Chiral Effective Field Theory Calculations of Weak Transitions in Light Nuclei
Abstract
We report Quantum Monte Carlo calculations of weak transitions in nuclei, based on the Norfolk two- and three-nucleon chiral interactions, and associated one- and two-body axial currents. We find that the contribution from two-body currents is at the 2–3% level, with the exception of matrix elements entering the rates of 8Li, 8B, and 8He beta decays. These matrix elements are suppressed in impulse approximation based on the (leading order) Gamow Teller transition operator alone; two-body currents provide a – correction, which is, however, insufficient to bring theory in agreement with experimental data. For the other transitions, the agreement with the data is satisfactory, and the results exhibit a negligible to mild model dependence when different combinations of Norfolk interactions are utilized to construct the nuclear wave functions. We report a complete study of two-body weak transition densities which reveals the expected universal behavior of two-body currents at short distances throughout the range of = to = systems considered here.
pacs:
21.10.Ky, 23.20.-g, 23.20.Js, 27.20.+nI Introduction
In this work, we present Quantum Monte Carlo (QMC) calculations, including both variational Monte Carlo (VMC) and Green’s function Monte Carlo (GFMC) calculations, of Gamow-Teller (GT) matrix elements entering beta-decay and electron-capture rates in =– nuclei. These observables are experimentally known (in most cases) at the sub-percent level, and are used here primarily to validate our microscopic theoretical modeling of the nucleus as a system of nucleons interacting with each other via effective interactions, and with electroweak probes via effective currents. Specifically, this modeling is based on local two- and three-nucleon interactions formulated in configuration space, and derived from a chiral effective field theory (EFT) that retains, in addition to nucleons and pions, -isobars as explicit degrees of freedom Piarulli et al. (2015, 2016); Baroni et al. (2016a, 2018). They are referred to below as the Norfolk interactions and are denoted as NV2+3. Accompanying these interactions are one- and two-body axial currents—local in configuration space—derived within the same EFT formulation Baroni et al. (2016b, a, 2018). In the calculations to follow, we include up to tree-level contributions at next-to-next-to-next-to-leading order (N3LO) in the chiral expansion, and disregard subleading corrections involving loops and higher order contact terms.
An analogous study was recently reported by some of the present authors in Ref. Pastore et al. (2018a). There, the QMC calculations were based on the Argonne- (AV18) two-nucleon Wiringa et al. (1995) and Illinois-7 (IL-7) three-nucleon Pieper (2008) interactions, in combination with the axial currents of Ref. Baroni et al. (2016b). We found agreement with the experimental Gamow-Teller matrix elements at the level for the = and systems, and at the level in 10C. The large uncertainty in the = transition is primarily systematic and results from the narrow energy separation between the first two = states in 10B, which makes it hard to precisely disentangle them Pastore et al. (2018a). The study of Ref. Pastore et al. (2018a) found that two-body currents generate an additive contribution of the order of and concluded that the agreement with the data is mainly attributable to the use of fully correlated nuclear wave functions, rather than two-body effects in the currents.
In the meantime, no-core shell-model calculations of weak matrix elements based on chiral interactions and currents Gysbers et al. (2019) found the sign of the overall correction generated by two-body currents to be opposite to that obtained in Ref. Pastore et al. (2018a) for the same systems (but in agreement with a hybrid calculation of the = decay reported in Ref. Gazit et al. (2009a)). This discrepancy was attributed to the hybrid nature of the calculation of Ref. Pastore et al. (2018a), i.e., to the mismatch between the two- and three-body correlations implemented to construct the nuclear wave functions—and induced by the AV18 and IL7—and those entering the axial currents which were instead derived from EFT.
In this work, by reexamining the evaluation of these weak matrix elements with the NV2+3 chiral interactions Piarulli et al. (2015, 2016); Baroni et al. (2016a, 2018), in combination with consistent chiral axial currents at tree-level Baroni et al. (2016b), we aim to address and explore the aforementioned claim. We investigate the sensitivity of the calculated matrix elements with respect to different choices of regulators and to different strategies adopted to constrain the three-body Norfolk interactions (NV3). This latter aspect is important in order to understand the interplay between these interactions and the axial currents, since the strength of the contact current and that of the three-body interaction of one-pion-range are rigorously related to each other by the symmetries imposed in the EFT formulation Gardestig and Phillips (2006); Gazit et al. (2009b); Schiavilla (2017).
This study has several merits. First, we report new GFMC results of the energy spectra of nuclei based on two classes of NV2+3 interactions. In addition to the systems studied in Ref. Pastore et al. (2018a), we study weak transitions in = nuclei where we find that two-body axial currents provide a large correction to the one-body results. Finally, we provide the first calculations of two-body weak transition densities which shed light on the role of short-range physics in these observables.
Searches for physics beyond the Standard Model (BSM) via beta decay are the focus of current and planned experimental programs carried out at the Facility for Rare Isotope Beams (FRIB), the the University of Washington, and Argonne National Laboratory (see, e.g., Ref. Gonzalez-Alonso et al. (2019) and references therein). Among the targets under consideration are 6He, 8Li, 8B and 10C. A systematic study of axial-current matrix elements in these systems is a prerequisite for all further investigations and BSM searches in beta decay. Beta decays are ideal processes to assess the validity of the dynamical inputs of ab inito calculations, namely many-body correlations and weak currents. The latter also impact calculations of neutrinoless double beta decay matrix elements, whose knowledge is critical to the neutrinoless double beta decay experimental program Engel and Menendez (2017).
II Quantum Monte Carlo method
The Quantum Monte Carlo methods used in this study have been described in detail in several review articles, the most recent of which being Refs. Carlson et al. (2015); Lynn et al. (2019). Here, we sketch the computational procedure and refer the interested reader to Refs. Carlson et al. (2015); Lynn et al. (2019) and references therein.
We seek accurate solutions of the many-nucleon Schrödinger equation
(1) |
where are the total angular momentum and parity of the state, and and are the total isospin and its projection, respectively. We use the Hamiltonian
(2) |
where is the non-relativistic kinetic energy, and and are the NV2 and NV3 local chiral interactions Piarulli et al. (2015, 2016); Baroni et al. (2016a, 2018), collectively denoted as NV2+3.
The VMC trial function for a given nucleus is constructed from products of two- and three-body correlation operators acting on an antisymmetric single-particle state of the appropriate quantum numbers. The correlation operators are designed to reflect the influence of the interactions at short distances, while appropriate boundary conditions are imposed at long range Wiringa (1991); Pudliner et al. (1997); Nollett et al. (2001); Nollett (2001). The contains variational parameters that are adjusted to minimize the expectation value
(3) |
which is evaluated by Metropolis Monte Carlo integration Metropolis et al. (1953). The lowest value for is then taken as the approximate ground-state energy of the exact lowest eigenvalue of , , for the specified quantum numbers.
A good trial wave function is given by
(4) |
The Jastrow wave function is fully antisymmetric and has the quantum numbers of the state of interest, while and are two- and three-body correlation operators.
The GFMC method Carlson et al. (2015) improves on the VMC wave functions by acting on with the operator . The operator is applied in a sequence of small imaginary-time steps to produce a propagated wave function
(5) |
Obviously = and =. Quantities of interest are evaluated in terms of a “mixed” expectation value between and :
(6) |
where the operator acts on the trial function . The desired expectation values would, of course, have on both sides; by writing and neglecting terms of order , we obtain the approximate expression
(7) | |||||
where is the variational expectation value.
For off-diagonal matrix elements required by the transitions we are interested here, the generalized mixed estimate is given by the expression
(8) | |||||
where
(9) |
and is defined similarly. For more details see Eqs. (19)–(24) and the accompanying discussions in Ref. Pervin et al. (2007).
