SciELO - Scientific Electronic Library Online

 
vol.68 issue3A comprehensive analysis of 19F + 9Be, 12C, 16O, 19F, 27Al, 28,30Si, 40Ca, 54,56Fe, 208Pb, 232Th fusion reactionsFirst-principles investigations of electronic and optical properties of Er-doped GaN involved in ErGaN/ErN quantum well heterostructures author indexsubject indexsearch form
Home Pagealphabetic serial listing  

Services on Demand

Journal

Article

Indicators

Related links

  • Have no similar articlesSimilars in SciELO

Share


Revista mexicana de física

Print version ISSN 0035-001X

Rev. mex. fis. vol.68 n.3 México May./Jun. 2022  Epub Apr 14, 2023

https://doi.org/10.31349/revmexfis.68.031203 

Research

Nuclear Physics

Microscopic spin orbit analysis for proton+9Be scattering

Zakaria M. M. Mahmouda 

aPhysics Department, Faculty Of Science, King Khalid University, Suadi Arabia. Physics Department, Faculty of Science, New Valley University, Egypt. e-mail: zamahmoud@kku.edu.sa


Abstract

We simultaneously reanalyzed the elastic scattering differential cross-sections (dσ/dΩ) and the vector analyzing power (Ay ) of p+9Be system. This analysis was performed using microscopic folded potentials for both the real central and the spin-orbit. For the imaginary central, we used surface Woods-Saxon (WS) potential. We aimed to test the microscopic spin orbit potential based on the M3Y effective nucleon-nucleon interaction for the light system p+9Be. The present calculation showed that the microscopic spin orbit potential satisfactory reproduce A y above 8 MeV and qualitatively reproduced Ay below 8 MeV. In addition, we found that the calculated real central potentials successfully reproduced the dσ/dΩ for all the considered energies. From the present analysis, we excepted that the present microscopic spin orbit potential could reproduce successfully the Ay for p+nucleus as the incident proton energy increases above 10 MeV.

Keyword: Proton+nucleus scattering; p+9Be scattering; p+9Be analyzin power; microscopic spin orbit potential; folding model

1 Introduction

Proton-projectile is one of the well-known nuclear probes. Using the proton as a projectile, we could obtain useful information about the nuclear structure and nuclear interaction. As an example, we could obtain information about the radial distribution of proton, neutron, and nuclear matter inside the nucleus. Moreover, from the proton-nucleus scattering, we could test the reliability of any theoretical model for nuclear structure or interaction.

Systematic studies on proton scattering from light weakly bound He, Li, and Be isotopes were carried out at wide energy ranges [1,2,3,4,5,6,7,8,9]. Polarized proton scattering from nuclei adds another constraint to the scattering problem and reflects various aspects of nuclear structures and reaction mechanisms. For example, Sakaguchi et al. [10] used the polarized proton for scattering from 6He to find the appropriate structure of this exotic nucleus. Uesaka et al. 11], found through analyzing Sakaguchi’s et al. data that dσ/dΩ favor the existence of 6He two neutrons halo at backward angles and that the cluster structure reproduced reasonably well the experimental data. In addition, they refered to the indirect effect of neutrons halo on Ay calculation. Mahmoud et al. [12] reanalyzed the dσ/dΩ and A y for the p+6He system in the framework of the optical model potential using microscopically real central and spin orbit (SO) potentials. They used CDM3Y6 energy and density-dependent version of the effective M3Y nucleon-nucleon (NN) interaction [13]. Their SO potential was calculated using the formalism reported in [13]. They aimed to study the validity of the microscopic SO potential based on CDM3Y6 effective NN interaction. In addition, they aimed to study the structure effect on the scattering observable for p + halo-nuclei system. As expected from these studies, the scattering of polarized protons shows different behavior in exotic nuclei compared to stable ones.

