Fact-checked by Grok 2 weeks ago

Microscopic reversibility

Microscopic reversibility is a fundamental principle in and asserting that, at , every microscopic process occurs at the same rate as its exact reverse, ensuring in molecular transitions. This principle applies specifically to elementary steps in reaction mechanisms, where the forward and reverse pathways involve identical sequences of molecular configurations, derived from the time-reversibility of classical or quantum mechanical . The concept was formalized by in 1925, building on earlier contributions from researchers such as Marcelin (1915), Langmuir (1916), and Einstein (1917), who explored molecular rates and conditions. Tolman postulated that in , the number of molecules traversing any specific from one to another equals the number traversing the reverse in unit time, excluding self-sustaining cycles. Although not directly derivable from the second law of , it aligns with it by prohibiting and supporting the equality of forward and reverse fluxes. In practice, microscopic reversibility underpins the validity of constants for reactions, as it requires that mechanisms be mirror images in both directions, influencing fields from to simulations. It has been extended to , where it maintains balance in traversals despite wave-like behaviors. This principle also connects to broader thermodynamic concepts, such as in nonequilibrium processes, highlighting its role in bridging microscopic dynamics and macroscopic observables.

Historical Development

Origins in Kinetic Theory

The concept of microscopic reversibility emerged in the late within the framework of kinetic theory, particularly through Ludwig Boltzmann's analysis of gas molecule collisions. In his 1872 paper "Weitere Studien über das Wärmegleichgewicht unter Gasmolekülen," Boltzmann examined the dynamics of gases composed of molecules undergoing random collisions, positing that the probability of a specific sequence of collisions is equal to that of its reverse sequence, assuming identical initial conditions for the reversed paths. This insight underscored the time-reversibility inherent in molecular interactions, laying early groundwork for understanding how microscopic processes could underpin macroscopic behavior. Boltzmann's derivation of the H-theorem in the same work appeared to demonstrate irreversibility, as it showed that the function H—defined as an integral over the velocity distribution—decreases monotonically toward equilibrium, corresponding to an increase in entropy. However, this result relied on the stosszahlansatz, or molecular chaos assumption, which posits that incoming velocities to a collision are uncorrelated, introducing a statistical element that reconciles the theorem's apparent irreversibility with the underlying time-reversible nature of individual collisions. Without this assumption, the H-theorem would not hold, highlighting how probabilistic interpretations bridge microscopic reversibility and macroscopic irreversibility in dilute gases. A concrete illustration of this reversibility arises in elastic collisions between , the model Boltzmann employed for gas molecules. In such collisions, conservation of and ensures that the post-collision velocities can be exactly reversed by negating all velocity components, yielding time-reversible trajectories that retrace the original paths under time reversal. This property holds for pairwise interactions in kinetic theory, where spheres interact only upon contact without dissipative forces, emphasizing the of forward and backward processes at the microscopic level. These ideas faced a significant challenge from the Loschmidt paradox, articulated by Josef Loschmidt in 1876. Loschmidt argued that since the laws governing molecular motions are time-symmetric, reversing all velocities in a at should produce a reversed motion leading back to a non-equilibrium state, thereby contradicting the H-theorem's prediction of monotonic increase and questioning the derivation of irreversibility from reversible dynamics. Boltzmann responded by emphasizing the extreme improbability of such reversed initial conditions occurring naturally, thus preserving the statistical validity of his approach within kinetic theory.

