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

Enhanced optomechanical interaction in coupled microresonators

Open Access Open Access

Abstract

We investigate the optomechanical interaction in two coupled microresonators. Compared to the single resonator optomechanical system where the input light is required to detune from the cavity resonance to generate two asymmetrical sidebands and thus large mechanical damping/amplification, the coupled resonator system can allow both the input light and its frequency sideband to be on resonance. In this configuration, we find that the optomechanical interaction can be enhanced and optically induced energy transfer between different mechanical oscillators is possible.

© 2012 Optical Society of America

1. Introduction

Optomechanical systems have attracted great interest for the ability to combine nanomechanics and nanophotonics [16]. These systems exploit optical forces arising from the perturbation of light fields by mechanical displacements [7, 8]. Typically the optomechanical interaction is weak and strong optical power is needed to generate detectable optical forces [9]. By resonance enhancement of both optical field and mechanical motion, the optomechanical system can go beyond mesoscale [10] and allow applications in both fundamental physics [1116] and applied engineering [1724], including quantum ground state cooling [12], optomechanically induced transparency [3, 21], microwave oscillator [2426], integrated pressure sensing [23], etc.

Generally, most of the reported optomechanical systems are based on the interaction between mechanical modes and a single high-Q optical resonance. When the input light is coupled to an optical resonator, the mechanical motion of the resonator modulates the intra-cavity optical field with the mechanical resonance frequency and generates two frequency sidebands. To achieve effective energy transfer between the light field and mechanical motions, the frequency of the input light must be detuned from the cavity resonance so that the resonator will favor one sideband which is closer to the optical resonance frequency. However, this also indicates that only a fraction of input power enters the resonator and contributes to the optomechanical interaction.

Compound resonator systems consist of two or more coupled resonators and generally exhibit multiple resonances. Recent progress in the fabrication of extremely high-Q microresonators and high-frequency mechanical oscillators makes it possible to couple multiple optical resonances via optomechanical interaction. The major benefit for building optomechanical system based on the compound resonator system is its controllable resonance functions. The frequency spacing between optical resonances can be continuously tuned by changing the coupling rate between the resonators [15], which provides additional freedom for designing optomechanical systems. Moreover, in contrast to the single resonator optomechanical system where mechanical modes of one mechanical resonator interact with the same optical field, mechanical modes of separate mechanical resonators can couple to the optical field inside individual resonator. This feature presents the opportunity for light-induced energy transfer and synchronization between mechanical oscillators [2729].

In this paper, we investigate the optomechanical interaction in two coupled microresonators while utilizing the composite resonances of the system. By taking advantages of the split resonance, we can bypass the requirement of frequency detuning for the input light in a single resonator system. For the thermally driven system, we show that the optomechanical interaction can be greatly enhanced compared to the single resonator system, and a controllable energy transfer between separate mechanical oscillators is possible. For the external modulation driven system, the power efficiency to obtain the Optomechanically Induced Transparency (OMIT) phenomenon can be greatly increased.

2. Conceptual model

The schematics of optomechanical systems based on a single resonator and two coupled resonators are shown in Fig. 1. The optical field inside each resonator is perturbed by the mechanical motion and consequently exerts a modulated optical force on the mechanical mode. The modulated force has two components: one is in phase of the motion, which corresponds to the optical spring effect and changes the mechanical resonance frequency; the other is out of phase of the motion, which modifies the mechanical damping. Both negative and positive damping can be obtained by detuning the input light with respect to the optical resonance.

 figure: Fig. 1

Fig. 1 The schematics and sideband picture of the optomechanical system based on (a) a single resonator; (b) two coupled resonators. The mechanical motion modifies the optical resonance frequency.

Download Full Size | PDF

The dynamic back-action can be also simply understood by using the sideband picture [1]. In Fig. 1(a), the input light with the angular frequency ω is coupled to the resonator. The optical field inside the cavity is modulated by the mechanical motion with the mechanical resonance frequency Ωm and this modulation generates two sidebands at ω ± Ωm. If the input frequency is not equal to the optical resonance frequency, the resonator will favor one sideband whose frequency is closer to the resonance. Thus the two sidebands will have different amplitudes. Because of the different photon energies of the two sidebands, the total optical energy inside the resonator will not be conserved, corresponding to an energy transfer between the optical field and the mechanical motion. If the input light is red (blue) detuned from the resonance, energy is transferred from (to) the mechanical motion and the input light produces a damping (amplification) force on the mechanical motion.

