1 Introduction

The observed deviations from Standard Model (SM) predictions in semileptonic B-meson decays persist as some of the most significant experimental hints for the presence of possible New Physics (NP) beyond the SM at the TeV scale. One set of deviations regards Lepton Flavor Universality (LFU) ratios of charged-current semileptonic B decays between the third and lighter lepton families, \(R(D^{(*)}) = {\mathcal {B}}(B \rightarrow D^{(*)} \tau \nu ) / {\mathcal {B}}(B \rightarrow D^{(*)} \ell \nu )\) [1,2,3,4,5,6,7,8,9,10,11], with a combined significance of approximately \(3\sigma \) [12]. Another set of deviations from the SM are observed in LFU ratios between second and first lepton families in neutral-current B decays, \(R_{K^{(*)}} = {\mathcal {B}}(B \rightarrow K^{(*)} \mu ^+ \mu ^-) / {\mathcal {B}}(B \rightarrow K^{(*)} e^+ e^-)\) [13,14,15,16,17], as well as in \(B_s \rightarrow \mu ^+ \mu ^-\), in angular observables of the \(B \rightarrow K^* \mu ^+ \mu ^-\) process as well as in branching ratios of other decay processes which involve the \(b \rightarrow s \mu ^+ \mu ^-\) transition [18,19,20,21,22,23]. The global significance of these deviations, obtained with very conservative estimates of SM uncertainties, is \(3.9\sigma \) [24]. On the other hand, global fits show that along the preferred directions in effective field theory (EFT) space the pulls from the SM can be even up to \(5-7\sigma \) [25,26,27,28,29,30,31].

When attempting to address at the same time both sets of anomalies, leptoquark (LQ) mediators are by far the preferred candidates. This is mainly due to the fact that, while semileptonic operators required for the B-anomalies can be induced at the tree-level, four-quark and four-lepton operators, that are strongly constrained by meson mixing or LFV, are induced only at one loop and thus automatically suppressed.

An interesting scenario for a combined explanation of the anomalies involves the two scalar LQs \(S_1 = ({ {{\bar{\mathbf{3}}}}}, \mathbf{1}, 1/3)\) and \(S_3 = ({{{\bar{\mathbf{3}}}}}, \mathbf{3}, 1/3)\) [32,33,34,35,36,37,38,39,40,41,42]. Interestingly enough, the \(S_1\) couplings to right-handed fermions allow also an explanation of the observed deviation in the muon anomalous magnetic moment \((g-2)_\mu \) [43,44,45]. As a possible ultraviolet (UV) completion for this model, the two scalars could arise, together with the Higgs boson, as pseudo-Nambu-Goldstone bosons from a new strongly coupled sector at the multi-TeV scale, which would also address the hierarchy problem of the electroweak scale in a composite Higgs framework [34, 41] (see also [46,47,48] for related works).

The solution to B-anomalies involves NP couplings to second and third generations of quarks and leptons, while the couplings to first generation could in principle be very small, which is also required by strong experimental constraints [49,50,51,52,53]. Nevertheless, the natural expectation is that NP should couple to all generations, possibly with some flavor structure dictated by a dynamical mechanism or a symmetry. In light of this, a question one can pose is: what are the expected effects in Kaon physics and electron observables for models that address the B-anomalies? This is the main question we aim at addressing with this paper, in the context of \(S_1 + S_3\) solutions. Analyses in the same spirit were performed in Refs. [54, 55], in an EFT context, and in Ref. [56] for single leptoquarks and in connection with \(R(K^{(*)})\) only. Specifically, the golden channels of rare Kaon decays, \(K^+ \rightarrow \pi ^+ \nu \nu \) and \(K_L \rightarrow \pi ^0 \nu \nu \), are now actively being investigated by the NA62 [57] and KOTO [58] experiments, respectively, and substantial improvements are expected in the future. Furthermore, an improvement by a few orders of magnitude in sensitivity is expected for \(\mu \rightarrow e\) conversion experiments COMET and Mu2e [59,60,61,62] as well as in \(\mu \rightarrow 3 e\) from the Mu3e experiment at PSI [63]. Can these experiments expect to observe a signal, possible related to the B-anomalies?

Our starting point is the analysis of how this setup can address the B-anomalies (as well as the \((g-2)_\mu \)) performed in Ref. [40]. In that work several scenarios were considered and, for each of them, a global analysis was performed including all the relevant observables computed at one-loop accuracy using the complete one-loop matching between the two mediators and the SMEFT obtained in Ref. [64]. In particular, two scenarios able to address both sets of B-anomalies were found:

  • LH couplings. If the \(S_1\) couplings to right-handed fermions are zero, then both charged and neutral-current B-anomalies can be addressed, while the muon magnetic moment deviation cannot. It was also observed that the preferred values of the couplings to second generation was compatible with the structure hinted to by an approximate \(U(2)^5\) flavor symmetry [40] (see also [33, 34]).

  • All couplings. If, instead, all couplings are allowed, then both the B-anomalies and the \((g-2)_\mu \) can be addressed, but the coupling structure is not compatible with the \(U(2)^5\) flavor symmetry since a large coupling to \(c_R \tau _R\) is required, with a small coupling to \(t_R \tau _R\) instead [40].

