Perfect optomechanically induced transparency in two-cavity optomechanics

Lai-Bin Qian , Xiao-Bo Yan

Front. Phys. ›› 2023, Vol. 18 ›› Issue (5) : 52301

PDF (3981KB)
Front. Phys. ›› 2023, Vol. 18 ›› Issue (5) : 52301 DOI: 10.1007/s11467-023-1279-1
RESEARCH ARTICLE

Perfect optomechanically induced transparency in two-cavity optomechanics

Author information +
History +
PDF (3981KB)

Abstract

Here, we study the controllable optical responses in a two-cavity optomechanical system, especially on the perfect optomechanically induced transparency (OMIT) in the model which has never been studied before. The results show that the perfect OMIT can still occur even with a large mechanical damping rate, and at the perfect transparency window the long-lived slow light can be achieved. In addition, we find that the conversion between the perfect OMIT and optomechanically induced absorption can be easily achieved just by adjusting the driving field strength of the second cavity. We believe that the results can be used to control optical transmission in modern optical networks.

Graphical abstract

Keywords

perfect optomechanically induced transparency / slow light / optomechanically induced absorption / cavity optomechanics

Cite this article

Download citation ▾
Lai-Bin Qian, Xiao-Bo Yan. Perfect optomechanically induced transparency in two-cavity optomechanics. Front. Phys., 2023, 18(5): 52301 DOI:10.1007/s11467-023-1279-1

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

The interaction between light and matter is an interesting and important research subject in quantum optics. It is an important means to understand the microstructure of matter. Cavity optomechanics [1] can provide such a research platform where the macroscopic mechanical resonators and light fields interact with each other. With the development of nanotechnology, various physical systems which can exhibit such interaction have been proposed and investigated, such as Fabry−Perot cavities [2,3], whispering-gallery microcavities [4-6], superconducting circuits [7,8] and membranes [9-12]. The optomechanical interaction can strongly affect the motion of mechanical oscillator and the optical properties in these systems, and then various interesting quantum phenomena can be generated, such as ground-state cooling of mechanical modes [13-20], quantum entanglement [21-28], mechanical squeezing [29], unconventional photon blockade [30], and optomechanically induced transmission and absorption [31-42].

Recently, the researches on optomechanically induced transparency (OMIT) [43-57] and the associated slow light in optomechanical systems have attracted much attentions. A remarkable feature of OMIT is that there is a deep dip in the absorption curve accompanied with a steep dispersion behavior at the transparency window. The steep dispersion behavior can generate a drastic reduction in the group velocity of light passing through the system [56]. According to this effect, many schemes have been proposed to slow or stop light [56-65], which is very meaningful in the construction of quantum information networks. However, the ideal depth of the transparency window cannot be achieved due to the nonzero mechanical damping rate in the usual OMIT theory [48,54,56], which will cause a very limited slow light effect (usually on the order of milliseconds) [58-60]. Until recently, the perfect OMIT can be easily achieved using the mechanism of non-rotating wave approximation [47,57], and the slow light effect at the perfect transparency window can be greatly improved [57]. The perfect OMIT theory above is studied in standard single-cavity optomechanics. In fact, there are more abundant and interesting quantum phenomena in multi-mode optomechanical systems, such as quantum entanglement [23-25], optical nonreciprocity [66,67] and quantum nonlinearity [68,69]. Therefore, it is necessary to generalize the perfect OMIT theory to multi-mode optomechanical systems.

Here, we theoretically study the optical responses including perfect OMIT, long-lived slow light and optomechanically induced absorption in a multi-mode optomechanical system (comprising two optical modes and one mechanical mode), see Fig.1 [25,68-70]. First, we give the conditions under which the perfect OMIT will occur, and the window width expression of the perfect OMIT. As long as the conditions are satisfied, the perfect OMIT can be achieved even with a large mechanical damping rate. Secondly, the dispersion curve becomes very steep at the perfect transparency window where it can be proofed that the negative dispersion curve slope is exactly equal to the value of time delay (slow light) in the model. It means that the long-lived slow light can be achieved at the perfect transparency window. Thirdly, the driving strength and dissipation of the second cavity have a great effect on the optical response of the system. Especially, the conversion between perfect OMIT and optomechanically induced absorption in the model can be easily achieved just by adjusting the driving field strength of the second cavity, which can be used as an optical switch in modern optical networks.

