15
Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts 02215, USA (Dated: July 17, 2019) The possibility that anyons — quantum particles other than fermions or bosons — can emerge in condensed matter systems has motivated generations of physicists. In addition to being of funda- mental scientific importance, so-called non-Abelian anyons are particularly sought-after for potential applications to quantum computing. However, experimental evidence of anyons in electronic systems remains inconclusive. In this paper, we propose to demonstrate non-Abelian braiding statistics by injecting coherent states of light into “topological guided modes” in specially-fabricated photonic waveguide arrays. These modes are photonic analogues of topological zero modes in electronic systems. Light traveling inside spatially well-separated topological guided modes can be braided, leading to the accumulation of non-Abelian phases. We propose an optical interference experiment to probe this non-Abelian braiding statistics directly. For decades, it was thought that the spin-statistics theorem 1 constrained all quantum particles to fall into one of two categories: bosons, with integer spin and many-body wavefunctions that are symmetric un- der particle interchange, and fermions, with half- integer spin and antisymmetric many-body wavefunc- tions. However, in a revolution of fundamental quantum mechanics, it was realized 2 that, in two spatial dimen- sions, many-body wavefunctions can acquire any phase between 0 (bosons) and π (fermions) when two parti- cles are exchanged. It was later understood that such particles, called (Abelian) anyons, 3,4 naturally emerge as fractionally-charged quasiparticles in the fractional quantum Hall (FQH) effect. 5,6 Later, it was found that another type of exchange statistics, dubbed non- Abelian, is possible, in which particle exchange does not yield merely a phase factor, but instead gives rise to a nontrivial unitary transformation in a space of de- generate many-body ground states. 7,8 Subsequent the- ory work found that such non-Abelian anyons exist in certain model FQH states, 9–11 and, in the past decade, non-Abelian anyons have become particularly desirable for potential applications to topological quantum com- puting. 12–14 However, while indirect experimental signa- tures have emerged, 15,16 there has not yet been any con- clusive observation of non-Abelian statistics in a physi- cal system. Besides FQH physics, another route to non-Abelian statistics lies in the remarkable physics of zero modes. As was pointed out in pioneering work by R. Jackiw with C. Rebbi 17 and P. Rossi, 18 localized zero-energy fermionic states can bind to topological defects in an order parameter, such as kinks in one dimension and vortices in two dimensions. In systems where electric charge is a good quantum number, these zero modes carry charges that are fractions of the “fundamental” electron charge. 17,19,20 In chiral superconductors, where charge conservation is broken, these localized modes are Majorana bound states with non-Abelian statis- tics. 18,21–23 In this paper, we propose a novel means of realizing topological zero modes in photonic, rather than elec- tronic, systems, and demonstrating their non-Abelian statistics directly. The realization that we present con- sists of non-interacting photons propagating in the non- trivial background of a photonic lattice with topologi- cal defects whose positions are controllable. Light chan- neled into “topological guided modes” localized at these defects can be braided, leading to the accumulation of non-Abelian phases. The following questions naturally spring to mind: How can a system of non-interacting photons possibly exhibit non-Abelian braiding? How does a manifold of degenerate multi-defect states, re- quired for non-Abelian statistics, emerge out of this non- interacting system? The answers to these questions will be simple, once the details of our proposal are offered below. As we shall see, it is the topological nature of the zero modes that underlies all non-Abelian properties that follow. The zero modes we propose to realize are photonic analogues of Kekul´ e zero modes in graphene, 20 which are bound to vortices in the complex order parameter Δ(r) describing a dimerization pattern in the hexagonal lattice (Fig. 1). The translation between electrons and photons is achieved by replacing the sites of the lattice with waveguides embedded in a bulk optical medium (e.g. fused silica), 24 which are extended in the z direc- tion and whose x-y positions mimic the 2D positions of the carbon atoms in the distorted graphene lattice. The wave equation for the paraxial propagation of light in such a waveguide array maps directly onto the time- dependent Schr¨ odinger equation (SE), where the time coordinate t in the SE is replaced by the coordinate z along the direction of light propagation. This wave equation can be further mapped, using coupled mode theory, 25 to a linear differential equation that is in one- to-one correspondence with the noninteracting tight- binding model of the electronic system. The waveguides themselves can therefore be thought of as the world lines of the carbon atoms, with straight waveguides corre- arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015

arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

Braiding light quanta

Thomas Iadecola, Thomas Schuster, and Claudio ChamonPhysics Department, Boston University, Boston, Massachusetts 02215, USA

(Dated: July 17, 2019)

The possibility that anyons — quantum particles other than fermions or bosons — can emerge incondensed matter systems has motivated generations of physicists. In addition to being of funda-mental scientific importance, so-called non-Abelian anyons are particularly sought-after for potentialapplications to quantum computing. However, experimental evidence of anyons in electronic systemsremains inconclusive. In this paper, we propose to demonstrate non-Abelian braiding statistics byinjecting coherent states of light into “topological guided modes” in specially-fabricated photonicwaveguide arrays. These modes are photonic analogues of topological zero modes in electronicsystems. Light traveling inside spatially well-separated topological guided modes can be braided,leading to the accumulation of non-Abelian phases. We propose an optical interference experimentto probe this non-Abelian braiding statistics directly.

For decades, it was thought that the spin-statisticstheorem1 constrained all quantum particles to fallinto one of two categories: bosons, with integer spinand many-body wavefunctions that are symmetric un-der particle interchange, and fermions, with half-integer spin and antisymmetric many-body wavefunc-tions. However, in a revolution of fundamental quantummechanics, it was realized2 that, in two spatial dimen-sions, many-body wavefunctions can acquire any phasebetween 0 (bosons) and π (fermions) when two parti-cles are exchanged. It was later understood that suchparticles, called (Abelian) anyons,3,4 naturally emergeas fractionally-charged quasiparticles in the fractionalquantum Hall (FQH) effect.5,6 Later, it was foundthat another type of exchange statistics, dubbed non-Abelian, is possible, in which particle exchange doesnot yield merely a phase factor, but instead gives riseto a nontrivial unitary transformation in a space of de-generate many-body ground states.7,8 Subsequent the-ory work found that such non-Abelian anyons exist incertain model FQH states,9–11 and, in the past decade,non-Abelian anyons have become particularly desirablefor potential applications to topological quantum com-puting.12–14 However, while indirect experimental signa-tures have emerged,15,16 there has not yet been any con-clusive observation of non-Abelian statistics in a physi-cal system.

Besides FQH physics, another route to non-Abelianstatistics lies in the remarkable physics of zero modes.As was pointed out in pioneering work by R. Jackiwwith C. Rebbi17 and P. Rossi,18 localized zero-energyfermionic states can bind to topological defects in anorder parameter, such as kinks in one dimension andvortices in two dimensions. In systems where electriccharge is a good quantum number, these zero modescarry charges that are fractions of the “fundamental”electron charge.17,19,20 In chiral superconductors, wherecharge conservation is broken, these localized modesare Majorana bound states with non-Abelian statis-tics.18,21–23

In this paper, we propose a novel means of realizingtopological zero modes in photonic, rather than elec-tronic, systems, and demonstrating their non-Abelianstatistics directly. The realization that we present con-sists of non-interacting photons propagating in the non-trivial background of a photonic lattice with topologi-cal defects whose positions are controllable. Light chan-neled into “topological guided modes” localized at thesedefects can be braided, leading to the accumulation ofnon-Abelian phases. The following questions naturallyspring to mind: How can a system of non-interactingphotons possibly exhibit non-Abelian braiding? Howdoes a manifold of degenerate multi-defect states, re-quired for non-Abelian statistics, emerge out of this non-interacting system? The answers to these questions willbe simple, once the details of our proposal are offeredbelow. As we shall see, it is the topological nature ofthe zero modes that underlies all non-Abelian propertiesthat follow.