We extend our previous work by considering also Kaon and D decays, as well as all processes sensitive to the \(\mu \rightarrow e\) lepton flavor violating (LFV) transition.

To find possible connections between B-anomalies and Kaon physics we take two different approaches for the two scenarios listed above. For the first scenario we impose from the beginning a concrete assumption on the flavor structure, in particular the \(U(2)^5\) flavor symmetry [65,66,67] and perform the first complete study of B-anomalies and Kaon physics with \(S_1\) and \(S_3\) within this context. The main feature of this approximate symmetry is that strict relations between the LQ coupling to first and second generations are predicted, implying that Kaon decays become strictly connected with B decays.

Our choice of the flavor symmetry is motivated by the observation that the approximate \(U(2)_q \times U(2)_{\ell }\) flavor symmetry, that acts on the light-generations of SM quarks and leptons and is a subset of \(U(2)^5\), appears to provide a consistent picture of all low-energy data. In fact, it was shown in the general EFT context [33, 68,69,70,71], as well as also in concrete UV realizations [34, 40, 72,73,74,75,76,77], that not only it can reproduce the observed hierarchies in the SM Yukawa sector, but also successfully control the strength of NP couplings allowing for sufficiently large effects in processes involving third-generation fermions. We mention that the flavor symmetry does not need to be a fundamental property of the UV theory, but it could arise as an accidental low-energy symmetry.

For the second scenario, since no evident flavor structure emerges from the B-anomalies fit when also right-handed couplings are included, we let vary the couplings relevant for \(K \rightarrow \pi \nu \nu \), while keeping the other couplings fixed to the best-fit values required by the B-anomalies and muon magnetic moment. In this way we can find the allowed values for \(K \rightarrow \pi \nu \nu \) that are compatible with the B-anomalies, in this general setup.

The paper is structured as follows. In Sect. 2 we briefly introduce the model, the setup, and the statistical tool used for the analysis. In Sect. 3 we discuss the structure of the LQ couplings predicted by the minimally broken \(U(2)^5\) flavor symmetry and study the results of the global fit in this framework. Section 4 is instead devoted to the study of rare Kaon decays and electron LFV processes in the general case where the flavor assumption is lifted. We conclude in Sect. 5. Details on the observables not discussed in [40] are collected in Appendix A.

2 Setup

We consider in this work the two scalar LQs \(S_1 = ({ {{\bar{\mathbf{3}}}}}, \mathbf{1}, 1/3)\) and \(S_3 = ({ {{\bar{\mathbf{3}}}}}, \mathbf{3}, 1/3)\), where the quantum numbers under the SM gauge group \(SU(3)_c \times SU(2)_L \times U(1)_Y\) are indicated. The Lagrangian to be added to the SM one, assuming baryon and lepton number conservation, is

$$\begin{aligned} {\mathcal {L}}_{\text {LQ}}= & {} |D_\mu S_1|^2 + |D_\mu S_3|^2 - M_{1}^2 |S_1|^2 \nonumber \\&- M_{3} ^2 |S_3|^2 - V_{\text {LQ}}(S_1, S_3, H)+\nonumber \\&+ \left( (\lambda ^{1L})_{i\alpha } \, \overline{q^c}_i \,\epsilon \, \ell _\alpha + (\lambda ^{1R})_{i\alpha } \, \overline{u^c}_i e_\alpha \right) S_1 \nonumber \\&+ (\lambda ^{3L})_{i\alpha } \, \overline{q^c}_i \,\epsilon \, \sigma ^I \ell _\alpha S_3^I + \text {h.c.} ~, \end{aligned}$$
(2.1)

where \(\epsilon =i\sigma _2\), \((\lambda ^{1L})_{i\alpha }, (\lambda ^{1R})_{i\alpha }, (\lambda ^{3L})_{i\alpha } \in {\mathbb {C}}\), and \(V_{\text {LQ}}\) includes LQ self-couplings and interactions with the Higgs boson, which are omitted since they are not relevant for the phenomenology studied here. We denote SM quark and lepton fields by \(q_i\), \(u_i\), \(d_i\), \(\ell _\alpha \), and \(e_\alpha \), while the Higgs doublet is H. We adopt latin letters (\(i,\,j,\,k,\,\dots \)) for quark flavor indices and greek letters (\(\alpha ,\,\beta ,\,\gamma ,\,\dots \)) for lepton flavor indices. We work in the down-quark and charged-lepton mass eigenstate basis, where

$$\begin{aligned} q_i = \left( \begin{array}{c} V^*_{ji} u^j_L \\ d^i_L \end{array} \right) \,, \qquad \ell _\alpha = \left( \begin{array}{c} \nu ^\alpha _L \\ e^\alpha _L \end{array} \right) ~, \end{aligned}$$
(2.2)

and V is the CKM matrix. We use the same conventions as Refs. [40, 64], to which we refer for further details.

Our goal is to study the phenomenology of the \(S_1+S_3\) model extending the list of observables already accounted for in Ref. [40] to all the other relevant Kaon, D-meson, and electron LFV ones. We do this by employing the same multistep procedure: the LQ model is matched at one-loop level into the SMEFT [64], which is then matched into the Low Energy EFT (LEFT) [78, 79]; within any EFT the renormalization group evolution (RGE) is taken into account, as well as the one-loop rational contributions within the LEFT, in terms of whose coefficients the observables and pseudo-observables are expressed (barring the case of observables that are measured at the electroweak scale).

