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

Manipulation of optomechanically induced transparency and absorption by indirectly coupling to an auxiliary cavity mode

Open Access Open Access

Abstract

We theoretically study the optomechanically induced transparency (OMIT) and absorption (OMIA) phenomena in a single microcavity optomechanical system, assisted by an indirectly coupled auxiliary cavity mode. We show that the interference effect between the two optical modes plays an important role and can be used to control the multiple-pathway induced destructive or constructive interference effect. The three-pathway interference could induce an absorption dip within the transparent window in the red sideband driving regime, while we can switch back and forth between OMIT and OMIA with the four-pathway interference. The conversion between the transparency peak and absorption dip can be achieved by tuning the relative amplitude and phase of the multiple light paths interference. Our system proposes a new platform to realize multiple pathways induced transparency and absorption in a single microcavity and a feasible way for realizing all-optical information processing.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

In the past decades, electromagnetically induced transparency (EIT) has been studied both theoretically and experimentally [13]. And the potential applications range from ultraslow light propagation [2], quantum information storage [4,5] to the enhancement of nonlinear processes [6,7]. Normally, EIT is a coherent phenomenon due to the destructive quantum interference of two excitation pathways in a three-level system. On the other hand, the electromagnetically induced absorption (EIA) [8,9] is the result of constructive interference between different pathways. Recently, the EIT effects were widely studied in the optical microresonators [1013], i.e. the so called EIT-like effect [1427]. Usually, this effect can be generated through a high quality factor (Q) cavity mode directly or indirectly coupling to a low Q one. In this case, a sharp transparent window at their original resonant frequency region appears when the two modes are frequency overlapped.

Based on the strong optical and mechanical interaction, optomechanical systems have potential applications in the fundamental research [2836] and provide a promising platform for exploring quantum nonlinear phenomena [34,37,38], such as ground state cooling [3946], entanglement in cavity optomechanical system [4749]. Meanwhile, the cavity optomechanics could enable the exploration of a variety of optical processes. The system has been studied and applied in OMIT [5060], OMIA [53,61,62], multimode optomechanical interactions [6365], quantum information processing [6668] and amplification [6971].

Analog to the EIT (EIA) in the atomic system, the OMIT (OMIA) effect is the result of destructive (constructive) interference between different pathways in optomechanical system. The OMIT effect has been investigated both theoretically and experimentally [5057,72,73]. Compared with OMIT, the OMIA system generally consists of more degrees of freedom or more than two transition pathways for the probe [13,61,7481]. The OMIT could be switched to OMIA through multiple-pathway interference effects, accomplished by coupling to an additional microcavity or adding more mechanical resonators. However, these approaches make the system more complicated and fragile to control.

In this paper, we study the optomechanical system and multiple-pathway interference in a single microcavity, in which one optical cavity mode indirectly couples to an auxiliary cavity mode. As shown schematically in Fig. 1, the optomechanical system contains two optical modes and one mechanical mode. The auxiliary cavity mode $a_2$ is adjacent to the dominant optical mode $a_1$ in frequency domain and can be tuned to overlap with $a_1$ [17,82]. Different from the directly coupled cavity case [74,77], the two optical modes here have no direct coupling. The interference effect between the two optical modes are mediated by the tapered fiber and could induce the EIT-like phenomenon without the optomechanical interaction. Combining the EIT-like and OMIT effect, we can easily realize multiple-pathway interference. When the auxiliary cavity mode decouples to the phonons, the three-pathway interference will induce constructive interference effect. We find in the red sideband driving regime, the transmission rate of the system is sensitive to the auxiliary cavity mode. While the auxiliary cavity mode couples to the mechanical mode, the four-pathway interference could induce destructive or constructive interference effect. And the bi-directional transformations between OMIT and OMIA can be achieved. We also study the transmission of the three-pathway interference case in the blue sideband regime. The optical amplification mainly depends on the optomechanical gain and could be enhanced by the auxiliary cavity mode. Therefore, the output spectrum can be precisely controlled by tuning the auxiliary cavity mode and the optomechanical coupling strength.

 figure: Fig. 1.

Fig. 1. (a) Schematic of the optomechanical system consisting of two optical modes and a mechanical resonator. The two optical modes are excited by the waveguide simultaneously. (b) The auxiliary cavity mode $a_2$ is adjacent to the dominant optical mode $a_1$ in frequency domain. The optical mode $a_2$ can be tuned to overlap with $a_1$, which could lead to electromagnetically-induced transparency-like window [17]. (c) Illustration of the three-pathway interference effect when the auxiliary cavity mode decouples with the mechanical mode. (d) Illustration of the four-pathway interference effect when both optical modes couples to the mechanical mode.

Download Full Size | PDF

2. Model

As schematically shown in Fig. 1, the system we considered here is a whispering-gallery-mode optomechanical microresonator, where two optical modes and one mechanical mode have been excited. The dominant optical mode $a_1$ has the resonant frequency $\omega _1$ and the internal loss $\kappa _{10}$. The auxiliary cavity mode $a_2$ has the resonant frequency $\omega _2$ and the internal loss $\kappa _{20}$. The mechanical resonator has the frequency $\omega _m$ with the effective mass $m$. The optical modes $a_1$ and $a_2$ are two independent WGMs. However, the cavity mode $a_1$ can couple to the mode $a_2$ indirectly through the waveguide [17]. The interference between the two optical pathways results in EIT-like line shape. Both optical modes could couple to the mechanical resonator with different optomechanical coupling rate $g_1$ and $g_2$. When the optomechanical coupling $g_2$ can be ignored, the system shows three pathways interference. Without loss of generality, the system shows four pathways interference when both optical modes couple to the mechanical mode. The Hamiltonian can be described as