2 Model and equations

We consider an optomechanical system consisting of a mechanical membrane and two cavities (see Fig.1). The mechanical membrane with frequency ωm and mass m is coupled to the two cavities via radiation pressure effects. The frequencies of the two cavities are described as ω1 and ω2, and the annihilation operators of the two cavities are denoted by a^1 and a^2, respectively. The position and momentum operators of the mechanical membrane are represented by q^ and p^, respectively. The radiation pressure effect can be described by q^(g2 a^2 a^2 g1 a^1 a^1) with the optomechanical coupling rate g i=ω i/L i where Li is the geometric length of the cavity i. The cavity a^1/a^2 is driven by a strong coupling/driving field with frequency ωc/ω d and amplitude εc/ε d. In addition, a weak probe field with frequency ωp and amplitude εp is injected into cavity a^1. For simplicity and without loss of generality, we can set ω1= ω2=ω 0 and L1=L2= L. Then, the Hamiltonian of the system Hs is

H s=ω0 a^1 a^1+ ω 0 a^2 a^2+ p ^22m+12mω m2 q^2+q ^g 0( a^2 a^2 a^1 a^1)+ i( a^1 εpeiω pta^1εp eiω pt) +iεc( a^1 eiω cta^1 eiω ct)+ iεd( a^2 eiω dta^2 eiω dt)

with g 0=ω 0/L.

Since in quantum mechanics the unitary transformation will not change the physics properties of the system, we can take the unitary transformation U^=ei(ω c a^1 a^1+ ωd a^2 a^2)t using the formula H= U^ HsU ^iU^U^ t. After this transformation, the Hamiltonian Hs will turn to H which can be given by

H=Δca^1 a^1+ Δ da^2 a^2+ p ^22m+12mω m2 q^2 + q^g0( a^2 a^2 a^1 a^1)+ i( a^1 εpeiδ ta^1εp eiδt) +iε c( a^1a^1)+ i εd(a^2a^2).

Here, Δ c/ d=ω 0ωc /d is the detuning between cavity field a^1/a^2 and coupling/driving field, and δ =ω pωc is the detuning between the probe field and coupling field.

In this paper, we deal with the mean response of the system to the probe field in the presence of the coupling and driving field, hence we do not include quantum fluctuations. According to the Heisenberg-Langevin equation and from Eq. (2), we can obtain the motion equations of mean values of the system operators as follow

a^ ˙1= [κ1+ i(Δc g0q^ )]a^1 +ε c+ε p eiδ t, a^ ˙ 2= [κ2+ i(Δd+ g0q^ )]a^2 +ε d, p^ ˙= γm p^m ωm2q^ +g0( a^1 a^1 a^2a^2 ), q^ ˙= p^m.

Here, we have used the usual factorization assumption, i.e., X^ Y^= X^ Y^ (which holds in the case of single-photon weak coupling, i.e., g0 ωm [71]), and γm, κ1 and κ2 are the damping rates of the mechanical membrane, cavity a^1 and a^2, respectively.

It is very difficult to obtain the exact solution of Eq. (3) because it is a nonlinear equation. However, in general the strength of the probe field is much smaller than that of the coupling field in cavity optomechanics. Hence, we can solve Eq. (3) by perturbation method. To this end, we assume that the solution of the mean values of the operators in Eq. (3) has the following form

s^ =s 0+ε p eiδ ts++εp eiδt s,

where s={q ,p,a 1,a 2}. It means that we can safely ignore the higher order terms of the probe field εp.

We are particularly interested in the properties of the field with frequency ωp in the output field of cavity a^1, which can be determined by the term a1+ in Eq. (4). The term a1 denotes the anti-Stokes effect and we don’t care here. Substituting Eq. (4) in Eq. (3) and ignoring the higher order terms of εp, we can obtain the important expression for the term a1+ (See Appendix A for detailed calculations) as

a 1+=1κ1i(δΔ1)+β 1 δ 2ω m2 +iδγm2iω m β 1κ1i(δ+Δ1)+ β 2κ2i(δΔ2) β 2κ2i(δ+Δ2).

Here,

β1= g02|a 10 | 22mω m,β2= g02| a20|22m ωm,

and Δ 1=Δ cg0q0, Δ2=Δ d+g0q0, and the expressions of q0, a10 and a20 can be found in Appendix A.

Because it is known that the coupling between the cavity and the resonator is strong at the near-resonant frequency, in this paper, we consider Δ 1Δ 2ω m and δω m. Then we have δ2ω m2 2ω m(δωm) and δ+ Δ1 δ+Δ22ω m. If we set x=δ ωm, Eq. (5) can be simplified as

a 1+=1κ1ix+β 1 γ m2 ixβ1 κ12iω m+β 2 κ2ixβ 2 κ22iω m.

Next, based on Eq. (7), we will study the optical response of the optomechanical system to the probe field, including the perfect optomechanically induced transparency, slow light and optomechanically induced absorption, respectively.

