Next Article in Journal
The Relation between General Relativity’s Metrics and Special Relativity’s Gravitational Scalar Generalized Potentials and Case Studies on the Schwarzschild Metric, Teleparallel Gravity, and Newtonian Potential
Previous Article in Journal
Azimuthal Correlations of D Mesons with Charged Particles in Simulations with the ALICE Experiment
Previous Article in Special Issue
Radiative Corrections to Semileptonic Beta Decays: Progress and Challenges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Beta Decay in Medium-Mass Nuclei with the In-Medium Similarity Renormalization Group

by
Steven Ragnar Stroberg
Physics Division, Argonne National Laboratory, Lemont, IL 60439, USA
Submission received: 6 September 2021 / Revised: 19 October 2021 / Accepted: 27 October 2021 / Published: 18 November 2021
(This article belongs to the Special Issue Beta-Decay Processes in Nuclear Systems)

Abstract

:
We review the status of ab initio calculations of allowed beta decays (both Fermi and Gamow–Teller), within the framework of the valence-space in-medium similarity renormalization group approach.

1. Introduction

Beta decays in atomic nuclei have long been a source of fundamental discoveries in physics [1,2,3], and precise measurements of beta decays continue to be a promising path to search for physics beyond the Standard Model (BSM) [4,5,6,7]. A major challenge in the search for a signal of new physics is understanding the Standard Model “background”, especially the effects of low-energy quantum chromodynamics which manifest as nuclear structure. The situation is aggravated by the fact that nuclei which are preferred experimentally are often difficult to treat theoretically in a framework that allows quantified uncertainties.
Nevertheless, progress has been made over the past few decades so that the internucleon interaction can be systematically constructed within an effective field theory framework [8,9,10]. Simultaneously, advances in many-body theory and computational resources have enabled ab initio treatment of the medium-mass nuclei which are often relevant for BSM searches [11,12,13,14,15,16,17]. Of course, more work remains to be done, both on the effective field theory side [18,19,20] and on understanding how approximation schemes in ab initio calculations impact the observables in question.
In this paper, I will focus on one particular many-body method—the valence-space in-medium similarity renormalization group (VS-IMSRG)—and consider two topics in allowed beta decay, the quenching in Gamow–Teller decays and correction factors for superallowed 0 + 0 + Fermi decays.

2. IMSRG Formalism

There are several review articles detailing both the free-space SRG [21,22] and the in-medium SRG [15,23,24,25,26], and so here I will review only what is needed for our present purposes.

2.1. Similarity Renormalization Group

The basic idea of the SRG is to perform a unitary transformation U on the Hamiltonian H (and all other operators) in such a way that the resulting nuclear wave function is simpler. This is achieved by performing a sequence of infinitessimal unitary transformations, labeled by a flow parameter s, so that
H ( s ) = U ( s ) H ( 0 ) U ( s )
with U ( 0 ) = 1 . The way in which U changes with s is specified by the action of an operator η , called the generator:
d U ( s ) d s = η ( s ) U ( s ) .
We are free to choose η however we like—it can depend on s—as long as it is anti-hermitian, i.e., η ( s ) = η ( s ) . Combining (1) and (2) we obtain a flow equation for the Hamiltonian in terms of a commutator with the generator
d H d s = [ η ( s ) , H ( s ) ] .
The flow equation for any other operator O is obtained by replacing H O on both sides of (3) [27,28,29,30].
It remains to specify η ( s ) . In the free-space SRG, we choose
η S R G ( s ) = [ T , V ( s ) ]
where T is the kinetic energy and V ( s ) is the potential so H ( s ) = T + V ( s ) . This generator drives V ( s ) towards a band-diagonal form in momentum space, with a width λ S R G s 1 / 4 .
When the SRG flow Equation (3) is formulated in Fock-space (i.e., in terms of creation and annihilation operators), many-body forces are inevitably induced, and these must be truncated in order to make the calculation tractable. For this reason, the “free-space” SRG evolution is typically performed out to λ S R G 2 fm 1 .

2.2. In-Medium SRG

The truncation of many-body forces is rendered less severe if all operators are normal-ordered with respect to a reference state | Φ , which should be a reasonable first approximation of the exact wave function | Ψ . This approach is called the in-medium SRG (IMSRG). If we choose the generator to suppress the parts of H which lead to excitations out of the reference | Φ , then for s | Φ becomes an eigenstate of H ( s ) with an eigenvalue corresponding to the energy of the exact wave function | Ψ , up to approximation errors in solving (3). In all calculations presented here, I neglect three-body operators after the initial normal-ordering step, resulting in the IMSRG(2) approximation.
One possible choice for the generator which achieves the desired suppression was proposed by White [31]
a | η Wh | b a | H od | b E a E b
where the “off-diagonal” part of the Hamiltonian, denoted H od , is any part of H which connects | Φ to a different state. The states | a and | b are elements of the basis used to express the operators, and the denominator E a E b is the energy difference between the basis states. (The denominators are typically defined with Epstein–Nesbet or Møller–Plesset partitioning. In this work, I use Epstein–Nesbet denominators, and a modification of (5)—also suggested by White—called the arctangent generator (see Ref. [26] for more details).)
As a further generalization, we may define a valence space (e.g., the sd-shell above an 16O core), and define H od as any part of H which connects a valence configuration to a non-valence configuration (specifically, we partition all single-particle states into core, valence, and excluded orbits. A “valence configuration” is one with all core orbits occupied and all excluded orbits unoccupied). The Hamiltonian is then driven to a block-diagonal form and we may diagonalize in the (typically much smaller) sub-space of valence configurations. Such a diagonalization directly corresponds to a standard large-scale shell model calculation with an effective interaction defined by H ( ) . This approach is referred to as the valence-space IMSRG (VS-IMSRG), and it is used for all calculations presented.
Generally, the states we wish to target in a valence space approach are not well-described by a single closed-shell configuration, and the choice of | Φ becomes less clear. In this work I use the ensemble normal-ordering (ENO) approach [32], which amounts to taking fractional occupation numbers such that the reference has spherical symmetry and the correct number of particles on average. In addition, I use the Magnus formulation of the IMSRG [33], in which we write
U ( s ) e Ω ( s )
where Ω = Ω is the Magnus operator. Equations (6) and (2) may be combined to obtain a flow equation for Ω ( s ) in terms of η ( s ) , and operators (including the Hamiltonian), are transformed as
O ( s ) = e Ω ( s ) O ( 0 ) e Ω ( s ) = O ( 0 ) + [ Ω ( s ) , O ( 0 ) ] + 1 2 ! Ω ( s ) , [ Ω ( s ) , O ( 0 ) ] +
In (7), each commutator is truncated at the normal-ordered two-body level, and the series is computed iteratively until the size of a term falls below a numerical threshold.

2.3. Aspects Relevant to Beta Decay

Two additional details of the calculation are relevant for β decays. The first is the choice of single-particle basis in which we express the operators at s = 0 . In this work I use a Hartree–Fock basis with Coulomb and isospin-breaking strong forces included, so that for a given set of single-particle quantum numbers { n , , j } , the proton and neutron radial wave functions are not identical. The second is the choice of reference. Because we wish to compute the initial state, with N neutrons and Z protons, consistently with the final state with N ± 1 neutrons and Z 1 protons, there is some ambiguity about which reference | Φ should be used. If we retain all induced many-body terms during the SRG evolution, the choice of reference is irrelevant. However, the accuracy of the IMSRG(2) approximation depends on the choice of reference. The two natural choices for β decay are to use the N , Z of the initial state or the final state. I will discuss this in more detail in Section 4.