III Norfolk interaction models
We base our calculations of weak transitions in =6–10 on the local NV2 and NV3 interactions developed in Refs. Piarulli et al. (2015, 2016); Baroni et al. (2016a, 2018). The NV2 model has been derived from a EFT that uses pions, nucleons and ’s as fundamental degrees of freedom. It consists of a long-range part, , mediated by one- and two-pion exchanges, and a short-range part, , described in terms of contact interactions with strengths specified by unknown low-energy constants (LECs). The strength of the long-range part is fully determined by the nucleon and nucleon-to- axial coupling constants and , the pion decay amplitude , and the LECs , , , , and , constrained by reproducing scattering data Krebs et al. (2007). The LECs entering the contact interactions are fixed by fitting nucleon-nucleon scattering data from the most recent and up-to-date database collected by the Granada group Navarro Pérez et al. (2013, 2014); Navarro Perez et al. (2014). The value for is taken from the large expansion or strong-coupling model Green (1976). The value of the nucleon axial coupling constant used to construct the nucleon-nucleon interaction accounts for the Goldberger-Treiman discrepancy Arndt et al. (1994); Stoks et al. (1993) and, to distinguish it from the experimental value of = Tanabashi et al. (2018) entering the axial currents, we denote it with . For completeness, Tables 1 and 2 report the values of these constants, along with the pion and nucleon masses, the -nucleon mass difference, the electron mass, and the fine structure constant used in the NV2 interactions (these last two characterize the electromagnetic part of the NV2s Piarulli et al. (2015)).
The contact terms are implemented using a Gaussian representation of the three dimensional delta function, with denoting the Gaussian parameter Piarulli et al. (2016, 2015); Baroni et al. (2016a, 2018). The pion-range operators are strongly singular at short range in configuration space, and are regularized by a radial function characterized by a cutoff Piarulli et al. (2016, 2015); Baroni et al. (2016a, 2018). There are two classes (I and II) of NV2s, differing only in the range of energy over which they are fitted to the database—class I up to 125 MeV, and class II up to 200 MeV. For each class, two combinations of short- and long-range regulators have been used, namely (, )=(0.8, 1.2) fm (models NV2-Ia and NV2-IIa) and (, )=(0.7, 1.0) fm (models NV2-Ib and NV2-IIb). Class I (II) fits about 2700 (3700) data points with a /datum () Piarulli et al. (2016, 2015).

The NV2 models were found to provide insufficient attraction in GFMC calculations of the binding energies of light nuclei Piarulli et al. (2016). To remedy this shortcoming, a consistent three-body interaction was constructed up to N2LO in the chiral expansion. It consists of a long-range part mediated by two-pion exchange and a short-range part parametrized in terms of two contact interactions van Kolck (1994); Epelbaum et al. (2002) proportional to the LECs and . These LECs have been obtained by fitting either observables that involve exclusively strong interactions Lynn et al. (2016); Tews et al. (2016); Lynn et al. (2017); Piarulli et al. (2018) or a combination of observables that involve both strong and weak interactions Gazit et al. (2009b); Marcucci et al. (2012); Baroni et al. (2018). This last strategy is feasible because of the relation established in EFT Gardestig and Phillips (2006) that links with the LECs entering the contact axial current at N3LO Gazit et al. (2009b); Marcucci et al. (2012); Schiavilla (2017) (see next section for details).
In Ref. Piarulli et al. (2018), and were determined by simultaneously reproducing the experimental trinucleon ground-state energies and doublet scattering length. These first-generation NV2+3 interactions, denoted with NV2+3-Ia/b and NV2+3-IIa/b, have been implemented in both VMC and GFMC codes and used to study static properties of light nuclei Piarulli et al. (2016, 2018); Lynn et al. (2019); Cirigliano et al. (2019); Piarulli and Tews (2020); Gandolfi et al. (2020), and in auxiliary-field diffusion Monte Carlo (AFDMC) Schmidt and Fantoni (1999), Brueckner-Bethe-Goldstone (BBG) Day (1967); Baldo and Burgio (2012) and Fermi hypernetted chain/single-operator chain (FHNC/SOC) Fantoni and Rosati (1975); Pandharipande and Wiringa (1979) approaches to investigate the equation of state of neutron matter Piarulli et al. (2019); Bombaci and Logoteta (2018).
In more recent work Baroni et al. (2018), and were constrained by fitting, in addition to the trinucleon energies, the empirical value of the GT matrix element in tritium decay. These second-generation NV2+3 interactions were designated as NV2+3-Ia∗/b∗ and NV2+3-IIa∗/b∗. These two different procedures for fixing and produced rather different values for these LECs. They are reported in Table 3.
Ia (Ia*) | Ib (Ib*) | IIa (IIa*) | IIb (IIb*) | |
---|---|---|---|---|
3.666 (–0.635) | –2.061 (–4.71) | 1.278 (–0.61) | –4.480 (–5.25) | |
–1.638 (–0.090) | –0.982 ( 0.55) | –1.029 (–0.35) | –0.412 (0.05) | |
0.090 ( 1.035) | 2.013 (2.881) | 0.615 (1.03) | 2.806 (3.059) |
IV EFT axial currents

Many-body axial currents have been first examined within EFT by Park and collaborators in Ref. Park et al. (1993). In that work, the authors retained pions and nucleons in their effective theory and calculated the two-body axial currents up to one-loop terms. The derivation neglected, for example, pion-pole contributions. More recently, two-body axial currents with pions and nucleons have been derived by the Bonn group Krebs et al. (2017) using the unitary transformation method, and by the JLab-Pisa group using time-ordered perturbation theory Baroni et al. (2016b, a). The two derivations differ in the treatment of reducible diagrams. When calculating box diagrams entering the electromagnetic charge and current operators Pastore et al. (2008, 2009, 2011); Kolling et al. (2009, 2011), the two methods lead to results that are in agreement. However, as discussed at length in Refs. Baroni et al. (2016a); Krebs et al. (2020), the two groups find different results for the box diagrams in the two-body axial current operator at N4LO. The numerical impact of this difference has been investigated in Refs. Baroni et al. (2016a, 2018), where both the JLab-Pisa and Bonn versions of the N4LO current operators have been implemented to calculate the GT matrix element in triton beta decay. In those studies, it was found that the corrections generated by the JLab-Pisa and Bonn N4LO operators are qualitatively in agreement (they both quench the GT matrix element at leading order), and provide, respectively, a and contribution to the total GT matrix element.
Here, we consider two-body axial currents derived within the same EFT used to construct the NV2+3 interactions Baroni et al. (2018). Moreover, we base our calculations on tree-level corrections only, and disregard the (problematic) N4LO loop contributions discussed above. This choice is advantageous also because it allows for a clearer comparison with the no-core shell model and coupled-cluster calculations of Ref. Gysbers et al. (2019), which are also based on tree-level axial currents alone (albeit derived in a -less EFT). Corrections at N4LO in the present formulation are, in practice, subsumed in the LECs of the theory, which have been determined by fits to experimental data.

Before moving on to a (brief) discussion of these axial currents, it is worthwhile pointing out that many-body corrections to leading one-body transition operators have been shown to be crucial for providing a quantitatively successful description of many nuclear electroweak observables Bacca and Pastore (2014), such as nuclear electromagnetic form factors Piarulli et al. (2013); Schiavilla et al. (2019); Nevo Dinur et al. (2019); Marcucci et al. (2016), low-energy electroweak transitions Pastore et al. (2009); Girlanda et al. (2010); Pastore et al. (2011, 2013); Datar et al. (2013); Pastore et al. (2014); Pastore et al. (2018a), and electroweak scattering Pastore et al. (2019). They have also been used in studies of double beta decay matrix elements Pastore et al. (2018b); Cirigliano et al. (2018, 2019); Wang et al. (2019).
