A high-sensitivity gyroscope is vital for both investigation of the fundamental physics and monitoring of the subtle variation of Earth’s behaviors. However, it is a challenge to realize a portable gyroscope with sensitivity approaching a small fraction of the Earth’s rotation rate. Here, we theoretically propose a method for implementing a table-top gyroscope with remarkably high sensitivity based on photon drag in a rotating dielectric object. By inserting an Er3+-doped glass rod in a Fabry–Pérot optical cavity with only 20 cm length, we theoretically show that the giant group refractive index and the narrowing cavity linewidth due to slow light can essentially increase the nonreciprocal phase shift due to the photon drag to achieve a rotation sensitivity of 26 frad/s/Hz. This work paves the way to accurately detect tiny variations of the Earth’s rotation rate and orientation and even can test the geodetic and frame-dragging effects predicted by the general relativity with small-volume equipment.

The sensitive gyroscope measuring the rotation of an object or the frame can be used to test fundamental physical effects1–3 and also promise important applications such as directional positioning,4,5 inertial navigation systems,6–9 and sensing length-of-day variation of the Earth.10,11 Currently, various methods are developed to measure slow rotation by using microelectromechanical gyroscopes,12–16 surface acoustic wave gyroscopes,17–20 ring laser gyroscopes (RLGs),21–25 fiber optic gyroscopes (FOGs),26–28 atomic-optical hybrid gyroscopes,29,30 and nanophotonic optical gyroscopes.31 Among them, optical gyroscopes based on the Sagnac effect are the most reliable and show the highest sensitivity thus far. Its principle involves two counter-propagating waves in the same closed loop, generating a beat frequency reflecting the rotation rate.32 

A Sagnac-effect-based gyroscope can reach a high sensitivity but typically requires a large size. The ROMY RLG with six 12 m arms displays a sensitivity of 80prad/s/Hz.21,33 Thus, it can analyze small ground disturbances caused by earthquakes and oceans. The GINGERINO with four 3.6 m long sides shows a sensitivity up to the level of prad/s/Hz.34 The ring gyroscope G using a 16-m-circumference ring cavity achieves 12prad/s/Hz and a resolution of several milliseconds in detecting disturbances on the Earth’s surface after 120 days of continuous measurements.35 These RLGs utilize mirrors with ultra-low optical loss to construct high finesse optical cavities and use two counter-propagating light beams to reduce bias drift errors.36,37 Despite great success in rotation measurement, these RLGs are large in size thus far, costly, and fixed on a basis. In contrast, FOGs shrink the volume by circling a long optical fiber to a many-turn coil and, therefore, are portable, but the obtained sensitivity is relatively lower. The state-of-the-art FOG can only achieve a sensitivity of 9.7×108rad/s/Hz,38 several orders lower than the RLG counterparts. Because of the long optical path, the application of FOGs is constrained by the asymmetric excitation in the middle of the fiber circuit caused by temporal and local perturbation, resulting in the non-negligible dynamic errors.39,40 Such asymmetric perturbation is a common challenge in the Sagnac-effect-based gyroscope with a big ring. Precise measurement of daily variations of the Earth’s rotation rate (Ωe = 72.92 μrad/s) requires a sensitivity approaching the level of tens prad/sHz, which has only been achieved by the large-size RLG.2,21,34,35 Therefore, it is highly desirable to develop a portable gyroscope with such sensitivity but remains a challenge.41 

Propagation of linearly polarized light in a spinning medium results in the change of polarization plane, known as the photon drag effect, because the rotation of the medium causes circular birefringence.42 The slow-light medium can significantly enhance the photon drag and narrow the linewidth of a cavity by several orders.43,44 However, it is thus far yet to explore the interplay of slow light and the photon drag effect for a sensitive gyroscope.

In this work, we theoretically propose a photon-drag-based gyroscope by embedding a slow-light medium in a 20 cm long Fabry–Pérot (FP) cavity. By measuring the transmitted light power of the cavity, we can, in principle, achieve the remarkable gyro sensitivity. Our proposal provides a table-top device for measuring daily variations in the Earth’s rotation rate and testing the effects of Lorentz violation and general relativity on Earth.

The paper is organized as follows: The system and model are described and developed in Sec. II. Then, the results are presented in Sec. III. The feasibility of implementing our system is discussed in Sec. IV. In the end, this paper is concluded in Sec. V.

The system of our gyroscope is depicted in Fig. 1(a). An Er3+-doped glass rod is inserted into the FP cavity by two highly reflective mirrors, M1 and M2. By detecting the field component orthogonal in polarization to the input field, this system can work as a gyroscope. We assume that the two mirrors have the same reflectivity R and cause a decay rate κe to the cavity each. We also assume that the cavity has an intrinsic loss with rate κi. The overall decay rate of the cavity is then given by κ = 2κe + κi. The two ends of the rod are coated with anti-reflective film. A horizontally polarized (HP) probe laser beam, denoted as âinH, is input to the cavity via the M1 mirror. âout denotes the transmitted field. When the frame or the glass rod spins at a rate Ω, the photon drag causes a difference Δn between the refractive indices nl and nr “seen” by the left-hand circularly polarized (LCP) and the right-hand circularly polarized (RCP) cavity eigenmodes, respectively. Because of the unidirectional mechanical angular momentum, this difference Δn leads to a nonreciprocal phase shift and lifts the resonance frequencies of the LCP and RCP eigenmodes to non-degenerate. As a result, the LCP and RCP components of the input field accumulate a non-zero relative phase after transmitting through the cavity. A vertically polarized (VP) component âoV appears in the transmitted field. It is then separated by a polarization beam splitter (PBS). By detecting its power with a photodetector, we can determine the spinning rate Ω. In addition, quantum noise (êin,1 and êin,2) also enters the photodetector and causes errors.

