modular invariance and the strong CP problem
We present simple effective theory of quark masses, mixing and CP violation with level () modular symmetry, which provides solution to the strong CP problem without the need for an axion. The vanishing of the strong CP-violating phase is ensured by assuming CP to be a fundamental symmetry of the Lagrangian of the theory. The CP symmetry is broken spontaneously by the vacuum expectation value (VEV) of the modulus . This provides the requisite large value of the CKM CP-violating phase while the strong CP phase remains zero or is tiny. Within the considered framework we discuss phenomenologically viable quark mass matrices with three types of texture zeros, which are realized by assigning both the left-handed and right-handed quark fields to singlets , and with appropriate weights. The VEV of is restricted to reproduce the observed CKM parameters. We discuss cases in which the modulus VEV is close to the fixed points , and . In particular, we focus on the VEV of , which gives the absolute minima of the supergravity-motivated modular- and CP-invariant potentials for the modulus , so called, modulus stabilisation. We present a successful model, which is consistent with the modulus stabilisation close to . )
1 Introduction
The strong CP problem has been a puzzle in particle physics since the QCD Lagrangian violates CP due to instanton effects [1, 2, 3, 4]. The CP violation in QCD is described by the strong CP phase
(1) |
where and denote the mass matrices of the up-type and the down-type quarks, respectively, and is the coefficient of the topological charge term in the QCD Lagrangian:
(2) |
Here is the gluon field stress tensor and is the multiplet of quark fields with definite masses. While and are transformed into each other via chiral transformation, the sum is invariant.
The upper-bound of is derived from the experimental upper bound on the electric dipole moment (edm) of the neutron as [5]. This bound is much smaller than the weak CP violation, for example, the Jarlskog rephasing invariant [6, 7]. Therefore, the strong CP problem is the problem of understanding why is so small.
The most well known solution is the axion solution [8] (see also, e.g., [9]), where is a dynamical degree of freedom which is set to a small value by a scalar potential. However, there have been no experimental hints for the existence of the axion so far.
Another well known solution is provided by the Nelson-Barr model [10, 11], where the CP symmetry is violated only by the mixings of Standard model (SM) quarks with hypothetical extra heavy quarks, and the extended quark mass matrix has a special structure with vanishing entries such that CP-violating terms do not contribute to its determinant. Specific models, in which quark mass matrices with real determinant and still generate the weak CP-violating (CPV) phase in the CKM quark mixing matrix have been proposed in [12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24] as well.
The modular invariance opened up a new promising approach to the flavour problem of quarks and leptons [25] (see also Refs.[26, 27, 28]). The strong CP problem has been also discussed recently within the modular invariance approach in Ref.[29]. In this study a simple effective theory of flavour and CP symmetry breaking with vanishing without the need for an axion has been proposed. The analysis was done using the level full modular group SL. A numerical example has shown that the modular symmetry allows to solve the strong CP problem and to reproduce correctly the quark masses and the CKM mixing matrix.
In this article we present alternative effective quark flavour models with finite modular symmetry of level (), which provide axion-less solution to the strong CP problem. The modular symmetry has been extensively used for understanding the origins of the quark and lepton flavours [25, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44]. Finite groups including finite modular groups have been widely employed in flavour model building (see, e.g., [26, 27, 45, 28, 46, 47, 48, 49, 50, 51, 52, 53] 222A rather complete list of articles on the modular invariance approach to lepton and quark flavour problems is given in [53]. and the reviews [54, 55, 56, 58, 59, 60, 61, 62, 63, 57, 64, 65]).
If CP is a fundamental symmetry of the Lagrangian, the QCD and the strong CPV phases and will vanish. However, in order to explain CP violation in weak interactions, the spontaneous breaking of the CP symmetry has to generate the large measured value of the CPV phase in the CKM matrix, while at the same time the strong CPV phase should still be zero or be tiny enough to be in agreement with experimental data on the edm of the neutron. In other words, we have to look for a texture with with a realistic value for the CKM CPV phase. Furthermore, as far as there is no accidental cancellation between phases in the up-type and down-type mass matrices, and should be real, and positive by themselves.
Since constraints of are very stringent, we need to take into account the corrections to this parameter. The most important corrections are 333 The corrections to from SM are known to be negligible [66, 67] [19, 29]:
-
•
Higher dimensional operators that spoil the structure of the mass matrices.
-
•
Corrections which are induced from supersymmetry (SUSY) breaking terms.
The first point has been addressed by introducing the modular symmetry within the modular invariance approach to the quark (lepton) flavour problem [25], in which the elements of the quark (charged lepton and neutrino) mass matrices are modular forms of certain levels and weights that are holomorphic functions of the modulus . The modular forms are essentially non-perturbative ones. Once the weight of the modular form is fixed there are no the higher dimensional operator contributions. The symmetry is sharp.
The second point is the correction due to the SUSY breaking terms. This correction has been discussed in the modular symmetry of flavours [29]. Since the modular symmetry is in the framework of SUSY, the SUSY breaking could contribute to significantly. As far as SUSY is unbroken, cannot be generated radiatively. On the other hand, the SUSY breaking sector, in principle, can introduce new sources of CP violation which are model dependent. Assuming that SUSY is broken via gauge-mediation or anomaly-mediation below the mass scale of the modulus , the renomalization group and threshold corrections due to the SUSY breaking have same flavour and CP structure as the SM ones [13]. Therefore, the corrections to are safely under control.
In the case of modular symmetry considered by us relevant texture zeros of the quark mass matrices are realized if both left-handed (LH) and right-handed (RH) quark fields are assigned to be singlets [68]. The CP violation is generated only by the vacuum expectation value (VEV) of the modulus [69]. By performing statistical analysis we obtain the values of the VEV of that allow to reproduce the observed CP violation in the quark sector.
It is known that the VEV of the modulus could be obtained by the modulus stabilisation analysing relatively simple (supergravity-motivated) modular- and CP-invariant potentials for the modulus . In the modulus stabilisation studies values of the VEV of close to the fixed points [45] , and were found [70, 71, 72, 73, 74, 75, 76, 77, 78]. In view of this we search in this work for examples of modular solutions to the strong CP problem for values of the VEV of close to the fixed points.
