Magnetic correction to the anomalous magnetic moment of electrons

Fan Lin, Mei Huang

Communications in Theoretical Physics ›› 2022, Vol. 74 ›› Issue (5) : 55202.

PDF(512 KB)
Welcome to visit Communications in Theoretical Physics, June 3, 2025
PDF(512 KB)
Communications in Theoretical Physics ›› 2022, Vol. 74 ›› Issue (5) : 55202. DOI: 10.1088/1572-9494/ac5b5e
Particle Physics and Quantum Field Theory

Magnetic correction to the anomalous magnetic moment of electrons

Author information +
History +

Abstract

We investigate the leading order correction of anomalous magnetic moment (AMM) to electrons in a weak magnetic field and find that the magnetic correction is negative and magnetic field dependent, indicating a magnetic catalysis effect for the electron gas. In the laboratory, to measure the g − 2, the magnitude of the magnetic field B is several T, and correspondingly the magnetic correction to the AMM of electron/muon is around 10−34/10−42, therefore the magnetic correction can be safely neglected in the current measurement. However, when the magnitude of the magnetic field strength is comparable with the electron mass, the magnetic correction of the electron's AMM will become considerable. This general magnetic correction to the charged fermion's AMM can be extended to study quantum chromodynamic matter under a strong magnetic field.

Key words

anomalous magnetic moment / magnetic correction / magnetic field

Cite this article

Download Citations
Fan Lin, Mei Huang. Magnetic correction to the anomalous magnetic moment of electrons[J]. Communications in Theoretical Physics, 2022, 74(5): 55202 https://doi.org/10.1088/1572-9494/ac5b5e

1. Introduction

Recently, the anomalous magnetic moment of the quark [110] has attracted much interest in investigating quantum chromodynamic (QCD) matter under a strong magnetic field created through non-central heavy ion collisions [11, 12]. It has been known that the magnetic field catalyzes the chiral condensate of a spin-0 quark–antiquark pair which carries a net magnetic moment and triggers a dynamical anomalous magnetic moment (AMM) of quarks [1315]. The exact value of a quark's AMM under a magnetic field is unknown but very important for magnetized QCD matter. In this work, we will gain some experience by investigating the AMM of electrons in a magnetic field, which can be calculated from theory and measured experimentally with high precision.
It is well known that a charged fermion can interact with an external magnetic field through an intrinsic magnetic momentum ( = 1, c = 1)
μ=g(q2m)s,
where q, m, and s are the charge, mass, and spin of the fermion, respectively. For an electron, the Lande´ factor ge is believed to be exactly equal to 2 according to Dirac relativistic quantum mechanics. Until a small derivation is observed in an elaborate experiment [16], the deviation of ge from 2 is defined as the anomalous magnetic moment ae = (ge − 2)/2. Meanwhile, quantum electrodynamics (QED) was in its ascendancy and has deepened our understanding of the physical vacuum, where the creation and annihilation of virtual particles contribute additional loop-diagram corrections. The leading order correction ae = α/2π (α ≃ 1/137 is the fine structure constant of QED) calculated by J. Schwinger [17] with renormalizable QED matches the experimental measurement perfectly, powerfully manifesting the validity of quantum field theory.
In the standard model, the theoretical contribution to ae comes from three types of interactions, electromagnetic, hadronic, and electroweak:
ae(theory)=ae(QED)+ae(hadronic)+ae(electroweak),
where the QED contribution can be expressed with perturbative expansion:
ae(QED)=n=1αnae(2n).
J. Schwinger's consequence gives ae(2)=1/2π and the latest calculation from Aoyama et al [18, 19] has reached up to the tenth-order ae(10) with the help of the automatic code generator numerically. The accuracy of experimental electron's AMM ae(experiment) also enhances continually with measurements in a one-electron quantum cyclotron [2022], which provides us with surprising consistency [19, 22]
ae(experiment)ae(theory)=1.06(0.82)×1012.
In essence, the emergence of the electron's AMM is related to slightly broken chiral symmetry because of electron mass me [13], implying any dynamical mechanisms can further break the chiral symmetry always inducing an extra AMM. Naturally, the external magnetic field existing in a one-electron quantum cyclotron is worthy of consideration, for the study of massless spinor QED explicitly shows that a magnetic field reinforces chiral symmetry breaking by endowing the electron with a dynamical mass, known as the magnetic catalysis effect [23], which suggests that the magnetic field could provide an electron with an extra correction. Recently, the latest experimental measurement of muon gμ − 2 confirms the disagreement between experiment and theory [2428], for the potential possibility as a window to pry into new physics, attracting much attention. All of these motivate us to investigate an electron's AMM in a magnetic field and extend to general fermions, which can deepen our understanding of quantum field theory in a magnetic field.
In this work, we investigate the leading order correction of the magnetic field on the AMM of an electron, which can be calculated with relatively high precision. There have been some relevant works to calculate the magnetic correction of an electron by the mass operator and the eigenfunction method [2932], but these do not consider the Feynman diagram of photon–electron vertex correction as done in usual quantum field theory [33]. The motivation of this work is two-fold: firstly, we would like to check how the magnetic field will modify the AMM of electrons; secondly, the magnetic field dependent on an electron's AMM will shed light on the AMM of quarks under a magnetic field.

2. The leading order correction of electron–photon vertex

The electron-photon coupling in QED is described by ieψ¯γμψAμ. Virtual particle loops provide the vertex with corrections ordered by the order shown in figure 1 by Feynman diagrams:
Figure 1. The Feynman diagram of total electron–photon vertex.

