This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Coulomb-nuclear dynamics in the weakly-bound 8Li breakup

B. Mukeru Department of Physics, University of South Africa, P.O. Box 392, Pretoria 0003, South Africa    J. Lubian Instituto de Física, Universidade Federal Fluminense, Niterói, RJ, 24210-340, Brazil    Lauro Tomio Instituto de Física Téorica, Universidade Estadual Paulista, 01140-070 São Paulo, SP, Brazil
Abstract

A detailed study of total, Coulomb and nuclear breakup cross sections dependence on the projectile ground-state binding energy εb\varepsilon_{b} is presented, by considering the 8Li+12C and 8Li+208Pb breakup reactions. To this end, apart from the experimental one-neutron separation energy of 8Li nucleus (εb=2.03\varepsilon_{b}=2.03 MeV), lower values of εb\varepsilon_{b} down to εb=0.01\varepsilon_{b}=0.01 MeV, are also being considered. It is shown that all breakup processes become peripheral as εb0.01\varepsilon_{b}\to 0.01 MeV, which is understood as due to the well-known large spacial extension of ground-state wave functions associated to weakly-bound projectiles. The Coulomb breakup cross section is found to be more strongly dependent on εb\varepsilon_{b} than the nuclear breakup cross section, such that the Coulomb breakup becomes more significant as εb\varepsilon_{b} decreases, even in a naturally nuclear-dominated reaction. This is mainly due to the long-range nature of the Coulomb forces, leading to a direct dependence of the Coulomb breakup on the electromagnetic transition matrix. It is also highlighted the fact that the nuclear absorption plays a minor role for small binding when the breakup becomes more peripheral.

Keywords: Nuclear fusion reactions; Cross sections; halo nuclei; 8Li; 208Pb; 12C

pacs:
24.10.Eq,24.10.-i

I Introduction

In the breakup of weakly-bound projectiles against heavy nuclei targets, a relevant phenomenon which has been investigated is the Coulomb-nuclear dynamics, such that considerable efforts have been made to understand the role of Coulomb-nuclear interference and the dynamics of fragments absorption in the breakup process. The established studies in this matter with the corresponding most relevant works can be found in Refs. 2003Suzuki ; Thompson100 ; Chat10 . For other complementary studies done in past two decades on the Coulomb-nuclear dynamics involving weakly-bound projectiles, with particular interest to our present investigation, we select the Refs. Thomp20 ; 1999Th ; 2002Margueron ; 2003Capel ; Tarutina10 ; Hussein20 ; 2006Canto ; 2009Lubian ; 2009Canto , as well as more recent works (among which we include contributions by some of us) in Refs. Kucuk10 ; 2014Capel ; Hussein200 ; 2015Lubian ; Mukeru10 ; Mukeru20 ; 2016Manjeet ; Pierre100 ; 2017Mukeru ; MukeruPRC2020 . Despite the advances verified by these studies, the question on how both Coulomb and nuclear forces interfere to produce a total breakup remains far from being fully established. Some of the challenges emanate from the fact that, in a Coulomb-dominated reaction, small contribution of the nuclear breakup does not automatically imply insignificant Coulomb-nuclear interference Nakamura10 ; Noc10 ; Aum10 ; Fukuda10 ; Abu10 . It could be interesting to verify what happens in nuclear-dominated reactions.

In view of the long-range nature of Coulomb forces, a low breakup threshold is expected to lead to peripheral collisions, where the Coulomb breakup prevails over the nuclear breakup. In this peripheral region, the Coulomb breakup cross section depends on the projectile structure through the electromagnetic matrix elements of the projectile. Although not a general rule, according to the Coulomb dissociation method Bertulani10 ; Winther10 ; Baur10 ; Baur20 , the breakup cross section is simply the product of the reaction parameters and the projectile dipole electric transition probability. As the binding energy decreases, the reaction becomes more peripheral, with the ratio between the Coulomb breakup cross section to the nuclear counterpart being expected to rise significantly, regardless the target mass. Intuitively, in this case, one would expect that the total breakup cross section becomes comparable to the Coulomb one, owing to both dynamic and static breakup effects. From the fact that lower is the ground-state binding, longer is the tail of the associated wave function, the nuclear forces are fairly stretched beyond the projectile-target radius. Therefore, for a projectile with very weak binding energy, even the nuclear breakup can be assumed to be a peripheral phenomenon, with the Coulomb-nuclear interference becoming stronger in the peripheral region.

The dependence of various reaction observables on the projectile ground-state binding energy has being studied recently in Refs. Wang10 ; Rath100 ; 2016Rangel ; Lei100 ; Mukeru15 ; 2020Mukeru , in which different projectiles with different binding energies have been considered. One of the drawbacks being that all the projectiles do not have the same ground-state structure, mass and charge. Among other ways to circumvent such shortcomings, at least theoretically, one could artificially consider different binding energies for the same projectile (i.e., with nucleon-number AA and charge ZZ unchanged), within an approach that has been adopted for instance in Refs. 2016Rangel ; Lei100 ; Mukeru15 ; 2020Mukeru . Even though a given nucleus has fixed ground-state energy, this is a convenient theoretical approach to unambiguously establish the dependence of the reaction observables on the projectile binding energy.

Another important aspect in breakup dynamics, relies on possible effects on other reaction observables, such as on fusion cross sections. While it is widely understood that the complete fusion suppression strongly depends on the projectile breakup threshold (see Ref. 2020Jha , for recent related studies), strong charge clustering has recently been identified as the main factor responsible for such suppression, in the breakup of 8Li on a heavy target Cook20 . Some behavior in the breakup of this nucleus, has also been reported in Refs. Pak20 ; Gum20 . Unlike several other loosely bound nuclei (such as 8B, 6,7Li, and 11Be), not much has been reported on the breakup dynamics of the Li8{}^{8}{\rm Li} nucleus.

In view of the above discussion, we are motivated to study the breakup of Li8{}^{8}{\rm Li} nucleus, within a model in which a valence neutron (n) is loosely bound to the Li7{}^{7}{\rm Li} nucleus by a binding energy εb=2.03\varepsilon_{b}=2.03 MeV Nut10 , by considering the light and heavy targets 12C and 208Pb. The present study on the 8Li+208+^{208}Pb breakup reaction is also extending a previous recent analysis for this reaction done in Ref. 2020Mukeru , where a critical angular momentum for complete fusion was also considered. We are particularly interested in analyzing the dependence of the resulting total, Coulomb and nuclear breakup cross sections, as well as the Coulomb-nuclear interference, on the projectile ground-state binding energy, in order to test the validity of the assumptions presented in the previous paragraphs. Within a more detailed investigation, we expect to show that for a much weaker projectile binding energy, the Coulomb breakup becomes dominant regardless the target mass, and the nuclear breakup becomes relatively peripheral, leading to a peripheral Coulomb-nuclear interference. Since both Coulomb and nuclear breakup cross sections increase with the decrease of the binding energy, a clear separation of their effects is not a simple task. The choice of 12C and 208Pb as the targets is motivated by the fact that, in the former case, the reaction should be dominated by the nuclear breakup, whereas it is dominated by Coulomb breakup in the latter case. If fact, 12C was also used in Ref. Fukuda10 , as a reference target when studying the 11Be Coulomb dissociation on 208Pb target. In our approach to obtain the corresponding total, Coulomb and nuclear breakup cross sections, we adopt the Continuum Discretized Coupled Channels (CDCC) formalism Aust100 , with the Fresco code 1988Thompson being used for the numerical solutions.

The next sections are organized as follows: Sect. II provides some details on the model approach, with a summary of the CDCC formalism. Sect. III contains the main results for elastic and breakup cross sections, together with our analysis on the Coulomb-nuclear interference and possible absorption contributions. Finally, the Sect. IV presents a summary with our conclusions.

II Formalism and computational approach

II.1 Brief description of the CDCC formalism

As mentioned in the introduction, in our numerical approach we use the CDCC formalism, in which we model the projectile Li8{}^{8}{\rm Li} as Li7{}^{7}{\rm Li} core nucleus, to which a neutron is loosely bound with ground-state energy εb=2.03MeV\varepsilon_{b}=2.03\,{\rm MeV}. This state is defined in the core-neutron centre-of-mass (c.m.) by n=1n=1, 0=\ell_{0}=1, ȷ~0π=2+\tilde{\j}_{0}^{\pi}=2^{+} quantum numbers, where nn stands for the radial state, 0\ell_{0} the orbital angular momentum and ȷ~0π\tilde{\j}_{0}^{\pi} the projectile total angular momentum with parity π\pi. It is obtained by applying the usual spin-orbit coupling 𝒋0=0+𝟏/𝟐{\bm{j}}_{0}={\bm{\ell}}_{0}+{\bf 1/2}; ȷ~0=𝒋0+𝑰c\tilde{\bm{\j}}_{0}={\bm{j}}_{0}+{\bm{I}}_{c}, with the core spin Ic=3/2I_{c}={3}/{2}. In addition to the ground state, an excited bound state with energy εex=0.98MeV\varepsilon_{\rm ex}=0.98\,{\rm MeV} (located in the ȷ~0π=1+\tilde{\j}_{0}^{\pi}=1^{+} state Nut10 ) was also considered in our coupling scheme. We would like to emphasize that we are not considering possible core excitations in our calculations.