The N3LO axial currents used in this work are represented diagrammatically Fig. 1. We refer the interested reader to Refs. Baroni et al. (2016b, 2018) for additional details and explicit expressions of the operators; here, we only note that we do not show diagrams that lead to vanishing contributions as well as pion-pole terms which give negligible corrections to the matrix elements under study.
The LO term, which scales as in the power counting ( denotes generically a low-momentum scale), is shown in panel (a) and reads
(10) |
where is the nucleon axial coupling constant (= Tanabashi et al. (2018)), is the pion-decay amplitude (= MeV), and are the spin and isospin Pauli matrices of nucleon , is the external field momentum, is the position of nucleon , and the subscript specifies the isospin component (=).
At N2LO there are two contributions (scaling as ). The first one is a relativistic correction to the single-nucleon operator at LO and is diagrammatically illustrated in panel (b), while the second involves the excitation of a nucleon into a by pion exchange, as illustrated in panel (c). In the tables and figures below, we will denote these two contributions with N2LO-RC and N2LO-, respectively. We use the same notation introduced in Ref. Baroni et al. (2018) and write the cumulative N2LO contribution as
(11) |
At N3LO (or in the chiral expansion), there is a term of one-pion range illustrated in panel (d), and a contact term shown in panel (e), which together give the following N3LO correction
(12) |
We will denote the individual terms with N3LO-OPE and N3LO-CT, respectively.
The configuration-space expressions of these currents are given in Eqs. (2.7)–(2.10) of Ref. Baroni et al. (2018). Here, we limit ourselves to report the expression of the N3LO contact term used in this work to explicitly show the relation between the LECs entering this axial current and the LEC in the three-nucleon interaction. In -space the N3LO-CT current reads
(13) |
where
(14) |
and is the interparticle distance. The -function in the contact axial current has been smeared by replacing it with a Gaussian cutoff of range ,
(15) |
as previously done for the contact-like terms of the NV2 interactions. The adimensional LEC (reported in Table 3) is given by
(16) | |||||
where is the LEC multiplying one of the contact terms in the three-nucleon interaction Epelbaum et al. (2002) given in Table 3, GeV is the chiral symmetry breaking scale, and are given in Table 1, and and are the average nucleon and pion masses. It has recently been realized Schiavilla (2017) that the relation between and had been given erroneously in the original reference Gazit et al. (2009b), a – sign and a factor 1/4 were missing in the term proportional to .
s.s. | [MeV] | |||
---|---|---|---|---|
Ia* | Ia* | Expt. | ||
6Li | [42] | –31.97(6) | –31.06(8) | –31.99 |
6He | [42] | –29.32(4) | –28.46(5) | –29.27 |
7Li | [43] | –39.25(15) | –38.27(14) | –39.25 |
7Li | [43] | –39.18(15) | –37.66(15) | –38.76 |
7Be | [43] | –37.75(8) | –36.56(10) | –37.60 |
8He | [422] | –31.33(7) | –28.53(6) | –31.40 |
8Li | [431] | –41.59(10) | –38.89(7) | –41.28 |
8Li | [431] | –40.59(7) | –37.78(7) | –40.30 |
8B | [431] | –37.87(8) | –35.63(8) | –37.74 |
8Be | [44] | –54.07(7) | –53.16(11) | –53.47 |
10B | [442] | –64.61(41) | –60.46(30) | –64.03 |
10C | [442] | –61.01(50) | –56.65(22) | –60.32 |
Transition | Model | LO | N2LO-(RC+) | N3LO-OPE | N3LO-CT | Total–LO | Total | Expt. |
---|---|---|---|---|---|---|---|---|
6He(;) 6Li(;) | IaI (Ib) | 2.200 (2.254) | 0.022 (0.056) | 0.039 (0.064) | –0.005 (–0.068) | 0.056 (0.052) | 2.256 (2.306) | 2.1609(40) |
[42][42] | IIa (IIb) | 2.207 (2.212) | 0.027 (0.043) | 0.043 (0.055) | –0.034 (–0.082) | 0.036 (0.016) | 2.243 (2.228) | |
Ia*I (Ib*) | 2.192 (2.256) | 0.021 (0.056) | 0.038 (0.063) | –0.054 (–0.097) | 0.005 (0.022) | 2.197 (2.279) | ||
IIa* (IIb*) | 2.202 (2.218) | 0.027 (0.044) | 0.043 (0.056) | –0.057 (–0.090) | 0.014 (0.010) | 2.216 (2.228) | ||
7Be(;)Li(;) | IaI (Ib) | 2.317 (2.294) | 0.099 (0.162) | 0.076 (0.118) | –0.010 (–0.148) | 0.165 (0.133) | 2.482 (2.427) | 2.3556(47) |
[43][43] | IIa (IIb) | 2.293 (2.309) | 0.102 (0.153) | 0.078 (0.113) | –0.070 (–0.190) | 0.110 (0.076) | 2.403 (2.385) | |
Ia*I (Ib*) | 2.327 (2.307) | 0.098 (0.161) | 0.076 (0.117) | –0.121 (–0.212) | 0.053 (0.066) | 2.380 (2.373) | ||
IIa* (IIb*) | 2.296 (2.316) | 0.103 (0.154) | 0.078 (0.114) | –0.120 (–0.210) | 0.061 (0.058) | 2.357 (2.374) | ||
7Be(;)Li(;) | IaI (Ib) | 2.157 (2.119) | 0.066 (0.122) | 0.063 (0.100) | –0.009 (–0.125) | 0.121 (0.096) | 2.278 (2.215) | 2.1116(57) |
[43][43] | IIa (IIb) | 2.128 (2.145) | 0.069 (0.111) | 0.065 (0.095) | –0.059 (–0.162) | 0.074 (0.044) | 2.202 (2.189) | |
Ia*I (Ib*) | 2.158 (2.124) | 0.065 (0.119) | 0.063 (0.099) | –0.103 (–0.180) | 0.025 (0.038) | 2.183 (2.162) | ||
IIa* (IIb*) | 2.131 (2.148) | 0.067 (0.111) | 0.064 (0.095) | –0.101 (–0.178) | 0.030 (0.028) | 2.161 (2.176) | ||
8Li(;1)Be(;) | IaI (Ib) | 0.147 (0.092) | 0.032 (0.028) | 0.011 (0.011) | –0.001 (–0.014) | 0.041 (0.031) | 0.188 (0.123) | 0.284 Ref. Warburton (1986) |
[431][44] | IIa (IIb) | 0.144 (0.101) | 0.031 (0.033) | 0.010 (0.011) | –0.008 (–0.019) | 0.033 (0.025) | 0.177 (0.126) | 0.190 Ref. Pritychenko et al. (2011) |
Ia*I (Ib*) | 0.148 (0.099) | 0.032 (0.033) | 0.010 (0.012) | –0.016 (–0.020) | 0.026 (0.025) | 0.174 (0.124) | ||
IIa* (IIb*) | 0.124 (0.121) | 0.032 (0.037) | 0.010 (0.013) | –0.014 (–0.023) | 0.028 (0.027) | 0.152 (0.148) | ||
8B(;1)Be(;) | IaI (Ib) | 0.146 (0.092) | 0.032 (0.032) | 0.011 (0.011) | –0.001 (–0.014) | 0.042 (0.030) | 0.188 (0.122) | 0.269(20) |
[431][44] | IIa (IIb) | 0.144 (0.102) | 0.031 (0.033) | 0.010 (0.011) | –0.008 (–0.019) | 0.033 (0.026) | 0.177 (0.128) | |
Ia*I (Ib*) | 0.148 (0.098) | 0.032 (0.034) | 0.010 (0.012) | –0.016 (–0.020) | 0.026 (0.025) | 0.174 (0.123) | ||
