Brought to you by:
Paper

Isospin-forbidden electric dipole capture and the α(d, γ)6Li reaction

and

Published 28 June 2018 © 2018 IOP Publishing Ltd
, , Citation D Baye and E M Tursunov 2018 J. Phys. G: Nucl. Part. Phys. 45 085102 DOI 10.1088/1361-6471/aacbfa

0954-3899/45/8/085102

Abstract

At the long-wavelength approximation, E1 transitions are forbidden between isospin-zero states. Hence E1 radiative capture is strongly hindered in reactions involving N = Z nuclei but the E1 astrophysical S factor may remain comparable to, or larger than, the E2 one. Theoretical expressions of the isoscalar and isovector contributions to E1 capture are analyzed in microscopic and three-body approaches in the context of the α(d, γ)6Li reaction. The lowest non-vanishing terms of the operators are derived and the dominant contributions to matrix elements are discussed. The astrophysical S factor computed with some of these contributions in a three-body α + n + p model is in agreement with the recent low-energy experimental data of the LUNA collaboration. This confirms that a correct treatment of the isovector E1 transitions involving small isospin-one admixtures in the wave functions should be able to provide an explanation of the data without adjustable parameter. The exact-masses prescription which is often used to avoid the disappearance of the E1 matrix element in potential models is not found at the microscopic level and should not be used for reactions of N = Z nuclei. The importance of capture components from an initial S scattering wave is also discussed.

This is a board member communicated paper

Export citation and abstract BibTeX RIS

1. Introduction

In some radiative capture reactions between light nuclei, electric dipole transitions are strongly suppressed [1]. This effect is due to an isospin selection rule: E1 transitions are isospin-forbidden in capture reactions involving N = Z nuclei [2].

At the long-wavelength approximation, which is a good approximation for this type of reactions, the isoscalar part of the E1 operator vanishes and transitions take place via its isovector part. Matrix elements of isovector operators vanish between isospin-zero states. However, except for the deuteron, realistic wave functions of N = Z nuclei are not pure eigenstates of the isospin operator and E1 transitions are not exactly forbidden. Their strength may keep an order of magnitude similar to the strength of the usually much weaker E2 transitions. This effect is particularly spectacular for the 12C(α, γ)16O reaction where the isospin-forbidden E1 component is enhanced by resonances (see reference in [1]). Disentangling the E1 and E2 strengths is experimentally very difficult and the theoretical calculations of the E1 component are still quite uncertain. The role of E1 transitions is also complicated in other reactions of astrophysical interest such as d(d, γ)4He, 4He(d, γ)6Li, and 16O(α, γ)20Ne. It may also play some role in the triple α mechanism generating 12C.

An ab initio description of the two lightest cases is in principle possible at present. The astrophysical S factor of the d(d, γ)4He reaction has been computed with an ab initio calculation in [3]. The E1 component is mainly obtained from T = 1 isospin components in 4He introduced by coupled p + 3H and n + 3He configurations. Its largest contribution reaches at most 4% near the center-of-mass energy 0.01 MeV and thus remains quite small with respect to E2 [4]. For the 4He(d, γ)6Li reaction, the problem is more difficult because of the larger numbers of nucleons and of possible configurations. An ab initio study of the α + d elastic scattering has been performed in [5] with a realistic nucleon–nucleon (NN) force. A study of the E2 capture component could be based on that work but the study of the E1 component would require much additional computer time with the introduction of T = 1 isospin components in the initial and final wave functions. Such a calculation is thus not available yet.

Preliminary attempts to calculate isospin-forbidden E1 cross sections for heavier systems have been performed in microscopic cluster models. In [6], an α cluster with a small T = 1 component in its ground state has been used to explore E1 capture in the 16O(α, γ)20Ne reaction but a similar component would at least have been necessary in the 16O cluster. In [7], E1 capture in the 12C(α, γ)16O reaction was studied by coupling 12C + α configurations with 15N + p and 15O + n configurations which introduced some T = 1 contributions in the 16-nucleon wave functions but some properties of the E1 resonances had to be modified phenomenologically. These attempts provide qualitative information but remain too limited for quantitative predictions.

Since realistic microscopic calculations are not available yet, most calculations of isospin-forbidden E1 capture have been performed in the two-body or potential model based on the cluster idea. The isospin quantum number does not appear in this model. The nuclei are only represented by their atomic numbers Z1 and Z2, their mass numbers A1 and A2, and their spin and parity quantum numbers. The physics arises from the interaction between them. Electric dipole transitions are nevertheless forbidden because of the presence of a factor Z1/A1 − Z2/A2 in E1 transition matrix elements, which vanishes for N = Z colliding nuclei since both ratios Z1/A1 and Z2/A2 are equal to 1/2. Indeed, this factor in the effective E1 operator is of microscopic origin and thus involves integer mass numbers.