Many theoretical models were conducted to describe nuclear clustering in the 9Be [14,15]. The exotic 9Be nucleus attracted attention because of its Borromean structure and its cluster breakup. For four decades, the elastic scattering of protons from 9Be at low energies was extensively studied both experimentally and theoretically. Farag et al. [16] analyzed proton elastic scattering observables of 9,10,11,12Be at a wide range of energy between 3 and 200 MeV using the optical model. They calculated dσ/dΩ and A y and reaction cross sections σR using single folding (SF) real potential based on the density and iso-spin dependent M3Y effective NN interaction and imaginary part based on the high energy approximation. They used the Thomas form with a radial form for the SO potential based on the real folded potential. They claimed that the SF potential reproduced the scattering observables for energies up to 100 MeV using the non-relativistic Schrodinger equation. For higher energies, they found that the high energy approximation or the eikonal approximation could reproduce the scattering observables better than the optical model of the non-relativistic Schrödinger equation. Meridi [17] analyzed the elastic scattering of protons from 9Be nucleus at incident energies up to 1000 MeV/nucleon. He used energy-dependent microscopic optical model potential based on the density- and isospin-dependent M3Y-Paris nucleon-nucleon (NN) interaction for the real and spin-orbit parts. For the imaginary part, he used the NN-scattering amplitude of the high-energy approximation. His microscopic complex spin-orbit was taken within Breiva-Rook approximation [18]. He found that the optical model potential fails to reproduce the differential cross-section data at energies larger than 100 MeV/nucleon. In addition, he found that a good improvement is obtained by including the surface contribution to the imaginary part. Recently, Maridi et al. [19] analyzed elastic scattering data for p+9Be at proton incident energy below 30 MeV by using two techniques. Their two techniques lead to similar normalized values for the existing data and consistently validate that low-energy data to be safely used for further theoretical studies.

Bingham et al. [20] measured the p+9Be scattering differential cross sections at eleven bombarding energies between 5 and 15.1 MeV. Their data cover a wide angular range from 15° to 170° in the center of mass (c.m.) system. These low energy data were found to differ by 15-20% between different measurements [21]. These data [20] were investigated by Keely et al. [21] to trace and remove any normalization inconsistencies using a coherent coupled reaction channels (CRC) method. The results of Keely et al. [21] motivated Pakou et al. 9] to reanalyze the data for this system using the microscopic JLM complex potential. The results of Keeley [21] and Pakou[9] support the conclusion of negligible or no compound elastic contribution to the elastic scattering dσ/dΩ at low energies.

The measured data of the elastic scattering Ay in conjunction with available dσ/dΩ of protons by complex nuclei extended the scope of the optical model. The analysis of A y data is essential and unique in obtaining information about the nuclear SO interaction. The availability of dσ/dΩ data makes the central parts of the optical model well defined where the SO potential has only a small effect upon the calculated dσ/dΩ. So, A y measurements and analysis are required to determine the SO interaction in a systematic way [22].

The results of Refs. [12,21,9] motivated us to reanalyze the elastic scattering of the proton from the exotic 9Be. The aim is to study the applicability of the microscopic SO potential based on the CDM3Y6 effective NN interaction for this system. In addition, we aimed to examine the success of the central SF real potential for analyzing scattering data at the considered low energy.

2 Theoretical formalism

2.1 Central real potential

In the present work, we used the SF model to calculate the p+9Be central real potential. This real central potential is calculated through the folding procedure from the following relation,

V(E,R)=ρ(r)νNN(|s|,ρ,E)dr,        s=R-r (1)

As shown from Eq. (1), the SF model has two essential ingredients: 1) a realistic NN effective interaction and, 2) target point nucleon density distribution. The energy- and density-dependent CDM3Y6 effective NN interaction [23] is used as an effective NN interaction , νNN(|s|,ρ,E). This effective interaction has the following form,

νNN(|s|,ρ,E)=g(E)F(ρ)υ00(01)D(Ex)(|s|)+υSO0(1)(|s|), (2)

where the intrinsic energy g(E) and density F(ρ) dependent factors [13] have the following forms,

g(E)=1.0-0.0026EN, (3)