Full size|PPT slide

where q=pp is the momentum transformation and not on-shell q2 ≠ 0, the electromagnetic current changes: ψ¯γμψψ¯Γμψ. A complete set of Dirac matrices contain 14×4,γμ,σμν=i2[γμ,γν],γμγ5,γ5, the expansion of Γμ in this basis has 16 components. However, the parity symmetry of QED, initial and final electron on-shell, electromagnetic current, and momentum conservation excludes all other terms leaving Γμ in momentum space taking the following form [33]:
Γμ=γμF1(q2)+iσμνqν2mF2(q2),
(1)
where m is the electron mass and F1, F2 are form factors and ae(QED) = F2(0) [33]. A point-particle in the absence of radiative corrections has F1 ≡ 1, F2 ≡ 0, which derives the Dirac value for the magnetic moment. Schwinger's calculation of the leading order correction shown in figure 2 gives F2(0) = α/2π, which contributes to the biggest correction.
Figure 2. Leading order correction of electron–photon vertex.

Full size|PPT slide

More carefully, the AMM term u¯(p)σμνu(p) denotes a helicity-flipping interaction, not invariant under chiral transformation ψ(x)exp(iθγ5)ψ(x), which indicates the AMM effect emerges from chiral symmetry breaking. Actually, Gordon's identity is derived from the Dirac equation:
2mu¯(p)γμu(p)=u¯(p)×[2(p+p)μ+iσμνqν]u(p),
(2)
this formula explicitly exhibits mass term, causing the left-handed fermion to entangle with the right-handed one. If we naively set m = 0, it leads to the massless spinor QED, which is an infrared-free theory, where the photon and electron are not coupled [34], then there will be no observable AMM, i.e., F2 ≡ 0. Nevertheless, this does not conflict with the perturbative and mass-independent consequence F2(0) = α/2π in the one-loop calculation, because QED is not an asymptotically free theory and does not possess a well-defined chiral limit, so the perturbative expression and computation of F2(0) cannot be applied to the case of m = 0. In contrast, the QCD is asymptotically free, whose chiral limit could be rigorously defined nonperturbatively, so a quark's AMM as a continuous function in the chiral limit is expected. This has been confirmed by the Dyson–Schwinger equation calculation [13], which reveals a dressed-quark with a mass of Mq ∼ 0.5 GeV possessing an AMM κqMq. In addition, κq ∼ 0.1 is much bigger than an intrinsic one due to the strong breaking of the QCD's chiral symmetry, dramatically changing the nature of quark matter in the magnetic field [10]. The presence of a magnetic field induces a magnetic catalysis effect [35], but it is unknown how this would affect the AMM of quarks. It is quite challenging to do the direct calculation in QCDs because of the nonperturbative nature, therefore we will check the magnetic field-dependent behavior of an electron's AMM and use it for reference in the case of quarks.
To calculate the magnetic field correction to the AMM of an electron, we put the Lagrangian density of QED in an external magnetic field,
L=ψ¯(iγμDμm)ψ14FμνFμν
(3)
where the covariant derivative Dμ=μie(Aμext+Aμ) and the vector potential Aμext introduce an external and uniform magnetic field B = (0, 0, B) along the z direction, we choose the symmetric gauge Aμext=B2(0,y,x,0) for convenience in later calculations. The presence of a magnetic field breaks translation invariance and the amplitude of the process has to be computed in coordinate space and subsequently integrated over space-time [36]. The photon does not carry a charge so the propagator keeps:
Gμν(p)=igμνp2+iϵ.
(4)
For the electron propagator under a uniform magnetic field, the proper-time approach is adopted in coordinate space representation [37]:
S(x,x)=Φ(x,x)d4k(2π)4eik(xx)S~(k),
(5)
where
Φ(x,x)=exp{i|q|×xxdξμ[Aμ+12Fμν(ξx)ν]}
(6)
is the Schwinger's phase factor, the Fourier transformation of translation invariant part S~(k) is given as:
S~(k)=0dseis(k2m2k2tan(|qB|s)|qB|s+iϵ)×[/k+m+(k1γ2k2γ1)tan(|qB|s)]×[1γ1γ2tan(|qB|s)],
(7)
where s is Schwinger's proper-time, the basic notations are defined as:
kμ(0,k1,k2,0),kμ(k0,0,0,k3),k2k12+k22,k2k02k32,k2=k2k2.
Here q represents the charge of the fermion and we assume eB > 0 for convenience.
By using the perturbative method, we can expand the total vertex amplitude with respect to α in leading-order
u¯(p)Γμu(p)=u¯(p)γμu(p)+u¯(p)Mμu(p)+O(α).
(8)
As is shown in figure 2, the leading-order correction Mμ reads:
Mμ=d4xd4yd4zΦ(x,y)Φ(y,z)d4t(2π)4d4k(2π)4d4k(2π)4×eit(zx)eik(xy)eik(yz)eipxeiqyeipzGνρ(t)×(ieγν)S~(k)γμS~(k)(ieγρ),=d4t(2π)4d4k(2π)4d4k(2π)4×ie2t2×0dr0dseir(k2m2k2tan(eBr)eBr)eis(k2m2k2tan(eBr)eBs)×γν[/k+m+(k1γ2k2γ1)tan(eBr)][1γ1γ2×tan(eBr)]γμ×[/k+m+(k1γ2k2γ1)tan(eBs)][1γ1γ2tan(eBs)]γν×d4xd4yd4zΦ(x,y)Φ(y,z)eit(zx)eik(xy)×eik(yz)eipxeiqyeipz,
(9)
where x, y, z are the three space-time vertexes with corresponding momentum, s, r and Φ(x, y), Φ(y, z) are proper-time and phase factors emerging from two-electron propagators. The expression of Mμ is quite tedious, in the following we separate Mμ into three parts: 1) the phase factor P(t,k,k), 2) the momentum integral I, and 3) the γ matrix part Θμ(r, s) with their explicit expressions given below:
P(t,k,k)=d4xd4yd4zΦ(x,y)Φ(y,z)×eit(zx)eik(xy)eik(yz)eipxeiqyeipz,
(10)
I=d4t(2π)4d4k(2π)4d4k(2π)4P(t,k,k)×ie2t20dr0dseir(k2m2k2tan(eBr)eBr)×eis(k2m2k2tan(eBr)eBs)
(11)
Θμ(r,s)=γν[/k+m+(k1γ2k2γ1)tan(eBr)]×[1γ1γ2tan(eBr)]γμ×[/k+m+(k1γ2k2γ1)tan(eBs)]×[1γ1γ2tan(eBs)]γν.
(12)
In the next subsections, we will carefully handle these three parts, respectively.

