entangled-state generation and bell inequality violations in nanomechanical resonators

11
Entangled-state generation and Bell inequality violations in nanomechanical resonators J. Robert Johansson, 1 Neill Lambert, 2 Imran Mahboob, 3 Hiroshi Yamaguchi, 3 and Franco Nori 2, 4 1 iTHES Research Group, RIKEN, Saitama 351-0198, Japan 2 CEMS, RIKEN, Saitama 351-0198, Japan 3 NTT Basic Research Laboratories, Nippon Telegraph and Telephone Corporation, Atsugi 243-0198, Japan 4 Physics Department, University of Michigan, Ann Arbor, Michigan, 48109, USA (Dated: March 4, 2014) We investigate theoretically the conditions under which a multi-mode nanomechanical resonator, operated as a purely mechanical parametric oscillator, can be driven into highly nonclassical states. We find that when the device can be cooled to near its ground state, and certain mode matching conditions are satisfied, it is possible to prepare distinct resonator modes in quantum entangled states that violate Bell inequalities with homodyne quadrature measurements. We analyze the parameter regimes for such Bell inequality violations, and while experimentally challenging, we believe that the realization of such states lies within reach. This is a re-imagining of a quintessential quantum optics experiment by using phonons that represent tangible mechanical vibrations. I. INTRODUCTION Reaching the quantum regime with mechanical res- onators have been a long-standing goal in the field of nanomechanics 1–4 . In recent experiments, such devices have been successfully cooled down to near their quan- tum ground states 5–7 , and in the future may be used for quantum metrology 8 , as quantum transducers and cou- plers between hybrid quantum systems 9–13 , for quantum information processing 14 , and for exploring the limits of quantum mechanics with macroscopic objects. In many of these applications it is essential to both prepare the nanomechanical system in highly nonclassical states and to unambiguously demonstrate the quantum nature of the produced states. Nonclassical states of harmonic resonators can be achieved by introducing time-dependent parametric modulation 15 or via nonlinearities. The latter can be realized by a variety of techniques, for example by cou- pling to a superconducting qubit 5 , coupling to additional optical cavity modes 16–18 , applying external nonlinear potentials 14 , or via intrinsic mechanical nonlinearities in the resonator itself 19–22 . Using such nonlinearites, spe- cific modes of a nanomechanical resonator could poten- tially be prepared in a rich variety of different nonclas- sical states, such as quadrature squeezed states 16,23–27 , subpoissonian phonon distributions 28–30 , Fock states 31 , and quantum superposition states 5,32,33 . Quantum cor- relations and entanglement between states of distinct os- cillator modes could also be potentially generated, typi- cally taking the form of entangled phonon states and two- mode quadrature correlations and squeezing 34–37 . Ex- perimentally, nonlinear interactions between modes of nanomechanical resonators have already been used 38,39 for parametric amplification and noise squeezing. Vari- ous schemes 31,40,41 have proposed using nonlinear dissi- pation processes to realizing steady state entanglement. In a similar direction, a recent proposal 42 looked at ways to couple different internal mechanical modes of a nanomechanical system via ancillary optical cavities. Also, Rips et al. [31] looked at ways to prepare nonclas- sical states using enhanced intrinsic mechanical nonlin- earities. Here we consider the generation of nonclassical states and the subsequent violation of Bell inequalities by the use of similar intrinsic mechanical nonlinearities 14,20–22 . We focus on a model relevant to a recent experimental realization 43 of a phonon laser, where a single mechani- cal device exhibits significant coupling between three in- ternal modes of deformation, due to asymmetries in the beam 22 , and selective activation using external driving. Here we examine that same intrinsic inter-mode inter- action in the quantum limit. A schematic illustration of the device considered here is shown in Fig. 1, though this is not intended to be representative of the ideal realiza- tion or measurement scheme for operating in the quan- tum limit. In most of our discussion we do not con- sider an explicit physical setup but rather focus on set- FIG. 1: (Color online) Schematic illustration of a conceptual nanomechanical resonator with two homodyne measurement setups that probe two different modes of oscillation. The beam can oscillate in a large number of vibrational and flex- ural modes with different frequencies. The two homodyne detectors measure the mode quadratures X1(φ) and X2(θ) with frequencies ω1/2π and ω2/2π, respectively. By analyzing the correlations between the X1(φ) and X2(θ) quadratures, it is possible to determine whether or not the mode states are quantum-mechanically entangled. arXiv:1402.4900v1 [quant-ph] 20 Feb 2014

Upload: independent

Post on 16-Nov-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

Entangled-state generation and Bell inequality violations in nanomechanicalresonators

J. Robert Johansson,1 Neill Lambert,2 Imran Mahboob,3 Hiroshi Yamaguchi,3 and Franco Nori2, 4

1iTHES Research Group, RIKEN, Saitama 351-0198, Japan2CEMS, RIKEN, Saitama 351-0198, Japan

3NTT Basic Research Laboratories, Nippon Telegraph and Telephone Corporation, Atsugi 243-0198, Japan4Physics Department, University of Michigan, Ann Arbor, Michigan, 48109, USA

(Dated: March 4, 2014)

We investigate theoretically the conditions under which a multi-mode nanomechanical resonator,operated as a purely mechanical parametric oscillator, can be driven into highly nonclassical states.We find that when the device can be cooled to near its ground state, and certain mode matchingconditions are satisfied, it is possible to prepare distinct resonator modes in quantum entangled statesthat violate Bell inequalities with homodyne quadrature measurements. We analyze the parameterregimes for such Bell inequality violations, and while experimentally challenging, we believe thatthe realization of such states lies within reach. This is a re-imagining of a quintessential quantumoptics experiment by using phonons that represent tangible mechanical vibrations.

I. INTRODUCTION

Reaching the quantum regime with mechanical res-onators have been a long-standing goal in the field ofnanomechanics1–4. In recent experiments, such deviceshave been successfully cooled down to near their quan-tum ground states5–7, and in the future may be used forquantum metrology8, as quantum transducers and cou-plers between hybrid quantum systems9–13, for quantuminformation processing14, and for exploring the limits ofquantum mechanics with macroscopic objects. In manyof these applications it is essential to both prepare thenanomechanical system in highly nonclassical states andto unambiguously demonstrate the quantum nature ofthe produced states.

