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

Squeezing transfer of light in a two-mode optomechanical system

Open Access Open Access

Abstract

The squeezing transfer from a squeezed vacuum injected in one cavity to the output spectrum of the other cavity in an optomechanical system is investigated. By calculating the noise spectrum of the output field, it is found that two squeezing dips appear symmetrically located about the resonant point. Besides the contribution from the destructive interference between the noise fluctuation of the input field and its optomechanically modified one, the major part of the squeezing is transferred from the squeezed vacuum injected in the cavity. Additionally, it is shown that the adverse effects of the environment temperature on the output spectrum can be strongly suppressed by the injected squeezed field. This study can be useful in quantum communications via the optomechanical interface.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

A canonical optomechanical system (OMS) is composed by a fixed mirror and a movable mirror which is coupled to the optical field in the cavity by the radiation pressure [1–3]. The OMS has attracted much attention due to its many important applications in quantum optics and information science, such as in the stationary entanglement [4], the cooling of mechanical oscillators to their quantum ground states [5–8], the detection of ultrahigh precision [9,10], the demonstration of the quantum nonlinearities [11,12], the generation of the macroscopic quantum superposition [13], and optomechanically induced transparency (OMIT) [14–17].

In particular, the squeezings in the optomechanical system have attracted researchers’ attention. It is well known that the squeezed light, which was proposed nearly three decades ago, has less fluctuation in one quadrature than that in a coherent state at the expense of increased fluctuation in the other quadrature [18,19]. The squeezing plays an important role in improving the identification sensitivity of gravitational-wave detection [20] and in continuous-variable information processing [21]. The squeezings in the optomechanical systems have been studied extensively. These investigations are mainly devoted to the mechanical and optical squeezings in the optomechanical system. For example, large degrees of squeezing of a mechanical micromirror in periodically amplitude-modulated optomechanical systems were achieved [22]. The squeezing of the nanoresonator motion in a system composed of a parametrically driven nanomechanical resonator capacitively coupled to a microwave cavity was studied [23]. The generation and detection of the measurement-induced squeezing of a cantilever in optomechanics were studied by using back-action free single-quadrature detection [24]. The momentum quadrature squeezing of a waveguide by injecting a quantum field and laser into the resonator through the waveguide in a dissipative optomechanical system was investigated [25]. Also, the quadrature squeezing of the mechanical oscillator in an optomechanical system can be created by using an field with a large detuning to drive the cavity [26]. The generation of the two-mode squeezed vacuum of the mechanical oscillators in the two-mode optomechanical system was discussed [27,28].

Besides the mechanical squeezing in the optomechanical system was studied, the squeezing of the cavity field in an optomechanical system has been investigated. Firstly, the generation of the output quadrature squeezing, i.e. ponderomotive squeezing produced in an optomechanical system was analysed in Refs. [29,30]. Subsequently, the phase-noise power spectrum of the output light in a cavity with a movable mirror subject to quantum Brownian motion was studied by using the quantum Langevin approach [31]. The quadrature squeezing of the cavity output field, which is induced by the combined effects of dispersive and dissipative couplings, in an optomechanical system was theoretically demonstrated [32]. And, the generation of a two-mode broadband squeezed light was investigated in the cascading double-cavity optomechanical systems [33,34]. In a quadratically coupled optomechanical system, the squeezing of the mechanical oscillator and that of the optical output field were discussed [35,36].

Correspondingly, the experimental investigations of the squeezing in the optomechanical system have been demonstrated. The ponderomotive squeezing induced by the radiation-pressure rigidity was experimentally analyzed [37]. Recently, the ponderomotive squeezing, produced by the back-action of the ultracold atoms’ collective motion which is analogous to the motion of a mirror in a cavity optomechanics was detected [38]. The squeezing of the reflected light’s fluctuation spectrum at the few percent level around the mechanical resonance frequency in the MHz range was observed through homodyne detection in an optomechanical system comprising a micromechanical resonator coupled to a nanophotonic cavity [39], and the production of ponderomotive squeezing in the kilohertz range was demonstrated experimentally in an optomechanical system formed by a Fabry-Pérot cavity with a micromechanical mirror [40]. A larger squeezing of 1.7 dB below the shot-noise level from the optomechanical system composed of a semitransparent membrane placed inside an optical cavity was observed [41].

Besides the optomechanical squeezings discussed above, the reversible transfer of the quantum information between the mechanical oscillator and the electromagnetic field is another remarkable property of the optomechanical system, which attracts much attention due to its potential applications for enabling the distribution of quantum information among distant quantum systems. One kind of the coherent optomechanical transfer is the quantum state transfer between the mechanical oscillator and the electromagnetic field via optomechanical interface. The scheme based on the radiation pressure effects for transferring quantum states as well as entanglement from the propagating light fields to macroscopic, collective vibrational degree of freedom of a massive mirror was proposed [42]. Additionally, the coherent state swapping between two spatially and frequency separated resonators was demonstrated [43]. And, Tian has investigated the transfer of the quantum states between different cavity modes by adiabatically varying the effective optomechanical couplings [44]. The coherent transfer of optical or microwave photon states to mechanical modes was experimentally demonstrated [45–47].

Another kind of coherent optomechanical transfer is the coherent wavelength conversion between photons with vastly differing wavelengths, which is mediated by a mechanical mode coupling to two cavity fields by radiation pressures in a two-optical-mode optomechanical system [48–50]. To achieve efficient conversion fidelity, Wang et al proposed an effective mechanically dark mode, which is immune to mechanical dissipation, to perform high-fidelity quantum conversion between optical and microwave cavities [49]. And, the coherent wavelength conversion of optical photons using photon-phonon translation in a cavity-optomechanical system was theoretically proposed and experimentally demonstrated [50].

Although the coherent optomechanical transfer including quantum state and wavelength conversion has been extensively studied, it is necessary to consider the squeezing transfer between two cavity fields via optomechanical interface due to the fact that the optical squeezing is one of essential properties of light and has application in continuous-variable information processing. The squeezing of a mechanical oscillator within an optical cavity generated from a squeezed vacuum under the condition of resolved-sideband ground state cooling [51], and in the non-resolved-sideband regime [52] has been investigated. To our knowledge, the squeezing transfer between two electromagnetic cavities (e.g., optical and microwave) via optomechanical interface has not been considered yet. The squeezing transfer can facilitate the development of scalable quantum information processors with applications in conversion of information between optical and microwave photons.

In the present paper, we shall investigate the squeezing transfer from a squeezed vacuum injected in one cavity (left-hand cavity) to the output spectrum from the other cavity (right-hand cavity) in a two-mode optomechanical system, in which the two cavity are coupled to a common mechanical oscillator by radiation pressure. It is shown that the squeezing of the output spectrum in two symmetrical region is partially contributed by the destructive interference between the noise fluctuation of the input field and its optomechanically modified one, and mainly by the injected squeezed vacuum. This paper is organized as follows. In Sec.2, we present the model and give the quantum Langevin equations of the system. In Sec.3, we study the squeezing of the output field in the two-mode optomechanical system by calculating the quadrature spectrum of the output field from the right-hand cavity with a shot noise injected into the left-hand cavity. We investigate the quadrature squeezing of the output field in the case where a squeezed vacuum is injected into the left-hand cavity in Sec.4. A brief conclusion is presented in Sec.5.

