Investigating the effect of entrance channel mass asymmetry on fusion reactions using the Skyrme energy density formalism

NUCLEAR PHYSICS AND INTERDISCIPLINARY RESEARCH

Investigating the effect of entrance channel mass asymmetry on fusion reactions using the Skyrme energy density formalism

M. M. Hosamani
A. Vinayak
N. M. Badiger
Nuclear Science and TechniquesVol.31, No.9Article number 89Published in print 01 Sep 2020Available online 29 Aug 2020
6500

In the present investigations, the fusion cross-sections for the formation of 200Pb compound nucleus (CN) using 16O + 184W, 30Si + 170Er, and 40Ar + 160Gd nuclear reactions at energies above the Coulomb barrier were calculated to understand the effect of entrance channel mass asymmetry (α) on the fusion reactions; the Skyrme energy density formalism (SEDF) was used for this calculation. The SEDF uses the Hartree–Fock–Bogolyubov (HFB) computational program with Skyrme forces such as SkM*, SLy4, and SLy5 to obtain the nucleus-nucleus potential parameters for the above reactions. Using the SEDF model with SkM*, SLy4, and SLy5 interaction forces, the theoretical fusion cross-sections were determined to be above the barrier energy and compared with the available experimental fusion cross-sections. The results show a close agreement between the theoretical and experimental values for all selected systems at energies well above the barrier. However, near the barrier energies, the theoretical values were observed to be higher than the experimental values.

Skyrme forceEnergy density formalismHartree–Fock–BogolyubovThomas–Fermi modelCoupled-channel calculation

1. Introduction

The inter-nuclear potential is one of the most important factors to describe the heavy-ion nuclear reaction dynamics [1]. It is well known that the study of heavy-ion fusion reactions, in which the interacting nuclei effectively interact with each other, will probe the pre-saddle region as well as the field inside the Coulomb barrier extensively [2]. Further, the study of heavy-ion nuclear reactions can provide major information about the barrier between the interacting nuclei, which hinders the synthesis of superheavy elements. Conversely, by studying the shape and size of the barrier distribution, the dynamics of the interacting nuclei can be investigated [3]. For a given system, the height and width of the barrier—in accordance with the semi-classical theory—remain same, and it is interesting to evaluate the effect of incident energy on such a barrier. Furthermore, theoretically, the reaction barrier is not very well defined and to some extent, it depends on the way it is calculated [4, 5]. For classical simplification, we are assuming that the interaction barrier is estimated from the interaction energy of the interacting nuclei, as a function of the relative distance (R) between them. However, the dynamical calculations of nuclear reactions were also performed by solving the classical equations of motion and using the liquid drop model [6, 7]. The estimated fusion barrier is also a deciding parameter for evaluating the probability of interaction of the heavy ions. In other words, the fusion barrier height and width are the deciding factors to evaluate the probability of fusion reactions. Therefore, the aim of the present investigations is to calculate the interaction potential using the energy density formalism (EDF) [8] with Skyrme forces [9]. Considering this, the total energy of the fermions of the system is given as a function of the one-body density [10]. Further, different Skyrme force gives different barrier height, position, and curvature [11, 12]. Moreover, this force is a zero-range force and density dependent by nature. This model can be applied for inter-nuclear potential, which consistently considers the saturation property of the nuclear matter and it is known as the EDF. This model assumes that the density of each colliding ion remains the same at all distances during the collision. Using this formalism and various types of Skyrme forces, the earlier studies have shown that this method can account for the elastic scattering and fusion barrier height of 16O + 16O reaction [13]. Brink and Stancu [14] have investigated the applicability of this method using the Skyrme energy functional. They have shown that both the EDF and proximity potential are consistent with each other. The same work has also been performed by Dobrowolski et al. [15] using a higher order Thomas–Fermi approximation (TFA) for the kinetic energy and spin-orbit-densities. Recently, the EDF has been applied to simplify the coupled-channel calculations for heavy-ion nuclear fusion reactions [16-18]. The Skyrme energy density formalism (SEDF) is given as a sum of the spin-orbit-densities dependent universal functions; the universal function parameters are obtained for different Skyrme forces [19]. This method is advantageous in solving the barrier modification problems by introducing the effects of either modifying the Fermi density parameters (the half-density radii and surface thickness for exact SEDF calculations) or constants of the parameterized universal functions, as mentioned in reference [14]. Further, for nuclear systems, the semi-classical extended Thomas–Fermi (ETF) approach is extended at finite temperature [20, 21] and is limited for free energy and entropy density functions only. In the semi-classical formalism, we have included the temperature effects, in the following spin-orbit density J (ρ) via the nuclear density, taken as the two-parameter Fermi-density function [22]. This follows from the fact that the HF density approximately takes a shape of the Fermi type at higher temperatures; the density becomes flat in the interior region of smaller "r" values. Furthermore, the shell model density distributions, the Fermi density distributions, and the total density interaction potential are approximately the same, at least in the surface region of relevance for the heavy-ion collision. In actual case, the barrier height decreases with increase in temperature. However, there is a vital point of difference in temperature dependence of the two approaches, whereas the order of the barrier remains the same in the semi-classical case. The different filling of the shell model states changes the approach of the same microscopic shell models. Therefore, we have applied the EDF with Skyrme forces SkM*, SLy4 and SLy5 to obtain the fusion reaction cross-sections for the systems that form the same compound nucleus (CN) in the mass region ~ 200 amu, without adjusting any parameters in the Skyrme forces. The nuclear reactions 16O + 184W, 30Si + 170Er, and 40Ar + 160Gd, which were studied in the present investigations, form the same CN 200Pb with different entrance channel mass asymmetry (α). The effect of mass asymmetry on the fusion reaction and the role of incompressibility (K) in nuclear dynamics are discussed in the following sections. The calculated fusion cross-sections are compared with the others experimental values and with cross-section obtained by the coupled-channel CCFULL code [23]. The CCFULL code calculates the fusion cross-section and mean-angular momentum of the CN under the influence of coupling between the relative motion and several nuclear collective motions. Using this formalism, the fusion cross-sections for the above systems in the given energy range well above the barrier were determined.