Nonclassical states of harmonic resonators can beachieved by introducing time-dependent parametricmodulation15 or via nonlinearities. The latter can berealized by a variety of techniques, for example by cou-pling to a superconducting qubit5, coupling to additionaloptical cavity modes16–18, applying external nonlinearpotentials14, or via intrinsic mechanical nonlinearities inthe resonator itself19–22. Using such nonlinearites, spe-cific modes of a nanomechanical resonator could poten-tially be prepared in a rich variety of different nonclas-sical states, such as quadrature squeezed states16,23–27,subpoissonian phonon distributions28–30, Fock states31,and quantum superposition states5,32,33. Quantum cor-relations and entanglement between states of distinct os-cillator modes could also be potentially generated, typi-cally taking the form of entangled phonon states and two-mode quadrature correlations and squeezing34–37. Ex-perimentally, nonlinear interactions between modes ofnanomechanical resonators have already been used38,39

for parametric amplification and noise squeezing. Vari-ous schemes31,40,41 have proposed using nonlinear dissi-pation processes to realizing steady state entanglement.In a similar direction, a recent proposal42 looked atways to couple different internal mechanical modes of

a nanomechanical system via ancillary optical cavities.Also, Rips et al. [31] looked at ways to prepare nonclas-sical states using enhanced intrinsic mechanical nonlin-earities.

Here we consider the generation of nonclassical statesand the subsequent violation of Bell inequalities by theuse of similar intrinsic mechanical nonlinearities14,20–22.We focus on a model relevant to a recent experimentalrealization43 of a phonon laser, where a single mechani-cal device exhibits significant coupling between three in-ternal modes of deformation, due to asymmetries in thebeam22, and selective activation using external driving.Here we examine that same intrinsic inter-mode inter-action in the quantum limit. A schematic illustration ofthe device considered here is shown in Fig. 1, though thisis not intended to be representative of the ideal realiza-tion or measurement scheme for operating in the quan-tum limit. In most of our discussion we do not con-sider an explicit physical setup but rather focus on set-

FIG. 1: (Color online) Schematic illustration of a conceptualnanomechanical resonator with two homodyne measurementsetups that probe two different modes of oscillation. Thebeam can oscillate in a large number of vibrational and flex-ural modes with different frequencies. The two homodynedetectors measure the mode quadratures X1(φ) and X2(θ)with frequencies ω1/2π and ω2/2π, respectively. By analyzingthe correlations between the X1(φ) and X2(θ) quadratures, itis possible to determine whether or not the mode states arequantum-mechanically entangled.

arX

iv:1

402.

4900

v1 [

quan

t-ph

] 2

0 Fe

b 20

14

2

ting bounds on the fundamental system parameters nec-essary to realize the phenomena we discuss. The modelwe derive consists of an adiabatically-eliminated pumpmode which drives the interaction between two lower-frequency signal and idler modes. We show that in thetransient regime one can obtain violations of a Bell in-equality based on correlations between quadrature mea-surements of the signal and idler modes. This is a re-imagining of a quintessential quantum optics experimentby using phonons that represent tangible mechanical vi-brations.

This paper is organized as follows: In Sec. II we intro-duce the general model and the Hamiltonian for a non-linear nanomechanical device. In Sec. II A we consider aregime in which a parametric oscillator is realized usingthree modes in the mechanical system, and in Sec. II Bwe introduce an effective two-mode model, valid when thepump mode can be adiabatically eliminated, and we ana-lyze the types of nonclassical states that can be generatedin this system. In Sec. III, we review Bell’s inequality us-ing quadrature measurements, and in Sec. IV we analyzethe conditions for realizing a violation of this quadrature-based Bell inequality with the mechanical system in theparametric oscillator regime studied in Sec. II B. Finallywe discuss the outlook for an experimental implementa-tion using either intrinsic nonlinearities in Sec. V, or, asan alternative, optomechanical nonlinearities in Sec. V A.We summarize our results in Sec. VII.

II. MODEL

The general Hamiltonian for a nonlinear multimoderesonator, describing both the self-nonlinearities andmultimode couplings up to fourth-order, can be writtenas21

H =∑k

ωka†kak +

∑klm

βklmxkxlxm

+∑klmn

ηklmnxkxlxmxn +O(x5), (1)

where ωk is the frequency, ak is the annihilation oper-

ator, and xk = ak + a†k is the quadrature of the me-chanical mode k. Here the basis has been chosen sothat linear two-mode coupling terms are eliminated. Thethird-order mode-coupling tensor βklm describes the odd-term self-nonlinearity and the trilinear multimode inter-action. The fourth-order terms describes the even-termself-nonlinearity and fourth-order multimode coupling.In symmetric systems the fourth-order terms dominate(odd terms vanish due to symmetry), and it has been pro-posed elsewhere that they can be used to create effectivemechanical qubits14. The possible combination of boththird and fourth-order terms will be briefly consideredin the final section. The strength of the nonlinearity de-pends on the fundamental frequency (length) of the res-onator, and can be enhanced by a range of techniques14.

0

ω1 ω2 ω0

ω1 + ω2ωL

∆L ∆0

FIG. 2: A visualization of the mode-matching condition re-quired to obtain an effective three-mode system. The modesare represented as solid vertical lines, and the driving fre-quency and the sum of the signal and idler frequencies, whichshould sum up to a frequency close to ω0, are represented bydashed vertical lines.

In this work we focus on the three-mode coupling terms,as these are necessary to generate the states that violatecontinous variable Bell inequalities. Such terms vanishin symmetric systems and thus depend on the degree ofasymmetry in the mechanical device21,22, which againcan be enhanced with fabrication techniques. Our ap-proach in the following is to identify the ideal situationunder which one can realize these rare Bell inequality vi-olating states. Ultimately these states will be degradedby losses (which we investigate), but also by unwantednonlinearities from the above hamiltonian.

A. Parametric oscillator regime

Nanomechanical devices of the type described in theprevious section have a large number of modes with dif-ferent frequencies which depend on the microscopic struc-tural properties of the beam. Here we focus on threesuch modes (relabelled as k = 0, 1, 2) which are cho-sen such that they satisfy the phase-matching conditionω1 + ω2 = ω0 + ∆0, where ∆0 � ω0. In this case we canperform a rotating-wave approximation to single out theslowly-oscillating coupling terms, and obtain the desiredeffective three-mode system, neglecting any higher-ordernon-linearities. In the original frame, the Hamiltonianwith this rotating-wave approximation is

H =

2∑k=0

ωka†kak + iκ(a†1a

†2a0 − a1a2a

†0), (2)

where a1 and a2 are the signal and idler modes, respec-tively, and a0 is the pump mode. Furthermore, we applya driving force that is nearly resonant with ω0, with fre-quency ωL = ω0 − ∆L, |∆L| � ω0, and transform theabove Hamiltonian to the rotating frame where the res-onant drive terms are time-independent,

H = ∆La†0a0 +

∑k=1,2

∆ka†kak

+ iκ(a†1a†2a0 − a1a2a

†0)− i(Ea†0 − E∗a0). (3)

Here ∆1 = ∆2 = (∆0−∆L)/2, κ = β012 is the inter-modeinteraction strength, and E is the driving amplitude of

3

mode a0. See Fig. II A for a visual representation of themode-matching condition and the detuning parameters∆0 and ∆L.