FIG. 1.

(a) Schematic of gyroscope based on the slow-light enhanced photon drag in a high-Q FP cavity embedded with an Er3+-doped glass rod spinning at a rate Ω. A HP light is input to the cavity. The VP transmitted laser field is detected by a photodetector (PD). (b) Group refractive index vs the input photon flux under different mirror reflectivity when δ = 10 Hz. (c) Group refractive index with input photon flux at various frequency detuning when the mirror reflectivity R = 0.998 and κ = 5 MHz. Other parameters: L = 10 cm and L0 = 20 cm, yielding η = L/L0 = 0.5.

FIG. 1.

(a) Schematic of gyroscope based on the slow-light enhanced photon drag in a high-Q FP cavity embedded with an Er3+-doped glass rod spinning at a rate Ω. A HP light is input to the cavity. The VP transmitted laser field is detected by a photodetector (PD). (b) Group refractive index vs the input photon flux under different mirror reflectivity when δ = 10 Hz. (c) Group refractive index with input photon flux at various frequency detuning when the mirror reflectivity R = 0.998 and κ = 5 MHz. Other parameters: L = 10 cm and L0 = 20 cm, yielding η = L/L0 = 0.5.

Close modal

We consider the slowly rotating frame, e.g., Earth’s rotation. Then, the photon drag is very weak. To amplify the signal, another control laser beam b̂inH is used to increase the group velocity index of the probe field to ng ≫ 1 in the glass rod via the coherent population oscillation (CPO) process of the Er3+ ions. The enhancement resulting from slow light is twofold: enhancing the photon drag by several orders43 and significantly narrowing the cavity linewidth.44 The combination of both allows us to make a table-top gyroscope with sensitivity exceeding the large RLGs. Here, we neglected the influence of the photon drag due to the spin of the mirrors and the PBS because it is much smaller than the slow-light enhanced photon drag in the Er3+-doped glass rod.

Now we discuss how to enhance the gyroscope sensitivity by using slow light in a FP cavity. In the absence of rotation, the control and probe laser beams act on the Er3+-doped glass rod together. At this time, Er3+ ions can be approximated as a two-level system with transition I15/24I13/24 at λ ≈ 1536 nm.45,46 The transition frequency is ω. T1 is the relaxation time of the excited state. The burnt hole can appear in the probe spectrum via the CPO process. It can cause a giant group refractive index ng given by47,48 (see the  Appendix)
(1)
where n0 is the refractive index of the host medium, α0 is the unsaturated absorption coefficient, and δ = ωpωs with ωp and ωs being the frequencies of the probe and control fields, respectively. Isat is the saturated absorption intensity of the ions. Ip and Is denote the intensity of the probe and control field, respectively. ng is crucially dependent on the ratio I0 = Is/Isat. The saturated absorption intensity of the ions is given by
(2)
The operators σ12 and σ21 are the absorption cross section and emission cross section, respectively; σ12σ21. The intensities of the control fields in the cavity are given by
(3)
where E0=ωs/ε0V, ɛ0 is the dielectric constant, and V is the mode volume. The input powers of the control and probe fields are Ps,in=ωs|binH|2 and Pp, in=ωp|ainH|2, respectively. Throughout the following investigation, we fix the power ratio to Pp,in/Ps,in = Ip/Is = 0.08, which is accessible in the experiment.49 In an experiment, one can modulate the input laser beam to form the control and probe fields. The power or intensity ratio between the control beam and the probe beam can be tuned via modulation. In our system, the modulation frequency is very small, such that ωpωs guarantees a giant group refractive index ng. For an Er3+-doped glass rod, we have n0 ≈ 1.48, T1 = 10.5 ms, and α0 = 0.16 cm−1.46 

Figure 1(b) reflects the group refractive index ng as a function of the power of the control laser field for various mirror reflectivity. The peak values of ng are very close for different reflectivities, but the required control laser power, Ps,in, increases as the reflectivity increases. When R = 0.99, the peak ng ≈ 3.7 × 106 is attained at Ps,in = 440 mW. When the reflectivity is improved to R = 0.999, which is much harder, the peaked ng is about 3.7 × 106, and the optimal power Ps,in decreases to 274mW, corresponding to 2.1 × 1018. Thus, it is experimentally convenient to choose the reflectivity between 0.99 and 0.999. Figure 1(c) shows the dependence of ng on the input photon flux of the control light for various frequency differences δ. The peak ng first increases rapidly and then decreases after reaching the maximum value as the control laser power increases. The peak value also decreases as the modulation frequency δ increases. From Figs. 1(b) and 1(c), the group refractive index can reach the level of 106.

