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

\WarningFilter

latexWriting or overwriting \WarningFilterrevtex4-2Failed to recognize

Pairwise Helicity in Higher Dimensions

Yale Fan Weinberg Institute for Theoretical Physics, Department of Physics, University of Texas at Austin, Austin, TX 78712, USA
Abstract

Studies of scattering amplitudes for electric and magnetic charges have identified previously overlooked multiparticle representations of the Poincaré group in four dimensions. Such representations associate nontrivial quantum numbers (known as pairwise helicities) with asymptotically separated pairs of particles, and thus cannot be described as tensor products of one-particle states. We extend this construction to sources and spacetimes of higher dimension. We first establish the dynamical origin of pairwise helicity in pp-form electrodynamics coupled to mutually nonlocal branes. We then interpret this pairwise helicity as a quantum number under an SO(2)SO(2) pairwise little group associated with pairs of distinct branes. We further characterize the “higher” little groups that could in principle be used to induce multiparticle or multi-brane representations of the Lorentz group.

I Introduction

A foundational assumption in most discussions of the SS-matrix in relativistic quantum field theory is that asymptotic states should factorize as tensor products of single-particle states, as classified by Wigner [1]. However, this assumption breaks down for the scattering of particles with both electric and magnetic charge, for which the angular momentum in the electromagnetic field remains finite even at asymptotic separation [2]. This effect is codified by the notion of pairwise helicity, or the charge under the pairwise little group, defined as the subgroup of the Lorentz group that leaves the momenta of a given pair of particles invariant [3, 4]. This group has nontrivial representations on certain multiparticle states, leading to fundamentally new multiparticle representations of the Poincaré group in four dimensions.

The pairwise little group naturally generalizes to the “nn-particle” little group for n>2n>2, which becomes nontrivial in sufficiently high dimension. One might wonder whether such a group can be represented nontrivially in any physical theory. While this compelling question remains open, we show in this paper that there does exist a physically manifest generalization of the pairwise little group to higher dimensions: the pairwise little group of mutually nonlocal branes.

To motivate this construction, we first recall how the pairwise helicity of four-dimensional multiparticle states arises dynamically from the electromagnetic angular momentum of dyons. While irreducible single-particle representations of the Poincaré group are induced by representations of the corresponding little group, representations of the pairwise little group induce multiparticle representations that go beyond tensor products of single-particle representations.

We next compute the angular momentum in the electromagnetic field between mutually nonlocal branes of arbitrary (complementary) dimensions. It admits a simple expression in terms of the Dirac-Schwinger-Zwanziger (DSZ) pairing [5, 6, 7] within pp-form electrodynamics. To provide a framework for this higher-dimensional analogue of pairwise helicity, we define representations of the Lorentz group that describe infinitely extended, isotropic branes. (For reasons that will become clear, such representations do not generally lift to representations of the Poincaré group.) A (p1)(p-1)-brane representation of the Lorentz group in dd spacetime dimensions is induced by a representation of the corresponding little group, which contains the group of transverse rotations SO(dp)SO(d-p). Multi-brane representations can further carry nontrivial quantum numbers under little groups associated with pairs of distinct branes. Such pairwise little groups include rotations in the directions mutually transverse to both branes. When this subgroup of mutually transverse rotations is SO(2)SO(2), we refer to the associated charge as the pairwise helicity.

Finally, we offer some thoughts on the higher “nn-brane” little groups that could in principle induce more general multi-brane representations of the Lorentz group.

II Pairwise Helicity for Particles

Our analysis of (p1)(p-1)-branes in dd dimensions builds on the familiar case d=4d=4, p=1p=1. Recall that particles (0-branes) furnish unitary irreducible representations of the Poincaré group. Such representations can be classified by their mass and their representation with respect to the little group, the subgroup of the Lorentz group that leaves some reference momentum kk of that mass invariant [1]. A general Lorentz transformation Λ\Lambda acts on a single-particle momentum eigenstate |p;σ|p;\sigma\rangle as

U(Λ)|p;σ=Dσσ(W)|Λp;σ,U(\Lambda)|p;\sigma\rangle=D_{\sigma^{\prime}\sigma}(W)|\Lambda p;\sigma^{\prime}\rangle, (1)

where W=W(Λ,p,k)W=W(\Lambda,p,k) is a little group transformation, DD is a unitary representation of WW, and the single-particle quantum number σ\sigma labels a little group representation. A particle representation of the Poincaré group (and in particular, of the Lorentz group) is thus induced by a unitary irreducible representation of the little group.

On the other hand, multiparticle representations of the Poincaré group may carry additional quantum numbers that characterize the Lorentz transformation properties of multiparticle states relative to tensor products of one-particle states [3, 4]. Given a generic pair of reference four-momenta (either massive or massless), the pairwise little group is defined as the SO(2)SO(2) subgroup of the Lorentz group that preserves both. Since three-particle and higher little groups are trivial in four dimensions, a general nn-particle state takes the form

|p1,,pn;σ1,,σn;q12,,qn1,n,|p_{1},\ldots,p_{n};\sigma_{1},\ldots,\sigma_{n};q_{12},\ldots,q_{n-1,n}\rangle, (2)

where pip_{i} is the momentum of particle ii, the σi\sigma_{i} are single-particle quantum numbers (spins or helicities), and the (n2)\binom{n}{2} quantized pairwise helicities qijq_{ij} are associated with pairs of particles i,ji,j. Under a Lorentz transformation Λ\Lambda, these states may transform with extra phases arising from nontrivial pairwise little group representations:

U(Λ)|p1,,pn;σ1,,σn;q12,,qn1,n=i<jeiqijϕij\displaystyle U(\Lambda)|p_{1},\ldots,p_{n};\sigma_{1},\ldots,\sigma_{n};q_{12},\ldots,q_{n-1,n}\rangle=\prod_{i<j}e^{iq_{ij}\phi_{ij}}
×i=1nDσiσii|Λp1,,Λpn;σ1,,σn;q12,,qn1,n.\displaystyle\times\prod_{i=1}^{n}D_{\sigma_{i}^{\prime}\sigma_{i}}^{i}|\Lambda p_{1},\ldots,\Lambda p_{n};\sigma_{1}^{\prime},\ldots,\sigma_{n}^{\prime};q_{12},\ldots,q_{n-1,n}\rangle. (3)

The angles ϕij=ϕ(Λ,pi,pj)\phi_{ij}=\phi(\Lambda,p_{i},p_{j}) parametrize pairwise little group transformations, while the DD’s are unitary representations of one-particle little group transformations.

There is precisely one setting in which nontrivial pairwise helicities are known to arise: the dynamics of electric and magnetic charges. If particles i,ji,j are dyons of charges (ei,gi)(e_{i},g_{i}) and (ej,gj)(e_{j},g_{j}), then qijq_{ij} is simply the quantized angular momentum 14π(eigjejgi)\frac{1}{4\pi}(e_{i}g_{j}-e_{j}g_{i}) in their mutual electromagnetic field. This is easiest to see in a Lorentz frame where their spatial momenta are equal and opposite. Indeed, for a spinless two-dyon state, qijq_{ij} is the charge under rotations about the axis of the two asymptotically separated dyons. A careful derivation of this result for non-static sources is given in [2].

III Pairwise Helicity for Branes

The above discussion makes clear that pairwise helicity in four dimensions is a reflection of the mutual nonlocality of the underlying electromagnetic sources.

A natural starting point for generalizing pairwise helicity to higher dimensions is to consider pp-form electrodynamics in dd-dimensional Minkowski space. In this theory, an electrically charged (p1)(p-1)-brane couples directly to the pp-form gauge potential AA. On the other hand, a magnetic (dp3)(d-p-3)-brane [8, 9, 10] can be thought of as coupling to AA through a (dp1)(d-p-1)-form current GG localized to the worldvolume of a (dp2)(d-p-2)-dimensional Dirac brane (or generalized Dirac string [5]) that it bounds. The total (p+1)(p+1)-form field strength is then given by F=dA+GF=dA+\ast G.

Let us label the time coordinate of d1,1\mathbb{R}^{d-1,1} by x0x_{0} and the spatial coordinates by

(x,y,z)(x1,x2,x3,y1,,yp1,z1,,zdp3).(\vec{x},\vec{y},\vec{z})\equiv(x_{1},x_{2},x_{3},y_{1},\ldots,y_{p-1},z_{1},\ldots,z_{d-p-3}). (4)

Suppose we have an electric (p1)(p-1)-brane lying along y\vec{y} and a magnetic (dp3)(d-p-3)-brane lying along z\vec{z}. For a magnetic brane (“ii”) of charge density gig_{i} and an electric brane (“jj”) of charge density eje_{j} in a configuration such as that of (4), we define their pairwise helicity to be

qij=14π{(1)pejgiif d2(p+1),eigj+(1)pejgiif d=2(p+1).q_{ij}=\frac{1}{4\pi}\begin{cases}(-1)^{p}e_{j}g_{i}&\text{if $d\neq 2(p+1)$},\\ e_{i}g_{j}+(-1)^{p}e_{j}g_{i}&\text{if $d=2(p+1)$}.\end{cases} (5)

Note that the branes can be dyonic when d=2(p+1)d=2(p+1), in which case we additionally assign the magnetic brane an electric charge density eie_{i} and the electric brane a magnetic charge density gjg_{j}.

The quantity qijq_{ij} in (5) is quantized as a half-integer. This statement is precisely the Dirac quantization condition when d2(p+1)d\neq 2(p+1) [8, 9, 10, 11] and the more refined DSZ quantization condition when d=2(p+1)d=2(p+1) [12, 13, 14, 15, 16]. Indeed, the trajectories of electric and magnetic branes can link in (d1)(d-1)-dimensional space. Hence amplitudes involving these objects incur an Aharonov-Bohm phase, and quantum-mechanical consistency of the theory places constraints on the allowed electric and magnetic charges.

More directly, however, qijq_{ij} is nothing other than the angular momentum in the electromagnetic field of branes ii and jj. To see this, we focus on the spatial directions d1\mathbb{R}^{d-1} in (4). We take brane ii (which fills the z\vec{z}-directions) to lie at the origin in x\vec{x}, while we take brane jj (which fills the y\vec{y}-directions) to lie at the point R=(0,0,R3)\smash{\vec{R}}=(0,0,R_{3}) in x\vec{x}. The angular momentum corresponding to rotations in the x1,x2x_{1},x_{2} directions is

J12=dd1x(x1T02x2T0)1,J_{12}=\int d^{d-1}x\,(x_{1}T^{0}{}_{2}-x_{2}T^{0}{}_{1}), (6)

where T0=i1p!F0Fii1ipi1ipT^{0}{}_{i}=\frac{1}{p!}F^{0}{}_{i_{1}\cdots i_{p}}F_{i}{}^{i_{1}\cdots i_{p}} is the linear momentum density of the gauge field. We can write the electric and magnetic fields sourced by the branes as differential forms EE and BB on d1\mathbb{R}^{d-1}, with components Ei1ip=F0i1ipE^{i_{1}\cdots i_{p}}=F^{0i_{1}\cdots i_{p}} and Bi1idp2=(1)p(F)0i1idp2B^{i_{1}\cdots i_{d-p-2}}=(-1)^{p}(\ast F)^{0i_{1}\cdots i_{d-p-2}}. They satisfy

dE=(1)p1ejΣ^j,dB=(1)dp3giΣ^i,d\ast E=(-1)^{p-1}e_{j}\widehat{\Sigma}_{j},\quad d\ast B=(-1)^{d-p-3}g_{i}\widehat{\Sigma}_{i}, (7)

where Σ^i,j\smash{\widehat{\Sigma}_{i,j}} denote the Poincaré duals of the brane submanifolds Σi,jd1\Sigma_{i,j}\subset\mathbb{R}^{d-1} (the operations dd, \ast, and ^\widehat{\phantom{m}} in (7) are all defined with respect to d1\mathbb{R}^{d-1}). Using (7), we find:

J12=(1)pejgi4πR3|R3|,J_{12}=\frac{(-1)^{p}e_{j}g_{i}}{4\pi}\frac{R_{3}}{|R_{3}|}, (8)

which exactly reproduces (5). When d=2(p+1)d=2(p+1), the eigje_{i}g_{j} term in (5) follows from a symmetry argument. Crucially, the result (8) is independent of the separation |R|=|R3||\smash{\vec{R}}|=|R_{3}|, so it holds equally well for branes that are asymptotically separated in the x3x_{3}-direction.111Our mathematical conventions, details on the derivation of (8), and additional context on pp-form electrodynamics can be found in Appendices B, C, and D.

Having explained the dynamical meaning of q12q_{12}, we would like to show that it deserves the moniker “pairwise helicity,” i.e., to interpret it as an SO(2)SO(2) pairwise little group charge. We propose that states of multiple branes can transform with nontrivial pairwise helicities when they contain pairs of mutually nonlocal branes with SO(2)SO(2) pairwise little groups. To formalize this statement, we now suggest a framework in which to understand pairwise helicity more systematically.

IV Brane Representations of the Lorentz Group

It is interesting to ask whether branes, like particles, can be described as representations of an appropriate group of spacetime symmetries. Since branes have infinitely many more degrees of freedom than particles, making such a question tractable requires certain simplifying assumptions. For this reason, we concentrate on (p1)(p-1)-branes of positive tension in dd spacetime dimensions that are completely isotropic [17]. The tangent space to a point on the brane worldvolume is a pp-dimensional linear subspace of Minkowski space that contains a timelike direction. This “brane subspace” is characterized by a decomposition

ημν=Πμν+(Π)μν,\eta^{\mu\nu}=\Pi^{\mu\nu}+(\Pi^{\perp})^{\mu\nu}, (9)

where Π\Pi and Π\Pi^{\perp} are orthogonal projection operators onto the parallel and transverse directions. They have rank pp and dpd-p, respectively. The projection tensor Π\Pi is manifestly invariant under internal SO(p1,1)SO(p-1,1) Lorentz transformations, i.e., independent of the choice of basis for the brane subspace.