2. Model and dynamical equation

A two-mode optomechanical system under our consideration, illustrated in Fig. 1, consists of two optical cavities which are coupled to a common mechanical oscillator by the radiation pressures. To enhance the optomechanical interactions of the two cavities with the oscillator, two coupling fields with power Pk (k = 1, 2) and amplitude Ek=2Pkκk/ωk are, respectively, used to drive the two cavities, in which ωk and κk denote the frequency of coupling fields and the loss rate of the cavity fields. Here, âk(a^k) is the annihilation (creation) operator of the kth cavity field with resonance frequency ωck. And, () denotes the annihilation (creation) operator of the movable mechanical mode with its resonance frequency, effective mass and quality factor denoted by ωm, m and Qm, respectively. The Hamiltonian of the system in the rotating frame at the frequency ωk (k = 1, 2) of the coupling fields reads (ħ = 1) [33,34,48–50]

H=k=1,2Δckaka^k+ωmb^b^(g1a^1a^1g2a^2a^2)(b^+b^)+k=1,2iEk(a^ka^k).
Here, Δck = ωckωk(k = 1, 2) is the detuning of the kth cavity field from the corresponding coupling field. The coupling between the kth cavity field and the movable mechanical oscillator via radiation pressure is described by the term with the coupling parameter gk, which is given by gk = ωck xzpt /Lk with xzpt being the zero-point fluctuation amplitude.

 figure: Fig. 1

Fig. 1 Schematic of the two-cavity optomechanical system. The system consists of two optical cavities which are coupled to a common mechanical oscillator by the radiation pressures. Two coupling fields with power Pk (k = 1, 2) and frequency ωk are used to drive the two cavities, respectively.

Download Full Size | PDF

The dynamical evolution of the system can be described by the quantum Heisenberg-Langevin equations of the cavity and oscillator variables, which are given by

a^˙1=(iΔc1+κ1)a^1+ig1a^1(b^+b^)+E1+2κ1a^in,1,
a^˙2=(iΔc2+κ2)a^2ig2a^2(b^+b^)+E2+2κ2a^in,2,
b^˙=(iωm+γm)b^+i(g1a^1a^1g2a^2a^2)+b^in
where the operator âin,k(k = 1, 2) describes the input field in the kth cavity with zero mean value. γm=ωmQm is the damping rate and in describes the Brownian force of the mechanical resonator. The equations of the cavity and oscillator variables are nonlinear ones because of the terms with gk contributed by the optomechanical interactions. Due to the driving of the strong coupling fields on the cavities, we can divide the cavity and mechanical operators in Eqs. (2)(4) into the steady-state parts and fluctuation ones, i.e. ô = os + δô with ô = âk, . The dynamical behavior of the system can be exhibited by solving the equations of motion for the fluctuations (δâk, δb̂) around their steady-state parts (aks, bs). By setting the derivatives of the dynamical equations to be zero, the steady-state solutions of the system can be obtained as
aks=Ekκk+iΔk,(k=1,2)
bs=ig1a1sa1s*ig2a2sa2s*γm+iωm
where Δk = Δck + (−1)k gk(bs+bs*), (k = 1, 2) is an effective detuning. By neglecting the higher-order terms of the fluctuation operators, the linearized quantum Langevin equations of the fluctuation parts are given by
δa^˙k=(iΔk+κk)δa^k(1)kigkaks(δb^+δb^)+2κkδa^in,k,(k=1,2)
δb^˙=(iωm+γm)δb^+i[g1(a1s*δa^1+a1sδa^1)g2(a2s*δa^2+a2sδa^2)]+δb^in.
The input squeezed vacuum noise operator δâin,k and the mechanical input operator δb̂in in Eqs. (7) and (8) satisfy the following correlation relations: 〈δâin,j(ω)δâin,k(Ω)〉 = 2πMδ(ω+Ω−2ωm)δj,k, δa^in,j(ω)δa^in,k(Ω)=2π(N+1)δ(ω+Ω)δj,k, and δb^in(ω)δb^in(Ω)=2πγm(nm+1)δ(ω+Ω). Here, N = sinh[r]2 and M = sinh[r] cosh[r]e in which r is the squeezing parameter and φ the phase of the squeezed vacuum light. And, nm=(eωmκBT1) is the mean thermal photon occupation number of energy ħωm at environment temperature T with κB being Boltzmans constant. For simplicity, we choose φ = 0. Here, we mainly consider the squeezing transfer from the squeezed field injected into the left-hand cavity to the output spectrum in the right-hand cavity, then we assume that the input field into the right-hand cavity is in a vacuum (white) noise, i.e. 〈δâin,2(ω)δâin,2(Ω)〉 = 0 and δa^in,2(ω)δa^in,2(Ω)=2πδ(ω+Ω).

When transforming the Langevin equations of the fluctuation parts shown in Eqs. (7) and (8) into the frequency domain by using the Fourier transform and substituting the fluctuations into the input-output formula δa^out,2(ω)=2κ2δa^2(ω)δa^in,2(ω), the fluctuation of the output field from the second cavity is obtained as

δa^out,2(ω)=C1(ω)δa^in,1+C2(ω)δa^in,1(ω)+V1(ω)δa^in,2+V2(ω)δa^in,2(ω)+W1(ω)δb^in(ω)+W2(ω)δb^in(ω)
where
C1(ω)=2ig1g2a1s*a2sκ1κ2Γmm[κ1+i(Δ1ω)][κ2+i(Δ2ω)],
C2(ω)=2ig1g2a1sa2sκ1κ2Γmm[κ1+i(Δ1ω)][κ2+i(Δ2ω)],
V1(ω)=2ig22|a2s|2κ2Γmm[κ2+i(Δ2ω)][κ2+i(Δ2ω)]+2κ2κ2+i(Δ2ω)1,
V2(ω)=2ig22a2s2κ2Γmm[κ2+i(Δ2ω)][κ2+i(Δ2+ω)],
W1(ω)=ig2a2s2κ2Γmbκ2+i(Δ2ω),
W2(ω)=ig2a2s2κ2Γmaκ2+i(Δ2ω),
Γma=Γ+Γ+Γ2iωm(Λm+iδm),
Γmb=ΓΓ+Γ2iωm(Λm+iδm),
Γmm=2ωmΓ+Γ2iωm(Λm+iδm)
with Γ+ = γm + i(ωmω), Γ = γmi(ωm + ω), δm=j=1,2Im[1κj+i(Δjω)1κji(Δj+ω)], and Λm=j=1,2Re[1κj+i(Δjω)1κji(Δj+ω)]. The output spectrum of the optical field from the right-hand cavity is defined by
Sθ(ω)=12πdΩδX^θout(ω)δX^θout(Ω),
in which the quadrature of the output field is given by δX^θout(ω)=eiθδa^out,2(ω)+eiθδa^out,2+(ω). From the Eq. (9), we get the noise spectrum of the output field Sθ(ω), which is given by
Sθ(ω)=MF1+(N+1)F2+NF3+M*F4+G1+γm(nb+1)H1+γmnbH2,
where
F1=C1(ω)C1(Ω1)e2iθ+C1(ω)C2*(Ω1)+C2*(ω)C1(Ω1)+C2*(ω)C2*(Ω1)e2iθ,
F2=C1(ω)C2(Ω1)e2iθ+C1(ω)C1*(Ω2)+C2*(ω)C2(Ω2)+C2*(ω)C1*(Ω2)e2iθ,
F3=C2(ω)C1(Ω2)e2iθ+C2(ω)C2*(Ω2)+C1*(ω)C1(Ω2)+C1*(ω)C2*(Ω2)e2iθ,
F4=C2(ω)C2(Ω3)e2iθ+C2(ω)C1*(Ω3)+C1*(ω)C2(Ω3)+C1*(ω)C1*(Ω3)e2iθ,
G1=V1(ω)V2(Ω2)e2iθ+V1(ω)V1*(Ω2)+V2*(ω)V2(Ω2)+V2*(ω)V1*(Ω2)e2iθ,
H1=W1(ω)W2(Ω2)e2iθ+W1(ω)W1*(Ω2)+W2*(ω)W2(Ω2)+W2*(ω)W1*(Ω2)e2iθ,
H2=W2(ω)W1(Ω2)e2iθ+W2(ω)W2*(Ω2)+W1*(ω)W1(Ω2)+W1*(ω)W2*(Ω2)e2iθ
with Ω1 = −ω + 2ωm, Ω2 = −ω and Ω3 = −ω − 2ωm. The output field from the right-hand cavity is squeezed if the quadrature spectrum meets Sθ(ω) < 1.

