Manipulating Type-I and Type-II Dirac in Cavity-Embedded Honeycomb Metasurfaces

Charlie-Ray Mann,1 Thomas J. Sturges,1 Guillaume Weick,2 William L. Barnes,1 and Eros Mariani1 1EPSRC Centre for Doctoral Training in Metamaterials (XM2), Department of Physics and Astronomy, University of Exeter, Exeter, EX4 4QL, United Kingdom. 2Universite´ de Strasbourg, CNRS, Institut de Physique et Chimie des Materiaux´ de Strasbourg, UMR 7504, F-67000 Strasbourg, France.

Pseudorelativistic Dirac quasiparticles have emerged in a plethora of artificial systems that mimic the underlying honeycomb symmetry of graphene. However, it is notoriously difficult to manipulate their properties without modifying the lattice structure. Here we theoretically investigate polaritons supported by honeycomb metasurfaces and, despite the trivial nature of the resonant elements, we unveil rich Dirac physics stemming from a non-trivial winding in the light-matter interaction. A new kind of type-II Dirac point emerges which simultaneously exists with its conventional type-I counterpart. By modifying only the photonic environment via an enclosing cavity, one can manipulate the location of the type-II Dirac points, leading to distinct phases. This enables one to alter the fundamental properties of the emergent Dirac polaritons while preserv- ing the lattice structure—a unique scenario which has no analog in real or artificial graphene systems so far. Exploiting the photonic environment will thus give rise to unexplored Dirac physics at the subwavelength scale. arXiv:1707.04503v2 [physics.optics] 4 Mar 2018 2

The groundbreaking discovery of monolayer graphene [1] has inspired an extensive quest to emulate massless Dirac quasi- particles in a myriad of distinct artificial graphene systems [2–11], ranging from ultracold atoms in optical lattices [3] to evanescently coupled photonic waveguide arrays [4]. Owing to their honeycomb symmetry, linear band-degeneracies man- ifest in the quasiparticle spectrum which we call conventional Dirac points (CDPs). These belong to the ubiquitous type- I class of two-dimensional (2D) Dirac points that are char- acterized by Dirac cones with closed isofrequency contours. As a result, the corresponding quasiparticles are described by the rather exotic 2D massless Dirac Hamiltonian [12], in stark contrast to the more conventional Schrodinger¨ equation. Thus, these emergent Dirac quasiparticles offer fundamental insight into pseudorelativistic phenomena such as the iconic Klein Figure 1. Schematic of a cavity-embedded honeycomb metasur- paradox, corresponding to the perfect transmission through ar- face. The honeycomb array of meta-atoms is composed of two in- bitrarily large potential barriers [13]. The latter is responsible equivalent (A and√ B) hexagonal sublattices—defined√ by lattice vec- for the suppression of backscattering and antilocalization of 3 3 3 3 tors a1 = a(− , ) and a2 = a( , )—which are connected 2 2 2 2 √ Dirac quasiparticles, which are highly desirable properties for 3 1 by nearest-neighbor vectors e1 = a(0, −1), e2 = a( , ) and √ 2 2 efficient quasiparticle propagation in novel devices. 3 1 e3 = a(− 2 , 2 ), where a is the subwavelength nearest-neighbor Since the existence of type-I CDPs is intrinsically linked separation. Each meta-atom is modelled as an electric dipole, ori- to the honeycomb structure, the fundamental properties of the ented normal to the plane of the lattice. The honeycomb metasurface massless Dirac quasiparticles are notoriously robust and diffi- is embedded inside a photonic cavity of height L, which is composed cult to manipulate. However, by exploiting meticulous control of two perfectly conducting metallic plates, enabling one to modify over the lattice structure, artificial graphene systems have en- the photonic environment while preserving the lattice structure. This abled the exploration of Dirac quasiparticles in new regimes general model can be realized across the electromagnetic spectrum, that are difficult, if not impossible to achieve in graphene from arrays of plasmonic nanorods to microwave helical resonators. itself [14–19]. An archetypal example which has attracted considerable interest is the paradigm of strain-engineering, where it has been shown that lattice anisotropy can induce the non-local interactions, mediated by the electromagnetic en- merging and annihilation of type-I CDPs [3, 14–16, 20–23], vironment, give rise to rich and unique Dirac physics at the and that aperiodicity can generate large pseudomagnetic fields subwavelength scale. In particular, we unveil the emergence [17, 24]. of a new kind of type-II Dirac point, which simultaneously ex- This concept of Hamiltonian engineering, and the hunt for ists with the type-I CDPs. In contrast to the latter, the former exotic quasiparticles, has recently entered the realm of polari- are characterized by critically tilted Dirac cones with open, tonics [25–30]. The true potential of polaritons lies in their hyperbolic isofrequency contours. hybrid nature, where their light and matter constituents can Recently, the discovery of type-II Dirac/Weyl semimet- be manipulated independently, thereby providing additional als has sparked a growing interest in the prospect of tunable degrees of freedom. As an example, recent work has type-II Dirac/Weyl points since the corresponding ‘Lorentz- shown the tantalizing prospect of engineering novel topologi- violating’ Dirac quasiparticles exhibit markedly different cal polaritons by introducing a winding coupling between or- properties from their type-I counterparts [31–35]. After dinary photons and [27]. Shortly thereafter, these their realization, the pursuit for electromagnetic analogs com- non-trivial interactions were exploited to gap out photonic menced [36–42]. However, just like type-I CDPs, the ex- Dirac points as an alternative route to topological polaritons istence of type-II Dirac/Weyl points is usually tied to judi- [29]. ciously engineered lattice structures, and are thus impossible In this work we exploit the hybrid nature of polaritons in to manipulate without modifying the structure [36–42]. a different setting, namely metamaterials. We explore Dirac In stark contrast, here the existence of the type-II Dirac physics in an entirely new scenario, shifting the focus from the points is intrinsically linked to the hybrid nature of the po- lattice structure and its deformations to the effect of manipu- laritons, emerging from a non-trivial winding in the light- lating the surrounding photonic environment. In particular, matter interaction. Crucially, the combination of this unique we theoretically study the polaritons supported by crystalline physical origin and the truly 2D nature of the metasurface metasurfaces consisting of a honeycomb array of resonant, results in an unprecedented tunability of emergent Dirac po- dipolar meta-atoms. Such a structure is imminently realizable laritons by modifying only the surrounding photonic environ- across the electromagnetic spectrum from arrays of plasmonic ment. Specifically, we show that by embedding the honey- nanoparticles to microwave helical resonators. Despite the el- comb metasurface inside a planar photonic cavity and sim- ementary nature of these metasurfaces, we show that complex ply changing the cavity height, one can induce novel phase 3 transitions, including the multi-merging of type-I and type- Emergence of Type-I Dirac Points II Dirac points and the annihilation of type-II Dirac points. This striking tunability results in distinct polariton phases, de- The matter Hamiltonian within the nearest-neighbor approxi- spite the preserved lattice structure. In particular, we unveil mation reads a morphing between a linear and a parabolic spectrum ac- X † †  ˜ X †  companied by a change in topological Berry phase, and an Hmat = ~ω˜0 aqaq + bqbq + ~Ω fqbqaq + H.c. , unprecedented environment-induced inversion of chirality, all q q of which have no analog in graphene or artificial graphene (1) systems studied thus far. Therefore, this unique paradigm of where, for brevity, we have not presented the non-resonant ω˜ exploiting the photonic environment will give rise to novel terms (see Methods for derivation). In equation (1), 0 is the Ω˜ Dirac-related phenomena at the subwavelength scale, such as renormalized resonant frequency and is the renormalized anomalous Klein tunnelling, negative refraction and pseudo- Coulombic interaction strength due to the cavity-induced im- magnetic Landau levels, which can all be tuned via the pho- age dipoles (see Methods for their dependence on the cav- a† b† tonic environment alone. ity height). The bosonic operators q and q create quanta of the quasistatic collective-dipole modes that extend across the A and B sublattices, respectively, with wavevector q in the first Brillouin zone (see Fig. 2a). Finally, the function P3  fq = j=1 exp iq · ej encodes the honeycomb geometry RESULTS of the lattice with nearest-neighbor vectors ej (see Fig. 1). We diagonalize the matter Hamiltonian (1) as Hmat = P P ωmatβ† β β† = Hamiltonian Formulation τ=± q ~ qτ qτ qτ where the bosonic operators qτ † ψq|ψqτ i create quasistatic collective-dipole normal modes with dispersion ωmat =ω ˜ + τΩ˜|f |. Here, τ indexes the While metamaterials have traditionally been described in qτ 0 q upper (τ = +1) and lower (τ = −1) bands and ψ† = terms of macroscopic effective properties [36, 39, 43], the q (a† , b† ) is a spinor creation operator. The spinors |ψ i = importance of crystallinity is becoming increasingly appar- q q √ qτ iϕ T ent [44]. Therefore, to capture the essential physics related (1, τe q ) / 2 describe an emergent pseudospin degree of to complex non-local effects that arise from strong multiple- freedom where the two components encode the relative ampli- scattering [45], here we study the properties of the cavity- tude and phase of the dipolar oscillations on the two inequiv- embedded honeycomb metasurface by means of a micro- alent A and B sublattices, respectively, with ϕq = arg(fq). scopic Hamiltonian formalism. This allows us to clearly iden- These spinors can be represented by a pseudospin vector on tify the distinct physical origins of the type-I and type-II Dirac the Bloch sphere which reads Sqτ = τ(cos ϕq, sin ϕq, 0). 0 points. At the high symmetry K and K points (see Fig. 2a) the sublattices decouple with no well-defined relative phase (i.e., The full polariton Hamiltonian of this system reads Hpol = fq = 0), giving rise to two inequivalent CDPs located at Hmat+Hph+Hint, where the interaction Hamiltonian Hint cou-   ±K = ± √4π , 0 as observed in Fig. 2b. As expected from ples the matter and photonic subspaces whose free dynamics 3 3a are governed by Hmat and Hph, respectively. We employ the the symmetry of the metasurface, the existence of CDPs is Coulomb gauge, where the instantaneous Coulomb interac- robust against long-range Coulomb interactions as shown in tion between the meta-atoms is incorporated within the matter Supplementary Note 1. In fact, for small cavity heights, the Hamiltonian Hmat [46]. The effects of the dynamic photonic image dipoles quench long-range Coulomb interactions and environment, described by the transverse vector potential, is the nearest-neighbor approximation becomes increasingly ac- included through the principle of minimal-coupling [46]. curate as shown in Supplementary Fig. 1. The CDPs corre- We model each meta-atom by a single dynamical degree spond to vortices in the pseudospin vector field Sqτ , which of freedom describing the electric-dipole moment associated give rise to topological singularities in the Berry curvature with its (non-degenerate) fundamental eigenmode with res- [47]. Therefore, CDPs are sources of quantized Berry flux wπ, where w = ±1 is the topological charge of the Dirac onant frequency ω0. These meta-atoms are then oriented such that their dipole moments point normal to the plane of point corresponding to the winding number of the vortex. the lattice. Furthermore, we consider subwavelength nearest- To quadratic order in k = q − K (ka  1), the eff neighbor separation a such that the light cone intersects the effective matter Hamiltonian near the K point is HK = P † eff † † † Brillouin zone edge above ω0, ensuring the existence of k ψkHK,kψk, with spinor creation operator ψk = (ak, bk) evanescently bound, subwavelength polaritons. The strength and Bloch Hamiltonian of the Coulomb dipole-dipole interaction between neighbor- eff ˜ ∗ ◦2 HK,k = ~ω˜012 − ~v˜σ · k + ~t (σ · k) . (2) ing meta-atoms is parametrized by Ω. Finally, the metasurface is embedded at the center of a planar photonic cavity of height Here, 12 is the 2 × 2 identity matrix, σ = (σx, σy) and ∗ ◦2 L, as depicted in Fig. 1, where the cavity walls are assumed to σ = (σx, −σy) are vectors of Pauli matrices, and rep- be lossless and perfectly conducting metallic plates. resents the Hadamard (element-wise) square. Note that the 4

