-
To investigate the various nuclear mechanisms the Quantum Mechanical Fragmentation Theory (QMFT) based [37−39] Dynamical cluster-decay model (DCM) [12, 13, 18−25] is worked out in the terms of collective coordinates of mass and charge asymmetry, relative separation coordinate 'R', deformations
$ \beta_{\lambda i }$ ($\lambda$ = 2, 3, 4 and i = 1, 2), orientations of the deformed fragments$ \theta_i $ (i=1,2), and the neck parameter$ (\Delta R) $ . The mass and charge asymmetry are stated as follows, respectively:$ \eta_A = \frac{A_1-A_2}{A_{CN}} ;\quad \eta_Z = \frac{Z_1-Z_2}{Z_{CN}} $
(1) where
$ A_i $ and$ Z_i $ (where i=1,2) represent the mass and charge numbers of the respective fragments and ACN and ZCN, is mass and charge of Compound nucleus. The temperature-dependent collective potential energy, or fragmentation potential, can be expressed using the relative spacing R and$ \eta $ -coordinates as:$ \begin{aligned}[b] V(R,\eta, T) &={\sum\limits_{i=1}^{2}V_{LDM}(A_i, Z_i, T)+ \sum\limits_{i=1}^{2} \delta U_i \exp \left(-\frac{T^2}{T_{0}^2}\right)}\\ &{+ V_C(R, Z_i, \beta_{\lambda i}, \theta_i, T)+ V_N(R, Z_i, \beta_{\lambda i}, \theta_i, T)} \\ &{+ V_{\ell}(R, Z_i, \beta_{\lambda i}, \theta_i, T).} \end{aligned} $
(2) Here
$ V_{LDM} $ corresponds to the liquid drop part of the binding energy of Davidson et al. [40] and$ \delta U $ is the shell corrections from Myers and Swiatecki [41], the value of$ T_0 $ =1.5 MeV is taken from classical work of Jensen and Damgaard [42],$ V_C $ ,$ V_N $ and$ V_\ell $ is the Coulomb, nuclear interaction and angular momentum-dependent potential for deformed and oriented nuclei.The preformation probability of decaying fragments in
$ \eta $ -coordinates at$ R=R{_a} $ is determined by solving the stationary Schrodinger equation as:$ \begin{equation} P_0 = \sum\limits_{\nu = 0}^{\infty} |\psi^{\nu}(\eta(A_{i}))|^2\sqrt{B_{\eta \eta}}\frac{2}{A_{CN}}exp(-E_{\eta}^{\nu}/T) \end{equation} $
(3) with the ground and excited state solutions given by
$ {\nu} $ = 0,1,2.... and the smooth hydrodynamical mass parameter represented by$ B_{\eta \eta} $ .[43].On the other hand, the barrier penetration probability P of decaying fragments is determined by WKB integral
$ P = exp\left[-\frac{2}{\hbar}\int_{R_a}^{R_b}{[2\mu(V(R)-Q_{eff})]}^{1/2}dR\right] $
(4) with
$ V(R_a,T) = V(R_b, T) = TKE(T) = Q_{eff} $
(5) regarding the two turning points. TKE denotes the total kinetic energy and
$ Q_{eff} $ is the effective Q value.In context to the compound nucleus decay, the following postulate is employed to describe the occurrence of the initial turning point.
$ R_a(T) = R_1(T) + R_2(T) + \Delta R (T) $
(6) $\qquad\;\, = R_t(T) + \Delta R (T), $
(7) The influence of neck formation, the neck length parameter denoted as
$ \Delta R (T) $ is accounted by [45−48]. The radii are taken from [59−63].The temperature T is related to the excitation energy
$ E^*_{CN} $ , through the semi-empirical statistical relation as [44] :$ E^*_{CN} = E_{c.m.} + Q_{in} = \frac{1}{a} A_{CN}T^2 - T \ (MeV). $
(8) For this system, we have used a = 9, the entrance channel Q-value, denoted as
$ Q_{in} $ , is calculated using the equation$ Q_{in} $ =$ B_1 + B_2 - B_{CN} $ , where$ B_1 $ ,$ B_2 $ , and$ B_{CN} $ represent the binding energies of the target, projectile, and compound nucleus, respectively [49].For the multipole-multipole interaction between two separated nuclei, the Coulomb potential can be expressed as given in references [50−52].