In order to have a non-vanishing E1 astrophysical S factor, the traditional prescription is to replace the integer mass numbers A1 and A2 by non-integer values deduced from the experimental masses of the colliding nuclei. This replacement can be justified in a two-body non-microscopic cluster model. It leads to a non-vanishing dipole moment of the nucleus in the cluster picture. This 'exact-masses' prescription, however, has no microscopic foundation at the nucleon level. As discussed below, it may give a plausible order of magnitude for the capture cross section but the possible agreement or disagreement with experimental data has no physical meaning. The energy dependence of the cross section may also be plausible but is not founded microscopically.

In this paper, we discuss various theoretical aspects of the forbidden E1 transitions. To fix ideas, we take the $\alpha +d\,\to $ 6Li + γ capture process as an example. This reaction was first studied experimentally at energies around and above the 0.711 MeV 3+ resonance [8, 9]. Until recently, the lower-energy data resulted from indirect measurements with Coulomb breakup reactions of 6Li on lead [10, 11]. The presence of nuclear breakup makes difficult the extraction of information on radiative capture from the data. Recently, the α(d, γ)6Li reaction was studied at the LUNA facility by direct measurements at the two astrophysical energies 94 and 134 keV [12].

From the theoretical side, calculations of astrophysical S factors have been developed within different two-body potential models [1321], three-body potential models [2224], and with semi-microscopic [25, 26] and microscopic [27, 28] models. Early models focused on the then existing data [8] at energies around and beyond the 3+ resonance where the main contribution to the capture process comes from E2 transitions. At low energies, the dominant contribution is expected to come from the E1 transition operator since the E2 cross section is smaller than the data in all models. The recent LUNA data have renewed the interest for theoretical calculations of the S factor at astrophysical energies [20, 21, 24].

In the theoretical literature, the E1 capture is treated in various ways, but the exact-masses prescription is in general used in potential models [14, 1621, 23, 24] and even in partly microscopic approaches [2527], sometimes combined with various other corrections. These calculations raise questions about the foundation of the exact-masses prescription and about the validity of its combination with other corrections.

The aim of the present study is to discuss theoretical aspects of the forbidden E1 transitions and question the validity of the exact-masses prescription. We analyze theoretically different contributions to the E1 S factor of the α(d, γ)6Li capture process and emphasize the main ones that should be necessarily included in a realistic model. A model able to take all these contributions into account in a consistent way is beyond our reach. We evaluate some of these contributions to the S factor and discuss their importance with the simplest model where an isospin mixing effect appears, the three-body α + n + p model of [24]. This allows us to suggest key points that should be studied in future model calculations.

In section 2, the microscopic expression of the electric dipole operator and the corresponding matrix elements for isospin-forbidden transitions are presented. In section 3, the expressions are specialized to a three-body model. The initial wave function is the product of a two-body deuteron wave function and an α + d scattering wave function. The final 6Li(1+) ground state is described with an α + n + p three-body wave function in hyperspherical coordinates [29, 30]. The model involves n + p, α + n/p, and α + d potentials. In section 4, results are presented and commented. The exact-masses prescription is discussed in section 5 as well as the possible role of capture from an initial S wave. Section 6 is devoted to a conclusion.

2. Microscopic treatment of isospin-forbidden E1 transitions

2.1. Microscopic electric multipole operators

Since the energies of the emitted photons are usually not large at astrophysical energies, their wavelengths are large with respect to typical dimensions of the system and the photon wavenumbers

Equation (1)

can be considered as small. The long-wavelength approximation can be used. Let  rj be the coordinate of the jth nucleon. At the long-wavelength approximation, the translation-invariant electric transition operators of multipolarity λ are given to a good approximation by

Equation (2)

where tj3 is the third component of the isospin operator ${{\boldsymbol{t}}}_{j}$ of the jth nucleon related to its charge by $e(\tfrac{1}{2}-{t}_{j3})$, and

Equation (3)

is its coordinate with respect to the center-of-mass

Equation (4)

of the A-nucleon system. The functions Yλμ (${{\rm{\Omega }}}_{j}^{{\prime} }$) are spherical harmonics depending on the angular part of ${{\boldsymbol{r}}}_{j}^{{\prime} }=(r{{\prime} }_{j},{\rm{\Omega }}{{\prime} }_{j})$.

The orbital angular momentum with respect to the center-of-mass and spin of nucleon j are denoted as ${\boldsymbol{L}}{{\prime} }_{j}$ and ${{\boldsymbol{S}}}_{j}$, respectively. The total orbital momentum operator of the system is ${\boldsymbol{L}}={\sum }_{j=1}^{A}{\boldsymbol{L}}{{\prime} }_{j}$, the total spin is ${\boldsymbol{S}}={\sum }_{j=1}^{A}{{\boldsymbol{S}}}_{j}$ and the total angular momentum is ${\boldsymbol{J}}={\boldsymbol{L}}+{\boldsymbol{S}}$. The total isospin operator of the system is ${\boldsymbol{T}}={\sum }_{j=1}^{A}{{\boldsymbol{t}}}_{j}$.