Figure 2. Evolution of the√ polariton dispersion as√ the cavity height is reduced. a, First Brillouin zone defined by primitive reciprocal 2π 2π lattice vectors b1 = 3a (− 3, −1) and b2 = 3a ( 3, −1). b, Quasistatic dispersion of the collective-dipole normal modes, where the upper band corresponds to a bright, symmetric dipole configuration (↑↑) and the lower band corresponds to a dark, antisymmetric dipole configuration (↑↓). The light-cone (shaded region) is bounded by the linear dispersion of the TEM mode. Due to the non-trivial winding in the light-matter interaction (see Fig. 3), the band crossings are expected to result in large (band crossings ‘1’ and ‘2’) or small (band crossings ‘3’ and ‘4’) direction-dependent anticrossings in the polariton spectrum. c-e, Polariton dispersion obtained from the polariton Hamiltonian Hpol ¯ (solid black lines) and the two-band Hamiltonian Hmat (orange dashed lines), for c subcritical (L = 5a), d critical (L = Lc = 1.75a), and e supercritical (L = a) cavity heights, respectively. While type-I CDPs with an isotropic (see inset of c) exist even in the quasistatic dispersion (see b), new type-II SDPs with a critically tilted Dirac cone (see inset in c) emerge due to the vanishing light-matter interaction for 0 the dark quasistatic band along the Γ − K(K ) directions (see Fig. 3). At the critical cavity height Lc, three type-II SDPs merge with the type-I CDPs (see Fig. 5) resulting in a quadratic band-degeneracy at K(K0) (see inset in d). After criticality, the type-II SDPs annihilate one another and the massless Dirac cone re-emerges at the type-I CDPs (see inset in e) accompanied by an inversion of chirality (see Fig. 5). Plots obtained ph with parameters ωK0 = 2.5ω0 and Ω = 0.01ω0.

image dipoles do not qualitatively affect the physics, but sim- Hmat capture the essential physics. In fact, we will show that ply lead to a renormalization of the group velocity v˜ = 3Ω˜a/2 this quasistatic description misses the profound influence of and trigonal warping parameter t˜ = 3Ω˜a2/8. Apart from a the surrounding photonic environment, which has a remark- global energy shift, equation (2) is equivalent to a 2D mass- ably non-trivial effect on the Berry curvature and, therefore, less Dirac Hamiltonian to leading order in k, with an isotropic on the corresponding nature of the polaritons. mat Dirac cone spectrum ωkτ =ω ˜0 + τv˜|k| that is character- Crucially, the metallic cavity supports a fundamental trans- ized by closed isofrequency contours. Therefore, as expected verse electromagnetic (TEM) mode whose polarization (par- from the honeycomb symmetry, the CDP belongs to the type- allel to the dipole moments) and linear dispersion (see Fig. 2b) I class of 2D Dirac points,√ and the corresponding spinors are independent of the cavity height. For brevity, in what fol- iθk T |ψkτ i = (1, −τe ) / 2, where θk = arctan(ky/kx), rep- lows we do not present the contributions from the other cavity resent massless Dirac collective-dipoles with a topological modes since the essential physics emerges from the interac- Berry phase of π. The effective Hamiltonian near the K0 point tion with the fundamental TEM mode (see Methods for full eff eff ∗ is given by HK0,k = (HK,−k) , where the CDP is a source of expressions). In fact, the higher order cavity modes become −π Berry flux as required by time-reversal symmetry. increasingly negligible for smaller cavities as they are pro- gressively detuned from the dipole resonances. The effects of the photonic environment are encoded in the Hybridization with the Photonic Environment free photonic Hamiltonian

Given the subwavelength nearest-neighbor separation, it is X ph † Hph = ~ ωqncqncqn (3) tempting to assert that the near-field Coulomb interactions in qn 5

(1) and in the light-matter interaction Hamiltonian Hint = Hint + (2) Hint , with

(1) X ∗  † † †  Hint = ~ iξqnφn aqcqn + aqc−qn qn (4) X  † † †  +~ iξqnφn bqcqn + bqc−qn + H.c. qn and

(2) X 2ξqnξqn0  ∗  H = Re φ φ 0 int ~ ω n n qnn0 0 (5)  † † †  cqncqn0 + cqnc−qn0 + H.c., Figure 3. Non-trivial winding in the light-matter interaction. a, Dependence of the magnitude of the light-matter coupling constant iϕq 0 −1/2 |Λq0τ | ∝ |e + τ| on the direction of q from Γ − M to Γ − K(K ) where ξqn ∝ L parametrizes the strength of the light-matter interaction (see Methods for analytical expres- (see inset), for the upper (red line) and lower (blue line) quasistatic † bands. Plots obtained with |q| = |K|/2. b-e, Schematics of the sion). The bosonic operator cqn creates a TEM photon with bright (↑↑) and dark (↑↓) configurations of the two sublattices inter- ph wavevector q in the first Brillouin zone and dispersion ωqn = acting with the photonic mode, indicated by the field profile. Panels b and d represent the crossings labeled ‘1’ and ‘3’ in Fig. 2 along the c|q − Gn|, where n indexes the set of reciprocal lattice vec- Γ − K(K0) directions, respectively, while panels c and e represent tors Gn. The complex phase factors φn = exp (iaGn · yˆ) are associated with Umklapp processes that arise due to the the crossings labelled ‘2’ and ‘4’ along the Γ − M directions, respec- tively. Crucially, the light-matter interaction strength for the dark discrete, in-plane translational symmetry of the metasurface, 0 mode vanishes (|Λq0−| = 0) along the Γ − K(K ) directions due and must be retained as they are critical for maintaining the to the complete destructive interference between the two sublattices point-group symmetry of the polariton Hamiltonian. (see d), leading to the emergence of six inequivalent type-II Dirac points in the polariton spectrum. We diagonalize Hpol using a generalized Hopfield- Bogoliubov transformation [48] (see Methods for details), and in Fig. 2c-e we present the resulting polariton dispersion for different cavity heights. Also, in Supplementary Fig. 2 we Hamiltonian, present the full polariton dispersion that includes long-range Coulomb interactions. For small cavity heights, the full po- (1) X X  † † †  Hint = ~ iΛqnτ βqτ cqn + βqτ c−qn + H.c., (6) lariton dispersion is almost indistinguishable from that ob- τ=± qn tained in the nearest-neighbor approximation, and therefore one can conclude that long-range Coulomb interactions do we find that complex non-local interactions, which arise from not qualitatively affect the physics presented here. It is im- strong multiple scattering in the bipartite structure, result in a portant to stress that our general model captures the essential non-trivial winding of the light-matter coupling as a function physics that will emerge in a variety of different experimental of wavevector direction setups. To show this, in Supplementary Fig. 3 and Supple- ∗ iϕq  mentary Fig. 4 we present the polariton dispersions obtained Λqnτ ∝ ξqn φne + τφn . (7) from full-wave numerical simulations of a honeycomb array Naively, one may expect all of the band crossings in Fig. 2b of plasmonic nanorods and microwave helical resonators, re- to be avoided as a result of the hybridization between the spectively. These different physical realizations indeed show collective-dipole and photonic modes, as it is a characteris- the same evolution of the polariton spectrum as shown in tic feature of polaritonic systems [48]. Indeed, this is the Fig. 2c-e and Supplementary Fig. 2. case for the crossings with the upper quasistatic band where iϕq Λq0+ ∝ (e + 1) (see red line in Fig. 3a) due to the con- structive interference between the sublattices of this bright Emergence of Type-II Dirac Points (↑↑) configuration (see Fig. 3b-c). This results in a large anti- crossing for all wavevector directions, as observed in Fig. 2c. Given the elementary nature of the individual resonant ele- In stark contrast, for the lower quasistatic band the coupling iϕq ments, it is tempting to assume that nothing peculiar could constant is significantly reduced Λq0− ∝ (e − 1) (see blue emerge from the ordinary dipole-dipole interactions between line in Fig. 3a) due to the destructive interference between the the meta-atoms that are mediated by the electromagnetic field. sublattices of this dark (↑↓) configuration (see Fig. 3e). Con- However, by expressing the interaction Hamiltonian (4) in sequently, this results in a small anticrossing for a general † terms of the βqτ and βqτ operators that diagonalize the matter wavevector direction. 6

Crucially, however, the light-matter interaction for the lower quasistatic band completely vanishes (Λq0− = 0) along 0 the high-symmetry Γ − K(K ) directions, where ϕq = 0, due to the complete destructive interference between the two sub- lattices (see Fig. 3d). As a result, along these high-symmetry directions the crossings are protected, leading to six inequiva- lent Dirac points emerging in the polariton spectrum—we call these satellite Dirac points (SDPs) to distinguish them from the CDPs. As we will see below, these SDPs belong to the type-II class of 2D Dirac points where the dispersion takes the form of a critically tilted Dirac cone (see inset of Fig. 2c), characterized by open, hyperbolic isofrequency contours.

