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

Corrections to the Gyromagnetic Factor
in Very Special Relativity

Benjamin Koch [email protected] Institut für Theoretische Physik and Atominstitut, Technische Universität Wien, Wiedner Hauptstrasse 8–10, A-1040 Vienna, Austria Facultad de Física, Pontificia Universidad Católica de Chile, Vicuña Mackenna 4860, Santiago, Chile    Enrique Muñoz [email protected] Facultad de Física, Pontificia Universidad Católica de Chile, Vicuña Mackenna 4860, Santiago, Chile    Alessandro Santoni [email protected] Facultad de Física, Pontificia Universidad Católica de Chile, Vicuña Mackenna 4860, Santiago, Chile
Abstract

We consider the corrections arising from the SIM(2) invariant realization of Very Special Relativity to the energy spectrum of a CC-invariant Dirac Fermion in a static and homogeneous magnetic field 𝑩\bm{B}. First, we analyze the case of 𝑩\bm{B} parallel to the spatial VSR preferred direction 𝒏\bm{n}, finding that the expression for the energy spectrum stays the same, except for a mass shift arising from the VSR contribution. Then, we relax the parallelism condition, finding a new equation for the energy spectrum. We solve this equation perturbatively. With a Penning trap’s experiment in mind, we derive the first order VSR corrections to the electron’s g2g-2 factor. Finally, using the most accurate electron’s gg-factor measurements in Penning trap’s experiments, we obtain an upper bound to the VSR electron mass parameter, and therefore also to the VSR electronic neutrino mass, of 1eV~{}1\,eV. This result does not contradict the possibility for VSR to be the origin of neutrino masses.

I Introduction

Very Special Relativity (VSR) Cohen and Glashow (2006a) is a theory where the flat spacetime symmetries are reduced to a subgroup of the Lorentz group plus the group of spacetime translations, which it is kept unchanged. While its classical consequences are, to our current knowledge, identical to the ones implied by special relativity, the non-classical consequences have already been analyzed in several physics’ areas. The original idea that motivated Cohen and Glashow to formulate VSR was a new mechanism for the emergence of Neutrino’s masses Cohen and Glashow (2006b). Further applications of VSR that have been studied regard Supersymmetric extensions Cohen and Freedman (2007); Vohánka (2012), the Gravitational sector Gibbons et al. (2007); Kouretsis et al. (2009); Alfaro and Santoni (2022), Quantum Electrodynamics Alfaro and Soto (2019); Cheon et al. (2009), the Standard Model of Particle Physics Dunn and Mehen (2006); Alfaro (2019), and much more.
A minimal candidate for VSR is the T(2)T(2) subgroup, that contains the generators T1=K1+J2T_{1}=K_{1}+J_{2}, and T2=K2J1T_{2}=K_{2}-J_{1}, that combine boosts 𝐊\mathbf{K} and rotations 𝐉\mathbf{J}. When T(2)T(2) is enlarged to include J3J_{3}, the resulting subgroup is isomorphic to the 3-parameter group of Euclidean translations, E(2)E(2). On the other hand, if T(2)T(2) is enlarged to incorporate K3K_{3}, the resulting 3-parameter subgroup is isomorphic to the homotheties group HOM(2)HOM(2). Finally, if T(2)T(2) is equipped with both K3K_{3} and J3J_{3} generators, the resulting four parameter subgroup is isomorphic to the similitudes group SIM(2)SIM(2). Each of these four VSR subgroups can be expanded to the full Lorentz group by the addition of discrete symmetries PP, TT, CPCP, or CTCT. In the rest of this work, we shall focus on the SIM(2)SIM(2) subgroup of the Lorentz group, since it is the biggest proper Lorentz subgroup and it directly preserves CPT symmetry Cohen and Glashow (2006a), an important ingredient in all the known quantum field theories. Therefore, in the rest of the paper we will refer to this particular SIM(2)SIM(2)-realization of VSR just as VSR.
The main feature of the VSR theory is the introduction of a light-like preferred spacetime direction nμ=(n0,𝒏)n^{\mu}=(n^{0},\bm{n}), which under SIM(2)SIM(2) transformations changes only by a scaling factor

nμSIM(2)eϕnμ.n^{\mu}\underset{SIM(2)}{\longrightarrow}e^{\phi}n^{\mu}\,. (1)

Therefore, ratios of scalar products with nμn^{\mu} both in the numerator and denominator are invariant under SIM(2)SIM(2) transformations, but not under the full Lorentz group. The VSR idea, as it is implemented into the Dirac equation (2), is a non-local violation of Lorentz symmetry and thus different from the more well known approach proposed by Kostelecky Colladay and Kosteleckỳ (1998); Kosteleckỳ and Russell (2011), where Lorentz violation is generally parametrized by spurion fields, which in this particular realization of VSR are not present.
The focus of this paper will be to consider the effect of a static and homogeneous magnetic field 𝑩\bm{B} on the energy spectrum of a Dirac Fermion in the framework of VSR. Furthermore, since transition energies between these energy levels are measured in experiments with Penning traps to calculate the gyromagnetic factor of electrons Hanneke et al. (2011, 2008); Gabrielse et al. (2006), we can estimate the effect of VSR corrections to the g-factor as well.
The remaining of the paper is structured as follows: in Section II, we analyze the case in which the VSR spatial vector 𝒏\bm{n} is parallel to the direction of the magnetic field 𝑩\bm{B}, thus finding the corresponding exact energy eigenvalues. In section III, we study the more general case, allowing an angular spacing, labelled by the angle θ\theta, between the direction of 𝒏\bm{n} and the magnetic field. In section IV, considering the experimental setup of an electron Penning trap, we study the possible consequences of VSR on the measured gyromagnetic moment of the electron, finding corrections even at zero order in the magnetic field. Finally, in section V we state our conclusions about the results and the possible future applications. In the appendices A, B, C and D we show, respectively, the explicit calculations involved in the equation of motion for the upper spinor φ(x1)\varphi(x^{1}), the derivation of the integrals I1(n,k)I_{1}(n,k), I2(n,k)I_{2}(n,k), the calculation of the perturbation matrix elements Vα,αn¯V^{\bar{n}}_{\alpha,\alpha^{\prime}}, and some details on Borel regularization.

II Magnetic Field parallel to VSR direction

Let’s start working in the VSR framework by considering the C-symmetric Dirac equation for a charged fermion Dunn and Mehen (2006)

(i∂̸m+iM22)ψ(x)=0,\displaystyle\left(i\not{\partial}-m+i\frac{M^{2}}{2}\not{N}\right)\psi(x)=0\,, (2)

where Nμ=nμ/nN^{\mu}=n^{\mu}/n\cdot\partial. The VSR correction in this equation is controlled by the parameter MM such that for M0M\rightarrow 0 the full Lorentz symmetry is recovered. Now, we include the external, constant and uniform magnetic field 𝐁=B𝐞^3\mathbf{B}=B\,\mathbf{\hat{e}}_{3} by the usual minimal substitution μμ+ieAμ\partial_{\mu}\to\partial_{\mu}+ieA_{\mu}, being AμA_{\mu} the electromagnetic four-potential

(i∂̸em+iM22)ψ(x)=0,\displaystyle\left(i\not{\partial}-e\not{A}-m+i\frac{M^{2}}{2}\not{N}\right)\psi(x)=0\,, (3)

such that the VSR operator becomes

Nμ=nμn(+ieA).N^{\mu}=\frac{n^{\mu}}{n\cdot\left(\partial+ieA\right)}\,. (4)

Due to the nμn^{\mu}-rescaling simmetry of Eq. (2), we can always choose it in the form nμ=(1,𝐧^)n^{\mu}=(1,\mathbf{\hat{n}}), with |𝐧^|2=1|\mathbf{\hat{n}}|^{2}=1. Here, in particular, we define the “preferred” VSR null vector as

nμ=(1,0,0,1)nn=0.\displaystyle n^{\mu}=\left(1,0,0,1\right)\rightarrow n\cdot n=0\,. (5)

In what follows, we choose the metric g00=+1g^{00}=+1, gij=δijg^{ij}=-\delta_{ij} for i=1,2,3i=1,2,3. To represent the uniform magnetic field 𝐁\mathbf{B}, we choose the gauge

A0=A1=A3=0,A2(x1)=Bx1,\displaystyle A^{0}=A^{1}=A^{3}=0,\,A^{2}(x^{1})=Bx^{1}\,, (6)

such that translational symmetry along 𝐞^2\mathbf{\hat{e}}_{2} and 𝐞^3\mathbf{\hat{e}}_{3} is preserved

[Aμ(x),p3]=[Aμ(x),p2]=0,\displaystyle\left[A^{\mu}(x),p_{3}\right]_{-}=\left[A^{\mu}(x),p_{2}\right]_{-}=0\,, (7)

where the components of the momentum operator are defined as pj=pj=ijp_{j}=-p^{j}=i\partial_{j}. By taking this into account, the 2 and 3-components of the momentum can be susbstituted by their eigenvalues, i.e. p3k3p_{3}\rightarrow k_{3} and p2k2p_{2}\rightarrow k_{2}. Moreover, we have that an eigenstate of (3) is of the form

ψ(x)=eiEtei(k3x3+k2x2)(φ(x1)χ(x1)).\psi(x)=e^{-iEt}e^{i\left(k^{3}x^{3}+k^{2}x^{2}\right)}\left(\begin{array}[]{c}\varphi(x^{1})\\ \chi(x^{1})\end{array}\right)\,. (8)

For the gauge (6) we also have that nA=0n\cdot A=0 automatically, and hence the VSR operator (4) for an eigenstate of the form (8) simplifies to

iM22M22γ0γ3Ek3,\displaystyle i\frac{M^{2}}{2}\not{N}\rightarrow-\frac{M^{2}}{2}\frac{\gamma^{0}-\gamma^{3}}{E-k^{3}}\,, (9)

where γμ\gamma^{\mu} are the Dirac matrices, for which we choose the standard representation

γ0=(𝟏00𝟏),γi=(0σiσi0),\displaystyle\gamma^{0}=\left(\begin{array}[]{cc}\mathbf{1}&0\\ 0&-\mathbf{1}\end{array}\right),\,\,\gamma^{i}=\left(\begin{array}[]{cc}0&\sigma^{i}\\ -\sigma^{i}&0\end{array}\right)\,, (14)

where 𝟏\mathbf{1} is a two by two identity matrix and 𝝈\bm{\sigma} the usual Pauli’s matrices.

II.1 Equations of motion for φ(x1)\varphi(x^{1})

After inserting Eq. (8) and Eq. (9) into Eq. (3), and dividing the bi-spinor ψ\psi in its upper and lower components as

ψ(x1)=(φ(x1)χ(x1)),\psi(x^{1})=\left(\begin{array}[]{c}\varphi(x^{1})\\ \chi(x^{1})\end{array}\right)\,, (15)

we obtain the following equation

[(EmM22(Ek3))𝟏eBx1σ2+M22(Ek3)σ3σipieBx1σ2M22(Ek3)σ3+σipi(E+mM22(Ek3))𝟏](φ(x1)χ(x1))=0,\displaystyle\left[\begin{array}[]{cc}\left(E-m-\frac{M^{2}}{2\left(E-k^{3}\right)}\right)\mathbf{1}&eBx^{1}\sigma^{2}+\frac{M^{2}}{2\left(E-k^{3}\right)}\sigma^{3}-\sigma^{i}p^{i}\\ -eBx^{1}\sigma^{2}-\frac{M^{2}}{2\left(E-k^{3}\right)}\sigma^{3}+\sigma^{i}p^{i}&-\left(E+m-\frac{M^{2}}{2\left(E-k^{3}\right)}\right)\mathbf{1}\end{array}\right]\left(\begin{array}[]{c}\varphi(x^{1})\\ \chi(x^{1})\end{array}\right)=0\,, (20)

From the system (20), we solve for the lower spinor χ\chi in terms of the upper one φ\varphi, to obtain the expression

χ(x1)=eBx1σ2+M22(Ek3)σ3σipiE+mM22(Ek3)φ(x1).\displaystyle\chi(x^{1})=-\frac{eBx^{1}\sigma^{2}+\frac{M^{2}}{2(E-k^{3})}\sigma^{3}-\sigma^{i}p^{i}}{E+m-\frac{M^{2}}{2(E-k^{3})}}\varphi(x^{1})\,. (21)

Inserting Eq. (21) back into the first system’s equation of (20), we obtain an expression for the upper spinor φ\varphi

[(EM22(Ek3))2m2\displaystyle\left[\left(E-\frac{M^{2}}{2\left(E-k^{3}\right)}\right)^{2}-m^{2}\right. (22)
(eBx1σ2+M22(Ek3)σ3σipi)2]φ(x1)=0.\displaystyle\left.-\left(eBx^{1}\sigma^{2}+\frac{M^{2}}{2(E-k^{3})}\sigma^{3}-\sigma^{i}p^{i}\right)^{2}\right]\varphi(x^{1})=0\,.

By applying the standard properties of the SU(2) algebra, we can calculate the square of the differential operator in Eq. (22), to obtain

[(EM22(Ek3))2m2(M22(Ek3)k3)2\displaystyle\left[\left(E-\frac{M^{2}}{2\left(E-k^{3}\right)}\right)^{2}-m^{2}-\left(\frac{M^{2}}{2(E-k^{3})}-k^{3}\right)^{2}\right.
(eBx1k2)2+eBσ3p12]φ(x1)=0.\displaystyle\left.-\left(eBx^{1}-k^{2}\right)^{2}+eB\sigma^{3}-p_{1}^{2}\right]\varphi(x^{1})=0\,. (23)

Clearly, the equation above is diagonal in the two components of the spinor φ(x1)\varphi(x^{1})

φ(x1)=(f1(x1)f2(x1)).\displaystyle\varphi(x^{1})=\left(\begin{array}[]{c}f^{1}(x^{1})\\ f^{2}(x^{1})\end{array}\right)\,. (26)

Therefore, the eigenvalue problem will have two independent solutions, that we define by

𝒇+1(x1)(f+1(x1)0),𝒇1(x1)(0f1(x1)).\bm{f}_{+1}(x^{1})\equiv\left(\begin{array}[]{c}f_{+1}(x^{1})\\ 0\end{array}\right)\,,\,\,\,\bm{f}_{-1}(x^{1})\equiv\left(\begin{array}[]{c}0\\ f_{-1}(x^{1})\end{array}\right)\,. (27)

The fα(x1)f_{\alpha}(x^{1}) introduced above, with α=±1\alpha=\pm 1 representing the two eigenvalues of σ3\sigma^{3}, are given by the solution to the two independent differential equations

[12+(eBx1k2)2αeBeBa(k3,E)]fα(x1)=0,\displaystyle\left[-\partial_{1}^{2}+\left(eBx^{1}-k^{2}\right)^{2}-\alpha eB-eBa(k_{3},E)\right]f_{\alpha}(x^{1})=0\,,
(28)

where we defined the coefficient

a(k3,E)=1eB\displaystyle a(k^{3},E)=\frac{1}{eB} [(EM22(Ek3))2m2\displaystyle\left[\left(E-\frac{M^{2}}{2(E-k^{3})}\right)^{2}-m^{2}\right. (29)
(k3M22(Ek3))2].\displaystyle\left.-\left(k^{3}-\frac{M^{2}}{2(E-k^{3})}\right)^{2}\right]\,.

