NatComm Editor v2

download NatComm Editor v2

of 21

Transcript of NatComm Editor v2

  • 7/28/2019 NatComm Editor v2

    1/21

    Realistic loophole-free Bell test with atom-photon

    entanglement

    C. Teo1, M. Arajo2,5, M. T. Quintino2,6, J. Min1, D.

    Cavalcanti1,7, V. Scarani1,3, M. Terra Cunha4, and M. Frana

    Santos2

    1Centre for Quantum Technologies, National University of Singapore, Singapore2Departamento de Fsica, Universidade Federal de Minas Gerais, Caixa Postal 702,

    30123-970 Belo Horizonte, MG, Brazil3

    Department of Physics, National University of Singapore, Singapore4Departamento de Matemtica, Universidade Federal de Minas Gerais, Caixa Postal

    702, 30123-970 Belo Horizonte, MG, Brazil5Faculty of Physics, University of Vienna, Boltzmanngasse 5, 1090 Vienna, Austria

    6Dpartement de Physique Thorique, Universit de Genve, 1211 Genve,Switzerland

    7ICFO-Institut de Ciencies Fotoniques, 08860 Castelldefels (Barcelona), Spain

    Monday 10th June, 2013

    Abstract

    The establishment of nonlocal correlations, guaranteed through theviolation of a Bell inequality, is not only important from a fundamentalpoint of view, but constitutes the basis for device-independent quantuminformation technologies. Although several nonlocality tests have beenperformed so far, all of them suffered from either the locality or the de-tection loopholes. Among the proposals to overcome these problems arethe use of atom-photon entanglement and hybrid photonic measurements(eg. photo-detection and homodyning). Recent studies have suggestedthat the use of atom-photon entanglement can lead to Bell inequality vi-olations with moderate transmission and detection efficiencies. Here wecombine these ideas and propose an experimental setup realizing a sim-ple atom-photon entangled state that can be used to obtain nonlocalitywhen considering realistic experimental parameters including detection

    efficiencies and losses due to required propagation distances.Electronic address: [email protected] address: [email protected]

    1

  • 7/28/2019 NatComm Editor v2

    2/21

    Introduction

    In recent years several applications of quantum nonlocality have been proposed[1, 2, 3, 4] (see a recent review on Bell nonlocality [5]). These proposals are basedon the fact that nonlocal correlations can be certified without any assumptionon the internal mechanisms of the devices used in the experiment. Thus, onceestablished, nonlocal correlations can be used in what is now referred as, device-independent protocols.

    Nonlocal correlations can be obtained by measuring entangled quantum sys-tems in appropriately chosen local observables. This is called a Bell test sincethe nonlocal nature of the measurement outcomes can be certified by the viola-tion of certain constraints known as Bell inequalities [5]. Many Bell tests havebeen performed in the last few decades, but no nonlocal correlations have beenstrictly established so far. This is because all of the performed experimentssuffered either from the detection loophole or the locality loophole [5].

    Experiments using entangled photons have reported Bell inequality viola-tions closing separetly the locality [6, 7, 8] and the detection [9] loopholes. Onthe other hand, the detection loophole has been closed with stationary systemslike atoms, ions and circuits [10, 11, 12]. The main technological challengeto close both loopholes simultaneously is to have both efficient detection long-distance entanglement.

    In this work, we propose an experimental setup involving available technol-ogy to implement a loophole-free Bell test. It uses a single atom coupled to thefield of a cavity in order to produce a specific entangled state between the atomand the light emitted by the cavity and combines efficient detection schemeson the atomic side and the coherent nature of the light field to perform hybriddetection on the photonic side [13, 14]. Our scheme considers experimental ef-fects neglected in previous proposals [15, 16] and thus puts current atom-photonsystems as good candidates to demonstrate loophole-free nonlocal correlations

    The paper is structured as follows: first, we introduce the target state andthe measurement settings used in the Bell test. We then give an explanation ofhow this state can be produced by means of cavity quantum electrodynamics(QED) techniques and discuss the relevant and feasible parameter regime whichyields the desired state. We also discuss an implementation of our scheme in op-tical cavities, and optimize the Clauser-Horne-Shimony-Holt (CHSH) inequality[17] violation for currently available experimental parameters. Finally, we dis-cuss possible circuit QED implementations in the final section and compare ourproposal to previous ones.

    Results

    The ideal case

    Our proposal takes advantage of two facts first noticed in [18] and [19]. First,highly efficient detections are typically available for the electronic levels of single

    2

  • 7/28/2019 NatComm Editor v2

    3/21

    atoms [20]. At the same time, photons propagating either in free space or inlow loss optical fibers are excellent candidates for carriers of information over

    long distances. Furthermore, the combination of high efficient homodyne detec-tion with photodetection greatly reduces the required minimum photodetectionefficiency [13, 14, 15, 16, 21].

    Our starting point is to consider a state of the form

    | = cos |s, 0 + sin |g, , (1)

    where |g, |s are two atomic states and |0 and | denote states of the electro-magnetic field (vacuum and a coherent state respectively) with

    | := e ||2

    2

    n=0

    nn!

    |n, (2)

    where |n are the energy eigenstates of the field. As we are going to see, this stateviolates a Bell inequality even for low efficiency photodetection and can be wellapproximated with current technology. Although we motivated this work byconsidering a real atom, the state (1) can be in principle devised using differentplatforms, e.g., superconducting qubits, quantum dots, nitrogen-vacancy centerand equivalent systems. We consider the CHSH Bell inequality [17] which statesthat the correlations obtained are nonlocal if the following inequality is violated:

    B = Tr(B) 2, (3)

    where is the quantum state under scrutiny and B denotes the Bell operator

    B = A0 B0 + A0 B1 + A1 B0 A1 B1, (4)

    with Ai and Bj being observables with outcomes 1. Here, we set the atomicobservables as

    A0 = cos z + sin x, A1 = cos z sin x, (5)

    where i are the usual Pauli matrices. In the photonic part we consider di-chotomized photodetection and Xquadrature operators (homodyne detection)[13,14]:

    B0 = 2 |00| 11, B1 = 2bb

    dx |xx| 11, (6)

    where |x is the eigenstate of the quadrature operator X = a+a2

    , and a is the

    annihilation operator of the field.Since quadrature measurements can be made very efficient (nearly perfect),there will be an asymmetry in the total efficiency of the photonic measurements:the total homodyning efficiency will be basically determined by transmissionlosses, while the total photodetection efficiency will be composed by transmis-sion and typical photodetector efficiency. We thus model the transmission losses

    3

  • 7/28/2019 NatComm Editor v2

    4/21

    0 0.2 0.4 0.6 0.8 1

    0.5

    0.6

    0.7

    0.8

    0.9

    1 A

    B

    C

    D

    |

    tline

    |

    2

    |tline|2

    = 2/3

    Ref. [16]

    |

    Figure 1: Critical line above which CHSH violations are obtained Theblue line corresponds to the ideal state of our proposal. The parameters , , , bwere optimized for each point. For comparison, we include the curve |tline|2 =2/3 (red) that results from the Eberhard bound [22, 23], and the curve fromthe best experimental proposal to date in Ref [16] (green). For the sake ofillustration, we give the specific numbers for the points represented by crossesA, B, C and D in Table 1.

