KEK-CP-393
OU-HET-1186
semileptonic form factors from lattice QCD with Möbius domain-wall quarks
Abstract
We calculate the form factors for the decay in 2+1 flavor lattice QCD. For all quark flavors, we employ the Möbius domain-wall action, which preserves chiral symmetry to a good precision. Our gauge ensembles are generated at three lattice cutoffs , 3.6 and 4.5 GeV with pion masses as low as MeV. The physical lattice size satisfies the condition to control finite volume effects (FVEs), while we simulate a smaller size at the smallest to directly examine FVEs. The bottom quark masses are chosen in a range from the physical charm quark mass to 0.7 to control discretization effects. We extrapolate the form factors to the continuum limit and physical quark masses based on heavy meson chiral perturbation theory at next-to-leading order. Then the recoil parameter dependence is parametrized using a model independent form leading to our estimate of the decay rate ratio between the tau () and light lepton () channels in the Standard Model. A simultaneous fit with recent data from the Belle experiment yields , which is consistent with previous exclusive determinations, and shows good consistency in the kinematical distribution of the differential decay rate between the lattice and experimental data.
I Introduction
The semileptonic decay, where and represents the corresponding neutrino, plays a key role in stringent tests of the Standard Model (SM) and searches for new physics. The channels associated with light leptons provide a determination of the Cabibbo-Kobayashi-Maskawa (CKM) matrix element , which is a fundamental parameter of the SM. On the other hand, the channel is expected to be a good probe of new physics, since, for instance, it is expected to strongly couple to the charged Higgs boson predicted by supersymmetric models. Indeed, there is an intriguing tension between the SM and experiments on the decay rate ratio
(1) |
describing the lepton-flavor universality violation Amhis et al. (2023).
However, there has been a long-standing tension between the determinations from exclusive and inclusive decays . The Heavy Flavor Averaging Group reported that analysis of the inclusive decay yields Amhis et al. (2023)
(2) |
which shows a (9 %) tension with the determination from the exclusive decays
(5) |
It has been argued Crivellin and Pokorski (2015) that such a tension can be explained by introducing a tensor interaction beyond the SM, which, however, largely distorts the decay rate measured precisely by experiments. Therefore, it is likely that the tension is due to our incomplete understanding of the theoretical and/or experimental uncertainty.
The largest theoretical uncertainty in the exclusive determination comes from form factors, which describe non-perturbative QCD effects to the decay amplitude through the matrix elements
(6) | |||||
(7) |
where is the recoil parameter defined by four velocities and , and is the polarization vector of , which satisfies . Note that, here and in the following, kinematical variables (momentum, velocity, polarization vector, but not space-time coordinates) with and without the symbol “” represent those for and , respectively.
Lattice QCD is a powerful method to provide a first-principles calculation of the form factors with systematically improvable accuracy Vaquero (2022); Kaneko (2023). However, until recently, only at the zero-recoil limit had been calculated in unquenched lattice QCD Bernard et al. (2009); Bailey et al. (2014); Harrison et al. (2018). In the previous determination of , therefore, other information of the form factors in addition to had to be determined by fitting experimental data of the differential decay rate to a theoretical expression Amhis et al. (2023); Bernlochner et al. (2017); Bigi et al. (2017a); Grinstein and Kobach (2017). While a model independent parametrization of the form factors by Boyd, Grinstein and Lebed (the BGL parametrization) is available Boyd et al. (1997), its large number of parameters due to the poor knowledge on the form factors makes the fit unstable Bigi et al. (2017b); Bernlochner et al. (2019); Gambino et al. (2019); Ferlewicz et al. (2021) even with experimental data of differential distribution with respect to all kinematical variables, namely the recoil parameter and three decay angles, from the KEKB/Belle experiment Waheed et al. (2019); Abdesselam et al. (2017) 111 We note that a more recent data on the normalized distribution from the full Belle data set Prim et al. (2023) as well as a recent report from the succeeding SuperKEKB/Belle II experiment Abudinén et al. (2023) are available in a different event-reconstruction approach..
One of the largest problems in the lattice study of meson physics is that the simulation cost rapidly increases as the lattice spacing decreases. In spite of recent progress in computer performance and development of simulation algorithms, it is still difficult to simulate lattice cutoffs sufficiently larger than the physical bottom quark mass to suppress discretization errors to, say, the 1 % level. For the time being, a practical approach to physics on the lattice is to employ a heavy quark action based on an effective field theory, such as the heavy quark effective theory (HQET), to directly simulate at currently available cutoffs . Another straightforward approach is the so-called relativistic approach, where a lattice QCD action is used also for heavy quarks but with their masses sufficiently smaller than the lattice cutoff to suppress discretization effects. The Fermilab/MILC reported the first lattice calculation of all form factors at zero and non-zero recoils Bazavov et al. (2022) using the Fermilab approach for heavy quarks, namely a HQET interpretation of the anisotropic Wilson-clover quark action El-Khadra et al. (1997), and staggered light quarks. It was recently followed by the HPQCD Collaboration with a fully relativistic approach Harrison and Davies (2024), where the highly improved staggered quark action was used both for light and heavy flavors.
In this paper, we present an independent study with a fully relativistic approach using the Möbius domain-wall quark action Brower et al. (2017) for all relevant quark flavors. This preserves chiral symmetry to good precision at finite lattice spacing, reducing the leading discretization errors to second order, i.e. . We can also use chiral perturbation theory in the continuum limit () to guide the chiral extrapolation to the physical pion mass. Note also that we do not need explicit renormalization of the weak axial and vector currents on the lattice to calculate relevant form factors, because the renormalization constants of these currents coincide with each other and are canceled by taking appropriate ratios of relevant correlation functions Hashimoto et al. (1999). Our preliminary analyses and discussions on these features can be found in our earlier reports Kaneko et al. (2018, 2019, 2022).
This paper is organized as follows. After a brief introduction to our choice of simulation parameters, the form factors are extracted from correlation functions at the simulation points in Sec. II. These results are extrapolated to the continuum limit and physical quark masses in Sec. III. In Sec. IV, we generate synthetic data of the form factors and fit them to a model-independent parametrization in terms of the recoil parameter to make a comparison with the previous lattice and phenomenological studies as well as to estimate the SM value of . We also perform a simultaneous fit with experimental data to estimate . Our conclusions are given in Sec. V.
II Form factors at simulation points
II.1 Gauge ensembles
Our gauge ensembles are generated for 2+1-flavor lattice QCD, where dynamical up and down quarks are degenerate and dynamical strange quark has a distinct mass. We take three lattice spacings of and 0.080 fm Colquhoun et al. (2022), which correspond to the lattice cutoffs , 3.610(9) and 4.496(9) GeV, respectively. They are determined using the Yang-Mills gradient flow Borsanyi et al. (2012). The tree-level improved Symanzik gauge action Weisz (1983); Weisz and Wohlert (1984); Luscher and Weisz (1985) and the Möbius domain-wall quark action Brower et al. (2017) are used to control discretization errors and to preserve chiral symmetry to a good precision. We refer the interested reader to Refs. Brower et al. (2017); Colquhoun et al. (2022); Boyle (2015) for the five dimensional formulation of the quark action and our implementation. Its four-dimensional effective Dirac operator is given by
(8) |
where represents the quark mass. We employ the kernel operator , where is the Wilson-Dirac operator with a negative mass with and with stout smearing Morningstar and Peardon (2004) applied three successive times to the gauge links. With our choice of the staple weight (), the smearing radius is at the scale of the lattice spacing and vanishes in the limit of . We therefore expect that the link smearing does not change the continuum limit of the form factors. It may induce additional discretization effects, which, however, do not turn out to be large in our continuum and chiral extrapolation in Sec. III. The polar decomposition approximation in Eq. (8) is realized for the sign function in the five-dimensional domain-wall implementation. With this choice, the residual quark mass, which is a measure of the chiral symmetry violation, is suppressed to the level of 1 MeV at fm, and 0.2 MeV or less on finer lattices, which are significantly smaller than the physical up and down quark masses Colquhoun et al. (2022). With reasonably small values of the fifth-dimensional size – 12, the computational cost is largely reduced from that of our previous large-scale simulation Aoki et al. (2008a); Aoki et al. (2012) using the overlap formulation that also preserves chiral symmetry Kaneko et al. (2014).
lattice parameters | [MeV] | [MeV] | |||||
---|---|---|---|---|---|---|---|
, GeV, | 0.0190 | 0.0400 | 499(1) | 618(1) | 12, 16, 24, 28 | 100 | 1 |
0.0120 | 0.0400 | 399(1) | 577(1) | 12, 16, 22, 26 | 100 | 1 | |
0.0070 | 0.0400 | 309(1) | 547(1) | 12, 16, 22, 26 | 100 | 2 | |
0.0035 | 0.0400 | 230(1) | 527(1) | 8, 12, 16, 20 | 100 | 2 | |
, | 0.0035 | 0.0400 | 226(1) | 525(1) | 8, 12, 16, 20 | 100 | 2 |
, GeV, | 0.0120 | 0.0250 | 501(2) | 620(2) | 18, 24, 36, 42 | 50 | 1 |
0.0080 | 0.0250 | 408(2) | 582(2) | 18, 24, 33, 39 | 50 | 1 | |
0.0042 | 0.0250 | 300(1) | 547(2) | 18, 24, 33, 39 | 50 | 2 | |
, GeV, | 0.0030 | 0.0150 | 284(1) | 486(1) | 16, 24, 32, 40 | 50 | 2 |
4.17 | 0.44037, 0.55046, 0.68808 |
4.35 | 0.27287, 0.34109, 0.42636, 0.53295, 0.66619 |
4.47 | 0.210476, 0.263095, 0.328869, 0.4110859, 0.5138574, 0.6423218 |
We employ the Möbius domain-wall action for all relevant quark flavors. Our choices of the degenerate up and down quark mass correspond to pion masses from MeV down to 230 MeV. Chiral symmetry allows us to use heavy meson chiral perturbation theory (HMChPT) Randall and Wise (1993); Savage (2002) for our chiral extrapolation of simulation results to the physical pion mass without introducing terms to describe discretization effects. We take a strange quark mass close to its physical value fixed from . The charm quark mass is set to its physical value fixed from the spin-averaged charmonium mass Nakayama et al. (2016). Depending on the lattice spacing , we take three to six bottom quark masses satisfying the condition to avoid large discretization errors. We note that, with chiral symmetry, the leading discretization error is reduced to . As we will see in the next section, and dependences of the form factors are mild, and the extrapolation to the continuum limit and physical quark masses is under good control. Our simulation parameters are summarized in Tables 1 and 2.
The spatial lattice size is chosen to satisfy the condition to suppress finite volume effects (FVEs). At our smallest on the coarsest lattice, we also simulate a smaller physical lattice size to directly examine the FVEs.
The statistics at the lattice cutoffs , 3.6 and 4.5 GeV, are 5,000, 2,500 and 2,500 Hybrid Monte Carlo trajectories with a unit trajectory length of 1, 2 and 4 times the molecular dynamics time unit, respectively, which is increased to take account of the longer auto correlation toward the continuum limit. On each ensemble, we take configurations in an equal trajectory interval to calculate correlation functions relevant to the decay. Our choice of is listed in Table 1. More details on our simulation setup can be found in Ref. Colquhoun et al. (2022).
II.2 Ratio method
The matrix elements can be extracted from the three- and two-point functions
(9) | |||||
(10) |
through their ground state contribution
(11) |
where is the weak vector () or axial () current, and we omit the symbol “” on the momentum variable and the argument for the two-point function for simplicity. The meson interpolating field is denoted by (), where Gaussian smearing is applied to enhance their overlap with the ground state . The meson is at rest () throughout this paper. The dependence of the form factors is studied by varying the three-momentum of as (in this paper, we denote the momentum on the lattice in units of ). To this end, we also calculate two-point functions with the local sink operator
(12) |
which are used in a correlator ratio (15) below.
The three-point functions are calculated by using the sequential source method, where the total temporal separation between the source and sink operators is fixed and the temporal location of the weak current is varied. In order to control the contamination from the excited states, we repeat our measurement for four values of the source-sink separation listed in Table 1.
The statistical accuracy of the two- and three-point functions are improved by averaging over the spatial location of the source operator as indicated in Eqs. (9) and (10). To this end, we employ a volume source with noise. At MeV, we repeat our measurement over two values of the source time-slices and to take the average of the correlators. Namely in Eqs. (9) and (10) is 2 at MeV, and 1 otherwise. The correlators with non-zero momentum are also averaged over all possible ’s based on parity and rotational symmetries on the lattice.

