Thermodynamic uncertainty relation for quantum first passage process
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 and ends at a fixed time (). 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.

II Methods
The Loschmidt echo is an indicator of quantum chaos in many-body systems. Let be a Hamiltonian and let be a perturbed Hamiltonian. When the system is closed, given the initial state , the Loschmidt echo is defined by
(1) |
The Loschmidt echo defined in Eq. (1) can be regarded as the fidelity between and , which are time-evolved states at induced by and , respectively, as shown in Fig. 1(a). If the system is sensitive to the perturbation, the fidelity decays quickly to , 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:
(2) |
where is a density operator and is a Kraus operator satisfying a completeness relation with being the identity operator in . This time evolution induced by the Kraus representation can be expressed by a unitary time evolution in a larger space comprising the system and environment :
(3) |
where constitutes an orthonormal basis of , and are initial states of and , respectively, and is a unitary operator acting on . Tracing out in Eq. (3), we recover the original Kraus representation of Eq. (2). Suppose that a Kraus operator for the perturbed dynamics is . Then the fidelity can be calculated between and , as shown in Fig. 1(b). Throughout this manuscript, we use 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 and be two pure states and let be an Hermitian observable that is applied to or . We define the mean and standard deviation of by and , respectively. and should be evaluated for instead of . Then the following relation holds [45]:
(4) |
where . 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 jump events, where . Let be a density operator of the principal system at time . We assume that the dynamics of is governed by the Lindblad equation:
(5) |
where is the Lindblad super-operator, is a system Hamiltonian, represents jump operators, is the number of jump operators, and is the dissipator defined by
(6) |
Here, denotes the anti-commutator. The solution of Eq. (5) corresponds to the dynamics when we do not measure the environment . However, when we measure the environment , 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:
(7) |
where is an effective Hamiltonian:
(8) |
Equation (7) shows that the dynamics of is governed by two contributions: the effective Hamiltonian term , which induces a continuous smooth state change, and the jump operator term , which causes a discontinuous jump event. Suppose that the state of right after a jump event is . Continuous evolution in the quantum trajectory is given by , which is not normalized, and the next jump is induced by the th jump operator after a waiting time . Therefore, from the state after the first jump event to the state after the second jump event, the state is transformed as
(9) |
where . satisfies a completeness relation , which can be shown by
(10) |
Therefore, the Kraus representation describing the evolution from the first to the second jump events is given by
(11) |

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 in Eq. (2) by , the Kraus representation of Eq. (11) can be represented as an interaction between the principal system and the environment as follows:
(12) |
where constitutes an orthonormal basis in as .

We consider a quantum Markov chain that stops after jump events. Let be a composite state in after jump events. Let be the output of the th jump, be the time stamp of the th jump (), and be the waiting time between the th and th jump events (i.e., ). Figures 3(a) and (b) are examples of the process, where the trajectories stop when they undergo jump events. We observe that the first passage time is a random variable, and it is different for these two cases. Repeating Eq. (12) times, is expressed by
(13) |
where is an abbreviation for . 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 and their time stamp , as shown in Fig. 3. Because knowing the waiting time is equivalent to the time stamp , , and can be obtained by applying the projector , where , to in Eq. (13). We consider the following Hermitian observable on .
(14) |
where is an arbitrary real function of and .
The Loschmidt echo considers the fidelity between the original and the perturbed composite states in , as shown in Fig. 1(b). Let and be the perturbed Hamiltonian and jump operators, respectively. The perturbed state is expressed by Eq. (13), where should be replaced by the perturbed operator with . Then the Loschmidt echo becomes
(15) |
where is the initial state in and is a mapping defined by
(16) |
Here, denotes exponentiation of a mapping , i.e.,
(17) |
where denotes a composition. Note that the mapping is not a Kraus representation, because in general. 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 from the Hilbert space to the Liouville space (Appendix A), thereby making it easy to calculate .
