anderson localization of composite excitations in ... fileanderson localization of composite...

13
Anderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque, 1 Vittorio Peano, 2 Oleg M. Yevtushenko, 2 and Florian Marquardt 2, 3 1 Instituto de F´ ısica Gleb Wataghin, Universidade Estadual de Campinas, 13083-859 Campinas, S˜ ao Paulo, Brazil 2 Institute for Theoretical Physics II, University Erlangen-N¨ urnberg, D-91058 Erlangen, Germany 3 Max Planck Institute for the Science of Light, G¨ unther-Scharowsky-Straße 1, D-91058 Erlangen, Germany (Dated: July 15, 2016) Optomechanical arrays are a promising future platform for studies of transport, many-body dy- namics, quantum control and topological effects in systems of coupled photon and phonon modes. We introduce disordered optomechanical arrays, focusing on features of Anderson localization of hybrid photon-phonon excitations. It turns out that these represent a unique disordered system, where basic parameters can be easily controlled by varying the frequency and the amplitude of an external laser field. We show that the two-species setting leads to a non-trivial frequency depen- dence of the localization length for intermediate laser intensities. This could serve as a convincing evidence of localization in a non-equilibrium dissipative situation. PACS numbers: 42.50.Wk, 71.55.Jv, 42.65.Sf Introduction: Optomechanics is a rapidly evolving re- search field at the intersection of condensed matter and quantum optics [1, 2]. By exploiting radiation forces, light can be coupled to the mechanical motion of vibra- tion modes. The interplay of light and motion is now being used for a range of applications, from sensitive mea- surements to quantum communication, while it also turns out to be of significant interest for fundamental studies of quantum physics. This rapidly developing area has so far mostly ex- ploited the interaction between a single optical mode and a single mechanical mode. Going beyond this, recent theoretical research indicates the substantial promise of so-called optomechanical arrays, where many modes are arranged in a periodic fashion. In such systems, a large variety of new phenomena and applications is predicted to become accessible in the future. These include the quantum many-body dynamics of photons and phonons [3], classical synchronization and nonlinear pattern for- mation [4–6], tunable long-range coupling of phonon modes [7–9], photon-phonon polariton bandstructures and transport [10, 11], artificial magnetic fields for pho- tons [12], and topological transport of sound and light [13]. A first experimental realization of a larger-scale op- tomechanical array has recently been presented, involv- ing seven coupled optical microdisks [14]. Even greater potential is expected for implementations based on op- tomechanical crystals [15–17], i.e photonic crystals that can be patterned specifically to generate localized photon and phonon modes. Given these promising predictions and the rapid exper- imental progress towards larger arrays, the question of disorder effects now becomes of urgent importance. For example, in the case of optomechanical crystals, experi- ments indicate fluctuations in the geometry of about 1%, which translate into equally large relative fluctuations of both the mechanical and optical resonance frequencies. This will invariably have a very significant impact on the transport properties. However, gaining a better un- derstanding of disorder effects in the various envisaged applications is only one motivation of the research to be presented here. Of equal, possibly even greater, impor- tance is the opportunity that is offered by optomechanics to create a highly tuneable novel platform for deliberately studying fundamental physical concepts such as Ander- son localization [18]. Localization of waves in a random potential is one of the most remarkable and nontrivial interference effects. Initially, it has been studied in electronic disordered sys- tems [19], though this effect applies equally to other types of quantum and even classical waves [20]. By now, local- ization and related phenomena have been discovered and investigated in photonic systems [21–27], coupled res- onator optical waveguides [28], cold atomic gases [29, 30], in propagation of acoustic waves [31] and in Josephson junction chains [32]. Localization can even play a con- structive role, namely in random lasing [27, 33]. In spite of extensive theoretical efforts, the unambiguous inter- pretation of experimental manifestations of localization often remains a challenge, even in situations where the ideal version of Anderson localization applies. Optomechanical arrays enable controlled optical exci- tation and readout and at the same time promise signif- icant flexibility in their design. However, it is the op- tical tuneability of the interaction between two differ- ent species (photons and phonons) that makes optome- chanical systems a unique platform. As we will show in the present Letter, this offers an opportunity to study effects in Anderson localization physics which currently represent a significant challenge even on the theoretical level and will thus open the door towards exploring novel physics that has not been observed so far. The model: We consider a 1D array of optomechanical cells (OMA), see Fig.1, driven by a single bright laser. The cell j contains an optical and a vibrational mode that are coupled via the standard linearized optomechanical arXiv:1607.04159v1 [cond-mat.mes-hall] 14 Jul 2016

Upload: nguyendien

Post on 09-Aug-2019

226 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

Anderson Localization of Composite Excitations in Disordered Optomechanical Arrays

Thales Figueiredo Roque,1 Vittorio Peano,2 Oleg M. Yevtushenko,2 and Florian Marquardt2, 3

1Instituto de Fısica Gleb Wataghin, Universidade Estadual de Campinas, 13083-859 Campinas, Sao Paulo, Brazil2Institute for Theoretical Physics II, University Erlangen-Nurnberg, D-91058 Erlangen, Germany

3Max Planck Institute for the Science of Light, Gunther-Scharowsky-Straße 1, D-91058 Erlangen, Germany(Dated: July 15, 2016)

Optomechanical arrays are a promising future platform for studies of transport, many-body dy-namics, quantum control and topological effects in systems of coupled photon and phonon modes.We introduce disordered optomechanical arrays, focusing on features of Anderson localization ofhybrid photon-phonon excitations. It turns out that these represent a unique disordered system,where basic parameters can be easily controlled by varying the frequency and the amplitude of anexternal laser field. We show that the two-species setting leads to a non-trivial frequency depen-dence of the localization length for intermediate laser intensities. This could serve as a convincingevidence of localization in a non-equilibrium dissipative situation.

PACS numbers: 42.50.Wk, 71.55.Jv, 42.65.Sf

Introduction: Optomechanics is a rapidly evolving re-search field at the intersection of condensed matter andquantum optics [1, 2]. By exploiting radiation forces,light can be coupled to the mechanical motion of vibra-tion modes. The interplay of light and motion is nowbeing used for a range of applications, from sensitive mea-surements to quantum communication, while it also turnsout to be of significant interest for fundamental studiesof quantum physics.

This rapidly developing area has so far mostly ex-ploited the interaction between a single optical mode anda single mechanical mode. Going beyond this, recenttheoretical research indicates the substantial promise ofso-called optomechanical arrays, where many modes arearranged in a periodic fashion. In such systems, a largevariety of new phenomena and applications is predictedto become accessible in the future. These include thequantum many-body dynamics of photons and phonons[3], classical synchronization and nonlinear pattern for-mation [4–6], tunable long-range coupling of phononmodes [7–9], photon-phonon polariton bandstructuresand transport [10, 11], artificial magnetic fields for pho-tons [12], and topological transport of sound and light[13]. A first experimental realization of a larger-scale op-tomechanical array has recently been presented, involv-ing seven coupled optical microdisks [14]. Even greaterpotential is expected for implementations based on op-tomechanical crystals [15–17], i.e photonic crystals thatcan be patterned specifically to generate localized photonand phonon modes.

