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

Topology Obstructing Anderson Localization of Light

Tobias Micklitz Centro Brasileiro de Pesquisas Físicas, Rua Xavier Sigaud 150, 22290-180, Rio de Janeiro, Brazil    Alexander Altland Institut für Theoretische Physik, Universität zu Köln, Zülpicher Str. 77, 50937 Cologne, Germany Centro Brasileiro de Pesquisas Físicas, Rua Xavier Sigaud 150, 22290-180, Rio de Janeiro, Brazil Institut für Theoretische Physik, Universität zu Köln, Zülpicher Str. 77, 50937 Cologne, Germany
(March 10, 2025)
Abstract

Light propagation in random media is notorious for its resilience to (Anderson) wave localization. We here argue that in materials with slow helicity relaxation previously unnoticed principles of topology reinforce delocalization. We show that the effective electromagnetic action is governed by principles, identical to those protecting delocalization at topological insulator surface states. The length scales over which these exert a protective influence depend on the conservation of helicity, leading to the prediction that light scattering in media with, e.g., strong magnetic and electric dipolar resonances evades localization over large scales.

Abstract

In this supplemental material we discuss details on the derivation of the ‘Dirac’ equation for photons and the effective field theory action.

pacs:
05.45.Mt, 72.15.Rn, 71.30.+h

Introduction:—Electromagnetic wave propagation in random media is a problem with many facets, including the prediction of Anderson localization Anderson (1958) with promising applications in photocatalysis Gomard et al. (2013); Cernuto et al. (2011), lasing Cao (2005); Wiersma (2008); Andreasen et al. (2011), and sensing Oliver et al. (2018). However, despite an extensive body of theoretical work, its experimental observation remains ‘stubbornly elusive’ Skipetrov and Page (2016); Yamilov et al. (2023): unlike with acoustic Condat and Kirkpatrick (1987); Hu et al. (2008), elastic Flores et al. (2013); Ángel et al. (2019), quantum matter Chabé et al. (2008); Lemarié et al. (2010); Billy et al. (2008); Roati et al. (2008); White et al. (2020), or optical fiber Schwartz et al. (2007); De Raedt et al. (1989) waves, the unambiguous observation of Anderson localization of light in three dimensions (3d3d) is pending. Recently, it has been proposed that the elusiveness of localization may originate in light’s vector nature Skipetrov and Sokolov (2014); Cherroret et al. (2016); van Tiggelen and Skipetrov (2021). We here bring an independent alternative into play, delocalization protected by topology.

Conceptually, Anderson localization is a highly universal phenomenon, backed by powerful concepts of statistical mechanics, and described in terms of effective field theories Efetov (1997); Evers and Mirlin (2008). The one principle powerful enough to trump localization in low spatial dimensions d3d\leq 3 is topology. A case in point are the d4d\leq 4-dimensional surfaces of topological insulators Ryu et al. (2010), for which the one-parameter scaling paradigm would require localization by strong disorder, while topological principles demand otherwise. Topology wins, and the ways in which it does can be described in different manners. One option starts with the observation that the band structure of a topological insulator is governed by a single (more generally, an odd number) of Dirac cones. The non-vanishing helicity of the latter requires that surface states remain extended. A more explicit approach reconsiders the universal field theories of localization to observe that in the presence of single Dirac cones they pick up topological Chern-Simons or θ\theta-terms blocking the renormalization group flow into a localized phase Evers and Mirlin (2008).

We here argue that similar principles are at work for light scattering. Our starting point will be a representation of the light scattering problem in terms of a Dirac equation, linear in space and time derivatives, as a direct descendant of Maxwell’s equations. (For our present discussion, this formulation is more natural than the effective second order differential equations more commonly employed in light-localization theory.) On general grounds, the Dirac problem splits into two sectors (‘Weyl-fermions’) distinguished by their helicity. Within the topological insulator analogy, these would correspond to opposite surfaces, separated in real space. Presently, the separation is in momentum space, a scenario otherwise realized in the Weyl-semimetal Armitage et al. (2018) Either way, as long as the two sectors are not coupled by helicity non-conserving scattering, delocalization is granted. The concrete question then is to what degree they remain isolated in the light scattering context, and how efficiently this principle extends localization lengths. Naturally, the answer depends on the microscopic realization of scattering channels, where the vectorial nature of light implies a rich spectrum of scenarios: from helicity conserving dual-scattering enjoying perfect protection to the irrelevance of the topological principle in the case of strong helicity-mixing. In the following, we combine elements of the theory of disordered Weyl quantum matter Altland and Bagrets (2015); Altland et al. (2002) with that of light scattering to a first principle analysis of the situation.

Dirac approach to light scattering:—Our starting point are Maxwell’s equations in absence of external sources (summation convention)

iDi\displaystyle\partial_{i}D_{i} =0,ϵijkjEk=tBi,\displaystyle=0,\qquad\epsilon_{ijk}\partial_{j}E_{k}=-\partial_{t}B_{i}, (1)
iBi\displaystyle\partial_{i}B_{i} =0,ϵijkjHk=tDi,\displaystyle=0,\qquad\epsilon_{ijk}\partial_{j}H_{k}=\partial_{t}D_{i}, (2)

where the presence of a scattering medium enters through the constitutive relations Bi=μHiB_{i}=\mu H_{i}, Di=ϵEiD_{i}=\epsilon E_{i} with randomly fluctuating magnetic permeability μ=μ0μ𝕩\mu=\mu_{0}\mu_{\mathbb{x}} and permittivity ϵ=ϵ0ϵ𝕩\epsilon=\epsilon_{0}\epsilon_{\mathbb{x}}, respectively (c0=1/ϵ0μ0=1c_{0}=1/\sqrt{\epsilon_{0}\mu_{0}}=1 throughout). As an alternative to the standard strategy John (1984); Lagendijk and van Tiggelen (1996) of combining Eqs. (1) and (2) to a single random wave equation for the electric field, we here follow Ref. Bialynicki-Birula (1996) and introduce field components i=ϵEi{\cal E}_{i}=\sqrt{\epsilon}E_{i} and i=μHi{\cal H}_{i}=\sqrt{\mu}H_{i}, to define the six-component ‘single photon wave function’ Φ=(Φ+,Φ)\Phi=(\Phi^{+},\Phi^{-}), comprising the complex linear combinations Φ±=(±i)/2n\Phi^{\pm}=({\cal E}\pm i{\cal H})/\sqrt{2n} associated to vacuum solutions of opposite circular polarization, or helicity, with n=ϵμn=\sqrt{\epsilon\mu} the (local) refractive index. Expressed in terms of this ‘single photon wave function’ Bialynicki-Birula (1996), Maxwell’s equations for solutions with monochromatic time dependence of frequency ω0\omega_{0} assume the form (see supplemental material)

(τ3τ1Z,𝕩+V𝕩)Φ\displaystyle\left(\tau_{3}\otimes\not{p}-\tau_{1}\otimes\not{a}_{Z,\mathbb{x}}+V_{\mathbb{x}}\right)\Phi =ω0Φ.\displaystyle=\omega_{0}\Phi. (3)

Here, the matrices τi\tau_{i} act between the helical subspaces of the Φ\Phi-vector and we use a variant of the Feynman slash notation, =viSi\not{v}=v_{i}S_{i}, where (Sj)αβ=iϵjαβ(S_{j})_{\alpha\beta}=-i\epsilon_{j\alpha\beta} are three-dimensional matrices representing the components of a spin-1 operator, with commutation relations [Si,Sj]=iϵijkSk[S_{i},S_{j}]=i\epsilon_{ijk}S_{k} and SiSi=2𝟙𝟛S_{i}S_{i}=2\openone_{3}. Randomness enters this equation through a scalar potential V=ω0(1n)V=\omega_{0}(1-n), and a vector potential, Z=12∂̸lnZ\not{a}_{Z}=\frac{1}{2}\not{\partial}\ln Z, involving the local refractive index n𝕩n_{\mathbb{x}} and impedance, Z𝕩=μ𝕩/ϵ𝕩Z_{\mathbb{x}}=\sqrt{\mu_{\mathbb{x}}/\epsilon_{\mathbb{x}}}, respectively. We assume a short range correlated Gaussian distribution of the former V𝕩=0\langle V_{\mathbb{x}}\rangle=0, V𝕩V𝕩=w0δ(𝕩𝕩)\langle V_{\mathbb{x}}V_{\mathbb{x}^{\prime}}\rangle=w_{0}\delta(\mathbb{x}-\mathbb{x}^{\prime}), and leave the latter unspecified for now. Finally, the linear operator in Eq. (3) possesses time-reversal symmetry, H=THτ3HTτ3H=\textrm{T}H\equiv\tau_{3}H^{T}\tau_{3}, squaring to unity, T2=𝟙\textrm{T}^{2}=\openone, placing the problem into the orthogonal Wigner-Dyson class-AI{\rm AI}.

Before turning to the analysis of the disorder, we note two key algebraic structures of Eq. (3). Helicity: in the absence of impedance variations, Z=0\not{a}_{Z}=0, the equation decouples into equations (χω0)Φχ=0(\chi\not{p}-\omega_{0})\Phi_{\chi}=0, for the two sectors χ=±\chi=\pm. These play a role analogous to the states of opposite chirality in the Weyl-semimetal, and enjoy individual topological protection against Anderson localization. The second structure, which has no analog in the condensed matter context, is transversality: Unlike vacuum, inhomogeneous media admit longitudinal field components, which turn out to be key to the delocalization problem.