Effective Hamiltonian in the Matter Subspace

To explore the nature of the polaritons in the vicinity of the Figure 4. Tunable parameters in the effective Hamiltonian. De- different Dirac points, we first neglect non-resonant terms pendence of the parameters in the effective polariton Hamiltonian (9) in the matter Hamiltonian and perform a unitary Schrieffer- on the inverse cavity height. The blue dashed line shows the varia- tion of the group velocity v¯ which changes sign at the critical cav- Wolff transformation [49] on H to integrate out the pho- pol ity height Lc, leading to the inversion of chirality. The orange dot- tonic degrees of freedom (see Methods for details). Finally, dashed line shows the variation of the trigonal warping parameter we extract the two-band Hamiltonian in the matter sublattice t¯ which becomes dominant close to criticality. These parameters space have been normalized to v = 3Ωa/2 and t = 3Ωa2/8 which are the group velocity and trigonal warping parameters, respectively, in 2 ph the absence of image dipoles and light-matter interactions. The or- X ξqnωqn H¯ =H − 2 ange dot-dashed line in the inset shows the variation of the CDP fre- mat mat ~ ph 2 2 qn (ωqn) − ω˜0 (8) quency ω¯0, while the blue dashed line in the inset shows the variation † † 2 † ∗2 †  of the wavevector-dependent diagonal term D. Plots obtained with aqaq + bqbq + φnbqaq + φn aqbq . ph ωK0 = 2.5ω0 and Ω = 0.01ω0. Diagonalizing the two-band Hamiltonian (8) leads to an effec- tive dispersion (see Methods) which provides an excellent de- scription of the polaritons as indicated by the orange dashed (¯v/t,¯ 0) and obtain lines in Fig. 2c-e. Finally, we expand the two-band Hamil-  2  tonian (8) up to quadratic order in k and obtain the effective eff Dv¯ 2Dv¯ 0 ∗ 0 H¯ 0 = ω¯ − − k 1 + σ ·¯v · k , (10) ¯ eff P † ¯eff KS,k ~ 0 t¯2 t¯ x 2 ~ Hamiltonian near the K point HK = k ψkHK,kψk (see Supplementary Note 2 for derivation) with Bloch Hamiltonian 0 1 0 where k measures the deviation from KS and ¯v =v ¯ ( 0 3 ) ¯eff ¯ ∗ ◦2 2 is the velocity tensor. Apart from a global energy shift, HK,k = ~ω¯012 − ~v¯σ · k + ~t (σ · k) − ~D|k| 12. (9) the effective Hamiltonian (10) near the SDP takes the form 0 Similarly, the effective Hamiltonian near the K point is given of a generalized 2D massless Dirac Hamiltonian Hk = ¯eff ¯eff ∗ P u k 1 + P v k σ . If the parameters u and by HK0,k = (HK,−k) . In equation (9), the resonant fre- i=x,y ~ i i 2 i=x,y ~ i i i i 2 2 2 2 quency ω¯0, group velocity v¯, and trigonal warping parameter vi satisfy the condition ux/vx + uy/vy < 1, then the Dirac t¯, now encode non-trivial contributions from the hybridiza- cone becomes tilted and anisotropic [50] but still belongs to tion with the photonic environment. There is also an addi- the type-I class with closed isofrequency contours. However, 2 2 2 2 tional wavevector-dependent diagonal term parametrized by the condition ux/vx + uy/vy > 1 defines a distinct type-II D, which breaks the symmetry between the upper and lower class of 2D Dirac points, giving rise to a critically tilted Dirac polariton bands. The dependence of these parameters on the cone with open, hyperbolic isofrequency contours. Hence, cavity height is shown in Fig. 4 (see Methods for analytical the distinct type-I and type-II classes are related via a Lif- expressions). To leading order in k, one can observe that the shitz transition in the topology of the isofrequency contours. 2 2 2 ¯2 effective Hamiltonian (9) near the CDP is equivalent to a 2D Indeed, since we have uy = 0 and ux/vx = 4D /t > 1, massless Dirac Hamiltonian. Therefore, the polariton CDPs the SDPs belong to the type-II class of 2D Dirac points. Fur- remain in the type-I class and are robust against the coupling thermore, since the Hamiltonian (10) is expressed in terms of with the photonic environment—this is not surprising given σ∗, the pseudospin winds in the opposite direction around the that their physical origin is intrinsically linked to the lattice SDPs as compared to the CDP, and therefore the SDPs located structure alone, which is preserved here. along the Γ − K directions are sources of −π Berry flux. As To elucidate the nature of the SDPs, we expand the effec- required by time-reversal symmetry, the SDPs located along 0 tive Hamiltonian (9) near one of the SDPs located at KS = the Γ − K directions are sources of π Berry flux (opposite to 7 the CDP located at the K0 point) To the authors knowledge, rection and have opposite Berry flux. This topological tran- this is the first realization of type-II Dirac points in a truly 2D sition leaves only the type-I CDP remaining in the spectrum system with subwavelength dimension in the transverse direc- with π Berry flux (see Fig. 2e and Fig. 5c). tion. In this supercritical phase, we recover the linear dispersion near the type-I CDP to leading order in k (see Fig. 2e), and the effective Hamiltonian (9) is equivalent to a 2D massless Dirac Manipulation of Type-I and Type-II Dirac Points |ψ i = Hamiltonian√ with corresponding spinor eigenstates kτ iθ T (1, τe k ) / 2. Remarkably, massless Dirac polaritons thus We have demonstrated the simultaneous existence of type-I re-emerge past criticality with an unprecedented environment- and type-II Dirac points in the 2D polariton spectrum. Since ˆ induced inversion of chirality hψkτ |σ · k|ψkτ i = τ (see the latter are intrinsically linked to the hybridization between Fig. 5f). Physically, this corresponds to a π-rotation in the rel- the light and matter degrees of freedom, one can manipulate ative phase between the dipole oscillations on the two sublat- their location within the Brillouin zone by simply modifying tices, which is also accompanied by a π-rotation in the isofre- the light-matter interaction via the cavity height. As a re- quency domains (compare Fig. 5a and Fig. 5c). We emphasize sult, the polariton spectrum evolves into qualitatively distinct that it is the chirality of massless Dirac that is re- phases as highlighted in Fig. 2c-e. To elucidate the differences sponsible for most of the remarkable properties of monolayer between these phases, we study the spinor eigenstates (see graphene, including the Klein tunnelling phenomenon [13]. Methods) of the two-band Hamiltonian (8). In Fig. 5a-c we Consequently, this novel environment-induced inversion of plot the pseudospin vector field near the K point for differ- chirality could give rise to unconventional phenomena such as ent cavity heights and schematically depict the location of the anomalous Klein transport. For example, near the K point, the Dirac points, along with their associated Berry flux. Finally, right-propagating polaritons correspond√ to an antisymmetric in Fig. 5d-f we illustrate the corresponding effective polariton dipole configuration |ψ i = (1, −1)T/ 2 in the subcritical spectrum to leading order in k. Note that similar analysis can kτ √ phase and to a symmetric configuration |ψ i = (1, 1)T/ 2 be performed for the K0 point. kτ in the supercritical phase. Thus, due to the orthogonality be- In the subcritical phase (L > L ), three type-II SDPs are c tween these two eigenstates, the inversion of chirality removes located along the Γ − K directions, each with −π Berry flux the channel responsible for the perfect transmission in the con- surrounding a type-I CDP with π Berry flux (see Fig. 5a). ventional Klein tunnelling effect [13] (see Fig. 5d and Fig. 5f). To leading order in k, the polariton spectrum disperses lin- Such a scenario could be realized in a simple setup character- early about the type-I CDPs (see Fig. 2c) forming an isotropic ized by two neighboring regions with different cavity heights. Dirac cone with a group velocity v¯ that is tunable via the cav- ity height (see Fig. 4). Here, the effective Hamiltonian (9) is As a side remark, we note that the polariton spectrum near criticality bears some resemblance with the low-energy spec- equivalent to a 2D massless Dirac√ Hamiltonian with spinor iθ T eigenstates |ψ i = (1, −τe k ) / 2. These represent mass- trum of bilayer graphene with its central type-I Dirac point kτ and three type-I SDPs [51, 52]. However, given the type- less Dirac polaritons with chirality hψ |σ · kˆ|ψ i = −τ, kτ kτ II nature of the polariton SDPs, the topology of the polari- resulting in a pseudospin that winds once around the CDP and ton isofrequency contours are markedly different from that of a topological Berry phase of π (see Fig. 5d). the bilayer spectrum. This is further highlighted at criticality At the critical cavity height (L = L ), the group velocity of c where the polariton bands have the same curvature, which is the massless Dirac polaritons vanishes v¯(L ) = 0 (see Fig. 4) c in stark contrast to the electronic bands in bilayer graphene. as the type-II SDPs merge with the type-I CDP, forming a quadratic band-degeneracy (see Fig. 2d) with combined −2π We also note that recent works explored the possibility to Berry flux (see Fig. 5b). The leading order term in the effec- manipulate the (3+1) type-I Dirac points in bilayer graphene through the application of lattice deformations [53–56], lead- tive Hamiltonian (9) is now quadratic in k with√ correspond- −i2θk T ing to the merging and annihilation of CDPs and SDPs. ing spinor eigenstates |ψkτ i = (1, −τe ) / 2. There- fore, during this critical merging transition, the massless Dirac In addition, a multi-merging transition of all (3+1) type-I polaritons morph into massive chiral polaritons with qualita- Dirac points has been proposed theoretically within tight- tively distinct physical properties. These include a parabolic binding models involving the artificial tuning of third-nearest- ∗ ◦2 neighbour hopping amplitudes in a graphene-like honeycomb spectrum and chirality hψkτ | (σ · k) |ψkτ i = −τ, result- ing in a pseudospin that winds twice as fast compared to the structure [57–60]. However, these proposals have no physical subcritical phase and a topological Berry phase of −2π (see realization so far. In stark contrast, the imminently-realizable Fig. 5e). metasurfaces in our work enable the exploration of rich Dirac Since the point-group symmetry is preserved, the type-II phases with ease by simply modifying the photonic environ- SDPs do not annihilate the type-I CDP, but they re-emerge ment via an enclosing cavity. and continue to migrate along the K − M directions as the As a final remark, we briefly comment on how one might cavity height is reduced past criticality (L < Lc) (see inset of probe these novel Dirac phases. Given that the Dirac points Fig. 5c). After a small decrease in cavity height, these SDPs exist in a polaritonic excitation spectrum, one must drive the annihilate with other SDPs that migrate along the opposite di- system with photons at the required frequency in order to 8

