This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quantum Magnetometer with Dual-Coupling Optomechanics

Gui-Lei Zhu School of Physics, Huazhong University of Science and Technology, Wuhan 430074, China    Jing Liu School of Physics, Huazhong University of Science and Technology, Wuhan 430074, China MOE Key Laboratory of Fundamental Physical Quantities Measurement and PGMF, Huazhong University of Science and Technology, Wuhan 430074, China    Ying Wu School of Physics, Huazhong University of Science and Technology, Wuhan 430074, China    Xin-You Lü [email protected] School of Physics, Huazhong University of Science and Technology, Wuhan 430074, China
Abstract

An experimentally feasible magnetometer based on a dual-coupling optomechanical system is proposed, where the radiation-pressure coupling transduces the magnetic signal to the optical phase, and the quadratic optomechanical interaction induces a periodic squeezing effect. The latter not only amplifies the signal to be measured, but also accelerates the signal transducing rate characterized by an experimentally observable phase accumulation efficiency. In the vicinity of opto-mechanical decoupled time, the ultimate bound to the estimability of magnetic signal is proportional to exp(6r)\exp(-6r), and then the optimized accuracy of estimation can be enhanced nearly 3 orders with a controllable squeezing parameter r<1r<1. Moreover, our proposal is robust against the mechanical thermal noise, and the sensitivity of a specific measurement can reach to the order of 1017T/Hz10^{-17}{\rm T/\sqrt{Hz}} in the presence of dissipations and without ground state cooling of mechanical oscillator. Our proposal fundamentally broadens the fields of quantum metrology and cavity optomechanics, with potential application for on-chip magnetic detection with high precision.

Ultrasensitive magnetic detection has contributed immensely to a wide range of scientific areas from fundamental physics to advanced technologies, such as geological exploration, aerospace Bennett et al. (2021), biomedical imaging and diagnostics Hämäläinen et al. (1993); Lee et al. (2015); Murzin et al. (2020). Over the past few decades, various magnetometers have been developed, including the superconducting quantum interference devices (SQUID) based on superconducting effects Jaklevic et al. (1964); Erné et al. (1976); Kleiner et al. (2004), spin-exchange relaxation-free atomic magnetometers Allred et al. (2002); Kominis et al. (2003); Savukov et al. (2005); Xia et al. (2006), NV center magnetometers Taylor et al. (2008); Maze et al. (2008) and Hall-effect sensors Bending (1999). Normally, they require the elaborated operating conditions, such as the associated denoising technology and/or the complex signal read-out schemes Robbes (2006), which reduces their capability of on-chip integration.

Cavity optomechanical system (OMS) Aspelmeyer et al. (2014); Kippenberg and Vahala (2008); Aspelmeyer et al. (2012); Meystre (2013) offers an alternative platform for the precision measurements of mass Li and Zhu (2012); Lin et al. (2017); Bin et al. (2019), weak forces Clerk et al. (2010); Tsang and Caves (2010); Pontin et al. (2014); Armata et al. (2017); Qvarfort et al. (2018), and magnetic fields Forstner et al. (2012, 2014); Yu et al. (2016); Wu et al. (2017); Zhu et al. (2017); Li et al. (2018). Particularly, optomechanical magnetometer with high sensitivity has excellent quality of on-chip integration Li et al. (2021). Recently, the YIG sphere-based optomechanical magnetometer has been experimentally demonstrated in Ref. Colombano et al. (2020), which have attained extremely low sensitivity values. In such magnetometers, the existence of Joule and Villari effects of magnetostrictive transducer Olabi and Grunwald (2008), allows one to directly extract magnetic information by reading out the optical frequency shift (or transmission spectrum). Moreover, quantum metrology Giovannetti et al. (2004, 2006, 2011); Dowling and Seshadreesan (2015) points out that the quantum squeezing or entanglement Caves (1981); Ma et al. (2011); Baumgratz and Datta (2016); Degen et al. (2017); Engelsen et al. (2017); Nagata et al. (2007); Israel et al. (2014); Luo et al. (2017) could improve the accuracy of parameter estimation in physical systems from the shot-noise limit to the Heisenberg limit, i.e., the optimal precision scales from 1/N1/\sqrt{N} to 1/N1/N with NN being the number of resources employed in the measurements. This has stimulated enormous interests in exploiting quantum resources in the atoms (or spins) Jones et al. (2009); Tanaka et al. (2015); Kómár et al. (2014); Hou et al. (2020) and optical systems Holland and Burnett (1993); Boto et al. (2000); Anisimov et al. (2010); Joo et al. (2011) for high-precision physical quantity measurements.

By applying quantum metrology to the detection of a static magnetic field, here we propose a quantum magnetometer based on a dual-coupling OMS, which has a periodic decoupling behavior between the optical and mechanical modes. The OMS supports two optical modes coupled simultaneously to the same mechanical mode, with radiation-pressure and quadratic optomechanical interactions, respectively Thompson et al. (2008); Sankey et al. (2010); Bhattacharya et al. (2008); Zhu et al. (2018). The radiation-pressure coupling acts as a signal transducer, encoding the magnetic signal received by the mechanical oscillator into the optical phase. The quadratic optomechanical interaction amplifies both the signal to be measured and the signal transducing rate via inducing a periodic squeezing effect on the mechanical oscillator, whose maximum squeezing strength is determined by a controllable squeezing parameter rr. By performing a homodyne detection on the optical phase within a wide time window around the first decoupled time τ1\tau_{1}, the magnetic signal could be estimated with high precision.

To qualitatively characterize the precision of magnetometer, we define a displaced phase accumulation efficiency (PAE) that is experimentally observable via state tomography technique. By presenting the exponentially increased quantum Fisher information (QFI), i.e., q(τ1)e12r\mathcal{F}_{q}(\tau_{1})\propto e^{12r}, we quantitatively demonstrate that the fundamental bound of measurement precision can be dramatically reduced even with a small squeezing parameter rr. The periodic opto-mechanical decoupling makes the classical Fisher information (CFI) robust against the mechanical environment at the decoupled time, which in turn allows the sensitivity of a specific measurement to saturate the fundamental bound and reach to the order of 1017T/Hz10^{-17}{\rm T/\sqrt{Hz}} in the presence of system dissipations. Moreover, the sensitivity to the order of 1015T/Hz\sim 10^{-15}{\rm T/\sqrt{Hz}} is predicted even in the case of the thermal phonon number n¯th103\bar{n}_{\rm th}\sim 10^{3}. Our work establishes a connection between quantum metrology and dual-coupling optomechanics, which is suitable for detecting various fields that linearly interact with the mechanical oscillator.

Refer to caption
Figure 1: Schematic illustration of quantum magnetometer based on a dual-coupling OMS with “membrane-in-the-middle” configuration. The middle SiN membrane (green), acting as the mechanical mode bb, is welded to the ends of magnetostrictive Terfenol-D rods (gray) Hong (2013). On application of a static magnetic field, the Terfenol-D rods expands, which displaces the equilibrium position of the mechanical mode, and then the magnetic signal is encoded into the phase Φ\Phi of optical mode a1a_{1} via radiation-pressure coupling. Besides, cavity mode a2a_{2} offers the periodic squeezing for mode bb via the quadratic optomechanical interaction [see the shaded area]. The optical phase is detected by the balanced homodyne detection scheme with a local oscillator (LO) pulse.

System and periodic mechanical squeezing.— We consider a dual-coupling optomechanical system depicted in Fig. 1 with Hamiltonian

H\displaystyle H =ω1a1a1+ω2a2a2+ωmbbλ1a1a1(b+b)\displaystyle=\hbar\omega_{1}a^{\dagger}_{1}a_{1}+\hbar\omega_{2}a^{\dagger}_{2}a_{2}+\hbar\omega_{m}b^{\dagger}b-\hbar\lambda_{1}a^{\dagger}_{1}a_{1}(b^{\dagger}+b)
λ2a2a2(b+b)2+Bzcactz,\displaystyle\,\,\,\,\,\,\,-\hbar\lambda_{2}a^{\dagger}_{2}a_{2}(b^{\dagger}+b)^{2}+B_{z}c_{\rm act}z, (1)

where aja_{j} (j=1,2)(j=1,2) and bb are the annihilation operators of the cavity mode with frequency ωj\omega_{j} and the mechanical mode with frequency ωm\omega_{m}, respectively. The mechanical membrane is placed at a node (antinode) of cavity mode a1a_{1} (a2a_{2}), and then the fourth (fifth) term in Eq. (1) describes the radiation-pressure (quadratic optomechanical) interaction between modes a1a_{1} (a2a_{2}) and bb with strength λ1\lambda_{1} (λ2\lambda_{2}). In the presence of a static magnetic field Bz{B_{z}} (along zz direction), the field-sensitive Terfenol-D expands, which leads to the change of the equilibrium position of the mechanical oscillator, thus generating an effective magnetic potential on the Hamiltonian, i.e., the last term of Equation (1) SM . The mechanical motion modulates the optical cavity field via the nonlinear radiation-pressure coupling. Meanwhile, the phase shift of the mechanical motions encoded with magnetic signal is transferred to the optical field. By reading out the phase shift of optical field via homodyne detection, we can extract the original magnetic information. Here z=/(2mωm)(b+b)z=\sqrt{\hbar/(2m\omega_{m})}(b^{{\dagger}}+b) is the position operator of the mechanical oscillator with mass mm, and cact=mωm2Lαmag/Ec_{\rm act}=m\omega_{m}^{2}L{\alpha_{\rm mag}}/{E} is the magnetic actuation constant charactering how well the magnetic field is converted into a force applied on the oscillator. The symbol LL denotes the length of Terfenol-D rods, αmag\alpha_{\rm mag} is magnetostrictive coefficient and EE is the Young’s modulus SM .

Refer to caption
Figure 2: (a) Wigner functions of the mechanical mode within 2TT given by a1a_{1} in vaccum state |0|0\rangle, ωm=1\omega_{m}=1 and 𝒩2=1414\sqrt{\mathcal{N}_{2}}=1414. The instantaneous squeezing degrees are marked. (b) Tomography of the state |Ψ(τ1)|\Psi(\tau_{1})\rangle projected on the subsapce {|,|}\{|\!\!\downarrow\rangle,|\!\!\uparrow\rangle\} with l=3l=3 and 𝒩2=1500\sqrt{\mathcal{N}_{2}}\!=\!1500. The dependence of σz(τ1)\sigma_{z}(\tau_{1}) on the maximum squeezing degree SmaxS_{\rm max} for l=1l=1 and f=ωmf=\omega_{m} is shown in the insert of (c). The main panel of (c) shows the absolute value of displaced PAE |𝒫~n||\widetilde{\mathcal{P}}_{n}| as a function of 𝒩2\sqrt{\mathcal{N}_{2}} for several values of nn. The parameters are ωm=2π×134kHz,f=0.01ωm\omega_{m}=2\pi\times 134\,{\rm kHz},f=0.01\omega_{m}, α=1\alpha=1, λ1=0.01ωm\lambda_{1}=0.01\omega_{m}, and λ2=107ωm\lambda_{2}=10^{-7}\omega_{m}.

By considering the ancillary mode a2a_{2} in the coherent state |ξ|\xi\rangle, the number operator a2a2a_{2}^{\dagger}a_{2} can be approximately replaced by an algebraic number 𝒩2=|ξ|2\mathcal{N}_{2}=|\xi|^{2} in the case of 𝒩21\mathcal{N}_{2}\gg 1SM . Assuming the modes a1a_{1} and bb are initially in the coherent state |Ψ(0)=|α|β|\Psi(0)\rangle=|\alpha\rangle\otimes|\beta\rangle with β=βRe+iβIm\beta=\beta_{\rm Re}+i\beta_{\rm Im}, the instantaneous state of system is given by SM

|Ψ(t)=e|α|2/2n=0αnn!exp[iη2(ωstsinωst)]\displaystyle\!\!\!|\Psi(t)\rangle\!\!=\!e^{-{|\alpha|^{2}}/{2}}\sum_{n=0}^{\infty}\frac{\alpha^{n}}{\sqrt{n!}}{\rm exp}\left[i\eta^{2}(\omega_{s}t-\sin\omega_{s}t)\right]
×exp{iη[βResinωsterβIm(cosωst1)er]}|n|φs(t),\displaystyle\!\!\!\times\!{\rm exp}\{{i\eta[\beta_{\rm Re}\sin\omega_{s}t\,e^{-r}\!\!-\!\!\beta_{\rm Im}(\cos\omega_{s}t\!-\!\!1)e^{r}]}\}|n\rangle|\varphi_{s}(t)\rangle,\! (2)

where η=λsn/ωsfs/ωs\eta=\lambda_{s}n/\omega_{s}-f_{s}/\omega_{s} with λs=λ1er\lambda_{s}=\lambda_{1}e^{r}, ωs=ωme2r\omega_{s}=\omega_{m}e^{-2r}, fs=fer=Bzcact1/(2mωm)erf_{s}=fe^{r}=B_{z}c_{\rm act}\sqrt{1/(2m\hbar\omega_{m})}e^{r}, r=(1/4)ln(14λ2𝒩2/ωm)r=-(1/4)\ln(1-4\lambda_{2}\mathcal{N}_{2}/\omega_{m}), and a rotating frame with exp(iω1a1a1/ωs){\rm exp}(-i\omega_{1}a_{1}^{{\dagger}}a_{1}/\omega_{s}) was adopted. The mechanical state reads |φs(t)=S(r)S(r)|φn(t)|\varphi_{s}(t)\rangle=S^{\dagger}(r)S(r^{\prime})|\varphi_{n}(t)\rangle with the defined squeezing operator S(r)=exp[r(b2b2)/2]S(r)=\exp[r(b^{2}-b^{\dagger 2})/2] and the squeezing parameter r=re2iωstr^{\prime}=re^{-2i\omega_{s}t}. Here |φn(t)=|eiωstβ+ημ¯|\varphi_{n}(t)\rangle=|e^{-i\omega_{s}t}\beta+\eta\bar{\mu}\rangle is a displaced coherent state with μ¯=(1eiωst)(coshreiωstsinhr)\bar{\mu}=(1-e^{-i\omega_{s}t})(\cosh r-e^{-i\omega_{s}t}\sinh r). The expression of |φs(t)|\varphi_{s}(t)\rangle clearly shows that the mechanical mode is dynamically squeezed with the period T=π/ωsT=\pi/\omega_{s}, whose squeezing degree of quadrature X=1/2(b+b)X=1/\sqrt{2}(b+b^{\dagger}) is defined by S(t)=10log10(δX2(t)/δXmin2)dBS(t)=10\log_{10}(\delta X^{2}(t)/\delta X_{\rm min}^{2}){\rm dB}. The maximum squeezing degree Smax=10log10(e4r)dBS_{\rm max}=10\log_{10}(e^{4r}){\rm dB} occurs at T/2T/2 with the state S(2r)|βS^{\dagger}(2r)|\beta\rangleSM . This periodic squeezing effect on the mechanical oscillator is induced by the quadratic optomechanical coupling, and can be qualitatively presented in the case of α=η=0\alpha=\eta=0 [see Fig. 2(a)].

