Welcome to visit Communications in Theoretical Physics,
Quantum Physics and Quantum Information

Quantum metrology with optimal control under an arbitrary non-Markovian bosonic environment

  • Likun Zhou ,
  • Wen Yang
Expand
  • Beijing Computational Science Research Center, Beijing 100193, China

Received date: 2024-06-18

  Revised date: 2024-09-25

  Accepted date: 2024-09-26

  Online published: 2024-11-27

Copyright

© 2024 Institute of Theoretical Physics CAS, Chinese Physical Society and IOP Publishing. All rights, including for text and data mining, AI training, and similar technologies, are reserved.

Abstract

Finding the optimal control is of importance to quantum metrology under a noisy environment. In this paper, we tackle the problem of finding the optimal control to enhance the performance of quantum metrology under an arbitrary non-Markovian bosonic environment. By introducing an equivalent pseudomode model, the non-Markovian dynamic evolution is reduced to a Lindblad master equation, which helps us to calculate the gradient of quantum Fisher information and perform the gradient ascent algorithm to find the optimal control. Our approach is accurate and circumvents the need for the Born–Markovian approximation. As an example, we consider the frequency estimation of a spin with pure dephasing under two types of non-Markovian environments. By maximizing the quantum Fisher information at a fixed evolution time, we obtain the optimal multi-axis control, which results in a notable enhancement in quantum metrology. The advantage of our method lies in its applicability to the arbitrary non-Markovian bosonic environment.

Cite this article

Likun Zhou , Wen Yang . Quantum metrology with optimal control under an arbitrary non-Markovian bosonic environment[J]. Communications in Theoretical Physics, 2025 , 77(2) : 025108 . DOI: 10.1088/1572-9494/ad7fd4

1. Introduction

Quantum metrology is of importance to multiple areas of scientific research, including gravitational wave detection [1], biological sensing [2], atomic clocks [3], magnetometry [4, 5], and so on. The task is to estimate unknown parameters encoded in quantum systems. The achievable estimation precision realized in classical physics is subject to the shot noise limit (SNL) $1/\sqrt{N}$, where N is the number of resources employed in the measurements. If one adopts nonclassical resources such as squeezing and entanglement, the achievable estimation precision can beat the SNL [6, 7]. Especially, for some states such as Greenberger–Horne–Zeilinger states, the estimation precision can reach the Heisenberg limit (HL) 1/N [8, 9]. However, in real experiments, the system of interest inevitably couples to its surrounding environment, which is known as an open quantum system (OQS). The dynamics of an OQS are characterized by decoherence, which inevitably deteriorates the estimation precision.
The achievable precision under the noisy environment has been investigated previously. For frequency estimation under a Markovian pure dephasing noise, the estimation precision returns to the SNL [10]. For a relaxation noise, the estimation precision scales asymptotically as N−5/6 [11]. Some estimation protocols under the non-Markovian environment have also been investigated. It was predicted in the Ramsey spectroscopy that the precision can reach N−3/4 [12], which was argued to be a consequence of the quantum Zeno effect at short time scales. This prediction was recently confirmed in a real experiment [13]. Afterwards, a novel limit on the attainable precision in frequency estimation was derived by [14], which holds for all forms of phase covariant uncorrelated noise. In recent years, Bai et al proposed that if the probe and environment formed a bound state [1517], the estimation precision can asymptotically recover to the ideal case in a long-time condition, where the dissipation was exactly studied by non-Markovian dynamic evolution. Tan et al have found that using coherent spin states can reach a 1/(Nt) scale estimation precision of the spectral density for super-Ohmic environment [18]. Those studies highlight the advantages of non-Markovian quantum metrology.
To improve the estimation precision, researchers have proposed using proper controls to mitigate the harmfulness of noises [1922]. Recently, some researchers focused on finding optimal controls based on different algorithms [2328] to enhance the metrology performance. A particularly efficient method is gradient-based algorithms such as Gradient Ascent Pulse Engineering [2325]. Other methods based on the Pontryagin Maximum Principle [26, 27] are another option. In those works, they assume that the system is dissipated in the Markovian environment. Under the premise of the Born–Markov approximation, the reduced system dynamics can be described by a Lindblad master equation [29, 30]. However, such approximation is valid only for a memoryless environment. In many practical scenarios with long-time environmental correlation, Markovian approximation is no longer valid, leading to distinct non-Markovian behavior [1217, 3134]. Therefore, finding the optimal control under non-Markovian environments becomes a subject of interest.
In this work, we tackle the problem of finding the optimal control under an arbitrary non-Markovian bosonic environment, where the environment is characterized by its spectral density. By introducing a pseudomode model, the original environment can be replaced by few dissipated pseudomode harmonic oscillators with proper parameters. The evolution of the system's density matrix can be equivalently and accurately described by a Lindblad master equation, which provides a simple description for an OQS. With the help of this pseudomode model, we derivate the gradient of quantum Fisher information (QFI), which allows us to use the gradient ascent algorithm to find the optimal control making QFI maximal. To show the effectiveness of our method, we consider the frequency estimation of a spin with multi-axis control under Lorentz and Ohmic spectral environments. Our results demonstrate a significant enhancement in QFI, which shows the effectiveness of our method in optimizing quantum metrology protocols under a non-Markovian environment.

