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

Thermodynamic uncertainty relation for quantum first passage process

Yoshihiko Hasegawa [email protected] Department of Information and Communication Engineering, Graduate School of Information Science and Technology, The University of Tokyo, Tokyo 113-8656, Japan
Abstract

We derive a thermodynamic uncertainty relation for first passage processes in quantum Markov chains. We consider first passage processes that stop after a fixed number of jumps, which contrasts with typical quantum Markov chains which end at a fixed time. We obtain bounds for the observables of the first passage processes in quantum Markov chains by the Loschmidt echo, which quantifies the extent of irreversibility in quantum many-body systems. Considering a particular case, we show that the lower bound corresponds to the quantum Fisher information, which plays a fundamental role in uncertainty relations in quantum systems. Moreover, considering classical dynamics, our bound reduces to a thermodynamic uncertainty relation for classical first passage processes.

I Introduction

The thermodynamic uncertainty relation (TUR) provides a fundamental limit of thermodynamic machines. It states that the precision of a machine, which is quantified by fluctuations of thermodynamic currents, is bounded from below by the thermodynamic cost, such as entropy production and dynamical activity. TURs have been derived for classical systems [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17] and quantum systems [18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28]. Recently, TURs have become a central topic in nonequilibrium thermodynamics. Besides their theoretical significance, TURs have practical advantages and are used in estimating the entropy production of thermodynamic machines solely from their stochastic trajectories [29, 30, 31, 32].

The first passage process is a stochastic process and has been extensively studied in various fields ranging from physics to finance [33, 34]. Recently, it is attracting much attention in quantum dynamics as well [35, 36]. Conventionally, we consider the dynamics of stochastic processes that starts at 0 and ends at a fixed time τ\tau (τ>0\tau>0). However, in first passage processes, the dynamics stops when the system satisfies some predefined conditions, and therefore the end time of the dynamics, known as the first passage time, is considered a random variable. The predefined conditions could be the system reaching some practically meaningful states, e.g., absorbing states such as extinction, or the number of jumps surpassing a threshold. Recently, first passage processes have become increasingly important in stochastic thermodynamics [37, 38, 39]. For instance, it is possible to extract work by a gambling demon, which monitors its state and determines when to stop the dynamics [40]. For TURs in first passage processes, relations for classical stochastic thermodynamics were derived in Refs. [6, 41, 42]. These relations show that the fluctuations of the first passage time are bounded from below by dynamical activity or entropy production. TURs for first passage processes are particularly important in biochemical clocks [1, 43]. The precision of such clocks is ideally evaluated through the first passage time for them to accomplish a single chemical reaction cycle. Although various studies on the first passage problem have been conducted in classical stochastic thermodynamics, studies of quantum variants are in a very early stage. It was recently proposed that quantum clocks can be implemented by quantum heat engines [18, 44], which strongly demands a TUR for the quantum first passage processes.

In this manuscript, we consider a TUR for quantum Markov chains that stop after a fixed number of jumps, which is a particular case of first passage processes. Using the techniques developed in Ref. [45], we obtain a TUR for the first passage time in quantum Markov chains, whose lower bound comprises the Loschmidt echo. The obtained bound concerns two dynamics: the original and perturbed dynamics. When the perturbed dynamics is identical to the original dynamics, except that the time scale of the perturbed dynamics is slightly faster or slower than that of the original dynamics, the Loschmidt echo reduces to the quantum Fisher information. This information plays a fundamental role in several uncertainty relations in quantum systems. Considering the classical limit of the derived TUR, we show that our relation reduces to a classical TUR for the classical first passage processes derived so far. Considering a two-level atom driven by a classical laser field as an application, we show that the fluctuations of the first passage time for a quantum Markov chain become smaller than its classical counterpart.

Refer to caption
Figure 1: Loschmidt echo for (a) closed and (b) open dynamics. SS represents the system and EE represents the environment. (a) Loschmidt echo defined by the fidelity between eiHτ|Ψ(0)e^{-iH\tau}\ket{\Psi(0)} and eiHτ|Ψ(0)e^{-iH_{\star}\tau}\ket{\Psi(0)} at t=τt=\tau, where HH and HH_{\star} are the original and perturbed Hamiltonians, respectively. (b) Loschmidt echo defined by the fidelity between mVm|ψS|em\sum_{m}V_{m}\ket{\psi_{S}}\otimes\ket{e_{m}} and mV,m|ψS|em\sum_{m}V_{\star,m}\ket{\psi_{S}}\otimes\ket{e_{m}}, where VmV_{m} and V,mV_{\star,m} are the Kraus operators of the original and perturbed dynamics, respectively.

II Methods

The Loschmidt echo is an indicator of quantum chaos in many-body systems. Let HH be a Hamiltonian and let HH_{\star} be a perturbed Hamiltonian. When the system is closed, given the initial state |Ψ(0)\ket{\Psi(0)}, the Loschmidt echo η\eta is defined by

η|Ψ(0)|eiHτeiHτ|Ψ(0)|2.\eta\equiv|\braket{\Psi(0)}{e^{iH_{\star}\tau}e^{-iH\tau}}{\Psi(0)}|^{2}. (1)

The Loschmidt echo η\eta defined in Eq. (1) can be regarded as the fidelity between eiHτ|Ψ(0)e^{-iH\tau}\ket{\Psi(0)} and eiHτ|Ψ(0)e^{-iH_{\star}\tau}\ket{\Psi(0)}, which are time-evolved states at τ\tau induced by HH and HH_{\star}, respectively, as shown in Fig. 1(a). If the system is sensitive to the perturbation, the fidelity decays quickly to 0, indicating that the Loschmidt echo is conceptually similar to the Lyapunov exponent. Because the highly susceptible fidelity under a small perturbation is related to the irreversibility of the dynamics, the Loschmidt echo can be used to quantify the irreversibility.

The Loschmidt echo in Eq. (1) is for closed quantum systems. However, we can define the Loschmidt echo for open quantum systems. Let us consider a Kraus representation:

ρSmVmρSVm,\rho_{S}\to\sum_{m}V_{m}\rho_{S}V_{m}^{\dagger}, (2)

where ρS\rho_{S} is a density operator and VmV_{m} is a Kraus operator satisfying a completeness relation mVmVm=𝕀S\sum_{m}V_{m}^{\dagger}V_{m}=\mathbb{I}_{S} with 𝕀S\mathbb{I}_{S} being the identity operator in SS. This time evolution induced by the Kraus representation can be expressed by a unitary time evolution in a larger space comprising the system SS and environment EE:

USE|ψS|e0=mVm|ψS|em,U_{SE}\ket{\psi_{S}}\otimes\ket{e_{0}}=\sum_{m}V_{m}\ket{\psi_{S}}\otimes\ket{e_{m}}, (3)

where |em\ket{e_{m}} constitutes an orthonormal basis of EE, |ψS\ket{\psi_{S}} and |e0\ket{e_{0}} are initial states of SS and EE, respectively, and USEU_{SE} is a unitary operator acting on S+ES+E. Tracing out EE in Eq. (3), we recover the original Kraus representation of Eq. (2). Suppose that a Kraus operator for the perturbed dynamics is V,mV_{\star,m}. Then the fidelity can be calculated between mVm|ψS|em\sum_{m}V_{m}\ket{\psi_{S}}\otimes\ket{e_{m}} and mV,m|ψS|em\sum_{m}V_{\star,m}\ket{\psi_{S}}\otimes\ket{e_{m}}, as shown in Fig. 1(b). Throughout this manuscript, we use \star in subscripts to express the perturbed dynamics.

