This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Lepton-flavor changing decays and non-unitarity in the inverse seesaw mechanism

Adrián González-Quiterio    Héctor Novales-Sánchez Facultad de Ciencias Físico Matemáticas, Benemérita Universidad Autónoma de Puebla, Apartado Postal 1152 Puebla, Puebla, México
Abstract

The pursuit of the genuine fundamental description, governing nature at some high-energy scale, must invariably consider the yet-unknown mechanism behind the generation of neutrino mass. Lepton-flavor violating decays lαγlβl_{\alpha}\to\gamma\,l_{\beta}, allowed in the presence of neutrino mass and mixing, provide a mean to look for physics beyond the Standard Model. In the present work we consider the inverse seesaw mechanism and then revisit the calculation of its contributions to the branching ratios of the aforementioned decay processes, among which we find μγe\mu\to\gamma\,e to be more promising, in the light of current bounds by the MEG Collaboration. Deviations from unitarity in the mixing of light neutrinos are related to the branching ratios Br(lαγlβ){\rm Br}\big{(}l_{\alpha}\to\gamma\,l_{\beta}\big{)} in a simple manner, which we address, then finding that, while experimental data are consistent with current bounds on non-unitarity effects, the upcoming MEG II update shall be able to improve restrictions on such effects by a factor 13\sim\frac{1}{3}.

I Introduction

The key role played by gauge symmetry in the description of fundamental physics and the origin of mass through the occurrence of spontaneous symmetry breaking have been greatly supported by the measurement, by the ATLAS and CMS Collaborations at the CERN, of the long-awaited Higgs boson HiggsATLAS ; HiggsCMS . In the minimal theoretical scheme, provided by the scalar sector of the Standard Model SMGlashow ; SMSalam ; SMWeinberg , the Brout-Englert-Higgs mechanism breaks the electroweak gauge symmetry group SU(2)LU(1)Y{\rm SU}(2)_{L}\otimes{\rm U}(1)_{Y} into the electromagnetic group U(1)e{\rm U}(1)_{e} EnBr ; Higgs , which comes along with the definition of the masses of all the currently measured particles, except for the neutrinos, assumed to be massless in this framework. Neutrino masses have not ever been measured PDG , and yet the widely accepted interpretation that neutrino oscillations Pontecorvo imply massiveness of these particles calls for an explanation emerging from some high-energy formulation of physics beyond the Standard Model. Though the addition of three sterile right-handed neutrino fields to the Standard-Model particle spectra, together with a set of Dirac-type Yukawa terms, does the job, the resulting neutrino masses are determined by ad hoc “unnaturally” small values of Yukawa constants. In view of the massiveness and electromagnetic neutrality of neutrinos, the proper description of these particles could rather correspond to Majorana spinor fields Majorana , in which case global U(1){\rm U}(1) invariance is lost and, therefore, lepton number is not conserved. If lepton-number non-preservation is assumed, the effective Lagrangian for the electroweak Standard Model LLR ; BuWy ; Wudka gets extended, then allowing for the emergence of Majorana mass-terms Weinbergoperator ; BaLe ; CSV , presumably originating in some description of fundamental physics characterized by a high-energy scale, Λ\Lambda. In particular, the Weinberg operator Weinbergoperator yields neutrino masses matching the mass profile that characterizes the seesaw mechanism MoSe1 ; MoSe2 , in which, besides the known neutrinos, a set of heavy-neutral-lepton partners, dubbed “heavy neutrinos”, comes about. The heavy-neutrino masses, mNjm_{N_{j}}, determined by the high-energy scale as mNjΛm_{N_{j}}\sim\Lambda, are linked to the masses, mνjm_{\nu_{j}}, of the light neutrinos, in this scheme given by mνjv2Λm_{\nu_{j}}\sim\frac{v^{2}}{\Lambda}, with vv the vacuum expectation value of the Standard-Model Higgs potential. In this context, current upper bounds on light-neutrino masses Planck ; cosmonumass ; KATRIN push the masses of heavy neutrinos towards enormous values, thus leaving direct production of heavy neutrinos off the table and also severely attenuating their contributions, as virtual particles, to Standard-Model observables.

So even though the seesaw mechanism provides a nice explanation for the tininess of neutrino mass, in connection with fundamental physics beyond the Standard Model, it is quite difficult to probe. This inconvenience has motivated the realization of seesaw variants, aimed at a relaxation of the restriction on the energy scale Λ\Lambda, in order to bring it closer to current experimental sensitivity. Ref. CHLR provides a review on the seesaw mechanism and its variants. The framework for the present paper is defined by the neutrino-mass generation approach known as the “inverse seesaw mechanism” MoVa ; GoVa ; DeVa . More concretely, we consider a realization of the inverse seesaw in which the neutrino sector is enriched, as compared to the case of the Standard Model, by the introduction of three right-handed neutrino fields, together with a set of three further left-handed lepton fields, all of them singlets with respect to the SU(2)LU(1)Y{\rm SU}(2)_{L}\otimes{\rm U}(1)_{Y} gauge-symmetry group. These new fermion fields introduce a slight violation of lepton number through Majorana-like mass terms characterized by two small-valued matrices, here denoted by μS\mu_{S} and μR\mu_{R}, both proportional to an energy scale vσv_{\sigma}, of spontaneous symmetry breaking, which abide by the hierarchy condition vσvΛv_{\sigma}\ll v\ll\Lambda. In the mass-eigenspinor basis, this extended neutrino sector yields three light neutrinos and a total of six heavy neutral leptons. The mass matrix of light neutrinos, mνm_{\nu}, is proportional to μS\mu_{S}, which therefore attenuates the pressure on Λ\Lambda, thus allowing for smaller and more reasonable values of the heavy-neutrino masses, as compared to what happens in the original version of the seesaw mechanism.

In the electroweak Standard Model, the absence of right-handed neutrinos and the assumption of lepton-number conservation prevents neutrino-mass terms from being generated. Moreover, lepton-flavor-violating processes, also forbidden in such a context, are well-motivated means to search for traces pointing towards the presence of new physics. Despite the large amount of experimental work dedicated to the search for charged-lepton-flavor violation, no process of this kind has ever been observed PDG , though notice that neutrino mixing, required for the occurrence of neutrino oscillations Pontecorvo , implies that such sort of processes are actually allowed, in which case their measurement could be eventually achieved in experimental facilities. The present investigation readdresses the one-loop contributions, generated by the whole set of massive neutral leptons defined in the context of the inverse seesaw, to the charged-lepton-flavor-changing decays μeγ\mu\to e\,\gamma, τeγ\tau\to e\,\gamma, and τμγ\tau\to\mu\,\gamma. These decay processes have been formerly examined in Refs. SFLYC ; SoRu ; GPH . Currently, the most stringent experimental constraints on these decay processes are the ones reported by the MEG, Belle, and BaBar collaborations MEGlfvbound ; Bellelfvbound ; BaBarlfvbound . The MEG II and Belle II upgrades, expected to improve experimental sensitivities by 1\sim 1 order of magnitude MEG2lfvestimation ; Belle2lfvestimation , are also to be borne in mind. Our estimations show that, in accordance with such experimental works, the inverse-seesaw contribution to μeγ\mu\to e\,\gamma yields the most promising result among the charged-lepton-flavor-violating decay processes under consideration. A worthwhile aspect of the inverse seesaw regards the matrix that characterizes the mixing of light neutrinos with charged leptons. In the simplest scenarios, as the Standard Model endowed with three singlet right-handed Dirac neutrino fields GiKi or the one given by the introduction of the Weinberg operator with the assumption that lepton number is not preserved Weinbergoperator , light-neutrino mixing is characterized by the Pontecorvo-Maki-Nakgawa-Sakata (PMNS) matrix MNSmatrix ; Pontecorvomatrix , which is unitary. However, the presence of a light-neutrino mixing matrix which is not unitary occurs in a large class of models aimed at generating neutrino masses FGLY . Constraints imposed by the charged-lepton-flavor-violating decays μeγ\mu\to e\,\gamma, τeγ\tau\to e\,\gamma, and τμγ\tau\to\mu\,\gamma on these non-unitary effects have been estimated in Refs. FHL ; BFHLMN . In the present work, we analyze the aforementioned lepton-flavor-changing decays in terms of their relation with such non-unitary effects. The most restrictive constraints proceed from the contributions to μeγ\mu\to\,e\gamma and its comparison with the MEG limit. Furthermore, we find that the sensitivity expectation estimated for MEG II should yield a potential improvement, given by a factor 13\sim\frac{1}{3}, of restrictions for the non-unitary effects.

The reminder of the paper has been organized as follows: in Section II, the inverse seesaw mechanism is reviewed, which includes the main expressions for the phenomenological calculation to be carried out later; we describe the execution of our analytical calculation of the charged-lepton-flavor-violating decays μeγ\mu\to e\,\gamma, τeγ\tau\to e\,\gamma, and τμγ\tau\to\mu\,\gamma at one loop in Section III; our estimations and analyses of the contributions are developed throughout Section IV; we conclude the paper by providing, in Section V, a summary.

II The inverse seesaw mechanism

In this section, we discuss the inverse seesaw mechanism MoVa ; GoVa ; DeVa , which is the framework behind our phenomenological investigation. Aiming at a more definite presentation, we address this task by considering a completion in which lepton number is spontaneously broken CMP . Think of an extension of the Standard Model, characterized by a Lagrangian invariant under the electroweak gauge group SU(2)LU(1)Y{\rm SU}(2)_{L}\otimes{\rm U}(1)_{Y}. Assume that the Standard-Model field content has been increased by the introduction of a set of three chiral fermion fields ν1,R\nu_{1,R}, ν2,R\nu_{2,R}, and ν3,R\nu_{3,R}, which are right handed with lepton number L(νR)=+1L(\nu_{R})=+1, and by three more fermion fields S1,LS_{1,L}, S2,LS_{2,L}, and S3,LS_{3,L}, with definite left chirality and lepton number L(SL)=+1L(S_{L})=+1. All these fields are assumed to be singlets with respect to electroweak gauge transformations. Then, a complex scalar field, σ\sigma, assumed to be an SU(2)LU(1)Y{\rm SU}(2)_{L}\otimes{\rm U}(1)_{Y} singlet and to have lepton number L(σ)=+2L(\sigma)=+2, is introduced as well. The addition of these new-physics fields translates into extensions of both the lepton-Yukawa sector and the scalar sector of the Standard Model.

The Standard-Model Higgs doublet Φ\Phi and the scalar singlet σ\sigma define a scalar potential, nested within the scalar sector, which reads

V(Φ,σ)=μΦ2ΦΦμσ2σσ+λΦ(ΦΦ)2\displaystyle V(\Phi,\sigma)=-\mu_{\Phi}^{2}\Phi^{\dagger}\Phi-\mu_{\sigma}^{2}\sigma^{\dagger}\sigma+\lambda_{\Phi}\big{(}\Phi^{\dagger}\Phi\big{)}^{2}
+λσ(σσ)2+λΦσ(ΦΦ)(ΦΦ),\displaystyle\hskip 36.98866pt+\lambda_{\sigma}\big{(}\sigma^{\dagger}\sigma\big{)}^{2}+\lambda_{\Phi\sigma}\big{(}\Phi^{\dagger}\Phi\big{)}\big{(}\Phi^{\dagger}\Phi\big{)}, (1)

where μΦ2\mu_{\Phi}^{2}, μσ2\mu_{\sigma}^{2}, λΦ\lambda_{\Phi}, λσ\lambda_{\sigma}, and λΦσ\lambda_{\Phi\sigma}, are positive quantities111A discussion on the vacuum stability of V(Φ,σ)V\big{(}\Phi,\sigma\big{)} can be found in Ref. MRSV .. Two stages of spontaneous symmetry breaking come about, the first of which we assume to occur at the energy scale v=2(2λσμΦ2λΦσμσ2)4λΦλσλΦσ2v=\sqrt{\frac{2(2\lambda_{\sigma}\mu_{\Phi}^{2}-\lambda_{\Phi\sigma}\mu_{\sigma}^{2})}{4\lambda_{\Phi}\lambda_{\sigma}-\lambda_{\Phi\sigma}^{2}}}, identified as the electroweak vacuum expectation value. By this symmetry breaking, the gauge group SU(2)LU(1)Y{\rm SU}(2)_{L}\otimes{\rm U}(1)_{Y} is broken down to the electromagnetic group U(1)e{\rm U}(1)_{e}. Then, at vσ=2(2λΦμσ2λΦσμΦ2)4λΦλσλΦσ2v_{\sigma}=\sqrt{\frac{2(2\lambda_{\Phi}\mu_{\sigma}^{2}-\lambda_{\Phi\sigma}\mu_{\Phi}^{2})}{4\lambda_{\Phi}\lambda_{\sigma}-\lambda_{\Phi\sigma}^{2}}}, the global symmetry U(1){\rm U}(1) associated to the preservation of lepton number is broken. After this, two massive scalar fields are defined, namely, the Higgs boson field, hh, and a novel scalar field, s0s_{0}, with masses given by

mh2=λΦv2+λσvσ2+(λΦv2λσvσ2)2+v2vσ2λΦσ2,m_{h}^{2}=\lambda_{\Phi}v^{2}+\lambda_{\sigma}v_{\sigma}^{2}+\sqrt{\big{(}\lambda_{\Phi}v^{2}-\lambda_{\sigma}v_{\sigma}^{2}\big{)}^{2}+v^{2}v_{\sigma}^{2}\lambda_{\Phi\sigma}^{2}}, (2)
ms02=λΦv2+λσvσ2(λΦv2λσvσ2)2+v2vσ2λΦσ2.m_{s_{0}}^{2}=\lambda_{\Phi}v^{2}+\lambda_{\sigma}v_{\sigma}^{2}-\sqrt{\big{(}\lambda_{\Phi}v^{2}-\lambda_{\sigma}v_{\sigma}^{2}\big{)}^{2}+v^{2}v_{\sigma}^{2}\lambda_{\Phi\sigma}^{2}}. (3)

Moreover, as a consequence of the rupture of the U(1){\rm U}(1) global symmetry, a physical Goldstone boson JJ, widely known as the Majoron, emerges CMP . Originated from the symmetry breaking of a global group, the Majoron is born massless. Yet, authors have evoked mechanisms to give it a mass, thus turning this scalar into a viable dark matter candidate ABMS ; BeVa ; RBS ; BLRV ; GMS ; EJJRTV ; GaHe ; BrPa .