The number of measurements on each ensemble is given as in Table 1. As mentioned above, correlation functions for each configuration are averaged over values of the source time-slice . Then, simulation data of configurations are divided into 50 bins: namely, the bin size is two (one) configurations for (4.35 and 4.47). The statistical error is estimated by the bootstrap method with 5,000 replicas. Figure 1 shows the bin size dependence of the relative statistical error of a three-point function at our largest cutoff, where the topological charge changes much less frequently leading to larger auto-correlation compared to the coarser lattices Colquhoun et al. (2022). Our choice of the bin size on this finest lattice is one configuration, and we do not observe significant increase of the statistical error toward larger bin sizes. This suggests that our bin size is sufficiently large partly due to our choice of larger unit trajectory length towards the continuum limit. We also note that the error estimate is stable when we vary the number of bootstrap replicas as well as when we employ the jackknife method instead.

As observed in our previous simulation in the trivial topological sector Aoki et al. (2008a), the local topological fluctuation are active even if we fix the “global” topological charge , namely the whole-volume integral of the topological charge density. In Ref. Aoki et al. (2008b), we demonstrated that the topological susceptibility is calculable within the fixed topology setup. Our result for shows quark mass dependence consistent with ChPT, and its value extrapolated to the chiral limit is in good agreement with that from our simulation in the -regime Fukaya et al. (2007a); Fukaya et al. (2007b). References Brower et al. (2003); Aoki et al. (2007) argued that the bias due to the freezing of the global topology can be considered as FVEs suppressed by the inverse space-time volume. In our study of the decay Colquhoun et al. (2022), we confirmed that such FVEs on the pion effective mass are well suppressed, so that its dependence is not significant. Figure 2 shows the statistical average of a three-point function calculated for each . We do not observe any significant dependence of the average at . At larger , we do not have enough data for a reliable error estimate, but individual data plotted by crosses are consistent with the averages at . We therefore conclude that the topology freezing effect is insignificant within our statistical accuracy.


To precisely extract the form factors, we construct ratios of the correlation functions, in which unnecessary overlap factors and exponential damping factors cancel for the ground-state contribution Hashimoto et al. (1999). The statistical fluctuation is also expected to partly cancel in such a ratio. With and mesons at rest, the vector current matrix element (6) vanishes, and the axial vector one (7) is sensitive only to the axial vector form factor at zero recoil, which is the fundamental input to determine . In order to precisely determine , we employ a double ratio
(13) |
where the polarization is chosen as along the polarization of the current inserted, i.e. . The left panel of Fig. 3 shows our results for this ratio at our smallest and with . Note that the ratio is symmetrized as , since we use the same smearing function for both the source and sink operators. We observe a reasonable consistency among data at intermediate values of the source-sink separation (circles) and 16 (squares). The excited state contribution could potentially be significant but not large ( %) for the smallest (diamonds). Data at the largest (triangles) show a long plateau and are consistent with those at smaller within the large statistical errors. These observations suggest that the data at , which is roughly 1 fm, are dominated by the ground state contribution at least at the midpoint . The situation is similar in the right panel of the same figure for our largest and with , where data around midpoint are consistent among all simulated source-sink separations within 2 .


We find that all data are well described by the following fitting form including effects from the first excited states
(14) |
with for a wide range of the values of and . Here represents the energy difference between the meson ground state and the first excited state of the same quantum numbers, and is set to the value estimated from a two-exponential fit to the two-point function . The fit range is chosen such that all the data of with the source-to-current (current-to-sink) separation equal to or larger than a lower cut are included irrespective of the source-sink separation . As shown in the left panel of Fig. 4, our result for is stable against the choice of the lower cuts and .
The right panels of Figs. 3 and 4 for our largest also show ground state dominance with fm and stability of against the choice of the fit range, respectively. The situation is similar for other simulation parameters.


