specifically, WZ1,N1,β;Z1',N1,β (WZ1,N1,β,;Z1,N1',β) is the mean transition probability from channel(Z1,N1,E1,β) to (Z1',N1,E1',β), [or (Z1,N1,E1,β) to (Z1,N1',E1',β)], and dZ1,Z1 denotes the microscopic dimension corresponding to the macroscopic state (Z1,N1,E1). The sum is considered over all possible proton and neutron numbers that fragment (Z1', N1') may take; however, only one nucleon transfer is considered in the model, with the relations Z1'=Z1±1 and N1'=N1±1. The quasi-fission width Λ qf and fission width Λfis were calculated using Kramers’ formula [45, 46].

The excited DNS creates a valence space where the valence nucleons are symmetrically distributed around the Fermi surface. Only particles in states within this valence space participate actively in nucleon transfer. The local excitation energy and nucleon transfer influence the transition probability. These are microscopically determined from the interaction potential within the valence space, as detailed in [47, 48]. WZ1,N1,β;Z1',N1,β'=τmem(Z1,N1,β,E1;Z1',N1β',E1')dZ1,N1dZ1',N12×ii|Z1',N1,E1',i'|V|Z1,N1,E1,i|2. (4) The neutron transition coefficient has the same formula. The relaxation time is calculated using the deflection function method [49]. The memory time τmem and V interaction elements can be found in [47].

The motion of the nucleons in the interacting potential is governed by a single-particle Hamiltonian[41, 50]: H(t)=H0(t)+V(t) (5) with H0(t)=KνKενK(t)ανK+(t)ανK(t) (6) V(t)=K,K'αK,βKuαK,βKααK+(t)αβK(t)=K,KVK,K(t) (7) Here, indices K and K’ (K,K’ = 1, 2) denote fragments 1 and 2, respectively. Furthermore, quantities εν K and uαK,βK represent the single-particle energy and interaction matrix elements, respectively. Single-particle states are defined with respect to the centers of the interacting nuclei and are assumed to be orthogonalized in the overlap region. Thus, the annihilation and creation operators depend on the time. The single-particle matrix elements are parameterized as follows: uαK,βK=UK,K(t){exp[12(εαK(t)εβK(t)ΔK,K(t))2]δαK,βK}. (8) Detailed calculations of these parameters and the mean transition probabilities are described in Ref. [41, 50]. ΔεK=4εK*gK, εK*=ε*AKA, gK=AK/12, (9) where ε* denotes the local excitation energy of the DNS. The microscopic dimensions of fragment (ZK,NK) were evaluated using the valence states NK = gKΔεK and valence nucleons mK = NK/2 (K=1,2) as: d(m1,m2)=(N1m1)(N2m2). (10) The local excitation energy E1 was derived from the dissipation energy coupled to the potential energy surface (PES) of the relative motion of the DNS. The excitation energy in the equilibrium stage was determined by dividing the fragments by their mass. The angular momentum of the main fragment was determined by the moment of inertia. The local excitation energy was evaluated using [47, 48] ε(t)=Ediss(t)(U({a})U({αEN})). (11) The quantities of the entrance channel, denoted as αEN, encompass proton and neutron numbers, quadrupole deformation parameters, and orientation angles. Specifically, they are represented as ZP, NP, ZT, NT, R, βP, βT, θP, and θT for the projectile and target nuclei by symbols P and T, respectively. The interaction time τint is obtained using the deflection function method [51]. The energy dissipated in the DNS exhibits an exponential increase. The potential energy surface (PES) of the DNS can be evaluated as follows: Udr(t)=Qgg+VC(Z1,N1;β1,Z2,N2,β2,t)+VN(Z1,N1,β1;Z2,N2,β2,t)+Vdef(t) (12) with Vdef(t)=12C1(β1βT(t))2+12C2(β2βP(t))2 (13) Ci=(λ1)(λ+2)RN2δ32πZ2e2RN(2λ+1). (14) where Qgg, derived from the negative binding energies of fragments (Zi,Ni), was calculated using the liquid drop model plus shell correction[11]. θi denotes the angle between the collision orientations and symmetry axes of the deformed nuclei. Furthermore, VC and VN were calculated using Wong’s formula [52] and the double-folding potential[53], respectively. The quadrupole deformations of the ground- state nuclei were obtained from Ref. [54]. Additionally, Vdef(t) denotes the deformation energy of DNS at reaction time t. The evolutions of quadrupole deformations of the projectile- and target-like fragments occurs from the initial configuration as βT(t)=βTexp(t/τβ)+β1[1exp(t/τβ)],βP(t)=βPexp(t/τβ)+β2[1exp(t/τβ)] (15) where the deformation relaxation was τβ=4×1021 s.

The total kinetic energy (TKE-mass) of the primary fragment was evaluated using the following expression: TKE=Ec.m.+QggEdiss, (16) where Qgg=MP+MTMPLFMTLF and Ec.m. denote the incident energy in the center-of-mass frame. Masses MP, MT, MPLF, and MTLF correspond to the projectile, target, projectile-like fragment, and target-like fragment, respectively.