2.4. Comparison with Other Ab Initio Methods

Besides the VS-IMSRG, there are a number of ab initio methods available which can be used to treat beta decay, and it is worth briefly outlining them for context. For a more detailed account of ab initio many-body methods for nuclei, the reader is referred to [15]. Quantum Monte Carlo [11] and no-core shell model [34] approaches have well-understood approximation errors and can often yield quasi-exact solutions. The price is that the computational cost of these methods scales exponentially with the number of particles, and so are generally limited to the p-shell, with A 16 .
Polynomially-scaling methods, like coupled cluster (CC) [12], self-consistent Green’s function [35], IMSRG [23], and many-body perturbation theory (MBPT) [36], can access heavier nuclei at the cost of approximation errors that, while systematically improvable, are more difficult to assess. With the exception of quantum Monte Carlo (which requires a local potential), all of these methods can use the same input Hamiltonian and operators, enabling benchmark comparisons to better understand many-body errors [37,38].
Perhaps the most significant difference between the VS-IMSRG approach and the other polynomially-scaling methods is that the latter methods are generally constructed to target a single (ground) state of a given nucleus, and other states are expressed in terms of particle-hole excitations out of that state (notable exceptions being the shell model couple- cluster approach [39], which decouples a valence space in a similar manner to the VS-IMSRG, and the in-medium generator coordinate method [40] in which initial and final states are expressed by superpositions of shapes). This can in principle lead to trouble for transitions between states which are not, to a good approximation, related by a few particle-hole excitations. In contrast, the VS-IMSRG, by virtue of an exact diagonalization in the valence space, can describe two states which are related by many-particle-many-hole excitations—so long as those excitations live in the valence space.
It bears emphasis that all the above methods are systematically improvable and, given the same input Hamiltonian, should all converge to the same answer as their respective truncations are relaxed. Which truncation scheme converges most rapidly will in general depend on the state and observable in question. Consequently, until the truncation errors of polynomial-scaling methods are more quantitatively understood, benchmarks and cross-checks between methods are essential for making robust predictions.

