Cluster effective field theory and nuclear reactions
Abstract
An effective field theory (EFT) for a nuclear reaction at low energies is studied. The astrophysical -factor of radiative capture on 12C at the Gamow-peak energy, MeV, is a fundamental quantity in nuclear-astrophysics, and we construct an EFT for the reaction. To fix parameters appearing in the effective Lagrangian, the EFT is applied to the study for three reactions: elastic -12C scattering at low energies, transition of radiative capture on 12C, and delayed emission from 16N. We report an estimate of the -factor of the reaction through the transition at by employing the EFT. We also discuss applications of EFTs to nuclear reactions at low energies.
pacs:
24.10.-iNuclear reaction models and methods and 24.40.LwRadiative capture and 24.55.CiElastic and inelastic scattering and 23.40.-s decay; double decay; electron and muon capture and 23.60.+e decay1 Introduction
In the late 1970s, a phenomenological Lagrangian method is suggested by Weinberg w-pa79 as an alternative of current algebra to calculate a hadronic matrix element at low energies. Skillful techniques are required for calculations of the current algebra to derive a hadronic reaction amplitude, e.g., the electro pion production on a nucleon, considering CVC and PCAC and satisfying the results from the low energy theorem and the relations among vertex functions required by Ward-Takahashi identities o-prc92 ; o-prc93 . Those results, including the radiative and non-radiative vector and axial-vector matrix elements of nucleon, nonetheless, are straightforwardly calculated from the chiral Lagrangian, which embodies the chiral symmetry and the explicit symmetry breaking patterns, in heavy-baryon chiral perturbation theory bklm-pr94 ; am-plb98 ; amk-prc01 . The idea of the method has been developed in various ways, and the framework of those methods are now known as effective field theories (EFTs). (For a review of EFTs, see, e.g., Refs. bv-arnps02 ; bh-pr06 ; m-15 ; hjp-17 ; hkvk-19 .) In the present work, we discuss a construction of an EFT for nuclear reactions at low energies based on our recent works a-epja16 ; a-prc18 ; a-jkps18 ; ya-19 ; a-18 .
When one studies a reaction by means of EFT, one may expect the characteristics for an EFT listed below: (1) It is a model independent approach. (2) One needs to introduce a momentum scale to separate relevant degrees of freedom at low energies from irrelevant degrees of freedom at high energies. (3) The theory provides us a perturbative expansion scheme around a specified theoretical limit; counting rules in powers of , where denotes a typical momentum scale of a reaction and does a high momentum scale, will be available. By using the counting rules, one can expand the amplitude order by order, and, up to a given order, one has a finite number of diagrams to calculate. (4) Coefficients appearing in an effective Lagrangian may not be constrained by its mother theory but can be fixed by using experimental data.
One might regard the construction of an EFT for a nuclear reaction as a formidable task because a nucleus is a many-body nucleon system and its structure is not still easily described from first principles. At low energies, e.g., a long wavelength limit for an external probe, however, an amplitude constructed from an effective Lagrangian, which is represented in terms of relevant low energy degrees of freedom and embodies symmetry requirements, exhibits a specific expression in the low energy limit and may describe well the reaction through the external probe. This might be a consequence of the low energy theorem em-cnpp95 for nuclear reactions. As discussed above, to construct an EFT for a nuclear reaction is possible; the important two conditions are (2) a clear separation scale and (4) an availability of experimental data for a reaction. For application in nuclear-astrophysics, nonetheless, one encounters an additional task. Because experimental data are available at relatively high energies due to the Coulomb barrier, it is necessary to work out an extrapolation of a reaction rate to a low energy.
In the following sections, we discuss and evaluate the factor of radiative capture on 12C at stellar energy through the transition by constructing an EFT a-18 . Our approach is rather new compared to popular methods for nuclear-astrophysics studies, such as -matrix or -matrix approach lt-rmp58 and potential models. Many studies of the transition of radiative capture on 12C have actually been carried out by using the conventional methods, in which a crucial observation has been made: The importance of indirect measurement of a reaction, delayed emission from 16N, was pointed out bb-npa06 ; detal-17 . An interference between a subthreshold state and a resonant state (also including so called background levels) for channel is sensitive to a secondary peak at lower energy side of the energy spectrum from -delayed emission from 16N. That is an essential input to deduce the factor when one employs the -matrix approach. Meanwhile, this feature appears quite differently in our approach: the subthreshold and resonant states for the channel are represented within a single dressed 16O propagator while the secondary peak of the energy spectrum from delayed emission from 16N is described by an interference between the amplitudes from a pole diagram and a non-pole one. Because one might argue about the applicability of the present approach to studies of nuclear reactions (e.g., because of no interference between the subthreshold and resonant states for the channel), the related reactions to estimate the -factor should be studied simultaneously using the same formalism. In addition, because the -factor has recently been estimated by using mainly a single method, the -matrix approach, the model dependence of the method is worth questioning. An examination employed by another method would be called for.
In the present work, we review a series of the calculations for the transition of radiative capture on 12C by constructing an EFT, in which three reactions, elastic -12C scattering a-epja16 ; a-prc18 ; a-jkps18 ; ya-19 , factor of the radiative capture on 12C a-18 , and delayed emission from 16N, are discussed. The construction of the reaction amplitudes for those three sectors, 1) nuclear elastic scattering, 2) an electromagnetic probe, and 3) an electroweak probe for nuclear reactions, is systematically carried out by introducing external vector (or minimally coupled photon) and axial-vector fields, which preserve symmetry requirement for constructing nuclear reaction amplitudes. In addition, a part of the reaction amplitudes, a dressed 16O propagator, is shared by those three reactions, which can easily be identified by drawing Feynman diagrams. Once parameters for the dressed 16O propagator are fitted to the elastic scattering data, it can be used for making nuclear reaction amplitudes for the radiative capture on 12C and the delayed emission from 16N. Additional coupling constants appearing in the reaction amplitudes are fixed by using experimental data for the corresponding reactions, and the factor is extrapolated to stellar energies. Reviewing those calculations, as an example of application of EFTs to nuclear reactions, one may see the reliability of the present approach.
This paper is organized as follows. In Sec. 2, a short overview of EFTs for the present work is presented. In Sec. 3, a construction of an EFT for the radiative capture on 12C at stellar energy is discussed; counting rules of the reaction are mentioned, and an effective Lagrangian is displayed. In Sec. 4, an application of the EFT to the study of the elastic -12C scattering at low energies is discussed, and in Sec. 5, the EFT is applied to the study of the transition of radiative capture on 12C. In Sec. 6, an application of the EFT to the delayed emission from 16N is explored. In Sec. 7, results and discussion of the present work are summarized. In Appendix, the propagators and the vertex functions for the elastic -12C scattering are displayed.
2 EFTs for nuclear reactions at low energies
EFTs are now a popular method for the studies of hadron and nuclear physics. In this section, we review the progress of EFTs as regards to their application to nuclear reactions at low energies.
The most representative example of an EFT is chiral perturbation theory (), a low energy effective field theory of QCD w-pa79 ; gl-ap84 ; gl-npb85 ; dgh-92 : Hadrons are composite particles consisting of quarks and gluons, which are described by color gauge theory. The quarks and gluons, however, are confined in the hadrons and never appear as free particles. Meanwhile, the QCD Lagrangian approximately has a global flavor symmetry because of the light quark masses. The symmetry is spontaneously broken down to involving eight massless Goldstone bosons in the ground state of QCD. Because of the nature of the Goldstone bosons, its interactions vanish in the zero momentum and zero light quark mass limits, and one can expand a reaction amplitude perturbatively in powers of the number of derivative and/or light meson mass factor around the vanishing interaction limit. One may notice that the picture of a hadron is completely altered between QCD and PT; a hadronic system, for example, a proton is described by QCD as a many body system consisting of strongly interacting quarks and gluons and by PT as a heavy core surrounded by a cloud of the massless Goldstone bosons in the chiral limit.
Weinberg suggested the first application of PT to nuclear physics w-plb90 ; w-npb91 : Because of an enhancement effect from two nucleon propagation, which can alter the counting rules, one may construct a nuclear potential according to the chiral order counting rules from two-nucleon irreducible diagrams in the time ordered perturbation theory. To obtain a reaction amplitude, one solves the Lippmann-Schwinger equation with the chiral potential. This approach matches to the traditional view of the nuclear potential; the long range part of the nuclear potential consists of the one-pion exchange, the intermediate range does of the two-pion exchange, and the short range part is parameterized by models and has a hard repulsion core. This picture can be reproduced order by order in the Weinberg’s counting scheme; at leading order (LO) the nuclear potential consists of the two nucleon contact interactions and the one-pion exchange contribution, at next-to leading order (NLO) one includes the two pion exchange contributions and contact interactions with two derivatives or two meson mass factors, and so on. However, because of a singular interaction due to the tensor force, modification of the Weinberg’s counting rules is necessary in order to make the scattering amplitudes cutoff-independent ntvk-prc05 .