Periodic-squeezing-enhanced phase accumulation efficiency.— As shown in Eq. (Quantum Magnetometer with Dual-Coupling Optomechanics), the magnetic signal is transduced into the optical phase during the evolution of system via the radiation-pressure interaction. Interestingly, at time τm=2mπ/ωs\tau_{m}\!=\!2m\pi/\omega_{s} (m=1,2,.m=1,2,....), Eq. (Quantum Magnetometer with Dual-Coupling Optomechanics) can be reduced to SM

|Ψ(τm)=e|α|2/2n=0αnn!eiΦn(τm)|n|β,\displaystyle|\Psi(\tau_{m})\rangle=e^{-{|\alpha|^{2}}/{2}}\sum_{n=0}^{\infty}\frac{\alpha^{n}}{\sqrt{n!}}e^{i\Phi_{n}(\tau_{m})}|n\rangle|\beta\rangle, (3)

which demonstrates that the optical and mechanical modes are periodically decoupled, meanwhile the magnetic signal to be measured is periodically encoded into the accumulated optical phase Φn(τm)=2mπ(λsn/ωsfs/ωs)2\Phi_{n}(\tau_{m})=2m\pi(\lambda_{s}n/\omega_{s}-f_{s}/\omega_{s})^{2} for a Fock state |n|n\rangle. Here the opto-mechanical decoupled period is double of the period of dynamical squeezing, i.e., τ1=2π/ωs=2T\tau_{1}=2\pi/\omega_{s}=2T, and hence the mechanical squeezing also disappears at the decoupled time τm\tau_{m}.

The above unique property allows us to estimate the magnetic field BzB_{z} by performing a homodyne detection [see Fig. 1] on the optical mode a1a_{1} at the first decoupled time τ1\tau_{1}. To qualitatively describe the detection precision, here we define a displaced PAE

𝒫~n=Φn(τ1)Φ0(τ1)τ1=λ1nωm(λ1n2f)e4r.\displaystyle\widetilde{\mathcal{P}}_{n}=\frac{\Phi_{n}(\tau_{1})-\Phi_{0}(\tau_{1})}{\tau_{1}}=\frac{{\lambda_{1}n}}{\omega_{m}}{{(\lambda}_{1}n-2{f})}{e^{4r}}. (4)

Obviously, the larger 𝒫~n\widetilde{\mathcal{P}}_{n}, i.e., the faster phase accumulation rate on the optical Fock state |n|n\rangle, the higher measurement accuracy of magnetometer should be obtained. As shown in Fig. 2(c), the 𝒫~n\widetilde{\mathcal{P}}_{n} is exponentially enhanced by increasing the coherent amplitude 𝒩2\sqrt{\mathcal{N}_{2}} of the ancillary mode a2a_{2}. This enhancement originally comes from the periodic squeezing of mechanical mode during one opto-mechanical decoupled period [see Fig. 2(a)]. The system Hamiltonian in the squeezed frame shown in Eq. (S17) of supplementary material SM clearly demonstrates that the mechanical squeezing effect not only accelerates the signal tranducing rate from phonon to photon (i.e., the radiation-pressure interaction λsa1a1(b+b)\lambda_{s}a^{\dagger}_{1}a_{1}(b+b^{\dagger})Lü et al. (2015); Lemonde et al. (2016), but also amplifies the magnetic signal to be measured (i.e., the magnetic potential fs(b+b)f_{s}(b^{{\dagger}}+b)).

Refer to caption
Figure 3: (a) The QFI q(τ1)\mathcal{F}_{q}(\tau_{1}) versus 𝒩2\sqrt{\mathcal{N}_{2}} (main) and rr (inset). (b) Contourf plot of the optimal sensitivity Δ𝔹z(τ1)\Delta\mathbb{B}_{z}(\tau_{1}) in units of T/Hz{\rm T/\sqrt{Hz}} as functions of 𝒩1\mathcal{N}_{1} and 𝒩2\mathcal{N}_{2}. (c) Δ𝔹z(τ1)\Delta\mathbb{B}_{z}(\tau_{1}) (solid curve) and the specific sensitivity ΔBz(τ1)\Delta B_{z}(\tau_{1}) in the presence of dissipations (circles) versus rr for 𝒩1=106\mathcal{N}_{1}=10^{6}. Here we have chosen m=4×108g,m=4\times 10^{-8}{\rm g}, L=630μL=630\,\mum, αmag=5×108NT1m1\alpha_{\rm mag}=5\times 10^{8}{\rm NT^{-1}m^{-1}}, E=30GPaE=30{\rm\,GPa}, κ/ωm=0.01,γ/ωm=0.001\kappa/\omega_{m}=0.01,\gamma/\omega_{m}=0.001, n¯th=10\bar{n}_{\rm th}=10 and other parameters are same as Fig. 2.

More importantly, this well-defined PAE is experimentally observable via state tomography in the proper basis vectors. Specifically, expanding the system state |Ψ(τ1)|\Psi(\tau_{1})\rangle in the subspace {|,|}\{|\!\!\downarrow\rangle,|\!\!\uparrow\rangle\} with basis vectors |:=(1/2)(|0|l)|\!\downarrow\rangle:=(1/\sqrt{2})(|0\rangle-|l\rangle) and |:=(1/2)(|0+|l)|\!\uparrow\rangle:=(1/\sqrt{2})(|0\rangle+|l\rangle) (l=1,2,.l=1,2,....), we obtain four elements of density matrix SM , as shown in Fig. 2(b). The difference between two diagonal elements is denoted by σz(τ1)=ρ(τ1)ρ(τ1)=2e|α|2(αl/l!)cos(ΔΦl(τ1))\sigma_{z}(\tau_{1})=\rho_{\uparrow\uparrow}(\tau_{1})-\rho_{\downarrow\downarrow}(\tau_{1})=2e^{-|\alpha|^{2}}(\alpha^{l}/\sqrt{l!})\cos(\Delta\Phi_{l}(\tau_{1})), where ΔΦl(τ1)=Φl(τ1)Φ0(τ1)\Delta\Phi_{l}(\tau_{1})=\Phi_{l}(\tau_{1})-\Phi_{0}(\tau_{1}). Then one can easily obtain the values of phase difference ΔΦl(τ1)\Delta\Phi_{l}(\tau_{1}) by directly measuring ρ(τ1)\rho_{\downarrow\downarrow}(\tau_{1}) and ρ(τ1)\rho_{\uparrow\uparrow}(\tau_{1}), which ultimately leads to 𝒫~l=ΔΦl(τ1)/τ1\widetilde{\mathcal{P}}_{l}=\Delta\Phi_{l}(\tau_{1})/\tau_{1} being experimentally observable. Moreover, the oscillation with increasing frequency of σz(τ1)\sigma_{z}(\tau_{1}), shown in the insert of Fig. 2(c), is another evidence for the enhanced 𝒫~n\widetilde{\mathcal{P}}_{n} along with increasing the mechanical squeezing.

Quantum and classical Fisher information.— From a quantitative point of view, the fundamental bound to the sensitivity and the measurement-specific sensitivity of the proposed magnetometer are respectively decided by the QFI q\mathcal{F}_{q} and CFI c\mathcal{F}_{c} based on the Cramér-Rao inequality ΔBz1/Mj{\Delta}B_{z}\geq 1/\sqrt{M\mathcal{F}_{j}} (j=q,c)(j=q,c), where ΔBz\Delta B_{z} is the standard deviation with respect to an unbiased estimator (Bz)est(B_{z})_{\rm est}, and MM is the number of repetition of the experiments Braunstein and Caves (1994).

Considering system state |Ψ(t)|\Psi(t)\rangle, the QFI with respect to the parameter BzB_{z} reads q(t)=4(BzΨ(t)|BzΨ(t)|Ψ(t)|BzΨ(t)|2)\mathcal{F}_{q}(t)\!=\!4(\langle\partial_{B_{z}}\Psi(t)|\partial_{B_{z}}\Psi(t)\rangle\!-\!\left|\langle\Psi(t)|\partial_{B_{z}}\Psi(t)\rangle\right|^{2}), where Bz=/Bz\partial_{B_{z}}=\partial/\partial_{B_{z}}Paris (2009); Tóth and Apellaniz (2014); Liu et al. (2019); Lu and Wang (2021); Liu et al. (2022). At the first decoupled time τ1\tau_{1}, the QFI reduces to SM

q(τ1)\displaystyle\mathcal{F}_{q}(\tau_{1}) =32π2mλ12L2αmag2ωmE2𝒩1e12r,\displaystyle=\frac{32\pi^{2}m\lambda_{1}^{2}L^{2}\alpha^{2}_{\rm mag}}{\hbar\omega_{m}E^{2}}\mathcal{N}_{1}e^{12r},
=32π2mλ12L2αmag2ωmE2𝒩1[1(4λ2/ωm)𝒩2]3.\displaystyle=\frac{32\pi^{2}m\lambda_{1}^{2}L^{2}\alpha^{2}_{\rm mag}}{\hbar\omega_{m}E^{2}}\frac{\mathcal{N}_{1}}{[1-(4\lambda_{2}/\omega_{m})\mathcal{N}_{2}]^{3}}. (5)

where 𝒩1=|α|2\mathcal{N}_{1}=|\alpha|^{2} is the mean photon number of cavity mode a1a_{1}. It can be seen that the QFI is exponentially enhanced with power of 12r12r, and hence increasing a small value of rr by changing 𝒩2\mathcal{N}_{2} can give rise to a large enhancement of the QFI. Figure 3(a) shows an enhancement with 7 orders of magnitude for the QFI, corresponding to a dramatic reduction of the optimal sensitivity Δ𝔹z(τ1)=1/Mq(τ1)\Delta\mathbb{B}_{z}(\tau_{1})=1/\sqrt{M\mathcal{F}_{q}{(\tau_{1})}}. As shown in Fig. 3(b), the optimal sensitivity Δ𝔹z(τ1)\Delta\mathbb{B}_{z}(\tau_{1}) can reach to the order of 10151017(T/Hz)10^{-15}-10^{-17}\rm{(T/\sqrt{Hz})} for a wide parameter region in terms of the mean photon numbers 𝒩1{\mathcal{N}}_{1} and 𝒩2{\mathcal{N}}_{2}. Equation (Quantum Magnetometer with Dual-Coupling Optomechanics) also indicates that the resources 𝒩1{\mathcal{N}}_{1} and 𝒩2{\mathcal{N}}_{2} exert different influences on the ultimate lower bound of sensitivity SM . We note that the QFI at τ1\tau_{1} does not depend on the actual value of BzB_{z}.

Next, let us calculate the CFI related to a specific measurement on the quadrature Xθ=(a1eiθ+a1eiθ)/2X_{\theta}=(a_{1}e^{-i\theta}+a_{1}^{{\dagger}}e^{i\theta})/\sqrt{2}, where θ\theta is the phase of local oscillator. At the first decoupled time τ1\tau_{1}, the CFI is given by SM

c(τ1)=32π2mλ12L2αmag2ωmE2e12r(αResinθαImcosθ)2.\displaystyle\!\!\!\mathcal{F}_{c}(\tau_{1})\!=\!\frac{32\pi^{2}m\lambda_{1}^{2}L^{2}\alpha_{\rm mag}^{2}}{\hbar\omega_{m}E^{2}}e^{12r}(\alpha_{\rm Re}\sin\theta\!-\!\alpha_{\rm Im}\cos\theta)^{2}.\! (6)

Evidently, it consists precisely with the QFI shown in Eq. (Quantum Magnetometer with Dual-Coupling Optomechanics) when θ=π/2\theta=\pi/2 (θ=0\theta=0) and α\alpha is a real (imaginary) number, which means that the momentum (position) measurement can saturate the optimal sensitivity in the absence of system dissipation.

In the practical situation, the dissipation caused by the system-bath coupling should be taken into account. Then the dynamics of system is dominated by the master equation

ρ˙=i[H,ρ]+κ𝒟[a1]ρ+γ(n¯th+1)𝒟[b]ρ+γn¯th𝒟[b]ρ,\displaystyle\!\!\!\dot{\rho}\!=\!-\frac{i}{\hbar}[H,\rho]\!+\!\kappa\mathcal{D}[a_{1}]\rho\!+\!\gamma(\bar{n}_{\rm th}\!\!+\!1)\mathcal{D}[b]\rho\!+\!\gamma\bar{n}_{\rm th}\mathcal{D}[b^{{\dagger}}]\rho, (7)

where κ(γ)\kappa(\gamma) is the cavity (mechanical) decay rate, n¯th\bar{n}_{\rm th} is the thermal phonon number of the mechanical mode, and 𝒟[o]ρ=oρo(ooρ+ρoo)/2\mathcal{D}[o]\rho=o\rho o^{{\dagger}}-(o^{{\dagger}}o\rho+\rho o^{{\dagger}}o)/2. Here we have considered the cavity mode a2a_{2} being in the coherent state |ξ|\xi\rangle for Hamiltonian HH. Performing a momentum homodyne measurement (i.e., θ=π/2\theta=\pi/2) on cavity a1a_{1}, in Fig. 3(c) and Fig. 4, we numerically demonstrate the influence of system dissipation on the sensitivity limit, i.e., ΔBz(τ1)=1/Mc(τ1)\Delta B_{z}(\tau_{1})=1/\sqrt{M\mathcal{F}_{c}{(\tau_{1})}}, of magnetometer Johansson et al. (2012).

With the practical experimental parameters, Figure 3(c) shows that the specific sensitivity ΔBz(τ1)\Delta B_{z}(\tau_{1}) obtained in the presence of system decay and noise, still can fit well with the optimal one Δ𝔹z(τ1)\Delta\mathbb{B}_{z}(\tau_{1}) from the QFI without system dissipation. This consistency is only broken weakly when one increases the squeezing parameter rr to a large value. This can be explained as follows. On the one hand, system dissipation has little effect on the CFI c(τ1)\mathcal{F}_{c}(\tau_{1}) in the case of weak squeezing parameter rr [see the inserts of Fig. 4]. More importantly, the periodic optomechanical decoupling makes the CFI robust against the mechanical thermal noise. As shown in Fig. 4(b), a high sensitivity reaching to the order of 1015T/Hz\sim 10^{-15}{\rm T/\sqrt{Hz}} is allowed even when n¯th=103\bar{n}_{\rm th}=10^{3}. This means that the mechanical ground state cooling is not necessary for obtaining high-precision magnetometer in our proposal. On the other hand, the strong mechanical squeezing amplifies the effect of mechanical dissipation on CFI via effectively heating the environment (see the insert of Fig. 4(b) and Fig. S5 in supplementary material SM ), which leads to the weak disagreement between the measurement-specific sensitivity limit and the fundamental bound of sensitivity in the case of large values of rr.

Refer to caption
Figure 4: (a) Time evolution of c(t)\mathcal{F}_{c}(t) for different values of κ/ωm\kappa/\omega_{m} in the presence of mechanical dissipation γ/ωm=0.01\gamma/\omega_{m}=0.01, n¯th=10\bar{n}_{\rm th}=10 and considering r=0.8r=0.8. The shaded area indicates the time-window of detection with nearly flat c(t)\mathcal{F}_{c}(t) in the vicinity of the first decoupled time τ1\tau_{1}. (b) CFI c(τ1)\mathcal{F}_{c}(\tau_{1})(left y-axis) and specific sensitivity ΔBz(τ1)\Delta B_{z}(\tau_{1}) in units of T/Hz{\rm T/\sqrt{Hz}} (right y-axis) versus mechanical thermal phonon number n¯th\bar{n}_{\rm th} for κ/ωm=0.01\kappa/\omega_{m}=0.01 and γ/ωm=0.001\gamma/\omega_{m}=0.001. The insets show the influence of κ\kappa and γ\gamma on c(τ1)\mathcal{F}_{c}(\tau_{1}) for different values of squeezing parameter rr.