    A B C D|tline|2 1 0.8 1 0.55 1 0.8 0.15 1|| 2.1 2.33 3.35 2.38 0.55 0.34 0.14 0.03 0.77 0.66 0.16 0.33b 0.53 0.53 0.34 0.44

    B 2.32 2.07 2+ 2+

    Table 1: Optimized parameters for 4 points on Fig. 1. The state andmeasurement parameters, ||, , and b, are each numerically optimized, givensome detector efficiency , and some transmission |tline|2.

    with a beam splitter with transmittance tline (affecting both the photodetectorand the homodyning apparatus) and the photodetector intrinsic efficiency as abeam splitter with transmittance

    followed by a perfect detector.

    Fig. 1 summarizes the results. In ideal conditions (i.e. point A in Table

    4

  • 7/28/2019 NatComm Editor v2

    5/21

    1), the CHSH value can reach B = 2.32 for || = 2.1, which translates toan average of 4.41 photons in the coherent state. Moreover, a violation can be

    found even for = 0.15 with |tline|2 = 1 (Point C in Table 1), or transmittance|tline|2 = 0.55 with = 1 (Point D in Table 1). Also, for a detection efficiency of = 0.8 and a transmission of|tline|2 = 0.8 we find a CHSH value of2.07 (Point Bin Table 1). For comparison, we have also included the curve |tline|2 = 2/3 thatresults from the Eberhard bound [22, 23] , which is the required efficiency andtransmission to perform a loophole-free experiment with photon polarization,and the curve from the best experimental proposal to date involving an atomand a photonic mode [16].

    Realistic scenario

    We now show a scheme aiming to produce the state (1) in a cavity QED scenario.The first part of the scheme is depicted in Fig. 2a. An input field is incidenton a cavity with a single three level atom prepared initially in a superpositionof the states g and s: |atom = cos |s + sin |g. Fig. 2b shows the levelstructure of the atom. Transition |g |e is coupled dispersively to the cavitywith detuning = ge c, where c is the cavity frequency and ge is thefrequency of the transition |g |e. The cavity is asymmetric with decay ratesof the mirrors b c and the cavity decay rate = b + c. This means thatthe field leaks out of the cavity essentially only through the right mirror (cf.Fig. 2a). The fact that one needs an asymmetric cavity is not a priori obviousand will be explained below. We assume that the level |s is detuned far enoughsuch that it does not interact with the cavity field.

    Since the cavity transmission depends on the atomic state, the atom-cavitysystem acts as a filter that transforms the initial state |in|atom in the fol-lowing way:

    |in |g|rginrefl |g |tgintrans (7)|in |s|rsinrefl |s |tsintrans (8)

    where |refl and |trans represent the reflected and transmitted photonic states,in is the complex amplitude of the input coherent state and rj and tj (j = s, g)are the amplitude reflectivity and transmitivity of the empty cavity (atom in thestate |s, i.e. j = s) and of the cavity with the atom (atom in the state |g,i.e.

    j = g) and |rj |2 + |tj |2 = 1. Notice that we assume that both the reflected andtransmitted fields are still coherent states (see Methods for details).

    In order to produce a state of the form (1), one needs to meet the followingconditions: ts = 0 with rsin|rgin = 1 and |tgin = |, where | is the

    desired coherent state in (1). An alternative is to apply a displacement on thetransmitted field [24]. The displacement is just the result of combining a fieldexiting the atom-cavity system (cout mode) with a coherent local oscillator |(dLO mode) on a beam splitter, as depicted in Fig. 2c. The output port ofinterest is the port with the output field |tBScout rBSdLO, where rBS and tBSare the amplitude reflectivity and transmitivity of the beam splitter respectively.

    5

  • 7/28/2019 NatComm Editor v2

    6/21

    Figure 2: State preparation. a. An input field is incident on a cavity withan atom in the dispersive regime. This causes some reflection and some trans-mission of the cavity field. We assume that the mirror on the right has a lowerreflectivity than the mirror on the left, such that the field predominantly leavesthe cavity to the right. The final beam splitter performs a displacement oper-ation which creates a superposition of propagating vacuum and coherent state.

    b. Level structure of the atom. The cavity field is dispersively coupled to the|g |e transition, and the |s state is assumed to be far detuned such thatit does not interact with the cavity field. c. Beam splitter convention used.A displacement operation is applied on the mode labeled cout by combining itwith a local oscillator (mode labeled dLO).

    For a coherent state, |x, in the cout mode, one can achieve a zero amplitudein this output port by tuning the amplitude of | such that rBS = tBSx. Inour case, x = tsin and one can obtain the state |s, 0 contained in (1).

    The full Bell test setup is shown in Fig. 3. The beam splitters are identicalwith the property |tBS|2 |rBS|2. Note that, in order to satisfy the conditionrBS = tBSx, the spatiotemporal modes of the two coherent states x and

    must be identical. This can be achieved e.g. by using an auxiliary cavity similarto the one containing the atom (see Fig. 3), where x = tsin. Eventually,instead of using the auxiliary cavity, one might look for an atom-cavity-laserconfiguration in which the coupling of the field of polarization connecting the|s state with some excited state is negligible compared to the dispersive couplingof the |g |e transition, and the orthogonal polarization coupling to neither

    6

  • 7/28/2019 NatComm Editor v2

    7/21

  • 7/28/2019 NatComm Editor v2

    8/21

    and the visibility V is (see Methods for details),

    V = exp

    F ||22t2BS

    , (12)

    F = r2BS + fcav + IsL1

    4C(1 + fcav), (13)

    IsL =

    d|sL() 1D() |2

    d|sL() 1D() 11+i2(c)/ |2, (14)

    D() = (

    2+ i( ge))(/2 + i( c)) + g2, (15)

    where C is the usual single-atom cooperativity C = g2

    and fcav =b+L

    cis

    a factor which describes the asymmetry of the cavity. Here, the continuousfrequency coherent state,

    |{

    }, is

    |{} = exptBSin

    d sL() (tg() ts()) cout() + h.c.|0, (16)

    ||2 = |tBSin|2

    dsL()(tg() ts())2, (17)

    with the transmission coefficients

    ts() =

    bc

    /2 + i( c) , (18)

    tg() =

    bc

    D()

    2+ i( ge)

    , (19)

    and the atomic state is initially prepared with such that the final state is ofthe form (10). In the above, sL() is the frequency spectrum of the laser field,g is the coupling constant of the cavity mode to the |g |e transition, isthe transverse decay rate of the |e state and L is the loss rate of the cavitymirrors.