Instead of a perturbative expansion of the nuclear potential, Kaplan, Savage and Wise(KSW) suggested to expand the two-nucleon scattering amplitudes ksw-plb98 ; ksw-npb98 . An elusive point to construct a perturbation theory for the two-nucleon systems is how to treat unnaturally small scales appearing in the -wave states, the scattering length for channel, fm, and the deuteron binding energy for channel, MeV, compared to the typical scale of PT, the pion mass, MeV. To take account of those small scales in the -wave two-nucleon scattering amplitudes, the LO two-nucleon contact interactions are resummed up to infinite order while the one-pion exchange is perturbatively included. 111 There are other approaches, e.g., which treat the one-pion-exchange piece non-perturbatively while the two-pion-exchange is treated perturbatively. See, e.g., Ref. ly-prc12 . This scheme is known as the KSW counting scheme. For studies of low energy reactions, the theory is easily simplified as a pionless EFT in which the pions are regarded as irrelevant degrees of freedom at high energy and integrated out of the theory crs-npa99 .
The pionless EFT is subsequently applied to the study of three nucleon systems. Bedaque, Hammer and van Kolck studied the triton system by using the pionless theory bhvk-npa00 222 3He system in the pionless EFT, including the Coulomb interaction, is studied in Ref. ab-jpg10 .. They found a cyclic singularity, the so-called limit cycle, in the triton channel when scaling a momentum cutoff introduced in the coupled integral equations. Along with the limit cycle, a universal feature emerges in the three-body system, known as the Efimov effect e-sjnp71 ; e-sjnp79 : an infinite number of three-body bound states, whose binding energies appear as a geometrical series, are accumulated at the threshold. The Efimov sates in the triton system appear in a theoretical limit, the so-called unitary limit, where the -wave scattering lengths become infinite and the two-body binding energy vanish in the two-body propagators. Thus, one can make a perturbative expansion around the unitary limit, which appears in the inverse of the two-body propagators. It actually coincides with the well-known effective range expansion b-pr49 .333 A dibaryon field, which has baryon number 2, is introduced by Kaplan k-npb97 , and it is useful to implement the effective range expansion in a theory without pions bs-npa01 ; ah-prc05 and with perturbative pions st-prc10 ; ah-prc12 . 444 Recently, a study for nuclear few-body systems around the unitary limit was reported by König et al. ketal-prl17 . It is conjectured that this vanishing point appears in QCD at a slightly large unphysical pion mass, 200 MeV, (see, e.g., Figs. 11 and 12 in Ref. emg-npa03 ). Such an infrared point as a function of the pion mass for the triton system is studied by Braaten and Hammer bh-prl03 . To renormalize the cyclic singularity, one needs to promote the three-body contact interaction at LO. For the study of the triton system, one can fix the coupling constant of the three-body contact interaction by using the triton binding energy. In addition, one can choose a triton wavefunction as a LO constituent of an amplitude and make a perturbative expansion around it j-12 ; jp-fbs13 ; v-prc13 ; v-prc17 .
Along with various theoretical developments of EFTs (a part of which we have briefly discussed above), EFTs are also employed for phenomenological studies at low energies, for which error estimates are important. An error estimate for a reaction can be controlled by using the perturbative expansion scheme of EFTs. EFTs have been employed for the phenomenological studies of, for example, neutron -decay aetal-plb04 ; amt-plb09 , solar neutrino reactions on the deuteron bck-prc01 ; aetal-plb03 ; ash-20 , radiative neutron capture on a proton at BBN energies r-npa00 ; achh-prc06 , proton-proton fusion kr-npa99 ; bc-plb01 ; ashhk-plb08 ; cly-plb13 , 3He(,)7Be hrv-epja18 ; znp-18 , and 7Be(,)7B znp-prc14 ; rfhp-epja14 in the Sun. In the following sections, we discuss an application of an EFT to nuclear reactions for heavier nuclei at low energies.
3 EFT for radiative capture on 12C at
In this section, we discuss a construction of an EFT for the study of radiative capture on 12C at MeV. We first briefly review previous studies and then construct an EFT for this reaction. In the following subsections, we analyze the typical and high energy-momentum scales for the radiative capture reaction at , we also discuss which are the relevant and irrelevant physical degrees of freedom, and from those we derive the power counting rules for our EFT description. We also write down the effective Lagrangian for three reactions: elastic -12C scattering, transition of the radiative capture on 12C, and delayed emission from 16N.
3.1 Introduction to the radiative capture on 12C
The radiative capture on 12C, 12C(, )16O, is one of the fundamental reactions in nuclear astrophysics, which determines the ratio 12C/16O produced in helium burning f-rmp84 . The reaction rate, equivalently the astrophysical -factor, of the process at the Gamow peak energy, MeV 555 See the footnote 6. , however, cannot experimentally be determined due to the Coulomb barrier. A theoretical model is necessary to be employed in order to extrapolate the cross section down to by fitting model parameters to available experimental data measured at a few MeV or larger. During the last half century, a lot of experimental and theoretical studies for the process have been carried out. For reviews, see, e.g., Refs. bb-npa06 ; detal-17 ; chk-epja15 and references therein.
In constructing a model for the radiative capture process, one needs to take into account the excited states of 16O bb-npa06 , particularly, two excited bound states for and just below the -12C breakup threshold at and MeV 666 The kinetic energy denotes that of the -12C system in the center of mass frame. , respectively, as well as and resonant (second excited) states at and MeV, respectively. Thus the capture reaction to the ground state of 16O at is expected to be and transitions dominant due to the subthreshold and states. While the resonant and states play a dominant role in the available experimental data at low energies, typically MeV. A minor contribution to the -factor appears from so-called cascade transitions in which the initial and 12C form an excited state of 16O emitting a photon, and it subsequently decays to the ground state of 16O. Experimental data pertaining to processes for nuclear astrophysics are compiled, known as NACRE-II compilation xetal-npa13 , in which the -factor of the 12C(,)16O reaction is estimated employing a potential model, and reported uncertainty of the process is less than 20 %.
Theoretical frameworks employed for the study are categorized mainly into two detal-17 : the cluster models using generalized coordinate method dbh-npa84 or potential model lk-npa85 and the phenomenological models using the parameterization of Breit-Wigner, -matrix lt-rmp58 , or -matrix hdz-npa76 . A recent trend of the study is to rely on intensive numerical analysis, in which a larger amount of the experimental data relevant to the study are accumulated, such as those from 12C(,)16O, radiative proton capture on 15N to the ground state of 16O, 15N(,)16O, delayed emission from 16N, 16N()12C, elastic -12C scattering, 12C(,)12C, and transfer reactions, 12C(6Li,)16O, 12C(7Li,)16O, 6Li(12C,)16O, and 7Li(12C,)16O, up to relatively high energies, MeV, and a significant number of parameters of the models are fitted to the data by using computing power detal-17 ; xetal-npa13 ; aetal-prc15 . In the following, we discuss an alternative approach to estimate the -factor at ; we discuss counting rules of the EFT at and display an effective Lagrangian for the reactions.
3.2 An EFT for the radiative capture on 12C
In the study of the radiative capture on 12C at MeV employing an EFT, at such a low energy, one may regard the ground states of and 12C as point-like particles whereas the first excited states of and 12C are chosen as irrelevant degrees of freedom, by which a large momentum scale of the theory is determined. An effective Lagrangian for the process is constructed by using two spinless scalar fields for and 12C, and terms of the Lagrangian are expanded in terms of the number of derivatives. The expansion parameter of the theory is where denotes a typical momentum scale : is the Gamow peak momentum, MeV, where is the reduced mass of and 12C.777 A typical length scale of the -12C system is, thus, fm. This is comparable to the sum of the radii of and 12C obtained from a nuclear radius formula, fm; fm. Meanwhile, the emitted photon carries away almost all of the released energy, MeV, and the length scale of the photon is larger than the other length scales, fm. Thus, the photon may recognize the nuclear system as to be point-like. denotes a large momentum scale or MeV where is the reduced mass of one and three-nucleon system and is that of four and eight-nucleon system. and are the first excited energies of and 12C, respectively: MeV for state of 4He and MeV for state of 12C. By including the terms up to next-to-next-to-leading order (NNLO), for example, one may obtain about 10% theoretical uncertainty for the process.
Coefficients appearing in an effective Lagrangian are fixed by using the experimental data which are measured at significantly higher energies than . In the following sections, we employ the data from three processes, elastic -12C scattering measured at MeV, factor for 12C(,)16O at MeV, and delayed emission from 16N at MeV. Thus, the perturbative scheme of the theory may not be reliable at the energies where those experimental data are measured. We first fix some parameters of the dressed 16O propagators, parameterized in terms of effective range expansion, by using the binding energies of the excited 16O states and fit the other parameters of the propagators to the phase shift data of the elastic scattering measured at MeV. Because of non-perturbative nature of a propagator we treat it as a non-perturbative quantity; the dressed 16O propagator commonly appears in the three processes, and we keep it as a non-perturbative one and expand reaction amplitudes perturbatively around it. For the radiative capture process, two additional parameters for the transition amplitude of the radiative capture process are fitted to the factor data measured at MeV, and a value of the factor is estimated at MeV.