Discussion of experimental feasibility.— Regarding experimental implementations, while we have considered here a Febry-Pérot cavity with the membrane-in-the-middle configuration, our versatile proposal is not limited to this particular architecture. Based on the excellent controllability of the SQUID You and Nori (2011); Xiang et al. (2013), the transmission-line (TL) resonator coupled to a SQUID-terminated TL resonator is a promising platform to realize dual-coupling optomechanical system Johansson et al. (2014); Kim et al. (2015). Recently, the dual-coupling optomechanics was also demonstrated in photonic crystal cavities Kalaee et al. (2016); Brunelli et al. (2018) and whispering gallery microcavities Li et al. (2012). Based on recent optomechanical experiments Thompson et al. (2008); Sankey et al. (2010); Ockeloen-Korppi et al. (2018), here the system parameters can be chosen as m=4×108m=4\times 10^{-8} g, ωm=2π×134\omega_{m}=2\pi\times 134 kHz, λ1=8.4\lambda_{1}=8.4 kHz, λ2=0.08\lambda_{2}=0.08 Hz, κ=8.4\kappa=8.4 kHz, γ=840\gamma=840 Hz, and 𝒩1=106\mathcal{N}_{1}=10^{6}. Then, with a cycle time on tens of μ\mus, our work theoretically predicts that the sensitivity in the range of 10151017T/Hz10^{-15}\!-\!10^{-17}\,{\rm T/\sqrt{Hz}} can be realized with the achievable squeezing parameter r[0,0.9]r\in[0,0.9]. Note that most of the above results are obtained in the case of performing the homodyne measurement at the first decoupled time τ1\tau_{1}. Fortunately, our proposal is robust against the detection time, i.e., the CFI in the vicinity of τ1\tau_{1} is nearly flat as shown in Figure 4(a). In other words, our proposal exhibits a wide time-window to perform measurement with sensitivity reaching to the order of 1017T/Hz10^{-17}{\rm T}/\sqrt{\rm Hz}.

Moreover, the experimental implementation of our proposal relies on the thin membrane held by Terfenol-D rods, which gives an effective magnetic potential on the Hamiltonian. The conversion efficiency from the applied magnetic field to an effective force on the mechanical oscillator, is determined by the magnetic actuation constant cactc_{\rm act}. The precise experimental determination of the magnetic actuation constant is of key importance for measuring afterwards accurate values of the magnetic field. This magnetic constant depends heavily on the specific geometrical configuration of the rods with respect to the membrane. To optimize the magnetostrictive effect, the applied magnetic field needs to be paralleled to the magnetostricitive direction of the Terfenol-D rods. In addition, here we considered the Terfenol-D rods of the length 105\sim 10^{-5} m, thus it is approximatively valid to consider the magnetostrictive material to be in the whole homogeneous magnetic field, which also allows the system to have a large actuation constant.

Despite here we focus on detecting a static magnetic field, our proposal, in principle, can also be applied to probe the alternating magnetic fields. The corresponding frequency response characteristics are discussed in Sec. VII of the supplementary material SM . Referring to the parameters used in optomechanical experiments  Li et al. (2021); Thompson et al. (2008), we numerically simulate the displacement noise power spectrum and the corresponding force sensitivity at different probe powers (see Fig. S6 in the supplementary material). We find that the thermal-noise-limited frequency range covers 03800\sim 380 kHz with the central resonant frequency 250 kHz in the case of the probe power at 20nW20\,{\rm nW}. Therefore, similar as a general optomechanical system Li et al. (2021), here the dual-coupling magnetometer also has a broad bandwidth, when it is used as a resonant sensor.

Conclusions.– We have presented a protocol to measure the weak magnetic field using dual-coupling optomechanics. The sensitivity could be enhanced to the order of 10151017T/Hz10^{-15}-10^{-17}\,{\rm T/\sqrt{Hz}} by adjusting the photon numbers of two cavity modes. This enhancement originally comes from the periodic mechanical squeezing, which greatly amplifies both the signal to be measured and the transducing rate of signal. We stress out that this periodic squeezing effect is self-sustained, which avoids the complicated process of preparing squeezed states. Our proposal, with wide time-window of detection, is robust against the mechanical thermal noise, and hence the ground state cooling of mechanical mode is not necessary. This work might inspire the studies of high-precision measurements of various physical quantities based on dual-coupling optomechanical systems.

Acknowledgments.–We thank Bei-Bei Li for fruitful discussions and valuable comments. This work is supported by the National Key Research and Development Program of China grant 2021YFA1400700 and the National Science Foundation of China (Grant Nos. 11822502, 11974125, 11875029, 12175075 and 11805073).

References

  • Bennett et al. (2021) J. S. Bennett, B. E. Vyhnalek, H. Greenall, E. M. Bridge, F. Gotardo, S. Forstner, G. I. Harris, F. A. Miranda, and W. P. Bowen, Precision magnetometers for aerospace applications: A review, Sensors 21, 5568 (2021).
  • Hämäläinen et al. (1993) M. Hämäläinen, R. Hari, R. J. Ilmoniemi, J. Knuutila, and O. V. Lounasmaa, Magnetoencephalography-theory, instrumentation, and applications to noninvasive studies of the working human brain, Rev. Mod. Phys. 65, 413 (1993).
  • Lee et al. (2015) H. Lee, T.-H. Shin, J. Cheon, and R. Weissleder, Recent developments in magnetic diagnostic systems, Chem. Rev. 115, 10690 (2015).
  • Murzin et al. (2020) D. Murzin, D. J. Mapps, K. Levada, V. Belyaev, A. Omelyanchik, L. Panina, and V. Rodionova, Ultrasensitive magnetic field sensors for biomedical applications, Sensors 20, 1569 (2020).
  • Jaklevic et al. (1964) R. C. Jaklevic, J. Lambe, A. H. Silver, and J. E. Mercereau, Quantum interference effects in josephson tunneling, Phys. Rev. Lett. 12, 159 (1964).
  • Erné et al. (1976) S. N. Erné, H.-D. Hahlbohm, and H. Lübbig, Theory of rf-biased superconducting quantum interference device for nonhysteretic regime, J. Appl. Phys. 47, 5440 (1976).
  • Kleiner et al. (2004) R. Kleiner, D. Koelle, F. Ludwig, and J. Clarke, Superconducting quantum interference devices: State of the art and applications, Proceedings of the IEEE 92, 1534 (2004).
  • Allred et al. (2002) J. C. Allred, R. N. Lyman, T. W. Kornack, and M. V. Romalis, High-sensitivity atomic magnetometer unaffected by spin-exchange relaxation, Phys. Rev. Lett. 89, 130801 (2002).
  • Kominis et al. (2003) I. K. Kominis, T. W. Kornack, J. C. Allred, and M. V. Romalis, A subfemtotesla multichannel atomic magnetometer, Nature 422, 596 (2003).
  • Savukov et al. (2005) I. M. Savukov, S. J. Seltzer, M. V. Romalis, and K. L. Sauer, Tunable atomic magnetometer for detection of radio-frequency magnetic fields, Phys. Rev. Lett. 95, 063004 (2005).
  • Xia et al. (2006) H. Xia, A. Ben-Amar Baranga, D. Hoffman, and M. V. Romalis, Magnetoencephalography with an atomic magnetometer, Appl. Phys. Lett. 89, 211104 (2006).
  • Taylor et al. (2008) J. M. Taylor, P. Cappellaro, L. Childress, L. Jiang, D. Budker, P. R. Hemmer, A. Yacoby, R. Walsworth, and M. D. Lukin, High-sensitivity diamond magnetometer with nanoscale resolution, Nat. Phys. 4, 810 (2008).
  • Maze et al. (2008) J. R. Maze, P. L. Stanwix, J. S. Hodges, S. Hong, J. M. Taylor, P. Cappellaro, L. Jiang, M. V. G. Dutt, E. Togan, A. S. Zibrov, A. Yacoby, R. L. Walsworth, and M. D. Lukin, Nanoscale magnetic sensing with an individual electronic spin in diamond, Nature 455, 644 (2008).
  • Bending (1999) S. J. Bending, Local magnetic probes of superconductors, Advances in Physics 48, 449 (1999).
  • Robbes (2006) D. Robbes, Highly sensitive magnetometers—a review, Sens. Actu. A: Phys. 129, 86 (2006).
  • Aspelmeyer et al. (2014) M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavity optomechanics, Rev. Mod. Phys. 86, 1391 (2014).
  • Kippenberg and Vahala (2008) T. J. Kippenberg and K. J. Vahala, Cavity optomechanics: Back-action at the mesoscale, Science 321, 1172 (2008).
  • Aspelmeyer et al. (2012) M. Aspelmeyer, P. Meystre, and K. Schwab, Quantum optomechanics, Physics Today 65, 29 (2012).
  • Meystre (2013) P. Meystre, A short walk through quantum optomechanics, Ann. Phys. 525, 215 (2013).
  • Li and Zhu (2012) J.-J. Li and K.-D. Zhu, Nonlinear optical mass sensor with an optomechanical microresonator, Appl. Phys. Lett. 101, 141905 (2012).
  • Lin et al. (2017) Q. Lin, B. He, and M. Xiao, Mass sensing by detecting the quadrature of a coupled light field, Phys. Rev. A 96, 043812 (2017).
  • Bin et al. (2019) S.-W. Bin, X.-Y. Lü, T.-S. Yin, G.-L. Zhu, Q. Bin, and Y. Wu, Mass sensing by quantum criticality, Opt. Lett. 44, 630 (2019).
  • Clerk et al. (2010) A. A. Clerk, M. H. Devoret, S. M. Girvin, F. Marquardt, and R. J. Schoelkopf, Introduction to quantum noise, measurement, and amplification, Rev. Mod. Phys. 82, 1155 (2010).
  • Tsang and Caves (2010) M. Tsang and C. M. Caves, Coherent quantum-noise cancellation for optomechanical sensors, Phys. Rev. Lett. 105, 123601 (2010).
  • Pontin et al. (2014) A. Pontin, M. Bonaldi, A. Borrielli, F. S. Cataliotti, F. Marino, G. A. Prodi, E. Serra, and F. Marin, Detection of weak stochastic forces in a parametrically stabilized micro-optomechanical system, Phys. Rev. A 89, 023848 (2014).
  • Armata et al. (2017) F. Armata, L. Latmiral, A. D. K. Plato, and M. S. Kim, Quantum limits to gravity estimation with optomechanics, Phys. Rev. A 96, 043824 (2017).
  • Qvarfort et al. (2018) S. Qvarfort, A. Serafini, P. F. Barker, and S. Bose, Gravimetry through non-linear optomechanics, Nat. Commun. 9, 3690 (2018).
  • Forstner et al. (2012) S. Forstner, S. Prams, J. Knittel, E. D. van Ooijen, J. D. Swaim, G. I. Harris, A. Szorkovszky, W. P. Bowen, and H. Rubinsztein-Dunlop, Cavity optomechanical magnetometer, Phys. Rev. Lett. 108, 120801 (2012).
  • Forstner et al. (2014) S. Forstner, E. Sheridan, J. Knittel, C. L. Humphreys, G. A. Brawley, H. Rubinsztein-Dunlop, and W. P. Bowen, Ultrasensitive optomechanical magnetometry, Adv. Mater. 26, 6348 (2014).
  • Yu et al. (2016) C. Yu, J. Janousek, E. Sheridan, D. L. McAuslan, H. Rubinsztein-Dunlop, P. K. Lam, Y. Zhang, and W. P. Bowen, Optomechanical magnetometry with a macroscopic resonator, Phys. Rev. Appl. 5, 044007 (2016).
  • Wu et al. (2017) M. Wu, N. L.-Y. Wu, T. Firdous, F. Fani Sani, J. E. Losby, M. R. Freeman, and P. E. Barclay, Nanocavity optomechanical torque magnetometry and radiofrequency susceptometry, Nat. Nanotech. 12, 127 (2017).
  • Zhu et al. (2017) J. Zhu, G. Zhao, I. Savukov, and L. Yang, Polymer encapsulated microcavity optomechanical magnetometer, Sci. Rep. 7, 8896 (2017).
  • Li et al. (2018) B.-B. Li, J. Bílek, U. B. Hoff, L. S. Madsen, S. Forstner, V. Prakash, C. Schäfermeier, T. Gehring, W. P. Bowen, and U. L. Andersen, Quantum enhanced optomechanical magnetometry, Optica 5, 850 (2018).
  • Li et al. (2021) B.-B. Li, L. Ou, Y. Lei, and Y.-C. Liu, Cavity optomechanical sensing, Nanophotonics 10, 2799 (2021).
  • Colombano et al. (2020) M. F. Colombano, G. Arregui, F. Bonell, N. E. Capuj, E. Chavez-Angel, A. Pitanti, S. O. Valenzuela, C. M. Sotomayor-Torres, D. Navarro-Urrios, and M. V. Costache, Ferromagnetic resonance assisted optomechanical magnetometer, Phys. Rev. Lett. 125, 147201 (2020).
  • Olabi and Grunwald (2008) A. Olabi and A. Grunwald, Design and application of magnetostrictive materials, Materials & Design 29, 469 (2008).
  • Giovannetti et al. (2004) V. Giovannetti, S. Lloyd, and L. Maccone, Quantum-enhanced measurements: Beating the standard quantum limit, Science 306, 1330 (2004).
  • Giovannetti et al. (2006) V. Giovannetti, S. Lloyd, and L. Maccone, Quantum metrology, Phys. Rev. Lett. 96, 010401 (2006).
  • Giovannetti et al. (2011) V. Giovannetti, S. Lloyd, and L. Maccone, Advances in quantum metrology, Nat. Photon. 5, 222 (2011).
  • Dowling and Seshadreesan (2015) J. P. Dowling and K. P. Seshadreesan, Quantun optical technologies for metrology, sensing, and imaging, J. Lightwave Technol. 33, 2359 (2015).
  • Caves (1981) C. M. Caves, Quantum-mechanical noise in an interferometer, Phys. Rev. D 23, 1693 (1981).
  • Ma et al. (2011) J. Ma, X. Wang, C. Sun, and F. Nori, Quantum spin squeezing, Phys. Rep. 509, 89 (2011).
  • Baumgratz and Datta (2016) T. Baumgratz and A. Datta, Quantum enhanced estimation of a multidimensional field, Phys. Rev. Lett. 116, 030801 (2016).
  • Degen et al. (2017) C. L. Degen, F. Reinhard, and P. Cappellaro, Quantum sensing, Rev. Mod. Phys. 89, 035002 (2017).
  • Engelsen et al. (2017) N. J. Engelsen, R. Krishnakumar, O. Hosten, and M. A. Kasevich, Bell correlations in spin-squeezed states of 500 000 atoms, Phys. Rev. Lett. 118, 140401 (2017).
  • Nagata et al. (2007) T. Nagata, R. Okamoto, J. L. O’Brien, K. Sasaki, and S. Takeuchi, Beating the standard quantum limit with four-entangled photons, Science 316, 726 (2007).
  • Israel et al. (2014) Y. Israel, S. Rosen, and Y. Silberberg, Supersensitive polarization microscopy using noon states of light, Phys. Rev. Lett. 112, 103604 (2014).
  • Luo et al. (2017) X.-Y. Luo, Y.-Q. Zou, L.-N. Wu, Q. Liu, M.-F. Han, M. K. Tey, and L. You, Deterministic entanglement generation from driving through quantum phase transitions, Science 355, 620 (2017).
  • Jones et al. (2009) J. A. Jones, S. D. Karlen, J. Fitzsimons, A. Ardavan, S. C. Benjamin, G. A. D. Briggs, and J. J. L. Morton, Magnetic field sensing beyond the standard quantum limit using 10-spin noon states, Science 324, 1166 (2009).
  • Tanaka et al. (2015) T. Tanaka, P. Knott, Y. Matsuzaki, S. Dooley, H. Yamaguchi, W. J. Munro, and S. Saito, Proposed robust entanglement-based magnetic field sensor beyond the standard quantum limit, Phys. Rev. Lett. 115, 170801 (2015).
  • Kómár et al. (2014) P. Kómár, E. M. Kessler, M. Bishof, L. Jiang, A. S. Sørensen, J. Ye, and M. D. Lukin, A quantum network of clocks, Nat. Phys. 10, 582 (2014).
  • Hou et al. (2020) Z. Hou, Z. Zhang, G.-Y. Xiang, C.-F. Li, G.-C. Guo, H. Chen, L. Liu, and H. Yuan, Minimal tradeoff and ultimate precision limit of multiparameter quantum magnetometry under the parallel scheme, Phys. Rev. Lett. 125, 020501 (2020).
  • Holland and Burnett (1993) M. J. Holland and K. Burnett, Interferometric detection of optical phase shifts at the heisenberg limit, Phys. Rev. Lett. 71, 1355 (1993).
  • Boto et al. (2000) A. N. Boto, P. Kok, D. S. Abrams, S. L. Braunstein, C. P. Williams, and J. P. Dowling, Quantum interferometric optical lithography: Exploiting entanglement to beat the diffraction limit, Phys. Rev. Lett. 85, 2733 (2000).
  • Anisimov et al. (2010) P. M. Anisimov, G. M. Raterman, A. Chiruvelli, W. N. Plick, S. D. Huver, H. Lee, and J. P. Dowling, Quantum metrology with two-mode squeezed vacuum: Parity detection beats the heisenberg limit, Phys. Rev. Lett. 104, 103602 (2010).
  • Joo et al. (2011) J. Joo, W. J. Munro, and T. P. Spiller, Quantum metrology with entangled coherent states, Phys. Rev. Lett. 107, 083601 (2011).
  • Thompson et al. (2008) J. D. Thompson, B. M. Zwickl, A. M. Jayich, F. Marquardt, S. M. Girvin, and J. G. E. Harris, Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane, Nature 452, 72 (2008).
  • Sankey et al. (2010) J. C. Sankey, C. Yang, B. M. Zwickl, A. M. Jayich, and J. G. E. Harris, Strong and tunable nonlinear optomechanical coupling in a low-loss system, Nat. Phys. 6, 707 (2010).
  • Bhattacharya et al. (2008) M. Bhattacharya, H. Uys, and P. Meystre, Optomechanical trapping and cooling of partially reflective mirrors, Phys. Rev. A 77, 033819 (2008).
  • Zhu et al. (2018) G.-L. Zhu, X.-Y. Lü, L.-L. Wan, T.-S. Yin, Q. Bin, and Y. Wu, Controllable nonlinearity in a dual-coupling optomechanical system under a weak-coupling regime, Phys. Rev. A 97, 033830 (2018).
  • Hong (2013) C. Hong, Application of a magnetostrictive actuator, Materials and Design 46, 617 (2013).
  • (62) See more information in Supplementary Material.
  • Lü et al. (2015) X.-Y. Lü, Y. Wu, J. R. Johansson, H. Jing, J. Zhang, and F. Nori, Squeezed optomechanics with phase-matched amplification and dissipation, Phys. Rev. Lett. 114, 093602 (2015).
  • Lemonde et al. (2016) M.-A. Lemonde, N. Didier, and A. A. Clerk, Enhanced nonlinear interactions in quantum optomechanics via mechanical amplification, Nat. Commun. 7, 11338 (2016).
  • Braunstein and Caves (1994) S. L. Braunstein and C. M. Caves, Statistical distance and the geometry of quantum states, Phys. Rev. Lett. 72, 3439 (1994).
  • Paris (2009) M. G. A. Paris, Quantum esitimation for quantum technology, Int. J. Quantum Inf. 07, 125 (2009).
  • Tóth and Apellaniz (2014) G. Tóth and I. Apellaniz, Quantum metrology from a quantum information science perspective, J. Phys. A: Math. Theor. 47, 424006 (2014).
  • Liu et al. (2019) J. Liu, H. Yuan, X.-M. Lu, and X. Wang, Quantum fisher information matrix and multiparameter estimation, J. Phys. A: Math. Theor. 53, 023001 (2019).
  • Lu and Wang (2021) X.-M. Lu and X. Wang, Incorporating heisenberg’s uncertainty principle into quantum multiparameter estimation, Phys. Rev. Lett. 126, 120503 (2021).
  • Liu et al. (2022) J. Liu, M. Zhang, H. Chen, L. Wang, and H. Yuan, Optimal scheme for quantum metrology, Advanced Quantum Technologies 5, 2100080 (2022).
  • Johansson et al. (2012) J. Johansson, P. Nation, and F. Nori, Qutip: An open-source python framework for the dynamics of open quantum systems, Comput. Phys. Commun. 183, 1760 (2012).
  • You and Nori (2011) J. Q. You and F. Nori, Atomic physics and quantum optics using superconducting circuits, Nature 474, 589 (2011).
  • Xiang et al. (2013) Z.-L. Xiang, S. Ashhab, J. Q. You, and F. Nori, Hybrid quantum circuits: Superconducting circuits interacting with other quantum systems, Rev. Mod. Phys. 85, 623 (2013).
  • Johansson et al. (2014) J. R. Johansson, G. Johansson, and F. Nori, Optomechanical-like coupling between superconducting resonators, Phys. Rev. A 90, 053833 (2014).
  • Kim et al. (2015) E.-j. Kim, J. R. Johansson, and F. Nori, Circuit analog of quadratic optomechanics, Phys. Rev. A 91, 033835 (2015).
  • Kalaee et al. (2016) M. Kalaee, T. K. Paraïso, H. Pfeifer, and O. Painter, Design of a quasi-2d photonic crystal optomechanical cavity with tunable, large x2x^{2}-coupling, Opt. Express 24, 21308 (2016).
  • Brunelli et al. (2018) M. Brunelli, O. Houhou, D. W. Moore, A. Nunnenkamp, M. Paternostro, and A. Ferraro, Unconditional preparation of nonclassical states via linear-and-quadratic optomechanics, Phys. Rev. A 98, 063801 (2018).
  • Li et al. (2012) H.-K. Li, Y.-C. Liu, X. Yi, C.-L. Zou, X.-X. Ren, and Y.-F. Xiao, Proposal for a near-field optomechanical system with enhanced linear and quadratic coupling, Phys. Rev. A 85, 053832 (2012).
  • Ockeloen-Korppi et al. (2018) C. F. Ockeloen-Korppi, E. Damskägg, J.-M. Pirkkalainen, M. Asjad, A. A. Clerk, F. Massel, M. J. Woolley, and M. A. Sillanpää, Stabilized entanglement of massive mechanical oscillators, Nature 556, 478 (2018).

