Skip to main content

REVIEW article

Front. Phys., 24 September 2020
Sec. Nuclear Physics​
This article is part of the Research Topic The Future of Nuclear Structure: Challenges and Opportunities in the Microscopic Description of Nuclei View all 11 articles

Self-Consistent Green's Function Theory for Atomic Nuclei

  • IRFU, CEA, Université Paris-Saclay, Gif-sur-Yvette, France

Nuclear structure theory has recently gone through a major renewal with the development of ab initio techniques that can be applied to a large number of atomic nuclei, well-beyond the light sector that had been traditionally targeted in the past. Self-consistent Green's function theory is one among these techniques. The present work aims to give an overview of the self-consistent Green's function approach for atomic nuclei, including examples of recent applications and a discussion on the perspectives for extending the method to nuclear reactions, doubly open-shell systems, and heavy nuclei.

1. Introduction

The theoretical description of atomic nuclei is particularly challenging, for several reasons. First, different energy scales are at play, which is manifest in the rich set of observables and phenomena one is confronted with1. As a consequence, the choice of relevant degrees of freedom might not be universal but depend on the particular properties one is interested in. The standard description in terms of nucleons, i.e., protons and neutrons, leads to a many-body Schrödinger equation for up to a few hundred particles, which are too many to be easily treated exactly but too few to be amenable to a statistical treatment2. Furthermore, interactions between nucleons should be derived from quantum chromodynamics (QCD) in its nonperturbative regime, which prevents direct calculations and requires an additional modelization effort3. In the past, all these features have hindered full solutions of the many-body Schrödinger equation and favored the development of a plentitude of models, following different strategies and characterized by different ranges of application. Although successful in reproducing experimental observations, these models are not always comparable to each other and do not provide a coherent and unified description of nuclear systems.

Only relatively recently so-called ab initio calculations of atomic nuclei, i.e., systematically improvable solutions of the many-body Schrödinger equation that start solely from the knowledge of inter-nucleon interactions, have become available for a large number of isotopes. These advances were made possible by the concomitance of different factors. First, new formal developments of many-body techniques were carried out. Second, chiral effective field theory (χ-EFT) was introduced in nuclear physics [35], providing a systematic and consistent framework in which the nuclear Hamiltonian can be modeled. Third, similarity renormalization group (SRG) techniques were applied to χ-EFT Hamiltonians [6], which largely improves the convergence of actual calculations with respect to basis truncations, thus enabling the handling of heavier nuclei. Finally, the availability of computational tools and in particular high-performance computing resources largely increased. As a result, at present, ab initio calculations (with two- and three-nucleon interactions) can be performed for a considerable fraction of the nuclei with mass number A ≲ 100.

Among the techniques applicable to the nuclear many-body problem, one can distinguish “virtually exact” and “approximate” approaches. Approaches in the first category do not impose any formal approximation on the solution of the many-body Schrödinger equation, which is thus affected only by basis truncation and numerical errors. Typical examples of such methods are Quantum Monte Carlo [7, 8], no-core shell model (or configuration interaction) [9] or nuclear lattice EFT [10]. The second category includes techniques that do approximate the solution of the Schrödinger equation in a systematic way. This solution can be improved, in principle, up to the recovery of the exact one. This is typically achieved by first selecting a reference state and then defining an expansion on top of it, usually in terms of particle-hole correlations (whence the denomination correlation-expansion methods). In doing so, the main advantage resides in the scaling with the basis size: while exact methods scale factorially or exponentially with the system size, correlation-expansion methods only scale polynomially and can be thus applied to a much wider set of nuclei.

The simplest (and most inexpensive) among correlation-expansion approaches is many-body perturbation theory (MBPT) [11]. Following the belief that the nuclear many-body problem is intrinsically non-perturbative, MBPT was put aside for several decades after its development in the 1950s. Only relatively recently, with the advent of EFTs and, specially, SRG techniques applied to nuclear Hamiltonians, MBPT was revived [12, 13] and it is now considered as a convenient approach for large-scale systematic calculations. Moreover, it can be easily complemented with resummation techniques like Padé or eigenvector continuation [14, 15] (see [16] for a recent review). Other methods rely on the MBPT concept but are built such that infinite subsets of MBPT contributions are resummed by construction. The three typical examples employed in the nuclear context are in-medium similarity renormalization group (IMSRG), coupled-cluster (CC) and self-consistent Green's function (SCGF) methods. While IMSRG has been originally designed for the nuclear many-body problem not long ago [17, 18], CC and SCGF, although initially proposed for nuclear systems, have been largely developed in quantum chemistry and solid-state physics before being reimported into nuclear physics starting from the 1990s [19, 20] (see [21] for a recent CC review). IMSRG [22], CC [23], as well as no-core shell model [24] have also been adapted to derive a valence-space interaction that can be used in standard shell-model codes, thus further enlarging the reach of ab initio calculations.

The present manuscript deals with one of such correlation-expansion approaches, the self-consistent Green's function method. Although GFs have been and are implemented in different ways in the context of atomic nuclei (see e.g., [25]), here the focus is on ab initio SCGFs. Early SCGF calculations with realistic nucleon-nucleon potentials dealt mainly with infinite nuclear matter [19]. Formal developments setting the bases for modern implementations in finite nuclei date back to the early 2000s [2628]. By the end of the decade advanced SCGF implementations were routinely applied [29, 30]. In 2011, standard (i.e., Dyson) SCGF theory was generalized to a particle-number-breaking (i.e., Gorkov) framework [31, 32]. In 2013 the inclusion of three-body interactions was formalized in Dyson theory [33]. In 2018, working equations for the state-of-the-art many-body truncation used in nuclear structure calculations (algebraic diagrammatic construction at third order, see section 3.1) were derived [34].

Building on these formal advances, several applications (based on either Dyson or Gorkov frameworks) have been carried out in the past decade. Typical examples are ground-state properties of medium-mass nuclei, going from the oxygen region [35, 36] to silicon and sulfur [37], calcium [38, 39], and nickel [40] regions, up to the very recent computation of tin and xenon isotopes [41]. As discussed in sections 2 and 4.2, excited spectra of odd-even nuclei are also easily accessible in SCGF theory and were studied in particular in [36, 37, 40]. Excited states of even-even systems were addressed in the form of the electromagnetic dipole response in [42]. In [43], ab initio optical potentials were computed and applied in elastic scattering off 16O and 40Ca. Barbieri et al. [44] instead discussed lepton-nucleus scattering, with particular attention to neutrino scattering off 40Ar. Dedicated applications studied effective charges (typically employed in shell-model calculations) [45] and the scale dependence of effective single-particle energies and other non-observable quantities [46]. The possibility of using nucleon-nucleon interactions derived from lattice QCD calculations was also explored in [2]. Last but not least, several collaborations with experimental groups have led to testing the method on e.g., state-of-the-art measurements of nuclear masses [47, 48], energies and spins of excited states [4952], charge and matter radii [53].

The present article discusses a few of these examples, without any pretension of exhaustivity but with the aim of giving the reader a perception of the reach and versatility of SCGF method, as well we the great deal of possible applications to atomic nuclei. A short introduction to the formalism and implementation in finite nuclei is presented beforehand; however, again, it is far from being complete. The reader interested in a comprehensive treatment of GF theory in a modern form is referred e.g., to the book of Dickhoff and Van Neck [54]. Older, yet valuable sources are the books of Nozières [55] and Abrikosov et al. [56]. An interesting work, although tailored to solid-state physics, is the one of Economou [57]. An extensive review covering SCGF applications to nuclear physics appeared in 2004 [28]. A more recent pedagogical introduction to the basics of SCGF formalism and implementation in both finite nuclei and infinite nuclear matter (including computational details and examples of numerical codes) can be found in [58]. A numerical code including a second-order evaluation of the self-energy in the Dyson framework is publicly available [59].

This manuscript is organized as follows. In section 2, the most important concepts and equations of GF theory are introduced. Section 3 describes the actual implementation of the methods in calculations of finite nuclei, briefly reviewing the most commonly used self-energy truncations, the working equations, the choice of basis and interaction. In section 4, some representative applications to ground-state properties, excitation spectra, and lepton-nucleus collisions are discussed. Finally, section 5 addresses perspectives and present as well as future challenges, focusing on three directions of research: the consistent treatment of nuclear structure and reactions, the generalization to doubly open-shell systems and the extension to heavy nuclei.

2. Basic Concepts and Equations

Many-body Green's function4 theory comprises a set of techniques that originated in quantum field theory (QFT) and have been subsequently imported in the (non-relativistic) quantum many-body problem. The late 1950's and the 1960's marked the beginning of the field, with flow of QFT ideas and development of formalism. Since the 1970's technical developments were realized and the approach was applied throughout several disciplines and types of many-body problems, ranging from many-electron systems such as crystals, molecules, and atoms to many-nucleon systems such as nuclei and nuclear matter. Starting from the 1990's such techniques were implemented as an ab initio method in nuclear physics.

As for other many-body methods, the purpose of such techniques is to achieve an approximate (yet systematically improvable) solution of the A-body Schrödinger equation. In standard (or Dyson) many-body Green's function theory, this is realized by rewriting the Schrödinger equation in terms of one-, two-, …, A-body objects gI(= g), gII, …, gA named propagators or, indeed, Green's functions (GFs). Each of these objects is then expanded in a perturbation series, which in practical applications is truncated to include a subset of all possible contributions. In self-consistent schemes such series are themselves expressed in terms of the exact GFs, which requires an iterative solution and makes the method intrinsically non-perturbative, effectively resumming an infinite subsets of perturbative terms. The x-body GF gx allows one to compute all x-body observables in the A-body ground state5. Therefore, for most applications one is mainly interested in the one-body GF.

Formally g is defined as the expectation value of a time-ordered product of annihilation and creation operators in the A-body ground state |Ψ0A

gαβ(tα,tβ)-i Ψ0A|T[aα(tα)aβ(tβ)]|Ψ0A,    (1)

with Greek indices labeling basis states of the one-body Hilbert space H1. One usually works with the Fourier transform of Equation (1) in the energy domain and recasts its perturbation series into the Dyson equation

gαβ(ω)=gαβ(0)(ω)+γδgαγ(0)(ω)Σγδ(ω)gδβ(ω),    (2)

where g(0) represents some initial ansatz for g, e.g., stemming from the solution of Hartree-Fock (HF) equations. The (irreducible) self-energy Σ encodes all terms of the expansion, which is truncated in actual calculations.

Once the one-body GF is obtained as the solution of Equation (2), the expectation value of any one-body operator Ô1BαβOαβ1Baαaβ can be computed as

Ô1B=αβCdω2πiOαβ1Bgβα(ω)=αβOαβ1Bρβα,    (3)

where C↑ denotes an integral closed on the upper imaginary plane and the one-body density matrix

ραβΨ0A|aβaα|Ψ0A=Cdω2πigαβ(ω)    (4)

has been introduced.

Additionally, for a Hamiltonian with one- and two-body operators, the one-body propagator gives access to the total energy by means of the Galitski-Migdal-Koltun (GMK) sum-rule [60, 61]

E0A=αβ12Cdω2πi[Tαβ+ωδαβ]gβα(ω),    (5)

where Tαβ denote the matrix elements of the one-body operator. Nowadays, realistic nuclear structure calculations require the inclusion of at least a three-body interaction in the starting Hamiltonian. In this case, the GMK sum rule needs to be generalized to [33]

E0A=αβ12Cdω2πi[Tαβ+ωδαβ]gβα(ω)-12W,    (6)

where the ground-state expectation value of the three-nucleon operator Ŵ has to be evaluated. Such a term requires in principle the knowledge of the three-body propagator gIII. This is however currently out of reach and in most of practical applications the last term in Equation (6) is computed as

W16αβμγδνWαβμ,γδν ργα ρδβ ρνμ,    (7)

i.e., by approximating the full three-body density matrix with the antisymmetrized product of the one-body one. In [35], this was shown to introduce errors smaller than 250 keV for the binding energy of oxygen isotopes.

In addition to giving access to the ground-state properties of the A-body system, the one-body GF contains information on neighboring (A ± 1) nuclei. It becomes evident when rewriting the propagator (Equation 1) in the Lehmann representation

gαβ(ω) = nΨ0A|aα|ΨnA+1ΨnA+1|aβ|Ψ0Aω-En++iη                 +kΨ0A|aβ|ΨkA-1ΨkA-1|aα|Ψ0Aω-Ek--iη,    (8)

where |ΨiA±1 represent eigenstates of (A ± 1)-body systems while En+(EnA+1-E0A) and Ek-(E0A-EkA-1) are one-nucleon addition and removal energies, respectively.

