Efficacy of the symmetry-adapted basis for ab initio nucleon-nucleus interactions
for light- and intermediate-mass nuclei
Abstract
We study the efficacy of a new ab initio framework that combines the symmetry-adapted (SA) no-core shell-model approach with the resonating group method (RGM) for unified descriptions of nuclear structure and reactions. We obtain ab initio neutron-nucleus interactions for 4He, 16O, and 20Ne targets, starting with realistic nucleon-nucleon potentials. We discuss the effect of increasing model space sizes and symmetry-based selections on the SA-RGM norm and direct potential kernels, as well as on phase shifts, which are the input to calculations of cross sections. We demonstrate the efficacy of the SA basis and its scalability with particle numbers and model space dimensions, with a view toward ab initio descriptions of nucleon scattering and capture reactions up through the medium-mass region.
pacs:
Valid PACS appear hereI Introduction
Ab initio descriptions of spherical and deformed nuclei up through the calcium region are now possible within a no-core shell-model framework, by utilizing emerging symplectic symmetry in nuclei. In particular, the symmetry-adapted no-core shell-model (SA-NCSM) Launey et al. (2016); Dytrych et al. (2020) uses a physically relevant symmetry-adapted (SA) basis that can achieve significantly reduced model spaces compared to the corresponding complete ultra-large model spaces, without compromising the accuracy of results for various observables Dytrych et al. (2015); Launey et al. (2016); Baker et al. (2020a). This enables the SA-NCSM to accommodate contributions from more shells and to describe heavier nuclei, such as 20Ne Dytrych et al. (2020), 21Mg Ruotsalainen et al. (2019), 22Mg Henderson et al. (2018), 28Mg Williams et al. (2019), as well as 32Ne and 48Ti Launey et al. (2018, 2021). The access to higher-lying shells makes the SA basis suitable for describing nuclear reactions Launey et al. (2021), the processes that are typically studied in experiments and govern stellar evolution. Remarkable progress has been made in first-principle descriptions to scattering and nuclear reactions for light nuclei (for an overview, see Johnson et al. (2020); Quaglioni and Navrátil (2020)), including studies of elastic scattering Nollett et al. (2007); Hagen et al. (2007); Quaglioni and Navrátil (2008a); Elhatisari et al. (2015); Zhang et al. (2020); Mercenne et al. (2019); Burrows et al. (2020), photoabsorption Bacca et al. (2014), transfer Navrátil and Quaglioni (2012) and capture reactions Girlanda et al. (2010), widths Kravvaris and Volya (2017); Dreyfuss et al. (2020) and resonant states Lazauskas et al. (2019), as well as thermonuclear fusion Hupin et al. (2019). In this paper, we show that expanding the reach of ab initio reactions to deformed and heavier targets is now feasible with the SA basis.
Microscopic approaches to nuclear reactions take into account nucleon degrees of freedom along with their correlations within and among the reaction fragments. Coupled with realistic inter-nucleon interactions, such as the ones derived in the framework of chiral effective field theory Bedaque and van Kolck (2002); Epelbaum et al. (2002); Entem and Machleidt (2003); Epelbaum (2006), these approaches provide ab initio calculations of reaction observables. One of the earliest and very successful microscopic approaches to nuclear reactions is the resonating-group method (RGM) Wildermuth and Tang (1977); Tang et al. (1978). In the RGM, nucleons are organized within different groups, or clusters, “resonating” through the inter-cluster exchange of nucleons. Most importantly, the cluster system is translationally invariant, and the Pauli exclusion principle is enforced by the antisymmetrization between the different clusters. All of these features make this method particularly suitable for unified descriptions of nuclear structure and reaction observables. Following the success of the Elliott model Elliott (1958a, b), showing that a leading (most deformed) shell-model configuration describes reasonably well the ground-state rotational band in intermediate-mass nuclei, the RGM has been extensively used with an basis and its no-core shell-model extension, the symplectic basis Hecht (1977); Hecht and Zahn (1978); Hecht and Suzuki (1982). Applications of the model with Gaussian interactions have successfully calculated and 8Be cluster amplitudes, spectroscopic amplitudes for heavy-fragment clusters, and sub-Coulomb 12C+12C resonances Hecht et al. (1981); Hecht and Braunschweig (1978); Suzuki and Hecht (1982). The formalism has been extended by utilizing a mixed no-core shell-model plus RGM cluster basis Suzuki (1986); Suzuki and Hecht (1986, 1987), and applied to studies of the monopole and quadrupole strengths in light nuclei Suzuki (1987); Suzuki and Hara (1989), as well as the +12C cluster system Suzuki (1976a, b); Suzuki et al. (2003).
