Three-mode optomechanical system for angular velocity detection
Li Kai, Davuluri Sankar, Li Yong
Beijing Computational Science Research Center, Beijing 100193, China

 

† Corresponding author. E-mail: liyong@csrc.ac.cn

Project supported by the National Key Research and Development Program of China (Grant No. 2016YFA0301200), the National Basic Research Program of China (Grant No. 2014CB921403), the Science Challenge Project of China (Grant No. TZ2017003), and the National Natural Science Foundation of China (Grant Nos. 11774024, 11534002, and U1530401).

Abstract

We propose a scheme for measuring the angular velocity of absolute rotation using a three-mode optomechanical system in which one mode of the two-dimensional (2D) mechanical resonator is coupled to an optical cavity. When the total system rotates, the Coriolis force acting on the 2D mechanical resonator due to the absolute rotation will affect the mechanical motion and thus change the phase of the output field from the cavity. The angular velocity of the absolute rotation can be estimated by monitoring the spectrum of the output field from the cavity via homodyne measurement. The minimum measurable angular velocity, which is determined by the noise spectrum, is calculated. The working range of the gyroscope for measuring angular velocity is discussed.

1. Introduction

Inertial navigation[1,2] systems have many applications such as guiding missiles, satellites, and so on. At a fundamental level, an inertial navigation system contains an accelerometer, which is used to measure instantaneous velocity, and a gyroscope which plays an important role as a direction sensor. Different kinds of gyroscopes like mechanical gyroscopes [e.g., gyroscopes based on microelectromachanical systems (MEMSs)[3,4] utilizing the Coriolis force], optical gyroscopes (e.g., ring-laser gyroscopes and fiber-optic gyroscopes, which utilize the Sagnac effect,[5]) and nuclear magnetic resonance gyroscopes[6,7] or atom interferometer gyroscopes[8,9] which utilize the quantum effect, have been studied extensively.

On the other hand, because of the recent advances in nano-fabrication techniques for designing high-quality optical and mechanical resonators, optomechanics has emerged as a rapidly developing field for exploring the coupling between optical fields and mechanical oscillations. One of the most promising applications of optomechanics is for designing highly sensitive optical detection of small forces,[1013] displacements[1417] or masses[1820] by manipulating the mechanical motion of a macroscopic object in the quantum regime.[2128] Hence, optomechanical systems are used to detect the angular velocity of absolute rotation,[29] by optically monitoring the Coriolis force acting on the two-dimensional (2D) mechanical resonator.

Gyroscopes based on the Coriolis force detection were discussed in Refs. [3] and [29]. The analysis provided in Ref. [29] is based on classical mechanics and the effects of quantum noise, such as optomechanical back-action noise, are neglected. Recently, we presented theoretical analysis for the application of the optomechanical cavities for absolute rotation detection[3032] by considering the effects of shot noise, back-action noise, and thermal noise.[33] In this work, we propose an optomechanical gyroscope model, where one mode of the 2D mechanical resonator is coupled to an optical cavity. Our model is similar to that in Refs. [31] and [32] except that here we do not use an external drive on the mechanical resonator. A strong external mechanical drive used in previous optomechanical gyroscopes[29,31,32] and thus the driven mode of the 2D mechanical resonator is treated classically by ignoring the effects of the related thermal noise and Coriolis force. In the present work, without using such an external drive, we treat all the three-modes of the system quantum mechanically and provide an alternative method for measuring angular velocity.

2. Model and Hamiltonian

The gyroscope is made of a three-mode optomechanical system, shown in Fig. 1(a), consisting of two parts: a 2D mechanical oscillator which does simple harmonic motion along the x axis and y axis, and a single-mode cavity field which is used to monitor the 2D mechanical oscillator’s displacement along the x axis via optomechanical coupling of the radiation pressure force. The gyroscope is placed on a platform rotating with angular velocity Ω. Ω is along the z axis and goes through the original point, e.g., the center of the platform.

Fig. 1. (color online) (a) Schematics of the gyroscope containing the three-mode optomechanical system. A 2D mechanical oscillator (without mechanical drive) is placed on a platform rotating with angular velocity Ω (Ω is parallel to the z axis). The two mechanical modes are coupled by the Coriolis effect due to the absolute rotation. The mechanical motion in the x direction is coupled to an optical cavity by radiation pressure. (b) Schematics of the balanced homodyne measurement. The output field from the cavity is mixed with a much stronger classical reference laser at the input of a 50:50 beam splitter. The output from the beam splitter is fed to two photo-detectors D1 and D2. The difference in the photodetector readings gives the phase of the output field from the cavity.

