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

Detailed fluctuation theorem from the one-time measurement scheme

Kenji Maeda Department of Physics, University of Massachusetts, Boston, Massachusetts 02125, USA    Tharon Holdsworth Department of Physics, University of Massachusetts, Boston, Massachusetts 02125, USA    Sebastian Deffner Department of Physics, University of Maryland, Baltimore County, Baltimore, Maryland 21250, USA    Akira Sone [email protected] Department of Physics, University of Massachusetts, Boston, Massachusetts 02125, USA
Abstract

We study the quantum fluctuation theorem in the one-time measurement (OTM) scheme, where the work distribution of the backward process has been lacking and which is considered to be more informative than the two-time measurement (TTM) scheme. We find that the OTM scheme is the quantum nondemolition TTM scheme, in which the final state is a pointer state of the second measurement whose Hamiltonian is conditioned on the first measurement outcome. Then, by clarifying the backward work distribution in the OTM scheme, we derive the detailed fluctuation theorem in the OTM scheme for the characteristic functions of the forward and backward work distributions, which captures the detailed information about the irreversibility and can be applied to quantum thermometry. We also verified our conceptual findings with the IBM quantum computer. Our result clarifies that the laws of thermodynamics at the nanoscale are dependent on the choice of the measurement and may provide experimentalists with a concrete strategy to explore laws of thermodynamics at the nanoscale by protecting quantum coherence and correlations.

One of the most significant conceptual factors distinguishing quantum physics and classical physics is measurement [1]. In quantum mechanics, measurements typically destroy quantum coherences and correlations that could be utilized as the resources for many quantum engineering tasks, such as quantum computing and quantum metrology. Compared to classical systems, one has many degrees of freedom in choosing the basis of the measurement based on their task on the quantum system. Particularly, the eigenbasis of the observable of the measurement apparatus is comprised of the so-called pointer states [2, 3, 4, 5, 6], which are immune to decoherence due to the corresponding measurement.

Quantum thermodynamics [7, 8, 9] is a rapidly growing field exploring the laws of thermodynamics from the perspective of quantum information science. Fluctuation theorems [10, 11] in both quantum and classical systems are regarded as one of the most significant laws to date [12] because many significant thermodynamic principles can be derived, such as the second law of thermodynamics [13] and response theory [14, 15]. The standard approach toward quantum fluctuation theorem is the so-called two-time measurement (TTM) scheme [16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31].

The TTM scheme is constructed by two energy projection measurements at the beginning and the end of a quantum process. In the standard setup of the time-varying Hamiltonian system, the initial state is prepared in the Gibbs state defined by its initial Hamiltonian H0H_{0}. Then, one performs an energy measurement on the initial state with H0H_{0}, which projects the system onto one of the eigenstates |Ei|E_{i}\rangle of H0H_{0} based on the initial measurement outcome EiE_{i}. Then, one evolves the system under the unitary operator UU during time τ\tau and measures the evolved state U|EiU|E_{i}\rangle with the final Hamiltonian HτH_{\tau}. Finally, the system will be projected again onto an eigenstate |Ej|E_{j}^{\prime}\rangle of HτH_{\tau} based on the final measurement outcome EjE_{j}^{\prime}.

The work performed on the system in a single realization is defined by the difference between the final and initial measurement outcome, WijEjEiW_{i\to j}\equiv E_{j}^{\prime}-E_{i} , which recovers the standard fluctuation theorem, also known as the TTM fluctuation theorem resembling the classical Jarzynksi equality [10]. Therefore, the TTM scheme can be regarded as a semiclassical approach, which has been experimentally implemented in various systems [32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42], including a demonstration on the DWave machine [30]. However, the second projection measurement usually destroys the quantum coherence and correlations generated through the dynamics, which means that the TTM cannot fully capture the peculiar features of the quantum systems when one analyzes its thermodynamic behaviors [29].

To address the thermodynamic contribution of quantum correlations, Ref. [43] proposed the so-called one-time measurement (OTM) scheme. In this scheme, the second measurement is considered to be avoided, and the work is determined by the energy difference conditioned on the initial energy measurement outcome. Within this paradigm, the corresponding Jarzynski equality includes the additional information contribution stemming from the quantum relative entropy of the conditional thermal state [44, 45] with respect to the Gibbs state defined by the final Hamiltonian. This additional term provides a tighter maximum work relation and captures the quantum coherence or correlations generated through the dynamics in the formalism. Therefore, the OTM scheme can be regarded as more informative than the TTM scheme. This has been elucidated in various contexts, including quantum thermometry [45], work as an external quantum observable [46], distinguishability of heat and work in an open quantum system [47], heat exchange [48], classical correspondence of the OTM scheme [44], quantum ergotropy  [49], and information production [50]. However, the backward process in the OTM scheme has not been considered yet, which has made the detailed quantum fluctuation theorem of the OTM scheme elusive.

In the present Letter, we first prove that the OTM scheme is the quantum nondemolition (QND) TTM scheme, where the pointer states of the second measurement (conditional Hamiltonian) are the evolved states conditioned on the initial measurement outcome. From this, we construct the backward work distribution and derive the detailed quantum fluctuation theorem of the OTM scheme, which we call OTM fluctuation theorem. Then, we propose a quantum circuit to compute the symmetric relation of the characteristic functions of the forward and backward work distributions. We explore the physical meaning of the OTM fluctuation theorem by associating it to the concept of irreversibility and demonstrate the potential application of the derived formalism to state preparation for low-temperature quantum thermometry. Finally, we verify the derived detailed fluctuation theorem with IBM quantum computer to demonstrate the experimental implementability of the OTM scheme. These results emphasize that the laws of quantum thermodynamics are strictly determined by the choice of measurements by the observers.

OTM detailed fluctuation theorem

Our first result is the derivation of the OTM fluctuation theorem. We consider a finite-dimensional closed quantum system described by a dd-dimensional Hilbert space. Let the initial state be a Gibbs state ρ0eqexp(βH0)/Z0\rho_{0}^{\text{eq}}\equiv\exp(-\beta H_{0})/Z_{0}, where H0H_{0} is the initial Hamiltonian and Z0tr{exp(βH0)}Z_{0}\equiv\mathrm{tr}\left\{\exp(-\beta H_{0})\right\} is the partition function. In a closed quantum system, the time evolution is described by a unitary operator UU. In the OTM scheme, the work for a single realization of the protocol is defined as

W~iEi|UHτU|EiEi,\widetilde{W}_{i}\equiv\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle-E_{i}, (1)

which also has been called conditional work [44]. This is the energy difference between the final energy conditioned on the initial measurement outcome and itself. Then, the forward conditional work distribution is simply given by [43]

P~f(W)=i=1deβEiZ0δ(WW~i),\widetilde{P}_{f}(W)=\sum_{i=1}^{d}\frac{e^{-\beta E_{i}}}{Z_{0}}\delta\left(W-\widetilde{W}_{i}\right), (2)

which is consistent with the exact average work

W=WP~f(W)𝑑W=tr{(UHτUH0)ρ0eq}\langle W\rangle=\int W\widetilde{P}_{f}(W)dW=\mathrm{tr}\left\{\left(U^{\dagger}H_{\tau}U-H_{0}\right)\rho_{0}^{\text{eq}}\right\} (3)

and yields the generalized Jarzynski equality [43]

eβWP~=Z~τZ0=eβΔFeS(ρ~τ||ρτeq).\langle e^{-\beta W}\rangle_{\widetilde{P}}=\frac{\widetilde{Z}_{\tau}}{Z_{0}}=e^{-\beta\Delta F}e^{-S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})}. (4)

In Eq. (4), the conditional partition function, Z~τi=1dexp(βEi|UHτU|Ei)\widetilde{Z}_{\tau}\equiv\sum_{i=1}^{d}\exp{\left(-\beta\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle\right)}, is the normalization factor used to construct the conditional thermal state,

ρ~τi=1deβEi|UHτU|EiZ~τU|EiEi|U.\widetilde{\rho}_{\tau}\equiv\sum_{i=1}^{d}\frac{e^{-\beta\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle}}{\widetilde{Z}_{\tau}}U|E_{i}\rangle\!\langle E_{i}|U^{\dagger}. (5)

Finally, S(ρ~τ||ρτeq)=tr{ρ~τlnρ~τ}tr{ρ~τlnρτeq}S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})=\mathrm{tr}\left\{\widetilde{\rho}_{\tau}\ln\widetilde{\rho}_{\tau}\right\}-\mathrm{tr}\left\{\widetilde{\rho}_{\tau}\ln\rho_{\tau}^{\text{eq}}\right\} is the quantum relative entropy of the conditional thermal state with respect to the Gibbs state ρτeqexp(βHτ)/Zτ\rho_{\tau}^{\text{eq}}\equiv\exp(-\beta H_{\tau})/Z_{\tau} of the final Hamiltonian HτH_{\tau}.

By comparing with the TTM scheme, we demonstrate that the OTM scheme is equivalent to the TTM scheme with a carefully chosen final Hamiltonian (conditional Hamiltonian) based on the information about the initial measurement outcome and the dynamics of the system. To see this point, let us define the conditional Hamiltonian GτG_{\tau},

Gτi=1dEi|UHτU|EiU|EiEi|U,G_{\tau}\equiv\sum_{i=1}^{d}\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle U|E_{i}\rangle\!\langle E_{i}|U^{\dagger}, (6)

where EiE_{i} is the eigenenergy of the initial Hamiltonian H0H_{0} with its corresponding eigenstate |Ei|E_{i}\rangle. At t=0t=0 we perform a projective energy measurement H0H_{0} on the system initially prepared in ρ0eq\rho_{0}^{\text{eq}}. Then, the post-measurement state will be projected onto |Ei|E_{i}\rangle with the corresponding energy EiE_{i}. After the evolution, the state is U|EiU|E_{i}\rangle. At t=τt=\tau we perform the second measurement GτG_{\tau}. Since the final state is a pointer state of GτG_{\tau}, it is not destroyed by the measurement, so that the observer obtains the final energy measurement outcome Ei|UHτU|Ei\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle, while the system remains as U|EiU|E_{i}\rangle. The corresponding quantum work is simply given by Eq. (1).

Then, the equivalent work distribution within the TTM paradigm, the forward work distribution is computed as

P~f(W)=i=1deβEi|H0|EiZ0|Ei|UU|Ei|2δ(WW~i),\widetilde{P}_{f}(W)=\sum_{i=1}^{d}\frac{e^{-\beta\langle E_{i}\hskip 1.0pt|H_{0}|\hskip 1.0ptE_{i}\rangle}}{Z_{0}}|\langle E_{i}\hskip 1.0pt|U^{\dagger}U|\hskip 1.0ptE_{i}\rangle|^{2}\delta(W-\widetilde{W}_{i}), (7)

which is identical to Eq. (2).

This is our first main result, namely we have that the OTM scheme is exactly the QND TTM scheme, in which the second projection measurement does not destroy the evolved state conditioned on the initial measurement outcome  (see Fig. 1 and [51]). We will now exploit this insight to construct the conditional work distribution for the backward process within the OTM paradigm.

