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

Quantum squeezing in a modulated optomechanical system

Open Access Open Access

Abstract

Quantum squeezing, as a typical quantum effect, is an important resource for many applications in quantum technologies. Here we propose a scheme for generating quantum squeezing, including the ponderomotive squeezing and the mechanical squeezing, in an optomechanical system, in which the radiation-pressure coupling and the mechanical spring constant are modulated periodically. In this system, the radiation-pressure interaction can be enhanced remarkably by the modulation-induced mechanical parametric amplification. Moreover, the effective phonon noise can be suppressed completely by introducing a squeezed vacuum reservoir. This ultimately leads to that our scheme can achieve a controllable quantum squeezing. Numerical calculations show that our scheme is experimentally realizable with current technologies.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Cavity optomechanics [1, 2], since it was put forward, has been a very hot research field. A prototypical cavity optomechanical system consists of a Fabry-Pérot cavity with one movable end mirror, which couples mechanical freedom with optical freedom through radiation pressure and provides a powerful research platform for a wide variety of theoretical and experimental investigations. For example, a considerable quantum effects have been found and confirmed including quantum squeezing [3–21], quantum ground-state cooling of the mechanical mirror [22–24], entanglement between light and the mirror [25], photon anti-bunching [26,27] and so on.

Quantum squeezing, as a typical quantum effect, is an important resource for many applications, such as gravitational wave detection [28, 29], quantum limited displacement sensing [30,31] and even biological measurements [32]. It is well known that optomechanical system, in its steady state, can mimic a Kerr medium [3] and the fluctuations in the system can give rise to optical quadrature squeezing termed the ponderomotive squeezing [3–5]. Many schemes for creating the ponderomotive squeezing in different physical systems have been proposed. For example, Hichem Eleuch et al. theoretically and experimentally investigated the ponderomotive squeezing in semiconductor microcavities [6–8]. Moreover, Marino et al. experimentally generated the ponderomotive squeezing by using classical intensity noise in optomechanical system [9], and Brooks et al. also experimentally demonstrated that the backaction of the motion of an ultracold atomic gas on the cavity light field produces the ponderomotive squeezing [10]. Besides ponderomotive squeezing, many schemes for creating quantum squeezing of the moving mirror in different physical systems have been proposed. For example, Jahne et al. theoretically demonstrated the strong squeezing of the mechanical oscillator via injecting a nonclassical light [11], and the experimental squeezing of a nanomechanical object has been achieved by Almog et al. through a nonlinear Duffing resonator [12]. Moreover, the mechanical resonator can also be squeezed by coupling it to an auxiliary nonlinear system, such as a superconducting quantum interference device loop [13], a Cooper-pair box circuit [14], or an optical cavity containing an atomic medium [15].

In this paper, we propose a scheme for generating quantum squeezing including ponderomotive squeezing and mechanical squeezing in an optomechanical system, where the radiation-pressure coupling and mechanical spring constant are modulated periodically [33–36]. In particular, the mechanical parametric amplification was theoretically and experimentally demonstrated by Rugar et al. and Szorkovszky et al. through the periodic modulation of the mechanical spring constant [33–35]. Moreover, together with the sinusoidal modulation of the radiation-pressure coupling, the resonant nonlinear photon-phonon interaction enhanced to the single-photon strong-coupling regime was also theoretically demonstrated through mechanical parametric amplification [36].

Besides applying the modulated optomechanical system, we assume that the mechanical resonator is coupled to a squeezed vacuum reservoir. It could be realized in principle by introducing an ancillary cavity mode, which has been deeply investigated in a great deal of studies [36–39], so we do not discuss it further in this paper. Our scheme is different from Ref. [16], in which the resonator is coupled to a thermal reservoir, so that the degree of squeezing will be affected by the thermal reservoir temperature even if the system can achieve a strong and robust squeezing. Moreover, our scheme is also different from Ref. [17], in which the cavity mode is coupled to a squeezed vacuum reservoir but the degree of squeezing is still affected by the effective phonon noise. Comparing with previous schemes for generating quantum squeezing, the main novelties of our scheme are the resonant radiation-pressure interaction can be enhanced remarkably by the modulation-induced mechanical parametric amplification, and the effective phonon noise can be suppressed completely by introducing a squeezed vacuum reservoir. Thus, under the conditions of enhanced radiation-pressure interaction and suppressed phonon noise, a controllable quantum squeezing can be achieved by modulating the amplitude of the mechanical parametric driving, which allows us to find the optimal quantum squeezing in the modulated optomechanical system.

This paper is organized as follows: In Sec. 2, we describe the model and solve the dynamical equation of the system. The ponderomotive squeezing and the mechanical squeezing are displayed in Sec. 3. Finally, we summarize our main results in Sec. 4.

2. Model and dynamical equation

As shown in Fig. 1, our system consists of an optomechanical cavity with modulated radiation-pressure coupling and mechanical spring constant, pumped by a laser with frequency ωp and amplitude ϵp=2κPp/ωp. Here Pp is the power of the pump field, and κ is the decay rate of the cavity field. A mechanical parametric amplification is induced due to the periodic modulation of the mechanical spring constant at 2ωd. Then, in the rotating frame with the frequency (ωd + ωp) and under the rotating wave approximation, the total Hamiltonian of the system can be written as (ħ = 1) [33–36]

Ht=Δccc+Δmbb+12ϵb(b2+b2)12g0cc(b+b)+iϵp(cc),
where c(b) and c(b) are the annihilation and creation operators of the cavity (mechanical) mode, respectively. Δc = ωcωpm = ωmωd) is the frequency detuning with the optical and mechanical resonant frequencies ωc and ωm. ϵb characterizes the amplitude of the mechanical parametric driving and g0 is the single-photon optomechanical coupling.

 figure: Fig. 1

Fig. 1 Sketch of the system. A cavity mode pumped by a laser at the red-detuned mechanical sideband, couples to a mechanical resonator with a modulating coupling strength g(t) = g0 cos (ωdt), in addition, the mechanical resonator is parametrically driven by modulating its spring constant, i.e., k(t) = k0kr cos(2ωdt).