F(ρ)=0.26581+3.8033exp(-1.4099ρ)-4.0ρ (4)

The radial forms of the iso-scalar (isospin T = 0) and iso-vector (isospin T = 1) components of the central M3Y-Paris interaction [13] have the following Yukawa forms,

υ00(01)D(Ex)(|s|)=i=13Y00(01)D(Ex)(i)exp(-Ri|s|)Ri|s|. (5)

Similarly, the SO components are represented in the Yukawa forms as,

υSOT(|s|)=i=13YSOT(i)exp(-Ri|s|)Ri|s|. (6)

The explicit ranges and strengths parameters of these Yukawa forms as given by Khoa et al. [13] are presented in Table I. The direct part of the real central folded potential is computed from,

V0TDE,R=ρTrυ0TDs,ρ,Ed3r (7)

Table I Yukawa ranges and strengths of the central and SO components of the M3Y-Paris effective NN interaction.  

i Ri fm-1 Y00D(i) MeV Y01D(i) MeV Y00Ex(i) MeV Y01Ex(i) MeV YSO0(i) MeV YSO1(i) MeV
1 4.0 11061.625 313.625 -1524.25 -4118.0 -5101.0 -1897.0
2 2.5 -2537.5 223.5 -518.75 1054.75 -337.0 -632.0
3 0.7072 0.0 0.0 -7.8474 2.6157 0.0 0.0

The exchange part of the real central folded potential is computed from,

V0TEX(E,R)=ρT(R,r)υ0TD(|s|,ρ,E)j0(k(Ec.m.R)|s|)dr, (8)

j0(x) is the zero order spherical Bessel function. ρ0,1(r) are the iso-scaler and iso-vector densities, respectively, which are defined as,

ρ0(r)=ρp(r)+ρn(r), (9)

ρ1(r)=ρp(r)-ρn(r), (10)

ρ0(R,r)=ρp(R,r)+ρn(R,r), (11)

ρ1(R,r)=ρp(R,r)-ρn(R,r), (12)

and k(E,R) is the relative momentum, which has the form,

kE,R=2M2Ec.m-VE,R-VCR (13)

Here M stands for the reduced nucleon mass, VC (R) is the Coulomb potential and V(E, R) is the total real folded nuclear potential. The density matrix ρk (R, r), (k = p, n), is considered using the following approximation,

ρi(R,r)=ρ|R+s2|j1kfi|R+s2|s, (14)

j1(x)=3sin(x)-xcos(x)x3, (15)

kfi(r) is the Fermi momentum and approximated as,

kfi(r)=53ρi(r)τi(r)-142ρi(r), (16)

τi(r) (the kinetic energy density) has the Thomas-Fermi approximation form [16] as,

τi(r)=3(3π2)2/35ρi(r)5/3+ρi(r)236ρi(r)+2ρi(r)3,i=p,n. (17)

Where p, n stand for proton and neutron, respectively.

2.2 Central SO potential

The formalism described in an earlier report[13] is used to calculate the SO-potential in the present work. According to this formalism and using the SO component of CDM3Y6 effective NN interaction and the target nuclear matter density the SO potential is computed microscopically as,

VSO(E,R)=-gEFρR2Φp(E,R)1RdρpRdR+Φn(E,R)1Rdρn(R)dR, (18)

Φp(E,R)=0υSO1(s)[1+j1(k(E,R)s)]s4ds, (19)

Φn(E,R)=120(υSO1(s)[1+j1{k(E,R)s}]+υSO0(s)[1-j1{k(E,R)s}])s4ds (20)

For comparison, we used a Thomas form SO potential with radial form factor based on WS or the SF real central potentials, respectively. Thus the SO potentials used in this work are written formally as,

VSO(E,R)=NsoVSO(E,R),mπc)2VSORddR11+exp{R-Rso}aso,(λπ)2NsoRddRV(E,R). (21)