Prior to electroweak symmetry breaking, the Yukawa sector, Y{\cal L}_{\rm Y}, is given by the Lagrangian

Y=j=13k=13(YjkνLj,L¯Φ~νk,R\displaystyle{\cal L}_{\rm Y}=\sum_{j=1}^{3}\sum_{k=1}^{3}\Big{(}-Y^{\nu}_{jk}\overline{L_{j,L}}\tilde{\Phi}\nu_{k,R}
MjkSj,L¯νk,R12λjk(S)σSj,L¯Sk,Lc\displaystyle\hskip 22.76228pt-M_{jk}\overline{S_{j,L}}\nu_{k,R}-\frac{1}{2}\lambda_{jk}^{(S)}\,\sigma\,\overline{S_{j,L}}S_{k,L}^{\rm c}
12λjk(N)σνj,Rc¯νk,R+H.c.)+,\displaystyle\hskip 22.76228pt-\frac{1}{2}\lambda_{jk}^{(N)}\,\sigma^{*}\overline{\nu_{j,R}^{\rm c}}\nu_{k,R}+{\rm H.c.}\Big{)}+\cdots,
(4)

where

Lj,L=(νj,Llj,L)L_{j,L}=\left(\begin{array}[]{c}\nu^{\prime}_{j,L}\vspace{0.15cm}\\ l^{\prime}_{j,L}\end{array}\right) (5)

is the j-thj\textrm{-th} SU(2)L{\rm SU}(2)_{L} lepton doublet, with j=1,2,3j=1,2,3, and where the definition Φ~=iσ2Φ\tilde{\Phi}=i\sigma_{2}\Phi^{*}, with σ2\sigma_{2} the imaginary Pauli matrix, has been used. This expression for Y{\cal L}_{\rm Y} shows explicitly only those Yukawa couplings involving the fermion fields that extend the field content of the Standard Model, around which we develop our discussion, whereas the ellipsis denote further Yukawa terms, including those made exclusively of Standard-Model fields. The lagrangian terms explicitly displayed in Eq. (4) incorporate four sorts of Yukawa constants, each of which defines a 3×33\times 3 matrix: YνY^{\nu}, MM, λ(S)\lambda^{(S)}, and λ(N)\lambda^{(N)}. In particular, we assume the matrix MM to originate from a stage of spontaneous symmetry breaking taking place at a high-energy scale Λ\Lambda, of some gauge group that characterizes a high-energy fundamental description beyond the Standard Model, in which case we work under the premise that vΛv\ll\Lambda. Further, MΛM\sim\Lambda is reasonably assumed. The lepton-number assignments for the fields Sj,LS_{j,L}, νj,R\nu_{j,R}, and σ\sigma guarantee invariance of Y{\cal L}_{\rm Y} with respect to the global group U(1){\rm U}(1). Next, electroweak gauge symmetry is broken at vv, which allows us to get to the charged-lepton mass-eigenspinor basis by implementing the unitary transformations lL=VLlLl_{L}=V^{\ell{\dagger}}_{L}l^{\prime}_{L} and lR=VRlRl_{R}=V^{\ell{\dagger}}_{R}l^{\prime}_{R}, with lj,Rl^{\prime}_{j,R} the jj-th right-handed SU(2)L{\rm SU}(2)_{L} lepton singlet, then defining the α\alpha-flavor charged-lepton field as lα=lα,L+lα,Rl_{\alpha}=l_{\alpha,L}+l_{\alpha,R}. This results in the charged-lepton mass terms mlαlα¯lα-m_{l_{\alpha}}\overline{l_{\alpha}}\,l_{\alpha}. After breaking the SU(2)LU(1)Y{\rm SU}(2)_{L}\otimes{\rm U}(1)_{Y} gauge symmetry and then using σ=vσ2+\sigma=\frac{v_{\sigma}}{\sqrt{2}}+\cdots, to break the global U(1){\rm U}(1) group, the Yukawa Lagrangian Y{\cal L}_{\rm Y}, Eq. (4), acquires the form

Y=α=e,μ,τk=13((mD)αkνα,L¯νk,R+H.c.)\displaystyle{\cal L}_{\rm Y}=\sum_{\alpha=e,\mu,\tau}\sum_{k=1}^{3}\Big{(}-(m_{\rm D})_{\alpha k}\,\overline{\nu_{\alpha,L}}\,\nu_{k,R}+{\rm H.c.}\Big{)}
+j=13k=13(MjkSj,L¯νk,R(μS)jk2Sj,L¯Sk,Lc\displaystyle\hskip 22.76228pt+\sum_{j=1}^{3}\sum_{k=1}^{3}\Big{(}-M_{jk}\overline{S_{j,L}}\nu_{k,R}-\frac{(\mu_{S})_{jk}}{2}\overline{S_{j,L}}S_{k,L}^{\rm c}
(μR)jk2νj,Rc¯νk,R+H.c.)+.\displaystyle\hskip 22.76228pt-\frac{(\mu_{R})_{jk}}{2}\overline{\nu_{j,R}^{\rm c}}\nu_{k,R}+{\rm H.c.}\Big{)}+\cdots.
(6)

By considering the left-handed unitary-transformation matrix VLV^{\ell}_{L}, we found it convenient to define νL=VLνL\nu_{L}=V^{\ell{\dagger}}_{L}\nu^{\prime}_{L}. Then the first term in the right-hand side of this equation features the 3×33\times 3 complex matrix mD=v2VLYνm_{\rm D}=\frac{v}{\sqrt{2}}V^{\ell{\dagger}}_{L}Y^{\nu}. Moreover, the 3×33\times 3 matrices μS=vσ2λ(S)\mu_{S}=\frac{v_{\sigma}}{\sqrt{2}}\lambda^{(S)} and μR=vσ2λ(N)\mu_{R}=\frac{v_{\sigma}}{\sqrt{2}}\lambda^{(N)} emerge as well. Note that both these matrices are symmetric. The μS\mu_{S} and μR\mu_{R} terms are no longer invariant under global U(1){\rm U}(1) and therefore spoil lepton-number conservation. Assuming, on the grounds of naturalness tHooftnaturalness , that these matrices are small, which can be achieved if the condition vσvv_{\sigma}\ll v is fulfilled, U(1){\rm U}(1) global symmetry is only slightly violated. We follow such an assumption from here on. Notice that, from the assumed origin for the MM matrix, we have the inverse-seesaw hierarchy condition

vσvΛ,v_{\sigma}\ll v\ll\Lambda, (7)

among the three involved energy scales.

The Yukawa sector can be rearranged as

Y=12(νL¯νRc¯SL¯)ν(νLcνRSLc)+h.c.+,{\cal L}_{\rm Y}=\frac{1}{2}\left(\begin{array}[]{ccc}\overline{\nu_{L}}&\overline{\nu_{R}^{\rm c}}&\overline{S_{L}}\end{array}\right){\cal M}_{\nu}\left(\begin{array}[]{c}\nu_{L}^{\rm c}\vspace{0.1cm}\\ \nu_{R}\vspace{0.1cm}\\ S_{L}^{\rm c}\end{array}\right)+{\rm h.c.}+\cdots, (8)

where ν{\cal M}_{\nu} is a 9×99\times 9 symmetric matrix given, in terms of the 3×33\times 3 block matrices mDm_{\rm D}, MM, μR\mu_{R}, and μS\mu_{S}, as

ν=(0mD0mDTμRMT0MμS).{\cal M}_{\nu}=\left(\begin{array}[]{ccc}0&m_{\rm D}&0\vspace{0.1cm}\\ m_{\rm D}^{\rm T}&\mu_{R}&M^{\rm T}\vspace{0.1cm}\\ 0&M&\mu_{S}\end{array}\right). (9)

Moreover, the definitions

MM=(μRMTMμS),M_{\rm M}=\left(\begin{array}[]{cc}\mu_{R}&M^{\rm T}\vspace{0.1cm}\\ M&\mu_{S}\end{array}\right), (10)
MD=(mD0),M_{\rm D}=\left(\begin{array}[]{cc}m_{\rm D}&0\end{array}\right), (11)

where the sizes of MMM_{\rm M} and MDM_{\rm D} respectively are 6×66\times 6 and 3×63\times 6, allow one to give ν{\cal M}_{\nu} the form of the neutral-lepton mass matrix that characterizes the type-1 seesaw mechanism. Since ν{\cal M}_{\nu} is symmetric, the existence of a unitary-diagonalization matrix Ω\Omega is ensured, with the diagonalization proceeding as Takagi

ΩTνΩ=(Mn000MN000MX).\Omega^{\rm T}{\cal M}_{\nu}\,\Omega=\left(\begin{array}[]{ccc}M_{n}&0&0\vspace{0.1cm}\\ 0&M_{N}&0\\ 0&0&M_{X}\end{array}\right). (12)

In this equation, MnM_{n} is a 3×33\times 3 diagonal matrix, whose in-diagonal elements are the light-neutrino masses, which we denote as mnjm_{n_{j}}, with j=1,2,3j=1,2,3. Furthermore, the matrix MNM_{N} is 3×33\times 3 sized and diagonal, with its diagonal entries corresponding to the masses mNjm_{N_{j}} of three heavy neutrinos , so j=1,2,3j=1,2,3 as well. Finally, the 3×33\times 3 diagonal matrix MXM_{X} comprises, nested within its diagonal, the masses mXjm_{X_{j}} of three further heavy neutral leptons. This diagonalization procedure entails the following change of basis:

(nLNLXL)=ΩT(νLνRcSL),(nRNRXR)=Ω(νLcνRSLc).\left(\begin{array}[]{c}n_{L}\vspace{0.1cm}\\ N_{L}\vspace{0.1cm}\\ X_{L}\end{array}\right)=\Omega^{\rm T}\left(\begin{array}[]{c}\nu_{L}\vspace{0.1cm}\\ \nu_{R}^{\rm c}\vspace{0.1cm}\\ S_{L}\end{array}\right),\hskip 8.5359pt\left(\begin{array}[]{c}n_{R}\vspace{0.1cm}\\ N_{R}\vspace{0.1cm}\\ X_{R}\end{array}\right)=\Omega^{\dagger}\left(\begin{array}[]{c}\nu_{L}^{\rm c}\vspace{0.1cm}\\ \nu_{R}\vspace{0.1cm}\\ S_{L}^{\rm c}\end{array}\right). (13)

Here, nLn_{L} and nRn_{R} are 3×13\times 1 matrices for chiral left- and right-handed neutrino fields, respectively. Meanwhile, NLN_{L}, NRN_{R}, XLX_{L}, and XRX_{R} are also 3×13\times 1 sized, comprised by left-handed and right-handed neutral spinor fields. Putting all the pieces together, the Yukawa-sector Lagrangian is written as

Y=j=13k=13(12mnjnj¯nj\displaystyle{\cal L}_{\rm Y}=\sum_{j=1}^{3}\sum_{k=1}^{3}\Big{(}-\frac{1}{2}m_{n_{j}}\overline{n_{j}}n_{j}
12mNjNj¯Nj12mXjXj¯Xj)+\displaystyle\hskip 17.07182pt-\frac{1}{2}m_{N_{j}}\overline{N_{j}}N_{j}-\frac{1}{2}m_{X_{j}}\overline{X_{j}}X_{j}\Big{)}+\cdots (14)

for which the non-chiral fermion fields nj=nj,L+nj,Rn_{j}=n_{j,L}+n_{j,R}, Nj=Nj,L+Nj,RN_{j}=N_{j,L}+N_{j,R}, and Xj=Xj,L+Xj,RX_{j}=X_{j,L}+X_{j,R} have been defined.

The diagonalization matrix Ω\Omega is conveniently expressed as

Ω=UV,\Omega=U\,V, (15)

with UU and VV a couple of 9×99\times 9 unitary matrices. The matrix UU is intended to block-diagonalize the mass matrix ν{\cal M}_{\nu}. This unitary matrix can be expressed in block-matrix form as

U=(U11U12U21U22),U=\left(\begin{array}[]{cc}U_{11}&U_{12}\vspace{0.1cm}\\ U_{21}&U_{22}\end{array}\right), (16)

where U11U_{11} and U22U_{22} are square matrices, the former 3×33\times 3 sized and the latter 6×66\times 6 sized. Meanwhile, U12U_{12} is a 3×63\times 6 matrix and U21U_{21} is 6×36\times 3. While the last equation is largely generic, the unitary character of UU allows for the following matrix-block parametrization KPS ; DePi :

U=((𝟏3+ξξT)12ξ(𝟏6+ξTξ)12ξT(𝟏3+ξξT)12(𝟏6+ξTξ)12),U=\left(\begin{array}[]{cc}\big{(}{\bf 1}_{3}+\xi^{*}\xi^{\rm T}\big{)}^{-\frac{1}{2}}&\xi^{*}\big{(}{\bf 1}_{6}+\xi^{\rm T}\xi^{*}\big{)}^{-\frac{1}{2}}\vspace{0.1cm}\\ -\xi^{\rm T}\big{(}{\bf 1}_{3}+\xi^{*}\xi^{\rm T}\big{)}^{-\frac{1}{2}}&\big{(}{\bf 1}_{6}+\xi^{\rm T}\xi^{*}\big{)}^{-\frac{1}{2}}\end{array}\right), (17)

where 𝟏3{\bf 1}_{3} and 𝟏6{\bf 1}_{6} stand for the 3×33\times 3 and the 6×66\times 6 identity matrices, respectively. Further, ξ\xi is a 3×63\times 6 matrix fulfilling MDξMDTξξM,M=0M_{\rm D}-\xi M_{\rm D}^{\rm T}\xi^{*}-\xi M_{\rm,M}=0. If the entries of ξ\xi are small, this condition yields the expression

ξ=MDMM1,\xi=M_{\rm D}M_{\rm M}^{-1}, (18)

corresponding to a large suppression provided by MM1M_{\rm M}^{-1}. From here on, we assume Eq. (18) to hold. Then, the block parametrization for UU, Eq. (17), can be approximated as