In our global analysis for the two LQs we add the following observables to those observables already studied in Ref. [40]: rare and LFV Kaon decays, \(\epsilon _K'/\epsilon _K\), rare D-meson decays, \(b\rightarrow d\ell \ell \) decays, \(\mu \rightarrow e\) conversion in nuclei, and the neutron EDM. In Tables 1, 2, 3, and 4, we show the complete list of observables that we analyze, together with their SM predictions and experimental bounds. In Appendix A we collect details on the low-energy observables that were not considered in Ref. [40], together with others which turn out to set relevant constraints in the fit, such as limits from Z couplings measurements and \(K\rightarrow \pi \nu \nu \), that we repeat for sake of completeness. For all the observables, the full set of one-loop corrections is considered in the numerical analysis.

We perform a \(\chi ^2\) fit, thus defining the likelihood as

$$\begin{aligned} - 2 \log {\mathcal {L}}\equiv \chi ^2(\lambda _x, M_x) = \sum _i \frac{ \left( {\mathcal {O}}_i(\lambda _x, M_x) - \mu _i\right) ^2}{\sigma _i^2},\quad \end{aligned}$$
(2.3)

where \({\mathcal {O}}_i(\lambda _x, M_x)\) is the expression of the observable as function of the model parameters, \(\mu _i\) its central measured value, and \(\sigma _i\) the associated standard deviation. The correlations between \(\Delta C_9^{sb\mu \mu }\) and \(C_9^\mathrm{univ}\), as well as between \(R_{D}\) and \(R_{D^*}\) are taken into account (see gray lines in the plots of Fig. 1). In the analysis presented in this paper, 73 observables are taken into account, for which, within the SM, the \(\chi ^2\) is \(\chi ^2_\mathrm{SM}=104.0\). In each scenario we first find the best-fit point by minimizing the \(\chi ^2\). We then perform a numerical scan over the parameter space using a Markov Chain Monte Carlo algorithm (Hastings-Metropolis), to select points that are within the 68 or 95% CL from the best-fit point, with final samples of size \({\mathcal {O}}(10^4)\). These scans are used to obtain preferred regions in parameter space or for selected pairs of interesting observables by projecting the obtained points onto the corresponding plane, which corresponds to profiling over the parameters not plotted.

3 Scalar leptoquarks and \(U(2)^5\) flavor symmetry

In the limit where only third generaton fermions are massive, the SM enjoys the global flavor symmetry [65,66,67]

$$\begin{aligned} G_F = U(2)_q \times U(2)_\ell \times U(2)_u \times U(2)_d \times U(2)_e~. \end{aligned}$$
(3.1)

Masses of the first two generations of fermions and their mixing break this symmetry. In the quark sector the largest breaking is of size \(\epsilon \approx y_t |V_{ts}| \approx 0.04\) [71]. Formally, the symmetry breaking terms in the Yukawa matrices can be described in terms of spurions transforming under representations of \(G_F\). The minimal set of spurions that can reproduce the observed masses and mixing angles isFootnote 1

$$\begin{aligned} \begin{aligned}&\mathbf{V}_q \sim (\mathbf{2}, \mathbf{1}, \mathbf{1}, \mathbf{1}, \mathbf{1})~, \quad \mathbf{V}_\ell \sim (\mathbf{1}, \mathbf{2}, \mathbf{1}, \mathbf{1}, \mathbf{1})~, \\&{{\varvec{\Delta }}}_u \sim (\mathbf{2}, \mathbf{1}, {{{\bar{\mathbf{2}}}}}, \mathbf{1}, \mathbf{1})~, \quad {{\varvec{\Delta }}}_d \sim (\mathbf{2}, \mathbf{1}, \mathbf{1}, {{{\bar{\mathbf{2}}}}}, \mathbf{1})~,\\&\quad {{\varvec{\Delta }}}_e \sim (\mathbf{1}, \mathbf{2}, \mathbf{1}, \mathbf{1}, {{{\bar{\mathbf{2}}}}})~. \end{aligned}\end{aligned}$$
(3.2)

In terms of these spurions the SM Yukawa matrices can be written as

$$\begin{aligned} Y_{u(d)}= & {} y_{t(b)} \left( \begin{array}{c c} {{\varvec{\Delta }}}_{u(d)} &{} x_{t(b)} \mathbf{V}_q \\ 0 &{} 1 \end{array}\right) ~,\nonumber \\ Y_{e}= & {} y_\tau \left( \begin{array}{c c} {{\varvec{\Delta }}}_{e} &{} x_\tau \mathbf{V}_\ell \\ 0 &{} 1 \end{array}\right) ~, \qquad \end{aligned}$$
(3.3)

with \(x_{t,b,\tau }\) are \({\mathcal {O}}(1)\) complex numbers, \({ {\varvec{\Delta }}}\)’s are \(2\times 2\) matrices, and \(\mathbf{V}_{q,\ell }\) are 2-component vectors.