The operators defined by equation (2) contain isoscalar (IS) and isovector (IV) parts. At the long-wavelength approximation, the E1 operator is special. It mainly contains an isovector component,

Equation (5)

The lowest-order term of the isoscalar part vanishes since ${\sum }_{j=1}^{A}{{\boldsymbol{r}}}_{j}^{{\prime} }=0$. This operator connects eigenstates of the total isospin operator with initial and final isospin quantum numbers differing by one unit, ${T}_{{\rm{f}}}=| {T}_{{\rm{i}}}\pm 1| $. It also connects states with Ti = Tf, but only for $N\ne Z$. Transitions from Ti = 0 to Tf = 0 are forbidden.

The isoscalar part of the E1 operator is however not exactly zero. It might play a non-negligible role in some cases. The first non-vanishing term reads using the Siegert theorem [31]

Equation (6)

where mp is the proton mass, and gp and gn are the proton and neutron gyromagnetic factors, respectively. The vector function $[{\boldsymbol{L}}{Y}_{1\mu }]({\rm{\Omega }})$ is the result of the action of the orbital momentum operator on the spherical harmonics Y1μ (Ω) with l = 1. This operator connects components with the same initial and final isospins, Ti = Tf. When it acts on a wave function with a largely dominant component with zero total orbital momentum and small intrinsic spin, the first term of equation (6) should give a reasonable approximation.

2.2. Transition matrix elements

We consider transitions in N = Z systems between an initial scattering state and a final bound state with dominant zero-isospin components. Their wave functions can be written symbolically as

Equation (7)

The T = 1 components ${{\rm{\Psi }}}_{i,f}^{{JM};1}$ are much smaller than the T = 0 components ${{\rm{\Psi }}}_{i,f}^{{JM};0}$. Possible admixtures of larger isospin values are neglected.

To a good approximation, three types of matrix elements must be calculated. Two of them involve an isovector transition, i.e., between the dominant Ti = 0 component in the initial scattering state and the Tf = 1 admixture in the final bound state

Equation (8)

and between the Ti = 1 admixture in the initial scattering state and the dominant Tf = 0 component in the final bound state

Equation (9)

An isoscalar transition is also possible, essentially between the dominant components,

Equation (10)

The E1 transition matrix element is the coherent sum of these three contributions.

2.3. α(d, γ)6Li E1 capture in resonating-group notation

To fix ideas we consider the α(d, γ)6Li reaction. We use the notation of the resonating-group method (RGM) [32, 33]. This notation is also valid for ab initio descriptions. We limit ourselves to α + n + p configurations. Realistic calculations might also include 3H + 3He configurations, for example, that we neglect to simplify the presentation. The wave functions that we now describe display the main components expected to play a significant role in E1 transitions. Many other smaller components are of course possible.

In the RGM, a partial wave of the initial scattering wave function (7) is written as

Equation (11)

where ${ \mathcal A }$ is the six-nucleon antisymmetrizer and ${\boldsymbol{R}}=(R,{{\rm{\Omega }}}_{R})$ is the relative coordinate between the centers of mass of the α and deuteron clusters. The functions ${\phi }_{\alpha }^{00+}$ and ${\phi }_{d}^{1m+}$ are translation-invariant internal wave functions of the ground states of the 4He nucleus with angular momentum 0 and positive parity and of the deuteron with angular momentum 1 and positive parity, respectively. The 4He wave function depends on three internal coordinates. The deuteron wave function depends on the relative coordinate ${\boldsymbol{r}}=(r,{{\rm{\Omega }}}_{r})$ between the proton and neutron. The total parity π is equal to ( − 1)L. The 4He ground state internal wave function may contain a small T = 1 admixture

Equation (12)

The T = 1 component is mainly due to the Coulomb interaction between the protons. The neutron–proton mass difference and isospin non-conserving terms in the nuclear force also contribute but to a lesser extent. The deuteron ground state wave function is purely T = 0. In reactions of α particles with heavier N = Z nuclei, a T = 1 admixture also appears in the second cluster.

Various corrections may also appear in the scattering wave function to take distortion of the initial state at short distances into account. They may involve sums over pseudo-states of the deuteron and/or of the α particle. The most important ones should arise from deuteron pseudo-states which can simulate its Coulomb polarizability [15]. They may also include additional shell-model-like 6Li terms [32]. We do not display these corrections here to simplify the discussion but they can be treated in the same way as similar terms displayed below in the final state.

Under some simplifying assumptions, the main components of the final bound state wave function of the 1+ ground state of 6Li can be approximated as

Equation (13)

The ${\phi }_{{d}^{* }n}^{1{\pi }_{n};{T}_{n}}$ with Tn = 0 or 1 are excited pseudo-states of the deuteron. The relative orbital momentum is Ln = 0 for πn = + and Ln = 1 for πn = −. The ${\phi }_{{\alpha }^{* }n}^{1-;1}$ are excited pseudo-states of the 4He nucleus with angular momentum 1 and isospin 1. The channel spin I can take the values 0–2.