Supplementary Material for
“Quantum Magnetometer with Dual-Coupling Optomechanics”

Gui-Lei Zhu1, Jing Liu1,2, Ying Wu1, and Xin-You Lü1,{}^{1,^{*}}

1School of Physics, Huazhong University of Science and Technology, Wuhan 430074, China
2MOE Key Laboratory of Fundamental Physical Quantities Measurement and PGMF,
Huazhong University of Science and Technology, Wuhan 430074, China

Overview of the Supplemental Material

In this Supplementary Material, we present the technical details of the dual-coupling optomechanical magnetometer considered in the main text. In Sec. S1, we list the main symbols and parameters used in this work. In Sec. S2, we give a detailed description of the mechanical responses of the magnetostrictive material exposed to external magnetic fields, and present the system Hamiltonian in the form of second quantization. We also derive the effective Hamiltonian in the squeezed frame when the ancillary cavity mode a2a_{2} is in the coherent state. In Sec. S3, we present the detailed derivations of system evolution as well as the periodic squeezing effect during the dynamical evolution. The detailed derivation of state tomography at the first decoupled time τ1\tau_{1} is presented in Sec. S4. Moreover, in Sec. S5, we derive the QFI and the fundamental bound to the sensitivity of the proposed magnetometer. In Sec. S6, considering a momentum measurement on the cavity mode a1a_{1}, we derive the CFI corresponding to the specific sensitivity limit of the proposed magnetometer. By numerically solving the master equation, we illustrate the influence of system dissipations on the CFI and the associated specific sensitivity. In Sec. S7, we discuss the frequency response characteristics (e.g., the bandwidth), when the proposed dual-coupling optomechanical magnetometer is used to detect the alternating magnetic fields as a resonant sensor.

S1 System Parameters

In Table 1, we list the main parameters used in our proposal. All of them are experimentally feasible in the state-of-the-art setups.

Table 1: Main symbols and parameters have been used in this work. If not specified, these parameters are used in all calculations.
Symbols Parameters Value
mm Effective mass of mechanical oscillator 4×108g4\times 10^{-8}{\rm g}Thompson et al. (2008)
ωm\omega_{m} Mechanical frequency 2π×134kHz2\pi\times 134\,{\rm kHz}Thompson et al. (2008)
λ1\lambda_{1} Radiation-pressure coupling strength 8.4kHz8.4\,{\rm kHz}
λ2\lambda_{2} Quadratic optomechanical coupling strength 0.08Hz0.08{\rm Hz}
γ\gamma Mechanical decay rate (2.512)kHz(2.5\sim 12)\,{\rm kHz}
κ\kappa Cavity decay rate (8.484)kHz(8.4\sim 84)\,{\rm kHz}
n¯th\bar{n}_{\rm th} Thermal phonon occupation 10210310^{-2}\sim 10^{3}
k=mωm2k=m\omega_{m}^{2} Spring constant 28Nm128\,{\rm Nm^{-1}}Thompson et al. (2008)
LL The length of Terfenol-D rods 630 μm\mu{\rm m}
αmag\alpha_{\rm mag} Magnetostrictive coefficient along zz axis 5×108NT1m25\times 10^{8}\,{\rm NT^{-1}m^{-2}}Verhoeven et al. (1990)
EE Young’s modulus of Terfenol-D 30GPa30\,{\rm GPa}
cactc_{\rm act} Magnetic actuation constant 2.98×104NT12.98\times 10^{-4}\,{\rm NT^{-1}}
𝒩1=a1a1\mathcal{N}_{1}=\langle a_{1}^{{\dagger}}a_{1}\rangle Mean photon number of cavity mode a1a_{1} 10610^{6}
𝒩2=a2a2\mathcal{N}_{2}=\langle a_{2}^{{\dagger}}a_{2}\rangle Mean photon number of cavity mode a2a_{2} 02.45×1060\sim 2.45\times 10^{6}
r=0.25ln(14λ2𝒩2/ωm)r=-0.25{\rm ln}(1-4\lambda_{2}\mathcal{N}_{2}/\omega_{m}) Squeezing parameter 00.90\sim 0.9

S2 System Hamiltonian

The quantum system considered in this work is a dual-coupling optomechanical system, where two optical modes simultaneously coupled to the same mechanical oscillator, with the Hamiltonian

H=ω1a1a1+ω2a2a2+ωmbbλ1a1a1(b+b)λ2a2a2(b+b)2+Bzcactz,\displaystyle H=\hbar\omega_{1}a_{1}^{{\dagger}}a_{1}+\hbar\omega_{2}a_{2}^{{\dagger}}a_{2}+\hbar\omega_{m}b^{{\dagger}}b-\hbar\lambda_{1}a_{1}^{{\dagger}}a_{1}(b^{{\dagger}}+b)-\hbar\lambda_{2}a_{2}^{{\dagger}}a_{2}(b^{{\dagger}}+b)^{2}+B_{z}c_{\rm act}z, (S1)

where a1a_{1}, a2a_{2} and bb are the annihilation operators of optical cavity modes (with frequency ω1,ω2\omega_{1},\omega_{2}) and the mechanical oscillator (with frequency ωm\omega_{m}). Here λ1\lambda_{1} and λ2\lambda_{2} are the radiation-pressure coupling and quadratic optomechanical coupling strengths, respectively. The magnetostrictive potential has been included in the last term on the right-hand side. Let us first discuss in detail the generation of magnetostrictive potential.

Refer to caption
Figure S1: Infinitesimal volume element taken from interior of Terfenol-D rod. (a-c) show the variation of stress due to the presence of body force density fmag\emph{{f}}_{\rm mag} along x,y,zx,y,z axes, respectively. Stress tensor elements Tij(i=j)T_{ij}(i=j) and Tij(ij)T_{ij}(i\neq j) represent the pure stress and the shear stress, respectively.

S2.1 Generation of magnetostrictive potential

Schematic setup of our proposal is shown in Fig. 1 of the main text. We start by describing the connection between strain and the magnetic force acting on Terfenol-D rods. The Terfenol-D rods are subjected to a magnetic field with body force density fmag=fxx+fyy+fzz\emph{{f}}_{\rm mag}=f_{x}\vec{x}+f_{y}\vec{y}+f_{z}\vec{z}, where x,y\vec{x},\vec{y} and z\vec{z} are unit vectors along x,yx,y and zz axes, respectively. We divide up this object into infinitesimally small cubic volume element, with its edges aligned with the coordinate axes, as shown in Fig. S1. The small volume element has the size of dxdydz{\rm d}x{\rm d}y{\rm d}z. Essentially, the cube is acted upon the forces from the surrounding solid, and the magnetostrictive stress will vary with position. Therefore, we expand the stress tensor T in a Taylor series as a function of position within the cube. In a static equilibrium condition, the forces along x,y,zx,y,z coordinate axes are balanced [see Fig. S1], which is given by Cle

(Txx+Txxxdx)dydzTxxdydz+(Tyx+Tyxydy)dxdzTyxdxdz+(Tzx+Tzxzdz)dxdyTzxdxdy+fxdxdydz=0,\displaystyle(T_{xx}\!+\!\frac{\partial T_{xx}}{\partial{x}}{\rm d}x){\rm d}y{\rm d}z\!-\!T_{xx}{\rm d}y{\rm d}z\!+\!(T_{yx}\!+\!\frac{\partial T_{yx}}{\partial{y}}{\rm d}y){\rm d}x{\rm d}z\!-\!T_{yx}{\rm d}x{\rm d}z\!+\!(T_{zx}\!+\!\frac{\partial T_{zx}}{\partial{z}}{\rm d}z){\rm d}x{\rm d}y\!-\!T_{zx}{\rm d}x{\rm d}y\!+\!f_{x}{\rm d}x{\rm d}y{\rm d}z\!=\!0, (S2)
(Txy+Txyxdx)dydzTxydydz+(Tyy+Tyyydy)dxdzTyydxdz+(Tzy+Tzyzdz)dxdyTzydxdy+fydxdydz=0,\displaystyle(T_{xy}\!+\!\frac{\partial T_{xy}}{\partial{x}}{\rm d}x){\rm d}y{\rm d}z\!-\!T_{xy}{\rm d}y{\rm d}z\!+\!(T_{yy}\!+\!\frac{\partial T_{yy}}{\partial{y}}{\rm d}y){\rm d}x{\rm d}z\!-\!T_{yy}{\rm d}x{\rm d}z\!+\!(T_{zy}\!+\!\frac{\partial T_{zy}}{\partial{z}}{\rm d}z){\rm d}x{\rm d}y\!-\!T_{zy}{\rm d}x{\rm d}y\!+\!f_{y}{\rm d}x{\rm d}y{\rm d}z\!=\!0, (S3)
(Txz+Txzxdx)dydzTxzdydz+(Tyz+Tyzydy)dxdzTyzdxdz+(Tzz+Tzzzdz)dxdyTzzdxdy+fzdxdydz=0.\displaystyle(T_{xz}\!+\!\frac{\partial T_{xz}}{\partial{x}}{\rm d}x){\rm d}y{\rm d}z\!-\!T_{xz}{\rm d}y{\rm d}z\!+\!(T_{yz}\!+\!\frac{\partial T_{yz}}{\partial{y}}{\rm d}y){\rm d}x{\rm d}z\!-\!T_{yz}{\rm d}x{\rm d}z\!+\!(T_{zz}\!+\!\frac{\partial T_{zz}}{\partial{z}}{\rm d}z){\rm d}x{\rm d}y\!-\!T_{zz}{\rm d}x{\rm d}y\!+\!f_{z}{\rm d}x{\rm d}y{\rm d}z\!=\!0. (S4)

