477
Views
1
CrossRef citations to date
0
Altmetric
Research Article

The folding potential description of 9C + 208Pb elastic scattering at 227 MeV

ORCID Icon
Article: 2175577 | Received 10 Sep 2022, Accepted 29 Jan 2023, Published online: 21 Feb 2023

Abstract

The sensitivity of the 9C + 208Pb elastic scattering reaction at Elab = 227 MeV (three times the barrier energy) towards the yielded potential combinations (9C density distributions + effective nucleon–nucleon interactions), bringing about a decent portrayal of the data, was estimated all through the optical double folding model. The merits and demerits of these combinations with regard to experimental data were handled, other than the comparability of results to those acquired by others. For the well-being of culmination, the S-matrix elements modulus, notch perturbation test and alteration of potential parameters are all used to highlight and reiterate some noteworthy aspects of the examined scattering reaction.

1. Introduction

Other than 3He, 9C has the most proton-rich nucleus known to be particle bound, with a proton-to-neutron ratio of Z/N = 2. 9C's ground state decays to 9B through beta decay with a half-life of 126.5 ms. It is a member of the T = 3/2, A = 9 isobaric quartet together with 9Li,9Be and 9B (T = 3/2). It is thought to have two decay channels: one proton (1p) and two protons (2p). Because the 1p separation energy from 9C (9C → 8B + p) is 1.29 MeV and from 8B (8B → 7Be + 2p) being just 0.136 MeV, the 2p channel is all but favoured. Accordingly, recent research has zeroed in on experimental and theoretical studies of its exotic nature [Citation1–7] as it is particularly significant in astrophysics [Citation8–10]. In relation to this, we give a speedy synopsis of the main ones.

The removal reaction cross section of either 1p (σ1p) or 2p (σ2p) for 8B and 9C is estimated by Blank et al. [Citation1], on a set of targets with different mass (A = 12-208) at the energy of 285 MeV/n. Strong evidence supports a 2p configuration for 9C given the higher values of deliberated σ2p compared to σ1p over two times, which are somewhat dependent upon the various target nuclei masses. Enders et al. [Citation2] and Warner et al.[Citation3] assessed the σ2p / σ1p ratio and viewed it as deliberately understanding with Blank et al. [Citation1] results.

Furthermore, thorough analyses of the measured and computed scattering reaction (absorption) cross section (σR) and the break-up reaction cross section for the angular distributions of our investigated system 9C + 208Pb and others were provided [Citation11–15] at various energies. Merely, the only available measurements and analysis of the elastic scattering and break-up processes were done at high incident energies for the 9C + 208Pb system. It is worth noting that breakup reactions are not within the scope of this investigation.

Assessing the elastic scattering data for the reactions induced by proton-rich nuclei might reveal some insight into their unusual structures and reaction mechanisms. In this instance, many models may be used to investigate how sensitive the measured elastic scattering data and cross section are to various factors, particularly the nuclear target-projectile potential elements (projectile/target density, effective nucleon–nucleon (NN) interaction) and breakup effects.

For example, utilizing the elastic scattering angular distribution of the 9C + p reaction at 277–290 MeV, Matsuda et al. [Citation16] determined the phenomenological two-parameters Fermi (2pF) density distributions and root-mean-square radius of 9C nuclear matter by using the relativistic folding model formulated by Murdock and Horowitz [Citation17]. This investigation was unable to provide a trustworthy density distribution in 9C or a satisfactory prediction of the reported cross sections. Rafi et al. [Citation18] then reanalysed the same scattered data using the Argonnev-18 NN interaction [Citation19] with the relativistic mean field (RMF) density in the Brueckner–Hartree–Fock model [Citation20]. The derived potential, according to the authors, was able to offer a good level of agreement with experimental results. Additionally, it was necessary to corroborate their findings using experimental data collected over a considerably larger area. Within a novel analytic approach, Bonacorsso et al. [Citation21] evaluated the sensitivity of σR for both 9C and its mirror nucleus 9Be among a series of microscopic densities; Variational Monte Carlo (VMC) [Citation22,Citation23], antisymmetrized molecular dynamics (AMD) [Citation5], relativistic Hatree–Fock (HF) [Citation24] and cluster model (CM) [Citation25] with Jeukenne, Lejeune and Mahaux (JLM) NN interaction [Citation26]. Based on the computed potentials from the optical glauber model, none of those distributions could represent the low energy increase of the σR, except when the surface term correction was included.

Further measurement on 9C,8B,7Be + 208Pb elastically scattering at 227,178 and 130 MeV, in that sequence, were reported by Yang et al. [Citation11]. As expected for this incident energy, very small break-up effects were observed. Therefore, the calculated break-up cross sections by using the continuum discrete coupled channel (CDCC) method are viewed as sensitive to the structure of proton-rich nuclei at high incident energies rather than the elastic scattering one as checked by the authors. Hence, for the two 9C cluster condurations underlined above, investigations of the measured elastic scattering cross sections on 208Pb are in like manner practically same.

These in-depth measurements and analyses paved the way for more research including proton/neutron-rich unstable nuclei and stable nuclei as well.

Yong Li et al. [Citation12] put upon a systematic global optical model potential (OMP) of 12C and 9Be to explore the elastic scattering data and the σR of 9–11,13,14C projectiles. The results show that this global potential adequately describes the elastic scattering for the triggered reactions by 9–11,13C, with disagreement at high angles in the case of 14C due to break-up effects. What is more, they claimed that this discrepancy could be fine-tuned by varying the potential parameters. Rong et al.[Citation13] examined the responsivity of the Sao-Paulo (SP) [Citation27,Citation28] and Akyüz Winther (AW) [Citation29,Citation30] NN potential along with phenomenological density (2pF) form in 17F observed elastic scattering on 208Pb at 94.5 MeV (just above the barrier) within the double folded potential (DFP). It is discovered that the extracted potentials exhibit similar behaviour. A fair description of the experimental data was provided by the obtained potentials up to a scattering angle 100o, however, for large angles, the data were under-predicted. After that, fitting quality enhancement in the intermediate angles was achieved by applying two coupled channel methods. As they stated, the 17F break up has a negligible influence on the elastic scattering and is independent of target mass. They also summarized other publications on the issue. In the optical DFP calculations based on the density-independent Michigan three-Yukawa (M3Y) NN interaction [Citation31], El-Hammamy et al. [Citation32] examined three density forms, namely Gaussian–Oscillator (GO), Gaussian–Gaussian (GG) and Gaussian (G) [Citation33–36] for the 17F nucleus scattered by 208Pb at 90.4 MeV (just below the barrier). Once again, by varying the 17F density forms, the resultant optical DFPs showed almost the same behaviour. However, in their research, the authors attempted to provide a plausible debate of the chosen combination (GO + M3Y). Moreover, the behaviour of the extracted σR values is found to be much more similar in comparison with other investigated systems 17O,16O + 208Pb than 19F + 208Pb at nearly the same energies. This proved the weak break-up effect at the energy range around the coulomb barrier.