The zero modes we propose to realize are photonicanalogues of Kekule zero modes in graphene,20 whichare bound to vortices in the complex order parameter∆(r) describing a dimerization pattern in the hexagonallattice (Fig. 1). The translation between electrons andphotons is achieved by replacing the sites of the latticewith waveguides embedded in a bulk optical medium(e.g. fused silica),24 which are extended in the z direc-tion and whose x-y positions mimic the 2D positionsof the carbon atoms in the distorted graphene lattice.The wave equation for the paraxial propagation of lightin such a waveguide array maps directly onto the time-dependent Schrodinger equation (SE), where the timecoordinate t in the SE is replaced by the coordinatez along the direction of light propagation. This waveequation can be further mapped, using coupled modetheory,25 to a linear differential equation that is in one-to-one correspondence with the noninteracting tight-binding model of the electronic system. The waveguidesthemselves can therefore be thought of as the world linesof the carbon atoms, with straight waveguides corre-

arX

iv:1

509.

0540

8v3

[co

nd-m

at.m

es-h

all]

20

Nov

201

5

Page 2: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

2

sponding to a lattice that is static in time.Waveguide arrays of this type have already been real-

ized, written into bulk materials with exquisite precisionusing femtosecond lasers,26 (see Ref. 24 for a review).They have been used successfully to mimic graphene andother electronic systems with both static lattices27–30

and ones that change as a function of time.31,32 Theexperimental protocol that we suggest here hinges onthis capability to execute a braiding procedure in whichthree vortices are wound around one another. As weshow below, a protocol can be chosen that reveals thenon-Abelian nature of the braiding directly via interfer-ence between different braiding patterns.

Our starting point is the coupled-mode equation forparaxial light propagation through the waveguide array,

i ∂z ψr(z) =

3∑j=1

Hr,r±sj ψr±sj (z). (1)

Here, the vector r is defined on a hexagonal lattice di-vided into two interpenetrating triangular sublatticesthat we call A and B. The vectors +sj (j=1,2,3) con-nect the point r in sublattice A to its three nearestneighbors in sublattice B, which are located a distancea away, and −sj connect points in B to their neigh-bors in A. For simplicity, we assume the light in thearray to be monochromatic, so that Eq. (1) is frequency-independent. Eq. (1) is formally identical to a nonin-teracting tight-binding model in which the time coordi-nate t has been replaced by the longitudinal coordinatez. Its solutions can be decomposed into eigenmodesψν,r(z) = ψν,r e

−iκνz, where ν = 0, . . . , 2N − 1, with Nthe number of unit cells in the waveguide array, labelsthe “energy eigenvalues” κν . The “wavefunction” ψr(z)in Eq. (1) is related to the amplitude of the local electricfield Er(z, t) on the “site” r. It is useful to work withthe quantized electric field operator33

Er(z, t) =∑ν

ψν,r ei[(kω−κν)z−ωt] bν n + H.c. (2)

where n is a unit vector in the x-y plane describing thepolarization of the field and kω satisfies the dispersionrelation of a single waveguide. The ladder operators

bν satisfy the bosonic commutation relations [bν , b†ν′ ] =

δν,ν′ . [See Supplemental Material (SM) for details onthe quantization procedure.]

The “Hamiltonian” in Eq. (1) depends exponen-tially on distances between nearest-neighbor waveg-uides (we neglect longer-range couplings for the mo-ment). We take Hr,r+sj = −t − δtr,j , where δtr,j =

∆(r) eiK+·sj e2iK+·r/2 + c.c. (we denote by K± the lo-cations of the two inequivalent Dirac points, at oppo-site corners of the hexagonal Brillouin zone). Here, theparameter t describes the evanescent couplings of thewaveguides, and the position-dependent function δtr,j

B

A

FIG. 1: (A) Uniform Kekule distortion in a hexagonal waveg-uide lattice at a slice of constant z. The faint circles repre-sent the original hexagonal lattice, while darker circles repre-sent the distorted lattice. Sublattice A is colored red, whilesublattice B is colored blue. The overlaid vector field rep-resents the magnitude and direction of the order parameter∆(r). The background is an image of a low-lying eigenmodeof Eq. (1). The intensity pattern represents the amplitudeof the electric field in each waveguide. (B) Waveguide lat-tice in the presence of a vortex in the Kekule pattern, withan overlaid order-parameter vector field showing circulationaround the vortex core. The background is an intensity plotof the localized zero-mode wavefunction. We have chosen avortex profile in which only sublattice A is displaced.

describes modulations of these couplings due to dis-placements of the waveguides from their original x-ypositions. The complex-valued Kekule order parame-ter ∆(r) controls the distortion of the lattice.20 A vor-tex in ∆(r) is a defect in this distortion pattern, butnot the lattice itself. The order parameter in the pres-ence of a single vortex centered at the origin reads

Page 3: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

3

∆(r) = ∆0(r) ei(α−θ) in polar coordinates r = (r, θ).Here, ∆0(r) = ∆ tanh(r/`0) describes the spatial pro-file of the vortex, which has a core radius `0, and α isthe phase of the order parameter.

The zero mode in the presence of this vortex profilecan be found by setting the left-hand side of Eq. (1)to zero and solving for ψr. The zero-mode solution istightly localized near the core of the vortex, with a sizeof order 1/∆, and has support on sublattice A only [seeFig. 1(B)]. [If we send θ → −θ in ∆(r), the zero modehas support on sublattice B instead.20] This means thatlight propagating in the zero mode travels as if confinedto an “optical fiber” located at the vortex core, albeitwith evanescent decay into neighboring waveguides inthe same sublattice. However, the zero mode differscrucially from a mode in an optical fiber, both becauseit takes the distortion of an entire waveguide lattice tocreate, and because it depends on the topological natureof this distortion. These “topological guided modes” areresponsible for the non-Abelian effects described below.

The above discussion neglects the presence of var-ious lattice-dependent effects that can shift the zeromode eigenvalue to a finite momentum. In principle,this in turn makes possible the accumulation of path-dependent (i.e. non-topological) dynamical phases dur-ing braiding. However, as we show in detail in theSM, these effects can be neglected to resonable preci-sion without fine-tuning. There, we estimate (conserva-tively) that we can perform thousands of braids beforethe effects of dynamical phases begin to manifest them-selves.

To facilitate our discussion of the non-Abelian braid-ing of vortices like the one above, we will work in thecontinuum limit, where the analytical form of the vor-tex wavefunction is known. (Exact diagonalization cal-culations corroborating these results are reviewed in theSM.) On length scales much longer than a, the operatorthat annihilates a photon in the zero mode is

b0 =

∫dr[u(r) b+(r) + u∗(r) b−(r)

], (3)

where the operators b±(r) annihilate a photon at theDirac points K±. (These operators are assumed to benormalized such that the canonical commutation rela-tion [b0, b

†0] = 1 holds.) The function u(r) ≡ u(r) is

given, up to normalization, by20

u(r) = ei(α/2+π/4) e−∫ r0

dr′∆0(r′). (4)

The quantum states created by b†0 can be connected tothe macroscopic state of light in the waveguide array bydefining the coherent state

|λ〉 = e−|λ|2/2 eλb

†0 |0〉 (5)

with mean photon number 〈b†0b0〉 = |λ|2. These statesprovide a faithful semiclassical description of the electric

A B

z

B2B1B1B2

R1R2

R3R1

R2R3

FIG. 2: Depiction of the braids (A) B1B2 and (B) B2B1

used in the proposed experiment. The strands being braidedrepresent the positions Ri(z) of the three vortex cores, intowhich coherent light has been injected. The order of thestrands coincides on the input and output facets at the bot-tom and top, respectively.

field in the waveguide array in the large-|λ| limit whenthe input light is a coherent, monochromatic laser spotcentered on the vortex core. The amplitude and phaseof the complex number λ correspond, in this limit, tothe amplitude and phase of the classical electric field inthe vicinity of the vortex core.

