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

Cooling mechanical motion via vacuum effect of an ensemble of quantum emitters

Open Access Open Access

Abstract

We design a hybrid optomechanical setup, in which an ensemble of quantum emitters is coupled with a movable mirror through vacuum interaction. The optical cavity is driven along with the quantum emitters and therefore the coupling between the cavity field and the ensemble determines the dynamics of the coupled system. In particular, we investigated the influence of the vacuum coupling strength on the effective frequency and the effective damping rate of the movable mirror, which shows that the vacuum interaction enhances greatly the effective damping rate. Further, the cooling characteristics of the mechanical resonator is analyzed in detail by counting the effective phonon number in the mirror’s motion. It is found that the ground-state cooling of the mechanical motion can be approached in the bad cavity limit when the vacuum coupling is included. The dependence of the cooling of the mechanical motion on the parameters of the cavity and the quantum emitter is investigated in detail numerically.

© 2015 Optical Society of America

1. Introduction

Recently, Cavity optomechanical system with movable mechanical oscillator has received extensive attention, in which a significant interaction between the light and the mechanical system via radiation pressure can be generated [1–3]. The optomechancial system can be used for demonstrating the quantum-mechanical effect of macroscopic resonator, such as the generation of the continuous-variable optomechanical entanglement [4–11], quantum squeezing and optical cooling of mechanical mode [12–20], and non-classical state preparation in mechanical resonators [21–24]. The classical dynamics in an optomechanical system [25,26] and the relation between the classical and quantum characteristics of a mechanical motion are discussed [27]. Further, quantum-classical transition and the quantum information processing based on the optomechanical system have been investigated in detail [28,29], where the ground-state cooling of mechanical resonator plays a role of the precondition.

So far, many theoretical and experimental efforts are devoted to achieve the ground-state cooling of mechanical resonator using various cooling schemes [30–37]. In general, the ground-state cooling of a mechanical resonator demands that the variances for the quantum fluctuation of the mechanical resonator’s position and momentum should both quickly tend to 1/2. In order to reach this goal, the effective mechanical damping rate under the perturbation of the radiation pressure should increase significantly [36], which can be achieved by enhancing the cooling anti-Stokes process and suppressing the heating Stokes’ in the resolved sideband limit, that is, the frequency of the mechanical resonator is greater than the decay rate of the optical field [30,31]. However, it is noted that the resolved sideband condition is a bit tight except for a few special optomechanical systems [38,39], which is hard to fulfill in experiment. In this regard, one way to circumvent the limitation is to use electromagnetically-induced-transparency (EIT) cooling scheme for a mechanical resonator coupled to a superconducting flux qubit, where the quantum interference can cancel unwanted heating effect [40]. The EIT ground-state cooling scheme in another hybrid optomechanical systems with a three-level atomic medium or a single electronic spin qubit are discussed in detail [41,42]. The ground-state cooling in the regime of bad cavity is achieved by adding a good cavity with a small cavity decay rate, which can vary the effective optical response of the whole cavity system and thus satisfy the effective resolved sideband condition for the ground-state cooling of a mechanical resonator [43–45].

In this work, we investigate the ground-state cooling of a mechanical resonator in a hybrid optomechanical system composed of a standard optomechanical cavity and an ensemble of two-level quantum emitters trapped inside the cavity. We consider that the ensemble is positioned near the cavity mirror and therefore the motion of the cavity mirror changes the transition rate of each quantum emitter via vacuum fluctuation, which induces a direct coupling between the internal states of the quantum emitters and the mechanical mode [46] which changes significantly the dynamics of the whole system and its steady-state characteristics compared with those in the absence of the vacuum force. We investigated the influences of the vacuum coupling strength on the effective frequency and the effective damping rate of the movable mirror by studying the dynamics of the quantum fluctuation of the system. Further, we focus on the effect of the vacuum fluctuation on the ground-state cooling of the mechanical resonator, in the presence of which we found that the ground-state cooling of the mechanical motion can be achieved in the bad cavity limit. Further, we also analyze in detail the effect of the cavity and atomical detuning, the coupling strength between the optical field and the quantum emitter and the decay rate of the quantum emitter on the motional ground-state cooling.

2. Model and Hamiltonian

The hybrid optomechanical system investigated here is shown in Fig. 1, in which a thin ensemble of N quantum emitters is positioned inside a standard optomechanical cavity consisting of one fixed mirror and one movable mirror. Further, the ensemble of the quantum emitters (position q0) is modeled by a two-level system with the excited-state |e〉 and ground-state |g〉, which is driven by an external field with frequency ωf. Thus, the quantum emitters can mediate a linear driving of the single-mode field inside the optical cavity. Correspondingly, the movable mirror interacts with the driven field via the optical radiation pressure. In general, the movable mirror in an optomechanical system is treated as a quantum-mechanical harmonic oscillator with mass m, frequency ωm, and damping rate γm. In order to realize the direct coupling between the two-level ensemble and the mechanical mirror, we assume that the ensemble is positioned near the right movable mirror, i.e. the distance between the emitter and movable mirror is of the order of nanometers. In this case, the vacuum energy fluctuation around the quantum emitters leads to a modification of its electronic state. Consequently, the transition rate of each emitter depends strongly on the position x of the movable mirror in the direction of the cavity axis, i.e. ωa = ωa(d + x), where d (dx) is the distance between the quantum emitter and the mirror in a static emitter-plane geometry [47,48]. The coupling strength between the quantum emitter and the cavity mirror can be calculated as λ0=ωa(d+x)x|x=0. The total Hamiltonian of the hybrid optomechanical system can be written as [49–51]

H=h¯ωcaa+p22m+12mωm2x2+12h¯ωa(d+x)i=1Nσz(i)+h¯g(ai=1Nσ+(i)+ai=1Nσ(i))h¯χcaax+h¯(Ωi=1Nσ+(i)eiωft+Ω*i=1Nσ(i)eiωft)
where a is the photon annihilation operator of the cavity mode with frequency ωc and decay rate κ, which satisfies the commutation relation [a, a] = 1. x and p are the position and momentum operator of the movable mirror with commutation relation [x, p] = ih̄. The two-level system can be described by the pseudospin-1/2 operators σz, σ+ and σ, which satisfy the commutation relations [σ+, σ] = σz and [σz, σ±] = ±2σ±. g (χc = ωc/L) is the coupling strength between the cavity field and the quantum emitter (movable mirror). Ω is the driving strength of the ensemble and L is the cavity length. In particular, we consider that the number of the quantum emitters is very large, i.e. N ≫ 1, but the excitation probability of a single quantum emitter is very small. In this low excitation limit, the collective operators can be defined as i=1Nσ+(i)=NS+ and i=1Nσ(i)=NS, where the operators S+ and S satisfy the fundamental commutation relation [S,S+] ≃ 1 [41,52]. In addition, the condition of low excitation limit should not be appreciably altered by the driving field and the interaction with the cavity, so that i=1Nσz(i)=2S+SNN.

 figure: Fig. 1

Fig. 1 A hybrid optomechanical system model. An ensemble of quantum emitters is positioned inside a standard optomechanical cavity, which is modeled by a two-level system with the excited-state |e〉 and ground-state |g〉 and driven by an external field with frequency ωf. The motion of the mirror nearby the quantum emitter changes its transition rate via vacuum fluctuation and therefore leads to the coupling between the quantum emitter and the mechanical mode.