λπ=/mπc is the pion wavelength. In this work, the optical potentials based on the microscopic SO potential is denoted as MI-SO, while that based on WS Thomas form is denoted as PH-SO and that based on the central real is denoted as CE-SO. The 9Be density is based on the experimental charge density [24] as,

ρi(r)=ρ0i1+ωr2exp(-βr2). (22)

This density form is known as a modified Gaussian and has charge root-mean-square radius rch21/2=2.519. The point nucleon density is obtained from this form by unfolding the finite proton size. The obtained density gives ρ0p=0.069941,ρ0n=0.0874263 with point nucleon mean square radius ri2=6.345-rp,ch2, where rp,ch2=0.76-0.11(N/Z) [25].

3 Results and discussion

We used the auto-search optical model computer code HERMES [26] to optimize our calculated elastic scattering dσ/dΩ and A y to the experimental data. The optimization are carried out by minimizing the χ2 value,

    χ2=1Ni=1Nσth(θi)-σex(θi)Δσex(θi)2, (23)

σth is the calculated, σex is the experimental cross sections at angle θi in the c.m. system, Δσex is the experimental error, and N is the number of data points. For the experimental errors, we adopted an average overall value of 10% for all the considered data. Each part of our calculated potential must be multiplied by a re-normalization factor to reproduce the experimental data. Hence, the total nuclear potential U(E, R), which are used to calculate the scattering observables, can be formally written as,

U(E,R)=NrV(E,R)+iW(R)+NsoVSO(E,R)(2LS), (24)

where, L is the relative angular momentum and S is the spin of the proton. Nr , Nso are the re-normalization factors of the central real and SO potentials, respectively.

WR=4aiW0ddR11+expR-Riai,            Ri=riAT1/3. (25)

W0, Ri and ai are the depth and shape parameters of the imaginary potential, respectively. In the optimization procedure, we searched for optimizing Nr, Nso and the WS shape parameters for surface imaginary and Thomas form SO potentials, respectively. Our calculated elastic scattering dσ/dΩ and Ay are shown in Figs. 1, 2. The optical model best fit parameters and the corresponding calculated quantities are listed in Tables II-IV.

Figure 1 Calculated dσ/dΩ (left panels) and Ay (right panels) for p+9Be system at energies between 3 and 7 MeV.  

Figure 2 Calculated dσ/dΩ (left panels) and Ay (right panels) for p+9Be system at energies between 8 and 15 MeV.  

Table II Phenomenological optical model fitting parameters for p+9Be system at energies between 3 and 15 MeV.  

Ep MeV Nr W0 MeV ri fm ai fm V0oso MeV rso fm aso fm
3 1.362 4.365 2.512 0.497 5.719 1.736 0.114
5 1.163 7.218 1.674 0.511 10.403 1.362 0.104
6 1.198 10.732 1.399 0.511 9.635 1.282 0.106
7 1.157 11.484 1.509 0.513 8.471 1.302 0.104
8 1.116 9.926 1.596 0.509 7.252 1.372 0.200
9 1.081 8.788 1.423 0.550 7.225 1.378 0.190
10 1.048 10.080 1.368 0.553 4.564 1.211 0.158
15 1.027 7.496 1.273 0.552 3.833 1.164 0.152

Table III Optical model fitting parameters for p+9Be system at energies between 3 and 15 MeV.  