    From equation (12), it is easy to see that for F 0, we have V 1.Also for F 0, we need the 3 conditions, rBS 0, fcav 0 and C . In practice, the first condition means that one should use a small value ofthe beam splitter reflectivity and adjust the amplitude of the local oscillator,such that the condition rBS = tBStsin is still satisfied. The second conditionis precisely the requirement of the asymmetric cavity we have mentioned atthe beginning of this section and is dependent only on the transmission andlosses of the cavity mirrors used. For completeness, we have also included non-

    radiative mirror losses. The third condition means that one needs large single-atom cooperativity. One remarkable feature of this result is that for C 1,the visibility V (but of course not the state) is independent of the details of thespectrum of the input pulse.

    The final point to note is that to satisfy the approximations used in ourderivation, we require a negligible probability of exciting the atom throughout

    8

  • 7/28/2019 NatComm Editor v2

    9/21

    the duration of the input pulse. This is important, since, to maintain somespecific amplitude of the output coherent state

    |{

    }after the beam splitter,

    one might have to use a very large amplitude input laser if the integral inequation (17) is small. For the sake of this theoretical analysis, we consider thecase of an input Gaussian pulse with the normalized spectrum

    |sL()|2 = 1L

    eLL

    2

    , (20)

    where L is the central laser frequency, and L is the bandwidth of the pulse.To close the locality loophole, one has to propagate the outgoing pulse for

    some minimum distance, given by the larger of the measurement times of theatom and the pulse, which depends on the pulse duration. The pulse durationis given by the relevant time scale of the system, i.e., either by the input pulseduration or the cavity lifetime, whichever is greater. This means that, from the

    point of view of closing the locality loophole, it is useful to push the duration ofthe input pulse down to the cavity lifetime, but not necessarily further. Sincethe pulse duration and the cavity lifetime are proportional to the inverse of thepulse bandwidth L and the cavity decay rate respectively, the best case wouldbe ifL .

    Figure 4: Log plots of the maximum probability of atomic excitationThis figure shows plots of log10[maxt Pe(t)] as a function of the parameter

    g ,

    assuming that the central laser frequency is on resonance with the bare cavity,for different values of g . a. The case with input pulse with minimal bandwidthsetting g/L = max(g/). b. The case with input pulse bandwidth set equalto the cavity bandwidth (g/L = g/). For g/ small enough, the probabilityof exciting the atom becomes insensitive to the precise value of g/. Thisfact is illustrated by the overlap of curves corresponding to g/ = 1/100 and

    g/ = 1/1000. The parameters used in this plot are || = 2.1, fcav = 4/100and g/ = 5/3.

    Fig. 4 shows plots oflog10(maxt Pe(t)), the maximum probability of excitingthe atom as a function of the parameter g (the coupling strength over the totalcavity decay rate), assuming that the central laser frequency is on resonance

    9

  • 7/28/2019 NatComm Editor v2

    10/21

    with the bare cavity, for g = 1/10, 1/100, 1/1000. Fig. 4a is the case for a fixedminimal bandwidth of the input pulse in the sense that we choose L such thatg/L = max(g/). Fig. 4b corresponds to the situation when the input pulsebandwidth is set equal to the cavity bandwidth in order to minimize the pulseduration. In the following, we will consider the regions where (maxt Pe(t)) 0.1,to be the parameter regimes when our analysis is suitable. As is evident fromFig. 4, for some large g/L, there always exists some g/ for which our solutionsare self-consistent, which is not the case when L = , the situation minimizingthe pulse duration. We can thus look for some intermediate value ofL (havingall the other parameters fixed), which minimizes the pulse duration while stillrespecting the validity of the used approximation. This is the strategy we employin the next section, where we will discuss a possible implementation of ourscheme using realistic experimental parameters.

    Experimental feasibilitySo far, we theoretically described a general cavity scenario. In the following, wewill focus our attention on possible implementations of our scheme in opticalcavities. This implementation has the advantage of allowing large propagationdistances due to the availability of low loss optical fibers.

    We start by general constraints on measurement times and propagation dis-tances imposed by the locality loophole. To close the locality loophole, onerequires the start of the measurement event on one side to be space-like sepa-rated from the end of the measurement event on the other side. In other words,we need that the space-time coordinates of the end of the atomic measurementand the choice of measurement basis (photon detection or homodyning) on thephotonic side be space-like separated, and vice versa. This constraint translatesinto the minimum propagation distance

    d c (max ((tph,c + tph,m), (tat,c + tat,m))) , (21)

    where c is the speed of light in free-space, tat/ph,c is the time required on theatomic/photonic side to choose the measurement settings, which also includesthe time to change the measurement basis, tat/ph,m is the time required onthe atomic/photonic side to perform the actual measurement, which necessarilyincludes the duration of the light pulse. We assume that the main time con-straint is set by the atomic measurement, which is typically slower than themeasurement of photons. The above equation shows, that even if we used avery short pulse, we would still need to propagate it at least to a distance, givenby the sum of atomic measurement time and the time required to choose a basis(a rotation in the space of

    |s

    and|g, typically using RF pulses), multiplied

    by c. The best trade-off in this case, is thus to have tph,m tat,m + tat,c.This is possible because the time needed to choose a basis on the photonic sidecan be very fast [7].

    One possibility to perform a fast atomic state detection would be to usea 2-photon ionization technique [20]. This technique has been performed in

    10

  • 7/28/2019 NatComm Editor v2

    11/21

    free-space configurations and can achieve 98% efficiency in < 1s. Althoughthe restriction of a small cavity might make such a photo-ionization technique

    technically challenging, proposals to make focusing cavities with large physicalvolumes (lengths of 1 cm) while still maintaining large coupling constants(g 100MHz) exist [26, 27] potentially making such techniques compatible.Here, we assume tat,m + tat,c = 1 s. We should thus seek to produce a pulsewhich requires a total measurement duration up to 1s. For the Gaussian pulseexample considered in our calculation, we take the total pulse measurementduration to be tph,m = 6

    2/L,m. The factor 6

    2 is chosen such that if

    the bandwidth of the pulse incident on the cavity is L,m, then the neglectedtails of the outgoing pulse correspond to L,m (shorter pulses), the totalmeasurement time is set by tat,m+tat,c = 1 s and the minimum propagationdistance is 300 m.

