Fact-checked by Grok 2 weeks ago

Crooks fluctuation theorem

The Crooks fluctuation theorem is a equality in that relates the probability distributions of work performed during a forward nonequilibrium process to those of the time-reversed process in systems governed by , microscopically reversible dynamics. Formally, it states that the ratio of the probability P_F(W) of observing work W in the forward process to the probability P_R(-W) of observing the negative work -W in the reverse process equals e^{\beta (W - \Delta F)}, where \beta = 1/(k_B T) is the inverse , k_B is Boltzmann's constant, T is the , and \Delta F is the difference between the initial and final states. This theorem, derived for finite systems coupled to a heat bath and driven arbitrarily far from , bridges microscopic reversibility with the emergence of irreversibility by imposing symmetry constraints on fluctuation statistics. Proposed by Gavin E. Crooks in 1999, the theorem generalizes earlier fluctuation relations, such as those by Evans and Searles (1994) for transient fluctuations and Jarzynski's 1997 equality for average exponential work, which it succinctly proves as a corollary: \langle e^{-\beta W} \rangle_F = e^{-\beta \Delta F}. Underpinning these results is the dissipation function (or entropy production) \Omega, an odd quantity under time reversal that quantifies irreversibility along a trajectory, with the theorem asserting P_F(\Omega)/P_R(-\Omega) = e^\Omega. Applicable to classical Markovian processes, the CFT has been extended to quantum regimes, where it governs work fluctuations in driven quantum systems while preserving the core symmetry. The theorem's implications extend to extracting thermodynamic quantities from nonequilibrium experiments, enabling the computation of free energy differences without reaching equilibrium—a feat impossible under classical thermodynamics. For instance, it has been experimentally verified in biopolymer folding, such as RNA unfolding driven by optical tweezers, where forward and reverse work distributions directly yield folding free energies even under fast, far-from-equilibrium conditions. In nanoscale systems, like colloidal particles or molecular motors, the CFT quantifies efficiency limits and fluctuation symmetries in driven transport. More broadly, it informs stochastic thermodynamics in living systems, including human sensorimotor adaptation tasks where nonequilibrium work relations align with behavioral data. These applications underscore the theorem's role in unifying nonequilibrium physics across scales, from quantum devices to biological processes.

Introduction

Statement of the theorem

The Crooks fluctuation theorem provides a fundamental relation in nonequilibrium , stating that for a system initially in A and driven to B via a nonequilibrium process, the ratio of the probability of observing work W in the forward process to the probability of observing work -W in the reverse process is given by \frac{P_F(W)}{P_R(-W)} = e^{\beta (W - \Delta F)}, where P_F(W) is the probability distribution of the work W performed on the system in the forward process, P_R(-W) is the probability distribution of the work -W in the reverse process, \beta = 1/(k_B T) with k_B the and T the of the heat bath, and \Delta F is the difference between states B and A. The forward process involves coupling the system, which begins in canonical at state A, to a time-dependent external that evolves the system over a finite time to state B, where it relaxes back to . The reverse process is defined as the time-reversal of the forward , starting from at state B and driving the system back to state A, ensuring that trajectories in the reverse process are the time-reversed counterparts of those in the forward process. This relation holds under specific assumptions, including microscopic reversibility of the underlying dynamics, which requires that the transition probabilities satisfy in and that the are invariant under time reversal (apart from momentum signs for classical systems). The system must start from a distribution, and the dynamics are assumed to be and Markovian, with the system weakly coupled to a single heat bath at constant . These conditions ensure that the work W is well-defined as the integral of the external parameter's change along the trajectory.