Download Full Size | PDF

In particular, the quadratic part of Ht can be diagonalized with the squeezing transformation, S(rd)HtS(rd), where S(rd) = exp[rd(b2b†2)] is the squeezing operator and rd=14lnΔm+ϵbΔmϵb is the squeezing amplitude. By using the relation

S(rd)bS(rd)=bcosh(rd)+bsinh(rd),
we obtain
Heff=S(rd)HtS(rd)=Δccc+ωmbbgcc(b+b)+iϵp(cc),
where ωm=Δm/cosh(2rd) is the transformed mechanical frequency, hence the coherent cavity driving on the red-detuned mechanical sideband corresponds to Δc=ωm in our system. And g=12g0erd is the transformed optomechanical coupling, which can be exponentially enhanced with the increase of rd. Specifically, a large value of rd can be obtained by adjusting the system parameter ϵb to approaching ∆m, which is feasible with current technologies [40, 41]. The Heisenberg-Langevin equations can be derived as
c˙=(κ+iΔc)c+igc(b+b)+ϵp+2κcin,
b˙=(γm+iωm)b+igcc+2γmbin,
where κ and γm are the decay rates of the cavity and mechanical modes. cin and bin are the Langevin noise operators for the cavity and mechanical modes, respectively.

Here we assume that the mechanical resonator is coupled to a squeezed vacuum reservoir with the center squeezing parameter rb and reference phase Φe, and the cavity is coupled to a vacuum reservoir. Thus, both noise operators have zero mean, that is, 〈cin〉 = 〈bin〉 = 0, and their only nonzero correlation functions are

cin(t)cin(t)=δ(tt),
bin(t)bin(t)=(Neff+1)δ(tt),
bin(t)bin(t)=Neffδ(tt),
bin(t)bin(t)=Meffδ(tt),
bin(t)bin(t)=Meffδ(tt),
where Neff and Meff represent the effective thermal noise and the two-phonon correlation interaction [42,43] with the expressions of
Neff=sinh2(re)cosh2(rd)+sinh2(rd)cosh2(re)+12cos(Φe)sinh(2re)sinh(2rd),
Meff=[cosh(re)sinh(rd)+eiΦesinh(re)cosh(rd)]×[cosh(re)cosh(rd)+eiΦesinh(re)sinh(rd)].

As shown in Fig. 2, the effective thermal noise Neff is plotted as a function of the squeezing parameters re and reference phase Φe for different values of the amplitude ϵb (note that the amplitude of Meff has the same behavior as Neff, and is not plotted here). From the figures, we can easily observe that the effective thermal noise Neff will increase exponentially (periodically) when the squeezing parameter re (phase Φe) deviates from the ideal parameters (re = rd and Φe = ± (n = 1, 3, 5, …)). That is, the effective thermal noise and the two-photon correlation interaction can be suppressed completely (i.e., Neff = Meff = 0) under the ideal parameter conditions, which shows that the squeezed vacuum reservoir, coupled to mechanical resonator, corresponds to an effective vacuum reservoir. Based on our scheme, it is necessary to suppress the effective thermal noise and two-phonon correlation interaction to observe the ponderomotive squeezing and mechanical squeezing as shown in the following sections.

 figure: Fig. 2

Fig. 2 The effective thermal noise Neff is plotted as a function of (a) the squeezing parameters re, (b) the reference phase Φe for different values of the amplitude ϵb, where ϵb = 3999.50κ (solid black curve), ϵb = 3999.65κ (dashed blue curve), ϵb = 3999.80κ (dotted red curve), ϵb = 3999.95κ (dot-dashed green curve). Other parameters are ∆m = 4000κ and κ = 0.1MHz.

Download Full Size | PDF

By using o = os + δo (o = c, b), Eqs. (4)–(5) can be divided into the steady parts and the fluctuation ones. Substituting the division forms into Eqs. (4)–(5), we can obtain the following steady-state solutions of the system operators

cs=ϵpiΔ+κ,
bs=g|cs|2ωmiγm,
where Δ=Δcg(bs+bs*) is the effective detuning.

As shown in Fig. 3, in our system, the strong red-detuned driving on the cavity generates large steady-state amplitudes in both the cavity and mechanical modes. For example, with a pump field amplitude ϵp ≈ 2000κ, |cs|2 ≈ 10 can ensure the validity of our following assumptions for linearization. Thus, the Heisenberg-Langevin equations for the fluctuations δo (o = c, b) can be given by

δc˙=(κ+iΔ)δc+igcs(δb+δb)+2κcin,
δb˙=(γm+iωm)δb+ig(cs*δc+csδc)2γmbin.

 figure: Fig. 3

Fig. 3 The steady-state amplitudes |cs| and |bs| vs pump amplitude ϵp. The parameters are κ = 0.1MHz, γm = 0.01κ, g0 = 0.005κ, ∆m = 4000κ, ϵb = 3999.95κ, Δc=ωm.

Download Full Size | PDF

We rewrite the above equations in the compact form as

μ˙(t)=Aμ(t)+ν(t),
with the column vector of fluctuations in the system being µT = δc, δc, δb, δb) and the column vector of noise being νT=(2κcin,2κcin,2γmbin,2γmbin). The matrix A is given by
A=((κ+iΔ)0igcsigcs0(κiΔ)igcs*igcs*igcs*igcs(γm+iωm)0igcs*igcs0(γmiωm)).