In this formalism, we first consider the expansion of the three-body wave function on the projectile internal states. After that, by introducing the three-body expansion into the corresponding Schrödinger equation, a one-dimensional radial set of coupled differential equations can be derived for the radial wave-function components χαLJ(R)\chi_{\alpha}^{LJ}(R), in terms of the projectile-target c.m. coordinate RR, which is given by

[22μpt(d2dR2L(L+1)R2)+UααLLJ(R)]χαLJ(R)\displaystyle\left[-\frac{\hbar^{2}}{2\mu_{pt}}\bigg{(}\frac{d^{2}}{dR^{2}}-\frac{L(L+1)}{R^{2}}\bigg{)}+U_{\alpha\alpha}^{LLJ}(R)\right]\chi_{\alpha}^{LJ}(R)
+αL(αα)UααLLJ(R)χαLJ=(Eεα)χαLJ,\displaystyle+\sum_{\alpha^{\prime}L^{\prime}(\alpha^{\prime}\neq\alpha)}U_{\alpha\alpha^{\prime}}^{LL^{\prime}J}(R)\chi_{\alpha^{\prime}}^{L^{\prime}J}=(E-\varepsilon_{\alpha})\chi_{\alpha}^{LJ}, (1)

where LL is the orbital angular momentum associated with RR, JJ is the total angular momentum, and μpt\mu_{pt} the projectile-target (ptpt) reduced mass. The total energy is given by EE, with εα\varepsilon_{\alpha} being the projectile bin energies. The index α\alpha appearing in the equation is representing a set of quantum numbers describing the projectile states, as given by α(i,,s,j,Ic,ȷ~)\alpha\equiv(i,\ell,s,j,I_{c},\tilde{\j}), i=0,1,2,,Nbi=0,1,2,\ldots,N_{b} (Nb=N_{b}= number of bins).

With the projectile-target potential given as a sum of the core-target (ctct) and neutron-target (ntnt) terms, i.e., Upt(𝒓,𝑹)=Uct(𝑹ct)+Unt(𝑹nt)U_{pt}({\bm{r}},{\bm{R}})=U_{ct}({\bm{R}}_{ct})+U_{nt}({\bm{R}}_{nt}), where 𝑹ct𝑹+18𝒓{\bm{R}}_{ct}\equiv{\bm{R}}+\frac{1}{8}{\bm{r}} and 𝑹nt𝑹78𝒓{\bm{R}}_{nt}\equiv{\bm{R}}-\frac{7}{8}{\bm{r}} (with 𝒓{\bm{r}} being the projectile internal coordinate), the potential matrix elements UααLLJ(R)U_{\alpha\alpha^{\prime}}^{LL^{\prime}J}(R) in (II.1) are given by its Coulomb and nuclear parts, such that

UααLLJ(R)\displaystyle U_{\alpha\alpha^{\prime}}^{LL^{\prime}J}(R) =\displaystyle= 𝒴αL(𝒓,ΩR)|VctCoul(𝑹ct)|𝒴αL(𝒓,ΩR)\displaystyle\langle\mathcal{Y}_{\alpha L}({\bm{r}},\Omega_{R})|V_{ct}^{Coul}({\bm{R}}_{ct})|\mathcal{Y}_{\alpha^{\prime}L^{\prime}}({\bm{r}},\Omega_{R})\rangle
+\displaystyle+ 𝒴αL(𝒓,ΩR)|Uctnucl(𝑹ct)|𝒴αL(𝒓,ΩR)\displaystyle\langle\mathcal{Y}_{\alpha L}({\bm{r}},\Omega_{R})|U_{ct}^{nucl}({\bm{R}}_{ct})|\mathcal{Y}_{\alpha^{\prime}L^{\prime}}({\bm{r}},\Omega_{R})\rangle
+\displaystyle+ 𝒴αL(𝒓,ΩR)|Untnucl(𝑹nt)|𝒴αL(𝒓,ΩR),\displaystyle\langle\mathcal{Y}_{\alpha L}({\bm{r}},\Omega_{R})|U_{nt}^{nucl}({\bm{R}}_{nt})|\mathcal{Y}_{\alpha^{\prime}L^{\prime}}({\bm{r}},\Omega_{R})\rangle,

where 𝒴αL(𝒓,ΩR)[Φ^α(𝒓)iLYLΛ(ΩR)]JM\mathcal{Y}_{\alpha L}({\bm{r}},\Omega_{R})\equiv[\hat{\Phi}_{\alpha}({\bm{r}})\otimes{\rm i}^{L}Y_{L}^{\Lambda}(\Omega_{R})]_{JM} is the direct product of the angular part of 𝑹{\bm{R}} with the projectile channel wave function, Φ^α(𝒓)\hat{\Phi}_{\alpha}({\bm{r}}), which contains the square integrable discretized bin wave functions. The nuclear terms express the sums of real and imaginary parts. The former are responsible for the nuclear dissociation, whereas the latter accounts for the nuclear absorption. These nuclear terms are, respectively, given by Uctnucl(𝑹ct)=Vctnucl(𝑹ct)+iWctnucl(𝑹ct)U_{ct}^{nucl}({\bm{R}}_{ct})=V_{ct}^{nucl}({\bm{R}}_{ct})+{\rm i}W_{ct}^{nucl}({\bm{R}}_{ct}) and Untnucl(𝑹nt)=Vntnucl(𝑹nt)U_{nt}^{nucl}({\bm{R}}_{nt})=V_{nt}^{nucl}({\bm{R}}_{nt})+ iWntnucl(𝑹nt){\rm i}W_{nt}^{nucl}({\bm{R}}_{nt}), with the Woods-Saxon shape being adopted for both components. The diagonal coupling matrix elements UααLLJ(R)U_{\alpha\alpha}^{LLJ}(R), contain the monopole nuclear term in the projectile-target c.m., which we denote by Vβ0β0LJ(R)=Φβ0(𝒓)|Uctnucl+Untnucl|Φβ0(𝒓)V_{\beta_{0}\beta_{0}}^{LJ}(R)=\langle\Phi_{\beta_{0}}({\bm{r}})|U_{ct}^{nucl}+U_{nt}^{nucl}|\Phi_{\beta_{0}}({\bm{r}})\rangle, where β0\beta_{0} represents the set of ground-state projectile quantum numbers, β0(k0,0,s,j0,Ic,ȷ~0)\beta_{0}\equiv(k_{0},\ell_{0},s,j_{0},I_{c},\tilde{\j}_{0}). The imaginary part accounts for the absorption in the projectile-target c.m. motion.

The separation of the Coulomb and nuclear interactions to obtain the Coulomb and nuclear breakup cross sections (σCoul\sigma_{Coul} and σnucl\sigma_{nucl}, respectively) remains a challenge in nowadays theories, making an accurate description of the Coulomb-nuclear interference a more tricky task. For that, in this work we resort to an approximate approach, as follows: The nuclear breakup cross sections, defined as σnucl\sigma_{nucl}, are obtained by including in the coupling matrix elements, the nuclear components of UctU_{ct} and UvtU_{vt} potentials, plus the diagonal monopole Coulomb potential. On the other hand, the Coulomb breakup cross sections, defined as σCoul\sigma_{Coul}, are obtained by including in the matrix elements the Coulomb component of the projectile-target potential, i.e., VctCoul(Rct)V_{ct}^{Coul}(R_{ct}) (as VntCoul=0V_{nt}^{Coul}=0), plus the monopole nuclear potential. The total breakup cross sections σtot\sigma_{tot} are obtained by including the full UptU_{pt} potential in the calculations.

Since the early works on Coulomb and nuclear breakup studies Thomp20 ; 1999Th , this approach has been widely adopted to study Coulomb and nuclear breakup cross sections, as one can follow from the review 2015Canto (and references therein). In Ref. Pierre100 , where different methods are considered in order to decompose the total breakup into its Coulomb and nuclear components, this approach is also referred as weak-coupling approximation. Two methods emerged from their discussion, which they refer to as method 1 and method 2. The weak-coupling approximation is very close to method 1 for nuclear breakup, and close to method 2 for Coulomb breakup. While this approximate procedure will not completely eliminate the ambiguities surrounding the separation of the total breakup cross section into its Coulomb and nuclear components (as also outlined in Ref.Pierre100 ), we believe that it is particularly justified in the present work, since by using the 12C target, the breakup is naturally dominated by nuclear dissociation, whereas by using the 208Pb target the breakup is dominated by Coulomb dissociation.