2.1. The phase factor part

In this section, we deal with the Schwinger part P(t,k,k). With the symmetric gauge Aμ=B2(0,y,x,0) and equation (6), the only non-zero terms of Fμν are F12 = − F21 = − B , so the product of the phase factor can be written as
Φ(x,y)Φ(y,z)=eieB2(ϵijxjyi+ϵijyjzi)=eieB2ϵijyi(xz)j.
(13)
Introducing a variable h = xz, one has x = h + z, d4x = d4h and after the integration of the space-time, we have:
P(t,k,k)=d4xd4yd4zΦ(x,y)Φ(y,z)eit(zx)eik(xy)×eik(yz)eipxeiqyeipz=(2π)24l4ei2(p+tk)1(q+kk)2/|eB|×ei2(p+tk)2(q+kk)1/|eB|×(2π)4δ4(pk+kp)(2π)2×δ2(p+tk)(2π)2δ2(q+kk),
(14)
where l=1/|eB| is the magnetic length. More details of this calculation are featured in appendix A. It is worth mentioning that the longitudinal momentum is strictly conserved but the transverse momentum expresses a distribution function. The magnetic field causes dimension reduction leaving the longitudinal physics unchanged but the transverse remains a little subtle. Now Mμ becomes:
Mμ=(dtdkdk)(2π)2δ2(ppq)×(2π)2δ2(pkp+k)×(2π)24l4ei2(p+tk)1(q+kk)2/|eB|×ei2(p+tk)2(q+kk)1/|eB|×Aμ(t,k,k)=Δ(dtdkdk)(2π)24l4ei2(tk)1(kk)2/|eB|×ei2(tk)2(kk)1/|eB|×Aμ(t,k,k),
(15)
where
Δ=(2π)2δ2(ppq)(2π)2×δ2(pkp+k)
(16)
and
Aμ(t,k,k)=Gνρ(t)(ieγν)S~(k)γμS~(k)(ieγρ)=ie2t20dr0ds×eir(k2m2k2tan(eBr)eBr)×eis(k2m2k2tan(eBr)eBs)×Θμ(r,s),
(17)
with (dk)d4k(2π)4,(dk)d2k(2π)2,(dk)d2k(2π)2 for simplification. The AMM is defined when q is very small and the measurement of AMM proceeds in a low-energy condition, which suggests a reasonable approximation: q → 0, p → 0 applied in the last line of equation (15), so the momentum conservation conditions: t=kp,k=k+q become t=k,k=k.

2.2. The momentum integral part

Actually, the AMM part ΘAMMμ can be extracted from Θμ(r, s) later and irrelevant to the momenta t,k,k, we can integrate all three momenta in equation (15), which exactly results in integral part I :
I=(2π)24l40dr0ds×(dtdkdk)ie2k2t2×ei2(tk)1(kk)2/|eB|ei2(tk)2(kk)1/|eB|×eir(k2m2k2tan(eBr)eBr)eis(k2m2k2tan(eBr)eBs),
(18)
where t2=k2t2 by momentum conservation. In order to remove the pole of momentum integral, we do a Wick rotation transferring Minkowski space-time into a Euclidean one [23]: k0 → ik0, r → − ir, s → − is. So the integral becomes
I=(2π)24l4×(i)20dr0ds×(i)(dtdkdkE)ie2kE2+t2×ei2(tk)1(kk)2/|eB|ei2(tk)2(kk)1/|eB|× er(kE2 + m2)k2 l2tanh(eBr)×es(kE2 + m2)k2 l2tanh(eBs),
(19)
here, we use tan(ix)=itanh(x),kE=k02+k32, the integral about proper-time r or s, momentum k0 and photon propagator contribute −i, i, −1 under Wick rotation, respectively. Even in Euclidean space, the integral calculation is still tedious and complicated, therefore we put all details of this calculation in appendix B. Finally, we obtain a much more transparent expression:
I=α4πl200drds×e(r+s)m2[tanh(eBr)+tanh(eBs)]lnvv1
(20)
with fine structure constant α = e2/4π and v=(eBr+eBs)/tanh(eBr+eBs) .