When dxdydz0{\rm d}x{\rm d}y{\rm d}z\neq 0, the above equations can be simplified as

Txxx+Tyxy+Tzxz+fx=0,\displaystyle\frac{\partial T_{xx}}{\partial{x}}+\frac{\partial T_{yx}}{\partial{y}}+\frac{\partial T_{zx}}{\partial{z}}+f_{x}=0, (S5)
Txyx+Tyyy+Tzyz+fy=0,\displaystyle\frac{\partial T_{xy}}{\partial{x}}+\frac{\partial T_{yy}}{\partial{y}}+\frac{\partial T_{zy}}{\partial{z}}+f_{y}=0, (S6)
Txzx+Tyzy+Tzzz+fz=0.\displaystyle\frac{\partial T_{xz}}{\partial{x}}+\frac{\partial T_{yz}}{\partial{y}}+\frac{\partial T_{zz}}{\partial{z}}+f_{z}=0. (S7)

Next, we assume a homogeneous magnetic field BzB_{z} oriented in the zz direction, and neglect the shear strain in the direction perpendicular to the tension. Then the magnetostrictive material only stretch along zz direction with magnetostrictive coefficient αmag\alpha_{\rm mag}, and the magnetostrictive-induced stress tensor T has only a single component Tzz=αmagBzT_{zz}=\alpha_{\rm mag}B_{z}. Simplifying Eqs. (S5-S7), we obtain the body force density

fz=Tzzz.\displaystyle f_{z}=-\frac{\partial T_{zz}}{\partial z}. (S8)

Under the action of the magnetic field, the magnetostrictive actuators stretch and further move the mount of suspended membrane (see Fig. 1 shown in main text). The displacement of an infinitesimally small cubic volume element at initial position r and time tt is Briant et al. (2003)

u(r,t)=Ψq(r)Xq(t),\displaystyle{\emph{{u}}}(\emph{{r}},t)={\Psi}_{q}(\emph{{r}})X_{q}(t), (S9)

where Ψq(r){\Psi}_{q}(\emph{{r}}) is the position-dependent mode shape function of eigenmode qq, which is normalized by VΨp(r)Ψq(r)dxdydz=δpqV\int_{V}{\Psi}_{p}(\emph{{r}})\cdot{\Psi}_{q}(\emph{{r}}){\rm d}x{\rm d}y{\rm dz}=\delta_{pq}V with VV being the spatial volume of oscillator. In addition, Xq(t)X_{q}(t) depends on the force applied on the membrane. Here, we consider the membrane having a single mechanical eigenmode with frequency ωm\omega_{m} and effective mass mm. The driving force received by the mechanical oscillator is Forstner et al. (2012)

Fmag=fmag(r)u(r)dV=fzΨz(r)dV=Bzcact.\displaystyle F_{\rm mag}=\int\emph{{f}}_{\rm mag}(\emph{{r}})\cdot\emph{{u}}(\emph{{r}}){\rm d}V=\int f_{z}{\Psi}_{z}(\emph{{r}}){\rm d}V=B_{z}c_{\rm act}. (S10)

where

cact=mωm2LαmagE,\displaystyle c_{\rm act}=m\omega_{m}^{2}L\frac{\alpha_{\rm mag}}{E}, (S11)

is the magnetic actuation constant charactering how well the magnetic field is converted into an applied force on the oscillator. Here LL is the length of Terfenol-D rod and EE is the Young’s modulus of Terfenol-D. The potential caused by the magnetic field can be written as

ΔU=Bzcactz,\displaystyle\Delta U=B_{z}c_{\rm act}z, (S12)

where z=/2mωm(b+b)z=\sqrt{\hbar/2m\omega_{m}}(b^{{\dagger}}+b) is the mechanical position operator. So far, we have obtained the magnetostrictive potential.

S2.2 Hamiltonian in the squeezed frame

By considering the ancillary mode a2a_{2} in a coherent state |ξ|\xi\rangle, the system Hamiltonian can be reduced to

H=ω1a1a1+ω2|ξ|2+ωmbbλ1a1a1(b+b)λ2|ξ|2(b+b)2+Bzcactz.\displaystyle H=\hbar\omega_{1}a_{1}^{{\dagger}}a_{1}+\hbar\omega_{2}|\xi|^{2}+\hbar\omega_{m}b^{{\dagger}}b-\hbar\lambda_{1}a_{1}^{{\dagger}}a_{1}(b^{{\dagger}}+b)-\hbar\lambda_{2}|\xi|^{2}(b^{{\dagger}}+b)^{2}+B_{z}c_{\rm act}z. (S13)

In deriving Hamiltonian (S13), we have assumed ξ|a2a2|ξ|ξ|2\langle\xi|a_{2}^{{\dagger}}a_{2}|\xi\rangle\approx|\xi|^{2}. Even though the coherent state |ξ|\xi\rangle is not the eigenstate of photon number operator a2a2a_{2}^{{\dagger}}a_{2}, the relative fluctuation is

Fre\displaystyle F_{\rm re} =ξ|a2a2a2a2|ξξ|a2a2|ξ2|ξ|2=1|ξ|.\displaystyle=\frac{\sqrt{\langle\xi|a_{2}^{{\dagger}}a_{2}a_{2}^{{\dagger}}a_{2}|\xi\rangle-\langle\xi|a_{2}^{{\dagger}}a_{2}|\xi\rangle^{2}}}{|\xi|^{2}}=\frac{1}{|\xi|}. (S14)

It is clearly shown that the higher ξ\xi, the smaller relative fluctuation. In our calculations, we considered ξ1550\xi\approx 1550, whose fluctuation is much smaller than mean-photon number and can be neglected safely.

Refer to caption
Figure S2: Effective mechanical frequency ωs\omega_{s} (left-axis), optomechanical coupling strength λs{\lambda}_{s}(right-axis) and fsf_{s}(inset) versus the squeezing parameter rr with ωm=1\omega_{m}=1.

Next, we apply a Schrieffer-Wolff transformation Hs=S(r)HS(r)H_{s}=S(r)HS^{{\dagger}}(r) to Eq. (S13) with squeezing operator S(r)=exp[r(b2b2)/2]S(r)=\exp{[r(b^{2}-b^{{\dagger}2})/2]} and squeezing parameter r=(1/4)ln(14λ2|ξ|2/ωm)r=-(1/4)\ln(1-4\lambda_{2}|\xi|^{2}/\omega_{m}). In view of

S(r)bS(r)=bcoshr+bsinhr,\displaystyle S(r)bS^{{\dagger}}(r)=b\cosh r+b^{{\dagger}}\sinh r, (S15)
S(r)bS(r)=bcoshr+bsinhr,\displaystyle S(r)b^{{\dagger}}S^{{\dagger}}(r)=b^{{\dagger}}\cosh r+b\sinh r, (S16)

we obtain the Hamiltonian in the squeezed frame is

Hs/=ω1a1a1+ωsbbλsa1a1(b+b)+fs(b+b)+E0.\displaystyle H_{s}/\hbar=\omega_{1}a_{1}^{{\dagger}}a_{1}+\omega_{s}b^{{\dagger}}b-\lambda_{s}a_{1}^{{\dagger}}a_{1}(b^{{\dagger}}+b)+f_{s}(b^{{\dagger}}+b)+E_{0}. (S17)

where ωs=ωme2r\omega_{s}=\omega_{m}e^{-2r}, λs=λ1er\lambda_{s}=\lambda_{1}e^{r} and fs=fer=1/(2mωm)Bzcacterf_{s}=fe^{r}=\sqrt{{1}/{(2m\hbar\omega_{m})}}B_{z}c_{\rm act}e^{r}. Here E0=ωm(e2r1)/2+ω2|ξ|2E_{0}=\omega_{m}(e^{-2r}-1)/2+\omega_{2}|\xi|^{2} is the constant term and will be omitted in the following discussions for the sake of simplicity. In Fig. S2 we plot the effective mechanical frequency ωs\omega_{s} and optomechanical coupling strength λs\lambda_{s} versus the squeezing parameter rr. Evidently, increasing the parameter rr enables an exponential reduction of mechanical frequency, but an exponential enhancement of the radiation-pressure optomechanical coupling. Moreover, from the fourth term of Hamiltonian HsH_{s} we observe that the magnetic signal to be measured is exponentially amplified [also see the insert of Fig. S2], which directly gives rise to the enhancement of estimation precision of magnetic field.

S3 System dynamics with periodic mechanical squeezing

S3.1 Periodic opto-mechanical decoupling

Starting from the original Hamiltonian Eq. (S13) under the condition of applying the mean field approximation for mode a2a_{2}, we can write the evolution operator of system as

U(t)\displaystyle U(t) =ei(λ~sa1a1f~s)2[ωstsin(ωst)]S(r)e(λ~sa1a1f~s)(μbμb)eibbωstS(r),\displaystyle=e^{i(\tilde{\lambda}_{s}a_{1}^{{\dagger}}a_{1}-\tilde{f}_{s})^{2}[\omega_{s}t-\sin(\omega_{s}t)]}S^{{\dagger}}(r)e^{(\tilde{\lambda}_{s}a_{1}^{{\dagger}}a_{1}-\tilde{f}_{s})(\mu b^{{\dagger}}-\mu^{*}b)}e^{-ib^{{\dagger}}b\omega_{s}t}S(r),
=ei(λ~sa1a1f~s)2[ωstsin(ωst)]S(r)S(r)e(λ~sa1a1f~s)(μ¯bμ¯b)eibbωst,\displaystyle=e^{i(\tilde{\lambda}_{s}a_{1}^{{\dagger}}a_{1}-\tilde{f}_{s})^{2}[\omega_{s}t-\sin(\omega_{s}t)]}S^{{\dagger}}(r)S(r^{\prime})e^{(\tilde{\lambda}_{s}a_{1}^{{\dagger}}a_{1}-\tilde{f}_{s})(\bar{\mu}b^{{\dagger}}-\bar{\mu}^{*}b)}e^{-ib^{{\dagger}}b\omega_{s}t}, (S18)

where

λ~s\displaystyle\tilde{\lambda}_{s} =λs/ωs=λ1ωme3r,\displaystyle=\lambda_{s}/\omega_{s}=\frac{\lambda_{1}}{\omega_{m}}e^{3r}, (S19)
f~s\displaystyle\tilde{f}_{s} =fs/ωs=Bzcact12mωm3e3r,\displaystyle=f_{s}/\omega_{s}=B_{z}c_{\rm act}\sqrt{\frac{1}{2m\hbar\omega_{m}^{3}}}e^{3r}, (S20)

are the rescaled parameter, and μ¯=(1eiωst)(coshreiωstsinhr)\bar{\mu}=(1-e^{-i\omega_{s}t})(\cosh r-e^{-i\omega_{s}t}\sinh r), S(r)=exp[(rb2rb2)/2]S(r^{\prime})=\exp{[(r^{\prime*}b^{2}-r^{\prime}b^{{\dagger}2})/2]} with r=re2iωstr^{\prime}=re^{-2i\omega_{s}t}. In derivating Eq. (S3.1), we have adopted a rotating frame with exp(iω~1a1a1)\exp(-i\tilde{\omega}_{1}a_{1}^{{\dagger}}a_{1}) where ω~1=ω1/ωs\tilde{\omega}_{1}=\omega_{1}/\omega_{s}. Under the condition that the optical and mechanical modes are initially in the coherent state |α|β|\alpha\rangle|\beta\rangle with β=βRe+iβIm\beta=\beta_{\rm Re}+i\beta_{\rm Im}, the state at time tt is

|Ψ(t)\displaystyle|\Psi(t)\rangle =ei(λ~sa1a1f~s)2[ωstsin(ωst)]S(r)S(r)e(λ~sa1a1f~s)(μ¯bμ¯b)eibbωst|α|β\displaystyle=e^{i(\tilde{\lambda}_{s}a_{1}^{{\dagger}}a_{1}-\tilde{f}_{s})^{2}[\omega_{s}t-\sin(\omega_{s}t)]}S^{{\dagger}}(r)S(r^{\prime})e^{(\tilde{\lambda}_{s}a_{1}^{{\dagger}}a_{1}-\tilde{f}_{s})(\bar{\mu}b^{{\dagger}}-\bar{\mu}^{*}b)}e^{-ib^{{\dagger}}b\omega_{s}t}|\alpha\rangle|\beta\rangle
=e|α|2/2n=0αnn!ei(λ~snf~s)2[ωstsin(ωst)]|n(S(r)S(r)e(λ~snf~s)(μ¯bμ¯b)|eiωstβ)\displaystyle=e^{-|\alpha|^{2}/2}\sum_{n=0}^{\infty}\frac{\alpha^{n}}{\sqrt{n!}}e^{i(\tilde{\lambda}_{s}n-\tilde{f}_{s})^{2}[\omega_{s}t-\sin(\omega_{s}t)]}|n\rangle\left(S^{{\dagger}}(r)S(r^{\prime})e^{(\tilde{\lambda}_{s}n-\tilde{f}_{s})(\bar{\mu}b^{{\dagger}}-\bar{\mu}^{*}b)}|e^{-i\omega_{s}t}\beta\rangle\right)
=e|α|2/2n=0αnn!ei(λ~snf~s)2[ωstsin(ωst)]|n(S(r)S(r)D[(λ~snf~s)μ¯]D(eiωstβ)|0)\displaystyle=e^{-|\alpha|^{2}/2}\sum_{n=0}^{\infty}\frac{\alpha^{n}}{\sqrt{n!}}e^{i(\tilde{\lambda}_{s}n-\tilde{f}_{s})^{2}[\omega_{s}t-\sin(\omega_{s}t)]}|n\rangle\left(S^{{\dagger}}(r)S(r^{\prime})D[(\tilde{\lambda}_{s}n-\tilde{f}_{s})\bar{\mu}]D(e^{-i\omega_{s}t}\beta)|0\rangle\right)
=e|α|2/2n=0αnn!ei(λ~snf~s)2[ωstsin(ωst)]ei(λ~snf~s)er[βResinωstβIm(cosωst1)e2r]|n(S(r)S(r)|φn(t)),\displaystyle=e^{-|\alpha|^{2}/2}\sum_{n=0}^{\infty}\frac{\alpha^{n}}{\sqrt{n!}}e^{i(\tilde{\lambda}_{s}n-\tilde{f}_{s})^{2}[\omega_{s}t-\sin(\omega_{s}t)]}e^{i(\tilde{\lambda}_{s}n-\tilde{f}_{s})e^{-r}[\beta_{\rm Re}\sin\omega_{s}t-\beta_{\rm Im}(\cos\omega_{s}t-1)e^{2r}]}|n\rangle\left(S^{{\dagger}}(r)S(r^{\prime})|\varphi_{n}(t)\right\rangle), (S21)

where φn(t)=eiωstβ+(λ~snf~s)μ¯\varphi_{n}(t)=e^{-i\omega_{s}t}\beta+(\tilde{\lambda}_{s}n-\tilde{f}_{s})\bar{\mu}. Here D(β)=e|β|2/2eβbeβbD(\beta)=e^{-|\beta|^{2}/2}e^{\beta b^{{\dagger}}}e^{-\beta^{*}b} is the displacement operator, and we have used D(α+β)=D(α)D(β)exp[iIm(αβ)]D(\alpha+\beta)=D(\alpha)D(\beta)\exp[-i{\rm Im}(\alpha\beta^{*})]. By tracing out the mechanical part, we obtain the reduced cavity state