In the context of the B-anomalies, this flavor symmetry was introduced as a possible explanation for the LFU breaking hints, that point to largest effects for \(\tau \) leptons, smaller for muons, and even smaller for electrons. Furthermore, it was observed in Refs. [33, 34, 40, 68,69,70,71] that the LQ couplings to second and third generations, required to fit the anomalies, were consistent with the expectations given by this symmetry. In this Section we study if, indeed, a complete implementation of \(U(2)^5\) flavor symmetry for the \(S_1\) and \(S_3\) scalar LQs, including the couplings to first generation fermions, is consistent with the observed anomalies.

In the same flavor basis used to write the Yukawa couplings of Eq. (3.3), the \(S_1\) and \(S_3\) LQ couplings have the following structure:

$$\begin{aligned} \lambda ^{1(3) L}= & {} \lambda ^{1(3)} \left( \begin{array}{c c} {\tilde{x}}^{1(3) L}_{q \ell } \mathbf{V}_q^* \times \mathbf{V}_\ell ^\dagger &{} {\tilde{x}}^{1(3) L}_{q} \mathbf{V}_q^* \\ {\tilde{x}}^{1(3) L}_{\ell } \mathbf{V}_\ell ^\dagger &{} {\tilde{x}}^{1(3) L}_{b\tau } \end{array}\right) ~, \end{aligned}$$
(3.4)
$$\begin{aligned} \lambda ^{1 R}= & {} \lambda ^{1}_R \left( \begin{array}{c c} {\mathcal {O}}({{\varvec{\Delta }}}_u \mathbf{V}_q {{\varvec{\Delta }}}_e \mathbf{V}_\ell ) &{} {\tilde{x}}^{1 R}_{u} {{\varvec{\Delta }}}_u^\dagger \mathbf{V}_q^* \\ {\tilde{x}}^{1 R}_{e} \mathbf{V}_\ell ^\dagger {{\varvec{\Delta }}}_e^* &{} {\tilde{x}}^{1 R}_{t \tau } \end{array}\right) \nonumber \\\approx & {} \lambda ^{1}_R \left( \begin{array}{c c} 0 &{} 0 \\ 0 &{} {\tilde{x}}^{1 R}_{t \tau } \end{array}\right) ~, \end{aligned}$$
(3.5)

where \(\lambda ^{1(3)}\) and \(\lambda ^1_R\) are overall couplings, all \({\tilde{x}}\) are \({\mathcal {O}}(1)\) parameters, and in the last step in \(\lambda ^{1R}\) we neglected all the terms that give too small couplings to have a significant influence to our observables. In the following we can thus neglect the presence of the \(\lambda ^{1R}\) couplings in the \(U(2)^5\) scenario since the \(t_R \tau _R\) coupling does not affect in a relevant way the phenomenology.

By diagonalizing the SM lepton and down-quark Yukawa matrices one can put in relation some of the parameters in Eq. (3.3) with observed masses and CKM elements, we refer to Ref. [71] for a detailed discussion on this procedure. For our purposes, the main result is that in this basis the quark doublet spurion is fixed by the CKM up to an overall \({\mathcal {O}}(1)\) factor, \(\mathbf{V}_q = \kappa _q (V_{td}^*, V_{ts}^*)^T\), while the size of the leptonic doublet spurion \(V_\ell \equiv |\mathbf{V}_\ell |\) as well as the angle that rotates left-handed electrons and muons, \(s_e \equiv \sin \theta _e\), are free. The same rotations that diagonalize the (lepton and down quark) Yukawas also apply to the LQ couplings. The final result of this procedure is the following structure for the LQ couplings in the mass basis:

$$\begin{aligned} \lambda ^{1(3) L} = \lambda ^{1(3)} \left( \begin{array}{c c c} x^{1(3) }_{q \ell } s_e V_\ell V_{td} &{} x^{1(3) }_{q \ell } V_\ell V_{td} &{} x^{1(3) }_{q} V_{td} \\ x^{1(3) }_{q \ell } s_e V_\ell V_{ts} &{} x^{1(3) }_{q \ell } V_\ell V_{ts} &{} x^{1(3) }_{q} V_{ts} \\ x^{1(3) }_{\ell } s_e V_\ell &{} x^{1(3) }_{\ell } V_\ell &{} 1 \end{array}\right) . \nonumber \\ \end{aligned}$$
(3.6)

All \( x^{1(3)}\) parameters are expected to be \({\mathcal {O}}(1)\) complex numbers and we absorbed the \({\mathcal {O}}(1)\) coefficient in the 3-3 component inside the definition of \(\lambda ^{1(3)}\). It directly follows from Eq. (3.6) that the flavor symmetry imposes some strict relations between families:

$$\begin{aligned} \lambda ^{1(3) L}_{1 \alpha } = \lambda ^{1(3) L}_{2 \alpha } \frac{V_{td}}{V_{ts}}~, \qquad \lambda ^{1(3) L}_{i 1} = \lambda ^{1(3) L}_{i 2} s_e ~. \qquad \end{aligned}$$
(3.7)

For the two LQs we thus remain with the two overall couplings \(\lambda ^{1(3)}\), that we can always take to be positive, six \({\mathcal {O}}(1)\) complex parameters (\( x^{1(3) }_{q \ell }, x^{1(3) }_{q}, x^{1(3) }_{\ell }\)), one small angle \(s_e\) that regulates the couplings to electrons compared to the muon ones, and finally \(V_{\ell }\) that sets the size of muon couplings compared to tau ones. With the coupling structure of Eq. (3.6) the two leptoquarks decay dominantly to third generation fermions, since couplings to lighter generations are suppressed by the spurion factors.