Given these promising predictions and the rapid exper-imental progress towards larger arrays, the question ofdisorder effects now becomes of urgent importance. Forexample, in the case of optomechanical crystals, experi-ments indicate fluctuations in the geometry of about 1%,which translate into equally large relative fluctuations ofboth the mechanical and optical resonance frequencies.This will invariably have a very significant impact on

the transport properties. However, gaining a better un-derstanding of disorder effects in the various envisagedapplications is only one motivation of the research to bepresented here. Of equal, possibly even greater, impor-tance is the opportunity that is offered by optomechanicsto create a highly tuneable novel platform for deliberatelystudying fundamental physical concepts such as Ander-son localization [18].

Localization of waves in a random potential is one ofthe most remarkable and nontrivial interference effects.Initially, it has been studied in electronic disordered sys-tems [19], though this effect applies equally to other typesof quantum and even classical waves [20]. By now, local-ization and related phenomena have been discovered andinvestigated in photonic systems [21–27], coupled res-onator optical waveguides [28], cold atomic gases [29, 30],in propagation of acoustic waves [31] and in Josephsonjunction chains [32]. Localization can even play a con-structive role, namely in random lasing [27, 33]. In spiteof extensive theoretical efforts, the unambiguous inter-pretation of experimental manifestations of localizationoften remains a challenge, even in situations where theideal version of Anderson localization applies.

Optomechanical arrays enable controlled optical exci-tation and readout and at the same time promise signif-icant flexibility in their design. However, it is the op-tical tuneability of the interaction between two differ-ent species (photons and phonons) that makes optome-chanical systems a unique platform. As we will show inthe present Letter, this offers an opportunity to studyeffects in Anderson localization physics which currentlyrepresent a significant challenge even on the theoreticallevel and will thus open the door towards exploring novelphysics that has not been observed so far.

The model: We consider a 1D array of optomechanicalcells (OMA), see Fig.1, driven by a single bright laser.The cell j contains an optical and a vibrational mode thatare coupled via the standard linearized optomechanical

arX

iv:1

607.

0415

9v1

[co

nd-m

at.m

es-h

all]

14

Jul 2

016

Page 2: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

2

0.8

1

1.2

Quasimomentum 0-

1

1.2

Ener

gy/

m

c

b

a

d

JoJm

g

FIG. 1. (color on-line) Scheme, implementations and bandstructure of an optomechanical array. (a) An array of pho-ton (blue dots) and phonon (yellow) modes can be viewedas a ladder: photons and phonons either can hop to nearestneighbor sites or can be interconverted on the same site. Thissystem can be implemented by (b) an array of microdisks,or (c) an array of co-localized optical and mechanical de-fect modes in an optomechanical crystal. (d) Optomechan-ical band structure for two different values of the couplingg/Ωm = 0.01, 0.1 (red/black lines). For the larger interactionstrength the upper and lower polariton bands are separatedby a complete band gap (grey region). The other parametersare Ωo = 1.1Ωm, Jo = 0.1Ωm, Jm = 0.01Ωm.

Hamiltonian

Hj =∑ν=o,m

ων,j nν,j − gj(co,j + c†o,j

)(cm,j + c†m,j

); (1)

see Refs.[1, 34] for details. Here nν,j ≡ c†ν,j cν,j , andcν,j is the bosonic annihilation operator of either optical,ν = “o”, or mechanical, ν = “m”, excitations (we set~ = 1). Due to disorder, the frequencies ων,j fluctuatearound mean values 〈ων,j〉d = Ων . We assume that ων,jare independent Gaussian random variables with vari-ances 〈(ων,j − Ων)(ων′,j′ − Ων′)〉d = σ2

νδj,j′δν,ν′ . Eq. (1)is defined in a rotating frame, where the optical frequen-cies ωo,j are counted off from the laser frequency, ωL[1]. Thus, Ωo indicates the average detuning and can betuned in situ by varying the laser frequency. The optome-chanical couplings gj are proportional to the mean am-plitude of the light circulating in the cavity j [1]. Hence,they are also tunable by varying the laser power.

The presence of two-mode squeezing interactions in Eq.(1) can in principle lead to instabilities. We choose Ωosuch that these terms are off-resonant and disorder con-figurations with optical [35] or vibrational instabilitiesare very rare. We leave their study for a forthcomingpaper.

We can describe the full OMA by a Hamiltonian with

nearest-neighbor optical, Jo, and mechanical, Jm, hop-ping amplitudes:

H =∑j

Hj − Hh, Hh =∑j,ν

Jν c†ν,j+1cν,j +H.c. (2)

Our model is time-reversal symmetric [36].Clean polariton bands: In a clean OMA without dissi-

pation (and without squeezing interaction), the photon–phonon hybridization leads to a pair of bands with ener-gies

Ω± = Ω− 2J cos(k)±√

[δΩ/2− δJ cos(k)]2

+ g2, (3)

where Ω = (Ωo + Ωm)/2, δΩ = (Ωo − Ωm) and like-wise for J , δJ . We refer to Ω± as upper/lower polaritonband, respectively. k denotes the wave-vector of polari-tonic Bloch states. We focus on the regime where theuncoupled bands overlap, δΩ < 4J . In this case, thepolariton bands are separated by a gap if the couplingbecomes large enough, g > gmin [37], see Fig.1.Anderson localization of uncoupled excitations: It is

well known that even weak disorder leads to a crucial ef-fect in a 1D system: the eigenstates become localized. Ifg = 0, each subsystem (photon/phonon), is individuallydescribed by the 1D Anderson model [18]. The local-ized states decay exponentially away from their center,

∼ exp(−|j−j0|/ξ(0)ν ). Here ξ

(0)o,m are the bare localization

lengths for photons and phonons (for g = 0), measuredin units of the lattice constant. Using the theory of 1Dlocalization [38], we can approximate the frequency de-pendence of the localization length:

ξ(0)ν (Ω) ' 2(2 sin[kν(Ω)]/χν)2; (4)

here the dimensionless quantities χν ≡ σν/Jν and2 sin[kν ] are the disorder strength and the bare group ve-locity, respectively. Eq.(4) is valid for weak (up to mod-erately strong) disorder [39]. The comparison of Eq.(4)with numerical results is shown below in Fig.3.

In any experiment, localization can be detected if pho-tons and phonons explore the localization length be-fore leaking out, at a rate κo,m. This holds true ifξν < 2Jν | sin(kν)|/κν , which allows us to neglect dissi-pation in a first approximation [40]. In addition, thesample size L should be larger than the localization

length, L max(ξ(0)ν ). For typical L ∼ 100 we need

max(ξ(0)ν ) ∼ 10, corresponding to χν ∼ 1.