We can derive an uncertainty relation from the Loschmidt echo. Using the lower bound for the Hellinger distance [46], we obtain the bounds for observables and the Loschmidt echo [45]. Let |Ψ\ket{\Psi} and |Ψ\ket{\Psi_{\star}} be two pure states and let \mathcal{F} be an Hermitian observable that is applied to |Ψ\ket{\Psi} or |Ψ\ket{\Psi_{\star}}. We define the mean and standard deviation of \mathcal{F} by Ψ||Ψ\braket{\mathcal{F}}\equiv\braket{\Psi}{\mathcal{F}}{\Psi} and 22\llbracket\mathcal{F}\rrbracket\equiv\sqrt{\braket{\mathcal{F}^{2}}-\braket{\mathcal{F}}^{2}}, respectively. \braket{\mathcal{F}}_{\star} and \llbracket\mathcal{F}\rrbracket_{\star} should be evaluated for |Ψ\ket{\Psi_{\star}} instead of |Ψ\ket{\Psi}. Then the following relation holds [45]:

(+)21η11,\left(\frac{\llbracket\mathcal{F}\rrbracket+\llbracket\mathcal{F}\rrbracket_{\star}}{\braket{\mathcal{F}}-\braket{\mathcal{F}}_{\star}}\right)^{2}\geq\frac{1}{\eta^{-1}-1}, (4)

where η=|Ψ|Ψ|2\eta=|\braket{\Psi_{\star}}{\Psi}|^{2}. Equation (4) is a tighter version of the inequality derived in Ref. [47]. We obtained a quantum TUR for quantum Markov chains that end at a fixed time using Eq.  (4), where the observable is a counting observable and the lower bound is defined by the Loschmidt echo. The results obtained in Ref. [45] can be regarded as a quantum analog of TUR, because the entropy production in stochastic thermodynamics also characterizes the extent of irreversibility due to the time-reversal operation.

Now, we derive a TUR for first passage processes in quantum Markov chains using Eq. (4). In particular, we consider quantum Markov chains that stop after KK jump events, where K{1,2,3,}K\in\{1,2,3,\ldots\}. Let ρS(t)\rho_{S}(t) be a density operator of the principal system SS at time tt. We assume that the dynamics of ρS(t)\rho_{S}(t) is governed by the Lindblad equation:

ρ˙S=ρSi[HS,ρS]+m=1M𝒟(ρS,Lm),\dot{\rho}_{S}=\mathcal{L}\rho_{S}\equiv-i\left[H_{S},\rho_{S}\right]+\sum_{m=1}^{M}\mathcal{D}(\rho_{S},L_{m}), (5)

where \mathcal{L} is the Lindblad super-operator, HSH_{S} is a system Hamiltonian, LmL_{m} represents jump operators, MM is the number of jump operators, and 𝒟\mathcal{D} is the dissipator defined by

𝒟(ρS,L)LρSL{LL,ρS}2.\mathcal{D}(\rho_{S},L)\equiv L\rho_{S}L^{\dagger}-\frac{\left\{L^{\dagger}L,\rho_{S}\right\}}{2}. (6)

Here, {,}\{\bullet,\bullet\} denotes the anti-commutator. The solution of Eq. (5) corresponds to the dynamics when we do not measure the environment EE. However, when we measure the environment EE, the system exhibits stochastic dynamics depending on the measurement record. Such stochastic dynamics conditioned on the measurement record is referred to as a quantum trajectory. Figures 2(a) and (b) show trajectories of classical and quantum Markov chains, respectively. For the classical trajectory, the state only varies via stochastic jumps, whereas the system undergoes smooth continuous evolution and sudden discontinuous jumps for the quantum trajectory. We can represent Eq. (5) by the following alternative expression:

ρ˙S=i(HeffρSρSHeff)+m=1MLmρSLm,\dot{\rho}_{S}=-i(H_{\mathrm{eff}}\rho_{S}-\rho_{S}H_{\mathrm{eff}}^{\dagger})+\sum_{m=1}^{M}L_{m}\rho_{S}L_{m}^{\dagger}, (7)

where HeffH_{\mathrm{eff}} is an effective Hamiltonian:

HeffHSi2m=1MLmLm.H_{\mathrm{eff}}\equiv H_{S}-\frac{i}{2}\sum_{m=1}^{M}L_{m}^{\dagger}L_{m}. (8)

Equation (7) shows that the dynamics of ρS\rho_{S} is governed by two contributions: the effective Hamiltonian term i(HeffρSρSHeff)-i(H_{\mathrm{eff}}\rho_{S}-\rho_{S}H_{\mathrm{eff}}^{\dagger}), which induces a continuous smooth state change, and the jump operator term m=1MLmρSLm\sum_{m=1}^{M}L_{m}\rho_{S}L_{m}^{\dagger}, which causes a discontinuous jump event. Suppose that the state of SS right after a jump event is |ψS\ket{\psi_{S}}. Continuous evolution in the quantum trajectory is given by eiHefft|ψSe^{-iH_{\mathrm{eff}}t}\ket{\psi_{S}}, which is not normalized, and the next jump is induced by the mmth jump operator LmL_{m} after a waiting time ww. Therefore, from the state after the first jump event to the state after the second jump event, the state is transformed as

|ψSY(w,m)|ψSψS|Y(w,m)Y(w,m)|ψS,\ket{\psi_{S}}\to\frac{Y(w,m)\ket{\psi_{S}}}{\sqrt{\braket{\psi_{S}}{Y^{\dagger}(w,m)Y(w,m)}{\psi_{S}}}}, (9)

where Y(w,m)LmeiHeffwY(w,m)\equiv L_{m}e^{-iH_{\mathrm{eff}}w}. Y(w,m)Y(w,m) satisfies a completeness relation m=1M0𝑑wY(w,m)Y(w,m)=𝕀S\sum_{m=1}^{M}\int_{0}^{\infty}dw\,Y^{\dagger}(w,m)Y(w,m)=\mathbb{I}_{S}, which can be shown by

m=1M0𝑑weiwHeffLmLmeiwHeff\displaystyle\sum_{m=1}^{M}\int_{0}^{\infty}dw\,e^{iwH_{\mathrm{eff}}^{\dagger}}L_{m}^{\dagger}L_{m}e^{-iwH_{\mathrm{eff}}}
=0𝑑wddw[eiwHeffeiwHeff]\displaystyle=-\int_{0}^{\infty}dw\,\frac{d}{dw}\left[e^{iwH_{\mathrm{eff}}^{\dagger}}e^{-iwH_{\mathrm{eff}}}\right]
=𝕀SeiwHeffeiwHeff|w\displaystyle=\mathbb{I}_{S}-\left.e^{iwH_{\mathrm{eff}}^{\dagger}}e^{-iwH_{\mathrm{eff}}}\right|_{w\to\infty} (10)

Therefore, the Kraus representation describing the evolution from the first to the second jump events is given by

𝒵()m=1M0𝑑wY(w,m)Y(w,m).\mathcal{Z}(\bullet)\equiv\sum_{m=1}^{M}\int_{0}^{\infty}dw\,Y(w,m)\bullet Y^{\dagger}(w,m). (11)
Refer to caption
Figure 2: Comparison of trajectories in (a) classical and (b) quantum Markov chains. In a quantum Markov chain, between the iith and i+1i+1th jump events, the state can change via eiHeff(tti)e^{-iH_{\mathrm{eff}}(t-t_{i})}, where tit_{i} is time stamp of the iith jump. On the other hand, the state remains unchanged between consecutive jump events for a classical Markov chain.