Figure 5. Merging of type-I and type-II Dirac points with chirality inversion. a-c, Pseudospin vector field and isofrequency contours for the upper polariton band near the K point for a subcritical (L = 2.3a), b critical (L = Lc = 1.78a), and c supercritical (L = 1.4a) ¯ cavity heights, respectively, as obtained from Hmat. The Dirac points (corresponding to vortices) are depicted by orange circles along with their associated Berry flux. Before criticality, three type-II SDPs (−π Berry flux) are driven towards the type-I CDP (π Berry flux) along the Γ − K directions as the cavity height is reduced (see inset in a). At the critical cavity height Lc, they merge together forming a quadratic band-degeneracy with combined Berry flux of −2π (see b). After criticality, the type-II SDPs re-emerge and are driven past the type-I CDP along the K − M directions (see inset in c). After a small decrease in cavity height, these SDPs annihilate other SDPs that migrate along the opposite direction and have opposite Berry flux, leaving only the type-I CDP remaining in the spectrum with π Berry flux (see c). d-f, Effective polariton spectrum near the K point to leading order in k. The colors of the bands correspond to the chirality of the Dirac polaritons as defined in the main text, where the orange and blue bands indicate a chirality of +1 and −1, respectively. The spinor eigenstates, represented by pseudospin vectors (gold arrows), describe d massless Dirac polaritons with linear dispersion and Berry phase π, e massive chiral polaritons with parabolic dispersion and Berry phase −2π, and f massless Dirac polaritons with linear dispersion and Berry phase π, but with inverted ph chirality. All pseudospin and contour plots are obtained with ωK0 = 2.5ω0 and Ω = 0.01ω0.

probe them. Both classes of Dirac points lie outside of the physics, since one can directly probe the field distributions us- light-line and thus one must overcome the momentum mis- ing near-field scanning techniques [39], and thus one could match with photons. The specific experimental technique that directly probe the environment-induced chirality inversion. one would employ will depend on the nature of the meta- surface and the corresponding frequency regime. For exam- ple, techniques for plasmonic systems have traditionally in- DISCUSSION volved coupling via evanescent waves with prisms, gratings, and local scatterers [61], or more recent techniques such as To conclude, we have revealed rich and unique Dirac physics non-linear wave-mixing [62]. In contrast, realizations in the that emerges even in the most elementary honeycomb meta- microwave regime can be probed using point-like antenna surfaces. In particular, we have shown that the polariton spec- sources and detectors [39]. In fact, microwave metamaterials trum simultaneously exhibits type-I and type-II Dirac points, are proving to be a versatile platform for exploring Dirac/Weyl where the latter emerge from a non-trivial winding in the light- 9

matter interaction. We would like to emphasize that it is this Hamiltonian within the nearest-neighbor approximation reads unique physical origin together with the truly 2D nature of 2 ! X X ΠR M 2 2 the metasurface that enables one to qualitatively modify the H = s + ω h mat 2M 2 0 Rs fundamental properties of the emergent Dirac polaritons by s=A,B Rs 2 3 manipulating the surrounding photonic environment. This Q X X + h h 3 RB RB+ej 4πε0a stands in stark contrast to conventional artificial graphene sys- RB j=1

tems where the fundamental properties are dictated by the lat- 2 +∞ 3 Q X X X0 a 2 − 2 h tice structure alone. Therefore, exploiting the rich tunabilty 3 Rs 8πε0a lL (12) of the polariton spectrum with the environment offers a new s=A,B Rs l=−∞ 2 3 +∞ lL 2 Q X X X0 2 − 1 paradigm that opens a variety of opportunities to explore novel − a h h 8πε a3   5 RB RB+ej 0 R j=1 l=−∞ lL 2 2 Dirac-related physics at the subwavelength scale. B 1 + a

For example, one can simultaneously probe the dynamics 2 3 +∞ lL 2 Q X X X0 2 − 1 − a h h , of type-I and type-II Dirac quasiparticles, where the latter 3 5 RA RA−ej 8πε0a  lL 2 2 RA j=1 l=−∞ 1 + are predicted to exhibit intriguing anomalous refraction be- a

haviour [40]. Furthermore, the environment-induced redshift l = 0 Π where the primed summations exclude the term. Here, Rs is the conjugate h of the CDP frequency ω¯0 (see Fig. 4) will allow the investiga- momentum to the dynamical coordinate Rs corresponding to the meta-atom located at tion of polaritonic Klein tunnelling through interfaces separat- Rs, and M is an effective mass. Next we introduce the bosonic operators ing regions with different cavity heights. Consequently, neg- s r Mω 1 a = 0 h + i Π (13) ative refraction can be induced by simple variations in cavity RA RA RA 2~ 2~Mω0 height, which could be exploited in novel schemes for guiding and manipulating light at the subwavelength scale, including and

polaritonic Veselago lensing [63, 64]. Moreover, the tunable s r Mω 1 group velocity will also enable the exploration of velocity bar- b = 0 h + i Π (14) RB 2 RB 2 Mω RB riers for the unprecedented guiding and localization of mass- ~ ~ 0

less Dirac quasiparticles [65, 66], which is extremely difficult that annihilate quanta of the fundamental eigenmode on the meta-atom located at RA and R , respectively, and satisfy the commutation relations [a , a† ] = δ , to achieve in real graphene. Furthermore, one could also com- B R R0 RR0 [b , b† ] = δ , and [a , b† ] = 0. In terms of these operators, the matter bine the effects of the environment with inhomogeneous strain R R0 RR0 R R0 deformations, giving rise to unique pseudomagnetic-related Hamiltonian (12) reads

X † X † effects, including the intriguing ability to induce a pseudo- H = ω a a + ω b b mat ~ 0 RA RA ~ 0 RB RB Landau level spectrum for polaritons that can be qualitatively RA RB tuned via the cavity height. Finally, the ability to controllably 3 X X h †  †  i + Ω(1 − I) b a + a + H.c. invert the chirality of Dirac polaritons opens new perspectives ~ RB RB+ej RB+ej RB j=1 for anomalous pseudorelativistic transport through interfaces (15) X h †  †  i − ~ ΩS a a + a + H.c. separating regions in distinct polaritonic phases. 2 RA RA RA RA

X h †  †  i − ~ ΩS b b + b + H.c. 2 RB RB RB RB METHODS 2 3 where Ω = Q /8πε0Mω0a parametrizes the strength of the nearest-neighbor Coulomb interaction, and the parameters Derivation of the polaritonic Hamiltonian ∞ ∞ lL 2 X  a 3 X 2 − 1 S = 4 , I = 2 a , (16) lL h i 5 The cavity-embedded metasurface is composed of a honeycomb array of identical meta- l=1 l=1 1 + lL 2 2 L L a atoms located at RA = R + ayˆ + 2 zˆ and RB = R − ayˆ + 2 zˆ on the A and B sublattices, respectively. Here, R = l1a1+l2a2 is an in-plane lattice translation vector with primitive vectors a1 and a2 (see Fig. 1) and integers l1 and l2. Each meta-atom is encode renormalizations due to the cavity-induced image dipoles. We apply Born-von modelled by a single dynamical degree of freedom h (with dimensions of length), where Karm´ an´ boundary conditions over a lattice with N  1 unit cells and introduce −1/2 P iq·R the electric-dipole moment associated with its fundamental eigenmode is p = −Qhzˆ, the Fourier transform of the bosonic operators a = N aqe A and RA q with effective charge Q. The Coulomb potential energy between two dipole moments p −1/2 P iq·R 0 0 b = N bqe B , which transforms the matter Hamiltonian (15) into and p located at generic positions r and r , respectively, is given by RB q the local and block-diagonal form

0 0 p · p − 3(p · nˆ)(p · nˆ) X  † †  V = , (11) H = (ω − ΩS) a a + b b Coul 4πε |r − r0|3 mat ~ 0 q q q q 0 q

X h †  †  i + Ω(1 − I) f b a + a + H.c. 0 0 ~ q q q −q (17) where nˆ = (r − r )/|r − r | and ε0 is the vacuum permittivity. q The presence of the perfectly conducting metallic plates, placed at z = 0 and z = X 1  † † † †  L, modifies the boundary conditions on the scalar potential and, therefore, the Coulomb − ΩS a a + b b + H.c. . ~ 2 q −q q −q interaction between the meta-atoms. Using the method of images to ensure the vanishing q of the scalar potential at the cavity walls [67], we introduce an infinite series of image dipoles located outside the cavity at positions Rs + lLzˆ, where s = A, B indexes the † † two sublattices and l is a non-zero integer. Noting that the Coulomb potential energy In the main text we do not present the non-resonant terms (e.g. bqa−q), leading to ˜ between a real and image dipole is 1/2 of that given by equation (11) [68], the matter equation (1) where ω˜0 = ω0 − ΩS and Ω = Ω (1 − I). 10