The survival probability Wsur(E1,J1,s) is particularly important in the evaluation of the cross-section, which is usually calculated using a statistical model. The physical process for understanding excited nuclei is straightforward. However, the magnitude of the survival probability is strongly dependent on the ingredients in the statistical model, such as the level density parameter[55], separation energy [54], shell correction[54], and fission barrier[56, 57], and others. The excited fragments were cooled by evaporating γ-rays and light particles (including neutrons, protons, and α particles) in competition with fission[44]. The probability in the channel for evaporating the x-th neutron, y-th proton, and z- alpha particle can be expressed as follows: Wsur(E1*,x,y,z,J)=P(E1*,x,y,z,J)×i=1xΓn(Ei*,J)Γtot(Ei*,J)j=1yΓp(Ej*,J)Γtot(Ej*,J)k=1zΓα(Ek*,J)Γtot(Ek*,J). (17) where E1* and J denote the excitation energies evaluated from the mass tables in Refs. [11] and spin of the excited nucleus, respectively. The total width Γtot is the sum of the partial widths of the particle evaporation, γ emission, and fission. The excitation energy Es* before evaporating the sth particle can be evaluated as follows: Es+1*=Es*BinBjpBkα2Ts, (18) where the initial conditions are E1* and s=i+j+k. Furthermore, Bin, Bjp, and Bkα are the separation energies of the ith neutron, jth proton, kth alpha, respectively. The nuclear temperature Ti is obtained by Ei*=aTi2Ti, where a denotes the level-density parameter. The fission and particle decay widths were calculated using the Weisskopf evaporation theory and Bohr–Wheeler formula, respectively. The realization probability P(E1*,x,y,z,J) was calculated using Jackson’s formula[58].

3

Results and Discussion.

We calculated the production cross-sections of actinide isotope chains with atomic number Z = 93-100 in the collisions of 132,136Xe + 248Cm at incident energy Elab = 699-790 MeV, as shown in Fig. 1. Compared with the available experimental data for 129,132,136Xe + 248Cm, which are represented by a solid red star, solid blue circle, and solid black squares with error bars, respectively, our calculation of 136Xe + 248Cm, marked by solid black lines, and 132Xe + 248Cm, marked by dashed red lines, could reproduce the tendency of the cross-sectional distribution of actinide isotopic chains. From experimental data[59, 60], it was determined that projectile 129,132,136Xe isotopes induced reactions with the target 248Cm to provide actinide products that have a large overlap distribution area in the neutron-rich region. This was not clearly distinguishable as expected. From our calculation in terms of the deep-inelastic mechanism, relative proton-rich projectile 132 Xe-induced reactions tend to shift to the proton-rich region when compared to the experimental results. Based on the data presented in Figs. 1, target-like fragments have production cross-sections of magnitude levels from 100 millibarns to 10 nanobarns. When significantly distanced from the target, the formation cross-section of products below the target declines more gradually than that of trans-target products. This suggests that quasi-fission plays a more dominant role in these collisions. It is worth noting that our calculations have limitations: they rely on a model with free parameters both for calculating the primary fragment cross-section and for estimating the survival probability against fission.

Fig. 1
(Color online) Calculation and experiment results of production cross-sections of actinide isotopic chains with Z = 93–100 in reactions of 129,132,136Xe + 248Cm at Elab = 699-790 MeV. The available experimental data are considered from [59, 60], marked by solid black square for 136Xe induced reactions, solid red star for 132Xe induced reactions, solid blue circle for 129Xe induced reactions Our calculations for 136Xe induced reactions were shown by solid black lines, 132Xe induced reactions shown by dashed red lines.
pic

To investigate the competition between the Coulomb repulsive potential and shell effect in MNT reactions, we calculated the reactions of isobaric projectiles with A=208 bombarding targets 248Cm and 232Th at incident energy Ec.m.=1.1× VB. The calculation details of these collisions are presented below. The interaction potential between the colliding partners was combined with the Coulomb and nuclear potentials. In Fig. 2(a), interaction potential VCN of 208Pt + 248Cm, 208Hg + 248Cm, 208Pb + 248Cm, 208Po + 248Cm, and 208Rn + 248Cm reactions were marked by solid black, dashed red, dash-dot blue, dash-dot-dot green, and short dashed olive lines, respectively. The tendencies of VCN distributions for these collisions were similar. A larger Coulomb potential results in a larger interaction potential VCN. Specifically, VCN increases exponentially with decreasing distance R decreasing in the attraction region of the nuclear force, where it increases slowly. Nucleon transfer occurred in the touch configuration. Based on the deflection function, the sticking times of the colliding partners are calculated for all impact parameters [49], as shown in Fig. 2(b), which decreases exponentially as angular momentum increases. In these collisions, a relatively large Coulomb potential resulted in a longer sticking time with a fixed impact parameter. During the sticking time, the kinetic energy dissipates into the composite system to heat with the internal excitation energy, which increases exponentially with reaction time and reaches equilibrium at approximately 2× 10-21 s, as shown in Fig. 2(c).

Fig. 2
(Color online) In panel (a), solid black, red, blue, green, olive lines indicate the interaction potential of the tip-tip collisions as a function of surface distance in reactions induced by projectiles 208Hg, 208Pb, 208Po, 208Pt, and 208Rn, respectively, with target 248Cm; Panel (b) shows distributions of reaction time to the angular momentum of collisions for these five reaction systems at incident energy Ec.m. = 1.1 × VB, which decreases exponentially as angular momentum increases. In panel(c), for a given angular momentum L=50 across these five collision systems, the internal excitation energies are shown to increase exponentially with reaction time.
pic

After the colliding partners are captured, the dissipated kinetic energy, combined with the angular momentum in the DNS, allows them to diffuse along the potential energy surface (PES). This is followed by nucleon rearrangement between the colliding entities, determined by solving a set of master equations. The PES and driving potential were derived using Eqs.(12). These are composed of the Coulomb potential, binding energy, and nuclear potential. They are calculated using the Wong formula, the liquid-drop model with shell correction, and the double folding method, respectively, as cited in [42]. The driving potential of projectiles 208Hg, 208Pb, and 208Po on targets 248Cm and 232Th during tip-tip collision at a fixed distance is plotted as a function of mass asymmetry, denoted as η. Here, η=(ATAP)/(AT+AP). This is illustrated in Fig. 3(a)(e), represented by solid black, dashed red, and dash-dot blue lines, respectively.