ρc(t)\displaystyle\rho_{c}(t) =e|α|2n,n[αn(α)nn!n!ei[(λ~snf~s)2(λ~snf~s)2][ωstsin(ωst)]eiλ~s(nn)er[βResinωstβIm(cosωst1)e2r]\displaystyle=e^{-|\alpha|^{2}}\sum_{n,n^{\prime}}\left[\frac{\alpha^{n}(\alpha^{*})^{n^{\prime}}}{\sqrt{n!n^{\prime}!}}e^{i\left[(\tilde{\lambda}_{s}n-\tilde{f}_{s})^{2}-(\tilde{\lambda}_{s}n^{\prime}-\tilde{f}_{s})^{2}\right][\omega_{s}t-\sin(\omega_{s}t)]}e^{i\tilde{\lambda}_{s}(n-n^{\prime})e^{-r}[\beta_{\rm Re}\sin\omega_{s}t-\beta_{\rm Im}(\cos\omega_{s}t-1)e^{2r}]}\right.
×e(|φn|2+|φn|2)/2+φnφn|nn|].\displaystyle\left.\times e^{-(|\varphi_{n}|^{2}+|\varphi_{n^{\prime}}|^{2})/2+\varphi_{n^{\prime}}^{*}\varphi_{n}}|n\rangle\langle n^{\prime}|\right]. (S22)

Interestingly, at time τm=2mπ/ωs(m=1,2,3)\tau_{m}=2m\pi/\omega_{s}(m=1,2,3...), the parameters μ¯=0,r=r,S(r)S(r)=1\bar{\mu}=0,r^{\prime}=r,S^{{\dagger}}(r^{\prime})S(r)=1 and |φn(τm)=|β|\varphi_{n}(\tau_{m})\rangle=|\beta\rangle, and then Eq. (S21) is reduced to

|Ψ(τm)\displaystyle|\Psi(\tau_{m})\rangle =e|α|2/2n=0αnn!ei2mπ(λ~snf~s)2|n|β,\displaystyle=e^{-|\alpha|^{2}/2}\sum_{n=0}^{\infty}\frac{\alpha^{n}}{\sqrt{n!}}e^{i2m\pi(\tilde{\lambda}_{s}n-\tilde{f}_{s})^{2}}|n\rangle|\beta\rangle,
=e|α|2/2n=0αnn!eiΦn(τm)|n|β.\displaystyle=e^{-|\alpha|^{2}/2}\sum_{n=0}^{\infty}\frac{\alpha^{n}}{\sqrt{n!}}e^{i\Phi_{n}(\tau_{m})}|n\rangle|\beta\rangle. (S23)

By tracing out the mechanical mode, the corresponding reduced density matrix becomes

ρc(τm)=e|α|2n,nαnαnn!n!ei[Φn(τm)Φn(τm)]|nn|.\displaystyle\rho_{c}(\tau_{m})=e^{-|\alpha|^{2}}\sum_{n,n^{\prime}}\frac{\alpha^{n}\alpha^{*n^{\prime}}}{\sqrt{n!n^{\prime}!}}e^{i[\Phi_{n}(\tau_{m})-\Phi_{n^{\prime}}(\tau_{m})]}|n\rangle\langle n^{\prime}|. (S24)

It is shown from Eq. (S3.1) that, at time τm\tau_{m}, the state of mechanical part back to its initial state (i.e., |β|\beta\rangle). This demonstrates that the mechanical and optical modes completely decoupled at this time. Meanwhile the signal to be measured has been transferred into the phase Φn(τm)=2mπ(λ~snf~s)2\Phi_{n}(\tau_{m})=2m\pi(\tilde{\lambda}_{s}n-\tilde{f}_{s})^{2}. The period of optical-mechanical decoupling is τ1=2π/ωs\tau_{1}=2\pi/\omega_{s}.

S3.2 Periodic mechanical squeezing effect

As shown in the main text, there is a periodic squeezing effect on the mechanical mode associated with the system evolution, which is also demonstrated in Eq. (S21). To qualitatively illustrate this periodic squeezing effect, for simplicity, we consider cavity a1a_{1} in vacuum state and the auxiliary cavity a2a_{2} with amplitude ξ=1\xi=1, and neglect the magnetostrictive potential for a while. Then system Hamiltonian (S13) can be simplified to

H/=ωmbbλ2(b+b)2.\displaystyle H/\hbar=\omega_{m}b^{\dagger}b-\lambda_{2}\left(b^{\dagger}+b\right)^{2}. (S25)

Applying a squeezing transformation Hs=S(r)HS(r)H_{s}=S(r)HS^{{\dagger}}(r) with r=0.25ln(14λ2/ωm)r=-0.25{\rm ln}(1-4\lambda_{2}/\omega_{m}), the resulting Hamiltonian (S25) in squeezed frame is then of the form

Hs/=ωsbb,\displaystyle H_{s}/\hbar=\omega_{s}b^{\dagger}b, (S26)

where ωs=e2rω\omega_{s}=e^{-2r}\omega. Accordingly, the time evolution operator reads

U(t)=S(r)eiωstbbS(r),\displaystyle U(t)=S^{\dagger}\left(r\right)e^{-i\omega_{s}tb^{\dagger}b}S\left(r\right),

where S(r)=e(1/2)(rb2rb2)S(r)=e^{\left(1/2\right)\left(rb^{2}-rb^{\dagger 2}\right)}. The instantaneous state of system is

|Ψ(t)=\displaystyle|\Psi(t)\rangle= S(r)eiωstbbS(r)|β,\displaystyle S^{\dagger}\left(r\right)e^{-i\omega_{s}tb^{\dagger}b}S\left(r\right)|\beta\rangle,
=\displaystyle= S(r)S(r)|β(t),\displaystyle S^{\dagger}\left(r\right)S\left(r^{\prime}\right)|\beta(t)\rangle, (S27)

where S(r)=e(1/2)(rb2rb2)S\left(r^{\prime}\right)=e^{\left(1/2\right)\left(r^{\prime\ast}b^{2}-r^{\prime}b^{\dagger 2}\right)} with r=re2iωstr^{\prime}=re^{-2i\omega_{s}t}, and β(t)=eiωstβ\beta(t)=e^{-i\omega_{s}t}\beta. To visualize the periodic squeezing behavior, we illustrated the instantaneous squeezing degree of mechanical mode in Fig. 2(a) of the main text. The detailed derivation is shown in the following. Instantaneous mean values of X2\langle X\rangle^{2} and X2\langle X^{2}\rangle with X=(b+b)/2X=(b+b^{{\dagger}})/\sqrt{2} being a dimensionless position operator, are

Ψ(t)|X|Ψ(t)2=2e2r(cosh2rRe[β(t)]22sinhrcoshrRe[β(t)]Re[(β(t)e2iωst)]+sinh2rRe[β(t)e2iωst]2),\displaystyle\langle\Psi(t)|X|\Psi(t)\rangle^{2}={2}e^{2r}\left(\cosh^{2}r{\rm Re}[\beta(t)]^{2}-2\sinh r\cosh r{\rm Re}[\beta(t)]{\rm Re}[(\beta(t)e^{2i\omega_{s}t})]+\sinh^{2}r{\rm Re}[\beta(t)e^{2i\omega_{s}t}]^{2}\right), (S28)
Ψ(t)|X2|Ψ(t)2=e2r(2(Re[β(t)])2cosh2r+2Re[β(t)e2iωst]2sinh2r4Re[β(t)e2iωst]Re[β(t)]sinhrcoshr+12(sinh2r+cosh2r)12(e2iωst+e2iωst)sinhrcoshr).\displaystyle\langle\Psi(t)|X^{2}|\Psi(t)\rangle^{2}={e^{2r}}\left(\begin{array}[]{c}2({\rm Re}[\beta(t)])^{2}\cosh^{2}r+2{\rm Re}[\beta(t)e^{2i\omega_{s}t}]^{2}\sinh^{2}r-4{\rm Re}[\beta(t)e^{2i\omega_{s}t}]{\rm Re}[\beta(t)]\sinh r\cosh r\\ +\frac{1}{2}(\sinh^{2}r+\cosh^{2}r)-\frac{1}{2}(e^{-2i\omega_{s}t}+e^{2i\omega_{s}t})\sinh r\cosh r\end{array}\right). (S31)

Then we obtain the variance of XX

δ2X(t)\displaystyle\langle\delta^{2}X(t)\rangle =X(t)2X(t)2=e2r2[cosh(2r)cos(2ωst)sinh(2r)].\displaystyle=\langle X(t)^{2}\rangle-\langle X(t)\rangle^{2}=\frac{e^{2r}}{2}\left[\cosh(2r)-\cos(2\omega_{s}t)\sinh(2r)\right]. (S33)

Fig. S3 clearly shows that δ2X(t)\langle\delta^{2}X(t)\rangle evolves with a period T=π/ωsT=\pi/\omega_{s}. It reaches the minimum δX2min=1/2\langle\delta X^{2}\rangle_{\rm min}=1/2 at t=mπ/ωs(m=0,1,2,)t=m\pi/\omega_{s}(m=0,1,2,...) and the maximum δX2max=e4r/2\langle\delta X^{2}\rangle_{\rm max}=e^{4r}/2 at t=mπ/(2ωs)(m=1,2,3,)t=m\pi/(2\omega_{s})(m=1,2,3,...). For a certain squeezing parameter rr, we define the instantaneous squeezing degree S(t)=10log10(δ2X(t)/δ2Xmin)dBS(t)=10\log_{10}\left({\langle\delta^{2}X(t)\rangle}/{\langle\delta^{2}X\rangle_{\rm min}}\right){\rm dB}, and its maximum value is

Smax\displaystyle S_{\rm max} =10log10(δ2Xmaxδ2Xmin)dB,\displaystyle=10\log_{10}\left(\frac{\langle\delta^{2}X\rangle_{\rm max}}{\langle\delta^{2}X\rangle_{\rm min}}\right){\rm dB},
=10log10(e4r)dB.\displaystyle=10\log_{10}(e^{4r})\,{\rm dB}. (S34)
Refer to caption
Figure S3: Variance of XX versus time for different squeezing parameter rr.

S4 State tomography

As discussed in the main text, the well-defined displaced phase-accumulation-efficiency 𝒫~n\widetilde{\mathcal{P}}_{n} is experimentally observable. Specifically, by projecting the system density matrix ρc(τ1)\rho_{c}(\tau_{1}) on the basis vectors {|,|}\{|\!\uparrow\rangle,|\!\downarrow\rangle\}, where |=(1/2)(|0+|l)|\!\uparrow\rangle=(1/\sqrt{2})(|0\rangle+|l\rangle), |=(1/2)(|0|l)|\!\downarrow\rangle=(1/\sqrt{2})(|0\rangle-|l\rangle) with Fock state of cavity mode |l|l\rangle, we obtain

ρ(τ1)\displaystyle\rho_{\downarrow\downarrow}(\tau_{1}) =12(0|l|)(e|α|2n,nαnαnn!n!ei(Φn(τ1)Φn(τ1))|nn|)(|0|l),\displaystyle=\frac{1}{2}(\langle 0|-\langle l|)\left(e^{-|\alpha|^{2}}\sum_{n,n^{\prime}}\frac{\alpha^{n}\alpha^{*n^{\prime}}}{\sqrt{n!n^{\prime}!}}e^{i(\Phi_{n}(\tau_{1})-\Phi_{n^{\prime}}(\tau_{1}))}|n\rangle\langle n^{\prime}|\right)(|0\rangle-|l\rangle),
=e|α|22[1+α2ll!αll!(eiΦ0(τ1)iΦl(τ1)+eiΦl(τ1)iΦ0(τ1))],\displaystyle=\frac{e^{-|\alpha|^{2}}}{2}\left[1+\frac{\alpha^{2l}}{l!}-\frac{\alpha^{l}}{\sqrt{l!}}\left(e^{i\Phi_{0}(\tau_{1})-i\Phi_{l}(\tau_{1})}+e^{i\Phi_{l}(\tau_{1})-i\Phi_{0}(\tau_{1})}\right)\right],
=e|α|22[1+α2ll!αll!2cos(2π(λ~s2l22f~sλ~sl))].\displaystyle=\frac{e^{-|\alpha|^{2}}}{2}\left[1+\frac{\alpha^{2l}}{l!}-\frac{\alpha^{l}}{\sqrt{l!}}2\cos\left(2\pi(\tilde{\lambda}_{s}^{2}l^{2}-2\tilde{f}_{s}\tilde{\lambda}_{s}l)\right)\right]. (S35)

As above we can obtain the other elements

ρ(τ1)\displaystyle\rho_{\uparrow\uparrow}(\tau_{1}) =12(0|+l|)(e|α|2n,nαnαnn!n!ei(Φn(τ1)Φn(τ1))|nn|)(|0+|l),\displaystyle=\frac{1}{2}(\langle 0|+\langle l|)\left(e^{-|\alpha|^{2}}\sum_{n,n^{\prime}}\frac{\alpha^{n}\alpha^{*n^{\prime}}}{\sqrt{n!n^{\prime}!}}e^{i(\Phi_{n}(\tau_{1})-\Phi_{n^{\prime}}(\tau_{1}))}|n\rangle\langle n^{\prime}|\right)(|0\rangle+|l\rangle),
=e|α|22[1+α2ll!+αll!2cos(2π(λ~s2l22f~sλ~sl))].\displaystyle=\frac{e^{-|\alpha|^{2}}}{2}\left[1+\frac{\alpha^{2l}}{l!}+\frac{\alpha^{l}}{\sqrt{l!}}2\cos\left(2\pi(\tilde{\lambda}_{s}^{2}l^{2}-2\tilde{f}_{s}\tilde{\lambda}_{s}l)\right)\right]. (S36)
ρ(τ1)\displaystyle\rho_{\downarrow\uparrow}(\tau_{1}) =12(0|l|)(e|α|2n,nαnαnn!n!ei(Φn(τ1)Φn(τ1))|nn|)(|0+|l),\displaystyle=\frac{1}{2}(\langle 0|-\langle l|)\left(e^{-|\alpha|^{2}}\sum_{n,n^{\prime}}\frac{\alpha^{n}\alpha^{*n^{\prime}}}{\sqrt{n!n^{\prime}!}}e^{i(\Phi_{n}(\tau_{1})-\Phi_{n^{\prime}}(\tau_{1}))}|n\rangle\langle n^{\prime}|\right)(|0\rangle+|l\rangle),
=e|α|22[1α2ll!+αll!2isin(2π(λ~s2l22f~sλ~sl))].\displaystyle=\frac{e^{-|\alpha|^{2}}}{2}\left[1-\frac{\alpha^{2l}}{l!}+\frac{\alpha^{l}}{\sqrt{l!}}2i\sin\left(2\pi(\tilde{\lambda}_{s}^{2}l^{2}-2\tilde{f}_{s}\tilde{\lambda}_{s}l)\right)\right]. (S37)
ρ(τ1)\displaystyle\rho_{\uparrow\downarrow}(\tau_{1}) =12(0|l|)(e|α|2n,nαnαnn!n!ei(Φn(τ1)Φn(τ1))|nn|)(|0+|l),\displaystyle=\frac{1}{2}(\langle 0|-\langle l|)\left(e^{-|\alpha|^{2}}\sum_{n,n^{\prime}}\frac{\alpha^{n}\alpha^{*n^{\prime}}}{\sqrt{n!n^{\prime}!}}e^{i(\Phi_{n}(\tau_{1})-\Phi_{n^{\prime}}(\tau_{1}))}|n\rangle\langle n^{\prime}|\right)(|0\rangle+|l\rangle),
=e|α|22[1α2ll!αll!2isin(2π(λ~s2l22f~sλ~sl))].\displaystyle=\frac{e^{-|\alpha|^{2}}}{2}\left[1-\frac{\alpha^{2l}}{l!}-\frac{\alpha^{l}}{\sqrt{l!}}2i\sin\left(2\pi(\tilde{\lambda}_{s}^{2}l^{2}-2\tilde{f}_{s}\tilde{\lambda}_{s}l)\right)\right]. (S38)

These matrix elements are shown in Fig. 2(b) of the main text. The difference of diagonal elements is

σz(τ1)=ρ(τ1)ρ(τ1)=2e|α|2αll!cos[Φl(τ1)Φ0(τ1)].\displaystyle\sigma_{z}(\tau_{1})=\rho_{\uparrow\uparrow}(\tau_{1})-\rho_{\downarrow\downarrow}(\tau_{1})=2\,e^{-|\alpha|^{2}}\frac{\alpha^{l}}{\sqrt{l!}}\cos\left[\Phi_{l}(\tau_{1})-\Phi_{0}(\tau_{1})\right]. (S39)

Evidently, we can obtain the phase difference ΔΦl(τ1)=Φl(τ1)Φ0(τ1)\Delta\Phi_{l}(\tau_{1})=\Phi_{l}(\tau_{1})-\Phi_{0}(\tau_{1}) by directly measuring ρ(τ1)\rho_{\downarrow\downarrow}(\tau_{1}) and ρ(τ1)\rho_{\uparrow\uparrow}(\tau_{1}), and then 𝒫~l=ΔΦl(τ1)/τ1\widetilde{\mathcal{P}}_{l}=\Delta\Phi_{l}(\tau_{1})/\tau_{1} is experimentally observable.