2.3. The AMM term extracted from Θμ(r, s)

The presence of the magnetic field breaks the original parity symmetry, thus the leading-order vertex structure equation (12) becomes more complicated than the original one in equation (1). Fortunately, we just need to extract and calculate the term related to the AMM and to not get stuck in the swamp of total Θμ(r, s). As is seen in equation (1), the linear term of q corresponds to the AMM effect in absence of a magnetic field. We naturally have the momentum constraint in the longitudinal direction q=kk which helps to separate the linear term of q but the momentum constraint in the transverse direction emerges as a distribution function equation (14). Therefore, it is impossible to have the same definition of AMM as in the vacuum, but considering the phase factor equation (14) is a highly oscillating function in a weak magnetic field region, this indicates that the momentum q=kk substantially dominates the distribution. So under the approximation qkk, the linear part of q in Θμ can be subtracted and takes the following form:
Θμ(r,s)qlinear=γν/q~[1γ1γ2tan(eBr)]γμ×[/k+m+(k1γ2k2γ1)tan(eBs)]×[1γ1γ2tan(eBs)]γν,
(21)
where q~=q+q^,q^:=(0,q2tan(eBr),q1tan(eBr),0). It is noticed that only the terms consisting of the product of even γ matrices contribute to the AMM, thus we directly get rid of other terms and obtain:
Θμ(r,s)AMM=γν/q~[1γ1γ2tan(eBr)]×γμm[1γ1γ2tan(eBs)]γν=mγν/q~[γμγμγ1γ2tan(eBs)γ1γ2γμtan(eBr)×+γ1γ2γμγ1γ2tan(eBr)tan(eBs)]γν.
(22)
The presence of the magnetic field creates anisotropy between the xy plane and z direction so that the AMM correction becomes different in the longitudinal and transverse directions. While a magnetic field along the z axis is imposing, the classical interaction term − μ · B, μ = (μx, μy, μz) implies only μz is measurable, corresponding to μ = 1, 2 components of the magnetic vector potential Aμ. Thus the magnetic interaction amplitude is given by
u¯(p)Miu(p)A~i=u¯(p)M1u(p)A~1+u¯(p)M2u(p)A~2
(23)
where A~μ is the Fourier transformation of Aμ, which guides us to concentrate on Θμ with μ = 1, 2 components. Because two transverse directions share the same physics, therefore we just need to display the case μ = 1.
When μ = 1, we can simplify the expression into a more friendly form by the Dirac algebra:
Θ1(r,s)AMM=iσ1αqα2m{8m2[1+tan(eBr)tan(eBs)]}+iσ2αqα2m{8m2[tan(eBs)tan(eBr)]}+iσ1αq^α2m{8m2[1+tan(eBr)tan(eBs)]}+iσ2αq^α2m{8m2[tan(eBs)tan(eBr)]}+4m/q[γ1+γ2tan(eBs)γ2tan(eBr)+γ1tan(eBr)tan(eBs)]+4m/q^[γ1+γ2tan(eBs)γ2tan(eBr)+γ1tan(eBr)tan(eBs)].
(24)
The presence of the magnetic field breaks the concise structure of the electron-photon vertex equation (1) in the vacuum. From the above-complicated expression, it is difficult to read explicit information about the magnetic correction of the AMM term.
Comparing with equation (1), we find the first term of equation (24) possesses the standard form contributing to the magnetic correction of AMM [33], but the other terms are more veiled, whose physical meaning will be explained by seeking the corresponding non-relativistic limit analyzed in appendix C. Finally, the last line also contributes corrections to AMM. Extracting the coefficient of the total magnetic correction term of the electron in equation (24) gives
8m2[tan2(eBr)2tan(eBr)tan(eBs)].
(25)
Some non-invariant terms under gauge transformation that cannot present observable phenomena are neglected.

3. Conclusion and discussion

For a measurable AMM correction, we only need to calculate the major contribution of Mμ in the x, y direction. Extracting the relevant term from equation (24) and doing a Wick rotation r → − ir, s → − is,
8m2[tan2(eBr)2tan(eBr)tan(eBs)] 8m2[2tanh(eBr)tanh(eBs)tanh2(eBr)].
(26)
Therefore, one multiplies the meaningful part with integral I in equation (20) obtaining the magnetic correction to the AMM of the electron:
κB=4m2α2πl200e(r+s)m2×[2tanh(eBr)tanh(eBs)tanh2(eBr)][tanh(eBr)+tanh(eBs)]lnvv1drds=4m2l2α2π00e(R+S)m2l2×[2tanh(R)tanh(S)tanh2(R)][tanh(R)+tanh(S)]lnVV1dRdS
(27)
with v=eBr+eBstanh(eBr+eBs),V=R+Stanh(R+S),l=1/|eB|, and variable transformations R = eBr, S = eBs done. First of all, there is one very interesting point that two parameters r, s (or R, S) in equation (27) do not keep symmetric as in equation (20) due to the presence of the magnetic field. As such, the time reversal invariance no longer maintains as QED in the vacuum, so the two parameters r, s, corresponding propagators with a time order, have different positions. Moreover, we see that the final expression of κB manifesting the magnetic correction is negative prompting us to check where the minus sign originates. Let us recall the equation (19), there are two electron propagators, one photon propagator, two vertex, respectively, contributing (i)2, − i × ( − 1), (i)2 after Wick rotation with an extra i in differential dk, which results in a minus sign (i)2×i×(i)2×i=1. Here, we would like to point out that the result in equation (27) vanishes when eB=0 so cannot continuously go back to the Schwinger result ae=α2π , one can regard this magnetic correction as an extra contribution from the magnetic field. After taking the vacuum contribution into consideration, the leading order correction to AMM of the electron is:
ae=α2π+κB,=α2π(14m2l200e(R+S)m2l2×[2tanh(R)tanh(S)tanh2(R)][tanh(R)+tanh(S)]lnVV1dRdS).
(28)
Analytically, m2l2 is the only dimensionless combination with eB, m, which appears twice in the expression: 1) 4m2l2 as a coefficient in front of the whole expression confirms that the magnetic correction to AMM κBm2, which indicates that the magnetic correction is a nonperturbative contribution similar to the AMM induced by the chiral symmetry breaking [1315]. This also naturally explains why the magnetic correction results cannot smoothly go back to α2π when eB=0. The nonperturbative nature of chiral symmetry breaking induced by a constant magnetic field was also been discussed in the 1990s [3841]. 2)m2l2 appears in the exponential factor which strongly depresses the integral so that the correction κB will only become considerable when eBm.
For the magnetic field magnitude B ∼ 1 T ↔ eB ∼ 2 × 10−10MeV2 used in the laboratory to measure the electron g − 2, ∣κB∣ < 10−33, is much smaller than the measurement ae = 0.00115965218073(28) [22, 24]. For the muon, whose mass is much heavier than the electron (mμ/me)2=43000, under the same magnetic field magnitude B ∼ 1 T ↔ eB ∼ 2 × 10−10MeV2, we find that the corresponding magnetic correction to the AMM is ∣κB∣ < 10−42, which is also much smaller than the experimental measurement of muon aμ = 0.00116592040(54) [24]. Thus one can safely neglect the magnetic correction to the AMM of the electron/muon in the vacuum.
The magnetic correction κB will become considerable when eBm, which can be seen explicitly from the numerical result shown in figure (3) on the magnetic correction κB in equation (27) to the electron and muon as a function of eB/m with m = me and m = mμ. To see how the negative magnetic correction κB affects physics, we consider the electron or a fermion with AMM κ + κB travels in the magnetic field, one can derive its dispersion as below:
El,s2={pz2+[m2+(2l+1+s)|eB|s(κ+κB)eB]2,l1.pz2+[m(κ+κB)eB]2,l=0.
(29)
where l, s = ± 1 are quantum numbers of orbital and spin angular momentum, respectively, eB > 0 for convenience. As is seen, the normal AMM κ causes the splitting of degenerate energy levels known as the Zeeman effect in the spectrum, which is used in the accurate measurement of the AMM [18, 19]. However, the energy level splitting will be suppressed with a negative magnetic correction taken into account κκ + κB, which will lead to a slight diamagnetism. κB increases the energy of the system expressing a trivial magnetic catalysis effect as usual [23]. We will explicitly show these properties in a separate paper [42]. Actually, our calculation mainly concentrates on the leading order correction to the AMM of the electron, the next to leading order and higher orders should also contribute to the correction, but the small coupling of QED αO(1/100) guarantees that we can safely ignore higher-order influence. Moreover, due to the vacuum polarization, the photon with a mass Mγ2O(α|eB|) results in a very strong magnetic field m2|q2||eB| and |q2||eB| [23], which depresses the higher-order contribution further.
Figure 3. The magnetic correction to the AMM of electron/muon as a function of the dimensionless quantity eB/m, with m = me or mμ.