$$\begin{aligned} H &= \hbar \omega_1 a^\dagger_1 a_1+\hbar\omega_2 a^\dagger_2 a_2+\frac{1}{2}m\omega^2_m x^2+\frac{P^2}{2m}\\ & +\int_{-\infty}^{+\infty} \hbar \omega c^\dagger(\omega)c(\omega)\, d\omega+\hbar g_1 x a^\dagger_1 a_1 +\hbar g_2 x a^\dagger_2 a_2\\ & +i\hbar \int_{-\infty}^{+\infty} \sum_{j=1,2} \kappa_{ex,j}(\omega)[c^\dagger(\omega)a_j-a^\dagger_j c(\omega)]\,d\omega\end{aligned}$$
where the operators $x$ and $P$ represent the position and momentum of the mechanical mode, respectively. The annihilation operator $c(\omega )$ indicates the waveguide mode, which satisfies $[c(\omega ),c^\dagger (\omega ^{'}) ]=\delta (\omega -\omega ^{'})$. $\kappa _{ex,j}(\omega )$ (j=1,2) describes the coupling constant between the cavity mode $a_j$ and waveguide modes. The first four terms represent the free Hamiltonians of the optical and mechanical modes. The fifth term is to describe the Hamiltonian of the waveguide modes. And the last term indicates the coupling between the optical modes and the waveguide modes. In our scheme, the system is driven by a strong control laser field with the amplitudes $\varepsilon _p$ and frequency $\omega _p$, respectively. Meanwhile, a weak probe laser field, with the amplitude $\varepsilon _{s}$ and the frequency $\omega _s$, is applied on the system. The dynamics of the optomechanical system can thus be described in the rotating frame at the pump frequency $\omega _p$
$$\frac{\mathrm{d}a_1}{\mathrm{d}t} = i \Delta_1 a_1 -\frac{\kappa_1}{2}a_1 - i g_1 x a_1-\frac{\sqrt{\kappa_{ex1}\kappa_{ex2}}}{2} a_2+\sqrt{\kappa_{ex1}} \varepsilon_p +\sqrt{\kappa_{ex1}} \varepsilon_s e^{-i\delta t}+\sqrt{\kappa_{ex1}}\xi_1$$
$$\frac{\mathrm{d}a_2}{\mathrm{d}t} = i \Delta_2 a_2 -\frac{\kappa_2}{2}a_2 - i g_2 x a_2-\frac{\sqrt{\kappa_{ex1}\kappa_{ex2}}}{2} a_1+\sqrt{\kappa_{ex2}} \varepsilon_p +\sqrt{\kappa_{ex2}} \varepsilon_s e^{-i\delta t} +\sqrt{\kappa_{ex2}}\xi_2$$
$$\frac{\mathrm{d}x}{\mathrm{d}t} =\frac{P}{m}$$
$$\frac{\mathrm{d}P}{\mathrm{d}t} = -m \omega^2_m x -\hbar( g_1 a^\dagger_1a_1+ g_2 a^\dagger_2a_2) -\frac{\Gamma_m}{2} P +\delta F$$
where $\Delta _1 = \omega _p-\omega _1$ and $\Delta _2 = \omega _p-\omega _2$ represent the detuings of optical modes with respect to the driving field. $\kappa _j = \kappa _{j0}+ \kappa _{exj}$ indicates the total decay rate of optical mode $a_j$ (j=1,2). $\delta = \omega _s-\omega _p$ is the frequency detuning between the probe field and the control field. And $\xi _1$ ($\xi _2$) is the external noise of the optical mode $a_1$ ($a_2$) introduced by the waveguide. $\delta F$ is the thermal noise of the mechanical oscillator and $\Gamma _m$ is the decay rate of the mechanical mode. In our case, the pump laser field is much stronger than the probe field. By using the linearization approach, thus the Heisenberg operators can be divided into the steady parts and the fluctuation ones, i.e., $a_j = \overline {a}_j + \delta a_j$ (j = 1,2) and $x = \overline {x} + \delta x$. Substituting the division forms into Eqs. (2)–(5), the steady solutions of the above dynamical equations can be obtained as
$$\overline{a}_1 = \frac{-\sqrt{\kappa_{ex1}}-\frac{\sqrt{\kappa_{ex1}}\kappa_{ex2}}{2(i\Delta_2-ig_2\overline{x}-\kappa_2/2)}}{i\Delta_1-ig_1\overline{x}-\kappa_1/2-\frac{\kappa_{ex1}\kappa_{ex2}}{4(i\Delta_2-ig_2\overline{x}-\kappa_2/2)}}\varepsilon_p$$
$$\overline{a}_2 = \frac{\frac{ \sqrt{\kappa_{ex1}\kappa_{ex2}}}{2}\overline{a}_1 -\sqrt{\kappa_{ex2}}\varepsilon_p}{i\Delta_2-ig_2\overline{x}-\kappa_2/2}$$
$$\overline{x} = -\hbar\frac{g_1|\overline{a}_1|^2+g_2|\overline{a}_2|^2}{m\omega^2_m}$$
Then, we only keep the first-order terms in the small fluctuation ones $\delta a_1$, $\delta a^\dagger _1$, $\delta a_2$, $\delta a^\dagger _2$ and $\delta x$. Under these conditions, we can obtain the linearized Langevin equations as follows:
$$\frac{\mathrm{d}\delta a_1}{\mathrm{d}t} = (i\overline{\Delta}_1-\frac{\kappa_1}{2}) \delta a_1-ig_1\overline{a}_1 \delta x-\frac{\sqrt{\kappa_{ex1}\kappa_{ex2}}}{2} \delta a_2+\sqrt{\kappa_{ex1}}\varepsilon_s e^{-i\delta t}+\sqrt{\kappa_{ex1}}\xi_1$$
$$\frac{\mathrm{d}\delta a_2}{\mathrm{d}t} = (i\overline{\Delta}_2-\frac{\kappa_2}{2}) \delta a_2-ig_2\overline{a}_2 \delta x-\frac{\sqrt{\kappa_{ex1}\kappa_{ex2}}}{2} \delta a_1+\sqrt{\kappa_{ex2}}\varepsilon_s e^{-i\delta t}+\sqrt{\kappa_{ex2}}\xi_2$$
$$m\frac{\mathrm{d^2}}{\mathrm{d}t^2}\delta x = -\frac{m\Gamma_m}{2} \frac{\mathrm{d}}{\mathrm{d}t} \delta x-m\omega^2_m \delta x-\hbar g_1\overline{a}_1 (\delta a_1+\delta a^\dagger_1)-\hbar g_2\overline{a}_2 (\delta a_2+\delta a^\dagger_2)+\delta F$$
where $\overline {\Delta }_1=\Delta _1-g_1\overline {x}$ and $\overline {\Delta }_2=\Delta _2-g_2\overline {x}$ denote the effective detuning between the cavity modes and the control laser beam, including the frequency shift caused by the mechanical motion.

Combining the above analysis and Figs. 1(c) and 1(d), we can better understand the multiple light paths for interference. Figure 1 displays a schematic of quantum interference between different light paths. The photons at the output port could come from four different pathways. The different paths are: ① the probe photons excite the cavity mode $a_1$ and pass to the output port; ② the photons generated by the sideband transition through the optomechanical interaction in mode $a_1$ are coupled out the cavity through the waveguide; ③ the probe field passes through the cavity mode $a_2$ directly. In our case, the system only shows three different pathways when the optomechanical coupling rate $g_2$ can be neglected. Otherwise, the fourth pathway is: ④ the photons genetated by anti-Stokes process in the optical mode $a_2$ are coupled out the cavity through the waveguide. Thus the photons in the output beam is the sum of four different paths. For convenience, we introduce $x=x_{ZPF}(b+b^\dagger )$ and $P=-i m \omega _m x_{ZPF} (b-b^\dagger )$ to get the transmission of the system, where $x_{ZPF}$ is the zero point fluctuation. We also neglect the quantum noise $\xi _1$, $\xi _2$ and $\delta F$. Without loss of generality, in the following discussions the frequency of the auxiliary cavity mode is tuned to satisfy $\overline {\Delta }_1 =\overline {\Delta }_2 =\Delta$. According to the input-output theory [83], we can get the output of the probe field: $a_{out} = \varepsilon _s-\sqrt {\kappa _{ex1}} a_1-\sqrt {\kappa _{ex2}} a_2$. With neglecting the high order sideband effect, the normalized transmission coefficient in the red sideband driving regime can be simplified to