More recently, a successful first-principle description of scattering and reactions has been realized by implementing the RGM using ab initio NCSM Navrátil et al. (2009); Barrett et al. (2013) wave functions for the clusters in a formalism known as NCSM/RGM Quaglioni and Navrátil (2008b, 2009); Hupin et al. (2013) and, later, by fully combining the two approaches into the generalized ab initio cluster expansion of the no-core shell model with continuum (NCSMC) Baroni et al. (2013a, b). These methods, which have enabled predictions of nucleon Hupin et al. (2014); Calci et al. (2016), deuteron Hupin et al. (2015) and alpha Kravvaris et al. (2020) scattering off light targets, as well as polarized deuterium-tritium fusion Hupin et al. (2019) from chiral nucleon-nucleon (NN) and three—nucleon (3N) forces, are reviewed in Refs. Navrátil et al. (2016); Quaglioni and Navrátil (2020).
In addition, the Gamow shell model coupled-channel approach combines the RGM with a continuum core-valence shell approach and allows for descriptions of nuclear reactions of heavier systems Jaganathen et al. (2014); Fossez et al. (2015); Mercenne et al. (2019).
The goal of this paper is to show the efficacy of a new approach that can extend the study of ab initio reactions to medium mass nuclei by using the SA-NCSM approach Launey et al. (2016); Dytrych et al. (2020). The SA framework takes advantage of symmetries inherent to nuclei and of group theoretical algorithms, and reorganizes the model space into a physically relevant basis. This allows us to account for the relevant correlations within only a few dominant components and, hence, achieve manageable Hamiltonian matrix sizes. In this paper, we present a new formalism of the RGM, one that admits the use of the SA basis, and we demonstrate the capability and potential of the approach for light and intermediate-mass nuclei. The formalism of the SA-RGM framework is presented in Sec. II, where we discuss RGM kernels computed using the SA basis. The sensitivity of the kernels on different selected model spaces and model space sizes is discussed in Sec. III.1 for a 4He target and in Sec. III.2 for intermediate-mass 16O and 20Ne targets. Section III.3 presents an analysis of the basis dimension and its scaling with model space sizes and particle numbers. Finally, Sec. IV outlines the conclusions.
II Theoretical framework
Traditionally, the RGM adopts microscopic cluster wave functions as basis functions to describe the motion of a system of two or more clusters (see, e.g., Refs. Quaglioni and Navrátil (2009); Quaglioni et al. (2013)). We consider two nuclear fragments, or binary-cluster nuclear reactions. For two clusters and , the cluster states for a channel are defined as:
(1) | |||||
where the cluster system is defined for a channel , which is labeled by the angular momentum (spin) and parity of each of the clusters and the total spin of the clusters (the labels and denote all other quantum numbers needed to fully characterize their respective states), and the orbital momentum . For particle laboratory coordinates (used in this study), the separation distance between the center-of-mass of the two clusters is determined from . The distance between the clusters defines the cluster states and the RGM kernels, as shown below, and as an integration variable facilitates the treatment of the inter-cluster antisymmetrization. Namely, the nuclear wave function is expressed in terms of the cluster states as
(2) |
with unknown amplitudes that are determined by solving the integral Hill-Wheeler equations for a given total energy in the center-of-mass frame:
(3) |
Here, is the Hamiltonian kernel and is the norm kernel, where is the inter-cluster antisymmetrizer. The kernels are computed by using the microscopic wave functions of the clusters that can be obtained in the ab initio NCSM and SA-NCSM. Once the kernels are computed, Eq. (3) can then be solved using the microscopic -matrix approach Descouvemont and Baye (2010); Descouvemont (2016).
In the SA-RGM, the target nucleus of particles is described by SA-NCSM many-body wave functions. In the SA-NCSM, the many-body basis is labeled by irreducible representations (irreps) according to the group chain Elliott (1958a, b):
(4) |
The quantum numbers label an irrep and can be related to the average deformation through the established link with the well-known parameters, deformation and triaxiality Castaños et al. (1988); Mustonen et al. (2018). The label distinguishes multiple occurrences of the same orbital momentum in the parent irrep , and is the projection. These quantum numbers define the spatial degrees of freedom, which can then be coupled to the intrinsic spin () to yield a good total angular momentum.
Specifically, a target state with total angular momentum and parity (and projection ) is constructed in terms of the SA basis:
(5) |
where the labels are defined as and deformation (it is understood that the coefficients are for given , which is omitted from labeling). Protons and neutrons are labeled by p and n, respectively, and labels the intrinsic spin (“” denotes all additional quantum numbers including ). The outer multiplicity Draayer et al. (1989) results from the coupling of the proton deformation with that of neutrons to total deformation . labels the total HO excitations above the valence-shell configuration and is truncated at a maximum value (), which determines the model space size.