Once the matrix elements (II.1) are computed, the coupled Eq. (II.1) is solved with the usual asymptotic conditions, which for kα(2μpt/2)(Eεα)k_{\alpha}\equiv\sqrt{{(2\mu_{pt}/\hbar^{2})(E-\varepsilon_{\alpha})}} is given by

χαLJ(R)Ri2[Hα(kαR)δααHα+(kαR)SααLLJ],\displaystyle\chi_{\alpha}^{LJ}(R)\stackrel{{\scriptstyle R\to\infty}}{{\longrightarrow}}\frac{\rm i}{2}\left[H_{\alpha}^{-}(k_{\alpha}R)\delta_{\alpha\alpha^{\prime}}-H_{\alpha}^{+}(k_{\alpha}R)S_{\alpha\alpha^{\prime}}^{LL^{\prime}J}\right], (3)

where Hα(kαR)H_{\alpha}^{\mp}(k_{\alpha}R) are the usual incoming (-) and outgoing (+) Coulomb Hankel functions Abramo100 , with Sαα(kα)S_{\alpha\alpha^{\prime}}(k_{\alpha}) being the scattering S-matrix elements. Due to the short-range nature of nuclear forces, the matrix elements corresponding to the nuclear interaction in Eq. (II.1) will vanish at large distances, RRn{R\gg R_{n}}, where

Rnr0(Ap1/3+At1/3)+δR(εb)R0+δR(εb)R_{n}\equiv r_{0}(A_{p}^{1/3}+A_{t}^{1/3})+\delta_{R}(\varepsilon_{b})\equiv R_{0}+\delta_{R}(\varepsilon_{b}) (4)

determines the range of the nuclear forces (r0r_{0} being the nucleon size, with r0Ap1/3r_{0}A_{p}^{1/3} and r0At1/3r_{0}A_{t}^{1/3} the projectile and target sizes, respectively). The function δR(εb)\delta_{R}(\varepsilon_{b}) is introduced to take into account the well-known effect which occurs in weakly-bound systems (low breakup thresholds), as in halo nuclei, in which the nuclear forces can be stretched beyond R0=r0(Ap1/3+At1/3)R_{0}=r_{0}(A_{p}^{1/3}+A_{t}^{1/3}). The various breakup cross sections are obtained by using the relevant S-matrix, as outlined for example in Ref. Thompson100 .

At large distance (RR\to\infty), Eq.(II.1) contains only the Coulomb interaction, which can be expanded as Hussein50

VCoul(𝒓,𝑹)R4πZteλ=0λmax2λ+1Rλ+1[𝒪λϵ(𝒓)Yλ(ΩR)]0,\displaystyle V^{Coul}({\bm{r}},{\bm{R}})\stackrel{{\scriptstyle R\to\infty}}{{\longrightarrow}}4\pi Z_{t}e\sum_{\lambda=0}^{\lambda_{\rm max}}\frac{\sqrt{2\lambda+1}}{R^{\lambda+1}}\left[\mathcal{O}_{\lambda}^{\epsilon}({\bm{r}})\otimes Y_{\lambda}(\Omega_{R})\right]^{0}, (5)

where ZteZ_{t}e is the target charge, with λ\lambda the multipole order truncated by λmax\lambda_{\rm max}. 𝒪λϵ(𝒓)\mathcal{O}_{\lambda}^{\epsilon}({\bm{r}}) is the projectile electric operator, given by

𝒪λμϵ(𝒓)\displaystyle\mathcal{O}_{\lambda\mu}^{\epsilon}({\bm{r}}) =\displaystyle= [Zce(AnAp)λ]rλYλμ(Ωr)=ZλrλYλμ(Ωr),\displaystyle\left[Z_{c}e\left(-\frac{A_{n}}{A_{p}}\right)^{\lambda}\right]r^{\lambda}Y_{\lambda}^{\mu}(\Omega_{r})=Z_{\lambda}r^{\lambda}Y_{\lambda}^{\mu}(\Omega_{r}), (6)

where ZceZ_{c}e is the charge of the projectile core, with ZλZ_{\lambda} being defined as the effective charge. The projectile electric transition probability for the transition from the projectile ground-state to the continuum states can be obtained through 𝒪λϵ(𝒓)\mathcal{O}_{\lambda}^{\epsilon}({\bm{r}}) Bertulani50 . For excitation energies ε\varepsilon, the corresponding variation of the electric transition probability B(Eλ)B(E\lambda) can be written as

dB(Eλ)dε\displaystyle\frac{dB(E\lambda)}{d\varepsilon} =\displaystyle= μcn2kȷ~(2ȷ~+1)|Φβ0(𝒓)|𝒪λϵ(𝒓)|Φβ(𝒓)|2,\displaystyle\frac{\mu_{cn}}{\hbar^{2}k}\sum_{\tilde{\j}}(2{\tilde{\j}}+1)\left|\langle\Phi_{\beta_{0}}({\bm{r}})|\mathcal{O}_{\lambda}^{\epsilon}({\bm{r}})|\Phi_{\beta}({\bm{r}})\rangle\right|^{2}, (7)

where (β\beta) refers to the set of quantum numbers in the continuum states β(k,,s,j,Ic,ȷ~)\beta\equiv(k,\ell,s,j,I_{c},\tilde{\j})], k=2μcnε/2k=\sqrt{2\mu_{cn}\varepsilon/\hbar^{2}}, k0=2μcn|ε0|/2k_{0}=\sqrt{2\mu_{cn}|\varepsilon_{0}|/\hbar^{2}}, with μcn\mu_{cn} the core-neutron reduced mass. By defining l^2l+1\hat{l}\equiv\sqrt{2l+1} for general angular quantum numbers, from the above we obtain

dB(Eλ)dε\displaystyle\frac{dB(E\lambda)}{d\varepsilon} =\displaystyle= μcn2kȷ~(2ȷ~+1)|λ,ȷ~|2,with\displaystyle\frac{\mu_{cn}}{\hbar^{2}k}\sum_{\tilde{\j}}(2{\tilde{\j}}+1)|\mathcal{F}_{\lambda,{\tilde{\j}}}|^{2},\;\;{\rm with} (8)
λ,j\displaystyle\mathcal{F}_{\lambda,j} \displaystyle\equiv 14πZλ^0^λ^2j^0j^(1)0++s+j+j0+Ic+ȷ~\displaystyle\frac{1}{4\pi}Z_{\lambda}\hat{\ell}_{0}\hat{\ell}\hat{\lambda}^{2}\hat{j}_{0}\hat{j}(-1)^{\ell_{0}+\ell+s+j+j_{0}+I_{c}+{\tilde{\j}}} (15)
×\displaystyle\times (λ0000)(jλj0000){s0j0λj}\displaystyle\left(\begin{array}[]{ccccc}\ell&\lambda&\ell_{0}\\ 0&0&0\end{array}\right)\left(\begin{array}[]{ccccc}j&\lambda&j_{0}\\ 0&0&0\end{array}\right)\left\{\ \begin{array}[]{cccc}s&\ell_{0}&j_{0}\\ \lambda&j&\ell\end{array}\right\}\
×\displaystyle\times {Icj0ȷ~0λȷ~j}0𝑑ruk00ȷ~0(r)rλukȷ~(r),\displaystyle\left\{\ \begin{array}[]{cccc}I_{c}&j_{0}&\tilde{\j}_{0}\\ \lambda&\tilde{\j}&j\end{array}\right\}\ \int_{0}^{\infty}dr\,u_{k_{0}\ell_{0}}^{\tilde{\j}_{0}}(r)r^{\lambda}u_{k\ell}^{\tilde{\j}}(r), (18)

where uk00ȷ~0(r)u_{k_{0}\ell_{0}}^{\tilde{\j}_{0}}(r), and ukȷ~(r)u_{k\ell}^{\tilde{\j}}(r) are the ground-state and continuum radial wave functions. The Eqs. (5)-(8) are indicating how the Coulomb breakup is being affected by the projectile structure.

II.2 Computational details

The energies and corresponding wave functions which appear in the set of coupled differential equations (II.1), for the bound and continuum states of the 7Li+n system, are obtained by considering a two-body Woods-Saxon potential as input, whose parameters are the same as in Ref. Moro200 . The depth V0V_{0} of the central part of the potential was adjusted to reproduce the ground and excited bound-state energies. These parameters are summarized in Table 1.