$$t_r = \frac{(L_1-\kappa_{ex1})(L_2-\kappa_{ex2})-(A-\sqrt{\kappa_{ex1}\kappa_{ex2}})^2}{L_2(L_1-\frac{A^2}{L_2})}$$
where $L_1=i(-\Delta -\delta )+\frac {\kappa _1}{2}+\frac {G^2_1}{i(-\Delta -\delta )+\frac {\Gamma _m}{2}}$, $L_2=i(-\Delta -\delta )+\frac {\kappa _2}{2}+\frac {G^2_2}{i(-\Delta -\delta )+\frac {\Gamma _m}{2}}$ and $A=\frac {G_1 G_2 }{i(-\Delta -\delta )+\frac {\Gamma _m}{2}}+\frac {\sqrt {\kappa _{ex1}\kappa _{ex2}}}{2}$. Here $G_1 = g_1 x_{ZPF} |\overline {a}_1|$ and $G_2 = g_2 x_{ZPF} |\overline {a}_2|$ are the effective optomechanical coupling strength, When the system is driven by the blue-detuned pump field, the normalized transmission coefficient is
$$t_b =\frac{(R_1+\kappa_{ex1})(R_2+\kappa_{ex2})-(B-\sqrt{\kappa_{ex1}\kappa_{ex2}})^2}{R_2(R_1-\frac{B^2}{R_2})}$$
where $R_1=i(\Delta +\delta )-\frac {\kappa _1}{2}-\frac {G^2_1}{i(\Delta +\delta )-\frac {\Gamma _m}{2}}$, $R_2=i(\Delta +\delta )-\frac {\kappa _2}{2}-\frac {G^2_2}{i(\Delta +\delta )-\frac {\Gamma _m}{2}}$ and $B=\frac {G_1 G_2 }{i(\Delta +\delta )-\frac {\Gamma _m}{2}}+\frac {\sqrt {\kappa _{ex1}\kappa _{ex2}}}{2}$. And the corresponding power transmission coefficient is given by $T = |t|^2$

3. Transmission rate with three pathways interference

Before we discuss the three-pathway interference effect, the two-pathway induced EIT-like effect is considered. The all-optical analogues of the EIT effect are widely studied in optical resonators systems. And this EIT-like effect contains two coupling mechanism: directly coupled-resonator-induced transparency(DCRIT) [14,15] and indirectly coupled-resonator-induced transparency(ICRIT) [17,19]. By combining the DCRIT and OMIT effect, the three-pathway induced EIT and EIA effects have been studied in different systems [13,74,7880]. For the coupled cavity system, the two cavities should be tuned precisely to couple, requiring the complicated control and fabrication for the experiments. Thus here we study the three-pathway interference effect by combining the ICRIT effect and optomechanical interaction in a single cavity. In this section, we consider the situation that the auxiliary cavity mode $a_2$ decouples to the mechanical mode, i.e., $g_2=0$. We study the power transmission coefficient with the effect of the auxiliary cavity mode both in the red and blue sideband driving regime. We find that the three mode interaction could be tuned by mode $a_2$, which can be used to control the transmission rate.

3.1 Three-pathway interference with red-detuned driving

We have discussed the physical origins of the three pathway interference phenomena in Fig. 1(c). The photons at the output port with frequency equaling to the probe signal come from the three pathways. When one light path is changed, the relative amplitude and phase of the multiple light paths interference are adjusted. This influence of the three-pathway interference effect will change the output spectra. To investigate the influence of path ③ on the three-pathway interference effect, we plot the output spectrum in Fig. 2 with different external coupling rate $\kappa _{ex2}$. Here the parameters are feasible in the experiments [23,55,56]: $\kappa _{10}=5$ MHz, $\kappa _{ex1}=15$ MHz, $\kappa _{20}=40$ MHz, $\Gamma _m=50$ KHz, $\omega _m=100$ MHz. In Fig. 2(a), the transmission shows the OMIT window without the auxiliary mode. This is the result of the two-pathway destructive interference effect. When the optical mode $a_2$ couples with tapered fiber, the transparent peak maintains but becomes really minute plotted by blue line (@ $G_1 / \kappa _{10} =0.7$) in (b). While the absorption dip starts appearing represent by the red solid line with the weaker optomechanical coupling rate $G_1=1.5$ MHz.

 figure: Fig. 2.

Fig. 2. The transmission rate are plotted with the different external coupling strength $\kappa _{ex2}$ of the auxiliary cavity mode $a_2$ in the red sideband regime. The parameters used here are: $\kappa _{10}=5$ MHz, $\kappa _{ex1}=15$ MHz, $\kappa _{20}=40$ MHz, $\Gamma _m=50$ KHz, $\omega _m=100$ MHz and $\Delta = -\omega _m$.

Download Full Size | PDF

Further increasing the external coupling rate $\kappa _{ex2}$ of the auxiliary mode in Fig. 2(c), the output spectra show the evident absorption dips within an EIT-like window for both lines. The power transmission drops to zero due to the constructive interference. By tuning the external coupling rate $\kappa _{ex2}/\kappa _{10}$ to $75$, the tapered fiber mediated interference effect is strong compared with the optomechanical coupling strength. Both the transmission spectrum and the absorption window become wider. While the transmission ($\delta =\omega _m$) increases to 0.33 even with the absorption dip. From Fig. 2, we can obtain that the interference between the two optical modes could lead the transmission rate from OMIT peak to OMIA dip. The reason for the absorption windows is the constructive interference effect induced by the auxiliary cavity mode. Without the auxiliary cavity mode $a_2$, the output field only connects with mode $a_1$ and shows a OMIT peak in Fig. 2(a). When the auxiliary cavity mode is introduced, the photons with frequency equaling to the probe signal come from the optical modes $a_1$ and $a_2$ both. While the anti-stokes process leads to destructive interference for the intracavity field $a_1$, suppressing the photons population of $a_1$. The constructive interference leads an absorption dip and the transmission rate of central dip drops to zero.

Then we study the influence of the optomechanical interaction on the three-pathway interference. The transmission rate is plotted in Fig. 3 with the different optomechanical coupling strength $G_1$. In Figs. 3(a)–3(c), the external coupling strength $\kappa _{ex2}$ is fixed to 60 MHz and $G_2=0$. The transmission displays the EIT-like window with $G_1=0$ plotted by the blue dashed lines. When the optomechanical coupling strength $G_1$ is increased to 0.2 MHz , there is an small absorption window within the EIT-like window in Fig. 3(a). In Fig. 3(c), the transmission of the central dip drops to 0.1 with stronger optomechanical coupling rate $G_1$. It is obvious that with increasing the optomechanical coupling rate, the transmission shows more prominent absorption effect.

 figure: Fig. 3.

