• Photonics Research
  • Vol. 11, Issue 2, 150 (2023)
Lin Jiao and Jun-Hong An*
Author Affiliations
  • Lanzhou Center for Theoretical Physics, Key Laboratory of Theoretical Physics of Gansu Province, Lanzhou University, Lanzhou 730000, China
  • show less
    DOI: 10.1364/PRJ.469779 Cite this Article Set citation alerts
    Lin Jiao, Jun-Hong An. Noisy quantum gyroscope[J]. Photonics Research, 2023, 11(2): 150 Copy Citation Text show less

    Abstract

    Gyroscope for rotation sensing plays a key role in inertial navigation systems. Developing more precise gyroscopes than the conventional ones bounded by the classical shot-noise limit by using quantum resources has attracted much attention. However, existing quantum gyroscope schemes suffer severe deterioration under the influence of decoherence, which is called the no-go theorem of noisy metrology. Here, by using two quantized optical fields as the quantum probe, we propose a quantum gyroscope scheme breaking through the constraint of the no-go theorem. Our exact analysis of the non-Markovian noise reveals that both the evolution time as a resource in enhancing the sensitivity and the achieved super-Heisenberg limit in the noiseless case are asymptotically recoverable when each optical field forms a bound state with its environment. The result provides a guideline for realizing high-precision rotation sensing in realistic noisy environments.

    1. INTRODUCTION

    High-performance gyroscopes for rotation sensing are of pivotal significance for navigation in many types of air, ground, marine, and space applications. Based on the Sagnac effect, i.e., two counter-propagating waves in a rotating loop accumulate a rotation-dependent phase difference, gyroscopes have been realized in optical [16] and matter–wave [714] systems. However, the precision of a purely Sagnac gyroscope, which is proportional to the surface area enclosed by the optical path, is theoretically limited by the classical shot-noise limit (SNL). It dramatically constrains their practical application and further performance improvement. The records for precision and stability of commercial gyroscopes are held by optical gyroscopes [15,16]. To reduce the noise effect, the practical operation of fiber optical gyroscopes generally modulates the optical signal and measures the ratios of harmonics instead of the phase difference [17], where the classical SNL model is not widely used. Building a purely Sagnac optical gyroscope beating the SNL from the fundamental principle is highly desired.

    Pursuing more precise measurement to physical quantities than the classical SNL by using quantum resources [1822], such as squeezing [2325] and entanglement [26,27], quantum metrology supplies a way toward achieving gyroscopes with ultimate sensitivity limits. Based on this idea, many schemes of quantum gyroscopes have been proposed. It was found that the entanglement in N00N states [28,29], continuous-variable squeezing [3032], and optical nonlinearity [33] can enhance the sensitivity of optical gyroscopes beyond the SNL. A quantum-enhanced sensitivity can also be achieved in matter–wave gyroscopes [3437] by using spin squeezing [3840] or entanglement. However, quantum gyroscopes are still at the stage of the proof-of-principle study, and their superiority over the conventional ones in the absolute value of sensitivity still has not been exhibited [22,36]. One key obstacle is that the stability of the quantum gyroscope is challenged by the decoherence caused by inevitable noise in microscopic world, which generally makes the quantum resources degraded. It was found that the metrology sensitivity using entanglement [4143] and squeezing [44,45] exclusively returns to or even becomes worse than the SNL; thus, their quantum superiority completely disappears when the photon loss is considered. This is called the no-go theorem of noisy quantum metrology [46,47] and is one difficulty in achieving a high-precision quantum gyroscope in practice.

    In this paper, we propose a scheme of quantum gyroscope and discover a mechanism to overcome the constraint of the no-go theorem on our scheme. A super-Heisenberg limit (HL) on the sensitivity is achieved in the ideal case by using two-mode squeezed vacuum state. Our exact analysis on the photon dissipation reveals that the performance of the quantum gyroscope in the dissipative environments intrinsically depends on the energy-spectrum feature of the total system formed by the probe and its environments. The encoding time as resource and the super-HL in the ideal case are asymptotically recovered when each optical field forms a bound state with its environment, which means that the no-go theorem is efficiently avoided. It supplies a guideline to engineer the optimal working condition of our quantum gyroscope in dissipative environments.

    2. IDEAL QUANTUM GYROSCOPE SCHEME

    To measure a physical quantity of certain systems, three processes, i.e., the initialization of the quantum probe, the quantity encoding via the probe-system coupling, and the measurement, are generally required. In our quantum gyroscope, we choose two beams of quantized optical fields as the quantum probe. They propagate in opposite directions and are input into a 50:50 beam splitter and split into clockwise and counterclockwise prorogating beams [see Fig. 1(a)]. The setup rotates with an angular velocity Ω about the axis perpendicular to its plane. Thus, the two beams accumulate a phase difference Δθ=N4πkR2Ω/c when they reencounter the beam splitter after N rounds of propagation in the circular path [48]. Here k is the wave vector, c is the speed of light, and R is the radius of the quantum gyroscope. Remembering the standing-wave condition kR=n (nZ) of the optical fields propagating along the circular path and defining ΔtN2πR/c, we have ΔωΔθ/Δt=2nΩ. Therefore, the quantum gyroscope can be equivalently treated as two counter-propagating optical fields with a frequency difference Δω along the circular path. For concreteness, we choose the basic mode n=1. Then the optical fields in the quantum gyroscope can be quantum mechanically described by (=1) [49] H^S=ω0l=1,2a^la^l+Ω(a^1a^1a^2a^2),where a^l is the annihilation operator of the lth field with frequency ω0. The optical fields couple to the beam splitter twice and output in the state |Ψout=V^U^0(Ω,t)V^|Ψin, where U^0(Ω,t)=exp(iH^St) is the evolution operator of the fields and V^=exp[iπ4(a^1a^2+a^2a^1)] describes the action of the beam splitter. Thus, the angular velocity Ω is encoded into the state |Ψout of the optical probe via the unitary evolution.

    (a) Schematic diagram of quantum gyroscope. (b) Evolution of the error δΩBA (cyan solid line) in the presence of photon loss under the Born–Markovian approximation. The blue dashed line is the local minima of δΩBA. The global minimum is marked by the red dot. (c) Numerical fitting reveals that the global minimum scales with the photon number as minδΩBA=5.4κN−0.23. The parameters are N=100, Ω=ω0, and κ=0.2ω0.