Table 1: Woods-Saxon potential parameters for the projectile (n7-^{7}Li) ground and excited bound-state energies.
ȷ~π\tilde{\j}^{\pi} V0V_{0} r0r_{0} a0a_{0} VSOV_{\rm SO} rSOr_{\rm SO} aSOa_{\rm SO}
(MeV) (fm) (fm) (MeV/fm2) (fm) (fm)
2+2^{+} 37.22 1.25 0.52 4.89 1.25 0.52
1+1^{+} 46.65 1.25 0.52 4.89 1.25 0.52

Similarly, the other binding energies considered in this work are obtained by adjusting V0V_{0}. The same ground-state potential parameters are adopted to calculate the corresponding continuum wave functions. With these potential parameters, we first calculate the electric transition probability B(E1)B(E1) variation with the excitation energy ε\varepsilon, given by Eq.(8), corresponding to the transition from the ground-state to continuum ss- plus dd-states, for the binding energies εb=0.01MeV,1.0MeV\varepsilon_{b}=0.01\,{\rm MeV},1.0\,{\rm MeV} and 2.03MeV2.03\,{\rm MeV}. The results are shown in the upper panel of Fig.1. One notices that B(E1)B(E1) varies substantially for εb=0.01MeV\varepsilon_{b}=0.01\,{\rm MeV} as compared with values obtained for larger εb\varepsilon_{b}. These results highlight the strong dependence of the Coulomb breakup on the projectile internal structure, particularly in the asymptotic region. In this regard, it is also instructive to verify how the projectile root-mean-square radii r2\sqrt{\langle r^{2}\rangle} vary with the projectile ground-state binding energies. For that, we add the lower panel of Fig. 1, with the corresponding root-mean-square radii, obtained for the projectile ground-state Φβ0(𝒓)\Phi_{\beta_{0}}(\bm{r}). As expected, the root-mean-square radii behavior is reflecting the large increasing of the wave function as the binding energy comes close to zero. Also, for εb=2.033\varepsilon_{b}=2.033 MeV, we note that we obtain r2=2.39\sqrt{\langle r^{2}\rangle}=2.39 fm, in very close agreement with the corresponding values reported in Refs. 2015Fan and 1988Tanihata (respectively, r2=2.39±0.05\sqrt{\langle r^{2}\rangle}=2.39\pm 0.05 fm and r2=2.37±0.02\sqrt{\langle r^{2}\rangle}=2.37\pm 0.02 fm).

Refer to caption
Figure 1: In panel (a), considering three different 7Li-n ground-state binding energies εb\varepsilon_{b}, it is shown how the derivative of the electric transition probability, given by (8), varies with the excitation energy ε\varepsilon, for transitions from ground to continuum ss- plus dd-states. In panel (b), the root-mean-square radii is shown as a function of the binding energy εb\varepsilon_{b}.

In order to evaluate the coupling matrix elements of Eq. (II.1), fragments-target optical potentials are needed. The 7Li+12+^{12}C optical potential parameters were taken from Ref. Barioni100 , whereas the 7Li+208+^{208}Pb optical potential parameters were obtained from the 7Li global potential of Ref. Cook300 , with the depth of the real part slightly modified to fit the elastic scattering experimental data. For the nn-target optical potentials, we adopted the global potential of Ref. 1969Green . The CDCC limiting values of the model space parameters, used for the numerical solution of Eq.(II.1), are listed in Table 2, for the two targets we are considering, 12C and 208Pb, where max\ell_{\rm max} is the maximum angular momentum between Li7{}^{7}{\rm Li} and the neutron, λmax\lambda_{\rm max} is the maximum order of the potential multipole expansion, εmax\varepsilon_{\rm max} is the maximum bin energies, rmaxr_{\rm max} is the maximum matching radius for bin potential integration, LmaxL_{\rm max} is the maximum angular momentum of the relative c.m. motion, and RmaxR_{\rm max} is the maximum matching radius of the integration for the coupled differential equations, with ΔR\Delta R the corresponding RR-step size. The reported main values are found to give enough converged results for εb\varepsilon_{b}\geq0.4 MeV. However, as we decrease the projectile binding energy, for εb0.08\varepsilon_{b}\leq 0.08MeV, to guarantee enough good convergence and precision of the results we found necessary to increase the maximum values for the projectile matching radius rmaxr_{\rm max}, for the matching radius RmaxR_{\rm max}, and for the relative angular momentum of the c.m. motion, LmaxL_{\rm max}, correspondingly to each of the target. These values for smaller εb\varepsilon_{b} are shown within parenthesis, below the respective values obtained for larger εb\varepsilon_{b}. The adopted bin widths were, Δε=0.5MeV\Delta\varepsilon=0.5\,{\rm MeV}, for ss- and pp-states, Δε=1.0MeV\Delta\varepsilon=1.0\,{\rm MeV}, for ff- and dd-states and Δε=1.5MeV\Delta\varepsilon=1.5\,{\rm MeV} for g{\rm g}-states.

Table 2: Maximum model space parameters, for optimal numerical convergence of Eq. (II.1) for both 12C and 208Pb targets. The main reported values are for εb0.4\varepsilon_{b}\geq 0.4MeV, with the corresponding ones within parenthesis for εb0.08\varepsilon_{b}\leq 0.08MeV.
Target max\ell_{\rm max} λmax\lambda_{\rm max} εmax\varepsilon_{\rm max} rmaxr_{\rm max} LmaxL_{\rm max} RmaxR_{\rm max} ΔR\Delta R
(\hbar) (MeV) (fm) (\hbar) (fm) (fm)
C12{}^{12}{\rm C} 3 3 6 80 300 300 0.08
(100) (1000) (500)
Pb208{}^{208}{\rm Pb} 4 4 10 80 1000 600 0.03
(100) (10000) (1000)
Refer to caption
Figure 2: 8Li+12C elastic scattering cross sections for the incident energies, Elab=E_{lab}=14 MeV and Elab=E_{lab}=23.9 MeV. The model results are for different 8Li binding energies εb\varepsilon_{b} (in MeV units), as indicated inside panel (a) for both panels. The available experimental data, converted to Rutherford σR\sigma_{R} units, are from Refs. 1993Becchetti [panel (a)] and Barioni100 [panel (b)], as indicated in the database reported in Ref. Jinr20 .
Refer to caption
Figure 3: 8Li+208Pb elastic scattering cross sections (in units of the Rutherford σR\sigma_{R}), obtained for the incident energies Elab=E_{lab}=36 MeV [panel (a)] and Elab=E_{lab}=60 MeV [panel (b)]. As in Fig. 2, the results are for the same set of εb\varepsilon_{b} (in MeV). From Ref. 2002Kolata , we included in (a) the closest available experimental data, which are for Elab=E_{lab}=30.6 MeV, as indicated in the database reported in Ref. Jinr20 .

III Results and Discussion

III.1 Elastic scattering cross sections

We start the first part of this section by analyzing the dependence of the elastic scattering cross sections on the projectile ground-state binding energy. These cross sections are displayed in Fig. 2 for C12{}^{12}{\rm C} target; and in Fig. 3 for Pb208{}^{208}{\rm Pb} target, considering two incident energies. In both the cases, we assume different values of εb\varepsilon_{b}, from the experimental one down to 0.01 MeV. In the case of C12{}^{12}{\rm C} target, which is a nuclear-dominated reaction, from the results shown in Fig. 2 one can observe a weak dependence on εb\varepsilon_{b} in the range 0.4MeVεb2.03MeV0.4\,{\rm MeV}\leq\varepsilon_{b}\leq 2.03\,{\rm MeV}, for both incident energies, Elab=E_{lab}=14 MeV [panel (a)] and 24 MeV [panel (b)]. However, it becomes relatively significant for εb0.08MeV\varepsilon_{b}\leq 0.08\,{\rm MeV} [see panel (a)]. Also shown in Fig. 2 is that the experimental data are well reproduced by the model for both incident energies.

For the Coulomb-dominated reaction with Pb208{}^{208}{\rm Pb}, the results given in Fig. 3 for Elab=E_{lab}=36 MeV (a) and 60 MeV (b)] are indicating strong dependence of the elastic scattering cross sections on all binding energies at forward angles (asymptotic region), where the Coulomb breakup is particularly dominant. However, at backward angles (short distance), where the nuclear breakup is expected to provide meaningful effects, the elastic cross sections become almost independent of the binding energy.

These results lead to a conclusion that, when the nuclear breakup is dominant or relatively significant, the effect of the binding energy on the elastic scattering cross section is rather small, whereas it is more pronounced when the Coulomb breakup is dominant. Therefore, since a relatively significant effect for the Li8+C12{}^{8}{\rm Li}+{}^{12}{\rm C} reaction is observed when εb0.08MeV\varepsilon_{b}\leq 0.08\,{\rm MeV}, it is possible that the Li8+C12{}^{8}{\rm Li}+{}^{12}{\rm C} reaction is already dominated by the Coulomb breakup for εb0.08MeV\varepsilon_{b}\leq 0.08\,{\rm MeV}. As the binding energy decreases, the Coulomb breakup becomes dominant over its nuclear counterpart, as anticipated. It also follows that the probability of the projectile to fly on the outgoing trajectory unbroken decreases, diminishing the corresponding elastic scattering cross section. In the next section, we will look into this observation in more detail.