In order to extract form factors at non-zero recoils, the momentum and polarization vector need to be chosen appropriately. The axial vector matrix element (7) is sensitive only to with a polarization vector and a momentum that satisfies . We note that, here and in the following, three dimensional polarization vector is accompanied by its temporal component to be fixed from the convention , which is assumed for Eqs. (6) and (7). Then, the recoil parameter dependence of may be studied from the following ratio
(15) |
Here we use the two-point function with the local sink (12) as in our study of the decay Aoki et al. (2017). The vector form factor can be extracted from the following ratio through the vector current matrix element (6)
(16) |
Here we use two different polarization vectors, and , and momenta and that satisfy . Note that to share the same recoil parameter , and . We emphasize that renormalization factors of the weak vector and axial currents cancel thanks to chiral symmetry preserved in our simulations.


Figures 5 and 6 show an example of our results for and at our largest cutoff GeV and . Around the mid-point , we observe good consistency in these ratios among all simulated values of the source-sink separation fm. A similar ground state dominance is also observed at other simulation points. We carry out a simultaneous fit using a fitting form that takes account of the first excited state contribution as
(17) |
to extract the form factor ratio from the ground state contribution , and a similar form for to extract . Our data are well described by this fitting form with . From these form factor ratios and from , we calculate and at simulated values of .


There is a 2 bump of for (black triangles) around () in Fig. 5. Since the global topological charge does not fluctuate very much at the corresponding cutoff, one may expect bump(s) in the exponential decay of relevant correlators due to instantons frozen at the temporal separation for a certain period of our simulation time. However, Fig. 7 shows that there is no statistically significant bump of the relevant correlators at and . As mentioned above, the local topological fluctuation is active even when the global topological charge is fixed. We attribute 1 – 2 bumps of correlator ratios in Figs. 3, 5 – 6, 8 – 9 to large statistical fluctuations of correlators at the largest temporal separation , and accidental anti correlation of the statistical fluctuations. Since these data have larger statistical error than those at smaller , these bumps do not change results of the fits (14) and (17) to extract the ground state contribution.
The axial vector form factors and can be extracted from the axial vector matrix element (7) with the spatial momentum not perpendicular to the spatial polarization vector. The matrix element of , for instance, has non-zero sensitivity to and . We use the following ratio to extract
(18) | |||||
where , , , and . The factor appears, since the overlap factor of depends on whether the polarization vector is perpendicular to the momentum, even if . While can be estimated from individual fit to the relevant two-point functions, we employ a ratio , which shows better stability against the temporal separation. The reference temporal separation is chosen as by inspecting the statistical accuracy and the ground state saturation of . A similar ratio but with in the numerator is also sensitive to
(19) | |||||




We plot our results for and on our finest lattice in Figs. 8 and 9, respectively. Similar to other ratios, the excited state contamination is not large. The fit similar to (17) yields the ground state contributions, and , which are stable against the choice of the fit range: we do not observe dependence beyond 1 level when , that is our smallest value of the source-sink separation . The error rapidly increases at larger values of , because less data are available for the fit. From , and extracted above, we can calculate and .
For one of the momenta , it is not straightforward to use the above-mentioned correlator ratios, which use with as a correlator sensitive only to . In this case, we use a linear combination
(20) |
which only involves with and .
The double ratio enables us to calculate with 1 % statistical error or better, as it only involves correlators with zero momentum and the statistical fluctuation partially cancels between the numerator and denominator. At non-zero recoil, the numerator of () is exclusively sensitive to (), and the statistical error of and is typically a few %. This is not the case for and with our setup where the meson is at rest. The numerator of involves and , and the typical precision for is 10 – 20 %. The statistical error increases to % for , since the central value is suppressed by heavy quark symmetry and is a linear combination of all axial form factors.
We note that, thanks to chiral symmetry, finite renormalization factors for the weak vector and axial currents cancel in the above correlator ratios (13), (15), (16), (18) and (19). While discretization effects to the wave function renormalization are not small Fahy et al. (2016), these also cancel in the ratios.
III Extrapolation to continuum limit and physical quark masses
III.1 Choice of fitting form




Our results for the form factors obtained at various simulation parameters are plotted as a function of the recoil parameter in Fig. 10. We find that the results with different heavy and light quark masses are all close to each other. It also turns out that the discretization effects are not large.
We extrapolate these results to the continuum limit and physical quark masses based on next-to-leading order heavy meson chiral perturbation theory (NLO HMChPT) Randall and Wise (1993); Savage (2002). The extrapolation form is generically written as
(21) | |||||
which comprises the constant term, chiral logarithm at and leading corrections in small expansion parameters
(22) |
with , , and nominal values of MeV.
We denote the one-loop integral function for the chiral logarithm by , where is the - mass splitting. While the loop function can be approximated as , we use its explicit form in Ref. Savage (2002). For the coupling, we take a value of , which was employed in the Fermilab/MILC paper Bailey et al. (2014) and covers previous lattice estimates Detmold et al. (2012); Becirevic and Sanfilippo (2013); Can et al. (2013); Bernardoni et al. (2015); Flynn et al. (2016). This choice does not lead to a large systematic uncertainty, because the chiral logarithm is suppressed by the small mass splitting squared .


In Fig. 11, we plot and at our smallest value of as a function of . The dependence is rather mild, possibly because there is no valence pion in this decay. We note that there is a non-trivial concave structure at small due to the opening of the decay. The effect estimated in NLO HMChPT is, however, not large compared to our statistical accuracy.
For our chiral extrapolation, we employ the so-called expansion in and : namely, we use the measured meson masses squared, and , instead of quark masses and , respectively, as well as quark-mass dependent for instead of the decay constant in the chiral limit. This is motivated by our observation Noaki et al. (2008) that the -expansion shows better convergence of the chiral expansion for light meson observables compared to the -expansion with , where gives the lowest order relation .


The extrapolation form explicitly includes the one-loop radiative correction for the matching of the heavy-heavy currents between QCD and HQET Falk et al. (1990); Falk and Grinstein (1990); Neubert (1992). Remaining heavy quark mass dependence of the form factors is then parametrized by polynomials of and . The dependence is mild as shown in Fig. 12. Our results are well described by the linear terms and in Eq. (21). We note that, for , the term is modified as to be consistent with Luke’s theorem Luke (1990), and we add the quadratic term to take account of possible dependence at .
The discretization effects in the form factors also turn out to be mild with our choice of parameters, GeV and . These are described by an expansion, i.e. linear in the independent and dependent parameters and , respectively. Figure 12 shows that the dependent discretization error is significant in our results. However, the net dependence is described reasonably well with our fitting form (21). We note that the dependence for and is insignificant due to their large relative uncertainty. As we will mention later, fit results do not change significantly if we exclude data at in our continuum and chiral extrapolation.
In this study, we first extrapolate the form factors to the continuum limit and physical quark masses, and then parametrize the dependence by using the model independent BGL parametrization, because it may not be simply combined with the chiral and heavy quark expansions. Concerning the dependence, therefore, the fit form (21) with an expansion in is used to interpolate our results. Figure 10 shows that our data do not show strong curvature in the simulation region of , and can be well described with only up to the quadratic term.
. /d.o.f 0.14(11) 0.898(54) 0.68(19) 0.08(16) -0.18(32) 0.59(17) 0.37(26) 0.0376(89) -0.953(70) 0.67(22) 0.18(13) -0.5(1.0) 0.9(3.6) -0.0(2.6) 0.13(49) – 3.3(3.0) -0.022(67) -0.41(86) 10.3(7.1) 0.18(13) 0.91(90) -0.4(3.3) 0.3(2.4) 1.25(47) – -3.4(2.7) 0.117(62) -1.72(75) 1.9(6.1) 0.15(14) 0.87(13) 1.47(45) 0.39(37) 0.33(13) – -0.51(63) 0.092(18) -1.32(14) 0.6(1.3)
Numerical results of our continuum and chiral extrapolation are summarized in Table 3. The polynomial coefficients turn out to be or less with our choice of the dimensionless expansion parameters (22), which are normalized by appropriate scales. Many of them are consistent with zero reflecting the mild dependence on the lattice spacing and quark masses mentioned above. As a result, our choice of the fitting forms (21) results in small values of , which, however, are only a rough measure of the goodness of the fits, because we employ approximated estimators of the covariance matrix as discussed in the next subsection.
III.2 Systematic uncertainties
III.2.1 Correlated fit
With limited statistics, the covariance matrix of our form factor data has exceptionally small eigenvalues with large statistical error. In this study, we test two methods to take the correlation into account in the combined continuum and chiral extrapolation. First, as in the study of Colquhoun et al. (2022), we introduce a lower cut of the eigenvalue (singular value) . Effects of the exceptionally small eigenvalues are suppressed by replacing eigenvalues smaller than by in the singular value decomposition of as
(23) |
where represents the eigenvector corresponding to the eigenvalue . Here eigenvalues are sorted in ascending order () and satisfies . We choose such that all eigenvalues below have statistical error larger than 100 %. The second method is the so-called shrinkage estimate of the covariance matrix Stein (1956); Ledoit and Wolf (2004), which amounts to replacing by
(24) |
With , the shrinkage is an interpolation between the uncorrelated () and correlated () fits. An optimal value of can be estimated by minimizing a loss function Ledoit and Wolf (2004). We observed, however, that results of the continuum and chiral extrapolation do not strongly depend on the choice of and . In this study, we therefore employ the singular value cut for our main results, since this is a conservative approach, which only tends to increase the final error. The shift of the fit results obtained with the shrinkage estimator is treated as a systematic error due to the estimate of .

