Home Maxwell-Like Equations for Free Dirac Electrons
Article Publicly Available

Maxwell-Like Equations for Free Dirac Electrons

  • S.A. Bruce EMAIL logo
Published/Copyright: February 6, 2018

Abstract

In this article, we show that the wave equation for a free Dirac electron can be represented in a form that is analogous to Maxwell’s electrodynamics. The electron bispinor wavefunction is explicitly expressed in terms of its real and imaginary components. This leads us to incorporate into it appropriate scalar and pseudo-scalar fields in advance, so that a full symmetry may be accomplished. The Dirac equation then takes on a form similar to that of a set of inhomogeneous Maxwell’s equations involving a particular self-source. We relate plane wave solutions of these equations to waves corresponding to free Dirac electrons, identifying the longitudinal component of the electron motion, together with the corresponding Zitterbewegung (“trembling motion”).

1 Introduction

There have been many attempts to construct classical models for a spin-1/2 particle [1], [2], [3], [4], [5], [6], [7], [8], [9]. This interest arises from the need to more fully understand the somehow intriguing internal structure of the electron. In the Barut-Zanghi theory [9], the electron is characterised by a Lagrangian whose equations of motion exhibit a classical analogue to the phenomenon of Zitterbewegung (ZB). Based on these ideas, Salesi and Recami [10], [11], [12], [13] present a field theory of the spinning electron that constructs a classically intelligible description of the electron. Different approaches have also been proposed [14], [15], [16], [17], [18]. In particular, Campolattaro [19] shows that the Dirac equation is equivalent to Maxwell’s equations for the electromagnetic field generated by two currents: one electric in nature and one magnetic-monopolar.

This article gives a different representation to this problem. It is well known [20], [21], [22], [23], [24] that Maxwell’s equations can be written in spinor notation which resembles that used for the Dirac electron. Here, we want to proceed in the opposite direction, by writing down Maxwell-like equations for the free Dirac electron bispinor components. If this is really possible, we need to find (electric-like) vector and (magnetic-like) pseudo-vector fields describing the electron bispinor wavefunction in a way similar to electrodynamic formalism. At first sight, this appears quite reasonable in view of the fact that the eigenvalue of the electron velocity operator is ±c, which resembles a photon-like behaviour.

2 Maxwellian Description of a Free Dirac Electron

To begin with, we consider the well-known [25] basic free-field Lorentz scalar Dirac Lagrangian density from which the field equations are derived (hereafter ħ=c=1):

(1)(Ψ,Ψ¯,Ψ,Ψ¯)=i2(Ψ¯(x)γααΨ(x)(αΨ¯(x))γαΨ(x))+meΨ¯(x)Ψ(x),

where x=(t, x), with metric signature g(+−−−). Here, we choose the standard Dirac representation

(2)γ0=(I00I),γi=(0σiσi0),γ5=(0II0),

in which σi, i=1, 2, 3, are Pauli spin matrices. Varying ℒ with respect to Ψ¯=Ψγ0, we find the (Euler-Lagrange) Dirac equation for a free electron of mass me:

(3)(iγααme)Ψ(x)=0.

The Dirac bispinor Ψ can be written in terms of two-component spinors ϕ and χ:

(4)Ψ(x)=12(ϕ(x)χ(x))4,

where, in general, we can write

(5)ϕ(x)=ϕR(x)+iϕI(x),χ(x)=χR(x)+iχI(x),

with ϕR, I and χR, I spinors with real components. Let ψ denote either ϕ or χ. Thus, we define

(6)ψNM(t,x)=12(ψR(t,x)±ψR(t,x))+i2(ψI(t,x)±ψI(t,x)),

with complex spinors ψN and ψM symmetric and antisymmetric under space inversion x→−x, respectively.

If the spinor (4) describes either a positive-energy free electron or a negative-energy free electron, ϕ and χ must have opposite parity. Motivated by [26], [27], for a state of positive parity (an electron), we set

(7)ϕN(x)=(N3(x)+iUS(x)N1(x)+iN2(x)),χM(x)=i(M3(x)+iUP(x)M1(x)+iM2(x)),

where Mi, Nj, US, and UP, i=1, 2, 3, are eight independent real fields, which under space inversion transform as

(8)Mi(t,x)=Mi(t,x),Ni(t,x)=Ni(t,x),UP(t,x)=UP(t,x),US(t,x)=US(t,x).