To organize the field solutions into transverse and longitudinal components, we consider the retarded vacuum field propagators of definite helicity, defined as (ω0+iδχ)Gχ,p0=𝟙𝟛\left(\omega_{0}+i\delta-\chi\not{p}\right)G^{0}_{\chi,p}=\openone_{3}, where p=(ω0+iδ,𝕡)p=(\omega_{0}+i\delta,\mathbb{p}) is the four-momentum and δ0\delta\searrow 0. Introducing the projection operators π¯𝕡=𝕡𝕡t/(pipi)\bar{\pi}_{\mathbb{p}}=\mathbb{p}\mathbb{p}^{t}/(p_{i}p_{i}), π𝕡=1π¯𝕡\pi_{\mathbb{p}}=1-\bar{\pi}_{\mathbb{p}}, a straightforward computation shows that the propagator assumes the form Gp0=Gt,p0π𝕡+Gl,p0π¯𝕡G^{0}_{p}=G^{0}_{t,p}\pi_{\mathbb{p}}+G^{0}_{l,p}\bar{\pi}_{\mathbb{p}}, with a propagating transverse (tt) and non-propagating longitudinal (ll) contribution given by

Gχ,t,p0\displaystyle G^{0}_{\chi,t,p} =ω0+iδ+χ(ω0+iδ)2p2,Gχ,l,p0=1ω0+iδ.\displaystyle=\frac{\omega_{0}+i\delta+\chi\not{p}}{(\omega_{0}+i\delta)^{2}-p^{2}},\qquad G^{0}_{\chi,l,p}=\frac{1}{\omega_{0}+i\delta}. (4)

As a first step towards the inclusion of disorder, we consider these propagators averaged over realizations of VV, G0G0GG_{0}\to\langle G_{0}\rangle\equiv G. Referring to the supplemental material for details, we assume weak disorder and compute the propagator within a self-consistent Born approximation (SCBA), leading to Gs=(Gs,01Σ)1G_{s}=(G_{s,0}^{-1}-\Sigma)^{-1}, with s(χ,t/l)s\equiv(\chi,t/l). For the simple disorder model used here, and at length scales ω01\gtrsim\omega_{0}^{-1} exceeding the wave length 111For a detailed discussion of the averaged propagator at short length scales, <ω0<\omega_{0}, see Ref. van Tiggelen and Skipetrov (2021)., the self energy Σ=iκ\Sigma=-i\kappa becomes a constant κ=κ(w0,ω0,Λ)ω0\kappa=\kappa(w_{0},\omega_{0},\Lambda)\ll\omega_{0} depending on frequency, disorder strength, and an ultraviolett cutoff, Λ\Lambda, set by the inverse width of the effective δ\delta-function determining the disorder correlation length. Physically, κ\kappa, describes the damping of the averaged propagator at length scales exceeding the mean free path κ1\sim\kappa^{-1}.

Anderson localization shows in observables involving disorder averaged products of advanced and retarded propagators, such as the inverse participation ratio of the electromagnetic energy density Mirlin (2000); Efetov (1997). It is the result of multiple constructive self interference of diffusion modes formed by co-propagating amplitudes of opposite causality, at length scales parametrically exceeding the mean free path. While ‘weak localization’ precursors of the phenomenon can be computed in diagrammatic perturbation theory, the understanding of strong localization (or the absence thereof) requires field theoretical analysis in terms of nonlinear σ\sigma-models Mirlin (2000); Efetov (1997).

Field theory:—The application of the replica field theory formalism to light localization goes back to Refs. John (1987, 1984, 1985). Referring to the original reference and the supplemental material for details, we note that prior to the inclusion of helicity mixing, this theory is described by a functional integral 𝒵=D(T+,T)eS+[T+]S[T]{\cal Z}=\int D(T_{+},T_{-})e^{-S_{+}[T_{+}]-S_{-}[T_{-}]}, split into two helicity sectors χ=±\chi=\pm with

Sχ[Tχ]\displaystyle S_{\chi}[T_{\chi}] =12Trln(ω0χ+iκTχσ3raTχ1).\displaystyle=-\frac{1}{2}{\rm Tr}\ln\left(\omega_{0}-\chi\not{p}+i\kappa T_{\chi}\sigma_{3}^{\rm ra}T_{\chi}^{-1}\right). (5)

Treating retarded and advanced propagators in a unified fashion, the action contains a Pauli matrix σ3ra\sigma_{3}^{\textrm{ra}} multiplying the causal symmetry breaking imaginary self energy iκi\kappa. The fact that prior to averaging the retarded and advanced action were identical (except of the infinitesimal iδi\delta), reflects in the presence of Goldstone modes, acting on the self energy operator iκσ3rai\kappa\sigma_{3}^{\mathrm{ra}} as slowly varying rotation matrices T(x)Sp(4R)T(\textbf{x})\in\textrm{Sp}(4R), where RR is the number of replicas, the symplectic structure is a consequence Efetov (1997) of the class AI time reversal symmetry, and the trace in Eq. (5) extends over both, the 4R4R-dimensional matrix space of the Goldstone modes, and Hilbert space. With the definition Qχ=Tχσ3raTχ1Q_{\chi}=T_{\chi}\sigma_{3}^{\mathrm{ra}}T_{\chi}^{-1}, we have identified the effective degrees of freedom of our theory, whose long range fluctuation behavior will decide over its localization behavior.

Eq. (5) is almost identical to the action of a single node in the Brillouin zone of a Weyl-semimetal in the presence of smoothly varying disorder (i.e., disorder excluding the scattering between nodes) Altland and Bagrets (2016). The differences — the matrices SiS_{i} acting in a spin-1 representation of SU(2)SU(2), instead of spin-1/2 Pauli matrices in the Weyl context, and the Weyl Goldstone mode matrices QQ lacking the symplectic structure representing symmetry class AI — take no influence on the next step in the construction of the theory: expansion of the ‘Tr ln’ to leading order in derivatives acting on the slow differences. Referring to Refs. Altland et al. (2002); Altland and Bagrets (2016) and the supplemental material for details, this exercise leads to the effective action

Sχ[Tχ]=Sd[Tχ]+χSCS[Tχ],\displaystyle S_{\chi}[T_{\chi}]=S_{\rm d}[T_{\chi}]+\chi S_{\rm CS}[T_{\chi}], (6)

where

Sd[Tχ]\displaystyle S_{\rm d}[T_{\chi}] =σ16d3xtr(iQχiQχ),\displaystyle=\frac{\sigma}{16}\int d^{3}x\,{\rm tr}\left(\partial_{i}Q_{\chi}\partial_{i}Q_{\chi}\right), (7)

is the action of the nonlinear σ\sigma model for three-dimensional disordered media. In Eq. (7), ‘tr{\rm tr}’ denotes the trace over internal degrees of freedom, and the coupling constant σ=(ω02+3κ2)(ω0213κ2)/(πκ(ω02+κ2))\sigma=(\omega_{0}^{2}+3\kappa^{2})\left(\omega_{0}^{2}-\frac{1}{3}\kappa^{2}\right)/\left(\pi\kappa(\omega_{0}^{2}+\kappa^{2})\right) plays the role of a bare conductivity characterizing the transport properties of the medium at length scales of the order of the scattering mean free path κ1\kappa^{-1}. Depending on whether σκ1\sigma\kappa^{-1} is large or small compared to unity, this action describes the flow into a diffusive or a localized phase, the two scenarios being separated by the Anderson localization transition Evers and Mirlin (2008).

The reason why the effective action SχS_{\chi} predicts different behavior has to do with the presence of the second term, SCS[Tχ]SCS[Aχ+]SCS[Aχ]S_{\textrm{CS}}[T_{\chi}]\equiv S_{\textrm{CS}}[A_{\chi}^{+}]-S_{\textrm{CS}}[A_{\chi}^{-}], where

SCS[A]\displaystyle S_{\rm CS}[A] =iϵijk8πd3xtr(AijAk+23AiAjAk),\displaystyle=-\frac{i\epsilon_{ijk}}{8\pi}\int d^{3}x\,{\rm tr}\left(A_{i}\partial_{j}A_{k}+\frac{2}{3}A_{i}A_{j}A_{k}\right), (8)

is a Chern-Simons action for the ‘vector potentials’ Aχs=χTχ1TχPsA^{s}_{\chi}=\chi T^{-1}_{\chi}\nabla T_{\chi}P^{s}, and Ps=12(𝟙ar+sσ3ar)P^{s}=\frac{1}{2}(\mathds{1}^{\textrm{ar}}+s\sigma_{3}^{\textrm{ar}}) projects onto retarded or advanced space. The presence of this term in the effective action follows from high-level reasons, namely the equivalence of Eq. (5) to the action of three-dimensional linearly dispersive fermions coupled to ‘gauge fluctuations’ TχT_{\chi}. For this system, the parity anomaly Redlich (1984) requires the presence of a Chern-Simons term. In the specific case, where the fields TχT_{\chi} are generated by disorder averaging, this term assumes the specific form Eq. (8Altland and Bagrets (2016).

The absence of localization in a system governed by the full action S=Sd+SCSS=S_{\textrm{d}}+S_{\textrm{CS}} follows from another high-level argument: the prototypical action Eq. (5) describes a three-dimensional linearly dispersive and time reversal invariant system, coupled to disorder. In condensed matter physics, such systems are conceptualized as surface states of four dimensional topological insulators in symmetry class AI, and topological obstructions prevent the localization of their microscopic quantum states. This statement includes strongly disordered cases, where the conductance σκ1\sigma\kappa^{-1} becomes of order unity and the derivation of the effective action SS is no longer under parametric control.

Helicity mixing:—

Our so-far analysis provides robust evidence for the absence of Anderson localization in helicity conserving light scattering. Within the context of the condensed matter analogy, sectors of definite helicity correspond to isolated Weyl nodes, or single topological insulator surfaces. Helicity non-conservation is analogous to inter-node scattering, or coupling between opposite surfaces, and hence compromises topological protection.