Based on the sideband picture, the origin of the mechanical damping/amplification is the imbalanced amplitudes of sidebands. The input light is required to be detuned from the cavity resonance to generate asymmetric sidebands in a single resonator optomechanical system. In other words, the input light is not resonance-enhanced. In the resolved-sideband regime, where the mechanical resonance frequency is much larger than the cavity linewidth (Γ), only a small fraction of the input power can enter the cavity and interact with the mechanical motion.

The motivation of introducing coupled resonator system is to eliminate the detuning requirement of the input light. A conceptual model of optomechanical systems based on two coupled resonators is shown in Fig. 1(b). The first resonator is evanescently coupled to the bus waveguide and the second resonator. The system can exhibit two split resonances in the transmission spectrum when the two resonators are strongly coupled. By controlling the coupling strength between the resonators, we can set the spacing between two split resonance frequencies close to the mechanical frequency. Therefore, if the frequency of the input light equals one optical resonance frequency, both the input light and one of its sidebands can be simultaneously resonance-enhanced. Note that here the first order sidebands are mainly generated by phase modulation of the optical field. The large asymmetry of sideband amplitudes will increase the back-action force, leading to an enhanced optomechanicam interaction. In the following sections, by using a perturbation method, we demonstrate that the coupled resonator system can provide much larger optomechanical interaction compared to the single resonator system.

3. Quantitative formalism and simulation results

As shown in Fig. 1(b), the optomechanical system consists of two optical resonators. The optical field inside each resonator is modulated by one mechanical mode. Here we assume the two mechanical modes have no direct coupling. The dynamic response of the system can be described using the coupled equations:

da1dt=(iΔ1Γ1/2ig1x1)a1+iΓeAin+iκa2
da2dt=(iΔ2Γ2/2ig2x2)a2+iκa1
d2xidt2+Γmidxidt+Ωmi2xi=Fimi+Fthmi
where Ain is the power amplitude of the input light; ai (i = 1, 2) is the energy amplitude of the optical field inside each resonator; κ is the decay rate for the coupling between the resonators; Γ1, Γe are the loaded decay rate and the external decay rate of the first resonator; Γ2 is the intrinsic decay rate of the second resonator; xi, Γmi, Ωmi, Fi, Fth, mi (i = 1, 2) are the displacement, decay rate, resonance frequency, optically induced force, thermally induced force and effective mass of each mechanical mode. In the general case, the two resonators have different resonance frequencies ω1 and ω2, and the frequency detuning of the input light from the resonance frequency is Δi = ωωi. The optomechanical coupling coefficient gi is defined as gi = dωi/dxi. The above equations can be solved in the steady-state for optical field amplitudes and mechanical displacements:
a¯1=iΓeAiniΔ¯1Γ1+κ2iΔ¯2Γ2/2
a¯2=iκa¯1iΔ¯2Γ2/2
x¯i=gi|a¯i|2miΩmi2ω
We assume the optomechanical coupling is not too strong to generate bistability and the input optical power is small so that the modified detuning, Δ̄i = Δigii, is approximately equal to Δi. The thermal forces are neglected as well for the equilibrium solution. The system provides two types of optical transmission based on the optical coupling rate (κ) between two resonators, as shown in Fig. 2.

 figure: Fig. 2

Fig. 2 Optical field energies of the two resonators in the: (a) weak coupling regime, κ = Γ2; (b) strong coupling regime, κ = 10Γ2. The intrinsic Q-factors are Qi1 = Qi2 = 106, and the external Q-factor of the first resonator is Qe1 = 105. The input optical power is 1mW. For simplicity, the two resonators are assumed to have the same intrinsic resonance, Δ = Δ1 = Δ2.

Download Full Size | PDF

In the weak coupling regime (Fig. 2(a)), because the second resonator has a larger loaded Q-factor, the circulating power inside the first resonator is absorbed into the second resonator in the vicinity of the resonance. In the strong coupling regime (Fig. 2(b)), because the coupling rate between two resonators is larger than their linewidth, the system exhibits two split resonances. The frequency spacing between the split resonances is approximately equal to 2κ for κ ≫ Γ2.