Localization in Optomechanical Arrays: At finite pho-ton-phonon coupling, we encounter an Anderson modelwith two channels. Localization in the symmetric versionof this model (with equal parameters of each channel) iswell studied and understood [41, 42]. However, OMAsdo not fall into this universality class since the mechani-cal band is generically much narrower than the optical

Page 3: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

3

-50 0 50position

5

4

3

2

1

0

log

10

⟨ n⟩

-50 0 50

6

4

2

0

FIG. 2. (color on-line) Upper panel: Typical shape of aneigenstate, at Ωo = Ωm, g = 0.001Ωm, without disorder av-eraging (inset) and after disorder averaging over 500 realiza-tions (main picture). Other parameters are explained in thetext. Two different localization lengths are clearly visible.The profiles of nj , obtained from different disorder realiza-tions, have been shifted in space, such that their maxima arealways located at j = 0. Lower panel: Local density of states(frequency- and position-resolved spectrum) at g = 0.05Ωm;other parameters as in the upper panel. The color rangesfrom orange to blue, depending on whether a given eigenstatehas a stronger mechanical or optical component, respectively.

one, Jm Jo. Thus, the hybrid excitations consistof two components with very different velocities. Sim-ilar composite quasiparticles are not uncommon, anotherexample is given by cavity polaritons [43, 44] includingpolaritons in a disordered potential [45]. Developing thetheory of localization for such non-symmetric systems re-mains a real challenge, cf. Ref.[46]. The hybrid localizedstates typically have two localization lengths, ξ1 < ξ2,see the upper panel of Fig.2. For small systems, L < ξ1,the excitations do not feel localization and propagateballistically. Their transmission decays as exp(−L/ξ1)in the range ξ1 < L < ξ2 and becomes suppressed asexp(−L/ξ2) at L > ξ2. Our numerical analysis showsthat the space region where ξ1 dominates quickly shrinkswith increasing g. Therefore, ξ2 seems more interestingexperimentally, and we will focus on this ’large’ localiza-tion length in the following. We start from a numericalanalysis for relatively strong disorder. At the first stage,we neglect disorder-induced fluctuations of gj [47] anduse its homogeneous mean value g = const.

The method: The localization length can be ob-tained, e.g., from the photon-photon transmission,

Too(j, k; Ω) ∝∣∣GRoo(j, k; Ω)

∣∣2 where GRoo(j, k; Ω) =

−i∫∞

0dt exp(iΩt)[co,j(t), c

†o,k(0)] is the frequency-resol-

ved retarded Green’s function. Too is defined via the op-tical power detected on site j at frequency ωL + Ω while

a probe laser of the same frequency is impinging on adifferent site k [48]. For x = |j − k| → ∞, we expectToo(j, k; Ω) ∝ exp(−2x/ξ2). Thus, the expression for theaveraged (inverse) localization length reads

ξ−12 (Ω) = − lim

x→∞

(⟨ln(Too(j, k; Ω)

)⟩d

/2x). (5)

We note that the value of ξ2(Ω) is the same for othertransmission processes (e.g. photon-phonon transmis-sion) [48].

Eq.(5) can be used as a definition even in the presenceof dissipation. In the absence of dissipation and insta-bilities, there is a simpler alternative, namely extractingthe localization length directly from the spatial profile ofeigenstates [48]. To ensure reliability of results, we havecombined both approaches in numerical simulations.Analysis of numerical results: The upper panel of Fig.2

shows a typical optomechanical eigenstate in the case ofsmall coupling. The excitation frequency has been se-lected from the tail of the pure mechanical band. Twodifferent slopes, which correspond to two different local-ization lengths ξ1,2, are clearly visible. When g increasesand the other parameters of the upper Fig.2 remain un-changed, the region where ξ1 dominates shrinks [49] andbecomes invisible very quickly. In the following, we willconcentrate on ξ2 and will denote it as ξ for the sakeof brevity. The lower panel of Fig.2 illustrates the dis-tribution of optomechanical excitations in space and fre-quency, including the character of excitations (photon vs.phonon).

Here, and in the following, we have displayed numer-ical results for an illustrative set of parameters: Jo =0.1Ωm Jm = 10−3Ωm. Localization of the optome-chanical excitations becomes pronounced at χo,m ∼ 1.For concreteness, we have chosen equal relative disorderstrength, χo = χm = 1. In real samples, Jo ranges from1GHz to 10THz (Jm: from 100kHz to 1GHz) with theoptical disorder being of order 100GHz to 1THz (me-chanical: from 10MHz to 100MHz). Thus, our choice ofχo,m falls into the range of experimentally relevant pa-rameters. The optomechanical coupling in our numericsranges from weak, g = 10−3Ωm, to strong, g = 0.05Ωm.To suppress finite size effects, we employed large systems,L = 103 ξ, during exact diagonalization. The Green’sfunctions method has allowed us to explore even muchlarger sizes.

In Fig.3, we display the frequency-dependence of thelocalization length of hybrid optomechanical excitationsin a disordered array, one of the central numerical resultsof this article. For comparison, we also show the situ-ation for the uncoupled systems, including the (scaled)

analytical expression for ξ(0)o , Eq.(4) [50] (green solid line

in Fig.3a). Once the subsystems are coupled, significantchanges of ξ(Ω) occur in the vicinity of the unperturbed(narrow) mechanical band where the optomechanical hy-bridization is most efficient. Firstly we note that, if

Page 4: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

4

0 < g < ∆(m)loc , the coupling between the optical and

the mechanical systems is perturbatively weak even inthe middle of the mechanical band [region I in Fig.3(f)].On the other hand, when the optomechanical coupling

becomes large, g >√JoJm =

√σoσm ∼ ∆

(o)loc for our

choice of parameters, a gap opens around the resonantfrequency Ω = Ωm and remaining excitations inside thegap tend to become localized [Fig.3(e)].

0.9 1.0 1.1 1.2 1.3frequency Ω/Ωm

0

4

8

12

loc.

length

ξ

a g= 0. 001Ωm

0.98 1.0 1.02

b g= 0. 004Ωm

0.98 1.0 1.02

c g= 0. 008Ωm

0.98 1.0 1.02

d g= 0. 02Ωm

0.9 1.0 1.1 1.2 1.3frequency Ω/Ωm

e g= 0. 05Ωm

0 0.04 0.08coupling g/Ωm

0.8

0.9

1.0

freq. Ω/Ωm

fgap

∆(o)loc

JmI II III

Weak coupling

Strong coupling

FIG. 3. (color on-line) Frequency-dependence of the localization length: Dashed lines show bare (g = 0) optical (ξ(0)o - blue)

and mechanical (ξ(0)m - black) localization lengths. The red solid line shows the localization length of hybrid excitations, ξ,

calculated at several values of the optomechanical coupling, 0.001Ωm ≤ g ≤ 0.05Ωm, and Ωo = 1.1Ωm. Green solid lines

describing ξ(0)o in panel (a) and ξ in panel (e) are obtained from Eqs.(4,6), respectively, after scaling by a constant factor. Panel

(f) illustrates schematically the different regimes as a function of coupling and frequency.