2. Theoretical background

It is well known that the ion-ion interaction potential can be generated using the SEDF [24]. Using the SEDF, it is possible to understand the fusion of two nuclei at various energies, predict the mass of nuclei as accurate as the semi-empirical mass formula, predict the ground state properties of the nuclei, and predict the fission barrier heights for all nuclei available in the periodic table [25]. Generally, in nuclear reactions the total potential is a combination of the nuclear, Coulomb and spin parts. However, in the SEDF, the total potential VT(R) is the combination of the nuclear part VN(R), which is the inclusion of spin-orbit interaction, and the Coulomb part VC(R). This total potential is given as:

VT(R) =VN(R) +VC(R),   (1) VT(R)=VB12μ2ω2(RRB)2, (2)

where equation (2) is a parabolic approximation of the total potential described by equation (1) around the barrier. The height VB and position RB of the fusion barrier can be obtained by applying the conditions [26]:

(dVT(r)dr)r=RB=0 and (d2VT(r)dr2)(r=RB)0. (3)

In the SEDF, VN (R) between the projectile and target is a function of the distance between the two nuclei. This potential is determined using the Hamiltonian H that depends on the nucleonic density ρ, kinetic energy density τ, and spin density J. It is given by

VN(R) =H12(R) H1(R) H2(R), (4)  H12(R)=H[ρ1p(r)+ρ2p(rR),ρ1n(r)+ρ2n(rR)]dr, (5) H1(R)= H[ρ1p(r),ρ1n(r)]dr, (6) H2(R)= H[ρ2p(r),ρ2n(r)]dr, (7)

In these equations, ρ1n, ρ2n are the frozen neutron densities and ρ1p, ρ2p are the frozen proton densities of the interacting nuclei. The H (r)  is associated with the Skyrme interaction and is given by:

H(r)=2τ2m+t02[1+x02ρ2(x0+12)(ρn2+ρp2)]+14(t1+t2)ρτ+18(t2+t1)(ρnτn+ρpτp)+116(t23t1)ρ2ρ+132(3t1+t2)(ρn2ρn+ρp2ρp)+116(t1t2)(Jn2+Jp2)+14t3ρnρpρ12W0(ρJ+ρnJn+ρpJp), (8)

where ρ = ρn + ρp, τ = τn + τp, and J=Jn+Jp. The parameters ρ, τ, and J are the local densities of the nucleons, kinetic energy, and spin-orbit, respectively. The subscripts n and p represent neutron and proton, respectively. The parameters of different Skyrme forces t0, x0, t1, t2, t3 and W0 are estimated by fitting the different properties of nuclei [27]. The mass of the nucleons is indicated by m. The parameters τ and J are estimated by the semi-classical ETF model [28]. Further, the fusion barrier is a combination of the nuclear and coulomb potentials, as given in equation (1). Therefore, the coulomb potential between the two-interacting heavy-ions can be expressed as:

Vc(R,θ)=Z1Z2e2R+920π Z1Z2e2R3 i=12Ri2β2iP2(cosθi)+ 37π Z1Z2e2R3  i=12Ri2[β2iP2(cosθi)]2, (9)

where  θi, Ri, and β2i are the angle between radius vector and symmetry axis, effective sharp radius, and the quadruple deformation of ith nucleus, respectively. Ri=1.28 Ai1/30.76+0.8 Ai1/3, as given by Blocki [29, 30].

During collisions, the nuclei may have several orientations. Therefore, it is preferred to consider the average over all possible relative orientations from 0o to 180o for both deformed nuclei in the study of fusion reactions [31]. Since both the nuclei are in the same plane i.e. P2(cosθi)=12(3cosθi21), the average possible orientation is unity [32].

In SEDF, the nucleon density distribution is determined using the ion-ion interaction potential. The density distribution of the interacting nuclei can be obtained by different approximation methods. However, the self-consistent Hartree–Fock–Bogolyubov (HFB) method can precisely determine the Skyrme effective interaction using a complete quantum approach [33]. The density distribution of the nucleons in the nuclei is calculated by the code HFBRAD(v1.00), which does not include the deformation of the nucleus [33].

ρi(r)=ρ0i[1+exp(rR0iai)]  (10)

The central density, ρ0i, is given by

ρ0i=3Ai4πR0i3[1+π2ai2R0i2]1 (11)

The half density radii R0i  and the surface thickness ai are determined by applying a polynomial fit to the experimentally determined values of the nuclei in the mass range from 4 to 238 amu [21, 34].

3. Result and discussions

Heavy ion-induced fusion reaction is a sensitive tool to probe the shape, structure, and size of an atomic nucleus as well as the reaction dynamics of a di-nuclear system (DNS) [34]. In the fusion of heavy ions, the relative kinetic energy of an entrance channel converts into the internal excitation energy of the CN and this internal excitation energy is distributed among all the nucleons in the nucleus. The fusion reactions can be explained based on the energy density distribution of the interacting nucleons in the nuclei. Therefore, in the present work, calculations of the SEDF were carried out using energy density distributions. Systems, such as 16O + 184W, 30Si + 170Er, and 40Ar + 160Gd were used, which form the same CN, 200Pb. The ion-ion potential parameters are determined using compatible Skyrme forces like SkM*, SLy4, and SLy5. It is important to notice that in the present investigations, the reactions with different entrance channels and α were selected deliberately to form the same CN. The α are different with respect to the Businaro–Gallone mass asymmetry (αBG) [35]. Some fitting parameters are required for the corresponding Skyrme forces to calculate the ion-ion potential parameters using the HFB method. These parameters are tabulated in Table 1 with their respective forces. The Woods–Saxon potential parameters, calculated using the Akyuz–Winther (AW) formalism, along with α and the deformation parameters for the corresponding systems are given in Table 2. The local ρ, τ, and J density distributions of the nuclei for all the above reactions were determined using the HFB program. A typical local densities distribution of 16O + 184W for both neutron and proton are shown in Figs. 1 and 2. The Skyrme energy density function H(r), as given in Eq. (8), was evaluated using local densities, which are obtained from the Skyrme forces of the HFB program. The interaction nuclear potential VN(R), as given in Eq. (4), was determined using this H(r). The nuclear potential for all the Skyrme forces for the selected reaction systems is shown in Figure 3. In this figure, we notice that the depth of the potential depends on the K of nuclear matter and the entrance channel α of the reaction; K is a significant component of the equation of state (EOS) of nuclear matter and it is one of the interesting subjects in the study of heavy-ion fusion reactions. For the investigation of the fusion process, different K values were used, which result from the corresponding effective interactions [36, 37]. The results obtained in our calculations show that the theoretical values are sensitive to the values of K. Therefore, it is very interesting to describe the nuclear fusion reaction using different effective interactions with different K and α values. Consequently, by knowing the K value, we can understand the variations in K of nuclear matter for various energies of projectile.