In the coordinate system fixed to the rotating platform, the Hamiltonian of the 2D mechanical oscillator is given as where x and y represent the displacements of the mechanical oscillator along the x axis and y axis, respectively. m and ωx (ωy) are the mass and the frequency of the mechanical oscillator along the x axis (y axis), respectively. The equilibrium position of the mechanical oscillator is placed at the original point. The terms depending on Ω in Eq. (1) result from the Coriolis force acting, in the xy plane, on the 2D mechanical oscillator.

Noteworthily, we have omitted the centrifugal force in Eq. (1) by assuming it is much smaller than the Coriolis force, that is . Actually, even though the centrifugal force is not negligible compared with the Coriolis force, it will also not bring a significant effect on our result of angular velocity detection. As we will show later, e.g., in Eqs. (31)–(39), the angular velocity is deduced by detecting the optical output spectrum. Though the Coriolis force will change the output spectrum in all the frequency range, we only focus on the spectrum around ω = ωy for a realistic detection. In contrast, the centrifugal force will only change the zero-frequency component of the output spectrum (for a constant angular velocity) and thus will not affect the detection result of angular velocity. For the sake of simplicity, we do not take the centrifugal force into account in this work.

The canonical momenta conjugate to x and y are given as[34,35] where and represent the mechanical momenta of the mechanical oscillator along the x axis and y axis, respectively. Hence, the corresponding commutation relations are [x,px] = i and [y,py] = i with the reduced Planck constant.

The model under consideration is a three-mode optomechanical system involving two mechanical modes and one cavity field mode. After taking the cavity field into account, the Hamiltonian for the whole system is given as where a is the annihilation operator for the cavity field, and g = ωc/L is the single-photon radiation pressure coupling strength with L the length of the cavity and ωc the eigen-frequency of the cavity. The cavity is coupled to an external driving field whose frequency and the driving strength are given as ωd and , respectively.

After making a rotating frame at drive frequency ωd, the Heisenberg–Langevin equations of motion are given as with κ being the decay rate of the cavity field and γx(γy) being the decay rate of the mechanical oscillation along the x axis (y axis). Δ0 = ωcωd is the detuning between the cavity field mode and the drive field. ξx,y and δain are the noise operators with zero mean values, and their non-zero correlations[36,37] are given as where kB is the Boltzmann constant and T is the bath temperature. We assumed that kBTℏωx,y in Eq. (10).

Note that the present optomechanical gyroscope model is different from the ones used in previous works[29,31,32] in the following way. In the previous works,[29,31,32] a classical external drive is applied on the mechanical resonator along the y axis. In the presence of such a strong classical external drive, the mechanical motion along the y axis is treated classically (e.g., the classical motion with sinusoidal velocity[29,32] or constant velocity[31] along the y direction) by ignoring the effects of the related thermal noise and Coriolis force. Thus, the corresponding Heisenberg–Langevin equations of motion contain only a single cavity mode and a single mechanical mode in the x axis. In the present work, due to the absence of such a classical external drive in our model, the Heisenberg–Langevin equations of motion for the complete system includes the motion of all the three modes (one cavity mode and two mechanical modes) as given in Eqs. (4)–(8).

3. The spectrum of fluctuation and homodyne measurement

Equations (4)–(8) can be linearized by writing all the operators as (A = a,x,y,px,py), where and δA are the classical mean value and quantum fluctuation of A. Thus the steady state solutions for the mean value equations of motion of Eqs. (4)–(8) can be obtained by setting as where the effective detuning is As can be seen from Eq. (13), the mean photon number, which is given by , inside the cavity is independent of Ω. Hence we cannot estimate Ω by monitoring the mean intensity of the cavity output field. Thus, we shall focus on the equations of motion of the fluctuation operators in order to measure Ω. When Ω ≠ 0, the Coriolis force coupling between the two mechanical modes modifies the spectrum of the fluctuations in the output optical field via radiation-pressure optomechanical coupling. Thus the absolute rotation of the platform can be monitored from the spectrum of the fluctuations in the output optical field.

The linearized equations of motion for the fluctuation operators in Eqs. (4)–(8) are where the effective coupling strength has been assumed to be real without loss of generality. Note that in writing Eqs. (15)–(19), the higher-order terms like δxδa are neglected. This is reasonable since we can make by changing the intensity of the driving laser, .

By defining the Fourier transform of operator A as with , we obtain the solution to the linearized equations (15)–(19) in the frequency domain as with Here we defined Making use of the input–output relationship[38] , we can obtain the fluctuation and steady-state mean value of the output cavity field as with and .

