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

Coupling mechanical motion of a single atom to a micromechanical cantilever

Open Access Open Access

Abstract

We design and analyze a hybrid optomechanical setup to achieve an effective coupling between the mechanical motions of a micromechanical cantilever and a single atom trapped inside a cavity, which is mediated by a direct interaction between the micromechanical cantilever and the atomic internal state via the quantum vacuum effect. Moreover, the optomechanical coupling between the mechanical motion of the cantilever and the cavity field can be mediated by the interaction with the atom. Their couplings are demonstrated in detail by analyzing the normal-mode splitting of the mechanical motion and the optical response of the hybrid optomechanical system. It is found that double optomechanically-induced transparency can be observed in the output probe field in the presence of the mediated coupling. In particular, both the width of the splitting peaks and the separation between the two absorption dips increase with the increasing strength of the vacuum coupling and with the decreasing trapped position of the atom. These characteristics can be used to study the strong coupling between a single atom and a massive micromechanical cantilever.

© 2017 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Recently a lot of attention has been drawn to the coherent manipulation of the dynamics of micromechanical oscillators with cavity optomechanics [1–3]. The general setup consists of movable parts with small effective mass and high quality factor, such as oscillating cavity wall [4–7], levitated nanoparticle or nanodiamond [8–13], membrane [14–17], microscale cantilever [18–20] and a cavity field so that the interaction between the electromagnetic fields and a macroscale mechanical oscillator can be enabled by shining directly a light on the mechanical system. Various quantum characteristics of mechanical motion in the optomechanical systems, i.e., ground state cooling of mechanical motion [4, 15, 21–24], non-classical state preparation in mechanical resonators [25–27], squeezing of mechanical mode [28, 29] and optomechanical normal-mode splitting (NMS) [30, 31], have been explored extensively, where the optomechanical coupling via radiation pressure of light plays a determined role.

On the other hand, the light propagation in an optomechanical system can be manipulated by tailoring the interaction between the mechanical oscillator and the light field. A special example of controlling light is optomechanically induced transparency (OMIT) [6, 32–41], where the width of transparency window is related directly to the strength of the optomechanical coupling. Further, upon application of the cavity quantum electrodynamics (QED) [42] to the optomechanical systems, the effective optomechanical coupling between the mechanical motion and the cavity field can be promoted further by the nonlinearity of atomic ensemble [43–49], which changes the quantum characteristics of mechanical system and the light propagation properties. The influence of a single atom or qubit on the response properties of light has been investigated in detail [50, 51]. In addition, the optomechanical coupling between a light field and a collective mode of cold atomic ensemble inside a cavity is explored theoretically [52–55].

Apart from the direct optomechanical coupling via radiation pressure, the internal state of a two-level system, such as a single atom or an ultracold atomic ensemble [56–58], a Josephson-junction qubit [59, 60] or an electron spin accumulated in a carbon nanotube (CNT) [61], is used for mediating the optomechancial coupling between a light field and a mechanical motion. In particular, a strong coupling between a mechanical membrane and the mechanical motion of a single atom trapped inside an optical cavity can be achieved by two quantized fields [56], where the quantum cavity fields are coupled simultaneously to the internal state of the single atom and its mechanical motion. In this setup, the mass ratio of the atom M and the mechanical oscillator m can be maneuvered not to enter the effective coupling and therefore the strong coupling can be realized by adjusting the detunings of cavity fields. It is noted that the internal state of a two-level atom or emitter can be coupled to a nearby plane via the vacuum fluctuation force when they are very close [62–64]. This also enables a direct coupling between the internal state of a single atom and a mechanical membrane, i.e., suspended graphene sheet [63, 64].

In this work we consider a hybrid optomechanical setup consisting of a micromechanical cantilever and a single trapped atom, which are both placed in a Fabry-Pérot cavity. In particular, the micromechanical cantilever plane approaches the atom and therefore the cantilever plane shifts the energy level of the atom via vacuum effect between them, which leads to couple the mechanical motions of the single atom and the oscillating cantilever despite the smallness of the scale factor M/m~106 [20]. Moreover, an effective optomechanical coupling between the micromechanical cantilever and the cavity field enabled via the internal state of the single atom. We analyze the occurrence of NMS in the displacement spectrum of the micromechanical cantilever and discuss in detail influences of the atom-cantilever coupling strength and the steady-state position of atom on the NMS. Further, we investigate the response of the system to a weak probe beam and show that a double OMIT behavior is generated in the presence of the atom-cantilever coupling. These results can be used for demonstrating the effective coupling between the mechanical motions of a single atom and a massive mechanical cantilever despite the fact that the mass of atom is much less than the effective mass of the micromechanical cantilever.

2. Optomechanical model

The atom-cantilever optomechanical system studied here is depicted in Fig. 1, where we consider an atom of mass M, which is confined inside a standing-wave cavity by a external harmonic potential of frequency ωM0 [65]. The atom’s center-of-mass (c.m.) motion is dominated along the cavity axis x, while the other transverse motions, i.e., y-direction or z-direction in Fig. 1, are confined tightly so that the corresponding degrees of freedom of motion have been neglected. The relevant internal degrees of freedom of the atom are the ground state |g〉 and the excited state |e〉 with the bare transition frequency ωa0 and the free-space spontaneous emission rate Γ0. Further, the two-level atom interacts strongly with a cavity mode of frequency ωcωa0. The system is coherently driven by a classical control field with frequency ωL and driving amplitude εL.

 figure: Fig. 1

Fig. 1 Schematic of the setup studied in the paper. The optomechanical system consists of a micromechanical cantilever and a single two-level atom both placed inside a Fabry-Pérot cavity. The system is driven by an external field with frequency ωL and the cavity mode with frequency ωc is probed by a weak probe field with frequency ωp. The distance between the atom and the cantilever plane is so small that the internal state of the atom is influenced by the nearby plane due to the vacuum energy fluctuation.

Download Full Size | PDF

In addition, we assume that the atom approaches a micromechanical cantilever with an effective mass m and oscillation frequency ωm0 placed in the cavity (xz plane). That is, the distance between the atom and the cantilever is very small (i.e., the order of the nanometes) so that the level shift of the atom induced by the zero-point energy fluctuation should be considered [62, 63]. The vacuum fluctuation leads to a direct coupling between the internal states of the two-level atom and the oscillation cantilever. It is noted that in the presence of the cantilever, the vacuum force between the atom and the cantilever is along the y-direction and therefore does not affect the mechanical motion and the oscillating frequency of the atom along the x-direction. That is, the transverse vacuum potential or force created by the nearby dielectric material is unnecessary for studying the dynamics of the atom in the x-direction. Furthermore, the micromechanical cantilever is placed in the xz plane and geometrically different from a movable mirror [4] or dielectric membrane [14]. Therefore, it is not necessary to take into account the light radiation pressure on the cantilever because light in the cavity is reflected back inefficiently [66]. Despite of the fact, an effective optomechanical coupling between the cavity field and the oscillation cantilever can be mediated by the internal states of the single atom. In particular, the effective coupling between the mechanical motions of a single atom and a micromechanical cantilever can be generated despite the scale factor M/m1.

