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

Classical perspectives on the Newton–Wigner position observable

Philip K. Schwartz Institute for Theoretical Physics, Leibniz University Hannover, Appelstraße 2, 30167 Hannover, Germany [email protected] Domenico Giulini Institute for Theoretical Physics, Leibniz University Hannover, Appelstraße 2, 30167 Hannover, Germany Center of Applied Space Technology and Microgravity, University of Bremen, Am Fallturm 1, 28359 Bremen, Germany [email protected]
Abstract

This paper deals with the Newton–Wigner position observable for Poincaré-invariant classical systems. We prove an existence and uniqueness theorem for elementary systems that parallels the well-known Newton–Wigner theorem in the quantum context. We also discuss and justify the geometric interpretation of the Newton–Wigner position as ‘centre of spin’, already proposed by Fleming in 1965 again in the quantum context.

1 Introduction

Even though we shall in this paper exclusively deal with classical (i.e. non-quantum) aspects of the Newton–Wigner position observable, we wish to start with a brief discussion of its historic origin, which is based in the early history of relativistic quantum field theory (RQFT). After that we will remark on its classical importance and give an outline of this paper.

The conceptual problem of how to properly ‘localise’ a physical system ‘in space’ has a very long history, the roots of which extend to pre-Newtonian times. Newtonian concepts of space, time, and point particles allowed for sufficiently useful localisation schemes, either in terms of the position of the particle itself if an elementary (i.e. indecomposable) systems is considered, or in terms of weighted convex sums of instantaneous particle positions for systems composed of many particles, like, e.g., the centre of mass. These concepts satisfy the expected covariance properties under spatial translations and rotations and readily translate to ordinary, Galilei-invariant Quantum Mechanics, where concepts like ‘position operators’ and the associated projection operators for positions within any measurable subset of space can be defined, again fulfilling the expected transformation rules under spatial motions.

However, serious difficulties with naive localisation concepts arose in attempts to combine Quantum Mechanics with Special Relativity. For example, as already observed in 1928 by Breit [1] and again in 1930–31 by Schrödinger [2, 3], a naive concept of ‘position’ for the Dirac equation leads to unexpected and apparently paradoxical results, like the infamous ‘Zitterbewegung’. It soon became clear that naive translations of concepts familiar from non-relativistic Quantum Mechanics did not result in satisfactory results. This had to do with the fact that spatially localised wave functions (e.g. those of compact spatial support) necessarily contained negative-energy modes in their Fourier decomposition and that negative-energy modes would necessarily be introduced if a ‘naive position operator’ (like multiplying the wave function with the position coordinate) were applied to a positive-energy state. Physically this could be seen as an inevitable result of pair production that sets in once the bounds on localisation come close to the Compton wavelength. Would that argument put an end to any further attempt to define localised states in a relativistic context?

This question was analysed and answered in the negative in 1949 by Newton and Wigner [4]. Their method was to write down axioms for what it meant that a system is ‘localised in space at a given time’ and then investigate existence as well as uniqueness for corresponding position operators. It turned out that existence and uniqueness are indeed given for elementary systems (fields being elements of irreducible representations of the Poincaré group), except for massless fields of higher helicity. A more rigorous derivation was later given by Wightman [5] who also pointed out the connection with the representation-theoretic notion of ‘imprimitivity systems’111A good text-book reference explaining the notion of imprimitivity systems is [6]..

It should be emphasised that the Newton–Wigner notion of localisation still suffers from the acausal spreading of localisation domains that is typical of fields satisfying special-relativistic wave equations, an observation made many times in the literature in one form or another; see, e.g., [7, 8, 9]. This means that if a system is Newton–Wigner localised at a point in space at a time tt, it is not strictly localised anymore in any bounded region of space at any time later than tt [4, 10]. In other words, the spatial bounds of localisation do not develop in time within the causal future of the original domain. Issues of that sort, and related ones concerning, in particular, the relation between Newton–Wigner localisation and the Reeh-Schlieder theorem in RQFT have been discussed many times in the literature even up to the more recent past, with sometimes conflicting statements as to their apparent paradoxical interpretations; see, e.g., [11] and [12, 13]. For us, these issues are not in the focus of our interest.

Clearly, due to its historical development, most discussions of Newton–Wigner localisation put their emphasis on its relevance for RQFT. This sometimes seems to mask the fact that the problem of localisation is likewise present for classical systems, in particular if they are ‘relativistic’ in the sense of Special Relativity. In fact, special-relativistic systems react with characteristic ambiguities if one tries to introduce the familiar notions of ‘centre of mass’ that one uses successfully in Newtonian physics. A first comprehensive discussion on relativistic notions of ‘position’ of various ‘centres’ was given by Pryce in his 1948 paper [14]. He starts with a list of no less than six different definitions, which Pryce labelled alphabetically from (a) through (f) and which for systems of point masses may briefly be characterised as follows: (a) and (c) correspond to taking the convex affine combination of spatial positions in each Lorentz frame with weights being equal to the rest masses and dynamical masses respectively, whereas (b) and (d) correspond to restricting this procedure to the zero-momentum frame and then transforming this position to other frames by Lorentz boosts. Possibility (e) is a combination of (c) and (d), determined by the condition that the spatial components of the ensuing position observable shall (Poisson) commute. This combination is, in fact, the Newton–Wigner position in its classical guise. Finally, possibility (f) is a variant of (b) in which the distinguished frame is not that of zero-momentum but that in which the ‘mass centre’ as defined by (a) is at rest.

In 1965, Fleming gave a more geometric discussion in [15] that highlighted the group-theoretic properties (regarding the group of spacetime automorphisms) underlying the constructions and thereby clarified many of the sometimes controversial issues regarding ‘covariance’. Fleming focussed on three position observables which he called ‘centre of inertia’, ‘centre of mass’, and the Newton–Wigner position observable, for which he, at the very end of his paper and almost in passing, suggested the name ‘centre of spin’. In our paper we shall give a more detailed geometric justification for that name.

Pryce, Fleming, and other contemporary commentators mainly had RQFT in mind as the main target for their considerations, presumably because the study of deeply relativistic classical systems was simply not considered relevant at that time. But that has clearly changed with the advent of modern relativistic astrophysics. For example, modern analytical studies of close compact binary-star systems also make use of various definitions of ‘centre of mass’ in an attempt to separate the ‘overall’ from the ‘internal’ motion as far as possible. Note that, as is well known, special-relativistic many-particle systems will generally show dynamical couplings of internal and external degrees of freedom which cannot be eliminated altogether by more clever choices of external coordinates. But, in that respect, it turns out that modern treatments of gravitationally interacting two-body systems within the theoretical framework of Hamiltonian General Relativity show a clear preference for the Newton–Wigner position [16, 17], emphasising once more its distinguished role, now in a purely classical context. In passing we mention the importance and long history connected with the ‘problem of motion’ in General Relativity, i.e. the problem of how to associate a timelike worldline with the field-theoretic evolution of an extended and structured body, a glimpse of which may be obtained by the recent collection [18]. A concise account of the various definitions of ‘centres’ that have been used in the context of General Relativity is given in [19], which also contains most of the original references in its bibliography. In our opinion, all this provides sufficient motivation for further attempts to work out the characteristic properties of Newton–Wigner localisation in the classical realm.

The plan of our investigation is as follows. After setting up our notation and conventions in section 2, where we also introduce some mathematical background, we prove a few results in section 3 which are intended to explain in what sense the Newton–Wigner position is indeed a ‘centre of spin’ and in what sense it is uniquely so (theorem 3.12). We continue in section 4 with the statement and proof of a classical analogue of the Newton–Wigner theorem, according to which the Newton–Wigner position is the unique observable satisfying a set of axioms. The result is presented in theorem 4.6 and in a slightly different formulation in theorem 4.7. They say that for a classical elementary Poincaré-invariant system with timelike four-momentum (as classified by Arens [20, 21]), there is a unique observable transforming ‘as a position should’ under translations, rotations, and time reversal, having Poisson-commuting components, and satisfying a regularity condition (being C1C^{1} on all of phase space). This observable is the Newton–Wigner position.

2 Notation and conventions

This section is meant to list our notation and conventions in the general sense, by also providing some background material on the geometric and group-theoretic setting onto which the following two sections are based.

2.1 Minkowski spacetime and the Poincaré group

We use the ‘mostly plus’ (+++)(-{++}+) signature convention for the spacetime metric and stick, as indicated, to four dimensions. This is not to say that our analysis cannot be generalised to other dimensions. In fact, as will become clear as we proceed, many of our statements have an obvious generalisation to other, in particular higher dimensions. On the other hand, as will also become clear, there are a few constructions which would definitely look different in other dimensions, like, e.g., the use of the Pauli–Lubański ‘vector’ in section 2.5, which becomes an (n3)(n-3)-form in nn dimensions, or the classification of elementary systems.

The velocity of light will be denoted by cc, and not set equal to 11. Affine Minkowski spacetime will be denoted by MM, and the corresponding vector space of ‘difference vectors’ will be denoted by VV. The Minkowski metric will be denoted by η:V×V\eta\colon V\times V\to\mathbb{R}. The isomorphism of VV with its dual space VV^{*} induced by η\eta (‘index lowering’) will be denoted by a superscript ‘flat’ symbol \flat, i.e. for a vector vVv\in V the corresponding one-form is v:=η(v,)Vv^{\flat}:=\eta(v,\cdot)\in V^{*}. The inverse isomorphism (‘index raising’) will be denoted by a superscript sharp symbol \sharp. Note that under a Lorentz transformation Λ\Lambda, vVv\in V transforms under the defining representation, (Λ,v)Λv(\Lambda,v)\mapsto\Lambda v, whereas its image vVv^{\flat}\in V^{*} under the η\eta-induced isomorphism transforms under the inverse transposed, (Λ,v)(Λ1)v=vΛ1(\Lambda,v^{\flat})\mapsto(\Lambda^{-1})^{\top}v^{\flat}=v^{\flat}\circ\Lambda^{-1}.

We fix an orientation and a time orientation on MM. The (homogeneous) Lorentz group, i.e. the group of linear isometries of (V,η)(V,\eta), will be denoted by :=𝖮(V,η)\mathcal{L}:=\mathsf{O}(V,\eta). The Poincaré group, i.e. the group of affine isometries of (M,η)(M,\eta), will be denoted by 𝒫\mathcal{P}. The proper orthochronous Lorentz and Poincaré groups (i.e. the connected components of the identity) will be denoted by +\mathcal{L}_{+}^{\uparrow} and 𝒫+\mathcal{P}_{+}^{\uparrow}, respectively222Note that speaking of just orthochronous or proper Lorentz / Poincaré transformations does not make invariant sense without specifying a time direction..

We employ standard index notation for Minkowski spacetime, using lowercase Greek letters for spacetime indices. When working with respect to bases, we will, unless otherwise stated, assume them to be positively oriented and orthonormal, and we will use 0 for the timelike and lowercase Latin letters for spatial indices. We will adhere to standard practice in physics where lowering and raising of indices are done while keeping the same kernel symbol; i.e. for a vector vVv\in V with components vμv^{\mu}, the components of the corresponding one-form vVv^{\flat}\in V^{*} will be denoted simply by vμv_{\mu}. For the sake of notational clarity, we will sometimes denote the Minkowski inner product of two vectors u,vVu,v\in V simply by

uv:=η(u,v)=uμvμ.u\cdot v:=\eta(u,v)=u_{\mu}v^{\mu}. (2.1)

We fix, once and for all, a reference point / origin oMo\in M in (affine) Minkowski spacetime, allowing us to identify MM with its corresponding vector space VV (identifying the reference point oMo\in M with the zero vector 0V0\in V, i.e. via Mx(xo)VM\ni x\mapsto(x-o)\in V), which we will do most of the time. Using the reference point oMo\in M, the Poincaré group splits as a semidirect product

𝒫=V\mathcal{P}=\mathcal{L}\ltimes V (2.2)

where the Lorentz group factor in this decomposition arises as the stabiliser of the reference point – i.e. a Poincaré transformation is considered a homogeneous Lorentz transformation if and only if it leaves oo invariant. Thus, a homogeneous Lorentz transformation Λ\Lambda\in\mathcal{L} acts on a point xMVx\in M\equiv V as (Λx)μ=Λνμxν(\Lambda x)^{\mu}=\Lambda^{\mu}_{\hphantom{\mu}\nu}x^{\nu}, and a Poincaré transformation (Λ,a)𝒫(\Lambda,a)\in\mathcal{P} acts as ((Λ,a)x)μ=Λνμxν+aμ((\Lambda,a)\cdot x)^{\mu}=\Lambda^{\mu}_{\hphantom{\mu}\nu}x^{\nu}+a^{\mu}.

We will sometimes make use of the set of spacelike hyperplanes in (affine) Minkowski spacetime MM, which we will denote by

𝖲𝗉𝖧𝖯:={ΣM:Σspacelike hyperplane}.\mathsf{SpHP}:=\{\Sigma\subset M:\Sigma\;\text{spacelike hyperplane}\}. (2.3)

Since the image of a spacelike hyperplane under a Poincaré transformation is again a spacelike hyperplane, there is a natural action of the Poincaré group on 𝖲𝗉𝖧𝖯\mathsf{SpHP}, which we will denote by ((Λ,a),Σ)(Λ,a)Σ((\Lambda,a),\Sigma)\mapsto(\Lambda,a)\cdot\Sigma and spell out in more detail in equation (3.4) below.

2.2 The Poincaré algebra

When considering the Lie algebra 𝔭\mathfrak{p} of the Poincaré group (or symplectic representations thereof), we will denote the generators of translations by PμP_{\mu} such that aμPμa^{\mu}P_{\mu} is the ‘infinitesimal transformation’ corresponding to the translation by aVa\in V, and the generators of homogeneous Lorentz transformations (with respect to the chosen origin oo) by JμνJ_{\mu\nu}, such that 12ωμνJμν-\frac{1}{2}\omega^{\mu\nu}J_{\mu\nu} is the ‘infinitesimal transformation’ corresponding to the Lorentz transformation exp(ω)+𝖦𝖫(V)\exp(\omega)\in\mathcal{L}_{+}^{\uparrow}\subset\mathsf{GL}(V) for ω𝔩=Lie()End(V)\omega\in\mathfrak{l}=\mathrm{Lie}(\mathcal{L})\subset\mathrm{End}(V).

Since we are using the (+++)(-{++}+) signature convention, the minus sign in the expression 12ωμνJμν-\frac{1}{2}\omega^{\mu\nu}J_{\mu\nu} is necessary in order that JabJ_{ab} generate rotations in the eae_{a}ebe_{b} plane from eae_{a} towards ebe_{b}, which is the convention we want to adopt. A detailed discussion of these issues regarding sign conventions for the generators of special orthogonal groups can be found in appendix A. Moreover, if uVu\in V is a future-directed unit timelike vector, then cPμuμcP_{\mu}u^{\mu} (i.e. cP0cP_{0} in the Lorentz frame defined by u=e0u=e_{0}), which is minus the energy in the frame defined by uu, is the generator of active time translations in the direction of uu. Therefore, with our conventions, for the case of causal four-momentum PVP\in V the energy (with respect to future-directed time directions) is positive if and only if PP is future-directed.