Download Full Size | PDF

In order to derive the coupling strength between the quantum emitter and the movable mirror, the specific expression of the transition rate of the quantum emitter is needed. We first decompose the transition rate of the emitter ωa(d + x) into two parts, the bare frequency ω0 plus the total frequency shift Δωa(d + x). Thus, in the linear approximation, we have ωa(d + x) = ω0 + Δωa(d) + λ0x. In the following, we derive the coupling strength λ0 in terms of the frequency shift. The frequency shift of an effective isotropic two-level system at position r near a plane can be expressed as [46,50,51]

Δωa(r)=δωae(r)δωag(r),
where δωag(r) and δωae(r) are, respectively, the frequency shifts of the emitter at positon r in its ground state and excited state. The two frequency shifts can be calculated straightforwardly in terms of the classical dyadic Green’s function G(r, r, iu) evaluated at the imaginary frequency ω = iu [46,50,51], i.e.
δωag(r)=3cΓ0ω020duu2ω02+u2Tr{G(r,r,iu)},
and
δωae(r)=δωag(r)3πcΓ0ω0TrRe{G(r,r,ω0)},
where c is the speed of light and Γ0 is the free-space spontaneous emission rate of the two-level system. Further, we assume that the quantum emitters sit near the center of the movable mirror. In this case, the Green’s function of the movable mirror can be approximated by that of an infinite plane. Moreover, a reflection component from a plane located at z = 0 determines the value of the Green’s function, which describes the interaction of a two-level system with its own field reflected by this surface. In the vacuum side with z > 0, the trace of this reflected component is TrG(z,z,ω)=ic24πω20dkkK0e2izK0[(ωc)2rs+(k2K02)rp], where K0=(ωc)2k2 and k is the parallel wavevector components, rs and rp are the Fresnel refection coefficients for s and p-polarized waves. Using Eqs. (2)(4), the coupling strength λ0 can be calculated as
λ0(d)=2c3Γ0πω020duω02+u20dkke2idK0[(ωc)2rs+(k2K02)rp]+Re{c3Γ02ω030dkke2id(ω0c)2k2[(ω0c)2rs+(2k2(ω0c)2)rp]}.
It is noted that the influence of the resonant cavity on the Casimir-Polder interaction of a two-level system nearby a cavity wall has been neglected when the cavity length L is large enough. This is because that the change of the density of states of the electromagnetic field inside the cavity does not significantly alter the Casimir-Polder force [53]. Furthermore, the dynamical Casimir-Polder effect due to the dissipation of the electromagnetic fields inside the cavity can be neglected safely with a large optical cavity [54,55].

In a rotating frame at the laser frequency ωf, the total Hamiltonian of the hybrid optomechanical system becomes

H=h¯Δc0aa+12h¯ωm(p2+q2)+h¯(Δa0+λq)S+S+h¯gN(aS++aS)h¯χaaq+h¯N(ΩS++Ω*S),
where Δc0 = ωcωf is the detuning between the cavity mode and the driving field of the emitter, and Δa0 = ω0ωf + Δωa(d) is the detuning between the quantum emitter and its driving field. The position and momentum operators x and p have been nondimensionalized as mωmh¯xq and 1h¯mωmpp in Eq. (6), and therefore the coupling strengths become λ=λ0h¯/mωm and χ=χch¯/mωm. During the derivation of Eq. (6) we neglected a constant term of N since it does not influences the system dynamics.

3. Dynamical evolution of the system

Here we focus mainly on the quantum dynamics of the nonlinear optomechanical system and investigate further the cooling characteristics of the mechanical oscillator in the presence of the vacuum effect. In particular, the ground-state cooling of the mechanical motion is expected to be possible even in the unresolved sideband regime i.e. k > ωm by adjusting the decay rate of the quantum emitter. Using the Hamiltonian (6) and taking into account the effects of the noises and damping, the full dynamics of the system can be described by the following Heisenberg-Langevin equation of motion:

q˙=ωmp,p˙=ωmqλN/2+χaaγmp+ξ(t),a˙=(κ+iΔc0)a+iχaqiGS+2κain(t),S˙=(Γa+iΔa0)SiGaiλqSiη+2ΓaΓ(t),
where G=gN and η=ΩN are the collective coupling strength of the two-level ensemble. The term −γmp depicts a damping force acting on the movable mirror with damping rate γm, which results mainly from the contact with the thermal bath through the external support. Further, we consider a high-frequency oscillator operating in the MHz range and thus its damping rate (γm = ωm/Q) can be evaluated to be of the order of 102 Hertz with a mechanical quality factor Q ∼ 105 [56]. ξ(t) is the Brownian noise force with zero mean acting on the movable mirror, which satisfies the correlation [57] ξ(t)ξ(t)=γmωmdω2πeiω(tt)ω(1+cothω2kBT), where kB is the Boltzmann constant and T is the thermal bath temperature related to the movable mirror. ain is the optical input-noise operator characterized fully by its correlation, which is given in the Markovian approximation as ain(t)ain(t)=δ(tt). Γ(t) is the environmental noises corresponding to the operator S, which obeys a similar correlation, Γ(t)Γ(t)=δ(tt). Note that a nearby boundary can modify the emission rate of the two-level system, i.e. Γa(z)=Γ0+2πcΓ0ω0TrIm{G(z,z,ω0)} [46,50,51]. In the linear approximation, we have Γa(d + x) ≃ Γa(d) + Λ0x, where Λ0=Γa(d+x)x|x=0. However, the coupling term Λ0x is always much smaller than the zeroth order term Γa(d) and the coupling term λ0x. Therefore the term Γa(d + x) ≃ Γa(d).

We first solve the steady-state expectation of the system, which can be obtained by replacing the operators by their averages in Eq. (7) and then setting all the time derivatives to zero. Further, using the mean field approximation (factorization assumption) 〈aq〉 = 〈a〉 〈q〉 [58], the steady-state values can be derived as follows:

ps=0,qs=χ|αs|2λN/2ωm,αs=iGSsκ+iΔc,Ss=iηΓa+G2κκ2+Δc2+i(ΔaG2Δcκ2+Δc2),
where Δc = Δc0χqs and Δa = Δa0 + λqs are, respectively, the effective detunings of the cavity mode and the quantum emitter. In the following, we focus on the dynamical evolution of the quantum fluctuation around these steady-state expectation, which is described by a set of linearized equations of motion for the fluctuation operators δq, δp, δa and δS. This can be derived by spliting the each operator in Eq. (7) into the steady-state value and corresponding quantum fluctuation, e.g. q = qs + δq, p = ps + δp, a = αs + δa and S = Ss + δS as
δq˙=ωmδp,δp˙=ωmδq+χ(αs*δa+αsδa)γmp+ξ(t),δa˙=(κ+iΔc)δa+iχαsδqiGδS+2κain(t),δS˙=(Γa+iΔa)δSiGδaiλSsδq+2ΓaΓ(t),
where we have considered the case |αs| ≫ 1 and therefore all the terms higher than linear order in the fluctuations δq, δp, δa, δσ and δσz in Eq. (7) are neglected safely.