Anwar et al. [Citation37] examined semi-phenomenological (SP) [Citation38,Citation39], HF, GO and CM [Citation40] density distributions for 8B nucleus scattered from 58Ni at 20.7–29.3 MeV enhanced by SPP and CDM3Y6 NN interactions [Citation41–43] within the optical DFP. The ability of their established potentials, especially SPP with SP/HF for halo 8B nucleus, is found to reproduce successfully the experimental data. Kassem et al. [Citation44] investigated the best 11Li density distribution forms for SP, cluster-orbital shell model approximation (COSMA) [Citation45], and HF in conjunction with several NN interactions (DDM3Y [Citation46], CDM3Y6-RT [Citation47] and SPP) via the optical DFP for 11Li scattered from 12C and 28Si at 29, 50 and 60 MeV/n. The major conclusion of this work is that the considered densities and NN interactions predict the best match with experimental data. Similar studies are presented in Refs. [Citation48–55] and authors debated the features of their assorted ingredients of the theoretical models that were used.

Regarding all of the prior research, it is still not clear whether single Yukawa (S1Y) [Citation56] or Knyzakov and Hefter (KH) [Citation57] effective NN interactions with a straightforward density forms for proton-rich nuclei (projectile/target) can be used to produce nuclear potentials in the same way they can for other types of nuclei.

For case, Satchler [Citation56] showed the success of his suggested NN interaction S1Y in analysing 36 sets of heavy ion (HI) elastic scattering at intermediate energies by DFP. Despite his limitations for light HI scattering [Citation56], Farid and Hassanian tested the validity of using four NN interactions, known as S1Y, M3Y, KH (3 terms) and JLM in DFP of 6,7Li [Citation58,Citation59] projectiles at intermediate energies. They concluded that successful predictions of the data by different NN interactions occurred in the following order: M3Y and S1Y, KH, JLM, with only one chosen published density form for each projectile. In Ref. [Citation60], the KH NN interaction was also used in the potential of several projectiles. Esmael and Allam [Citation61] analysed the elastic scattering of proton from 16O target at 135 and 200 MeV utilizing a potential built with three forms of KH NN interaction and three densities for 16O based on optical single folded potential. The analysis confirmed that the potential sets obtained with a single G term (KH1) NN potential give the best description of the data. Anwar [Citation53] reanalysed the elastic scattering cross sections of 6Li from various targets at 240 MeV with S1Y and M3Y and three density forms named 2pF, GO and G density distributions for 6Li nucleus. The derived potential combination (2pF density + S1Y) predicts an excellent agreement of the data over all measured angular ranges in DF calculations.

Hence, stretching out the assessment of these NN interactions (S1Y and KH) in contrast to the most popular and successful ones (M3Y and CDM3Y6) to the instance of 9C isotope seems epochal. Along these lines, the ongoing study centres around gauging the sensitivity of the recently published elastic scattering angular distributions of 9C + 208Pb reaction at Elab = 227 MeV [Citation11] above the coulomb barrier to the constructed potentials by using numerous combinations of 9C density distributions and effective NN interactions within the optical DFP which produces a successful description of the data. The outcomes were compared to those obtained for thoroughness. The corresponding techniques for these calculations; S-matrix elements modulus, notch perturbation test, and alteration of potential parameters are also employed to emphasize and restate some remarkable aspects of the examined scattering reaction. So, it should be mentioned that this work supplements our earlier research for different nuclear systems [Citation32,Citation62,Citation63].

This piece of work is broken into four parts. This section 1 started with a concise introduction, followed by the formalism of Section 2. Section 3 renders the results and discussions, while Section 4 provides the conclusion.

2. Formalism

The nuclear potential of the standard OM is separated into two portions for theoretical calculations: the real V and imaginary W potentials, which are specified as (1) U(R)=V(R)+iW(R).(1) The following phenomenological Woods–Saxon (WS) form was proposed for both real and imaginary components, (2) X(R)=X01+exp((RRX)/aX)and(X=V,W),(2) where Xo, aX and RX=rX(Ap1/3+AT1/3) are potential depth, diffuseness and radius with AP and AT mass numbers of projectile and target nuclei, individually. Furthermore, the Coulomb potential UC(R) for uniformly charge distributions of the colliding nuclei with the radius RC=1.3(Ap1/3+AT1/3) is concluded in Equation (1) [Citation64].

It is well known that the phenomenological representation does exclude the description of the structure of a projectile or target. The present elastic-scattering data were analysed using the program HIOPTIM-94 code [Citation65], in which the phenomenological real potential was supplanted by the DFP. As regards the nucleus–nucleus interaction, we adopt the same DFP used in Refs. [Citation31,Citation66] (3) VF(R)=Nρ1(r1)ρ2(r2)vNN(s)dr1dr2,(s=R+r2r1)(3) where ρ1 and ρ2 are nucleon densities of the two colliding nuclei, R signifies the distance between the nuclei’s centres of mass, (s) is the relative vector between the interacting nucleon pair, N is renormalization factor and vNN (s) stands for the effective NN interaction used in the calculations. Several vNN (s) variants can be implemented for ascertaining the DFPs via the DFPOT code [Citation67]. In this piece of work, four density-independent variants and one density-dependent variant are explored.