Helicity non-conservation in light scattering is caused by local variations of the impedance, 𝕒Z,𝕩=12lnZ𝕩\mathbb{a}_{Z,\mathbb{x}}=\frac{1}{2}\nabla\ln Z_{\mathbb{x}}. In the Dirac equation (3), 𝕒Z\mathbb{a}_{Z} is coupled to τ1\tau_{1} hybridizing the τ3\tau_{3}-eigenspaces χ=±1\chi=\pm 1. To explore the consequences, we assume fluctuations of the helicity mixing field short range correlated as aZ,i,xaZ,j,y=γδijδ(xy)\langle a_{Z,i,\textbf{x}}a_{Z,j,\textbf{y}}\rangle=\gamma\delta_{ij}\delta(\textbf{x}-\textbf{y}). Expansion of the ‘Tr ln’ in Eq. (5) to second order in aZa_{Z} followed by averaging then leads to the coupling action (see the supplemental material)

S±[Q]=6γκ2w02d3xtr(Q+Q).\displaystyle S_{\pm}[Q]=\frac{6\gamma\kappa^{2}}{w_{0}^{2}}\int d^{3}x\,\textrm{tr}(Q_{+}Q_{-}). (9)

The effect of this action is a locking of the two previously independent soft modes TχT_{\chi}. (For T+=TTT_{+}=T_{-}\equiv T, the coupling action vanishes because tr(Q+Q)tr(Q2)=tr(𝟙)0\mathrm{tr}(Q_{+}Q_{-})\to\mathrm{tr}(Q^{2})=\mathrm{tr}(\mathds{1})\to 0, in the replica limit.) To understand at what length scales the locking becomes effective, we compare the gradient term (7) to the coupling action (9). At length scales, l0\sim l_{0}, the lowest lying fluctuations cost a gradient action σl0\sim\sigma l_{0} comparable to the conductance at these scales. The coupling action at the same length scales is γ(κ/w0)2l03\sim\gamma(\kappa/w_{0})^{2}l_{0}^{3}. Comparison of the two estimates defines lh(σ/γ)1/2(w0/κ)l_{\textrm{h}}\equiv(\sigma/\gamma)^{1/2}(w_{0}/\kappa) as the crossover length scale beyond which the locking to a single mode TT has become effective.

The most important consequence of this field reduction is the cancellation of the Chern-Simons actions at length scales exceeding lhl_{\textrm{h}}. The reason is the helicity sign χ\chi multiplying SCSS_{\textrm{CS}} in Eq. (6). For T+=TT_{+}=T_{-}, the two helical Chern-Simons actions cancel, and we are left with (twice) the action SdS_{\textrm{d}}. The topological action now being absent, localization at length scales >lh>l_{\textrm{h}} may result, provided the disorder is strong enough to push the system below the Anderson transition threshold.

Light scattering:—The effectiveness of the mechanism discussed above depends on the degree of helicity mixing, an extreme case being the absence of it, 𝕒Z=0\mathbb{a}_{Z}=0. This condition is realized in the limit of identical electric and magnetic polarizabilities, known as “first Kerker condition” Kerker et al. (1983). Scatterers satisfying the first Kerker condition respect the duality between electric and magnetic field otherwise characterizing the vacuum. This condition is met, e.g., in the case of strong magnetic and electric dipolar resonances, where spectral overlap between quadrupole and higher-order modes is required to generate residual helicity mixing.

As a case in point, we mention dielectric sub-wavelength spheres whose scattering profile was found García-Etxarri et al. (2011); Geffrin et al. (2012) to be indistinguishable from that of dual magneto-dielectric spheres. Silicon (and likely other semiconductor materials such as germanium and rutile-TiO2 Geffrin et al. (2012)) show strong magnetic and electric dipolar resonances in the nearly absorptionless visible, telecom, and near-infrared frequencies range, likewise realizing a near-dual limit. Similarly, Refs. MacKintosh et al. (1989); Bicout et al. (1994); Gorodnichev et al. (2017) noted the persistence of circular polarization indicative of a dual limit in resonant dielectric Mie particles with high magnetic polarizability. In these cases, the randomization of circular polarization requires an order of magnitude more scattering events than the complete randomization of the wave’s direction foo , indicating approximate helicity conservation over large distances. Slow helicity relaxation was recently discussed Gorodnichev et al. (2022) to be responsible for experimentally observed anomalous features in coherent backscattering (a precursor of strong localization) in porous magnetoactive glass Lenke et al. (2000). Our present analysis shows that the protection against localization in scattering media with conserved helicity runs even deeper than that, as it is rooted in a topological principle.

We finally note that the correspondence between helicity conserving light scattering and single Weyl nodes entails phenomenological consequences beyond the mere absence of Anderson localization: Both renormalized perturbation theory Fradkin (1986) and a detailed analysis of the self-consistent Born self energies Klier et al. (2019) suggest a segregation of the parameter plane spanned by disorder, w0w_{0}, and frequency, ω0\omega_{0}, into a low-disorder/low-frequency wedge dominated by the clean Weyl semimetallic fixed and a strong disorder regime governed by diffusive transport. While these predictions must be taken with a grain of salt (analytical approaches around a single Weyl point lack parametric control, and exponentially rare strong disorder fluctuations Pixley and Wilson (2021) may compromise transitions between the purported phases), the existence of a parameter regime with maintained semimetallic behavior is backed by numerics Altland and Bagrets (2016). Translated to the light scattering context, the observable consequence may be that in the low frequency limit ω00\omega_{0}\to 0, and for disorder concentrations w0Λ/c031w_{0}\Lambda/c_{0}^{3}\lesssim 1, where the momentum scale Λ\Lambda is set by the inverse correlation length of the disorder potential, helicity conserving scattering regions look effectively transparent at large length scales. However, the quantitative exploration of such scenarios requires further study.

Discussion:—We have shown that light scattering in helicity conserving scattering media is protected by the same topological principle otherwise operational on topological insulator surfaces, or Weyl semimetals. These analogies were made manifest by mapping the light scattering problem to the same Chern-Simons action, Eqs. (7), (8), describing the condensed matter systems. The construction relied on a number of assumptions, notably the modelling of both helicity conserving and non-conserving scattering by uncorrelated Gaussian disorder. While these simplified the analysis, more general scattering potentials of, e.g., finite range correlation, would affect the value of the coupling σ\sigma, but not the structure of the action, and hence not the expected phenomenology at large distance scales.

We finally mention a few connections to related work that are not fully explored at this point: While our present theory indicates that combinations of strong disorder and helicity mixing may restore Anderson localization, Ref. Makhfudz (2018) (using the same formalism, but for time-reversal invariant Weyl semimetals) argues that non-perturbative effects of vortex-loop formation may provide an even stronger instance of topological protection. To what extent this construction carries over to light-scattering remains to be explored.

Alongside the mechanisms discussed here, longitudinal electromagnetic modes, which have no analog in the condensed matter framework, have been discussed as a potential suppressor of Anderson localization, too, see Ref. van Tiggelen and Skipetrov (2021) and numerical evidence in Ref. Yamilov et al. (2023). The point there is that longitudinal modes dominate the solution of the scattering self energy equations in the sub-wavelength regime, thus opening a transport channel (“Förster-resonance”) abesent in scalar scattering media. It is interesting to note that our mechanism, too, crucially relies on longitudinal modes: technically, the Chern-Simons action is obtained via a third order expansion of the logarithm Eq. (5) in photon propagators, and one of these has to be a longitudinal propagator (see the supplemental material). A second point to notice is that the action is obtained by integration over all wavelength, where the independendence of the coupling constant of the scattering strength indicates that the dominant contributions come from large momenta, i.e. the parametric domain dominantly supporting the longitudinal propagator. It remains to be explored if this is a coincidence or an actual connection to the physics mentioned above.

We finally mention that in condensed matter the physics of Weyl nodes manifests itself in unique transport phenomena, such as the anomalous Hall effect, and the chiral magnetic effect. It will be interesting to consider tunable electromagnetic meta-materials close to the dual limit and explore if, and in what sense there are analogous phenomena in light scattering.

Acknowledgement:—We thank Antonio Zelaquett Khoury and Felipe Pinheiro for stimulating discussions and helpful comments on the manuscript. Financial support by Brazilian agencies CNPq and FAPERJ and the Deutsche Forschungsgemeinschaft (DFG) project grant 277101999 within the CRC network TR 183 (subproject A03) is acknowledged.