2. Accurate description of the non-Markovian dissipation model

As illustrated in figure 1(a), we consider a system coupling to a bosonic environment. The total Hamiltonian is given by
$\begin{eqnarray}{\hat{H}}_{\mathrm{tot}}={\hat{H}}_{{\rm{s}}}+{\hat{H}}_{\mathrm{env}}+\hat{V},\end{eqnarray}$
where ${\hat{H}}_{{\rm{s}}}$, ${\hat{H}}_{\mathrm{env}}$, $\hat{V}$ are the Hamiltonians of the system, environment and interaction, respectively. The environment Hamiltonian reads
$\begin{eqnarray}{\hat{H}}_{\mathrm{env}}=\displaystyle \sum _{k}{\omega }_{k}{\hat{b}}_{k}^{\dagger }{\hat{b}}_{k},\end{eqnarray}$
where bk is kth boson mode's annihilation operator of the environment, with the characteristic frequency ωk. Initially, the environment is in a thermal state with a temperature of T. We consider the interaction Hamiltonian
$\begin{eqnarray}\hat{V}=\hat{S}\otimes \hat{B},\end{eqnarray}$
where $\hat{S}$ is the operator of system and $\hat{B}={\sum }_{k}{g}_{k}{\hat{b}}_{k}+{g}_{k}^{* }{\hat{b}}_{k}^{\dagger }$ is the operator of environment.
Figure 1. Two equivalent open quantum system models used for noisy quantum metrology. Under the influence of control and environment, the system evolves to the final state ρ(θ) encoded by an unknown parameter θ. (a) The original open quantum system model. The system couples to a non-Markovian bosonic environment characterized by a certain spectrum. (b) An equivalent pseudomode model. The system couples to few pseudomode harmonic oscillators, but the oscillators are collectively dissipated in a Markovian bosonic environment. The two models are equivalent if their correlation functions are equal.
The effects of the system-environment interaction are characterized by the correlation function (COF), namely
$\begin{eqnarray}C(t)\equiv \langle \hat{B}(t)\hat{B}\rangle ,\end{eqnarray}$
where $\hat{B}(t)\equiv {{\rm{e}}}^{{\rm{i}}{\hat{H}}_{\mathrm{env}}t}\hat{B}{{\rm{e}}}^{-{\rm{i}}{\hat{H}}_{\mathrm{env}}t}$ and ⟨⋯⟩ denotes the expectation value with respect to the initial state of the environment. Defining the spectral density $J(\omega )\equiv {\sum }_{k}{\left|{g}_{k}\right|}^{2}\delta \left(\omega -{\omega }_{k}\right)$ in continuum limit, the COF is given by [35]
$\begin{eqnarray}C(t)=\int {\rm{d}}\omega J(\omega )\left[\coth \left(\displaystyle \frac{\beta \omega }{2}\right)\cos (\omega t)-{\rm{i}}\,\sin (\omega t)\right],\end{eqnarray}$
where β = 1/(KbT) and Kb is the Boltzmann constant. If the system's initial state $\rho \left(0\right)$ and the operator $\hat{S}$ are given, the reduced density matrix of the system is fully dominated by the COF [3640], which is the characteristic of Gaussian bosonic environment.
The original model can also be exactly simulated by an equivalent pseudomode model shown in figure 1(b). The system couples to M pseudomode harmonic oscillators, but the oscillators are collectively dissipated in a pseudo environment. The total Hamiltonian of the pseudomode model is given by
$\begin{eqnarray}{\hat{H}}_{\mathrm{pse}}={\hat{H}}_{{\rm{s}}}+\hat{S}\otimes {\hat{B}}_{\mathrm{pse}}+{\hat{H}}_{{\rm{B}},\mathrm{pse}},\end{eqnarray}$
with
$\begin{eqnarray}{\hat{B}}_{\mathrm{pse}}=\displaystyle \sum _{j=1}^{M}{\eta }_{j}{\hat{c}}_{j}^{\dagger }+{\eta }_{j}^{* }{\hat{c}}_{j},\end{eqnarray}$
and
$\begin{eqnarray}\begin{array}{rcl}{\hat{H}}_{{\rm{B}},\mathrm{pse}} & = & \displaystyle \sum _{j=1}^{M}{{\rm{\Omega }}}_{j}{\hat{c}}_{j}^{\dagger }{\hat{c}}_{j}+\displaystyle \sum _{k}{\omega }_{k}{\hat{b}}_{k}^{\dagger }{\hat{b}}_{k}\\ & & +\displaystyle \sum _{k}\displaystyle \sum _{j=1}^{M}{g}_{j,k}{\hat{b}}_{k}{\hat{c}}_{j}^{\dagger }+{g}_{j,k}^{* }{\hat{b}}_{k}^{\dagger }{\hat{c}}_{j}.\end{array}\end{eqnarray}$
Here, ηj denotes the interaction strength between the system and jth harmonic oscillator, gj,k denotes the coupling strength between the jth harmonic oscillator and kth boson mode of the pseudo environment. For convenience, we consider gj,k = μj, i.e., the pseudo environment is Markovian, which gives a simplified description of the original model. If we regard the system and harmonic oscillators as a whole, it undergoes dissipation within a Markovian environment, which can be described by a Lindblad master equation [40]
$\begin{eqnarray}\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}}{\partial t}=-{\rm{i}}\left[\hat{H},{\rho }_{\mathrm{sc}}\right]+{ \mathcal D }\left[\displaystyle \sum _{j=1}^{M}{\mu }_{j}{\hat{c}}_{j}\right]{\rho }_{\mathrm{sc}}\equiv { \mathcal L }{\rho }_{\mathrm{sc}},\end{eqnarray}$
where ${ \mathcal L }$ is the Liouvillian superoperator and
$\begin{eqnarray}\hat{H}={\hat{H}}_{s}+\displaystyle \sum _{j=1}^{M}{{\rm{\Omega }}}_{j}{\hat{c}}_{j}^{\dagger }{\hat{c}}_{j}+\hat{S}\otimes {\hat{B}}_{\mathrm{pse}},\end{eqnarray}$
$\begin{eqnarray}{ \mathcal D }[\hat{A}]{\rho }_{\mathrm{sc}}\equiv \hat{A}{\rho }_{\mathrm{sc}}{\hat{A}}^{\dagger }-\displaystyle \frac{{\rho }_{\mathrm{sc}}{\hat{A}}^{\dagger }\hat{A}+{\hat{A}}^{\dagger }\hat{A}{\rho }_{\mathrm{sc}}}{2}.\end{eqnarray}$
Using the matrix representation of the superoperator [41], the solution of equation (9) is given by
$\begin{eqnarray}{\rho }_{\mathrm{sc}}\left(t\right)={{\rm{e}}}^{{ \mathcal L }t}{\rho }_{\mathrm{sc}}\left(0\right),\end{eqnarray}$
with the initial state ${\rho }_{\mathrm{sc}}\left(0\right)={\rho }_{{\rm{s}}}\left(0\right)\otimes {\rho }_{\mathrm{vac}}^{\otimes M}$, where ${\rho }_{{\rm{s}}}\left(0\right)$ is the initial state of system and ρvac is the vacuum state of the harmonic oscillator. Note that ρsc is the total density matrix of the system and the pseudomode oscillators, the reduced density matrix of the system is obtained by tracing out the pseudomode oscillators, i.e.,
$\begin{eqnarray}\rho \left(t\right)={\mathrm{Tr}}_{{\rm{c}}}\left[{\rho }_{\mathrm{sc}}\left(t\right)\right].\end{eqnarray}$
The two models in figures 1(a) and (b) are equivalent if their COFs are equal. In other words, given a non-Markovian model shown in figure 1(a), we can always find an equivalent pseudomode model shown in figure 1(b) with appropriate parameters (Ωj, μj, ηj). The procedures of determining the parameters (Ωj, μj, ηj) are shown in appendix A. In the following, we will use the pseudomode model to exactly simulate a certain non-Markovian environment and find the optimal control to improve the performance of quantum metrology.