Since \(S_1\) does not mediate \(d_i \rightarrow d_j \bar{\ell }_\alpha \ell _\beta \) at tree level, its contributions proportional to \({x}^1_\ell \) and \({x}^1_{q\ell }\) are only very weakly constrained. For this reason, to simplify the numerical scan we fix them to be equal to 1.Footnote 2

We provide here some simplified expressions for the most relevant NP effects in this setup, deferring for details to Appendix A:

$$\begin{aligned}&\frac{\Delta R(D^{(*)})}{R(D^{(*)})_\mathrm{SM}} \approx v^2 \left( 1.09 \frac{ |\lambda ^1|^2 (1 - x^{1 *}_q V^*_{tb} )}{2 M_1^2}\right. \nonumber \\&\left. \quad - 1.02 \frac{ |\lambda ^3|^2 (1 - x^{3 *}_q V^*_{tb} )}{2 M_3^2} \right) ~, \end{aligned}$$
(3.8)
$$\begin{aligned}&\Delta C_9^{sb\mu \mu } = - \Delta C_{10}^{sb\mu \mu } \approx \frac{\pi }{\sqrt{2} G_F \alpha V_{tb}} \frac{|\lambda ^3|^2 |V_\ell |^2 x^3_\ell x^{3*}_{q\ell }}{M_3^2}~, \end{aligned}$$
(3.9)
$$\begin{aligned}&\Delta C_9^{sd\mu \mu } = - \Delta C_{10}^{sd\mu \mu } \approx \frac{\pi V_{ts}^* V_{td}}{\sqrt{2} G_F \alpha } \frac{|\lambda ^3|^2 |V_\ell |^2 | x^{3}_{q\ell }|^2 }{M_3^2}~, \end{aligned}$$
(3.10)
$$\begin{aligned}&\left[ L^{VLL}_{\nu d} \right] _{\nu _\tau \nu _\tau s b} \approx V_{ts}^* \left( \frac{|\lambda ^1|^2 x^{1 *}_q}{2 M_1^2} + \frac{ |\lambda ^3|^2 x^{3 *}_q }{2 M_3^2} \right) ~, \end{aligned}$$
(3.11)
$$\begin{aligned}&\left[ L^{VLL}_{\nu d} \right] _{\nu _\tau \nu _\tau d s} \approx V_{td}^* V_{ts} \left( \frac{|\lambda ^1|^2 |x^1_q|^2}{2 M_1^2} + \frac{|\lambda ^3|^2 |x^3_q|^2}{2 M_3^2} \right) ~, \end{aligned}$$
(3.12)
$$\begin{aligned}&10^3 \delta g^Z_{\tau _L} \approx 0.59 \frac{|\lambda ^1|^2}{M_1^2 / \text { TeV}^2} + 0.80 \frac{|\lambda ^1|^2}{M_1^2 / \text { TeV}^2} ~, \end{aligned}$$
(3.13)
$$\begin{aligned}&C^1_K \approx \frac{V_{ts}^* V_{td}}{128 \pi ^2} \left( \frac{|\lambda ^1|^4 |x^1_q|^4}{M_3^2} + 5 \frac{|\lambda ^3|^4 |x^3_q|^4}{M_3^2}\right. \nonumber \\&\quad \left. + \frac{|\lambda ^1|^2 |\lambda ^3|^2 |x^1_q|^2|x^3_q|^2 \log M_3^2/M_1^2}{M_3^2 - M_1^2}\right) ~, \end{aligned}$$
(3.14)

where \(\left[ L^{VLL}_{\nu d} \right] _{\nu _\tau \nu _\tau d_i d_j}\) are the Wilson coefficients (WCs) of the low-energy operators \(({\bar{\nu }}_\tau \gamma _\mu \nu _\tau )({\bar{d}}^i_L \gamma ^\mu d^j_L)\), \(\delta g^Z_{\tau _L}\) describes the deviation in the Z couplings to \(\tau _L\), and \(C^1_K\) is the coefficient of the \(({{\bar{s}}} \gamma _\mu P_L d)^2\) operator. The leading contribution to \(s \rightarrow d \mu \mu \) transitions has a phase fixed to be equal to the SM one, so no large effect in \(K_S \rightarrow \mu \mu \) can be expected. Analogously, also in \(s \rightarrow d \nu \nu \) the NP coefficients have the same phase as in the SM, since the x coefficients enter with the absolute value squared. This implies that no cancellation between the two LQs can take place in this channel and that we expect a tight correlation between \(K_L \rightarrow \pi ^0 \nu \nu \) and \(K^+ \rightarrow \pi ^+ \nu \nu \), independently on the phases of the couplings. Similar considerations apply for all \(s \rightarrow d\) transitions. On the other hand, non-trivial phases can appear in \(b \rightarrow s\) transitions and a mild cancellation can alleviate the \(B\rightarrow K^{(*)}\nu \nu \) bound [33]. Since real couplings are favored by the B-anomalies and since in any case the phases in Kaon physics observables are fixed by the \(U(2)^5\) flavor structure, in our numerical analysis we only consider real values for all parameters.