3. Gamow–Teller Decays

In a Gamow–Teller decay, the leptons carry one unit of angular momentum and leave the parity of the nucleus unchanged. The relevant nuclear transition operator is obtained from the space-like part of the hadronic axial-vector current. The leading term in the non-relativistic reduction is g A σ τ , where g A 1.27 is the axial coupling constant, and σ and τ are the spin and isospin Pauli matrices.
Historically, when the leading operator was combined with shell model wave functions, a systematic “quenching” of the decay strength was observed, i.e., experimental matrix elements were smaller than the predicted ones, with a similar effect (tt should be noted that the quenching in M 1 observables is less clear experimentally, see e.g., Ref. [41]. Two body currents have been shown to improve agreement with experiment in quantum Monte Carlo calculations of p-shell nuclei [42], but these currents have not yet been studied in ab initio calculations of heavier systems) in isovector M 1 observables [43,44,45,46,47,48]. It was quickly surmised that the source of the discrepancy should be some combination of inadequate wave functions (missing correlations) and an inadequate transition operator (missing currents), that neither of these obviously dominated and that the two effects were not independent [44,49,50]. It was also suggested that pions ought to have something to do with the renormalization of the axial current in the nuclear medium [51].
These physics arguments survive in the modern EFT point of view, which organizes the nuclear interaction and coupling to external fields in powers of a ratio of scales. The distinction between short and long-distance physics is made by a cutoff, and the arbitrariness of the cutoff is reflected in the requirement that observables be independent of its value. The relationship between pions and the axial current arises as a consequence of broken chiral symmetry [52]. Importantly, chiral EFT enables a systematic and consistent construction of three-nucleon forces and two-body currents [53,54,55,56].
The result in the limit that the momentum carried by the leptons vanishes, up to order Q 0 (leading order is Q 3 ) is [53,57,58]
J = J 1 b + J 2 b ; cont + J 2 b ; 1 π
where
J 1 b ± = g A σ τ ±
J 2 b ; cont ± = 1 2 c D Λ χ f π 2 ( σ 1 τ 1 ± + σ 2 τ 2 ± )
J 2 b ; 1 π = g A f π 2 σ 2 · q 2 k 2 2 + M π 2 i p 1 2 m τ × ± + 2 c 3 τ 2 ± k 2 + ( c 4 + 1 4 m ) τ × ± ( σ 1 × k 2 ) + ( 1 2 )
where p i , p i are the incoming and outgoing momenta of the ith nucleon, k i = p i p i , τ × = τ 1 × τ 2 , and f π is the pion decay constant. The low-energy constants c 3 , c 4 and c D also enter into the NN and 3N forces, and so are not additional free parameters. (The constant c D is dimensionless and expected to be O ( 1 ) , while c i ( i = 3 , 4 ) have mass dimension −1. The dimensionless quantities c ^ i m N c i are expected to be O ( 1 ) .) Equations (9)–(11) correspond to diagrams (a), (b), and (c) in Figure 1, respectively. There are also corrections to the one-body operator of order p i 2 / m N 2 . Depending on how the nucleon mass is counted, these corrections will enter at different orders. In the counting of e.g., Park et al. [53], these corrections are also Q 0 , while in the counting used by other authors [37,56,58], including the calculations in this paper, these corrections are Q 1 . It is also worth noting that, because terms of order Q 2 vanish, some authors label Q 1 terms as next-to-leading-order (NLO) [56], while others call this next-to-next-to-leading-order (NNLO) [53]. To avoid confusion, I shall explicitly refer to powers of Q.
In Refs. [57,58], these currents were normal ordered with respect to uniform nuclear matter to obtain an in-medium quenching factor for the one-body operator. In Ref. [59], the full two-body current was constructed, consistently (in [59], the relationship between the two-body currents and three-body force contained an erroneous factor of 1 / 4 [55]) with the NN + 3N force, and the normal-ordered one-body operator (with respect to a Hartree–Fock reference) was used to compute Gamow–Teller decays of 14C, 22O, and 24O with the coupled-cluster method. (I also note that while the decay of 14C is interesting due to the anomalously long half-life [60,61], the small matrix element makes it difficult to draw conclusions regarding systematic quenching effects.) The effect of the residual normal-ordered two-body part of the operator was estimated and found to be small. In all three of these cases, a quenching of about the right size was obtained. In Ref. [62], axial currents up to Q 1 were used in quantum Monte Carlo calculations of A = 6–10 nuclei, where it was found that correlations beyond the shell model accounted for most of the quenching, with subleading currents playing a minor role. In Ref. [37], the full two-body current up to Q 0 was constructed consistently with the NN + 3N force, consistently SRG evolved, and evaluated, with the normal-ordered two-body operator fully included, in a range of nuclei in the p, s d , and p f shells, as well as 100Sn, using no-core shell model, coupled cluster, or VS-IMSRG to solve the many-body problem. Here, I will provide some additional calculations not presented in [37], and some further discussion. Specifically, I use an interaction for which no VS-IMSRG results were presented in [37] (though the conclusions drawn are the same), and I select additional transitions with large matrix elements to better emphasize the quenching “signal” over valence-space configuration-mixing “noise”.
The experimental Gamow–Teller matrix elements are obtained from the f t values by
f t = K f V f A B ( F ) + B ( G T ) G V 2
with K ( 2 π 3 7 ln 2 ) / ( m e 5 c 4 ) , and K / G V 2 6140 s. The Gamow–Teller matrix element is defined as
M ( G T ) ( 2 J i + 1 ) B ( G T ) 1 / 2 .
Note different definitions have been used in the literature, e.g., one may divide the right hand side by g A , as was done in [37]. The definition (13) leaves the experimental value independent of the adopted form of the current, or the adopted value of g A . The theoretical matrix element is given by M G T = Ψ f J Ψ i , with the current J as defined in (8), with or without the two-body part.
I consider Gamow–Teller transitions in nuclei in the p, s d , and p f shells, with experimental data taken from Refs. [46,47,48]. I have selected transitions with large transition matrix elements, with the goal of reducing sensitivity to fine-tuned cancellations. I also consider the decay of 100Sn, which was treated with equations-of-motion coupled cluster in [37], and for which the experimental picture is still somewhat conflicted [63,64,65,66]. I adopt the average value presented in [66]. In the VS-IMSRG calculation of 100Sn, I use valence space consisting of the 0 f 5 , 1 p 3 , 1 p 3 , 0 g 9 orbits for protons and 0 f 7 , 1 d 5 , 1 d 3 , 2 s 1 , 0 h 11 for neutrons.
For the theoretical calculations, I use the NN + 3N (lnl) interaction developed by Navrátil [35]. The interaction and current are consistently SRG-evolved to a scale λ S R G = 2.0 fm 1 and evaluated in an oscillator space defined by 2 n + e max = 12 and ω = 16 . The 3N matrix elements are further truncated with e 1 + e 2 + e 3 E 3 max = 14 . All operators are transformed to the Hartree–Fock basis, and then the residual 3N operators are truncated (the NO2B approximation). Next, a VS-IMSRG calculation is performed using the code imsrg++ [67], yielding and effective valence space interaction and operator. The valence space diagonalization is carried out either using NuShellX@MSU [68] with operators evaluating using the code nutbar [69], or with KSHELL [70]. The results are listed in Table A1 and plotted in Figure 2.
Table A1 in the appendix contains the numerical results. The column labeled M ( G T ) exp lists the experimental Gamow–Teller matrix elements defined by (13) (experimental uncertainties are not listed). The column labeled σ τ bare is the obtained by evaluating the operator σ τ (assuming identical radial wave functions for protons and neutrons) between valence space wave functions obtained using the VS-IMSRG evolved interaction. The column labeled σ τ IMSRG is obtained by consistently SRG and VS-IMSRG evolving the σ τ operator (including the radial mismatch due to the Hartree–Fock basis). Finally, M ( G T ) th also includes the two-body currents, consistently SRG and VS-IMSRG evolved. In a few cases, the listed strength is summed over multiple final states with the same spin and parity.
In Figure 2, panel (a) shows a scatter plot of M ( G T ) exp vs. M ( G T ) bare , while panel (b) shows M G T ) th vs. M ( G T ) exp . The solid line shows y = x corresponding to the perfect agreement between theory and experiment. The dashed line shows a best-fit slope, which is indicated as a quenching factor at the top of the figure. For this quenching factor, I only include s d and p f shell nuclei because the p shell nuclei have a large scatter due to nuclear structure details. The quantity in parenthesis indicates the standard deviation about the best-fit line. If I include p shell nuclei in the fit, the full theory quenching factor changes to q = 0.99 , but the standard deviation increases to 0.21.
It is evident from Table A1 that both the correlations included in σ τ IMSRG and the two-body currents lead to a reduction of the Gamow–Teller matrix element. As discussed in Ref. [37], the detailed breakdown of the quenching into correlations and currents is scheme- and scale-dependent; some of the effects attributed to correlations when using a hard interaction get shuffled into currents, when using a soft interaction. Thus the smaller impact of currents found in quantum Monte Carlo calculations [62] is consistent with the harder interactions used. However, even with soft interactions, the impact of currents for p-shell nuclei is less than for heavier systems.
It is also evident from Figure 2 that the systematic quenching effect, observed when using the bare σ τ operator, essentially vanishes when using the consistently-evolved operator including two-body currents.
In the right panel of Figure 2 I highlight the transition 12 N 12 C 0 + as an illustration of the cancellation effects in the p shell which wash out the quenching signal. When evaluating the bare σ τ operator (the matrix element of the bare operator is not an observable, so strictly there’s no reason different Hamiltonians should agree on it. On the other hand one might expect some degree of universality within a low-resolution picture like the shell model [21,71]) (including g A ), there are four terms that contribute, corresponding to proton-to-neutron transitions p 3 p 3 , p 3 p 1 , p 1 p 3 , and p 1 p 1 . The contributions are +0.366, −0.955, +1.592, and −0.173, respectively, totaling to 0.830. Evidently, there is significant cancellation so that a relatively small change of the individual terms can lead to a relatively large change on the final matrix element. For example, if I use valence-space wave functions obtained with the phenomenological Cohen–Kurath interaction [72], the bare operator yields a matrix element of 1.219. This difference, due to configuration mixing within the valence space, is larger than the systematic quenching effect of interest.
We may view this “noise” in the quenching from another perspective. Figure 3 shows the theory-to-theory quenching factor M ( G T ) th / M ( G T ) bare as a function of mass number. This indicates the quenching factor needed if we wanted to approximately account for the correlations and currents contained in M ( G T ) th . The main point here is to emphasize that the quenching is not a smooth function of A, but in fact has considerable state dependence.
It is clear that including two-body currents consistently with the NN + 3N interaction helps to remove the ambiguity of empirical quenching factors. Moving forward, the analogous calculation should be pursued for double beta decay, where the quenching effect is an important source of uncertainty [58,73]. In single beta decay, it will be important to explore the impacts of currents from the next order in the chiral expansion to confirm the convergence.