References

  • Anderson (1958) P. W. Anderson, Phys. Rev. 109, 1492 (1958).
  • Gomard et al. (2013) G. Gomard, R. Peretti, E. Drouard, X. Meng, and C. Seassal, Opt. Express 21, A515-A527 (2013).
  • Cernuto et al. (2011) G. Cernuto, N. Masciocchi, A. Cervellino, G. M. Colonna, and A. Guagliardi, Journal of the American Chemical Society 133, 3114 (2011).
  • Cao (2005) H. Cao, Journal of Physics A: Mathematical and General 38, 10497 (2005).
  • Wiersma (2008) D. S. Wiersma, Nature Physics 4, 359 (2008).
  • Andreasen et al. (2011) J. Andreasen, A. A. Asatryan, L. C. Botten, M. A. Byrne, H. Cao, L. Ge, L. Labonté, P. Sebbah, A. D. Stone, H. E. Türeci, et al., Adv. Opt. Photon. 3, 88 (2011).
  • Oliver et al. (2018) J. T. Oliver, T. Crane, and L. Sapienza, in SPIE photonics europe (22/04/18 - 26/04/18) (2018).
  • Skipetrov and Page (2016) S. E. Skipetrov and J. H. Page, New Journal of Physics 18, 021001 (2016).
  • Yamilov et al. (2023) A. Yamilov, S. E. Skipetrov, T. W. Hughes, M. Minkov, Z. Yu, and H. Cao, Nature Physics (2023).
  • Condat and Kirkpatrick (1987) C. A. Condat and T. R. Kirkpatrick, Phys. Rev. Lett. 58, 226 (1987).
  • Hu et al. (2008) H. Hu, A. Strybulevych, J. H. Page, S. E. Skipetrov, and B. A. van Tiggelen, Nature Physics 4, 945 (2008).
  • Flores et al. (2013) J. Flores, L. Gutiérrez, R. A. Méndez-Sánchez, G. Monsivais, P. Mora, and A. Morales, Europhysics Letters 101, 67002 (2013).
  • Ángel et al. (2019) J. C. Ángel, J. C. T. Guzmán, and A. D. de Anda, Scientific Reports 9, 3572 (2019).
  • Chabé et al. (2008) J. Chabé, G. Lemarié, B. Grémaud, D. Delande, P. Szriftgiser, and J. C. Garreau, Phys. Rev. Lett. 101, 255702 (2008).
  • Lemarié et al. (2010) G. Lemarié, H. Lignier, D. Delande, P. Szriftgiser, and J. C. Garreau, Phys. Rev. Lett. 105, 090601 (2010).
  • Billy et al. (2008) J. Billy, V. Josse, Z. Zuo, A. Bernard, B. Hambrecht, P. Lugan, D. Clément, L. Sanchez-Palencia, P. Bouyer, and A. Aspect, Nature 453, 891 (2008).
  • Roati et al. (2008) G. Roati, C. D’Errico, L. Fallani, M. Fattori, C. Fort, M. Zaccanti, G. Modugno, M. Modugno, and M. Inguscio, Nature 453, 895 (2008).
  • White et al. (2020) D. H. White, T. A. Haase, D. J. Brown, M. D. Hoogerland, M. S. Najafabadi, J. L. Helm, C. Gies, D. Schumayer, and D. A. W. Hutchinson, Nature Communications 11, 4942 (2020).
  • Schwartz et al. (2007) T. Schwartz, G. Bartal, S. Fishman, and M. Segev, Nature 446, 52 (2007).
  • De Raedt et al. (1989) H. De Raedt, A. Lagendijk, and P. de Vries, Phys. Rev. Lett. 62, 47 (1989).
  • Skipetrov and Sokolov (2014) S. E. Skipetrov and I. M. Sokolov, Phys. Rev. Lett. 112, 023905 (2014).
  • Cherroret et al. (2016) N. Cherroret, D. Delande, and B. A. van Tiggelen, Phys. Rev. A 94, 012702 (2016).
  • van Tiggelen and Skipetrov (2021) B. A. van Tiggelen and S. E. Skipetrov, Phys. Rev. B 103, 174204 (2021).
  • Efetov (1997) K. B. Efetov, Sypersymmetry in Disorder and Chaos (Cambridge University Press, Cambridge, 1997).
  • Evers and Mirlin (2008) F. Evers and A. D. Mirlin, Rev. Mod. Phys. 80, 1355 (2008).
  • Ryu et al. (2010) S. Ryu, A. P. Schnyder, A. Furusaki, and A. W. W. Ludwig, New Journal of Physics 12, 65010 (2010).
  • Armitage et al. (2018) N. P. Armitage, E. J. Mele, and A. Vishwanath, Rev. Mod. Phys. 90, 015001 (2018).
  • Altland and Bagrets (2015) A. Altland and D. Bagrets, Phys. Rev. Lett. 114, 257201 (2015).
  • Altland et al. (2002) A. Altland, B. Simons, and M. Zirnbauer, Physics Reports 359, 283 (2002).
  • John (1984) S. John, Phys. Rev. Lett. 53, 2169 (1984).
  • Lagendijk and van Tiggelen (1996) A. Lagendijk and B. A. van Tiggelen, Physics Reports 270, 143 (1996).
  • Bialynicki-Birula (1996) I. Bialynicki-Birula (Elsevier, 1996), vol. 36 of Progress in Optics, pp. 245–294.
  • Mirlin (2000) A. D. Mirlin, Physics Reports 326, 259 (2000).
  • John (1987) S. John, Phys. Rev. Lett. 58, 2486 (1987).
  • John (1985) S. John, Phys. Rev. B 31, 304 (1985).
  • Altland and Bagrets (2016) A. Altland and D. Bagrets, Phys. Rev. B 93, 075113 (2016).
  • Redlich (1984) A. N. Redlich, Phys. Rev. D 29, 2366 (1984).
  • Kerker et al. (1983) M. Kerker, D.-S. Wang, and C. L. Giles, J. Opt. Soc. Am. 73, 765 (1983).
  • García-Etxarri et al. (2011) A. García-Etxarri, R. Gómez-Medina, L. S. Froufe-Pérez, C. López, L. Chantada, F. Scheffold, J. Aizpurua, M. Nieto-Vesperinas, and J. J. Sáenz, Opt. Express 19, 4815 (2011).
  • Geffrin et al. (2012) J. M. Geffrin, B. García-Cámara, R. Gómez-Medina, P. Albella, L. S. Froufe-Pérez, C. Eyraud, A. Litman, R. Vaillon, F. González, M. Nieto-Vesperinas, et al., Nature Communications 3, 1171 (2012).
  • MacKintosh et al. (1989) F. C. MacKintosh, J. X. Zhu, D. J. Pine, and D. A. Weitz, Phys. Rev. B 40, 9342 (1989).
  • Bicout et al. (1994) D. Bicout, C. Brosseau, A. S. Martinez, and J. M. Schmitt, Phys. Rev. E 49, 1767 (1994).
  • Gorodnichev et al. (2017) E. E. Gorodnichev, A. I. Kuzovlev, and D. B. Rogozkin, Journal of Physics: Conference Series 788, 012039 (2017).
  • (44) See e.g. Refs. Bicout et al. (1994); MacKintosh et al. (1989) for studies of circular depolarization in multiple scattering media composed of optically inactive spherical particles, showing large decay lengths in the Mie regime.
  • Gorodnichev et al. (2022) E. E. Gorodnichev, K. A. Kondratiev, and D. B. Rogozkin, Phys. Rev. B 105, 104208 (2022).
  • Lenke et al. (2000) R. Lenke, R. Lehner, and G. Maret, Europhysics Letters 52, 620 (2000).
  • Fradkin (1986) E. Fradkin, Phys. Rev. B 33, 3263 (1986).
  • Klier et al. (2019) J. Klier, I. V. Gornyi, and A. D. Mirlin, Phys. Rev. B 100, 125160 (2019).
  • Pixley and Wilson (2021) J. Pixley and J. H. Wilson, Annals of Physics 435, 168455 (2021).
  • Makhfudz (2018) I. Makhfudz, Scientific Reports 8, 6719 (2018).

Supplemental Material for “Topology Obstructing Anderson Localization of Light”

Tobias Micklitz

Alexander Altland

March 10, 2025

I ‘Dirac’ equation for photons

We start by defining the six component vector Ψ=EH\Psi=E\oplus H, and the diagonal matrix D=ϵ𝟙μ𝟙D=\epsilon\mathds{1}\oplus\mu\mathds{1}, to write Maxwell’s equations as

itDΨ=τ2i∂̸Ψ,\displaystyle i\partial_{t}D\Psi=\tau_{2}i\not{\partial}\Psi,

where τ2\tau_{2} acts in the electric/magnetic subspace of Ψ\Psi, and the action of the differential operator on three-component vectors is defined through the slash notation (∂̸X)k=iϵklmlXm(\not{\partial}X)_{k}=i\epsilon_{klm}\partial_{l}X_{m}, or ∂̸X=SllX\not{\partial}X=S_{l}\partial_{l}X, with the Hermitian matrices (Sl)mkiϵmlk(S_{l})_{mk}\equiv i\epsilon_{mlk}. Assuming harmonic time dependence of Ψ\Psi with characteristic frequency ω0\omega_{0}, the equation reduces to ω0Ψ=D1τ2(i∂̸)Ψ\omega_{0}\Psi=D^{-1}\tau_{2}(-i\not{\partial})\Psi. We next apply a similarity transformation, Ψ=D1/2Φ~\Psi=D^{-1/2}\tilde{\Phi}, to obtain the representation

ω0Φ~=D1/2τ2(i∂̸)D1/2Φ~.\displaystyle\omega_{0}\tilde{\Phi}=D^{-1/2}\tau_{2}(-i\not{\partial})D^{-1/2}\tilde{\Phi}.

The operator on the r.h.s. is manifestly hermitian (safeguarding the existence of solutions with real frequencies ω0\omega_{0}) but inconvenient to work with as the scattering disorder is hiding in the factor matrices DD. We aim to bring it into a more customary form ‘derivative operator + disorder potential’.

We first represent the matrix DD in terms of the refractive index and impedance as D=nexp(lnZτ3)D=n\exp(\ln Z\tau_{3}), leading to

ω0nΦ\displaystyle\omega_{0}n\Phi =τ2e12lnZτ3(i∂̸)e12lnZτ3Φ\displaystyle=\tau_{2}e^{\frac{1}{2}\ln Z\tau_{3}}(-i\not{\partial})e^{-\frac{1}{2}\ln Z\tau_{3}}\Phi
=(i∂̸τ2τ12(∂̸lnZ))Φ,\displaystyle=\left(-i\not{\partial}\tau_{2}-\frac{\tau_{1}}{2}(\not{\partial}\ln Z)\right)\Phi,

where Φ=n1/2Φ~\Phi=n^{-1/2}\tilde{\Phi}. Finally, with V=ω0(1n)V=\omega_{0}(1-n), and Z12∂̸lnZ\not{a}_{Z}\equiv\frac{1}{2}\not{\partial}\ln Z we obtain the desired form of the equation,

ω0Φ=(i∂̸τ2τ1Z,x+Vx)Φ.\displaystyle\omega_{0}\Phi=\left(-i\not{\partial}\tau_{2}-\tau_{1}\not{a}_{Z,\textbf{x}}+V_{\textbf{x}}\right)\Phi.

Finally, performing a unitary rotation around τ1\tau_{1} to the helical eigenstates Φ(Φ+,Φ)\Phi\to(\Phi^{+},\Phi^{-}), we arrive at Dirac-like equation stated in the text.

II Replica field theory

The complex interference processes underlying Anderson localization manifest on length scales exceeding the mean free path. Their description is most efficiently formulated in terms of the diffusion modes of the disordered system. Specifically, interference processes can be organized in terms of diffusion modes scattering off each other, and corresponding interaction vertices are encoded in a nonlinear sigma-model. Starting point of the field theory description is the replica partition function,

𝒵[j]\displaystyle{\cal Z}[j] =Dψeid3xψ¯(iδσ3ra+ω0τ3Vj)ψdis,\displaystyle=\int D\psi\left\langle e^{i\int d^{3}x\,\bar{\psi}(i\delta\sigma_{3}^{\rm ra}+\omega_{0}-\tau_{3}\otimes\not{p}-V-j)\psi}\right\rangle_{\rm dis}, (10)

where ψ={ψs,σ,a,ν,χ(𝕩)}\psi=\{\psi_{s,\sigma,a,\nu,\chi}(\mathbb{x})\} is a (2×2×R×3×2)(2\times 2\times R\times 3\times 2) component Grassmann field. Here s=1,2s=1,2 is a two-component index distinguishing between retarded and advanced indices, σ=1,2\sigma=1,2 has been introduced to account for time-reversal symmetry of the class AI{\rm AI} system, a=1,,Ra=1,\ldots,R is a replica index, ν=1,2,3\nu=1,2,3 parametrizes the spin-1 degree of freedom, χ=±\chi=\pm is the helicity, and σ3ra=δss(1)s+1\sigma_{3}^{\rm ra}=\delta_{ss^{\prime}}(-1)^{s+1} a Pauli matrix causing infinitesimal symmetry breaking in the advanced-retarded sector. Time-reversal symmetry manifests in the relation ψ¯=(iσ2trψ)T\bar{\psi}=(i\sigma^{\rm tr}_{2}\psi)^{T}, with σ2tr\sigma^{\rm tr}_{2} acting in the space of time-reversal indices σ=1,2\sigma=1,2. Differentiation of Eq. (10) with respect to the source jj, generates ensemble averages of products of retarded and advanced Green’s functions entering e.g. the inverse participation ratio I2I_{2} defined in the main text app_MIRLIN2000259 ; app_Efetov-book . We here do not consider jj further, but notice that it can be chosen structureless in helical τ\tau-space.