Physical significance

The Crooks fluctuation theorem provides a profound into the statistical nature of thermodynamic processes far from , demonstrating that the probability of a forward process yielding a particular work value is exponentially related to the probability of the time-reversed process yielding the negative of that work. This relation underscores that , such as spontaneous reversal where the extracts work from the surroundings without input, are exponentially suppressed compared to their counterparts, ensuring the second law of holds on average while permitting temporary violations through fluctuations. Such suppression arises because the theorem quantifies the relative likelihoods via the dissipation function, which weights improbable outcomes by factors of e^\Omega, where \Omega measures the deviation from reversibility. This interpretation aligns with the second law by showing that average entropy production remains non-negative, \langle \Omega \rangle \geq 0, yet fluctuations enable scenarios where entropy decreases locally, reflecting the probabilistic underpinnings of irreversibility in small-scale systems. The dissipation function itself acts as a key metric for irreversibility, capturing the irreversible work or entropy generation that distinguishes nonequilibrium trajectories from quasistatic ones, and it directly ties to total entropy production as \Omega = \beta (W - \Delta F), where negative values indicate apparent second-law violations that are statistically rare. In this way, the theorem bridges with macroscopic irreversibility, explaining why processes like gas expansion appear irreversible despite underlying time-reversible dynamics. Furthermore, the theorem's physical implications extend to practical by enabling the computation of differences, \Delta F, directly from distributions of nonequilibrium work performed in forward and reverse protocols, without requiring slow, quasistatic . This connection arises because the in work probabilities intersects at W = \Delta F, allowing experimentalists to infer thermodynamic potentials from fast, driven processes that would otherwise be inaccessible. Thus, it transforms fluctuations from mere into a resource for probing properties, with broad relevance to mesoscopic systems where is significant.

Historical context

Discovery and key publications

The Crooks fluctuation theorem was discovered by Gavin E. Crooks in 1999 during his graduate studies in the Department of Chemistry at the . Crooks' research was driven by the challenges in nonequilibrium statistical mechanics, particularly the desire to establish general relations governing fluctuations in systems driven far from equilibrium. The seminal publication introducing the appeared in 1999 as "Entropy production fluctuation theorem and the nonequilibrium work for free energy differences," published in Physical Review E. In this work, Crooks derived a connecting the probabilities of forward and reverse nonequilibrium processes, while also proving a nonequilibrium work for estimating differences. The paper built directly on earlier developments in during the 1990s, including the Evans-Searles dissipation introduced in 1994 and 1995, which addressed transient fluctuations in steady-state systems, and Jarzynski's 1997 equality relating nonequilibrium work to equilibrium .

Early experimental validations

The first experimental validation of the Crooks fluctuation theorem was achieved by Collin et al. in 2005, using to manipulate single molecules. They focused on an RNA hairpin targeting HIV receptor mRNA and an RNA three-helix junction from E. coli 16S rRNA, subjecting these molecules to repeated nonequilibrium unfolding and refolding cycles by varying the applied force at controlled pulling speeds, such as 1.5, 7.5, and 20 nm s⁻¹. Work values were computed by integrating the force-extension curves for forward (unfolding) and reverse (refolding) processes, enabling the construction of work probability distributions. Key results demonstrated the theorem's prediction through the exponential ratio of forward to reverse work probabilities, with the logarithm of this ratio yielding a straight line of slope approximately 1.06 across a range of dissipated work up to about 100 k_B T, confirming the relation under both weak and strong nonequilibrium conditions. The forward and reverse work histograms exhibited the expected , overlapping and crossing at the difference ΔG, which was accurately recovered—for instance, ΔG = 62.8 ± 1.5 k_B T for the RNA hairpin at 7.5 nm s⁻¹—allowing direct extraction of folding free energies without assuming . This approach also quantified effects like magnesium stabilization, yielding ΔG = -31.7 ± 2 k_B T. Challenges in these early validations included managing large statistical uncertainties in far-from-equilibrium regimes due to the rarity of low-dissipation events, as well as instrumental drifts at low loading rates below a few s⁻¹, which complicated near-equilibrium measurements. Precise control of nonequilibrium protocols and high-resolution tracking of single-molecule trajectories were essential to accumulate sufficient data for reliable distributions, limiting applicability to systems with measurable mechanical extensions and moderate energy dissipation.