For a single-particle projectile, the SA-RGM channel basis states can thus be defined for a channel [related to channel in Eq. (19)] as:
(6) |
where the basis states for the target are coupled to the HO single-particle states of the projectile with SU(3) quantum numbers and spin (we will omit the parity from the notation throughout the paper for simplicity). We note that the SU(3) outer multiplicity associated with the coupling of and is 1, and hence, omitted from the notations. Remarkably, there is no dependence on the orbital momentum of the projectile, only on the shell number it occupies, . Furthermore, the summation over implies that the SA-RGM basis requires only a part of the information present in the SA basis.
The SA-RGM basis is used to calculate the RGM kernels, which is the main computational task in RGM Quaglioni and Navrátil (2009). These include the norm kernel, which is the overlap between antisymmetrized non-orthogonal RGM basis states. It consists of a direct part (a Dirac delta function), which dominates at large relative distances, and an exchange part that takes into account the Pauli principle at short distances. The exchange norm kernel is related to the permutation operator that exchanges the nucleon projectile with another nucleon within the target, thereby ensuring antisymmetrization (cf. Quaglioni and Navrátil (2009)):
(7) |
The exchange norm kernel in the SA-RGM basis is thus reduced to evaluating the following (similarly, for the Hamiltonian kernels):
(10) | |||||
(11) |
where is the 6- recoupling coefficient Draayer and Akiyama (1973), analogous to the SU(2) 6- symbol, , and the one-body density matrix elements are defined as:
(12) |
where and creates and annihilates, respectively, a particle of spin 1/2 in the -th HO shell, and is the annihilation SU(3) tensor operator. The matrix elements of the density can be quickly computed on the fly in the SA basis. Their computation can utilize an efficacious algorithm that exploits organization of SA basis states in terms of subspaces of SU(3) irreps and the factorization of spatial SU(3) and SU(2) spin degrees of freedom Dytrych et al. (2007); Launey et al. (2016), and this can be done prior to the computation of the kernels. It is notable that, as a result of the Kronecker delta function in Eq. (11), the exchange part of the norm kernel turns out to be block-diagonal in the SA-RGM basis. The reason is that the operator is an SU(3) scalar and spin scalar, and therefore preserves deformation and spin of the composite system (note that it may change the deformation of the target itself).
Eq. (11) allows the kernels to be calculated, for each , through the SA-RGM channel basis of Eq. (6) that only depends on the deformation, rotation, and spin of the target (that is, ), and the deformation, rotation, and spin of the target-projectile system (that is, ). From this, it is clear that the SA offers two main advantages: first, calculations utilize group-theoretical algorithms that use a reduced subset of quantum numbers and , and second, the number of configurations in the target wave function, we find, is a manageable number when compared to the complete model-space size. This results in a manageable number of configurations for the target-projectile system based on and SU(2) selection rules, namely, and (for further details on scalability, see Sec. III.3).
Another advantage of the SA scheme is that the dependence on the orbital momentum is recovered in the very last step:
(15) | |||
(19) |
This wave function is then used in a microscopic -matrix approach Descouvemont and Baye (2010) to calculate phase shifts and cross sections.
To study the efficacy of the SA scheme, we focus on the norm and potential kernels. For the potential kernel, we consider only the part that involves the projectile and a single nucleon in the target (similarly to Ref. Burrows et al. (2020)), that is, the potential kernel of particle-rank one, denoted here as (cf. Quaglioni and Navrátil (2009)):
(20) |
Note that the exchange of two nucleons that interact with each other is part of this kernel. We do not consider the particle-rank two potential kernel that accounts for the projectile exchanging with one nucleon in the target and interacting with another nucleon (called exchange potential kernel in Quaglioni and Navrátil (2009)). Since the goal of this study is to validate the use of the SA scheme against the use of complete model spaces, we expect that the particle-rank two potential kernel will benefit from advantages similar to those shown in the next section. The reason is that the main advantage stems from the reductions of the number of basis states needed to describe the target wavefunctions. Such a reduction ensures that one-body densities, along with the two-body densities that will be needed for the particle-rank two potential kernels, are computed for wavefunctions that span only a fraction of the complete model space (as discussed in Sec. III.3).
The derivation of the potential kernel in the SA-RGM basis follows a procedure similar to that for the norm kernel:
(27) | |||
(36) | |||
(39) | |||
(40) |
where , , and denote the SU(3) rank of the operator that transforms the initial state to the final state of the projectile, target, and the system, respectively.

III Results and discussions
It is important to validate the use of the SA basis in the SA-RGM, by comparing selected model spaces with the corresponding complete , to ensure that the selection does not remove configurations relevant for the reaction processes under consideration. For this, we study single-projectile scattering off the spherical 4He and 16O nuclei, as well as for the deformed 20Ne nucleus. We present kernels that use target ground state (g.s.) wavefunctions computed with the SA basis in a complete model space (equivalent to NCSM/RGM calculations Quaglioni and Navrátil (2009)) and we compare these to the results that use wavefunctions calculated in a selected SA model space. In general, SA selections are denoted as . For example, the model space includes the complete set of excitations up to and selected excitations in the - subspaces, following a prescription detailed in Ref. Launey et al. (2020). This allows the mixing of all possible shapes within the complete subspaces, whereas the higher selected subspaces accommodate spatially expanded collective modes Launey et al. (2021).