EpMeV Potential Nr Nso a W0 MeV ri fm ai fm
3 MI-SO 1.342 1.143 4.220 2.600 0.526
CE-SO 1.336 0.277 4.035 2.600 0.540
5 MI-SO 1.224 1.555 7.330 1.725 0.511
CE-SO 1.211 0.328 8.095 1.741 0.511
6 MI-SO 1.135 1.345 11.034 1.607 0.533
CE-SO 1.112 0.245 12.631 1.589 0.533
7 MI-SO 1.067 1.550 13.140 1.567 0.503
CE-SO 1.097 0.301 14.392 1.539 0.503
8 MI-SO 1.065 1.242 8.530 1.210 0.545
CE-SO 1.085 0.255 10.746 1.572 0.505
9 MI-SO 1.070 1.409 7.731 1.164 0.560
CE-SO 1.169 0.247 14.112 1.117 0.508
10 MI-SO 1.134 1.405 12.946 1.126 0.550
CE-SO 1.115 0.260 13.565 1.078 0.550
15 MI-SO 1.056 1.200 10.764 1.141 0.540
CE-SO 1.078 0.275 10.630 1.128 0.556

Table IV Volume integrals for central real, central imaginary and SO potentials.  

Ep MeV Potential Jr MeV·fm3 Ji MeV·fm3 Jso MeV·fm3 σR mb
3 PH-SO 855.4 340.5 57.67 992.3
MI-SO 843.1 373.7 40.26 1027.0
CE-SO 839.6 367.7 50.90 1033.0
5 PH-SO 719.1 267.4 82.30 659.0
MI-SO 756.7 287.3 54.24 791.9
CE-SO 748.5 322.8 58.87 820.2
6 PH-SO 734.5 285.6 71.75 676.9
MI-SO 696.0 397.8 46.68 818.9
CE-SO 681.9 446.0 44.02 817.4
7 PH-SO 703.8 352.8 64.07 730.6
MI-SO 649.0 422.9 53.53 784.6
CE-SO 667.3 447.7 53.68 779.5
8 PH-SO 673.7 335.2 57.79 749.6
MI-SO 642.6 189.6 42.68 717.0
CE-SO 654.7 349.7 45.14 774.4
9 PH-SO 647.2 263.5 57.85 707.0
MI-SO 640.4 166.5 48.18 671.6
CE-SO 699.6 250.0 43.39 718.8
10 PH-SO 622.3 283.4 32.11 724.8
MI-SO 673.2 257.6 47.8 705.4
CE-SO 662.0 250.5 45.34 709.1
15 PH-SO 586.1 185.0 25.92 572.3
MI-SO 602.4 213.8 39.85 599.3
CE-SO 615.0 215.0 46.19 611.0

In Fig. 1 we present the calculated dσ/dΩ (left panels) and A y (right panels) for p+9Be system at energies between 3 and 7 MeV. As shown from this figure, the calculated Ph-, MI- and CE-SO potentials are able to reproduce the experimental dσ/dΩ with equal success. For A y the MI- and CE-SO potentials failed to reproduce the experimental data successfully but have the same angular distribution pattern. The PH-SO potential is successfully reproduced both the dσ/dΩ and A y angular distributions for all the considered energies.

Moreover, we found that the success of MI-SO and CE-SO potentials are improved in reproducing A y angular distributions as energy increases.

Our calculated dσ/dΩ (left panels) and A y (right panels) at energies between 8 and 15 MeV are shown in Fig. 2. As shown in this figure, the MI- and CE-SO potentials reproduced both the dσ/dΩ and A y reasonably well. The improvement of the calculated A y started at 8 MeV, where both MI- and CE-SO potentials reproduced the angular distribution successfully at the foreword angles up to θ80 and overestimated it for larger angles but kept the same angular pattern. As energy increases above 8 MeV, the two potentials reasonably reproduced the A y over the full considered angular ranges.

The energy dependence of optical model searched parameters, and calculated quantities are shown in Figs. 3, 4. From Fig. 3, we see that the real J r for the three potentials globally decrease exponential-like with increasing energy. A fine look at this figure shows that the N r has a small hill at 5, 10, and 9 MeV for Ph-, MI- and CE-SO potentials, respectively. The energy dependence of J i has a peak at around 6.5 MeV for MI- and CE-SO and at 7 MeV for PH-SO potentials. Moreover, MI-SO potential has a minimum of around 8.5 MeV.

Figure 3 The obtained volume integrals J for p+9Be system at energies between 3 and 15 MeV.  