We learn from Ref. 44 that slow light can narrow the linewidth of the cavity. The total loss can be approximated as κ′ ≈ κ/ng and κeκe/ng. According to the input–output relationship of the optical cavity,50,51 the probe field output operator can be expressed as
(4)
where Δ = ωcωp and ωc is the cavity frequency. In the case of rotation, the rotation of the polarization state is caused by the refractive index difference between LCP and RCP light.42 The refractive index difference between LCP and RCP light is given as
(5)
where nϕ is the phase refractive index, Ω is the medium rotation rate, and nr and nl are the refractive indices for RCP and LCP light, respectively. Based on the resonance condition of the cavity, the phase change can be expressed as
(6)
where η = L/L0 represents the duty ratio, and L0 and L are the cavity length and medium length, respectively. At this point, the probe field output operator can be approximated as
(7)
where ξ = 2κe/κ. Subsequently, the output light from the FP cavity is split into the HP and VP components after passing through the PBS. Using the relationship between circularly and linearly polarized light, we obtain the output of the VP component is
(8)
where Δϕ=Δω+Δω/2κ and êin=êin,1+êin,2/2 is the quantum fluctuation. The output power of the VP component is PoutV=ωpâinVâinV. We define the VP component output photon number operator as
(9)
Based on the error propagation analysis,30,52 the rotation sensitivity is defined as
(10)
where
(11a)
(11b)
with Nin=|ainH|2 being the input probe photon flux. According to the above formula, our system sensitivity ΔΩ is given by
(12)

The gyro sensitivity relies on the input photon flux (Nin), the group refractive index (ng), the duty ratio (η), and the decay rate (κ). In order to obtain optimal sensitivity, the optimal operation is discussed next.

The performance of our gyro crucially relies on the power of the input probe light. Figure 2 shows the photon flux of the output VP component and the gyroscope sensitivity. According to the photon drag effect, the deflection phase caused by a rotating medium is proportional to the group refractive index ng.53 Thus, the giant group refractive index induced in the Er3+-doped glass rod can essentially enhance the polarization rotation of light due to the photon drag,43 greatly amplifying the output power to a detectable level. Figure 2(a) shows the output photon flux as a function of the probe power and the rotation rate. Taking the experimentally accessible values for parameters R = 0.998, L = 10 cm, L0 = 20 cm, and η = 0.5, the maximum output photon flux of the VP component can reach PoutV/ωp1011photons/s(13nW) when Ω/Ωe ≈ 10−9. In this case, the total transmitted photon flux is approximately Pout/ℏωp ≈ 5.5 × 1016 photons/s (≈7 mW). This level of photon flux can be well detected experimentally. The peak output photon flux is attainable for Ω/Ωe < 1.5 × 10−9. The required input power is less than Pp,in/ℏωp < 4 × 1017 photons/s, corresponding to a probe field power <52mW. The performance of our gyroscope is determined by the duty ratio η but not the cavity length L0. Throughout the following investigation, we fix L0 = 20 cm.

FIG. 2.

(a) Photon flux (color bar) of the VP transmitted light as a function of the rotation rate of the rod and the input photon flux. (b) Sensitivity of the gyroscope as a function of the input photon flux. Other parameters as in Fig. 1 except for R = 0.998.

FIG. 2.

(a) Photon flux (color bar) of the VP transmitted light as a function of the rotation rate of the rod and the input photon flux. (b) Sensitivity of the gyroscope as a function of the input photon flux. Other parameters as in Fig. 1 except for R = 0.998.

Close modal

According to Eq. (12), it can be seen that the gyro sensitivity is proportional to κ and inversely proportional to η, ng2, and Nin. Therefore, the sensitivity tends to be improved as the input power increases, as shown in Fig. 2(b), but reaches the optimal value of ΔΩmin=26frad/s/Hz when the input photon flux Pp,in/ℏωp ≈ 2 × 1017 photons/s (≈26 mW), and then reduces. The reason is that ng first increases rapidly and then decreases after reaching the maximum value as the input power increases; see Figs. 1(b) and 1(c). It implies that to obtain the best sensitivity, the gyro needs to operate at the optimal input power.

According to Fig. 1(b), we are interested in R = 0.99–0.999, which allows us to achieve the maximum ng with low control laser power. Figure 3(a) shows the gyro sensitivity vs the mirror reflectivity R and the probe power Pp,in. The gyro is most sensitive when R ≈ 0.998 and Pp,in/ℏωp ≈ 2 × 1017 photons/s; see Fig. 3. Figure 3(b) clearly shows the obtained optimal sensitivity of ΔΩmin26frad/s/Hz at R ≈ 0.998. A higher reflectivity of the cavity mirrors not only leads to a lower cavity loss54 but also significantly affects ng, as shown in Fig. 1(b). As a trade-off, the mirror reflectivity is optimal at R ≈ 0.998. Note that the slow-light enhances ng and, thus, the photon drag effect by a factor of 10643 and significantly narrows the intracavity linewidth.44 Thanks to this double improvement, the gyro sensitivity can reach nine orders of magnitude higher than Earth’s rotation rate, enabling precise detection of tiny variations of the Earth’s rotation and its wobbling.35 

FIG. 3.

(a) Scaled sensitivity (color bar, in a unit of frad/sHz) as a function of the mirror reflectivity and the probe photon flux. (b) Highest sensitivity obtained at different mirror reflectivity when changing the probe light power. Other parameters: L = 10 cm, L0 = 20 cm, η = 0.5, and κ ≈ 5 MHz.