The phase of the output field, represented by aout, from the cavity is measured using a homodyne setup as shown in Fig. 1(b). The output field aout is mixed with a classical reference laser field,[39] whose amplitude is αR, at a 50:50 beam splitter as shown in Fig. 1(b). The output from the beam splitter is fed onto photo-detectors D1 and D2. The readings in D1 and D2 are given as and , respectively, with and . The difference in the photo detector readings is given as where αR has been taken to be real.

We rewrite I as a sum of its mean value and fluctuation by writing . So we obtain which is independent of Ω. The fluctuation δI in I is given as The spectrum of δI is given as By using Eqs. (9), (10), (20), (25), and (29), we can obtain the spectrum with In Eq. (31), Sadd includes both the optical shot noise and the optomechanical back-action noise, Sxx gives the thermal noise in the mechanical oscillator’s motion along the x axis, and Syy gives the effect of the Coriolis force on the mechanical oscillator’s motion along the x axis due to the thermal vibrations along the y axis. Note that χ(ω) is dependent on Ω, which means Sxx and Syy (as well as Sadd) include the effect of the Coriolis forces.

4. Estimation of angular velocity and its minimum detectable value

In principle, the fact that all the three parts (Sadd, Sxx, and Syy) in the spectrum SII depend on Ω, allows one to measure Ω by monitoring the optical output spectrum. Here, we will just focus on the following simple case When equation (35) is satisfied, χ(ω) is independent of Ω and is given as Thus, both Sadd(ω) and Sxx(ω) are independent of Ω, and Syy(ω) can be rewritten as with the transmission coefficient

The fact that Syy is proportional to Ω2 provides us with a method to estimate the angular velocity as , from the homodyne measurement results of the optical output spectrum. The other two parts in the spectrum SII(ω), i.e., Sadd and Sxx, determine the noise in the optical output spectrum. Hence the minimum measurable angular velocity Ωnoise is given as In principle, the above minimum measurable angular velocity will be obtained when the measurement time tm → ∞. Note that the signal spectrum Syy(ω) as well as the transmission coefficient σ(ω) is a Lorentz peak with full width at half maximum being of the order of γy. In a realistic detection, when the measurement time is long enough such that tm ≫ 1/γy, one can obtain approximately the minimum measurable angular velocity Ωnoise.

In what follows, we will only consider two limiting cases: Sadd(ω) ≪ Sxx(ω) and Sadd(ω) ≫ Sxx(ω).

When the minimum detectable angular velocity is independent of T, Δ, and ωx.

According to Eq. (41), Ωnoise(ω) reaches its minimum value when ω = ωy. For simulation, we consider the parameters:[40]m = 6.9 × 10−4 kg, ωy/2π = 84.8 Hz, γx/2π = γy/2π = 1.9 × 10−3 Hz, κ/2π = 4 × 106 Hz, and g/2π = 1.46 ×1017 Hz/m. Temperature is as T = 300 K and the effective optomechanical coupling strength is G/2π = 2.19 × 1022 Hz/m (with the corresponding ). The minimum detectable angular velocity is shown in Fig. 2(a) with the optimal value Ωnoise(ωy)/2π = 9.5 × 10−4 Hz. The corresponding transmission coefficient σ(ωy) is optimized when Δ = 0 as shown in Fig. 2(b).

Fig. 2. (color online) (a) The minimum detectable angular velocity in Eq. (41). (b) The corresponding transmission coefficient σ(ωy) versus Δ for different values of ωx. (c) Simulation of the condition given in Eq. (35) with ω = ωy for different ωx. (d) The ratio of Sadd(ωy) in Eq. (32) to Sxx(ωy) in Eq. (33) as a function of ωx and Ω with fixed Δ = 0 and T = 300 K. The other parameters are:[40]m = 6.9 × 10−4 kg, ωy/2π = 84.8 Hz, γx/2π = γy/2π = 1.9 × 10−3 Hz, κ/2π = 4 × 106 Hz, and g/2π = 1.46 × 1017 Hz/m. The effective optomechanical coupling strength is taken as G/2π = 2.19 × 1022 Hz/m (with the corresponding ). For these parameters, the optimal value of Ωnoise(ωy)/2π = 9.5 × 10−4 Hz, also see the vertical line in panel (c).