    Figure 1.(a) Schematic diagram of quantum gyroscope. (b) Evolution of the error δΩBA (cyan solid line) in the presence of photon loss under the Born–Markovian approximation. The blue dashed line is the local minima of δΩBA. The global minimum is marked by the red dot. (c) Numerical fitting reveals that the global minimum scales with the photon number as minδΩBA=5.4κN0.23. The parameters are N=100, Ω=ω0, and κ=0.2ω0.

    To exhibit the quantum superiority, we employ two-mode squeezed vacuum state as the input state |Ψin=S^|0,0, where S^=exp[r(a^1a^2a^1a^2)] is the squeezing operator, with r being the squeezing parameter. The total photon number of this input state is N=2sinh2r, which is the quantum resource of our scheme. The parity operator Π^=exp(iπa^1a^1) is measured at the output port [50]. To the output state |Ψout, we can calculate Π¯Ψout|Π^|Ψout=[1+N(2+N)cos2(2Ωt)]1/2 and δΠ=(1Π¯2)1/2, where Π^2=1 has been used. Then the sensitivity of sensing Ω can be evaluated via the error propagation formula δΩ=δΠ|ΩΠ¯| as minδΩ=[2tN(2+N)]1,when Ωt=(2n+1)π/4 with nZ. It is remarkable to find that the best sensing error achieved in our scheme is even smaller than the HL ΔΩ(tN)1, which reflects the quantum superiority of the used squeezing and measured observable in our scheme. It can be verified that this measurement scheme saturates the Cramér–Rao bound governed by quantum Fisher information. We call such a sensitivity surpassing the HL the super-HL [51,52]. It is noted that a phase estimation error smaller than the inverse of the mean photon number was called the sub-HL in Refs. [50,53,54]. The outstanding performance of quantum squeezing has been found in gravitational wave detection [55].

    3. EFFECTS OF DISSIPATIVE ENVIRONMENTS

    The superiority of the quantum sensor is challenged by the decoherence of the quantum probe due to the inevitable interactions with its environment. Depending on whether the probe has energy exchange with the environment or not, the decoherence can be classified into dissipation and dephasing. The main decoherence in our quantum gyroscope is the photon dissipation. The previous works phenomenologically treat the photon dissipation by introducing an imperfect transmission to the beam splitter [41,44,45,56,57], which is equivalent to the Born–Markovian approximate description. Such an approximation is convenient, but it might miss important physics. It has been found that the system–environment interplay caused by the inherent non-Markovian nature would induce diverse characters absent in the Born–Markovian approximation [5862]. To reveal the practical performance of our quantum gyroscope, we, going beyond the Born–Markovian approximation and paying special attention to the non-Markovian effect, investigate the impact of the photon dissipation on the scheme.

    We consider that the encoding process is influenced by the photon dissipation, which is caused by the energy exchange between the two optical fields and two independent environments [48]. The Hamiltonian of the total system is H^=H^S+l=1,2k[ωk,lb^k,lb^k,l+gk,l(a^lb^k,l+H.c.)],where b^k,l is the annihilation operator of the kth mode with frequency ωk,l of the environment felt by the lth optical field and gk,l is their coupling strength. The coupling is further characterized by the spectral density Jl(ω)=kgk,l2δ(ωωk) in the continuous limit of the environmental frequencies. We consider the Ohmic-family spectral density J1(ω)=J2(ω)J(ω)=ηωsωc1seω/ωc for both environments, where η is a dimensionless coupling constant, ωc is a cutoff frequency, and s is an Ohmicity index. Under the condition that the environments are initially in the vacuum state, we can derive an exact master equation for the encoding process using the Feynman–Vernon influence functional method [63], ρ˙(t)=l=1,2{iϖl(t)[a^la^l,ρ(t)]+γl(t)D^lρ(t)},where D^l·=2a^l·a^la^la^l··a^la^l is the Lindblad superoperator, ϖl(t)=Im[u˙l(t)/ul(t)] is the renormalized frequency, and γl(t)=Re[u˙l(t)/ul(t)] is the dissipation rate. The time-dependent functions ul(t) satisfy u˙l(t)+iωlul(t)+0tf(tτ)ul(τ)dτ=0,under ul(0)=1, where ω1,2=ω0±Ω and f(x)=0J(ω)eiωxdω is the environmental correlation function. Equation (4) indicates that all the non-Markovian effects induced by the environmental backactions have been incorporated into these time-dependent coefficients self-consistently. Solving Eq. (4), we obtain (see Appendix A) Π¯(t)=x[4m1(m2*m1*p22)+4m2(m1*m2*p12)+(1p1p2)2+16|m1m2|2]1/2,where x=(A1A2cosh2r)1, ml=iul(t)2tanhr2Al, and pl=|ul(t)|2(1Al1), with Al=1[|ul(t)|21]2tanh2r. The analytical form of δΩ can then be calculated in a similar manner as the ideal case.

    In the special case when the probe–environment coupling is weak and the time scale of f(tτ) is smaller than the typical time scale of the probe, we can apply the Born–Markovian approximation in Eq. (5) [41,57]. Their approximate solutions read ul,BA(t)=e{κl+i[ωl+Δ(ωl)]}t, with κl=πJ(ωl) and Δ(ωl)=P0J(ω)ω1ωdω. Substituting them into Eq. (6) and using the error propagation formula, we obtain (see Appendix B) δΩBA(t)=(2e2tκ+NCe2tκ)C8N(N+2)t|sin(4Ωt)|,where C=4e2tκ+N2+(N+2)cos(4Ωt). Here we have chosen κ1=κ2κ. We plot in Fig. 1(b) the evolution of δΩBA(t). It can be found that δΩBA(t) experiences an obvious oscillation with time. However, the best sensitivity manifested by the profile of its local minima tends to be divergent with time. Thus, being in sharp contrast to the ideal case in Eq. (2), the superiority of time as a resource in enhancing the precision of the quantum gyroscope disappears. After optimizing the encoding time, we obtain the global minimum δΩ [see the red dot in Fig. 1(b)]. The numerical fitting reveals minδΩ=5.4κN0.23 [see Fig. 1(c)], which is even worse than the SNL. Therefore, being consistent with the previous quantum sensing schemes [41,44,45,56,57], the photon dissipation under the Born–Markovian approximation makes the quantum advantages of our scheme completely vanish. It is called the no-go theorem of noisy quantum metrology [46,47] and is the main obstacle to achieve a high-precision quantum sensing in practice.