† † † † † In the Coulomb gauge, the light-matter interaction is described by the minimal- Cq = (cq1, cq2, . . . , cqp, . . . , cqN ) is the vector of TM photon creation operators, coupling Hamiltonian [46] which, within the dipole approximation, reads where p indexes the set of ordered triplets of integers (n1, n2, m), and N is the total number of photonic operators considered. The Hermitian [2(N + 2)] × [2(N + 2)] 2 pol Q X X Q X X 2 matrix H can be written in block form as H = Π A (R ) + A (R ), (18) q int M Rs z s 2M z s s=A,B Rs s=A,B Rs H+ H− − W ! | {z } | {z } pol q q q (1) (2) † ∗ H H Hq =  −   +  , (25) int int Hq − Wq H−q

where we have used ΠR = ΠRzˆ. The vector potential can be decomposed into trans- verse electric (TE) and transverse magnetic (TM) modes of the cavity. However, the where photons corresponding to the TE modes have an in-plane polarization, and therefore   only TM modes contribute to the z-component of the vector potential ph ph ph ph Wq = ~ Diag ω0, ω0, ωq1, ωq2, . . . , ωqp, . . . , ωqN (26) s |q − G |  mπ  X ~ n is the (N +2)×(N +2) diagonal matrix of resonant frequencies of the free oscillators. Az (r, z) = cos z ε N NALωph |q − G + mπ zˆ| L ± qmn 0 m qmn n L The (N + 2) × (N + 2) block matrices Hq can be sub-divided into block matrices

h i(q−Gn)·r † −i(q−Gn)·ri cqmne + cqmne , Hmat Hint! ± q q (19) † Hq =  int ph , (27) ± Hq Hq √ 2 where A = 3 3a /2 is the area of a unit cell and Nm = 1 + δm0. The bosonic † operator cqmn creates a TM photon with wavevector q in the first Brillouin zone and where the upper-diagonal block ph dispersion ωqmn = c|q − Gn + zˆmπ/L|. Here, Gn = n1b1 + n2b2 is a reciprocal lattice vector with primitive vectors b and b , where n indexes the set ˜ ∗! 1 2 mat ω˜ Ωf of ordered pairs of integers (n , n ), and m is a non-negative integer indexing the H = 0 q (28) 1 2 q ~ Ω˜f ω˜ different TM cavity modes. Only TM photons with even m couple to the dipoles due to q 0 the parity selection rule at the centre of the cavity. Substituting the vector potential (19) into equation (18) we obtain the light-matter ph is the 2 × 2 matrix in the matter subspace, and the lower-diagonal block Hq is the interaction Hamiltonian N × N matrix in the photonic subspace with components

(1) X ∗  † † †  Hint = ~ iξqmnφn aqcqmn + aqc−qmn ξ ξ 0 qmn ph ph qp qp n ∗ o (H ) 0 = ω δ + 4 Re φ φ . (29) (20) q pp ~ qp pp0 ~ p p0 X  † † †  ω0 +~ iξqmnφn bqcqmn + bqc−qmn + H.c. qmn int Finally, the off-diagonal block Hq in equation (27) is the 2 × N interaction matrix, and where the pth column reads

! 2ξ ξ 0 0 ∗ (2) X qmn qm n  ∗   int iξqpφp Hint = ~ Re φ φ 0 Hq = ~ . (30) ω n n p iξ φ qmm0nn0 0 (21) qp p

 † † †  c c 0 0 + c c 0 0 + H.c. . qmn qm n qmn −qm n The polariton Hamiltonian Hpol is diagonalized by a generalized Hopfield-Bogoliubov † † T † transformation [48] Ψq = TqXq, where Xq = (χq, χ−q) and χq = The strength of the light-matter interaction is parametrized by † † † † (γq1, γq2, . . . , γqν , . . . , γqN+2). To ensure the invariance of the bosonic commu-

1 tation relations for the transformed operators, Tq must be a [2(N+2)]×[2(N+2)] pa- ph ! 2 ω 8π a Ω † † ξ = ω F(ωph ) q0n √ , raunitary matrix [69] that satisfies T η T = T η T = η , where η = σ ⊗ 1 qmn 0 qmn ph ph (22) q z q q z q z z z 2 ωqmn 3 3Nm L ωqmn † † and σz is the Pauli matrix. The transformed bosonic operators γqν = Ψqηz |Ψqν i and γ = hΨ |η Ψ that diagonalize the polariton Hamiltonian as where, to take into account the finite size of the meta-atoms, we have introduced a phe- qν qν z q ph nomenological function F(ωqmn) that provides a smooth cut-off for the interaction X pol † with short-wavelength photonic modes where the dipole approximation breaks down. Hpol = ~ωqν γqν γqν , (31) We choose the phenomenological cut-off function to be of the Fermi-Dirac distribution qν form th 1 create and annihilate polaritons in the ν band, respectively. The polariton dispersion F(ωph ) = , (23) pol qmn ph ωqν (black solid lines in Fig. 2c-e) and the corresponding linearly-independent eigen- 1 + e2(ωqmn−3ω0)/ω0 vectors |Ψqν i (first two columns of Tq) are determined from the positive eigenvalue pol pol which is smooth enough to avoid spurious artifacts appearing in the polariton spectrum. solutions to the non-Hermitian eigenvalue equation ηz Hq |Ψqν i = ~ωqν |Ψqν i. Finally, the free photonic Hamiltonian of the cavity reads

X ph † Hph = ~ ωqmncqmncqmn. (24) qmn Schrieffer-Wolff transformation

In equations (3), (4), (5) in the main text we only present the contribution from the TEM To obtain an effective two-band Hamiltonian in the matter sublattice space, we neglect mode (m = 0), dropping the corresponding index. In Supplementary Note 1 we discuss non-resonant terms in the matter Hamiltonian (since Ω/ω0  1 for practical realiza- the effect of the higher order (m 6= 0) TM cavity modes for larger cavities. tions of the metasurface), but not in the light-matter interaction Hamiltonian since the photons are not resonant with the collective-dipoles near the corners of the Brillouin zone (see Fig. 2b). Next, we perform a unitary transformation

Hopfield-Bogoliubov diagonalization h i 1 h h ii H¯ = eS H e−S = H + S, H + S S, H + ... (32) pol pol pol pol 2 pol

The polariton Hamiltonian Hpol = Hmat + Hph + Hint, where Hmat is given by equa- tion (17), Hph by equation (24), and Hint by equations (20) and (21), can be recast into and impose the Schrieffer-Wolff condition [49] 1 P † pol † † † T T matrix form as Hpol = 2 q ΨqHq Ψq where Ψq = (ψq,Cq, ψ−q,C−q). † † † h i (1) Here, ψq = (aq, bq) is the spinor creation operator in the matter sublattice space and S, Hmat + Hph = −Hint (33) 11

which eliminates the light-matter interaction to first order in ξqmn. From equation (33) and the particular form of the anti-Hermitian operator S reads 2 ! D Ω X 4π (2) 1 (1) = Cn∆Kmn − ∆Kmn , (43) X iξqmn ∗ † † ω a2 ω 27 2 S = − (φ a + φ b )c 0 0 mn ph n q n q qmn qmn ωqmn − ω˜0 (34) with X iξqmn ∗ † † † + (φnaq + φnbq)c−qmn − H.c. √ ωph +ω ˜ 3  4π  qmn qmn 0 A = (2 − 3n ) cos (n + n ) n 2 1 3 1 2 (44) ph   where we have used the approximation |ωqmn ± ω0|  Ω|fq| that is valid near the 1 4π 0 + (6n2 − 3n1) sin (n1 + n2) , K and K points. Retaining leading-order terms in ξqmn, the transformed polariton 2 3 Hamiltonian (32) reads

1 h i   ¯ (1) (2)  2 2  4π Hpol ' Hmat + S, Hint + Hph + Hint , (35) B = 3n − 6n − 6n + 6n n + 2 cos (n + n ) 2 n 1 2 1 1 2 3 1 2 | {z } | {z } H¯ (45) H¯ ph √    4π  mat + 3 2n − 4n + 6n n − 3n2 sin (n + n ) , 1 2 1 2 1 3 1 2

where the matter and photonic subspaces are decoupled to quadratic order in ξqmn. Calculating the commutator in equation (35) and extracting the Hamiltonian within the and matter sublattice space we obtain the two-band Hamiltonian

Cn = 1 + 3n1 (n1 − 1) + 3n2 (n2 − n1) . (46) 2 ph X ξqmnωqmn † † 2 † ∗2 † H¯ = H −2 a a +b b +φ b a +φ a b . mat mat ~ ph 2 2 q q q q n q q n q q For brevity, we retain only the dominant (m = 0) TEM contribution for the plots in qmn (ωqmn) − ω˜ 0 Fig. 4, where the coefficients ∆(υ) in equation (39) are given by (36) K0n In equation (8) in the main text we only present the contribution from the TEM mode     " 2 # (m = 0), dropping the corresponding index. We can recast the Hamiltonian (36) into (0) 4π a ω0 2 ph ¯ P † ¯mat ∆K0n = 2 √ F (ωK0n), (47) matrix form as Hmat = q ψqHq ψq, with Bloch Hamiltonian L ph 2 2 3 3 (ωK0n) − ω˜0  ∗  ¯mat Wq Fq Hq = ~ . (37) Fq Wq ph !2 " #2  4π −1  a  ω ω2 ∆(1) = 4 √ K00 0 K0n ph 2 2 P ˜ P 2 3 3 L ω0 (ω ) − ω˜ Here Wq =ω ˜0 − Ω mn ∆qmn and Fq = Ωfq − Ω mn ∆qmnφn with K0n 0  h ph i  (48) ∂ F 2(ω ) ph 2 2 2  2 ph K0n (ω ) − ω˜0  2 ωph ! F (ω ) − K0n , 16π a ω0 q0n 2 ph K0n ph ph ∆ = √ F (ω ). (38)  ∂ωK0n 2ωK0n  qmn L ph 2 2 ph qmn 3 3Nm (ωqmn) − ω˜0 ωqmn and ¯ mat Diagonalizing Hmat leads to the two-band dispersion ω¯qτ = Wq + τ|Fq|, which is indicated by the orange-dashed lines in Fig. 2c-e. The corresponding spinor eigenstates  −3   ph !4 " 2 #3 √ (2) 4π a ω ω |ψ i = (1, τeiϕ ¯q )T/ 2, where ϕ¯ = arg(F ), can be represented by the pseu- ∆ = 16 √ K00 0 qτ q q K0n L ω ph 2 2 3 3 0 (ωK0n) − ω˜0 dospin vector Sqτ = τ(cosϕ ¯q, sinϕ ¯q, 0) from which we obtain the pseudospin vector field plots in Fig 5a-c.  h 2 ph i h ph 2 2i h ph 2 2i  ∂ F (ωK0n) 5(ωK0n) − ω˜0 (ωK0n) − ω˜0 F 2(ωph ) − K0n ph ph 3  ∂ωK0n 8(ωK0n) 2  2 h 2 ph i h ph 2 2i Expansion of the effective two-band Hamiltonian ∂ F (ωK0n) (ωK0n) − ω˜0  + . (∂ωph )2 8(ωph )2 K0n K0n  Near the K point, the function ∆ , given by equation (38), expands as qmn (49)

(0) 2 (1) h i ∆kmn ' ∆Kmn − a ∆Kmn (K − Gn)xkx + (K − Gn)y ky

1 h 2 (1) 4 (2) 2 i 2 Data Availability + −a ∆Kmn + a ∆Kmn (K − Gn)x kx 2 (39) 1 h i + −a2∆(1) + a4∆(2) (K − G )2 k2 All relevant data are available from the corresponding authors upon reasonable request. 2 Kmn Kmn n y y 4 (2) + a ∆Kmn (K − Gn)x (K − Gn)y kxky