The paper is organised as follows. In Section 2, we present quark mass matrices with three types of texture zeros, which satisfy . In Section 3, we propose quark mass matrices with the considered texture zeros in the case of modular symmetry and discuss their CP violating phase structure. In Section 4, we show numerical examples which allow to reproduce the observed quark masses, CKM mixing angles and CP-violating phase for the VEV of close to the fixed points , and . In Section 5, we focus on the VEV of , which gives the absolute minima of the supergravity-motivated modular- and CP-invariant potentials for the modulus , so called, modulus stabilisation. We present a successful model, which is consistent with the modulus stabilisation close to [73]. Section 6 is devoted to the summary and discussions. In Appendix A the modular forms of weights up to 16 are presented. In Appendix B we list the observed values of the quark masses, the CKM elements and CPV phase, which serve as input for our numerical analyses. In Appendix C we present numerically three examples of phenomenologically viable quark mass matrices with texture zeros that allow to describe the data on quark masses, mixing and CP violation.
2 Texture zeros and Real
The texture zeros approach has a long history. In Ref.[80] Weinberg, assuming the existence of two families of quarks, considered a mass matrix for the down-type quark sector with zero (1,1) entry in the basis in which the up-type quark mass matrix is diagonal. He further supposed that the down-type quark mass matrix is symmetric. In this case the number of free parameters is reduced to only two and hence he succeeded to predict the Cabibbo angle to be , which is the so-called Gatto, Sartori, Tonin relation [81]. Fritzsch extended the above approach to the three family case [82, 83]. Ramond, Roberts and Ross presented a systematic analysis with four or five zeros for symmetric or hermitian quark mass matrices [84]. Their textures are not viable today since they cannot describe the current rather precise data on the CKM quark mixing matrix. However, the texture zero approach to the quark mass matrices is still promising [86, 85] 444By making a suitable weak basis transformation, one can obtain some sets of zeros of the quark mass matrices. This issue was discussed in Refs. [87, 88, 89].
A systematic study of texture zeros has been presented for the down-type quark mass matrix in the basis of diagonal up-type quark mass matrix in Ref.[90] from the standpoint of “Occam’s Razor approach” [91], in which a minimum number of parameters is allowed. The down-type quark mass matrix was arranged to have the minimum number of parameters by setting three of its elements to zero, while at the same time requiring that it describes successfully the CKM mixing and CP violation without assuming it to be symmetric or hermitian. Some of the texture zeros considered in Ref. [90] lead to real . For the purpose of our study, where the modular invariance of the quark mass matrices determines the texture zeros, we discuss three sets of texture zeros for both the down-type and up-type mass matrices, , which are texture zeros of the corresponding quark Yukawa couplings 555There are thirteen available textures with three zeros, of which six textures lead to real [90]. They give three different relations between the CKM angles and the quark masses as seen in Table 1 of Ref.[90]. The three textures in Eq. (3) are representatives of the indicated six textures. : , , being the VEVs of the down-type and up-type doublet Higgs fields. In the right-left (RL) convention for the mass matrices , which we are going employ in our analysis, the three sets of textures we were referring to above are:
(3) |
where , and are real coefficients. The texture (1) is used in Ref. [29] for the discussion of the strong CP problem. The textures (2) and (3) are obtained by exchange of columns and of the matrix (1), respectively. The physics (the CKM matrix) is different among them. On the other hand, the physics is not changed by the exchange of the rows of (1) since it respects only the right-handed sector. The CPV phases and are assigned to specific entries so that the mass matrices in Eq. (3) coincide in form with those obtained from modular invariance in our further analysis. As we will see, the modular invariance constraints the down-type and up-type quark mass matrices we will consider to have each only one CPV phase originating from the VEV of the modulus . The determinants of the considered mass matrices are real:
(4) |
We show typical examples of numerical fits for the three textures of Eq. (3) in Appendix C, for which the input data are given in Appendix B.
3 Realization of texture zeros in modular symmetry
We will present next modular invariant mass matrices with level () modular symmetry. They have the texture zeros of the matrices given in Eq. (3). In order to realize the desirable texture zeros, all quarks should be assigned to singlets. However, the modular forms representing trivial singlets 1 are just equivalent to the Eisenstein series, which correspond to the modular forms of level modular symmetry, as shown in Eq. (57) of Appendix A. An example of modular invariant mass matrix with level modular symmetry was presented in Ref. [29]. In this work, we employ modular forms that are non-trivial singlets and in addition to forms which are trivial singlets 1. In order to reproduce the observed CP violation of the quark sector, two singlet modular forms with same weight furnishing the same singlet representation are required to avoid the vanishing of the CPV phase due to cancellation between the contributions for the down-type and up-type quark mass matrices. The smallest weight at which there exists two modular forms of the same singlet representation is , but the forms are trivial 1 singlets. Two non-trivial singlet modular forms of the same weight exist at and at larger weights, as seen in Appendix A. In this section, using two singlet modular forms with weight together with additional singlet modular forms, we construct quark mass matrices with modular symmetry, which have the general forms given in Eq. (3).
3.1 Quark mass matrix of model (1) and its CPV phase structure
In order to keep to be real, the following condition of the weights are required [29]:
(5) |
where and denote weights for the left-handed quarks and right-handed down (up)-type quarks, respectively. The weights of Higgs are set to be zero. The condition of Eq.(5) guarantees that the modular symmetry has no QCD anomaly.
We assign the representations and the weights for quarks and Higgs as follows:
-
•
quark doublet (left-handed) : singlets with weight .
-
•
quark singlets (righ-handed) and : singlets with weight , respectively.
-
•
Higgs fields of down-type and up-type quark sectors : singlet with weight 0.
The assigned weights satisfy the condition in Eq.(5). These assignments are summarized in Table 1.
2 | 1 | 2 | 2 | |
0 | 0 |
Taking account of the following tensor products of singlets [55, 56, 57],
(6) |
the superpotential terms and of the down-type and up-type quark superfields with weights as given in Table 1 read:
where are modular forms of with weight as seen in Appendix A. Parameters and are constants. These constants are real as a consequence of the imposed CP invariance of the Lagrangian of the theory. We note that and involve three modular forms that are singlets, , and , in addition to the trivial singlet one .
The Yukawa matrices of the down-type and up-type quarks follow from the expressions for and and are given by:
(8) |
We take the modular invariant kinetic terms simply as 666Possible non-minimal additions to the Kähler potential, compatible with the modular symmetry, may jeopardise the predictive power of the approach [92]. However, those do not affect as discussed in Ref.[29].
(9) |
where denotes a chiral superfield with weight , and is the anti-holomorphic modulus. The (anti-)holomorphic modulus is a dynamical field. It becomes a complex number after the modulus takes a VEV. Then, one can set .