Analytical methods which would allow one to explorelocalization in strongly disordered systems are not avail-able in general. Nevertheless, it turns out that our op-tomechanical array corresponds to a certain two-channelsystem, which was studied analytically in Ref.[46] for thelimit of weak disorder and large coupling. Remarkably,the shape of our numerically extracted ξ(Ω) at large gagrees with the predictions of Ref.[46], even though weare here dealing with strong disorder, χν ∼ 1 [42]. Thetheory of Ref.[46] is valid if g is large compared withthe (bare) mean level spacing in the localization volume,

∆(ν)loc , which holds true for the parameters of our numer-

ical study at g ≥ 0.05Ωm [51]. If g > gmin, (i.e., if theclean polariton bands are separated by the gap of thewidth Ω+(k = 0)−Ω−(k = π)) we can use the following(leading in χν) expression for the localization length [46]:

ξ(Ω) ' 4(2 sin [k±(Ω)]

)2/(χ2[1 + cos2(γ)

]); (6)

tan(γ) = 2√JoJmg/δJ(Ω− Ωr), Ωr ≡ JoδΩ/δJ.

Here χ = χo/C = χm/C and k±(Ω) denotes the in-verted dispersion relation Ω±(k). The quantity V± ≡2 sin [k±(Ω)] is called “rapidity”. It coincides with thegroup velocity of the excitations for g = 0, and accordingto Eq.(6) it governs the frequency-dependence of ξ(Ω) inthe coupled case. The factor C reflects renormalization ofthe disorder strength caused by the optomechanical cou-pling. Calculation of C is beyond the scope of Ref.[46] andwe have found its approximate value C ' 1.16 by fittingthe analytically calculated maximal value of ξ(Ω > Ωm)to the numerical one. Fig.3e shows the comparison of theanalytical and numerical results. They differ noticeablyonly close to edges of the clean band where the analytical

theory looses its validity because ξ → 1. In addition, thegap is smeared by the relatively strong disorder.

We have discovered that, at Ω ' Ωm, the crossoverbetween small and large values of g is highly non-trivial(and it is outside the scope of the analytical theory):when the optomechanical coupling increases from g ∼∆

(m)loc to g ∼ Jm [region II in Fig.3(f)], the single max-

imum of ξ [cf. Fig.3(a)→(b)] grows sublinearly in g[52]. This growth stops and turns into a decrease wheng Jm. Simultaneously, a new local maximum developsat the frequency corresponding to the maximum of therapidity [Fig.3(b)→(c) and region III in Fig.3(f)]. Fi-nally, the new local maximum becomes the global one

and a dip appears close to Ωm at g ≥ ∆(o)loc [Fig.3(c)→(d)].

This non-trivial dependence of the localization length onthe coupling constant, i.e., on the tuneable intensity ofthe external laser, could help to distinguish localizationand trivial dissipation effects in real experiments.

We have checked that the shape of ξ(Ω) is robust withrespect to dissipation effects as long as the mean levelspacing in the localization volume of the hybrid excita-tions is larger than the optical and mechanical decayrates, κν [53]. Propagation of the excitations is sup-pressed due to their finite life time which is reflected bythe frequency-independent decrease of ξ. Typical profilesξ(Ω) are shown in Fig.4 where the optical dissipation rateincreases until κo = 0.1Ωm. These profiles are also robustwith respect to the spatial inhomogeneity of gj which re-sults from randomness of the cell frequencies [47, 54].

Conclusions and discussion: Disordered OMAs belongto a new class of disordered systems where composite(photon-phonon) excitations are localized and the most

Page 5: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

5

0.9 1.0 1.1 1.2 1.3frequency Ω/Ωm

0

2

4

6

8

10lo

c. length

ξa g= 0. 001Ωm

0.9 1.0 1.1 1.2 1.3frequency Ω/Ωm

b g= 0. 05Ωm

FIG. 4. (color on-line) Frequency-dependence of the local-ization length at Ωo = 1.1Ωm and g/Ωm = 0.001, 0.05 cal-culated for different values of the optical decay rate κo/Ωm =0, 0.01, 0.05, 0.1 (red-, blue-, green-, and brown lines, re-spectively). Note that the peak at g = 0.001Ωm and Ω = Ωmis almost insensitive to the optical dissipation since the corre-sponding wavefunctions have mainly mechanical components.

important parameters can be easily fine-tuned. Thus,OMAs provide a unique opportunity to study Andersonlocalization of composite particles in real experiments.Moreover, they should allow to reliably distinguish lo-calization from trivial dissipation effects. Future stud-ies may address the additional novel physics that willarise when two-mode squeezing processes become rele-vant. At strong driving, this could involve the inter-play between instabilities and localization, with interest-ing connections to random lasing, extending the new re-search domain of disordered optomechanical arrays intothe nonlinear regime.

We acknowledge support from the EU Research Coun-cil through the grant EU-ERC OPTOMECH 278320.TFR acknowledges support from FAPESP. We are grate-ful to Vladimir Kravtsov and Igor Yurkevich for usefuldiscussions.

[1] M. Aspelmeyer, T. Kippenberg, and F. Marquardt, Rev.Mod. Phys. 86, 1391 (2014).

[2] M. Aspelmeyer, S. Groeblacher, K. Hammerer, andN. Kiesel, Journal of the Optical Society of America B27, A189 (2010).

[3] M. Ludwig and F. Marquardt, Phys. Rev. Lett. 111,073603 (2013).

[4] G. Heinrich, M. Ludwig, J. Qian, B. Kubala, and F. Mar-quardt, Phys. Rev. Lett. 107, 043603 (2011).

[5] C. A. Holmes, C. P. Meaney, and G. J. Milburn, Phys.Rev. E 85, 066203 (2012).

[6] R. Lauter, C. Brendel, S. J. M. Habraken, and F. Mar-quardt, Phys. Rev. E 92, 012902 (2015).

[7] A. Xuereb, C. Genes, and A. Dantan, Phys. Rev. Lett.109, 223601 (2012).

[8] A. Xuereb, C. Genes, G. Pupillo, M. Paternostro, andA. Dantan, Phys. Rev. Lett. 112, 133604 (2014).

[9] M. Schmidt, M. Ludwig, and F. Marquardt, New J.Phys. 14, 125005 (2012).

[10] W. Chen and A. A. Clerk, Phys. Rev. A 89, 033854(2014).

[11] M. Schmidt, V. Peano, and F. Marquardt, New Journalof Physics 17, 023025 (2015).

[12] M. Schmidt, S. Kessler, V. Peano, O. Painter, andF. Marquardt, Optica 2, 635 (2015).

[13] V. Peano, C. Brendel, M. Schmidt, and F. Marquardt,Phys. Rev. X 5, 031011 (2015).

[14] M. Zhang, S. Shah, J. Cardenas, and M. Lipson, Phys.Rev. Lett. 115, 163902 (2015).

[15] A. H. Safavi-Naeini, T. P. M. Alegre, M. Winger, andO. Painter, Appl. Phys. Lett. 97, 181106 (2010).