With our conventions, the commutation relations for the Poincaré generators are as follows:

[Pμ,Pν]\displaystyle[P_{\mu},P_{\nu}] =0\displaystyle=0 (2.4a)
[Jμν,Pρ]\displaystyle[J_{\mu\nu},P_{\rho}] =ημρPνηνρPμ\displaystyle=\eta_{\mu\rho}P_{\nu}-\eta_{\nu\rho}P_{\mu} (2.4b)
[Jμν,Jρσ]\displaystyle[J_{\mu\nu},J_{\rho\sigma}] =ημρJνσ+(antisymm.)\displaystyle=\eta_{\mu\rho}J_{\nu\sigma}+\text{(antisymm.)}
=(ημρJνσ(μν))(ρσ)\displaystyle=\Big{(}\eta_{\mu\rho}J_{\nu\sigma}-(\mu\leftrightarrow\nu)\Big{)}-\Big{(}\rho\leftrightarrow\sigma\Big{)} (2.4c)

As indicated, the abbreviation ‘antisymm.’, which we shall also use in the sequel of this paper, stands for the additional three terms that one obtains by first antisymmetrising (without a factor of 1/21/2) in the first pair of indices on the left hand side, here (μν)(\mu\nu), and then the ensuing combination once more in the second set of indices, here (ρσ)(\rho\sigma), again without a factor 1/21/2.

For later reference we already point out here that this Lie algebra has several convenient features, one of which being that it is perfect. This means that it equals its own derived algebra or, in other words, that each of its element is expressible as a linear combination of Lie brackets. This is easy to see directly from (2.4). Indeed, contraction of (2.4b) and (2.4) with ημρ\eta^{\mu\rho} gives (dimV1)Pν(\dim V-1)P_{\nu} in the first and (dimV2)Jνσ(\dim V-2)J_{\nu\sigma} in the second case, showing that each basis element PνP_{\nu} and JνσJ_{\nu\sigma} is a linear combination of Lie brackets if dimV>2\dim V>2. Being perfect implies that its first cohomology is trivial. Moreover, the second cohomology is also trivial. Being perfect and of trivial second cohomology will later allow us to conclude that symplectic actions are necessarily Poisson actions.

2.3 Symplectic geometry

We employ the following sign conventions for symplectic geometry (as used by Abraham and Marsden in [22], but different to those of Arnold in [23]). Let (Γ,ω)(\Gamma,\omega) be a symplectic manifold. For a smooth function fC(Γ)f\in C^{\infty}(\Gamma), we define the Hamiltonian vector field XfST(Γ)X_{f}\in ST(\Gamma) (STST denoting sections in the tangent bundle) corresponding to ff by

ιXfω:=ω(Xf,)=df,\iota_{X_{f}}\omega:=\omega(X_{f},\cdot)=\mathrm{d}f, (2.5)

where ι\iota denotes the interior product between vector fields and differential forms. The Poisson bracket of two smooth functions f,gC(Γ)f,g\in C^{\infty}(\Gamma) is then defined as

{f,g}:=ω(Xf,Xg)=df(Xg)=ιXgdf.\{f,g\}:=\omega(X_{f},X_{g})=\mathrm{d}f(X_{g})=\iota_{X_{g}}\mathrm{d}f. (2.6)

These conventions give the usual coordinate forms of the Hamiltonian flow equations and the Poisson bracket if the symplectic form ω\omega takes the coordinate form (sign-opposite to that in [23])

ω=dqadpa.\omega=\mathrm{d}q^{a}\wedge\mathrm{d}p_{a}\,. (2.7)

It is important to note that C(Γ)C^{\infty}(\Gamma) as well as ST(Γ)ST(\Gamma) are (infinite dimensional) Lie algebras with respect to the Poisson bracket and the commutator respectively, and that, with respect to these Lie structures, the map C(Γ)ST(Γ),fXfC^{\infty}(\Gamma)\to ST(\Gamma),f\mapsto X_{f} is a Lie anti-homomorphism, that is,

X{f,g}=[Xf,Xg].X_{\{f,g\}}=-\,[X_{f},X_{g}]. (2.8)

The proof is simple once one recalls from (2.5) that the Lie derivative of ω\omega with respect to any Hamiltonian vector field vanishes: LXfω=d(ιXfω)+ιXfdω=d2f=0L_{X_{f}}\omega=\mathrm{d}(\iota_{X_{f}}\omega)+\iota_{X_{f}}\mathrm{d}\omega=\mathrm{d}^{2}f=0. Therefore, d{f,g}=d(ιXgdf)=LXgdf=LXg(ιXfω)=ι[Xf,Xg]ω\mathrm{d}\{f,g\}=\mathrm{d}(\iota_{X_{g}}\mathrm{d}f)=L_{X_{g}}\mathrm{d}f=L_{X_{g}}(\iota_{X_{f}}\omega)=-\iota_{[X_{f},X_{g}]}\omega.

By saying that a one-parameter group ϕs:ΓΓ\phi_{s}\colon\Gamma\to\Gamma of symplectomorphisms is generated by a function gC(Γ)g\in C^{\infty}(\Gamma), we mean that ϕs\phi_{s} is the flow of the Hamiltonian vector field to gg, i.e. that

ddsϕs(γ)=Xg(ϕs(γ))\frac{\mathrm{d}}{\mathrm{d}s}\phi_{s}(\gamma)=X_{g}(\phi_{s}(\gamma)) (2.9)

for γΓ\gamma\in\Gamma, or equivalently

dds(fϕs)\displaystyle\frac{\mathrm{d}}{\mathrm{d}s}(f\circ\phi_{s}) =(df(Xg))ϕs\displaystyle=\Big{(}\mathrm{d}f(X_{g})\Big{)}\circ\phi_{s}
={f,g}ϕs\displaystyle=\{f,g\}\circ\phi_{s} (2.10)

for fC(Γ)f\in C^{\infty}(\Gamma). Here both sides of (2.3) are to be understood as evaluated pointwise.

2.4 Poincaré-invariant Hamiltonian systems and their momentum maps

A classical Poincaré-invariant system will be described by a phase space (Γ,ω)(\Gamma,\omega) – i.e. a symplectic manifold – with a symplectic action

Φ:𝒫×ΓΓ,((Λ,a),γ)Φ(Λ,a)(γ)\Phi\colon\mathcal{P}\times\Gamma\to\Gamma,\;((\Lambda,a),\gamma)\mapsto\Phi_{(\Lambda,a)}(\gamma) (2.11)

of the Poincaré group (in fact, for most of our purposes an action of 𝒫+\mathcal{P}_{+}^{\uparrow} is enough). We will take Φ\Phi to be a ‘left’ action, i.e. to satisfy333We refer to [24] for a detailed discussion of left versus right actions and the corresponding sign conventions that will also play an important role in the sequel of this paper.

Φ(Λ1,a1)Φ(Λ2,a2)=Φ(Λ1Λ2,a1+Λ1a2).\Phi_{(\Lambda_{1},a_{1})}\circ\Phi_{(\Lambda_{2},a_{2})}=\Phi_{(\Lambda_{1}\Lambda_{2},a_{1}+\Lambda_{1}a_{2})}\;. (2.12)

We will denote such systems as (Γ,ω,Φ)(\Gamma,\omega,\Phi).

The left action Φ\Phi of 𝒫\mathcal{P} on Γ\Gamma induces vector fields VξV_{\xi} on Γ\Gamma (the so-called ‘fundamental vector fields’), one for each ξ\xi in the Lie algebra 𝔭\mathfrak{p} of 𝒫\mathcal{P}. They are given by

Vξ(γ):=ddsΦexp(sξ)(γ)|s=0,V_{\xi}(\gamma):=\left.\frac{\mathrm{d}}{\mathrm{d}s}\Phi_{\exp(s\xi)}(\gamma)\right|_{s=0}\;, (2.13)

so that the map 𝔭ST(Γ),ξVξ\mathfrak{p}\to ST(\Gamma),\xi\mapsto V_{\xi}, given by the differential of Φ\Phi with respect to its first argument and evaluated at the group identity, is clearly linear. In fact, it is straightforward to show that it is an anti-homomorphism from the Lie algebra 𝔭\mathfrak{p} into the Lie algebra ST(M)ST(M)444Had we chosen Φ\Phi to be a right action, we would have obtained a proper Lie homomorphism; compare [24, appendix B]., i.e.

[Vξ1,Vξ2]=V[ξ1,ξ2].\bigl{[}V_{\xi_{1}},V_{\xi_{2}}\bigr{]}=-V_{[\xi_{1},\xi_{2}]}. (2.14)

Moreover, a similar calculation shows [24, appendix B]

(DΦ(Λ,a))Vξ=VAd(Λ,a)(ξ)Φ(Λ,a),(\mathrm{D}\Phi_{(\Lambda,a)})\circ V_{\xi}=V_{\mathrm{Ad}_{(\Lambda,a)}(\xi)}\circ\Phi_{(\Lambda,a)}\;, (2.15)

where DΦ(Λ,a):TΓTΓ\mathrm{D}\Phi_{(\Lambda,a)}\colon T\Gamma\to T\Gamma denotes the differential of Φ(Λ,a):ΓΓ\Phi_{(\Lambda,a)}\colon\Gamma\to\Gamma.

As 𝒫\mathcal{P} acts by symplectomorphisms, we clearly have

LVξω=0for allξ𝔭.L_{V_{\xi}}\omega=0\quad\text{for all}\;\xi\in\mathfrak{p}. (2.16)

As ω\omega is closed, the latter equation implies that ιVξω\iota_{V_{\xi}}\omega is likewise closed. Hence, by Poincaré’s lemma, locally (i.e. in a neighbourhood of each point) there exists a local function fξf_{\xi} such that dfξ=ιVξω\mathrm{d}f_{\xi}=\iota_{V_{\xi}}\omega. This function is unique up to the addition of a ξ\xi-dependent constant. Again by Poincaré’s lemma we could argue that fξf_{\xi} existed globally if Γ\Gamma were simply connected. But, fortunately, we do not need that extra assumption.

In fact, since we are dealing with a special group, the function fξf_{\xi} always exists globally, irrespective of Γ\Gamma’s topology, so that each VξV_{\xi} is a globally defined Hamiltonian vector field (i.e. each one-parameter group Φexp(sξ):ΓΓ\Phi_{\exp(s\xi)}\colon\Gamma\to\Gamma of symplectomorphisms is generated, in the sense of (2.3), by the corresponding function fξf_{\xi}). Moreover, the constants up to which the collection of fξf_{\xi} is defined can be chosen in such a way that the map ξfξ\xi\mapsto f_{\xi} from the Lie algebra 𝔭\mathfrak{p} to the Lie algebra C(Γ)C^{\infty}(\Gamma) (the Lie product of the latter being the Poisson bracket) is a Lie homomorphism, i.e.

{fξ1,fξ2}=f[ξ1,ξ2].\left\{f_{\xi_{1}},f_{\xi_{2}}\right\}=f_{[\xi_{1},\xi_{2}]}. (2.17)

This clearly fixes the constants uniquely. Note that, according to (2.14) and (2.8), both maps, ξVξ\xi\mapsto V_{\xi} and VξfξV_{\xi}\mapsto f_{\xi}, are Lie anti-homomorphisms. Hence their combination ξfξ\xi\mapsto f_{\xi} is a proper Lie homomorphism (no minus sign on the right-hand side of (2.17)).

A symplectic action of a group whose generating vector fields are globally Hamiltonian and satisfy (2.17) is called a Poisson action. The statement made here is that if dimV>2\dim V>2, any symplectic action of the Poincaré of group is always a Poisson action. This is a non-trivial statement depending crucially on properties of the groups’s Lie algebra. For example, it would fail to hold for the Galilei group (homogeneous as well as inhomogeneous) which, despite being just a contraction of the Poincaré group, behaves quite differently in that matter and, consequently, also as regards the problem of localisation [25, 5].

The underlying reason for why fξf_{\xi} exists globally is that 𝔭\mathfrak{p} is perfect, as already shown above. Indeed, the proof is quite simple: Since ξ=[ξ1,ξ2]\xi=[\xi_{1},\xi_{2}] (or sums of such commutators) we have Vξ=[Vξ1,Vξ2]V_{\xi}=-[V_{\xi_{1}},V_{\xi_{2}}] and hence dfξ=ι[Vξ1,Vξ2]ω=LVξ1(ιVξ2ω)=d(ω(Vξ1,Vξ2))\mathrm{d}f_{\xi}=-\iota_{[V_{\xi_{1}},V_{\xi_{2}}]}\omega=-L_{V_{\xi_{1}}}(\iota_{V_{\xi_{2}}}\omega)=\mathrm{d}(\omega(V_{\xi_{1}},V_{\xi_{2}})), so that fξ=ω(Vξ1,Vξ2)+const.f_{\xi}=\omega(V_{\xi_{1}},V_{\xi_{2}})+\text{const.} which is globally defined. The other statement concerning the choice of constants that guarantee (2.17) is an immediate consequence of the triviality of the second cohomology of 𝔭\mathfrak{p}, the proof of which may, e.g., be looked up in [26, § 3.3].

Having established global existence and uniqueness of the generators fξf_{\xi} satisfying ω(Vξ,)=dfξ\omega(V_{\xi},\cdot)=\mathrm{d}f_{\xi}, we can now deduce the transformation property of fξf_{\xi} under the action of 𝒫\mathcal{P}. Taking the pullback of the equation ω(Vξ,)=dfξ\omega(V_{\xi},\cdot)=\mathrm{d}f_{\xi} with Φ(Λ,a)1\Phi_{(\Lambda,a)^{-1}} and using the invariance of ω\omega as well as (2.15), we immediately deduce

Φ(Λ,a)1fξ:=fξΦ(Λ,a)1=fAd(Λ,a)(ξ),\Phi_{(\Lambda,a)^{-1}}^{*}f_{\xi}:=f_{\xi}\circ\Phi_{(\Lambda,a)^{-1}}=f_{\mathrm{Ad}_{(\Lambda,a)}(\xi)}\;, (2.18)

which may also be read as the invariance of the real-valued function f:𝔭×Γf\colon\mathfrak{p}\times\Gamma\to\mathbb{R}, (ξ,γ)fξ(γ)(\xi,\gamma)\mapsto f_{\xi}(\gamma), under the combined left action of 𝒫\mathcal{P} on 𝔭×Γ\mathfrak{p}\times\Gamma given by Ad×Φ\mathrm{Ad}\times\Phi. Alternatively, since ξfξ\xi\mapsto f_{\xi} is linear, we may regard ff as 𝔭\mathfrak{p}^{*}-valued function on Γ\Gamma, where 𝔭\mathfrak{p}^{*} denotes the vector space dual to 𝔭\mathfrak{p}. This map is called the momentum map555See [22, chap. 4.2] for a general discussion on the notion of ‘momentum map’ and also [24] for an account of its use and properties restricted to the case of Poincaré-invariant systems. for the given system (Γ,ω,Φ)(\Gamma,\omega,\Phi), which according to (2.18) is then Ad\mathrm{Ad}^{*}-equivariant:

fΦ(Λ,a)=Ad(Λ,a)fAd(Λ,a)fΦ(Λ,a)1=ff\circ\Phi_{(\Lambda,a)}=\mathrm{Ad}^{*}_{(\Lambda,a)}\circ f\iff\mathrm{Ad}^{*}_{(\Lambda,a)}\circ f\circ\Phi_{(\Lambda,a)^{-1}}=f (2.19)

The second expression is again meant to stress that the condition of equivariance is equivalent to the invariance of the function ff under the combined left actions in its domain and target spaces (invariance of the graph). Note that Ad\mathrm{Ad}^{*} denotes the co-adjoint representation of 𝒫\mathcal{P} on 𝔭\mathfrak{p}^{*}, given by Ad(Λ,a):=(Ad(Λ,a)1)\mathrm{Ad}^{*}_{(\Lambda,a)}:=(\mathrm{Ad}_{(\Lambda,a)^{-1}})^{\top} with superscript \top denoting the transposed map.