U(𝟏312MD(MM1)MM1MDTMD(MM1)MM1MDT𝟏612MM1MDTMD(MM1)).U\approx\left(\begin{array}[]{cc}{\bf 1}_{3}-\frac{1}{2}M_{\rm D}^{*}\big{(}M_{\rm M}^{-1}\big{)}^{*}M_{\rm M}^{-1}M_{\rm D}^{\rm T}&M_{\rm D}^{*}\big{(}M_{\rm M}^{-1}\big{)}^{*}\vspace{0.1cm}\\ -M_{\rm M}^{-1}M_{\rm D}^{\rm T}&{\bf 1}_{6}-\frac{1}{2}M_{\rm M}^{-1}M_{\rm D}^{\rm T}M_{\rm D}^{*}\big{(}M_{\rm M}^{-1}\big{)}^{*}\end{array}\right). (19)

The afore-announced block-matrix diagonalization, driven by UU, goes as follows:

UTνU=(Mlight00Mheavy),U^{\rm T}{\cal M}_{\nu}U=\left(\begin{array}[]{cc}M_{\rm light}&0\vspace{0.1cm}\\ 0&M_{\rm heavy}\end{array}\right), (20)

where, the matrix Mlight=MDMM1MDT+𝒪((MM1)3)M_{\rm light}=-M_{\rm D}M_{\rm M}^{-1}M_{\rm D}^{\rm T}+{\cal O}\big{(}(M_{\rm M}^{-1})^{3}\big{)} is 3×33\times 3 sized, whereas Mheavy=MM+𝒪(MM1)M_{\rm heavy}=M_{\rm M}+{\cal O}\big{(}M_{\rm M}^{-1}\big{)} is a 6×66\times 6 matrix. The matrix VV, introduced in Eq. (15), is written as

V=(V^00V~),V=\left(\begin{array}[]{cc}\hat{V}&0\vspace{0.1cm}\\ 0&\tilde{V}\end{array}\right), (21)

with the 3×33\times 3 matrix V^\hat{V} and the 6×66\times 6 matrix V~\tilde{V} respectively diagonalizing MlightM_{\rm light} and MheavyM_{\rm heavy} as

V^TMlightV^=Mn,\hat{V}^{\rm T}M_{\rm light}\,\hat{V}=M_{n}, (22)
V~TMheavyV~=(MN00MX),\tilde{V}^{\rm T}M_{\rm heavy}\tilde{V}=\left(\begin{array}[]{cc}M_{N}&0\vspace{0.1cm}\\ 0&M_{X}\end{array}\right), (23)

in accordance with Eq. (12). The inverse of MMM_{\rm M} turns out to be DePi

MM1=((μRMTμS1M)1(μRMTμS1M)1MTμS1(μSMμR1MT)1MμR1(μSMμR1MT)1),M_{\rm M}^{-1}=\left(\begin{array}[]{cc}\big{(}\mu_{R}-M^{\rm T}\mu_{S}^{-1}M\big{)}^{-1}&-\big{(}\mu_{R}-M^{\rm T}\mu_{S}^{-1}M\big{)}^{-1}M^{\rm T}\mu_{S}^{-1}\vspace{0.1cm}\\ -\big{(}\mu_{S}-M\mu_{R}^{-1}M^{\rm T}\big{)}^{-1}M\,\mu_{R}^{-1}&\big{(}\mu_{S}-M\mu_{R}^{-1}M^{\rm T}\big{)}^{-1}\end{array}\right), (24)

from which the inverse-seesaw light-neutrino mass formula,

MnV^TmDM1μS(MT)1mDV^,M_{n}\approx\hat{V}^{\rm T}m_{\rm D}M^{-1}\mu_{S}\,(M^{\rm T})^{-1}m_{\rm D}\hat{V}, (25)

follows. According to Eq. (25), the tinniness characterizing the masses of light neutrinos does not rely only on the suppression introduced by M1M^{-1}, as in the canonical seesaw mechanism, because the smallness of μS\mu_{S} establishes a further suppression, thus reducing the stress on the high-energy scale Λ\Lambda, associated to MM, which in this manner evades huge values.

As shown in Eq. (23), the unitary matrix V~\tilde{V}, which diagonalizes the heavy-neutral-lepton mass matrix MheavyM_{\rm heavy}, yields the diagonal mass matrices MNM_{N} and MXM_{X}, which correspond to the sets of heavy-neutral leptons {N1,N2,N3}\{N_{1},N_{2},N_{3}\} and {X1,X2,X3}\{X_{1},X_{2},X_{3}\}, respectively. In pursuit of some insight regarding the mass spectra for these fields, let us assume that the matrices MM, μR\mu_{R}, and μS\mu_{S} are diagonal, in which case V~\tilde{V} can be written as

V~=12(i(𝟏3+14χ)𝟏314χi(𝟏314χ)𝟏3+14χ),\tilde{V}=\frac{1}{\sqrt{2}}\left(\begin{array}[]{rcr}i({\bf 1}_{3}+\frac{1}{4}\chi)&&{\bf 1}_{3}-\frac{1}{4}\chi\vspace{0.2cm}\\ -i({\bf 1}_{3}-\frac{1}{4}\chi)&&{\bf 1}_{3}+\frac{1}{4}\chi\end{array}\right), (26)

where χ=M1(μSμR)\chi=M^{-1}\big{(}\mu_{S}-\mu_{R}\big{)} and ii is de imaginary unity. Thus, from Eqs. (23) and (26), the diagonal matrices

MN=MμS2μR2,M_{N}=M-\frac{\mu_{S}}{2}-\frac{\mu_{R}}{2}, (27)
MX=M+μS2+μR2,M_{X}=M+\frac{\mu_{S}}{2}+\frac{\mu_{R}}{2}, (28)

emerge, which then lead us to the neutral-lepton mass terms displayed in Eq. (14). Now consider the difference MXMN=μS+μRM_{X}-M_{N}=\mu_{S}+\mu_{R}, according to which masses mNjm_{N_{j}} and mXjm_{X_{j}}, all of them large as dictated by the hierarchy condition shown in Eq. (7), are very similar to each other. Thus the complete heavy-neutral-lepton mass spectrum is quasi-degenerate for each pair of fields NjN_{j} and XjX_{j}, with j=1,2,3j=1,2,3.

To recapitulate, after the spontaneous symmetry breaking of the Standard-Model electroweak gauge group SU(2)LU(1)Y{\rm SU}(2)_{L}\otimes{\rm U}(1)_{Y} into the electromagnetic group U(1)e{\rm U}(1)_{e}, at vv, and the ulterior breaking of the global U(1){\rm U}(1) symmetry at vσv_{\sigma}, the neutral-lepton field content comprehends three light-neutrino fields, denoted by njn_{j}, also three heavy-neutrino fields, which we have referred to as NjN_{j}, and three further heavy-neutral-lepton fields, represented by XjX_{j}, with mNjmXjm_{N_{j}}\approx m_{X_{j}}, for each pair labeled by j=1,2,3j=1,2,3. Let us jointly denote all the heavy-neutral-lepton fields as

fj={Nk, if j=1,2,3,Xk, if j=4,5,6,f_{j}=\left\{\begin{array}[]{l}N_{k},\textrm{ if }j=1,2,3,\vspace{0.1cm}\\ X_{k},\textrm{ if }j=4,5,6,\end{array}\right. (29)

where Nk=N1,N2,N3N_{k}=N_{1},N_{2},N_{3} and Xk=X1,X2,X3X_{k}=X_{1},X_{2},X_{3}. With this in mind, the charged-currents lagrangian term can be written as

Wνl=α=e,μ,τ(j=13(αnjWμlα¯γμPLnj+h.c.)\displaystyle{\cal L}_{W\nu l}=\sum_{\alpha=e,\mu,\tau}\Big{(}\sum_{j=1}^{3}\big{(}{\cal B}_{\alpha n_{j}}W^{-}_{\mu}\overline{l_{\alpha}}\gamma^{\mu}P_{L}n_{j}+{\rm h.c.}\big{)}\vspace{0.1cm}
+j=16(αfjWμlα¯γμPLfj+h.c.)),\displaystyle\hskip 34.14322pt+\sum_{j=1}^{6}\big{(}{\cal B}_{\alpha f_{j}}W^{-}_{\mu}\overline{l_{\alpha}}\gamma^{\mu}P_{L}f_{j}+{\rm h.c.}\big{)}\Big{)}, (30)

for which the definitions

αnj=(U11V^)αj,{\cal B}_{\alpha n_{j}}=\big{(}U_{11}^{*}\hat{V}^{*}\big{)}_{\alpha j}, (31)
αfj=(U12V~)αj,{\cal B}_{\alpha f_{j}}=\big{(}U_{12}^{*}\tilde{V}^{*}\big{)}_{\alpha j}, (32)

have been used. The factors (U11)αβ\big{(}U_{11}\big{)}_{\alpha\beta} and (U12)αk\big{(}U_{12}\big{)}_{\alpha k}, in Eqs. (31) and (32), are components of the matrices U11U_{11} and U12U_{12}, which are part of the generic block-matrix form of UU, given in Eq. (16). Moreover, V^βj\hat{V}_{\beta j} and V~kj\tilde{V}_{kj} are entries of the unitary matrices introduced in Eq. (21). The whole set of quantities αnj{\cal B}_{\alpha n_{j}} and αfj{\cal B}_{\alpha f_{j}} constitute the 3×33\times 3 matrix n{\cal B}_{n} and the 3×63\times 6 matrix f{\cal B}_{f}, respectively, which can be accommodated into the 3×93\times 9 matrix =(n,f){\cal B}=\big{(}{\cal B}_{n}\,,\,{\cal B}_{f}\big{)}, with entries

αj={αnk, if j=1,2,3,αfk, if j=4,5,6,7,8,9,{\cal B}_{\alpha j}=\left\{\begin{array}[]{l}{\cal B}_{\alpha n_{k}},\textrm{ if }j=1,2,3,\vspace{0.2cm}\\ {\cal B}_{\alpha f_{k}},\textrm{ if }j=4,5,6,7,8,9,\end{array}\right. (33)

where nk=n1,n2,n3n_{k}=n_{1},n_{2},n_{3} and fk=f1,f2,f3,f4,f5,f6f_{k}=f_{1},f_{2},f_{3},f_{4},f_{5},f_{6}, for any fixed α\alpha. The matrix {\cal B} fulfills a sort of semi-unitarity property, that is

j=19αjβj=δαβ,\displaystyle\displaystyle\sum_{j=1}^{9}{\cal B}_{\alpha j}{\cal B}_{\beta j}^{*}=\delta_{\alpha\beta},\vspace{0.1cm} (34)
α=e,μ,ταjαk=𝒞jk.\displaystyle\displaystyle\sum_{\alpha=e,\mu,\tau}{\cal B}_{\alpha j}^{*}{\cal B}_{\alpha k}={\cal C}_{jk}. (35)

The quantities 𝒞jk{\cal C}_{jk}, involved in Eq. (35), form a 9×99\times 9 matrix, 𝒞{\cal C}, which can be written in block-matrix form as

𝒞=(𝒞nn𝒞nf𝒞fn𝒞ff),{\cal C}=\left(\begin{array}[]{cc}{\cal C}_{nn}&{\cal C}_{nf}\vspace{0.1cm}\\ {\cal C}_{fn}&{\cal C}_{ff}\end{array}\right), (36)

where 𝒞nn{\cal C}_{nn} is 3×33\times 3, the size of 𝒞ff{\cal C}_{ff} is 6×66\times 6, 𝒞nf{\cal C}_{nf} is a 3×63\times 6 matrix, and 𝒞fn{\cal C}_{fn} is 6×36\times 3 sized. Furthermore, the entries of 𝒞{\cal C} relate to those of these block matrices as

𝒞jk={Cnlni, if j=1,2,3 and k=1,2,3,Cnlfi, if j=1,2,3 and k=4,5,6,7,8,9,Cflni, if j=4,5,6,7,8,9 and k=1,2,3,Cflfi, if j=4,5,6,7,8,9 and k=4,5,6,7,8,9,{\cal C}_{jk}=\left\{\begin{array}[]{l}C_{n_{l}n_{i}},\textrm{ if }j=1,2,3\textrm{ and }k=1,2,3,\vspace{0.2cm}\\ C_{n_{l}f_{i}},\textrm{ if }j=1,2,3\textrm{ and }k=4,5,6,7,8,9,\vspace{0.2cm}\\ C_{f_{l}n_{i}},\textrm{ if }j=4,5,6,7,8,9\textrm{ and }k=1,2,3,\vspace{0.2cm}\\ C_{f_{l}f_{i}},\textrm{ if }j=4,5,6,7,8,9\textrm{ and }k=4,5,6,7,8,9,\end{array}\right. (37)

where nl,ni=n1,n2,n3n_{l},n_{i}=n_{1},n_{2},n_{3} and fl,fi=f1,f2,f3,f4,f5,f6f_{l},f_{i}=f_{1},f_{2},f_{3},f_{4},f_{5},f_{6}. These block matrices are defined as

𝒞njnk=(Ω11TΩ11)jk,{\cal C}_{n_{j}n_{k}}=\big{(}\Omega_{11}^{\rm T}\Omega_{11}^{*}\big{)}_{jk}, (38)
𝒞njfk=(Ω11TΩ12)jk,{\cal C}_{n_{j}f_{k}}=\big{(}\Omega_{11}^{\rm T}\Omega_{12}^{*}\big{)}_{jk}, (39)
𝒞fjnk=(Ω12TΩ11)jk,{\cal C}_{f_{j}n_{k}}=\big{(}\Omega_{12}^{\rm T}\Omega_{11}^{*}\big{)}_{jk}, (40)
𝒞fjfk=(Ω12TΩ12)jk.{\cal C}_{f_{j}f_{k}}=\big{(}\Omega_{12}^{\rm T}\Omega_{12}^{*}\big{)}_{jk}. (41)

From the UU block parametrization displayed in Eq. (19), the expression for MM1M^{-1}_{\rm M} as given in Eq. (24), and the explicit form shown in Eq. (26) for V~\tilde{V}, the matrices n{\cal B}_{n} and f{\cal B}_{f} can be written as

n=(𝟏312mDM2mD)V^,{\cal B}_{n}=\Big{(}{\bf 1}_{3}-\frac{1}{2}m_{\rm D}M^{-2}m_{\rm D}^{\dagger}\Big{)}\hat{V}^{*}, (42)
f=(i2mDM112mDM1).{\cal B}_{f}=\Big{(}\begin{array}[]{lcr}\frac{i}{\sqrt{2}}m_{\rm D}M^{-1}&&\frac{1}{\sqrt{2}}m_{\rm D}M^{-1}\end{array}\Big{)}. (43)