Defining new quadrature fluctuation operators for the nonlinear optomechanical system δX=(δa+δa)/2, δY=(δaδa)/2i, δU=(δS+δS+)/2, δV=(δSδS+)/2i and the corresponding Hermitian input noise operators δXin=(δain+δain,)/2, δYin=(δainδain,)/2i, δu=(δΓ+δΓ)/2, δv=(δΓδΓ)/2i, the Eq. (9) can be written in a more compact form, (t) = Jf(t) + n(t), with the column vector of fluctuation operator fT(t) = (δp(t), δq(t), δX(t), δY(t), δU(t), δV(t)) and the corresponding column vector of noise nT(t)=(0,ξ,2κδXin,2κδYin,2Γaδu,2Γaδv). J is the drift matrix, which is given by

J=(0ωm0000ωmγmχpχn00χn0κΔc0Gχp0ΔcκG0Gn00GΓaΔaGp0G0ΔaΓa),
with χp=χ(αs*+αs)/2, χn=χ(αsαs*)/2i, Gp=λ(Ss*+Ss)/2, Gn=λ(SsSs*)/2i. For simplify, we choose properly the phase of the quantum emitter so that Ss is real. Further, the stability of the system demands that all the eigenvalues of the drift matrix J have negative real parts. These stability conditions for the hybrid optomechanical system can be obtained by using the Routh-Hurwitz criteria [59]. In the following section, We always resort to numerical calculation to check the stability of the steady-state solutions because the explicit inequalities satisfying stability condition are quite cumbersome. Correspondingly, all the external parameters used by us are also chosen to satisfy the stability condition.

4. Cooling of the mechanical motion

We are interested in the cooling characteristics of the mechanical resonator in the hybrid optomechanical system, where the two-level ensemble is driven by an external laser. In particular, we investigate the important role of the vacuum coupling in the ground-state cooling of the mechanical motion. In order to achieve this goal, we solve Eq. (9) by taking its Fourier transform to obtain the position fluctuation of the movable mirror around the steady state as δq(ω) = χ(ω)F(ω), where F is the Fourier transform of the total force acting on the movable mirror. The effective mechanical susceptibility of the movable mirror introduced here is defined as χ(ω)=ωm/(ωeff2ω2+iωγeff), where the effective resonance frequency and effective damping rate of the movable mirror are given by

ωeff=ωmRe[A(ω)+B(ω)+C(ω)D(ω)]
and
γeff=γm+ωmωIm[A(ω)+B(ω)+C(ω)D(ω)],
respectively, in which we define
D(ω)=[G2Γaκi(Γa+κ)ω+ω2]2+2G2ΔaΔc+(Γa+iω)2Δc2+Δa2β1,A(ω)=G2β3(Γa+iω)+(Γa+iω)2[β1ωmΔc(χn2+χp2)],B(ω)=Δa[(G2Δc+β1Δa)ωm(G2+ΔaΔc)χn2ΔcΔcχp2GGpβ2],C(ω)=G(Γa+iω)(Gβ4+Gpβ5)+G2[G2ωmGGpχp+Δa(Δcωmχp2)],β1(ω)=(κ+iω)2+Δc2,β2(ω)=(κ+iω)χn+Δcχp,β3(ω)=(κ+iω)ωm+χnχp,β4(ω)=χnχp(κ+iω)ωm,β5(ω)=(κ+iω)χpχnΔc.
In order to show the influences of the vacuum effect on the cooling of the movable mirror and how the ground-state cooling of the movable mirror is realized, we can count the effective phonon number in the mirror’s motion,
Nph=(δp2+(δq2)1)/2,
where 〈δp2〉 and 〈δq2〉 are the momentum and displacement variances of the movable mirror in the steady state, and expressed as δp2=12πSp(ω)dω and δq2=12πSq(ω)dω, respectively [36]. In view of definition of the spectrum Sq(ω) of fluctuation in position of the movable mirror 〈δq(ω′)δq(ω)〉s = Sq(ω)δ(ω′ + ω), we have
Sq(ω)=|χ(ω)|2[St(ω)+Sf(ω)+Sa(ω)],
where
St(ω)=γmωωm(1+cothh¯ω2kBT),Sf(ω)=2κ|D(ω)|2[|h1(ω)|2+|h2(ω)|2],Sa(ω)=2Γa|D(ω)|2[|h3(ω)|2+|h4(ω)|2],h1(ω)=(G2Δa+β6Δc)χn+[(Γ1+iω)β7+(κ+iω)Δa2]χp,h2(ω)=(G2Δa+β6Δc)χp+[(Γa+iω)β7(κ+iω)Δa2]χn,h3(ω)=Gβ8χnGχp[(κ+iω)Δa+(Γa+iω)Δc],h4(ω)=G[(Γa+iω)β5G2χpΔaβ2],β6(ω)=(κ+iω)2+Δa2,β7(ω)=G2+(ωiΓa)(ωiκ),β8(ω)=G2Γaκiω(Γa+κ)+ω2+ΔaΔc.
The spectrum Sp(ω) of the momentum fluctuation of the movable mirror can be written as Sp(ω) = (ω/ωm)2 Sx1(ω). We next investigate the cooling of the movable mirror in the absence or the presence of the vacuum coupling. The dependence of the effective frequency and the damping rate of the movable mirror on the vacuum coupling strength is also studied.

To illustrate numerically the effective oscillation frequency of the movable mirror, we need to calculate the frequency shift and emission decay rate of the two-level system near the movable mirror and then derive the vacuum coupling strength λ0. We assume the movable mirror is a thick dielectric (SiN) membrane coated with copper. Then, the reflection coefficient characterizing copper-vacuum interface is given by the standard Fresnel coefficients rs(iu,k)=K3k2+ε(iu)u2/c2K3+k2+ε(iu)u2/c2 and rp(iu,k)=ε(iu)K3k2+ε(iu)u2/c2ε(iu)K3+k2+ε(iu)u2/c2, where K3=k2+u2/c2. Further, the frequency shift Δωa(d) can be calculated with Green’s function. In addition, we need the electromagnetic response of copper, which is described by a Drude model through the dielectric constant ε(iu)=1+ωP2u2+uγ, where ωP is the plasma frequency proportional to the density of conducting electrons in the metal, and γ is a damping parameter satisfying γωP and therefore their contributions to the vacuum coupling strength and the frequency shift are marginal.

