Quantum Many-Body Dynamics in Optomechanical...

9
Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig 1, * and Florian Marquardt 1,2 1 Institute for Theoretical Physics, Universita ¨t Erlangen-Nu ¨rnberg, Staudtstraße 7, 91058 Erlangen, Germany 2 Max-Planck-Institute for the Science of Light, Gu ¨nther-Scharowsky-Straße 1/Bau 24, 91058 Erlangen, Germany (Received 1 August 2012; revised manuscript received 30 May 2013; published 16 August 2013) We study the nonlinear driven dissipative quantum dynamics of an array of optomechanical systems. At each site of such an array, a localized mechanical mode interacts with a laser-driven cavity mode via radiation pressure, and both photons and phonons can hop between neighboring sites. The competition between coherent interaction and dissipation gives rise to a rich phase diagram characterizing the optical and mechanical many-body states. For weak intercellular coupling, the mechanical motion at different sites is incoherent due to the influence of quantum noise. When increasing the coupling strength, however, we observe a transition towards a regime of phase-coherent mechanical oscillations. We employ a Gutzwiller ansatz as well as semiclassical Langevin equations on finite lattices, and we propose a realistic experimental implementation in optomechanical crystals. DOI: 10.1103/PhysRevLett.111.073603 PACS numbers: 42.50.Wk Introduction.—Recent experimental progress has brought optomechanical systems into the quantum regime: A single mechanical mode interacting with a laser-driven cavity field has been cooled to the ground state [1,2]. Several of these setups, in particular optomechanical crystals, offer the po- tential to be scaled up to form optomechanical arrays. Applications of such arrays for quantum information pro- cessing [3,4] have been proposed. Given these develop- ments, one is led to explore quantum many-body effects in optomechanical arrays. In this work, we analyze the nonlinear photon and phonon dynamics in a homogeneous two-dimensional optomechanical array. In contrast to earlier works [36], here we study the array’s quantum dynamics beyond a quadratic Hamiltonian. To tackle the nonequilib- rium many-body problem of this nonlinear dissipative sys- tem, we employ a mean-field approach for the collective dynamics. First, we discuss photon statistics in the array, in particular, how the photon blockade effect [7] is altered in the presence of intercellular coupling. The main part of the article focuses on the transition of the collective mechanical motion from an incoherent state (due to quantum noise) to an ordered state with phase-coherent mechanical oscillations. For these dynamics, the dissipative effects induced by the optical modes play a crucial role. On the one hand, they allow the mechanical modes to settle into self-induced oscillations [8] once the optomechanical amplification rate exceeds the intrinsic mechanical damping. On the other hand, the funda- mental quantum noise (e.g., cavity shot noise) diffuses the mechanical phases and prevents the mechanical modes from synchronizing. This interplay leads to an elaborate phase diagram characterizing the transition. We develop a semi- classical model to describe the effective dynamics of the mechanical phases and to study the system on finite lattices. While true long-range order is prohibited for a two- dimensional system with continuous symmetry, at least for equilibrium systems, a Beresinskii-Kosterlitz-Thouless transition towards a state with quasi-long-range order is possible. The ordered mechanical phase thus resembles the superfluid phase in two-dimensional cold atomic gases [9] or Josephson junction arrays [10]. Notably, optomechan- ical arrays combine the tunability of optical systems with the robustness and durability of an integrated solid-state device. Other driven dissipative systems that have been studied with regard to phase transitions recently include cold atomic gases [1114], nonlinear cavity arrays [15,16] and optical fibers [17]. In a very recent work and along the lines of [11], the preparation of long-range order for photonic modes was proposed using the linear dissipative effects in an optomechanical array [5]. Our work adds the novel aspect of a mechanical transition to the studies of driven dissipative many-body systems. Model.—We study the collective quantum dynamics of a two-dimensional homogeneous array of optomechanical cells (Fig. 1). Each of these cells consists of a mechanical displacement electric field 0 1 0 1 FIG. 1 (color online). Example implementation of an optome- chanical array: A two-dimensional snowflake optomechanical crystal [38,39] supports localized optical and mechanical modes around defect cavities. Here, we propose arranging them in a superstructure, forming the array. The insets show electric field ~ E and displacement field ~ u of an isolated defect cavity (obtained from finite element simulations). Due to the finite overlap between modes of neighboring sites [29], photons and phonons can hop through the array, see Eq. (2). A wide laser beam drives the optical modes of the array continuously and the reflected light is read out. PRL 111, 073603 (2013) PHYSICAL REVIEW LETTERS week ending 16 AUGUST 2013 0031-9007= 13=111(7)=073603(5) 073603-1 Ó 2013 American Physical Society

Transcript of Quantum Many-Body Dynamics in Optomechanical...

Page 1: Quantum Many-Body Dynamics in Optomechanical Arraysthp2.nat.fau.de/images/5/56/2013_Ludwig_QuantumManyBodyDynamics.pdf · Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig1,*

Quantum Many-Body Dynamics in Optomechanical Arrays

Max Ludwig1,* and Florian Marquardt1,2

1Institute for Theoretical Physics, Universitat Erlangen-Nurnberg, Staudtstraße 7, 91058 Erlangen, Germany2Max-Planck-Institute for the Science of Light, Gunther-Scharowsky-Straße 1/Bau 24, 91058 Erlangen, Germany

(Received 1 August 2012; revised manuscript received 30 May 2013; published 16 August 2013)

We study the nonlinear driven dissipative quantum dynamics of an array of optomechanical systems. At

each site of such an array, a localized mechanical mode interacts with a laser-driven cavity mode via

radiation pressure, and both photons and phonons can hop between neighboring sites. The competition

between coherent interaction and dissipation gives rise to a rich phase diagram characterizing the optical

and mechanical many-body states. For weak intercellular coupling, the mechanical motion at different

sites is incoherent due to the influence of quantum noise. When increasing the coupling strength, however,

we observe a transition towards a regime of phase-coherent mechanical oscillations. We employ a

Gutzwiller ansatz as well as semiclassical Langevin equations on finite lattices, and we propose a realistic

experimental implementation in optomechanical crystals.

