Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Detection of genuine tripartite entanglement and steering in hybrid optomechanics

Open Access Open Access

Abstract

Multipartite quantum entanglement is a key resource for ensuring security in quantum network. We show that by using a unified parameter in terms of reduced noise variances one can determine different types of tripartite entanglement of a given state generated in a hybrid optomechanical system, where an atomic ensemble is located inside a single-mode cavity with a movable mirror, with different thresholds for each type. In particular, the special quantum states which allow both entanglement and steering genuinely shared among atom-light-mirror modes can be observed, even though there is no direct interaction between the mirror and the atomic ensemble. We further show the robustness against mechanical thermal noise and damping, the relaxation time of atomic ensemble, as well as the effect of gain factors involved in the criteria. Our analysis provides an experimentally achievable method to determine the type of tripartite quantum correlation in a way.

© 2015 Optical Society of America

1. Introduction

The multipartite entangled state has attracted great interest, mainly due to its fundamental significance for understanding the nature of transition from quantum to classical regime, and potential application in certain quantum information tasks involving more than two parties [1–4]. Van Loock and Furusawa first developed the concept of fully N–partite inseparable state for continuous variable systems [5], which is demonstrated if the system cannot be described by any one of the biseparable forms. Another important type of multipartite quantum correlations is known as genuine multipartite entanglement [6–12]. By definition, a total number of N systems are genuinely N–partite entangled if and only if the density operator of the N–partite system cannot be represented in any possible mixtures of the bipartition. Recently, Shalm et al. pointed out that the full N–partite inseparability does not always imply genuine N–partite entanglement for general mixed states [13]. They derived criteria for detecting genuine tripartite entanglement, and experimentally generated genuine three-photon entangled state using the nonlinear process of cascaded spontaneous parametric downconversion. In addition, Armstrong et al. reported the realization of genuine tripartite entanglement of three intense light beams with respect to the quadrature phase amplitudes of distinct fields, setting a milestone for testing mesoscopic quantum mechanics [14].

However, it has been realized that entanglement in terms of inseparability does not demonstrate either Bell nonlocality [15] or “steering” [16, 17]. The concept of steering was first introduced by Schrödinger [18] to describe the “spooky action-at-a-distance” nonlocality raised in the Einstein–Podolsky–Rosen (EPR) paradox [19]. In 2007, Wiseman, Jones and Doherty formalized the concept of steering, and pointed that it is fundamentally different from both entanglement and Bell’s nonlocality in terms of violations of local hidden state (LHS) models [16, 17]. In fact, steering gives a way to quantify how measurements by Alice on her local particle A can collapse the wave packet of Bob’s particle B. Note the inherent asymmetry of steering that a system can be steerable in one way but not the other [20–22]. The realization of steering in experiments reveals an EPR paradox [23, 24]. Besides, steering is also potentially less susceptible to noise and decoherence than the strongest form of nonlocality, Bell’s nonlocality, in which all local hidden variable (LHV) models are falsified [25]. Therefore, the generation and observation of steering is more accessible to experiment [20, 26–33]. What’s more, steering is closely related to certain quantum information task such as one-sided device-independent quantum key distribution [34].

Recently, the concept of genuine N–partite EPR steering has been developed by He and Reid [35], where a steering nonlocality is necessarily shared among all N parties. To demonstrate it, one must eliminate the steering created by mixing states with two-party steering across different bipartitions. In 2014, Teh and Reid derived various criteria that can be used to demonstrate genuine tripartite entanglement and steering [36]. We notice that a latest experimental demonstration of genuine multipartite EPR steering that was defined in a different way by Cavalcanti and his co-workers [37]. In their scenario, one can certify different kinds of multi-partite entanglement of the ordinal state via the implementation of quantum state tomography by trusted parties. In this work, we will follow the definition given in Ref. [14, 35, 36]. It has been shown that the special multipartite entangled and steerable states can be created in the multimode optical system [14, 38] and multimode pulsed cavity optomechanical system [39]. We then ask can we generate robust genuine multipartite entanglement and steering in other promising systems, e.g. in a hybrid massive system.

Thanks to the rapid progresses of experimental techniques, optomechanical systems cooled near their ground states [40–42] provide a new physical platform to study quantum effects in mesoscopic massive systems. Particularly, the hybrid systems combining nano-mechanical oscillator and atomic ensemble or Bose-Einstein condensate and propagating fields via radiation pressure offer new possibilities of local information processing and network construction [43–47]. Recently, the generation of bipartite or tripartite quantum entangled states in optomechanical systems has been proposed [48–57].

In this paper, we propose a scheme to generate the genuine tripartite entanglement and genuine tripartite steering in the context of pulsed cavity optomechanics, and test them by a unified signature in terms of an EPR-type variance which manifests as reduced noise variance. We consider an ensemble of N identical two-level atoms located inside an optical cavity with a movable mirror. The cavity mode is driven by a short laser pulse, which couples with the mirror by a nonlinear parametric type interaction and with the atomic ensemble via a linear beamsplitter-type interaction. Using a set of experimentally realistic parameters, we can easily distinguish various types of quantum correlations among light-mirror-atom modes in this system, including the tripartite inseparability, genuine tripartite entanglement, the steering of the mirror by the collaboration of cavity field and atomic mode, and genuine tripartite steering. We further study the effects of mechanical thermal noise and damping, the relaxation time of the atomic mode, and the gains included in the criteria on the observation of those tripartite quantum correlations.

2. Dynamics of the system

The schematic diagram of the three-mode atomic optomechanical system is illustrated in Fig. 1, which can be described by the following Hamiltonian [49, 52, 53]

H=Δcacac+ωmbmbm+Δacaca+g0acac(bm+bm)+ga(caac+acca)+i[E(t)acE*(t)ac].
The first three terms give the energies of cavity field mode ac, mirror mode bm, and atomic mode ca, where Δc = ωcωL, Δa = ωaωL are the detunings of the driving laser with respect to the cavity resonance and the atomic transition frequency, respectively. The fourth term denotes the optomechanical interaction, where g0 is the single-photon coupling strength. This interaction between the cavity and mechanical modes involves the nonlinear optomechanical coupling, which can create quantum entanglement. The fifth term indicates a beamsplitter-type interaction between the cavity mode and atomic mode with coupling constant ga. The last term represents the driving energy, where we assume the cavity is driven by a rectangular pulse with constant amplitude of E(t) over the short time interval τ and zero outside this interval. Note that there is no direct interaction between the mechanical mode and atomic mode.

 figure: Fig. 1

Fig. 1 Schematic diagram of a hybrid atomic optomechanical system. The cavity mode has a frequency ωc and a decay rate κ, driven by a laser pulse of frequency ωL. The movable mirror oscillates with frequency ωm and dissipation rate γm. Each of the N identical two-level atoms includes a ground state |gj〉 and an excited state |ej〉, with transition frequency ωa and relaxation time γa. The collective dipole lowering, raising, and population difference operators can be defined as S=j|gjej|, S+=j|ejgj|, and Sz=j(|ejej||gjgj|), respectively. We assume that all the atoms are initially prepared in their ground state, so that Sz ≃ 〈Sz〉 ≃ −N. At the same time, the initial assumption κga assures that the photons inside the cavity may escape from the cavity quickly before making an effective Rabi cycle between the two levels. Thus, most of atoms stay in the ground state such that the condition of low atomic excitation limit still stands. In this case, we may represent the collective dipole lowering and raising operators in terms of annihilation and creation operators ca=S/|Sz| and ca=S+/|Sz|, respectively.

