Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

No Proca Photons

F. Steininger1,2 , T. B. Mieling1,2 , and P. T. Chruściel2


1
University of Vienna, Faculty of Physics, Vienna Doctoral School in Physics, Boltzmanngasse 5, 1090 Vienna, Austria
2 University of Vienna, Faculty of Physics and Research platform TURIS, Boltzmanngasse 5, 1090 Vienna, Austria
arXiv:2212.12408v2 [physics.class-ph] 13 Mar 2023

March 14, 2023

Abstract
We show that the Proca equation in vacuum, as well as its plausible modifications in dielectric
media, is incompatible with experimental evidence, no matter how small the Proca mass is.

1 Introduction
Both Maxwell’s equations and the standard model of particle physics assume photons to be massless. The
question whether this is so remains intriguing, as mechanisms have been described [1, 2] which provide
photons with an effective mass. However, a non-zero photon mass would imply corrections to Coulomb’s
inverse-square law, and would challenge the current picture of photons being gauge bosons, see, e.g.,
[2, 4, 7] and references therein.
The standard modification of the microscopic Maxwell equations to include a photon mass µ > 0 is
given by Proca’s equation

µ2 c2
η αβ ∂α Fβγ − A = −ηγα j α , with Fαβ = ∂α Aβ − ∂β Aα , (1)
~2 γ
where ηαβ = diag(−1, +1, +1, +1) is the Minkowski metric and j α the electric current, which is assumed
to satisfy the continuity equation, ∂α j α = 0. Henceforward, we will set ~ and c equal to 1.
Experiments testing this equation have produced more and more stringent upper bounds on the
photon mass µ, with the current Particle Data Group bound [8] being µ < 10−18 eV/c2 ; compare [1] for
a discussion of more stringent bounds based on astrophysical considerations.
While it is generally assumed that a small photon mass µ in (1) produces only small deviations from
solutions to Maxwell’s equations [4], we show that this is not the case. In particular, we show that (1)
implies matching conditions at material interfaces which deviate significantly from those of Maxwell’s
equations. Considering coaxial cables as an example, we show the following result:
Assuming charge conservation, and that the field strength contains no Dirac-delta distributions, the
Proca equation (1) in a coaxial cable consisting of vacuum enclosed by ideal conductors rules out the
existence of transverse electromagnetic (TEM) modes, no matter how small µ is.
The condition that Fαβ contains no Dirac-delta distributions is motivated by the requirement of a
well-defined energy-momentum tensor, which would otherwise contain squares of Dirac-deltas.
We further demonstrate that the same result remains valid for two natural generalizations of equation
(1) to linear isotropic dielectrics.
The argument proceeds by showing that µ 6= 0 in (1) implies an inequality, which forbids TEM modes,
between the propagation constant of the electromagnetic field and its frequency. It should be kept in
mind that the TEM modes are the main propagation modes in coaxial cables.
This leaves one with two options of either discarding Proca’s equation, or calling into question the
assumptions leading to these results. While we indicate below a possible modification of this equation
which restores TEM modes at the cost of introducing a preferred vector field associated with the symmetry
axis of the coaxial cable, no such modification is available in the absence of spatial symmetry.

1
2 Proca Modes in Coaxial Cables
Recall that coaxial cables are composed of a solid copper core, surrounded by a dielectric insulator, which
is in turn surrounded by woven copper shielding, and covered by a plastic sheath. Copper is modelled as
a perfect conductor, meaning that the electromagnetic field has to vanish within the metal. The field is
confined between the conductors, hence no field is present in the sheath and ambient air. The dielectric’s
refractive index is often close to unity, so that the region between the core and shielding can be modelled
as a vacuum as far as the electromagnetic properties are concerned. We also make this assumption in
this section, but allow for non-trivial refractive indices in the next section.
To proceed, we start by determining the boundary conditions satisfied by the fields. By definition,
the field Fαβ vanishes inside ideal conductors. There is an induced current j α which is non-zero only on
the interfaces. Thus, by (1), Aα vanishes within the conductors as well.
Let us show that this, together with our hypotheses listed above, implies continuity of Aα at the
interface. For this, let us denote by U × R ⊂ R3 the vacuum region between the conductors, where
U ⊂ R2 is an annulus with interior radius r1 and exterior radius r2 , and let ∂U be the boundary of
U . In cylindrical coordinates (t, r, θ, z), the requirement that Fαβ contains no Dirac-delta distributions
immediately implies continuity of At , Aθ and Az at ∂U .
To show continuity of Ar , note that (1) implies the “Lorenz-gaugelike” constraint

η αβ ∂α Aβ = 0 , (2)

by anti-symmetry of Fαβ and ∂α j α = 0. Since the remaining components have already been shown to
be continuous at the interfaces, this can only hold if Ar is continuous as well.
Thus, the boundary condition induced by ideal conductors is

