logo

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

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

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

Peng-Hui Chen
Hao Wu
Zu-Xing Yang
Xiang-Hua Zeng
Zhao-Qing Feng
Nuclear Science and TechniquesVol.34, No.1Article number 7Published in print Jan 2023Available online 13 Jan 2023
40800

In the framework of the dinuclear system model, the synthesis mechanism of the superheavy nuclides with atomic numbers Z=112, 114, 115 in the reactions of projectiles 40,48Ca bombarding on targets 238U, 242Pu, and 243Am within a wide interval of incident energy has been investigated systematically. Based on the available experimental excitation functions, the dependence of calculated synthesis cross-sections on collision orientations has been studied thoroughly. The total kinetic energy (TKE) of these collisions with fixed collision orientation shows orientation dependence, which can be used to predict the tendency of kinetic energy diffusion. The TKE is dependent on incident energies, as discussed in this paper. We applied the method based on the Coulomb barrier distribution function in our calculations. This allowed us to approximately consider all the collision orientations from tip-tip to side-side. The calculations of excitation functions of 48Ca + 238U, 48Ca + 242Pu, and 48Ca + 243Am are in good agreement with the available experimental data. The isospin effect of projectiles on production cross-sections of moscovium isotopes and the influence of the entrance channel effect on the synthesis cross-sections of superheavy nuclei are also discussed in this paper. The synthesis cross-section of new moscovium isotopes 278-286Mc was predicted to be as large as hundreds of pb in the fusion-evaporation reactions of 35,37Cl + 248Cf, 38,40Ar + 247Bk, 39,41K + 247Cm, 40,42,44,46Ca + 243Am, 45Sc + 244Pu, and 46,48,50Ti + 237Np, 51V + 238U at some typical excitation energies.

Dinuclear system modelSuperheavy nucleiComplete fusion reactionsProduction cross-section
1

Introduction

Since the "island of stability" of superheavy nuclei was predicted by the shell model in the 1960s[1], the synthesis of superheavy nuclei has been an exciting frontier field in the laboratories that could provide a unique tool to explore the properties of nuclei and nuclear structure under extremely strong Coulomb force. However, owing to the extremely low production cross-sections, the synthesis of superheavy nuclei in current experiments is time-consuming and costly. Therefore, it is particularly necessary to make reliable theoretical calculations that provide a reasonable reference for experiments. In recent years, synthesizing superheavy elements through low-energy heavy ion collisions near the Coulomb barrier has attracted extensive attention from theorists and experimentalists.

On the experimental side, in the past half century, fifteen superheavy elements characterized by values of Z in the range 104-118 have been synthesized and identified in laboratories all over the world[2]. Generally, superheavy synthesis methods are classified by the excitation energy of compound nuclei as cold fusion and hot fusion, resulting in compound nuclei surviving by emitting 1-2 neutrons and 3-5 neutrons, respectively, against fission. Elements Rf, Db, Sg, Fl, Mc, Lv, Ts, and Og were synthesized first in hot-fusion reactions. 257,258,259Rf (Z=104) was discovered simultaneously in Dubna[3] and Berkeley[4] in the reactions of 249Cf(12,13C, 3-4n)257,258,259Rf at incident energy Elab = 10.4 MeV/nucleon. 260,261Db (Z=105) was discovered simultaneously in Dubna[5] and Berkeley[6] in the reactions of 249Cf(15N,4n)260Ds at Elab = 85 MeV and 243Am(22Ne, 4n)261Ds at Elab = 114 MeV. 263Sg (Z=106) was discovered at Berkeley[7] in the reactions of 249Cf(18O,4n)263Sg at Elab = 95 MeV. 286-289Fl was essentially discovered at Dubna[8] in the reactions of 244Pu(48Ca, 3-6n)286-289Fl at Elab = 352.6 MeV. 288Mc (Z=115) was essentially discovered at Dubna[9] in the reactions of 243Am(48Ca, 3n)288Mc at Elab = 248, 253 MeV. 286-289Lv (Z=116) was discovered at Dubna[10] in the reactions of 245Cm(48Ca, xn)293-xFl at Elab = 243 MeV. 293-294Ts (Z=117) was essentially discovered at Dubna[11] in the reactions of 249Bk(48Ca, 3-4n)293-294Fl at Elab = 247, 252 MeV. 294Og (Z=118) was essentially discovered at Dubna[12] in the reactions of 249Cf(48Ca, 3n)294Og at Elab = 251 MeV. Elements Sg, Bh, Hs, Mt, Ds, Rg, Cn, and Nh were synthesized first in cold-fusion reactions. 259Sg (Z=106) was discovered at Dubna[13] in the reactions of 207Pb(54Cr,2n)259Sg at Elab = 262 MeV. 262Bh (Z=107) was essentially discovered at Gesellschatt Für Schwerionenforschung (GSI)[14] in the reactions of 209Bi(54Cr, 1n)262Bh at Elab = 4.85 MeV/u. 263-265Hs (Z=108) was synthesized at GSI[15] in the reactions of 208Pb(58Fe, 2n)265Hs at Elab = 5.02 MeV/u. 266Mt (Z=109) was synthesized at GSI[16] in the reactions of 209Bi(58Fe, 1n)266Mt at Elab = 5.15 MeV/u. 269Ds (Z=110) was synthesized at GSI[17] in the reactions of 208Pb(62Ni, 1n)269Ds at Elab = 311 MeV. 272Rg (Z=111) was synthesized at GSI[18] in the reactions of 209Bi(64Ni, 1n)272Rg at Elab = 318, 320 MeV. 277Cn (Z=112) was synthesized at GSI[19] in the reactions of 208Pb(70Zn, 1n)277Cn at Elab = 344 MeV. 278Nh (Z=113) was synthesized at RIKEN[20] in the reactions of 209Bi(70Zn, 1n)278Nh at Elab = 352.6 MeV. Synthesis information of the most neutron-rich and proton-rich superheavy nuclei with atomic numbers Z=104-118, including elements, isotopes, reactions, channels, laboratories, and year, is provided in Table 1. Chinese superheavy nuclei group synthesized the superheavy isotopes of 258,259Db[21], 264,265,266Bh[22], and 271Ds[23] at the Institute of Modern Physics (IMP) (Lanzhou, China).

Table 1
Synthesis information of the most neutron-rich and proton-rich superheavy isotopes with atomic numbers Z=104-118: production reactions, evaporation channel, laboratory, year, and reference.
Element Isotopes Reactions Channel Lab Year Ref.
Rf(104) 253 Rf 50 Ti + 204 Pb 1n GSI 1997 [24]
(13) 267Rf 48Ca + 242Pu α Dubna 2004 [25]
Db(105) 256Db 50Ti + 209Bi 3n GSI 2001 [26]
(11) 270Db 48Ca + 249Bk 3nα Berkeley 2010 [11]
Sg(106) 258Sg 51V + 209Bi 2n GSI 1997 [24]
(12) 271Sg 48Ca + 238U α Dubna 2004 [25]
Bh(107) 260Bh 52Cr + 209Bi α Berkeley 2008 [27]
(10) 274Bh 48Ca + 249Bk 3nα Dubna 2010 [11]
Hs(108) 263Hs 56Fe + 208Pb 1n Berkeley 2009 [28]
(12) 277Hs 48Ca + 244Pu 3nα GSI 2010 [29]
Mt(109) 266Mt 58Fe + 209Bi 1n GSI 1982 [30]
(7) 278Mt 48Ti + 249Bk 3nα Dubna 2010 [11]
Ds(110) 267Ds 59Co + 209Bi 1n Berkeley 1995 [31]
(8) 281Ds 48Ca + 244Pu 3nα Dubna 2004 [10]
Rg(111) 272Rg 64Ni + 209Bi 1n GSI 1995 [32]
(7) 282Rg 48Ca + 249Bk 3nα Dubna 2010 [11]
Cn(112) 277Cn 70Zn + 208Pb 1n GSI 1996 [19]
(6) 285Cn 48Ca + 244Pu 3nα Dubna 2004 [10]
Nh(113) 278Nh 70Zn + 209Bi 1n RIKEN 2004 [33]
(6) 286Nh 48Ca + 249Bk 3nα Dubna 2010 [11]
FI(114) 285FI 48Ca + 242Pu 5n Berkeley 2010 [8]
(5) 289FI 48Ca + 244Pu 3n Dubna 2004 [10]
Mc(115) 287Mc 48Ca + 243Am 4n Dubna 2004 [9]
(4) 290Mc 48Ca + 249Bk 3nα Dubna 2010 [11]
Lv(116) 290Lv 48Ca + 245Cm 3n Dubna 2004 [10]
(4) 293Lv 48Ca + 245Cm 1n Dubna 2004 [10]
Ts(117) 293Ts 48Ca + 249Bk 4n Dubna 2010 [11]
(2) 294Ts 48Ca + 249Bk 3n Dubna 2010 [11]
Og(118) 294Og 48Ca + 249Cf 3n Dubna 2006 [12]
Show more

The mechanism of fusion-evaporation (F.E.) cannot easily reach the next new period in the periodic table of elements because of the limited available combinations of projectile-target. With the development of suitable separation and detection techniques, the multinucleon transfer (MNT) mechanism might be the most promising method to synthesize unknown superheavy elements. This mechanism has been applied to produce massive heavy and superheavy isotopes[34]. Laboratories all over the world such as IMP [35], GSI [36, 37], Dubna[38, 39], RIKEN[40-42], and Lawrence Berkeley National Laboratory (LBNL)[43, 43] are focused on synthesizing new superheavy elements and their isotopes. From the chart of nuclei, in the superheavy region, there are substantial isotopes of superheavy elements that are still unknown. One of the objectives of the present study was to predict the production cross-sections of moscovium isotopes in F.E. reactions based on different combinations of projectile-target.