Fig. 1
figure 1

Results of a parameter scan in the \(U(2)^5\) scenario. The green (yellow) points are within the 68% (95%) CL from the best-fit point (shown in black). We also overlay 2\(\sigma \) constraints from single observables, where other parameters are fixed to the best-fit point (3.15)

3.1 Analysis and discussion

Using the global likelihood we find the best-fit point in parameter space, allowing the x’s to vary in the range \(|x|<5\), while \(\lambda ^{1(3)}, V_\ell > 0\). Fixing \(M_1 = M_3 = 1.1\text { TeV}\) we get \(\chi ^2_{\text {SM}} - \chi ^2_\mathrm{best-fit} = 47.6\), for:Footnote 3

$$\begin{aligned} \text {best-fit } U(2)^5: \quad \begin{array}{l l l l} \lambda ^{1} \approx 0.79~, &{} \lambda ^{3} \approx 0.72~, &{} V_\ell \approx 0.071~, &{} s_e \approx 0~, \\ x^1_q \approx -0.97~, &{} x^3_q \approx 1.6~, &{} x^3_{\ell } \approx 3.6~, &{} x^3_{q\ell } \approx -2.0~. \end{array} \nonumber \\ \end{aligned}$$
(3.15)

We then perform a numerical scan on all the parameters in Eq. (3.15), selecting only points with a \(\Delta \chi ^2 = \chi ^2 - \chi ^2_\mathrm{best-fit}\) corresponding to a 68% (green points) or 95% (yellow) confidence level. The results are shown in Fig. 1, where we project the points to several 2D planes. We also show \(2\sigma \) constraints from single observables, obtained by fixing the parameters not in the plot to the corresponding best-fit values, Eq. (3.15).Footnote 4 We observe from Fig. 1 (bottom-left) that values \(V_\ell \approx 0.1\) and \(|s_e| \lesssim 0.02\) are preferred. We can also see that all the x’s can be of \({\mathcal {O}}(1)\), with no tendency towards parametrically smaller or larger values, hence the structure of the \(U(2)^5\) symmetry is respected.

Fig. 2
figure 2

Results, for the B-anomalies, of the parameter scan in the \(U(2)^5\) scenario shown in Fig. 1. The green (yellow) points are within the 68% (95%) CL from the best-fit point (3.15). The gray lines describe different CL regions in the fit of the anomalies (see Appendix A for details)

Fig. 3
figure 3

Results, for rare Kaon decays and \(\mu \rightarrow e\) conversion processes, of the parameter scan in the \(U(2)^5\) scenario shown in Fig. 1. The green (yellow) points are within the 68% (95%) CL from the best-fit point. In the upper plot, the gray region is excluded by the Grossman-Nir bound [80], the red solid and dashed lines represent the present measurement from NA62, the dotted brown line the sensitivity prospect for KOTO after stage-I, while the dotted purple one the final sensitivity expected from NA62 and KOTO (stage-II). In the lower plot the red lines describe the present 95% CL bound (see Appendix A for details)

In Figs. 2 and 3 we show the values of particularly interesting pairs of observables obtained with the same sets of parameter-space points. From Fig. 2 we observe that, while neutral-current B-anomalies can be addressed entirely, this setup can enhance \(R(D^{(*)})\) only by \(\lesssim 7\%\) of the SM size, i.e. at the \(2\sigma \) level of the present combination. This situation should be compared with the result of the analogous similar fit with \(S_1\) and \(S_3\) with only couplings to left-handed fermions shown in [40] (c.f. Fig. 5), where both anomalies can be satisfied in a scenario where couplings to the second generation quarks were compatible with a \(U(2)^5\) flavor structure, \(|\lambda ^{1(3)L}_{s \ell }| \sim |V_{ts}/V_{tb}| |\lambda ^{1(3)L}_{b \ell }|\), but couplings to first generation were set to zero.

Therefore, the reason for the inability of the \(U(2)^5\)-symmetric scenario to fully address charged-current anomalies must be found in first-generation constraints, specifically Kaon physics. Indeed, this can be seen in the first row of Fig. 1, where we observe that the bounds from \(K^+ \rightarrow \pi ^+ {\bar{\nu }} \nu \), Eq. (3.12), and \(\epsilon _K\) (i.e. \(\mathop {\mathrm{Im}}{C^1_K}\)), Eq. (3.14), in combination with the constraints on \(\lambda ^{1,3}\) from \(Z\rightarrow {\bar{\tau }}\tau \), Eq. (3.13), don’t allow the fit to enter the region preferred by \(R(D^{(*)})\), due to the precise relations between couplings to the first and the second generation, derived from the flavor structure, i.e. Eq. (3.7).

