We review the formulation of a Lorentz-covariant bispinorial wave function and wave equation for a single photon on a flat background. We show the existence of a 10-dimensional set of conservation laws for this equation, and prove that 8 of these can be used to obtain global, gauge-invariant, ADM-like quantities that together define a covariantly constant self-dual bispinor.
Citation: Michael K.-H. Kiessling, A. Shadi Tahvildar-Zadeh. Noetherian conservation laws for photons[J]. Communications in Analysis and Mechanics, 2025, 17(2): 606-629. doi: 10.3934/cam.2025024
[1] | Sergey A. Rashkovskiy . Quantization of Hamiltonian and non-Hamiltonian systems. Communications in Analysis and Mechanics, 2023, 15(2): 267-288. doi: 10.3934/cam.2023014 |
[2] | Long Ju, Jian Zhou, Yufeng Zhang . Conservation laws analysis of nonlinear partial differential equations and their linear soliton solutions and Hamiltonian structures. Communications in Analysis and Mechanics, 2023, 15(2): 24-49. doi: 10.3934/cam.2023002 |
[3] | Sixing Tao . Lie symmetry analysis, particular solutions and conservation laws for the dissipative (2 + 1)- dimensional AKNS equation. Communications in Analysis and Mechanics, 2023, 15(3): 494-514. doi: 10.3934/cam.2023024 |
[4] | Andrea Brugnoli, Ghislain Haine, Denis Matignon . Stokes-Dirac structures for distributed parameter port-Hamiltonian systems: An analytical viewpoint. Communications in Analysis and Mechanics, 2023, 15(3): 362-387. doi: 10.3934/cam.2023018 |
[5] | Anthony Bloch, Marta Farré Puiggalí, David Martín de Diego . Metriplectic Euler-Poincaré equations: smooth and discrete dynamics. Communications in Analysis and Mechanics, 2024, 16(4): 910-927. doi: 10.3934/cam.2024040 |
[6] | Xueqi Sun, Yongqiang Fu, Sihua Liang . Normalized solutions for pseudo-relativistic Schrödinger equations. Communications in Analysis and Mechanics, 2024, 16(1): 217-236. doi: 10.3934/cam.2024010 |
[7] | Floyd L. Williams . From a magnetoacoustic system to a J-T black hole: A little trip down memory lane. Communications in Analysis and Mechanics, 2023, 15(3): 342-361. doi: 10.3934/cam.2023017 |
[8] | Sergey A. Rashkovskiy . Nonlinear Schrödinger equation with a short-range compensating field. Communications in Analysis and Mechanics, 2025, 17(2): 520-549. doi: 10.3934/cam.2025021 |
[9] | Antoine Bendimerad-Hohl, Ghislain Haine, Laurent Lefèvre, Denis Matignon . Stokes-Lagrange and Stokes-Dirac representations of $ N $-dimensional port-Hamiltonian systems for modeling and control. Communications in Analysis and Mechanics, 2025, 17(2): 474-519. doi: 10.3934/cam.2025020 |
[10] | Velimir Jurdjevic . Time optimal problems on Lie groups and applications to quantum control. Communications in Analysis and Mechanics, 2024, 16(2): 345-387. doi: 10.3934/cam.2024017 |
We review the formulation of a Lorentz-covariant bispinorial wave function and wave equation for a single photon on a flat background. We show the existence of a 10-dimensional set of conservation laws for this equation, and prove that 8 of these can be used to obtain global, gauge-invariant, ADM-like quantities that together define a covariantly constant self-dual bispinor.
Of central importance to the modern development of the theory of partial differential equations has been the notion of conservation laws and their connection with continuous symmetries of the equations, as developed in utmost generality by E. Noether in her two celebrated theorems of 1918 [1]. Noether's two theorems are so general and all-encompassing, and in their theoretical significance so ahead of their time that, more than fifty years after their discovery, manuscripts were still being submitted and published that purported to be "a generalization of Noether's theorem", while in reality they were special cases* [3]. The search for such conservation laws, whether exact or approximate, whether local or global or even microlocal, and the a priori estimates they imply for solutions of hyperbolic (or more generally, evolutionary) partial differential equations, have been an idée fixe in the works of many of the giants of this field, chief among them F. John, C. Morawetz, W. Strauss, S. Klainerman, D. Christodoulou, J. Shatah, and T. C. Sideris.
*For a fascinating account of the reception of Noether's theorems by the mathematics and physics communities, and their influence on the development of these sciences, see [2].
Indeed, a recurrent theme in many of Sideris's seminal works in hyperbolic PDEs is how to compensate for the lack of a particular symmetry in a system, whose associated conservation law—had it been available—would have made the deriving of an a priori estimate straightforward by appealing to previously known results. The case in point is the pioneering work of Klainerman [4], in which he obtained the dispersive estimates that are now named after him, for solutions of the linear wave equation on a flat background. Klainerman made use of local conservation laws associated with the full 15-dimensional group of conformal isometries of 3+1 dimensional Minkowski space, which is generated by 4 translations, 3 spatial rotations, 3 boosts, 4 inverted translations, and one scaling vectorfield. He used them all because they were all there, not because they were all absolutely necessary. Sideris was among the first to realize that one could get away with fewer symmetries. In his work on the Klein-Gordon equation [5] he noted that the scaling is no longer available, and showed that boosts can be used instead. Later, in his collaboration with Klainerman on equations of elastodynamics [6], the reverse was the case: boosts were not available, so scaling had to be used, combined with an ingenious use of weighted norms. This observation in turn became the cornerstone in Sideris's celebrated works on nonlinear elastodynamics [7] and [8], published in Inventiones and the Annals, respectively, as well as the path-breaking work on systems of wave equations with multiple speeds [9].
It is worth noting that, the oft-repeated phrase about "Noether's Theorem"† being a result in classical field theory stating that "isometries of spacetime give you conservation laws, for Lagrangians depending on the fields and their first derivatives, " is but a cartoon version of Noether's actual result, which concerns itself with systems of partial differential equations in any number of dependent and independent variables and any number of derivatives, without the need for the existence of a metric for the domain or any notion of isometry. The notion of "symmetry" developed by Noether was likewise completely general, allowing for symmetries acting only on the independent variables (i.e. symmetries of the domain), or only on the dependent variables (so-called target symmetries), as well as those with a combined action on the Cartesian product of the domain and the target. The latter play a key role in Sideris's works on elastodynamics, where the dynamical object is the vector-valued mapping between the "undeformed" reference state of a solid and its "deformed" state in physical space. Thus, Sideris's extensive use of simultaneous rotations is an instance of such "product" Noether currents.
†When used in singular form, what is meant most often is Noether's First Theorem.
Another key situation arises in tensorial equations, in which the target space is a vector or principal bundle over the domain manifold. This is the realm of Noether's Second Theorem, which concerns itself with what nowadays are called "gauge symmetries." These lie at the heart of today's fundamental physical theories, such as the Standard Model of particle physics. For a careful expository account of Noether's work using modern terminology that includes the important case of Lagrangian theory for sections of vector bundles, see Christodulou's beautiful monograph [10].
It is a fair question to ask, what to do when there are no symmetries available? This is most famously the case when one studies Einstein's equations of general relativity as a system of PDEs, where the unknown is the spacetime itself, which—when not vacuum—has no a priori reason to have any symmetries. For the important subset of asymptotically flat spacetimes, however, there is a notion of symmetry available "at infinity." It was already known to Einstein [11] and Weyl [12], and later Landau and Lifschitz [13], that such spacetimes will have global conserved quantities, and it was shown by Arnowitt, Deser, and Misner [14] that these are in the form of integrals over the "sphere at infinity" of a spatial slice, and that they are independent of the asymptotic time variable. For a rigorous derivation of the ADM conservation laws for Einstein's equations using Noether's Theorems see [15].