The equation accounts for the influence of nuclear deformation on the radius vector
$ R_i $ is$ R_i(\alpha_i, T) = R_{0i}(T)\left[1+\sum\limits_{\lambda}\beta_{\lambda i}Y_{\lambda}^{(0)}(\alpha_i)\right]. $
(9) Here i=1,2,
$ \lambda $ =2,3,4 and the variable$ \alpha_i $ represents the angle formed between the symmetry axis and the radius vector$ R_i $ of the colliding nuclei.In the above expression, the T-dependence nuclear radius term
$ R_{0i} (T) $ is given as$ R_{0i}(T) = R_{0i}[1 + 0.007T^2]. $
(10) Here,
$ R_{0i} $ =1.28$ A_{i}^{1/3} $ −0.76 + 0.8$ A_{i}^{-1/3} $ in fm.The angular momentum effects impart additional energy to the rotational motion, and the corresponding rotational potential is computed in the references as [53, 54]. Finally, in terms of
$ P_0 $ and$ P $ co-ordinates, the decay cross-sections are computed.$ \sigma =\frac{\pi}{k^2} \sum\limits_{\ell=0}^{\ell_{max}}(2\ell+1)P_0P ; k = \sqrt{\frac{2\mu E_{c.m.}}{\hbar^2}}, $
(11) where
$ \mu $ is the reduced mass.The collective clusterization process within the domain of DCM is used to calculate the cross-section of values compound nucleus (CN) processes such as evaporation residue and fusion-fission (i.e
$ \sigma_{ER} $ and$ \sigma_{ff} $ ) as$ \qquad\qquad\quad \sigma_{ER} = \sum\limits_{A_2=1}^{4} \sigma(A_1, A_2) ; $
(12) $ \qquad\qquad\quad \sigma_{ff} = 2\sum\limits_{A_2=A/2-20}^{A/2} \sigma(A_1, A_2) $
(13) and for the non-compound nucleus (nCN) processes such as Quasi fission and fast fission using the formula as
$ \sigma_{QF} =\frac{\pi}{k^2} \sum\limits_{\ell=0}^{\ell_{max}}(2\ell+1)P_{ic} $
(14) Where
$ P_{ic} $ is the penetration probability.$ \sigma_{FF} =\frac{\pi}{k^2} \sum\limits_{\ell_{Bf}}^{\ell_{max}}(2\ell+1)P_{0} $
(15) Here,
$ P_0 $ is calculated by solving Schrodinger wave equation for fission fragments for angular momentum varying from$ \ell_{Bf} $ to$ \ell_{max} $ and barrier penetration probability is considered to be maximum (i.e. P=1).The
$ V_C $ and$ V_\ell $ are widely comprehended within the field, however the$ V_N $ lacks a specific definition. Numerous theoretical frameworks exist for the computation of nuclear interaction potentials. In this study, the Skyrme energy density formalism (SEDF) based$ V_N $ is intended to examine the stability of the heavy and superheavy mass area. -
The semi-classical extended Thomas Fermi (ETF) approach [55] based, nucleus-nucleus interaction potential in SEDF is described as
$ V_N(R) = E(R) - E(\infty), $
(16) i.e.,the potential of the interaction between two nuclei can be characterised as a function of the separation distance.
$ V_N(R) $ denotes the difference in the expected energy value, referred to as E, between the colliding nuclei when they are overlapping at a finite separation distance R, and when they are completely separated at R =$ \infty $ .$ E = \int H(r)dr. $
(17) The Skyrme Hamiltonian density is precisely given as [26, 57]
$ \begin{aligned}[b] H(\rho,\tau,{\bf{J}})&={\frac{\hbar^2}{2m} \tau + \frac{1}{2}t_0\left[\left(1+\frac{1}{2}x_0\right)\rho^2-\left(x_0+\frac{1}{2}\right)\left(\rho_n^2+\rho_p^2\right)\right]}\\ & {+\frac{1}{2}\sum\limits_{i=1}^{3} t_{3i}\rho^{\alpha i}\left[\left(1+\frac{1}{2}x_{3i}\right)\rho^2-\left(x_{3i}+\frac{1}{2}\right)\left(\rho_n^2+\rho_p^2\right)\right]}\\ & {+\frac{1}{4}\left[t_1\left(1+\frac{1}{2}x_1\right)+t_2\left(1+\frac{1}{2}x_2\right)\right]\rho\tau }\\ &{ -\frac{1}{4}\left[t_1\left(x_1+\frac{1}{2}\right)-t_2\left(x_2+\frac{1}{2}\right)\right]\left(\rho_n\tau_n+\rho_p\tau_p\right) }\\ & {+\frac{1}{16}\left[3t_1\left(1+\frac{1}{2}x_1\right)-t_2\left(1+\frac{1}{2}x_2\right)\right](\nabla\rho)^2 }\\ & { -\frac{1}{16}\left[3t_1\left(x_1+\frac{1}{2}\right)+t_2\left(x_2+\frac{1}{2}\right)\right] }\\ &{ \times\left[(\nabla\rho_n)^2+(\nabla\rho_n)^2\right]}\\ & {-\frac{1}{2}W_0\left[\rho\nabla .{\bf{J}} + \rho_n\nabla .{\bf{J}}_n + \rho_p\nabla . {\bf{J}}_p\right] }\\ & -\left[\frac{1}{16}(t_1x_1+t_2x_2){\bf{J}}^2 -\frac{1}{16}(t_1-t_2)({\bf{J}}^2_p+{\bf{J}}^2_n)\right]. \end{aligned} $
(18) Here, the nuclear density, kinetic energy density, and spin-orbit density are depicted by
$ \rho = \rho_n + \rho_p $ ,$ \tau = \tau_n + \tau_p $ and$ {\bf{J}} = {\bf{J}}_n + {\bf{J}}_p $ and m denotes the nucleon mass. The Skyrme force parameters such as$ \alpha_i $ ,$ x_1 $ ,$ x_2 $ ,$ t_1 $ ,$ t_2 $ ,$ t_3 $ ,$ W_0 $ and$ A $ are fitted by Agrawal et al. [56, 57] referring to the modified version implemented for Skyrme interactions, including GSkI, GSkII, and SSk Skyrme interactions.The densities in this study are determined using the frozen density approximation [58].