Aα |∂U = 0 . (3)

Proca’s equation (1) can be equivalently restated as a wave equation

(η − µ2 )Aα = −ηαβ j β , (4)

keeping in mind that we seek solutions satisfying (2). Any solution of (4) can be Fourier-decomposed as
a sum of fields of the form
Aα = Ãα (x, y)ei(βz−ωt) , (5)
leading to the following boundary-value problem for Ãα :

−∆⊥ Ãα = ξ Ãα in U , (6)


Ãα = 0 on ∂U , (7)

with
ξ := ω 2 − µ2 − β 2 , (8)
and where ∆⊥ is the two-dimensional Laplacian on the annulus U . We see that ξ belongs to the spectrum
of the operator −∆⊥ with vanishing Dirichlet data on ∂U . It is a standard fact that this spectrum is a
discrete subset of (0, ∞); let ξ0 > 0 denote the smallest eigenvalue. For any non-trivial solution of (6)–(7)
with µ > 0 we have
ω 2 − β 2 = µ2 + ξ > ξ0 =⇒ ω 2 > β 2 + ξ0 , (9)
with the last inequality holding independently of µ as soon as µ 6= 0. Since the observed TEM modes
satisfy
ω = ±β (10)
within the precision of measurements, and since ξ0 > 0, we see that (9) can only be consistent with
experiment if ξ0 itself is not larger than the inaccuracy in determining ω 2 and β 2 . Now, assuming that
the inner and outer radii of the conductors are of the same order of magnitude, elementary scaling shows
that the smallest eigenvalue ξ0 is of order r1−2 . For example, a numerical approximation for the smallest
eigenvalue ξ0 for r2 = 2r1 is
ξ0 ≈ 9.75332 r1−2 , (11)
which makes clear the incompatibility of (9) with existence of TEM modes.
Our no-go results can be compared with the discussion in [3], where an upper bound on the contribu-
tion of the mass of the photon to the perturbation of the Maxwell field has been claimed. The estimates

2
there are not consistent with our analysis, which shows that some modes disappear completely. (It follows
from (8) that the surviving ones acquire a phase change of order ∼ µ2 L/β|µ=0 , where L is the length of
a coaxial cable and β|µ=0 is the propagation constant for µ = 0.) The discrepancy can be traced back to
the fact that the analysis of [3] overlooks the contribution of boundary currents, which are essential in
the problem at hand.

3 Extension to Dielectric Coaxial Cables


So far we considered wave propagation in vacuum, and the question arises whether the considerations
above are affected in the presence of dielectric media. For this, recall that the Proca equation (1) is
variational, with Lagrangian

LProca = − 41 η αβ η γδ Fαγ Fβδ − 21 µ2 η αβ Aα Aβ + j α Aα . (12)

To extend the theory to macroscopic electrodynamics in linear isotropic media, we orient ourselves on
Maxwell’s equations which, for such media, are derived from the Lagrangian
1 αβ γδ
LMaxwell = − 4µ γ γ Fαγ Fβδ + j α Aα , (13)

where γ αβ is Gordon’s optical metric [5]

γ αβ = η αβ + (1 − n2 )uα uβ , (14)

with uα being the four-velocity of the medium and n = µε the refractive index expressed in terms
of the permittivity ε and permeability µ. To pass to a theory of massive photons for macroscopic
electrodynamics, we thus consider the Lagrangian

L = − 4µ γ γ Fαγ Fβδ − 12 µ2 mαβ Aα Aβ + j α Aα ,


1 αβ γδ
(15)

where a choice of the “mass matrix” mαβ has to be made. Obvious candidates here are either η αβ , or
γ αβ . We will see that, as before, neither of them allows TEM modes if µ 6= 0. Anticipating, choosing a
specific projection matrix as mαβ will allow for such modes, but at the cost of introducing a preferred
spatial direction in the equations.
In regions where µ is constant, the Lagrangian (15) leads to the following field equations:

∂α (γ αρ γ βσ Fρσ ) − µµ2 mαβ Aα = −µj β . (16)

In order not to clutter the notation we will from now assume that µ = 1; alternatively, µ can be absorbed
in a redefinition of µ and of j β .
Clearly, there is again a constraint obtained by taking the divergence of (16),

∂α (mαβ Aβ ) = 0 , (17)

now “Gordon-Lorenz-gaugelike”. Additionally, if det mαβ 6= 0, the same arguments as in Section 2 show
that a) Aα ≡ 0 in the conductor region, and b) the boundary condition Aα |∂U = 0 must hold.
Let us start the analysis with the simplest case, which turns out to be mαβ = γ αβ . In regions of
constant n, Equation (16) can then be equivalently rewritten as

(γ − µ2 )Aα = −γαβ j β . (18)

Using the Fourier modes (5), one is again led to a spectral problem

−∆⊥ Ãα = ξ Ãα in U , (19)


Ãα = 0 on ∂U , (20)

except that now we have ξ = n2 ω 2 − µ2 − β 2 .