3. Quadrature squeezing generated by optomechanical interactions

We investigate the squeezing transfer between the cavities via optomechanical interface by calculating the noise spectrum of the output field from the right-hand optomechanical cavity. To distinguish the squeezing of the output field transferred from the input squeezed field in the left-hand cavity field from that originated by the optomechanical interactions in the system, we firstly investigate the variation of the noise spectrum with the two coupling fields in the case where the left-hand cavity is in a quantum vacuum noise, i.e. 〈δâin,2(ω)δâin,2(Ω)〉 = 0 and δa^in,2(ω)δa^in,2(Ω)=2πδ(ω+Ω). To make the following results within experimental realizations, we use the parameters in a recent experiment for demonstrating coherent wavelength conversion of photons using photon-phonon translation in a cavity-optomechanical system [50]: ωm = 2π × 3.993 × 109Hz, Qm = 87 × 103, g1 = 2π × 960 × 103, g2 = 2π × 430 × 103, κ1 = 2π ×520×106Hz, κ2 = 1.73×109Hz, ω1 = 2π ×205.3×1012Hz, ω2 = 2π ×194.1×1012Hz. The following discussions are explored under the resonance condition of Δ1 = Δ2 = ωm. The resonance relation is the key condition for transferring the squeezing between the cavity fields in the present optomechanical system, which is addressed in the following section.

First, we consider the effects of the left coupling field (driving the left-hand cavity) on the noise spectrum Sθ(ω) when the left-hand cavity is injected by the quantum vacuum noise. In Fig. 2(a), the power of the left coupling field is set by different values as 1mW, 5mW, 10mW and that of the right coupling field (driving the right-hand cavity) is fixed at 1mW. It is shown that there appears a symmetric noise spectrum, in which the spectra around ω = ±ωm display dispersion-like behavior and the other parts is in shot-noise. The dips with Sθ(ω) < 1 near ω = ±ωm indicate the production of the squeezing of the output field, and the peaks with Sθ(ω) > 1 describe the optomechanically induced heating. Additionally, the squeezing degree is decreased with the left coupling filed. Second, the effects of the right coupling field on the noise spectrum are displayed in Fig. 2(b). At this time, the two symmetric dips become deeper and the squeezing of the output field is increased with the right coupling field. Correspondingly, the optomechanically induced heating becomes stronger and makes the noise peaks higher.

 figure: Fig. 2

Fig. 2 The squeezing spectra Sθ (ω) as a function of the normalized frequency ω/ωm with P2 = 1 mW and for the different powers of the left coupling field (a): P1 = 1 mW (black, solid curve); 5 mW (red, dashed curve); 10 mW (blue, dotted curve). The squeezing spectra Sθ (ω) as a function of the normalized frequency ω/ωm with P1 = 1 mW and for the different powers of the right coupling field (b): P2 = 1 mW (black, solid curve); 5 mW (red, dashed curve); 10 mW (blue, dotted curve). The values of the parameters are given by: ωm = 2π × 3.993 × 109Hz, g1 = 2π × 960 × 103, g2 = 2π × 430 × 103, T = 10K, r = 0, Qm = 87 × 103, κ1 = 2π × 520 × 106Hz, κ2 = 1.73 × 109Hz, ω1 = 2π × 205.3 × 1012Hz, ω2 = 2π × 194.1 × 1012Hz, θ = π/2.

Download Full Size | PDF

The optomechanically induced heating describes the thermal phonon excitation, which is induced by the photonic excitation in the optomechanical interaction and is related to the thermal phonons (or the environment temperature T). This can be verified by Fig. 6(a) in the following section, in which the peaks are dramatically increased with the environment temperature T. Specifically, the thermal phonon excitation is described by the last two terms in Eq. (20), which are related to the thermal phonon occupation number at environment temperature T. These two terms are respectively originated from the last two terms in Eq. (9) with the coefficients addressed by the linearized coupling g2a2s, which describe the thermal noise from the environment of the mechanical oscillator through the optomechanical interaction between the second cavity and the mechanical oscillator. Also, this can explain why the peaks in Fig. 2(b) are remarkably increased with the power P2.

The squeezing of the output field induced by the optomechanical interactions can be understood by using the quadrature fluctuations of cavity and mechanical variables, which are defined as A^k=δa^k+δa^k+ (k = 1, 2), B^k=1i(δa^kδa^k+), = δb̂ + δb̂+ and P^=1i(δb^δb^+). In fact, the quadrature δX^θout(ω) with θ = π/2 adopted in the present paper is the phase quadrature, i.e. δX^θ=π/2out(ω)=B^2(ω). Without loss of generality, the steady-state variable aks is assumed to be real. Then the phase quadrature 2 in the frequency domain is