Mathematical formulation

System setup and definitions

The Crooks fluctuation theorem applies to a classical in contact with a heat bath maintained at constant inverse β, evolving in the under either deterministic dynamics or stochastic processes. The is specified by coordinates x, which include positions and momenta, and a time-dependent control parameter λ(t) that externally drives the process from an initial A (λ(-τ) = λ_A) to a final B (λ(τ) = λ_B) over a finite 2τ. The H(x, λ(t)) governs the , and the λ(t) is typically smooth and symmetric in time for the forward process. A forward trajectory x(t) = {x(-τ), ..., x(τ)} begins with x(-τ) sampled from the canonical equilibrium distribution ρ_A(x) ∝ exp[-β H(x, λ_A)] at state A and evolves forward in time according to the system's dynamics under λ(t). The corresponding reverse trajectory \tilde{x}(t) = {x(τ), ..., x(-τ)} is the time-reversed path, starting from the equilibrium distribution ρ_B(x) ∝ exp[-β H(x, λ_B)] at state B, driven by the time-reversed protocol λ(-t), and with time-odd variables (such as momenta) reversed in sign to ensure consistency with microscopic reversibility. The probability of observing a forward trajectory is P_F[x(t)] = ρ_A(x(-τ)) Π P[x(t+Δt)|x(t), λ(t)], where the transition probabilities satisfy detailed balance in equilibrium. The work W performed on the along a forward is defined via as W = ΔE - Q, where ΔE is the change in the 's and Q is the absorbed from the ; equivalently, in terms of the protocol, it corresponds to the energy input due to changes in λ(t). The dissipation function, denoted as the ω (or σ = β ω in some notations), quantifies irreversibility and is given by ω = β(W - ΔF), where ΔF = F_B - F_A is the difference between states B and A. This follows from ω = ln[ρ_A(x(-τ))/ρ_B(x(τ))] - β Q, linking changes to contributions. The framework assumes Markovian dynamics with microscopic reversibility, ensuring that transition probabilities obey P[x(t)|x(t'), λ(t)] / P[x(t')|x(t), λ(t')] = exp[β (H(x(t'), λ(t')) - H(x(t), λ(t))) ] in , which enforces . Initial conditions require sampling from the exact at the start (or end for reverse processes), with the system assumed to relax quickly to before the protocol begins if needed; of the underlying dynamics ensures representative trajectory ensembles. These setups hold for both systems with weak coupling to the bath and overdamped models, such as Langevin equations.