The Hamiltonian of the system can be written as [63, 65, 67, 68]

H=ωcaa+py22m+12mωm02y2+px22M+12MωM02x2+[ωa0+Δωa(d+y)]σ+σig(x)(σ+aσa)+i(εLaeiωLtH.c.),
where the first term in the first line is the free Hamiltonian of the driven cavity field and a the photon annihilation operator in the cavity mode of frequency ωc = kc, satisfying the commutation relation [a, a] = 1. c is the speed of light in vacuum. The other terms in the first line describe the kinetic and potential energies of the cantilever and the atom, x(y) and px (py) are the position and momentum operators for the atom (cantilever) with commutation relation [x, px] = iℏ ([y, py] = iℏ). The first term in the second line is the free Hamiltonian of the atom, where Δωa(d + y) is the frequency shift due to vacuum fluctuation [24, 62, 63] and related to the distance between the atom and the cantilever. Here d is the distance between the two-level atom and the cantilever in the static atom-plane geometry, and y the position of the cantilever. σ = |g〉〈e| is the lowering operator of the atom, σ its adjoint. The second term in the second line describes the interaction of the atom with the driven cavity field, where g (x) represents the cavity atom coupling strength at x, with g(x) = g0cos(kx) [68, 69]. Finally, the last terms describe the interaction of the cavity field with the control field, with amplitude |εL|=2Pcκ/ωL. κ is the decay rate of the cavity field and Pc is the laser power.

In a rotating frame with the laser frequency ωL, the Hamiltonian (1) of the atom-cantilever optomechanical system is

H=Δc0aa+py22m+12mωm02y2+px22M+12MωM02x2+Δa0(d+y)σ+σig0cos(kx)(σ+aσa)+i(εLaH.c.),
where Δc0 = ωLωc is the detuning of the pump laser from the cavity and Δa0 (d + y) = ωLωa0 −∆ωa(d + y) is the detuning from the atomic resonance. The corresponding Heisenberg equation for the atomic polarization and the cavity field operator can be written as [69–71]
σ˙=(γaiΔa0)σ+g(x)aσz+σz2γaΓin,
a˙=(κiΔc0)a+g(x)aσ+εL+2κain.
It is noted that a nearby boundary can modify the spontaneous emission rate of the two-level system, i.e. γa = γa(d + y). In the present model we consider a large atom-pump detuning and low saturation, i.e., ωLωa and Δa0(d + y) ≫ γa, so that the excited atomic state can be eliminated adiabatically [68, 69]. This leads to σs=g(x)aγaiΔa0. Then, the effective Hamiltonian of the atom-membrane optomechanical system become
H=py22m+12mωm02y2+px22M+12MωM02x2Δc0aa+g02aacos2(kx)Δa(d)λ0y+i(εLaH.c.),
where Δa(d) = ωLωa0 − Δωa(d) is the effective detuning of the pump laser from the atomic resonance in the static atom-plane geometry. In the derivation of Eq. (5) we used a first-order approximation of the frequency shift Δωa(d + y) because y is very small compared to the distance d, i.e., Δωa(d + y) = Δωa(d) + λ0y, with λ0(d)=Δωa(d+y)y|y=0 being the vacuum induced coupling between the single atom and the nearby plane, which is related to the level shifts of the atom in its ground and excited states [62, 63, 72, 73].

In order to calculate the vacuum induced coupling λ0, we consider an effective isotropic two-level system at position r near a plane. The shift of the transition rate due to the vacuum effect is written as Δωa(r) = δωae(r) − δωag(r), where δωag (r) and δωae (r) are, respectively, the frequency shifts of the two-level system in its ground state and excited state. Using the classical dyadic Green’s function G(r,r,iu) evaluated at the imaginary frequency ω = iu [62, 63], the two frequency shifts read

δωag(r)=3cΓ0ωa020duu2ωa02+u2Tr{G(r,r,iu)},
and
δωae(r)=δωag(r)3πcΓ0ωa0Tr Re{G(r,r,ωa0)},
respectively. Further, we assume that the single atom sits near the center of the cantilever. In this case, the Green’s function of the cantilever can be approximated by that of an infinite plane. We mainly care a reflection component from a plane located at z = 0, which determines the value of the Green’s function describing the interaction of a two-level system with its own field reflected by this plane. In the vacuum regime of z > 0, the trace of this reflected component is TrG(z,z,ω)=ic24πω20dk||k||K0e2izK0[(ωc)2rs+(k||2K02)rp], where K0=(ωc)2k||2 and k is the parallel wavevector components; rs and rp are the Fresnel refection coefficients for the s and p-polarized waves. Using the specific expression of the shift of the transition rate, the coupling strength λ0 between a two-level atom and a plane can be calculated as
λ0(d)=2c3Γ0πωa020duωa02+u20dk||k||e2idK0×[(ωc)2rs+(k||2K02)rp]+Rec3Γ02ωa030dk||k||e2id(ωa0c)2k||2×[(ωa0c)2rs+(2k||2(ωa0c)2)rp].

We can illustrate numerically the vacuum coupling strength λ0 by assuming that the mechanical oscillator is a dielectric (SiN) cantilever coated with a thin copper [24]. The Fresnel reflection coefficients in the case of vacuum-metal interface are given by rs(iu,k||)=K3k||2+ε(iu)u2/c2K3+k||2+ε(iu)u2/c2 and rp(iu,k||)=ε(iu)K3k||2+ε(iu)u2/c2ε(iu)K3+k||2+ε(iu)u2/c2 respectively, where K3=k||2+u2/c2 and ε is the dielectric constant describing the electromagnetic response of the metallic material. In general, the electromagnetic response of copper 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. In the calculation, we use the plasma wavelength (λP = 136nm) of copper corresponding to the plasma frequency ωP = 2πc/λP and γ = 0.0033ωP [74]. We choose the parameters of the atom, i.e., the transition wavelength λa0 = 780 nm and the free-space spontaneous emission rate Γ0 = 2π × 6.1 MHz, which corresponds to the decay of 87Rb atom from its excited state |5P3/2〉 to the ground state |5S1/2〉 [75]. Figure 2 plots the vacuum coupling strength λ0 as a function of the dimensionless distance ωa0d/c. From Fig. 2 we see that the vacuum induced coupling can be evaluated to be the order of 1012 Hz/nm with a small distance.

 figure: Fig. 2

Fig. 2 The vacuum coupling strength λ0 as a function of the dimensionless distance ωa0d/c. The free-space spontaneous emission rate and the transition wavelength of the atom are Γ0 = 2π × 6.1 MHz and λa0 = 780 nm, respectively. The plasma frequency of copper λP = 136nm and γ = 0.0033ωP.

Download Full Size | PDF

3. Dynamical equation and quantum fluctuations

Using the Hamiltonian (5) and taking into account the decay of the cavity mode and thermal noises affecting the atom and the micromechanical cantilever [76], the quantum dynamics of the system can be described by the following Heisenberg-Langevin equation:

x˙=px/M,p˙x=MωM02x+kg02aasin(2kx)Δaλ0yγMpx+ξM,y˙=py/m,p˙y=mωm02yλ0g02aacos2(kx)(Δaλ0y)2γmpy+ξm,a˙=(iΔc0+κ)aig02cos2(kx)aΔaλ0y+εL+2κain,
where γM and γm are the damping rate of the atom and the cantilever, respectively. ξj(t)(j = m, M) is the Brownian noise term with zero mean value, which is characterized by the correlation function [77] ξj(t)ξj(t)=γjj2πdωeiω(tt)ω[1+coth(ω2kBTj)], where kB is the Boltzmann constant and Tj is the temperature surrounding the atom or the cantilever. ain is the environmental input-noise corresponding to the operator a, which is fully characterized by the nonzero correlation function ain(t)ain(t)=δ(tt).