    In the general non-Markovian case, the analytical solution of Eq. (5) can be found by the method of Laplace transform, which converts Eq. (5) into u˜l(zl)=[zl+iωl+0J(ω)dωzl+iωdω]1. Then ul(t) is obtained by making the inverse Laplace transform to u˜l(zl), which can be done by finding its poles from Yl(El)ωl0J(ω)ωEldω=El,El=izl.Here, El are also the eigenenergies in the single-excitation subspace of the total systems formed by each optical field and its environment. To see this, we expand the eigenstate as |Φl=(xla^l+kyk,lb^k,l)|0,{0k,l}. From the stationary Schrödinger equation, we have [El(ω0±Ω)]xl=kgk,lyk,l and yk,l=gk,lxl/(Elωk,l) with El being the eigenenergies. The two equations readily result in Eq. (8) in the continuous limit of the environmental frequencies. It implies that the dissipation of the optical probe is intrinsically determined by the energy-spectrum character of the probe–environment system in the single-excitation subspace, even though the subspaces with any excitation numbers are involved. Due to Yl(El) being decreasing functions in the regime El<0, each of Eq. (8) has one isolated root Eb,l in this regime provided Yl(0)<0. While Yl(El) are ill-defined when El>0 due to the poles in the integrand, they have infinite roots in this regime, which form a continuous energy band. We call the eigenstates of the isolated eigenenergies Eb,l bound states [58]. Making the inverse Laplace transform, we obtain ul(t)=ZleiEb,lt+0Θ(E)eiEtdE, where Zl=[1+0J(ω)(Eb,lω)2dω]1 and Θ(E)=J(E)[EωlΔ(E)]2+[πJ(E)]2. The integral in ul(t) is from the energy band and tends to zero in the long-time limit due to the out-of-phase interference. Thus, when the bound state is formed, we have limtul(t)=ZleiEb,lt, characterizing the suppressed dissipation; otherwise, we have limtul(t)=0, meaning a complete dissipation. It can be determined that the bound state is formed for the Ohmic-family spectral density when ωl<ηωcΓ(s), where Γ(s) is the Euler’s Γ function.

    We have three parameter regimes where zero, one, and two bound states are formed, respectively. It is natural to expect that δΩ in the former two regimes is qualitatively consistent with the Born–Markovian approximate result Eq. (7) due to the complete dissipation in either two or one optical field. Focusing on the case in the presence of two bound states and substituting the asymptotic solution ZleiEb,lt into Eq. (6), we obtain (see Appendix C) limtδΩ(t)=F2F42N(2+N)Z12Z22t(Z1+Z2)|sin(2Gt)|,where F=2+NlZl2(2Zl2)+NZ12Z22[N+(2+N)·cos(2Gt)] with G=Eb,1Eb,2. We have used ΩEb,l=(1)l1Zl derived from Eq. (8). Equation (9) exhibits a t1-dependence on time, which is as perfect as the ideal result Eq. (2). Another finding from Eq. (9) is that it tends to the exactly same form as the ideal result Eq. (2) in the limit Zl tending to 1. Therefore, the formation of two bound states overcomes the problem of no-go theorem and asymptotically retrieves the ideal sensitivity.

    4. NUMERICAL RESULTS

    We now numerically verify our general result by choosing the Ohmic spectral density. Figure 2(a) shows the energy spectrum of the total system consisting of the optical fields and their environments. It can be seen that the two branches of bound states divide the energy spectrum into three regimes: without bound state when ωc<19.8ω0, one bound state when ωc(19.8,20.2)ω0, and two bound states when ωc>20.2ω0. The result confirms our analytical criterion that the bound states are formed when ωc>ωl/[ηΓ(s)]. Numerically solving Eq. (5) and using Eq. (6), we obtain the exact evolution of δΩ(t) in the three regimes. When no or one bound state is formed, the local minima profile of δΩ(t) tends to divergence in the long-time limit, and the quantum superiority of the scheme completely disappears [see Figs. 2(b) and 2(c)], which is qualitatively similar to the Born–Markovian result. However, as long as two bound states are formed, the profile of the local minima becomes a decreasing function of the encoding time. The matching of the numerical result with the long-time behavior Eq. (9) verifies the validity of the result in Eq. (9). Thus, the encoding time as a resource in sensing Ω is recovered as perfectly as the ideal case by the formation of the two bound states. Our above result also gives a direct proof on that whether the Born–Markovian approximation is applicable or not depends sensitively on the feature of the energy spectrum of the total probe–environment system. Whenever a bound state is formed in the energy spectrum, the decoherence would be suppressed, and the Born–Markovian approximation would no long be valid anymore. This result refreshes our general belief on applicability of the Born–Markovian approximation.

    (a) Energy spectrum of the total system formed by the two optical fields and their environments. Non-Markovian dynamical evolution of δΩ(t) multiplied by a magnification factor P when (ωc/ω0,P)=(2,10−1) in (b), (20, 10−2) in (c), and (25, 10−3) in (d). The blue dashed line in (d) is obtained by numerically solving Eq. (5), and the cyan solid line is obtained from the analytical form Eq. (9). We use s=1, η=0.05, Ω=10−2ω0, and N=100.

    Figure 2.(a) Energy spectrum of the total system formed by the two optical fields and their environments. Non-Markovian dynamical evolution of δΩ(t) multiplied by a magnification factor P when (ωc/ω0,P)=(2,101) in (b), (20,102) in (c), and (25,103) in (d). The blue dashed line in (d) is obtained by numerically solving Eq. (5), and the cyan solid line is obtained from the analytical form Eq. (9). We use s=1, η=0.05, Ω=102ω0, and N=100.