B^2(ω)=2(κ2iω)g2a2s(κ2iω)2+Δ22Q^(ω)Δ22κ2(κ2iω)2+Δ22A^in,2+(κ2iω)2κ2(κ2iω)2+Δ22B^in,2.
The first term comes from the optomechanical interaction between the right-hand cavity and the mechanical oscillator, and the next two terms are resulted from the noise of the input field in the right-hand cavity. Correspondingly, the position operator is obtained as
Q^(ω)=1D(2κ1K1Ω1A^in,1+2κ1Δ1Ω1B^in,12κ2K2Ω2A^in,22κ2Δ2Ω2B^in,2+ΓmR1R2Q^in+ωmR1R2P^in)
where D = RmR1R2 − 2(G1Δ1Ω1 + G2Δ2Ω2) with Ωi = 2ωmGiRj, (i, j = 1, 2, ij), Rm=Γm2+ωm2, Ri=Ki2+Δi2, Gi = giais, Γm = γm, Ki = ki. The first (next) two terms in Eq. (29) are contributed from the input vacuum in left(right)-hand cavity via their optomechanical interactions. The last two terms describe the contributions from the mechanical damping and the quantum Brownian noise. Substitution of Eq. (29) into Eq. (28) leads to the phase quadrature as
B^2(ω)=a11A^in,1+b11B^in,1+a22A^in,2+b22B^in,2+q11Q^in+p11P^in
where
a11=1R2D×2K1K2G2Ω12κ1,
b11=1R2D×2Δ1K2G2Ω12κ1,
a22=1R2D×2κ2(Δ2D2K22G2Ω2),
b22=1R2D×2κ2K2(D+2Δ22G2Ω2),
q11=1R2D×2K2G2R1R2Γm,
p11=1R2D×2K2G2ωmR1R2.

In above discussions, the squeezing of the output field from the right-hand cavity with a quantum vacuum injected in the left-hand cavity is increased with the right coupling field, while it is decreased with the left coupling field. This can be explained by using quadrature spectrum shown in Eq. (30). In the present system, there exist the optical noises from the cavities and the mechanical thermal noise affecting the squeezing of the output spectrum. The optical noise associated with the second cavity is fed into the output spectrum through two ways: (i) it directly enters into the second cavity; and (ii) it is indirectly transferred into the second cavity as a modified noise through the optomechanical interaction between the mechanical oscillator and the second cavity field. This can be verified by the third and fourth terms in Eq. (30), in each coefficient of which there are two terms: the first part in Eqs. (33) and (34) is directly contributed from the noise and the second one comes from the optomechanically modified noise dressed by the linearized coupling G2. It is the destructive interference between the two paths that leads to the squeezing in the output spectrum. Consequently, the squeezing is enhanced by the right coupling field through the optomechanically modified noise, which is shown by the Figs. 2(b) and 3(b).

 figure: Fig. 3

Fig. 3 The same settings are used as those in Fig. 2 except for the presence of squeezed vacuum injected in the left cavity with squeezing parameter r = 1.

Download Full Size | PDF

On the hand, the optical noise from the left cavity is transferred to the output spectrum only through the optomechanical interactions and enhances the fluctuation in the noise spectrum. Specifically, the modified noise from the left cavity by the optomechanical interactions is described by the first two terms in Eq. (30), in each of which there is only one term of the modified fluctuation by the optomechanical interactions. When the left coupling field is increased, the modified noise is stronger and the squeezing in the output field is suppressed. This is why the squeezing is decreased with the left coupling field, which are shown in Figs. 2(a) and 3(a).

4. Squeezing transfer of light between two cavities via the optomechanical interactions

In the following, we shall discuss the squeezing of the output field from the right-hand cavity when the left-hand cavity is injected by a squeezed vacuum. To effectively contrast the squeezing properties between the cases of the squeezed field and the quantum vacuum injected in the left-hand cavity, we use same settings in Fig. 3 as those in Fig. 2 except that a squeezed vacuum with squeezing parameter r = 1 is injected in the left-hand cavity. In Fig. 3(a), the power of the left coupling field is set by different values as 1mW, 5mW, 10mW and that of the right coupling field is fixed at 1mW. It is shown that the maximum squeezing around ω = ±ωm becomes smaller when the left coupling field becomes stronger. This is similar to that shown in Fig. 2(a). However, the squeezing dips become deeper and wider than the corresponding curves in Fig. 2(a) due to the effects from the injected squeezed field. The effects of the right coupling field on the squeezing of the output spectrum in the case of the squeezed field injected into the left-hand cavity are exhibited in Fig. 3(b). It is shown that the maximal squeezing is increased with the power of the right coupling. By comparing with Fig. 2(b), the squeezing amount in Fig. 3(b) is much larger than that in the corresponding curve in the case of injected vacuum, which is shown in Fig. 2(b). To see clearly the enhancement of the squeezing of the output field due to the squeezed vacuum injected in the left-hand cavity, we plot the noise spectra with the quantum vacuum and the squeezed vacuum injected in the left-hand cavity, which are shown by the solid and dashed curves in Fig. 4, respectively. It is seen that in the case of the squeezed vacuum with r = 1, the amount of the maximal squeezing reaches larger than 60% for P1 = 1mW and P2 = 10mW, while the squeezing shown by the solid curve is 8% for the vacuum injected in the left-hand cavity.

 figure: Fig. 4

Fig. 4 The squeezing spectra Sθ(ω) as a function of the normalized frequency ω/ωm with P1 = 1 mW, P2 = 5 mW for the injected vacuum field r = 0 (black, solid curve), and for the injected squeezed field r = 1 (red, dashed curve). The others are set with the same values as in Fig. 2.

Download Full Size | PDF

By comparing the noise spectra in Fig. 3 with those in Fig. 2, it is shown that the enhanced squeezing in the Fig. 3 mainly comes from the squeezed field injected in the left cavity. To provide a specific demonstration of this squeezing transfer, we shall display the dependence of the squeezing of the output field from the right-hand cavity on the squeezing parameter r of the injected squeezed field. In Fig. 5, the squeezing parameter r is set by different values: r = 0.3, 0.6, 1, which correspond to solid, dashed and dotted curves, respectively. It is shown that the squeezing of the output field from the right-hand cavity is increased with the squeezing parameter r of the injected squeezed field, which maps the essential property of the variation of the injected squeezed field with the squeezing parameter on the output spectrum.

 figure: Fig. 5

Fig. 5 The squeezing spectra Sθ(ω) as a function of the normalized frequency ω/ωm with P1 = 1 mW, P2 = 5 mW and for different squeezing parameter r of the squeezed field injected in the left cavity: r = 0.3(black, solid curve); r = 0.6(red, dashed curve); r = 1(blue, dotted curve). The others are set with the same values as in Fig. 2.

Download Full Size | PDF

On the other hand, the injected squeezed field can mitigate the adverse effects of the temperature on the squeezing of the output field. By comparing with Fig. 2 for a vacuum injected in the left-hand cavity, the peaks in Fig. 3 are dramatically decreased in the case of the injected squeezed field. Now we see the effects of the temperature of the environment on the noise spectra in the cases of a vacuum and a squeezed field injected in the left-hand cavity, which are shown by Figs. 6(a) and 6(b), respectively. Although the noise spectrum in Fig. 6(a) behaves like that in Fig. 3(b), there are different physics mechanisms: the right coupling field in Fig. 3(b) leads to optomechanically induced heating and makes the noise peaks higher, meanwhile it induces the destructive interference of quantum noise and makes the squeezing dips deeper. While the temperature of the environment surrounding the mechanical oscillator in Fig. 6(a) provides the thermal phonons of the mechanical oscillator to increase the noise peaks and suppresses the squeezing induced by the optomechanical interactions. When a squeezed field is injected in the left-hand cavity, shown in Fig. 6(b), the noise spectra are almost invariant with the temperature. This is different from the case of a injected vacuum in the left-hand cavity shown in Fig. 6(a), in which the noise spectrum varies drastically with the temperature. It implies that the adverse effects of the environment temperature on the output spectra can be strongly mitigated by the injected squeezed field.

 figure: Fig. 6

Fig. 6 The spectrum Sθ(ω) (a) as a function of the normalized frequency ω/ωm for the injected vacuum r = 0 (a) and the injected squeezed vacuum r = 1 (b) in the left-hand cavity. The temperature for the mechanical environment are given by different values: T1 = 10K (black, solid curve); 50 K (red, dashed curve); 100 K (blue, dotted curve). The others are set with the same values as in Fig. 2.