FIG. 3.

(a) Scaled sensitivity (color bar, in a unit of frad/sHz) as a function of the mirror reflectivity and the probe photon flux. (b) Highest sensitivity obtained at different mirror reflectivity when changing the probe light power. Other parameters: L = 10 cm, L0 = 20 cm, η = 0.5, and κ ≈ 5 MHz.

Close modal

Figure 4 presents the dependence of the gyro sensitivity on the duty ratio η. For a given η, the gyro sensitivity first improves as the probe photon flux increases. After reaching an optimal point, its value increases. As predicted by Eq. (12), the sensitivity ΔΩ is inversely proportional to the duty ratio η. An example for Pp,in/ℏωp = 2 × 1017 photons/s is presented in Fig. 4(b). Obviously, a larger duty ratio yield is preferable for achieving high performance. However, considering practical experimental losses and operational, we chose a spinning medium length of L = 10 cm. On this basis, the gyro sensitivity can reach 26frad/s/Hz. This level of sensitivity is accurate enough to detect small variations in the Earth’s rotation rate and the effects of Lorentz violation and general relativity in gravity research.

FIG. 4.

(a) Sensitivity (color bar, in a unit of frad/sHz) as a function of the duty ratio and the probe photon flux. (b) Optimal sensitivity obtained at different duty ratios by tuning the probe power. Other parameters: L0 = 20 cm, R = 0.998, and κ ≈ 5 MHz.

FIG. 4.

(a) Sensitivity (color bar, in a unit of frad/sHz) as a function of the duty ratio and the probe photon flux. (b) Optimal sensitivity obtained at different duty ratios by tuning the probe power. Other parameters: L0 = 20 cm, R = 0.998, and κ ≈ 5 MHz.

Close modal

Figure 5 illustrates the relationship between gyro sensitivity and cavity length L0 with a fixed duty ratio (η = 0.5). By extending both the cavity length and medium length proportionally, sensitivity is further enhanced, reaching 4.3frad/s/Hz at a cavity length of L0 = 120 cm. This indicates that greater cavities and medium lengths are advantageous for achieving high sensitivity. However, the medium length is subject to technical limitations. Moreover, longer cavity lengths may affect stability in practical applications. There are still significant challenges in experimental implementation. Therefore, we chose a cavity length of L0 = 20 cm and a medium length of L = 10 cm to achieve a high-sensitivity gyroscope.

FIG. 5.

Optimal sensitivity by fixing the duty ratio but increasing the cavity length and the medium length. Other parameters: R = 0.998 and κ ≈ 5 MHz.

FIG. 5.

Optimal sensitivity by fixing the duty ratio but increasing the cavity length and the medium length. Other parameters: R = 0.998 and κ ≈ 5 MHz.

Close modal

Note that we measure the VP component of the transmitted field. Because the accumulated relative phase of the LCP and RCP fields is very small, this VP component is only a fraction of the total transmitted field. Most of the transmitted photons are HP. Moreover, some intracavity photons are dissipated into the environment and the cavity does not store many photons.

In this work, we use a FP cavity embedded with a spinning medium to construct a sensitive gyro. The medium is an Er3+-doped glass rod, and slow light can be obtained with a modulated laser beam. The length of the Er3+-doped glass rod can be set to L = 10 cm, for example. The refractive index of the glass rod is about n0 ≈ 1.48. According to Ref. 46, the relaxation time of the excited state is T1 = 10.5 ms, and the corresponding unsaturated absorption coefficient is α0 = 0.16 cm−1. We set the cavity mirror reflectivity of the FP cavity R = 0.998. The cavity decay rate is calculated to be κ ≈ 5 MHz. The input photon flux Pp,in/ℏωp ≈ 2 × 1017 photons/s (≈26 mW). This laser field at wavelength 1536nm is sinusoidally modulated by the function generator. We fix Pp/Ps = 0.08, which is experimentally accessible.49 We can obtain the output photon flux of the VP component is PoutV/ωp1011photons/s(13nW) and total transmitted photon flux is approximately Pout/ℏωp ≈ 5.5 × 1016 photons/s (≈7 mW). On this basis, the resulting gyro sensitivity can reach 26frad/s/Hz. Moreover, this method is also applicable to a ruby. Using an argon ion laser functioning at 514.5 nm, the population is pumped from the ground state to the wide absorption band of 4F2. Then, slow light based on spectral holes can be observed.47 Actually, the slow-light enhanced photon drag effect using a ruby has been experimentally observed.43 Therefore, our theoretically proposed scheme can be implemented with an Er3+-doped glass rod or a ruby. To measure the modulated sideband light, the output VP component can be measured via the homodyne detection. In particular, the output VP light is first interfered with by a local light field to generate an electric signal oscillating at frequency δ. Then, this oscillating signal is filtered by an electronic circuit and measured.

Vibration of the Earth’s rotation rate is currently measured mainly by using a large ring-laser gyroscope or fiber optic gyroscopes based on the Sagnac effect. However, these gyroscopes are large, costly, and subject to errors induced by asymmetric perturbation and common-mode noises. Here, we theoretically propose to implement a table-top gyroscope with unprecedented high sensitivity up to 26frad/s/Hz by exploring the nonreciprocity of photon drag in a rotating medium. The proposed gyroscope is small in volume and, thus, cost-effective. The two circularly polarized components of the laser beam co-propagate in the system, therefore eliminating the effects of the asymmetric perturbation and common-mode noise. This work opens up the door for a portable gyroscope measuring the fluctuation of Earth’s motion.