Figure 2(a) plots the normalized effective oscillation frequency ωeffm as a function of the normalized frequency ω/ωm with different distances d. For the explicit calculation of the vacuum coupling strength, we use the plasma wavelength (λP = 136nm) of copper corresponding to the plasma frequency ωP = 2πc/λP and γ = 0.0033ωP [60]. In addition, we select the accessible parameters in experiment, i.e. the intrinsic frequency of the movable mirror ωm = 2π × 2 × 106 Hz, the mass of the movable mirror m = 5 ng and the decay rate of the movable mirror γm = 2π × 100 Hz. Other parameters i.e. the effective detunings Δc = −ωm and Δa = −ωm; the cavity decay rate κ = 5ωm and the free-space spontaneous emission rate Γ0 = 0.1ωm; The cavity length is L = 0.01 m and the collective coupling strength G = ωm. We consider that the two-level ensemble is driven by a laser with wavelength λL = 1064 nm and the collective coupling strength |η| = 1012 Hz. We see from Fig. 2(a) that in the high-frequency regime of the mechanical oscillator, the frequency of the movable mirror is not altered significantly and therefore the optical spring effect [36,61] of the system is very small. However, the change of the effective frequency becomes significant with decreasing distance d, which means that the stronger the vacuum coupling, the larger the optical spring effect. In particular, the frequency shift of the movable mirror due to the optical spring effect becomes relevant with a small mechanical frequency [61] and therefore contributes to the measurement of the CP force by the optomechanical device.

 figure: Fig. 2

Fig. 2 The normalized effective oscillation frequency ωeffm and the normalized effective damping rate γeffm are plotted as a function of the normalized frequency ω/ωm with different distances. We use the plasma wavelength λP = 136 nm of copper corresponding to the plasma frequency ωP = 2πc/λP and γ = 0.0033 ωP to calculate the coupling strength between the quantum emitter and the nearby mirror. Other parameters are, respectively, ωm = 2π × 2 × 106 Hz, m = 5 ng, γm = 2π × 100 Hz, Δc = −ωm, Δa = −ωm, κ = 5ωm, Γ0 = 0.1ωm and L = 0.01 m. In addition, the quantum emitter is driven by a laser with wavelength λL = 1064 nm and the collective coupling strength |η| = 1012 Hz. The collective coupling strength G = ωm.

Download Full Size | PDF

Figure 2(b) depicts the normalized effective damping rate γeffm as a function of the normalized frequency ω/ωm with different distances d. We can see from Fig. 2(b) that in the presence of the vacuum coupling, the effective damping rate increases significantly with decreasing d. When the distance is very large, i.e. ω0d/c → ∞, the vacuum coupling disappears and the effective damping rate is minimal. Indeed, in the presence of the vacuum coupling, the decay rate of the two-level ensemble Γa and the coupling strength λ increase with decreasing d. Thus, the effective damping rate [see Eq. (12)] depends strongly on the distance, which can be made very large by decreasing d in the chosen parameter regime. The significant increase in the effective damping rate is essential for cooling mechanical motion which drives the movable mirror close to the quantum ground state. Physically, the increase of the decay rate Γa helps dissipate the phonon in the mechanical motion via the decay channel of the atom ensemble, which enables the ground-state cooling of the movable mirror (see below). The ground state of a mechanical system is approached as the effective phonon number nph < 1, which corresponds generally to the situation of the energy equipartition with both variances 〈δq2〉 ≃ 1/2 and 〈δp2〉 ≃ 1/2. This demands that the initial mean-thermal excitation quantum number n = [exp(h̄ωm/kBT) − 1]−1 is not excessively large, which is possible when the temperature is low enough, e.g., T = 0.1 K. We focus mainly on the ground-state cooling of the mechanical part in the bad cavity limit, i.e. κ = 5ωm, in which the condition of the effective resolved sideband is satisfied by adjusting the bare decay rate of the quantum emitter, i.e. Γ0ωm.

In Fig. 3, we plot the effective phonon number as a function of the dimensionless effective cavity detuning Δcm with different distances d. It is clear that the effective phonon number is always larger than one in the absence of the vacuum coupling i.e. ω0d/c → ∞. Therefore, the ground state of the mechanical resonator can not be achieved. In the presence of the vacuum coupling, the effective phonon number nph < 1 in a finite interval of Δc, which shrinks gradually and vanishes with increasing distance d. Correspondingly, the minimum of the effective phonon number nphmin increases with increasing distance d. The cooling mechanism can be understood from the interference and the level configuration in Fig. 4. In general, in order to cool the mechanical mode, the heating process in Fig. 4 should be suppressed as much as possible [30,62,63]. In the absence of vacuum coupling, the destructive quantum interference exists between the two different excitation pathways, from |n, na, m〉 → |n + 1, na, m + 1〉 directly and from |n, na, m〉 → |n + 1, na, m + 1〉 → |n, na + 1, m + 1〉 → |n + 1, na, m + 1〉 indirectly, which leads to the suppression of the heating process and therefore the cooling of the mechanical mode. However, when the vacuum coupling is included, the additional excitation pathway, i.e. from |n, na, m〉 → |n, na +1, m+1〉, will also take place. Then, the two excitation pathways, from |n, na, m〉 → |n + 1, na, m + 1〉 directly and from |n, na, m〉 → |n, na + 1, m + 1〉 → |n + 1, na, m + 1〉 indirectly, destructively interfere. In addition, the destructive interference also takes place between two indirect pathways, from |n, na, m〉 → |n + 1, na, m + 1〉 → |n, na + 1, m + 1〉 → |n + 1, na, m + 1〉 through the radiation pressure coupling and from |n, na, m〉 → |n, na + 1, m + 1〉 → |n + 1, na, m + 1〉 through the vacuum coupling. Therefore, increase of the interference channels in contrast to the case with only optomechanical coupling helps suppress the heating process in the hybrid system and enhance ground state cooling of the mechanical motion. Consequently, the effective phonon number nph is related to the strength of vacuum coupling and increases with increasing distance d. Then, the vacuum coupling may be used for controlling the cooling of the mechanical motion.

 figure: Fig. 3

Fig. 3 The effective phonon number Nph is plotted as a function of the dimensionless effective cavity detuning Δcm with different distances d. The collective coupling strength |η| = 0.8 × 1012 Hz and the temperature of the movable mirror is selected as T = 0.1 K. Other parameter values are the same as those in Fig. 2.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Plot the energy level diagram in the hybrid optomechanical system, where |n, na, m〉 denotes the state of n photons in optical mode, na atomic excitations in collective excitation mode and m phonons in mechanical mode. The black double arrow denotes the coupling between states |n, na + 1, m + 1〉 and |n + 1, na, m + 1〉 through the atom-field coupling. The black solid (dashed) arrow denotes the cooling (heating) process of the mechanical mode through the radiation pressure coupling and the red solid (dashed) arrow denotes the cooling (heating) process of the mechanical mode through vacuum coupling.

Download Full Size | PDF

Figure 5 shows the effective phonon number as a function of the dimensionless effective detuning of the emitter Δam with different distances d. The effective detuning Δc = −ωm and therefore the cavity is tuned to the Stokes sideband. We can see that the curve for the effective phonon number has two minima, which are, respectively, reached at the Anti-Stokes sideband resonance, i.e. Δaωm and Stokes sideband resonance, i.e. Δa ≃ −ωm. In the absence of the vacuum coupling, the effective phonon number is always larger than one. Things changed when the vacuum coupling between the quantum emitter and the movable mirror is included. For example, when ω0d/c = 0.10, the two minimal values of the effective phonon number are nph = 0.07 and nph = 0.23, respectively. Moreover, the minimum of the effective phonon number also increases with increasing distance d. This is because in the presence of vacuum coupling, the destructive quantum interference appearing between the additional excitation pathways through vacuum coupling further suppresses the heating transition in the system, where one pair of pathways are |n, na, m〉 → |n + 1, na, m + 1〉 and |n, na, m〉 → |n, na + 1, m + 1〉 → |n + 1, na, m + 1〉, and the other pair of pathways are |n, na, m〉 → |n + 1, na, m + 1〉 → |n, na + 1, m + 1〉 → |n + 1, na, m + 1〉 and |n, na, m〉 → |n, na + 1, m + 1〉 → |n + 1, na, m + 1〉. These results suggest that a strong vacuum coupling is preferable for achieving the ground-state cooling of a mechanical oscillator.

 figure: Fig. 5

Fig. 5 The effective phonon number Nph is plotted as a function of the dimensionless effective detuning of the emitter Δam with different distances d. The effective detuning Δc = −ωm. Other parameter values are the same as those in Fig. 3.