Our goal in this paper was to obtain a set of gauge-independent conservation laws for a particular quantum mechanical equation that we have previously proposed [16] as the relativistic equation of evolution for the wave function of a single photon, i.e. a massless spin-one particle with no longitudinal modes. Our hope was to obtain these on a spacetime background with no assumed symmetries. One can see that the requirement of gauge-independence limits one at best to global quantities that can be defined on an asymptotically flat spacetime, in analogy to the ADM quantities of General Relativity. Further restrictions coming from the need for the existence of covariantly constant tensorfields at the moment seem to limit our conclusions to flat spacetimes, yet we end up with new global conserved quantities for our photon wave equation on Minkowski space. We believe this is further evidence that, nearly 100 years after their discovery, Noether's Theorems continue to remain relevant to mathematical physics and be a source of inspiration (and surprise) to its practitioners, just as they have been to Sideris and many others before him.
Of central importance to the relativistic formulation of quantum mechanics in d space dimensions is the spacetime algebra A, defined as the complexification of the real Clifford algebra Cl1,d(R) associated with the Minkowski quadratic form of signature (+,−,…,−). For d=3, this algebra is easily seen to be isomorphic to the algebra of 4×4 complex matrices, as well as the algebra of 2×2 matrices with complex quaternionic entries:
A:=Cl1,3(R)C≅M4(C)≅M2(P)≅P⊗P, |
where P:=C⊗H is the algebra of complex quaternions, also known as the Pauli algebra [17]. (All tensor products are over C unless otherwise noted.)
The isomorphism of A with M4(C) can be realized by choosing a basis for Cl1,3(R) and taking complex linear combinations of the basis elements. A convenient basis for the real algebra is formed by the so-called Dirac gamma-matrices (in their Weyl representation) and their products: Let 1n denote the n×n identity matrix, and define
γ0=(012120),γk=(0−σkσk0), k=1,2,3, | (2.1) |
where the σk∈{1,2,3} are the three conventional Pauli matrices. The γ-matrices satisfy the Clifford algebra relations
γμγν+γνγμ=2ημν14; | (2.2) |
where the ημν are the components of the Minkowski metric tensor
η=diag(1,−1,−1,−1). | (2.3) |
Any Clifford algebra A associated with a vector space V over a field F contains a subspace that is isomorphic to V. The elements of that subspace are called 1-vectors. Note that (2.2) simply states that the four matrices {γμ}3μ=0 form a Lorentz-orthonormal set of 1-vectors in the Clifford algebra Cl1,3(R).
By definition, a k-vector is the (Clifford) product of k elements, each one of which is a 1-vector. Let {ej}nj=1 be a basis for V. Every Clifford number a∈A has a k-vector expansion of the form
a=aS1+n∑k=1∑1≤i1<⋯<ik≤nai1…ikei1…eik, | (2.4) |
where n=dimFV and the coefficients aS,ai1…ik are in F. It follows that the following is a basis for the (16-dimensional) algebra Cl1,3(R):
B:={14;γ0,γ1,γ2,γ3;γ0γ1,γ0γ2,…;γ0γ1γ2,…;γ0γ1γ2γ3}, | (2.5) |
and therefore its complexification can be obtained by taking the coefficients of the expansion to be complex numbers: A=SpanCB.
The complexified algebra A in particular includes the pseudoscalar
γ5:=iγ0γ1γ2γ3=(12 00−12), | (2.6) |
and therefore the projections
Π±:=12(14±γ5). | (2.7) |
Using these projections, it follows right away that A contains all 4×4 matrices, and it is easy to verify that A and M4(C) are indeed isomorphic as algebras, with the Clifford multiplication given by matrix multiplication.
Let a=aS1+∑IaIγI denote the k-vector expansion of a∈A. Thus, aS,aI∈C, and each γI is a k-fold product of gamma matrices, for some k. Two important operations on Clifford numbers are the following:
● The scalar part aS of an element a∈A is by definition the coefficient of the unit element 1 in the expansion of a in any basis (such as B.) Using the isomorphism above, we can view a as a 4×4 matrix, and we then have
aS=14tra, | (2.8) |
where tr denotes the usual operation of taking the trace of a matrix.
● The conjugate reversion (a.k.a. Dirac adjoint) ¯a of a∈A is by definition the element obtained by reversing the order of multiplication of the 1-vectors in the expansion of a in terms of k-vectors, and taking the complex conjugate of the coefficients in that expansion. Thus, ¯a=a∗S1+∑Ia∗I˜γI with ˜γI=γik…γi1 whenever γI=γi1…γik. Using the isomorphism A≅M4(C), it is not hard to see that, when a is viewed as a 4×4 matrix,
¯a=γ0a†γ0, | (2.9) |
where a†=(a∗)T denotes the conjugate-transpose of a. (Here and elsewhere, ∗ denotes the complex conjugation i→−i, while T denotes the matrix transpose.)
The isomorphism A≅P⊗P, on the other hand, can be realized by first noting that the complexification of the real Clifford algebra Cl1,3(R) yields the complex Clifford algebra Cl(4), which is known to be isomorphic to Cl(2)⊗Cl(2), and that Cl(2) is in turn isomorphic to P.
Finally, since P≅M2(C), by partitioning a 4×4 matrix into four 2×2 blocks in the obvious way, we may view an element Φ∈A≅M4(C) as a 2×2 matrix with entries in P:
Φ=(ϕ+χ−χ+ϕ−),ϕ±,χ±∈P≅M2(C). | (2.10) |
In [16] it was shown that there is a correspondence between rank-one spinors and 1-forms, and similarly between trace-free rank-two spinors and 2-forms. Specifically, for every rank-two bispinor ψph=(ϕ+χ−χ+ϕ−) that satisfies the trace condition trϕ+=trϕ−=0, there exists two real-valued 2-form f± and two complex-valued 1-forms a± such that
ϕ+=Σ(f+),ϕ−=Σ′(f−),χ+=σ′(a+),χ−=σ(a−), |
where the (invertible) mappings σ,σ′,Σ,Σ′ are defined as follows: Let σ0=σ′0=12×2. Let σ1,σ2,σ3 denote the Pauli matrices, and let σ′k:=−σk for k=1,2,3. Then, for contravariant tensors A and F,
σ(A):=σμAμ,σ′(A):=σ′μAμ,Σ(F):=i4σμσ′νFμν,Σ′(F):=i4σ′μσνFμν. |
The action on other tensor types can then be defined via the musical isomorphism ♯, i.e. raising of subscript indices using the (inverse) spacetime metric.
The group of spacetime rotations of R1,3 is the Lorentz group O(1,3). This group is disconnected: it has four connected components. The connected component that contains the identity element is called the proper Lorentz group. Viewed as a matrix group, the proper Lorentz group is identified with SO0(1,3), the subgroup of matrices Λ∈O(1,3) with detΛ=1, and uTΛu>0 for all future-directed timelike vectors u∈R1,3.
The full Lorentz group O(1,3) is generated by elements of SO0(1,3) together with the space-reflection
P:=(100−13) | (2.11) |
and time-reversal T:=−P=(−10013). For x∈R1,3, let
γ(x):=γμxμ=(0x01+xiσix01−xiσi0)=(0σ(x)σ′(x)0)∈A | (2.12) |
be the image of x under the standard embedding of the Minkowski spacetime into its Clifford algebra (Note that indices are raised and lowered using the Minkowski metric η, and that we are using Einstein's summation convention). It is a standard result of the representation theory of the Lorentz group that, for every Λ∈SO0(1,3), there exists A=A(Λ)∈SL(2,C)⊂P such that
γ(Λx)=LΛγ(x)L−1Λ |
where
LΛ=(A00A−†)∈A. | (2.13) |
It is also easy to check that
γ(Px)=γ0γ(x)γ0 |
so that, as an operator on C4, we can set
LP=γ0. | (2.14) |
There are of course other choices for LP, for example LP=iγ0. In fact we know that, since the Lorentz group is not connected, it does not have a unique covering group: it has eight non-isomorphic coverings. In four of these, the representation of the time-reversal operator T is unitary, and in the other four it is anti-unitary (see [18], Thm. 3. 10.) Here we are going to make a convenient choice for LT that makes it anti-unitary, while keeping the Dirac equation (see below) covariant under the action of the full Lorentz group:
LT:=−iS2C, | (2.15) |
where
Sk:=(σk00σk), | (2.16) |
are components of the spin operator, and C denotes the complex-conjugation operator in A≅M2(P), i.e.,
Cψ:=(ϕ∗+χ∗−χ∗+ϕ∗−). | (2.17) |
We now have the projective representation of O(1,3) as the group generated by matrices of the form LΛ as in (2.13) together with LP and LT.