The quantum states obtained by channeling light intothe topological guided mode are excited states of theHamiltonian in Eq. (1). These excitations are stablebecause the photons are non-interacting, and becausethe photonic zero modes are separated from other eigen-states by a gap. More accurately, the photons are veryweakly interacting inside the waveguides, guaranteeingthat these excitations are sufficiently long-lived for lightto propagate throughout the depth of the photonic crys-tal without decaying into other states.

Let us now study the statistics of zero modes in aninfinite system with v vortices at z-dependent positionsRi(z) (i = 1, . . . , v). The order parameter in the pres-ence of this vortex configuration is given by

∆(r) = ∆

n∏j=1

tanh(|r −Rj |/`0) ei[αj−arg(r−Rj)] , (6)

and the zero-mode Hilbert space is spanned by the op-

erators b†0,i. A clockwise adiabatic exchange of vor-tices i and i+ 1 is implemented by winding the vortex-core coordinates Ri(z) and Ri+1(z) around one anotheras functions of increasing z (see Fig. 2). The non-trivial nature of this winding process stems from thefact that the zero mode amplitude u(r) is double-valuedunder α → α − 2π. (We verified the double valued-ness of the zero-mode wavefunction on the lattice via

Page 4: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

4

a numerical tight-binding calculation, as we describe inthe SM.) This double-valuedness is a hallmark of thetopological guided modes we consider, but not of non-topological light-trapping modes. The effect of this ex-

change, up to a gauge choice, is to map b†0,i → b†0,i+1

and b†0,i+1 → −b†0,i while leaving all other vortex opera-

tors unchanged, similarly to the case of Majorana22 andDirac zero modes34 in electronic systems. The operatorτi that implements this exchange is given by

τi = eπ (b†0,i+1b0,i−b†0,ib0,i+1)/2. (7)

One verifies by direct calculation that these operatorssatisfy [τi, τj ] = 0 for |i− j| > 1 and τiτj τi = τj τiτj for|i− j| = 1, and therefore form a unitary representationof Bv, the braid group on v strands.14 The action ofthese generators on a state

|n1, . . . , nv〉 =

v∏i=1

(b†0,i)ni

√ni!|0〉, (8)

representing a system with v vortices, each with a fixednumber of photons, is given by

τi | . . . , ni, ni+1, . . . 〉 = (−1)ni+1 | . . . , ni+1, ni, . . . 〉.(9)

We now explain the consequences of the above phasefactor, which arises purely from braiding the vortices.Consider the case of two vortices, and the operationof winding one around the other, which restores themto their initial locations. We begin by considering i)an example where this phase factor does not lead tounitary operations on a multi-dimensional space of de-generate states, and then turn to ii) an example whereit does. For case i), consider any eigenstate of occu-pation number, |n1, n2〉, which upon winding goes tothe state (−1)n1+n2 |n1, n2〉; clearly, the initial and finalstates are equal up to a phase, and the braiding opera-tion acts on a one-dimensional space only. In contrast,consider case ii), where we take an initial state that isnot an eigenstate of occupation number, and instead isa superposition of states with occupation 0 and 1, forexample. Then winding takes

|0〉+|1〉√2⊗ |0〉+|1〉√

2−→ |0〉−|1〉√

2⊗ |0〉−|1〉√

2. (10)

Clearly, the initial and final states are different (here,they are orthogonal). This is a crucial hallmark of thenon-Abelian nature of the action of the braiding gen-erators on these states. Thus, the essential ingredi-ent necessary for braiding to connect states in a multi-dimensional Hilbert space is that the initial quantumstates loaded into the zero modes are superpositions ofstates with even and odd numbers of photons.

Such superpositions do not only exist within the do-main of quantum optics. The coherent states defined in

Eq. (5) consist of a superposition of photon states withall occupation numbers, and can be created by shininga coherent laser beam centered on a single vortex core.Let us now consider a system of v vortices, into each ofwhich is loaded a coherent state of photons:

|λ1, . . . , λv〉 =

v∏i=1

e−|λi|2/2 eλb

†0,i |0〉. (11)

The action of the braiding generators on these states isfound from Eq. (9) to be

τi| . . . , λi, λi+1, . . . 〉 = | . . . ,−λi+1, λi, . . . 〉. (12)

To see that this braiding operation connects quantumstates in a multi-dimensional space of degenerate states,consider again the winding of two vortices, which re-stores their initial locations. For a coherent state, thisoperation takes

|λ1, λ2〉 −→ | − λ1,−λ2〉 . (13)

The overlap of these two states is 〈λ1, λ2| − λ1,−λ2〉 =

e−2(|λ1|2+|λ2|2), whose magnitude is smaller than one, in-dicating that the states are linearly independent. Thisdemonstrates that the Hilbert space spanned by thebraiding operations is multi-dimensional. Moreover, inthe limit where a large number of photons are loadedinto the zero modes (large |λ1,2|, which is to be expectedfrom a laser), the overlap is exponentially small, and theinitial and final states are effectively orthogonal. If onerestricts oneself to braids that bring all vortices backto their original positions, the dimension of the Hilbertspace that can be accessed by braiding the vortices is2v−1. This is because such braids always take λi → ±λi,subject to the parity constraint that the number of mi-nus signs must always be even, as in Eq. (13).

At this point, it is worthwhile to reflect on the signif-icance of the fact that the effect of braiding the vorticesmanifests itself at the level of coherent states, whichare essentially classical. While the braiding of thesevortices can be understood at the quantum (i.e. few-photon) level, its effects permeate the entire zero-modeHilbert space for arbitrary occupation numbers. Conse-quently, the quantum action of the braiding generatorssurvives the limit of large occupations, so that macro-scopic effects of this braiding can be seen. In particular,observe that λi → −λi under a braid corresponds, inthe limit of large |λi|, to a change in the sign of theelectric field Ei near the core of the ith vortex. Thisobservation forms the basis of our discussion below.

We now propose an experiment that could provide di-rect and unambiguous evidence for non-Abelian statis-tics at the level of coherent states. The photonic non-Abelian interferometer (PNAI) that we propose consistsof two separate photonic lattices, each with three vor-tices, with waveguides written into the host medium

Page 5: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

5

B1B2

B2B1

FIG. 3: Schematic of the proposed photonic non-Abelianinterferometer. A coherent laser beam (one of three, onefor each vortex in a lattice) passes through a 50/50 beamsplitter (blue) and simultaneously enters two separate pho-tonic lattices fabricated with the two braids to be compared.The two output beams are reflected by mirrors (grey) intoanother beam splitter that interferes the two signals and out-puts the sum and difference to separate screens (black). Forthe choice of braids used here, one screen shows a brightand two dark spots, while the other screen shows the “logi-cal complement” of the first, with one dark and two brightspots.

in such a way that the vortex cores wind adiabati-cally around one another according to a specific pattern.(The adiabatic condition here corresponds to demand-ing that any change in the position of each waveguideas a function of z occur on length scales much largerthan the inverse photonic bandgap.) In one lattice, the

waveguide pattern executes a braid B1B2 (see Fig. 2),while in the other, the waveguide pattern implementsthe braid B2B1. Interfering the light output from thetwo lattices (see Fig. 3) reveals the defining featureof non-Abelian statistics, namely that performing thesame braids in different orders yields different results.We now proceed through each stage of the interferome-ter setup.