Regarding Kaon physics observables, from Fig. 3 we see that \({\mathcal {B}}(K^+ \rightarrow \pi ^+ {\bar{\nu }} \nu )\) can take all values currently allowed by the NA62 bound [57] (we show with vertical lines the best-fit and the \({\pm }1\sigma \) intervals) and therefore any future update on this observable will put further strong constraints on this scenario. Furthermore, since the phase in \(s \rightarrow d \nu \nu \) is fixed by the corresponding CKM phase, Eq. (3.12), a correlation between this mode and \({\mathcal {B}}(K_L \rightarrow \pi ^0 {\bar{\nu }} \nu )\) is obtained, with values \(\sim 10^{-10}\) also for the latter.Footnote 5 Therefore, even by the end of stage-I the KOTO experiment won’t be able to reach the sensitivity to test this model (brown horizontal dotted line). However, the future sensitivity goals by NA62 (\(10\%\) [82]) and KOTO stage-II , or KLEVER, (\(20\%\) [83, 84]) would be able to completely test this scenario (purple ellipse).

The model also predicts short-distance contributions to \({\mathcal {B}}(K_L\rightarrow \mu \mu )\) of the order of the present bound, although it remains challenging to improve this constraint in the future due to non-perturbative contributions to the long-distance component. Also in this channel the phase of the NP WC is fixed to the one of \(V_{ts}^* V_{td}\), see Eq. (3.103.11). The \(K_S \rightarrow \mu \mu \) mode is thus completely correlated with the \(K_L\) decay and the New Physics effect adds constructively to the SM short-distance amplitude, the expected size is however below the SM long-distance contribution of \(\approx 5\times 10^{-12}\). The expected relative effect in the tree-level transitions \(s \rightarrow u \mu {\bar{\nu }}_\mu \) and \(d \rightarrow u \mu {\bar{\nu }}_\mu \) is of order \(10^{-7}\) and \(10^{-9}\), respectively, excluding possible signatures from these decays.

Finally, the rotation angle between electrons and muons, \(s_e\), is constrained mainly by LFV \(\mu \rightarrow e\) processes, the strongest bound presently given by \(\mu \rightarrow e\) conversion in gold atoms, while the predicted effect in titanium is strictly correlated in our model, with an approximate relation \({\mathcal {B}}(\mu \rightarrow e)_\mathrm{Au} \approx 1.3 {\mathcal {B}}(\mu \rightarrow e)_\mathrm{Ti}\). In Fig. 3 we show the correlation with \(\mu \rightarrow 3 e\), while the expected branching ratio for \(\mu \rightarrow e \gamma \) is \(\lesssim 0.5 \times 10^{-13}\), thus below than the present precision. These measurements are expected to improve substantially in the future, reaching limits of the order of \(10^{-16}\) in \(\mu \rightarrow e\) conversion in nuclei from COMET and Mu2e [59,60,61,62] or even \(10^{-18}\) by the PRISM proposal, and also a level of \(10^{-16}\) in \(\mu \rightarrow 3 e\) from the Mu3e experiment [63]. These will put further constraints on the \(s_e\) parameter, or a signal could be observed if this mixing angle is large enough.

For what regards LFV in Kaon decays, given the largest allowed values for \(s_e\) and \(V_\ell \) from present limits, we obtain at most \({\mathcal {B}}(K_L \rightarrow \mu e) \lesssim 10^{-15}\) and \({\mathcal {B}}(K^+ \rightarrow \pi ^+ \mu e) \lesssim 10^{-18}\).

4 General case with right-handed couplings

In this section we depart from the flavor symmetry assumption and examine the fit of the \(S_1 + S_3\) model in the general case where all the couplings, including the right-handed ones, are allowed. As investigated in Ref. [40], if the couplings are a priori uncorrelated, there is enough freedom to accommodate both B-physics anomalies as well as the discrepancy in the anomalous magnetic moment of the muon.

Due to the high dimensionality of the parameter space, it is computationally too expensive to perform a \(\chi ^2\) minimization by random search techniques. For the purposes of our discussion it suffices then to use the best-fit point of Eq. (3.11) in Ref. [40]Footnote 6 in order to fix the relevant couplings and let only the additional couplings vary. A further simplification constitutes in switching off the couplings to electrons. This choice is justified by the fact that the necessary suppression required to pass the stringent bounds from LFV \(\mu \rightarrow e\) processes is of the same order in both the flavor-symmetry motivated case and the general one.

Fig. 4
figure 4

Allowed region obtained by varying the couplings \(\lambda _{s\tau }^{1(3)L}\) and \(\lambda _{d\tau }^{1(3)L}\),relevant for the \(B \rightarrow K \nu \nu \) decays (left), and the couplings \(\lambda _{s\mu }^{3L}\) and \(\lambda _{d\mu }^{3L}\) relevant for the \(K_{L,S} \rightarrow \mu \mu \) decays (right). The rest of the couplings are fixed to the best-fit point couplings in Eq. (3.11) of Ref. [40] and compatibility with the global fit is retained at 68% (green) and 95% (yellow) CL. On the left plot, the gray region is excluded by the Grossman-Nir bound, the red solid and dashed lines represent the present measurement from NA62, the dotted brown line the sensitivity prospect for KOTO after stage-I, while the dotted purple one the final sensitivity expected from NA62 and KOTO (stage-II). On the right plot, the red solid line represents the bound set in Ref. [85] and the red dashed line the future prospects by LHCb [86]