We employ the input-output formalism for the first passage process with a fixed number of jumps as in Ref. [48]. The input-output formalism is also known as the continuous matrix product state. If we replace VmV_{m} in Eq. (2) by Y(w,m)Y(w,m), the Kraus representation of Eq. (11) can be represented as an interaction between the principal system SS and the environment EE as follows:

|Ψ=m=1M0𝑑wY(w,m)|ψS|w,m,\ket{\Psi}=\sum_{m=1}^{M}\int_{0}^{\infty}dw\,Y(w,m)\ket{\psi_{S}}\otimes\ket{w,m}, (12)

where |w,m\ket{w,m} constitutes an orthonormal basis in EE as w,m|w,m=δ(ww)δmm\braket{w^{\prime},m^{\prime}}{w,m}=\delta(w^{\prime}-w)\delta_{m^{\prime}m}.

Refer to caption
Figure 3: Quantum Markov chain that stops after a fixed number of jumps (K=4K=4). (a) and (b) are different realizations of the process, both stopping when they undergo K=4K=4 jump events. mim_{i}, tit_{i}, and wiw_{i} denote the iith output, time stamp of the iith output, and waiting time between the (i1)(i-1)th and iith jump events, respectively.

We consider a quantum Markov chain that stops after KK jump events. Let |ΨK\ket{\Psi_{K}} be a composite state in S+ES+E after KK jump events. Let mim_{i} be the output of the iith jump, tit_{i} be the time stamp of the iith jump (t0=0t_{0}=0), and wiw_{i} be the waiting time between the i1i-1th and iith jump events (i.e., wititi1w_{i}\equiv t_{i}-t_{i-1}). Figures 3(a) and (b) are examples of the process, where the trajectories stop when they undergo K=4K=4 jump events. We observe that the first passage time is a random variable, and it is different for these two cases. Repeating Eq. (12) KK times, |ΨK\ket{\Psi_{K}} is expressed by

|ΨK=mK,,m10𝑑𝒘Y(wK,mK)Y(w1,m1)|ψS|(wK,mK),,(w1,m1),\ket{\Psi_{K}}=\sum_{m_{K},\cdots,m_{1}}\int_{0}^{\infty}d\boldsymbol{w}\,Y(w_{K},m_{K})\cdots Y(w_{1},m_{1})\ket{\psi_{S}}\otimes\ket{(w_{K},m_{K}),\cdots,(w_{1},m_{1})}, (13)

where 0𝑑𝒘\int_{0}^{\infty}d\boldsymbol{w} is an abbreviation for 0𝑑w10𝑑w20𝑑wK\int_{0}^{\infty}dw_{1}\int_{0}^{\infty}dw_{2}\cdots\int_{0}^{\infty}dw_{K}. We used Eq. (5) in Ref. [45] for expressing the input-output state for constant-time quantum Markov chains. Equation (13) shows a constant jump case of Eq. (5) in Ref. [45].

When we perform a continuous measurement, we obtain records of jump events 𝒎[m1,m2,,mK]\boldsymbol{m}\equiv[m_{1},m_{2},\ldots,m_{K}] and their time stamp 𝒕[t1,t2,,tK]\boldsymbol{t}\equiv[t_{1},t_{2},\ldots,t_{K}], as shown in Fig. 3. Because knowing the waiting time 𝒘[w1,w2,,wK]\boldsymbol{w}\equiv[w_{1},w_{2},\ldots,w_{K}] is equivalent to the time stamp 𝒕\boldsymbol{t}, 𝒎\boldsymbol{m}, and 𝒕\boldsymbol{t} can be obtained by applying the projector |𝒘,𝒎𝒘,𝒎|\ket{\boldsymbol{w},\boldsymbol{m}}\bra{\boldsymbol{w},\boldsymbol{m}}, where |𝒘,𝒎|(wK,mK)|(w1,m1)\ket{\boldsymbol{w},\boldsymbol{m}}\equiv\ket{(w_{K},m_{K})}\otimes\cdots\otimes\ket{(w_{1},m_{1})}, to |ΨK\ket{\Psi_{K}} in Eq. (13). We consider the following Hermitian observable on EE.

𝒪𝒎0𝑑𝒘h(𝒘,𝒎)|𝒘,𝒎𝒘,𝒎|,\mathcal{O}\equiv\sum_{\boldsymbol{m}}\int_{0}^{\infty}d\boldsymbol{w}\,h(\boldsymbol{w},\boldsymbol{m})\ket{\boldsymbol{w},\boldsymbol{m}}\bra{\boldsymbol{w},\boldsymbol{m}}, (14)

where h(𝒘,𝒎)h(\boldsymbol{w},\boldsymbol{m}) is an arbitrary real function of 𝒘\boldsymbol{w} and 𝒎\boldsymbol{m}.

The Loschmidt echo considers the fidelity between the original |ΨK\ket{\Psi_{K}} and the perturbed |Ψ,K\ket{\Psi_{\star,K}} composite states in S+ES+E, as shown in Fig. 1(b). Let H,SH_{\star,S} and L,mL_{\star,m} be the perturbed Hamiltonian and jump operators, respectively. The perturbed state |Ψ,K\ket{\Psi_{\star,K}} is expressed by Eq. (13), where Y(w,m)Y(w,m) should be replaced by the perturbed operator Y(w,m)L,meiH,effwY_{\star}(w,m)\equiv L_{\star,m}e^{-iH_{\star,\mathrm{eff}}w} with H,effH,Si2m=1ML,mL,mH_{\star,\mathrm{eff}}\equiv H_{\star,S}-\frac{i}{2}\sum_{m=1}^{M}L_{\star,m}^{\dagger}L_{\star,m}. Then the Loschmidt echo η\eta becomes

η=|Tr[𝒵K(ρS(0))]|2,\eta=\left|\mathrm{Tr}\left[\mathcal{Z}_{\star}^{K}(\rho_{S}(0))\right]\right|^{2}, (15)

where ρS(0)\rho_{S}(0) is the initial state in SS and 𝒵\mathcal{Z}_{\star} is a mapping defined by

𝒵()m0𝑑wY(w,m)Y(w,m).\mathcal{Z}_{\star}(\bullet)\equiv\sum_{m}\int_{0}^{\infty}dw\,Y(w,m)\bullet Y_{\star}^{\dagger}(w,m). (16)

Here, 𝒜K\mathcal{A}^{K} denotes exponentiation of a mapping 𝒜\mathcal{A}, i.e.,

𝒜K𝒜𝒜𝒜K,\mathcal{A}^{K}\equiv\underbrace{\mathcal{A}\circ\mathcal{A}\circ\cdots\circ\mathcal{A}}_{K}, (17)

where \circ denotes a composition. Note that the mapping 𝒵\mathcal{Z}_{\star} is not a Kraus representation, because Tr[𝒵(ρS)]1\mathrm{Tr}[\mathcal{Z}_{\star}(\rho_{S})]\neq 1 in general. 𝒵\mathcal{Z}_{\star} is the fixed jump number analog of a two-sided Lindblad operator [49, 50], which is used to calculate the Loschmidt echo in the fixed end time system. Composition of the mapping can be reduced to a matrix multiplication when we convert 𝒵\mathcal{Z}_{\star} from the Hilbert space to the Liouville space (Appendix A), thereby making it easy to calculate η\eta.