In a given SO(d1,1)SO(d-1,1) Lorentz frame, the relativistic dd-velocity of the brane is the normalized timelike vector in this subspace given by

uμΠ0μΠ00.u^{\mu}\equiv\frac{\Pi^{0\mu}}{\sqrt{\Pi^{00}}}. (10)

We use “mostly minus” signature, so that uμuμ=1u_{\mu}u^{\mu}=1. The spatial orientation of the brane is then characterized by a spatial projection tensor Ξ\Xi of rank p1p-1:

ΞμνΠμνuμuν.\Xi^{\mu\nu}\equiv\Pi^{\mu\nu}-u^{\mu}u^{\nu}. (11)

In any Lorentz frame, the brane’s dd-velocity is parallel to its worldvolume directions (Πμuνν=uμ\Pi^{\mu}{}_{\nu}u^{\nu}=u^{\mu}) but orthogonal to its p1p-1 purely spatial directions (Ξμuνν=0\Xi^{\mu}{}_{\nu}u^{\nu}=0).

Note that, despite our notation, uu and Ξ\Xi do not generally transform as Lorentz tensors. On the other hand, because the brane orientation Π\Pi transforms as a symmetric two-index Lorentz tensor, we can use it to construct representations of the Lorentz group SO(d1,1)SO(d-1,1). First, we observe that the subgroup of the Lorentz group that leaves Π\Pi invariant is SO(p1,1)×SO(dp)SO(p-1,1)_{\parallel}\times SO(d-p)_{\perp}, corresponding to rotations or boosts in the directions parallel (\parallel) and transverse (\perp) to the brane. Indeed, we can always choose a Lorentz frame in which the worldvolume projection tensor takes the standard form

Π~μν=diag(1,1,,1p,0,,0dp).\tilde{\Pi}^{\mu\nu}=\operatorname{diag}(\underbrace{1,-1,\ldots,-1}_{p},\underbrace{0,\ldots,0}_{d-p}). (12)

This can be achieved by boosting to the brane’s rest frame, where u~μ=(1,0,,0)\tilde{u}^{\mu}=(1,0,\ldots,0), and performing an appropriate spatial rotation. For simplicity, we consider only brane representations of the Lorentz group in which the noncompact SO(p1,1)SO(p-1,1)_{\parallel} factor is represented trivially. We refer to the compact SO(dp)SO(d-p)_{\perp} factor as the single-brane little group, dropping the subscript \perp from here on. To specify a representation of the SO(dp)SO(d-p) little group, we introduce a collective label σ\sigma that includes all necessary Casimirs. We call σ\sigma the “spin” under SO(dp)SO(d-p).

Following Wigner, we introduce a collection of single-brane states |Π;σ|\Pi;\sigma\rangle that transform in an infinite-dimensional unitary representation of the dd-dimensional Lorentz group according to

U(Λ)|Π;σ=Dσσ(W)|ΛΠΛ1;σ,U(\Lambda)|\Pi;\sigma\rangle=D_{\sigma^{\prime}\sigma}(W)|\Lambda\Pi\Lambda^{-1};\sigma^{\prime}\rangle, (13)

where W=W(Λ,Π,Π~)W=W(\Lambda,\Pi,\tilde{\Pi}) is a little group transformation. (Here, we write Π\Pi for the matrix Πμν\Pi^{\mu}{}_{\nu} without indices.) These brane states, unlike particle states, are not constructed as eigenstates of the translation generators of the Poincaré algebra. Indeed, while the dd-momentum per unit volume is given by pμ=Tuμp^{\mu}=Tu^{\mu} in terms of the brane tension TT (with mass dimension [T]=p[T]=p), pμp^{\mu} does not correspond to the integrated Poincaré charge PμP^{\mu} (which diverges in states of infinite branes), nor does it transform as a Lorentz vector.222Further details on this construction and its subsequent generalization to multiple branes are given in Appendix A.

In the case of a particle (p=1p=1), we have T=m>0T=m>0 and Πμν=uμuν\Pi^{\mu\nu}=u^{\mu}u^{\nu}. Hence Π\Pi and uu contain precisely the same information, under the assumption that the timelike component of uu is positive. In this case, (10) becomes a tautology and uu transforms as a genuine Lorentz vector. The little group reduces to the expected SO(d1)SO(d-1) for a massive particle. In the opposite case of a space-filling brane (p=dp=d), we have Πμν=ημν\Pi^{\mu\nu}=\eta^{\mu\nu} in any frame, and uu transforms as a Lorentz scalar.

We now illustrate how the definitions of pairwise little group and pairwise helicity generalize to a state of two branes relevant to the calculation of the preceding section. We then turn to multi-brane states in generality.

Consider an electric (p1)(p-1)-brane and a magnetic (dp3)(d-p-3)-brane that span orthogonal directions in space, as in (4). In other words, we suppose that there exists a frame in which their spatial projection tensors Ξ1\Xi_{1} and Ξ2\Xi_{2} take the standard forms

(Ξ~1)μν\displaystyle(\tilde{\Xi}_{1})^{\mu\nu} =diag(0,0,0,0,1,,1p1,0,,0dp3),\displaystyle=-\operatorname{diag}(0,0,0,0,\underbrace{1,\ldots,1}_{p-1},\underbrace{0,\ldots,0}_{d-p-3}), (14)
(Ξ~2)μν\displaystyle(\tilde{\Xi}_{2})^{\mu\nu} =diag(0,0,0,0,0,,0p1,1,,1dp3).\displaystyle=-\operatorname{diag}(0,0,0,0,\underbrace{0,\ldots,0}_{p-1},\underbrace{1,\ldots,1}_{d-p-3}). (15)

Suppose also that the branes have generic (spatially transverse) dd-velocities u1u_{1}, u2u_{2}. Via an SO(3,1)SO(d1,1)SO(3,1)\subset SO(d-1,1) Lorentz transformation acting only on (x0,x)(x_{0},\vec{x}), which preserves the spatial projection tensors (14)–(15), any u1u_{1}, u2u_{2} may be brought to the standard forms

u~1μ=(u10,0,0,+uc,0(p1),u1(dp3)),\displaystyle\tilde{u}_{1}^{\mu}=(u_{1}^{0},0,0,+u_{c},\vec{0}^{(p-1)},\vec{u}_{1}^{(d-p-3)}), (16)
u~2μ=(u20,0,0,uc,u2(p1),0(dp3)),\displaystyle\tilde{u}_{2}^{\mu}=(u_{2}^{0},0,0,-u_{c},\vec{u}_{2}^{(p-1)},\vec{0}^{(d-p-3)}), (17)

where (ui0)2uc2(ui)2=1(u_{i}^{0})^{2}-u_{c}^{2}-(\vec{u}_{i})^{2}=1 for i=1,2i=1,2. This frame makes manifest that there exists an SO(2)SO(2) subgroup of the Lorentz group that preserves Ξ~1\tilde{\Xi}_{1}, Ξ~2\tilde{\Xi}_{2}, and any two generic transverse dd-velocities simultaneously.333Although the full pairwise little group of two branes in the configuration (14)–(17) is SO(2)×SO(p2)×SO(dp4)SO(2)\times SO(p-2)\times SO(d-p-4), we focus on the SO(2)SO(2) subgroup of external rotations because it will turn out to admit a transparent physical interpretation. When the x\vec{x}-components of their velocities are equal and opposite along the x3x_{3}-axis as in (16)–(17), the SO(2)SO(2) pairwise little group of these two branes consists of rotations R12R_{12} in the plane of x1,x2x_{1},x_{2}.

We can now define a two-brane representation of the Lorentz group with respect to the reference projectors (Π~i)μν=(u~i)μ(u~i)ν+(Ξ~i)μν(\tilde{\Pi}_{i})^{\mu\nu}=(\tilde{u}_{i})^{\mu}(\tilde{u}_{i})^{\nu}+(\tilde{\Xi}_{i})^{\mu\nu} defined by (14)–(17). This representation is carried by states |Π1,Π2;σ1,σ2;σ12|\Pi_{1},\Pi_{2};\sigma_{1},\sigma_{2};\sigma_{12}\rangle, where the σi\sigma_{i} label single-brane little group representations and σ12q12\sigma_{12}\equiv q_{12} labels a representation of the aforementioned SO(2)SO(2) pairwise little group. These states transform according to

U(Λ)|Π1,Π2;σ1,σ2;q12=\displaystyle U(\Lambda)|\Pi_{1},\Pi_{2};\sigma_{1},\sigma_{2};q_{12}\rangle= (18)
eiq12ϕ12Dσ1σ1Dσ2σ2|ΛΠ1Λ1,ΛΠ2Λ1;σ1,σ2;q12,\displaystyle e^{iq_{12}\phi_{12}}D_{\sigma_{1}^{\prime}\sigma_{1}}D_{\sigma_{2}^{\prime}\sigma_{2}}|\Lambda\Pi_{1}\Lambda^{-1},\Lambda\Pi_{2}\Lambda^{-1};\sigma_{1}^{\prime},\sigma_{2}^{\prime};q_{12}\rangle,

where the DD’s are unitary representations of single-brane little group transformations. As any pairwise little group transformation takes the form W12=R12(ϕ12)W_{12}=R_{12}(\phi_{12}) for some angle ϕ12=ϕ(Λ,Π1,Π2)\phi_{12}=\phi(\Lambda,\Pi_{1},\Pi_{2}), we have written D(W12)=eiq12ϕ12D(W_{12})=e^{iq_{12}\phi_{12}}. The SO(2)SO(2) charge q12q_{12}, or the pairwise helicity, can thus be identified with the electromagnetic angular momentum of these mutually nonlocal branes (which are asymptotically separated along the x3x_{3}-axis). Furthermore, we have seen that q12q_{12} satisfies a quantization condition that gives rise to a bona fide representation of SO(2)SO(2).

V Multi-Brane Representations of the Lorentz Group

Finally, we define general states of asymptotically separated, non-intersecting branes in d1,1\mathbb{R}^{d-1,1}, in the spirit of the four-dimensional multiparticle states of [3, 4]. Such states may have quantum numbers that are not captured by tensor products of single-brane states.

We define a general NN-brane Lorentz representation by a set of reference projection tensors {Π~i|i=1,,N}\{\tilde{\Pi}_{i}\,|\,i=1,\ldots,N\} and a set of representations {σ|}\{\sigma_{\ell}\,|\,\ell\in\mathcal{L}\}, where \mathcal{L} indexes all possible little groups GG_{\ell}, which are subgroups of the Lorentz group that stabilize the reference data. Explicitly, we take \mathcal{L} to consist of all 2N12^{N}-1 nontrivial subsets of {1,,N}\{1,\ldots,N\}. For each \ell\in\mathcal{L}, we take GG_{\ell} to consist of those Lorentz transformations λSO(d1,1)\lambda\in SO(d-1,1) for which

λΠ~iλ1=Π~ifor all i∈ℓ.\lambda\tilde{\Pi}_{i}\lambda^{-1}=\tilde{\Pi}_{i}\quad\text{for all {\hbox{i\in\ell}}}. (19)

Every state in this representation is defined in terms of some canonical Lorentz transformations relating it to the reference state |Π~1,,Π~N;{σ|}|\tilde{\Pi}_{1},\ldots,\tilde{\Pi}_{N};\{\sigma_{\ell}\,|\,\ell\in\mathcal{L}\}\rangle, and an arbitrary state transforms as

U(Λ)|Π1,,ΠN;{σ|}=\displaystyle U(\Lambda)|\Pi_{1},\ldots,\Pi_{N};\{\sigma_{\ell}\,|\,\ell\in\mathcal{L}\}\rangle= (20)
[Dσσ(W)]|ΛΠ1Λ1,,ΛΠNΛ1;{σ|},\displaystyle\left[\prod_{\ell\in\mathcal{L}}D_{\sigma_{\ell}^{\prime}\sigma_{\ell}}(W_{\ell})\right]|\Lambda\Pi_{1}\Lambda^{-1},\ldots,\Lambda\Pi_{N}\Lambda^{-1};\{\sigma_{\ell}^{\prime}\,|\,\ell\in\mathcal{L}\}\rangle,

where each WGW_{\ell}\in G_{\ell} is a little group transformation that depends on {Πi|i}\{\Pi_{i}\,|\,i\in\ell\} as well as on Λ\Lambda. Those NN-brane representations that are induced by specifying nontrivial representations σ\sigma_{\ell} only for ||=1|\ell|=1 are tensor products of single-brane representations. In principle, however, we are free to specify a nontrivial representation of the entire little group G\prod_{\ell\in\mathcal{L}}G_{\ell}. We have shown that the induced representations arising from nontrivial σ\sigma_{\ell} for ||=2|\ell|=2 have physical relevance, as was known to be the case even for particles [4].

In d=4d=4, all higher little groups beyond the pairwise ones are generically trivial. This is no longer true in higher dimensions. Therefore, it is natural to ask whether the higher little groups with ||>2|\ell|>2 have any significance in d>4d>4. We content ourselves with addressing a simpler question: given a set of reference data for nn branes, what is the nn-brane little group (for any n=||n=|\ell|)?