It is important to address the transformation needed to get the kinetic terms of matter superfields in canonical form because the terms in Eq. (9) are not canonical. The canonical form is obtained by an overall re-normalization of the quark superfields of interest:
(10) |
This changes some of the constant parameters in the quark mass matrices in the following way:
(11) |
The constants , , remain unchanged. Thus, the mass matrices of the down-type and up-type quarks take the form:
(12) |
The determinants of the quark mass matrices and given in the preceding equation are real:
(13) |
The phase structure of the matrices in Eq. (12) is rather simple. The CP violating phases originate from the VEV of the modulus via the modular forms , , and . The phases of and in the and elements can be factorised as follows:
(14) |
The phase matrices and are given by:
(15) |
where
(16) |
The matrix does not appear in the CKM mixing matrix. Since is common in down-type quark and up-type quark mass matrices, the phase matrix is cancelled out in the CKM matrix due to . Thus, the CP violation originates from the phases of the elements of and in Eq. (14):
(17) |
These phases are fixed by the values of the VEV of modulus and of the real parameters and . Due to the two singlet weight 16 modular forms, non-trivial CP-violating phases appear in down-type and up-type quark mass matrices as well as in the CKM matrix.
In Appendix C.1 we show that the mass matrices , , given in Eq. (14), which correspond to the first (upper) matrix of Yukawa couplings in Eq. (3), are completely consistent with the observed quark masses and CKM quark mixing and CP violation when their elements have the numerical values reported in Eq. (81) of Appendix C.1 and the CPV phases in the (3,3) elements in and have the values and , respectively. This implies that the quark matrices in Eq. (14), obtained using the modular invariance, could be consistent with the data on quark masses, mixing and CP violation if the phases and in Eq. (17) have the indicated values. If we will be able to adjust the VEV of the modulus and the values of the real constants and so that the phases and get these values, we will have viable quark mass matrices since the constants and can be used to reproduce the numerical values in the matrices in Eq. (81). These numerical values are derived by fitting the quark masses, the CKM matrix elements and the Jarlskog rephasing invariant quoted in Eqs. (73), (75) and (78) in Appendix B.
We plot in Fig. 2 the values of the modulus 777By “values of the modulus ” here and in what follows we mean “values of the VEV of the modulus ”. , which allow to reproduce the quoted numerical values of the CPV phases and within deviation by scanning and in the ranges of (blue points), (red points) and (magenta points). These ranges are chosen in order to illustrate the and dependences. As is indicated by the figure, there is a viable region of values of up to . We find a region near the fixed points that is also viable. Indeed, in subsection 4.1 we easily obtain an example of a parameter set near the fixed points which provides a good description of the quark data. There is scarcely a viable point close to .
In Fig. 2 we show the corresponding region in the plane for values of . The imaginary part of can have relatively large values in the region of , the maximal value being reached for . We do not show the analogous plot in the plane since it is almost the same as the one in Fig. 2.
We will present in subsection 4.1 the results of the fits of the the quark masses, CKM elements and the Jarlskog rephasing invariant quoted in Eqs. (73), (75) and (78) of Appendix B.
![]() ![]() |
Let us compare the allowed region obtained by us in the modular model with quark mass matrices given in Eq. (14) with that in the model proposed in Ref. [29], in which modular forms of level modular symmetry are used. These modular forms are SL singlets - the Eisenstein series with weight , and , , and , . They coincide with the singlet modular forms of the modular group of the same weights: , , , , where and , , , and are given in Appendix A. The parameters and appear in the quark mass matrices of the model in the same way they appear in the Yukawa matrices in Eq. (8):
(18) |
The results of this analysis performed by us are shown in Fig. 4 and Fig. 4. The distributions of and are almost the same as those shown in Figs. 2 and 2 because the textures of and are the same in the two models. We will see later that the distributions of and depend significantly on the textures of the quark mass matrices.
![]() ![]() |
3.2 Quark mass matrix of model (2) and its CPV phase structure
We present the second model (2), in which the assignments of the weights for the relevant chiral superfields are as follows:
-
•
quark doublet (left-handed) : singlets with weight .
-
•
quark singlets (righ-handed) and : singlets with weight .
-
•
Higgs fields of down-type and up-type quark sectors : singlet with weight 0.
These assignments satisfy the weight condition of Eq. (5). Those are summarized in Table 2.
2 | 1 | 2 | 2 | |
0 | 0 |
The mass matrices of the down-type and up-type quarks read:
(19) |
The CP violating phases come from the modulus in modular forms of , , and . The phases of and in the and elements can be factorised as follows:
(20) |
The phase matrices and are given by:
(21) |
where and are defined in Eq.(16). As in the model (1), the matrix does not contribute to the CKM mixing matrix, while , being common to down-type quark and up-type quark mass matrices, is cancelled out in the CKM matrix. The CP violation is generated by the phases of the elements of and in Eq. (20),
(22) |
which are determined by the VEV of the modulus and the real parameters and .
The analysis which follows is analogous to that performed for the model (1) in the preceding section. The second (middle) matrix of Yukawa couplings in Eq. (3) generates the quark mass matrices reported in Eq. (86) of Appendix C.2, which have the same structure as the the mass matrices in Eq. (20), obtained by employing the modular symmetry. It is shown in C.2 that these mass matrices successfully describe the data on the quark masses, mixing and CP violation if their elements have the numerical values given in Eq. (88) and the CPV phases in the (3,3) elements in and posses the values and , respectively. Reproducing the real numerical values of the elements of and reported in Eq. (88) using the constants and does not pose a problems. Thus, if we find values of the VEV of and of the real constants and that generate CPV phases and , then the modular matrices and in Eq. (20) will be phenomenologically viable for the so found values of the modulus VEV and and .
We have performed the relevant analysis and show in Fig. 6 the region of values of the VEV of for which the phases and are reproduced within deviation by scanning and in the ranges of (blue points), (red points) and (magenta points). The allowed points of are considerably reduced as compared with those in Fig. 2. However, there is still a viable region of values of close to . We also find a region near the fixed points that is also viable. Indeed, in subsection 4.2 we present an example of a parameter set near the fixed points as well as the ones close to and at the large .
In Fig. 6 we show the corresponding region in the plane for values of . The imaginary part of reaches the value of for . The analogous plot of plane is almost the same one as the one in Fig. 6 and we do not show it here.
We will present in subsection 4.2 the results of the fits of the the quark masses, CKM elements and the Jarlskog rephasing invariant quoted in Eqs. (73), (75) and (78) of Appendix B.