DOI: 10.1103/PhysRevLett.111.073603 PACS numbers: 42.50.Wk

Introduction.—Recent experimental progress has broughtoptomechanical systems into the quantum regime: A singlemechanical mode interacting with a laser-driven cavity fieldhas been cooled to the ground state [1,2]. Several of thesesetups, in particular optomechanical crystals, offer the po-tential to be scaled up to form optomechanical arrays.Applications of such arrays for quantum information pro-cessing [3,4] have been proposed. Given these develop-ments, one is led to explore quantum many-body effectsin optomechanical arrays. In this work, we analyze thenonlinear photon and phonon dynamics in a homogeneoustwo-dimensional optomechanical array. In contrast to earlierworks [3–6], here we study the array’s quantum dynamicsbeyond a quadratic Hamiltonian. To tackle the nonequilib-rium many-body problem of this nonlinear dissipative sys-tem, we employ a mean-field approach for the collectivedynamics. First, we discuss photon statistics in the array, inparticular, how the photon blockade effect [7] is altered inthe presence of intercellular coupling. The main part of thearticle focuses on the transition of the collective mechanicalmotion from an incoherent state (due to quantum noise) to anordered state with phase-coherent mechanical oscillations.For these dynamics, the dissipative effects induced by theopticalmodes play a crucial role. On the one hand, they allowthe mechanical modes to settle into self-induced oscillations[8] once the optomechanical amplification rate exceeds theintrinsic mechanical damping. On the other hand, the funda-mental quantum noise (e.g., cavity shot noise) diffuses themechanical phases and prevents the mechanical modes fromsynchronizing. This interplay leads to an elaborate phasediagram characterizing the transition. We develop a semi-classical model to describe the effective dynamics of themechanical phases and to study the system on finite lattices.

While true long-range order is prohibited for a two-dimensional system with continuous symmetry, at leastfor equilibrium systems, a Beresinskii-Kosterlitz-Thouless

transition towards a state with quasi-long-range order ispossible. The ordered mechanical phase thus resemblesthe superfluid phase in two-dimensional cold atomic gases[9] or Josephson junction arrays [10]. Notably, optomechan-ical arrays combine the tunability of optical systems with therobustness and durability of an integrated solid-state device.Other driven dissipative systems that have been studied withregard to phase transitions recently include cold atomicgases [11–14], nonlinear cavity arrays [15,16] and opticalfibers [17]. In a very recent work and along the lines of [11],the preparation of long-range order for photonic modeswas proposed using the linear dissipative effects in anoptomechanical array [5]. Our work adds the novel aspectof a mechanical transition to the studies of driven dissipativemany-body systems.Model.—We study the collective quantum dynamics of a

two-dimensional homogeneous array of optomechanicalcells (Fig. 1). Each of these cells consists of a mechanical

displacement electric field

0 1 0 1

FIG. 1 (color online). Example implementation of an optome-chanical array: A two-dimensional snowflake optomechanicalcrystal [38,39] supports localized optical and mechanical modesaround defect cavities. Here, we propose arranging them in asuperstructure, forming the array. The insets show electric field ~Eand displacement field ~u of an isolated defect cavity (obtainedfromfinite element simulations). Due to the finite overlap betweenmodes of neighboring sites [29], photons and phonons can hopthrough the array, see Eq. (2). Awide laser beam drives the opticalmodes of the array continuously and the reflected light is read out.

PRL 111, 073603 (2013) P HY S I CA L R EV I EW LE T T E R Sweek ending

16 AUGUST 2013

0031-9007=13=111(7)=073603(5) 073603-1 2013 American Physical Society

Page 2: Quantum Many-Body Dynamics in Optomechanical Arraysthp2.nat.fau.de/images/5/56/2013_Ludwig_QuantumManyBodyDynamics.pdf · Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig1,*

mode and a laser driven optical mode that interact via theradiation pressure coupling at a rate g0 (@ ¼ 1):

Hom;j ¼ ayj aj þbyj bj g0ðbyj þ bjÞayj ajþ Lðayj þ ajÞ: (1)

The mechanical mode (bj) is characterized by a frequency

. The cavity mode (aj) is transformed into the frame

rotating at the laser frequency ( ¼ !laser !cav) anddriven at the rate L. In the most general case, bothphotons and phonons can tunnel between neighboring siteshiji at rates J=z and K=z, where z denotes the coordinationnumber. The full Hamiltonian of the array is given by

H ¼ PjHom;j þ Hint, with

Hint ¼ J

z

Xhi;ji

ðayi aj þ aiayj Þ

K

z

Xhi;ji

ðbyi bj þ bibyj Þ: (2)

To bring this many-body problem into a treatable form, we

apply the Gutzwiller ansatz Ayi Aj hAy

i iAj þ Ayi hAji

hAyi ihAji to Eq. (2). The accuracy of this approximation

improves if the number of neighboring sites z increases.For identical cells, the index j can be dropped and theHamiltonian reduces to a sum of independent contribu-tions, each of which is described by

Hmf ¼ HomJðayhaiþ ahayiÞKðbyhbiþ bhbyiÞ: (3)

Hence, a Lindblad master equation for the single cell density

matrix , d=dt ¼ i½Hmf; þ D½aþ D½b can

be employed. The Lindblad terms D½A ¼ A Ay AyA =2 AyA=2 take into account photon decay at arate and mechanical dissipation (here assumed dueto a zero temperature bath) at a rate .

Photon statistics.—Recently, it was shown that the effectof the photon blockade [7] can appear in a single optome-chanical cell: The interaction with the mechanical modeinduces an optical nonlinearity of strength g20= [7,18] and

the presence of a single photon can hinder other photonsfrom entering the cavity. To observe this effect, the non-linearity must be comparable to the cavity decay rate, i.e.,g20= * , and the laser drive weak (L ) [7,19].

To study nonclassical effects in the photon statistics,we analyze the steady-state photon correlation function

gð2ÞðÞ ¼ hayðtÞayðtþ Þaðtþ ÞaðtÞi=haðtÞyaðtÞi2 [20] atequal times ( ¼ 0). Here (Fig. 2), we probe the influence ofthe collective dynamics by varying the optical couplingstrength J, while keeping the mechanical coupling K zerofor clarity. We note that, when increasing J, the opticalresonance effectively shifts: ! þ J. To keep the photonnumber fixed while increasing J, the detuning has to beadapted [21]. In this setting, we observe that the interactionbetween the cells suppresses antibunching [Fig. 2(b)]. Thephoton blockade is lost if the intercellular coupling becomeslarger than the effective nonlinearity, 2J * g20=. Above

this value, the photon statistics shows bunching, and ulti-mately reaches Poissonian statistics for large couplings.