Fig. 3
(Color online) Potential energy surface and driver potentials of projectiles 208Hg, 208Pb, and 208Po induced reactions with targets 248Cm and 232Th at tip-tip collisions are listed in Fig. 3. Specifically, 208Hg, 208Pb, and 208Po induced reactions were represented by solid black, dashed red, and dash-dot blue lines in panels (a) and (e), respectively. Potential energy surfaces for these collisions are shown in panels (b), (c), (d), (f), (g), (h), respectively. Open stars denote projectile–target injection points. These solid black lines represent valley trajectories on the two-dimensional potential energy surface.
pic

Open circles and stars represent the projectile–target injection points. Panels (a) and (e) show that the tendencies of the driving potential trajectories for these collisions are similar. Two pockets appeared at η = 0.2, 0 to derive the potentials of the target 248Cm-based reactions. One pocket in the driving potentials for the target 232Th-based reactions appears at η = 0.2. The neutron subshell number N=162 can potentially play a crucial role in pocket formation. For projectiles, such as 208Po, which are distant from the β- stable line, their initial points of interaction lie significantly away from their respective driving potential trajectories. As diffusion starts, they quickly converge toward the driving potential path. Generally, using the PES, one can broadly predict the spectral distribution trend across each isotope chain.

By solving a set of master equations, we derived the production probabilities of primary fragments with their respective excitation energies. These equations categorized the fragments based on mass number and kinetic energy as delineated in Eq. (16). These distributions are illustrated in Figs. 4, with driving potential trajectories represented as solid grey lines. From Fig. 4, we observe two prominent peaks in the high kinetic regions, situated around the projectile–target injection points. Additionally, cross-sections tend to concentrate in the valleys of the driving potential trajectories. Reactions involving projectiles 208Hg, 208Pb, and 208Po with targets 248Cm and 232Th at the incident energy Ec.m. = 1.1 × VB exhibited TKE-mass distributions that were both symmetric and expansive. The TKE-mass distribution spans broadly within the kinetic range of 500–800 MeV and mass range of 160–-280 MeV, suggesting a potential transfer of more than 30 nucleons.

Fig. 4
(Color online) Panels (a) through (f) display the calculated TKE-mass distribution of primary reaction products from head-on collisions involving projectiles 208Hg, 208Pb, and 208Po with targets 248Cm and 232Th at Ec.m. = 1.1 × VB. Driving potential trajectories are also incorporated.
pic

Utilizing the statistical evaporation program, we calculated the survival probability of the excited primary fragments, which in turn determined the production cross-sections of the secondary fragments. The production cross-sections of primary and secondary fragments, delineated by mass and charge numbers in the collisions from projectiles 208Hg, 208Pb, and 208Po with target 248Cm at Ec.m. = 1.1 × VB, are depicted in Fig. 5 panels (a) through (f). The solid blue and dashed red lines represent secondary and primary fragments, respectively, while regions of superheavy nuclei are highlighted with rectangular shading. Our findings show that primary fragments span a vast charge spectrum, even reaching the superheavy regions. Conversely, secondary fragment production was significantly dampened by de-excitation. This reduction is attributed to the fact that highly excited primary trans-target fragments, having minimal fission barriers, are prone to undergoing fission. Our predictions indicate cross-sections for superheavy nuclei (with atomic numbers Z = 104-116) exceeding 10 picobarns. The neutron subshell N=162 may be particularly influential, especially in the 208Po + 248Cm collision.

Fig. 5
(Color online) The calculated primary and secondary yields mass and charge distribution for 208Hg, 208Pb, and 208 Po-induced reactions with targets 248Cm at Ec.m. = 1.1 × VB were listed in panels (a), (b), (c), (d), (e), and (f), respectively. Dashed red and solid blue lines represented primary and secondary yields. The superheavy region (Z > 104) are shown by a rectangular shadow.
pic

Secondary production cross-sections for actinide target-like fragments, including isotopes of Actinium, Thorium, Protactinium, Uranium, Neptunium, Plutonium, Americium, Curium, Berkelium, Californium, Einsteinium, Fermium, Mendelevium, Nobelium, and Lawrencium, have been calculated. These calculations pertain to collisions involving projectiles 208Pt, 208Hg, 208Pb, 208Po, 208Rn, and 208Ra bombarding on targets 248Cm at Ec.m. = 1.1 × VB. The line representations for these projectiles in Fig. 6 are as follows: solid black for 208Pt, dashed red for 208Hg, dash-dot green for 208Pb, dashed-dot-dot blue for 208Po, and short dashed olive for 208Rn. Observations indicate that collisions characterized by a smaller Coulomb force lean towards the neutron-rich region, whereas those with a more substantial Coulomb force gravitate to neutron-deficient areas. Numerous previously unidentified actinide isotopes are predicted from reactions 208Pt+248Cm, 208Hg+248Cm, 208Pb+248Cm, 208Po+248Cm, 208Rn+248Cm, and 208Ra+248Cm. These predictions are detailed in Table 1. For the new neutron-rich actinide isotopes, 208Pt+248Cm reactions produce the largest cross-sections. However, 208Pt remains unclear. Notably, unknown actinide products are highly dependent on the Coulomb potential. The 208Rn+248Cm reactions result in the largest cross-sections for new neutron-deficient actinide isotopes. The open circles represent the new neutron-rich actinide nuclides.