This is an all-mechanical realization of the generalthree-mode parametric oscillator model in nonlinearoptics44,45, where mode a0 is the quantized pump mode,and modes a1 and a2 are the signal and idler modes,respectively.

B. Effective two-mode model

We assume that in this purely nanomechanical real-ization of the parametric oscillator model, Eq. (3), allthree mechanical modes interact with independent envi-ronments. We describe these processes with a standardLindblad master equation on the form

ρ = −i[H, ρ] +∑k

γk

{(Nk + 1)D[ak] +NkD[a†k]

}ρ (4)

where D[ak]ρ = akρa†k −

12a†kakρ −

12ρa†kak is the dis-

sipator of mode ak, γk is the corresponding dissipa-tion rate, and the average thermal occupation numberis Nk = (exp(~ωkβ)− 1)

−1. Here β = 1/kBT is the in-

verse temperature T , and kB is Boltzmann’s constant.Assuming that the pump mode is strongly damped

compared to the signal and idler modes, γ0 � γ1, γ2,and that the pump-mode dissipation dominates over the

coherent interaction, γ0 � 〈H〉 ∼ κ〈a†0a1a2〉, one canadiabatically eliminate45,47,48 the pump mode from themaster equation given above. Here we also assume thatthe high-frequency pump mode is at zero temperature,N0 = 0, while the temperatures of modes a1 and a2 canremain finite. This results in a two-mode master equationthat includes correlated two-phonon dissipation, whereone phonon from each mode dissipates to the environ-ment through the pump mode, in addition to the originalsingle-phonon losses in each mode:

ρ = − i[H ′, ρ] + γD[a1a2]ρ

+∑k=1,2

γk

{(Nk + 1)D[ak] +NkD[a†k]

}ρ, (5)

where the effective two-phonon dissipation rate is

γ =κ2γ0/2

|γ0/2 + i∆L|2. (6)

The reduced two-mode Hamiltonian is given by

H ′ =∑k=1,2

∆ka†kak + i(µa†1a

†2 − µ∗a1a2) + χa†1a1a

†2a2,

(7)

with the two-mode interaction strength

µ =Eκ

γ0/2 + i∆L, (8)

(a) (b)

(c)

FIG. 3: (Color online) Visualization of the steady state ofthe effective two-mode system. (a) The Fock state distribu-tion of mode a1 and a2. (b) The single-mode Wigner functionof mode a1 and a2. Both the Fock state distribution and theWigner function are identical for both modes a1 and a2, dueto the symmetric two-phonon processes and equal dissipationrates and initial states. The single-mode Wigner function ispositive, and the single-mode state can therefore be consid-ered classical. However, strong correlations exists betweenquadratures of two different modes, as shown in the jointquadrature probability distribution PX1,X2(0, 0) in (c). Herewe have used the parameters κ = 0.15, E = 0.094, γ0 = 1.0,γ1 = γ2 = 0, and ∆0 = ∆L = 0. These parameters werechosen to produce a steady state that closely corresponds tothe ideal state46 for quadrature Bell inequality violation, i.e.with r = 1.12 (see Sec. IV.A).

and the effective cross-Kerr interaction strength

χ = − κ2∆L

|γ0/2 + i∆L|2, (9)

which vanishes when the driving field is at exact reso-nance with the pump mode. In the following we willgenerally assume that this resonance condition can bereached, and ∆L will be set to zero in the equationsabove.

In this resonant limit the Hamiltonian H ′ describesan ideal two-mode parametric amplifier, which is well-known to be the generator of two-mode squeezed states49.When applied to the vacuum state, or a low-temperaturethermal state, the resulting two-modes squeezed statesare nonclassically correlated50, but when viewed individ-ually, both modes appear to be in thermal states. In

4

0

1P

hoto

n n

um

ber

⟨n1

⟩⟨n2

0

1

2

Sin

gle

-mod

eva

rian

ces

Var(x1 )

Var(p1 )

1

3

5

Tw

o-m

od

eva

rian

ces

Var(x1−x2 )

Var(x1−p2 )

Var(p1−p2 )

0 1 2 3 4 5 6

time tγ0/2

0

1

Log

. n

eg

ati

vity

(a)

(b)

(c)

(d)

FIG. 4: (Color online) The time evolution of the phononnumber (a), single-mode quadrature variances (b), and thevariances of the two-mode quadrature differences (c), the log-arithmic negativity (d), for the case when the state is ini-tially in the zero-temperature ground state. In the large-time limit the state approaches the steady state that is vi-sualized in Fig. 3. The single-mode variances increase abovethe vacuum limit as time increases, but the variance of cross-quadrature difference Var(x1 − x2) decrease below the vac-uum limit, which is a characteristic of two-mode squeezing,and the nonzero logarithmic negativity demonstrates that thesteadystate is nonclassical. Here we used the same parametersas those given in the caption of Fig. 3.

spite of being quantum mechanically entangled, thesetwo-mode squeezed states have a positive Wigner func-tion and cannot violate the quadrature binning Bellinequalities51 that we consider below. One must con-sider the effect of the two-phonon dissipation in Eq. [5]to induce such violations.