Formalization in Equilibrium Thermodynamics

The formalization of microscopic reversibility within equilibrium thermodynamics emerged in the late 19th and early 20th centuries, building on precursors from kinetic theory to establish it as a cornerstone principle for chemical systems at . Ludwig Boltzmann's earlier work in the 1870s on provided a foundational precursor by modeling as a balance in molecular collisions, though it focused primarily on gases. This set the stage for applications to broader thermodynamic contexts, particularly in chemical equilibria. A pivotal contribution came from Jacobus van't Hoff in 1884, who applied the concept to chemical by viewing them as dynamic balances between forward and reverse molecular processes. In his seminal work, van't Hoff linked this reversibility to in dilute solutions, demonstrating that the pressure exerted by solutes follows the , implying reversible akin to molecular in gases. This insight, later elaborated in his studies, underscored how osmotic phenomena reflect underlying reversible chemical affinities, enabling quantitative predictions of constants from thermodynamic data. In 1901, Rudolf Wegscheider provided a rigorous thermodynamic statement of the principle, asserting that at , the rate of every elementary chemical process must equal the rate of its reverse. Wegscheider derived this from the conditions for simultaneous equilibria in homogeneous systems, showing it as a necessary consequence of the second law to avoid in reaction cycles. His formulation extended reversibility beyond simple binaries to , ensuring consistency between kinetic rates and thermodynamic potentials. Building on these ideas, René Marcelin in 1915 applied the principle to chemical reaction rates, treating reactions as probabilistic events balanced at equilibrium. Irving Langmuir in 1916 explored its implications for adsorption and surface reactions, emphasizing equal forward and reverse molecular rates. Albert Einstein in 1917 further connected it to diffusion and equilibrium in his work on molecular kinetics. Gilbert N. Lewis further clarified the principle in 1925, emphasizing its application to chemical reactions where forward and reverse paths must follow identical microscopic mechanisms. Lewis argued that equilibrium arises from the exact balancing of these symmetric pathways, rejecting any asymmetry that would violate thermodynamic stability. This clarification resolved ambiguities in reaction kinetics, affirming that detailed reversibility governs the approach to equilibrium. Richard C. Tolman, also in 1925, discussed the principle of microscopic reversibility and its close connection to the second of thermodynamics, demonstrating its necessity for maintaining entropy increase in isolated systems without perpetual motion. Tolman showed that any violation at the microscopic level would lead to inconsistencies with macroscopic irreversibility, thus proving the principle essential for thermodynamic consistency across scales, though not directly derivable from the second . His proof bridged and equilibrium theory, solidifying reversibility as a fundamental postulate.

Theoretical Foundations

Time-Reversibility of Microscopic Dynamics

Time-reversibility, also known as , refers to the invariance of physical laws under the transformation t \to -t, meaning that if a process evolves forward in time according to the governing equations, the time-reversed process evolves equally validly backward in time. This symmetry underpins the microscopic dynamics of particles in isolated systems, ensuring that trajectories can be exactly reversed without violating the fundamental laws of motion. In , Newton's second law, \mathbf{F} = m \mathbf{a}, exemplifies time-reversibility, as \mathbf{a} = d^2 \mathbf{x}/dt^2 depends on even powers of , causing trajectories to reverse precisely under time inversion while forces remain unchanged. The Hamiltonian formulation further reinforces this, where the dq/dt = \partial H / \partial p and dp/dt = -\partial H / \partial q generate reversible flows in , preserving the incompressibility of phase volume as stated by : d(\rho)/dt = 0, where \rho is the phase space . This theorem implies that the evolution of an ensemble of particles is deterministic and reversible in a , with no net loss of over time. In , the time-dependent , i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi, exhibits T-invariance for time-independent Hamiltonians in closed systems, as the time-reversal operator (anti-unitary, involving complex conjugation) maps solutions to valid reversed-time solutions. Conservation laws for energy and linear momentum in closed systems further guarantee this reversibility, as these quantities remain constant under , prohibiting irreversible changes in isolated microscopic dynamics.

Principle of Microscopic Reversibility at Equilibrium

The principle of microscopic reversibility at equilibrium states that, in a at , every microscopic process occurs with the same average frequency as its exact reverse. This pairwise equality ensures that no net flux exists between any pair of microstates, maintaining the . Formally, if A and B represent distinct microstates, the principle implies : the transition rate from A to B times the equilibrium probability of A, w(A \to B) P_{eq}(A), equals the reverse, w(B \to A) P_{eq}(B). In the , P_{eq} \propto e^{-E/kT}, so w(A \to B)/w(B \to A) = e^{-(E_B - E_A)/kT}. This principle follows from the time-reversibility of microscopic dynamics, ensuring that forward and reverse paths are equally probable at . The in assumes that an explores all accessible microstates in with equal probability over sufficiently long times, supporting the in the . Under the equal a priori probability postulate of the , P_{eq}(A) is uniform across the energy surface, so P_{eq}(A) = P_{eq}(B) for states A and B at the same , and simplifies to w(A \to B) = w(B \to A). The ergodic theorem reinforces this by guaranteeing that, for time-reversible microscopic dynamics, long-time averages of observables equal ensemble averages over the . This equivalence relies on the dynamics satisfying to preserve the . Unlike balance, which merely requires the total influx to a to equal its outflux (i.e., \sum_C w(C \to A) P_{eq}(C) = P_{eq}(A) \sum_C w(A \to C)), microscopic reversibility demands for each individual pair of transitions, excluding cyclic or collective compensations. This stricter condition arises directly from the symmetry of reversible dynamics in at .

