Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for...

9
Cram´ er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni 1 1 Quantum Technology Lab, Dipartimento di Fisica, Universit` a degli Studi di Milano, 20133 Milano, Italy We describe a compact and reliable method to calculate the Fisher information for the estimation of a dy- namical parameter in a continuously measured linear Gaussian quantum system. Unlike previous methods in the literature, which involve the numerical integration of a stochastic master equation for the corresponding density operator in a Hilbert space of infinite dimension, the formulas here derived depends only on the evo- lution of first and second moments of the quantum states, and thus can be easily evaluated without the need of any approximation. We also present some basic but physically meaningful examples where this result is exploited, calculating analytical and numerical bounds on the estimation of the squeezing parameter for a quan- tum parametric amplifier, and of a constant force acting on a mechanical oscillator in a standard optomechanical scenario. I. INTRODUCTION Parameter estimation via quantum probes and quantum measurements will lead to a new generation of detectors char- acterised by sensitivities not achievable through only classical means [1, 2]. The promised quantum enhancement is typically however lost as soon as some decoherence affects the system [3, 4]. On the other hand, the information leaking into the en- vironment can be in principle used for parameter estimation as well, in particular via time-continuous monitoring of the environment itself [5, 6]. While several strategies based on time-continuous measurements and feedback have been pro- posed for quantum state engineering, in particular with the main goal of generating steady-state squeezing and entangle- ment [5, 7–15] or to study and exploit trajectories of super- conducting qubits [16, 17], less attention has been devoted to parameter estimation. Notable exceptions are the estimation of a magnetic field via a continuously monitored atomic en- semble [18], the tracking of a varying phase [19–21], the esti- mation of Hamiltonian and environmental parameters [22–29] and optimal state estimation for a cavity optomechanical sys- tem [30]. The ultimate precision achievable by quantum metrological strategies is determined by the classical and quantum Cram` er- Rao bounds [31–33], which are expressed in terms of, re- spectively, classical and quantum Fisher information (FI). Re- cently, methods have been proposed to calculate these quanti- ties in the stationary regime for certain relevant setups [22, 23, 28], or in the dynamical regime in the case of time-continuous homodyne and photon-counting measurements [24, 25]. In particular, in order to evaluate the FI corresponding to a continuous homodyne detection, the method presented in Ref.[24] relies on the integration of stochastic master equa- tions for operators characterizing the quantum state and the measurement performed. While this can be straightforwardly accomplished in the case of finite-dimensional quantum sys- tems, such as two-level atoms and superconducting qubits, the method becomes computationally very expensive and less re- liable in the case of large or even infinite-dimensional sys- tems, such as the electromagnetic field, atomic ensembles and mechanical oscillators. In fact, in these cases, one has to trun- cate the corresponding Fock space, posing a constraint on the maximum energy of the system. The goal of this article is to provide an efficient and reli- able method to calculate the FI for parameter estimation via time-continuous measurements in linear Gaussian quantum systems. Gaussian systems represents a subclass of infinite- dimensional bosonic systems, whose properties and dynamics can be univocally described in terms of first and second mo- ments only [34–36]. In order to observe such a dynamics, one has to consider Hamiltonians at most quadratic in the canon- ical operators, a linear coupling with the environment and time-continuous monitoring via Gaussian measurements [36– 38]. This restriction allows one to greatly simplify the analy- sis of infinite-dimensional quantum systems, and at the same time describe several state-of-the-art experimental setups in the area of quantum optics, opto-mechanics, trapped ions and atomic ensembles. In particular, we remind the reader how optimal state estimation via time-continuous monitoring has been very recently accomplished for a Gaussian cavity op- tomechanical systems [30], showing the timeliness and rele- vance of this approach. In detail, the manuscript is structured as follows. In Sec. II we provide a basic introduction on Gaussian systems and their diffusive and conditional dynamics, while in Sec. III we revise the Cram´ er-Rao bound, with a focus on a posteriori Gaussian distributions. In Sec. IV we present the main result of the manuscript, that is a method for the calculation of the FI for Gaussian systems depending only on the evolution of first and second moments and that does not need any approximation or limit on the energy of the quantum states in exam. To show the potential of our results, in Sec. V we provide two examples, calculating numerical and analytical bounds on the estimation precision for the squeezing parameter in a quantum paramet- ric amplifier, and for a constant force acting on a mechanical oscillator in a standard opto-mechanical setup. II. DIFFUSIVE AND CONDITIONAL DYNAMICS IN LINEAR GAUSSIAN SYSTEMS We consider a set of n bosonic modes described by a vector of quadrature operators ˆ r T = (ˆ x 1 , ˆ p 1 ,..., ˆ x n , ˆ p n ), satisfying the canonical commutation relation r, ˆ r T ]= iΩ (with Ω jk = δ k,j+1 - δ k,j-1 ) [39]. We define a quantum state % Gaussian, if and only if can be written as a ground or thermal state of a arXiv:1608.08429v3 [quant-ph] 7 Jun 2017

Transcript of Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for...

Page 1: Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni1 1Quantum Technology

Cramer-Rao bound for time-continuous measurements in linear Gaussian quantum systems

Marco G. Genoni11Quantum Technology Lab, Dipartimento di Fisica, Universita degli Studi di Milano, 20133 Milano, Italy

We describe a compact and reliable method to calculate the Fisher information for the estimation of a dy-namical parameter in a continuously measured linear Gaussian quantum system. Unlike previous methods inthe literature, which involve the numerical integration of a stochastic master equation for the correspondingdensity operator in a Hilbert space of infinite dimension, the formulas here derived depends only on the evo-lution of first and second moments of the quantum states, and thus can be easily evaluated without the needof any approximation. We also present some basic but physically meaningful examples where this result isexploited, calculating analytical and numerical bounds on the estimation of the squeezing parameter for a quan-tum parametric amplifier, and of a constant force acting on a mechanical oscillator in a standard optomechanicalscenario.