III.2 Breakup cross sections

Refer to caption
Figure 4: For incident energies Elab=14MeVE_{lab}=14\,{\rm MeV} (left column) and Elab=24MeVE_{lab}=24\,{\rm MeV} (right column), with fixed different εb\varepsilon_{b} (shown inside the panels), the 8Li+12+^{12}C angular distributions for the total, Coulomb and nuclear differential breakup cross sections dσ/dΩd\sigma/d\Omega (identified inside the upper panels) are shown as functions of the c.m. angle θ\theta.
Refer to caption
Figure 5: Convergence sample results for the 8Li+208+^{208}Pb, total (upper frames), Coulomb (middle frames) and nuclear (bottom frames) breakup angular distributions, dσ/dΩd\sigma/d\Omega, at Elab=36MeVE_{lab}=36\,{\rm MeV}, considering different maximum projectile internal angular momenta max\ell_{\rm max} (indicated in the upper-left frame). The left set [(a)-(c)] is for εb=0.01\varepsilon_{b}=0.01MeV, with the right set [(d)-(f)] for εb=2.03\varepsilon_{b}=2.03MeV.
Refer to caption
Figure 6: For incident energies Elab=36MeVE_{lab}=36\,{\rm MeV} (left column) and Elab=60MeVE_{lab}=60\,{\rm MeV} (right column), with fixed different εb\varepsilon_{b} (shown inside the panels), the Li8+Pb208{}^{8}{\rm Li}+{}^{208}{\rm Pb} angular distributions for the total, Coulomb and nuclear dσ/dΩd\sigma/d\Omega (identified in the upper panels) are shown as functions of the c.m. angle θ\theta.

The differential total, Coulomb and nuclear breakup cross sections, for the C12{}^{12}{\rm C} target, are depicted in Fig. 4, for Elab=E_{lab}=14 MeV [(a)-(e) panels] and Elab=E_{lab}=24 MeV [(f)-(j) panels]. As anticipated, in the case of nuclear-dominated reactions, for both incident energies, dσnucl/dΩd\sigma_{nucl}/d\Omega dσtot/dΩ\simeq d\sigma_{tot}/d\Omega dσCoul/dΩ\gg d\sigma_{Coul}/d\Omega as εb2.03MeV\varepsilon_{b}\to 2.03\,{\rm MeV}, with dσCoul/dΩ0d\sigma_{Coul}/d\Omega\to 0. However, it is interesting to notice that as εb\varepsilon_{b} decreases, the Coulomb breakup increases rapidly, such that for εb0.01MeV\varepsilon_{b}\to 0.01\,{\rm MeV}, dσnucl/dΩd\sigma_{nucl}/d\Omega dσCoul/dΩ\ll d\sigma_{Coul}/d\Omega dσtot/dΩ\simeq d\sigma_{tot}/d\Omega at forward angles, for both incident energies. On the light of these results it follows that as the binding energy further decreases, the Coulomb breakup becomes more relevant, and comparable with the total breakup even in such a naturally nuclear-dominated reaction. This can be attributed to the fact that the breakup becomes more peripheral as εb\varepsilon_{b} decreases, where only Coulomb forces are available. Hence, the importance of the Coulomb breakup in this case relies mainly on the long-range behavior of the Coulomb forces, and on its direct dependence on the electromagnetic transition matrix elements, in agreement with our assessment in Sect. III.1. Furthermore, these results show that the “nuclear-dominated reaction” concept may be relative to the projectile binding energy.

As the projectile binding energy varies from 2.03 MeV down to 0.01 MeV, one may wonder how relevant higher-order partial-waves (\ell) are in the breakup process for such very low binding energy, particularly for heavy targets. In order to verify the importance of higher-order partial-waves in this case, we performed a convergence test of the total, Coulomb and nuclear differential breakup cross sections for 208Pb target at Elab=36E_{lab}=36 MeV. The different breakup cross sections are shown in Fig. 5, as functions of the c.m. angle θ\theta, for different maximum projectile internal angular momenta max\ell_{\rm max}, and only for εb=0.01\varepsilon_{b}=0.01 MeV and 2.03 MeV binding energies. As evidenced by the results in this figure, there is no meaningful difference between max=4\ell_{\rm max}=4 and max=7\ell_{\rm max}=7, regardless the binding energy. This implies that, by reducing the ground-state binding energy, the convergence of the breakup cross sections is not affected, in respect to the maximum core-neutron orbital angular momentum max\ell_{\rm max}.

In Fig.6, displays, the total, Coulomb and nuclear breakup angular cross-section distributions as functions of the c.m. angle θ\theta, for the different binding energies εb\varepsilon_{b}, for 8Li+208+^{208}Pb reaction. We first observe that as εb\varepsilon_{b} decreases, the peaks of dσtot/dΩd\sigma_{tot}/d\Omega and dσCoul/dΩd\sigma_{Coul}/d\Omega are shifted to forward angles. In fact, for εb\varepsilon_{b}\leq 0.08 MeV, the peaks are located close to zero degree. This is a clear display of the peripheral nature of the breakup process as εb\varepsilon_{b} decreases. A careful look at this figure also indicates that as εb\varepsilon_{b} decreases, even the peak of dσnucl/dΩd\sigma_{nucl}/d\Omega is shifted to forward angles, which may suggest that even the nuclear breakup process becomes peripheral as εb\varepsilon_{b}\to 0.01 MeV. The peripherality of the nuclear breakup in this case, can be understood by considering the function δR(εb)\delta_{R}(\varepsilon_{b}), which appears in Eq.(4). The nuclear breakup dynamics require that δR(εb)0\delta_{R}(\varepsilon_{b})\to 0, as εb\varepsilon_{b} increases, implying that RnR0R_{n}\to R_{0}, due to the short-range nature of nuclear forces. However, as εb0\varepsilon_{b}\to 0, δR(εb)\delta_{R}(\varepsilon_{b}) increases and so does RnR_{n}, leading to a significant nuclear effect in the peripheral region. Therefore, the function δR(εb)\delta_{R}(\varepsilon_{b}) is introduced to take into account the well-known effect which occurs in weakly-bound systems, as in halo nuclei, in which the nuclear forces are stretched beyond the usual range.

Quantitatively, since this 8Li+208+^{208}Pb reaction is Coulomb-dominated, we observe that at forward angles both dσtot/dΩd\sigma_{tot}/d\Omega and dσCoul/dΩd\sigma_{Coul}/d\Omega are substantially larger than dσnucl/dΩd\sigma_{nucl}/d\Omega (about three orders of magnitude as εb\varepsilon_{b} decreases). A further inspection of this figure shows that for Elab=60E_{lab}=60 MeV, we notice that the total and Coulomb breakup cross sections are more similar compared to Elab=36E_{lab}=36 MeV, with the difference coming from the competition between the nuclear and Coulomb interactions above the barrier (for a discussion on the role of the diagonal Coulomb interaction, see also Ref. MukeruPRC2020 ).

Refer to caption
Figure 7: For the 8Li+208+^{208}Pb reaction, by considering Elab=E_{lab}=36 MeV (a) and 60 MeV (b), it is shown the integrated breakup cross sections (BU) (when both WctnuclW_{ct}^{nucl} and WntnuclW_{nt}^{nucl} are contributing), the breakup cross section without absorption (NA) (when Wctnucl=Wntnucl=0W_{ct}^{nucl}=W_{nt}^{nucl}=0), and the total fusion cross section (TF), as functions of the projectile binding energy εb\varepsilon_{b}.

In order to better elucidate the importance of the nuclear absorption in the breakup process, we present in Fig. 7, for the 8Li+208Pb reaction, the integrated total breakup cross section as well as the total fusion cross sections as functions of εb\varepsilon_{b}. In this regard, we are extending a previous analysis done for this reaction in Ref. 2020Mukeru , in which the total fusion cross sections are shown as functions of the incident energy for different projectile binding energies. The breakup cross section obtained in the presence of nuclear absorption (i.e., Wctnucl0W_{ct}^{nucl}\neq 0, Wntnucl0W_{nt}^{nucl}\neq 0), are indicated by the label “BU”. The breakup cross section obtained in the absence of nuclear absorption (i.e, Wct=Wnt=0W_{ct}=W_{nt}=0), are indicated by “NA”. The total fusion cross section is labeled as “TF”. By observing this figure, it follows that, as εb2.03\varepsilon_{b}\to 2.03 MeV, the nuclear absorption contributes to largely reduce the breakup cross section about one order magnitude in the log-scale. However, we observe that the nuclear absorption plays a minor role on the breakup cross section for smaller binding energies, being negligible for εb0.01\varepsilon_{b}\to 0.01 MeV, in particular at Elab=60E_{lab}=60 MeV. In this case, σNAσBUσTF\sigma_{\rm NA}\simeq\sigma_{\rm BU}\gg\sigma_{\rm TF}, ( where σBU\sigma_{\rm BU} is the breakup cross section followed by fragments absorption, and σNA\sigma_{\rm NA} is the breakup cross section without fragments absorption after breakup). A larger breakup cross section over the total fusion cross section can be understood as due to the fact that, when the breakup occurs where classically the trajectory is far away from the target, the projectile fragments have no easy access to the absorption region, thus significantly reducing the flux that contributes to the fusion cross section. However, as expected, as εb2.03\varepsilon_{b}\to 2.03 MeV, where the breakup process occurs closer to the target, where the probability for the projectile fragments to survive absorption is significantly reduced, we observe that σBUσNA<σTF\sigma_{\rm BU}\ll\sigma_{\rm NA}<\sigma_{\rm TF}. A weak dependence of the total fusion cross section on the binding energy compared to the breakup cross section is also observed. The energy region well above the Coulomb barrier is particularly dominated by the complete fusion process. As shown in Ref. Lei100 , the complete fusion cross section is insignificantly dependent on the projectile εb\varepsilon_{b} for 7Li+209+^{209}Bi reaction. We believe that these observations would be valid for any loosely bound projectile, and hence there is nothing unusual in the breakup of the Li8{}^{8}{\rm Li} nucleus.