With the approximated estimators of , is only a rough measure of the goodness of the fit. The impact is shown in Fig. 13. We observe that, in the case of our analysis, tends to be slightly smaller than that for the unquenched fit, and rather stable against the choice of the estimator (singular value cut, or shrinkage) and parameter . While is not invertible for the correlated fit (with for the shrinkage estimator), we may expect that does not change drastically and serves as a “rough” measure of the goodness of the fit even with the approximated estimators of . We also note that, in this paper, the approximated estimators are used only for the continuum and chiral extrapolation of form factors due to the large size of with different choices of the quark masses (including for the relativistic approach), lattice cutoff and recoil parameter.
III.2.2 Fitting form
The highest order is quadratic in the recoil parameter expansion and heavy quark expansion for , and linear for other expansions including chiral and finite expansions. The systematic error due to this choice is estimated by repeating the continuum and chiral extrapolation without one of the highest order terms, if the corresponding coefficient is poorly determined, otherwise by adding a yet higher order term.
Possible higher order corrections in the chiral expansion are also examined by testing the -expansion. We also test a multiplicative form
(25) | |||||
to examine the potential effect of cross terms. We find that these alternative fits yield a change of the fit results well below the statistical error.
III.2.3 Inputs
The lattice cutoff , physical meson masses , , , and the coupling are inputs to the continuum and chiral extrapolation. Since the masses have been measured very precisely, we take account of the uncertainty due to and by repeating the extrapolation with each input shifted by 1 .
We set and to those in the isospin limit as in our determination of Colquhoun et al. (2022). To examine the isospin breaking effects to the form factors, we test the continuum and chiral extrapolation to and as well as and . We also test the heavy quark expansion parameter replaced with and compare the extrapolations using and . The difference in the form factor can be considered as an estimate of the strong isospin breaking effect. It turns out to be small as suggested by the mild quark mass dependence of the form factors. We include this in the systematic uncertainty so that our synthetic data for the form factors can be used for both the neutral and charged meson decays.




III.2.4 Finite volume effects (FVEs)
Our spatial lattice size satisfies . At our smallest pion mass on the coarsest lattice, we also simulate a smaller size to directly examine FVEs. Comparison between the two volumes at that simulation point does not show any significant deviation: for instance, at is consistent within statistical error of %. We also repeat the continuum and chiral extrapolation by replacing the form factors at the smallest with those on the smaller . The change in the fit results is included in the systematic uncertainty, but turns out to be well below the statistical uncertainty. We conclude that finite volume effects are not significant in our simulation region of MeV.
III.2.5 Accuracy of continuum and chiral extrapolation
Figure 14 shows the estimate of various uncertainties for the form factors in the continuum limit and at the physical quark masses as a function of . As discussed in Sec. II, the axial (vector) matrix element is exclusively sensitive to () with an appropriate choice of the meson polarization vector and momentum. Combined with a precise determination at zero-recoil, the total uncertainty of is about 2 % in our simulation region of , and the largest uncertainty comes from statistics and discretization effects. On the other hand, either or needs to have non-zero momentum to determine , which is, therefore, less accurate. The total accuracy is % with 3 –4 % statistical and % discretization errors.
Since the simulated bottom quark masses are not far below the lattice cut-off (), we repeat the continuum and chiral extrapolation only using the data with . The change in the fit results is well below the statistical uncertainty. We therefore conclude that our continuum extrapolation in the relativistic approach is controlled.
On the other hand, the axial form factors, and , have larger uncertainty, i.e. 40 % and 15 %, respectively, dominated by the statistical error. Towards a more precise determination, it would be advantageous to simulate moving mesons so as to have matrix elements more sensitive to these form factors.