On the theoretical side, to describe the production mechanism of superheavy nuclei, some theoretical models were built, for example, the time-dependent Hartree-Fock model [44-46], the improved quantum molecular dynamics model[47-49], a dynamical approach based on Langevin equations[50, 51], and the dinuclear system (DNS) model [52-56]. The calculations resulting from these theoretical models are in good agreement with the available experimental data, which have their own features. In this study, the DNS model has been applied. This model has some advantages such as better consideration of the shell effect, dynamical deformation, fission, quasi-fission, deep-inelastic and odd-even effects, and high calculation efficiency. In previous studies[53, 52, 56-62], the DNS model accurately reproduced the available experimental results and predicted the synthesis production cross-sections of superheavy elements and exotic heavy nuclei in the mechanisms of F.E. and MNT reactions.

In this study, we investigated the dependence of the evaporation residue cross-sections on collision orientations and the influence of entrance channel effect on the evaporation residue cross-sections. We propose a Gaussian-like barrier distribution function for treating the problem of collision orientation dependence. The article is organized as follows. Section 2 briefly describes the DNS model. Calculated results and discussions are presented in Sect. 3. Section 4 concludes the paper with a summary.

2

Model Description

Initially, the DNS concept was proposed to describe the deep-inelastic reaction mechanism, which is a molecular-like configuration of two colliding partners, keeping their own individuality in the collision process. The DNS model has been widely used to describe F.E. and multinucleon transfer reactions. The complete fusion evaporation reaction can be described in terms of three processes. First, the colliding partners overcome the Coulomb barrier to form the composite system. Second, the kinetic energy and angular momentum dissipate into the composite system to enable the nucleon transfer between the touching colliding partners. Finally, all the nucleons are transferred from projectile nuclei to the target nuclei, which could form the compound nuclei with excitation energy and angular momentum. The highly excited compound nuclei will be de-excited by evaporation of the light particles (i.e., neutrons, γ-rays, and light-charged particles) or fission. Based on the DNS model, the evaporation residual cross-sections of superheavy nuclei can be expressed as σER(Ec.m.)=π22μEc.m.J=0Jmax(2J+1)T(Ec.m.,J)PCN(Ec.m.,J)Wsur(Ec.m.,J) (1) where the penetration probability T(Ec.m., J) is the probability that the collision system overcomes the Coulomb barrier, which is calculated using the empirical coupling channel model[52]. The fusion probability PCN(Ec.m., J) is the probability to form compound nuclei[63, 64]. The survival probability Wsur is the probability that the highly excited compound nuclei survive by evaporating light particles against fission. The maximal angular momentum is set as Jmax = 30-50 because the fission barrier for the superheavy nuclei may vanish at high spin [65].

2.1
Capture probability

The capture cross-sections of the two colliding partners are expressed as σcap(Ec.m.)=π22μEc.m.J(2J+1)T(Ec.m.,J). (2) Here, the penetration probability T(Ec.m.,J) is evaluated by the Hill-Wheeler formula [66] using the barrier distribution function: T(Ec.m.,J)=f(B)11+exp{2πω(J)[Ec.m.B2J(J+1)2μRB2(J)]}dB, (3) where ω(J) is the width of the parabolic barrier at position RB(J). The normalization constant is with respect to the relation f(B)dB=1. The barrier distribution function is assumed to be in an asymmetric Gaussian form[52, 67]: f(B)={1Nexp[(BBmΔ1)]B<Bm,1Nexp[(BBmΔ2)]B>Bm, (4) where Δ2 = (B0-Bs)/2, Δ1= Δ2-2 MeV, Bm=(B0+Bs)/2, and B0 and Bs are the Coulomb barriers of the side-side collision and saddle-point barriers in dynamical deformations[67]. The nucleus-nucleus interaction potential is expressed as V({α})=VC({α})+VN({α})+Vdef (5) with Vdef=12C1(β1β10)2+12C2(β2β20)2). where 1 and 2 represent the projectile and target, respectively; R=R1+R2+s and s are the distances between the center and surface of the projectile and target, respectively; R1 and R2 are the radii of the projectile and target, respectively; β1(2)0 represents the static deformation of the projectile-target; β1(2) represents the adjustable quadrupole deformation, which is varied to find the minimal V({α}); and {α} stands for {R,β1,β1,β2,β1,β2}. To reduce the number of deformation variables, we assume that the deformation energy of the colliding system is proportional to its mass[67], that is, C1β12/C2β22=A1/A2. Thus, only one deformation parameter, β=β1+β2, is required. The stiffness parameters Ci(i=1,2) are calculated using the liquid-drop model[68] through the following parameterization formula: Ci=(λ1)[(λ1)Ri2σ32πZ2e2Ri(2λ+1)] (6) where Ri is the radius of the spheroidal nucleus given by Ri=1.18 Ai1/3 (i=1,2). In this study, the quadrupole deformation was taken into account (λ=2). Note that σ is the coefficient of surface tension that fits 4πi2 σ=asAi2/3, where as = 18.32 MeV is the surface energy. The nuclear potential is calculated using the double-folding method [63-65]: VN=C0{FinFexρ0[ρ12(r)ρ2(rR)dr+ρ1(r)ρ22(rR)dR]+Fexρ1(r)ρ2(rR)dr} (7) where Fin(ex)=fin(ex)+fin(ex),N1Z1A1N2Z2A2. Note the dependence on the nuclear density and orientation of the deformed colliding partners. We set the following parameter values in our calculations: C0 = 300 MeV fm3, fin = 0.09, fex=-2.59, fin, = 0.42, fex, = 0.54, and ρ0 = 0.16 fm-3. The Woods-Saxon density distribution is expressed as ρ1(r)=ρ01+exp((r1(θ1))/a1) (8) and ρ2(rR)=ρ01+exp[(|rR|2(θ2))/a2] (9) where i(θi)(i=1,2) denotes the surface radii of the nuclei given by i(θi)= [1+βiY20(θi)], where Ri is the spheroidal nuclei radius and ai is the surface diffuseness coefficient, which was set as 0.55 fm in our calculations. The Coulomb potential was derived by Wong’s formula as follows [69]: VC({α})=Z1Z2e2r+(920π)1/2(Z1Z2e2r3)i=12Ri2βiP2(cosθi)+(37π)(Z1Z2e2r3)i=12Ri2(βiP2cosθi)2 (10) where θi, βi, Ri, and P2(cosθi) are the angle between the symmetry axis of the deformed projectile-target and collision axis, quadrupole deformation, radius of the projectile-target, and Legendre polynomial, respectively. Wong’s formula is in good agreement with the double-folding method.