Figure 1:
Typical local densities distribution (ρ, τ, and J) for neutrons and protons of the target 184 W.
pic
Figure 2:
Typical local densities distribution (ρ, τ, and J) for neutrons and protons of projectile 16 O.
pic
Figure 3:
The calculated nuclear potential for 16 O + 184 W, 30 Si + 170 Er, and 40 Ar + 160 Gd using the Skyrme forces SkM*, SLy4, and SLy5.
pic
Table 1:
Fitting parameters for the different Skyrme forces used in the SEDF.
Parameters SkM* SLy4 SLy5
t0 -2645.0 -2488.913 -2488.354
t1 410.0 486.818 484.230
t2 -135.0 -546.395 -556.690
t3 15595.0 13777.0 13757.0
x0 0.09 0.834 0.776
x1 0.0 -0.344 -0.317
x2 0.0 -1.000 -1.000
x3 0.0 1.354 1.263
γ 1/6 1/6 1/6
W0 120.0 123.0 125.0
Show more

For the calculation of EDF based on each Skyrme force, we have estimated the Fermi density distributions using the parameters obtained from the fitting results of the HFB. The calculated density distributions’ diffuseness parameters of the nucleons for the projectiles and targets of the systems 16O + 184W, 30Si + 170Er, and 40Ar + 160Gd are shown in Figs. 1 and 2 using Skyrme forces SkM*, SLy4 and SLy5. The interaction potentials were obtained for above mentioned reactions using the calculated densities with the respective Skyrmes. The properties of a calculated fusion barrier potential, viz. V0, a0, and RB, are given in Table 3 for the corresponding Skyrmes. The results clearly indicate that V0 increases with the increasing K value. In other words, the increase in the value of K increases the depth of reaction potential pocket. This observed increase in the potential depth may be due to a vital role of the peripheral nucleons in the heavy-ion reaction. Another set of potential parameters for the above systems was calculated using the AW formalism for the CCFULL calculation (Table 2) [38]. It is also to be noted that the system 16O + 184W has α ˃ αBG, and the systems 30Si + 170 Er and 40Ar + 160Gd have α ˂ αBG. In case of incompressibility, SkM* [39] has lesser incompressibility as compared to SLy4 and SLy5 [40]. The nuclear potential parameters obtained from different Skyrme forces and with different incompressibility of the HFB are given in Table 3. The Coulomb part of the nuclear reaction was also estimated using equation (9) for all the systems. The total potential for the three Skyrme forces were determined for all the selected systems using the nuclear and coulomb potentials. There exists a potential pocket that is responsible for the capture of the nuclei. The deformation of the interacting nuclei is neglected here to avoid the long duration of the calculation [41, 42]. With respect to the α, the system 40Ar + 160Gd may have less capture probability; therefore, the fission barrier height appears to be very less compared to the remaining systems, and hence, the contribution of the quasi-fission may lead to less fusion probability. To confirm this, our group is planning to experimentally measure these type of systems in the above mentioned energy range using a recoil mass separator. By incorporating these potential parameters—derived from the different Skyrme forces—in the CCFULL code as an input, we have calculated the fusion cross-section for reactions 16O + 184W, 30Si + 170Er, and 40Ar + 160Gd. The fusion cross-sections of the CN are calculated using the following equation:

Table 3:
Calculated potential parameters using the Skyrme forces.
Skyrme forces Incompressibility K (MeV) Systems V0 (MeV) a0 (fm) RB (fm)
SkM* 216.65 16O+184 W 49.50 0.41 6.9
30Si+170Er 48.00 0.40 7.2
40Ar+160Gd 46.00 0.41 7.1
SLy4 229.90 16O+184 W 69.50 0.459 7.3
30Si+170Er 66.80 0.45 7.2
40Ar+160Gd 64.10 0.45 7.19
SLy5 229.90 16O+184 W 70.20 0.45 7.3
30Si+170Er 67.10 0.455 7.2
40Ar+160Gd 64.80 0.46 7.19
Show more
Table 2:
The parameters α, β2, and the potential parameters calculated using the AW formalism for the corresponding systems.
Reaction Systems Bass Barrier (MeV) Mass asymmetry Deformation Potential parameters using AW
α αBG β2p β2T V0 (MeV) R0 (fm) a0(fm)
16O+184W 74.19 0.84 0.83 0.364 0.236 63.90 1.20 0.65
30Si+170Er 113.24 0.70 0.83 0.315 0.336 72.93 1.20 0.67
40Ar+160Gd 100.43 0.60 0.83 0.251 0.353 77.57 1.20 0.68
Show more
σfus(E)=πk2l(2l+1)Pl(E) (12)

where Pl(E) is the transmission probability or penetrability for the l-wave scattering and k is the local wave number for the lth partial wave [23]. It has been used to explain the fusion cross-section in CCFULL at energies near and above the barrier [38]. The theoretical fusion cross-section at various excitation energies have been compared with the available experimental values [43, 44] and the theoretical fusion cross-sections obtained by the CCFULL code with Woods–Saxon potential parameters of the AW formalism as the inputs. Figure 4 gives the comparison of fusion cross-sections. As shown in this figure, the density distributions of the heavy ions, which were calculated by different methods and using different parameters of Skyrmes, influence the fusion barrier, which are calculated using the EDF. With this motivation, we have carried out a comparative study of the fusion barrier using SkM*, SLy4, and SLy5 forces, in addition to the Fermi density distribution including the parameters determined in the SEDF. The total fusion cross-sections were computed from the estimated potentials and compared with the available experimental data. By varying the K values with different α of the same CN, a good agreement was obtained between the calculated fusion cross-sections derived from different Skyrme forces and the available experimental fusion cross-sections at different bombarding energies. It was also observed that as the K value increases, the depth of the reaction capture pocket increases. In other words, it increases the probability of the fusion and level density. Similarly, with the increase in α, the fusion probability would also increase. It implies that α between the projectile and the target nuclei is responsible for the deformation of the DNS during the dynamic process. If α value is large, the DNS will be less deformed and consequently, the contribution of the fission and fission-like probabilities will be less. Conversely, if α value is small, the DNS will be more deformed, and therefore, the contribution of the fission or fission-like probability will be more. Further, K also depends on the nucleons lying on the surface or on the periphery of the nuclei, which are more actively involved during the interaction, even though the system is less asymmetric. In other words, K would enhance the probability of the capture cross-section even for less asymmetric systems.

Figure 4:
The fusion cross-sections as a function of the lab energies for the reactions 16 O + 184 W, 30 Si + 170 Er, and 40 Ar + 160 Gd. Theoretical values have been compared with the available experimental values.
pic

4. Conclusion

In the present investigations, different Skyrme forces were used to estimate the fusion cross-sections at energies well above the Coulomb barrier for the reactions 16O + 184W, 30Si + 170Er, and 40Ar + 160Gd. In this study, three systems were selected such that the first reaction has ααBG, the second one has α ˂ αBG, and the third one has α ˂˂ αBG. The results of the SEDF calculations show that the calculated values are in close agreement with the experimental values of 16O + 184W system, indicating that whenever α is close to αBG the fusion cross-section values, based on Skyrme forces, closely agree with the experimental values. On the other hand, in the case of 30Si + 170Er (α ˂ αBG), the calculated fusion cross-section based on Skyrme forces were found to be higher than the experimental values. It is interesting to note that as the energy increases the deviation between the theoretical and experimental values decreases. In the case of 40Ar + 160Gd, there are no experimental data on fusion cross-sections. Therefore, we have compared the calculated fusion cross-sections based on Skyrme forces with the values obtained from the CCFULL code. Interestingly, the calculated theoretical values closely agree with the experimental values in the higher energy region, but in the lower energy region, the values predicted by CCFULL are higher than the theoretical values. Therefore, we may conclude that the fusion cross-section values, obtained with different Skyrme forces, can be used not only to estimate the fusion cross-section for any projectile and target combination, but it can also throw light on the effect of entrance channel α on the fusion cross-section. Further, there is no need to change the parameters of Skyrme force to modify the barrier as the theoretical framework based on the Skyrme forces can reproduce the corresponding experiment values. For the fusion cross-sections at sub-barrier energies, different Skyrme forces are required for the different reactions to reproduce the experimental data. A single Skyrme force cannot reproduce the fusion cross-sections over the entire energy range for the selected reactions.

