Paper The following article is Open access

Experimental test of fluctuation relations for driven open quantum systems with an NV center

, , and

Published 7 June 2021 © 2021 The Author(s). Published by IOP Publishing Ltd on behalf of the Institute of Physics and Deutsche Physikalische Gesellschaft
, , Focus on Microscopic Engines and Refrigerators: Theory and Experiments from Classical to Quantum Citation Santiago Hernández-Gómez et al 2021 New J. Phys. 23 065004 DOI 10.1088/1367-2630/abfc6a

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1367-2630/23/6/065004

Abstract

The experimental verification of quantum fluctuation relations for driven open quantum system is currently a challenge, due to the conceptual and operative difficulty of distinguishing work and heat. The nitrogen-vacancy (NV) center in diamond has been recently proposed as a controlled test bed to study fluctuation relations in the presence of an engineered dissipative channel, in absence of work (Hernández-Gómez et al 2020 Phys. Rev. Res. 2 023327). Here, we extend those studies to exploring the validity of quantum fluctuation relations in a driven-dissipative scenario, where the spin exchanges energy both with its surroundings because of a thermal gradient, and with an external work source. We experimentally prove the validity of the quantum fluctuation relations in the presence of cyclic driving in two cases, when the spin exchanges energy with an effective infinite-temperature reservoir, and when the total work vanishes at stroboscopic times—although the power delivered to the NV center is non-null. Our results represent the first experimental study of quantum fluctuation relation in driven open quantum systems.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Quantum fluctuation relations [14] are exact relations in quantum statistical mechanics that allow for a compact and effective description of a plethora of non-equilibrium phenomena such as the statistical nature of the second law of thermodynamics and the emergence of an arrow of time [5], Onsager relations, Green–Kubo relations and high order relations between responses and multi-point correlation functions [6, 7], the microscopic mechanisms at the basis of Clausius inequality [8], and the Carnot theorem imposing the celebrated Carnot efficiency as the universal upper bound to efficiency of heat engines [9, 10]. Recently, they have been shown to be at the basis of thermodynamic uncertainty relations expressing lower bounds on entropy production in terms of the precision (signal-to-noise ratio) of fluctuating non-equilibrium charges (i.e., integrated non-equilibrium currents) [11, 12], or the asymmetry of their statistics [13].

Generally speaking, quantum fluctuation relations refer to a scenario where a quantum system, prepared in thermal equilibrium at some inverse temperature β1, is then placed in contact (sequentially or at the same time) with several thermal baths at various temperatures β2, ..., βN while being possibly acted upon by a work source resulting in a time-dependent forcing [8]. In integral form, the quantum fluctuation relation reads [4]:

Equation (1)

where ΔFi is the free energy change between the (hypothetical) equilibrium state of component i of the global system + bath compound at the protocol's end time, and its initial free energy 7 , while ΔEi is the according stochastic energy change as obtained by taking the difference of projectively measured eigen-energies of component i at the beginning and end of the driving protocol (here the so called two-point measurement scheme is implied [4]). The symbol ⟨⋅⟩ denotes average over the statistics of the joint probability distribution pE1, ..., ΔEN ) for the energy change of each component.

The above relation encompasses a number of typical situations encountered in non-equilibrium thermodynamics. For example, a system initially in equilibrium and in contact with a thermal bath at some inverse temperature β is acted upon by a time dependent force: in this case β1 = β2 = β, and since the work done by the external source is given by W = ΔE1 + ΔE2, the Jarzynski relation follows ⟨eβW ⟩ = eβΔF , with ΔF the global free energy variation [14]. This relation has been verified experimentally with closed quantum systems (namely in absence of a thermal bath) [15, 16], and for a quantum system in contact with a dephasing environment that does not modify the system energy [17]. However, the verification of this relation with an open quantum system that exchanges energy with its environment is still elusive, due to the fact that the measurement of total energy injected in system and bath is typically very challenging. Possible solutions to this challenge are offered by the interferometric method for the measurement of work statistics [18, 19] as discussed in [20], or with reference to superconducting circuit setups, by calorimetric measurements of energy exchanges between a qubit and a resistor [21].

A complementary situation is when two quantum systems prepared initially at different temperatures, are placed in thermal contact for some time τ. Provided the interaction energy is negligible, no work is associated to this process, and the general relation takes the form of the so called exchange fluctuation relation ⟨eΔβΔE ⟩ = 1 [22], where Δβ = β2β1 and ΔE = ΔE1 = −ΔE2. This relation has been recently verified in the quantum regime with NMR setup [23] and with nitrogen vacancy (NV) centers in diamond [24].

A situation of central interest in quantum thermodynamics combines the previous two scenarios, namely when not only a system exchanges energy with its surroundings because of a thermal gradient, but also with an external work source. That is indeed what happens during the operation of a heat engine. The experimental verification of equation (1) when there is a simultaneous exchange of heat and work poses significant challenges, as typical quantum platforms allow to experimentally access the system energy change ΔE = WQ, but not the individual terms W and Q.