In the highly idealized case when single phonon dissipa-tion in the a1 and a2 modes is absent, i.e., γ1 = γ2 = 0,but with γ0 > 0, the model Eq. (5) produces a steadystate47 of the form

ρ =1

I0(2r2)

∑m,n

r2m+2n

m!n!|m,m〉〈n, n|, (10)

where I0 is the zeroth order modified Bessel function andr =

√2E/κ. The special structure of this steady state,

with equal number of phonons in each mode, is becauseboth the Hamiltonian and two-phonon dissipator con-

serve the phonon-number difference a†2a2 − a†1a1. How-

ever, this symmetry is broken if the single-phonon dissi-pation processes are included in the model, i.e. γ1, γ2 >0. The state Eq. (10) is visualized in Fig. 3, for the spe-cific set of parameters given in the figure caption. Figure3(a-b) show the Fock-state distribution and the Wignerfunction for the modes a1 and a2 (because of symmetrythe states of both modes are identical in this case, andonly one is shown). In this case the states of the twomodes no longer appear to be thermal when viewed indi-vidually, but the reduced single-mode Wigner functionsare positive and thus, on their own, each mode appearsclassical. However, together, the two-mode Wigner func-tion can be negative. For example, there is a strong cross-quadrature correlation, as shown in Fig. 3(c). The vari-ances of the cross-quadrature differences, in the transientapproach to the steady state, are shown in Fig. 4, andexactly in the steady state the variance of the squeezedtwo-mode operator difference is

Var(x1 − x2) = 1 + 2r2I1(2r2)

I0(2r2)− 2r2, (11)

which in the limit of large r approaches 1/2, but hasa local minimum of about 0.4 at r ≈ 0.92. We notethat for the vacuum state Var(x1 − x2) = 1, and thusthis quadrature difference variance is therefore squeezedbelow the vacuum level for any r > 0. The logarithmicnegativity52 shown in Fig. 4(d) further demonstrates thenonclassical nature of this state.

These intermode quadrature correlations, with squeez-ing below the vacuum level of fluctuations, are non-classical and it has been shown that this particularstate can violate Bell inequalities based on quadraturemeasurements51, as we will discuss in the next section. Infact, this steady state is, for a certain value of r, a goodapproximation to the ideal two-mode quantum state46

for these kind of Bell inequalities. However, it has alsobeen shown that in the presence of single phonon dissipa-tion the steady state two-mode Wigner function is alwayspositive, and thus exhibits a hidden-variables descriptionand cannot violate any Bell inequalities44. Fortunately,this is only the case for the steady state, and there canbe a significant transient period in which the two-modesystem is in a state that can give a violation.

In the following we consider two regimes; the steadystate, and the slow transient dynamics of a system thatis originally in the ground state, and approaches the newsteady state after the driving field has been turned on.

III. BELL INEQUALITIES FORNANOMECHANICAL SYSTEMS

Verifying that a nanomechanical system is in the quan-tum regime, and that the states produced in the sys-tem are nonclassical, can be sometimes be experimen-

5

tally challenging, largely because of the difficulty in im-plementing single-phonon detectors in nanomechanicalsystems. As has been done in circuit QED53,54, measur-ing a nonlinear energy spectrum14 would be a convincingindication that the system is operating in the quantumregime, although it does not imply that the state of thesystem is nonclassical, and all quantum nanomechanicalsystems need not necessarily be nonlinear. A number oftechniques could be used to demonstrate that the stateis nonclassical55, for example reconstructing the Wignerfunction using state tomography and looking for negativevalues, or evaluating entanglement measures such as thelogarithmic negativity (for Gaussian states) or entangle-ment entropy (suitable only at zero temperature).

Here we are interested in a nonclassicality test thatcan be evaluated using joint two-mode quadrature mea-surements. The two-mode squeezing shown in the pre-vious section can be considered as an entanglementwitness56,57, and was recently investigated experimen-tally in an opto-mechanical device25,26. The quadrature-based Bell inequality can be seen as another, stricter,example of a nonclassicality test, and in the following wefocus on the possibility of violating these Bell inequalitieswith the nanomechanical system outlined in the previoussection. Even though one cannot rule out the locality-loophole in such a system, and thus a violation wouldlack any meaning as a strict test of Bell nonlocality58, itwould still serve as a very satisfying test for two-modeentanglement.

The original Bell inequalities are formulated for di-chotomic measurements, with two possible outcomes.However, dichotomic measurements are not normallyavailable in harmonic systems like the nanomechanicalsystems considered here, where the measurement out-comes are, for the most part, continuous and unbound.In this continuous-variable limit one must choose howto perform a Bell inequality test with care. General-ized inequalities for unbound measurements exist59, butare both extremely challenging to implement and hardto violate. Fortunately one can implement CHSH-typeBell inequality by binning quadrature measurements, andthus obtaining a dichotomic bound observable. Munro46

showed that, while in general it is hard to generate stateswhich can violate such an inequality, it is possible to gen-erate precisely the type of states which do cause a vio-lation with a nondegenerate parametric oscillator, whichis analogous to the system we investigate here.

One possible binning strategy51,60 for the continuousoutcomes of quadrature measurements of the mechanicalmodes is to classify the outcomes as 1 if the measurementoutcome is Xθ > 0, and 0 otherwise. The probability ofthe outcomes 0 and 1 for the two modes can then bewritten

Pαβ(θ, φ) =

∫ U(α)

L(α)

∫ U(β)

L(β)

d2Xp(Xθ1 , X

φ2 )[ρ] (12)

where

L(α) =

{0 if α = 1

−∞ if α = 0, U(α) =

{∞ if α = 1

0 if α = 0. (13)

Here ρ is the two-mode density matrix and p(Xθ1 , X

φ2 )[ρ]

is the probability distribution for obtaining the measure-

ment outcomes Xθ1 and Xφ

2 for the signal and idler modequadratures

xθ1 = a1e−iθ + a†1e

iθ, (14)

xφ2 = a2e−iφ + a†2e

iφ, (15)

respectively. This probability distribution is given by

p(Xθ1 , X

φ2 )[ρ] =

⟨Xθ

1 , Xφ2

∣∣∣ ρ ∣∣∣Xθ1 , X

φ2

⟩=

∑m,n,p,q

ρ(m,n),(p,q)e−i(mθ+nφ)ei(pθ+qφ)

π√

2m+n+p+qm!n!p!q!

×e−X21 e−X

22Hm(X1)Hn(X2)Hp(X1)Hq(X2)

(16)

where Hn(x) is the Hermite polynomial of nth order, andwhere we have written the density matrix in the two-mode Fock basis,

ρ =∑

m,n,p,q

ρ(m,n),(p,q) |m,n〉〈p, q| . (17)

The integral in Eq. (12) can be evaluated analytically46,but in general the sum in Eq. (16) cannot.