Figure 4 The obtained Ri21/2 for p+9Be system at energies between 3 and 15 MeV.  

The corresponding Ri21/2 (see Fig. 4) has shoulders between 5-8 MeV and 5-9 for MI- CE-SO potentials, respectively. Ri21/2 corresponding to PH-SO potential shows a peak around 8.5 MeV. In general, the R21/2 for the three potentials shows an exponential-like decrease with increasing energy. The SO J so could be approximated with linear relations for MI- and CE-SO potential with a small ripple in the energy interval between 5-9 MeV. For PH-SO potential, J so has a complex energy dependence all over the considered energy range. It is sharply increasing from 3 to 5 MeV and then linearly decreasing from 5-8 MeV. It shows a shoulder from 8 to 9 MeV and then weakly decreases from 10-15 MeV. The Rso21/2 for MI-, CE-SO potentials are approximately constant with values 2.660 and 3.197 fm, respectively. The Rso21/2 for PH-SO potential is similar in energy behavior to the corresponding Ri21/2 one. That means the Rso21/2 for MI-SO potential is less than that of PH-SO potentials for most energies. In contrary, the Rso21/2 of CE-SO potential is larger than that of PH-SO potentials for most energies. For compensation, the MI-SO potential has to be re-normalized by more than unity, and CE-SO potential has to be re-normalized by less than unity for all energies to fit the calculated data with the experimental ones.

In Fig. 5, we present the energy dependence of the total reaction cross sections σR in comparison with the available experimental data close to the considered energies [27]. As shown in this figure, the calculated σR are very close to the experimental ones. Also, the calculated σR is very close to the calculated one based on the four body CDCC at energy around 6 MeV [28]. This agreement indicates the success of the present microscopic potential. In addition, we found that the σR for MI-, CE-SO potentials has a sharp decrease from 3-5 MeV and then decreases approximately linear with increasing energy from 5-15 MeV. For PH-SO potential, this quantity has a pocket with a minimum of around 5.5 MeV and then has a linear decrease above 8 MeV.

Figure 5 Calculated total reaction cross sections σR for p+9Be system at energies between 3 and 15 MeV.  

From these energy dependences, we could expect the influence of 9Be breakup channel at the energy interval between 5-8 MeV. In addition, we could conclude that the present data are not entirely free from the normalization problem. This conclusion comes from the data analysis at 5 and 6 MeV, where a theory normalization for PH-SO calculation 1.276, 1.123 has to be introduced to reproduce the data, respectively.

4 Conclusion

In the present work, we analyzed the p+9Be elastic scattering at 3-15 MeV. Both the differential dσ/dΩ and A y are analyzed simultaneously in the framework of the optical model. The real part of the optical model potential is computed using the SF procedure. For the SO potential, we adopted microscopic and phenomenological Thomas form methods. For microscopic SF and SO potentials, the CDM3Y6 effective NN interaction is used. In the phenomenological Thomas form method for the SO potential, the radial form factor is chosen in the WS-form (phenomenological form) or in the form of the calculated SF real potential (semi-microscopic form). The optical potential imaginary part is adopted in the conventional surface WS form throughout this analysis.

The optimization of the calculated dσ/dΩ and A y is done using the spherical optical model code HERMES [26]. We found that the SF real potential with the different versions of SO potentials can reproduce the dσ/dΩ and the A y with the phenomenological SO potential for all energies. Also, we found that the microscopic SO potential and the semi-microscopic SO potentials cannot reproduce the A y at energies below 8 MeV, while they are reasonably and successfully reproduced A y at energies ≥ 8 MeV. The success of this microscopic SO potential is increased while increasing energy.

In conclusion, we found that the microscopic SO potential is successful in reproducing the analyzing power A y for the p+9Be system contrary to the founding of p+6He [12]. The present analysis provides a good application for using the microscopic spin-orbit potential. The success of the microscopic SO potential motivates us to study the structural effects of 9Be on the scattering observable and other calculated quantities. Also, it encourages us to extend it to other nuclei where experimental A y data exist over a wide energy range.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through Research Groups Program under grant no. R.G.P.1/78/42.