In Dyson GF theory the expansion starts from a particle-number conserving (e.g., a Hartree-Fock) reference state providing g(0). On top of this, spherical symmetry is typically imposed. While such an expansion can suitably address closed-shell systems, it becomes inefficient or even breaks down as soon as pairing and/or quadrupole correlations become important. If one wishes to stick with a single-reference method, a possible solution consists in working, from the outset, with a symmetry-breaking reference state. In particular, breaking U(1) symmetry associated with particle number conservation6 while maintaining spherical symmetry allows one to efficiently capture pairing correlations, thus gaining access to singly open-shell nuclei.

Dyson GFs were thus generalized to a U(1) symmetry-breaking (typically Hartree-Fock-Bogolyubov) reference state originally by [62]. The formalism was then adapted and implemented for applications to finite nuclei in [31]. Technically, the extension is achieved by working with an A-body ground state that is a linear combination of states with different particle numbers

Ψ0AΨ0=AcAΨ0A.    (9)

This leads to the definition of four one-body propagators

gαβ11(t,t)-iΨ0|T[aα(t)aβ(t)]|Ψ0,    (10a)
gαβ12(t,t)-iΨ0|T[aα(t)āβ(t)]|Ψ0,    (10b)
gαβ21(t,t)-iΨ0|T[āα(t)aβ(t)]|Ψ0,    (10c)
gαβ22(t,t)-iΨ0|T[āα(t)āβ(t)]|Ψ0,    (10d)

two of which (g11 and g22) involve normal combinations of a and a and are associated to the standard density matrix (Equation 3). The remaining two propagators (g12 and g21) invoke so-called anomalous contributions of a and a (interpreted as the annihilation or the creation of a nucleon pair) and lead to the definition of an anomalous (or pairing) density matrix

ρ~αβΨ0|āβaα|Ψ0=Cdω2πigαβ12(ω).    (11)

In Equation (10), creation operators {āα} define a one-body basis dual to {aα} and are obtained via

āa(t)ηaaã(t),  āa(t)ηaaã(t),    (12)

which correspond to exchanging the state a with its time-reversal partner ã up to the phase ηa [31]. The four Gorkov propagators (Equation 10) can be conveniently recast in a 2 × 2 matrix notation via Nambu's formalism [63]

gαβ(t,t)(gαβ11(t,t)gαβ12(t,t)gαβ21(t,t)gαβ22(t,t)).    (13)

All quantities (operators, self-energy, …) can be generalized in an analogous fashion such that one ends up with the Gorkov equation

gαβ(ω)=gαβ(0)(ω)+γδgαγ(0)(ω)Σγδ(ω)gδβ(ω).    (14)

Similarly, all standard GF equations including Equations (3)–(8) are rewritten in a matrix form. Last but not least, a chemical potential λ needs to be introduced to guarantee that the number of particles is the correct one on average. This amounts to replacing the Hamiltonian Ĥ with the grand potential

Ω^Ĥ-λÂ.    (15)

As a consequence of the symmetry breaking, observables might be contaminated by components associated to different particle numbers. Even if in practice the variance is expected to remain small7, the broken symmetry has to be eventually restored. While symmetry-restored formalism has been developed for other (post-Hartree-Fock-Bogolyubov) many-body methods [64, 65], it remains to be formulated for Gorkov GFs.

3. Implementation for Atomic Nuclei

3.1. Choice of Approximation Scheme

Self-consistent GF approximation schemes are defined by the content of the irreducible self-energy, which is expressed as a function of the exact GFs and encodes its perturbative expansion. There exist several ways of approximating the self-energy. The most basic one simply amounts to truncating the perturbative expansion at a certain order. More refined techniques resort to including infinite subsets of perturbation-theory terms via the definition of implicit equations. This is the case, e.g., of the so called ladder or in-medium T-matrix approximation that resums all multiple particle-particle scattering contributions8. This approximation scheme gained considerable attention in early ab initio applications because of its ability of tackling nucleon-nucleon interactions with strong short-range components [6772]. These truncations and resummations are typically conveniently expressed in terms of Feynman diagrams, which facilitate the manipulation of the various terms and give an insight in their physical content.

An alternative, although not orthogonal, route was proposed in the context of quantum chemistry [73, 74] and instead exploits the analytical structure of the self-energy. Similarly to the one-body GF, the exact (dynamical, i.e., energy-dependent, part of the) self-energy displays a Lehmann representation

Σαβ(ω)=nnMαn[1ω-(E>+C)+iη]nnMnβ                 +kkNαk[1ω-(E<+D)-iη]kkNkβ,    (16)

where the matrices M, N couple the single-particle motion of the nucleons (i.e., the one-body propagator in which the self-energy is inserted) to intermediate multiparticle-multihole configurations, whose energies (“bare” and resulting from the interference between them) are encoded in the matrices E>, E<, and C, D, respectively. The algebraic diagrammatic construction at order n [ADC(n)] is built by demanding that, in addition to including all perturbation-theory contributions up to a given order n, the approximated self-energy has the same analytical structure as the exact one, i.e., in particular, is the same function of the energy. The latter condition requires the self-energy to contain additional sets of contributions, e.g., infinite resummations that would necessitate ad-hoc procedures are in this way automatically included in the ADC formalism. The first order, ADC(1), is simply the standard Hartree-Fock (or Hartree-Fock-Bogolyubov, in the case of Gorkov GFs) approximation. ADC(2) introduces lowest-order dynamical correlations in terms of two particle-one hole and two hole-one particle contributions. ADC(3) builds couplings between such configurations and, as a result, includes infinite-order resummations of both particle-particle/hole-hole and particle-hole ladders. Higher orders build on higher-rank particle-hole excitations in a similar but not identical fashion as in other popular many-body methods like coupled-cluster (CC) or in-medium SRG (see e.g., [75] for a connection between GF and CC formalisms). Importantly, by preserving at each order the analytical properties of the exact self-energy, the ADC expansion ensure that causality is not violated. Moreover, this form allows the derivation of an energy-independent auxiliary eigenvalue problem that significantly simplifies the numerical solution of Dyson and Gorkov equations, as discussed in section 3.2.

The ADC scheme has been developed in the context of finite nuclei in the past few years. At present, ADC(1), ADC(2), and ADC(3) self-energies are implemented in the Dyson framework [35, 37], while ADC(1) and ADC(2) are available in the Gorkov case [31, 32]. Gorkov-ADC(3) is under construction within a project whose goal is to automatize the generation of the associated self-energy diagrams. In Dyson theory, the ADC formalism has been recently generalized to the presence of three-body interactions [34].

Generally speaking, ADC(n) defines a truncation scheme that is systematically improvable in the sense that going to higher orders should provide results closer and closer to the exact solution, recovered in the limit of ADC(∞). Nevertheless, in the case of perturbation theory (and, by consequence, resummation methods building on MBPT), there does not exist a well-defined expansion parameter informing on the accuracy associated to a certain truncation level. Hence, there are two ways of assessing the accuracy of a given approximation: (i) by comparing successive orders in the expansion and (ii) via benchmarks with exact methods. Concerning the first possibility, a typical convergence behavior for ADC(1-3) is shown in Figure 1 for ground-state energies and root-mean-square (rms) charge radii of two representative medium-mass nuclei. Two different interactions are used, the “soft” NN+3N(lnl) and the “harder” NNLOsat (see section 3.4 for details). One observes a clear pattern going from ADC(1) to ADC(3) for all quantities. For total energies, while ADC(2) already yields a qualitatively good account, additional correlations introduced at the ADC(3) level are deemed necessary for an accurate description. In contrast, charge radii are already largely converged at the ADC(2) level. After analysing the ADC convergence behavior, one expects ADC(4) corrections to be small for total energies and negligible for charge radii. The accuracy of the ADC(3) scheme is also confirmed by a direct benchmark against no-core shell model in 16O, with total energies computed in the two approaches differing by less than 1% for any tested interaction [40].

FIGURE 1
www.frontiersin.org

Figure 1. Ground-state energies (top) and rms charge radii (bottom) of 36Ca and 68Ni computed within different ADC(n) truncation schemes. Results for the NNLOsat and NN+3N(lnl) interactions are displayed. Reprinted figure with permission from [40], copyright (2020) by the American Physical Society.

3.2. Working Equations

In practice, solving Dyson or Gorkov equations in their forms Equations (2) and (14) is problematic, specially for finite systems where a solution has to be achieved for numerous (discrete) values of the energy ω. A method to overcome the problem was proposed once again in quantum chemistry [76], and consists of two steps. In the first one, after exploiting the analytical energy dependence of the propagator, Equation (2) is transformed into an eigenvalue equation where the (energy-dependent) self-energy acts as an effective one-body potential. The second steps makes use of the analytical energy dependence of the self-energy in the form of Equation (16) and rewrites the problem as an energy-independent eigenvalue equation. The latter constitutes the working equation to be solved iteratively, and whose solutions directly provide amplitudes and energies entering Equation (8) for the one-body propagator (see [58] for a more detailed discussion). The derivation of this energy-independent eigenvalue problem has been generalized to Gorkov theory in [31].

One disadvantage of the energy-independent formulation is that the number of energy poles, i.e., the eigenvalues and the dimensionality of the energy-independent Dyson or Gorkov matrix, increases at each self-consistent iteration. In practice, this growth is reduced via the application of Krylov projection techniques [76, 77], typically implemented by means of a Lanczos algorithm (see [32] for a detailed discussion and a numerical study).

3.3. Choice of Basis

Equations presented in section 2 are general, i.e., are valid in any basis of choice9. In an actual calculation, one needs to specify a basis in which operators, together with all relevant quantities, are expanded on. In the case of atomic nuclei one typically employs a one-body spherical harmonic oscillator (HO) basis, whose eigenfunctions are well suited to the description of a confined system10. A HO basis is characterized by two parameters: the oscillator inverse length ℏΩ and the number of considered HO wave functions. The latter is usually determined by the parameter emax ≡ max(2n + l), which sets the energy threshold of a basis eigenfunction. Many-body bases are subsequently built as direct products of one-body bases. While naturally for a k-body operator one would set ekmax = k · emax, the storage of three-body matrix elements for realistic values of emax presently constitutes an issue and obliges one to work with e3max〈〈3 · emax. In current state-of-the-art implementations, typical values of emax = 12 − 15, e2max = 2 · emax, and e3max = 14 − 18 are used.

The first step of a calculation consists in studying the convergence properties (of the observables of interest) with respect to the basis size and oscillator parameter. An example is shown in Figure 2, where the basis dependence of ground-state energy and charge radius is investigated for two different nuclei and two different Hamiltonians with ADC(2) calculations. One sees that in 36Ca an emax = 13 model space guarantees sufficiently converged results. In 68Ni and for the higher-cutoff Hamiltonian NNLOsat (see next section for details) the convergence is not yet optimal.

FIGURE 2
www.frontiersin.org

Figure 2. Ground-state energies (top) and rms charge radii (bottom) of 36Ca and 68Ni computed with the NN+3N(lnl) and NNLOsat Hamiltonians as a function of the harmonic oscillator parameter ℏΩ and for increasing size emax of the one-body basis. In all cases e3max = 16 was used. Reprinted figure with permission from [40], copyright (2020) by the American Physical Society.

Combining uncertainties from basis truncation, discussed here, and from many-body expansion, discussed in section 3.1, one can evaluate the total theoretical error of the method for a given input interaction. To give an example [40], for soft potentials, like the NN+3N(400) and the NN+3N(lnl) Hamiltonians presented in the next section, one estimates an error of about 2% (< 0.5%) associated to ground-state energies (radii) of medium-mass nuclei. For harder potentials, like the NNLOsat Hamiltonian, the errors rise to 4 and 1% for energies and radii, respectively. To a good extent, these uncertainties cancel out when differential observables are considered, as in the case of one- or two-nucleon separation energies. In the results presented in the following such systematic uncertainties are not explicitly reported in the figures (except for Figures 8, 10) and should be kept in mind by the reader.

3.4. Choice of Input Interaction

Like most of ab initio techniques, self-consistent GFs can take in principle any nucleon-nucleon plus three-nucleon (NN+3N) Hamiltonian as input. In the early 2000s, nearly all applications were performed using semi-phenomenological11 potentials like CD-Bonn [84] or Nijmegen [85], possibly complemented with three-nucleon forces [72, 86]. Although these interactions had proven successful in the description of light nuclei, their “hard” character, i.e., the associated large momentum cutoff, required the use of sophisticated resummation schemes, thus hindering applications in medium-mass nuclei. Starting from 2010, the use of Hamiltonians derived in the context of chiral effective field theory (χ-EFT) [5] began to spread. Compared to the previous phenomenological models, χ-EFT interactions present several advantages:

1. By explicitly taking into account only low-energy degrees of freedom, they have a much smaller associated cutoff;

2. All many-body (i.e., 2N, 3N, 4N, …) operators and currents are derived consistently without any ad-hoc assumption, which augments the predictive power;

3. A theoretical error can be associated to a given χ-EFT Hamiltonian (which relates to the employed truncation in the EFT expansion).