Table 1
Calculated cross-sections of unknown actinide isotopes with Z=89–103 in the reactions of projectiles 208Pt, 208Hg, and 208Pb induced MNT reactions with target 248Cm at incident energy Ec.m. = 1.1× VB.
248Cm + 208Pt 208Hg 208Pb 248Cm+ 208Pt 208Hg 208Pb 248Cm+ 208Pt 208Hg 208Pb
237Ac 6.8 μb 24 nb   254Pu 1.3 μb     261Es 16 μb 2.2 μb 7.6 nb
238Ac 2.6 μb 6.9 nb   255Pu 0.2 μb     262Es 0.2 μb 0.1 μb 0.3 nb
239Ac 1 μb 1.8 nb   256Pu 8.4 nb     263Es 0.3 μb 22 nb 0.1 nb
240Ac 28 nb 0.1 nb   248Am 11 mb 50 mb 56 μb 264Es 23 nb 1.1 nb 30 pb
241Ac 2.8 nb 0.5 pb   249Am 4.5 mb 1 mb 1.1 μb 265Es 9.2 nb 30 pb
239Th 20 μb 0.2 μb 0.34 nb 250Am 3.9 mb 0.3 mb 74 nb 260Fm 1.3 μb 2.1 μb 0.3 μb
240Th 8.4 μb 64.8 nb 0.04 nb 251Am 1.1 mb 14 μb 6.6 nb 261Fm 2.4 μb 1.4 μb 0.1 μb
241Th 5 μb 9.6 nb   252Am 0.4 mb 1.5 μb   262Fm 1.6 μb 0.4 μb 16 nb
242Th 0.7 μb 0.94 nb   253Am 0.3 mb 0.2 μb   263Fm 0.7 μb 0.1 μb 2.8 nb
243Th 0.1 μb 0.02 nb   254Am 72 μb 16 nb   264Fm 0.1 μb 5.8 nb 0.08 nb
244Th 26 nb     255Am 17 μb 1.9 nb   265Fm 34 nb 0.9 nb
245Th 2.8 nb     256Am 2.7 μb 40 pb   266Fm 7 nb 0.1 nb
246Th 0.3 nb     257Am 0.3 μb     267Fm 0.9 nb 10 pb
240Pa 0.5 mb 7.8 μb 20. nb 258Am 9.7 nb     261Md 0.9 μb 1.9 μb 1.7 μb
241Pa 0.4 mb 4.9 μb 5.5 nb 252Cm 0.8 mb 0.6 mb 43 nb 262Md 0.6 μb 0.6 μb 0.2 μb
242Pa 0.2 mb 1.1 μb 0.5 nb 253Cm 0.2 mb 47 μb 1.9 nb 263Md 2.3 μb 1.2 μb 0.5 μb
243Pa 10 μb 0.2 μb 5 pb 254Cm 0.1 mb 9.7 μb 0.2 nb 264Md 0.5 μb 0.2 μb 62 nb
244Pa 28 μb 17 nb   255Cm 97 μb 2.8 μb 2 pb 265Md 0.4 μb 74 nb 18 nb
245Pa 3.3 μb 1 nb   256Cm 28 μb 0.3 μb   266Md 35 nb 3.7 nb 0.9 nb
246Pa 0.4 μb 20 pb   257Cm 11 μb 40 nb   267Md 23 nb 1.1 nb 40 pb
247Pa 74 nb     258Cm 1.4 μb 1.7 nb   268Md 1.6 nb 40 pb
248Pa 7.3 nb     259Cm 0.1 μb 40 pb   269Md 0.7 nb 8 pb
249Pa 0.6 nb     260Cm 0.8 nb     261No 6.3 nb 54 nb 47 nb
250Pa 9 pb     252Bk 0.2 mb 0.7 mb 36 μb 262No 12 nb 64 nb 219 nb
243U 0.9 mb 9.1 μb 18.4 nb 253Bk 0.3 mb 0.9 mb 1.9 μb 263No 0.1 nb 206 nb 68 nb
244U 0.7 mb 2.2 μb 2.2 nb 254Bk 0.2 mb 0.1 mb 0.1 μb 264No 0.2 nb 131 nb 47 nb
245U 2.2 mb 0.5 μb 70 pb 255Bk 0.1 mb 46 μb 14 nb 265No 0.1 nb 86 nb 5.4 nb
246U 46 μb 29 nb   256Bk 81 μb 12 μb 1.5 nb 266No 62 nb 16 nb 0.15 nb
247U 10 μb 2.6 nb   257Bk 0.1 mb 5.2 μb 0.3 nb 267No 16 nb 2.6 nb 0.01 nb
248U 1.7 μb 0.3 nb   258Bk 36 μb 0.4 μb 3 pb 268No 6 nb 0.4 nb
249U 0.2 μb     259Bk 22 μb 46 nb   269No 1.3 nb 70 pb
250U 14 nb     260Bk 2.5 μb 2. nb   267Lr 86 nb 31 nb 15 nb
251U 1.1 nb     261Bk 0.1 μb 20 pb   268Lr 12 nb 2.1 nb 0.6 nb
245Np 3.3 μb 1 nb   261Bk 6.1 nb     269Lr 17 nb 1.5 nb 0.1 nb
246Np 4.3 μb 20 pb   257Cf 27 μb 12 μb 0.3 μb 270Lr 1.2 nb 0.1 nb 6 pb
247Np 74 nb     258Cf 24 μb 4.5 μb 36 nb 271Lr 0.9 nb 80 pb 1 pb
248Np 7.3 nb     259Cf 23 μb 1.6 μb 6.5 nb
249Np 0.7 nb     260Cf 9.2 μb 0.1 μb 0.4 nb
248Pu 17 mb 41 μb 12.8 nb 261Cf 1.4 μb 15 nb 7 pb
249Pu 2.7 mb 11 μb 1.3 nb 262Cf 6.1 nb 0.3 nb
250Pu 0.3 mb 0.2 μb 0.019 nb 263Cf 0.6 nb
251Pu 0.1 mb 30 nb   258Es 7.4 μb 6.2 μb 1.1 μb
252Pu 48 μb 2.4 nb   259Es 29 μb 13 μb 0.3 μb
253Pu 13 μb 10 pb   260Es 11 μb 3 μb 34 nb        
Show more
Fig. 6
(Color online) Predicted isotopic distribution cross-sections for target-like fragments with atomic numbers Z = 89-103 resulting from the collisions of projectiles Pt, Hg, Pb, Po, and Rn with mass number A = 208 bombarding on target 248Cm at Ec.m. = 1.1 × VB. The patterns correspond to the following projectiles: solid black for Pt, dashed red for Hg, dash-dot green for Pb, dashed-dot-dot blue for Po, and short dashed olive for Rn. Predicted new actinide isotopes are highlighted with open circles.
pic