There are two variants with a single term, either in Gaussian or Yukawa form. The first is the S1Y interaction, which is indicated by [Citation56] (4) vNN(s)=60exp(1.428s)1.428s(10.0005EAP),(4) with a depth V0 of ≈60 MeV, where E and AP are the incident energy and the projectile mass number, respectively. KH-1 interaction is employed for comparative purposes as Knyzakov and Hefter [Citation57]. This is chosen as a second option, and its configuration is as per the following: (5) vNN(s)=20.97exp(0.68s)2.(5)

The other two variants are comprised of two terms plus Zero range exchange term (3 terms). The third interaction, KH-3 [Citation57] is expressed as (6) vNN(s)=601.99exp(1.25s)2+2256.4exp(2s)2276[10.005E/AP]δ(s)(6) Then again, the most common three Yukawa terms M3Y-Reid effective NN interaction [Citation66] is formed as (7) vNN(s)=7999exp(4s)4s2134exp(2.5s)2.5s276[10.005E/AP]δ(s)(7)

The last variant consists of two direct terms plus three finite range exchange terms. The fourth interaction, known as CDM3Y6, depends on the M3Y-Paris density-dependent interactions and is described as [Citation42,Citation43] (8) vNN(s)=F(ρ,E)[(11602exp(4s)4s2538exp(2.5s)2.5s)+(1524exp(4s)4s518.8exp(2.5s)2.5s7.847exp(0.707s)0.707s)](8) F(ρ,E)=[10.003(E/Ap][0.2658(1+3.8033exp(1.41ρ)4ρ)]

Because the density distribution of colliding nuclei is basic in folding calculations (see Equation 3), three alternative matter density distributions for the exotic nucleus 9C ground state were employed, considering various models of the internal nuclear structure.

The first is phenomenological Gupta1-two parameters Fermi (G1-2pF) [Citation68,Citation69] (9) ρ(r)=0.237(1+exp((r1.787)0.436))1,(9) with 9C rms radius r21/2=2.131 fm, which varies from the value in Ref. [Citation16].

In the second one, we look at the (7Be + 2p) cluster structure of the 9C nucleus [Citation11]. In this case, we treat 9C density as follows: (10) ρ9C(r)=ρ7Be(r)+ρ2P(r).(10) The core (7Be) and halo (2p) are presented with different spatial distributions in some phenomenological different forms. The density of the core nucleons is selected to be Gaussian (G), whereas the density of the halo nucleons is characterized by Harmonic Oscillator (HO) [Citation33–35]. The accompanying equations show the density distribution shapes that were used: (11) ρ7Be(r)=Cexp(r2α7Be2),(11) and (12) ρp(r)=23C(r2αp2)exp(r2αp2),(12) where α7Be2=2R7Be23,αp2=2Rp25,C=Njρjandρj=(1παj2)3/2withj=7Be,p.where R and Nj are the r.m.s radii and number of the core or halo nucleons, in that order, depending on the density used form. The r.m.s radii for 7Be and 2p equal 2.23 fm [Citation70] and 3.542 fm, separately, from this density. For this situation, the resultant Gaussian Oscillator (GO) density has a radius of 2.58 fm (9C r.m.s) [Citation5].

Finally, the proton and neutron densities of 9C nucleus have been obtained in a microscopic way by the RMF computations as the third density distribution form. Proton and neutron densities are acquired from Ref. [Citation18] and the derived 9C r.m.s radius = 2.522 fm.

Figure depicts the three particular densities of the 9C nucleus. The (2pF) form is employed to achieve the desired target 208Pb density [Citation71] (13) ρ(r)=ρ1+exp(r6.6210.551),(13) and the deduced r.m.s radius is 5.52 fm.

Figure 1. Panels (a) and (b) show the nuclear matter density distributions of 9C as logarithmic and linear scales, respectively.

Figure 1. Panels (a) and (b) show the nuclear matter density distributions of 9C as logarithmic and linear scales, respectively.

The elastic-scattering differential cross sections shaped by the five effective interactions S1Y, KH1, KH3, M3Y and CDM3Y6 with three density distributions not entirely set in stone, as shown in Figures and compared to the experimental results in Figures ,. A computerized search is performed to maximize the fits to data by diminishing the prescribed χ2 (14) χ2=1Ni=1N[σth(θi)σexp(θi)Δσexp(θi)]2,(14) where σth(θi) and σexp(θi) are the theoretical and experimental cross sections at angle (θi) respectively, Δσexp(θi) is the experimental error, and N is the difference number of data points and fit parameters. The χ2 values are calculated using a 10% statistical error for all analysed data.

Figure 2. The variation of real folded potential depths by using different density forms of different NN interactions for 9C + 208Pb system at 227 MeV.

Figure 2. The variation of real folded potential depths by using different density forms of different NN interactions for 9C + 208Pb system at 227 MeV.

Figure 3. The real folded potentials by using RMF density form and different NN interactions for 9C + 208Pb system at 227 MeV.

Figure 3. The real folded potentials by using RMF density form and different NN interactions for 9C + 208Pb system at 227 MeV.

Figure 4. The normalized real folded potentials by using different-NN interactions form and different 9C densities for 9C + 208Pb system at 227 MeV in FM-1 scenario.

Figure 4. The normalized real folded potentials by using different-NN interactions form and different 9C densities for 9C + 208Pb system at 227 MeV in FM-1 scenario.

Figure 5. The elastic-scattering differential cross sections in comparison with the experimental data of the 9C+ 208Pb reaction at 227 MeV [Citation11] by using different combination of densities and NN interactions within DFP of scenario FM-1.Labels indicated on the figures.

Figure 5. The elastic-scattering differential cross sections in comparison with the experimental data of the 9C+ 208Pb reaction at 227 MeV [Citation11] by using different combination of densities and NN interactions within DFP of scenario FM-1.Labels indicated on the figures.

Figure 6. The elastic-scattering differential cross sections in comparison with the experimental data of the 9C+ 208Pb reaction at 227 MeV [Citation11] by using different combination of densities and NN interactions within DFP of scenario FM-2. Labels indicated on the figures

Figure 6. The elastic-scattering differential cross sections in comparison with the experimental data of the 9C+ 208Pb reaction at 227 MeV [Citation11] by using different combination of densities and NN interactions within DFP of scenario FM-2. Labels indicated on the figures

The real renormalization factor (NR), as well as the three imaginary WS parameters, are looked for in the first scenario of the double folding model (FM-1). Notwithstanding, there are just two parameters for the second scenario (FM-2): the real (NR) and imaginary (NI) renormalization factors, with the real and imaginary potential parts having the same folded shape. Table shows the calculated parameters, the volume integrals per pair of interacting nucleons for the real (JR) and imaginary (JI) parts of the potential, the reaction cross-section (σR), and the best-fit χ2 for every scenario. In the succeeding segments, a thorough discussion is given to exist results.