Treating the binned quadrature measurements as di-chotomic observables we can write the standard Bell’sinequalities in the Clauser-Horne (CH)61 form

BCH =P11(θ, φ)− P11(θ, φ′) + P11(θ′, φ) + P11(θ′, φ′)

P1(θ′) + P1(φ)(18)

which for a classical state satisfies |BCH| ≤ 1, and in theClauser-Horne-Shimony-Holt (CHSH)62 form

BCHSH = E(θ, φ)− E(θ′, φ) + E(θ, φ′) + E(θ′, φ′),(19)

E(θ, φ) = P11(θ, φ) + P00(θ, φ)

− P10(θ, φ)− P01(θ, φ), (20)

which for a classical state satisfies |BCHSH| ≤ 2. Here wehave also used

P1(θ) =

∫ ∞0

∫ ∞−∞

d2Xp(Xθ1 , X

φ2 )[ρ]. (21)

Both BCH and BCHSH are in general functions of thefour angles θ, φ, θ′, and φ′. However, to reduce the num-ber of parameters here we consider the angle parameter-ization θ = −2ϕ, φ = 3ϕ, θ′ = 0, and φ′ = ϕ, which onlyleaves a single free angle parameter ϕ. In principle thiscan reduce the magnitude of violation one can observe,but as we will see this parameterization still allows viola-tions to occur for the types of states we are interested inhere. In the following we evaluate both BCH and BCHSH

using this angle parameterization.

6

IV. VIOLATION OF BELL’S INEQUALITYWITH NANOMECHANICAL RESONATORS

In this section we investigate the conditions underwhich the states formed in the multimode nanomechan-ical system may violate Bell’s inequality. We empha-size again that in this context we are interested in Bell’sinequality as a test that can demonstrate entanglementbetween different mechanical modes. We begin with ananalysis of the steadystate for the idealized model withγ1, γ2 = 0, and then turn our attention to the transientbehaviour for finite γ1 and γ2.

A. Steady state

With γ1, γ2 = 0, the steady state is given by Eq. (10),and inserting this state in the Bell inequalities Eqs. (18-19) gives an expression as a function of the steady stateparameter r and the angle ϕ that can be optimized formaximum Bell violation. The optimal value of the angleturns out to be ϕ = π/4, and the resulting equation foroptimal r is

I0(2r2)dG(r)

dr= 4r2I1(2r2)G(r), (22)

but the sum over Fock-state basis that comes fromEq. (16) cannot to our knowledge be evaluated in a sim-ple analytical form, so we have

G(r) =∑n

∑m>n

8(2r2)n+mπ

(n!m!)2(n−m)2[F(n,m)−F(m,n)]

2 ×

{3 cos [(n−m)ϕ]− cos [3ϕ(n−m)]} , (23)

and

F(n,m) =

(1

2− n

2

)Γ(−m

2

)]−1, (24)

as given in Ref. 46. Solving Eq. (22) numerically givesropt ≈ 1.12, as reported in Ref. 51. The correspondingsteady state Eq. (10) for ropt is visualized in Fig. 3. Wenote that for this optimal Bell violating state the meanphonon number in each mode is only 〈n〉 ≈ 0.94, whichhighlights the need to operate the system near its groundstate. If fact, when 〈n〉 � 1 no Bell inequality violationcan be observed.

In our nanomechnical model this translates to an op-timal driving strength, Eopt = r2optκ/2, that maximizesthe Bell inequality violation for a given nonlinearity κ.This optimal driving amplitude Eopt applies to the steadystate of the idealized model without single-phonon dissi-pation. With finite single-phonon dissipation, the steadystate does not violate any of the Bell inequalities. How-ever, as we will see in the following section, Eopt still givesa good approximation for the optimal transient violation.While these transients are harder to capture, recent ex-periments on opto-mechanical systems have shown they

0 1 2 3 4 5 6time t 0 / 2

0.9

1.0

1.1

BCH

BCHSH/2

(a)

0 /4 /2 3 /4Bell angle

0.6

0.8

1.0

BCH

BCHSH/2

(b)

FIG. 5: (Color online) (a) The normalized CH and CHSHBell quantities (violation above 1) as a function of time,for the pumped two-mode nanomechanical resonator, underthe ideal condition without signal and idler mode dissipation(dashed) and for the case including signal and idler dissipa-tion. The initial state is the vacuum state, which we assumecan be prepared to a good approximation using cooling. Att = 0, the parametric amplification is turned on by the ac-tivation of the driving field with amplitude E. In the idealcase, the steady state violates both the CH and CHSH Bell in-equalities, but there is no steady state violation when single-phonon dissipation is included. However, there is a periodof time during the transient where both inequalities are vi-olated. (b) The angle ϕ dependence for the normalized CHand CHSH Bell quantities for tγ0/κ

2 = 1.8, where solid linesinclude dissipation and dashed lines are the ideal case. Theoptimal value of ϕ for the states produced in the model weinvestigate here is ϕ = π/4, which was used in (a). The pa-rameters used here are the same as in Fig. 3, and for the solidlines we used γ1 = γ2 = 0.001.

are in principle possible25,26, and relevant for the alterna-tive proposal in the final section below. How far one cango with using multiple ancilla optical or microwave cavi-ties to perform similar measurements on different internalmodes of a single mechanical device is not yet clear.

7

(a) (b) (c)

(d) (e) (f)

FIG. 6: (Color online) Violation of the normalized quadrature CHSH Bell inequality (redish region) as a function of time tand the inter-mode coupling κ (a,d), the pump mode driving amplitude E (b,e), and the pump-mode dissipation rate γ0 (c,f).The ideal case without signal and idler mode dissipation is shown in (a-c), and (d-e) include signal and idler mode dissipationwith equal dissipation rates γ1 = γ2 = 0.001. In (a-c) there is a parameter window for κ and E which results in a violationfor sufficiently large t, as well as in the steady state. However, in (d-e) there is no violation in the steady state, but during atransient time a violation may still occur for suitably chosen parameters. Apart from the parameters on the vertical axes, theparameters were kept fixed at the same values as given in Fig. 3, and denoted by a bar over the symbol in the axes.

B. Transient

Since the more realistic model, with finite single-phonon dissipation processes, does not produce a steadystate that violates any of the Bell inequalities, we are leadto investigate transient dynamics. Here we focus on thetransient which occurs when the driving field E is turnedon after the relevant modes have been cooled to theirground states. The state of the system then evolves fromthe ground state to the steady state that does not vio-late the Bell inequalities. However, if the single phonondissipation processes are sufficiently slow there can bea significant time interval during which the state of thesystem does violate the Bell inequalities.