Fig. 3. The transmission rate are plotted with the different optomechanical coupling rate between the optical modes and the phonons. The parameters used here are: $\kappa _{ex1}=15$ MHz, $\kappa _{20}=40$ MHz, $\kappa _{ex2}=60$ MHz, $\Gamma _m=50$ KHz, $\omega _m=100$ MHz and $\Delta = -\omega _m$.

Download Full Size | PDF

From the Fig. 2 and Fig. 3, we can obtain that the three-pathway interference effect could vary from the transparent window to an absorption dip in our system. The output spectrum is sensitive to the auxiliary cavity mode in Fig. 2, which can be used to control the three mode interaction. While in Fig. 3, the depth of the absorption dip can be manipulated by the path ②. This system offers two effective methods to tune the optical response. The transmission is plotted in Fig. 4 as the function of the optomechanical coupling rate $G_1$ and the external coupling strength $\kappa _{ex2}$ of the auxiliary cavity mode. It can be obtained that with the weak optomechanical coupling rate, the effect of two optical modes interference becomes more evident and leads to transparency-like window. While the transmission shows OMIT window when $\kappa _{ex2}$ is weak. Thus the transmission can be manipulated through tuning the relative amplitudes and phases of the three-pathway interference.

 figure: Fig. 4.

Fig. 4. The transmission is plotted as the function of the optomechanical coupling rate $G_1$ and the external coupling strength $\kappa _{ex2}$ of the auxiliary cavity mode. The parameters used here are: $\kappa _{10}=5$ MHz, $\kappa _{ex1}=15$ MHz, $\kappa _{20}=40$ MHz, $\Gamma _m=50$ KHz, $\omega _m=100$ MHz and $\Delta = -\omega _m$.

Download Full Size | PDF

3.2 Three-pathway induced amplification with blue-detuned driving

Now we consider the situation that the system is driven by a blue-detuned driving field and the auxiliary optical mode decouples to the mechanical mode. In this scheme we limit our parameters to the stable driving regime. And the stability condition can be obtained by analyzing the Lyapunov exponents [84] of the Jacobian matrix. We have plotted the border between the stable and unstable condition in the inset of Fig. 5(c). In Fig. 5, we have shown the normalized output spectra of optomechanical system with a blue-detuned pump field. In Fig. 5(a), we plot the transmission rate as a function of $\delta$ for different values of external coupling rate $\kappa _{ex2}$. The red solid line indicates the transmission with optomechanical coupling rate $G_1=0.3$ MHz while $\kappa _{ex2}=0$. Without the auxiliary optical mode, it is clear that the transmission line shows an amplified peak. With the enhancement of the two optical mode interference through increasing $\kappa _{ex2}$, the transmission becomes broader and shows the EIT-like window. As the inset of Fig. 5(a) shows, the transmission at the central peak could enhanced with stronger external coupling strength of the auxiliary optical mode.

 figure: Fig. 5.

Fig. 5. The transmission rate are plotted in the blue sideband driving regime when the auxiliary cavity mode decouples to the mechanical mode. (a) The external coupling rate $\kappa _{ex2}$ varies from 0 to $8\kappa _{10}$ when $G_1$ is fixed to 0.3 MHz. (b) Here $\kappa _{ex2}$ is fixed to 60 MHz while the $G_1$ varies from 0 to 0.3 MHz. (c) The transmission of the central peak is presented as the function of the optomechanical coupling strength $G_1$. The parameters used here are: $\kappa _{10}=5$ MHz, $\kappa _{ex1}=20$ MHz, $\kappa _{20}=40$ MHz, $\Delta =\omega _m$, $\Gamma _m=50$ KHz and $G_2 = 0$. The inset of figure (c) shows the border between the stable and unstable area when $\kappa _{ex2}= 60$ MHz.

Download Full Size | PDF

We also plot the transmission amplification effect for different values of the coupling strength $G_1$ in Fig. 5(b). Here $\kappa _{ex2}$ is fixed to 60 MHz, and $G_1$ is increased to 0.3 MHz. The red solid line shows a EIT-like window without optomechanical interaction. It is clear that the transmission rate of central peak is effectively amplified with $G_1=0.3$ MHz compared with red line. In Fig. 5(b), the transmission amplification effect of the probe field becomes stronger with increasing the optomechanical coupling rate $G_1$. Figure 5(c) plots the transmission rate of the central peak as a function of the coupling rate $G_1$ when the probe frequency detuning $\delta =\omega _m$. The blue (red) solid line plots the amplification effect when $\kappa _{ex2}$ equals to 10 (60) MHz. When $G_1$ is small, the fiber mediated two optical interference is strong, the transmission rate depends on $\kappa _{ex2}$. For strong optomechanical coupling rate, the amplification depends on $G_1$, while the transmission can be affected by the auxiliary cavity mode.

By increasing the optomechanical coupling rate $G_1$, the stokes process is enhanced and emits photons and phonons. The photons generated by the optomechanical interaction are degenerate with the probe field, which accelerates the circulating power of $a_1$. The two optical interference is weak compared with the optomechanical coupling. Thus the optomechanical amplification plays a dominant role in the three pathways. The closer to the stability border, the more prominent the optomechanical gain effect is. However, even the central peak is not sensitive to $\kappa _{ex2}$, the amplification effect could be enhanced by the two optical interference.

4. Transmission rate with four-pathway interference effect

In this section, we will consider the situation that the auxiliary cavity mode couples to the mechanical mode. Here we assume that the effective optomechanical coupling strength $G_1$ and $G_2$ can be tuned separately by using the extra tapered fiber to introduce the extra pump field. To investigate the effect of optomechanical coupling rate $G_1$, $G_2$ on the transmission spectrum, we plot the transmission with the different optomechanical couping rates $G_1$ and $G_2$ in Fig. 6 with the red-detuned driving. Under the three pathways situation in red sideband regime, the transmission rates show an absorption window, which have been presented in Fig. 2. Through increasing the optomechanical coupling rate $G_1$, the absorption dip becomes deeper and wider. Different with three pathways induced absorption, four pathways interference effect can switch the transmission spectrum back and forth between OMIT and OMIA depending on the coupling strength $G_1$ and $G_2$. In Fig. 6(a), when $G_2=0$, the transmission of the central dip is 0.13 which has been presented by the blue dashed line. While the absorption dip becomes more evident in the middle of transmission window when $G_1=0.6$ MHz and $G_2=0.2$ MHz. The transmission drops to 0 due to constructive interference effect. By keeping $G_1 =0.6$ MHz unchanged, the transmission rate is plotted in Fig. 6(b) when the coupling rate $G_2$ is enlarged to 1 MHz. In contrast to the absorption dip for the coupling rate $G_2=0.2$ MHz, here the OMIT peak appears for the resonant case, making the transmission go up to 0.75. By increasing the coupling rate $G_2$ to 1.8 MHz, the constructive interference between different pathways results in the OMIA again. To show the manipulation between OMIA and OMIT through tuning the ratio between $G_1$ and $G2$, the height of the central peak is plotted in Fig. 6(c). It is clear that when the optomechanical coupling strength $G_2$ is weak, the four-pathway interference will enhance the constructive interference and lead to the OMIA effect. Further increasing $G_2$, the four pathways will result in the destructive interference and OMIT window. While $G_2$ is larger than 1.5 MHz, the constructive interference occurs, resulting in the absorption window again. In Fig. 6(d), we plot the transmission rate as a function of different coupling rate $G_1$ and $G_2$. We can obtain that the conversion between the constructive and destructive interference can be achieved by tuning the coupling strength of $G_1$ and $G_2$. Compared with $T=0.51$ when the coupling rate $G_1 = G_2 = 0$, transmission can be switched to 0 or 1 due to the OMIA or OMIT effect.