Moreover, in the measurement, the photon dragging effect is proportional to the group refractive index of the rotating part. The group refractive index of the glass rod is enhanced by slow-light to an order of 106, far larger than other components of the system. Thus, we only need to consider the photon drag effect of the glass rod and can neglect the influence of other components. There are a few technical challenges in the implementation of our gyroscope. Inserting a solid slow-light medium in a high-quality cavity can be difficult. Nevertheless, slow-light in an Er3+-doped glass rod and a ruby has been experimentally demonstrated.43,47 Slow-light with atomic ensembles in a high-quality cavity has also been realized.55 Second, the proposed gyroscope needs a stable laser, the cavity, and the ambient temperature. Despite these challenges, the concept of exploring the photon drag in a solid slow-light medium embedded in a stable, high-quality cavity for a gyroscope is experimentally feasible.

This work was supported by the National Natural Science Foundation of China (Grant Nos. 92365107 and 12305020), the National Key R&D Program of China (Grant No. 2019YFA0308700), the Program for Innovative Talents and Teams in Jiangsu (Grant No. JSSCTD202138), the China Postdoctoral Science Foundation (Grant No. 2023M731613), and the Jiangsu Funding Program for Excellent Postdoctoral Talent (Grant No. 2023ZB708). We thank the High Performance Computing Center of Nanjing University for allowing the numerical calculations on its blade cluster system.

The authors have no conflicts to disclose.

Min She: Conceptualization (equal); Investigation (equal); Methodology (equal); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal). Jiangshan Tang: Conceptualization (equal); Methodology (equal); Validation (equal); Writing – review & editing (equal). Keyu Xia: Conceptualization (lead); Methodology (equal); Project administration (equal); Supervision (equal); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

Based on the CPO in a two-level system,47,48 we can derive the density matrix equation for the Er3+-doped glass rod. Er3+ can be excited by a laser beam with wavelength λ ≈ 1536 nm from the ground state 4I15/2 to the substable state 4I13/2 as a two-level system.45,46 Due to the long lifetime of metastable level 4I13/2 of Er3+, the system exhibits a capacity for rapid nonlinear response within a narrow spectrum. When the control and probe fields with a small frequency difference are input simultaneously, the medium exhibits the phenomenon of saturable absorption. In this situation, the number of ground state particles undergoes periodic oscillations, resulting in spectral hole burning in the absorption spectrum. At the burning point, the refractive index of light rapidly changes. According to the Kramers–Kronig relationship, the group velocity also undergoes drastic changes.