Concerning our approach to total fusion (TF) and absorption, let us clarify that: In the standard CDCC method, the optical potentials are chosen to describe the elastic scattering of the fragments by the target. So, their imaginary parts account for the absorption to fusion and other direct channels (surface reactions). Nevertheless, as direct reaction cross sections are expected to be small for the interactions between fragments and targets selected in this work, the TF cross section provides the major contribution to this absorption.

Refer to caption
Figure 8: For the 8Li+12C breakup reaction, the angular-integrated total, Coulomb and nuclear breakup cross sections are given as functions of the projectile binding energy εb\varepsilon_{b}, for the incident energies Elab=E_{lab}=14 MeV (a) and 24 MeV (b).
Refer to caption
Figure 9: The angular-integrated total, Coulomb and nuclear breakup cross sections are given for the 8Li+208Pb breakup reaction as functions of εb\varepsilon_{b}, with nuclear absorption in the panels (a) and (b); and without absorption in the panels (c) and (d). As indicated, the incident energies are Elab=E_{lab}=36 MeV [panels (a) and (c)] and 60 MeV [panels (b) and (d)].

For a better quantitative assessment of these results, we consider the integrated total (σtot\sigma_{tot}), Coulomb (σCoul\sigma_{Coul}), and nuclear (σnucl\sigma_{nucl}) breakup cross sections, which are displayed as functions of εb\varepsilon_{b} in Fig. 8 (for the C12{}^{12}{\rm C} target), and in Fig. 9 (for the Pb208{}^{208}{\rm Pb} target). The results in both figures confirm the conclusions already drawn from Figs.4 and 6. For example, both panels of Fig.8, show that as εb\varepsilon_{b}\to 0.01 MeV, σCoul>σnucl\sigma_{Coul}>\sigma_{nucl} (σCoulσtot\sigma_{Coul}\simeq\sigma_{tot}), whereas σCoul<σnuclσtot\sigma_{Coul}<\sigma_{nucl}\simeq\sigma_{tot} (σnuclσtot\sigma_{nucl}\simeq\sigma_{tot}) as εb2.03\varepsilon_{b}\to 2.03 MeV. For Pb208{}^{208}{\rm Pb} target, the results are shown in the presence of nuclear absorption. When the nuclear absorption is taken into account [panels (a) and (b)], we notice that σcoulσtotσnucl\sigma_{coul}\simeq\sigma_{tot}\gg\sigma_{nucl} and this is independent of εb\varepsilon_{b}. In the absence of the nuclear absorption [panels (c) and (d)], while σcoulσtotσnucl\sigma_{coul}\simeq\sigma_{tot}\gg\sigma_{nucl} remains valid for εb0.01\varepsilon_{b}\to 0.01 MeV, it is noticed that σtotσnucl>σCoul\sigma_{tot}\simeq\sigma_{nucl}>\sigma_{Coul}, for εb2.03\varepsilon_{b}\to 2.03 MeV, which further highlights the importance of the nuclear absorption for large binding energies. The results in this figure further support the fact that strong nuclear absorption in the inner region is the main factor that dictates the importance of the Coulomb breakup cross section over its nuclear counterparts.

Table 3: Coulomb, nuclear and interference cross-sections for the Li8+C12{}^{8}{\rm Li}+{}^{12}{\rm C} and 8Li+208+^{208}Pb, considering n7-^{7}Li binding energies εb=\varepsilon_{b}=0.01 MeV and 2.03 MeV. For each target, we present our results, in terms of ratios, for two colliding energies. For 208Pb target, with no-nuclear absorption (NA) the results are shown within parenthesis below the ones with absorption.
Target Elab{E_{lab}} εb=\varepsilon_{b}= 2.03 MeV εb=\varepsilon_{b}= 0.01 MeV
(MeV) σCoulσtot\frac{\sigma_{Coul}}{\sigma_{tot}} σnuclσtot\frac{\sigma_{nucl}}{\sigma_{tot}} σCoulσnucl\frac{\sigma_{Coul}}{\sigma_{nucl}} σintσnucl\frac{\sigma_{int}}{\sigma_{nucl}} σintσtot\frac{\sigma_{int}}{\sigma_{tot}} σCoulσtot\frac{\sigma_{Coul}}{\sigma_{tot}} σnuclσtot\frac{\sigma_{nucl}}{\sigma_{tot}} σCoulσnucl\frac{\sigma_{Coul}}{\sigma_{nucl}} σintσnucl\frac{\sigma_{int}}{\sigma_{nucl}} σintσtot\frac{\sigma_{int}}{\sigma_{tot}}
C12{}^{12}{\rm C} 14 0.024 0.824 0.029 0.186 0.153 0.836 0.268 3.123 -0.387 -0.104
24 0.048 0.808 0.059 0.190 0.154 0.701 0.339 2.069 -0.118 -0.040
Pb208{}^{208}{\rm Pb} 36 1.800 0.150 12.00 -6.333 -0.950 1.033 0.012 90.16 -3.850 -0.044
(0.344) (1.481) (0.232) (-0.557) (-0.825) (1.032) (0.029) (33.97) (-2.061) (-0.063)
60 1.326 0.087 15.25 -4.750 -0.413 1.015 0.010 104.6 -2.540 -0.025
(0.198) (0.783) (0.253) (0.025) (0.019) (1.000) (0.059) (17.03) (-0.991) (-0.058)

In Table 3, we provide more quantitative results, given as fractions from σtot\sigma_{tot} and σnucl\sigma_{nucl}, reflecting the competition between the different cross sections, by selecting the two limiting binding energies we are studying, i.e., εb=\varepsilon_{b}=0.01 MeV and εb=\varepsilon_{b}=2.03 MeV. We are also including σint\sigma_{int}, as defined by

σint=σtot(σCoul+σnucl),\displaystyle\sigma_{int}=\sigma_{tot}-(\sigma_{Coul}+\sigma_{nucl}), (19)

which we naively regard as the Coulomb-nuclear interference and that will be discussed in the next subsection. From this table, it becomes evident that, when εb\varepsilon_{b} decreases, σCoul\sigma_{Coul} (approaching to σtot\sigma_{tot}) becomes substantially larger than σnucl\sigma_{nucl}. Also, for the light 12C target, at Elab=14E_{lab}=14 MeV and 24 MeV, we note that σCoul/σnucl\sigma_{Coul}/\sigma_{nucl} rapidly grows, when varying εb\varepsilon_{b} from 2.03 MeV down to 0.01 MeV. As shown, in this energy interval, σCoul/σnucl\sigma_{Coul}/\sigma_{nucl} increases from 0.03 to 3.12 for 14 MeV, and from 0.06 to 2.10 for 24 MeV. This indicates that, as the binding energy decreases, the 8Li+12+^{12}C reaction becomes like a “Coulomb-dominated reaction”, with the emergence of a long-range behavior. Moreover, with the heavy target at Elab=E_{lab}=36 MeV, in the presence of nuclear absorption, for εb=\varepsilon_{b}=2.03 MeV, σcoul/σnucl=12\sigma_{coul}/\sigma_{nucl}=12, whereas σcoul/σnucl90\sigma_{coul}/\sigma_{nucl}\simeq 90 for εb=\varepsilon_{b}= 0.01 MeV. It is noticed in this case that this ratio is substantially affected in the absence of nuclear absorption (NA), becoming σcoul/σnucl0.06\sigma_{coul}/\sigma_{nucl}\simeq 0.06 (εb=2.03\varepsilon_{b}=2.03 MeV), and σcoul/σnucl34\sigma_{coul}/\sigma_{nucl}\simeq 34 (εb=0.01\varepsilon_{b}=0.01 MeV).