3 Perfect optomechanically induced transparency

According to input-output relation [48], the quadrature of the optical component with frequency ω p in the output field can be defined as εT= 2κ 1a1+ [48]. The real part Re[εT] and imaginary part Im [εT] represent the absorptive and dispersive behavior of the optomechanical system to the probe field, respectively. We first give the conditions of perfect OMIT in the model. According to the conclusions in Ref. [47], the conditions of perfect OMIT are determined by the pole location of the subfraction in Eq. (7), i.e.,

γm2ixβ1 κ12iω m+β 2 κ2ixβ 2 κ22iω m=0.

It can be known from Eq. (8) that the perfect OMIT cannot occur at the resonant frequency x=0 because in this case β 2=γ m κ2(κ2 2 +4ωm2)2(κ1κ2+4ω m2)<0 which is in contradiction with the definition in Eq. (6).

We first study the case of large κ2, i.e., the cavity damping rate κ2 is much larger than the detuning x where the perfect OMIT appears (κ2 | x|), and from Eq. (8) we can obtain the conditions of perfect OMIT as

x=2κ2ω m[ γmκ2(κ12+ 4ωm2)2 β1(κ 1 κ2+4ω m2)] (κ 12+ 4ωm2)ξ2β 1κ1(κ22+ 4ωm2),β2= κ2(κ22+ 4ωm2)[2 β1κ1 γm( κ12+4ω m2)] 8ωm2(κ12+ 4ωm2),

with ξ =8κ2 ωm2+γm(κ22+ 4ωm2).

The perfect OMIT can be achieved if the conditions in Eq. (9) are satisfied. In Fig. (2), we plot the real part Re[εT] (red-solid) vs. frequency detuning x with ωm=κ 1=κ 2=10 4, γm=1, β1= 3×104 and β2= 1250 according to Eq. (9). With these parameters, according to Eq. (9), the transparency window will appear at x 1.25 which is consistent with the numerical result in Fig. (2). The inset in Fig. (2) shows the OMIT profile in a large scale, from which it can be seen that the width of the transparency window is very narrow. From Eq. (7), we can obtain the expression of the width ΓOMIT (full width at half maximum) of transparent window, but it is too lengthy to be reported here. However, if we take the case of equal cavity damping rate (κ1= κ2=κ), the width ΓOMIT can be given as

Γ OMIT= κ{32β 1 ωm2η+κ [η+ ωm(4 β12κ γm)]2}2η+ κ{32β 1 ωm2η+κ [η ωm(4 β12κ γm)]2}2η κ,

with η =8κω m2 2β 1κ+ γm(κ2+4ω m2). With the parameters above, the width Γ OMIT5.998 which shows an excellent agreement with the numerical result (see the blue-dashed line) in Fig.2.

One of the advantages of the perfect OMIT theory is the perfect transparency window can still appear even with a large mechanical damping rate γ m. In Fig.3, we plot the real part Re[εT] vs. frequency detuning x for large mechanical damping rate γm= 10 (red-solid) with β1= 3×105, and β2= 1.25×104 according to Eq. (9), and for γm=100 (blue-dashed) with β1=3× 106, and β2=1.25× 105 according to Eq. (9). The other parameters are ωm=κ1= κ2=104. It can be clearly seen from Fig.3 that the perfect OMIT can do occur with large mechanical damping rate γm as long as the conditions in Eq. (9) are satisfied. In addition, according to Eq. (10), the width ΓOMIT=59.83 (red-solid) and Γ OMIT=583.79 (blue-dashed), which are consistent with the results in Fig.3. In Fig.4, we plot the imaginary part Im[εT] vs. frequency detuning x with the same parameters as Fig.3. It can be seen from Fig.4 that the dispersion curve becomes very steep at the perfect transparency window and the slope is negative there.

The above discussion is based on the condition that the cavity damping rate κ2 is much larger than the transparent window position x. While if the second cavity is the microwave cavity, the cavity damping rate κ2 can be very small [72,73]. If κ2 is very small, the above calculation would be a little more complicated. However, we can always obtain the window position x and driving strength β2 through numerical methods according to Eq. (7). For example, for parameters κ2= 10, ωm=104, κ1= 4×103, γm= 1 and β1=105, we numerically obtain the window position x5.55 and β2=5.91. With these parameters, in Fig.5 we plot the real part Re[εT] vs. frequency detuning x, and the inset in Fig.5 shows a zoom-in of the transparency window at x5.55. It can be clearly seen from Fig.5 that the perfect OMIT can still occur with small cavity damping rate κ2.

4 Ultraslow light

The time delay (slow light effect) of the probe field with frequency ωp in the output field can be determined by [54,56,59]

τ=arg(εT 1)ωp.

The positive (negative) value of the time delays represents slow (fast) light [74] in the system. According to Eqs. (7), (9) and (11), we can obtain the analytic expressions of time delays τ, but it is too tedious to be reported here. However, the time delay at the transparent window can be obtained as

τ =κ1[8κ 2 ωm2+γm(κ22+ 4ωm2)] 4β1κ2ω m2 κ 12(κ2 2 +4ωm2)2κ2ω m2(κ1 2 +4ωm2).