(υ) to quadratic order in k, where the real parameters ∆Kmn (υ = 0, 1, 2) depend only ph on the photon frequencies ωKmn at the K point. Collecting the contributions from the degenerate photons (see Supplementary Note 2 for details) we obtain the effective [1] Novoselov, K. S. et al. Electric field effect in atomically thin carbon Hamiltonian (9), where parameters are given by films. Science 306, 666-669 (2004). [2] Polini, M., Guinea, F., Lewenstein, M., Manoharan, H. C. & Pellegrini, ω¯0 Ω Ω X (0) V. Artificial honeycomb lattices for , atoms and photons. Nat. = 1 − S − ∆Kmn, (40) ω0 ω0 ω0 mn Nanotechnol. 8, 625-633 (2013). [3] Zhu, S.-L., Wang, B. & Duan, L.-M. Simulation and detection of Dirac fermions with cold atoms in an optical lattice. Phys. Rev. Lett. 98, v¯ 4π X (1) 260402 (2007). = 1 − I − An∆Kmn, (41) v 27 mn [4] Peleg, O. et al. Conical diffraction and gap solitons in honeycomb pho- tonic lattices. Phys. Rev. Lett. 98, 103901 (2007). [5] Han, D., Lai, Y., Zi, J., Zhang, Z. Q. & Chan, C. T. Dirac spectra 2 t¯ 8π X (2) and edge states in honeycomb plasmonic lattices. Phys. Rev. Lett. 102, = 1 − I − Bn∆Kmn, (42) t 81 mn 123904 (2009). 12

[6] Gibertini, M. et al. Engineering artificial graphene in a two-dimensional [34] Huang, H., Zhou, S. & Duan, W. Type-II Dirac fermions in the PtSe2 gas. Phys. Rev. B 79, 241406(R) (2009). class of transition metal dichalcogenides. Phys. Rev. B 94, 121117(R) [7] Torrent, D. & Sanchez-Dehesa,´ J. Acoustic analogue of graphene: Ob- (2016). servation of Dirac cones in acoustic surface waves. Phys. Rev. Lett. 108, [35] Yan, M. et al. Lorentz-violating type-II Dirac fermions in transition 174301 (2012). metal dichalcogenide PtTe2. Nat. Commun. 8, 257 (2017). [8] Gomes, K. K., Mar, W., Ko, W., Guinea, F. & Manoharan, H. C. De- [36] Xiao, M., Lin, Q. & Fan, S. Hyperbolic Weyl point in reciprocal chiral signer Dirac fermions and topological phases in molecular graphene. metamaterials. Phys. Rev. Lett. 117, 057401 (2016). Nature 483, 306-310 (2012). [37] Chen, W.-J., Xiao, M. & Chan, C. T. Photonic crystals possessing mul- [9] Bellec, M., Kuhl, U., Montambaux, G. & Mortessagne, F. Tight-binding tiple Weyl points and the experimental observation of robust surface couplings in microwave artificial graphene. Phys. Rev. B 88, 115437 states. Nat. Commun. 7, 13038 (2016). (2013). [38] Noh, J. et al. Experimental observation of optical Weyl points and Fermi [10] Weick, G., Woollacott, C., Barnes, W. L., Hess, O. & Mariani, E. Dirac- arc-like surface states. Nat. Phys. 13, 611?618 (2017). like in honeycomb lattices of metallic nanoparticles. Phys. [39] Yang, B. et al. Direct observation of topological surface-state arcs in Rev. Lett. 110, 106801 (2013). photonic metamaterials. Nat. Commun. 8, 97 (2017). [11] Yu, S. et al. Surface phononic graphene. Nat. Mater. 15, 1243?1247 [40] Wang, H.-X., Chen, Y., Hang, Z. H., Kee, H.-Y. & Jiang, J.-H. Type-II (2016). Dirac photons. npj Quantum Mater. 2, 54 (2017). [12] Castro Neto, A. H., Peres, N. M. R., Novoselov, K. S., Geim, A. K. & [41] Lin, J. Y., Hu, N. C., Chen, Y. J., Lee, C. H. & Zhang, X. Line nodes, Guinea, F. The electronic properties of graphene. Rev. Mod. Phys. 81, Dirac points, and Lifshitz transition in two-dimensional nonsymmor- 109-162 (2009). phic photonic crystals. Phys. Rev. B 96, 075438 (2017). [13] Katsnelson, M. I., Novoselov, K. S. & Geim, A. K. Chiral tunnelling [42] Pyrialakos, G. G., Nye, N. S., Kantartzis, N. V & Christodoulides, D. and the Klein paradox in graphene. Nat. Phys. 2, 620-625 (2006). N. Emergence of type-II Dirac points in graphynelike photonic lattices. [14] Tarruell, L., Greif, D., Uehlinger, T., Jotzu, G. & Esslinger, T. Creat- Phys. Rev. Lett. 119, 113901 (2017). ing, moving and merging Dirac points with a Fermi gas in a tunable [43] Alu,` A. First-principles homogenization theory for periodic metamate- honeycomb lattice. Nature 483, 302-305 (2012). rials. Phys. Rev. B 84, 075153 (2011). [15] Bellec, M., Kuhl, U., Montambaux, G. & Mortessagne, F. Topological [44] Yves, S. et al. Crystalline metamaterials for topological properties at transition of Dirac points in a microwave experiment. Phys. Rev. Lett. subwavelength scales. Nat. Commun. 8, 16023 (2017). 110, 033902 (2013). [45] Kaina, N., Lemoult, F., Fink, M. & Lerosey, G. Negative refractive [16] Rechtsman, M. C. et al. Topological creation and destruction of edge index and acoustic superlens from multiple scattering in single negative states in photonic graphene. Phys. Rev. Lett. 111, 103901 (2013). metamaterials. Nature 525, 77?81 (2015). [17] Rechtsman, M. C. et al. Strain-induced pseudomagnetic field and pho- [46] Craig, D. P. & Thirunamachandran, T. Molecular Quantum Electro- tonic Landau levels in dielectric structures. Nat. Photonics 7, 153-158 dynamics: An Introduction to Radiation-Molecule Interactions. (Aca- (2013). demic Press, London, 1984). [18] Rechtsman, M. C. et al. Photonic Floquet topological insulators. Nature [47] Fuchs, J. N., Piechon,´ F., Goerbig, M. O. & Montambaux, G. Topolog- 496, 196-200 (2013). ical Berry phase and semiclassical quantization of cyclotron orbits for [19] Ni, X., Purtseladze, D., Smirnova, D. A., Slobozhanyuk, A. & Alu,` A. two dimensional electrons in coupled band models. Eur. Phys. J. B 77, Spin and valley polarized one-way Klein tunneling in photonic topo- 351-362 (2010). logical insulators. Preprint at https://arxiv.org/abs/1707.05802 (2017). [48] Hopfield, J. J. Theory of the contribution of excitons to the complex [20] Hasegawa, Y., Konno, R., Nakano, H. & Kohmoto, M. Zero modes dielectric constant of crystals. Phys. Rev. Lett. 112, 1555-1567 (1958). of tight-binding electrons on the honeycomb lattice. Phys. Rev. B 74, [49] Schrieffer, J. R. & Wolff, P. A. Relation between the Anderson and 033413 (2006). Kondo Hamiltonians. Phys. Rev. 149, 491-492 (1966). [21] Wunsch, B., Guinea, F. & Sols, F. Dirac-point engineering and topolog- [50] Goerbig, M. O., Fuchs, J. N., Montambaux, G. & Piechon,´ F. ical phase transitions in honeycomb optical lattices. New J. Phys. 10, Tilted anisotropic Dirac cones in quinoid-type graphene and α − 103027 (2008). (BEDT-TTF)2I3. Phys. Rev. B 78, 045415 (2008). [22] Pereira, V. M., Castro Neto, A. H. & Peres, N. M. R. Tight-binding [51] McCann, E. & Fal’ko, V. I. Landau-level degeneracy and quantum Hall approach to uniaxial strain in graphene. Phys. Rev. B 80, 045401 (2009). effect in a graphite bilayer. Phys. Rev. Lett. 96, 086805 (2006). [23] Montambaux, G., Piechon,´ F., Fuchs, J. N. & Goerbig, M. O. Merging [52] McCann, E. & Koshino, M. The electronic properties of bilayer of Dirac points in a two-dimensional crystal. Phys. Rev. B 80, 153412 graphene. Reports Prog. Phys. 76, 056503 (2013). (2009). [53] Mucha-Kruczynski,´ M., Aleiner, I. L. & Fal’ko, V. I. Strained bilayer [24] Guinea, F., Katsnelson, M. I. & Geim, A. K. Energy gaps and a zero- graphene: Band structure topology and Landau level spectrum. Phys. field in graphene by strain engineering. Nat. Phys. Rev. B 84, 041404(R) (2011). 6, 30-33 (2009). [54] Mariani, E., Pearce, A. J. & Von Oppen, F. Fictitious gauge fields in [25] Jacqmin, T. et al. Direct observation of Dirac cones and a flatband in a bilayer graphene. Phys. Rev. B 86, 165448 (2012). honeycomb lattice for polaritons. Phys. Rev. Lett. 112, 116402 (2014). [55] Verberck, B., Partoens, B., Peeters, F. M. & Trauzettel, B. Strain- [26] Nalitov, A. V., Solnyshkov, D. D. & Malpuech, G. Polariton Z topolog- induced band gaps in bilayer graphene. Phys. Rev. B 85, 125403 (2012). ical insulator. Phys. Rev. Lett. 114, 116401 (2015). [56] Pearce, A. J., Cavaliere, F. & Mariani, E. Conductance and shot noise in [27] Karzig, T., Bardyn, C. E., Lindner, N. H. & Refael, G. Topological strained bilayer graphene. J. Phys. Condens. Matter 25, 375301 (2013). polaritons. Phys. Rev. X 5, 031001 (2015). [57] Bena, C. & Simon, L. Dirac point metamorphosis from third-neighbor [28] Bardyn, C. E., Karzig, T., Refael, G. & Liew, T. C. H. Topological couplings in graphene and related materials. Phys. Rev. B 83, 115404 polaritons and excitons in garden-variety systems. Phys. Rev. B 91, (2011). 161413(R) (2015). [58] Montambaux, G. An equivalence between monolayer and bilayer hon- [29] Yi, K. & Karzig, T. Topological polaritons from photonic Dirac cones eycomb lattices. Eur. Phys. J. B 85, 375 (2012). coupled to excitons in a magnetic field. Phys. Rev. B 93, 104303 (2016). [59] Sticlet, D. & Piechon,´ F. Distant-neighbor hopping in graphene and [30] Yuen-Zhou, J. et al. Dirac points and topological modes. Nat. Haldane models. Phys. Rev. B 87, 115402 (2013). Commun. 7, 11783 (2016). [60] Bhattacharya, U., Hutchinson, J. & Dutta, A. Quenching in Chern insu- [31] Soluyanov, A. A. et al. Type-II Weyl semimetals. Nature 527, 495?498 lators with satellite Dirac points: The fate of edge states. Phys. Rev. B (2015). 95, 144304 (2017). [32] Deng, K. et al. Experimental observation of topological Fermi arcs in [61] Maier, S. A. Plasmonics: Fundamentals and Applications (Springer- type-II MoTe2. Nat. Phys. 12, 1105?1110 (2016). Verlag, Berlin, 2007). [33] Huang, L. et al. Spectroscopic evidence for a type II Weyl semimetallic [62] Renger, J., Quidant, R., Van Hulst, N., Palomba, S. & Novotny, L. Free- state in MoTe2. Nat. Mater. 15, 1155-1160 (2016). space excitation of propagating surface polaritons by nonlinear 13