3. Quantum metrology theory with optimal control

Given the system's reduced density matrix ρ(t, θ) encoding an unknown parameter θ, the estimation precision about θ obeys the quantum Cramér–Rao inequality [9]
$\begin{eqnarray}\delta \theta \geqslant \displaystyle \frac{1}{\sqrt{{F}_{Q}}},\end{eqnarray}$
where
$\begin{eqnarray}{F}_{Q}=\mathrm{Tr}\left(\rho {l}^{2}\right)\end{eqnarray}$
is the QFI and l is the symmetric logarithmic derivative determined by
$\begin{eqnarray}\displaystyle \frac{\partial \rho }{\partial \theta }=\displaystyle \frac{\rho l+l\rho }{2}.\end{eqnarray}$
Considering a non-Markovian model shown in figure 1(a) with a certain environmental spectrum, we can design an equivalent pseudomode model with proper parameters (Ωj, μj, ηj) shown in figure 1(b). Thus, the system's reduced density matrix ρ(t) can be solved by equations (9)–(13). After getting ρ(t) and its derivative, one can calculate ${F}_{Q}\left(t\right)$.
To improve the performance of metrology, we consider a controlled Hamiltonian
$\begin{eqnarray}\hat{H}\left(t\right)=\hat{H}+\displaystyle \sum _{k}{f}_{k}\left(t\right){\hat{H}}_{k},\end{eqnarray}$
where $\hat{H}$ is the free Hamiltonian given by equation (10), $\{{\hat{H}}_{k}\}$ is a set of control applied to the system, and ${f}_{k}\left(t\right)$ is the control functions to be optimized. For such a time-dependent Hamiltonian, the dynamical evolution can be described by the Lindblad master equation
$\begin{eqnarray}\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}}{\partial t}=-{\rm{i}}\left[\hat{H}\left(t\right),{\rho }_{\mathrm{sc}}\right]+{ \mathcal D }\left[\displaystyle \sum _{j=1}^{M}{\mu }_{j}{\hat{c}}_{j}\right]{\rho }_{\mathrm{sc}}\equiv { \mathcal L }\left(t\right){\rho }_{\mathrm{sc}}.\end{eqnarray}$
We emphasize that equation (18) is accurate and does not involve any approximations. The above equation can be solved by the time discretization method. Specifically, we set a time range $t\in \left[0,T\right]$ and separate the range into N segments, which gives a succession of ${f}_{k}\left({t}_{i}\right)$ and ${{ \mathcal L }}_{i}\equiv { \mathcal L }\left({t}_{i}\right)$ with i = 1, ⋯ ,N. The size of each segment is Δt = T/N. Next, performing the iterations
$\begin{eqnarray}{\rho }_{\mathrm{sc}}\left({t}_{i+1}\right)={{\rm{e}}}^{{{ \mathcal L }}_{i}{\rm{\Delta }}t}{\rho }_{\mathrm{sc}}\left({t}_{i}\right),\end{eqnarray}$
and
$\begin{eqnarray}\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}({t}_{i+1})}{\partial \theta }=\displaystyle \frac{\partial {{\rm{e}}}^{{{ \mathcal L }}_{i}{\rm{\Delta }}t}}{\partial \theta }{\rho }_{\mathrm{sc}}\left({t}_{i}\right)+{{\rm{e}}}^{{{ \mathcal L }}_{i}{\rm{\Delta }}t}\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}\left({t}_{i}\right)}{\partial \theta },\end{eqnarray}$
for i = 1 to N, we can obtain ${\rho }_{\mathrm{sc}}\left({t}_{N+1}\right)$ and $\partial {\rho }_{\mathrm{sc}}\left({t}_{N+1}\right)/\partial \theta $. Tracing out the pseudomodes, we obtain $\rho \left(T\right)\,={\mathrm{Tr}}_{{\rm{c}}}\left[{\rho }_{\mathrm{sc}}\left({t}_{N+1}\right)\right]$ and $\partial \rho \left(T\right)/\partial \theta ={\mathrm{Tr}}_{{\rm{c}}}\left[\partial {\rho }_{\mathrm{sc}}\left({t}_{N+1}\right)/\partial \theta \right]$. Using equations (15)–(16), we can calculate the symmetric logarithmic derivative $l\left(T\right)$ and ${F}_{Q}\left(T\right)$.
Due to ${F}_{Q}\left(T\right)$ being highly related to the control fk(ti), we need to find the optimal fk(ti) making ${F}_{Q}\left(T\right)$ maximal. A powerful method is the gradient ascent method, which requires the gradient of ${F}_{Q}\left(T\right)$ with respect to fk(ti). We regard ${f}_{k}\left({t}_{i}\right)$ as variables to be optimized and calculate the corresponding gradients (details are shown in appendix B)
$\begin{eqnarray}\displaystyle \frac{\partial {F}_{Q}\left(T\right)}{\partial {f}_{k}\left({t}_{i}\right)}={\mathrm{Tr}}_{{\rm{s}}}\left[2l{\mathrm{Tr}}_{{\rm{c}}}\displaystyle \frac{{\partial }^{2}{\rho }_{\mathrm{sc}}\left(T\right)}{\partial {f}_{k}\left({t}_{i}\right)\partial \theta }-{l}^{2}{\mathrm{Tr}}_{{\rm{c}}}\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}\left(T\right)}{\partial {f}_{k}\left({t}_{i}\right)}\right].\end{eqnarray}$
Next, we set the initial values of ${f}_{k}\left({t}_{i}\right)\ $ to zero and perform a gradient-based iteration
$\begin{eqnarray}{f}_{k}^{\left(n+1\right)}\left({t}_{i}\right)={f}_{k}^{\left(n\right)}\left({t}_{i}\right)+\epsilon \displaystyle \frac{\partial {F}_{Q}\left(T\right)}{\partial {f}_{k}^{\left(n\right)}\left({t}_{i}\right)},\end{eqnarray}$
where ${f}_{k}^{\left(n\right)}\left({t}_{i}\right)$ denotes the control function fk at time ti and at the nth iteration, ε is a proper positive number (i.e., the so called learning rate). For each step in equation (22), $\partial {F}_{Q}\left(T\right)/\partial {f}_{k}^{\left(n\right)}\left({t}_{i}\right)$ should be recalculated by equation (21). If ${F}_{Q}\left(T\right)$ converges after a sufficient number of iteration steps, we terminate the iteration. Concurrently, the optimal control is obtained.