In this work, we address this challenge and experimentally explore the validity the quantum fluctuation relation for a driven-dissipative dynamics (equation (1)) realized by the ground state of a single negatively-charged NV center in diamond. The NV center is well established as a biocompatible quantum sensor [2527], single-photon source [28], as well as building block for quantum registers [2931], thanks to the coherent control of its ground state spin and optical addressability. The capability of coherently manipulating a two or three level spin system, and to engineer dissipative channels to mimic the thermal contact with a quantum reservoir, make the NV center also an excellent platform to explore quantum thermodynamics processes [24, 32]. Here, we use an NV center spin qubit to prove the validity of the fluctuation relation in equation (1) in two special cases of driven-dissipative dynamics, where the measurement of the spin energy change ΔE suffices: (i) when the environment with which the system exchanges energy has infinite temperature; (ii) when the duration of the process is chosen in such a way that the total work done on the system amounts to zero, although generally the power delivered to it is non-null.

The results obtained in this work expand the experimental basis over which quantum fluctuation relations have been tested, and prepare the ground for a fully fledged testing of equation (1) in the general case when energy is exchanged both in the form of heat and work.

2. Experimental setup and protocol

To engineer a driven-dissipative dynamics, we use the electronic spin ground state of a single NV center in diamond at room temperature. An external magnetic field, B = 38.9 mT, is aligned along the NV spin quantization axis to remove the degeneracy between the spin projections mS = ±1. The two states $\left\vert {m}_{\text{S}}=+1\right\rangle \equiv \left\vert 1\right\rangle $ and $\left\vert {m}_{\text{S}}=0\right\rangle \equiv \left\vert 0\right\rangle $ form the basis of a two-level system. The two-level system is optically initialized into $\left\vert 0\right\rangle $ before each experiment, and its state is read out by measuring the NV spin-selective photoluminescence intensity [33, 34].

We apply a continuous resonant microwave (mw) driving field, described by a Hamiltonian that, in the mw rotating frame, reads

Equation (2)

where σi are Pauli matrices, ω(t) is the Rabi frequency that is set by the mw amplitude, and θ is an angular frequency of the mw phase variation. We denote the instantaneous eigenstates of the Hamiltonian in equation (2) as ${\rho }_{{\pm}}\left(t\right)\equiv \left\vert {E}_{{\pm}}\left(t\right)\right\rangle \left\langle {E}_{{\pm}}\left(t\right)\right\vert $, and their corresponding eigenvalues as E±(t) = ±ℏω(t)/2.

Besides the unitary evolution, the system interacts with a train of NL temporally-equidistant short laser pulses, as depicted in figure 1. In a previous work [24], we have demonstrated that a photon absorption from each laser pulse results in the application of a quantum projective measurement and the opening of a dissipative channel. The application of a train of short laser pulses brings the spin into an asymptotic Gibbs state with inverse (pseudo)temperature βR [24] and, as such, is analogous to the dissipation channel caused by the interaction with a heat reservoir R. The value of βR is determined by a combination of factors, which are the relative orientation of the driving Hamiltonian operator with respect to the NV quantization axis (σz ), the Rabi frequency of the driving field, the laser absorption probability pa, and the time between pulses [24].

Figure 1.

Figure 1. Scheme of the experimental protocol. The NV spin qubit is initially prepared in one of the Hamiltonian eigenstates. Then, the map $\mathcal{M}$ is applied, given by a combination of coherent microwave driving (blue) and a train of NL equidistant short laser pulses that result in quantum projective measurements and a dissipative channel (green). We vary NL, and thus the total evolution time tf under the map $\mathcal{M}$, while keeping fixed the interpulse delay time τ. At the end of the protocol, the spin energy is measured.

Standard image High-resolution image

To measure the statistics of the spin energy change, and to verify the fluctuation relation in equation (1), we implement an experimental protocol consisting of three steps [24]:

  • (a)  
    Initialization. The system is initialized into one of the initial Hamiltonian eigenstates {ρ±(0)}, say {ρ+(0)}, by applying a long laser pulse that optically pumps the spin in $\left\vert 0\right\rangle $, followed by an opportune spin-rotating mw gate.
  • (b)  
    Evolution. The system evolves during a time tf under the action of the mw driving, and is intermittently perturbed by short laser pulses.
  • (c)  
    Readout. At time tf, the spin is mapped back to the $\left\{\left\vert 0\right\rangle ,\left\vert 1\right\rangle \right\}$ basis and the NV photoluminescence intensity under laser illumination is recorded to measure the spin state.

The protocol is repeated several times and the integrated photoluminescence signal gives the probability for the spin to end up in state ρ+(t). As a result, we experimentally construct the conditional probability P+|+(tf) that the state ρ+(t) is detected at time t = tf given that the two-level-system was initialized into ρ+(0) at time t = 0. The same procedure is applied with the system initialized into ρ(0) to give P+|−(tf).

Formally, the conditional probability is expressed as

Equation (3)

where $\mathcal{M}\left({t}_{\text{f}}\right)\left[{\rho }_{{\pm}}\left(0\right)\right]$ is the density operator at time tf, expressed as a quantum map $\mathcal{M}\left({t}_{\text{f}}\right)$ applied to the initial state ρ±(0). Notice that, by definition, P−|± = 1 − P+|±. Therefore, measuring P+|+(tf) and P+|−(tf) is enough to completely reconstruct the energy jump probabilities.

Together with the initial probability Pi (0), the conditional probability Pj|i (tf), with j = ± and i = ±, provides all necessary information to construct the two-level-system energy change probability distribution function:

Equation (4)

During the statistical analysis of the experimental data, the results for each of the initial states are weighted with the according Gibbs probability ${P}_{{\pm}}\left(0\right)={\text{e}}^{-\beta {E}_{{\pm}}\left(0\right)}/{Z}_{0}$ hence emulating a projective measurement being performed onto a thermal mixture at inverse temperature β. Given that we measure the conditional probabilities Pj|i (tf) for a high number of experimental realizations (∼106) and for each Hamiltonian eigenstate, this protocol is equivalent to the two-time measurement protocol applied to a thermal mixed state with inverse temperature β.

