Fermion Scattering in a Gravitational Background:

arXiv:1312.7657v1 [hep-ph] 30 Dec 2013

Electroweak Corrections and Flavour Transitions

Claudio Corian` oa , Luigi Delle Rosea , Emidio Gabriellib,c∗ and Luca Trentadued

(a) Dipartimento

di Matematica e Fisica ”Ennio De Giorgi”, Universit` a del Salento and † INFN-Lecce, Via Arnesano, 73100 Lecce, Italy

(b)

(c)

(d)

NICPB, R¨ avala 10, Tallinn 10143, Estonia INFN, Sezione di Trieste, Via Valerio 2, I-34127 Trieste, Italy

Dipartimento di Fisica e Scienze della Terra ”Macedonio Melloni”, Universit` a di Parma and INFN, Sezione di Milano Bicocca, Milano, Italy

Abstract We investigate the role of the electroweak corrections to the scattering cross section of Standard Model fermions with gravity. We use both an approach of scattering off an external potential, where the gravitational field is treated as a classical background generated by a heavy source, and the usual interaction based on the one-graviton-exchange. In the potential appoach we consider the fields both of a localized and of a distributed gravitational source of spherical symmetry and uniform density, separating the cases of interactions taking place both in the inner and external regions of the source. This allows to make a distinction between interactions involving neutrinos and dark matter particles with a realistic gravity source, which cover the inner region, and the rest of the Standard model fermions. The role of the gravitationally induced flavourchanging transitions, as well as the flavour diagonal ones, are investigated in the limit of both large and small momentum transfers, deriving the structure of the corresponding Hamiltonian.





On leave of absence from Dipartimento di Fisica Universit` a di Trieste, Strada Costiera 11, I-34151 Trieste [email protected], [email protected], [email protected], [email protected]

1

1

Introduction

Radiative corrections to the graviton/matter vertex have been investigated since the 70’ s by Berends and Gastmans [1] in the QED case, who quantified the impact of these corrections to the bending of photons and fermions in an external localized gravitational field. These results were based on the study of 3-point functions (gravitational form factors) with two fermions (T f f ) or two photons (T V V ) in external gravity, and one insertion of the energy momentum tensor (EMT) (T ) of the matter fields. The very small size of these corrections, which affect the angles of deflection, combined with the experimental difficulty to improve on their direct measurements, call into question the possibility of using these effects as a possible test of General Relativity. Perhaps, it is for this reason that the study of the electroweak corrections to the bending angles have not drawn any attention on these processes until the analysis of [2] and, more recently, of [3, 4], which concern the fermion/graviton vertex. In the photon case, the T V V vertex in QED has been re-investigated in the analysis of the conformal anomaly action in [5, 6], later extended to the electroweak case in [7, 8]. In the T f f case, the study has been separated into the flavour-changing [2, 4] and flavour diagonal sectors [3], the first of the two playing a significant role in the analysis of possible flavour transitions in the presence of a scalar component in the external gravitational field, as pointed out in [2]. This aspect will be re-addressed in our current study. The issue concerning the size of these corrections and their relevance can be resolved by noticing that the very same interactions become significant in the presence of sizeable gravitational backgrounds. Such are those due to black holes, which are ubiquitous in galactic centers, and surging to remarkable relevance in the analysis of galaxy structure formation. Obviously our formalism, which simply extends to the electroweak case the original analysis of [1], remains valid only in the case of small fluctuations around a flat spacetime metric, where curvature effects are ignored, and does not allow to handle backgrounds characterized by a strong curvature. It could, however, be extended to include further corrections related to the curvature scale of certain backgrounds, in the limit in which gravity is treated as an external classical source. For this reason, we think that our analysis allows to highlight some of the basic features of the interaction of the Standard Model Lagrangian, once this is immersed in a weak gravitational background, clarifying some of the specific features of this coupling at a perturbative level. In particular, the renormalizability of the theory, which remains intact in this extended framework, is an important feature of these computations, not shared by other backgrounds. One of the specific features of our results is their possible application to the neutrino sector, which has received some attention in recent years (see for instance [9, 10, 11]). We will provide indeed the expression for the QED and electroweak corrections to the scattering of a neutrino off an external source. These results are the starting point in the derivation of the equation relating the angle of deflection to the impact parameter of the collision, in the case of a static heavy source of the gravitational field. The use of an impact parameter formalism allows to merge the usual quantum corrections, computed in the exact electroweak theory, with the semiclassical method in which the incoming asymptotic quantum state is described by a particle approaching the source at a distance b, with b denoting the impact parameter. This method has been used in [1] and in the more recent literature on the computation of corrections to the gravitational scattering of quantum fields both 2