Download Full Size | PDF

The collective coupling strength G will influence significantly the effective phonon number and therefore the ground-state cooling of the mechanical motion. In the following, we focus on the two cases in the absence and the presence (ω0d/c = 0.1) of the vacuum coupling. The effective phonon number is shown in Fig. 6 as a function of the coupling strength G. We see from Fig. 6 that the curve for the effective phonon number has only one minimum in the absence of vacuum coupling. The optimal coupling strength G for cooling mechanical motion results from a pair of excitation pathways (|n, na, m〉 → |n + 1, na, m+ 1〉 and |n, na, m〉 → |n + 1, na, m + 1〉 → |n, na + 1, m + 1〉 → |n + 1, na, m + 1〉) through the radiation pressure. In particular, in this case the ground-state cooling of the mechanical motion can not be obtained with the selected parameters. When the vacuum coupling is included, i.e. ω0d/c = 0.1, there exist two minima of the effective phonon number, which are smaller than one and therefore the ground-state cooling of the mechanical motion can be achieved with a proper coupling strength G. Similarly, these optimal coupling strengths G for the cooling of mechanical motion result from the transition |n, na, m〉 → |n + 1, na, m + 1〉 through radiation pressure coupling and the transition |n, na, m〉 → |n, na + 1, m + 1〉 through vacuum coupling, both of which contribute to the suppression of the heating process in the mechanical mode.

 figure: Fig. 6

Fig. 6 The effective phonon number Nph is plotted as a function of the collective coupling strength g in absence and presence of the vacuum coupling (ω0d/c = 0.1). The effective detuning Δc = −ωm. Other parameter values are the same as those in Fig. 3.

Download Full Size | PDF

The effective phonon number is shown in Fig. 7 as a function of the free-space spontaneous emission rate of the quantum emitter Γ0 in the absence and presence of the vacuum coupling. The aim of Fig. 7 is to present clearly the important role of the quantum emitter for the cooling of the movable mirror in the bad cavity limit. We see from Fig. 7 that the effective phonon number is smaller than one only in a finite region near Γ0 = 0.1ωm when the vacuum coupling is included, which means that the decay rate Γ0 of the quantum emitter helps satisfy the effective resolved sideband condition and therefore enhances the cooling process of the mechanical motion in the bad cavity limit, which is similar to the ground-state cooling of a mechanical resonator in a coupled cavity optomechanical system via an electromagnetically-induced- transparency (EIT)-like mechanism [43]. In the absence of the vacuum coupling, the effective phonon number is also always larger than one and therefore the ground state of the movable mirror can not be approached. These results suggest that we should select proper coupling strength G and decay rate Γ0 in order to get optimal cooling for the movable mirror.

 figure: Fig. 7

Fig. 7 The effective phonon number Nph is plotted as a function of the free-space spontaneous emission rate of the quantum emitter Γ0 in the absence and presence of the vacuum coupling (ω0d/c = 0.1). Other parameter values are the same as those in Fig. 6.

Download Full Size | PDF

5. Conclusion

In conclusion,we studied the cooling of a mechanical resonator in a hybrid optomechanical system with an ensemble of two-level quantum emitters, which is coupled to the nearby movable mirror via vacuum force. The hybrid optomechanical cavity is driven along with the atom ensemble. We adopt the quantum Langevin equation to deal with the hybrid system and focus on the linearized dynamics around the steady state of the system. Further, by applying the Fourier transform method, the effective resonance frequency and effective damping rate of the movable mirror are derived. We also investigated the cooling characteristics of the mechanical motion by counting the effective phonon number in the absence and presence of the vacuum effect. We showed that the ground-state cooling of the mechanical motion can be approached in the bad cavity regime, when the vacuum coupling between the emitter ensemble and the movable mechanical resonator is included. The influences of the parameters of the cavity and the quantum emitter on the cooling of the mechanical motion are investigated in detail numerically. The results obtained here are not only useful for realizing the quantum control and manipulation of macroscopic mechanical motion but also able to provide one possible utilization of the vacuum force in an optomechanical device.

Acknowledgments

This research is supported by the National Natural Science Foundation of China (NSFC) under Grant No. 11304010, No. 11565014, No. 11365009 and No. 11375093, the startup Foundation for Doctors of East China Jiaotong University, under Grant No. 26541001.

References and links

1. T. J. Kippenberg and K. J. Vahala, “Cavity optomechanics: back-action at the mesoscale,” Science 321(5893), 1172–1176 (2008). [CrossRef]   [PubMed]  

2. T. J. Kippenberg and K. J. Vahala, “Cavity opto-mechanics,” Opt. Express 15(25), 17172–17205 (2007). [CrossRef]   [PubMed]  

3. M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, “Cavity optomechanics,” Rev. Mod. Phys. 86(4), 1391–1452 (2014). [CrossRef]  

4. D. Vitali, S. Gigan, A. Ferreira, H. R. Böhm, P. Tombesi, A. Guerreiro, V. Vedral, A. Zeilinger, and M. Aspelmeyer, “Optomechanical entanglement between a movable mirror and a cavity field,” Phys. Rev. Lett. 98(3), 030405 (2007). [CrossRef]   [PubMed]  

5. C. Genes, A. Mari, P. Tombesi, and D. Vitali, “Robust entanglement of a micromechanical resonator with output optical fields,” Phys. Rev. A 78(3), 032316 (2008). [CrossRef]  

6. M. J. Hartmann and M. B. Plenio, “Steady state entanglement in the mechanical vibrations of two dielectric membranes,” Phys. Rev. Lett. 101(20), 200503 (2008). [CrossRef]   [PubMed]  

7. M. C. Kuzyk, S. J. van Enk, and H. Wang, “Generating robust optical entanglement in weak-coupling optomechanical systems,” Phys. Rev. A 88(6), 062341 (2013). [CrossRef]  

8. Y.-D. Wang and A. A. Clerk, “Reservoir-engineered entanglement in optomechanical systems,” Phys. Rev. Lett. 110(25), 253601 (2013). [CrossRef]   [PubMed]  

9. W. J. Nie, Y. H. Lan, Y. Li, and S. Y. Zhu, “Effect of the Casimir force on the entanglement between a levitated nanosphere and cavity modes,” Phys. Rev. A 86(6), 063809 (2012). [CrossRef]  

10. W. J. Nie, Y. H. Lan, Y. Li, and S. Y. Zhu, “Generating large steady-state optomechanical entanglement by the action of Casimir force,” Sci. China Phys. Mech. 57(12), 2276–2284 (2014). [CrossRef]  

11. W. Ge, M. Al-Amri, H. Nha, and M. S. Zubairy, “Entanglement of movable mirrors in a correlated-emission laser,” Phys. Rev. A 88(2), 022338 (2013). [CrossRef]  