Substituting =𝕀S𝒪\mathcal{F}=\mathbb{I}_{S}\otimes\mathcal{O} in Eq. (14) and η\eta in Eq. (15) into Eq. (4), we obtain

(𝒪+𝒪𝒪𝒪)21|Tr[𝒵K(ρS(0))]|21,\left(\frac{\llbracket\mathcal{O}\rrbracket+\llbracket\mathcal{O}\rrbracket_{\star}}{\braket{\mathcal{O}}-\mathcal{\braket{\mathcal{O}}_{\star}}}\right)^{2}\geq\frac{1}{\left|\mathrm{Tr}\left[\mathcal{Z}_{\star}^{K}(\rho_{S}(0))\right]\right|^{-2}-1}, (18)

which is the main result in this manuscript. Equation (18) is a TUR for first passage processes in quantum Markov chains. The above calculations assume an initially pure state ρS(0)=|ψSψS|\rho_{S}(0)=\ket{\psi_{S}}\bra{\psi_{S}}, but it is straightforward to generalize to an initially mixed state through purification as was done in Ref. [45]. It is important to note that the Loschmidt echo can be computed using information of the principal system only, i.e., HSH_{S}, H,SH_{\star,S}, LmL_{m}, and L,mL_{\star,m}. Moreover, by using an ancilla qubit, we can show that the Loschmidt echo of Eq. (15) becomes a measurable quantity of a physical process (Appendix B).

Equation (18) holds for arbitrary perturbed dynamics specified by H,SH_{\star,S} and L,mL_{\star,m}. Next, we consider a specific perturbed dynamics where Eq. (18) reduces to a more intuitive form. Let us consider the following perturbed operators:

L,m\displaystyle L_{\star,m} =1+εLm,\displaystyle=\sqrt{1+\varepsilon}L_{m}, (19)
H,S\displaystyle H_{\star,S} =(1+ε)HS,\displaystyle=(1+\varepsilon)H_{S}, (20)

where ε\varepsilon is a real value parameter. We take the ε0\varepsilon\to 0 limit. Because the Lindblad equation for the perturbed dynamics specified by Eqs. (19) and (20) is given by ρ˙=(1+ε)(ρS)\dot{\rho}=(1+\varepsilon)\mathcal{L}(\rho_{S}), the perturbed dynamics is the same as the original dynamics, except for the time scale. We may define a first passage time observable 𝒪f\mathcal{O}_{f}, which is a subset of 𝒪\mathcal{O}, where the subscript ff denotes the capital of the first passage time. The time scale of the perturbed dynamics is (1+ε)(1+\varepsilon) times faster than that of the original dynamics. We assume that the first passage time observable 𝒪f\mathcal{O}_{f} scales as

𝒪f\displaystyle\braket{\mathcal{O}_{f}}_{\star} =11+ε𝒪f,\displaystyle=\frac{1}{1+\varepsilon}\braket{\mathcal{O}_{f}}, (21)
𝒪f\displaystyle\llbracket\mathcal{O}_{f}\rrbracket_{\star} =11+ε𝒪f.\displaystyle=\frac{1}{1+\varepsilon}\llbracket\mathcal{O}_{f}\rrbracket. (22)

For instance, when h(𝒘,𝒎)=i=1Kwi=tKh(\boldsymbol{w},\boldsymbol{m})=\sum_{i=1}^{K}w_{i}=t_{K} in Eq. (14), the observable corresponds to the first passage time to reach KK jump events. Apparently, this first passage time satisfies Eqs. (21) and (22). With this scaling relation, the left hand side of Eq. (18) becomes

𝒪f+𝒪f𝒪f𝒪f=ε+2ε𝒪f𝒪f.\frac{\llbracket\mathcal{O}_{f}\rrbracket+\llbracket\mathcal{O}_{f}\rrbracket_{\star}}{\braket{\mathcal{O}_{f}}-\braket{\mathcal{O}_{f}}_{\star}}=\frac{\varepsilon+2}{\varepsilon}\frac{\llbracket\mathcal{O}_{f}\rrbracket}{\braket{\mathcal{O}_{f}}}. (23)

Next, we evaluate the right-hand side of Eq. (18). When we employ the perturbed Hamiltonian and jump operators in Eqs. (20) and (19), respectively, the ε\varepsilon dependence of the composite state of the perturbed dynamics can be expressed as

|Ψ,K=|ΨK(ε).\ket{\Psi_{\star,K}}=\ket{\Psi_{K}(\varepsilon)}. (24)

When ε=0\varepsilon=0, Eq. (24) reduces to the unperturbed state, |ΨK(ε=0)=|ΨK\ket{\Psi_{K}(\varepsilon=0)}=\ket{\Psi_{K}}. Let us consider the quantum Fisher information [51, 52], which is obtained by the following calculation:

𝒥K(α)\displaystyle\mathcal{J}_{K}(\alpha) =4αΨK(α)|αΨK(α)\displaystyle=4\braket{\partial_{\alpha}\Psi_{K}(\alpha)}{\partial_{\alpha}\Psi_{K}(\alpha)}
+4(αΨK(α)|ΨK(α))2,\displaystyle+4(\braket{\partial_{\alpha}\Psi_{K}(\alpha)}{\Psi_{K}(\alpha)})^{2}, (25)

where |αΨK(α)(d/dα)|ΨK(α)\ket{\partial_{\alpha}\Psi_{K}(\alpha)}\equiv(d/d\alpha)\ket{\Psi_{K}(\alpha)}. It is known that the fidelity and quantum Fisher information are related via [53]

𝒥K(α)=8ε2[1|ΨK(α+ε)|ΨK(α)|],\mathcal{J}_{K}(\alpha)=\frac{8}{\varepsilon^{2}}\left[1-\left|\braket{\Psi_{K}(\alpha+\varepsilon)}{\Psi_{K}(\alpha)}\right|\right], (26)

where we consider the ε0\varepsilon\to 0 limit. Substituting Eqs. (23) and (26) into Eq. (18), we obtain

𝒪f2𝒪f21𝒥K(0).\frac{\llbracket\mathcal{O}_{f}\rrbracket^{2}}{\braket{\mathcal{O}_{f}}^{2}}\geq\frac{1}{\mathcal{J}_{K}(0)}. (27)

Equation (27) provides a TUR for the first passage time in quantum Markov chains. As will be shown later, when we only consider classical stochastic processes, we obtain 𝒥K(0)K\mathcal{J}_{K}(0)\to K. The quantum Fisher information gives the fundamental limit of the quantum parameter estimation [54, 55, 51, 52]. Moreover, it plays a central role in the quantum speed limit [56, 57, 58] and a quantum TUR [24]. If we measure quantum systems and obtain a measurement output, the obtained output can be treated classically, and we observe that the quantum nature does not come into play. Therefore, we let K(α;SE)\mathcal{I}_{K}(\alpha;\mathcal{M}_{SE}) be the classical Fisher information, which is obtained by applying a positive operator-valued measure (POVM) SE\mathcal{M}_{SE} to |ΨK(α)\ket{\Psi_{K}(\alpha)}. Because a record of continuous measurement can be obtained by measuring |ΨK(α)\ket{\Psi_{K}(\alpha)} by the projector |𝒘,𝒎𝒘,𝒎|\ket{\boldsymbol{w},\boldsymbol{m}}\bra{\boldsymbol{w},\boldsymbol{m}}, Kcm(α)\mathcal{I}_{K}^{\mathrm{cm}}(\alpha) is represented as