4. Superallowed 0 + 0 + Fermi Decays

For transitions between J = 0 states, B ( G T ) = 0 by conservation of angular momentum. Furthermore, in the limit in which isospin is a perfect symmetry, a “superallowed” transition between T = 1 isobaric analogue states yields B ( F ) = 2 , and so (12) reduces to (I am also neglecting here radiative corrections, which have a non-negligible impact)
f t = K 2 G V 2 ( isospin limit )
(where here f = f V ) . This implies all superallowed 0 + 0 + should have the same f t value, and that from this one may measure the coupling constant for semileptonic decay G V , which is in turn related to the constant G F obtained from muon decay by G V = V u d G F , where V u d is the up-down element of the Cabibbo–Kobayashi–Maskawa (CKM) quark mixing matrix. Consequently, precise f t measurements of superallowed 0 + 0 + decays provide a sensitive test of the Standard Model: non-universality of superallowed f t values, or non-unitarity of the CKM matrix would be signs of new physics.
Of course, isospin is not a perfect symmetry of the Standard Model. It is broken by the quark electric charges, and the up-down mass difference. This is manifested at the nuclear level as the Coulomb force between protons and isospin-violating strong interactions. The Standard Model corrections to (14) have been parameterized by Towner and Hardy [6] as
F t f t ( 1 + δ R ) ( 1 + δ NS δ C ) = K 2 G V 2 Δ R V .
In (15) Δ R V is a process-independent radiative correction [74], δ R is a radiative correction only depending on the electron energy and the charge of the daughter nucleus, and δ NS is a radiative correction depending on the detailed nuclear structure. The isospin-symmetry-breaking correction δ C accounts for the fact that the final state is not exactly an isospin rotation of the initial state.
Consequently, only δ NS and δ C are the purview of nuclear structure theory. To draw an analogy with the situation for Gamow–Teller decays, δ C corresponds to including the effects of correlations for the leading operator τ , while the radiative corrections correspond to sub-leading corrections to the operator, with δ NS corresponding to two-body currents. The difference here is that the corrections are sub-leading in the fine structure constant α 1 / 137 (or Z α ), as opposed to the chiral EFT expansion parameter Q 1 / 4 . The various corrections are illustrated in Figure 4.
In this paper, we focus on the δ C correction, for no better reason than the operator is the simplest to implement. Towner and Hardy decompose δ C into a correction due to isospin-breaking configuration mixing effects, and a correction due to the mismatch in single-particle wave functions between protons and neutrons. As we will be treating both within a consistent calculation, such a decomposition is not necessary (and ambiguous) and we will simply use
δ C = 1 | M F | 2 / 2
where M F = Ψ f τ Ψ i is the result of the ab initio calculation.
Nevertheless, it is useful to keep the two mechanisms (configuration mixing and wave function mismatch) in mind when considering the impact of various approximations. The wave function mismatch effect is taken into account primarily by the fact that we use a Hartree–Fock single-particle basis, with Coulomb and nuclear ISB effects included in the potential (see also [75]).
To get an idea of what the VS-IMSRG framework produces for the isospin-breaking correction δ C , I consider three transitions spanning the p, s d , and f p shells: 14 O 14 N, 34 Ar 34 Cl, and 46 Cr 46 V. I take the 1.8/2.0 (EM) interaction [76] with oscillator frequency ω = 16 MeV and E 3 max = 16 . The resulting δ C values are plotted in Figure 5 as a function of the e m a x truncation. I show results where the normal-ordering reference | Φ is taken to be either the initial or the final nucleus, and also the results from including only the one-body part of the evolved operator. For reference, I also indicate the δ C values adopted by Towner and Hardy [6].
If the calculation were under control, we should observe the following: convergence with respect to e max ; independence of the choice of reference; and a relatively small correction from including induced two-body terms, indicating a converging hierarchy of the cluster expansion. For the 46 Cr 46 V transition, we observe reference independence and a small two-body correction, but only a hint of convergence in e max . For the lighter nuclei, the situation is worse, especially for 14 O 14 N. In all cases, it appears that the large e max behavior will need to be incorporated in some manner, possibly by utilizing natural orbitals [77,78], or by obtaining an extrapolation formula [79,80]. However, I leave this for future work. It appears that, at least for the near term, such calculations will tell us more about the IMSRG than about physics beyond the Standard Model.
Figure 5. Isospin-breaking correction δ C for three transitions, computed with the VS-IMSRG as a function of the e max truncation. Dashed lines show the result using only the one-body part of the evolved operator while solid lines also include induced two-body terms. The black circles indicate the preferred values reported by Towner and Hardy [6]. For the 14 O decay, we also include coupled-cluster points [81].
Figure 5. Isospin-breaking correction δ C for three transitions, computed with the VS-IMSRG as a function of the e max truncation. Dashed lines show the result using only the one-body part of the evolved operator while solid lines also include induced two-body terms. The black circles indicate the preferred values reported by Towner and Hardy [6]. For the 14 O decay, we also include coupled-cluster points [81].
Particles 04 00038 g005
The transition 14 O 14 N warrants a closer inspection, because it is light enough that it can be benchmarked against the no core shell model [34], although the observed e max dependence suggests converged results may be challenging. Moreover, the dramatic reference-dependence and contribution of two-body terms make this a good system for studying such effects, which are also important (in double beta decay, the bare operator is already two-body, and we are concerned with induced three-body terms, making this a more challenging case) in neutrinoless double beta decay [38,73,82,83]. The two references used are shown schematically in Figure 6.
Because 14 O is a closed shell, it may also be treated by the coupled-cluster method [12] with the transition handled by the isospin-breaking equation-of-motion approach [59]. The strong e max dependence suggests sensitivity to infrared physics, and this may be probed by including the continuum in the coupled-cluster calculation [84]. The results at e max = 12 are indicated with yellow diamonds in Figure 5. The large δ C correction obtained with coupled cluster, as well as the substantial effect of including the continuum reinforces the notion that the pathology is not unique to the IMSRG.
The reference dependence persists even down to e max = 3 , which is sufficiently small that we may directly perform truncated configuration interaction (CI) calculations and extrapolate to the full CI result. At e max = 3 , the VS-IMSRG yields δ C values of 0.839% and 0.056% with 14 O and 14 N references, respectively. The full CI result is 0.081%, which is considerably closer to the 14 N reference value. This reinforces our suspicion that the calculation with the 14 O reference is misbehaving.
As I mentioned above, results would be independent of the reference if all induced many-body terms were retained during the IMSRG evolution. In practice, we only retain up to two-body operators, so any reference dependence indicates the impact of discarded three-body or (if we are especially unlucky) higher-body terms. If I truncate the expansion (7) for the transformed operator at two nested commutators, the reference dependence remains, while truncating at one nested commutator eliminates the effect. This suggests that three-body operators—which first show up at two nested commutators if O ( 0 ) is purely one-body—are the culprit.
A more complete investigation would benefit from calculations at the IMSRG(3) level, which are becoming available [85], but which will not be pursued here. For the moment, I will speculate. It appears that using the 14 O reference produces the largest error on the δ C value out of all the calculations presented in Figure 5. This is naively counterintuitive because 14 O is a closed shell and should be best approximated by the single-configuration reference. However, we are interested in the extent to which isospin symmetry is violated. Choosing a reference that approximates 14 O well but does a poor job for 14 N artificially breaks this symmetry, leading to an overestimate of the correction δ C . On the other hand, the ensemble 14 N reference used is not a great approximation of the state of interest in 14 N, which is actually open shell. Using the 14 N ensemble reference, we make an error in our description of 14 N, and we make a similar error in the description of 14 O, so the artificial breaking of isospin symmetry is reduced. In support of this, using a 12 C reference, which is a poor approximation of both 14 O and 14 N, results in δ C values in very close agreement to those obtained with the 14 N reference.
This is in line with the behavior observed with other relative quantities, namely excitation energies and separation energies. It is a robust finding that the VS-IMSRG predicts too-high 2 + excitation energies for closed-shell nuclei [86,87]. This can be understood by considering that the reference is a good approximation of the 0 + ground state, and a worse approximation for the 2 + excited state. The truncation of three-body terms has a more severe impact on the 2 + state, leading to missed correlation energy, and consequently an excitation energy that is too high. For open-shell nuclei, the reference is a mediocre approximation of both the ground and excited states, and so they are treated on more equal footing, leading to a more accurate excitation energy. Likewise, it was observed that separation energies are more accurately obtained when using the same valence space for both the mass A and mass A 1 nuclei, even if a different valence space might produce a more accurate ground state for one of the nuclei [88].
Preliminary studies have found that including three-body terms ameliorates the issue with 2 + states, and an analogous effect is observed in coupled cluster [89]. Presumably retaining three-body terms would also reduce the dependence of separation energies on the choice of valence space, but this has not yet been explored. One hopes that then including three-body terms will also help with the δ C calculation.