Table 1. Best-fit parameters of the double folded potentials (DFPs) using the two scenarios; FM-1 and FM-2 for 9C + 208Pb scattering data at 227 MeV.

3. Results and discussions

3.1. Comparability of density parameterizations and real nuclear potentials

The two elements that comprise the optical DFP of Equation (3) have been examined in the field of the elastic scattering angular distributions of 9C + 208Pb reaction at energy 227 MeV.

To begin, Figure conveys the radial dependency of the three selected density parameterizations of 9C, meant as G1-2pF, GO and RMF. The GO and RMF density distributions are near to one other and have a lower value in the nucleus's centre than the phenomenological G1-2pF, as displayed in Figure panel (a). Simply put, as the radius rises, they tumble to the most elevated. As well, it has been shown that the microscopic density RMF draws out astoundingly farther than the other densities, as seen in Figure panel (b). As an aftereffect, the r.m.s radius of 9C is lengthy from 2.13 to 2.58 fm.

Then, with the chosen densities of both the 9C projectile and the 208Pb target, Equation (3) was coordinated across S1Y, KH1, KH3, M3Y and CDM3Y6-NN interaction potentials. DFPOT code [Citation67] was used to generate real DFPs. Table and Figure show the noticed parameters related to the determined potentials.

Just a single figure, Figure is shown when the RMF density is in use, because the estimated DFPs with the other densities have about the same depths and ways of behaving with all NN interactions. At the point when the radial dependency of the five built DFPs is examined, it is discovered that S1Y is the shallowest of the five. Nonetheless, the extracted depths of M3Y and KH3 potentials are similar in value and fairly greater than that of KH1 potential. The CDM3Y6 potential is almost in the middle of the preceding potentials. Thus the KH3 potential generated results are eliminated from the analysis.

Additionally, the comparison of the normalized real DFPs for various NN interactions demonstrates the effect of the used densities. Hence, this is shown in Figure in the framework of FM-1 scenario, for instance. It is obvious from four panels (a–d) that the three densities have a slight effect on the S1Y potentials as shown in (c). For M3Y, KH1 and CDM3Y6, the situation is different, the effect of used densities rises in the radial distance region from 0 to 8 fm. These divergences are reflected on the calculated parameters in Table .

3.2. Fitting quality of elastic scattering folding models

At 227 MeV, the FM-1 is used to investigate elastic scattering cross-section data of 9C + 208Pb reaction. In this scenario, the real DFP of Equation (3) was created using the DFPOT code [Citation67] by consolidating the three density distributions of the 9C projectile, the five grouped designated NN interactions and 2pF density for the 208Pb target. In addition to the phenomenological WS imaginary potentials, the subsequent potentials were fed into the HIOPTIM-94 [Citation65]. The standout accord between theoretical calculations and experimental data is uncovered in Figure , and the resulting parameters are catalogued in Table . To execute the fitting method, just one free parameter NR for the real potential and three parameters for the imaginary potential were used.

The S1Y, KH1, M3Y and CDM3Y6 interaction potentials with G1-2pF, GO and RMF densities demonstrate the comparable way of behaving with each other with approximately identical χ2 values. Also, there is a minor underrating of the data at θ∼(12–15)o for those potentials that correspond to the phenomenological G1-2pF density. As found in Table , this is mirrored by the shallowest imaginary potential values. In any case, none of the computations could identify the experimental data in the forward angle θ∼(9–12)o.

For comparison, the resultant scattering cross section with data is clearly on par with those detailed in Ref. [Citation51] in the framework of density-independent M3Y-NN interaction coupled with RMF density for 10,11C + 208Pb reactions at 226 and 256 MeV, one by one. Similarly, the experimental data in forward angle θ ∼ (7–11) o couldn't be achieved in that analysis.

As far as density distribution sensitivity, any of the 9C may be compensated by the NR and imaginary parameters, regardless of the NN interaction type. It is observed that the strength of the real potential should be reduced by ∼12–28% (see Table ) to describe the experimental data. the extracted average NR is 0.88 ± 0.146, 0.87 ± 0.144, 0.94 ± 0.027 and 0.72 ± 0.125 from M3Y, KH1, S1Y and CDM3Y6 calculations. The best match, nonetheless, is obtained for the associated parameters of S1Y interaction potential added to either GO or RMF density. Rather than KH1 and M3Y interaction potentials, this is upheld by a minor deviation of NR values from unity and a shallow imaginary depth. However, it has almost the same imaginary depths as CDM3Y6 potential. The S1Y potential is marginally sensitive to various kinds of densities, with NRG12pF,NRGO,NRRMF values of 0.96, 0.95 and 0.91, respectively. As an outcome, the values of the real JR and imaginary JI volume integrals made by KH1and M3Y potentials are in magnificent agreement and greater than those for S1Y and CDM3Y6 potentials within the same density.

The same elastic scattering cross-section data has been reanalysed employing the FM-2 scenario to look at the strength of the model. In this scenario, the imaginary potential, as well as the real one, were obtained in a DFP type of Equation (2) with two separate renormalization factors. So, just two free parameters; NR and NI were uninhibitedly adjusted to get the greatest amicability between theoretical predictions and experimental data, as shown in Figure and Table . It is ascertained that the strength of the real potential should be reduced by ∼8–37% to describe the experimental data. the extracted average NR is 0.92 ± 0.247, 0.84 ± 0.225, 1.08 ± 0.116 and 0.63 ± 0.139 from M3Y, KH1, S1Y and CDM3Y6 calculations.

Applying this scenario, the experimental data in forward angle θ ∼ (7–11)o couldn't be described once again, as perceived in Figure , add-on to under estimation at θ > (17) o for all density and potential combinations with the exception of G1-2pF for S1Y and M3Y potentials, which misjudges experimental results at θ ∼ (12–17) o. likewise, this is same as indicated in Ref. [Citation11] by using JLM interaction potential and the σR value got close to those premeditated for 9C (3396 mb).