Substituting in Eq. (14) and in Eq. (15) into Eq. (4), we obtain
(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 , 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., , , , and . 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 and . Next, we consider a specific perturbed dynamics where Eq. (18) reduces to a more intuitive form. Let us consider the following perturbed operators:
(19) | ||||
(20) |
where is a real value parameter. We take the limit. Because the Lindblad equation for the perturbed dynamics specified by Eqs. (19) and (20) is given by , the perturbed dynamics is the same as the original dynamics, except for the time scale. We may define a first passage time observable , which is a subset of , where the subscript denotes the capital of the first passage time. The time scale of the perturbed dynamics is times faster than that of the original dynamics. We assume that the first passage time observable scales as
(21) | ||||
(22) |
For instance, when in Eq. (14), the observable corresponds to the first passage time to reach jump events. Apparently, this first passage time satisfies Eqs. (21) and (22). With this scaling relation, the left hand side of Eq. (18) becomes
(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 dependence of the composite state of the perturbed dynamics can be expressed as
(24) |
When , Eq. (24) reduces to the unperturbed state, . Let us consider the quantum Fisher information [51, 52], which is obtained by the following calculation:
(25) |
where . It is known that the fidelity and quantum Fisher information are related via [53]
(26) |
where we consider the limit. Substituting Eqs. (23) and (26) into Eq. (18), we obtain
(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 . 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 be the classical Fisher information, which is obtained by applying a positive operator-valued measure (POVM) to . Because a record of continuous measurement can be obtained by measuring by the projector , is represented as
(28) |
Because the quantum Fisher information is larger than the classical counterpart, we have
(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].

Next, we obtain the classical limit of Eq. (27). We consider a classical Markov chain with states, , and take to be a time-independent transition rate from to . Eq. (27) then reduces to the following simple relation:
(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 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 and be the excited and ground states, respectively. The Hamiltonian and a jump operator are given by
(31) | ||||
(32) |
where is the detuning between the laser field and the atomic-transition frequencies, is the Rabi-oscillation frequency, and is the decay rate. induces a jump from to . We first calculate the parameter dependence of the quantum Fisher information in Eq. (27) on the decay rate . Figure 4(a) shows as a function of , with the parameters given in the caption. The dashed line denotes and the solid line shows , which is the classical limit of the lower bound, as shown in Eq. (30). We see that reduces to as . Because jumps in the Lindblad dynamics become dominant compared to the dynamics induced by for , the dynamics becomes closed to a classical Markov chain. Moreover, we observe that is always larger than , 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 is smaller for the quantum Markov chain, we perform a computer simulation. We randomly generate , , and and calculate . Specifically, we consider with , which gives the first passage time needed to undergo jump events. We plot as a function of the quantum Fisher information 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 . Therefore, we also check whether can be bounded from below by . We plot as a function of with circles, where the dashed line now describes , as shown in Fig. 4(b). We observe that some circles are below the dashed line, which indicates that the observable cannot be bounded from below by . As explained in Fig. 4(a), this enhancement of precision can be ascribed to the fact that is larger than for smaller , where the state change in the dynamics is mostly induced by . 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 in the Hilbert space is shown in Eq. (16). Because the calculation of is a computationally expensive task, we can use the Liouville space representation of . The Liouville space representation is advantageous because the mapping can be realized via matrix multiplication. In the Hilbert space, the density operator is
(33) |
where is an orthonormal basis. It can be represented in the Liouville space as follows [63]:
(34) |
Let , , and be arbitrary matrices in the Hilbert space. Then the following relation holds:
(35) |
where denotes the transpose. When the dimensions of the operators in the Hilbert space is , their corresponding Liouville space representations have the dimensions of . Using Eq. (35), we obtain [Eq. (16)] in the Liouville space as follows:
(36) |
where the superscript denotes complex conjugate. Then, exponentiation of a mapping, , is simply realized by a matrix power .
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 and be Hamiltonian operators of the original and the perturbed dynamics, respectively, with which we want to calculate the Loschmidt echo, and let and 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:
(37) |
The initial state of the composite system is , where is the initial state of the original system. The state of the ancilla after the time evolution is given by
(40) |
where is the trace operation with respect to the principal system , , and . From Eq. (40), we see that the Loschmidt echo can be obtained by measuring the (or ) 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:
(43) | ||||
(46) |
We define the density operator in the ancilla and system as follows:
(47) |
where the initial state is set to be . Similar to the closed quantum case, let evolve through [Eq. (43)] and [Eq. (46)]. Specifically, suppose that is governed by the following Lindblad equation:
(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
(49) |
where with . Using Eqs. (43) and (46), becomes
(50) |
Substituting Eqs. (47) and (50) into Eq. (49) yields
(53) |
Taking the trace with respect to the principal system in Eq. (53), we obtain the ancilla state after a single jump event:
(56) |
Notably, the element of the state of the ancilla is identical to given in Eq. (16) (when multiplied by ). Therefore, the Loschmidt echo for can be obtained by simply measuring the (or ) element of the ancilla after a single jump. Although the procedure explained above is for a single jump event, i.e., , it is straightforward to generalize to cases. Therefore, we can measure the Loschmidt echo [Eq. (15)] by first considering the ancilla qubit and performing a continuous measurement on the composite system comprising the ancilla and system. After jump events, we measure the (or ) element of the ancilla. By repeating this procedure many times and calculating the average of the measurements, we obtain 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 as a function of 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, is calculated by the element of the ancilla state after jump events. We can see that the ancilla approach yields very close values for , which numerically confirm the validity of the ancilla approach. The disagreements for are due to undersampling and hence can be improved by increasing the number of samples.