Points in Γ\Gamma faithfully represent the state of the physical system whereas observables correspond to functions on Γ\Gamma. In order to implement time evolution we shall employ a ‘classical Heisenberg picture’, in which the phase space point remains the same at all times, whereas the evolution will correspond to the changes of observables according to their association to different spacelike hyperplanes in spacetime. Although this is different from the (‘Schrödinger picture’) approach usually taken in classical mechanics (where the state of the system is given by a phase space point changing in ‘time’, which is an external parameter), this point of view is clearly better adapted to the Poincaré-relativistic framework, in which there simply is no absolute notion of time.

Choosing a set of ten basis vectors (Pμ,Jμν)(P_{\mu},J_{\mu\nu}) for 𝔭\mathfrak{p} obeying (2.4) (compare appendix A), we can contract the 𝔭\mathfrak{p}^{*}-valued momentum map with each of these basis vectors in order to obtain the corresponding ten real-valued component functions of the momentum map. By some abuse of notation we shall call these component functions by the same letters (Pμ,Jμν)(P_{\mu},J_{\mu\nu}) as the Lie algebra elements themselves. Equation (2.17) now says that the map that sends the Lie algebra elements PμP_{\mu} and JμνJ_{\mu\nu} in 𝔭\mathfrak{p} to the corresponding component functions of the momentum map is a Lie homomorphism from 𝔭\mathfrak{p} to the Lie algebra C(Γ,)C^{\infty}(\Gamma,\mathbb{R}) (the latter with Poisson bracket as Lie multiplication):

{Pμ,Pν}\displaystyle\{P_{\mu},P_{\nu}\} =0\displaystyle=0 (2.20a)
{Jμν,Pρ}\displaystyle\{J_{\mu\nu},P_{\rho}\} =ημρPνηνρPμ\displaystyle=\eta_{\mu\rho}P_{\nu}-\eta_{\nu\rho}P_{\mu} (2.20b)
{Jμν,Jρσ}\displaystyle\{J_{\mu\nu},J_{\rho\sigma}\} =ημρJνσ+(antisymm.)\displaystyle=\eta_{\mu\rho}J_{\nu\sigma}+\text{(antisymm.)} (2.20c)

The Ad\mathrm{Ad}^{*}-equivariance of the momentum map can now be written down in component form if we first set ξ=Pμ\xi=P_{\mu} and then ξ=Jμν\xi=J_{\mu\nu}. Indeed, considering (2.18) and recalling our abuse of notation in denoting the real-valued phase space functions fPμf_{P_{\mu}} and fJμνf_{J_{\mu\nu}} again with the letters PμP_{\mu} and JμνJ_{\mu\nu}, we can immediately read from equation (B.8) of appendix B, in which we need to replace eae_{a} with PμP_{\mu} and BabB_{ab} with Jμν-J_{\mu\nu} according to (A.15) of appendix A, that

PμΦ(Λ,a)\displaystyle P_{\mu}\circ\Phi_{(\Lambda,a)} =(Λ1)μνPν,\displaystyle=(\Lambda^{-1})^{\nu}_{\phantom{\nu}\mu}\,P_{\nu}\;, (2.21a)
JμνΦ(Λ,a)\displaystyle J_{\mu\nu}\circ\Phi_{(\Lambda,a)} =(Λ1)μρ(Λ1)νσJρσ+aμ(Λ1)νρPρaν(Λ1)μρPρ.\displaystyle=(\Lambda^{-1})^{\rho}_{\phantom{\rho}\mu}(\Lambda^{-1})^{\sigma}_{\phantom{\sigma}\nu}\,J_{\rho\sigma}+a_{\mu}(\Lambda^{-1})^{\rho}_{\phantom{\rho}\nu}\,P_{\rho}-a_{\nu}(\Lambda^{-1})^{\rho}_{\phantom{\rho}\mu}\,P_{\rho}\;. (2.21b)

Note that the left-hand sides of (2.21) are precisely what we need; that is, we need the composition with Φ(Λ,a)\Phi_{(\Lambda,a)} rather than Φ(Λ,a)1\Phi_{(\Lambda,a)^{-1}} to evaluate the momenta PμP_{\mu} and JμνJ_{\mu\nu} on the actively Poincaré-displaced phase space points. Note also that if we had put the indices upstairs and had used, e.g., Pμ=ημνPνP^{\mu}=\eta^{\mu\nu}P_{\nu} rather than PμP_{\mu} then the right-hand side of (2.21a) would read ΛνμPν\Lambda^{\mu}_{\phantom{\mu}\nu}\,P^{\nu}, and correspondingly in (2.21b). Finally recall that the last term on the right-hand side of (2.21b) just reflects the familiar transformation of angular momentum (the momentum associated to spatial rotations) under spatial translations, which is typical for the co-adjoint representation, which here gets extended to the momentum associated to boost transformations666One easily checks that the signs are right: translating a system whose momentum points in yy-direction by a positive amount into the xx-direction should enhance the angular momentum in zz-direction. This is just what (2.21b) implies..

2.5 The Pauli–Lubański vector

Given a classical Poincaré-invariant system, the Pauli–Lubański vector WW is the VV-valued phase space function defined in components by

Wμ=12εμνρσPνJρσW_{\mu}=-\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}P^{\nu}J^{\rho\sigma} (2.22)

where ε\varepsilon denotes the volume form of Minkowski space (whose components in a positively oriented orthonormal basis are just given by the usual totally antisymmetric symbol, with ε0123=+1\varepsilon_{0123}=+1). The sign convention in this definition can be understood as follows. We imagine a situation in which PP is timelike and future-directed (positive energy, see above), and consider the spatial components of WW with respect to an orthonormal basis {e0,,e3}\{e_{0},\dots,e_{3}\} of VV with (e0)μ=Pμ/PνPν(e_{0})^{\mu}=P^{\mu}/\sqrt{-P_{\nu}P^{\nu}} (‘momentum rest frame’). For those, we obtain

WaPμPμ=12εa0ρσJρσ=12εabc(3)Jbc\frac{W_{a}}{\sqrt{-P_{\mu}P^{\mu}}}=-\frac{1}{2}\varepsilon_{a0\rho\sigma}J^{\rho\sigma}=\frac{1}{2}{{}^{(3)}\varepsilon}_{abc}J^{bc} (2.23)

where the εabc(3){{}^{(3)}\varepsilon}_{abc} is the three-dimensional antisymmetric symbol / the components of the spatial volume form. Thus, since Jbc=JbcJ^{bc}=J_{bc} generates rotations from ebe_{b} towards ece_{c}, we see that Wa/PμPμW_{a}/\sqrt{-P_{\mu}P^{\mu}} generates rotations ‘along the eae_{a} axis’ in the usual, three-dimensional sense. Thus, W/PμPμW/\sqrt{-P_{\mu}P^{\mu}} can be interpreted as the ‘spatial spin vector’ in the momentum rest frame, which is the usual interpretation of the Pauli–Lubański vector.

Rewriting the definition of WW as

Wμ=12εμνρσPνJρσ=12ενρσμPνJρσ=13!ενρσμ(PJ)νρσ,W_{\mu}=-\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}P^{\nu}J^{\rho\sigma}=\frac{1}{2}\varepsilon_{\nu\rho\sigma\mu}P^{\nu}J^{\rho\sigma}=\frac{1}{3!}\varepsilon_{\nu\rho\sigma\mu}(P^{\flat}\wedge J)^{\nu\rho\sigma}, (2.24)

we see that in the language of exterior algebra

W=((PJ))W=(*(P^{\flat}\wedge J))^{\sharp} (2.25)

where * is the Hodge star operator. Here we use the standard sign conventions for the Hodge operator, i.e. the definition αβ=η(α,β)ε\alpha\wedge*\beta=\eta(\alpha,\beta)\,\varepsilon; see for example [27] or [24, appendix A].

3 The Newton–Wigner position as a ‘centre of spin’

In this section we will explain our understanding and present our geometric clarification of Fleming’s statement in [15] that the Newton–Wigner position may be understood as a ‘centre of spin’. To this end, we introduce Fleming’s geometric framework for special-relativistic position observables, and then discuss the definition of position observables by spin supplementary conditions (SSCs). Finally, we introduce the notion of a position observable being a ‘centre of spin’, and prove that the Newton–Wigner position is the only continuous position observable defined by an SSC that represents a centre of spin in that sense.

3.1 Position observables on spacelike hyperplanes

We start by describing the general framework developed by Fleming in [15] and also [28] for the description of special-relativistic position observables, translated to our case of classical systems from Fleming’s quantum language. Consider a classical Poincaré-invariant system (Γ,ω,Φ)(\Gamma,\omega,\Phi). By a position observable χ\chi for this system we understand a ‘procedure’ which, given any spacelike hyperplane Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP} in (affine) Minkowski spacetime, allows us to ‘localise’ the system on Σ\Sigma. More precisely, this means that for any Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP}, we have an MM-valued phase space function

χ(Σ):ΓM\chi(\Sigma)\colon\Gamma\to M (3.1)

with image contained in Σ\Sigma, whose value χ(Σ)(γ)\chi(\Sigma)(\gamma) for γΓ\gamma\in\Gamma is to be interpreted as the ‘χ\chi-position’ of our system in state γ\gamma on the hyperplane Σ\Sigma.

Any spacelike hyperplane Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP} is uniquely characterised by its (timelike) future-directed unit normal uVu\in V and its distance τ\tau\in\mathbb{R} to the origin oMo\in M, measured along the straight line through oo in direction uu. In terms of these, it has the form

Σ={xM:uμxμ=τ},\Sigma=\{x\in M:u_{\mu}x^{\mu}=-\tau\}, (3.2)

where we identified MM with VV. From now on, whenever convenient, we will identify Σ\Sigma with the tuple (u,τ)(u,\tau). The condition that the image of χ(Σ)\chi(\Sigma) be contained in Σ\Sigma then takes the form

uμχμ(u,τ)(γ)=τ.u_{\mu}\chi^{\mu}(u,\tau)(\gamma)=-\tau. (3.3)

We can now also spell out explicitly the left action of 𝒫\mathcal{P} on 𝖲𝗉𝖧𝖯\mathsf{SpHP} that is induced from the left action of 𝒫\mathcal{P} on MM (as already mentioned below equation (2.3)):

(Λ,a)(u,τ)=(Λu,τΛua)(\Lambda,a)\cdot(u,\tau)=(\Lambda u,\tau-\Lambda u\cdot a) (3.4)

One easily checks that this indeed defines a left action, i.e. (Λ1,a1)[(Λ2,a2)(u,τ)]=(Λ1Λ2,a1+Λ1a2)(u,τ)(\Lambda_{1},a_{1})\cdot[(\Lambda_{2},a_{2})\cdot(u,\tau)]=(\Lambda_{1}\Lambda_{2},a_{1}+\Lambda_{1}a_{2})\cdot(u,\tau).

Fixing uu and varying τ\tau in (3.2), we obtain the spacelike hyperplanes corresponding to different ‘instants of time’ τ\tau in the Lorentz frame corresponding to uu. Thus, for a fixed state γΓ\gamma\in\Gamma and fixed frame uu, the set

{χ(u,τ)(γ):τ}M\{\chi(u,\tau)(\gamma):\tau\in\mathbb{R}\}\subset M (3.5)

gives the ‘worldline’ of the χ\chi-position of the system. Following Fleming [15], who says that this is a requirement ‘easily agreed upon’, we require that this worldline should be parallel to the four-momentum777This assumption is natural for closed systems as we consider here. For non-closed systems, i.e. systems without local energy–momentum conservation, the four-velocity is in general not parallel to the four-momentum; see, e.g., the discussion at the beginning of section 2.6 in [29]., i.e. χ(u,τ)τP\frac{\partial\chi(u,\tau)}{\partial\tau}\propto P. Together with (3.3), this implies condition (3.8) in the definition below, which is meant to sum up all the preceding considerations.

Definition 3.1.

A position observable for a classical Poincaré-invariant system (Γ,ω,Φ)(\Gamma,\omega,\Phi) with causal four-momentum is a map

χ:𝖲𝗉𝖧𝖯×ΓM,(Σ,γ)χ(Σ)(γ)\chi\colon\mathsf{SpHP}\times\Gamma\to M,\;(\Sigma,\gamma)\mapsto\chi(\Sigma)(\gamma) (3.6)

satisfying

χ(Σ)(γ)Σ\chi(\Sigma)(\gamma)\in\Sigma (3.7)

for all Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP} and all γΓ\gamma\in\Gamma (or, equivalently, (3.3)), as well as

χμ(u,τ)τ=1(uP)Pμ.\frac{\partial\chi_{\mu}(u,\tau)}{\partial\tau}=\frac{1}{(-u\cdot P)}P_{\mu}\;. (3.8)

For fixed Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP}, we will often view χ(Σ):ΓM\chi(\Sigma)\colon\Gamma\to M as a phase space function in its own right.

Note that (3.8) and (3.3) imply that the four-momentum must be causal for such a position observable to exist.

In addition to the demands of the positions χ(Σ)\chi(\Sigma) being located on Σ\Sigma and of ‘worldlines’ in direction of the four-momentum, Fleming also introduces the following covariance requirement (which we, different to Fleming, do not include in the definition of a position observable):

Definition 3.2.

A position observable for a classical Poincaré-invariant system (Γ,ω,Φ)(\Gamma,\omega,\Phi) is said to be covariant if and only if

χ((Λ,a)Σ)(Φ(Λ,a)(γ))=(Λ,a)(χ(Σ)(γ))\chi\Big{(}(\Lambda,a)\cdot\Sigma\Big{)}\Big{(}\Phi_{(\Lambda,a)}(\gamma)\Big{)}=(\Lambda,a)\cdot\Big{(}\chi(\Sigma)(\gamma)\Big{)} (3.9)

for all Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP}, γΓ\gamma\in\Gamma and (Λ,a)𝒫(\Lambda,a)\in\mathcal{P}. This can be read concisely as saying that the map (3.6) is invariant under the natural left action induced from those on the domain and target spaces (invariance of χ\chi’s graph):

χ=(Λ,a)χ((Λ,a)1×Φ(Λ,a)1).\chi=(\Lambda,a)\circ\chi\circ\left((\Lambda,a)^{-1}\times\Phi_{(\Lambda,a)^{-1}}\right). (3.10)

This is indeed a sensible notion of covariance: it demands that, for any Poincaré transformation (Λ,a)(\Lambda,a), the χ\chi-position of the transformed system Φ(Λ,a)(γ)\Phi_{(\Lambda,a)}(\gamma) on the transformed hyperplane (Λ,a)Σ(\Lambda,a)\cdot\Sigma be the transform of the ‘original position’ χ(Σ)(γ)\chi(\Sigma)(\gamma). In terms of components, (3.9) assumes the form

χμ(Λu,τΛua)Φ(Λ,a)=Λνμχν(u,τ)+aμ,\chi^{\mu}(\Lambda u,\tau-\Lambda u\cdot a)\circ\Phi_{(\Lambda,a)}=\Lambda^{\mu}_{\hphantom{\mu}\nu}\chi^{\nu}(u,\tau)+a^{\mu}\,, (3.11)

taking into account (3.4).