For the delayed emission from 16N, parameters of the decay amplitudes are fitted to the experimental data measured at MeV. This study would explore a validity of the present approach. One reason is that the delayed emission data are covered with the small energy region compared to that for the elastic scattering data. The other reason is, as mentioned before, to test a different parameterization for the dressed 16O propagator compared to that in the -matrix or -matrix analysis. In the conventional -matrix analysis, the subthreshold state and the broad resonant state of 16O are represented by the Breit-Wigner formula and are linearly combined in the reaction matrix along with a background contribution; a secondary maximum of the delayed emission data is known to be important to constrain an interference pattern among those levels bd-npa88 ; jfhk-prc90 ; hfk-prc91 . In the present approach, the interference between the subthreshold and resonant states doesn’t exist because the subthreshold state and the broad resonant state of 16O is represented by a single dressed 16O propagator (in terms of effective range expansion); instead, we will see that the secondary peak can be reproduced by an interference between the amplitudes from a non-pole diagram and a pole diagram.
An effective Lagrangian for the present study is written as a-epja16 ; a-18
(1) |
where is the Lagrangian for the elastic scattering process a-epja16 , is for additional terms for the radiative capture process a-18 , and is for additional terms for the -delayed emission from 16N.
The Lagrangian may be written by using the composite 16O fields consisting of and 12C as a-epja16 ; a-jkps18 ; bs-npa01 ; ah-prc05
(2) | |||||
where () and () are fields (masses) of and 12C, respectively. is a covariant derivative, where is a charge operator and is the photon field. The dots denote higher order terms. represent and 12C composite fields of angular momentum : for , for , for , and for where those spin states are represented by the subscripts , , and as Cartesian tensors j-prb70 . (We have suppressed those indices for the Cartesian tensors in the Lagrangian in Eq. (2).) Those composite fields are introduced to make an expansion around the unitary limit. are coupling constants for the propagation of the -12C composite fields for the channels, and are related to effective range parameters along with common multiplicative factors . As we will discuss later, because of a modification of the counting rules for the elastic -12C scattering we include the terms up to for and those up to for . are projection operators by which the -12C system is projected to the -th partial wave states. Thus one has
(3) |
The Lagrangian for additional terms to the study for the transition of the radiative capture reaction may be written down as a-18
(4) | |||||
with
(5) |
where is the mass of 16O in the ground state. We note that because the 16O ground state appears only in the final state, () is introduced as a source field for the 16O ground state in the final (initial) state. In the first term in Eq. (4), for example, () destroys (creates) the 16O ground state, and [()] fields create (destroy) a -wave -12C state. The transition rate between the 16O ground state and the -wave -12C state is parameterized by the coupling constant , which is fixed by using experimental data. A contact interaction, the term, is introduced to renormalize divergence from loop diagrams for the radiative capture reaction.
A weak decay amplitude is described in terms of the type current-current interaction for semi-leptonic decay process, and its interaction Hamiltonian density is given as
(6) |
where is the Fermi constant, is a nuclear current to be constructed by using the external vector and axial-vector fields in the effective Lagrangian while is a lepton current. In the energy-momentum space it is given as
(7) |
where is the momentum transfer, .
The Lagrangian is for additional terms for the study of the delayed emission from 16N. We construct for interactions where the initial ground state of 16N decaying to the and states of 16O and the -12C breakup state for and 3 channels through the Gamow-Teller transition, which may be written as
(8) | |||||
where is the external axial-vector field, which generates an axial nuclear current for the Gamow-Teller transition, and is the source field for the initial ground state of 16N. , , , and are coefficients for and at LO, and and are those for at NNLO; the indices of the squared operators, and , for the terms at NNLO are suppressed in the above equation.
4 Elastic -12C scattering in the cluster EFT
In this section, we construct dressed composite 16O propagators. We first review the formalism of the effective range expansion for elastic -12C scattering. We then discuss a modification of the counting rules for a low momentum expansion around the unitary limit, based on the observation in a comparison between experimental data and terms generated from the Coulomb self-energy. After including broad resonant states of 16O, we fit the effective range parameters to phase shift data. We then calculate asymptotic normalization coefficients (ANCs) for the and states of 16O and compare them to the previous results.
4.1 Differential cross section of the elastic scattering
The differential cross section of the elastic -12C scattering (for two spin-0 charged particles) in terms of the phase shifts is given by (see, e.g., Ref. lt-rmp58 )
(9) | |||||
where is the scattering amplitude including both pure Coulomb part and Coulomb modified strong interaction part, is the scattering angle, is the absolute relative momentum, and . is the inverse of the Bohr radius, where and are the number of protons in and 12C, respectively, and is the fine structure constant. In addition, is the Coulomb scattering phase, for to ; are the Coulomb phase shifts, with , and
(10) |
are real scattering phase shifts. are the Legendre polynomials. The scattering amplitudes are represented in terms of as lt-rmp58 888 One can see the relation between and through the relation
(11) |


In the EFT, the amplitudes of the elastic scattering are calculated from diagrams depicted in Figs. 1 and 2. We obtain the scattering amplitudes for -th partial wave states as a-epja16 ; aetal-prc07 ; a-epja07
(12) |
with
(13) |
and
(14) |
where is the digamma function. Note that the function, , the Coulomb self-energy term, in the denominator of the amplitude is obtained from the Coulomb bubble diagram for the dressed propagator of 16O in Fig. 1, and the factor, , in the numerator is from the initial and final state Coulomb interactions between and 12C in Fig. 2. In Appendix A, we display the renormalized dressed three-point vertices for the initial and final Coulomb interactions and the renormalized dressed composite 16O propagators for .
The function represents the interaction due to the short range nuclear force (compared with the long range Coulomb force), which is obtained in terms of the effective range parameters as
(15) |
One can find that the expression obtained in Eq. (12) reproduces well the previous results reported in Refs. klh-jpg13 ; hot-npb73 ; h-jmp77 ; scb-prc10 .
By comparing the two expressions of the amplitudes in Eqs. (11) and (12), one has a relation between the phase shift and the effective range parameters in the denominator of the scattering amplitudes, , as
(16) |
where
(17) |
To estimate the ANC, , for the , , , states of 16O, we employ the definition of ir-prc84 ,
(18) |
where ; , and are binding momenta which will be given in the next subsection.
4.2 Effective range expansion and modification of the counting rules
We now discuss a modification of the counting rules a-prc18 , based on an observation in Eq. (16). The term on the left-hand-side of Eq. (16) is significantly suppressed by the factor at low energies. Meanwhile, the Coulomb self-energy term, , on the right-hand-side of Eq. (16) turns out to be two orders of magnitude larger than the term on the left-hand-side of the equation. Thus, we introduce terms in as counter terms at LO, by which those unnaturally large terms from the Coulomb self-energy term are subtracted. We will discuss it in detail below.
The effective range parameters in are expanded in powers of whereas the real part of the function can be expanded in powers of as well. For the function in , one has
(19) | |||||
where is the inverse of the Bohr radius, MeV, and regarded as another large scale of the theory. Thus, the right-hand-side of equation, , in Eq. (16) can be expanded as a power series of for both and . Meanwhile, the left-hand-side of Eq. (16) is suppressed by the factor , due to the Gamow factor .
In the case of the -wave, for example, the reported phase shift at the smallest energy, MeV, where is the kinetic energy of in the laboratory frame 999 The kinetic energies, in the center of mass frame and in the laboratory frame, are related by a relation . , is tetal-prc09 . The factor becomes at MeV which corresponds to MeV, and the left-hand-side of Eq. (16) numerically becomes MeV. The function is expanded as
(20) | |||||
at MeV. The numerical values in the third line of Eq. (20) correspond to the terms appearing in the second line of the equation in order. One can see that the power series converges well, but the first and second terms are two and one order of magnitude larger than a value estimated by using the experimental data in the left-hand-side of Eq. (16), 0.019 MeV. Those terms are unnaturally large, and thus it is necessary to introduce a new renormalization method, in which the counter terms remove the unnaturally large terms and make the terms in a natural size. In other words, we assume that fitting polynomial functions are represented as a natural power series at the low energy region, and to maintain such polynomial functions, large cancellations for the first and second terms with the and effective terms, respectively, are expected. So we include the three effective range parameters, , , and , for the channel as the counter terms. The same tendency can be seen in the channels whereas one needs to include four effective range parameters for the channel. Thus, we employ the three effective range parameters, , , for the channels and the four effective range parameters, , , , for the channel when fitting the parameters to the phase shift data.