The system is stable and reaches its steady state when all of the eigenvalues of A have negative real parts. The stability conditions can be derived by applying the Routh-Hurwitz criterion [44], yielding the following two nontrivial conditions on the system parameters,

4γmκ{[(γm+κ)2+Δ2]2+2[(γm+κ)2Δ2]ωm2+ωm4}+16(γm+κ)2Δωmg2|cs|2>0,
(Δ2+κ2)(γm2+ωm2)4Δωmg2|cs|2>0,
which will be considered to be satisfied from now on.

3. Quantum squeezing

3.1. The ponderomotive squeezing

In this section, we investigate the squeezing properties of the transmitted field, which is accessible to experiment and useful for practical applications. It is well known that fluctuations of the electric field are more convenient to measure in the frequency domain than in the time domain experimentally. Therefore, by using the definition of the Fourier transform for an operator o = (δc, δb, cin, bin),

o(ω)=12πo(t)eiωtdt,o(ω)=[o(ω)],
the correlations of noise operators in the frequency domain can be defined as
cin(ω)cin(ω)=δ(ω+ω),
bin(ω)bin(ω)=(Neff+1)δ(ω+ω),
bin(ω)bin(ω)=Neffδ(ω+ω)
bin(ω)bin(ω)=Meff*δ(ω+ω),
bin(ω)bin(ω)=Meffδ(ω+ω).

After solving the matrix Eq. (17) in the frequency domain, we obtain

δc(ω)=ζ1(ω)cin+ζ2(ω)cin+ζ3(ω)bin+ζ4(ω)bin,
with
ζ1(ω)=2κ[Θ1(g2|cs|2+Θ2Λ2)Θ2g2|cs|2]d(ω)
ζ2(ω)=2κ(Θ1Θ2)g2cs2d(ω),
ζ3(ω)=i2γmΘ2Λ2gcsd(ω)
ζ4(ω)=i2γmΘ1Λ2gcsd(ω)
d(ω)=Θ1[g2|cs|2(Λ1Λ2)+Θ2Λ1Λ2]+Θ2g2|cs|2(Λ2Λ1),
where Θ1,2 = γmi (ωωm) and Λ1,2 = κi (ω ∓ ∆).

By using the results above and the input-output relation [45], δcout=2κδccin, we can obtain the intensity spectrum of the transmitted field given by

Sout(ω)=δcout(ω)δcout(ω),
and the stationary squeezing spectrum of the transmitted field given by [3,4,18,19]
Sθ(ω)=δXθout(ω)δXθout(ω),
where δXθout(ω)=eiθδcout(ω)+eiθδcout(ω) with θ being the measurement phase angle. Since [δXθout,δXθ+π/2out]=2i, the quadrature squeezing occurs when δXθout(ω)δXθout(ω)<1, i.e., Sθ (ω) < 1. Then, the squeezing spectrum can be put in the form
Sθ(ω)=1+2Bcc(ω)+[e2iθBcc(ω)+c.c],
where
δcout(ω)δcout(ω)=δ(ω+ω)Bcc(ω),
δcout(ω)δcout(ω)=δ(ω+ω)Bcc(ω).

As the degree of squeezing depends on the direction of the measurement of the quadratures, we choose the optimal phase angle θopt by solving dSθ(ω)/dθ = 0, then we obtain

e2iθopt=±Bcc(ω)|Bcc(ω)|.

We need to choose the solution with a negative sign as it minimizes the spectrum function, therefore, we have

Sopt(ω)=1+2Bcc(ω)2|Bcc(ω)|,
where
Bcc(ω)=2κ[|ζ2(ω)|2+|ζ3(ω)|2Neff+ζ3*(ω)ζ4(ω)Meff+ζ4*(ω)ζ3(ω)Meff*+|ζ4(ω)|2(Neff+1)],
Bcc(ω)=2κ[ζ1(ω)ζ2(ω)+ζ3(ω)ζ3(ω)Meff*+ζ3(ω)ζ4(ω)(Neff+1)+ζ4(ω)ζ3(ω)Neff+ζ4(ω)ζ4(ω)Meffζ2(ω)2κ].

Note that under the ideal parameter conditions re = rd and Φe = ± (n = 1, 3, 5, …, Neff = Meff = 0. Thus, Bcc(ω) and Bcc(ω) simplify to

Bcc(ω)=2κ[|ζ2(ω)|2+|ζ4(ω)|2],
Bcc(ω)=2κ[ζ1(ω)ζ2(ω)+ζ3(ω)ζ4(ω)ζ2(ω)2κ],
respectively.

The parameters used in this work are experimentally realizable with current technologies. The single-photon optomechanical coupling strength we choose is g0 = 0.005κ, however, for the required optomechanical coupling strength (i.e., g=12g0erd), our scheme is applicable in principle to use a smaller g0 by employing a larger squeezing parameter rd in the critical parameter regime ϵb → ∆m [40,41]. Moreover, our scheme is general and applicable to the optomechanical system in the optical wave range [46] or electromechanical system in the microwave range [47]. To implement our scheme, we can choose the ultrahigh-Q toroid microcavity [48] as an example and the system parameters are analogous to Ref. [36]: the cavity damping rate κ = 0.1MHz, the mechanical damping rate γm = 0.01κ, the single-photon optomechanical coupling g0 = 0.005κ, the frequency detuning ∆m = 4000κ and the pump field amplitude ϵp = 2150κ. It should be pointed out that ωmκ, i.e., the system operates in the resolved sideband regime.