IIa* (IIb*) | 0.126 (0.118) | 0.032 (0.037) | 0.010 (0.013) | –0.014 (–0.022) | 0.028 (0.027) | 0.154 (0.145) | ||
8He(;2)Li(;) | IaI (Ib) | 0.386 (0.363) | 0.030 (0.034) | 0.009 (0.012) | –0.001 (–0.014) | 0.038 (0.032) | 0.424 (0.396) | 0.512(6) |
[422][431] | IIa (IIb) | 0.465 (0.370) | 0.032 (0.034) | 0.012 (0.011) | –0.009 (–0.017) | 0.035 (0.028) | 0.500 (0.398) | |
Ia*I (Ib*) | 0.362 (0.377) | 0.031 (0.034) | 0.009 (0.014) | –0.010 (–0.022) | 0.029 (0.026) | 0.391 (0.402) | ||
IIa* (IIb*) | 0.481 (0.364) | 0.033 (0.035) | 0.012 (0.012) | –0.017 (–0.019) | 0.029 (0.028) | 0.510 (0.391) | ||
10C(;1)B(;) | IaI (Ib) | 1.940 (2.118) | 0.003 (0.026) | 0.044 (0.068) | –0.006 (–0.081) | 0.041 (0.011) | 1.981 (2.129) | 1.8331(34) |
[442][442] | IIa (IIb) | 2.157 (2.176) | -0.002 (0.026) | 0.046 (0.072) | –0.042 (–0.120) | 0.002 (–0.022) | 2.159 (2.154) | |
Ia*I (Ib*) | 2.015 (2.123) | 0.024 (0.022) | 0.059 (0.069) | –0.119 (–0.117) | –0.037 (–0.028) | 1.978 (2.095) | ||
IIa* (IIb*) | 2.071 (2.143) | 0.015 (0.031) | 0.061 (0.072) | –0.126 (–0.123) | –0.050 (–0.020) | 2.021 (2.123) |
Transition | Model | s.s. | LO | Total–LO | Total | Expt. |
---|---|---|---|---|---|---|
6He(;) 6Li(;) | IaI* | [42][42] | 2.130(2.200) | 0.070(0.056) | 2.201(2.256) | 2.1609(40) |
Ia*I | 2.107(2.192) | 0.011(0.005) | 2.118(2.197) | |||
Ref. Pastore et al. (2018a) | 2.168(2.174) | 0.037(0.030) | 2.205(2.211) | |||
7Be(;)Li(;) | IaI* | [43][43] | 2.273(2.317) | 0.164(0.165) | 2.440(2.482) | 2.3556(47) |
Ia*I | 2.286(2.327) | 0.052(0.053) | 2.338(2.380) | |||
Ref. Pastore et al. (2018a) | 2.294(2.334) | 0.061(0.050) | 2.355(2.384) | |||
7Be(;)Li(;) | IaI* | [43][43] | 2.065(2.157) | 0.103(0.121) | 2.168(2.278) | 2.1116(57) |
Ia*I | 2.061(2.158) | 0.009(0.025) | 2.070(2.183) | |||
Ref. Pastore et al. (2018a) | 2.083(2.150) | 0.046(0.046) | 2.129(2.196) | |||
8Li(;1)Be(;) | IaI* | [431][44] | 0.074(0.147) | 0.029(0.041) | 0.103(0.188) | 0.284 Ref. Warburton (1986) |
Ia*I | 0.096(0.148) | 0.025(0.026) | 0.120(0.174) | 0.190 Ref. Pritychenko et al. (2011) | ||
8B(;1)Be(;) | IaI* | [431][44] | 0.091(0.146) | 0.035(0.042) | 0.125(0.188) | 0.269(20) |
Ia*I | 0.102(0.148) | 0.024(0.026) | 0.126(0.174) | |||
8He(;2)Li(;) | IaI* | [422][431] | 0.262(0.386) | 0.040(0.038) | 0.302(0.424) | 0.512(6) |
Ia*I | 0.297(0.362) | 0.025(0.029) | 0.322(0.391) | |||
10C(;1)B(;) | IaI* | [442][442] | 1.928(1.940) | 0.050(0.041) | 1.978(1.981) | 1.8331(34) |
Ia*I | 2.086(2.015) | –0.031(–0.037) | 2.055(1.978) | |||
Ref. Pastore et al. (2018a) | 2.032(2.062) | 0.016(0.015) | 2.048(2.077) |
V Results
The GFMC energies of the nuclei of interest calculated using the NV2+3-Ia and NV2+3-Ia* models are listed in Table 4 along with the dominant spatial symmetry (s.s.) of the variational wave functions Wiringa (2006). The energies are obtained using walkers, and are all well converged by 30 unconstrained steps Wiringa et al. (2000a). All the GFMC results presented in this article (but for the two cases discussed below) are averages over the imaginary time from 0.2 to 0.82 MeV-1. Results obtained with the NV2+3-Ia interaction are in statistical agreement with those published in Ref. Piarulli et al. (2018) based on the same nuclear Hamiltonian. Model NV2+3-Ia leads to predictions that are in excellent agreement with the data. We also report for the first time results based on the second generation of NV2+3 interactions, specifically model NV2+3-Ia*, whose three-nucleon interaction has been constrained by fits to the experimental trinucleon binding energies and tritium GT matrix element. Results obtained with the NV2+3-Ia∗ Hamiltonian display a somewhat less satisfactory agreement with the experimental data, but still less than 4% away from them.
A typical imaginary-time evolution of the GFMC transition matrix elements is shown in Fig. 2. As can be seen, there is a rapid drop of 3% from the initial VMC estimate at = that reaches a stable value around 0.2 MeV-1. The results for all transitions presented in this article are averages over from 0.2 to 0.82 MeV-1, as indicated by the dashed lines, with statistical errors denoted by the solid lines. The calculations of weak transitions involving the state of 8Be and the ground state of 8B are treated differently. For these two states, we observe that the binding energy, magnitude of the quadrupole moment, and point-proton radius all increase monotonically as the imaginary time increases. This can be appreciated in Fig. 3 where we show the point-proton radii of the two nuclear states. We interpret this behavior as an indication that the resonant excited state of 8Be is dissolving into two separated ’s, while 8B is breaking into Be. In the case of 8Be, this issue has been addressed already in Refs. Wiringa et al. (2000b); Pastore et al. (2014); Datar et al. (2013). Here, we use similar techniques to treat these systems and extract matrix elements from the GFMC data. In particular, we note that the ground-state energy of 8Be drops very quickly as the imaginary time increases and reaches stability around MeV-1. The transition matrix elements involving the two dissolving states have been determined assuming that, also for these states, MeV-1 is the point at which spurious contaminations in the nuclear wave functions have been removed by the GFMC propagation. We then average in a small interval around this point, typically between 0.06 and 0.14 MeV-1. Calculations involving these two systems are clearly affected by a systematic error. To have a rather rough estimate of this error, we study the sensitivity of the extracted matrix elements with respect to variations in the imaginary time interval selected for the averaging. We find that such a procedure generates an additional uncertainty of which is added in quadrature to the statistical one, and quoted in Table 6 below of GFMC results.