Using the perturbation method we can obtain the dynamic equations for perturbed optical field amplitudes δai and mechanical displacements δxi (i = 1, 2)

dδa1dt=(iΔ¯Γ12)δa1ig1a¯1δx1+iκδa2+iΓeδAindδa2dt=(iΔ¯Γ22)δa2ig2a¯2δx2+iκδa1d2δxidt2+Γmidδxidt+Ωmi2δxi=gimiω(a¯iδai*+a¯i*δai)+Fthmi

Based on the source of the driven force, optomechanical systems can be categorized into: (1) thermal force driven system (2) external modulation driven system. For the thermally driven system, the mechanical oscillation is driven by the thermal reservoir and the amplitude of the input light is constant. The system response can be found by solving Eq. (7) in the frequency domain with δAin = 0. For the external modulation driven system, the input optical field is modulated near the mechanical resonance. The back-action force generated by the modulated optical field is much larger than the thermal forces, which means we can neglect Fth and solve Eq. (7) with a time-dependent δAin. Next we solve these equations in frequency domain under three different situations.

3.1. Strong coupling under thermal force driven

First we consider that the optical resonators are strongly coupled with each other, which creates two split resonances as shown in Fig. 2(b). We set δAin = 0 for the thermally driven system. The optomechanical coupling is assumed to occur only within the first resonator (δx2 = 0). Then Eq. (7) lead to a modified equation for the mechanical oscillation

[(Ωm12Ω2iΓm1Ω)+ig12|a¯1|2m1ω[f+(Δ¯,Ω)+f(Δ¯,Ω)]]δx˜1(Ω)=F˜th(Ω)m1
where the cavity factors f±(Δ̄, Ω) are defined as
f+(Δ¯,Ω)=[i(Δ¯+Ω)Γ1/2+κ2i(Δ¯+Ω)Γ2/2]1f(Δ¯,Ω)=[i(Δ¯Ω)+Γ1/2+κ2i(Δ¯Ω)+Γ2/2]1

The dressed mechanical motion exhibits the modified resonance frequency Ω′m1 and damping rate Γ′m1

(Ωm1')2=Ωm12+{ig12|a¯1|2m1ω[f+(Δ¯,Ω)+f(Δ¯,Ω)]}
Γm1'=Γm1{ig12|a¯1|2m1ωΩ[f+(Δ¯,Ω)+f(Δ¯,Ω)]}

The results are similar to the previous report [1] except for different expressions of the cavity factors. If the frequency spacing between two optical resonances equals the mechanical resonance frequency, the input light and one of its sidebands can be simultaneously on resonance for Δ̄ ≈ ±Ωm/2. We plot the modified mechanical damping rate and resonance frequency in Fig. 3 for κ = 0.5Ωm. For the single resonator system (Fig. 3(a) and 3(c)), the modification of the mechanical damping rate is maximized at Δ̄ = ±Ωm, and the maximum optical spring effect is obtained around Δ̄ ≈ ±0.4Ωm. While for the coupled resonator system (Fig. 3(b) and 3(d)), both the mechanical damping modification and optical spring effect are maximized when the input light is on resonance with one of the split resonances. The results agree with previous analysis using the sideband picture. Because of the enhancement of both the input light and its sideband, the modifications of the damping rate and resonance frequency for the mechanical oscillation increase by a percentage of about 100% and 75%, respectively.

 figure: Fig. 3

Fig. 3 Calculated (a)(b) optomechanical amplification (Γ′m1 < Γm1) or cooling (Γ′m1 > Γm1) and (c)(d) optical spring effect for (1) single optical resonator with Qe = Qi = 106, Ωm = 2π × 500MHz; (2) coupled resonator with Qe1 = Qi1 = Qi2 = 106, κ = 0.5Ωm. The blue curves denote the normalized optical energy inside the resonators. The input power is 1mW. The optomechanical coupling coefficient is gi = 10GHz/nm. The effective mass and Q-factor of the mechanical oscillator are set at mi = 10pg and Qm=1000, respectively.