It is convenient to define the dimensionless coordinate

ξ=eB(x1k2eB),\displaystyle\xi=\sqrt{eB}\left(x^{1}-\frac{k^{2}}{eB}\right)\,, (30)

such that Eq. (28) becomes (for α=±1\alpha=\pm 1)

[d2dξ2+ξ2α]fα(ξ)=a(k3,E)fα(ξ).\displaystyle\left[-\frac{d^{2}}{d\xi^{2}}+\xi^{2}-\alpha\right]f_{\alpha}(\xi)=a(k^{3},E)f_{\alpha}(\xi)\,. (31)

The only L2L^{2}-normalizable solutions of Eq. (31) are the functions

fn,α(ξ)=Ceξ2/2Hn(ξ),\displaystyle f_{n,\alpha}(\xi)=Ce^{-\xi^{2}/2}H_{n}(\xi)\,, (32)

where Hn(ξ)H_{n}(\xi) are the Hermite polynomials of order n0n\in\mathbb{N}_{0}, provided the following quantization condition is satisfied

a(k3,E)+α=2n+1,n0,α=±1,\displaystyle a(k^{3},E)+\alpha=2n+1,\,n\in\mathbb{N}_{0},\,\alpha=\pm 1\,, (33)

while, the normalization coefficient CC in Eq. (32) is obtained from the orthonormality condition for the Hermite polynomials

+𝑑ξeξ2Hn(ξ)Hm(ξ)=2nn!πδnm.\int_{-\infty}^{+\infty}d\xi\;e^{-\xi^{2}}H_{n}(\xi)H_{m}(\xi)=2^{n}n!\sqrt{\pi}\delta_{nm}\,. (34)

II.2 Energy Spectrum

Following Eq. (33), the energy spectrum EE±(k3,n,α)E\equiv E_{\pm}\left(k^{3},n,\alpha\right) is defined by the roots of the algebraic equation

(EM22(Ek3))2\displaystyle\left(E-\frac{M^{2}}{2(E-k^{3})}\right)^{2} \displaystyle- (k3M22(Ek3))2=\displaystyle\left(k^{3}-\frac{M^{2}}{2(E-k^{3})}\right)^{2}= (35)
=\displaystyle= eB(2n+1α)+m2.\displaystyle eB\left(2n+1-\alpha\right)+m^{2}\,.

This equation can be solved explicitly, to obtain the exact energy eigenvalues

E±(0)(k3,n,α)\displaystyle E^{(0)}_{\pm}(k^{3},n,\alpha) =\displaystyle= ±eB(2n+1α)+(k3)2+mf2,\displaystyle\pm\sqrt{eB(2n+1-\alpha)+(k^{3})^{2}+m^{2}_{f}}\,,

where

mf2=m2+M2.m^{2}_{f}=m^{2}+M^{2}. (37)

One confirms that when the full Lorentz symmetry is restored in limit M20M^{2}\rightarrow 0, the spectrum in Eq. (LABEL:eq_VSRspectrum) reduces to the well known “unperturbed” solutions

E±(0)(k3,n,α)|M=0=±E(u)(k3,n,α),\displaystyle\left.E^{(0)}_{\pm}(k^{3},n,\alpha)\right|_{M=0}=\pm E^{(u)}(k^{3},n,\alpha)\,, (38)

with

E(u)(k3,n,α)=eB(2n+1α)+(k3)2+m2.\displaystyle E^{(u)}(k^{3},n,\alpha)=\sqrt{eB(2n+1-\alpha)+(k^{3})^{2}+m^{2}}\,.

Moreover, it is also clear from Eq. (LABEL:eq_VSRspectrum) that for this configuration, where the field 𝐁\mathbf{B} is parallel to the direction of 𝐧^{\mathbf{\hat{n}}}, the sole effect of the VSR term in the single-particle spectrum is to shift the particle’s mass mmfm\rightarrow m_{f}.
Following the analysis in the previous section, we remark that, apart from the ground state energy with n=0n=0 and α=+1\alpha=+1, each energy eigenvalue is degenerate since we can obtain it with the two combinations (n,α=1)(n\,,\alpha=-1) and (n+1,α=+1)(n+1\,,\alpha=+1).
Therefore, introducing the system’s eigenstates |𝒇(0)\ket{\bm{f}^{(0)}} such that 𝒇(0)(ξ)=ξ|𝒇(0)\bm{f}^{(0)}(\xi)=\langle\xi|\bm{f}^{(0)}\rangle, our eigenvector’s basis will look like

{|𝒇(0),0,+1,\displaystyle\bigg{\{}\ket{\bm{f}^{(0)},0,+1}, [|𝒇(0),0,1,|𝒇(0),1,+1],,\displaystyle\left[\ket{\bm{f}^{(0)},0,-1},\ket{\bm{f}^{(0)},1,+1}\right]\,,\,...\,,
[|𝒇(0),n,1,|𝒇(0),n+1,+1],},\displaystyle\left[\ket{\bm{f}^{(0)},n,-1},\ket{\bm{f}^{(0)},n+1,+1}\right],\,...\bigg{\}}\,,

with the square brackets highlighting the 2-dimensional degenerate eigenspaces. To simplify the notation in the rest of the calculations and to label each degenerate eigenspace, we re-order the eigenstates by |𝒇(0),n,α|𝒇(0),n¯,α\ket{\bm{f}^{(0)},n,\alpha}\to\ket{\bm{f}^{(0)},\bar{n},\alpha}, where

{n¯=nforα=+1n¯=n+1forα=1.\left\{\begin{array}[]{l}\bar{n}=n\;\;\;\;\;\;\;\;\;for\;\alpha=+1\\ \bar{n}=n+1\;\;\;for\;\alpha=-1\;.\end{array}\right. (41)

The eigenvector’s basis then becomes

{|𝒇(0),0¯,+1,\displaystyle\bigg{\{}\ket{\bm{f}^{(0)},\bar{0},+1}, [|𝒇(0),1¯,1,|𝒇(0),1¯,+1],,\displaystyle\left[\ket{\bm{f}^{(0)},\bar{1},-1},\ket{\bm{f}^{(0)},\bar{1},+1}\right]\,,\,...\,,
[|𝒇(0),n¯,1,|𝒇(0),n¯,+1],},\displaystyle\left[\ket{\bm{f}^{(0)},\bar{n},-1},\ket{\bm{f}^{(0)},\bar{n},+1}\right],\,...\bigg{\}}\,,

where each degenerate eigenspace is spanned by the two orthogonal eigenvectors

ξ|𝒇(0),n¯,+1\displaystyle\langle\xi|\bm{f}^{(0)},\bar{n},+1\rangle =\displaystyle= 1π1/42n¯2n¯!(eξ22Hn¯(ξ)0),\displaystyle\frac{1}{\pi^{1/4}2^{\frac{\bar{n}}{2}}\sqrt{\bar{n}!}}\left(\begin{array}[]{c}e^{-\frac{\xi^{2}}{2}}H_{\bar{n}}(\xi)\\ 0\end{array}\right)\,, (45)
ξ|𝒇(0),n¯,1\displaystyle\langle\xi|\bm{f}^{(0)},\bar{n},-1\rangle =\displaystyle= 1π1/42n¯12(n¯1)!(0eξ22Hn¯1(ξ)),\displaystyle\frac{1}{\pi^{1/4}2^{\frac{\bar{n}-1}{2}}\sqrt{(\bar{n}-1)!}}\left(\begin{array}[]{c}0\\ e^{-\frac{\xi^{2}}{2}}H_{\bar{n}-1}(\xi)\end{array}\right)\,, (48)

while the ground state n¯=n=0\bar{n}=n=0 is defined by

ξ|𝒇(0),0¯,+1=1π1/4(eξ22H0(ξ)0),\langle\xi|\bm{f}^{(0)},\bar{0},+1\rangle=\frac{1}{\pi^{1/4}}\left(\begin{array}[]{c}e^{-\frac{\xi^{2}}{2}}H_{0}(\xi)\\ 0\end{array}\right)\,, (49)

and its energy is not degenerate.

III Magnetic Field not parallel to VSR direction

Let us now consider a magnetic field 𝐁\mathbf{B} oriented with an angle θ[0,π]\theta\in[0,\pi] with respect to the VSR unit vector 𝐧^{\mathbf{\hat{n}}}, i.e. 𝐁𝐧^=Bcosθ\mathbf{B}\cdot{\mathbf{\hat{n}}}=B\cos\theta. Therefore, without loss of generality, we choose the coordinate system such that 𝐁=B𝐞^𝟑\mathbf{B}=B\,\mathbf{\hat{e}_{3}}, and

nμ=(1,𝐧^)=(1,sinθ,0,cosθ).\displaystyle n^{\mu}=(1,{\mathbf{\hat{n}}})=(1,\sin\theta,0,\cos\theta)\,. (50)

We still choose the same gauge as in the previous case, Aμ=(0,0,Bx1,0)A^{\mu}=(0,0,Bx^{1},0), so that the translational invariance along the x2x^{2} and x3x^{3} directions allows us to choose the same separation of variables as in Eq. (8), while preserving the light-cone condition nA=0n\cdot A=0.
In this case, however, the VSR term in the equation of motion reduces to the form

iM22M22γ0Eγ1sinθγ3cosθEk3cosθp1sinθ.\displaystyle i\frac{M^{2}}{2}\not{N}\rightarrow-\frac{M^{2}}{2}\frac{\gamma^{0}E-\gamma^{1}\sin\theta-\gamma^{3}\cos\theta}{E-k^{3}\cos\theta-p^{1}\sin\theta}\,. (51)

III.1 Equations of motion for φ(x1)\varphi(x^{1})

Substituting Eq. (51) into Eq. (3), we obtain for this general case the expression

[γ0E+γ1p1γ2(k2eBx1)γ3k3m\displaystyle\left[\gamma^{0}E+\gamma^{1}p_{1}-\gamma^{2}\left(k^{2}-eBx^{1}\right)-\gamma^{3}k^{3}-m\right. (52)
M22γ0Eγ1sinθγ3cosθEk3cosθp1sinθ](φ(x1)χ(x1))=0.\displaystyle\left.-\frac{M^{2}}{2}\frac{\gamma^{0}E-\gamma^{1}\sin\theta-\gamma^{3}\cos\theta}{E-k^{3}\cos\theta-p^{1}\sin\theta}\right]\left(\begin{array}[]{c}\varphi(x^{1})\\ \chi(x^{1})\end{array}\right)=0\,. (55)

From this linear system, the lower spinor χ(x1)\chi(x^{1}) can be solved in terms of the upper component φ(x1)\varphi(x^{1}), as done in the case 𝐁𝐧^\mathbf{B}\parallel\mathbf{\hat{n}}, and then replaced again to obtain the following single equation for the upper spinor φ(x1)\varphi(x^{1})

[(EM2/2Ek3cosθp1sinθ)2m2\displaystyle\left[\left(E-\frac{M^{2}/2}{E-k^{3}\cos\theta-p^{1}\sin\theta}\right)^{2}-m^{2}\right.
((p1M2sinθ/2Ek3cosθp1sinθ)σ1+(k2eBx1)σ2+(k3M2cosθ/2Ek3cosθp1sinθ)σ3)2]φ(x1)=0.\displaystyle\left.-\left(\left(p^{1}-\frac{M^{2}\sin\theta/2}{E-k^{3}\cos\theta-p^{1}\sin\theta}\right)\,\sigma^{1}+\left(k^{2}-eBx^{1}\right)\sigma^{2}+\left(k^{3}-\frac{M^{2}\cos\theta/2}{E-k^{3}\cos\theta-p^{1}\sin\theta}\right)\sigma^{3}\right)^{2}\right]\varphi(x^{1})=0\,. (56)

By expanding the squares, we observe that, due to the anti-commutation relations satisfied by the Pauli matrices, among the terms with mixed σ\sigma’s only the ones that involve operators acting on x1φ(x1)x^{1}\varphi(x^{1}) can have a chance to generate a surviving part after summing up the sigma products with interchanged indices.
For example, for terms involving p1p^{1}, we can exploit the fact that, for any function g(x1)g(x^{1})

[σ1p1,g(x1)σ2]+φ(x1)=σ31g(x1)φ(x1).\left[\sigma^{1}p^{1},g(x^{1})\sigma^{2}\right]_{+}\varphi(x^{1})=-\sigma^{3}\partial_{1}g(x^{1})\varphi(x^{1})\,. (57)

Applying σ\sigma’s anticommutation rules, the relation Eq. (57) and keeping in mind the representation Eq. (75) for the inverse operator, Eq. (56) is reduced to the expression (see appendix A for an explicit calculation)

[E2(k3)2m2M2(p1)2+eBσ3(k2eBx1)2\displaystyle\left[E^{2}-(k^{3})^{2}-m^{2}-M^{2}-(p^{1})^{2}+eB\sigma^{3}-(k^{2}-eBx^{1})^{2}\right.
M22eBsin2θ(Ek3cosθp1sinθ)2σ3+M22eBsinθcosθ(Ek3cosθp1sinθ)2σ1]φ(x1)=0.\displaystyle\left.-\frac{M^{2}}{2}eB\frac{\sin^{2}\theta}{(E-k^{3}\cos\theta-p^{1}\sin\theta)^{2}}\sigma^{3}+\frac{M^{2}}{2}eB\frac{\sin\theta\cos\theta}{(E-k^{3}\cos\theta-p^{1}\sin\theta)^{2}}\sigma^{1}\right]\varphi(x^{1})=0\,. (58)