Similar physics has recently been analyzed for coupledqubit-cavity arrays, [21]. For large coupling strengths,though, Fig. 2(a) reveals signs of the collective mechanicalmotion (hatched area). There we observe the correlationfunction to oscillate (at the mechanical frequency) and toshow bunching. We will now investigate this effect.Collective mechanical quantum effects.—To describe the

collective mechanical motion of the array, we focus on thecase of purely mechanical intercellular coupling (K > 0,J ¼ 0) for simplicity. Note, though, that the effect is alsoobservable for optically coupled arrays, as discussed above.As our main result, Figs. 3(a) and 3(d) show the

sharp transition between incoherent self-oscillations and aphase-coherent collective mechanical state as a function ofboth laser detuning and coupling strength K: In theregime of self-induced oscillations, the phonon number

hbybi reaches a finite value. Yet, the expectation value

hbi remains small and constant in time. When increasing

the intercellular coupling, though, hbi suddenly startsoscillating and reaches a steady state

hbiðtÞ ¼ bþ reieff t: (4)

Here, we introduced the mechanical coherence r and theoscillation frequency eff , which is shifted by the opticalfields and the intercellular coupling, cf. Eq. (6).Our more detailed analysis (see below) indicates

that this transition results from the competition betweenthe fundamental quantum noise of the system and thetendency of phase locking between the coupled nonlinearoscillators. Below threshold, the quantum noise from thephonon bath and the optical fields diffuses the mechanicalphases at different sites and drives the mechanical motion

detuning

optic

al c

oupl

ing

stre

ngth

photon correlations

-1 0-0.5-1.5

1

2

4

3

coupling strength0.001 0.01 0.1 1

1

3

phot

on c

orre

latio

ns

(a) (b)

0.5

0

1

FIG. 2 (color online). Loss of the photon blockade for increas-ing optical coupling in an array of optomechanical cavities.(a) The equal time photon correlation function shows antibunch-ing (gð2Þð0Þ< 1) and bunching (gð2Þð0Þ> 1) as a function ofdetuning and optical coupling strength J. The smallest valuesof gð2Þð0Þ are found for a detuning 0 ¼ g20=. (b) When

increasing the coupling J while keeping the intracavity photonnumber constant, i.e., along the dashed line in panel (a), thephoton blockade is lost (black solid line). For a smaller drivingpower (blue solid line, L ¼ 5 105), anti-bunching is morepronounced and the behavior is comparable to that of a nonlinearcavity (dashed line). The hatched area in (a) outlines a regionwhere a transition towards coherent mechanical oscillations hasset in. ¼ 0:3 , L ¼ 0:65, g0 ¼ 0:5 , ¼ 0:074 .

PRL 111, 073603 (2013) P HY S I CA L R EV I EW LE T T E R Sweek ending

16 AUGUST 2013

073603-2

Page 3: Quantum Many-Body Dynamics in Optomechanical Arraysthp2.nat.fau.de/images/5/56/2013_Ludwig_QuantumManyBodyDynamics.pdf · Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig1,*

into an incoherent mixed state. The reduced density

matrix ðmÞ is predominantly occupied on the diagonal, see

Fig. 3(b), and the Wigner distribution, Wðx;pÞ¼ð1=@ÞR11hxyjðmÞjxþyie2ipy=@dy, has a ringlike shape, reflect-

ing the fact that the mechanical phase is undetermined[22,23]. Above threshold, the mechanical motion at differentsites becomes phase locked, and the coherence parameter rreaches a finite value. The emergence of coherence also

becomes apparent from the off-diagonal elements of ðmÞ[Fig. 3(c)]. The corresponding Wigner function assumes theshape of a coherent state with a definite phase oscillating inphase space. Thus, this transition spontaneously breaks thetime-translation symmetry. In a two-dimensional implemen-tation, true long range order is excluded, but the coherencebetween different sites is expected to decay as a power lawwith distance.We also note that this transition is the quantummechanical analogon of classical synchronization, whichwas studied for optomechanical systems in [24–26]. Animportant difference is, though, that the classical nonlineardynamics was analyzed for an inhomogeneous (with disor-deredmechanical frequencies) system in the absenceof noise

[24–26], while in our case disorder is only introduced viafundamental quantum noise. Quantum synchronization hasalso been discussed in the context of linear oscillators [27]and nonlinear cavities [28] recently.The laser detuning determines both the strength of the

self-oscillations and the influence of the cavity shot noiseon the mechanical motion. It turns out that the diffusion ofthe mechanical phases is pronounced close to the onset ofself-oscillations and at the mechanical sideband [29]. Aswe will show below, even the coherent coupling betweenthe mechanical phases (ultimately leading to synchroniza-tion) is tunable via the laser frequency. As a result, thesynchronization threshold depends nontrivially on thedetuning parameter , see Fig. 3(a).Langevin dynamics on finite lattices.—In order to gain

further insight into the coupling and decoherence mecha-nisms as well as effects of geometry and dimensionality,we analyze the semi-classical Langevin equations of thefull optomechanical array:

_i ¼i

2

i þ ig0jij2 þ i

K

z

Xhiji

j þffiffiffiffi

2

s

_i ¼ iþ ig0ði þ i Þ

2

i iL þ

ffiffiffiffi

2

r: (5)

The fluctuating noise forces ¼;ðtÞ mimic the effects

of the zero temperature phonon bath and the cavity shotnoise, respectively. They are independent at each site andobey hi ¼ 0 and hðtÞ

ðt0Þi ¼ ðt t0Þ. In this con-text, h. . .i denotes the average over different realizations ofthe stochastic terms. This Langevin approach is equivalentto the truncated Wigner approximation (see [30] for areview), and it has shown good qualitative agreementwith the full quantum dynamics for a single optomechan-ical cell [22,31]. It allows us to treat the effects of quantumfluctuations at all wavelengths on the spatial phase corre-lations via numerical simulations. At this point, a fullquantum treatment for sufficiently large systems remainsa challenging problem for future studies.First, we study the onset of quasi-long-range order in a

finite system. To this end we evaluate the correlationsCðd ¼ ji jjÞ ¼ hei’iei’ji, where ei’i ¼ i=jij.Numerical calculations on a 30 30 square lattice [seeFig. 4(a)] indicate that for weak intercellular coupling themechanical phases at different sites are uncorrelated even forsmall distances d. When increasing the coupling strength,however, themechanicalmotion becomes correlated over thewhole array with only a slow decrease with distance. Thecoupling threshold, here defined by setting a lower bound ofCð14Þ> 0:01, varies with the coordination number, seeFig. 4(b). Within the mean-field approximation, i.e., for alattice with global coupling of all sites, fluctuations betweenneighboring sites and hence the threshold value are under-estimated. The coupling threshold grows with the quantumparameter [22], i.e., the ratio of optomechanical coupling andcavity decay rate, g0=, see Fig. 4(c): For g0 , single

0 0.25 0.50

0.5

1

0

15

15 0

15 0

5

0

-5-5 0 5

5

0

-5-5 0 5

mec

hani

cal c

oupl

ing

detuning

mechanical coherence

incoherent

coherent

onse

t of

self

-osc

illat

ions

0

0.5

1.5

1

-1 0-0.5-1.5

0.25

0

0.5

10.5

bc

b b

c c

mec

hani

cal

coupling

opticalreadout

cohe

renc

e

(a)