in static [12, 13] and in rotating (Lense-Thirring) backgrounds [14]. Except for the case of the QED corrections, already discussed in [1] and that we recompute, a general analysis of the weak corrections and of the modified lense equations, which may impact the structure of cosmic shears, is left to future work. The primary goal of our work is to extend our previous analysis of the T f f vertex, by computing the radiative corrections to the cross section in an external gravitational field, in the limit of a source of large mass, i.e. in a scattering off a potential. In this case we discuss both the limits of low and high momentum transfers. At low momentum transfers we derive an effective Hamiltonian which describes the flavour-changing transitions. We also discuss the cross section for the exchange of a scalar component for the gravitational field, with the inclusion of the electroweak effects. Our expressions for fermion scattering are specialized both to ordinary and to weakly interacting fermions, such as neutrinos or dark matter particles, by the inclusion of suitable form factors which account for interactions in the inner core of the source. We conclude our analysis with some perspectives and possible extensions of our work, to be left for further studies. Being this work the third in a sequel [3, 4], most of our previous notations and results will be necessary in order to proceed with the applications discussed here.

2

The Standard Model Lagrangian in a gravitational background: the fermion sector

In this section we briefly review the main features of the coupling of the Standard Model in an external gravitational field, focusing our attention on the fermion sector. We recall that the dynamics of the Standard Model plus gravity is described by the action S = SSM + SG + SI with SSM containing the Standard Model (SM) Lagrangian, while Z √ 1 SG = − 2 d4 x −g R κ Z √ SI = χ d4 x −g R H † H

(1)

(2)

denote respectively the Einstein gravitational term of the action and a term of improved action SI , involving the Higgs doublet H. The latter is responsible for generating a symmetric and traceless energy-momentum tensor (EMT) [15]. SSM is obtained by extending the ordinary Lagrangian of the Standard Model to a curved metric background. SI vanishes in the flat space time limit, due to the vanishing of the Ricci scalar in the same limit. Its corresponding EMT (TI µν ), however, is non-vanishing. With the inclusion of SI , varying the action S with respect to the fields of the SM allows to generate an EMT which is symmetric without any additional symmetrization. We remark that χ is conformal at the value χ ≡ χc = 1/6 and guarantees the renormalizability of the model, including all Green functions at any order in k containing only insertions of external graviton fields h, with κ2 = 16πG.

3

We will be using the flat metric ηµν = diag(1, −1, −1, −1), with an expansion of the form gµν = ηµν + κhµν + O(κ2 ) gµν = η µν − κhµν + O(κ2 ) √ κ −g = 1 + h + O(κ2 ), 2

(3)

where h ≡ hµν ηµν is the trace of the metric fluctuations. The interaction between the gravitational field and matter, at this order, is mediated by diagrams containing a single power of the EMT T µν and multiple fields of the Standard Model. The tree level coupling is summarized by the action Z κ Sint = − d4 x Tµν hµν , (4) 2

where Tµν denotes the symmetric and covariantly conserved EMT of the Standard Model Lagrangian, embedded in a curved space-time background and defined as 2 δ (SSM + SI ) Tµν = √ . (5) −g δgµν g=η

The complete EMT of the Standard Model, including ghost and gauge-fixing contributions can be found in [8]. The fermionic part of the EMT is obtained using the vielbein formalism. Indeed the fermions are coupled to gravity by using the spin connection Ω induced by the curved metric gµν . This allows the introduction of a derivative Dµ which is covariant both under gauge and diffeomorphism transformations. Generically, the Lagrangian for a fermion (f ) takes the form   √ i¯ µ i µ ¯ ¯ LF = −g (6) ψγ (Dµ ψ) − (Dµ ψ)γ ψ − m ψψ , 2 2 where the covariant derivative is defined as Dµ = ∂µ + Aµ + Ωµ , with Aµ denoting the gauge field. The spin connection takes the form Ωµ =