Refer to caption
Figure 1: Comparison between the TTM and OTM scheme. In (a) the standard TTM scheme, in which the second energy measurement is the final Hamiltonian, the projection measurement projects the evolved state U|EiU|E_{i}\rangle onto |Ej|E_{j}^{\prime}\rangle the eigenstate of HτH_{\tau}. In (b) the OTM scheme, the final Hamiltonian Gτi=1dEi|UHτU|EiU|EiEi|UG_{\tau}\equiv\sum_{i=1}^{d}\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle U|E_{i}\rangle\!\langle E_{i}|U^{\dagger} is the conditional Hamiltonian with its pointer state U|EiU|E_{i}\rangle, which is equivalent to the evolved state conditioned on the initial measurement outcome. Therefore, this measurement preserves the state U|EiU|E_{i}\rangle. In this sense, this measurement is a QND measurement.

The backward process is initialized from the state ρ~τexp(βGτ)/Z~τ\widetilde{\rho}_{\tau}\equiv{\exp{\left(-\beta G_{\tau}\right)}}/{\widetilde{Z}_{\tau}}. After the backward evolution described by UU^{\dagger}, the measurement H0H_{0} is performed on the final state of the backward process. Since the final state is UU|Ei=|EiU^{\dagger}U|E_{i}\rangle=|E_{i}\rangle, which is a pointer state of H0H_{0}, H0H_{0} does not destroy the state. Moreover, the outcome is always EiE_{i}.

Then, by following the TTM scheme, the conditional work distribution of the backward process is given by

P~b(W)i=1deβEi|UHτU|EiZ~τδ(W+W~i).\widetilde{P}_{b}(-W)\equiv\sum_{i=1}^{d}\frac{e^{-\beta\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle}}{\widetilde{Z}_{\tau}}\delta(-W+\widetilde{W}_{i}). (8)

From Eq. (8) we now derive the fluctuation theorem [52, 53, 54] between the forward conditional work distribution and backward distribution in the characteristic function form.

The characteristic functions are defined as the Fourier transform of the work distributions

C~f(u)𝑑WP~f(W)eiuWC~b(u)𝑑WP~b(W)eiuW.\begin{split}\widetilde{C}_{f}(u)&\equiv\int dW\widetilde{P}_{f}(W)e^{iuW}\\ \widetilde{C}_{b}(u)&\equiv\int dW\widetilde{P}_{b}(-W)e^{-iuW}.\end{split} (9)

By applying the approach to the TTM scheme in Ref. [52, 53, 54], the characteristic functions are equivalent to

C~f(u)=tr{UeiuGτUeiuH0ρ0eq}C~b(u)=tr{UeiuH0UeiuGτρ~τ}.\begin{split}\widetilde{C}_{f}(u)&=\mathrm{tr}\left\{U^{\dagger}e^{iuG_{\tau}}Ue^{-iuH_{0}}\rho^{\text{eq}}_{0}\right\}\\ \widetilde{C}_{b}(u)&=\mathrm{tr}\left\{Ue^{iuH_{0}}U^{\dagger}e^{-iuG_{\tau}}\widetilde{\rho}_{\tau}\right\}.\end{split} (10)

Thus, we obtain the following symmetry relation 111In Supplemental Material [51], we provide the proof for the case that the initial state is also the conditional thermal state defined by G0i=1dψi|H0|ψi|ψiψi|G_{0}\equiv\sum_{i=1}^{d}\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle|\psi_{i}\rangle\!\langle\psi_{i}| with |ψi|\psi_{i}\rangle being not necessarily the eigenstate of H0H_{0}. Equation. (11) can be regarded as its corollary. , which is our second main result,

C~f(u)C~b(u+iβ)=Z~τZ0=eβΔFS(ρ~τ||ρτeq),\frac{\widetilde{C}_{f}(u)}{\widetilde{C}_{b}(-u+i\beta)}=\frac{\widetilde{Z}_{\tau}}{Z_{0}}=e^{-\beta\Delta F-S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})}, (11)

where

C~b(u+iβ)=tr{UeiuH0eβH0UeiuGτeβGτρ~τ}.\widetilde{C}_{b}(-u+i\beta)=\mathrm{tr}\left\{Ue^{-iuH_{0}}e^{-\beta H_{0}}U^{\dagger}e^{iuG_{\tau}}e^{\beta G_{\tau}}\widetilde{\rho}_{\tau}\right\}. (12)

The characteristic functions can be determined directly from quantum circuits, and hence our results permit the demonstration of the experimental implementability of the OTM scheme in the single qubit interferometry.

Single-qubit interferometry approach.

By employing the single-qubit interferometry approach developed in Refs. [53, 54], we now construct a quantum algorithm to verify Eq. (11). This indicates that the OTM scheme is experimentally implementable, which is our third main result.

Let us define |0(10)T|0\rangle\equiv\begin{pmatrix}1&0\end{pmatrix}^{T} and |1(01)T|1\rangle\equiv\begin{pmatrix}0&1\end{pmatrix}^{T}. We denote by 𝟙\openone the 2×22\times 2 identity matrix and XX, YY, ZZ as the usual Pauli matrices. Also, we write the Hadamard gate as 𝐇12(1111)\mathbf{H}\equiv\frac{1}{\sqrt{2}}\begin{pmatrix}1&1\\ 1&-1\end{pmatrix}. Then, the characteristic function of the forward process C~f(u)\widetilde{C}_{f}(u) can be computed by the quantum circuit depicted in Fig. 2. In this circuit, the ancilla qubit is initially prepared in |0|0\rangle. The target system is prepared in the Gibbs state ρ0eq\rho_{0}^{\text{eq}}. To obtain the characteristic function C~f(u)\widetilde{C}_{f}(u), we measure the output state of the ancilla qubit with XX and YY, whose expectation values become X=Re[C~f(u)]\langle X\rangle=\text{Re}[\widetilde{C}_{f}(u)] and Y=Im[C~f(u)]\langle Y\rangle=\text{Im}[\widetilde{C}_{f}(u)].

\Qcircuit@C=1em @R=1.2em \lstick|0⟩& \gateH \ctrl0 \qwx[1] \qw \ctrl0 \qwx[1] \meter   XX, YY
\lstickρ_0^eq\qw \gatee^-i u H_0 \gateU \gatee^iu G_τ \qw

Figure 2: Quantum circuit for computing C~f(u)\widetilde{C}_{f}(u). The expectation values of XX and YY obtained by measuring the final state of the ancilla qubit are Re[C~f(u)]\text{Re}[\widetilde{C}_{f}(u)] and Im[C~f(u)]\text{Im}[\widetilde{C}_{f}(u)], respectively.

Next, we consider the quantum circuit to compute the characteristic function of the backward process C~b(u+iβ)\widetilde{C}_{b}(-u+i\beta). From Eq. (12), we need to first decompose exp(βGτ)\exp(\beta G_{\tau}) and exp(βH0)\exp(-\beta H_{0}) with Pauli string {σk}k=1d2\{\sigma_{k}\}_{k=1}^{d^{2}}, where σ1\sigma_{1} is the d×dd\times d identity matrix 222When d=4d=4, the Pauli string is {𝟙,𝕏𝟙,𝕐𝟙,𝟙,𝟙𝕏,𝟙𝕐,𝟙,𝕏𝕏,𝕏𝕐,𝕏,𝕐𝕏,𝕐𝕐,𝕐,𝕏,𝕐,}\{\openone,X\otimes\openone,Y\otimes\openone,Z\otimes\openone,\openone\otimes X,\openone\otimes Y,\openone\otimes Z,X\otimes X,X\otimes Y,X\otimes Z,Y\otimes X,Y\otimes Y,Y\otimes Z,Z\otimes X,Z\otimes Y,Z\otimes Z\}, which has 42=164^{2}=16 elements, and these elements are the bases constructing any 4×44\times 4 matrix.. Here, note that 1dtr{σkσ}=δk\frac{1}{d}\mathrm{tr}\left\{\sigma_{k}\sigma_{\ell}\right\}=\delta_{k\ell} becomes Kronecker’s delta. Then, we can write exp(βH0)=k=1d2αk(0)σk\exp(-\beta H_{0})=\sum_{k=1}^{d^{2}}\alpha_{k}^{(0)}\sigma_{k} and exp(βGτ)=k=1d2αk(τ)σk\exp(\beta G_{\tau})=\sum_{k=1}^{d^{2}}\alpha_{k}^{(\tau)}\sigma_{k}, where αk(0)=1dtr{eβH0σk}\alpha_{k}^{(0)}=\frac{1}{d}\mathrm{tr}\left\{e^{-\beta H_{0}}\sigma_{k}\right\} and αk(τ)=1dtr{eβGτσk}\alpha_{k}^{(\tau)}=\frac{1}{d}\mathrm{tr}\left\{e^{\beta G_{\tau}}\sigma_{k}\right\}. Note that the coefficients {αk(0),αk(τ)}k=1d2\{\alpha_{k}^{(0)},\alpha_{k}^{(\tau)}\}_{k=1}^{d^{2}} are computable via a classical computer since we have full knowledge of H0H_{0}, HτH_{\tau}, and UU if dd is smaller. Therefore, we can write

C~b(u+iβ)=k,αk(0)α(τ)Fk,\widetilde{C}_{b}(-u+i\beta)=\sum_{k,\ell}\alpha_{k}^{(0)}\alpha_{\ell}^{(\tau)}F_{k\ell}, (13)

where we define

Fktr{UσkeiuH0UσeiuGτρ~τ}.F_{k\ell}\equiv\mathrm{tr}\left\{U\sigma_{k}e^{-iuH_{0}}U^{\dagger}\sigma_{\ell}e^{iuG_{\tau}}\widetilde{\rho}_{\tau}\right\}. (14)

Then, we can employ the quantum circuit depicted in Fig. 3 to compute FkF_{k\ell}. In this circuit, the ancilla qubit is prepared in |0|0\rangle. The target system is prepared in the conditional thermal state ρ~τ\widetilde{\rho}_{\tau}. Here, the expectation values become X=Re[Fk]\langle X\rangle=\text{Re}[F_{k\ell}] and Y=Im[Fk]\langle Y\rangle=\text{Im}[F_{k\ell}]. Given the fact that we have already known {αk(0),αk(τ)}k=1d2\{\alpha_{k}^{(0)},\alpha_{k}^{(\tau)}\}_{k=1}^{d^{2}} via a classical computer, from Eq. (13), we can finally obtain C~b(u+iβ)\widetilde{C}_{b}(-u+i\beta).

\Qcircuit@C=1em @R=1.2em \lstick|0⟩& \gateH \ctrl0 \qwx[1] \ctrl0 \qwx[1] \qw \ctrl0 \qwx[1] \ctrl0 \qwx[1] \meter   XX, YY
\lstick~ρ_τ\qw \gatee^iu G_τ\gateσ_ℓ \gateU^† \gatee^-iuH_0 \gateσ_k \qw

Figure 3: Quantum circuit for computing FkF_{k\ell}: The expectation values of XX and YY obtained by measuring the final state of the ancilla qubit are Re[Fk]\text{Re}[F_{k\ell}] and Im[Fk]\text{Im}[F_{k\ell}], respectively.