The steady-state expectation values of the atom, the cantilever and the cavity field can be obtained by setting the time derivatives to 0 in Eq. (9) as

px,ys=0,as=εLκiΔc,xs=kg02|as|2sin(2kxs)MωM02Δa,ys=λ0g02|as|2cos2(kxs)mωm02Δa 2,
where Δa=Δaλ0ys and Δc=Δc0g02cos2(kxs)/Δa. In general, for a large coherent driving of the system, the photon number in the cavity as ≫ 1. Thus, we linearize the problem by considering the case with large photon number. For example, one can split the operators in Eq. (9) into their steady-state values and quantum fluctuations O = Os + δO(O = x, px, y, py, a). By inserting the ansatz O = Os + δO into Eq. (9) and neglecting all the terms higher than linear order in the fluctuations, i.e. δOδO, the quantum Langevin equations for the fluctuations are given by
δx˙=ωMδpx,δp˙x=ωMδx+G12δy+G2a(δa+δa)γMδpx+ξM,δy˙=ωmδpy,δp˙y=ωmδy+G12δx+G1a(δa+δa)γmδpy+ξm,δa˙=(iΔcκ)δaiG1aδy+iG2aδx+2κain,
where ωm=ωm02+2λ02g02|as|2cos2(kxs)/(mΔa 3) and ωM=ωM022k2g02|as|2cos(2kxs)/(MΔa) are the effective frequencies associated with c.m. oscillations of the cantilever and the atom, respectively. G12=λKg02|as|2sin(2kxs)/Δa 2 is the effective coupling between the mechanical motion of the atom and the cantilever; G1a=λg02ascos2(kxs)/Δa 2 is the effective coupling between the cantilever and the cavity field; G2a=Kg02sin(2kxs)as/Δa the effective coupling between the atomical motion and the cavity field. It is found that the coupling G12 is independent of the mass ratio M/m and therefore the atom will affect significantly the dynamics of the cantilever. λ=λ0mωm and K=kMωM. In deriving Eq. (11), we redefined the dimensionless position and momentum fluctuations δx (δy) and δpx (δpy) as MωMδxδx(mωmδyδy) and 1MωMδpxδpx(1mωmδpyδpy) 1. In addition, the phase of the cavity field is chosen so that as is real. Correspondingly, the Brownian noises with zero mean ξj are also renormalized, and have the correlation function ξj(t)ξj(t)=γj2πωjdωeiω(tt)ω[1+coth(ω2kBTj)].

The coupling strengths G12 and G1a can be demonstrated through the normal mode splitting (NMS) of the mechanical motion. To this end, we transform to the quadratures: δX=(δa+δa)/2, δY=i(δaδa)/2, δXin=(δain+δain)/2 and δYin=i(δainδain)/2. Further, introducing the column vector of fluctuation operators f(t) with fT(t) = (δy, δpy, δx, δpx, δX, δY) and the corresponding column vector of noise sources n(t) with nT(t)=(0,ξm,0,ξM,2κδXin,2κδYin), Eq. (11) can be written in the following compact form:

f˙(t)=Jf(t)+n(t),
where the matrix J is given by
J=(0ωm0000ωmγmG1202G1a0000ωM00G120ωMγM2G2a00000κΔc2G1a02G2a0Δcκ).
The stability conditions for the system demands that the real parts of all the eigenvalues of the matrix J are negative. We can use the Routh-Hurwitz criteria [78] to derive these stability conditions. Here, the explicit inequalities are quite cumbersome and therefore we always resort to numerics. Hereafter, we restrict the selected external parameters to the stable regime.

The expression for the position fluctuation of the oscillating cantilever in Fourier space is given by

δy(ω)=ωmd(ω)[A1(ω)ξm(ω)+A2(ω)ξM(ω)]+2κωmd(ω)[A3(ω)δXin(ω)+A4(ω)δYin(ω)],
where
A1(ω)=(κ+iω)2(ω2+iγMω+ωM2)Δc(ω2Δc2ωMG2a2ΔcωM2iωγMΔc),A2(ω)=ωM[G12(κ+iω)2+2G1aG2aΔc+G12Δc2],A3(ω)=2(κ+iω)[G1a(ω2ωM2iωγM)+G12G2aωM],A4(ω)=2Δc[G1a(ωM2ω2+iωγM)G12G2aωM],d(ω)=(κ+iω)2B0ΔcB1,
with B0=ω4ω2ωm2iγM(ω3ωωm2)G122ωmω2ω2ωM2ωm2ωM2iωγm(ω2iωγMωM2) and B1=2G2a2ω(ωiγm)ωM2G2a2ωm2ωM+2G1aωm(2G1aω22iG1aωγM+22G2aG12ωM2G1aωM2)ΔcB0. In Eq. (13) for δy(ω), the terms proportional to ξm and ξM arise from thermal noises, and the terms proportional to δXin and δYin originate from radiation pressure. The spectrum of fluctuation in the position of the oscillating cantilever is obtained from
Sy(ω)=dΩ2πei(ω+Ω)tδy(ω)δ(Ω)+δy(Ω)δ(ω)2,
together with the nonzero correlation functions of the noise sources in the frequency domain:
ξj(ω)ξj(Ω)=2πγjωjω[1+coth(ω2kBTj)]δ(ω+Ω),δXin(ω)δXin(Ω)=δYin(ω)δYin(Ω)=πδ(ω+Ω).
The spectrum of fluctuation in the cantilever position in Fourier space is finally obtained as
Sy(ω)=|χ(ω)|2[γmωωmβm+γMωωM|A2(ω)A1(ω)|2βM]+|χ(ω)|2κ[|A3(ω)A1(ω)|2+|A4(ω)A1(ω)|2],
where βj=1+cothω2kBTj and the effective mechanical coefficient for cantilever is written as χ(ω)=ωm/[(ωmeff)2ω2+iωγmeff];
ωmeff=Re[d(ω)/A1(ω)+ω2]
and
γmeff=Im[d(ω)/A1(ω)]/ω
are the effective resonance frequency and effective damping rate of the cantilever, respectively. This spectrum Sy(ω) is characterized by the effective mechanical coefficient χ(ω) and any information about the modified motion of the cantilever can be obtained from the study of Sy(ω). It is noted that the modification of mechanical frequency due to radiation pressure in Eq. (17) is the so-called optical spring effect [21], which is negligibly small for a mechanical oscillator with high frequency (~ MHz), as shown in Fig. 3(a), where the normalized effective oscillation frequency ωmeff/ωm is plotted as a function of the normalized frequency ω/ωm. Here we select the accessible parameters in optomechanical systems. For example, the wavelength of the driving field λL ≃ 780 nm, the total cavity length L = 1 cm and the cavity decay rate κ = 2π × 105 Hz. The mass of a single 87Rb atom M ≈ 1.42 × 10−25 Kg, the effective oscillation frequency of atom ωM = 2π ×1.9 MHz and the decay rate γM = 2π ×1 Hz due to the collisions with the background gas [76]. The single-photon Rabi frequency g0 = 2π × 10.9 MHz and the effective pump-atom detuning Δa =2π×30GHz [52]. The effective mass of the cantilever m = 0.05 × 10−12Kg, the effective frequency of the cantilever ωm = ωM and the damping rate γM = 2π × 300 Hz [20, 79]. The temperatures surrounding the atom and the cantilever are same, i.e., Tm = TM = 0.1K. The steady-stat position of atom satisfies sin(2kxs) = 0.3 and the vacuum coupling between the atom and the cantilever λ = 2π × 9.5 MHz. We also consider that the cavity is driven on its red sideband, i.e., Δc = −ωm. The driving strength εL and the steady-state photon number can be calculated in terms of the steady-state position xs. In contrast to the optical spring effect, the increase of the effective mechanical damping rate plays an important role in the cooling of the mechanical motion [21]. Figure 3(b) depicts the normalized effective damping rate γmeff/ωm as a function of the normalized frequency ω/ωm. It is clearly seen that the effective mechanical damping rate increases significantly at ωωm and therefore helps drive the cantilever close to the quantum ground state.