    Figure 3(a) shows the evolution of the local minima of δΩ(t) in Eq. (9) at different ωc when two bound states are formed. The formation of the bound state causes an abrupt increase of the corresponding |ul()| from zero to a finite value exactly matching with Zl [see the inset of Fig. 3(b)]. It is interesting to find that not only the encoding time as a resource is retrieved, but also the ideal precision is asymptotically recovered. This is double confirmed by the long-time behavior of minδΩ(t) as a function of the photon number N in Fig. 3(b). A similar performance is found by changing the coupling constant η (see Fig. 4). All the results demonstrate the constructive role played by the two bound states and the non-Markovian effect in retrieving the quantum superiority of our quantum gyroscope. It offers us a guideline to achieve a noise-tolerant rotation sensing by manipulating the formation of the bound states. It is noted that, according to the condition of forming the bound states, we see that what really matters is the relative value ωc/ω0. The equivalent result is achievable by tuning ω0 for given ωc and η.

    Local minima of δΩ(t) as a function of (a) time and (b) N when t=2.5×104ω0−1 at different ωc. Steady-state |ul(∞)| marked by dots, which match with Zl depicted by lines, and the energy spectrum are shown in the inset of (b). We use s=1, η=7×10−4, Ω=10−2ω0, and N=100.

    Figure 3.Local minima of δΩ(t) as a function of (a) time and (b) N when t=2.5×104ω01 at different ωc. Steady-state |ul()| marked by dots, which match with Zl depicted by lines, and the energy spectrum are shown in the inset of (b). We use s=1, η=7×104, Ω=102ω0, and N=100.

    Local minima of δΩ(t) as a function of (a) time and (b) N when t=2.5×104ω0−1 at different η. Steady-state |ul(∞)| marked by dots, which match with Zl depicted by lines, and the energy spectrum are shown in the inset of (b). We use s=1, ωc=5×103ω0, Ω=10−2ω0, and N=100.

    Figure 4.Local minima of δΩ(t) as a function of (a) time and (b) N when t=2.5×104ω01 at different η. Steady-state |ul()| marked by dots, which match with Zl depicted by lines, and the energy spectrum are shown in the inset of (b). We use s=1, ωc=5×103ω0, Ω=102ω0, and N=100.

    5. DISCUSSION AND CONCLUSION

    Our scheme is independent of the form of the spectral density. Although only the Ohmic form is considered, our scheme can be generalized to other spectra. Given the rich way of controlling the spectral density in the setting of quantum reservoir engineering [64,65], we deem that our scheme is realizable in state-of-the-art quantum-optical experiments. The Ohmic-family spectral densities for the electromagnetic noise are well cotrolled in circuit QED systems [6668]. Actually, the non-Markovian effect has been observed in the linear optical systems [69,70]. The bound state and its dynamical effect have been observed in circuit QED [71] and ultracold atom [72] systems. A squeezing parameter r2.5, which corresponds to N73, has been realized [73]. Inspired by these experimental achievements in circuit QED systems, we can design a realizable microwave Sagnac interferometer to test our result. We prepare two quantized optical fields propagating in opposite directions in a 1D transmission line, which is coupled via a capacitance or inductance to another wide-band transmission line acting as a structured environment [71]. The squeezed state of the two fields can be generated by the Josephson traveling-wave parametric amplifier [73]. When the fields reencounter after several rounds of rotation, a phase difference depending on the measured angular velocity is accumulated. The structured environments, on one hand, exert a strong non-Markovian effect on the dynamics of the fields and, on the other hand, protect the scheme from the photon loss according to our mechanism.

    In summary, we have proposed a quantum gyroscope scheme by using two quantized fields as the quantum probe, which achieves a super-HL sensitivity in measuring the angular velocity. However, the photon dissipation under the conventional Born–Markovian approximation forces this sensitivity to even be worse than the classical SNL. To overcome this problem, we have presented a mechanism to retrieve the ideal sensitivity by relaxing this approximation. It is found that the ideal sensitivity is asymptotically recoverable when each optical field forms a bound state with its environment, which can be realized by the technique of quantum reservoir engineering. Exhibiting the optimal working condition of the quantum gyroscope, our mechanism breaks through the constraint of the no-go theorem of noisy quantum metrology and supplies a guideline for developing high-precision rotation sensing for next-generation inertial navigation systems.

    APPENDIX A: EXPECTATION VALUE OF PARITY OPERATOR

    In this section, we give the derivation of Eq. (6). The Feynman and Vernon’s influence-functional theory enables us to derive the evolution of the reduced density matrix of the quantum probe formed by two quantized optical fields exactly. By expressing the forward and backward evolution operators of the density matrix of the probe and the environments as a double path integral in the coherent-state representation and performing the integration over the environmental degrees of freedom, we incorporate all the environmental effects on the probe in a functional integral named influence functional. The reduced density matrix fully describing the encoding dynamics of the probe is given by [63] ρ(α¯f,αf;t)=dμ(αi)dμ(αi)J(α¯f,αf;t|α¯i,αi;0)×ρ(α¯i,αi;0),where ρ(α¯f,αf;t)=α¯f|ρ(t)|αf is the reduced density matrix expressed in coherent-state representation and J(α¯f,αf;t|α¯i,αi;0) is the propagating function. In the derivation of Eq. (A1), we have used the coherent-state representation |α=l=12|αl with |αl=exp(αla^l)|0l, which are the eigenstates of annihilation operators, i.e., a^l|αl=αl|αl, and obey the resolution of identity dμ(α)|αα¯|=1 with dμ(α)=leα¯lαldα¯ldαl2πi. α¯ denotes the complex conjugate of α. The propagating function J(α¯f,αf;t|α¯i,αi;0) is expressed as the path integral governed by an effective action, which consists of the free actions of the forward and backward propagators of the optical probe and the influence functional obtained from the integration of environmental degrees of freedom. After evaluation of the path integral, its final form reads J(α¯f,αf;t|α¯i,αi;0)=exp{l=12{ul(t)α¯lfαli+u¯l(t)α¯liαlf+[1|ul(t)|2]α¯liαli}},where ul(t) satisfies u˙l(t)+iωlul(t)+0tf(tτ)ul(τ)=0,with f(x)0J(ω)eiωxdω and ul(0)=1. We have assumed that the spectral density of the two environments are identical.