We start with the input stage of the PNAI. We assumethat the light input into the interferometer comes fromthree monochromatic, quantum-coherent laser sources,each focused on the core of a single vortex so that the in-put light has large overlap with the zero modes. Each ofthe three input beams is split by a 50/50 dielectric beamsplitter, so that the light entering each vortex core comesfrom the same source beam. We will denote the lightentering the upper branch of the interferometer (i.e. af-

ter the beam splitter) as |λ1, λ2, λ3〉, and |λ1, λ2, λ3〉 forthe lower branch (see Fig. 3). (The first beam splitter

enforces the phase relation λj = −i λj .)For the braiding stage of the interferometer, we

choose, for example, B1 = τ2τ1 and B2 = τ2τ−11 τ2τ

−11

(see Fig. 2), which ensures that the vortex cores on theoutput facets of both lattices are in the same order.35 Inthe SM, we provide more information about how thesebraiding patterns can be written into the waveguide ar-ray. The output states from the two braids are

B1B2 |λ1, λ2, λ3〉 = | − λ1, λ2,−λ3〉 (14)

and

B2B1 |λ1, λ2, λ3〉 = | − λ1,−λ2, λ3〉. (15)

In the final stage of the interferometer, the outputbeams from the braiding stage are combined at anotherbeam splitter. The sign differences between the coher-ent states exiting the two branches of the interferometercause the light to interfere constructively at one detec-tor and destructively at the other. Which detector thisis for each vortex depends on the relative signs of λiand λi, which in turn depend on the braids (see Fig. 3,and SM for more details). Since the effects of dynamicalphases can be heavily suppressed in a controlled manner,as discussed earlier, the only source of this interferenceis the noncommutativity of braiding the vortices.

In summary, we have demonstrated in this papera means to realize topological zero modes with non-Abelian statistics in photonic lattices. Owing to thebosonic statistics of photons, the non-Abelian statisticsof the zero modes survives the limit of macroscopic oc-cupation, so long as a coherent state of light propagatesin the waveguide array. We further proposed a photonicnon-Abelian interferometer to detect unambiguous sig-natures of non-Abelian statistics. This device can befabricated with current technology and would providethe first direct and conclusive evidence of non-Abelianstatistics.

Acknowledgments

We thank Luca Dal Negro for inspiring discussions,and Chang-Yu Hou, Christopher Mudry, Titus Neu-pert, and Luiz H. Santos for valuable feedback on themanuscript. T.I. was supported by the National ScienceFoundation Graduate Research Fellowship Program un-der Grant No. DGE-1247312, and C.C. was supportedby DOE Grant DEF-06ER46316.

Page 6: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

6

1 W. Pauli, Phys. Rev. 58, 716 (1940), URL http://link.

aps.org/doi/10.1103/PhysRev.58.716.2 J. Leinaas and J. Myrheim, Il Nuovo Cimento B 37, 1

(1977), URL http://dx.doi.org/10.1007/BF02727953.3 F. Wilczek, Phys. Rev. Lett. 48, 1144 (1982), URL http:

//link.aps.org/doi/10.1103/PhysRevLett.48.1144.4 F. Wilczek, Phys. Rev. Lett. 49, 957 (1982), URL http:

//link.aps.org/doi/10.1103/PhysRevLett.49.957.5 B. I. Halperin, Phys. Rev. Lett. 52, 1583 (1984), URLhttp://link.aps.org/doi/10.1103/PhysRevLett.52.

1583.6 D. Arovas, J. R. Schrieffer, and F. Wilczek, Phys. Rev.

Lett. 53, 722 (1984), URL http://link.aps.org/doi/

10.1103/PhysRevLett.53.722.7 J. Frohlich, in Nonperturbative Quantum Field Theory,

edited by G. ‘t Hooft, A. Jaffe, G. Mack, P. Mit-ter, and R. Stora (Springer US, 1988), vol. 185of NATO ASI Series, pp. 71–100, ISBN 978-1-4612-8053-8, URL http://dx.doi.org/10.1007/

978-1-4613-0729-7_4.8 E. Witten, Communications in Mathematical Physics

121, 351 (1989), ISSN 0010-3616, URL http://dx.doi.

org/10.1007/BF01217730.9 X. G. Wen, Phys. Rev. Lett. 66, 802 (1991), URL http:

//link.aps.org/doi/10.1103/PhysRevLett.66.802.10 G. Moore and N. Read, Nuclear Physics B

360, 362 (1991), ISSN 0550-3213, URL http:

//www.sciencedirect.com/science/article/pii/

055032139190407O.11 N. Read and E. Rezayi, Phys. Rev. B 59, 8084 (1999),

URL http://link.aps.org/doi/10.1103/PhysRevB.

59.8084.12 A. Kitaev, Annals of Physics 303, 2 (2003), ISSN 0003-

4916, URL http://www.sciencedirect.com/science/

article/pii/S0003491602000180.13 M. Freedman, A. Kitaev, M. Larsen, and Z. Wang,

Bulletin of the American Mathematical Society 40, 31(2002).

14 C. Nayak, S. H. Simon, A. Stern, M. Freedman, andS. Das Sarma, Rev. Mod. Phys. 80, 1083 (2008),URL http://link.aps.org/doi/10.1103/RevModPhys.

80.1083.15 S. An, P. Jiang, H. Choi, W. Kang, S. H. Simon,

L. N. Pfeiffer, K. W. West, and K. W. Baldwin,arXiv:1112.3400.

16 R. L. Willett, C. Nayak, K. Shtengel, L. N.Pfeiffer, and K. W. West, Phys. Rev. Lett. 111,186401 (2013), URL http://link.aps.org/doi/10.

1103/PhysRevLett.111.186401.17 R. Jackiw and C. Rebbi, Phys. Rev. D 13, 3398 (1976),

URL http://link.aps.org/doi/10.1103/PhysRevD.

13.3398.18 R. Jackiw and P. Rossi, Nuclear Physics B

190, 681 (1981), ISSN 0550-3213, URL http:

//www.sciencedirect.com/science/article/pii/

0550321381900444.19 W. P. Su, J. R. Schrieffer, and A. J. Heeger, Phys. Rev.

Lett. 42, 1698 (1979), URL http://link.aps.org/doi/

10.1103/PhysRevLett.42.1698.20 C.-Y. Hou, C. Chamon, and C. Mudry, Phys. Rev. Lett.

98, 186809 (2007), URL http://link.aps.org/doi/10.

1103/PhysRevLett.98.186809.21 N. Read and D. Green, Phys. Rev. B 61, 10267 (2000),

URL http://link.aps.org/doi/10.1103/PhysRevB.

61.10267.22 D. A. Ivanov, Phys. Rev. Lett. 86, 268 (2001), URL http:

//link.aps.org/doi/10.1103/PhysRevLett.86.268.23 A. Y. Kitaev, Physics-Uspekhi 44, 131 (2001), URL

http://stacks.iop.org/1063-7869/44/i=10S/a=S29.24 A. Szameit and S. Nolte, Journal of Physics B: Atomic,

Molecular and Optical Physics 43, 163001 (2010).25 F. Lederer, G. I. Stegeman, D. N. Christodoulides, G. As-

santo, M. Segev, and Y. Silberberg, Physics Reports 463,1 (2008), ISSN 0370-1573.