2.2
Fusion probability

The composite system is formed after the capture process in which the dissipation of kinetic energy and angular momentum takes place to activate the transfer of nucleons in the touching configuration of the projectile target that results in mass probability diffusion. The mass probability of the formed fragments was evaluated by solving a set of master equations. The term of mass probability P(Z1,N1,E1,t) contains the proton number, neutron numbers of Z1 and N1, and internal excitation energy of E1 for a given fragment A1. The master equation is [64, 57, 70] dP(Z1,N1,E1,t)dt=Z1WZ1,N1;Z1,N1(t)[dZ1,N1P(Z1,N1,E1,t)dZ1,N1P(Z1,N1,E1,t)]+N1WZ1,N1;Z1,N1(t)[dZ1,N1P(Z1,N1,E1,t)dZ1,N1P(Z1,N1,E1,t)][ΛA1,E1,tqf(Θ)+ΛA1,E1,tfis(Θ)]P(Z1,N1,E1,t). (11) where WZ1,N1,Z1,N1 (WZ1,N1,Z1,N1) denotes the mean transition probability from the channel (Z1,N1,E1) to (Z1,N1,E1) [or (Z1,N1,E1) to (Z1,N1,E1)]; dZ1,N1 denotes the microscopic dimension corresponding to the macroscopic state (Z1,N1,E1). The sum contains all possible numbers of protons and neutrons for fragment (Z1, N1). However, only one nucleon transfer at one time is assumed in the model with relations Z1 = Z1 ± 1, and N1 = N1 ± 1. The excitation energy E1 is the local excitation energy #x03B5;1* for fragment (Z1, N1), which is derived by the dissipation of the relative motion along with the potentical energy surface (PES) of the DNS [71]. The time of the dissipation process is evaluated by the parameterization classical deflection function [72]. The motion of nucleons in the interaction potential is governed by the single-particle Hamiltonian: H(t)=H0(t)+V(t) (12) where the total single-particle energy and interaction potential are H0(t)=KνKενK(t)ανK+(t)ανK(t) (13) V(t)=K,K'αKβKuαKβKααK+(t)αβK(t)=K,KVK,K(t). (14) where ενK and uαK,βK represent the single-particle energies and interaction matrix elements, respectively, in which the single-particle state is defined as center of colliding nuclei assumed to be orthogonal in the overlapping region. The annihilation and creation operators are time-dependent. The single-particle matrix elements are parameterized as uαK,βK=UK,K(t)×{exp[12(εαK(t)εβK(t)ΔK,K(t))2]δαK,βK} (15) where UK,K(t) and δαK,βK(t) are described in Ref.[73]. The proton transition probability is microscopically derived from WZ1,N1;Z1',N1=τmem(Z1,N1,E1;Z1',N1,E1')dZ1,N1dZ1',N12×ii'|Z1',N1,E1',i'VZ1,N1,E1,i|2. (16) The neutron transition probability has a similar formula. The memory time and interaction elements V are described in Ref.[63].

The evolution of the DNS along distance R leads to quasi-fission. The decay probability of quasi-fission is calculated based on the one-dimensional Kramers equation as [73, 74] ΛA1,E1,tqf(Θ)=ω2πωBqf[(Γ2)2+(ωBqf)2Γ2]×exp[Bqf(A1)Θ(A1,E1,t)] (17) where Bqf(A1) is the quasi-fission barrier; ω and ωBqf are the frequencies of the harmonic oscillator approximation at the bottom and top of the interaction potential pocket, which are constants expressed as ωBqf = 2.0 MeV and ω = 3.0 MeV in this study; Γ = 2.8 MeV is the quantity characterizing the average double width of the single-particle state. The local temperature is expressed using the Fermi gas model, i.e., Θ=(ε*/(A/12))1/2. In the nuclear collision process, heavy fragments might lead to fission; the fission probability is calculated by the Kramers formula: ΛfisA1,E1,t(Θ)=ωg.s2πωf[(Γ02)2+(ωf)2Γ02]×exp[Bf(A1)Θ(A1,E1,t)], (18) where ωg.s. and ωf are the frequencies of the oscillators approximating the fission-path potential at the ground state and top of the fission barrier for fragment A1, respectively, which were set as ωg.s. = ωf = 1.0 MeV and Γ0 = 2 MeV. The fission barrier is calculated by the macroscopic part plus the shell correction energy. In the relaxation process of the relative motion, the DNS is excited by the dissipation of the relative kinetic energy and angular momentum. The excited composite system opens a valence space Δ εK in fragment K (K = 1, 2) that has a symmetrical distribution around the Fermi surface. The nucleons in the valence space are actively enabled to be excited and transferred. The average of these quantities is performed in the valence space: ΔεK=4εK*gK,εK*=ε*AKA,gK=AK/12, (19) where ε* is the local excitation energy of the DNS, which provides the excitation energy for the mean transition probability. There are NK = gKΔεK valence states and mK = NK/2 valence nucleons in the valence space ΔεK, leading to dimensions d(m1,m2)=(N1m1)(N2m2). (20) The local excitation energy is expressed as ε*=Ex(Udr(A1,A2)Udr(AP,AT)), (21) where Udr(A1,A2) and Udr(AP,AT) are the driving potentials of fragments A1, A2 and AP, AT, respectively. The detailed calculations of the driving potentials are given by Eq. 22. The excitation energy Ex of the composite system is converted from the relative kinetic energy dissipation[64]. The PES of the DNS is expressed as Udr(A1,A2;J,θ1,θ2)=B1(N1,Z1,β1)+B2(N2,Z2,β2)BCN(N,Z,β)+UC(Z1,Z2,β1,β2,R,θ1,θ2)+UN(Z1,Z2,β1,β2,R,θ1,θ2), (22) where Bi (i = 1, 2) and BCN are the negative binding energies of fragment Ai and compound nucleus A = A1+A2, respectively, where the shell and pairing corrections are reasonably included; βi represents the quadrupole deformations of binary fragments; θi denotes collision orientations; and UC and UN are derived from Eqs. 10 and 7, respectively.

By solving a set of master equations, the probability of all possible formed fragments is obtained. The hindrance in the fusion process is named inner fusion barrier, Bfus, which is defined by the difference from the injection position to the B.G. point. In the DNS model, these fragments overcome the inner barrier that is considered to lead to fusion. Therefore, the fusion probability is evaluated by adding all the fragments that could penetrate the inner fusion barrier: PCN(Ec.m.,J,B)=Z=1ZBGN=1NBGP(N1,Z1,E1(J),τint(J)). (23) Here, the interaction time τint(Ec.m., J, B) is obtained from the deflection function method [71]. We calculated the fusion probability as PCN(Ec.m.,J)=f(B)PCN(Ec.m.,J,B)dB. (24) The Coulomb barrier distribution function f(B) is given by Eq. 4. Therefore, the fusion cross-section is expressed as σfus(Ec.m.)=σcap(Ec.m.)PCN(Ec.m.,J). (25)

2.3
Survival probability

The compound nuclei are formed by all the transfers of nucleons from projectile nuclei to target nuclei that have a few excitation energies. The excited compound nuclei are extremely unstable and can be de-excited by evaporating γ-rays, neutrons, protons, α, etc., against fission. The survival probability of the channel x-th neutron, y-th proton, and z-th α is expressed as [75, 57, 70] Wsur(ECN*,x,y,z,J)=P(ECN*,x,y,z,J)×i=1xΓn(Ei*,J)Γtot(Ei*,J)j=1yΓp(Ej*,J)Γtot(Ei*,J)k=1zΓα(Ek*,J)Γtot(Ek*,J) (26) where ECN* and J denote the excitation energy and spin of the excited nucleus, respectively. The total width Γtot is the sum of the partial widths of particle evaporation, γ-rays, and fission. The excitation energy Es* before evaporating the s-th particles is evaluated by Es+1*=Es*BinBjpBkα2Ts, (27) with initial condition Ei*= ECN* and s=i+j+k; Bin, Bjp, and Bkα denote the separation energies of the i-th neutron, j-th proton, and k-th alpha, respectively. The nuclear temperature Ti is defined by Ei*=αTi2Ti with level density a. The decay width of the γ-rays and the particle decay were evaluated with a method similar to that reported in Ref. [75]. We set E*Bvδδn to the term ϱ.

The widths of particles decay are evaluated using the Weisskopf evaporation theory as Γv(E*,J)=(2sv+1)mvπ22ρ(E*,J)×0ϱ1aερ(ϱ+δErotε,J)σinv(ε)dε. (28) Here, sv, mv, and Bv are the spin, mass, and binding energy of the particle, respectively. The pairing correction energy δ was set to be 12/A, 0, and 12/A for even-even, even-odd, and odd-odd nuclei, respectively. The inverse cross-section is expressed as σinv=πRν2T(ν). The penetration probability was set to 1 for neutrons and T(ν)=(1+exp(π(VC(ν)ε)/))1 for charged particles with ω=5 and 8 MeV for proton and α, respectively. The Coulomb barrier of the emitting charge particles and daughter nuclei is expressed as VC=(ZCNi)Zie2ri(ACNi1/3+Ai1/3). (29) In this study, we set proton emitting rp = 1.7 fm and α emitting α = 1.75 fm; for further information, please refer to [76]. The fission width was calculated using the Bohr-Wheeler formula, as in Ref.[63, 64]. We set E*BfErotδδf to the term κ. Γf(E*,J)=12πρf(E*,J)0κ1αfρf(κε+δ,J)dε1+exp[2π(κε+δ+δf)/ω] (30) For heavy fragments, the fission width is usually set as α 2.2 MeV [77], and δf denotes the pairing correction for the fission barrier. The fission barrier is divided into microscopic and macroscopic parts: Bf(E*,J)=BfLD+BfM(E*=0,J)exp(E*/ED), (31) where the macroscopic part is derived from the liquid-drop model: BfLD={0.38(0.75x)ES0,(1/3<x<2/3)0.83(1x)3ES0,(2/3<x<1) (32) with x=EC02ES0. (33) Here, EC0 and ES0 are the Coulomb energy and surface energy of the spherical nuclear, respectively, which could be taken from the Myers-Swiatecki formula: ES0=17.944[11.7826(NZA)2]A2/3MeV (34) and EC0=0.7053Z2A1/3MeV. (35) The microcosmic shell correction energy was taken from [78]. The shell-damping energy was ED=5.48A1/31+1.3A1/3MeV (36) or ED=0.4A4/3/aMeV (37) where a is the energy level density parameter. The fission level density was set as af = 1.1a. The moments of inertia of fission compound nuclei at ground state (gs) and saddle point (sd) configurations are expressed as ζgs(sd)=k×25mr2(1+β2gs(sd)/3). (38) Here, k = 0.4 is the correction factor of the rigid body and β2 is the quadrupole deformation taken from Ref. [78]; β2sd = β2gs + 0.2 was the quadrupole deformation at the saddle point calculated by the relativistic mean field theory. Based on the Fermi gas model, the energy level density could be expressed as [78] ρ(E*,J)=Kcoll×2J+1242σ3a1/4(E*δ)5/4×exp[2a(E*δ)(J+1/2)22σ2] (39) where σ2=6m¯2a(E*δ)/π2 and m¯0.24A2/3; Kcoll is the collective enhancement factor, which contains the rotational and vibration effects. The level density parameter was set as a=A/12, af=1.1a for the fission-level density parameter.