To investigate this transient dynamics we numericallyevolve the effective two-mode system described by themaster equation, Eq. (5), and evaluate the BCH andBCHSH quantities as a function of time and the angleϕ. The results shown in Fig. 5, for the situations withand without signal and idler mode dissipation and at zerotemperature, demonstrate that the nanomechanical sys-tem we consider can indeed be driven into a transientstate that violates both types of Bell inequalities. Withlosses the onset of violation is proportional to γ0/κ

2, andthe time at which the violation cease is proportional toγ−11 , γ−12 , so if

γ1, γ2 � κ2/γ0, (25)

we expect a significant period of time during the transientwhere the inequalities will be violated. We note that inFig. 5(a), the regions of violation for the CH and CHSHinequalities are identical, and this is, according to ourobservations, always the case for this model and angleparametrization. Because of this, in the following weonly show the results for the CHSH inequality.

To further explore the parameter space that can pro-duce a Bell-inequality violation we evolve the masterequation as a function of time and the parameters E,κ and γ0, for both the ideal case with dissipation-lesssignal and idler modes, γ1 = γ2 = 0, and for the caseincluding signal and idler mode dissipation, γ1, γ2 > 0.In these simulations the initial state is always the groundstate, and we take the temperature of the signal and idlermodes to be zero. The results are shown in Fig. 6(a-c)and (e-f), respectively. From Fig. 6 it is clear that forthe case γ1 = γ2 = 0, there exist optimal values of κ andE, given that other parameters are fixed, that producesteady states that maximally violates the Bell inequality(marked with dashed lines in the figures). However, im-portantly, we also note that the optimal values for κ andE for the steady state of the ideal model also give a goodindicator for the optimal regime for the Bell violation inthe transient of the case with finite single-phonon dissi-pation, when additionally taking into account the timescales for the transient given in Eq. (25).

When the signal and idler modes have finite temper-

8

FIG. 7: (Color online) The regions where a transient Bellinequality violation can be achieved, as a function of nonlin-earity κ and time t, and for different temperatures. The re-gion of violation at zero temperature is shown in redish color.The contours mark the regions of violation for finite signaland idler mode temperatures (assumed equal), labeled by theaverage number N of thermal phonons. We note that evena very small number of initial thermal phonons inhibits theBell inequality violation, which suggest that excellent ground-state cooling of both modes is a prerequisite to obtaining aviolation.

ature the region of Bell inequality violation is furtherreduced, as shown in Fig. 7. The detrimental effects ofthermal phonons are two-fold: It reduces the transienttime-interval during which a violation can be observed,and the nonlinear interaction strength required to be ableto see any violation at all increases. In fact, to observea transient Bell inequality violation, the average ther-mal occupation number must be very small: An averagethermal occupation number of even 0.1 phonon in thesignal and idler mode is sufficient to inhibit any Bell vio-lation with the system we have considered here. Excellentground state cooling is therefore a prerequisite to violat-ing a Bell inequality tests in a nanomechanical resonator.

V. EXPERIMENTAL OUTLOOK

As can be seen in Fig. 6 and Fig. 7, the violation of aBell inequality in the system we consider here requires, asexpected, a combination of low temperature, large non-linearity, and transient quadrature measurements. Theseconditions can all be rather challenging to satisfy in anexperimental system, but on the other hand they are ex-actly the type of conditions that one can expect wouldhave to be satisfied for realistic quantum mechanical ap-plications in these devices. The Bell inequality violationcan therefore be seen as a benchmark that indicates thatentangled quantum states can be generated and detectedwith high precision.

While one can imagine cryogenics and side-band cool-ing techniques can satisfy the first criteria5–7,63, the ul-timate upper limit of the strength of intrinsic nonlinear-ities in mechanical systems is not clear. In a recent ex-periment extremely large nonlinear intra-mode couplingwas observed in a carbon nanotube system when themodes had frequencies which were integer multiples ofeach other64. In that case a strong effective mode-modecoupling was also found, which is required for generat-ing the Bell inequality violating states we consider here.Nonlinear mode coupling has also been demonstrated andanalyzed in doubly-clamped beam resonators19–22, andcircular graphene membrane resonators65. Also, recentstudies14 have proposed enhancing the nonlinearity perphonon by reducing the fundamental frequency of themechanical oscillator, which essentially amounts to in-creasing the ground state displacement. Thus, there isprogress in realizing nonlinear mode interaction in severaltypes of nanomechanical systems, and sufficiently strongnonlinearities to produce Bell inequality violating statesshould be obtainable in these devices, although furtherprogress in this experimental work in this direction maybe required.

Performing transient quadrature measurements of se-lected modes of the mechanical resonator is an anotherexperimental challenge. However, the displacement of ananomechanical resonator can be converted to electricalsignals and measured for example using a range of dif-ferent techniques, for example piezoelectric schemes66,coupling to auxiliary optical modes13,67, or by capacitivecoupling to a microwave circuit8. In recent experiments,transient quadrature measurements of a nanomechanicalsystem were carried out with high precision and level ofcontrol25,26. Aslo, in microwave electronics, quadraturemeasurements in the quantum regime have been appliedto measure two-mode squeezing68,69, state tomography70,and entanglement71. Given a sufficiently efficient trans-ducer from mechanical displacement to electrical signals,the outlook for the required measurements for evaluatingthe quadrature Bell inequality is therefore encouraging.

A. Optomechanical realization

As an alternative to the purely mechanical scheme dis-cussed so far one could observe the similar quadrature-based Bell inequality violations in an optomechanicalsetup akin to that proposed in Refs. 17 and 72, where asingle mechanical mode is coupled to two optical cavities,e.g., in a membrane-in-the-middle geometry or within aphotonic crystal cavity. The most straightforward im-plementation would be to use the mechanical mode asthe pump mode which then acts to entangle the opticalmodes. The optical modes are coupled due to a photontunneling, and the resulting hybridized modes replace themechanical signal and idler modes a1 and a2 we discussedin this work. On resonance this again leads to the sameinteraction we use in Eq. [2]. The main motivation of

9

inducing this interaction in these earlier works was toengineer anharmonic energy levels. This anharmonicityallows specific transitions to be addressed with externallaser fields allowing one to use such devices as single-phonon/photon transistors and for non-demolition mea-surements of phonons or photons. In the limit where themechanical pump mode can be driven and adiabaticallyeliminated, one in principle observe Bell inequality vio-lations in the (hybridized) quadrature measurements ofthe two optical cavities.