4. Example: frequency estimation of a spin with optimal two-axis control

As an example, we now consider a 1/2-spin dissipated in a bosonic environment with the spin Hamiltonian
$\begin{eqnarray}{\hat{H}}_{{\rm{s}}}=\displaystyle \frac{{\omega }_{0}}{2}{\hat{\sigma }}_{z},\end{eqnarray}$
where ω0 is the frequency to be estimated and ${\hat{\sigma }}_{z}$ is the Pauli-z operator. The total Hamiltonian and interaction Hamiltonian are given by equation (1) and equation (3) with $\hat{S}={\hat{\sigma }}_{z}/2$. Given the environmental temperature T and the spectral density function J(ω), the COF can be calculated by equation (5). If we take the initial state $\left(\left|\uparrow \right\rangle +\left|\downarrow \right\rangle \right)/\sqrt{2}$, the exact solution of the reduced density matrix is given by [35]
$\begin{eqnarray}\begin{array}{rcl}\rho (t) & = & \displaystyle \frac{\left|\uparrow \right\rangle \left\langle \,\uparrow \,\right|+\left|\downarrow \right\rangle \left\langle \,\downarrow \,\right|}{2}\\ & + & \displaystyle \frac{{{\rm{e}}}^{-\gamma \left(t\right)}}{2}\left({{\rm{e}}}^{-{\rm{i}}{\omega }_{0}t}\left|\uparrow \right\rangle \left\langle \,\downarrow \,\right|+{{\rm{e}}}^{{\rm{i}}{\omega }_{0}t}\left|\downarrow \right\rangle \left\langle \,\uparrow \,\right|\right),\end{array}\end{eqnarray}$
where
$\begin{eqnarray}\gamma \left(t\right)=\int {\rm{d}}\omega J(\omega )\coth (\displaystyle \frac{\beta \omega }{2})\displaystyle \frac{1-\cos (\omega t)}{{\omega }^{2}}.\end{eqnarray}$
The estimation precision of the frequency ω0 is bounded by $1/\sqrt{{F}_{Q}}$. Using equations (15) and (16), one can calculate the QFI [12]
$\begin{eqnarray}{F}_{Q}={t}^{2}{{\rm{e}}}^{-2\gamma \left(t\right)}.\end{eqnarray}$
To improve the estimation precision of ω0, we consider a two-axis control strategy with a controlled Hamiltonian
$\begin{eqnarray}\hat{H}\left(t\right)=\hat{H}+{f}_{x}\left(t\right){\hat{\sigma }}_{x}+{f}_{y}\left(t\right){\hat{\sigma }}_{y}.\end{eqnarray}$
The dynamical evolution is described by equation (18) with
$\begin{eqnarray}\hat{H}={\hat{H}}_{s}+\displaystyle \sum _{j=1}^{M}{{\rm{\Omega }}}_{j}{\hat{c}}_{j}^{\dagger }{\hat{c}}_{j}+\displaystyle \frac{{\hat{\sigma }}_{z}}{2}\otimes \displaystyle \sum _{j=1}^{M}\left({\eta }_{j}{\hat{c}}_{j}^{\dagger }+{\eta }_{j}^{* }{\hat{c}}_{j}\right),\end{eqnarray}$
where the values of Ωj, μj, ηj depend on the specific environmental spectrum. In the subsequent discussion, we will examine two types of environmental spectra and demonstrate the optimal control that maximizes the QFI.

4.1. Lorentz spectral environment

For a Lorentz spectral environment, the spectral density function is given by
$\begin{eqnarray}J(\omega )=\displaystyle \frac{{\chi }^{2}}{\pi }\displaystyle \frac{{\rm{\Gamma }}}{{\left(\omega -{\omega }_{c}\right)}^{2}+{{\rm{\Gamma }}}^{2}},\end{eqnarray}$
where χ is the coupling strength and Γ is the spectral width. At zero temperature, we have
$\begin{eqnarray}\gamma \left(t\right)={\int }_{-\infty }^{\infty }{\rm{d}}\omega J(\omega )\displaystyle \frac{1-\cos (\omega t)}{{\omega }^{2}}.\end{eqnarray}$
For the non-control case, the QFI can be calculated by equations (26) and (30), which is shown in figure 2(a). The red dashed line is the numerical result, obtained by our pseudomode method (equations (9) and (15)). One can find that the numerical result coincides with the analytical one, which verifies the validity of our pseudomode method.
Figure 2. The QFI and optimal control under a Lorentz spectral environment. (a) The comparison of the QFI between the non-control case (blue line) and the optimal control case (green triangles). The blue line is the analytical result of the QFI given by equation (26). The red dashed line is obtained by numerical method. (c) The QFI at a fixed time T = 17/Γ exhibits an increase as the number of iteration steps. (b) and (d) The optimal control fx(t) and fy(t) under Lorentz spectral environment. Environment parameters: Γ = 0.17ω0, ωc = 0.6ω0 and χ = 0.5ω0.
For the control case, we need to exactly calculate the dynamical evolution under the Lorentz spectral environment. We utilize the pseudomode model depicted in figure 1(b), which comprises only one pseudomode harmonic oscillator with a frequency Ω1 = ωc, a dissipation rate ${\mu }_{1}=\sqrt{2{\rm{\Gamma }}}$ and a coupling strength η1 = χ. To obtain the optimal control, we use the gradient ascent algorithm shown in equations (21) and (22). As shown in figure 2(c), the QFI gradually converges after 40 iterations. The optimal control functions ${f}_{x}\left(t\right)$ and ${f}_{y}\left(t\right)$ are shown in figures 2(b) and (d). The black lines are the initial values of ${f}_{x}\left(t\right)$ and ${f}_{y}\left(t\right)$, which are taken as zero. Using optimal ${f}_{x}\left(t\right)$ and ${f}_{y}\left(t\right)$, we show the comparison of controlled QFI and non-controlled QFI in figure 2(a), which indicates a significant improvement in QFI.
At the long-time limit, we observe that the QFI no longer exhibits any increase. This is because the system evolves into a steady state under the influence of control. In figures 3(a)–(c), we plot the expectation values of ${\hat{\sigma }}_{x}$, ${\hat{\sigma }}_{y}$ and ${\hat{\sigma }}_{z}$, which indicate that the off-diagonal elements of the density matrix are gradually decaying to zero. The presence of the control field not only decelerates the decay of off-diagonal elements, but also encodes the information about ω0 into the diagonal elements. When the steady state is reached, the information about ω0 is only encoded in diagonal elements, which explains why the QFI remains constant. Even if the control is removed, the QFI remains unchanged (see figure 3(d)).
Figure 3. (a)-(c) The expectation values of ${\hat{\sigma }}_{x}$, ${\hat{\sigma }}_{y}$, ${\hat{\sigma }}_{z}$ for both the non-control and optimal control cases. (d) The comparison of QFI between non-control and optimal control protocol, with the control field being applied within the time region where Γt < 17. The control field is removed when Γt ≥ 17, as indicated by the pink region.