The Dirac opertor D̸ on R1,d is by definition
D̸:=γ(∂)=γμ∂∂xμ. | (2.18) |
Thus, for d=3, we have
D̸=(012∂0−σ⋅∇12∂0+σ⋅∇0)=:(0D−D+0). | (2.19) |
Thus,
D−:=σ(∂),D+:=σ′(∂) |
(since ∂k=−∂k for k=1,2,3.) Note that D̸2=14□ and D+D−=D−D+=12□ where □:=∂20−ΔR3 is the three-dimensional wave operator.
According to Kiessling & Tahvildar-Zadeh [16], in d space dimensions, the wave function of a single photon is a rank-two bi-spinor field on R1,d which, when viewed as a linear transformation, has trace-free diagonal blocks. In the case d=3, rank-two bi-spinors are the same as Clifford numbers, i.e. general elements of the algebra A, and thus of the form (2.10), while the trace-free condition implies
trϕ+=trϕ−=0. | (2.20) |
Here, tr means the matrix trace, or equivalently, twice the scalar part of the quaternion.
Remark 2.1. The trace condition (2.20) projects out the spin-zero sector. An arbitrary element of A is a mixture of spin-zero and spin-one fields.
Thus, in three space dimensions, the wave function of a single photon has four quaternionic components, or 16 complex components, two of which have to be zero because of (2.20):
ψph=(ϕ+χ−χ+ϕ−). | (2.21) |
Moreover, according to [16], the photon wave function satisfies a Dirac-type equation with a projection term:
−iℏD̸ψph+m↯Πψph=0, | (2.22) |
where Π is the projection onto diagonal blocks (defined in terms of the projections Π± previously defined in (2.7)):
ΠΨ:=Π+ΨΠ++Π−ΨΠ−,∀Ψ∈A, | (2.23) |
ℏ is Planck's constant, and m↯>0 a dimensional constant to be determined. (The speed of light has been set equal to one.)
The equation (2.22) is the Euler–Lagrange equation for an action functional with the real scalar Lagrangian density function given by
ℓ ph=ℏ16πitr(¯ψphγμ∂μψph−∂μ¯ψphγμψph)+m↯8πtr(¯ψphΠψph). | (2.24) |
Remark 2.2. The massive version of (2.22), i.e. where the projection operator is replaced by the identity, was already written down by M. Riesz a long time ago [19]:
−iℏD̸ψ+mψ=0,ψ∈M4(C). | (2.25) |
This is the Euler–Lagrange equation for an action functional with real scalar Lagrangian density given by
ℓ=ℏ16πitr(¯ψγμ∂μψ−∂μ¯ψγμψ)+m8πtr(¯ψψ). | (2.26) |
See the Appendix at the end of this paper for a brief account of the genesis of the photon wave function, and its precursors in the works of Riesz and Harish-Chandra. For a fuller account, and a review of all the previous attempts at constructing a photon wave function and wave equation, see [16].
In terms of the wave function's quaternionic constituents, (2.22) can be written as
{D±χ∓=−im↯ℏϕ∓D±ϕ±=0. | (2.27) |
Note that the above implies that all components of ψph satisfy the massless linear Klein-Gordon (a.k.a. classical wave) equation
□ϕ±=0,□χ±=0. | (2.28) |
Let
ρ:O(1,3)→SL(2,C)×SL(2,C),ρ(Λ)=LΛ, |
denote the projective representation of the Lorentz group given in the above. It is easy to check that (2.22) is covariant with respect to the action of the full Lorentz group, i.e., given ψph(x) any solution of (2.22) and any Λ∈O(1,3), the spinor
ψph′(x):=LΛψel(Λ−1x)L−1Λ | (2.29) |
is also a solution of (2.22). This implies that the components ϕ±,χ± of ψph transform in the following way: For Λ∈SO0(1,3), we have
ϕ+→Aϕ+A−1,ϕ−→A−†ϕ−A†,χ+→A−†χ+A−1,χ−→Aχ−A†, | (2.30) |
where A∈SL(2,C) is such that LΛ=(A00A−†) is the projective representation of Λ, while under Λ=P, the space reflection (parity) transformation, using (2.29, 2.14) we have
ϕ±→ϕ∓,χ±→χ∓. | (2.31) |
Finally, under Λ=T, the time-reversal transformation, using (2.29, 2.15) we have
ϕ±→ϕ∗±,χ±→χ∗±. | (2.32) |
Note that this in particular implies that ψph∗:=(ϕ∗+χ∗−χ∗+ϕ∗−) solves the time-reversed version of (2.22).
Let ∘ψph=(∘ϕ+∘χ−∘χ+∘ϕ−) be the initial data supplied on the Cauchy hypersurface {x0=0} for (2.22) that is subject to compatibility conditions tr∘ϕ±=tr((σ⋅∇)∘ϕ±)=0. The corresponding initial value problem for ψph is equivalent to the following Cauchy problems for the classical wave equation:
{□ϕ±=0ϕ±|x0=0=∘ϕ±∂0ϕ±|x0=0=∓σ⋅∇∘ϕ±{□χ±=0χ±|x0=0=∘χ±∂0χ±|x0=0=∓im↯ℏ∘ϕ±±σ⋅∇∘χ± | (2.33) |
We now establish that the diagonal blocks Πψph of the photon wave function propagate only in transversal modes: As explained in [16], by fixing a Lorentz frame for the Minkowski space, one can find e±,b±:R1,3→R3 such that
ϕ+=iσ⋅(e++ib+),ϕ−=−iσ⋅(e−−ib−). | (2.34) |
Thus, setting f±:=e±+ib±, in this frame the equations for the diagonal blocks become
ℏ(∂t+σ⋅∇)(σ⋅f+)=0,ℏ(∂t−σ⋅∇)(σ⋅f∗−)=0. | (2.35) |
Now, it is well-known [20,21,22] that the equation
(∂t+σ⋅∇)(σ⋅f)=0, | (2.36) |
for f:=e+ib:R1,3→C3, is formally equivalent to the Maxwell system of equations for a source-free electric field e and a magnetic induction field b in the given Lorentz frame, viz.
∂te−∇×b=0,∇⋅e=0,∂tb+∇×e=0,∇⋅b=0. | (2.37) |
Thus, the first equation in (2.35) and the second equation in (2.35) are each separately equivalent to (2.37). This proves the absence of longitudinal modes in (2.35). It also establishes that to describe a photon wave function fully, even in a fixed frame, a pair of Maxwell fields (solutions of (2.37)) are needed, one will not be enough (cf. [23]).
More generally, it was established in [16] that
χ+=σ′(a+),χ−=σ(a−), | (2.38) |
where a± are complex-valued 1-forms on the configuration Minkowski space. Even though this appears to imply that the off-diagonal terms have 8 complex, or 16 real, degrees of freedom, it turns out that at least half of those are due to gauge freedom, which for (2.22) consists of
ψph↦ψph+(1−Π)Υ,D̸Υ=0. | (2.39) |
More precisely, the following was shown in [16]:
Proposition. Let ψph=(ψ+χ−χ+ψ−) be a solution of (2.22). There exists a gauge transformation (2.39) such that, after applying it, the χ± are Hermitian matrices (equivalently, the a± are real-valued.)