![]() ![]() |
3.3 Quark mass matrix of model (3) and its CPV phase structure
For the model (3), the assignments of the weights for the relevant chiral superfields is given as follows:
-
•
quark doublet (left-handed) : singlets with weight .
-
•
quark singlets (righ-handed) and : singlets with weight .
-
•
Higgs fields of down-type and up-type quark sectors : singlet with weight 0.
These assignments satisfy the weight condition of Eq.(5). Those are summarized in Table 3.
2 | 1 | 2 | 2 | |
0 | 0 |
The mass matrices of the down-type and up-type quarks have the form:
(23) |
Since the number of parameters are 14, four parameters are redundant for reproducing the quark masses and CKM elements (10 observables). In this case, following Ref.[90], we set and we investigate whether one can describe successfully the quark data with this additional phenomenological assumption. As a consequence of setting the constants , and to zero the up-type quark mass matrix is diagonal.
The CP violating phases originate from the modulus and appear in the down-type quark mass matrix via the modular forms , , and . The phases of and in the and elements can be factored out as follows:
(24) |
The phase matrices and are
(25) |
where and are given in Eq. (16).
On the other hand, the up-type quark mass matrix is diagonal. Therefore, and do not contribute to the CKM matrix. Thus, the CP violation is generated by the phase of the element of in Eq. (23),
(26) |
which is determined by the VEV of and the real parameter .
As shown in Eq. (93) of Appendix C.3, the quark mass matrices in Eq. (23) could be completely consistent with observed masses and the CKM matrix if the phase of Eq. (26) has the value . By adjusting the value of the VEV of the modulus and of the real constant to get , we will obtain a phenomenologically viable down-quark mass matrix . The real constants can be used to reproduce the real numerical values in in Eq. (95).
We plot in Fig. 8 the values of the VEV of for which one can generate within deviation by scanning and again in the ranges of (blue points), (red points) and (magenta points). In the considered case it is possible to reproduce the observed CP violation in the quark sector even for close to and “close” to . Indeed, we present examples of viable parameter sets for and in subsection 4.3.
In Fig. 8, we show the values of leading (within deviation) to the requisite value of versus . The magnitude of increases rapidly when decreases towards .
![]() ![]() |
Let us emphasise that in the case of the discussed model (3) we have set the redundant parameters “by hand” to get diagonal mass matrix for the up-type quarks. This set-up is not guaranteed within the framework of the modular invariance approach without additional (symmetry) assumptions.
4 Reproducing quark masses and CKM parameters
We present next numerical examples of successfully reproducing the observed quark masses, the CKM mixing angles and CPV phase in the cases of the three models considered by us.
4.1 Fitting model (1)
We present three examples of model (1) corresponding to the being relatively close to the fixed points , and , respectively.
4.1.1 close to
The first one is an example of model (1), where is rather close to . This case is comparable to the example in Ref.[29]. We show the numerical values of parameters obtained in the fit of the quark masses, CKM mixing angles, the CPV phase and of the Jarlskog rephasing invariant [6], as defined in [7] and quoted in Eqs. (73), (75) and (78) of Appendix B:
(27) |
where denotes a measure of goodness of the fit. By employing the sum of one-dimensional for eight observable quantities , it is defined as . The result of the fit of the quark mass ratios, three CKM elements , , , the phase and of the factor are collected in Table 4.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
We comment on the values of parameters in Eq. (27). The parameters , , () are real constants that do not couple to modular forms. On the other hand, , , () are constants multiplying the modular forms. Since the normalizations of each of the three modular forms present in the expressions for the quark mass matrices and are arbitrary, the magnitudes of the parameters , , and respectively those of , , should be compared separately among themselves. The constants multiplying the modular forms are of the same order in magnitude:
(28) |
In what concerns the constants, , , , only the ration is somewhat smaller than the other ratios , and :
(29) |
4.1.2 close to
We present the second example of successful fit of the quark data, in which the VEV of the modulus is close to the fixed point (the left cusp). The numerical values of parameters read:
(30) |
In Table 5 we present the results of the fit of the quark mass ratios, the three CKM elements , , , of the CPV phase and of the factor.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
The ratios of the constants multiplying the modular forms are of the same order of magnitude:
(31) |
In what concerns the other constants, only the ratio of is again somewhat smaller than the other ratios:
(32) |
4.1.3 “close” to
We present the third example of successful fit of quark data, in which the VEV of the modulus has a relatively large imaginary part “close” to the fixed point . The numerical values of parameters read:
(33) |
In Table 6 we present the results of the fit of the quark mass ratios, the three CKM elements , , , of the CPV phase and of the factor.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
The constants multiplying the modular forms present in and in are of the same order of magnitude except for , which is somewhat smaller than :
(34) |
In what concerns the other constants in and in , the ratio of is somewhat smaller than , the other ratios and being of the same order of magnitude:
(35) |
4.2 Fitting model (2)
We present three examples of model (2) corresponding to the being relatively close to the fixed points , and , respectively.
4.2.1 close to
The first example is close to . For the numerical values of parameters resulting from the fit, we obtain in this case:
(36) |
The results of the fit of the observables are shown in Table 7.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
The constants multiplying the modular forms are of the same order of magnitude except for , which is significantly smaller:
(37) |
The other constants are hierarchical:
(38) |
4.2.2 close to
We present the second example, in which the VEV of the modulus is close to the fixed point . The numerical values of parameters obtained in the fit are:
(39) |
The results of the fit of the observables are shown in Table 8.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
The constants multiplying the modular forms are of the same order of magnitude except for , which is significantly smaller also in this case:
(40) |
The other constants are hierarchical:
(41) |
4.2.3 ”close” to
We present the third example at the VEV of having a relatively large imaginary part “close” to . The numerical values of parameters obtained in the fit are:
(42) |
The results of the fit of the observables are shown in Table 9.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
The constants multiplying the modular forms are of the same order of magnitude for the down-type quarks, but hierarchical for up-type quarks:
(43) |
The other constants are hierarchical:
(44) |
4.3 Fitting model (3)
We present three examples of model (3) corresponding to the being relatively close to the fixed points , and , respectively.
4.3.1 close to
The first one is an example of model (3), where is rather close to . For the numerical values of parameters resulting from the fit, we obtain in this case:
(45) |
Both the constants multiplying the modular forms present in and the other constants in and have very different values with the their relevant ratios exhibiting strong hierarchies. The results of the fit of the observables are shown in Table 10.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
4.3.2 close to
We present the second example, in which the VEV of the modulus is close to the fixed point . The numerical values of parameters obtained in the fit are:
(46) |
As in the previous examples, there is a strong hierarchy among both the constants multiplying the modular forms present in and the other constants in and .