Figure 7 shows secondary production cross-sections of all the formed fragments in collisions of 208Os+248Cm, 208Pt+248Cm, 208Hg+248Cm, 208Pb+248Cm, 208Po+248Cm, 208Rn+248Cm, 208Ra+248Cm, and primary production cross-sections of 208Pb+248Cm at the incident energy Ec.m. = 1.1 × VB as N-Z panel. Panels (g) and (h) clearly show the de-excitation effects. Panels (a), (b), (c), (d), (e), (f), and (h) show that many new isotopes are predicted, including neutron-rich and neutron-deficient isotopes and even superheavy nuclei. The projectile-target injection points and all existing isotopes in the nuclide chart are represented by solid black triangles and open squares, respectively.

Fig. 7
(Color online) The production cross-sections of all secondary fragments formed in the collisions involving 208Pt + 248Cm, 208Hg + 248Cm, 208Pb + 248Cm, 208Po + 248Cm 208Rn + 248Cm, and 208Ra + 248Cm at the incident energy Ec.m. = 1.1 × VB, along with the primary fragments from 208Pb + 248Cm are listed in N-Z panels. Open stars denote projectile–target injection points.
pic
4

Conclusion

Using the DNS model framework, we systematically calculated the production cross-sections of MNT fragments in reactions involving projectiles such as 208Os, 208Pt, 208Hg, 208Pb,208Po, 208Rn, 208Ra, and 132,136Xe colliding with targets 232Th and 248Cm around Coulomb barrier energies. To investigate the isospin diffusion on the formation of actinide products during the MNT process, the same number of projectiles with A=208 were selected. Our calculation for 132,136Xe + 248Cm is consistent with the available experimental data. The sticking time for these colliding systems, inferred from deflection functions, was significantly influenced by the Coulomb force, especially at smaller impact parameters. Furthermore, PES and TKE of these reactions, which can contribute to predicting the tendency of cross-sectional diffusion, are discussed. A relatively large cross-section from TKE appears around the pockets in PES, where the neutron subshell N=162 is evident. The de-excitation process strongly depresses the primary cross-section of actinide isotopes up to four magnitude levels. The production cross-section of the new actinides is highly dependent on the N/Z ratio of the isobaric projectile. It was determined that the Coulomb force coupled with the shell effect plays a crucial role in the production of cross-sections of actinides products in MNT reactions. These five colliding systems predicted a wide array of previously unknown heavy isotopes, with accessible cross-sectional values even for superheavy nuclei within the charge numbers Z=104-110.

References
1. K. Auranen, A.D. Briscoe, L.S. Ferreira, et al.,

Nanosecond-scale proton emission from strongly oblate-deformed 149Lu

. Phys. Rev. Lett. 128, 112501 (2022). doi: 10.1103/PhysRevLett.128.112501
Baidu ScholarGoogle Scholar
2. H.B. Yang, Z.G. Gan, Z.Y. Zhang, et al.,

New isotope 207Th and odd-even staggering in α-decay energies for nuclei with Z>82 and N<126

. Phys. Rev. C 105, L051302 (2022). doi: 10.1103/PhysRevC.105.L051302
Baidu ScholarGoogle Scholar
3. T. Huang, D. Seweryniak, B.B. Back, et al.,

Discovery of the new isotope 251Lr: Impact of the hexacontetrapole deformation on single-proton orbital energies near the z=100 deformed shell gap

. Phys. Rev. C 106, L061301 (2022). doi: 10.1103/PhysRevC.106.L061301
Baidu ScholarGoogle Scholar
4. Y.T. Oganessian, V.K. Utyonkov, N.D. Kovrizhnykh, et al.,