Given the angular momentum and parity of the final state, the initial state for E1 transitions corresponds to J = 0–2 and a negative parity. This is realized by choosing L = 1 in equation (11). Within these assumptions, let us write the various matrix elements. Matrix element (8) reads for an initial wave with L = 1,

Equation (14)

and matrix element (9) reads

Equation (15)

where J can be equal to 0–2. Other contributions appear when the initial state is distorted. Matrix element (10) reads

Equation (16)

As the operator is much smaller here, only the dominant T = 0 components are kept.

3. Three-body model of isospin-forbidden E1 transitions

3.1. Three-body operators

We now consider the three-body α + n + p model. The 4He nucleus is treated as a structureless particle. Its properties appear in the interaction with the nucleons. They may also appear in some parameters of the model.

Let us start from the isovector microscopic operator (5). Let us assume that the first four coordinates ${{\boldsymbol{r}}}_{j}$ correspond to the α particle and that the last two correspond to the deuteron. In vector notation, operator (5) reads

Equation (17)

The deuteron internal coordinate is

Equation (18)

and the α-deuteron relative coordinate is given by

Equation (19)

where ${{\boldsymbol{R}}}_{\mathrm{cm}}^{\alpha }=\tfrac{1}{4}{\sum }_{j=1}^{4}{{\boldsymbol{r}}}_{j}$ is the center-of-mass coordinate of the α particle.

Then, the E1 operator can be rewritten as

Equation (20)

where the first term

Equation (21)

is the E1 operator for the α particle. The second term is the E1 operator for the deuteron and the last term corresponds to the relative motion. The operators ${{\boldsymbol{T}}}_{\alpha }={\sum }_{j=1}^{4}{{\boldsymbol{t}}}_{j}$ and ${{\boldsymbol{T}}}_{d}={{\boldsymbol{t}}}_{5}+{{\boldsymbol{t}}}_{6}$ are the isospin operators of the α particle and deuteron, respectively. Hence, in multipolar form, one has

Equation (22)

with

Equation (23)

For more general clusters with mass numbers A1 and A2, the factor in front of $-e{{ \mathcal Y }}_{1\mu }({\boldsymbol{R}})$ in the last term becomes $({A}_{2}{T}_{{A}_{1}3}-{A}_{1}{T}_{{A}_{2}3})/A$. Its eigenvalue contains the factor ${Z}_{1}/{A}_{1}-{Z}_{2}/{A}_{2}$ mentioned in the introduction.

In a similar way, the first term of the isoscalar E1 operator (6) becomes

Equation (24)

where ${R}_{\alpha }^{2}=\tfrac{1}{4}{\sum }_{j=1}^{4}{({{\boldsymbol{r}}}_{j}-{{\boldsymbol{R}}}_{\mathrm{cm}}^{\alpha })}^{2}$, and the E2 operator reads

Equation (25)

In the simplest version of a three-body model, the α particle is in its ground state ${\phi }_{\alpha }^{00+}$. Effective multipole operators are obtained by taking the mean value of the above expressions,

Equation (26)

The eigenvalue of Tα3 is zero, as well as the mean value of ${{ \mathcal M }}_{\alpha ,\mu }^{E\lambda }$. The eigenvalue of Td3 vanishes for the neutron–proton system. Hence, for E1, one obtains from (22) and (24), with the neutron as particle 5 and the proton as particle 6,

Equation (27)

and

Equation (28)

where ${r}_{\alpha }^{2}=\langle {\phi }_{\alpha }^{00+}| {R}_{\alpha }^{2}| {\phi }_{\alpha }^{00+}\rangle $ is the mean square radius of the α particle. With (25), the E2 operator is given by

Equation (29)

This expression can also be deduced from equation (B2) of [29]. The first two terms are also derived in [22].

3.2. Transition matrix elements

In the present α + n + p three-body model, the initial scattering wave function is defined by coupling the ground state deuteron wave function with partial waves describing the relative motion. The polarizability of the deuteron and other distortion effects of the initial wave are thus neglected. The deuteron wave function is defined as a pure s state (except in section 5.2 below) by

Equation (30)

with l = 0 and S = j = 1. The spinor χS is the total spin state of the neutron and proton. The initial scattering functions for partial wave L read

Equation (31)

with π = (−1)L and

Equation (32)

since the α particle has spin 0 and positive parity.

The final 6Li(1+) ground state is described by a three-body wave function defined in the hyperspherical basis as

Equation (33)

where $\rho =\sqrt{\tfrac{1}{2}{r}^{2}+\tfrac{4}{3}{R}^{2}}$ is the hyperradius and Ω5 represents five angles, the orientations Ωr of ${\boldsymbol{r}}$ and ΩR of ${\boldsymbol{R}}$, and the hyperangle $\alpha =\arctan (\sqrt{8/3}\,R/r)$ (see [29, 30] for details). Number K is the hypermomentum. Notation γ represents the other quantum numbers of the problem, i.e., the orbital momentum l and spin S of the proton–neutron pair, and the orbital momentum L of the α − (n + p) relative motion. The functions ${{ \mathcal Y }}_{\gamma K}^{{JM}}({{\rm{\Omega }}}_{5})$ are hyperspherical harmonics and the functions χγK(ρ) are hyperradial functions. The positive parity requires l + L even.