The results of the fit of the observables are shown in Table 11.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
4.3.3 ”close” to
Finally, we present an example at the VEV of having a relatively large imaginary part “close” to . The numerical values of parameters obtained in the fit are:
(47) |
In this example, there is also a strong hierarchy among both the constants multiplying the modular forms present in and the other constants in and .
The results of the fit of the observables are shown in Table 12.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
5 A texture realizing modulus stabilisation
5.1 Modulus stabilisation
We have performed statistical analyses of the three models and have shown that they are phenomenologically viable. We have shown that model (1) (Eq. (12)), model (2) (Eq. (19) and model (3) (Eq. (23) and the related discussion) can describe well the quark data for values of close to , close to (the left cusp) and “close” to . The values of the VEV of the modulus , , and , as is well known, are the only fixed points of the modular group in its fundamental domain. Values of close to three fixed points have been found in studies of the modulus stabilisation (see, e.g., [70, 71, 72, 73, 74, 75, 76, 77, 78]), in which is obtained by analysing the absolute minima of relatively simple (supergravity-motivated) modular- and CP-invariant potentials for the modulus . In this section, we focus on the solution, which give the absolute minima of the supergravity-motivated modular- and CP-invariant potentials for the modulus , obtained in Ref.[73].
, | |
, | |
, |
The values of are very close to the fixed point for the non-negative integer , which denote the power indices in the modular-invariant function, as seen in Table 13.
In the next subsection, we present a model, which is consistent with in Table 13.
5.2 An alternative texture zero
In the standpoint of ”Occam’s Razor approach” of the quark mass matrix (minimum number of parameters) [90], we consider the following texture zeros for mass matrices of the down-type quarks with the diagonal up-type quark mass matrix.
(48) |
where and are real, and is the CP phase. For down-type quark sector, we obtain those numerical values by inputting observed ones in Appendix B as follows:
(49) |
Fit | * | * | |||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
It is emphasized that the numerical values of the parameter set are unique in these mass matrices because the number of parameters is ten while input data are also ten.
5.3 Variant of model (1)
We consider the variant model of the model (1) in Table 1. We show the assignment of the representations and weights for the relevant quarks in Table 15, where the assignments of are changed:
2 | 1 1 | 2 | 2 | |
0 | 0 |
Since the representations and the weights of the right-handed up-type quarks are different from the down-type ones, the mass matrices are different each other. We have real and . The mass matrices of the down- and up-type quarks are given as:
(50) |
In order to reproduce the texture zeros in (48), we choose or by hand.
5.4 Numerical result of the variant model
In the quark mass matrices of Eq.(50), the allowed region is obtained by fitting the phase of the texture in (48) with error bar as discussed in section 3. We show the full allowed region in Fig.10 and the restricted region close to the fixed point in Fig.10.
![]() ![]() |
It is remarked that the three points in Table 13 are almost on the line, which is very narrow allowed region. These three points are never inside the allowed region in the previous models as seen in Figs.2, 4, 6, 8.
Indeed, we obtain a successful example fitting quark masses and CKM parameters by fixing , which corresponds to in Table 13 as follows:
(51) |
where the up-type quark mass matrix is diagonal. The output is given in Table 16.
Fit | * | * | |||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
6 Summary and discussions
We have discussed the strong CP problem within the modular invariance approach to flavour. Working with () modular symmetry we have constructed simple models which provide solution to the strong CP problem without the need for an axion. In these models it is assumed that CP is a fundamental symmetry of the Lagrangian. As a consequence, the strong CPV phase . The CP symmetry is broken spontaneously by the VEV of the modulus (), so the large CPV phases in the CKM matrix is generated and at the same time remains zero or gets a tiny value compatible with the existing stringent experimental limit .
To be more specific, we have considered three models, i.e., three types of down-type and up-type mass matrices and , which have, as a consequence of the modular symmetry, three zero elements, or three texture zeros, each (Eqs. (12), (19) and (23)). The position of the zeros and the requirement of CP invariance ensures that and are real quantities. The quark mass matrices and contain three modular forms of weights 12 and 16 which are singlets, , and , and one modular form of weight 4 which is singlet, . The presence of two modular forms of the same weight furnishing the same singlet representation of is required in order to describe correctly the observed CP violation in the quark sector when the CP symmetry is broken spontaneously by . In the case of model (3) we have considered phenomenologically the presence of three additional zero elements in to reduce the number of redundant constant parameters in the model. This set-up, in which is a diagonal matrix, is not guaranteed by modular invariance without additional (symmetry) assumptions.
Our work is an extension of the pioneering work of Ref.[29] where one pattern of texture zeros is considered in the framework of the level modular symmetry. We have presented phenomenologically viable models with finite modular symmetry of level () by using three patterns of texture zeros. We have studied the distributions of the VEV of the modulus allowing to reproduce the observed quark masses CKM mixing angles and CP violation. And we have found examples of the models which are viable for values of close to the fixed points of the modular group.
In particular, we focus on the VEV of , which gives the absolute minima of the supergravity-motivated modular- and CP-invariant potentials for the modulus , so called, modulus stabilisation [73]. We present a successful model, which is consistent with this results of the modulus stabilisation.
Our work together with Ref.[29] promotes the modular invariance as a successful approach and framework providing solutions not only to the quark and lepton flavour problems but also to the strong CP problem.
Acknowledgments
The work of S. T. P. was supported in part by the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No. 860881-HIDDeN, by the Italian INFN program on Theoretical Astroparticle Physics and by the World Premier International Research Center Initiative (WPI Initiative, MEXT), Japan. The authors would like to thank Kavli IPMU, University of Tokyo, where part of this study was done, for the kind hospitality.
Appendix
Appendix A Modular forms of with higher weights
The lowest weight 2 triplet modular forms of are given as:
(52) |
where . They satisfy also the constraint [25]:
(53) |
For weight 4, five modular forms are given as:
where vanishes due to the constraint of Eq. (53).
For weigh 6, there are seven modular forms as:
For weigh 8, there are nine modular forms as:
For weigh 10, there are eleven modular forms as:
(54) |
For weigh 12, there are thirteen modular forms as:
For weigh 14, there are fifteen modular forms as:
(55) |
Four triplets are obtained by and .
For weigh 16, there are seventeen modular forms as:
(56) |
Four triplets are obtained by and .