1 ab ν σ Va Vbν;µ 2

(7)

where V is the vielbein, the semicolon denotes the gravitationally covariant derivative and σ ab are the generators of the Lorentz group in the spinorial representation. The latin indices are Lorentz indices of a local free-falling frame. The connection can be expanded as   1 mn 1 ρ σ l ν Ωµ = σ (8) Vm (∂µ Vnν − ∂ν Vnµ ) + Vm Vn (∂σ Vlρ − ∂ρ Vlσ ) Vµ − (m ↔ n) . 4 2 with

κ m h + O(κ2 ). 2 µ At leading order the interacting Dirac Lagrangian (6) generates the terms Vµm = δµm +

LF

= L0 − 4

κ hµν T (0)µν 2

(9)

(10)

ψ¯ k1 hµν T ff′ k2 ψ (a)

(b)

Figure 1: The leading order (a) graviton/fermion/fermion vertex. Fig. (b) symbolizes the radiative corrections. with L0 denoting the free Dirac term i L0 = 2



←  → ¯ ¯ ¯ ψ ∂/ ψ − ψ ∂/ ψ − mψψ

(11)

and the EMT of a Dirac fermion of mass mf (0) Tµν =

 i ¯ ¯ µ ψ) + (µ ↔ ν) − ηµν L0 (ψγµ ∂ν ψ − ∂ν ψγ 4

(12)

that we couple to the external field hext µν . In momentum space the corresponding vertex takes the form (see Fig. 118)  i µ γ (p1 + p2 )ν + γ ν (p1 + p2 )µ − 2η µν (/p1 + /p2 − 2m) , (13) V (0)µν = 4 from which we have omitted an overall factor (−κ/2) coming from the interaction Lagrangian, which will be reinserted at the level of the transition amplitude. Only diagonal transition amplitudes in flavour space are induced at leading order, while the inclusion of the electroweak corrections allows flavour-changing transitions, which will play a key role in our analysis. We will denote with Tˆ µν the transition amplitude between the intial and the final state induced by the insertions of the EMT vertex, with initial and final wave functions included, while V µν will be denoting the corresponding vertex in momentum space, i.e. Tˆµν = u ¯(p2 )V µν u(p1 ),

(14)

having labeled the momenta of the incoming and outgoing fermion with p1 and p2 respectively. We will similarly use the notation ˆ≡u O ¯(p2 )Ou(p1 ) (15) to denote the matrix element of any operator O in momentum space. Using these notations, the same matrix element can be expressed in position space in terms of the operator T µν (x), inserted on the Dirac eigenstates given by incoming (ψi ) and outgoing (ψf ) plane waves hp2 |T µν (x)|p1 i = ψ¯f (p2 )V µν ψi (p1 )eiq·x , 5

(16)

where q = p1 − p2 is the 4-momentum transfer. We have introduced plane waves given by r m1 ψi (p1 ) = Ni u(p1 ), Ni = , u ¯(p1 )u(p1 ) = 1, E1 Vl

(17)

and similarly for ψf , while Vl denotes a finite volume normalization of the two scattering states. The E1 (E2 ) and m1 (m2 ) are the energy and corresponding mass of the incoming (outgoing) particle respectively. The scattering matrix element is then written as Z κ iSif = − d4 xhp2 |hµν (x)T µν (x)|p1 i, (18) 2 V where V denotes the relevant region of integration and with the gravitational field being external, which gives ¯ 2 )V µν ψ(p1 )eiq·x . hp2 |hµν (x)T µν (x)|p1 i = hµν (x)ψ(p

(19)

The scattering amplitude which can be immediately expressed in momentum space as κ ¯ 2 )V µν ψ(p1 ) iSf i = − hµν (q)ψ(p 2 κ = − hµν (q)Ni Nf Tˆµν 2

(20)

in terms of the gravitational fluctuations in momentum space hµν (q), defined as in (32). For a static external field the energies of the incoming/outgoing fermions are conserved (E1 = E2 ≡ E) both in the flavour-diagonal and non-diagonal cases, as one can immediately realize from (31). It is obvious that the interaction region depends both on the type of scattering particle and on the geometry of the source. We will consider the potential scattering of a Dirac fermion off an external static source, which acts as a perturbation ext and the fluctuations on the otherwise flat spacetime background. The source is characterized by an EMT Tµν are determined by solving the linearized equations of General Relativity.