Download Full Size | PDF

3.2. Strong coupling under external modulation driven

For the external modulation driven system, the Optomechanically Induced Transparency (OMIT) effect has been reported for the single resonator optomechanical system, where a control light induces a narrow transparency window for a weak, probe light [3, 21]. To investigate the OMIT effect in the coupled resonator system, the control light (Ain) is set at one of the resonances with smaller frequency. A frequency modulation of the control light generates two sidebands, one of which serves as the probe light. The modulation frequency is then scanned to obtain the transmission spectrum of the probe light. We also assume that the mechanical oscillation only occurs in the first resonator (δx2 = 0). Because the back-action force generated by the modulated optical field is much larger than the thermal force, we can neglect Fth and solve Eq. (7) in the frequency domain:

δa˜1(Ω)=iΓef+(Ω){δA˜in(Ω)+ig12ω1χ(Ω)a¯1f(Ω)[a¯1*δA˜in(Ω)a¯1δA˜in*(Ω)]}1+ig12ω1χ(Ω)|a¯1|2[f+(Ω)+f(Ω)]
where δÃin is the power amplitude of the probe light, δã1(Ω) is the optical field amplitude at the probe frequency inside the first resonator, and the tilde symbol denotes the Fourier transform of optical fields. The cavity factors f±(Δ̄, Ω) are defined in Eq. (9). The mechanical susceptibility χ(Ω) is defined as χ(Ω)=[m(Ωm2Ω2iΓmΩ)]1.

From Eq. (12), the transmission of the probe light can be calculated as a function of its frequency detuning from the control light. When the detuning gets close to the mechanical frequency, the intra-cavity field of the probe light experiences destructive interference with the anti-Stokes sideband of the control light generated by the mechanical oscillation. Therefore, a narrow transparency window arises, as shown in Fig. 4(a), corresponding to the OMIT effect [3, 21]. If the mechanical frequency is much larger than the linewidth of the split resonances, one of the cavity factors is much smaller than the other and can be neglected consequently, leading to a similar expression as derived in [3]. In Fig. 4(b), we calculate the OMIT window peak transmission with different control light powers. The peak transmission is normalized as t = (tptr)/(1 − tr) for both the single resonator system and the coupled resonator system, where tp is the peak value of the transmission of the transparency window, and tr is the residue transmission in the absence of the control light. Because of the resonance-enhancement of the control light, the pump power efficiency is greatly increased for generating a transparency window in the coupled resonator optomechanical system.

 figure: Fig. 4

Fig. 4 (a) Normalized power transmission in the coupled resonator system for OMIT effect. The control light is on resonance with one of the optical resonances with smaller frequency. The frequency detuning of the probe light from the control light is scanned to obtain its transmission spectrum. The input power of control and probe light are 1mW and 10μW, respectively. Qi1 = Qi2 = Qe1 = 106, m1 = 10pg, g1 = 20GHz/nm, Qm = 1000, Ωm = 2π × 500MHz, κ = 0.5Ωm. (b) Comparison of the normalized peak transmission for OMIT effect in a single resonator and coupled resonator system. The single resonator system has the same optical decay rates and properties of mechanical oscillation as the first resonator in the coupled resonator system.

Download Full Size | PDF

3.3. Weak coupling under thermal force driven

In the weak coupling regime, when the frequency of the input light is scanned near the intrinsic resonance, the optical energy inside of the first (second) resonator decreases (increases) when reducing the frequency detuning, as shown in Fig. 2(a). This indicates that the dynamic back action force will cool the mechanical mode of one resonator but simultaneously amplify the mechanical mode of the other one (here we consider that there is a mechanical mode coupled to the optical resonance in both resonators).

We assume that both optical resonators are perturbed by mechanical motions with harmonic functions, i.e. δxi(t) = δxi cos(Ωmit). The perturbation of optical fields inside each resonator can be calculated by solving Eq. (7). For simplicity, we assume the mechanical modes have the same resonance frequency Ωm

dδa1dt=(iΔ1Γ1)δa1+iκδa2ig1x1cos(Ωmt)a¯1
dδa2dt=(iΔ2Γ2)δa2+iκδa1ig2x2cos(Ωmt)a¯2
Rewrite δai=δai+eiΩmt+δaieiΩmt and plug δai back into the equations above. We obtain
δa1±=12ig1x1a¯1[i(Δ2Ωm)Γ2/2]+g2x2a¯2κ[i(Δ1Ωm)Γ1/2][i(Δ2Ωm)Γ2/2]+κ2
δa2±=12ig2x2a¯2[i(Δ1Ωm)Γ1/2]+g1x1a¯1κ[i(Δ1Ωm)Γ1/2][i(Δ2Ωm)Γ2/2]+κ2