References

[1] V. K. Lukyanov et al., Calculations of 8 He + p elastic cross sections using a microscopic optical potential, Phys. Rev. C 80 (2009) 024609, https://doi.org/10.1103/PhysRevC.80.024609. [ Links ]

[2] M. Y. H. Farag, E. H. Esmael, and H. M. Maridi, Microscopic study on proton elastic scattering of helium and lithium isotopes at energy range up to 160 MeV/nucleon., EPJ Web of Conferences, 66 (2014) 03025, https://doi.org/10.1051/epjconf/20146603025. [ Links ]

[3] P. Egelhof, Nuclear-matter distributions of halo nuclei from elastic proton scattering in inverse kinematics Eur Phys J A 15, (2002) 27, https://doi.org/10.1140/epja/i2001-10219-7. [ Links ]

[4] A. Pakou, Global description of the 7 L1 + p reaction at 5.44 MeV/u in a continuum-discretized coupled-channels approach, Phys. Rev. C 96 (2017) 034615, https://doi.org/10.1103/PhysRevC.96.034615. [ Links ]

[5] A. Pakou et al., Probing the cluster structure of 7Li via elastic scatter- ing on protons and deuterons in inverse kinematics, Phys. Rev. C 94 (2016) 014604, https://doi.org/10.1103/PhysRevC.94.014604. [ Links ]

[6] S. Sakaguchi et al., Analyzing Power in Elastic Scattering of Polarized Protons from Neutron-rich Helium Isotopes Few-Body Syst. 54 (2013) 1393, https://doi.org/10.1007/s00601-013-0617-1. [ Links ]

[7] V. Soukeras et al., Reexamination of 6 Li + p elastic scattering in inverse kinematics, Phys. Rev. C 91 (2015) 057601, https://doi.org/10.1103/PhysRevC.91.057601. [ Links ]

[8] S. P. Weppner, A nucleon-nucleus optical model for A ( 13 nuclei at 65-75 MeV projectile energy, J Phys G: Nucl Part Phys 45 (2018) 095102, https://doi.org/10.1088/1361-6471/aad53d. [ Links ]

[9] A. Pakou et al., A Microscopic Approach for p + 9 Be at Energies Between 1.7 to 15 MeV/nucleon, Acta Phys. Polon. B50 (2019) 1547, https://doi.org/10.5506/aphyspolb.50.1547. [ Links ]

[10] S. Sakaguchi, et al., Analyzing Power in Elastic Scattering of 6 He from a Polarized Proton Target at 71 Mev/nucleon, Phys. Rev. C 84, (2011) 024604, https://doi.org/10.1103/PhysRevC.84.024604. [ Links ]

[11] T. Uesaka, et al., Analyzing power for proton elastic scattering from the neutron-rich 6 He nucleus, Phys. Rev. C 82 (2010) 021602, https://doi.org/10.1103/PhysRevC.82.021602. [ Links ]

[12] Zakaria M. M. Mahmoud, and Awad A. Ibraheem, and M. A. Hassanain, Microscopic spin-orbit potential for p +6 He elastic scattering, Int. J. Mod. Phys. E 28 (2019) 1950074, https://doi.org/10.1142/S0218301319500745. [ Links ]

[13] D. T. Khoa, E. Khan, G. Coló and N. Van Giai, Folding model analysis of elastic and inelastic proton scattering on sulfur isotopes, Nucl. Phys. A 706, (2002) 61, https://doi.org/10.1016/S0375-9474(02)00866-7. [ Links ]

[14] T. A. D. Brown, et al., Decay studies for states in 9 Be up to 11 MeV: Insights into the n + 8 Be and α + 5 He cluster structure, Phys. Rev. C 76, (2007) 054605, https://doi.org/10.1103/PhysRevC.76.054605. [ Links ]