To approach this question, it is simplest to begin with the case of particles. We define the nn-particle little group as the subgroup of the dd-dimensional Lorentz group that leaves invariant an arbitrary set of nn momenta (or velocities). For generic momenta, this group is SO(dn)SO(d-n). The reasoning is simple: given nn generic (i.e., linearly independent) dd-vectors, this SO(dn)SO(d-n) preserves the orthogonal complement of their span.

Now consider nn branes with projection tensors Πi\Pi_{i} for i=1,,ni=1,\ldots,n. Let ViV_{i} denote the corresponding linear subspaces of Minkowski space d1,1\mathbb{R}^{d-1,1}, i.e., the images of the Πi\Pi_{i}. Lorentz transformations that preserve a given subspace of Minkowski space act purely within the subspace or purely within its orthogonal complement. Therefore, those Lorentz transformations acting within any of the subspaces

V1i1VninV_{1}^{i_{1}}\cap\cdots\cap V_{n}^{i_{n}} (21)

for (i1,,in){0,1}n(i_{1},\ldots,i_{n})\in\{0,1\}^{n}, where Vi0ViV_{i}^{0}\equiv V_{i} and Vi1ViV_{i}^{1}\equiv V_{i}^{\perp}, preserve all subspaces ViV_{i} simultaneously. The nn-brane little group is thus at least as large as

(i1,,in){0,1}nSO(dim(V1i1Vnin)),\prod_{(i_{1},\ldots,i_{n})\in\{0,1\}^{n}}SO(\dim(V_{1}^{i_{1}}\cap\cdots\cap V_{n}^{i_{n}})), (22)

where the group SO(m)SO(m) is understood to be trivial for integers m1m\leq 1.444Note that the subspace (21) can contain a timelike direction only when (i1,,in)=(0,,0)(i_{1},\ldots,i_{n})=(0,\ldots,0). If it does, then the signature of the corresponding SOSO factor in (22) should be adjusted accordingly. For our branes of interest, we would expect the corresponding SOSO factor to be represented trivially, since it consists only of Lorentz transformations internal to all branes. The actual little group may in fact be larger (in particular, we have ignored possible discrete subgroups).

Let us see what the formula (22) gives in special cases. If all nn branes are particles with generic dd-momenta, then none of the momenta are collinear. Hence the only factor that contributes to (22) is that with (i1,,in)=(1,,1)(i_{1},\ldots,i_{n})=(1,\ldots,1), giving back the nn-particle little group

SO(dim(V1Vn))=SO(dn),SO(\dim(V_{1}^{\perp}\cap\cdots\cap V_{n}^{\perp}))=SO(d-n), (23)

where we have used the fact that the intersection of nn generic hyperplanes ((d1)(d-1)-dimensional subspaces) has dimension dnd-n. In the other extreme case that all nn branes are space-filling, we get simply

SO(dim(V1Vn)1,1)=SO(d1,1).SO(\dim(V_{1}\cap\cdots\cap V_{n})-1,1)=SO(d-1,1). (24)

For a single (p1)(p-1)-brane (n=1n=1) spanning the subspace VV, we get the expected little group

SO(dimV1p1,1)×SO(dimVdp),SO(\underbrace{\dim V-1}_{p-1},1)\times SO(\underbrace{\dim V^{\perp}}_{d-p}), (25)

including the internal Lorentz transformations that we take to be represented trivially. Finally, consider a (p11)(p_{1}-1)-brane and a (p21)(p_{2}-1)-brane. Generically, the intersection of d1d_{1}- and d2d_{2}-dimensional linear subspaces has dimension max(d1+d2d,0)\max(d_{1}+d_{2}-d,0). So for generic configurations of these two branes, the corresponding pairwise little group always contains an SO(dp1p2)SO(d-p_{1}-p_{2}) factor.555Note that the specific two-brane configuration in (14)–(17) was chosen to simplify the angular momentum analysis for p1=dp22pp_{1}=d\linebreak[1]-\linebreak[1]p_{2}\linebreak[1]-2\equiv p, so the corresponding pairwise little group is larger than would be expected generically.

We have not presented an exhaustive classification of multi-brane representations of the Lorentz group. While the subgroup of SO(d1,1)SO(d-1,1) that preserves any given set of brane subspaces always contains a subgroup isomorphic to (22), we have not attempted to find the full stabilizer subgroup or to catalogue all such groups that could arise. In addition, we have excluded non-generic “tensionless” brane representations with null (lightlike) dd-velocities.

VI Discussion

In this paper, we have constructed representations of the Lorentz group that describe ideal branes whose worldvolume directions span linear subspaces of Minkowski spacetime. We have illustrated how quantum numbers associated to pairwise little groups of multi-brane representations can arise from the coupling of branes to abelian gauge fields in higher dimensions.

This study raises many more questions than answers. We highlight some of these questions below.

VI.1 Origin of Brane Representations

The representations of the Lorentz group constructed herein should be regarded as “phenomenological” models of a highly idealized class of brane states. While these representations allow one to proceed by formal analogy to the construction of multiparticle states, their physical status remains unclear. For instance, how might these brane representations embed into representations of the Poincaré group (or perhaps a different extension of the Lorentz group), and how might they be realized as physical states in a Poincaré-invariant theory of dynamical branes?

Unlike for particles, the usual Poincaré charges (energy and momentum) diverge in states of infinite branes, and are difficult to interpret as meaningful quantum numbers. Said differently, because states of infinite branes occupy different superselection sectors than particle states, it is not at all obvious that a Hilbert space of branes should carry a representation of the Poincaré group, even in a theory that is classically Poincaré-invariant.666It should be kept in mind that this paper does not treat branes as defects with fixed asymptotic boundary conditions that spontaneously break dd-dimensional Poincaré symmetry (such as flat defects p1,1d1,1\mathbb{R}^{p-1,1}\subset\mathbb{R}^{d-1,1} that preserve ISO(p1,1)×SO(dp)ISO(d1,1)ISO(p-1,1)\times SO(d-p)\subset ISO(d-1,1)), but rather as objects whose orientation in spacetime changes under Lorentz transformations. It may make more sense to ask whether there exists a local version of the Poincaré algebra, generated by currents rather than charges, that describes the Lorentz transformation properties of isotropic branes. For example, the brane states defined here carry representations not only of the Lorentz algebra, but of an extension of the Lorentz algebra by two-index tensor generators.777The Poincaré group is a group extension of the Lorentz group by a vector representation thereof, while the states that we have defined form a representation of a group extension of the Lorentz group by a symmetric two-index tensor representation thereof. The corresponding Lie algebra is [Πμν,Πρσ]\displaystyle[\Pi_{\mu\nu},\Pi_{\rho\sigma}] =0,\displaystyle=0, [Mμν,Πρσ]\displaystyle[M_{\mu\nu},\Pi_{\rho\sigma}] =i(ημρΠνσηνρΠμσ)+(ρσ),\displaystyle=i(\eta_{\mu\rho}\Pi_{\nu\sigma}-\eta_{\nu\rho}\Pi_{\mu\sigma})+(\rho\leftrightarrow\sigma), [Mμν,Mρσ]\displaystyle[M_{\mu\nu},M_{\rho\sigma}] =i(ημρMνσηνρMμσ)(ρσ).\displaystyle=i(\eta_{\mu\rho}M_{\nu\sigma}-\eta_{\nu\rho}M_{\mu\sigma})-(\rho\leftrightarrow\sigma). These extra generators carry two Lorentz indices regardless of the dimensionality of the brane, unlike the brane charges that occur in supersymmetry algebras [18, 19]. They take finite values in states of infinite branes, but unlike currents, they have no position dependence. Such an extended Lorentz algebra could conceivably be interpreted as a partially integrated version of the commutator algebra of the energy-momentum tensor TμνT^{\mu\nu} [20].

Alternatively, to directly assess the physical content of the abstract states |Πμν|\Pi^{\mu\nu}\rangle, one could ask whether the geometrical data contained in Πμν\Pi^{\mu\nu} admit an interpretation in terms of a maximal set of canonical variables that specify the state of an unexcited brane [21].

VI.2 Relevance to Brane Dynamics

For our discussion, it is useful to make a distinction between “kinematics” (the definition of certain representations of the Lorentz group independently of any physical theory whose states might give rise to them) and “dynamics” (the interpretation of certain quantum numbers that label these representations in terms of parameters appearing in a Lagrangian).

Our treatments of brane kinematics and brane dynamics have been largely disjoint, and strengthening their connection remains an outstanding task. On the dynamical side, we have presented a calculation of electromagnetic angular momentum for arbitrary branes related by electric-magnetic duality, generalizing earlier calculations for dyonic branes [13]. This calculation assumes a static configuration that suffices for deriving the form of the pairwise helicity. On the kinematical side, we have presented a minimal mathematical model of how (not necessarily static) flat brane sources transform under the subgroups of the Lorentz group that realize these pairwise helicities as well as higher quantum numbers. We have left open the question of whether this formalism has applications to dynamical questions about branes, and in particular, whether it provides a useful or meaningful way to describe brane states that are analogous to asymptotic states of particles.888The answer to this question may very well be “no.” For example, if the algebra of Footnote 7 indeed arises as a partially integrated current algebra, then one would not expect the resulting Πμν\Pi^{\mu\nu} to be conserved charges in the same sense as the Poincaré generators PμP^{\mu}. In particular, one would not expect these quantities to be conserved in interactions where branes collide and fluctuate. For instance, could the pairwise quantum numbers of asymptotic brane states impose constraints on the scattering of electrically and magnetically charged branes [22, 23, 24]? (Some related comments appear in [25].)

VI.3 Higher Little Groups for Particles and Branes

In closing, we reiterate that while higher little groups are kinematically allowed in d>4d>4, it would be fascinating to identify (or to rule out) any dynamical effects that might lead to the nn-particle or nn-brane little group being represented nontrivially for n>2n>2.

VII Acknowledgements

I thank Csaba Csáki, Ofri Telem, and John Terning for initial collaboration. I thank Jacques Distler, Dan Freed, Vadim Kaplunovsky, Andreas Karch, Liam McAllister, and Aaron Zimmerman for helpful discussions. Finally, I thank Steven Weinberg [26] and Jacques Distler [27] for providing the intellectual impetus to contemplate representations of the Lorentz group in higher dimensions. This work was supported in part by the NSF grant PHY-1914679.

References

Appendix A Pairwise Little Group

A.1 Definition for Particles

Wigner [1] classified the unitary irreducible (or one-particle) representations of the Poincaré group999More precisely, these are projective representations of the inhomogeneous proper orthochronous Lorentz group [28, 29]. according to their mass and their little group representation. The little group is the subgroup of the Lorentz group that leaves a particular reference momentum kk invariant. The single-particle Hilbert space is spanned by momentum eigenstates |p;σ|p;\sigma\rangle, where the single-particle quantum numbers σ\sigma fix the little group transformations of |k;σ|k;\sigma\rangle. General states are obtained from this reference state by |p;σU(Lp)|k;σ|p;\sigma\rangle\equiv U(L_{p})|k;\sigma\rangle, where U(Lp)U(L_{p}) is a unitary representation of the Lorentz boost LpL_{p} such that p=Lpkp=L_{p}k. Under a general boost Λ\Lambda, the transformation of |p;σ|p;\sigma\rangle is induced from that of |k;σ|k;\sigma\rangle as follows:

U(Λ)|p;σ=U(LΛp)U(W)|k;σ=Dσσ(W)|Λp;σ,U(\Lambda)|p;\sigma\rangle=U(L_{\Lambda p})U(W)|k;\sigma\rangle=D_{\sigma^{\prime}\sigma}(W)|\Lambda p;\sigma^{\prime}\rangle, (26)

where WLΛp1ΛLpW\equiv L_{\Lambda p}^{-1}\Lambda L_{p} is a little group transformation and DD is a unitary representation of WW.

Multiparticle representations of the Poincaré group are not limited to tensor products of single-particle representations. They may involve additional quantum numbers that characterize the Lorentz transformations of multiparticle states relative to tensor products of one-particle states [3, 4].