$ \begin{aligned}[b] &\rho = \rho_1 + \rho_2, \\ &\tau(\rho) = \tau_1(\rho_1) + \tau_2(\rho_2), \\ &{\bf{J}}(\rho) = {\bf{J}}(\rho_1) + {\bf{J}}(\rho_2), \end{aligned} $
(19) with
$ \rho_i = \rho_{in} + \rho_{ip} $ ,$ \tau(\rho_i) = \tau_1in(\rho_{in}) + \tau_ip(\rho_{ip}) $ and$ {\bf{J}}(\rho_i) = {\bf{J}}(\rho_{in}) + {\bf{J}}(\rho_{ip}) $ .Nuclear density
$ {\rho}_i $ is calculated using the two-parameter Fermi density distribution, as shown in [59, 60]$ \rho_i(r) = \rho_{0i}(T) \left[ 1+\exp\left(\frac{r - R_{i}(T)}{a_{i}(T)}\right)\right]^{-1}, $
(20) with central density
$ \rho_{0i}(T) = \frac{3A_{i}}{4\pi R_{i}^3(T)}\left[ 1+\frac{\pi^2 a_{i}^2(T)}{R_{i}^2(T)}\right]^{-1}. $
(21) Here
$ R_{i} $ is the nuclear radius and$ a_{i} $ is the surface thickness parameters [59−63]. Further, the T-dependence in the nuclear radii$ R_{i} $ has been discussed in Eq.(9) and the T-dependence in the surface thickness parameter$ a_{i} $ is introduced as [64, 65]$ a_{i}(T) = a_{i}(T = 0) [1 + 0.01T^2]. $
(22) In the context of the
$ V_N $ , we adopt the slab approximation of semi-infinite nuclear matter with parallel surfaces in the x-y plane. The slab is in motion along the z-direction and is separated by a distance s, with a minimum separation value denoted as$ s_{0} $ [66, 67].The expression for the interaction potential$ V_N $ (R) between two distant nuclei, where$ R $ =$ R_1 + R_2 + s $ , is provided as$ \begin{aligned}[b] V_N(R) &={ 2\pi\overline{R}\int_{s_0}^{\infty} e(s)ds } \\ &= {2\pi\overline{R}\int H(\rho , \tau , j) - [H(\rho_1 , \tau_1 , {\bf{J}}_1) + H(\rho_2 , \tau_2 , {\bf{J}}_2)]} \\ &= {V_{P}(R) + V_{J}(R),}\end{aligned} $
(23) $ \overline{R} $ is the mean curvature radius, and$ e (s) $ is the interaction energy per unit area between the two slabs.Moreover,
$ V_{P}(R) $ and$ V_{J}(R) $ represent the components of the nuclear interaction potential that are independent and dependent on the spin density, respectively.In this work, we have used two different approximations in order to calculate the reaction cross-sections. The WKB approximation and the Hill-wheeler approximations are two different approaches developed to calculate the barrier transmission probabilities. The Hill-wheeler approximation is a purely parabolic barrier and is widely appreciated for its simplicity and numerical efficiency in calculations. Although, at above barrier energies the cross-sections marge for both the approximations. Hence, in next subsection, we have applied Hill-Wheeler approximation to compute the capture cross-sections.
-
The cross-section for fusion/capture between two orientated and deformed nuclei can be determined by considering the orientation angles
$ \theta_i $ and the center of mass energy$ E_{c.m.} $ of the collision in$ \ell $ -summed Wong model [68] is calculated as follows in terms of angular momentum$ \ell $ partial waves:$ \sigma_{(E_{c.m.},\theta)} = \frac{\pi}{k^2} \sum\limits_{\ell=0}^{\ell_{max}}(2\ell+1) P_\ell(E_{c.m.},\theta), $
(24) $ P_\ell $ is the transmission coefficient for each$ \ell $ , which characterizes the penetration of barrier, and$ \ell_{max} $ is the maximum angular momentum, with$ k = \sqrt{\dfrac{2\mu E_{c.m.}}{\hbar^2}} $ and$ \mu $ as the reduced mass [70].Probability of Compound nucleus formation PCN:
Probability of completely fused compound system after the capture stage is referred as Probability of Compound nucleus formation (PCN). In the superheavy mass region, the probability of formation of compound nucleus diminish as the atomic number increases. Here, the energy dependence of fusion probability approximated by a simple relation as [33, 71, 72].