I. INTRODUCTION

Parameter estimation via quantum probes and quantummeasurements will lead to a new generation of detectors char-acterised by sensitivities not achievable through only classicalmeans [1, 2]. The promised quantum enhancement is typicallyhowever lost as soon as some decoherence affects the system[3, 4]. On the other hand, the information leaking into the en-vironment can be in principle used for parameter estimationas well, in particular via time-continuous monitoring of theenvironment itself [5, 6]. While several strategies based ontime-continuous measurements and feedback have been pro-posed for quantum state engineering, in particular with themain goal of generating steady-state squeezing and entangle-ment [5, 7–15] or to study and exploit trajectories of super-conducting qubits [16, 17], less attention has been devoted toparameter estimation. Notable exceptions are the estimationof a magnetic field via a continuously monitored atomic en-semble [18], the tracking of a varying phase [19–21], the esti-mation of Hamiltonian and environmental parameters [22–29]and optimal state estimation for a cavity optomechanical sys-tem [30].The ultimate precision achievable by quantum metrologicalstrategies is determined by the classical and quantum Cramer-Rao bounds [31–33], which are expressed in terms of, re-spectively, classical and quantum Fisher information (FI). Re-cently, methods have been proposed to calculate these quanti-ties in the stationary regime for certain relevant setups [22, 23,28], or in the dynamical regime in the case of time-continuoushomodyne and photon-counting measurements [24, 25]. Inparticular, in order to evaluate the FI corresponding to acontinuous homodyne detection, the method presented inRef.[24] relies on the integration of stochastic master equa-tions for operators characterizing the quantum state and themeasurement performed. While this can be straightforwardlyaccomplished in the case of finite-dimensional quantum sys-tems, such as two-level atoms and superconducting qubits, themethod becomes computationally very expensive and less re-liable in the case of large or even infinite-dimensional sys-tems, such as the electromagnetic field, atomic ensembles andmechanical oscillators. In fact, in these cases, one has to trun-cate the corresponding Fock space, posing a constraint on themaximum energy of the system.

The goal of this article is to provide an efficient and reli-able method to calculate the FI for parameter estimation viatime-continuous measurements in linear Gaussian quantumsystems. Gaussian systems represents a subclass of infinite-dimensional bosonic systems, whose properties and dynamicscan be univocally described in terms of first and second mo-ments only [34–36]. In order to observe such a dynamics, onehas to consider Hamiltonians at most quadratic in the canon-ical operators, a linear coupling with the environment andtime-continuous monitoring via Gaussian measurements [36–38]. This restriction allows one to greatly simplify the analy-sis of infinite-dimensional quantum systems, and at the sametime describe several state-of-the-art experimental setups inthe area of quantum optics, opto-mechanics, trapped ions andatomic ensembles. In particular, we remind the reader howoptimal state estimation via time-continuous monitoring hasbeen very recently accomplished for a Gaussian cavity op-tomechanical systems [30], showing the timeliness and rele-vance of this approach.In detail, the manuscript is structured as follows. In Sec. IIwe provide a basic introduction on Gaussian systems and theirdiffusive and conditional dynamics, while in Sec. III we revisethe Cramer-Rao bound, with a focus on a posteriori Gaussiandistributions. In Sec. IV we present the main result of themanuscript, that is a method for the calculation of the FI forGaussian systems depending only on the evolution of first andsecond moments and that does not need any approximation orlimit on the energy of the quantum states in exam. To show thepotential of our results, in Sec. V we provide two examples,calculating numerical and analytical bounds on the estimationprecision for the squeezing parameter in a quantum paramet-ric amplifier, and for a constant force acting on a mechanicaloscillator in a standard opto-mechanical setup.

II. DIFFUSIVE AND CONDITIONAL DYNAMICS INLINEAR GAUSSIAN SYSTEMS

We consider a set of n bosonic modes described by a vectorof quadrature operators rT = (x1, p1, . . . , xn, pn), satisfyingthe canonical commutation relation [r, rT] = iΩ (with Ωjk =δk,j+1 − δk,j−1) [39]. We define a quantum state % Gaussian,if and only if can be written as a ground or thermal state of a

arX

iv:1

608.

0842

9v3

[qu

ant-

ph]

7 J

un 2

017

Page 2: Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni1 1Quantum Technology

2

quadratic Hamiltonian, i.e.

% =exp−βHG

Z, β ∈ R , (1)

where HG = (1/2)rTHG r and HG ≥ 0 [34, 36]. Gaussianstates can be univocally described by the vector of first mo-ments R and the covariance matrix σ [39]:

R = Tr[%r] , σ = Tr[%r−R, (r−R)T] . (2)

We also recall that, in order to describe a proper Gaussianquantum state, the covariance matrix has to satisfy the physi-cality condition σ + iΩ ≥ 0 [40].

We now consider a dynamics generated by a Hamiltonianwith linear and quadratic terms of the form

Hs =1

2rTHsr− rTΩu , (3)

where Hs is a matrix of dimension 2n× 2n, while u is a 2n-dimensional vector. We also assume that the system is coupledto a large Markovian environment described by a train of in-coming modes rb(t), each of which interacts with the systemat a given time t. The correlations characterizing the environ-ment are specified through the white-noise condition

〈rb(t), rTb (t′)〉 = σbδ(t− t′) , σb + iΩ ≥ 0. (4)

The interaction in a interval of time dt between the system andthe environment is ruled by the Hamiltonian

HC dt = rTC rb(t) dt = rTC drb(t) = rTC r′b(t) dW.