The realization probability of evaporation channels is an important component in the survival probability equation. The realization probability of one particle evaporation is expressed as P(ECN*,J)=exp[(ECN*BsErot2T)22σ2] (40) where σ is the half-height width of the excitation function of the residual nucleon in the F.E. reactions, which was set as 2.5 MeV in our calculations, and Erot is the rotation energy. For the multiple neutron evaporation channels (x >1), the realization probability can be derived using the Jackson formula: P(ECN*,s,J)=I(Δs,2s3)I(Δs+1,2s1) (41) where I and Δ are given by I(z,m)=1m!0zumeudu (42) Δs=ECN*i=1sBivTi (43) where Biv is the separation energy of the evaporation of the i-th particle and s=x+y+z. The spectrum of realization probabilities determines the distribution shape of survival probability in the evaporation channels.

3

Results and Discussion

In the framework of the DNS model involving all the collision orientations, we calculated the excitation functions of 2n-, 3n-, 4n-, and 5n-evaporation channels for the collisions of 48Ca+243Am, 48Ca+242Pu, and 48Ca+238U, marked by solid olive, dash red, dash-dot blue, and orange short-dash lines, respectively, in Fig. 1. In panel (a), the olive-filled up-triangle, red-filled square, and blue-filled circle represent experimental results of 2n-, 3n-, and 4n-evaporation channels for 48Ca+243Am taken from Ref. [9, 83, 84]. According to Ref. [9], the experiments concerning 48Ca+243Am at incident energies Elab = 248, 253 MeV were carried out at FLNR, JINR. At Elab = 248 MeV, three similar decay chains consisting of five consecutive α decays were identified. At Elab = 253 MeV, the decay properties of these synthesized nuclei are consistent with consecutive α decays originating from the parent isotopes of the new element, Mc, i.e., 287Mc and 288Mc, produced in the 3n- and 4n-evaporation channels with cross-sections of approximately 3 pb and 1 pb, respectively. According to Ref. [83], the cross-section for the 3n-evaporation channel reaches its maximum, σ3n = 8.53.7+6.4 pb, at E* = 34.0 - 38.3 MeV and decreases with further increase of the excitation energy of the compound nucleus 291Mc. At excitation energy, E* = 44.8 ± 2.3 MeV, not a single event indicating the formation of 288Mc was detected. The upper cross-section limit can thus be set at level σ3n 1 pb. At excitation energy in the range of E* = 31.1 - 36.4 MeV, the cross-sections for the formation of ERs in the 3n- and 2n-evaporation channels were approximately 3.51.5+2.7 pb and 2.51.5+2.7 pb, respectively. At energies E* 36 MeV, which could be expected for the 2n-evaporation product, 289Mc was not detected. The upper cross-section limit can be set at level σ2n 3 pb. According to Ref. [84], the cross-sections for the formation of ERs in the 3n- and 2n-evaporation channels are approximately 3.21.2+0.8 pb and 0.30.2+0.7 pb at energies E* = 33 MeV, respectively. In Fig. 1 (b), the olive up-triangle, red square, blue circle, and orange down-triangle represent the experimental results of 2n-, 3n-, 4n-, and 5n-evaporation channels for 48Ca+242Pu, respectively, where filled, half-filled, and open symbols represent three experiments for 48Ca+242Pu [8, 79, 80, 86]. According to Ref. [79], a maximum cross-section of 10.42.1+3.5 pb was measured for the 242Pu(48Ca, 3n)287Fl reaction. According to Ref. [8], at excitation energy E* = 50 MeV, the 242Pu(48Ca, 5n)285Fl cross-section was 0.60.5+0.9 pb. The no-observation of a 3n-evaporation product led to an upper limit for the 242Pu(48Ca, 3n)287Fl reaction of 1.1 pb. The 3n and 4n cross-section values measured at E* = 41 MeV were 3.12.6+4.9 pb. In Fig. 1 (c), the red square and blue circle stand for the experimental results of 3n- and 4n-evaporation channels for 48Ca+238U, respectively, where filled, half-filled, and open symbols represent three experiments for 48Ca+238U [79, 81-82]. According to Ref. [80], the maximum cross-section values of the xn-evaporation channels for the reaction 238U(48Ca, xn)286-xCn were measured to be σ3n = 2.51.1+1.8 pb and σ4n = 0.60.5+1.6 pb. At the excitation energies of the compound nucleus E* = 34.5 MeV, two decay events from 283Cn were observed, resulting in a cross-section of 2.01.3+2.7 pb[82]. The cross-section deduced from all four events was 0.720.35+0.58 pb, measured at an excitation energy of 34.6 MeV of the compound nucleus 286Cn[81]. From the above three panels, we can conclude that our calculations are in good agreement with the available experimental excitation functions of the reactions 48Ca+243Am, 48Ca+242Pu, and 48Ca+238U.

Fig. 1
(Color online) The calculated excitation functions of 2n-, 3n-, 4n-, and 5n-evaporation channels for the reactions of 48Ca+243Am, 48Ca+242Pu, and 48Ca+238U are marked by solid olive, dash red, dash-dot blue, and orange short-dash lines, respectively. The experimental measurement results of excitation functions for 2n-, 3n-, 4n-, and 5n-evaporation channels are represented by an up-triangle, square, circle, and down-triangle. Vertical error bars correspond to total uncertainties. Symbols with arrows show upper cross-section limits. Data marked by open, half-closed, and filled symbols are taken from [8, 9, 79-85], respectively.
pic

To investigate the dependence of the production cross-section of superheavy nuclei in F.E. reactions on collision orientation, we exported four configurations of the collision orientations from our calculations for the reaction of 48Ca+243Am as (0°, 0°), (30°, 30°), (60°, 60°) and (90°, 90°), marked by the solid black, red dash, olive dash-dot, and blue short-dash lines, respectively, in Fig. 2. The projectile nuclei 48Ca and target nuclei 243Am have theoretical quadrupole deformation values βP = 0. and βT = 0.224, respectively. In Fig. 2, panel (a) shows the distributions of interaction potential energy with respect to the distance between the surfaces of projectile nuclei and target nuclei. The interaction potential VCN consists of Coulomb potential VC and nucleus-nucleus potential VN, which were calculated by Wong’s formula [69] and the double-folding method [87], respectively. The interaction potential energies were increased with the large collision orientations because of the large effective interaction face. Panel (b) displays the distributions of radial kinetic energy with respect to the interaction time. The kinetic energy decreased exponentially with increasing reaction time at the prescribed impact parameter, i.e., L=20 . These evolutions of kinetic energy reached equilibrium at approximately 2 × 10-21s. These equilibrium kinetic energies were 225 MeV, 228 MeV, 235MeV, and 239 MeV, corresponding to collision orientations (0°, 0°), (30°, 30°), (60°, 60°), and (90°, 90°), respectively. The kinetic energy dissipated into the internal excitation of the composite system, which correspondingly increased exponentially with the reaction time for the same relaxation time, as illustrated in panel (c). According to Fig. 2, we can conclude that the interaction potential and evolution of kinetic energy and internal excitation energy were highly dependent on the orientations. These were the basic reasons causing the dependence of the final synthesis cross-sections of superheavy nuclei on collision orientations.