5. Conclusions and Outlook

We are in an exciting time in nuclear structure theory, in which it is becoming possible to address issues that were long plagued by ambiguities arising from inconsistent modeling. On the question of Gamow–Teller decays, considerable progress has been made and it appears the issue is understood. However, in order to fully put the question to rest one should properly assess theoretical errors from the EFT truncation and the many-body solution, and demonstrate systematic improvement in both.
In addition, a similar quenching is observed in strong-interaction charge exchange reactions [90], where it is assumed that the transition operator for the target nucleus is proportional to σ τ . The quenching is often ascribed to missing strength at higher energies (the equivalent of “correlations” in beta decay). However, three-body forces should play an analogous role as two-body currents do for beta decay (replace the axial current in the diagrams in Figure 1 with an additional nucleon). From another point of view, the three-body forces correspond to an effective density-dependent two-nucleon interaction, which has also been considered in this context [91]. In practice, charge exchange cross sections are often analyzed by normalizing to a low-lying beta decay transition [92], thus implicitly assuming that any quenching effects in charge exchange scale identically to those in beta decay. It will be interesting to quantitatively investigate this parallel.
For superallowed Fermi decays, the work of Towner and Hardy [6] has laid a clear path, but there is more work to do on the many-body side. A more detailed understanding of how errors creep into the calculations will be essential.

Funding

This work was supported by the U.S. Department of Energy, Office of Science, Office of Nuclear Physics, contract no. DEAC02-06CH11357.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

I thank my collaborators P. Gysbers, G. Hagen, J. D. Holt, G. R. Jansen, T. D. Morris, P. Navrátil, T. Papenbrock, S. Quaglioni, A. Schwenk, and K. A. Wendt for providing matrix elements of the chiral NN + 3N interactions and axial-vector currents. I also thank T. Miyagi for helpful discussions regarding the KSHELL code and various related scripts, and M. Rho for some enlightening comments.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Table of Gamow–Teller Matrix Elements

Table A1. Matrix elements for the Gamow–Teller transitions plotted in Figure 2 and Figure 3. In the last column, † indicates the lowest 3 states of the listed J π in the final nucleus are summed in the matrix element, while * indicates that 4 states are summed.
Table A1. Matrix elements for the Gamow–Teller transitions plotted in Figure 2 and Figure 3. In the last column, † indicates the lowest 3 states of the listed J π in the final nucleus are summed in the matrix element, while * indicates that 4 states are summed.
A Z i Z f 2 J i 2 J f M ( GT ) exp σ τ bare σ τ IMSRG M ( GT ) th
623022.7482.9952.8172.776
743332.8823.0882.8892.808
743312.6782.9072.7092.639
1276201.1840.8300.6370.616
1276222.3701.8161.7281.687
1587110.8891.0371.0771.035
1798553.1683.6813.5043.424
1898202.2092.8602.4832.430
19109112.2732.9442.6412.540
2089021.0581.4191.3071.266
201110422.4032.4302.2082.124
241312882.8863.0232.7312.639 †
251312551.9712.4332.2732.139
261413023.0553.6483.1283.006 *
271112531.3611.7181.4351.365
271413551.6881.9341.7031.630
301213021.0901.4881.2811.181
351516111.0331.2991.0600.999
371918331.1691.7681.6331.565
392019331.3081.9671.7241.628
412120772.9994.0733.7333.548
42212014122.4973.3053.0062.836
422221022.0382.7132.4152.237
452221771.1231.5131.3621.252
452322771.8012.1771.9821.844
472423330.9421.1901.0770.967
482524883.5964.0813.4943.340 *
492524551.3641.7681.5251.466
492524570.7640.7680.6560.629
1005049022.8705.3553.7173.471