We next summarize the key construction steps underlying the nonlinear sigma-model: The action in Eq. (10) is invariant under uniform rotations ψTψ\psi\rightarrow T\psi commuting with τ3\tau_{3}\otimes\not{p}. That is, homogeneous rotations that are structureless in spin-indices ν=1,2,3\nu=1,2,3, and diagonal in helical τ\tau-space. We also notice that TT=σ2trTσ2trT^{T}=\sigma_{2}^{\rm tr}T\sigma_{2}^{\rm tr} to guarantee the time-reversal symmetry constraint on ψ¯\bar{\psi}, ψ\psi. Averaging over Gaussian fluctuations of the light potential generates a ψ4\psi^{4}-contribution, local in space and non-local in ‘internal’ indices s,σ,a,ν,χs,\sigma,a,\nu,\chi. The latter is decoupled by a matrix field, playing the role of a self-energy, via Hubbard-Stratonovich transformation. Integration over ψ\psi leads to a representation of 𝒵{\cal Z} entirely in the matrix field, which is subjected to a stationary point analysis. The SCBA self-energy of the previous paragraph provides a saddle point, consistent with causality of the theory. The physics on long length scales is encoded in soft rotations around the latter, iκTσ3T1i\kappa T\sigma_{3}T^{-1}, and corresponding soft mode action is derived from a ‘trace-log’ expansion. We next detail the above outlined construction steps.

III Effective action

Throughout this section, we closely follow the analysis of Ref. appAltlandBagrets2016 on strongly disordered Weyl semimetals. We point out differences for 3d3d light scattering related to the presence of longitudinal and transverse modes when relevant, and refer to the former for further details when differences turn out irrelevant.

III.1 Hubbard-Stratonovich decoupling and mean-field analysis

Starting out from the replica partition function Eq. (10), the Gaussian average over fluctuations V𝕩V_{\mathbb{x}} results in a quartic ‘interaction term’, that is decoupled by a matrix-field via Hubbard-Stratonovich transformation,

eid3xψ¯Vψdis\displaystyle\langle e^{i\int d^{3}x\,\bar{\psi}V\psi}\rangle_{\rm dis} =ew02d3xψ¯ψψ¯ψ=𝑑Qe12w0TrQ2+d3xψ¯Qψ.\displaystyle=e^{-\frac{w_{0}}{2}\int d^{3}x\,\bar{\psi}\psi\bar{\psi}\psi}=\int dQ\,e^{-\frac{1}{2w_{0}}{\rm Tr}Q^{2}+\int d^{3}x\bar{\psi}Q\psi}. (11)

Integration over ψ¯,ψ\bar{\psi},\psi, then gives the alternative representation of the partition function in terms of the matrix-field Z=𝑑QeS[Q]Z=\int dQe^{-S[Q]}, with

S[Q]\displaystyle S[Q] =12w0TrQ212Trln(ω0τ3+iδσ3raj+iQ),\displaystyle=\frac{1}{2w_{0}}{\rm Tr}\,Q^{2}-\frac{1}{2}{\rm Tr}\ln\left(\omega_{0}-\tau_{3}\otimes\not{p}+i\delta\sigma_{3}^{\rm ra}-j+iQ\right), (12)

where the factor 1/21/2 reflects that ψ¯\bar{\psi}, ψ\psi are dependent variables. Variation of the action leads to the saddle point equation,

Q\displaystyle Q =iw02(dp)1ω0τ3+iδσ3ra+iQ,\displaystyle=\frac{iw_{0}}{2}\int(dp)\,\frac{1}{\omega_{0}-\tau_{3}\otimes\not{p}+i\delta\sigma_{3}^{\rm ra}+iQ}, (13)

where we neglected the source jj, and introduced (dp)d3p/(2π)3(dp)\equiv d^{3}p/(2\pi)^{3}. We then focus on the simplest real solution, noting that a finite imaginary part can be absorbed by a renormalization of ω0\omega_{0}. Inserting the ansatz Q=κσ3raQ=\kappa\sigma_{3}^{\rm ra}, dictated by causality, we arrive at

κ\displaystyle\kappa =w02Im(dp)(23ω0+iκ(ω0+iκ)2p2+131ω0+iκ),\displaystyle=-\frac{w_{0}}{2}{\rm Im}\int(dp)\,\left(\frac{2}{3}\frac{\omega_{0}+i\kappa}{(\omega_{0}+i\kappa)^{2}-p^{2}}+\frac{1}{3}\frac{1}{\omega_{0}+i\kappa}\right), (14)

which is Dyson equation for the (imaginary part of the) self-energy in the self-consistent Born approximation. As stated in the main text, we here keep things simple and only consider identical self-energies for transverse and longitudinal modes. Generalization to different self-energies is straightforward, but does not change our main conclusions (see also main text). Integrals Eq. (14) are formally UV divergent, which is related to the approximation of uncorrelated spatial fluctuations w(x)=w0δ(x)w(x)=w_{0}\delta(x). A model with uncorrelated spatial fluctuations is only reasonable when the photon wave length 1/ω01/\omega_{0} exceeds the actual correlation length of fluctuations, and a cut-off Λ\Lambda thus has to be introduced. The solution κ\kappa then depends on the ratios of w0w_{0}, ω0\omega_{0}, and Λ\Lambda. For our purposes, however, it is sufficient to continue with an unspecified general κ\kappa.

III.2 Soft mode action and trace-log expansion

Recalling that soft rotations are diagonal in helical space, we parametrize T=diag(T+,T)T={\rm diag}(T_{+},T_{-}). The effective action controlling the low energy partition function is then a sum of contributions from the two helical modes. That is, 𝒵=D[T+,T]eS[T+]S[T]{\cal Z}=\int D[T_{+},T_{-}]e^{-S[T_{+}]-S[T_{-}]} with

S[Tχ]\displaystyle S[T_{\chi}] =12Trln(ω0χ+iκTχσ3raTχ1).\displaystyle=-\frac{1}{2}{\rm Tr}\ln\left(\omega_{0}-\chi\not{p}+i\kappa T_{\chi}\sigma_{3}^{\rm ra}T_{\chi}^{-1}\right). (15)

Notice that contributions from both helicitites χ=±\chi=\pm are related by the parity transformation 𝕡𝕡\mathbb{p}\leftrightarrow-\mathbb{p}. The soft rotation mode Qχ=Tχσ3raTχ1Q_{\chi}=T_{\chi}\sigma_{3}^{\rm ra}T_{\chi}^{-1} here spans the Goldstone manifold Sp(4R)/Sp(2R)×Sp(2R){\rm Sp}(4R)/{\rm Sp}(2R)\times{\rm Sp}(2R) of class AI{\rm AI} systems, fixed by the symmetry constraint QχT=σ2trQχσ2trQ_{\chi}^{T}=\sigma_{2}^{\rm tr}Q_{\chi}\sigma_{2}^{\rm tr}, and ‘Tr{\rm Tr}’ is the trace over internal degrees of freedom and space. Similarity transformation under the ‘trace-log’, Trln()Trln(Tχ1()Tχ){\rm Tr}\ln\left(\ldots\right)\to{\rm Tr}\ln\left(T_{\chi}^{-1}(\ldots)T_{\chi}\right), then leads to an expression in terms of χχSiAχ,iχSiTχ1iTχ\not{A}_{\chi}\equiv\chi S_{i}A_{\chi,i}\equiv\chi S_{i}T_{\chi}^{-1}\partial_{i}T_{\chi}, which can be exposed to a gradient-expansion. However, to apply the transformation, we need to account for UV-divergencies typical for Dirac-like operators. Following the regularization scheme of Refs. appAltlandBagrets2016 ; appALTLAND2002283 this leads to

S[Tχ]\displaystyle S[T_{\chi}] =12Trln(ω0+iκσ3raχiχ)Sη[Tχ],\displaystyle=-\frac{1}{2}{\rm Tr}\ln\left(\omega_{0}+i\kappa\sigma_{3}^{\rm ra}-\chi\not{p}-i\not{A}_{\chi}\right)-S_{\eta}[T_{\chi}], (16)

where Sη[Tχ]=12Trln(iησ3raiχ)S_{\eta}[T_{\chi}]=\frac{1}{2}{\rm Tr}\ln\left(i\eta\sigma_{3}^{\rm ra}-\not{p}-i\not{A}_{\chi}\right) with η0\eta\to 0 is introduced to regularize divergencies.