Kcm(α)=K(α;𝕀S{|𝒘,𝒎𝒘,𝒎|}).\mathcal{I}_{K}^{\mathrm{cm}}(\alpha)=\mathcal{I}_{K}(\alpha;\mathbb{I}_{S}\otimes\{\ket{\boldsymbol{w},\boldsymbol{m}}\bra{\boldsymbol{w},\boldsymbol{m}}\}). (28)

Because the quantum Fisher information is larger than the classical counterpart, we have

Kcm(α)𝒥K(α).\mathcal{I}_{K}^{\mathrm{cm}}(\alpha)\leq\mathcal{J}_{K}(\alpha). (29)

We have derived Eq. (27) using the lower bound of the Hellinger distance and considering a particular perturbed dynamics specified by Eqs. (19) and (20). As shown in Appendix D, Eq. (27) can also be derived through the classical Cramér–Rao inequality, which was employed to derive classical TURs [10, 59].

Refer to caption
Figure 4: (a) Quantum Fisher information 𝒥K(0)\mathcal{J}_{K}(0) as a function of κ\kappa. The dashed line denotes 𝒥K(0)\mathcal{J}_{K}(0) and the solid line denotes KK, which is the classical limit of 𝒥K(0)\mathcal{J}_{K}(0). The parameters are Δ=Ω=1.0\Delta=\Omega=1.0 and K=1K=1. (b) Precision 𝒪f2/𝒪f2\llbracket\mathcal{O}_{f}\rrbracket^{2}/\braket{\mathcal{O}_{f}}^{2} as a function of the quantum Fisher information 𝒥K(0)\mathcal{J}_{K}(0) (triangles) and the jump number KK (circles) for random realizations. The dashed line corresponds to 1/𝒥K(0)1/\mathcal{J}_{K}(0) for the triangles and 1/K1/K for the circles. The parameters are randomly selected from Δ[0.1,3.0]\Delta\in[0.1,3.0], Ω[0.1,3.0]\Omega\in[0.1,3.0], and κ[1.0,3.0]\kappa\in[1.0,3.0].

Next, we obtain the classical limit of Eq. (27). We consider a classical Markov chain with NSN_{S} states, B1,B2,,BNSB_{1},B_{2},...,B_{N_{S}}, and take γji\gamma_{ji} to be a time-independent transition rate from BiB_{i} to BjB_{j}. Eq. (27) then reduces to the following simple relation:

𝒪f2𝒪f21K,\frac{\llbracket\mathcal{O}_{f}\rrbracket^{2}}{\braket{\mathcal{O}_{f}}^{2}}\geq\frac{1}{K}, (30)

as shown in Appendix C. Equation (30) corresponds to a specific case of the result reported in Ref. [42]. Equation (30) shows that for first passage processes that stop after KK jumps, the lower bound does not depend on the details of the dynamics or the initial distribution.

III Example

We apply the main result, Eq. (18), to a two-level atom driven by a classical laser field. Let |ϵe\ket{\epsilon_{e}} and |ϵg\ket{\epsilon_{g}} be the excited and ground states, respectively. The Hamiltonian HSH_{S} and a jump operator LL are given by

HS\displaystyle H_{S} =Δ|ϵeϵe|+Ω2[|ϵeϵg|+|ϵgϵe|],\displaystyle=\Delta\ket{\epsilon_{e}}\bra{\epsilon_{e}}+\frac{\Omega}{2}\left[\ket{\epsilon_{e}}\bra{\epsilon_{g}}+\ket{\epsilon_{g}}\bra{\epsilon_{e}}\right], (31)
L\displaystyle L =κ|ϵgϵe|,\displaystyle=\sqrt{\kappa}\ket{\epsilon_{g}}\bra{\epsilon_{e}}, (32)

where Δ\Delta is the detuning between the laser field and the atomic-transition frequencies, Ω\Omega is the Rabi-oscillation frequency, and κ\kappa is the decay rate. LL induces a jump from |ϵe\ket{\epsilon_{e}} to |ϵg\ket{\epsilon_{g}}. We first calculate the parameter dependence of the quantum Fisher information 𝒥K(0)\mathcal{J}_{K}(0) in Eq. (27) on the decay rate κ\kappa. Figure 4(a) shows 𝒥K(0)\mathcal{J}_{K}(0) as a function of κ\kappa, with the parameters given in the caption. The dashed line denotes 𝒥K(0)\mathcal{J}_{K}(0) and the solid line shows KK, which is the classical limit of the lower bound, as shown in Eq. (30). We see that 𝒥K(0)\mathcal{J}_{K}(0) reduces to KK as κ\kappa\to\infty. Because jumps in the Lindblad dynamics become dominant compared to the dynamics induced by eiHeffte^{-iH_{\mathrm{eff}}t} for κ\kappa\to\infty, the dynamics becomes closed to a classical Markov chain. Moreover, we observe that 𝒥K(0)\mathcal{J}_{K}(0) is always larger than KK, which indicates that the lower bound of the TUR is smaller for this quantum Markov chain.

To confirm that the fluctuation of the first passage observable 𝒪f\mathcal{O}_{f} is smaller for the quantum Markov chain, we perform a computer simulation. We randomly generate Δ\Delta, Ω\Omega, and κ\kappa and calculate 𝒪f2/𝒪f2\llbracket\mathcal{O}_{f}\rrbracket^{2}/\braket{\mathcal{O}_{f}}^{2}. Specifically, we consider 𝒪f\mathcal{O}_{f} with h(𝒘,𝒎)=i=1Kwi=tKh(\boldsymbol{w},\boldsymbol{m})=\sum_{i=1}^{K}w_{i}=t_{K}, which gives the first passage time needed to undergo KK jump events. We plot 𝒪f2/𝒪f2\llbracket\mathcal{O}_{f}\rrbracket^{2}/\braket{\mathcal{O}_{f}}^{2} as a function of the quantum Fisher information 𝒥K(0)\mathcal{J}_{K}(0) with triangles, where the dashed line denotes the lower bound, as shown in Fig. 4(b). We confirm that all realizations are above the dashed line, which numerically verifies Eq. (27). We have shown that Eq. (27) reduces to Eq. (30) in a classical case, where the lower bound is given by 1/K1/K. Therefore, we also check whether 𝒪f2/𝒪f2\llbracket\mathcal{O}_{f}\rrbracket^{2}/\braket{\mathcal{O}_{f}}^{2} can be bounded from below by 1/K1/K. We plot 𝒪f2/𝒪f2\llbracket\mathcal{O}_{f}\rrbracket^{2}/\braket{\mathcal{O}_{f}}^{2} as a function of KK with circles, where the dashed line now describes 1/K1/K, as shown in Fig. 4(b). We observe that some circles are below the dashed line, which indicates that the observable 𝒪f\mathcal{O}_{f} cannot be bounded from below by 1/K1/K. As explained in Fig. 4(a), this enhancement of precision can be ascribed to the fact that 𝒥K(0)\mathcal{J}_{K}(0) is larger than KK for smaller κ\kappa, where the state change in the dynamics is mostly induced by HeffH_{\mathrm{eff}}. Therefore, the precision enhancement in the quantum Markov chain is due to the state change via the effective Hamiltonian. Quantum-induced precision enhancement has been reported for quantum Markov chains with a fixed end time [20, 24]. Recently, it was reported that quantum coherence can improve the precision of a quantum heat engine [28].