Here we have introduced the so-called quantum Wiener incre-ment [41]

drb(t) = rb(t) dt = r′b(t) dW,

with dW being a real Wiener increment such that dW 2 =dt, and where r′b(t) is a vector of “proper” dimensionlessfield operators (that can be associated with detector clicksin the laboratory, and formally with POVM operators in theHilbert space) satisfying the canonical commutation relation[r′b(t), r

′b(t)

T] = iΩ (we refer to Ref. [36] and to Appendix Bfor more details on these definitions and on the derivation ofthe following formulas). By tracing out the degrees of free-dom of the environment, the dynamics of the Gaussian state isthen described by the following equations

dRt

dt= ARt + u (5)

dσtdt

= Aσt + σtAT +D (6)

where we have introduced the drift matrix A = ΩHs +(ΩCΩCT)/2 and the diffusion matrix D = ΩCσbC

TΩT.One should notice that, as expected, the linear term in theHamiltonian (3) is responsible for only a displacement of thefirst moments vector, while the evolution of the covariancematrix is not affected.

We now assume that the environment is continuously moni-tored at each time via a Gaussian measurement described bya matrix σm (s.t. σm + iΩ ≥ 0 ), and whose measurementoutcome corresponds to a vector xm. In this scenario, the con-ditional state is still Gaussian, and the dynamics is describedby a stochastic equation for the first moment vector and by adeterministic Riccati equation for the covariance matrix [36](see also Appendix B):

dRt = ARt dt+ u dt+

(σtB +N√

2

)dw ,

dσtdt

= Aσt + σtAT +D − (σtB +N)(σtB +N)T ,

(7)

where dw is a vector of independent Wiener increments (s.t.dwjdwk = δjkdt) and we have introduced the matrices B =

CΩ(σb + σm)−1/2 and N = ΩCσb(σb + σm)−1/2.The dynamics we have just presented in terms of first andsecond moments for Gaussian states, can be equivalently de-scribed by the following family of stochastic master equationsfor the (infinite-dimensional) density operator

d% = −i[H, %] dt+

L∑j=1

D[cj ]% dt+ dz†∆c%+ %∆c†dz ,

(8)

where c = C r, D[o]% = o%o†−o†o, %/2, ∆o = o−Tr[%o]and dz is a vector of complex Wiener increments [37, 38].For our purposes is important to recall that the outcomes xmof the measurement performed on the bath operators r′b(t) aredistributed according to a Gaussian multi-variate distributionwith mean value xm = ΩCTRt dW and covariance matrix(σb + σm)/2. Typically, the results of time-continuous mea-surements are formulated as a real current with uncorrelatednoise [38], i.e.

dy := (σb + σm)−1/2 xm dW (9)

= −BTRt dt+dw√

2. (10)

III. FISHER INFORMATION FROM A MULTIVARIATEGAUSSIAN DISTRIBUTION

Let us consider an a posteriori distribution p(x|θ), wherethe vector x corresponds to the outcomes of the measurementperformed and θ is a parameter we want to estimate. In par-ticular we assume p(x|θ) to be a multi-variate Gaussian dis-tribution with mean value xθ and covariance matrix Σ, whereonly the mean value depends on the parameter θ. The ultimatelimit on how accurate we can estimate θ is determined by theCramer-Rao bound [42]

Var(θ) ≥ 1

MF (θ), (11)

Page 3: Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni1 1Quantum Technology

3

where Var(θ) is the variance of an unbiased estimator θ, M isthe number of measurements, and

F (θ) = Ep

[(∂ ln p(x|θ)

∂θ

)2]

(12)

is the FI corresponding to the distribution p(x|θ). As it is clearfrom the inequality (11), the FI F (θ) quantifies how well wecan infer the value of the parameter θ from the measurementoutcomes. By explicitly writing the square of the derivativeof the likelihood function l(x|θ) = log p(x|θ) and exploitingthe property Ep[(x − xθ)j(x − xθ)k] = Σjk , one can eas-ily prove that the FI corresponding to a Gaussian a posterioridistribution reads

F (θ) = (∂θxθ)TΣ−1(∂θxθ) (13)

IV. FISHER INFORMATION FOR TIME-CONTINUOUSMEASUREMENTS IN LINEAR GAUSSIAN QUANTUM

SYSTEMS

Let us assume that we want to estimate the value of a pa-rameter θ that characterizes the dynamic of a Gaussian linearquantum system described by Eqs. (7). In particular we alsoassume that only the system Hamiltonian Hs depends on θ,and, thus only the drift matrix Aθ and/or the vector uθ de-pends on the parameter [43].We stated above that the probability distribution p(xm|θ) cor-responding to the measurement performed at time t + dton the environment is a Gaussian distribution centered inΩCTRt dW , and with covariance matrix (σb + σm)/2. Asonly the mean value depends on the parameter θ, via the firstmoment vector Rt, we can exploit Eq. (13) and write the cor-responding infinitesimal FI as

dF(traj)t (θ) = 2(∂θRt)

TCΩT(σb + σm)−1 ΩCT(∂θRt) dt .(14)