26 K. M. Davis, K. Miura, N. Sugimoto, and K. Hirao,Opt. Lett. 21, 1729 (1996), URL http://ol.osa.org/

abstract.cfm?URI=ol-21-21-1729.27 O. Peleg, G. Bartal, B. Freedman, O. Manela,

M. Segev, and D. N. Christodoulides, Phys. Rev. Lett.98, 103901 (2007), URL http://link.aps.org/doi/10.

1103/PhysRevLett.98.103901.28 O. Bahat-Treidel, O. Peleg, and M. Segev, Opt. Lett.

33, 2251 (2008), URL http://ol.osa.org/abstract.

cfm?URI=ol-33-19-2251.29 M. C. Rechtsman, J. M. Zeuner, A. Tunnermann,

S. Nolte, M. Segev, and A. Szameit, Nature Photonics7, 153 (2012).

30 S. Mukherjee, A. Spracklen, D. Choudhury, N. Goldman,P. Ohberg, E. Andersson, and R. R. Thomson, Phys. Rev.Lett. 114, 245504 (2015), URL http://link.aps.org/

doi/10.1103/PhysRevLett.114.245504.31 M. C. Rechtsman, J. M. Zeuner, Y. Plotnik, Y. Lumer,

D. Podolsky, F. Dreisow, S. Nolte, M. Segev, and A. Sza-meit, Nature (London) 496, 196 (2013).

32 Y. E. Kraus, Y. Lahini, Z. Ringel, M. Verbin, andO. Zilberberg, Phys. Rev. Lett. 109, 106402 (2012), URLhttp://link.aps.org/doi/10.1103/PhysRevLett.109.

106402.33 W. Vogel and D.-G. Welsch, Quantum Optics (Wiley-

VCH, Weinheim, GER, 2006).34 S. Yasui, K. Itakura, and M. Nitta, Nuclear

Physics B 859, 261 (2012), ISSN 0550-3213, URLhttp://www.sciencedirect.com/science/article/

pii/S0550321312000831.35 Of course, this is not the only braid we could have chosen.

For example, the choice B1 = τ1τ2 and B2 = τ2τ1 alsosatisfies the requirement that vortices end up in the sameorder in both braids.

Supplemental Material: Braiding light quanta

Page 7: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

7

Appendix A: Quantized electromagnetism in a waveguide lattice

We seek to quantize the electromagnetic field in the presence of a spatially varying relative permittivity ε(r), andfollow the notation of Quantum Optics, by Vogel and Welsch.33 (We adopt the convention here that r is a three-dimensional vector when it enters the argument of a function, and a two-dimensional vector in the x-y plane whenit appears as a subscript.) We denote the vector potential by A(r, t), and work in the Coulomb gauge ∇ ·A = 0.In the absence of free charges we also have A0 = 0. We work under the assumption that ∇ε(r) ·A is small. Withthis the classical Hamiltonian is

H =

∫d3r

(1

ε(r)ε0Π2 +

1

µ0[∇×A]2

)(A1)

with Π(r, t) the momentum conjugate to A(r, t):

Π(r, t) = ε(r)ε0 A(r, t). (A2)

We then have the canonical Poisson brackets:

{A(r, t),Π(r′, t)} = δ(r − r′). (A3)

We quantize Eq. (A1) by promoting A and Π to operators obeying equal-time commutation relations

[A(r, t), Π(r′, t)] = ih δ(r − r′) (A4)

with all other commutators vanishing. Our operator equations of motion,

˙A(r, t) =

1

ih[A(r, t), H(t)] =

1

ε(r)ε0Π(r, t), (A5)

˙Π(r, t) =

1

ih[Π(r, t), H(t)] =

1

µ0∇2A(r, t), (A6)

lead to the second-order equation of motion for A(r, t) :

∇2A(r, t)− ε(r)

c2∂2A(r, t)

∂t2= 0. (A7)

For a quasi-two-dimensional array of waveguides, our relative permittivity takes the form ε(x, y, z) = εr[1+∆ε(x, y)].

We can solve this using separation of variables with A(r, t) ≡ ∑ν,ωAν,ω(r)Tν,ω(t) bν,ω + H.c., where we are

anticipating two indices ν and ω which will label solutions of the resulting equations for A(r) and T (t):

Tν,ω(t) + ω2 Tν,ω(t) = 0 (A8)

and

∇2Aν,ω(x, y, z) +ω2εrc2

Aν,ω(r) +ω2εr∆ε(x, y)

c2Aν,ω(r) = 0. (A9)

The first equation is easily solved with Tν,ω(t) = T (0)e−iωt, and restricting ω > 0. The second equation is thefamiliar Helmholtz equation, which the classical vector potential would also obey. We are interested in wavespropagating primarily in the positive z-direction, and thus consider solutions of the form

Aν,ω(x, y, z) = −iω−1 ψν,ω(x, y, z) eiωv z n. (A10)

Here v = c/√εr and the factor −iω−1 has been chosen so that ψν,ω here corresponds to ψν,ω in equation (2) in the

main text. This leads to an equation of motion which is first-order in z (assuming the envelope varies slowly),

i∂zψν,ω(x, y, z) = −1

2

v

)−1

∇2⊥ψν,ω(x, y, z)− 1

2

v

)∆ε(x, y)ψν,ω(x, y, z), (A11)

Page 8: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

8

where ∇2⊥ is the Laplacian operator acting in the x-y plane. This resembles a two-dimensional Schrodinger equation

with the z-direction taking the place of time, and with a potential ∆ε(x, y). Motivated by this analogy, we translatethe Schrodinger-like equation into a lattice model by writing ψ(x, y, z) (we will suppress the ν and ω indices fornow) as a sum over “Wannier states” of the lattice: ψ(x, y, z) =

∑r ψr(z)wr(x, y). Using the orthonormality of

Wannier states we find

i∂zψr(z) = ψr(z) 〈wr|H|wr〉+∑r′ 6=r

ψr′(z) 〈wr|H|wr′〉, (A12)

where we have switched to bra-ket notation for the Wannier states, and H = − 12 (ωv )−1∇2

⊥ − 12 (ωv )∆ε(x, y). Matrix

elements between different Wannier states decay exponentially with the spacing between lattice sites, and so weneglect all off-diagonal elements except Hr,r+sj

= 〈wr|H|wr+sj〉. We remove the diagonal elements, which adjust

the wavevector to satisfy the massive dispersion relation of the waveguide at the “site” r, by defining kω,r ≡ω/v + 〈wr|H|wr〉 and replacing ω/v → kω in equation (A10). With these adjustments, we arrive at the tight-binding equation for ψr(z),

i∂zψr(z) =∑

j=1,2,3

Hr,r±sjψr±sj

(z). (A13)

Here the sign choice is positive if r is on sublattice A and negative if on sublattice B. Restoring indices, and nowusing ν to index eigenvectors of this Hamiltonian, we have

ψν,ω(x, y, z) =∑r

e−iκν,ωz ψν,ω,r wr(x, y). (A14)

where ψν,ω,r is an eigenstate of the Hamiltonian Hω,r,r±sjwith eigenvalue κν,ω. Our full expressions for the vector

potential and electric field are

A(r, t) =∑ν,ω

(−iω−1)∑r

ei(kω,r−κν,ω)z−iωt ψν,ω,r wr(x, y) bν,ω n + H.c., (A15)

E(r, t) = − 1

ε(r)ε0Π =

∑ν,ω

∑r

ei(kω,r−κν,ω)z−iωt ψν,ω,r wr(x, y) bν,ω n + H.c. (A16)

With these solutions in hand, we can derive commutation relations of the bν,ω and b†ν,ω by enforcing our commutation

relations for A and Π. As expected, we find that:

[bν,ω, bν′,ω′ ] = 0 (A17)

[b†ν,ω, b†ν′,ω′ ] = 0 (A18)

[bν,ω, b†ν′,ω′ ] = δν,ν′δω,ω′ . (A19)

Lastly, it is convenient to decompose E(r, t) in the Wannier basis, and define an electric field operator at eachlattice site using:

E(r, t) =∑r

Er(z, t)wr(x, y), (A20)

Er(z, t) =∑ν,ω

ei(kω,r−κν,ω)z−iωt ψν,ω,r bν,ω n + H.c. (A21)

In the main text, we have specialized to the case of monochromatic input light, which selects a single frequency ωin the sum above.