(b)

(c)

(d)

0

15

FIG. 3 (color online). Transition from the incoherent to thesynchronized (coherent) phase: (a) Mechanical coherence r[Eq. (4)] as a function of laser detuning andmechanical couplingK. At weak coupling, the self-oscillations are incoherent, r ¼ 0,due to quantum noise. When increasing the coupling strength, thesystems shows a sharp transition towards the ordered regime,where the mechanical oscillations are phase coherent, r > 0. (b),(c) Modulus of the density matrix elements (in Fock space) andWigner density of the collective mechanical state in the incoherent(b) and the coherent regime (c), as marked in (a). (d) Mechanicalcoherence r as a function of coupling strength K along the dashedline in (a). The dotted line shows the optical readout of coherence,i.e., the oscillating component of the photon number hayai,proportional to the intensity of the reflected beam and thusdirectly accessible in experiment. g0 ¼ ¼ 0:3 , L ¼ 1:1, ¼ 0:074 .

PRL 111, 073603 (2013) P HY S I CA L R EV I EW LE T T E R Sweek ending

16 AUGUST 2013

073603-3

Page 4: Quantum Many-Body Dynamics in Optomechanical Arraysthp2.nat.fau.de/images/5/56/2013_Ludwig_QuantumManyBodyDynamics.pdf · Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig1,*

photons and phonons interact strongly and quantum fluctua-tions hamper synchronization.

Synchronization threshold.—For an analytical approach,the complexity of the Langevin equations can be reducedby integrating out the dynamics of the optical modes andthe mechanical amplitudes and by going back to the mean-field approximation [29]. The resulting equation describesthe coupling of the mechanical phase on a single site, ’, toa mean field :

_’ ¼ ð AÞ þ KR cosð ’Þ þ K1R sinð ’Þþ

ffiffiffiffiffiffiffiffiffiffi2D’

q’ þOðR2Þ: (6)

Here, the order parameter is defined as hei’ji Rei. Therate K1 ¼ ðd K=2ÞK= determines the coupling ofphases mediated by slow amplitude modulations betweenneighboring sites. These beatmodes couple back to the phasedynamics via the amplitude dependent optical spring effect,ð AÞ þ dðA AÞ= A, where d ¼ Aðd=dAÞjA¼ A, andthe bare mechanical coupling K, leading to two opposingterms in K1. Here, A denotes the steady state mechanicalamplitude and the amplitude decay rate set by the optical

field. The fluctuating noise force’ comprises the effects

of mechanical fluctuations and radiation pressure noise andis characterized by a diffusion constant D’ [29,31].

Equation (6) reveals the close connection to theKuramoto model [32] and the two-dimensional xy model.In the incoherent regime, the order parameter R is zero andthe phase fluctuates freely. In the coherent regime, therestoring force K1R leads the phase ’ towards a fixedrelation with . The cosine term only renormalizes theoscillation frequency. This statement can be clarified by alinear stability analysis, see [29,33]. It turns out that theincoherent phase becomes unstable for

K1 ¼ 2D’; (7)

defining the threshold of the transition. Moreover, if K1

becomes negative, no stable phase synchronization ispossible. This situation arises if d< 0, or for largeintercellular coupling rates K > 2d, see Fig. 3(d).Experimental prospects.—We note that observation of

the mechanical phase transition does not require singlephoton strong coupling (g0 * ): The quantum fluctua-tions of the light field will dominate over thermal fluctua-tions as long as 4g20jj2= > kBT=Q. This is essentially

the condition for ground-state cooling, which has beenachieved using high-Q mechanical resonators and cryo-genic cooling [1,2], see Table I. In contrast, the photon-blockade effect (Fig. 2) requires low temperatures T andg20 * , or at least, in a slightly modified setup [35,36],

g0 * . While still challenging, optomechanical systemsare approaching this regime [37].Microfabricated optomechanical systems such as micro-

resonators (e.g., [34]), optomechanical crystals (e.g., [2])or microwave-based setups (e.g., [1]) lend themselves toextensions to optomechanical arrays. Here, we focus onoptomechanical crystals, which are well suited due totheir extremely small mode volumes. The properties oftwo-dimensional optomechanical crystals have been ana-lyzed in [38]. The finite overlap of the evanescent tails ofadjacent localized modes [29] results in a coupling of theform of Eq. (2), in analogy to the tight-binding descriptionof electronic states in solids. Sufficiently strong opticaland mechanical hopping rates are feasible, see [24] forone-dimensional and [29] for two-dimensional structures.The simultaneous optical driving of many cells may berealized by a single broad laser beam irradiating the slab,see Fig. 1. Alternatively, similar physics may be observedfor many mechanical modes coupling to one extended

0 0.5 1

0.1 0.110 4 8 12

corr

elat

ions

corr

elat

ions

coupling

thre

shol

d

thresholddistance on lattice

quantum parameter

mean- eld

mean- eld

square

hexag.

square

mean- eld,quantum

(a) (b)

(c)

0

0.5

1

0

0.5

1fi

fifi

0

0.1

0.2

FIG. 4 (color online). Langevin dynamics on finite lattices:(a) Correlations Cðd ¼ ji jjÞ ¼ jhei’iei’j ij in a 30 30optomechanical array. Quasi-long-range order sets in forsufficiently large coupling strengths. K ¼ f0:09; 0:105; 0:107;0:12; 0:15g . (b) Correlations over a distance of d ¼ 14 as afunction of mechanical coupling strength K for a square lattice(z ¼ 4, squares), a hexagonal lattice (z ¼ 6, triangles) slightlybelow the mean-field result (circles). (c) Coupling threshold as afunction of quantum parameter g0= (squares: square lattice,empty (filled) circles: semiclassical (quantum) mean-field ap-proach). þ g20= ¼ 0:34, g0 ¼ 0:1 in (a),(b), g0L ¼ 0:33in (c), other parameters as in Fig. 3.

TABLE I. Parameters of optomechanical systems [1,2,34]: Temperature of phonon bath T,strength of mechanical fluctuations nth kBT=Q, strength of cavity shot noise opt 4g20j j2=, quantum parameter g0= and approximate size L.