[16] J. Chan, A. H. Safavi-Naeini, J. T. Hill, S. Meenehan,and O. Painter, Applied Physics Letters 101, 081115(2012).

[17] A. H. Safavi-Naeini, J. T. Hill, S. Meenehan, J. Chan,S. Groblacher, and O. Painter, Phys. Rev. Lett. 112,153603 (2014).

[18] P. W. Anderson, Phys. Rev. 109, 1492 (1958).[19] B. Kramer and A. MacKinnon, Reports on Progress in

Physics 56, 1469 (1993).[20] M. C. W. van Rossum and T. M. Nieuwenhuizen, Rev.

Mod. Phys. 71, 313 (1999).[21] A. A. Chabanov, M. Stoytchev, and A. Z. Genack, Na-

ture 404, 850 (2000).[22] T. Schwartz, G. Bartal, S. Fishman, and M. Segev, Na-

ture 446, 52 (2007).[23] S. Mookherjea, J. S. Park, S.-H. Yang, and P. R. Ban-

daru, Nature Photonics 2, 90 (2008).[24] Y. Lahini, A. Avidan, F. Pozzi, M. Sorel, R. Morandotti,

D. N. Christodoulides, and Y. Silberberg, Phys. Rev.Lett. 100, 013906 (2008).

[25] V. Savona, Phys. Rev. B 83, 085301 (2011).[26] M. Segev, Y. Silberberg, and D. N. Christodoulides, Na-

ture Photonics 7, 197 (2013).[27] J. Liu, P. D. Garcia, S. Ek, N. Gregersen, T. Suhr,

M. Schubert, J. Mørk, S. Stobbe, and P. Lodahl, Na-ture Nanotechnology 9, 285 (2014).

[28] M. Hafezi, E. A. Demler, M. D. Lukin, and J. M. Taylor,Nature Physics 7, 907912 (2011).

[29] J. Billy, V. Josse, Z. Zuo, A. Bernard, B. Hambrecht,P. Lugan, D. Clement, L. Sanchez-Palencia, P. Bouyer,and A. Aspect, Nature 453, 891 (2008).

[30] G. Roati, C. D’Errico, L. Fallani, M. Fattori, C. Fort,M. Zaccanti, G. Modugno, M. Modugno, and M. Ingus-cio, Nature 453, 895 (2008).

[31] H. F. Hu, A. Strybulevych, J. H. Page, S. E. Skipetrov,and B. A. van Tiggelen, Nature Physics 4, 945 (2008).

[32] D. M. Basko and F. W. J. Hekking, Phys. Rev. B 88,094507 (2013).

[33] D. S. Wiersma, Nature 4, 359 (2008).[34] Derivation of the standard optomechanical Hamiltonian

is briefly reviewed in Suppl.Mat.1.[35] L. Mandel and E. Wolf, Optical coherence and quantum

optics (Cambridge Univ. Press, Cambridge, 2008).[36] This means that Jo,m, g0 and g are real.[37] The value of gmin is found from the condition Ω+(k =

0) = Ω−(k = π)⇒ gmin = 2Re√JoJm

(1− [δΩ/4J ]2

).

[38] V. I. Melnikov, Sov. Phys. Solid State 23, 444 (1981).[39] Strictly speaking, one must require χν < 1 though a

softer condition ξν 1 sufficies for practical purposes.[40] We have taken into account that, in 1D systems, the exci-

tations propagate ballistically in the localization volume.[41] O. N. Dorokhov, Solid State Communications 44, 915

(1982).

Page 6: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

6

[42] P. A. Mello and N. Kumar, Quantum transport in meso-scopic systems (Oxford Univ. Press, Oxford, 2004).

[43] J. Kasprzak, M. Richard, S. Kundermann, A. Baas,P. Jeambrun, J. M. J. Keeling, F. M. Marchetti, M. H.Szymanska, R. Andre, J. L. Staehli, V. Savona, P. B. Lit-tlewood, B. Deveaud, and L. S. Dang, Nature 443, 409(2006).

[44] E. Wertz, L. Ferrier, D. D. Solnyshkov, R. Johne, D. San-vitto, A. Lemaitre, I. Sagnes, R. Grousson, A. V. Ka-vokin, P. Senellart, G. Malpuech, and J. Bloch, NaturePhysics 6, 860 (2010).

[45] F. Manni, K. G. Lagoudakis, B. Pietka, L. Fontanesi,M. Wouters, V. Savona, R. Andre, and B. Deveaud-Pledran, Phys. Rev. Lett. 106, 176401 (2011).

[46] H.-Y. Xie, V. E. Kravtsov, and M. Muller, Phys. Rev.B 86, 014205 (2012).

[47] gj is proportional to the mean occupation number of thephotons on the site j while the cells with smaller opticalfrequencies host more photons. This locally enchances gjon these sites.

[48] Algebraic details for the derivation of the localization ra-dius can be found in Suppl.Mat.2.

[49] If g increases and the other parameters of the upper Fig.2remain unchanged, ξ2 slightly increases. This might lead

to suppression of the region where ξ1 dominates becausethe eigenstate is normalized.

[50] The accuracy of Eq.(4) is insufficient to reproduce the

maximal value of ξ(0)o (Ωo) at the relatively strong optical

disorder, χo = 1. Therefore, we have scaled the analyticalanswer by a constant factor to adjust the heights.

[51] The mean level spacing in the localization volume isobtained from the relation ∆loc/∆ = N/ξ . Using theestimate for the mean level spacing in a band of ex-tremely localized states ∆ ∼ δ/N [55, 56], we find

∆(ν)loc ∼ δν/ξ

(0)ν ∼ Jν/ξ

(0)ν . Here ξ

(0)ν denotes bare (at

g = 0) localization lengths. For the chosen parameters,this yields the condition g > 10−1maxJν ∼ 10−2Ωm.Thus, the analytical theory from Ref.[46] can be used tounderstand the case g = 0.05Ωm while smaller values of gare beyond the validity range of the analytic expressions.

[52] The increase of ξ2(Ω ' Ωm) when the optomechanicalcoupling is small, g < Jm, is explained in Suppl.Mat.3.

[53] Equivalently, we can require that the localization lengthis much smaller than the escape length.

[54] This robustness is exemplified in Suppl.Mat.4.[55] O. Yevtushenko and V. E. Kravtsov, J. Phys. A 36, 8265

(2003).[56] O. Yevtushenko and V. E. Kravtsov, Phys. Rev. E 69,

026104 (2004).

Page 7: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

7

SUPPLEMENTAL MATERIALS

1. Standard optomechanical Hamiltonian

The linearized Hamiltonian, Eq.(2) of the main text, is derived starting from the Hamiltonians of a phononic anda photonic array. In the tight-binding approximation, both Hamiltonians take the same form,

Hν =∑j

ων,j c

†ν,j cν,j − Jν

(c†ν,j+1cν,j +H.c.