We then identify the inverse SCBA Green’s function, Gχ1=ω0+iκσ3raχG_{\chi}^{-1}=\omega_{0}+i\kappa\sigma_{3}^{\rm ra}-\chi\not{p}, and expose the action to a gradient expansion of the trace-log to third order in GχχG_{\chi}\not{A}_{\chi}. For simplicity, we now concentrate on the χ=+\chi=+ mode, drop the index χ\chi, and obtain the result for the χ=\chi=- mode by parity transformation 𝕡𝕡\mathbb{p}\mapsto-\mathbb{p}. We then notice that for n3n\leq 3 a decomposition in transverse and longitudinal components of the Green’s function gives

tr([G]n)\displaystyle{\rm tr}\left([G\not{A}]^{n}\right) =tr((1nπ¯𝕡)[Gt]n)+ntr(π¯𝕡Gl[Gt]n1),\displaystyle={\rm tr}\left((1-n\bar{\pi}_{\mathbb{p}})[G_{t}\not{A}]^{n}\right)+n{\rm tr}\left(\bar{\pi}_{\mathbb{p}}G_{l}\not{A}[G_{t}\not{A}]^{n-1}\right), (17)

with

Gps\displaystyle G^{s}_{p} =Gt,psπ𝕡+Gl,psπ¯𝕡,Gt,ps=Nps(ωs+),Gl,ps=1ωs,\displaystyle=G^{s}_{t,p}\pi_{\mathbb{p}}+G^{s}_{l,p}\bar{\pi}_{\mathbb{p}},\qquad G^{s}_{t,p}=N^{s}_{p}\left(\omega_{s}+\not{p}\right),\qquad G^{s}_{l,p}=\frac{1}{\omega_{s}}, (18)

where ωs=ω0+isκ\omega_{s}=\omega_{0}+is\kappa and Nps=(ωs2p2)1N^{s}_{p}=(\omega_{s}^{2}-p^{2})^{-1}. Specifically, for n=3n=3 we can substitute π¯𝕡=13𝟙𝟛\bar{\pi}_{\mathbb{p}}=\frac{1}{3}\openone_{3} in the first term, and find that

tr([G]3)\displaystyle{\rm tr}\left([G\not{A}]^{3}\right) =3tr(π¯𝕡Gl[Gt]2).\displaystyle=3{\rm tr}\left(\bar{\pi}_{\mathbb{p}}G_{l}\not{A}[G_{t}\not{A}]^{2}\right). (19)

That is, only triangle graphs involving one longitudinal and two transverse modes contribute, as stated in the main text.

III.2.1 Linear order

The first order expansion in GG\not{A} can be organized as,

S(1)\displaystyle S^{(1)} =i2tr(G)=S0(1)+S1(1),\displaystyle=-\frac{i}{2}{\rm tr}\left(G\not{A}\right)=S_{0}^{(1)}+S_{1}^{(1)}, (20)

where

S0(1)\displaystyle S_{0}^{(1)} =i2sfisd3xtr(PsAi),fis=(dp)tr(GpsSi),\displaystyle=-\frac{i}{2}\sum_{s}f_{i}^{s}\int d^{3}x\,{\rm tr}\left(P^{s}A_{i}\right),\qquad f_{i}^{s}=\int(dp)\,{\rm tr}\left(G^{s}_{p}S_{i}\right), (21)
S0(2)\displaystyle S_{0}^{(2)} =i2sfikssd3xtr(PsAkAi),fikss=i2(dp)tr(GpsSkGpsSi),\displaystyle=\frac{i}{2}\sum_{s}f_{ik}^{ss}\int d^{3}x\,{\rm tr}\left(P^{s}A_{k}A_{i}\right),\qquad f_{ik}^{ss}=-\frac{i}{2}\int(dp)\,{\rm tr}\left(G^{s}_{p}S_{k}G^{s}_{p}S_{i}\right), (22)

and we employed that xkAi=AkAi\partial_{x_{k}}A_{i}=-A_{k}A_{i} and pkGps=GpsSkGps\partial_{p_{k}}G^{s}_{p}=G_{p}^{s}S_{k}G_{p}^{s}.

III.2.2 Second order

To second order expansion of the trace-log in GG\not{A},

S(2)\displaystyle S^{(2)} =14tr(GG)=S0(2)+S1(2),\displaystyle=-\frac{1}{4}{\rm tr}\left(G\not{A}G\not{A}\right)=S_{0}^{(2)}+S_{1}^{(2)}, (23)

with

S0(2)\displaystyle S_{0}^{(2)} =14ssfssijd3xtr(PsAiPsAj),fssij=(dp)tr(GpsSiGpsSj),\displaystyle=-\frac{1}{4}\sum_{ss^{\prime}}f^{ij}_{ss^{\prime}}\int d^{3}x\,{\rm tr}\left(P^{s}A_{i}P^{s^{\prime}}A_{j}\right),\qquad f^{ij}_{ss^{\prime}}=\int(dp)\,{\rm tr}\left(G_{p}^{s}S_{i}G_{p}^{s^{\prime}}S_{j}\right), (24)
S1(2)\displaystyle S_{1}^{(2)} =i4ssfssijkd3xtr(Ps[xkAi]PsAj),fssijk=(dp)tr([pkGps]SiGpsSj).\displaystyle=\frac{i}{4}\sum_{ss^{\prime}}f^{ijk}_{ss^{\prime}}\int d^{3}x\,{\rm tr}\left(P^{s}[\partial_{x_{k}}A_{i}]P^{s^{\prime}}A_{j}\right),\qquad f^{ijk}_{ss^{\prime}}=\int(dp)\,{\rm tr}\left([\partial_{p_{k}}G_{p}^{s}]S_{i}G_{p}^{s^{\prime}}S_{j}\right). (25)

III.2.3 Third order

To third order expansion of the trace-log in GG\not{A},

S(3)\displaystyle S^{(3)} =i6tr(GGG),\displaystyle=\frac{i}{6}{\rm tr}\left(G\not{A}G\not{A}G\not{A}\right), (26)

where for our purposes it is sufficient to approximate,

S0(3)=i6s1s2s3fijks1s2s3d3xtr(Ps1AiPs2AjPs3Ak),fijks1s2s3=(dp)tr(Gps1SiGps2SjGps3Sk).\displaystyle S_{0}^{(3)}=\frac{i}{6}\sum_{s_{1}s_{2}s_{3}}f_{ijk}^{s_{1}s_{2}s_{3}}\int d^{3}x\,{\rm tr}\left(P^{s_{1}}A_{i}P^{s_{2}}A_{j}P^{s_{3}}A_{k}\right),\qquad f_{ijk}^{s_{1}s_{2}s_{3}}=\int(dp)\,{\rm tr}\left(G^{s_{1}}_{p}S_{i}G^{s_{2}}_{p}S_{j}G^{s_{3}}_{p}S_{k}\right). (27)

III.3 Traces

We next perform traces, using that

tr(SiSjSk)\displaystyle{\rm tr}\left(S_{i}S_{j}S_{k}\right) =iϵijk,tr(π¯𝕡SiSjSk)=i3ϵijk,tr(SiSjSk)=ip23ϵijk,tr(π¯𝕡SiSjSk)=ip23ϵijk.\displaystyle=i\epsilon_{ijk},\quad{\rm tr}\left(\bar{\pi}_{\mathbb{p}}S_{i}S_{j}S_{k}\right)=\frac{i}{3}\epsilon_{ijk},\quad{\rm tr}\left(S_{i}\not{p}S_{j}\not{p}S_{k}\right)=\frac{ip^{2}}{3}\epsilon_{ijk},\quad{\rm tr}\left(\bar{\pi}_{\mathbb{p}}S_{i}\not{p}S_{j}\not{p}S_{k}\right)=\frac{ip^{2}}{3}\epsilon_{ijk}. (28)

We further employ low order expansions of Moyal products of functions AA and BB depending only on coordinates, respectively, momenta

(AB)(𝕩,𝕡)\displaystyle(AB)(\mathbb{x},\mathbb{p}) =A(𝕩)B(𝕡)+i2𝕩A(𝕩)𝕡B(𝕡)+,\displaystyle=A(\mathbb{x})B(\mathbb{p})+\frac{i}{2}\partial_{\mathbb{x}}A(\mathbb{x})\partial_{\mathbb{p}}B(\mathbb{p})+..., (29)
(BA)(𝕩,𝕡)\displaystyle(BA)(\mathbb{x},\mathbb{p}) =A(𝕩)B(𝕡)i2𝕩A(𝕩)𝕡B(𝕡)+.\displaystyle=A(\mathbb{x})B(\mathbb{p})-\frac{i}{2}\partial_{\mathbb{x}}A(\mathbb{x})\partial_{\mathbb{p}}B(\mathbb{p})+...\,. (30)

It is convenient to start with the expression resulting from the third order trace-log expansion and then turn to the second and first order contributions.

III.3.1 Third order term

Separating different contributions from longitudinal and transverse modes, we introduce

fijks1s2s3\displaystyle f_{ijk}^{s_{1}s_{2}s_{3}} =(dp)(fijk,tts1s2s3+fijk,lts1s2s3),\displaystyle=\int(dp)\,\left(f_{ijk,tt}^{s_{1}s_{2}s_{3}}+f_{ijk,lt}^{s_{1}s_{2}s_{3}}\right), (31)
fijk,tts1s2s3\displaystyle f_{ijk,tt}^{s_{1}s_{2}s_{3}} =tr((13π¯𝕡)Gt,ps1SiGt,ps2SjGt,ps3Sk),fijk,lts1s2s3=3tr(π¯𝕡Gl,ps1AiGt,ps2SjGt,ps3Sk),\displaystyle={\rm tr}\left((1-3\bar{\pi}_{\mathbb{p}})G_{t,p}^{s_{1}}S_{i}G_{t,p}^{s_{2}}S_{j}G_{t,p}^{s_{3}}S_{k}\right),\qquad f_{ijk,lt}^{s_{1}s_{2}s_{3}}=3{\rm tr}\left(\bar{\pi}_{\mathbb{p}}G_{l,p}^{s_{1}}A_{i}G_{t,p}^{s_{2}}S_{j}G_{t,p}^{s_{3}}S_{k}\right), (32)

where the factor three in 3π¯𝕡3\bar{\pi}_{\mathbb{p}} in the last term sums the three contributions related by symmetrization under cyclic exchange of (s1,i)(s_{1},i), (s2,j)(s_{2},j), and (s3,k)(s_{3},k). We then find that

fijk,tts1s2s3\displaystyle f_{ijk,tt}^{s_{1}s_{2}s_{3}} =0,\displaystyle=0, (33)

and

fijk,lts1s2s3\displaystyle f_{ijk,lt}^{s_{1}s_{2}s_{3}} =3tr(π¯𝕡Gl,ps1SiGt,ps2SjGt,ps3Sk)\displaystyle=3{\rm tr}\left(\bar{\pi}_{\mathbb{p}}G_{l,p}^{s_{1}}S_{i}G_{t,p}^{s_{2}}S_{j}G_{t,p}^{s_{3}}S_{k}\right)
=3ωs1Nps2Nps3(ωs2ωs3tr(π¯𝕡SiSjSk)+tr(π¯𝕡SiSjSk))\displaystyle=\frac{3}{\omega_{s_{1}}}N_{p}^{s_{2}}N_{p}^{s_{3}}\left(\omega_{s_{2}}\omega_{s_{3}}{\rm tr}\left(\bar{\pi}_{\mathbb{p}}S_{i}S_{j}S_{k}\right)+{\rm tr}\left(\bar{\pi}_{\mathbb{p}}S_{i}\not{p}S_{j}\not{p}S_{k}\right)\right)
=iϵijkωs1Nps2Nps3(ωs2ωs3+p2),\displaystyle=\frac{i\epsilon_{ijk}}{\omega_{s_{1}}}N_{p}^{s_{2}}N_{p}^{s_{3}}\left(\omega_{s_{2}}\omega_{s_{3}}+p^{2}\right), (34)

leading to

fijks1s2s3\displaystyle f_{ijk}^{s_{1}s_{2}s_{3}} =iϵijkωs1(dp)Nps2Nps3(ωs2ωs3+p2).\displaystyle=\frac{i\epsilon_{ijk}}{\omega_{s_{1}}}\int(dp)\,N_{p}^{s_{2}}N_{p}^{s_{3}}\left(\omega_{s_{2}}\omega_{s_{3}}+p^{2}\right). (35)