References
[1]. W. Kohn, L. J. Sham,

Quantum density oscillations in an inhomogeneous electron gas

. Phys. Rev. 137, A1697 (1965). https://doi.org/10.1103/PhysRev.137.A1697
Baidu ScholarGoogle Scholar
[2]. P. Hohenberg, W. Kohn,

Inhomogeneous electron gas

. Phys. Rev. 136, B864 (1964). https://doi.org/10.1103/PhysRev.136.B864
Baidu ScholarGoogle Scholar
[3]. R. Kumar, M. K. Sharma, R. K. Gupta,

Fusion reaction cross-sections using the Wong model within Skyrme energy density based semiclassical extended Thomas Fermi approach

. Nucl. Phys. A 870-871, 42-57 (2011). https://doi.org/10.1016/j.nuclphysa.2011.09.010
Baidu ScholarGoogle Scholar
[4]. R. K. Gupta, Dalip SinghRaj Kumar et al.,

Universal functions of nuclear proximity potential for Skyrme nucleus-nucleus interaction in a semiclassical approach

. J. Phys. G: Nucl. Part. Phys. 36, 075104 (2009).
Baidu ScholarGoogle Scholar
[5]. K. A. Brueckner, J. R. Buchler, R. C. Clark et al., Phys. Rev. 181, 1534 (1969).
[6]. H. Holm, W. Greiner,

Influence of nuclear forces on the Coulomb barrier in heavy-ion reactions

. Phys. Rev. Lett. 24, 404 (1970). https://doi.org/10.1103/PhysRevLett.24.404
Baidu ScholarGoogle Scholar
[7]. A. S. Jensen, C. Y. Wong,

Distortion of nuclei in heavy ion reactions

. Phys. Lett. B 32, 567-570 (1970). https://doi.org/10.1016/0370-2693(70)90545-9
Baidu ScholarGoogle Scholar
[8]. H. Holm, D. Scharnweber, W. Scheid et al.,

Coulomb distortion in heavy-ion reactions

. Z. Phys. 231, 450 (1970). https://doi.org/10.1007/BF01642535
Baidu ScholarGoogle Scholar
[9]. P. W. Riesenfeld, T. D. Thomas,

Effect of nuclear deformability on reaction cross sections

. Phys. Rev. C 2, 711 (1970). https://doi.org/10.1103/PhysRevC.2.711
Baidu ScholarGoogle Scholar
[10]. L. Wilets. A. Goldberg, F. H. Lewis,

New theoretical approach to nuclear heavy-ion scattering

. Phys. Rev. 2, 1576 (1970). https://doi.org/10.1103/PhysRevC.2.1576
Baidu ScholarGoogle Scholar
[11]. K. A. Brueckner, J. R. Buchler, M. M. Kelly,

New theoretical approach to nuclear heavy-ion scattering

. Phys. Rev. 173, 944 (1968). https://doi.org/10.1103/PhysRev.173.944
Baidu ScholarGoogle Scholar
[12]. T. H. R. Skyrme et al.,

Phil

. Mag. 1, 1043 (1956);
T. H. R. Skyrme et al.,

Phil

. Nucl. Phys. A 9, 615 (1959).
Baidu ScholarGoogle Scholar
[13]. C. Ngow, B. Tamain, J. Galin et al,

Calculation of interaction barriers using the energy density formalism

. Nucl. Phys. A 240, 353 (1975). https://doi.org/10.1016/0375-9474(75)90335-8
Baidu ScholarGoogle Scholar
[14]. D. M. Brink, FL. Stancu,