4.2. Ohmic spectral environment

The Ohmic spectral environment is widely considered for electron transfer dynamics [42] or biomolecular complexes [43], as well as Josephson junctions qubits [44]. The spectral density function is given by $J(\omega )=\chi \omega {{\rm{e}}}^{-\omega /{\omega }_{c}}$ (ω > 0), where χ is the coupling strength and ωc is the cut-off frequency. At zero temperature, we have
$\begin{eqnarray}\gamma \left(t\right)=\displaystyle \frac{\chi }{2}\mathrm{ln}\ \left(1+{\omega }_{c}^{2}{t}^{2}\right).\end{eqnarray}$
Under the short time condition, we have $\gamma \left(t\right)\approx \tfrac{\chi }{2}{\omega }_{c}^{2}{t}^{2}\propto {t}^{2}$, which indicates a quantum Zeno effect [12]. The QFI of non-control case is given by
$\begin{eqnarray}{F}_{Q}={t}^{2}{{\rm{e}}}^{-2\gamma \left(t\right)}=\displaystyle \frac{{t}^{2}}{{\left(1+{\omega }_{c}^{2}{t}^{2}\right)}^{\chi }}.\end{eqnarray}$
To find the optimal control, we need to exactly simulate the dynamical evolution under Ohmic spectral environment. Different to the Lorentz spectral case, we use the pseudomode model shown in figure 1(b) containing 3 pseudomode harmonic oscillators with the frequencies (Ω1, Ω2, Ω3), the dissipation rates (μ1,μ2, μ3 ) and interaction strengths (η1, η2, η3). These pseudomode parameters can be determined by the procedures shown in appendix A. In figure 4, we show the comparison of COF between the original model (solid lines) and the pseudomode model (dashed lines), given by equation (5) and equation (A2). The coincidence of COFs indicates that the two models are nearly equivalent. If one expects a higher accuracy, it is possible to increase the number of pseudomode harmonic oscillators, albeit at the cost of increasing the computational cost.
Figure 4. The real part and the imaginary part of the COF for the original Ohmic spectral model and the pseudomode model. We take the coupling strength χ = 3 and the cut-off frequency ωc = 0.25. The pseudomode parameters: (Ω123) = (0.241, 1.267 31, −0.203), (μ1,μ2,μ3) = (0.5648, 1.1465, 0.8938) and (η1, η2, η3) = (0.0289+0.3711i, −0.0177 − 0.2272i, 0.002 28+0.02931i).
As demonstrated in figure 5(a), the effectiveness of the pseudomode method can be confirmed by the agreement between the QFI values in the non-control scenario. Using the gradient ascent algorithm, we derive the optimal control strategy under an Ohmic environment. The iterative process is shown in figure 5(c), and the optimal control functions ${f}_{x}\left(t\right)$ and ${f}_{y}\left(t\right)$ are shown in figures 5(b) and (d). The comparison of the controlled QFI and the non-controlled QFI are shown in figure 5(a), which indicates a significant improvement in QFI.
Figure 5. The QFI and optimal control under the Ohmic spectral environment. (a) The comparison of QFI between non-control case (blue line) and optimal control case (green triangles). The blue line is the analytical result of QFI given by equation (32). The red dashed line is obtained by numerical method. (c) The iterative process of QFI at time T = 3.9/ωc. (b) and (d) The optimal control fx(t) and fy(t) under Ohmic spectral environment. Environmental parameters: ωc = 0.25ω0 and χ = 0.5.

5. Conclusion

In summary, we have used a pseudomode model method to find the optimal control to improve the performance of the quantum metrology under an arbitrary non-Markovian bosonic environment. The original model is replaced by a pseudomode model which contains few harmonic oscillators with proper parameters dissipated in one Markovian environment. The time evolution is equivalently described by a time-dependent Lindblad master equation. By exactly solving the dynamics, we calculate the gradient of the QFI and use the gradient ascent algorithm to find the optimal control.
To demonstrate its validity, we examine the frequency estimation of a spin system with pure dephasing under two distinct non-Markovian environments. To simulate the evolution of the spin, we introduce one pseudomode harmonic oscillator for the case of a Lorentz spectral environment and three pseudomode harmonic oscillators for the case of an Ohmic spectral environment. For the non-control case, numerical results of the QFI align with analytical findings. By maximizing the QFI at a fixed evolution time, we numerically obtain the optimal multi-axis control, which yields a significant enhancement in the QFI compared to the non-control scenario.

Appendix A. Determining the pseudomode parameters

For the pseudomode model shown in figure 1(b), the COF is given by

$\begin{eqnarray}{C}_{\mathrm{pse}}\left(t\right)\equiv \langle {\hat{B}}_{\mathrm{pse}}(t){\hat{B}}_{\mathrm{pse}}\rangle ,\end{eqnarray}$
with ${\hat{B}}_{\mathrm{pse}}(t)\equiv {{\rm{e}}}^{{\rm{i}}{\hat{H}}_{{\rm{B}},\mathrm{pse}}t}\hat{B}{{\rm{e}}}^{-{\rm{i}}{\hat{H}}_{{\rm{B}},\mathrm{pse}}t}$, where ${\hat{B}}_{\mathrm{pse}}$ and ${\hat{H}}_{{\rm{B}},\mathrm{pse}}$ are given by equations (7) and (8), respectively. Considering the pseudomode oscillators are initially in vacuum states, we can exactly calculate ${C}_{\mathrm{pse}}\left(t\right)$ by the Heisenberg motion equation, i.e.,
$\begin{eqnarray}{C}_{\mathrm{pse}}\left(t\right)=\displaystyle \sum _{j=1}^{M}{\lambda }_{j}{{\rm{e}}}^{{\gamma }_{j}t},\end{eqnarray}$
where γj is determined by the equation group [40]
$\begin{eqnarray}\displaystyle \sum _{l=1}^{M}{P}_{n,j}{\left|{\mu }_{l}\right|}^{2}=-2,\end{eqnarray}$
λj is given by
$\begin{eqnarray}{\lambda }_{j}=\displaystyle \frac{{\sum }_{k=1}^{M}{\sum }_{l=1}^{M}{\eta }_{k}{\eta }_{l}^{* }{P}_{k,j}{P}_{l,j}{\mu }_{k}{\mu }_{l}^{* }}{{\sum }_{n=1}^{M}{P}_{n,j}^{2}{\left|{\mu }_{n}\right|}^{2}},\end{eqnarray}$
and we have defined
$\begin{eqnarray}{P}_{n,j}:= \displaystyle \frac{1}{{\rm{i}}{{\rm{\Omega }}}_{n}+{\gamma }_{j}}.\end{eqnarray}$

Given an original open quantum system model, one can calculate $C\left(t\right)$ by equation (5). To equivalently simulate the original model, we need to find the pseudomode parameters (Ωj, μj, ηj) making ${C}_{\mathrm{pse}}\left(t\right)\approx C\left(t\right)$, which can be done by the following steps.