 figure: Fig. 6.

Fig. 6. The output spectra are shown in (a) and (b) as a function of $\delta$ with different $G_2$. The effective optomechanical coupling strength $G_1$ is fixed to 0.6 MHz. The central peak is plotted in (c) as a function of $G_2$. The power transmission coefficient T as a function of $G_2$ and $G_1$ in (d). Here $\kappa _{10}=5$ MHz, $\kappa _{ex1}=20$ MHz, $\kappa _{20}=40$ MHz, $\kappa _{ex2}=60$ MHz, $\Gamma _m=5$ KHz and $\Delta = -\omega _m$.

Download Full Size | PDF

5. Conclusion

In summary, we have explored an optomechanical system with multiple light paths interference effect in a single optical resonator. We give the explicit physical explanations and detailed calculations in this paper. By combining the EIT-like and OMIT effect, the system shows destructive or constructive interference effect under different conditions. The auxiliary cavity mode offers the additional light pathway to control the output spectrum. Through enhancing the optomechanical coupling strength or the external coupling rate of the auxiliary cavity mode, we can tune the optical response of the probe field effectively. Moreover, our model paves an easy way for realization of multiple interference and manipulation of OMIT and OMIA.

Funding

Ministry of Science and Technology of the People's Republic of China (2017YFA0303700); National Natural Science Foundation of China (61727801); Beijing Innovation Center for Future Chip; Key R&D Program of Guangdong Province (2018B030325002).

Disclosures

The authors declare no conflicts of interest.

References

1. J. P. Marangos, “Electromagnetically induced transparency,” J. Mod. Opt. 45(3), 471–503 (1998). [CrossRef]  

2. M. M. Kash, V. A. Sautenkov, A. S. Zibrov, L. Hollberg, G. R. Welch, M. D. Lukin, Y. Rostovtsev, E. S. Fry, and M. O. Scully, “Ultraslow group velocity and enhanced nonlinear optical effects in a coherently driven hot atomic gas,” Phys. Rev. Lett. 82(26), 5229–5232 (1999). [CrossRef]  

3. M. Fleischhauer, A. Imamoglu, and J. P. Marangos, “Electromagnetically induced transparency: Optics in coherent media,” Rev. Mod. Phys. 77(2), 633–673 (2005). [CrossRef]  

4. D. Phillips, A. Fleischhauer, A. Mair, R. Walsworth, and M. D. Lukin, “Storage of light in atomic vapor,” Phys. Rev. Lett. 86(5), 783–786 (2001). [CrossRef]  

5. C. Liu, Z. Dutton, C. H. Behroozi, and L. V. Hau, “Observation of coherent optical information storage in an atomic medium using halted light pulses,” Nature 409(6819), 490–493 (2001). [CrossRef]  

6. S. E. Harris, J. Field, and A. Imamoğlu, “Nonlinear optical processes using electromagnetically induced transparency,” Phys. Rev. Lett. 64(10), 1107–1110 (1990). [CrossRef]  

7. M. Jain, H. Xia, G. Yin, A. Merriam, and S. Harris, “Efficient nonlinear frequency conversion with maximal atomic coherence,” Phys. Rev. Lett. 77(21), 4326–4329 (1996). [CrossRef]  

8. A. Lezama, S. Barreiro, and A. Akulshin, “Electromagnetically induced absorption,” Phys. Rev. A 59(6), 4732–4735 (1999). [CrossRef]  

9. A. Taichenachev, A. Tumaikin, and V. Yudin, “Electromagnetically induced absorption in a four-state system,” Phys. Rev. A 61(1), 011802 (1999). [CrossRef]  

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

11. H. Jing, H. Lü, S. Özdemir, T. Carmon, and F. Nori, “Nanoparticle sensing with a spinning resonator,” Optica 5(11), 1424–1430 (2018). [CrossRef]  

12. X.-F. Liu, F. Lei, T.-J. Wang, G.-L. Long, and C. Wang, “Gain lifetime characterization through time-resolved stimulated emission in a whispering-gallery mode microresonator,” Nanophotonics 8(1), 127–134 (2018). [CrossRef]  

13. Y.-L. Liu, R. Wu, J. Zhang, Ş. K. Özdemir, L. Yang, F. Nori, and Y.-x. Liu, “Controllable optical response by modifying the gain and loss of a mechanical resonator and cavity mode in an optomechanical system,” Phys. Rev. A 95(1), 013843 (2017). [CrossRef]  

14. K. Totsuka, N. Kobayashi, and M. Tomita, “Slow light in coupled-resonator-induced transparency,” Phys. Rev. Lett. 98(21), 213904 (2007). [CrossRef]  

15. M. F. Yanik, W. Suh, Z. Wang, and S. Fan, “Stopping light in a waveguide with an all-optical analog of electromagnetically induced transparency,” Phys. Rev. Lett. 93(23), 233903 (2004). [CrossRef]  

16. Q. Xu, S. Sandhu, M. L. Povinelli, J. Shakya, S. Fan, and M. Lipson, “Experimental realization of an on-chip all-optical analogue to electromagnetically induced transparency,” Phys. Rev. Lett. 96(12), 123901 (2006). [CrossRef]  

17. Y. Xiao, L. He, J. Zhu, and L. Yang, “Electromagnetically induced transparency-like effect in a single polydimethylsiloxane-coated silica microtoroid,” Appl. Phys. Lett. 94(23), 231115 (2009). [CrossRef]  

18. L. Zhou, T. Ye, and J. Chen, “Coherent interference induced transparency in self-coupled optical waveguide-based resonators,” Opt. Lett. 36(1), 13–15 (2011). [CrossRef]  

19. C.-H. Dong, C.-L. Zou, Y.-F. Xiao, J.-M. Cui, Z.-F. Han, and G.-C. Guo, “Modified transmission spectrum induced by two-mode interference in a single silica microsphere,” J. Phys. B: At., Mol. Opt. Phys. 42(21), 215401 (2009). [CrossRef]  