Nowadays, χ-EFT interactions constitute the standard for ab initio nuclear structure calculations. The first feature, i.e., their “softness,” is often further amplified by the use of similarity renormalization group (SRG) techniques [6], i.e., unitary transformations of the Hamiltonian that further decouple low- and high-momentum modes, leading to much improved convergence properties. The third point implies that the EFT error can (and should) be subsequently propagated to many-body observables. This has been done in practice only very recently (see e.g., [16, 8792]) and more formal and technical developments along these lines will be required in the future.

Mainly three different χ-EFT Hamiltonians have been employed in recent GF calculations, all of which are discussed here. The first one, labeled NN+3N(400), is based on the next-to-next-to-next-to-leading order (N3LO) nucleon-nucleon potential from Entem and Machleidt [4, 93] combined with the N2LO 3N interaction with a local regulator [94]. The 2N interaction of [93] was built with a cutoff of 500 MeV/c, however, a 400 MeV/c regulator was used for the 3N sector [95]. This Hamiltonian has been systematically applied to p- and sd-shell nuclei and yields a good reproduction of oxygen, nitrogen and fluorine binding energies [35, 36, 96]. Nevertheless, it leads to overbinding in medium-mass nuclei starting in the calcium chain and underpredicts nuclear radii even for O isotopes [39, 53, 97].

With the main objective of improving on the description of radii, a chiral Hamiltonian with terms up to N2LO was developed in [98]. It is characterized by a simultaneous fit of 2N and 3N LECs that does not rely solely on two-nucleon and A = 3, 4 data, but also on binding energies of 14C and 16,22,24,25O as well as charge radii of 14C and 16O. The resulting interaction, named NNLOsat, successfully describes the saturation of infinite nuclear matter [98] as well as various observables in mid-mass nuclei, including charge radii [4244, 53]. Unlike the NN+3N(400) interaction, NNLOsat employs a non-local regulator.

Motivated by the success of NNLOsat, a novel interaction named NN+3N(lnl) was presented recently [40]. The goal was to amend the original NN+3N(400) interaction, and in particular its 3N part. While the latter has been shown to be problematic, its 2N part is instead believed to perform relatively well and thus is kept unchanged. Being based on the N3LO potential, which provides a better description of nucleon-nucleon data compared to the lower-order NNLOsat, it yields superior features in light systems, e.g., a better reproduction of spectroscopy of natural parity states in p- and light sd-shell nuclei. NN+3N(lnl) has been also shown to provide a very good description of energy observables (ground-state energies and energy spectra) in medium-mass nuclei up to mass A ~ 60 [40].

3.5. Computational Requirements

As any other state-of-the-art ab initio nuclear structure approach, the SCGF method requires the development of an advanced numerical code. The computational cost of a simulation strongly depends on (i) the size of the model space, (ii) the chosen level of approximation12, and (iii) the two- or three-body character of the input Hamiltonian. ADC(1) calculations with only NN forces can be easily performed on a laptop also in large bases (in few CPU minutes). Going to ADC(2) limits a laptop calculation to a small, yet (semi-)realistic model space, typically emax = 8 − 9 (doable in a few CPU hours). If larger bases are needed (to ensure model-space convergence, e.g., typically emax = 12 − 13), then one has to resort to a dedicated computer cluster (with a corresponding cost of few hundred CPU hours). ADC(3) is doable on a laptop only for very small model spaces and any realistic calculation requires the implementation of MPI parallelization and the use of a high-performance computing center (with running times of several thousand CPU hours).

The inclusion of 3N forces results into an increase of both CPU time (due to the higher rank of the tensors at play) and memory usage (due to the larger amount of matrix elements to be stored). As a consequence, on a laptop and for realistic bases, even an ADC(1) calculation becomes heavy in terms of CPU and one quickly reaches the limits in terms of available RAM13. ADC(2) and ADC(3) calculations require optimized implementations and the use of a high-performance computing center, with typical running times of a few thousand and tens of thousands CPU hours, respectively.

4. Recent Applications

4.1. Ground-State Properties

The total ground-state energy, or binding energy, of a nucleus constitutes the most basic nuclear structure observable. In Green's function theory, total energies are preferably computed via the generalized GMK sum-rule (Equation 6). While earlier applications made use of a 2N-only Hamiltonian, possibly complemented by a phenomenological correction to compensate for missing 3N [29, 30, 38], starting from 2013 calculations with realistic 2N + 3N interactions could be routinely performed. A representative example concerns the oxygen chain [35] and is shown in Figure 3 (left). In the bottom panel, ADC(3) ground-state energies are displayed for closed-shell oxygen isotopes, computed with the NN+3N(400) interaction, respectively excluding and including original 3N operators14. One notices that the addition of 3N forces is crucial for a quantitative reproduction of experimental data. In particular, when only a 2N interaction is considered, the neutron dripline is wrongly located at N = 20, while it is correctly reproduced at 24O in the presence of 3N forces.

FIGURE 3
www.frontiersin.org

Figure 3. (Left) ADC(3) calculations of closed-shell oxygen isotopes performed with the NN+3N(400) Hamiltonian (blue points and lines). For reference, results with only original 2N operators are displayed (red points and lines). (Top) One-neutron addition and removal energies associated to dominant quasiparticle peaks. (Bottom) Ground-state energies compared with experimental values (gray bars). Adapted figure with permission from [35], copyright (2013) by the American Physical Society. (Right) ADC(2) calculations of even-even calcium isotopes performed with the NN+3N(400), NNLOsat, and NN+3N(lnl) Hamiltonians (colored points and lines), compared with measured and extrapolated data (black points). ADC(3) results are depicted as horizontal lines when available. (Top) Total ground-state energies. (Bottom) Two-neutron separation energies. Adapted figure with permission from [40], copyright (2020) by the American Physical Society.

In such a context it can be instructive to inspect one-neutron addition and removal energies associated to dominant quasiparticle peaks, see top-left panel of Figure 3. One sees that the 3/2+ fragment becomes bound in neutron-rich isotopes when the 2N-only Hamiltonian is employed. When 3N interactions are switched on, it is instead pushed up and remains unbound all the way to 28O, thus explaining the position of the dripline. This observation confirmed the repulsive character of the Fujita-Miyazawa 3N interaction, as previously discussed in [99].

This result was one of the first successful applications of ab initio techniques beyond light nuclei. The oxygen chain also constituted, for a few years, a testbed where calculations from various approaches could be benchmarked, demonstrating the reliability of the different many-body truncations [23, 96, 100, 101]. More recently, the availability of new Hamiltonians prompted calculations of heavier systems, from calcium up to the nickel chain. An example is constituted by Gorkov GF calculations exploring the performance of three interactions along few semi-magic chains, as reported in [40]. A typical outcome is the one displayed in Figure 3 (right), where ground-state energies of calcium isotopes as well as their differences, two-neutron separation energies, are shown as a function of neutron number. The overbinding generated by the older NN+3N(400) clearly stands out. In contrast, the newer Hamiltonians NNLOsat and NN+3N(lnl) yield an excellent reproduction of total as well as differential ground-state energies, specially once ADC(3) corrections are taken into account. A good performance is found also in the nickel chain, up to the point where current computational limitations hinder a complete model-space convergence of the calculations [40]. In addition to semi-magic chains, the theory was tested in relation to novel experimental measurements in potassium [47], titanium [48], and argon [102] chains.

The performance of these three Hamiltonians was further investigated on nuclear radii in [40]. It was found that NNLOsat provides a good account of rms charge radii all the way up to nickel. Specifically, the bulk contributions are well-described already at the ADC(2) level, while finer details (e.g., the parabolic behavior observed between 40Ca and 48Ca) need further improvement of the many-body truncation and/or the interaction. Density distributions provide even further insight into the way nucleons arrange themselves in the correlated nuclear medium. The nuclear charge density distribution is typically obtained as a sum of three contributions [103],

ρch(r)=ρchp(r)+ρchn(r)+ρchls(r),    (17)

where ρchp (ρchn) is determined by folding the point-proton (point-neutron) density with the finite charge distribution of the proton (neutron) and ρchls is a relativistic correction that depends on spin-orbit terms. In addition, center-of-mass and relativistic Darwin-Foldy corrections are taken into account by employing an effective position variable [104]. Finite nucleon charge distributions can be expressed as a sum of Gaussians, with the parameters adjusted to reproduce form factors from electron scattering data [105]. The relativistic spin-orbit correction is usually computed within the factorization approximation introduced in [106]. Even though ρchp largely dominates, the other two contributions can visibly alter the total charge distribution in some cases.

As for radii, densities are computed directly from the one-body GF and can therefore be routinely evaluated for medium-mass systems. An interesting example relates to the possible presence of a depletion in the central part of the charge density profile, usually referred to as bubble. One of the most likely candidates has been identified in the nucleus 34Si [107113]. In [37] this system, together with its Z + 2 partner 36S, has been thoroughly investigated by means of both Dyson and Gorkov GF calculations. The resulting point-proton and charge density distributions are shown in Figure 4 (left). The effect of folding with the finite size of the proton is specially visible in the center, with an attenuation of oscillations in the charge profiles. The agreement between the computed and the measured charge distribution for the stable nucleus 36S is excellent, which gives confidence in the prediction for the unstable 34Si, yet unknown experimentally. For the latter, a depletion in the region below 2 fm is indeed found. Its magnitude is comparable or even larger than what previously obtained in EDF [111, 112] or shell model calculations [110].

FIGURE 4
www.frontiersin.org

Figure 4. (Left) Point-proton and charge densities of 34Si and 36S computed at the ADC(3) level with the NNLOsat interaction. The experimental charge density of 36S is taken from [114]. (Right) Angular dependence of the charge form factor computed for 300 MeV electron scattering. Modified figure with permission from [37], copyright (2017) by the American Physical Society.

A measurement of the charge distribution of 34Si would require electron scattering on radioactive ions. These experiments are becoming feasible only now [115], with first results on the heavier 132Xe obtained by the SCRIT collaboration [83]. In the coming years, the case of 34Si will thus constitute an interesting objective for electron-nucleus scattering facilities. Such a measurement would extract the electromagnetic charge form factor, related to the charge profile via

F(q)=drρch(r)e-iq·r,    (18)

where q is the transferred momentum, itself related to the incident momentum p and the scattering angle θ via q = 2p sinθ/2. The calculated charge form factors for 300 MeV electron scattering on 34Si and 36Si are displayed in Figure 4 (right). Clear differences appear in the angular dependence for the two systems, with a higher magnitude and a displaced position of the second minimum for 34Si. This analysis gives indications on what range of transferred momenta, and consequently which luminosities, are necessary for identifying possible depleted density profiles in this mass region.

4.2. Excited-State Properties

An asset of GF theory resides in the rich content of the one-body propagator, which does not solely provide information on the targeted (even-even) system with mass number A but also on the four neighboring (odd-even) A ± 1 nuclei. This information is explicit in the Lehmann representation of the GF (Equation 8). The poles Ei± of the function correspond to one-nucleon addition and removal energies, as schematically depicted in Figure 5 (top left). In addition, the associated amplitudes in the numerator represent the probabilities to reach a specific eigenstate |ΨnA+1 (|ΨkA-1) of the A + 1 (A − 1) system by adding (removing) a nucleon in a single-particle state to (from) the ground state |Ψ0A of the even-even system. Those amplitudes can be expanded in a single-particle basis {aα} according to

UnαΨ0A|aα|ΨnA+1,    (19a)
VkαΨ0A|aα|ΨkA-1.    (19b)
FIGURE 5
www.frontiersin.org

Figure 5. (Top left) Schematic view of the A ± 1 states (and associated energies) reached via one-nucleon addition and removal. (Top right) Neutron spectral strength distribution in 34Si computed at the ADC(1) level with the NNLOsat interaction. The four panels display different angular momentum and parity channels. The dashed vertical line represents the Fermi energy, separating the spectra of 33Si and 35Si. (Bottom) Same as above but for ADC(2) and ADC(3) calculations. Data originates from the results of [37].

Next, spectroscopic probability matrices for the nucleon addition and removal can be built15, Sn+UnUn and Sk-Vk*VkT, respectively. Their elements read as

Sn+αβΨ0A|aα|ΨnA+1ΨnA+1|aβ|Ψ0A,    (20a)
Sk-αβΨ0A|aβ|ΨkA-1ΨkA-1|aα|Ψ0A.    (20b)

Taking the trace over the one-body Hilbert space H1 leads to spectroscopic factors

SFn+TrH1[Sn+]=αH1|Unα|2,    (21a)
SFk-TrH1[Sk-]=αH1|Vkα|2,    (21b)

which are the norms of the spectroscopic amplitudes. A spectroscopic factor thus sums the probabilities that an eigenstate of the A + 1 (A − 1) system can be described as a nucleon added to (removed from) a single-particle state on top of the ground state of the A-nucleon system.