To see this, we specialize to 4D, where these less familiar multiparticle states find their natural home in the study of scattering amplitudes for electric and magnetic charges. For a single massive or massless particle, we choose k=(m,0,0,0)k=(m,0,0,0) or k=(E,0,0,E)k=(E,0,0,E), respectively, in which case the little group is SU(2)SU(2) or ISO(2)ISO(2). For a pair of particles with momenta (p1,p2)(p_{1},p_{2}), we can always101010Except when both momenta are null and parallel, a case that we ignore. boost to a center-of-momentum (COM) frame in which the momenta take the form

k~1,2=(m1,22+pc2,0,0,±pc),\tilde{k}_{1,2}=\left(\sqrt{m_{1,2}^{2}+p_{c}^{2}},0,0,\pm p_{c}\right), (27)

where pcp_{c} is the Lorentz-invariant COM momentum. Regardless of whether the particles are massive or massless, there exists a U(1)U(1) pairwise little group of rotations about the zz-axis that preserves these reference momenta.111111Again, we ignore non-generic configurations of momenta in which the pairwise little group is enhanced to SO(3)SO(3). To apply Wigner’s method to multiparticle states, we relate the momenta of any pair of particles to the above reference momenta via Lorentz transformations:

pi\displaystyle p_{i} =Lpiiki,\displaystyle=L^{i}_{p_{i}}k_{i}, (28)
(p1,p2)\displaystyle(p_{1},p_{2}) =(Lp1,p212k~1,Lp1,p212k~2),\displaystyle=\left(L^{12}_{p_{1},p_{2}}\tilde{k}_{1},L^{12}_{p_{1},p_{2}}\tilde{k}_{2}\right), (29)

where i=1,2i=1,2. The condition that Lp1,p212L^{12}_{p_{1},p_{2}} takes k~1p1\smash{\tilde{k}_{1}}\rightarrow p_{1} and k~2p2\smash{\tilde{k}_{2}}\rightarrow p_{2} determines it up to a U(1)U(1) rotation. We then define the single-particle and pairwise little group transformations

Wi\displaystyle W_{i} (LΛpii)1ΛLpii,\displaystyle\equiv\left(L^{i}_{\Lambda p_{i}}\right)^{-1}\Lambda L^{i}_{p_{i}}, (30)
W12\displaystyle W_{12} (LΛp1,Λp212)1ΛLp1,p212,\displaystyle\equiv\left(L^{12}_{\Lambda p_{1},\Lambda p_{2}}\right)^{-1}\Lambda L^{12}_{p_{1},p_{2}}, (31)

where the latter is a rotation: Rz(ϕ12)W12R_{z}(\phi_{12})\equiv W_{12}. Finally, since three-particle and higher little groups are trivial in 4D, a general nn-particle state takes the form

|p1,,pn;σ1,,σn;q12,,qn1,n,|p_{1},\ldots,p_{n};\sigma_{1},\ldots,\sigma_{n};q_{12},\ldots,q_{n-1,n}\rangle, (32)

where pip_{i} is the momentum of particle ii, the σi\sigma_{i} are single-particle quantum numbers (spins or helicities), and there are (n2)\binom{n}{2} quantized pairwise helicities, or U(1)U(1) charges, qijq_{ij} associated with pairs of particles i,ji,j. Under a Lorentz transformation Λ\Lambda, it transforms as

U(Λ)|p1,,pn;σ1,,σn;q12,,qn1,n=i<jeiqijϕij\displaystyle U(\Lambda)|p_{1},\ldots,p_{n};\sigma_{1},\ldots,\sigma_{n};q_{12},\ldots,q_{n-1,n}\rangle=\prod_{i<j}e^{iq_{ij}\phi_{ij}}
×i=1nDσiσii|Λp1,,Λpn;σ1,,σn;q12,,qn1,n.\displaystyle\times\prod_{i=1}^{n}D_{\sigma_{i}^{\prime}\sigma_{i}}^{i}|\Lambda p_{1},\ldots,\Lambda p_{n};\sigma_{1}^{\prime},\ldots,\sigma_{n}^{\prime};q_{12},\ldots,q_{n-1,n}\rangle. (33)

The rotation angles ϕij=ϕ(Λ,pi,pj)\phi_{ij}=\phi(\Lambda,p_{i},p_{j}) parametrize pairwise little group transformations, while the DD’s are unitary representations of one-particle little group transformations: Dσiσii=Dσiσi(Wi)D_{\sigma_{i}^{\prime}\sigma_{i}}^{i}=D_{\sigma_{i}^{\prime}\sigma_{i}}(W_{i}).121212Note that using rest- or COM-frame reference momenta is not essential to this construction. Alternatively, one could specify a set of nn reference momenta p~1,,p~n\tilde{p}_{1},\ldots,\tilde{p}_{n} at the outset, which fix all one-particle and pairwise little groups.

The pairwise helicities qijq_{ij} depend on the dynamics of the theory. If particles i,ji,j are dyons, then qijq_{ij} is simply the quantized angular momentum 14π(eigjejgi)\frac{1}{4\pi}(e_{i}g_{j}-e_{j}g_{i}) in their electromagnetic field.

A.2 Definition for Branes

We now turn to branes in any dimension. We define (p1)(p-1)-brane states that, for p>1p>1, yield representations of the Lorentz group but not of the full Poincaré group, as they are not obtained by diagonalizing the translation generators of the Poincaré algebra.

The state of a single brane at rest that extends along the first p1p-1 spatial axes is written as |Π~;σ|\tilde{\Pi};\sigma\rangle, with the reference projection tensor Π~\tilde{\Pi} given in (12). A little group rotation WW acts on this reference state as

U(W)|Π~;σ=Dσσ(W)|Π~;σ,U(W)|\tilde{\Pi};\sigma\rangle=D_{\sigma^{\prime}\sigma}(W)|\tilde{\Pi};\sigma^{\prime}\rangle, (34)

where the DD’s are matrix elements of unitary representations of WW. A general spacetime orientation Π\Pi is related to (12) by a Lorentz transformation, which we denote by L(Π)L(\Pi):

Π=LΠ~L1.\Pi=L\tilde{\Pi}L^{-1}. (35)

A generic brane state is defined by

|Π;σ=U(L)|Π~;σ.|\Pi;\sigma\rangle=U(L)|\tilde{\Pi};\sigma\rangle. (36)

Finally, a general Lorentz transformation Λ\Lambda acts as

U(Λ)|Π;σ=Dσσ(W)|ΛΠΛ1;σ,U(\Lambda)|\Pi;\sigma\rangle=D_{\sigma^{\prime}\sigma}(W)|\Lambda\Pi\Lambda^{-1};\sigma^{\prime}\rangle, (37)

where the little group rotation WW is given by

W(Λ,Π)=L1(ΛΠΛ1)ΛL(Π).W(\Lambda,\Pi)=L^{-1}(\Lambda\Pi\Lambda^{-1})\Lambda L(\Pi). (38)

The single-brane states |Π;σ|\Pi;\sigma\rangle furnish an infinite-dimensional unitary representation of the dd-dimensional Lorentz group.

We define a two-brane representation of the Lorentz group by specifying a reference state

|Π~1,Π~2;σ1,σ2;σ12|\tilde{\Pi}_{1},\tilde{\Pi}_{2};\sigma_{1},\sigma_{2};\sigma_{12}\rangle (39)

for some reference projectors Π~i\tilde{\Pi}_{i}, single-brane little group representations σi\sigma_{i}, and pairwise little group representation σ12\sigma_{12}. Given these data, we define a general state in this representation by its transformation property

U\displaystyle U (Λ)|Π1,Π2;σ1,σ2;σ12\displaystyle(\Lambda)|\Pi_{1},\Pi_{2};\sigma_{1},\sigma_{2};\sigma_{12}\rangle
=Dσ1σ1(W1)Dσ2σ2(W2)Dσ12σ12(W12)\displaystyle=D_{\sigma_{1}^{\prime}\sigma_{1}}(W_{1})D_{\sigma_{2}^{\prime}\sigma_{2}}(W_{2})D_{\sigma_{12}^{\prime}\sigma_{12}}(W_{12}) (40)
×|ΛΠ1Λ1,ΛΠ2Λ1;σ1,σ2;σ12,\displaystyle\phantom{==}\times|\Lambda\Pi_{1}\Lambda^{-1},\Lambda\Pi_{2}\Lambda^{-1};\sigma_{1}^{\prime},\sigma_{2}^{\prime};\sigma_{12}^{\prime}\rangle,

where

Wi\displaystyle W_{i} =Li(ΛΠiΛ1)1ΛLi(Πi),\displaystyle=L_{i}(\Lambda\Pi_{i}\Lambda^{-1})^{-1}\Lambda L_{i}(\Pi_{i}), (41)
W12\displaystyle W_{12} =L12(ΛΠ1Λ1,ΛΠ2Λ1)1ΛL12(Π1,Π2)\displaystyle=L_{12}(\Lambda\Pi_{1}\Lambda^{-1},\Lambda\Pi_{2}\Lambda^{-1})^{-1}\Lambda L_{12}(\Pi_{1},\Pi_{2}) (42)

are (pairwise) little group transformations and Li(Πi)L_{i}(\Pi_{i}), L12(Π1,Π2)L_{12}(\Pi_{1},\Pi_{2}) are some standard Lorentz transformations satisfying

Πi=LiΠ~iLi1,Πi=L12Π~iL121\Pi_{i}=L_{i}\tilde{\Pi}_{i}L_{i}^{-1},\qquad\Pi_{i}=L_{12}\tilde{\Pi}_{i}L_{12}^{-1} (43)

for i=1,2i=1,2.

In more detail, a two-brane reference state of the form (39) should be thought of as a tensor product

|Π~1;σ1|Π~2;σ2|(Π~1,Π~2);σ12,|\tilde{\Pi}_{1};\sigma_{1}\rangle\otimes|\tilde{\Pi}_{2};\sigma_{2}\rangle\otimes|(\tilde{\Pi}_{1},\tilde{\Pi}_{2});\sigma_{12}\rangle, (44)

where the first two factors in (44) are single-brane reference states and the last factor defines a family of pairwise states, which are auxiliary objects that transform analogously to single-brane states under the Lorentz group, but with respect to Lorentz transformations that act simultaneously on a pair of brane projectors (Π1,Π2)(\Pi_{1},\Pi_{2}). Given (44), we define a generalized two-brane state as

|Π1,Π2,(Θ1,Θ2);σ1,σ2;σ12\displaystyle|\Pi_{1},\Pi_{2},(\Theta_{1},\Theta_{2});\sigma_{1},\sigma_{2};\sigma_{12}\rangle (45)
(i=12U(Li)|Π~i;σi)U(L12)|(Π~1,Π~2);σ12,\displaystyle\equiv\left(\bigotimes_{i=1}^{2}U(L_{i})|\tilde{\Pi}_{i};\sigma_{i}\rangle\right)\otimes U(L_{12})|(\tilde{\Pi}_{1},\tilde{\Pi}_{2});\sigma_{12}\rangle,

where Li(Πi)L_{i}(\Pi_{i}) and L12(Θ1,Θ2)L_{12}(\Theta_{1},\Theta_{2}) are some standard Lorentz transformations satisfying (for i=1,2i=1,2)

Πi=LiΠ~iLi1,Θi=L12Π~iL121.\Pi_{i}=L_{i}\tilde{\Pi}_{i}L_{i}^{-1},\qquad\Theta_{i}=L_{12}\tilde{\Pi}_{i}L_{12}^{-1}. (46)

Note that while each LiL_{i} takes Π~i\tilde{\Pi}_{i} to Πi\Pi_{i} individually, L12L_{12} takes the pair (Π~1,Π~2)(\tilde{\Pi}_{1},\tilde{\Pi}_{2}) to (Θ1,Θ2)(\Theta_{1},\Theta_{2}). The definitions of the LiL_{i} and of L12L_{12} are unique up to single-brane and pairwise little group transformations, respectively. An arbitrary Lorentz transformation Λ\Lambda therefore acts on generalized two-brane states as

U(Λ)|Π1,Π2,(Θ1,Θ2);σ1,σ2;σ12\displaystyle U(\Lambda)|\Pi_{1},\Pi_{2},(\Theta_{1},\Theta_{2});\sigma_{1},\sigma_{2};\sigma_{12}\rangle
=(i=12U(Wi)|ΛΠiΛ1;σi)\displaystyle=\left(\bigotimes_{i=1}^{2}U(W_{i})|\Lambda\Pi_{i}\Lambda^{-1};\sigma_{i}\rangle\right) (47)
U(W12)|(ΛΘ1Λ1,ΛΘ2Λ1);σ12\displaystyle\phantom{==}\otimes U(W_{12})|(\Lambda\Theta_{1}\Lambda^{-1},\Lambda\Theta_{2}\Lambda^{-1});\sigma_{12}\rangle
=(i=12Dσiσi(Wi)|ΛΠiΛ1;σi)\displaystyle=\left(\bigotimes_{i=1}^{2}D_{\sigma_{i}^{\prime}\sigma_{i}}(W_{i})|\Lambda\Pi_{i}\Lambda^{-1};\sigma_{i}^{\prime}\rangle\right) (48)
Dσ12σ12(W12)|(ΛΘ1Λ1,ΛΘ2Λ1);σ12,\displaystyle\phantom{==}\otimes D_{\sigma_{12}^{\prime}\sigma_{12}}(W_{12})|(\Lambda\Theta_{1}\Lambda^{-1},\Lambda\Theta_{2}\Lambda^{-1});\sigma_{12}^{\prime}\rangle,

where

Wi\displaystyle W_{i} =Li(ΛΠiΛ1)1ΛLi(Πi),\displaystyle=L_{i}(\Lambda\Pi_{i}\Lambda^{-1})^{-1}\Lambda L_{i}(\Pi_{i}), (49)
W12\displaystyle W_{12} =L12(ΛΘ1Λ1,ΛΘ2Λ1)1ΛL12(Θ1,Θ2)\displaystyle=L_{12}(\Lambda\Theta_{1}\Lambda^{-1},\Lambda\Theta_{2}\Lambda^{-1})^{-1}\Lambda L_{12}(\Theta_{1},\Theta_{2}) (50)

are (pairwise) little group transformations. Finally, we define a physical two-brane state as one satisfying Θi=Πi\Theta_{i}=\Pi_{i} for i=1,2i=1,2:

|Π1,Π2;σ1,σ2;σ12\displaystyle|\Pi_{1},\Pi_{2};\sigma_{1},\sigma_{2};\sigma_{12}\rangle
|Π1,Π2,(Π1,Π2);σ1,σ2;σ12.\displaystyle\equiv|\Pi_{1},\Pi_{2},(\Pi_{1},\Pi_{2});\sigma_{1},\sigma_{2};\sigma_{12}\rangle. (51)

Such states transform as written in (40), with WiW_{i} and W12W_{12} as in (49)–(50) but with Θi=Πi\Theta_{i}=\Pi_{i}. The collection of generalized two-brane states defines a unitary representation of the Lorentz group, and that of physical two-brane states defines a subrepresentation.