20. B. Peng, Ş. K. Özdemir, W. Chen, F. Nori, and L. Yang, “What is and what is not electromagnetically induced transparency in whispering-gallery microcavities,” Nat. Commun. 5(1), 5082 (2014). [CrossRef]  

21. Y. Wang, K. Zhang, S. Zhou, Y. Wu, M. Chi, and P. Hao, “Coupled-mode induced transparency in a bottle whispering-gallery-mode resonator,” Opt. Lett. 41(8), 1825–1828 (2016). [CrossRef]  

22. A. Naweed, G. Farca, S. I. Shopova, and A. T. Rosenberger, “Induced transparency and absorption in coupled whispering-gallery microresonators,” Phys. Rev. A 71(4), 043804 (2005). [CrossRef]  

23. T. Wang, Y.-Q. Hu, C.-G. Du, and G.-L. Long, “Multiple eit and eia in optical microresonators,” Opt. Express 27(5), 7344–7353 (2019). [CrossRef]  

24. Y.-F. Xiao, X.-F. Jiang, Q.-F. Yang, L. Wang, K. Shi, Y. Li, and Q. Gong, “Tunneling-induced transparency in a chaotic microcavity,” Laser Photonics Rev. 7(5), L51–L54 (2013). [CrossRef]  

25. G. Li, X. Jiang, S. Hua, Y. Qin, and M. Xiao, “Optomechanically tuned electromagnetically induced transparency-like effect in coupled optical microcavities,” Appl. Phys. Lett. 109(26), 261106 (2016). [CrossRef]  

26. Q. Huang, Z. Shu, G. Song, J. Chen, J. Xia, and J. Yu, “Electromagnetically induced transparency-like effect in a two-bus waveguides coupled microdisk resonator,” Opt. Express 22(3), 3219–3227 (2014). [CrossRef]  

27. Y. Yang, S. Saurabh, J. M. Ward, and S. N. Chormaic, “Coupled-mode-induced transparency in aerostatically tuned microbubble whispering-gallery resonators,” Opt. Lett. 40(8), 1834–1837 (2015). [CrossRef]  

28. M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, “Cavity optomechanics,” Rev. Mod. Phys. 86(4), 1391–1452 (2014). [CrossRef]  

29. I. Pikovski, M. R. Vanner, M. Aspelmeyer, M. Kim, and Č. Brukner, “Probing planck-scale physics with quantum optics,” Nat. Phys. 8(5), 393–397 (2012). [CrossRef]  

30. X.-F. Liu, T.-J. Wang, Y.-P. Gao, C. Cao, and C. Wang, “Chiral microresonator assisted by rydberg-atom ensembles,” Phys. Rev. A 98(3), 033824 (2018). [CrossRef]  

31. Y. Liu, T. Hao, W. Li, J. Capmany, N. Zhu, and M. Li, “Observation of parity-time symmetry in microwave photonics,” Light: Sci. Appl. 7(1), 38 (2018). [CrossRef]  

32. X. Chen, C. Chardin, K. Makles, C. Caër, S. Chua, R. Braive, I. Robert-Philip, T. Briant, P.-F. Cohadon, A. Heidmann, T. Jacqmin, and S. Deléglise, “High-finesse fabry–perot cavities with bidimensional si 3 n 4 photonic-crystal slabs,” Light: Sci. Appl. 6(1), e16190 (2017). [CrossRef]  

33. K. Cai, R. Wang, Z. Yin, and G. Long, “Second-order magnetic field gradient-induced strong coupling between nitrogen-vacancy centers and a mechanical oscillator,” Sci. China: Phys., Mech. Astron. 60(7), 070311 (2017). [CrossRef]  

34. X.-R. Xiong, Y.-P. Gao, X.-F. Liu, C. Cao, T.-J. Wang, and C. Wang, “The analysis of high-order sideband signals in optomechanical system,” Sci. China: Phys., Mech. Astron. 61(9), 90322 (2018). [CrossRef]  

35. M. AbuGhanem, A. Homid, and M. Abdel-Aty, “Cavity control as a new quantum algorithms implementation treatment,” Front. Phys. 13(1), 130303 (2018). [CrossRef]  

36. M. Tian, Y.-G. Huang, S.-S. Wen, H. Yang, X.-Y. Wang, J.-Z. Peng, and H.-P. Zhao, “Finite-element method for obtaining the regularized photon green function in lossy material,” Europhys. Lett. 126(1), 13001 (2019). [CrossRef]  

37. Y.-F. Jiao, T.-X. Lu, and H. Jing, “Optomechanical second-order sidebands and group delays in a kerr resonator,” Phys. Rev. A 97(1), 013843 (2018). [CrossRef]  

38. Y. Jiao, H. Lü, J. Qian, Y. Li, and H. Jing, “Nonlinear optomechanics with gain and loss: amplifying higher-order sideband and group delay,” New J. Phys. 18(8), 083034 (2016). [CrossRef]  

39. 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]  

40. F. Marquardt, J. P. Chen, A. A. Clerk, and S. Girvin, “Quantum theory of cavity-assisted sideband cooling of mechanical motion,” Phys. Rev. Lett. 99(9), 093902 (2007). [CrossRef]  

41. 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]  

42. J. Chan, T. 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(7367), 89–92 (2011). [CrossRef]  

43. Y.-S. Park and H. Wang, “Resolved-sideband and cryogenic cooling of an optomechanical resonator,” Nat. Phys. 5(7), 489–493 (2009). [CrossRef]  

44. Y.-C. Liu, Y.-F. Xiao, X. Luan, and C. W. Wong, “Dynamic dissipative cooling of a mechanical resonator in strong coupling optomechanics,” Phys. Rev. Lett. 110(15), 153606 (2013). [CrossRef]  

45. S. Al-Awfi, M. Al-Hmoud, and S. Bougouffa, “Ground-state cooling in cavity optomechanics with unresolved sidebands,” Europhys. Lett. 123(1), 14005 (2018). [CrossRef]  

46. H. Jing, Ş. Özdemir, H. Lü, and F. Nori, “High-order exceptional points in optomechanics,” Sci. Rep. 7(1), 3386 (2017). [CrossRef]  

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

48. T. Palomaki, J. Teufel, R. Simmonds, and K. Lehnert, “Entangling mechanical motion with microwave fields,” Science 342(6159), 710–713 (2013). [CrossRef]  

49. J.-Q. Liao, Q.-Q. Wu, and F. Nori, “Entangling two macroscopic mechanical mirrors in a two-cavity optomechanical system,” Phys. Rev. A 89(1), 014302 (2014). [CrossRef]  

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

51. H. Xiong and Y. Wu, “Fundamentals and applications of optomechanically induced transparency,” Appl. Phys. Rev. 5(3), 031305 (2018). [CrossRef]  