The NV center is an open quantum system: during its evolution it exchanges energy in the form of heat with the laser field (mimicking a thermal reservoir) and in the form of work with the external time dependent driving. Using the symbol Q to denote the energy gained by the reservoir, equation (1) becomes in the present case:

Equation (5)

where βR denotes the reservoir inverse (pseudo)temperature, and ΔF = −β−1 ln Z(tf)/Z(0) is the free energy difference of the NV center, between its reference equilibrium state at time tf and its initial equilibrium state. Note that the external driving only applies to the NV center and does not affect the reservoir, therefore the reservoir free energy is constant (ΔFR = 0) and hence it does not appear in the above equation. In equation (5) the symbol ⟨⋅⟩ denotes the average over the joint probability distribution function pE, Q), while we only have experimental access to its marginal pE). So, in principle, in order to check the validity of equation (5) one needs to perform measurements of both Q and ΔE. The marginal distribution pE) is however sufficient to check the validity of equation (1) in two special cases, namely when the reservoir temperature is infinite (βR = 0), and when the total work W = ΔE + Q performed on the system is null, and accordingly ΔE = −Q.

In the first case equation (5) reduces to:

Equation (6)

Since the heat Q does not explicitly appear in equation (6), the average reduces to the one over the marginal pE). Note that although this expression appears formally similar to the Jarzynski relation [35], in an open system as considered here the energy exchange takes into account not only work but also heat flux. In the second case, equation (5) reduces to:

Equation (7)

where, for the same reason as above, the average can be understood as an average over the marginal pE). Note that in our experiment the power delivered to the NV spin is non-null at all times, namely the NV spin continuously exchanges energy with the work source. However, we will focus on specific times tf at which the total work amounts to zero, in which case equation (7) holds.

3. Reservoir at infinite pseudo-temperature

In this section we report on the experimental verification of equation (6) as a special case of the general fluctuation relation in equation (1). To that end we set an infinite pseudo-temperature of the reservoir that mimics the thermal environment (βR = 0). That is achieved by applying the following Hamiltonian:

Equation (8)

Applying a sequence of short laser pulses while the system evolves under this Hamiltonian results in our NV center qubit reaching the completely mixed state, corresponding to βR = 0, as we demonstrate in the experiment and with numerical simulations, as discussed below. Work is provided via a periodic temporal variation of the Hamiltonian amplitude ω(t), which in takes the form:

Equation (9)

with period set to τA = 616 ns, and ω0 = π/τA. Under these conditions, the eigenstates of the Hamiltonian remain constant in time, while the eigenvalues are time-dependent.

We applied the protocol described in the previous section to measure the conditional probabilities Pj|i (tf), considered as functions of tf. Since the Hamiltonian commutes with itself at different time values, Pj|i (tf) remain constant during the time in-between laser pulses, while they present discontinuities caused by the dissipation at the times when laser pulses are applied, as shown in figure 2(a). Each experimental point shown in this figure was obtained by averaging over an ensemble of ∼106 repetitions of the same protocol.

Figure 2.

Figure 2. Spin evolution under a driven-dissipative map, with a reservoir at infinite pseudo-temperature. The spin is driven by a Hamiltonian aligned along σx with time-varying amplitude, and short laser pulses are applied. (a) Conditional probabilities P+|i of the spin to be measured in the eigenstate $\left\vert {E}_{+}\right\rangle $, after initialization in $\left\vert i\right\rangle =\left\vert {E}_{{\pm}}\right\rangle $ (red dots and black squares, respectively), as a function of time tf. The vertical dashed lines denote the times at which each laser pulse is applied (interpulse time τ = 410 ns). The top axis reports the number of applied pulses NL. Blue line: numerical simulation, as detailed in appendix A. Notice how the conditional probabilities have a jump when a laser pulse is applied. For long evolution time P+|i approach the constant value ${P}_{+}^{\infty }=0.5$. (b)–(c) Simulated mean trajectories of the spin after initialization in each of the eigenstates $\left\vert i\right\rangle =\left\vert {E}_{{\pm}}\right\rangle $, shown on the Bloch sphere, respectively in red and black (top (b) and front (c) view). The mean trajectories reveal the interplay of two different kinds of dynamics: (i) the coherent evolution under the Hamiltonian $\mathcal{H}\left({t}_{\text{f}}\right)$ aligned along σx , and (ii) the dissipative evolution directed toward the center of the Bloch sphere, due to the presence of laser pulses. The markers (red dots and black squares) indicate tf = . The simulated conditional probabilities P+|i shown in (a) (blue lines) are obtained from projecting these simulated mean spin trajectories onto the Hamiltonian axis.

Standard image High-resolution image

The asymptotic behavior is such that, in the limit of tf, the conditional probabilities Pj|i (tf) → 0.5 for any i and j, meaning that our dissipative channel is for practical purposes analogous to a thermal reservoir of inverse temperature βR = 0, as anticipated above. In figures 2(b) and (c), we show the simulated mean trajectory obtained by propagating the density operators $\left\vert {E}_{+}\right\rangle \left\langle {E}_{+}\right\vert $, and $\left\vert {E}_{-}\right\rangle \left\langle {E}_{-}\right\vert $, with the dissipative map $\mathcal{M}$ that describes the dynamics. See appendix A for details on the simulation. The concatenation of coherent driving and dissipation results in a map that intermittently pushes the state toward the center of the Bloch sphere.