Appendix C Classical bound
We derive the classical limit of Eq. (27). We can represent states, , in a classical Markov chain by an orthonormal basis . To emulate classical Markov chains with quantum Markov chains, we consider the following Hamiltonian and jump operators in the Lindblad equation:
(57) | ||||
(58) |
Let be a small perturbation parameter as defined in Eqs. (19) and (20). The Hamiltonian and the jump operators for the perturbed dynamics become
(59) | ||||
(60) |
We also consider the following initial density operator which emulates the classical probability distribution:
(61) |
where is the classical probability distribution of being at time . Let (with being defined in the same way). Substituting Eqs. (57)–(61) into Eq. (16), in Eq. (16) becomes
(62) |
where and is a matrix defined by
(63) |
Therefore we obtain
(64) |
From Eq. (63), is a stochastic matrix , yielding
(65) |
in the classical systems also satisfies the scaling condition of Eq. (23). From Eqs. (23) and (65), Eq. (18) becomes
(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
(67) |
where is defined in Eq. (14), and and . When we only consider the first passage time observable , the left hand side of Eq. (67) becomes
(68) |
References
- Barato and Seifert [2015] A. C. Barato and U. Seifert, Thermodynamic uncertainty relation for biomolecular processes, Phys. Rev. Lett. 114, 158101 (2015).
- Gingrich et al. [2016] T. R. Gingrich, J. M. Horowitz, N. Perunov, and J. L. England, Dissipation bounds all steady-state current fluctuations, Phys. Rev. Lett. 116, 120601 (2016).
- Pietzonka et al. [2016] P. Pietzonka, A. C. Barato, and U. Seifert, Universal bounds on current fluctuations, Phys. Rev. E 93, 052145 (2016).
- Horowitz and Gingrich [2017] J. M. Horowitz and T. R. Gingrich, Proof of the finite-time thermodynamic uncertainty relation for steady-state currents, Phys. Rev. E 96, 020103 (2017).
- Pigolotti et al. [2017] S. Pigolotti, I. Neri, E. Roldán, and F. Jülicher, Generic properties of stochastic entropy production, Phys. Rev. Lett. 119, 140604 (2017).
- Garrahan [2017] J. P. Garrahan, Simple bounds on fluctuations and uncertainty relations for first-passage times of counting observables, Phys. Rev. E 95, 032134 (2017).
- Dechant and Sasa [2018] A. Dechant and S.-i. Sasa, Current fluctuations and transport efficiency for general Langevin systems, J. Stat. Mech: Theory Exp. 2018, 063209 (2018).
- Barato et al. [2018] A. C. Barato, R. Chetrite, A. Faggionato, and D. Gabrielli, Bounds on current fluctuations in periodically driven systems, New J. Phys. 20 (2018).
- Terlizzi and Baiesi [2019] I. D. Terlizzi and M. Baiesi, Kinetic uncertainty relation, J. Phys. A: Math. Theor. 52, 02LT03 (2019).
- Hasegawa and Van Vu [2019a] Y. Hasegawa and T. Van Vu, Uncertainty relations in stochastic processes: An information inequality approach, Phys. Rev. E 99, 062126 (2019a).
- Hasegawa and Van Vu [2019b] Y. Hasegawa and T. Van Vu, Fluctuation theorem uncertainty relation, Phys. Rev. Lett. 123, 110602 (2019b).
- Van Vu and Hasegawa [2019] T. Van Vu and Y. Hasegawa, Uncertainty relations for underdamped Langevin dynamics, Phys. Rev. E 100, 032130 (2019).
- Van Vu and Hasegawa [2020] T. Van Vu and Y. Hasegawa, Thermodynamic uncertainty relations under arbitrary control protocols, Phys. Rev. Research 2, 013060 (2020).
- Dechant and Sasa [2020] A. Dechant and S.-i. Sasa, Fluctuation–response inequality out of equilibrium, Proc. Natl. Acad. Sci. U.S.A. 117, 6430 (2020).
- Vo et al. [2020] V. T. Vo, T. Van Vu, and Y. Hasegawa, Unified approach to classical speed limit and thermodynamic uncertainty relation, Phys. Rev. E 102, 062132 (2020).