III.1 Validation of the SA scheme
We study the SA efficacy for the potential kernel of Eq. (20) for 4He() + n (Fig. 1), for which calculations in the complete space (no SA selection) are available up to with other interactions ( denotes the model space for the target) Quaglioni and Navrátil (2009). For two NN interactions, Ref. Quaglioni and Navrátil (2009) has shown that the results are sufficient to achieve converged phase shifts for the 4He() + n and channels. In the present study, we use the complete model space for the target, and we compare to the model space. The 4He wavefunctions in these model spaces, calculated with the JISP16 NN interactions Shirokov et al. (2007), have been shown to converge for the binding energy and the g.s. root-mean-square (rms) matter radius, as well as to yield various electromagnetic sum rules Baker et al. (2020b) that agree with those calculated in the hyperspherical harmonics approach Bacca et al. (2013).
We explore the potential kernel of Eq. (20) for as a function of the distance between the clusters, which is used to describe the resonant g.s. in 5He (Fig. 1a), as well as for for a description of the scattering states of 5He (Fig. 1b). We find that the SA space yields results that are indistinguishable from those in the complete space. While Fig. 1 shows the comparison only for fm, the results remain indistinguishable for any . In addition, the norm kernels exhibit the same behavior, namely, the outcomes for the SA and complete model spaces coincide. These results demonstrate that the SA wavefunctions account for the relevant correlations necessary to describe the norm and the direct component of the non-local potentials that govern the resonant ground state and low-energy scattering states in 5He. Because the kernels are used as the input for calculating phase shifts, the findings show that the SA model spaces are sufficient to reproduce the corresponding and phase shifts calculated in the complete model spaces (see Fig. 2 for the comparison between selected and complete model spaces for the wave). We emphasize that this comparison focuses on the effect of SA model spaces benchmarked against the corresponding complete model spaces, not on reproducing experimental phase shifts with all RGM kernels.
III.2 Application to intermediate-mass nuclei
To illustrate the capability of the SA-RGM, we present the first ab initio calculations of RGM norm and leading-order potential kernels in the intermediate mass region, namely, for neutron scattering off 20Ne(). The SA 20Ne wave functions are calculated using the NNLOopt NN interaction Ekström et al. (2013) and have been shown to reproduce observables, such as excitation energies and B(E2) strengths Dytrych et al. (2020). The NNLOopt is used without 3N forces, which have been shown to contribute minimally to the 3- and 4-nucleon binding energies Ekström et al. (2013). Furthermore, the NNLOopt NN potential has been found to reproduce various observables, including the 4He electric dipole polarizability Baker et al. (2020a); the challenging analyzing power for elastic proton scattering on 4He, 12C, and 16O Burrows et al. (2019); along with B(E2) transition strengths for 21Mg and 21F Ruotsalainen et al. (2019).
As expected, the exchange norm kernel for 20Ne( manifests itself at short distances and vanishes at long distances (see Fig. 3a & b, for the case of fm). This reflects the short-range nature of the Pauli exclusion principle. We find that the change in the model space size from to has only a small effect on the exchange norm of the partial wave (Fig. 3a) and partial wave (Fig. 3b). The largest deviations are observed at short distances, where the kernels have the largest magnitude. As the model space increases, the kernels start to converge, and the exchange kernel maximum slightly increases in magnitude for , whereas it slightly decreases for . Note that even though the deviation seems larger for , the magnitude of the exchange kernel maximum is smaller by a factor of 3.5 than that of . Hence, these outcome indicates that the selection of dominant SU(3) components at (see also Fig. 3 in Ref. Dytrych et al. (2020)) is sufficient to incorporate the relevant correlations needed to describe the short-range Pauli effect.
The potential kernels of Eq. (20) for 20Ne(, calculated with the NNLOopt NN, are also studied with increasing model space sizes (see Fig. 3c & d, for fm). Similarly to the exchange norm kernel, the increase in the model space from to has a much smaller effect on this potential kernel when compared to the increase from to , for both and partial waves, suggesting converging results. When compared to the potentials for 4He + n of Fig. 3, the 20Ne( case shows a slightly larger deviation around the kernel maximum when varying the model space. This effect might be a result of the open-shell structure of the ground-state wave function of 20Ne compared to that of 4He. In addition, the small changes in these kernels result in only very little deviations in the phase shifts for the low-energy neutron scattering off 20Ne(, with a relative difference of the order of - compared to the largest model space used (Fig. 4).