Similarly, the JR and JI volume integrals values made by KH1 and M3Y potentials are higher than those generated by S1Y and CDM3Y6 potentials within the same density, as written in Table . It displays also, a small disparity among between NR and NI values. Therefore, the highest accord with experimental data is demonstrated by either GO or RMF and S1Y combinations, as it is cleared in Figure and Table .

The processed results lay out a high sensitivity to the NR and W/NI, which is most likely due to the breakdown of the 9C projectile in the field of the heavy target nucleus, 208Pb. All of the σR values got are as per each other and close to those distributed in Ref. [Citation11] for 9C and [Citation51] for 10,11C + 208Pb reactions at 226 and 256 MeV, on an individual basis. As a result of this investigation, the FM-1 scenario associated with S1Y-NN interaction and GO or RMF density for 9C projectile is preferred.

Sadly, because of an absence of additional energies for the analysed reaction, the overall trend of depths or volume integrals changing with energies still up in the air.

3.3. S-matrix

The modulus of partial wave elastic scattering matrix (S-matrix) elements |SL| [Citation72–74] as a function of projectile-target orbital angular momentum (L) is introduced in Figure . This curve, otherwise called an absorption profile [Citation72], offers a measure of absorption degree and, as a result, the σR, the two of which are given by: (15) SL=e2iδL(15) and (16) σR=2πk2L(2L+1)(1|SL|2),(16) in terms of the nuclear reflection coefficient ηL=|SL| and nuclear phase shift δL, for complex optical potentials. Where, (1|SL|2)=TL is the nuclear transmission coefficient and k is the wave number. The magnitude |SL|is considered the most important quantity for describing elastic scattering data. It is powerfully related to L, where |SL|0 which means almost (complete absorption) at small L, increasing |SL| for intermediate L (partial absorption), whereas |SL|1 at large L, which explicit (no absorption). Then TL0is taken as complete transmission while TL1 is taken as complete reflection. This usual behaviour is shown in Figure and is tantamount to that of the 11Be + 208Pb reaction elastic scattering at 210 MeV, which is roughly 5.2 times the coulomb barrier as portrayed in Ref. [Citation14].

Figure 7. Modulus of the scattering matrix |SL| to the calculated real folded potential by using RMF density and S1Y-NN interaction within DFP of scenarios; FM-1 and FM-2 versus the orbital angular momentum (L).

Figure 7. Modulus of the scattering matrix |SL| to the calculated real folded potential by using RMF density and S1Y-NN interaction within DFP of scenarios; FM-1 and FM-2 versus the orbital angular momentum (L).

The FM-2 potential has the biggest cross section because the |SL| attains the unitary value at a more prominent L than the FM-1 potential. This is coincidence with what has been stated in the literature [Citation21,Citation75]. But then, the FM-2 potential produces a more keen |SL|, coming about in rather lower critical angular momentum (L1/2) and strong absorption radius (R1/2) at which both |SL|2andTL=12. The R1/2 is commonly described as the distance of the coulomb trajectory's closest approach for the partial waves L1/2. Since elastic scattering cross sections are generally impacted by the surface region, changes of |SL| around L1/2 (surface collisions) result in fairly differing elastic scattering angular distribution characteristics.

Precisely, these alterations are believed to be brought about by the imaginary δL activity in the surface region nearbyL1/2. This is supported by the way that the zero values of the |SL|at the low angular momentum part mirrors the complete strong absorption effects connected with the higher value of the reduced imaginary potential (W/V)R1/2 [Citation72] for FM-1 potentials than the FM-2 ones as seen in Figure and Table . Thereupon, this confirms the preference for FM-1 over FM-2 scenario in the preceding debate.

Table 2. Values of some characteristic quantities accompanied by the best-fit parameters of the selected bold case in Table .

3.4. Notch test

Notch perturbation calculations [Citation53,Citation76–78] were used to determine the radial area of the potential that is susceptible to scattering. This technique involves scanning a radial perturbation across the potential. The χ2(R)/χo2 value of the fit remains steady in radial regions where there is no sensitivity, whereas it deteriorates considerably in sensitive parts. χ2(R) and χo2 are the best-fit values obtained for the perturbed and unperturbed real potentials, respectively.

In this work, we use a notch with an amplitude of 1 fm and a width of 0.5 fm to instigate a small perturbation at radius R in the radial distribution, where R fluctuates regularly from 0 to 18 fm in 0.5 fm increments. The results of such a scan for the 9C + 208Pb reaction with the best-selected potential within the FM-1scenario are presented in Figure . The sensitive zone in 9C elastic scattering goes from about (R = 12–14 fm). The ratio W/V = 1.9 (in the nuclear surface) at the radius of greatest sensitivity for the real potential (RS=12 fm) near the empirical strong absorption radius (R1/2 = 11.45 fm) as bestowed in Table attests the domination of the absorptive potential. As well as, this severely restricts the quantity of information that can be obtained about the real one.

Figure 8. Radial sensitivity of elastic-scattering differential cross sections to the calculated real folded potential by using RMF density and S1Y-NN interaction within DFP of scenario FM-1.

Figure 8. Radial sensitivity of elastic-scattering differential cross sections to the calculated real folded potential by using RMF density and S1Y-NN interaction within DFP of scenario FM-1.

Accordingly, this surface localized absorption suggests that the reaction mechanism is controlled by direct reactions, which is predicted at energies over the barrier. What is more, the somewhat larger RS radius than R1/2 radius result is consistent with that reported in Ref. [Citation78] for various weakly bound systems scattered by heavy targets near coulomb energies. By and large, these results point out that 9C acts as a heavy ion projectile rather than a light one, which is in agreement with the systematics found in other weakly bound systems such as 9Be [Citation79].

3.5. Sensitivity of the real/imaginary potential

To get some insights into the values expected to describe the experimental data well above the coulomb barrier energy, we modify either the NR or the imaginary potential depth Wo within the FM-1 scenario associated with S1Y-NN interaction potential and projectile RMF density. The adopted procedure is analogous to those delineated in Refs. [Citation13,Citation80,Citation81]. Figures and show a comparison of the experimental data and the computed results using various values of either NR or Wo.

Figure 9. The elastic scattering calculated with different values of real folded potential renormalization factor (NR) by using RMF density and S1Y-NN interaction added to WS imaginary potential within the folding model (FM-1) compared with experimental data as linear and logarithmic scales.