The complete spectroscopic information associated with one-nucleon addition and removal processes can be collected into the spectral function S(ω), defined as the energy-dependent matrix on H1

S(ω)nHA+1Sn+δ(ω-En+)+kHA-1Sk-δ(ω-Ek-),    (22)

where the first (second) sum runs over eigenstates of H in the Hilbert space HA+1 (HA-1) associated with the A + 1 (A − 1) system. Taking the trace of S(ω) gives the spectral strength distribution (SDD)

S(ω)TrH1[S(ω)]           =nHA+1SFn+δ(ω-En+)+kHA-1SFk-δ(ω-Ek-),    (23)

which is a basis-independent function of the energy. Equations (19)–(23) can be generalized to the Gorkov formalism [31].

An example of SDD computed at three different levels of approximation, ADC(1-3), is shown in Figure 5 for the nucleus 34Si. ADC(1), i.e., Hartree-Fock, is a mean-field (or independent-particle) approximation, which translates into a series of quasiparticle peaks with unity spectroscopic factors, i.e., there is a one-to-one correspondence between the single-particle basis {aα} and the many-body states |ΨiA±1 accessible via the process of adding or removing a nucleon. One can nevertheless identify the main qualitative features of the SDD, e.g., associate the ground state of 33Si (35Si) with the first peak on the left (right) of the Fermi energy characterized by spin and parity 3/2+ (7/2). ADC(2) introduces the lowest-order dynamical correlations that lead to a first fragmentation of the spectral distribution. A number of fragments with small spectroscopic factors appear in the vicinity of the ADC(1) peak, which is now shifted in energy and reduced in strength, i.e., it has SFi±<1. ADC(3) correlations further fragment the quasiparticle strength, giving rise to a large number of small peaks and a further reduction of the main-peak spectroscopic factor. One notices that, around the Fermi energy, fragments with a good quasiparticle character, i.e., large spectroscopic factors, survive, in accordance to Landau's Fermi liquid theory. In contrast, away from the Fermi level the strength is spread over a wide energy interval and one can hardly identify single-particle-like excitations.

From the SDD depicted in Figure 5 one can extract an excitation spectrum of the A − 1 (A + 1) system, by looking at increasingly negative (increasingly positive) energies on the left (right) of the Fermi surface (see top-left panel for a schematic representation). By comparing ADC(2) and ADC(3) truncations, one concludes that the latter is necessary for an accurate description of the main energy peaks [37, 40], in accordance with analogous calculations in quantum chemistry [116119]. Conservatively, one can associate uncertainty bands with ADC(2)-ADC(3) differences, as done in Figure 6 where spectra resulting from one-nucleon addition and removal from and to 54Ca are displayed. These four systems sit at the limits of what can be presently studied in radioactive ion beam facilities. Yet, not many experimental data are available, such that these calculations mostly represent predictions that can be presumably tested in the near future. In 53Ca, where the ground state and two excited states are identified, NN+3N(lnl) GF calculations succeed in reproducing the measured spectrum in with good accuracy. NNLOsat instead mildly overestimates the splitting between the two excited states. In 53K, the difference between the two interactions is even more striking. While NN+3N(lnl) reproduces the measured first excited state with good precision, NNLOsat predicts a wrong ordering of the 3/2+ and 1/2+ states, resulting in a ground state with incorrect spin. The potassium chain indeed constitutes an interesting case because of an unusual inversion (at N = 28) and re-inversion (at N = 32) of the spins of the ground- and first excited states [50, 120]. As discussed in [40, 52], Gorkov GF calculations with the NN+3N(lnl) Hamiltonian yield ax excellent reproduction of the whole trend, from 37K to 53K. Such studies, complementary to those focusing on ground-state observables, are not only relevant in relation to current experimental programs, but also provide a unique testing ground for the development of nuclear interactions.

FIGURE 6
www.frontiersin.org

Figure 6. One-nucleon addition (panels B and D) and removal (panels A and C) spectra starting from 54Ca. Available experimental data appear in the left column of each panel. GF calculations performed with the NNLOsat and NN+3N(lnl) interactions are displayed in the central and right column of each panel, respectively. Red lines constitute results obtained at the ADC(3) level. Shaded areas connect ADC(2) and ADC(3) where available. In all panels, states with Ex < 5 MeV and SF < 10% are shown. Modified figure with permission from [40], copyright (2020) by the American Physical Society. Additional new experimental data are from [51, 52].

The spectral representation of the one-body GF thus gives access to spectroscopic studies of odd-even nuclei. If one is interested in the excitations of an even-even system, the two-body GF (in one of its possible time orderings) has to be considered instead. A typical example is the polarization propagator, which provides the response of the nuclear system to an external operator. In analogy to Equation (8), its Lehmann representation reads as

Πγδ,αβ(ω) =nπ0Ψ0A|aδaγ|ΨnπAΨnπA|aαaβ|Ψ0Aω-(EnπA-E0A)+iη                      -nπ0Ψ0A|aαaβ|ΨnπAΨnπA|aδaγ|Ψ0Aω+(EnπA-E0A)-iη,    (24)

where nπ label the excited states of the A-body system. In the numerators, the residues

ZαβnπΨnπA|aαaβ|Ψ0A.    (25)

represent particle-hole matrix elements between excited states of the A-nucleon system. The poles appearing in the denominator

ϵnππEnπA-E0A    (26)

instead constitute energy differences between excited states of the A-nucleon system and its ground-state. The polarization propagator (Equation 24) is obtained as a solution of the Bethe-Salpeter equation,

Πγδ,αβ(ω) =Πγδ,αβ(0)(ω)                              +μρνσΠγδ,μρ(0)(ω)Kμρ,νσ(ph)(ω)Πνσ,αβ(ω),    (27)

where Π(0)(ω) is the free polarization propagator, and the particle-hole irreducible interaction K(ph) plays a similar role as that of the self-energy in Equation (2) for the single-particle propagator. While the corresponding formalism has been developed and implemented for Dyson GFs, the generalization to the Gorkov framework remains to be carried out.

In [42, 121], the polarization propagator was calculated with the aim of accessing the nuclear isovector electric dipole (E1) response. A dressed random phase approximation (DRPA) was adopted for the computation of Π(ω). The scheme makes use of a correlated, e.g., ADC(3), one-body propagator as the starting point for the RPA equations, thus going beyond a simple HF-based particle-hole resummation. Figure 7 shows results for the integrated isovector E1 photoabsorption cross section, which is directly obtained from the polarization propagator [42]. Two representative examples are shown here, 16O and 48Ca. Calculations make use of the NNLOsat interaction, ensuring that nuclei have the correct size, which is a crucial property for this application. In 16O the peak associated to the giant dipole resonance (GDR) is well-reproduced. At higher excitation energies, the calculation underestimates the experimental spectrum, presumably due to missing correlations beyond the simple RPA. A similar picture emerges for 48Ca, with the GDR peak in good agreement with recent experimental measurements [124] and the high-energy tail missing some strength. Several other closed-shell nuclei, for some of which experimental data are not yet available, are discussed in [42]. A generalization of this formalism to the Gorkov framework would allow us to extend these promising results to a large number of open-shell nuclei.

FIGURE 7
www.frontiersin.org

Figure 7. Isovector E1 photoabsorbption cross section of 16O (A) and 48Ca (B) computed with the NNLOsat interaction. Theoretical curves are obtained by folding the discrete spectra with Lorentzian widths Γ = 3 MeV. Experimental data are from [122124]. Adapted figure with permission from [42], copyright (2019) by the American Physical Society.

4.3. Lepton-Nucleus Scattering

The spectral function introduced in Equations (22)–(23) carries information about the energy-momentum distribution of the correlated nucleons. Certain scattering processes, where an external probe scatters off the nucleus, under certain kinematical conditions (e.g., characterized by a sufficiently large momentum transfer), can be described as an incoherent sum of scattering amplitudes on bound nucleons. Then, the cross section can be computed in the so-called impulse approximation and expressed as a sum of one-body terms containing a convolution with the nucleon spectral function.

This is the case, e.g., of electron and neutrino scattering in the region of the quasielastic peak. Here the double differential cross section for inclusive lepton-nucleus scattering can be written as [125]

(dσdEdΩ)=CEkEkLμνWμν,    (28)

where Lμν is the leptonic tensor and k = (Ek, k) and k=(Ek,k) are the laboratory four-momenta of the incoming and outgoing leptons, respectively. The factor C=α/(k-k)4 for electrons and C=G/8π2 for neutrinos, where G = GF for neutral current (NC) and G = GF cosθc for charged current (CC) processes. The electroweak coupling constants are α≃1/137, GF=1.1803×10-5GeV-2 [126] and cosθc = 0.97425 [127]. The hadron tensor Wμν incorporates the transition matrix elements from the target ground state |Ψ0A to the final states |ΨfA due to the hadronic currents, including additional axial terms for neutrino scattering. The impulse approximation consists in factorizing |ΨfA|p|ΨnA-1, i.e. it allows to work with the outgoing nucleon of momentum p′ and the residual nucleus, left in a state |ΨnA-1. This leads to [128, 129]

W1bμν(q,ω)=d3pdE(2π)3mN2e(p)e(p-q)δ(ω+E-e(p))     ×sSsh(p-q,E)p|jsμ|p-qp-q|jsν|p,    (29)

where ω represents the energy transfer, MN is the nucleon mass, e(p) the energy of a nucleon with momentum p. The one-body current operators jsμ depend on the spin-isospin degrees of freedom s and Ssh(p,E) is the one-body spectral function normalized to the total number of nucleons. For two-body currents and hadron production, Equation (29) extends non trivially in terms of one- and two-body spectral functions [129132].

Following the above formalism, electron and neutrino scattering off argon and titanium isotopes was investigated in [44]. The interest in studying lepton scattering off these nuclei resides in the fact that future-generation neutrino experiments (e.g., DUNE [133]) will use liquid-argon time-projection chambers, which rely on scattering of neutrinos off 40Ar. The nuclear component of the cross section has to be well determined for a meaningful interpretation of the measured events, in particular to reconstruct the neutrino energy with sufficient accuracy. In this respect, a tailored electron scattering experiment16 was designed and recently performed at JLab [81, 82, 134]. In [44], nuclear spectral functions of stable argon, calcium and titanium isotopes were computed at the ADC(2) level with the NNLOsat interactions. A good reproduction of available charge radii and density distributions was found. Starting from these spectral functions, inclusive electron-48Ti and electron-40Ar cross sections were computed. They are shown in Figure 8 (left) as a function of the energy transfer and compared to the recent experimental data from the JLab E12-14-012 collaboration [81, 82]. Calculations closely follow the quasi elastic peak, thus validating the theoretical approach and the impulse approximation in particular. Next, the quasielastic neutral and charged current cross sections were studied for 1 GeV neutrino scattering. Results are displayed in the right panel of Figure 8, also compared to scattering off 12C. The quasielastic peak is found at a similar energy, with an increase in the magnitude of the cross section consistent with super scaling properties of inclusive reactions. The use of a Ti proton spectral function instead of an Ar neutron spectral function was also tested and found to be an excellent approximation. Further studies will be needed to thoroughly assess theoretical uncertainties.

FIGURE 8
www.frontiersin.org

Figure 8. (Left) Inclusive Ti(e, e′) (top) and Ar(e, e′) (bottom) cross sections at 2.2 GeV and 15.5° scattering angle. The solid (dashed) line shows the quasielastic cross section with (without) the inclusion of finals-state interactions (FSI, see [44] for details). For the FSI results, the theoretical uncertainties coming from model-space convergence are also shown as a shaded band. Experimental data are taken from [81, 82] and show both the quasielastic peak and the contribution from meson production at larger missing energies. (Right) Quasielastic neutral (top) and charged current (bottom) cross sections for 1 GeV neutrino scattering. Dot-dashed lines refer to a 12C target and solid lines refer to 40Ar. The colored band represents theoretical uncertainties due to model-space convergence. The dotted lines result from using the 48Ti proton spectral function as an approximation for neutrons in 40Ar. The insets show the difference between the latter and calculations where the full spectral distribution of 40Ar is used. Adapted figure with permission from [44], copyright (2019) by the American Physical Society.

5. Perspectives and Challenges

5.1. Toward the Description of Nuclear Reactions

While the good knowledge of electro(weak) interactions facilitate the modeling of lepton-nucleus scattering, nuclear reactions, because of the complexity of strong interactions, require a more involved theoretical description. As a consequence, very few ab initio methods are presently capable of going beyond structure properties and directly computing reaction observables. Nevertheless, nucleon-nucleus and nucleus-nucleus reactions constitute nowadays the tools of choice to study the properties of atomic nuclei, such that progress on their ab initio description would be highly valuable.