As another illustrative example, we present the potential kernel of Eq. (20) for 16O( (Fig. 5), which is feasible for no-core shell-model calculations with the importance truncation using other interactions Navrátil et al. (2010). In our study, we use the NNLOsat Ekström et al. (2015), for which the three-nucleon (3N) forces are included in the SA-NCSM as averages Launey et al. (2021). Namely, in these calculations, the 3N forces are included as a mass-dependent monopole interaction Launey et al. (2012), which has an effect on binding energies. or the 16O ground-state energy, the 7-shell 3N contribution is 20.46 MeV, resulting in MeV total energy for and =16 MeV, which agrees with the experimental value of MeV. In this case, we compare calculations within a selected model space to those in the complete model space. The results of the two model spaces are practically indistinguishable, despite the largely reduced SA model space used here and the addition of SU(3) dominant configurations in the 8 subspace. For 16O, this outcome could be understood by the fact that % of the ground state is composed of a spherical shape and low model spaces are able to account for its vibrations.

III.3 Efficacy and scalability of the SA scheme
In this section we explore the scalability of the SA-RGM calculations with increasing model space sizes and particle number. The SA-RGM channel basis (6) is used to compute the kernels of Eqs. (11) and (40). These channels are constructed from the unique quantum numbers of the target state, resulting in a manageable number of SA-RGM basis states that scale polynomially with , as shown in Fig. 6 for several nuclear systems.
For example, for proton- or neutron-nucleus interaction for N+20Ne (), there are only about - SA-RGM basis states for 7 to 13 shells, and only about for 23Mg when more target states are used (with channels for , ), which is still manageable (Fig. 6). Interestingly, the number of unique deformed configurations for heavier targets, such as Ne and Mg, decrease in larger model spaces, as dominant shapes are allowed to develop, thereby reducing shape mixing. As a consequence, in such cases the SA-RGM basis can become smaller when increasing .
Furthermore, there is a large reduction in the number of SU(3) basis states needed for the target wave functions, as one eliminates negligible contributions identified in the target eigenfunctions. Namely, for the illustrative example of the 23Mg target (Fig. 6), we show the number of the SA-RGM channels after retaining basis states that contribute with a probability amplitude [see Eq. (5)] greater than a certain value . We find that the number of the 23Mg+N SA-RGM states continues to scale polynomially with increasing for each and largely decreases for higher reduction cutoffs. We note that the cutoff uses the 23Mg basis states with a probability greater that and results in no reduction. In the SA-RGM calculations, the cutoff for the SA selection is decreased until convergence of results is achieved.
We note that an important step for computing the kernels from the many-body wavefunctions is the calculation of the operator of Eq. (12). Its calculation can be compared to the one-body density matrix elements, namely, they need to be calculated only once for a given set of target wavefunctions, and, as mentioned above, can utilize an efficient algorithm that exploits SU(3) SA subspaces and the factorization of spatial and spin degrees of freedom. As for the kernels, these calculations are also facilitated by the large reduction in the number of SU(3) basis states needed to describe the target wave functions, as compared to the complete model space. These same reductions are observed for two-body densities that will be needed for the particle-rank two potential kernels.

IV Conclusions
In this paper, we have studied the efficacy of the new ab initio SA-RGM approach that combines the SA-NCSM and RGM frameworks. We have discussed nucleon-nucleus interactions and the use of the SA framework for 4He and 16O targets, as well as the intermediate-mass 20Ne and 23Mg targets feasible in the SA-NCSM. We have shown that the SU(3) selection of the model space has almost negligible effect on the SA-RGM norm and particle-rank one potential kernels that provide the input to calculations of phase shifts and cross sections. The results demonstrate that the nonnegligible components that are included in the calculations account for the correlations needed to describe the single nucleon scattering process in this mass region.
In addition, we have studied the scalability of the SA-RGM approach, showing its computational advantages that stem from the largely reduced number of SU(3) basis states needed to describe the target, as well as the manageable number of the SA-RGM basis states for the target+N system that scale polynomially with the increase in the model space size. The demonstrated efficacy of the SA basis and its scalability with particle numbers and model space dimensions opens the way to ab initio calculations up through the medium-mass region of nucleon-nucleus interactions that enter nucleon scattering and nucleon capture reactions.
Acknowledgements.
We acknowledge useful discussions with Petr Navratil, as well as Linda Hlophe for calculating the NCSM/RGM kernel matrix elements for model benchmarks. This work was supported in part by the U.S. National Science Foundation (PHY-1913728), SURA, the Czech Science Foundation (16-16772S), and the U.S. Department of Energy, Office of Science, Office of Nuclear Physics, under Work Proposals No. SCW0498 and No. SC0019521. A portion of this work was performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory under Contract No. DE-AC52-07NA27344 with support from LDRD Project No. 19-ERD-017. It benefited from high performance computational resources provided by LSU (www.hpc.lsu.edu), the National Energy Research Scientific Computing Center (NERSC), a U.S. Department of Energy Office of Science User Facility operated under Contract No. DE-AC02-05CH11231, as well as the Frontera computing project at the Texas Advanced Computing Center, made possible by National Science Foundation award OAC-1818253.References
- Launey et al. (2016) K. D. Launey, T. Dytrych, and J. P. Draayer, Prog. Part. Nucl. Phys. 89, 101 (review) (2016).