Thanks to the antisymmetry of the deuteron wave function, it is possible to associate an isospin to the different parts of the three-body wave function,

Equation (34)

For the neutron–proton system in the isospin formalism, antisymmetry imposes that l + S + T must be odd. Hence it is possible to perform the separation (34) of the final wave function according to the deuteron isospin T. The component with l + S odd corresponds to Tf = 0 while the component with l + S even corresponds to Tf = 1. The wave function (33) can be interpreted as corresponding to the first two terms of equation (13). Indeed, while the α particle is frozen in its ground state, the deuteron can be fully distorted or excited and Tf = 1 admixtures can appear in the neutron–proton system.

Matrix element (14) becomes with (20)

Equation (35)

where J can be equal to 0–2. Matrix element (9) vanishes,

Equation (36)

Matrix element (10) reads

Equation (37)

When comparing with the microscopic expressions, one observes that important components are missing in the α + n + p model. The last term of equation (14) suggests that the transition matrix elements involving a virtual excitation of the α particle described by

Equation (38)

could play a significant role. Indeed, such a matrix element is related to the giant dipole resonance of the α particle. This effect occurs for an initial relative orbital momentum L = 1.

Simulating the effect of matrix element (38) is not possible in the present three-body model. Indeed, while the value of matrix element (38) might be estimated, the radial component ${g}_{{\alpha }^{* }{In}}^{1+}(R)$ of the relative wave function in equation (14) is unknown.

4. Numerical results

4.1. Conditions of the calculations

The determination of the final 6Li(1+) ground state wave function in a variational calculation is explained in [29]. The central Minnesota NN potential is employed as neutron–proton interaction [34]. For the α + N nuclear interaction, the potentials of Voronchev et al [35] and of Kanada et al [36] are employed. They are slightly renormalized by respective scaling factors 1.014 and 1.008 to reproduce the experimental binding energy 3.70 MeV of 6Li with respect to the α + n + p threshold. The Coulomb interaction between α and proton is taken as 2e2 erf(0.83 R)/R [37]. The coupled hyperradial equations are solved with the Lagrange-mesh method [29, 38]. The hypermomentum expansion includes terms up to Kmax = 24, which ensures a good convergence of the energy and of the T = 1 component of 6Li. The ground state is essentially S = 1 (96%). The matter r.m.s. radius of the ground state (with 1.4 fm as α radius) is found as $\sqrt{{r}^{2}}\approx 2.25$ fm with the potential of [35] or 2.24 fm with the potential of [36], i.e. values slightly lower than the experimental value 2.32 ± 0.03 fm [39]. The isotriplet component in the 6Li ground state has a squared norm 5.3 × 10−3 with the potential of [35] and 4.2 × 10−3 with the potential of [36].

For the initial scattering waves, the radial wave function u011(r) of the deuteron is the ground state solution of the Schrödinger equation with the Minnesota potential with ${{\hslash }}^{2}/2{m}_{N}=20.7343\,\mathrm{MeV}$ fm2. The Schrödinger equation is solved by using the Lagrange–Laguerre mesh method [38]. The converged deuteron energy is Ed = −2.202 MeV with 40 mesh points and a scaling parameter hd = 0.40. The scattering wave functions ${g}_{i}^{L\pi }(R)$ of the α + d relative motion are calculated with the deep potential of [19] adapted from the potential of [40].

4.2. Astrophysical S factors

The astrophysical S factor for multipolarity is defined in terms of the cross section σ(E) as [41]

Equation (39)

where η is the Sommerfeld parameter.

First, we evaluate the role of the two contributions to SE1 that are calculable in the present model, i.e. the isovector transition involving operator (27) from the L = 1 initial partial wave to the Tf = 1 component of the 6Li ground state and the isoscalar transition involving operator (28) to the Tf = 0 component. These two contributions add coherently. The transition operator given by the first term of equation (28) differs from the ones studied in several earlier works [6, 14, 15]. Indeed, it is argued in [31] that a neglected term in the matrix element may be rather large in these works. In the isoscalar operator (6) based on a Siegert transformation from which expression (28) is deduced, the second term should be negligible in the present case. The resulting difference is that the coefficient of the first term of equation (6) is smaller by a factor 4 than in the operators considered in [6, 14, 15].

In table 1, the resulting isovector and isoscalar ${S}_{E1}^{\mathrm{IV}+\mathrm{IS}}$ factor is compared at three energies with the purely isovector ${S}_{E1}^{\mathrm{IV}}$ factor. The isoscalar correction represents about 2%. It can be neglected as long as the isovector part is not better known. Notice that the isoscalar correction should be more important in the d(d, γ)4He capture reaction since the photon wavenumber kγ is much larger at low scattering energy.

Table 1.  E1 astrophysical S factor with the isovector (IV) and isovector + isoscalar (IV+IS) models. The α + N interaction of [35] is used.