    The input state of the probe is a two-mode squeezed vacuum state |Ψin=exp[r(a^1a^2a^1a^2)]|00, where r is the squeezing parameter. After passing the first beam splitter of the quantum optical gyroscope, the state changes into |Ψ(0)V^|Ψin, with V^=exp[iπ4(a^1a^2+a^2a^1)], which acts as the initial state of the encoding dynamics. In the coherent-state representation, this initial state is given by ρ(α¯i,αi;0)=1cosh2rexp[itanhr2l(α¯li2αli2)].The time-dependent reduced density matrix is obtained by integrating the propagating function over the initial state of Eq. (A1). It reads ρ(α¯f,αf;t)=xexp[l(mlα¯lf2+m¯lαlf2+plα¯lfαlf)],where x=(A1A2cosh2r)1, ml=iul(t)2tanhr2Al, and pl=|ul(t)|2(1Al1), with Al=1(|ul(t)|21)2tanh2r. Remembering ρ(t)=dμ(αf)dμ(αf)ρ(α¯f,αf;t)|α1f,α2f·α¯1f,α¯2f| and ρout=V^ρ(t)V^, we obtain ρout=dμ(αf)dμ(αf)ρ(α¯f,αf;t)×|α1f+iα2f2,α2f+iα1f2α¯1fiα¯2f2,α¯2fiα¯2f2|.Then the expectation value Π¯=Tr[Π^ρout] of the parity operator Π^=exp(iπa^1a^1) can be calculated as Π¯=x[4m1(m2*m1*p22)+4m2(m1*m2*p12)+(1p1p2)2+16|m1m2|2]1/2,where Π^|α,β=|eiπα,β has been used. The sensing sensitivity of Ω is calculated by δΩ=1Π¯2|ΩΠ¯|.

    In the ideal limit, the solution of Eq. (A3) reads ul(t)=exp(iωlt); thus, Al=1, pl=0, and ml=iei2ωlttanhr2. Then Eq. (A7) reduces to Π¯=[1+N(2+N)·cos2(2Ωt)]1/2 with N=2sinh2r.

    APPENDIX B: SENSITIVITY UNDER THE BORN–MARKOVIAN APPROXIMATION

    Defining ul(t)=eiωltul(t), we can rewrite Eq. (A3) as u˙l(t)+0tdτ0dωJ(ω)ei(ωωl)(tτ)ul(τ)=0.When the probe–environment coupling is weak and the time scale of the environmental correlation function is much smaller than the one of the probe, we can apply the Born–Markovian approximation to Eq. (B1) by neglecting the memory effect, i.e., u(τ)u(t), and extending the upper limit of the integral to infinity, i.e., 0tdτ0dτ. The utilization of the identity limt0tdτei(ωω0)(tτ)=πδ(ωω0)+iP1ω0ω, with P being the Cauchy principal value, results in ul,MA(t)=e[κl+iΔ(ωl)]t, where κl=πJ(ωl) and Δ(ωl)=P0J(ω)ωlωdω. We, thus, have the Born–Markovian approximate solution of ul(t) as ul,MA(t)=e{κl+i[ωl+Δ(ωl)]}t.

    Substituting ul,MA(t) into Eq. (A7) and using the error propagation formula, we analytically obtain the sensitivity under the Born–Markovian approximation as δΩBA(t)=(2e2tκ+NCe2tκ)C8N(N+2)t|sin(4Ωt)|,where C=4e2tκ+N2+(N+2)cos(4Ωt). Here we have chosen κ1=κ2κ and neglected the constant Δ(ωl), which is generally renormalized into ω0. We readily see from Eq. (B2) that the sensitivity under the Born–Markovian approximation tends to be divergent in the long-time limit.

    APPENDIX C: SENSITIVITY IN THE NON-MARKOVIAN DYNAMICS

    In the non-Markovian case, Eq. (A3) can be analytically solvable by the method of Laplace transform, which converts Eq. (A3) into u˜l(zl)=[zl+iωl+0J(ω)dωzl+iωdω]1. Then ul(t) is obtained by applying the inverse Laplace transform on u˜l(zl), we obtain ul(t)=12πiiσ+iσeiEltElωl+0J(ω)ωEldωdE,where El=izl and σ is chosen to be larger than all the poles of the integrand. We find the pole of Eq. (C1) from Yl(El)ωl0J(ω)ωEldω=El.It is noted that El also represents the eigenenergy in the single-excitation subspace of the total systems formed by each optical field and its environment. To see this, we expand the eigenstate as |Φl=(xla^l+kyk,lb^k,l)|0,{0k,l}. From the stationary Schrödinger equation, we have [El(ω0±Ω)]xl=kgk,lyk,l and yk,l=gk,lxl/(Elωk,l), with El being the eigenenergy. These two equations readily result in Eq. (C2) in the continuous limit of the environmental frequencies. According to the residue theorem, we have ul(t)=ZleiEb,lt+0Θ(E)eiEtdE,where Zl=[1+0J(ω)(Eb,lω)2dω]1and Θ(E)=J(E)[EωlΔ(E)]2+[πJ(E)]2. The first and the second terms of Eq. (C3) are the residues contributed from the poles of Eq. (C2) in the regime El<0 and El>0, respectively. It can be found that Yl(El) are decreasing functions in the regime El<0, and each of Eq. (C2) has one isolated root Eb,l in this regime provided Yl(0)<0. While Yl(El) are not well analytic in the regime El>0, they have infinite roots in this regime, which form a continuous energy band. We call the eigenstates of the isolated eigenenergies Eb,l bound states. We plot in Fig. 5 the solution of Eq. (C2) obtained by the graphical method. It verifies the three typical features on the solutions, i.e., no bound state when Y±(0)>0 in Fig. 5(a), one bound state when Y(0)<0 but Y+(0)>0 in Fig. 5(b), and two bound states when Y±(0)<0 in Fig. 5(c).

    (a)–(c) Solution of Eq. (C2) determined by the intersectors of two curves of y(E)=E (red dashed lines) and y(E)=Y+(E) (blue solid lines) or y(E)=Y−(E) (magenta dotted lines). In the regime E>0, both Y±(E) have infinite intersections with E, which form a continuous energy band. As long as either Y−(0)<0 or Y+(0)<0, an isolated eigenenergy corresponding to a bound state is formed in the regime E<0. (d)–(f) Corresponding behaviors of |u+(t)| (blue dotted lines) and |u−(t)| (red dashed lines) determined by numerically solving Eq. (A3). The light blue dotted and light red dashed lines in (d) and (e) show Zl determined by Eq. (C4). Accompanying the formation of a bound state, the corresponding |ul(t)| approaches a finite value, which exactly matches with Zl. The parameters are s=1, η=0.05, Ω=10−2ω0, ωc=2ω0 in (a) and (d), 20ω0 in (b) and (e), and 25ω0 in (c) and (f).