The modular form is the holomorphic normalized Eisenstein series with weight , which is given
(57) |
where and are integers.
We show the values of singlets of modular form at the fixed points, , , in Table 17.
0 | ||||
0 | ||||
0 | 0 | |||
0 | ||||
{0, 0} | {0, 0} | |||
{0, 0} | ||||
0 | 0 | |||
Appendix B Input data of quark masses and CKM elements
The modulus breaks the modular invariance by obtaining a VEV at some high mass scale. We assume this to be the GUT scale. Correspondingly, the values of the quark masses and CKM parameters at the GUT scale play the role of the observables that have to be reproduced by the considered quark flavour models. They are obtained using the renormalisation group (RG) equations which describe the “running” of the observables of interest from the electroweak scale, where they are measured, to the GUT scale. In the analyses which follow we adopt the numerical values of the quark Yukawa couplings at the GUT scale GeV derived in the framework of the minimal SUSY breaking scenarios with [79]:
(73) |
The quark masses are given as with GeV. The choice of relatively small value of allows us to avoid relatively large -enhanced threshold corrections in the RG running of the Yukawa couplings. We set these corrections to zero.
The quark flavour mixing is given by the CKM matrix, which has three independent mixing angles and one CP violating phase. These mixing angles are given by the absolute values of the three CKM elements , and . We take the present data on the three CKM elements in Particle Data Group (PDG) edition of Review of Particle Physics [7] as:
(74) |
By using these values as input and we obtain the CKM mixing angles at the GUT scale of GeV [79]:
(75) |
The tree-level decays of are used as the standard candles of the CP violation. The latest world average of the CP violating phase is given in PDG2022 [7] as:
(76) |
Since the phase is almost independent of the evolution of RG equations, we refer to this value in the numerical discussions. The rephasing invariant CP violating measure [6] is also given in [7]:
(77) |
Taking into account the RG effects on the mixing angles for , we have at the GUT scale GeV:
(78) |
Appendix C Quark mass matrices with texture zeros
In this Appendix, we present three numerical examples of the quark mass matrix with the texture zeros, which are consistent with the observed masses and CKM elements given in Appendix B.
C.1 Texture (1)
Consider the quark mass matrices with the texture zeros of model (1):
(79) |
where are real constants, and are CP violating (CPV) phases. The determinants of the mass matrices and are real and are given in terms of the real parameters , and :
(80) |
If these mass matrices keep strictly their zero structures and describe correctly the observed CP violation, they are candidates for solving the strong CP problem.
Since the number of parameters in and are 14, four parameters are redundant from the point of view of reproducing the six quark masses and the four independent CKM mixing angles and CPV phase (10 observables ). We present the mass matrices in the case of . Using as input the observed quark masses and the CKM parameters and performing a statistical analysis we show a typical numerical example of the mass matrices which reproduce the quark data in Appendix B:
(81) |
where are approximately satisfied with only being somewhat smaller than the other five constants 888The magnitudes of the constants, including , and , , which in the modular model multiply the modular forms present in and , are discussed in section 4. . The matrices and with the numerical values of the real constants and of the CPV phases and as given in Eq. (81), provide a good quality of the fit of the quark mass ratios, the CKM mixing angles and the CPV phase ( C.L.), as seen in Table 18.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
C.2 Texture (2)
Consider next the quark mass matrices with the texture zeros of model (2):
(86) |
where again the constants are real, and are CPV phases. The determinants of the mass matrices and are given in terms of and real:
(87) |
If these mass matrices keep their zero elements strictly zero and describe the the observed CP violation, they are candidates for solving the strong CP problem.
Performing a statistical analysis we have found a typical numerical example of the mass matrices and which describe the quark data in Appendix B:
(88) |
The quark mass ratios and the CKM matrix obtained by diagonalising the matrices and given in Eq. (88) are completely consistent with observed one including the CP violating phase ( C.L.), as seen in Table 19.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
C.3 Texture (3)
Let us analyse finally the quark mass matrices with the texture zeros of model (3):
(93) |
where as in the previous two cases the constants , , , , , are real, and are CPV phases. Also in this case the determinants of the quark mass matrices and are given in terms of real parameters, namely of , and are real:
(94) |
As in the previous two cases we note that if the zero elements of these mass matrices remain strictly zero, and the mass matrices describe correctly the observed quark masses, and especially the quark mixing angles and CP violation in the quark sector, they are a candidate for solving the strong CP problem.
We will present next a typical numerical example of the mass matrices which describe the quark data. Since the number of parameters is 14, four parameters are redundant for reproducing the quark masses and CKM elements (10 observables). In this case, following Ref.[90], we set (also ). Performing a statistical analysis we get a good description of the quark data in Appendix B with:
(95) |
The quark mass ratios, and especially the CKM mixing angles, the factor and the CPV phase obtained using the numerical matrices given in Eq. (95) are completely consistent with observed one ( C.L.), as seen in Table 20.
Fit | |||||||||
---|---|---|---|---|---|---|---|---|---|
Exp | |||||||||
References
- [1] A. A. Belavin, A. M. Polyakov, A. S. Schwartz and Y. S. Tyupkin, Phys. Lett. B 59 (1975) 85-87.
- [2] R. Jackiw and C. Rebbi, Phys. Rev. Lett. 37 (1976) 172-175.
- [3] C. G. Callan, Jr., R. F. Dashen and D. J. Gross, Phys. Lett. B 63 (1976) 334-340.
- [4] G. ’t Hooft, Phys. Rev. Lett. 37 (1976) 8-11.
- [5] C. Abel, S. Afach, N. J. Ayres, C. A. Baker, G. Ban, G. Bison, K. Bodek, V. Bondar, M. Burghoff and E. Chanel, et al. Phys. Rev. Lett. 124 (2020) no.8, 081803 [arXiv:2001.11966 [hep-ex]].
- [6] C. Jarlskog, Phys. Rev. Lett. 55 (1985) 1039.
- [7] R. L. Workman et al. [Particle Data Group], PTEP 2022 (2022) 083C01.
- [8] R. D. Peccei and H. R. Quinn, Phys. Rev. Lett. 38 (1977) 1440-1443.
- [9] L. Di Luzio, M. Giannotti, E. Nardi and L. Visinelli, Phys. Rept. 870 (2020) 1-117 [arXiv:2003.01100 [hep-ph]].
- [10] A. E. Nelson, Phys. Lett. B 136 (1984) 387-391.