In the so-called “minimally extended Standard Model” GiKi , characterized by the addition of three right-handed Dirac-neutrino fields, Yukawa terms for right-handed neutrinos give rise to neutrino mass terms. In the neutrino mass-eigenspinor basis, the lepton charged currents, jSM,Wμ=2lL¯UPMNSγμnLj^{\mu}_{{\rm SM},W}=2\,\overline{l_{L}}\,U_{\rm PMNS}\gamma^{\mu}\,n_{L}, involve lepton-flavor change, driven by the PMNS matrix, UPMNSU_{\rm PMNS}, which is the lepton-sector analogue of the Kobayashi-Maskawa matrix, lying in the quark sector. The neutrino-flavor basis is then defined by the transformation νL=UPMNSnL\nu_{L}=U_{\rm PMNS}\,n_{L}, by means of which the charged currents are simply expressed as jSM,Wμ=2lL¯γμνLj^{\mu}_{{\rm SM},W}=2\,\overline{l_{L}}\gamma^{\mu}\,\nu_{L}. An important feature of UPMNSU_{\rm PMNS} is its unitary property. A similar discussion can be developed if neutrino masses are rather defined through the inclusion, in the framework of a lepton-number-violating effective Lagrangian for Standard Model, of the Weinberg operator. Getting back to the theoretical framework of the present paper, the charged currents in which light neutrinos participate, displayed in the first line of Eq. (30), carry neutrino mixing characterized not by the UPMNSU_{\rm PMNS} unitary matrix, but by the 3×33\times 3 matrix n{\cal B}_{n}, which is not unitary. Expressing this matrix as n=(1η)V^{\cal B}_{n}=\big{(}1-\eta\big{)}\hat{V}^{*}, the 3×33\times 3 matrix

η=12mDM2mD\eta=\frac{1}{2}m_{\rm D}M^{-2}m_{\rm D}^{\dagger} (44)

is understood as an object characterizing non-unitary effects in light-neutrino mixing.

III Lepton-flavor change induced by the inverse seesaw at one loop

In this section, we present and discuss our analytical calculation of contributions from virtual light neutrinos neutrinos njn_{j}, heavy neutrinos NjN_{j}, and heavy neutral leptons XjX_{j}, to the charged-lepton-flavor-violating decay lαlβγl_{\alpha}\to l_{\beta}\,\gamma, where the indices α\alpha and β\beta label charged-lepton flavors, so α=μ,τ\alpha=\mu,\tau and β=e,μ\beta=e,\mu. In the presence of massive neutrinos, and as long as neutrino mixing takes place, such a process receives contributions from Feynman diagrams since the one-loop order, whatever Standard-Model extension is considered. Note that, in contraposition, this decay process is forbidden in the Standard Model. In general, physical processes which are either not allowed by the Standard Model or suppressed in such a framework bear great relevance, because manifestations from the high-energy description might be more easily detected.

In the context of the inverse seesaw mechanism, the contributing diagrams emerge at one loop as a result of lepton mixing in the charged currents displayed in Eq. (30), which are characterized by the matrix =(n,f){\cal B}=\big{(}{\cal B}_{n},{\cal B}_{f}\big{)}. In general, the set of Feynman rules by which Majorana fermions abide differ from those for fermions described by Dirac fields. Refs. DEHK ; GlZr provide detailed discussions on the matter. For instance, the assumption of Majorana fermions opens the possibility of having a larger number of contributing diagrams, as compared with the Dirac case. However, by following the Wick’ theorem Wick , we found no extra diagrams. Moreover, all the determined contributing diagrams were found not to distinguish among the Majorana and Dirac cases.

While gauge invariance is a key element in the construction of field theories, as it provides a criterion to build Lagrangian terms, the quantization of gauge theories imperiously requires the elimination of this symmetry, which is achieved by fixing the gauge. Then, gauge freedom ensures that any physical quantity must be gauge independent. In particular, the gauge choice must be innocuous to the decay amplitude we are calculating. The linear and the non-linear gauge-fixing approaches FLS ; Shore ; EiWu ; MeTo , where the choice of the gauge is parametrized by some gauge-fixing parameter, are well-known gauge-fixing schemes. Another, often considered, gauge choice is the “unitary gauge”, distinguished by the absence of pseudo-Goldstone bosons, which are born in the process of spontaneous symmetry breaking but which lack physical degrees of freedom. We carry out the calculation of the amplitude for lαlβγl_{\alpha}\to\,l_{\beta}\gamma in the unitary gauge, so no contributing Feynman diagrams featuring pseudo-Goldstone-boson lines are to be taken into account, in contraposition to what happens in other gauges. Therefore, the amplitude for lαlβγl_{\alpha}\to\,l_{\beta}\gamma is given by i=uβ¯Γμβαuαεμi{\cal M}=\overline{u_{\beta}}\,\Gamma_{\mu}^{\beta\alpha}u_{\alpha}\,\varepsilon^{\mu*}, where uαu_{\alpha} and uβu_{\beta} are momentum-space spinors, whereas εμ\varepsilon^{\mu} is the polarization vector associated to the electromagnetic field. The vertex function Γμβα\Gamma_{\mu}^{\beta\alpha} is expressed as

Γμβα=j=13ψ=n,N,X([Uncaptioned image]+[Uncaptioned image]+[Uncaptioned image]).\Gamma_{\mu}^{\beta\alpha}=\sum_{j=1}^{3}\sum_{\psi=n,N,X}\Bigg{(}\begin{gathered}\vspace{-0.2cm}\includegraphics[width=65.44142pt]{diag1}\end{gathered}+\begin{gathered}\vspace{-0.2cm}\includegraphics[width=65.44142pt]{diag2}\end{gathered}+\begin{gathered}\vspace{-0.2cm}\includegraphics[width=65.44142pt]{diag3}\end{gathered}\Bigg{)}. (45)

The presence of loop-momentum integrals opens the possibility of having ultraviolet-divergent contributions. About this, note that the superficial degrees of divergence of the diagrams shown in Eq. (45) are 0, 1, and 1, respectively, so these pieces might bear ultraviolet divergences growing as fast as linearly. To deal with this, a regularization method must be implemented. Among the different options, we followed the dimensional-regularization approach BoGi ; tHVe . This regularization method has the advantage of preserving gauge symmetry, ensured by fulfillment of Ward identities Ward , and is also well suited for its implementation through software tools. In the dimensional-regularization method, the dimension of spacetime is assumed to be DD, with 1 time-like dimension and D1D-1 space-like dimensions. Then, loop integrals are modified as d4k(2π)4F(k)μR4DdDk(2π)DF(k)\int\frac{d^{4}k}{(2\pi)^{4}}F(k)\to\mu_{\rm R}^{4-D}\int\frac{d^{D}k}{(2\pi)^{D}}F(k), where μR\mu_{\rm R} is the renormalization scale, which has units of mass and whose task is to preserve the units of loop integrals, whereas F(k)F(k) represents some function of the loop 4-momentum kk. An analytic continuation of the DD-dimensional loop integrals is defined by assuming the dimension DD to be a complex quantity, with D4D\to 4.

To perform the analytic calculation of the amplitudes corresponding to the contributing diagrams of Eq. (45), use has been made of the software packages FeynCalc SMO1 ; SMO2 ; MBD and Package-X Patel , implemented through Mathematica, by Wolfram. With these software tools, calculations have been carried out by following the tensor-reduction method PaVe ; DeSt . Furthermore, the momenta conventions displayed in Fig. 1

Refer to caption
Figure 1: Conventions for momenta used to carry out the calculation of the lαlβγl_{\alpha}\to\,l_{\beta}\gamma amplitude.

have been used to perform the calculation, where q=pαpβq=p_{\alpha}-p_{\beta}, due to 4-momentum conservation. The expression we found for the vertex function Γμβα\Gamma_{\mu}^{\beta\alpha} has the gauge-invariant and Lorentz-covariant structure

Γμβα=μβασμνqν+dβασμνqνγ5,\Gamma_{\mu}^{\beta\alpha}=\mu^{\beta\alpha}\,\sigma_{\mu\nu}q^{\nu}+d^{\beta\alpha}\,\sigma_{\mu\nu}q^{\nu}\gamma_{5}, (46)

where μβα\mu^{\beta\alpha} and dβαd^{\beta\alpha} are the transition magnetic form factor and the transition electric form factor, respectively NPR ; BGS . The corresponding branching ratio is then given by

Br(lαlβγ)=(mα2mβ2)38πmα3Γtot.(|μβα|2+|dβα|2),{\rm Br}(l_{\alpha}\to\,l_{\beta}\gamma)=\frac{\big{(}m_{\alpha}^{2}-m_{\beta}^{2}\big{)}^{3}}{8\pi m_{\alpha}^{3}\Gamma_{\rm tot.}}\Big{(}|\mu^{\beta\alpha}|^{2}+|d^{\beta\alpha}|^{2}\Big{)}, (47)

where mαm_{\alpha} and mβm_{\beta} respectively denote the masses of the charged-leptons lαl_{\alpha} and lβl_{\beta}, whereas Γtot.\Gamma_{\rm tot.} is the total decay rate for lαl_{\alpha}.

The transition magnetic and electric form factors, μβα\mu^{\beta\alpha} and dβαd^{\beta\alpha}, can be written as

μβα=j=13βnjαnjμnjβα+j=16βfjαfjμfjβα,\mu^{\beta\alpha}=\sum_{j=1}^{3}{\cal B}_{\beta n_{j}}{\cal B}_{\alpha n_{j}}^{*}\mu^{\beta\alpha}_{n_{j}}+\sum_{j=1}^{6}{\cal B}_{\beta f_{j}}{\cal B}_{\alpha f_{j}}^{*}\mu^{\beta\alpha}_{f_{j}}, (48)
dβα=j=13βnjαnjdnjβα+j=16βfjαfjdfjβα.d^{\beta\alpha}=\sum_{j=1}^{3}{\cal B}_{\beta n_{j}}{\cal B}_{\alpha n_{j}}^{*}d^{\beta\alpha}_{n_{j}}+\sum_{j=1}^{6}{\cal B}_{\beta f_{j}}{\cal B}_{\alpha f_{j}}^{*}d^{\beta\alpha}_{f_{j}}. (49)

The μnjβα\mu^{\beta\alpha}_{n_{j}} and dnjβαd^{\beta\alpha}_{n_{j}} factors, corresponding to the sum of contributing Feynman diagrams which exclusively involve the jj-th virtual light neutrino njn_{j}, are functions of the mass mnjm_{n_{j}}, whereas μfjβα\mu^{\beta\alpha}_{f_{j}} and dfjβαd^{\beta\alpha}_{f_{j}}, associated to the sum of diagrams in which only the jj-th virtual heavy-neutral lepton fjf_{j} participates, is either mNjm_{N_{j}} or mXjm_{X_{j}} dependent. All the aforementioned form-factor contributions depend on the masses mαm_{\alpha} and mβm_{\beta}, of the external leptons, while they also depend on the WW-boson mass, mWm_{W}. Moreover, since use has been made of the tensor-reduction method, the resulting expressions for all these contributing form factors are given in terms of Passarino-Veltman scalar functions PaVe . More precisely, 2-point and 3-point Passarino-Veltman scalar functions, respectively defined as

B0(p12,m02,m12)=(2πμR)4Diπ2dDk1(k2m02)((k+p1)2m12),B_{0}(p_{1}^{2},m_{0}^{2},m_{1}^{2})=\frac{(2\pi\mu_{\rm R})^{4-D}}{i\pi^{2}}\int d^{D}k\frac{1}{\big{(}k^{2}-m_{0}^{2}\big{)}\big{(}(k+p_{1})^{2}-m_{1}^{2}\big{)}}, (50)
C0(p12,(p1p2)2,p22,m02,m12,m22)=(2πμR)4Diπ2dDk1(k2m02)((k+p1)2m12)((k+p2)2m22),C_{0}(p_{1}^{2},(p_{1}-p_{2})^{2},p_{2}^{2},m_{0}^{2},m_{1}^{2},m_{2}^{2})=\frac{(2\pi\mu_{\rm R})^{4-D}}{i\pi^{2}}\int d^{D}k\frac{1}{\big{(}k^{2}-m_{0}^{2}\big{)}\big{(}(k+p_{1})^{2}-m_{1}^{2}\big{)}\big{(}(k+p_{2})^{2}-m_{2}^{2}\big{)}}, (51)

are involved. The scalar functions B0B_{0} are the sources of ultraviolet divergences, whereas 3-point functions C0C_{0} are finite in this sense. A latent drawback of our gauge choice, regarding the ultraviolet behavior of the amplitude Γμβα\Gamma_{\mu}^{\beta\alpha}, is that the growth of ultraviolet divergences might be worsened by the unitary gauge because, by this election, the gauge-boson propagators increase the superficial degree of divergence of the diagrams and, thus, has the potential of complicating the elimination of ultraviolet divergences. However, let us point out that any factor μnjβα\mu^{\beta\alpha}_{n_{j}}, μfjβα\mu^{\beta\alpha}_{f_{j}}, dnjβαd^{\beta\alpha}_{n_{j}}, or dfjβαd^{\beta\alpha}_{f_{j}} has the following structure: kB0(k)f(k)+nC0(n)g(n)+h\sum_{k}B_{0}^{(k)}f^{(k)}+\sum_{n}C_{0}^{(n)}g^{(n)}+h, where f(k)f^{(k)} is a mass-dependent function accompanied by some 2-point scalar function, B0(k)B_{0}^{(k)}, whereas g(n)g^{(n)} is another function depending on masses, which appears multiplied by a 3-point scalar function, here referred to as C0(n)C_{0}^{(n)}. The sums k\sum_{k} and n\sum_{n} run over all the 2- and 3-point Passarino-Veltman scalar functions featured by the partial contribution under consideration. Finally, hh is another mass-dependent function, which is not multiplied by any loop scalar function. We have been able to verify that the first term, kB0(k)f(k)\sum_{k}B_{0}^{(k)}f^{(k)}, is written as a sum made exclusively of terms of the form (B0(r)B0(s))w(r,s)\big{(}B_{0}^{(r)}-B_{0}^{(s)}\big{)}w^{(r,s)}, with w(r,s)w^{(r,s)} some combination of the aforementioned mass-dependent functions f(k)f^{(k)}. In other words, any partial form-factor contribution μnjβα\mu^{\beta\alpha}_{n_{j}}, μfjβα\mu^{\beta\alpha}_{f_{j}}, dnjβαd^{\beta\alpha}_{n_{j}}, or dfjβαd^{\beta\alpha}_{f_{j}} can be written in such a way that the presence of 2-point functions exclusively occurs as differences among pairs of them. Keep in mind that all the B0B_{0} functions, no matter what their arguments are, share the same divergent part, that is, B0(k)=Δdiv.+ηfin.(k)B_{0}^{(k)}=\Delta_{\rm div.}+\eta^{(k)}_{\rm fin.} for all kk, where Δdiv.\Delta_{\rm div.} gathers all the divergent contribution, whereas ηfin.(k)\eta^{(k)}_{\rm fin.} represents the non-divergent part, which is determined by the specific arguments of the 2-point function under consideration. Then, differences B0(r)B0(s)B_{0}^{(r)}-B_{0}^{(s)} are ultraviolet finite, thus implying that μnjβα\mu^{\beta\alpha}_{n_{j}}, μfjβα\mu^{\beta\alpha}_{f_{j}}, dnjβαd^{\beta\alpha}_{n_{j}}, or dfjβαd^{\beta\alpha}_{f_{j}} are free of ultraviolet divergences.