At the binding energies of the excited , , , states of 16O, the amplitudes should have a pole at where are the binding momenta, ; denote the binding energies of excited states of 16O. Thus the denominator of the scattering amplitude, , should vanish at . Using this condition, a first effective range parameter, , is related to other effective range parameters as
(21) | |||||
and the remaining effective range parameters are fixed by using the phase shift data of the elastic scattering.101010 In Ref. a-jkps18 , we have also studied an inclusion of the ground state of 16O in the parameter fitting for the elastic -12C scattering at low energies.
4.3 Numerical results for the elastic scattering
The effective range parameters, , , for and , , , for are fitted, employing the standard fit 111111 We employ a python package, emceefmetal-12 , for the fitting. , to the experimental phase shift data of the elastic -12C scattering reported by Tischhauser et al. tetal-prc09 ; epaps . 121212 Input data for the parameter fitting are the phase shifts of the elastic -12C scattering for tetal-prc09 ; epaps , which have been generated from the -matrix analysis of the elastic scattering data tetal-prl02 . In the input data files epaps , there are four column data: the first column is the alpha energy, the second one the phase shift as derived from the globalized Monte Carlo simulations, the third one is the same phase shift randomized by the error from the Monte Carlo simulations, and the fourth one is the error of the phase shifts from the Monte Carlo simulations. We have used the second column of the phase shift data in our previous works a-epja16 ; a-prc18 and the third column of the phase shift data in our other works a-jkps18 ; ya-19 ; a-18 for fitting. The reported energy range of the data is MeV ( MeV.) tetal-prc09 . Though we have discussed that a large energy scale of the theory is that for the first excited state of 12C, MeV ( MeV) in the previous section, other large scales can emerge from resonate energies of 16O. As suggested by Teichmann t-pr51 , below the resonance energies, the Breit-Wigner-type parameterization for the resonances can be expanded in powers of the energy, and one can obtain an expression of the amplitude in terms of the effective range expansion. Therefore, in the previous works a-epja16 ; a-prc18 , we choose the energies of the resonant states, MeV, MeV, MeV, MeV, as a large scale of the theory for each partial wave state. After fitting the parameters to the data below the resonant energies, we confirm large cancellations between the fitted parameters and those generated from the Coulomb self-energy terms, e.g., in Eq. (20) while the perturbative series of the effective range expansions at converges well as expected by the theory. See Table V in Ref. a-prc18 and Table 2 in Ref. a-jkps18 . The perturbative series, however, does not converge at the large, experimental energies. We will come back to the issue in the next section.
In our recent works ya-19 ; a-18 , we include broad resonant states in the fitting because one may expect that because a relatively deep pole for a broad resonance state in the complex energy plane is located at a distance from the real axis, it can possibly be represented by a polynomial function in terms of the effective range expansion. We find that the broad resonant and states with the decay widths keV and keV can be well incorporated into the fitting of the effective range parameters whereas the narrow resonant and states with the decay widths keV and keV cannot. In addition, a tail-like structure from a higher resonant state at MeV with the width keV cannot be incorporated into the fitting for the high energy region of the data at MeV for either. Thus we use the maximum energies, , 6.0, 3.2, and 6.62 MeV for , and 3, respectively, of the data sets for the fitting where we assume that the resonant state of 12C, which could appear at MeV, would be negligible for and .
4.3.1 channel
As discussed above, we use the phase shift data for up to the maximum energy MeV due to the narrow resonant state at MeV with keV. To study the dependence on the choice of the data sets, we employ four data sets up to different maximum energies, , 4.5, 5.0, and 5.5 MeV.
(MeV) | (fm) | (fm3) | (fm5) | (fm) | (fm-1/2) |
---|---|---|---|---|---|
4.0 | |||||
4.5 | |||||
5.0 | |||||
5.5 |
In Table 1, fitted values and errors of the effective range parameters, , , and , are displayed, and calculated values of by using Eq. (21) and those of the ANC, , for the state of 16O by using Eq. (18) are also displayed. and numbers of the data () for the fitting are , 0.393(149), 0.450(167), 0.552(230) for , 4.5, 5.0, 5.5 MeV, respectively. As seen in the table, the errors of those quantities become smaller as the value of increases (mainly because the number of the data increases) while the center values of them are changing. The change of the center values of the coefficients may indicate an effect of truncation error in the fitting and/or that of the narrow resonance. If we include the data at high energies larger than MeV into the fitting, we have large values: with those up to MeV, and with those up to MeV. 131313 In Ref. a-20 , we recently studied an inclusion of the sharp resonant state of 16O and the first excited state of 12C in the study of elastic -12C scattering for up to MeV. We found that the state of 12C is redundant for fitting the phase shift data. To investigate its precise role, the inelastic open channel, C, would be necessary to be included in the study of the elastic scattering above the excited energy of 12C.
4.3.2 channel
For , as discussed above, we cannot incorporate the narrow resonant state at MeV with keV in the fitting. Thus we choose the maximum energy of the data set for the fitting as MeV. Using the data of the data set we fit effective range parameters as
(22) |
We calculate a value of and that of the ANC, , for the subthreshold state of 16O as fm5 and fm-1/2, respectively, where the number of the data is and . One can see above that because the value of is quite small while the error bars of those fitted quantities are large, we cannot deduce a meaningful result for from the fitting.
4.3.3 and channels
For , as discussed above, we can incorporate the broad resonant state at MeV with keV in the fitting for the effective range parameters but cannot do a tail from the next broad resonant state at MeV with keV. Thus the highest energy of the data for the fitting is MeV for . To study the dependence on the choice of the data sets, we employ four data sets up to the different maximum energies, , 4.0, 5.0, and 6.0 MeV.
(MeV) | (fm-1) | (fm) | (fm3) | (fm3) | (fm-1/2) |
---|---|---|---|---|---|
3.0 | |||||
4.0 | |||||
5.0 | |||||
6.0 |
In Table 2, fitted values and errors of the effective range parameters, , , and are displayed, and calculated values of and those of the ANC, , for the state of 16O are displayed as well. and numbers of the data () for the fitting are , 0.450(80), 0.509(167), 0.738(273) for , 4.0, 5.0, 6.0, respectively. As seen in the table, the errors of those quantities decrease as increases. It is worth pointing out that the center values of them are almost not altered even though is changed. Because the value for the data up to MeV is still smaller than 1, an effect from the resonant state of 12C is not significant for this channel. 141414 Though the maximum energy of the data for fit is larger than the energy of the first excited state of 12C, MeV, there is no indication of a need to include the state of 12C. See the footnote 13 as well. If we include the data at high energies larger than MeV into the fitting, we have a large value, with the data up to MeV.
For , as discussed above, we can incorporate the broad resonant state at MeV with keV in the fitting of the effective range parameters. Thus we take available all energy range of the experimental data for the parameter fitting; the highest energy of the data is MeV for . To study the choice of the data sets, we employ four data sets up to different maximum energies, , 5.0, 6.0, and 6.62 MeV.
(MeV) | (fm-5) | (fm-3) | (fm-1) | (fm) | (fm7) | (fm-1/2) |
---|---|---|---|---|---|---|
4.6 | ||||||
5.0 | ||||||
6.0 | ||||||
6.62 |
In Table 3, fitted values and errors of the effective range parameters, , , , and are displayed, and calculated values of and those of the ANC, , for the state of 16O are also displayed. and numbers of the data () for the fitting are , 0.45(167), 0.49(273), 0.50(354) for , 5.0, 6.0, 6.62 MeV, respectively. As seen in the table, the errors of those quantities decrease as increases. It is interesting to point out that, as the same as that we have seen for , the center values of them are almost not altered even though is changed except for the deviations of and and a large error of for MeV compared to the corresponding values for the other . That may stem from negative correlations between the parameters for MeV and almost no correlations between those for MeV. (One can find the correlations between the parameters in Table VI in Ref. ya-19 .) The effect from the resonant state of 12C is not significant either because of the small value for the data up to MeV. 151515 See the footnote 14.
4.3.4 Comparison of the ANCs
(fm-1/2) |
---|
In Table 4, we summarize our result of the ANCs for the and states of 16O where we show the results obtained from the largest data sets for the fitting. Because of the change of the center values and the large uncertainty for the ANCs for the and states of 16O, respectively, we do not include those results. We suppress the comparison about the ANCs for the and states below.
The ANCs for the and states of 16O have intensively been studied because the major contribution of the -factor of the radiative capture on 12C at come from the and transitions due to the subthreshold and states of 16O. In our previous work a-prc18 , we obtained fm-1/2 for the state, which agrees well with the result presented above. Our result underestimates, by about 10%, compared to the other theoretical estimates: fm-1/2 obtained from a potential model calculation by Katsuma k-prc08 , and fm-1/2 and fm-1/2 from a new parameterization method by Ramirez Suarez and Sparenberg rss-prc17 and by Orlov et al. oetal-prc17 , respectively. Our result, on the other hand, also underestimates or agrees well with the experimental results within the reported errors: fm-1/2 obtained from the 6Li(12C,)16O reaction by Avila et al. aetal-prl15 , fm-1/2 from the 12C(7Li,)16O reaction by Oulebsir et al. oetal-prc12 , and fm-1/2 from the 12C(6Li,)16O and 12C(7Li,)16O reactions by Brune et al. betal-prl99 .