For small perturbation, the dynamic back-action force generated by the optical field is equal to F = −2g𝔕(ā*δai)/ω, and contains two components:

FIi=2giωi[a¯i*(δai++δai)]
FQi=2giωi[a¯i*(δai+δai)]

The in-phase component FI (proportional to cos(Ωmt)) relates to the optical spring effect that changes the mechanical frequency, while the quadrature component FQ (proportional to sin(Ωmt)) relates to the change of damping rate of the mechanical motion. The net power transfered from the optical field to the mechanical motion equals 〈P〉 = 〈Fẋ(t)〉 = −FQδxsin(Ωmt). Therefore, positive (negative) FQ corresponds to a cooling (amplifying) process of the mechanical modes. During the calculation, the properties of the mechanical modes are set at mi = 10pg, gi = 10GHz/nm, Qmi = 1000, Ωmi = 2π × 200MHz, and we assume the two resonators have the same intrinsic resonance, Δ1 = Δ2.

First, we consider that the loaded Q-factor of the first resonator is much smaller than the second resonator. Fig. 5(a) shows the optical energy inside each resonator depending on the normalized detuning of the input light. The calculated FQi are shown in Fig. 5(b). The red and blue curves represent the quadrature component of the force inside the first and second resonator, respectively. For red (blue) detuning, the optical fields amplify (cool) the mechanical mode in the first resonator and cool (amplify) the mechanical mode in the second resonator. We further compare the results to the cases where only one mechanical motion exists in the system, as shown in Fig. 5(b). The optical force in the first resonator (FQ1) is greatly enhanced with the existence of the mechanical motion (δx2) in the second resonator, while FQ2 changes very little with or without δx1. In other words, because most of the optical energy accumulates in the second resonator around the resonance, δx1(t) is controlled by δx2(t) via the optomechanical interaction.

 figure: Fig. 5

Fig. 5 (a) The optical energy inside the first (red) and second (blue) resonator. (b) The calculated quadrature force component generated by optical field in the first (red) and second (blue) resonator. The solid lines are calculated when mechanical oscillation exists in both resonators. The dotted lines are calculated when only one resonator has mechanical motion. The Q-factors of the two resonators are Qi1 = Qi2 = 106, Qe = 105, Qκ = 106.

Download Full Size | PDF

When the loaded Q-factor of the two resonators are similar, the maximum optical energies in both resonators are on the same level near the resonance (Fig. 6(a)). Thus, the optical forces in both resonators are influenced with the existence of mechanical motion of the other, as shown in Fig. 6(b). When the input light is red detuned from the resonance, both the optical field and the mechanical motion δx1(t) generate additional damping on δx2(t). In Fig. 6(b), we also compare the calculated backaction forces with the ones where only one mechanical mode exists. The result indicates that the optical field transfers part of the energy in δx2(t) to δx1(t) when Δ̄ ≈ −Ωm. The energy transfer process is reversed for blue detuning. Although the two mechanical modes do not have direct coupling, the input optical field serves as an optical spring that facilitates the energy transfer process. With proper optimization, this phenomenon can be used to manipulate phonon energy among different mechanical oscillators, and provides the potential for employing optomechanical systems for signal storage and processing.

 figure: Fig. 6

Fig. 6 (a) The optical energy and (b) the calculated quadrature force component generated by optical field in the first (red) and second (blue) resonator. The Q-factors of the two resonators are: Qi1 = Qκ = 106, Qi2 = Qe = 5 × 105,.

Download Full Size | PDF

To form a complete analysis, we also investigate the coupled resonator system under external modulation in the weak coupling regime. The external modulation is usually used to drive the mechanical motion or to generate the OMIT effect. However, due to the lack of mode splitting in the weak coupling regime, the coupled resonator system does not exhibit noticeable advantages over a single resonator with respect to the enhancement of optomechanical force or power efficiency of OMIT effect.