3

Leading order cross section

We start analyzing the flavour-conserving transitions at leading order. In the rest of the paper, for simplicity, we will omit the flavour index from each spinor, together with the explicit flavour dependence on the masses of the internal flavours, which appear at 1-loop order. The linearized equations take the form   1 ext  hµν − ηµν h = −κTµν (21) 2 and can be rewritten as hµν = κSµν ,

Sµν = −



ext Tµν

 1 ext − ηµν T , 2

(22)

where we have denoted with T ext the EMT trace of the external source. The external field is obtained by convoluting the static source with the retarded propagator GR (x, y) =

1 δ(x0 − |~x − ~y| − y0 ) 4π |~x − ~y | 6

(23)

normalized as GR (x, y) = δ4 (x − y).

(24)

The solution of (22) takes the form hext µν (x)



Z

d4 yGR (x, y)Sµν (y),

(25)

Pµ Pν 3 δ (~x), P0

(26)

with the EMT of the external localized source ext Tµν =

which for a particle of mass M at rest at the origin with Pµ = (M, ~0) takes the form ext Tµν = M δµ0 δν0 δ3 (~x).

(27)

This gives

M¯ Sµν S¯µν ≡ ηµν − 2δµ0 δν0 (28) 2 and the field generated by a local (point-like, L) mass distribution has a typical 1/r (r ≡ |~x|) behaviour Sµν =

2GM ¯ Sµν . κ|~x|

hL µν (x) =

(29)

The fluctuations are normalized in such a way that hµν has mass dimension 1, as an ordinary bosonic field, with κ of mass dimension −1. The Fourier transform of hµν in momentum space is given by Z hµν (q0 , ~q) = d4 xeiq·x hµν (x) (30)

which for a static field reduces to the form hµν (q0 , ~q) = 2πδ(q0 )hµν (~q),

(31)

hµν (~q) ≡ h0 (~q)S¯µν

(32)

with

defined in terms of the scalar form factor which is related to the geometrical structure of the external source. In the case of a point-like source, we obtain Z 2GM ¯ ei~q·~x hµν (q0 , ~ q ) = 2πδ(q0 ) × Sµν d3 ~x κ |~x|   κM ¯ Sµν , (33) = 2πδ(q0 ) × 2~q2 7

which gives h0 (~ q) ≡



κM 2~q2



hµν (~q) ≡

,



κM 2~q2



S¯µν .

(34)

At the same time, the leading order interaction is redefined as (0)µν

V (0)µν ≡ iOV

,

(35)

with

 1 µ ν p γ + pν γ µ − 2η µν (p / − 2m) , 4 which will turn useful in the computation of the radiative corrections. Using the matrix element  κ (0)µν u(p1 )eiq·x × Ni Nf × i¯ u(p2 )OV ψ¯f (x)T (0)µν ψi (x) = − 2 (0)µν

OV

=

the leading order scattering amplitude can be rewritten as Z (0) d4 xhµν (x)ψ¯f (x)T (0)µν ψi (x) iSif =  κ   (0)µν = − u(p1 ) × 2πδ(q0 ). × Ni Nf × ihµν (~q)¯ u(p2 )OV 2

(36)

(37)

(38)

The averaged squared amplitude then takes the form (0) h|iSif |2 i

   κ 2 κM 2 1 2 = − (Ni Nf ) × (2πδ(q0 )T ) × × Y0 , 2 2~q2 2

(39)

where we have introduced a factor (1/2) for the fermion spin average. As usual, we have extracted, from the square of the delta function, the transition time T , using (2πδ(q0 ))2 = 2πδ(q0 )T . We have defined i h 1 (0)αβ ¯ ¯ (0)µν Sµν Sαβ ( p + m)O T r ( p + m)O / / V V 1 2 4m2      p/2 + m m p/1 + m m 2γ 0 − 2γ 0 − = E 2 Tr 2m E 2m E