Download Full Size | PDF

Following the standard linearization method [49,52,53], we find that under the rotating wave approximation and putting the blue detuned laser pulse to both cavity and atomic resonance, the slowly varying parts of the fluctuation operators satisfy the following motion equations

δa˙c=κδacigaδcaig(δbm+δbm)2κain,δb˙m=γmδbmig(δac+δac)2γmbin,δc˙a=γaδcaigaδac2γacin,
where g and ga are the effective coupling strengths of the cavity mode to the mirror and to the atomic mode, respectively. The Langevin noise operators corresponding to the light decay rates κ (associated with the coupling at the input cavity mirror, and assume no further loss processes inside the cavity), the mechanical dispassion γm, and the relaxation of atoms γa, have nonzero δ correlated functions, ain(t)ain(t)=(n¯a+1)δ(tt), bin(t)bin(t)=(n¯m+1)δ(tt), cin(t)cin(t)=(n¯c+1)δ(tt), i.e., the cavity field, the mirror, and the atomic mode are in a nonzero temperature T with mean number of the thermal photons or phonons a,m,c = [exp(h̄ωc,m,a/kBT) − 1]−1. Hereafter, we will drop the label δ on the fluctuation operators to simplify the notation.

In the bad cavity limit κga, g, the cavity mode can be adiabatically eliminated from the dynamics of the mirror and atomic modes to leave all modes depending only on the input quantities. The adiabatic solutions for the fluctuation operator of the cavity mode is

ac(t)igaκca(t)igκbm(t)2κain(t).
By substituting this result into the equations for m and ċa in Eq. (2), we get
c˙a=(Ga+γa)caGGabm+i2Gaain2γacin,b˙m=(Gγm)bm+GGacai2Gain2γmbin,
where G = g2/κ ( Ga=ga2/κ) is the effective coupling strength between the cavity mode and the mechanical (atomic) mode.

We may solve that set of equations by matrix diagonalization which gives two orthogonal superposition modes. The procedure is as follows. Equation (4) can be written in a matrix form as

Y˙(t)=MY(t)+η(t),
where Y(t)=(ca(t),bm)(t))T, the drift matrix M is given by
M=(GaγaGGaGGaGγm),
and η(t)=(i2Gaain2γacin,i2Gain2γmbin)T. A diagonalization of the matrix above results in two real eigenvalues λu=12[GGaγaγm+ε] and λv=12[GGaγaγmε], where ε=G2+(Ga+γaγm)22G(Gaγa+γm). The coefficients of the corresponding eigenvectors are (u1,u2)T=(1,(λuGaγa)/GGa)T and (v1,v2)T=(1,(λv+Ga+γa)/GGa)T. We can then directly obtain the time-dependent solutions of the modes ca(t) and bm(t) based on the two eigenvalues and their corresponding eigenvectors, as given by
ca(t)=eλutv2+eλvtu2v2+u2ca(0)+eλuteλvtv2+u2bm(0)+iv22Ga2Gv2+u2eλut0tdtain(t)eλut+iu22Ga+2Gv2+u2eλvt0tdtain(t)eλvt2γav2+u2[u2eλvt0tdtcin(t)eλvt+v2eλut0tdtcin(t)eλut]+2γmv2+u2[eλvt0tdtbin(t)eλvteλut0tdtbin(t)eλut],bm(t)=(eλuteλvt)u2v2v2+u2ca(0)+eλvtv2+eλutu2v2+u2bm(0)+iu2v22Ga2Gv2+u2eλut0tdtain(t)eλutiv2u22Ga+2Gv2+u2eλvt0tdtain(t)eλvt+u2v22γav2+u2[eλvt0tdtcin(t)eλvteλut0tdtcin(t)eλut]2γmv2+u2[u2eλut0tdtbin(t)eλut+v2eλvt0tdtbin(t)eλvt].
Substituting Eq. (7) into Eq. (3), we can get the time-dependent solution of the cavity field ac(t).

Equations (3) and (7) manifest that the motions of all the three modes get correlated to the input noises ain, bin, cin with exponentially shaped envelopes, indicating that it is convenient to introduce a set of normalized temporal pulse-shape amplitudes. We can introduce a set of normalized temporal modes and determine how each mode at the end of the pulse t = τ is related to the initial modes. All technical details are illustrated in the Appendix.

To explain how to define the temporal modes, we consider a simplified model where we neglect the existence of the mechanical and atomic decoherence effects, i.e., γa = γm = 0. Note that this approximation is valid when the total duration τ of protocol is much smaller than the effective mechanical and atomic decoherence time 1/γmm and 1/γac. By applying the standard cavity input-output relations, aout(t)=ain(t)+2κac(t), we can define a set of normalized temporal light modes (assuming G > Ga) [49, 52, 53],

Ain=2(GGa)1e2(GGa)τ0τdtain(t)e(GGa)t,Aout=2(GGa)e2(GGa)τ10τdtaout(t)e(GGa)t,
which obey the canonical commutation relations [Ai,Ai]=1 with i = (in, out). To evaluate the input and output quadratures of the oscillator and atomic ensemble, we define Bin = bm(0), Bout = bm(τ), Cin = ca(0), Cout = ca(τ). Then, we can determine how the output field of each mode is related to the input fields after an interaction time τ:
Aout=erAinie2r1(αBin+βCin),Bout=(α2erβ2)Bin+αβ(er1)Cin+iαe2r1Ain,Cout=(α2β2er)Cinαβ(er1)Bin+iβe2r1Ain,
where α=G/(GGa), β=Ga/(GGa) are the parameters related to the coupling strength, and r = (GGa)τ is the normalized interaction time parameter as an effective squeezing parameter. The general expressions for Aout, Bout, and Cout including mechanical damping γm, thermal noise n0, m, and relaxation time of atoms γa are given in the Appendix.