Previous determinations of often employed a HQET-based parametrization by Caprini, Lellouch and Neubert of and form factor ratios
(26) |
where higher order HQET corrections are expected to be partially canceled Caprini et al. (1998). Figure 15 shows that these ratios also have mild parameter dependences and can be safely extrapolated to the continuum limit and physical quark masses. We, however, do not employ the model-dependent CLN parametrization in the following analysis.
IV Form factors as a function of the recoil-parameter
IV.1 Fit to lattice data
From the continuum and chiral extrapolation in the previous section, we generate synthetic data of the form factors in the relativistic convention,
(27) | |||||
(28) |
with which the BGL parametrization was originally proposed. A kinematic invariant is defined with the maximum value of the momentum transfer , which corresponds to the zero recoil limit, and the threshold for the production channel. Instead of and , it is common to use linear combinations
(29) | |||||
(30) |
These are related to the form factors in the HQET convention as
(31) | |||||
(32) | |||||
(33) | |||||
(34) |
We note that there are two kinematical constraints to be satisfied at the minimum () and maximum values of ( with )
(35) | |||||
(36) |
With these form factors, we can write the helicity amplitudes for a given helicity of the intermediate boson, and the differential decay rate with respect to in simple forms
(37) | |||||
(38) | |||||
(39) |
and
(40) | |||||
From the continuum and chiral extrapolation presented in Sec. III, synthetic data of , and are calculated at reference values of the recoil parameter through the relations (31) – (34). We take , 1.060 and 1.100, which span the whole simulated region of , since our continuum and chiral extrapolation interpolates and in . Three values are chosen, because the statistical covariance matrix of the synthetic data develops ill-determined eigenvalues with four or more values of . This might be partly because, in our continuum and chiral extrapolation, we parametrize the dependence of each form factor using a quadratic form with three independent parameters. The numerical values together with their covariance matrix are presented in Table 4.
. 5.6928(936) 1.00000 0.98385 0.93944 0.94354 0.71378 0.49823 0.24518 0.22123 0.21755 0.22911 0.22641 0.22699 5.5879(992) 0.98385 1.00000 0.98338 0.92328 0.70780 0.49984 0.22221 0.20401 0.20620 0.22112 0.22194 0.22694 5.473(109) 0.93944 0.98338 1.00000 0.87531 0.68095 0.48820 0.19269 0.18105 0.18953 0.20747 0.21181 0.22233 18.612(324) 0.94354 0.92328 0.87531 1.00000 0.88532 0.70711 0.36140 0.35455 0.37432 0.20808 0.20305 0.20067 18.308(410) 0.71378 0.70780 0.68095 0.88532 1.00000 0.94630 0.58476 0.60631 0.65922 0.15762 0.15344 0.15170 17.983(545) 0.49823 0.49984 0.48820 0.70711 0.94630 1.00000 0.66784 0.70838 0.78248 0.11626 0.11328 0.11262 2.1704(962) 0.24518 0.22221 0.19269 0.36140 0.58476 0.66784 1.00000 0.98622 0.96645 0.05204 0.04696 0.04049 2.097(106) 0.22123 0.20401 0.18105 0.35455 0.60631 0.70838 0.98622 1.00000 0.98356 0.04418 0.03974 0.03438 1.992(109) 0.21755 0.20620 0.18953 0.37432 0.65922 0.78248 0.96645 0.98356 1.00000 0.04246 0.03872 0.03398 0.3326(212) 0.22911 0.22112 0.20747 0.20808 0.15762 0.11626 0.05204 0.04418 0.04246 1.00000 0.99764 0.98991 0.3178(213) 0.22641 0.22194 0.21181 0.20305 0.15344 0.11328 0.04696 0.03974 0.03872 0.99764 1.00000 0.99622 0.3013(215) 0.22699 0.22694 0.22233 0.20067 0.15170 0.11262 0.04049 0.03438 0.03398 0.98991 0.99622 1.00000
The synthetic data are fitted to the model-independent BGL parametrization
(41) |
where the parameter
(42) |
maps the whole semileptonic kinematical region () () into a small parameter region of , .
. input-A input-B 4 6.739, 6.750, 7.145, 7.150 4 6.730, 6.736, 7.135, 7.142 3 6.329, 6.920, 7.020 4 6.337, 6.899, 7.012, 7.280 3 6.275, 6.842, 7.250 3 6.274, 6,871, 7220
The Blaschke factors
(43) |
factor out the pole singularities outside the semileptonic region but below the threshold due to the resonances for a given channel of mesons. The explicit form is given as
(44) |
where
(45) |
is the -parameter corresponding to the -th resonance (in term of rather than ). In our main analysis, we employ resonance masses denoted as “input-A” in Table 5. This choice consists of slightly old experimental measurement Patrignani et al. (2016) ( for ) and estimates based on lattice QCD Dowdall et al. (2012); Colquhoun et al. (2015) ( for and for ) and quark potential models Godfrey (2004); Rai and Devlani (2013) (other ’s). This input has been used in some of previous phenomenological analysis using the BGL parametrization Bigi et al. (2017a); Gambino et al. (2019) and lattice calculations of form factors Bazavov et al. (2022); Harrison and Davies (2024). Another choice (“input-B” in Table 5) in the literature is used to estimate the systematics uncertainty due to the use of input-A (see below).
The outer function is chosen such that a constraint on the expansion coefficients derived from unitarity of the theory has a simple form
(46) |
The explicit form is given as
(47) | |||||
(48) | |||||
(49) | |||||
(50) |
where is the number of the spectator quarks with a correction due to SU(3) breaking. By , we denote the derivative of the vacuum polarization function of the weak current at zero momentum for a given channel and polarization . As in the literature, we replace by , from which the one-particle state contribution due to is subtracted for a better saturation of the unitarity bound (46). While a lattice QCD estimate is available Martinelli et al. (2021), we employ a perturbative estimate Bigi et al. (2017b); Grigo et al. (2012); Bigi and Gambino (2016) summarized in Table 6. We note that our choice of the resonance masses (input-A) and derivatives are the same as previous lattice studies to allow a direct comparison of the BGL parametrization with them. It is known that the unitarity bound (46) is not well saturated by the channel, and hence is rather weak. In this study, we do not impose this bound, but confirm that fit results satisfy it within the error.
. [GeV-2] [GeV-2] perturbation Bigi et al. (2017b); Grigo et al. (2012); Bigi and Gambino (2016) lattice QCD Martinelli et al. (2021)

Figure 16 shows the regular part of the axial form factor , which is to be expanded in . We observe rather mild dependence of all form factors. While there is no clear sign of curvature in our simulation region, we employ a quadratic form to safely suppress the truncation error. Among twelve coefficients, constants and are fixed to fulfill the kinematical constraints (35) at , where and the linear and quadratic terms in the -parameter expansion (41) vanish, and (36) at , where is well below unity, and hence the constant term gives rise to a large contribution to . To be consistent with the unitarity bound (46), we impose the bound () for all form factors and . Table 7 shows fit results, which describe the synthetic data with an acceptable value of .
. 0.0291(18) 1.00000 0.46208 0.00157 0.23345 0.01973 0.00342 0.23345 0.00136 0.01362 0.08685 0.00553 0.01628 0.045(35) 0.46208 1.00000 0.01024 0.06133 0.21056 0.06890 0.06133 0.01804 0.02956 0.03921 0.00394 0.03000 1.0(1.7) 0.00157 0.01024 1.00000 0.01137 0.02396 0.01746 0.01137 0.00789 0.01312 0.00205 0.01056 0.00708 0.01198(19) 0.23345 0.06133 0.01137 1.00000 0.06216 0.12057 1.00000 0.07462 0.09380 0.39328 0.19208 0.13641 0.018(11) 0.01973 0.21056 0.02396 0.06216 1.00000 0.49732 0.06216 0.49162 0.24351 0.15088 0.45046 0.25008 0.10(45) 0.00342 0.06890 0.01746 0.12057 0.49732 1.00000 0.12057 0.33668 0.20338 0.18421 0.23045 0.19828 0.002006(31) 0.23345 0.06133 0.01137 1.00000 0.06216 0.12057 1.00000 0.07462 0.09380 0.39328 0.19208 0.13641 0.0013(41) 0.00136 0.01804 0.00789 0.07462 0.49162 0.33668 0.07462 1.00000 0.51671 0.17631 0.84705 0.49244 0.03(21) 0.01362 0.02956 0.01312 0.09380 0.24351 0.20338 0.09380 0.51671 1.00000 0.19200 0.48706 0.98112 0.0484(16) 0.08685 0.03921 0.00205 0.39328 0.15088 0.18421 0.39328 0.17631 0.19200 1.00000 0.07882 0.12820 0.059(87) 0.00553 0.00394 0.01056 0.19208 0.45046 0.23045 0.19208 0.84705 0.48706 0.07882 1.00000 0.55376 0.9(1.1) 0.01628 0.03000 0.00708 0.13641 0.25008 0.19828 0.13641 0.49244 0.98112 0.12820 0.55376 1.00000
. choice of input-B, or i) for 0.32 0.04 0.03 0.01 0.00 ii) for – – 0.00 0.00 0.00 0.57 iii) for – – 0.00 0.00 0.00 0.10 iv) from lattice 0.08 0.02 0.03 v) 0.12 0.31 vi) 0.30 vii) 2 ’s 0.01 0.11
We test the stability of the fit with different choices of the input and reference values of . Input-B in Table 5 of the resonance masses for the Blaschke factors consists of the pseudo-scalar masses from recent experimental measurements Workman et al. (2022), and other resonance masses from phenomenological studies different from input-A: namely, Ref. Li et al. (2023) for and , and Ref. Eichten and Quigg (1994) for . We note that these vector and axial masses have been used in a phenomenological analysis Grinstein and Kobach (2017) and Belle analyses Abdesselam et al. (2017); Waheed et al. (2019) of the decay. For the derivatives of the vacuum polarization functions , we test the lattice QCD estimate in Ref. Martinelli et al. (2021). Shift of the expansion coefficients in units of their uncertainty is summarized in Table 8, where we test four choices of the input: namely, i) vector and axial resonance masses from Ref. Eichten and Quigg (1994), ii) ground and first excited pseudo-scalar masses from Ref. Workman et al. (2022), iii) second excited pseudo-scalar mass from Ref. Li et al. (2023), and iv) lattice estimate of . The shift of is at the 2- level or typically well below it. A large shift () in with the lattice QCD estimate of can be attributed to about 20% shift of for the axial channel. As suggested in Eq. (48), this simply leads to rescaling of all for the corresponding channel. Also for other coefficients, the changes in the Blaschke factors and outer functions due to the different inputs are largely absorbed into the shift in to reproduce the synthetic data of the form factors. As a result, the shift in physical observables, and in this paper, is well below 1 as we will see later. This also means that our results for in Table 7 should be used with our choice of the input, namely input-A for resonance masses in Table 5 and perturbative estimate of in Table 6, otherwise the uncertainty suggested in Table 8 should be taken into account.
In Table 8, we also test three different choices of the reference values : namely, v) decreased maximum value , vi) increased minimum value and vii) two values of against our main choice of . Table 8 shows that the shift of the coefficients is well below 1 . Their relative accuracy is slightly worse (typically by 5 – 10 %) with the above choices of suggesting that our main choice is reasonably good.
We confirm the unitarity bound for the axial vector channel as , which is poorly saturated by . The bounds for other channels are satisfied as and . The large uncertainty comes from the poorly-determined quadratic coefficients. The sums up to the linear term, and , suggest that the unitarity bound is rather weak as mentioned above.