A.3 Hilbert Space Interpretation

We now make some overdue comments on the Hilbert space structure of the representations that we have defined, and in particular, on the inner products with respect to which they are unitary.

Taking a broad perspective, for a classical system with abstract configuration space XX, we are often interested in the corresponding quantum system with Hilbert space L2(X)L^{2}(X).131313More generally, we may wish to consider the space of L2L^{2} sections of a complex vector bundle over XX, where the fiber VV parametrizes internal degrees of freedom. For a trivial bundle, the Hilbert space is simply L2(X)VL^{2}(X)\otimes V. We can write the inner product on L2(X)L^{2}(X) as

f1|f2=𝑑xf1(x)f2(x),\langle f_{1}|f_{2}\rangle=\int dx\,f_{1}(x)^{\ast}f_{2}(x), (52)

where f1,f2:Xf_{1},f_{2}:X\to\mathbb{C} and xXx\in X. If the group GG acts on XX and the associated measure dxdx is GG-invariant, then L2(X)L^{2}(X) carries a unitary representation UU of GG induced by the action of GG on XX. For gGg\in G, we have

U(g)|f=|fgU(g)|f\rangle=|f^{g}\rangle (53)

where fg(x)=f(g1x)f^{g}(x)=f(g^{-1}x). In terms of “position eigenstates” that satisfy x|x=δ(xx)\langle x|x^{\prime}\rangle=\delta(x-x^{\prime}) with respect to the measure dxdx (and are therefore not elements of L2(X)L^{2}(X)), this action is equivalent to

U(g)|x=|gx,U(g)|x\rangle=|gx\rangle, (54)

where |f=𝑑xf(x)|x|f\rangle=\int dx\,f(x)|x\rangle.

As a special case of the above, the carrier space of a unitary irreducible representation of the Poincaré group corresponding to a spinless particle of mass m>0m>0 is the space of L2L^{2} functions on the positive mass shell Sm+S_{m}^{+}, i.e., the orbit of the fiducial timelike dd-vector kμ=(m,0,,0)k^{\mu}=(m,0,\ldots,0) in d1,1\mathbb{R}^{d-1,1} under proper orthochronous Lorentz transformations. Including spin, the carrier space is the space of L2L^{2} sections of a vector bundle with fiber VV over Sm+S_{m}^{+}, where VV is the carrier space of a unitary irreducible representation of the little group SO(d1)SO(d-1). Sm+S_{m}^{+} inherits a Lorentz-invariant integration measure from the flat measure on d1,1\mathbb{R}^{d-1,1} via the embedding Sm+d1,1S_{m}^{+}\subset\mathbb{R}^{d-1,1}, which we use to define an inner product:

Ψ|Φ=ddpδ(p2m2)θ(p0)σΨp,σΦp,σ.\langle\Psi|\Phi\rangle=\int d^{d}p\,\delta(p^{2}-m^{2})\theta(p^{0})\sum_{\sigma}\Psi_{p,\sigma}^{\ast}\Phi_{p,\sigma}. (55)

Lorentz transformations act on states as U(Λ)|Ψ=|ΨΛU(\Lambda)|\Psi\rangle=|\Psi^{\Lambda}\rangle with

Ψp,σΛ=σDσσ(W1(Λ1,p))ΨΛ1p,σ,\Psi_{p,\sigma}^{\Lambda}=\sum_{\sigma^{\prime}}D_{\sigma\sigma^{\prime}}(W^{-1}(\Lambda^{-1},p))\Psi_{\Lambda^{-1}p,\sigma^{\prime}}, (56)

where WW is a little group transformation with respect to kμk^{\mu} and we have used W(Λ,Λ1p)=W1(Λ1,p)W(\Lambda,\Lambda^{-1}p)=W^{-1}(\Lambda^{-1},p). We write the wavefunctions Ψp,σ\Psi_{p,\sigma} in terms of dd-momenta pμp^{\mu} rather than (d1)(d-1)-momenta p\vec{p}, with the understanding that the dd-momenta are constrained to lie on Sm+S_{m}^{+}.

Similarly, ignoring spin, the carrier space of a single-brane representation of the Lorentz group as defined in this paper is the space of L2L^{2} functions on the Lorentz orbit of a reference projection tensor within the space of all two-index Lorentz tensors. Such an orbit inherits a Lorentz-invariant integration measure from the ambient space, and this measure allows one to define an inner product on the corresponding Hilbert space.

More explicitly, Lorentz transformations Λ\Lambda act as ΠΛΠΛ1\Pi\to\Lambda\Pi\Lambda^{-1} on the set of all two-index Lorentz tensors Π\Pi (thought of as matrices), which is isomorphic to d2\mathbb{R}^{d^{2}} as a vector space. This space has a natural Lorentz-invariant flat measure, which we denote by dΠd\Pi.141414Whereas ddpd^{d}p transforms with a factor of |detΛ||{\det\Lambda}|, dΠd\Pi transforms with a factor of |detΛ|2d|{\det\Lambda}|^{2d}. This is easily seen by considering the vectorization of Π\Pi and the behavior of the determinant under Kronecker products. Let OpO_{p} denote the orbit of the fiducial rank-pp projection tensor

Π~=diag(1,,1p,0,,0dp)\tilde{\Pi}=\operatorname{diag}(\underbrace{1,\ldots,1}_{p},\underbrace{0,\ldots,0}_{d-p}) (57)

under the action of the Lorentz group, and let dΠpd\Pi_{p} denote the measure on OpO_{p} that is induced by the flat measure dΠd\Pi on the space of all Π\Pi. Then we obtain an infinite-dimensional unitary representation of the Lorentz group on L2(Op)L^{2}(O_{p}) given (in terms of “position eigenstates”) by U(Λ)|Π=|ΛΠΛ1U(\Lambda)|\Pi\rangle=|\Lambda\Pi\Lambda^{-1}\rangle, or equivalently (in terms of normalizable states) by U(Λ)|Ψ=|ΨΛU(\Lambda)|\Psi\rangle=|\Psi^{\Lambda}\rangle with ΨΠΛ=ΨΛ1ΠΛ\Psi_{\Pi}^{\Lambda}=\Psi_{\Lambda^{-1}\Pi\Lambda}. We have written the wavefunctions as functions of Π\Pi, with Π\Pi implicitly constrained to OpO_{p}. One can also add spin indices, using the fact that (57) has a nontrivial stabilizer within the Lorentz group, in which case the carrier space is enlarged and the inner product takes the form

Ψ|Φ=𝑑ΠpσΨΠ,σΦΠ,σ.\langle\Psi|\Phi\rangle=\int d\Pi_{p}\sum_{\sigma}\Psi_{\Pi,\sigma}^{\ast}\Phi_{\Pi,\sigma}. (58)

In general, we would expect such a representation to be reducible.

Finally, while massive particle representations of the Poincaré group and these unitary representations of the Lorentz group share the common feature of being induced by unitary representations of compact subgroups of the Lorentz group, these brane representations admit no natural action of the translation generators of the Poincaré group. Indeed, the total energies and momenta of infinite branes do not provide meaningful quantum numbers for labeling states, while their finite energy and momentum densities transform non-tensorially (unlike the generators of the Poincaré algebra).

Appendix B Mathematical Conventions

We denote the dimension of spacetime by dd. We assume that spacetime is topologically trivial and boundaryless. For our subsequent calculations (unlike in the main text), we work in “mostly plus” Lorentzian signature. We define the Hodge star operator by

(dxμ1dxμp)\displaystyle\ast(dx^{\mu_{1}}\wedge\cdots\wedge dx^{\mu_{p}}) (59)
=1(dp)!ϵμ1μpdν1νdpxν1dxνdp.\displaystyle=\frac{1}{(d-p)!}\epsilon^{\mu_{1}\cdots\mu_{p}}{}_{\nu_{1}\cdots\nu_{d-p}}dx^{\nu_{1}}\wedge\cdots\wedge dx^{\nu_{d-p}}.

In particular, dxμ1dxμd=ϵμ1μd1dx^{\mu_{1}}\wedge\cdots\wedge dx^{\mu_{d}}=-\epsilon^{\mu_{1}\cdots\mu_{d}}\ast 1 with 1=ddxg\ast 1=d^{d}x\sqrt{-g}. With this convention, for a pp-form

α=1p!αμ1μpdxμ1dxμp,\alpha=\frac{1}{p!}\alpha_{\mu_{1}\cdots\mu_{p}}dx^{\mu_{1}}\wedge\cdots\wedge dx^{\mu_{p}}, (60)

the components of its dual are given by

(α)μp+1μd=1p!ϵμ1μdαμ1μp.(\ast\alpha)_{\mu_{p+1}\cdots\mu_{d}}=\frac{1}{p!}\epsilon_{\mu_{1}\cdots\mu_{d}}\alpha^{\mu_{1}\cdots\mu_{p}}. (61)

For any two pp-forms α\alpha and β\beta, we have

αβ=βα=1p!αμ1μpβμ1μp1,\alpha\wedge\ast\beta=\beta\wedge\ast\alpha=\frac{1}{p!}\alpha_{\mu_{1}\cdots\mu_{p}}\beta^{\mu_{1}\cdots\mu_{p}}\ast 1, (62)

as ϵμ1μpρ1ρdpϵν1νpρ1ρdp=p!(dp)!δμ1μpν1νp\epsilon_{\mu_{1}\cdots\mu_{p}\rho_{1}\cdots\rho_{d-p}}\epsilon^{\nu_{1}\cdots\nu_{p}\rho_{1}\cdots\rho_{d-p}}=-p!(d-p)!\delta_{\mu_{1}\cdots\mu_{p}}^{\nu_{1}\cdots\nu_{p}}.

Poincaré duality associates (cohomology classes of) pp-forms with (homology classes of) codimension-pp submanifolds [30]. Operationally, given a pp-dimensional submanifold MM, its Poincaré dual M^\smash{\widehat{M}} has delta-function support on MM and is defined as satisfying

MA=AM^\int_{M}A=\int A\wedge\widehat{M} (63)

for all pp-forms AA, where the integral on the right is taken over all of spacetime. For a pp-manifold MM, we have

M^+(1)dpdM^=0.\widehat{\partial M}+(-1)^{d-p}d\widehat{M}=0. (64)

Poincaré duality is an involution, unlike Hodge duality.

If MM and NN are submanifolds with dimM+dimN=d\dim M+\dim N=d that intersect transversally at a finite number of points, then the (signed) intersection number is an integer given by

I(M,N)=M^N^,I(M,N)=\int\widehat{M}\wedge\widehat{N}, (65)

reflecting the fact that cup product in cohomology is Poincaré-dual to intersection pairing in homology. It satisfies I(M,N)=(1)dimMdimNI(N,M)I(M,N)=(-1)^{\dim M\dim N}I(N,M).

If CC and DD are two non-intersecting trivial cycles in spacetime with dimC+dimD=d1\dim C+\dim D=d-1, where C=CC=\partial C^{\prime} and D=DD=\partial D^{\prime}, then we define the linking number of CC and DD as the intersection number of CC and the “Stokes surface” DD^{\prime}:

L(C,D)=C^D^.L(C,D)=\int\widehat{C}\wedge\widehat{D^{\prime}}. (66)

It satisfies L(C,D)=(1)(dimC+1)(dimD+1)L(D,C)L(C,D)=(-1)^{(\dim C+1)(\dim D+1)}L(D,C) as a consequence of (64).

Appendix C Pairwise Helicity From Angular Momentum

In this appendix, we elaborate on the calculation of the angular momentum (8).

We first recall the argument for the Dirac(-Schwinger-Zwanziger) quantization condition. Consider an electric brane with pp-dimensional worldvolume VV and a spacelike submanifold Σdp\Sigma^{d-p}. The quantity

Q=ΣdpJ=(Σ)dp1FQ=\int_{\Sigma^{d-p}}\ast J=\int_{(\partial\Sigma)^{d-p-1}}\ast F (67)

defines a conserved charge in the sense that it depends only on the linking homology class of (Σ)dp1(\partial\Sigma)^{d-p-1} with respect to VV. Here, we have applied Stokes’ theorem to the electric brane current JJ: dF=Jd\ast F=\ast J. Therefore, the charge densities of an electric (p1)(p-1)-brane and a magnetic (dp3)(d-p-3)-brane, as measured by spheres Sdp1S^{d-p-1} and Sp+1S^{p+1} that link them in space, are given by

e=Sdp1F,g=Sp+1F,e=\int_{S^{d-p-1}}\ast F,\qquad g=\int_{S^{p+1}}F, (68)

where we assume that 1pd31\leq p\leq d-3. More generally, (67) evaluates to Q=eI(V,Σ)=eL(V,Σ)Q=eI(V,\Sigma)=eL(V,\partial\Sigma), where II and LL denote intersection and linking numbers, respectively.