Refer to caption
Figure 10: The 8Li+12C integrated Coulomb-nuclear interference σint\sigma_{int} [panel (a)], given by (19), with the respective ratio σint/σtot\sigma_{int}/\sigma_{tot} [panel (b)], are shown as functions of εb\varepsilon_{b}, for the colliding energies Elab=E_{\rm lab}=14 and 24 MeV.
Refer to caption
Figure 11: The Li8+Pb208{}^{8}{\rm Li}+{}^{208}{\rm Pb} integrated Coulomb-nuclear interference σint\sigma_{int} [panels (a) and (b)], given by (19), with their ratios σint/σtot\sigma_{int}/\sigma_{tot} [panels (c) and (d)], are shown as functions of εb\varepsilon_{b}, for Elab=E_{\rm lab}=36 and 60 MeV (upper and lower frames, respectively). σintWA\sigma_{int}^{\rm WA} (solid lines) denotes the interference when the breakup is followed by nuclear absorption, with σintNA\sigma_{int}^{\rm NA} (dot-dashed lines) denoting interference with no nuclear absorption.

III.3 Coulomb-nuclear interference

It is well-known that the incoherent sum of the Coulomb and nuclear breakup cross section (σCoul+σnucl)(\sigma_{Coul}+\sigma_{nucl}) is always different from their coherent sum, σtot\sigma_{tot}, due to the Coulomb-nuclear interference effect. To assess this effect in the context of very weak ground-state binding energy, we consider σint\sigma_{int} as defined in Eq. (19) to estimate the Coulomb-nuclear interference. Given that, for the two limiting binding energies, the quantitative results for σint\sigma_{int} are already furnished in Table 3 as ratios with respect to σtot\sigma_{tot} and σnucl\sigma_{nucl}. In Figs. 10 and 11 (respectively, for 12C and 208Pb targets), we provide the exact σint\sigma_{int} behaviors, together with their respective ratios σint/σtot\sigma_{int}/\sigma_{tot}, as functions of εb\varepsilon_{b}, in a way to clarify that the differences between σtot\sigma_{tot} and (σCoul+σnucl)(\sigma_{Coul}+\sigma_{nucl}) are quite large in both the cases, with the amount varying with ElabE_{lab} (|σint||\sigma_{int}| decreasing with increasing ElabE_{lab}). The Coulomb-nuclear interference is strongly dependent on εb\varepsilon_{b}. As one can notice, it appears to increase as εb\varepsilon_{b} decreases, and becomes quite small as εb2.03\varepsilon_{b}\to 2.03 MeV.

For the 8Li+208Pb reaction, nuclear absorption which is already shown to reduce the breakup cross section (Fig.7), is expected to be more relevant on the Coulomb-nuclear interference. The Coulomb-nuclear interference obtained when the breakup is followed by nuclear absorption (i.e., Wctnucl0,Wntnucl0W_{ct}^{nucl}\neq 0,W_{nt}^{nucl}\neq 0), is denoted by σintWA\sigma_{int}^{\rm WA} (WA standing for “with absorption”), and by σintNA\sigma_{int}^{\rm NA} the Coulomb-nuclear interference obtained when Wctnucl=Wntnucl=0W_{ct}^{nucl}=W_{nt}^{nucl}=0. Therefore, in order to assess the relevance of the nuclear absorption on this interference, we compare σintWA\sigma_{int}^{\rm WA} with σintNA\sigma_{int}^{\rm NA}. The results are presented in Fig.11. In this figure, the panels (a) and (b) are for the exact σint\sigma_{int} results, whereas in panels (c) and (d) we have the respective ratios σint/σtot\sigma_{int}/\sigma_{tot}. The upper panels are for Elab=36E_{lab}=36 MeV, and the lower panels for Elab=60E_{lab}=60 MeV. The absorption contribution to σint\sigma_{int} is verified by the observed difference |σintNAσintWA||\sigma_{int}^{\rm NA}-\sigma_{int}^{\rm WA}|, which are clearly shown for both ElabE_{lab} energies, as εb\varepsilon_{b} varies.

Besides the fact that the Coulomb-nuclear interference is shown to be larger in the very small binding energy limits, such larger values may also be influenced by the large magnitudes of the total and Coulomb breakup cross sections, which are shown in Figs. 9 and 10. However, as verified from Table 3, the ratios σCoul/σtot\sigma_{Coul}/\sigma_{tot} for the smaller binding are enough deviating from one (when full absorption is considered, in the 208Pb case). Consistently, we also noticed from the results given in Table 3, that σint/σtot\sigma_{int}/\sigma_{tot} is larger for εb=2.03\varepsilon_{b}=2.03 MeV when we have the usual cross section values with absorption. Further investigation may be required to clarify the εb\varepsilon_{b} dependence of Coulomb-nuclear interference, in support to the actual results that are shown an overall significant effect of nuclear absorption.

In the case of such weakly-bound projectiles, a better understanding of the function δR(εb)\delta_{R}(\varepsilon_{b}), which appears in Eq.(4), could shed more light on the complexity of the Coulomb-nuclear interference. In such cases, RnR_{n} can significantly deviate from R0R_{0}, since the nuclear breakup dynamics requires that δR(εb)0\delta_{R}(\varepsilon_{b})\to 0 for larger values of εb\varepsilon_{b}. Particularly, the main characteristics of this function could show up in a study with charged projectiles, considering that strong Coulomb/nuclear interference has been observed for the reaction of proton halo 8B with 58Ni target 2001Tostevin ; 2002Margueron ; Tarutina10 ; 2009Lubian , in which we have a very weakly-bound projectile with breakup threshold of 0.137 MeV.

IV Conclusion

We have presented a study on the breakup of the weakly-bound 8Li (n7-^{7}Li) projectile on light and heavy target masses, namely, 12C and 208Pb. Our main objective was to investigate the dependence of the total, Coulomb and nuclear breakup cross sections, on the 8Li ground-state binding energy εb\varepsilon_{b}, in order to study the peripherality of the total, Coulomb and nuclear breakup processes, which are associated to the weaker binding energy of the projectile. To this end, apart from the experimentally-known ground-state binding energy of the n7-^{7}Li system, we artificially considered four other binding energies, below the experimental value, down to εb=0.01\varepsilon_{b}=0.01 MeV, which is much smaller than the experimental value, εb=2.03\varepsilon_{b}=2.03 MeV. From our analysis it is shown that the total, Coulomb and nuclear breakup processes become peripheral as εb0.01\varepsilon_{b}\to 0.01 MeV, regardless the target mass. We argue that the peripherality of the nuclear breakup in this case is primarily related to the spacial extension of the corresponding ground-state wave function, which is related to weaker binding energy. The peripheral region is determined by the range of the nuclear forces R0R_{0}, and the corresponding extension of the ground-state wave function, which is associated to a function δR(εb)\delta_{R}(\varepsilon_{b}), expressed by RnR_{n} defined in Eq. (4). By taking into account the fact that close to the nn-core εb0\varepsilon_{b}\to 0 limit, a long-range interaction is expected to emerge between projectile and target (similar as for three-body halo-nuclei systems 2012Frederico ), the size of the associated wave function will increase significantly in this limit. So, a detailed investigation of this function δR(εb)\delta_{R}(\varepsilon_{b}) (which should go to zero by increasing εb\varepsilon_{b}) can shed more light into the dynamics of nuclear breakups induced by loosely bound projectiles. It is also noticed that the variation of εb\varepsilon_{b} strongly affects the Coulomb breakup, as compared to the nuclear breakup, such that as εb\varepsilon_{b}\to 0.01 MeV, the Coulomb breakup becomes dominant even for the C12{}^{12}{\rm C} target, which is known to be naturally dominated by nuclear breakup. Therefore, in view of this binding-energy dependence, one may infer that the expression “naturally-dominated by nuclear breakup” may be relative to the projectile binding energy. It is also verified that the nuclear absorption has an insignificant effect on the total and nuclear breakup cross sections when decreasing the binding energy to small binding such as εb\varepsilon_{b}\to 0.01 MeV. In this small binding energy region, we found that the total breakup cross section is larger than the calculated total fusion cross section, while as expected, the opposite is observed as εb\varepsilon_{b}\to 2.03 MeV.

Acknowledgements

We thank T. Frederico, B.V. Carlson and L. F. Canto for useful discussions. B.M. is also grateful to the South American Institute of Fundamental Research (ICTP-SAIFR) for local facilities. For partial support, we also thank Conselho Nacional de Desenvolvimento Científico e Tecnológico [INCT-FNA Proc.464898/2014-5 (LT and JL), Proc. 304469/2019-0(LT) and Proc. 306652/2017-0(JL)], and Fundação de Amparo à Pesquisa do Estado de São Paulo [Projs. 2017/05660-0(LT)].