The BGL parametrization and synthetic data of the form factors are plotted in Fig. 17. We observe reasonable consistency with the Fermilab/MILC results Bazavov et al. (2022) except the vector form factor near zero-recoil and the axial form factor extrapolated to large recoils. This can be seen in a more detailed comparison of the expansion coefficients in Fig. 18, where there is slight tension in , and . Recent HPQCD results show better consistency with ours, except for and .







In the same figure, we also plot a phenomenological estimate Gambino et al. (2019) obtained from the Belle data Waheed et al. (2019) combined with a lattice input of the FLAG average of Aoki et al. (2020). Note that this analysis employs the same BGL parametrization with the input ( and ) and orders (, , ), which are the same as our main analysis. The phenomenological result of is well determined and shows good consistency with lattice results, simply because it is mainly fixed from the lattice input. Other poorly determined coefficients suggest the importance of lattice calculations to obtain controlled BGL parametrization.
On the other hand, the slope is well determined from the phenomenological analysis of the experimental data, which are precise at large recoils. This study observes a reasonable agreement in and, hence, in the differential decay rate as discussed in the next subsection (see Fig. 19). We note, however, that the Fermilab/MILC and HPQCD studies reported tension in the dependence of the differential decay rate with experiment.
Since describes the contribution suppressed by to the decay rate (40), its expansion coefficients are poorly constrained by the experimental data for the light lepton channels (). The lattice determination of is, therefore, helpful towards a precision new physics search using the channel. Through numerical integration of Eq. (40) for and with fit results in Table 7, we obtain a pure SM value
(51) |
which does not resort to phenomenological models nor experimental data. The second error comes from the test of the stability of the BGL parametrization summarized in Table 8. It is mainly comes from the shift in and , but smaller than statistical and other systematic errors of the BGL parametrization. This estimate can be improved as by imposing the unitarity bound (46) on the expansion coefficients. These are consistent with Fermilab/MILC and HPQCD estimates, 0.265(13) and 0.279(13), respectively, and to be compared with the experimental average 0.295(14) Amhis et al. (2023). For a more precise and reliable estimate of , it is helpful to simulate larger recoils as well as resolving the tensions in shown in Fig. 18.
IV.2 Fit including Belle data
In this subsection, we carry out a simultaneous fit to our lattice and Belle data to estimate . The differential decay rate with respect to and three decay angles , and is given as
(52) | |||||
with the helicity amplitudes given in (37) and (38). For a simultaneous fit with the Belle data for in Ref. Waheed et al. (2019), we set , and include the Coulomb factor in the right-hand side. The decay angles are chosen the same as for the Belle paper, where the full kinematical distribution is described by four differential decay rates , , and , which are integrated over the other three variables. The differential decay rates are provided at ten equal-size bins in the range of each variable, , and .
To estimate , we fit our synthetic data of the form factors to the BGL parametrization (41) alongside the Belle data of the calculated differential decay rates where the expansion coefficients are shared parameters and is an additional parameter determined from the fit. The expression (40) is used for . We calculate and by analytically integrating over two other decay angles and by numerically integrating over via Simpson’s rule with 800 steps within each bin as in the Belle paper Waheed et al. (2019).
. 0.02896(84) 0.01196(18) 0.002003(30) 0.0489(14) 0.050(23) 0.0227(92) 0.0040(14) 0.018(24) 1.0(1.9) 0.46(24) 0.040(27) 1.0(2.0)