Only some studies for the ANCs for the and states of 16O have been reported so far, though those ANCs are important to fix the -factor of the radiative capture reaction at through the cascade transitions for the -matrix analysis. Our result of the ANC of the state is large compared to that in our previous result, fm-1/2 a-prc18 . This is because of an error of the numerical code in the previous study and the use of the different data set of the phase shift for fitting. Our result of the ANC for the state overestimates an experimental result, fm-1/2 obtained by Avila et al. aetal-prl15 .
5 The -factor of radiative capture on 12C
In this section, we discuss how to estimate the -factor of radiative capture on 12C by using an EFT. First, we explain the formalism to calculate the and then we derive the transition amplitudes up to NLO from the effective Lagrangian. We then discuss a suppression of the transition amplitudes between nuclear states, and, after fitting two parameters to the data, we estimate the -factor at .
5.1 factor and radiative capture amplitude
The -factor is represented by using the total cross section from the transition as
(23) |
with
(24) |
where is the kinetic energy of the initial -12C state in the center of mass frame, ; is the magnitude of relative momentum between and 12C. is the photon energy,
(25) |
where is the -12C breakup energy from the ground state of 16O; MeV. One may notice that is larger than the large energy scale of the theory, the first excited energy of 12C, MeV. Because the released large energy is carried away by the outgoing photon, the final nuclear state remains in a state with a typical energy scale. We will discuss how one can avoid invoking a resonant state originated from the first excited state of 12C below. is a transition amplitude which will also be shown in the following.

In Fig. 3, diagrams of the radiative capture process from the initial -12C state for to the 16O ground () state up to NLO are depicted, in which the Coulomb interaction between and 12C is taken into account, and in Fig. 1, those for dressed composite propagators of 16O consisting of and 12C for are depicted. The propagator is obtained in the previous section. (See Appendix as well.)
The radiative capture amplitude for the initial state is presented as
(26) |
where is the polarization vector of outgoing photon and ; is the relative momentum of the initial and 12C. The amplitude is decomposed as
(27) |
where those amplitudes correspond to the diagrams depicted in Fig. 3.
We follow the calculation method suggested by Ryberg et al. rfhp-prc14 , in which Coulomb Green’s functions are represented in the coordinate space satisfying appropriate boundary conditions. Thus we obtain the expression of those amplitudes in the center of mass frame as
(28) | |||||
(29) | |||||
(31) | |||||
where is the magnitude of outgoing photon momentum, is the number of protons in 16O, and is the binding momentum of the ground state of 16O, . and are gamma function and spherical Bessel function, respectively, while and are regular Coulomb function and Whittaker function, respectively. The functions, , (where ), and , have been introduced in Eqs. (13), (14), and (15), respectively.
Regarding the divergence from the loop integrals, the loops of the diagrams (a) and (b) in Fig. 3 are finite while those of the diagrams (d) and (e) lead to a log divergence in in the limit, . We introduce a short range cutoff in the integral in Eq. (LABEL:eq;Xde), and the divergence is renormalized by the counter term, . The loop of the diagram (f) diverges and is renormalized by the term as well. Thus we have
(32) |
where is the divergence term from the diagrams (d) and (e) and is that from the diagram (f); we have
(33) |
where, when we derive the expression of , the dimensional regularization in space-time dimensions is used; and is a scale factor from the dimensional regularization. is a renormalized coupling constant which is fixed by experiment.
We have used the two regularization methods when calculating the loop diagrams (d) and (e) and the loop diagram (f). Some different regularization methods result in different expressions for divergent terms and constant terms but the same expression for functional terms (such as mass or momentum dependence terms) r-npb96 . Thus the different regularization methods may be adapted by adjusting a value of the coefficient when we send a value of sufficiently small. Recently, Higa, Rupak, and Vaghani reported that the divergent terms are exactly canceled with each other among those diagrams when they calculate one-photon-exchange diagrams for the -order terms using the dimensional regularization hrv-epja18 .
We now discuss how one can avoid invoking a resonant state due to the released energy from the radiative capture reaction. As mentioned above, the both initial and final nuclear states remain in the states at the typical energies because the photon carries away almost all of the released energy. A large energy gap, then, appears in the intermediate state, i.e., in the loop diagrams; the initial -wave -12C state is in a typical energy state while, after the photon is emitted, the energy gap occurs as a deep binging energy for the ground state of 16O in the -wave -12C propagation. Because the binding energy for the ground state of 16O is far below the typical energies for the -12C propagation, its physical effect will not be significant. Moreover, the excited energy for the resonant state of 12C is located at farther above than that for the -12C breakup threshold, thus an effect from the resonant state of 12C will hardly be seen. Meanwhile the loop integrals may pick up the deep momentum scale. From the loop diagram (f), for example, when the Coulomb interaction is ignored, the large momentum scale MeV is picked up in the numerator of the amplitude. It causes the emergence of a term which does not obey the counting rules. 161616 A method to renormalize a term which does not obey counting rules in manifestly Lorentz invariant baryon chiral perturbation theory, is known as the extended on mass shell (EOMS) scheme fgjs-prd03 ; af-prd07 . One can similarly renormalize the term proportional to in the counter term, , even when the Coulomb interaction does not exist. In the present case, the large momentum scale from the ground state energy of 16O appears as a ratio , due to the non-perturbative Coulomb interaction, where is another large momentum scale, MeV. The finite term in from the loop of the diagram (f) with is reduced to a typical momentum scale, MeV.
5.2 Suppression of the transition and mixture of isospin state
Before fitting the parameters to available experimental data, we discuss three issues: non-perturbative treatment of the dressed 16O propagator for , suppression of the transition amplitude, and a mixture of isospin state.
For the radiative capture amplitudes, whose expressions are displayed in Eqs (28), (29), (LABEL:eq;Xde), (31), we have two limits for perturbative expansion: The one appears in the denominator of the transition amplitudes; the dressed 16O propagator is expanded around the unitary limit. The other appears in the numerator of the transition amplitudes as loop and vertex corrections. As discussed above, the perturbative expansion in the denominator in terms of the effective range expansion is valid at and does not converge at the energies where the experimental data are available for fitting. Meanwhile, because the phase shift data for are reproduced very well (as we will see in Fig. 4) by means of the effective range expansion, we treat the dressed 16O propagator as a non-perturbative quantity and perturbatively expand the transition amplitudes around it.
An order of an amplitude from each of the diagrams is found by counting the number of momenta of vertices and propagators in a Feynman diagram; one has a LO amplitude from the diagram (c) because the contact -- vertex of the term does not have a momentum dependence, and NLO amplitudes from the other diagrams in Fig. 3. One may notice a large suppression factor, , appearing in ; . Similar suppression effect can be found in and as well; we denote those amplitudes as , and when changing the minus sign to the plus one in the front of the spherical Bessel function in Eqs. (28) and (LABEL:eq;Xde), we do them as . We thus have and at the energy range, MeV, at which we fit the parameters to the experimental data in the next subsection. The suppression effect is common among those amplitudes from the diagrams (a), (b), (d), (e), (f) at NLO. Thus, the radiative capture rate will be well controlled by the coefficient from the diagram (c) at LO.
The strong suppression effect mentioned above is well known; the transition is strongly suppressed between isospin-zero () nuclei. This mechanism is recently reviewed and studied for Li reaction by Baye and Tursunov bt-jpg18 . In the standard microscopic calculations with the long-wavelength approximation, the term proportional to vanishes because of the standard choice of mass of nuclei as where is the mass number of -th nucleus and is the nucleon mass. We have strongly suppressed but non-zero contribution above because of the use of the physical masses for and 12C. The small but non-vanishing transition for the cases has intensively been studied in the microscopic calculations and can be accounted by two effects: The one is the second order term of the multipole operator in the long-wavelength approximation db-plb83 , and the other is due to the mixture of the small configuration in the actual nuclei db-npa86 . In the present approach, the first one may be difficult to incorporate in the point-like particles while the second one could be introduced from a contribution at high energy: At MeV and 8.5 MeV above the -12C breakup threshold, -15N and -15O breakup channels, respectively, are open, and resonant states of 16O start emerging (along with the isobars, 16N, 16O, and 16F). We might have introduced the -15N and -15O fields as relevant degrees of freedom in the theory. The -15N and -15O fields, then, appear in the intermediate states, as -15N or -15O propagation, in the loop diagrams (d), (e), (f) in Fig. 3 instead of the -12C propagation. One may introduce a mixture of the isospin and states in the -15N or -15O propagation, and the strong suppression is circumvented in the loops. (The contribution from the -15N and -15O channels for the 12C(,)16O reaction has already been studied in the microscopic approach db-prc87 .) In our work, however, the -14N and -15O fields are regarded as irrelevant degrees of freedom at high energy and integrated out of the effective Lagrangian. Its effect, thus, is embedded in the coefficient of the contact interaction, the term, in the diagram (c) while the term is fitted to the experimental data in the next subsection.