Full size|PPT slide

As addressed in the introduction, the AMM of a quark plays a vital role for QCD matter under a strong magnetic field when the magnitude of the magnetic field becomes comparable with the quark mass square, such as in the early evolution of cosmological phase transition [43], non-central heavy ion collisions [12], and neutron-star merger [44]. The exact value of the AMM of quarks under a magnetic field remains unknown because of its nonperturbative nature. In most literature, the AMM of quarks under a magnetic field was taken as a positive constant. We can extend the magnetic field-dependent AMM of electrons to the case of quarks and investigate its effect on the magnetized QCD matter. We will report the results in the forthcoming work [42].

References

1
Ferrer E J de la Incera V 2010 Nucl. Phys. B 824 217 238
2
Fayazbakhsh S Sadooghi N 2014 Phys. Rev. D 90 105030
3
Strickland M Dexheimer V Menezes D P 2012 Phys. Rev. D 86 125032
4
Mao S Rischke D H 2019 Phys. Lett. B 792 149
5
Mei J Mao S 2020 Phys. Rev. D 102 114035
6
Ghosh S Chaudhuri N Roy P Sarkar S 2021 Phys. Rev. D 103 116008
7
Chaudhuri N Ghosh S Sarkar S Roy P 2019 Phys. Rev. D 99 116025
8
Ferrer E J de la Incera V Manreza Paret D Pérez Martínez A Sanchez A 2015 Phys. Rev. D 91 085041
9
Chaudhuri N Ghosh S Sarkar S Roy P 2020 Eur. Phys. J. A 56 213
10
Xu K Chao J Huang M 2021 Phys. Rev. D 103 076015
11
Skokov V Illarionov A Y Toneev V 2009 Int. J. Mod. Phys. A 24 5925
12
Deng W-T Huang X-G 2012 Phys. Rev. C 85 044907
13
Chang L Liu Y Roberts C D 2011 Phys. Rev. Lett. 106 072001
14
Ferrer E J de la Incera V 2009 Phys. Rev. Lett. 102 050402
15
Ferrer E J de la Incera V Portillo I Quiroz M 2014 Phys. Rev. D 89 085034
16
Kusch P Foley H M 1948 Phys. Rev. 74 250
17
Schwinger J S 1948 Phys. Rev. 73 416 417
18
Aoyama T Hayakawa M Kinoshita T Nio M 2007 Phys. Rev. Lett. 99 110406
19
Aoyama T Hayakawa M Kinoshita T Nio M 2012 Phys. Rev. Lett. 109 111807
20
Van Dyck R S Schwinberg P B Dehmelt H G 1987 Phys. Rev. Lett. 59 26 29
21
Odom B C Hanneke D D'Urso B Gabrielse G 2006 Phys. Rev. Lett. 97 030801
22
Hanneke D Fogwell S Gabrielse G 2008 Phys. Rev. Lett. 100 120801
23
Miransky V A Shovkovy I A 2015 Phys. Rept. 576 1 209
24
Abi B (Muon g-2) 2021 Phys. Rev. Lett. 126 141801
25
Bennett G W (Muon g-2) 2006 Phys. Rev. D 73 072003
26
Davier M Hoecker A Malaescu B Zhang Z 2011 Eur. Phys. J. C 71 1515
Davier M Hoecker A Malaescu B Zhang Z 2012 Eur. Phys. J. C 72 1874 erratum:
27
Davier M Hoecker A Malaescu B Zhang Z 2017 Eur. Phys. J. C 77 827
28
Davier M Hoecker A Malaescu B Zhang Z 2020 Eur. Phys. J. C 80 241
Davier M Hoecker A Malaescu B Zhang Z 2020 Eur. Phys. J. C 80 410 erratum:
29
Newton R G 1954 Phys. Rev. 96 523 528
30
Jancovici B 1969 Phys. Rev. 187 2275 2276
31
Ternov I M 1969 Sov. Phys. JETP 28 1206
32
Ritus V I 2001 J. Exp. Theor. Phys. 93 860 864
Ritus V I 1978 Zh. Eksp. Teor. Fiz. 75 1560 1583
33
Peskin M E Schroeder D V 1995 An Introduction to Quantum Field Theory Reading, MA Addison-Wesley
34
Shifman M 2012 Advanced Topics in Quantum Field Theory: A Lecture Course Cambridge Cambridge University Press 11 39
35
Shovkovy I A 2013 Lect. Notes Phys. 871 13 49
36
Ayala A Castano-Yepes J D Dominguez C A Hernandez L A Hernandez-Ortiz S Tejeda-Yeomans M E 2017 Phys. Rev. D 96 014023
Ayala A Castano-Yepes J D Dominguez C A Hernandez L A Hernandez-Ortiz S Tejeda-Yeomans M E 2017 Phys. Rev. D 96 119901 erratum
37
Schwinger J S 1951 Phys. Rev. 82 664 679
38
Gusynin V P Miransky V A Shovkovy I A 1994 Phys. Rev. Lett. 73 3499 3502
Gusynin V P Miransky V A Shovkovy I A 1996 Phys. Rev. Lett. 76 1005 erratum:
39
Gusynin V P Miransky V A Shovkovy I A 1995 Phys. Rev. D 52 4718 4735
40
Gusynin V P Miransky V A Shovkovy I A 1995 Phys. Lett. B 349 477 483
41
Gusynin V P Miransky V A Shovkovy I A 1995 Phys. Rev. D 52 4747 4751
42
Lin F Xu K Huang M in preparation
43
Vachaspati T 1991 Phys. Lett. B 265 258 261
44
Kiuchi K Cerdá-Durán P Kyutoku K Sekiguchi Y Shibata M 2015 Phys. Rev. D 92 124034
45
Gradshteyn I S Ryzhik I M 2014 Table of Integrals, Series, and Products Cambridge Academic