Fig. 2
(Color online) Panel (a) shows the interaction potential for the collisions of 48Ca+243Am as a function of distance with different collision angles. The collision orientations (0°, 0°), (30°, 30°), (60°, 60°), and (90°, 90°) correspond to solid black, red dash, olive dash-dot, and short dash lines, respectively. Panel (b) represents the radial kinetic energy decreases along with the reaction time at different collision orientations under an angular momentum L=20 . Panel (c) shows that the internal excitation energy of the composite system varies with the sticking time for the given angular momentum L=20 .
pic

The PES and driving potential (DP) of the reaction 48Ca+243Am were calculated by Eq. 22 for the collision orientations of sphere-sphere, (0°, 0°), (30°, 30°), (60°, 60°) and (90°, 90°), as illustrated in Fig. 3. The PES and DP are listed in the upper and lower panels, respectively. Panels (a) and (f) show the PES and DP of the no-deformation of projectile-target nuclei. The minimum trajectories and injection points are attached to the PESs, which are represented by solid black lines and filled black stars. The structure effect is clearly shown in the PESs and DPs by the comparison of no-deformation collision with quadrupole deformation collision. The inner fusion barrier was set as the difference between the injection points and Businaro-Gallone (B.G.) points, which were 8 MeV, 11.5 MeV, 10.5 MeV, 7.1 MeV, and 6 MeV corresponding to collision orientations of no-deformation, (0°, 0°), (30°, 30°), (60°, 60°), and (90°, 90°), respectively. It was found that the inner fusion barrier was highly dependent on the collision orientations, which could reveal the fusion probability directly. The inner fusion barriers were decreased with the increased collision orientation, reaching its minimum at the waist-waist collision. Sketches of collision orientations are shown at the top of Fig. 3. The potential energy of the symmetry field in the PES increased with increasing collision orientations because the corresponding Coulomb force increased as well.

Fig. 3
(Color online) Panels (a), (b), (c), (d), and (e) represent the PES of 48Ca+243Am at collision orientations of sphere-sphere, (0°, 0°), (30°, 30°), (60°, 60°), and (90°, 90°), respectively. Panels (f), (g), (h), (i), and (j) correspond to their collision orientation-based valley trajectories in PES along with the mass asymmetry η with respect to η=(ATAP)/(AT+AP). Their inner barrier value is given by B_BG. The arrow lines represent injection points.
pic

In the collision process, when overcoming the Coulomb barrier, the kinetic energies of the colliding partners rapidly dissipate into the composite system. The probability of projectile and target diffusing along the PES was calculated by solving a set of master equations. The TKE of binary fragments was related to the incident energy, ground-state binding energy, and internal excitation energy as TKE=Ec.m.VCNQggE*. Fig. 4 presents the TKE-mass distributions for collision orientations of no-deformation, (0°, 0°), (30°, 30°), (60°, 60°), and (90°, 90°) at incident energy Ec.m. = 1.1 × VB, as shown in panels (a), (b), (c), (d), and (e), respectively. The TKE term could be rewritten as TKE=Ec.m.UdrE*. The TKE-mass distribution shape is highly dependent on the driving potential. The TKE-mass distribution for no-deformation collision in panel (a) was smoother than others in panels (b), (c), (d), and (e), thereby showing the structure effect in the TKE-mass distribution. The fragments in the black square passing the B.G. points are supposed to lead to fusion. The fusion probability was calculated by summing all the formation probabilities passing B.G. points. Fig. 4 reveals that it is difficult to evaluate the dependence of fusion probability on the collision orientations; the reason is that only one incident energy, Ec.m. = 1.1 × VB, is shown.

Fig. 4
(Color online) Panels (a), (b), (c), (d), and (e) display the calculations of TKE-mass distribution of the primary fragments in the collisions of 48Ca+243Am at E_c.m. = 1.1× VB for their collision orientations of sphere-sphere, (0°, 0°), (30°, 30°), (60°, 60°), and (90°, 90°), respectively.
pic

Figure 5 shows the TKE-mass distributions at excitation energies ECN* = 10 MeV, 40 MeV, 70 MeV, and 100 MeV for the tip-tip collisions of 48Ca+243Am, as illustrated in panels (a), (b), (c), and (d), respectively. Fig. 5 shows that the TKE-mass distribution was broader for increasing incident energy. It is evident that the fusion probability increased with larger excitation energy. However, the compound nuclei with large excitation energy could easily lead to fission. The maximum evaporation residue cross-section of the high-excitation compound nuclei was the balance between fusion probability and survival probability.

Fig. 5
(Color online) Panels (a), (b), (c), and (d) show the calculations of TKE-mass distribution of the primary fragments in the head-on collisions of 48Ca+243Am at incident energies corresponding to excitation energies of compound nuclei, that is, 10 MeV, 40 MeV, 70 MeV, and 100 MeV, respectively.
pic

To approximate the real collision process as much as possible, we propose Gaussian-like barrier distributions to consider all the collision orientations. Eq. (4) can be employed for this purpose. The olive solid, red dash, blue dot-dash, and orange short-dash lines represent the calculated excitation function of the 2n-, 3n-, 4n-, and 5n-evaporation channels. The olive-filled up-triangle, red-filled square, and blue-filled circle represent the experimental excitation function of the 2n-, 3n-, and 4n-evaporation channels, respectively. For the reactions 48Ca+243Am at excitation energies within the interval E* = 20–100 MeV, the excitation functions of the 2n-, 3n-, 4n-, and 5n-evaporation channels were calculated by the DNS model involving the barrier distribution, as shown in panel (a); these functions are in good agreement with the experimental data [83, 9]. The calculated excitation functions of 48Ca+243Am for the collision orientations (0°, 0°), (30°, 30°), (60°, 60°), (90°, 90°), and no-deformation are shown in panels (b), (c), (d), (e), (f), respectively. It was found that the (0°, 0°) collisions underestimate the experimental results. Collisions (30°, 30°) agree with the experimental results relatively well. Collisions (30°, 30°), (60°, 60°), and (90°, 90°) overestimate the experimental data. Fig. 6 shows thatthe DNS model involving barrier distributions could reproduce the experimental results relatively well.

Fig. 6
(Color online) For the collisions of 48Ca+243Am, the panels show the calculations of excitation functions in 2n-, 3n-, 4n-, and 5n-evaporation channels corresponding to solid olive, red dash, blue dash-dot, and orange short-dash lines, respectively. Panels (b), (c), (d), (e), and (f) display the excitation functions at orientations (0°, 0°), (30°, 30°), (60°, 60°), (90°, 90°), and sphere to sphere, respectively. Panel (a) shows the total excitation function when considering all the collision orientations using the method of Gaussian distribution. Experimental data are marked by filled up-triangle, square, circle, and down-triangle symbols, as in [83, 9].
pic

Based on the DNS model involving barrier distribution, to investigate the dependence of evaporation residue cross-section on the isospin of the projectile, we systematically calculated the reactions of 42Ca+243Am, 44Ca+243Am, 46Ca+243Am, 48Ca+243Am, 44Ti+237Np, 46Ti+237Np, 48Ti+237Np, and 50Ti+237Np at excitation energies within the interval E* = 1 - 80 MeV. Fig. 7 shows that the excitation functions of the evaporation residue cross-section are highly dependent on the isospin of the projectile. Regarding the isotopes of Ca-induced reactions, the cross-sections of 2n- and 3n-evaporation channels decreased with the projectile of Ca isotopes with large N/Z, which might be caused by fusion probability. The ratio of σ3n/σ2n increased with increasing N/Z, which implies that more-neutron-rich compound nuclei are prone to evaporating more neutrons. The existing moscovium isotopes are 287-290Mc. The predictions of maximum cross-sections of the new 281-286Mc were 4 pb, 45 pb, 150 pb, 50 pb, 101 pb, and 30 pb, respectively, in calcium-isotope-induced F.E. reactions. The maximum synthesis cross-section of new moscovium isotopes was 283Mc as 0.15 nb in the reactions 42Ca+243Am. Concerning Ti-isotope induced reactions, the 2n-evaporation channel was dominant in the evaporation residue cross-sections. The maximum synthesis cross-section of Mc was 281Mc as 0.2 nb in the reactions 46Ti+237Np. The new moscovium isotopes of 278-286Mc were evaluated as 0.5 pb, 9 pb, 12 pb, 10.5 pb, 150 pb, 11 pb, 100 pb, 10 pb, and 31 pb, respectively, in titanium-isotope-induced F.E. reactions.

Fig. 7
(Color online) The calculations of excitation functions for the collisions of 42Ca+243Am, 44Ca+243Am, 46Ca+243Am, 48Ca+243Am, 44Ti+237Np, 46Ti+237Np, 48Ti+237Np, and 50Ti+237Np are shown in panels (a), (b), (c), (d), (e), (f), (g), and (h), respectively. The 2n-, 3n-, 4n-, and 5n-evaporation channels correspond to black solid, red dash, blue dash-dot, and olive short-dash lines, respectively
pic

To investigate the influence of the entrance effect on the synthesis cross-section of superheavy moscovium in the F.E. reactions, we systematically calculated the collisions of 35Cl + 248Cf (η = 0.75), 40Ar + 247Bk (η = 0.72), 39K + 247Cm (η = 0.73), 40Ca + 243Am (η = 0.72), 48Ca + 243Am (η = 0.67), 45Sc + 244Pu (η = 0.69), 48Ti + 237Np (η = 0.66), and 51V + 238U (η = 0.65) based on the DNS model, as illustrated in panels (a), (b), (c), (d), (e), (f), (g), and (h), respectively. Fig. 8 shows that the reaction systems with large η are prone to producing large production cross-sections because the large mass asymmetry reactions are in turn prone to fusion. In these calculations, the new moscovium 278-286Mc was predicted with production cross-section values of 1 pb, 10 pb, 130 pb, 50 pb, 15 pb, 100 pb, 30 pb, 200 pb, 40 pb, respectively. The 2n- or 3n-evaporation residue channels were dominant in the evaporation survival process. The ratio σ3n/σ2n illustrates the role of the odd-even effect on the production cross-section of superheavy nuclei. The maximum production cross-section of moscovium isotopes was predicted as 200 pb in the reaction 247Cm(39K, 3n)283Mc.