Having measured the conditional probabilities of energy variation, we can obtain the energy variation probability distribution function in equation (4). This is enough to reconstruct the mean value of a generic function of ΔE.

The mean value of the energy variation ⟨ΔE⟩ is shown in figure 3. The probability associated with the initial Gibbs state was set to P+(0) = 1/(1 + e2), hence the initial inverse temperature was β = 2/ℏω0. The experimental values retrieved for ⟨ΔE⟩ are in agreement with the theoretical values of the sum of the mean work applied on the system plus the mean heat flux, ⟨W⟩ + ⟨Q⟩, as expected from the first law of thermodynamics. These mean values are calculated as $\langle W\rangle ={\int }_{0}^{{t}_{\text{f}}}\mathrm{d}{t}^{\prime }\enspace \mathrm{Tr}\left[\rho \left({t}^{\prime }\right)\frac{\mathrm{d}\mathcal{H}\left({t}^{\prime }\right)}{\mathrm{d}{t}^{\prime }}\right]$, and $\langle Q\rangle ={\int }_{0}^{{t}_{\text{f}}}\mathrm{d}{t}^{\prime }\enspace \mathrm{Tr}\left[\frac{\mathrm{d}\rho \left({t}^{\prime }\right)}{\mathrm{d}{t}^{\prime }}\mathcal{H}\left({t}^{\prime }\right)\right]$ [36]. More details on this calculation can be found in appendix B.

Figure 3.

Figure 3. Energy exchange statistics for a driven-dissipative qubit dynamics, with a reservoir at infinite pseudo-temperature. Measured mean energy variation ⟨ΔE⟩ (blue dots); calculated mean work ⟨W⟩ (orange line), and mean heat flux ⟨Q⟩ (green line). The sum ⟨W⟩ + ⟨Q⟩ (red line) results in agreement with the measured ⟨ΔE⟩. The free energy variation ΔF (gray line) depends only on the amplitude of the Hamiltonian (see text). Interpulse delay time: (a) τ = 410 ns, and (b) τ = 616 ns, respectively.

Standard image High-resolution image

In the case considered here, the Hamiltonian time dependency is reflected in the oscillations of ⟨W⟩. On the other hand, the dissipation induced by the interaction with the laser pulses appears as discrete jumps in ⟨Q⟩. As mentioned before, the dissipation tends to equilibrate the population of the Hamiltonian eigenstates, thus the energy variation induced by the Hamiltonian driving is gradually reduced, which explains the damping on ⟨W⟩. For a time between laser pulses τ = τA carefully chosen to be equal to the Hamiltonian period (figure 3(b)), the flow of heat is perfectly synchronized to the times when work vanishes. Instead, when the synchronization is not perfect ττA (figure 3(a)), the mean work accumulates and tends to a finite value. In these experiments, the Hamiltonian is such that $Z\left(t\right)\equiv \mathrm{Tr}\left[{\text{e}}^{-\beta \mathcal{H}\left(t\right)}\right]=2\enspace \mathrm{cosh}\left[\beta \enspace \hslash \omega \left(t\right)/2\right]$, with ω(t) defined in equation (9). Therefore, the values of ΔF = −β−1 ln(Z(tf)/Z(0)) can be computed analytically (gray lines in figure 3). The periodic time-variation of the Hamiltonian induces an oscillation with constant amplitude on ΔF. Notice that, before the very first laser pulse, when the system is closed, the difference ⟨Wirr ≡ ⟨W⟩−ΔF ⩾ 0 is caused by the inner friction during the unitary evolution [37]. Given that the Hamiltonian eigenstates remain constant in time, the initial state remains unaltered by the unitary evolution. However, since the Hamiltonian eigenvalues change for a time 0 < tf < τA, then the system is, on average, in a non-equilibrium state ${\rho }_{{t}_{\text{f}}}={\text{e}}^{-\beta \mathcal{H}\left(0\right)}/Z\left(0\right)$ (only describing an equilibrium thermal state for tf = 0, τA, 2τA, ...). The quantum relative entropy between ${\rho }_{{t}_{\text{f}}}$ and the actual thermal state ${\rho }_{{t}_{\text{f}}}^{\text{th}}={\text{e}}^{-\beta \mathcal{H}\left({t}_{\text{f}}\right)}/Z\left({t}_{\text{f}}\right)$ defines the value of ⟨Wirr [37, 38]. After the first laser pulse the system is no longer closed, and ΔF can exceed ⟨W⟩. In contrast, the free energy variation is always smaller or equal than the mean energy variation, as expected from Jensen's inequality applied to equation (6).

In order to effectively verify the validity of the FR in equation (6), we used the experimental values of the energy variation probability distribution to obtain ⟨eβΔE ⟩, and we compared it with the values of eβΔF = Z(tf)/Z(0) computed analytically, as shown in figure 4. Notice that the oscillating eβΔF depends only on the time-varying Hamiltonian, therefore, its values are the same in figures 4(a) and (b). In contrast, ⟨eβΔE ⟩ depends on the time-varying Hamiltonian, and on the energy variation probability distribution function, which in turns depends on the time between laser pulses, and the strength of the dissipation, i.e., the mean amplitude of the energy jumps induced by the interaction with a single laser pulse. The fact that work and heat flux are perfectly compensated to allow equation (6) to be valid, is a consequence of the infinite pseudo-temperature of the environment. The agreement between the predicted values of eβΔF and the experimental values of ⟨eβΔE ⟩ verifies the validity of the FR in equation (6).