S5 Quantum Fisher information and ultimate bound of sensitivity

In this section, we will present the detailed derivation of QFI in our model. From Eq. (S3.1), we can obtain the state at time τ1\tau_{1} reads

|Ψ(τ1)\displaystyle|\Psi(\tau_{1})\rangle =e|α|2/2nαnn!ei2π(λ~snf~s)2|n|β.\displaystyle=e^{-|\alpha|^{2}/2}\sum_{n}\frac{\alpha^{n}}{\sqrt{n!}}e^{i2\pi(\tilde{\lambda}_{s}n-\tilde{f}_{s})^{2}}|n\rangle|\beta\rangle. (S40)

As discussed before, the optical and mechanical modes absolutely decoupled at time τ1\tau_{1}, and all information about BzB_{z} is transduced into optical phase Φn(τ1)\Phi_{n}(\tau_{1}). Starting from the system state |Ψ(τ1)|\Psi(\tau_{1})\rangle, the QFI is given by Paris (2009); Tóth and Apellaniz (2014); Liu et al. (2019)

q(τ1)=4(BzΨ(τ1)|BzΨ(τ1)|Ψ(τ1)|BzΨ(τ1)|2).\displaystyle\mathcal{F}_{q}({\tau_{1}})=4\left(\langle\partial_{B_{z}}\Psi(\tau_{1})|\partial_{B_{z}}\Psi(\tau_{1})\rangle-|\langle\Psi(\tau_{1})|\partial_{B_{z}}\Psi(\tau_{1})\rangle|^{2}\right). (S41)

Taking the partial derivative of Eq. (S40) with respect to BzB_{z}, we obtain

Bz|Ψ(τ1)=|Ψ(τ1)Bz=i4πcact12mωm3e3re|α|2nαnn!(λ~snf~s)ei2π(λ~snf~s)2|n|β,\displaystyle\partial_{B_{z}}|\Psi(\tau_{1})\rangle=\frac{\partial|\Psi(\tau_{1})\rangle}{\partial B_{z}}=-i4\pi c_{\rm act}\sqrt{\frac{1}{2m\hbar\omega_{m}^{3}}}e^{3r}e^{-|\alpha|^{2}}\sum_{n}\frac{\alpha^{n}}{\sqrt{n!}}(\tilde{\lambda}_{s}n-\tilde{f}_{s})e^{i2\pi(\tilde{\lambda}_{s}n-\tilde{f}_{s})^{2}}|n\rangle|\beta\rangle, (S42)

and

BzΨ(τ1)|BzΨ(τ1)\displaystyle\langle\partial_{B_{z}}\Psi(\tau_{1})|\partial_{B_{z}}\Psi(\tau_{1})\rangle =8π2cact2e6rmωm3e|α|2(n(λ~snf~s)2α2nn!),\displaystyle=\frac{8\pi^{2}c_{\rm act}^{2}e^{6r}}{m\hbar\omega_{m}^{3}}e^{-|\alpha|^{2}}\left(\sum_{n}\frac{(\tilde{\lambda}_{s}n-\tilde{f}_{s})^{2}\alpha^{2n}}{n!}\right),
=8π2cact2e6rmωm3[λ~s2α4+(λ~s22λ~sf~s)α2+f~s2].\displaystyle=\frac{8\pi^{2}c_{\rm act}^{2}e^{6r}}{m\hbar\omega_{m}^{3}}\left[\tilde{\lambda}_{s}^{2}\alpha^{4}+(\tilde{\lambda}_{s}^{2}-2\tilde{\lambda}_{s}\tilde{f}_{s})\alpha^{2}+\tilde{f}_{s}^{2}\right]. (S43)

In derivating Eq. (S5), we have used the following Taylor series expansion

nn2α2nn!=α2(1+α2)eα2,\displaystyle\sum_{n}{\frac{n^{2}\alpha^{2n}}{n!}}=\alpha^{2}(1+\alpha^{2})e^{\alpha^{2}}, (S44)
nnα2nn!=α2eα2,\displaystyle\sum_{n}\frac{n\alpha^{2n}}{n!}=\alpha^{2}e^{\alpha^{2}}, (S45)
nα2nn!=eα2.\displaystyle\sum_{n}\frac{\alpha^{2n}}{n!}=e^{\alpha^{2}}. (S46)

Following the same method introduced above, we can obtain

|Ψ(τ1)|BzΨ(τ1)|2=8π2cact2e6rmωm3(λ~s2α42λ~sf~sα2+f~s2).\displaystyle|\langle\Psi(\tau_{1})|\partial_{B_{z}}\Psi(\tau_{1})|^{2}=\frac{8\pi^{2}c_{\rm act}^{2}e^{6r}}{m\hbar\omega_{m}^{3}}(\tilde{\lambda}_{s}^{2}\alpha^{4}-2\tilde{\lambda}_{s}\tilde{f}_{s}\alpha^{2}+\tilde{f}_{s}^{2}). (S47)

Substituting Eqs. (S5) and (S47) to Eq. (S41) and using cact=mωm2Lαmag/Ec_{\rm act}=m\omega_{m}^{2}L{\alpha_{\rm mag}/}{E}, we obtain

q(τ1)=32π2mλ12L2αmag2ωmE2𝒩1e12r,\displaystyle\mathcal{F}_{q}(\tau_{1})=\frac{32\pi^{2}m\lambda_{1}^{2}L^{2}\alpha_{\rm mag}^{2}}{\hbar\omega_{m}E^{2}}\mathcal{N}_{1}e^{12r}, (S48)

where 𝒩1=α2\mathcal{N}_{1}=\alpha^{2}. To clearly see the dependence of QFI on the mean photon number, we do the following expansion

e12r\displaystyle e^{12r} =1[1(4λ2/ωm)𝒩2]3,\displaystyle=\frac{1}{\left[1-{(4\lambda_{2}/\omega_{m})\mathcal{N}_{2}}\right]^{3}}, (S49)

where 𝒩2=|ξ|2\mathcal{N}_{2}=|\xi|^{2}. Then q(τ1)\mathcal{F}_{q}(\tau_{1}) can also be expressed as

q(τ1)=32π2mλ12L2αmag2ωmE2𝒩1[1(4λ2/ωm)𝒩2]3.\displaystyle\mathcal{F}_{q}(\tau_{1})=\frac{32\pi^{2}m\lambda_{1}^{2}L^{2}\alpha_{\rm mag}^{2}}{\hbar\omega_{m}E^{2}}\frac{\mathcal{N}_{1}}{\left[1-{(4\lambda_{2}/\omega_{m})\mathcal{N}_{2}}\right]^{3}}. (S50)

According to the Cramér-Rao inequality, the QFI gives the ultimate lower limit of parameter estimation. Obtaining this lower bound requires the adoption of optimal measurements for the system, but it does not reveal which specific measurement is required to achieve it. The ultimate bound of sensitivity is given by

Δ𝔹z(τ1)\displaystyle\Delta\mathbb{B}_{z}(\tau_{1}) =Ee5r4πλ1Lαmagπm𝒩1(T/Hz),\displaystyle=\frac{Ee^{-5r}}{4\pi\lambda_{1}L\alpha_{\rm mag}}\sqrt{\frac{\pi\hbar}{m\mathcal{N}_{1}}}({\rm T/\sqrt{Hz}}),
=E4πλ1Lαmagπ[1(4λ2/ωm)𝒩2]52m𝒩1(T/Hz).\displaystyle=\frac{E}{4\pi\lambda_{1}L\alpha_{\rm mag}}\sqrt{\frac{\pi\hbar[1-(4\lambda_{2}/\omega_{m})\mathcal{N}_{2}]^{\frac{5}{2}}}{m\mathcal{N}_{1}}}({\rm T/\sqrt{Hz}}). (S51)

Fig. S4 clearly shows the different scaling of Δ𝔹z(τ1)\Delta\mathbb{B}_{z}(\tau_{1}) with respect to the resources 𝒩1\mathcal{N}_{1} and 𝒩2\mathcal{N}_{2}. Evidently, resources 𝒩1\mathcal{N}_{1} and 𝒩2\mathcal{N}_{2} jointly determine the ultimate sensitivity.

Refer to caption
Figure S4: Ultimate bound of sensitivity Δ𝔹z(τ1)\Delta\mathbb{B}_{z}(\tau_{1}) in units of T/Hz{\rm\sqrt{Hz}} versus the resources 𝒩1\mathcal{N}_{1} and 𝒩2\mathcal{N}_{2}. We considered 𝒩2=2.45×106\mathcal{N}_{2}=2.45\times 10^{6} in (a), 𝒩1=106\mathcal{N}_{1}=10^{6} and aa is a constant in (b). Other parameters are same as the main text.

S6 CFI and specific sensitivity

S6.1 CFI without dissipation

In this section, let us offer the detailed derivation of the CFI, i.e., Eq. (6) used in the main text. Normally, the CFI, corresponding to a specific measurement of BzB_{z}, is given by

c(t)=1P(X|Bz)(P(X|Bz)Bz)2dX.\displaystyle\mathcal{F}_{c}(t)=\int\frac{1}{P(X|B_{z})}\left(\frac{\partial P(X|B_{z})}{\partial B_{z}}\right)^{2}{\rm d}X. (S52)

where P(X|Bz)P(X|B_{z}) represents the conditional probability of measuring XX relied on parameter BzB_{z}. With the positive-operator valued measure elements {ΠX}\{\Pi_{X}\}, we obtain P(X|Bz)=Tr[ΠXρ(t)]P(X|B_{z})={\rm Tr}[\Pi_{X}\rho(t)]. Here we consider a general homodyne measurement on the traced-out cavity state ρc(t)\rho_{c}(t) with the observable operator Xθ=(a1eiθ+a1eiθ)/2X_{\theta}=(a_{1}e^{-i\theta}+a_{1}^{{\dagger}}e^{i\theta})/\sqrt{2}, where θ\theta is the phase of local oscillator. The cases of θ=0\theta=0 and θ=π/2\theta=\pi/2 correspond to the position and momentum measurements, respectively. Then the conditional probability becomes

P(Xθ|Bz)=Tr[|XθXθ|ρc(t)],\displaystyle P(X_{\theta}|B_{z})={\rm Tr}[|X_{\theta}\rangle\langle X_{\theta}|\rho_{c}(t)], (S53)

where |Xθ|X_{\theta}\rangle is the eigenstate of XθX_{\theta}. Using the inner product n|Xθ=π1/4[2n(n!)]1/2exp(Xθ2/2)Hn(Xθ)exp(inθ)\langle n|X_{\theta}\rangle=\pi^{-1/4}[2^{n}(n!)]^{-1/2}\exp(-X_{\theta}^{2}/2)H_{n}(X_{\theta})\exp(in\theta) with Hn(Xθ)=exp(Xθ2/2)(Xθd/dXθ)nexp(Xθ2/2)H_{n}(X_{\theta})=\exp(X_{\theta}^{2}/2)\left(X_{\theta}-{\rm d}/{\rm d}X_{\theta}\right)^{n}\exp(-X_{\theta}^{2}/2) being the Hermite polynomials of order nn, we can rewrite Eq. (S53) as

P(Xθ|Bz)\displaystyle P(X_{\theta}|B_{z}) =e|α|2n,n[αn(α)nn!n!ei[λ~s2(n2n2)2λ~sf~s(nn)](ωstsinωst)eXθ2πHn(Xθ)Hn(Xθ)eiθ(nn)2(n+n)/2n!n!\displaystyle=e^{-|\alpha|^{2}}\sum_{n,n^{\prime}}\left[\frac{\alpha^{n}(\alpha^{*})^{n^{\prime}}}{\sqrt{n!n^{\prime}!}}e^{i\left[\tilde{\lambda}_{s}^{2}(n^{2}-n^{\prime 2})-2\tilde{\lambda}_{s}\tilde{f}_{s}(n-n^{\prime})\right](\omega_{s}t-\sin\omega_{s}t)}\frac{e^{-X_{\theta}^{2}}}{\sqrt{\pi}}\frac{H_{n}(X_{\theta})H_{n^{\prime}}(X_{\theta})e^{-i\theta(n-n^{\prime})}}{2^{(n+n^{\prime})/2}\sqrt{n!n^{\prime}!}}\right.
×eiλ~s(nn)[erβResinωsterβIm(cosωst1)]e(|φn|2+|φn|2)/2+φnφn].\displaystyle\left.\times e^{i\tilde{\lambda}_{s}(n-n^{\prime})[e^{-r}\beta_{\rm Re}\sin\omega_{s}t-e^{r}\beta_{\rm Im}(\cos\omega_{s}t-1)]}e^{-(|\varphi_{n}|^{2}+|\varphi_{n^{\prime}}|^{2})/2+\varphi_{n^{\prime}}^{*}\varphi_{n}}\right]. (S54)

Substituting Eq. (S6.1) into Eq. (S52), we obtain the analytical expression of CFI at time τ1=2π/ωs\tau_{1}=2\pi/\omega_{s} is

c(τ1)=8π2mλ12L2αmag2ωmE2e12r[2sinθαRe2cosθαIm]2,\displaystyle\mathcal{F}_{c}(\tau_{1})=\frac{8\pi^{2}m\lambda_{1}^{2}L^{2}\alpha_{\rm mag}^{2}}{\hbar\omega_{m}E^{2}}e^{12r}[2\sin\theta\alpha_{\rm Re}-2\cos\theta\alpha_{\rm Im}]^{2}, (S55)

and it can be reduced further as

c(τ1)={32π2mλ12L2αmag2|αIm|2ωmE2e12r,forθ=0,32π2mλ12L2αmag2|αRe|2ωmE2e12r,forθ=π/2.\mathcal{F}_{c}(\tau_{1})=\begin{dcases}\frac{32\pi^{2}m\lambda_{1}^{2}L^{2}\alpha_{\rm mag}^{2}|\alpha_{\rm Im}|^{2}}{\hbar\omega_{m}E^{2}}e^{12r},&{\rm for}\,\,\,\theta=0,\\ \frac{32\pi^{2}m\lambda_{1}^{2}L^{2}\alpha_{\rm mag}^{2}|\alpha_{\rm Re}|^{2}}{\hbar\omega_{m}E^{2}}e^{12r},&{\rm for}\,\,\,\theta=\pi/2.\end{dcases} (S56)
Refer to caption
Figure S5: Classical Fisher information c(τ1)\mathcal{F}_{c}(\tau_{1}) versus squeezing parameter rr for different (a) cavity decay rates κ\kappa, and (b) mechanical decay rates γ\gamma. In (c,d), we show the corresponding sensitivity limit ΔBz(τ1)\Delta B_{z}(\tau_{1}) in units of T/Hz{\rm\sqrt{Hz}}. We have chosen the parameters as (a,c) γ/ωm=0.001\gamma/\omega_{m}=0.001, and (b,d) κ/ωm=0.01\kappa/\omega_{m}=0.01. Other parameters are ωm=2π×134kHz,f=0.01ωm\omega_{m}=2\pi\times 134\,{\rm kHz},f=0.01\omega_{m} and n¯th=10\bar{n}_{\rm th}=10.