12. K. Jähne, C. Genes, K. Hammerer, M. Wallquist, E. S. Polzik, and P. Zoller, “Cavity-assisted squeezing of a mechanical oscillator,” Phys. Rev. A 79(6), 063819 (2009). [CrossRef]  

13. W.-J. Gu, G.-X. Li, and Y.-P. Yang, “Generation of squeezed states in a movable mirror via dissipative optomechanical coupling,” Phys. Rev. A 88(1), 013835 (2013). [CrossRef]  

14. X.-Y. Lü, Y. Wu, J. R. Johansson, H. Jing, J. Zhang, and F. Nori, “Squeezed optomechanics with phase-matched amplification and dissipation,” Phys. Rev. Lett. 114(9), 093602 (2015). [CrossRef]   [PubMed]  

15. T. Li, S. Kheifets, and M. G. Raizen, “Millikelvin cooling of an optically trapped microsphere in vacuum,” Nat. Physics 7(7), 527–530 (2011). [CrossRef]  

16. D. E. Chang, C. A. Regal, S. B. Papp, D. J. Wilson, J. Ye, O. Painter, H. J. Kimble, and P. Zoller, “Cavity opto-mechanics using an optically levitated nanosphere,” PNAS 107(3), 1005–1010 (2010). [CrossRef]   [PubMed]  

17. Z.-q. Yin, A. A. Geraci, and T. Li, “Optomechanics of levitated dielectric particles,” Int. J. Mod. Phys. B 27(26), 1330018 (2013). [CrossRef]  

18. J. Restrepo, C. Ciuti, and I. Favero, “Single-polariton optomechanics,” Phys. Rev. Lett. 112(1), 013601 (2014). [CrossRef]   [PubMed]  

19. E. Verhagen, S. Deleglise, S. Weis, A. Schliesser, and T. J. Kippenberg, “Quantum-coherent coupling of a mechanical oscillator to an optical cavity mode,” Nature 482(7383), 63–67 (2012). [CrossRef]   [PubMed]  

20. Y. Li, Y.-D. Wang, F. Xue, and C. Bruder, “Quantum theory of transmission line resonator-assisted cooling of a micromechanical resonator,” Phys. Rev. B 78(13), 134301 (2008). [CrossRef]  

21. M. Abdi and M. J. Hartmann, “Entangling the motion of two optically trapped objects via time-modulated driving fields,” New J. Phys. 17(1), 013056 (2015). [CrossRef]  

22. O. Romero-Isart, A. C. Pflanzer, M. L. Juan, R. Quidant, N. Kiesel, M. Aspelmeyer, and J. I. Cirac, “Optically levitating dielectrics in the quantum regime: theory and protocols,” Phys. Rev. A 83(1), 013803 (2011). [CrossRef]  

23. H. Xiong, L. G. Si, X. Y L’u, X. X. Yang, and Y. Wu, “Review of cavity optomechanics in the weak-coupling regime: from linearization to intrinsic nonlinear interactions,” Sci. China Phys. Mech. 58(5), 1–13 (2015). [CrossRef]  

24. S. Barzanjeh, M. Abdi, G. J. Milburn, P. Tombesi, and D. Vitali, “Reversible optical-to-microwave quantum interface,” Phys. Rev. Lett. 109(13), 130503 (2012). [CrossRef]   [PubMed]  

25. J. Ma, C. You, L.-G. Si, H. Xiong, J. Li, X. Yang, and Y. Wu, “Formation and manipulation of optomechanical chaos via a bichromatic driving,” Phys. Rev. A 90(4), 043839 (2014). [CrossRef]  

26. K. Zhang, W. Chen, M. Bhattacharya, and P. Meystre, “Hamiltonian chaos in a coupled BEC-optomechanical-cavity system,” Phys. Rev. A 81(1), 013802 (2010). [CrossRef]  

27. G. Wang, L. Huang, Y.-C. Lai, and C. Grebogi, “Nonlinear dynamics and quantum entanglement in optomechanical systems,” Phys. Rev. Lett. 112(11), 110406 (2014). [CrossRef]   [PubMed]  

28. K. Stannigel, P. Rabl, A. S. Sørensen, P. Zoller, and M. D. Lukin, “Optomechanical transducers for long-distance quantum communication,” Phys. Rev. Lett. 105(22), 220501 (2010). [CrossRef]  

29. S. Rips and M. J. Hartmann, “Quantum information processing with nanomechanical qubits,” Phys. Rev. Lett. 110(12), 120503 (2013). [CrossRef]   [PubMed]  

30. I. Wilson-Rae, N. Nooshi, W. Zwerger, and T. J. Kippenberg, “Theory of ground state cooling of a mechanical oscillator using dynamical backaction,” Phys. Rev. Lett. 99(9), 093901 (2007). [CrossRef]   [PubMed]  

31. F. Marquardt, J. P. Chen, A. A. Clerk, and S. M. Girvin, “Quantum theory of cavity-assisted sideband cooling of mechanical motion,” Phys. Rev. Lett. 99(9), 093902 (2007). [CrossRef]   [PubMed]  

32. D. Kleckner and D. Bouwmeester, “Sub-kelvin optical cooling of a micromechanical resonator,” Nature 444(7115), 75–78 (2006). [CrossRef]   [PubMed]  

33. M. Bhattacharya and P. Meystre, “Trapping and cooling a mirror to its quantum mechanical ground state,” Phys. Rev. Lett. 99(7), 073601 (2007). [CrossRef]   [PubMed]  

34. S. Gigan, H. R. Bohm, M. Paternostro, F. Blaser, G. Langer, J. B. Hertzberg, K. C. Schwab, D. Bauerle, M. Aspelmeyer, and A. Zeilinger, “Self-cooling of a micromirror by radiation pressure,” Nature 444(7115), 67–70 (2006). [CrossRef]   [PubMed]  

35. Y.-C. Liu, Y.-F. Xiao, X. Luan, and C. W. Wong, “Dynamic dissipative cooling of a mechanical resonator in strong coupling optomechanics,” Phys. Rev. Lett. 110(15), 153606 (2013). [CrossRef]   [PubMed]  

36. C. Genes, D. Vitali, P. Tombesi, S. Gigan, and M. Aspelmeyer, “Ground-state cooling of a micromechanical oscillator: comparing cold damping and cavity-assisted cooling schemes,” Phys. Rev. A 77(3), 033804 (2008). [CrossRef]  

37. W. Nie, Y. Lan, Y. Li, and S. Zhu, “Dynamics of a levitated nanosphere by optomechanical coupling and Casimir interaction,” Phys. Rev. A 88(6), 063849 (2013). [CrossRef]  

38. J. D. Teufel, T. D. Donner, 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(7356), 359–363 (2011). [CrossRef]   [PubMed]  

39. J. Chan, T. P. M. Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Krause, S. Groblacher, M. Aspelmeyer, and O. Painter, “Laser cooling of a nanomechanical oscillator into its quantum ground state,” Nature 478(7367), 89–92 (2011). [CrossRef]   [PubMed]  

40. K. Xia and J. Evers, “Ground state cooling of a nanomechanical resonator in the nonresolved regime via quantum interference,” Phys. Rev. Lett. 103(22), 227203 (2009). [CrossRef]  