Y0 =

= 8

p~1 4 (0) F (x, θ) m2

(40)

in the flavour-diagonal case, with F (0) (x, θ) given by F (0) (x, θ) = cos2

θ θ x x2 3 + + + x cos2 , 2 4 4 4 2

(41)

with x = m2 /p~1 2 and p~1 the 3-momentum of the incoming fermion. We have used the relations p~f ≡ |~ p1 | = |~ p2 | 2 2 2 2 2 in the elastic limit, while ~ q = 4p~1 sin θ/2 and q = −~q . To compute the differential cross section we define the transition probability from the initial state (i) into a set of final states (f ) of differential dnf given by dW =

iSif 2 dnf ji 8

(42)

normalized with respect to the incident flux ji = p~1 /(Ei V ) with the final state density, with a volume normalization V given by V V dnf = d3 ~pf = |p~2 |E2 dE2 dΩ, (43) 3 (2π) (2π)3 which allows to define the cross section as the differential transition rate per unit time (dσ ≡ dW /T ). Combining all the contributions together and integrating over the energy of the final state we obtain dσ dΩ



dσ |L dΩ GM sin2 θ2

=

!2

F (0) (x, θ).

(44)

The result in Eq. (44) is in agreement with the corresponding one evaluated in Ref. [16]. In the following sections we will refer to this cross section as to the localized one, denoted in (44) also as dσ/dΩ|L , being generated by a gravitational point-like source located at the origin. This will turn useful in order to make a distinction between the cross sections of localized and of extended sources. The computation of the angle of deflection for the fermion involves a simple semiclassical analysis, in which one introduces the impact parameter representation of the same cross section. Assuming that the incoming particle is moving along the z direction, with the source localized at the origin, and denoting with θ the azymuthal angle, we have the relation dσ b db = sin θ dθ dΩ

(45)

between the impact parameter b and the scattering angle θ, measured from the z−direction. This semiclassical approach allows to relate the quantum interaction between the particle and the source. We recall, if not obvious, that in the classical description the relation between b and the deflection angle θd is obtained from energy and angular momentum conservation. In this respect Eq. (45) takes a similar role. The relation between b and θ as a function of the incoming energy (b = b(E, θ)), at least for small deflection angles, which correspond to large impact parameters, can be found analytically, but it should be integrated numerically otherwise. In the case of the point-like cross section (L), for instance, one obtains the differential relation !2 db2 GM F (0) (x, θ) sin θ (46) = −2 2 θ dθ sin 2 which gives b2 (θ) = (GM )2

!  (2 + x)2 θ . + 2(4 + 3x) log sin( ) 2 sin2 ( θ2 )

In the small θ limit we get the relations   x 4 2x + + (1 + )θ log θ + O(x2 θ log θ) + O(θ), b ∼ GM θ θ 4

(47)

(48)

which allows us to identify the deflection angle as θ ≡ θd ∼ 4 9

GM . b

(49)

Notice that in general x ≪ 1 for a scattering at high energy, being the fermion mass small compared to the incoming energy or three-momentum. Onbviously, the bending angle of neutrinos, charged leptons and dark matter particles will be affected by the geometry of the source and by the strength of the fermion interaction with the source. This will vary if, for instance, the particles are allowed to enter the core of the gravitational source, as in the neutrino and the dark matter cases. For this reason we will proceed by determining the structure of these corrections, providing more general cross sections beyond the point-like approximation.