At first, we analyze the intensity spectrum of the transmitted field. As shown in Eq. (3), due to the periodic modulation of the radiation-pressure coupling and mechanical spring constant, the mechanical frequency ωm and the single-photon optomechanical coupling g0 are modified to the amplitude ϵb-dependent frequency ωm and coupling g′, respectively. These modifications affect the spectrum of the transmitted field reflected from the oscillating mirror, that is, some changes would occur in the amplitude and position of the spectral peaks. As shown in Fig. 4, we plot the intensity spectrum of the transmitted field Sout(ω) as a function of the normalized frequency (ωωc)/ωm for different values of the amplitude ϵb. From the curves, with the increase of the amplitude ϵb, we can easily observe that the amplitude of the intensity spectrum increases remarkably. Moreover, the splitting phenomenon occurs in the intensity spectrum, which also becomes more and more obvious with the increase of amplitude ϵb. Specifically, the splitting phenomenon can be analyzed by d(ω) in Eq. (32) quantitatively, whose real and imaginary parts give the position and width of the spectral peaks, respectively. Thus, the explicit expressions for the frequencies at which spectral peaks occur can be derived as

ω1=±z1z12+z22,
ω2=±z1+z12+z22,
where
z1=γ2+ωm2+κ2+Δ2+4γκ,
z2=16Δωm+g2|cs|24(Δ2+κ2)(γ2+ωm2).

 figure: Fig. 4

Fig. 4 The intensity spectrum of the transmitted field is plotted as a function of normalized frequency (ωωc)/ωm for different values of the amplitude ϵb. Other parameters are the same as in Fig. 3.

Download Full Size | PDF

From above equations, we can easily find that the shift in the spectral peaks strongly depends on the amplitude ϵb, which is clearly manifested for higher values of ϵb as shown in Fig. 4. As the enhanced and splitting spectral features of the transmitted field are observed, it is evident that one can also expect the same to follow for the squeezing spectrum.

Now we focus on how the amplitude ϵb affects the squeezing degree of the transmitted field. As shown in Fig. 5, we plot the squeezing spectrum Sopt (ω) as a function of the normalized frequency (ωωc)/ωm for different values of the amplitude ϵb. As can be seen from figures, the transmitted field exhibits squeezing with the degree of squeezing strongly relying on the amplitude ϵb. Moreover, a considerable amount of squeezing at frequencies (ω1, ω2) is observed as expected.

 figure: Fig. 5

Fig. 5 The squeezing spectrum of the transmitted field is plotted as a function of normalized frequency (ωωc)/ωm for different values of the amplitude ϵb. Other parameters are the same as in Fig. 3.

Download Full Size | PDF

The spectral features of the squeezing spectrum Sopt (ω) can be analyzed as follows: As shown in Eq. (3), the effective optomechanical coupling can be exponentially enhanced in the modulated optomechanical system, hence our system can be seen as an enhanced Kerr medium [3–5]. Further, as shown in Fig. 2, in our system, the radiation pressure force noise can completely dominate the effects of thermal noise under the ideal parameters. Thus, under the conditions of enhanced radiation-pressure interaction and suppressed phonon noise, the amount of squeezing of the transmitted field is ultimately enhanced remarkably.

In order to observe the features of quantitative change when deviating from the ideal parameters, the squeezing spectrum Sopt (ω1) is shown as a function of the squeezing parameter re and reference phase Φe as shown in Fig. 6 (note that Sopt (ω2) has the same behavior as Sopt(ω1) in our numerical calculations, and is not plotted here). From the figures, it is clearly shown that the optimal squeezing occurs in the vicinity of the ideal parameters, in which the thermal noise for the mechanical mode is completely suppressed. Thus, the above results show that a squeezed vacuum environment with appropriate squeezing parameter re and reference phase Φe is required to observe the ponderomotive squeezing, which provides us an effective method for controlling the degree of squeezing and finding the optimal squeezing.

 figure: Fig. 6

Fig. 6 The squeezing spectrum is plotted as a function of (a) the squeezing parameter re, (b) the reference phase Φe. Other parameters are the same as in Fig. 3.

Download Full Size | PDF

3.2. The mechanical squeezing

Next, we analyze the squeezing properties of the mechanical resonator in the same picture, Heff = S(rd) HtS(rd), by evaluating the variances of the transformed quadrature operators, X+ = S(rd) (δb + δb) S(rd) and X = S(rd)[i(δbδb)] S(rd), given by

ΔX±2=[1+2δbδb±(δb2+δb2)(δb±δb)2]e±2rd.

Thus, the mechanical squeezing takes place when the variances satisfy the condition that either ΔX+2<1 or ΔX2<1.

From Eqs. (15)–(16), we can obtain the linearized Hamiltonian, and under the rotating wave approximation, it can be written as

Heff,L=Δδcδc+ωmδbδbgcsδcδbgcs*δcδb.

In the new rotating frame, the Heisenberg-Langevin equations for the fluctuations can be given by

δc˙=(κ+iΔ)δc+igcsδb+2κcin,
δb˙=(γm+iωm)δb+igcs*δc+2γmbin.

We can solve the above equations and obtain the analytical result for the fluctuation δb(t) using the Laplace transform [17,20,21],

δb(t)=f1(t)δc(0)+2κ0tf1(tt)cin(t)dt+f2(t)δb(0)+2γm0tf2(tt)bin(t)dt,
where
f1(t)=igcs*[e(uv)t/2e(u+v)t/2]u,
f2(t)=[u+v2(γm+iωm)]e(uν)t/2+[u+v2(κ+iΔ)]e(u+v)t/22u,
u=[γmκi(Δωm)]24g2|cs|2,
v=γm+κ+i(Δ+ωm).