The simultaneous coupling of all three modes can result in a tripartite quantum correlation [52, 53]. We investigate the possibility to generate different types of quantum correlations (tripartite inseparability, genuine tripartite entanglement, tripartite steerability, and genuine tripartite steering) in this hybrid atomic optomechanics, and analyze the effects of imperfections on the quantum correlation. The type of quantum correlation can be determined by using the reduced noise variance of the quadrature of mirror mode based on the outcomes of measurements performed on the cavity field and atomic modes, compared with different thresholds for each type. By examining properties of the variances of suitable linear combinations of the quadrature operators for the three modes, from the solution for output field operators given by Eq. (9) and the general one with finite γm, γa and thermal environment noise m shown in Appendix Eq. (18), we find that the strongest correlations between the mirror mode and cavity field occur in the XP combination of the corresponding quadratures, whereas the strongest correlations between the mirror and atomic modes occur in the XX combination of the corresponding quadratures [52, 53]. With these combinations we can test an EPR-type variance parameter

Em|ac(g)=Δinf,ac2Xm+Δinf,ac2Pm=Δ2(Xm+gaPa+gcXc)+Δ2(Pm+haXa+hcPc),
which can be minimized by optimizing the gain factors ga,c and ha,c. Here, the quadrature operators of the output modes are defined by Xj=(Oout+Oout)/2 and Pj=(OoutOout)/2i, where j ∈ {a, m, c}, O ∈ {A, B, C}, with the Heisenberg uncertainty relation ΔXjΔPj ≥ 1/2. We can similarly define the quadratures for the input temporal modes, and express the quadrature components Xj and Pj of the output modes in terms of those associated with the input modes. We assume that the input modes are in thermal states characterized by the variances Δ2Xain=Δ2Pain=na0+1/2, Δ2Xcin=Δ2Pcin=nc0+1/2, and Δ2Xmin=Δ2Pmin=nm0+1/2, where the variance is defined as Δ2O = 〈O2〉 − 〈O2. Here, na0, nc0, and nm0 are the initial occupation numbers of thermal photons or phonons in the cavity field, atomic mode, and mirror, respectively.

3. Different types of tripartite entanglement

First, a tripartite system {m, n, l} is genuinely entangled iff the density operator of the system cannot be represented in the biseparable form ρ=P1iηi(1)ρmniρli+P2jηj(2)ρnljρmj+P3kηk(3)ρmlkρnk, where ∑j Pj = 1 and iηi(j)=1. Pj = {P1, P2, P3} are probabilities for the system to be in a state with given bipartition. The bipartition ρmnρl indicates the subsystems m and n may be entangled or not, but both of them are separated with subsystem l. To show the existence of genuine tripartite entanglement among light-mirror-atom (amc) in the proposed system, we need verify the elimination of all the possible mixtures of biseparable model by violating the following inequality [14, 36]

Em|ac(g)min{1+|gaha+gchc|,|gaha|+|1+gchc|,|gchc|+|1+gaha|}.

Second, to demonstrate the steering genuinely shared among three parties m, n, and l, we need to falsify the following description of the statistics based on the hybrid LHS model, where the averages are given as XmXnXl=P1iηi(1)XmXnXlρ+P2jηj(2)XnXlXmρ+P3kηk(3)XmXlXnρ [36]. The subscript ρ denotes that the average statistics is consistent with that of a quantum density matrix, i.e., 〈Xρ = Tr(ρX), and its uncertainty variances must satisfy the Heisenberg uncertainty relation. Whereas there are no such constraints for products of moments, written without the subscript. Thus, to demonstrate the genuine tripartite steering shared among light-mirror-atom, the following inequality should be violated to eliminate the steering created by mixing states with two-party steering across different bipartitions [14,35,36]

Em|ac(g)min{1,|gaha|,|gchc|}.

As the parameter Em|ac(g) must be smaller than the minimum of three bounds given by the three possible biseparable forms to demonstrate the quantum states that the entanglement or steering genuinely shared among three parties, it is in general difficult to realize in practical experiments. We can, however, still have the other two types of tripartite quantum correlation: tripartite inseparability, which is certified by the violation of the quantum separable state model [5, 14]

Em|ac(g)1+|gaha|+|gchc|,
and tripartite steering of the mirror mode m by the cavity mode a and atomic mode c, revealed by the violation of LHS model [14, 36]
Em|ac(g)1.

If genuine tripartite steering can be observed, the state necessarily acquires the property of tripartite inseparability, genuine tripartite entanglement, and steering of one mode by the other two. Note that all criteria above are sufficient but not necessary to certify the corresponding types of tripartite quantum correlation [58]. The parameter Em|ac(g) in each inequality can be minimized by different optimal gain factors ga,c and ha,c. In other words, one must set different gain factors varying with time in the experiments to achieve different types of tripartite quantum entanglement tested by the violation of inequalities (1114).

To save the trouble for real-time adjustment of gain factors in practice, we can test the reduced variance parameter with fixed gain factors ga=gc=1/2 and ha=hc=1/2, i.e. change inequality (10) to

Em|ac=Δ2(Xm+Pa+Xc2)+Δ2(Pm+XaPc2).
Then the three modes are verified to exhibit tripartite inseparability, genuine tripartite entanglement, tripartite steering of the mirror by the collaboration of the other two modes, and genuine tripartite quantum steering, if the EPR-type tripartite entanglement parameter given by Eq. (15) satisfy Em|ac < 2, Em|ac < 1, Em|ac < 1, and Em|ac < 0.5, respectively. As a result, one can simply test a single parameter to determine four types of tripartite entanglement of a given state, with different thresholds for each type. Notice that the linear combinations of the quadrature components involved in the variances satisfy the commutation relation [Xm+(Pa+Xc)/2,Pm+(XaPc)/2]=0, which indicates the variances of the position and momentum of the mirror mode can be inferred simultaneously with a 100% precision with Em|ac = 0, and hence demonstrate a perfect EPR paradox for the macroscopic and massive oscillator.

4. The effects of damping and thermal noise

Useful quantum resource must be immune to the dissipation and thermal noise. In this section we will investigate how the tripartite entanglement measured by the parameter shown in Eq. (15) is affected by the initial mechanical thermal noise nm0, and the decoherence induced by the reservoirs with decay rates γa, γm with or without additional mechanical thermal noise m. We will also show that the influence of the optimal gain factors ga,c, ha,c within the tripartite entanglement parameter Em|ac(g) given in Eq. (10). All those imperfections will decrease the created tripartite entanglement. However, we will see that within a proper parameter regime, all these detrimental effects can be significantly suppressed such that a maximal entanglement can be sustained.

4.1. Sensitivity to the initial thermal noise

We first illustrate the influence of the initial state preparation by considering the sensitivity to the initial thermal noise when the mirror and atomic modes are not coupled to reservoirs, i.e., γm = γa = 0. The tripartite entanglement parameter can be obtained as given by

Em|ac=[γe2r1er2]2(2na0+1)+[αγ(er1)α2e2r1+1]2(2nm0+1)+[βγ(er1)β2e2r1+12]2(2nc0+1),
where γ=αβ/2.