41. C. Genes, H. Ritsch, M. Drewsen, and A. Dantan, “Atom-membrane cooling and entanglement using cavity electromagnetically induced transparency,” Phys. Rev. A 84(5), 051801 (2011). [CrossRef]  

42. S. Zhang, J.-Q. Zhang, J. Zhang, C.-W. Wu, W. Wu, and P.-X. Chen, “Ground state cooling of an optomechanical resonator assisted by a Λ-type atom,” Opt. Express 22(23), 28118–28131 (2014). [CrossRef]   [PubMed]  

43. Y. J. Guo, K. Li, W. J. Nie, and Y. Li, “Electromagnetically-induced-transparency-like ground-state cooling in a double-cavity optomechanical system,” Phys. Rev. A 90(5), 053841 (2014). [CrossRef]  

44. F. Bariani, S. Singh, L. F. Buchmann, M. Vengalattore, and P. Meystre, “Hybrid optomechanical cooling by atomic Λ systems,” Phys. Rev. A 90(3), 033838 (2014). [CrossRef]  

45. T. Ojanen and K. Børkje, “Ground-state cooling of mechanical motion in the unresolved sideband regime by use of optomechanically induced transparency,” Phys. Rev. A 90(1), 013824 (2014). [CrossRef]  

46. D. E. Chang, K. Sinha, J. M. Taylor, and H. J. Kimble, “Trapping atoms using nanoscale quantum vacuum forces,” Nat. Commun. 5, 4343 (2014). [CrossRef]   [PubMed]  

47. C. A. Muschik, S. Moulieras, A. Bachtold, F. H. L. Koppens, M. Lewenstein, and D. E. Chang, “Harnessing vacuum forces for quantum sensing of graphene motion,” Phys. Rev. Lett. 112(22), 223601 (2014). [CrossRef]   [PubMed]  

48. M. Antezza, C. Braggio, G. Carugno, A. Noto, R. Passante, L. Rizzuto, G. Ruoso, and S. Spagnolo, “Optomechanical Rydberg-atom excitation via dynamic Casimir-Polder coupling,” Phys. Rev. Lett. 113(2), 023601 (2014). [CrossRef]   [PubMed]  

49. C. K. Law, “Interaction between a moving mirror and radiation pressure: a hamiltonian formulation,” Phys. Rev. A 51(3), 2537–2541 (1995). [CrossRef]   [PubMed]  

50. S. Y. Buhmann, L. Knöll, D.-G. Welsch, and H. T. Dung, “Casimir-Polder forces: a nonperturbative approach,” Phys. Rev. A 70(5), 052117 (2004). [CrossRef]  

51. S. Y. Buhmann and D.-G. Welsch, “Dispersion forces in macroscopic quantum electrodynamics,” Prog. Quant. Electron. 31(2), 51–130 (2007). [CrossRef]  

52. C. P. Sun, Y. Li, and X. F. Liu, “Quasi-spin-wave quantum memories with a dynamical symmetry,” Phys. Rev. Lett. 91(14), 147903 (2003). [CrossRef]   [PubMed]  

53. A. M. Alhambra, A. Kempf, and E. Martín-Martínez, “Casimir forces on atoms in optical cavities,” Phys. Rev. A 89(3), 033835 (2014). [CrossRef]  

54. T. Tian, T. Y. Zheng, Z. H. Wang, and X. Zhang, “Dynamical Casimir-Polder force in a one-dimensional cavity with quasimodes,” Phys. Rev. A 82(1), 013810 (2010). [CrossRef]  

55. H. Yang, T. Zheng, X. Zhang, X. Shao, and S. Pan, “Dynamical Casimir-Polder force on a partially dressed atom in a cavity comprising a dielectric,” Ann. Phys. 344, 69–77 (2014). [CrossRef]  

56. S. Gröblacher, K. Hammerer, M. R. Vanner, and M. Aspelmeyer, “Observation of strong coupling between a micromechanical resonator and an optical cavity field,” Nature 460(7256), 724–727 (2009). [CrossRef]   [PubMed]  

57. V. Giovannetti and D. Vitali, “Phase-noise measurement in a cavity with a movable mirror undergoing quantum Brownian motion,” Phys. Rev. A 63(2), 023812 (2001). [CrossRef]  

58. G. S. Agarwal and S. Huang, “Electromagnetically induced transparency in mechanical effects of light,” Phys. Rev. A 81(4), 041803 (2010). [CrossRef]  

59. E. X. DeJesus and C. Kaufman, “Routh-Hurwitz criterion in the examination of eigenvalues of a system of nonlinear ordinary differential equations,” Phys. Rev. A 35(12), 5288–5290 (1987). [CrossRef]   [PubMed]  

60. A. Lambrecht and S. Reynaud, “Casimir force between metallic mirrors,” Eur. Phys. J. D 8(3), 309–318 (2000). [CrossRef]  

61. T. Corbitt, Y. Chen, E. Innerhofer, H. Müller-Ebhardt, D. Ottaway, H. Rehbein, D. Sigg, S. Whitcomb, C. Wipf, and N. Mavalvala, “An all-optical trap for a gram-scale mirror,” Phys. Rev. Lett. 98(15), 150802 (2007). [CrossRef]   [PubMed]  

62. X. Chen, Y.-C. Liu, P. Peng, Y. Zhi, and Y.-F. Xiao, “Cooling of macroscopic mechanical resonators in hybrid atom-optomechanical systems,” Phys. Rev. A 92(3), 033841 (2015). [CrossRef]  

63. W.-J. Gu and G.-X. Li, “Quantum interference effects on ground-state optomechanical cooling,” Phys. Rev. A 87(2), 025804 (2013). [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 (7)

Fig. 1
Fig. 1 A hybrid optomechanical system model. An ensemble of quantum emitters is positioned inside a standard optomechanical cavity, which is modeled by a two-level system with the excited-state |e〉 and ground-state |g〉 and driven by an external field with frequency ωf. The motion of the mirror nearby the quantum emitter changes its transition rate via vacuum fluctuation and therefore leads to the coupling between the quantum emitter and the mechanical mode.
Fig. 2
Fig. 2 The normalized effective oscillation frequency ωeffm and the normalized effective damping rate γeffm are plotted as a function of the normalized frequency ω/ωm with different distances. We use the plasma wavelength λP = 136 nm of copper corresponding to the plasma frequency ωP = 2πc/λP and γ = 0.0033 ωP to calculate the coupling strength between the quantum emitter and the nearby mirror. Other parameters are, respectively, ωm = 2π × 2 × 106 Hz, m = 5 ng, γm = 2π × 100 Hz, Δc = −ωm, Δa = −ωm, κ = 5ωm, Γ0 = 0.1ωm and L = 0.01 m. In addition, the quantum emitter is driven by a laser with wavelength λL = 1064 nm and the collective coupling strength |η| = 1012 Hz. The collective coupling strength G = ωm.
Fig. 3
Fig. 3 The effective phonon number Nph is plotted as a function of the dimensionless effective cavity detuning Δcm with different distances d. The collective coupling strength |η| = 0.8 × 1012 Hz and the temperature of the movable mirror is selected as T = 0.1 K. Other parameter values are the same as those in Fig. 2.
Fig. 4
Fig. 4 Plot the energy level diagram in the hybrid optomechanical system, where |n, na, m〉 denotes the state of n photons in optical mode, na atomic excitations in collective excitation mode and m phonons in mechanical mode. The black double arrow denotes the coupling between states |n, na + 1, m + 1〉 and |n + 1, na, m + 1〉 through the atom-field coupling. The black solid (dashed) arrow denotes the cooling (heating) process of the mechanical mode through the radiation pressure coupling and the red solid (dashed) arrow denotes the cooling (heating) process of the mechanical mode through vacuum coupling.
Fig. 5
Fig. 5 The effective phonon number Nph is plotted as a function of the dimensionless effective detuning of the emitter Δam with different distances d. The effective detuning Δc = −ωm. Other parameter values are the same as those in Fig. 3.
Fig. 6
Fig. 6 The effective phonon number Nph is plotted as a function of the collective coupling strength g in absence and presence of the vacuum coupling (ω0d/c = 0.1). The effective detuning Δc = −ωm. Other parameter values are the same as those in Fig. 3.
Fig. 7
Fig. 7 The effective phonon number Nph is plotted as a function of the free-space spontaneous emission rate of the quantum emitter Γ0 in the absence and presence of the vacuum coupling (ω0d/c = 0.1). Other parameter values are the same as those in Fig. 6.