Notice that this FI corresponds to a specific trajectory, as it iscalculated via the vector (∂θRt) whose evolution is in princi-ple stochastic and described by the equations

d(∂θRt) = (∂θA)Rt dt+A(∂θRt) dt+ (∂θu) dt+(∂θσt)B√

2dw +

(σtB +N√

2

)(∂θdw) ,

d(∂θσt)

dt= (∂θA)σt + σt(∂θA)T +A(∂θσt) + (∂θσt)A

T − (∂θσt)B(σtB +N)T − (σtB +N)((∂θσt)B)T, (15)

where, from Eq. (10), we obtain that ∂θdw =√2BT(∂θRt) dt. As a consequence, the actual FI for the mea-

surement performed at time t + dt is evaluated by averagingover all the possible trajectories, i.e.:

dFt(θ) = Edw[dF(traj)t (θ)] . (16)

Finally, if one consider the whole data set, i.e. the continuousstream of measurement outcomes d = xmtt′=0 obtainedup to time t, the FI corresponding to the whole a posterioridistribution p(d|θ) can be calculated, exploiting its additiveproperty [44], by integrating it (numerically or analytically)as

Ft(θ) =

∫ t

t′=0

dFt′(θ). (17)

V. EXAMPLES

A. Estimation of the squeezing parameter for a quantumparametric amplifier

Let us consider the Hamiltonian for a degenerate paramet-ric amplifier, which in the cavity mode rotating frame readsHs = −χ(xp + px)/2, and assume that the correspondingcavity mode is weakly interacting with a bath at zero tem-perature, such that the interaction matrix corresponds to abeam-splitter with C =

√κΩ and the correlations of the bath

are described by σb = 12. The corresponding master equa-tion for the density operator reads % = −i[Hs, %] + κD[a]%

where a = (x + ip)/√

2 is the bosonic annihilation operator.In our formalism the dynamics is described by the followingdrift and diffusion matrices: A = diag(−χ− κ/2, χ− κ/2),D = κ12.We now assume to perform a time-continuous measurementon the cavity output, i.e. on the environmental modes afterthe interaction with the cavity mode, via a Gaussian measure-ment, in order to estimate the squeezing coupling constant χ.In our analysis we will focus on two type of measurements:a time-continuous homodyne measurement of a quadrature

Page 4: Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni1 1Quantum Technology

4

0 10 20 30 400

50

100

150

200

κt

F t(χ)

FIG. 1. FI Ft(χ) for continuous homodyne and heterodyne detec-tion as a function of time and for χ = −0.2κ (numerical evaluationwith 2000 trajectories).Green dotted line: homodyne detection of quadrature x; red dashedline: homodyne detection of quadrature p; blue solid line: hetero-dyne detection.

r(φ) = (cosφ x+ sinφ p)/√

2 and a time-continuous hetero-dyne detection, i.e. the projection on single-mode coherentstates (the details on the Gaussian description of these mea-surements can be found in Appendix A).Under these assump-tions, Eq. (14) for the infinitesimal FI simplifies to the follow-ing equations

dFt = 2κEdw[(∂χ〈rφ〉t)2] dt homodyne quadrature rφdFt = κEdw[(∂χ〈x〉t)2 + (∂χ〈p〉t)2] dt heterodyne

In Fig. 1 we plot the FI Ft(χ) as a function of time. We noticethat the best performances, between the three measurementshere considered, are obtained via a time-continuous homo-dyne measurement of the quadrature x, which, for the parame-ters we have chosen, is the quadrature being anti-squeezed bythe Hamiltonian Hs. One can also observe how at long times,all three curves present a linear behaviour, indicating that theinfinitesimal FI takes the form dFt(χ) = K dt, and that thehomodyne monitoring of p overcomes the performance of het-erodyne detection.

B. Estimation of a constant force on a mechanical oscillator

We consider a standard cavity optomechanical setup wherea mechanical oscillator oscillating at frequency ωm, is cou-pled to a cavity mode characterized by a resonance frequencyωc and driven with a laser at frequency ωl = ∆ + ωc. The in-teraction Hamiltonian is linearized as Hint = g xmxc, and asusual we consider cavity decay rate κ, while the mechanicaloscillator is coupled to a phononic Markovian bath charac-terized by nth thermal phonons, and decoherence rate γ [45].A constant force is exerted on the mechanical oscillator, de-scribed by the Hamiltonian Hλ = λxm, where λ is the param-

eter we want to estimate. The details of the master equationfor the two-mode density operator and of the correspondingGaussian description can be found in Appendix C.In order to estimate the force parameter λ, the environment ofthe cavity field, i.e. the cavity output field, is measured contin-uously as in [30] (it is possible to include also the continuousmeasurement on the oscillator environment, which can be per-formed experimentally in certain opto-mechanical setups, forexample by monitoring the light scattered from a levitatingnanosphere [14, 46]).Here we consider continuous homodyne detection with finiteefficiency η, whose FI can be easily evaluated via the formu-las presented in the previous example. It is important to noticehow in this case, only the vector u depends on the parameter.As a consequence, since ∂λA = 0, also the matrix ∂λσt = 0at any time, and the evolution of the vector ∂λRt is completelydeterministic, reading

d(∂λRt)

dt= [A+ (σtB +N)BT](∂λRt) + ∂λu . (18)

It is then not necessary then to average the infinitesimal FIin Eq. (14), and its value can be easily obtained numericallywithout needing to average over thousands of trajectories. Wefind that, as one could expect, the optimal measurement is ob-tained for φ = π/2, i.e. for homodyne monitoring of quadra-ture pc; moreover in Fig. 2 we report the behaviour of theFisher information as a function of time and for different val-ues of the loss parameter κ and we observe that the Fisherinformation is monotonically increasing with κ for all the val-ues of ωmtwe have investigated. Also this results is somehowexpected as κ represents in this picture the strength of the mea-surement performed via the environmental modes.This example clearly shows the potential of our method: infact, in order to evaluate the FI of this estimation problemwith the method described in [24], it would have been nec-essary to integrate numerically a stochastic master equation,over around thousands trajectories, for two-mode operators,and thus corresponding to approximated matrices of dimen-sion (dmdc)× (dmdc) (dm and dc being the truncated dimen-sions of the Fock space for respectively the mechanical oscil-lator and the cavity field).