- Dytrych et al. (2020) T. Dytrych, K. D. Launey, J. P. Draayer, D. J. Rowe, J. L. Wood, G. Rosensteel, C. Bahri, D. Langr, and R. B. Baker, Phys. Rev. Lett. 124, 042501 (2020).
- Dytrych et al. (2015) T. Dytrych, A. C. Hayes, K. D. Launey, J. P. Draayer, P. Maris, J. P. Vary, D. Langr, and T. Oberhuber, Phys. Rev. C 91, 024326 (2015).
- Baker et al. (2020a) R. B. Baker, K. D. Launey, S. Bacca, N. N. Dinur, and T. Dytrych, Phys. Rev. C 102, 014320 (2020a).
- Ruotsalainen et al. (2019) P. Ruotsalainen, J. Henderson, G. Hackman, G. H. Sargsyan, K. D. Launey, A. Saxena, P. C. Srivastava, S. R. Stroberg, T. Grahn, J. Pakarinen, G. C. Ball, R. Julin, P. T. Greenlees, J. Smallcombe, C. Andreoiu, N. Bernier, M. Bowry, M. Buckner, R. Caballero-Folch, A. Chester, S. Cruz, L. J. Evitts, R. Frederick, A. B. Garnsworthy, M. Holl, A. Kurkjian, D. Kisliuk, K. G. Leach, E. McGee, J. Measures, D. Mücher, J. Park, F. Sarazin, J. K. Smith, D. Southall, K. Starosta, C. E. Svensson, K. Whitmore, M. Williams, and C. Y. Wu, Phys. Rev. C 99, 051301 (2019).
- Henderson et al. (2018) J. Henderson et al., Phys. Lett. B782, 468 (2018), arXiv:1709.03948 [nucl-ex] .
- Williams et al. (2019) J. Williams, G. C. Ball, A. Chester, T. Domingo, A. B. Garnsworthy, G. Hackman, J. Henderson, R. Henderson, R. Krücken, A. Kumar, K. D. Launey, J. Measures, O. Paetkau, J. Park, G. H. Sargsyan, J. Smallcombe, P. C. Srivastava, K. Starosta, C. E. Svensson, K. Whitmore, and M. Williams, Phys. Rev. C 100, 014322 (2019).
- Launey et al. (2018) K. D. Launey, A. Mercenne, G. H. Sargsyan, H. Shows, R. B. Baker, M. E. Miora, T. Dytrych, and J. P. Draayer, in Proceedings of the 4th International Workshop on ’State of the Art in Nuclear Cluster Physics’ (SOTANCP4), May 2018, Galveston, Texas, Vol. 2038 (AIP Conference Proceedings, 2018).
- Launey et al. (2021) K. D. Launey, A. Mercenne, and T. Dytrych, Annu. Rev. Nucl. Part. Sci. 71, 253 (2021).
- Johnson et al. (2020) C. W. Johnson, K. D. Launey, et al., J. Phys. G 47, 23001 (2020), arXiv:1912.00451.
- Quaglioni and Navrátil (2020) S. Quaglioni and P. Navrátil, Nuclear Physics News 30, 12 (2020).
- Nollett et al. (2007) K. Nollett, S. Pieper, R. Wiringa, J. Carlson, and G. Hale, Phys. Rev. Lett. 99, 022502 (2007).
- Hagen et al. (2007) G. Hagen, D. Dean, M. Hjorth-Jensen, and T. Papenbrock, Phys. Lett. B 656, 169 (2007).
- Quaglioni and Navrátil (2008a) S. Quaglioni and P. Navrátil, Phys. Rev. Lett. 101, 092501 (2008a).
- Elhatisari et al. (2015) S. Elhatisari, D. Lee, G. Rupak, E. Epelbaum, et al., Nature 528, 111 (2015).
- Zhang et al. (2020) X. Zhang, S. R. Stroberg, P. Navrátil, C. Gwak, J. A. Melendez, R. J. Furnstahl, and J. D. Holt, Phys. Rev. Lett. 125, 112503 (2020).
- Mercenne et al. (2019) A. Mercenne, N. Michel, and M. Płoszajczak, Phys. Rev. C 99, 044606 (2019).
- Burrows et al. (2020) M. Burrows, R. B. Baker, C. Elster, S. P. Weppner, K. D. Launey, P. Maris, and G. Popa, Phys. Rev. C 102, 034606 (2020).