Step 1: Firstly, given a $C\left(t\right)$, we use Prony's method [40, 45] to decompose $C(t)\approx {\sum }_{j=1}^{M}{\tilde{\lambda }}_{j}{{\rm{e}}}^{{\tilde{\gamma }}_{j}t}$ with all Re$[{\tilde{\gamma }}_{j}]\lt 0$ and ${\sum }_{j=1}^{M}{\tilde{\lambda }}_{j}=C\left(0\right)$. Here, ${\tilde{\gamma }}_{j}$ and ${\tilde{\lambda }}_{j}$ are usually complex numbers.

Step 2: Secondly, solving the equation group

$\begin{eqnarray}\displaystyle \sum _{l=1}^{M}\displaystyle \frac{-\tfrac{1}{2}{\left|{\mu }_{l}\right|}^{2}}{{\rm{i}}{{\rm{\Omega }}}_{l}+{\tilde{\gamma }}_{j}}=1,\ \mathrm{with}\ j=1,2,\cdots ,M,\end{eqnarray}$
we can easily determine one choice of Ωj, μj. Note that Ωj is real number (positive or negative), but μj can be complex number. For convenience, we can take a positive μj.

Step 3: Next, replace γj by ${\tilde{\gamma }}_{j}$ in equation (A4) and hence directly calculate ${\lambda }_{j}\left({\eta }_{1},{\eta }_{2},\cdots ,{\eta }_{M}\right)$. Finally, solving the equations group ${\lambda }_{j}\left({\eta }_{1},{\eta }_{2},\cdots ,{\eta }_{M}\right)={\tilde{\lambda }}_{j}$ with j = 1, 2, ⋯ ,M, which gives the optimal values of η1, η2, ⋯ ,ηM. Here, those ηj are complex numbers.

After determining those pseudomode parameters, the original model can be described by equation (9). Tracing out the pseudomode oscillators, we obtain the reduced density matrix of the system, i.e., $\rho ={\mathrm{Tr}}_{c}\left({\rho }_{\mathrm{sc}}\right)$.

Appendix B. Gradient of QFI

Given a fixed evolution time T, the QFI is

$\begin{eqnarray}{F}_{Q}\left(T\right)={\mathrm{Tr}}_{{\rm{s}}}\left[\rho \left(T\right){l}^{2}\left(T\right)\right],\end{eqnarray}$
where $l\left(T\right)$ is determined by
$\begin{eqnarray}\displaystyle \frac{\partial \rho \left(T\right)}{\partial \theta }=\displaystyle \frac{\rho \left(T\right)l\left(T\right)+l\left(T\right)\rho \left(T\right)}{2},\end{eqnarray}$
and θ is the parameter to be estimated. In the following, we always use the simplified symbol $l\equiv l\left(T\right)$. Next, we will calculate the gradient of the QFI with respect to an arbitrary variable f.

Calculating the partial derivative of equation (B1) and (B2), we obtain

$\begin{eqnarray}\begin{array}{rcl}\displaystyle \frac{\partial {F}_{Q}(T)}{\partial f} & = & {\mathrm{Tr}}_{{\rm{s}}}\left[\displaystyle \frac{\partial \rho (T)}{\partial f}{l}^{2}+\rho (T)\displaystyle \frac{\partial l}{\partial f}l+\rho (T)l\displaystyle \frac{\partial l}{\partial f}\right]\\ & = & {\mathrm{Tr}}_{{\rm{s}}}\left[\displaystyle \frac{\partial \rho (T)}{\partial f}{l}^{2}+\left(\rho (T)\displaystyle \frac{\partial l}{\partial f}+\displaystyle \frac{\partial l}{\partial f}\rho (T)\right)l\right].\end{array}\end{eqnarray}$
and
$\begin{eqnarray}\rho (T)\displaystyle \frac{\partial l}{\partial f}+\displaystyle \frac{\partial l}{\partial f}\rho (T)=2\displaystyle \frac{{\partial }^{2}\rho (T)}{\partial f\partial \theta }-\displaystyle \frac{\partial \rho (T)}{\partial f}l-l\displaystyle \frac{\partial \rho (T)}{\partial f}.\end{eqnarray}$
Substituting equation (B4) into equation (B3), we have
$\begin{eqnarray}\begin{array}{rcl}\displaystyle \frac{\partial {F}_{Q}(T)}{\partial f} & = & {\mathrm{Tr}}_{{\rm{s}}}\left[\displaystyle \frac{\partial \rho (T)}{\partial f}{l}^{2}+\left(2\displaystyle \frac{{\partial }^{2}\rho (T)}{\partial f\partial \theta }\right.\right.\\ & & \left.\left.-\displaystyle \frac{\partial \rho (T)}{\partial f}l-l\displaystyle \frac{\partial \rho (T)}{\partial f}\right)l\right]\\ & = & {\mathrm{Tr}}_{{\rm{s}}}\left[2l\displaystyle \frac{{\partial }^{2}\rho (T)}{\partial f\partial \theta }-{l}^{2}\displaystyle \frac{\partial \rho (T)}{\partial f}\right].\end{array}\end{eqnarray}$
Using the relation $\rho \left(T\right)={\mathrm{Tr}}_{{\rm{c}}}\left[{\rho }_{\mathrm{sc}}\left(T\right)\right]$, we have
$\begin{eqnarray}\displaystyle \frac{\partial {F}_{Q}(T)}{\partial f}={\mathrm{Tr}}_{{\rm{s}}}\left[2l{\mathrm{Tr}}_{{\rm{c}}}\displaystyle \frac{{\partial }^{2}{\rho }_{\mathrm{sc}}(T)}{\partial f\partial \theta }-{l}^{2}{\mathrm{Tr}}_{{\rm{c}}}\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}(T)}{\partial f}\right].\end{eqnarray}$
Taking $f={f}_{k}\left({t}_{i}\right)$, we obtain equation (21) in the main text. Here, ${\partial }^{2}{\rho }_{\mathrm{sc}}\left(T\right)/\partial {f}_{k}\left({t}_{i}\right)\partial \theta $ and $\partial {\rho }_{\mathrm{sc}}\left(T\right)/\partial {f}_{k}\left({t}_{i}\right)$ can be calculated by iterations from j = 1 to N + 1 (note that T = tN+1). Specifically, taking the partial derivative of equation (19) with respect to ${f}_{k}\left({t}_{i}\right)$, we obtain the iteration
$\begin{eqnarray}\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}({t}_{j+1})}{\partial {f}_{k}\left({t}_{i}\right)}=\displaystyle \frac{\partial {{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}}{\partial {f}_{k}\left({t}_{i}\right)}{\rho }_{\mathrm{sc}}({t}_{j})+{\rm{\Theta }}\left(j-i\right){{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}({t}_{j})}{\partial {f}_{k}\left({t}_{i}\right)}.\end{eqnarray}$
Taking the partial derivative of equation (B7) with respect to θ, we obtain
$\begin{eqnarray}\begin{array}{rcl}\displaystyle \frac{{\partial }^{2}{\rho }_{\mathrm{sc}}({t}_{j+1})}{\partial {f}_{k}\left({t}_{i}\right)\partial \theta } & = & \displaystyle \frac{{\partial }^{2}{{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}}{\partial {f}_{k}\left({t}_{i}\right)\partial \theta }+\displaystyle \frac{\partial {{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}}{\partial {f}_{k}\left({t}_{i}\right)}\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}({t}_{j})}{\partial \theta }\\ & & +{\rm{\Theta }}\left(j-i\right)\left(\displaystyle \frac{\partial {{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}}{\partial \theta }\displaystyle \frac{\partial {\rho }_{\mathrm{sc}}({t}_{j})}{\partial {f}_{k}\left({t}_{i}\right)}\right.\\ & & \left.+{{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}\displaystyle \frac{{\partial }^{2}{\rho }_{\mathrm{sc}}({t}_{j})}{\partial {f}_{k}\left({t}_{i}\right)\partial \theta }\right),\end{array}\end{eqnarray}$
where ${\rm{\Theta }}\left(\cdots \right)$ is the Heaviside function. The iterations of equations (19, 20, B7, B8) should be executed simultaneously.