3.2 Spin supplementary conditions

The most important and widely used procedure to define special-relativistic position observables is by so-called spin supplementary conditions. Suppose we are given a causal, future-directed vector PVP\in V and an antisymmetric 2-tensor J2VJ\in\bigwedge^{2}V^{*}, describing the four-momentum and the angular momentum (with respect to the origin oMo\in M) of some physical system. For any future-directed timelike vector fVf\in V, we then consider the equation

0=Sμνfν0=S_{\mu\nu}f^{\nu} (3.12)

with Sμν:=JμνxμPν+xνPμS_{\mu\nu}:=J_{\mu\nu}-x_{\mu}P_{\nu}+x_{\nu}P_{\mu}, which we view as an equation for xMx\in M. Since SS is the angular momentum tensor with respect to the reference point xx (instead of the origin oo as for JJ), or the spin tensor with respect to xx, (3.12) is called the spin supplementary condition (SSC) with respect to ff. As is well-known (and easily verified), the set of its solutions xx is a line in MM with tangent PP, namely

{xM:0=Sμνfν}={xM:xμ=JμρfρfP+λPμwithλ}.\{x\in M:0=S_{\mu\nu}f^{\nu}\}=\left\{x\in M:x_{\mu}=\frac{J_{\mu\rho}f^{\rho}}{f\cdot P}+\lambda P_{\mu}\;\text{with}\;\lambda\in\mathbb{R}\right\}. (3.13)

This line can be given the interpretation of the ‘centre of energy’ worldline of our system with respect to the Lorentz frame defined by ff. See [19] and references therein for further discussion on the interpretation and impact of various SSCs as regards equations of motion in General Relativity.

The idea is now to explicitly combine the SSC-based approach with Fleming’s geometric ideas, thereby introducing the two independent parameters ff from (3.13) and uu from (3.2). We define a position observable in the sense of definition 3.1 in the following way: given a classical Poincaré-invariant system (Γ,ω,Φ)(\Gamma,\omega,\Phi) with causal four-momentum and a state γΓ\gamma\in\Gamma, we consider the SSC worldline defined by (3.12) where we now take Pμ(γ)P_{\mu}(\gamma) for the four-momentum and Jμν(γ)J_{\mu\nu}(\gamma) for the angular momentum tensor. We then simply define χ(Σ)(γ)\chi(\Sigma)(\gamma) to be the intersection of this worldline with the hyperplane Σ=(u,τ)\Sigma=(u,\tau). This means that we take the x(λ)x(\lambda) from (3.13) and determine the parameter λ\lambda from (3.3), i.e. from x(λ)u+τ=0x(\lambda)\cdot u+\tau=0. Inserting the λ=λ(u,τ)\lambda=\lambda(u,\tau) so determined leads to

Definition 3.3.

The SSC position observable with respect to ff is given by

χμ(u,τ)=JμρfρfP+τPμ(uP)Jλρuλfρ(fP)Pμ(uP).\chi_{\mu}(u,\tau)=\frac{J_{\mu\rho}f^{\rho}}{f\cdot P}+\frac{\tau P_{\mu}}{(-u\cdot P)}-\frac{J_{\lambda\rho}u^{\lambda}f^{\rho}}{(-f\cdot P)}\,\frac{P_{\mu}}{(-u\cdot P)}\;. (3.14)

Let us again stress the interpretation of this expression: it is the SSC position with respect to ff (i.e. a point on the ‘centre of energy’ worldline with respect to ff) as localised on the hyperplane characterised by unit normal uu and distance τ\tau to the origin, i.e. as seen in the Lorentz frame with respect to uu at ‘time’ τ\tau.

Note that for this definition to make sense, ff does not have to be a fixed timelike future-directed vector: it can depend on the normal uu (and could even depend on τ\tau), and it can also depend on phase space888Various choices for ff were given distinguished names in the literature. The main ones, different from the Newton–Wigner condition to be discussed here, are as follows. If ff is meant to just characterise a fixed ‘laboratory frame’, which may be preferred for any reason, like rotational symmetries in that frame, the SSC is named after Corinaldesi & Papapetrou [30]. If ff is proportional to the total linear momentum of the system, the SSC is named after Tulczyjew [31] and Dixon [32]. If ff is chosen in a somewhat self-referential way to be the four-velocity of the worldline that is to be determined by the very SSC containing that ff, the condition is named after Frenkel [33], Mathisson [34, 35], and Pirani [36, 37].. Of course this means that according to this dependence of ff, we will possibly be considering different worldlines for different choices of uu.

Example 3.4.
  1. (i)

    Choosing f=uf=u, we are considering, for each uu, the SSC worldline with respect to uu, i.e. the centre of energy worldline999Note that it was called ‘centre of mass’ by Fleming [15]. with respect to uu. Using (3.14), the centre of energy position observable has the form

    χμCE(u,τ)=JμρuρuP+τPμ(uP).\chi^{\mathrm{CE}}_{\mu}(u,\tau)=\frac{J_{\mu\rho}u^{\rho}}{u\cdot P}+\frac{\tau P_{\mu}}{(-u\cdot P)}\;. (3.15)
  2. (ii)

    In the case of timelike four-momentum, we can choose f=Pf=P the four-momentum (the Tulczyjew–Dixon SSC), such that the corresponding SSC worldline is the centre of energy worldline in the momentum rest frame of the system. This worldline, which is obviously independent of uu, was called the centre of inertia worldline by Fleming [15]. The centre of inertia has the form

    χμCI(u,τ)=JμρPρm2c2+τPμ(uP)JλρuλPρm2c2Pμ(uP),\chi^{\mathrm{CI}}_{\mu}(u,\tau)=-\frac{J_{\mu\rho}P^{\rho}}{m^{2}c^{2}}+\frac{\tau P_{\mu}}{(-u\cdot P)}-\frac{J_{\lambda\rho}u^{\lambda}P^{\rho}}{m^{2}c^{2}}\;\frac{P_{\mu}}{(-u\cdot P)}\;, (3.16)

    where m=P2/cm=\sqrt{-P^{2}}/c is the mass of the system.

  3. (iii)

    Choosing f=u+Pmcf=u+\frac{P}{mc} where m=P2/cm=\sqrt{-P^{2}}/c is the mass of the system (again only possible in the case of timelike four-momentum), we obtain the Newton–Wigner position observable. Evaluating (3.14), it has the form

    χμNW(u,τ)=Jμρ(uρ+Pρmc)mcuP+τPμ(uP)JλρuλPρmc(mcuP)Pμ(uP).\chi_{\mu}^{\mathrm{NW}}(u,\tau)=-\frac{J_{\mu\rho}\left(u^{\rho}+\frac{P^{\rho}}{mc}\right)}{mc-u\cdot P}+\frac{\tau P_{\mu}}{(-u\cdot P)}-\frac{J_{\lambda\rho}u^{\lambda}P^{\rho}}{mc(mc-u\cdot P)}\,\frac{P_{\mu}}{(-u\cdot P)}\;. (3.17)

Of course, the SSC position observable (3.14) will generally not be covariant in the sense of definition 3.2 unless ff is also assumed to transform appropriately. If ff depends on Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP} and γΓ\gamma\in\Gamma and takes values in VV it seems obvious that for the resulting position to be covariant ff itself must be a covariant function under the combined actions on its domain and target spaces. Indeed, we have

Proposition 3.5.

If the vector ff defining the SSC position observable χ\chi is a function

f:𝖲𝗉𝖧𝖯×ΓV,(Σ,γ)f(Σ)(γ),f\colon\mathsf{SpHP}\times\Gamma\to V,\quad(\Sigma,\gamma)\mapsto f(\Sigma)(\gamma), (3.18)

such that

f((Λ,a)Σ)(Φ(Λ,a)(γ))=Λ(f(Σ)(γ))f\Big{(}(\Lambda,a)\cdot\Sigma\Big{)}\Big{(}\Phi_{(\Lambda,a)}(\gamma)\Big{)}=\Lambda\cdot\Big{(}f(\Sigma)(\gamma)\Big{)} (3.19)

for all Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP}, γΓ\gamma\in\Gamma, and (Λ,a)𝒫(\Lambda,a)\in\mathcal{P}, then χ\chi is a covariant position observable. Again we note that, just like in the transition from (3.9) to (3.10), we may rewrite (3.19) equivalently as expressing the invariance of ff (i.e. its graph) under simultaneous actions on its domain and target spaces (using that translations act trivially on the target space VV):

f=Λf((Λ,a)1×Φ(Λ,a)1)f=\Lambda\circ f\circ\left((\Lambda,a)^{-1}\times\Phi_{(\Lambda,a)^{-1}}\right) (3.20)
Proof.

At first, suppose we are given a future-directed timelike four-momentum PVP\in V and an angular momentum tensor J2VJ\in\bigwedge^{2}V^{*}, as well as a future-directed timelike vector ff for the definition of an SSC. In addition, fix a Poincaré transformation (Λ,a)𝒫(\Lambda,a)\in\mathcal{P}. If we now consider (a) the SSC worldline for PP and JJ with respect to ff, and (b) the SSC worldline for the transformed four-momentum P=ΛPP^{\prime}=\Lambda P and angular momentum J=((Λ1)(Λ1))J+a(Λ1)PJ^{\prime}=((\Lambda^{-1})^{\top}\otimes(\Lambda^{-1})^{\top})J+a^{\flat}\wedge(\Lambda^{-1})^{\top}P^{\flat} (compare (2.21b)) with respect to the transformed vector Λf\Lambda f, it is easy to check that the second worldline is the Poincaré transform by (Λ,a)(\Lambda,a) of the first. That is, by Poincaré transforming the four-momentum and angular momentum of the system as well as the ‘direction vector’ for the SSC, we Poincaré transform the SSC worldline.

Now, the SSC position χ(Σ)(γ)\chi(\Sigma)(\gamma) is defined to be the intersection of the hyperplane Σ\Sigma with the SSC worldline of γ\gamma with respect to f(Σ)(γ)f(\Sigma)(\gamma). Thus, the ‘new position’

χ((Λ,a)Σ)(Φ(Λ,a)(γ))\chi\Big{(}(\Lambda,a)\cdot\Sigma\Big{)}\Big{(}\Phi_{(\Lambda,a)}(\gamma)\Big{)} (3.21)

is the intersection of the transformed hyperplane (Λ,a)Σ(\Lambda,a)\cdot\Sigma with the SSC worldline of the transformed system Φ(Λ,a)(γ)\Phi_{(\Lambda,a)}(\gamma) with respect to the transformed vector Λ(f(Σ)(γ))\Lambda\cdot\Big{(}f(\Sigma)(\gamma)\Big{)}, where we used the covariance requirement (3.19). But according to our earlier considerations, this means that the ‘new position’ is the intersection of the transformed hyperplane with the transform of the original SSC worldline – i.e. the transform of the original position χ(Σ)(γ)\chi(\Sigma)(\gamma). This means that the position observable is covariant. ∎

Since the vectors defining the centre of energy, the centre of inertia and the Newton–Wigner position satisfy (3.19), all of these are covariant position observables. We stress once more that for this to be true we need to take into account the action of the Poincaré group on 𝖲𝗉𝖧𝖯\mathsf{SpHP}. This remark is particularly relevant in the Newton–Wigner case, in which ff is the sum of two vectors, uu and P/(mc)P/(mc), the first being associated to an element of 𝖲𝗉𝖧𝖯\mathsf{SpHP} and the second to an element of Γ\Gamma. Covariance cannot be expected to hold for non-trivial actions on Γ\Gamma alone. In the next section we will offer an insight as to why this somewhat ‘hybrid’ combination for ff in terms of an ‘external’ vector uu and an ‘internal’ vector P/(mc)P/(mc) appears. The latter is internal, or dynamical, in the sense that it is defined entirely by the physical state of the system, i.e. a point in Γ\Gamma, while the former is external, or kinematical, in the sense that it refers to the choice of Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP}, which is entirely independent of the physical system and its state.

Finally, we will need the following well-known result for SSCs with respect to different vectors ff, which was first shown by Møller in 1949 in [38]; see also [24, theorem 17] for a recent and more geometric discussion:

Theorem 3.6 (Møller disc and radius).

Suppose we are given the future-directed timelike four-momentum vector PVP\in V and the angular momentum tensor J2VJ\in\bigwedge^{2}V^{*} of some physical system. Consider the bundle of all possible SSC worldlines (3.13) for this system, defined by considering all future-directed timelike vectors ff. The intersection of this bundle with any hyperplane Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP} orthogonal to PP is a two-dimensional disc (the so-called Møller disc) in the plane orthogonal to the Pauli–Lubański vector W=((PJ))W=(*(P^{\flat}\wedge J))^{\sharp}, whose centre is the centre of inertia on Σ\Sigma and whose radius is the Møller radius

RM=Smc,R_{M}=\frac{S}{mc}\;, (3.22)

where S=W2/(mc)S=\sqrt{W^{2}}/(mc) is the spin of the system and m=P2/cm=\sqrt{-P^{2}}/c its mass.

3.3 The centre of spin condition

For a system with timelike four-momentum, the Pauli–Lubański vector WW has the interpretation of being (mcmc times) the spin vector in the momentum rest frame. We now define the spin vector in an arbitrary Lorentz frame by boosting W/(mc)W/(mc) to the new frame:

Definition 3.7.

Given the timelike four-momentum PVP\in V and the Pauli–Lubański vector WPW\in P^{\perp} of a physical system, its spin vector in the Lorentz frame given by the future-directed unit timelike vector uu is

s(u):=B(u)Wmc,s(u):=B(u)\cdot\frac{W}{mc}\;, (3.23)

where B(u)+B(u)\in\mathcal{L}_{+}^{\uparrow} is the unique Lorentz boost with respect to Pmc\frac{P}{mc} (i.e. containing Pmc\frac{P}{mc} in its timelike 2-plane of action) that maps Pmc\frac{P}{mc} to uu, with m=P2/cm=\sqrt{-P^{2}}/c being the mass. In terms of components, this boost is given by101010Generally, given two unit timelike future-pointing vectors n1n_{1} and n2n_{2}, then the boost that maps n1n_{1} onto n2n_{2} and fixes the spacelike plane orthogonal to span{n1,n2}\mathrm{span}\{n_{1},n_{2}\} is given by the combination ρn1+n2ρn1\rho_{n_{1}+n_{2}}\circ\rho_{n_{1}} of two hyperplane-reflections, where ρn:=idV2nnn2\rho_{n}:=\mathrm{id}_{V}-2\frac{n\otimes n^{\flat}}{n^{2}} is the reflection at the hyperplane orthogonal to nn. Setting n1=P/(mc)n_{1}=P/(mc) and n2=un_{2}=u gives (3.24).

Bνμ(u)=δνμ+(Pμmc+uμ)(Pνmc+uν)1uPmc2uμPνmc.B^{\mu}_{\hphantom{\mu}\nu}(u)=\delta^{\mu}_{\nu}+\frac{\left(\frac{P^{\mu}}{mc}+u^{\mu}\right)\left(\frac{P_{\nu}}{mc}+u_{\nu}\right)}{1-u\cdot\frac{P}{mc}}-2\frac{u^{\mu}P_{\nu}}{mc}\;. (3.24)
Definition 3.8.