VI. COMBINING EVEN AND ODDNONLINEARITIES: COUPLING MECHANICAL

QUBITS

In the previous calculations we have been exclusivelyconsidering the effect of odd nonlinearities which can onlyarise in asymmetric mechanical systems. In purely sym-metric devices the even order terms dominate, arguablythe most important of which is the x4 Duffing nonlinear-ity. Recent works14 have examined how this induces ananharmonic energy spectrum in the fundamental modeof a nanomechanical system, and outlined how this an-harmonic spectrum can be used as an effective qubitfor quantum computation. Naturally one can considerthe effect of both the third order coupling we have out-lined here, and the third and fourth order Duffing self-anharmonicity. Ultimately the relative strengths of thesedifferent terms depend strongly on the overlap betweenthe different mode shapes within the device, the geom-etry of the device, and the effect of various nonlinear-ity enhancing mechanisms. A naive investigation of thecontributions from these quartic terms suggest they onlywork to degrade the Bell inequality violation we discusshere. However, going beyond the regime we have outlinedthus far, one may note that, by changing the frequencyof the driving field in Eq. (1) one can get an excitation-preserving beam-splitter type of interaction between thesignal and idler modes.

Hcint = µ(a†1a2 + a†2a1). (26)

If this is combined with a sufficiently strong third orfourth-order self nonlinearity, such that the lowest ly-ing energy states of each mode can be considered as atwo-level system, one has a means to couple differentmechanical qubits in a single device. It may be possi-

ble to construct similar interactions with ancilla cavitiesand optomechanical interactions17,42. The original para-metric interaction described in Eq. (3) is not useful forthis purpose as it takes one out of a single excitationsubspace, as does the two-phonon dissipation.

VII. CONCLUSIONS

We have investigated a regime of a multimode nanome-chanical resonator, with intrinsic nonlinear mode cou-pling, in which three selected modes realize a paramet-ric oscillator. In the regime where the pump modeof the parametric oscillator can be adiabatically elimi-nated, we have investigated the generation of entangledstates between two distinct modes of oscillation in thenanomechanical resonator, and the possibility of detect-ing this entanglement using quadrature-based Bell in-equality tests. Our results demonstrate that while re-alistically it will not be possible to violate any Bell in-equality in the steady state, there can be a significant du-ration of time in which the transient evolution from theground state (prepared by cooling) to the steady statewhere the state of the system violates Bell inequalities.However, to achieve this transient violation requires a rel-atively large nonlinear mode coupling, excellent groundstate cooling, and fast and efficient quadrature measure-ments. These are, of course, very challenging experimen-tal requirements, but we believe that if a quadrature Bellinequality violation is realized experimentally it would bea very strong demonstration of quantum entanglement ina macroscopic mechanical system.

Acknowledgements

The numerical simulations were carried out usingQuTiP73,74, and the source code for the simulationsare available in Ref. [75]. We acknowledge W. Munroand T. Brandes for discussions and feedback. Thiswork was partly supported by the RIKEN iTHESProject, MURI Center for Dynamic Magneto-Optics,JSPS-RFBR No. 12-02-92100, Grant-in-Aid for ScientificResearch (S), MEXT Kakenhi on Quantum Cybernetics,the JSPS-FIRST program, and JSPS KAKENHI GrantNo. 23241046.

1 A. Cleland, Foundations of Nanomechanics, AdvancedTexts in Physics (Springer, 2002).

2 M. Blencowe, Physics Reports 395, 159 (2004).3 M. Poot and H. S. van der Zant, Physics Reports 511, 273

(2012).4 M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt,

arXiv:1303.0733 (2013).

5 A. D. O’Connell, M. Hofheinz, M. Ansmann, R. C. Bial-czak, M. Lenander, E. Lucero, M. Neeley, D. Sank,H. Wang, M. Weides, et al., Nature 464, 697 (2010).

6 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 475, 359 (2011).

7 J. Chan, T. P. M. Alegre, A. H. Safavi-Naeini, J. T. Hill,

10

A. Krause, S. Groblacher, M. Aspelmeyer, and O. Painter,Nature 478, 89 (2011).

8 C. A. Regal, J. D. Teufel, and K. W. Lehnert, Nat Phys4, 555 (2008).

9 C. P. Sun, L. F. Wei, Y.-x. Liu, and F. Nori, Phys. Rev. A73, 022318 (2006).

10 M. Wallquist, K. Hammerer, P. Rabl, M. Lukin, andP. Zoller, Physica Scripta 2009, 014001 (2009).

11 A. H. Safavi-Naeini and O. Painter, New Journal of Physics13, 013017 (2011).

12 Z.-L. Xiang, S. Ashhab, J. Q. You, and F. Nori, Rev. Mod.Phys. 85, 623 (2013).

13 J. Bochmann, A. Vainsencher, D. D. Awschalom, and A. N.Cleland, Nat Phys 9, 712 (2013).

14 S. Rips and M. J. Hartmann, Phys. Rev. Lett. 110, 120503(2013).

15 L. Tian, M. S. Allman, and R. W. Simmonds, New Journalof Physics 10, 115001 (2008).

16 K. Jahne, C. Genes, K. Hammerer, M. Wallquist, E. S.Polzik, and P. Zoller, Phys. Rev. A 79, 063819 (2009).

17 K. Stannigel, P. Rabl, A. S. Sørensen, M. D. Lukin, andP. Zoller, Phys. Rev. A 84, 042341 (2011).

18 Y.-D. Wang and A. A. Clerk, Phys. Rev. Lett. 110, 253601(2013).

19 H. J. R. Westra, M. Poot, H. S. J. van der Zant, and W. J.Venstra, Phys. Rev. Lett. 105, 117205 (2010).

20 K. J. Lulla, R. B. Cousins, A. Venkatesan, M. J. Patton,A. D. Armour, C. J. Mellor, and J. R. Owers-Bradley, NewJournal of Physics 14, 113040 (2012).

21 R. Khan, F. Massel, and T. T. Heikkila, Phys. Rev. B 87,235406 (2013).

22 H. Yamaguchi and I. Mahboob, New Journal of Physics15, 015023 (2013).

23 A. A. Clerk, F. Marquardt, and K. Jacobs, New Journalof Physics 10, 095010 (2008).

24 J.-Q. Liao and C. K. Law, Phys. Rev. A 83, 033820 (2011).25 T. A. Palomaki, J. W. Harlow, J. D. Teufel, R. W. Sim-

monds, and K. W. Lehnert, Nature 495, 210 (2013).26 T. A. Palomaki, J. D. Teufel, R. W. Simmonds, and K. W.

Lehnert, Science 8, 710 (2013).27 M.-A. Lemonde, N. Didier, and A. A. Clerk, Phys. Rev.