Interaction potential between two 16O nuclei derived from the Skyrme interaction

. Nucl. Phys. A 243, 175-188 (1975). https://doi.org/10.1016/0375-9474(75)90027-5
Baidu ScholarGoogle Scholar
[15]. A. Dobrowolski, K. Pomorski, J. Bartel,

Mean-field description of fusion barriers with Skyrme`s interaction

. Nucl. Phys. A 729, 713 (2003). https://doi.org/10.1016/j.nuclphysa.2003.09.008
Baidu ScholarGoogle Scholar
[16]. L. Kaur, R. Kumari,

On the role of deformed Coulomb potential in fusion using energy density formalism

. Pramana-Journal of Physics 85, 639 (2015). https://doi.org/10.1007/s12043-014-0898-z
Baidu ScholarGoogle Scholar
[17]. O. N. Ghodsi, F. Torabi,

Comparative study of fusion barriers using Skyrme interactions and the energy density functional

. Phys. Rev. C 92, 064612 (2015). https://doi.org/10.1103/PhysRevC.92.064612
Baidu ScholarGoogle Scholar
[18]. O. N. Ghodsi, F. Torabi,

Effect of nuclear matter incompressibility on the 16O + 208Pb system

. Phys. Rev. C 93, 064611 (2016). https://doi.org/10.1103/PhysRevC.93.064611
Baidu ScholarGoogle Scholar
[19]. R. J. Lombard,

The energy density formalism in nuclei

. Annals of Physics 77, 380-413 (1973). https://doi.org/10.1016/0003-4916(73)90422-3
Baidu ScholarGoogle Scholar
[20]. M. Brack, C. Guet, H. B. Hakansson,

Selfconsistent semiclassical description of average nuclear properties - A link between microscopic and macroscopic models

. Phys. Rep. 123, 275 (1985).
Baidu ScholarGoogle Scholar
[21]. J. Bartel, M. Brack, M. Durand,

Extended Thomas-Fermi theory at finite temperature

. Nucl. Phys. A 445, 263-303 (1985). https://doi.org/10.1016/0375-9474(85)90071-5
Baidu ScholarGoogle Scholar
[22]. D. Singh et al., Proceedings of DAE-BRNS symposium on Nuclear Physics, Vol B46, P-254 (2003).
[23]. K. Hagino, N. Rowley, A. T. Kruppa,

A program for coupled-channel calculations with all order couplings for heavy-ion fusion reactions

. Comp. Phys. Comm. 123, 143-152 (1999). https://doi.org/10.1016/S0010-4655(99)00243-X
Baidu ScholarGoogle Scholar
[24]. A. S. Umar, V. E. Oberacker,

Three-dimensional unrestricted time-dependent Hartree-Fock fusion calculations using the full Skyrme interaction

. Phys. Rev. C 73, 054607 (2006). https://doi.org/10.1103/PhysRevC.73.054607
Baidu ScholarGoogle Scholar
[25]. J. W. Negele,

The mean-field theory of nuclear structure and dynamics

. Rev. Mod. Phys. 54, 913 (1982). https://doi.org/10.1103/RevModPhys.54.913
Baidu ScholarGoogle Scholar
[26]. K. T. R. Davies et al., in Treatise on heavy ion science, edited by D. A. Bromley (Plenum, New York, 1985), Vol.03 P.03.
[27]. J. P. Svenne,

Fourteen years of self-consistent field calculations: What has been learned

. Adv. Nucl. Phys. 11, 179 (1979).
Baidu ScholarGoogle Scholar
[28]. R. K. Gupta, D. Singh, W. Greiner,

Semiclassical and microscopic calculations of the spin-orbit density part of the Skyrme nucleus-nucleus interaction potential with temperature effects included

. Phys. Rev. C 75, 024603 (2007). https://doi.org/10.1103/PhysRevC.75.024603
Baidu ScholarGoogle Scholar
[29]. J. Blocki, J. Randrup, W. J. Swiatecki et al.,

Proximity forces

. Ann. Phys. (NY) 105, 427 (1977). https://doi.org/10.1016/0003-4916(77)90249-4
Baidu ScholarGoogle Scholar
[30]. R. K. Puri, R. K. Gupta,

Analytical formulation of the ion-ion interaction potential including spin density term in energy density formalism