Macroscopic Implications

Detailed Balance Condition

The detailed balance condition, a key macroscopic consequence of microscopic reversibility, asserts that at , the forward and reverse transition fluxes between any pair of microstates balance exactly. Mathematically, for discrete states i and j, this is expressed as P_i w_{ij} = P_j w_{ji}, where P_k denotes the probability of state k and w_{kl} is the transition rate from state k to state l. This pairwise equality ensures no net exists between states, reflecting the time-reversibility of underlying dynamics at . Ludwig Boltzmann first formalized a version of this condition in his 1872 analysis of gas molecular collisions, deriving it within the to prove the H-theorem. By assuming for pairwise collisions—such that the collision rate from velocity distribution f(v_1) f(v_2) to f(v_1') f(v_2') equals the reverse—he showed that the equilibrium solution is the Maxwell-Boltzmann distribution, f(v) \propto \exp(-mv^2 / 2kT). For discrete-state systems, Boltzmann's approach extends analogously: assuming transition rates obey w_{ij}/w_{ji} = \exp(-(E_j - E_i)/kT), where E_k is the energy of state k, yields the canonical distribution P_i \propto \exp(-E_i / kT), directly linking microscopic rates to probabilities. In the framework of the master equation, which governs the time evolution of probabilities in Markovian systems, detailed balance provides a sufficient condition for the steady-state solution. The master equation takes the form \frac{dP_i}{dt} = \sum_{j \neq i} \left( P_j w_{ji} - P_i w_{ij} \right), where the sum runs over all states j. At steady state, dP_i/dt = 0 for all i, implying global balance: the total incoming flux equals the total outgoing flux for each state. Under detailed balance, this strengthens to pairwise cancellation, P_j w_{ji} - P_i w_{ij} = 0 for every pair, guaranteeing the equilibrium distribution without net currents anywhere in the state space. A prominent application appears in models and spin systems, such as the , where simulations rely on to sample equilibrium configurations. The algorithm proposes state changes with acceptance probability \min(1, \exp(-(E_j - E_i)/kT)), ensuring P_i w_{ij} = P_j w_{ji} and thus convergence to the ; this design traces directly to the 1953 original formulation for equation-of-state calculations. implies the system reaches equilibrium, as it satisfies the steady-state term-by-term, but the converse does not hold: global balance alone is weaker and permits steady states with circulating currents, such as in closed cycles where pairwise fluxes cancel collectively but not individually (e.g., a three-state ring with unequal forward/reverse rates yet zero net flux per state). This distinction underscores detailed balance's role in enforcing true equilibrium under time-reversible dynamics.

Onsager Reciprocal Relations

The Onsager reciprocal relations form a cornerstone of linear irreversible thermodynamics, establishing symmetries among the phenomenological coefficients that describe coupled processes in systems close to . These relations arise directly from of microscopic reversibility, which ensures that the underlying microscopic are under time reversal. In phenomenological theory, the fluxes J_i (such as heat or particle currents) are linearly related to the thermodynamic forces X_k (such as temperature or gradients) via the J_i = \sum_k L_{ik} X_k, where L_{ik} are the transport coefficients. Onsager's 1931 theorem asserts that these coefficients satisfy L_{ij} = L_{ji}, implying a symmetry in the response matrix that reflects the reciprocity of cause and effect in near-equilibrium transport. The derivation of this reciprocity stems from the time-reversal invariance of the system's Hamiltonian, which governs the microscopic equations of motion, combined with the fluctuation-dissipation theorem. This theorem links the linear response of the system to external perturbations with the equilibrium fluctuations of the fluxes, ensuring that the correlation functions underlying the transport coefficients exhibit the required symmetry under time reversal. Microscopic reversibility guarantees that the probability of a fluctuation and its reverse path are equal in the equilibrium ensemble, leading to the macroscopic reciprocity when averaged over many particles. In the equilibrium limit, this connects to the detailed balance condition, where forward and reverse transition rates balance without net fluxes. These relations manifest in specific transport phenomena. For instance, thermal conductivity is captured by the diagonal coefficient L_{qq}, relating heat flux to temperature gradient, while cross terms like L_{q\mu} couple heat flow to diffusion driven by chemical potential gradients, as in the Soret effect. In thermoelectric effects, the reciprocity equates the Seebeck coefficient \alpha (which generates voltage from a temperature difference) and the Peltier coefficient \pi (which produces heat absorption at junctions under electric current) via \pi = \alpha T, where T is temperature; this follows from the symmetry L_{qe} = L_{eq} between heat and electric fluxes. When external are present, which break time-reversal , the relations generalize to L_{ij}(\mathbf{B}) = L_{ji}(-\mathbf{B}), accounting for the odd behavior of magnetic effects under reversal; this extension preserves reciprocity while incorporating phenomena like the Hall effect.80108-4) The Onsager coefficients can also be expressed through the Green-Kubo formulas, which integrate time-correlation functions of the fluxes: for example, L_{ij} = \frac{1}{T} \int_0^\infty \langle J_i(t) J_j(0) \rangle \, dt, where the angular brackets denote the equilibrium ensemble average. The reciprocity L_{ij} = L_{ji} emerges naturally from these formulas because time-reversal invariance implies \langle J_i(t) J_j(0) \rangle = \epsilon_i \epsilon_j \langle J_j(t) J_i(0) \rangle, with \epsilon = \pm 1 depending on whether the flux is even or odd under reversal (e.g., currents are odd, while entropy fluxes are even). This microscopic origin underscores the deep connection between equilibrium fluctuations and nonequilibrium transport.