For analyzing the slow light, we denote the ground state level as g and the metastable level as e. According to the calculation steps in Ref. 48, the density matrix in a frame rotating at frequency ωs is given by
(A1a)
(A1b)
where T1 is the relaxation time of the excited state, T2 is the decoherence time of off-diagonal elements in the density matrix, and ρeeρgg0 is the inversion of particle number in thermal equilibrium state. Veg is the interaction energy, given by Veg=μegEs+Epeiδt, with δ = ωpωs, where Es and Ep are the amplitudes of the control and probe light, ωs and ωp are the frequencies of the control light and probe light, respectively, and μeg is the dipole matrix element. When 1/T2 is larger than both the detuning frequency δ and the Rabi frequency of the control laser Ωs=2|μeg||Es|/, the diagonal elements of the density matrix can be expressed as
(A2)
Following Ref. 48, the overall particle number inversion is given by
(A3)
where ρeeρggdc is the direct current component of the overall inversion, and ρeeρgg±δ are the amplitude components oscillating at δ. When δ ≤ 1/T1, particle number oscillations are evident. We consider the resonance case: ωeg = ωs. The response at the probe light frequency can be expressed as
(A4)
where β=1/T11+Ωs2T1T2. In the two-level system, the linear susceptibility is given by χδ=Nμegρegωp/Ep. The refractive index and absorption at the hole burning location are, respectively, given by
(A5)
(A6)
where I is the intensity of input power, Isat is the saturated absorption intensity, and α0 is the unsaturated absorption coefficient. Using the formula ng = n(δ) + ωcdn/, the group refractive index ng can be calculated as47 
(A7)
where n0 is the refractive index of the host medium at δ = 0 when only the control light passes through.
1.
M.
Cerdonio
,
G. A.
Prodi
, and
S.
Vitale
, “
Dragging of inertial frames by the rotating Earth: Proposal and feasibility for a ground-based detection
,”
Gen. Relativ. Gravitation
20
,
83
(
1988
).
2.
A. D. V.
Di Virgilio
,
A.
Basti
,
N.
Beverini
,
F.
Bosi
,
G.
Carelli
,
D.
Ciampini
,
F.
Fuso
,
U.
Giacomelli
,
E.
Maccioni
,
P.
Marsili
,
A.
Ortolan
,
A.
Porzio
,
A.
Simonelli
, and
G.
Terreni
, “
Underground Sagnac gyroscope with sub-prad/s rotation rate sensitivity: Toward general relativity tests on Earth
,”
Phys. Rev. Res.
2
,
032069
(
2020
).
3.
J.
Wang
,
Q.
Zhang
, and
H.
Jing
, “
Quantum weak torque sensing with squeezed optomechanics
,”
Phys. Rev. A
110
,
033505
(
2024
).
4.
I. P.
Prikhodko
,
S. A.
Zotov
,
A. A.
Trusov
, and
A. M.
Shkel
, “
What is MEMS gyrocompassing? Comparative analysis of maytagging and carouseling
,”
J. Microelectromech. Syst.
22
,
1257
(
2013
).
5.
H.
Xu
,
L.
Wang
,
Y.
Zu
,
W.
Gou
, and
Y.
Hu
, “
Application and development of fiber optic gyroscope inertial navigation system in underground space
,”
Sensors
23
,
5627
(
2023
).
6.
L.
Zhang
,
W.
Gao
,
Q.
Li
,
R.
Li
,
Z.
Yao
, and
S.
Lu
, “
A novel monitoring navigation method for cold atom interference gyroscope
,”
Sensors
19
,
222
(
2019
).
7.
J.
Ban
,
G.
Chen
,
Y.
Meng
, and
J.
Shu
, “
Calibration method for misalignment angles of a fiber optic gyroscope in single-axis rotational inertial navigation systems
,”
Opt. Express
30
,
6487
(
2022
).
8.
J. M.
Silver
,
L.
Del Bino
,
M. T. M.
Woodley
,
G. N.
Ghalanos
,
A.
Svela
,
N.
Moroney
,
S. Y.
Zhang
,
K. T. V.
Grattan
, and
P.
Del’Haye
, “
Nonlinear enhanced microresonator gyroscope
,”
Optica
8
,
1219
(
2021
).
9.
H. L.
Zhang
,
M. Y.
Peng
,
X. W.
Xu
, and
H.
Jing
, “
Anti-PT-symmetric Kerr gyroscope
,”
Chin. Phys. B
31
,
014215
(
2022
).
10.
D.
He
,
Y.
Cao
,
T.
Zhou
,
C.
Peng
, and
Z.
Li
, “
Sensitivity enhancement through RIN suppression in dual-polarization fiber optic gyroscopes for rotational seismology
,”
Opt. Express
28
,
34717
(
2020
).
11.
Y. H.
Lai
,
M. G.
Suh
,
Y. K.
Lu
,
B.
Shen
,
Q. F.
Yang
,
h.
Wang
,
J.
Li
,
S. H.
Lee
,
K. Y.
Yang
, and
K.
Vahala
, “
Earth rotation measured by a chip-scale ring laser gyroscope
,”
Nat. Photonics
14
,
345
(
2020
).
12.
D.
Guo
,
C.
Ren
,
L.
Zhang
, and
F.
Lu
, “
Research on improving the north-seeking precision of MEMS gyroscope
,”
Meas. Sci. Technol.
34
,
027005
(
2022
).
13.
M. S.
Weinberg
and
A.