Physical meaning of OTM fluctuation theorem.

By considering the backward process, we can find that the OTM fluctuation theorem can capture the detailed information about the irreversibility and be applied to state preparation for quantum thermometry in the low-temperature limit. To quantify the irreversibility of a quantum process, we consider the Kullback-Leibler (KL) divergence D[P~f||P~b]D[\widetilde{P}_{f}||\widetilde{P}_{b}] of P~f(W)\widetilde{P}_{f}(W) with respect to P~b(W)\widetilde{P}_{b}(-W), which is defined as

D[P~f||P~b]dWP~f(W)ln(P~f(W)P~b(W)).\displaystyle D[\widetilde{P}_{f}||\widetilde{P}_{b}]\equiv\int dW\widetilde{P}_{f}(W)\ln\left(\frac{\widetilde{P}_{f}(W)}{\widetilde{P}_{b}(-W)}\right). (15)

From Eqs. (2) and (8), we obtain [51]

D[P~f||P~b]=S(ρ0eq)+βtr{Uρ0eqUHτ}+lnZ~τ,\displaystyle\!\!\!\!D[\widetilde{P}_{f}||\widetilde{P}_{b}]\!=\!-S(\rho_{0}^{\text{eq}})+\beta\mathrm{tr}\left\{U\rho_{0}^{\text{eq}}U^{\dagger}H_{\tau}\right\}+\ln\widetilde{Z}_{\tau}, (16)

where S(ρ0eq)tr{ρ0eqlnρ0eq}S(\rho_{0}^{\text{eq}})\equiv-\mathrm{tr}\left\{\rho_{0}^{\text{eq}}\ln\rho_{0}^{\text{eq}}\right\} is the von-Neumann entropy of ρ0eq\rho_{0}^{\text{eq}}. Given that the exact averaged work Wtr{Uρ0eqUHτ}tr{ρ0eqH0}\langle W\rangle\equiv\mathrm{tr}\left\{U\rho_{0}^{\text{eq}}U^{\dagger}H_{\tau}\right\}-\mathrm{tr}\left\{\rho_{0}^{\text{eq}}H_{0}\right\}, from Eq. (11), the excess work WexWΔF\langle W_{\text{ex}}\rangle\equiv\langle W\rangle-\Delta F can be written as

βWex=D[P~f||P~b]+S(ρ~τ||ρτeq).\displaystyle\beta\langle W_{\text{ex}}\rangle=D[\widetilde{P}_{f}||\widetilde{P}_{b}]+S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}}). (17)

This means that the excess work Wex\langle W_{\text{ex}}\rangle is a sum of the KL divergence D[P~f||P~b]D[\widetilde{P}_{f}||\widetilde{P}_{b}], which characterizes the irreversible process, and βS(ρ~τ||ρτeq)\beta S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}}), which is the energy dissipated into the heat bath when the system is thermalized from ρ~τ\widetilde{\rho}_{\tau}. Therefore, βS(ρ~τ||ρτeq)\beta S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}}) can be interpreted as a heatlike quantity.

Furthermore, D[P~f||P~b]D[\widetilde{P}_{f}||\widetilde{P}_{b}] can be employed in quantum thermometry in the low-temperature limit. In Ref. [45], it was demonstrated that the conditional thermal state ρ~τ\widetilde{\rho}_{\tau} can outperform the Gibbs state ρτeq\rho_{\tau}^{\text{eq}} in the low-temperature limit. Therefore, preparing ρ~τ\widetilde{\rho}_{\tau} is a desired task for quantum thermometry. First the excess work can be written as βWex=S(Uρ0eqU||ρτeq)\beta\langle W_{\text{ex}}\rangle=S(U\rho_{0}^{\text{eq}}U^{\dagger}||\rho_{\tau}^{\text{eq}}) [57, 58, 59]. In Ref. [45], we derived the so-called thermodynamic triangle equality S(Uρ0eqU||ρ~τ)+S(ρ~τ||ρτeq)=S(Uρ0eqU||ρτeq)S(U\rho_{0}^{\text{eq}}U^{\dagger}||\widetilde{\rho}_{\tau})+S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})=S(U\rho_{0}^{\text{eq}}U^{\dagger}||\rho_{\tau}^{\text{eq}}). Therefore, from Eq. (17), we obtain

D[P~f||P~b]=S(Uρ0eqU||ρ~τ),\displaystyle D[\widetilde{P}_{f}||\widetilde{P}_{b}]=S(U\rho_{0}^{\text{eq}}U^{\dagger}||\widetilde{\rho}_{\tau}), (18)

which measures the distinguishability of the exact final state Uρ0eqUU\rho_{0}^{\text{eq}}U^{\dagger} and the conditional thermal state ρ~τ\widetilde{\rho}_{\tau}. This quantity can be used to design the unitary process UU that minimizes S(Uρ0eqU||ρ~τ)S(U\rho_{0}^{\text{eq}}U^{\dagger}||\widetilde{\rho}_{\tau}) for the final exact state Uρ0eqUU\rho_{0}^{\text{eq}}U^{\dagger} to be closer to ρ~τ\widetilde{\rho}_{\tau}. Also, note that the distinguishability measure S(Uρ0eqU||ρ~τ)S(U\rho_{0}^{\text{eq}}U^{\dagger}||\widetilde{\rho}_{\tau}) can be computed by a quantum computer. From Eqs. (11) and (16), we have

D[P~f||P~b]=βW+ln(C~f(u)C~b(u+iβ)),\displaystyle D[\widetilde{P}_{f}||\widetilde{P}_{b}]=\beta\langle W\rangle+\ln\left(\frac{\widetilde{C}_{f}(u)}{\widetilde{C}_{b}(-u+i\beta)}\right), (19)

where W\langle W\rangle, C~f(u)\widetilde{C}_{f}(u), and C~b(u+iβ)\widetilde{C}_{b}(-u+i\beta) can be computed by a quantum computer.

Finally, we emphasize that these analyses are hard to conduct within the TTM scheme [31]. Therefore, our detailed fluctuation demonstrates an additional advantage of the OTM scheme.

Verification with IBM quantum computers

To conclude our analysis, we employ the IBM cloud-based quantum computer [60] to verify the detailed fluctuation theorem (11). Our setup is the following. The initial Hamiltonian H0H_{0} is H0=ω(Z𝟙+𝟙)H_{0}=\omega(Z\otimes\openone+\openone\otimes Z) with the corresponding eigenbasis |E1=(1000)T|E_{1}\rangle=\begin{pmatrix}1&0&0&0\end{pmatrix}^{T}, |E2=(0100)T|E_{2}\rangle=\begin{pmatrix}0&1&0&0\end{pmatrix}^{T}, |E3=(0010)T|E_{3}\rangle=\begin{pmatrix}0&0&1&0\end{pmatrix}^{T}, and |E4=(0001)T|E_{4}\rangle=\begin{pmatrix}0&0&0&1\end{pmatrix}^{T}. The final Hamiltonian is set to be Hτ=J(XX)H_{\tau}=J(X\otimes X). The unitary operator that describes the evolution is set as U=exp(iΩτ2(Y𝟙+𝟙𝕐))U=\exp\left(-i\frac{\Omega\tau}{2}(Y\otimes\openone+\openone\otimes Y)\right).

For the initial Gibbs state preparation, we consider the decomposition of the input mixed state. This is because we can only prepare pure states on the IBM quantum computers and {|Ei}i=14\{|E_{i}\rangle\}_{i=1}^{4} can be prepared. For the weights {exp(βEi)/Z0}i=14\{\exp(-\beta E_{i})/Z_{0}\}_{i=1}^{4}, since we already know the initial Hamiltonian H0H_{0}, we assume that the weights are also known and we can compute C~f(u)\widetilde{C}_{f}(u).

For the backward process, we need to prepare the conditional thermal state ρ~τ\widetilde{\rho}_{\tau} as the initial state. In our simulation, we suppose that we already know UU; therefore, we can prepare {U|Ei}i=14\{U|E_{i}\rangle\}_{i=1}^{4} with the quantum computer. Similarly, because we assume that UU, H0H_{0} and HτH_{\tau} are known, the weights {exp(βEi|UHτU|Ei)/Z~τ}i=14\{\exp\left(-\beta\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle\right)/\widetilde{Z}_{\tau}\}_{i=1}^{4} are considered to be also known, which we use to compute FkF_{k\ell}. Thus the coefficients αk(0)\alpha_{k}^{(0)} and α(τ)\alpha_{\ell}^{(\tau)} are also regarded as known values, which enables one to compute C~b(u+iβ)\widetilde{C}_{b}(-u+i\beta).

By setting the parameters as β=0.5\beta=0.5, ω=2\omega=2, Ω=3\Omega=3, J=1J=1, and τ=π/4\tau=\pi/4, we obtain the theoretical value of the ratio

RtrueC~f(u)C~b(u+iβ)=0.433167R_{\text{true}}\equiv\frac{\widetilde{C}_{f}(u)}{\widetilde{C}_{b}(-u+i\beta)}=0.433167 (20)

for any uu. Here, we particularly focus on the case u=1u=1, and verify Eq. (20) with the IBM cloud-based quantum computer [60].

To determine X\langle X\rangle and Y\langle Y\rangle, for each quantum circuit in Figs. 2 and 3, we perform the single-shot measurement 2000020000 times. The median of the errors of the gates and the single-qubit readout error in our setup are around 10410210^{-4}\sim 10^{-2} with the T1T_{1} and T2T_{2} ranging from around 17 μs232μs\upmu\text{s}\sim 232\upmu\text{s} (for complete information of the IBM machine, refer to the Supplemental Material [51]). Because of the error, the computed ratio becomes complex; therefore, we consider the absolute value of the ratio R|C~f(1)/C~b(1+0.5i)|R\equiv|\widetilde{C}_{f}(1)/\widetilde{C}_{b}(-1+0.5i)|. To obtain more precise values, we run the whole process NN times (number of trials) and compare the true value with the average value RN1Nj=1NRj\langle R\rangle_{N}\equiv\frac{1}{N}\sum_{j=1}^{N}R_{j}, where RjR_{j} is the value of RR at the jjth trial. Then, we compute the error rate as eN=|1RN/Rtrue|×100[%]e_{N}=|1-\langle R\rangle_{N}/R_{\text{true}}|\times 100~{}[\%] for each NN.

We have achieved a very high accuracy in our simulation. In Fig. 4, we plot the relation between RN\langle R\rangle_{N} for each number of trials N=10,15,20,,100N=10,15,20,\cdots,100, where the error bars represent the 99% confidence interval [61]. As we can see, as we increase the number of trials, the average value converges to the true value with explicit plateau starting from N=75N=75. When N=100N=100, R100\langle R\rangle_{100} records 0.433706.