Note that the above analysis for estimating the minimum detectable angular velocity as well as the optimal transmission coefficient is valid only when equations (35) and (40) are satisfied. As shown in Fig. 2(c), the condition in Eq. (35) is not fulfilled for the resonance case of ωx = ωy when Ω is larger than the optimal minimum detectable angular velocity Ωnoise. For the non-resonance case of ωx = 2ωy or ωx = 0.5ωy, the condition in Eq. (35) is satisfied well when Ω/2π is less than about 0.1 Hz (∼10−3ωy). Since the condition of Eq. (40) is always fulfilled for the above parameters [see Fig. 2(d)], we can know that for the typical simulation parameters as given in the caption of Fig. 2 and the non-resonance case of ωx ≃ 2ωy, the optimal minimum detectable angular velocity of our optomechanical gyroscope is Ωnoise(ωy)/2π = 9.5 × 10−4 Hz with the corresponding working range of the angular velocity being Ωnoise/2πΩ/2π ≲ 0.1 Hz.

Now we consider the second limiting case: when In this case the minimum detectable angular velocity reduces to which is dependent on T, Δ, and ωx.

In Fig. 3(a), we plot the minimum detectable angular velocity in Eq. (43) for the parameters:[41]m = 1.9 × 10−7 kg, ωy/2π = 8 × 105 Hz, γx/2π = γy/2π = 80 Hz, κ/2π = 1 × 106 Hz, and g/2π = 1.17 × 1017 Hz/m. Temperature is T = 300 K and the effective optomechanical coupling strength is G/2π = 3.51 × 1020 Hz/m (with the corresponding ). The optimal minimum detectable angular velocity appears at ω = ωy [as similar to the case in Fig. 2(a)] with Ωnoise/2π = 92.6 Hz, 40.0 Hz, and 336.4 Hz for ωx = 0.5ωy, ωy, and 2ωy, respectively. The corresponding transmission coefficient σ(ωy) is shown in Fig. 3(b) with its optimal value appearing at Δ = 0.

Fig. 3. (color online) (a) The minimum detectable angular velocity in Eq. (43) for different values of ωx. (b) The corresponding transmission coefficient σ(ωy) versus Δ for different ωx. (c) Simulation of the condition given in Eq. (35) with ω = ωy for different ωx. (d) The ratio of Sadd(ωy) in Eq. (32) to Sxx(ωy) in Eq. (33) as a function of ωx and Ω. Here Δ = 0 in panels (a) and (d). The other parameters are:[41]m = 1.9 × 10−7 kg, ωy/2π = 8 × 105 Hz, γx/2π = γy/2π = 80 Hz, κ/2π = 1 × 106 Hz, T = 300 K, and g/2π = 1.17 × 1017 Hz/m. The effective optomechanical coupling strength is taken as G/2π = 3.51 × 1020 Hz/m (with the corresponding ).

The above analysis of the minimum detectable angular velocity as well as the optimal transmission coefficient is only valid when equations (35) and (42) are satisfied. It is shown in Fig. 3(c) that the condition in Eq. (35) is not fulfilled for the resonance case of ωx = ωy when Ω is larger than the optimal Ωnoise. However, this condition is satisfied well for the non-resonance case of ωx = 2ωy or ωx = 0.5ωy when Ω/2π is less than 800 Hz (∼10−3ωy). It is shown in Fig. 3(d) that the condition of Eq. (42) is always fulfilled for the above considered parameters. Thus we know for the typical parameters as given in the caption of Fig. 3, and under the non-resonance case ωx ≃ 0.5ωy, the optomechanical gyroscope works well when Ωnoise(ωy)/2πΩ/2π ≲ 800 Hz and the optimal minimum detectable angular velocity Ωnoise(ωy)/2π = 92.6 Hz.

5. Discussion and conclusion

In conclusion, we studied the application of a three-mode optomechanical system as a gyroscope for detecting absolute rotation. Our model of an optomechanical gyroscope, where a cavity field mode is coupled to one of the two mechanical modes of a 2D resonator, is similar to the previous ones[2931] except that here we do not use any additional external mechanical drive on the mechanical resonator. When our optomechanical system is placed on a rotating platform, the mechanical oscillation modes of the 2D mechanical oscillator are coupled due to the Coriolis force. Using the Heisenberg–Lengivin equations of motion, the dynamics of all the three modes (one cavity field mode and two mechanical oscillator modes) are studied. The angular velocity of the absolute rotation is estimated by monitoring the spectrum of the output field from the cavity via homodyne measurement. The noises in the output spectrum, which determine the minimum detectable angular velocity, include the thermal noise of the mechanical oscillator (Sxx), and the additional noise (Sadd) from the optical shot noise and optomechanical back-action. We further analyzed the cases when SxxSadd and SxxSadd, and estimated the optimal minimum detectable angular velocity. The corresponding working ranges of the angular velocity in these two cases are also discussed.