Fixing a Lorentz frame, we can therefore set
χ+=φ+σ0−σ⋅a+,χ−=φ−σ0+σ⋅a−, | (2.40) |
for φ±:R1,3→R and a±:R1,3→R3. Let us also denote the components of ϕ± as in (2.34). Let ψph be a solution of (2.22). Writing (2.22) out in components, we then obtain
∂tφ±+∇⋅a±=0,e±=ℏm↯(−∇φ±−∂ta±),b±=ℏm↯∇×a±. | (2.41) |
Thus, it appears that the relationship of the off-diagonal terms in the photon wave function ψph to its diagonal terms is formally the same as that of electromagnetic potentials (in Lorenz gauge) to their corresponding electromagnetic fields‡.
‡We stress that this is only a formal, mathematical correspondence. The wave function of any quantum mechanical system of particles must be defined on the configuration space of those particles, a space whose dimension goes up with the number of particles, while classical fields such as the electric and magnetic fields are defined on physical space, which always has the same dimension regardless of how many particles are in it.
Equations (2.35) also imply that φ± and a± must satisfy the classical wave equation
∂2tφ±−Δφ±=0,∂2ta±−Δa±=0. | (2.42) |
The existence of a conserved probability current is of profound importance to the understanding of the dynamics of a quantum particle. In [16] we showed that the photon wave function ψph described above has an intrinsically-defined conserved probability current, one which we constructed in two steps: First, we showed that given any Killing field X of Minkowski space, the manifestly covariant current
jμX:=14tr(¯ϕphγμϕphγ(X)), | (2.43) |
where ϕph:=Πψph, is conserved, i.e.,
∂μjμX=0. | (2.44) |
Here, ¯ψ:=γ0ψ†γ0 is the Dirac adjoint for rank-two bispinors, and γ(X):=γμXμ.
The proof of (2.44) relies on the fact that for ψph satisfying the photon wave equation (2.22), its projection onto the diagonal blocks ϕph will satisfy the massless Dirac equation, which is the same as the massless version of Riesz's equation (2.25), for which one has the conservation laws (3.14). We then set jμX:=τμνXν.
We next proved that when X is causal and future-directed, then so is jX, i.e. η(jX,jX)≥0, and j0X≥0. We then showed that there exists a distinguished, constant (and therefore Killing) vectorfield X that is completely determined by the wave function ψph (in fact, given a Cauchy surface Σ, it depends only on the initial value of ψph on Σ.) We defined X by considering any given Lorentz frame {e(μ)} of Minkowski space and defining
π(μ):=∫Rdj0e(μ)dx. | (2.45) |
Conservation law (2.44) then implies that the π(μ) are constant, while the definitions of π(μ) and jX imply that as a four-component object, π:=(π(μ)) transforms correctly, i.e. like a Lorentz 4-vector. Moreover, π is a future-directed causal vectorfield, and is typically timelike. We set
X:=π/|π|2,|π|2:=η(π,π) |
and defined the photon probability current jph to be jX for this particular choice of constant vectorfield X.
Finally, we showed that jph satisfies the appropriate generalization of the Born rule, i.e., if one defines
ρph:=j0ph,vkph:=jkphj0ph | (2.46) |
then one has the continuity equation
∂tρph+∂k(ρphvkph)=0, | (2.47) |
and moreover, in the Lorentz frame where§ X=(1/|π|,0,0,0)T, one has
§We note that there was a slip of pen in eq. (7.15) of [16].
j0ph=Cϕtr(ϕph†ϕph)=Cϕtr(ϕ†+ϕ++ϕ†−ϕ−), | (2.48) |
jkph=Cϕtr(ϕph†αkϕph)=Cϕtr(ϕ†+σkϕ+−ϕ†−σkϕ−). | (2.49) |
where
Cϕ:=1∫Rdtr(ϕph†ϕph)dx | (2.50) |
is a normalization constant (i.e. it's time-independent) that depends only on the initial values of the photon wave function. Therefore, ρph:=j0ph is for all practical purposes a probability density, which (for normalized wave functions with Cϕ=1) depends quadratically on the wave function¶.
¶Note that this is the same situation as in the standard Born rule for non-relativistic quantum mechanics, namely, that the Schrödinger wave function needs to be normalized so that its ρ:=ψ†ψ integrates to one, and the probability current is quadratic in the normalized wave function.
Let (M,g) be a 4-dimensional smooth, connected, orientable, and time-orientable Lorentzian manifold. We are going to additionally assume that (M,g) has a Cauchy hypersurface Σ, i.e. a complete spacelike submanifold of codimension 1 with the property that every inextendible past-directed timelike curve in M intersects it at one and only one point. In other words, we assume that (M,g) is globally hyperbolic.
At every point p∈M, the tangent space TpM is a copy of the Minkowski space, with gp a quadratic form of signature (+,−,−,−) on it. Therefore, virtually all of the above constructions on Minkowski space extend to the tangent spaces at every point of M. In particular, the Clifford algebra A has a basis generated by g-dependent Dirac matrices γμ satisfying
γμγν+γνγμ=2gμν14. | (3.1) |
Since A≅M2(P), a general element G∈A can be written as
G=(g+h−h+g−),h±,g±∈P | (3.2) |
The photon wave function ψph is defined to be an A-valued field with trace-free diagonal blocks, as in (2.21). It is assumed to satisfy the equation
−iℏD̸ψph+m↯Πψph=0, | (3.3) |
where D̸ is the Dirac operator of (M,g), i.e.
D̸:=γμDμ. | (3.4) |
Here, Dμ is the covariant spin connection on M, which is a lifting of the Levi-Civita connection of the metric g to the spin bundle over M. It follows that both the metric g and the γμ are covariantly constant with respect to D differentiation. The projection Π is defined as before (2.23).
Even though the domain (M,g) is not assumed to have any symmetries, one may still wonder if there are conservation laws for (3.3) associated with symmetries of the fiber A. Because A is an algebra, it acts on itself, and since dimCA=16, there could in theory be up to a 16-dimensional set of symmetry generators, with a conserved Noether current associated to each one. It however turns out that the dynamics generated by (3.3) preserve at most 5 of these symmetries. In particular, the projection term in (3.3) is not invariant under the full symmetry group. For a massive spin-one field satisfying (2.25), where the projection Π in (2.22) is replaced with the identity operator, there will be a larger symmetry group, and the resulting conservation laws in fact include those found by Riesz [19]. We will follow the general procedure outlined in [10] for finding Noetherian conservation laws that are due to the symmetries of the fiber, for Lagrangian field theories defined on sections of a vector bundle.
Let G be a section of the bundle B of rank-2 bi-spinors over the configuration Minkowski space, and let Z denote the generator of the following right action of G on rank-two bi-spinors ψ∈B:
ψs:=ψeisG,s∈(−ϵ,ϵ). | (3.5) |
Z is thus a section of the bundle E:=∪x∈ML(Bx,Bx) where Bx:=π−1B,M(x) is the fiber in B over x∈M, and
Z⋅ψ=dds|s=0ψs=iψG. | (3.6) |
Let ϵ=ϵ[g] denote the volume form of M with respect to the metric g, let the Lagrangian density function ℓ ph be defined as in (2.24), except with ∂μ replaced by the spin connection Dμ, and let Lph:=ℓ phϵ be the 4-form corresponding to it. According to [10], the conserved Noether current corresponding to Z is the 3-form
Jμνλ=pμνλa(Z⋅ψ)a, |
where p=(pμνλa) are the canonical momenta (i.e. the derivative of the Lagrangian density Lph with respect to the canonical velocities v=(Dμψa).) Furthermore, we have (see [10, Chap. 4, (1.148)])
dJ=LZLph+((DZ)⋅ψ)∧p. |
It follows that J is a conserved Noether current if the right-hand side of the above vanishes on solutions ψ of the Euler-Lagrange equations for the action with Lagrangian Lph.