$ P_{CN} = \frac{P^{'}_{0}}{1 + exp\left(\dfrac{V_{B}^{*}-E^{*}}{\Delta}\right)}. $
(25) where
$ V_{B}^{*} $ is the compound nucleus excitation energy at$ E_{c.m.} $ $ \approx $ Coulomb barrier,$ E^{*} $ is the compound nucleus excitation energy and$ \Delta $ = 4 MeV considered for these calculations. Moreover, the parameters used in calculating the value$ P^{'}_{0} $ is taken from Ref. [72].Fusion-fission (ff) and Quasi fission (QF) lifetimes:
Further, the lifetimes for fusion-fission (ff) and quasi fission (QF) are examined using the theoretical approach and is given by [73, 74].
$ \tau_{ff \mid QF} = \frac{1}{\lambda_{ff \mid QF}} $
(26) where
$ \lambda_{ff \mid QF} $ is the fission fusion or quasi fission decay constant and is expressed as$ \begin{aligned}[b] \lambda_{ff \mid QF}& = \frac{\omega_m }{2\pi \omega_{ff \mid QF}} {\left(\sqrt{\left(\frac{\Gamma}{2\hbar}\right)^2+\omega_{ff \mid QF}^2} - \frac{\Gamma}{2\hbar}\right) }\\ &{ \times \exp \frac{- B_{ff \mid QF}}{T}} \end{aligned} $
(27) Here,
$ \omega_m $ is the frequency of the harmonic oscillator,$ \omega_{ff \mid QF} $ refers to the frequency of the inverted harmonic oscillator,$ B_{ff \mid QF } $ is the barrier corresponding to the fusion-fission and Quasi fission,$ \Gamma $ denotes an average width taken as 2 MeV and T is the temperature taken in MeV.The DCM equations are employed for the computation of cross-sections pertaining to different CN and nCN processes, as well as the determination of lifetimes associated with fusion-fission (ff) and quasi fission (QF), as elaborated in section III.
-
During heavy ion processes, the nuclei make contact with one other as a result of Coulomb interactions. In context to the centre-of-mass system, if the projectile possesses sufficient energy and the appropriate angular momentum, there exists a possibility that the nuclei could penetrate the Coulomb barrier and become confined within the potential well. This results in the attainment of the stage where the compound nucleus is in a state of complete equilibrium, also known as the CN process. Alternatively, if the captured system does not undergo significant evolution within the fusion pocket, mechanisms for example quasi fission (QF) and fast fission (FF) become relevant. In the present analysis, we have carried out our calculations to investigate the decay mechanisms of 248No* and 250No* isotopes of
$ Z=102 $ Nobelium nuclei over the broad range of centre-of-mass energies near and above the Coulomb barrier. The Dynamical cluster-decay model (DCM) is used to examine the contributions of CN (compound nucleus) and nCN (non-compound nucleus) in fission. The interaction potential is obtained by applying the Skyrme energy density formalism (SEDF) with GSkI force parameters. The included deformations extend up to the quadrupole ($ \beta_2 $ ) moment, with the optimum orientation$ \theta_{i}^{opt} $ . The detailed analysis to study the decay mechanism (ff, QF, FF) of the potential energy surfaces (PES), preformation factor$ P_0 $ , penetrability P, neck length parameters and scattering potential V(R) is carried out. Further, the capture cross-sections, are studied using the$ \ell $ -summed Wong model was used to compare it with the existing experimental data. Moreover, the decay cross-sections for the above stated processes is obtained and compared with the available experimental findings [33, 34]. Beside this, fusion-fission and quasi fission lifetime are estimated and the compound nucleus formation probability PCN is worked out.Here, we will discuss the decay of the 248No* composite system formed via 40Ca + 208Pb reaction. Fig. 1 shows the scattering potential at
$ \ell = 0\hbar $ for 40Ca + 208Pb reaction at centre mass energy$ E_{c.m.} = 187.03 $ MeV with respect to range R (fm). It is crucial to note that the first turning point$ R_a $ (which is equal to$ R_1 + R_2 + \Delta R $ ) represents the distance between the nuclei at which the fragments are assumed to have already preformed and begin to penetrate the interaction barrier. Similarly,$ R_b $ , second turning point is the point at which the process of penetrating through the interaction barrier is fully completed. The quasi fission barriers is marked and is defined as the potential difference between the barrier$ V_B $ and the potential at the first turning point$ V(R_a) $ , which depends on the angular momentum of the incoming channel at the specified incident energy.Figure 1. (color online) The calculated Scattering potential V (MeV) as a function of range R (fm) for the entrance channel of 248No* nuclei at
$ \ell = 0 \hbar $ at$ E_{c.m} $ = 187.03 MeV.To evaluate the impact of different mechanisms on the superheavy nuclei synthesis, we have computed the probability of compound nucleus formation