 figure: Fig. 3

Fig. 3 The normalized effective oscillation frequency ωmeff/ωm (a) and the normalized effective damping rate γmeff/ωm (b) as functions of the normalized frequency ω/ωm. We select atomic parameters, i.e. the free-space spontaneous emission rate Γ0 = 2π × 6.1 MHz, the transition wavelength λa0 = 780 nm, the atomic mass M ≈ 1.42 × 10−25 Kg, the effective mass of the cantilever m = 0.05 × 10−12 Kg, the effective oscillation frequency of atom ωM = 2π × 1.9 MHz, the decay rate γM = 2π × 1 Hz, the Rabi frequency g0 = 2π × 10.9 MHz, the effective pump-atom detuning Δa=2π×30GHz, the steady-state position sin(2kxs) = 0.3 and λ = 2π × 9.5 MHz. The other parameters are λL ≃ 780nm, L = 1 cm, κ = 2π × 105 Hz, ωm = ωM, γm = 2π × 300 Hz, Tm=TM = 0.1K and Δc = −ωm.

Download Full Size | PDF

In Fig. 4, we show the plot of Sy(ω) as a function of normalized frequency ω/ωm with different effective coupling λ’s at fixed position sin(2kxs) = 0.3 [Fig. 4(a)] and different xs’s at fixed effective coupling λ = 2π × 3.8 MHz [Fig. 4(b)]. We used the detuning Δc = 0. Other parameter values are the same as those in Fig. 3. In the absence of the vacuum induced coupling λ = 0 and therefore G12 = 0 and G1a = 0, the effective coupling G2a between the atomic motion and the cavity field does not lead to NMS. This situation is changed when the vacuum interaction between the atom and the nearby cantilever is included. From Fig. 4(a) and 4(b), we observe that NMS appears with sufficiently large vacuum coupling λ. This is because that the vacuum effect in the hybrid system not only induces a directed effective coupling G12 between the cantilever and the atomic motion, but also a non-directed effective coupling G1a between the cantilever and the cavity field, which both contribute to the splitting of spectrum lines. Therefore, the NMS in the system is associated to the mixing between the mechanical mode and the atomic motion and the fluctuation of the cavity field around the steady state. Furthermore, with increasing vacuum coupling strength, the width of the splitting peaks increases. In addition, the position of the atom xs also has an important impact on NMS. From Fig. 4(b), we observe that the width between two peaks increases with decreasing xs. The gradual increase in the width of the splitting peaks results from the increase in the amplitudes of the effective couplings G12 and G1a, which leads to an intriguing mixing of the mechanical mode and the atomical motion and the cavity field. As we know, NMS indicates strong coupling. In particular, the displacement spectrum of the cantilever and the NMS can be obtained by homodyne detection scheme [5, 80]. Thus, the present characteristics may be used for demonstrating the strong coupling between the mechanical motions of a single atom and a massive cantilever despite the fact that the mass of atom is much less than the effective mass of the cantilever, i.e., M/m1.

 figure: Fig. 4

Fig. 4 The displacement spectrum of the cantilever as a function of normalized frequency ω/ωm with different λ’s for sin(2kxs) = 0.3 (a) and different xs’s for λ = 2π × 3.8 MHz (b). Δc = 0 and other parameter values are the same as those in Fig. 3.

Download Full Size | PDF

4. OMIT in the output field

In order to study the optical response of the hybrid system, we consider that a weak probe field with frequency ωp and amplitude εp is sent into the cavity (see Fig. 1). Thus, the Hamiltonian of the whole system should include the terms related to the interaction between the driven field and the probe field, i.e., i(εpaeiωptH.c.). Correspondingly, the dynamics of the cavity field in the quantum Langevin equations [Eqs. (9) and (11)] should include the role of the probe field, i.e., δa˙(iΔcκ)δaiG1aδy+iG2aδx+εpeiδt+2κain, where δ = ωpωL is the detuning of the probe laser from the cavity. Then, using the dynamics of the quantum fluctuations around the steady-state, we can study the optical properties of the output field in the atom-cantilever optomechanical system, i.e., optomechanically induced transparency, which also demonstrates the coupling between the mechanical oscillator and the atom. Here we use the ansatz δO = O+eiδt + Oe to solve Eq. (11). After substituting this ansatz into Eq. (11), the following solution of interest for the response of the optomechanical system is attained,

a+=εp[β3iG1aiG2a(G1aα2+G2aα1)2β4+iG1aα2+iG2aα1]1,
where α1=G2aβ1G12G1aβ1β2G122, α2 = (G1aG12α1)/β1, β1 = ωm2δ2iδγm, β2 = ωM2δ2iδγM, β3 = κi(∆c + δ), β4 = κ + i(∆cδ).

Here we investigate mainly the component of the output field oscillating at the probe frequency ωp, and therefore the expression of a is not necessary to describe the four-wave mixing with frequency ωp−2ωf for the driving field and the weak probe field. Using the standard input-output theory [50] aout(t)+εL+εpeiδt2κ=2κa and the ansatz δa = a+eiδt + a_ei, we can express the mean value of the output field as

2κaout(t)=2καsεL+(2κa+εp)eiδt+2κ(a)eiδt.
It is stressed that the term related to a+ in the above expression corresponds directly to the output field at the probe frequency ωp via the detung δ. In order to examine the total output field at the frequency ωp, we define an amplitude of the rescaled output field:
εout=2κa+/εp,
where the constant term has been omitted. The real and imaginary parts of the amplitude of this term, Re(εout) and Im(εout), respectively, account for in-phase and out-phase quadratures of the output field spectrum and describe the absorption and dispersion by the whole system of the weak probe field, which can be measured by homodyne detections [80].

In the present hybrid optomechanical system, the mass of the atom is much less than the effective mass of the cantilever. However, the mechanical motions of a single atom and a massive cantilever oscillator can be coupled effectively by the effective coupling coefficient G12, which is independent of their mass ratio M/m. Thus, the motion of a single atom may affect significantly the optical response of the system. In this section, we investigate in detail the generation of dips in the absorption spectrum induced by the vacuum-assisted atom-cantilever coupling G12 and optomechanical coupling G1a.