Let the vectorfield jμ=ϵμνλκJνλκ denote the Hodge dual of J with respect to the volume form ϵ[g]. If dJ=0, it follows that j is divergence-free. Thus, defining
jμG:=(¯ψphγμψphG)S, | (3.7) |
(using the notation (2.8), ) we have
∇μjμG=0, | (3.8) |
when ψph satisfies (3.3).
From (3.8) it follows that, if one introduces coordinates (t,s) in a neighborhood of the Cauchy hypersurface Σ⊂M such that t is a time function on that neighborhood with level sets Σt and the original Cauchy hypersurface Σ=Σ0, then by the divergence theorem one has
ddt∫Σtj0Gd3s=0. |
For the photon Lagrangian (2.24), we have
jμ=(∗p)μa(Z⋅ψph)a=∂ℓ ph∂(Dμψph)(iψphG)=tr[116πi¯ψphγμ(iψphG)]=116πtr(¯ψphγμψphG). | (3.9) |
We look for conditions on G such that the current j is conserved, i.e.,
LZLph[ψph]+((DZ)⋅ψph)∧p=0, | (3.10) |
for ψph satisfying (3.3). This does not seem to be an easy task, and may require G to be ψph-dependent. If one is however content with deriving sufficient conditions on G, one can simplify matters considerably by assuming that G is covariantly constant, i.e.
DμG=0. | (3.11) |
(We will investigate nontrivial solutions of this in a moment.) This implies that DZ=0. We next compute
LZℓ ph(ψph)=dds|s=0ℓ ph(ψpheisG)=I+II, |
where I contains the contribution of the first term in the Lagrangian, the one involving Dψph, and II has the contribution of the projection term. We find
I=ℏ16πtr(−¯GN+NG),N:=¯ψphγμDμψph−Dμ¯ψphγμψph. |
Thus, we have I=0 for all ψph provided
¯G=G | (3.12) |
(since tr(NG)=tr(GN)). Assuming (3.12), we then have
II=m↯i8πtr(¯ψph(Π(ψphG)−(Πψph)G))=m↯i8πtr[(ϕ†−χ−−χ†−ϕ−)h++(ϕ†+χ+−χ†+ϕ+)h−], |
where we have used (2.21) and (3.2). Therefore II=0 for all ψph provided we further restrict G such that
h±∈C. | (3.13) |
Incidentally, we can also see that the Lagrangian (2.26) of the massive version of the theory has a larger symmetry group since the term corresponding to II in that case will be zero without the need for assumption (3.13). In particular, on Minkowski space and for the massive case studied by Riesz in [19], one can set G=γν for ν=0,…,3, in which case one recovers the conservation laws of the Riesz tensor‖:
‖Riesz does not mention Noether or refer to her work in his paper. He derives his conservation laws using the standard multiplier method, i.e., multiplying the equation through by something and integrating by parts. This led us in [16] to erroneously state that Riesz's conservation laws are non-Noetherian. We are glad to have the opportunity here to correct this slip of mind.
∇μτμν=0,τμν:=14tr(¯ψγμψγν), | (3.14) |
whenever ψ satisfies (2.25). To connect back with what was shown in [16], we observe that if ψph satisfies (2.22), then Πψph will satisfy the massless Dirac equation, which is (2.25) with m=0. Thus, (3.14) holds with ψ=Πψph, and this is the conservation law used in [16] to define a probability current for the photon wave equation on Minkowski spacetime.
The assumption (3.11) allowed us to find several conserved quantities, but the question is, on a curved spacetime, does it have any nontrivial solutions? Recall that G, being a section of the bundle B, at every point x∈M has an expansion in the basis (2.5) for the Clifford algebra A. It follows that there are p-forms {G(p)} for p=0,1,2,3,4 such that G=G(0)1+G(1)μγμ+⋯+G(4)γ5. Thus, G(0) is a function, and G(4)=Gϵ[g] for some function G on M. The condition (3.11) implies that G(0) and G are two complex constants. On the other hand, on a 4-dimensional manifold with no continuous symmetries, there are in general no nontrivial covariantly constant 1-forms, 2-forms, or 3-forms, so G(1)=G(2)=G(3)=0, which, together with the other restrictions (3.12) and (3.13), on a general spacetime leaves us only with
G=a1+ibγ5,a,b∈R | (3.15) |
(cf. [16, (5.10)].)
If, on the other hand, we are on a Minkowski background, where every vectorfield whose components are constant (in the standard frame given by Cartesian coordinates on R1,3) is both a Killing field and is hypersurface orthogonal, (3.11) just means that the components of G (in the standard Cartesian frame) are constant, and we have the full 5C- or 10R-dimensional set of conserved quantities implied by conditions (3.12) and (3.13) on the generator G. Substituting in (3.7) for ψph in terms of its quaternionic components ϕ± and χ± as in (2.21), one then obtains the explicit form of the conserved charge, i.e. the time component j0G of the current to be
j0G=14tr[(χ†−ϕ++ϕ†−χ+)g++(χ†+ϕ−+ϕ†+χ−)g−+(χ†−χ−+ϕ†−ϕ−)h++(χ†+χ++ϕ†+ϕ+)h−], |
where, by virtue of the assumption (3.12), we have g+=g†−∈P and h±∈R. Thus, we have 4 complex-valued and 2 real-valued conserved currents. One of the complex conserved charges, the one corresponding to (3.15), had already been identified in [16]. On Minkowski space, using the decompositions (2.40) and (2.34), it can be written as
z:=14tr(χ†−ϕ++ϕ†−χ+)=12(a+⋅b−−a−⋅b+)+i2(a−⋅e+−a+⋅e−), | (3.16) |
while the real-valued charges correspond to h±∈R, g±=0, and are:
r±:=14tr(χ†±χ±+ϕ†±ϕ±)=12(φ2±+|a±|2)+12(|e±|2+|b±|2). | (3.17) |
In what follows, we are going to focus on three more complex charges, which correspond to g+=g†−=c⋅σ for an arbitrary c∈C3. Here, σ denotes the vector of Pauli matrices (Recall that the Pauli matrices span the subspace of imaginary quaternions, i.e. trace-free Hermitian matrices). Thus, we have a conserved, C3-valued charge
J:=14tr[(χ†−ϕ++ϕ†−χ+)σ], | (3.18) |
that is to say, each component of J is a conserved complex charge. Using the decompositions (2.40) and (2.34), one has
J=−12(φ+b−+φ−b++a+×e−+a−×e+)−i2(−φ+e−−φ−e++a+×b−+a−×b+). |
On a flat background, the quantities (r±,z,J) together form a (real) ten-dimensional set of conserved charges for the photon wave function in the sense that their integrals over constant t slices Σt are independent of t.
For the remainder of this paper, we are going to restrict ourselves to the case of a Minkowski background, and investigate the gauge-dependence of the new conserved charges we have just found.
Finding all these new conservation laws may seem significant, but there is a potential snag with the above procedure, namely, we know that the photon wave equation (3.3) is invariant under a rather large group of gauge transformations: It was shown in [16] that if one adds to a solution ψph of (3.3) the off-diagonal part of any bispinor field Υ that solves the massless Dirac equation
D̸Υ=0, |
then the resulting wave function ψph+(1−Π)Υ is still a solution, and is physically equivalent to ψph. One important consequence of this fact is that the off-diagonal components of ψph, namely the χ± can change a lot under a gauge transformation. For example, already on Minkowski space, χ± transforms to
χ′±=χ±+D±v±,□v±=0 | (3.19) |
with v± two arbitrary solutions of the linear classical wave equation, while the diagonal blocks ϕ± remain invariant. Thus, any expression in which χ± explicitly appear is not likely to be gauge-invariant. A conserved quantity that is not invariant under a gauge transformation would be of little value for physical studies. None of the quantities we have produced so far are manifestly gauge-invariant, since they all have the χ± appearing in them**.