V.1 Weak Transitions in =6–10 Nuclei
In this section, we present results for the GT reduced matrix element (RME) defined as
(17) |
where is the -component (at vanishing momentum transfer) of the charge-raising/lowering current with =, and is a Clebsch-Gordan coefficient.
Results for the GT RMEs in =–10 nuclei based on variational wave functions are reported in Table 5 for the eight different NV2+3 interactions discussed above, namely the NV2+3-Ia/b, NV2+3-IIa/b and corresponding starred models. The one-body axial current at LO, illustrated in panel (a) of Fig. 1, leads to contributions to the matrix elements reported in the third column of Table 5. One-body relativistic corrections (N2LO-RC) and two-body currents of one-pion range (N2LO-) at N2LO, displayed in panels (b) and (c) of Fig. 1, are added up and given in the fourth column of Table 5 labeled with N2LO-(RC+). A rough estimate of the size of the RC corrections can be obtained suppressing each LO term by a factor of , where we used a “typical” nucleon’s low-momentum MeV. Contributions at N3LO are given in the columns labeled by N3LO-OPE and N3LO-CT, corresponding to the one-pion range and contact currents (and displayed in panels (d) and (e) of Fig. 1). The cumulative contributions are given in the next to last column, while the contributions beyond LO only in the column labeled “Total–LO”. Experimental data from Refs. Knecht et al. (2012); Suzuki et al. (2003); Chou et al. (1993); Warburton (1986); Pritychenko et al. (2011) are reported in the last column of Table 5.
All the calculations use axial currents at tree-level which are consistent with the specific NV2+3 model used to generate the VMC wave functions. VMC results based on different nuclear Hamiltonians are qualitatively in agreement. In particular, for the and systems the LO contribution provides about of the total matrix elements with currents beyond LO giving the remaining correction. This correction adds up constructively to the LO contribution for all nuclei being considered, but for the = transition. For this last transition, we find that the contributions beyond LO give a correction that quenches the LO results obtained with all the starred models, and with the un-starred NV2+3-IIb interaction. More details about this calculation will be given in the following section. We emphasize that the “Total–LO” column includes, in addition to two-body contributions, also a small correction resulting from the one-body N2LO-RC current.
Transitions involving = nuclei exhibit a large suppression at LO. This behavior is attributable to the fact that the initial and final VMC wave functions are characterized by different dominant spatial symmetries, which make their overlap small compared to cases in which both the initial and final states display the same dominant spatial symmetry. As a consequence, in these cases the LO term is only about 40–50% of the total matrix element, with two-body currents providing a large correction. Two-body currents, while improving the agreement with the experimental values, are insufficient to fully explain them. Because of the reduced overlap between dominant components in the wave functions, these matrix elements are particularly sensitive to small components, which are poorly constrained and model dependent. This can be appreciated by looking at the one-body transition densities, , defined as
(18) |
where is the distance of nucleon from the center-of-mass of the system.
In Figs. 4 and 5, we show one-body densities for two transitions, namely the -capture of the 7Be ground state to the 7Li ground state and the 8Li -decay. The former, involves initial and final states with the same [43] (dominant) spatial symmetry, while for the latter the initial state is in a [431] spatial symmetry configuration and the final state is in a [44] one. The densities are calculated using the NV2+3-Ia/b and NV2+3-IIa/b interactions. From the figures we can see that the = one-body densities are well constrained and essentially model independent, while the = ones are particularly sensitive to the nuclear Hamiltonian used to generate the wave functions. Of course, these considerations are based on VMC results. A GFMC propagation might mitigate the observed model dependence.

In all the cases we studied, the contact-like current at N3LO provides a correction that quenches the LO terms, while the currents of one-pion range at N2LO and N3LO add up constructively to the LO contributions (see Table 5). The main difference between the un-starred and starred calculations is observed in the size of the contact contribution. In particular, starred models are characterized by a larger value of (see Table 3), which in turns leads to a larger (in magnitude) N3LO-CT correction.
We performed GFMC propagations only for the NV2+3-Ia and NV2+3-Ia* models. GFMC results are reported in Table 6, where, for completeness, we also show the corresponding VMC values in parentheses along with the GFMC results from Ref. Pastore et al. (2018a). We summarize the GFMC results in Fig. 6 and compare them (where possible) to the results of Ref. Pastore et al. (2018a) based on the AV18+IL7 nuclear Hamiltonian.
The effect of the GFMC propagation in imaginary time is to reduce the VMC results by in all selected transitions (but for the = transition obtained with the NV2+3-Ia* model). The agreement with the data, after the inclusion of two-body currents, is at the () level for the = transition with the NV2+3-Ia (NV2+3-Ia*) model; and at the () level for the = cases with the NV2+3-Ia (NV2+3-Ia*) model. These results are in agreement with those obtained for the same transitions in the calculations of Ref. Pastore et al. (2018a) which were based on the AV18+IL7 interactions. The NV2+3 models lead to a more satisfactory agreement with the data for the = RME primarily because, with these interactions, the LO term is smaller than obtained using AV18-IL7 model.
The largest discrepancy generated by the use of different nuclear Hamiltonians, including AV18+IL7, is observed in the = transition. This can be appreciated looking at both Table 5 and Table 6. From the former, we observe a rather large cutoff dependence (models a vs models b), and also a large sensitivity to the class (either I or II) used to generate the nuclear wave functions. From Table 6, we see that the results of Ref. Pastore et al. (2018a), based on the AV18+IL7 Hamiltonian, lie between models Ia and Ia*. This large model and cutoff dependence can be traced back to the existence of two nearby = excited states in 10B, the lower one a predominantly 3S1[442] state and the upper one a 3D1[442] state (in coupling), which are only 1 MeV apart. The transition from the 10C(0+) state, which is predominantly 1S0[442], is large in the components, but about five times smaller in the components. This makes the GT matrix element particularly sensitive to the exact mixing of the 3S1 and 3D1 components in the two 10B(1+) states produced by a given Hamiltonian. It would appear that none of the interactions models studied here gets quite the right mixing of these components. In particular, results based on the NV2+3-Ia and NV2+3-Ia* interactions over-predict the data by and , respectively, which gives an indication of the spread of the theoretical estimates.
Predictions for the RMEs of = transitions are the first QMC calculations for these systems that include corrections from two-body axial currents. As discussed above, RMEs are suppressed at leading order which gives only and of the experimental values for the 8BBe and 8HeLi transitions. Two-body currents provide about 20–30% correction in the right direction which is, however, still insufficient to reach agreement with the experimental data. These transitions are challenging not only from the theoretical but also from the experimental point of view. For example, in Tables 5 and 6 we quote two results for the RME of the 8LiBe decay obtained from the log()-values of Ref. Warburton (1986) and Ref. Pritychenko et al. (2011) via the following formula Chou et al. (1993)
(19) |
where is the angular momentum of the parent nucleus. The Fermi transition strength is small enough in this case that it can be neglected in the above formula. We then obtain two values, namely, RME(Ref. Warburton (1986))=0.284, and RME(Ref. Pritychenko et al. (2011))=0.190. We note that Refs. Suzuki et al. (2003); Hardy and Towner (2015), report a different overall factor of instead of in the formula given above. In our estimate we used =. Despite this additional uncertainty in the deduced experimental values, our predictions still severely underestimate the data. For example, the calculated Ia* RME provides only and of the experimental values given in Ref. Warburton (1986) and Ref. Pritychenko et al. (2011), respectively.

V.2 Two–Body Transition Densities
In order to have a better understanding of the two-body terms in the axial current and their contributions, it is helpful to study the associated two-body transition densities, which we define as Schiavilla et al. (1998)
(20) |
where is the interparticle distance, and = N2LO-, N2LO-OPE, and N2LO-CT.