). (7)

Here, the index ν = “o” and ν = “m” refers to the optical and mechanical degrees of freedom. The operators cν,jdenote the annihilation operator for the site j; ων,j and Jν denote random on-site frequencies and constant overlapintegrals, respectively. The optical and mechanical modes co-localized on the same site are coupled by the radiationpressure force. The resulting interaction reads

Hom = −g0

∑j

c†o,j co,j(c†m,j + cm,j) , (8)

where g0 is the eigenfrequency shift of a localized optical mode by a single phonon on the same site. The presence ofa laser drive of frequency ωL is described by the additional Hamiltonian term

Hlaser = αL∑j

co,j exp(iωLt) +H.c. (9)

The dynamics of the OMA with the Hamiltonian H = Ho + Hm + Hom + Hlaser is most conveniently described in therotating frame defined by the unitary transformation

H → UHU† − i U d

dtU† , U =

∑j

exp(iωLt c

†0,j c0,j

). (10)

We decompose the operators cν,j as sums of their mean filed values and new displaced operators, δcν,j , incorporating

the fluctuations, cν,j = 〈cν,j〉 + δcν,j . After inserting this decomposition into H, all linear terms in δcν,j cancel outin the Hamiltonian. We, thus, reproduce Eq.(7) but with the fluctuation operators δcν,j and the detunings ωo,j − ωLreplacing the bare operators cν,j and the optical frequencies ωo,j , respectively. The eigenfrequencies ωo,j of the opticallocalized modes include a small power-dependent frequency shift due to a static displacement ∝ 2Re[〈cm,j〉] of thecorresponding mechanical oscillators, ωo,j = ωo,j − 2g0Re[〈cm,j〉]. In the limit where the fluctuations δcν,j are smallcompared to the mean values (for a strong enough drive), we can neglect all cubic terms and arrive to the linearizedopto-mechanical interaction [1]

H(L)om = −

∑j

(g∗j δco,j +H.c.

)(δcm,j +H.c.) . (11)

Here gj ≡ g0 〈co,j〉 are the couplings of the linearized interaction. In the main text, we have investigated a parameterregime where the laser is red-detuned compared to all optical resonances. We have also focused on the OMAs wherethe broadening of the resonances (set by the typical decay rate κo) is smaller than the minimal detuning. In thiscase, all gj are real valued, consequently, the time-reversal symmetry is preserved by the OM interaction. Summing

all contributions, we obtain the Hamiltonian H of the main text. There, for brevity, we use cν,j and ωo,j for thefluctuation operators and the detunings, respectively. We follow this convention also below. In the main text, wehave also assumed that all linearized couplings gj are approximately equal, gj ≈ g. This approximation holds whenthe mean value of the onsite detuning is much larger than its typical fluctuations, the optical hopping rate, and thetypical optical decay rate. Below, we go beyond this approach investigating fluctuating coupling constants gj .

2. Calculation of the localization length

2.1 Input/Output formalism

Let us express the elastic part of the photon-photon transmission Too(k, j; Ω) in terms of the retarded (photon-photon) Green’s function.

Page 8: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

8

The response of the OMA to an additional probe field is described by the standard Langevin equations [S1]:

˙cν,j = i[H, cν,j ]− κν,j cν,j/2 +√κν,j c

inν,j ; (12)

where cinν,j is the input (probe) field. The corresponding output field coutν,j is given by the input/output relations

coutν,j = cinν,j −

√κν,j cν,j . (13)

For a probe laser of frequency ωL + Ω (corresponding to the frequency Ω in the rotating frame) applied at site k wehave

cino,l = δl,kαpe−iΩt, cinm,l = 0, (14)

where αp is the amplitude of the probe laser.The transmission is defined with the help of the ratio

couto,j

/cino,k = δj,k −

√κo,j co,j

/cino,k. (15)

The Hamiltonian has been linearized, hence, the response of co,j to the input field in Eq. (12) is linear. Moreover, forthe purpose of calculating co,j , we can replace the operators cinν,j in the source terms of the Langevin equation (12)

with their mean values cinν,j . We can even formally replace all the input terms with the coherent interaction

HI = i√κo,k

(c†o,kαp exp[−iΩt]− co,kα∗p exp[iΩt]

). (16)

Thus, co,j is given by the Kubo formula where HI plays the role of the perturbation. We note that the optomechanicalcoupling does not conserve the number of excitations and, therefore, co,j has both an elastic (frequency Ω in therotating frame or Ω + ωL in the laboratory frame) and an inelastic (frequency −Ω in the rotating frame or −Ω + ωLin the laboratory frame) components. The elastic part of transmission is obtained after time averaging:

Too(j, k; Ω) =

∣∣∣∣∣ Ω

∫ 2πΩ

0

couto,j (t)

cino,k(t)dt

∣∣∣∣∣2

. (17)

Using Eqs.(15–16) and the Kubo formula, we find

Too(j, k; Ω) =∣∣δj,k − i√κo,jκo,kGRoo(j, k; Ω)

∣∣2 ; Goo(j, k; Ω) = −i∫ ∞

0

dt eiΩt [co,j(t), c†o,k(0)]. (18)

For j 6= k, we recover the formula for Too(j, k; Ω) given in the main text.If the eigenstates of the OMA are localized then Too(j, k; Ω) decays exponentially on large distances and the inverse

localization length can defined as follows:

1

ξ(Ω)= − lim

x→∞

ln[Too(j, k; Ω)

]2x

= − limx→∞

ln |GRoo(j, k; Ω)|x

; x ≡ |j − k|. (19)

We note that Eq.(19) contains only matrix elements of the Green’s function relating operators from sites j and k.These elements can be calculated iteratively with the help of the Dyson’s equation which is similar to that suggestedin Ref.[S2] for the transfer matrix (details of the algorithm can be found in Ref.[S3]).

Generically, one can introduce 4 transmissions in the elastic channel, Tνν′(j, k; Ω), and 4 transmissions in theinelastic one, Tνν′(j, k; Ω) (e.g., the photon-photon transmission, inelastic photon-photon transmission). Similar toEq.(18), these transmissions are described by entries of the matrix Green’s function constructed from four-component”Opto-mechanical×Nambu”-spinors Cj :

CTj (t) =

co,j(t), cm,j(t), c

†o,j(t), c

†m,j(t)

; (20)

GR(j, t; j′, t′) = iθ(t− t′)Cj(t)⊗ C†j′(t′)− C∗j′(t′)⊗ CTj (t). (21)

GR has 16 entries which can be obtained from a straightforward generalization of the Dyson equation for the matrixGreen’s function in frequency space, GR(j, j′; Ω). We have solved the generalized Dyson equation numerically for the

Page 9: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

9

0.9 1.0 1.1 1.2 1.3frequency Ω/Ωm

0

2

4

6

8

10

loc.