We find that the tripartite entanglement is less sensitive to the initial thermal noise in the steering modes na0 and nc0, and the influence is mainly induced by the initial thermal noise nm0 in the steered mechanical mode [53]. Hence, we can focus on the reduced tripartite entanglement parameter Em|ac=2(ere2r1)2(nm0+1) with na0 = nc0 = 0. Figure 2 illustrates the behavior of the tripartite entanglement with respect to the relative strength of the coupling constants r = (GGa)t and the initial thermal phonon number nm0 for a fixed coupling strength α=2. We see that even when the initial thermal noise is large, the tripartite entanglement parameter can be made negligibly small by increasing the squeezing parameter r. Tripartite inseparability (Em|ac < 2), genuine tripartite entanglement and the steering of mirror by the cavity field and atomic modes (Em|ac < 1), genuine tripartite steering (Em|ac < 0.5) can be observed when r>ln2+nm021+nm0, r>ln3+2nm022+2nm0, and r>ln5+4nm041+nm0, respectively. For nm0 ≫ 1, the thresholds of squeezing parameter are reduced as r>lnnm0/4, r<lnnm0/2, and r>lnnm0, respectively.

 figure: Fig. 2

Fig. 2 The effect of initial mechanical thermal noise nm0 on the EPR-type tripartite entanglement parameter Em|ac. Left: the 3D plot of Em|ac given in Eq. (16) as function of nm0 and squeezing parameter r = (GGA)τ for fixed coupling strength α=2; Right: the contour plot of Em|ac. The value of Em|ac smaller than the thresholds 2, 1, and 0.5, reveals the tripartite inseparability, genuine tripartite entanglement and the steering of the mirror by the cavity field and atomic modes, and genuine tripartite steering, respectively. Here, the cavity field and the atomic mode are assumed initially in a vacuum state na0 = nc0 = 0.

Download Full Size | PDF

4.2. Sensitivity to the decoherence

We then add the relaxation of the atom γa and the mechanical damping γm in expressions of output modes (see Appendix for the expressions with relaxation times), which usually arise because the system is coupled to a bath, to examine the robustness of the tripartite entanglement against the mechanical and atomic decoherence.

Figure 3 shows that both γa and γm in general have a destructive effect on the tripartite entanglement even in the absence of thermal noise in the reservoir m,a,c = 0 (i.e. temperature T = 0). However, it is interesting to see that in the presence of decoherence, different types of tripartite entanglement are preserved at smaller squeezing parameter r rather than at larger r. This can be understood by noticing that a strong squeezing leads to a high sensitivity of the variances to the external noise. Even for very large decay rates γa and γm, one still finds the observation of genuine tripartite steering Em|ac < 0.5 and genuine tripartite entanglement Em|ac < 1 at a relatively low level of squeezing nearby r ∼ 0.5. This indicates that the observation of genuine tripartite entanglement and steering immune to the mechanical damping to the heat bath becomes possible for lower and more accessible values of squeezing parameter r.

 figure: Fig. 3

Fig. 3 The effect of the relaxation of the atom γa when γm = 0 (a) and the mechanical damping γm when γa = 0 (b) on the EPR-type tripartite entanglement parameter Em|ac. Here, we set the coupling strength α=2, the initial thermal noise nm0,a0,c0 = 0 and the reservoir bath m,a,c = 0.

Download Full Size | PDF

A crucial requirement to quantum information processing is the preservation of multipartite entanglement within a noisy environment. Figure 4 shows the decoherence effect due to the decay of the mirror into a general thermal bath with nonzero excitation number m. As we expect, the detection of the tripartite entanglement is clearly more challenging. However, one can still finds that in the presence of the thermal noise or damping, the genuine tripartite steering is preserved at small r rather than at large r. This is also because the detection of entanglement is less sensitive to the decoherence for lower value of squeezing parameter.

 figure: Fig. 4

Fig. 4 Same to Fig. 2 but the mirror in contact with a nonzero temperature T reservoir with mean number of the thermal phonons m = nm0 = n, showing the decoherence effects due to mechanical coupling to a heat bath. Here, we set α=2, γm = 6 × 10−5κ, and γa = 10−4κ. The cavity field and the atomic modes are in the ordinary zero temperature environment a,c = na0,c0 = 0.

Download Full Size | PDF

5. The effect of optimal gain factors included in the criteria

We see that the test using the tripartite entanglement parameter Em|ac with fixed gains ha=gc=hc=1/2 is very sensitive to the mechanical thermal noise and decoherence. Those fixed gains are not the optimal factors to indicate the existence of tripartite quantum correlations. The following results however indicate a more promising prediction of four types of tripartite entanglement if the criteria given in inequalities (1114) involving an asymmetric choice of amplitude weightings are used.

The sensitivity of the tripartite entanglement to the decoherence decay rate γm, γa is greatly reduced, as shown in Fig. 5(a). With fixed gains, the reduced variance Em|ac goes up at larger squeezing parameter r because of a large sensitivity to the decoherence for a strong squeezing, as shown by the blue dashed curve. Therefore, it may lead to missing the test of the existence of tripartite entanglement. However, Em|ac(g) → 0 (which manifests a perfect EPR paradox for the mirror mode) can be observed over a wide range of large r with optimal gain factors ga,c, ha,c (shown in the inset of the figure). This makes the detection of tripartite entanglement becomes more feasible.

To confirm the presence of the four types tripartite entanglement, we need to compare the value of Em|ac(g) (black solid) with four different thresholds for each type with respect to the gain factors, to get falsification of inequalities (1114). Notice from Fig. 5(b) that with optimal choice of gains, we can always produce inseparable tripartite entangled state (evidenced by the black curve is always lower than the red solid curve), genuine tripartite entangled state (confirmed by the black curve is lower than the cyan solid curve), the tripartite steering of the mirror mode by the measurements on the cavity field and atomic modes (tested by the blue dotted bound) in the entire range of r > 0. The genuine tripartite steering (lower than the yellow dashed bound) can be also observed for a relatively relaxed range of r compared with that showing Em|ac < 0.5 (blue dashed curve) in Fig. 5(a).

 figure: Fig. 5

Fig. 5 (a) The effect of gains in inequalities (1114) on the tripartite entanglement parameter Em|ac(g) (black solid) comparing with the parameter Em|ac shown in Fig. 3 when n = 0 (blue dashed). The inset shows the optimal gains that can minimize the reduced noise variance of the mirror Em|ac(g), in terms of the measurements performed on the cavity field and atomic modes. (b) the tripartite entanglement parameter Em|ac(g) (black solid) versus four different thresholds determined by the corresponding gains for each type of tripartite entanglement. Here, we set α=2, γm = 6 × 10−5κ, and γa = 10−4κ, a,m,c = na0,m0,c0 = 0.

Download Full Size | PDF

According to different experimental systems, we could choose different gain factors to achieve the best genuine tripartite entangled and steerable states. But considering the implementation of experiments, optimizing all gain factors varying with time brings significant complication. In the scheme discussed here, choosing a combination of ga=ha=1/2 and gc=hc=1/2, one can simply measure a single parameter Em|ac to distinguish which kind of tripartite quantum correlations created in the experiment, with different bounds 2, 1, 0.5 for each type. In fact, the optimal gain factors to minimize Em|ac(g) are close to ±1/2, as shown in the inset of Fig. 5(a), so that setting ga=ha=1/2 and gc=hc=1/2 is a good choice to get a smaller value of Em|ac. We can see that from the blue curve in Fig. 5(a), for lower, but more accessible values of squeezing parameter r, one can observe all four types tripartite quantum entanglement and determine the type of tripartite entangled states by simply testing a unified parameter Em|ac, with different thresholds for each type.