    Next, we use experimental parameters given in [28, 29, 30]. These exper-iments use 852 nm and 780 nm light respectively, which are subject to about2 dB/km loss in optical fibers. In the following, we compare both experimentsand possible modifications. As discussed previously, we will work in the largedetuning regime, g. We choose an arbitrary, but subjected to experimentalconstraints, value of g/ = 1/10. This ensures that we respect all approxima-tions used, as long as we also satisfy condition maxt Pe(t) 0.1. The figure ofmerit in both cases will be 0.1, which is the largest acceptable laser bandwidththat satisfies condition maxt Pe(t) 0.1, and the state production visibility(12), which is computed using the actual laser bandwidth min(, 0.1, L,m).We also include the effect of finite pulse measurement time on the visibility,

    using (12)-(15). We assume || = 2.1 and |rBS|2

    = 0.001. Table 2 summarizesthe results.The parameters of Ref. [28, 29] are (g/(2), /(2), /(2), fcav) = (34 MHz,

    4.1 MHz, 2.6 MHz, 10/4) (we assume that the experiment performed implementsa symmetric cavity). Due to the symmetry of the cavity, one obtains a smalleffective visibility. Assuming that the cavity could be made asymmetric reducingfcav to the current value in the experiment of Ref. [30], the visibility dramaticallyincreases to 72.7% (second row). Neglecting the mirror losses, we show in thethird row that V can be as high as 90.8%.

    The parameters of Ref. [30] are (g/(2), /(2), /(2), fcav) = (5 MHz,3 MHz, 3 MHz, 14/100). As the setup stands, the visibility is 56.3%. However,neglecting the mirror losses, V increases to 71.3% (fifth row). If it were furtherpossible to reduce the total cavity decay rate by a factor of 2, thus increasing

    the cooperativity, while maintaining the same asymmetry, the visibility furtherincreases to 77.2% (sixth row). The required incident photon number can becalculated from (17). For the parameters in Table 2 and requiring the resultingphoton number, ||2 = 2.12, one requires |in|2 25 400 input photons.

    For the specific case of 87Rb, one might also identify possible states playingthe role of |g, |s and |e. We may choose for example the |s state to be the

    11

  • 7/28/2019 NatComm Editor v2

    12/21

    g/(2) /(2) /(2) fcav 0.1/(2) V dmin34 MHz 4.1 MHz 2.6 MHz 10/4 13.3 MHz 0.4% 300 m

    34 MHz 4.1 MHz 2.6 MHz 14/100 63.3 MHz 72.7% 300 m34 MHz 4.1 MHz 2.6 MHz 4/100 65.2 MHz 90.8% 300 m5 MHz 3 MHz 3 MHz 14/100 1.1 MHz 56.3% 370 m5 MHz 3 MHz 3 MHz 4/100 1.3 MHz 71.3% 310 m5 MHz 1.5 MHz 3 MHz 4/100 3.1 MHz 77.2% 300 m

    Table 2: Derived visibilities with available experimental parameters.All parameters (g/(2), /(2), /(2), fcav) in the first and fourth row areactual cavity parameters (including mirror losses) obtained from [28, 29, 30].The second row shows the effect on the visibility and 0.1 by reducing fcav tothe current value in the experiment of Ref. [30]. The third row is obtained byneglecting the mirror losses, which further decreases the value of fcav. The fifthrow shows the effect of neglecting mirror losses, and the last row shows the effectof increasing g/. Notice that 0.1 > 2 1.35 MHz in the first three and lastrows, which means that the propagation distance is limited by the detectiontime on the atomic side and not by the pulse duration and corresponds to thepropagation distance of 300 m. The state production visibility V is computedassuming || = 2.1, |rBS|2 = 0.001, using the laser bandwidth min(, 0.1, L,m)and taking into account the truncation of the pulse due to finite measurementtime.

    hyperfine state |5S1/2, F = 1, mF = 1, the |g state |5S1/2, F = 2, mF = 2 andthe |e state |5P3/2, F = 3, mF = 3. In this case, the input pulse and cavityfield would have a + polarization coupling the

    |g

    |e

    transition. Due to the

    large detuning of the hyperfine states (6.8 GHz), and the fact that the s stateis far-detuned to any other + transitions, these states are possible candidatesfor the experiment.

    A B C|tline|2 1 0.9 0.83

    1 0.75 0.63 0.42 0.33 0.02 0.75 0.6 0.03b 0.53 0.56 0.58

    B 2.17 2.05 2+

    Table 3: Optimized parameters for 3 points on Fig. 5. These parametersare results from numerical optimization of the state (9), given some detectorefficiency and some transmission |tline|2. We have fixed || = 2.1, and usedthe Visibility V (12) from the second row of Table 2.

    We now maximize B as a function of transmission and detector inefficiency

    12

  • 7/28/2019 NatComm Editor v2

    13/21

    | |

    |

    0.2 0.4 0.6 0.8 1

    0.5

    0.6

    0.7

    0.8

    0.9

    12 2.05 2.1 2.15|

    A

    B

    C

    |

    tline

    |

    2

    Figure 5: Contour lines of B as a function of and |tline|2. In this plot wefixed || = 2.1 and optimized the measurement and state parameters , b, foreach point. The parameters used are (g/(2), /(2), /(2), fcav) = (34MHz,4.1MHz, 2.6MHz, 14/100) and |rBS|2 = 0.001 (second row of Table 2). Theresult for the ideal state in (1) (dotted line) is included for comparison. For thesake of illustration we give the specific numbers for the points represented bycrosses A, B and C in Table 3.

    for a set of realistic parameters. We thus set || = 2.1, (g/(2), /(2), /(2), fcav)= (34 MHz, 4.1 MHz, 2.6 MHz, 14/100) and |rBS|2 = 0.001 (second row of Table2). Notice that we have included possible non-radiative losses from cavity mir-rors. Given these constraints, the maximal CHSH violation is 2.17. Moreover,we can attain a violation even for = 0.37 with |tline|2 = 1, or transmittance|tline|2 = 0.68 with = 1. Fig. 5 summarizes these results. Note that the higherthe visibility of the produced state, the closer one gets to the ideal scenario(dashed line in Fig. 5)

    Discussion

    In this work, we have shown that using current technology, it is possible toproduce a hybrid atom-photon entangled state that still violates the CHSHinequality up to a value of 2.17. We also showed that we can attain a violationeven for a low photon counting efficiency of 37% with perfect transmission, or aline transmission of 68% with perfect detection efficiency. Moreover, assumingthat it is possible to perform photoionization measurements < 1s, the required

    13

  • 7/28/2019 NatComm Editor v2

    14/21

    propagation distances to close the locality loophole are of the order of 300m foroptical setups. This gives a very good outlook for optical systems in eventually

    performing a loophole-free Bell test.In principle, one may also consider circuit QED setups. The advantages of

    such setups are twofold. First, the very fast and efficient qubit state detection,on the order of 10 ns [12] potentially lowers the required propagation distancesto laboratory scale distances (10 m or less). Second, in this system, large ratiosof coupling constant to cavity decay, g/, can be achieved. The drawbacks withcurrent technology are the limited efficiencies of both photodetection and homo-dyne detection (private communication with Steve Girvin), and the requirementto cool the propagation line down to cryogenic temperatures. The hope is thatboth these drawbacks, being of technological nature, can be eventually overcomein near future experiments.