Derivation outline

The derivation of the Crooks fluctuation theorem proceeds from the framework of stochastic dynamics satisfying , where the probability of observing a specific (path) in is considered for both forward and reverse processes. Consider a classical in with a heat bath at inverse temperature \beta = 1/(k_B T), driven by a time-dependent \lambda(t) that changes the H_\lambda(x) from initial value \lambda_A to final value \lambda_B over time interval [0, \tau] in the forward process. The reverse process uses the time-reversed \bar{\lambda}(t) = \lambda(\tau - t), starting from \lambda_B to \lambda_A. The forward process begins with the sampled from the equilibrium \rho_A(x) = e^{-\beta H_A(x)} / Z_A at \lambda = \lambda_A, while the reverse begins from \rho_B(x) = e^{-\beta H_B(x)} / Z_B at \lambda = \lambda_B, with difference \Delta F = F_B - F_A = -\frac{1}{\beta} \ln(Z_B / Z_A). The probability of a forward path x(t) is P_F[x(t)] = \rho_A(x(0)) \, P[x(t) \mid \lambda(t)], where P[x(t) \mid \lambda(t)] is the (conditional path probability) given the . Similarly, for the reverse path \tilde{x}(t) = x(\tau - t), P_R[\tilde{x}(t)] = \rho_B(\tilde{x}(0)) \, P[\tilde{x}(t) \mid \bar{\lambda}(t)] = \rho_B(x(\tau)) \, P[\tilde{x}(t) \mid \bar{\lambda}(t)]. , a consequence of in the underlying dynamics (e.g., Langevin or Markovian jumps), implies that the ratio of the propagators satisfies \frac{P[x(t) \mid \lambda(t)]}{P[\tilde{x}(t) \mid \bar{\lambda}(t)]} = e^{-\beta Q[x(t)]}, where Q[x(t)] is the absorbed by the system along the forward path x(t). This relation arises because time reversal flips the direction of irreversible currents, leaving the reversible parts unchanged, and the dissipative contribution corresponds to the heat flow. The work W[x(t)] done on the system along the path is defined as W = \int_0^\tau \frac{\partial H_{\lambda(t)}(x(t))}{\partial \lambda} \dot{\lambda}(t) \, dt. By the first law of thermodynamics applied to the path, the change in the system's energy satisfies H_B(x(\tau)) - H_A(x(0)) = Q[x(t)] + W[x(t)], so Q[x(t)] = H_B(x(\tau)) - H_A(x(0)) - W[x(t)]. Substituting into the propagator ratio gives \frac{P[x(t) \mid \lambda(t)]}{P[\tilde{x}(t) \mid \bar{\lambda}(t)]} = e^{-\beta \left( H_B(x(\tau)) - H_A(x(0)) - W[x(t)] \right)}. The full path probability ratio is then \frac{P_F[x(t)]}{P_R[\tilde{x}(t)]} = \frac{\rho_A(x(0))}{ \rho_B(x(\tau)) } \cdot e^{-\beta \left( H_B(x(\tau)) - H_A(x(0)) - W[x(t)] \right)}. Inserting the equilibrium forms, \rho_A(x(0)) / \rho_B(x(\tau)) = e^{\beta \left( H_B(x(\tau)) - H_A(x(0)) \right)} e^{-\beta \Delta F}, the boundary terms cancel exactly: \frac{P_F[x(t)]}{P_R[\tilde{x}(t)]} = e^{\beta \left( H_B(x(\tau)) - H_A(x(0)) \right)} e^{-\beta \Delta F} \cdot e^{-\beta \left( H_B(x(\tau)) - H_A(x(0)) - W[x(t)] \right)} = e^{\beta \left( W[x(t)] - \Delta F \right)}. This ratio depends only on the work W[x(t)], not on other path details. To obtain the theorem, consider the work distributions: the forward work probability is P_F(W) = \int \mathcal{D}x(t) \, P_F[x(t)] \, \delta \big( W - W[x(t)] \big), where \int \mathcal{D}x(t) denotes path integration (or summation for discrete cases). For the reverse, P_R(-W) = \int \mathcal{D}\tilde{x}(t) \, P_R[\tilde{x}(t)] \, \delta \big( -W - W_R[\tilde{x}(t)] \big). Changing variables to the forward path via \tilde{x}(t) \mapsto x(t) (with Jacobian 1 under time reversal for even variables), and noting that the reverse work satisfies W_R[\tilde{x}(t)] = -W[x(t)], yields P_R(-W) = \int \mathcal{D}x(t) \, P_R[\tilde{x}(t)] \, \delta \big( W + W[x(t)] \big) = \int_{\{x : W=W\}} \mathcal{D}x(t) \, P_R[\tilde{x}(t)]. Thus, \frac{P_F(W)}{P_R(-W)} = \frac{ \int_{\{x : W=W\}} \mathcal{D}x(t) \, P_F[x(t)] }{ \int_{\{x : W=W\}} \mathcal{D}x(t) \, P_R[\tilde{x}(t)] } = \left\langle \frac{P_F[x(t)]}{P_R[\tilde{x}(t)]} \right\rangle_{\{x : W=W\}} = e^{\beta (W - \Delta F)}, where the average is over all paths with fixed work W. Since the ratio is identical for every such path, the distributions satisfy the symmetry relation \frac{P_F(W)}{P_R(-W)} = e^{\beta (W - \Delta F)}. This establishes the Crooks fluctuation theorem. Integrating over W further implies the Jarzynski equality \langle e^{-\beta W} \rangle_F = e^{-\beta \Delta F}, but the direct symmetry follows from the path-level relation.