Refer to caption
Figure 4: Verification with IBM quantum computers. The dashed line is the exact value Rtrue=0.433167R_{\text{true}}=0.433167. The error bars represent 99%99\% confidence interval. As we increase the number of trials NN, the averaged ratios RN\langle R\rangle_{N} approach the exact value with a clear plateau starting from N=75N=75. When N=100N=100, we have R100=0.433706\langle R\rangle_{100}=0.433706.

In Fig. 5, we show the relation between the error rate eNe_{N} and the number of trials NN. As we can see, as the number of trials increases, the error rate becomes smaller. Actually, when N=100N=100, e100e_{100} records around 0.12%0.12\%, which is accurate enough to claim that the OTM fluctuation theorem Eq. (11) is verified with the IBM quantum computer.

Refer to caption
Figure 5: Error rate vs number of trials. As we increase the number of trials NN, the error rate eNe_{N} becomes smaller. When N=100N=100, we have e1000.12%e_{100}\simeq 0.12\%.

Conclusion.

In conclusion, we have derived the detailed fluctuation theorem of the OTM scheme by clarifying the backward work distribution. This has been enabled by the insight that the OTM scheme can be regarded as a QND TTM scheme, where the second measurement is constructed by the pointer states conditioned on the initial energy measurement outcome. We have related its physical meaning to the irreversibility and quantum thermometry in the low-temperature limit. We have also demonstrated its experimental implementability and we have verified the derived fluctuation theorem on the IBM quantum computer by introducing the corresponding quantum circuit to compute the symmetric relation between the characteristic functions of the forward and backward work distributions. These results not only provide the solutions to the open problems regarding the OTM scheme, but also clarify that the laws of thermodynamics at the nanoscale are strictly dependent on the choice of the measurement of the observer. From a practical point of view, these results provide experimentalists with a concrete strategy to study laws of thermodynamics at the nanoscale by protecting quantum coherence and correlations.

Acknowledgements.
This work is supported by the NSF under Grant No. MPS-2328774. A.S. is grateful to S. Endo for helpful discussions. K.M. is supported by the Goldwater scholarship. T.H. is supported by the graduate study program at the University of Massachusetts Boston. S.D. acknowledges support from the U.S. National Science Foundation under Grant No. DMR-2010127 and the John Templeton Foundation under Grant No. 62422.

References

  • Nielsen and Chuang [2010] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information: 10th Anniversary Edition, 10th ed. (Cambridge University Press, New York, NY, USA, 2010).
  • Zurek [1981] W. H. Zurek, Pointer basis of quantum apparatus: Into what mixture does the wave packet collapse?, Phys. Rev. D 24, 1516 (1981).
  • Schlosshauer [2007] M. A. Schlosshauer, Decoherence: and the quantum-to-classical transition (Springer Science & Business Media, 2007).
  • Brasil and de Castro [2015] C. A. Brasil and L. A. de Castro, Understanding the pointer states, Eur. J. Phys. 36, 065024 (2015).
  • Touil et al. [2022] A. Touil, B. Yan, D. Girolami, S. Deffner, and W. H. Zurek, Eavesdropping on the decohering environment: quantum Darwinism, amplification, and the origin of objective classical reality, Phys. Rev. Lett. 128, 010401 (2022).
  • Zurek [2003] W. H. Zurek, Decoherence, einselection, and the quantum origins of the classical, Rev. Mod. Phys. 75, 715 (2003).
  • Deffner and Campbell [2019] S. Deffner and S. Campbell, Quantum Thermodynamics (Morgan and Claypool Publishers, San Rafael, 2019).
  • Vinjanampathy and Anders [2016] S. Vinjanampathy and J. Anders, Quantum thermodynamics, Contemp. Phys. 57, 545 (2016).
  • Binder et al. [2019] F. Binder, L. A. Correa, C. Gogolin, J. Anders, and G. Adesso, Thermodynamics in the Quantum Regime (Springer, 2019).
  • Jarzynski [1997] C. Jarzynski, Nonequilibrium equality for free energy differences, Phys. Rev. Lett. 78, 2690 (1997).
  • Crooks [1999] G. E. Crooks, Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences, Phys. Rev. E 60, 2721 (1999).
  • de Zárate [2011] J. M. O. de Zárate, Interview with Michael E. Fisher, Europhysics News 42, 14 (2011).
  • Jarzynski [2011] C. Jarzynski, Equalities and inequalities: Irreversibility and the second law of thermodynamics at the nanoscale, Annu. Rev. Condens. Matter Phys. 2, 329 (2011).
  • Andrieux and Gaspard [2008] D. Andrieux and P. Gaspard, Quantum work relations and response theory, Phys. Rev. Lett. 100, 230404 (2008).
  • Andrieux et al. [2009] D. Andrieux, P. Gaspard, T. Monnai, and S. Tasaki, The fluctuation theorem for currents in open quantum systems, New J. Phys. 11, 043014 (2009).
  • [16] H. Tasaki, Jarzynski relations for quantum systems and some applications, arXiv preprint arXiv:cond-mat/0009244 .
  • [17] J. Kurchan, A quantum fluctuation theorem, arXiv preprint arXiv:cond-mat/0007360 .
  • Talkner et al. [2007] P. Talkner, E. Lutz, and P. Hänggi, Fluctuation theorems: Work is not an observable, Phys. Rev. E 75, 050102(R) (2007).
  • Jarzynski and Wójcik [2004] C. Jarzynski and D. K. Wójcik, Classical and quantum fluctuation theorems for heat exchange, Phys. Rev. Lett. 92, 230602 (2004).
  • Morikuni et al. [2017] Y. Morikuni, H. Tajima, and N. Hatano, Quantum jarzynski equality of measurement-based work extraction, Phys. Rev. E 95, 032147 (2017).
  • Rastegin [2013] A. E. Rastegin, Non-equilibrium equalities with unital quantum channels, J. Stat. Mech. Theo. Exp. 2013, P06016 (2013).
  • Jarzynski et al. [2015] C. Jarzynski, H. Quan, and S. Rahav, Quantum-classical correspondence principle for work distributions, Phys. Rev. X 5, 031038 (2015).
  • Zhu et al. [2016] L. Zhu, Z. Gong, B. Wu, and H. T. Quan, Quantum-classical correspondence principle for work distributions in a chaotic system, Phys. Rev. E 93, 062108 (2016).
  • [24] R. Pan, Z. Fei, T. Qiu, J.-N. Zhang, and H. Quan, Quantum-classical correspondence of work distributions for initial states with quantum coherence, arXiv preprint arXiv:1904.05378 .
  • Kafri and Deffner [2012] D. Kafri and S. Deffner, Holevo’s bound from a general quantum fluctuation theorem, Phys. Rev. A 86, 044302 (2012).
  • Goold et al. [2015] J. Goold, M. Paternostro, and K. Modi, Nonequilibrium quantum landauer principle, Phys. Rev. Lett. 114, 060602 (2015).
  • Rastegin and Życzkowski [2014] A. E. Rastegin and K. Życzkowski, Jarzynski equality for quantum stochastic maps, Phys. Rev. E 89, 012127 (2014).
  • [28] J. Goold and K. Modi, Energetic fluctuations in an open quantum process, arXiv preprint arXiv: 1407.4618 .
  • Perarnau-Llobet et al. [2017] M. Perarnau-Llobet, E. Bäumer, K. V. Hovhannisyan, M. Huber, and A. Acin, No-go theorem for the characterization of work fluctuations in coherent quantum systems, Phys. Rev. Lett. 118, 070601 (2017).
  • Gardas and Deffner [2018] B. Gardas and S. Deffner, Quantum fluctuation theorem for error diagnostics in quantum annealers, Sci. Rep. 8, 17191 (2018).
  • Kiely et al. [2023] A. Kiely, E. O’Connor, T. Fogarty, G. T. Landi, and S. Campbell, Entropy of the quantum work distribution, Phys. Rev. Research 5, L022010 (2023).
  • Huber et al. [2008] G. Huber, F. Schmidt-Kaler, S. Deffner, and E. Lutz, Employing Trapped Cold Ions to Verify the Quantum Jarzynski Equality, Phys. Rev. Lett. 101, 070403 (2008).
  • Smith et al. [2018] A. Smith, Y. Lu, S. An, X. Zhang, J.-N. Zhang, Z. Gong, H. T. Quan, C. Jarzynski, and K. Kim, Verification of the quantum nonequilibrium work relation in the presence of decoherence, New J. Phys. 20, 013008 (2018).
  • An et al. [2015] S. An, J.-N. Zhang, M. Um, D. Lv, Y. Lu, J. Zhang, Z.-Q. Yin, H. T. Quan, and K. Kim, Experimental test of the quantum jarzynski equality with a trapped-ion system, Nat. Phys. 11, 193 (2015).
  • Batalhão et al. [2014] T. B. Batalhão, A. M. Souza, L. Mazzola, R. Auccaise, R. S. Sarthour, I. S. Oliveira, J. Goold, G. DeChiara, M. Paternostro, and R. M. Serra, Experimental reconstruction of work distribution and study of fluctuation relations in a closed quantum system, Phys. Rev. Lett. 113, 140601 (2014).
  • Hernández-Gómez et al. [2020] S. Hernández-Gómez, S. Gherardini, F. Poggiali, F. S. Cataliotti, A. Trombettoni, P. Cappellaro, and N. Fabbri, Experimental test of exchange fluctuation relations in an open quantum system, Phys. Rev. Research 2, 023327 (2020).
  • Hernández-Gómez et al. [2021] S. Hernández-Gómez, N. Staudenmaier, M. Campisi, and N. Fabbri, Experimental test of fluctuation relations for driven open quantum systems with an NV center, New J. Phys. 23, 065004 (2021).
  • Collin et al. [2005] D. Collin, F. Ritort, C. Jarzynski, S. B. Smith, I. Tinoco, and C. Bustamante, Verification of the crooks fluctuation theorem and recovery of rna folding free energies, Nature 437, 231 (2005).
  • Parrondo et al. [2015] J. M. R. Parrondo, J. M. Horowitz, and T. Sagawa, Thermodynamics of information, Nat. Phys. 11, 131 (2015).
  • Toyabe et al. [2010] S. Toyabe, T. Sagawa, M. Ueda, E. Muneyuki, and M. Sano, Experimental demonstration of information-to-energy conversion and validation of the generalized jarzynski equality, Nat. Phys. 6, 988 (2010).
  • Zhang et al. [2018] Z. Zhang, T. Wang, L. Xiang, Z. Jia, P. Duan, W. Cai, Z. Zhan, Z. Zong, J. Wu, L. Sun, Y. Yin, and G. Guo, Experimental demonstration of work fluctuations along a shortcut to adiabaticity with a superconducting xmon qubit, New J. Phys. 20, 085001 (2018).
  • [42] D. Hahn, M. Dupont, M. Schmitt, D. J. Luitz, and M. Bukov, Verification of the Quantum Jarzynski Equality on Digital Quantum Computers, arXiv preprint arXiv:2207.14313 .
  • Deffner et al. [2016] S. Deffner, J. P. Paz, and W. H. Zurek, Quantum work and the thermodynamic cost of quantum measurements, Phys. Rev. E 94, 010103(R) (2016).
  • Sone and Deffner [2021a] A. Sone and S. Deffner, Jarzynski equality for stochastic conditional work, J. Stat. Phys 183, 11 (2021a).
  • Sone et al. [a] A. Sone, D. Soares-Pinto, and S. Deffner, Conditional quantum thermometry–enhancing precision by measuring less, arXiv preprint arXiv: 2304.13595  (a).
  • Beyer et al. [2020] K. Beyer, K. Luoma, and W. T. Strunz, Work as an external quantum observable and an operational quantum work fluctuation theorem, Phys. Rev. Research 2, 033508 (2020).
  • Sone et al. [2020] A. Sone, Y.-X. Liu, and P. Cappellaro, Quantum Jarzynski equality in open quantum systems from the one-time measurement scheme, Phys. Rev. Lett. 125, 060602 (2020).
  • Sone et al. [b] A. Sone, D. O. Soares-Pinto, and S. Deffner, Exchange fluctuation theorems for strongly interacting quantum pumps, arXiv preprint arXiv:2209.12927  (b).
  • Sone and Deffner [2021b] A. Sone and S. Deffner, Quantum and classical ergotropy from relative entropies, Entropy 23, 1107 (2021b).
  • Sone et al. [2023] A. Sone, N. Yamamoto, T. Holdsworth, and P. Narang, Jarzynski-like Equality of Nonequilibrium Information Production Based on Quantum Cross Entropy, Phys. Rev. Research 5, 023039 (2023).
  • [51] See Supplemental Material for (1) the explanation of why OTM scheme is a QND TTM scheme, (2) the derivation of the detailed fluctuation theorem and the symmetric relation for the general case that the initial state is also the conditional thermal state, (3) the derivation of the KL divergence D[P~f||P~b]D[\widetilde{P}_{f}||\widetilde{P}_{b}] , (4) the details of the experimental verification of the symmetric relation in 2-qubit case and (5) the detailed information about the IBM machines that we used.
  • Campisi et al. [2011] M. Campisi, P. Hänggi, and P. Talkner, Colloquium: Quantum fluctuation relations: Foundations and applications, Rev. Mod. Phys. 83, 771 (2011).
  • Mazzola et al. [2013] L. Mazzola, G. De Chiara, and M. Paternostro, Measuring the characteristic function of the work distribution, Phys. Rev. Lett. 110, 230602 (2013).
  • Dorner et al. [2013] R. Dorner, S. R. Clark, L. Heaney, R. Fazio, J. Goold, and V. Vedral, Extracting quantum work statistics and fluctuation theorems by single-qubit interferometry, Phys. Rev. Lett. 110, 230601 (2013).
  • Note [1] In Supplemental Material [51], we provide the proof for the case that the initial state is also the conditional thermal state defined by G0\sum@\slimits@i=1dψi|H0|ψi|ψiψi|G_{0}\equiv\sum@\slimits@_{i=1}^{d}\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle|\psi_{i}\rangle\!\langle\psi_{i}| with |ψi|\psi_{i}\rangle being not necessarily the eigenstate of H0H_{0}. Equation. (11) can be regarded as its corollary.
  • Note [2] When d=4d=4, the Pauli string is {𝟙,𝕏𝟙,𝕐𝟙,𝟙,𝟙𝕏,𝟙𝕐,𝟙,𝕏𝕏,𝕏𝕐,𝕏,𝕐𝕏,𝕐𝕐,𝕐,𝕏,𝕐,}\{\openone,X\otimes\openone,Y\otimes\openone,Z\otimes\openone,\openone\otimes X,\openone\otimes Y,\openone\otimes Z,X\otimes X,X\otimes Y,X\otimes Z,Y\otimes X,Y\otimes Y,Y\otimes Z,Z\otimes X,Z\otimes Y,Z\otimes Z\}, which has 42=164^{2}=16 elements, and these elements are the bases constructing any 4×44\times 4 matrix.
  • Deffner and Lutz [2010] S. Deffner and E. Lutz, Generalized Clausius Inequality for Nonequilibrium Quantum Processes, Phys. Rev. Lett. 105, 170402 (2010).
  • Kawai et al. [2007] R. Kawai, J. M. R. Parrondo, and C. V. den Broeck, Dissipation: The phase-space perspective, Phys. Rev. Lett. 98, 080602 (2007).
  • Vaikuntanathan and Jarzynski [2009] S. Vaikuntanathan and C. Jarzynski, Dissipation and lag in irreversible processes, Europhys. Lett. 87, 60005 (2009).
  • Cross [2018] A. Cross, The IBM Q experience and QISKit open-source quantum computing software, in APS March meeting abstracts, Vol. 2018 (2018) pp. L58–003.
  • Dekking et al. [2005] F. M. Dekking, C. Kraaikamp, H. P. Lopuhaä, and L. E. Meester, A Modern Introduction to Probability and Statistics: Understanding why and how, Vol. 488 (Springer, 2005).