Figure 9. The elastic scattering calculated with different values of real folded potential renormalization factor (NR) by using RMF density and S1Y-NN interaction added to WS imaginary potential within the folding model (FM-1) compared with experimental data as linear and logarithmic scales.

Figure 10. The elastic scattering was calculated with different values of imaginary potential depth (Wo) added to real folded potential by using RMF density and S1Y-NN interaction within folding model (FM-1) compared with experimental data as linear and logarithmic scales.

Figure 10. The elastic scattering was calculated with different values of imaginary potential depth (Wo) added to real folded potential by using RMF density and S1Y-NN interaction within folding model (FM-1) compared with experimental data as linear and logarithmic scales.

As should be conferred in Figure , a comparison of the calculations with the best fit NR  = 0.91 to those with NR = 0.5 and 0.25 supplemented with fixed best suitable imaginary parameters reveals that a slight reduction of NR to 0.91 than its default value ( = 1) is sufficient to reproduce the experimental data. By raising the reduction to 0.25, the theoretical curve goes slightly downward at θ ∼ (12–14)o and upward at θ > (16o) without changing its behaviour.

Perversely in Figure , a comparability of the results with the foremost acceptable imaginary potentials depth to those with Wo/2, Wo/10 and Wo≈ 0 further with fixed NR  = 0.91 are presented. A search on the radius (r) and diffuseness (a) parameters was performed to evaluate their sensitivity to the quality of the imaginary potential. The diffuseness parameter was found to be more sensitive than the radius parameter at Wo < Wo/2. This agrees with the end in Ref. [Citation80], merely with different behaviour ascribed to lower energies of their inspected system.

Furthermore, using Wo/10 and Wo≈ 0, oscillating curves are existed. Only, as Wo ≈ 0 broad oscillations emerge in the back angle cross sections. These formations are assigned to the manifestation of the real potential’s refractive power. The absorption necessary to match the data dampens these features in the produced angular distributions. This pattern was connected to the use of deep potentials rather than shallow potentials [Citation79,Citation82,Citation83].

Broadly speaking, this implies that a weak real part contrasted with its relating strong imaginary part in the surface region provokes reasonable agreement with the experimental data. Accordingly, no apparent impacts of associated phenomena (rainbow scattering), which is typical of well-bound heavy nuclei scattering can be detected [Citation11,Citation84]. Homologous investigations are conferred for various systems with miscellaneous conclusions, such as 6Li + 209Bi [Citation80], 17F + 208Pb [Citation13], 17F + 14N [Citation83] and 7li + 138Ba [Citation85].

4. Conclusion

Throughout this study, we have compared our results to the current elastic scattering cross sections of the 9C + 208Pb reaction at 227 MeV (3 times over the barrier energy) using two basic scenarios of the optical DFM (FM-1 and FM-2) of the elastic scattering process. Various DFPs in light of the combination of 9C density distribution (G1-2pF,GO and RMF), 208Pb (2pF) density distribution and the effective NN interaction (density-independent S1Y, KH1, KH3, M3Y and density-dependent CDM3Y6) were examined. Principally, this analysis glares the undermentioned four significant aspects:

  1. All resultant potentials give a similar outstanding prediction of the data except in forward angle θ ∼ (9–12)o. The strength of the real potential should be reduced by ∼12–28% and ∼8-37% within the framework of FM-1 and FM-2 scenarios, in that order, to reproduce the data.

  2. The ascertained scattering matrix elements |SL| from the data analysis agree with one another using the concerned scenarios. It verifies preferring of the FM-1 scenario over the other one and focuses on an absorptive and a peripheral nature of the investigated scattering reaction.

  3. The calculated notch test shows the radius of maximum sensitivity (RS  = 12 fm) which is likewise near the empirical strong absorption radius (R1/2 = 11.45 fm). This is joined by a higher reduced potential value (W/V)R1/2=1.9 than unity. In our opinion, the explanation may be owing to the weak binding of 9C nucleus.