Fig. 8
(Color online) The calculations of excitation functions in the collisions of 35Cl+248Cf, 40Ar+247Bk, 39K+247Cm, 40Ca+243Am, 48Ca+243Am, 45Sc+244Pu, 48Ti+237Np, and 51V+238U are shown in panels (a), (b), (c), (d), (e), (f), (g), and (h), respectively. The 2n-, 3n-, 4n-, and 5n-evaporation channels correspond to black solid, red dash, blue dash-dot, and olive short-dash lines, respectively.
pic
4

Conclusion

As a summary, to simulate the real collision process, we propose a Gaussian-like barrier distribution function used to include all collision orientations. To investigate the dependence of the production cross-section of superheavy isotopes on the collision orientations, we systematically calculated the reactions of 48Ca+243Am at excitation energies within the interval 0-100 MeV for the collision orientations of no-deformation, i.e., (0°, 0°), (30°, 30°), (60°, 60°), and (90°, 90°). In the DNS model, for a given collision orientation, some physical quantities such as interaction potential, radial kinetic energy, internal excitation energy, TKE-mass, PES, DP, and inner fusion barrier were exported to show the influence of collision orientations; the conclusion is that these quantities are highly dependent on the collision orientations. We compared the calculated excitation functions of 48Ca+243Am at some fixed collision orientations with available experimental results. We found that large collision orientations showed an overestimated value compared to experimental data. The collision orientation nearby (30°, 30°) fit the experimental data very well. The barrier-distribution-based excitation function was in good agreement with the experimental data. To test the barrier distribution function, we calculated the reactions of 48Ca+243Pu and 48Ca+238U, which reproduced the experimental excitation functions well. Based on the DNS model involving the barrier distribution function, we systematically calculated the reactions of projectiles 42-48Ca bombarding on target 243Am and projectiles 42-48Ca on target 237Np. The influence of the isospin of a projectile on the production cross-section was studied. For Ca-induced F.E. reactions, σ2n and σ3n were dominant in the evaporation residue cross–sections, which decreased with increasing N/Z in projectiles. The ratio σ3n/σ2n increased with increasing N/Z in projectiles, which might be caused by neutron-rich compound nuclei prone to losing neutrons. For Ti-induced F.E. reactions, the maximum cross-section was 150 pb for 283Mc in the reaction 237Np(46Ti, 2n)283Mc. The reactions of 35Cl+248Cf, 40Ar+247Bk, 39K+247Cm, 40Ca+243Am, 48Ca+243Am, 45Sc+244Pu, 48Ti+237Np, and 51V+238U were calculated to investigate the entrance channel effect on production cross-sections of superheavy nuclei. Large mass asymmetry systems lead to large production cross-section. We also found that the odd-even effect might play a role in the evaporation residue cross-section. We predicted the new moscovium isotopes 278-286Mc with maximum cross-sections of 0.5 pb, 9 pb, 12 pb, 10.5 pb, 150 pb, 11 pb, 100 pb, 10 pb, and 31 pb in the collisions of 35,37Cl + 248Cf, 38,40Ar + 247Bk, 39,41K + 247Cm, 40,42,44,46Ca + 243Am, 45Sc + 244Pu, and 46,48,50Ti + 237Np, 51V + 238U at some typical excitation energies.

References
[1] Y.T. Oganessian, V.K. Utyonkov,

Super-heavy element research

. Reports on Progress in Physics 78, 036301 (2015). doi: 10.1088/0034-4885/78/3/036301
Baidu ScholarGoogle Scholar
[2] M. Thoennessen,

Discovery of isotopes of elements with Z≥100

. Atomic Data and Nuclear Data Tables 99, 312-344 (2013). doi: 10.1016/j.adt.2012.03.003
Baidu ScholarGoogle Scholar
[3] G. Heinlein, K. Bächmann, J. Rudolph, in Superheavy Elements, Model experiments with short-lived nuclides for the investigation of the quantitative chemical behaviour of new elements. (Pergamon, 1978), pp. 407414 doi: 10.1016/B978-0-08-022946-1.50051-7
[4] A. Ghiorso, M. Nurmia, K. Eskola, et al.,

261rf; new isotope of element 104

. Physics Letters B 32, 95-98 (1970). doi: 10.1016/0370-2693(70)90595-2
Baidu ScholarGoogle Scholar
[5] G.N. Flerov, Y.T. Oganesyan, Y.V. Lobanov, et al.,

Synthesis of element 105

. Sov. At. Energy 29, 967 (1970).
Baidu ScholarGoogle Scholar
[6] A. Ghiorso, M. Nurmia, K. Eskola, et al.,

New element hahnium, atomic number 105

. Phys. Rev. Lett. 24, 1498-1503 (1970). doi: 10.1103/PhysRevLett.24.1498
Baidu ScholarGoogle Scholar
[7] A. Ghiorso, J.M. Nitschke, J.R. Alonso, et al.,

Element 106

. Phys. Rev. Lett. 33, 1490-1493 (1974). doi: 10.1103/PhysRevLett.33.1490
Baidu ScholarGoogle Scholar
[8] P.A. Ellison, K.E. Gregorich, J.S. Berryman, et al.,

New superheavy element isotopes: 242Pu(48 Ca, 5n)285114

. Phys. Rev. Lett. 105, 182701 (2010). doi: 10.1103/PhysRevLett.105.182701
Baidu ScholarGoogle Scholar
[9] Y.T. Oganessian, V.K. Utyonkoy, Y.V. Lobanov, et al.,

Experiments on the synthesis of element 115 in the reaction 243Am(48 Ca, xn)291-x115

. Phys. Rev. C 69, 021601 (2004). doi: 10.1103/PhysRevC.69.021601
Baidu ScholarGoogle Scholar
[10] Y.T. Oganessian, V.K. Utyonkov, Y.V. Lobanov, et al.,

Measurements of cross sections for the fusion-evaporation reactions 244 Pu(48Ca, xn)292-x 114 and 245Cm(48Ca, xn)293-x116

. Phys. Rev. C 69, 054607 (2004). doi: 10.1103/PhysRevC.69.054607
Baidu ScholarGoogle Scholar
[11] Y.T. Oganessian, F.S. Abdullin, P.D. Bailey, et al.,

Synthesis of a new element with atomic number Z=117

. Phys. Rev. Lett. 104, 142502 (2010). doi: 10.1103/PhysRevLett.104.142502
Baidu ScholarGoogle Scholar
[12] Y.T. Oganessian, V.K. Utyonkov, Y.V. Lobanov, et al.,

Synthesis of the isotopes of elements 118 and 116 in the 249Cf and 245Cm+48Ca fusion reactions

. Phys. Rev. C 74, 044602 (2006). doi: 10.1103/PhysRevC.74.044602
Baidu ScholarGoogle Scholar
[13] Y.T. Oganesyan, Y.P. Tretyakov, A.S. Ilinov, et al.,

Synthesis of neutron-deficient isotopes of fermium, kurchatovium, and element 106

. JETP Lett. (USSR) 20, 265 (1975).
Baidu ScholarGoogle Scholar
[14] G. Münzenberg, S. Hofmann, F.P. Heßberger, et al.,

Identification of element 107 by α correlation chains

. Zeitschrift fur Physik A Hadrons and Nuclei 300, 107-108 (1981). doi: 10.1007/BF01412623
Baidu ScholarGoogle Scholar
[15] G. Münzenberg, P. Armbruster, H. Folger, et al.,

The identification of element 108

. Zeitschrift für Physik A Atoms and Nuclei 317, 235-236 (1984). doi: 10.1007/BF01421260
Baidu ScholarGoogle Scholar
[16] G. Münzenberg, P. Armbruster, F.P. Hessberger, et al.,

Observation of one correlated α-decay in the reaction58fe on209bi →267109

. Zeitschrift für Physik A Atoms and Nuclei 309, 89-90 (1982).
Baidu ScholarGoogle Scholar
[17] S. Hofmann, V. Ninov, F. Heßberger, et al.,

Production and decay of269110

. Zeitschrift für Physik A Hadrons and Nuclei 350, 277-280 (1995). doi: 10.1007/BF01291181
Baidu ScholarGoogle Scholar
[18] S. Hofmann, V. Ninov, F. Heßberger, et al.,