- Bacca et al. (2014) S. Bacca, N. Barnea, G. Hagen, M. Miorelli, G. Orlandini, and T. Papenbrock, Phys. Rev. C 90, 064619 (2014).
- Navrátil and Quaglioni (2012) P. Navrátil and S. Quaglioni, Phys. Rev. Lett. 108, 042503 (2012).
- Girlanda et al. (2010) L. Girlanda, A. Kievsky, L. E. Marcucci, S. Pastore, R. Schiavilla, and M. Viviani, Phys. Rev. Lett. 105, 232502 (2010).
- Kravvaris and Volya (2017) K. Kravvaris and A. Volya, Phys. Rev. Lett. 119, 062501 (2017).
- Dreyfuss et al. (2020) A. C. Dreyfuss, K. D. Launey, J. E. Escher, G. H. Sargsyan, R. B. Baker, T. Dytrych, and J. P. Draayer, Phys. Rev. C 102, 044608 (2020).
- Lazauskas et al. (2019) R. Lazauskas, E. Hiyama, and J. Carbonnell, Phys. Lett. B 791, 335 (2019).
- Hupin et al. (2019) G. Hupin, S. Quaglioni, and P. Navrátil, Nature Communications 10, 351 (2019).
- Bedaque and van Kolck (2002) P. F. Bedaque and U. van Kolck, Annu. Rev. Nucl. Part. Sci. 52, 339 (2002).
- Epelbaum et al. (2002) E. Epelbaum, A. Nogga, W. Glöckle, H. Kamada, U.-G. Meißner, and H. Witala, Phys. Rev. C 66, 064001 (2002).
- Entem and Machleidt (2003) D. R. Entem and R. Machleidt, Phys. Rev. C 68, 041001 (2003).
- Epelbaum (2006) E. Epelbaum, Prog. Part. Nucl. Phys. 57, 654 (2006).
- Wildermuth and Tang (1977) K. Wildermuth and Y. Tang, A Unified Theory of The Nucleus (Braunschweig, Vieweg, 1977).
- Tang et al. (1978) Y. C. Tang, M. LeMere, and D. R. Thompson, Phys. Rep. 47, 167 (1978).
- Elliott (1958a) J. P. Elliott, Proc. Roy. Soc. A 245, 128 (1958a).
- Elliott (1958b) J. P. Elliott, Proc. Roy. Soc. A 245, 562 (1958b).
- Hecht (1977) K. T. Hecht, Nucl. Phys. A 283, 223 (1977).
- Hecht and Zahn (1978) K. T. Hecht and W. Zahn, Nucl. Phys. A 318, 1 (1978).
- Hecht and Suzuki (1982) K. T. Hecht and Y. Suzuki, J. Math. Phys. 24, 785 (1982).
- Hecht et al. (1981) K. T. Hecht, E. J. Reske, T. H. Seligman, and W. Zahn, Nucl. Phys. A 356, 146 (1981).
- Hecht and Braunschweig (1978) K. T. Hecht and D. Braunschweig, Nucl. Phys. A 295, 34 (1978).
- Suzuki and Hecht (1982) Y. Suzuki and K. T. Hecht, Nucl. Phys. A 388, 102 (1982).
- Suzuki (1986) Y. Suzuki, Nucl. Phys. A 448, 395 (1986).
- Suzuki and Hecht (1986) Y. Suzuki and K. T. Hecht, Nucl. Phys. A 455, 315 (1986).
- Suzuki and Hecht (1987) Y. Suzuki and K. T. Hecht, Prog. Theor. Phys. 77, 190 (1987).
- Suzuki (1987) Y. Suzuki, Nucl. Phys. A 470, 119 (1987).
- Suzuki and Hara (1989) Y. Suzuki and S. Hara, Phys. Rev. C 39, 658 (1989).
- Suzuki (1976a) Y. Suzuki, Prog. Theor. Phys. 55, 1751 (1976a).
- Suzuki (1976b) Y. Suzuki, Prog. Theor. Phys. 56, 111 (1976b).
- Suzuki et al. (2003) Y. Suzuki, R. G. Lovas, K. Yabana, and K. Varga, Structure and reactions of exotic nuclei (Taylor & Francis, London and New York, 2003).
- Navrátil et al. (2009) P. Navrátil, S. Quaglioni, I. Stetcu, and B. R. Barrett, J. Phys. G: Nucl. Part. 36, 083101 (2009).
- Barrett et al. (2013) B. Barrett, P. Navrátil, and J. Vary, Prog. Part. Nucl. Phys. 69, 131 (2013).
- Quaglioni and Navrátil (2008b) S. Quaglioni and P. Navrátil, Phys. Rev. Lett. 101, 092501 (2008b).
- Quaglioni and Navrátil (2009) S. Quaglioni and P. Navrátil, Phys. Rev. C 79, 044606 (2009).