Download Full Size | PDF

In the squeezing transfer between the two cavities via the optomechanical interactions, the key condition is the resonance conditions of Δ1 = Δ2 = ωm, which leads the two-mode optomechanical system to be a double-beam-splitter form. In above discussions, we use the linearization approach of quantum optics and get the linearized quantum Heisenberg-Langevin equations for the fluctuations in Eqs. (7) and (8), which corresponds to the linearized Hamiltonian: ĤEĤ1 + Ĥ2 with H^1=k=1,2(1)kGk(δa^kδb^+δa^kδb^) and H^2=k=1,2(1)kGk(δa^kδb^+δa^kδb^). The optomechanical coupling Gk=gknk is enhanced from the single-photon optomechanical coupling gk by the intracavity photon number nk = |aks|2 (k = 1, 2). The first part Ĥ1 of the linearized Hamiltonian describes the two-mode squeezing interaction form, which has been employed for generating the entanglement between optical and mechanical modes and two-mode squeezing [4, 27, 53, 54]. The second part Ĥ2, taking the beam-splitter-like form, can enable quantum transfer between the optical and the mechanical systems, such as state swapping [43,45–47] and optical wavelength conversion [44,48–50].

When the two intense red-detuned coupling fields are applied on the two cavities according to the conditions of Δ1 = Δ2 = ωm, this makes the beam-splitter-like interactions dominant over the two-mode squeezing ones. Thus the optomechanical interactions in Eq. (1) can be effectively characterized by the beam-splitter Hamiltonian shown in Ĥ2. As is proposed that a reversible transfer of light squeezing conveyed to the membrane or mirror via optomechanical interface can be realized under this conditions [51,52]. Then, the squeezing of the squeezed field injected into the left-hand cavity is transferred to the squeezing of the mechanical oscillator, which is in turn conveyed to that of the right-hand cavity. Additionally, the beam-splitter-like interaction, for red-detuned driving at the resonance conditions of Δ1 = Δ2 = ωm, makes energy conversion between the intracavity field and the movable mirror and results in ground-state cooling of the mechanical resonator by using the optical vacuum bath to successively extract the energy. This can suppress the damages of the mechanical heating excitation on the squeezing of the output spectrum.

To further demonstrate the key role played by the beam-splitter-like interaction in the squeezing transfer, we numerically investigate three cases: noise spectra in the single optomechanical cavity under the conditions of Δ2 = ωm and Δ2 = −ωm, and in the two-mode optomechanical system under the condition of Δ2 = −ωm, which are shown in Figs. 7(a)–7(c), respectively. In Fig. 7(a), we investigate the squeezing properties in the single-mode optomechanical system under the condition of Δ2 = ωm by setting g1 = 0. There appear two noise peaks lie in the inner sides of ω = ±ωm and two squeezing dips on their outsides, which become deeper with the right coupling field. The noise spectrum behaves like the case of two-mode optomechanical system, which is shown in Fig. 2(b). When we adjust the right coupling to detune at Δ2 = −ωm in Fig. 7(b), the spectrum properties are different from the case of Δ2 = ωm, shown in Fig. 7(a). Here, the positions of the squeezing dips are changed in the inner sides of ω = ±ωm, and those of the noise peaks are located on their outsides. Additionally, the noise peaks become shorter and the squeezing dips turn slightly deeper with the right coupling field. Thus, the squeezing of the output spectrum with a vacuum injected in the left-hand cavity is mainly resulted from the destructive interference between the noise and the modified one by the optomechanical interaction in the right-hand cavity, which can be described by the beam-splitter-like Hamiltonian δa^2δb^+δa^2δb^ at Δ2 = ωm.

 figure: Fig. 7

Fig. 7 The spectrum Sθ(ω) (a) as a function of the normalized frequency ω/ωm for g1 = 0, r = 0 at Δ2 = ωm (a) and Δ2 = −ωm (b), and for g1 = 2π × 960 × 103, r = 1 at Δ2 = −ωm (c). The power of the right coupling field is set as: P2 = 1 mW (black, solid curve); 3 mW (red, dashed curve); 6 mW (blue, dotted curve). The others are set with the same values as in Fig. 2.

Download Full Size | PDF

In the following, we consider the squeezing transfer in the two-mode optomechanical system under the condition of Δ2 = −ωm when a squeezed vacuum is injected in the first cavity by turning on its optomechanical interaction. Now, the squeezing dips disappear, instead there appear two noise peaks symmetrically located at ω = ±ωm, in which the magnitudes of the peaks are increased with the right coupling field. The comparison in these three cases can show that the squeezing transfer mainly benefits from the beam-splitter-like part of the linearized Hamiltonian δa^2δb^+δa^2δb^ by setting the detuning of the coupling field to the red side of the optomechanical cavity resonance at Δ2 = ωm, which is used to cool the mechanical oscillator as well as to generate a state swap between the mechanical and the optical mode. On the contrary, two-mode squeezing interaction (counter-rotating interaction term) δa^2δb^+δa^2δb^, which can be dominated in the linearized Hamiltonian by tuning the coupling field to the blue side of the optomechanical cavity resonance at Δ2 = −ωm and leads to the optomechanical induced amplification [15], destroys the squeezing transfer in the present model.

5. Conclusions

We have investigated the squeezing of the output field from the right-hand cavity when the left cavity is injected by a squeezed field in a two-mode optomechanical system, and found that there appear two squeezing regions symmetrically located at ω = ±ωm. By comparing the squeezing properties in the case of a squeezed field with that for a vacuum injected in the left-hand cavity, we find that the squeezing of the output field from the right-hand cavity partially comes from the optomechanical interactions, and it is mainly transferred from the squeezed field injected in the left-hand cavity. Additionally, the squeezing amount is increased with the right coupling field but it is decreased with left coupling field. Finally, we find that the injected squeezed field can mitigate the effects of the temperature. Combination of the quantum transfer and the continuous-variable information processing related to the squeezed state can be useful in quantum communications in the optomechanical interface.

Funding

National Natural Science Foundation of China (10647007, 11874273); Science Foundation of Sichuan Province of China (2018JY0180).