We now numerically evaluate the phase quadratures Re(εout) and Im(εout)through the corresponding output field a+. Further, we analyze in detail the OMIT phenomena in the system, induced by the mechanical motion of the atom and the cantilever. In Fig. 5, we show the absorption Re(εout) and dispersion Im(εout) in the output field, as a function of δ/ωm with different effective coupling strength λ. In the absence of the vacuum action, i.e., λ = 0 in Fig. 5(a), we found that the system is transparent and the single-mode OMIT behavior appears when δ = ωm. In this case the effective coupling coefficients G12 = 0 and G1a = 0 due to the disappeared λ and the generation of the single absorption dip is induced directly by the effective coupling G2a between the atomic motion and the cavity field. Physically, This can be understood since the probe beam interferes destructively with the anti-Stokes field generated by the atom [6, 32], which plays a role of mechanical oscillator. In particular, we may find from Fig. 5(b)5(d) that the double absorption dips for the output light of the probe beam appear when the vacuum induced couplings, i.e., G12 and G1a, are included. Indeed, in the presence of the vacuum action between the atom and the nearby plane, the oscillating cantilever is coupled to the atom directly via the effective coupling G12 and to the cavity field indirectly via the effective coupling G1a. These couplings leads to more interference channels and therefore break down the symmetry in the single OMIT interference [41]. Furthermore, the two absorption dips become more and more separated with increasing the effective coupling λ, which determines the amplitude of the vacuum induced couplings, G12 and G1a. We see from Fig. 5 that Im(εout) exhibits abnormal dispersion, induced by the atom-field and the optomechanical couplings. In Fig. 6, we show the interval δ12/ωm between the two absorption dips as a function of normalized coupling λ/ωm, where δ12 = δ1δ2 and δ1 (δ2) is the detuning corresponding to left (right) absorption dips in Fig. 5 and can be evaluated by dRe(εout)/dδ|δ=δ1,2=0. It is clearly seen from Fig. 6 that δ12 increases linearly with increasing the coupling λ, which can be used to detect the atom-plane vacuum force with the hybrid optomechanical system.

 figure: Fig. 5

Fig. 5 The absorption Re(εout) (solid line) and dispersion Im εout (dash line) in the output field are plotted as a function of δ/ωm with different λ’s. Other parameter values are the same as those in Fig. 3.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 The interval δ12/ωm between the two absorption dips is plotted as a function of the effective coupling strength λ. Other parameter values are the same as those in Fig. 3.

Download Full Size | PDF

In the presence of the vacuum force, the coupling coefficients G12 and G1a depend strongly on the steady-state position of atom xs. Fig. 7 presents the variation of the absorption Re(εout) and dispersion Im(εout) of the output field with respect to δ/ωm with different xs’s at given λ = 2π × 3.8 MHz. From Fig. 7, we can observe that the positions of the two absorption dips separate gradually with decreasing xs. Correspondingly, the left and right transparency windows become more and more separated. This leads to that the interval δ12 between the two absorption dips increases monotonously with decreasing xs due to the increased effective coupling coefficients G12 and G1. In Fig. 8, we show the number of the excited atom η=g02cos2(kxs)|as|2/(Δa )2 as a function of sin(2kxs). From Fig. 8, we see clearly that the smaller the position xs, the larger the excited number of atom. This means that in the regime of a small value xs, the low-excitation limit of the atom will be broken. Thus, in order to keep away from the regime and distinguish the double absorption dips as much as possible, we should select optimally the position of the trapped atom.

 figure: Fig. 7

Fig. 7 The absorption Re(εout) (solid line) and dispersion Im(εout) (dash line) of the output field are plotted as a function of δ/ωm with different xs’s at given λ = 2π × 3.8 MHz. Other parameter values are the same as those in Fig. 3.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 The number of the excited atom η are plotted as a function of sin (2kxs). Other parameter values are the same as those in Fig. 7.

Download Full Size | PDF

The general case is considered with different frequencies for the trapped atom and the oscillating cantilever, i.e., ωmωM. In Fig. 9, we show the absorption Re(εout) in the output field, as a function of δ/ωm with identical [Fig. 9(b)] and different [Fig. 9(a)) and 9(c)] oscillation frequencies. The effective coupling strength λ = 2π × 3.8 MHz and the steady-state position of atom sin(2kxs) = 0.3. In Fig. 9, we observe that with respect to the case of identical frequency ωm = ωM, the additional absorption peak induced by the vacuum effect moves rightward (leftward) in the case of ωM < ωm (ωM > ωm). In particular, in the case of different frequencies, the absorption spectrum profile becomes more asymmetric and the dips in absorption become shallower when the two frequencies are further apart. The increase of the interval δ12 in the different frequencies enhances its sensitivity to the vacuum interaction.

 figure: Fig. 9

Fig. 9 The absorption Re(εout) of the output field is plotted as a function of δ/ωm with identical [Fig. 9(b)] and different [Fig. 9(a) and 9(c)] oscillation frequencies. The effective coupling strength λ = 2π×3.8 MHz and the steady-state position of the atom sin(2kxs)= 0.3. Other parameter values are the same as those in Fig. 7.

Download Full Size | PDF

Discussion

In this work, we have designed a hybrid optomechanical system for the realization of vacuum-assisted coupling among a macromechanical cantilever, a single atom trapped inside a cavity and a single mode of a Fabry-Pérot cavity. In the presence of atom-vacuum interaction, the effective mechanical coupling between a massive cantilever and a single atom as well as an effective optomechanical coupling between the massive cantilever and the cavity field are studied with the quantum Langevin equation and its linearized dynamics around semiclassical steady states. Furthermore, we have analyzed the occurrence of NMS in the displacement spectrum of the oscillating cantilever, which is associated to the mixing between the mechanical mode and the atomic motion and the fluctuation of the cavity field. We discuss in detail numerically influences of the vacuum induced coupling and the position of the atom on the NMS, which indicates the effective strength of coupling between the mechanical motion of a single atom and a massive cantilever. In addition, we also focused our attention to the optical-response properties of the hybrid system and showed that the double-OMIT behavior in the absorption spectrum of the probe field can be observed when the effective optomechanical coupling and mechanical coupling are large enough. In particular, the separation between the two absorption dips is related directly to the vacuum coupling between the atom and the cantilever as well as the position of the trapped atom. These results have a potential application to the realization of the effective coupling between the mechanical motions of a mechanical oscillator and a single atom.

Funding

National Natural Science Foundation of China (NSFC) (11565014, 11775190, 11375093); Natural Science Foundation of Jiangxi Province (20161BAB211023 and 20171BAB201015).

References and links

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

2. P. Meystre, “A short walk through quantum optomechanics,” Ann. Phys. 525, 215–233 (2013). [CrossRef]  

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

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

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

6. S. Weis, R. Rivière, S. Deléglise, E. Gavartin, O. Arcizet, A. Schliesser, and T. J. Kippenberg, “Optomechanically induced transparency,” Science 330, 1520 (2010). [CrossRef]   [PubMed]  

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

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

9. 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, 013803 (2011). [CrossRef]  

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

11. M. L. Juan, G. Molina-Terriza, T. Volz, and O. Romero-Isart, “Near-field levitated quantum optomechanics with nanodiamonds,” Phys. Rev. A 94, 023841 (2016). [CrossRef]  

12. T. M. Hoang, Y. Ma, J. Ahn, J. Bang, F. Robicheaux, Z.-Q. Yin, and T. Li, “Torsional optomechanics of a levitated nonspherical nanoparticle,” Phys. Rev. Lett. 117, 123604 (2016). [CrossRef]   [PubMed]  

13. T. M. Hoang, J. Ahn, J. Bang, and T. Li, “Electron spin control of optically levitated nanodiamonds in vacuum,” Nat. Commun. 7, 12250 (2016). [CrossRef]  

14. J. D. Thompson, B. M. Zwickl, A. M. Jayich, Florian Marquardt, S. M. Girvin, and J. G. E. Harris, “Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane,” Nature (London) 452, 72–75 (2008). [CrossRef]  