- Hupin et al. (2013) G. Hupin, J. Langhammer, P. Navrátil, S. Quaglioni, A. Calci, and R. Roth, Phys. Rev. C 88, 054622 (2013).
- Baroni et al. (2013a) S. Baroni, P. Navrátil, and S. Quaglioni, Phys. Rev. Lett. 110, 022505 (2013a).
- Baroni et al. (2013b) S. Baroni, P. Navrátil, and S. Quaglioni, Phys. Rev. C 87, 034326 (2013b).
- Hupin et al. (2014) G. Hupin, S. Quaglioni, and P. Navrátil, Phys. Rev. C 90, 061601 (2014).
- Calci et al. (2016) A. Calci, P. Navrátil, R. Roth, J. Dohet-Eraly, S. Quaglioni, and G. Hupin, Phys. Rev. Lett. 117, 242501 (2016).
- Hupin et al. (2015) G. Hupin, S. Quaglioni, and P. Navrátil, Phys. Rev. Lett. 114, 212502 (2015).
- Kravvaris et al. (2020) K. Kravvaris, S. Quaglioni, G. Hupin, and P. Navratil, arXiv preprint arXiv:2012.00228 (2020).
- Navrátil et al. (2016) P. Navrátil, S. Quaglioni, G. Hupin, C. Romero-Redondo, and A. Calci, Physica Scripta 91, 053002 (2016).
- Jaganathen et al. (2014) Y. Jaganathen, N. Michel, and M. Płoszajczak, Phys. Rev. C 89, 034624 (2014).
- Fossez et al. (2015) K. Fossez, N. Michel, M. Płoszajczak, Y. Jaganathen, and R. I. Betan, Phys. Rev. C 91, 034609 (2015).
- Quaglioni et al. (2013) S. Quaglioni, C. Romero-Redondo, and P. Navrátil, Phys. Rev. C 88, 034320 (2013), [Erratum: Phys. Rev. C 94, 019902 (2016)].
- Descouvemont and Baye (2010) P. Descouvemont and D. Baye, Rep. Prog. Phys. 73, 3 (2010).
- Descouvemont (2016) P. Descouvemont, Comput. Phys. Commun. 200, 199 (2016).
- Castaños et al. (1988) O. Castaños, J. P. Draayer, and Y. Leschber, Z. Phys. A 329, 33 (1988).
- Mustonen et al. (2018) M. T. Mustonen, C. N. Gilbreth, Y. Alhassid, and G. F. Bertsch, Phys. Rev. C 98, 034317 (2018).
- Draayer et al. (1989) J. P. Draayer, Y. Leschber, S. C. Park, and R. Lopez, Comput. Phys. Commun. 56, 279 (1989).
- Draayer and Akiyama (1973) J. P. Draayer and Y. Akiyama, J. Math. Phys. 14, 1904 (1973).
- Dytrych et al. (2007) T. Dytrych, K. D. Sviratcheva, C. Bahri, J. P. Draayer, and J. P. Vary, Phys. Rev. Lett. 98, 162503 (2007).
- Launey et al. (2020) K. D. Launey, T. Dytrych, G. H. Sargsyan, R. B. Baker, and J. P. Draayer, Eur. Phys. J. Spec. Top. 229, 2429 (2020).
- Shirokov et al. (2007) A. Shirokov, J. Vary, A. Mazur, and T. Weber, Phys. Lett. B 644, 33 (2007).
- Baker et al. (2020b) R. B. Baker, K. D. Launey, S. Bacca, N. N. Dinur, and T. Dytrych, Phys. Rev. C 102, 014320 (2020b).
- Bacca et al. (2013) S. Bacca, N. Barnea, G. Hagen, G. Orlandini, and T. Papenbrock, Phys. Rev. Lett. 111, 122502 (2013).
- Ekström et al. (2013) A. Ekström, G. Baardsen, C. Forssén, G. Hagen, M. Hjorth-Jensen, G. R. Jansen, R. Machleidt, W. Nazarewicz, et al., Phys. Rev. Lett. 110, 192502 (2013).
- Burrows et al. (2019) M. Burrows, C. Elster, S. P. Weppner, K. D. Launey, P. Maris, A. Nogga, and G. Popa, Phys. Rev. C 99, 044603 (2019).
- Navrátil et al. (2010) P. Navrátil, R. Roth, and S. Quaglioni, Phys. Rev. C 82, 034609 (2010).
- Ekström et al. (2015) A. Ekström, G. R. Jansen, K. A. Wendt, G. Hagen, T. Papenbrock, B. D. Carlsson, C. Forssén, M. Hjorth-Jensen, P. Navrátil, and W. Nazarewicz, Phys. Rev. C 91, 051301 (2015).
- Launey et al. (2012) K. D. Launey, T. Dytrych, and J. P. Draayer, Phys. Rev. C 84, 044003 (2012).