References

1. P. Meystre, ”A short walk through quantum optomechanics,” Ann. Phys. 525(3), 215–233 (2013). [CrossRef]  

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, and A. Ferreira, ”Optomechanical entanglement between a movable mirror and a cavity field,” Phys. Rev. Lett. 98(3), 030405 (2007). [CrossRef]   [PubMed]  

5. 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]  

6. J. D. Teufel, T. Donner, D. Li, J. W. Harlow, and M. S. Allman, ”Sideband cooling of micromechanical motion to the quantum ground state,” Nature(London) 475(7356), 359–363 (2011). [CrossRef]  

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

8. D.-G. Lai, F. Zou, B.-P. Hou, Y.-F. Xiao, and J.-Q. Liao, “Simultaneous cooling of coupled mechanical resonators in cavity optomechanics,” Phys. Rev. A 98(2), 023860 (2018). [CrossRef]  

9. M. D. LaHaye, O. Buu, B. Camarota, and K. C. Schwab, ”Approaching the quantum limit of a nanomechanical resonator,” Science 304(5667), 74–77 (2004). [CrossRef]   [PubMed]  

10. A. A. Geraci, S. B. Papp, and J. Kitching, ”Short-range force detection using optically cooled levitated microspheres,” Phys. Rev. Lett. 105(10), 101101 (2010). [CrossRef]   [PubMed]  

11. P. Rabl, ”Photon blockade effect in optomechanical systems,” Phys. Rev. Lett. 107(6), 063601 (2011). [CrossRef]   [PubMed]  

12. A. Nunnenkamp, K. Børkje, and S. M. Girvin, ”Single-photon optomechanics,” Phys. Rev. Lett. 107(6), 063602 (2011). [CrossRef]   [PubMed]  

13. J. Q. Liao and L. Tian, “Macroscopic quantum superposition in cavity optomechanics,” Phys. Rev. Lett. 116(16), 163602 (2016). [CrossRef]   [PubMed]  

14. S. Weis, R. Rivière, S. Deléglise, E. Gavartin, O. Arcizet, A. Schliesser, and T. J. Kippenberg, “Optomechanically induced transparency,” Science 330(6010), 1520–1523 (2010). [CrossRef]   [PubMed]  

15. A. H. Safavi-Naeini, T. P. Mayer Alegre, and J. Chan, ”Electromagnetically induced transparency and slow light with optomechanics,” Nature(London) 472(7341), 69–73 (2011). [CrossRef]  

16. G. S. Agarwal and S. Huang, ”Normal mode splitting and antibunching in stokes and anti-stokes processes in cavity optomechanics: radiation pressure induced four-wave mixing cavity optomechanics,” Phys. Rev. A 81(3), 033830 (2010). [CrossRef]  

17. B. P. Hou, L. F. Wei, and S. J. Wang, “Optomechanically induced transparency and absorption in hybridized optomechanical systems,” Phys. Rev. A 92(3), 033829 (2015). [CrossRef]  

18. C. M. Caves, “Quantum-mechanical noise in an interferometer,” Phys. Rev. D 23(8), 1693 (1981). [CrossRef]  

19. D. F. Walls, “Squeezed states of light,” Nature(London) 306, 141–146 (1983). [CrossRef]  

20. R. X. Adhikari, “Gravitational radiation detection with laser interferometry,” Rev. Mod, Phys. 86(1), 121–151 (2014). [CrossRef]  

21. S. L. Braunstein and P. V. Loockand, “Quantum information with continuous variables,” Rev. Mod. Phys. 77(2), 513–577 (2005). [CrossRef]  

22. A. Mari and J. Eisert, “Gently modulating optomechanical systems,” Phys. Rev. Lett. 103(21), 213603 (2009). [CrossRef]  

23. M. J. Woolley, A. C. Doherty, G. J. Milburn, and K. C. Schwab, “Nanomechanical squeezing with detection via a microwave cavity,” Phys. Rev. A 78(6), 062303 (2008). [CrossRef]  

24. A. A. Clerk, F. Marquardt, and K. Jacobs, “Back-action evasion and squeezing of a mechanical resonator using a cavity detector,” New J. Phys. 10, 095010 (2008). [CrossRef]  

25. S. Huang and G. S. Agarwal, “Reactive coupling can beat the motional quantum limit of nanowaveguides coupled to a microdisk resonator,” Phys. Rev. A 82(3), 033811 (2010). [CrossRef]  

26. J.-Q. Liao and C. K. Law, “Parametric generation of quadrature squeezing of mirrors in cavity optomechanics,” Phys. Rev. A 83(3), 033820 (2011). [CrossRef]  

27. H. Tan, G. Li, and P. Meystre, “Dissipation-driven two-mode mechanical squeezed states in optomechanical systems,” Phys. Rev. A 87(3), 033829 (2013). [CrossRef]  

28. M. J. Woolley and A. A. Clerk, “Two-mode squeezed states in cavity optomechanics via engineering of a single reservoir,” Phys. Rev. A 89(6), 063805 (2014). [CrossRef]  

29. S. Mancini and P. Tombesi, “Quantum noise reduction by radiation pressure,” Phys. Rev. A 49(5), 4055 (1994). [CrossRef]   [PubMed]  

30. C. Fabre, M. Pinard, S. Bourzeix, A. Heidmann, E. Giacobino, and S. Reynaud, “Quantum-noise reduction using a cavity with a movable mirror,” Phys. Rev. A 49(2), 1337 (1994). [CrossRef]   [PubMed]  

31. 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]  

32. K. Qu and G. S. Agarwal, “Generating quadrature squeezed light with dissipative optomechanical coupling,” Phys. Rev. A 91(6), 063815 (2015). [CrossRef]  

33. K. Qu and G. S. Agarwal, “Strong squeezing via phonon mediated spontaneous generation of photon pairs,” New J. Phys. 16, 113004 (2014). [CrossRef]  

34. Z. Li, S.-l. Ma, and F.-l. Li, “Generation of broadband two-mode squeezed light in cascaded double-cavity optomechanical systems,” Phys. Rev. A 92(2), 023856 (2015). [CrossRef]  

35. A. Nunnenkamp, K. Borkje, J. G. E. Harris, and S. M. Girvin, “Generation of broadband two-mode squeezed light in cascaded double-cavity optomechanical systems,” Phys. Rev. A 82(2), 021806(R) (2010). [CrossRef]  

36. M. Asjad, G. S. Agarwal, M. S. Kim, P. Tombesi, G. D. Giuseppe, and D. Vitali, “Robust stationary mechanical squeezing in a kicked quadratic optomechanical system,” Phys. Rev. A 89(2), 023849 (2014). [CrossRef]  

37. T. Corbitt, Y. Chen, F. Khalili, D. Ottaway, S. Vyatchanin, S. Whitcomb, and N. Mavalvala, “Squeezed-state source using radiation-pressure-induced rigidity,” Phys. Rev. A 73(2), 023801 (2006). [CrossRef]  

38. D. W. C. Brooks, T. Botter, S. Schreppler, T. P. Purdy, N. Brahms, and D. M. Stamper-Kurn, “Non-classical light generated by quantum-noise-driven cavity optomechanics,” Nature(London) 488, 476–480 (2012). [CrossRef]  

39. A. H. Safavi-Naeini, S. Groblacher, J. T. Hill, J. Chan, M. Aspelmeyer, and O. Painter, “Squeezed light from a silicon micromechanical resonator,” Nature(London) 500, 185–189 (2013). [CrossRef]  

40. A. Pontin, C. Biancofiore, E. Serra, A. Borrielli, F. S. Cataliotti, F. Marino, G. A. Prodi, M. Bonaldi, F. Marin, and D. Vitali, “Frequency-noise cancellation in optomechanical systems for ponderomotive squeezing,” Phys. Rev. A 89(3), 033810 (2014). [CrossRef]  

41. T. P. Purdy, P.-L. Yu, R. W. Peterson, N. S. Kampel, and C. A. Regal, “Strong optomechanical squeezing of light,” Phys. Rev. X 3(3), 031012 (2013).