While the matrix {\cal B} is not unitary, Eq. (34) allows for a sort of Glashow-Iliopoulos-Maiani mechanism to operate GIM . An adequate implementation of this mechanism, during numerical evaluation, is imperative in order to avoid inaccurately large contributions. By usage of Eq. (34), with the flavor-change assumption αβ\alpha\neq\beta, the transition electromagnetic moments μβα\mu^{\beta\alpha} and dβαd^{\beta\alpha} can be conveniently written as

μβα=j=13βnjαnj(μnjβαμf6βα)\displaystyle\mu^{\beta\alpha}=\sum_{j=1}^{3}{\cal B}_{\beta n_{j}}{\cal B}_{\alpha n_{j}}^{*}\big{(}\mu_{n_{j}}^{\beta\alpha}-\mu_{f_{6}}^{\beta\alpha}\big{)}
+j=15βfjαfj(μfjβαμf6βα),\displaystyle\hskip 22.76228pt+\sum_{j=1}^{5}{\cal B}_{\beta f_{j}}{\cal B}_{\alpha f_{j}}^{*}\big{(}\mu_{f_{j}}^{\beta\alpha}-\mu_{f_{6}}^{\beta\alpha}\big{)}, (52)
dβα=j=13βnjαnj(dnjβαdf6βα)\displaystyle d^{\beta\alpha}=\sum_{j=1}^{3}{\cal B}_{\beta n_{j}}{\cal B}_{\alpha n_{j}}^{*}\big{(}d_{n_{j}}^{\beta\alpha}-d_{f_{6}}^{\beta\alpha}\big{)}
+j=15βfjαfj(dfjβαdf6βα).\displaystyle\hskip 22.76228pt+\sum_{j=1}^{5}{\cal B}_{\beta f_{j}}{\cal B}_{\alpha f_{j}}^{*}\big{(}d_{f_{j}}^{\beta\alpha}-d_{f_{6}}^{\beta\alpha}\big{)}. (53)

Now notice that all contributions μn1βα\mu^{\beta\alpha}_{n_{1}}, μn2βα\mu^{\beta\alpha}_{n_{2}}, μn3βα\mu^{\beta\alpha}_{n_{3}}, μf1βα\mu^{\beta\alpha}_{f_{1}}, μf2βα\mu^{\beta\alpha}_{f_{2}}, μf3βα\mu^{\beta\alpha}_{f_{3}}, μf4βα\mu^{\beta\alpha}_{f_{4}}, μf5βα\mu^{\beta\alpha}_{f_{5}}, μf6βα\mu^{\beta\alpha}_{f_{6}} come from Feynman diagrams sharing the very same structure, only distinguished among each other by the virtual neutral lepton flowing through the loop, which can be appreciated from the diagrammatic expression displayed in Eq. (45). For this reason, these partial contributions to μβα\mu^{\beta\alpha} differ of each other only by their neutral-lepton-mass dependence. For instance, if the change mn3mf5m_{n_{3}}\to m_{f_{5}} is implemented in μn3βα\mu_{n_{3}}^{\beta\alpha}, the resulting expression is the one for μf5βα\mu_{f_{5}}^{\beta\alpha}. The same argumentation applies for the electric-dipole partial contributions dn1βαd^{\beta\alpha}_{n_{1}}, dn2βαd^{\beta\alpha}_{n_{2}}, dn3βαd^{\beta\alpha}_{n_{3}}, df1βαd^{\beta\alpha}_{f_{1}}, df2βαd^{\beta\alpha}_{f_{2}}, df3βαd^{\beta\alpha}_{f_{3}}, df4βαd^{\beta\alpha}_{f_{4}}, df5βαd^{\beta\alpha}_{f_{5}}, df6βαd^{\beta\alpha}_{f_{6}}. Then note that terms which do not depend on heavy-neutral-lepton mass vanish from the differences (μnjβαμf6βα)\big{(}\mu_{n_{j}}^{\beta\alpha}-\mu_{f_{6}}^{\beta\alpha}\big{)}, (μfjβαμf6βα)\big{(}\mu_{f_{j}}^{\beta\alpha}-\mu_{f_{6}}^{\beta\alpha}\big{)}, (dnjβαdf6βα)\big{(}d_{n_{j}}^{\beta\alpha}-d_{f_{6}}^{\beta\alpha}\big{)}, and (dfjβαdf6βα)\big{(}d_{f_{j}}^{\beta\alpha}-d_{f_{6}}^{\beta\alpha}\big{)}, in Eqs. (52) and (53). Moreover, notice that further cancellations from these differences take place, thus yielding a delicate balance in which a fine suppression of contributions happens.

IV Estimations and discussion of the contributions

Now we turn to our numerical estimations of contributions. Charged-lepton-flavor-violating decays lαlβγl_{\alpha}\to\,l_{\beta}\gamma serve as means to search for new-physics traces. On the other hand, in the presence of nonzero neutrino masses and lepton mixing, these decays can be generated since the one-loop level. Up to these days, such decay processes have never been observed, while stringent bounds are available PDG . In Ref. MEGlfvbound , the MEG Collaboration reported their results on a study of the decay μ+e+γ\mu^{+}\to e^{+}\gamma, from which the upper limit Br(μ+e+γ)MEG<4.2×1013{\rm Br}\big{(}\mu^{+}\to e^{+}\gamma\big{)}_{\rm MEG}<4.2\times 10^{-13}, at the 90%C.L.90\%\,{\rm C.L.}, was set on the branching ratio for this decay. Moreover, according to Ref. MEG2lfvestimation , the upcoming MEG II detector will be able to search for μ+e+γ\mu^{+}\to e^{+}\gamma with an improved sensitivity of 6×10146\times 10^{-14}. A search for the tau lepton-flavor-violating decays τeγ\tau\to e\gamma and τμγ\tau\to\mu\gamma was reported in Ref. BaBarlfvbound , by the BaBar Collaboration, where the bounds Br(τeγ)BaBar<3.3×108{\rm Br}\big{(}\tau\to e\gamma\big{)}_{\rm BaBar}<3.3\times 10^{-8} and Br(τμγ)BaBar<4.4×108{\rm Br}\big{(}\tau\to\mu\gamma\big{)}_{\rm BaBar}<4.4\times 10^{-8} were derived, both at the 90%C.L.90\%\,{\rm C.L.} A more recent analysis on these tau decays has been carried out by the Belle Collaboration, which in Ref. Bellelfvbound presented the limits Br(τeγ)Belle<5.6×108{\rm Br}\big{(}\tau\to e\gamma\big{)}_{\rm Belle}<5.6\times 10^{-8} and Br(τμγ)Belle<4.2×108{\rm Br}\big{(}\tau\to\mu\gamma\big{)}_{\rm Belle}<4.2\times 10^{-8} at the 90%C.L.90\%\,{\rm C.L.} Note that the Belle II Collaboration has projected an increased sensitivity to these tau decays of order 10910^{-9}, namely, bounds as stringent as Br(τeγ)BelleII<9.0×109{\rm Br}\big{(}\tau\to e\gamma\big{)}_{\rm Belle\,II}<9.0\times 10^{-9} and Br(τμγ)BelleII<6.9×109{\rm Br}\big{(}\tau\to\mu\gamma\big{)}_{\rm Belle\,II}<6.9\times 10^{-9} are expected from this upgrade Belle2lfvestimation .

The mystery of whether neutrinos were massive prevailed for quite some time, since their introduction, in 1930, until the first measurements of neutrino oscillations, at the Super-Kamiokande and at the Sudbury Neutrino Observatory, reported in 1998 and 2002 nuoscillationsSKamiokande ; nuoscillationsSNO . Neutrino oscillations is a quantum phenomenon by which the probability of measuring a neutrino with a lepton flavor different from the one that originally characterized it is nonzero, after the neutrino has traveled across some distance from its source. Pontecorvo . Massiveness of neutrinos is a necessary condition for neutrino oscillations to occur, so the observation of this phenomenon has been interpreted as solid evidence supporting nonzero neutrino masses. Clearly, this experimental fact contradicts the Standard Model, where neutrinos are massless, and thus incarnates a manifestation of new physics. While valuable data on quadratic neutrino-mass differences Δmjk2=mnj2mnk2\Delta m_{jk}^{2}=m_{n_{j}}^{2}-m_{n_{k}}^{2} have been extracted from several experimental facilities focused on measurements of neutrino oscillations KamLANDdmass ; SKamiokandedmass1 ; SKamiokandedmass2 ; RENOdmass ; MINOSdmass ; NOvAdmass ; IceCubedmass ; T2Kdmass ; DayaBaydmass , the absolute neutrino-mass scale cannot be determined by this mean. An upper limit on the sum of light-neutrino masses as stringent as jmnj<0.12eV\sum_{j}m_{n_{j}}<0.12\,{\rm eV} has been determined, at 95%C.L.95\%\,{\rm C.L.}, from cosmological observations ApacheObservatorynumass ; Plancknumass . Moreover, under the assumption that neutrinos are Majorana fermions, and in view of the lack of measurements of the elusive neutrinoless double beta decay nohayndbd ; CUPIDMOndbd ; CUOREndbd ; GERDAndbd ; Majoranandbd ; EXO200ndbd ; KamLANDZenndbd , estimations of upper bounds on the neutrino effective mass mββ=|j(UPMNS)ejmnj|m_{\beta\beta}=\big{|}\sum_{j}\big{(}U_{\rm PMNS}\big{)}_{ej}m_{n_{j}}\big{|}, lying within the 102eV101eV\sim 10^{-2}\,{\rm eV}-10^{-1}\,{\rm eV} energy range, have been established through exploration of different isotopes CUOREndbd ; KamLANDZenndbd ; GERDAndbd . A constraint on the effective electron anti-neutrino mass mνe2(eff)=j|(Uν)ej|2mνj2m_{\nu_{e}}^{2({\rm eff})}=\sum_{j}|\big{(}U_{\nu}\big{)}_{ej}|^{2}m_{\nu_{j}}^{2}, established by the KATRIN Collaboration KATRIN , has been translated into the upper limit 0.8eV0.8\,{\rm eV} on neutrino mass. While the KATRIN result is not the most stringent constraint, a worth comment on this bound regards its generality, as it is independent on cosmological assumptions and on whether neutrinos are Dirac or Majorana. In what follows, we take the neutrino-mass upper bound by KATRIN as reference. Neutrino masses, in normal hierarchy (NH) or inverted hierarchy (IH), can be related to quadratic mass differences as

NH{m1=m32Δm312,m2=m32Δm322,\textrm{NH}\left\{\begin{array}[]{l}m_{1}=\sqrt{m_{3}^{2}-\Delta m_{31}^{2}},\vspace{0.3cm}\\ m_{2}=\sqrt{m_{3}^{2}-\Delta m_{32}^{2}},\end{array}\right. (54)
IH{m1=m22Δm212,m3=m22+Δm322,\textrm{IH}\left\{\begin{array}[]{l}m_{1}=\sqrt{m_{2}^{2}-\Delta m_{21}^{2}},\vspace{0.3cm}\\ m_{3}=\sqrt{m_{2}^{2}+\Delta m_{32}^{2}},\end{array}\right. (55)

where either Δm312<m3<0.8eV\sqrt{\Delta m_{31}^{2}}<m_{3}<0.8\,{\rm eV}, in the NH scheme, or Δm322<m2<0.8eV\sqrt{-\Delta m_{32}^{2}}<m_{2}<0.8\,{\rm eV}, if IH is considered. The PDG recommends the values

NH{Δm212=(7.53±0.18)×105eV2,Δm322=(2.455±0.028)×103eV2,\textrm{NH}\left\{\begin{array}[]{l}\Delta m_{21}^{2}=\big{(}7.53\pm 0.18\big{)}\times 10^{-5}\,{\rm eV}^{2},\vspace{0.3cm}\\ \Delta m_{32}^{2}=\big{(}2.455\pm 0.028\big{)}\times 10^{-3}\,{\rm eV}^{2},\end{array}\right. (56)
IH{Δm212=(7.53±0.18)×105eV2,Δm322=(2.529±0.029)×103eV2,\textrm{IH}\left\{\begin{array}[]{l}\Delta m_{21}^{2}=\big{(}7.53\pm 0.18\big{)}\times 10^{-5}\,{\rm eV}^{2},\vspace{0.3cm}\\ \Delta m_{32}^{2}=\big{(}-2.529\pm 0.029\big{)}\times 10^{-3}\,{\rm eV}^{2},\end{array}\right. (57)

for the light-neutrino quadratic-mass differences Δm212\Delta m_{21}^{2} and Δm322\Delta m_{32}^{2}. We have observed that our results do not appreciably change if either of the light-neutrino mass orderings is assumed, so from here on all our estimations are carried out by taking the NH.