For the sake of simplicity, we have neglected in the main text the r-dependence of the waveguide dispersion kω,r.This is rigorously justifiable when the variation in the relative permittivity ∆ε(x, y) has the discrete translationalsymmetry of the undistorted hexagonal lattice, or that of the translationally-invariant Kekule pattern shown inFig. 1(A). However, when translational symmetry is broken by the presence of a vortex, kω,r is not independent ofthe waveguide position r. This is problematic for systems with multiple vortices, as it can lead to the accumulationof dynamical phases as the vortices are braided (see Appendix D). We argue in Appendix D that the effects ofthese dynamical phases on the zero modes can be eliminated for an appropriate class of distortions of the waveguidelattice, presented in Appendix C.

Page 9: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

9

Appendix B: Zero-mode statistics and exact diagonalization results

1. Hamiltonian

We begin by reviewing the geometry of the hexagonal waveguide lattice. The vectors connecting a site onsublattice A with its 3 nearest-neighbors are

s1 =(0, −a

), s2 =

(√3

2a,

1

2a), and s3 =

(−√3

2a,

1

2a). (B1)

A unit cell consists of one site on sublattice A and a single nearest-neighbor on sublattice B. Our lattice vectorsare then

a1 = s2 − s1 =(√3

2a,

3

2a), a2 = s2 − s3 =

(√3 a, 0

). (B2)

We have chosen a to be the distance between nearest-neighbors, whereupon the distance between unit cells is√

3 a.In reciprocal space, the Dirac points lie at opposite corners of the hexagonal Brillouin zone, K± = ±

(4π

3√

3a, 0).

The Hamiltonian for the class of waveguide lattices that we study is given by

Hr,r+sj= −t− δtr,i, (B3)

Hr+sj ,r = Hr,r+sj(B4)

for r in sublattice A and all other matrix elements zero. In the presence of a Kekule texture, δtr,i is given in termsof the complex-valued order parameter ∆(r) by:

δtr,i =1

2∆(r) eiK+·si ei(K+−K−)·r + c.c. (B5)

A vortex centered at position R is described by ∆(r;R) = ∆0(|r − R|)ei[α−arg(r−R)], where ∆0(|r − R|) is areal-valued function which vanishes at |r −R| = 0 and approaches a constant value when |r −R| � `0, for somelength scale `0. For simplicity, we take ∆0(|r−R|) = ∆ tanh(|r−R|/`0), as in the main text. For multiple vortices,the order parameter is proportional to the product of order parameters for each individual vortex; for v identicalvortices at positions R1, . . . ,Rv, we have

∆(r) = ∆

[ v∏j=1

tanh(|r −Rj |/`0)

]ei

∑vj=1[αj−arg(r−Rj)]. (B6)

2. Vortex statistics

The Kekule texture with v vortices admits v zero modes, one localized at each vortex. We assume vortices arekept far enough apart to prevent mixing between zero modes (see Appendix D). Near a single vortex, say near R1,we can approximate r ≈ R1 in equation (B6) for all contributions from vortices Rj , j > 1. So the zero mode at thefirst vortex sees an order parameter

∆(r) ≈ ∆ tanh(|r −R1|/`0)

[ v∏j=2

tanh(|R1 −Rj |/`0)

]e−i arg(r−R1)eiα1+i

∑vj=2[αj−arg(R1−Rj)]

≈ ∆ tanh(|r −R1|/`0) e−i arg(r−R1)eiα1+i∑vj=2[αj−arg(R1−Rj)], (B7)

where we have used that, for `0 much smaller than the distance between vortices, we have tanh(|Ri −Rj |/`0) ≈ 1for all i, j. Thus, the only effect of the n− 1 other vortices on the first zero mode is to shift the phase of its orderparameter by a constant, i.e. taking α1 → α1 +

∑vj=2[αj − arg(R1 −Rj)].

Page 10: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

10

Now consider the operation of counterclockwise braiding two vortices R1 and R2. The braiding operation takesarg(R1 −R2) → arg(R2 −R1) + π and arg(R1 −R2) → arg(R2 −R1) + π, while leaving all other arg(Ri −Rj)unchanged. Using the local approximation of the order parameter near R1 in equation (B7), we see that braidingvortices R1 and R2 is equivalent to taking α1 → α1 − π in the single vortex wavefunction. Using a similarapproximation near the vortex at R2, we also have α2 → α2 − π. Relabeling the exchanged vortices 1 ↔ 2 andusing arg(R1 − R2) = arg(R2 − R1) + π, we finally have α1 → α2 and α2 → α1 − 2π. Because our zero modewavefunction (4) is double-valued with respect to α, u(r;α− 2π) = −u(r;α) and we arrive at the statistics in Eq.(7).

3. Exact diagonalization results

We here review exact diagonalization calculations of the tight-binding Hamiltonian (B3) on a finite hexagonallattice. Upon introduction of a constant ∆(r) = ∆, a band gap opens among the bulk states, although a smallnumber of edge states with zero and near-zero energy remain. Introduction of a vortex in the order parameter leadsto a zero mode localized at the vortex core, and an additional edge mode, which are decoupled from one anotherand at zero energy for sufficiently large lattices. The zero mode has support only on sublattice A, and decaysexponentially with distance from the vortex core.

Exploiting the mapping between the Helmholtz equation (A9) and the tight-binding equation (A13), we simulatethe effect of braiding in our waveguide array by promoting our Hamiltonian to a function of time, H(t) (recall that“time” in the tight-binding equation is actually the spatial z-direction in the waveguide array - we refer to it as t inthis section to make it clear that our simulations were performed in a two-dimensional system). We begin with aHamiltonian as in equation (B6) with two well-seperated vortices, and promote the centers of the vortices R1 andR2 to be functions of time:

R1(t) = RB(

cos(2πt/T ), sin(2πt/T )), (B8)

R2(t) = −RB(

cos(2πt/T ), sin(2πt/T )), (B9)

where RB is the radius of the circle traced out by the vortices during the braid and T is the total time of the process(chosen large enough for the process to be adiabatic). Taking our initial state to be the sum of the zero mode statescentered on each vortex, we then discretize the time interval into N steps of duration δt, and act with the matrixeiH(tj)δt on our state at each time tj = j δt, j = 1, . . . , N . Upon completion of a braid of radius RB = 9 a over atotal time T = 60/∆, for a 34× 35-unit-cell lattice with ∆ = .6 t, we find that the inner product of the initial andfinal states is approximately −0.95, consistent with the expected value of −1.