In Fig.6, we plot the time delay τ vs. frequency detuning x for γm= 1 (red-solid) with β1= 3×104 and β2= 1250 (same as Fig.2), and for γm= 10 (blue-dashed) with β1= 3×105 and β2= 1.25×104 (same as Fig.3). The other parameters are ωm=κ 1=κ 2=10 4 (same as Fig.2). It can be seen from Fig.6 that the time delay τ exactly takes the maximum at the transparency window where the steepest dispersion appears. In fact, it can be proofed that the time delay at the transparency window is exactly equal to the negative dispersion curve slope there. It means that the steeper the slope of dispersion curve is, the larger the slow light effect becomes. From Fig.6, the maximum delay τmax0.67 for red-solid line and τmax0.067 for blue-dashed line, which are very consistent with the results according to Eq. (12).

We also study the effect of cavity damping rate ( κ1,κ 2) on the time delay τ. For simplicity, we also take κ1=κ 2 here. In Fig.7, we plot the time delay τ vs. frequency detuning x for κ1= κ2=2× 104 (red-solid) with β1= 3×104 and β2= 104 according to Eq. (9), and for κ1= κ2=8× 103 (blue-dashed) with β1= 3×104 and β2= 160 according to Eq. (9). The other parameters are ωm=104 and γm= 1. It can be seen from Fig.7 that the maximum time delay (at the transparency window) in the unresolved sideband regime (κ1> ωm) will larger than that in the resolved sideband regime ( κ1<ω m). With the parameters, the maximum time delay τmax1.33 in the unresolved sideband regime (see red-solid). It means that the τmax1.33 s (if the units of physical quantities above are Hertz), which is actually a long-lived slow light. This long-lived slow light may be used for OMIT-based memories in the future.

5 Optomechanically induced absorption

Compared with Re[εT]=0 at the window of the perfect OMIT, the phenomenon of optomechanically induced absorption will appear at the near resonance position x=0 ( Re[εT] shows a noticeable increase at position x=0), if some conditions are satisfied. We first do some qualitative analyses of these conditions. At the position x=0, we have

ε T(x=0 )=2κ 1κ 1+β1 γ m2β1κ 12iω m+β 2κ 2 β 2κ22iω m,

which means that Re[εT] will display a noticeable increase at x=0 if the ratio β2/κ 2 is large enough. In other words, if κ2 is small enough or the driving strength β2 is large enough, the phenomenon of optomechanically induced absorption at x=0 can occur.

In Fig.8, we plot the Re [εT] vs. the frequency detuning x for κ2= 10 (blue-dashed) and κ2= 1 (red-solid) with parameters κ1= 4×103, ωm= 104, γm=1, β1= 105 and β2=100. It can be seen from Fig.8 that the phenomenon of optomechanically induced absorption at x=0 becomes very significant with the decrease of κ2. In addition, according to Eq. (13), we have εT(x=0) 0.58 for κ2=10 and εT(x=0 )1.60 for κ2=1, which are clearly consistent with the numerical results in Fig.8.

Generally, once a quantum device has been manufactured, its characteristic parameters are fixed. Therefore, it is not easy to achieve optomechanically induced absorption by changing the dissipation rate κ2. While it is very convenient to adjust the driving strength β2 in experiments. In Fig.9, we plot the Re[εT] vs. the frequency detuning x for β2=105 (red-solid) with parameters κ2=10, κ1= 4×103, ωm= 104, γm=1, and β1= 105. With the parameters, Re[εT]=1.9995 at x=0 according to Eq. (13), which is very much in agreement with the results in Fig.9. Hence the optomechanically induced absorption can be achieved at x=0 with the driving strength β2=105. In addition, the absorption curve is approximately a horizontal line over a very wide frequency range, see the red-solid line in the inset in Fig.9. For comparison, we also plot the curve of perfect OMIT with the same parameters but β2= 5.91 (blue-dashed), the transparency window x 5.55 (see the blue-dashed line in the inset of Fig.9). It means that the conversion between perfect OMIT and optomechanically induced absorption of the probe field can be achieved simply by adjusting the driving strength β2. These results can be used to achieve the optical switch in modern optical networks.

6 Conclusion

In summary, we have theoretically studied the controllable optical responses in a two-cavity optomechanical system, especially on the perfe ct optomechanically induced transparency (OMIT), long-lived slow light and optomechanically induced absorption in the model. From the theoretical results, we can draw some conclusions. First, the perfect OMIT can be still achieved even with a large mechanical damping rate γm, which is difficult to be realized in the usual OMIT theory. Second, at the transparency window of the perfect OMIT, the long-lived slow light can be achieved, which can be used for OMIT-based memories in the future. Third, an optical switch taking advantage of the conversion between perfect OMIT and optomechanically induced absorption can be achieved just by adjusting the driving field strength of the second cavity. We believe that the results can be used to control optical transmission in quantum information processing.