  4. An examination of the favoured real deep potential bespeaks that there are no apparent impacts of associated phenomena (rainbow scattering). As per our assessment, there is sufficient absorption to match the experimental data and eliminate the recognized oscillations in the computed angular distributions.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Blank B, Marchand C, Pravlikoff MS, et al. Total interaction and proton-removal cross-section measurements for the proton-rich isotopes 7Be, 8B, and 9C. Nucl Phys A. 1997;624:242–256.
  • Enders J, Baumann T, Brown BA, et al. Spectroscopic factors measured in inclusive proton-knockout reactions on 8B and 9C at intermediate energies. Phys Rev C. 2003;67:064301.
  • Warner RE, Becchetti FD, Brown JA, et al. Phys Rev C. 2004;69:024612.
  • Fukui T, Ogata K, Minomo K, et al. Determination of the 8B (p,γ) 9C reaction rate from 9C breakup. Phys Rev C. 2012;86:022801.
  • Furutachi N, Kimura M, Doté A, et al. Structures of light halo nuclei. Prog Theor Phys. 2009;122:865–880.
  • Chilug AI, Panin V, Tudor D, et al. AIP conference proceedings. AIP Conf Proceedings. 2019;2076:060001.
  • Hooker JL. Ph.D Thesis,Texas A and M University. 2019.
  • Tanihata I, Hamagaki H, Hashimoto O, et al. Measurements of interaction cross sections and nuclear radii in the light p-shell region. Phys Rev Lett. 1985;55:2676.
  • Al-Khalili S. Lecture notes in physics. Lect Notes Phys. 2004;651:77–112.
  • Wiescher M, Görres J, Graff S, et al. The hot proton-proton chains in low-metallicity objects. Astrophys J. 1989;343:352.
  • Yang YY, Liu X, Pang DY, et al. Elastic scattering of the proton drip line nuclei 7Be 8B, and 9C on a lead target at energies around three times the Coulomb barriers. Phys Rev C. 2018;98:044608.
  • Xu Y-L, Han Y-L, Su X-W, et al. Description of elastic scattering induced by the unstable nuclei 9,10,11,13,14C *. Ch Phys C. 2021;45(11):114103.
  • Rong CH, Rangel J, Wu YS, et al. Study of quasi-elastic scattering of 17F+208Pb at energies around Coulomb barrier. Eur Phys J A. 2021;57:143.
  • Duan FF, Yang YY, Lei J, et al. Elastic scattering and breakup reactions of neutron-rich nucleus 11Be on 208Pb at 210 MeV. Phys Rev C. 2022;105:034602.
  • Heo K, Cheoun M, Choi K, et al. Coulomb breakup reaction of loosely bound 17F with dynamic polarization potentials. Phys Rev C. 2022;105:014601.
  • Matsuda Y, Sakaguchi H, Takeda H, et al. Elastic scattering of protons from 9C with a 290 MeV/nucleon 9C beam. Phys Rev C. 2013;87:034614.
  • Murdock DP, Horowitz CJ. Microscopic relativistic description of proton-nucleus scattering. Phys Rev C. 1987;35:1442–1462.
  • Rafi S, Bhagwat A, Haider W, et al. Phys Rev C. 2014;89:067601.
  • Wiringa RB, Stoks VGJ, Schiavilla R. Accurate nucleon-nucleon potential with charge-independence breaking. Phys Rev C. 1995;51:38–51.
  • Gambhir YK, Ring P, Thimet A. Relativistic mean field theory for finite nuclei. Ann Phys. 1990;198:132–179. and references there in.
  • Bonaccorso A, Carstoiu F, Charity RJ. Imaginary part of the 9C−9Be single-folded optical potential. Phys Rev C. 2016;94:034604.
  • Wiringa RB. http://www.phy.anl.gov/theory/research/density/.
  • Pieper SC, Wiringa RB. Quantum Monte Carlo calculations of light nuclei. Annu Rev Nucl Part Sci. 2001;51:53–90.
  • Typel S, Ropke G, Klahn T, et al. Composition and thermodynamics of nuclear matter with light clusters. Phys Rev C. 2010;81:015803.
  • Descouvement P. Microscopic study of proton-capture reactions on unstable nuclei. Nucl Phys A. 1999;646:261–273.
  • Jeukenne JP, Lejeune A, Mahaux C. Optical-model potential in finite nuclei from Reid’s hard core interaction. Phys Rev C. 1977;16:80–96.
  • Chamon LC, Pereira D, Hussein MS, et al. Nonlocal description of the nucleus-nucleus interaction. Phys Rev Lett. 1997;79:5218–5221.
  • Chamon LC, Carlson BV, Gasques LR, et al. Toward a global description of the nucleus-nucleus interaction. Phys Rev C. 2002;66:014610.
  • Broglia RA, Winther A. Heavy Ion reactions. Boulder: West view Press; 2004.
  • Akyüz O, Winther A. Nuclear structure and heavy Ion reactions. In: RA Broglia, CH Dasso, RA Ricci, editor. Proc. Enrico Fermi Int. summer school of physics, 1979. Amsterdam, North Holland; 1981:491.
  • Bertsch G, Borysowicz J, McManus H, et al. Interactions for inelastic scattering derived from realistic potentials. Nucl Phys A. 1977;284:399–419.
  • El-Hammamy MN, El-Nohy NA, El-Azab Farid M, et al. Systematic analysis of 17,19F and 16,17O elastic scattering on 208Pb just below the Coulomb barrier. Chinese J Phys. 2021;73:136–146.
  • Ilieva S. PhD thesis, Johannes Gutenberg-Universitat in Mainz. 2008).
  • Alkhazov GD, Dobrovolsky AV, Egelhof P, et al. Nuclear matter distributions in the 6He and 8He nuclei from differential cross sections for small-angle proton elastic scattering at intermediate energy. Nucl Phys A. 2002;712:269–299.
  • Bush MP, Al-Khalili JS, Tostevin JA, et al. Sensitivity of reaction cross sections to halo nucleus density distributions. Phys Rev C. 1996;53:3009–3013.
  • Alkhazov GD, Novikov IS, Shabelski YM. Nuclear radii of unstable nuclei. Int J Mod Phys E. 2011;20:583–627.
  • Anwar M, El-Naggar B, Behairy KO. Microscopic analysis of the 8B +58 Ni elastic scattering at energies from 20.7 to 29.3 MeV. J Phys Soc of Japan. 2022;91:014201.
  • Barioni A, Zamora JC, Guimarães V, et al. Elastic scattering and total reaction cross sections for the 8B, 7Be, and 6Li+12C systems. Phys Rev C. 2011;84:014603.
  • Bhagwat A, Gambhir YK, Patil SH. Nuclear densities in the neutron-halo region. Eur Phys J A. 2000;8:511–520.
  • Ibraheem AA, El-Azab Farid M, Al-Hajjaji AS. Analysis of 8B proton halo nucleus scattered from 12C and 58Ni at different energies. Braz J Phys. 2018;48:507–512.
  • Anantaraman N, Toki H, Bertsch GF. An effective interaction for inelastic scattering derived from the Paris potential. Nucl Phys A. 1983;398:269–278.
  • Gao-Long Z, Hao L, Xiao-Yun L. Nucleon–nucleon interactions in the double folding model for fusion reactions. Chin Phys B. 2009;18:136–141.
  • Khoa DT, Satchler GR, von Oertzen W. Nuclear incompressibility and density dependent NN interactions in the folding model for nucleus-nucleus potentials. Phys Rev C. 1997;56:954–969.