Lett. 111, 053602 (2013).28 J. Qian, A. A. Clerk, K. Hammerer, and F. Marquardt,

Phys. Rev. Lett. 109, 253601 (2012).29 P. D. Nation, Phys. Rev. A 88, 053828 (2013).30 N. Lorch, J. Qian, A. Clerk, F. Marquardt, and K. Ham-

merer, arXiv:1310.1298 (2013).31 S. Rips, M. Kiffner, I. Wilson-Rae, and M. J. Hartmann,

New Journal of Physics 14, 023042 (2012).32 L. Tian, Phys. Rev. B 72, 195411 (2005).33 A. Voje, J. M. Kinaret, and A. Isacsson, Phys. Rev. B 85,

205415 (2012).34 F. Xue, Y.-x. Liu, C. P. Sun, and F. Nori, Phys. Rev. B

76, 064305 (2007).35 G. Z. Cohen and M. Di Ventra, Phys. Rev. B 87, 014513

(2013).36 H. Tan, G. Li, and P. Meystre, Phys. Rev. A 87, 033829

(2013).37 X.-W. Xu, Y.-J. Zhao, and Y.-x. Liu, Phys. Rev. A 88,

022325 (2013).38 J. Suh, M. D. LaHaye, P. M. Echternach, K. C. Schwab,

and M. L. Roukes, Nano Letters 10, 3990 (2010).39 F. Massel, T. T. Heikkila, J.-M. Pirkkalainen, S. U. Cho,

H. Saloniemi, P. J. Hakonen, and M. A. Sillanpaa, Nature

480, 351 (2011).40 A. Eichler, J. Moser, J. Chaste, M. Zdrojek, I. Wilson-Rae,

and A. Bachtold, Nat. Nano. 6, 339 (2011).41 A. Voje, A. Isacsson, and A. Croy, Phys. Rev. A 88, 022309

(2013).42 K. Jacobs, arXiv:1209.2499 (2012).43 I. Mahboob, K. Nishiguchi, A. Fujiwara, and H. Yam-

aguchi, Phys. Rev. Lett. 110, 127202 (2013).44 K. V. Kheruntsyan and K. G. Petrosyan, Phys. Rev. A 62,

015801 (2000).45 K. J. McNeil and C. W. Gardiner, Phys. Rev. 28, 1560

(1983).46 W. J. Munro, Phys. Rev. A 59, 4197 (1999).47 M. D. Reid and L. Krippner, Phys. Rev. A 47, 552 (1993).48 H. M. Wiseman and G. J. Milburn, Phys. Rev. A 47, 642

(1993).49 M. D. Reid and P. D. Drummond, Phys. Rev. Lett. 60,

2731 (1988).50 P. Marian, T. A. Marian, and H. Scutaru, Phys. Rev. A

68, 062309 (2003).51 A. Gilchrist, P. Deuar, and M. D. Reid, Phys. Rev. Lett.

80, 3169 (1998).52 G. Adesso and F. Illuminati, Journal of Physics A: Math-

ematical and Theoretical 40, 7821 (2007).53 J. M. Fink, M. Goppl, M. Baur, R. Bianchetti, P. J. Leek,

A. Blais, and A. Wallraff, Nature 454, 315 (2008).54 I. Schuster, A. Kubanek, A. Fuhrmanek, T. Puppe,

P. W. H. Pinkse, K. Murr, and G. Rempe, Nature Physics4, 382 (2008).

55 A. Miranowicz, M. Bartkowiak, X. Wang, Y.-x. Liu, andF. Nori, Phys. Rev. A 82, 013824 (2010).

56 R. Simon, Phys. Rev. Lett. 84, 2726 (2000).57 L.-M. Duan, G. Giedke, J. I. Cirac, and P. Zoller, Phys.

Rev. Lett. 84, 2722 (2000).58 N. Brunner, D. Cavalcanti, S. Pironio, V. Scarani, and

S. Wehner, arXiv:1303.2849 (2013).59 E. G. Cavalcanti, C. J. Foster, M. D. Reid, and P. D.

Drummond, Phys. Rev. Lett. 99, 210405 (2007).60 J. Wenger, M. Hafezi, F. Grosshans, R. Tualle-Brouri, and

P. Grangier, Phys. Rev. A 67, 012105 (2003).61 J. F. Clauser and M. A. Horne, Phys. Rev. D 10, 526

(1974).62 J. F. Clauser, M. A. Horne, A. Shimony, and R. A. Holt,

Phys. Rev. Lett. 23, 880 (1969).63 M. Grajcar, S. Ashhab, J. R. Johansson, and F. Nori, Phys.

Rev. B 78, 035406 (2008).64 S. Sapmaz, Y. M. Blanter, L. Gurevich, and H. S. J. van der

Zant, Phys. Rev. B 67, 235414 (2003).65 A. M. Eriksson, D. Midtvedt, A. Croy, and A. Isacsson,

Nanotechnology 24, 395702 (2013).66 I. Mahboob, K. Nishiguchi, H. Okamoto, and H. Yam-

aguchi, Nat Phys 8, 387 (2012).67 M. Paternostro, D. Vitali, S. Gigan, M. S. Kim,

C. Brukner, J. Eisert, and M. Aspelmeyer, Phys. Rev. Lett.99, 250401 (2007).

68 C. Eichler, D. Bozyigit, C. Lang, M. Baur, L. Steffen, J. M.Fink, S. Filipp, and A. Wallraff, Phys. Rev. Lett. 107,113601 (2011).

69 N. Bergeal, F. Schackert, L. Frunzio, and M. H. Devoret,Phys. Rev. Lett. 108, 123902 (2012).

70 F. Mallet, M. A. Castellanos-Beltran, H. S. Ku, S. Glancy,E. Knill, K. D. Irwin, G. C. Hilton, L. R. Vale, and K. W.Lehnert, Phys. Rev. Lett. 106, 220502 (2011).

71 E. Flurin, N. Roch, F. Mallet, M. H. Devoret, and

11

B. Huard, Phys. Rev. Lett. 109, 183901 (2012).72 M. Ludwig, A. H. Safavi-Naeini, O. Painter, and F. Mar-

quardt, Phys. Rev. Lett 109, 063601 (2012).73 J. R. Johansson, P. D. Nation, and F. Nori, Comp. Phys.

Comm. 183, 1760 (2012).74 J. R. Johansson, P. D. Nation, and F. Nori, Comp. Phys.

Comm. 184, 1234 (2013).75 The source code for the numerical

simulations is available on Figshare,http://dx.doi.org/10.6084/m9.figshare.936910 (2014).