E (MeV) ${S}_{E1}^{\mathrm{IV}}$ (MeV b) ${S}_{E1}^{\mathrm{IV}+\mathrm{IS}}$ (MeV b)
0.01 6.38 × 10−10 6.23 × 10−10
0.1 1.17 × 10−9 1.15 × 10−9
1 1.45 × 10−8 1.41 × 10−8

With the α + N potential of [35], the present IV+IS SE1 is represented in figure 1 as a dotted line. We have reanalyzed SE2 calculated with the E2 operator of equation (29) within the three-body model of [24], depicted as a dashed line in figure 1. At low energies, the cross section is very sensitive to the asymptotic behavior of the overlap integrals between the deuteron and the three-body wave functions for partial waves L = 0 and 2,

Equation (40)

up to large α − d distances R. In the model of [24], IL(R) follows over the interval [5, 10] fm the expected asymptotic behavior ${C}_{L}{W}_{-{\eta }_{b},L+1/2}(2{k}_{b}R)/R$, where ηb and kb are the Sommerfeld parameter and wavenumber calculated at the separation energy 1.474 MeV of the 6Li bound state into α and d. The L = 0 asymptotic normalization coefficient is C0 ≈ 2.05 fm−1/2 in reasonable agreement with the value C0 ≈ 2.30 fm−1/2 extracted in [42] from experimental data on α + d scattering. However, beyond about 10 fm, the absolute value of IL(R) decreases faster than the correct asymptotics. Hence, within that model, SE2 is underestimated at low collision energies. To solve this problem, beyond R0 = 7.5 fm, we replace IL(R) by the exact asymptotic expression with CL calculated at 7.5 fm. This corrected S factor is denoted as SE2c and is represented as a full line in figure 1. It is significantly larger than SE2 because the cross section is sensitive to R values up to about 50 fm at E = 0.1 MeV. From now on, we only use SE2c. Around the resonance, the S factor is dominated by E2 transitions. Dipole transitions should be dominant below about 0.1 MeV.

Figure 1.

Figure 1. Present E1 S factor, E2 S factor of [24] and corrected E2 S factor calculated with the α + N potential of [35] (Model A). The experimental data are from [8] (triangles), [10] (squares), [9] (open circles), and [12] (full circles).

Standard image High-resolution image

The total S factors SE1 + SE2c calculated with the potentials of [35] (Model A) and [36] (Model B) are presented in figure 2. They are compared with the direct data of [8] above the resonance (triangles), of [9] on resonance (open circles), and of [12] around 0.1 MeV (full circles). The indirect breakup data of [10] are indicated as squares. At low energies, the total S factor obtained in Model A (full line) nicely agrees with the LUNA data. The total S factor in Model B (dotted line) is lower by about 35% than in Model A (full line) but remains within the experimental error bars. This relative smallness is related with a smaller Tf = 1 component in Model B.

Figure 2.

Figure 2. Total E1 + E2c astrophysical S factor within the present three-body models A and B. The experimental data are from [8] (triangles), [10] (squares), [9] (open circles) and [12] (full circles).

Standard image High-resolution image

Despite that several possibly important T = 1 contributions are not included in the present discussion, i.e. mainly the whole Ti = 1 component in the initial wave and the Tf = 1 dipole excitation of the α core in the final wave function, one may nevertheless conjecture that a consistent treatment of all isovector E1 transitions can explain the low-energy experimental data. One can assume that the Ti = 1 component in the initial wave has the same order of magnitude as the one of the final bound state. Even a rough estimation of the role of the Tf = 1 dipole excitation of the α core is not possible because of a lack of knowledge of the functions ${g}_{{\alpha }^{* }{In}}^{1+}(R)$ appearing in equation (13). The present conjecture assumes, however, that the different contributions do not interfere destructively.

5. Discussion

5.1. Comparison to the exact-masses prescription

The developments of the previous sections now allow us to discuss the validity of the exact-masses prescription for reactions between N = Z nuclei. We have seen that one can conjecture that isovector E1 transitions are able to explain the low-energy S factor with a good accuracy. This is incompatible with the exact-masses prescription as we now show.

To simplify the discussion, let us consider E1 transitions in the two-cluster case. When the isospin formalism is not used, the E1 operator reads

Equation (41)

where ${{\boldsymbol{r}}}_{{pj}}^{{\prime} }$ is the coordinate of proton j with respect to the center-of-mass of the system. For a system of two clusters with masses A1mN and A2mN, where ${m}_{N}=\tfrac{1}{2}({m}_{n}+{m}_{p})$ is the nucleon mass, and charges Z1e and Z2e, this operator can be rewritten as [43]

Equation (42)

For the two structureless clusters of the potential model, the internal E1 operators ${{ \mathcal M }}_{1\mu }^{E1}$ and ${{ \mathcal M }}_{2\mu }^{E1}$ do not contribute and matrix elements are obtained from the final term. It vanishes if Z1/A1 = Z2/A2, and in particular in the case of two N = Z clusters.