In the presence of a dielectric, the Maxwell TEM mode propagates with phase velocity vp = n−1 ,
corresponding to nω = ±β at the level of experimental precision. Since (19)–(20) is of the same form as
(6)–(7), an identical argument as in Section 2 leads to the conclusion that µ 6= 0 is incompatible with
experiment.

3
The case of mαβ = η αβ requires more work. The arguments given so far lead to the following
boundary-value problem for each Fourier mode (5):

−∆⊥ Ã0 = ξ Ã0 , (21)


−∆⊥ div à = ξ div à , (22)
2
−∆⊥ Ãi = ξ̃ Ãi + (n − 1)∂i div à , (23)

in U and
Ãα = 0 on ∂U , (24)
with

div à = ∂x Ãx + ∂y Ãy + iβ Ãz , (25)


2 2 −2 2
ξ =ω −β −n µ , (26)
2 2 2 2
ξ̃ = n ω − β − µ . (27)

If Ã0 or div à does not vanish, we obtain a contradiction as before based on the spectral problem
associated with (21) or (22), respectively. The alternative is that Ã0 ≡ 0 ≡ div Ã, in which case (23)
becomes
− ∆⊥ Ãi = ξ˜Ãi , (28)
and a contradiction with existence of TEM modes is obtained from the inequalities

n2 ω 2 − β 2 = µ2 + ξ˜ > ξ0 =⇒ n2 ω 2 > β 2 + ξ0 , (29)

with ξ0 as in (9).
Summarizing, we have shown that the Proca field equations are not compatible with TEM modes in a
coaxial cable, whether in vacuum or in dielectric media, contradicting the fact that TEM modes constitute
the main propagating mode in coaxial cables. Thus, neither the Proca equation, nor its variation involving
the Gordon metric, correctly describe photons with small non-zero mass in this setting.
Our analysis provides yet another piece of evidence that photons are massless. The question then
arises, whether there exist an equation which sidesteps the problem raised above. The requirement of
Lorentz covariance allows for modifications of the Proca Lagrangian which involve only the Lorentz metric
and the four-velocity uα of the medium. For reasons that will become clear shortly, we also need a unit
vector field ℓα , orthogonal to uα and directed along the axis of symmetry of the medium, and consider
the Lagrangian (15) with mαβ given by

mαβ = η αβ + puα uβ − qℓα ℓβ , (30)

for some constants p and q to be determined. This is the most general form of mαβ which is compatible
with Lorentz invariance, time-reversal symmetry, the existence of a preferred four-velocity determined
by the rest frame of the medium, and the existence of a preferred spacelike direction determined by the
geometry of the medium.
In regions where ∂σ uα , ∂σ ℓα abd ∂σ n vanish, the resulting field equations read

γ αβ ∂α Fβγ − µ2 Aα mαβ γβγ = 0 . (31)

A repetition of the arguments above shows that, if µ 6= 0, the vanishing of Fαβ in the conducting region
implies Aα mαβ γβγ = 0 there. The contradictions reached above can only be avoided if we can find field
configurations for which Aα 6≡ 0 in the conducting region. This will be the case if and only if the matrix
 
1−p
mαβ γβγ = δγα + 1 − uα uβ ηβγ − q ℓα ℓβ ηβγ (32)
n2

is singular. The vanishing of the determinant of mαβ γβγ is equivalent to (p − 1)(q − 1) = 0. We show
in an Appendix that inconsistencies arise again unless p = 1 = q, which leads to an equation which does
now admit modes approximating the observed TEM ones:

γ αβ ∂α Fβi = 0 with i ∈ {0, 3} , (33)

and
γ αβ ∂α Fβj − µ2 Aj = 0 with j ∈ {1, 2} . (34)

4
In such a theory there is a residual gauge freedom Aµ → Aµ + ∂µ λ, where λ is an arbitrary function of t
and z, and only the transverse components of the field acquire a mass.
But, none of the equations above are compatible with experimental evidence if µ 6= 0, no matter how
small, without introducing extra structures in the equation.
One might wonder whether an alternative solution could be found by introducing in (15) a new optical
metric
γ αβ = η αβ + σuα uβ + τ ℓα ℓβ , (35)
with constants σ and τ restricted to a range so that γ αβ is Lorentzian, to guarantee a well-posed Cauchy
problem. But a linear rescaling of the coordinates together with obvious field redefinitions reduces
the resulting equations to the ones already analyzed above, leading to the same incompatibilities with
experiment.