7 Derivation of a1+

Substituting Eq. (4) in Eq. (3), just keeping the constant term and the first order term of εp, and then comparing the coefficients of the terms eiδt and eiδ t on both sides of the equation, we can obtain

q0= g0( |a10|2 |a20|2)mωm2,

q=g0(a10a1++a10 a 1 a 20a2+a20 a2)m (ωm2 δ2+ iγmδ ) ,

q+=g0(a 10 a 1++ a10a 1a20 a2+ a20 a2 )m( ωm2 δ2 iγmδ ) ,

a10= εc κ1+ iΔ1,

a1+ = ig0q+a10+ 1κ1i( δΔ1),

a1= ig0qa10 κ1+ i(δ+Δ1),

a20= εd κ2+ iΔ2,

a2+ = ig0q+a 20κ2 i(δΔ2),

a2= i g 0q a20κ2+i( δ+Δ2),

with Δ 1=Δ cg0q0 and Δ2=Δ d+g0q0.

In order to obtain the expression of a1+, we need to give the expression of q+ in Eq. (A2). To this end, according to Eqs. (A3)–(A9) and using the fact q=q+, we have

Aq += a10a1+ +a10 a1 a 20a2+ a20 a2 ,

a10 a 1+=B( ig0|a10 | 2 q+ +a10),

a10 a1 = Cig0| a10 | 2q +,

a20 a 2+=Dig0|a20 | 2 q+ ,

a20 a2 =E ig0| a20 | 2q +,

with

A=m(ωm2 δ2 iγmδ )g0,

B=1κ 1i(δΔ1),

C=1 κ1 i(δ+Δ1),

D=1 κ2 i(δΔ2),

E=1 κ2 i(δ+Δ2).

From Eqs. (A11)–(A14), we can obtain the expression of a10 a1 ++a 10 a1 a20 a2+ a20a2 , and then combining Eq. (A10), the expression of q+ can be obtained as

q+=B a10A ig0[(B C) | a 10 | 2+(DE)|a20|2].

Substituting Eq. (A20) into Eq. (A5), we obtain a1+ as

a1+ =ig0B2| a10 | 2Aig0[(B C) | a 10 | 2+(DE)|a20|2] +B =[Ai g 0(C|a10|2+(DE)|a20 | 2)]BAi g 0[(BC )|a10|2+(DE)|a20 | 2] =11 B ig0| a10 | 2Aig0(C| a10 | 2+D| a20 | 2E | a 20 | 2).

Substituting Eqs. (A15)–(A19) into Eq. (A21) and setting β1= g0 2 | a10|22m ωm and β2= g0 2 | a20|22m ωm, we can obtain Eq. (5) in text for a1 +.

References

[1]

M. Aspelmeyer, T. J. Kippenberg, F. Marquardt. Cavity optomechanics. Rev. Mod. Phys., 2014, 86(4): 1391

[2]

S. Gigan, H. Böhm, M. Paternostro, F. Blaser, G. Langer, J. Hertzberg, K. Schwab, D. Bäuerle, M. Aspelmeyer, A. Zeilinger. Self-cooling of a micromirror by radiation pressure. Nature, 2006, 444(7115): 67

[3]

O. Arcizet, P. F. Cohadon, T. Briant, M. Pinard, A. Heidmann. Radiation-pressure cooling and optomechanical instability of a micromirror. Nature, 2006, 444(7115): 71

[4]

T. J. Kippenberg, H. Rokhsari, T. Carmon, A. Scherer, K. J. Vahala. Analysis of Radiation-Pressure Induced Mechanical Oscillation of an Optical Microcavity. Phys. Rev. Lett., 2005, 95(3): 033901

[5]

M. Tomes, T. Carmon. Photonic micro-electromechanical systems vibrating at X-band (11-GHz) rates. Phys. Rev. Lett., 2009, 102(11): 113601

[6]

X. Jiang, Q. Lin, J. Rosenberg, K. Vahala, O. Painter. High-Q double-disk microcavities for cavity optomechanics. Opt. Express, 2009, 17(23): 20911

[7]

C. A. Regal, J. D. Teufel, K. W. Lehnert. Measuring nanomechanical motion with a microwave cavity interferometer. Nat. Phys., 2008, 4(7): 555

[8]

J. D. Teufel, D. Li, M. S. Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, R. W. Simmonds. Circuit cavity electromechanics in the strong-coupling regime. Nature, 2011, 471(7337): 204