Now note that in a background with nonzero magnetic charge, AA cannot be globally defined, since otherwise, we would have F=dAF=dA and g=0g=0. Hence the action of an electric brane is ambiguous in such a background. If an electric brane sweeps out a closed Wilson surface along Σp\Sigma^{p} in the field of a magnetic brane, then the amplitude for this process is given by the exponentiated action

exp(ieΣpA)=exp(ieMp+1F),\exp\left(ie\int_{\Sigma^{p}}A\right)=\exp\left(ie\int_{M^{p+1}}F\right), (69)

where Mp+1M^{p+1} is any (p+1)(p+1)-manifold with boundary Σp\Sigma^{p}. The expression on the right is unambiguous because FF is globally defined, but it depends nonlocally on Σp\Sigma^{p}. To preserve locality, we demand that it be independent of the choice of Mp+1M^{p+1}, or equivalently, that

exp(ieCp+1F)=exp(iegL(V,Cp+1))=1\exp\left(ie\int_{C^{p+1}}F\right)=\exp(iegL(V^{\prime},C^{p+1}))=1 (70)

for any closed Cp+1C^{p+1}, where VV^{\prime} is the worldvolume of the magnetic brane. In particular, choosing Cp+1C^{p+1} to have unit linking number with the magnetic brane in space yields exp(ieg)=1\exp(ieg)=1, or the Dirac quantization condition

eg2π.eg\in 2\pi\mathbb{Z}. (71)

This condition can be derived in many alternative ways, such as via Dirac branes [8, 9] or by considering the transition functions between gauge patches [10] (generalizing the Wu-Yang construction [11]).

While (71) holds for generic dd and pp, in the special case that d=2(p+1)d=2(p+1), brane sources can carry both electric and magnetic charge. In this case, given any two dyonic (p1)(p-1)-branes of charge densities (e1,g1)(e_{1},g_{1}) and (e2,g2)(e_{2},g_{2}), we instead have the condition

e1g2+(1)pe2g12π,e_{1}g_{2}+(-1)^{p}e_{2}g_{1}\in 2\pi\mathbb{Z}, (72)

which we refer to as the DSZ quantization condition. The derivation of (72) is more subtle than that of (71) [12, 13, 14, 15, 16]. The presence of the sign (1)p(-1)^{p} is closely related to the (non)existence of theta terms and the structure of the electric-magnetic duality group in the corresponding dd.151515The theta angle modifies the electric charges of dyons via the Witten effect [31] while preserving the DSZ pairing. Only the latter enters into pairwise helicities, not the individual charges. But scattering processes of dyonic particles or branes are sensitive to the charges themselves and therefore also to the theta angle. See Appendix D for details.

The conditions (71) and (72) motivate the definition (5) for the pairwise helicity of two branes ii and jj related by electric-magnetic duality. We have shown that any two such branes admit a kinematic configuration in which their pairwise little group is SO(2)SO(2). We now show that (5) arises dynamically as the SO(2)SO(2) charge. Specifically, we show that the pairwise helicity qijq_{ij} is the angular momentum in the electromagnetic field sourced by dual branes. This angular momentum shows up as a phase in the pairwise little group transformation of asymptotic brane states. As a byproduct of our analysis, we also (re)derive the conditions (71) and (72) in a uniform way.

C.1 Pairwise Helicity for Particles

Our derivation proceeds by analogy to the case d=4d=4, p=1p=1, which we now review. In 4D, particles (0-branes) can carry both electric and magnetic charge. The DSZ quantization condition (72) becomes

q12e1g2e2g14π12.q_{12}\equiv\frac{e_{1}g_{2}-e_{2}g_{1}}{4\pi}\in\tfrac{1}{2}\mathbb{Z}. (73)

We first recall how q12q_{12} appears as the angular momentum of a charge-monopole field, setting (e1,g1)=(0,g)(e_{1},g_{1})=(0,g) and (e2,g2)=(e,0)(e_{2},g_{2})=(e,0). For ease of generalization, we work in terms of differential forms.

We place the monopole at the origin in x=(x1,x2,x3)\vec{x}=(x_{1},x_{2},x_{3}) and the electric charge at R=(R1,R2,R3)\smash{\vec{R}}=(R_{1},R_{2},R_{3}). We normalize their field strengths and potentials as follows:

E=e(xR)4π|xR|3,B=gx4π|x|3,\vec{E}=\frac{e(\vec{x}-\vec{R})}{4\pi|\vec{x}-\vec{R}|^{3}},\qquad\vec{B}=\frac{g\vec{x}}{4\pi|\vec{x}|^{3}}, (74)

where EiF0iE^{i}\equiv F^{0i} and Bi(F)0i=12ϵijkFjkB^{i}\equiv-(\ast F)^{0i}=\frac{1}{2}\epsilon^{ijk}F_{jk}. In terms of the one-forms E=EidxiE=E_{i}\,dx^{i} and B=BidxiB=B_{i}\,dx^{i}, the angular momentum pseudovector (Ji=12ϵijkJjkJ^{i}=\frac{1}{2}\epsilon^{ijk}J_{jk}) becomes

Jij=(xidxjxjdxi)EB.J_{ij}=\int(x_{i}\,dx_{j}-x_{j}\,dx_{i})\wedge E\wedge B. (75)

Substituting B=g4π|x|3xkdxkB=\frac{g}{4\pi|\vec{x}|^{3}}x_{k}\,dx^{k} into (75) gives

Jij=g4πE|x|3(|x|2dxidxjϵijkxk(xdx)),J_{ij}=\frac{g}{4\pi}\int\frac{E}{|\vec{x}|^{3}}\wedge(|\vec{x}|^{2}\,dx_{i}\wedge dx_{j}-\epsilon_{ijk}x^{k}\ast^{\prime}(x_{\ell}\,dx^{\ell})), (76)

where we use \ast^{\prime} to denote the Hodge star with respect to the (Euclidean) spatial dimensions. This simplifies to

Jij=g4πEd(ϵijkxk|x|)=g4πϵijkxk|x|d(E),J_{ij}=\frac{g}{4\pi}\int E\wedge\ast^{\prime}d\left(\frac{\epsilon_{ijk}x^{k}}{|\vec{x}|}\right)=-\frac{g}{4\pi}\int\frac{\epsilon_{ijk}x^{k}}{|\vec{x}|}\,d(\ast^{\prime}E), (77)

where we have used symmetry of the inner product on differential forms and integrated by parts. Using d(E)=(E)d3x=eδ3(xR)d3xd(\ast^{\prime}E)=(\nabla\cdot\vec{E})\,d^{3}x=e\delta^{3}(\vec{x}-\vec{R})\,d^{3}x then gives

Jij=eg4πϵijkR^kJ_{ij}=-\frac{eg}{4\pi}\epsilon_{ijk}\hat{R}^{k} (78)

where R^R/|R|\hat{R}\equiv\vec{R}/|\vec{R}|, which is independent of the separation |R|\smash{|\vec{R}|}. The Dirac quantization condition now follows from the semiclassical expectation that the angular momentum (78) should be half-integrally quantized.161616To repeat this derivation in non-covariant language [32], substituting B\vec{B} into J=d3x[x×(E×B)]\vec{J}=\int d^{3}x\,[\vec{x}\times(\vec{E}\times\vec{B})] gives J=g4πd3x(E)x^=g4πd3x(E)x^=eg4πR^,\vec{J}=\frac{g}{4\pi}\int d^{3}x\,(\vec{E}\cdot\nabla)\hat{x}=-\frac{g}{4\pi}\int d^{3}x\,(\nabla\cdot\vec{E})\hat{x}=-\frac{eg}{4\pi}\hat{R}, (79) where x^x/|x|\hat{x}\equiv\vec{x}/|\vec{x}| and we have dropped a surface term since the fields vanish at infinity.

C.2 Pairwise Helicity for Branes

We now calculate the angular momentum in the electromagnetic field of two mutually nonlocal branes in d1,1\mathbb{R}^{d-1,1}, obtaining precisely the pairwise helicities (5).

To start, the gauge field carries a momentum density

Pi=T0=i1p!F0Fii1ip,i1ipP_{i}=T^{0}{}_{i}=\frac{1}{p!}F^{0}{}_{i_{1}\cdots i_{p}}F_{i}{}^{i_{1}\cdots i_{p}}, (80)

where Tμν=2gμν+gμνT_{\mu\nu}=-2\frac{\partial\mathcal{L}}{\partial g^{\mu\nu}}+g_{\mu\nu}\mathcal{L} is the stress tensor of the pp-form Maxwell theory with Lagrangian density

=12(p+1)!Fμ1μp+1Fμ1μp+1.\mathcal{L}=-\frac{1}{2(p+1)!}F_{\mu_{1}\cdots\mu_{p+1}}F^{\mu_{1}\cdots\mu_{p+1}}. (81)

The corresponding angular momentum density is

𝒥ij=xiPjxjPi=xiT0jxjT0,i\mathcal{J}_{ij}=x_{i}P_{j}-x_{j}P_{i}=x_{i}T^{0}{}_{j}-x_{j}T^{0}{}_{i}, (82)

which is a rank-two antisymmetric tensor of SO(d1)SO(d-1). The angular momentum Jij=dd1x𝒥ijJ_{ij}=\int d^{d-1}x\,\mathcal{J}_{ij} is dimensionless in natural units.

We write the electric and magnetic fields as (spatial) forms EE and BB, with components Ei1ip=F0i1ipE^{i_{1}\cdots i_{p}}=F^{0i_{1}\cdots i_{p}} and

Bi1idp2\displaystyle B^{i_{1}\cdots i_{d-p-2}} =(1)p(F)0i1idp2\displaystyle=(-1)^{p}(\ast F)^{0i_{1}\cdots i_{d-p-2}} (83)
=1(p+1)!ϵj1jp+1i1idp2Fj1jp+1.\displaystyle=\frac{1}{(p+1)!}\epsilon^{j_{1}\cdots j_{p+1}i_{1}\cdots i_{d-p-2}}F_{j_{1}\cdots j_{p+1}}. (84)

Note that ϵ0i1id1=ϵi1id1\epsilon_{0i_{1}\cdots i_{d-1}}=\epsilon_{i_{1}\cdots i_{d-1}} and ϵ0i1id1=ϵi1id1\epsilon^{0i_{1}\cdots i_{d-1}}=-\epsilon^{i_{1}\cdots i_{d-1}}. We denote the spatial Hodge star by \ast^{\prime}, so that

dxi1dxid1=ϵi1id11.dx^{i_{1}}\wedge\cdots\wedge dx^{i_{d-1}}=\epsilon^{i_{1}\cdots i_{d-1}}\ast^{\prime}1. (85)

We then compute that

(xidxjxjdxi)EB=(xiT0jxjT0)i1.(x_{i}\,dx_{j}-x_{j}\,dx_{i})\wedge E\wedge B=(x_{i}T^{0}{}_{j}-x_{j}T^{0}{}_{i})\ast^{\prime}1. (86)

It follows from (86) that the formula (75) for the angular momentum corresponding to SO(2)SO(2) rotations in the i,ji,j directions holds in arbitrary dd.

We restrict our attention to the spatial directions d1\mathbb{R}^{d-1}. We take the electric (p1)(p-1)-brane to fill the directions

R3×p1×0dp3\vec{R}_{3}\times\mathbb{R}^{p-1}\times\vec{0}_{d-p-3} (87)

and the magnetic (dp3)(d-p-3)-brane to fill the directions

03×0p1×dp3.\vec{0}_{3}\times\vec{0}_{p-1}\times\mathbb{R}^{d-p-3}. (88)

We label the coordinates as in (4), and we denote by

Ωn=2πn+12Γ(n+12)\Omega_{n}=\frac{2\pi^{\frac{n+1}{2}}}{\Gamma(\frac{n+1}{2})} (89)

the area of the unit nn-sphere. The electric and magnetic potentials are

ϕE\displaystyle\phi_{E} =e(dp2)Ωdp1((xR)2+z2)(dp2)/2,\displaystyle=-\frac{e}{(d-p-2)\Omega_{d-p-1}((\vec{x}-\vec{R})^{2}+\vec{z}^{2})^{(d-p-2)/2}},
ϕB\displaystyle\phi_{B} =gpΩp+1(x2+y2)p/2.\displaystyle=-\frac{g}{p\Omega_{p+1}(\vec{x}^{2}+\vec{y}^{2})^{p/2}}. (90)

The corresponding electric and magnetic fields are

E=ex^eΩdp1|xe|dp1,B=gx^gΩp+1|xg|p+1,\vec{E}=\frac{e\hat{x}_{e}}{\Omega_{d-p-1}|\vec{x}_{e}|^{d-p-1}},\qquad\vec{B}=\frac{g\hat{x}_{g}}{\Omega_{p+1}|\vec{x}_{g}|^{p+1}}, (91)

where we have set

xe\displaystyle\vec{x}_{e} (xR)×0p1×z,\displaystyle\equiv(\vec{x}-\vec{R})\times\vec{0}_{p-1}\times\vec{z}, (92)
xg\displaystyle\vec{x}_{g} x×y×0dp3.\displaystyle\equiv\vec{x}\times\vec{y}\times\vec{0}_{d-p-3}. (93)

The constants are important for obtaining the precise normalization factors in the Dirac quantization condition.