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

16. L. F. Buchmann, L. Zhang, A. Chiruvelli, and P. Meystre, “Macroscopic tunneling of a membrane in an optomechanical double-well potential,” Phys. Rev. Lett. 108, 210403 (2012). [CrossRef]   [PubMed]  

17. R. A. Norte, J. P. Moura, and S. Gröblacher, “Mechanical resonators for quantum optomechanics experiments at room temperature,” Phys. Rev. Lett. 116, 147202 (2016). [CrossRef]   [PubMed]  

18. C. Metzger, M. Ludwig, C. Neuenhahn, A. Ortlieb, I. Favero, K. Karrai, and F. Marquardt, “Self-induced oscillations in an optomechanical system driven by bolometric backaction,” Phys. Rev. Lett. 101, 133903 (2008). [CrossRef]   [PubMed]  

19. K. Srinivasan, H. Miao, M. T. Rakher, M. Davanço, and V. Aksyuk, “Optomechanical transduction of an integrated silicon cantilever probe using a microdisk resonator,” Nano Letters 11, 791–797 (2011). [CrossRef]   [PubMed]  

20. C. Doolin, P. H. Kim, B. D. Hauer, A. J. R. MacDonald, and J. P. Davis, “Multidimensional optomechanical cantilevers for high-frequency force sensing,” New J. Phys. 16, 035001 (2014). [CrossRef]  

21. 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, 033804 (2008). [CrossRef]  

22. 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, 153606 (2013). [CrossRef]   [PubMed]  

23. P. Stadler, W. Belzig, and G. Rastelli, “Ground-state cooling of a carbon nanomechanical resonator by spin-polarized current,” Phys. Rev. Lett. 113, 047201 (2014). [CrossRef]   [PubMed]  

24. W. J. Nie, A. X. Chen, and Y. H. Lan, “Cooling mechanical motion via vacuum effect of an ensemble of quantum emitters,” Opt. Express 23, 30970–30984 (2015). [CrossRef]   [PubMed]  

25. X. W. Xu, Y. J. Zhao, and Y. X. Liu, “Entangled-state engineering of vibrational modes in a multimembrane optomechanical system,” Phys. Rev. A 88, 022325 (2013). [CrossRef]  

26. M. Abdi and M. J. Hartmann, “Entangling the motion of two optically trapped objects via time-modulated driving fields,” New J. Phys. 17013056 (2015). [CrossRef]  

27. R. Riedinger, S. Hong, R. A. Norte, J. A. Slater, J. Shang, A. G. Krause, V. Anant, M. Aspelmeyer, and S. Gröblacher, “Non-classical correlations between single photons and phonons from a mechanical oscillator,” Nature (London) 530, 313–316 (2016). [CrossRef]  

28. 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, 063819 (2009). [CrossRef]  

29. E. E. Wollman, C. U. Lei, A. J. Weinstein, J. Suh, A. Kronwald, F. Marquardt, A. A. Clerk, and K. C. Schwab, “Quantum squeezing of motion in a mechanical resonator,” Science 349, 952–955 (2015). [CrossRef]   [PubMed]  

30. J. M. Dobrindt, I. Wilson-Rae, and T. J. Kippenberg, “Parametric normal-mode splitting in cavity optomechanics,” Phys. Rev. Lett. 101, 263602 (2008). [CrossRef]   [PubMed]  

31. S. Huang and G. S. Agarwal, “Normal-mode splitting in a coupled system of a nanomechanical oscillator and a parametric amplifier cavity,” Phys. Rev. A 80, 033807 (2009). [CrossRef]  

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

33. S. Huang and G. S. Agarwal, “Electromagnetically induced transparency from two-phonon processes in quadratically coupled membranes,” Phys. Rev. A 83, 023823 (2011). [CrossRef]  

34. S. Huang and G. S. Agarwal, “Electromagnetically induced transparency with quantized fields in optocavity mechanics,” Phys. Rev. A 83, 043826 (2011). [CrossRef]  

35. A. H. Safavi-Naeini, T. P. M. Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and O. Painter, “Electromagnetically induced transparency and slow light with optomechanics,” Nature (London) 472, 69–73 (2011). [CrossRef]  

36. K. Qu and G. S. Agarwal, “Fano resonances and their control in optomechanics,” Phys. Rev. A 87, 063813 (2013). [CrossRef]  

37. M. Karuza, C. Biancofiore, M. Bawaj, C. Molinelli, M. Galassi, R. Natali, P. Tombesi, G. Di Giuseppe, and D. Vitali, “Optomechanically induced transparency in a membrane-in-the-middle setup at room temperature,” Phys. Rev. A 88, 013804 (2013). [CrossRef]  

38. A. Kronwald and F. Marquardt, “Optomechanically induced transparency in the nonlinear quantum regime,” Phys. Rev. Lett. 111, 133601 (2013). [CrossRef]   [PubMed]  

39. H. Jing, Ş. K. Özdemir, Z. Geng, J. Zhang, X.-Y. Lü, B. Peng, L. Yang, and F. Nori, “Optomechanically-induced transparency in parity-time-symmetric microresonators,” Sci. Rep. 5, 9663 (2015). [CrossRef]   [PubMed]  

40. H. Xiong, Y.-M. Huang, L.-L. Wan, and Y. Wu, “Vector cavity optomechanics in the parameter configuration of optomechanically induced transparency,” Phys. Rev. A 94, 013816 (2016). [CrossRef]  

41. P.-C. Ma, J.-Q. Zhang, Y. Xiao, M. Feng, and Z.-M. Zhang, “Tunable double optomechanically induced transparency in an optomechanical system,” Phys. Rev. A 90, 043825 (2014). [CrossRef]  

42. S. Haroche and J. M. Raimond, Exploring the Quantum: Atoms cavities and Photons (Oxford University, 2006). [CrossRef]  

43. H. Ian, Z. R. Gong, Y. X. Liu, C. P. Sun, and F. Nori, “Cavity optomechanical coupling assisted by an atomic gas,” Phys. Rev. A 78, 013824 (2008). [CrossRef]  

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

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

46. Y. Xiao, Y.-F. Yu, and Z.-M. Zhang, “Controllable optomechanically induced transparency and ponderomotive squeezing in an optomechanical system assisted by an atomic ensemble,” Opt. Express 22, 17979–17989 (2014). [CrossRef]   [PubMed]  

47. A. Jöckel, A. Faber, T. Kampschulte, M. Korppi, M. T. Rakher, and P. Treutlein, “Sympathetic cooling of a membrane oscillator in a hybrid mechanical-atomic system,” Nat. Nanotech. 10, 55 (2015). [CrossRef]  

48. W. J. Nie, A. Chen, and Y. H. Lan, “ Optical-response properties in levitated optomechanical systems beyond the low-excitation limit,” Phys. Rev. A 93, 023841 (2016). [CrossRef]  

49. L. Li, W. J. Nie, and A. Chen, “Transparency and tunable slow and fast light in a nonlinear optomechanical cavity,” Sci. Rep. 6, 35090 (2016). [CrossRef]   [PubMed]  

50. M. J. Akram, F. Ghafoor, and F. Saif, “Electromagnetically induced transparency and tunable fano resonances in hybrid optomechanics,” J. Phys. B: At. Mol. Opt. Phys. 48, 065502 (2015). [CrossRef]  