[9]

J. D. Thompson, B. M. Zwickl, A. M. Jayich, F. Marquardt, S. M. Girvin, J. G. E. Harris. Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane. Nature, 2008, 452(7183): 72

[10]

A. M. Jayich, J. C. Sankey, B. M. Zwickl, C. Yang, J. D. Thompson, S. M. Girvin, A. A. Clerk, F. Marquardt, J. G. E. Harris. Dispersive optomechanics: A membrane inside a cavity. New J. Phys., 2008, 10(9): 095008

[11]

J. C. Sankey, C. Yang, B. M. Zwickl, A. M. Jayich, J. G. E. Harris. Strong and tunable nonlinear optomechanical coupling in a low-loss system. Nat. Phys., 2010, 6(9): 707

[12]

M. Karuza, C. Biancofiore, M. Bawaj, C. Molinelli, M. Galassi, R. Natali, P. Tombesi, G. Di Giuseppe, D. Vitali. Optomechanically induced transparency in a membrane-in-the-middle setup at room temperature. Phys. Rev. A, 2013, 88(1): 013804

[13]

F. Marquardt, J. P. Chen, A. A. Clerk, S. M. Girvin. Quantum theory of cavity-assisted sideband cooling of mechanical motion. Phys. Rev. Lett., 2007, 99(9): 093902

[14]

I. Wilson-Rae, N. Nooshi, W. Zwerger, T. J. Kippenberg. Theory of ground state cooling of a mechanical oscillator using dynamical backaction. Phys. Rev. Lett., 2007, 99(9): 093901

[15]

B. He, L. Yang, Q. Lin, M. Xiao. Radiation pressure cooling as a quantum dynamical process. Phys. Rev. Lett., 2017, 118(23): 233604

[16]

D. Y. Wang, C. H. Bai, S. Liu, S. Zhang, H. F. Wang. Optomechanical cooling beyond the quantum backaction limit with frequency modulation. Phys. Rev. A, 2018, 98(2): 023816

[17]

J. Y. Yang, D. Y. Wang, C. H. Bai, S. Y. Guan, X. Y. Gao, A. D. Zhu, H. F. Wang. Ground-state cooling of mechanical oscillator via quadratic optomechanical coupling with two coupled optical cavities. Opt. Express, 2019, 27(16): 22855

[18]

J. Wang. Ground-state cooling based on a three-cavity optomechanical system in the unresolved-sideband regime. Chin. Phys. B, 2021, 30(2): 024204

[19]

Q. He, F. Badshah, Y. Song, L. Wang, E. Liang, S. L. Su. Force sensing and cooling for the mechanical membrane in a hybrid optomechanical system. Phys. Rev. A, 2022, 105(1): 013503

[20]

J. Yang, C. Zhao, Z. Yang, R. Peng, S. Chao, L. Zhou. Nonreciprocal ground-state cooling of mechanical resonator in a spinning optomechanical system. Front. Phys., 2022, 17(5): 52507

[21]

J. Li, G. Li, S. Zippilli, D. Vitali, T. Zhang. Enhanced entanglement of two different mechanical resonators via coherent feedback. Phys. Rev. A, 2017, 95(4): 043819

[22]

C. H. Bai, D. Y. Wang, H. F. Wang, A. D. Zhu, S. Zhang. Robust entanglement between a movable mirror and atomic ensemble and entanglement transfer in coupled optomechanical system. Sci. Rep., 2016, 6(1): 33404

[23]

X. B. Yan. Enhanced output entanglement with reservoir engineering. Phys. Rev. A, 2017, 96(5): 053831

[24]

Z. J. Deng, X. B. Yan, Y. D. Wang, C. W. Wu. Optimizing the output-photon entanglement in multimode optomechanical systems. Phys. Rev. A, 2016, 93(3): 033842

[25]

X. B. Yan, Z. J. Deng, X. D. Tian, J. H. Wu. Entanglement optimization of filtered output fields in cavity optomechanics. Opt. Express, 2019, 27(17): 24393

[26]

Y. T. Chen, L. Du, Y. Zhang, J. H. Wu. Perfect transfer of enhanced entanglement and asymmetric steering in a cavity-magnomechanical system. Phys. Rev. A, 2021, 103(5): 053712

[27]

S. Y. Guan, H. F. Wang, X. X. Yi. Cooperative-effect-induced one-way steering in open cavity magnonics. npj Quantum Inf., 2022, 8: 102

[28]

Y. Zeng, B. Xiong, C. Li. Suppressing laser phase noise in an optomechanical system. Front. Phys., 2022, 17(1): 12503

[29]

C. H. Bai, D. Y. Wang, S. Zhang, S. Liu, H. F. Wang. Engineering of strong mechanical squeezing via the joint effect between Duffing nonlinearity and parametric pump driving. Photon. Res., 2019, 7(11): 1229