A centre of spin position observable for a classical Poincaré-invariant system (Γ,ω,Φ)(\Gamma,\omega,\Phi) with timelike four-momentum is a position observable χ\chi satisfying

sμ(u)=12εμνρσuνSρσ(u),s_{\mu}(u)=-\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}u^{\nu}S^{\rho\sigma}(u), (3.25)

where Sρσ(u):=Jμνχμ(u,τ)Pν+χν(u,τ)PμS_{\rho\sigma}(u):=J_{\mu\nu}-\chi_{\mu}(u,\tau)P_{\nu}+\chi_{\nu}(u,\tau)P_{\mu} is the spin tensor111111Since χ(u,τ)τ\frac{\partial\chi(u,\tau)}{\partial\tau} is proportional to PP, the spin tensor is independent of τ\tau. with respect to χ\chi. Expressed in terms of the Hodge operator, this condition reads

s(u)=((uS(u))).s(u)=(*(u^{\flat}\wedge S(u)))^{\sharp}. (3.26)

With respect to an orthonormal basis {u=e0,,e3}\{u=e_{0},\dots,e_{3}\} adapted to uu, the centre of spin condition takes the form

s0(u)=0,sa(u)=12εa0ρσSρσ(u)=12εabc(3)Sbc(u),s_{0}(u)=0,\quad s_{a}(u)=-\frac{1}{2}\varepsilon_{a0\rho\sigma}S^{\rho\sigma}(u)=\frac{1}{2}{{}^{(3)}\varepsilon}_{abc}S^{bc}(u), (3.27)

through which it acquires an immediate interpretation: a position observable is a centre of spin if and only if, for any Lorentz frame uu, the spin vector defined by boosting the Pauli–Lubański vector to uu really generates spatial rotations around the point given by the position observable.

We will now rewrite the centre of spin condition. Since S(u)=J(χ(u,τ))PS(u)=J-(\chi(u,\tau))^{\flat}\wedge P^{\flat}, we can rewrite the Pauli–Lubański vector as W=[(PmcJ)]=[(PmcS(u))]W=\left[*\left(\frac{P^{\flat}}{mc}\wedge J\right)\right]^{\sharp}=\left[*\left(\frac{P^{\flat}}{mc}\wedge S(u)\right)\right]^{\sharp}. Thus, the centre of spin condition takes the form

(B(u)1)[(PmcS(u))]=(uS(u)).(B(u)^{-1})^{\top}\left[*\left(\frac{P^{\flat}}{mc}\wedge S(u)\right)\right]=*(u^{\flat}\wedge S(u)). (3.28)

Since B(u)B(u) is a Lorentz transformation, i.e. an isometry of (V,η)(V,\eta), and it maps P/(mc)P/(mc) to uu, this is equivalent to

u((B(u)1)(B(u)1))(S(u))=uS(u).u^{\flat}\wedge\left((B(u)^{-1})^{\top}\otimes(B(u)^{-1})^{\top}\right)(S(u))=u^{\flat}\wedge S(u). (3.29)

Using the explicit form (3.24) of B(u)B(u), we see that

((B(u)1)(B(u)1))(S(u))=S(u)+Pmc(ιu+PmcS(u))1uPmc+u().\left((B(u)^{-1})^{\top}\otimes(B(u)^{-1})^{\top}\right)(S(u))=S(u)+\frac{\frac{P^{\flat}}{mc}\wedge\left(\iota_{u+\frac{P}{mc}}S(u)\right)}{1-u\cdot\frac{P}{mc}}+u^{\flat}\wedge(\ldots). (3.30)

Thus, we have the following:

Lemma 3.9.

The centre of spin condition is equivalent to

uP(ιu+PmcS(u))=0.u^{\flat}\wedge P^{\flat}\wedge\left(\iota_{u+\frac{P}{mc}}S(u)\right)=0.\qed (3.31)

Since the Newton–Wigner position observable is defined by the SSC ιu+PmcS(u)=0\iota_{u+\frac{P}{mc}}S(u)=0, the preceding result immediately implies

Theorem 3.10.

The Newton–Wigner position observable χNW\chi^{\mathrm{NW}} is a centre of spin. ∎

Further rewriting the centre of spin condition, we see that (3.31) is equivalent to

ιu+PmcS(u)span{u,P}.\iota_{u+\frac{P}{mc}}S(u)\in\mathrm{span}\{u^{\flat},P^{\flat}\}. (3.32)

Due to the antisymmetry of S(u)S(u), this is equivalent to

ιu+PmcS(u)span{uPmc}.\iota_{u+\frac{P}{mc}}S(u)\in\mathrm{span}\left\{u^{\flat}-\frac{P^{\flat}}{mc}\right\}. (3.33)

Using this, we can show:

Lemma 3.11.

χ\chi is a centre of spin \iff χ(u,τ)χNW(u,τ)span{u,P}\chi(u,\tau)-\chi^{\mathrm{NW}}(u,\tau)\in\mathrm{span}\{u,P\}.

Proof.

Writing D:=χ(u,τ)χNW(u,τ)D:=\chi(u,\tau)-\chi^{\mathrm{NW}}(u,\tau), the spin tensor of χ\chi may be expressed as S(u)=SNW(u)DPS(u)=S^{\mathrm{NW}}(u)-D^{\flat}\wedge P^{\flat}. Thus, (3.33) is equivalent to

ιu+Pmc(DP)span{uPmc}.\iota_{u+\frac{P}{mc}}(D^{\flat}\wedge P^{\flat})\in\mathrm{span}\left\{u^{\flat}-\frac{P^{\flat}}{mc}\right\}. (3.34)

We have ιu+Pmc(DP)=(Du+DPmc)P(Pumc)D\iota_{u+\frac{P}{mc}}(D^{\flat}\wedge P^{\flat})=(D\cdot u+\frac{D\cdot P}{mc})P^{\flat}-(P\cdot u-mc)D^{\flat}, and thus (3.34) implies that for all vuPv\in u^{\perp}\cap P^{\perp}, we have

vD=0.v\cdot D=0. (3.35)

But this means D(uP)=span{u,P}D\in(u^{\perp}\cap P^{\perp})^{\perp}=\mathrm{span}\{u,P\}.

Conversely, if Dspan{u,P}D\in\mathrm{span}\{u,P\}, we have ιu+Pmc(DP)span{ιu+Pmc(uP)}\iota_{u+\frac{P}{mc}}(D^{\flat}\wedge P^{\flat})\in\mathrm{span}\left\{\iota_{u+\frac{P}{mc}}(u^{\flat}\wedge P^{\flat})\right\}. But now

ιu+Pmc(uP)=(1+uPmc)P(uPmc)u=(mcuP)(uPmc),\iota_{u+\frac{P}{mc}}(u^{\flat}\wedge P^{\flat})=\left(-1+u\cdot\frac{P}{mc}\right)P^{\flat}-(u\cdot P-mc)u^{\flat}=(mc-u\cdot P)\left(u^{\flat}-\frac{P^{\flat}}{mc}\right), (3.36)

and thus we have (3.34), i.e. χ\chi is a centre of spin. ∎

We can now prove the main result of this section.

Theorem 3.12.

The Newton–Wigner position observable χNW\chi^{\mathrm{NW}} is the only centre of spin position observable that is continuous and defined by an SSC.

Proof.

Let χ\chi be an SSC position observable. Writing D(u,τ):=χ(u,τ)χNW(u,τ)D(u,\tau):=\chi(u,\tau)-\chi^{\mathrm{NW}}(u,\tau), we know by the Møller disc theorem (theorem 3.6) that the projection of D(u,τ)D(u,\tau) orthogonal to PP is orthogonal to the Pauli–Lubański vector WW. Thus, since PP itself is orthogonal to WW, we have

D(u,τ)WD(u,\tau)\perp W (3.37)

for any (u,τ)𝖲𝗉𝖧𝖯(u,\tau)\in\mathsf{SpHP}. In addition, we know that D(u,τ)uD(u,\tau)\perp u; in particular, D(u,τ)D(u,\tau) is spacelike for any (u,τ)𝖲𝗉𝖧𝖯(u,\tau)\in\mathsf{SpHP}.

Now suppose that χ\chi is a centre of spin. By lemma 3.11 this means that

D(u,τ)span{u,P}D(u,\tau)\in\mathrm{span}\{u,P\} (3.38)

for all (u,τ)𝖲𝗉𝖧𝖯(u,\tau)\in\mathsf{SpHP}. Using (3.37) and PWP\perp W, we conclude that

for alluwithuW0:D(u,τ)span{P}.\text{for all}\;u\;\text{with}\;u\cdot W\neq 0:D(u,\tau)\in\mathrm{span}\{P\}. (3.39)

Since D(u,τ)D(u,\tau) has to be spacelike, we thus have shown

D(u,τ)=0for alluwithuW0.D(u,\tau)=0\;\text{for all}\;u\;\text{with}\;u\cdot W\neq 0. (3.40)

If W0W\neq 0, the set of future-directed unit timelike uu satisfying uW0u\cdot W\neq 0 is dense in the hyperboloid of all possible uu, and thus assuming continuity of χ\chi, we conclude that D(u,τ)=0D(u,\tau)=0 for all uu, finishing the proof.

If W=0W=0, then by the Møller disc theorem all SSC worldlines coincide, and thus we also have χ=χNW\chi=\chi^{\mathrm{NW}}. ∎

Looking back into the various steps of the proofs it is interesting to note how the ‘extrinsic–intrinsic’ combination u+P/(mc)u+P/(mc) for ff came about. It entered through the unique boost transformation (3.24) that was needed in order to transform an intrinsic quantity to an externally specified rest frame. The intrinsic quantity is the spin vector in the momentum rest frame, i.e. the Pauli–Lubański vector, which is a function of Γ\Gamma only, and the externally specified frame is defined by uu, which is independent of Γ\Gamma and determined through the choice of Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP}.

4 A Newton–Wigner theorem for classical elementary systems

For elementary Poincaré-invariant quantum systems – i.e. quantum systems with an irreducible unitary action of the Poincaré group – the Newton–Wigner position operator is uniquely characterised by transforming ‘as a position should’ under translations, rotations and time reversal, having commuting components and satisfying a regularity condition. This has been well-known since the original publication by Newton and Wigner [4]. As advertised in the introduction, we shall now prove an analogous statement for classical systems.

For the whole of this section, we fix a future-directed unit timelike vector uu defining a Lorentz frame, and an adapted positively oriented orthonormal basis {u=e0,,e3}\{u=e_{0},\dots,e_{3}\}. Unless otherwise stated, phrases such as ‘temporal’, ‘spatial’ and the like refer to the preferred time direction given by uu. We will raise and lower spatial indices by the Euclidean metric δ\delta induced by the Minkowski metric η\eta on the orthogonal complement of uu; the components of δ\delta in the adapted basis are simply given by the usual Kronecker delta. We denote the spatial volume form by ε(3)=ιuε{}^{(3)}\varepsilon=\iota_{u}\varepsilon.

We will employ a ‘three-vector’ notation for spatial vectors, for example writing A=(Aa)\vec{A}=(A^{a}). We then use the usual three-vector notations for the Euclidean scalar product AB=AaBa\vec{A}\cdot\vec{B}=A_{a}B^{a}, the Euclidean norm |A|:=A2|\vec{A}|:=\sqrt{\vec{A}^{2}} and the vector product (A×B)a=εabc(3)AbBc(\vec{A}\times\vec{B})_{a}={{}^{(3)}\varepsilon_{abc}}A^{b}B^{c}.

4.1 Classical elementary systems

In the quantum case, an elementary system is given by a Hilbert space with an irreducible unitary action of the Poincaré group – i.e. each state of the system is connected to any other by a Poincaré transformation. In direct analogy, we define the notion of a classical elementary system:

Definition 4.1.

A classical elementary system is a classical Poincaré-invariant system (Γ,ω,Φ)(\Gamma,\omega,\Phi), where Φ\Phi is a transitive action of the proper orthochronous Poincaré group 𝒫+\mathcal{P}_{+}^{\uparrow}.

Note the we only assumed an action of the identity connected component of the Poincaré group, whereas Arens in [21] considered the whole Poincaré group. In the classical context, simple transitivity replaces irreducibility in the quantum case.

Arens classified the classical elementary systems121212In fact, Arens classified what he called one-particle elementary systems (systems that admit a map from Γ\Gamma to the set of lines in Minkowski space which is equivariant with respect to a certain subgroup of 𝒫+\mathcal{P}_{+}^{\uparrow}). However, he also proved that this ‘one-particle’ condition is fulfilled for an elementary system if and only if the four-momentum is not zero. in [21]; the classification proceeds in terms of the system’s four-momentum and Pauli–Lubański vector (similar to the Wigner classification in the quantum case [39]). We are only interested in the case of timelike four-momentum. For this case, the phase space can be explicitly constructed as follows:

Theorem 4.2 (Phase space of a classical elementary system).

Any classical elementary system with timelike four-momentum is equivalent (in the sense of a symplectic isomorphism respecting the action of 𝒫+\mathcal{P}_{+}^{\uparrow}) to precisely one of the following two cases:

  1. (i)

    (Spin zero, one parameter m+m\in\mathbb{R}_{+})

    • Phase space Γ=T3\Gamma=T^{*}\mathbb{R}^{3} with coordinates (x,p)(\vec{x},\vec{p}), symplectic form ω=dxadpa\omega=\mathrm{d}x^{a}\wedge\mathrm{d}p_{a}

    • Poincaré generators (i.e. component functions of the momentum map):

      spatial translationsPa\displaystyle\text{spatial translations}\quad P_{a} =pa\displaystyle=p_{a} (4.1a)
      time translationP0\displaystyle\text{time translation}\quad P_{0} =m2c2+p2\displaystyle=-\sqrt{m^{2}c^{2}+\vec{p}^{2}} (4.1b)
      rotationsJab\displaystyle\text{rotations}\quad J_{ab} =xapbxbpa\displaystyle=x_{a}p_{b}-x_{b}p_{a} (4.1c)
      boostsJa0\displaystyle\text{boosts}\quad J_{a0} =P0xa\displaystyle=P_{0}x_{a} (4.1d)
  2. (ii)

    (Spin non-zero, two parameters m,S+m,S\in\mathbb{R}_{+})

    • Phase space Γ=T3×𝖲2\Gamma=T^{*}\mathbb{R}^{3}\times\mathsf{S}^{2} with coordinates (x,p)(\vec{x},\vec{p}) for T3T^{*}\mathbb{R}^{3}, symplectic form ω=dxadpa+SdΩ2\omega=\mathrm{d}x^{a}\wedge\mathrm{d}p_{a}+S\cdot\mathrm{d}\Omega^{2} where dΩ2\mathrm{d}\Omega^{2} is the standard volume form on 𝖲2\mathsf{S}^{2}. We denote the phase space function projecting onto the second factor 𝖲2\mathsf{S}^{2} by s^:Γ𝖲23\hat{s}\colon\Gamma\to\mathsf{S}^{2}\subset\mathbb{R}^{3}. The spin vector observable is the 𝖲S2\mathsf{S}^{2}_{S}-valued phase space function s:=Ss^\vec{s}:=S\cdot\hat{s}; its components satisfy the Poisson bracket relations

      {sa,sb}=εabc(3)sc.\{s_{a},s_{b}\}={{}^{(3)}\varepsilon_{abc}}s^{c}. (4.2)

      Here 𝖲S23\mathsf{S}^{2}_{S}\subset\mathbb{R}^{3} denotes the 2-sphere of radius SS in 3\mathbb{R}^{3}.