Two-body transition densities, calculated for selected nuclei with variational wave functions corresponding to the Hamiltonian models NV2+3-Ia (Ia) and NV2+3-Ia* (Ia*), are presented in Fig. 7. These models produce similar N2LO- and N3LO-OPE densities, since they are based on the same underlying NV2 interaction and only differ in the NV3 interaction. Specifically, what differs is the strength of the contact current—linked to the (contact) three-nucleon interaction via the relation in Eq. (16). Because of the different methodologies adopted in constraining and (the LECs parameterizing the contact piece of the three-nucleon interaction), turns out to be much larger for Ia* than for Ia, see Table 3. Since the contact current is proportional to , this also explains why the corresponding density for Ia* is much larger (in magnitude) than for Ia. Note that they are both negative. As a consequence, the total density (black symbols in Fig. 7) develops a node at around 1 fm in the case of model Ia*.
Another interesting feature of Fig. 7 is the difference between the N2LO- and N3LO-OPE densities at separations fm for the transitions in the larger systems, especially those involving the = resonant states. In the limit of vanishing momentum transfer we are considering here, the corresponding currents have the same operator structure Baroni et al. (2018), up to a momentum dependent term, absent in the N2LO- current, which, however, we have explicitly verified to give a numerically small contribution by direct calculation. Examination of Eqs. (2.9) and (2.10) of Ref. Baroni et al. (2018) shows that this common operator structure involves two independent correlations functions and with =, proportional to different combinations of LECs, denoted by () and () in the OPE () current, with
(21) |
using the values in Tables 1 and 2 ( is the average nucleon mass). Here, = and =. Indeed, the N2LO- current reads
(22) |
with
(23) |
A similar expression holds for with replacing . There are cancellations between the terms proportional to and in each of these currents, and these cancellations are sensitive to the values of the ratios and , and to the overlap between the wave functions of the states involved in the transition.

To gain insight into how short-range physics impacts these weak transitions across different nuclei, we display in Fig. 8 the densities corresponding to the individual two-body contributions, each normalized as , where is the maximum attained value (in magnitude); so all curves peak at 1. We also display the total densities and note that, since both a positive peak and a negative valley are present in this case, each curve is normalized so that the value of the positive peak is 1.
The universal behavior exhibited by the N2LO-, N3LO-OPE, N3LO-CT densities is quite striking, as the curves corresponding to different nuclei and different Hamiltonian models, essentially overlap for . (It is even more striking when the weighing factor is not included). Such behavior can be understood as follows.


In a charge-raising process the two-body weak transition operators primarily convert a pair with total spin/isospin = 1/0 ( pair with = 0/1) to a pair with = 0/1 ( pair with = 1/0) Schiavilla et al. (1998)—of course, similar considerations apply to a charge lowering process. These operators, at least in light systems, do not couple = and to = in a significant way, since P-wave components are small in that case. At separations , where these transitions operators are most effective, the pair wave functions with = 1/0 and 0/1 in different nuclei are similar in shape and only differ by a scale factor Forest et al. (1996)—those corresponding to the Ia* model are illustrated in Fig. 9. This is the origin for the universal behavior observed in Fig. 8. Note that in Fig. 9 we also show the densities of pairs in =0/0 and , which do not scale. A complete analysis and interpretation of these results—in particular, of the role played by the tensor component of the nuclear interaction in shaping these densities—can be found in Ref. Forest et al. (1996).
However, at separations , the N2LO- and N3LO-OPE densities (especially the former) do not scale, particularly in the case of the heavier systems with , presumably due to delicate cancellations between the terms proportional to the operators and present in these currents (P-wave components in the wave functions of these systems may also play a significant role). As a matter of fact, because of the different long-range behavior present in the transitions (note, e.g., how the N2LO- density in the 10C transition assumes negative values at large separations, all the other densities being positive), and because of the rather large size of the contact contribution in the Ia* model, the total densities exhibit nodes when calculated with the starred interactions, which leads to non-trivial cancellations.
VI Conclusions
In this work, we reported on a detailed study of weak matrix elements in =–10 systems based on chiral (two- and three-nucleon) interactions and associated (one- and two-body) axial currents at high orders in the chiral expansion. A summary of our results is displayed in Fig. 6. Agreement with the experimental data is obtained when correlated nuclear wave functions are adopted. For these transitions the contribution of corrections beyond LO in the axial current is typically at the level of the value calculated with the LO Gamow-Teller operator. These findings are in line with those reported in the hybrid study of Ref. Pastore et al. (2018a) for the same transitions. Here, we also present calculations of matrix elements entering the rates of the 8Li, 8B, and 8He beta decays. These matrix elements are found to be suppressed at LO, and N2LO and N3LO currents provide a large correction (–30%) which is, however, insufficient to explain the experimental data. We attribute the large suppression at LO to the fact that the Gamow-Teller operator is, in these transitions, connecting large to small components of the initial and final wave functions. Improving on these calculations will require the development of more sophisticated wave functions with better constrained small components.
Finally, we also reported on a careful analysis of one- and two-body transition densities shown in Figs. 4–8. The latter are especially interesting because they allow us to understand the spatial distributions of the various two-body current operators, that is their behavior as functions of interparticle distance. We have shown that, for each set of interactions and consistent currents (either NV2+3-Ia or NV2+3-Ia*), the two-body transition densities exhibit a universal behavior at short distance across all nuclei we have considered in the present study.
Acknowledgments
We would like to thank V. Cirigliano, R. Charity, W. Dekens, J. Engel, A. Hayes, E. Mereghetti, and L. Sobotka for useful discussions at various stages of this work. S. P. and G. K. would like to thank Grigor Sargsyan for useful discussions and for sharing information on the experimental log value of 8Li. The many-body calculations were performed on the parallel computers of the Laboratory Computing Resource Center, Argonne National Laboratory, the computers of the Argonne Leadership Computing Facility (ALCF) via the 2019/2020 ALCC grant “Low energy neutrino-nucleus interactions” for the project NNInteractions. The work of J.C., S.G., and R.B.W. has been supported by the NUclear Computational Low-Energy Initiative (NUCLEI) SciDAC project. This research is supported by the U.S. Department of Energy, Office of Science, Office of Nuclear Physics, under contracts DE-AC05-06OR23177 (R.S.) and DE-AC02-06CH11357 (R.B.W.), DE-AC52-06NA25396 (S.G. and J.C.) and U.S. Department of Energy funds through the FRIB Theory Alliance award DE-SC0013617 (M.P. and S.P.).
References
- Piarulli et al. (2015) M. Piarulli, L. Girlanda, R. Schiavilla, R. Navarro Pérez, J. E. Amaro, and E. Ruiz Arriola, Phys. Rev. C91, 024003 (2015), eprint 1412.6446.
- Piarulli et al. (2016) M. Piarulli, L. Girlanda, R. Schiavilla, A. Kievsky, A. Lovato, L. E. Marcucci, S. C. Pieper, M. Viviani, and R. B. Wiringa, Phys. Rev. C94, 054007 (2016), eprint 1606.06335.
- Baroni et al. (2016a) A. Baroni, L. Girlanda, A. Kievsky, L. E. Marcucci, R. Schiavilla, and M. Viviani, Phys. Rev. C94, 024003 (2016a), [Erratum: Phys. Rev.C95,no.5,059902(2017)], eprint 1605.01620.
- Baroni et al. (2018) A. Baroni et al., Phys. Rev. C98, 044003 (2018), eprint 1806.10245.