51. H. Wang, X. Gu, . Y.-x. Liu, A. Miranowicz, and F. Nori, “Optomechanical analog of two-color electromagnetically induced transparency: Photon transmission through an optomechanical device with a two-level system,” Phys. Rev. A 90, 023817 (2014). [CrossRef]  

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

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

54. L. Zhou, H. Pu, K. Zhang, X.-D. Zhao, and W. P. Zhang, “Cavity-induced switching between localized and extended states in a noninteracting Bose-Einstein condensate,” Phys. Rev. A 84, 043606 (2011). [CrossRef]  

55. K. A. Yasir, L. Zhuang, and W.-M. Liu, “Spin-orbit-coupling-induced backaction cooling in cavity optomechanics with a Bose-Einstein condensate,” Phys. Rev. A 95, 013810 (2017). [CrossRef]  

56. K. Hammerer, M. Wallquist, C. Genes, M. Ludwig, F. Marquardt, P. Treutlein, P. Zoller, J. Ye, and H. J. Kimble, “Strong coupling of a mechanical oscillator and a single atom,” Phys. Rev. Lett. 103, 063005 (2009). [CrossRef]   [PubMed]  

57. J.-M. Pirkkalainen, S. U. Cho, F. Massel, J. Tuorila, T. T. Heikkilä, P. J. Hakonen, and M. A. Sillanpää, “Cavity optomechanics mediated by a quantum two-level system,” Nat. Commun. 6, 6981 (2015). [CrossRef]   [PubMed]  

58. S. Camerer, M. Korppi, A. Jöckel, D. Hunger, T. W. Hänsch, and P. Treutlein, “Realization of an optomechanical interface between ultracold atoms and a membrane,” Phys. Rev. Lett. 107, 223001 (2011). [CrossRef]   [PubMed]  

59. L. F. Wei, Y.-X. Liu, C. P. Sun, and F. Nori, “Probing tiny motions of nanomechanical resonators: classical or quantum mechanical?” Phys. Rev. Lett. 97, 237201 (2006). [CrossRef]  

60. W. Y. Huo and G. L. Long, “Generation of squeezed states of nanomechanical resonator using three-wave mixing,” Appl. Phys. Lett. 92, 133102 (2008). [CrossRef]  

61. A. V. Parafilo, S. I. Kulinich, L. Y. Gorelik, M. N. Kiselev, R. I. Shekhter, and M. Jonson, “Spin-mediated photomechanical coupling of a nanoelectromechanical shuttle,” Appl. Phys. Lett. 117, 057202 (2016). [CrossRef]  

62. 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]  

63. 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, 223601 (2014). [CrossRef]   [PubMed]  

64. S. Ribeiro and H. Terças, “Sympathetic laser cooling of graphene with Casimir-Polder forces,” Phys. Rev. A 94, 043420 (2016). [CrossRef]  

65. G. Morigi, J. Eschner, S. Mancini, and D. Vitali, “Coherent generation of EPR-entangled light pulses mediated by a single trapped atom,” Phys. Rev. A 73, 033822 (2006). [CrossRef]  

66. X. Z. Yuan, “Entangling an optical cavity and a nanomechanical resonator beam by means of a quantum dot,” Phys. Rev. A 88, 052317 (2013). [CrossRef]  

67. D. J. Alton, N. P. Stern, T. Aoki, H. Lee, E. Ostby, K. J. Vahala, and H. J. Kimble, “Strong interactions of single atoms and photons near a dielectric boundary,” Nat. Phys. 7, 159–165 (2011). [CrossRef]  

68. C. Maschler and H. Ritsch, “Cold atom dynamics in a quantum optical lattice potential,” Phys. Rev. Lett. 95, 260401 (2005). [CrossRef]  

69. C. Maschler and H. Ritsch, “Quantum motion of laser-driven atoms in a cavity field,” Opt. Commun. 243, 145 (2004). [CrossRef]  

70. C.W. Gardiner and P. Zoller, Quantum Noise (Springer, 2000). [CrossRef]  

71. Z.-Q. Yin, W. L. Yang, L. Sun, and L. M. Duan, “Quantum network of superconducting qubits through an optomechanical interface,” Phys. Rev. A 91, 012333 (2015). [CrossRef]  

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

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

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

75. C. S. Hofmann, G. Günter, H. Schempp, M. Robert-de-Saint-Vincent, M. Gärttner, J. Evers, S. Whitlock, and M. Weidemüller, “Sub-Poissonian statistics of Rydberg-interacting dark-state polaritons,” Phys. Rev. Lett. 110, 203601 (2013). [CrossRef]   [PubMed]  

76. D. Hungera, S. Camerera, M. Korppia, A. Jöckela, T. W. Hänscha, and P. Treutleina, “Coupling ultracold atoms to mechanical oscillators,” Comptes Rendus Physique 12, 871 (2011). [CrossRef]  

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

78. 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, 5288 (1987). [CrossRef]  

79. G. Anetsberger, O. Arcizet, Q. P. Unterreithmeier, R. Rivière, A. Schliesser, E. M. Weig, J. P. Kotthaus, and T. J. Kippenberg, “Near-field cavity optomechanics with nanomechanical oscillators,” Nature Phys. 5, 909–914 (2009). [CrossRef]  

80. D. F. Walls and G. J. Milburn, Quantum Optics (Springer, 1994).

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 (9)

Fig. 1
Fig. 1 Schematic of the setup studied in the paper. The optomechanical system consists of a micromechanical cantilever and a single two-level atom both placed inside a Fabry-Pérot cavity. The system is driven by an external field with frequency ωL and the cavity mode with frequency ωc is probed by a weak probe field with frequency ωp. The distance between the atom and the cantilever plane is so small that the internal state of the atom is influenced by the nearby plane due to the vacuum energy fluctuation.
Fig. 2
Fig. 2 The vacuum coupling strength λ0 as a function of the dimensionless distance ωa0d/c. The free-space spontaneous emission rate and the transition wavelength of the atom are Γ0 = 2π × 6.1 MHz and λa0 = 780 nm, respectively. The plasma frequency of copper λP = 136nm and γ = 0.0033ωP.
Fig. 3
Fig. 3 The normalized effective oscillation frequency ω m e f f / ω m (a) and the normalized effective damping rate γ m e f f / ω m (b) as functions of the normalized frequency ω/ωm. We select atomic parameters, i.e. the free-space spontaneous emission rate Γ0 = 2π × 6.1 MHz, the transition wavelength λa0 = 780 nm, the atomic mass M ≈ 1.42 × 10−25 Kg, the effective mass of the cantilever m = 0.05 × 10−12 Kg, the effective oscillation frequency of atom ωM = 2π × 1.9 MHz, the decay rate γM = 2π × 1 Hz, the Rabi frequency g0 = 2π × 10.9 MHz, the effective pump-atom detuning Δ a = 2 π × 30 GHz, the steady-state position sin(2kxs) = 0.3 and λ = 2π × 9.5 MHz. The other parameters are λL ≃ 780nm, L = 1 cm, κ = 2π × 105 Hz, ωm = ωM, γm = 2π × 300 Hz, Tm=TM = 0.1K and Δc = −ωm.
Fig. 4
Fig. 4 The displacement spectrum of the cantilever as a function of normalized frequency ω/ωm with different λ’s for sin(2kxs) = 0.3 (a) and different xs’s for λ = 2π × 3.8 MHz (b). Δc = 0 and other parameter values are the same as those in Fig. 3.
Fig. 5
Fig. 5 The absorption Re(εout) (solid line) and dispersion Im εout (dash line) in the output field are plotted as a function of δ/ωm with different λ’s. Other parameter values are the same as those in Fig. 3.
Fig. 6
Fig. 6 The interval δ12/ωm between the two absorption dips is plotted as a function of the effective coupling strength λ. Other parameter values are the same as those in Fig. 3.
Fig. 7
Fig. 7 The absorption Re(εout) (solid line) and dispersion Im(εout) (dash line) of the output field are plotted as a function of δ/ωm with different xs’s at given λ = 2π × 3.8 MHz. Other parameter values are the same as those in Fig. 3.
Fig. 8
Fig. 8 The number of the excited atom η are plotted as a function of sin (2kxs). Other parameter values are the same as those in Fig. 7.
Fig. 9
Fig. 9 The absorption Re(εout) of the output field is plotted as a function of δ/ωm with identical [Fig. 9(b)] and different [Fig. 9(a) and 9(c)] oscillation frequencies. The effective coupling strength λ = 2π×3.8 MHz and the steady-state position of the atom sin(2kxs)= 0.3. Other parameter values are the same as those in Fig. 7.