We want to remark that analytical solutions can also beobtained in a similar experimentally relevant scenarios. Wereport here the example of single-mode force estimationwith time-continuous monitoring described by the followingstochastic master equation, that for example describes con-tinuous monitoring of a levitated nanosphere undergoing mo-mentum diffusion in situations where the interaction with thecavity mode can be neglected [15, 47],

d% = −iλ[p, %] dt+ κD[x]% dt+√ηκH[x]% dw . (19)

The covariance matrix of the conditional state evolves deter-ministically as

σt =

(1

1+2ηκt 0

0 1 + 2κt

). (20)

Page 5: Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni1 1Quantum Technology

5

0 20 40 60 80 1000

2

4

6

8

ωm t

F t(λ)

FIG. 2. FI Ft(λ) for continuous homodyne detection of the cavityfield quadrature pc as a function of time and for different values ofthe cavity decay rate: κ = ωm/2 - red solid line; κ = ωm/10 - greendashed line; κ = ωm/20 - blue dotted line (the other parameters arechosen as: g = ωm/2, γ = ωm/3, η = 1). The inset shows thebehaviour of the FI at small times.

In order to evaluate the Fisher information corresponding tothe estimation of the parameter λ, one only needs the vector∂λ〈x〉t, whose evolution is described by the equation

d(∂λ〈x〉t)dt

= − 2ηκ

1 + 2ηκt(∂λ〈x〉t)− 1 , (21)

and that can be analytically solved as

∂λ〈x〉t = − 1 + ηκt

1 + 2ηκtt . (22)

The corresponding infinitesimal and total Fisher informationcan be evaluated straightforwardly and one obtains

dFt(λ) =4t2ηκ(1 + tηκ)2

(1 + 2tηκ)2,

Ft(λ) =2t3ηκ(2 + tηκ)

3(1 + 2tηκ). (23)

In this case one can analytically check that Ft(λ) is monoton-ically increasing with κ, and one also observes that, for largemonitoring times, the Fisher information presents a remark-able t3-scaling.

VI. DISCUSSION AND OUTLOOKS

We have presented a reliable method for the calculation ofthe FI for dynamical parameter estimation in continuouslymeasured linear Gaussian quantum systems. As shown in thetwo examples here described, our method greatly simplifies,in terms of computation complexity, the calculation of boundson the estimation of parameters for such a rich and physically

relevant family of quantum systems, compared to both themethod presented in Ref. [24] and the ones derived forgeneral classical Gaussian systems [48–51]. It can alsoprovide analytical results, allowing one to investigate in moredetail the role played by the different physical parametersand to compare easily the efficiency of different measurementstrategies. Furthermore, as Eqs. (7) are formally equivalentto the classical continuous-time Kalman filter, our methodcan be generalized to the classical case (a more detaileddiscussion can be found in Appendix D).Our results will find applications in assessing the perfor-mances of quantum sensors in several physical systems. Itis worth mentioning the estimation of a magnetic field, incases where an atomic spin ensemble can be approximatedby a bosonic field via the Holstein-Primakovv approximation[18], and in several other quantum optomechanics setups andestimation problems, in particular with the aim of testingfundamental theories as corrections to Newtonian gravity[52] or to the Schroedinger equation [15, 53]. Moreover thisapproach can be generalized to the estimation of stochasticparameters, e.g. stochastic forces on mechanical oscillators,and of parameters characterizing the interaction with theenvironment and its temperature.

Notice that this arXiv version of the manuscript includesthe Erratum of the version published in Phys. Rev. A.

ACKNOWLEDGMENTS

The author thanks S. Conforti, M. Paris and A. Serafini fordiscussions and constant support, A. Mari for useful discus-sions regarding the additive property of the Fisher informa-tion and F. Albarelli for several discussions that contributed tofind the error in the previous version of the manuscript. Theauthor acknowledges support from Marie Skłodowska-CurieAction H2020-MSCA-IF-2015 (project ConAQuMe, grant nr.701154).

Appendix A: Gaussian (general-dyne) measurements

We here briefly present the parametrization of Gaussianmeasurements that are discussed and used in the article.A Gaussian measurement is univocally described by a ma-trix σm, s.t. σm + iΩ ≥ 0 and the corresponding measure-ment outcomes are described by a vector xm. We start byfocusing on single-mode projective Gaussian measurements,which thus correspond in the Hilbert space to projection ontoa single-mode state |ψG〉. As the most general single-modeGaussian state is a displaced squeezed vacuum state, the cor-responding general matrix σm can be written as

σm(s, φ) = R(φ)

(s 00 1/s

)R(φ)T , (A1)

Page 6: Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni1 1Quantum Technology

6

with

R(φ) =

(cosφ sinφ− sinφ cosφ

). (A2)

In the case of homodyne measurement of the quadrature op-erator rφ = cosφ x + sinφ p, one has to evalulate the limitσ

(hom)m = lims→0 σm(s, φ), while for heterodyne detec-

tion, i.e. projection onto coherent states one has σ(het)m =

σm(1, φ).In order to take into account of inefficient detection, the mea-surement matrix σ

(η)m is calculated via the action of the dual

noisy map on the projective measurement covariance matrixof Eq. (A1) as [36]

σ(ineff)m = X∗σmX

∗T + Y ∗ , (A3)

where

X∗ = 12/√η ,

Y ∗ =1− ηη

12 ,

and η quantifies the detection efficiency.Notice that if one wants to consider a two-mode (or evenmulti-mode) local measurement, the overall measurement ma-trix σm is obtained by taking the direct sum of the single-mode measurements, e.g. σm = σm,(a) ⊕ σm,(b). If one ofthe modes is not monitored, then one has to use the inefficientmeasurement matrix σ

(ineff)m as in Eq. (A3) and take the limit

for the corresponding efficiency parameter η → 0.