3.1

Modified leading order cross sections

Once we allow a finite size for the mass distribution, expression (44) gets modified. This is the case, for instance, for beams of neutrinos or dark matter particles which may have, in principle, impact parameters smaller than the source radius. In this case, it is intuitively clear that in the limit of a weak gravitational background, Gauss’ theorem in gravity implies that only the region of the source of radius smaller than the particle impact parameter is relevant in the scattering process. This can be shown rigorously directly from the spherically symmetric Schwarzschild metric. However, the proof - in the weak field limit - of factorization of the 3-dimensional euclidean metric from the Schwarzschild solution, requires some technical steps which are left to an appendix (appendix A). This allows to avoid the use of polar projectors in the scattering amplitude - for spherically symmetric backgrounds - which would render the treatment more involved from the computational side. Our discussion, here, is limited to the case of a uniform mass distribution, but it can be generalized to any distribution of spherical symmetry, in the weak field limit. Using this result, it follows that the metric fluctuations of a distributed source can be described (in c = 1 units) by the field hµν = −

2Φd ¯ Sµν κ

(50)

with the potential of the distributed source Φd , given by Φd (r) = −

GM GM r2 θ(r − R) − (3 − )θ(R − r). r′ 2R R2

(51)

Notice that in the expression above we have separated the contribution of the external core region from the internal one, with R being the radius of the source. The modifications to the expression of the cross section in the various cases, due to the different nature of the scattering particles, can be taken into account by the insertion of the appropriate form of the fluctuation tensor in momentum space, obtained from this generalized geometrical setting. • Non-weakly interacting fermions For this reason, as in the case of the localized gravitational source, for non-weakly interacting fermions we restrict the integration region of the scattering amplitude in the form iSif

κ =− 2

Z

dx0

Z

>R

d3 ~xhp2 |hµν (x)T µν (x)|p1 i,

10

(52)

where we have restricted the integration region to |~x| > R, i.e. outside the radius of the gravitational source. Therefore, according to (51), it is convenient to denote the contributions to the gravitational field coming from the external region as hEµν . At leading order, for interactions which involve only the region of the external core we have κ iSf i = − hEµν (q0 , ~q)Ni Nf Tˆ µν 2

(53)

with hEµν (q0 , ~q) = 2πδ(q0 )hE (~q)S¯µν ,   Z 2Φd 3 ei~q·~x d ~x − hE (~q) = κ >R κM = cos(|~q|R) 2~q2

(54)

obtained as Fourier transform of the potential in the restricted region. It is then clear that, if we allow only interactions out of the core region, as we should for non-weakly interacting fermions, the modification of the cross section and of the expression of the bending angle, at leading order, are given by dσ dσ |E = cos2 (|~q|R) |L , dΩ dΩ

(55)

having used for dσ/dΩ|L the result given in (44). • Neutrinos and dark matter fermions Moving to the case of neutrino/ dark matter scatterings, as in the previous cases, the interaction region covers all the 3-dimensional space. Beside hµν E (x), it is convenient to introduce also the expression of the field generated µν inside the core of the source hI (x) and its corresponding form factor hI (q). In this case the toal form factor h(~q) is given by h(~q) = hI (~q) + hE (~q) (56) with 

 2ΦD d ~x − hI (~ q) = ei~q·~x κ

(179)

which requires a regularization at large momentum, due to the presence of an oscillating factor. The simplest way avoid any regularization is to redefine it as difference between the contribution of the entire region and that of the interior region (r < R), the latter being given by Z 4π ei~q·~x θ(R − |~x|) = 2 (1 − cos(|~q|R)) . (180) d3 ~x |~x| ~q R< Therefore we obtain Z

ei~q·~x d ~x θ(|~x| − R) = |~x| R> 3

= =

B

Z ei~q·~x ei~q·~x d3 ~x d ~x − θ(R − |~x|) |~x| |~x| R< 4π 4π − 2 (1 − cos(|~q|R)) ~q2 ~q 4π cos(|~q|R). ~q2

Z

3

(181)

List of g functions and c coefficients

The c coefficients of the flavour-changing cross section Eq. (97), in the case of equal mass fermions, are given by c1 = 72m2 − 18q 2

c3 = 32m2 − 8q 2

c4 = 12 (4m2 − q 2 )3

c7 = 8q 2

c8 = 21 q 2 (−4m2 + q 2 )2

c9 = 12 (q 2 )3

d1 3 = 96m2 − 24q 2

d1 4 = 6(−4m2 + q 2 )2

d1 5 = 6(4m2 − q 2 )q 2

d3 4 = 4(−4m2 + q 2 )2 d4 11 = 16 d1 11 d8 9 = (q 2 )2 (4m2 − q 2 ).