$ (P_{CN}) $ for both 248No* and 250No* nuclei. If the value of PCN ~ 1, then the reaction is classified as a compound nucleus (CN) reaction. The deviation of PCN from unity impart the potential to investigate the significance of the non-compound nucleus (nCN) process. The calculated PCN for the two isotopes of i.e. 248No* and 250No* of Z=102 nuclei with three different entrance channels i.e 40Ca+208Pb, 44Ca+206Pb and 64Ni+186W comes out to be 3.40 × 10−5, 1.94 × 10−5 and 1.06 × 10−5 respectively. The value of PCN being less than 1 in the calculated data suggests the existence of nCN processes. Hence, the contributions of$ \sigma_{ff} $ ,$ \sigma_{QF} $ and$ \sigma_{FF} $ , are obtained so that dynamics of superheavy system be understood. -
The experimental findings are employed to assess the ff, QF and FF cross-sections for the 248No* and 250No* nuclei. These calculations are performed via the DCM framework. Additionally, the
$ \sigma_{capture} $ are studied by utilizing the$ \ell $ -summed Wong model. The calculations are carried out by taking into account the hot optimum orientations at the energies around the barrier of the decay fragments. Initially, the research and discussion is carried out for 248No* nucleus. Fig. 2, illustrates the fragmentation potential V (MeV) regarding the decay of 248No* nucleus at three$ E_{c.m.} $ = 187.03,209.67 and 239.03 MeV for the$ \ell_{max} $ values of angular momentum obtained from the most probable fragment for which the penetrability becomes equal to one (i.e. P = 1). The T-dependent collective potential energy calculation provides information about the relative contributions of potential decay fragments. (i) From the figure, it is evident that with increase in temperature, the magnitude of fragmentation potential enhances whereas the structure remains similar as we move from lower energies to the higher excitation energies. (ii) The most probable decaying fragments are clearly indicated in the figure and can be seen that the decay fragments remains same independent of expectation energy. (iii) The angular momentum for the highest$ E_{c.m.} $ is more than for the other$ E_{c.m.'s} $ , which could be attributed to the fact that higher$ E_{c.m.} $ takes more angular momentum to decay. (iv) The configuration of fragmentation potential for light mass fragments (LPs) and intermediate mass fragments (IMFs) and the fission region remains nearly similar at extreme energies.Figure 2. (color online) Fragmentation potential V
$ (A_2) $ for the nuclear system 248No* formed in 40Ca + 208Pb reaction system at$ E_{c.m.} $ = 187.03,209.67 and 239.03 MeV, using the fixed value pf$ \Delta R $ 's for the maximum$ \ell_{max} $ values of angular momentum.Fig. 3, delves deeper into the examination of decay by plotting the preformation probability
$ (P_{0}) $ based on the fragment mass$ A_{i} $ (i = 1,2). The analysis illustrates that the fission contribution becomes more pronounced as the$ \ell $ values increase. When examining the preformation profile at different$ E_{c.m.} $ , it is clear that the value of$ P_0 $ varies, while the distribution of mass for the fission fragments remains nearly equal and exhibits an asymmetric nature, regardless of the$ E_{c.m.} $ . It is crucial to note that these secondary peaks can be linked to the potential occurrence of QF. Further, the most probable fragments and their complimentary fragments observed on the asymmetric peaks remain similar as we move from the lowest to the highest$ E_{c.m.} $ . Also, It is noteworthy to emphasize that the fragment with maximum probability to be preformed is 120Sn and its complementary fragment 128Te. Both emitted fragments are in close proximity to the magic numbers Z = 50 and N = 82, and hence the shell effects are instrumental for the asymmetric fission distribution.Figure 3. (color online) Fragment preformation probability
$ P_0 $ against the fragment mass$ A_i $ (i=1,2) for the decay of 248No* nuclei by including the$ \beta_2 $ -deformation effects, plotted at fixed neck-length parameter and highest value of angular momentum.Hence, following our understanding of the potential for fragmentation and the analysis of preformation, our next objective is to examine the conflicting processes of CN and nCN decay in the 248No* nucleus. The recent investigation involved conducting an experiment on Z = 102 nucleus using 40Ca + 208Pb reaction, and different decay mechanisms were explored and addressed by DCM. The