A relatively clean process that can be used as a training field for ab initio reaction theory is quasifree nucleon knockout [135]. Quasifree scattering represents a process in which an incident nucleon (typically a proton17) with an energy of few hundred MeV knocks out a bound nucleon from the isotope of interest (see Figure 9, left panel, for a schematic illustration). Kinematical conditions are chosen such that the process can be preferentially described by a single, localized interaction between the incident and the struck nucleons, thus minimizing multiple collisions for the incoming nucleon. This sudden removal mechanism suggests that the remaining A − 1 nucleons can be treated as spectators, which translates into an impulse approximation analogous to the one discussed in section 4.3. As a result, the total cross section can be separated into a structure and a reaction part. The latter involves nucleons that are usually described in terms of distorted waves to account for their propagation under the influence of the nuclear medium. Under these assumptions the differential cross section, labeled distorted-wave impulse approximation (DWIA), can be schematically written as

(dσd3Q)DWIA=SFN×dσpNdΩ×SpASp(A-1)SN(A-1)ΨN.    (30)

The first factor on the right-hand side is the spectroscopic factor of the struck nucleon, encoding the structure properties of the nucleus. The second term represents an in-medium proton-nucleon cross section, determining the probability of the collision between the projectile and one of the bound nucleons. The third term contains the scattering matrices for the (effective) degrees of freedom at play, i.e., respectively (i) proton and initial A-body system, (ii) proton and final (A − 1)-body system, (iii) struck nucleon and final (A − 1)-body system, plus the scattering wave function of the outgoing nucleon.

FIGURE 9
www.frontiersin.org

Figure 9. (Left) Schematic representation of quasifree nucleon knockout: an incoming proton p scatters off an A-body nucleus, knocking out a nucleon N. (Right) One-proton overlap functions (squared) computed at the ADC(3) level using NNLOsat and NN+3N(lnl) interactions [136]. OFs corresponding to the main quasiparticle fragments for three angular momentum and parity channels JΠ = 1/2+, 3/2+, 5/2+ are shown. For comparison, EDF calculations [137, 138] performed with the SLy4 parameterization are displayed. These OFs were employed in the calculation of one-proton knockout cross sections in [52].

An ab initio calculations for some of the above quantities was recently performed within the GF formalism, in connection with a quasifree neutron knockout study on 54Ca [51]. Specifically, spectroscopic factors and one-body overlap functions for the struck neutron were computed. The latter enter the evaluation of the scattering matrices, where they are convoluted with nucleon-nucleus optical potentials. Overlaps functions (OFs) for three different neutron states, corresponding to the first three (ground and excited) states in 53Ca, are shown in Figure 9 (right). GF calculations were performed at the ADC(3) level with two different interactions, NNLOsat and NN+3N(lnl). For comparison, OFs from an EDF calculation with the SLy4 Skyrme parameterization are displayed. An overall good agreement is found between the three sets of calculations. Looking more in detail, one notices that NNLOsat better reproduces the maximum of the overlap at around 2–5 fm, i.e., the region of the nuclear surface. This clearly relates to the ability of this interaction to well describe nuclear sizes (cf. discussion in section 4.1). In contrast, NN+3N(lnl) is in better agreement in the tail of the OF. This is in line with the more accurate description of the low-lying spectrum of 53Ca (cf. Figure 6).

In order to compute the cross section (Equation 30) for the 54Ca(p, pn)53Ca reaction, in [51] the GF input was complemented by phenomenological optical potentials and in-medium nucleon-nucleon cross sections, yielding results in good agreement with shell model calculations. An analogous study had been performed previously for the 14O(d, t)13O and 14O(d,3He)13N reactions, leading to similar conclusions [49]. These applications might be seen as a first step toward a consistent approach to structure and reaction, and show that ab initio ingredients can be as efficient as phenomenological ones. In fact, the in-medium nucleon-nucleon cross section could be already extracted from nuclear matter calculations (see e.g., [139]). Moreover, a nucleon-nucleus potential can be directly computed from the one-body self-energy (see [43]), for the first applications to oxygen and calcium isotopes. Thus, in the future the full ab initio calculation of the cross section (30) can be envisaged.

5.2. Toward Doubly Open-Shell Nuclei

The development of Gorkov GF theory [31], and its subsequent implementation to finite nuclei [32, 38] proved that symmetry breaking can be a powerful tool in the context of ab initio calculations. The generalization of other many-body techniques to a symmetry-breaking scheme [13, 96, 140] further confirmed the validity of this strategy. Such advances have allowed to extend the reach of ab initio calculations in mid-mass systems from a few closed-shell nuclei to a large number of open-shell isotopes, e.g., to complete semi-magic isotopic or isotonic chains. In their current implementation, however, these methods do not break rotational symmetry. This results in an inefficient account of quadrupole correlations, such that the description of (doubly) open-shell systems displaying significant deformation can be problematic.

As opposed to pairing, where the strong static correlations at the Fermi surface cause the breakdown of the particle-hole expansion, in the presence of deformation one can usually produce converged calculations, i.e., compute few orders in the many-body expansion. Nevertheless, one expects the accuracy to deteriorate with the strength of the deformation. This has been indeed observed in Gorkov GF calculations around the calcium chain, in particular for titanium and chromium isotopes characterized by mid-shell protons. For instance, by studying neutron gaps18 at the neutron traditional magic number N = 28 one finds an excellent agreement with experiment up to Z = 22, after which symmetry-restricted GF calculations clearly depart from data [102]. Furthermore, one can identify a correlation between the deviation to experimental data and the amount of deformation (e.g., estimated by EDF calculations [141]). This situation suggests that the additional breaking of the SU(2) symmetry associated to rotational invariance will be needed in the extension of correlation-expansion methods to doubly open-shell nuclei. Some of the existing approaches are indeed being generalized along these lines [142]. In the case of Gorkov GF such an extension will presumably require the use of importance truncation [143] and/or tensor factorization techniques [144].

5.3. Toward Heavy Nuclei

Over the past years, GF ab initio calculations have extended their reach across the Segrè chart, going from the first application to the oxygen chain (A ~ 20) [35] to recent computations of nickel isotopes (A ~ 70) [40]. Such calculations rely on sophisticated numerical codes that make extensive use of available high-performance computing resources. Although the management of the computing time19 could be problematic, the bottleneck that currently prevents (converged) calculations beyond A ~ 100 is related to the storage of the matrix elements of 3N operators. Indeed, presently employed truncations on the three-body basis of e3max = 14 − 18 allow keeping the size of 3N matrix elements below 100 GB, which is roughly the order of magnitude of the available memory on a single node of a state-of-the-art supercomputer. Going considerably beyond this size thus constitutes an issue. At the same time, such values of e3max are enough to achieve reasonably converged results in the (A ~ 60 − 70), while they become insufficient for larger isotopes [40, 145].

Different strategies are being explored to overcome this issue. One possibility would be to discard beforehand, for a given e3max, a subset of the initial 3N matrix elements. While performing a selection20 directly on the original set might be problematic, a promising technique based on tensor factorization algorithms has been put forward recently [143]. Since in the majority of applications 3N forces are included in the normal-ordered two-body approximation21, another option could consist in performing the normal ordering procedure in a different (smaller) basis than the HO one, e.g., in momentum space.

Even with current limitations, however, some meaningful results can be produced for nuclei above mass A = 100. Indeed, as discussed in section 3.3, not all observables show the same convergence pattern. In particular, while ground-state energy curves get lower and lower as the basis size is increased, radii tilt around a fixed point that can be assumed to correspond to the infinite-basis result. This allow to provide an estimate of the radius (and, correspondingly, of the density distribution) in bases for which the energy is far from being converged. Based on this observation, charge densities of closed- and open-shell tin and xenon isotopes have been recently calculated within GF theory [41]. Examples are reported in Figure 10, where the NNLOsat interaction has been employed. In the left panel, the charge distribution of several nuclei is shown, exemplifying the typical range of system that can be presently accessed. In the right panel, the charge density of 132Xe is displayed and compared to a two-point Fermi distribution fitted on the recent experimental data from the SCRIT collaboration [83]. The two are in very good agreement at the surface and in the tail of the distribution. In the interior, the two-point Fermi distribution and the lack of high-momentum transfer data lead to flat behavior for the experimental distribution, whereas the computed density shows a well defined oscillation pattern. This example shows that even in present implementations of GF calculations it is possible to provide relevant predictions in the region A = 100 − 150.

FIGURE 10
www.frontiersin.org

Figure 10. (Left) Charge density distributions of selected Sn and Xe isotopes computed with the NNLOsat interaction in Gorkov GF theory at the ADC(2) level. Each curve is shifted upwards with respect to the one below by 0.025 fm−3 for clarity purposes. Colored band account for uncertainties stemming from model-space convergence. (Right) Charge density distributions of 132Xe computed with the NNLOsat interaction in Gorkov GF theory at the ADC(2) level. The dotted line with gray band corresponds to the two-point Fermi distribution with parameter and error bars extracted from [83]. Adapted figure with permission from [41].

Author Contributions

The author confirms being the sole contributor of this work and has approved it for publication.

Conflict of Interest

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

The author wishes to thank F. Raimondi and Y. L. Sun for providing the results appearing in Figure 9, all collaborators that contributed to the results presented in this work, and additionally C. Barbieri and T. Duguet for useful remarks on the manuscript.

Footnotes

1. ^One could further recall the large variety of experimental probes used to study nuclear properties, as well as the diversity of processes that can concern atomic nuclei, for which three of the four fundamental interactions (strong, electromagnetic and weak) are involved.

2. ^This characteristic is usually associated to the definition of nuclei as mesoscopic systems.

3. ^Recently, pioneering lattice QCD calculations have produced the first usable nucleon-nucleon potentials [1, 2]. However, complications still exist in the derivation of three-nucleon forces such that the systematic implementation of lattice QCD interactions is not envisaged in the very near future.

4. ^Let us remark that many-body Green's functions and Green's function Monte Carlo are not the same thing. The latter refers to a (virtually exact) technique that aims at projecting out the ground-state wave function typically from a variational solution of the Schrödinger equation [7, 8].

5. ^One notable exception is the ground-state energy E0A: for a Hamiltonian containing up to y-body operators the knowledge of one- up to (y − 1)-body Green's functions is sufficient (see Equations 5, 6).

6. ^In the case of atomic nuclei proton and neutron numbers are conserved individually, therefore it is always intended U(1)N⊗U(1)Z where one of the two or both are broken.

7. ^Away from closed-shell systems, contributions from components with A′ ≠ A are assumed to cancel out to some extent. The largest contamination is expected in differential observables across closed shells, where one of the two systems does not spontaneously break particle-number symmetry.

8. ^Other types of resummation are employed in other domains. For instance, in solid-state physics the resummation of particle-hole (i.e., ring) diagrams (typically in the so-called GW approximation) allows resolving the long-range features of the Coulomb force [66].

9. ^Since the present formalism is written in single-nucleon coordinates, the appearance of center-of-mass (c.o.m.) contributions might pollute the computation of different quantities/observables and must be taken care of. As for energies, the c.o.m. kinetic energy T^c.o.m. is subtracted from the total Hamiltonian from the outset, such that one eventually works with the intrinsic Hamiltonian ĤintĤ-T^c.o.m.. Although in the present truncation scheme this does not lead to an exact factorization of c.o.m. and intrinsic wave functions, it has been demonstrated in similar correlation-expansion methods [18, 78] that such an approximate c.o.m. correction is sufficient for all practical purposes. Furthermore, specific a posteriori corrections are applied for radii and densities [see discussion in section 4.1 and, e.g., [36] for more details].

10. ^On the other hand, the asymptotic behavior of HO wave functions do not correctly account for the fall-off of the nuclear wave function. As a consequence, a HO basis is not well-suited to describe states near the particle continuum, where the long-range part of the wave function is particularly important. In this case, a possibility consists in complementing the HO with a basis specifically designed to account for resonances and non-resonant continuum, e.g., the Berggren basis [79, 80].

11. ^For example, based on a one-boson exchange model plus phenomenological corrections.

12. ^For a given approximation, a Gorkov calculation is more costly than a Dyson one because of the increased dimensionality of the quasiparticle space. Estimates presented in this section refer to Dyson calculations of typical medium-mass nuclei.

13. ^See also discussion in section 5.3 on storage of 3N matrix elements.

14. ^The SRG evolution described in section 3.4 and standardly applied to nuclear Hamiltonians induces additional many-body operators that need to be taken into account [6]. Hence, one has to distinguish between original and induced e.g., 3N forces.

15. ^Here bold symbols denote matrices in the one-body Hilbert space H1.

16. ^Since the neutron spectral function is not easily accessed by electrons, a complementary study with titanium, which has as many protons (22) as the neutrons in 40Ar, was carried out.

17. ^Since many current experiments concern unstable nuclei, they are performed in inverse kinematics, whence the use of a proton target. In the following, the case of an incident proton will be thus considered.

18. ^Neutron gaps, defined as differences of two-neutron separation energies, are one of possible observables that meaningfully estimate the “magic” character of a neutron number.

19. ^In this respect, while the numerical cost of Dyson GF calculations grows with the mass number, the one of Gorkov GFs solely depend on the basis dimension [32].