Connections to other fluctuation theorems

Relation to Jarzynski equality

The Crooks fluctuation theorem implies the Jarzynski equality through a mathematical integration over the forward work distribution. Specifically, the ratio of probabilities from the Crooks relation, \frac{P_F(W)}{P_R(-W)} = e^{\beta (W - \Delta F)}, where P_F(W) and P_R(W) are the probability densities of work W in forward and reverse processes, \beta = 1/(k_B T), and \Delta F is the free energy difference, leads to the average \langle e^{-\beta W} \rangle_F = e^{-\beta \Delta F} upon integrating both sides over W. This derivation, presented in the original Crooks paper, demonstrates that the distributional symmetry of the theorem directly yields the exponential average characterizing the Jarzynski equality. While the Jarzynski equality, formulated in 1997, relates the exponential average of work solely from forward nonequilibrium processes to the equilibrium free energy difference, \Delta F = -k_B T \ln \langle e^{-\beta W} \rangle_F, the Crooks theorem extends this by symmetrically linking the full distributions of forward and reverse work. This bidirectional relation in Crooks provides a deeper insight into fluctuation symmetries, complementing the one-sided averaging in Jarzynski by revealing how rare reverse events balance forward ones. Historically, Jarzynski's 1997 result preceded Crooks' 1999 theorem by two years, yet the latter's framework unifies and generalizes the former by establishing the underlying path probability symmetry that underpins the . Together, they form complementary tools in , with Crooks enabling verification of Jarzynski through direct comparison of forward and reverse ensembles.

Comparison with Evans-Searles theorem

The Evans-Searles dissipation fluctuation theorem, introduced in 1994, provides a relation for the probability ratio of observing a given value of the time-averaged function σ and its negative counterpart during transient processes in systems starting from nonequilibrium initial states. Specifically, it states that \frac{P(\sigma)}{P(-\sigma)} = e^{\sigma t}, where t is the duration of the process and σ quantifies the average rate of irreversible scaled appropriately. This theorem applies to a broad class of deterministic dynamics with even parity under time reversal, emphasizing the emergence of the second law from in finite systems driven out of . In contrast, the Crooks fluctuation theorem focuses on nonequilibrium processes that connect two states, relating the distributions of work performed in forward and reverse protocols to the difference between those states. A primary distinction lies in their scopes: while the Evans-Searles theorem accommodates arbitrary initial distributions and uses the dissipation function to capture general irreversibility in transient evolutions, the Crooks theorem is tailored to protocols starting and ending at , leveraging work as the central quantity to enable estimation from nonequilibrium data. This makes the Crooks theorem particularly suited for thermodynamic computations in controlled switching processes, whereas the Evans-Searles formulation offers a more general framework for analyzing dissipation without requiring endpoints. Despite these differences, both theorems belong to the class of integral fluctuation theorems, which express the bias toward positive or work in nonequilibrium through probability ratios. The Crooks can be viewed as a specialized instance of the Evans-Searles relation when applied to equilibrium-initialized systems where the function aligns with the work done, allowing direct links to differences. This overlap underscores their shared role in unifying with macroscopic irreversibility across nonequilibrium .

Applications and experimental tests

Biophysical systems

The Crooks fluctuation theorem has been instrumental in studying the nonequilibrium thermodynamics of RNA folding, enabling the extraction of equilibrium free energy differences from distributions of mechanical work performed during folding and unfolding processes. In pioneering experiments using optical tweezers, researchers manipulated a small RNA hairpin and an RNA three-helix junction, verifying the theorem across weak and strong nonequilibrium driving speeds. These measurements confirmed the symmetry relation between forward and reverse work distributions, allowing precise recovery of folding free energies, such as the thermodynamic stabilization contributed by magnesium ions to the three-helix structure. Extensions of this approach have mapped detailed folding landscapes for more complex RNA motifs, validating free energy differences (ΔF) along reaction coordinates by analyzing work histograms at varying extension points. For instance, nonequilibrium pulling experiments on RNA three-way junctions have reconstructed position-dependent free energies, revealing barriers and minima that align with equilibrium models while accounting for force-induced pathway alterations. These studies demonstrate the theorem's robustness in probing hierarchical folding mechanisms, where secondary structures form prior to tertiary assembly, providing insights into RNA's evolutionary design principles. In protein unfolding, optical tweezer experiments on —a with two lobes connected by a central —have quantified nonequilibrium work to elucidate conformational changes under mechanical force. By tethering the protein via engineered residues and applying controlled stretching and relaxation cycles, researchers constructed a folding network with four intermediates, including off-pathway states driven by non-native interdomain contacts. The Crooks theorem was applied to work distributions from these transitions, revealing cooperative and anticooperative domain interactions that influence the overall folding barrier, with landscapes showing energetic dependencies on pulling geometry. This approach highlighted how force selectively unfolds one lobe while monitoring the other's state, uncovering anisotropic mechanical stability. Applications to enzyme cycles, particularly in molecular motors, leverage the theorem to estimate free energies from fluctuation data during ATP hydrolysis-driven processes. For the rotary motor F1-ATPase, fluctuation relations were tested on single-molecule rotations, where ATP binding and hydrolysis power 120° substeps; analysis of angular work distributions confirmed the theorem's predictions and quantified the torque generated, linking it to the free energy released per hydrolysis event (approximately 20 kBT under physiological conditions). These measurements illustrate how the theorem dissects energy transduction efficiency in cyclic nonequilibrium processes, distinguishing dissipative losses from useful mechanical work in motor stepping.