- Baroni et al. (2016b) A. Baroni, L. Girlanda, S. Pastore, R. Schiavilla, and M. Viviani, Phys. Rev. C93, 015501 (2016b), [Erratum: Phys. Rev.C95,no.5,059901(2017)], eprint 1509.07039.
- Pastore et al. (2018a) S. Pastore, A. Baroni, J. Carlson, S. Gandolfi, S. C. Pieper, R. Schiavilla, and R. B. Wiringa, Phys. Rev. C97, 022501 (2018a), eprint 1709.03592.
- Wiringa et al. (1995) R. B. Wiringa, V. G. J. Stoks, and R. Schiavilla, Phys. Rev. C51, 38 (1995), eprint nucl-th/9408016.
- Pieper (2008) S. C. Pieper, AIP Conference Proceedings 1011, 143 (2008), eprint https://aip.scitation.org/doi/pdf/10.1063/1.2932280, URL https://aip.scitation.org/doi/abs/10.1063/1.2932280.
- Gysbers et al. (2019) P. Gysbers et al., Nature Phys. 15, 428 (2019), eprint 1903.00047.
- Gazit et al. (2009a) D. Gazit, S. Vaintraub, and N. Barnea, in Particles and nuclei. Proceedings, 18th International Conference, PANIC08, Eilat, Israel, November 9-14, 2008 (2009a), pp. 1005–1007, eprint 0901.2670.
- Gardestig and Phillips (2006) A. Gardestig and D. R. Phillips, Phys. Rev. Lett. 96, 232301 (2006), eprint nucl-th/0603045.
- Gazit et al. (2009b) D. Gazit, S. Quaglioni, and P. Navratil, Phys. Rev. Lett. 103, 102502 (2009b), [Erratum: Phys. Rev. Lett. 122, 029901 (2019)].
- Schiavilla (2017) R. Schiavilla, unpublished (2017).
- Gonzalez-Alonso et al. (2019) M. Gonzalez-Alonso, O. Naviliat-Cuncic, and N. Severijns, Prog. Part. Nucl. Phys. 104, 165 (2019), eprint 1803.08732.
- Engel and Menendez (2017) J. Engel and J. Menendez, Rept. Prog. Phys. 80, 046301 (2017), eprint 1610.06548.
- Carlson et al. (2015) J. Carlson, S. Gandolfi, F. Pederiva, S. C. Pieper, R. Schiavilla, K. E. Schmidt, and R. B. Wiringa, Rev. Mod. Phys. 87, 1067 (2015), eprint 1412.3081.
- Lynn et al. (2019) J. E. Lynn, I. Tews, S. Gandolfi, and A. Lovato, Submitted to: Ann. Rev. Nucl. Part. Sci. (2019), eprint 1901.04868.
- Wiringa (1991) R. B. Wiringa, Phys. Rev. C43, 1585 (1991).
- Pudliner et al. (1997) B. S. Pudliner, V. R. Pandharipande, J. Carlson, S. C. Pieper, and R. B. Wiringa, Phys. Rev. C56, 1720 (1997), eprint nucl-th/9705009.
- Nollett et al. (2001) K. M. Nollett, R. B. Wiringa, and R. Schiavilla, Phys. Rev. C63, 024003 (2001), eprint nucl-th/0006064.
- Nollett (2001) K. M. Nollett, Phys. Rev. C63, 054002 (2001), eprint nucl-th/0102022.
- Metropolis et al. (1953) N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller, and E. Teller, J. Chem. Phys. 21, 1087 (1953).
- Pervin et al. (2007) M. Pervin, S. C. Pieper, and R. B. Wiringa, Phys. Rev. C76, 064319 (2007), eprint 0710.1265.
- Krebs et al. (2007) H. Krebs, E. Epelbaum, and U.-G. Meissner, Eur. Phys. J. A32, 127 (2007), eprint nucl-th/0703087.
- Navarro Pérez et al. (2013) R. Navarro Pérez, J. E. Amaro, and E. Ruiz Arriola, Phys. Rev. C88, 064002 (2013), [Erratum: Phys. Rev.C91,no.2,029901(2015)], eprint 1310.2536.
- Navarro Pérez et al. (2014) R. Navarro Pérez, J. E. Amaro, and E. Ruiz Arriola, Phys. Rev. C 89, 024004 (2014), eprint 1310.6972.
- Navarro Perez et al. (2014) R. Navarro Perez, J. E. Amaro, and E. Ruiz Arriola, Phys. Rev. C 89, 064006 (2014), eprint 1404.0314.
- Green (1976) A. M. Green, Rept. Prog. Phys. 39, 1109 (1976).
- Arndt et al. (1994) R. A. Arndt, R. L. Workman, and M. M. Pavan, Phys. Rev. C49, 2729 (1994).
- Stoks et al. (1993) V. G. J. Stoks, R. Timmermans, and J. J. de Swart, Phys. Rev. C47, 512 (1993), eprint nucl-th/9211007.
- Tanabashi et al. (2018) M. Tanabashi, K. Hagiwara, K. Hikasa, K. Nakamura, Y. Sumino, F. Takahashi, J. Tanaka, K. Agashe, G. Aielli, C. Amsler, et al. (Particle Data Group), Phys. Rev. D 98, 030001 (2018), URL https://link.aps.org/doi/10.1103/PhysRevD.98.030001.
- van Kolck (1994) U. van Kolck, Phys. Rev. C49, 2932 (1994).
- Epelbaum et al. (2002) E. Epelbaum, A. Nogga, W. Gloeckle, H. Kamada, U. G. Meissner, and H. Witala, Phys. Rev. C66, 064001 (2002), eprint nucl-th/0208023.
- Lynn et al. (2016) J. E. Lynn, I. Tews, J. Carlson, S. Gandolfi, A. Gezerlis, K. E. Schmidt, and A. Schwenk, Phys. Rev. Lett. 116, 062501 (2016), eprint 1509.03470.
- Tews et al. (2016) I. Tews, S. Gandolfi, A. Gezerlis, and A. Schwenk, Phys. Rev. C93, 024305 (2016), eprint 1507.05561.
- Lynn et al. (2017) J. E. Lynn, I. Tews, J. Carlson, S. Gandolfi, A. Gezerlis, K. E. Schmidt, and A. Schwenk, Phys. Rev. C96, 054007 (2017), eprint 1706.07668.
- Piarulli et al. (2018) M. Piarulli et al., Phys. Rev. Lett. 120, 052503 (2018), eprint 1707.02883.
- Marcucci et al. (2012) L. E. Marcucci, A. Kievsky, S. Rosati, R. Schiavilla, and M. Viviani, Phys. Rev. Lett. 108, 052502 (2012), [Erratum: Phys. Rev. Lett.121,no.4,049901(2018)], eprint 1109.5563.
- Cirigliano et al. (2019) V. Cirigliano, W. Dekens, J. De Vries, M. L. Graesser, E. Mereghetti, S. Pastore, M. Piarulli, U. Van Kolck, and R. B. Wiringa (2019), eprint 1907.11254.
- Piarulli and Tews (2020) M. Piarulli and I. Tews, Frontiers in Physics 7, 245 (2020), ISSN 2296-424X, URL https://www.frontiersin.org/article/10.3389/fphy.2019.00245.
- Gandolfi et al. (2020) S. Gandolfi, D. Lonardoni, A. Lovato, and M. Piarulli (2020), eprint 2001.01374.
- Schmidt and Fantoni (1999) K. E. Schmidt and S. Fantoni, Phys. Lett. B446, 99 (1999).
- Day (1967) B. D. Day, Rev. Mod. Phys. 39, 719 (1967).
- Baldo and Burgio (2012) M. Baldo and G. F. Burgio, Rep. Progr. Phys. 75, 026301 (2012).