    Figure 5.(a)–(c) Solution of Eq. (C2) determined by the intersectors of two curves of y(E)=E (red dashed lines) and y(E)=Y+(E) (blue solid lines) or y(E)=Y(E) (magenta dotted lines). In the regime E>0, both Y±(E) have infinite intersections with E, which form a continuous energy band. As long as either Y(0)<0 or Y+(0)<0, an isolated eigenenergy corresponding to a bound state is formed in the regime E<0. (d)–(f) Corresponding behaviors of |u+(t)| (blue dotted lines) and |u(t)| (red dashed lines) determined by numerically solving Eq. (A3). The light blue dotted and light red dashed lines in (d) and (e) show Zl determined by Eq. (C4). Accompanying the formation of a bound state, the corresponding |ul(t)| approaches a finite value, which exactly matches with Zl. The parameters are s=1, η=0.05, Ω=102ω0, ωc=2ω0 in (a) and (d), 20ω0 in (b) and (e), and 25ω0 in (c) and (f).

    (a) Global behavior of the local minima of the steady-state δΩ(t) as a function of N. (b) Threshold Nc in (a) as a function of Z1Z2. We use the same parameter values as the ones of the blue solid line in Fig. 4(b) of the main text.

    Figure 6.(a) Global behavior of the local minima of the steady-state δΩ(t) as a function of N. (b) Threshold Nc in (a) as a function of Z1Z2. We use the same parameter values as the ones of the blue solid line in Fig. 4(b) of the main text.

    Experimentally, a squeezing parameter r2.5, which corresponds to N73, has been realized [73]. We see from Figs. 3(b) and 4(b) that the threshold is still absent even when N is as large as 100. Therefore, although we have to face the balance between the bound-state favored restoring superiority and the photon-dissipation caused destruction to the sensitivity, our mechanism still supplies us with a sufficient space to rescue the ideal sensitivity from the noise using the experimentally accessible numbers of quantum resources.

    References

    [1] P. P. Khial, A. D. White, A. Hajimiri. Nanophotonic optical gyroscope with reciprocal sensitivity enhancement. Nat. Photonics, 12, 671-675(2018).

    [2] Y.-H. Lai, M.-G. Suh, Y.-K. Lu, B. Shen, Q.-F. Yang, H. Wang, J. Li, S. H. Lee, K. Y. Yang, K. Vahala. Earth rotation measured by a chip-scale ring laser gyroscope. Nat. Photonics, 14, 345-349(2020).

    [3] S. Srivastava, D. S. Shreesha Rao, H. Nandakumar. Novel optical gyroscope: proof of principle demonstration and future scope. Sci. Rep., 6, 34634(2016).

    [4] 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, 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).

    [5] 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, G. Terreni. Underground Sagnac gyroscope with sub-prad/s rotation rate sensitivity: toward general relativity tests on Earth. Phys. Rev. Res., 2, 032069(2020).

    [6] G. A. Sanders, A. A. Taranta, C. Narayanan, E. N. Fokoua, S. A. Mousavi, L. K. Strandjord, M. Smiciklas, T. D. Bradley, J. Hayes, G. T. Jasion, T. Qiu, W. Williams, F. Poletti, D. N. Payne. Hollow-core resonator fiber optic gyroscope using nodeless anti-resonant fiber. Opt. Lett., 46, 46-49(2021).

    [7] A. Lenef, T. D. Hammond, E. T. Smith, M. S. Chapman, R. A. Rubenstein, D. E. Pritchard. Rotation sensing with an atom interferometer. Phys. Rev. Lett., 78, 760-763(1997).

    [8] T. L. Gustavson, P. Bouyer, M. A. Kasevich. Precision rotation measurements with an atom interferometer gyroscope. Phys. Rev. Lett., 78, 2046-2049(1997).

    [9] D. S. Durfee, Y. K. Shaham, M. A. Kasevich. Long-term stability of an area-reversible atom-interferometer Sagnac gyroscope. Phys. Rev. Lett., 97, 240801(2006).

    [10] P. Berg, S. Abend, G. Tackmann, C. Schubert, E. Giese, W. P. Schleich, F. A. Narducci, W. Ertmer, E. M. Rasel. Composite-light-pulse technique for high-precision atom interferometry. Phys. Rev. Lett., 114, 063002(2015).

    [11] R. Trubko, J. Greenberg, M. T. S. Germaine, M. D. Gregoire, W. F. Holmgren, I. Hromada, A. D. Cronin. Atom interferometer gyroscope with spin-dependent phase shifts induced by light near a tune-out wavelength. Phys. Rev. Lett., 114, 140404(2015).

    [12] I. Dutta, D. Savoie, B. Fang, B. Venon, C. L. Garrido Alzar, R. Geiger, A. Landragin. Continuous cold-atom inertial sensor with 1 nrad/sec rotation stability. Phys. Rev. Lett., 116, 183003(2016).

    [13] Y. Zhao, X. Yue, F. Chen, C. Huang. Extension of the rotation-rate measurement range with no sensitivity loss in a cold-atom gyroscope. Phys. Rev. A, 104, 013312(2021).

    [14] E. R. Moan, R. A. Horne, T. Arpornthip, Z. Luo, A. J. Fallon, S. J. Berl, C. A. Sackett. Quantum rotation sensing with dual Sagnac interferometers in an atom-optical waveguide. Phys. Rev. Lett., 124, 120403(2020).

    [15] B. Culshaw. The optical fibre Sagnac interferometer: an overview of its principles and applications. Meas. Sci. Technol., 17, R1-R16(2005).

    [16] H. C. Lefèvre. The fiber-optic gyroscope, a century after Sagnac’s experiment: the ultimate rotation-sensing technology?. C. R. Phys., 15, 851-858(2014).