52. A. Kronwald and F. Marquardt, “Optomechanically induced transparency in the nonlinear quantum regime,” Phys. Rev. Lett. 111(13), 133601 (2013). [CrossRef]  

53. A. H. Safavi-Naeini, T. M. Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and O. Painter, “Electromagnetically induced transparency and slow light with optomechanics,” Nature 472(7341), 69–73 (2011). [CrossRef]  

54. M.-A. Lemonde, N. Didier, and A. A. Clerk, “Nonlinear interaction effects in a strongly driven optomechanical cavity,” Phys. Rev. Lett. 111(5), 053602 (2013). [CrossRef]  

55. J. Kim, M. C. Kuzyk, K. Han, H. Wang, and G. Bahl, “Non-reciprocal brillouin scattering induced transparency,” Nat. Phys. 11(3), 275–280 (2015). [CrossRef]  

56. Z. Shen, Y.-L. Zhang, Y. Chen, C.-L. Zou, Y.-F. Xiao, X.-B. Zou, F.-W. Sun, G.-C. Guo, and C.-H. Dong, “Experimental realization of optomechanically induced non-reciprocity,” Nat. Photonics 10(10), 657–661 (2016). [CrossRef]  

57. C.-H. Dong, Z. Shen, C.-L. Zou, Y.-L. Zhang, W. Fu, and G.-C. Guo, “Brillouin-scattering-induced transparency and non-reciprocal light storage,” Nat. Commun. 6(1), 6193 (2015). [CrossRef]  

58. H. Zhang, F. Saif, Y. Jiao, and H. Jing, “Loss-induced transparency in optomechanics,” Opt. Express 26(19), 25199–25210 (2018). [CrossRef]  

59. H. Lü, C. Wang, L. Yang, and H. Jing, “Optomechanically induced transparency at exceptional points,” Phys. Rev. Appl. 10(1), 014006 (2018). [CrossRef]  

60. H. Jing, Ş. K. Özdemir, Z. Geng, J. Zhang, X.-Y. Lü, B. Peng, L. Yang, and F. Nori, “Optomechanically-induced transparency in parity-time-symmetric microresonators,” Sci. Rep. 5(1), 9663 (2015). [CrossRef]  

61. K. Qu and G. Agarwal, “Phonon-mediated electromagnetically induced absorption in hybrid opto-electromechanical systems,” Phys. Rev. A 87(3), 031802 (2013). [CrossRef]  

62. F. Hocke, X. Zhou, A. Schliesser, T. J. Kippenberg, H. Huebl, and R. Gross, “Electromechanically induced absorption in a circuit nano-electromechanical system,” New J. Phys. 14(12), 123037 (2012). [CrossRef]  

63. M. C. Kuzyk and H. Wang, “Controlling multimode optomechanical interactions via interference,” Phys. Rev. A 96(2), 023860 (2017). [CrossRef]  

64. H. Zhang, X.-K. Song, Q. Ai, H. Wang, G.-J. Yang, and F.-G. Deng, “Fast and robust quantum control for multimode interactions using shortcuts to adiabaticity,” Opt. Express 27(5), 7384–7392 (2019). [CrossRef]  

65. Y.-P. Gao, Z.-X. Wang, T.-J. Wang, and C. Wang, “Optomechanically engineered phononic mode resonance,” Opt. Express 25(22), 26638–26650 (2017). [CrossRef]  

66. K. Stannigel, P. Komar, S. Habraken, S. Bennett, M. D. Lukin, P. Zoller, and P. Rabl, “Optomechanical quantum information processing with photons and phonons,” Phys. Rev. Lett. 109(1), 013603 (2012). [CrossRef]  

67. C. Dong, V. Fiore, M. C. Kuzyk, and H. Wang, “Optomechanical dark mode,” Science 338(6114), 1609–1613 (2012). [CrossRef]  

68. X.-Y. Lü, Y. Wu, J. Johansson, H. Jing, J. Zhang, and F. Nori, “Squeezed optomechanics with phase-matched amplification and dissipation,” Phys. Rev. Lett. 114(9), 093602 (2015). [CrossRef]  

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

70. Y. Li, Y. Huang, X. Zhang, and L. Tian, “Optical directional amplification in a three-mode optomechanical system,” Opt. Express 25(16), 18907–18916 (2017). [CrossRef]  

71. T.-X. Lu, Y.-F. Jiao, H.-L. Zhang, F. Saif, and H. Jing, “Selective and switchable optical amplification with mechanical driven oscillators,” Phys. Rev. A 100(1), 013813 (2019). [CrossRef]  

72. X. Jiang, M. Wang, M. C. Kuzyk, T. Oo, G.-L. Long, and H. Wang, “Chip-based silica microspheres for cavity optomechanics,” Opt. Express 23(21), 27260–27265 (2015). [CrossRef]  

73. H. Lü, Y. Jiang, Y.-Z. Wang, and H. Jing, “Optomechanically induced transparency in a spinning resonator,” Photonics Res. 5(4), 367–371 (2017). [CrossRef]  

74. F.-C. Lei, M. Gao, C. Du, Q.-L. Jing, and G.-L. Long, “Three-pathway electromagnetically induced transparency in coupled-cavity optomechanical system,” Opt. Express 23(9), 11508–11517 (2015). [CrossRef]  

75. P.-C. Ma, J.-Q. Zhang, Y. Xiao, M. Feng, and Z.-M. Zhang, “Tunable double optomechanically induced transparency in an optomechanical system,” Phys. Rev. A 90(4), 043825 (2014). [CrossRef]  

76. C. Jiang, Y. Cui, X. Bian, F. Zuo, H. Yu, and G. Chen, “Phase-dependent multiple optomechanically induced absorption in multimode optomechanical systems with mechanical driving,” Phys. Rev. A 94(2), 023837 (2016). [CrossRef]  

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

78. C. Bai, B. Hou, D. Lai, and D. Wu, “Tunable optomechanically induced transparency in double quadratically coupled optomechanical cavities within a common reservoir,” Phys. Rev. A 93(4), 043804 (2016). [CrossRef]  

79. L.-G. Si, H. Xiong, M. S. Zubairy, and Y. Wu, “Optomechanically induced opacity and amplification in a quadratically coupled optomechanical system,” Phys. Rev. A 95(3), 033803 (2017). [CrossRef]  

80. X. Zhang, Y. Guo, P. Pei, and X. Yi, “Optomechanically induced absorption in parity-time-symmetric optomechanical systems,” Phys. Rev. A 95(6), 063825 (2017). [CrossRef]  

81. K. Ullah, H. Jing, and F. Saif, “Multiple electromechanically-induced-transparency windows and fano resonances in hybrid nano-electro-optomechanics,” Phys. Rev. A 97(3), 033812 (2018). [CrossRef]  

82. X.-S. Xu, H. Zhang, M. Wang, D. Ruan, and G.-L. Long, “Arbitrary function resonance tuner of the optical microcavity with sub-mhz resolution,” Opt. Lett. 44(13), 3250–3253 (2019). [CrossRef]  

83. C. Gardiner and P. Zoller, Quantum noise: a handbook of Markovian and non-Markovian quantum stochastic methods with applications to quantum optics, vol. 56 (Springer Science & Business Media, 2004).