The parity 𝒫 and the charge conjugation 𝒞 operators for the Dirac field Ψ are defined as usual:

(9)Ψ(t,x)Ψp(t,x)=γ0Ψ(t,x)=+Ψ(t,x),Ψ(t,x)ΨC(t,x)=γ2Ψ(t,x)=12(ϕC(t,x)χC(t,x)),

with ϕC(x)=σ2χM(x) and χC(x)=σ2ϕN(x).

Inserting (7) and (4) into (1), we can rewrite the Lagrangian density in the form

(10)(G,G˜,G,G˜)=12(qαGβα(x)γG˜γβ(x)qαG˜βα(x)γGγβ(x))+18me(Gβα(x)Gαβ(x)G˜βα(x)G˜αβ(x)),

where Gαβ(x) and G˜αβ(x) are second-rank tensors defined by

(11)Gαβ(x)=Kαβ(x)gαβUS(x),G˜αβ(x)=K˜αβ(x)gαβUP(x).

Here, Kαβ=gαγKγδgδβ is the analogue of the antisymmetric EM field-strength tensor whose components are

(12)K0j(x)=Mj(x),Kij(x)=ϵijkNk(x),

with K˜αβ=(1/2)ϵαβγδKγδ, where ϵαβγδ is the Levi-Civita pseudo-tensor (ϵ0123=1). The symmetric fields (in the indices α and β) gαβUS and gαβUP transform like tensor and pseudo-tensor, respectively. In (10), qα is a constant spacelike four-vector (it does not depend on the space-time coordinates) under restricted Lorentz transformations, which will be specified for a particular inertial reference frame below. Thus, the electromagnetic-like tensors (11) are no longer traceless:

(13)tr(Gαβ(x))=4US(x),tr(G˜αβ(x))=4UP(x).

We derive the (coupled) differential equations of motion for Mi, Nj, US, and UP from (3) by directly using the bispinor representation (7). After a straightforward algebra, we find a set of eight (formally) simple Maxwell-like equations:

(14)M(x)tUS(x)=meqN(x),×M(x)+tN(x)+UP(x)=me(q×N(x)qUS(x)),N(x)+tUP(x)=meqM(x),×N(x)+tM(x)US(x)=me(M(x)×q+qUP(x)).

In the above equations, we have chosen a reference frame where (qα)=(0, 0, 0, 1)=(0, q), which is consistent with the Dirac representation (2). From (3), we find that Fa={Mi, Nj, US, UP} also satisfies the Klein-Gordon equations:

(15)(αα+me2)Fa(x)=0.

Note that on similar grounds, the massless case is discussed in [26], [27].

Lorentz invariance of (14) can easily be checked as one does in the case of Maxwell’s equations for dyons [27]. From (14), we define electric and magnetic-like (self) four-currents

(16)4π(jMα(x))=me(qN(x),M(x)×q+qUP(x)),4π(jNα(x))=me(qM(x),q×N(x)qUS(x)),

where, as represented above, q transforms as the spatial component of a four-vector. Hence, we can write (14) in the covariant form

(17)αGαβ(x)=4πjMβ(x)=meqαG˜αβ(x),αG˜αβ(x)=4πjNβ(x)=meqαGαβ(x).

These equations of motion can be derived by varying the Lagrangian (10) with respect to the four-vectors χη=qαGηα and χ˜η=qαG˜ηα, so that (17) are deduced from the Euler-Lagrange equations

(18)μ(δδ(μχη))δδχη=0,μ(δδ(μχ˜η))δδχ˜η=0.

Here, we should emphasise that, in general, the following terms do not vanish

(19)μ(Gβ3δ(γG˜γβ)δ(μGη3))0μ(G˜β3δ(γGγβ)δ(μG˜η3)).