Moreover, we assume that the cavity and the mechanical modes are initially in the vacuum state, that is,

δc(0)δc(0)=0,δb(0)δb(0)=0.

Then, using Eq. (52) and taking into account the correlations of noise operators in the time domain, we can obtain the steady-state variances after some tedious calculations [17, 20, 21],

ΔX+2ss=[1+4γmNeffη1±2γmMeff*η2±2γmMeffη2*]e±2rd,
where
η1=κ[(γm+κ)2+(Δωm)2]+(γm+κ)g2|cs|22γmκ[(γm+κ)2+(Δωm)2]+2(γm+κ)2g2|cs|2,
η2=(κ+iΔ)[γm+κ+i(Δ+ωm)]+g2|cs|22[(κ+iΔ)(γm+ωm)+g2|cs|2][γm+κ+i(Δ+ωm)].

From the above expressions, it is apparent that it can only be present in ΔX2ss if there is any squeezing in mechanical resonator. Further, the steady-state variance ΔX2ss crucially depends on the squeezed vacuum reservoir. We confirm this behavior by plotting the steady-state variance ΔX2ss as a function of the squeezing parameter re and reference phase Φe as shown in Fig. 7. From the figures, we can easily observe that a considerable amount of squeezing occurs in the mechanical resonator even if the squeezing parameter re is very small, i.e., rerd. Moreover, similar to the ponderomotive squeezing discussed above, higher squeezing in the reservoir doesn’t lead to better squeezing of the mechanical resonator and it would lead to the vanishment of squeezing instead, but we can achieve the optimal squeezing in the vicinity of the ideal parameters.

 figure: Fig. 7

Fig. 7 The steady-state variance ΔX2ss is plotted as a function of (a) the squeezing parameter re, (b) the reference phase Φe. Other parameters are the same as in Fig. 3.

Download Full Size | PDF

4. Conclusion

In conclusion, we have theoretically investigated the squeezing properties of the cavity and the mechanical modes in an optomechanical system, where the radiation-pressure coupling and the mechanical spring constant are modulated periodically. We have shown that the resonant radiation-pressure interaction can be enhanced remarkably by the modulation-induced mechanical parametric amplification. Moreover, the effective phonon noise can be suppressed completely by introducing a squeezed vacuum reservoir. Specifically, a controllable quantum squeezing is demonstrated by modulating the amplitude of the mechanical parametric driving, then the optimal quantum squeezing can be found in the modulated optomechanical system. This work presents an alternative approach to generate quantum squeezing with current experimental technologies, which can play an important role in the ultrahigh precision measurement and other related fields.

Funding

National Natural Science Foundation of China (Grant Nos. 11574092, 61775062, 61378012, 91121023, and 60978009); National Basic Research Program of China (Grant No. 2013CB921804); the Innovation Project of Graduate School of South China Normal University (2017LKXM090).

References and links

1. V. B. Braginskii and A. B. Manukin, “Ponderomotive effects of electromagnetic radiation,” Sov. Phys. JETP 25(4), 653–655 (1967).

2. V. B. Braginskii, A. B. Manukin, and M. Y. Tikhonov, “Investigation of dissipative ponderomotive effects of electromagnetic radiation,” Sov. Phys. JETP 31, 829 (1970).

3. 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), 1337–1343 (1994). [CrossRef]   [PubMed]  

4. S. Mancini and P. Tombesi, “Quantum noise reduction by radiation pressure,” Phys. Rev. A 49(5), 4055–4065 (1994). [CrossRef]   [PubMed]  

5. F. Marino, F. S. Cataliotti, A. Farsi, M. S. de Cumis, and F. Marin, “Classical signature of ponderomotive squeezing in a suspended mirror resonator,” Phys. Rev. Lett. 104(7), 073601 (2010). [CrossRef]   [PubMed]  

6. H. Eleuch, J. M. Courty, G. Messin, C. Fabre, and E. Giacobino, “Cavity QED effects in semiconductor microcavities,” J. Opt. B: Quantum Semiclass. Opt. 1(1), 1–7 (1999). [CrossRef]  

7. G. Messin, J. P. Karr, H. Eleuch, J. M. Courty, and E. Giacobino, “Squeezed states and the quantum noise of light in semiconductor microcavities,” J. Phys.: Condens. Matter 11(31), 6069–6078 (1999).

8. E. Giacobino, J. P. Karr, G. Messin, H. Eleuch, and A. Baas, “Quantum optical effects in semiconductor microcavities,” C. R. Physique 3(1), 41–52 (2002). [CrossRef]  

9. F. Marino, F. S. Cataliotti, A. Farsi, M. S. de Cumis, and F. Marin, “Classical signature of ponderomotive squeezing in a suspended mirror resonator,” Phys. Rev. Lett. 104(7), 073601 (2010). [CrossRef]   [PubMed]  

10. D. W. 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 488(7412), 476–480 (2012). [CrossRef]   [PubMed]  

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

12. R. Almog, S. Zaitsev, O. Shtempluck, and E. Buks, “Noise squeezing in a nanomechanical Duffing resonator,” Phys. Rev. Lett. 98(7), 078103 (2007). [CrossRef]   [PubMed]  

13. J. Zhang, Y. X. Liu, and F. Nori, “Cooling and squeezing the fluctuations of a nanomechanical beam by indirect quantum feedback control,” Phys. Rev. A 79(5), 052102 (2009). [CrossRef]  

14. P. Rabl, A. Shnirman, and P. Zoller, “Generation of squeezed states of nanomechanical resonators by reservoir engineering,” Phys. Rev. B 70(20), 205304 (2004). [CrossRef]  

15. 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(1), 013824 (2008). [CrossRef]  