- Koyuk and Seifert [2020] T. Koyuk and U. Seifert, Thermodynamic uncertainty relation for time-dependent driving, Phys. Rev. Lett. 125, 260604 (2020).
- Dechant and ichi Sasa [2020] A. Dechant and S. ichi Sasa, Continuous time-reversal and equality in the thermodynamic uncertainty relation, arXiv:2010.14769 (2020).
- Erker et al. [2017] P. Erker, M. T. Mitchison, R. Silva, M. P. Woods, N. Brunner, and M. Huber, Autonomous quantum clocks: Does thermodynamics limit our ability to measure time?, Phys. Rev. X 7, 031022 (2017).
- Brandner et al. [2018] K. Brandner, T. Hanazato, and K. Saito, Thermodynamic bounds on precision in ballistic multiterminal transport, Phys. Rev. Lett. 120, 090601 (2018).
- Carollo et al. [2019] F. Carollo, R. L. Jack, and J. P. Garrahan, Unraveling the large deviation statistics of Markovian open quantum systems, Phys. Rev. Lett. 122, 130605 (2019).
- Liu and Segal [2019] J. Liu and D. Segal, Thermodynamic uncertainty relation in quantum thermoelectric junctions, Phys. Rev. E 99, 062141 (2019).
- Guarnieri et al. [2019] G. Guarnieri, G. T. Landi, S. R. Clark, and J. Goold, Thermodynamics of precision in quantum nonequilibrium steady states, Phys. Rev. Research 1, 033021 (2019).
- Saryal et al. [2019] S. Saryal, H. M. Friedman, D. Segal, and B. K. Agarwalla, Thermodynamic uncertainty relation in thermal transport, Phys. Rev. E 100, 042101 (2019).
- Hasegawa [2020] Y. Hasegawa, Quantum thermodynamic uncertainty relation for continuous measurement, Phys. Rev. Lett. 125, 050601 (2020).
- Friedman et al. [2020] H. M. Friedman, B. K. Agarwalla, O. Shein-Lumbroso, O. Tal, and D. Segal, Thermodynamic uncertainty relation in atomic-scale quantum conductors, Phys. Rev. B 101, 195423 (2020).
- Hasegawa [2021a] Y. Hasegawa, Thermodynamic uncertainty relation for general open quantum systems, Phys. Rev. Lett. 126, 010602 (2021a).
- Sacchi [2021] M. F. Sacchi, Thermodynamic uncertainty relations for bosonic Otto engines, Phys. Rev. E 103, 012111 (2021).
- Kalaee et al. [2021] A. A. S. Kalaee, A. Wacker, and P. P. Potts, Violating the thermodynamic uncertainty relation in the three-level maser, arXiv:2103.07791 (2021).
- Li et al. [2019] J. Li, J. M. Horowitz, T. R. Gingrich, and N. Fakhri, Quantifying dissipation using fluctuating currents, Nat. Commun. 10, 1666 (2019).
- Manikandan et al. [2020] S. K. Manikandan, D. Gupta, and S. Krishnamurthy, Inferring entropy production from short experiments, Phys. Rev. Lett. 124, 120603 (2020).
- Van Vu et al. [2020] T. Van Vu, V. T. Vo, and Y. Hasegawa, Entropy production estimation with optimal current, Phys. Rev. E 101, 042138 (2020).
- Otsubo et al. [2020] S. Otsubo, S. Ito, A. Dechant, and T. Sagawa, Estimating entropy production by machine learning of short-time fluctuating currents, Phys. Rev. E 101, 062106 (2020).
- Redner [2001] S. Redner, A guide to first-passage processes (Cambridge University Press, 2001).
- Ding and Rangarajan [2004] M. Ding and G. Rangarajan, First passage time problem: A Fokker-Planck approach, in New Directions in Statistical Physics: Econophysics, Bioinformatics, and Pattern Recognition, edited by L. T. Wille (Springer Berlin Heidelberg, Berlin, Heidelberg, 2004) pp. 31–46.
- Friedman et al. [2017] H. Friedman, D. A. Kessler, and E. Barkai, Quantum walks: The first detected passage time problem, Phys. Rev. E 95, 032141 (2017).
- Yin et al. [2019] R. Yin, K. Ziegler, F. Thiel, and E. Barkai, Large fluctuations of the first detected quantum return time, Phys. Rev. Research 1, 033086 (2019).
- Saito and Dhar [2016] K. Saito and A. Dhar, Waiting for rare entropic fluctuations, EPL 114, 50004 (2016).
- Neri et al. [2017] I. Neri, E. Roldán, and F. Jülicher, Statistics of infima and stopping times of entropy production and applications to active molecular processes, Phys. Rev. X 7, 011019 (2017).