Setup T½K nth= opt= g0= L½mMicrowave based 25 mK 104 7 103 103 100Optomech. crystal 20 K 103 4 103 2 103 4Microtoroid 650 mK 7 102 6 102 5 104 30

PRL 111, 073603 (2013) P HY S I CA L R EV I EW LE T T E R Sweek ending

16 AUGUST 2013

073603-4

Page 5: Quantum Many-Body Dynamics in Optomechanical Arraysthp2.nat.fau.de/images/5/56/2013_Ludwig_QuantumManyBodyDynamics.pdf · Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig1,*

in-plane optical mode [6,24,25] (thereby effectively real-izing global coupling).

The transition towards the synchronized phase can bedetected by probing the light reflected from the optome-chanical array and measuring the component oscillating atthe mechanical frequency, see Fig. 3(d). To read out corre-lations between individual sites, the intensities of individualdefect cavities may be analyzed [29], for example by evan-escently coupling them to tapered fibers or waveguides.

We expect the transition to be robust against disorder[24]. One may also study the formation of vortices andother topological defects induced by engineered irregular-ities and periodic variations, and explore various differentlattice structures or the possibility of other order phases(e.g., antiferromagnetic order). Thus, optomechanicalarrays provide a novel, integrated, and tunable platformfor studies of quantum many body effects.

The authors would like to thank Oskar Painter and BjornKubala for valuable discussions. This work was supportedby the DFG Emmy-Noether program, an ERC startinggrant, the DARPA/MTO ORCHID program, and the ITNnetwork cQOM.

*[email protected][1] J. D. Teufel, T. Donner, D. Li, J.W. Harlow, M. S. Allman,

K. Cicak, A. J. Sirois, J. D. Whittaker, K.W. Lehnert, andR.W. Simmonds, Nature (London) 475, 359 (2011).

[2] J. Chan, T. P.M. Alegre, A. H. Safavi-Naeini, J. T. Hill,A. Krause, S. Groblacher, M. Aspelmeyer, and O. Painter,Nature (London) 478, 89 (2011).

[3] D.E. Chang, A.H. Safavi-Naeini,M. Hafezi, and O. Painter,New J. Phys. 13, 023003 (2011).

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

[5] A. Tomadin, S. Diehl, M.D. Lukin, P. Rabl, and P. Zoller,Phys. Rev. A 86, 033821 (2012).

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

[7] P. Rabl, Phys. Rev. Lett. 107, 063601 (2011).[8] T. Carmon, H. Rokhsari, L. Yang, T. J. Kippenberg, and

K. J. Vahala, Phys. Rev. Lett. 94, 223902 (2005); F.Marquardt, J. G. E. Harris, and S.M. Girvin, ibid. 96,103901 (2006); C. Metzger, M. Ludwig, C. Neuenhahn,A. Ortlieb, I. Favero, K. Karrai, and F. Marquardt, ibid.101, 133903 (2008); Q. Lin, J. Rosenberg, X. Jiang, K. J.Vahala, and O. Painter, ibid. 103, 103601 (2009); M.Bagheri, M. Poot, M. Li, W. P. H. Pernice, and H.X.Tang, Nat. Nanotechnol. 6, 726 (2011).

[9] Z. Hadzibabic, P. Kruger, M. Cheneau, B. Battelier, andJ. Dalibard, Nature (London) 441, 1118 (2006).

[10] R. Fazio and H. van der Zant, Phys. Rep. 355, 235 (2001).[11] S. Diehl, A. Micheli, A. Kantian, B. Kraus, H. P. Buchler,

and P. Zoller, Nat. Phys. 4, 878 (2008).

[12] D. Nagy, G. Konya, G. Szirmai, and P. Domokos, Phys.Rev. Lett. 104, 130401 (2010).

[13] K. Baumann, C. Guerlin, F. Brennecke, and T. Esslinger,Nature (London) 464, 1301 (2010).

[14] T. E. Lee, H. Haffner, and M.C. Cross, Phys. Rev. A 84,031402 (2011).

[15] M. J. Hartmann, F. G. S. L. Brandao, and M.B. Plenio,Nat. Phys. 2, 849 (2006).

[16] A. D. Greentree, C. Tahan, J. H. Cole, and L. C. L.Hollenberg, Nat. Phys. 2, 856 (2006).

[17] D. E. Chang, V. Gritsev, G. Morigi, V. Vuletic, M.D.Lukin, and E.A. Demler, Nat. Phys. 4, 884 (2008).

[18] A. Nunnenkamp, K. Børkje, and S.M. Girvin, Phys. Rev.Lett. 107, 063602 (2011).

[19] A. Kronwald, M. Ludwig, and F. Marquardt, Phys. Rev. A87, 013847 (2013).

[20] L. Mandel and E. Wolf, Optical Coherence and QuantumOptics (Cambridge University Press, Cambridge, England,1995).

[21] F. Nissen, S. Schmidt, M. Biondi, G. Blatter, H. E. Tureci,and J. Keeling, Phys. Rev. Lett. 108, 233603 (2012).

[22] M. Ludwig, B. Kubala, and F. Marquardt, New J. Phys. 10,095013 (2008).

[23] J. Qian, A.A. Clerk, K. Hammerer, and F. Marquardt,Phys. Rev. Lett. 109, 253601 (2012).

[24] G. Heinrich, M. Ludwig, J. Qian, B. Kubala, and F.Marquardt, Phys. Rev. Lett. 107, 043603 (2011).

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

[26] M. Zhang, G. S. Wiederhecker, S. Manipatruni, A.Barnard, P. McEuen, and M. Lipson, Phys. Rev. Lett.109, 233906 (2012).

[27] G. L. Giorgi, F. Galve, G. Manzano, P. Colet, and R.Zambrini, Phys. Rev. A 85, 052101 (2012).

[28] T. E. Lee and M.C. Cross, arXiv:1209.0742.[29] See Supplemental Material at http://link.aps.org/

supplemental/10.1103/PhysRevLett.111.073603 for detailson the effective phase equations, possible experimentalimplementations and the numerical methods used.

[30] A. Polkovnikov, Ann. Phys. (Amsterdam) 325, 1790 (2010).[31] D. A. Rodrigues and A.D. Armour, Phys. Rev. Lett. 104,

053601 (2010).[32] Y. Kuramoto, Prog. Theor. Phys. Suppl. 79, 223 (1984).[33] S.H. Strogatz and R.E. Mirollo, J. Stat. Phys. 63, 613

(1991).[34] E. Verhagen, S. Deleglise, S. Weis, A. Schliesser, and T. J.

Kippenberg, Nature (London) 482, 63 (2012).[35] K. Stannigel, P. Komar, S. J.M. Habraken, S. D. Bennett,

M. D. Lukin, P. Zoller, and P. Rabl, Phys. Rev. Lett. 109,013603 (2012).

[36] M. Ludwig, A. H. Safavi-Naeini, O. Painter, and F.Marquardt, Phys. Rev. Lett. 109, 063601 (2012).