$ \sigma_{capture} $ incorporate the contributions from CN and nCN process, i.e$ \sigma_{capture} = \sigma_{CN} + \sigma_{nCN} $ . The current study focuses on the$ \sigma_{capture} $ for the 248No* nucleus corresponding the$ E_{c.m.} $ is calculated using the$ \ell $ -summed Wong Model and the$ \ell_{max} $ values are determined via the sharp cutoff model [82]. Table 1, clearly demonstrates that$ \sigma_{capture} $ exhibit an increase as the$ E_{c.m.} $ increases. The conclusions derived within the theoretical approach align with the experimental data. Additionally, the formation of a compound system involves two components: the evaporation residue (ER) cross-section and the fusion-fission (ff) cross-section. Mathematically, this can be expressed as$ \sigma_{CN} = \sigma_{ER} + \sigma_{ff} $ . Alternatively, we can address the hindrance in the CN formation by considering the nCN cross sections$ (\sigma_{nCN}) $ , which take into account the contributions of both QF and FF processes. In other words, we can express$ \sigma_{nCN} $ as the sum of$ \sigma_{QF} $ and$ \sigma_{FF} $ . An effort is put forth to examine the CN-fission. The fragments chosen for 248No* nuclei within the limits of A/2 ± 20, which indicate a favorable correspondence with the existing data.$ E_{c.m.} $
(MeV)$ E^*_{CN} $
(MeV)T
(MeV)$ \ell_{max} $ $ (\hbar) $ $ \Delta R $
(fm)$ \sigma_{2n}^{DCM} $
(nb)$ \sigma_{ff}^{DCM} $
(mb)$ \sigma_{ff}^{Expt.} $
(mb)$ \Delta R_{QF} $
(fm)$ \sigma_{QF}^{DCM} $
(mb)$ \sigma_{QF}^{Expt.} $
(mb)$ \Delta R_{FF} $
(fm)$ \sigma_{FF}^{DCM} $
(mb)$ \sigma_{FF}^{Expt.} $
(mb)$ \sigma_{capt.}^{DCM} $
(mb)$ \sigma_{capt.}^{Expt.} $
(mb)187.03 49 1.35 123 2.14 0.00771 160.11 159 2.27 53.49 53 - - - 212.0 212 209.67 73 1.64 134 2.21 11.3 306.25 305 2.29 62.90 62 1.58 253.35 253 627.35 620 238.19 101 1.93 147 2.22 442 280.23 280 2.30 79.10 79 1.77 575.44 572 939.96 931 Table 1. The DCM measured Evaporation residue cross-section
$ \sigma_{2n} $ , fusion-fission$ \sigma_{ff} $ , quasi fission$ \sigma_{QF} $ , fast fission$ \sigma_{FF} $ cross sections and capture$ \sigma_{capt.} $ cross section calculated using$ \ell $ -summed Wong Model for 248No* nucleus at different centre of mass energies$ E_{c.m.} $ along with relevant fitted neck length$ \Delta R $ , Temperatures T and$ \ell_{max} $ values compared with experimental data.The phenomenon known as QF, the projectile is captured by the target nucleus and a non-equilibrated compound system is formed. This system remains confined within the potential well for a brief duration. The QF contributions are calculated by taking in consideration the most probable fragments that appears on the shoulders of the preformation probability
$ P_0 $ from the Fig. 3, and their complementary fragments, and further taking into account the preformation probability of each fragment on the peak and distributing the probability among all the considered fragments. The FF process results in the formation of a mononucleus that has successfully withstood the QF process. The angular momentum of the mononucleus is significant. At high angular momentum, the rotating system's fission barrier becomes insignificant due to the enhanced rotational energy. Therefore, a highly energetic and rapidly rotating nucleus experiences rapid fission, resulting in the production of two fission fragments which have a resemblance to those produced in the fast fission process. For fission fragments ($ A_2 $ = 90-124 and the complementary fragments), the Schrodinger equation must be solved in order to find the preformation probability$ P_0 $ for FF. The$ \ell $ values range from$ \ell_{Bf} $ to$ \ell_{max} $ , where$ \ell_{Bf} $ denotes the angular momentum at which the fission barrier ceases to exist. In this case, the possibility of barrier penetration is deemed to be maximal, i.e. P = 1. Clearly, one can observe from Table 1, that the contribution of CN process of ff first increases and then decreases as we move from lowest to the highest$ E_{c.m.} $ whereas in nCN processes i.e. QF and FF contribution is large at higher energies. Further, Table 1 provides the estimated cross sections for the DCM, together with the associated values of$ \Delta R $ , temperatures T,$ \ell_{max} $ values,$ E^*_{CN} $ and$ E_{c.m.} $ , and the$ \sigma_{capture} $ by employing the$ \ell $ -summed Wong model for the decay of the 248No* nucleus. The DCM-derived cross sections processes such as of ff, QF and FF and capture cross section, demonstrate excellent concordance with the experimentally obtained data at all energy levels. Also, we have obtained the 2n channel evaporation cross-sections for 248No* nuclei. The aforementioned observed cross sections are determined by the optimization of the$ \Delta R $ . Accounting for the contribution of$ \Delta R $ in the decay process is crucial because it leads to shape elongation in the compound system, resulting in the development of a neck between the nascent fragments. The presence of a neck region in the dinuclear system allows for a free movement of nucleons between the nuclei. This creates an opportunity for significant exit channels by altering the interaction barrier [85, 86]. The flow of mass drift and the adjustment of the barrier are governed by the neck length$ \Delta R $ . According to Fig. 4, there is a clear correlation between an increase in the$ \Delta R $ with an increased$ E_{c.m.} $ . Further, as a result of its lower barrier characteristics, the extended GSkI force necessitates a greater$ \Delta R $ value, but it remains within the maximum allowable value. The$ \Delta R $ may give an idea about the temporal scale of the fragments reaction time. The reaction time will be faster when the value of$ \Delta R $ is higher. As the QF process takes place faster than the ff and FF, hence$ \Delta R $ is slightly higher for QF than the ff and FF.Figure 4. (color online) Neck length parameter