Figure 4.

Figure 4. Experimental verification of the quantum fluctuation relation for a driven qubit in contact with a reservoir at infinite pseudo-temperature. Blue dots: experimental values of the average exponentiated energy variation ⟨eβΔE ⟩. Orange line: calculated eβΔF . As in the previous figure, (a) τ = 410 ns, and (b) τ = 616 ns.

Standard image High-resolution image

4. Work vanishing at stroboscopic times

Now we consider a driving Hamiltonian with fixed amplitude ω(t) ≡ ω0 and a time varying phase such that,

Equation (10)

In contrast to the case treated in the previous section, here the instantaneous eigenstates $\left\vert {E}_{{\pm}}\left(t\right)\right\rangle $ of $\mathcal{H}\left(t\right)$ are changing in time while the eigenvalues remain constant in time. The Hamiltonian, and accordingly its eigenvectors, are periodic functions of time, with period τθ = 2π/θ. As long as only the discrete times tn = θ are concerned, the discrete-time dynamics of the system, due to its periodicity, can be conveniently expressed in terms of an effective Floquet-type Hamiltonian, reading, in this case

Equation (11)

The according unitary ${U}_{\theta }={\text{e}}^{-\text{i}{\mathcal{H}}_{\theta }{\tau }_{\theta }/\hslash }$ advances the system from time θ to (n + 1)τθ . As opposed to the original Hamiltonian $\mathcal{H}\left(t\right)$, the discrete-time Hamiltonian ${\mathcal{H}}_{\theta }$ is time-independent. Accordingly, in the discrete-time picture no work source appears to be at play.

In our experiment the two-level system is prepared in one of the two eigenstates of ${\mathcal{H}}_{\theta }$, reading

Equation (12)

Equation (13)

where $\alpha =-\mathrm{arctan}\left({\omega }_{0}/\theta \right)$. The probability P↑|↑ to start in the $\left\vert {\uparrow}\right\rangle $ state and find the system again in the $\left\vert {\uparrow}\right\rangle $ state after a time tf reads ${P}_{{\uparrow}\vert {\uparrow}}=1-\frac{{\omega }_{0}^{2}}{{\omega }_{0}^{2}+{\theta }^{2}}\enspace {\mathrm{sin}}^{2}\enspace \frac{t\theta }{2}$. Similarly, it is ${P}_{{\uparrow}\vert {\downarrow}}=\frac{{\omega }_{0}^{2}}{{\omega }_{0}^{2}+{\theta }^{2}}\enspace {\mathrm{sin}}^{2}\enspace \frac{t\theta }{2}$. Note that these quantities are, as expected, periodic with period τθ . In addition, P↑|↑ = 1 and P↓|↑ = 0 at stroboscopic times tn = θ , in accordance with the fact that $\left\vert {E}_{\theta ,{\uparrow}}\right\rangle $ and $\left\vert {E}_{\theta ,{\downarrow}}\right\rangle $ are eigenvectors of the discrete-time Hamiltonian ${\mathcal{H}}_{\theta }$. Figure 5(a) shows the conditional probabilities P↑|↑ and P↑|↓ measured in absence of laser pulses. Note the agreement with the analytic expressions, confirming that $\mathcal{H}\left(t\right)$ well describes the dynamics in absence of laser pulses.

Figure 5.

Figure 5. Conditional probabilities P↑|i for the spin qubit to be measured in the Floquet eigenstate $\left\vert {E}_{\theta ,{\uparrow}}\right\rangle $, for an initial state $\left\vert i\right\rangle =\left\vert {E}_{\theta ,{\uparrow}}\right\rangle $ (blue circles) or $\left\vert i\right\rangle =\left\vert {E}_{\theta ,{\downarrow}}\right\rangle $ (green squares), for a Hamiltonian rotating in the xy-plane with period τθ = 2π/θ. (a) Without laser pulses. τθ = 616 ns, and α = arctan(1/2). The spin qubit evolves in time performing Rabi oscillations. For synchronized stroboscopic measurements, the spin does not evolve (see black markers). (b)–(d) With short laser pulses. Interpulse time τ = τθ . Data are reported as a function of discrete times tf = NL τθ . (b) τθ = 1296 ns, α = π/4; (c) τθ = 616 ns, α = arctan(1/2); and (d) τθ = 308 ns, α = arctan(1/4). For long times, P↑|i approaches a constant value ${P}_{{\uparrow}}^{\infty }$ that determines the reservoir inverse temperature: (b) ${P}_{{\uparrow}}^{\infty }=0.276{\pm}0.003$; (c) ${P}_{{\uparrow}}^{\infty }=0.138{\pm}0.002$; and (d) ${P}_{{\uparrow}}^{\infty }=0.050{\pm}0.003$. The crosses represent the result of Monte-Carlo simulations. Inset of (b): numerical simulation of the evolution between laser pulses. The bare Rabi frequency of the driving mw field is ω0 ≃ (2π)800 kHz in all the cases.