Appendix B: Diffusive and conditional evolution undertime-continuous measurements

In this section we will briefly describe the formalism andthe calculations developed in Ref. [36] that lead to the equa-tions describing the evolution of Gaussian states under time-continuous general-dyne measurements on the environment.To keep the presentation easier, we consider the case where noHamiltonian for the system is present, i.e. Hs = 0 and we fo-cus on the interaction between the system and the bath degrees

of freedom. This interaction is described by the Hamiltonian

HC = rTC rb(t) =1

2rTsbHC rsb =

1

2rTsb

(0 CCT 0

)rsb ,

(B1)

where rb(t) are the bath operators, defined by the white-noisecondition

〈rb(t), rTb (t′)〉 = σbδ(t− t′) , (B2)

and rTsb = (rT, rb(t)

T). From Eq. (B2) one notices that theoperators rb(t) have the dimensions of the square root of afrequency. One can then define the so called quantum Wienerincrement drb(t) as [5, 41]:

drb(t) = rb(t) dt = r′b(t) dW (B3)and impose that r′b(t) is a vector of dimensionless field opera-tors (that can be associated with detector clicks in the labora-tory, and formally with POVM operators in the Hilbert space),satisfying the canonical commutation relations,

[r′b(t), r′b(t)] = iΩ . (B4)

By observing that

[drb(t),drb(t)] = [r′b(t) dW, r′b(t) dW ] = iΩ dW 2 , (B5)

= [rb(t) dt, rb(t) dt] = iΩ dt , (B6)

one obtains the relationship dW 2 = dt. One can then in-terpret dW as a stochastic Wiener increment, which is indeedresponsible for the diffusive behaviour of the dynamics (in theHeisenberg picture the system and bath operators show in facta random-walk like evolution). It is important to remark thatthe properties of dW are a consequence of the white-noisecondition describing the input operators rb(t) (other types ofcorrelations would lead to a different stochastic behaviour).We can now apply the Gaussian formalism to the vector of(well-defined) canonical operators r′Tsb = (rT, r′b(t)

T). Un-der the coupling described in Eq. (B1), the dynamics overan interval dt is generated by the operator rT

sbHC rsb dt =r′TsbHC r′sbdW . By expanding the corresponding symplectictransformation as

eΩHCdW ≈(1+ ΩHC dW +

(ΩHC)2

2dt

), (B7)

one calculates the evolution of the system-bath covariance ma-trix as

eΩHCdW (σ ⊕ σb) e(ΩHC)TdW ≈ (σ ⊕ σb) +(Aσ + σAT +D

)⊕ σb dt+ σsb dW , (B8)

where A = (ΩCΩCT)/2 and D = ΩCσbCTΩT are typi-

cally addressed as drift and diffusion matrices, while the othermatrices are

σsb =

(0 ΩCσb + σCΩT

σbCTΩT + ΩCTσ 0

), (B9)

σb =ΩCTΩCσb + σbC

TΩCΩ

2+ ΩTCTσCΩ . (B10)

Page 7: Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni1 1Quantum Technology

7

The same procedure can be applied in order to obtain the evo-lution of the first moments, s.t.

eΩHCdW

(RRb

)≈(1+ ΩHCdW +

(ΩHC)2

2dt

)(RRb

)(B11)

≈(

R +AR dtΩCTR dW

). (B12)

In the last equation we have assumed Rb = 0, i.e. that thebath operators have first moments equal to zero.The probability density for the outcomes xm of a general-dyne measurement on the bath degrees of freedom, describedby the covariance matrix σm, is a multivariate Gaussian dis-tribution

p(xm) =e−

12 (xm−xm)TΣ−1(xm−xm)

(2π)n√

Det[Σ](B13)

with first moments vector xm = ΩCTR dW , and with co-variance matrix Σ = (σb + σm)/2. One can then define thevector of random variables

dw =

(σb + σm

2

)−1/2

(xm − xm) dW (B14)

and, by exploiting the statistical properties of the measure-ment outcomes xm, one observe that dw is in fact a vectorof uncorrelated real Wiener increments, i.e. it is distributedaccording to a Gaussian with first moments E[dw] = 0 andcovariance matrix E[dw, dwT/2] = 1dt. It is often cus-tomary to express the measurement outcomes as a real currentwith uncorrelated noise [38]:

dy := (σb + σm)−1/2 xm dW

= (σb + σm)−1/2xm dW +dw√

2

= (σb + σm)−1/2ΩCTRt dt+dw√

2. (B15)

Finally, starting from Eqs. (B8) and (B12), and by apply-ing the formulas for the evolution of Gaussian states underconditional measurements on one of the two subsystems, oneobtains the equations [36]

dRt = ARt dt+ u dt+

(σtB +N√

2

)dw ,

dσtdt

= Aσt + σtAT +D − (σtB +N)(σtB +N)T .