Note that the presence of US and UP (identified as scalar and pseudo-scalar fields, respectively) induces longitudinal components in M and N, since from (14) we can obtain the (Klein-Gordon) second-order differential equations

(20)2M2Mt2=4π(×JN+JMt+JM0)=me2M,2N2Nt2=4π(×JM+JNt+JN0)=me2N.

On the left-hand side of (20), the terms in US and UP vanish because [∇, ∂t]=0.

It is worth mentioning that jMβ(x) and jNβ(x) are, generally, not conserved four-currents as it is easily determined from (17). In fact

(21)4πβjMβ(x)=me2US(x),4πβjNβ(x)=me2UP(x).

However, from (15), we can define conserved four-currents by making

(22)4πjMβ(x)4πjMβ(x)βUS(x),4πjNβ(x)4πjNβ(x)βUP(x).

Starting with (17), it follows that the scalar product of jMβ(x) and jNβ(x) is the pseudo-scalar (invariant) quantity

(23)jMjNMN+USUP=inv.

3 Plane Wave Solutions and ZB

Next, we want to study plane wave solutions for (14). First, we choose a positive-energy solution of (3):

(24)ΨEp±(x)=me+Ep2Ep(u±(me+Ep)1σpu±)expiφp(x)4,

where

(25)u+=(10),u=(01),

for spin-up (down) wavefunctions, respectively, where φp(x)=p·xEpt is the invariant phase, with Ep=+me2+p2. Identifying ΨΨEp±, we get, for spin-up solutions,

(26)N1(x)=0=N2(x),N3(x)=Cpcosφp(x),US(x)=Cpsinφp(x),M1(x)=Cpme+Ep(p1sinφp(x)+p2cosφp(x)),M2(x)=Cpme+Ep(p1cosφp(x)p2sinφp(x)),M3(x)=Cpme+Epp3sinφp(x),UP(x)=Cpme+Epp3cosφp(x),

with Cp=(me+Ep)/Ep. Correspondingly, for spin-down solutions,

(27)N1(x)=Cpcosφp(x),N2(x)=Cpsinφp(x),N3(x)=0=US(x),M1(x)=Cpme+Epp3sinφp(x),M2(x)=Cpme+Epp3cosφp(x),M3(x)=Cpme+Ep(p1sinφp(x)p2cosφp(x)),UP(x)=Cpme+Ep(p1cosφp(x)+p2sinφp(x)),

where p→−p under space inversion x→−x. Thus, in both cases, jM·jN=0. For the latter case, M·N=0 in any inertial reference frame because US(x)=0.

In what follows, we regard a particularly interesting expression for the flux density associated with the Dirac field, as represented by (4) and (7). Consider a basic non-definite parity Dirac bispinor Φ as a linear superposition of positive- and negative-energy contributions:

(28)Φ(x)=12(Ψ(x)+ΨC(x)).

The corresponding flux density is given by

(29)S(x)=Φ(x)αΦ(x)=S(x)+SZB(x)=12{Ψ(x)αΨ(x)+ΨC(x)αΨC(x)}+12[Ψ(x)αΨC(x)+ΨC(x)αΨ(x)]={M(x)×N(x)+UP(x)N(x)US(x)M(x)}+[(M1M2+N1N2M3UPN3US)ϵ^1+(UP2+US2+2(M22+N22)MMNN)ϵ^2+(M2M3+N2N3+M1UP+N1US)ϵ^3].

Here, αi are the Dirac matrices and ϵ^i,i=1, 2, 3, is the usual Cartesian canonical basis. The flux S(x) contains two pieces: (i) a longitudinal component in curl bracket {…}, and (ii) a ZB component in squared bracket […]. Note that S(x) represents the analogue of the Poynting vector for the electron [27]. Let us consider the motion of a spin-down electron along the z axis (p1=0=p2). From either (24) or (26) together with (29), we find that

(30)S(z,t)=S(z,t)+SZB(z,t)=ϵ^3p3Ep3+ϵ^1meEp3sin2φp3(z,t)+ϵ^2meEp3cos2φp3(z,t).