Supplementary Material for
“Detailed Fluctuation Theorem from the One-Time Measurement Scheme”

Appendix A OTM scheme as a QND TTM scheme

In the OTM scheme, focusing on the forward process, when the final measurement is

Gτi=1dEi|UGτU|EiU|EiEi|U\displaystyle G_{\tau}\equiv\sum_{i=1}^{d}\langle E_{i}\hskip 1.0pt|U^{\dagger}G_{\tau}U|\hskip 1.0ptE_{i}\rangle U|E_{i}\rangle\!\langle E_{i}|U^{\dagger} (S1)

and the state before the final measurement is U|EiU|E_{i}\rangle, the state does not change due to GτG_{\tau} as [1]

GτU|EiEi|UGτGτU|Ei=U|Ei.\displaystyle\frac{G_{\tau}U|E_{i}\rangle}{\sqrt{\langle E_{i}\hskip 1.0pt|U^{\dagger}G_{\tau}^{\dagger}G_{\tau}U|\hskip 1.0ptE_{i}\rangle}}=U|E_{i}\rangle. (S2)

This demonstrates that GτG_{\tau} is a QND measurement, so that OTM scheme can be classified as the QND TTM scheme.

Appendix B Derivation of the detailed fluctuation theorem and the symmetric relation

To generalize our formalism, we consider the case that the initial state is also a conditional thermal state

ρ~0i=1deβψi|H0|ψiZ~0|ψiψi|.\widetilde{\rho}_{0}\equiv\sum_{i=1}^{d}\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{0}}|\psi_{i}\rangle\!\langle\psi_{i}|. (S3)

Here, note that |ψi|\psi_{i}\rangle is not necessarily the eigenstate of H0H_{0}. Defining

G0i=1dψi|H0|ψi|ψiψi|,G_{0}\equiv\sum_{i=1}^{d}\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle|\psi_{i}\rangle\!\langle\psi_{i}|, (S4)

where ψi|H0|ψi\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle and |ψi|\psi_{i}\rangle are the eigenvalue and the corresponding eigenstate of G0G_{0}, we can write

ρ~0=eβG0Z~0\widetilde{\rho}_{0}=\frac{e^{-\beta G_{0}}}{\widetilde{Z}_{0}} (S5)

with Z~0tr{exp(βG0)}\widetilde{Z}_{0}\equiv\mathrm{tr}\left\{\exp(-\beta G_{0})\right\} the normalization factor. Here, note that when |ψi=|Ei|\psi_{i}\rangle=|E_{i}\rangle, we have G0=H0G_{0}=H_{0} and ρ~0=ρ0eq\widetilde{\rho}_{0}=\rho_{0}^{\text{eq}}.

We first consider the following TTM scheme. At time t=0t=0, we measure the system with G0G_{0}. Suppose that the outcome is ψi|H0|ψi\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle; therefore, the post-measurement state will be projected onto |ψi|\psi_{i}\rangle. After the evolution UU, we obtain a final state U|ψiU|\psi_{i}\rangle. At time t=τt=\tau we perform a measurement GτG_{\tau} defined as

Gτi=1dψi|UHτU|ψiU|ψiψi|U.G_{\tau}\equiv\sum_{i=1}^{d}\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle U|\psi_{i}\rangle\!\langle\psi_{i}|U^{\dagger}. (S6)

Here, note that GτG_{\tau} does not destroy the final state U|ψiU|\psi_{i}\rangle because U|ψiU|\psi_{i}\rangle itself is the pointer state of GτG_{\tau}. Therefore, the outcome of GτG_{\tau} is always ψi|UHτU|ψi\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle. We define

W~iψi|UHτU|ψiψi|H0|ψi,\widetilde{W}_{i}\equiv\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle-\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle, (S7)

which is a random variable determined in a single measurement trajectory. Then, applying the way of constructing the work distribution of the forward process in the TTM scheme, for the forward process, we can write

P~f(W)=i=1deβψi|H0|ψiZ~0|ψi|UU|ψi|2δ(WW~i)=i=1deβψi|H0|ψiZ~0δ(WW~i).\begin{split}\widetilde{P}_{f}(W)&=\sum_{i=1}^{d}\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{0}}|\langle\psi_{i}\hskip 1.0pt|U^{\dagger}U|\hskip 1.0pt\psi_{i}\rangle|^{2}\delta(W-\widetilde{W}_{i})\\ &=\sum_{i=1}^{d}\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{0}}\delta(W-\widetilde{W}_{i}).\end{split} (S8)

For the backward process, we start from the state

ρ~τeβGτZ~τ=i=1deβψi|UHτU|ψiZ~τU|ψiψi|U.\!\!\!\widetilde{\rho}_{\tau}\equiv\frac{e^{-\beta G_{\tau}}}{\widetilde{Z}_{\tau}}=\sum_{i=1}^{d}\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}U|\psi_{i}\rangle\!\langle\psi_{i}|U^{\dagger}. (S9)

Then, after the backward evolution described by UU^{\dagger}, we perform the measurement G0G_{0} on the final state of the backward process. Again, since the final state is UU|ψi=|ψiU^{\dagger}U|\psi_{i}\rangle=|\psi_{i}\rangle, which is the pointer state of G0G_{0}, G0G_{0} does not destroy the state |ψi|\psi_{i}\rangle. Therefore, the outcome is always ψi|H0|ψi\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle. Then, the work distribution of the backward process is given by

P~b(W)i=1deβψi|UHτU|ψiZ~τ|ψi|UU|ψi|2δ(W+W~i)=i=1deβψi|UHτU|ψiZ~τδ(W+W~i).\begin{split}\widetilde{P}_{b}(-W)\equiv&\sum_{i=1}^{d}\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}|\langle\psi_{i}\hskip 1.0pt|U^{\dagger}U|\hskip 1.0pt\psi_{i}\rangle|^{2}\delta(-W+\widetilde{W}_{i})\\ =&\sum_{i=1}^{d}\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}\delta(-W+\widetilde{W}_{i}).\end{split} (S10)