Kourepenis
, “
Error sources in in-plane silicon tuning-fork MEMS gyroscopes
,”
J. Microelectromech. Syst.
15
,
479
(
2006
).
14.
T.
Tsukamoto
and
S.
Tanaka
, “
Fully differential single resonator FM gyroscope using CW/CCW mode separator
,”
J. Microelectromech. Syst.
27
,
985
(
2018
).
15.
M. M.
Torunbalci
,
S. E.
Alper
, and
T.
Akin
, “
Wafer level hermetic sealing of MEMS devices with vertical feedthroughs using anodic bonding
,”
Sens. Actuators, A
224
,
169
(
2015
).
16.
P.
Minotti
,
S.
Dellea
,
G.
Mussi
,
A.
Bonfanti
,
S.
Facchinetti
,
A.
Tocchio
,
V.
Zega
,
C.
Comi
,
A. L.
Lacaita
, and
G.
Langfelder
, “
High scale-factor stability frequency-modulated MEMS gyroscope: 3-axis sensor and integrated electronics design
,”
IEEE Trans. Ind. Electron.
65
,
5040
(
2018
).
17.
F.
Xu
,
W.
Wang
,
X.
Shao
,
X.
Liu
, and
Y.
Liang
, “
Optimization of surface acoustic wave-based rate sensors
,”
Sensors
15
,
25761
(
2015
).
18.
M.
Mahmoud
,
A.
Mahmoud
,
L.
Cai
,
M.
Khan
,
T.
Mukherjee
,
J.
Bain
, and
G.
Piazza
, “
Novel on chip rotation detection based on the acousto-optic effect in surface acoustic wave gyroscopes
,”
Opt. Express
26
,
25060
(
2018
).
19.
W.
Wang
,
W.
Wang
,
J.
Liu
,
M.
Liu
, and
S.
Yang
, “
Wireless and passive gyroscope based on surface acoustic wave gyroscopic effect
,”
Appl. Phys. Express
4
,
086601
(
2011
).
20.
L.
Tian
,
Q.
Shen
, and
H. L.
Chang
, “
A novel progressive wave gyroscope based on acousto-optic effects
,”
Microsyst. Nanoeng.
8
,
95
(
2022
).
21.
A.
Gebauer
,
M.
Tercjak
,
K. U.
Schreiber
,
H.
Igel
,
J.
Kodet
,
U.
Hugentobler
,
J.
Wassermann
,
F.
Bernauer
,
C.-J.
Lin
,
S.
Donner
,
S.
Egdorf
,
A.
Simonelli
, and
J. P. R.
Wells
, “
Reconstruction of the instantaneous earth rotation vector with sub-arcsecond resolution using a large scale ring laser array
,”
Phys. Rev. Lett.
125
,
033605
(
2020
).
22.
M. P.
Hokmabadi
,
A.
Schumer
,
D. N.
Christodoulides
, and
M.
Khajavikhan
, “
Non-Hermitian ring laser gyroscopes with enhanced Sagnac sensitivity
,”
Nature
576
,
70
(
2019
).
23.
K.
Liu
,
F. L.
Zhang
,
Z. Y.
Li
,
X. H.
Feng
,
K.
Li
,
Z. H.
Lu
,
K. U.
Schreiber
,
J.
Luo
, and
J.
Zhang
, “
Large-scale passive laser gyroscope for earth rotation sensing
,”
Opt. Lett.
44
,
2732
(
2019
).
24.
H. M.
Wang
,
Y. H.
Lai
,
Z. Q.
Yuan
,
M. G.
Suh
, and
K.
Vahala
, “
Petermann-factor sensitivity limit near an exceptional point in a Brillouin ring laser gyroscope
,”
Nat. Commun.
11
,
1610
(
2020
).
25.
Y. H.
Lai
,
Y. K.
Lu
,
M. G.
Suh
,
Z. Q.
Yuan
, and
K.
Vahala
, “
Observation of the exceptional-point-enhanced Sagnac effect
,”
Nature
576
,
65
(
2019
).
26.
S.
Liu
,
J.
Hu
,
Y.
Wang
,
Q.
Liu
,
H.
Ma
, and
Z.
He
, “
Navigation-grade three-axis resonant fiber-optic gyroscope employing a multiplexed source
,”
Opt. Lett.
48
,
2639
(
2023
).
27.
X. S.
Yao
,
H.
Xuan
,
X.
Chen
,
H.
Zou
,
X.
Liu
, and
X.
Zhao
, “
Polarimetry fiber optic gyroscope
,”
Opt. Express
27
,
19984
(
2019
).
28.
A.
Yertutanol
,
Ö.
Akçaalan
,
S.
Öğüt
,
E.
Özbay
, and
A.
Ceylan
, “
Fiber-optic gyroscope for the suppression of a Faraday-effect-induced bias error
,”
Opt. Lett.
46
,
4328
(
2021
).
29.
E. R.
Moan
,
R. A.
Horne
,
T.
Arpornthip
,
Z.
Luo
,
A. J.
Fallon
,
S. J.
Berl
, and
C. A.
Sackett
, “
Quantum rotation sensing with dual Sagnac interferometers in an atom-optical waveguide
,”
Phys. Rev. Lett.
124
,
120403
(
2020
).
30.
Y.
Wu
,
J.
Guo
,
X.
Feng
,
L. Q.
Chen
,
C.-H.
Yuan
, and
W.
Zhang
, “
Atom-light hybrid quantum gyroscope
,”
Phys. Rev. Appl.
14
,
064023
(
2020
).
31.
P. P.
Khial
,
A. D.
White
, and
A.
Hajimiri
, “
Nanophotonic optical gyroscope with reciprocal sensitivity enhancement
,”
Nat. Photonics
12
,
671
(
2018
).
32.
G.
Sagnac
, “
The luminiferous ether demonstrated by the effect of the relative motion of the ether in an interferometer in uniform motion
,”
C. R. Acad. Sci.
157
,
708
(
1913
).
33.
H.
Igel
,
K. U.
Schreiber
,
A.
Gebauer
,
F.
Bernauer
,
S.
Egdorf
,
A.
Simonelli
,
C.-J.
Lin
,
J.
Wassermann
,
S.
Donner
,
C.
Hadziioannou
,
S.
Yuan
,
A.
Brotzer
,
J.
Kodet
,
T.
Tanimoto
,
U.
Hugentobler
, and
J. P. R.
Wells
, “
ROMY: A multicomponent ring laser for geodesy and geophysics
,”
Geophys. J. Int.
225
,
684
(
2021
).
34.
A. D. V.