The new element 111

. Zeitschrift für Physik A Hadrons and Nuclei 350, 281-282 (1995). doi: 10.1007/BF01291182
Baidu ScholarGoogle Scholar
[19] JOUR, S. Hofmann, V. Ninov, et al.,

The new element 112

. Zeitschrift für Physik A Hadrons and Nuclei 354,. doi: 10.1007/BF02769517
Baidu ScholarGoogle Scholar
[20] K. Morita, K. Morimoto, D. Kaji, et al.,

Experiment on the synthesis of element 113 in the reaction 209bi(70zn,n)278113

. Journal of the Physical Society of Japan 73, 2593-2596 (2004). doi: 10.1143/JPSJ.73.2593
Baidu ScholarGoogle Scholar
[21] Z. Gan, Z. Qin, H. Fan, et al.,

A new alpha-particle - emitting isotope 259Db

. European Physical Journal A 10, 21-25 (2001). doi: 10.1007/s100500170140
Baidu ScholarGoogle Scholar
[22] Q. Z., W.X. L., D.H. J., et al.,

Alpha decay properties of 266Bh

. Nuclear Physics Review 23, 404 (2006). doi: 10.11804/NuclPhysRev.23.04.404
Baidu ScholarGoogle Scholar
[23] Z.Y. ZHANG, Z.G. GAN, L. MA, et al.,

Observation of the superheavy nuclide 271Ds

. Chin. Phys. Lett. 58,. doi: 10.1088/0256-307X/29/1/012502
Baidu ScholarGoogle Scholar
[24] F.P. Heßberger, S. Hofmann, V. Ninov, et al.,

Spontaneous fission and alpha-decay properties of neutron deficient isotopes 257-253104 and 258106

. Zeitschrift für Physik A Hadrons and Nuclei 359, 415-425 (1997). doi: 10.1007/s002180050422
Baidu ScholarGoogle Scholar
[25] U. Tippawan, S. Pomp, A. Ataç, et al.,

Light-ion production in the interaction of 96 MeV neutrons with silicon

. Phys. Rev. C 69, 064609 (2004). doi: 10.1103/PhysRevC.69.064609
Baidu ScholarGoogle Scholar
[26] JOUR, F. Heßberger, S. Hofmann, et al.,

Decay properties of neutron-deficient isotopes 256, 257Db, 255Rf, 252, 253Lr

. The European Physical Journal A - Hadrons and Nuclei. http://dx.doi.org/10.1007/s100500170039 doi:10.1007/s100500170039
Baidu ScholarGoogle Scholar
[27] S.L. Nelson, K.E. Gregorich, I. Dragojević, et al.,

Lightest isotope of bh produced via the 209Bi(52Cr, n)260 reaction

. Phys. Rev. Lett. 100, 022501 (2008). doi: 10.1103/PhysRevLett.100.022501
Baidu ScholarGoogle Scholar
[28] I. Dragojević, K.E. Gregorich, C.E. Düllmann, et al.,

New isotope 263Hs

. Phys. Rev. C 79, 011602 (2009). doi: 10.1103/PhysRevC.79.011602
Baidu ScholarGoogle Scholar
[29] C.E. Düllmann, M. Schädel, A. Yakushev, et al.,

Production and decay of element 114: High cross sections and the new nucleus 277Hs

. Phys. Rev. Lett. 104, 252701 (2010). doi: 10.1103/PhysRevLett.104.252701
Baidu ScholarGoogle Scholar
[30] G. Münzenberg, P. Armbruster, F. Heßberger, et al.,

Observation of one correlated α-decay in the reaction 58Fe on 209Bi→ 267109

. Zeitschrift für Physik A Hadrons and Nuclei 309, 89-90 (1982).
Baidu ScholarGoogle Scholar
[31] A. Ghiorso, D. Lee, L.P. Somerville, et al.,

Evidence for the possible synthesis of element 110 produced by the 59Co+209Bi reaction

. Phys. Rev. C 51, R2293-R2297 (1995). doi: 10.1103/PhysRevC.51.R2293
Baidu ScholarGoogle Scholar
[32] JOUR, S. Hofmann, V. Ninov, et al.,

The new element 111

. Zeitschrift für Physik A Hadrons and Nuclei 350,. doi: 10.1007/BF01291182
Baidu ScholarGoogle Scholar
[33] K. Morita, K. Morimoto, D. Kaji, et al.,

Experiment on the synthesis of element 113 in the reaction 209Bi(70Zn,n)278113

. Journal of the Physical Society of Japan 73, 2593-2596 (2004). doi: 10.1143/jpsj.73.2593
Baidu ScholarGoogle Scholar
[34] 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
[35] Z. Gan, W. Huang, Z. Zhang, et al.,

Results and perspectives for study of heavy and super-heavy nuclei and elements at IMP/CAS

. The European Physical Journal A 58,. doi: 10.1140/epja/s10050-022-00811-w
Baidu ScholarGoogle Scholar
[36] S. Heinz, H. Devaraja,

Nucleosynthesis in multinucleon transfer reactions

. The European Physical Journal A 58,. doi: 10.1140/epja/s10050-022-00771-1
Baidu ScholarGoogle Scholar
[37] M. Block, F. Giacoppo, F.P. Heßberger, et al.,

Recent progress in experiments on the heaviest nuclides at SHIP

. Nuovo Cimento Rivista Serie 45, 279-323 (2022). doi: 10.1007/s40766-022-00030-5
Baidu ScholarGoogle Scholar
[38] B. Sergey, S. Dmitriev, N. Kazarinov, et al., She factory: Cyclotron facility for super heavy elements research. 2020. doi: 10.18429/JACoW-Cyclotrons2019-THC01
[39] G. Adamyan, N. Antonenko, A. Diaz-Torres, et al.,

How to extend the chart of nuclides?

European Physical Journal A 56, 47-1 (2020). doi: 10.1140/epja/s10050-020-00046-7
Baidu ScholarGoogle Scholar
[40] D. Kaji, K. Morimoto, A. Yoneda, et al.,

Target for the heaviest element production at riken

. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 590, 198-203 (2008). Toward the Realization of Target and Stripper Foil Technologies for High-Power Proton and Radioactive Ion Accelerators. doi: 10.1016/j.nima.2008.02.090
Baidu ScholarGoogle Scholar
[41] H. Haba,

A new period in superheavy-element hunting

. Nature Chemistry 11, 10-13 (2019). doi: 10.1038/s41557-018-0191-8
Baidu ScholarGoogle Scholar
[42] T. Tanaka, K. Morita, K. Morimoto, et al.,

Study of quasielastic barrier distributions as a step towards the synthesis of superheavy elements with hot fusion reactions

. Phys. Rev. Lett. 124, 052502 (2020). doi: 10.1103/PhysRevLett.124.052502
Baidu ScholarGoogle Scholar
[43] J.L. Pore, J.M. Gates, R. Orford, et al.,

Identification of the new isotope 244Md

. Phys. Rev. Lett. 124, 252502 (2020). doi: 10.1103/PhysRevLett.124.252502
Baidu ScholarGoogle Scholar
[44] L. Guo, C. Simenel, L. Shi, et al.,

The role of tensor force in heavy-ion fusion dynamics

. Physics Letters B 782, 401-405 (2018). doi: 10.1016/j.physletb.2018.05.066
Baidu ScholarGoogle Scholar
[45] K. Sekizawa,

Tdhf theory and its extensions for the multinucleon transfer reaction: A mini review

. Frontiers in Physics 7,. doi: 10.3389/fphy.2019.00020
Baidu ScholarGoogle Scholar
[46] J. Maruhn, P.G. Reinhard, P. Stevenson, et al.,

The tdhf code sky3d

. Computer Physics Communications 185, 2195-2216 (2014). doi: 10.1016/j.cpc.2014.04.008
Baidu ScholarGoogle Scholar
[47] N. Wang, Z. Li, X. Wu,

Improved quantum molecular dynamics model and its applications to fusion reaction near barrier

. Phys. Rev. C 65, 064608 (2002). doi: 10.1103/PhysRevC.65.064608
Baidu ScholarGoogle Scholar
[48] X. Jiang, S. Yan, J.A. Maruhn,

Dynamics of energy dissipation in heavy-ion fusion reactions

. Phys. Rev. C 88, 044611 (2013). doi: 10.1103/PhysRevC.88.044611
Baidu ScholarGoogle Scholar
[49] K. Zhao, Z. Li, X. 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
[50] A.V. Karpov, V.V. Saiko,

Modeling near-barrier collisions of heavy ions based on a langevin-type approach

. Phys. Rev. C 96, 024618 (2017). doi: 10.1103/PhysRevC.96.024618
Baidu ScholarGoogle Scholar
[51] V.I. Zagrebaev, A.V. Karpov, W. Greiner,

Possibilities for synthesis of new isotopes of superheavy elements in fusion reactions

. Phys. Rev. C 85, 014608 (2012). doi: 10.1103/PhysRevC.85.014608
Baidu ScholarGoogle Scholar
[52] Z.Q. Feng, G.M. Jin, F. Fu, et al.,

Production cross sections of superheavy nuclei based on dinuclear system model

. Nuclear Physics A 771, 50-67 (2006). doi: 10.1016/j.nuclphysa.2006.03.002
Baidu ScholarGoogle Scholar
[53] X.J. Bao, Y. Gao, J.Q. Li, et al.,

Influence of the nuclear dynamical deformation on production cross sections of superheavy nuclei

. Phys. Rev. C 91, 011603 (2015). doi: 10.1103/PhysRevC.91.011603
Baidu ScholarGoogle Scholar
[54] L. Zhu, W.J. Xie, F.S. Zhang,

Production cross sections of superheavy elements Z=119 and 120 in hot fusion reactions

. Phys. Rev. C 89, 024615 (2014). doi: 10.1103/PhysRevC.89.024615
Baidu ScholarGoogle Scholar
[55] G. Adamyan, N. Antonenko, A. Diaz-Torres, et al.,

How to extend the chart of nuclides?