Equations (16)

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

H = h ¯ ω c a a + p 2 2 m + 1 2 m ω m 2 x 2 + 1 2 h ¯ ω a ( d + x ) i = 1 N σ z ( i ) + h ¯ g ( a i = 1 N σ + ( i ) + a i = 1 N σ ( i ) ) h ¯ χ c a a x + h ¯ ( Ω i = 1 N σ + ( i ) e i ω f t + Ω * i = 1 N σ ( i ) e i ω f t )
Δ ω a ( r ) = δ ω ae ( r ) δ ω ag ( r ) ,
δ ω ag ( r ) = 3 c Γ 0 ω 0 2 0 d u u 2 ω 0 2 + u 2 Tr { G ( r , r , i u ) } ,
δ ω ae ( r ) = δ ω ag ( r ) 3 π c Γ 0 ω 0 Tr Re { G ( r , r , ω 0 ) } ,
λ 0 ( d ) = 2 c 3 Γ 0 π ω 0 2 0 d u ω 0 2 + u 2 0 d k k e 2 i d K 0 [ ( ω c ) 2 r s + ( k 2 K 0 2 ) r p ] + Re { c 3 Γ 0 2 ω 0 3 0 d k k e 2 i d ( ω 0 c ) 2 k 2 [ ( ω 0 c ) 2 r s + ( 2 k 2 ( ω 0 c ) 2 ) r p ] } .
H = h ¯ Δ c 0 a a + 1 2 h ¯ ω m ( p 2 + q 2 ) + h ¯ ( Δ a 0 + λ q ) S + S + h ¯ g N ( a S + + a S ) h ¯ χ a a q + h ¯ N ( Ω S + + Ω * S ) ,
q ˙ = ω m p , p ˙ = ω m q λ N / 2 + χ a a γ m p + ξ ( t ) , a ˙ = ( κ + i Δ c 0 ) a + i χ aq i GS + 2 κ a in ( t ) , S ˙ = ( Γ a + i Δ a 0 ) S i Ga i λ q S i η + 2 Γ a Γ ( t ) ,
p s = 0 , q s = χ | α s | 2 λ N / 2 ω m , α s = i GS s κ + i Δ c , S s = i η Γ a + G 2 κ κ 2 + Δ c 2 + i ( Δ a G 2 Δ c κ 2 + Δ c 2 ) ,
δ q ˙ = ω m δ p , δ p ˙ = ω m δ q + χ ( α s * δ a + α s δ a ) γ m p + ξ ( t ) , δ a ˙ = ( κ + i Δ c ) δ a + i χ α s δ q i G δ S + 2 κ a in ( t ) , δ S ˙ = ( Γ a + i Δ a ) δ S i G δ a i λ S s δ q + 2 Γ a Γ ( t ) ,
J = ( 0 ω m 0 0 0 0 ω m γ m χ p χ n 0 0 χ n 0 κ Δ c 0 G χ p 0 Δ c κ G 0 G n 0 0 G Γ a Δ a G p 0 G 0 Δ a Γ a ) ,
ω eff = ω m Re [ A ( ω ) + B ( ω ) + C ( ω ) D ( ω ) ]
γ eff = γ m + ω m ω Im [ A ( ω ) + B ( ω ) + C ( ω ) D ( ω ) ] ,
D ( ω ) = [ G 2 Γ a κ i ( Γ a + κ ) ω + ω 2 ] 2 + 2 G 2 Δ a Δ c + ( Γ a + i ω ) 2 Δ c 2 + Δ a 2 β 1 , A ( ω ) = G 2 β 3 ( Γ a + i ω ) + ( Γ a + i ω ) 2 [ β 1 ω m Δ c ( χ n 2 + χ p 2 ) ] , B ( ω ) = Δ a [ ( G 2 Δ c + β 1 Δ a ) ω m ( G 2 + Δ a Δ c ) χ n 2 Δ c Δ c χ p 2 G G p β 2 ] , C ( ω ) = G ( Γ a + i ω ) ( G β 4 + G p β 5 ) + G 2 [ G 2 ω m G G p χ p + Δ a ( Δ c ω m χ p 2 ) ] , β 1 ( ω ) = ( κ + i ω ) 2 + Δ c 2 , β 2 ( ω ) = ( κ + i ω ) χ n + Δ c χ p , β 3 ( ω ) = ( κ + i ω ) ω m + χ n χ p , β 4 ( ω ) = χ n χ p ( κ + i ω ) ω m , β 5 ( ω ) = ( κ + i ω ) χ p χ n Δ c .
N ph = ( δ p 2 + ( δ q 2 ) 1 ) / 2 ,
S q ( ω ) = | χ ( ω ) | 2 [ S t ( ω ) + S f ( ω ) + S a ( ω ) ] ,
S t ( ω ) = γ m ω ω m ( 1 + coth h ¯ ω 2 k B T ) , S f ( ω ) = 2 κ | D ( ω ) | 2 [ | h 1 ( ω ) | 2 + | h 2 ( ω ) | 2 ] , S a ( ω ) = 2 Γ a | D ( ω ) | 2 [ | h 3 ( ω ) | 2 + | h 4 ( ω ) | 2 ] , h 1 ( ω ) = ( G 2 Δ a + β 6 Δ c ) χ n + [ ( Γ 1 + i ω ) β 7 + ( κ + i ω ) Δ a 2 ] χ p , h 2 ( ω ) = ( G 2 Δ a + β 6 Δ c ) χ p + [ ( Γ a + i ω ) β 7 ( κ + i ω ) Δ a 2 ] χ n , h 3 ( ω ) = G β 8 χ n G χ p [ ( κ + i ω ) Δ a + ( Γ a + i ω ) Δ c ] , h 4 ( ω ) = G [ ( Γ a + i ω ) β 5 G 2 χ p Δ a β 2 ] , β 6 ( ω ) = ( κ + i ω ) 2 + Δ a 2 , β 7 ( ω ) = G 2 + ( ω i Γ a ) ( ω i κ ) , β 8 ( ω ) = G 2 Γ a κ i ω ( Γ a + κ ) + ω 2 + Δ a Δ c .
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.