[30]

L. J. Feng, L. Yan, S. Q. Gong. Unconventional photon blockade induced by the self-Kerr and cross-Kerr nonlinearities. Front. Phys., 2023, 18(1): 12304

[31]

K. Quand G. S. Agarwal, Phonon-mediated electromagnetically induced absorption in hybrid opto-electromechanical systems, Phys. Rev. A 87, 031802(R) (2013)

[32]

G. S. Agarwal, S. Huang. Nanomechanical inverse electromagnetically induced transparency and confinement of light in normal modes. New J. Phys., 2014, 16(3): 033023

[33]

J. Wang. Optomechanically induced tunable ideal nonreciprocity in optomechanical system with Coulomb interaction. Quantum Inform. Process., 2022, 21(7): 238

[34]

B. He, L. Yang, and M. Xiao, Dynamical phonon laser in coupled active-passive microresonators, Phys. Rev. A 94, 031802(R) (2016)

[35]

L. Du, Y. M. Liu, B. Jiang, Y. Zhang. All-optical photon switching, router and amplifier using a passive-active optomechanical system. EPL, 2018, 122(2): 24001

[36]

C. C. Xia, X. B. Yan, X. D. Tian, F. Gao. Ideal optical isolator with a two-cavity optomechanical system. Opt. Commun., 2019, 451: 197

[37]

X. B. Yan, C. L. Cui, K. H. Gu, X. D. Tian, C. B. Fu, J. H. Wu. Coherent perfect absorption, transmission, and synthesis in a double-cavity optomechanical system. Opt. Express, 2014, 22(5): 4886

[38]

X. B. Yan, H. L. Lu, F. Gao, L. Yang. Perfect optical nonreciprocity in a double-cavity optomechanical system. Front. Phys., 2019, 14(5): 52601

[39]

Y. T. Chen, L. Du, Y. M. Liu, Y. Zhang. Dual-gate transistor amplifier in a multimode optomechanical system. Opt. Express, 2020, 28(5): 7095

[40]

T. Wang, C. H. Bai, D. Y. Wang, S. Liu, S. Zhang, H. F. Wang. Optomechanically induced Faraday and splitting effects in a double-cavity optomechanical system. Phys. Rev. A, 2021, 104(1): 013721

[41]

L. Qi, G. L. Wang, S. Liu, S. Zhang, H. F. Wang. Dissipation-induced topological phase transition and periodic-driving-induced photonic topological state transfer in a small optomechanical lattice. Front. Phys., 2021, 16(1): 12503

[42]

Z. R. Zhong, L. Chen, J. Q. Sheng, L. T. Shen, S. B. Zheng. Multiphonon-resonance quantum Rabi model and adiabatic passage in a cavity-optomechanical system. Front. Phys., 2022, 17(1): 12501

[43]

S. Shahidani, M. H. Naderi, M. Soltanolkotabi. Control and manipulation of electromagnetically induced transparency in a nonlinear optomechanical system with two movable mirrors. Phys. Rev. A, 2013, 88(5): 053813

[44]

Y. X. Liu, M. Davanco, V. Aksyuk, K. Srinivasan. Electromagnetically induced transparency and wideband wavelength conversion in silicon nitride microdisk optomechanical resonators. Phys. Rev. Lett., 2013, 110(22): 223603

[45]

G. Pan, R. Xiao, H. Chen, J. Gao. Multicolor optomechanically induced transparency in a distant nano-electro-optomechanical system assisted by two-level atomic ensemble. Laser Phys., 2021, 31(6): 065202

[46]

H. Xiong, L. G. Si, A. S. Zheng, X. Yang, Y. Wu. Higher-order sidebands in optomechanically induced transparency. Phys. Rev. A, 2012, 86(1): 013815

[47]

X. B. Yan. Optomechanically induced transparency and gain. Phys. Rev. A, 2020, 101(4): 043820

[48]

G. S. Agarwaland S. Huang, Electromagnetically induced transparency in mechanical effects of light, Phys. Rev. A 81, 041803(R) (2010)

[49]

T. Bodiya, V. Sudhir, C. Wipf, N. Smith, A. Buikema, A. Kontos, H. Yu, N. Mavalvala. Sub-Hertz optomechanically induced transparency with a kilogram-scale mechanical oscillator. Phys. Rev. A, 2019, 100(1): 013853

[50]

A. Kronwald, F. Marquardt. Optomechanically induced transparency in the nonlinear quantum regime. Phys. Rev. Lett., 2013, 111: 133601

[51]

H. Jing, S. K. Özdemir, Z. Geng, J. Zhang, X. Y. Lü, B. Peng, L. Yang, F. Nori. Optomechanically-induced transparency in parity−time-symmetric microresonators. Sci. Rep., 2015, 5(1): 9663