First experiment at the super heavy element factory: High cross section of 288Mc in the 243Am+48Ca reaction and identification of the new isotope 264Lr

. Phys. Rev. C 106, L031301 (2022). doi: 10.1103/PhysRevC.106.L031301
Baidu ScholarGoogle Scholar
5. G.G. Kiss, A. Vitez-Sveiczer, Y. Saito, et al.,

Measuring the beta-decay properties of neutron-rich exotic pm, sm, eu, and gd isotopes to constrain the nucleosynthesis yields in the rare-earth region

. The Astrophysical Journal 936, 107 (2022). doi: 10.3847/1538-4357/ac80fc
Baidu ScholarGoogle Scholar
6. M. Huang, Z. Gan, Z. Zhang, et al.,

alpha decay of the new isotope 204Ac

. Physics Letters B 834, 137484 (2022). doi: 10.1016/j.physletb.2022.137484
Baidu ScholarGoogle Scholar
7. D.S. Ahn, J. Amano, H. Baba, et al.,

Discovery of 39Na

. Phys. Rev. Lett. 129, 212502 (2022). doi: 10.1103/PhysRevLett.129.212502
Baidu ScholarGoogle Scholar
8. Y.T. Oganessian, V.K. Utyonkov, N.D. Kovrizhnykh, et al.,

New isotope 286Mc produced in the 243Am+48Ca reaction

. Phys. Rev. C 106, 064306 (2022). doi: 10.1103/PhysRevC.106.064306
Baidu ScholarGoogle Scholar
9. M. Thoennessen, The Discovery of Isotopes, (Springer Nature, 2016)
10. 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

. Atom. Data Nucl. Data 144, 101488 (2022). doi: 10.1016/j.adt.2022.101488
Baidu ScholarGoogle Scholar
11. P. Möller, A.J. Sierk, T. Ichikawa, et al.,

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

. Atom. Data Nucl. Data 109-110, 1-204 (2016). doi: 10.1016/j.adt.2015.10.002
Baidu ScholarGoogle Scholar
12. P. Jachimowicz, M. Kowal, J. Skalski,

Properties of heaviest nuclei with 98≤Z≤126 and 134≤N≤192

. Atom. Data Nucl. Data 138, 101393 (2021). doi: 10.1016/j.adt.2020.101393
Baidu ScholarGoogle Scholar
13. A. Såmark-Roth, D.M. Cox, D. Rudolph, et al.,

Spectroscopy along flerovium decay chains: Discovery of 280Ds and an excited state in 282Cn

. Phys. Rev. Lett. 126, 032503 (2021). doi: 10.1103/PhysRevLett.126.032503
Baidu ScholarGoogle Scholar
14. Z.Y. Zhang, H.B. Yang, M.H. Huang, et al.,

New α-emitting isotope 214U and abnormal enhancement of α-particle clustering in lightest uranium isotopes

. Phys. Rev. Lett. 126, 152502 (2021). doi: 10.1103/PhysRevLett.126.152502
Baidu ScholarGoogle Scholar
15. J. Khuyagbaatar, F.P. Heßberger, S. Hofmann, et al.,

α decay of 243Fm143 and 245Fm145, and of their daughter nuclei

. Phys. Rev. C 102, 044312 (2020). doi: 10.1103/PhysRevC.102.044312
Baidu ScholarGoogle Scholar
16. N. Fukuda, T. Kubo, D. Kameda, et al.,

Identification of new neutron-rich isotopes in the rare-earth region produced by 345 MeV/nucleon 238U

. J. Phys. Soci. J. 87, 014202 (2018). doi: 10.7566/JPSJ.87.014202
Baidu ScholarGoogle Scholar
17. H.M. Devaraja, S. Heinz, O. Beliuskina, et al.,

Observation of new neutron-deficient isotopes with Z ≥ 92 in multinucleon transfer reactions

. Phys. Lett. B 748, 199-203 (2015). doi: 10.1016/j.physletb.2015.07.006
Baidu ScholarGoogle Scholar
18. F.P. Hessberger, V. Ninov, D. Ackermann, et al.,

Production of heavy residues in nuclear collisions leading to Zp + Zt ≥ 75 at bombarding energies of E/A = (11.4-15.3)MeV/u

. Nucl. Phys. A 568, 121-140 (1994). doi: 10.1016/0375-9474(94)90007-8
Baidu ScholarGoogle Scholar
19. A. Winther,

Dissipation, polarization and fluctuation in grazing heavy-ion collisions and the boundary to the chaotic regime

. Nucl. Phys. A 594, 203-245 (1995). doi: 10.1016/0375-9474(95)00374-A
Baidu ScholarGoogle Scholar
20. A. Winther,

Grazing reactions in collisions between heavy nuclei

. Nucl. Phys. A 572, 191-235 (1994). doi: 10.1016/0375-9474(94)90430-8
Baidu ScholarGoogle Scholar
21. Z.Q. Feng, G.M. Jin, J.Q. Li,

Production of heavy isotopes in transfer reactions by collisions of 238U+238U

. Phys. Rev. C 80, 067601 (2009). doi: 10.1103/PhysRevC.80.067601
Baidu ScholarGoogle Scholar
22. Z.Q. Feng,

Production of neutron-rich isotopes around N = 126 in multinucleon transfer reactions

. Phys. Rev. C 95, 024615 (2017). doi: 10.1103/PhysRevC.95.024615
Baidu ScholarGoogle Scholar
23. M. Huang, Z. Gan, X. Zhou, et al.,