    Before finishing let us compare our present results with previous proposalsinvolving similar setups. In Ref. [16] a Bell test involving the production of anentangled state between an atom and a photonic field created by atomic decaywas studied (see also [31]). It was shown that Bell violations with photodetectionefficiency of39% (see green curve in Fig. 1) can be achieved. Note however thatthis number refers to the ideal state, considering that all light emitted by theatom is collected. Our scheme overcomes this problem since the photonic partof the state is the transmitted mode, and, moreover, requires lower efficiencies.

    We believe that our proposal will trigger possible implementations of loophole-free Bell tests with atom-photon interfaces, which are particularly important inquantum communication and cryptographic applications.

    Methods

    Input-output relations

    Here we derive the input-output relations [25] for a two-sided cavity, clearlystating the approximations used, and their validity. We consider a cavity mode(a mode) which couples to a left mode (b mode) and a right mode (c mode).We also include loss in the mirrors of the cavity as the coupling of the cavity amode to an additional L mode.These give the following Heisenberg equation of motion for the cavity field,

    ta(t) = i

    [Hsys , a(t)] 2

    a(t)

    +

    bbin(t) +

    ccin(t) +

    LLin(t), (22)

    where = b + c + L and Hsys is the system Hamiltonian (empty cavity orcavity with atom) without considering the baths, which is either

    Hempty = caa or, (23)

    Hatomcavity = caa + a + g(a + a). (24)

    14

  • 7/28/2019 NatComm Editor v2

    15/21

    Equation (22) can also be written in the equivalent way

    ta(t) = i

    [Hsys , a(t)] kc2

    a(t)

    +

    bbin(t) ccout(t) + LLin(t), (25)

    where k = 2 and we now consider the output operator for the c mode,where the input and output operators are defined as,

    Oin(t) =1

    2

    d O(t0)e

    i(tt0), (26)

    Oout(t) =12

    d O(t1)e

    i(tt1). (27)

    For the case of an empty cavity, these equations are just a linear system of

    differential equations, and can be solved simply by Fourier transforms definedas

    f(t) =12

    d eitf(), (28)

    f() =12

    dt eitf(t), (29)

    to give bin()cin()

    Lin()

    =Us

    bout()cout()

    Lout()

    , (30)

    where

    Us =rb() tb() lb()tb() rc() lc()

    lb() lc() rL()

    (31)

    r() = k

    2 + i( c)2 + i( c)

    (32)

    tb() =

    bc

    2 + i( c)

    (33)

    lb/c() =

    Lb/c

    2 + i( c)

    (34)

    These are relations used in the main text when the atom is in the |s state,since it is assumed to be decoupled from all cavity and environmental modes.However, when the atom is in the |g state, it is assumed to couple to the cavitya mode. In this situation, one has to include atomic spontaneous emission. Thiscan be done by including the coupling of the |g |e transition to an additional

    15

  • 7/28/2019 NatComm Editor v2

    16/21

    bath, E mode, different from the cavity. This gives the set of equations

    ta(t) = i

    [Hatomcavity , a(t)] 2

    a(t)

    +

    bbin(t) +

    ccin(t) +

    LLin(t), (35)

    t(t) = i

    [Hatomcavity , (t)] 2

    (t)

    zEin(t), (36)

    where describes the rate of emission into modes other than the cavity modes,and in is the input operator of the environment. Note that can be madesmall if the physical cavity mode has a large spatial overlap with the emissionpattern of the |g |e transition.

    In the following, we will investigate the system dynamics in the low excitationregime. The reason is twofold. First, we want to avoid exciting the atom inorder to prevent the decay into the environment (spontaneous emission with

    the decay rate ). Second, populating the excited state would induce a morecomplex dynamics producing in general a nontrivial entangled state betweenthe atom and input, output and cavity fields. This is certainly an interestingregime to investigate in the context of quantum state engineering, but in thepresent paper we focus on the more intuitive picture in the spirit of ( 8). Theassumption of the atom occupying mostly the ground state can be translatedas z 11. With this assumption, the above set of equations can be solvedanalytically. Proceeding analogously to the empty cavity case, we have

    bin()cin()Lin()Ein()

    =Ug

    bout()cout()Lout()Eout()

    , (37)

    where the unitary matrix

    Ug = 1D()

    Us(

    2 + ia)(

    2 + ic) ig

    ig

    T (2 ia)(2 + ic)

    g211

    , (38)

    with a = ge, c = c,Us is defined in equation (31), the vector reads

    =

    bcL

    (39)

    and

    D() = ( 2 + ia)( 2 + i

    c) + g2. (40)

    Notice that this set of equations reduces to (30) for the b and c modes, wheng = 0. Finally, we also assume that all output operators (bout, cout, Lout, Eout)commute with each other, which is not true in general. This approximation isrequired, for an input coherent field in the bin mode to transform into a coherent

    16

  • 7/28/2019 NatComm Editor v2

    17/21

    reflected field in the bout mode and coherent fields in the cout, Lout and Eoutmodes.

    Lastly, as a consistency check of our work, we note that the approxima-tion z 11 necessarily requires that the atom remains in the ground statethroughout its evolution. Solving for (), one obtains

    () ig

    bbin()

    (2 ia)(2 ic) + g2, (41)

    where we have used the fact that the cin and in modes represent thermal noiseand are thus taken to be negligible compared to the bin mode. Taking the inversefourier transform, one obtains

    Pe(t) = (t) =

    1

    2

    dig

    bsL()

    (2

    i(

    a

    ))(2

    i(

    c

    )) + g2eit

    2

    .