If the masses of the clusters are M1 and M2, the center-of-mass coordinate ${{\boldsymbol{R}}}_{\mathrm{cm}}$ replacing (4) becomes

Equation (43)

as a function of the center-of-mass coordinates ${{\boldsymbol{R}}}_{\mathrm{cm}}^{1}$ and ${{\boldsymbol{R}}}_{\mathrm{cm}}^{2}$ of the clusters, with M = M1 + M2. When this expression is introduced in (41), one recovers (42) except for the dimensionless factor multiplying ${{ \mathcal Y }}_{1\mu }({\boldsymbol{R}})$ which is exactly replaced by

Equation (44)

The exact-masses prescription consists in using this factor in (42), where M1 and M2 are the experimental masses of the colliding nuclei. For N = Z nuclei, this factor does not vanish in reactions between different nuclei. While this prescription is exact in the context of the two-cluster model with masses M1 and M2 differing from A1mN and A2mN, it is in contradiction with the treatment of a collision between N = Z nuclei within the isospin formalism as we show below.

The factor (44) is sometimes also justified as a relativistic correction [26]. If one replaces the center-of-mass coordinates of the clusters by center-of-energy coordinates, the electric dipole moment becomes closer to expression (44). Though it is true that relativistic corrections could play a role, the argument is weakened by the fact that the original factor ${Z}_{1}/{A}_{1}-{Z}_{2}/{A}_{2}$ is based on a microscopic description in terms of nucleons while the center-of energy argument is based on a two-cluster structure. Consistent relativistic corrections should also be based on nucleons.

Let us first notice that the factor (44) still vanishes in collisions between identical nuclei. It predicts that E1 transitions are strictly forbidden in collisions between two identical N = Z nuclei, in contrast with microscopic and ab initio calculations [3, 4]. It would for example be ineffective to try to describe the forbidden E1 deuteron-deuteron capture.

Second, let us show that there is no relation of this factor with isospin mixing. The mass of a nucleus ${}_{Z}^{A}$XN can be written as

Equation (45)

where B(A, Z) is the binding energy. As the binding energy per nucleon is small with respect to the nucleon mass energy, factor (44) can be approximated for a capture involving nuclei with N = Z as

Equation (46)

This correction is small since the binding energy per nucleon does not vary much from one nucleus to another. In the α + d case, it is about 4 × 10−4. This factor is quite small and is fortuitously able to reproduce a plausible order of magnitude of forbidden E1 transitions. However, there is no physical relation between this correction and the dominant isovector transitions when the E1 transition is isospin-forbidden. Indeed, the binding energy per nucleon of an N = Z nucleus mainly depends on the dominant T = 0 component of its ground state. It is in no appreciable way sensitive to T = 1 admixtures as E1 matrix elements describing an isospin-forbidden capture should be.

Moreover, for the α + d capture, the prescription involves a matrix element which is sensitive to the long distance α + d tail of the 6Li ground state wave function. This is not the case in the isospin formalism where the main part of isospin mixing occurs when the two clusters overlap.

Can the exact-masses prescription give a realistic energy dependence of the S factor below the 711 keV resonance? Since the dominant initial orbital momentum is L = 1, the low-energy dependence of the initial relative scattering wave (equation (11)) is close to the dependence of the regular Coulomb function F1 (see equation (7) of [44]),

Equation (47)

In any model, the coefficients fi(R) are calculable functions of R. For Coulomb waves, they are given by equation (22) of [44]. The integral M(E) over R appearing in matrix element (35) and its various corrections can thus be written at very low energies as

Equation (48)

where coefficient Mi is an integral involving fi(R), the radial operator R, and the overlap integral IL(R) of the bound state wave function with the internal cluster wave functions (such as equation (40) in the three-body case). This last factor is quite different in the exact-masses prescription (where it is just given by the final bound state wave function with Tf = 0) and in isovector matrix elements (where it corresponds to a small Tf = 1 admixture of the final wave function). In particular, it is quite different at large distances since the Tf = 1 admixture does not have an α + d asymptotic behavior. Hence M0 and M1 may be quite different in both descriptions.

The low-energy behavior of the S factor is given by the expansion

Equation (49)

where the slope s1 depends on the ratio of M1 and M0 [44, 45]. At sufficiently low energies, this ratio computed with the exact-masses prescription is not related to the one in the isovector transition picture. We have checked numerically that the S factors of both descriptions have different low-energy dependences. The prescription does not reproduce the physical energy slope of SE1 near zero energy.

5.2. Role of S wave capture

The E1 S factor which is dominant below about 0.1 MeV decreases with decreasing energy since it is due to a transition from an initial P wave. As transitions from S waves have an almost flat energy dependence at very low energies, an energy (possibly very low) may exist where transitions from an initial S wave dominate.