Now, let us define:

a(k3,E)=1eB(E2(k3)2m2M2),\displaystyle a(k^{3},E)=\frac{1}{eB}\left(E^{2}-(k^{3})^{2}-m^{2}-M^{2}\right)\,,
E~=Ek3cosθ,\displaystyle\tilde{E}=E-k^{3}\cos\theta\,, (59)

along with the change of variables

ξ=eB(x1k2/(eB)).\displaystyle\xi=\sqrt{eB}\left(x^{1}-k^{2}/(eB)\right)\,. (60)

Therefore, dividing Eq. (III.1) by eBeB and using the above definitions, we obtain

[ξ2+ξ2σ3a(k3,E)\displaystyle\left[-\partial_{\xi}^{2}+\xi^{2}-\sigma^{3}-a(k^{3},E)\right. (61)
+M2sinθ/2(E~eBpξsinθ)2(sinθσ3cosθσ1)]φ(x1)=0,\displaystyle+\left.\frac{M^{2}\sin\theta/2}{(\tilde{E}-\sqrt{eB}p^{\xi}\sin\theta)^{2}}(\sin\theta\sigma^{3}-\cos\theta\sigma^{1})\right]\varphi(x^{1})=0\,,

which, defining the operator PξE~+ieBsinθξP_{\xi}\equiv\tilde{E}+i\sqrt{eB}\sin\theta\partial_{\xi}, can be expressed as an eigenvalue equation

[ξ2+ξ2σ3+M2sinθ2Pξ2(sinθσ3\displaystyle\left[-\partial_{\xi}^{2}+\xi^{2}-\sigma^{3}+\frac{M^{2}\sin\theta}{2P^{2}_{\xi}}(\sin\theta\,\sigma^{3}\right.
cosθσ1)](f1(ξ)f2(ξ))=a(f1(ξ)f2(ξ)).\displaystyle\left.-\cos\theta\,\sigma^{1})\right]\left(\begin{array}[]{c}f^{1}(\xi)\\ f^{2}(\xi)\end{array}\right)=a\left(\begin{array}[]{c}f^{1}(\xi)\\ f^{2}(\xi)\end{array}\right)\,. (66)

III.2 VSR Perturbative Scheme

Thinking of VSR as a correction to special relativity, we can consider MM as a small parameter compared to the other system’s energy scales, like mm or eB\sqrt{eB}, so that we identify an unperturbed operator H0H_{0} and a perturbation operator VV

H0=ξ2+ξ2σ3\displaystyle H_{0}=-\partial_{\xi}^{2}+\xi^{2}-\sigma^{3}
V=sinθPξ2(sinθσ3cosθσ1).\displaystyle V=\frac{\sin\theta}{P^{2}_{\xi}}(\sin\theta\sigma^{3}-\cos\theta\sigma^{1})\,. (67)

Thus, in terms of the system’s eigenstates |𝒇\ket{\bm{f}}, we can write Eq. (III.1) as

(H0+M22V)|𝒇=a|𝒇.(H_{0}+\frac{M^{2}}{2}V)\ket{\bm{f}}=a\ket{\bm{f}}\,. (68)

We can then approach the problem in a perturbative scheme. Let’s define λ:=M2/2\lambda:=M^{2}/2 and expand in a power series |𝒇\ket{\bm{f}} and aa so that

a\displaystyle a =\displaystyle= a(0)+λa(1)+O(λ2),\displaystyle a^{(0)}+\lambda a^{(1)}+O(\lambda^{2})\,,
|𝒇\displaystyle\ket{\bm{f}} =\displaystyle= |𝒇(0)+λ|𝒇(1)+O(λ2).\displaystyle\ket{\bm{f}^{(0)}}+\lambda\ket{\bm{f}^{(1)}}+O(\lambda^{2})\,. (69)

Therefore, we are considering a perturbation to the solution already obtained in the previous section

H0|𝒇(0),n¯,α=an¯,α(0)|𝒇(0),n¯,α,H_{0}\ket{\bm{f}^{(0)},\bar{n},\alpha}=a^{(0)}_{\bar{n},\alpha}\ket{\bm{f}^{(0)},\bar{n},\alpha}\,, (70)

with ξ|𝒇(0),n¯,α\braket{\xi}{\bm{f}^{(0)},\bar{n},\alpha} defined by Eqs. (45), (49), and an,α(0)=2n+1αa^{(0)}_{n,\alpha}=2n+1-\alpha, or in the n¯\bar{n}-basis an¯,α(0)=2n¯a^{(0)}_{\bar{n},\alpha}=2\bar{n}. At first order in the perturbation V{V}, we find

(H0an¯,α(0)𝟏)|𝒇(1),n¯,α=(an¯,α(1)𝟏V)|𝒇(0),n¯,α(H_{0}-a^{(0)}_{\bar{n},\alpha}\mathbf{1})\ket{\bm{f}^{(1)},\bar{n},\alpha}=(a^{(1)}_{\bar{n},\alpha}\mathbf{1}-V)\ket{\bm{f}^{(0)},\bar{n},\alpha}\, (71)

III.2.1 Perturbative correction to the state n¯=0\bar{n}=0

For the case n¯=0\bar{n}=0, we necessarily have α=+1\alpha=+1 and no degenerate eigenvectors, corresponding then to ordinary perturbation theory. Multiplying (71) by f(0),0¯,+1|\bra{f^{(0)},\bar{0},+1}, we see that

a0¯,+1(1)\displaystyle a^{(1)}_{\bar{0},+1} =\displaystyle= 𝒇(0),0¯,+1|V|𝒇(0),0¯,+1\displaystyle\bra{\bm{f}^{(0)},\bar{0},+1}{V}\ket{\bm{f}^{(0)},\bar{0},+1} (72)

Inserting two identities through the completeness relation 1=𝑑ξ|ξξ|1=\int d\xi\ket{\xi}\bra{\xi} and using (49) we obtain

a0¯,+1(1)=\displaystyle a^{(1)}_{\bar{0},+1}= 1π𝑑ξeξ2/2H0(ξ)sin2θPξ2(eξ2/2H0(ξ))\displaystyle\frac{1}{\sqrt{\pi}}\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{0}(\xi)\frac{\sin^{2}\theta}{P^{2}_{\xi}}(e^{-\xi^{2}/2}H_{0}(\xi))
=\displaystyle= sin2θπE~2ddAdξeξ2/2H0(ξ)×\displaystyle-\frac{\sin^{2}\theta}{\sqrt{\pi}\tilde{E}^{2}}\frac{d}{dA}\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{0}(\xi)\times (73)
×1A+iηsinθξ(eξ2/2H0(ξ))|A=1,\displaystyle\times\frac{1}{A+i\eta\sin\theta\partial_{\xi}}(e^{-\xi^{2}/2}H_{0}(\xi))\bigg{|}_{A=1}\,,

where we have defined the dimensionless quantity

η=eB/E~.\eta=\sqrt{eB}/{\tilde{E}}. (74)

For calculation purposes, we introduce a Schwinger-type integral representation for the inverse operator in Eq. (III.2.1) (valid for A>0A>0)

1A+iηsinθξ=0𝑑tet(A+iηsinθξ).\displaystyle\frac{1}{A+i\eta\sin\theta\partial_{\xi}}=\int_{0}^{\infty}dt\,e^{-t\left(A+i\eta\sin\theta\partial_{\xi}\right)}. (75)

With the integral form in Eq. (75), we get

a0¯,+1(1)=\displaystyle a^{(1)}_{\bar{0},+1}= sin2θπE~2ddA0dtdξeξ2/2H0(ξ)×\displaystyle-\frac{\sin^{2}\theta}{\sqrt{\pi}\tilde{E}^{2}}\frac{d}{dA}\int_{0}^{\infty}dt\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{0}(\xi)\times
×eAt(1+iηsinθξ)(eξ2/2H0(ξ))|A=1=\displaystyle\times e^{-At(1+i\eta\sin\theta\partial_{\xi})}(e^{-\xi^{2}/2}H_{0}(\xi))\bigg{|}_{A=1}=
=\displaystyle= sin2θπE~2ddA0eAtdtdξeξ2/2H0(ξ)×\displaystyle-\frac{\sin^{2}\theta}{\sqrt{\pi}\tilde{E}^{2}}\frac{d}{dA}\int_{0}^{\infty}e^{-At}dt\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{0}(\xi)\times (76)
×eitηsinθξ(eξ2/2H0(ξ))|A=1,\displaystyle\times e^{-it\eta\sin\theta\partial_{\xi}}(e^{-\xi^{2}/2}H_{0}(\xi))\bigg{|}_{A=1}\,,

where clearly we take the limit A1A\rightarrow 1 at the end.

Expanding the exponential operator

a0¯,+1(1)=\displaystyle a^{(1)}_{\bar{0},+1}= sin2θπE~2k=0(iηsinθ)kk!ddA0eAttkdt×\displaystyle-\frac{\sin^{2}\theta}{\sqrt{\pi}\tilde{E}^{2}}\sum_{k=0}^{\infty}\frac{(-i\eta\sin\theta)^{k}}{k!}\frac{d}{dA}\int_{0}^{\infty}e^{-At}t^{k}dt\times
×dξeξ2/2H0(ξ)ξk(eξ2/2H0(ξ))|A=1\displaystyle\times\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{0}(\xi)\partial_{\xi}^{k}(e^{-\xi^{2}/2}H_{0}(\xi))\bigg{|}_{A=1}
=\displaystyle= sin2θπE~2k=0(k+1)(iηsinθ)kI1(0,k),\displaystyle\frac{\sin^{2}\theta}{\sqrt{\pi}\tilde{E}^{2}}\sum_{k=0}^{\infty}(k+1)(i\eta\sin\theta)^{k}I_{1}(0,k)\,, (77)

where we defined the integral

I1(n¯,k)=𝑑ξeξ2/2Hn¯(ξ)ξk(eξ2/2Hn¯(ξ)).I_{1}(\bar{n},k)=\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{\bar{n}}(\xi)\partial_{\xi}^{k}(e^{-\xi^{2}/2}H_{\bar{n}}(\xi))\,. (78)

As shown in Appendix B, the parity properties of the integrand imply that it vanishes for odd values of k.
Moreover, as shown in detail in Appendix B, the analytical expression for the n¯=0\bar{n}=0 case is

I1(0,2k)=π 22k(1)kΓ(2k+1)Γ(k+1),\displaystyle I_{1}(0,2k)=\sqrt{\pi}\,2^{-2k}(-1)^{k}\frac{\Gamma(2k+1)}{\Gamma(k+1)}\,, (79)

while for n¯>0\bar{n}>0 we have

I1(n¯,2k)=π(1)kn¯! 2n¯2kΓ(2k+1)Γ(k+1)F(k,n¯;1;2),\displaystyle I_{1}(\bar{n},2k)=\sqrt{\pi}(-1)^{k}\bar{n}!\,2^{\bar{n}-2k}\frac{\Gamma(2k+1)}{\Gamma(k+1)}F(-k,-\bar{n};1;2)\,,

where F(a,b;c;z)F(a,b;c;z) is the Hypergeometric function. Therefore, we can re-write Eq. (III.2.1) as

a0¯,+1(1)\displaystyle a^{(1)}_{\bar{0},+1} =\displaystyle= sin2θπE~2k=0(2k+1)(1)k(ηsinθ)2kI1(0,2k)\displaystyle\frac{\sin^{2}\theta}{\sqrt{\pi}\tilde{E}^{2}}\sum_{k=0}^{\infty}(2k+1)(-1)^{k}\left(\eta\sin\theta\right)^{2k}I_{1}(0,2k) (81)
=\displaystyle= sin2θE~2k=0(ηsinθ2)2kΓ(2k+2)Γ(k+1),\displaystyle\frac{\sin^{2}\theta}{\tilde{E}^{2}}\sum_{k=0}^{\infty}\left(\frac{\eta\sin\theta}{2}\right)^{2k}\frac{\Gamma(2k+2)}{\Gamma(k+1)},

which is a completely real expression as one would expect. The kk-sum in Eq. (81) is not convergent in the standard sense. However, it can be regularized, for example, by the Borel prescription, to obtain a closed form in terms of the incomplete Gamma function Γ(1/2,z)\Gamma(-1/2,z) as follows (see Appendix D for details)

a0¯,+1(1)=sin2θE~e1η2sin2θ(η2sin2θ)3/2Γ(12,1η2sin2θ).\displaystyle a^{(1)}_{\bar{0},+1}=\frac{\sin^{2}\theta}{\tilde{E}}\frac{e^{-\frac{1}{\eta^{2}\sin^{2}\theta}}}{\left(-\eta^{2}\sin^{2}\theta\right)^{3/2}}\,\Gamma\left(-\frac{1}{2},-\frac{1}{\eta^{2}\sin^{2}\theta}\right)\,. (82)

At the lowest orders for η1\eta\ll 1, both the power series Eq.(81) and the regularized Borel Eq.(82) reduces to,

a0¯,+1(1)sin2θE~2[1+32η2sin2θ+O(η4)].\displaystyle a^{(1)}_{\bar{0},+1}\simeq\frac{\sin^{2}\theta}{\tilde{E}^{2}}\left[1+\frac{3}{2}\eta^{2}\sin^{2}\theta+O(\eta^{4})\right]\,. (83)

We remark that, at this order η2\eta^{2}, the result is unique regardless of the regularization prescription. Then, to go further in our analysis, we assume to be in a situation where the magnetic field is small respect to other energy scales η=eBE~1\eta=\frac{\sqrt{eB}}{\tilde{E}}\ll 1.
In this weak field approximation, we have that at first order in λ\lambda, going back to the nn-notation (that does not make any difference for the ground state)

a0,+1=a0,+1(0)+λa0,+1(1)λsin2θE~2[1+32η2sin2θ].a_{0,+1}=a^{(0)}_{0,+1}+\lambda a^{(1)}_{0,+1}\approx\lambda\frac{\sin^{2}\theta}{\tilde{E}^{2}}\left[1+\frac{3}{2}\eta^{2}\sin^{2}\theta\right]\,. (84)