Equation (30) shows that the electron motion follows, as one would expect, a helical path [1], [9], [10], with the radius of the circle decreasing as the energy increases (see Fig. 1). Thus, for Ep3, ZB fades out, although its frequency increases linearly with Ep3. Notice that, by inserting ħ and c into the phase 2φp3(z,t) in the rest frame system (p3=0), we get 2|tφp3(z,t)|=2mec2/, which is the ZB frequency.

Figure 1: Schematic graph (not to scale) of the flux density (30) when the electron is in the superposition state Φ(x) for spin-down solutions ΨEp3−(x).$\Psi _{{E_{{p_3}}}}^ - (x).$
Figure 1:

Schematic graph (not to scale) of the flux density (30) when the electron is in the superposition state Φ(x) for spin-down solutions ΨEp3(x).

4 Concluding Remarks

To summarise, we have derived Maxwell-like equations for free Dirac electrons starting with the associated Lagrangian density. For completeness, it was necessary to incorporate appropriate scalar and pseudo-scalar fields US and UP into the Dirac bispinor (4), according to both (7) and (8). We then studied plane wave solutions associated to this system. To this end, we defined a Dirac bispinor as a linear superposition of positive- and negative-energy contributions obtaining the corresponding flux density, which exhibits a helical trajectory due to the electron ZB, as depicted in Figure 1.

A natural ensuing task is to attempt (second) quantisation of the free fields (11) which satisfy the equations of motion (14). Here, we have to follow, as much as possible, the quantisation procedure we know for the electron case, considering that the occupation number of a particular electron state is at most one. This problem will be treated in a future article.

Finally, note that (to some extent conversely) for a realistic, classical version of ZB, see [28], where this “trembling” motion at constant speed c is used to derive the Schrödinger equation. For the physical origin of ZB as a due to self-reaction, see [29], [30], [31]. It is noteworthy to mention that ZB of a free particle has never been directly observed [32], [33], [34], [35]. For one author [36], ZB is not even observable. Nonetheless, it has been already simulated. First, it was simulated with a trapped ion, by putting it in an environment such that the non-relativistic Schrödinger equation for the ion has a similar mathematical form as the Dirac equation (although the physical condition is different) [32], [33], [34]. Second, it was simulated in a configuration with Bose-Einstein condensates [37]. Further research could lead to applications of (14) in these and some other physical systems whose Hamiltonians are listed in [33].

References

[1] K. Huang, Am. J. Phys.20, 479 (1952).10.1119/1.1933296Search in Google Scholar

[2] M. Bunge, Nuovo Cimento 1, 977 (1955).10.1007/BF02731409Search in Google Scholar

[3] F. Gursey, Nuovo Cimento 5, 784 (1957).10.1007/BF02903205Search in Google Scholar

[4] H. C. Corben, Phys. Rev. 121, 1833 (1961).10.1103/PhysRev.121.1833Search in Google Scholar

[5] B. Liebowitz, Nuovo Cimento A63, 1235 (1969).10.1007/BF02754933Search in Google Scholar

[6] D. Gutkowski, M. Moles, and J. P. Vigier, Nuovo Cimento B39, 193 (1977).10.1007/BF02738189Search in Google Scholar

[7] A. O Barut, Z. Naturforsch. 33a, 993 (1978).10.1515/zna-1978-0819Search in Google Scholar

[8] W. A. Rodrigues, J. Vaz, and E. Recami, Found. Phys. 23, 459 (1993).10.1007/BF01883725Search in Google Scholar

[9] A. O. Barut and N. Zanghi, Phys. Rev. Lett. 52, 2009 (1984).10.1103/PhysRevLett.52.2009Search in Google Scholar

[10] G. Salesi and E. Recami, Phys. Lett. A190, 137 (1994).10.1016/0375-9601(94)90066-3Search in Google Scholar

[11] M. Pavsic, E. Recami, and W. Rodrigues Jr., Phys. Lett. B318, 481 (1993).10.1016/0370-2693(93)91543-VSearch in Google Scholar

[12] W. Rodrigues Jr and J. Vaz Jr., Phys. Lett. B318, 623 (1993).10.1016/0370-2693(93)90464-SSearch in Google Scholar