References

  1. Fermi, E. Versuch einer Theorie der β-Strahlen. I. Z. Phys. 1934, 88, 161–177. [Google Scholar] [CrossRef]
  2. Lee, T.D.; Yang, C.N. Question of Parity Conservation in Weak Interactions. Phys. Rev. 1956, 104, 254–258. [Google Scholar] [CrossRef]
  3. Wu, C.S.; Ambler, E.; Hayward, R.W.; Hoppes, D.D.; Hudson, R.P. Experimental test of parity conservation in beta decay. Phys. Rev. 1957, 105, 1413–1415. [Google Scholar] [CrossRef]
  4. Avignone, F.T.; Elliott, S.R.; Engel, J. Double beta decay, Majorana neutrinos, and neutrino mass. Rev. Mod. Phys. 2008, 80, 481–516. [Google Scholar] [CrossRef] [Green Version]
  5. Holstein, B.R. Precision frontier in semileptonic weak interactions: Theory. J. Phys. G Nucl. Part. Phys. 2014, 41, 114001. [Google Scholar] [CrossRef]
  6. Hardy, J.C.; Towner, I.S. Superallowed 0+ → 0+ nuclear β decays: 2014 critical survey, with precise results for Vud and CKM unitarity. Phys. Rev. C 2015, 91, 025501. [Google Scholar] [CrossRef] [Green Version]
  7. Hayen, L.; Severijns, N.; Bodek, K.; Rozpedzik, D.; Mougeot, X. High precision analytical description of the allowed β spectrum shape. Rev. Mod. Phys. 2018, 90, 015008. [Google Scholar] [CrossRef] [Green Version]
  8. Epelbaum, E.; Hammer, H.W.; Meißner, U.G. Modern theory of nuclear forces. Rev. Mod. Phys. 2009, 81, 1773–1825. [Google Scholar] [CrossRef]
  9. Machleidt, R.; Entem, D.R. Chiral effective field theory and nuclear forces. Phys. Rep. 2011, 503, 1–75. [Google Scholar] [CrossRef] [Green Version]
  10. Hammer, H.W.; König, S.; van Kolck, U. Nuclear effective field theory: Status and perspectives. Rev. Mod. Phys. 2020, 92, 025004. [Google Scholar] [CrossRef]
  11. Carlson, J.; Gandolfi, S.; Pederiva, F.; Pieper, S.C.; Schiavilla, R.; Schmidt, K.E.E.; Wiringa, R.B.B. Quantum Monte Carlo methods for nuclear physics. Rev. Mod. Phys. 2015, 87, 1067–1118. [Google Scholar] [CrossRef]
  12. Hagen, G.; Papenbrock, T.; Hjorth-Jensen, M.; Dean, D.J. Coupled-cluster computations of atomic nuclei. Rep. Prog. Phys. 2014, 77, 096302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Binder, S.; Langhammer, J.; Calci, A.; Roth, R. Ab initio path to heavy nuclei. Phys. Lett. B 2014, 736, 119. [Google Scholar] [CrossRef] [Green Version]
  14. Morris, T.D.; Simonis, J.; Stroberg, S.R.; Stumpf, C.; Hagen, G.; Holt, J.D.; Jansen, G.R.; Papenbrock, T.; Roth, R.; Schwenk, A. Structure of the Lightest Tin Isotopes. Phys. Rev. Lett. 2018, 120, 152503. [Google Scholar] [CrossRef] [Green Version]
  15. Hergert, H. A Guided Tour of ab initio Nuclear Many-Body Theory. Front. Phys. 2020, 8, 1–33. [Google Scholar] [CrossRef]
  16. Somà, V. Self-Consistent Green’s Function Theory for Atomic Nuclei. Front. Phys. 2020, 8, 1–31. [Google Scholar] [CrossRef]
  17. Lee, D. Recent Progress in Nuclear Lattice Simulations. Front. Phys. 2020, 8, 1–7. [Google Scholar] [CrossRef]
  18. Van Kolck, U. The Problem of Renormalization of Chiral Nuclear Forces. Front. Phys. 2020, 8, 79. [Google Scholar] [CrossRef]
  19. Phillips, D.R. What hath Weinberg wrought? Reflections on what Weinberg’s papers on ‘Nuclear Forces from Chiral Lagrangians’ did and did not accomplish. arXiv 2021, arXiv:2107.03558. [Google Scholar]
  20. Cirigliano, V.; Dekens, W.; de Vries, J.; Hoferichter, M.; Mereghetti, E. Toward Complete Leading-Order Predictions for Neutrinoless Double β Decay. Phys. Rev. Lett. 2021, 126, 172002. [Google Scholar] [CrossRef]
  21. Bogner, S.; Furnstahl, R.; Schwenk, A. From low-momentum interactions to nuclear structure. Prog. Part. Nucl. Phys. 2010, 65, 94–147. [Google Scholar] [CrossRef] [Green Version]
  22. Furnstahl, R.J.; Hebeler, K. New applications of renormalization group methods in nuclear physics. Rep. Prog. Phys. 2013, 76, 126301. [Google Scholar] [CrossRef] [Green Version]
  23. Hergert, H.; Bogner, S.K.; Morris, T.D.; Schwenk, A.; Tsukiyama, K. The In-Medium Similarity Renormalization Group: A Novel Ab Initio Method for Nuclei. Phys. Rep. 2016, 621, 165–222. [Google Scholar] [CrossRef] [Green Version]
  24. Hergert, H. In-Medium Similarity Renormalization Group for Closed and Open-Shell Nuclei. Phys. Scr. 2017, 92, 023002. [Google Scholar] [CrossRef] [Green Version]
  25. Hergert, H.; Bogner, S.K.; Lietz, J.G.; Morris, T.D.; Novario, S.J.; Parzuchowski, N.M.; Yuan, F. In-Medium Similarity Renormalization Group Approach to the Nuclear Many-Body Problem. In An Advanced Course in Computational Nuclear Physics; Hjorth-Jensen, M., Lombardo, M.P., van Kolck, U., Eds.; Springer: Cham, Switzerland, 2017; pp. 477–570. [Google Scholar] [CrossRef] [Green Version]
  26. Stroberg, S.R.; Bogner, S.K.; Hergert, H.; Holt, J.D. Nonempirical Interactions for the Nuclear Shell Model: An Update. Annu. Rev. Nucl. Part. Sci. 2019, 69, 307–362. [Google Scholar] [CrossRef] [Green Version]
  27. Anderson, E.R.; Bogner, S.K.; Furnstahl, R.J.; Perry, R.J. Operator evolution via the similarity renormalization group: The deuteron. Phys. Rev. C 2010, 82, 054001. [Google Scholar] [CrossRef] [Green Version]
  28. Schuster, M.D.; Quaglioni, S.; Johnson, C.W.; Jurgenson, E.D.; Navrátil, P. Operator evolution for ab initio theory of light nuclei. Phys. Rev. C 2014, 90, 011301. [Google Scholar] [CrossRef] [Green Version]
  29. Parzuchowski, N.M.; Stroberg, S.R.; Navrátil, P.; Hergert, H.; Bogner, S.K. Ab initio electromagnetic observables with the in-medium similarity renormalization group. Phys. Rev. C 2017, 96, 034324. [Google Scholar] [CrossRef] [Green Version]
  30. Tropiano, A.J.; Bogner, S.K.; Furnstahl, R.J. Operator evolution from the similarity renormalization group and the Magnus expansion. Phys. Rev. C 2020, 102, 034005. [Google Scholar] [CrossRef]
  31. White, S.R. Numerical canonical transformation approach to quantum many-body problems. J. Chem. Phys. 2002, 117, 7472. [Google Scholar] [CrossRef] [Green Version]
  32. Stroberg, S.R.; Calci, A.; Hergert, H.; Holt, J.D.; Bogner, S.K.; Roth, R.; Schwenk, A. Nucleus-Dependent Valence-Space Approach to Nuclear Structure. Phys. Rev. Lett. 2017, 118, 032502. [Google Scholar] [CrossRef] [Green Version]
  33. Morris, T.D.; Parzuchowski, N.M.; Bogner, S.K. Magnus expansion and in-medium similarity renormalization group. Phys. Rev. C 2015, 92, 034331. [Google Scholar] [CrossRef] [Green Version]
  34. Barrett, B.R.; Navrátil, P.; Vary, J.P. Ab initio no core shell model. Prog. Part. Nucl. Phys. 2013, 69, 131–181. [Google Scholar] [CrossRef] [Green Version]
  35. Somà, 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. [Google Scholar] [CrossRef] [Green Version]
  36. Tichai, A.; Roth, R.; Duguet, T. Many-Body Perturbation Theories for Finite Nuclei. Front. Phys. 2020, 8, 1–29. [Google Scholar] [CrossRef]
  37. Gysbers, P.; Hagen, G.; Holt, J.D.; Jansen, G.R.; Morris, T.D.; Navrátil, P.; Papenbrock, T.; Quaglioni, S.; Schwenk, A.; Stroberg, S.R.; et al. Discrepancy between experimental and theoretical β-decay rates resolved from first principles. Nat. Phys. 2019, 15, 428–431. [Google Scholar] [CrossRef]