Notice that upon symmetrization under cyclic exchange of (s1,i)(s_{1},i), (s2,j)(s_{2},j),

fijks1s2s3\displaystyle f_{ijk}^{s_{1}s_{2}s_{3}} =i3ϵijk(dp)(Nps2Nps3ωs2ωs3+p2ωs1+Nps3Nps1ωs3ωs1+p2ωs2+Nps1Nps1ωs1ωs2+p2ωs3).\displaystyle=\frac{i}{3}\epsilon_{ijk}\int(dp)\,\left(N_{p}^{s_{2}}N_{p}^{s_{3}}\frac{\omega_{s_{2}}\omega_{s_{3}}+p^{2}}{\omega_{s_{1}}}+N_{p}^{s_{3}}N_{p}^{s_{1}}\frac{\omega_{s_{3}}\omega_{s_{1}}+p^{2}}{\omega_{s_{2}}}+N_{p}^{s_{1}}N_{p}^{s_{1}}\frac{\omega_{s_{1}}\omega_{s_{2}}+p^{2}}{\omega_{s_{3}}}\right). (36)

III.3.2 Second order term

Continuing with the term S0(2)S_{0}^{(2)} found in second order trace-log expansion, we write

fssij\displaystyle f^{ij}_{ss^{\prime}} =(dp)(fss,ttij+fss,ltij),fss,ttij=tr((12π¯𝕡)Gt,psSiGt,psSj),fss,ltij=2tr(π¯𝕡Gl,psSiGt,psSj),\displaystyle=\int(dp)\,\left(f^{ij}_{ss^{\prime},tt}+f^{ij}_{ss^{\prime},lt}\right),\qquad f^{ij}_{ss^{\prime},tt}={\rm tr}\left((1-2\bar{\pi}_{\mathbb{p}})G_{t,p}^{s}S_{i}G_{t,p}^{s^{\prime}}S_{j}\right),\quad f^{ij}_{ss^{\prime},lt}=2{\rm tr}\left(\bar{\pi}_{\mathbb{p}}G_{l,p}^{s}S_{i}G_{t,p}^{s^{\prime}}S_{j}\right), (37)

and use that

fss,ttij\displaystyle f^{ij}_{ss^{\prime},tt} =NpsNpstr((12π¯p)ωsSiωsSj+pm2SmSiSmSj)δij\displaystyle=N_{p}^{s}N_{p}^{s^{\prime}}{\rm tr}\left((1-2\bar{\pi}_{p})\omega_{s}S_{i}\omega_{s^{\prime}}S_{j}+p_{m}^{2}S_{m}S_{i}S_{m}S_{j}\right)\delta_{ij}
=13NpsNpstr((ωsωs+p2)Si2)δij\displaystyle=\frac{1}{3}N_{p}^{s}N_{p}^{s^{\prime}}{\rm tr}\left((\omega_{s}\omega_{s^{\prime}}+p^{2})S_{i}^{2}\right)\delta_{ij}
=23NpsNps(ωsωs+p2)δij,\displaystyle=\frac{2}{3}N_{p}^{s}N_{p}^{s^{\prime}}\left(\omega_{s}\omega_{s^{\prime}}+p^{2}\right)\delta_{ij}, (38)

and

fss,ltij\displaystyle f^{ij}_{ss^{\prime},lt} =2Npstr(π¯𝕡Gl,psSiωsSj)=4ωs3ωsNpsδij,\displaystyle=2N_{p}^{s^{\prime}}{\rm tr}\left(\bar{\pi}_{\mathbb{p}}G_{l,p}^{s}S_{i}\omega_{s^{\prime}}S_{j}\right)=\frac{4\omega_{s^{\prime}}}{3\omega_{s}}N_{p}^{s^{\prime}}\delta_{ij}, (39)

to arrive at

fssij\displaystyle f_{ss^{\prime}}^{ij} =2(dp)NpsNps(ωsωs+p23(12ωsωs)).\displaystyle=2\int(dp)\,N_{p}^{s}N_{p}^{s^{\prime}}\left(\omega_{s}\omega_{s^{\prime}}+\frac{p^{2}}{3}\left(1-\frac{2\omega_{s^{\prime}}}{\omega_{s}}\right)\right). (40)

For the contribution S1(2)S_{1}^{(2)}, found in second order trace-log expansion, we notice that

pkGps\displaystyle\partial_{p_{k}}G^{s}_{p} =GpsSkGps,\displaystyle=G^{s}_{p}S_{k}G^{s}_{p}, (41)

and thus

fssijk=(dp)tr(GpsSkGpsSiGpsSj)=fijksss,\displaystyle f^{ijk}_{ss^{\prime}}=\int(dp)\,{\rm tr}\left(G_{p}^{s}S_{k}G_{p}^{s}S_{i}G_{p}^{s^{\prime}}S_{j}\right)=f^{sss^{\prime}}_{ijk}, (42)

with fijks1s2s3f^{s_{1}s_{2}s_{3}}_{ijk} from the third order trace-log expansion calculated above.

III.3.3 First order term

First order terms vanish.

III.4 Integrations

For convenience, we summarize all relevant terms found above,

fijss\displaystyle f_{ij}^{ss} =2(dp)NpsNps(ωsωsp23),\displaystyle=2\int(dp)\,N_{p}^{s}N_{p}^{s}\left(\omega_{s}\omega_{s}-\frac{p^{2}}{3}\right), (43)
fij++fij+\displaystyle f_{ij}^{+-}+f_{ij}^{-+} =4(dp)Np+Np(ω02+κ2+p23p23ω02κ2ω02+κ2),\displaystyle=4\int(dp)\,N_{p}^{+}N_{p}^{-}\left(\omega_{0}^{2}+\kappa^{2}+\frac{p^{2}}{3}-\frac{p^{2}}{3}\frac{\omega_{0}^{2}-\kappa^{2}}{\omega_{0}^{2}+\kappa^{2}}\right), (44)
fijk++\displaystyle f_{ijk}^{++-} =i3ϵijk(dp)(2Np+Np(ω+p2ω+)+Np+Np+ω+ω++p2ω),\displaystyle=\frac{i}{3}\epsilon_{ijk}\int(dp)\,\left(2N_{p}^{+}N_{p}^{-}\left(\omega_{-}+\frac{p^{2}}{\omega_{+}}\right)+N_{p}^{+}N_{p}^{+}\frac{\omega_{+}\omega_{+}+p^{2}}{\omega_{-}}\right), (45)
fijk+\displaystyle f_{ijk}^{--+} =i3ϵijk(dp)(2NpNp+(ω++p2ω)+NpNpωω+p2ω+),\displaystyle=\frac{i}{3}\epsilon_{ijk}\int(dp)\,\left(2N_{p}^{-}N_{p}^{+}\left(\omega_{+}+\frac{p^{2}}{\omega_{-}}\right)+N_{p}^{-}N_{p}^{-}\frac{\omega_{-}\omega_{-}+p^{2}}{\omega_{+}}\right), (46)
fijk+++\displaystyle f_{ijk}^{+++} =iϵijk(dp)Np+Np+(ω++p2ω+),\displaystyle=i\epsilon_{ijk}\int(dp)\,N_{p}^{+}N_{p}^{+}\left(\omega_{+}+\frac{p^{2}}{\omega_{+}}\right), (47)
fijk\displaystyle f_{ijk}^{---} =iϵijk(dp)NpNp(ω+p2ω),\displaystyle=i\epsilon_{ijk}\int(dp)\,N_{p}^{-}N_{p}^{-}\left(\omega_{-}+\frac{p^{2}}{\omega_{-}}\right), (48)

which are next evaluated using that

(dp)[Nps]2p2\displaystyle\int(dp)\,[N_{p}^{s}]^{2}p^{2} 3isωs8π,(dp)Np+Npp2ω023κ28πκ,(dp)[Nps]2=is8πωs,(dp)Np+Np=18πκ,\displaystyle\to\frac{3is\omega_{s}}{8\pi},\qquad\int(dp)\,N_{p}^{+}N_{p}^{-}p^{2}\to\frac{\omega_{0}^{2}-3\kappa^{2}}{8\pi\kappa},\qquad\int(dp)\,[N_{p}^{s}]^{2}=\frac{is}{8\pi\omega_{s}},\qquad\int(dp)\,N_{p}^{+}N_{p}^{-}=\frac{1}{8\pi\kappa}, (49)

where the first two integrals used regularization by SηS_{\eta} in the limit η0\eta\to 0, see Ref. appAltlandBagrets2016 for further details.