[37] J. Chan, A. H. Safavi-Naeini, J. T. Hill, S. Meenehan, andO. Painter, Appl. Phys. Lett. 101, 081115 (2012).

[38] A. H. Safavi-Naeini and O. Painter, Opt. Express 18,14 926 (2010).

[39] A. H. Safavi-Naeini and O. Painter, New J. Phys. 13,013017 (2011).

PRL 111, 073603 (2013) P HY S I CA L R EV I EW LE T T E R Sweek ending

16 AUGUST 2013

073603-5

Page 6: Quantum Many-Body Dynamics in Optomechanical Arraysthp2.nat.fau.de/images/5/56/2013_Ludwig_QuantumManyBodyDynamics.pdf · Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig1,*

Supplemental Material for “Quantum many-body dynamics in optomechanicalarrays”

Max Ludwig1, ∗ and Florian Marquardt1, 2

1Institute for Theoretical Physics, Universitat Erlangen-Nurnberg, Staudtstraße 7, 91058 Erlangen, Germany2Max Planck Institute for the Science of Light, Gunther-Scharowsky-Straße 1/Bau 24, 91058 Erlangen, Germany

MEAN-FIELD PHASE EQUATION

In this section, we provide details for the derivation ofthe mean-field phase equation, Eq. (S.4), starting fromthe equations of motion of the optomechanical array, Eqs.(5). Introducing phases ϕj and amplitudes Aj (in unitsof the mechanical ground state width) as new coordinatesof mechanical motion and omitting fast oscillating terms,the so-called Hopf equations are derived directly from (5).We follow [1], but add noise terms:

ϕi = −Ω(A) +K

zAi

∑〈ij〉

Aj cos(ϕj − ϕi) +ξϕAi

Ai = −γ (Ai − A)− K

z

∑〈ij〉

Aj sin(ϕj − ϕi) + ξA.(S.1)

The steady amplitude A and the amplitude decay rateγ are determined by the optical field: γ(A − A) =

(Γ +

Γopt(A))A/2, where Γopt = −4g0〈|α|2 sinϕ〉T and where

〈...〉T denotes the average over one mechanical period.The mechanical oscillation frequency is modified via anamplitude dependent optical spring effect: Ω(A) = Ω −2g0〈|α|2 cosϕ〉T /A. The fluctuating noise forces ξϕ and

ξA comprise the effects of the phonon bath and the cavityshot noise, see [2] and below.

For weak coupling, K/z Ω, the fluctuations of themechanical amplitudes around the steady state value Aare given by [1]

δAi(t) ≈ −KA

∑〈ij〉

sin(ϕj(t)− ϕi(t)). (S.2)

These beat modes introduce an effective second ordercoupling between phases at different sites, as can be seenafter plugging Eq. (S.2) into the Hopf equation for ϕi(S.1) and performing the time averages 〈...〉T :

ϕi = −Ω(A) +K

z

∑〈ij〉

cos(ϕj − ϕi) +K dΩ

z γ

∑〈ij〉

sin(ϕj − ϕi)

+K2

2z2γ

∑〈ij〉

∑〈jk〉

(sin(2ϕj − ϕk − ϕi)− sin(ϕk − ϕi)

)+K2

2z2γ

∑〈ij〉

∑〈ik〉

sin(ϕj + ϕk − 2ϕi) + ξϕ, (S.3)

where we introduced dΩ = A dΩdA |A=A. This equation is

similar to the xy model and the Kuramoto model in thepresence of noise, but with additional terms that mainlyshift the frequency (the cos-term) and indicate higher-order coupling (the contributions of the double sums).

Ultimately, we apply a mean-field approximation: Wereplace eiϕj for neighboring cells by 〈eiϕj 〉 ≡ ReiΨ andei2ϕj by 〈ei2ϕj 〉 ≡ R2e

iΨ2 , where 〈...〉 denotes the aver-age over all sites [3], and arrive at the effective phaseequation, Eq. (S.4), with additional second order contri-butions:

ϕ = −Ω(A) +KR cos(Ψ− ϕ) +K1R sin(Ψ− ϕ)

+K2R2 sin(2Ψ− 2ϕ) +K2RR2 sin(Ψ2 −Ψ− ϕ) + ξϕ,

(S.4)

where we introduced the effective coupling rates

K1 = KdΩ/γ −K2, (S.5)

K2 = K2/2z2γ. (S.6)

PHASE DIFFUSION

Here, we list some more details of the phase diffusionin the system based on the analysis given by Rodriguesand Armour [2] for a single optomechanical cell. Thediffusion constant associated with ξϕ (see Eq. (S.3)) isgiven by

Dϕ =1

A2

(Dϕ +

δΩ2

γ2DA

), (S.7)

where Dϕ and DA correspond to the noise acting onphase and amplitude in Eqs. (S.1), and where the diffu-sion rates are defined as

2Dϕ,A =

∫ ∞−∞

dτ〈ξϕ,A(t+ τ)ξϕ,A(t)〉T (S.8)

and likewise for Dϕ. For the explicit expressions we referto [2]. In the limit of g0A Ω, and for Ω κ and ∆ =Ω, the maximum diffusion rates can be approximated by

2Dϕ,A ≈ Γ + Γopt, (S.9)

i.e. the sum of the intrinsic mechanical damping Γ andthe optomechanical damping rate at the mechanical side-band

Γopt ≈ 4g20 |α|2/κ. (S.10)

Page 7: Quantum Many-Body Dynamics in Optomechanical Arraysthp2.nat.fau.de/images/5/56/2013_Ludwig_QuantumManyBodyDynamics.pdf · Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig1,*

2

-1 0 10

0.1

0.2-1 0 1

123

a

bphase di!usion

detuning

phonon number

mechanical coupling0.2 0.4

1

0

2

3

4

optic

al c

oupl

ing

0

0.5

1.0

0

c mechanical coherence

Figure S1. Additional details for the quantum dynamics of theoptomechanical array within the mean-field approximation:(a) The phonon number 〈b†b〉 shows, as a function of detuning,maxima at the resonance and at the sideband (∆ ≈ Ω −g20/Ω). (b) The diffusion constant for the mechanical phase,Dϕ, for an uncoupled (K = 0, solid line) and a coupled array(K = 0.1 Ω, dash-dotted line). Other parameters as in Fig. 3.(c) Interplay of mechanical and optical coupling: Mechanicalcoherence is observed as a function of both mechanical andoptical coupling. ∆ +J = Ω/2, other parameters as in Fig. 2.