    (42)We then conclude that the self-consistency of our approximation requires that

    (t) 1, t, which implies that maxt Pe(t) 1. Notice that this condition

    is a necessary but not sufficient condition for the approximation z 11, sincethis condition itself is derived under the approximation. However, it should benoted that equation (41) can be written in terms of a() to obtain

    () ig2 i( a)

    a(). (43)

    =ig

    i(1 c + 2i)a() (44)

    Then, for c, , meaning that if the laser addresses only wavelengthsclose to the bare cavity resonance compared to the detuning between the atomand the cavity, and the atom-cavity detuning is many atomic linewidths awayfrom resonance, one has the condition

    (t)

    g

    2 aa(t)

    1, (45)which is the condition of validity of the dispersive approximation (see [32]). Thismeans that the condition c, , together with condition maxt Pe(t) 1, is sufficient to justify the approximation z 11.

    Derivation of visibility

    The state produced after the cavity and beam splitter is of the form

    cos |s, 0 |o, b, L, E + sin |g, |o , b , L , E (46)

    where o,b,L and E are the other port of the beam splitter, the reflected fieldfrom the cavity, the field loss in cavity mirrors and the spontaneous emission

    17

  • 7/28/2019 NatComm Editor v2

    18/21

    term respectively. If we measure only the atomic state and the a mode, we loseinformation in all the other modes. Tracing over all the rest of the modes, we

    obtain the state (9), where the visibility, V is,

    V = |o|o b|b L|L E|E | (47)We take only the magnitude of the inner products, since the total phase can, inprinciple, be compensated by suitable atomic state preparation. This gives theexpression

    V = exp

    F ||2

    2t2BS

    , (48)

    F = fcav + IsL1

    4C(1 + fcav) + r

    2BS, (49)

    IsL =d|

    sL()1

    D()

    |2

    d|sL() 1D() 11+i2(c)/ |2 , (50)

    C =g2

    , (51)

    fcav =b + L

    c, (52)

    where, as in the main text, tBS and rBS are the transmittivity and reflectivityof the beam splitter used in the displacement operation, sL() is the spectrumof the laser input field and i is the coupling rate of the cavity mode to the ithbath.

    References

    [1] A. Acn, N. Brunner, N. Gisin, S. Massar, S. Pironio, and V. Scarani,Device-Independent Security of Quantum Cryptography against CollectiveAttacks, Phys. Rev. Lett. 98, 230501 (2007).

    [2] S. Pironio, A. Acn, S. Massar, A. B. d. l. Giroday, D. N. Matsukevich,P. Maunz, S. Olmschenk, D. Hayes, L. Luo, T. A. Manning, and C. Monroe,Random numbers certified by Bells theorem, Nature 464, 10211024(2010).

    [3] C.-E. Bardyn, T. C. H. Liew, S. Massar, M. McKague, and V. Scarani,Device-independent state estimation based on Bells inequalities, Phys.Rev. A 80, 062327 (2009).

    [4] R. Rabelo, M. Ho, D. Cavalcanti, N. Brunner, and V. Scarani, Device-Independent Certification of Entangled Measurements, Phys. Rev. Lett.107, 050502 (2011).

    [5] N. Brunner, D. Cavalcanti, S. Pironio, V. Scarani, S. Wehner. Bell nonlo-cality, Preprint at http://arxiv.org/abs/1303.2849 (2013).

    18

    http://dx.doi.org/10.1103/PhysRevLett.98.230501http://dx.doi.org/10.1103/PhysRevLett.98.230501http://dx.doi.org/10.1103/PhysRevLett.98.230501http://dx.doi.org/10.1038/nature09008http://dx.doi.org/10.1038/nature09008http://dx.doi.org/10.1038/nature09008http://dx.doi.org/10.1038/nature09008http://dx.doi.org/10.1103/PhysRevA.80.062327http://dx.doi.org/10.1103/PhysRevA.80.062327http://dx.doi.org/10.1103/PhysRevA.80.062327http://dx.doi.org/10.1103/PhysRevA.80.062327http://dx.doi.org/10.1103/PhysRevLett.107.050502http://dx.doi.org/10.1103/PhysRevLett.107.050502http://dx.doi.org/10.1103/PhysRevLett.107.050502http://dx.doi.org/10.1103/PhysRevLett.107.050502http://dx.doi.org/10.1103/PhysRevLett.107.050502http://dx.doi.org/10.1103/PhysRevLett.107.050502http://dx.doi.org/10.1103/PhysRevA.80.062327http://dx.doi.org/10.1103/PhysRevA.80.062327http://dx.doi.org/10.1038/nature09008http://dx.doi.org/10.1038/nature09008http://dx.doi.org/10.1103/PhysRevLett.98.230501
  • 7/28/2019 NatComm Editor v2

    19/21

    [6] A. Aspect, J. Dalibard, and G. Roger, Experimental Test of Bells In-equalities Using Time- Varying Analyzers, Phys. Rev. Lett. 49, 18041807

    (1982).

    [7] G. Weihs, T. Jennewein, C. Simon, H. Weinfurter, and A. Zeilinger, Vio-lation of Bells Inequality under Strict Einstein Locality Conditions, Phys.Rev. Lett. 81, 50395043 (1998).

    [8] T. Scheidl, R. Ursin, J. Kofler, S. Ramelow, X. Ma, T. Herbst,L. Ratschbacher, A. Fedrizzi, N. K. Langford, T. Jennewein, andA. Zeilinger, Violation of Local Realism with Freedom of Choice, PNAS107, 1970819713 (2010).

    [9] M. Giustina, A. Mech, S. Ramelow, B. Wittmann, J. Kofler, J. Beyer,A. Lita, B. Calkins, T. Gerrits, S. W. Nam, R. Ursin, A. Zeilinger, Bellviolation using entangled photons without the fair-sampling assumption,Nature (2013).

    [10] D. N. Matsukevich, P. Maunz, D. L. Moehring, S. Olmschenk, and C. Mon-roe, Bell Inequality Violation with Two Remote Atomic Qubits, Phys.Rev. Lett. 100, 150404 (2008).

    [11] M. A. Rowe, D. Kielpinski, V. Meyer, C. A. Sackett, W. M. Itano, C. Mon-roe, and D. J. Wineland, Experimental violation of a Bells inequality withefficient detection, Nature 409, 791794 (2001).

    [12] M. Ansmann, H. Wang, R. C. Bialczak, M. Hofheinz, E. Lucero, M. Neeley,A. D. OConnell, D. Sank, M. Weides, J. Wenner, A. N. Cleland, and J. M.Martinis, Violation of Bells inequality in Josephson phase qubits, Nature

    461, 504506 (2009).[13] D. Cavalcanti, N. Brunner, P. Skrzypczyk, A. Salles, and V. Scarani, Large

    violation of Bell inequalities using both particle and wave measurements,Phys. Rev. A 84, 022105 (2011).