5.3 Numerical results for the radiative capture reaction
We have five parameters in the radiative capture amplitudes to fit to the data; three parameters, , , , are fitted to the phase shift data of the elastic scattering, and the other two parameters, and , are to the experimental data. The standard -fit is performed by employing a Markov chain Monte Carlo method for the parameter fitting. 171717 See the footnote 11. The phase shift data for are taken from Tischhauser et al.’s paper tetal-prc09 , and the experimental data are from the literature summarized in Tables V and VII in Ref. detal-17 : Dyer and Barnes (1974) ex1 , Redder et al. (1987) ex2 , Ouellet et al. (1996) ex3 , Roters et al. (1999) ex4 , Gialanella et al. (2001) ex5 , Kunz et al. (2001) ex6 , Fey (2004) ex7 , Makii et al. (2009) ex8 , and Plag et al. (2012) ex9 .
As discussed in Sec. 3, we fitted the effective range parameters to the phase shift data for at MeV and displayed the fitted values of , , and in Table 2 where the number of the data is and . The uncertainties of the fitted values stem from those of the experimental data. In Fig. 4, we plot a curve of the phase shift calculated by using the fitted effective range parameters as a function of . We display the experimental data in the figure as well. One can see that the theory curve reproduces well the experimental data at the energy range, MeV.

Because the pole structures for the subthreshold state and the first resonant state of 16O are well accounted by the dressed 16O propagator, and an effect from the first excited state of 12C at MeV is appeared weak, we now may regard that an irrelevant degree of freedom at high energy for comes from the state of 16O at MeV (with the width, keV).
We fit the parameters, and , to the experimental data of at the energy range, MeV using some values of the cutoff in the range, fm, in the integral in in Eq. (LABEL:eq;Xde). The number of the data is . We find a significant cutoff dependence of the couplings, and , as well as the factor at when varying the short range cutoff, fm; as the values of become larger, become larger while the values at become smaller. (See Table 1 in Ref. a-18 .) In Fig. 5, we plot a curve of calculated by using the fitted parameters, MeV4 and MeV1/2 with fm (where ). We display the experimental data in the figure as well. One can see that the theory curve reproduces well the experimental data.

In our work, we choose the results of with at fm for our estimate of at the Gamow-peak energy, MeV, thus, we have
(34) |
where the small, about 5%, uncertainty stems from those of and as well as that of the dependence of within . The previous estimates of the factor at are well summarized in Table IV in Ref. detal-17 . The reported vales are scattered from 1 to 340 keVb with various size of the error bars. Nonetheless it is worth pointing out that our result is about 30% smaller than those reported recently: by Tang et al. (2010) tetal-prc10 , 83.4 by Schurmann et al. (2012) setal-12 , by Oulebsir et al (2012) oetal-12 , by Xu et al. (2013) xetal-npa13 , by An et al. (2015) aetal-prc15 , and 86.3 by dwBoer et al. (2017) detal-17 .
6 delayed emission from 16N
In this section, we study the delayed emission from 16N by employing an EFT. For the -matrix analysis this is an important input to estimate the -factor at while for the EFT approach this is not the case; we will discuss that though the experimental data of the delayed emission are well described in the EFT approach, it is notably different from those in the -matrix analysis. In the following subsections, the formalism of -decay and delayed emission from 16N is first discussed, and the decay amplitudes up to NNLO are derived from the effective Lagrangian. After fitting parameters to each of existing two data sets for the energy distributions, we discuss numerical results we obtained.
6.1 -decay and delayed emission from 16N
14N and 15N are stable nuclei while 16N is radioactive whose half-lifetime is sec decaying through the Gamow-Teller transition for -decay a-pr58 ; bk-np64 ; twc-npa93 . It shows a first-forbidden character of the -decay, and the ground state of 16N is identified as . Branching ratios of the decaying channels to the , , , , states of 16O and the -12C breakup channel are experimentally known as , , , , , , respectively. - and -waves are dominant for the -12C breakup channel, and its value is MeV. Recently, the branching ratios are updated by experiment as for the bound state and for the delayed emission ketal-prl18 .
Yields at the energy bins for the kinetic energy, , for the delayed emission from 16N may be obtained as
where we have defined as a dimensionless quantity, and is an overall constant, (MeV-6), which is fitted to the experimental data later. In addition, a phase space integral is given as
(36) |
where is Fermi function, and
(37) | |||||
(38) |
with , and () is the electron energy (mass).

The decay amplitudes, and , to the final -12C state for the and channels, respectively, are calculated from the diagrams depicted in Fig. 6, and we have
(39) | |||||
(40) |
where we have introduced six parameters in those amplitudes; four of them, , , , and are coefficients of the contact vertices for (a) non-pole and (b) pole diagrams for the and channels at LO. We also introduce two coefficients, and , for the vertex corrections for the diagrams (a) and (b) for the channel at NNLO. Thus, we have seven additional parameters, including the overall constant , appearing in while the effective range parameters appearing in the functions in Eq. (39) and in Eq. (40) have already been obtained in Tables 2 and 3. We use those values for the effective range parameters in the following.

In Fig. 7, a diagram for the -decay from 16N is depicted. Here we assume the perturbation expansion for the vertex correction and include the leading ones, and only. Thus, the decay rates to the final and states of 16O are obtained as
(41) | |||||
(42) |
where is the same integral as that in Eq. (36) while
(43) | |||||
(44) |
where is the mass of 16N in the ground state, and are the masses of the excited and states of 16O. In addition, and are wave-function normalization factors for the and states of 16O, and we have
(46) | |||||
where are the poly-gamma function. In the following, we fix the two parameters, and , by using the branching ratios of the -decay, and fit the remaining five parameters to the delayed emission data.
6.2 Numerical results for the delayed emission from 16N
Using the experimental data for the branching ratios of the -decay from 16N, we obtain
(47) | |||||
(48) |
Thus, five unfixed parameters, , , , , and remain in , and we fit them to the experimental data. Two sets of the experimental data for the delayed emission from 16N are available; one is from a paper by Azuma et al. aetal-prc94 , and the other is from that by Tang et al. tetal-prc10 .
In Table 5, fitted values and errors of the parameters are displayed.181818 One may check a convergence of the weak vertex correction for the -decay, the term, which we ignored in Eq. (41). Because of where is the value of the -decay to the ground state of 16O, MeV, one has about 1 % correction, and 0.0099 for the two sets of the fitted parameters.
Exp. data set | Azuma et al. | Tang et al. |
---|---|---|
(MeV-6) | ||
(MeV-2) | ||
(MeV-2) | 2.61(9) | 3.36(3) |
(MeV) | ||
(MeV3) | ||
() | 4.06 (91) | 3.56 (93) |
We include the numbers of the data () and values of in the table as well.


In Figs. 8 and 9, curves of the delayed emission from 16N are plotted as a function of the energy by using the fitted parameters obtained in Table 5. The experimental data are also included in the figures. One can see that the fitted curves reproduce the experimental data reasonably well in the two figures though the values of the parameters fitted from the reported data sets by Azuma et al. and Tang et al. obtained in the table are significantly different. Thus, as pointed out, e.g., in Ref. detal-17 , to see a convergence of experimental data for the delayed emission from 16N is important by carrying out new experiments setal-aipcp15 .
The main peak in the figures appears due to the broad resonant state of 16O while the secondary peak, as discussed before, is important to determine an interference pattern between the levels in the -matrix or the -matrix analysis. In the present approach, on the other hand, it is reproduced by an interference between the amplitudes from the non-pole and pole diagrams displayed in Fig. 6. Though one can find smaller values, 130 and 116 for 89 data points for background state energies, MeV and 20 MeV, for -matrix fit in Ref. betal-prl93 , our values are comparable to those reported in a recent -matrix analysis, for Azuma et al.’s data and 466 (88) for Tang et al.’s data, i.e., and 5.26, respectively, in Ref. detal-17 .
7 Results and discussion
In the present work, we discuss the application of EFTs to nuclear reactions at low energies. We study the elastic scattering, the radiative capture reaction, and the delayed emission from 16N for the -12C systems by constructing an EFT. This is a typical reaction for an application of EFTs because one can have a separation scale and many experimental data are available. EFTs, thus, may provide us a new theoretical method, as an alternative of -matrix or potential model analysis, for the study of nuclear-astrophysics where one needs to extrapolate a reaction rate down to a low energy by fitting some parameters of theory to experimental data measured at higher energies.