Appendix C: Waveguide worldlines for vortex braiding

In this section, we provide explicit formulas for the x-y positions of the waveguides in the presence of v vorticesin the Kekule order parameter. The positions of the centers of the waveguides in the hexagonal lattice are given by

rA = m1 a1 +m2 a2, rB = rA + s1, (C1)

where m1,2 are integers, the vectors a1,2 spanning the triangular lattice are given in Eq. (B2), and the vector s1

connecting sublattices A and B is defined in Eq. (B1).A Kekule distortion, constant or spatially varying, displaces the equilibrium positions of the waveguides to new

positions rA,B + uA,B(rA,B), with

uA(r) =i

2γ∆(r) e−ir·K+

(1

+i

)+ c.c. (C2)

uB(r) =i

2γ∆(r) e−ir·K+

(1−i

)+ c.c., (C3)

where γ is a dimensionful constant with units of [length]2. The pattern of waveguide displacements in the presenceof v vortices is obtained by substituting Eq. (B6) into the above.

Page 11: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

11

In this work, we have modified slightly the above pattern of distortions in the presence of v vortices, in order tomitigate the effect of the position-dependence of the waveguide dispersion kω,r on the braiding of the associatedzero modes. The modified pattern of distortions used to generate Fig. 1(B), which leads to a hopping modulationof the form (B5) with an order parameter of the form (B6), is

uA(r) =i

2γ∆(r) e−ir·K+

(1

+i

)+ c.c., uB(r) = 0. (C4)

We show in Appendix D that, for small displacements uA,B , this choice of lattice distortion leads to an onsitepotential that is uniform for all sites in sublattice A, so that the localized states in the vortex core, which havesupport only in sublattice A, experience an onsite potential that is uniform everywhere. (For a vortex with theopposite vorticity, whose associated zero mode has support on sublattice B rather than A, one would need to freezesublattice A instead of sublattice B.)

To write into some bulk medium a photonic lattice that implements an arbitrary braid B, one simply promotesthe displacements uA,B(r) → uA,B(r, z) by introducing a z-dependence of the vortex positions Ri ≡ Ri(z) inEq. (B6). The Ri(z) can be defined piecewise on intervals of length LB (see Appendix D) for each segment ofthe braid corresponding to the application of a single braiding generator. On each of these intervals, Ri(z) can bechosen to be an arbitrary parametric function of z that interpolates between the initial and final positions of the i-thvortex during that segment of the braid (albeit sufficiently slowly that the adiabatic condition holds). Examples ofsuch piecewise parametric functions are shown in Fig. 2 of the main text, where the strands being braided representthe vortex-core positions Ri(z) for i = 1, 2, 3.

Appendix D: Justifications for neglecting nonuniversal effects

In this section we discuss the requirements imposed by dynamical phases, adiabaticity, wavefunction mixing, andfinite size effects, and use these to estimate the number of braids we can perform in our system.

1. Dynamical phases

In our tight-binding Hamiltonian, the sublattice symmetry (SLS) ψr → ψr, ψr+sj → −ψr+sj , with r in sublatticeA, ensures that a zero mode will be at precisely zero energy. However, the presence of a position-dependentwaveguide dispersion kω,r (referred to below as an onsite potential) or next-nearest-neighbor hopping breaks SLSand can cause the zero mode energy to depend on both α and the position of the vortex center. This could cause thedynamical phase accumulation of the zero mode to depend on the path taken by the vortex, foiling measurementof the geometric phase due to braiding. We show that, and estimate the total dynamical phase accumulated in oursystem.

a. Onsite potential

We show here that the onsite potential kω,r can be made uniform in sublattice A by “freezing” (i.e., not displacing)the sites on sublattice B. Suppose that kω,r ≡ kω in the absence of displacements of the waveguides from thehexagonal lattice sites rA and rB . When the waveguides are displaced, one can show that, for r in sublattice A,

kω,r = kω exp

−3∑j=1

[∣∣∣1a

(sj − uA(r) + uB(r + sj)

)∣∣∣− 1

] . (D1)

For r in sublattice B, one simply takes A→ B and sj to −sj in the above. Let us now expand this expression toleading order for small displacements, i.e. |uA,B | � a:

kω,rkω≈ 1 +

1

a2

3∑j=1

sj · [uA(r)− uB(r + sj)] (D2)

Page 12: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

12

ap3 a

r0

FIG. 4: Schematic showing the length scales used in Sec. 3b. of this Appendix. Blue circles represent waveguides.

Noting that∑3j=1 sj = 0 [c.f. Eq. (B1)], we see that kω,r depends solely on uB in this limit. Therefore, setting

uB = 0, as in Eq. (C4), renders kω,r ≡ kω for r in sublattice A, to leading order in the displacements of thewaveguides. (Likewise, setting uA = 0 renders kω,r ≡ kω for r in sublattice B.) Keeping this in mind, we proceedto second order in the expansion of kω,r for uB(r) ≡ 0, finding

kω,rkω≈ 1− 1

2a2

3∑j=1

[|uA(r)|2 − 1

a2

(sj · uA(r)

)2]. (D3)

If |uA(r)| is constant as a function of r, then this leading correction is in fact independent of r, i.e. the distortionof the waveguide lattice leads to a position-independent renormalized dispersion kω,r ≈ kω + δkω. The leadingposition-dependent correction then appears at third order in |uA(r)|. A position-independent |uA(r)| can easily beachieved if the vortex width `0 is sufficiently small, and if the vortex centers Ri(z) never approach a lattice site towithin a distance `0 during braiding.

It is also worth pointing out that it is possible to choose the magnitudes |uA(r)| of the displacements in such away as to generate a position-independent kω,r, so long as we take uB ≡ 0. To see this, note that, if uB(r) = 0,Eq. (D1) is “ultra-local” in r, as its value at a point r is completely determined by uA(r). It is thus sufficient totune one parameter, namely |uA(r)|, in order to fix kω,r to a fixed value for each r.

b. Next-nearest-neighbor hopping

We now outline how the geometry of the waveguide lattice can be tuned such that next-nearest-neighbor hoppingt′ is strongly suppressed relative to nearest-neighbor hopping t. We can estimate t and t′ by recalling our Schrodingerequation (A11) for ψ. We expect our Wannier states to decay exponentially with decay length l0 = λ/(2π

√∆ε),

where λ is the wavelength of light in the material λ/(2π) = v/ω. For circular waveguides of radius r0 and nearest-neighbor spacing a (see Fig. 4), we estimate the hoppings to be

t ∼ 1

ae−(a−2r0)/l0 (D4)

and

t′ ∼ 1

ae−(√

3a−2r0)/l0 . (D5)

The waveguide radius r0 gives us an additional degree of freedom in choosing t and t′. We see that if

a− 2r0 ≈ l0 � (√

3− 1)a (D6)

holds, one can increase l0 to exponentially suppress the ratio t′/t while only algebraically suppressing t.

Page 13: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

13

c. Dynamical phase accumulation

We now estimate the total dynamical phase accumulated during propagation over a vertical distance L. Aswe can exponentially suppress dynamical phases due to next-nearest neighbor hopping, we assume the dominantcontribution to the dynamical phase is from a varying onsite potential. Consider a disorder potential that at somewaveguide i takes the form µi(z) = ηi(z) ∆, where ηi(z) is dimensionless. The energy shift of the zero mode (in thecontinuum limit) is

δε(z) =

∫d2x δµ(x, z) |ψ0(x)|2 ,

and the phase shift after propagation past a distance L is

φ =

∫ L

0

δε(z)dz

hc.

The root mean square of the phase shift is

δφ = η

√(`xyξ

)2 √L `z

hc

≈ η`xyξ

√L `zξ

≈ η

√`2xy `z

ξ3

√L

ξ.

where `xy ∼ a is the correlation length of the disorder in the plane, `z the correlation length in the paraxial direction,and ξ is the horizontal extent of the zero mode.