    • Poincaré generators (i.e. component functions of the momentum map):

      spatial translationsPa\displaystyle\text{spatial translations}\quad P_{a} =pa\displaystyle=p_{a} (4.3a)
      time translationP0\displaystyle\text{time translation}\quad P_{0} =m2c2+p2\displaystyle=-\sqrt{m^{2}c^{2}+\vec{p}^{2}} (4.3b)
      rotationsJab\displaystyle\text{rotations}\quad J_{ab} =xapbxbpa+εabc(3)sc\displaystyle=x_{a}p_{b}-x_{b}p_{a}+{{}^{(3)}\varepsilon_{abc}}s^{c} (4.3c)
      boostsJa0\displaystyle\text{boosts}\quad J_{a0} =P0xa(p×s)amcP0\displaystyle=P_{0}x_{a}-\frac{(\vec{p}\times\vec{s})_{a}}{mc-P_{0}} (4.3d)

Note that in fact the explicit construction of the systems in [21] as co-adjoint orbits of 𝒫+\mathcal{P}_{+}^{\uparrow} is quite different in appearance to the forms given above. However, one can show that the above systems are indeed elementary systems (i.e. that the action of 𝒫+\mathcal{P}_{+}^{\uparrow} is transitive), and thus due to Arens’ uniqueness result they are possible representatives of their respective classes. We will use the forms given above, which were anticipated by Bacry in [40], since they will be easier to explicitly work with. To unify notation, we let S=0,s:=0S=0,\vec{s}:=0 in the case of zero-spin systems. Furthermore, we introduce the open subset of phase space Γ:=Γ{|P|=0}\Gamma^{*}:=\Gamma\setminus\{|\vec{P}|=0\} and the 𝖲2\mathsf{S}^{2}-valued function P^:=P|P|\hat{P}:=\frac{\vec{P}}{|\vec{P}|} on Γ\Gamma^{*}.

Using the explicit form of the systems given in theorem 4.2, one directly checks:

Lemma 4.3.

For a classical elementary system with timelike four-momentum, the functions Pa,P^sP_{a},\hat{P}\cdot\vec{s} (or just the PaP_{a} in the case of zero spin) form a complete involutive set on Γ\Gamma^{*} (or the whole of Γ\Gamma in the case of zero spin). ∎

The behaviour of the momentum and spin vectors under translations and rotations is also easily obtained:

Lemma 4.4.

For a classical elementary system with timelike four-momentum, P\vec{P} and s\vec{s} are invariant under translations and ‘transform as vectors’ under spatial rotations, i.e. we have

{Pa,Vb}=0,{Jab,Vc}=δacVbδbcVaforV=P,s.\{P_{a},V_{b}\}=0,\quad\{J_{ab},V_{c}\}=\delta_{ac}V_{b}-\delta_{bc}V_{a}\quad\text{for}\quad\vec{V}=\vec{P},\vec{s}. (4.4)
Proof.

For P\vec{P}, these are part of the Poincaré algebra relations and thus true by definition. For s\vec{s}, they are easily confirmed using the explicit form of the Poincaré generators. ∎

For our considerations, we will need to know how the time reversal operation with respect to the hyperplane in MM through the origin oMo\in M and orthogonal to u=e0u=e_{0} is implemented on phase space. In order to get this right, we recall that the incorporation of time reversal in the context of Special Relativity corresponds, by its very definition, to a particular upward 2\mathbb{Z}_{2} extension131313Here we are using the terminology of [41, p. xx], according to which a group GG with normal subgroup AA and quotient G/ABG/A\cong B is either called an upward extension of AA by BB or a downward extension of BB by AA. of 𝒫+\mathcal{P}_{+}^{\uparrow}, i.e. the formation of a new group called 𝒫+𝒫\mathcal{P}_{+}^{\uparrow}\cup\mathcal{P}_{-}^{\downarrow} of which 𝒫+\mathcal{P}_{+}^{\uparrow} is a normal subgroup with (𝒫+𝒫)/𝒫+2(\mathcal{P}_{+}^{\uparrow}\cup\mathcal{P}_{-}^{\downarrow})/\mathcal{P}_{+}^{\uparrow}\cong\mathbb{Z}_{2}. It is the particular nature of this extension that eventually defines what is meant by time reversal: it consists in the requirement that the outer automorphism induced by the only non-trivial element of 2\mathbb{Z}_{2} on the Lie algebra 𝔭\mathfrak{p} of 𝒫+\mathcal{P}_{+}^{\uparrow} shall be the one which reverses the sign of spatial translations and rotations and leaves invariant boosts and time translations; see, e.g., [42]. Implementing time reversal on phase space then means to extend the action of 𝒫+\mathcal{P}_{+}^{\uparrow} to an action of 𝒫+𝒫\mathcal{P}_{+}^{\uparrow}\cup\mathcal{P}_{-}^{\downarrow}.

Now, according to this scheme, we can immediately write down how our particular time reversal transformation on phase space, Tu:ΓΓT_{u}\colon\Gamma\to\Gamma, acts on the Poincaré generators, i.e. the component functions of the momentum map:

PaTu=Pa,JabTu=Jab,Ja0Tu=Ja0,P0Tu=P0P_{a}\circ T_{u}=-P_{a}\;,\quad J_{ab}\circ T_{u}=-J_{ab}\;,\quad J_{a0}\circ T_{u}=J_{a0}\;,\quad P_{0}\circ T_{u}=P_{0} (4.5)

From this the well-known result follows that time reversal (as defined above) necessarily corresponds to an anti-symplectomorphism. Hence, in the process of extending our symplectic action of 𝒫+\mathcal{P}_{+}^{\uparrow} on Γ\Gamma to an action of 𝒫+𝒫\mathcal{P}_{+}^{\uparrow}\cup\mathcal{P}_{-}^{\downarrow} satisfying the time reversal criterion above, we had to generalise to possibly anti-symplectomorphic actions. This is akin to the situation in Quantum Mechanics, where, as is well-known, time reversal necessarily corresponds to an anti-unitary transformation.

It is now clear how time reversal is implemented in the case at hand:

Lemma 4.5.

For an elementary system as in theorem 4.2, time reversal with respect to the hyperplane through the origin and orthogonal to u=e0u=e_{0} is given by

Tu:(x,p,s^)(x,p,s^).T_{u}\colon(\vec{x},\vec{p},\hat{s})\mapsto(\vec{x},-\vec{p},-\hat{s}).\qed (4.6)

Unless otherwise stated, in the following we will always mean time reversal with respect to the hyperplane through the origin and orthogonal to u=e0u=e_{0} when saying ‘time reversal’.

4.2 Statement and interpretation of the Newton–Wigner theorem

The classical Newton–Wigner theorem we are going to prove can be formulated very similar to the quantum case:

Theorem 4.6 (Classical Newton–Wigner theorem).

For a classical elementary system with timelike four-momentum, there is a unique 3\mathbb{R}^{3}-valued phase space function X\vec{X} that

  1. (i)

    is C1C^{1},

  2. (ii)

    has Poisson-commuting components,

  3. (iii)

    satisfies the canonical Poisson relations {Xa,Pb}=δba\{X^{a},P_{b}\}=\delta^{a}_{b} with the generators of spatial translations with respect to u=e0u=e_{0},

  4. (iv)

    transforms ‘as a (position) vector’ under spatial rotations with respect to u=e0u=e_{0}, i.e. satisfies {Jab,Xc}=δacXbδbcXa\{J_{ab},X^{c}\}=\delta_{a}^{c}X_{b}-\delta_{b}^{c}X_{a}, and

  5. (v)

    is invariant under time reversal with respect to the hyperplane through the origin and orthogonal to u=e0u=e_{0}, i.e. satisfies XTu=X\vec{X}\circ T_{u}=\vec{X}.

In terms of the Poincaré generators, it is given by

Xa=Ja0mcJabPbmc(mcP0)Jb0PbP0mc(mcP0)Pa,X_{a}=-\frac{J_{a0}}{mc}-\frac{J_{ab}P^{b}}{mc(mc-P_{0})}-\frac{J_{b0}P^{b}}{P_{0}mc(mc-P_{0})}P_{a}\;, (4.7)

where m=P02P2/cm=\sqrt{P_{0}^{2}-\vec{P}^{2}}/c is the mass of the system.

Before proving the theorem in the next section, we will now discuss the interpretation of the ‘position’ X\vec{X} it characterises. We want to interpret the value of X\vec{X} (in some state γΓ\gamma\in\Gamma) as the spatial components of a point in Minkowski spacetime MM. Since X\vec{X} is invariant under time reversal with respect to the hyperplane through the origin and orthogonal to u=e0u=e_{0}, it can be interpreted as defining a point on this hyperplane. Thus, if we want to use the phase space function from the Newton–Wigner theorem to define a position observable χ\chi in the sense of section 3.1, we should set (in our basis adapted to uu)

χa(u,τ=0):=Xa,χ0(u,τ=0):=0.\chi^{a}(u,\tau=0):=X^{a}\;,\quad\chi^{0}(u,\tau=0):=0. (4.8)

The transformation behaviour of X\vec{X} under spatial translations and rotations (i.e. assumptions (iii) and (iv) of theorem 4.6) will then ensure that the position observable χ\chi be covariant (in the sense of definition 3.2) regarding these transformations.

In fact, comparing (4.7) to the expression (3.17) for the Newton–Wigner position observable χNW\chi^{\mathrm{NW}}, we see that we have (in our adapted basis)

χNW,a(u,τ=0)=Xa,χNW,0(u,τ=0)=0:\chi^{\mathrm{NW},a}(u,\tau=0)=X^{a}\;,\quad\chi^{\mathrm{NW},0}(u,\tau=0)=0\colon (4.9)

the position X\vec{X} characterised by theorem 4.6 is the one given by the Newton–Wigner position observable χNW\chi^{\mathrm{NW}} on the hyperplane (u,0)𝖲𝗉𝖧𝖯(u,0)\in\mathsf{SpHP} (which is a covariant position observable due to proposition 3.5). Let us also remark that since any position observable’s dependence on τ\tau is fixed by (3.8), a position observable satisfying (4.8) is equal to the Newton–Wigner observable χNW\chi^{\mathrm{NW}} on the whole family of hyperplanes Σ𝖲𝗉𝖧𝖯\Sigma\in\mathsf{SpHP} with normal vector uu.

Combining this identification with the observation that we can freely choose the origin oMo\in M, we can restate the Newton–Wigner theorem in the following form:

Theorem 4.7 (Classical Newton–Wigner theorem, version 2).

For a classical elementary system with timelike four-momentum, given any hyperplane Σ=(u,τ)𝖲𝗉𝖧𝖯\Sigma=(u,\tau)\in\mathsf{SpHP}, there is a unique Σ\Sigma-valued phase space function χNW(Σ)\chi^{\mathrm{NW}}(\Sigma) that

  1. (i)

    is C1C^{1},

  2. (ii)

    has Poisson-commuting components, i.e.

    {χNW,μ(Σ),χNW,ν(Σ)}=0,\left\{\chi^{\mathrm{NW},\mu}(\Sigma),\chi^{\mathrm{NW},\nu}(\Sigma)\right\}=0, (4.10a)
  3. (iii)

    satisfies the canonical Poisson relations with the generators of spatial translations with respect to uu, i.e.

    vμwν{χNW,μ(Σ),Pν}=vwforv,wu,v_{\mu}w^{\nu}\left\{\chi^{\mathrm{NW},\mu}(\Sigma),P_{\nu}\right\}=v\cdot w\;\text{for}\;v,w\in u^{\perp}, (4.10b)
  4. (iv)

    transforms ‘as a position’ under spatial rotations with respect to uu, i.e. satisfies

    vμv~νwρ{Jμν,χNW,ρ(Σ)}=vμv~νwρ[δμρχνNW(Σ)δνρχμNW(Σ)]forv,v~,wu,v^{\mu}\tilde{v}^{\nu}w_{\rho}\left\{J_{\mu\nu},\chi^{\mathrm{NW},\rho}(\Sigma)\right\}=v^{\mu}\tilde{v}^{\nu}w_{\rho}\left[\delta_{\mu}^{\rho}\chi^{\mathrm{NW}}_{\nu}(\Sigma)-\delta_{\nu}^{\rho}\chi^{\mathrm{NW}}_{\mu}(\Sigma)\right]\;\text{for}\;v,\tilde{v},w\in u^{\perp}, (4.10c)

    and

  5. (v)

    is invariant under time reversal with respect to Σ\Sigma.

These χNW(Σ)\chi^{\mathrm{NW}}(\Sigma) together form the Newton–Wigner observable as given by (3.17). ∎

4.3 Proof of the Newton–Wigner theorem

Proof of theorem 4.6.

For the whole of the proof, we will work with the explicit form of the phase space of our elementary system given in theorem 4.2. It is easily verified that in this explicit form, x\vec{x} (i.e. the coordinate of the base point in T3T^{*}\mathbb{R}^{3}) is a phase space function with the properties demanded for X\vec{X}. Thus we need to prove uniqueness. Our proof will follow the proof of the quantum-mechanical Newton–Wigner theorem given by Jordan in [43], some parts of which can be applied literally to the classical case.

We will several times need the following.

Lemma 4.8.

Consider a classical elementary system with timelike four-momentum, with phase space Γ\Gamma, and some open subset Γ~\tilde{\Gamma} of Γ=Γ{|P|=0}\Gamma^{*}=\Gamma\setminus\{|\vec{P}|=0\}. Let ff be an \mathbb{R}-valued C1C^{1} function defined on Γ~\tilde{\Gamma} that is invariant under spatial translations and rotations, i.e. {Pa,f}=0={Jab,f}\{P_{a},f\}=0=\{J_{ab},f\}. Then ff is a function of |P|,P^s|\vec{P}|,\hat{P}\cdot\vec{s}. 141414By ‘ff is a function of |P|,P^s|\vec{P}|,\hat{P}\cdot\vec{s} ’ we mean that ff depends on phase space only via |P|,P^s|\vec{P}|,\hat{P}\cdot\vec{s}, i.e. that there is a C1C^{1} function F:UF\colon U\to\mathbb{R}, U={(|P|(γ),(P^s)(γ)):γΓ~}+×[S,S]U=\left\{(|\vec{P}|(\gamma),(\hat{P}\cdot\vec{s})(\gamma)):\gamma\in\tilde{\Gamma}\right\}\subset\mathbb{R}_{+}\times[-S,S] satisfying f(γ)=F(|P|(γ),(P^s)(γ))for allγΓ~.f(\gamma)=F(|\vec{P}|(\gamma),(\hat{P}\cdot\vec{s})(\gamma))\;\text{for all}\;\gamma\in\tilde{\Gamma}.

Proof.

ff Poisson-commutes with P\vec{P} and JabJ_{ab}. Therefore it also Poisson-commutes with P\vec{P} and 12εabc(3)P^aJbc=P^s\frac{1}{2}{{}^{(3)}\varepsilon^{abc}}\hat{P}_{a}J_{bc}=\hat{P}\cdot\vec{s}. Now P,P^s\vec{P},\hat{P}\cdot\vec{s} form a complete involutive set on Γ\Gamma^{*} (lemma 4.3), so since ff Poisson-commutes with them, it must be a function of P,P^s\vec{P},\hat{P}\cdot\vec{s}. Since ff and P^s\hat{P}\cdot\vec{s} are rotation invariant (by lemma 4.4), ff must be a function of |P|,P^s|\vec{P}|,\hat{P}\cdot\vec{s}. ∎

Let now X\vec{X} be an observable as in the statement of theorem 4.6, and consider the difference d:=Xx\vec{d}:=\vec{X}-\vec{x}. Due to the assumptions of theorem 4.6, d\vec{d} is C1C^{1}, is invariant under translations (i.e. {da,Pb}=0\{d^{a},P_{b}\}=0), transforms as a vector under spatial rotations (i.e. {Jab,dc}=δacdbδbcda\{J_{ab},d^{c}\}=\delta_{a}^{c}d_{b}-\delta_{b}^{c}d_{a}) and is invariant under time reversal with respect to the hyperplane through the origin and orthogonal to uu (i.e. dTu=d\vec{d}\circ T_{u}=\vec{d}).