Nanoscale and colloidal experiments

Experiments with colloidal particles have provided key validations of the Crooks fluctuation theorem in driven nonequilibrium systems. In a seminal study, a single overdamped colloidal sphere was confined in a time-dependent nonharmonic optical trap, where the potential was modulated to drive the particle out of equilibrium. By tracking the particle's trajectories and computing the work distributions for forward and reverse protocols, researchers verified the relation of the theorem with high accuracy, demonstrating its applicability to under external driving. Nanomechanical resonators have enabled precise measurements of work fluctuations in harmonic traps, testing the theorem at the nanoscale. Using a torsion pendulum as a mechanical oscillator coupled to a thermal bath, experiments quantified the work performed during cyclic driving of the trap stiffness. The resulting work probability distributions for forward and backward processes obeyed the Crooks relation, confirming the theorem's predictions for harmonic systems and highlighting finite-time corrections to the fluctuations. Recent advances have integrated techniques to analyze trajectories in out-of-equilibrium colloidal assemblies, enhancing tests of the theorem. Neural network-based feedback protocols have been developed to control fluctuating nanosystems, such as colloidal particles in optical traps, optimizing work extraction while adhering to fluctuation symmetries. These approaches, applied to self-assembling colloidal structures under nonequilibrium driving, allow for efficient estimation of and validation of Crooks' relation without exhaustive trajectory sampling. In 2024, the theorem was experimentally tested in a single-ion Paul trap, confirming its validity for nonequilibrium processes at different speeds and effective temperatures, further demonstrating applicability to nanoscale .

Extensions and generalizations

Quantum formulations

The quantum formulation of the Crooks fluctuation theorem adapts the classical relation to , where work is defined through protocols that account for the non-commutativity of observables and the effects of . A key approach, as detailed by Deffner and Campbell, employs the two-point (TPM) scheme, in which projective measurements of the system's energy are performed at the initial and final times of the driving protocol to determine the work W = E_m - E_n, where E_n and E_m are the measured eigenvalues of the at times t=0 and t=\tau, respectively. Alternatively, full statistics via the G(\lambda) = \langle e^{i\lambda H(\tau)} U(\tau,0) e^{-i\lambda H(0)} U^\dagger(\tau,0) \rangle provides the for the work without explicit measurements, enabling the extraction of probability densities through . The theorem itself retains its classical form in the quantum setting: \frac{P_F(W)}{P_R(-W)} = e^{\beta (W - \Delta F)}, where P_F(W) and P_R(-W) are the probability densities of work for the forward and reverse processes, \beta = 1/kT is the inverse temperature, and \Delta F is the equilibrium free energy difference between initial and final states. This holds for both inclusive work, which incorporates the full energy change including measurement-induced transitions, and exclusive work, which isolates the coherent evolution contribution by marginalizing over intermediate states. The TPM protocol ensures the theorem's validity for isolated quantum systems driven out of equilibrium, bridging classical nonequilibrium thermodynamics with quantum mechanics. Significant challenges arise in quantum implementations due to inherent irreversibility and backaction. Unlike classical systems, quantum evolutions are unitary and reversible, but projective s collapse the wavefunction, introducing irreversibility that alters the system's and beyond classical predictions. backaction further complicates work statistics by correlating initial and final s, potentially violating in the absence of careful design, such as using weak or continuous monitoring to mitigate effects. These issues necessitate distinctions between inclusive and exclusive frameworks to accurately capture quantum-specific fluctuations. In applications to quantum heat engines, the quantum Crooks theorem quantifies work and fluctuations during cyclic processes like the quantum , revealing bounds on distributions influenced by nonadiabatic driving and coherence. For instance, it predicts symmetric probability ratios for forward and reverse strokes, enabling diagnostics of engine performance and verification of thermodynamic consistency in finite-time operations, though backaction can induce divergent peaks that require experimental averaging over ensembles. This framework supports the design of fluctuation-minimized quantum devices, such as those using oscillators or qubits.