Acknowledgments

We thank Xinyang Wang and Anatoly V. Borisov for their helpful discussions. This work is supported by the NSFC under Grant Nos. 11 725 523 and 11 735 007, Chinese Academy of Sciences under Grant No.XDPB15 and XDB34000000, the start-up funding from University of Chinese Academy of Sciences(UCAS), and the Fundamental Research Funds for the Central Universities.

Appendix A. Calculation details in section 2.1

In order to to manage the Schwinger phase Φ(x, y)Φ(y, z), we have

xyξidξj=12(x+y)i(yx)jxyAμdξμ=B2ϵijxjyi(i,j=1,2),
(A.1)
and the only non-zero terms of Fμν are F12 = − F21 = − B, which gives
xy12Fμν(ξx)νdξμ=B2ϵij(yx)i(yx)j2=0.
(A.2)
Then
P(t,k,k)=d4xd4yd4zΦ(x,y)Φ(y,z)eit(zx)eik(xy)×eik(yz)eipxeiqyeipz=d4hd4yd4zeieB2ϵijyihj×eih(p+tk)eiy(q+kk)eiz(ppk+k)=(2π)4δ4(pk+kp)×d4hd4yeieB2ϵijyihjeih(p+tk)eiy(q+kk)=(2π)4δ4(pk+kp)×d2hd2yeih(p+tk)eiy(q+kk)
×d2hd2yeih(p+tk)×eiy(q+kk)×eieB2ϵijyihj=(2π)4δ4(pk+kp)×(2π)2δ2(p+tk)(2π)2δ2(q+kk)×d2hd2yeih(p+tk)×eiy(q+kk)×eieB2ϵijyihj=(2π)4δ4(pk+kp)(2π)2δ2(p+tk)×(2π)2δ2(q+kk)×d2heih(p+tk)dy1dy2eiy1((q+kk)1eB2h2)×eiy2((q+kk)2eB2h1)=(2π)4δ4(pk+kp)(2π)2δ2(p+tk)×(2π)2δ2(q+kk)×d2heih(p+tk)(2π)2×δ((q+kk)1eB2h2)δ((q+kk)2eB2h1)=(2π)4δ4(pk+kp)(2π)2δ2(p+tk)×(2π)2δ2(q+kk)×(2π)24l4ei2(p+tk)1(q+kk)2/|eB|×ei2(p+tk)2(q+kk)1/|eB|,
(A.3)
where the new conventions are defined: p · kp0k0p3k3, p · k ≡ − p1k1p2k2, so p · k = p · k + p · k. we can integrate t||,k||:
Mμ=d4t(2π)4d4k(2π)4d4k(2π)4P(t,k,k)×Aμ(t,k,k)=(dtdkdk)(2π)4δ4(pk+kp)×(2π)2δ2(p+tk)(2π)2δ2(q+kk)×(2π)24l4ei2(p+tk)1(q+kk)2/|eB|×ei2(p+tk)2(q+kk)1/|eB|×Aμ(t,k,k)=(dtdkdk)(2π)2δ2(ppq)×(2π)2δ2(pkp+k)×(2π)24l4ei2(p+tk)1(q+kk)2/|eB|×ei2(p+tk)2(q+kk)1/|eB|×Aμ(t,k,k).
(A.4)