length

ξ

a g= 0. 001Ωm

0.9 1.0 1.1 1.2 1.3frequency Ω/Ωm

b g= 0. 05Ωm

FIG. 5. Blue dots: 16 localization lengths obtained after inserting each entry of GR into Eq.(19). Red dots are obtainedfrom Eq.(23). Ten different values of the frequency have been taken to demonstrate that all components of the matrix Green’sfunction show the same spatial decay. Parameters of the OMA are the same as in the main text: Ωo = 1.1Ωm, Jo = σo =0.1Ωm, Jm = σm = 0.001Ωm, g = 0.001Ωm (left panel) and g = 0.05Ωm (right panel). Numerics were done in the absence ofdissipation.

parameters of the OMA given in the main text and compared 16 largest localization lengths governed by insertingeach component of GR into Eq.(19). These localization lengths coincide up to small numerical errors ≤ 1%, see bluedots in Fig.5. This is the related to the symmetries, namely, particle-antiparticle symmetry, time-reversal symmetry,and space-inversion symmetry. The latter appears effectively in the long disordered OMAs due to the self-averaging.The equivalence of the different transmissions on large distances allows one to find the largest localization lengthsof the OMA from any convenient linear combination of |GRab(j, j′; Ω)| ensuring a good convergence of the numericalalgorithm. In particular, we can use “the generalized transmission” of the opto-mechanical excitations

T (j, k; Ω) = Tr(|GR(j, k; Ω)|2

); (22)

and, after disorder averaging, arrive at:

1

ξ2(Ω)= − lim

x→∞

〈ln[T (j, k; Ω)

]〉d

2x, x ≡ |j − k|. (23)

Eq.(23) has been used in the numerical code with the disorder averaging being substituted by the self-averaging of ξin very long systems, see red dots in Fig.5.

2.2 Bogoluibov eigenstates

In the absence of dissipation, there is a simple method which allows one to find the localization length directly fromthe average number of the excitation. This approach is realized after diagonalizing the Hamiltonian (or, equally, theHeisenberg equations of motion) with the help of the Bogoliubov transformation. Let us define eigenmode operators

ds. Generically, ds can be written as follows:

ds =

N∑j=1

∑ν

[u

(ν)j,s cν,j + v

(ν)j,s c

†ν,j

], 1 ≤ s ≤ 2N . (24)

Here u(ν)j,s and v

(ν)j,s are the Bogoliubov coefficients. The transformation matrix that diagonalize the Heisenberg

equations reads as:

T =

U (o)

[V (o)

]∗U (m)

[V (m)

]∗V (o)

[U (o)

]∗V (m)

[U (m)

]∗ , (25)

Page 10: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

10

where U (ν) and V (ν) are N×2N matrices whose entries are the coefficients u(ν)j,s and v

(ν)j,s from Eq.(24). In the absence

of dissipation and instabilities, these coefficients satisfy the following relation:

N∑j=1

∑ν

u

(ν)j,s

[u

(ν)j,s′

]∗− v(ν)

j,s

[v

(ν)j,s′

]∗= δs,s′ . (26)

The minus sign in front of summands v(ν)j,s

[v

(ν)j,s′

]∗is caused by the bosonic commutation relations of the operators

cν,j . As a consequence, T is pseudounitary with the inverse matrix

T−1 =

[ [U (o)

]† [U (m)

]† −[V (o)

]† − [V (m)]†

−[V (o)

]T − [V (m)]T [

U (o)]T [

U (m)]T

]. (27)

The time evolution of the operators cν,j can be obtained using Eq. (27), and it is given by

cν,j(t) =

2N∑s=1

[u

(ν)j,s

]∗ds(0)e−iεst − v(ν)

j,s d†s(0)eiεst

. (28)

Here εs is the frequency of the hybrid (opto-mechanical) eigenmode s. Using Eqs.(28), one can derive

〈0|dsnν,j d†s|0〉 =∣∣∣u(ν)j,s

∣∣∣2 +∣∣∣v(ν)j,s

∣∣∣2 +∑s

∣∣∣v(ν)j,s

∣∣∣2 ; 〈0|nν,j |0〉 =∑s

∣∣∣v(ν)j,s

∣∣∣2 ; (29)

and find the total average number of excitations at a given site j after the eigenmode s is excited:

nj(s) = 〈0|dsnj d†s|0〉 − 〈0|nj |0〉 =∑ν

(∣∣∣u(ν)j,s

∣∣∣2 +∣∣∣v(ν)j,s

∣∣∣2) , nj = no,j + nm,j . (30)

We have subtracted the (background) fluctuations of nj in the ground state since the number of excitation in theOMA always fluctuates due to the optomechanical coupling, see the second term in the RHS of Eq.(1).

The eigenmodes of the OMA can be found via the numerical diagonalization of the Heisenberg equations. Now wesubstitute nj(s) for Too(j, k,Ω) in Eq.(5) and associate the frequency Ω with εs and the origin k with the coordinatewhere the eigenmode s has maximal amplitude. This yields the second expression for ξ2. Thus, the localization lengthcan be estimated from a log-linear fit of nj(s), see the discussion of Eq.(5) in the main text.

3. Weak coupling regime, ∆(m)loc ≤ g ≤ Jm

√JmJo

Let us analyze the behavior of ξ2(g) for the case g ≤ Jm √JmJo where the influence of the optomechanical

coupling on the band structure is negligible. The numerical analysis shows that ξ2(g) is sub-linear, see the left panelof Fig.6, which indicates the presence of non-perturbative contributions. The full theory for this is missing and wegive only phenomenological arguments which are similar to those of the Mott theory [S4] and allow one to explainthe sub-linear growth of ξ2 when g increases up to Jm. For simplicity, we concentrate on the transmission Tom in the

regime ∆(m)loc ∼ Jm/ξ

(0)m . g Jm. Other parameters correspond to Fig.3a in the main text.

Finite transmission Tom requires hybridization of bare optical and mechanical states. The main idea of the phe-nomenological approach is to find a pair of the optical- and the mechanical- states which, being strongly hybridized,provides the largest possible increase of ξ2. In other words, we have to estimate the maximal distance between barelocalization centers which does not violate the necessary condition for the strong hybridization.

Consider an optical state with the frequency inside the unperturbed mechanical band, Ωm − Jm < εo < Ωm + Jm,see the blue wave-function in the right panel of Fig.6. The space coordinates will be counted from the localizationcenter of this optical state. Such a state can be strongly hybridized with the mechanical states if inequality

g 〈Mj |O〉 ≥ |εo − εm,j |, (31)

holds true. Here j is the number of the mechanical state with the localization center at xj > 0 and with the frequencyεm,j ; 〈Mj |O〉 is the overlap between the localized optical and the localized mechanical states

〈Mj |O〉 ∼[ξ(0)m exp

(−xj/ξ(0)

m

)− ξ(0)

o exp(−xj/ξ(0)

o

)]/(ξ(0)m − ξ(0)

o

). (32)

Page 11: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

11

0 0.2 0.4 0.6 0.8 1.0

g/Jm

9.0

9.4

9.8

10.2

10.6

Loc.

length

ξ

X

1 2

X3

0 x1 x2

0 x3

X

xopt

FIG. 6. Left panel: Numerically obtained dependence ξ2(g) for g ≤ Jm, Ω = Ωm (shadowed dots). The red line is an exampleof the fitting which demonstrates the sub-linear nature of this dependence. Right panel: a bare optical state (with blue filling)can be hybridized with different bare mechanical states (with orange filling). The hybridization with the state No.1 is strongbut it is unable to change the largest localization length substantially. The hybridization with the state No.2 is negligible.The hybridization with the (optimal) state No.3 is also strong and is responsible for the increase of ξ2. The optical statebeing hybridized with the mechanical states No.1,3 yields a “double-hump” optomechanical state (with brown filling) which isresponsible for the transmission Tom on large distances.