The E2 capture cross section mainly corresponds to a transition between an initial D-wave and the 6Li ground state. In the present α + n + p model, an E2 capture from an initial S wave exists but is smaller than the other E2 contributions by several orders of magnitude in the energy range of figures 1 and 2 [24]. However, other transitions starting from the S wave are possible, which are not considered here. Since the 6Li, 4He, and 2H ground states contain a D-wave component due to the NN tensor force, several types of E2 transition from an initial S wave can contribute. As the energy dependence of transition matrix elements from an initial S wave is much weaker than for a D-wave, this contribution could become dominant below some low-energy. This mechanism is well illustrated by the d(d, γ)4He capture reaction [3, 4]. The main contribution to the capture at low energies is due to the small D-wave components of the α particle and of the deuterons. For 4He(d, γ)6Li, earlier works indicate that this component is small [13, 22] but they are restricted to energies above the 711 keV resonance. It is thus not possible for the moment to estimate the energy below which this mechanism would be important nor the order of magnitude of its contribution to the cross section at low energies.

We have performed a partial test within the α + n + p three-body model by including a D-wave component in the initial deuteron wave function. With the full deuteron wave function obtained with the soft-core potential of [46], the S wave contribution to SE2 is negligible above 10 keV. The resulting S wave capture remains very small in agreement with previous studies. Full confirmation requires a calculation taking simultaneous account of the 6Li, 4He, and 2H D components. Such a calculation requires extensions of the three-body model but is within the reach of present-day ab initio calculations.

The magnetic dipole capture is another case where capture from the S wave can occur. The microscopic M1 operator can be written as a sum of a term proportional to the total angular momentum and a residual spin term. The matrix elements of the first term must vanish in any model because of the orthogonality between the initial and final wave functions [26, 27, 47]. It is thus meaningless to evaluate M1 capture in models (like the present one) where the initial scattering partial waves and the final bound state wave function are not derived from the same Hamiltonian. When the matrix element of the residual spin term is small, M1 transitions are strongly hindered. The energy below which M1 transitions might dominate E1 transitions must be very small.

6. Conclusion

In this paper, we discuss the properties expected for a realistic treatment of the isospin-forbidden E1 component of a radiative-capture reaction between N = Z nuclei and we study the α(d, γ)6Li reaction. Since such a calculation is presently not available at the nucleon microscopic level, we evaluate some contributions that are accessible with a three-body model. The higher-order contribution from the isoscalar part of the operator is found to be small and could be neglected in future calculations of this reaction to a good approximation. The isotriplet component of the final 6Li(1+) ground state due to deuteron virtual excitations leads to a total E1 + E2 astrophysical S factor compatible with the experimental data at the low energies of [12]. Other E1 components of the S factor due to similar distortions of the initial scattering wave and to Tα = 1 virtual excitations of the α particle in the 6Li ground state are not accessible within the present model. We conjecture that, with these other contributions, isovector transitions are able to explain the data without adjustable parameter. We also emphasize the need for correct α + d asymptotics of the three-body wave function to correctly describe the E2 component of the astrophysical S factor.

We have questioned the exact-masses prescription of the potential model and shown that it is not founded at the microscopic level. It is incompatible with an explanation of the low-energy data in terms of isovector E1 transitions. Its order of magnitude and energy dependence may be accidentally correct but this prescription does not seem to have a physical meaning for capture reactions between N = Z nuclei. Its use should be avoided in capture reactions such as α(d, γ)6Li or 12C(α, γ)16O.

Radiative capture from the S wave could become dominant below some unknown low-energy. It is not completely established that this type of transition is too weak to contribute to the capture process at the lowest energies where experiments are available. This initial partial wave can play a role in M1 and E2 transitions. While M1 transitions are strongly hindered by the orthogonality between the initial and final states, it could be worth reexamining the E2 radiative capture at very low energies to evaluate the role of the various D-wave components in the initial and final clusters. Indeed such components in 2H, 4He, and 6Li render possible transitions from an initial S wave with a much weaker energy dependence at very low energies as obtained in the d(d, γ)4He reaction [3].

As long as ab initio calculations or advanced microscopic cluster calculations involving various forms of isospin mixing are not available, the importance of E1 transitions in the $\alpha +d\,{\to }^{6}$Li + γ reaction will remain poorly known. The three-body model is interesting as it offers simpler physical interpretations than more elaborate models. Some aspects of the present three-body study, however, limit its predictive power. Extensions are possible which should be considered in the future. The first one is to improve the α + d asymptotics of the final 6Li wave function. A second one is to replace the frozen-deuteron description in the initial wave by a flexible three-body description allowing distortions of the deuteron and, in particular, the appearance of isotriplet admixtures which will contribute to E1 capture in a consistent way with those of the final 6Li ground state. A third, more difficult, extension would involve core excitations, i.e., additional configurations for the α particle. We expect that a significant component of E1 capture could come from T = 1 virtual excitations of the α particle corresponding to its giant dipole resonance. Future three-body but also microscopic calculations of E1 α + d capture should usefully include this kind of configuration.

Acknowledgments

EMT thanks the Fonds de la Recherche Scientifique—FNRS (Belgium) for a grant. He is grateful to P Descouvemont for his kind invitation and welcome. He also acknowledges useful discussions with LD Blokhintsev and AS Kadyrov.

Please wait… references are loading.