42. J. Zhang, K. Peng, and S. L. Braunstein, “Quantum-state transfer from light to macroscopic oscillators,” Phys. Rev. A 68(1), 013808 (2003). [CrossRef]  

43. M. J. Weaver, F. Buters, F. Luna, H. Eerkens, K. Heeck, S. de Man, and D. Bouwmeester, “Coherent optomechanical state transfer between disparate mechanical resonators,” Nat. Commun. 8, 824 (2017). [CrossRef]  

44. L. Tian, “Adiabatic state conversion and pulse transmission in optomechanical systems,” Phys. Rev. Lett. 108(15), 153604 (2012). [CrossRef]   [PubMed]  

45. V. Fiore, Y. Yang, M. C. Kuzyk, R. Barbour, L. Tian, and H. Wang, “Storing optical information as a mechanical excitation in a silica optomechanical resonator,” Phys. Rev. Lett. 107(13), 133601 (2011). [CrossRef]   [PubMed]  

46. 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 (London) 482, 63–67 (2012). [CrossRef]  

47. T. A. Palomaki, J. W. Harlow, J. D. Teufel, R. W. Simmonds, and K. W. Lehnert, “Coherent state transfer between itinerant microwave fields and a mechanical oscillator,” Nature(London) 495, 210–214 (2013). [CrossRef]  

48. L. Tian and H. Wang, “Optical wavelength conversion of quantum states with optomechanics,” Phys. Rev. A 82(5), 053806(2010). [CrossRef]  

49. Y.-D. Wang and A. A. Clerk, “Using interference for high fidelity quantum state transfer in optomechanics,” Phys. Rev. Lett. 108(15), 153603 (2012). [CrossRef]   [PubMed]  

50. J. T. Hill, A. H. Safavi-Naeini, J. Chan, and O. Painter, “Coherent optical wavelength conversion via cavity optomechanics,” Nat. Commun. 3, 1196 (2012). [CrossRef]  

51. K. Jahne, 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]  

52. 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]  

53. M. Paternostro, D. Vitali, S. Gigan, M. S. Kim, C. Brukner, J. Eisert, and M. Aspelmeyer, “Creating and probing multipartite macroscopic entanglement with light,” Phys. Rev. Lett. 99(25), 250401 (2007). [CrossRef]  

54. M. J. Woolley and A. A. Clerk, “Two-mode back-action-evading measurements in cavity optomechanics,” Phys. Rev. A 87(6), 063846 (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 Schematic of the two-cavity optomechanical system. The system consists of two optical cavities which are coupled to a common mechanical oscillator by the radiation pressures. Two coupling fields with power Pk (k = 1, 2) and frequency ωk are used to drive the two cavities, respectively.
Fig. 2
Fig. 2 The squeezing spectra Sθ (ω) as a function of the normalized frequency ω/ωm with P2 = 1 mW and for the different powers of the left coupling field (a): P1 = 1 mW (black, solid curve); 5 mW (red, dashed curve); 10 mW (blue, dotted curve). The squeezing spectra Sθ (ω) as a function of the normalized frequency ω/ωm with P1 = 1 mW and for the different powers of the right coupling field (b): P2 = 1 mW (black, solid curve); 5 mW (red, dashed curve); 10 mW (blue, dotted curve). The values of the parameters are given by: ωm = 2π × 3.993 × 109Hz, g1 = 2π × 960 × 103, g2 = 2π × 430 × 103, T = 10K, r = 0, Qm = 87 × 103, κ1 = 2π × 520 × 106Hz, κ2 = 1.73 × 109Hz, ω1 = 2π × 205.3 × 1012Hz, ω2 = 2π × 194.1 × 1012Hz, θ = π/2.
Fig. 3
Fig. 3 The same settings are used as those in Fig. 2 except for the presence of squeezed vacuum injected in the left cavity with squeezing parameter r = 1.
Fig. 4
Fig. 4 The squeezing spectra Sθ(ω) as a function of the normalized frequency ω/ωm with P1 = 1 mW, P2 = 5 mW for the injected vacuum field r = 0 (black, solid curve), and for the injected squeezed field r = 1 (red, dashed curve). The others are set with the same values as in Fig. 2.
Fig. 5
Fig. 5 The squeezing spectra Sθ(ω) as a function of the normalized frequency ω/ωm with P1 = 1 mW, P2 = 5 mW and for different squeezing parameter r of the squeezed field injected in the left cavity: r = 0.3(black, solid curve); r = 0.6(red, dashed curve); r = 1(blue, dotted curve). The others are set with the same values as in Fig. 2.
Fig. 6
Fig. 6 The spectrum Sθ(ω) (a) as a function of the normalized frequency ω/ωm for the injected vacuum r = 0 (a) and the injected squeezed vacuum r = 1 (b) in the left-hand cavity. The temperature for the mechanical environment are given by different values: T1 = 10K (black, solid curve); 50 K (red, dashed curve); 100 K (blue, dotted curve). The others are set with the same values as in Fig. 2.
Fig. 7
Fig. 7 The spectrum Sθ(ω) (a) as a function of the normalized frequency ω/ωm for g1 = 0, r = 0 at Δ2 = ωm (a) and Δ2 = −ωm (b), and for g1 = 2π × 960 × 103, r = 1 at Δ2 = −ωm (c). The power of the right coupling field is set as: P2 = 1 mW (black, solid curve); 3 mW (red, dashed curve); 6 mW (blue, dotted curve). The others are set with the same values as in Fig. 2.