Aiming at getting further insight regarding the relevance on lαlβγl_{\alpha}\to\,l_{\beta}\gamma of the non-unitary effects of light-neutrino mixing, comprised by η\eta, we define the dimensionless ratios xnj=mnj2mW2x_{n_{j}}=\frac{m_{n_{j}}^{2}}{m_{W}^{2}} and xNj=mNj2mW2x_{N_{j}}=\frac{m_{N_{j}^{2}}}{m_{W}^{2}}, and then we consider the electromagnetic-moment contributions μβα\mu^{\beta\alpha} and dβαd^{\beta\alpha}, shown in Eqs. (48) and (49), in the limit as xnj0x_{n_{j}}\to 0 and xNjx_{N_{j}}\to\infty. Note that, as shown in Eq. (45), each of the one-loop Feynman diagrams contributing to the amplitude for lαlβγl_{\alpha}\to\,l_{\beta}\gamma involves only one neutral-lepton loop line, so the complete set of diagrams can be split into those in which light neutrinos njn_{j} participate and diagrams with heavy-neutral leptons fj=Nj,Xjf_{j}=N_{j},X_{j}. Thus, the two limits of interest can be taken separately in the corresponding terms. The resulting expressions read

limxnj0xNjμβα=μ^βαηβα,\lim_{\begin{subarray}{c}x_{n_{j}}\to 0\\ x_{N_{j}}\to\infty\end{subarray}}\mu^{\beta\alpha}=\hat{\mu}^{\beta\alpha}\,\eta_{\beta\alpha}, (58)
limxnj0xNjdβα=d^βαηβα,\lim_{\begin{subarray}{c}x_{n_{j}}\to 0\\ x_{N_{j}}\to\infty\end{subarray}}d^{\beta\alpha}=\hat{d}^{\beta\alpha}\,\eta_{\beta\alpha}, (59)

both of which are given in terms of the αβ\alpha\beta-th entry of η\eta. Further, the μ^βα\hat{\mu}^{\beta\alpha} and d^βα\hat{d}^{\beta\alpha} factors have the following expressions:

μ^αβ=χ^αβmα+mβ(2mα2(mα+mβ)2\displaystyle\hat{\mu}_{\alpha\beta}=\frac{\hat{\chi}_{\alpha\beta}}{m_{\alpha}+m_{\beta}}\Big{(}-2m_{\alpha}^{2}(m_{\alpha}+m_{\beta})^{2}
+3mαmβ(mαmβmW2)(mαmβ+2mW2)\displaystyle\hskip 11.38092pt+3\frac{m_{\alpha}}{m_{\beta}}(m_{\alpha}m_{\beta}-m_{W}^{2})(m_{\alpha}m_{\beta}+2m_{W}^{2})
+3(mW2mα2)(mβ22mW4mα3(2mα+mβ)\displaystyle\hskip 11.38092pt+3(m_{W}^{2}-m_{\alpha}^{2})\Big{(}m_{\beta}^{2}-2\frac{m_{W}^{4}}{m_{\alpha}^{3}}(2m_{\alpha}+m_{\beta})
+mW2mα2(2mα+mβ)(2mα+3mβ))log(mW2mW2mα2)\displaystyle\hskip 11.38092pt+\frac{m_{W}^{2}}{m_{\alpha}^{2}}(2m_{\alpha}+m_{\beta})(2m_{\alpha}+3m_{\beta})\Big{)}\log\bigg{(}\frac{m_{W}^{2}}{m_{W}^{2}-m_{\alpha}^{2}}\bigg{)}
+6mW2mα2(2mα23mαmβ2mβ2+2mW2)C0)\displaystyle\hskip 11.38092pt+6m_{W}^{2}m_{\alpha}^{2}(-2m_{\alpha}^{2}-3m_{\alpha}m_{\beta}-2m_{\beta}^{2}+2m_{W}^{2})C_{0}\Big{)}
+(αβ),\displaystyle\hskip 11.38092pt+\big{(}\alpha\longleftrightarrow\beta\big{)}, (60)
d^αβ=iχ^αβmαmβ(2mα2(mαmβ)2\displaystyle\hat{d}_{\alpha\beta}=\frac{i\hat{\chi}_{\alpha\beta}}{m_{\alpha}-m_{\beta}}\Big{(}-2m_{\alpha}^{2}(m_{\alpha}-m_{\beta})^{2}
3mαmβ(mαmβ+mW2)(mαmβ2mW2)\displaystyle\hskip 11.38092pt-3\frac{m_{\alpha}}{m_{\beta}}(m_{\alpha}m_{\beta}+m_{W}^{2})(m_{\alpha}m_{\beta}-2m_{W}^{2})
+3(mW2mα2)(mβ22mW4mα3(2mαmβ)\displaystyle\hskip 11.38092pt+3(m_{W}^{2}-m_{\alpha}^{2})\Big{(}m_{\beta}^{2}-2\frac{m_{W}^{4}}{m_{\alpha}^{3}}(2m_{\alpha}-m_{\beta})
+mW2mα2(2mαmβ)(2mα3mβ))log(mW2mW2mα2)\displaystyle\hskip 11.38092pt+\frac{m_{W}^{2}}{m_{\alpha}^{2}}(2m_{\alpha}-m_{\beta})(2m_{\alpha}-3m_{\beta})\Big{)}\log\bigg{(}\frac{m_{W}^{2}}{m_{W}^{2}-m_{\alpha}^{2}}\bigg{)}
+6mW2mα2(2mα2+3mαmβ2mβ2+2mW2)C0)\displaystyle\hskip 11.38092pt+6m_{W}^{2}m_{\alpha}^{2}(-2m_{\alpha}^{2}+3m_{\alpha}m_{\beta}-2m_{\beta}^{2}+2m_{W}^{2})C_{0}\Big{)}
(αβ),\displaystyle\hskip 11.38092pt-\big{(}\alpha\longleftrightarrow\beta\big{)}, (61)

where we have defined

χ^αβ=e33(8π)2sW2mW2(mα2mβ2),\hat{\chi}_{\alpha\beta}=\frac{e^{3}}{3(8\pi)^{2}s_{\rm W}^{2}m_{W}^{2}(m_{\alpha}^{2}-m_{\beta}^{2})}, (62)

with sW=sinθWs_{\rm W}=\sin\theta_{\rm W} the sine of the weak-mixing angle, whereas the short-hand notation

C0=C0(0,mα2,mβ2,mW2,mW2,0)C_{0}=C_{0}(0,m_{\alpha}^{2},m_{\beta}^{2},m_{W}^{2},m_{W}^{2},0) (63)

has been used. As displayed by Eqs. (60) and (61), the factors μ^αβ\hat{\mu}_{\alpha\beta} and d^αβ\hat{d}_{\alpha\beta} are, respectively, symmetric and antisymmetric with respect to their lepton-flavor indices. In the limit considered for Eqs. (58) and (59), the branching ratio for lαlβγl_{\alpha}\to\,l_{\beta}\gamma, previously given in Eq. (47), is expressed as

Br(lαlβγ)=(mα2mβ2)38πmα3Γtot.(|μ^βα|2+|d^βα|2)|ηβα|2,{\rm Br}\big{(}l_{\alpha}\to\,l_{\beta}\gamma\big{)}=\frac{\big{(}m_{\alpha}^{2}-m_{\beta}^{2}\big{)}^{3}}{8\pi m_{\alpha}^{3}\Gamma_{\rm tot.}}\Big{(}|\hat{\mu}_{\beta\alpha}|^{2}+|\hat{d}_{\beta\alpha}|^{2}\Big{)}|\eta_{\beta\alpha}|^{2}, (64)

from which the link among off-diagonal non-unitary effects ηβα\eta_{\beta\alpha}, with βα\beta\neq\alpha, and the branching ratios for lepton-flavor-changing decays lαlβγl_{\alpha}\to\,l_{\beta}\gamma can be appreciated.

The matrix η\eta is Hermitian, so only six of its entries are independent. A dedicated global fit, comprehending several scenarios in which heavy-neutral leptons are added to the Standard-Model particle content, was carried out by the authors of Ref. FHL . An ulterior work BFHLMN , which features the same authors and further collaborators, has improved and updated this analysis, establishing the preferable regions

|ηeμ|<1.2×105,\displaystyle|\eta_{e\mu}|<1.2\times 10^{-5}, (65)
|ηeτ|<8.8×104,\displaystyle|\eta_{e\tau}|<8.8\times 10^{-4}, (66)
|ημτ|<1.8×104,\displaystyle|\eta_{\mu\tau}|<1.8\times 10^{-4}, (67)

at 95%C.L.95\%\,{\rm C.L.}, on the non-unitarity parameters which are considered for the present investigation. To perform our estimations, the matrix V^\hat{V}, first introduced in Eq. (21), is related to the PMNS matrix as V^=UPMNS\hat{V}^{*}=U_{\rm PMNS}. The set of parameters defining the PMNS neutrino-mixing matrix, UPMNSU_{\rm PMNS}, is therefore another aspect to take into account for our estimations. The number of such parameters depends on whether the neutrinos are described by Dirac or Majorana fields. In the case of Majorana neutrinos, this matrix is expressed as UPMNS=UDUMU_{\rm PMNS}=U_{\rm D}U_{\rm M}, where UMU_{\rm M} is a 3×33\times 3 diagonal matrix, given in terms of the so-called Majorana phases, ϕ1\phi_{1} and ϕ2\phi_{2}, as UM=diag(1,eiϕ1,eiϕ2)U_{\rm M}={\rm diag}\big{(}1,e^{i\phi_{1}},e^{i\phi_{2}}\big{)}. For the sake of simplicity, from here on we take the values ϕ1=0\phi_{1}=0 and ϕ2=0\phi_{2}=0 for the Majorana phases. The other matrix factor in UPMNSU_{\rm PMNS} is written as

UD=(c12c13s12s13s13eiδDs12c23c12s23s13eiδDc12c23s12s23s13eiδDs23c13s12s23c12c23s13eiδDc12s23s12c23s13eiδDc23c13),U_{\rm D}=\left(\begin{array}[]{ccc}c_{12}c_{13}&s_{12}s_{13}&s_{13}e^{-i\delta_{\rm D}}\vspace{0.2cm}\\ -s_{12}c_{23}-c_{12}s_{23}s_{13}e^{i\delta_{\rm D}}&c_{12}c_{23}-s_{12}s_{23}s_{13}e^{i\delta_{\rm D}}&s_{23}c_{13}\vspace{0.2cm}\\ s_{12}s_{23}-c_{12}c_{23}s_{13}e^{i\delta_{\rm D}}&-c_{12}s_{23}-s_{12}c_{23}s_{13}e^{i\delta_{\rm D}}&c_{23}c_{13}\end{array}\right), (68)

where mixing angles θ12\theta_{12}, θ23\theta_{23}, and θ13\theta_{13} reside within trigonometric functions, briefly denoted as sjk=sinθjks_{jk}=\sin\theta_{jk} and cjk=cosθjkc_{jk}=\cos\theta_{jk}, whereas a CPCP-violating complex phase, given the name “Dirac phase”, is denoted by δD\delta_{\rm D}. Using experimental data by a number of experimental collaborations, the Particle Data Group (PDG) reports the following values for the mixing angles PDG :

sin2θ12=0.307±0.013,\displaystyle\sin^{2}\theta_{12}=0.307\pm 0.013, (69)
sin2θ23=0.546±0.0021,\displaystyle\sin^{2}\theta_{23}=0.546\pm 0.0021, (70)
sin2θ13=0.0220±0.0007.\displaystyle\sin^{2}\theta_{13}=0.0220\pm 0.0007. (71)

The PDG considered Ref. SKamiokandedmass1 , by the Super-Kamiokande Collaboration, to set the s122s_{12}^{2} value shown in Eq. (69). For s232s_{23}^{2}, Eq. (70), the PDG took the results by Super-Kamiokande SKamiokandedmass2 , Minos+ MINOSdmass , T2K T2Ks23 , NOvA NOvAs23 , and IceCube IceCubes23 . The results followed by the PDG to give the value of s132s_{13}^{2}, which is displayed in Eq. (71), were provided by DayaBay 1DayaBays13 ; 2DayaBays13 , RENO RENOdmass ; RENOs13 , and Double Chooz DoubleChoozs13 . A value for the Dirac phase has also been set by the PDG, namely,

δD=1.23±0.21πrad,\delta_{\rm D}=1.23\pm 0.21\pi\,{\rm rad}, (72)

which was obtained on the grounds of the results reported by Super-Kamiokande SKamiokandedmass2 , NOvA NOvAdmass , and T2K T2Kdmass .