III.2.2 Perturbative correction to the states n¯>0\bar{n}>0

For n¯>0\bar{n}>0, instead, we notice that Eq. (71) corresponds to degenerate perturbation theory within the subspace spanned by the two degenerate spinors defined by Eq. (45), i.e. we have

|𝒇(1),n¯,α=α=±Cα,αn¯|𝒇(0),n¯,α.\displaystyle\ket{\bm{f}^{(1)},\bar{n},\alpha}=\sum_{\alpha^{\prime}=\pm}C^{\bar{n}}_{\alpha,\alpha^{\prime}}\ket{\bm{f}^{(0)},\bar{n},\alpha^{\prime}}. (85)

Therefore, projecting Eq. (71) over each of the spinors 𝒇(0),n¯,α|\bra{\bm{f}^{(0)},\bar{n},\alpha^{\prime}}, and applying the zero-order property 𝒇(0),n¯,α|(H0an¯,α(0)𝟏)=0\bra{\bm{f}^{(0)},\bar{n},\alpha}(H^{0}-a^{(0)}_{\bar{n},\alpha}\mathbf{1})=0, we obtain the linear eigenvalue system

[V+1,+1n¯V+1,1n¯V1,+1n¯V1,1n¯](Cα,+1n¯Cα,1n¯)=an¯,α(1)(Cα,+1n¯Cα,1n¯),\left[\begin{array}[]{cc}V^{\bar{n}}_{+1,+1}&V^{\bar{n}}_{+1,-1}\\ V^{\bar{n}}_{-1,+1}&V^{\bar{n}}_{-1,-1}\end{array}\right]\left(\begin{array}[]{c}C^{\bar{n}}_{\alpha,+1}\\ C^{\bar{n}}_{\alpha,-1}\end{array}\right)=a^{(1)}_{\bar{n},\alpha}\left(\begin{array}[]{c}C^{\bar{n}}_{\alpha,+1}\\ C^{\bar{n}}_{\alpha,-1}\end{array}\right), (86)

where we defined the matrix elements of the perturbation within the subspace of unperturbed degenerate states

Vα,αn¯=𝒇(0),n¯,α|V|𝒇(0),n¯,α.\displaystyle V^{\bar{n}}_{\alpha,\alpha^{\prime}}=\bra{\bm{f}^{(0)},\bar{n},\alpha}{V}\ket{\bm{f}^{(0)},\bar{n},\alpha^{\prime}}. (87)

Up to second order in η\eta, the matrix is explicitly given by

Vn¯=sin2θE~2[1+3(1+2n¯)2η2sin2θiηn¯2cosθiηn¯2cosθ(1+3(1+2n¯)2η2sin2θ)].V^{\bar{n}}=\frac{\sin^{2}\theta}{\tilde{E}^{2}}\left[\begin{array}[]{cc}1+\frac{3(1+2\bar{n})}{2}\eta^{2}\sin^{2}\theta&i\eta\sqrt{\frac{\bar{n}}{2}}\cos\theta\\ -i\eta\sqrt{\frac{\bar{n}}{2}}\cos\theta&-(1+\frac{3(1+2\bar{n})}{2}\eta^{2}\sin^{2}\theta)\end{array}\right]\,. (88)

The calculation of the Vα,αn¯V^{\bar{n}}_{\alpha,\alpha^{\prime}}, that follows from a similar procedure as in the previous n¯=0\bar{n}=0 case, is shown in appendix C.
Therefore, the first-order correction an¯,α(1)a^{(1)}_{\bar{n},\alpha}, defined up to order η2\eta^{2}, is obtained from the two eigenvalues of the linear system Eq. (86), i.e. from the characteristic equation det(V𝟏an¯(1))=0\det(V-\mathbf{1}\,a_{\bar{n}}^{(1)})=0, which reads

(an¯,α(1)E~2sin2θ+1+3(1+2n¯)2η2sin2θ)(an¯,α(1)E~2sin2θ1\displaystyle\left(\frac{a^{(1)}_{\bar{n},\alpha}\tilde{E}^{2}}{\sin^{2}\theta}+1+\frac{3(1+2\bar{n})}{2}\eta^{2}\sin^{2}\theta\right)\left(\frac{a^{(1)}_{\bar{n},\alpha}\tilde{E}^{2}}{\sin^{2}\theta}-1\right.
3(1+2n¯)2η2sin2θ)n¯2η2cos2θ=0.\displaystyle\left.\;\;\;\;\;\;\;\;\;\;\;\;-\frac{3(1+2\bar{n})}{2}\eta^{2}\sin^{2}\theta\right)-\frac{\bar{n}}{2}\eta^{2}\cos^{2}\theta=0\,. (89)

Solving for the two eigenvalues an,α(1)a_{n,\alpha}^{(1)} up to order η2\eta^{2}, we obtain

an¯,±1(1)\displaystyle a^{(1)}_{\bar{n},\pm 1} =\displaystyle= ±sin2θE~2(1+3(1+2n¯)2η2sin2θ)2+n¯2η2cos2θ\displaystyle\pm\frac{\sin^{2}\theta}{\tilde{E}^{2}}\sqrt{\left(1+\frac{3(1+2\bar{n})}{2}\eta^{2}\sin^{2}\theta\right)^{2}+\frac{\bar{n}}{2}\eta^{2}\cos^{2}\theta} (90)
\displaystyle\approx ±sin2θE~2(1+η2sin2θ(32+3n¯+n¯4cot2θ)).\displaystyle\pm\frac{\sin^{2}\theta}{\tilde{E}^{2}}\left(1+\eta^{2}\sin^{2}\theta(\frac{3}{2}+3\bar{n}+\frac{\bar{n}}{4}\cot^{2}\theta)\right)\,.

In this perturbative scheme, we can identify the correction corresponding to each unperturbed eigenvector. Consider, for example, the positive eigenvalue an¯,+1(1)a^{(1)}_{\bar{n},+1} of Vn¯V^{\bar{n}}, for which we have the equation

[(1+3(1+2n¯)2η2sin2θ)iηn¯2cosθiηn¯2cosθ(1+3(1+2n¯)2η2sin2θ)](C+1n¯C1n¯)\displaystyle\left[\begin{array}[]{cc}(1+\frac{3(1+2\bar{n})}{2}\eta^{2}\sin^{2}\theta)&i\eta\sqrt{\frac{\bar{n}}{2}}\cos\theta\\ -i\eta\sqrt{\frac{\bar{n}}{2}}\cos\theta&-(1+\frac{3(1+2\bar{n})}{2}\eta^{2}\sin^{2}\theta)\end{array}\right]\left(\begin{array}[]{c}C^{\bar{n}}_{+1}\\ C^{\bar{n}}_{-1}\end{array}\right) (95)
=(1+η2sin2θ(32+3n¯+n¯4cot2θ))(C+1n¯C1n¯).\displaystyle=\left(1+\eta^{2}\sin^{2}\theta(\frac{3}{2}+3\bar{n}+\frac{\bar{n}}{4}\cot^{2}\theta)\right)\left(\begin{array}[]{c}C^{\bar{n}}_{+1}\\ C^{\bar{n}}_{-1}\end{array}\right). (98)

Here, applying the L2L^{2}-normalization |C+1n¯|2+|C1n¯|2=1|C_{+1}^{\bar{n}}|^{2}+|C_{-1}^{\bar{n}}|^{2}=1 and choosing the arbitrary phase such that C+1n¯C_{+1}^{\bar{n}} is real, we obtain

C+1n¯=(11+n¯8η2cos2θ)1/2,\displaystyle C_{+1}^{\bar{n}}=\left(\frac{1}{1+\frac{\bar{n}}{8}\,\eta^{2}\cos^{2}\theta}\right)^{1/2}, (100)
C1n¯=iηcosθ(n¯81+n¯8η2cos2θ)1/2,\displaystyle C_{-1}^{\bar{n}}=-i\eta\cos\theta\left(\frac{\frac{\bar{n}}{8}}{1+\frac{\bar{n}}{8}\,\eta^{2}\cos^{2}\theta}\right)^{1/2},

implying |C1n¯|<<|C+1n¯||C_{-1}^{\bar{n}}|<<|C_{+1}^{\bar{n}}|. Therefore, we can identify the positive eigenvalue correction of Vn¯V^{\bar{n}} as corresponding to the unperturbed |𝒇(0),n¯,+1\ket{\bm{f}^{(0)},\bar{n},+1}, while the negative one will correspond to the unperturbed |𝒇(0),n¯,1\ket{\bm{f}^{(0)},\bar{n},-1}.
Thus, going back from the n¯\bar{n}-notation to the nn-notation, we can write the eigenvalue’s first-order corrections

{an,+1(1)=sin2θE~2(1+η2sin2θ(32+3n+n4cot2θ)),n>0an,1(1)=sin2θE~2(1+η2sin2θ(92+3n+n+14cot2θ)),n0\displaystyle\left\{\begin{array}[]{ll}a^{(1)}_{n,+1}=\frac{\sin^{2}\theta}{\tilde{E}^{2}}\left(1+\eta^{2}\sin^{2}\theta(\frac{3}{2}+3n+\frac{n}{4}\cot^{2}\theta)\right)\,,\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;n>0\\ \\ a^{(1)}_{n,-1}=-\frac{\sin^{2}\theta}{\tilde{E}^{2}}\left(1+\eta^{2}\sin^{2}\theta(\frac{9}{2}+3n+\frac{n+1}{4}\cot^{2}\theta)\right)\,,\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;n\geq 0\end{array}\right. (106)
(107)

thus implying, from Eq. (107) and Eq. (84), that at first order in λ\lambda we have

an,α=2n\displaystyle a_{n,\alpha}=2n +1α+λan,α(1),\displaystyle+1-\alpha+\lambda a^{(1)}_{n,\alpha}\,, (108)

with

an,α(1)\displaystyle a^{(1)}_{n,\alpha} =\displaystyle= αsin2θE~2(1+3(n+12+δα,1)η2sin2θ\displaystyle\alpha\frac{\sin^{2}\theta}{\tilde{E}^{2}}\left(1+3(n+\frac{1}{2}+\delta_{\alpha,-1})\eta^{2}\sin^{2}\theta\right. (109)
+n+δα,14η2cos2θ).\displaystyle\left.+\frac{n+\delta_{\alpha,-1}}{4}\eta^{2}\cos^{2}\theta\right)\,.

Here α=±1\alpha=\pm 1, with δα,1\delta_{\alpha,-1} being the Kronecker delta.

III.3 Perturbative corrections to the energy spectrum

Remembering the definition Eq. (III.1) and the above Eq. (108), we can solve for the corrected energy eigenvalues, to find

En,α2(k3)=\displaystyle E^{2}_{n,\alpha}(k^{3})= mf2+(k3)2+eB(2n+1α+M22an,α(1))\displaystyle\,m^{2}_{f}+(k^{3})^{2}+eB\left(2n+1-\alpha+\frac{M^{2}}{2}a^{(1)}_{n,\alpha}\right)
=\displaystyle= En,α(0)2(k3)+eBM22an,α(1),\displaystyle\,{{E^{(0)}_{n,\alpha}}^{2}(k^{3})}+\frac{eBM^{2}}{2}a^{(1)}_{n,\alpha}\,, (110)

where mfm_{f} was defined in Eq. (37). Substituting Eq.(109), and consistently retaining terms up to order η2\eta^{2}, we obtain the modified energy eigenvalues for the VSR system in the general configuration 𝐁𝐧^=Bcosθ\mathbf{B}\cdot{\mathbf{\hat{n}}}=B\cos\theta

En,α±(k3)\displaystyle E^{\pm}_{n,\alpha}(k^{3}) =\displaystyle= ±[En,α(0)2(k3)+M22αeBsin2θ(En,α(0)k3cosθ)2\displaystyle\pm\left[{{E^{(0)}_{n,\alpha}}^{2}(k^{3})}+\frac{M^{2}}{2}\frac{\alpha eB\sin^{2}\theta}{({{E^{(0)}_{n,\alpha}}}-k^{3}\cos\theta)^{2}}\right. (111)
×(1+3(n+12+δα,1)eBsin2θ(En,α(0)k3cosθ)2\displaystyle\left.\times\left(1+3(n+\frac{1}{2}+\delta_{\alpha,-1})\frac{eB\sin^{2}\theta}{({{E^{(0)}_{n,\alpha}}}-k^{3}\cos\theta)^{2}}\right.\right.
+n+δα,14eBcos2θ(En,α(0)k3cosθ)2)]12.\displaystyle\left.\left.+\frac{n+\delta_{\alpha,-1}}{4}\frac{eB\cos^{2}\theta}{({{E^{(0)}_{n,\alpha}}}-k^{3}\cos\theta)^{2}}\right)\right]^{\frac{1}{2}}.

IV Gyromagnetic Factor and Penning Traps

In this section we discuss the above results in the context of Penning traps. For this purpose, notation and results from Koch et al. (2022a); Hanneke et al. (2011) will be used. Furthermore, while our work until now has been independent of the fermionic or leptonic family taken into account, in the following analysis we will refer to electrons.
Since the first experimental observation of the anomalous magnetic moment of the electron g2g\neq 2 Foley and Kusch (1948); Kusch and Foley (1948), the measurements have been continuously improved. The currently most precise direct measurement of the electron’s gg factor has 13 significant digits Hanneke et al. (2011, 2008); Gabrielse et al. (2006). These experiments are based on Penning traps. The motion of an electron in a Penning trap has four eigenfrequencies, known as the spin-, cyclotron-, axial-, and magnetron-frequency. These four frequencies can be combined in a suitable ratio to extract an experimental value for the gyromagnetic factor of the electron, as explained in Eq. (10)(10) from Hanneke et al. (2011). However, for the calculations in this paper we will consider a simplified setup without electric fields, magnetron nor cavity shifts effects.