6. Conclusion

In summary, we demonstrate a scheme for the generation of tripartite entangled states showing tripartite inseparability, genuine tripartite entanglement, steering of the mirror mode by the cavity field and atomic modes, and genuine tripartite steering in a hybrid atomic optomechanical system. Using the linearization approach, we derive analytical expressions for the input-output relations between the quadratures of the three modes in the presence of mechanical damping, the relaxation of atoms, and thermal noises. We can determine the type of the tripartite entanglement in terms of reduced noise variance of the mirror mode by the measurements performed on the other two modes, with different thresholds for each type. This allow us to distinguish various types of tripartite quantum correlations of a given state in a way that is easily measured in experiment. We also analyze the robustness of tripartite entanglement against different kinds of imperfect factors. The realization of the proposed scheme may be helpful for the study of quantum interfaces and quantum memories [59].

Appendix: the solutions for the output modes with decoherence

We now introduce a set of normalized temporal atomic output and input modes

Ain1=Γ10τdtain(t)eλut,Ain2=Γ20τdtain(t)eλvt,Ain3=Γ30τdtain(t)eλut,Bm1=Γ10τdtbin(t)eλut,Bm2=Γ20τdtbin(t)eλvt,Bm3=Γ30τdtbin(t)eλut,C1=Γ10τdtcin(t)eλut,C2=Γ20τdtcin(t)eλvt,C3=Γ30τdtcin(t)eλut,
where Γ1=2λu/(1e2λuτ), Γ2=2λv/(1e2λvτ), Γ3=2λu/(e2λuτ1).

To evaluate the input and output quadratures of the oscillator and atomic ensemble, we define Bin = bm(0), Bout = bm(τ), Cin = ca(0), Cout = ca(τ). Then, we can determine how the output field of each mode is related to the input fields after an interaction time τ:

Cout=eλuτv2+eλvτu2v2+u2Cin+eλuτeλvτv2+u2Bin+iv22Ga2GΓ3(v2+u2)Ain1+ieλvτ(u22Ga+2G)Γ2(v2+u2)Ain21Γ3(v2+u2)(v22γaC1+2γmBm1)+eλvτΓ2(v2+u2)(2γnBm2u22γaC2),Bout=(eλuτeλvτ)u2v2v2+u2Cin+eλvτv2+eλvτu2v2+u2Biniu2v22Gau22GΓ3(v2+u2)Ain1+ieλvτ(v2u22Ga+v22G)Γ2(v2+u2)Ain2u2Γ3(v2+u2)(v22γaC1+2γmBm1)+v2eλvτΓ2(v2+u2)(u22γaC22γmBm2),Aout=1u2+v2(i2f1v2Γ3+i2f2u2Γ3Γ5)Cin+1u2+v2(i2f1Γ3i2f2Γ3Γ5)Bin+eλuτλu(u2+v2)(h1Ain1if1v2γaC1if1γmBm1)+1λu(u2+v2)(h1Ain3+if1v2γaC3+if1γmBm3)+2eλuτΓ4(λu+λv)(u2+v2)(h2Ain2if2u2γaC2+if2γmBm2)+2(λu+λv)(u2+v2)(h2Ain3if2u2γaC3if2γmBm3)Ain3,
where Γ4=λu(e2λvτ1)λv(e2λuτ1), Γ5=e(λu+λv)τ1λu+λv, f1=Gu2Ga, f2=Gv2Ga, h1=Gav2Gu2GGa(1v2u2), h2=Gau2+Gv2GGa(1v2u2).

Acknowledgments

This work is supported by the National Natural Science Foundation of China (NSFC) under Grants No. 11274025 and No. 61475006. We thank fruitful discussions with Z. Ficek and M. D. Reid on the properties of multipartite steering. Q.Y.H. also thanks Prof. C. P. Sun at Beijing Computational Science Research Center for hospitality.

References and links

1. M. Hillery, V. Bužek, and A. Berthiaume, “Quantum secret sharing,” Phys. Rev. A 59, 1829–1834 (1999). [CrossRef]  

2. R. Cleve, D. Gottesman, and H.-K. Lo, “How to share a quantum secret,” Phys. Rev. Lett. 83, 648–651 (1999). [CrossRef]  

3. P. van Loock and S. L. Braunstein, “Multipartite entanglement for continuous variables: A quantum teleportation network,” Phys. Rev. Lett. 84, 3482–3485 (2000). [CrossRef]   [PubMed]  

4. H. Yonezawa, T. Aoki, and A. Furusawa, “Demonstration of a quantum teleportation network for continuous variables,” Nature 431, 430–433 (2004). [CrossRef]   [PubMed]  

5. P. van Loock and A. Furusawa, “Detecting genuine multipartite continuous-variable entanglement,” Phys. Rev. A 67, 052315 (2003). [CrossRef]  

6. R. Gallego, L. E. Würflinger, A. Acín, and M. Navascués, “Operational framework for nonlocality,” Phys. Rev. Lett. 109, 070401 (2012). [CrossRef]   [PubMed]  

7. J. D. Bancal, N. Gisin, Y. C. Liang, and S. Pironio, “Device-independent witnesses of genuine multipartite entanglement,” Phys. Rev. Lett. 106, 250404 (2011). [CrossRef]   [PubMed]  

8. S. M. Giampaolo and B. C. Hiesmayr, “Genuine multipartite entanglement in the XY model,” Phys. Rev. A 88, 052305 (2013). [CrossRef]  

9. J. D. Bancal, J. Barrett, N. Gisin, and S. Pironio, “Definitions of multipartite nonlocality,” Phys. Rev. A 88, 014102 (2013). [CrossRef]  

10. F. Levi and F. Mintert, “Hierarchies of multipartite entanglement,” Phys. Rev. Lett. 110, 150402 (2013). [CrossRef]   [PubMed]  

11. D. L. Deng, Z. S. Zhou, and J. L. Chen, “Svetlichny’s approach to detecting genuine multipartite entanglement in arbitrarily-high-dimensional systems by a Bell-type inequality,” Phys. Rev. A 80, 022109 (2009). [CrossRef]  

12. M. Huber, H. Schimpf, A. Gabriel, C. Spengler, D. Brup, and B. C. Hiesmayr, “Experimentally implementable criteria revealing substructures of genuine multipartite entanglement,” Phys. Rev. A 83, 022328 (2011). [CrossRef]  

13. L. K. Shalm, D. R. Hamel, Z. Yan, C. Simon, K. J. Resch, and T. Jennewein, “Three-photon energy–time entanglement,” Nat. Phys. 9, 19–22 (2013). [CrossRef]  

14. S. Armstrong, M. Wang, R. Y. Teh, Q. H. Gong, Q. Y. He, J. Janousek, H. - A. Bachor, M D. Reid, and P. K. Lam, “Multipartite Einstein–Podolsky–Rosen steering and genuine tripartite entanglement with optical networks,” Nat. Phys. 11, 167–172 (2015). [CrossRef]  

15. R. F. Werner, “Quantum states with Einstein-Podolsky-Rosen correlations admitting a hidden-variable model,” Phys. Rev. A 40, 4277–4281 (1989). [CrossRef]   [PubMed]  