**Except in one space dimension (d=1), where ϕ±≡0 and there is no gauge freedom left in χ±. In that case, the r± that were defined in (3.17) are conserved quantities that were not previously derived.
The situation may however be analogous to the Einstein equations of general relativity, for which it is possible to find a set of conserved currents with densities that are not manifestly invariant under a diffeomorphism, since they depend on various components of the metric written in a specific coordinate system. And yet, in that case one is able to find global conserved quantites, the so-called ADM quantities, obtained by integrating those densities on the "sphere at spatial infinity". This is in the context of asymptotically flat spacetimes, i.e., those for which outside a compact set there is a Cartesian coordinate system (xμ)=(x0,x) with the property that the metric in those coordinates approaches the Euclidean metric (often at a specified rate) as |x|→∞ for fixed x0. It is then shown that these integral quantities are both independent of x0 and diffeomorphism invariant. For Einstein's equations, the existence of these quantities is connected with the asymptotic symmetries of such spacetimes, i.e. symmetries at infinity, and there are indeed 10 such quantities, corresponding to the generators of the Poincare group of isometries of Minkowski space, aptly named mass (or equivalently energy), linear momentum, angular momentum, and center of mass integrals (see e.g [15].) One also notes that these quantities are obtained by integrating certain surface densities (i.e. 2-forms) on large coordinate spheres that lie in a constant time slice, and then taking the limit as the radius of the sphere goes to infinity.
For this analogy to work for the photon wave equation on Minkowski background, we need to construct the 2-forms that are to be integrated on coordinate spheres Sr lying in a constant time slice, and we need to verify that the limit as r→∞ of those integrals is both independent of t and also invariant under our gauge transformations. Since we have already found several conservation laws for the photon wave equation, we can investigate whether a conserved charge already constructed, say Ji, is itself a complete 3-divergence, so that its integral on a large ball of radius r in Σt can be turned into a surface integral on the large sphere that is the boundary of that ball, so that in the limit as r→∞ we end up with the total charge of that slice, which we already know is conserved. We would then need to verify that even though the integrand is not manifestly gauge invariant, the limiting value of the integral is.
In [16] we carried out this program for one of the two conserved charges we had constructed, obtaining an ADM-like quantity that we tentatively termed "cross helicity". Here in this paper we would like to do the same with the vector-valued charge J∈C3. It turns out however that its components Ji are in fact not complete divergences. To remedy this, we will construct a separate set of conserved quantities Ki, with the property that the components of L=J−K are complete divergences, Li=∇⋅Yi, thus arriving at complex-valued ADM-like conserved quantities
fi:=limr→∞∫Br⊂ΣtLi d3s=limr→∞∫Sr=∂BrYi⋅n dS. |
We will then conclude by showing that adding a solution of the massless Dirac equation to ψph does not change fi since its contribution to the integrand falls off faster than 1/r2.
Finally, viewing this complex vector as f=e+ib allows us to identify f with a constant bispinor Ψ∞ that is self-dual, i.e., one whose diagonal blocks satisfy ϕ−=ϕ†+.
Boundary current is the name used by Christodoulou [15, p. 50] to denote the general procedure of obtaining ADM-like conserved quantities in theories with gauge freedom, such as Einstein's equations.
In our setting, we need to correct the conserved charges J so that they become boundary currents. We begin by recalling that on Minkowski space, the quaternionic constituents of ψph satisfy the massless wave equation (2.28), since we have
D̸2=14□. |
Now recall that D̸=(0D−D+0) and thus D̸2=(D−D+00D+D−). Also, from (2.27) it follows that D±D∓χ±=0. Thus, we have
□χ±=∇μ∇μχ±=0. | (3.20) |
Consider now the 1-form κ=Kμdxμ with
Kμ:=iℏm↯(χ†−∇μχ+−∇μχ†−χ+)S. | (3.21) |
From (3.20) it follows that
∇μKμ=0 | (3.22) |
i.e. κ is another conserved current for ψph. (In the Riemannian setting, (3.22) is sometimes referred to as the second Green identity). In particular, its time component
K0=iℏm↯(χ†−∇tχ+−∇tχ†−χ+)S | (3.23) |
is a conserved charge. Clearly, the same is true for
K:=iℏm↯[(χ†−∇tχ+−∇tχ†−χ+)σ]S. | (3.24) |
We now claim the following.
Proposition 3.1. Let z, K0, J, and K be as in (3.16), (3.23), (3.18), and (3.24), respectively. Let
w:=z−K0,L:=J−K. | (3.25) |
Then, the complex scalar w and each component of L are boundary currents, i.e., there exist vectorfields X and Yi, i=1,2,3 such that
w=∇⋅X,Li=∇⋅Yi,i=1,2,3. | (3.26) |
Proof. Let D±:=∇t1±σ⋅∇s as before. The quaternionic components of ψph satisfy equations (2.27). Thus, we can write
χ†−ϕ+=iℏm↯χ†−D−χ+=iℏm↯(χ†−∇tχ+−χ†−(σ⋅∇)χ+)=iℏm↯(χ†−∇tχ+−∇⋅(χ†−σχ+)+((σ⋅∇)χ−)†χ+)=iℏm↯(χ†−∇tχ+−∇⋅(χ†−σχ+)−(∇tχ−)†χ++(D+χ−)†χ+)=iℏm↯(χ†−∇tχ+−(∇tχ−)†χ+−∇⋅(χ†−σχ+))−ϕ†−χ+. | (3.27) |
Thus, it follows that
(χ†−ϕ++ϕ†−χ+)S−K0=−iℏm↯(∇⋅(χ†−σχ+))S=−iℏm↯∇⋅(χ†−σχ+)S, |
which is a complete divergence. Thus, defining
X:=−iℏm↯(χ†−σχ+)S | (3.28) |
and recalling (3.16), we obtain
w:=z−K0=∇⋅X. | (3.29) |
Let c∈C3 be a fixed vector. Repeating the calculation in (3.27) with σ⋅c multiplied on the right gives us the desired vector version, i.e.,
J−K=−iℏm↯∇⋅[(χ†−σχ+)σ⋅c]S. |
Thus, setting
Yk:=−iℏm↯[(χ†−σχ+)σk]S,k=1,2,3, | (3.30) |
establishes the claim.
Let Sr denote the coordinate sphere of radius r lying in the constant-time slice Σt of the globally hyperbolic, asymptotically flat spacetime (M,g). Let ψph be a photon wave function, i.e. a rank two bispinor satisfying (3.3). Let ϕ±,χ± denote the quaternionic components of ψph, as in (2.21), and let the vectorfields X and Yj be defined as in (3.28) and (3.30), respectively. Let
μ(ψph):=limr→∞∫SrX⋅ndS,fi(ψph):=limr→∞∫SrYi⋅ndS | (3.31) |
define the complex scalar μ and components of the complex vector f=e+ib. It follows that μ and fi are independent of t, and therefore constant. What remains to show is that, even though X and Yi are not manifestly gauge-invariant, μ and f are.