$ \Delta R $ (fm) in context to the centre of mass energy$ E_{c.m.} $ (MeV) optimized for fusion-fission (ff), quasi fission (QF) and fast fission (FF) using the GSkI Skyrme force.Further, the study examines the impact of different entrance channel mass asymmetry on the synthesis of 250No* nucleus. This is done by considering two different incoming channels: 44Ca + 206Pb and 64Ni + 186W, at different
$ E_{c.m.} $ = 187.04 MeV and 231.38 MeV. The comparison of the fragmentation potential V (in MeV) is presented with fragment mass, shown in Fig. 5. The fragmentation potential shows a roughly identical variation for both entrance channels, with a slightly greater magnitude seen for the 44Ca + 206Pb case compared to the 64Ni + 186W case. The deformation effect shows asymmetric nature of the fragmentation potential for both the considered entrance channels in the analysis. According to the calculations based on DCM, the fragmentation characteristics of ER, IMF, heavy HMF and fission fragments are found to be nearly identical. This means that the choice of entrance channel does not have any significant impact on the fragmentation behavior. Furthermore, the minima in the fragmentation potential for both entrance channels exhibit a similar pattern. The results are elucidated in relation to the relative preformation probability$ P_0 $ . Fig. 6 illustrates the computed preformation probability for the decay of 250No* at different$ \ell_{max} $ values and their corresponding$ E_{c.m.'s} $ . One can observe that preformation probability show slight variation in magnitude for the different entrance channels whereas the structure remains almost similar and even overlaps each other in fission region irrespective of the choice of entrance channels. Additionally, both cases demonstrate almost symmetrical fission peaks, and the contributing fission fragments remains same. Also, the fragments with maximum probability to be preformed i.e 122Sn and its complementary fragment 128Te are close to Z = 50 and N = 82 magic shell closure. Table 2, gives the information related to the various decay modes and their corresponding cross-sections,$ \ell_{max} $ values, neck length parameter for both the incoming channels using the GSkI force parameters. Table 2, clearly demonstrates that the$ \ell_{max} $ values and the$ \Delta R $ are comparable for both incoming channels. This suggests that the decay of 250No* is not influenced by the entrance channel effect. The calculations demonstrate the extent of the contribution of the CN (ff) process is higher in the case of the 44Ca + 206Pb reaction, while the nCN (QF) process appears to compete with ff in the 64Ni + 186W reaction channel. Ultimately, the investigation of the decay of 250No* resulting from the collision of 44Ca and 64Ni beams with 206Pb and 186W targets, respectively, was carried out using the DCM framework, taking into account the impact of deformation. From the findings, it can be concluded that the decay process remains unaffected by the specific approach of formation or the range of excitation energy.Figure 5. (color online) Variation of Fragmentation potential V
$ (A_2) $ for the parent nucleus 250No* formed in 44Ca + 206Pb and 64Ni + 186W reaction channels at$ \ell_{max} $ values and best fitted values of neck length parameter$ \Delta R $ .Figure 6. (color online) Same as Fig.5 but for Preformation probability
$ P_0 $ varies with fragment mass$ A_i $ (i=1,2).Reaction $ E_{c.m.} $
(MeV)$ E^*_{CN} $
(MeV)T
(MeV)$ \ell_{max} $ $ (\hbar) $ $ \Delta R_{ff} $
(fm)$ \sigma_{ff} $
(mb)$ \Delta R_{QF} $
(fm)$ \sigma_{QF} $
(mb)$ \sigma_{capt.} $
(mb)44Ca + 206Pb 187.04 38.69 1.19 85 2.19 109.34 2.33 27.23 140 64Ni + 186W 231.38 40 1.21 86 2.10 22.54 2.41 66.99 89.7 Table 2. The DCM-calculated fusion fission
$ \sigma_{ff} $ , quasi fission$ \sigma_{QF} $ , fast fission$ \sigma_{FF} $ cross sections and capture$ \sigma_{capt.} $ cross section calculated using$ \ell $ -summed Wong Model in the decay of$ ^{250}No^* $ nucleus formed formed in 44Ca + 206Pb and 64Ni + 186W reaction channels at different centre of mass energies$ E_{c.m.} $ with the best fitted neck length$ \Delta R $ , Temperatures T and$ \ell_{max} $ values. -