d3 5 = q 2 d3 4 d5 11 =

q2 4

d3 11

d3 11 = 23 d1 11

d7 8 = 4(4m2 − q 2 )q 2

c5 = 12 (4m2 − q 2 )q 2

2(m2 +2m2w )2 (4m2 −q 2 ) m4w 12(m2 +2m2w )(4m2 −q 2 ) d1 11 = m2w 2 2 d4 5 = q (−4m + q 2 )2 d7 9 = 4(q 2 )2

c11 =

(182) The functions gi (x) are given by ga (x) = +

gb (x) =

 1 2 3 4 4 44 − 194x + 243x − 98x + 5x 36 (x − 1)   6x 2 − 15x + 10x2 log (x)

   1 2 3 4 3 4 8 − 14x + 21x − 14x − x + 2x 4 + 3x + 2x log (x) 6 (x − 1) 34

gc (x) = − gd (x) = − gg (x) = + gh (x) = +

 1 2 3 4 5 6 2 + 9x − 152x + 88x + 54x − x 36 (x − 1)   12x −1 + 5x + 10x2 + x3 log (x)  1 2 3 4 5 6 6 − 83x + 200x + 12x − 142x + 7x 72 (x − 1)   12x 1 + 4x − 18x2 − 2x3 log (x)  1 2 3 4 5 6 106 + 245x − 240x + 20x + 70x − 201x 360 (x − 1)   12 2 + 23x + 18x2 + 8x3 + 18x4 + 6x5 log (x)   1 2 3 4 5 6 5 −6 + 41x + 64x − 180x + 70x + 11x 240 (x − 1)   4 −6 + 21x + 76x2 + 36x3 − 84x4 + 2x5 log (x) .

(183)

References [1] F. A. Berends and R. Gastmans, Ann. Phys. 98, 225 (1976). [2] G. Degrassi, E. Gabrielli, and L. Trentadue, Phys.Rev. D79, 053004 (2009), arXiv:0812.3262. [3] C. Corian`o, L. Delle Rose, E. Gabrielli, and L. Trentadue, Phys.Rev. D87, 054020 (2013), arXiv:1212.5029. [4] C. Corian`o, L. Delle Rose, E. Gabrielli, and L. Trentadue, Phys.Rev. D88, 085008 (2013), arXiv:1303.1305. [5] M. Giannotti and E. Mottola, Phys. Rev. D79, 045014 (2009), arXiv:0812.0351. [6] R. Armillis, C. Corian`o, and L. Delle Rose, Phys. Rev. D81, 085001 (2010), arXiv:0910.3381. [7] C. Corian`o, L. Delle Rose, A. Quintavalle, and M. Serino, Phys.Lett. B700, 29 (2011), arXiv:1101.1624. [8] C. Corian`o, L. Delle Rose, and M. Serino, Phys.Rev. D83, 125028 (2011), arXiv:1102.4558. [9] R. Escribano, J. Frere, D. Monderen, and V. Van Elewyck, arXiv:hep-ph/0105211.

Phys.Lett. B512, 8 (2001),

[10] O. Mena, I. Mocioiu, and C. Quigg, Astropart.Phys. 28, 348 (2007), arXiv:astro-ph/0610918. [11] E. F. Eiroa and G. E. Romero, Phys.Lett. B663, 377 (2008), arXiv:0802.4251. [12] A. Accioly and R. Paszko, Phys.Rev. D69, 107501 (2004). [13] F. Sorge and S. Zilio, Class.Quant.Grav. 25, 225004 (2008). [14] F. Sorge, Class.Quant.Grav. 29, 045002 (2012). [15] J. Callan, Curtis G., S. R. Coleman, and R. Jackiw, Annals Phys. 59, 42 (1970). 35

[16] J. Lawrence, Annals Phys. 58, 47 (1970). [17] G. Zatsepin and V. Kuzmin, JETP Lett. 4, 78 (1966). [18] K. Greisen, Phys.Rev.Lett. 16, 748 (1966).

36