. Int. J. Mod. Phys. E 1, 269 (1992). https://doi.org/10.1142/S0218301392000138
Baidu ScholarGoogle Scholar
[31]. G. B. Arfken et al, Mathematical Methods for Physicists: A comprehensive guide, 7th edition, Academic Press is an imprint of Elsevier 225 Wyman Street, Waltham, MA 02451, USA.
[32]. Ishita Sharma, R. Kumar, M. K. Sharma,

Probing the role of Skyrme interactions on the fission dynamics of the 6Li + 238U reaction

. Eur. Phys. J. A 53, 140 (2017). https://doi.org/10.1140/epja/i2017-12331-5
Baidu ScholarGoogle Scholar
[33]. K. Bennaceur,

Coordinate-space solution of the Skyrme-Hartree-Fock-Bogolyubov equations within spherical symmetry

. The program HFBRAD (v1.00). Comp. Phys. Comm. 168, 96 (2005). https://doi.org/10.1016/j.cpc.2005.02.002
Baidu ScholarGoogle Scholar
[34]. A. Shamlath, E. Prasad, N. Madhavan et al.,

Fusion and quasifission studies in reactions forming Rn via evaporation residue measurements

. Phys. Rev. C 95, 034610 (2017). https://doi.org/10.1103/PhysRevC.95.034610
Baidu ScholarGoogle Scholar
[35]. B. B. Back, R. R. Betts, J. E. Gindler et al.,

Angular distribution in heavy-ion induced fission

. Phys. Rev. C 32, 195 (1985). https://doi.org/10.1103/PhysRevC.32.195
Baidu ScholarGoogle Scholar
[36]. S. Misicu, H. Esbensen,

Signature of shallow potentials in deep sub-barrier fusion reactions

. Phys. Rev. C 75, 034606 (2007). https://doi.org/10.1103/PhysRevC.75.034606
Baidu ScholarGoogle Scholar
[37]. H. Esbensen, S. Misicu,

Hindrance of 16O + 208Pb fusion at extreme sub-barrier energies

. Phys. Rev. C 76, 054609 (2007). https://doi.org/10.1103/PhysRevC.76.054609
Baidu ScholarGoogle Scholar
[38]. K. Hagino, N. Takigawa,

Subbarrier fusion reactions and many-particle quantum tunneling

. Progress of Theoretical Physics, 128, 1061-1106 (2012). https://doi.org/10.1143/PTP.128.1061
Baidu ScholarGoogle Scholar
[39]. J. Bartel, P. Quentin,

Towards a better parameterization of Skyrme-like effective forces: A critical study of the SkM force

. Nucl. Phys. A 386, 79 (1982).
Baidu ScholarGoogle Scholar
[40]. E. Chabanat, P. Bonche, P. Haensel et al.,

A Skyrme parameterization from subnuclear to neutron star densities part II

. Nuclei far from stability. Nucl. Phys. A 635, 231 (1998). https://doi.org/10.1016/S0375-9474(98)00180-8
Baidu ScholarGoogle Scholar
[41]. J. Dobaczewski, J. Dudek,

Solution of the Skyrme-Hartree-Fock equations in the Cartesian deformed harmonic oscillator basis I

. The method. Comput. Phys. Comm. 102, 166 (1997).
Baidu ScholarGoogle Scholar
[42]. J. Dobaczewski, J. Dudek,

Solution of the Skyrme-Hartree-Fock equations in the Cartesian deformed harmonic oscillator basis I

. The program HFODD. Comput. Phys. Comm. 102, 183 (1997).
Baidu ScholarGoogle Scholar
[43]. P. D. Shidling, N. M. Badiger, S. Nath et al.,

Fission hindrance studies in 200Pb: Evaporation residue cross section and spin distribution measurements

. Phys. Rev. C 74, 064603 (2006). https://doi.org/10.1103/PhysRevC.74.064603
Baidu ScholarGoogle Scholar
[44]. G. Mohanto, N. Madhavan, S. Nath et al.,

Entrance channel effect on ER spin distribution

. Nucl. Phys. A 890, 62 (2012). https://doi.org/10.1016/j.nuclphysa.2012.07.004
Baidu ScholarGoogle Scholar