Equations (23)

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

H = ω c a a + p y 2 2 m + 1 2 m ω m 0 2 y 2 + p x 2 2 M + 1 2 M ω M 0 2 x 2 + [ ω a 0 + Δ ω a ( d + y ) ] σ + σ i g ( x ) ( σ + a σ a ) + i ( ε L a e i ω L t H . c . ) ,
H = Δ c 0 a a + p y 2 2 m + 1 2 m ω m 0 2 y 2 + p x 2 2 M + 1 2 M ω M 0 2 x 2 + Δ a 0 ( d + y ) σ + σ i g 0 cos ( k x ) ( σ + a σ a ) + i ( ε L a H . c . ) ,
σ ˙ = ( γ a i Δ a 0 ) σ + g ( x ) a σ z + σ z 2 γ a Γ i n ,
a ˙ = ( κ i Δ c 0 ) a + g ( x ) a σ + ε L + 2 κ a i n .
H = p y 2 2 m + 1 2 m ω m 0 2 y 2 + p x 2 2 M + 1 2 M ω M 0 2 x 2 Δ c 0 a a + g 0 2 a a cos 2 ( k x ) Δ a ( d ) λ 0 y + i ( ε L a H . c . ) ,
δ ω a g ( r ) = 3 c Γ 0 ω a 0 2 0 d u u 2 ω a 0 2 + u 2 Tr { G ( r , r , i u ) } ,
δ ω a e ( r ) = δ ω a g ( r ) 3 π c Γ 0 ω a 0 Tr Re { G ( r , r , ω a 0 ) } ,
λ 0 ( d ) = 2 c 3 Γ 0 π ω a 0 2 0 d u ω a 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 a 0 3 0 d k | | k | | e 2 i d ( ω a 0 c ) 2 k | | 2 × [ ( ω a 0 c ) 2 r s + ( 2 k | | 2 ( ω a 0 c ) 2 ) r p ] .
x ˙ = p x / M , p ˙ x = M ω M 0 2 x + k g 0 2 a a sin ( 2 k x ) Δ a λ 0 y γ M p x + ξ M , y ˙ = p y / m , p ˙ y = m ω m 0 2 y λ 0 g 0 2 a a cos 2 ( k x ) ( Δ a λ 0 y ) 2 γ m p y + ξ m , a ˙ = ( i Δ c 0 + κ ) a i g 0 2 cos 2 ( k x ) a Δ a λ 0 y + ε L + 2 κ a i n ,
p x , y s = 0 , a s = ε L κ i Δ c , x s = k g 0 2 | a s | 2 sin ( 2 k x s ) M ω M 0 2 Δ a , y s = λ 0 g 0 2 | a s | 2 cos 2 ( k x s ) m ω m 0 2 Δ a   2 ,
δ x ˙ = ω M δ p x , δ p ˙ x = ω M δ x + G 12 δ y + G 2 a ( δ a + δ a ) γ M δ p x + ξ M , δ y ˙ = ω m δ p y , δ p ˙ y = ω m δ y + G 12 δ x + G 1 a ( δ a + δ a ) γ m δ p y + ξ m , δ a ˙ = ( i Δ c κ ) δ a i G 1 a δ y + i G 2 a δ x + 2 κ a i n ,
f ˙ ( t ) = J f ( t ) + n ( t ) ,
J = ( 0 ω m 0 0 0 0 ω m γ m G 12 0 2 G 1 a 0 0 0 0 ω M 0 0 G 12 0 ω M γ M 2 G 2 a 0 0 0 0 0 κ Δ c 2 G 1 a 0 2 G 2 a 0 Δ c κ ) .
δ y ( ω ) = ω m d ( ω ) [ A 1 ( ω ) ξ m ( ω ) + A 2 ( ω ) ξ M ( ω ) ] + 2 κ ω m d ( ω ) [ A 3 ( ω ) δ X i n ( ω ) + A 4 ( ω ) δ Y i n ( ω ) ] ,
A 1 ( ω ) = ( κ + i ω ) 2 ( ω 2 + i γ M ω + ω M 2 ) Δ c ( ω 2 Δ c 2 ω M G 2 a 2 Δ c ω M 2 i ω γ M Δ c ) , A 2 ( ω ) = ω M [ G 12 ( κ + i ω ) 2 + 2 G 1 a G 2 a Δ c + G 12 Δ c 2 ] , A 3 ( ω ) = 2 ( κ + i ω ) [ G 1 a ( ω 2 ω M 2 i ω γ M ) + G 12 G 2 a ω M ] , A 4 ( ω ) = 2 Δ c [ G 1 a ( ω M 2 ω 2 + i ω γ M ) G 12 G 2 a ω M ] , d ( ω ) = ( κ + i ω ) 2 B 0 Δ c B 1 ,
S y ( ω ) = d Ω 2 π e i ( ω + Ω ) t δ y ( ω ) δ ( Ω ) + δ y ( Ω ) δ ( ω ) 2 ,
ξ j ( ω ) ξ j ( Ω ) = 2 π γ j ω j ω [ 1 + coth ( ω 2 k B T j ) ] δ ( ω + Ω ) , δ X i n ( ω ) δ X i n ( Ω ) = δ Y i n ( ω ) δ Y i n ( Ω ) = π δ ( ω + Ω ) .
S y ( ω ) = | χ ( ω ) | 2 [ γ m ω ω m β m + γ M ω ω M | A 2 ( ω ) A 1 ( ω ) | 2 β M ] + | χ ( ω ) | 2 κ [ | A 3 ( ω ) A 1 ( ω ) | 2 + | A 4 ( ω ) A 1 ( ω ) | 2 ] ,
ω m e f f = Re [ d ( ω ) / A 1 ( ω ) + ω 2 ]
γ m e f f = Im [ d ( ω ) / A 1 ( ω ) ] / ω
a + = ε p [ β 3 i G 1 a i G 2 a ( G 1 a α 2 + G 2 a α 1 ) 2 β 4 + i G 1 a α 2 + i G 2 a α 1 ] 1 ,
2 κ a o u t ( t ) = 2 κ α s ε L + ( 2 κ a + ε p ) e i δ t + 2 κ ( a ) e i δ t .
ε o u t = 2 κ a + / ε p ,
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.