IV.1 The g-factor analysis without VSR

For small values of the magnetic field, defining the “free-space” cyclotron frequency νc=eB2πm\nu_{c}=\frac{eB}{2\pi m} Hanneke et al. (2011), the energy-eigenvalues of the unperturbed Dirac-system are given by Brown and Gabrielse (1986)

En,±1(u)\displaystyle E^{(u)}_{n,\pm 1} \displaystyle\simeq 12(2n+1±1)hνc±a2hνc\displaystyle\frac{1}{2}\left(2n+1\pm 1\right)h\nu_{c}\pm\frac{a}{2}h\nu_{c}- (112)
18h2νc2mc2(2n+1±1)2.\displaystyle-\frac{1}{8}\frac{h^{2}\nu_{c}^{2}}{mc^{2}}\left(2n+1\pm 1\right)^{2}.

Here, the term a(g2)/2a\equiv(g-2)/2 with gg being the anomalous magnetic moment, arises from adding an additional a2σμνFμν\frac{a}{2}\sigma_{\mu\nu}F^{\mu\nu} term to the equations of motion, which we can think of as a perturbation arising from Quantum Field Theory (QFT) loop corrections. Higher orders in |𝑩||\bm{B}| are experimentally not relevant Koch et al. (2022a), since the magnetic field strength, used in Hanneke et al. (2011, 2008); Gabrielse et al. (2006), and which can be calculated from the measured cyclotron frequency of about νc149\nu_{c}\approx 149~{}GHz Hanneke et al. (2011), is too “weak”

|𝑩|=2πνcme5.3Tϵ=eBm2109,|\bm{B}|=\frac{2\pi\nu_{c}m}{e}\approx 5.3~{}T\to\epsilon=\frac{eB}{m^{2}}\sim 10^{-9}\,, (113)

where ee is the absolute value of the electron charge.
With our assumptions, the expression of aa only depends upon the anomaly frequency νa\nu_{a} and the relativistic cyclotron frequency fcf_{c} Hanneke et al. (2011), which in our notation correspond respectively to the transition energies E0,1(u)E1,+1(u)E^{(u)}_{0,-1}-E^{(u)}_{1,+1} and E1,1(u)E0,1(u)E^{(u)}_{1,-1}-E^{(u)}_{0,-1} (see also Fig.33 in Hanneke et al. (2011)). In fact, from (112), we directly have

E0,1(u)E1,+1(u)E1,1(u)E0,1(u)+32mϵ2ag22.\frac{E_{0,-1}^{(u)}-E_{1,+1}^{(u)}}{E_{1,-1}^{(u)}-E_{0,-1}^{(u)}+\frac{3}{2}m\epsilon^{2}}\rightarrow a\equiv\frac{g-2}{2}\,. (114)

Therefore, the idea is now to see what happens when considering the new VSR energy spectrum (111), which, along with the mass shift mmfm\to m_{f}, will introduce corrections to the value of the transitions energies in the ratio (114), giving

aVSRgVSR22E0,1E1,+1E1,1E0,1+32mfϵ2,a_{VSR}\equiv\frac{g_{VSR}-2}{2}\equiv\frac{E_{0,-1}-E_{1,+1}}{E_{1,-1}-E_{0,-1}+\frac{3}{2}m_{f}\epsilon^{2}}\,, (115)

so that, despite all the assumptions made, if VSR is correct, measuring the ratio in (115) would give aVSRaa_{VSR}\neq a already in this ideal and simplified experimental setup.

IV.2 Energy spectrum’s expansion for weak magnetic field

Let’s define the new perturbative parameters μ=M2/mf21\mu=M^{2}/m_{f}^{2}\ll 1 and ϵ=eB/mf21\epsilon=eB/m^{2}_{f}\ll 1. Starting from expression (111) in the particle’s rest frame, where we can neglect its momentum, we obtain

En,α\displaystyle E_{n,\alpha} =\displaystyle= mf[1+ϵ(2n+1α)+α2μϵsin2θ1+ϵ(2n+1α)+\displaystyle m_{f}\bigg{[}1+\epsilon(2n+1-\alpha)+\frac{\alpha}{2}\frac{\mu\epsilon\sin^{2}\theta}{1+\epsilon(2n+1-\alpha)}+
+32α(n+12+δα,1)μϵ2sin4θ(1+ϵ(2n+1α))2+\displaystyle\;\;\;\;\;\;\;\;+\frac{3}{2}\alpha(n+\frac{1}{2}+\delta_{\alpha,-1})\frac{\mu\epsilon^{2}\sin^{4}\theta}{(1+\epsilon(2n+1-\alpha))^{2}}+
+α(n+δα,1)8μϵ2sin2θcos2θ(1+ϵ(2n+1α))2]12\displaystyle\;\;\;\;\;\;\;+\frac{\alpha(n+\delta_{\alpha,-1})}{8}\frac{\mu\epsilon^{2}\sin^{2}\theta\cos^{2}\theta}{(1+\epsilon(2n+1-\alpha))^{2}}\bigg{]}^{\frac{1}{2}}
=\displaystyle= mf(1+ϵ(2n+1α))12×\displaystyle m_{f}(1+\epsilon(2n+1-\alpha))^{\frac{1}{2}}\times
×[1+α2μϵsin2θ(1+ϵ(2n+1α))2+\displaystyle\;\;\;\;\;\;\times\bigg{[}1+\frac{\alpha}{2}\frac{\mu\epsilon\sin^{2}\theta}{(1+\epsilon(2n+1-\alpha))^{2}}+
+32α(n+12+δα,1)μϵ2sin4θ(1+ϵ(2n+1α))3+\displaystyle\;\;\;\;\;\;\;+\frac{3}{2}\alpha(n+\frac{1}{2}+\delta_{\alpha,-1})\frac{\mu\epsilon^{2}\sin^{4}\theta}{(1+\epsilon(2n+1-\alpha))^{3}}+
+α(n+δα,1)8μϵ2sin2θcos2θ(1+ϵ(2n+1α))3]12,\displaystyle\;\;\;\;\;\;\;+\frac{\alpha(n+\delta_{\alpha,-1})}{8}\frac{\mu\epsilon^{2}\sin^{2}\theta\cos^{2}\theta}{(1+\epsilon(2n+1-\alpha))^{3}}\bigg{]}^{\frac{1}{2}}\,,

from which we expand in the parameter μ\mu to first order

En,α\displaystyle E_{n,\alpha} =\displaystyle= mf(1+ϵ(2n+1α))12×\displaystyle m_{f}(1+\epsilon(2n+1-\alpha))^{\frac{1}{2}}\times
×[1+α4μϵsin2θ(1+ϵ(2n+1α))2\displaystyle\;\;\;\;\;\;\times\bigg{[}1+\frac{\alpha}{4}\frac{\mu\epsilon\sin^{2}\theta}{(1+\epsilon(2n+1-\alpha))^{2}}
+34α(n+12+δα,1)μϵ2sin4θ(1+ϵ(2n+1α))3+\displaystyle\;\;\;\;\;\;\;+\frac{3}{4}\alpha(n+\frac{1}{2}+\delta_{\alpha,-1})\frac{\mu\epsilon^{2}\sin^{4}\theta}{(1+\epsilon(2n+1-\alpha))^{3}}+
+α(n+δα,1)16μϵ2sin2θcos2θ(1+ϵ(2n+1α))3],\displaystyle\;\;\;\;\;\;\;+\frac{\alpha(n+\delta_{\alpha,-1})}{16}\frac{\mu\epsilon^{2}\sin^{2}\theta\cos^{2}\theta}{(1+\epsilon(2n+1-\alpha))^{3}}\bigg{]}\,,

and now in the parameter ϵ\epsilon to second order

En,αmf=\displaystyle\frac{E_{n,\alpha}}{m_{f}}=  1+ϵ2(2n+1α)ϵ28(2n+1α)2+\displaystyle\,1+\frac{\epsilon}{2}(2n+1-\alpha)-\frac{\epsilon^{2}}{8}(2n+1-\alpha)^{2}+
+α4μϵsin2θ(132ϵ(2n+1α)+\displaystyle+\frac{\alpha}{4}\mu\epsilon\sin^{2}\theta\,\bigg{(}1-\frac{3}{2}\epsilon(2n+1-\alpha)+
+3ϵ(n+12+δα,1)sin2θ+ϵ4(n+δα,1)cos2θ).\displaystyle+3\epsilon(n+\frac{1}{2}+\delta_{\alpha,-1})\sin^{2}\theta+\frac{\epsilon}{4}(n+\delta_{\alpha,-1})\cos^{2}\theta\bigg{)}\,.

Adding usual radiative corrections from QFT as a perturbative term a2σμνFμν\frac{a}{2}\sigma_{\mu\nu}F^{\mu\nu} to the VSR Dirac equation, we obtain an additional αaϵmf/2-\alpha a\epsilon\,m_{f}/2 term in the energy spectrum Koch et al. (2022b), implying

En,αmf=\displaystyle\frac{E_{n,\alpha}}{m_{f}}= 1+ϵ2(2n+1α)+ϵ4μαsin2θϵ2αa\displaystyle 1+\frac{\epsilon}{2}(2n+1-\alpha)+\frac{\epsilon}{4}\mu\alpha\sin^{2}\theta-\frac{\epsilon}{2}\alpha a
ϵ28(2n+1α)238ϵ2αμsin2θ(2n+1α)\displaystyle-\frac{\epsilon^{2}}{8}(2n+1-\alpha)^{2}-\frac{3}{8}\epsilon^{2}\alpha\mu\sin^{2}\theta\,(2n+1-\alpha)
+34ϵ2αμ(n+12+δα,1)sin4θ\displaystyle+\frac{3}{4}\epsilon^{2}\alpha\mu(n+\frac{1}{2}+\delta_{\alpha,-1})\sin^{4}\theta
+ϵ216αμ(n+δα,1)sin2θcos2θ.\displaystyle+\frac{\epsilon^{2}}{16}\alpha\mu(n+\delta_{\alpha,-1})\sin^{2}\theta\cos^{2}\theta\,.

IV.3 Gyromagnetic factor’s corrections in VSR

As mentioned above, the ratio of energy differences in Eq. (114) is particularly useful since in the unperturbed scheme, starting from (112), it gives exactly the theoretical parameter aa.
Our goal is to calculate the expression in (115) for our VSR approximated energy spectrum (IV.2), to see if it can produce deviations from the unperturbed theoretical value aa, derived from Lorentz invariant QFT. Keeping terms up to first order in μ\mu and second order in ϵ\epsilon, we obtain

aVSRa\displaystyle a_{VSR}-a =\displaystyle= μ2[13ϵ+92ϵsin2θ+ϵ4cos2θ+\displaystyle-\frac{\mu}{2}\left[1-3\epsilon+\frac{9}{2}\epsilon\sin^{2}\theta+\frac{\epsilon}{4}\cos^{2}\theta+\right.
+a(32ϵ32ϵsin2θ18ϵcos2θ)]sin2θ,\displaystyle\;\;\;\;\;\;\;\;\left.+a\left(\frac{3}{2}\epsilon-\frac{3}{2}\epsilon\sin^{2}\theta-\frac{1}{8}\epsilon\cos^{2}\theta\right)\right]\sin^{2}\theta\,,

from which we see that the discrepancy between the VSR value gVSRg_{VSR} and the QFT’s one gg would be

gVSRg=μ[1118(23411sin2θacos2θ)ϵ]sin2θ.g_{VSR}-g=-\mu\left[1-\frac{11}{8}(2-\frac{34}{11}\sin^{2}\theta-a\cos^{2}\theta)\epsilon\right]\sin^{2}\theta\,. (121)

Being ϵ\epsilon already pretty small (113), in the following we will consider the relation at order ϵ0\epsilon^{0}

gVSRgμsin2θ.g_{VSR}-g\sim-\mu\sin^{2}\theta\,. (122)

IV.4 Estimation of an upper limit for μ\mu

Here, using (122), we want to put experimental bounds on the magnitude of the VSR perturbative parameter μ\mu and, consequently, the electronic VSR mass term MM.
Obviously, due to our assumptions, we cannot compare directly gVSRg_{VSR} with the experimental value of the gyromagnetic factor gexpg_{exp} measured in a non-ideal Penning Trap, as done in Hanneke et al. (2011, 2008). Nevertheless, one should not doubt that, if VSR is correct, the modification found in (122) would be hidden in the value of gexpg_{exp}, probably together with other terms deriving from the analysis in the complete scheme. Anyway, since our aim is just to give an upper bound for μ\mu, here we will make the further assumption that the whole discrepancy between the experimental gexpg_{exp} and theoretical value gg would be due to the VSR perturbation (122), i.e.

gexpgμsin2θ.g_{exp}-g\sim-\mu\sin^{2}\theta\,. (123)

Using the most precise so far current experimental value for gexp/2=1.00115965218073(28)g_{exp}/2=1.00115965218073(28) Tiesinga et al. (2018); Hanneke et al. (2011, 2008) and as the theoretical prediction g/2=1.001159652182032(720)g/2=1.001159652182032(720) Aoyama et al. (2018), we see that Eq. (123), first of all, is consistent with the electron g-factor discrepancy since gexpg<0g_{exp}-g<0.
Considering again the above-cited values for gg and gexpg_{exp}, and observing that clearly sin2θ1\sin^{2}\theta\leq 1, directly gives the following restriction for the μ\mu parameter

μ3×1012,\mu\lesssim 3\times 10^{-12}\,, (124)

that is comparable or even stronger than other upper bounds found in literature Dunn and Mehen (2006); Maluf et al. (2014). The only more stringent estimation would be μ<9.7×1019\mu<9.7\times 10^{-19} obtained in Alfaro (2019), which refers to electric dipole-like interaction terms 𝒏×𝑬𝝈\bm{n}\times\bm{E}\cdot\bm{\sigma}, where 𝑬\bm{E} represent an electric field here. Nevertheless, as already stated in Dunn and Mehen (2006), experiments in which the electric and magnetic fields are parallel, like the one collaboration et al. (2014) used in Alfaro (2019) to give the above upper bound, are insensitive to interaction terms of that type. Therefore, other experiments with no-parallel magnetic and electric field, should be used in this case to give a coherent upper limit from electric dipole-like VSR interactions.
At this point, by using for the electron an approximated mass of mf0.51MeVm_{f}\sim 0.51\,MeV, we can traslate the restriction in Eq. (124) to the following rough upper bound for the electron’s VSR parameter

M<1013mf21eV,M<\sqrt{10^{-13}m_{f}^{2}}\approx 1\,eV\,, (125)

Observe that, the electron’s VSR parameter MM and the electronic neutrino’s VSR parameter MνM_{\nu} in a C-symmetric VSR Dirac equation (2) would come from the same VSR parameter included in the VSR extension of the Standard Model for each leptonic family Dunn and Mehen (2006); Alfaro (2019), meaning Mν=MM_{\nu}=M. Therefore, we would also have an upper limit for the VSR electronic neutrino’s mass Mν1eVM_{\nu}\lesssim 1\,eV, which is interestingly similar to the upper bound actually known for the electronic neutrino neu (2022), and leaves open the possibility for VSR to be the mechanism or one of the mechanisms giving mass to neutrinos.
Lastly, let’s spend some words on the nature of the VSR θ\theta angle, which we have not discussed so far. In general, we can think of two different scenarios:

  • The VSR spatial preferred direction 𝐧^\mathbf{\hat{n}} represents some universal background effective property Ilderton (2016). Thus, we should take into account the orientation changes of the vector 𝒏\bm{n} respect to 𝑩\bm{B} due to Earth’s motions in the Universe. For example, if the experiment’s duration is of order days\sim days, we should consider the Earth’s rotation movement by averaging in θ(0,π)\theta\in(0,\pi), for which we would obtain the replacement sin2θO(1)<1\sin^{2}\theta\to O(1)<1.