When we execute the iterations, the partial derivatives of exponential ${{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}$ in equations (B7, B8) can be approximated. Considering Δt is very small, we expand ${{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}$ as

$\begin{eqnarray}{{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}\approx 1+{{ \mathcal L }}_{j}{\rm{\Delta }}t+\displaystyle \frac{1}{2}{{ \mathcal L }}_{j}^{2}{\rm{\Delta }}{t}^{2}.\end{eqnarray}$
Therefore, we can obtain the one order partial derivatives
$\begin{eqnarray}\displaystyle \frac{\partial {{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}}{\partial {f}_{k}\left({t}_{i}\right)}\approx -{\rm{i}}{\delta }_{i,j}\left[{\rm{\Delta }}t{\hat{H}}_{k}^{\times }+\displaystyle \frac{{\rm{\Delta }}{t}^{2}}{2}\left({{ \mathcal L }}_{j}{\hat{H}}_{k}^{\times }+{\hat{H}}_{k}^{\times }{{ \mathcal L }}_{j}\right)\right],\end{eqnarray}$
$\begin{eqnarray}\displaystyle \frac{\partial {{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}}{\partial \theta }\approx \displaystyle \frac{\partial {{ \mathcal L }}_{j}}{\partial \theta }{\rm{\Delta }}t+\displaystyle \frac{{\rm{\Delta }}{t}^{2}}{2}\left({{ \mathcal L }}_{j}\displaystyle \frac{\partial {{ \mathcal L }}_{j}}{\partial \theta }+\displaystyle \frac{\partial {{ \mathcal L }}_{j}}{\partial \theta }{{ \mathcal L }}_{j}\right),\end{eqnarray}$
and the mixed partial derivative
$\begin{eqnarray}\displaystyle \frac{{\partial }^{2}{{\rm{e}}}^{{{ \mathcal L }}_{j}{\rm{\Delta }}t}}{\partial {f}_{k}\left({t}_{i}\right)\partial \theta }\approx -{\delta }_{i,j}\displaystyle \frac{{\rm{\Delta }}{t}^{2}}{2}\left(\displaystyle \frac{\partial {\hat{H}}_{{\rm{s}}}^{\times }}{\partial \theta }{\hat{H}}_{k}^{\times }+{\hat{H}}_{k}^{\times }\displaystyle \frac{\partial {\hat{H}}_{{\rm{s}}}^{\times }}{\partial \theta }\right).\end{eqnarray}$
Here, we have defined the superoperator ${\hat{H}}^{\times }\rho \equiv \hat{H}\rho -\rho \hat{H}$ and used the fact of $\partial {{ \mathcal L }}_{j}/\partial {f}_{k}\left({t}_{i}\right)=-{\rm{i}}{\delta }_{i,j}{\hat{H}}_{k}^{\times }$ and $\partial {{ \mathcal L }}_{j}/\partial \theta =-{\rm{i}}\partial {\hat{H}}_{{\rm{s}}}^{\times }/\partial \theta $. Specially, for the example of frequency estimation in main text, we have ${\hat{H}}_{{\rm{s}}}=\tfrac{{\omega }_{0}}{2}{\hat{\sigma }}_{z}$ and $\partial {\hat{H}}_{{\rm{s}}}^{\times }/\partial \theta ={\hat{\sigma }}_{z}^{\times }/2$.

We now make a summary of calculating the gradient of QFI. Firstly, we execute the iterations of equations (19, 20, B7, B8) simultaneously, from j = 1 to N + 1, which gives the value of ${\partial }^{2}{\rho }_{\mathrm{sc}}(T)/\partial {f}_{k}\left({t}_{i}\right)\partial \theta $ and $\partial {\rho }_{\mathrm{sc}}(T)/\partial {f}_{k}\left({t}_{i}\right)$. Next, the gradient $\partial {F}_{Q}(T)/\partial {f}_{k}\left({t}_{i}\right)$ can be directly obtained by equation (B6).

This work was supported by the National Natural Science Foundation of China (Grant No. 12274019), and the NSAF Joint Fund (Grant No. U2230402). We acknowledge the computational support from the Beijing Computational Science Research Center.