Accounting for the dimensions of the branes, the electric potential is a spatial (p1)(p-1)-form

ΦE=ϕEdy1dyp1,\Phi_{E}=\phi_{E}\,dy_{1}\wedge\cdots\wedge dy_{p-1}, (94)

and the magnetic potential is a spatial (dp3)(d-p-3)-form

ΦB=ϕBdz1dzdp3.\Phi_{B}=\phi_{B}\,dz_{1}\wedge\cdots\wedge dz_{d-p-3}. (95)

We then have

E\displaystyle E =dΦE=Ei(dxe)idy1dyp1,\displaystyle=d\Phi_{E}=E^{i}\,(dx_{e})_{i}\wedge dy_{1}\wedge\cdots\wedge dy_{p-1}, (96)
B\displaystyle B =dΦB=Bi(dxg)idz1dzdp3.\displaystyle=d\Phi_{B}=B^{i}\,(dx_{g})_{i}\wedge dz_{1}\wedge\cdots\wedge dz_{d-p-3}. (97)

Using the identity

(r^rn)=Ωnδn+1(r),rn+1,\nabla\cdot\left(\frac{\hat{r}}{r^{n}}\right)=\Omega_{n}\delta^{n+1}(\vec{r}),\qquad\vec{r}\in\mathbb{R}^{n+1}, (98)

we see that

dE\displaystyle d\ast^{\prime}\!E =(1)p1eδdp(xe)(dy1dyp1),\displaystyle=(-1)^{p-1}e\delta^{d-p}(\vec{x}_{e})\ast^{\prime}\!(dy_{1}\wedge\cdots\wedge dy_{p-1}), (99)
dB\displaystyle d\ast^{\prime}\!B =(1)dp3gδp+2(xg)(dz1dzdp3).\displaystyle=(-1)^{d-p-3}g\delta^{p+2}(\vec{x}_{g})\ast^{\prime}\!(dz_{1}\wedge\cdots\wedge dz_{d-p-3}). (100)

Hence dEd\ast^{\prime}E and dBd\ast^{\prime}B are proportional to the Poincaré duals of the electric and magnetic branes, respectively.

In the angular momentum formula (75), we substitute

E\displaystyle E =iϕEdxidy1dyp1+,\displaystyle=\partial^{i}\phi_{E}\,dx_{i}\wedge dy_{1}\wedge\cdots\wedge dy_{p-1}+\cdots, (101)
B\displaystyle B =iϕBdxidz1dzdp3+,\displaystyle=\partial^{i}\phi_{B}\,dx_{i}\wedge dz_{1}\wedge\cdots\wedge dz_{d-p-3}+\cdots, (102)

where the omitted terms in EE involve dzidz_{i} and the omitted terms in BB involve dyidy_{i}. For simplicity, we take the separation R=(0,0,R3)\vec{R}=(0,0,R_{3}) to point in the x3x_{3}-direction. Then we have

J12\displaystyle J_{12} =(1)pdd1x((x11+x22)ϕE3ϕB\displaystyle=(-1)^{p}\int d^{d-1}x\,((x_{1}\partial_{1}+x_{2}\partial_{2})\phi_{E}\partial_{3}\phi_{B}
3ϕE(x11+x22)ϕB)\displaystyle\hskip 85.35826pt-\partial_{3}\phi_{E}(x_{1}\partial_{1}+x_{2}\partial_{2})\phi_{B}) (103)
=(1)pegΩdp1Ωp+1dd1x\displaystyle=\frac{(-1)^{p}eg}{\Omega_{d-p-1}\Omega_{p+1}}\int d^{d-1}x
R3(x12+x22)((xR)2+z2)(dp)/2(x2+y2)(p+2)/2.\displaystyle\hskip 28.45274pt\frac{R_{3}(x_{1}^{2}+x_{2}^{2})}{((\vec{x}-\vec{R})^{2}+\vec{z}^{2})^{(d-p)/2}(\vec{x}^{2}+\vec{y}^{2})^{(p+2)/2}}. (104)

By dimensional analysis, this integral is independent of R3R_{3}. Now we evaluate the integrals over y\vec{y} and z\vec{z} recursively using

dx(a+x2)p/2=1a(p1)/2Ωp1Ωp2,\int_{-\infty}^{\infty}\frac{dx}{(a+x^{2})^{p/2}}=\frac{1}{a^{(p-1)/2}}\frac{\Omega_{p-1}}{\Omega_{p-2}}, (105)

which holds for p>1p>1 and a>0a>0 (the aa-dependence follows from rescaling xx). Note that it is valid to apply (105) inside the integral because a>0a>0 almost everywhere. This gives

J12=(1)peg16π2d3xR3(x12+x22)((xR)2)3/2(x2)3/2,J_{12}=\frac{(-1)^{p}eg}{16\pi^{2}}\int d^{3}x\,\frac{R_{3}(x_{1}^{2}+x_{2}^{2})}{((\vec{x}-\vec{R})^{2})^{3/2}(\vec{x}^{2})^{3/2}}, (106)

which can be evaluated just as in the familiar case of d=4d=4, p=1p=1:

J12\displaystyle J_{12} =(1)peg16π2d3xxR((xR)2)3/2(x3|x|)\displaystyle=-\frac{(-1)^{p}eg}{16\pi^{2}}\int d^{3}x\,\frac{\vec{x}-\vec{R}}{((\vec{x}-\vec{R})^{2})^{3/2}}\cdot\nabla\left(\frac{x_{3}}{|\vec{x}|}\right)
=(1)peg4πR^3.\displaystyle=\frac{(-1)^{p}eg}{4\pi}\hat{R}_{3}. (107)

The result (107) encompasses our earlier result (78) for d=4d=4, p=1p=1, as well as that derived in [13] for the dyonic case d=2(p+1)d=2(p+1).171717The proof in [13] proceeds by integrating the second term of (103) by parts to transfer the x3x_{3}-derivative to ϕB\phi_{B} and the x1,2x_{1,2}-derivatives to ϕE\phi_{E}, leading to partial cancellation with the first term: J12=2(1)pdd1xϕE3ϕB.J_{12}=-2(-1)^{p}\int d^{d-1}x\,\phi_{E}\partial_{3}\phi_{B}. (108) After recursively applying (105), we obtain J12=(1)peg8π2d3x1((xR)2)1/231(x2)1/2,J_{12}=-\frac{(-1)^{p}eg}{8\pi^{2}}\int d^{3}x\,\frac{1}{((\vec{x}-\vec{R})^{2})^{1/2}}\partial_{3}\frac{1}{(\vec{x}^{2})^{1/2}}, (109) which is equivalent to our expression but less amenable to exact evaluation. It matches (5) exactly, where the first term of (5) in the case that d=2(p+1)d=2(p+1) follows from the exchange (anti)symmetry of the pp-forms EE and BB in the formula (75), so that qij=(1)pqjiq_{ij}=(-1)^{p}q_{ji}. As noted in [13], this argument automatically produces the sign (1)p(-1)^{p} in the DSZ quantization condition (72).

To summarize, we have identified the pairwise helicities qijq_{ij} defined in (5) with the angular momentum in the electromagnetic field sourced by dual branes in any d4d\geq 4. Hence the little group transformation of multi-brane states involves an extra phase of eiijqijϕije^{i\sum_{ij}q_{ij}\phi_{ij}}, just as for electric-magnetic multiparticle states in 4D [4].

Appendix D pp-Form Electrodynamics

We summarize here some additional useful facts about pp-form electrodynamics. Note that locality constrains pp-form gauge theories with p>1p>1 to be abelian [33].181818In modern language, these theories arise from gauging (p1)(p-1)-form global symmetries [34], which are abelian for p1>0p-1>0 because charge operators of codimension >1>1 have no canonical ordering.

D.1 Classical Action

Given a brane with pp-dimensional worldvolume VV and charge density μ\mu, we define its pp-form brane current JJ by

J=μV^,\ast J=\mu\widehat{V}, (110)

with V^\widehat{V} being the Poincaré dual of VV in spacetime. In components, we have

Jμ1μp(x)=μVδd(xX(σ))𝑑X(σ)μ1dX(σ)μp,J^{\mu_{1}\cdots\mu_{p}}(x)=\mu\int_{V}\delta^{d}(x-X(\sigma))\,dX(\sigma)^{\mu_{1}}\wedge\cdots\wedge dX(\sigma)^{\mu_{p}}, (111)

where xμx^{\mu} are spacetime coordinates, XμX^{\mu} are embedding coordinates, and σ1,,σp\sigma^{1},\ldots,\sigma^{p} are coordinates internal to VV.

The action of pp-form electrodynamics is obtained by coupling the Maxwell action of a pp-form gauge field AA to electric and magnetic sources. To write an action that yields the desired field equations

dF=Jm,dF=Je,dF=\ast J_{m},\qquad d\ast F=\ast J_{e}, (112)

we use a single singular gauge potential defined on all of spacetime rather than multiple regular potentials defined on patches of spacetime [9, 10]. For simplicity, we consider a single electric (p1)(p-1)-brane and a single magnetic (dp3)(d-p-3)-brane (the generalization to multiple branes is obvious). We denote their brane currents by

Je=eVe^,Jm=gVm^,\ast J_{e}=e\widehat{V_{e}},\qquad\ast J_{m}=g\widehat{V_{m}}, (113)

respectively. To allow for magnetic sources without violating the Bianchi identity d2A=0d^{2}A=0, we further introduce a (dp2)(d-p-2)-dimensional Dirac brane ending on the magnetic brane and a (dp1)(d-p-1)-form current GG localized to the Dirac brane worldvolume satisfying

dG=Jm.d\ast G=\ast J_{m}. (114)

We then define the field strength

FdA+G,F\equiv dA+\ast G, (115)

in terms of which the classical action is

S[A,G]=12FF(1)pAJe.S[A,G]=-\frac{1}{2}\int F\wedge\ast F-(-1)^{p}\int A\wedge\ast J_{e}. (116)

In natural units, the form fields AA and FF and the charge densities ee and gg have mass dimensions

[A]=[F]=[e]=[g]=d2(p+1).[A]=[F]=-[e]=[g]=\frac{d}{2}-(p+1). (117)

This action treats electric and magnetic sources highly asymmetrically. While the coupling of AA to the electric brane is manifest, the coupling of AA to the magnetic brane (via the Dirac brane) is hidden in the FFF\wedge\ast F term. To ensure that the Dirac brane introduces no independent dynamics (i.e., that the equation of motion for GG follows from that of AA), we require that the Dirac brane does not intersect the electric brane (“Dirac veto”).

We assume that the brane sources are non-dynamical. To make the branes dynamical, we would add a kinetic term for each brane of the form

TV1=TVddimVσ|h|,-T\int_{V}\ast 1=-T\int_{V}d^{\dim V}\sigma\sqrt{|h|}, (118)

where 1\ast 1 is the appropriate volume form on the worldvolume VV (with coordinates σ\sigma) and the tension TT has mass dimension [T]=dimV[T]=\dim V. The signature of the induced metric hab=σaXμσbXμh_{ab}=\partial_{\sigma^{a}}X_{\mu}\partial_{\sigma^{b}}X^{\mu} depends on the position of the brane. The action would then additionally be a functional of the brane embedding coordinates X(σ)X(\sigma).

Both fields AA and GG are singular along the Dirac brane, while FF is nonsingular except at the magnetic brane. The Dirac brane current GG cancels the magnetic flux that would otherwise be carried away from the magnetic brane by the singular part of the field due to AA. The Dirac brane is unphysical because its configuration is gauge-dependent. To see this, note that if spacetime is topologically trivial, then the existence of a Dirac brane worldvolume VDV_{D} satisfying VD=Vm\partial V_{D}=V_{m} follows from the fact that the magnetic brane worldvolume VmV_{m} is closed (Vm=0\partial V_{m}=0). This relation determines only the homology class of VDV_{D}, so the Dirac brane is ambiguous up to

VDVD+V.V_{D}\to V_{D}+\partial V. (119)

Via (64), we may take G=(1)pgVD^\ast G=(-1)^{p}g\widehat{V_{D}}, which then satisfies dG=Jmd\ast G=\ast J_{m} because VD=Vm\partial V_{D}=V_{m}. This ambiguity translates into a gauge freedom

GG+dΛ,AA+dλΛ,\ast G\to\ast G+d\ast\Lambda,\qquad A\to A+d\lambda-\ast\Lambda, (120)

where Λ=gV^\ast\Lambda=-g\widehat{V}. The gauge field AA must undergo a compensatory gauge transformation to leave FF invariant. Λ\ast\Lambda represents a singular gauge transformation of AA (on top of ordinary gauge transformations AA+dλA\to A+d\lambda) because Λ\ast\Lambda is only locally exact.

The field equations are then the modified Bianchi identity for FF and the equation of motion for AA, with variations taken from the left. Both JeJ_{e} and JmJ_{m} are conserved, as required by gauge invariance. Conservation of brane current (dJ=0d\ast J=0) is Poincaré dual to the statement that the worldvolume has no boundary (V=0\partial V=0).

D.2 Classical Duality Group

When d=2(p+1)d=2(p+1), dyonic branes exist, the charge density is dimensionless, and the pp-form Maxwell action is conformal: Tμ=μ(d2(p+1))=0T_{\mu}{}^{\mu}=(d-2(p+1))\mathcal{L}=0. In this case, the field equations have the same form degree,

d(FF)=(JmJe),d\left(\begin{array}[]{c}F\\ \ast F\end{array}\right)=\ast\left(\begin{array}[]{c}J_{m}\\ J_{e}\end{array}\right), (121)

and are therefore preserved by linear transformations

(FF)R(FF),(JmJe)R(JmJe)\left(\begin{array}[]{c}F\\ \ast F\end{array}\right)\to R\left(\begin{array}[]{c}F\\ \ast F\end{array}\right),\quad\left(\begin{array}[]{c}J_{m}\\ J_{e}\end{array}\right)\to R\left(\begin{array}[]{c}J_{m}\\ J_{e}\end{array}\right) (122)

for RGL(2,)R\in GL(2,\mathbb{R}) [35]. Further requiring such a transformation to respect the Hodge duality condition 2F=(1)pF\ast^{2}F=(-1)^{p}F and to preserve the stress tensor restricts RR to lie in SO(2)SO(2) for pp odd and 2×2\mathbb{Z}_{2}\times\mathbb{Z}_{2} for pp even, where the 2\mathbb{Z}_{2} factors act as FFF\leftrightarrow-F and FFF\leftrightarrow\ast F. This is the classical electric-magnetic duality group of pp-form electrodynamics. It can be promoted to an off-shell symmetry of the action using a two-potential formulation [35] in which

(FF)=d(AB)+(GH),\left(\begin{array}[]{c}F\\ \ast F\end{array}\right)=d\left(\begin{array}[]{c}A\\ B\end{array}\right)+\ast\left(\begin{array}[]{c}G\\ H\end{array}\right), (123)

where A,BA,B are the electric and magnetic potentials and G,HG,H are Dirac brane currents for the magnetic and electric branes satisfying dG=Jmd\ast G=\ast J_{m} and dH=Jed\ast H=\ast J_{e}. We can then imagine the usual minimal coupling to the brane worldvolumes:

(1)peVeA(1)dp2gVmB.-(-1)^{p}e\int_{V_{e}}A-(-1)^{d-p-2}g\int_{V_{m}}B. (124)

Such an action sacrifices manifest Lorentz invariance. The construction of a Lagrangian for both electric and magnetic charges that simultaneously manifests locality and Lorentz invariance is a long-standing problem [36, 37].