(B16)

Appendix C: Estimation of a constant force on a mechanicaloscillator

We consider a standard cavity optomechanical setup wherea mechanical oscillator, described by position and momentum

operators xm and pm and oscillating at frequency ωm, is cou-pled to a cavity mode, described by amplitude and phase op-erators xc and pc and with resonance frequency ωc. Assumingthat the cavity is strongly driven with a laser at frequency ωl,one can consider the following linearized Hamiltonian in aframe rotating with ωl:

Hom = ωm(x2m + p2

m)/2−∆(x2c + p2

c)/2 + g xmxc ,(C1)

where ∆ = ωl − ωc is the detuning of the driving laser withrespect to the cavity, and g is the effective optomechanicalcoupling strength. If a constant force is exerted on the me-chanical oscillator, then one has to add the linear HamiltonianHλ = λxm. We also assume that the cavity has a decay rateκ, while the mechanical oscillator is coupled to a phononicMarkovian bath characterized by nth thermal phonons, andthe corresponding width of the mechanical resonance is equalto γ. The master equation for the two-mode density operatorreads

d%

dt= −i[Hom + Hλ, %] + κD[a]%+ γ(nth + 1)D[b]%

+ γnthD[b†]% , (C2)

where a = (xc + pc)/√

2 and b = (xm + pm)/√

2.

In the Gaussian picture the interaction with the environmentand its correlations are described by the matrices

C =

0 −√κ 0 0√

κ 0 0 00 0 0

√γ

0 0 −√γ 0

σb =

1 0 0 00 1 0 00 0 1 + 2nth 00 0 0 1 + 2nth

.

By also assuming that a continuous homodyne measurementwith efficiency η is performed on the environment (in this caseon the cavity output field), the dynamics is equivalently de-scribed by the equations for the first moment vector and forthe covariance matrix (B16). The corresponding matrices andvectors read

A =

−κ/2 −∆ 0 0∆ −κ/2 −g 00 0 −γ/2 ωm−g 0 −ωm −γ/2

D =

κ 0 0 00 κ 0 00 0 γ(1 + 2nth) 00 0 0 γ(1 + 2nth)

B = −N =

√ηκ cos2 φ

√ηκ sinφ cosφ 0 0

−√ηκ sinφ cosφ√ηκ sin2 φ 0 0

0 0 0 00 0 0 0

uT = (0 0 0 − λ) .

Page 8: Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni1 1Quantum Technology

8

As only the vector u depends on the parameter λ, with∂λu

T = (0 0 0,−1), in order to calculate the infinitesimal FI,one has only to solve the following differential equation for(∂λRt):

d(∂λRt)

dt= [A+ (σtB +N)BT](∂λRt) + ∂λu . (C3)

Appendix D: Generalization to the classical continuous-timeKalman filter

The equations (7) describing the Gaussian quantum dynam-ics are in fact formally equivalent to the classical continuous-time Kalman filter. As a consequence we can generalize themethod to the classical case, obtaining similar equations asthe ones presented in Refs.[48–51].The goal of the Kalman filter is to obtain an estimate xt of acertain process xt, via a continuous measurement output yt

with uncorrelated noise. The corresponding equations read

dxt = Axt dt+ u dt+ (ΣtB +N) dw ,

dΣt

dt= AΣt + ΣtA

T +D − (ΣtB +N)(ΣtB +N)T ,

dyt = BTxt dt+ dw , (D1)

where Σt represents the mean squared error matrix of theestimate. As the increment of the time continuous outputdyt is a Gaussian random variable with mean value vector〈dyt〉 = BTxt dt and covariance matrix Γt = 1dt, the cor-responding infinitesimal FI (for a specific trajectory) can becalculated by using Eq. (13). In particular one obtains

dF(traj)t (θ) =

[∂θ(B

Txt)]T [

∂θ(BTxt)

]dt , (D2)

where the evolution of the vector ∂θ(BTxt) can be derivedas in Eqs. (15). In order to evaluate the FI for the clas-sical continuous-time Kalman filter, corresponding to thewhole stream of outcomes up to time t, one can then averagedF

(traj)t (θ) over the Wiener process and perform the integral

over time as in Eqs. (16) and (17).

[1] V. Giovannetti, S. Lloyd and L. Maccone, Nature Photon. 5,222 (2011).

[2] R. Demkowicz-Dobrzanski, M. Jarzyna and J. Kolodynsky,Progress in Optics 60, 345 (2015).

[3] B. M. Escher, R. L. de Matos Filho, L. Davidovich, NaturePhys. 7, 406 (2011).

[4] R. Demkowicz-Dobrzanski, J. Kolodynsky and M. Guta, Na-ture Communications 3, 1063 (2012).

[5] H. M. Wiseman and G. J. Milburn, Quantum Measurement andControl (Cambridge University Press, New York, 2010).

[6] K. Jacobs and D. A. Steck, Contemporary Physics 47, 279(2006).

[7] H. M. Wiseman and G. J. Milburn, Phys. Rev. A, 49, 1350(1994).

[8] S. Mancini and H. M. Wiseman, Phys. Rev. A, 75, 012330(2007).

[9] A. Serafini and S. Mancini, Phys. Rev. Lett., 104, 220501(2010).

[10] A. Szorkovszky, A. C. Doherty, G. I. Harris and W. P. Bowen,Phys. Rev. Lett. 107, 213603 (2011).

[11] M. G. Genoni, S. Mancini and A. Serafini, Phys. Rev. A 87,042333 (2013).

[12] M. G. Genoni, S. Mancini, H. M. Wiseman, and A. Serafini,Phys. Rev. A 90, 063826 (2014).

[13] S. G. Hofer and K. Hammerer, Phys. Rev. Lett. 91, 033822(2015).