16. H. M. Wiseman, S. J. Jones, and A. C. Doherty, “Steering, entanglement, nonlocality, and the Einstein-Podolsky-Rosen paradox,” Phys. Rev. Lett. 98, 140402 (2007). [CrossRef]   [PubMed]  

17. S. J. Jones, H. M. Wiseman, and A. C. Doherty, “Entanglement, Einstein-Podolsky-Rosen correlations, Bell nonlocality, and steering,” Phys. Rev. A 76, 052116 (2007). [CrossRef]  

18. E. Schrödinger, “Discussion of probability relations between separated systems,” Proc. Cambridge Philos. Soc. 31, 555–563 (1935). [CrossRef]  

19. A. Einstein, B. Podolsky, and N. Rosen, “Can quantum-mechanical description of physical reality be considered complete?” Phys. Rev. 47, 777–780 (1935). [CrossRef]  

20. V. Händchen, T. Eberle, S. Steinlechner, A. Samblowski, T. Franz, R. F. Werner, and R. Schnabel, “Observation of one-way Einstein–Podolsky–Rosen steering,” Nat. Photonics 6, 596–599 (2012). [CrossRef]  

21. Q. Y. He, Q. H. Gong, and M. D. Reid, “Classifying directional Gaussian entanglement, Einstein-Podolsky-Rosen steering, and discord,” Phys. Rev. Lett. 114, 060402 (2015). [CrossRef]   [PubMed]  

22. I. Kogias, A. R. Lee, S. Ragy, and G. Adesso, “Quantification of Gaussian quantum steering,” Phys. Rev. Lett. 114, 060403 (2015). [CrossRef]   [PubMed]  

23. M. D. Reid, “Demonstration of the Einstein-Podolsky-Rosen paradox using nondegenerate parametric amplification,” Phys. Rev. A 40, 913–923 (1989). [CrossRef]   [PubMed]  

24. E. G. Cavalcanti, S. J. Jones, H. M. Wiseman, and M. D. Reid, “Experimental criteria for steering and the Einstein-Podolsky-Rosen paradox,” Phys. Rev. A 80, 032112 (2009). [CrossRef]  

25. J. S. Bell, “On the Einstein–Podolsky–Rosen paradox,” Physics 1, 195–200 (1964).

26. B. Hage, A. Samblowski, and R. Schnabel, “Towards Einstein-Podolsky-Rosen quantum channel multiplexing,” Phys. Rev. A 81, 062301 (2010). [CrossRef]  

27. D. J. Saunders, S. J. Jones, H. M. Wiseman, and G. J. Pryde, “Experimental EPR-steering using Bell-local states,” Nat. Phys. 6, 845–849 (2010). [CrossRef]  

28. D. H. Smith, G. Gillett, M. P. de Almeida, C. Branciard, A. Fedrizzi, T. J. Weinhold, A. Lita, B. Calkins, T. Gerrits, H. M. Wiseman, S. W. Nam, and A. G. White, “Conclusive quantum steering with superconducting transition-edge sensors,” Nat. Commun. 3, 625 (2012). [CrossRef]   [PubMed]  

29. A. J. Bennet, D. A. Evans, D. J. Saunders, C. Branciard, E. G. Cavalcanti, H. M. Wiseman, and G. J. Pryde, “Arbitrarily loss-tolerant Einstein-Podolsky-Rosen steering allowing a demonstration over 1 km of optical fiber with no detection loophole,” Phys. Rev. X 2, 031003 (2012).

30. B. Wittmann, S. Ramelow, F. Steinlechner, N. K. Langford, N. Brunner, H. M. Wiseman, R. Ursin, and A. Zeilinger, “Loophole-free Einstein–Podolsky–Rosen experiment via quantum steering,” New J. Phys. 14, 053030 (2012). [CrossRef]  

31. S. Steinlechner, J. Bauchrowitz, T. Eberle, and R. Schnabel, “Strong Einstein-Podolsky-Rosen steering with unconditional entangled states,” Phys. Rev. A 87, 022104 (2013). [CrossRef]  

32. K. Sun, J. S. Xu, X. J. Ye, Y. C. Wu, J. L. Chen, C. F. Li, and G. C. Guo, “Experimental demonstration of the Einstein-Podolsky-Rosen steering game based on the all-versus-nothing proof,” Phys. Rev. Lett. 113, 140402 (2014). [CrossRef]   [PubMed]  

33. C. M. Li, K. Chen, Y. N. Chen, Q. Zhang, Y. A. Chen, and J. W. Pan, “Genuine high-order Einstein-Podolsky-Rosen steering,” Phys. Rev. Lett. 115, 010402 (2015). [CrossRef]   [PubMed]  

34. C. Branciard, E. G. Cavalcanti, S. P. Walborn, V. Scarani, and H. M. Wiseman, “One-sided device-independent quantum key distribution: Security, feasibility, and the connection with steering,” Phys. Rev. A 85, 010301(R) (2012). [CrossRef]  

35. Q. Y. He and M. D. Reid, “Genuine multipartite Einstein-Podolsky-Rosen steering,” Phys. Rev. Lett. 111, 250403 (2013). [CrossRef]  

36. R. Y. Teh and M. D. Reid, “Criteria for genuine N-partite continuous-variable entanglement and Einstein-Podolsky-Rosen steering,” Phys. Rev. A 90, 062337 (2014). [CrossRef]  

37. D. Cavalcanti, P. Skrzypczyk, G. H. Aguilar, R. V. Nery, P. H. Souto Ribeiro, and S. P. Walborn, “Detection of entanglement in asymmetric quantum networks and multipartite quantum steering,” Nat. Commun. 6, 7941 (2015). [CrossRef]   [PubMed]  

38. M. Wang, Q. H. Gong, and Q.Y. He, “Collective multipartite Einstein-Podolsky-Rosen steering: more secure optical networks,” Opt. Lett. 39, 6703–6706 (2014). [CrossRef]   [PubMed]  

39. M. Wang, Q. H. Gong, Z. Ficek, and Q. Y. He, “Efficient scheme for perfect collective Einstein-Podolsky-Rosen steering,” Sci. Rep. 5, 12346 (2015). [CrossRef]   [PubMed]  

40. A. D. O’Connell, M. Hofheinz, M. Ansmann, R. C. Bialczak, M. Lenander, E. Lucero, M. Neeley, D. Sank, H. Wang, M. Weides, J. Wenner, J. M. Martinis, and A. N. Cleland, “Quantum ground state and single-phonon control of a mechanical resonator,” Nature 464, 697–703 (2010). [CrossRef]  

41. J. Chan, T. P. Mayer Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Krause, S. Gröblacher, M. Aspelmeyer, and O. Painter, “Laser cooling of a nanomechanical oscillator into its quantum ground state,” Nature 478, 89–92 (2011). [CrossRef]   [PubMed]  

42. J. D. Teufel, T. Donner, D. L Li, J. W. Harlow, M. S. Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, K. W. Lehnert, and R. W. Simmonds, “Sideband cooling of micromechanical motion to the quantum ground state,” Nature 475, 359–363 (2011). [CrossRef]   [PubMed]  