Competing fusion and quasifission reaction mechanisms in the production of superheavy nuclei

. Phys. Rev. C 82, 044614 (2010). doi: 10.1103/PhysRevC.82.044614
Baidu ScholarGoogle Scholar
24. L. Zhu, P.W. Wen, C.J. Lin, et al.,

Shell effects in a multinucleon transfer process

. Phys. Rev. C 97, 044614 (2018). doi: 10.1103/PhysRevC.97.044614
Baidu ScholarGoogle Scholar
25. X.Q. Deng, S.G. Zhou,

Examination of promising reactions with 241Am and 244Cm targets for the synthesis of new superheavy elements within the dinuclear system model with a dynamical potential energy surface

. Phys. Rev. C 107, 014616 (2023). doi: 10.1103/PhysRevC.107.014616
Baidu ScholarGoogle Scholar
26. X.B. Yu, L. Zhu, Z.H. Wu, et al.,

Predictions for production of superheavy nuclei with Z = 105–112 in hot fusion reactions

. Nuclear Science and Techniques 29, 154 (2018). doi: 10.1007/s41365-018-0501-2
Baidu ScholarGoogle Scholar
27. P.H. Chen, F. Niu, Y.F. Guo, et al.,

Nuclear dynamics in multinucleon transfer reactions near coulomb barrier energies

. Nuclear Science and Techniques 29, 185 (2018). doi: 10.1007/s41365-018-0521-y
Baidu ScholarGoogle Scholar
28. F. Niu, P.H. Chen, H.G. Cheng, et al.,

Multinucleon transfer dynamics in nearly symmetric nuclear reactions

. Nuclear Science and Techniques 31, 59 (2020). doi: 10.1007/s41365-020-00770-1
Baidu ScholarGoogle Scholar
29. F. Niu, P.H. Chen, Z.Q. Feng,

Systematics on production of superheavy nuclei Z = 119-122 in fusion-evaporation reactions

. Nuclear Science and Techniques 32, 103 (2021). doi: 10.1007/s41365-021-00946-3
Baidu ScholarGoogle Scholar
30. P.H. Chen, H. Wu, Z.X. Yang, et al.,

Prediction of synthesis cross sections of new moscovium isotopes in fusion-evaporation reactions

. Nuclear Science and Techniques 34, 7 (2023). doi: 10.1007/s41365-022-01157-0
Baidu ScholarGoogle Scholar
31. V. Zagrebaev, W. Greiner,

Cross sections for the production of superheavy nuclei

. Nucl. Phys. A 944, 257-307 (2015). Special Issue on Superheavy Elements. doi: 10.1016/j.nuclphysa.2015.02.010
Baidu ScholarGoogle Scholar
32. D. Boilley, B. Cauchois, H. , et al.,

How accurately can we predict synthesis cross sections of superheavy elements

? Nucl. Sci. Tech. (2018). doi: 10.1007/s41365-018-0509-7
Baidu ScholarGoogle Scholar
33. C. Golabek, C. Simenel,

Collision dynamics of two 238U atomic nuclei

. Phys. Rev. Lett. 103, 042701 (2009). doi: 10.1103/PhysRevLett.103.042701
Baidu ScholarGoogle Scholar
34. C. Simenel,

Particle-number fluctuations and correlations in transfer reactions obtained using the balian-vénéroni variational principle

. Phys. Rev. Lett. 106, 112502 (2011). doi: 10.1103/PhysRevLett.106.112502
Baidu ScholarGoogle Scholar
35. X. Jiang, N. Wang,

Production mechanism of neutron-rich nuclei around N = 126 in the multi-nucleon transfer reaction 132Sn + 208Pb

. Chin. Phys. C 42, 104105 (2018). doi: 10.1088/1674-1137/42/10/104105
Baidu ScholarGoogle Scholar
36. J.L. Tian, X.Z. Wu, K. Zhao, et al.,

Properties of the composite systems formed in the reactions of 238U+238U and 232Th+250Cf

. Phys. Rev. C 77, 064603 (2008). doi: 10.1103/PhysRevC.77.064603
Baidu ScholarGoogle Scholar
37. K. Zhao, Z.X. Li, X.Z. Wu, et al.,

Production probability of superheavy fragments at various initial deformations and orientations in the 238U+238U reaction

. Phys. Rev. C 88, 044605 (2013). doi: 10.1103/PhysRevC.88.044605
Baidu ScholarGoogle Scholar
38. V.V. Volkov,

Deep inelastic transfer reactions — the new type of reactions between complex nuclei

. Phys. Rep. 44, 93-157 (1978). doi: 10.1016/0370-1573(78)90200-4
Baidu ScholarGoogle Scholar
39. G.G. Adamian, N.V. Antonenko, W. Scheid, et al.,

Treatment of competition between complete fusion and quasifission in collisions of heavy nuclei

. Nucl. Phys. A 627, 361-378 (1997). doi: 10.1016/S0375-9474(97)00605-2
Baidu ScholarGoogle Scholar
40. G.G. Adamian, N.V. Antonenko, W. Scheid, et al.,

Fusion cross sections for superheavy nuclei in the dinuclear system concept

. Nucl. Phys. A 633, 409-420 (1998). doi: 10.1016/S0375-9474(98)00124-9
Baidu ScholarGoogle Scholar
41. Z.Q. Feng, G.M. Jin, J.Q. Li, et al.,