1
T. L. S. Collaboration 2011 A gravitational wave observatory operating beyond the quantum shot-noise limit Nat. Phys. 7 962

DOI

2
Taylor M A, Bowen W P 2016 Quantum metrology and its application in biology Phys. Rep. 615 1

DOI

3
Ludlow A D, Boyd M M, Ye J, Peik E, Schmidt P O 2015 Optical atomic clocks Rev. Mod. Phys. 87 637

DOI

4
Jones J A, Karlen S D, Fitzsimons J, Ardavan A, Benjamin S C, Briggs G A D, Morton J J L 2009 Magnetic field sensing beyond the standard quantum limit using 10-spin noon states Science 324 1166

DOI

5
Len Y L, Gefen T, Retzker A, Kołodyński J 2022 Quantum metrology with imperfect measurements Nat. Commun. 13 6971

DOI

6
Caves C M 1981 Quantum-mechanical noise in an interferometer Phys. Rev. D 23 1693

DOI

7
Giovannetti V, Lloyd S, Maccone L 2004 Quantum-enhanced measurements: beating the standard quantum limit Science 306 1330

DOI

8
Leibfried D, Barrett M D, Schaetz T, Britton J, Chiaverini J, Itano W M, Jost J D, Langer C, Wineland D J 2004 Toward Heisenberg-limited spectroscopy with multiparticle entangled states Science 304 1476

DOI

9
Giovannetti V, Lloyd S, Maccone L 2006 Quantum metrology Phys. Rev. Lett. 96 010401

DOI

10
Huelga S F, Macchiavello C, Pellizzari T, Ekert A K, Plenio M B, Cirac J I 1997 Improvement of frequency standards with quantum entanglement Phys. Rev. Lett. 79 3865

DOI

11
Chaves R, Brask J B, Markiewicz M, Kołodyński J, Acín A 2013 Noisy metrology beyond the standard quantum limit Phys. Rev. Lett. 111 120401

DOI

12
Chin A W, Huelga S F, Plenio M B 2012 Quantum metrology in non-Markovian environments Phys. Rev. Lett. 109 233601

DOI

13
Long X 2022 Entanglement-enhanced quantum metrology in colored noise by quantum zeno effect Phys. Rev. Lett. 129 070502

DOI

14
Smirne A, Kołodyński J, Huelga S F, Demkowicz-Dobrzański R 2016 Ultimate precision limits for noisy frequency estimation Phys. Rev. Lett. 116 120801

DOI

15
Wang Y-S, Chen C, An J-H 2017 Quantum metrology in local dissipative environments New J. Phys. 19 113019

DOI

16
Bai K, Peng Z, Luo H-G, An J-H 2019 Retrieving ideal precision in noisy quantum optical metrology Phys. Rev. Lett. 123 040402

DOI

17
Bai S-Y, An J-H 2023 Floquet engineering to overcome no-go theorem of noisy quantum metrology Phys. Rev. Lett. 131 050801

DOI

18
Tan Q-S, Wu W, Xu L, Liu J, Kuang L-M 2022 Quantum sensing of supersensitivity for the Ohmic quantum reservoir Phys. Rev. A 106 032602

DOI

19
Tan Q-S, Huang Y, Yin X, Kuang L-M, Wang X 2013 Enhancement of parameter-estimation precision in noisy systems by dynamical decoupling pulses Phys. Rev. A 87 032102

DOI

20
Lang J E, Liu R B, Monteiro T S 2015 Dynamical-decoupling-based quantum sensing: Floquet spectroscopy Phys. Rev. X 5 041016

DOI

21
Sekatski P, Skotiniotis M, Dür W 2016 Dynamical decoupling leads to improved scaling in noisy quantum metrology New J. Phys. 18 073034

DOI

22
Zhai Y, Yang X, Tang K, Long X, Nie X, Xin T, Lu D, Li J 2023 Control-enhanced quantum metrology under Markovian noise Phys. Rev. A 107 022602

DOI

23
Liu J, Yuan H 2017 Quantum parameter estimation with optimal control Phys. Rev. A 96 012117

DOI

24
Liu J, Yuan H 2017 Control-enhanced multiparameter quantum estimation Phys. Rev. A 96 042114

DOI

25
Liu J, Zhang M, Chen H, Wang L, Yuan H 2022 Optimal scheme for quantum metrology Advanced Quantum Technologies 5 2100080

DOI

26
Boscain U, Sigalotti M, Sugny D 2021 Introduction to the Pontryagin maximum principle for quantum optimal control PRX Quantum 2 030203

DOI

27
Lin C, Ma Y, Sels D 2022 Application of Pontryagin's maximum principle to quantum metrology in dissipative systems Phys. Rev. A 105 042621

DOI

28
Ansel Q, Dionis E, Sugny D 2024 Optimal control strategies for parameter estimation of quantum systems SciPost Phys. 16 013

DOI

29
Gorini V, Kossakowski A, Sudarshan E C G 1976 Completely positive dynamical semigroups of n-level systems J. Math. Phys. 17 821

DOI

30
Lindblad G 1976 On the generators of quantum dynamical semigroups Commun. Math. Phys. 48 119

DOI

31
Gröblacher S, Trubarov A, Prigge N, Cole G D, Aspelmeyer M, Eisert J 2015 Observation of non-Markovian micromechanical brownian motion Nat. Commun. 6 7606

DOI

32
de Vega I, Porras D, Ignacio Cirac J 2008 Matter-wave emission in optical lattices: single particle and collective effects Phys. Rev. Lett. 101 260404

DOI

33
Weiss U 2012 Quantum Dissipative Systems World Scientific

34
Lambert N, Chen Y-N, Cheng Y-C, Li C-M, Chen G-Y, Nori F 2013 Quantum biology Nat. Phys. 9 10

DOI

35
Breuer H-P, Petruccione F 2002 The Theory of Open Quantum Systems Oxford University Press

36
Tanimura Y, Kubo R 1989 Time evolution of a quantum system in contact with a nearly Gaussian–Markoffian noise bath J. Phys. Soc. Jpn. 58 101

DOI

37
Tanimura Y 1990 Nonperturbative expansion method for a quantum system coupled to a harmonic-oscillator bath Phys. Rev. A 41 6676

DOI

38
Tanaka M, Tanimura Y 2009 Quantum dissipative dynamics of electron transfer reaction system: Nonperturbative hierarchy equations approach J. Phys. Soc. Jpn. 78 073802

DOI

39
Luo S, Lambert N, Liang P, Cirio M 2023 Quantum-classical decomposition of Gaussian quantum environments: A stochastic pseudomode modei PRX Quantum 4 030316

DOI

40
Zhou L K, Jin G R, Yang W 2024 Systematic and efficient pseudomode method to simulate open quantum systems under a bosonic environment Phys. Rev. A 110 022221

DOI

41
Havel T F 2003 Robust procedures for converting among Lindblad, Kraus and matrix representations of quantum dynamical semigroups J. Math. Phys. 44 534

DOI

42
Egger R, Mak C H 1994 Low-temperature dynamical simulation of spin-boson systems Phys. Rev. B 50 15210

DOI

43
Gilmore J B, McKenzie R H 2006 Criteria for quantum coherent transfer of excitations between chromophores in a polar solvent Chem. Phys. Lett. 421 266

DOI

44
Makhlin Y, Schön G, Shnirman A 2001 Quantum-state engineering with Josephson-junction devices Rev. Mod. Phys. 73 357

DOI

45
Beylkin G, Monzón L 2005 On approximation of functions by exponential sums Appl. Comput. Harmon. Anal. 19 17

DOI

Outlines

/