In the study of the elastic scattering, we discuss a modification of the counting rules for an expansion around the unitary limit due to the large suppression factor from the Coulomb interaction at low energies; we include up to order terms for and up to order terms for in the effective range expansion. We find that, after fixing the first effective range term by using the binding energies of 16O and fitting the remaining effective range parameters to the phase shift data, the broad resonances, and states of 16O can be described by the effective range parameters, but the narrow resonances, and states of 16O cannot. As has been studied in the previous works a-prc18 ; a-jkps18 , the expansion series converges well at while it doesn’t at the energies where the experimental data are available (though the data are well reproduced by means of the effective range expansion). In the studies for the radiative capture on 12C and the delayed emission from 16N, thus, we assume that the dressed 16O propagators for and are non-perturbative quantities, and the capture and decay amplitudes are expanded around them.
In the study of the radiative capture reaction, we calculate the radiative capture amplitudes up to NLO for the initial -wave -12C system and confirm the suppression of the transition between the iso-singlet () nuclei. We discuss that a mixture of the contribution could be introduced in the present approach by including the -15N and -15O channels, which are open at MeV and 8.5 MeV above the -12C breakup threshold. In the present work, those states are regarded as irrelevant degrees of freedom at high energy and integrated out of the Lagrangian, and its effect is embedded in the contact term, the term. After fitting the two parameters, and to the data, we make an estimate of the -factor at by employing an EFT.
In the study of the delayed emission from 16N, we calculate the decay amplitudes up to NNLO. We confirm that the primary peak of the data is accounted by the broad resonant state of 16O while we find that the secondary peak is obtained from an interference between a non-pole amplitude and a pole amplitude in the present approach. Though the study of the delayed emission from 16N is crucial in the -matrix analysis, now one may see that, except for sharing the dressed 16O propagators in the two reaction amplitudes, the coefficients of vertex functions for the radiative capture and the delayed emission are independently fixed by using the corresponding experimental data because the nuclear current for the radiative capture reaction is coupled to a vector current (or a minimally coupled photon) while that for the delayed emission reaction is to an axial-vector current. Thus, the study of the delayed emission cannot be a constraint on an estimate of the -factor in our approach. A remarkable difference between our approach and the -matrix approach can be seen in the concept for the non-pole contribution for delayed emission from 16N (whose diagram is displayed in the diagram (a) in Fig. 6). In the present approach, the non-pole contribution is systematically derived; the contact vertex functions are obtained from the effective Lagrangian in Eq. (4), and the reaction amplitudes in Eqs. (39) and (40) are calculated straightforwardly. In the -matrix or -matrix approach, there is no non-pole contribution while one necessarily introduces the so-called “background levels” jfhk-prc90 ; hfk-prc91 ; betal-prl93 . Though the coefficients of the background levels are fitted to the data as free parameters, they play the same role in reproducing the interference pattern for the secondary peak. Thus, our result may indicate that the non-pole contributions account for an origin of the background levels in the -matrix or -matrix calculations. We also find that the values of the coefficients in the delayed emission amplitudes are quite different when fitting the coefficients to each of the two experimental data sets. Thus, to see a convergence of the experimental data would be important by performing new experiments in the future.
Though we have reported a first result of the radiative capture on 12C employing an EFT, some issues remain to be explained: In the study of the elastic scattering, we have introduced a modification of the counting rules from the observation of anomaly of the expansion series compared to the experimental data while we have not investigated how the anomalous terms come out. It could appear because of our assumption of the point-like particles; a real nucleus has a finite size, and the short range contributions due to the assumption may need to be subtracted by introducing counter terms. In the study of the radiative capture reaction, we reproduce the suppression of the transition while the non-vanishing contribution, the term, is merely fitted to the experimental data. Thus it would be interesting to study a mixture of the state for the transition by including the -15N and -15O channels in the framework of EFT. It is also important to include higher order terms at NNLO in order to estimate a theoretical uncertainty of the -factor at .
As we have discussed above, an application of EFTs for nuclear reactions would be possible, provided that one can choose a clear separation scale for an observable of a reaction and the experimental data are available to fix coefficients appearing in an effective Lagrangian. Thus, those nuclear reactions at low energies, which are important in nuclear-astrophysics, are possible candidates for the application of EFTs, especially when the accuracy and the error estimate of a reaction are important. It would also be interesting to study the transition and the cascade transitions of the radiative capture on 12C at by employing an EFT. A study toward this direction is now underway.
Acknowledgements
The author would like to thank X. D. Tang, T. Kajino, A. Hosaka, and T. Sato for useful discussions and RCNP, Osaka University for hospitality during his stay when finalizing the work. This work was supported by the Basic Science Research Program through the National Research Foundation of Korea funded by the Ministry of Education of Korea (NRF-2016R1D1A1B03930122 and NRF-2019R1F1A1040362) and in part by the National Research Foundation of Korea (NRF) grant funded by the Korean government (NRF-2016K1A3A7A09005580).
Appendix
The elastic scattering amplitudes are calculated from the renormalized dressed three-point vertices and the renormalized dressed composite 16O propagators as
(49) |
where and , and we have suppressed the indices from the Cartesian tensors.
We have the renormalized dressed three-point vertices for the initial and final Coulomb interaction for as
(50) | |||||
(51) | |||||
(52) | |||||
where
(54) | |||||
(55) |
We have the renormalized dressed composite 16O propagators for as
(56) | |||||
(57) | |||||
(58) | |||||
(59) |
where , , and are the projection operators which satisfy the relation, , and we have
(60) | |||||
(61) | |||||
(62) | |||||
In addition, the couplings, are redundant when one fixes them by using the effective range parameters, conventionally one may choose them as
(63) |
References
- (1) S. Weinberg, Physica A 96, (1979) 327.
- (2) K. Ohta, Phys. Rev. C 46, (1992) 2519.
- (3) K. Ohta, Phys. Rev. C 47, (1993) 2344.
- (4) V. Bernard, N. Kaiser, T. S. H. Lee, and U.-G. Meißner, Phys. Rep. 246, (1994) 315.
- (5) S. Ando and D.-P. Min, Phys. Lett. B 417, (1998) 177.
- (6) S. Ando, F. Myhrer, and K. Kubodera, Phys. Rev. C 63, (2001) 015203.
- (7) P.F. Bedaque and U. van Kolck, Ann. Rev. Nucl. Part. Sci. 52, (2002) 339.
- (8) E. Braaten and H.-W. Hammer, Phys. Rept. 428, (2006) 259.
- (9) U.-G. Meißner, Phys. Scripta 91, (2016) 033005.
- (10) H.-W. Hammer, C. Ji, D.R. Phillips, J. Phys. G 44, (2017) 103002.
- (11) H.-W. Hammer, S. Konig, and U. van Kolck, Rev. Mod. Phys. 92, (2020) 025004.
- (12) S.-I. Ando, Eur. Phys. J. A 52, (2016) 130.
- (13) S.-I. Ando, Phys. Rev. C 97, (2018) 014604.
- (14) S.-I. Ando, J. Korean Phys. Soci. 73, (2018) 1452.
- (15) H.-E. Yoon and S.-I. Ando, J. Korean Phys. Soci. 75, (2019) 202.
- (16) S.-I. Ando, Phys. Rev. C 100, (2019) 015807.
- (17) C. Ecker and U.-G. Meißner, Comments Nucl. Part. Phys. 21, (1995) 347.
- (18) A. M. Lane and R. G. Thomas, Rev. Mod. Phys. 30, (1958) 257.
- (19) L. R. Buchmann and C. A. Barnes, Nucl. Phys. A 777, (2006) 254.
- (20) R. J. deBoer et al., Rev. Mod. Phys. 89, (2017) 035007.
- (21) J. Gasser and H. Leutwyler, Ann. Phys. (N.Y.) 158, (1984) 142.
- (22) J. Gasser and H. Leutwyler, Nucl. Phys. B 250, (1985) 465.
- (23) J. F. Donoghue, E. Golowich, and B. R. Holstein, Dynamics of the Standard Model (Cambridge University Press, 1992).
- (24) S. Weinberg, Phys. Lett. B 251, (1990) 288.
- (25) S. Weinberg, Nucl. Phys. B 363, (1991) 3.
- (26) A. Nogga, R. G. E. Timmermans, and U. van Kolck, Phys. Rev. C 72, (2005) 054006.
- (27) D. B. Kaplan, M. J. Savage, and M. B. Wise, Phys. Lett. B 424, (1998) 390.
- (28) D. B. Kaplan, M. J. Savage, and M. B. Wise, Nucl. Phys. B 534, (1998) 329.
- (29) B. Long and C.-J. Yang, Phys. Rev. C 85, (2012) 034002.
- (30) J.-W. Chen, G. Rupak, and M. J. Savage, Nucl. Phys. A 653, (1999) 386.
- (31) P. F. Bedaque, H.-W. Hammer, and U. van Kolck, Nucl. Phys. A 676, (2000) 367.
- (32) S. Ando and M. Birse, J. Phys. G 37, (2010) 105108.
- (33) V. N. Efimov, Sov. J. Nucl. Phys. 12, (1971) 589.
- (34) V. N. Efimov, Sov. J. Nucl. Phys. 29, (1979) 546.