Appendix B. Calculation details in section 2.2

After Wick rotation k0ik0, r → − ir, s → − is, we obtain

I=(2π)24l4×(i)20dr0ds×(i)(dtdkdkE)ie2kE2+t2×ei2(tk)1(kk)2/|eB|ei2(tk)2(kk)1/|eB|×er(kE2+m2)k2l2tanh(eBr)×es(kE2+m2)k2l2tanh(eBs),
(B.1)
the subscript E will be omitted from now on for convenience. Firstly, we do the integral related to t:
I=[e2(2π)24l4]0dr×0ds(dkdk)×er(k2+m2)k2l2tanh(eBr)×es(k2+m2)k2l2tanh(eBs)×T,
(B.2)
where
T=(dt)1k2+t2ei2(tk)1(kk)2/|eB|×ei2(tk)2(kk)1/|eB|=ei2k1(kk)2/|eB|ei2k2(kk)1/|eB|×dt1dt2(2π)2eiat1eibt21k2+t12+t22=ei2k1(kk)2/|eB|ei2k2(kk)1/|eB|×+dt22πeibt2ek2+t22|a|2k2+t22=ei2k1(kk)2/|eB|ei2k2(kk)1/|eB|×12π0+dt2ek2+t22|a|k2+t22cos(bt2)=ei2k1(kk)2/|eB|ei2k2(kk)1/|eB|×12πK0(|k|a2+b2).
(B.3)
where a=2l2(kk)2,b=2l2(kk)1 and Kν(z) is the Bessel function of imaginary argument, two formulae have been used [45] in the calculation:
12π+eiξxc2+x2dx=ec|ξ|2c[c>0,ξR],
(B.4)
0eβγ2+x2γ2+x2cosbxdx=K0(γβ2+b2)×[Reβ>0,Reγ>0,b>0].
(B.5)
Then, we obtain
I=0dr0ds(dkdk)×er(k2+m2)k2l2tanh(eBr)es(k2+m2)k2l2tanh(eBs)×[e2(2π)24l4]ei2k1(kk)2/|eB|×ei2k2(kk)1/|eB|12πK0(|k|a2+b2).
(B.6)
Making a variable transformation k~=kk,dk~=dk, it is easy to see a2+b2=2l2(kk)2=2l2|k~|, which simplifies the expression:
I=[e2(2π)4l4]0dr0ds×(dk~dk)ei2k1(kk)2/|eB|ei2k2(kk)1/|eB|×er(k2+m2)(k~+k)2l2tanh(eBr)×es(k2+m2)k2l2tanh(eBs)K0(2l2|k||k~|)=[e2(2π)4l4]0dr0ds×(dk~dk)e(r+s)(k2+m2)k~2l2tanh(eBr)×K0(2l2|k||k~|)×K,
(B.7)
where K is the part related to k:
K=(dk)ei2l2k1k~2ei2l2k2k~1×ek2l2tanh(eBr)2l2k~ktanh(eBr)×ek2l2tanh(eBs)=dk12π×ei2l2k1k~2ek12l2tanh(eBr)k12l2tanh(eBs)2k1k~1tanh(eBr)×dk22π×ei2l2k2k~1ek22l2tanh(eBr)k22l2tanh(eBs)2k2k~2tanh(eBr)=1(2π)2πl2[tanh(eBr)+tanh(eBs)]e4l4[k1[tanh(eBr)ik2]24tanh(eBr)+tanh(eBs)]×e4l4[l2[tanh(eBr)ik1]24tanh(eBr)+tanh(eBs)]=14πl21[tanh(eBr)+tanh(eBs)]×exp{k~2l2[tanh2(eBr)1]tanh(eBr)+tanh(eBs)},
(B.8)
an integral formula has been used [45]:
exp(p2x2±qx)dx=exp(q24p2)πp[Rep2>0].
(B.9)
We have
I=0dr0ds(dk~dk)×[e2(2π)4l4]e(r+s)(k2+m2)k~2l2tanh(eBr)×14πl2K0(2l2|k||k~|)[tanh(eBr)+tanh(eBs)]×exp{k~2l2[tanh2(eBr)1]tanh(eBr)+tanh(eBs)}=0dr0ds(dk~dk)[e22l2]×e(r+s)(k2+m2)k~2l2tanh(eBr+eBs)×1[tanh(eBr)+tanh(eBs)]×120dueuuel4k2k~2u=0dr0dse2l2[tanh(eBr)+tanh(eBs)]×(dk)e(r+s)(k2+m2)×U,
(B.10)
where
U=0dueuu(dk~)ek~2l2tanh(eBr+eBs)l4k2k~2u,
(B.11)
the formulae we used here are [45]:
K0(z)=120euz2/4uduu×[|argz|<π4,Rez2>0],
(B.12)
tanh(x+y)=tanhx+tanhy1+tanhxtanhy.
(B.13)
Define two new notations α, β:
α+βu=l2tanh(eBr+eBs)+l4k2u,
(B.14)
so that we can integrate the k~ :
U=0dueuu(dk~)ek~2l2tanh(eBr+eBs)l4k2k~2u=0dueuu(dk~)e(α+βu)k~2=0dueuu0d|k~|2π|k~|e(α+βu)|k~|2=14π0dueuαu+β=14π0dueuαu+l4k2,
(B.15)
which simplifies I again:
I=00[e2l2]drds[tanh(eBr)+tanh(eBs)]×(dk)e(r+s)(k2+m2)×14π0dueuαu+l4k2=e2l24π00e(r+s)m2drds[tanh(eBr)+tanh(eBs)]×(dk)e(r+s)k20dueuαu+l4k2=e2l24π00e(r+s)m2drds[tanh(eBr)+tanh(eBs)]×K.
(B.16)
The integral of k becomes:
K=(dk)e(r+s)k20dueuαu+l4k2=0dueu0d|k|2πl4×|ke(r+s)|k|2αu/l4+|k|2=14πl40du0dwewuw+vu=14πl4lnvv1,
(B.17)
where a variable transformation is applied w = ∣k∣∣2, and v=(r+s)α/l4=eB(r+s)/tanh(eBr+eBs)1. Here a necessary integral is used:
0xeax2b+x2dx=120ewdww+ab=12eabΓ(0,ab)[a,b>0],
(B.18)
Γ(x,z)=ztx1etdt is the incomplete Gamma function. Through tedious computation, we finally receive:
I=e2l24π00e(r+s)m2drds[tanh(eBr)+tanh(eBs)]×14πl4lnvv1=αQED4πl2×00e(r+s)m2drds[tanh(eBr)+tanh(eBs)]lnvv1
(B.19)
with v=eBr+eBstanh(eBr+eBs) .