2. Adiabatic condition

Braiding vortices in the vertical direction z must be done slowly enough to satisfy the adiabatic condition. Weintroduce factors of h and define t ≡ z/c to work in units more familiar from quantum mechanics. The amplitudeto be in the nth mode, given that the state starts in the zero mode, is given by

cn(t) = −∫ t

0

dt′〈n |H| 0 〉E0 − En

ei(E0−En)t′/h c0(t′)

= −h 〈n |H| 0 〉(E0 − En)2

ei12 (E0−En)t 2 sin

(1

2(E0 − En)t

).

One can estimate the probability pout(t) = 1− |c0(t)|2 to escape the zero mode. Using 1− |c0(t)|2 =∑n 6=0 |cn(t)|2,

we have

pout ≈ (2h)2∑n 6=0

〈 0 |H|n 〉 〈n |H| 0 〉(En − E0)4

≤ (2h)2

∆4

∑n 6=0

〈 0 |H|n 〉 〈n |H| 0 〉

=(2h)2

∆4

(〈 0 |H2| 0 〉 − 〈 0 |H| 0 〉2

)=

(2h)2

∆4〈 0 |H2| 0 〉c .

Thus we arrive at

pout<∼

(2h)2

∆4〈 0 |H2| 0 〉c .

Page 14: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

14

In our case, say we move the vortex in space with velocity R, then 〈 0 |H2| 0 〉c = R2 〈 0 |H ′2| 0 〉c ≈ R2 (∆/ξ)2,where ξ is again the horizontal extent of the zero mode. Consider the time τξ it takes to move the vortex a distanceξ. In the same time, light propagates a distance Lξ (in the z-direction) along the guided mode. Notice that inthe units that we use above, ∆ is energy, not wavenumber. From our tight-binding Hamiltonian we can estimate∆ ∼ hc/ξ. We can then write

pout<∼

(2h)2

∆4

τξ

)2 (∆

ξ

)2

≈ (2hc)2

∆4

)2 (∆

ξ

)2

,

or, finally,

pout<∼

)2

.

Therefore, the condition for adiabaticity is that the distance Lξ along the paraxial direction needed to move thevortex a distance ξ in the xy-plane must be large compared to ξ, i.e., Lξ � ξ.

3. Number of braids

The total length L needed to do N braids is L ≈ πN Lξ d/ξ, where d is the distance between the vortices.(Basically, it takes πd/ξ movements of the vortex by a distance ξ, and adiabaticity requires that the distance in theparaxial direction be Lξ for each move.)

The upper limit on the number of braids is reached when δφ ∼ π, hence

π ≈ δφ ≈ η

√`2xy `z

ξ3

√L

ξ

≈ η

√`2xy `z

ξ3

√Lξξ

√d

ξ

√πN ,

leading to

N ≈ π η−2 ξ

d

ξ

ξ3

`2xy `z.

Now, the ratio nd = d/ξ ∼ 20 guarantees that the zero modes in different vortices do not mix. The ratio ξ/Lξ =√p0,

and we want to keep p0 ∼ 0.01, say. So we can write

N ≈ π η−2 ξ

d

ξ

ξ3

`2xy `z

≈ π

√po

η2 nd

ξ3

`2xy `z.

A few comments are in order. First, notice that, as expected, if there is no disorder (η → 0), the number of braidsis unbounded. Second, even if η is non-zero, there is a controlled limit where the number of braids can be made aslarge as desired: one can take the limit of large vortex sizes ξ.

Say one takes a pessimistic estimation using η ∼ 5%, and ξ ∼ 5a. Then, N ≈ π/.5× 53 ≈ 800. For η ∼ 1% andξ ∼ 10a, N ≈ π/.02× 103 ≈ 160, 000. And for η ∼ 1% and ξ ∼ 50a, N ≈ π/.02× 503 ≈ 20, 000, 000.

One can also estimate the vertical length needed per braid, L/N = πndξ/√p0. Taking a ∼ 10µm and all other

estimates as above, we have L/N ≈ 3 cm if ξ ∼ 5a and L/N ≈ 6 cm if ξ ∼ 10a.Finally, it is noteworthy that the correlation length in the paraxial direction is not necessarily of the order of a,

it can be much smaller. This is because it comes from fluctuations in writing the waveguides during their growth.So `z can be possibly hundredths of times smaller than a, in which case the number of braids estimated above goesup by a factor of 100.

Page 15: arXiv:1509.05408v3 [cond-mat.mes-hall] 20 Nov 2015Braiding light quanta Thomas Iadecola, Thomas Schuster, and Claudio Chamon Physics Department, Boston University, Boston, Massachusetts

15

4. Finite size effects

In a lattice with edges, each vortex will create both a mode localized at the vortex and a mode localized atthe edge of the system. If the lattice is small enough, there will be an “energy” splitting between symmetric andantisymmetric linear combinations of our zero mode and edge mode. Let us assume the splitting to be proportional

to t e−∆dEta . To prevent the zero mode from tunneling to the edge (a distance dE away) during the course of braiding,

we require

L� π

te

∆dEta . (D7)

Appendix E: Beam splitters and interference of coherent states

In this Appendix, we elaborate on the use of beam splitters to interfere the coherent states of light used in ourexperimental proposal. A beam splitter can be thought of as an optical circuit element that interferes photonsentering two input channels and outputs the result to two output channels. At the level of single photons, a 50/50dielectric beam splitter acts as follows:(

b†1b†2

)−→ 1√

2

(1 −i−i 1

)(b†1b†2

)=

1√2

(b†1 − i b†2−i b†1 + b†2

). (E1)

The factors of i above arise from π/2 phase shifts of the photons in each channel upon reflection within the beamsplitter. This action on single photons in turn induces an action on coherent states of photons:

|λ1, λ2〉 = e−(|λ1|2+|λ2|2)/2 eλ1 b†1 eλ2 b

†2 |0, 0〉

−→ e−(|λ1|2+|λ2|2)/2 eλ1(b†1−i b†2)/√

2 eλ2(−i b†1+b†2)/√

2 |0, 0〉 =

∣∣∣∣λ1 − i λ2√2

,−i λ1 + λ2√

2

⟩. (E2)

Note that the output coherent state is not a superposition of coherent states, but rather a new coherent state witha 50/50 admixture of photons from the two channels.

The interferometer proposed in the main text uses two sets of three beam splitters, one for each vortex core: thefirst splits each incoming laser beam into two equal parts, and the second interferes light from the two branches ofthe interferometer. Let us now follow one of the laser beams through each optical circuit element. We denote theinitial state of light from one of the three laser beams by |λ′j , 0〉. The effect of the first beam splitter is to split thisbeam in two:

|λ′j , 0〉 −→∣∣∣∣ λ′j√2

,−i λ′j√

2

⟩≡ |λj , λj〉. (E3)

From here, the light in the coherent state |λj〉 enters the waveguide lattice executing the braid B1B2, while the

light in the coherent state |λj〉 enters the waveguide lattice executing the braid B2B1. The state of light entering

the second beamsplitter is then | ± λj ,±λj〉, where the choice of sign on each coherent state depends on the braidthat was performed on that state. Passing this state through the beamsplitter gives

| ± λj ,±λj〉 −→∣∣∣∣∣±λj ∓ i λj√

2,∓i λj ± λj√

2

⟩=

∣∣∣∣±λj ∓ λj√2

,−i(±λj ± λj)√

2

⟩. (E4)

Therefore, light exiting one output channel of the beam splitter experiences destructive interference, while lightexiting the other channel experiences constructive interference. Whether this occurs in the first or second outputchannel depends on the relative sign between λj and λj after each beam exits its respective waveguide lattice.