S6.2 CFI and specific sensitivity limit with system dissipation

In this subsection, let us present the detailed derivation for the CFI in the case of including system dissipation. In a practical experimental setup, the dissipation caused by the system-bath coupling should be considered. Then the full dynamics of the system satisfy the following master equation

ρ(t)t=\displaystyle\frac{\partial\rho(t)}{\partial t}= i[H,ρ(t)]+κ2[2a1ρa1ρa1a1a1a1ρ]+γ2(n¯th+1)[2bρbρbbbbρ]\displaystyle-\frac{i}{\hbar}[H,\rho(t)]+\frac{\kappa}{2}\left[2a_{1}\rho a_{1}^{{\dagger}}-\rho a_{1}^{{\dagger}}a_{1}-a_{1}^{{\dagger}}a_{1}\rho\right]+\frac{\gamma}{2}(\bar{n}_{\rm th}+1)\left[2b\rho b^{{\dagger}}-\rho b^{{\dagger}}b-b^{{\dagger}}b\rho\right]
+γ2n¯th[2bρbρbbbbρ],\displaystyle+\frac{\gamma}{2}\bar{n}_{\rm th}\left[2b^{{\dagger}}\rho b-\rho bb^{{\dagger}}-bb^{{\dagger}}\rho\right], (S57)

where κ(γ)\kappa(\gamma) is the cavity (mechanical) decay rate, and n¯th\bar{n}_{\rm th} is the thermal phonon number of mechanical mode. Hamiltonian HH is shown in Eq. (S13). Applying a squeezing transformation to Eq. (S6.2) with S(r)=exp[r(b2b2)/2]S(r)=\exp[r(b^{2}-b^{{\dagger}2})/2] and r=(1/4)ln(14λ2𝒩2/ωm)r=-(1/4)\ln(1-4\lambda_{2}\mathcal{N}_{2}/\omega_{m}), we obtain

ρs(t)t\displaystyle\frac{\partial{\rho}_{s}(t)}{\partial t} =\displaystyle= i[Hs,ρs]+κ2[2a1ρsa1ρsa1a1a1a1ρs]\displaystyle-\frac{i}{\hbar}\left[H_{s},\rho_{s}\right]+\frac{{\kappa}}{2}\left[2a_{1}\rho_{s}a_{1}^{\dagger}-\rho_{s}a_{1}^{\dagger}a_{1}-a_{1}^{\dagger}a_{1}\rho_{s}\right] (S58)
+γ(n¯th+1)(𝒟[b]ρscosh2r+𝒟[b]ρssinh2r+𝒢[b]ρssinhrcoshr+𝒢[b]ρssinhrcoshr)\displaystyle+{\gamma}(\bar{n}_{\rm th}+1)\left({\mathcal{D}}[b]\rho_{s}\cosh^{2}r+{\mathcal{D}}[b^{{\dagger}}]\rho_{s}\sinh^{2}r+{\mathcal{G}}[b]\rho_{s}\sinh r\cosh r+{\mathcal{G}}[b^{{\dagger}}]\rho_{s}\sinh r\cosh r\right)
+γn¯th(𝒟[b]ρscosh2r+𝒟[b]ρssinh2r+𝒢[b]ρssinhrcoshr+𝒢[b]ρssinhrcoshr),\displaystyle+{\gamma}\bar{n}_{\rm th}\left(\mathcal{D}[b^{{\dagger}}]\rho_{s}\cosh^{2}r+{\mathcal{D}}[b]\rho_{s}\sinh^{2}r+{\mathcal{G}}[b]\rho_{s}\sinh r\cosh r+{\mathcal{G}}[b^{{\dagger}}]\rho_{s}\sinh r\cosh r\right),

where ρs=S(r)ρS(r)\rho_{s}=S(r)\rho S^{{\dagger}}(r), Hs=S(r)HS(r)H_{s}=S(r)HS^{{\dagger}}(r) and the Lindblad superoperators read

𝒟[o]ρ=oρo(ooρ+ρoo)/2,𝒢[o]ρ=oρo(ooρ+ρoo)/2.\displaystyle\mathcal{D}[o]\rho=o\rho o^{{\dagger}}-(o^{{\dagger}}o\rho+\rho o^{{\dagger}}o)/2,\,\,\,\,\,\,\,\,\,\mathcal{G}[o]\rho=o\rho o-(oo\rho+\rho oo)/2. (S59)

We can numerically calculate ρ\rho by ρ=SρsS(r)\rho=S^{{\dagger}}\rho_{s}S(r). Once we numerically obtain the density matrix, we can further calculate the CFI.

Next we discuss the effect of system dissipations on the CFI and specific sensitivity limit ΔBz(τ1)\Delta B_{z}(\tau_{1}). By numerically solving the master equation  (S58) Johansson et al. (2012), we can obtain the CFI at time c(τ1)\mathcal{F}_{c}(\tau_{1}) and the sensitivity ΔBz(τ1)\Delta B_{z}(\tau_{1}) in the presence of system dissipations. In addition to the main results shown in the main text, here we supplement some numerical results indicating the influences of squeezing parameter rr and system decay rates on the CFI and sensitivity in Figs. S5. It is shown that, for a certain squeezing parameter rr, the increased cavity decay κ\kappa leads to a decline of CFI c(τ1)\mathcal{F}_{c}(\tau_{1}). In spite of this, its corresponding sensitivities still maintain at the order of 1017T/Hz10^{-17}\,{\rm T/\sqrt{Hz}} [see Fig. S5(c)].

In Fig. S5(b) we plot c(τ1)\mathcal{F}_{c}(\tau_{1}) varying with squeezing parameter rr, for different mechanical decays. Specifically, for a small squeezing parameter, mechanical dissipation exerts little influence on the CFI. The increasing squeezing parameter amplifies the noise coming from the mechanical bath which causes the influence of mechanical decay γ\gamma becomes larger. This is definitely different from the effect of optical decay κ\kappa exerted on the CFI [see Fig. S5(a)]. Even though the mechanical decay reduces achievable CFI, numerically simulation has clearly shown that the measurement maintains a high accuracy in the presence of dissipation [see Figs. S5(d)]. In other words, the measurement sensitivity of our proposal is insensitive to the system dissipation due to the mechanical oscillator and optical cavity decoupled at detection time τ1\tau_{1}.

S7 Frequency response of resonant dual-coupling magnetometer

In the main text, we proposed a scheme to measure the dc magnetic fields (or static magnetic fields) in a dual-coupling optomechanical system. We calculated the quantum and classical Fisher information FqF_{q} and FcF_{c}, and then obtained the fundamental bound to the sensitivity ΔBz\Delta B_{z}. This sensitivity gives the theoretical lower limit of the measurement precision based on our proposal, which is several orders of magnitude lower than that achieved in recent experiments Forstner et al. (2012); Li et al. (2021); Colombano et al. (2020). Note that, besides detecting a static magnetic field discussed in the main text, in principle, our proposal can also work as a resonant sensor to be applied to detect the alternating magnetic fields. In the following, we will estimate the frequency response characteristics including the bandwidth issue, when the proposed dual-optomechanical system is used to detect the alternating magnetic fields.

Refer to caption
Figure S6: (a) Displacement thermal noise power spectra Sxx(ω)S_{xx}(\omega) as a function of frequency ω\omega, for different values of powers, i.e., P=20pWP=20\,{\rm pW} (green), 200pW200\,{\rm pW} (blue), 2nW2\,{\rm nW} (magenta), and 20nW20\,{\rm nW} (black). (b) Force sensitivity SFF(ω)\sqrt{S_{FF}(\omega)} vs frequency. Plotting these figures, we referred to the parameters used in optomechanical experiments Thompson et al. (2008); Li et al. (2021), i.e., ωm=2π×1.34×105\omega_{m}=2\pi\times 1.34\times 10^{5}Hz, γ=2π×0.12\gamma=2\pi\times 0.12 Hz, m=4×1011m=4\times 10^{-11} kg, ωc=ωL=2π×1014\omega_{c}=\omega_{L}=2\pi\times 10^{14}\,Hz, κ=2π×108\kappa=2\pi\times 10^{8} Hz, κex=κ\kappa_{\rm ex}=\kappa, the optomechanical coupling G=500G=500  MHz/nm, η=0.8\eta=0.8, T=300KT=300\,{\rm K}, and r=0.6r=0.6.

In cavity optomechanical magnetometers, the magnetostrictive material expands, exerting a force upon the mechanical oscillator. The mechanical motion modulates the optical cavity field via the radiation-pressure coupling. Meanwhile, the phase shift of the mechanical motions encoded with magnetic signal is transferred to the optical field, which enables us to optically readout mechanical motions. The combination of mechanical and optical resonances provides enhanced mechanical response to applied forces and optical readout with high precision. Experimentally, the sensitivity of such a magnetometer is defined as the minimum detectable signal, and can be quantified by the noise equivalent signal. Thus, the performance of the optomechanical magnetometer relies on how efficiently the magnetic signal drives the mechanical modes above the experiment noise level. Generally speaking, the noise sources in the optomechanical systems mainly consist of the thermal noise from the thermal environment with nonzero temperature and the photon shot noise from the probe laser.

For simplicity, we consider a single mode of mechanical resonance, whose response to an external force in the frequency domain is quantified by the mechanical susceptibility of the resonator χ(ω)=1/(m(ωs2ω2iωγ))\chi(\omega)=1/(m(\omega_{s}^{2}-\omega^{2}-i\omega\gamma)) with effective mass mm, frequency ωs\omega_{s} and mechanical decay rate γ\gamma. The magnetometer sensor is essentially a force sensor, whose sensitivity of displacement at the frequency ω\omega is determined by Sxx(ω)\sqrt{S_{xx}(\omega)}, where the noise spectrum is defined as Sxx(ω)=+x(t)x(0)eiωt𝑑tS_{xx}(\omega)=\int_{-\infty}^{+\infty}\langle x(t)x(0)\rangle e^{i\omega t}dt. According to the fluctuation dissipation theorem, a mechanical resonator with frequency ωs\omega_{s} experiences a thermal noise force Fth=2mγkBTF_{\rm th}=\sqrt{2m\gamma k_{B}T}Aspelmeyer et al. (2014) at a temperature TT, and the corresponding thermal force power spectrum reads Li et al. (2021)

Sxxth=2γkBTm[(ωs2ω2)2ω2γ2].\displaystyle S_{xx}^{\rm th}=\frac{2\gamma k_{B}T}{m[(\omega_{s}^{2}-\omega^{2})^{2}-\omega^{2}\gamma^{2}]}. (S60)

The displacement noise power spectrum from the laser shot noise,

Sxxshot(ω)=κ16ηNG2(1+ω2κ2),\displaystyle S_{xx}^{\rm shot}(\omega)=\frac{\kappa}{16\eta NG^{2}}(1+\frac{\omega^{2}}{\kappa^{2}}), (S61)

where N=4Pκex/(ωLκ2)N=4P\kappa_{\rm ex}/(\hbar\omega_{L}\kappa^{2}) is the intracavity photon number Aspelmeyer et al. (2014), ωL\omega_{L} and PP are the frequency and power of the probe field, respectively. Here κ\kappa is the total cavity loss rate, κex\kappa_{\rm ex} refers to the loss rate associated with the input coupling, η\eta is the optical detection efficiency, and G=dωc/dtG={\rm d}\omega_{c}/{\rm d}t is the optomechanical coupling strength.

In Fig. S6, we numerically simulate the displacement noise power spectrum and the corresponding force sensitivity at different probe powers. It can be seen from Fig. S6(a) that these noise power spectra show a peak at resonance, i.e., ω=ωs\omega=\omega_{s}. With the increase of probe power, the shot noise is suppressed effectively [see Eq. (S61)]. Meanwhile, the peak of the force sensitivity also occurs at resonance [ses Fig. S6(b)], which is determined by the mechanical decay rate γ\gamma. The red dotted curve in Fig. S6(b) denotes the frequency range in which the sensitivity is better than twice of the peak sensitivity, which is defined as the bandwidth Li et al. (2021). It can be seen that, the larger the probe power is, the broader bandwidth is. For example, when the probe power is 20nW20\,{\rm nW}, the thermal-noise-limited frequency range covers 03800\sim 380 kHz with the central resonant frequency 250 kHz. Therefore, similar as a general optomechanical system Li et al. (2021), here the dual-coupling magnetometer also has a broad bandwidth, when it is used as a resonant sensor.

References

  • Thompson et al. (2008) J. D. Thompson, B. M. Zwickl, A. M. Jayich, F. Marquardt, S. M. Girvin, and J. G. E. Harris, Nature 452, 72 (2008).
  • Verhoeven et al. (1990) J. D. Verhoeven, E. D. Gibson, O. D. Mcmasters, and J. E. Ostenson, Metallurgical Transactions A 21, 2249 (1990).
  • (3) A. N. Cleland, Foundations of Nanomechanics: From Solid-State Theory to Device Applications (Springer, New York, 2003).
  • Briant et al. (2003) T. Briant, P.-F. Cohadon, A. Heidmann, and M. Pinard, Phys. Rev. A 68, 033823 (2003).
  • Forstner et al. (2012) S. Forstner, S. Prams, J. Knittel, E. D. van Ooijen, J. D. Swaim, G. I. Harris, A. Szorkovszky, W. P. Bowen, and H. Rubinsztein-Dunlop, Phys. Rev. Lett. 108, 120801 (2012).
  • Paris (2009) M. G. A. Paris, Int. J. Quantum Inf. 07, 125 (2009).
  • Tóth and Apellaniz (2014) G. Tóth and I. Apellaniz, J. Phys. A: Math. Theor. 47, 424006 (2014).
  • Liu et al. (2019) J. Liu, H. Yuan, X.-M. Lu, and X. Wang, J. Phys. A: Math. Theor. 53, 023001 (2019).
  • Johansson et al. (2012) J. Johansson, P. Nation, and F. Nori, Comput. Phys. Commun. 183, 1760 (2012).
  • Li et al. (2021) B.-B. Li, L. Ou, Y. Lei, and Y.-C. Liu, Nanophotonics 10, 2799 (2021).
  • Colombano et al. (2020) M. F. Colombano, G. Arregui, F. Bonell, N. E. Capuj, E. Chavez-Angel, A. Pitanti, S. O. Valenzuela, C. M. Sotomayor-Torres, D. Navarro-Urrios, and M. V. Costache, Phys. Rev. Lett. 125, 147201 (2020).
  • Aspelmeyer et al. (2014) M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Rev. Mod. Phys. 86, 1391 (2014).