Applications to steady states

The Crooks fluctuation theorem, originally formulated for transient processes between states, has been extended to nonequilibrium s (NESS) in driven systems, where forward and reverse fluxes exhibit a relation. In , Crooks introduced a steady-state version applicable to systems under time-symmetric periodic or constant external fields, such as sheared fluids, stating that the ratio of probabilities for observing positive and negative values of the dissipation function v (a measure of total ) satisfies \frac{P(+v)}{P(-v)} = e^{v}, where v \approx \beta Q for long observation times and Q is the dissipated to the at inverse \beta = 1/k_B T. This holds for trajectories in the steady state over integer cycles of the protocol, ensuring the theorem's validity without requiring initial conditions. Subsequent work by Crooks in 2000 generalized -ensemble averages to far-from-equilibrium driven systems, including open systems maintained in NESS by continuous external forcing, by defining ratios of probabilities that incorporate work and changes symmetrically under time reversal. Independently, Maes developed extensions for systems in NESS, relating forward and reverse transition rates to rates via a fluctuation relation \frac{P(\sigma)}{P(-\sigma)} = e^{\sigma t} (where \sigma is the average rate and t the observation time), applicable to Markov processes with local in open environments like periodically driven colloidal particles. These steady-state formulations have been applied to quantify fluctuations in and irreversibility in driven nanoscale devices. In thermoelectric systems, such as quantum point contacts operating in steady-state heat-to-work conversion, the constrains and fluctuations, revealing that efficiency variations obey a generalized fluctuation-dissipation , with average output bounded by , as demonstrated in models of mesoscopic conductors where charge and satisfy \langle \Delta W \rangle \geq 0 on average but fluctuate symmetrically around the second law. For , such as or tracers in living baths maintained in NESS by injection, the quantifies trajectory irreversibility through the log-ratio of forward and backward path probabilities \Delta \Sigma, which decomposes into thermal and with the active noise, enabling measurement of nonreciprocal in systems like harmonically trapped Brownian particles under Ornstein-Uhlenbeck active forcing, where active fluctuations reduce but do not eliminate irreversibility growth over time.