  38. Yao, J.M.; Bally, B.; Engel, J.; Wirth, R.; Rodríguez, T.R.; Hergert, H. Ab Initio Treatment of Collective Correlations and the Neutrinoless Double Beta Decay of 48Ca. Phys. Rev. Lett. 2020, 124, 232501. [Google Scholar] [CrossRef]
  39. Sun, Z.H.; Morris, T.D.; Hagen, G.; Jansen, G.R.; Papenbrock, T. Shell-model coupled-cluster method for open-shell nuclei. Phys. Rev. C 2018, 98, 054320. [Google Scholar] [CrossRef] [Green Version]
  40. Yao, J.M.; Engel, J.; Wang, L.J.; Jiao, C.F.; Hergert, H. Generator-coordinate reference states for spectra and 0νββ decay in the in-medium similarity renormalization group. Phys. Rev. C 2018, 98, 054311. [Google Scholar] [CrossRef] [Green Version]
  41. Matsubara, H.; Tamii, A.; Nakada, H.; Adachi, T.; Carter, J.; Dozono, M.; Fujita, H.; Fujita, K.; Fujita, Y.; Hatanaka, K.; et al. Nonquenched Isoscalar Spin- M 1 Excitations in s d -Shell Nuclei. Phys. Rev. Lett. 2015, 115, 102501. [Google Scholar] [CrossRef]
  42. Pastore, S.; Pieper, S.C.; Schiavilla, R.; Wiringa, R.B. Quantum Monte Carlo calculations of electromagnetic moments and transitions in A ≤ 9 nuclei with meson-exchange currents derived from chiral effective field theory. Phys. Rev. C 2013, 87, 1–15. [Google Scholar] [CrossRef] [Green Version]
  43. Wilkinson, D.H. Renormalization of the Axial-Vector Coupling Constant in Nuclear β Decay. Phys. Rev. C 1973, 7, 930–936. [Google Scholar] [CrossRef]
  44. Wilkinson, D. Renormalization of the axial-vector coupling constant in nuclear β-decay (II). Nucl. Phys. A 1973, 209, 470–484. [Google Scholar] [CrossRef]
  45. Brown, B.A.; Chung, W.; Wildenthal, B.H. Empirical Renormalization of the One-Body Gamow-Teller β -Decay Matrix Elements in the 1s-0d Shell. Phys. Rev. Lett. 1978, 40, 1631–1635. [Google Scholar] [CrossRef] [Green Version]
  46. Brown, B.; Wildenthal, B. Experimental and theoretical Gamow-Teller beta-decay observables for the sd-shell nuclei. At. Data Nucl. Data Tables 1985, 33, 347–404. [Google Scholar] [CrossRef]
  47. Chou, W.T.; Warburton, E.K.; Brown, B.A. Gamow-Teller beta-decay rates for A ≤ 18 nuclei. Phys. Rev. C 1993, 47, 163–177. [Google Scholar] [CrossRef] [PubMed]
  48. Martínez-Pinedo, G.; Poves, A.; Caurier, E.; Zuker, A.P. Effective gA in the pf shell. Phys. Rev. C 1996, 53, R2602. [Google Scholar] [CrossRef] [Green Version]
  49. Rho, M. Quenching of axial-vector coupling constant in β-decay and pion-nucleus optical potential. Nucl. Phys. Sect. A 1974, 231, 493–503. [Google Scholar] [CrossRef]
  50. Towner, I.S.; Khanna, F.C. Corrections to the single-particle M1 and Gamow-Teller matrix elements. Nucl. Phys. A 1983, 399, 334–364. [Google Scholar] [CrossRef]
  51. Ericson, M.; Figureau, A.; Thévenet, C. Pionic field and renormalization of the axial coupling constant in nuclei. Phys. Lett. B 1973, 45, 19–22. [Google Scholar] [CrossRef]
  52. Scherer, S.; Schindler, M.R. A Primer for Chiral Perturbation Theory; Lecture Notes in Physics; Springer: Berlin/Heidelberg, Germany, 2012; Volume 830. [Google Scholar] [CrossRef]
  53. Park, T.S.; Marcucci, L.E.; Schiavilla, R.; Viviani, M.; Kievsky, A.; Rosati, S.; Kubodera, K.; Min, D.P.; Rho, M. Parameter-free effective field theory calculation for the solar proton-fusion and hep processes. Phys. Rev. C 2003, 67, 055206. [Google Scholar] [CrossRef] [Green Version]
  54. Gårdestig, A.; Phillips, D.R. How Low-Energy Weak Reactions Can Constrain Three-Nucleon Forces and the Neutron-Neutron Scattering Length. Phys. Rev. Lett. 2006, 96, 232301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Gazit, D.; Quaglioni, S.; Navrátil, P. Three-Nucleon Low-Energy Constants from the Consistency of Interactions and Currents in Chiral Effective Field Theory. Phys. Rev. Lett. 2009, 103, 102502, Erratum in 2019, 122, 029901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Krebs, H.; Epelbaum, E.; Meißner, U.G.G. Nuclear axial current operators to fourth order in chiral effective field theory. Ann. Phys. 2017, 378, 317–395. [Google Scholar] [CrossRef] [Green Version]
  57. Park, T.S.; Jung, H.; Min, D.P. In-medium effective axial-vector coupling constant. Phys. Lett. Sect. B Nucl. Elem. Part. High-Energy Phys. 1997, 409, 26–32. [Google Scholar] [CrossRef] [Green Version]
  58. Menéndez, J.; Gazit, D.; Schwenk, A. Chiral Two-Body Currents in Nuclei: Gamow-Teller Transitions and Neutrinoless Double-Beta Decay. Phys. Rev. Lett. 2011, 107, 062501. [Google Scholar] [CrossRef]
  59. Ekström, A.; Jansen, G.R.; Wendt, K.A.; Hagen, G.; Papenbrock, T.; Bacca, S.; Carlsson, B.; Gazit, D. Effects of three-nucleon forces and two-body currents on Gamow-Teller strengths. Phys. Rev. Lett. 2014, 113, 262504. [Google Scholar] [CrossRef] [Green Version]
  60. Holt, J.W.; Brown, G.E.; Kuo, T.T.S.; Holt, J.D.; Machleidt, R. Shell model description of the 14C dating β decay with Brown-Rho-scaled NN interactions. Phys. Rev. Lett. 2008, 100, 1–4. [Google Scholar] [CrossRef] [Green Version]
  61. Maris, P.; Vary, J.P.; Navrátil, P.; Ormand, W.E.; Nam, H.; Dean, D.J. Origin of the Anomalous Long Lifetime of C 14. Phys. Rev. Lett. 2011, 106, 202502. [Google Scholar] [CrossRef] [Green Version]
  62. Pastore, S.; Baroni, A.; Carlson, J.; Gandolfi, S.; Pieper, S.C.; Schiavilla, R.; Wiringa, R.B. Quantum Monte Carlo calculations of weak transitions in A = 6–10 nuclei. Phys. Rev. C 2018, 97, 022501. [Google Scholar] [CrossRef] [Green Version]
  63. Faestermann, T.; Schneider, R.; Stolz, A.; Sümmerer, K.; Wafers, E.; Friese, J.; Geissel, H.; Hellström, M.; Kienle, P.; Körner, H.J.; et al. Decay studies of N ≈ Z nuclei from 75Sr to 102Sn. Eur. Phys. J. A 2002, 15, 185–188. [Google Scholar] [CrossRef]
  64. Batist, L.; Górska, M.; Grawe, H.; Janas, Z.; Kavatsyuk, M.; Karny, M.; Kirchner, R.; la Commara, M.; Mukha, I.; Plochocki, A.; et al. Systematics of Gamow-Teller beta decay “Southeast” of 100Sn. Eur. Phys. J. A 2010, 46, 45–53. [Google Scholar] [CrossRef] [Green Version]
  65. Hinke, C.B.; Böhmer, M.; Boutachkov, P.; Faestermann, T.; Geissel, H.; Gerl, J.; Gernhäuser, R.; Górska, M.; Gottardo, A.; Grawe, H.; et al. Superallowed Gamow–Teller decay of the doubly magic nucleus 100Sn. Nature 2012, 486, 341–345. [Google Scholar] [CrossRef] [Green Version]
  66. Lubos, D.; Park, J.; Faestermann, T.; Gernhäuser, R.; Krücken, R.; Lewitowicz, M.; Nishimura, S.; Sakurai, H.; Ahn, D.S.; Baba, H.; et al. Improved Value for the Gamow-Teller Strength of the Sn 100 Beta Decay. Phys. Rev. Lett. 2019, 122, 222502. [Google Scholar] [CrossRef]
  67. Stroberg, S.R. Imsrg++ Code. Available online: https://github.com/ragnarstroberg/imsrg (accessed on 27 October 2021).
  68. Brown, B.A.; Rae, W.D.M. The Shell-Model Code NuShellX@MSU. Nucl. Data Sheets 2014, 120, 115–118. [Google Scholar] [CrossRef]
  69. Stroberg, S.R. Nutbar Code. Available online: https://github.com/ragnarstroberg/nutbar (accessed on 27 October 2021).
  70. Shimizu, N. Nuclear shell-model code for massive parallel computation, “KSHELL”. arXiv 2013, arXiv:1310.5431. [Google Scholar]