We recall that ξ(0)m > ξ

(0)o for Ω ' εo, cf. Fig.3a.

Firstly we note, that, unlike the Mott theory, frequencies εo and εm,j are not correlated at g = 0. Therefore,|εo − εm,j | can be arbitrary small even if the localization centers of the bare states are close to each other, xj ξ

(0)o,m ⇒ 〈Mj |O〉 ∼ 1, see the orange state No.1 in the right panel of Fig.6. On the other hand, it is clear that the

1st mechanical state is unable to support an essential increase of the transmission Tom beyond the bare localizationlength.

Tom can become more long-ranged if ξ(0)m . xj . In the extreme case ξ

(0)o ξ

(0)m xj , the overlap becomes

exponentially small, 〈Mj |O〉 ∼ exp(−xj/ξ(0)m ). Distant mechanical states do not obey the condition Eq.(31) and,

therefore, are unimportant, cf. the orange state No.2 in the right panel of Fig.6. However, there is always an optimalstate for which xj is relatively large and the smallness of 〈Mj |O〉 in Eq.(31) is compensated by the smallness of thefrequency separation:

optimal state: 〈Mj |O〉 ∼|εo − εm,j |

g⇒ xopt ∼ ξ(0)

m log(g/

∆(m)loc

); (33)

cf. the orange state No.3 in the right panel of Fig.6. Now we can speculate that, if ∆(m)loc ∼ Jm/ξ

(0)m . g Jm, Tom on

large distances and, correspondingly, ξ2 are governed by the “double-hump” optomechanical state originating mainlyfrom hybridization of the optical state with the optimal mechanical one, see an example in the right panel of Fig.6.Therefore, the largest localization length can be estimated as

ξ2 ' ξ(0)m + const× xopt ' ξ(0)

m

[C1 + C2 log

(g/

∆(m)loc

)]. (34)

(C1,2 are constants of order O(1) which cannot be determined in the frame of the phenomenological approach). Theoptomechnical state may be “multiple-hump” if x3 covers several localization volumes of the bare mechanical states.

If g ∆(m)loc we expect a crossover to the purely perturbative regime.

Since ∆(o)loc Jm ∼ σm, the minimal space distance between two optical states belonging to the frequency range of

the mechanical band is large. We estimate it as ξ(0)o ∆

(o)loc/Jm xopt. This condition allows us to consider relevant

optical states independently.

The universal dependence Eq.(34) can be justified only if the frequency range from ∆(m)loc to Jm is broad and

xopt ξ(0)m . This is not the case for the parameters of the main text, in particular, because the disorder is strong.

Page 12: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

12

The fitting in the left panel of Fig.6 can be done equally by using either Eq.(34) or a power-law dependence withsome non-universal exponent α < 1. More rigorous theory of the weak coupling regime can be developed by exploitingbasic ideas of the virial expansion, see Refs.[55] and [S5,S6] for more details.

4. Localization of hybrid excitations in the case of fluctuating coupling constant

In the main text, we have concentrated on the case where the optomechanical coupling g is one and the same forall cells. In reality, the coupling fluctuates: gj depends on the mean occupation number of the photons on the sitej, gj ∝ co,j , while the cells with smaller optical frequencies host more photons, see Fig.7. This locally enhances gjon these sites. One can speculate that the coupling constant acquires an effective frequency dependence; g becomeslarger for smaller frequencies and it slightly decreases with increasing the frequency. We note that α(ω0,j) depictedin Fig.7 is defined as α(ω0,j) = co,j |g0=0. We do not distinguish α(ω0,j) and co,j since their difference is small,(α(ω0,j)− co,j) ∼ O(g2

0).

1.41.21.00.8local detuning, (ωo, j −ωL)/Ωm

0.8

1.0

1.2

1.4

α(ω

o,j)×

Ωm/α

L

FIG. 7. Dependence of α(ω0,j) = co,j |g0=0 on the optical frequency of the cells.

0.9 1.0 1.1 1.2 1.3frequency Ω/Ωm

0

2

4

6

8

loc.

length

ξ

a⟨g⟩d = 0. 001Ωm

0.9 1.0 1.1 1.2 1.3frequency Ω/Ωm

b⟨g⟩d = 0. 05Ωm

FIG. 8. Frequency dependence of the localization length at Ωo = 1.1Ωm and 〈g〉d = g = 0.001, 0.05Ωm calculated fordifferent values of the optical decay rate κo/Ωm = 0.01, 0.05, 0.1 (blue-, green-, and brown lines, respectively). Solid/dashedlines show profiles with/without fluctuations of the coupling constant.

We have recalculated curves from Fig.4 for fluctuating gj . Results are shown in Fig.8. The disorder averagedcoupling has been adjusted to the same values (a) 〈g〉d = g = 0.001Ωm; and (b) 〈g〉d = g = 0.05Ωm. Solid/dashedlines show the frequency dependence of the localization length with/without fluctuations of the coupling. The profilesat smaller g are almost intact by its fluctuations. When the mean coupling is larger, the left (right) maximum of ξ2(Ω)is suppressed (enhanced) by these fluctuations. This can be explained if, based on Fig.7, we assume that g becomeseffectively larger (smaller) at Ω < Ω (Ω > Ω) and notice that the peaks decrease when g increases, cf. Fig.3. Thus,the fluctuations of gj are able to modify the shape of ξ2(Ω) at relatively large values of the mean coupling constant gthough all qualitatively important features are expected to be robust.

Page 13: Anderson Localization of Composite Excitations in ... fileAnderson Localization of Composite Excitations in Disordered Optomechanical Arrays Thales Figueiredo Roque,1 Vittorio Peano,2

13

[S1] C. C. Gerry, P. L. Knight, Introductory quantum optics, Cambridge Univ. Press (2005).

[S2] A. MacKinnon and B. Kramer, Phys. Rev. Lett. 47, p.1546 (1981).

[S3] A. MacKinnon and B. Kramer, Zeitschrift fur Physik B Condensed Matter 53, pp.1-13 (1983).

[S4] N. F. Mott, Philosophical Magazine 22 pp.7-29 (1970).

[S5] O. Yevtushenko, and A. Ossipov, J. Phys. A: Math. Theor. 40, pp.4691-4716 (2007).

[S6] S. Kronmuller, O.M. Yevtushenko, and E. Cuevas, J. Phys. A: Math. Theor. 43, 075001 (2010).