Standard image High-resolution image

In the further experiments, short laser pulses are applied at the discrete times tn = θ , and final energy measurements are performed at a final time tf = NL τθ . Accordingly, the spin free energy difference ΔF is null, and equation (7) takes on the simpler form

Equation (14)

The experimental data are shown together with a Monte-Carlo simulation (see appendix A). Figures 5(b)–(d) shows the impact of the laser pulses on the conditional probabilities P↑|↑ and P↑|↓ as a function of NL. Note how these probabilities converge to the same asymptotic value ${P}_{{\uparrow}}^{\infty }$, meaning that, regardless of the initial condition, the system ends up in the $\left\vert {\uparrow}\right\rangle $ state with probability ${P}_{{\uparrow}}^{\infty }$. In other words, the system reaches a steady state which conforms to a Gibbs state of inverse temperature ${\beta }_{R}=-\frac{1}{\hslash \omega } \mathrm{ln} \left(\frac{{P}_{{\uparrow}}^{\infty }}{1-{P}_{{\uparrow}}^{\infty }}\right)$. The value of ${P}_{{\uparrow}}^{\infty }$, and therefore the value of βR , depends on the ratio ω0/θ. In the adiabatic regime, θω0, the Hamiltonian is approximately proportional to σx , and ${P}_{{\uparrow}}^{\infty }\simeq 1/2$. Hence, in the adiabatic regime the value of βR tends to zero. As the value of θ increases with respect to ω0, the dynamics no longer belongs to the adiabatic regime, and the value ${P}_{{\uparrow}}^{\infty }$ approaches to zero, hence increasing the value of βR . Note that, although the spin is driven by a time dependent Hamiltonian that delivers power continuously, at discrete times in which the energy is measured the system evolves as under the influence of an effective thermal reservoir of inverse temperature βR and as if it were not in contact with any work source. Given these conditions, the mean energy variation is equal to the mean heat flux ⟨Q⟩, as demonstrated in figures 6(a)–(c) by directly comparing ⟨ΔE⟩, extracted from the experimental data, with the mean heat flux calculated with the theoretical model. See appendix B for details on how to calculate ⟨Q⟩ in this case. Remarkably, even when ⟨ΔE⟩ is different from zero, we demonstrate that the energy variation statistics follows the quantum fluctuation relation $\langle {\text{e}}^{-\left(\beta -{\beta }_{R}\right){\Delta}E}\rangle =1$ (equation (14)), as the general theory predicts. Figures 6(d)–(f) shows the experimental values of the quantity $\langle {\text{e}}^{-\left(\beta -{\beta }_{R}\right){\Delta}E}\rangle $, as obtained under various experimental conditions (specifically with various values of τθ and α). The experimental data confirm the validity of the quantum fluctuation relation as expressed in equation (7).

Figure 6.

Figure 6. Measurement of the energy exchange statistics and experimental test of the quantum fluctuation relation in a driven-dissipative system with work vanishing at stroboscopic times. The driving Hamiltonian varies in phase with period τθ , and short laser pulses are applied with interpulse time τ = τθ . (a)–(c) Mean energy variation Δβ⟨ΔE⟩ (experiment, blue dots) and mean heat flux ΔβQ⟩ (numerical simulation, orange diamonds). (d)–(f) Experimental verification of the quantum fluctuation relation ⟨e−ΔβΔE ⟩ = 1: experimental data (blue dots) versus theoretical prediction (orange diamonds). The value of ΔβββR is obtained from the probabilities P(0) and ${P}_{{\uparrow}}^{\infty }$ to find the spin in $\left\vert {E}_{{\uparrow}}\right\rangle $ at initial and asymptotic times, respectively. (a) and (d) P(0) = 0.509, τθ = 1296 ns, α = π/4; (b) and (d) P(0) = 0.303, τθ = 616 ns, α = arctan(1/2); and (c) and (f) P(0) = 0.126, τθ = 308 ns, α = arctan(1/4). The values of ${P}_{{\uparrow}}^{\infty }$ are those reported in figures 5(b)–(d).

Standard image High-resolution image

5. Conclusions

The experimental verification of quantum fluctuation relations for open systems that simultaneously exchange heat—due to a thermal gradient—and work—due to a time-dependent driving—is a formidable challenge, owing to the general difficulty of measuring heat and work separately, while having typically experimental access to their difference ΔE.

With this work we establish a first milestone along the way toward a fully fledged experimental demonstration of fluctuation relations for driven-dissipative systems. The verification becomes feasible (i.e., it requires the statistics of ΔE only) in two special cases: (i) the external environment is at infinite temperature, and (ii) the system is subject to a periodic driving, so that, at multiples of the driving period the total work amounts to zero (while the power is generally non-null). In those cases the fluctuation relations boil down to the simplified expressions in equations (6) and (7), respectively, which we have experimentally verified.

Our experimental platform of choice was a single NV center in diamond. Due to its long coherence time, and the possibility to finely control its state and its interaction with engineered environments, this system naturally lend itself to study non-equilibrium thermodynamics in the quantum regime [24].

Our results represent the first experimental study of quantum fluctuation relation in driven open quantum systems.

Acknowledgments

The authors would like to thank P Cappellaro, F Poggiali, and S Gherardini for useful discussions, and F S Cataliotti and M Inguscio for continuous support.

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Appendix A.: Numerical model of the dynamics

In order to model the qubit dynamics, we separate the evolution into a unitary part, describing the dynamics between consecutive short laser pulses, and the irreversible dynamics that describe the interaction with laser pulses.