  • If the VSR special four-vector nμn_{\mu} is a dynamic ingredient of some more fundamental theory of nature, therefore an extension of VSR including, for example, gravitational and non-inertial effects may be needed to tackle the time evolution problem of the orientation between vector 𝐧^\mathbf{\hat{n}} and the magnetic field 𝑩\bm{B} accurately.

Obviously, depending on the choosen scenarios different consequencies could arise. However, since in both cases the angular effects will introduce a correction factor of O(1)O(1), our rough upper limit’s estimation (124) for the VSR parameter would remain valid anyway. We therefore leave a more exhaustive and precise analysis of the θ\theta-dependence, like the one done in Kostelecký and Lane (1999), for a future work.

V Conclusions

In this work we have analyzed, in the framework of VSR, the effect of an homogeneous magnetic field 𝑩\bm{B} on the energy spectrum of a Dirac Fermion. First, we found the exact solution to the energy eigenvalues problem in the case of 𝑩𝐧^\bm{B}\parallel\mathbf{\hat{n}}. After that, relaxing this parallelism condition, we found the expression for the energy spectrum at first order in the VSR parameter μ\mu. By considering usual radiative QFT’s corrections, we encountered an expression for the experimental parameter aexpa_{exp} in our scheme, which, confronted with the theoretical one, gave us a theory-experiment discrepancy proportional to μ=M2/mf2\mu=M^{2}/m_{f}^{2}. Therefore, from the current precision in Penning trap experiments that measure the electronic gg-factor Hanneke et al. (2011, 2008); Gabrielse et al. (2006), we found an upper bound for the VSR electron’s mass M1eVM\lesssim 1\,eV.
If VSR is realized in nature as modeled in Dunn and Mehen (2006); Alfaro (2019), the VSR parameter MM would affect leptonic doublets as a whole, generating in particular a mass Mν=MM_{\nu}=M for the electronic neutrino. The upper bound (125) we found, then, is compatible with the present acceptable mass range for the electronic neutrino neu (2022).
In conclusion, we note that this work could actually also apply to the case of the muonic gyromagnetic factor, which recently is getting attention due to the increasing discrepancy between the theoretical and experimental values of its gg-factor Abi et al. (2021). However, since the experiments that measures the muonic gg-factor use a semi-classical approach Keshavarzi et al. (2022), a link between our scheme and the experimental one must still be found, to make a coherent confront. We plan to include this analysis in a future work.

Acknowledgements.
A.S. acknowledges financial support from ANID Fellowship CONICYT-PFCHA/DoctoradoNacional/2020-21201387. E.M. acknowledges financial support from Fondecyt Grant No 1190361 and ANID PIA Anillo ACT192023.

Appendix A Calculation of the Equation for φ(x1)\varphi(x^{1})

For the sake of notation’s simplicity, we define

D=(np)1=(Ek3cosθp1sinθ)1,{D}=(n\cdot p)^{-1}=(E-k^{3}\cos\theta-p^{1}\sin\theta)^{-1}\,, (126)

to be used in this appendix. Let‘s expand the second squared bracket in (56), keeping in mind the relation (57) and the fact that mixed σ\sigma’s products which do not involve operators simply get cancelled due to their anticommutation relations

[(EM22D)2m2(p1M2sinθ2D)2(k2eBx1)2(k3M2cosθ2D)2\displaystyle\left[\left(E-\frac{M^{2}}{2}D\right)^{2}-m^{2}-(p^{1}-\frac{M^{2}\sin\theta}{2}D)^{2}-(k^{2}-eBx^{1})^{2}-(k^{3}-\frac{M^{2}\cos\theta}{2}D)^{2}\right. (127)
ieBσ1σ2eBM22sinθDx1σ1σ2eBM22sinθx1Dσ2σ1eBM22cosθDx1σ3σ2eBM22cosθx1Dσ2σ3]φ(x1)=0.\displaystyle\left.-ieB\sigma^{1}\sigma^{2}-eB\frac{M^{2}}{2}\sin\theta Dx^{1}\sigma^{1}\sigma^{2}-eB\frac{M^{2}}{2}\sin\theta x^{1}D\sigma^{2}\sigma^{1}-eB\frac{M^{2}}{2}\cos\theta Dx^{1}\sigma^{3}\sigma^{2}-eB\frac{M^{2}}{2}\cos\theta x^{1}D\sigma^{2}\sigma^{3}\right]\varphi(x^{1})=0\,.

Therefore, we now have to calculate D(x1)D(x^{1}). To do that, we will use the integral representation (75), and we define for simplicity E~=Ek3cosθ\tilde{E}=E-k^{3}\cos\theta, implying

D(x1φ)\displaystyle D(x^{1}\varphi) =\displaystyle= 0𝑑texp[t(E~p1sinθ)]x1φ\displaystyle\int_{0}^{\infty}dt\;exp[-t(\tilde{E}-p^{1}\sin\theta)]\,x^{1}\varphi
=\displaystyle= 0𝑑tetE~j=0(tsinθp1)jx1φ,\displaystyle\int_{0}^{\infty}dt\;e^{-t\tilde{E}}\sum_{j=0}^{\infty}(t\sin\theta p^{1})^{j}x^{1}\varphi\,,

which can be expanded as

D(x1φ)\displaystyle D(x^{1}\varphi) =\displaystyle= x10𝑑tetE~j=0(tsinθp1)jφ\displaystyle x^{1}\int_{0}^{\infty}dt\;e^{-t\tilde{E}}\sum_{j=0}^{\infty}(t\sin\theta p^{1})^{j}\,\varphi
isinθ0t𝑑tetE~j=1(tsinθp1)j1φ.\displaystyle-i\sin\theta\int_{0}^{\infty}tdt\;e^{-t\tilde{E}}\sum_{j=1}^{\infty}(t\sin\theta p^{1})^{j-1}\varphi\,.

Sending the index of the second sum to ii1i\rightarrow i-1, and “reversing” the use of the integral representation, we have the relation

D(x1φ)\displaystyle D(x^{1}\varphi) =\displaystyle= x1Dφ+isinθddE~(D)φ,\displaystyle x^{1}D\varphi+i\sin\theta\frac{d}{d\tilde{E}}(D)\varphi\,, (130)

and considering the definition (126)

D(x1φ)\displaystyle D(x^{1}\varphi) =\displaystyle= x1DφisinθD2φ.\displaystyle x^{1}D\varphi-i\sin\theta D^{2}\varphi\,. (131)

Therefore, since the terms x1Dφ(x1)x^{1}D\varphi(x^{1}) always get cancelled from the other terms with inverted σ\sigma product, and since

σiσj=iϵijkσk,\sigma^{i}\sigma^{j}=i\epsilon_{ijk}\sigma^{k}\,, (132)

Eq. (127) then becomes

[(EM22D)2m2(p1M2sinθ2D)2(k2eBx1)2(k3M2cosθ2D)2\displaystyle\left[\left(E-\frac{M^{2}}{2}D\right)^{2}-m^{2}-(p^{1}-\frac{M^{2}\sin\theta}{2}D)^{2}-(k^{2}-eBx^{1})^{2}-(k^{3}-\frac{M^{2}\cos\theta}{2}D)^{2}\right. (133)
+eBσ3eBM22sin2θD2σ3+eBM22sinθcosθD2σ1]φ(x1)=0.\displaystyle\left.+eB\sigma^{3}-eB\frac{M^{2}}{2}\sin^{2}\theta D^{2}\sigma^{3}+eB\frac{M^{2}}{2}\sin\theta\cos\theta D^{2}\sigma^{1}\right]\varphi(x^{1})=0\,.

Finally, expanding the squares involving the terms EE, p1p^{1} and k3k^{3}, and remembering that Nμpμ=Dnμpμ=1N_{\mu}p^{\mu}=D\,n_{\mu}p^{\mu}=1 while nμnμ=0n_{\mu}n^{\mu}=0, we get to Eq. (III.1), that we have already seen in the main text of this article.
Note that Eq. (133) has the correct limits: in fact, sending B0B\to 0, we re-obtain, as expected, the usual VSR dispersion relation

[E2p2m2M2]φ=0,[E^{2}-p^{2}-m^{2}-M^{2}]\,\varphi=0\,, (134)

while, for θ0\theta\to 0 we find again Eq. (23), the equation of motion for the case 𝑩𝐧^\bm{B}\parallel\mathbf{\hat{n}}.

Appendix B Calculation of the integrals I1(n¯,k)I_{1}(\bar{n},k) and I2(n¯,k)I_{2}(\bar{n},k)

In the main text, we defined the integrals

I1(n¯,k)=𝑑ξeξ2/2Hn¯(ξ)ξk(eξ2/2Hn¯(ξ)),\displaystyle I_{1}(\bar{n},k)=\int_{-\infty}^{\infty}d\xi\,e^{-\xi^{2}/2}H_{\bar{n}}(\xi)\partial_{\xi}^{k}\left(e^{-\xi^{2}/2}H_{\bar{n}}(\xi)\right),
(135)

where Hn¯(z)H_{\bar{n}}(z) are the Hermite polynomials, defined by

Hn¯(z)=(1)n¯ez2zn¯(ez2)\displaystyle H_{\bar{n}}(z)=(-1)^{\bar{n}}e^{z^{2}}\partial_{z}^{\bar{n}}\left(e^{-z^{2}}\right) (136)

In particular, we remark that H0(z)=1H_{0}(z)=1, and hence for the particular case n¯=0\bar{n}=0, Eq. (135) reduces to

I1(0,k)\displaystyle I_{1}(0,k) =\displaystyle= 𝑑ξeξ2/2ξk(eξ2/2)\displaystyle\int_{-\infty}^{\infty}d\xi\,e^{-\xi^{2}/2}\partial_{\xi}^{k}\left(e^{-\xi^{2}/2}\right) (137)
=\displaystyle= 2k/2(1)k𝑑ξeξ2Hk(ξ/2)\displaystyle 2^{-k/2}(-1)^{k}\int_{-\infty}^{\infty}d\xi\,e^{-\xi^{2}}\,H_{k}(\xi/\sqrt{2})

From the standard result (see for instance Gradshteyn-Rhizik Gradshteyn and Ryzhik (2014), p. 837)

e(xy)2Hk(αx)𝑑x=\displaystyle\int_{-\infty}^{\infty}e^{-(x-y)^{2}}H_{k}(\alpha x)dx= (138)
=π(1α2)k/2Hk(αy1α2),\displaystyle\;\;\;=\sqrt{\pi}(1-\alpha^{2})^{k/2}H_{k}\left(\frac{\alpha y}{\sqrt{1-\alpha^{2}}}\right),

we obtain (setting α=1/2\alpha=1/\sqrt{2} and y=0y=0)

I1(0,k)=π(1)k2kHk(0).\displaystyle I_{1}(0,k)=\sqrt{\pi}(-1)^{k}2^{-k}H_{k}(0). (140)

Given that H2n+1(0)=0H_{2n+1}(0)=0, while H2n(0)=(1)nΓ(2n+1)/Γ(n+1)H_{2n}(0)=(-1)^{n}\Gamma(2n+1)/\Gamma(n+1), we obtain

I1(0,2k)=π22k(1)kΓ(2k+1)Γ(k+1),\displaystyle I_{1}(0,2k)=\sqrt{\pi}2^{-2k}(-1)^{k}\frac{\Gamma(2k+1)}{\Gamma(k+1)}, (141)
I1(0,2k+1)=0.\displaystyle I_{1}(0,2k+1)=0\,.