Remark 3.2. Using the decomposition (2.40) with respect to the frame {∂∂xμ} associated with the Cartesian coordinates on Minkowski space, we obtain
μ=ℏ2m↯limr→∞∫Sr(a−×a++i(φ−a+−φ+a−))⋅n dSr | (3.32) |
b=ℏ2m↯limr→∞∫Sr(a−⋅n)a++(a+⋅n)a−−(a−⋅a++φ−φ+)n dSr | (3.33) |
e=ℏ2m↯limr→∞∫Sr(φ+a−+φ−a+)×n dSr | (3.34) |
Here we show how to construct photon wave functions ψph with a finite, nonzero bispinor "at infinity" in Minkowski space. Using gauge-invariance, we are going to look for solutions to (2.41–2.42) that are not just in Lorenz gauge but also in Coulomb gauge. This implies that, without loss of generality, φ±≡0, yielding the following equations:
□a±=0,∇⋅a±=0. | (3.35) |
It is also clear that for the quantities defined in (3.32) to be finite, one needs
|a±|∼c±|s|as |s|→∞. | (3.36) |
Equations (3.35) can be solved exactly using separation of variables, and a complete basis for "outgoing" solutions with finite L2 norm for e±,b± was found by Marchal [24]. These have the form a=∑∞n=1an with
an(t,s)=Pn(s)|s|ngn(t,|s|)+Pn−2(s)|s|n−2hn(t,|s|), | (3.37) |
where P−1=P0≡0, and, for n≥1, Pn,Pn−2∈R3 are two vector-valued homogeneous harmonic polynomials of degree n and n−2, respectively, with the property that
s⋅(Pn(s)+|s|2Pn−2(s))=0, | (3.38) |
gn(t,r):=(−r)n(1r∂∂r)nfn(t−r)r,hn(t,r):=(−r)n(1r∂∂r)nf″n(t−r)r, | (3.39) |
with fn∈Cn+2(R) arbitrary. To ensure that the L2 condition is satisfied, Marchal took fn to be compactly supported in an interval not containing zero, but this is clearly not necessary. We can see that for the asymptotic condition (3.36) as well as the L2 condition on e±,b± to be satisfied, it is enough that fn∈Cn+2(R+) with
fn(r)∼{rn+1r→0+rnr→∞. | (3.40) |
Marchal showed that these solutions, when appropriately normalized, form a complete orthonormal basis for the outgoing solutions of (3.35) (For incoming solutions, change t−r to t+r in (3.39)).
Armed with this decomposition, we construct an example of a photon wave function with a nontrivial bispinor at infinity by letting a+=a1 and a−=a2, as defined in (3.37), with the choices
P1(s)=(0,−s3,s2),P2(s)=(−s2s3,s1s3,0) |
and with f1,f2 satisfying the conditions (3.40). It then follows that e=0 by the Coulomb gauge condition, and, since our choice makes a+⋅n=a−⋅n=0,
b=−ℏ2m↯limr→∞∫Sr(a+⋅a−)n dSr=ℏ2m↯limr→∞g1(0,r)g2(0,r)r2∫S1(s1s23)(s1s2s3)dS1=Cℏm↯(100) |
for some nonzero numerical constant C, since by their definitions (3.39), g1,g2∼1r as r→∞.
THEOREM 3.3. Let Υ be a solution of the massless Dirac equation D̸Υ=0 with smooth compactly supported data, and let
ψph′:=ψph+(1−Π)Υ |
be the gauge-transformed version of ψph. Let ψ± denote the off-diagonal blocks of Υ. It follows that
ϕ′±=ϕ±,χ′±=χ±+ψ±,D∓ψ±=0. |
We then have
μ(ψph′)=μ(ψph),fi(ψph′)=fi(ψph). |
Proof. Recall that there exist two real-valued 2-forms f± and two real-valued 1-forms a± such that
ϕ+=Σ(f+),ϕ−=Σ′(f−),χ+=σ′(a+),χ−=σ(a−). |
Moreover, if the above ψph satisfies our photon wave equation (2.22), then we will have
da±=m↯ℏf±,δa±=0,df±=0,δf±=0. |
Here, δ=∗d∗ is the co-differential (or divergence) operator, and ∗ is the Hodge star operator with respect to the spacetime metric g.
By the same token, if Υ=(0ψ−ψ+0) satisfies the massless Dirac equation D̸Υ=0, it follows that there are real-valued 1-forms α± such that ψ+=σ′(α+), ψ−=σ(α−), and we have
dα±=0,δα±=0. |
Thus, by the Poincaré lemma, there are functions h± such that α±=dh±, and moreover we have
□gh±=0, | (3.41) |
i.e., the h± solve the classical wave equation on the background spacetime (M,g), which we have assumed to be Minkowski space again. Thus, the Cauchy problem for (3.41) with smooth initial data prescribed on Σ0 is uniquely solvable, the solution is smooth, and satisfies the domain of dependence property. In particular, if the data is compactly supported, so would be the solution at all times. In that case, since ψ± are compactly supported, they will not make a contribution to the boundary currents μ and fi.
We note that, using standard density arguments, the compact support assumption on the data of h± can be relaxed to include data belonging to appropriate Sobolev spaces (e.g. ˙H1) without changing the conclusion. This will allow the dynamics of the wave function to be well-defined on L2-based Hilbert spaces.
Michael K.-H. Kiessling and A. Shadi Tahvildar-Zadeh: Writing-original draft, Writing-review editing, Supervision, Formal Analysis.
This paper is dedicated to Prof. Thomas C. Sideris on the occasion of his 70th birthday.
We thank Prof. Premala Chandra for wonderful conversations, and reminiscences about her father Harish-Chandra. We are grateful to the anonymous referee for a careful reading of our manuscript and for the truly helpful suggestions on how to improve it.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors declare there is no conflict of interest.
As we remarked earlier, the massive version of our photon wave equation (2.22), i.e. where the projection operator is replaced by the identity, was already written down by M. Riesz a long time ago [19]:
−iℏD̸ψ+mψ=0,ψ∈M4(C). | (A.1) |
Riesz however did not address the question of the physical significance of ψ, beyond the fact that it was a 4×4 matrix-valued field. His main point was to either obtain new conservation laws for the Dirac equation for the electron, or see the known ones in a new light. He did this by first finding the conservation laws that hold for (A.1), and then using the fact that a 4-component spin-half Dirac wave function Ψel can be viewed as the only nonzero column of a 4×4 matrix ψ, and that the set of all such matrices forms a minimal left ideal in the space of all Clifford numbers (his term for elements of Cl(4)), to restrict those conservation laws back to electron wavefunctions Ψel. In particular, Riesz seems to be unaware of the fact that a Clifford-number-valued field can represent a spin-one object, and that the equation (A.1) has the interpretation of a relativistic wave equation for a massive spin-one particle. Indeed, (A.1) is equivalent to Kemmer's equation [25] for the same type of particle:
βμ∂μψ+mψ=0, | (A.2) |
with ψ∈C10 and matrices βμ∈M10(C) satisfying the relations βμβνβρ+βρβνβμ=gρνβμ+gμνβρ, first found by Duffin [26].
At roughly the same time as Riesz, Harish-Chandra was working with the Kemmer equation, trying to find what the corresponding equation would be for a massless spin-one particle, and realizing that setting the mass parameter m in (A.2) to zero yields only trivial solutions. He showed in [27] that instead one must replace the mass term with a projection operator onto a 6-dimensional subspace of C10. He then went on and found many conservation laws for the "massless Kemmer equation" he had obtained, but not the ones that would correspond to what Riesz had found.
With hindsight, for someone who would be aware of these two strands of thought, it seems an easy move to combine them and come up both with the photon wave equation (2.22) as well as its key conserved current (2.43), long before we accomplished this task in 2017, but apparently no one did (although many came close, see [16] for details.) This may be partly because Riesz's paper appeared in a very obscure publication, and also he was generally ignored by physicists who were his contemporaries††. He did not seem to succeed in convincing them of the significance of Clifford algebras for modern physics. Riesz's students also do not seem to have picked up on this particular aspect of his work‡‡.
††See for example Gårding's account of Riesz's years in Lund [28].