[13] J. Vaz Jr. and W. Rodrigues Jr., Phys. Lett. B319, 203 (1993).10.1016/0370-2693(93)90803-PSearch in Google Scholar

[14] M. Pavsic, Class. Quantum Grav. 7, L187 (1990).10.1088/0264-9381/7/8/006Search in Google Scholar

[15] A. O. Barut and M. Pavsic, Phys. Lett. B216, 297 (1989).10.1016/0370-2693(89)91118-0Search in Google Scholar

[16] D. Hestenes, Found. Phys. 15, 63 (1985).10.1007/BF00738738Search in Google Scholar

[17] D. Hestenes, Found. Phys. 20, 1213 (1990).10.1007/BF01889466Search in Google Scholar

[18] D. Hestenes, Found. Phys. 40, 1 (2010).10.1007/s10701-009-9360-3Search in Google Scholar

[19] A. Campolattaro, Int. J. Theor. Phys. 29, 141 (1990).10.1007/BF00671324Search in Google Scholar

[20] R. H. Good, Phys. Rev. 105, 1914 (1957).10.1103/PhysRev.105.1914Search in Google Scholar

[21] I. Saavedra, Prog. Theoret. Phys. 50, 1006 (1973).10.1143/PTP.50.1006Search in Google Scholar

[22] J. R. Oppenheimer, Phys. Rev. 38, 725 (1931).10.1103/PhysRev.38.725Search in Google Scholar

[23] H. E. Moses, Phys. Rev. 113, 1670 (1959).10.1103/PhysRev.113.1670Search in Google Scholar

[24] W. Greiner, Relativistic Quantum Mechanics, Springer-Verlag, Berlin 1994, p. 304.10.1007/978-3-642-57974-5Search in Google Scholar

[25] C. Itzykson and J. Zuber, Quantum Field Theory, McGraw-Hill, NY 1980, p. 143.Search in Google Scholar

[26] S. A. Bruce and P. Montenegro, Proceedings of the XXII International Colloquium on Group Theoretical Methods in Physics (ICGTMP), Hobart, Australia (International Press, 1998) p. 490.Search in Google Scholar

[27] S. A. Bruce, Z. Naturforsch. 72a, 885 (2017).10.1515/zna-2017-0079Search in Google Scholar

[28] G. Cavalleri and G. Mauri, Phys. Rev. B41, 6751 (1990).10.1103/PhysRevB.41.6751Search in Google Scholar

[29] G. Cavalleri and A. Zecca, Phys. Rev. B43, 3223 (1991).10.1103/PhysRevB.43.3223Search in Google Scholar PubMed

[30] G. Cavalleri, Nuovo Cimento B112, 1193 (1997).Search in Google Scholar

[31] G. Cavalleri and E. Tonni, in Late Proceedings of VII Int. Conf. on “Physical Interpretation of Relativity Theory” (London, 2000) ed. by M. C. Duffy (University of Suderland, UK) p. 52.Search in Google Scholar

[32] C. Wunderlich, Nature 463, 37 (2010).10.1038/463037aSearch in Google Scholar PubMed

[33] J. Cserti and G. Dávid, Phys. Rev. B74, 172305 (2006).10.1103/PhysRevB.74.172305Search in Google Scholar

[34] W. Zawadzki and T. M. Rusin, J. Phys. Condens. Matter 23, 143201 (2011) (a review).10.1088/0953-8984/23/14/143201Search in Google Scholar PubMed

[35] T. M. Rusin and W. Zawadzki, Physica E70, 1 (2015).10.1016/j.physe.2015.02.013Search in Google Scholar

[36] R. F. O’Connell, Mod. Phys. Lett. A26, 469 (2011).10.1142/S0217732311034967Search in Google Scholar

[37] L. J. LeBlanc, M. C. Beeler, K. Jiménez-García, A. R. Perry, S. Sugawa, et al., New J. Phys. 15, 073011 (2013).10.1088/1367-2630/15/7/073011Search in Google Scholar

Received: 2017-09-15
Accepted: 2018-01-12
Published Online: 2018-02-06
Published in Print: 2018-03-28

©2018 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 28.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/zna-2017-0328/html
Scroll to top button