Equations (36)

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

H = k = 1 , 2 Δ c k a k a ^ k + ω m b ^ b ^ ( g 1 a ^ 1 a ^ 1 g 2 a ^ 2 a ^ 2 ) ( b ^ + b ^ ) + k = 1 , 2 i E k ( a ^ k a ^ k ) .
a ^ ˙ 1 = ( i Δ c 1 + κ 1 ) a ^ 1 + i g 1 a ^ 1 ( b ^ + b ^ ) + E 1 + 2 κ 1 a ^ in , 1 ,
a ^ ˙ 2 = ( i Δ c 2 + κ 2 ) a ^ 2 i g 2 a ^ 2 ( b ^ + b ^ ) + E 2 + 2 κ 2 a ^ in , 2 ,
b ^ ˙ = ( i ω m + γ m ) b ^ + i ( g 1 a ^ 1 a ^ 1 g 2 a ^ 2 a ^ 2 ) + b ^ in
a k s = E k κ k + i Δ k , ( k = 1 , 2 )
b s = i g 1 a 1 s a 1 s * i g 2 a 2 s a 2 s * γ m + i ω m
δ a ^ ˙ k = ( i Δ k + κ k ) δ a ^ k ( 1 ) k i g k a k s ( δ b ^ + δ b ^ ) + 2 κ k δ a ^ in , k , ( k = 1 , 2 )
δ b ^ ˙ = ( i ω m + γ m ) δ b ^ + i [ g 1 ( a 1 s * δ a ^ 1 + a 1 s δ a ^ 1 ) g 2 ( a 2 s * δ a ^ 2 + a 2 s δ a ^ 2 ) ] + δ b ^ in .
δ a ^ out , 2 ( ω ) = C 1 ( ω ) δ a ^ in , 1 + C 2 ( ω ) δ a ^ in , 1 ( ω ) + V 1 ( ω ) δ a ^ in , 2 + V 2 ( ω ) δ a ^ in , 2 ( ω ) + W 1 ( ω ) δ b ^ in ( ω ) + W 2 ( ω ) δ b ^ in ( ω )
C 1 ( ω ) = 2 i g 1 g 2 a 1 s * a 2 s κ 1 κ 2 Γ m m [ κ 1 + i ( Δ 1 ω ) ] [ κ 2 + i ( Δ 2 ω ) ] ,
C 2 ( ω ) = 2 i g 1 g 2 a 1 s a 2 s κ 1 κ 2 Γ m m [ κ 1 + i ( Δ 1 ω ) ] [ κ 2 + i ( Δ 2 ω ) ] ,
V 1 ( ω ) = 2 i g 2 2 | a 2 s | 2 κ 2 Γ m m [ κ 2 + i ( Δ 2 ω ) ] [ κ 2 + i ( Δ 2 ω ) ] + 2 κ 2 κ 2 + i ( Δ 2 ω ) 1 ,
V 2 ( ω ) = 2 i g 2 2 a 2 s 2 κ 2 Γ m m [ κ 2 + i ( Δ 2 ω ) ] [ κ 2 + i ( Δ 2 + ω ) ] ,
W 1 ( ω ) = i g 2 a 2 s 2 κ 2 Γ m b κ 2 + i ( Δ 2 ω ) ,
W 2 ( ω ) = i g 2 a 2 s 2 κ 2 Γ m a κ 2 + i ( Δ 2 ω ) ,
Γ m a = Γ + Γ + Γ 2 i ω m ( Λ m + i δ m ) ,
Γ m b = Γ Γ + Γ 2 i ω m ( Λ m + i δ m ) ,
Γ m m = 2 ω m Γ + Γ 2 i ω m ( Λ m + i δ m )
S θ ( ω ) = 1 2 π d Ω δ X ^ θ out ( ω ) δ X ^ θ out ( Ω ) ,
S θ ( ω ) = M F 1 + ( N + 1 ) F 2 + N F 3 + M * F 4 + G 1 + γ m ( n b + 1 ) H 1 + γ m n b H 2 ,
F 1 = C 1 ( ω ) C 1 ( Ω 1 ) e 2 i θ + C 1 ( ω ) C 2 * ( Ω 1 ) + C 2 * ( ω ) C 1 ( Ω 1 ) + C 2 * ( ω ) C 2 * ( Ω 1 ) e 2 i θ ,
F 2 = C 1 ( ω ) C 2 ( Ω 1 ) e 2 i θ + C 1 ( ω ) C 1 * ( Ω 2 ) + C 2 * ( ω ) C 2 ( Ω 2 ) + C 2 * ( ω ) C 1 * ( Ω 2 ) e 2 i θ ,
F 3 = C 2 ( ω ) C 1 ( Ω 2 ) e 2 i θ + C 2 ( ω ) C 2 * ( Ω 2 ) + C 1 * ( ω ) C 1 ( Ω 2 ) + C 1 * ( ω ) C 2 * ( Ω 2 ) e 2 i θ ,
F 4 = C 2 ( ω ) C 2 ( Ω 3 ) e 2 i θ + C 2 ( ω ) C 1 * ( Ω 3 ) + C 1 * ( ω ) C 2 ( Ω 3 ) + C 1 * ( ω ) C 1 * ( Ω 3 ) e 2 i θ ,
G 1 = V 1 ( ω ) V 2 ( Ω 2 ) e 2 i θ + V 1 ( ω ) V 1 * ( Ω 2 ) + V 2 * ( ω ) V 2 ( Ω 2 ) + V 2 * ( ω ) V 1 * ( Ω 2 ) e 2 i θ ,
H 1 = W 1 ( ω ) W 2 ( Ω 2 ) e 2 i θ + W 1 ( ω ) W 1 * ( Ω 2 ) + W 2 * ( ω ) W 2 ( Ω 2 ) + W 2 * ( ω ) W 1 * ( Ω 2 ) e 2 i θ ,
H 2 = W 2 ( ω ) W 1 ( Ω 2 ) e 2 i θ + W 2 ( ω ) W 2 * ( Ω 2 ) + W 1 * ( ω ) W 1 ( Ω 2 ) + W 1 * ( ω ) W 2 * ( Ω 2 ) e 2 i θ
B ^ 2 ( ω ) = 2 ( κ 2 i ω ) g 2 a 2 s ( κ 2 i ω ) 2 + Δ 2 2 Q ^ ( ω ) Δ 2 2 κ 2 ( κ 2 i ω ) 2 + Δ 2 2 A ^ in , 2 + ( κ 2 i ω ) 2 κ 2 ( κ 2 i ω ) 2 + Δ 2 2 B ^ in , 2 .
Q ^ ( ω ) = 1 D ( 2 κ 1 K 1 Ω 1 A ^ in , 1 + 2 κ 1 Δ 1 Ω 1 B ^ in , 1 2 κ 2 K 2 Ω 2 A ^ in , 2 2 κ 2 Δ 2 Ω 2 B ^ in , 2 + Γ m R 1 R 2 Q ^ in + ω m R 1 R 2 P ^ in )
B ^ 2 ( ω ) = a 11 A ^ in , 1 + b 11 B ^ in , 1 + a 22 A ^ in , 2 + b 22 B ^ in , 2 + q 11 Q ^ in + p 11 P ^ in
a 11 = 1 R 2 D × 2 K 1 K 2 G 2 Ω 1 2 κ 1 ,
b 11 = 1 R 2 D × 2 Δ 1 K 2 G 2 Ω 1 2 κ 1 ,
a 22 = 1 R 2 D × 2 κ 2 ( Δ 2 D 2 K 2 2 G 2 Ω 2 ) ,
b 22 = 1 R 2 D × 2 κ 2 K 2 ( D + 2 Δ 2 2 G 2 Ω 2 ) ,
q 11 = 1 R 2 D × 2 K 2 G 2 R 1 R 2 Γ m ,
p 11 = 1 R 2 D × 2 K 2 G 2 ω m R 1 R 2 .
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.