4 Conclusions
We have shown that a non-zero photon mass µ in the Proca equation implies interface conditions for the
electromagnetic field which deviate strongly from those of Maxwell’s equations, irrespective of the mag-
nitude of µ. Combined with the assumptions listed in the Introduction, Proca’s equation thus predicts
an absence of TEM modes in coaxial cables, which are both predicted to exist by Maxwell’s equations
and observed in experiments. From this, we conclude that the following statements are incompatible: (i)
Proca’s equation holds, (ii) charge conservation holds, (iii) the electromagnetic field has an everywhere
well-defined energy-momentum density, (iv) ideal conductors are admissible idealizations in electrody-
namics.
It was argued in [3, 6] that the dispersion relation of the TEM mode is insensitive to the photon mass
µ. Our results here prove rigorously instead that the observed existence of TEM modes is incompatible
with the Proca equation.

Appendix
In this appendix we show that TE modes, a special case of which are the TEM modes, are incompatible
with (31) unless p = q = 1. For this, recall that the TE modes of Maxwell’s equations satisfy Ez =
−i(β Ã0 + ω Ãz )ei(βz−ωt) ≡ 0 in U . Hence, a solution of the Proca equation which approximates the
observed Maxwell TEM solutions has to satisfy

β Ã0 + ω Ãz = f , (36)

where f is a function of µ and all coordinates which is not larger than the experimental precision, say
f = O(ǫ). Using f , the system (31) can be written as

ˆ Ã = iq∂ f ,
(−∆⊥ − ξ) (37)
0 z
ˇ Ã = −i(1 − p − n2 )∂ f ,
(−∆⊥ − ξ) (38)
z 0
2
(−∆⊥ − ξ)Ãx = −i∂x (qβ Ãz − (1 − p − n )ω Ã0 ) , (39)
2
(−∆⊥ − ξ)Ãy = −i∂y (qβ Ãz − (1 − p − n )ω Ã0 ) , (40)

where ξ is as in (26) and

ξˆ = ω 2 (1 − p) − β 2 (1 − q) − µ2 n−2 (1 − p) , (41)
ξˇ = ω 2 (1 − p) − β 2 (1 − q) − µ2 (1 − q) . (42)

Take now p = 1 and q 6= 1. With this mass matrix mαβ we still need Ãi |∂U = 0 but now Ã0 |∂U can
be arbitrary. Then (38) implies that Ãz = O(ǫ) unless ξˇ is an eigenvalue. In this last case we obtain a
contradiction as in the main body of this paper, thus it must be that Ãz = O(ǫ) holds. But then (36)
shows that Ã0 = O(ǫ) as well. It follows that the source terms in (39)-(40) are O(ǫ), and the whole
solution will be O(ǫ) unless ξ is an eigenvalue. An inconsistency with experiment is obtained by (9).
The treatment for q = 1 but p 6= 1 is analogous. This establishes our claim.

5
Acknowledgements
Research supported in part by the Austrian Science Fund (FWF), Project P34274 and Grant TAI 483-N,
as well as the Vienna University Research Platform TURIS. TBM is a recipient of a DOC Fellowship of
the Austrian Academy of Sciences at the Faculty of Physics of the University of Vienna and acknowledges
partial support from the Vienna Doctoral School in Physics (VDSP).

References
[1] E. Adelberger, G. Dvali, and A. Gruzinov, Photon-Mass Bound Destroyed by Vortices, Phys. Rev. Lett.
98 (2007), no. 1, 010402.
[2] L. Bonetti, L.R. dos Santos Filho, J.A. Helayël-Neto, and A.D.A.M. Spallicci, Effective photon mass
by Super and Lorentz symmetry breaking, Physics Letters B 764 (2017), 203–206.
[3] A.S. Goldhaber and M.M. Nieto, How to Catch a Photon and Measure its Mass, Phys. Rev. Lett. 26
(1971), no. 22, 1390–1392.
[4] , Photon and graviton mass limits, Reviews of Modern Physics 82 (2010), no. 1, 939–979.
[5] W. Gordon, Zur Lichtfortpflanzung nach der Relativitätstheorie, Annalen der Physik 377 (1923),
no. 22, 421–456.
[6] N.M. Kroll, Theoretical Interpretation of a Recent Experimental Investigation of the Photon Rest
Mass, Phys. Rev. Lett. 26 (1971), no. 22, 1395–1398.
[7] L. Tu, J. Luo, and G.T. Gillies, The mass of the photon, Reports on Progress in Physics 68 (2005),
no. 1, 77–130.
[8] R. L. Workman et al., Review of Particle Physics, PTEP 2022 (2022), 083C01.

You might also like