    [14] M. Tlio Quintino, M. Arajo, D. Cavalcanti, M. Frana Santos, andM. Terra Cunha, Maximal violations and efficiency requirements for Belltests with photodetection and homodyne measurements, J. Phys. A:Math. Theor. 45, 215308 (2012).

    [15] M. Arajo, M. Tlio Quintino, D. Cavalcanti, M. Frana Santos, A. Ca-bello, and M. Terra Cunha, Bell tests with arbitrarily low photodetec-tion efficiency and homodyne measurements, Phys. Rev. A 86, 030101(R)

    (2012).[16] N. Sangouard, J.-D. Bancal, N. Gisin, W. Rosenfeld, P. Sekatski, M. Weber,

    and H. Weinfurter, Loophole-free Bell test with one atom and less thanone photon on average, Phys. Rev. A 84, 052122 (2011).

    19

    http://dx.doi.org/10.1103/PhysRevLett.49.1804http://dx.doi.org/10.1103/PhysRevLett.49.1804http://dx.doi.org/10.1103/PhysRevLett.49.1804http://dx.doi.org/10.1103/PhysRevLett.49.1804http://dx.doi.org/10.1103/PhysRevLett.81.5039http://dx.doi.org/10.1103/PhysRevLett.81.5039http://dx.doi.org/10.1103/PhysRevLett.81.5039http://dx.doi.org/10.1103/PhysRevLett.81.5039http://dx.doi.org/10.1103/PhysRevLett.81.5039http://dx.doi.org/10.1073/pnas.1002780107http://dx.doi.org/10.1073/pnas.1002780107http://dx.doi.org/10.1073/pnas.1002780107http://dx.doi.org/10.1073/pnas.1002780107http://dx.doi.org/10.1038/nature12012http://dx.doi.org/10.1038/nature12012http://dx.doi.org/10.1103/PhysRevLett.100.150404http://dx.doi.org/10.1103/PhysRevLett.100.150404http://dx.doi.org/10.1103/PhysRevLett.100.150404http://dx.doi.org/10.1103/PhysRevLett.100.150404http://dx.doi.org/10.1103/PhysRevLett.100.150404http://dx.doi.org/10.1038/35057215http://dx.doi.org/10.1038/35057215http://dx.doi.org/10.1038/35057215http://dx.doi.org/10.1038/35057215http://dx.doi.org/10.1038/nature08363http://dx.doi.org/10.1038/nature08363http://dx.doi.org/10.1038/nature08363http://dx.doi.org/10.1038/nature08363http://dx.doi.org/10.1103/PhysRevA.84.022105http://dx.doi.org/10.1103/PhysRevA.84.022105http://dx.doi.org/10.1103/PhysRevA.84.022105http://dx.doi.org/10.1103/PhysRevA.84.022105http://dx.doi.org/10.1088/1751-8113/45/21/215308http://dx.doi.org/10.1088/1751-8113/45/21/215308http://dx.doi.org/10.1088/1751-8113/45/21/215308http://dx.doi.org/10.1088/1751-8113/45/21/215308http://pra.aps.org/abstract/PRA/v86/i3/e030101http://pra.aps.org/abstract/PRA/v86/i3/e030101http://pra.aps.org/abstract/PRA/v86/i3/e030101http://pra.aps.org/abstract/PRA/v86/i3/e030101http://dx.doi.org/10.1103/PhysRevA.84.052122http://dx.doi.org/10.1103/PhysRevA.84.052122http://dx.doi.org/10.1103/PhysRevA.84.052122http://dx.doi.org/10.1103/PhysRevA.84.052122http://dx.doi.org/10.1103/PhysRevA.84.052122http://pra.aps.org/abstract/PRA/v86/i3/e030101http://pra.aps.org/abstract/PRA/v86/i3/e030101http://dx.doi.org/10.1088/1751-8113/45/21/215308http://dx.doi.org/10.1088/1751-8113/45/21/215308http://dx.doi.org/10.1103/PhysRevA.84.022105http://dx.doi.org/10.1038/nature08363http://dx.doi.org/10.1038/nature08363http://dx.doi.org/10.1038/35057215http://dx.doi.org/10.1103/PhysRevLett.100.150404http://dx.doi.org/10.1103/PhysRevLett.100.150404http://dx.doi.org/10.1038/nature12012http://dx.doi.org/10.1073/pnas.1002780107http://dx.doi.org/10.1073/pnas.1002780107http://dx.doi.org/10.1103/PhysRevLett.81.5039http://dx.doi.org/10.1103/PhysRevLett.81.5039http://dx.doi.org/10.1103/PhysRevLett.49.1804http://dx.doi.org/10.1103/PhysRevLett.49.1804
  • 7/28/2019 NatComm Editor v2

    20/21

    [17] J. F. Clauser, M. A. Horne, A. Shimony, and R. A. Holt, Proposed Ex-periment to Test Local Hidden-Variable Theories, Phys. Rev. Lett. 23,

    880884 (1969).

    [18] N. Brunner, N. Gisin, V. Scarani, and C. Simon, Detection Loophole inAsymmetric Bell Experiments, Phys. Rev. Lett. 98, 220403 (2007).

    [19] A. Cabello and J.-. Larsson, Minimum Detection Efficiency for aLoophole-Free Atom-Photon Bell Experiment, Phys. Rev. Lett. 98, 220402(2007).

    [20] F. Henkel, M. Krug, J. Hofmann, W. Rosenfeld, M. Weber, and H. We-infurter, Highly Efficient State-Selective Submicrosecond PhotoionizationDetection of Single Atoms, Phys. Rev. Lett. 105, 253001 (2010).

    [21] J. B. Brask, N. Brunner, D. Cavalcanti, and J. Leverrier, The use of am-

    plifiers in continuous-variable Bell tests, Phys. Rev. A 85, 042116 (2012).

    [22] P. H. Eberhard, Background level and counter efficiencies required fora loophole-free Einstein-Podolsky-Rosen experiment, Phys. Rev. A 47,R747R750 (1993).

    [23] J.-. Larsson and J. Semitecolos, Strict detector-efficiency bounds for n-site Clauser-Horne inequalities, Phys. Rev. A 63, 022117 (2001).

    [24] M. G. A. Paris, Displacement operator by beam splitter, Phys. Lett. A217, 7880 (1996).

    [25] D. Walls and G. J. Milburn, Quantum Optics. Springer, 2nd ed., Feb.,2008.

    [26] S. A. Aljunid, B. Chng, M. Paesold, G. Maslennikov and C. Kurtsiefer,Interaction of light with a single atom in the strong focusing regime,Preprint at http://arxiv.org/abs/1006.2191 (2010).