4. Summary

Based on the calculation of the dynamic back-action in two coupled resonators, we show that building optomechanical systems based on the compound resonator system can control and enhance the optomechanical interaction. For the modulation driven system, we show that the power efficiency of the control light to obtain OMIT effect can be greatly increased. For the thermally driven system, the optomechanical interaction is enhanced compared to the single resonator system. With properly choosing the parameters of the coupled optomechanical system, an optically controlled mechanical energy transfer between different mechanical oscillators is possible.

References and links

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

2. M. Li, W. H. P. Pernice, and H. X. Tang, “Reactive cavity optical force on microdisk-coupled nanomechanical beam waveguides,” Phys. Lett. 103, 223901 (2009). [CrossRef]  

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

4. W. H. P. Pernice, M. Li, and H. X. Tang, “Optomechanical coupling in photonic crystal supported nanomechanical waveguides,” Opt. Express 17, 12424–12432 (2009). [CrossRef]   [PubMed]  

5. D. V. Thourhout and J. Roels, “Optomechanical device actuation through the optical gradient force,” Nature Photon. 4, 211–217 (2010). [CrossRef]  

6. F. Marquardt and S. Girvin, “Optomechanics,” Physics 2, 40 (2009). [CrossRef]  

7. S. Johnson, M. Ibanescu, M. Skorobogatiy, O. Weisberg, J. Joannopoulos, and Y. Fink, “Perturbation theory for Maxwells equations with shifting material boundaries,” Phys. Rev. E 65, 066611 (2002). [CrossRef]  

8. M. L. Povinelli, M. Lončar, M. Ibanescu, E. J. Smythe, S. G. Johnson, F. Capasso, and J. D. Joannopoulos, “Evanescent-wave bonding between optical waveguides,” Opt. Lett. 30, 3042–3044 (2005). [CrossRef]   [PubMed]  

9. D. G. Grier, “A revolution in optical manipulation,” Nature 424, 810–816 (2003). [CrossRef]   [PubMed]  

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

11. A. Xuereb, T. Freegarde, P. Horak, and P. Domokos, “Optomechanical cooling with generalized interferometers,” Phys. Rev. Lett. 105, 013602 (2010). [CrossRef]   [PubMed]  

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

13. Q. Lin, J. Rosenberg, X. Jiang, K. Vahala, and O. Painter, “Mechanical oscillation and cooling actuated by the optical gradient force,” Phys. Rev. Lett. 103, 103601 (2009). [CrossRef]   [PubMed]  

14. Y.-S. Park and H. Wang, “Resolved-sideband and cryogenic cooling of an optomechanical resonator,” Nature Phys. 5, 489–493 (2009). [CrossRef]  

15. I. S. Grudinin, H. Lee, O. Painter, and K. J. Vahala, “Phonon laser action in a tunable two-level system,” Phys. Rev. Lett. 104, 083901 (2010). [CrossRef]   [PubMed]  

16. A. Schliesser, P. DelHaye, N. Nooshi, K. Vahala, and T. Kippenberg, “Radiation pressure cooling of a micromechanical oscillator using dynamical backaction,” Phys. Rev. Lett. 97, 243905 (2006). [CrossRef]  

17. M. Eichenfield, C. P. Michael, R. Perahia, and O. Painter, “Actuation of micro-optomechanical systems via cavity-enhanced optical dipole forces,” Nature Photon. 1, 416–422 (2007). [CrossRef]  

18. A. H. Safavi-Naeini and O. Painter, “Proposal for an optomechanical traveling wave phononphoton translator,” New J. Phys. 13, 013017 (2011). [CrossRef]  

19. M. Eichenfield, R. Camacho, J. Chan, K. J. Vahala, and O. Painter, “A picogram- and nanometre-scale photonic-crystal optomechanical cavity,” Nature 459, 550–555 (2009). [CrossRef]   [PubMed]  

20. M. Eichenfield, J. Chan, R. M. Camacho, K. J. Vahala, and O. Painter, “Optomechanical crystals,” Nature 462, 78–82 (2009). [CrossRef]   [PubMed]  

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

22. M. Li, W. H. P. Pernice, C. Xiong, T. Baehr-Jones, M. Hochberg, and H. X. Tang, “Harnessing optical forces in integrated photonic circuits,” Nature 456, 480–484 (2008). [CrossRef]   [PubMed]  