The unitary operators of the map correspond to

Equation (A.1)

with tN . We recall that τ is the time between laser pulses. In the case of the Hamiltonian in equation (8), considered in section 3, each unitary operator describes a rotation around the x axis, such that $U\left({t}_{{N}_{\text{L}}-1},{t}_{{N}_{\text{L}}}\right)=\mathrm{exp}\left(-i{\sigma }_{x}{\int }_{{t}_{{N}_{\text{L}}-1}}^{{t}_{{N}_{\text{L}}}}\mathrm{d}{t}^{\prime }\enspace \omega \left({t}^{\prime }\right)\right)$, with ω(t) defined by equation (9). In contrast, in the case studied in section 4, the unitary evolution is described by the Floquet Hamiltonian ${\mathcal{H}}_{\theta }=\frac{\hslash }{2}\left({\omega }_{0}{\sigma }_{x}+\theta {\sigma }_{z}\right)$ (equation (11)) only for discrete times tn = θ , such that, for interpulse times τ = τθ , the unitary operator corresponds to $U\left({t}_{{N}_{\text{L}}-1},{t}_{{N}_{\text{L}}}\right)={U}_{\theta }=\mathrm{exp}\left(-i{\mathcal{H}}_{\theta }{\tau }_{\theta }/\hslash \right)$. Both of these unitary operators can be solved analytically and then included in a numerical model of the dynamics.

The dissipative dynamics associated with each laser pulse consists instead in an energy jump that is modeled as a POVM, with (1 − pa) probability to leave the state unaffected, and a probability pa to absorb photons. Upon absorption, the spin qubit is subject to a projective measurement of σz followed by an optical pumping of the spin modeled via Lindbladian master equation, with a Lindbladian jump operator $\sqrt{{\Gamma}}\left\vert 0\right\rangle \left\langle 1\right\vert $ that transfers spin population to the state $\left\vert 0\right\rangle $. The numerical simulation is then performed by using a Monte-Carlo simulation to determine the result of the POVM, hence determining if the laser pulse is absorbed or not. If the pulse is absorbed, Lindbladian operator is applied to the state of the system. In this way, we simulate single trajectories followed by the system, and we finally average over several realizations to obtain the mean trajectories shown in figures 2 and 5.

Appendix B.: Mean work and mean heat flux

As described in the main text, the mean values of work and heat flux can be calculated as $\langle W\rangle ={\int }_{0}^{{t}_{\text{f}}}\mathrm{d}{t}^{\prime }\enspace \mathrm{Tr}\enspace \left[\rho \left({t}^{\prime }\right)\frac{\mathrm{d}\mathcal{H}\left({t}^{\prime }\right)}{\mathrm{d}{t}^{\prime }}\right]$, and $\langle Q\rangle ={\int }_{0}^{{t}_{\text{f}}}\mathrm{d}{t}^{\prime }\enspace \mathrm{Tr}\enspace \left[\frac{\mathrm{d}\rho \left({t}^{\prime }\right)}{\mathrm{d}{t}^{\prime }}\mathcal{H}\left({t}^{\prime }\right)\right]$ [36]. In this appendix we describe how to compute this quantities for the dynamics described in sections 3 and 4.

B.1. Mean work and mean heat flux for section 3

The spin qubit is driven by the Hamiltonian (equation (8)) $\mathcal{H}\left(t\right)=\frac{\hslash \omega \left(t\right)}{2}{\sigma }_{x}$, where ω(t) is a periodic function such that ω(t) = ω(t + τA). In addition the system interacts with short laser pulses applied at times tf = NL τ, where τ is the time between consecutive pulses and NL is the number of pulses. Assuming that the duration of every short laser pulse interaction is given by δ, then the mean work and mean heat flux after the first laser pulse is given by

Equation (B.1)

Equation (B.2)

The second term in equation (B.1) vanishes because we assume the interaction with the laser pulses to be instantaneous, δτ and δτA, hence the Hamiltonian is kept constant during the interaction with a single laser pulse. Instead, the first term in equation (B.2) is equal to zero because the system evolves under unitary dynamics during the integration interval, meaning that $\frac{\mathrm{d}\rho }{\mathrm{d}t}=\frac{1}{i\hslash }\left[\mathcal{H},\rho \right]$. Then, the cyclic property of the trace implies that $\mathrm{Tr}\left[\frac{\mathrm{d}\rho }{\mathrm{d}t}\mathcal{H}\right]=\frac{1}{i\hslash }\left(\mathrm{Tr}\left[\mathcal{H}\rho \mathcal{H}\right]-\mathrm{Tr}\left[\rho {\mathcal{H}}^{2}\right]\right)=0$. Therefore, for a given final time tf such that NL τ < tf < (NL + 1)τ we can separate the mean work and heat flux as

Equation (B.3)

Equation (B.4)

where

Equation (B.5)

Equation (B.6)

Note that the density operator at any given time t can be written as

Equation (B.7)

where 0 ⩽ P+(t) ⩽ 1 is the population on the eigenstate $\left\vert {E}_{+}\right\rangle $, and q(t) is a complex number describing the coherence of the state in the Hamiltonian basis. We recall that $\left\vert {E}_{{\pm}}\right\rangle $ are eigenstates of σx with eigenvalues ±1, which implies that Tr[ρ(t)σx ] = 2P+(t) − 1. Therefore, equation (B.6) can be rewritten as