Since

P~f(W)=i=1deβψi|H0|ψiZ~0δ(WW~i)=Z~τZ~0i=1deβ(Wψi|UHτU|ψi)Z~τδ(W+W~i)=Z~τZ~0eβWP~b(W),\begin{split}\widetilde{P}_{f}(W)&=\sum_{i=1}^{d}\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{0}}\delta(W-\widetilde{W}_{i})\\ &=\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}\sum_{i=1}^{d}\frac{e^{\beta(W-\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle)}}{\widetilde{Z}_{\tau}}\delta(-W+\widetilde{W}_{i})\\ &=\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}e^{\beta W}\widetilde{P}_{b}(-W),\end{split} (S11)

which yields

P~f(W)P~b(W)eβW=Z~τZ~0.\frac{\widetilde{P}_{f}(W)}{\widetilde{P}_{b}(-W)}e^{-\beta W}=\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}. (S12)

Let us compute the following quantum relative entropies S(ρ~τ||ρτeq)S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}}) and S(ρ~0||ρ0eq)S(\widetilde{\rho}_{0}||\rho_{0}^{\text{eq}}). From Ref. [43], it has already been proven that

S(ρ~τ||ρτeq)=ln(Z~τZτ).S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})=-\ln\left(\frac{\widetilde{Z}_{\tau}}{Z_{\tau}}\right). (S13)

Similarly, for S(ρ~0||ρ0eq)S(\widetilde{\rho}_{0}||\rho_{0}^{\text{eq}}), we have

S(ρ~0||ρ0eq)=ln(Z~0Z0).S(\widetilde{\rho}_{0}||\rho_{0}^{\text{eq}})=-\ln\left(\frac{\widetilde{Z}_{0}}{Z_{0}}\right). (S14)

Since the equilibrium Helmholtz free energy difference is given by

ΔF=1βln(ZτZ0),\Delta F=-\frac{1}{\beta}\ln\left(\frac{Z_{\tau}}{Z_{0}}\right), (S15)

we can write

S(ρ~τ||ρτeq)S(ρ~0||ρ0eq)=ln(ZτZ0)ln(Z~τZ~0)=βΔFln(Z~τZ~0).S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})-S(\widetilde{\rho}_{0}||\rho_{0}^{\text{eq}})=\ln\left(\frac{Z_{\tau}}{Z_{0}}\right)-\ln\left(\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}\right)=-\beta\Delta F-\ln\left(\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}\right). (S16)

Therefore, we can obtain the following detailed fluctuation theorem

P~f(W)P~b(W)eβW=eβΔFS(ρ~τ||ρτeq)+S(ρ~0||ρ0eq).\frac{\widetilde{P}_{f}(W)}{\widetilde{P}_{b}(-W)}e^{-\beta W}=e^{-\beta\Delta F-S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})+S(\widetilde{\rho}_{0}||\rho_{0}^{\text{eq}})}. (S17)

Note that P~f(W)\widetilde{P}_{f}(W) and P~b(W)\widetilde{P}_{b}(-W) cross at W=ΔF+β1(S(ρ~τ||ρτeq)S(ρ~0||ρ0eq))W=\Delta F+\beta^{-1}(S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})-S(\widetilde{\rho}_{0}||\rho_{0}^{\text{eq}})). Also, in the setup in Ref. [43] (i.e. |ψi=|Ei|\psi_{i}\rangle=|E_{i}\rangle), we have ρ~0=ρ0eq\widetilde{\rho}_{0}=\rho_{0}^{\text{eq}} so that S(ρ~0||ρ0eq)=0S(\widetilde{\rho}_{0}||\rho_{0}^{\text{eq}})=0. Therefore, we can recover the OTM fluctuation theorem and the corresponding Jarzynski equality in Ref. [43].

Next, we derive the symmetric relation by considering the characteristic functions of the work distributions [52, 53, 54]. The characteristic functions are defined as

C~f(u)𝑑WP~f(W)eiuWC~b(u)𝑑WP~b(W)eiuW.\begin{split}\widetilde{C}_{f}(u)&\equiv\int dW\widetilde{P}_{f}(W)e^{iuW}\\ \widetilde{C}_{b}(u)&\equiv\int dW\widetilde{P}_{b}(-W)e^{-iuW}.\end{split} (S18)

From Eq. (S8), we have

C~f(u)=i=1d𝑑Weβψi|H0|ψiZ~0eiuWδ(WW~i)=i=1deβψi|H0|ψiZ~0eiuW~i=i=1de(iu+β)ψi|H0|ψiZ~0eiuψi|UHτU|ψi=Z~τZ~0i=1deiuψi|UHτU|ψiZ~τe(iu+β)ψi|H0|ψi.\begin{split}\widetilde{C}_{f}(u)&=\sum_{i=1}^{d}\int dW\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{0}}e^{iuW}\delta(W-\widetilde{W}_{i})\\ &=\sum_{i=1}^{d}\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{0}}e^{iu\widetilde{W}_{i}}\\ &=\sum_{i=1}^{d}\frac{e^{-(iu+\beta)\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{0}}e^{iu\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}\\ &=\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}\sum_{i=1}^{d}\frac{e^{iu\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}e^{-(iu+\beta)\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}.\end{split} (S19)

Furthermore, from Eq. (S10) and δ(W+Wi~)=δ(WWi~)\delta(-W+\widetilde{W_{i}})=\delta(W-\widetilde{W_{i}}), we can obtain

C~b(u+iβ)=i=1d𝑑Weβψi|UHτU|ψiZ~τei(u+iβ)Wδ(W+Wi~)=i=1d𝑑Weβψi|UHτU|ψiZ~τe(iu+β)Wi~=i=1deiuψi|UHτU|ψiZ~τe(iu+β)ψi|H0|ψi\begin{split}\widetilde{C}_{b}(-u+i\beta)=&\sum_{i=1}^{d}\int dW\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}e^{-i(-u+i\beta)W}\delta(-W+\widetilde{W_{i}})\\ =&\sum_{i=1}^{d}\int dW\frac{e^{-\beta\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}e^{(iu+\beta)\widetilde{W_{i}}}\\ =&\sum_{i=1}^{d}\frac{e^{iu\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}e^{-(iu+\beta)\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}\end{split} (S20)

Therefore, we can obtain the symmetric relation

C~f(u)C~b(u+iβ)=Z~τZ~0.\frac{\widetilde{C}_{f}(u)}{\widetilde{C}_{b}(-u+i\beta)}=\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}. (S21)

From Eq. (S16), we can write

C~f(u)C~b(u+iβ)=eβΔFS(ρ~τ||ρτeq)+S(ρ~0||ρ0eq).\frac{\widetilde{C}_{f}(u)}{\widetilde{C}_{b}(-u+i\beta)}=e^{-\beta\Delta F-S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})+S(\widetilde{\rho}_{0}||\rho_{0}^{\text{eq}})}. (S22)

Equation. (S21) can be recovered from the quantum circuit representation. Following Refs. [52, 53, 54], the characteristic functions are

C~f(u)=tr{UeiuGτUeiuG0ρ~0}C~b(u)=tr{UeiuG0UeiuGτρ~τ}.\begin{split}\widetilde{C}_{f}(u)&=\mathrm{tr}\left\{U^{\dagger}e^{iuG_{\tau}}Ue^{-iuG_{0}}\widetilde{\rho}_{0}\right\}\\ \widetilde{C}_{b}(u)&=\mathrm{tr}\left\{Ue^{iuG_{0}}U^{\dagger}e^{-iuG_{\tau}}\widetilde{\rho}_{\tau}\right\}.\end{split} (S23)

From Eq. (S6), we have

UGτU=i=1dψi|UHτU|ψi|ψiψi|.U^{\dagger}G_{\tau}U=\sum_{i=1}^{d}\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle|\psi_{i}\rangle\!\langle\psi_{i}|. (S24)

Therefore,

C~f(u)=Z~τZ~0i=1deiuψi|UHτU|ψiZ~τe(iu+β)ψi|H0|ψi,\widetilde{C}_{f}(u)=\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}\sum_{i=1}^{d}\frac{e^{iu\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}e^{-(iu+\beta)\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}, (S25)

which is equivalent to Eq. (S19).

Next, for the backward process, because

UG0U=i=1dψi|H0|ψiU|ψiψi|U,UG_{0}U^{\dagger}=\sum_{i=1}^{d}\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle U|\psi_{i}\rangle\!\langle\psi_{i}|U^{\dagger}, (S26)

we can obtain

C~b(u)=i=1de(iuβ)ψi|UHτU|ψiZ~τeiuψi|H0|ψi,\widetilde{C}_{b}(u)=\sum_{i=1}^{d}\frac{e^{(-iu-\beta)\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}e^{iu\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}, (S27)

so that we can arrive at

C~b(u+iβ)=i=1deiuψi|UHτU|ψiZ~τe(iu+β)ψi|H0|ψi,\widetilde{C}_{b}(-u+i\beta)=\sum_{i=1}^{d}\frac{e^{iu\langle\psi_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0pt\psi_{i}\rangle}}{\widetilde{Z}_{\tau}}e^{-(iu+\beta)\langle\psi_{i}\hskip 1.0pt|H_{0}|\hskip 1.0pt\psi_{i}\rangle}, (S28)

which is equivalent to Eq. (S20). Hence, we can obtain

C~f(u)=tr{UeiuGτUeiuG0ρ~0}=tr{UeiuGτUeiuG0eβG0Z~0}=1Z~0tr{UeiuGτUeiuG0eβG0}=1Z~0tr{UeiuG0eβG0UeiuGτ}=1Z~0tr{UeiuG0eβG0UeiuGτeβGτeβGτ}=Z~τZ~0tr{UeiuG0eβG0UeiuGτeβGτeβGτZ~τ}=Z~τZ~0tr{Uei(u+iβ)G0Uei(u+iβ)Gτρ~τ}=Z~τZ~0C~b(u+iβ),\begin{split}\widetilde{C}_{f}(u)&=\mathrm{tr}\left\{U^{\dagger}e^{iuG_{\tau}}Ue^{-iuG_{0}}\widetilde{\rho}_{0}\right\}\\ &=\mathrm{tr}\left\{U^{\dagger}e^{iuG_{\tau}}Ue^{-iuG_{0}}\frac{e^{-\beta G_{0}}}{\widetilde{Z}_{0}}\right\}\\ &=\frac{1}{\widetilde{Z}_{0}}\mathrm{tr}\left\{U^{\dagger}e^{iuG_{\tau}}Ue^{-iuG_{0}}e^{-\beta G_{0}}\right\}\\ &=\frac{1}{\widetilde{Z}_{0}}\mathrm{tr}\left\{Ue^{-iuG_{0}}e^{-\beta G_{0}}U^{\dagger}e^{iuG_{\tau}}\right\}\\ &=\frac{1}{\widetilde{Z}_{0}}\mathrm{tr}\left\{Ue^{-iuG_{0}}e^{-\beta G_{0}}U^{\dagger}e^{iuG_{\tau}}e^{\beta G_{\tau}}e^{-\beta G_{\tau}}\right\}\\ &=\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}\mathrm{tr}\left\{Ue^{-iuG_{0}}e^{-\beta G_{0}}U^{\dagger}e^{iuG_{\tau}}e^{\beta G_{\tau}}\frac{e^{-\beta G_{\tau}}}{\widetilde{Z}_{\tau}}\right\}\\ &=\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}\mathrm{tr}\left\{Ue^{i(-u+i\beta)G_{0}}U^{\dagger}e^{-i(-u+i\beta)G_{\tau}}\widetilde{\rho}_{\tau}\right\}\\ &=\frac{\widetilde{Z}_{\tau}}{\widetilde{Z}_{0}}\widetilde{C}_{b}(-u+i\beta),\end{split} (S29)

which leads Eq. (S21). Again, for the case |ψi=|Ei|\psi_{i}\rangle=|E_{i}\rangle, because ρ~0=ρ0eq\widetilde{\rho}_{0}=\rho_{0}^{\text{eq}}, we can recover our first main result C~f(u)/C~b(u+iβ)=exp(βΔFS(ρ~τ||ρτeq))\widetilde{C}_{f}(u)/\widetilde{C}_{b}(-u+i\beta)=\exp(-\beta\Delta F-S(\widetilde{\rho}_{\tau}||\rho_{\tau}^{\text{eq}})).