[15] P. Papka, et al., Decay path measurements for the 2.429 MeV state in 9 Be: Implications for the astrophysical α + α + n reaction, Phys. Rev. C 75 (2007) 045803, https://doi.org/10.1103/PhysRevC.75.045803. [ Links ]

[16] M. Y. H. Farag, E. H. Esmael, and H. M. Maridi, Analysis of proton- 9,10,11,12 Be scattering using an energy-, density-, and isospin-dependent microscopic optical potential, Phys. Rev. C 90 (2014) 034615, https://doi.org/10.1103/PhysRevC.90.034615. [ Links ]

[17] H. M. Maridi, Energy dependence and surface contribution of the folding-model optical potential for nucleon-nucleus scattering at energies up to 1 GeV/nucleon, Phys. Rev. C 100 (2019) 014613, https://doi.org/10.1103/PhysRevC.100.014613. [ Links ]

[18] F. A. Brieva and J. R. Rook, Nucleon-nucleus optical model potential: (III). The spin-orbit component, Nucl. Phys. A 297 (1978) 206, https://doi.org/10.1016/0375-9474(78)90272-5. [ Links ]

[19] H. M. Maridi, A. Pakou, and K. Rusek, The p +9Be elastic scattering below 30 MeV: Optical model analysis and data normalization, Int. J. Mod. Phys. E 30, (2021) 2150024, https://doi.org/10.1142/S0218301321500245. [ Links ]

[20] F. W. Bingham, M. K. Brussel and J. D. Steben, Scattering of 5-15 MeV protons from Be 9, Nucl. Phys. 55 (1964) 265, https://doi.org/10.1016/0029-5582(64)90145-2. [ Links ]

[21] N. Keeley, et al., Coherent coupled-reaction-channels analysis of existing and new p + 9 Be data between 1.7 and 15 MeV/nucleon, Phys. Rev. C 99 (2019) 014615, https://doi.org/10.1103/PhysRevC.99.014615. [ Links ]

[22] D. J. Baugh, J. A. R. Griffith and S. Roman, Polarization of 17.8 MeV protons elastically scattered by nuclei, Nucl. Phys. B 83 (1966) 481, https://doi.org/10.1016/0029-5582(66)90106-4. [ Links ]

[23] Dao T. Khoa, G. R.Satchler, and W. von Oertzen, Folding analysis of the elastic 6 Li + 12 C scattering: Knock-on exchange effects, energy dependence, and dynamical polarization potential, Phys. Rev. C 51 (1995) 2069, https://doi.org/10.1103/PhysRevC.51.2069. [ Links ]

[24] H. De Vries, C. W. De Jager, and C. De Vries, Nuclear charge and magnetization density distribution parameters from elastic electron scattering, Atom. Data Nucl. Data Tabl. 36 (1987) 495. https://doi.org/10.1016/0092-640X(87)90013-1. [ Links ]

[25] J. Cook, DFPOT - A program for the calculation of double folded potentials, Comput. Phys. Commun. 25 (1982) 125, https://doi.org/10.1016/0010-4655(82)90029-7. [ Links ]

[26] J. Cook, Hermes - an optical model search program including tensor potentials for projectile spins 0 to 3/2 Comput Phys Commun 31 (1984) 363, https://doi.org/10.1016/0010-4655(84)90020-1. [ Links ]

[27] R. F. Carlson, Proton-Nucleus Total Reaction Cross Sections and Total Cross Sections Up to 1 GeV, At. Data. Nucl. Data Tables 63 (1996) 93, https://doi.org/10.1006/adnd.1996.0010. [ Links ]

[28] V. Soukeras et al., Global study of 9Be + p at 2.72 A MeV, Phys. Rev. C 102 (2020) 064622, https://doi.org/10.1103/PhysRevC.102.064622. [ Links ]

Received: September 27, 2021; Accepted: October 26, 2021

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License