Our results for the expansion coefficients are summarized in Table 9. These are consistent with those in Table 7, which are obtained without the experimental data, with slightly smaller uncertainty. We obtain
(53) |
where the first error comes from our main fit with input-A in Table 5, perturbative in Table 6, and . The second error is obtained by testing the stability of the fit against the different choices of the input and as in Table 8. This is in good agreement with the previous determination from the exclusive decay (5). In contrast to previous lattice studies, as shown in Fig. 19, we observe consistency between our lattice and Belle data leading to an acceptable value of . Towards the resolution of the tension, we need a more detailed analysis taking care of, for instance, bias in discussed in Refs. Gambino et al. (2019); Ferlewicz et al. (2021) due to D’Agostini effects D’Agostini (1994) and the strong systematic correlation of experimental data. That analysis is left for future work.
V Conclusion
In this article, we present our calculation of the semileptonic form factors in 2+1-flavor lattice QCD. The Möbius domain-wall action is employed for all relevant quark flavors to simulate lattices of cutoffs up to 4.5 GeV with chiral symmetry preserved to good accuracy. This removes the leading discretization errors and explicit renormalization is not necessary to calculate the SM form factors through the correlator ratios. The ground state saturation of the correlator ratios is carefully studied by simulating four source-sink separations. The bottom quark mass is limited to to suppress and higher discretization error. Note also that the chiral logarithm is suppressed by the small – mass splitting squared. As a result, the lattice spacing and quark mass dependence turns out to be mild in our simulation region leading to good control of the continuum and chiral extrapolation. We find finite size effects to be small by direct simulations of two volumes at our smallest pion mass.
One of the main results is the synthetic data of the form factors , , and in the relativistic convention at three reference values of the recoil parameter , 1.060 and 1.100, which can be used in future analyses to determine and to search for new physics. Through the BGL parametrization, we obtain in the SM. A combined analysis with the Belle data yields . These are consistent with previous lattice studies. However, the expansion coefficients for , and show significant tension among lattice studies. In particular, our slope for is slightly larger and consistent with a phenomenological analysis of the Belle data. As a result, our data show better consistency with the Belle data in the differential decay rates.
High statistics simulations on finer lattices would be helpful in resolving this tension. Simulating larger recoils can lead to a better estimate of as well as more detailed comparison with experiment in a wide range of the recoil parameter to search for new physics. In order to resolve the tension, it would also be very helpful to study the inclusive decay on the lattice for more direct comparison of the exclusive and inclusive analyses in the same simulations Hashimoto (2017). While it is an ill-posed problem to calculate the relevant hadronic tensor from discrete lattice data on a finite volume, a joint project of the JLQCD and RBC/UKQCD Collaborations along a strategy proposed in Ref. Hansen et al. (2017); Gambino and Hashimoto (2020) is in progress Barone et al. (2023a); Kellermann et al. (2023); Barone et al. (2023b) towards a realistic study of .
Acknowledgements.
We thank F. Bernlochner, B. Dey, D. Ferlewicz, P. Gambino, S. Iguro, M. Jung, T. Kitahara, P. Urquijo and E. Waheed for helpful discussions. This work used computational resources of supercomputer Fugaku provided by the RIKEN Center for Computational Science and Oakforest-PACS provided by JCAHPC through the HPCI System Research Projects (Project IDs: hp170106, hp180132, hp190118, hp210146, hp220140 and hp230122) and Multidisciplinary Cooperative Research Program (MCRP) in Center for Computational Sciences, University of Tsukuba (Project ID: xg18i016), SX-Aurora TSUBASA at the High Energy Accelerator Research Organization (KEK) under its Particle, Nuclear and Astrophysics Simulation Program (Project IDs: 2019L003, 2020-006, 2021-007, 2022-006 and 2023-004), Wisteria/BDEC-01 Odyssey at the University of Tokyo provided by MCRP (Project ID: Wo22i049), and Yukawa Institute Computer Facility. This work is supported in part by JSPS KAKENHI Grant Numbers JP18H01216, JP21H01085, JP22H01219, JP22H00138 and JP22K21347, by “Program for Promoting Researches on the Supercomputer Fugaku” (Simulation for basic science: approaching the new quantum era), by the Toshiko Yuasa France Japan Particle Physics Laboratory (TYL-FJPPL project FLAV_03 and FLAV_06), and by Joint Institute for Computational Fundamental Science (JICFuS). J.K. acknowledges support by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program through Grant Agreement No. 771971-SIMDAMA.References
- Amhis et al. (2023) Y. S. Amhis et al. (Heavy Flavor Averaging Group), Phys. Rev. D 107, 052008 (2023), eprint arXiv:2206.07501 [hep-ex].
- Crivellin and Pokorski (2015) A. Crivellin and S. Pokorski, Phys. Rev. Lett. 114, 011802 (2015), eprint arXiv:1407.1320 [hep-ph].
- Vaquero (2022) A. Vaquero, PoS LATTICE2022, 250 (2022), eprint arXiv:2212.10217 [hep-lat].
- Kaneko (2023) T. Kaneko, PoS LATTICE2022, 238 (2023), eprint arXiv:2304.01618 [hep-lat].
- Bernard et al. (2009) C. Bernard et al. (Fermilab/MILC Collaboration), Phys. Rev. D 79, 014506 (2009), eprint arXiv:0808.2519 [hep-lat].
- Bailey et al. (2014) J. A. Bailey et al. (Fermilab/MILC Collaboration), Phys. Rev. D 89, 114504 (2014), eprint arXiv:1403.0635 [hep-lat].
- Harrison et al. (2018) J. Harrison, C. Davies, and M. Wingate (HPQCD Collaboration), Phys. Rev. D 97, 054502 (2018), eprint arXiv:1711.11013 [hep-lat].
- Bernlochner et al. (2017) F. U. Bernlochner, Z. Ligeti, M. Papucci, and D. J. Robinson, Phys. Rev. D 95, 115008 (2017), [Erratum: Phys.Rev.D 97, 059902 (2018)], eprint arXiv:1703.05330 [hep-ph].
- Bigi et al. (2017a) D. Bigi, P. Gambino, and S. Schacht, Phys. Lett. B 769, 441 (2017a), eprint arXiv:1703.06124 [hep-ph].
- Grinstein and Kobach (2017) B. Grinstein and A. Kobach, Phys. Lett. B 771, 359 (2017), eprint arXiv:1703.08170 [hep-ph].
- Boyd et al. (1997) C. G. Boyd, B. Grinstein, and R. F. Lebed, Phys. Rev. D 56, 6895 (1997), eprint arXiv:hep-ph/9705252.
- Bigi et al. (2017b) D. Bigi, P. Gambino, and S. Schacht, JHEP 11, 061 (2017b), eprint arXiv:1707.09509 [hep-ph].
- Bernlochner et al. (2019) F. U. Bernlochner, Z. Ligeti, and D. J. Robinson, Phys. Rev. D 100, 013005 (2019), eprint arXiv:1902.09553 [hep-ph].
- Gambino et al. (2019) P. Gambino, M. Jung, and S. Schacht, Phys. Lett. B 795, 386 (2019), eprint arXiv:1905.08209 [hep-ph].
- Ferlewicz et al. (2021) D. Ferlewicz, P. Urquijo, and E. Waheed, Phys. Rev. D 103, 073005 (2021), eprint arXiv:2008.09341 [hep-ph].
- Waheed et al. (2019) E. Waheed et al. (Belle Collaboration), Phys. Rev. D 100, 052007 (2019), [Erratum: Phys.Rev.D 103, 079901 (2021)], eprint arXiv:1809.03290 [hep-ex].
- Abdesselam et al. (2017) A. Abdesselam et al. (Belle Collaboration) (2017), eprint arXiv:1702.01521 [hep-ex].
- Bazavov et al. (2022) A. Bazavov et al. (Fermilab/MILC Collaboration), Eur. Phys. J. C 82, 1141 (2022), [Erratum: Eur.Phys.J.C 83, 21 (2023)], eprint arXiv:2105.14019 [hep-lat].
- El-Khadra et al. (1997) A. X. El-Khadra, A. S. Kronfeld, and P. B. Mackenzie, Phys. Rev. D 55, 3933 (1997), eprint arXiv:hep-lat/9604004.
- Harrison and Davies (2024) J. Harrison and C. T. H. Davies (HPQCD, (HPQCD Collaboration)), Phys. Rev. D 109, 094515 (2024), eprint arXiv:2304.03137.