43. F. Brennecke, S. Ritter, T. Donner, and T. Esslinger, “Cavity optomechanics with a Bose-Einstein condensate,” Science 322, 235–238 (2008). [CrossRef]   [PubMed]  

44. T. P. Purdy, D. W. C. Brooks, T. Botter, N. Brahms, Z.-Y. Ma, and D. M. Stamper-Kurn, “Tunable cavity optomechanics with ultracold atoms,” Phys. Rev. Lett. 105, 133602 (2010). [CrossRef]  

45. R. Kanamoto and P. Meystre, “Optomechanics of a quantum-degenerate Fermi gas,” Phys. Rev. Lett. 104, 063601 (2010). [CrossRef]  

46. S. Singh, H. Jing, E. M. Wright, and P. Meystre, “Quantum-state transfer between a Bose-Einstein condensate and an optomechanical mirror,” Phys. Rev. A 86, 021801(R) (2012). [CrossRef]  

47. A. Carmele, B. Vogell, K. Stannigel, and P. Zoller, “Opto-nanomechanics strongly coupled to a Rydberg super-atom: coherent versus incoherent dynamics,” New J. Phys. 16, 063042 (2014). [CrossRef]  

48. C. Genes, D. Vitali, and P. Tombesi, “Emergence of atom-light-mirror entanglement inside an optical cavity,” Phys. Rev. A 77, 050307(R) (2008). [CrossRef]  

49. S. G. Hofer, W. Wieczorek, M. Aspelmeyer, and K. Hammerer, “Quantum entanglement and teleportation in pulsed cavity optomechanics,” Phys. Rev. A 84, 052327 (2011). [CrossRef]  

50. L. H. Sun, G. X. Li, and Z. Ficek, “First-order coherence versus entanglement in a nanomechanical cavity,” Phys. Rev. A 85, 022327 (2012). [CrossRef]  

51. L. Tian, “Robust photon entanglement via quantum interference in optomechanical interfaces,” Phys. Rev. Lett. 110, 233602 (2013). [CrossRef]   [PubMed]  

52. Q. Y. He and Z. Ficek, “Einstein-Podolsky-Rosen paradox and quantum steering in a three-mode optomechanical system,” Phys. Rev. A 89, 022332 (2014). [CrossRef]  

53. M. Wang, Q. H. Gong, Z. Ficek, and Q.Y. He, “Role of thermal noise in tripartite quantum steering,” Phys. Rev. A 90, 023801 (2014). [CrossRef]  

54. Y. Yan, G. X. Li, and Q. L. Wu, “Entanglement and Einstein-Podolsky-Rosen steering between a nanomechanical resonator and a cavity coupled with two quantum dots,” Opt. Express 23, 21306–21322 (2015). [CrossRef]   [PubMed]  

55. Y. D. Wang, S. Chesi, and A. A. Clerk, “Bipartite and tripartite output entanglement in three-mode optomechanical systems,” Phys. Rev. A 91, 013807 (2015). [CrossRef]  

56. Q. Lin and B. He, “Optomechanical entanglement under pulse drive,” Opt. Express 23, 24497–24507 (2015). [CrossRef]   [PubMed]  

57. T. A. Palomaki, J. D. Teufel, R. W. Simmonds, and K. W. Lehnert, “Entangling mechanical motion with microwave fields,” Science 342, 710–713 (2013). [CrossRef]   [PubMed]  

58. For two-mode Gaussian states, the condition EA|B(g) < 1 is necessary and sufficient to confirm steering of A by B [16, 17, 24], and the minimum value of EA|B(g) can be used to quantify the Gaussian EPR steering [22].

59. K. Hammerer, A. S. Sørensen, and E. S. Polzik, “Quantum interface between light and atomic ensembles,” Rev. Mod. Phys. 82, 1041 (2010). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Schematic diagram of a hybrid atomic optomechanical system. The cavity mode has a frequency ωc and a decay rate κ, driven by a laser pulse of frequency ωL. The movable mirror oscillates with frequency ωm and dissipation rate γm. Each of the N identical two-level atoms includes a ground state |gj〉 and an excited state |ej〉, with transition frequency ωa and relaxation time γa. The collective dipole lowering, raising, and population difference operators can be defined as S = j | g j e j |, S + = j | e j g j |, and S z = j ( | e j e j | | g j g j | ), respectively. We assume that all the atoms are initially prepared in their ground state, so that Sz ≃ 〈Sz〉 ≃ −N. At the same time, the initial assumption κga assures that the photons inside the cavity may escape from the cavity quickly before making an effective Rabi cycle between the two levels. Thus, most of atoms stay in the ground state such that the condition of low atomic excitation limit still stands. In this case, we may represent the collective dipole lowering and raising operators in terms of annihilation and creation operators c a = S / | S z | and c a = S + / | S z |, respectively.
Fig. 2
Fig. 2 The effect of initial mechanical thermal noise nm0 on the EPR-type tripartite entanglement parameter Em|ac. Left: the 3D plot of Em|ac given in Eq. (16) as function of nm0 and squeezing parameter r = (GGA)τ for fixed coupling strength α = 2; Right: the contour plot of Em|ac. The value of Em|ac smaller than the thresholds 2, 1, and 0.5, reveals the tripartite inseparability, genuine tripartite entanglement and the steering of the mirror by the cavity field and atomic modes, and genuine tripartite steering, respectively. Here, the cavity field and the atomic mode are assumed initially in a vacuum state na0 = nc0 = 0.
Fig. 3
Fig. 3 The effect of the relaxation of the atom γa when γm = 0 (a) and the mechanical damping γm when γa = 0 (b) on the EPR-type tripartite entanglement parameter Em|ac. Here, we set the coupling strength α = 2, the initial thermal noise nm0,a0,c0 = 0 and the reservoir bath m,a,c = 0.
Fig. 4
Fig. 4 Same to Fig. 2 but the mirror in contact with a nonzero temperature T reservoir with mean number of the thermal phonons m = nm0 = n, showing the decoherence effects due to mechanical coupling to a heat bath. Here, we set α = 2, γm = 6 × 10−5κ, and γa = 10−4κ. The cavity field and the atomic modes are in the ordinary zero temperature environment a,c = na0,c0 = 0.
Fig. 5
Fig. 5 (a) The effect of gains in inequalities (1114) on the tripartite entanglement parameter Em|ac(g) (black solid) comparing with the parameter Em|ac shown in Fig. 3 when n = 0 (blue dashed). The inset shows the optimal gains that can minimize the reduced noise variance of the mirror Em|ac(g), in terms of the measurements performed on the cavity field and atomic modes. (b) the tripartite entanglement parameter Em|ac(g) (black solid) versus four different thresholds determined by the corresponding gains for each type of tripartite entanglement. Here, we set α = 2, γm = 6 × 10−5κ, and γa = 10−4κ, a,m,c = na0,m0,c0 = 0.

Equations (18)

Equations on this page are rendered with MathJax. Learn more.