23. X. Zhao, J. M. Tsai, H. Cai, X. M. Ji, J. Zhou, M. H. Bao, Y. P. Huang, D. L. Kwong, and A. Q. Liu, “A nano-opto-mechanical pressure sensor via ring resonator,” Opt. Express 20, 8535–8542 (2012). [CrossRef]   [PubMed]  

24. M. Hossein-Zadeh and K. J. Vahala, “An optomechanical oscillator on a silicon chip,” IEEE J. Sel. Topics Quantum Electron. 16, 276–287 (2010). [CrossRef]  

25. X. Sun, K. Y. Fong, C. Xiong, W. H. P. Pernice, and H. X. Tang, “GHz optomechanical resonators with high mechanical Q factor in air,” Opt. Express 19, 22316–22321 (2011). [CrossRef]   [PubMed]  

26. X. Sun, X. Zhang, and H. X. Tang, “High-Q silicon optomechanical microdisk resonators at gigahertz frequencies,” Appl. Phys. Lett. 100, 173116 (2012) [CrossRef]  

27. Q. Lin, J. Rosenberg, D. Chang, R. Camacho, M. Eichenfield, K. J. Vahala, and O. Painter, “Coherent mixing of mechanical excitations in nano-optomechanical structures,” Nature Photon. 4, 236–242 (2010). [CrossRef]  

28. G. S. Wiederhecker, L. Chen, A. Gondarenko, and M. Lipson, “Controlling photonic structures using optical forces,” Nature 462, 633–637 (2009). [CrossRef]   [PubMed]  

29. S. Manipatruni, G. Weiderhecker, and M. Lipson, “Long-range synchronization of optomechanical structures,” OSA/CLEO (2011).

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 (6)

Fig. 1
Fig. 1 The schematics and sideband picture of the optomechanical system based on (a) a single resonator; (b) two coupled resonators. The mechanical motion modifies the optical resonance frequency.
Fig. 2
Fig. 2 Optical field energies of the two resonators in the: (a) weak coupling regime, κ = Γ2; (b) strong coupling regime, κ = 10Γ2. The intrinsic Q-factors are Qi1 = Qi2 = 106, and the external Q-factor of the first resonator is Qe1 = 105. The input optical power is 1mW. For simplicity, the two resonators are assumed to have the same intrinsic resonance, Δ = Δ1 = Δ2.
Fig. 3
Fig. 3 Calculated (a)(b) optomechanical amplification (Γ′m1 < Γm1) or cooling (Γ′m1 > Γm1) and (c)(d) optical spring effect for (1) single optical resonator with Qe = Qi = 106, Ωm = 2π × 500MHz; (2) coupled resonator with Qe1 = Qi1 = Qi2 = 106, κ = 0.5Ωm. The blue curves denote the normalized optical energy inside the resonators. The input power is 1mW. The optomechanical coupling coefficient is gi = 10GHz/nm. The effective mass and Q-factor of the mechanical oscillator are set at mi = 10pg and Qm=1000, respectively.
Fig. 4
Fig. 4 (a) Normalized power transmission in the coupled resonator system for OMIT effect. The control light is on resonance with one of the optical resonances with smaller frequency. The frequency detuning of the probe light from the control light is scanned to obtain its transmission spectrum. The input power of control and probe light are 1mW and 10μW, respectively. Qi1 = Qi2 = Qe1 = 106, m1 = 10pg, g1 = 20GHz/nm, Qm = 1000, Ωm = 2π × 500MHz, κ = 0.5Ωm. (b) Comparison of the normalized peak transmission for OMIT effect in a single resonator and coupled resonator system. The single resonator system has the same optical decay rates and properties of mechanical oscillation as the first resonator in the coupled resonator system.
Fig. 5
Fig. 5 (a) The optical energy inside the first (red) and second (blue) resonator. (b) The calculated quadrature force component generated by optical field in the first (red) and second (blue) resonator. The solid lines are calculated when mechanical oscillation exists in both resonators. The dotted lines are calculated when only one resonator has mechanical motion. The Q-factors of the two resonators are Qi1 = Qi2 = 106, Qe = 105, Qκ = 106.
Fig. 6
Fig. 6 (a) The optical energy and (b) the calculated quadrature force component generated by optical field in the first (red) and second (blue) resonator. The Q-factors of the two resonators are: Qi1 = Qκ = 106, Qi2 = Qe = 5 × 105,.