Let’s now focus on the case n¯>0\bar{n}>0. For this purpose, we notice that the functions

ξ|φn¯=φn¯(ξ)=12n¯n¯!πeξ2/2Hn¯(ξ)\displaystyle\langle\xi|\varphi_{\bar{n}}\rangle=\varphi_{\bar{n}}(\xi)=\frac{1}{\sqrt{2^{\bar{n}}\bar{n}!\sqrt{\pi}}}e^{-\xi^{2}/2}H_{\bar{n}}(\xi) (142)

constitute a complete orthonormal set (they are just the eigenfunctions of the 1D-harmonic oscilator)

n¯=0|φn¯φn¯|=𝟏\displaystyle\sum_{\bar{n}=0}^{\infty}|\varphi_{\bar{n}}\rangle\langle\varphi_{\bar{n}}|=\mathbf{1} (143)
φn¯|φn¯=𝑑ξφn¯|ξξ|φn¯=δn¯,n¯\displaystyle\langle\varphi_{\bar{n}}|\varphi_{\bar{n}^{\prime}}\rangle=\int_{-\infty}^{\infty}d\xi\langle\varphi_{\bar{n}}|\xi\rangle\langle\xi|\varphi_{\bar{n}^{\prime}}\rangle=\delta_{\bar{n},\bar{n}^{\prime}}

With these definitions, along with the momentum operator pξ=iξp_{\xi}=-i\partial_{\xi}, we have the rather simple expression

I1(n¯,k)=2n¯n¯!π(i)kφn¯|(pξ)k|φn¯\displaystyle I_{1}(\bar{n},k)=2^{\bar{n}}\bar{n}!\sqrt{\pi}(i)^{k}\langle\varphi_{\bar{n}}|(p_{\xi})^{k}|\varphi_{\bar{n}}\rangle (144)

Since the momentum (Fourier) eigenbasis is also complete, we have

𝑑p|pp|=𝟏,ξ|p=12πeiξp.\displaystyle\int_{-\infty}^{\infty}dp|p\rangle\langle p|=\mathbf{1}\,,\,\,\,\langle\xi|p\rangle=\frac{1}{\sqrt{2\pi}}e^{i\xi p}. (145)

Therefore, we can insert the identity in the basis of momentum eigenstates into Eq. (144) to obtain

I1(n¯,k)\displaystyle I_{1}(\bar{n},k) =\displaystyle= 2n¯n¯!π(i)k𝑑pφn¯|(pξ)k|pp|φn¯\displaystyle 2^{\bar{n}}\bar{n}!\sqrt{\pi}(i)^{k}\int_{-\infty}^{\infty}dp\langle\varphi_{\bar{n}}|(p_{\xi})^{k}|p\rangle\langle p|\varphi_{\bar{n}}\rangle (146)
=\displaystyle= 2n¯n¯!π(i)k𝑑ppk|p|φn¯|2\displaystyle 2^{\bar{n}}\bar{n}!\sqrt{\pi}(i)^{k}\int_{-\infty}^{\infty}dp\,p^{k}\left|\langle p|\varphi_{\bar{n}}\rangle\right|^{2}

Finally, notice that the functions p|φn¯\langle p|\varphi_{\bar{n}}\rangle are related to the functions ξ|φ\langle\xi|\varphi\rangle via Fourier representation, since

p|φn¯\displaystyle\langle p|\varphi_{\bar{n}}\rangle =\displaystyle= 𝑑ξp|ξξ|φn¯\displaystyle\int_{-\infty}^{\infty}d\xi\langle p|\xi\rangle\langle\xi|\varphi_{\bar{n}}\rangle (147)
=\displaystyle= 12π𝑑ξeiξpφn¯(ξ)\displaystyle\frac{1}{\sqrt{2\pi}}\int_{-\infty}^{\infty}d\xi\,e^{-i\xi p}\varphi_{\bar{n}}(\xi)
=\displaystyle= 12n¯n¯!π12π𝑑ξeiξpeξ2/2Hn¯(ξ)\displaystyle\frac{1}{\sqrt{2^{\bar{n}}\bar{n}!\sqrt{\pi}}}\frac{1}{\sqrt{2\pi}}\int_{-\infty}^{\infty}d\xi\,e^{-i\xi p}e^{-\xi^{2}/2}H_{\bar{n}}(\xi)
=\displaystyle= (i)n¯2n¯n¯!πep2/2Hn¯(p)\displaystyle\frac{(-i)^{\bar{n}}}{\sqrt{2^{\bar{n}}\bar{n}!\sqrt{\pi}}}e^{-p^{2}/2}H_{\bar{n}}(p)

Substituting this last result into Eq. (146), we have

I1(n¯,2k)=(1)k𝑑pp2kep2(Hn¯(p))2,\displaystyle I_{1}(\bar{n},2k)=(-1)^{k}\int_{-\infty}^{\infty}dp\,p^{2k}e^{-p^{2}}\left(H_{\bar{n}}(p)\right)^{2}, (148)

with I1(n,2k+1)=0I_{1}(n,2k+1)=0 trivially by parity.
Using the Hermite-Fourier series representation for the power p2kp^{2k}, we have

p2k=(2k)!22k=0kH2(p)(2)!(k)!,\displaystyle p^{2k}=\frac{(2k)!}{2^{2k}}\sum_{\ell=0}^{k}\frac{H_{2\ell}(p)}{(2\ell)!(k-\ell)!}, (149)

and inserting this expression into Eq. (148) we obtain

I1(n¯,2k)\displaystyle I_{1}(\bar{n},2k) =\displaystyle= (1)k(2k)!22k=0k1(2)!(k)!\displaystyle(-1)^{k}\frac{(2k)!}{2^{2k}}\sum_{\ell=0}^{k}\frac{1}{(2\ell)!(k-\ell)!}
×dpep2(Hn¯(p))2H2(p)\displaystyle\times\int_{-\infty}^{\infty}dp\,e^{-p^{2}}\left(H_{\bar{n}}(p)\right)^{2}H_{2\ell}(p)

We can evaluate this last integral by applying the identity (see Gradshteyn-Rhizik Gradshteyn and Ryzhik (2014), p.797, for s=(n¯+m+k)/2s=(\bar{n}+m+k)/2)

𝑑zez2Hk(z)Hm(z)Hn¯(z)=2sπk!m!n¯!(sk)!(sm)!(sn¯)!\displaystyle\int_{-\infty}^{\infty}dz\,e^{-z^{2}}H_{k}(z)H_{m}(z)H_{\bar{n}}(z)=\frac{2^{s}\sqrt{\pi}\,k!\,m!\,\bar{n}!}{(s-k)!(s-m)!(s-\bar{n})!}

Applying this result into Eq. (B), we obtain

I1(n¯,2k)\displaystyle I_{1}(\bar{n},2k) =\displaystyle= π(1)k(2k)!n¯! 2n¯2kk!=0k2(n¯)(k)\displaystyle\frac{\sqrt{\pi}(-1)^{k}(2k)!\,\bar{n}!\,2^{\bar{n}-2k}}{k!}\sum_{\ell=0}^{k}2^{\ell}\left(\begin{array}[]{c}\bar{n}\\ \ell\end{array}\right)\left(\begin{array}[]{c}k\\ \ell\end{array}\right)
=\displaystyle= π(1)kn¯! 2n¯2kΓ(2k+1)Γ(k+1)F(k,n¯;1;2)\displaystyle\sqrt{\pi}(-1)^{k}\bar{n}!\,2^{\bar{n}-2k}\frac{\Gamma(2k+1)}{\Gamma(k+1)}F(-k,-\bar{n};1;2)

where F(a,b;c;z)F(a,b;c;z) is the Hypergeometric function.
Furthermore, we want to calculate the integrals of the form

I2(n¯,k)=𝑑ξeξ2/2Hn¯(ξ)ξk(eξ2/2Hn¯1(ξ)).I_{2}(\bar{n},k)=\int_{-\infty}^{\infty}d\xi e^{-\xi^{2}/2}H_{\bar{n}}(\xi)\partial^{k}_{\xi}\left(e^{-\xi^{2}/2}H_{\bar{n}-1}(\xi)\right)\,. (158)

We can undergo all the previous steps done for the integrals I1I_{1} to find

I2(n¯,k)\displaystyle I_{2}(\bar{n},k) =\displaystyle= 2n¯12(n¯1)!πn(i)kφn¯|(pξ)k|φn¯1=\displaystyle 2^{\bar{n}-\frac{1}{2}}(\bar{n}-1)!\sqrt{\pi n}(i)^{k}\langle\varphi_{\bar{n}}|(p_{\xi})^{k}|\varphi_{\bar{n}-1}\rangle= (159)
=\displaystyle= ik+1𝑑ppkep2Hn¯(p)Hn¯1(p),\displaystyle i^{k+1}\int_{-\infty}^{\infty}dp\,p^{k}e^{-p^{2}}H_{\bar{n}}(p)H_{\bar{n}-1}(p)\,,

which, due to the parity of the integrand is easily seen to be 0 for even values of kk: I2(n¯,2k)=0I_{2}(\bar{n},2k)=0. Remembering that for odd powers of momentum pp

p2k+1=(2k+1)!22k+1=0kH2+1(p)(2+1)!(k)!,p^{2k+1}=\frac{(2k+1)!}{2^{2k+1}}\sum_{\ell=0}^{k}\frac{H_{2\ell+1}(p)}{(2\ell+1)!(k-\ell)!}, (160)

we find

I2(n¯,2k+1)=(1)k+1(2k+1)!22k+1×\displaystyle I_{2}(\bar{n},2k+1)=(-1)^{k+1}\frac{(2k+1)!}{2^{2k+1}}\times (161)
×=0k1(2+1)!(k)!dpep2H2l+1Hn¯Hn¯1.\displaystyle\;\;\;\;\;\;\times\sum_{\ell=0}^{k}\frac{1}{(2\ell+1)!(k-\ell)!}\int_{-\infty}^{\infty}dp\,e^{-p^{2}}H_{2l+1}H_{\bar{n}}H_{\bar{n}-1}\,.

Using again the expression (B) for the integral in I2I_{2}, we finally obtain

I2(n¯,2k+1)\displaystyle I_{2}(\bar{n},2k+1) =\displaystyle= π(1)k+1(2k+1)!k!(n¯1)! 2n¯2k1×\displaystyle\sqrt{\pi}(-1)^{k+1}\frac{(2k+1)!}{k!}(\bar{n}-1)!\,2^{\bar{n}-2k-1}\times
×=0k2(k)(n¯+1)=\displaystyle\times\sum_{\ell=0}^{k}2^{\ell}\left(\begin{array}[]{c}k\\ \ell\end{array}\right)\left(\begin{array}[]{c}\bar{n}\\ \ell+1\end{array}\right)=
=\displaystyle= π(1)k+1(n¯1)! 2n¯2k1Γ(2k+2)Γ(k+1)×\displaystyle\sqrt{\pi}(-1)^{k+1}(\bar{n}-1)!\,2^{\bar{n}-2k-1}\frac{\Gamma(2k+2)}{\Gamma(k+1)}\times
×F(k1;n;1;2)F(k;n;1;2)2.\displaystyle\times\frac{F(-k-1;-n;1;2)-F(-k;-n;1;2)}{2}\,.

Appendix C Calculation of Matrix elements Vα,αn¯V^{\bar{n}}_{\alpha,\alpha^{\prime}}

First of all, we observe that the perturbation VV is hermitian and therefore also Vn¯V^{\bar{n}} is hermitian, implying V+1,1n¯=(V1,+1n¯)V^{\bar{n}}_{+1,-1}=(V^{\bar{n}}_{-1,+1})^{*}. Furthermore, since the diagonal elements in VV are equal but opposite in sign, we will have V+1,+1n¯=V1,1n¯V^{\bar{n}}_{+1,+1}=-V^{\bar{n}}_{-1,-1}. Therefore, is sufficient to calculate only two matrix elements: V+1,+1n¯V^{\bar{n}}_{+1,+1} and V+1,1n¯V^{\bar{n}}_{+1,-1}.

C.1 Case V+1,+1n¯V^{\bar{n}}_{+1,+1}

We want to calculate the matrix element

V+1,+1n¯\displaystyle V^{\bar{n}}_{+1,+1} =\displaystyle= 𝒇(0),n¯,+1|V|𝒇(0),n¯,+1.\displaystyle\bra{\bm{f}^{(0)},\bar{n},+1}{V}\ket{\bm{f}^{(0)},\bar{n},+1}\,. (167)

Inserting two identities through the completeness relation 1=𝑑ξ|ξξ|1=\int d\xi\ket{\xi}\bra{\xi} we get

V+1,+1n¯=\displaystyle V^{\bar{n}}_{+1,+1}= sin2θ2n¯n¯!π𝑑ξeξ2/2Hn¯(ξ)1Pξ2(eξ2/2Hn¯(ξ))=\displaystyle\frac{\sin^{2}\theta}{2^{\bar{n}}\bar{n}!\sqrt{\pi}}\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{\bar{n}}(\xi)\frac{1}{P^{2}_{\xi}}(e^{-\xi^{2}/2}H_{\bar{n}}(\xi))= (168)
=sin2θ2n¯n¯!πE~2ddAdξeξ2/2Hn¯(ξ)×\displaystyle=-\frac{\sin^{2}\theta}{2^{\bar{n}}\bar{n}!\sqrt{\pi}\tilde{E}^{2}}\frac{d}{dA}\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{\bar{n}}(\xi)\times
×1A+iηsinθξ(eξ2/2Hn¯(ξ))|A=1,\displaystyle\;\;\;\times\frac{1}{A+i\eta\sin\theta\partial_{\xi}}(e^{-\xi^{2}/2}H_{\bar{n}}(\xi))\bigg{|}_{A=1}\,,

where we have defined the dimensionless quantity η=eB/E~\eta={\sqrt{eB}}/{\tilde{E}}. By representing the inverse operator in (168) with the integral form, we obtain

V+1,+1n¯=\displaystyle V^{\bar{n}}_{+1,+1}= sin2θ2n¯n¯!πE~2ddA0𝑑t𝑑ξeξ2/2Hn¯(ξ)\displaystyle-\frac{\sin^{2}\theta}{2^{\bar{n}}\bar{n}!\sqrt{\pi}\tilde{E}^{2}}\frac{d}{dA}\int_{0}^{\infty}dt\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{\bar{n}}(\xi)
×eAt(1+iηsinθξ)(eξ2/2Hn(ξ))|A=1=\displaystyle\;\;\;\times e^{-At(1+i\eta\sin\theta\partial_{\xi})}(e^{-\xi^{2}/2}H_{n}(\xi))\bigg{|}_{A=1}=
=\displaystyle= sin2θ2n¯n¯!πE~2ddA0eAt𝑑t𝑑ξeξ2/2Hn¯(ξ)\displaystyle-\frac{\sin^{2}\theta}{2^{\bar{n}}\bar{n}!\sqrt{\pi}\tilde{E}^{2}}\frac{d}{dA}\int_{0}^{\infty}e^{-At}dt\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{\bar{n}}(\xi) (169)
×eitηsinθξ(eξ2/2Hn¯(ξ))|A=1,\displaystyle\;\;\;\times e^{-it\eta\sin\theta\partial_{\xi}}(e^{-\xi^{2}/2}H_{\bar{n}}(\xi))\bigg{|}_{A=1}\,,

and expanding the exponential operator

V+1,+1n¯=\displaystyle V^{\bar{n}}_{+1,+1}= sin2θ2n¯n¯!πE~2k=0(iηsinθ)kk!ddA0eAttk𝑑t\displaystyle-\frac{\sin^{2}\theta}{2^{\bar{n}}\bar{n}!\sqrt{\pi}\tilde{E}^{2}}\sum_{k=0}^{\infty}\frac{(-i\eta\sin\theta)^{k}}{k!}\frac{d}{dA}\int_{0}^{\infty}e^{-At}t^{k}dt
×dξeξ2/2Hn¯(ξ)ξk(eξ2/2Hn¯(ξ))|A=1=\displaystyle\times\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{\bar{n}}(\xi)\partial_{\xi}^{k}(e^{-\xi^{2}/2}H_{\bar{n}}(\xi))\bigg{|}_{A=1}=
=\displaystyle= sin2θ2n¯n¯!πE~2k=0(k+1)(iηsinθ)kI1(n¯,k),\displaystyle\frac{\sin^{2}\theta}{2^{\bar{n}}\bar{n}!\sqrt{\pi}\tilde{E}^{2}}\sum_{k=0}^{\infty}(k+1)(i\eta\sin\theta)^{k}I_{1}(\bar{n},k)\,, (170)

where we introduced the integral I1(n¯,k)I_{1}(\bar{n},k) defined in (78), which vanish for odd values of k. Thus, we can re-write equation (C.1) as