- [11] S. M. Barr, Phys. Rev. Lett. 53 (1984) 329.
- [12] M. Dine and P. Draper, JHEP 08 (2015) 132 [arXiv:1506.05433 [hep-ph]].
- [13] G. Hiller and M. Schmaltz, Phys. Lett. B 514 (2001) 263-268 [arXiv:hep-ph/0105254 [hep-ph]].
- [14] K. S. Babu and R. N. Mohapatra, Phys. Rev. D 41 (1990) 1286.
- [15] L. Bento, G. C. Branco and P. A. Parada, Phys. Lett. B 267 (1991) 95-99.
- [16] R. Kuchimanchi, Phys. Rev. Lett. 76 (1996) 3486-3489. [arXiv:hep-ph/9511376 [hep-ph]].
- [17] S. M. Barr, D. Chang and G. Senjanovic, Phys. Rev. Lett. 67 (1991) 2765-2768.
- [18] Q. Bonnefoy, L. Hall, C. A. Manzari and C. Scherb, Phys. Rev. Lett. 131 (2023) no.22, 221802 [arXiv:2303.06156 [hep-ph]].
- [19] S. Antusch, M. Holthausen, M. A. Schmidt and M. Spinrath, Nucl. Phys. B 877 (2013) 752-771 [arXiv:1307.0710 [hep-ph]].
- [20] R. Harnik, G. Perez, M. D. Schwartz and Y. Shirman, JHEP 03 (2005) 068 [arXiv:hep-ph/0411132 [hep-ph]].
- [21] C. Cheung, A. L. Fitzpatrick and L. Randall, JHEP 01 (2008) 069 [arXiv:0711.4421 [hep-th]].
- [22] L. Vecchi, JHEP 04 (2017) 149 [arXiv:1412.3805 [hep-ph]].
- [23] Z. G. Berezhiani, R. N. Mohapatra and G. Senjanovic, Phys. Rev. D 47 (1993) 5565-5570 [arXiv:hep-ph/9212318 [hep-ph]].
- [24] A. Valenti and L. Vecchi, JHEP 07 (2021) no.152, 152 [arXiv:2106.09108 [hep-ph]].
- [25] F. Feruglio, in From My Vast Repertoire …: Guido Altarelli’s Legacy, A. Levy, S. Forte, Stefano, and G. Ridolfi, eds., pp.227–266, 2019, arXiv:1706.08749 [hep-ph].
- [26] T. Kobayashi, K. Tanaka and T. H. Tatsuishi, Phys. Rev. D 98 (2018) 016004 [arXiv:1803.10391].
- [27] J. T. Penedo and S. T. Petcov, Nucl. Phys. B939 (2019) 292 [arXiv:1806.11040].
- [28] P. P. Novichkov, J. T. Penedo, S. T. Petcov and A. V. Titov, JHEP 04 (2019) 174 [arXiv:1812.02158 [hep-ph]].
- [29] F. Feruglio, A. Strumia and A. Titov, JHEP 07 (2023) 027 [arXiv:2305.08908 [hep-ph]].
- [30] J. C. Criado and F. Feruglio, SciPost Phys. 5 (2018) 042 [arXiv:1807.01125 [hep-ph]].
- [31] T. Kobayashi, N. Omoto, Y. Shimizu, K. Takagi, M. Tanimoto and T. H. Tatsuishi, JHEP 1811 (2018) 196 [arXiv:1808.03012 [hep-ph]].
- [32] F. J. de Anda, S. F. King and E. Perdomo, Phys. Rev. D 101 (2020) no.1, 015028 [arXiv:1812.05620 [hep-ph]].
- [33] P. P. Novichkov, S. T. Petcov and M. Tanimoto, Phys. Lett. B 793 (2019) 247 [arXiv:1812.11289 [hep-ph]].
- [34] H. Okada and M. Tanimoto, Phys. Lett. B 791 (2019) 54-61 [arXiv:1812.09677 [hep-ph]].
- [35] G. J. Ding, S. F. King and X. G. Liu, JHEP 1909 (2019) 074 [arXiv:1907.11714 [hep-ph]].
- [36] H. Okada and M. Tanimoto, Eur. Phys. J. C 81 (2021) 52 [arXiv:1905.13421 [hep-ph]].
- [37] H. Okada and M. Tanimoto, Phys. Dark Univ. 40 (2023) 101204 [arXiv:2005.00775 [hep-ph]].
- [38] H. Okada and M. Tanimoto, Phys. Rev. D 103 (2021) 015005 [arXiv:2009.14242 [hep-ph]].
- [39] H. Okada and M. Tanimoto, JHEP 03 (2021) 010 [arXiv:2012.01688 [hep-ph]].
- [40] H. Okada, Y. Shimizu, M. Tanimoto and T. Yoshida, JHEP 07 (2021) 184 [arXiv:2105.14292 [hep-ph]].
- [41] T. Kobayashi, H. Otsuka, M. Tanimoto and K. Yamamoto, Phys. Rev. D 105 (2022) 055022 [arXiv:2112.00493 [hep-ph]].
- [42] T. Kobayashi, H. Otsuka, M. Tanimoto and K. Yamamoto, JHEP 08 (2022) 013 [arXiv:2204.12325 [hep-ph]].
- [43] S. T. Petcov and M. Tanimoto, Eur. Phys. J. C 83 (2023) 579 [arXiv:2212.13336 [hep-ph]].
- [44] S. T. Petcov and M. Tanimoto, JHEP 08 (2023) 086 [arXiv:2306.05730 [hep-ph]].
- [45] P. P. Novichkov, J. T. Penedo, S. T. Petcov and A. V. Titov, JHEP 04 (2019) 005 [arXiv:1811.04933 [hep-ph]].
- [46] X. G. Liu and G. J. Ding, JHEP 08 (2019) 134 [arXiv:1907.01488 [hep-ph]].
- [47] P. P. Novichkov, J. T. Penedo and S. T. Petcov, Nucl. Phys. B 963 (2021) 115301 [arXiv:2006.03058 [hep-ph]].
- [48] C. Y. Yao, X. G. Liu and G. J. Ding, Phys. Rev. D 103 (2021) 095013 [arXiv:2011.03501 [hep-ph]].
- [49] G. J. Ding, F. Feruglio and X. G. Liu, SciPost Phys. 10 (2021) 133 [arXiv:2102.06716 [hep-ph]].
- [50] P. P. Novichkov, J. T. Penedo and S. T. Petcov, JHEP 04 (2021) 206 [arXiv:2102.07488 [hep-ph]].