We then find

fijss\displaystyle f^{ss}_{ij} =0,fij++fij+=(ω02+3κ2)(ω0213κ2)2πκ(ω02+κ2),\displaystyle=0,\qquad f_{ij}^{+-}+f_{ij}^{-+}=\frac{(\omega_{0}^{2}+3\kappa^{2})\left(\omega_{0}^{2}-\frac{1}{3}\kappa^{2}\right)}{2\pi\kappa(\omega_{0}^{2}+\kappa^{2})}, (50)

and

fijk++\displaystyle f_{ijk}^{++-} =iϵijk6π(ω02κ2κω++iω+ω),fijk+=iϵijk6π(ω02κ2κωiωω+),fijk+++=ϵijk2π,fijk=ϵijk2π,\displaystyle=\frac{i\epsilon_{ijk}}{6\pi}\left(\frac{\omega_{0}^{2}-\kappa^{2}}{\kappa\omega_{+}}+\frac{i\omega_{+}}{\omega_{-}}\right),\quad f_{ijk}^{--+}=\frac{i\epsilon_{ijk}}{6\pi}\left(\frac{\omega_{0}^{2}-\kappa^{2}}{\kappa\omega_{-}}-\frac{i\omega_{-}}{\omega_{+}}\right),\quad f_{ijk}^{+++}=-\frac{\epsilon_{ijk}}{2\pi},\quad f_{ijk}^{---}=\frac{\epsilon_{ijk}}{2\pi}, (51)

and notice that

fijk+++fijk++\displaystyle f^{++-}_{ijk}+f^{-++}_{ijk} =0.\displaystyle=0. (52)

III.5 Action

We now have all ingredients to derive the different contributions to the effective action.

III.5.1 First order

We first recall that first order terms vanish, S0(1)=S1(1)=0S_{0}^{(1)}=S_{1}^{(1)}=0.

III.5.2 First term from second order

The leading contribution from second order is,

S0(2)\displaystyle S_{0}^{(2)} =(ω02+3κ2)(ω0213κ2)32πκ(ω02+κ2)d3xtr(AiAiσ3raAiσ3raAi).\displaystyle=-\frac{(\omega_{0}^{2}+3\kappa^{2})\left(\omega_{0}^{2}-\frac{1}{3}\kappa^{2}\right)}{32\pi\kappa(\omega_{0}^{2}+\kappa^{2})}\int d^{3}x\,{\rm tr}\left(A_{i}A_{i}-\sigma_{3}^{\rm ra}A_{i}\sigma_{3}^{\rm ra}A_{i}\right). (53)

Noting that T[Ai,σ3ra]T1=iQT[A_{i},\sigma^{\rm ra}_{3}]T^{-1}=\partial_{i}Q, with Q=Tσ3raT1Q=T\sigma_{3}^{\rm ra}T^{-1}, this leads to Eq. (13) in the main text.

III.5.3 Remaining contributions

The remaining second and third order term can be combined to

S1(2)+S0(3)\displaystyle S_{1}^{(2)}+S_{0}^{(3)} =i4fijk++d3xtr(P+AiAjPAk2P+AiP+AjPAk)\displaystyle=-\frac{i}{4}f^{++-}_{ijk}\int d^{3}x\,{\rm tr}\left(P^{+}A_{i}A_{j}P^{-}A_{k}-2P^{+}A_{i}P^{+}A_{j}P^{-}A_{k}\right)
i4fijk+d3xtr(PAiAjP+Ak2PAiPAjP+Ak)\displaystyle\quad-\frac{i}{4}f^{--+}_{ijk}\int d^{3}x\,{\rm tr}\left(P^{-}A_{i}A_{j}P^{+}A_{k}-2P^{-}A_{i}P^{-}A_{j}P^{+}A_{k}\right)
i4fijk+++d3xtr(P+AiAjP+Ak23P+AiP+AjP+Ak)\displaystyle\quad-\frac{i}{4}f^{+++}_{ijk}\int d^{3}x\,{\rm tr}\left(P^{+}A_{i}A_{j}P^{+}A_{k}-\frac{2}{3}P^{+}A_{i}P^{+}A_{j}P^{+}A_{k}\right)
i4fijkd3xtr(PAiAjPAk23PAiPAjPAk).\displaystyle\quad-\frac{i}{4}f^{---}_{ijk}\int d^{3}x\,{\rm tr}\left(P^{-}A_{i}A_{j}P^{-}A_{k}-\frac{2}{3}P^{-}A_{i}P^{-}A_{j}P^{-}A_{k}\right). (54)

We then notice that in the above expression the first two contributions simplify to

fijk++tr(P+AiAjPAk2P+AiP+AjPAk)+fijk+tr(PAiAjP+Ak2PAiPAjP+Ak)\displaystyle f^{++-}_{ijk}{\rm tr}\left(P^{+}A_{i}A_{j}P^{-}A_{k}-2P^{+}A_{i}P^{+}A_{j}P^{-}A_{k}\right)+f^{--+}_{ijk}{\rm tr}\left(P^{-}A_{i}A_{j}P^{+}A_{k}-2P^{-}A_{i}P^{-}A_{j}P^{+}A_{k}\right)
=14(fijk+++fijk+)tr(σ3raAiAjAkσ3raAiσ3raAjσ3raAk)=0,\displaystyle=-\frac{1}{4}\left(f^{++-}_{ijk}+f^{--+}_{ijk}\right){\rm tr}\left(\sigma_{3}^{\rm ra}A_{i}A_{j}A_{k}-\sigma_{3}^{\rm ra}A_{i}\sigma_{3}^{\rm ra}A_{j}\sigma_{3}^{\rm ra}A_{k}\right)=0, (55)

where to pass from the first to second line we used invariance of fijks1s2s3f^{s_{1}s_{2}s_{3}}_{ijk} under cyclic permutations of i,j,ki,j,k. The last two contributions, on the other hand,

fijk+++tr(P+AiAjP+Ak23P+AiP+AjP+Ak)+fijktr(PAiAjPAk23PAiPAjPAk)\displaystyle f^{+++}_{ijk}{\rm tr}\left(P^{+}A_{i}A_{j}P^{+}A_{k}-\frac{2}{3}P^{+}A_{i}P^{+}A_{j}P^{+}A_{k}\right)+f^{---}_{ijk}{\rm tr}\left(P^{-}A_{i}A_{j}P^{-}A_{k}-\frac{2}{3}P^{-}A_{i}P^{-}A_{j}P^{-}A_{k}\right)
=12fijk+++tr(σ3raAiAjAk13σ3raAiσ3raAjσ3raAk),\displaystyle=\frac{1}{2}f^{+++}_{ijk}{\rm tr}\left(\sigma_{3}^{\rm ra}A_{i}A_{j}A_{k}-\frac{1}{3}\sigma_{3}^{\rm ra}A_{i}\sigma_{3}^{\rm ra}A_{j}\sigma_{3}^{\rm ra}A_{k}\right), (56)

where we used that fijk=fijk+++f^{---}_{ijk}=-f^{+++}_{ijk} and invariance under cyclic permutations of i,j,ki,j,k. We thus arrive at the contribution,

S1(2)+S0(3)\displaystyle S_{1}^{(2)}+S_{0}^{(3)} =iϵijk16πd3xtr(σ3raAiAjAk13σ3raAiσ3raAjσ3raAk),\displaystyle=-\frac{i\epsilon_{ijk}}{16\pi}\int d^{3}x\,{\rm tr}\left(\sigma_{3}^{\rm ra}A_{i}A_{j}A_{k}-\frac{1}{3}\sigma_{3}^{\rm ra}A_{i}\sigma_{3}^{\rm ra}A_{j}\sigma_{3}^{\rm ra}A_{k}\right), (57)

which, using ϵijkAjAk=ϵijkjAk\epsilon_{ijk}A_{j}A_{k}=-\epsilon_{ijk}\partial_{j}A_{k}, gives Eq. (14) stated in the main text.

Finally we notice that employing different self-energies for transverse and longitudinal modes modifies the coupling σ\sigma of the diffusive sigma model action, but does not modify the topological term within the accuracy of the weak disorder approximation κω0\kappa\ll\omega_{0}.

IV Helicity breaking

Adding the helicity-mixing contribution τ1Z\tau_{1}\not{a}_{Z} from a locally varying impedance, we consider

S±[T]\displaystyle S_{\pm}[T] =12Trln(ω0+iκσ3raτ3+Z),\displaystyle=-\frac{1}{2}{\rm Tr}\ln\left(\omega_{0}+i\kappa\sigma_{3}^{\rm ra}-\tau_{3}\not{p}+\not{A}_{Z}\right), (58)

with ZT1τ1ZT\not{A}_{Z}\equiv T^{-1}\tau_{1}\not{a}_{Z}T, and we dropped the contribution χ\not{A}_{\chi} discussed in the previous sections. Expansion of the ‘Tr ln’ to quadratic order we use aZ,i,xaZ,j,y=γδijδ(xy)\langle a_{Z,i,\textbf{x}}a_{Z,j,\textbf{y}}\rangle=\gamma\delta_{ij}\delta(\textbf{x}-\textbf{y}), and arrive at

S±[T]\displaystyle S_{\pm}[T] =γ4d3xtr(TG(𝕩,𝕩)T1τ1SiTG(𝕩,𝕩)T1τ1Si)\displaystyle=-\frac{\gamma}{4}\int d^{3}x\,{\rm tr}\left(TG(\mathbb{x},\mathbb{x})T^{-1}\tau_{1}S_{i}TG(\mathbb{x},\mathbb{x})T^{-1}\tau_{1}S_{i}\right)
=(2κw0)2×3γ4d3xtr(Qτ1Qτ1),\displaystyle=\left(\frac{2\kappa}{w_{0}}\right)^{2}\times\frac{3\gamma}{4}\int d^{3}x\,{\rm tr}\left(Q\tau_{1}Q\tau_{1}\right), (59)

where in passing from the first to the second line we employed the mean-field equation Eq. (14). Tracing over helical space we then arrive at Eq. (9) in the main text.

References

  • (1) Alexander Altland, and Dmitry Bagrets, “Theory of the strongly disordered Weyl semimetal”, Phys. Rev. B 93, 075113 (2016).
  • (2) Alexander D. Mirlin, “Statistics of energy levels and eigenfunctions in disordered systems”, Physics Reports 326, 259 (2000).
  • (3) K. B. Efetov, “Sypersymmetry in Disorder and Chaos”, Cambridge University Press (1997).
  • (4) Alexander Altland, B.D. Simons and M.R. Zirnbauer, “Theories of low-energy quasi-particle states in disordered d-wave superconductors”, Physics Reports 359, 283 (2002).