IV Conclusion

In this manuscript, we derived a TUR for a first passage time in quantum Markov chains. Furthermore, we used a derivation technique developed in Ref. [45], where the fluctuation of observables is bounded from below by the Loschmidt echo. Our approach is quite general, and it can be applied to many quantum systems, which could not be handled with the previous derivations. A possible application of our approach is an extension to non-Markovian systems such as time-delayed systems [60]. Moreover, recently, the notion of quantum thermodynamics has been generalized to quantum field theory [61, 62]. Because the input-output formalism itself is a quantum field, we expect that uncertainty relations for quantum field theory can be derived using our technique.

Acknowledgements.
This work was supported by the Ministry of Education, Culture, Sports, Science and Technology (MEXT) KAKENHI Grant No. JP19K12153.

Appendix A Liouville space representation

The mapping 𝒵\mathcal{Z}_{\star} in the Hilbert space is shown in Eq. (16). Because the calculation of 𝒵K\mathcal{Z}_{\star}^{K} is a computationally expensive task, we can use the Liouville space representation of 𝒵\mathcal{Z}_{\star}. The Liouville space representation is advantageous because the mapping can be realized via matrix multiplication. In the Hilbert space, the density operator is

ρ=i,jϱij|ij|,\rho=\sum_{i,j}\varrho_{ij}\ket{i}\bra{j}, (33)

where |i\ket{i} is an orthonormal basis. It can be represented in the Liouville space as follows [63]:

vec(ρ)=i,jϱij|j|i.\mathrm{vec}(\rho)=\sum_{i,j}\varrho_{ij}\ket{j}\otimes\ket{i}. (34)

Let 𝖠\mathsf{A}, 𝖡\mathsf{B}, and 𝖢\mathsf{C} be arbitrary matrices in the Hilbert space. Then the following relation holds:

vec(𝖠𝖡𝖢)=(𝖢𝖠)vec(𝖡),\mathrm{vec}(\mathsf{A}\mathsf{B}\mathsf{C})=(\mathsf{C}^{\top}\otimes\mathsf{A})\mathrm{vec}(\mathsf{B}), (35)

where \top denotes the transpose. When the dimensions of the operators in the Hilbert space is d×dd\times d, their corresponding Liouville space representations have the dimensions of d2×d2d^{2}\times d^{2}. Using Eq. (35), we obtain 𝒵\mathcal{Z}_{\star} [Eq. (16)] in the Liouville space as follows:

m=1M0𝑑wY(w,m)Y(w,m),\mathfrak{Z}_{\star}\equiv\sum_{m=1}^{M}\int_{0}^{\infty}dw\,Y_{\star}^{*}(w,m)\otimes Y(w,m), (36)

where the superscript * denotes complex conjugate. Then, exponentiation of a mapping, 𝒵K(ρS)\mathcal{Z}_{\star}^{K}(\rho_{S}), is simply realized by a matrix power Kvec(ρS)\mathfrak{Z}_{\star}^{K}\mathrm{vec}(\rho_{S}).

Appendix B Measurement of Loschmidt echo

We here show a method for obtaining the Loschmidt echo given by Eq. (15) through measurement of a physical process.

We first review the well known approach for calculating the Loschmidt echo for closed quantum dynamics. We can measure the Loschmidt echo for closed quantum dynamics by introducing an ancilla qubit. Let HH and HH_{\star} be Hamiltonian operators of the original and the perturbed dynamics, respectively, with which we want to calculate the Loschmidt echo, and let |0\ket{\mathfrak{0}} and |1\ket{\mathfrak{1}} be the two states of the ancilla qubit. The Loschmidt echo for closed quantum dynamics is given by Eq. (1). We define the Hamiltonian in the composite system comprising the ancilla and system as follows:

H~|00|H+|11|H=[H00H].\tilde{H}\equiv\ket{\mathfrak{0}}\bra{\mathfrak{0}}\otimes H+\ket{\mathfrak{1}}\bra{\mathfrak{1}}\otimes H_{\star}=\left[\begin{array}[]{cc}H&0\\ 0&H_{\star}\end{array}\right]. (37)

The initial state of the composite system is |Ψ~(0)=12(|0+|1)|Ψ(0)\ket{\tilde{\Psi}(0)}=\frac{1}{\sqrt{2}}(\ket{\mathfrak{0}}+\ket{\mathfrak{1}})\otimes\ket{\Psi(0)}, where |Ψ(0)\ket{\Psi(0)} is the initial state of the original system. The state of the ancilla after the time evolution is given by

TrS[eiH~τ|Ψ~(0)Ψ~(0)|eiH~τ]\displaystyle\mathrm{Tr}_{S}\left[e^{-i\tilde{H}\tau}\ket{\tilde{\Psi}(0)}\bra{\tilde{\Psi}(0)}e^{i\tilde{H}\tau}\right]
=12[1Ψ(τ)|Ψ(τ)Ψ(τ)|Ψ(τ)1],\displaystyle=\frac{1}{2}\left[\begin{array}[]{cc}1&\braket{\Psi_{\star}(\tau)}{\Psi(\tau)}\\ \braket{\Psi(\tau)}{\Psi_{\star}(\tau)}&1\end{array}\right], (40)

where TrS\mathrm{Tr}_{S} is the trace operation with respect to the principal system SS, |Ψ(τ)=eiHτ|Ψ(0)\ket{\Psi(\tau)}=e^{-iH\tau}\ket{\Psi(0)}, and |Ψ(τ)=eiHτ|Ψ(0)\ket{\Psi_{\star}(\tau)}=e^{-iH_{\star}\tau}\ket{\Psi(0)}. From Eq. (40), we see that the Loschmidt echo |Ψ(τ)|Ψ(τ)||\braket{\Psi_{\star}(\tau)}{\Psi(\tau)}| can be obtained by measuring the |01|\ket{\mathfrak{0}}\bra{\mathfrak{1}} (or |10|\ket{\mathfrak{1}}\bra{\mathfrak{0}}) element of the ancilla.

The Loschmidt echo for a Lindblad equation can be calculated in a similar way. For a system that ends at a constant time, the Loschmidt echo can be calculated by the procedure shown in Ref. [50]. We here show how we can measure the Loschmidt echo for a constant jump case. Similar to Eq. (37), we prepare an ancilla qubit and define the Hamiltonian and jump operators in the ancilla and system as follows:

H~S\displaystyle\tilde{H}_{S} |00|HS+|11|H,S=[HS00H,S],\displaystyle\equiv\ket{\mathfrak{0}}\bra{\mathfrak{0}}\otimes H_{S}+\ket{\mathfrak{1}}\bra{\mathfrak{1}}\otimes H_{\star,S}=\left[\begin{array}[]{cc}H_{S}&0\\ 0&H_{\star,S}\end{array}\right], (43)
L~m\displaystyle\tilde{L}_{m} |00|Lm+|11|L,m=[Lm00L,m].\displaystyle\equiv\ket{\mathfrak{0}}\bra{\mathfrak{0}}\otimes L_{m}+\ket{\mathfrak{1}}\bra{\mathfrak{1}}\otimes L_{\star,m}=\left[\begin{array}[]{cc}L_{m}&0\\ 0&L_{\star,m}\end{array}\right]. (46)