  • Behairy KO, El-Azab Farid M, et al. Analysis of strong refractive effect within11Li projectile structure. Chinese Physics C. 2021;45:024101.
  • Antonov AN, Gaidarov MK, Kadrev DN, et al. Charge density distributions And related form factors In neutron-rich light exotic nuclei. Int J Mod Phys E. 2004;13:759–772.
  • El-Azab Farid M, Satchler GR. A density-dependent interaction in the folding model for heavy-ion potentials. Nucl Phys A. 1985;438:525–535.
  • Khoa DT, Phuc NH, Loan DT, et al. Nuclear mean field and double-folding model of the nucleus-nucleus optical potential. Phys Rev C. 2016;94:034612.
  • Chen X, Lui Y-W, Clark HL, et al. Giant resonances in 24Mg and 28Si from 240 MeV 6Li scattering. Phys Rev C. 2009;80:014312.
  • Yang YY, Wang JS, Wang Q, et al. Quasi-elastic scattering of 10,11C and 10B from a natPb target. Phys Rev C. 2014;90:014606.
  • Chamon LC, Gasques LR. Reinterpreting the energy dependence of the optical potential. J Phys G. 2016;43:015107.
  • Aygun M. Analysis with relativistic mean-field density distribution of elastic scattering cross-sections of carbon isotopes (10–14,16C) by various target nuclei. Pramana J Phys. 2019;93.
  • Ibraheem AA, Al-Hajjaj AS, El-Azab Farid M. Elastic scattering of one-proton halo nucleus 17F on different mass targets using semi microscopic potentials. Rev Mex Fis. 2019;65:168–174.
  • Anwar M. Semimicroscopic analysis of 6Li elastic scattering at 40 MeV/nucleon. Phys Rev C. 2020;101:064617.
  • Wang K, Yang YY, Moro AM, et al. Elastic scattering and breakup reactions of the proton drip-line nucleus 8B on 208Pb at 238 MeV. Phys Rev C. 2021;103:024606.
  • Olorunfunmi SD, Bahini A. Reanalysis of 10B+120Sn elastic scattering cross section using São Paulo potential version 2 and Brazilian nuclear potential. Braz J of Phys. 2022;52:11.
  • Satchler GR. A simple effective interaction for peripheral heavy-ion collisions at intermediate energies. Nucl Phys A. 1994;579:241–255.
  • Knyarkov OH, Hefter EF. An analytical folding potential for deformed nuclei. Z Phys A. 1981;301:277–282.
  • El-Azab Farid M, Hassanain MA. Density-independent folding analysis of the Li elastic scattering at intermediate energies. Nucl Phys A. 2000;678:39–75.
  • El-Azab Farid M, Hassanain MA. Folding model analysis of 6,7Li elastic scattering at 12.5–53 MeV/u. Nucl Phys A. 2002;697:183–205.
  • Esmael EH, Abou Stait SAH, Zedan MEM, et al. Density-dependent effect on alpha -12C elastic scattering. J Phys G. 1991;17( ):1755–1768.
  • Esmael EH, Allam MA. Acta Phys Pol B. 2008;39.
  • El-Hammamy MN, Attia A. 16C-elastic scattering examined using several models at different energies. Pramana J Phys. 2018;90(5):66.
  • El-Nohy NA, El-Hammamy MN, El-Azab Farid M, et al. Int Astronomy and Astrophysics Research Journal. 2019;1(1):1.
  • Poling JE, Norbeck E, Carlson RR. Elastic scattering of lithium by 9Be, 10B, 12C, 13C, 16O, and 28Si from 4 to 63 MeV. Phys Rev C. 1976;13:648–660.
  • Clarke NM. private communication.
  • Satchler GR, Love WG. Folding model potentials from realistic interactions for heavy-ion scattering. Phys Rep. 1979;55:183–254.
  • Cook J. DFPOT - A program for the calculation of double folded potentials. Comput Phys Comm. 1982;25:125–139.
  • Gupta RK, Singh D, Greiner W. Semiclassical and microscopic calculations of the spin-orbit density part of the skyrme nucleus-nucleus interaction potential with temperature effects included. Phys Rev C. 2007;75:024603.
  • Ghodsi ON, Torabi F. Comparative study of fusion barriers using skyrme interactions and the energy density functional. Phys Rev C. 2015;92:064612.
  •  Dobrovolsky AV, Korolev GA, Inglessi AG, et al. Nuclear-matter distribution in the proton-rich nuclei 7Be and 8B from intermediate energy proton elastic scattering in inverse kinematics. Nucl Phys A. 2019;989:40–58.
  • Umemoto Y, Hirenzaki S, Kume K, et al. Isotope dependence of deeply bound pionic states in Sn and Pb. Phys Rev C. 2000;62(2):024606.
  • Brandan ME, Satchler GR. The interaction between light heavy-ions and what it tells us. Phys Rep. 1997;258:143–243.
  • Perez-Torres R, Belyaeva TL, Aguilera EF. Fusion and elastic-scattering cross-section analysis of the 12C + 12C system at low energies. Phys At Nucl. 2006;69(8):1372–1382.
  • Korda VY, Molev AS, Klepikov VF, et al. Unified model-independent S-matrix description of nuclear rainbow, prerainbow, and anomalous large-angle scattering in 4He− 40Ca elastic scattering. Phys Rev C. 2015;91:024619.
  • Alharbi WR, Ibraheem AA, El-Azab Farid M. Theoretical investigation of 6,8He halo nuclei by using microscopic optical potentials. J Korean Phys Soc. 2013;63(5):965–969.
  • Rusek K, Moroz Z. Čaplar R. Egelhof P. Möbius K.-H.. Spin-orbit potentials for elastic scattering of polarized 6Li ions from 12C and 58Ni. Nucl Phys A. 1983;407:208–220.
  • El-Azab Farid M, Ibraheem AA, Al-Hajjaji AS. Investigation of 17F+p elastic scattering at near-barrier energies. Eur Phys J A. 2015;51:134.
  • Roubos D, Pakou A, Alamanos N, et al. Radial sensitivity of elastic scattering at near barrier energies for weakly bound and tightly bound nuclei. Phys Rev C. 2006;73(R):051603.
  • Zisman MS, Cramer JG, Goldberg DA, et al. Dominance of strong absorption in 9Be+28Si elastic scattering. Phys Rev C. 1980;21(6):2398–2416.
  • Santra S, Kailas S, Ramachandran K, et al. Reaction mechanisms involving weakly bound 6Li and 209Bi at energies near the Coulomb barrier. Phys Rev C. 2011;83:034616.
  • Alvarez MA, Fernández-García J. P. León-García J. L. Rodríguez-Gallardo M. Gasques L. R.. Systematic study of optical potential strengths in reactions on 120Sn involving strongly bound, weakly bound, and exotic nuclei. Phys Rev C. 2019;100:064602.
  • Hu L, Song Y, Hou Y, et al. The refractive scattering of 17F+12C. EPJ Web of Conferences. 2020;239:03010.
  • Blackmon JC, Carstoiu F, Trache L, et al. Elastic scattering of the proton drip-line nucleus F17. Phys Rev C. 2005;72:034606.
  • Frahn WE. Diffraction scattering of charged particles. Ann Phys. 1972;72:524–547.
  • Chen W-D, Guo H-R, Sun W-L, et al. Microscopic study of 7Li-nucleus potential *. Chin Phys C. 2020;44(5):054109.