European Physical Journal A 56, 47-1 (2020). doi: 10.1140/epja/s10050-020-00046-7
Baidu ScholarGoogle Scholar
[56] Z.Q. Feng, g.m. no, H. Ming-Hui, et al.,

Possible way to synthesize superheavy element Z=117

. Chinese Physics Letters 24,. doi: 10.1088/0256-307X/24/9/024
Baidu ScholarGoogle Scholar
[57] Z.Q. Feng, G.M. Jin, J.Q. Li,

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

. Nuclear Physics A 836, 82-90 (2010). doi: 10.1016/j.nuclphysa.2010.01.244
Baidu ScholarGoogle Scholar
[58] K.X. Cheng, J. Pu, Y.T. Wang, et al.,

Non-frozen process of heavy-ion fusion reactions at deep sub-barrier energies

. Nuclear Science and Techniques 33, 132 (2022). doi: 10.1007/s41365-022-01114-x
Baidu ScholarGoogle Scholar
[59] 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
[60] M.T. Jin, S.Y. Xu, G.M. Yang, et al.,

Yield of long-lived fission product transmutation using proton-, deuteron-, and alpha particle-induced spallation

. Nuclear Science and Techniques 32, 96 (2021). doi: 10.1007/s41365-021-00933-8
Baidu ScholarGoogle Scholar
[61] Y.Q. Xin, N.N. Ma, J.G. Deng, et al.,

Properties of Z=114 super-heavy nuclei

. Nuclear Science and Techniques 32, 55 (2021). doi: 10.1007/s41365-021-00899-7
Baidu ScholarGoogle Scholar
[62] 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
[63] Z.Q. Feng, G.M. Jin, J.Q. Li,

Production of new superheavy Z=108-114 nuclei with 238U, 244Pu, and 248,250Cm targets

. Phys. Rev. C 80, 057601 (2009). doi: 10.1103/PhysRevC.80.057601
Baidu ScholarGoogle Scholar
[64] 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
[65] G.G. Adamian, N.V. Antonenko, R.V. Jolos, et al.,

Effective nucleus-nucleus potential for calculation of potential energy of a dinuclear system

. International Journal of Modern Physics E 05, 191216 (1996).
Baidu ScholarGoogle Scholar
[66] D.L. Hill, J.A. Wheeler,

Nuclear constitution and the interpretation of fission phenomena

. Phys. Rev. 89, 1102-1145 (1953). doi: 10.1103/PhysRev.89.1102
Baidu ScholarGoogle Scholar
[67] V.I. Zagrebaev, Y. Aritomo, M.G. Itkis, et al.,

Synthesis of superheavy nuclei: How accurately can we describe it and calculate the cross sections? Phys

. Rev. C 65, 014607 (2001). doi: 10.1103/PhysRevC.65.014607
Baidu ScholarGoogle Scholar
[68] W.D. Myers, W.J. Swiatecki,

Nuclear masses and deformations

. Nuclear Physics 81, 1-60 (1966). doi: 10.1016/0029-5582(66)90639-0
Baidu ScholarGoogle Scholar
[69] 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
[70] Z.Q. Feng, G.M. Jin, J.Q. Li, et al.,

Production of heavy and superheavy nuclei in massive fusion reactions

. Nuclear Physics A 816, 33-51 (2009). doi: 10.1016/j.nuclphysa.2008.11.003
Baidu ScholarGoogle Scholar
[71] 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
[72] J. Li, X. Tang, G. Wolschin,

Anticorrelated angular-momentum distributions in heavy-ion collisions

. Physics Letters B 105, 107-110 (1981). doi: 10.1016/0370-2693(81)91000-5
Baidu ScholarGoogle Scholar
[73] 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
[74] 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
[75] P.H. Chen, Z.Q. Feng, J.Q. Li, et al.,

A statistical approach to describe highly excited heavy and superheavy nuclei

. Chinese Physics C 40, 091002 (2016). doi: 10.1088/1674-1137/40/9/091002
Baidu ScholarGoogle Scholar
[76] A.S. Zubov, G.G. Adamian, N.V. Antonenko, et al.,

Competition between evaporation channels in neutron-deficient nuclei

. Phys. Rev. C 68, 014616 (2003). doi: 10.1103/PhysRevC.68.014616
Baidu ScholarGoogle Scholar
[77] A. Zubov, G.G. Adamyan, N. Antonenko, et al.,

Survival probabilities of superheavy nuclei based on recent predictions of nuclear properties

. Eur. Phys. J. A 23, 249-256 (2005). doi: 10.1140/epja/i2004-10089-5
Baidu ScholarGoogle Scholar
[78] P. Moller, J. Nix, W. Myers, et al.,

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
[79] Y.T. Oganessian, V.K. Utyonkov, D. Ibadullayev, et al.,

Investigation of 48Ca-induced reactions with 242Pu and 238U targets at the JINR superheavy element factory

. Phys. Rev. C 106, 024612 (2022). doi: 10.1103/PhysRevC.106.024612
Baidu ScholarGoogle Scholar
[80] Y.T. Oganessian, V.K. Utyonkov, Y.V. Lobanov, et al.,

Measurements of cross sections and decay properties of the isotopes of elements 112, 114, and 116 produced in the fusion reactions 233,238U, 242Pu, and 248Cm+48Ca

. Phys. Rev. C 70, 064609 (2004). doi: 10.1103/PhysRevC.70.064609
Baidu ScholarGoogle Scholar
[81] S. Hofmann, D. Ackermann, S. Antalic, et al.,

The reaction 48Ca + 238U → 286112* studied at the GSI-SHIP

. European Physical Journal A 32, 251-260 (2007). doi: 10.1140/epja/i2007-10373-x
Baidu ScholarGoogle Scholar
[82] D. Kaji, K. Morimoto, H. Haba, et al.,

Decay measurement of 283Cn produced in the 238U(48Ca,3n) reaction using GARIS-II

. Journal of the Physical Society of Japan 86, 085001 (2017). doi: 10.7566/JPSJ.86.085001
Baidu ScholarGoogle Scholar
[83] Y.T. Oganessian, F.S. Abdullin, S.N. Dmitriev, et al.,

Investigation of the 243Am+48Ca reaction products previously observed in the experiments on elements 113, 115, and 117

. Phys. Rev. C 87, 014302 (2013). doi: 10.1103/PhysRevC.87.014302
Baidu ScholarGoogle Scholar
[84] Y.T. Oganessian, F.S. Abdullin, S.N. Dmitriev, et al.,

New insights into the 243Am+48Ca reaction products previously observed in the experiments on elements 113, 115, and 117

. Phys. Rev. Lett. 108, 022502 (2012). doi: 10.1103/PhysRevLett.108.022502
Baidu ScholarGoogle Scholar
[85] Y.T. Oganessian, V.K. Utyonkov, S.N. Dmitriev, et al.,

Synthesis of elements 115 and 113 in the reaction 243Am+48Ca

. Phys. Rev. C 72, 034611 (2005). doi: 10.1103/PhysRevC.72.034611
Baidu ScholarGoogle Scholar
[86] L. Stavsetra, K.E. Gregorich, J. Dvorak, et al.,

Independent verification of element 114 production in the 48Ca+242Pu reaction

. Phys. Rev. Lett. 103, 132502 (2009). doi: 10.1103/PhysRevLett.103.132502
Baidu ScholarGoogle Scholar
[87] I.I. Gontchar, D.J. Hinde, M. Dasgupta, et al.,

Double folding nucleus-nucleus potential applied to heavy-ion fusion reactions. Zhangjiang Laboratory, Shanghai 201210

, ChinaPhys. Rev. C 69, 024610 (2004). doi: 10.1103/PhysRevC.69.024610
Baidu ScholarGoogle Scholar