We define the density operator in the ancilla and system as follows:

ρ~(t)=12[ρ00(t)ρ01(t)ρ10(t)ρ11(t)],\tilde{\rho}(t)=\frac{1}{2}\left[\begin{array}[]{cc}\rho_{00}(t)&\rho_{01}(t)\\ \rho_{10}(t)&\rho_{11}(t)\end{array}\right], (47)

where the initial state is set to be ρ00(0)=ρ01(0)=ρ10(0)=ρ11(0)=ρS(0)\rho_{00}(0)=\rho_{01}(0)=\rho_{10}(0)=\rho_{11}(0)=\rho_{S}(0). Similar to the closed quantum case, let ρ~(t)\tilde{\rho}(t) evolve through H~S\tilde{H}_{S} [Eq. (43)] and L~m\tilde{L}_{m} [Eq. (46)]. Specifically, suppose that ρ~(t)\tilde{\rho}(t) is governed by the following Lindblad equation:

ρ~˙=i[H~S,ρ~]+m=1M𝒟(ρ~,L~m).\dot{\tilde{\rho}}=-i\left[\tilde{H}_{S},\tilde{\rho}\right]+\sum_{m=1}^{M}\mathcal{D}(\tilde{\rho},\tilde{L}_{m}). (48)

Let us consider a continuous measurement in Eq. (48). From Eq. (11), the Kraus representation describing the evolution from the first to the second jump events is

𝒵~()m=1M0𝑑wY~(w,m)Y~(w,m),\tilde{\mathcal{Z}}(\bullet)\equiv\sum_{m=1}^{M}\int_{0}^{\infty}dw\,\tilde{Y}(w,m)\bullet\tilde{Y}^{\dagger}(w,m), (49)

where Y~(w,m)L~meiH~effw\tilde{Y}(w,m)\equiv\tilde{L}_{m}e^{-i\tilde{H}_{\mathrm{eff}}w} with H~effH~Si2m=1ML~mL~m\tilde{H}_{\mathrm{eff}}\equiv\tilde{H}_{S}-\frac{i}{2}\sum_{m=1}^{M}\tilde{L}_{m}^{\dagger}\tilde{L}_{m}. Using Eqs. (43) and (46), Y~(w,m)\tilde{Y}(w,m) becomes

Y~(w,m)=[Y(w,m)00Y(w,m)].\tilde{Y}(w,m)=\left[\begin{array}[]{cc}Y(w,m)&0\\ 0&Y_{\star}(w,m)\end{array}\right]. (50)

Substituting Eqs. (47) and (50) into Eq. (49) yields

𝒵~(ρ~)\displaystyle\tilde{\mathcal{Z}}(\tilde{\rho}) =m=1M0𝑑w12[Y(w,m)ρ00Y(w,m)Y(w,m)ρ01Y(w,m)Y(w,m)ρ01Y(w,m)Y(w,m)ρ11Y(w,m)].\displaystyle=\sum_{m=1}^{M}\int_{0}^{\infty}dw\,\frac{1}{2}\left[\begin{array}[]{cc}Y(w,m)\rho_{00}Y^{\dagger}(w,m)&Y(w,m)\rho_{01}Y_{\star}^{\dagger}(w,m)\\ Y_{\star}(w,m)\rho_{01}Y^{\dagger}(w,m)&Y_{\star}(w,m)\rho_{11}Y_{\star}^{\dagger}(w,m)\end{array}\right]. (53)

Taking the trace with respect to the principal system in Eq. (53), we obtain the ancilla state after a single jump event:

TrS[𝒵~(ρ~)]\displaystyle\mathrm{Tr}_{S}\left[\tilde{\mathcal{Z}}(\tilde{\rho})\right] =m=1M0𝑑w12[1Tr[Y(w,m)ρ01Y(w,m)]Tr[Y(w,m)ρ01Y(w,m)]1].\displaystyle=\sum_{m=1}^{M}\int_{0}^{\infty}dw\,\frac{1}{2}\left[\begin{array}[]{cc}1&\mathrm{Tr}\left[Y(w,m)\rho_{01}Y_{\star}^{\dagger}(w,m)\right]\\ \mathrm{Tr}\left[Y_{\star}(w,m)\rho_{01}Y^{\dagger}(w,m)\right]&1\end{array}\right]. (56)

Notably, the |01|\ket{\mathfrak{0}}\bra{\mathfrak{1}} element of the state of the ancilla is identical to 𝒵()\mathcal{Z}_{\star}(\bullet) given in Eq. (16) (when multiplied by 22). Therefore, the Loschmidt echo for K=1K=1 can be obtained by simply measuring the |01|\ket{\mathfrak{0}}\bra{\mathfrak{1}} (or |10|\ket{\mathfrak{1}}\bra{\mathfrak{0}}) element of the ancilla after a single jump. Although the procedure explained above is for a single jump event, i.e., K=1K=1, it is straightforward to generalize to K>1K>1 cases. Therefore, we can measure the Loschmidt echo η\eta [Eq. (15)] by first considering the ancilla qubit and performing a continuous measurement on the composite system comprising the ancilla and system. After KK jump events, we measure the |01|\ket{\mathfrak{0}}\bra{\mathfrak{1}} (or |10|\ket{\mathfrak{1}}\bra{\mathfrak{0}}) element of the ancilla. By repeating this procedure many times and calculating the average of the measurements, we obtain 12Tr[𝒵K(ρS(0))]\frac{1}{2}\mathrm{Tr}\left[\mathcal{Z}_{\star}^{K}(\rho_{S}(0))\right] which directly gives the Loschmidt echo.

We numerically verify this approach of calculating the Loschmidt echo. We use a two-level atom driven by a classical laser field, the same model as in Section III. Figure 5 shows the Loschmidt echo η\eta as a function of KK obtained by the analytic and ancilla approaches, which are shown by circles and crosses, respectively. The settings for the calculation are shown in the caption of Fig. 5. For the ancilla approach, η\eta is calculated by the |01|\ket{\mathfrak{0}}\bra{\mathfrak{1}} element of the ancilla state after KK jump events. We can see that the ancilla approach yields very close values for K7K\leq 7, which numerically confirm the validity of the ancilla approach. The disagreements for K8K\geq 8 are due to undersampling and hence can be improved by increasing the number of samples.

Refer to caption
Figure 5: Loschmidt echo η\eta as a function of the number of jump events KK, calculated by the analytic approach (circles) based on Eq. (15) and the ancilla approach (crosses) introduced in Sec. B. The initial state is ρS(0)=|ϵgϵg|\rho_{S}(0)=\ket{\epsilon_{g}}\bra{\epsilon_{g}}. The parameters are Δ=1.0\Delta=1.0, Ω=1.0\Omega=1.0, and κ=2.0\kappa=2.0 for the original process and Δ=0.4\Delta=0.4, Ω=1.2\Omega=1.2, and κ=0.5\kappa=0.5 for the perturbed process. For the ancilla approach, the number of trials used to calculate each η\eta is 1000010000.