- Fantoni and Rosati (1975) S. Fantoni and S. Rosati, Nuovo Cimento A 25, 593 (1975).
- Pandharipande and Wiringa (1979) V. R. Pandharipande and R. B. Wiringa, Rev. Mod. Phys. 51, 821 (1979), URL https://link.aps.org/doi/10.1103/RevModPhys.51.821.
- Piarulli et al. (2019) M. Piarulli, I. Bombaci, D. Logoteta, A. Lovato, and R. B. Wiringa (2019), eprint 1908.04426.
- Bombaci and Logoteta (2018) I. Bombaci and D. Logoteta, Astron. Astrophys. 609, A128 (2018), eprint 1805.11846.
- Park et al. (1993) T.-S. Park, D.-P. Min, and M. Rho, Phys. Rept. 233, 341 (1993), eprint hep-ph/9301295.
- Krebs et al. (2017) H. Krebs, E. Epelbaum, and U. G. Meißner, Annals Phys. 378, 317 (2017), eprint 1610.03569.
- Pastore et al. (2008) S. Pastore, R. Schiavilla, and J. L. Goity, Phys. Rev. C78, 064002 (2008), eprint 0810.1941.
- Pastore et al. (2009) S. Pastore, L. Girlanda, R. Schiavilla, M. Viviani, and R. B. Wiringa, Phys. Rev. C80, 034004 (2009), eprint 0906.1800.
- Pastore et al. (2011) S. Pastore, L. Girlanda, R. Schiavilla, and M. Viviani, Phys. Rev. C84, 024001 (2011), eprint 1106.4539.
- Kolling et al. (2009) S. Kolling, E. Epelbaum, H. Krebs, and U. G. Meissner, Phys. Rev. C80, 045502 (2009), eprint 0907.3437.
- Kolling et al. (2011) S. Kolling, E. Epelbaum, H. Krebs, and U. G. Meissner, Phys. Rev. C84, 054008 (2011), eprint 1107.0602.
- Krebs et al. (2020) H. Krebs, E. Epelbaum, and U.-G. Meißner (2020), eprint 2001.03904.
- Bacca and Pastore (2014) S. Bacca and S. Pastore, J. Phys. G41, 123002 (2014), eprint 1407.3490.
- Piarulli et al. (2013) M. Piarulli, L. Girlanda, L. E. Marcucci, S. Pastore, R. Schiavilla, and M. Viviani, Phys. Rev. C87, 014006 (2013), eprint 1212.1105.
- Schiavilla et al. (2019) R. Schiavilla, A. Baroni, S. Pastore, M. Piarulli, L. Girlanda, A. Kievsky, A. Lovato, L. E. Marcucci, S. C. Pieper, M. Viviani, et al., Phys. Rev. C 99, 034005 (2019), URL https://link.aps.org/doi/10.1103/PhysRevC.99.034005.
- Nevo Dinur et al. (2019) N. Nevo Dinur, O. J. Hernandez, S. Bacca, N. Barnea, C. Ji, S. Pastore, M. Piarulli, and R. B. Wiringa, Phys. Rev. C99, 034004 (2019), eprint 1812.10261.
- Marcucci et al. (2016) L. E. Marcucci, F. Gross, M. T. Pena, M. Piarulli, R. Schiavilla, I. Sick, A. Stadler, J. W. Van Orden, and M. Viviani, J. Phys. G43, 023002 (2016), eprint 1504.05063.
- Girlanda et al. (2010) L. Girlanda, A. Kievsky, L. E. Marcucci, S. Pastore, R. Schiavilla, and M. Viviani, Phys. Rev. Lett. 105, 232502 (2010), eprint 1008.0356.
- Pastore et al. (2013) S. Pastore, S. C. Pieper, R. Schiavilla, and R. B. Wiringa, Phys. Rev. C87, 035503 (2013), eprint 1212.3375.
- Datar et al. (2013) V. M. Datar et al., Phys. Rev. Lett. 111, 062502 (2013), eprint 1305.1094.
- Pastore et al. (2014) S. Pastore, R. B. Wiringa, S. C. Pieper, and R. Schiavilla, Phys. Rev. C90, 024321 (2014), eprint 1406.2343.
- Pastore et al. (2019) S. Pastore, J. Carlson, S. Gandolfi, R. Schiavilla, and R. B. Wiringa (2019), eprint 1909.06400.
- Pastore et al. (2018b) S. Pastore, J. Carlson, V. Cirigliano, W. Dekens, E. Mereghetti, and R. B. Wiringa, Phys. Rev. C97, 014606 (2018b), eprint 1710.05026.
- Cirigliano et al. (2018) V. Cirigliano, W. Dekens, J. De Vries, M. L. Graesser, E. Mereghetti, S. Pastore, and U. Van Kolck, Phys. Rev. Lett. 120, 202001 (2018), eprint 1802.10097.
- Wang et al. (2019) X. B. Wang, A. C. Hayes, J. Carlson, G. X. Dong, E. Mereghetti, S. Pastore, and R. B. Wiringa, Phys. Lett. B798, 134974 (2019), eprint 1906.06662.
- Tilley et al. (2002) D. R. Tilley, C. M. Cheves, J. L. Godwin, G. M. Hale, H. M. Hofmann, J. H. Kelley, C. G. Sheu, and H. R. Weller, Nucl. Phys. A708, 3 (2002).
- Tilley et al. (2004) D. R. Tilley, J. H. Kelley, J. L. Godwin, D. J. Millener, J. E. Purcell, C. G. Sheu, and H. R. Weller, Nucl. Phys. A745, 155 (2004).
- Knecht et al. (2012) A. Knecht et al., Phys. Rev. C86, 035506 (2012).
- Suzuki et al. (2003) T. Suzuki, R. Fujimoto, and T. Otsuka, Phys. Rev. C 67, 044302 (2003), URL https://link.aps.org/doi/10.1103/PhysRevC.67.044302.
- Chou et al. (1993) W. T. Chou, E. K. Warburton, and B. A. Brown, Phys. Rev. C47, 163 (1993).
- Warburton (1986) E. K. Warburton, Phys. Rev. C 33, 303 (1986), URL https://link.aps.org/doi/10.1103/PhysRevC.33.303.
- Pritychenko et al. (2011) B. Pritychenko, E. Běták, M. Kellett, B. Singh, and J. Totans, Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 640, 213–218 (2011), ISSN 0168-9002, URL http://www.sciencedirect.com/science/article/pii/S0168900211005584.
- Wiringa (2006) R. B. Wiringa, Phys. Rev. C 73, 034317 (2006), URL https://link.aps.org/doi/10.1103/PhysRevC.73.034317.
- Wiringa et al. (2000a) R. B. Wiringa, S. C. Pieper, J. Carlson, and V. R. Pandharipande, Phys. Rev. C 62, 014001 (2000a), URL https://link.aps.org/doi/10.1103/PhysRevC.62.014001.
- Wiringa et al. (2000b) R. B. Wiringa, S. C. Pieper, J. Carlson, and V. R. Pandharipande, Phys. Rev. C62, 014001 (2000b), eprint nucl-th/0002022.
- Hardy and Towner (2015) J. C. Hardy and I. S. Towner, Phys. Rev. C91, 025501 (2015), eprint 1411.5987.
- Schiavilla et al. (1998) R. Schiavilla et al., Phys. Rev. C58, 1263 (1998), eprint nucl-th/9808010.
- Forest et al. (1996) J. L. Forest, V. R. Pandharipande, S. C. Pieper, R. B. Wiringa, R. Schiavilla, and A. Arriaga, Phys. Rev. C54, 646 (1996), eprint nucl-th/9603035.