- Neri [2020] I. Neri, Second law of thermodynamics at stopping times, Phys. Rev. Lett. 124, 040601 (2020).
- Manzano et al. [2021] G. Manzano, D. Subero, O. Maillet, R. Fazio, J. P. Pekola, and E. Roldán, Thermodynamics of gambling demons, Phys. Rev. Lett. 126, 080603 (2021).
- Pal et al. [2021] A. Pal, S. Reuveni, and S. Rahav, Thermodynamic uncertainty relation for first-passage times on Markov chains, Phys. Rev. Research 3, L032034 (2021).
- Hiura and Sasa [2021] K. Hiura and S.-i. Sasa, Kinetic uncertainty relation on first-passage time for accumulated current, Phys. Rev. E 103, L050103 (2021).
- Barato and Seifert [2016] A. C. Barato and U. Seifert, Cost and precision of Brownian clocks, Phys. Rev. X 6, 041053 (2016).
- Mitchison [2019] M. T. Mitchison, Quantum thermal absorption machines: refrigerators, engines and clocks, Contemp. Phys. 0, 1 (2019).
- Hasegawa [2021b] Y. Hasegawa, Irreversibility, Loschmidt echo, and thermodynamic uncertainty relation, Phys. Rev. Lett. 127, 240602 (2021b).
- Nishiyama [2020] T. Nishiyama, A tight lower bound for the Hellinger distance with given means and variances, arXiv:2010.13548 (2020).
- Holevo [1973] A. S. Holevo, A generalization of the Rao–Cramér inequality, Teoriya Veroyatnostei i ee Primeneniya 18, 371 (1973).
- Kiukas et al. [2015] J. Kiukas, M. Guţă, I. Lesanovsky, and J. P. Garrahan, Equivalence of matrix product ensembles of trajectories in open quantum systems, Phys. Rev. E 92, 012132 (2015).
- Gammelmark and Mølmer [2014] S. Gammelmark and K. Mølmer, Fisher information and the quantum Cramér-Rao sensitivity limit of continuous measurements, Phys. Rev. Lett. 112, 170401 (2014).
- Mølmer [2015] K. Mølmer, Hypothesis testing with open quantum systems, Phys. Rev. Lett. 114, 040401 (2015).
- Paris [2009] M. G. A. Paris, Quantum estimation for quantum technology, Int. J. Quantum Inf. 7, 125 (2009).
- Liu et al. [2019] J. Liu, H. Yuan, X.-M. Lu, and X. Wang, Quantum Fisher information matrix and multiparameter estimation, J. Phys. A: Math. Theor. 53, 023001 (2019).
- Braunstein and Caves [1994] S. L. Braunstein and C. M. Caves, Statistical distance and the geometry of quantum states, Phys. Rev. Lett. 72, 3439 (1994).
- Helstrom [1976] C. W. Helstrom, Quantum detection and estimation theory (Academic Press, New York, 1976).
- Hotta and Ozawa [2004] M. Hotta and M. Ozawa, Quantum estimation by local observables, Phys. Rev. A 70, 022327 (2004).
- Deffner and Campbell [2017] S. Deffner and S. Campbell, Quantum speed limits: from Heisenberg’s uncertainty principle to optimal quantum control, J. Phys. A: Math. Theor. 50, 453001 (2017).
- Fröwis [2012] F. Fröwis, Kind of entanglement that speeds up quantum evolution, Phys. Rev. A 85, 052127 (2012).
- Taddei et al. [2013] M. M. Taddei, B. M. Escher, L. Davidovich, and R. L. de Matos Filho, Quantum speed limit for physical processes, Phys. Rev. Lett. 110, 050402 (2013).
- Dechant [2019] A. Dechant, Multidimensional thermodynamic uncertainty relations, J. Phys. A: Math. Theor. 52, 035001 (2019).
- Grimsmo [2015] A. L. Grimsmo, Time-delayed quantum feedback control, Phys. Rev. Lett. 115, 060402 (2015).
- Bartolotta and Deffner [2018] A. Bartolotta and S. Deffner, Jarzynski equality for driven quantum field theories, Phys. Rev. X 8, 011033 (2018).
- Ortega et al. [2019] A. Ortega, E. McKay, A. M. Alhambra, and E. Martín-Martínez, Work distributions on quantum fields, Phys. Rev. Lett. 122, 240604 (2019).
- Landi [2018] G. T. Landi, Quantum Information and Quantum Noise, Tech. Rep. (University of São Paulo, http://www.fmt.if.usp.br/~gtlandi/quantum-information-and.html, 2018).