The electric-magnetic duality group can be extended to SL(2,)SL(2,\mathbb{R}) in the presence of a theta term for d=2(p+1)d=2(p+1) and pp odd.191919“Diagonal” theta terms are only allowed when 4|d4|d since FF=(1)p+1FFF\wedge F=(-1)^{p+1}F\wedge F by antisymmetry of the wedge product. However, mixed theta terms are allowed for pp even [15]. We do not discuss the inclusion of Chern-Simons terms in odd dd [10]. To see this, we rescale Ae1AA\to e^{-1}A so that the charge densities become e1e\to 1 and gg/eg\to g/e, resulting in a non-canonically normalized kinetic term:

S=12e2FF+θ4πFF(1)pVA,S=-\frac{1}{2e^{2}}\int F\wedge\ast F+\frac{\theta}{4\pi}\int F\wedge F-(-1)^{p}\int_{V}A, (125)

where all couplings are dimensionless. Assuming that the Dirac branes do not intersect (GG=0\ast G\wedge\ast G=0) and dropping boundary terms, we can write

S=12e2FF(1)pA(Je+θ2πJm).S=-\frac{1}{2e^{2}}\int F\wedge\ast F-(-1)^{p}\int A\wedge\left(\ast J_{e}+\frac{\theta}{2\pi}\ast J_{m}\right). (126)

The field equations are modified to

d(F1e2F)=(JmJe+θ2πJm),d\left(\begin{array}[]{c}F\\ \frac{1}{e^{2}}\ast F\end{array}\right)=\ast\left(\begin{array}[]{c}J_{m}\\ J_{e}+\frac{\theta}{2\pi}J_{m}\end{array}\right), (127)

thus manifesting the Witten effect (the theta angle shifts the electric charge of a dyonic brane by θg2π\frac{\theta g}{2\pi}) [13, 15]. In terms of the complex coupling and the complex field strength

τ=τ1+iτ2=θ2π+ie2,Fc=F+iF,\tau=\tau_{1}+i\tau_{2}=\frac{\theta}{2\pi}+\frac{i}{e^{2}},\qquad F^{c}=F+i\ast F, (128)

and defining the two-component objects

ψ(1τ),Re(ψFc),𝒥(JmJe),\psi\equiv\left(\begin{array}[]{c}1\\ -\tau\end{array}\right),\quad\mathcal{F}\equiv\operatorname{Re}(\psi F^{c}),\quad\mathcal{J}\equiv\left(\begin{array}[]{c}J_{m}\\ J_{e}\end{array}\right), (129)

the field equations become

d=𝒥.d\mathcal{F}=\ast\mathcal{J}. (130)

They are invariant under

ψ(R1)Tψ,𝒥(R1)T𝒥\psi\to(R^{-1})^{T}\psi,\qquad\mathcal{J}\to(R^{-1})^{T}\mathcal{J} (131)

for RGL(2,)R\in GL(2,\mathbb{R}). The requirement of preserving the self-duality condition

=γω,γ1τ2(1τ1τ1|τ|2),ω(0110)\mathcal{F}=\gamma\omega\ast\mathcal{F},\quad\gamma\equiv\frac{1}{\tau_{2}}\left(\begin{array}[]{cc}1&-\tau_{1}\\ -\tau_{1}&|\tau|^{2}\end{array}\right),\quad\omega\equiv\left(\begin{array}[]{cc}0&-1\\ 1&0\end{array}\right) (132)

in d=2(p+1)d=2(p+1) with pp odd restricts RSL(2,)R\in SL(2,\mathbb{R}). Indeed, writing

γ=ψψ+c.c.det(ψψ+c.c.)\gamma=\frac{\psi\psi^{\dagger}+\text{c.c.}}{\sqrt{\det(\psi\psi^{\dagger}+\text{c.c.})}} (133)

(following [38]), we see that the SL(2,)SL(2,\mathbb{R}) transformation

R=(abcd),adbc=1R=\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right),\qquad ad-bc=1 (134)

takes γ(R1)TγR1\gamma\to(R^{-1})^{T}\gamma R^{-1} while R1ω(R1)T=ωR^{-1}\omega(R^{-1})^{T}=\omega. Correspondingly, the complex coupling transforms as

τaτ+bcτ+d.\tau\to\frac{a\tau+b}{c\tau+d}. (135)

D.3 Dirac Quantization Condition

The Dirac quantization condition (DQC) follows by demanding invariance under singular gauge transformations [16]. The change in the action under a singular gauge transformation comes from the coupling of AA to JeJ_{e}:

ΔSΛJeegL(Ve,V),\Delta S\propto\int\ast\Lambda\wedge\ast J_{e}\propto egL(V_{e},\partial V), (136)

where the proportionality factors are signs. This immediately gives the DQC, as long as the variation in the Dirac brane worldvolume can link with the electric brane worldvolume.

Now consider dimensions d=2(p+1)d=2(p+1). Consider two dyonic branes of charge densities (e1,g1)(e_{1},g_{1}) and (e2,g2)(e_{2},g_{2}) as well as worldvolumes and corresponding Dirac branes V1=D1V_{1}=\partial D_{1} and V2=D2V_{2}=\partial D_{2}. A variation in the configurations of the Dirac branes,

D1D1D1+E1,D2D2D2+E2,D_{1}\to D_{1}^{\prime}\equiv D_{1}+\partial E_{1},\quad D_{2}\to D_{2}^{\prime}\equiv D_{2}+\partial E_{2}, (137)

is implemented by a singular gauge transformation that takes

AAΛ,Λ=g1E1^g2E2^.A\to A-\ast\Lambda,\qquad\ast\Lambda=-g_{1}\widehat{E_{1}}-g_{2}\widehat{E_{2}}. (138)

Again, the change in the action comes from the coupling of AA to JeJ_{e} (the theta term, if present, does not affect the argument because it is manifestly gauge-invariant). Using Je=e1V1^+e2V2^\ast J_{e}=e_{1}\smash{\widehat{V_{1}}}+e_{2}\smash{\widehat{V_{2}}}, we can write this variation as an intersection number:

ΔS\displaystyle\Delta S =(1)pΛJe\displaystyle=\textstyle(-1)^{p}\int\ast\Lambda\wedge\ast J_{e} (139)
=(e1g2I(V1,E2)+e2g1I(V2,E1))\displaystyle=-(e_{1}g_{2}I(V_{1},E_{2})+e_{2}g_{1}I(V_{2},E_{1})) (140)
=(e1g2I(D1,E2)+e2g1I(D2,E1))\displaystyle=-(e_{1}g_{2}I(\partial D_{1},E_{2})+e_{2}g_{1}I(\partial D_{2},E_{1})) (141)
=(e1g2I(E2,D1)+e2g1I(E1,D2)),\displaystyle=-(e_{1}g_{2}I(\partial E_{2},D_{1})+e_{2}g_{1}I(\partial E_{1},D_{2})), (142)

where we use that dimVi=dimDi1=dimEi2=p\dim V_{i}=\dim D_{i}-1=\dim E_{i}-2=p. For arbitrary variations of the Dirac branes, we clearly require e1g22πe_{1}g_{2}\in 2\pi\mathbb{Z} and e2g12πe_{2}g_{1}\in 2\pi\mathbb{Z} separately (the DQC), but for variations in which the Dirac branes in their initial and final configurations do not intersect, we have

I(D1,D2)=I(D1,D2)=0.I(D_{1},D_{2})=I(D_{1}^{\prime},D_{2}^{\prime})=0. (143)

The intersection number of two boundaries is zero since 2=0\partial^{2}=0, so I(E1,E2)=I(D1D1,D2D2)=0I(\partial E_{1},\partial E_{2})=I(D_{1}^{\prime}-D_{1},D_{2}^{\prime}-D_{2})=0, from which we see that

I(D1,D2)+I(D1,D2)=0.I(D_{1},D_{2}^{\prime})+I(D_{1}^{\prime},D_{2})=0. (144)

Hence we deduce that

ΔS=(1)p(e1g2+(1)pe2g1)I(D1,D2),\Delta S=(-1)^{p}(e_{1}g_{2}+(-1)^{p}e_{2}g_{1})I(D_{1},D_{2}^{\prime}), (145)

and we get e1g2+(1)pe2g12πe_{1}g_{2}+(-1)^{p}e_{2}g_{1}\in 2\pi\mathbb{Z}. The consistency of this DSZ quantization condition with dimensional reduction is discussed in [12, 15].

D.4 Quantum Duality Group

The DSZ pairing appearing in (72) is invariant under the action of the classical electric-magnetic duality group, namely SL(2,)SL(2,\mathbb{R}) for pp odd and the 2×2\mathbb{Z}_{2}\times\mathbb{Z}_{2} generated by {(0110),(1001)}\{\left(\begin{smallmatrix}0&1\\ 1&0\end{smallmatrix}\right),\left(\begin{smallmatrix}-1&0\\ 0&-1\end{smallmatrix}\right)\} for pp even. In general, however, only a discrete subgroup of the classical duality group preserves the quantum-mechanical charge lattice.

In d=2(p+1)d=2(p+1) with pp odd, (72) takes the form

e1g2e2g12π.e_{1}g_{2}-e_{2}g_{1}\in 2\pi\mathbb{Z}. (146)

The relative minus sign ensures that the quantization condition is invariant under the Witten effect. In a convenient normalization, the maximal charge lattice allowed by (72) is

(e,g)=(m+θn,2πn)(e,g)=(m+\theta n,2\pi n) (147)

for m,nm,n\in\mathbb{Z}. Letting q=e+igq=e+ig\in\mathbb{C}, this lattice defines a torus qq+1q\sim q+1 and qq+2πτq\sim q+2\pi\tau with τ=θ2π+i\tau=\frac{\theta}{2\pi}+i. The subgroup of the classical duality group SL(2,)SL(2,\mathbb{R}) that preserves the charge lattice is SL(2,)SL(2,\mathbb{Z}). If θ\theta is constrained to vanish, then the subgroup of the classical duality rotation group SO(2)SO(2) that preserves the resulting rectangular charge lattice is 4\mathbb{Z}_{4}.

In d=2(p+1)d=2(p+1) with pp even, (72) takes the form

e1g2+e2g12π,e_{1}g_{2}+e_{2}g_{1}\in 2\pi\mathbb{Z}, (148)

and theta terms are not allowed. The corresponding charge lattice (e,g)=(m,2πn)(e,g)=(m,2\pi n) for m,nm,n\in\mathbb{Z} is preserved by the full classical duality group 2×2\mathbb{Z}_{2}\times\mathbb{Z}_{2}.

The SL(2,)SL(2,\mathbb{Z}) duality is an ambiguity in the presentation of the theory, i.e., in the choice of which sources to call electric or magnetic. It can be thought of as a change of variables in the path integral. Of the generators of SL(2,)SL(2,\mathbb{Z}), the TT-transformation T:ττ+1T:\tau\mapsto\tau+1 simply takes θθ+2π\theta\to\theta+2\pi. The SS-transformation S:τ1/τS:\tau\mapsto-1/\tau also has a simple path integral interpretation. For d=2(p+1)d=2(p+1) with pp odd, consider the rescaled action with no sources:

S[A]=(12e2FF+θ4πFF),FdA.S[A]=\int\left(-\frac{1}{2e^{2}}F\wedge\ast F+\frac{\theta}{4\pi}F\wedge F\right),\quad F\equiv dA. (149)

Assuming that spacetime is topologically trivial, we can instead view the action as a functional of the fundamental fields BB and FF, where BB is a pp-form Lagrange multiplier that enforces the Bianchi identity, thus implying that FF is the field strength of a pp-form potential:

S[B,F]=(12e2FF+θ4πFFBdF).S[B,F]=\int\left(-\frac{1}{2e^{2}}F\wedge\ast F+\frac{\theta}{4\pi}F\wedge F-B\wedge dF\right). (150)

Integrating out FF then gives

S[B]=(12e2GG+θ4πGG),GdB,S[B]=\int\left(-\frac{1}{2e^{\prime 2}}G\wedge\ast G+\frac{\theta^{\prime}}{4\pi}G\wedge G\right),\quad G\equiv dB, (151)

where the new complex coupling is

τ=θ2π+ie2=1τ.\tau^{\prime}=\frac{\theta^{\prime}}{2\pi}+\frac{i}{e^{\prime 2}}=-\frac{1}{\tau}. (152)

Under the duality FGF\leftrightarrow G, the equation of motion for AA goes to the Bianchi identity for BB and vice versa. A similar argument establishes the quantum equivalence of abelian pp-form and (dp2)(d-p-2)-form gauge fields.