Appendix C. Calculation details in section 2.3

When μ = 1, we can simplify the expression into a more friendly form by Dirac algebra:

Θμ(r,s)AMM=4mq~α[gα1+gα2tan(eBs)×gα2tan(eBr)+gα1tan(eBr)tan(eBs)]=4mq~α[iσα1+iσα2tan(eBs)iσα2tan(eBr)+iσα1tan(eBr)tan(eBs)]+4mq~α[γαγ1+γαγ2tan(eBs)γαγ2tan(eBr)+γαγ1tan(eBr)tan(eBs)]=iσ1αqα2m{8m2[1+tan(eBr)tan(eBs)]}+iσ2αqα2m{8m2[tan(eBs)tan(eBr)]}+iσ1αq^α2m{8m2[1+tan(eBr)tan(eBs)]}+iσ2αq^α2m{8m2[tan(eBs)tan(eBr)]}+4m/q[γ1+γ2×tan(eBs)γ2tan(eBr)+γ1tan(eBr)tan(eBs)]+4m/q^[γ1+γ2×tan(eBs)γ2tan(eBr)+γ1tan(eBr)tan(eBs)],
(C.1)
Similarly, when μ = 2
Θμ(r,s)AMM=4mq~α[gα2gα1tan(eBs)+gα1×tan(eBr)+gα2tan(eBr)tan(eBs)]=4mq~α[iσα2iσα1tan(eBs)+iσα1tan(eBr)+iσα2tan(eBr)tan(eBs)]+4mq~α[γαγ2γαγ1×tan(eBs)+γαγ1tan(eBr)+γαγ2tan(eBr)tan(eBs)]=iσ2αqα2m{8m2[1+tan(eBr)tan(eBs)]}+iσ1αqα2m{8m2[tan(eBr)tan(eBs)]}+iσ2αq^α2m{8m2[1+tan(eBr)tan(eBs)]}+iσ1αq^α2m{8m2[tan(eBr)tan(eBs)]}+4m/q[γ2γ1tan(eBs)+γ1×tan(eBr)+γ2tan(eBr)tan(eBs)]+4m/q^[γ2γ1tan(eBs)+γ1×tan(eBr)+γ2tan(eBr)tan(eBs)],
(C.2)
with q~μ=qμ+q^μ,q^μ:=(0,q2tan(eBr),q1tan(eBr),0). Consider the last equal sign in equation (C.1, C.2), iσ1αqα2m,iσ2αqα2m terms are what we seek providing AMM with a magnetic correction, 4m/q terms actually are irrelevant with AMM by conversion in virtue of Gordon identity:
u¯(p)/qγμu(p)=u¯(p)[2mγμ2pμ]u(p),
(C.3)
the initial and final state of the electron are considered in infinitely far positions where there is no magnetic field so the Gordon identity holds validity. For the remaining terms, the analysis is a little tedious by seeking the corresponding non-relativistic limit as done in quantum field theory [33]:
u¯(p)Θiu(p)A~i=u¯(p)Θ1u(p)A~1+u¯(p)Θ2u(p)A~2.
(C.4)
Insert the non-relativistic expansion and keep the leading order of momentum:
u(p)=(pσξpσ¯ξ)m((1pσ/2m)ξ(1+pσ/2m)ξ).
(C.5)
We have
u¯(p)iσ1αq^α2mu(p)A~1+u¯(p)iσ2αq^α2mu(p)A~2i(q^2A~1q^1A~2)ξσ3ξ,
(C.6)
u¯(p)iσ2αq^α2mu(p)A~1+u¯(p)iσ1αq^α2mu(p)A~2i(q^2A~2q^1A~1)ξσ3ξ,
(C.7)
so taking q^μ=(0,q2tan(eBr),q1tan(eBr),0) into account, one can find the above two terms are gauge-dependent corresponding to operators ∂1A1 + ∂2A2 and ∂1A2 + ∂2A1 in coordinate space. So the fifth lines of equations (C.1), (C.2) make no sense for an observable quantity, whose gauge-dependent will eliminate after integration with q in a complete calculation of the Feynman diagram. Similarly, for the last line of equations (C.1), (C.2), we analyze in the same way and find that the principle contribution to the AMM is
4m(tan(eBs)tan(eBr))(q^2A~1+q^1A~1)γ1γ2,
(C.8)
after inserting the spinor expansion:
u¯(p)[4m(tan(eBs)tan(eBr))×(q^1A~1+q^2A~2)γ1γ2]u(p)8m2tan(eBr)[tan(eBr)tan(eBs)]×[i(q2A~1q1A~2)]ξσ3ξ.
(C.9)
we get the coefficient of the total magnetic correction term of the electron as follows
8m2tan(eBr)[tan(eBr)tan(eBs)]8m2[1+tan(eBr)tan(eBs)]=8m2[tan2(eBr)2tan(eBr)tan(eBs)1].
(C.10)
Because of the dimension reduction originating from the magnetic field, leaving the term − 8m2 after directly setting eB=0 in the above does not make sense and does not equate to the vacuum contribution. So our final magnetic correction of AMM is 8m2[tan2(eBr)2tan(eBr)tan(eBs)]. For the longitudinal directions μ = 0, 3:
Θμ(r,s)AMM=mγν/q~[γμγμγ1γ2tan(eBs)γ1γ2γμtan(eBr)γμtan(eBr)tan(eBs)]γν=iσμαq~α2m{8m2[1tan(eBr)tan(eBs)]}+4m/qγμ[1tan(eBr)tan(eBs)]+4m/q^γμ[1tan(eBr)tan(eBs)]+2mq~α(g2αγμγ1g1αγμγ2+iσμαγ1γ2)×[tan(eBr)+tan(eBs)],
(C.11)
the analysis is in the same way but the longitudinal directions will not contribute an observable AMM when a magnetic field is along the z axis, so we omit it.

Funding

Chinese Academy of Sciences(XDPB09)
National Natural Science Foundation of China(11725523)
National Natural Science Foundation of China(11735007)

RIGHTS & PERMISSIONS

© 2022 Institute of Theoretical Physics CAS, Chinese Physical Society and IOP Publishing
PDF(512 KB)

175

Accesses

0

Citation

1

Altmetric

Detail

Sections
Recommended

/