Lemma 4.9.

Let A\vec{A} be a 3\mathbb{R}^{3}-valued C1C^{1} phase space function on a classical elementary system with timelike four-momentum that is invariant under translations, transforms as a vector under spatial rotations and is invariant under time reversal. Then AP=0\vec{A}\cdot\vec{P}=0.

Proof.

Since P\vec{P} is invariant under translations and a vector under rotations, AP\vec{A}\cdot\vec{P} is invariant under translations and rotations. By lemma 4.8, AP|Γ\left.\vec{A}\cdot\vec{P}\right|_{\Gamma^{*}} is a function of |P|,P^s|\vec{P}|,\hat{P}\cdot\vec{s}. This means we have

AP|Γ=F(|P|,P^s)\left.\vec{A}\cdot\vec{P}\right|_{\Gamma^{*}}=F(|\vec{P}|,\hat{P}\cdot\vec{s}) (4.11)

for some function F:+×[S,S]F\colon\mathbb{R}_{+}\times[-S,S]\to\mathbb{R}.

Now considering time reversal TuT_{u}, on the one hand we have (using lemma 4.5)

|P|Tu=|PTu|=|P|=|P||\vec{P}|\circ T_{u}=|\vec{P}\circ T_{u}|=|-\vec{P}|=|\vec{P}| (4.12a)
and
(P^s)Tu\displaystyle(\hat{P}\cdot\vec{s})\circ T_{u} =(12εabc(3)P^aJbc)Tu\displaystyle=\left(\frac{1}{2}{{}^{(3)}\varepsilon^{abc}}\hat{P}_{a}J_{bc}\right)\circ T_{u}
=12εabc(3)(P^aTu)(JbcTu)\displaystyle=\frac{1}{2}{{}^{(3)}\varepsilon^{abc}}(\hat{P}_{a}\circ T_{u})(J_{bc}\circ T_{u})
=12εabc(3)(P^a)(Jbc)\displaystyle=\frac{1}{2}{{}^{(3)}\varepsilon^{abc}}(-\hat{P}_{a})(-J_{bc})
=12εabc(3)P^aJbc\displaystyle=\frac{1}{2}{{}^{(3)}\varepsilon^{abc}}\hat{P}_{a}J_{bc}
=P^s,\displaystyle=\hat{P}\cdot\vec{s}, (4.12b)

implying

F(|P|,P^s)Tu=F(|P|Tu,(P^s)Tu)=F(|P|,P^s).F(|\vec{P}|,\hat{P}\cdot\vec{s})\circ T_{u}=F(|\vec{P}|\circ T_{u},(\hat{P}\cdot\vec{s})\circ T_{u})=F(|\vec{P}|,\hat{P}\cdot\vec{s}). (4.13)

On the other hand, A\vec{A} is invariant under time reversal while P\vec{P} changes its sign, implying that (AP)Tu=AP(\vec{A}\cdot\vec{P})\circ T_{u}=-\vec{A}\cdot\vec{P}. Combining this with (4.11) and (4.13), we obtain AP|Γ=0\left.\vec{A}\cdot\vec{P}\right|_{\Gamma^{*}}=0, and continuity implies AP=0\vec{A}\cdot\vec{P}=0. ∎

For zero spin, we can easily complete the proof of the Newton–Wigner theorem. Since the difference vector d\vec{d} is translation invariant and the PaP_{a} form a complete involutive set on Γ\Gamma, d\vec{d} must be a function of P\vec{P}. Then since it is a vector under rotations, it must be of the form

d(P)=F(|P|)P\vec{d}(\vec{P})=F(|\vec{P}|)\vec{P} (4.14)

for some function FF of |P||\vec{P}|. Then, since according to lemma 4.9 dP\vec{d}\cdot\vec{P} is zero, d\vec{d} is zero. Thus, for the spin-zero case, we have proved the Newton–Wigner theorem without any use of the condition of Poisson-commuting components of the position observable.

For the non-zero spin case, we continue as follows.

Lemma 4.10.

Let A\vec{A} be a 3\mathbb{R}^{3}-valued C1C^{1} phase space function on a classical elementary system with timelike four-momentum and non-zero spin that is invariant under translations, transforms as a vector under spatial rotations and satisfies AP=0\vec{A}\cdot\vec{P}=0. Then it is of the form

A=BP^×s+CP^×(P^×s)\vec{A}=B\hat{P}\times\vec{s}+C\hat{P}\times(\hat{P}\times\vec{s}) (4.15)

on Γ{sP^}\Gamma^{*}\setminus\{\vec{s}\parallel\hat{P}\}, where BB and CC are C1C^{1} functions of |P||\vec{P}| and P^s\hat{P}\cdot\vec{s}, i.e. C1C^{1} functions

B,C:+×(S,S).B,C\colon\mathbb{R}_{+}\times(-S,S)\to\mathbb{R}.
Proof.

For the whole of this proof, we will work on Γ~:=Γ{sP^}\tilde{\Gamma}:=\Gamma^{*}\setminus\{\vec{s}\parallel\hat{P}\}. Since evaluated at each point of Γ~\tilde{\Gamma}, the 3\mathbb{R}^{3}-valued functions P^,P^×s,P^×(P^×s)\hat{P},\hat{P}\times\vec{s},\hat{P}\times(\hat{P}\times\vec{s}) form an orthogonal basis of 3\mathbb{R}^{3}, and since we have AP=0\vec{A}\cdot\vec{P}=0, we can write A\vec{A} in the form (4.15) with coefficients B,CB,C given by

B\displaystyle B =A(P^×s)|P^×s|,\displaystyle=\frac{\vec{A}\cdot(\hat{P}\times\vec{s})}{|\hat{P}\times\vec{s}|}\;, (4.16)
C\displaystyle C =A(P^×(P^×s))|P^×(P^×s)|.\displaystyle=\frac{\vec{A}\cdot(\hat{P}\times(\hat{P}\times\vec{s}))}{|\hat{P}\times(\hat{P}\times\vec{s})|}\;. (4.17)

Since A\vec{A}, P\vec{P} and s\vec{s} are invariant under translations and vectors under rotations, these equations imply that B,CB,C are invariant under translations and rotations. The result follows with lemma 4.8. ∎

Now we consider again the difference vector d=Xx\vec{d}=\vec{X}-\vec{x}. It satisfies dP=0\vec{d}\cdot\vec{P}=0 by lemma 4.9, and thus we have

XP=xP.\vec{X}\cdot\vec{P}=\vec{x}\cdot\vec{P}. (4.18)

Since we assume that the components of X\vec{X} Poisson-commute with each other and that {Xa,Pb}=δba\{X^{a},P_{b}\}=\delta^{a}_{b}, this implies

{Xa,xP}={Xa,XP}=Xa.\{X^{a},\vec{x}\cdot\vec{P}\}=\{X^{a},\vec{X}\cdot\vec{P}\}=X^{a}. (4.19)

Combining this with {xa,xP}=xa\{x^{a},\vec{x}\cdot\vec{P}\}=x^{a}, we obtain

{da,xP}=da.\{d^{a},\vec{x}\cdot\vec{P}\}=d^{a}. (4.20)

On the other hand, for any function FF of P\vec{P} and s\vec{s}, we have

{F(P,s),xP}={F(P,s),xa}Pa=F(P,s)PaPa=|P|F|P||P^=const.,s=const..\{F(\vec{P},\vec{s}),\vec{x}\cdot\vec{P}\}=\{F(\vec{P},\vec{s}),x^{a}\}P_{a}=-\frac{\partial F(\vec{P},\vec{s})}{\partial P_{a}}P_{a}=-|\vec{P}|\left.\frac{\partial F}{\partial|\vec{P}|}\right|_{\hat{P}=\mathrm{const.},\vec{s}=\mathrm{const.}}. (4.21)

This implies

d=|P|d|P||P^=const.,s=const..\vec{d}=-|\vec{P}|\left.\frac{\partial\vec{d}}{\partial|\vec{P}|}\right|_{\hat{P}=\mathrm{const.},\vec{s}=\mathrm{const.}}. (4.22)

Combining lemmas 4.9 and 4.10, we know that d\vec{d} has the form (4.15) on Γ{sP^}\Gamma^{*}\setminus\{\vec{s}\parallel\hat{P}\} for two functions B,C:+×(S,S)B,C\colon\mathbb{R}_{+}\times(-S,S)\to\mathbb{R}. Thus (4.22) implies the two equations

B(|P|,P^s)=|P|B(|P|,P^s)|P|,C(|P|,P^s)=|P|C(|P|,P^s)|P|B(|\vec{P}|,\hat{P}\cdot\vec{s})=-|\vec{P}|\frac{\partial B(|\vec{P}|,\hat{P}\cdot\vec{s})}{\partial|\vec{P}|},\;C(|\vec{P}|,\hat{P}\cdot\vec{s})=-|\vec{P}|\frac{\partial C(|\vec{P}|,\hat{P}\cdot\vec{s})}{\partial|\vec{P}|} (4.23)

on +×(S,S)\mathbb{R}_{+}\times(-S,S). These equations determine the |P||\vec{P}| dependence of BB and CC; they must be proportional to |P|1|\vec{P}|^{-1}. However, for d\vec{d} to be C1C^{1} on the whole of Γ\Gamma, in fact for (4.15) not to diverge as |P|0|\vec{P}|\to 0 even when coming from a single direction P^\hat{P}, we then need BB and CC to vanish. Continuity implies d=0\vec{d}=0 on all of Γ\Gamma. This finishes the proof of the Newton–Wigner theorem. ∎

5 Conclusion

In this paper we have studied the localisation problem for classical system whose phase space is a symplectic manifold. We focussed on the Newton–Wigner position observable and asked for precise characterisations of it in order to gain additional understanding, over and above that already known from its practical use for the solution of concrete problems of motion, e.g., in astrophysics [16, 17]. We proved two theorems that we believe advance our understanding in the desired direction: first we showed how Fleming’s geometric scheme [15] in combination with the characterisation of worldlines through SSCs (Spin Supplementary Conditions) allows to give a precise meaning to, and proof of, the fact that the Newton–Wigner position is the unique centre of spin. Given that interpretation, it also offers an insight as to why the Newton–Wigner SSC uses a somewhat unnatural looking ‘hybrid’ combination f=u+Pmcf=u+\frac{P}{mc}, where uu is ‘external’ or ‘kinematical’, and PP is ‘internal’ or ‘dynamical’. Then, restricting to elementary systems, i.e. systems whose phase space admits a transitive action of the proper orthochronous Poincaré group, we proved again a uniqueness result to the effect that the Newton–Wigner observable is the unique phase space function whose components satisfy the ‘familiar’ Poisson relations, provided it is continuously differentiable, time-reversal invariant, and transforms as a vector under spatial rotations. These properties seem to be the underlying reason for the distinguished rôle it plays in solution strategies like those of [16, 17]151515Despite the fact that on a more general level of theorisation other choices (characterised by other SSCs) are often considered more appropriate; see, e.g., [18]., which in recent years have lead to astounding progress in the Hamiltonian analytical understanding of the dynamics of binary systems of spinning compact objects: the calculations have been pushed to ever higher post-Newtonian orders, starting from the next-to-leading order for spin–orbit and spin–spin effects, i.e. order c4c^{-4}, in [44, 45, 46], and most recently reaching a complete description at ‘4.5th’ post-Newtonian order, i.e. order c9c^{-9}, in [47, 48]. We believe that our results add a conceptually clear and mathematically precise Hamiltonian underpinning of what the choice of the Newton–Wigner observable entails, at least in a special-relativistic context or, more generally, in general-relativistic perturbation theory around Minkowski space.

Acknowledgements

This work was supported by the Deutsche Forschungsgemeinschaft through the Collaborative Research Centre 1227 (DQ-mat), projects B08/A05.

References

Appendix A Sign conventions for generators of special orthogonal groups

Let VV be a finite-dimensional real vector space with a non-degenerate, symmetric bilinear form g:V×Vg\colon V\times V\to\mathbb{R}. Note that we do not assume anything about the signature of gg. We introduce the ‘musical isomorphism’

VV,vv:=g(v,)V\to V^{*},v\mapsto v^{\flat}:=g(v,\cdot) (A.1)

induced by gg.

We fix a basis {ea}a\{e_{a}\}_{a} of VV. As bases for its dual vector space VV^{*} we distinguish its natural dual basis {θa}a\{\theta^{a}\}_{a}, where θa(eb)=δba\theta^{a}(e_{b})=\delta^{a}_{b}, and the (gg-dependent) image of {ea}a\{e_{a}\}_{a} under (A.1), which is just {ea}a\{e^{\flat}_{a}\}_{a}, where ea=gabθbe^{\flat}_{a}=g_{ab}\theta^{b}, so that ea(eb)=gabe^{\flat}_{a}(e_{b})=g_{ab}. The reason for this will become clear now.

For each a,b{1,,dimV}a,b\in\{1,\dots,\dim V\} we introduce the endomorphism

Bab:=eaebebeaEnd(V),B_{ab}:=e_{a}\otimes e_{b}^{\flat}-e_{b}\otimes e_{a}^{\flat}\in\mathrm{End}(V), (A.2)

which satisfies

g(v,Bab(w))=g(v,ea)g(eb,w)g(v,eb)g(ea,w)=g(Bab(v),w).g(v,B_{ab}(w))=g(v,e_{a})g(e_{b},w)-g(v,e_{b})g(e_{a},w)=-g(B_{ab}(v),w). (A.3)

This means that BabB_{ab} is anti-self-adjoint with respect to gg and hence that it is an element of the Lie algebra 𝔰𝔬(V,g)\mathfrak{so}(V,g) of the Lie group 𝖲𝖮(V,g)\mathsf{SO}(V,g) of special orthogonal transformations of (V,g)(V,g):

Bab𝔰𝔬(V,g).B_{ab}\in\mathfrak{so}(V,g). (A.4)

As Bab=BbaB_{ab}=-B_{ba}, it is the set {Bab:1a<bdimV}\{B_{ab}:1\leq a<b\leq\dim V\} which is linearly independent and of the same dimension as 𝔰𝔬(V,g)\mathfrak{so}(V,g). Hence this set forms a basis of 𝔰𝔬(V,g)\mathfrak{so}(V,g) so that any ω𝔰𝔬(V,g)\omega\in\mathfrak{so}(V,g) can be uniquely written in the form

ω=1a<bdimVωabBab=12ωabBab,\omega=\sum_{1\leq a<b\leq\dim V}\omega^{ab}B_{ab}=\frac{1}{2}\omega^{ab}B_{ab}\;, (A.5)

where

ωab=ωba.\omega^{ab}=-\omega^{ba}\,. (A.6)

This representation can easily be compared to the usual one in terms of the metric-independent basis {eaθb:1a,bdimV}\{e_{a}\otimes\theta^{b}:1\leq a,b\leq\dim V\} of End(V)\mathrm{End}(V) in the following way: for ω=ωcaeaθc\omega=\omega^{a}_{\hphantom{a}c}\,e_{a}\otimes\theta^{c}, we have ω𝔰𝔬(V,g)\omega\in\mathfrak{so}(V,g) if and only if

ωcagcb=ωcbgca.\omega^{a}_{\hphantom{a}c}\,g^{cb}=-\omega^{b}_{\hphantom{a}c}\,g^{ca}\;. (A.7)

It is the obvious simplicity of (A.6) as opposed to (A.7) as conditions for ωEnd(V)\omega\in\mathrm{End}(V) being contained in 𝔰𝔬(V,g)End(V)\mathfrak{so}(V,g)\subset\mathrm{End}(V) that makes it easier to work with the basis eaebe_{a}\otimes e^{\flat}_{b} of End(V)\mathrm{End}(V) rather than eaθbe_{a}\otimes\theta^{b}. Note that the components of ω\omega with respect to the two bases considered above are connected by the equation

ωab=ωcagcb.\omega^{ab}=\omega^{a}_{\hphantom{a}c}\,g^{cb}\;. (A.8)

The basis elements BabB_{ab} satisfy the commutation relations

[Bab,Bcd]\displaystyle[B_{ab},B_{cd}] =gbcBad+gadBbcgacBbdgbdBac\displaystyle=g_{bc}B_{ad}+g_{ad}B_{bc}-g_{ac}B_{bd}-g_{bd}B_{ac}
=gbcBad+(antisymm.),\displaystyle=g_{bc}B_{ad}+\text{(antisymm.)}, (A.9)

where ‘antisymm.’ is as explained below equation (2.4).