16. E. A. Sete and H. Eleuch, “Strong squeezing and robust entanglement in cavity electromechanics,” Phys. Rev. A 89(1), 013841 (2014). [CrossRef]  

17. E. A. Sete, S. Das, and H. Eleuch, “External-field effect on quantum features of radiation emitted by a quantum well in a microcavity,” Phys. Rev. A 83(2), 023822 (2011). [CrossRef]  

18. M. J. Collett and D. F. Walls, “Squeezing spectra for nonlinear optical systems,” Phys. Rev. A 32(5), 2887–2892 (1985). [CrossRef]  

19. 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(15), 17979–17989 (2014). [CrossRef]   [PubMed]  

20. E. A. Sete and H. Eleuch, “Interaction of a quantum well with squeezed light: Quantum-statistical properties,” Phys. Rev. A 82(4), 043810 (2010). [CrossRef]  

21. Q. H. Cai, Y. Xiao, Y. F. Yu, and Z. M. Zhang, “Mechanical squeezing and photonic anti-bunching in a coupled two-cavity optomechanical system,” Opt. express 24(18), 20036–20048 (2016). [CrossRef]   [PubMed]  

22. O. Arcizet, P. F. Cohadon, T. Briant, M. Pinard, and A. Heidmann, “Radiation-pressure cooling and optomechanical instability of a micromirror,” Nature 444(7115), 71–74 (2006). [CrossRef]   [PubMed]  

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

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

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

26. R. Loudon, “Non-classical effects in the statistical properties of light,” Rep. Progr. Phys. 43(7), 913 (1980). [CrossRef]  

27. H. Paul, “Photon anti-bunching,” Rev. Mod. Phys. 54(4), 1061–1102 (1982). [CrossRef]  

28. T. Eberle, S. Steinlechner, J. Bauchrowitz, V. Handchen, H. Vahlbruch, M. Mehmet, H. Muller-Ebhardt, and R. Schnabel, “Quantum enhancement of the zero-area Sagnac interferometer topology for gravitational wave detection,” Phys. Rev. Lett. 104(25), 251102 (2010). [CrossRef]   [PubMed]  

29. The LIGO Scientific Collaboration, “A gravitational wave observatory operating beyond the quantum shot-noise limit,” Nat. Phys . 7(12), 962 (2011). [CrossRef]  

30. F. Y. Khalili and V. Braginski, Quantum Measurement(Cambridge University, Cambridge, 1995).

31. U. B. Hoff, G. I. Harris, L. S. Madsen, H. Kerdoncuff, M. Lassen, B. M. Nielsen, W. P. Bowen, and U. L. Andersen, “Quantum-enhanced micromechanical displacement sensitivity,” Opt. Lett. 38(9), 1413–1415 (2013). [CrossRef]   [PubMed]  

32. M. A. Taylor, J. Janousek, V. Daria, J. Knittel, B. Hage, H. A. Bachor, and W. P. Bowen, “Biological measurement beyond the quantum limit,” Nat. Photonics 7(3), 229 (2013). [CrossRef]  

33. D. Rugar and P. Grutter, “Mechanical parametric amplification and thermomechanical noise squeezing,” Phys. Rev. Lett. 67(6), 699–702 (1991). [CrossRef]   [PubMed]  

34. A. Szorkovszky, A. C. Doherty, G. I. Harris, and W. P. Bowen, “Mechanical squeezing via parametric amplification and weak measurement,” Phys. Rev. Lett. 107(21), 213603 (2011). [CrossRef]   [PubMed]  

35. A. Szorkovszky, G. A. Brawley, A. C. Doherty, and W. P. Bowen, “Strong thermomechanical squeezing via weak measurement,” Phys. Rev. Lett. 110(18), 184301 (2013). [CrossRef]   [PubMed]  

36. T. S. Yin, X. Y. Lü, L. L. Zheng, M. Wang, S. Li, and Y. Wu, “Nonlinear effects in modulated quantum optomechanics,” Phys. Rev. A 95(5), 053861 (2017). [CrossRef]  

37. 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]  

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

39. A. Nunnenkamp, V. Sudhir, A. K. Feofanov, A. Roulet, and T. J. Kippenberg, “Quantum-limited amplification and parametric instability in the reversed dissipation regime of cavity optomechanics,” Phys. Rev. Lett. 113(2), 023604 (2014). [CrossRef]   [PubMed]  

40. Y. Y. Jiang, A. D. Ludlow, N. D. Lemke, R. W. Fox, J. A. Sherman, L. S. Ma, and C. W. Oates, “Making optical atomic clocks more stable with 10-16-level laser stabilization,” Nat. Photonics 5(3), 158 (2011). [CrossRef]  

41. T. M. Fortier, M. S. Kirchner, F. Quinlan, J. Taylor, J. C. Bergquist, T. Rosenband, N. Lemke, A. Ludlow, Y. Jiang, C. W. Oates, and S. A. Diddams, “Generation of ultrastable microwaves via optical frequency division,” Nat. Photonics 5(7), 425 (2011). [CrossRef]  

42. H. P. Breuer and F. Petruccione, The Theory of Open Quantum Systems(Clarendon, Oxford, 2002), Chap. 3.

43. X. Y. Lü, Y. Wu, J. R. Johansson, H. Jing, J. Zhang, and F. Nori, “Squeezed Optomechanics with Phase-Matched Amplification and Dissipation,” Phys. Rev. Lett. 114(9), 093602 (2015). [CrossRef]   [PubMed]  

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

45. C. W. Gardiner and P. Zoller, Quantum noise: a handbook of Markovian and non-Markovian quantum stochastic methods with applications to quantum optics (Springer-Verlag, 2000). [CrossRef]  