  71. Bogner, S.K.; Kuo, T.T.S.; Schwenk, A. Model-independent low momentum nucleon interaction from phase shift equivalence. Phys. Rep. 2003, 386, 1–27. [Google Scholar] [CrossRef] [Green Version]
  72. Cohen, S.; Kurath, D. Effective interactions for the 1p shell. Nucl. Phys. 1965, 73, 1–24. [Google Scholar] [CrossRef]
  73. Engel, J.; Menéndez, J. Status and Future of Nuclear Matrix Elements for Neutrinoless Double-Beta Decay: A Review. Rep. Prog. Phys. 2017, 80, 046301. [Google Scholar] [CrossRef] [PubMed]
  74. Seng, C.Y.; Gorchtein, M.; Ramsey-Musolf, M.J. Dispersive evaluation of the inner radiative correction in neutron and nuclear β decay. Phys. Rev. D 2019, 100, 13001. [Google Scholar] [CrossRef] [Green Version]
  75. Miller, G.A.; Schwenk, A. Isospin-symmetry-breaking corrections to superallowed Fermi β decay: Radial excitations. Phys. Rev. C 2009, 80, 064319. [Google Scholar] [CrossRef] [Green Version]
  76. Hebeler, K.; Bogner, S.K.; Furnstahl, R.J.; Nogga, A.; Schwenk, A. Improved nuclear matter calculations from chiral low-momentum interactions. Phys. Rev. C 2011, 83, 031301. [Google Scholar] [CrossRef] [Green Version]
  77. Tichai, A.; Müller, J.; Vobig, K.; Roth, R. Natural orbitals for ab initio no-core shell model calculations. Phys. Rev. C Nucl. Phys. 2019, 99, 034321. [Google Scholar] [CrossRef] [Green Version]
  78. Hoppe, J.; Tichai, A.; Heinz, M.; Hebeler, K.; Schwenk, A. Natural orbitals for many-body expansion methods. Phys. Rev. C 2021, 103, 014321. [Google Scholar] [CrossRef]
  79. Furnstahl, R.J.; Hagen, G.; Papenbrock, T. Corrections to nuclear energies and radii in finite oscillator spaces. Phys. Rev. C 2012, 86, 031301. [Google Scholar] [CrossRef] [Green Version]
  80. Furnstahl, R.J.; Phillips, D.R.; Wesolowski, S. A recipe for EFT uncertainty quantification in nuclear physics. J. Phys. G Nucl. Part. Phys. 2014, 42, 034028. [Google Scholar] [CrossRef] [Green Version]
  81. Hagen, G.; (Oak Ridge National Laboratory). Private communication, 2019.
  82. Belley, A.; Payne, C.G.; Stroberg, S.R.; Miyagi, T.; Holt, J.D. Ab Initio Neutrinoless Double-Beta Decay Matrix Elements for Ca 48, Ge 76, and Se 82. Phys. Rev. Lett. 2021, 126, 042502. [Google Scholar] [CrossRef]
  83. Yao, J.M.; Belley, A.; Wirth, R.; Miyagi, T.; Payne, C.G.; Stroberg, S.R.; Hergert, H.; Holt, J.D. Ab initio benchmarks of neutrinoless double-β decay in light nuclei with a chiral Hamiltonian. Phys. Rev. C 2021, 103, 014315. [Google Scholar] [CrossRef]
  84. Hagen, G.; Papenbrock, T.; Hjorth-Jensen, M. Ab initio computation of the F17 proton halo state and resonances in A = 17 nuclei. Phys. Rev. Lett. 2010, 104, 5–8. [Google Scholar] [CrossRef] [Green Version]
  85. Heinz, M.; Tichai, A.; Hoppe, J.; Hebeler, K.; Schwenk, A. In-medium similarity renormalization group with three-body operators. Phys. Rev. C 2021, 103, 044318. [Google Scholar] [CrossRef]
  86. Simonis, J.; Stroberg, S.R.; Hebeler, K.; Holt, J.D.; Schwenk, A. Saturation with chiral interactions and consequences for finite nuclei. Phys. Rev. C 2017, 96, 014303. [Google Scholar] [CrossRef] [Green Version]
  87. Taniuchi, R.; Santamaria, C.; Doornenbal, P.; Obertelli, A.; Yoneda, K.; Authelet, G.; Baba, H.; Calvet, D.; Château, F.; Corsi, A.; et al. 78Ni revealed as a doubly magic stronghold against nuclear deformation. Nature 2019, 569, 53–58. [Google Scholar] [CrossRef] [Green Version]
  88. Stroberg, S.R.; Holt, J.D.; Schwenk, A.; Simonis, J. Ab initio Limits of Atomic Nuclei. Phys. Rev. Lett. 2021, 126, 022501. [Google Scholar] [CrossRef] [PubMed]
  89. Hagen, G.; Jansen, G.R.; Papenbrock, T. Structure of Ni 78 from First-Principles Computations. Phys. Rev. Lett. 2016, 117, 172501. [Google Scholar] [CrossRef] [Green Version]
  90. Gaarde, C. Gamow-Teller and M1 resonances. Nucl. Phys. Sect. A 1983, 396, 127–144. [Google Scholar] [CrossRef]
  91. Cheon, T.; Takayanagi, K. Isospin-Dependent Effective Interaction in Nucleon-Nucleus Scattering. Phys. Rev. Lett. 1992, 68, 1291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Zegers, R.G.T.; Akimune, H.; Austin, S.M.; Bazin, D.; Berg, A.M.D.; Berg, G.P.; Brown, B.A.; Brown, J.; Cole, A.L.; Daito, I.; et al. The (t, He3) and (He3, t) reactions as probes of Gamow-Teller strength. Phys. Rev. C Nucl. Phys. 2006, 74, 1–15. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Diagrams for (a) leading-order Gamow–Teller decay σ τ , (b) short-range two-body current, and (c) long-range two-body current.
Figure 1. Diagrams for (a) leading-order Gamow–Teller decay σ τ , (b) short-range two-body current, and (c) long-range two-body current.
Particles 04 00038 g001
Figure 2. Experimental M ( G T ) vs. M ( G T obtained with (a) the bare g A σ τ operator, (b) the SRG- and IMSRG-transformed transition operator including two-body currents. In both panels, the solid line shows y = x corresponding to perfect agreement, while the dashed line indicates the best-fit slope.
Figure 2. Experimental M ( G T ) vs. M ( G T obtained with (a) the bare g A σ τ operator, (b) the SRG- and IMSRG-transformed transition operator including two-body currents. In both panels, the solid line shows y = x corresponding to perfect agreement, while the dashed line indicates the best-fit slope.
Particles 04 00038 g002
Figure 3. Theory-to-theory “quenching” factors as a function of mass number A.
Figure 3. Theory-to-theory “quenching” factors as a function of mass number A.
Particles 04 00038 g003
Figure 4. Schematic high-resolution diagrams corresponding to (a) the leading operator τ , (b) the one-body correction leading to Δ R V and δ R , and (c) the two-body correction leading to δ N S .
Figure 4. Schematic high-resolution diagrams corresponding to (a) the leading operator τ , (b) the one-body correction leading to Δ R V and δ R , and (c) the two-body correction leading to δ N S .
Particles 04 00038 g004
Figure 6. Schematic illustration of reference states used for the 14 O 14 N decay in the left panel of Figure 5. The half-filled circles for 14 N indicate the equal filling of m-states used in the ensemble reference.
Figure 6. Schematic illustration of reference states used for the 14 O 14 N decay in the left panel of Figure 5. The half-filled circles for 14 N indicate the equal filling of m-states used in the ensemble reference.
Particles 04 00038 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Stroberg, S.R. Beta Decay in Medium-Mass Nuclei with the In-Medium Similarity Renormalization Group. Particles 2021, 4, 521-535. https://0-doi-org.brum.beds.ac.uk/10.3390/particles4040038

AMA Style

Stroberg SR. Beta Decay in Medium-Mass Nuclei with the In-Medium Similarity Renormalization Group. Particles. 2021; 4(4):521-535. https://0-doi-org.brum.beds.ac.uk/10.3390/particles4040038

Chicago/Turabian Style

Stroberg, Steven Ragnar. 2021. "Beta Decay in Medium-Mass Nuclei with the In-Medium Similarity Renormalization Group" Particles 4, no. 4: 521-535. https://0-doi-org.brum.beds.ac.uk/10.3390/particles4040038

Article Metrics

Back to TopTop