Appendix C Derivation of D[P~f||P~b]D[\widetilde{P}_{f}||\widetilde{P}_{b}]

The KL divergence D[P~f||P~b]D[\widetilde{P}_{f}||\widetilde{P}_{b}] is defined as

D[P~f||P~b]dWP~f(W)lnP~f(W)P~b(W).\displaystyle D[\widetilde{P}_{f}||\widetilde{P}_{b}]\equiv\int dW\widetilde{P}_{f}(W)\ln\frac{\widetilde{P}_{f}(W)}{\widetilde{P}_{b}(-W)}. (S30)

Because

P~f(W)\displaystyle\widetilde{P}_{f}(W) i=1deβEiZ0δ(WW~i)\displaystyle\equiv\sum_{i=1}^{d}\frac{e^{-\beta E_{i}}}{Z_{0}}\delta\left(W-\widetilde{W}_{i}\right) (S31)
P~b(W)\displaystyle\widetilde{P}_{b}(-W) i=1deβEi|UHτU|EiZ~τδ(W+W~i)\displaystyle\equiv\sum_{i=1}^{d}\frac{e^{-\beta\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle}}{\widetilde{Z}_{\tau}}\delta(-W+\widetilde{W}_{i}) (S32)

with

W~iEi|UHτU|EiEi,\displaystyle\widetilde{W}_{i}\equiv\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle-E_{i}, (S33)

we obtain

D[P~f||P~b]=𝑑WP~f(W)lnP~f(W)P~b(W)=𝑑Wi=1deβEiZ0δ(WW~i)ln(j=1deβEjZ0δ(WW~j))𝑑Wi=1deβEiZ0δ(WW~i)ln(j=1deβEj|UHτU|EjZ~τδ(W+W~j))=i=1deβEiZ0ln(eβEiZ0)+β(i=1deβEiZ0Ei|UHτU|Ei)+lnZ~τ=S(ρ0eq)+βtr{Uρ0eqUHτ}+lnZ~τ,\displaystyle\begin{split}D[\widetilde{P}_{f}||\widetilde{P}_{b}]=&\int dW\widetilde{P}_{f}(W)\ln\frac{\widetilde{P}_{f}(W)}{\widetilde{P}_{b}(-W)}\\ =&\int dW\sum_{i=1}^{d}\frac{e^{-\beta E_{i}}}{Z_{0}}\delta(W-\widetilde{W}_{i})\ln\left(\sum_{j=1}^{d}\frac{e^{-\beta E_{j}}}{Z_{0}}\delta(W-\widetilde{W}_{j})\right)\\ &-\int dW\sum_{i=1}^{d}\frac{e^{-\beta E_{i}}}{Z_{0}}\delta(W-\widetilde{W}_{i})\ln\left(\sum_{j=1}^{d}\frac{e^{-\beta\langle E_{j}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{j}\rangle}}{\widetilde{Z}_{\tau}}\delta(-W+\widetilde{W}_{j})\right)\\ =&\sum_{i=1}^{d}\frac{e^{-\beta E_{i}}}{Z_{0}}\ln\left(\frac{e^{-\beta E_{i}}}{Z_{0}}\right)+\beta\left(\sum_{i=1}^{d}\frac{e^{-\beta E_{i}}}{Z_{0}}\langle E_{i}\hskip 1.0pt|U^{\dagger}H_{\tau}U|\hskip 1.0ptE_{i}\rangle\right)+\ln\widetilde{Z}_{\tau}\\ =&-S(\rho_{0}^{\text{eq}})+\beta\mathrm{tr}\left\{U\rho_{0}^{\text{eq}}U^{\dagger}H_{\tau}\right\}+\ln\widetilde{Z}_{\tau}\,,\end{split} (S34)

where S(ρ0eq)S(\rho_{0}^{\text{eq}}) is the von-Neumann entropy of ρ0eq\rho_{0}^{\text{eq}}.

Appendix D Detailed Qiskit circuits to compute the characteristic functions

In this section, we provide the details of the quantum circuits to compute the characteristic functions C~f(1)\widetilde{C}_{f}(1) and C~b(1+0.5i)\widetilde{C}_{b}(-1+0.5i).

D.1 Forward process

To compute the characteristic function of the forward process C~f(1)\widetilde{C}_{f}(1), we need to run 4×2=84\times 2=8 quantum circuits because of the decomposition of ρ0eq{\rho}_{0}^{\text{eq}} into the four orthogonal states and the measurement of the ancilla qubit on two different bases. Figure. S1 is the quantum circuit for computing the real part of tr{UeiGτUeiH0|E4E4|}\mathrm{tr}\left\{U^{\dagger}e^{iG_{\tau}}Ue^{-iH_{0}}|E_{4}\rangle\!\langle E_{4}|\right\}.

Refer to caption
Figure S1: An example of the forward process circuits. After applying X gate on the bottom two qubits and Hadamard gate on the ancilla qubit, controlled exp(iH0)\exp(-iH_{0}) gate, UU gate, controlled exp(iGπ/4)\exp(iG_{\pi/4}) gate are implemented. Note that PP is a phase gate and UU is a custom unitary gate, which are defined by P(λ)=(100eiλ)P(\lambda)=\begin{pmatrix}1&0\\ 0&e^{i\lambda}\end{pmatrix} and U(θ,ϕ,λ)=(cos(θ2)eiλsin(θ2)eiϕsin(θ2)ei(ϕ+λ)cos(θ2))U(\theta,\phi,\lambda)=\begin{pmatrix}\cos(\frac{\theta}{2})&-e^{i\lambda}\sin(\frac{\theta}{2})\\ e^{i\phi}\sin(\frac{\theta}{2})&e^{i(\phi+\lambda)}\cos(\frac{\theta}{2})\end{pmatrix}

D.2 Backward process

Next, we explain the details of computing C~b(1+0.5i)\widetilde{C}_{b}(-1+0.5i). By computing αk(0)=14tr{eβH0σk}\alpha_{k}^{(0)}=\frac{1}{4}\mathrm{tr}\left\{e^{-\beta H_{0}}\sigma_{k}\right\} and αk(π/4)=14tr{e0.5Gπ/4σk}\alpha_{k}^{(\pi/4)}=\frac{1}{4}\mathrm{tr}\left\{e^{0.5G_{\pi/4}}\sigma_{k}\right\}, the decomposition of exp(0.5H0)\exp(0.5H_{0}) and exp(0.5Gτ)\exp(-0.5G_{\tau}) with the Pauli strings are given by

e0.5H0=2.3811(𝟙𝟙)1.81343(𝟙)1.81343(𝟙)+1.3811()e0.5Gτ=1.03141(𝟙𝟙)+0.126306(𝕏𝕏)0.126306(𝕏)0.126306(𝕏)+0.126306().\begin{split}e^{-0.5H_{0}}&=2.3811(\openone\otimes\openone)-1.81343(\openone\otimes Z)-1.81343(Z\otimes\openone)+1.3811(Z\otimes Z)\\ e^{0.5G_{\tau}}&=1.03141(\openone\otimes\openone)+0.126306(X\otimes X)-0.126306(X\otimes Z)-0.126306(Z\otimes X)+0.126306(Z\otimes Z).\end{split} (S35)

Therefore, we have 5×4=205\times 4=20 non-zero values of FkF_{k\ell}. Because ρ~π/4\widetilde{\rho}_{\pi/4} is the linear combination of four orthogonal states, taking into account the fact that we need to perform the measurement on two different bases, we finally need to make 4×5×4×2=1604\times 5\times 4\times 2=160 quantum circuits to compute C~b(1+0.5i)\widetilde{C}_{b}(-1+0.5i). Figure. S2 is the quantum circuit for computing the imaginary part of tr{UeiH0σkUeiGπ/4σU|E1E1|U}\mathrm{tr}\left\{Ue^{-iH_{0}}\sigma_{k}U^{\dagger}e^{iG_{\pi/4}}\sigma_{\ell}U|E_{1}\rangle\!\langle E_{1}|U^{\dagger}\right\}, where σk=Z𝟙\sigma_{k}=Z\otimes\openone and σ=XX\sigma_{\ell}=X\otimes X.

Refer to caption
Figure S2: An example of the backward process circuits. After applying UU on the two qubits and Hadamard gate on the ancilla qubit, we add controlled XXX\otimes X gate, controlled exp(iGπ/4)\exp(iG_{\pi/4}) gate, UU^{\dagger} gate, controlled Z𝟙Z\otimes\openone gate, and controlled exp(iH0)\exp(-iH_{0}) gate. Note that H in the circuit is Hadamard gate.

Appendix E Detailed information about IBM machines

In this section, we present detailed information about the IBM quantum machines that we used for this simulation. We used ibmq_lima, ibmq_belem, ibmq_quito, ibm_oslo, ibmq_manila, and ibm_lagos. The parameters of these machines are shown in Tables. S1, S2, S3, S4, and S5. Note that ibm_oslo has retired so that its information is no longer available. The qubit configurations of each machine are depicted in Figs. S3, S4, S5, S6 and S7.