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 482(7383), 63–67 (2012). [CrossRef]   [PubMed]  

47. J. D. Teufel, D. Li, M. S. Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, and R. W. Simmonds, “Circuit cavity electromechanics in the strong-coupling regime,” Nature 471(7337), 204–208 (2011). [CrossRef]   [PubMed]  

48. D. K. Armani, T. J. Kippenberg, S. M. Spillane, and K. J. Vahala, “Ultra-high-Q toroid microcavity on a chip,” Nature 421(6926), 925–928 (2003). [CrossRef]   [PubMed]  

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 Sketch of the system. A cavity mode pumped by a laser at the red-detuned mechanical sideband, couples to a mechanical resonator with a modulating coupling strength g(t) = g0 cos (ωdt), in addition, the mechanical resonator is parametrically driven by modulating its spring constant, i.e., k(t) = k0kr cos(2ωdt).
Fig. 2
Fig. 2 The effective thermal noise Neff is plotted as a function of (a) the squeezing parameters re, (b) the reference phase Φ e for different values of the amplitude ϵb, where ϵb = 3999.50κ (solid black curve), ϵb = 3999.65κ (dashed blue curve), ϵb = 3999.80κ (dotted red curve), ϵb = 3999.95κ (dot-dashed green curve). Other parameters are ∆ m = 4000κ and κ = 0.1MHz.
Fig. 3
Fig. 3 The steady-state amplitudes |cs| and |bs| vs pump amplitude ϵp. The parameters are κ = 0.1MHz, γm = 0.01κ, g0 = 0.005κ, ∆ m = 4000κ, ϵb = 3999.95κ, Δ c = ω m .
Fig. 4
Fig. 4 The intensity spectrum of the transmitted field is plotted as a function of normalized frequency ( ω ω c ) / ω m for different values of the amplitude ϵb. Other parameters are the same as in Fig. 3.
Fig. 5
Fig. 5 The squeezing spectrum of the transmitted field is plotted as a function of normalized frequency ( ω ω c ) / ω m for different values of the amplitude ϵb. Other parameters are the same as in Fig. 3.
Fig. 6
Fig. 6 The squeezing spectrum is plotted as a function of (a) the squeezing parameter re, (b) the reference phase Φ e . Other parameters are the same as in Fig. 3.
Fig. 7
Fig. 7 The steady-state variance Δ X 2 s s is plotted as a function of (a) the squeezing parameter re, (b) the reference phase Φ e . Other parameters are the same as in Fig. 3.

Equations (60)