‡‡During his long career in Stockholm and then Lund, M. Riesz had relatively few students, although this apparent lack of quantity was more than made up by quality, since those students became extremely well-known mathematicians: E. Hille, L. Gårding, and L. Hörmander, just to name three. Hille in particular is relevant to our story, since he came to the US and trained a whole generation of first-rate analysts at Yale, among them I. Segal, who did the same at Chicago and then at MIT. W. Strauss was Segal's first student at MIT, and continued this tradition at Brown, where his first student was R. Glassey. Glassey's first student at Indiana was Tom Sideris. The second author of this paper is another academic grandchild of Strauss.
Meanwhile, two years after Harish-Chandra (who was Dirac's assistant in Cambridge) published his work on Kemmer's equation, he accompanied Dirac to Princeton, where he switched to doing pure mathematics§§ (something for which apparently Dirac had a hard time forgiving him), eventually becoming one of the greatest algebraists of the 20th century, so he did not pursue this question further, either. In the ensuing decades, the empirical successes of Quantum Electrodynamics and Quantum Field Theory removed the urgency for physicists of addressing foundational questions such as what a photon is and how it interacts with the electron, so the search for a quantum-mechanical description of a photon as a particle was relegated to the theoretical backwaters. Various pronouncements by prominent physicists regarding either the futility or the impossibility of this task did not help matters either (see [30, p. 56], [31, p. 3]).
§§When asked for the reason for this switch, he is quoted to have said that now he could believe in what he was doing! [29]
[1] |
E. Noether, Invariant variation problems, Transport Theo. Stat. Phys., 1 (1971), 186–207 Translated from German (Nachr. Akad. Wiss. Göttingen Math. Phys. Kl. II, (1918), 235–257). https://doi.org/10.1080/00411457108231446 doi: 10.1080/00411457108231446
![]() |
[2] | Y. Kosmann-Schwarzbach, The Noether Theorems: Invariace and Conservation Laws in the Twentieth Century, translated by B. E. Schwarzbach, Springer, 2018. |
[3] | P. J. Olver, Applications of Lie Groups to Differential Equations, New York: Springer-Verlag, 1993. https://doi.org/10.1007/978-1-4612-4350-2 |
[4] |
S. Klainerman, Uniform decay estimates and the Lorentz invariance of the classical wave equation, Commun. Pur. Appl. Math., 38 (1985), 321–332. https://doi.org/10.1002/cpa.3160380305 doi: 10.1002/cpa.3160380305
![]() |
[5] |
T. C. Sideris, Decay estimates for the three-dimensional inhomogeneous Klein-Gordon equation and applications, Commun. Part. Differ. Equ., 14 (1989), 1421–1455. https://doi.org/10.1080/03605308908820660 doi: 10.1080/03605308908820660
![]() |
[6] |
S. Klainerman, T. C. Sideris, On almost global existence for nonrelativistic wave equations in 3D, Commun. Pur. Appl. Math., 49 (1996), 307–321. https://doi.org/10.1002/(SICI)1097-0312(199603)49:3<307::AID-CPA4>3.0.CO;2-H doi: 10.1002/(SICI)1097-0312(199603)49:3<307::AID-CPA4>3.0.CO;2-H
![]() |
[7] | T. C. Sideris, The null condition and global existence of nonlinear elastic waves, Invent. Math. 123 (1996), 323–342. https://doi.org/10.1007/BF01232380 |
[8] |
T. C. Sideris, Nonresonance and global existence of prestressed nonlinear elastic waves, Ann. of Math., 151 (2000), 849–874. https://doi.org/10.2307/121050 doi: 10.2307/121050
![]() |
[9] |
T. C. Sideris, S. Y. Tu, Global existence for systems of nonlinear wave equations in 3D with multiple speeds, SIAM J. Math. Anal., 33 (2001), 477–488. https://doi.org/10.1137/S0036141000378966 doi: 10.1137/S0036141000378966
![]() |
[10] | D. Christodoulou, The Action Principle and Partial Differential Equations, Princeton: Princeton University Press, 2000. https://doi.org/10.1515/9781400882687 |
[11] | A. Einstein, Hamiltonsches Prinzip und allgemeine Relativitätstheorie, Sitzungsber. preuss. Akad. Wiss., 2 (1916), 1111–1116. |
[12] | H. Weyl, Raum-Zeit-Materie, 4th ed., Berlin: Springer-Verlag, 1921. https://doi.org/10.1007/978-3-662-02044-9 |
[13] | L. Landau, E. Lifshitz, The Classical Theory of Fields, Cambridge: Addison-Wesley Press, 1951. |
[14] | R. Arnowitt, S. Deser, C. W. Misner, Energy and the criteria for radiation in general relativity, Phys. Rev., 118 (1960), 1100–1104. https://doi.org/10.1103/PhysRev.118.1100 |
[15] | D. Christodoulou, Mathematical Problems of General Relativity I, Zürich Lectures in Advanced Mathematics, Zürich: European Mathematical Society, 2008. https://doi.org/10.4171/005 |
[16] | M. K.-H. Kiessling, A. S. Tahvildar-Zadeh, On the quantum-mechanics of a single photon, J. Math. Phys., 59 (2018), 112302. https://doi.org/10.1063/1.5021066 |
[17] | D. Hestenes, Space-Time Algebra, 2nd ed., Birkhauser, 2015. https://doi.org/10.1007/978-3-319-18413-5 |
[18] | B. Thaller, The Dirac Equation, Berlin: Springer-Verlag, 1992. https://doi.org/10.1007/978-3-662-02753-0 |
[19] | M. Riesz, Sur certaines notions fondamentales en théorie quantique relativiste, In "Dixieme Congres Math. des Pays Scandinaves, " Copenhagen, 1946,123–148. |
[20] |
O. Laporte, G. E. Uhlenbeck, Application of spinor analysis to the Maxwell and Dirac equations, Phys. Rev., 37 (1931), 1380. https://doi.org/10.1103/PhysRev.37.1380 doi: 10.1103/PhysRev.37.1380
![]() |
[21] |
J. R. Oppenheimer, Note on light quanta and the electromagnetic field, Phys. Rev., 38 (1931), 725–748. https://doi.org/10.1103/PhysRev.38.725 doi: 10.1103/PhysRev.38.725
![]() |
[22] | I. Białynicki-Birula, Photon wave function, Prog. Opt., 36 (1996), 245–294. https://doi.org/10.1016/S0079-6638(08)70316-0 |
[23] |
R. Penrose, Nonlinear gravitons and curved twistor theory, Gen. Relat. Gravit., 7 (1976), 31–52. https://doi.org/10.1007/BF00762011 doi: 10.1007/BF00762011
![]() |
[24] | C. Marchal, Solution générale des équations de Maxwell dans le vide. Décomposition compléte des ondes électromagnétiques à énergie finie, Onde électrique, 71 (1991), 61–73. |
[25] |
N. Kemmer, The particle aspect of meson theory, Proc. R. Soc. London. A., 173 (1939), 91–116. https://doi.org/10.1098/rspa.1939.0131 doi: 10.1098/rspa.1939.0131
![]() |
[26] |
R. J. Duffin, On the characteristic matrices of covariant systems, Phys. Rev., 54 (1938), 1114. https://doi.org/10.1103/PhysRev.54.1114 doi: 10.1103/PhysRev.54.1114
![]() |
[27] |
Harish-Chandra, The correspondence between the particle and the wave aspects of the meson and the photon, Proc. R. Soc. London. A., 186 (1946), 502–525. https://doi.org/10.1098/rspa.1946.0061 doi: 10.1098/rspa.1946.0061
![]() |
[28] | L. Gårding, Mathematics and Mathematicians: Mathematics in Sweden before 1950, American Mathematical Society, 1994. |
[29] | A. Borel, Some recollections of Harish-Chandra, Current Science, 99 (2010), 69–71. |
[30] | M. Sargent, M. O. Scully, W. E. Lamb, Laser Physics, Addison-Wesley, 1974. |
[31] | S. Weinberg, The Quantum Theory of Fields, Vol. I, Cambridge University Press, 1995. https://doi.org/10.1017/CBO9781139644167 |