Using the best-fit point from Ref. [40] assures that all the anomalies are addressed within \(1\sigma \) (c.f. Fig. 5 therein). In order to assess the viable values of \({\mathcal {B}}(K_L \rightarrow \pi ^0 \nu \nu )\) and \({\mathcal {B}}(K^+ \rightarrow \pi ^+ \nu \nu )\) in this setup we perform a likelihood scan of the complex couplings \(\lambda _{s\tau }^{1(3)L}\) and \(\lambda _{d\tau }^{1(3)L}\) (see Eq. (A.10)) and keep only values that have a likelihood within the 68% or 95% CL from the best-fit point. The corresponding branching ratios are reported in the left plot of Fig. 4, from which we observe that, compared to the \(U(2)^5\) case, the viable values are significantly expanded. We also notice that the two decays are induced by the same short distance operators and are trivially related through isospin, leading to the so-called Grossman-Nir bound [80]. The LQ interactions fall under a category of models that may saturate the bound [87]. Yet this is not entirely possible due to the constraints on the magnitude of the couplings to muons and electrons. Nevertheless, we see that \({\mathcal {B}}(K_L \rightarrow \pi ^0 \nu \nu )\) could potentially take values that can be probed by the end of stage 1 of the KOTO experiment. For other studies of correlations between the \(K^+\) and the \(K_L\) modes see e.g. Refs. [54, 81, 88, 89].

By letting vary the \(\lambda _{s\mu }^{3L}\) and \(\lambda _{d\mu }^{3L}\) complex couplings we find that the short distance contribution to \(K_{S} \rightarrow \mu \mu \) can reach values of the order of the long-distance one, while that to the \(K_L\) mode saturate the theory-derived constraint [85]. In the right plot of Fig. 4 we present the 68% and 95% CL regions from the best-fit point in these two observables. Considering the \(\approx 30\%\) uncertainty on the SM prediction for the \(K_S\) mode, we notice that a significant part of the preferred region features NP effects that are distinguishable from the SM ones. As a matter of fact, the future prospects look promising, since the LHCb Upgrade II plans to exclude branching fractions down to near the SM prediction [86] (see dashed line in Fig. 4).

Regarding \(\mu \rightarrow e\) transition in nuclei we expect a very similar result to the one shown in Fig. 3, since even in that case the observables saturate the present bounds and there is additional freedom if the flavor symmetry assumption is removed.

5 Conclusions

The observed anomalies in B decays can potentially be addressed in LQ scenarios. While couplings to first generation of fermions are not required to describe these deviations, and could therefore be set to zero in a bottom-up approach, the typical expectation from UV models is that all couplings should be generated. In this work we study the impact that LQ couplings to first generation fermions can have on Kaon and electron LFV observables, in scenarios that aim at addressing the B-anomalies.

Correlations between couplings to different generations, and therefore between B and Kaon decays, can be obtained only if the flavor structure is specified. In this work we consider the approximate \(U(2)^5\) flavor symmetry, that is motivated from the observed SM fermions mass hierarchies and also predicts a pattern of deviations in B decays consistent with the observed one. With this assumption, the LQ coupling to left-handed d quarks are fixed to be equal to those to left-handed s quarks times the small CKM factor \(V_{td}/V_{ts}\) (see Eq. (3.7)). This structure correlates strongly Kaon physics with B decays. After performing a global likelihood analysis of a large set of observables in this scenario, we find that the \(S_1 + S_3\) LQs can address the \(R(D^{(*)})\) anomalies only at the \(2\sigma \) level. The most important observables preventing a successful fit are \({\mathcal {B}}(K^+ \rightarrow \pi ^+ \nu \nu )\), \(\epsilon _K\), and the Z couplings to \(\tau \). We thus expect future improved measurements of \(K^+ \rightarrow \pi ^+ \nu \nu \) from NA62 to further test this scenario. The LQ couplings to electrons, instead, are constrained mainly by the limits from \(\mu \rightarrow e\) transition in nuclei and \(\mu \rightarrow 3 e\) decay, that will be improved by several orders of magnitude in the near future.

Going beyond the flavor symmetric scenario, we also studied the allowed values for \({\mathcal {B}}(K^+ \rightarrow \pi ^+ \nu \nu )\), \({\mathcal {B}}(K_L \rightarrow \pi ^0 \nu \nu )\), and \(K_{L/S} \rightarrow \mu ^+ \mu ^-\) in the general case where no flavor structure is imposed. We find that, in this setup, values of \({\mathcal {B}}(K_L \rightarrow \pi ^0 \nu \nu )\) and \({\mathcal {B}}(K_S \rightarrow \mu ^+ \mu ^-)\) that could be potentially probed by KOTO stage-1 or LHCb, respectively, are allowed.

Table 1 Observables from B and D meson decays. Upper limits correspond to 95%CL. The correlations between \(\Delta C_9^{sb\mu \mu }\) and \(C_9^\mathrm{univ}\), as well as between \(R_{D}\) and \(R_{D^*}\) are taken into account
Table 2 D meson and Kaon physics observables with the corresponding SM predictions and experimental bounds. Upper limits correspond to 95% CL
Table 3 Meson-mixing and leptonic observables and EDMs, with their SM predictions and experimental bounds. Upper limits correspond to 95%CL

The B-anomalies are expected to receive further experimental inputs in the next few years by LHCb, Belle-II, as well as CMS and ATLAS experiments. However, even assuming these will be confirmed, in order to understand the flavor structure of the underlying New Physics the connection to Kaon physics and to observables sensitive to electron couplings will be crucial.