The diffusion of the mechanical phase can also be stud-ied using the full quantum simulations and evaluating thelinewidth of the correlator 〈b(t)b†(0)〉 ∼ e−(iΩeff+Dϕ)t, seeFig. S1(b). Close to the onset, for small amplitudes Aand weak amplitude damping γ, the mechanical phasesare very susceptible to quantum noise preventing syn-chronization. For finite coupling strengths, the diffusionis also enhanced, most strikingly at the mechanical side-band. As a result, the synchronization threshold showsa minimum between the onset of self-oscillations and thesideband, as observed in Fig. 3(a). From extended simu-lations, we find that this behavior is generic for systemsin the resolved sideband regime (Ω > κ).

STABILITY ANALYSIS

Here, we briefly recall details of the stability analysisleading to Eq. (7), which, for the case of the Kuramotomodel, has been given in [4, 5]. We consider the density of

the mechanical phases, %(ϕ). It is normalized,∫ 2π

0%(ϕ) =

1, and the order parameterR and the mean-field Ψ can be

computed from ReiΨ =∫ 2π

0eiϕ%(ϕ)dϕ (and likewise for

R2 and Ψ2). The Fokker-Planck equation correspondingto Eq. (S.4) is given by

∂t%+ ∂ϕ(%v)

= Dϕ∂2ϕ% (S.11)

with a velocity

v = −Ω(A) +K cos(Ψ− ϕ) +K1R sin(Ψ− ϕ) +O(R2),(S.12)

where second order contributions can be neglected forthis linear analysis. In the unsynchronized regime, themechanical phases are equally distributed over the inter-val [0, 2π], and % = (2π)−1. To study the time evolu-

tion of a small fluctuation on top of the incoherent back-ground, we employ the ansatz [5]:

%(ϕ) =1

2π+ c(t)eiϕ + c∗(t)e−iϕ, (S.13)

leading to

c = −(i(Ω− K

2

)+(Dϕ −

K1

2

))c. (S.14)

This equation reveals that the incoherent solution % =(2π)−1 becomes unstable for

K1=2Dϕ. (S.15)

and thus defines the coupling threshold for the synchro-nization transition [5].

EXPERIMENTAL IMPLEMENTATION

(a) Competing setups - Here, we provide some detailson possible experimental implementations suitable forobserving the mechanical transition:

• Microdisks [6–8] and microtoroids [9–11] fabricatedon microchips: Strong optical coupling betweenresonators is feasible via evanescent fields [8]. Forboth types of setups, scalability still has to beshown.

• Micro- and nanomechanical beams [12, 13] or mem-branes [14] coupling to superconducting microwavecavities: Mechanical interaction may be achievedvia a common support or, capacitively, by applyinga voltage bias between the mechanical resonators.Two-dimensional arrays of coupled microwave cavi-ties are starting to be developed [15]. Related elec-tromechanical systems (see, e.g., [16] for a setupcomprising a nanobeam coupling to a supercon-ducting single electron transistor) may also be em-ployed.

• Optomechanical crystals [17–19] feature smallmode volumes and are thus very suitable for ex-tensions to optomechanical arrays. Some details ofthe proposed implementation are given below (Fig.S2).

(b) Required parameters: According to our semiclassicalanalysis, the essential requirement is

KA2 & Γopt > Γnth. (S.16)

In this case, quantum noise (Γopt) dominates over ther-mal fluctuations (Γnth), which enter the model by replac-ing Γ → Γnth ≈ kBT/Q in Eqs. (5). The mechanicaltransition can then be studied by varying Γopt via thelaser detuning ∆.

Page 8: Quantum Many-Body Dynamics in Optomechanical Arraysthp2.nat.fau.de/images/5/56/2013_Ludwig_QuantumManyBodyDynamics.pdf · Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig1,*

3

1 2 3 4

9.10

9.15m

echa

nica

l fre

quen

cya

b

0

-1

1op

tical

freq

uenc

y

0

-1

1

1 2 3 4 5 6196

200

204 defect spacing

defect spacing

Figure S2. Variation of coupling strength with separationbetween defect cavities: (a) mechanical and (b) optical fre-quencies of symmetric (cross) and antisymmetric (diamond)normal modes for two defects on a snowflake optomechani-cal crystal (as discussed in the main text). The insets showthe displacement field (electrical field) components of the an-tisymmetric mode for a defect separation of 2. The barefrequencies of the localized eigenmodes are 9.12 GHz and200.5 THz, respectively. We note that the optical splittings(and likewise J) change sign when varying the defect separa-tion. Results obtained from finite element simulations.

Recent experiments have demonstrated Γopt > Γnth,see Table I. We note that two-dimensional optomechan-ical crystal devices are expected to show very good op-tomechanical properties [18], even exceeding those of ex-isting one-dimensional setups [17].

(c) Photon and phonon hopping amplitudes: Using fi-nite element simulations, we studied the hybridizationof the photon and phonon modes of two defect cavitiesinside a two-dimensional snowflake silicon optomechan-ical crystal to obtain the hopping constants as a func-tion of defect separation. Parameters: relative permit-tivity εr = 11.68, Poisson ratio 0.17, density 2329 kg/m3,Young’s modulus 170 GPa, snowflake design [20] withlattice constant 500 nm, snowflake radius 168 nm andsnowflake width 60 nm. The mechanical and optical split-tings (i.e. couplings 2K/z and 2J/z, respectively) reachvalues of up to 8 % and 4 % of the mechanical and opticaleigenfrequencies, respectively (Fig. S2). These values arecompatible with the requirements given by Eq. (S.16),even for very small oscillation amplitudes of the order ofthe mechanical zero-point width, A ≈ 1.

Other experimental approaches may realize differentcoupling terms, e.g. (xi−xj)2 [21]. Additional fast rotat-

ing terms like bibj are, however, negligible for K/z Ω.

We note that the transition from incoherent to syn-chronized dynamics is also observable for extended opti-cal modes (J Ω, κ), see Fig. S1(c).

(d) Optical drive: In principle, the methods used fordriving single defect modes via tapered and dimpledfibers [22] and optical waveguides [23, 24] can be extendedto larger scales. To limit experimental efforts, however,one may drive the array using a freestanding broad laserbeam, see the schematic picture in Fig. 1. The couplingbetween laser mode and the in-plane cavity modes willbe relatively weak, which, however, can be compensatedvia the laser power. Preliminary experimental results onfree space coupling to optomechanical crystals have beenreported in [25].