Di Virgilio
,
F.
Bajardi
,
A.
Basti
,
N.
Beverini
,
G.
Carelli
,
D.
Ciampini
,
G.
Di Somma
,
F.
Fuso
,
E.
Maccioni
,
P.
Marsili
,
A.
Ortolan
,
A.
Porzio
, and
D.
Vitali
, “
Noise level of a ring laser gyroscope in the femto-rad/s range
,”
Phys. Rev. Lett.
133
,
013601
(
2024
).
35.
K. U.
Schreiber
,
J.
Kodet
,
U.
Hugentobler
,
T.
Klügel
, and
J. P. R.
Wells
, “
Variations in the Earth’s rotation rate measured with a ring laser interferometer
,”
Nat. Photonics
17
,
1054
(
2023
).
36.
R. B.
Hurst
,
G. E.
Stedman
,
K. U.
Schreiber
,
R. J.
Thirkettle
,
R. D.
Graham
,
N.
Rabeendran
, and
J. P. R.
Wells
, “
Experiments with an 834 m2 ring laser interferometer
,”
J. Appl. Phys.
105
,
113115
(
2009
).
37.
A. D. V.
Di Virgilio
,
G.
Terreni
,
A.
Basti
,
N.
Beverini
,
G.
Carelli
,
D.
Ciampini
,
F.
Fuso
,
E.
Maccioni
,
P.
Marsili
,
J.
Kodet
, and
K. U.
Schreiber
, “
Overcoming 1 part in 109 of earth angular rotation rate measurement with the G Wettzell data
,”
Eur. Phys. J. C
82
,
824
(
2022
).
38.
Y. W.
Cao
,
X. D.
Ma
,
Y. J.
Chen
,
H. M.
Huang
,
L. X.
Zhu
,
W. B.
Wang
,
Y.
He
, and
Z. B.
Li
, “
Interferometric fiber-optic gyroscope based on mode-division multiplexing
,”
Opt. Lett.
48
,
3067
(
2023
).
39.
S.
Zhao
,
Y. L.
Zhou
, and
X. W.
Shu
, “
Analysis of fiber optic gyroscope dynamic error based on CEEMDAN
,”
Opt. Fiber Technol.
69
,
102835
(
2022
).
40.
I. A.
Andronova
and
G. B.
Malykin
, “
Physical problems of fiber gyroscopy based on the Sagnac effect
,”
Phys.-Usp.
45
,
793
(
2002
).
41.
E.
Hand
, “
Lord of the rings
,”
Science
356
,
236
(
2017
).
42.
M. A.
Player
, “
On the dragging of the plane of polarization of light propagating in a rotating medium
,”
Proc. R. Soc. London, Ser. A
349
,
441
(
1976
).
43.
S.
Franke-Arnold
,
G.
Gibson
,
R. W.
Boyd
, and
M. J.
Padgett
, “
Rotary photon drag enhanced by a slow-light medium
,”
Science
333
,
65
(
2011
).
44.
M.
Sabooni
,
Q.
Li
,
L.
Rippe
,
R. K.
Mohan
, and
S.
Kröll
, “
Spectral engineering of slow light, cavity line narrowing, and pulse compression
,”
Phys. Rev. Lett.
111
,
183602
(
2013
).
45.
F.
Arrieta-Yáñez
,
E.
Cabrera-Granado
,
J. M.
Ezquerro
,
O. G.
Calderón
, and
S.
Melle
, “
Pulse-width-dependent subluminal and superluminal propagation in highly doped erbium fibers
,”
J. Opt. Soc. Am. B
28
,
1172
(
2011
).
46.
S.
Melle
,
O. G.
Calderón
,
F.
Carreño
,
E.
Cabrera
,
M. A.
Antón
, and
S.
Jarabo
, “
Effect of ion concentration on slow light propagation in highly doped erbium fibers
,”
Opt. Commun.
279
,
53
(
2007
).
47.
M. S.
Bigelow
,
N. N.
Lepeshkin
, and
R. W.
Boyd
, “
Observation of ultraslow light propagation in a ruby crystal at room temperature
,”
Phys. Rev. Lett.
90
,
113903
(
2003
).
48.
R. W.
Boyd
,
M. G.
Raymer
,
P.
Narum
, and
D. J.
Harter
, “
Four-wave parametric interactions in a strongly driven two-level system
,”
Phys. Rev. A
24
,
411
(
1981
).
49.
A.
Schweinsberg
,
N. N.
Lepeshkin
,
M. S.
Bigelow
,
R. W.
Boyd
, and
S.
Jarabo
, “
Observation of superluminal and slow light propagation in erbium-doped optical fiber
,”
Europhys. Lett.
73
,
218
(
2006
).
50.
C. W.
Gardiner
and
M. J.
Collett
, “
Input and output in damped quantum systems: Quantum stochastic differential equations and the master equation
,”
Phys. Rev. A
31
,
3761
(
1985
).
51.
K.
Xia
and
J.
Twamley
, “
All-optical switching and router via the direct quantum control of coupling between cavity modes
,”
Phys. Rev. X
3
,
031013
(
2013
).
52.
Z.-D.
Chen
,
C.-H.
Yuan
,
H.-M.
Ma
,
D.
Li
,
L. Q.
Chen
,
Z. Y.
Ou
, and
W.
Zhang
, “
Effects of losses in the atom-light hybrid SU(1,1) interferometer
,”
Opt. Express
24
,
17766
(
2016
).
53.
J.
Leach
,
A. J.
Wright
,
J. B.
Götte
,
J. M.
Girkin
,
L.
Allen
,
S.
Franke-Arnold
,
S. M.
Barnett
, and
M. J.
Padgett
, “
“Aether drag” and moving images
,”
Phys. Rev. Lett.
100
,
153902
(
2008
).
54.
L.-F.
Tian
,
Y.-X.
Kuang
,
Z.-C.
Fan
, and
Z.-Y.
Li
, “
Low insertion loss silicon-based spatial light modulator with high reflective materials outside Fabry–Pérot cavity
,”
Chin. Phys. B
28
,
104209
(
2019
).
55.
Y.-J.
Guo
and
W.-J.
Nie
, “
Vacuum induced transparency and slow light phenomena in a two-level atomic ensemble controlled by a cavity
,”
Chin. Phys. B
24
,
094205
(
2015
).