Equation (B.8)

In addition, since the Hamiltonian is always parallel to σx , then the population P+(t) during the unitary evolution is kept constant. Therefore, equation (B.5) can be recast as

Equation (B.9)

This means that knowing the population at times tf = is enough to calculate ⟨Wn and ⟨Qn . The population, as we just mentioned, are unaltered by the unitary evolution, so we need to understand how they are affected by the short laser pulses. As described in appendix A, the interaction with a single laser pulse has a probability 1 − pa of leaving the system unaltered, and a probability pa to project the state of the system into one of the eigenstates of σz , and then applying a Lindbladian evolution that pumps population from $\left\vert 1\right\rangle \left\langle 1\right\vert $ to $\left\vert 0\right\rangle \left\langle 0\right\vert $. Hence, the state of the system after absorbing a laser pulse is diagonal in the σz basis, meaning that its populations in the σx basis must be equal to 1/2. Therefore, we can write a recursive equation for the population in the σx basis

Equation (B.10)

Notice that P+(0) = 1/(1 + e), as mentioned in the main text. Therefore, equation (B.10) can be used recursively to obtain

Equation (B.11)

Finally, we can recast equations (B.9) and (B.8) as

Equation (B.12)

Equation (B.13)

Equations (B.12) and (B.13) can be used to calculate the values of ⟨W⟩ and ⟨Q⟩ (equations (B.3) and (B.4)). The mean values of work and heat flux are therefore completely defined by the parameters NL, P+(0), tf, τ/τA, and pa, the latter being the only free parameter fitted to the experimental data in figure 2.

B.2. Mean heat flux for section 4

For the stroboscopic times described in section 4, the unitary evolution of the system is described by ${U}_{\theta }={\text{e}}^{-\text{i}{\mathcal{H}}_{\theta }{\tau }_{\theta }/\hslash }$, where ${\mathcal{H}}_{\theta }$ is the time independent Floquet Hamiltonian in equation (11), and τθ is the time between two consecutive laser pulses, and the time at which the energy of the system is measured. Given that ${\mathcal{H}}_{\theta }$ is time independent, then the mean work measured at stroboscopic times tf = τNL is equal to zero. On the other hand, in analogy to appendix B.1, the mean heat flux measured at those stroboscopic times is

Equation (B.14)

where ${E}_{\theta }=\hslash \sqrt{{\omega }_{0}^{2}+{\theta }^{2}}/2$.

Let us assume that the density operator ρ(τ) represents the state of the system after applying a single short laser pulse to the state ρ(0). As explained in appendix A, the qubit interaction with a short laser pulse is modeled as a POVM with probability (1 − pa) to leave unaltered the state of the qubit and a probability pa to apply a projective measurement of σz followed by Lindbladian dynamics under the jump operator $\sqrt{{\Gamma}}\left\vert 0\right\rangle \left\langle 1\right\vert $. Under this model, the probability to obtain Eθ,↑ = Eθ when measuring the energy of ρ(τ) is

Equation (B.15)

where pd ≡ 1 − e−Γδ is the optical pumping probability associated with the Lindbladian dynamics, $\alpha =-\mathrm{arctan}\left({\omega }_{0}/\theta \right)$ is the angle that defines the ratio between the terms in the Floquet Hamiltonian (equation (11)), ${P}_{{\uparrow}}\left(0\right)=\mathrm{Tr}\left[\left\vert {E}_{\theta ,{\uparrow}}\right\rangle \left\langle {E}_{\theta ,{\uparrow}}\right\vert \rho \left(0\right)\right]$ is the probability associated with the initial state, and where we have defined, to simplify notation, a factor

Equation (B.16)

The relation in equation (B.15) can be recursively applied in order to obtain

Equation (B.17)

Therefore, the mean heat flux in equation (B.14) can be recast as

Equation (B.18)

While pa and pd depend on the laser pulse duration and intensity, all the other parameters are precisely set in experiment, with negligible uncertainty. Note that pd can be extracted from the asymptotic state probability ${P}_{{\uparrow}}^{\infty }={\mathrm{lim}}_{{N}_{\text{L}}\to \infty }\enspace {P}_{{\uparrow}}\left({N}_{\text{L}}\tau \right)=\frac{1}{2}\left(1-\frac{{p}_{\mathrm{d}}}{k}\enspace \mathrm{cos}\enspace \alpha \right)$. Thus, the only free parameter in equation (B.18) is the laser absorption probability pa, which we fit to the experimental data (see figure 5).

Footnotes

  • Generally, as a consequence of the applied driving, the Hamiltonian of each subpart acquires a time dependence, Hi = Hi (t), and each subpart is driven away from the corresponding equilibrium, described by the density operator ${\text{e}}^{-{\beta }_{i}{H}_{i}\left(t\right)}/{Z}_{i}\left(t\right)$, with Zi (t) the according partition function. The free energy ΔFi appearing in equation (1), is the difference between ${F}_{i}\left({t}_{\text{f}}\right)=-{\beta }_{i}^{-1}\enspace \mathrm{ln}\enspace {Z}_{i}\left({t}_{\text{f}}\right)$ the free energy that the system would have if it hypothetically were at equilibrium at the protocol end-time tf, and its actual initial free energy ${F}_{i}\left(0\right)=-{\beta }_{i}^{-1}\enspace \mathrm{ln}\enspace {Z}_{i}\left(0\right)$.

Please wait… references are loading.