20. ^That is, using a different truncation than the e3max truncation.

21. ^It consists of two steps: (i) a normal ordering of the (3N) Hamiltonian with respect to a reference state and (ii) the disregard of operators of rank higher than two (see [146] for a pedagogical description and the generalization to the case of symmetry-breaking reference states).

References

1. Ishii N, Aoki S, Hatsuda T. Nuclear force from lattice QCD. Phys Rev Lett. (2007) 99:022001. doi: 10.1103/PhysRevLett.99.022001

PubMed Abstract | CrossRef Full Text | Google Scholar

2. McIlroy C, Barbieri C, Inoue T, Doi T, Hatsuda T. Doubly magic nuclei from lattice QCD forces at MPS=469 MeV/c2. Phys Rev C. (2018) 97:021303. doi: 10.1103/PhysRevC.97.021303

CrossRef Full Text | Google Scholar

3. Epelbaum E, Hammer HW, Meißner UG. Modern theory of nuclear forces. Rev Mod Phys. (2009) 81:1773. doi: 10.1103/RevModPhys.81.1773

CrossRef Full Text | Google Scholar

4. Machleidt R, Entem D. Chiral effective field theory and nuclear forces. Phys Rep. (2011) 503:1. doi: 10.1016/j.physrep.2011.02.001

CrossRef Full Text | Google Scholar

5. Hammer HW, König S, van Kolck U. Nuclear effective field theory: status and perspectives. Rev Mod Phys. (2020) 92:025004. doi: 10.1103/RevModPhys.92.025004

CrossRef Full Text | Google Scholar

6. Bogner SK, Furnstahl RJ, Schwenk A. From low-momentum interactions to nuclear structure. Prog Part Nucl Phys. (2010) 65:94–147. doi: 10.1016/j.ppnp.2010.03.001

CrossRef Full Text | Google Scholar

7. Carlson J, Gandolfi S, Pederiva F, Pieper SC, Schiavilla R, Schmidt KE, et al. Quantum Monte Carlo methods for nuclear physics. Rev Mod Phys. (2015) 87:1067–118. doi: 10.1103/RevModPhys.87.1067

CrossRef Full Text | Google Scholar

8. Gandolfi S, Lonardoni D, Lovato A, Piarulli M. Atomic nuclei from quantum Monte Carlo calculations with chiral EFT interactions. Front Phys. (2020) 8:117. doi: 10.3389/fphy.2020.00117

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Barrett BR, Navratil P, Vary JP. Ab initio no core shell model. Prog Part Nucl Phys. (2013) 69:131–81. doi: 10.1016/j.ppnp.2012.10.003

CrossRef Full Text | Google Scholar

10. Lähde TA, Epelbaum E, Krebs H, Lee D, Meißner UG, Rupak G. Lattice effective field theory for medium-mass nuclei. Phys Lett B. (2014) 732:110–5. doi: 10.1016/j.physletb.2014.03.023

CrossRef Full Text | Google Scholar

11. Shavitt I, Bartlett RJ. Many-Body Methods in Chemistry and Physics. New York, NY: Cambridge University Press (2009). doi: 10.1017/CBO9780511596834

CrossRef Full Text | Google Scholar

12. Tichai A, Langhammer J, Binder S, Roth R. Hartree-Fock many-body perturbation theory for nuclear ground-states. Phys Lett B. (2016) 756:283–8. doi: 10.1016/j.physletb.2016.03.029

CrossRef Full Text | Google Scholar

13. Tichai A, Arthuis P, Duguet T, Hergert H, Somá V, Roth R. Bogoliubov many-body perturbation theory for open-shell nuclei. Phys Lett B. (2018) 786:195–200. doi: 10.1016/j.physletb.2018.09.044

CrossRef Full Text | Google Scholar

14. Demol P, Duguet T, Ekström A, Frosini M, Hebeler K, König S, et al. Improved many-body expansions from eigenvector continuation. Phys Rev C. (2020) 101:041302. doi: 10.1103/PhysRevC.101.041302

CrossRef Full Text | Google Scholar

15. Demol P, Frosini M, Tichai A, Somá V, Duguet T. Bogoliubov many-body perturbation theory under constraint. (2020) arXiv:2002.02724.

Google Scholar

16. Tichai A, Roth R, Duguet T. Many-body perturbation theories for finite nuclei. Front Phys. (2020) 8:164. doi: 10.3389/fphy.2020.00164

CrossRef Full Text | Google Scholar

17. Tsukiyama K, Bogner SK, Schwenk A. In-medium similarity renormalization group for nuclei. Phys Rev Lett. (2011) 106:222502. doi: 10.1103/PhysRevLett.106.222502

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Hergert H, Bogner SK, Morris TD, Schwenk A, Tsukiyama K. The in-medium similarity renormalization group: a novel ab initio method for nuclei. Phys Rep. (2016) 621:165–222. doi: 10.1016/j.physrep.2015.12.007

CrossRef Full Text | Google Scholar

19. Dickhoff WH, Muther H. Nucleon properties in the nuclear medium. Rept Prog Phys. (1992) 55:1947–2023. doi: 10.1088/0034-4885/55/11/002

CrossRef Full Text | Google Scholar

20. Kowalski K, Dean DJ, Hjorth-Jensen M, Papenbrock T, Piecuch P. Coupled cluster calculations of ground and excited states of nuclei. Phys Rev Lett. (2004) 92:132501. doi: 10.1103/PhysRevLett.92.132501

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Hagen G, Papenbrock T, Hjorth-Jensen M, Dean DJ. Coupled-cluster computations of atomic nuclei. Rep Prog Phys. (2014) 77:096302. doi: 10.1088/0034-4885/77/9/096302

PubMed Abstract | CrossRef Full Text | Google Scholar

22. Bogner SK, Hergert H, Holt JD, Schwenk A, Binder S, Calci A, et al. Nonperturbative shell-model interactions from the in-medium similarity renormalization group. Phys Rev Lett. (2014) 113:142501. doi: 10.1103/PhysRevLett.113.142501

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Jansen GR, Engel J, Hagen G, Navratil P, Signoracci A. Ab initio coupled-cluster effective interactions for the shell model: application to neutron-rich oxygen and carbon isotopes. Phys Rev Lett. (2014) 113:142502. doi: 10.1103/PhysRevLett.113.142502

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Smirnova NA, Barrett BR, Kim Y, Shin IJ, Shirokov AM, Dikmen E, et al. Effective interactions in the sd shell. Phys Rev C. (2019) 100:054329. doi: 10.1103/PhysRevC.100.054329

CrossRef Full Text | Google Scholar

25. Broglia RA, Bortignon PF, Barranco F, Vigezzi E, Idini A, Potel G. Unified description of structure and reactions: implementing the nuclear field theory program. Phys Scripta. (2016) 91:063012. doi: 10.1088/0031-8949/91/6/063012

CrossRef Full Text | Google Scholar

26. Barbieri C, Dickhoff WH. Faddeev description of two-hole-one-particle motion and the single-particle spectral function. Phys Rev C. (2001) 63:034313. doi: 10.1103/PhysRevC.63.034313

CrossRef Full Text | Google Scholar

27. Barbieri C, Dickhoff WH. Faddeev treatment of long-range correlations and the one-hole spectral function of 16O. Phys Rev C. (2002) 65:064313. doi: 10.1103/PhysRevC.65.064313

CrossRef Full Text | Google Scholar

28. Dickhoff WH, Barbieri C. Self-consistent Green's function method for nuclei and nuclear matter. Prog Part Nucl Phys. (2004) 52:377. doi: 10.1016/j.ppnp.2004.02.038

CrossRef Full Text | Google Scholar

29. Barbieri C, Hjorth-Jensen M. Quasiparticle and quasihole states of nuclei around 56Ni. Phys Rev C. (2009) 79:064313. doi: 10.1103/PhysRevC.79.064313

CrossRef Full Text | Google Scholar

30. Barbieri C. Role of long-range correlations in the quenching of spectroscopic factors. Phys Rev Lett. (2009) 103:202502. doi: 10.1103/PhysRevLett.103.202502

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Somá V, Duguet T, Barbieri C. Ab initio self-consistent Gorkov-Green's function calculations of semimagic nuclei: formalism at second order with a two-nucleon interaction. Phys Rev C. (2011) 84:064317. doi: 10.1103/PhysRevC.84.064317

CrossRef Full Text | Google Scholar

32. Somá V, Barbieri C, Duguet T. Ab initio self-consistent Gorkov-Green's function calculations of semi-magic nuclei: numerical implementation at second order with a two-nucleon interaction. Phys Rev C. (2014) 89:024323. doi: 10.1103/PhysRevC.89.024323

CrossRef Full Text | Google Scholar

33. Carbone A, Cipollone A, Barbieri C, Rios A, Polls A. Self-consistent Green's functions formalism with three-body interactions. Phys Rev C. (2013) 88:054326. doi: 10.1103/PhysRevC.88.054326

CrossRef Full Text | Google Scholar

34. Raimondi F, Barbieri C. Algebraic diagrammatic construction formalism with three-body interactions. Phys Rev C. (2018) 97:054308. doi: 10.1103/PhysRevC.97.054308

CrossRef Full Text | Google Scholar

35. Cipollone A, Barbieri C, Navrátil P. Isotopic chains around oxygen from evolved chiral two- and three-nucleon interactions. Phys Rev Lett. (2013) 111:062501. doi: 10.1103/PhysRevLett.111.062501

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Cipollone A, Barbieri C, Navrátil P. Chiral three-nucleon forces and the evolution of correlations along the oxygen isotopic chain. Phys Rev C. (2015) 92:014306. doi: 10.1103/PhysRevC.92.014306

CrossRef Full Text | Google Scholar

37. Duguet T, Somá V, Lecluse S, Barbieri C, Navrátil P. Ab initio calculation of the potential bubble nucleus 34Si. Phys Rev C. (2017) 95:034319. doi: 10.1103/PhysRevC.95.034319

CrossRef Full Text | Google Scholar

38. Somá V, Barbieri C, Duguet T. Ab initio Gorkov-Green's function calculations of open-shell nuclei. Phys Rev C. (2013) 87:011303. doi: 10.1103/PhysRevC.87.011303

CrossRef Full Text | Google Scholar

39. Somá V, Cipollone A, Barbieri C, Navrátil P, Duguet T. Chiral two- and three-nucleon forces along medium-mass isotope chains. Phys Rev C. (2014) 89:061301. doi: 10.1103/PhysRevC.89.061301

CrossRef Full Text | Google Scholar

40. Somá V, Navrátil P, Raimondi F, Barbieri C, Duguet T. Novel chiral Hamiltonian and observables in light and medium-mass nuclei. Phys Rev C. (2020) 101:014318. doi: 10.1103/PhysRevC.101.014318

CrossRef Full Text | Google Scholar

41. Arthuis P, Barbieri C, Vorabbi M, Finelli P. Ab initio computation of charge densities for Sn and Xe isotopes. (2020) arXiv:2002.02214.