    [17] H. Zhang, X. Chen, X. Shu, C. Liu. Fiber optic gyroscope noise reduction with fiber ring resonator. Appl. Opt., 57, 7391-7397(2018).

    [18] V. Giovannetti, S. Lloyd, L. Maccone. Quantum-enhanced measurements: beating the standard quantum limit. Science, 306, 1330-1336(2004).

    [19] V. Giovannetti, S. Lloyd, L. Maccone. Quantum metrology. Phys. Rev. Lett., 96, 010401(2006).

    [20] V. Giovannetti, S. Lloyd, L. Maccone. Advances in quantum metrology. Nat. Photonics, 5, 222-229(2011).

    [21] C. L. Degen, F. Reinhard, P. Cappellaro. Quantum sensing. Rev. Mod. Phys., 89, 035002(2017).

    [22] L. Pezzè, A. Smerzi, M. K. Oberthaler, R. Schmied, P. Treutlein. Quantum metrology with nonclassical states of atomic ensembles. Rev. Mod. Phys., 90, 035005(2018).

    [23] C. M. Caves. Quantum-mechanical noise in an interferometer. Phys. Rev. D, 23, 1693-1708(1981).

    [24] N. J. Engelsen, R. Krishnakumar, O. Hosten, M. A. Kasevich. Bell correlations in spin-squeezed states of 500 000 atoms. Phys. Rev. Lett., 118, 140401(2017).

    [25] D. Gatto, P. Facchi, F. A. Narducci, V. Tamma. Distributed quantum metrology with a single squeezed-vacuum source. Phys. Rev. Res., 1, 032024(2019).

    [26] Y. Israel, S. Rosen, Y. Silberberg. Supersensitive polarization microscopy using NOON states of light. Phys. Rev. Lett., 112, 103604(2014).

    [27] X.-Y. Luo, Y.-Q. Zou, L.-N. Wu, Q. Liu, M.-F. Han, M. K. Tey, L. You. Deterministic entanglement generation from driving through quantum phase transitions. Science, 355, 620-623(2017).

    [28] M. Fink, F. Steinlechner, J. Handsteiner, J. P. Dowling, T. Scheidl, R. Ursin. Entanglement-enhanced optical gyroscope. New J. Phys., 21, 053010(2019).

    [29] F. De Leonardis, R. Soref, M. De Carlo, V. M. N. Passaro. On-chip group-IV Heisenberg-limited Sagnac interferometric gyroscope at room temperature. Sensors, 20, 3476(2020).

    [30] M. Mehmet, T. Eberle, S. Steinlechner, H. Vahlbruch, R. Schnabel. Demonstration of a quantum-enhanced fiber Sagnac interferometer. Opt. Lett., 35, 1665-1667(2010).

    [31] K. Liu, C. Cai, J. Li, L. Ma, H. Sun, J. Gao. Squeezing-enhanced rotating-angle measurement beyond the quantum limit. Appl. Phys. Lett., 113, 261103(2018).

    [32] M. R. Grace, C. N. Gagatsos, Q. Zhuang, S. Guha. Quantum-enhanced fiber-optic gyroscopes using quadrature squeezing and continuous-variable entanglement. Phys. Rev. Appl., 14, 034065(2020).

    [33] A. Luis, I. Morales, A. Rivas. Nonlinear fiber gyroscope for quantum metrology. Phys. Rev. A, 94, 013830(2016).

    [34] C. L. Garrido Alzar. Compact chip-scale guided cold atom gyrometers for inertial navigation: enabling technologies and design study. AVS Quantum Sci., 1, 014702(2019).

    [35] S. S. Szigeti, S. P. Nolan, J. D. Close, S. A. Haine. High-precision quantum-enhanced gravimetry with a Bose-Einstein condensate. Phys. Rev. Lett., 125, 100402(2020).

    [36] S. S. Szigeti, O. Hosten, S. A. Haine. Improving cold-atom sensors with quantum entanglement: prospects and challenges. Appl. Phys. Lett., 118, 140501(2021).

    [37] C. Luo, J. Huang, X. Zhang, C. Lee. Heisenberg-limited Sagnac interferometer with multiparticle states. Phys. Rev. A, 95, 023608(2017).

    [38] D. J. Wineland, J. J. Bollinger, W. M. Itano, F. L. Moore, D. J. Heinzen. Spin squeezing and reduced quantum noise in spectroscopy. Phys. Rev. A, 46, R6797-R6800(1992).

    [39] M. Kitagawa, M. Ueda. Squeezed spin states. Phys. Rev. A, 47, 5138-5143(1993).

    [40] S.-Y. Bai, J.-H. An. Generating stable spin squeezing by squeezed-reservoir engineering. Phys. Rev. Lett., 127, 083602(2021).

    [41] U. Dorner, R. Demkowicz-Dobrzanski, B. J. Smith, J. S. Lundeen, W. Wasilewski, K. Banaszek, I. A. Walmsley. Optimal quantum phase estimation. Phys. Rev. Lett., 102, 040403(2009).

    [42] R. Demkowicz-Dobrzanski, U. Dorner, B. J. Smith, J. S. Lundeen, W. Wasilewski, K. Banaszek, I. A. Walmsley. Quantum phase estimation with lossy interferometers. Phys. Rev. A, 80, 013825(2009).

    [43] J. Joo, W. J. Munro, T. P. Spiller. Quantum metrology with entangled coherent states. Phys. Rev. Lett., 107, 083601(2011).

    [44] T. Ono, H. F. Hofmann. Effects of photon losses on phase estimation near the Heisenberg limit using coherent light and squeezed vacuum. Phys. Rev. A, 81, 033819(2010).

    [45] Z. Huang, K. R. Motes, P. M. Anisimov, J. P. Dowling, D. W. Berry. Adaptive phase estimation with two-mode squeezed vacuum and parity measurement. Phys. Rev. A, 95, 053837(2017).

    [46] A. Smirne, J. Kołodyński, S. F. Huelga, R. Demkowicz-Dobrzański. Ultimate precision limits for noisy frequency estimation. Phys. Rev. Lett., 116, 120801(2016).

    [47] F. Albarelli, M. A. C. Rossi, D. Tamascelli, M. G. Genoni. Restoring Heisenberg scaling in noisy quantum metrology by monitoring the environment. Quantum, 2, 110(2018).