This subsection focus on the lifetimes in reference to ff and QF. Fission is a dynamic phenomenon where the nucleus undergoes deformation until it reaches a point of scission. An induced fission process has time scale of greatest significance, both theoretically and experimentally. Understanding the lifetime of this process is crucial for comprehending the nuclear reaction process. The overall duration of a fission process can be conceptually separated into two primary components: the time required for the nucleus to cross the saddle point, and the time it takes for the nucleus to deform from the saddle point to the scission point. While the QF barrier depends upon
$ Z_1Z_2 $ product, which in turn influence its lifetime. Hence, the available time may not be adequate for the conversion into a compound nucleus, resulting in the occurrence of the QF process. Hence, the duration of a partially equilibrated nuclear complex ought to be briefer compared to that of a fully equilibrated compound nuclear channel. The fission rate and fission lifetime for the asymmetric reaction such as 40Ca + 208Pb, 44Ca + 206Pb and symmetric 64Ni + 186W reactions are calculated which leads to the formation of 248No* and 250No* of Z=102 nucleus.Table 3, shows the comparison of ff and QF lifetime$ \tau_{ff} $ /$ \tau_{QF} $ using the excitation energy$ E^*_{CN} $ within the DCM and DNS approaches [73, 74]. The DCM and DNS approaches uses different parameters to calculate the lifetime thus leading to the difference in magnitude. In DCM the lifetime depends on three major factors i.e. Preformation probability$ P_0 $ , penetrability P and barrier assault frequency$ \nu_0 $ whereas in DNS approach is greatly influenced by the charge number of the projectile and target nuclei, beam energy etc. One can observe from the Table 3 that, the calculations carried out for the DNS cases are in agreement with the trend that lifetime goes on decreasing with the increase in the excitation energy$ E^*_{CN} $ whereas in DCM analysis the lifetime remains almost constant. There is a noticeable trend that ff and QF lifetime$ \tau_{ff} $ /$ \tau_{QF} $ decreases with increase in the$ E^*_{CN} $ . Therefore, the stability of a massive compound nucleus decreases as its excitation energy increases, primarily because the fission barrier is reduced. On comparing the lifetimes obtained using the DCM and DNS approaches one may differ of few magnitude is observed for the ff channel, whereas the quasi fission lifetimes are almost similar for both the approaches. Thus, The chance for survival of the large compound nucleus diminishes as the fission barrier falls with increasing$ E^*_{CN} $ of the resulting compound system.Reaction $ E_{c.m.} $
(MeV)$ E^*_{CN} $
(MeV)T
(MeV)$ \tau_{ff} $ (DCM)
(sec−1)$ \tau_{ff} $ (DNS)
(sec−1)$ \tau_{QF} $ (DCM)
(sec−1)$ \tau_{QF} $ (DNS)
(sec−1)40Ca + 208Pb 187.03 49 1.35 1.64 × 10−16 4.98 × 10−19 3.42 × 10−21 5.41 × 10−21 209.67 73 1.64 6.91 × 10−16 2.39 × 10−19 3.68 × 10−21 3.12 × 10−21 238.19 101 1.93 1.52 × 10−15 1.45 × 10−19 3.93 × 10−21 2.14 × 10−21 44Ca + 206Pb 187.04 38.69 1.19 5.49 × 10−17 1.50 × 10−18 3.01 × 10−21 1.46 × 10−20 64Ni + 186W 231.38 40 1.21 4.85 × 10−17 1.18 × 10−18 1.18 × 10−21 1.41 × 10−21 Table 3. Comparison of Fusion-fission lifetime
$ \tau_{ff} $ and Quasi fission lifetime$ \tau_{QF} $ for different reactions which are used for the formation of 248No* and 250No* at different excitation energies within the DCM and DNS approach.
Investigation of decay mechanisms and associated aspects of exotic Nobelium isotopes using the Skyrme energy density formalism
- Received Date: 2024-01-30
- Available Online: 2024-11-01
Abstract: Background: The search of the heavier elements has yielded many surprises and enhanced our knowledge in the direction of nuclear synthesis and associated dynamical aspects. Although new elements and their associated isotopes have been synthesized, the amount of information with the Z ≥ 102, remains somewhat scarce. Further, in the domain of transfermium elements, the nuclear shell structure is of significant relevance for ensuring nuclear stability. Hence, the shell effects become indispensable for such nuclei. Purpose: Persistent experimental and theoretical endeavors have been conducted to examine the reactions induced by heavy ions and the subsequent decay mechanisms in the realm of superheavy mass. In addition, the region of transfermium elements is itself of great interest because of the neutron / proton shell effects. Here, Our objective is to analyze the subsequent decay mechanisms of nuclides of Z = 102 nucleus, i.e. 248No* and 250No*. Methods: An extensive study is conducted using the dynamical cluster-decay model (DCM) based on Quantum Mechanical Fragmentation Theory (QMFT). The focus is on investigating compound nucleus (CN) and non-compound nucleus (nCN) mechanisms, including fusion-fission (ff), Quasi fission (QF), and fast fission (FF). The specific isotopes of interest are 248No* and 250No*, with attention given to the role of centre of mass energy