[14] M. G. Genoni, J. Zhang, J. Millen, P. F. Barker and A. Serafini,New J. Phys. 17, 073019 (2015).

[15] M. G. Genoni, O. S. Duarte and A. Serafini, New J. Phys. 18,103040 (2016).

[16] D. Tan, S. Weber, I. Siddiqi, K. Molmer, K. W. Murch, Phys.Rev. Lett. 114, 090403 (2015).

[17] N. Foroozani, M. Naghiloo, D. Tan, K. Molmer and K. W.Murch, Phys. Rev. Lett. 116, 110401 (2016)

[18] J. M. Geremia, J. K. Stockton, A. C. Doherty and H. Mabuchi,Phys. Rev. Lett. 91, 250801 (2003).

[19] D. W. Berry and H. M. Wiseman, Phys. Rev. A 65, 043803(2002).

[20] M. Tsang, H. M. Wiseman and C. M. Caves, Phys. Rev. Lett.106, 090401 (2011)

[21] H. Yonezawa et al., Science 337, 1514 (2012).[22] S. Z. Ang, G. I. Harris, W. P. Bowen, New J. Phys. 15, 103028

(2013).[23] K. Iwasawa, K. Makino, H. Yonezawa, M. Tsang, A. Davi-

dovic, E. Huntington and A. Furusawa, Phys. Rev. Lett. 111,163602 (2013).

[24] S. Gammelmark and K. Molmer, Phys. Rev. A 87, 032115(2013).

[25] S. Gammelmark and K. Molmer, Phys. Rev. Lett. 112, 170401(2014).

[26] A. H. Kiilerich and K. Molmer, Phys. Rev. A 91, 012119(2015).

[27] S. Ng, S. Z. Ang, T. A. Wheatley, H. Yonezawa, A. Furu-sawa, E. H. Huntington and M. Tsang, Phys. Rev. A 93, 042121(2016).

[28] M. Guta and N. Yamamoto, IEEE Trans. Automatic Control61-4, 921 (2016).

[29] L. F. Buchmann, S. Schreppler, J. Kohler, N. Spethmann, andD. M. Stamper-Kurn, Phys. Rev. Lett. 117, 030801 (2016).

[30] W. Wieczorek, S. G. Hofer, J. Hoelscher-Obermaier, R.Riedinger, K. Hammerer, and M. Aspelmeyer, Phys. Rev. Lett.114, 223601 (2015).

[31] C. W. Helstrom, Quantum Detection and Estimation Theory(Academic Press, New York, 1976).

[32] S. Braunstein and C. Caves, Phys. Rev. Lett. 72, 3439 (1994).[33] M. G. A. Paris, Int. J. Quant. Inf. 7, 125 (2009).[34] A. Ferraro, S. Olivares and M.G. A. Paris, Gaussian States in

Quantum Information, (Bibliopolis, Napoli, 2005)

Page 9: Cram´er-Rao bound for time-continuous measurements in ... fileCram´er-Rao bound for time-continuous measurements in linear Gaussian quantum systems Marco G. Genoni1 1Quantum Technology

9

[35] C. Weedbrook, S. Pirandola, R. Garca-Patron, N. J. Cerf, T. C.Ralph, J. H. Shapiro, and S. Lloyd, Rev. Mod. Phys. 84, 621(2012).

[36] M. G. Genoni, L. Lami and A. Serafini, Contemporary Physics57, 331 (2016).

[37] H. M. Wiseman and L. Diosi, Chem. Phys. 268, 91 (2001).[38] H. M. Wiseman and A. C. Doherty, Phys. Rev. Lett. 94, 070405

(2005).[39] Throughout the paper, the quantities inside commutator or anti-

commutator has to be considered as an outer product. In com-ponents, given a vector of hermitian operators a: [a, aT]jk =aj ak − akaj and a, aTjk = aj ak + akaj .

[40] R. Simon, N. Mukunda, and B. Dutta, Phys. Rev. A 49, 1567(1994).

[41] C. Gardiner and P. Zoller, Quantum Noise (Springer, Heidel-berg, 2010).

[42] H. Cramer, Mathematical methods of statistics, (Princeton Uni-versity Press, 1946)

[43] We remark that all the results we will present in the followingcan be generalized to the case where the parameter θ character-izes the dissipative dynamics, either via the interaction betweenthe system and the environment, or via the correlations of theenvironment itself.

[44] After having averaged the Fisher information over all the tra-jectories up to time t, the measurement at later times is com-

pletely independent on the previous ones and thus one can ex-ploit the additive property. It is useful to remark that one obtainsthe same exact result by calculating the overall Fisher informa-tion Ft(θ) directly starting from the conditional probability ofthe complete trajectory p(d|θ).

[45] M. Aspelmeyer, T. J. Kippenberg and F. Marquardt, Rev. Mod.Phys. 86, 1391 (2014).

[46] Y. L. Li, J. Millen and P. F. Barker, Optics Express 24, 1392(2016).

[47] A. C. Pflanzer, O. Romero-Isart and J. I. Cirac, Phys. Rev. A86, 013802 (2012).

[48] J. E. Cavanaugh and R. H. Shumway, Stat. Probabil. Lett. 26,347 (1996).

[49] A. Klein, G. Melard and T. Zahaf, Linear Algebra Appp. 321,209 (2000).

[50] A. Klein and H. Neudecker, Linear Algebra App. 321, 233,(2000).

[51] R. J. Ober, Syst. Control Lett. 47, 221 (2002).[52] A. A. Geraci, S. B. Papp and J. Kitching, Phys. Rev. Lett. 105,

101101 (2010).[53] A. Bassi et al., Rev. Mod. Phys. 85, 471 (2013).