As we stated above, the 3×33\times 3 matrix MM, first introduced in Eq. (4), is assumed to be connected with a high-energy scale Λ\Lambda, characteristic of some fundamental description of nature, beyond the Standard Model. We reasonably assume that such a relation goes as MΛζMM\simeq\Lambda\,\zeta_{M}, with ζM\zeta_{M} a diagonal real matrix, in accordance with our previous discussion on the diagonalization of the neutral-lepton mass matrix. Recall that both the μS\mu_{S} and the μR\mu_{R} matrices are proportional to the scale vσv_{\sigma}, at which the global U(1){\rm U}(1) symmetry is broken. With this in mind, and for the sake of simplicity, from here on we assume that these matrices are the same, that is, μS=μ\mu_{S}=\mu and μR=μ\mu_{R}=\mu, where the notation μ\mu is used in a generic manner. This means that λ(N)=λ(S)\lambda^{(N)}=\lambda^{(S)}, which we assume to be 𝒪(102){\cal O}\big{(}10^{-2}\big{)}, so μ=(vσ2×102)ζμ\mu=\big{(}\frac{v_{\sigma}}{\sqrt{2}}\times 10^{-2}\big{)}\zeta_{\mu}, where ζμ\zeta_{\mu} is a 3×33\times 3 matrix, both real and diagonal. On the other hand, we take the Yukawa matrix YνY^{\nu} to be Yν101Y^{\nu}\sim 10^{-1} MSV , so mD=(v2×101)m^Dm_{\rm D}=\big{(}\frac{v}{\sqrt{2}}\times 10^{-1}\big{)}\hat{m}_{\rm D}, with m^D\hat{m}_{\rm D} a 3×33\times 3 complex matrix. Putting all these ingredients together, we have the light-neutrino mass matrix, Eq. (25), expressed as

Mn=122(v2vσΛ2×104)V^Tm^DζM1ζμζM1m^DTV^.M_{n}=\frac{1}{2\sqrt{2}}\Big{(}\frac{v^{2}v_{\sigma}}{\Lambda^{2}}\times 10^{-4}\Big{)}\hat{V}^{\rm T}\hat{m}_{\rm D}\,\zeta_{M}^{-1}\,\zeta_{\mu}\,\zeta_{M}^{-1}\,\hat{m}_{\rm D}^{\rm T}\hat{V}. (73)

Furthermore, these assumptions allow us to write the matrix η\eta, defined in Eq. (44) and which describes non-unitary effects in light-neutrino mixing, as

η=(v24Λ2×102)m^DζM2m^D.\eta=\Big{(}\frac{v^{2}}{4\Lambda^{2}}\times 10^{-2}\Big{)}\hat{m}_{\rm D}\,\zeta^{-2}_{M}\,\hat{m}_{\rm D}^{\dagger}. (74)

A main objective of the present work is the estimation of η\eta off-diagonal entries. To this aim, information can be extracted from Eq. (73). To begin, let us assume that

v2vσΛ2×1041eV,\frac{v^{2}v_{\sigma}}{\Lambda^{2}}\times 10^{-4}\sim 1\,{\rm eV}, (75)

which gives, in passing, the value of the high-energy scale Λ\Lambda in terms of vσv_{\sigma}. Regarding the matrices ζM\zeta_{M} and ζμ\zeta_{\mu}, both of them real and diagonal, we consider values for their in-diagonal entries ranging within 0.5(ζM)jj1.50.5\leqslant\big{(}\zeta_{M}\big{)}_{jj}\leqslant 1.5 and 0.5(ζμ)jj1.50.5\leqslant\big{(}\zeta_{\mu}\big{)}_{jj}\leqslant 1.5. We also work under the assumption that m^D\hat{m}_{\rm D} is symmetric. Then, given a set of values for the masses of light neutrinos, the entries of m^D\hat{m}_{\rm D} can be determined from Eq. (73), for either the NH or the IH of light-neutrino masses PDG . It turns out that four symmetric matrix-texture solutions for m^D\hat{m}_{\rm D} are determined for each set of fixed light-neutrino masses.

With the previous discussion in mind, we provide the graphs shown in Fig. 2,

Refer to caption
Refer to caption
Refer to caption
Figure 2: Contributions from inverse-seesaw massive neutrinos to Br(lαlβγ){\rm Br}\big{(}l_{\alpha}\to l_{\beta}\gamma\big{)}, in base-10 logarithmic scale, as functions of vσv_{\sigma}. Two regions, one contained within the other, are displayed in each graph: smallest regions for heavy-neutral-lepton quasi-degenerate masses; largest regions for non-degenerate masses of heavy neutral leptons. Horizontal lines represent upper bounds from MEG and MEG II (upper panel), and from Belle, Belle II and BaBar (second and third panels).

where Br(μeγ){\rm Br}\big{(}\mu\to e\gamma\big{)}, Br(τeγ){\rm Br}\big{(}\tau\to e\gamma\big{)}, and Br(τμγ){\rm Br}\big{(}\tau\to\mu\gamma\big{)} are displayed for different values of the scale vσv_{\sigma}, varied within 0MeV<vσ10MeV0\,{\rm MeV}<v_{\sigma}\leqslant 10\,{\rm MeV}. In order to have a clearer appreciation of the orders of magnitude of the contributions, the values of the branching ratios have been plotted in base-10 logarithmic scale. Each of the graphs has two shadowed regions, which represent the resulting inverse-seesaw contributions to lαlβγl_{\alpha}\to l_{\beta}\gamma for different values of the parameters (m3,(ζM)jj,(ζμ)jj)\big{(}m_{3},(\zeta_{\rm M})_{jj},(\zeta_{\mu})_{jj}\big{)}. Keep in mind that, in each graph, the smallest of these regions is completely embedded within the other. To determine the largest region in each graph, we plotted a set of curves by giving the largest light-neutrino mass, in the NH, the values m3=Δm312m_{3}=\sqrt{\Delta m_{31}^{2}} and m3=0.8eVm_{3}=0.8\,{\rm eV}. We have also used the values (ζM)jj=0.5,1.5(\zeta_{\rm M})_{jj}=0.5,1.5 and (ζμ)jj=0.5,1.5(\zeta_{\mu})_{jj}=0.5,1.5. All these values have then been inserted into Eq. (73), to get, for each configuration defined by specific parameters (m3,(ζM)jj,(ζμ)jj)\big{(}m_{3},(\zeta_{\rm M})_{jj},(\zeta_{\mu})_{jj}\big{)}, four solutions for mDm_{\rm D}, thus getting a total of 512 curves per graph. All these curves fall within the largest regions shown in each of the graphs, which are bounded by two curves. The smallest regions in the graphs, on the other hand, correspond to the case ζM=κ𝟏3\zeta_{M}=\kappa\cdot{\bf 1}_{3}, where values for κ\kappa have been taken to range within [0.5,1.5][0.5,1.5]. This matrix texture for ζM\zeta_{M} corresponds to a quasi-degenerate spectrum of masses of the heavy neutral leptons, that is, where mfjmNm_{f_{j}}\approx m_{N}, for j=1,2,3,4,5,6j=1,2,3,4,5,6 and with mNm_{N} representing some generic mass. In this context, with less parameters involved, we have been able to generate a total of 1372 curves, all of them falling within the smallest region in each graph. Regarding the upper panel of Fig. 2, corresponding to Br(μeγ){\rm Br}\big{(}\mu\to e\gamma\big{)}, a solid horizontal line has been added, which represents the upper bound established by the MEG Collaboration, at 4.2×10134.2\times 10^{-13}. In the same graph, a dashed horizontal line at 6×10146\times 10^{-14} is displayed to represent the expected sensitivity for MEG II. In the case of the second and third panels of Fig. 2, which correspond to the tau decays, the dot-dashed line corresponds to the Belle bound, whereas the solid horizontal line stands for the limit by BaBar. Furthermore, the lower dashed line represents the expected sensitivity for Belle II. Note that the regions for our contributions to the tau decays fall well out of experimental sensitivity. For this reason, from now on we concentrate on the decay process μeγ\mu\to e\gamma.

Previously, we showed how the branching ratio for lαlβγl_{\alpha}\to\,l_{\beta}\gamma is practically determined by the non-unitarity parameters ηαβ\eta_{\alpha\beta} as far as mW2mnj2m_{W}^{2}\gg m_{n_{j}}^{2} and mfj2mW2m_{f_{j}}^{2}\gg m_{W}^{2} hold. Such a relation is depicted by the graph of Fig. 3,

Refer to caption
Figure 3: The branching ratio for μeγ\mu\to e\gamma as a function of |ημe||\eta_{\mu e}|, in comparison with the MEG current bound MEGlfvbound , the MEG II expected sensitivity MEG2lfvestimation , and upper limits on |ημe||\eta_{\mu e}| in accordance with the preferable region determined in Ref. BFHLMN .

where Br(μeγ){\rm Br}\big{(}\mu\to e\gamma\big{)} is shown as a function of the non-unitary parameter |ημe||\eta_{\mu e}|, in accordance with Eq. (64). Here, we have varied 0<|ημe|14×1060<|\eta_{\mu e}|\leqslant 14\times 10^{-6}, whereas the comportment of Br(μeγ){\rm Br}\big{(}\mu\to e\gamma\big{)} within 0 and 6×10136\times 10^{-13} is displayed. The solid horizontal line, which bounds the shadowed region from below, represents the MEG upper limit for this branching ratio, whereas the dashed horizontal line corresponds to the expected sensitivity for MEG II. Moreover, a vertical line, at |ημe|=12×106|\eta_{\mu e}|=12\times 10^{-6}, constraining from the left a further shadowed region, is the lower limit on this non-unitarity parameter, as reported in Ref. BFHLMN . From this graph, we observe that a slight improvement of the lower bound on |ημe||\eta_{\mu e}| is expected from the MEG II update, which would set the constraint |ημe|4.29×106|\eta_{\mu e}|\lesssim 4.29\times 10^{-6}, represented in this graph by the vertical dashed line.

V Summary

In this paper we have revisited the decay process lαγlβl_{\alpha}\to\gamma\,l_{\beta}, at the one-loop level, which features lepton-flavor change. Such a phenomenon is important as its observation would point towards the presence of new unknown fundamental physics, beyond the Standard Model. To this aim, the inverse seesaw mechanism, for the generation of neutrino masses, has been used as our framework. The inverse seesaw is a seesaw variant on which the explanation for tiny light-neutrino masses, currently bounded to be within the sub-eV scale, relies not only on some high-energy scale, linked to a formulation of new physics, but also on a second scale associated to the violation of lepton number, which is assumed to be quite small. The version of the inverse seesaw which we considered for the present work introduces a set of six heavy-neutral leptons, which turn out to have very similar masses by pairs. The analytic expression for the lαγlβl_{\alpha}\to\gamma\,l_{\beta} amplitude has been calculated, following the dimensional-regularization approach and tensor-reduction method, finding gauge-invariant and ultraviolet-finite results. For the resulting branching ratio, we have considered the scenario in which mW2mnj2m_{W}^{2}\gg m_{n_{j}}^{2} and mfj2mW2m_{f_{j}}^{2}\gg m_{W}^{2}, which unveils a direct link among such quantity and non-unitary effects in light-neutrino mixing, located in the charged currents featuring the Standard-Model WW boson. Using nowadays best constraints on the branching ratio of lαγlβl_{\alpha}\to\gamma\,l_{\beta} decay, established by the MEG, Belle and BaBar collaborations, we find results consistent with current bounds on non-unitarity effects. However, we have observed that projections for MEG II would be able to improve the limit on the non-unitarity parameter |ηeμ||\eta_{e\mu}| by a factor 13\sim\frac{1}{3}.

Acknowledgements

We acknowledge financial support from Conahcyt (México).