    [48] M. Scully, M. Zubairy. Quantum Optics(1997).

    [49] P. Kok, J. Dunningham, J. F. Ralph. Role of entanglement in calibrating optical quantum gyroscopes. Phys. Rev. A, 95, 012326(2017).

    [50] P. M. Anisimov, G. M. Raterman, A. Chiruvelli, W. N. Plick, S. D. Huver, H. Lee, J. P. Dowling. Quantum metrology with two-mode squeezed vacuum: parity detection beats the Heisenberg limit. Phys. Rev. Lett., 104, 103602(2010).

    [51] M. M. Rams, P. Sierant, O. Dutta, P. Horodecki, J. Zakrzewski. At the limits of criticality-based quantum metrology: apparent super-Heisenberg scaling revisited. Phys. Rev. X, 8, 021022(2018).

    [52] Z. Hou, Y. Jin, H. Chen, J.-F. Tang, C.-J. Huang, H. Yuan, G.-Y. Xiang, C.-F. Li, G.-C. Guo. ‘Super-Heisenberg’ and Heisenberg scalings achieved simultaneously in the estimation of a rotating field. Phys. Rev. Lett., 126, 070503(2021).

    [53] V. Giovannetti, L. Maccone. Sub-Heisenberg estimation strategies are ineffective. Phys. Rev. Lett., 108, 210404(2012).

    [54] L. Pezzé. Sub-Heisenberg phase uncertainties. Phys. Rev. A, 88, 060101(2013).

    [55] Y. Zhao, N. Aritomi, E. Capocasa, M. Leonardi, M. Eisenmann, Y. Guo, E. Polini, A. Tomura, K. Arai, Y. Aso, Y.-C. Huang, R.-K. Lee, H. Lück, O. Miyakawa, P. Prat, A. Shoda, M. Tacca, R. Takahashi, H. Vahlbruch, M. Vardaro, C.-M. Wu, M. Barsuglia, R. Flaminio. Frequency-dependent squeezed vacuum source for broadband quantum noise reduction in advanced gravitational-wave detectors. Phys. Rev. Lett., 124, 171101(2020).

    [56] P. A. Knott, T. J. Proctor, K. Nemoto, J. A. Dunningham, W. J. Munro. Effect of multimode entanglement on lossy optical quantum metrology. Phys. Rev. A, 90, 033846(2014).

    [57] J. J. Cooper, D. W. Hallwood, J. A. Dunningham, J. Brand. Robust quantum enhanced phase estimation in a multimode interferometer. Phys. Rev. Lett., 108, 130402(2012).

    [58] Q.-J. Tong, J.-H. An, H.-G. Luo, C. H. Oh. Mechanism of entanglement preservation. Phys. Rev. A, 81, 052330(2010).

    [59] W.-M. Zhang, P.-Y. Lo, H.-N. Xiong, M. W.-Y. Tu, F. Nori. General non-Markovian dynamics of open quantum systems. Phys. Rev. Lett., 109, 170402(2012).

    [60] H.-J. Zhu, G.-F. Zhang, L. Zhuang, W.-M. Liu. Universal dissipationless dynamics in Gaussian continuous-variable open systems. Phys. Rev. Lett., 121, 220403(2018).

    [61] H.-P. Breuer, E.-M. Laine, J. Piilo, B. Vacchini. Colloquium: non-Markovian dynamics in open quantum systems. Rev. Mod. Phys., 88, 021002(2016).

    [62] L. Li, M. J. Hall, H. M. Wiseman. Concepts of quantum non-Markovianity: a hierarchy. Phys. Rep., 759, 1-51(2018).

    [63] J.-H. An, W.-M. Zhang. Non-Markovian entanglement dynamics of noisy continuous-variable quantum channels. Phys. Rev. A, 76, 042127(2007).

    [64] C. J. Myatt, B. E. King, Q. A. Turchette, C. A. Sackett, D. Kielpinski, W. M. Itano, C. Monroe, D. J. Wineland. Decoherence of quantum superpositions through coupling to engineered reservoirs. Nature, 403, 269-273(2000).

    [65] D. Kienzler, H.-Y. Lo, B. Keitch, L. de Clercq, F. Leupold, F. Lindenfelser, M. Marinelli, V. Negnevitsky, J. P. Home. Quantum harmonic oscillator state synthesis by reservoir engineering. Science, 347, 53-56(2015).

    [66] N.-H. Tong, M. Vojta. Signatures of a noise-induced quantum phase transition in a mesoscopic metal ring. Phys. Rev. Lett., 97, 016802(2006).

    [67] P. Forn-Díaz, J. J. García-Ripoll, B. Peropadre, J.-L. Orgiazzi, M. A. Yurtalan, R. Belyansky, C. M. Wilson, A. Lupascu. Ultrastrong coupling of a single artificial atom to an electromagnetic continuum in the nonperturbative regime. Nat. Phys., 13, 39-43(2017).

    [68] E. Paladino, Y. M. Galperin, G. Falci, B. L. Altshuler. 1/f noise: implications for solid-state quantum information. Rev. Mod. Phys., 86, 361-418(2014).

    [69] B.-H. Liu, L. Li, Y.-F. Huang, C.-F. Li, G.-C. Guo, E.-M. Laine, H.-P. Breuer, J. Piilo. Experimental control of the transition from Markovian to non-Markovian dynamics of open quantum systems. Nat. Phys., 7, 931-934(2011).

    [70] N. K. Bernardes, A. Cuevas, A. Orieux, C. H. Monken, P. Mataloni, F. Sciarrino, M. F. Santos. Experimental observation of weak non-Markovianity. Sci. Rep., 5, 17520(2015).

    [71] Y. Liu, A. A. Houck. Quantum electrodynamics near a photonic bandgap. Nat. Phys., 13, 48-52(2012).

    [72] L. Krinner, M. Stewart, A. Pazmiño, J. Kwon, D. Schneble. Spontaneous emission of matter waves from a tunable open quantum system. Nature, 559, 589-592(2018).

    [73] C. Macklin, K. O’Brien, D. Hover, M. E. Schwartz, V. Bolkhovsky, X. Zhang, W. D. Oliver, I. Siddiqi. A near–quantum-limited Josephson traveling-wave parametric amplifier. Science, 350, 307-310(2015).