Appendix C Classical bound

We derive the classical limit of Eq. (27). We can represent NSN_{S} states, B1,B2,,BNSB_{1},B_{2},\ldots,B_{N_{S}}, in a classical Markov chain by an orthonormal basis {|b1,|b2,,|bNS}\{\ket{b_{1}},\ket{b_{2}},\ldots,\ket{b_{N_{S}}}\}. To emulate classical Markov chains with quantum Markov chains, we consider the following Hamiltonian and jump operators in the Lindblad equation:

HS\displaystyle H_{S} =0,\displaystyle=0, (57)
Lji\displaystyle L_{ji} =γji|bjbi|(ij).\displaystyle=\sqrt{\gamma_{ji}}\ket{b_{j}}\bra{b_{i}}\,\,\,\,\,(i\neq j). (58)

Let ε\varepsilon be a small perturbation parameter as defined in Eqs. (19) and (20). The Hamiltonian and the jump operators for the perturbed dynamics become

H,S\displaystyle H_{\star,S} =0,\displaystyle=0, (59)
L,ji\displaystyle L_{\star,ji} =1+εγji|bjbi|(ij).\displaystyle=\sqrt{1+\varepsilon}\sqrt{\gamma_{ji}}\ket{b_{j}}\bra{b_{i}}\,\,\,\,\,(i\neq j). (60)

We also consider the following initial density operator which emulates the classical probability distribution:

ρS(t)=ipi(t)|bibi|,\rho_{S}(t)=\sum_{i}p_{i}(t)\ket{b_{i}}\bra{b_{i}}, (61)

where pi(t)p_{i}(t) is the classical probability distribution of being |bi\ket{b_{i}} at time tt. Let Y(w,i,j)LjieiHeffwY(w,i,j)\equiv L_{ji}e^{-iH_{\mathrm{eff}}w} (with Y(w,i,j)Y_{\star}(w,i,j) being defined in the same way). Substituting Eqs. (57)–(61) into Eq. (16), 𝒵\mathcal{Z}_{\star} in Eq. (16) becomes

𝒵(ρS)\displaystyle\mathcal{Z}_{\star}(\rho_{S}) =ij0𝑑wY(w,i,j)ρSY(w,i,j)\displaystyle=\sum_{i\neq j}\int_{0}^{\infty}dwY(w,i,j)\rho_{S}Y_{\star}^{\dagger}(w,i,j)
=ij0𝑑w1+εγji\displaystyle=\sum_{i\neq j}\int_{0}^{\infty}dw\,\sqrt{1+\varepsilon}\gamma_{ji}
×exp[w2(2+ε)k(i)γki]pi|bjbj|\displaystyle\times\exp\left[-\frac{w}{2}(2+\varepsilon)\sum_{k(\neq i)}\gamma_{ki}\right]p_{i}\ket{b_{j}}\bra{b_{j}}
=ji(j)21+ε(2+ε)γjik(i)γkipi|bjbj|\displaystyle=\sum_{j}\sum_{i(\neq j)}\frac{2\sqrt{1+\varepsilon}}{(2+\varepsilon)}\frac{\gamma_{ji}}{\sum_{k(\neq i)}\gamma_{ki}}p_{i}\ket{b_{j}}\bra{b_{j}}
=j21+ε(2+ε)[𝒑]j|bjbj|.\displaystyle=\sum_{j}\frac{2\sqrt{1+\varepsilon}}{(2+\varepsilon)}[\mathcal{B}\boldsymbol{p}]_{j}\ket{b_{j}}\bra{b_{j}}. (62)

where 𝒑=[p1,p2,,pN]\boldsymbol{p}=[p_{1},p_{2},\ldots,p_{N}]^{\top} and \mathcal{B} is a matrix defined by

ij{0i=jγijk(j)γkjij.\mathcal{B}_{ij}\equiv\begin{cases}0&i=j\\ \frac{\gamma_{ij}}{\sum_{k(\neq j)}\gamma_{kj}}&i\neq j\end{cases}. (63)

Therefore we obtain

Tr[𝒵K(ρS)]\displaystyle\mathrm{Tr}\left[\mathcal{Z}_{\star}^{K}(\rho_{S})\right] =j(21+ε2+ε)K[K𝒑]j.\displaystyle=\sum_{j}\left(\frac{2\sqrt{1+\varepsilon}}{2+\varepsilon}\right)^{K}\left[\mathcal{B}^{K}\boldsymbol{p}\right]_{j}. (64)

From Eq. (63), \mathcal{B} is a stochastic matrix iij=1\sum_{i}\mathcal{B}_{ij}=1, yielding

Tr[𝒵K(ρS)]=(21+ε2+ε)K.\mathrm{Tr}\left[\mathcal{Z}_{\star}^{K}(\rho_{S})\right]=\left(\frac{2\sqrt{1+\varepsilon}}{2+\varepsilon}\right)^{K}. (65)

𝒪f\mathcal{O}_{f} in the classical systems also satisfies the scaling condition of Eq. (23). From Eqs. (23) and (65), Eq. (18) becomes

𝒪f2𝒪f21(ε+2ε)2[(21+ε2+ε)2K1]ε01K,\frac{\llbracket\mathcal{O}_{f}\rrbracket^{2}}{\braket{\mathcal{O}_{f}}^{2}}\geq\frac{1}{\left(\frac{\varepsilon+2}{\varepsilon}\right)^{2}\left[\left(\frac{2\sqrt{1+\varepsilon}}{2+\varepsilon}\right)^{-2K}-1\right]}\overset{\varepsilon\to 0}{\to}\frac{1}{K}, (66)

where we used l’Hôpital’s rule for calculating the limit. Equation (66) is a classical case of the main result [Eq. (30)].

Appendix D Derivation based on classical Cramér–Rao inequality

Equation (27), which is a particular case of the main result [Eq. (18)], can also be derived through the classical Cramér–Rao inequality. The classical Cramér–Rao inequality states

𝒪α2(α𝒪α)21Kcm(α),\frac{\llbracket\mathcal{O}\rrbracket_{\alpha}^{2}}{\left(\partial_{\alpha}\braket{\mathcal{O}}_{\alpha}\right)^{2}}\geq\frac{1}{\mathcal{I}_{K}^{\mathrm{cm}}(\alpha)}, (67)

where 𝒪\mathcal{O} is defined in Eq. (14), and 𝒪αΨK(α)|𝒪|ΨK(α)\braket{\mathcal{O}}_{\alpha}\equiv\braket{\Psi_{K}(\alpha)}{\mathcal{O}}{\Psi_{K}(\alpha)} and 𝒪α𝒪2α𝒪α2\llbracket\mathcal{O}\rrbracket_{\alpha}\equiv\sqrt{\braket{\mathcal{O}^{2}}_{\alpha}-\braket{\mathcal{O}}_{\alpha}^{2}}. When we only consider the first passage time observable 𝒪f\mathcal{O}_{f}, the left hand side of Eq. (67) becomes

𝒪fα2(α𝒪fα)2=1(1+α)2𝒪f2(α11+α𝒪f)2=(1+α)2𝒪f2𝒪f2.\frac{\llbracket\mathcal{O}_{f}\rrbracket_{\alpha}^{2}}{\left(\partial_{\alpha}\braket{\mathcal{O}_{f}}_{\alpha}\right)^{2}}=\frac{\frac{1}{(1+\alpha)^{2}}\llbracket\mathcal{O}_{f}\rrbracket^{2}}{\left(\partial_{\alpha}\frac{1}{1+\alpha}\braket{\mathcal{O}_{f}}\right)^{2}}=(1+\alpha)^{2}\frac{\llbracket\mathcal{O}_{f}\rrbracket^{2}}{\braket{\mathcal{O}_{f}}^{2}}. (68)

Using Eqs. (29) and (68), we obtain Eq. (27).

References