Equations (18)

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

d a 1 d t = ( i Δ 1 Γ 1 / 2 i g 1 x 1 ) a 1 + i Γ e A in + i κ a 2
d a 2 d t = ( i Δ 2 Γ 2 / 2 i g 2 x 2 ) a 2 + i κ a 1
d 2 x i d t 2 + Γ m i d x i d t + Ω m i 2 x i = F i m i + F th m i
a ¯ 1 = i Γ e A in i Δ ¯ 1 Γ 1 + κ 2 i Δ ¯ 2 Γ 2 / 2
a ¯ 2 = i κ a ¯ 1 i Δ ¯ 2 Γ 2 / 2
x ¯ i = g i | a ¯ i | 2 m i Ω m i 2 ω
d δ a 1 d t = ( i Δ ¯ Γ 1 2 ) δ a 1 i g 1 a ¯ 1 δ x 1 + i κ δ a 2 + i Γ e δ A in d δ a 2 d t = ( i Δ ¯ Γ 2 2 ) δ a 2 i g 2 a ¯ 2 δ x 2 + i κ δ a 1 d 2 δ x i d t 2 + Γ m i d δ x i d t + Ω m i 2 δ x i = g i m i ω ( a ¯ i δ a i * + a ¯ i * δ a i ) + F th m i
[ ( Ω m 1 2 Ω 2 i Γ m 1 Ω ) + i g 1 2 | a ¯ 1 | 2 m 1 ω [ f + ( Δ ¯ , Ω ) + f ( Δ ¯ , Ω ) ] ] δ x ˜ 1 ( Ω ) = F ˜ th ( Ω ) m 1
f + ( Δ ¯ , Ω ) = [ i ( Δ ¯ + Ω ) Γ 1 / 2 + κ 2 i ( Δ ¯ + Ω ) Γ 2 / 2 ] 1 f ( Δ ¯ , Ω ) = [ i ( Δ ¯ Ω ) + Γ 1 / 2 + κ 2 i ( Δ ¯ Ω ) + Γ 2 / 2 ] 1
( Ω m 1 ' ) 2 = Ω m 1 2 + { i g 1 2 | a ¯ 1 | 2 m 1 ω [ f + ( Δ ¯ , Ω ) + f ( Δ ¯ , Ω ) ] }
Γ m 1 ' = Γ m 1 { i g 1 2 | a ¯ 1 | 2 m 1 ω Ω [ f + ( Δ ¯ , Ω ) + f ( Δ ¯ , Ω ) ] }
δ a ˜ 1 ( Ω ) = i Γ e f + ( Ω ) { δ A ˜ in ( Ω ) + i g 1 2 ω 1 χ ( Ω ) a ¯ 1 f ( Ω ) [ a ¯ 1 * δ A ˜ in ( Ω ) a ¯ 1 δ A ˜ in * ( Ω ) ] } 1 + i g 1 2 ω 1 χ ( Ω ) | a ¯ 1 | 2 [ f + ( Ω ) + f ( Ω ) ]
d δ a 1 d t = ( i Δ 1 Γ 1 ) δ a 1 + i κ δ a 2 i g 1 x 1 cos ( Ω m t ) a ¯ 1
d δ a 2 d t = ( i Δ 2 Γ 2 ) δ a 2 + i κ δ a 1 i g 2 x 2 cos ( Ω m t ) a ¯ 2
δ a 1 ± = 1 2 i g 1 x 1 a ¯ 1 [ i ( Δ 2 Ω m ) Γ 2 / 2 ] + g 2 x 2 a ¯ 2 κ [ i ( Δ 1 Ω m ) Γ 1 / 2 ] [ i ( Δ 2 Ω m ) Γ 2 / 2 ] + κ 2
δ a 2 ± = 1 2 i g 2 x 2 a ¯ 2 [ i ( Δ 1 Ω m ) Γ 1 / 2 ] + g 1 x 1 a ¯ 1 κ [ i ( Δ 1 Ω m ) Γ 1 / 2 ] [ i ( Δ 2 Ω m ) Γ 2 / 2 ] + κ 2
F I i = 2 g i ω i [ a ¯ i * ( δ a i + + δ a i ) ]
F Q i = 2 g i ω i [ a ¯ i * ( δ a i + δ a i ) ]
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.