four-wave mixing. Phys. Rev. Lett. 103, 266802 (2009). MetaMat) and the Centre National de la Recherche Scientifique through the [63] Cheianov, V. V, Fal’ko, V. & Altshuler, B. L. The focusing of electron Projet International de Cooperation´ Scientifique program (Contract No. 6384 flow and a Veselago lens in graphene p-n junctions. Science 315, 1252- APAG). W.L.B. acknowledges the financial support from EPSRC (Grant No. 1255 (2007). EP/K041150/1) and EU ERC project Photmat (ERC-2016-ADG-742222). [64] Lee, G.-H., Park, G. & Lee, H. Observation of negative refraction of E.M. acknowledges financial support from the Leverhulme Trust (Research Dirac fermions in graphene. Nat. Phys. 11, 925-929 (2015). Project Grant RPG-2015-101), and the Royal Society (International Ex- [65] Raoux, A. et al. Velocity-modulation control of electron-wave propaga- change Grant No. IE140367, Newton Mobility Grants NI160073, and Theo tion in graphene. Phys. Rev. B 81, 073407 (2010). Murphy Award TM160190). [66] Downing, C. A. & Portnoi, M. E. Localization of massless Dirac par- ticles via spatial modulations of the Fermi velocity. J. Phys. Condens. Matter 29, 315301 (2017). AUTHOR CONTRIBUTIONS [67] Jackson, J. D. Classical Electrodynamics, 3rd ed. (Wiley, New York, 1999). [68] Power, E. A. & Thirunamachandran, T. Quantum electrodynamics in a C.-R.M. performed the theoretical calculations and wrote the manuscript with cavity. Phys. Rev. A 25, 2473-2484 (1982). E.M.; T.J.S. contributed to the theoretical calculations; E.M. and G.W. con- [69] Colpa, J. H. P. Diagonalization of the quadratic boson Hamiltonian. ceived the project; and E.M. and W.L.B. supervised the project. All authors Physica A 93, 327-353 (1978). commented on the manuscript.

ADDITIONAL INFORMATION ACKNOWLEDGMENTS Supplementary information accompanies this work. Correspondence and requests for materials should be addressed to C.-R.M (email: We would like to thank Ian Hooper for useful discussions and help with [email protected]) or to E.M. (email: [email protected]) the numerical simulations. C.-R.M. and T.J.S. acknowledge financial sup- port from the Engineering and Physical Sciences Research Council (EPSRC) of the United Kingdom. In addition, C.-R.M acknowledges the EPSRC COMPETING FINANCIAL INTERESTS Centre for Doctoral Training in Metamaterials (Grant No. EP/L015331/1) and QinetiQ for additional funding. G.W. acknowledges financial support from Agence Nationale de la Recherche (Project ANR-14-CE26-0005 Q- The authors declare no competing financial interests. 14

Supplementary Information: Manipulating Type-I and Type-II Dirac Polaritons in Cavity-Embedded Honeycomb Metasurfaces

Supplementary Figure 1. Quasistatic dispersion beyond the nearest-neighbor approximation. a, Quasistatic dispersion within the nearest- neighbor approximation for cavity heights of L = 1.3a (blue dashed line) and L = a (orange dot-dashed line). For comparison we also have plotted the quasistatic dispersion without a cavity (black solid line). One observes that the effect of the image dipoles is to renormalize the resonant frequency and the Coulomb interaction strength, resulting in the quasistatic dispersion shifting down in frequency and reducing in bandwidth. Panel b is the same as a but with the full quasistatic dispersion (S8) that includes Coulomb interactions beyond nearest-neighbors until convergence. Long-range Coulomb interactions introduce an asymmetry between the upper and lower quasistatic bands as shown in b, but the CDPs remain robust at the K and K0 points. As the cavity height is reduced, the image dipoles quench long-range Coulomb interactions, which become negligible for cavity heights on the order of the nearest-neighbor separation. All plots obtained with Ω = 0.01ω0.

Supplementary Figure 2. Evolution of the polariton dispersion beyond the nearest-neighbor approximation. Full polariton dispersion that includes long-range Coulomb interactions (solid black lines) compared to the polariton dispersion obtained within the nearest-neighbor approximation (orange dashed lines) for a very large (L = 15a), b sub-critical (L = 5a), c critical (L = 1.75a), and d super-critical (L = a) cavity heights, respectively. Higher-order cavity modes can be observed for very large cavities as seen in a. Along the Γ − K(K0) directions the band crossings between the photon dispersion and the lower quasistatic band are protected, which gives rise to many additional linear-band degeneracies within the light-cone. As the cavity height is decreased, these additional Dirac points are driven towards the Γ point (opposite direction to the SDPs emerging from the TEM mode) where they annihilate one another. Comparing panels a and b, one can see that the dispersion near the CDPs does not noticeably change, and qualitative changes outside the light-cone only occur for small cavity heights where the higher-order cavity modes become increasingly detuned from the dipole resonances (see panels b-d). For small cavity heights, the nearest- neighbor approximation becomes increasingly accurate and indistinguishable from the full polariton dispersion (see d for example). Plots ph obtained with parameters ωK0 = 2.5ω0 and Ω = 0.01ω0. 15

Supplementary Figure 3. Numerical simulation of the polariton dispersion for a cavity-embedded honeycomb array of plasmonic nanorods. We consider ellipsoidal nanorods (prolate spheroids) with semiminor axis and semimajor axis of lengths 10 nm and 20 nm, 16 −1 respectively, and model them using the simple Drude model permittivity with plasma frequency ωp = 1.32 × 10 rad s and nonradiative damping frequency γ = 6.9 × 1013 s−1. The plasmonic nanorods are arranged in a honeycomb array with nearest-neighbor separation a = 45 nm. We apply perfect electric conductor boundary conditions for the cavity walls and periodic boundary conditions surrounding the unit cell (see inset on panel c). a-c, Numerical simulations of the polariton dispersion for the lowest two bands, calculated using the eigenmode solver in COMSOL Multiphysics, for cavity heights of a L = 3a, b L = 1.7a, and c L = 1.3a, respectively. The orange dots represent 15 −1 numerical solutions and the dashed line is a guide for the eye. The frequencies are normalized with respect to ω0 = 5.3 × 10 rad s .

Supplementary Figure 4. Numerical simulation of the polariton dispersion for a cavity-embedded honeycomb array of microwave helical resonators. We consider microwave helical resonators with minor radius 2 mm, major radius 6 mm, pitch 6 mm, and 5 turns (see inset on panel c). We model the helices and cavity walls as perfect electric conductors, which is a good approximation at microwave frequencies, and we apply periodic boundary conditions surrounding the unit cell. The helices are arranged in a honeycomb array with nearest-neighbor separation a = 45 mm. a-c, Numerical simulations of the polariton dispersion for the lowest two bands, calculated using the eigenmode solver in COMSOL Multiphysics, for cavity heights of a L = 3.33a, b L = 1.56a, and c L = 1.11a, respectively. The orange dots represent 9 −1 numerical solutions and the dashed line is a guide for the eye. The frequencies are normalized with respect to ω0 = 6.2 × 10 rad s . 16

SUPPLEMENTARY NOTE 1: BEYOND THE NEAREST-NEIGHBOR APPROXIMATION

Including all Coulomb interactions beyond the nearest-neighbor approximation leads to the full matter Hamiltonian within the dipole approximation

X † †  X h AB †  †  i Hmat = ~ω0 aqaq + bqbq + ~Ω fq bq aq + a−q + H.c. q q (S1) Ω X h   i Ω X h   i + ~ f AAa† a + a† + H.c. + ~ f BBb† b + b† + H.c. , 2 q q q −q 2 q q q −q q q where 3 X  a  f AA = f BB = −S + 1 − IAA eiq·R (S2) q q |R| R R6=0 and  3 AB X a AB iq·(R+e ) f = 1 − I e 1 . (S3) q |R + e | R R 1 In equations (S2) and (S3), the parameters  2  2 ∞ 2 lL − 1 ∞ 2 lL − 1 AA X |R| AB X |R+e1| IR = 2 5 , IR = 2 5 (S4)   2 2   2 2 l=1 1 + lL l=1 1 + lL |R| |R+e1| encode the renomalizations due to the interaction with cavity-induced image dipoles of neighboring meta-atoms, and S, given by equation (16) in the Methods, encodes renormalizations due to interactions with their self-images. We can recast the matter 1 P † mat † † T † † † Hamiltonian (S1) into matrix form as Hmat = 2 q ΨqHq Ψq, where Ψq = (ψq, ψ−q) and ψq = (aq, bq). The Hermitian mat 4 × 4 matrix Hq can be written in block form as

 mat mat  mat Hq Hq − ~ω012 Hq = mat mat , (S5) Hq − ~ω012 Hq with

 AA AB∗  mat ω0 + Ω Re fq Ω fq Hq = ~ AB BB . (S6) Ωfq ω0 + Ω Re fq

† † T † † † We diagonalize Hmat via a Bogoliubov transformation Ψq = TqXq, where Xq = (χq, χ−q) and χq = (βq+, βq−), with Tq † † T being a paraunitary matrix (see Methods). The bosonic operators βqτ = (cosh ϑqτ ψq − sinh ϑqτ ψ−q)|ψqτ i diagonalize the matter Hamiltonian (S1) as