84. G. Benettin, L. Galgani, A. Giorgilli, and J.-M. Strelcyn, “Lyapunov characteristic exponents for smooth dynamical systems and for hamiltonian systems; a method for computing all of them. part 1: Theory,” Meccanica 15(1), 9–20 (1980). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. (a) Schematic of the optomechanical system consisting of two optical modes and a mechanical resonator. The two optical modes are excited by the waveguide simultaneously. (b) The auxiliary cavity mode $a_2$ is adjacent to the dominant optical mode $a_1$ in frequency domain. The optical mode $a_2$ can be tuned to overlap with $a_1$ , which could lead to electromagnetically-induced transparency-like window [17]. (c) Illustration of the three-pathway interference effect when the auxiliary cavity mode decouples with the mechanical mode. (d) Illustration of the four-pathway interference effect when both optical modes couples to the mechanical mode.
Fig. 2.
Fig. 2. The transmission rate are plotted with the different external coupling strength $\kappa _{ex2}$ of the auxiliary cavity mode $a_2$ in the red sideband regime. The parameters used here are: $\kappa _{10}=5$ MHz, $\kappa _{ex1}=15$ MHz, $\kappa _{20}=40$ MHz, $\Gamma _m=50$ KHz, $\omega _m=100$ MHz and $\Delta = -\omega _m$ .
Fig. 3.
Fig. 3. The transmission rate are plotted with the different optomechanical coupling rate between the optical modes and the phonons. The parameters used here are: $\kappa _{ex1}=15$ MHz, $\kappa _{20}=40$ MHz, $\kappa _{ex2}=60$ MHz, $\Gamma _m=50$ KHz, $\omega _m=100$ MHz and $\Delta = -\omega _m$ .
Fig. 4.
Fig. 4. The transmission is plotted as the function of the optomechanical coupling rate $G_1$ and the external coupling strength $\kappa _{ex2}$ of the auxiliary cavity mode. The parameters used here are: $\kappa _{10}=5$ MHz, $\kappa _{ex1}=15$ MHz, $\kappa _{20}=40$ MHz, $\Gamma _m=50$ KHz, $\omega _m=100$ MHz and $\Delta = -\omega _m$ .
Fig. 5.
Fig. 5. The transmission rate are plotted in the blue sideband driving regime when the auxiliary cavity mode decouples to the mechanical mode. (a) The external coupling rate $\kappa _{ex2}$ varies from 0 to $8\kappa _{10}$ when $G_1$ is fixed to 0.3 MHz. (b) Here $\kappa _{ex2}$ is fixed to 60 MHz while the $G_1$ varies from 0 to 0.3 MHz. (c) The transmission of the central peak is presented as the function of the optomechanical coupling strength $G_1$ . The parameters used here are: $\kappa _{10}=5$ MHz, $\kappa _{ex1}=20$ MHz, $\kappa _{20}=40$ MHz, $\Delta =\omega _m$ , $\Gamma _m=50$ KHz and $G_2 = 0$ . The inset of figure (c) shows the border between the stable and unstable area when $\kappa _{ex2}= 60$ MHz.
Fig. 6.
Fig. 6. The output spectra are shown in (a) and (b) as a function of $\delta$ with different $G_2$ . The effective optomechanical coupling strength $G_1$ is fixed to 0.6 MHz. The central peak is plotted in (c) as a function of $G_2$ . The power transmission coefficient T as a function of $G_2$ and $G_1$ in (d). Here $\kappa _{10}=5$ MHz, $\kappa _{ex1}=20$ MHz, $\kappa _{20}=40$ MHz, $\kappa _{ex2}=60$ MHz, $\Gamma _m=5$ KHz and $\Delta = -\omega _m$ .

Equations (13)

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

H = ω 1 a 1 a 1 + ω 2 a 2 a 2 + 1 2 m ω m 2 x 2 + P 2 2 m + + ω c ( ω ) c ( ω ) d ω + g 1 x a 1 a 1 + g 2 x a 2 a 2 + i + j = 1 , 2 κ e x , j ( ω ) [ c ( ω ) a j a j c ( ω ) ] d ω
d a 1 d t = i Δ 1 a 1 κ 1 2 a 1 i g 1 x a 1 κ e x 1 κ e x 2 2 a 2 + κ e x 1 ε p + κ e x 1 ε s e i δ t + κ e x 1 ξ 1
d a 2 d t = i Δ 2 a 2 κ 2 2 a 2 i g 2 x a 2 κ e x 1 κ e x 2 2 a 1 + κ e x 2 ε p + κ e x 2 ε s e i δ t + κ e x 2 ξ 2
d x d t = P m
d P d t = m ω m 2 x ( g 1 a 1 a 1 + g 2 a 2 a 2 ) Γ m 2 P + δ F
a ¯ 1 = κ e x 1 κ e x 1 κ e x 2 2 ( i Δ 2 i g 2 x ¯ κ 2 / 2 ) i Δ 1 i g 1 x ¯ κ 1 / 2 κ e x 1 κ e x 2 4 ( i Δ 2 i g 2 x ¯ κ 2 / 2 ) ε p
a ¯ 2 = κ e x 1 κ e x 2 2 a ¯ 1 κ e x 2 ε p i Δ 2 i g 2 x ¯ κ 2 / 2
x ¯ = g 1 | a ¯ 1 | 2 + g 2 | a ¯ 2 | 2 m ω m 2
d δ a 1 d t = ( i Δ ¯ 1 κ 1 2 ) δ a 1 i g 1 a ¯ 1 δ x κ e x 1 κ e x 2 2 δ a 2 + κ e x 1 ε s e i δ t + κ e x 1 ξ 1
d δ a 2 d t = ( i Δ ¯ 2 κ 2 2 ) δ a 2 i g 2 a ¯ 2 δ x κ e x 1 κ e x 2 2 δ a 1 + κ e x 2 ε s e i δ t + κ e x 2 ξ 2
m d 2 d t 2 δ x = m Γ m 2 d d t δ x m ω m 2 δ x g 1 a ¯ 1 ( δ a 1 + δ a 1 ) g 2 a ¯ 2 ( δ a 2 + δ a 2 ) + δ F
t r = ( L 1 κ e x 1 ) ( L 2 κ e x 2 ) ( A κ e x 1 κ e x 2 ) 2 L 2 ( L 1 A 2 L 2 )
t b = ( R 1 + κ e x 1 ) ( R 2 + κ e x 2 ) ( B κ e x 1 κ e x 2 ) 2 R 2 ( R 1 B 2 R 2 )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.