Google Scholar

42. Raimondi F, Barbieri C. Nuclear electromagnetic dipole response with the self-consistent Green's function formalism. Phys Rev C. (2019) 99:054327. doi: 10.1103/PhysRevC.99.054327

CrossRef Full Text | Google Scholar

43. Idini A, Barbieri C, Navrátil P. Ab initio optical potentials and nucleon scattering on medium mass nuclei. Phys Rev Lett. (2019) 123:092501. doi: 10.1103/PhysRevLett.123.092501

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Barbieri C, Rocco N, Somá V. Lepton scattering from 40Ar and 48Ti in the quasielastic peak region. Phys Rev C. (2019) 100:062501. doi: 10.1103/PhysRevC.100.062501

CrossRef Full Text | Google Scholar

45. Raimondi F, Barbieri C. Core-polarization effects and effective charges in O and Ni isotopes from chiral interactions. Phys Rev C. (2019) 100:024317. doi: 10.1103/PhysRevC.100.024317

CrossRef Full Text | Google Scholar

46. Duguet T, Hergert H, Holt JD, Somá V. Nonobservable nature of the nuclear shell structure: meaning, illustrations, and consequences. Phys Rev C. (2015) 92:034313. doi: 10.1103/PhysRevC.92.034313

CrossRef Full Text | Google Scholar

47. Rosenbusch M, Ascher P, Atanasov D, Barbieri C, Beck D, Blaum K, et al. Probing the N = 32 shell closure below the magic proton number Z = 20: mass measurements of the exotic isotopes 52, 53K. Phys Rev Lett. (2015) 114:202501. doi: 10.1103/PhysRevLett.114.202501

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Leistenschneider E, Reiter MP, Ayet San Andrés S, Kootte B, Holt JD, Navrátil P, et al. Dawning of the N = 32 shell closure seen through precision mass measurements of neutron-rich titanium isotopes. Phys Rev Lett. (2018) 120:062503. doi: 10.1103/PhysRevLett.120.062503

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Flavigny F, Gillibert A, Nalpas L, Obertelli A, Keeley N, Barbieri C, et al. Limited asymmetry dependence of correlations from single nucleon transfer. Phys Rev Lett. (2013) 110:122503. doi: 10.1103/PhysRevLett.110.122503

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Papuga J, Bissell ML, Kreim K, Barbieri C, Blaum K, Rydt MD, et al. Shell structure of potassium isotopes deduced from their magnetic moments. Phys Rev C. (2014) 90:034321. doi: 10.1103/PhysRevC.90.034321

CrossRef Full Text | Google Scholar

51. Chen S, Lee J, Doornenbal P, Obertelli A, Barbieri C, Chazono Y, et al. Quasifree neutron knockout from 54Ca corroborates arising N = 34 neutron magic number. Phys Rev Lett. (2019) 123:142501. doi: 10.1103/PhysRevLett.123.142501

PubMed Abstract | CrossRef Full Text | Google Scholar

52. Sun YL, et al. Restoration of the natural E(1/21+)-E(3/21+) energy splitting in odd-K isotopes towards N = 40. Phys Lett B. (2020) 802:135215. doi: 10.1016/j.physletb.2020.135215

CrossRef Full Text | Google Scholar

53. Lapoux V, Somá V, Barbieri C, Hergert H, Holt JD, Stroberg SR. Radii and binding energies in oxygen isotopes: a challenge for nuclear forces. Phys Rev Lett. (2016) 117:052501. doi: 10.1103/PhysRevLett.117.052501

PubMed Abstract | CrossRef Full Text | Google Scholar

54. Dickhoff WH, Van Neck D. Many-Body Theory Exposed! Singapore: World Scientific (2005). doi: 10.1142/5804

CrossRef Full Text | Google Scholar

55. Noziéres P. Theory of Interacting Fermi Sytems. Boulder: Westview Press (1964).

56. Abrikosov AA, Gorkov LP, Dzyaloshinski IE. Methods of Quantum Field Theory in Statistical Physics. Lowell, MA: Courier Corporation (1975).

Google Scholar

57. Economou EN. Green's Functions in Quantum Physics. Berlin: Springer (2006). doi: 10.1007/3-540-28841-4

CrossRef Full Text | Google Scholar

58. Barbieri C, Carbone A. Self-consistent green's function approaches. In: Hjorth-Jensen M, Lombardo M, van Kolck U, editors. An Advanced Course in Computational Nuclear Physics. Vol. 936 of Lecture Notes in Physics. Cham: Springer (2017). p. 571–644.

PubMed Abstract | Google Scholar

59. Barbieri C. Computational Many-Body Physics. Available online at: http://personal.ph.surrey.ac.uk/~cb0023/bcdor/bcdor/Comp_Many-Body_Phys.html

PubMed Abstract | Google Scholar

60. Galitskii VM, Migdal AB. Applications of quantum field theory methods to the many body problem. Z Eksp Teor Fiz. (1958) 34:139.

Google Scholar

61. Koltun DS. Total binding energies of nuclei, and particle-removal experiments. Phys Rev Lett. (1972) 28:182. doi: 10.1103/PhysRevLett.28.182

CrossRef Full Text | Google Scholar

62. Gorkov LP. On the energy spectrum of superconductors. Sov Phys JETP. (1958) 34:505. doi: 10.2307/2607987

CrossRef Full Text | Google Scholar

63. Nambu Y. Quasi-particles and gauge invariance in the theory of superconductivity. Phys Rev. (1960) 117:648. doi: 10.1103/PhysRev.117.648

CrossRef Full Text | Google Scholar

64. Duguet T. Symmetry broken and restored coupled-cluster theory: I. Rotational symmetry and angular momentum. J Phys G. (2015) 42:025107. doi: 10.1088/0954-3899/42/2/025107

CrossRef Full Text | Google Scholar

65. Duguet T, Signoracci A. Symmetry broken and restored coupled-cluster theory: II. Global gauge symmetry and particle number. J Phys G. (2017) 44:015103. doi: 10.1088/0954-3899/44/1/015103

CrossRef Full Text | Google Scholar

66. Onida G, Reining L, Rubio A. Electronic excitations: density-functional versus many-body Green's-function approaches. Rev Mod Phys. (2002) 74:601–59. doi: 10.1103/RevModPhys.74.601

CrossRef Full Text | Google Scholar

67. Bożek P. Self-consistent solution of Galitskii-Feynman equations at finite temperature. Phys Rev C. (1999) 59:2619–26. doi: 10.1103/PhysRevC.59.2619

CrossRef Full Text | Google Scholar

68. Frick T, Müther H. Self-consistent solution to the nuclear many-body problem at finite temperature. Phys Rev C. (2003) 68:034310. doi: 10.1103/PhysRevC.68.034310

CrossRef Full Text | Google Scholar

69. Somá V, Bożek P. Diagrammatic calculation of thermodynamical quantities in nuclear matter. Phys Rev C. (2006) 74:045809. doi: 10.1103/PhysRevC.74.045809

CrossRef Full Text | Google Scholar

70. Rios A, Polls A, Ramos A, Müther H. Entropy of a correlated system of nucleons. Phys Rev C. (2006) 74:054317. doi: 10.1103/PhysRevC.74.054317

CrossRef Full Text | Google Scholar

71. Rios A, Polls A, Ramos A, Müther H. Liquid-gas phase transition in nuclear matter from realistic many-body approaches. Phys Rev C. (2008) 78:044314. doi: 10.1103/PhysRevC.78.044314

CrossRef Full Text | Google Scholar

72. Somá V, Bożek P. In-medium T-matrix for nuclear matter with three-body forces: binding energy and single-particle properties. Phys Rev C. (2008) 78:054003. doi: 10.1103/PhysRevC.78.054003

CrossRef Full Text | Google Scholar

73. Schirmer J. Beyond the random-phase approximation: a new approximation scheme for the polarization propagator. Phys Rev A. (1982) 26:2395–416. doi: 10.1103/PhysRevA.26.2395

CrossRef Full Text | Google Scholar

74. Schirmer J, Cederbaum LS, Walter O. New approach to the one-particle Green's function for finite Fermi systems. Phys Rev A. (1983) 28:1237–59. doi: 10.1103/PhysRevA.28.1237

CrossRef Full Text | Google Scholar

75. Nooijen M, Snijders JG. Coupled cluster approach to the single-particle Green's function. Int J Quant Chem. (1992) 26:55. doi: 10.1002/qua.560440808

CrossRef Full Text | Google Scholar

76. Schirmer J, Angonoa G. On Green's function calculations of the static self-energy part, the ground state energy and expectation values. J Chem Phys. (1989) 91:1754. doi: 10.1063/1.457081

CrossRef Full Text | Google Scholar

77. Dewulf Y, Neck DV, Daele LV, Waroquier M. Long-range correlations in finite nuclei: comparison of two self-consistent treatments. Phys Lett B. (1997) 396:7–14. doi: 10.1016/S0370-2693(97)00135-4

CrossRef Full Text | Google Scholar

78. Hagen G, Papenbrock T, Dean DJ. Solution of the center-of-mass problem in nuclear structure calculations. Phys Rev Lett. (2009) 103:062503. doi: 10.1103/PhysRevLett.103.062503

PubMed Abstract | CrossRef Full Text | Google Scholar

79. Berggren T. On the use of resonant states in eigenfunction expansions of scattering and reaction amplitudes. Nucl Phys A. (1968) 109:265–87. doi: 10.1016/0375-9474(68)90593-9

CrossRef Full Text | Google Scholar

80. Berggren T. On the treatment of resonant final states in direct reactions. Nucl Phys A. (1971) 169:353–62. doi: 10.1016/0375-9474(71)90889-X

CrossRef Full Text | Google Scholar

81. Dai H, Murphy M, Pandey V, Abrams D, Nguyen D, Aljawrneh B, et al. First measurement of the Ti(e,′)X cross section at Jefferson Lab. Phys Rev C. (2018) 98:014617. doi: 10.1103/PhysRevC.98.014617

CrossRef Full Text | Google Scholar

82. Dai H, Murphy M, Pandey V, Abrams D, Nguyen D, Aljawrneh B, et al. First measurement of the Ar(e, e′)X cross section at Jefferson Laboratory. Phys Rev C. (2019) 99:054608. doi: 10.1103/PhysRevC.99.054608

CrossRef Full Text | Google Scholar

83. Tsukada K, Enokizono A, Ohnishi T, Adachi K, Fujita T, Hara M, et al. First elastic electron scattering from 132Xe at the SCRIT facility. Phys Rev Lett. (2017) 118:262501. doi: 10.1103/PhysRevLett.118.262501

PubMed Abstract | CrossRef Full Text | Google Scholar

84. Machleidt R. The high-precision, charge-dependent Bonn nucleon-nucleon potential (CD-Bonn). Phys Rev C. (2001) 63:024001. doi: 10.1103/PhysRevC.63.024001

CrossRef Full Text | Google Scholar

85. Stoks VGJ, Klomp RAM, Terheggen CPF, de Swart JJ. Construction of high quality NN potential models. Phys Rev C. (1994) 49:2950. doi: 10.1103/PhysRevC.49.2950

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Somá V, Bożek P. Thermodynamic properties of nuclear matter with three-body forces. Phys Rev C. (2009) 80:025803. doi: 10.1103/PhysRevC.80.025803

CrossRef Full Text | Google Scholar

87. Epelbaum E, Krebs H, Meißner UG. Improved chiral nucleon-nucleon potential up to next-to-next-to-next-to-leading order. Eur Phys J A. (2015) 51:53. doi: 10.1140/epja/i2015-15053-8

CrossRef Full Text | Google Scholar

88. Binder S, Calci A, Epelbaum E, Furnstahl RJ, Golak J, Hebeler K, et al. Few-nucleon and many-nucleon systems with semilocal coordinate-space regularized chiral nucleon-nucleon forces. Phys Rev C. (2018) 98:014002. doi: 10.1103/PhysRevC.98.014002

CrossRef Full Text | Google Scholar

89. Drischler C, Hebeler K, Schwenk A. Chiral interactions up to next-to-next-to-next-to-leading order and nuclear saturation. Phys Rev Lett. (2019) 122:042501. doi: 10.1103/PhysRevLett.122.042501

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Hüther T, Vobig K, Hebeler K, Machleidt R, Roth R. Family of chiral two- plus three-nucleon interactions for accurate nuclear structure studies. (2019) arXiv:1911.04955. doi: 10.1016/j.physletb.2020.135651

CrossRef Full Text | Google Scholar

91. Epelbaum E, Golak J, Hebeler K, Hüther T, Kamada H, Krebs H, et al. Few- and many-nucleon systems with semilocal coordinate-space regularized chiral two- and three-body forces. Phys Rev C. (2019) 99:024313. doi: 10.1103/PhysRevC.99.024313

CrossRef Full Text | Google Scholar

92. Epelbaum E, Krebs H, Reinert P. High-precision nuclear forces from chiral EFT: state-of-the-art, challenges and outlook. Front Phys. (2020) 8:98. doi: 10.3389/fphy.2020.00098

CrossRef Full Text | Google Scholar

93. Entem DR, Machleidt R. Accurate charge-dependent nucleon-nucleon potential at fourth order of chiral perturbation theory. Phys Rev C. (2003) 68:041001. doi: 10.1103/PhysRevC.68.041001

CrossRef Full Text | Google Scholar

94. Navrátil P. Local three-nucleon interaction from chiral effective field theory. Few-Body Syst. (2007) 41:117–40. doi: 10.1007/s00601-007-0193-3

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Roth R, Binder S, Vobig K, Calci A, Langhammer J, Navrátil P. Medium-mass nuclei with normal-ordered chiral NN+3N interactions. Phys Rev Lett. (2012) 109:052501. doi: 10.1103/PhysRevLett.109.052501

PubMed Abstract | CrossRef Full Text | Google Scholar

96. Hergert H, Binder S, Calci A, Langhammer J, Roth R. Ab initio calculations of even oxygen isotopes with chiral two- plus three-nucleon interactions. Phys Rev Lett. (2013) 110:242501. doi: 10.1103/PhysRevLett.110.242501

PubMed Abstract | CrossRef Full Text | Google Scholar

97. Binder S, Langhammer J, Calci A, Roth R. Ab initio path to heavy nuclei. Phys Lett B. (2014) 736:119–23. doi: 10.1016/j.physletb.2014.07.010

CrossRef Full Text | Google Scholar

98. Ekström A, Jansen GR, Wendt KA, Hagen G, Papenbrock T, Carlsson BD, et al. Accurate nuclear radii and binding energies from a chiral interaction. Phys Rev C. (2015) 91:051301. doi: 10.1103/PhysRevC.91.051301