[52]

W. Li, Y. Jiang, C. Li, H. Song. Parity−time-symmetry enhanced optomechanically-induced-transparency. Sci. Rep., 2016, 6(1): 31095

[53]

Y. C. Liu, B. B. Li, Y. F. Xiao. Electromagnetically induced transparency in optical microcavities. Nanophotonics, 2017, 6(5): 789

[54]

S. Weis, R. Rivie’re, S. Deléglise, E. Gavartin, O. Arcizet, A. Schliesser, T. J. Kippenberg. Optomechanically induced transparency. Science, 2010, 330(6010): 1520

[55]

X. B. Yan, K. H. Gu, C. B. Fu, C. L. Cui, R. Wang, J. H. Wu. Optical switching of optomechanically induced transparency and normal mode splitting in a double-cavity system. Eur. Phys. J. D, 2014, 68(5): 126

[56]

A. H. Safavi-Naeini, T. P. M. Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, O. Painter. Electromagnetically induced transparency and slow light with optomechanics. Nature, 2011, 472(7341): 69

[57]

X. B. Yan. Optomechanically induced optical responses with non-rotating wave approximation. J. Phys. At. Mol. Opt. Phys., 2021, 54(3): 035401

[58]

B. Chen, C. Jiang, K. D. Zhu. Slow light in a cavity optomechanical system with a Bose−Einstein condensate. Phys. Rev. A, 2011, 83(5): 055803

[59]

D. Tarhan, S. Huang, Ö. E. Müstecaplioğlu. Superluminal and ultraslow light propagation in optomechanical systems. Phys. Rev. A, 2013, 87(1): 013824

[60]

K. H. Gu, X. B. Yan, Y. Zhang, C. B. Fu, Y. M. Liu, X. Wang, J. H. Wu. Tunable slow and fast light in an atom-assisted optomechanical system. Opt. Commun., 2015, 338: 569

[61]

D. E. Chang, A. H. Safavi-Naeini, M. Hafezi, O. Painter. Slowing and stopping light using an optomechanical crystal array. New J. Phys., 2011, 13(2): 023003

[62]

M. J. Akram, M. M. Khan, F. Saif. Tunable fast and slow light in a hybrid optomechanical system. Phys. Rev. A, 2015, 92(2): 023846

[63]

X. B. Yan. Optomechanically induced ultraslow and ultrafast light. Physica E, 2021, 131: 114759

[64]

L. Wang, Y. T. Chen, K. Yin, Y. Zhang. Nonreciprocal transmission and asymmetric fast–slow light effect in an optomechanical system with two PT-symmetric mechanical resonators. Laser Phys., 2020, 30(10): 105205

[65]

Y. N. Zhao, T. Wang, D. Y. Wang, X. Han, S. Zhang, H. F. Wang. Optical amplification and fast-slow light in a three-mode cavity optomechanical system without rotating wave approximation. Photonics, 2021, 8(9): 384

[66]

X. W. Xu, Y. Li. Optical nonreciprocity and optomechanical circulator in three-mode optomechanical systems. Phys. Rev. A, 2015, 91(5): 053854

[67]

N. R. Bernier, L. D. Tóth, A. Koottandavida, M. A. Ioannou, D. Malz, A. Nunnenkamp, A. K. Feofanov, T. J. Kippenberg. Nonreciprocal reconfigurable microwave optomechanical circuit. Nat. Commun., 2017, 8(1): 604

[68]

M. Ludwig, A. H. Safavi-Naeini, O. Painter, F. Marquardt. Enhanced quantum nonlinearities in a two-mode optomechanical system. Phys. Rev. Lett., 2012, 109(6): 063601

[69]

X. Y. Lü, W. M. Zhang, S. Ashhab, Y. Wu, F. Nori. Quantum-criticality-induced strong Kerr nonlinearities in optomechanical systems. Sci. Rep., 2013, 3(1): 2943

[70]

L. Tian. Robust photon entanglement via quantum interference in optomechanical interfaces. Phys. Rev. Lett., 2013, 110(23): 233602

[71]

A. Nunnenkamp, K. Børkje, S. M. Girvin. Single-photon optomechanics. Phys. Rev. Lett., 2011, 107(6): 063602

[72]

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

[73]

Sh. Barzanjeh, M. Abdi, G. J. Milburn, P. Tombesi, D. Vitali. Reversible optical-to-microwave quantum interface. Phys. Rev. Lett., 2012, 109(13): 130503

[74]

M. S. Bigelow, N. N. Lepeshkin, R. W. Boyd. Superluminal and slow light propagation in a room-temperature solid. Science, 2003, 301(5630): 200

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (3981KB)

1479

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/