We would like to remark that our analysis is based on the condition in Eq. (35), which requires that the angular velocity should not be too strong. That is, the noises in the optical output spectrum are required to keep approximately unchanged when the system rotates. This limits the maximum value of the angular velocity to be detected in our model. Actually, similar conditions also exist in the previous optomechanical gyroscopes.[29,31,32] In future, we will further investigate the application of an optomechanical gyroscope in the case of large angular velocity which does not satisfy the condition in Eq. (35).

Reference
[1] Chatfield A B 2015 Fundamentals of High Accuracy Inertial Navigation American Institute of Aeronautics and Astronautic
[2] Britting K R 1971 Inertial Navigation Systems Analysis Wiley-Interscience
[3] Acar C Shkel A 2009 MEMS Vibratory Gyroscopes Springer
[4] Zaman M F Sharma A Hao Z Ayazi F 2008 Journal of Microelectromechanical System 17 1526
[5] Ezekiel S Arditty H J 1981 Fiber-Optic Rotation Sensors and Related Technologies Springer
[6] Karwacki F A 1980 Navigation 27 72
[7] Meyer D Larsen M 2014 Gyroscopy & Navigation 5 75
[8] Gustavson T L Bouyer P Kasevich M A 1997 Phys. Rev. Lett. 78 2046
[9] Müller T Gilowski M Zaiser M Berg P Schubert Ch Wendrich T Ertmer W Rasel E M 2009 Eur. Phys. J. 53 273
[10] The LIGO Scientific Collaboration 2013 Nat. Photon. 7 613
[11] Rugar D Budakian R Mamin H J Chui B W 2004 Nature 430 329
[12] Zhang J Q Li Y Feng M Xu Y 2012 Phys. Rev. 86 053806
[13] Xiong H Si L G Wu Y 2017 Appl. Phys. Lett. 110 171102
[14] Hoff U B Harris G I Madsen L S Kerdoncuff H Lassen M Nielsen B M Bowen W P Andersen U L 2013 Opt. Lett. 38 1413
[15] Anetsberger G Arcizet O Gavartin E Unterreithmeier Q P Weig E M Kotthaus J P Kippenberg T J 2009 Nat. Phys. 5 909
[16] Yin Z Q Li T C Feng M 2011 Phys. Rev. 83 013816
[17] Zhou L M Xiao K W Yin Z Q Chen J Zhao N 2017 arXiv: 1709.07177
[18] Lin Q He B Xiao M 2017 Phys. Rev. 96 043812
[19] Li J J Zhu K D 2012 Appl. Phys. Lett. 101 141905
[20] Zhao N Yin Z Q 2014 Phys. Rev. 90 042118
[21] Liao J Q Law C K 2011 Phys. Rev. 83 033820
[22] Wang Q Zhang J Q Ma P C Yao C M Feng M 2015 Phys. Rev. 91 063827
[23] Xiong H Si L G Zheng A S Yang X X Wu Y 2012 Phys. Rev. 86 013815
[24] Wilson-Rae I Nooshi N Zwerger W Kippenberg T J 2007 Phys. Rev. Lett. 99 093901
[25] Guo Y J Li K Nie W J Li Y 2014 Phys. Rev. 90 053841
[26] Fu H Ding J F Li Y Cao G Y 2015 Sci. China-Phys. Mech. Astron. 58 050304
[27] Zhang J Mu Q X Zhang W Z 2018 Chin. Phys. 27 040304
[28] Liu Y C Xiao Y F Luan X S Wong C W 2015 Sci. China-Phys. Mech. Astron. 58 050305
[29] Norgia M Donati S Hybrid 2001 Electron. Lett. 37 756
[30] Davuluri S 2016 Phys. Rev. 94 013808
[31] Davuluri S Li Y 2016 New J. Phys. 18 103047
[32] Davuluri S Li K Li Y 2017 New J. Phys. 19 113004
[33] Gardiner C W Zoller P 2000 Quantum Noise Springer
[34] Gazeau J P Koide T Murenzi R 2017 Europhys. Lett. 118 50004
[35] Arnold V I Kozlov V V Neishtadt A I 2006 Mathematical Aspects of Classical and Celestial Mechanics Springer
[36] Breuer H P Petruccione F 2006 The theory of open quantum systems Oxford
[37] Clerk A A Devoret M H Girvin S M Marquardt F Schoelkopf R J 2010 Rev. Mod. Phys. 82 1155
[38] Walls D F Milburn G J 2008 Quantum Optics Springer
[39] Loudon R Knight P L 1987 J. Mod. Opt. 34 709
[40] Mow-lowry C M Mullavey A J Gossler S Gray M B McClelland D E 2008 Phys. Rev. Lett. 100 010801
[41] Arcizet O Cohadon p -F Briant T Pinard M Heidmann A 2006 Nature 444 71