References

  • (1) Y. Suzuki, K. Yabana, R.G. Lovas, K. Varga, Structure and Reactions of Light Exotic Nuclei (Taylor & Francis Group, London and New York, 2003).
  • (2) I. J. Thompson and F. M. Nunes, Nuclear Reactions for Astrophysics (Cambridge University Press, New York, 2009).
  • (3) R. Chatterjee and R. Shyam, Prog. Part. Nucl. Phys. 103, 67 (2018).
  • (4) F. M. Nunes and I. J. Thompson, Phys. Rev. C 57, R2818 (1998).
  • (5) F. M. Nunes and I. J. Thompson, Phys. Rev. C 59, 2652 (1999).
  • (6) J. Margueron, A. Bonaccorso, and D. M. Brink, Nucl. Phys. A 703, 105 (2002).
  • (7) P. Capel, D. Baye and V. S. Melezhik, Phys. Rev. C 68, 014612 (2003).
  • (8) T. Tarutina and M. S. Hussein, Phys. Rev. C 70, 034603 (2004).
  • (9) M. S. Hussein, R. Lichtenthaler, F. M. Nunes and I. J. Thompson, Phys. Lett. B 640, 91 (2006).
  • (10) L. F. Canto, P. R. S. Gomes, R. Donangelo, and M. S. Hussein, Phys. Rep. 424, 1 (2006).
  • (11) J. Lubian, T. Correa, E. F. Aguilera, L. F. Canto, A. Gomez-Camacho, E. M. Quiroz, and P. R. S. Gomes, Phys. Rev. C 79, 064605 (2009).
  • (12) L. F. Canto, J. Lubian, P. R. S. Gomes, and M. S. Hussein, Phys. Rev. C 80, 047601 (2009).
  • (13) Y. Kucuk and A. M. Moro, Phys. Rev. C 86, 034601 (2012).
  • (14) P. Capel, J. Phys. G: Nucl. Part. Phys. 41, 094002 (2014).
  • (15) D. R. Otomar, P. R. S. Gomes, J. Lubian, L. F. Canto and M. S. Hussein, Phys. Rev. C 87, 014615 (2013).
  • (16) D. R. Otomar, P. R. S. Gomes, J. Lubian, L. F. Canto and M. S. Hussein, Phys. Rev. C 92, 064609 (2015).
  • (17) B. Mukeru, M. L. Lekala and A. S. Denikin, J. Phys. G: Nucl. Part. Phys. 42, 015109 (2015).
  • (18) B. Mukeru and M. L. Lekala, Phys. Rev. C 91, 064609 (2015).
  • (19) S. G. Manjeet, Chin. Phys. C 40, 054101 (2016).
  • (20) P. Descouvemont, L.F. Canto, M.S. Hussein, Phys. Rev. C 95, 014604 (2017).
  • (21) B. Mukeru and M.L. Lekala, Int. J. Mod. Phys. E 26,1750075 (2017).
  • (22) B. Mukeru, T. Frederico, and L. Tomio, Phys. Rev. C 102, 064623 (2020).
  • (23) T. Nakamura, N. Kobayashi, Y. Kondo,Y. Satou, N. Aoi, H. Baba, et al., Phys. Rev. Lett. 103, 262501 (2009).
  • (24) C. Nociforo et al., Phys. Lett. B 605, 79 (2005).
  • (25) T. Aumann and T. Nakamura, Phys. Scr. T 152, 014012 (2013).
  • (26) N. Fukuda, T. Nakamura,N. Aoi,N. Imai, M. Ishihara,T. Kobayashi, H. Iwasaki, T. Kubo, A. Mengoni, M. Notani, H. Otsu,H. Sakurai, S. Shimoura, T. Teranishi,Y. X. Watanabe,K. Yoneda, Phys. Rev. C 70, 054606 (2004).
  • (27) B. Abu-Ibrahim and Y. Suzuki, Progr. Theor. Phys. 112, 1013 (2004).
  • (28) C. A. Bertulani and G. Baur, Phys. Rep. 163, 299 (1988).
  • (29) A. Winther and K. Alder, Nucl. Phys. A 319, 518 (1979).
  • (30) G. Baur, K. Hencken, and D. Trautmann, Prog. Part. Nucl. Phys. 51, 487 (2003).
  • (31) G. Baur and H. Rebel, Annu. Rev. Nucl. Part. Sci. 46, 321 (1996).
  • (32) B. Wang, W. J. Zhao, P. R. S. Gomes, E. G. Zhao, and S.-G. Zhou, Phys. Rev. C 90, 034612 (2014).
  • (33) P. K. Rath, S. Santra, N. L. Singh, R. Tripathi, V. V. Parkar, B. K. Nayak, K. Mahata, R. Palit, S. Kumar, S. Mukherjee, S. Appannababu, R. K. Choudhury, Phys. Rev. C 79, 051601(R) (2009).
  • (34) J. Rangel, J. Lubian, L. F. Canto, and P. R. S. Gomes, Phys. Rev. C 93, 054610 (2016).
  • (35) J. Lei A. M. Moro, Phys. Rev. Lett. 122, 042503 (2019).
  • (36) B. Mukeru, J. Phys. G: Nucl. Part. Phys. 45, 065201 (2018).
  • (37) B. Mukeru, M.L. Lekala, J. Lubian and L. Tomio, Nucl. Phys. A 996, 121700 (2020).
  • (38) V. Jha, V. V. Parkar and S. Kailas, Phys. Rep. 845 1 (2020).
  • (39) K. J. Cook, I. P. Carter, E. C. Simpson, M. Dasgupta, D. J. Hinde,L. T. Bezzina, S. Kalkal, C. Sengupta, C. Simenel, B. M. A. Swinton-Bland, K. Vo-Phuoc,E. Williams, Phys. Rev. C 97 021601(R) (2018).
  • (40) A. Pakou et al., Eur. Phys. J. A 51, 55 (2015).
  • (41) V. Guimarães, J. Lubian, J. J. Kolata, E. F. Aguilera, M. Assunção, V. Morcelle, Eur. Phys. J. A 54, 223 (2018).
  • (42) M. Wang, et al., Chin. Phys. C 41, 030003 (2017). [See also at https://www.nndc.bnl.gov/nudat2/].
  • (43) N. Austern et al., Phys. Rep. 154,125 (1987).
  • (44) I. J. Thompson, Comput. Phys. Rep. 7, 167 (1988)
  • (45) L. F. Canto, P. R. S. Gomes, R. Donangelo, J. Lubian and M. S. Hussein, Phys. Rep. 596, 1 (2015).
  • (46) M. Abramowitz and I. Stegun, Handbook of Mathematical Functions: with Formulas,Graphs, and Mathematical Tables (Dover Publications, National Bureau of Standards, New York, 1964).
  • (47) L. F. Canto and M. S. Hussein, Scattering theory of molecules, atoms and nuclei (World Scientific Publishing Co. Pte. Ltd, Singapore, 2013)
  • (48) C. A. Bertulani, Comput. Phys. Commun. 156, 123 (2003).
  • (49) A. M. Moro, R. Crespo, H. Garcia-Martinez, E. F. Aguilera, E. Martinez-Quiroz, J. Gomez-Camacho, F. M. Nunes, Phys. Rev. C 68, 034614 (2003).
  • (50) G. W. Fan, M. Fukuda, D. Nishimura, X. L. Cai, S. Fukuda, I. Hachiuma, et al., Phys. Rev. C 91, 014614 (2015).
  • (51) I. Tanihata, T. Kobayashi, O. Yamakawa, S. Shimoura, K. Ekuni, K. Sugimoto, N. Takahashi, T. Shimoda, H. Sato, Phys. Lett. B206, 592 (1988).
  • (52) A. Barioni, V. Guimarães, A. Lépine-Szily, R. Lichtenthäler, D. R. Mendes, E. Crema, K. C. C. Pires, M.C. Morais, V. Morcelle, P. N. deFaria, R. P. Condori., A. M. Moro., D. S. Monteiro, J. M. B. Shorto, J. Lubian, M. Assunção, Phys. Rev. C 80, 034617 (2009).
  • (53) J. Cook, Nucl. Phys. A 388, 153 (1982).
  • (54) A. J. Koning and J. P. Delaroche, Nucl. Phys.A 713, 231 (2003).
  • (55) F. D. Becchetti, W. Z. Liu, K. Ashktorab, J. F. Bajema, J. A. Brown, J. W. Janecke, D. A. Roberts, J. J. Kolata, K. L. Lamkin, A. Morsad, R. J. Smith, X. J. Kong, R. E. Warner, Phys. Rev. C 48, 308 (1993).
  • (56) A. V. Karpov, A. S. Denikin, A. P. Alekseev, V. I. Zagrebaev, V. A. Rachkov, M. A. Naumenko, and V. V. Saiko, Phys. Atom. Nucl. 79,749 (2016) and http://nrv.jinr.ru/nrv/webnrv/expdata database.
  • (57) J. J. Kolata, V. Z. Goldberg, L. O. Lamm, M. G. Marino, C. J. OKeeffe, G. Rogachev, et al., Phys. Rev. C 65, 054616 (2002).
  • (58) J. A. Tostevin, F. M. Nunes, and I. J. Thompson, Phys. Rev. C 63, 024617 (2001).
  • (59) T. Frederico, A. Delfino, L. Tomio and M.T. Yamashita, Prog. Part. Nucl. Phys. 67, 939 (2012).