References

  1. [1]
    [cond-mat/9901352] The Entropy Production Fluctuation Theorem ...
    Jul 29, 1999 · The Entropy Production Fluctuation Theorem and the Nonequilibrium Work Relation for Free Energy Differences. Authors:Gavin E. Crooks.Missing: original | Show results with:original
  2. [2]
    Fully Quantum Fluctuation Theorems | Phys. Rev. X
    Feb 6, 2018 · The starting point is Crooks fluctuation theorem, which characterizes the randomness of the work needed for driving thermodynamic processes. A ...
  3. [3]
    Verification of the Crooks fluctuation theorem and recovery of RNA ...
    Sep 8, 2005 · Here we show that the Crooks fluctuation theorem can be used to determine folding free energies for folding and unfolding processes occurring in ...Missing: review - | Show results with:review -
  4. [4]
    [PDF] The Crooks Fluctuation Theorem Derived for Two-Dimensional Fluid ...
    The Crooks fluctuation theorem is a relation from nonequilibrium thermodynamics, which has its typical applications in nanoscale systems. It quantifies the ...
  5. [5]
    Thermodynamic fluctuation theorems govern human sensorimotor ...
    Jan 17, 2023 · Here, we test Jarzynski's equality and Crooks' fluctuation theorem experimentally in a human sensorimotor adaptation task. In this context, the ...<|control11|><|separator|>
  6. [6]
    Entropy production fluctuation theorem and the nonequilibrium work ...
    Sep 1, 1999 · Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences ... E 60, 2721 – Published 1 ...
  7. [7]
    Nonequilibrium Equality for Free Energy Differences | Phys. Rev. Lett.
    Nonequilibrium Equality for Free Energy Differences. C. Jarzynski ... 78, 2690 – Published 7 April, 1997. DOI: https://doi.org/10.1103/PhysRevLett.78.2690.
  8. [8]
  9. [9]
  10. [10]
    [PDF] Fluctuation Theorems - arXiv
    Sep 26, 2007 · Another, the Crooks Fluctuation Theorem (Crooks FT) ... In contrast Liphardt's RNA- unfolding experiment importantly demonstrated the application ...
  11. [11]
    Folding Free Energy Determination of an RNA Three-Way Junction ...
    A useful approach to estimate the folding free energy from nonequilibrium experiments is the Crooks fluctuation theorem (CFT) [23] and its corollary, the ...
  12. [12]
    The Complex Folding Network of Single Calmodulin Molecules
    ### Summary of Crooks Fluctuation Theorem in Calmodulin Folding/Unfolding
  13. [13]
    Thermodynamics of a Colloidal Particle in a Time-Dependent ...
    ... a detailed fluctuation theorem are verified with good accuracy ... Crooks, C. J. Bustamante, and J. Liphardt, Proc. Natl. Acad. Sci. U.S.A. 101 ...
  14. [14]
    Work Fluctuation Theorems for Harmonic Oscillators | Phys. Rev. Lett.
    Oct 5, 2006 · In this Letter, we investigate, within the context of the fluctuation theorems (FTs), the work fluctuations of a harmonic oscillator in contact ...Missing: nanomechanical resonators silicon Crooks
  15. [15]
    Demon in the Machine: Learning to Extract Work and Absorb ...
    Apr 10, 2023 · Driving a fluctuating molecular machine or nanoscale system out of equilibrium costs energy, on average, according to the second law of ...Missing: analysis post-
  16. [16]
  17. [17]
  18. [18]
    The Tasaki–Crooks quantum fluctuation theorem - IOPscience
    Jun 12, 2007 · The quantum version of the Crooks fluctuation theorem is shown to emerge almost immediately by the mere application of an inverse Fourier transformation.
  19. [19]
    Quantum fluctuation theorems and power measurements - IOPscience
    Work in the paradigm of the quantum fluctuation theorems of Crooks and Jarzynski is determined by projective measurements of energy at the beginning and end ...
  20. [20]
    Experimentally reducing the quantum measurement back action in ...
    Mar 1, 2019 · We report the experimental investigation of reducing quantum measurement back action in work distribution using CMs on two identically prepared qubit states.
  21. [21]
    [PDF] fluctuations and correlations of quantum heat engines
    ... heat fluctuation theorem. Then we move our attention to quantum heat engines. Stability is an important property, which one has to take into account when ...
  22. [22]
    [PDF] Entropy production fluctuation theorem and the nonequilibrium work ...
    The final result is the entropy production fluctuation theorem,. Eq. 共2兲. The theorem readily generalizes to other ensembles. As an example, consider an ...
  23. [23]
  24. [24]
    [PDF] THE FLUCTUATION THEOREM FOR STOCHASTIC SYSTEMS - arXiv
    The Fluctuation Theorem describes the probability ratio of observing trajectories that satisfy or violate the second law of thermodynamics.
  25. [25]
    [PDF] arXiv:1806.04956v2 [cond-mat.stat-mech] 23 Feb 2019
    Feb 23, 2019 · Irreversibility in active matter systems: Fluctuation theorem and mutual information. Lennart Dabelow. Fakultät für Physik, Universität ...