- [51] G. J. Ding, S. F. King and C. Y. Yao, Phys. Rev. D 104 (2021) 055034 [arXiv:2103.16311 [hep-ph]].
- [52] C. C. Li, X. G. Liu and G. J. Ding, JHEP 10 (2021) 238 [arXiv:2108.02181 [hep-ph]].
- [53] I. de Medeiros Varzielas, M. Levy, J. T. Penedo and S. T. Petcov, JHEP 09 (2023) 196 [arXiv:2307.14410 [hep-ph]].
- [54] G. Altarelli and F. Feruglio, Rev. Mod. Phys. 82 (2010) 2701 [arXiv:1002.0211 [hep-ph]].
- [55] H. Ishimori, T. Kobayashi, H. Ohki, Y. Shimizu, H. Okada and M. Tanimoto, Prog. Theor. Phys. Suppl. 183 (2010) 1 [arXiv:1003.3552 [hep-th]].
- [56] H. Ishimori, T. Kobayashi, H. Ohki, H. Okada, Y. Shimizu and M. Tanimoto, Lect. Notes Phys. 858 (2012) 1, Springer.
- [57] T. Kobayashi, H. Ohki, H. Okada, Y. Shimizu and M. Tanimoto, Lect. Notes Phys. 995 (2022) 1, Springer.
- [58] D. Hernandez and A. Y. Smirnov, Phys. Rev. D 86 (2012) 053014 [arXiv:1204.0445 [hep-ph]].
- [59] S. F. King and C. Luhn, Rept. Prog. Phys. 76 (2013) 056201 [arXiv:1301.1340 [hep-ph]].
- [60] S. F. King, A. Merle, S. Morisi, Y. Shimizu and M. Tanimoto, New J. Phys. 16 (2014) 045018 [arXiv:1402.4271 [hep-ph]].
- [61] M. Tanimoto, AIP Conf. Proc. 1666 (2015) 120002.
- [62] S. F. King, Prog. Part. Nucl. Phys. 94 (2017) 217 [arXiv:1701.04413 [hep-ph]].
- [63] S. T. Petcov, Eur. Phys. J. C 78 (2018) 709 [arXiv:1711.10806 [hep-ph]].
- [64] F. Feruglio and A. Romanino, Rev. Mod. Phys. 93 (2021) no.1, 015007 [arXiv:1912.06028 [hep-ph]].
- [65] T. Kobayashi and M. Tanimoto, [arXiv:2307.03384 [hep-ph]].
- [66] J. R. Ellis and M. K. Gaillard, Nucl. Phys. B 150 (1979) 141.
- [67] I. B. Khriplovich, Phys. Lett. B 173 (1986) 193.
- [68] D. Zhang, Nucl. Phys. B 952 (2020) 114935 [arXiv:1910.07869 [hep-ph]].
- [69] P. P. Novichkov, J. T. Penedo, S. T. Petcov and A. V. Titov, JHEP 07 (2019) 165 [arXiv:1905.11970].
- [70] T. Kobayashi, Y. Shimizu, K. Takagi, M. Tanimoto and T. H. Tatsuishi, Phys. Rev. D 100 (2019) 115045 [erratum: Phys. Rev. D 101 (2020) 039904] [arXiv:1909.05139 [hep-ph]].
- [71] H. Abe, T. Kobayashi, S. Uemura and J. Yamamoto, Phys. Rev. D 102 (2020) 045005 [arXiv:2003.03512 [hep-th]].
- [72] K. Ishiguro, T. Kobayashi and H. Otsuka, JHEP 03 (2021) 161 [arXiv:2011.09154 [hep-ph]].
- [73] P. P. Novichkov, J. T. Penedo and S. T. Petcov, JHEP 03 (2022) 149 [arXiv:2201.02020 [hep-ph]].
- [74] K. Ishiguro, H. Okada and H. Otsuka, JHEP 09 (2022) 072 [arXiv:2206.04313 [hep-ph]].
- [75] V. Knapp-Perez, X. G. Liu, H. P. Nilles, S. Ramos-Sanchez and M. Ratz, Phys. Lett. B 844 (2023) 138106 [arXiv:2304.14437 [hep-th]].
- [76] S. F. King and X. Wang, [arXiv:2310.10369 [hep-ph]].
- [77] T. Kobayashi, K. Nasu, R. Sakuma and Y. Yamada, Phys. Rev. D 108 (2023) 115038 [arXiv:2310.15604 [hep-ph]].
- [78] T. Higaki, J. Kawamura and T. Kobayashi, [arXiv:2402.02071 [hep-ph]].
- [79] S. Antusch and V. Maurer, JHEP 1311 (2013) 115 [arXiv:1306.6879 [hep-ph]].
- [80] S. Weinberg, Trans. New York Acad. Sci. 38 (1977) 185.
- [81] R. Gatto, G. Sartori and M. Tonin, Phys. Lett. B 28 (1968) 128.
- [82] H. Fritzsch, Phys. Lett. B 73 (1978) 317.
- [83] H. Fritzsch, Nucl. Phys. B 155 (1979) 189.
- [84] P. Ramond, R. G. Roberts and G. G. Ross, Nucl. Phys. B 406 (1993) 19 [hep-ph/9303320].
- [85] Z. z. Xing and Z. h. Zhao, Nucl. Phys. B 897 (2015) 302 [arXiv:1501.06346 [hep-ph]].
- [86] H. Fritzsch and Z. z. Xing, Phys. Lett. B 555 (2003) 63-70 [arXiv:hep-ph/0212195 [hep-ph]].
- [87] G. C. Branco, L. Lavoura and F. Mota, Phys. Rev. D 39 (1989) 3443.
- [88] G. C. Branco and J. I. Silva-Marcos, Phys. Lett. B 331 (1994) 390.
- [89] G. C. Branco, D. Emmanuel-Costa and R. Gonzalez Felipe, Phys. Lett. B 477 (2000) 147 [arXiv:hep-ph/9911418 [hep-ph]].
- [90] M. Tanimoto and T. T. Yanagida, PTEP 2016 (2016) 043B03 [arXiv:1601.04459 [hep-ph]].
- [91] K. Harigaya, M. Ibe and T. T. Yanagida, Phys. Rev. D 86 (2012) 013002 [arXiv:1205.2198 [hep-ph]].
- [92] M.-C. Chen, S. Ramos-Sanchez and M. Ratz, Phys. Lett. B801 (2020) 135153 [arXiv:1909.06910].