H = Δ c a c a c + ω m b m b m + Δ a c a c a + g 0 a c a c ( b m + b m ) + g a ( c a a c + a c c a ) + i [ E ( t ) a c E * ( t ) a c ] .
δ a ˙ c = κ δ a c i g a δ c a i g ( δ b m + δ b m ) 2 κ a in , δ b ˙ m = γ m δ b m i g ( δ a c + δ a c ) 2 γ m b in , δ c ˙ a = γ a δ c a i g a δ a c 2 γ a c in ,
a c ( t ) i g a κ c a ( t ) i g κ b m ( t ) 2 κ a in ( t ) .
c ˙ a = ( G a + γ a ) c a GG a b m + i 2 G a a in 2 γ a c in , b ˙ m = ( G γ m ) b m + GG a c a i 2 G a in 2 γ m b in ,
Y ˙ ( t ) = M Y ( t ) + η ( t ) ,
M = ( G a γ a GG a GG a G γ m ) ,
c a ( t ) = e λ u t v 2 + e λ v t u 2 v 2 + u 2 c a ( 0 ) + e λ u t e λ v t v 2 + u 2 b m ( 0 ) + i v 2 2 G a 2 G v 2 + u 2 e λ u t 0 t d t a in ( t ) e λ u t + i u 2 2 G a + 2 G v 2 + u 2 e λ v t 0 t d t a in ( t ) e λ v t 2 γ a v 2 + u 2 [ u 2 e λ v t 0 t d t c in ( t ) e λ v t + v 2 e λ u t 0 t d t c in ( t ) e λ u t ] + 2 γ m v 2 + u 2 [ e λ v t 0 t d t b in ( t ) e λ v t e λ u t 0 t d t b in ( t ) e λ u t ] , b m ( t ) = ( e λ u t e λ v t ) u 2 v 2 v 2 + u 2 c a ( 0 ) + e λ v t v 2 + e λ u t u 2 v 2 + u 2 b m ( 0 ) + i u 2 v 2 2 G a 2 G v 2 + u 2 e λ u t 0 t d t a in ( t ) e λ u t i v 2 u 2 2 G a + 2 G v 2 + u 2 e λ v t 0 t d t a in ( t ) e λ v t + u 2 v 2 2 γ a v 2 + u 2 [ e λ v t 0 t d t c in ( t ) e λ v t e λ u t 0 t d t c in ( t ) e λ u t ] 2 γ m v 2 + u 2 [ u 2 e λ u t 0 t d t b in ( t ) e λ u t + v 2 e λ v t 0 t d t b in ( t ) e λ v t ] .
A in = 2 ( G G a ) 1 e 2 ( G G a ) τ 0 τ d t a in ( t ) e ( G G a ) t , A out = 2 ( G G a ) e 2 ( G G a ) τ 1 0 τ d t a out ( t ) e ( G G a ) t ,
A out = e r A in i e 2 r 1 ( α B in + β C in ) , B out = ( α 2 e r β 2 ) B in + α β ( e r 1 ) C in + i α e 2 r 1 A in , C out = ( α 2 β 2 e r ) C in α β ( e r 1 ) B in + i β e 2 r 1 A in ,
E m | ac ( g ) = Δ inf , ac 2 X m + Δ inf , ac 2 P m = Δ 2 ( X m + g a P a + g c X c ) + Δ 2 ( P m + h a X a + h c P c ) ,
E m | ac ( g ) min { 1 + | g a h a + g c h c | , | g a h a | + | 1 + g c h c | , | g c h c | + | 1 + g a h a | } .
E m | ac ( g ) min { 1 , | g a h a | , | g c h c | } .
E m | ac ( g ) 1 + | g a h a | + | g c h c | ,
E m | ac ( g ) 1 .
E m | ac = Δ 2 ( X m + P a + X c 2 ) + Δ 2 ( P m + X a P c 2 ) .
E m | ac = [ γ e 2 r 1 e r 2 ] 2 ( 2 n a 0 + 1 ) + [ α γ ( e r 1 ) α 2 e 2 r 1 + 1 ] 2 ( 2 n m 0 + 1 ) + [ β γ ( e r 1 ) β 2 e 2 r 1 + 1 2 ] 2 ( 2 n c 0 + 1 ) ,
A in 1 = Γ 1 0 τ d t a in ( t ) e λ u t , A in 2 = Γ 2 0 τ d t a in ( t ) e λ v t , A in 3 = Γ 3 0 τ d t a in ( t ) e λ u t , B m 1 = Γ 1 0 τ d t b in ( t ) e λ u t , B m 2 = Γ 2 0 τ d t b in ( t ) e λ v t , B m 3 = Γ 3 0 τ d t b in ( t ) e λ u t , C 1 = Γ 1 0 τ d t c in ( t ) e λ u t , C 2 = Γ 2 0 τ d t c in ( t ) e λ v t , C 3 = Γ 3 0 τ d t c in ( t ) e λ u t ,
C out = e λ u τ v 2 + e λ v τ u 2 v 2 + u 2 C in + e λ u τ e λ v τ v 2 + u 2 B in + i v 2 2 G a 2 G Γ 3 ( v 2 + u 2 ) A in 1 + i e λ v τ ( u 2 2 G a + 2 G ) Γ 2 ( v 2 + u 2 ) A in 2 1 Γ 3 ( v 2 + u 2 ) ( v 2 2 γ a C 1 + 2 γ m B m 1 ) + e λ v τ Γ 2 ( v 2 + u 2 ) ( 2 γ n B m 2 u 2 2 γ a C 2 ) , B out = ( e λ u τ e λ v τ ) u 2 v 2 v 2 + u 2 C in + e λ v τ v 2 + e λ v τ u 2 v 2 + u 2 B in i u 2 v 2 2 G a u 2 2 G Γ 3 ( v 2 + u 2 ) A in 1 + i e λ v τ ( v 2 u 2 2 G a + v 2 2 G ) Γ 2 ( v 2 + u 2 ) A in 2 u 2 Γ 3 ( v 2 + u 2 ) ( v 2 2 γ a C 1 + 2 γ m B m 1 ) + v 2 e λ v τ Γ 2 ( v 2 + u 2 ) ( u 2 2 γ a C 2 2 γ m B m 2 ) , A out = 1 u 2 + v 2 ( i 2 f 1 v 2 Γ 3 + i 2 f 2 u 2 Γ 3 Γ 5 ) C in + 1 u 2 + v 2 ( i 2 f 1 Γ 3 i 2 f 2 Γ 3 Γ 5 ) B in + e λ u τ λ u ( u 2 + v 2 ) ( h 1 A in 1 i f 1 v 2 γ a C 1 i f 1 γ m B m 1 ) + 1 λ u ( u 2 + v 2 ) ( h 1 A in 3 + i f 1 v 2 γ a C 3 + i f 1 γ m B m 3 ) + 2 e λ u τ Γ 4 ( λ u + λ v ) ( u 2 + v 2 ) ( h 2 A in 2 i f 2 u 2 γ a C 2 + i f 2 γ m B m 2 ) + 2 ( λ u + λ v ) ( u 2 + v 2 ) ( h 2 A in 3 i f 2 u 2 γ a C 3 i f 2 γ m B m 3 ) A in 3 ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.