Qubit T1T_{1} (μ\upmus) T2T_{2} (μ\upmus) Frequency (GHz) Anharmonicity (GHz) Readout assignment error
0 166.56 232.22 5.030 -0.33574 2.130×102\times 10^{-2}
1 140.20 135.21 5.128 -0.31835 1.850×102\times 10^{-2}
2 94.23 106.02 5.247 -0.33360 1.860×102\times 10^{-2}
3 86.82 95.33 5.303 -0.33124 3.370×102\times 10^{-2}
4 17.48 24.33 5.092 -0.33447 4.780×102\times 10^{-2}
Qubit Prob meas 0 prep 1 Prob meas 1 prep 0 Readout length (ns) ID error X\sqrt{X} (SX) error
0 0.0338 0.0088 5912.889 2.946×104\times 10^{-4} 2.946×104\times 10^{-4}
1 0.0256 0.0114 5912.889 4.617×104\times 10^{-4} 4.617×104\times 10^{-4}
2 0.0288 0.0084 5912.889 5.618×104\times 10^{-4} 5.618×104\times 10^{-4}
3 0.0442 0.0232 5912.889 2.428×104\times 10^{-4} 2.428×104\times 10^{-4}
4 0.0760 0.0196 5912.889 8.183×104\times 10^{-4} 8.183×104\times 10^{-4}
Qubit Pauli-X error CNOT error Gate time (ns)
0 2.946×104\times 10^{-4} 0_1:0.00677 0_1:305.778
1 4.617×104\times 10^{-4} 1_0:0.00677 1_0:341.333
1_3:0.01360 1_3:497.778
1_2:0.00677 1_2:334.222
2 5.618×104\times 10^{-4} 2_1:0.00677 2_1:298.667
3 2.428×104\times 10^{-4} 3_4:0.01679 3_4:519.111
3_1:0.01360 3_1:462.222
4 8.183×104\times 10^{-4} 4_3:0.01679 4_3:483.556
Table S1: The parameter settings of ibmq_lima. In this simulation, we used either Qubit 0, 1, 2, Qubit 0, 1, 3, Qubit 1, 2, 3, or Qubit 1, 3, 4.
Qubit T1T_{1} (μ\upmus) T2T_{2} (μ\upmus) Frequency (GHz) Anharmonicity (GHz) Readout assignment error
0 144.96 139.86 5.090 -0.33612 1.990×102\times 10^{-2}
1 85.46 94.85 5.246 -0.31657 1.820×102\times 10^{-2}
2 69.38 51.70 5.361 -0.33063 2.510×102\times 10^{-2}
3 102.99 117.70 5.170 -0.33374 3.580×102\times 10^{-2}
4 130.83 155.00 5.258 -0.33135 1.900×102\times 10^{-2}
Qubit Prob meas 0 prep 1 Prob meas 1 prep 0 Readout length (ns) ID error X\sqrt{X} (SX) error
0 0.0330 0.0068 6158.222 1.737×104\times 10^{-4} 1.737×104\times 10^{-4}
1 0.0320 0.0044 6158.222 2.221×104\times 10^{-4} 2.221×104\times 10^{-4}
2 0.0436 0.0066 6158.222 2.661×104\times 10^{-4} 2.661×104\times 10^{-4}
3 0.0564 0.0152 6158.222 3.742×104\times 10^{-4} 3.742×104\times 10^{-4}
4 0.0292 0.0088 6158.222 5.762×104\times 10^{-4} 5.762×104\times 10^{-4}
Qubit Pauli-X error CNOT error Gate time (ns)
0 1.737×104\times 10^{-4} 0_1:0.01276 0_1:810.667
1 2.221×104\times 10^{-4} 1_3:0.00776 1_3:440.889
1_2:0.00884 1_2:419.556
1_0:0.01276 1_0:775.111
2 2.661×104\times 10^{-4} 2_1:0.00884 2_1:384.000
3 3.742×104\times 10^{-4} 3_4:0.00978 3_4:526.222
3_1:0.00776 3_1:405.333
4 5.762×104\times 10^{-4} 4_3:0.00978 4_3:490.667
Table S2: The parameter settings of ibmq_belem. In this simulation, we used either Qubit 0, 1, 2, Qubit 0, 1, 3, Qubit 1, 2, 3, or Qubit 1, 3, 4.
Qubit T1T_{1} (μ\upmus) T2T_{2} (μ\upmus) Frequency (GHz) Anharmonicity (GHz) Readout assignment error
0 121.37 151.04 5.301 -0.33148 5.870×102\times 10^{-2}
1 43.25 79.69 5.081 -0.31925 3.550×102\times 10^{-2}
2 111.43 116.63 5.322 -0.33232 6.710×102\times 10^{-2}
3 86.75 17.79 5.164 -0.33508 4.000×102\times 10^{-2}
4 89.82 107.98 5.052 -0.31926 3.290×102\times 10^{-2}
Qubit Prob meas 0 prep 1 Prob meas 1 prep 0 Readout length (ns) ID error X\sqrt{X} (SX) error
0 0.0838 0.0336 5351.111 4.604×104\times 10^{-4} 4.604×104\times 10^{-4}
1 0.0402 0.0308 5351.111 2.924×104\times 10^{-4} 2.924×104\times 10^{-4}
2 0.0646 0.0696 5351.111 2.564×104\times 10^{-4} 2.564×104\times 10^{-4}
3 0.0616 0.0184 5351.111 1.178e-03 1.178e-03
4 0.0466 0.0192 5351.111 3.567×104\times 10^{-4} 3.567×104\times 10^{-4}
Qubit Pauli-X error CNOT error Gate time (ns)
0 4.604×104\times 10^{-4} 0_1:0.01376 0_1:234.667
1 2.924×104\times 10^{-4} 1_3:0.00920 1_3:334.222
1_2:0.00675 1_2:298.667
1_0:0.01376 1_0:270.222
2 2.564×104\times 10^{-4} 2_1:0.00675 2_1:263.111
3 1.178e-03 3_4:0.01483 3_4:277.333
3_1:0.00920 3_1:369.778
4 3.567×104\times 10^{-4} 4_3:0.01483 4_3:312.889
Table S3: The parameter settings of ibmq_quito. In this simulation, we used either Qubit 0, 1, 2, Qubit 0, 1, 3, Qubit 1, 2, 3, or Qubit 1, 3, 4.
Qubit T1T_{1} (μ\upmus) T2T_{2} (μ\upmus) Frequency (GHz) Anharmonicity (GHz) Readout assignment error
0 72.22 18.13 4.962 -0.34463 4.840×102\times 10^{-2}
1 176.31 63.37 4.838 -0.34528 4.230×102\times 10^{-2}
2 147.72 17.26 5.037 -0.34255 2.810×102\times 10^{-2}
3 161.33 60.76 4.951 -0.34358 1.890×102\times 10^{-2}
4 47.92 37.82 5.065 -0.34211 2.360×102\times 10^{-2}
Qubit Prob meas 0 prep 1 Prob meas 1 prep 0 Readout length (ns) ID error X\sqrt{X} (SX) error
0 0.0778 0.0190 5351.111 5.700×104\times 10^{-4} 5.700×104\times 10^{-4}
1 0.0456 0.0390 5351.111 3.243×104\times 10^{-4} 3.243×104\times 10^{-4}
2 0.0380 0.0182 5351.111 2.435×104\times 10^{-4} 2.435×104\times 10^{-4}
3 0.0250 0.0128 5351.111 1.617×104\times 10^{-4} 1.617×104\times 10^{-4}
4 0.0352 0.0120 5351.111 4.111×104\times 10^{-4} 4.111×104\times 10^{-4}
Qubit Pauli-X error CNOT error Gate time (ns)
0 5.700×104\times 10^{-4} 0_1:0.00800 0_1:277.333
1 3.243×104\times 10^{-4} 1_2:0.01801 1_2:469.333
1_0:0.00800 1_0:312.889
2 2.435×104\times 10^{-4} 2_3:0.00652 2_3:355.556
2_1:0.01801 2_1:504.889
3 1.617×104\times 10^{-4} 3_4:0.00499 3_4:334.222
3_2:0.00652 3_2:391.111
4 4.111×104\times 10^{-4} 4_3:0.00499 4_3:298.667
Table S4: The parameter settings of ibmq_manila. In the simulation, we used wither Qubit 0, 1, 2, Qubit 1, 2, 3, or Qubit 2, 3, 4.
Qubit T1T_{1} (μ\upmus) T2T_{2} (μ\upmus) Frequency (GHz) Anharmonicity (GHz) Readout assignment error
0 120.31 43.15 5.235 -0.33987 1.860×102\times 10^{-2}
1 134.21 86.21 5.100 -0.34325 1.520×102\times 10^{-2}
2 198.82 137.93 5.188 -0.34193 6.700e-03
3 253.10 89.53 4.987 -0.34529 1.400×102\times 10^{-2}
4 66.93 35.95 5.285 -0.33923 2.260×102\times 10^{-2}
5 126.12 65.55 5.176 -0.34079 1.620×102\times 10^{-2}
6 189.10 107.66 5.064 -0.34276 1.420×102\times 10^{-2}
Qubit Prob meas 0 prep 1 Prob meas 1 prep 0 Readout length (ns) ID error X\sqrt{X} (SX) error
0 0.0138 0.0234 789.333 2.333×104\times 10^{-4} 2.333×104\times 10^{-4}
1 0.0174 0.0130 789.333 1.712×104\times 10^{-4} 1.712×104\times 10^{-4}
2 0.0056 0.0078 789.333 2.233×104\times 10^{-4} 2.233×104\times 10^{-4}
3 0.0166 0.0114 789.333 1.805×104\times 10^{-4} 1.805×104\times 10^{-4}
4 0.0208 0.0244 789.333 2.025×104\times 10^{-4} 2.025×104\times 10^{-4}
5 0.0180 0.0144 789.333 1.640×104\times 10^{-4} 1.640×104\times 10^{-4}
6 0.0146 0.0138 789.333 2.433×104\times 10^{-4} 2.433×104\times 10^{-4}
Qubit Pauli-X error CNOT error Gate time (ns)
0 2.333×104\times 10^{-4} 0_1:0.00783 0_1:576.000
1 1.712×104\times 10^{-4} 1_3:0.00453 1_3:334.222
1_2:0.00559 1_2:327.111
1_0:0.00783 1_0:611.556
2 2.233×104\times 10^{-4} 2_1:0.00559 2_1:291.556
3 1.805×104\times 10^{-4} 3_1:0.00453 3_1:298.667
3_5:0.00879 3_5:334.222
4 2.025×104\times 10^{-4} 4_5:0.00642 4_5:362.667
5 1.640×104\times 10^{-4} 5_4:0.00642 5_4:327.111
5_6:0.00690 5_6:256.000
5_3:0.00879 5_3:298.667
6 2.433×104\times 10^{-4} 6_5:0.00690 6_5:291.556
Table S5: The parameter settings of ibmq_lagos. In the simulation, we used either Qubit 0, 1, 2, Qubit 0, 1, 3, Qubit 1, 2, 3, Qubit 1, 3, 5, Qubit 3, 4, 5, or Qubit 3, 5, 6.
Refer to caption
Figure S3: Qubit configuration of ibm_lima.
Refer to caption
Figure S4: Qubit configuration of ibm_belem.
Refer to caption
Figure S5: Qubit configuration of ibm_quito.
Refer to caption
Figure S6: Qubit configuration of ibm_manila.
Refer to caption
Figure S7: Qubit configuration of ibm_lagos.