(e) Detection: To detect the mechanical transition,measurements of the optical field emitted from the arrayare sufficient, see Fig. 3(d). To determine correlationsbetween two separate lattice sites, the intensity emanat-ing from these defect cavities may be probed by two ta-pered fibers in the near-field of the selected cells [22] or byespecially designed waveguides [23, 24]. The correlationsbetween the mechanical sideband components of the op-

tical intensities Ii(t) =∫ t+2π/Ωeff

teiΩeff t

′ |αi|2(t′)dt′, arethen proportional to the mechanical correlations,

〈IiI?j 〉 ∝ 〈eiϕie−iϕj 〉. (S.17)

NUMERICAL METHODS

To study the quantum dynamics of the system, we nu-merically integrate the Lindblad master equation (seemain text) to the steady state (independent of initialconditions) using a fourth-order Runge-Kutta algorithm.

0.1 0.110

0.5

1

0.09 0.1 0.110

0.5

1

mechanical coupling

a

corr

elat

ions

mechanical coupling

corr

elat

ions

b

open boundary

periodic

1

0

0.5

Figure S3. Langevin dynamics on finite square lattices, con-firming results shown in Fig. 4: (a) Correlations C(d) as afunction of coupling strength for different lattice sizes N ×Nwith N = 5, 10, 20, 30 (d = 2, 4, 9, 14, periodic boundaryconditions). (b) Correlations C(d = 14) on a 30 × 30 squarelattice with periodic boundary conditions (empty squares)and open boundary conditions (filled squares). The insetshows the correlations |〈eiϕie−iϕj 〉| for i = (8, 15) (square)and K = 0.115 Ω and open boundary conditions. The crossmarks the lattice site j = (22, 15) for which the correlationsare shown in the main plot. Other parameters as in Fig. 4.

Page 9: Quantum Many-Body Dynamics in Optomechanical Arraysthp2.nat.fau.de/images/5/56/2013_Ludwig_QuantumManyBodyDynamics.pdf · Quantum Many-Body Dynamics in Optomechanical Arrays Max Ludwig1,*

4

The size of the Hilbert space is optimized to enable an ad-equate representation of the physical state while obtain-ing reasonable simulation times. Typically, 10− 20 pho-ton and phonon levels, respectively, were taken into ac-count. When slowly sweeping through parameter space,bistable behavior is revealed. This can be seen mostprominently in Fig. S1(c) from the cut line in the rightpart of the plot.

The numerical integration of the Langevin equations(5) was obtained using a fourth order Runge-Kuttamethod for stochastic differential equations [26]. Whenevaluating the dependence of the threshold value on thequantum parameter g0/κ (Fig. 4), the value of g0αL washeld constant. In this case, only the strength of quan-tum fluctuations changes (keeping the classcial solutionconstant). For very large values of g0/κ, fluctuations areoverestimated by the Langevin equations and one has torely on the exact quantum simulations.

The finite element simulations of Figs.1 and S2 wereperformed using COMSOL Multiphysics. We studied ahexagonal silicon slab with sides of length 5.6µm andrestricted our simulations to two space dimensions, i.e.in-plane elastic deformations and electromagnetic waves,see also [20]. Extended simulations in three dimensionshave to be employed to analyze the coupling of the cavitymodes to the out-of plane-modes.

[email protected][1] G. Heinrich, M. Ludwig, J. Qian, B. Kubala, and F. Mar-

quardt, Phys. Rev. Lett. 107, 043603 (2011).[2] D. A. Rodrigues and A. D. Armour, Phys. Rev. Lett.

104, 053601 (2010).[3] Y. Kuramoto, Progress of Theoretical Physics Supple-

ment 79, 223 (1984).[4] H. Sakaguchi, Progress of Theoretical Physics 79, 39

(1988).[5] S. H. Strogatz and R. E. Mirollo, Journal of Statistical

Physics 63, 613 (1991).[6] P. E. Barclay, K. Srinivasan, O. Painter, B. Lev, and

H. Mabuchi, Applied Physics Letters 89, 131108 (2006).

[7] L. Ding, C. Baker, P. Senellart, A. Lemaitre, S. Ducci,G. Leo, and I. Favero, Phys. Rev. Lett. 105, 263903(2010).

[8] M. Zhang, G. S. Wiederhecker, S. Manipatruni,A. Barnard, P. McEuen, and M. Lipson, Phys. Rev. Lett.109, 233906 (2012).

[9] D. K. Armani, T. J. Kippenberg, S. M. Spillane, andK. J. Vahala, Nature 421, 925 (2003).

[10] A. M. Armani, A. Srinivasan, and K. J. Vahala, NanoLetters 7, 1823 (2007).

[11] E. Verhagen, S. Deleglise, S. Weis, A. Schliesser, and T. J.Kippenberg, Nature 482, 63 (2012).

[12] C. A. Regal, J. D. Teufel, and K. W. Lehnert, NaturePhys. 4, 555 (2008).

[13] T. Rocheleau, T. Ndukum, C. Macklin, J. B. Hertzberg,A. A. Clerk, and K. C. Schwab, Nature 463, 72 (2010).

[14] J. D. Teufel, T. Donner, Dale Li, J. W. Harlow, M. S.

Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, K. W.Lehnert, and R. W. Simmonds, Nature 475, 359 (2011).

[15] A. A. Houck, H. E. Tureci, and J. Koch. Nature Phys. 8,292 (2012).

[16] A. Naik, O. Buu, M. D. LaHaye, A. D. Armour, A. A.Clerk, M. P. Blencowe, and K. C. Schwab, Nature 443,193 (2006).

[17] J. Chan, T. P. Mayer Alegre, A. H. Safavi-Naeini, J. T.Hill, A. Krause, S. Groblacher, M. Aspelmeyer, andO. Painter, Nature, 478, 89 (2011).

[18] A. H. Safavi-Naeini and O. Painter, Opt. Express 18,14926 (2010).

[19] E. Gavartin, R. Braive, I. Sagnes, O. Arcizet, A. Bever-atos, T. J. Kippenberg, and I. Robert-Philip, Phys. Rev.Lett. 106, 203902 (2011).

[20] A. H Safavi-Naeini and O. Painter, New J. Phys. 13,013017 (2011).

[21] M. L. Roukes and E. Buks, Journal of Microelectrome-chanical Systems 11, 6 (2002).

[22] C. P. Michael, M. Borselli, T. J. Johnson, C. Chrystal,and O. Painter, Opt. Express 15, 4745 (2007).

[23] A. H. Safavi-Naeini, S. Groeblacher, J. T. Hill, J. Chan,M. Aspelmeyer, and O. Painter, arxiv:1302.6179.

[24] J. D. Cohen, S. Meenehan, and O. Painter, Opt. Express21, 11227 (2013).

[25] J. Chan, PhD thesis, California Institute of Technology,2012.

[26] N.J. Kasdin, Proceedings of the IEEE, 83, 802 (1995).