CrossRef Full Text | Google Scholar

99. Otsuka T, Suzuki T, Holt JD, Schwenk A, Akaishi Y. Three-body forces and the limit of oxygen isotopes. Phys Rev Lett. (2010) 105:032501. doi: 10.1103/PhysRevLett.105.032501

PubMed Abstract | CrossRef Full Text | Google Scholar

100. Epelbaum E, Krebs H, Lähde TA, Lee D, Meißner UG, Rupak G. Ab initio calculation of the spectrum and structure of 16O. Phys Rev Lett. (2014) 112:102501. doi: 10.1103/PhysRevLett.112.102501

PubMed Abstract | CrossRef Full Text | Google Scholar

101. Hebeler K, Holt JD, Menéndez J, Schwenk A. Nuclear forces and their impact on neutron-rich nuclei and neutron-rich matter. Annu Rev Nucl Part Sci. (2015) 65:457–84. doi: 10.1146/annurev-nucl-102313-025446

CrossRef Full Text | Google Scholar

102. Mougeot M, Atanasov D, Barbieri C, Blaum K, Breitenfeld M, de Roubin A, et al. Examining the N = 28 shell closure through high-precision mass measurements of 46−48Ar. Phys Rev C. (2020) 102:014301. doi: 10.1103/PhysRevC.102.014301

CrossRef Full Text | Google Scholar

103. Brown BA, Massen SE, Hodgson PE. The charge distributions of the oxygen and calcium isotopes. Phys Lett B. (1979) 85:167–71. doi: 10.1016/0370-2693(79)90569-0

CrossRef Full Text | Google Scholar

104. Negele JW. Structure of finite nuclei in the local-density approximation. Phys Rev C. (1970) 1:1260–321. doi: 10.1103/PhysRevC.1.1260

CrossRef Full Text | Google Scholar

105. Chandra H, Sauer G. Relativistic corrections to the elastic electron scattering from 208Pb. Phys Rev C. (1976) 13:245–52. doi: 10.1103/PhysRevC.13.245

CrossRef Full Text | Google Scholar

106. Bertozzi W, Friar J, Heisenberg J, Negele JW. Contributions of neutrons to elastic electron scattering from nuclei. Phys Lett B. (1972) 41:408–14. doi: 10.1016/0370-2693(72)90662-4

CrossRef Full Text | Google Scholar

107. Richter WA, Brown BA. Nuclear charge densities with the Skyrme Hartree-Fock method. Phys Rev C. (2003) 67:034317. doi: 10.1103/PhysRevC.67.034317

CrossRef Full Text | Google Scholar

108. Todd-Rutel BG, Piekarewicz J, Cottle PD. Spin orbit splitting in low j neutron orbits and proton densities in the nuclear interior. Phys Rev C. (2004) 69:021301. doi: 10.1103/PhysRevC.69.021301

CrossRef Full Text | Google Scholar

109. Khan E, Grasso M, Margueron J, Van Giai N. Detecting bubbles in exotic nuclei. Nucl Phys A. (2008) 800:37–46. doi: 10.1016/j.nuclphysa.2007.11.012

CrossRef Full Text | Google Scholar

110. Grasso M, Gaudefroy L, Khan E, Niksic T, Piekarewicz J, Sorlin O, et al. Nuclear ‘bubble’ structure in Si-34. Phys Rev C. (2009) 79:034318. doi: 10.1103/PhysRevC.79.034318

CrossRef Full Text | Google Scholar

111. Yao JM, Baroni S, Bender M, Heenen PH. Beyond-mean-field study of the possible “bubble” structure of 34Si. Phys Rev C. (2012) 86:014310. doi: 10.1103/PhysRevC.86.014310

CrossRef Full Text | Google Scholar

112. Yao JM, Mei H, Li ZP. Does a proton “bubble” structure exist in the low-lying states of 34Si? Phys Lett B. (2013) 723:459–63. doi: 10.1016/j.physletb.2013.05.049

CrossRef Full Text | Google Scholar

113. Wu XY, Yao JM, Li ZP. Low-energy structure and anti-bubble effect of dynamical correlations in 46Ar. Phys Rev C. (2014) 89:017304. doi: 10.1103/PhysRevC.89.017304

CrossRef Full Text | Google Scholar

114. Rychel D, Emrich HJ, Miska H, Gyufko R, Wiedner CA. Charge distribution of the neven sulphur isotopes from elastic electron scattering. Phys Lett. (1983) 130B:5–8. doi: 10.1016/0370-2693(83)91051-1

CrossRef Full Text | Google Scholar

115. Suda T, Simon H. Prospects for electron scattering on unstable, exotic nuclei. Prog Part Nucl Phys. (2017) 96:1–31. doi: 10.1016/j.ppnp.2017.04.002

CrossRef Full Text | Google Scholar

116. von Niessen W, Schirmer J, Cederbaum LS. Computational methods for the one-particle Green's function. Comput Phys Rep. (1984) 1:57–125. doi: 10.1016/0167-7977(84)90002-9

CrossRef Full Text | Google Scholar

117. Danovich D. Green's function methods for calculating ionization potentials, electron affinities, and excitation energies. Wiley Interdisc Rev. (2011) 1:377–87. doi: 10.1002/wcms.38

CrossRef Full Text | Google Scholar

118. Degroote M, Van Neck D, Barbieri C. Faddeev random-phase approximation for molecules. Phys Rev A. (2011) 83:042517. doi: 10.1103/PhysRevA.83.042517

CrossRef Full Text | Google Scholar

119. Barbieri C, Van Neck D, Degroote M. Accuracy of the Faddeev random phase approximation for light atoms. Phys Rev A. (2012) 85:012501. doi: 10.1103/PhysRevA.85.012501

CrossRef Full Text | Google Scholar

120. Papuga J, Bissell ML, Kreim K, Blaum K, Brown BA, De Rydt M, et al. Spins and magnetic moments of 49K and 51K: establishing the 1/2+ and 3/2+ level ordering beyond N=28. Phys Rev Lett. (2013) 110:172503. doi: 10.1103/PhysRevLett.110.172503

PubMed Abstract | CrossRef Full Text | Google Scholar

121. Barbieri C, Raimondi F, McIlroy C. Recent applications of self-consistent Green's function theory to nuclei. J Phys. (2018) 966:012015. doi: 10.1088/1742-6596/966/1/012015

CrossRef Full Text | Google Scholar

122. Ahrens J, Borchert H, Czock KH, Eppler HB, Gimm H, Gundrum H, et al. Total nuclear photon absorption cross sections for some light elements. Nucl Phys A. (1975) 251:479–92. doi: 10.1016/0375-9474(75)90543-6

CrossRef Full Text | Google Scholar

123. Ishkhanov B, Kapitonov I, Lileeva E, Shirokov E, Erokhova V, Elkin M, et al. Cross Sections of Photon Absorption by Nuclei With Nucleon Numbers. Moscow: Institute of Nuclear Physics, Moscow State University (2002).

124. Birkhan J, Miorelli M, Bacca S, Bassauer S, Bertulani CA, Hagen G, et al. Electric dipole polarizability of 48Ca and implications for the neutron skin. Phys Rev Lett. (2017) 118:252501. doi: 10.1103/PhysRevLett.118.252501

PubMed Abstract | CrossRef Full Text | Google Scholar

125. Boffi S, Giusti C, Pacati Fd, Radici M. Electromagnetic Response of Atomic Nuclei. Vol. 20 of Oxford Studies in Nuclear Physics. Oxford: Clarendon Press (1996).

Google Scholar

126. Herczeg P, Hoffman CM, Klapdor-Kleingrothaus HV. Physics Beyond the Standard Model. Singapore: World Scientific (1999). doi: 10.1142/9789814527514

PubMed Abstract | CrossRef Full Text | Google Scholar

127. Nakamura K, et al. Review of particle physics. J Phys G. (2010) 37:075021. doi: 10.1088/0954-3899/37/7A/075021

CrossRef Full Text | Google Scholar

128. Rocco N, Barbieri C. Inclusive electron-nucleus cross section within the self-consistent Green's function approach. Phys Rev C. (2018) 98:025501. doi: 10.1103/PhysRevC.98.025501

CrossRef Full Text | Google Scholar

129. Rocco N, Barbieri C, Benhar O, De Pace A, Lovato A. Neutrino-nucleus cross section within the extended factorization scheme. Phys Rev C. (2019) 99:025502. doi: 10.1103/PhysRevC.99.025502

CrossRef Full Text | Google Scholar

130. Giusti C, Pacati FD, Schwamb M, Boffi S. On the treatment of the Δ-contributions in electromagnetic pp-knockout reactions. Eur Phys J A. (2005) 26:209–20. doi: 10.1140/epja/i2005-10162-7

CrossRef Full Text | Google Scholar

131. Barbieri C, Giusti C, Pacati FD, Dickhoff WH. Effects of nuclear correlations on the 16O(e, epN) reactions to discrete final states. Phys Rev C. (2004) 70:014606. doi: 10.1103/PhysRevC.70.014606

CrossRef Full Text | Google Scholar

132. Barbieri C. Final state interactions in electron scattering at high missing energies and momenta. Nucl Phys B. (2006) 159:174–9. doi: 10.1016/j.nuclphysbps.2006.08.035

CrossRef Full Text | Google Scholar

133. Acciarri R, et al. Long-baseline neutrino facility (LBNF) and deep underground neutrino experiment (DUNE). (2015) arXiv:1601.05471. doi: 10.2172/1250878

CrossRef Full Text | Google Scholar

134. Benhar O, et al. Measurement of the spectral function of 40Ar through the (e,e'p) reaction. (2014).

Google Scholar

135. Aumann T, Bertulani CA, Ryckebusch J. Quasifree (p,2p) and (p,pn) reactions with unstable nuclei. Phys Rev C. (2013) 88:064610. doi: 10.1103/PhysRevC.88.064610

CrossRef Full Text | Google Scholar

136. Raimondi F. Private communication (2019).

137. Bennaceur K, Dobaczewski J. Comp Phys Comm. (2005) 168:96. doi: 10.1016/j.cpc.2005.02.002

CrossRef Full Text

138. Sun YL (2019). Private communication.

139. Rios A, Somá V. Self-consistent Green's function calculation of the nucleon mean free path. Phys Rev Lett. (2012) 108:012501. doi: 10.1103/PhysRevLett.108.012501

PubMed Abstract | CrossRef Full Text | Google Scholar

140. Signoracci A, Duguet T, Hagen G, Jansen G. Ab initio Bogoliubov coupled cluster theory for open-shell nuclei. Phys Rev C. (2015) 91:064320. doi: 10.1103/PhysRevC.91.064320

CrossRef Full Text | Google Scholar

141. Bender M, Bertsch GF, Heenen PH. Global study of quadrupole correlation effects. Phys Rev C. (2006) 73:034322. doi: 10.1103/PhysRevC.73.034322

CrossRef Full Text | Google Scholar

142. Yao JM, Bally B, Engel J, Wirth R, Rodríguez TR, Hergert H. Ab initio treatment of collective correlations and the neutrinoless double beta decay of 48Ca. (2019) Phys Rev Lett. 124:232501. doi: 10.1103/PhysRevLett.124.232501

PubMed Abstract | CrossRef Full Text | Google Scholar

143. Tichai A, Ripoche J, Duguet T. Pre-processing the nuclear many-body problem: importance truncation versus tensor factorization techniques. Eur Phys J A. (2019) 55:90. doi: 10.1140/epja/i2019-12758-6

CrossRef Full Text | Google Scholar

144. Tichai A, Schutski R, Scuseria GE, Duguet T. Tensor-decomposition techniques for ab initio nuclear structure calculations. From chiral nuclear potentials to ground-state energies. Phys Rev C. (2019) 99:034320. doi: 10.1103/PhysRevC.99.034320

CrossRef Full Text | Google Scholar

145. Hagen G, Jansen GR, Papenbrock T. Structure of 78Ni from first-principles computations. Phys Rev Lett. (2016) 117:172501. doi: 10.1103/PhysRevLett.117.172501

PubMed Abstract | CrossRef Full Text | Google Scholar

146. Ripoche J, Tichai A, Duguet T. Normal-ordered k-body approximation in particle-number-breaking theories. Eur Phys J A. (2020) 56:40. doi: 10.1140/epja/s10050-020-00045-8

CrossRef Full Text | Google Scholar

Keywords: nuclear theory, many-body theory, ab initio nuclear structure, Green's function theory, open-shell nuclei

Citation: Somà V (2020) Self-Consistent Green's Function Theory for Atomic Nuclei. Front. Phys. 8:340. doi: 10.3389/fphy.2020.00340

Received: 18 March 2020; Accepted: 22 July 2020;
Published: 24 September 2020.

Edited by:

Saori Pastore, Washington University in St. Louis, United States

Reviewed by:

Pieter Maris, Iowa State University, United States
Mack Atkinson, TRIUMF, Canada

Copyright © 2020 Somà. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Vittorio Somà, vittorio.soma@cea.fr

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.