    [27] S. E. Morrin, C. C. Yu and T. W. Mossberg, Strong Atom-Cavity Couplingover Large Volumes and the Observation of Subnatural Intracacvity AtomicLinewidths, Phys. Rev. Lett., 73, 1489 (1994)

    [28] K. Birnbaum, A. Boca, R. Miller, A. Boozer, T. Northup, and H. Kimble,Photon blockade in an optical cavity with one trapped atom, Nature436,8790 (2005).

    [29] C. J. Hood, Real-time Measurement and Trapping of Single Atoms by

    Single Photons PhD dissertation, California Institute of Technology (2000)[30] S. Ritter, C. Nlleke, C. Hahn, A. Reiserer, A. Neuzner, M. Uphoff,

    M. Mcke, E. Figueroa, J. Bochmann, and G. Rempe, An elementaryquantum network of single atoms in optical cavities, Nature 484, 195200(2012).

    20

    http://dx.doi.org/10.1103/PhysRevLett.23.880http://dx.doi.org/10.1103/PhysRevLett.23.880http://dx.doi.org/10.1103/PhysRevLett.23.880http://dx.doi.org/10.1103/PhysRevLett.23.880http://dx.doi.org/10.1103/PhysRevLett.23.880http://dx.doi.org/10.1103/PhysRevLett.98.220403http://dx.doi.org/10.1103/PhysRevLett.98.220403http://dx.doi.org/10.1103/PhysRevLett.98.220403http://dx.doi.org/10.1103/PhysRevLett.98.220402http://dx.doi.org/10.1103/PhysRevLett.98.220402http://dx.doi.org/10.1103/PhysRevLett.98.220402http://dx.doi.org/10.1103/PhysRevLett.98.220402http://dx.doi.org/10.1103/PhysRevLett.105.253001http://dx.doi.org/10.1103/PhysRevLett.105.253001http://dx.doi.org/10.1103/PhysRevLett.105.253001http://dx.doi.org/10.1103/PhysRevA.84.052122http://dx.doi.org/10.1103/PhysRevA.84.052122http://dx.doi.org/10.1103/PhysRevA.84.052122http://dx.doi.org/10.1103/PhysRevA.47.R747http://dx.doi.org/10.1103/PhysRevA.47.R747http://dx.doi.org/10.1103/PhysRevA.47.R747http://dx.doi.org/10.1103/PhysRevA.47.R747http://dx.doi.org/10.1103/PhysRevA.63.022117http://dx.doi.org/10.1103/PhysRevA.63.022117http://dx.doi.org/10.1103/PhysRevA.63.022117http://dx.doi.org/10.1103/PhysRevA.63.022117http://dx.doi.org/10.1016/0375-9601(96)00339-8http://dx.doi.org/10.1016/0375-9601(96)00339-8http://dx.doi.org/10.1016/0375-9601(96)00339-8http://dx.doi.org/10.1016/0375-9601(96)00339-8http://arxiv.org/abs/1006.2191http://arxiv.org/abs/1006.2191http://dx.doi.org/10.1038/nature03804http://dx.doi.org/10.1038/nature03804http://dx.doi.org/10.1038/nature03804http://dx.doi.org/10.1038/nature03804http://dx.doi.org/10.1038/nature03804http://dx.doi.org/10.1038/nature11023http://dx.doi.org/10.1038/nature11023http://dx.doi.org/10.1038/nature11023http://dx.doi.org/10.1038/nature11023http://dx.doi.org/10.1038/nature11023http://dx.doi.org/10.1038/nature11023http://dx.doi.org/10.1038/nature03804http://dx.doi.org/10.1038/nature03804http://arxiv.org/abs/1006.2191http://dx.doi.org/10.1016/0375-9601(96)00339-8http://dx.doi.org/10.1016/0375-9601(96)00339-8http://dx.doi.org/10.1103/PhysRevA.63.022117http://dx.doi.org/10.1103/PhysRevA.47.R747http://dx.doi.org/10.1103/PhysRevA.47.R747http://dx.doi.org/10.1103/PhysRevA.84.052122http://dx.doi.org/10.1103/PhysRevLett.105.253001http://dx.doi.org/10.1103/PhysRevLett.98.220402http://dx.doi.org/10.1103/PhysRevLett.98.220402http://dx.doi.org/10.1103/PhysRevLett.98.220403http://dx.doi.org/10.1103/PhysRevLett.23.880http://dx.doi.org/10.1103/PhysRevLett.23.880
  • 7/28/2019 NatComm Editor v2

    21/21

    [31] J. B. Brask and R. Chaves, Robust nonlocality tests with displacement-based measurements, Phys. Rev. A 86, 010103(R) (2012).

    [32] M. Boissonneault, J. M. Gambetta, and A. Blais, Dispersive regime of cir-cuit QED: Photon-dependent qubit dephasing and relaxation rates, Phys.Rev. A 79, 013819 (2009).

    Acknowledgements

    We would like to thank Ignacio Cirac for pointing out the issue of cooperativityand also Gerhard Rempe, Stephan Ritter, Serge Haroche, Christian Kurtsiefer,Gleb Maslennikov, Matthias U. Staudt, Antonio Badolato, Dario Gerace andSteve Girvin for useful discussions. This work was supported by the Brazilianagencies Fapemig, Capes, CNPq, and INCT-IQ, the Swiss National Science

    Foundation, the John Templeton Foundation, the National Research Foundationand the Ministry of Education of Singapore.

    Authors contributions

    M. Arajo, M. T. Quintino, D. Cavalcanti, M. T. Cunha, and M. F. Santosworked on the ideal case and Bell test proposal. C. Teo, J. Min, M. F. Santos,and V. Scarani, developed the calculations for the realistic case. All authorsdiscussed the results and wrote the manuscript.

    Competing financial interests

    The authors declare no competing financial interests.

    21

    http://pra.aps.org/abstract/PRA/v86/i1/e010103http://pra.aps.org/abstract/PRA/v86/i1/e010103http://pra.aps.org/abstract/PRA/v86/i1/e010103http://pra.aps.org/abstract/PRA/v86/i1/e010103http://dx.doi.org/10.1103/PhysRevA.79.013819http://dx.doi.org/10.1103/PhysRevA.79.013819http://dx.doi.org/10.1103/PhysRevA.79.013819http://dx.doi.org/10.1103/PhysRevA.79.013819http://dx.doi.org/10.1103/PhysRevA.79.013819http://dx.doi.org/10.1103/PhysRevA.79.013819http://pra.aps.org/abstract/PRA/v86/i1/e010103