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

H t = Δ c c c + Δ m b b + 1 2 ϵ b ( b 2 + b 2 ) 1 2 g 0 c c ( b + b ) + i ϵ p ( c c ) ,
S ( r d ) b S ( r d ) = b cosh ( r d ) + b sinh ( r d ) ,
H e f f = S ( r d ) H t S ( r d ) = Δ c c c + ω m b b g c c ( b + b ) + i ϵ p ( c c ) ,
c ˙ = ( κ + i Δ c ) c + i g c ( b + b ) + ϵ p + 2 κ c i n ,
b ˙ = ( γ m + i ω m ) b + i g c c + 2 γ m b i n ,
c i n ( t ) c i n ( t ) = δ ( t t ) ,
b i n ( t ) b i n ( t ) = ( N e f f + 1 ) δ ( t t ) ,
b i n ( t ) b i n ( t ) = N e f f δ ( t t ) ,
b i n ( t ) b i n ( t ) = M e f f δ ( t t ) ,
b i n ( t ) b i n ( t ) = M e f f δ ( t t ) ,
N e f f = sinh 2 ( r e ) cosh 2 ( r d ) + sinh 2 ( r d ) cosh 2 ( r e ) + 1 2 cos ( Φ e ) sinh ( 2 r e ) sinh ( 2 r d ) ,
M e f f = [ cosh ( r e ) sinh ( r d ) + e i Φ e sinh ( r e ) cosh ( r d ) ] × [ cosh ( r e ) cosh ( r d ) + e i Φ e sinh ( r e ) sinh ( r d ) ] .
c s = ϵ p i Δ + κ ,
b s = g | c s | 2 ω m i γ m ,
δ c ˙ = ( κ + i Δ ) δ c + i g c s ( δ b + δ b ) + 2 κ c i n ,
δ b ˙ = ( γ m + i ω m ) δ b + i g ( c s * δ c + c s δ c ) 2 γ m b i n .
μ ˙ ( t ) = A μ ( t ) + ν ( t ) ,
A = ( ( κ + i Δ ) 0 i g c s i g c s 0 ( κ i Δ ) i g c s * i g c s * i g c s * i g c s ( γ m + i ω m ) 0 i g c s * i g c s 0 ( γ m i ω m ) ) .
4 γ m κ { [ ( γ m + κ ) 2 + Δ 2 ] 2 + 2 [ ( γ m + κ ) 2 Δ 2 ] ω m 2 + ω m 4 } + 16 ( γ m + κ ) 2 Δ ω m g 2 | c s | 2 > 0 ,
( Δ 2 + κ 2 ) ( γ m 2 + ω m 2 ) 4 Δ ω m g 2 | c s | 2 > 0 ,
o ( ω ) = 1 2 π o ( t ) e i ω t d t , o ( ω ) = [ o ( ω ) ] ,
c i n ( ω ) c i n ( ω ) = δ ( ω + ω ) ,
b i n ( ω ) b i n ( ω ) = ( N e f f + 1 ) δ ( ω + ω ) ,
b i n ( ω ) b i n ( ω ) = N e f f δ ( ω + ω )
b i n ( ω ) b i n ( ω ) = M e f f * δ ( ω + ω ) ,
b i n ( ω ) b i n ( ω ) = M e f f δ ( ω + ω ) .
δ c ( ω ) = ζ 1 ( ω ) c i n + ζ 2 ( ω ) c i n + ζ 3 ( ω ) b i n + ζ 4 ( ω ) b i n ,
ζ 1 ( ω ) = 2 κ [ Θ 1 ( g 2 | c s | 2 + Θ 2 Λ 2 ) Θ 2 g 2 | c s | 2 ] d ( ω )
ζ 2 ( ω ) = 2 κ ( Θ 1 Θ 2 ) g 2 c s 2 d ( ω ) ,
ζ 3 ( ω ) = i 2 γ m Θ 2 Λ 2 g c s d ( ω )
ζ 4 ( ω ) = i 2 γ m Θ 1 Λ 2 g c s d ( ω )
d ( ω ) = Θ 1 [ g 2 | c s | 2 ( Λ 1 Λ 2 ) + Θ 2 Λ 1 Λ 2 ] + Θ 2 g 2 | c s | 2 ( Λ 2 Λ 1 ) ,
S o u t ( ω ) = δ c o u t ( ω ) δ c o u t ( ω ) ,
S θ ( ω ) = δ X θ o u t ( ω ) δ X θ o u t ( ω ) ,
S θ ( ω ) = 1 + 2 B c c ( ω ) + [ e 2 i θ B c c ( ω ) + c . c ] ,
δ c o u t ( ω ) δ c o u t ( ω ) = δ ( ω + ω ) B c c ( ω ) ,
δ c o u t ( ω ) δ c o u t ( ω ) = δ ( ω + ω ) B c c ( ω ) .
e 2 i θ o p t = ± B c c ( ω ) | B c c ( ω ) | .
S o p t ( ω ) = 1 + 2 B c c ( ω ) 2 | B c c ( ω ) | ,
B c c ( ω ) = 2 κ [ | ζ 2 ( ω ) | 2 + | ζ 3 ( ω ) | 2 N e f f + ζ 3 * ( ω ) ζ 4 ( ω ) M e f f + ζ 4 * ( ω ) ζ 3 ( ω ) M e f f * + | ζ 4 ( ω ) | 2 ( N e f f + 1 ) ] ,
B c c ( ω ) = 2 κ [ ζ 1 ( ω ) ζ 2 ( ω ) + ζ 3 ( ω ) ζ 3 ( ω ) M e f f * + ζ 3 ( ω ) ζ 4 ( ω ) ( N e f f + 1 ) + ζ 4 ( ω ) ζ 3 ( ω ) N e f f + ζ 4 ( ω ) ζ 4 ( ω ) M e f f ζ 2 ( ω ) 2 κ ] .
B c c ( ω ) = 2 κ [ | ζ 2 ( ω ) | 2 + | ζ 4 ( ω ) | 2 ] ,
B c c ( ω ) = 2 κ [ ζ 1 ( ω ) ζ 2 ( ω ) + ζ 3 ( ω ) ζ 4 ( ω ) ζ 2 ( ω ) 2 κ ] ,
ω 1 = ± z 1 z 1 2 + z 2 2 ,
ω 2 = ± z 1 + z 1 2 + z 2 2 ,
z 1 = γ 2 + ω m 2 + κ 2 + Δ 2 + 4 γ κ ,
z 2 = 16 Δ ω m + g 2 | c s | 2 4 ( Δ 2 + κ 2 ) ( γ 2 + ω m 2 ) .
Δ X ± 2 = [ 1 + 2 δ b δ b ± ( δ b 2 + δ b 2 ) ( δ b ± δ b ) 2 ] e ± 2 r d .
H e f f , L = Δ δ c δ c + ω m δ b δ b g c s δ c δ b g c s * δ c δ b .
δ c ˙ = ( κ + i Δ ) δ c + i g c s δ b + 2 κ c i n ,
δ b ˙ = ( γ m + i ω m ) δ b + i g c s * δ c + 2 γ m b i n .
δ b ( t ) = f 1 ( t ) δ c ( 0 ) + 2 κ 0 t f 1 ( t t ) c i n ( t ) d t + f 2 ( t ) δ b ( 0 ) + 2 γ m 0 t f 2 ( t t ) b i n ( t ) d t ,
f 1 ( t ) = i g c s * [ e ( u v ) t / 2 e ( u + v ) t / 2 ] u ,
f 2 ( t ) = [ u + v 2 ( γ m + i ω m ) ] e ( u ν ) t / 2 + [ u + v 2 ( κ + i Δ ) ] e ( u + v ) t / 2 2 u ,
u = [ γ m κ i ( Δ ω m ) ] 2 4 g 2 | c s | 2 ,
v = γ m + κ + i ( Δ + ω m ) .
δ c ( 0 ) δ c ( 0 ) = 0 , δ b ( 0 ) δ b ( 0 ) = 0.
Δ X + 2 s s = [ 1 + 4 γ m N e f f η 1 ± 2 γ m M e f f * η 2 ± 2 γ m M e f f η 2 * ] e ± 2 r d ,
η 1 = κ [ ( γ m + κ ) 2 + ( Δ ω m ) 2 ] + ( γ m + κ ) g 2 | c s | 2 2 γ m κ [ ( γ m + κ ) 2 + ( Δ ω m ) 2 ] + 2 ( γ m + κ ) 2 g 2 | c s | 2 ,
η 2 = ( κ + i Δ ) [ γ m + κ + i ( Δ + ω m ) ] + g 2 | c s | 2 2 [ ( κ + i Δ ) ( γ m + ω m ) + g 2 | c s | 2 ] [ γ m + κ + i ( Δ + ω m ) ] .
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.