*

Appendix A

References

  • (1) G. Aad et al. (The ATLAS Collaboration), Observation of a new particle in the search for the Standard Model Higgs boson with the ATLAS detector at the LHC, Phys. Lett. B 716, 1 (2012).
  • (2) S. Chatrchyan et al. (The CMS Collaboration), Observa- tion of a new boson at mass of 125 GeV with the CMS experiment at the LHC, Phys. Lett. B 716, 30 (2012).
  • (3) S. L. Glashow, Partial-symmetries of weak interactions, Nucl. Phys. 22, 579 (1961).
  • (4) A. Salam, Weak and electromagnetic interactions, Conf. Proc. C 680519, 367 (1968).
  • (5) S. Weinberg, A model of leptons, Phys. Rev. Lett. 19, 1264 (1967).
  • (6) F. Englert and R. Brout, Broken symmetry and the mass of gauge vector mesons, Phys. Rev. Lett. 13, 321 (1964).
  • (7) P. W. Higgs, Broken symmetries and the masses of gauge bosons, Phys. Rev. Lett. 13, 508 (1964).
  • (8) B. Pontecorvo, Mesonium and anti-mesonium, Sov. Phys. JETP 6, 429 (1957).
  • (9) S. Navas et al. (Particle Data Group), Review of particle physics, to be published in Phys. Rev. D 110, 030001 (2024).
  • (10) E. Majorana, Teoria simmetrica dell’elettrone e del positrone, Nuovo Cimento 14, 171 (1937).
  • (11) C. N. Leung, S. T. Love, and S. Rao, Low-energy manifestations of a new interactions scale: Operator analysis, Z. Phys. C 31, 433 (1986).
  • (12) W. Buchmüller and D. Wyler, Effective lagrangian analysis of new interactions and flavour conservation, Nucl. Phys. B268, 621 (1986).
  • (13) J. Wudka, Electroweak effective Lagrangians, Int. J. Mod. Phys. A 09, 2301 (1994).
  • (14) S. Weinberg, Baryon- and lepton-nonconserving processes, Phys. Rev. Lett. 43, 1566 (1979).
  • (15) K. S. Babu and C. N. Leung, Classification of effective neutrino mass operators, Nucl. Phys. B619, 667 (2001).
  • (16) S. Centelles, R. Srivastava, and J. W. F. Valle, Seesaw roadmap to neutrino mass and dark matter, Phys. Lett. B 781, 122 (2018).
  • (17) R. N. Mohapatra and G. Senjanović, Neutrino mass and spontaneous parity nonconservation, Phys. Rev. Lett. 44, 912 (1980).
  • (18) R. N. Mohapatra and G. Senjanović, Neutrino masses and mixings in gauge models with spontaneous parity violation, Phys. Rev. D 23, 165 (1981).
  • (19) N. Aghanim et al. (Planck Collaboration), Planck 2018 results. VL. Cosmological parameters, Astron. Astrophys. 641, A6 (2020).
  • (20) S. Alam et al., Completed SDSS-IV extended Baryon Oscillation Spectroscopic Survey: Cosmological implications from two decades of spectroscopic surveys at the Apache Point Observatory, Phys. Rev. D 103, 083533 (2021).
  • (21) M. Aker et al. (The KATRIN Collaboration), Direct neutrino-mass measurement with sub-electronvolt sensitivity, Nature Phys. 18, 160 (2022).
  • (22) Y. Cai, T. Han, T. Li, and R. Ruiz, Lepton Number Violation: Seesaw Models and Their Collider Tests, Front. Phys. 6, 40 (2018).
  • (23) R. N. Mohapatra and J. W. F. Valle, Neutrino mass and baryon-number nonconservation in superstring models, Phys. Rev. D 34, 1642 (1986).
  • (24) M. C. Gonzalez-Garcia and J. W. F. Valle, Fast decaying neutrinos and observable flavour violation in a new class of majoron models, Phys. Lett. B 216, 360 (1989).
  • (25) F. Deppisch and J. W. F. Valle, Enhanced lepton flavor violation in the supersymmetric inverse seesaw model, Phys. Rev. D 72, 036001 (2005).
  • (26) K. -S Sun, T. -F. Feng, G. -H. Luo,, X. -Y. Yang, and J. -B . Chen, Lepton flavor violation in inverse seesaw model, Mod. Phys. Lett. A 28, 1350151 (2013).
  • (27) C. Soumya and M. Rukmani, Non-unitary lepton mixing in an inverse seesaw and its impact on the physics of long-baseline experiments, J. Phys. G: Nucl. Part. Phys. 45, 095003 (2018).
  • (28) J. C. Garnica and E. Peinado, Charged lepton-flavor violating processes and suppression of nonunitarity mixing in low-scale seesaw models, Phys. Rev. D 108 035033 (2023).
  • (29) A. M. Baldini et al. (MEG Collaboration), Search for the lepton flavour violating decay μ+e+γ\mu^{+}\to e^{+}\gamma with the full dataset of the MEG experiment, Eur. Phys. J. C 76, 434 (2016).
  • (30) K. Uno et al. (The BELLE colaboration), Search for lepton-flavor-violating tau-lepton decays to γ\ell\gamma at Belle, JHEP 10, 19 (2021).
  • (31) B. Aubert et al. (BaBar Collaboration), Searches for Lepton Flavor Violation in the Decays τ±e±γ\tau^{\pm}\to e^{\pm}\gamma and τ±μ±γ\tau^{\pm}\to\mu^{\pm}\gamma, Phys. Rev. Lett. 104, 021802 (2010).
  • (32) A. M. Baldini et al. (MEG II Collaboration), The design of the MEG II experiment, Eur. Phys. J. C 78, 380 (2018).
  • (33) L. Aggarwal et al. (Belle-II Collaboration), Snowmass white paper: Belle II physics reach and plans for the next decade and beyond, arXiv:2207.06307.
  • (34) C. Giunti and C.W. Kim, Fundamentals of Neutrino Physics and Astrophysics (Oxford University Press, New York, 2007).
  • (35) Z. Maki, M. Nakagawa, and S. Sakata, Remarks on the unified model of elementary particles, Prog. Theor. Phys. 28, 870 (1962).
  • (36) B. Pontecorvo, Neutrino experiments and the problem of conservation of leptonic charge, Sov. Phys. JETP 26, 984 (1968).
  • (37) E. Fernández-Martínez, M. B. Gavela, J. López-Pavón, and O. Yasuda, CP-violation from non-unitary leptonic mixing, Phys. Lett. B 649, 427 (2007).
  • (38) E. Fernandez-Martinez, J. Hernandez-Garcia, and J. Lopez-Pavon, Global constraints on heavy neutrino mixing, JHEP 08, 033 (2016).
  • (39) M. Blennow, E. Fernández-Martínez, J. Hernández-García, J. López-Pavón, X. Marcano, and D. Naredo-Tuero, Bounds on lepton non-unitarity and heavy neutrino mixing, JHEP 08, 030 (2023).
  • (40) Y. Chikashige, R. N. Mohapatra, and R. D. Peccei, Are there real Goldstone bosons associated with broken leton number?, Phys. Lett. B 98, 265 (1981).
  • (41) S. Mandal, J. C. Romão, R. Srivastava, and J. W. F. Valle, Dynamical inverse seesaw mechanism as a simple benchmark for electroweak breaking and Higgs boson studies, JHEP 07, 029 (2021).
  • (42) E. Kh. Akhmedov, Z. G. Berezhiani, R. N. Mohapatra, and G. Senjanović, Planck scale effects on the majoron, Phys. Lett. B 299, 90 (1993).
  • (43) V. Berezinsky and J. W. F. Valle, The keV majoron as a dark matter particle, Phys. Lett B 318, 360 (1993).
  • (44) I. Z. Rothstein, K. S. Babu, and D. Seckel, Plank scale symmetry breaking and majoron physics, Nuycl. Phys. B403, 725 (1993).
  • (45) F. Bazzocchi, M. Lattanzi, S. Riemer-Sørensen, J. W. F. Valle, X-ray photons from late-decaying majoron dark matter, JCAP 08, 013 (2008).
  • (46) P. -H. Gu, E. Ma, and U. Sarkar, Pseudo-Majoron as dark matter, Phys. Lett B 690, 145 (2010).
  • (47) J. N. Esteves, F. R. Joaquim, A. S. Joshipura, J. C. Romão, M. A. Tórtola, and J. W. F. Valle, A4A_{4}-based neutrino masses with Majoron decaying dark matter, Phys. Rev. D 82, 073008 (2010).
  • (48) C. Garcia-Cely and J. Heeck, Neutrino lines from majoron dark matter, JHEP 05 102 (2017).
  • (49) T. Brune and H. Päs, Massive Majorons and constraints on the Majoron-neutrino coupling, Phys. Rev. D 99, 096005 (2019).
  • (50) G. ’t Hooft, Naturalness, chiral symmetry, and spontaneous chiral symmetry breaking, NATO Sci. Ser. B 59 135 (1980).
  • (51) T. Takagi, On an algebraic problem related to an analytic theorem of Carathéodory and Fejér and on an allied theorem of Landau, Jpn. J. Math. 1, 83 (1925).
  • (52) J. G. Körner, A. Pilaftsis, and K. Schilcher, Leptonic CP asymmetries in flavor-changing H0H^{0} decays, Phys. Rev. D 47, 1080 (1993).
  • (53) P. S. B. Dev and A. Pilaftsis, Minimal radiative neutrino mass mechanism for inverse seesaw models, Phys. Rev. D 86, 113001 (2012).
  • (54) A. Denner, H. Eck, O. Hahn, and J. Küblbeck, Feynman rules for fermion-number-violating interactions, Nucl. Phys. B387, 467 (1992).
  • (55) J. Gluza and M. Zraek, Feynman rules for Majorana-neutrino interactions, Phys. Rev. D 45, 1693 (1992).
  • (56) G. C. Wick, The evaluation of the collision matrix, Phys. Rev. 80, 26 (1950).
  • (57) K. Fujikawa, B. W. Lee, and A. I. Sanda, Generalized renormalizable gauge formulation of spontaneously broken gauge theories, Phys. Rev. D 6, 2923 (1972).
  • (58) G. Shore, Symmetry restoration and the background field method in gauge theories, Ann. Phys. 137, 262 (1981).
  • (59) M.B. Einhorn and J. Wudka, Screening of heavy-Higgs-boson radiative effects, Phys. Rev. D 39, 2758 (1989).
  • (60) J. G. Méndez and J. J. Toscano, A nonlinear RξR_{\xi}-gauge for the electroweak theory, Rev. Mex. Fis. 50, 346 (2004).
  • (61) C. G. Bollini and J. J. Giambiagi, Dimensional renormalization: The number of dimensions as a regularizing parameter, Nuovo Cimento Soc. Ital. Fis. 12B, 20 (1972).
  • (62) G. ’t Hooft and M. Veltman, Regularization and renormalization of gauge fields, Nucl. Phys. B44, 189 (1972).
  • (63) J. C. Ward, An identity in quantum electrodynamics, Phys. Rev. 78, 182 (1950).
  • (64) V. Shtabovenko, R. Mertig, and F. Orellana, FeynCalc 9.3: New features and improvements, Comput. Phys. Commun. 256, 107478 (2020).
  • (65) V. Shtabovenko, R. Mertig, and F. Orellana, New developments in FeynCalc 9.0, Comput. Phys. Commun. 207, 432 (2016).
  • (66) R. Mertig, M. Bohm, and A. Denner, FeynCalc-Computer-algebraic calculation of Feynman amplitudes, Comput. Phys. Commun. 64, 345 (1991).
  • (67) H. H. Patel, textscPackage-X: A Mathematica package for the analytic calculation of one-loop integrals, Comput. Phys. Commun. 197, 276 (2015).
  • (68) G. Passarino and M. Veltman, One-loop corrections for e+ee^{+}e^{-} annihilation into μ+μ\mu^{+}\mu^{-} in the Weinberg model, Nucl. Phys. B160, 151 (1979).
  • (69) G. Devaraj and R. G. Stuart, Reduction of one-loop tensor form factors to scalar integrals: A general scheme, Nucl. Phys. B519, 483 (1998).
  • (70) M. Nowakowski, E. A. Paschos, and J. M. Rodríguez, All electromagnetic form factors, Eur. J. Phys. 26, 545 (2005).
  • (71) C. Broggini, C. Giunti, and A. Studenikin, Electromagnetic properties of neutrinos, Adv. High Energy Phys. 2012, 459526 (2012).
  • (72) S. L. Glashow, J. Iliopoulos and L. Maiani, Weak Interactions with Lepton-Hadron Symmetry, Phys. Rev. D 2, 1285 (1970).
  • (73) Y. Fukuda et al. (Super-Kamiokande Collaboration), Evidence for oscillation of atmospheric neutrinos, Phys. Rev. Lett. 81, 1562 (1998).
  • (74) Q. R. Ahmad et al. (SNO Collaboration), Direct evidence for neutrino flavor transformation from neutral-current interactions in the Sudbury Neutrino Observatory, Phys. Rev. Lett. 89, 011301 (2002).
  • (75) A. Gando et al. (KamLAND Collaboration), Reactor on-off antineutrino measurement with KamLAND, Pays. Rev. D 88, 033001 (2013).
  • (76) K. Abe et al. (Super-Kamiokande Collaboration), Solar neutrino measurements in Super-Kamiokande-IV, Phys. Rev. D 94, 052010 (2016).
  • (77) K. Abe et al. (Super-Kamiokande Collaboration), Atmospheric neutrino oscillation analysis with external constraints in Super-Kamiokande I-IV, Phys. Rev. D 97, 072001 (2018).
  • (78) G. Bak et al. (RENO Collaboration), Measurement of Reactor Antineutrino Oscillation Amplitude and Frequency at RENO, Phys. Rev. Lett. 121, 201801 (2018).
  • (79) P. Adamson et al. (MINOS+ Collaboration), Precision Constraints for Three-Flavor Neutrino Oscillations from the Full MINOS+ and MINOS Dataset, Phys. Rev. Lett. 125, 131802 (2020).
  • (80) M. A. Acero et al. (The NOvA Collaboration), Improved measurement of neutrino oscillation parameters by the NOvA experiment, Phys. Rev. D 106, 032004 (2022).
  • (81) R. Abbasi et al. (IceCube Collaboration), Measurement of atmospheric neutrino mixing with improved IceCube DeepCore calibration and data processing, Phys. Rev. D 108, 012014 (2023).
  • (82) K. Abe et al. (T2K Collaboration), Measurements of neutrino oscillation parameters from the T2K experiment using 3.6×10213.6\times 10^{21} protons on target, Eur. Phys. J. C 83, 782 (2023).
  • (83) F. P. An et al. (Daya Bay Collaboration), Precision Measurement of Reactor Antineutrino Oscillation at Kilometer-Scale Baselines by Daya Bay, Phys. Rev. Lett. 130, 161802 (2023).
  • (84) S. Alam et al., Completed SDSS-IV extended Baryon Oscillation Spectroscopic Survey: Cosmological implications from two decades of spectroscopic surveys at the Apache Point Observatory, Phys. Rev. D 103, 083533 (2021).
  • (85) N. Aghanim et al. (Planck Collaboration), Planck 2018 results. VL. Cosmological parameters, Astron. Astrophys. 641, A6 (2020).
  • (86) A. Gando et al. (KamLAND-Zen Collaboration), Limit on neutrinoless ββ\beta\beta decay of 136Xe from the first phase of KamLAND-Zen and comparison with the positive claim in 76Ge, Phys. Rev. Lett. 110, 062502 (2013).
  • (87) E. Armengaud et al. (CUPID-Mo Collaboration), New limit for neutrinoless double-beta decay of 100Mo from the CUPID-Mo experiment, Phys. Rev. Lett. 126, 181802 (2021).
  • (88) D.Q. Adams et al. (CUORE Collaboration), Improved limit on neutrinoless double-beta decay in 130Te with CUORE, Phys. Rev. Lett. 124, 122501 (2020).
  • (89) M. Agostini et al. (GERDA Collaboration), Final results of GERDA on the search for neutrinoless double-β\beta decay, Phys. Rev. Lett. 125, 252502 (2020).
  • (90) S.I. Alvis et al. (Majorana Collaboration), Search for neutrinoless double-β\beta decay in 76Ge with 26 kg yr of exposure from the Majorana demonstrator, Phys. Rev. C 100, 025501 (2019).
  • (91) G. Anton et al. (EXO-200 Collaboration), Search for neutrinoless double-β\beta decay with the complete EXO-200 dataset, Phys. Rev. Lett. 123, 161802 (2019).
  • (92) A. Gando et al. (KamLAND-Zen Collaboration), Search for Majorana neutrinos near the inverted mass hierarchy region with KamLAND-zen, Phys. Rev. Lett. 117, 082503 (2016).
  • (93) S. Mandal, R. Srivastava, and J. W. F. Valle, Electroweak symmetry breaking in the inverse seesaw mechanism, JHEP 03, 212 (2021).
  • (94) K. Abe et al. (The T2K Collaboration), Constraint on the matter-antimatter symmetry-violating phase in neutrino oscillations, Nature (London) 580, 339 (2020).
  • (95) M. A. Acero et al. (NOvA Collaboration), First Measurement of Neutrino Oscillation Parameters using Neutrinos and Antineutrinos by NOvA, Phys. Rev. Lett. 123, 151803 (2019).
  • (96) M. G. Aartsen et al. (IceCube Collaboration), Measurement of Atmospheric Neutrino Oscillations at 6-56 GeV with IceCube DeepCore, Phys. Rev. Lett. 120, 071801 (2018).
  • (97) D. Adey et al. (The Daya Bay Collaboration), Measurement of the Electron Antineutrino Oscillation with 1958 Days of Operation at Daya Bay, Phys. Rev. Lett. 121, 241805 (2018)
  • (98) F. P. An et al. (Daya Bay Collaboration), New measurement of θ13\theta_{13} via neutron capture on hydrogen at Daya Bay, Phys. Rev. D 93, 072011 (2016).
  • (99) C. D. Shin et al. (The RENO Collaboration), Observation of reactor antineutrino disappearance using delayed neutron capture on hydrogen at RENO, J. High Energy Phys. 04, 029 (2020).
  • (100) H. de Kerret et al. (The Double Chooz Collaboration), Double Chooz θ13\theta_{13} measurement via total neutron capture detection, Nature Phys. 16, 558 (2020).