V+1,+1n¯=sin2θ2n¯n¯!πE~2k=0(2k+1)(1)k(ηsinθ)2kI1(n¯,2k),V^{\bar{n}}_{+1,+1}=\frac{\sin^{2}\theta}{2^{\bar{n}}\bar{n}!\sqrt{\pi}\tilde{E}^{2}}\sum_{k=0}^{\infty}(2k+1)(-1)^{k}\left(\eta\sin\theta\right)^{2k}I_{1}(\bar{n},2k)\,, (171)

that is a completely real expression as expected.
Therefore the final expression for the V+1,+1(n¯)V^{(\bar{n})}_{+1,+1} matrix element will be

V+1,+1n¯=sin2θE~2k=0(ηsinθ2)2kΓ(2k+2)Γ(k+1)F(k,n¯;1;2),V^{\bar{n}}_{+1,+1}=\frac{\sin^{2}\theta}{\tilde{E}^{2}}\sum_{k=0}^{\infty}\left(\frac{\eta\sin\theta}{2}\right)^{2k}\frac{\Gamma(2k+2)}{\Gamma(k+1)}F(-k,-\bar{n};1;2)\,, (172)

which for weak magnetic field can be expanded as

V+1,+1n¯=sin2θE~2(1+(2n¯+1)32(ηsinθ)2+O(η4)).V^{\bar{n}}_{+1,+1}=\frac{\sin^{2}\theta}{\tilde{E}^{2}}\left(1+(2\bar{n}+1)\,\frac{3}{2}\left(\eta\sin\theta\right)^{2}+O(\eta^{4})\right)\,. (173)

C.2 Case V+1,1n¯V^{\bar{n}}_{+1,-1}

We want to calculate the matrix element

V+1,1n¯\displaystyle V^{\bar{n}}_{+1,-1} =\displaystyle= 𝒇(0),n¯,+1|V|𝒇(0),n¯,1\displaystyle\bra{\bm{f}^{(0)},\bar{n},+1}{V}\ket{\bm{f}^{(0)},\bar{n},-1}
=\displaystyle= sinθcosθπ2n¯12(n¯1)!n¯dξeξ2/2Hn¯(ξ)×\displaystyle\frac{\sin\theta\cos\theta}{\sqrt{\pi}2^{\bar{n}-\frac{1}{2}}(\bar{n}-1)!\sqrt{\bar{n}}}\int_{-\infty}^{\infty}d\xi\;e^{-\xi^{2}/2}H_{\bar{n}}(\xi)\times
×1Pξ2(eξ2/2Hn¯1(ξ)).\displaystyle\times\frac{1}{P^{2}_{\xi}}(e^{-\xi^{2}/2}H_{\bar{n}-1}(\xi))\,.

Following exactly the same procedure as before, we get to the expression

V+1,1n¯\displaystyle V^{\bar{n}}_{+1,-1} =\displaystyle= sinθcosθπ2n¯12(n¯1)!n¯1E~2\displaystyle\frac{\sin\theta\cos\theta}{\sqrt{\pi}2^{\bar{n}-\frac{1}{2}}(\bar{n}-1)!\sqrt{\bar{n}}}\frac{1}{\tilde{E}^{2}} (175)
×k(k+1)(iηsinθ)kI2(n¯,k),\displaystyle\times\sum_{k}(k+1)(-i\eta\sin\theta)^{k}I_{2}(\bar{n},k)\,,

where we introduced the integrals I2I_{2} defined in (158).
In the end, using Eq. (B) for I2I_{2} and remembering that it is non-zero only for odd values of kk, we have

V+1,1n¯\displaystyle V^{\bar{n}}_{+1,-1} =\displaystyle= i18n¯sinθcosθE~2k=0(ηsinθ2)2k+1Γ(2k+3)Γ(k+1)×\displaystyle i\sqrt{\frac{1}{8\bar{n}}}\frac{\sin\theta\cos\theta}{\tilde{E}^{2}}\sum_{k=0}^{\infty}\left(\frac{\eta\sin\theta}{2}\right)^{2k+1}\frac{\Gamma(2k+3)}{\Gamma(k+1)}\times
×[F(k1;n;1;2)F(k;n;1;2)].\displaystyle\times\left[F(-k-1;-n;1;2)-F(-k;-n;1;2)\right]\,.

For our purposes, when using the weak field approximation η1\eta\ll 1, it will be sufficient to consider only the first term of the kk-series, since the next one would already be order η3\propto\eta^{3}, obtaining

V+1,1n¯\displaystyle V^{\bar{n}}_{+1,-1} \displaystyle\approx iηn¯2sin2θcosθE~2.\displaystyle i\eta\sqrt{\frac{\bar{n}}{2}}\frac{\sin^{2}\theta\cos\theta}{\tilde{E}^{2}}\,. (177)

Appendix D Borel regularization

In this section, we show in detail the procedure to obtain a Borel regularization for the infinite series Eq. (81) defined in the main text

a0¯,+1(1)\displaystyle a^{(1)}_{\bar{0},+1} =\displaystyle= sin2θπE~2k=0(2k+1)(1)k(ηsinθ)2kI1(0,2k)\displaystyle\frac{\sin^{2}\theta}{\sqrt{\pi}\tilde{E}^{2}}\sum_{k=0}^{\infty}(2k+1)(-1)^{k}\left(\eta\sin\theta\right)^{2k}I_{1}(0,2k) (178)
=\displaystyle= sin2θE~2k=0(ηsinθ2)2kΓ(2k+2)Γ(k+1).\displaystyle\frac{\sin^{2}\theta}{\tilde{E}^{2}}\sum_{k=0}^{\infty}\left(\frac{\eta\sin\theta}{2}\right)^{2k}\frac{\Gamma(2k+2)}{\Gamma(k+1)}\,.

First, notice that for kk a positive integer, the ratio

Γ(2k+2)Γ(k+1)=(2k+1)!k!,\displaystyle\frac{\Gamma(2k+2)}{\Gamma(k+1)}=\frac{(2k+1)!}{k!}\,, (179)

and let us consider, with a simplified notation z=(ηsinθ/2)2z=\left(\eta\sin\theta/2\right)^{2}, the equivalent series

A(z)=k=0zk(2k+1)!k!,\displaystyle A(z)=\sum_{k=0}^{\infty}z^{k}\frac{(2k+1)!}{k!}\,, (180)

where clearly Eq.(178) corresponds to

a0¯,+1(1)=sin2θE~2A((ηsinθ/2)2).\displaystyle a^{(1)}_{\bar{0},+1}=\frac{\sin^{2}\theta}{\tilde{E}^{2}}A(\left(\eta\sin\theta/2\right)^{2})\,. (181)

The definition of the Borel transform of this series leads to the expression

BA(zt)\displaystyle BA(zt) =\displaystyle= k=0(zt)kk!(2k+1)!k!\displaystyle\sum_{k=0}^{\infty}\frac{(zt)^{k}}{k!}\frac{(2k+1)!}{k!} (182)
=\displaystyle= 1(14zt)3/2,\displaystyle\frac{1}{\left(1-4zt\right)^{3/2}}\,,

where the second result follows as an identity directly from the Taylor expansion. As the following step, we recover the regularized expression for the series A~(z)A(z)\tilde{A}(z)\simeq A(z) by performing the integral transform

A~(z)\displaystyle\tilde{A}(z) =\displaystyle= 0etBA(zt)𝑑t\displaystyle\int_{0}^{\infty}e^{-t}BA(zt)\,dt (183)
=\displaystyle= 0et(14zt)3/2𝑑t.\displaystyle\int_{0}^{\infty}\frac{e^{-t}}{\left(1-4zt\right)^{3/2}}\,dt\,.

In order to evaluate this integral, we perform the change of variables u=t14zu=t-\frac{1}{4z}, that implies 1/(4z)u<-1/(4z)\leq u<\infty, and du=dtdu=dt. Therefore, we have

A~(z)\displaystyle\tilde{A}(z) =\displaystyle= e14z(4z)3/214zu3/2eu𝑑u\displaystyle\frac{e^{\frac{-1}{4z}}}{(-4z)^{3/2}}\int_{-\frac{1}{4z}}^{\infty}u^{-3/2}e^{-u}du (184)
=\displaystyle= e14z(4z)3/2Γ(12,14z),\displaystyle\frac{e^{\frac{-1}{4z}}}{(-4z)^{3/2}}\Gamma\left(-\frac{1}{2},-\frac{1}{4z}\right),

where in the final result we applied the definition of the Incomplete Gamma function

Γ(s,x)=xus1eu𝑑u.\displaystyle\Gamma(s,x)=\int_{x}^{\infty}u^{s-1}e^{-u}du\,. (185)

References

  • Cohen and Glashow (2006a) A. G. Cohen and S. L. Glashow, Physical review letters 97, 021601 (2006a).
  • Cohen and Glashow (2006b) A. G. Cohen and S. L. Glashow, arXiv preprint hep-ph/0605036  (2006b).
  • Cohen and Freedman (2007) A. G. Cohen and D. Z. Freedman, Journal of High Energy Physics 2007, 039 (2007).
  • Vohánka (2012) J. Vohánka, Physical Review D 85, 105009 (2012).
  • Gibbons et al. (2007) G. Gibbons, J. Gomis,  and C. Pope, Physical Review D 76, 081701 (2007).
  • Kouretsis et al. (2009) A. Kouretsis, M. Stathakopoulos,  and P. Stavrinos, Physical Review D 79, 104011 (2009).
  • Alfaro and Santoni (2022) J. Alfaro and A. Santoni, Physics Letters B 829, 137080 (2022).
  • Alfaro and Soto (2019) J. Alfaro and A. Soto, Physical Review D 100, 055029 (2019).
  • Cheon et al. (2009) S. Cheon, C. Lee,  and S. J. Lee, Physics Letters B 679, 73 (2009).
  • Dunn and Mehen (2006) A. Dunn and T. Mehen, arXiv preprint hep-ph/0610202  (2006).
  • Alfaro (2019) J. Alfaro, Universe 5, 16 (2019).
  • Colladay and Kosteleckỳ (1998) D. Colladay and V. A. Kosteleckỳ, Physical Review D 58, 116002 (1998).
  • Kosteleckỳ and Russell (2011) V. A. Kosteleckỳ and N. Russell, Reviews of Modern Physics 83, 11 (2011).
  • Hanneke et al. (2011) D. Hanneke, S. F. Hoogerheide,  and G. Gabrielse, Physical Review A 83, 052122 (2011).
  • Hanneke et al. (2008) D. Hanneke, S. Fogwell,  and G. Gabrielse, Physical Review Letters 100, 120801 (2008).
  • Gabrielse et al. (2006) G. Gabrielse, D. Hanneke, T. Kinoshita, M. Nio,  and B. Odom, Physical review letters 97, 030802 (2006).
  • Koch et al. (2022a) B. Koch, F. Asenjo,  and S. Hojman, Phys. Rev. D 105, 053004 (2022a)arXiv:2110.05506 [hep-ph] .
  • Foley and Kusch (1948) H. Foley and P. Kusch, Physical Review 73, 412 (1948).
  • Kusch and Foley (1948) P. Kusch and H. Foley, Physical Review 74, 250 (1948).
  • Brown and Gabrielse (1986) L. S. Brown and G. Gabrielse, Reviews of Modern Physics 58, 233 (1986).
  • Koch et al. (2022b) B. Koch, F. Asenjo,  and S. Hojman, Physical Review D 105, 053004 (2022b).
  • Tiesinga et al. (2018) E. Tiesinga, J. P. Mohr, D. B. Newell,  and B. N. Taylor, Journal of Physical and Chemical Reference Data 50, 033105 (2018).
  • Aoyama et al. (2018) T. Aoyama, T. Kinoshita,  and M. Nio, Physical Review D 97, 036001 (2018).
  • Maluf et al. (2014) R. Maluf, J. Silva, W. Cruz,  and C. Almeida, Physics Letters B 738, 341 (2014).
  • collaboration et al. (2014) A. collaboration, J. Baron, W. C. Campbell, D. DeMille, J. Doyle, G. Gabrielse, Y. Gurevich, P. Hess, N. R. Hutzler, E. Kirilov, et al., Science 343, 269 (2014).
  • neu (2022) Nature Physics 18, 160 (2022).
  • Ilderton (2016) A. Ilderton, Physical Review D 94, 045019 (2016).
  • Kostelecký and Lane (1999) V. A. Kostelecký and C. D. Lane, Phys. Rev. D 60, 116010 (1999).
  • Abi et al. (2021) B. Abi, T. Albahri, S. Al-Kilani, D. Allspach, L. Alonzi, A. Anastasi, A. Anisenkov, F. Azfar, K. Badgley, S. Baeßler, et al., Physical Review Letters 126, 141801 (2021).
  • Keshavarzi et al. (2022) A. Keshavarzi, K. S. Khaw,  and T. Yoshioka, Nuclear Physics B , 115675 (2022).
  • Gradshteyn and Ryzhik (2014) I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products (Academic press, 2014).
  • Kim et al. (2021) G. Kim, T. Demircik, D. K. Hong,  and M. Järvinen, arXiv preprint arXiv:2110.01169  (2021).
  • Alfaro (2017) J. Alfaro, Physics Letters B 772, 100 (2017).
  • Alfaro and Rivelles (2014) J. Alfaro and V. O. Rivelles, Phys. Lett. B 734, 239 (2014)arXiv:1306.1941 [hep-th] .