Formation of superheavy nuclei in cold fusion reactions

. Phys. Rev. C 76, 044606 (2007). doi: 10.1103/PhysRevC.76.044606
Baidu ScholarGoogle Scholar
42. Z.Q. Feng, G.M. Jin, F. Fu, et al.,

Production cross sections of superheavy nuclei based on dinuclear system model

. Nucl. Phys. A 771, 50-67 (2006). doi: 10.1016/j.nuclphysa.2006.03.002
Baidu ScholarGoogle Scholar
43. P.H. Chen, F. Niu, W. Zuo, et al.,

Approaching the neutron-rich heavy and superheavy nuclei by multinucleon transfer reactions with radioactive isotopes

. Phys. Rev. C 101, 024610 (2020). doi: 10.1103/PhysRevC.101.024610
Baidu ScholarGoogle Scholar
44. P.H. Chen, Z.Q. Feng, J.Q. Li, et al.,

A statistical approach to describe highly excited heavy and superheavy nuclei

. Chin. Phys. C 40, 091002 (2016). doi: 10.1088/1674-1137/40/9/091002
Baidu ScholarGoogle Scholar
45. G.G. Adamian, N.V. Antonenko, W. Scheid,

Characteristics of quasifission products within the dinuclear system model

. Phys. Rev. C 68, 034601 (2003). doi: 10.1103/PhysRevC.68.034601
Baidu ScholarGoogle Scholar
46. P. Grangé, L. Jun-Qing, H.A. Weidenmüller,

Induced nuclear fission viewed as a diffusion process: Transients

. Phys. Rev. C 27, 2063-2077 (1983). doi: 10.1103/PhysRevC.27.2063
Baidu ScholarGoogle Scholar
47. P.H. Chen, Z.Q. Feng, F. Niu, et al.,

Production of proton-rich nuclei around Z = 84-90 in fusion-evaporation reactions

. Eur. Phys. J. A 53,. doi: 10.1140/epja/i2017-12281-x
Baidu ScholarGoogle Scholar
48. W. Nörenberg,

Quantum-statistical approach to gross properties of peripheral collisions between heavy nuclei

. Z. Phys. A 274, 241-250 (1975). doi: 10.1007/BF01437736
Baidu ScholarGoogle Scholar
49. J.Q. Li, G. Wolschin,

Distribution of the dissipated angular momentum in heavy-ion collisions

. Phys. Rev. C 27, 590-601 (1983). doi: 10.1103/PhysRevC.27.590
Baidu ScholarGoogle Scholar
50. W.F. Li, N. Wang, J.F. Li, et al.,

Fusion probability in heavy-ion collisions by a dinuclear-system model

. Eur. Phys. Lett. (EPL) 64, 750-756 (2003). doi: 10.1209/epl/i2003-00622-0
Baidu ScholarGoogle Scholar
51. G. Wolschin, W. Nörenberg,

Analysis of relaxation phenomena in heavy-ion collisions

. Z. Phys. A 284, 209-216 (1978). doi: 10.1140/epja/i2017-12281-x
Baidu ScholarGoogle Scholar
52. C.Y. Wong,

Interaction barrier in charged-particle nuclear reactions

. Phys. Rev. Lett. 31, 766-769 (1973). doi: 10.1103/PhysRevLett.31.766
Baidu ScholarGoogle Scholar
53. I.I. Gontchar, D.J. Hinde, M. Dasgupta, et al.,

Double folding nucleus-nucleus potential applied to heavy-ion fusion reactions

. Phys. Rev. C 69, 024610 (2004). doi: 10.1103/PhysRevC.69.024610
Baidu ScholarGoogle Scholar
54. P. Möller,

Nuclear ground-state masses and deformations

. Atomic Data and Nuclear Data Tables 59, 185-381 (1995). doi: 10.1006/adnd.1995.1002
Baidu ScholarGoogle Scholar
55. P. Demetriou, S. Goriely,

Microscopic nuclear level densities for practical applications

. Nuclear Physics A 695, 95-108 (2001). doi: 10.1016/S0375-9474(01)01095-8
Baidu ScholarGoogle Scholar
56. S. Cohen, W.J. Swiatecki,

The deformation energy of a charged drop: Part v: Results of electronic computer studies

. Annals of Physics 22, 406-437 (1963). doi: 10.1016/0003-4916(63)90385-3
Baidu ScholarGoogle Scholar
57. G.G. Adamian, N.V. Antonenko, S.P. Ivanova, et al.,

Analysis of survival probability of superheavy nuclei

. Phys. Rev. C 62, 064303 (2000). doi: 10.1103/PhysRevC.62.064303
Baidu ScholarGoogle Scholar
58. J.D. Jackson,

A schematic model for (p, xn) cross sections in heavy elements

. Can. J. Phys. 34, 767-779 (1956). doi: 10.1139/p56-087,
Baidu ScholarGoogle Scholar
59. R.B. Welch, K.J. Moody, K.E. Gregorich, et al.,

Dependence of actinide production on the mass number of the projectile: Xe +248Cm

. Phys. Rev. C 35, 204-212 (1987). doi: 10.1103/PhysRevC.35.204
Baidu ScholarGoogle Scholar
60. K.J. Moody, D. Lee, R.B. Welch, et al.,

Actinide production in reactions of heavy ions with 248Cm

. Phys. Rev. C 33, 1315-1324 (1986). doi: 10.1103/PhysRevC.33.1315
Baidu ScholarGoogle Scholar
Footnote

The authors declare that they have no competing interests.