- Brower et al. (2017) R. C. Brower, H. Neff, and K. Orginos, Comput. Phys. Commun. 220, 1 (2017), eprint arXiv:1206.5214 [hep-lat].
- Hashimoto et al. (1999) S. Hashimoto, A. X. El-Khadra, A. S. Kronfeld, P. B. Mackenzie, S. M. Ryan, and J. N. Simone, Phys. Rev. D 61, 014502 (1999), eprint arXiv:hep-ph/9906376.
- Kaneko et al. (2018) T. Kaneko, Y. Aoki, B. Colquhoun, H. Fukaya, and S. Hashimoto (JLQCD Collaboration), PoS LATTICE2018, 311 (2018), eprint arXiv:1811.00794 [hep-lat].
- Kaneko et al. (2019) T. Kaneko, Y. Aoki, G. Bailas, B. Colquhoun, H. Fukaya, S. Hashimoto, and J. Koponen (JLQCD Collaboration), PoS LATTICE2019, 139 (2019), eprint arXiv:1912.11770 [hep-lat].
- Kaneko et al. (2022) T. Kaneko, Y. Aoki, B. Colquhoun, M. Faur, H. Fukaya, S. Hashimoto, J. Koponen, and E. Kou (JLQCD Collaboration), PoS LATTICE2021, 561 (2022), eprint arXiv:2112.13775 [hep-lat].
- Colquhoun et al. (2022) B. Colquhoun, S. Hashimoto, T. Kaneko, and J. Koponen (JLQCD Collaboration), Phys. Rev. D 106, 054502 (2022), eprint arXiv:2203.04938 [hep-lat].
- Borsanyi et al. (2012) S. Borsanyi et al., JHEP 09, 010 (2012), eprint arXiv:1203.4469 [hep-lat].
- Weisz (1983) P. Weisz, Nucl. Phys. B 212, 1 (1983).
- Weisz and Wohlert (1984) P. Weisz and R. Wohlert, Nucl. Phys. B 236, 397 (1984), [Erratum: Nucl.Phys.B 247, 544 (1984)].
- Luscher and Weisz (1985) M. Luscher and P. Weisz, Commun. Math. Phys. 97, 59 (1985), [Erratum: Commun.Math.Phys. 98, 433 (1985)].
- Boyle (2015) P. A. Boyle (RBC/UKQCD Collaboration), PoS LATTICE2014, 087 (2015).
- Morningstar and Peardon (2004) C. Morningstar and M. J. Peardon, Phys. Rev. D 69, 054501 (2004), eprint arXiv:hep-lat/0311018.
- Aoki et al. (2008a) S. Aoki et al. (JLQCD Collaboration), Phys. Rev. D 78, 014508 (2008a), eprint arXiv:0803.3197 [hep-lat].
- Aoki et al. (2012) S. Aoki et al. (JLQCD and TWQCD Collaborations), PTEP 2012, 01A106 (2012).
- Kaneko et al. (2014) T. Kaneko, S. Aoki, G. Cossu, H. Fukaya, S. Hashimoto, and J. Noaki (JLQCD Collaboration), PoS LATTICE2013, 125 (2014), eprint arXiv:1311.6941 [hep-lat].
- Randall and Wise (1993) L. Randall and M. B. Wise, Phys. Lett. B 303, 135 (1993), eprint arXiv:hep-ph/9212315.
- Savage (2002) M. J. Savage, Phys. Rev. D 65, 034014 (2002), eprint arXiv:hep-ph/0109190.
- Nakayama et al. (2016) K. Nakayama, B. Fahy, and S. Hashimoto (JLQCD Collaboration), Phys. Rev. D 94, 054507 (2016), eprint arXiv:1606.01002 [hep-lat].
- Aoki et al. (2008b) S. Aoki et al. (JLQCD, TWQCD), Phys. Lett. B 665, 294 (2008b), eprint arXiv:0710.1130 [hep-lat].
- Fukaya et al. (2007a) H. Fukaya et al. (JLQCD), Phys. Rev. Lett. 98, 172001 (2007a), eprint hep-lat/0702003.
- Fukaya et al. (2007b) H. Fukaya, S. Aoki, T. W. Chiu, S. Hashimoto, T. Kaneko, H. Matsufuru, J. Noaki, K. Ogawa, T. Onogi, and N. Yamada (TWQCD), Phys. Rev. D 76, 054503 (2007b), eprint 0705.3322.
- Brower et al. (2003) R. Brower, S. Chandrasekharan, J. W. Negele, and U. J. Wiese, Phys. Lett. B 560, 64 (2003), eprint arXiv:hep-lat/0302005.
- Aoki et al. (2007) S. Aoki, H. Fukaya, S. Hashimoto, and T. Onogi, Phys. Rev. D 76, 054508 (2007), eprint arXiv:0707.0396 [hep-lat].
- Aoki et al. (2017) S. Aoki, G. Cossu, X. Feng, H. Fukaya, S. Hashimoto, T. Kaneko, J. Noaki, and T. Onogi (JLQCD Collaboration), Phys. Rev. D 96, 034501 (2017), eprint arXiv:1705.00884 [hep-lat].
- Fahy et al. (2016) B. Fahy, G. Cossu, and S. Hashimoto (JLQCD Collaboration), PoS LATTICE2016, 118 (2016), eprint arXiv:1702.02303 [hep-lat].
- Detmold et al. (2012) W. Detmold, C. J. D. Lin, and S. Meinel, Phys. Rev. D 85, 114508 (2012), eprint arXiv:1203.3378 [hep-lat].
- Becirevic and Sanfilippo (2013) D. Becirevic and F. Sanfilippo, Phys. Lett. B 721, 94 (2013), eprint arXiv:1210.5410 [hep-lat].
- Can et al. (2013) K. U. Can, G. Erkol, M. Oka, A. Ozpineci, and T. T. Takahashi, Phys. Lett. B 719, 103 (2013), eprint arXiv:1210.0869 [hep-lat].
- Bernardoni et al. (2015) F. Bernardoni, J. Bulava, M. Donnellan, and R. Sommer (ALPHA Collaboration), Phys. Lett. B 740, 278 (2015), eprint arXiv:1404.6951 [hep-lat].
- Flynn et al. (2016) J. M. Flynn, P. Fritzsch, T. Kawanai, C. Lehner, B. Samways, C. T. Sachrajda, R. S. Van de Water, and O. Witzel (RBC/UKQCD Collaboration), Phys. Rev. D 93, 014510 (2016), eprint arXiv:1506.06413 [hep-lat].
- Noaki et al. (2008) J. Noaki et al. (JLQCD and TWQCD Collaborations), Phys. Rev. Lett. 101, 202004 (2008), eprint arXiv:0806.0894 [hep-lat].
- Falk et al. (1990) A. F. Falk, H. Georgi, B. Grinstein, and M. B. Wise, Nucl. Phys. B 343, 1 (1990).
- Falk and Grinstein (1990) A. F. Falk and B. Grinstein, Phys. Lett. B 249, 314 (1990).
- Neubert (1992) M. Neubert, Nucl. Phys. B 371, 149 (1992).
- Luke (1990) M. E. Luke, Phys. Lett. B 252, 447 (1990).
- Stein (1956) C. Stein, Proceedings of the Third Berkeley Symposium on Mathematical Statistics and Probability, Volume 1: Contributions to the Theory of Statistics (University of California Press, Berkeley, CA) p. 197 (1956).
- Ledoit and Wolf (2004) O. Ledoit and M. Wolf, J. Multivariate Anal. 88, 365 (2004).
- Caprini et al. (1998) I. Caprini, L. Lellouch, and M. Neubert, Nucl. Phys. B 530, 153 (1998), eprint arXiv:hep-ph/9712417.
- Patrignani et al. (2016) C. Patrignani et al. (Particle Data Group), Chin. Phys. C 40, 100001 (2016).
- Dowdall et al. (2012) R. J. Dowdall, C. T. H. Davies, T. C. Hammant, and R. R. Horgan (HPQCD Collaboration), Phys. Rev. D 86, 094510 (2012), eprint arXiv:1207.5149 [hep-lat].
- Colquhoun et al. (2015) B. Colquhoun, C. T. H. Davies, R. J. Dowdall, J. Kettle, J. Koponen, G. P. Lepage, and A. T. Lytle (HPQCD Collaboration), Phys. Rev. D 91, 114509 (2015), eprint arXiv:1503.05762 [hep-lat].
- Godfrey (2004) S. Godfrey, Phys. Rev. D 70, 054017 (2004), eprint arXiv:hep-ph/0406228.
- Rai and Devlani (2013) A. K. Rai and N. Devlani, PoS Hadron2013, 045 (2013).
- Eichten and Quigg (1994) E. J. Eichten and C. Quigg, Phys. Rev. D 49, 5845 (1994), eprint arXiv:hep-ph/9402210.
- Workman et al. (2022) R. L. Workman et al. (Particle Data Group), PTEP 2022, 083C01 (2022).
- Li et al. (2023) X.-J. Li, Y.-S. Li, F.-L. Wang, and X. Liu, Eur. Phys. J. C 83, 1080 (2023), eprint 2308.07206.
- Martinelli et al. (2021) G. Martinelli, S. Simula, and L. Vittorio, Phys. Rev. D 104, 094512 (2021), eprint arXiv:2105.07851 [hep-lat].
- Grigo et al. (2012) J. Grigo, J. Hoff, P. Marquard, and M. Steinhauser, Nucl. Phys. B 864, 580 (2012), eprint arXiv:1206.3418 [hep-ph].
- Bigi and Gambino (2016) D. Bigi and P. Gambino, Phys. Rev. D 94, 094008 (2016), eprint arXiv:1606.08030 [hep-ph].
- Aoki et al. (2020) S. Aoki et al. (Flavour Lattice Averaging Group), Eur. Phys. J. C 80, 113 (2020), eprint arXiv:1902.08191 [hep-lat].
- D’Agostini (1994) G. D’Agostini, Nucl. Instrum. Meth. A 346, 306 (1994).
- Hashimoto (2017) S. Hashimoto, PTEP 2017, 053B03 (2017), eprint arXiv:1703.01881 [hep-lat].
- Hansen et al. (2017) M. T. Hansen, H. B. Meyer, and D. Robaina, Phys. Rev. D 96, 094513 (2017), eprint arXiv:1704.08993 [hep-lat].
- Gambino and Hashimoto (2020) P. Gambino and S. Hashimoto, Phys. Rev. Lett. 125, 032001 (2020), eprint arXiv:2005.13730 [hep-lat].
- Barone et al. (2023a) A. Barone, A. Jüttner, S. Hashimoto, T. Kaneko, and R. Kellermann, PoS LATTICE2022, 403 (2023a), eprint arXiv:2211.15623 [hep-lat].
- Kellermann et al. (2023) R. Kellermann, A. Barone, S. Hashimoto, A. Jüttner, and T. Kaneko, PoS LATTICE2022, 414 (2023), eprint arXiv:2211.16830 [hep-lat].
- Barone et al. (2023b) A. Barone, S. Hashimoto, A. Jüttner, T. Kaneko, and R. Kellermann, JHEP 07, 145 (2023b), eprint arXiv:2305.14092 [hep-lat].
- Prim et al. (2023) M. T. Prim et al. (Belle Collaboration), Phys. Rev. D 108, 012002 (2023), eprint arXiv:2301.07529.
- Abudinén et al. (2023) F. Abudinén et al. (Belle II Collaboration) (2023), eprint arXiv:2301.04716 [hep-ex].