From now on, we will assume the basis {ea}a\{e_{a}\}_{a} to be orthonormal. For notational convenience, for a,b{1,,dimV}a,b\in\{1,\dots,\dim V\} we define

εab:=gaagbb=±1\varepsilon_{ab}:=g_{aa}g_{bb}=\pm 1 (A.10)

which has the value +1+1 if gaa=g(ea,ea)g_{aa}=g(e_{a},e_{a}) and gbb=g(eb,eb)g_{bb}=g(e_{b},e_{b}) have the same sign, and 1-1 if they have opposite signs161616Note that repeated indices on the same level, i.e. both up or both down, are not to be summed over..

We now want to compute the exponential exp(αBab)𝖲𝖮(V,g)\exp(\alpha B_{ab})\in\mathsf{SO}(V,g). At first, we note that

(Bab)2\displaystyle(B_{ab})^{2} =gbbeaeagaaebeb\displaystyle=-g_{bb}e_{a}\otimes e_{a}^{\flat}-g_{aa}e_{b}\otimes e_{b}^{\flat}
=εabPrab,\displaystyle=-\varepsilon_{ab}\;\mathrm{Pr}_{ab}\,, (A.11)

where Prab:=Prspan{ea,eb}\mathrm{Pr}_{ab}:=\mathrm{Pr}_{\mathrm{span}\{e_{a},e_{b}\}} denotes the gg-orthogonal projector onto the plane span{ea,eb}\mathrm{span}\{e_{a},e_{b}\} in VV.171717In the general case of two linearly independent vectors v,wVv,w\in V, not necessarily orthonormal, the orthogonal projector is given by Prspan(v,w)=1g(v,v)g(w,w)(g(v,w))2[g(w,w)vv+g(v,v)wwg(v,w)(vw+wv)],\mathrm{Pr}_{\mathrm{span}(v,w)=\frac{1}{g(v,v)g(w,w)-(g(v,w))^{2}}}\left[g(w,w)\;v\otimes v^{\flat}+g(v,v)\;w\otimes w^{\flat}-g(v,w)\;(v\otimes w^{\flat}+w\otimes v^{\flat})\right], (A.12) implying (vwwv)2\displaystyle(v\otimes w^{\flat}-w\otimes v^{\flat})^{2} =g(w,w)vvg(v,v)ww+g(v,w)(vw+wv)\displaystyle=-g(w,w)\;v\otimes v^{\flat}-g(v,v)\;w\otimes w^{\flat}+g(v,w)\;(v\otimes w^{\flat}+w\otimes v^{\flat}) =[g(v,v)g(w,w)(g(v,w))2]Prspan(v,w).\displaystyle=-\left[g(v,v)g(w,w)-(g(v,w))^{2}\right]\mathrm{Pr}_{\mathrm{span}(v,w)}. (A.13) Using this and BabPrab=BabB_{ab}\circ\mathrm{Pr}_{ab}=B_{ab}, the exponential series evaluates to

exp(αBab)\displaystyle\exp(\alpha B_{ab}) =(idVPrab)+k=01(2k)!α2k(εab)kPrab\displaystyle=(\mathrm{id}_{V}-\mathrm{Pr}_{ab})+\sum_{k=0}^{\infty}\frac{1}{(2k)!}\,\alpha^{2k}(-\varepsilon_{ab})^{k}\,\mathrm{Pr}_{ab}
+k=01(2k+1)!α2k+1(εab)kBabPrab\displaystyle\quad+\sum_{k=0}^{\infty}\frac{1}{(2k+1)!}\,\alpha^{2k+1}(-\varepsilon_{ab})^{k}\,B_{ab}\circ\mathrm{Pr}_{ab}
=(idVPrab)+{cos(α)idV+sin(α)Bab,εab=+1cosh(α)idV+sinh(α)Bab,εab=1}Prab.\displaystyle=(\mathrm{id}_{V}-\mathrm{Pr}_{ab})+\left\{\!\!\begin{aligned} &\cos(\alpha)\,\mathrm{id}_{V}+\sin(\alpha)\,B_{ab}\,,&\varepsilon_{ab}=+1\\ &\cosh(\alpha)\,\mathrm{id}_{V}+\sinh(\alpha)\,B_{ab}\,,&\varepsilon_{ab}=-1\end{aligned}\right\}\circ\mathrm{Pr}_{ab}\;. (A.14)

Geometrically, this transformation is either a rotation by angle α\alpha (for εab=+1\varepsilon_{ab}=+1) or a boost by rapidity α\alpha (for εab=1\varepsilon_{ab}=-1) in the plane span{ea,eb}\mathrm{span}\{e_{a},e_{b}\}. The direction of the transformation depends on the signs of gaa,gbbg_{aa},g_{bb}:

  • εab=+1\varepsilon_{ab}=+1:

    1. (i)

      gaa=gbb=+1g_{aa}=g_{bb}=+1: We have Bab(ea)=eb,Bab(eb)=eaB_{ab}(e_{a})=-e_{b},B_{ab}(e_{b})=e_{a}. Thus, exp(αBab)\exp(\alpha B_{ab}) is a rotation by α\alpha from ebe_{b} towards eae_{a}.

    2. (ii)

      gaa=gbb=1g_{aa}=g_{bb}=-1: We have Bab(ea)=eb,Bab(eb)=eaB_{ab}(e_{a})=e_{b},B_{ab}(e_{b})=-e_{a}. Thus, exp(αBab)\exp(\alpha B_{ab}) is a rotation by α\alpha from eae_{a} towards ebe_{b}.

  • εab=1\varepsilon_{ab}=-1:

    1. (i)

      gaa=+1,gbb=1g_{aa}=+1,g_{bb}=-1: We have Bab(ea)=eb,Bab(eb)=eaB_{ab}(e_{a})=-e_{b},B_{ab}(e_{b})=-e_{a}. Thus, exp(αBab)\exp(\alpha B_{ab}) is a boost by α\alpha ‘away’ from ea+ebe_{a}+e_{b}.

    2. (ii)

      gaa=1,gbb=+1g_{aa}=-1,g_{bb}=+1: We have Bab(ea)=eb,Bab(eb)=eaB_{ab}(e_{a})=e_{b},B_{ab}(e_{b})=e_{a}. Thus, exp(αBab)\exp(\alpha B_{ab}) is a boost by α\alpha ‘towards’ ea+ebe_{a}+e_{b}.

Now we will apply the preceding considerations to the case of (the ‘difference’ vector space of) Minkowski spacetime, where for now we leave open the signature convention for the metric (either (+)(+{--}-) or (+++)(-{++}+)). We work with respect to a positively oriented orthonormal basis {eμ}μ=0,,3\{e_{\mu}\}_{\mu=0,\dots,3} where e0e_{0} is timelike. Latin indices will denote spacelike directions.

In the case of ‘mostly minus’ signature (+)(+{--}-), BabB_{ab} generates rotations from eae_{a} towards ebe_{b} and Ba0B_{a0} generates boosts (with respect to e0e_{0}) in direction of eae_{a}. In the case of ‘mostly plus’ signature (+++)(-{++}+), Bba=BabB_{ba}=-B_{ab} generates rotations from eae_{a} towards ebe_{b} and B0a=Ba0B_{0a}=-B_{a0} generates boosts (with respect to e0e_{0}) in direction of eae_{a}.

Thus, since we want to use the notation JabJ_{ab} for the spacelike rotational generator generating rotations from eae_{a} towards ebe_{b}, we have to set

Jμν={Bμνfor (+) signature,Bμνfor (+++) signatureJ_{\mu\nu}=\begin{cases}B_{\mu\nu}&\text{for $(+{--}-)$ signature},\\ -B_{\mu\nu}&\text{for $(-{++}+)$ signature}\end{cases} (A.15)

for the Lorentz generators. Adopting this convention, boosts in direction of eae_{a} are then generated by Ja0J_{a0}. The commutation relations for the JμνJ_{\mu\nu} are

[Jμν,Jρσ]={ημσJνρ+(antisymm.)for (+) signature,ημρJνσ+(antisymm.)for (+++) signature,[J_{\mu\nu},J_{\rho\sigma}]=\begin{cases}\eta_{\mu\sigma}J_{\nu\rho}+\text{(antisymm.)}&\text{for $(+{--}-)$ signature},\\ \eta_{\mu\rho}J_{\nu\sigma}+\text{(antisymm.)}&\text{for $(-{++}+)$ signature},\end{cases} (A.16)

and general Lorentz algebra elements ωLie()\omega\in\mathrm{Lie}(\mathcal{L}) can be written as

ω=±12ωμνJμνwithωμν=ωρμηρν\omega=\pm\frac{1}{2}\omega^{\mu\nu}J_{\mu\nu}\;\text{with}\;\omega^{\mu\nu}=\omega^{\mu}_{\hphantom{\mu}\rho}\;\eta^{\rho\nu} (A.17)

in terms of their components ωρμ\omega^{\mu}_{\hphantom{\mu}\rho} as endomorphisms, where the upper/lower sign holds for (+)(+{--}-)/(+++)(-{++}+) signature.

Appendix B Notes on the adjoint representation

Here we wish to make a few remarks and collect a few formulae concerning the adjoint and co-adjoint representation, which will be made use of in the main text.

In the defining representation on VV, an element Λ𝖦𝖫(V)\Lambda\in\mathsf{GL}(V) is given in terms of the basis {ea}a\{e_{a}\}_{a} by the coefficients Λba\Lambda^{a}_{\phantom{a}{}_{b}}, where

Λea=Λabeb.\Lambda e_{a}=\Lambda^{b}_{\phantom{b}a}\,e_{b}\;. (B.1)

This defines a left action of 𝖦𝖫(V)\mathsf{GL}(V) on VV. The corresponding left action of 𝖦𝖫(V)\mathsf{GL}(V) on the dual space VV^{*} is given by the inverse-transposed, i.e. 𝖦𝖫(V)×VV\mathsf{GL}(V)\times V^{*}\to V^{*}, (Λ,α)(Λ1)α:=αΛ1(\Lambda,\alpha)\mapsto(\Lambda^{-1})^{\top}\alpha:=\alpha\circ\Lambda^{-1}. For the basis {θa}a\{\theta^{a}\}_{a} of VV^{*} dual to the basis {ea}a\{e_{a}\}_{a} this means

θaΛ1=(Λ1)baθb.\theta^{a}\circ\Lambda^{-1}=(\Lambda^{-1})^{a}_{\phantom{a}b}\,\theta^{b}\;. (B.2)

In contrast, for the basis {ea}a\{e_{a}^{\flat}\}_{a} of VV^{*}, this reads in general

ebΛ1=gacgbd(Λ1)cdea,e_{b}^{\flat}\circ\Lambda^{-1}=g^{ac}g_{bd}(\Lambda^{-1})^{d}_{\phantom{d}c}\,e_{a}^{\flat}\;, (B.3)

which for isometries Λ𝖮(V,g)\Lambda\in\mathsf{O}(V,g) simply becomes

ebΛ1=Λbaea.e^{\flat}_{b}\circ\Lambda^{-1}=\Lambda^{a}_{\phantom{a}b}\,e^{\flat}_{a}\;. (B.4)

The adjoint representation of 𝖦𝖫(V)\mathsf{GL}(V) on End(V)VV\mathrm{End}(V)\cong V\otimes V^{*} or any Lie subalgebra of End(V)\mathrm{End}(V) is by conjugation, which for our basis (A.2) implies, using (B.1) and (B.4),

AdΛBab=ΛBabΛ1=ΛacΛbdBcdforΛ𝖮(V,g).\mathrm{Ad}_{\Lambda}B_{ab}=\Lambda\circ B_{ab}\circ\Lambda^{-1}=\Lambda^{c}_{\phantom{c}a}\Lambda^{d}_{\phantom{d}b}\,B_{cd}\quad\text{for}\;\Lambda\in\mathsf{O}(V,g). (B.5)

The adjoint representation of the inhomogeneous group 𝖦𝖫(V)V\mathsf{GL}(V)\ltimes V on its Lie algebra End(V)V\mathrm{End}(V)\oplus V is given by, for any XEnd(V)X\in\mathrm{End}(V) and yVy\in V,

Ad(Λ,a)(X,y)=(ΛXΛ1,Λy(ΛXΛ1)a).\mathrm{Ad}_{(\Lambda,a)}(X,y)=\left(\Lambda\circ X\circ\Lambda^{-1},\Lambda y-(\Lambda\circ X\circ\Lambda^{-1})a\right). (B.6)

In the main text we will use this formula for (Λ,a)(\Lambda,a) being replaced by its inverse (Λ,a)1=(Λ1,Λ1a)(\Lambda,a)^{-1}=(\Lambda^{-1},-\Lambda^{-1}a):

Ad(Λ,a)1(X,y)=(Λ1XΛ,Λ1y+(Λ1X)a)\mathrm{Ad}_{(\Lambda,a)^{-1}}(X,y)=\left(\Lambda^{-1}\circ X\circ\Lambda,\Lambda^{-1}y+(\Lambda^{-1}\circ X)a\right) (B.7)

Applied to the basis vectors separately, i.e. to (X,y)=(0,eb)(X,y)=(0,e_{b}) and (X,y)=(Bbc,0)(X,y)=(B_{bc},0), for Λ𝖮(V,g)\Lambda\in\mathsf{O}(V,g) we get

Ad(Λ,a)1(0,eb)\displaystyle\mathrm{Ad}_{(\Lambda,a)^{-1}}(0,e_{b}) =(0,(Λ1)bcec)\displaystyle=\left(0,(\Lambda^{-1})^{c}_{\phantom{c}b}e_{c}\right) (B.8a)
Ad(Λ,a)1(Bbc,0)\displaystyle\mathrm{Ad}_{(\Lambda,a)^{-1}}(B_{bc},0) =((Λ1)bd(Λ1)ceBde,ab(Λ1)cded+ac(Λ1)bded)\displaystyle=\left((\Lambda^{-1})^{d}_{\phantom{d}b}(\Lambda^{-1})^{e}_{\phantom{e}c}\,B_{de},-a_{b}(\Lambda^{-1})^{d}_{\phantom{d}c}e_{d}+a_{c}(\Lambda^{-1})^{d}_{\phantom{d}b}e_{d}\right) (B.8b)

where ab:=eb(a)=gbcaca_{b}:=e_{b}^{\flat}(a)=g_{bc}a^{c} in the second equation. From these equations we immediately deduce (2.21) in the case of four spacetime dimensions (greek indices) and signature mostly plus, in which case Jμν=BμνJ_{\mu\nu}=-B_{\mu\nu} according to (A.15).