X X mat † Hmat = ~ωqτ βqτ βqτ , (S7) τ=± q with quasistatic dispersion

r   mat AA AB ωqτ = ω0 + 2Ω< fq + 2τΩ fq . (S8) √ iϕq T As described in the main text, the spinors |ψqτ i = (1, τe ) / 2 describe a pseudospin degree of freedom where the two components encode the relative amplitude and phase of the dipolar oscillations on the two inequivalent A and B sublattices, re- AB mat p mat mat p mat spectively, where ϕq = arg(fq ). The coefficients cosh ϑqτ = (ω0 +ωqτ )/2 ω0ωqτ and sinh ϑqτ = (ω0 −ωqτ )/2 ω0ωqτ describe the contributions of the resonant and non-resonant terms, respectively. Neglecting non-resonant terms we obtain P † mat P P mat † † † Hmat = q ψqHq ψq which can be diagonalized as Hmat = τ=± q ~ωqτ βqτ βqτ , where the operators βqτ = ψq|ψqτ i mat AA AB and the corresponding dispersion ωqτ = ω0 + Ω< fq + τΩ fq take a simpler form. 17

In Supplementary Fig. 1a we plot the quasistatic dispersion within the nearest-neighbor approximation for different cavity heights and compare it to the full quasistatic dispersion in Fig. 1b, which is given by equation (S8). Indeed, due to the honeycomb symmetry, one observes that the type-I CDPs remain pinned at the K and K0 points. To prove this robustness against long-range AB Coulomb interactions, one must show that the function fq , given by equation (S3), vanishes at the corners of the Brillouin zone, i.e., one must show that the phase factors associated with the interactions with dipoles residing at the same separation distance and different sublattices sum to zero. To this aim, we define separation vectors Al l = l1a1 + l2a2 + e3 that connect √ 1 2 √ 3 3 meta-atoms residing on different sublattices, with primitive lattice vectors a1 = a( 3, 0) and a2 = a(− 2 , 2 ), and integers l1 and l2. We can then find six separation vectors with the same magnitude by exploiting the three-fold rotational symmetry Aj+ = O A , where the operator O rotates a vector by the angle 2πj/3 (j = 0, 1, 2), and the mirror symmetry about the l1l2 j l1l2 j y-axis Aj− = M Aj+ , where the operator M inverts the sign of the x component. Explicitly, the six separation vectors (not l1l2 x l1l2 x necessarily all different) read

"√ #   0± 3 1 Al l = ±a (2l1 − l2 − 1) xˆ + a (3l2 + 1) yˆ 1 2 2 2 "√ #   1± 3 1 Al l = ±a (−l1 − l2) xˆ + a (3l1 − 3l2 − 1) yˆ (S9) 1 2 2 2 "√ #   2± 3 1 Al l = ±a (−l1 + 2l2 + 1) xˆ + a (−3l1 + 1) yˆ 1 2 2 2

AB and evaluating equation (S3) at K one indeed finds fK = 0. AA Without a cavity, the fq term, given by equation (S2), introduces strong asymmetry between the upper and lower quasistatic bands. Since the lower band corresponds to a dark (↑↓) configuration of the two sublattices, the dispersion converges rapidly with increasing number of interactions taken into account as long-range Coulomb interactions quickly average to zero. The most prominent change is seen in the upper band where the group velocity increases towards the Γ point, which is in contrast to the dispersion within the nearest-neighbor approximation. This is because the upper quasistatic band corresponds to a bright (↑↑) configuration of the two sublattices and therefore long-range Coulomb interactions do not average to zero, especially near the Γ point. However, as the cavity height is reduced, the image dipoles quench long-range Coulomb interactions, which eventually become negligible for cavity heights on the order of the nearest-neighbor separation. The polariton dispersion plots presented in the main text (Fig. 2c-e) are within the nearest-neighbor approximation. In Sup- plementary Fig. 2a-d we present the polariton dispersion including all Coulomb interactions beyond nearest-neighbors until convergence, i.e., using equation (S6) instead of equation (28) in the polariton Hamiltonian (see Methods). As expected from the quasistatic dispersion plots in Supplementary Fig. 1b, long-range Coulomb interactions do not qualitatively affect the physics near the K and K0 points. In fact, for small cavity heights, the nearest-neighbor approximation becomes increasingly accurate and indistinguishable from the full polariton dispersion (see Supplementary Fig. 2d for example). As described in the main text, six inequivalent type-II SDPs emerge in the polariton spectrum due to the protected crossing with the TEM mode along the Γ − K(K0) directions. Protected crossings with higher-order cavity modes (m 6= 0) result in a large number of additional type-II Dirac points being generated within the light-cone for large cavity heights (see Supplemen- tary Fig. 2a). However, as the cavity height is decreased, these additional Dirac points migrate towards the Γ point (opposite direction to the SDPs), where they eventually merge with and annihilate one another. Comparing Supplementary Fig. 2a and Supplementary Fig. 2b, one observes that the polariton dispersion near the CDPs does not noticeably change, and qualitative changes outside the light-cone only occur at small cavity heights where the higher-order cavity modes become increasingly detuned with the dipole resonances.

SUPPLEMENTARY NOTE 2

In the Methods section we performed a unitary Schrieffer-Wolff transformation on the polariton Hamiltonian and obtained a two-band Hamiltonian in the matter sublattice space given by equation (36). In this section, we seek an effective Hamiltonian describing the polaritons near the corners of the Brillouin zone by expanding the two-band Bloch Hamiltonian (37) near the K (0) (1) (2) point (see Methods). To quadratic order in k = q − K, the diagonal matrix elements expand as Wq ' WK,k + WK,k + WK,k (0) (1) (2) and the off-diagonal elements as Fq ' FK,k + FK,k + FK,k, where the superscript index denotes the corresponding order in k. To determine the real and imaginary components in the expansion, we must sum the phase contributions from the degenerate photons at the K point. To this aim, we define reciprocal lattice vectors Gn = n1b1 + n2b2 with primitive vectors b1 = 18 √ 2π 4π 3a ( 3, −1) and b2 = 3a (0, 1). In analogy with the analysis in Supplementary Note 1, we can find the reciprocal lattice vectors associated with six degenerate photonic modes at K by exploiting the three-fold rotational symmetry of the reciprocal lattice j+ Gn = K − Oj(K − Gn), where the operator Oj rotates a vector by the angle 2πj/3 (j = 0, 1, 2), and the mirror symmetry j− j+ about the ky = 0 line Gn = MyGn , where the operator My inverts the sign of the y-component. Explicitly, the six reciprocal lattice vectors (not necessarily all different) read

 2π  2π  G0µ = K + √ (3n − 2) xˆ + µ (2n − n ) yˆ, n 3 3a 1 3a 2 1     1µ 2π 2π G = K + √ (1 − 3n ) xˆ + µ (2n − n2 − 1) yˆ, (S10) n 3 3a 2 3a 1  2π  2π  G2µ = K + √ (1 + 3n − 3n ) xˆ + µ (1 − n − n ) yˆ, n 3 3a 2 1 3a 1 2 with µ = ±1. Exploiting these symmetries one can rewrite the summations over photonic indices as

2 X (υ) 1 X (υ) X X ∆ · · · → ∆ ... (S11) Kmn 6 Kmn mn mn j=0 µ=± where the factor of 1/6 is introduced to avoid over counting. First, the constant (zeroth order) term in the expansion of Wq is

2 (0) Ω X (0) X X W = ω − ΩS − ∆ 1 =ω ¯ . (S12) K,k 0 6 Kmn 0 mn j=0 µ=±

The linear (first order) term vanishes as

2 2 (1) Ωa X (1) X X W = ∆ (K − Gjµ) k + (K − Gjµ) k  = 0, (S13) K,k 6 Kmn n x x n y y mn j=0 µ=± which illustrates that the light-matter interaction does not introduce anisotropic warping of the Dirac cone to first order. Finally, the quadratic (second order) term is

4 2 (2) Ωa X (2) X X W = − ∆ (K − Gjµ) (K − Gjµ) k k K,k 6 Kmn n x n y x y mn j=0 µ=± 4 2 Ωa X (2) X X − ∆ (K − Gjµ)2 k2 + (K − Gjµ)2k2 12 Kmn n x x n y y mn j=0 µ=± (S14) 2 2 Ωa X (1) X X + ∆ (k2 + k2) 12 Kmn x y mn j=0 µ=± 2 2 = − D(kx + ky),

The constant (zeroth order) term in the expansion of Fq vanishes at the high-symmetry points as

2 (0) Ω X (0) X X F = − ∆ φ2 = 0, (S15) K,k 6 Kmn njµ mn j=0 µ=±

jµ 0 where φnjµ = exp(iaGn · yˆ). This shows that the CDPs remain pinned at the K and K points and are completely robust against the coupling to the surrounding photonic environment. The linear (first order) term reads

2 2 (1) 3Ωa Ωa X (1) X X F = − (1 − I) k + ik  + ∆ φ2 (K − Gjµ) k + (K − Gjµ) k  K,k 2 x y 6 Kmn njµ n x x n y y mn j=0 µ=± (S16)

= −v¯(kx + iky), 19

Finally, the quadratic (second order) term is given by

2 2 2 (2) 3Ωa 2 Ωa X (1) X X F = (1 − I) k − ik  + ∆ φ2 (k2 + k2) K,k 8 x y 12 Kmn njµ x y mn j=0 µ=± 4 2 Ωa X (2) X X − ∆ φ2 (K − Gjµ)2 k2 + (K − Gjµ)2k2 12 Kmn njµ n x x n y y mn j=0 µ=± (S17) 4 2 Ωa X (2) X X − ∆ φ2 (K − Gjµ) (K − Gjµ) k k 6 Kmn njµ n x n y x y mn j=0 µ=± ¯ 2 = t kx − iky .

¯ eff P † ¯eff We can now identify the effective Hamiltonian HK = k ψkHK,kψk in the matter sublattice space, where the effective Bloch ¯eff Hamiltonian HK,k is given by equation (9) in the main text.