- (35) H. A. Bethe, Phys. Rev. 76, (1949) 38.
- (36) D. B. Kaplan, Nucl. Phys. B 494, (1997) 471.
- (37) S. R. Beane and M. J. Savage, Nucl. Phys. A 694, (2001) 511.
- (38) S. Ando and C. H. Hyun, Phys. Rev. C 72, (2005) 014008.
- (39) J. Soto and J. Tarrus, Phys. Rev. C 81, (2010) 014005.
- (40) S.-I. Ando and C. H. Hyun, Phys. Rev. C 86, (2012) 024002.
- (41) S. König, H. W. Grießhammer, H.-W. Hammer, and U. van Kolck, Phye. Rev. Lett. 118, (2017) 202501.
- (42) E. Epelbaum, Ulf.-G. Meißner, and W. Glockle, Nucl. Phys. A 714, (2003) 534.
- (43) E. Braaten and H.-W. Hammer, Phys. Rev. Lett. 91, (2003) 102002.
- (44) C. Ji, Ph. D. thesis, Ohio University (2012).
- (45) C. Ji and D. R. Phillips, Few Body Sys. 54, (2013) 2317.
- (46) J. Vanasse, Phys. Rev. C 88, (2013) 044001.
- (47) J. Vanasse, Phys. Rev. C 95, (2017) 024002.
- (48) S. Ando et al., Phys. Lett. B 595, (2004) 250.
- (49) S. Ando, J. McGovern, and T. Sato, Phys. Lett. B 677, (2009) 109.
- (50) M. Butler, J.-W. Chen, and X. Kong, Phys. Rev. C 63, (2001) 035501.
- (51) S. Ando, Y. H. Song, T.-S. Park, H. W. Fearing, and K. Kubodera, Phys. Lett. B 555, (2003) 49.
- (52) S.-I. Ando, Y.-H. Song, and C. H. Hyun, Phys. Rev. C 101, (2020) 054001.
- (53) G. Rupak, Nucl. Phys. A 678, (2000) 405.
- (54) S. Ando, R. H. Cyburt, S. W. Hong, and C. H. Hyun, Phys. Rev. C 74, (2006) 025809.
- (55) X. Kong and F. Ravndal, Nucl. Phys. A 656, (1999) 421.
- (56) M. Butler and J.-W. Chen, Phys. Lett. B 520, (2001) 87.
- (57) S. Ando, J. W. Shin, C. H. Hyun, S. W. Hong, and K. Kubodera, Phys. Lett. B 668, (2008) 187.
- (58) J.-W. Chen, C.-P. Liu, and S.-H. Yu, Phys. Lett. B 720, (2013) 385.
- (59) R. Higa, G. Rupak, and A. Vaghani, Eur. Phys. J. A 54, (2018) 89.
- (60) X. Zhang, K. M. Nollett, and D. R. Phillips, J. Phys. G 47, (2020) 054002.
- (61) X. Zhang, K. M. Nollett, and D. R. Phillips, Phys. Rev. C 89, (2014) 051602(R).
- (62) E. Ryberg, C. Forssen, H.-W. Hammer, and L. Platter, Eur. Phys. J. A 50, (2014) 170.
- (63) W. A. Fowler, Rev. Mod. Phys. 56, (1984) 149.
- (64) A. Coc, F. Hammache, and J. Kiener, Eur. Phys. J. A 51, (2015) 34.
- (65) Y. Xu et al., Nucl. Phys. A 918, (2013) 61.
- (66) P. Descouvemont, D. Baye, and P.-H. Heenen, Nucl. Phys. A 430, (1984) 426.
- (67) K. Langanke and S. E. Koonin, Nucl. Phys. A 439, (1985) 384.
- (68) J. Humblet, P. Dyer, and B. A. Zimmerman, Nucl. Phys. A 271, (1976) 210.
- (69) Z.-D. An et al., Phys. Rev. C 92, (2015) 045802.
- (70) D. Baye and P. Descouvemont, Nucl. Phys. A 481, (1988) 445.
- (71) X. Ji, B. W. Filippone, J. Humblet, and S. E. Koonin, Phys. Rev. C 41, (1990) 1736.
- (72) J. Humblet, B. W. Filippone, and S. E. Koonin, Phys. Rev. C 44, (1991) 2530.
- (73) J. Jerphagnon, Phys. Rev. B 2, (1970) 1091.
- (74) S. Ando, J.W. Shin, C.H. Hyun, S.W. Hong, Phys. Rev. C 76, (2007) 064001.
- (75) S.-I. Ando, Eur. Phys. J. A 33, (2007) 185.
- (76) S. Konig, D. Lee, H.-W. Hammer, J. Phys. G: Nucl. Part. Phys. 40, (2013) 045106.
- (77) J. Hamilton, I. Overbo, B. Tromborg, Nucl. Phys. B 60, (1973) 443.
- (78) H. van Haeringen, J. Math. Phys. 18, (1977) 927.
- (79) J.-M. Sparenberg, P. Capel, and D. Baye, Phys. Rev. C 81, (2010) 011601.
- (80) Z. R. Iwinski and L. Rosenberg, Phys. Rev. C 29, (1984) 349.
- (81) P. Tischhauser et al., Phys. Rev. C 79, (2009) 055803.
- (82) P. Tischhauser et al., Phys. Rev. Lett. 88, (2002) 072501.
- (83) EPAPS-document No. E-PRVCAN-70-032904 for tables of -12C phase shift.
- (84) D. Foreman-Mackey et al., Publ. Astron. Soc. Pac. 125, (2013) 306.
- (85) T. Teichmann, Phys. Rev. 83, (1951) 141.
- (86) S.-I. Ando, Phys. Rev. C 102, (2020) 034611.
- (87) M. Katsuma, Phys. Rev. C 78, (2008) 034606.
- (88) O. L. Ramirez Suarez, J.-M. Sparenberg, Phys. Rev. C 96, (2017) 034601.
- (89) Yu. V. Orlov, B. F. Irgaziev, and J.-U. Nabi, Phys. Rev. C 96, (2017) 025809.
- (90) M. L. Avila et al., Phys. Rev. Lett. 114, (2015) 071101.
- (91) N. Oulebsir et al., Phys. Rev. C 85, (2012) 035804.
- (92) C. R. Brune et al., Phys. Rev. Lett. 83, (1999) 4025.
- (93) E. Ryberg, C. Forssen, H.-W. Hammer, and L. Platter, Phys. Rev. C 89, (2014) 014325.
- (94) A. K. Rajantie, Nucl. Phys. B 480, (1996) 729.
- (95) T. Fuchs, J. Gegelia, G. Japaridze, and S. Scherer, Phys. Rev. D 68, (2003) 056005.
- (96) S. Ando and H. W. Fearing, Phys. Rev. D 75, (2007) 014025.
- (97) D. Baye and E. M. Tursunov, J. Phys. G: Nucl. Part. Phys. 45, (2018) 085102.
- (98) P. Descouvemont and D. Baye, Phys. Lett. B 127, (1983) 286.
- (99) P. Descouvemont and D. Baye, Nucl. Phys. A 459, (1986) 374.
- (100) P. Descouvemont and D. Baye, Phys. Rev. C 36, (1987) 1249.
- (101) P. Dyer and C. Barnes, Nucl. Phys. A 233, (1974) 495.
- (102) A. Redder et al., Nucl. Phys. A 462, (1987) 385.
- (103) J. M. L. Ouellet et al., Phys. Rev. C 54, (1996) 1982.
- (104) G. Roters et al., Eur. Phys. J. A 6, (1999) 451.
- (105) L. Gialanella et al., Eur. Phys. J. A 11, (2001) 357.
- (106) R. Kunz et al., Phys. Rev. Lett. 86, (2001) 3244.
- (107) M. Fey, Ph.D. thesis (Universitat Stuttgart) (2004).
- (108) H. Makii et al., Phys. Rev. C 80, (2009) 065802.
- (109) R. Plag et al., Phys. Rev. C 86, (2012) 015805.
- (110) X. D. Tang et al., Phys. Rev. C 81, (2010) 045809.
- (111) D. Schurmann et al., Phys. Lett. B 711, (2012) 35.
- (112) N. Oulebsir et al., Phys. Rev. C 85, (2012) 035804.
- (113) D. E. Alburger, Phys. Rev. 111, (1958) 1586.
- (114) J. K. Bienlein and E. Kalsch, Nucl. Phys. 50, (1964) 202.
- (115) D. R. Tilley, H. R. Weller, and C. M. Cheves, Nucl. Phys. A 564, (1993) 1.
- (116) O. S. Kirsebom et al., Phys. Rev. Lett. 121, (2018) 142701.
- (117) R. E. Azuma et al., Phys. Rev. C 50, (1994) 1194.
- (118) S. Sanfilippo et al., AIP Conf. Proc. 1645, (2015) 387.
- (119) L. Buchmann et al., Phys. Rev. Lett. 70, (1993) 726.