Applications and Extensions

In Chemical Reaction Mechanisms

In chemical reaction mechanisms, the principle of microscopic reversibility dictates that for any elementary step, the forward and reverse processes must follow pathways, ensuring that the sequence of states and intermediates is precisely mirrored in the opposite direction. This requirement stems from the time-reversibility of underlying at , preventing discrepancies that would violate thermodynamic consistency. Consequently, the energies for forward and reverse reactions are related through the standard change of the reaction, given by \Delta G^\circ = \Delta G^\ddagger_f - \Delta G^\ddagger_r = -[RT](/page/RT) \ln K_{eq}, where K_{eq} is the , R is the , and T is the temperature. A representative example is the reversible dissociation \ce{A ⇌ B + C}, where at equilibrium, the forward rate k_f [\ce{A}] equals the reverse rate k_r [\ce{B}][\ce{C}], leading to K_{eq} = k_f / k_r. This equality holds because the activation barrier for recombination of B and C mirrors that for the initial bond breaking in A, adjusted by the overall \Delta G^\circ. For multi-step mechanisms, microscopic reversibility implies that proposed pathways are invalid if the reverse route cannot retrace the of intermediates; for instance, in catalytic cycles, each step must allow precise reversal to avoid energy dissipation inconsistencies. In , this principle validates mechanisms by ensuring that substrate binding and product release steps are thermodynamically coupled, with rate constants satisfying for state probabilities at . Similarly, in reactions, it distinguishes SN1 and SN2 pathways: SN1 mechanisms, involving intermediates, require symmetric ion-pair collapse and formation for reversibility, while SN2 demands a collinear that inverts identically in reverse. Thermodynamic consistency is maintained as constants derive directly from the ratios of forward to reverse constants for each elementary step, K_{eq} = k_f / k_r, enforcing overall balance in complex networks. Historically, this has been instrumental in validating mechanisms by eliminating proposals where forward and reverse paths diverge, as seen in early studies of and reactions during the mid-20th century.

In Non-Equilibrium and Stochastic Systems

In non-equilibrium steady states, microscopic reversibility manifests through the principle of local (LDB), which applies to open systems coupled to multiple reservoirs. LDB posits that for any between mesoscopic states, the logarithm of the ratio of forward to reverse rates equals the associated flux to the environment, ensuring that microscopic dynamics remain time-reversible at the local level despite global irreversibility. This condition allows the total to decompose into a boundary term—arising from exchanges of energy and particles with external reservoirs—and an internal term reflecting changes in the system's configurational , with the boundary term dominating in steady states to satisfy on average. Fluctuation theorems extend this reversibility to stochastic trajectories in driven systems, quantifying the symmetry between forward and reverse paths. The , for instance, states that the ratio of probabilities for a forward trajectory and its time-reversed counterpart satisfies \frac{P_F(\Omega)}{P_R(-\Omega)} = e^{\Omega}, where \Omega is the dimensionless total along the path, directly linking microscopic reversibility to the exponential bias in trajectory likelihoods. This relation holds for arbitrary non-equilibrium processes, including those far from , and underpins second-law inequalities for fluctuating systems by showing that negative entropy production, while possible, is exponentially suppressed. In applications to molecular motors and Brownian ratchets, microscopic reversibility ensures that directed motion in driven, fluctuating environments adheres to thermodynamic consistency. For chemically powered motors like , LDB constrains the coupling between cycles and mechanical steps, allowing ratchet-like mechanisms to rectify thermal noise into net displacement without violating time-reversibility at the single-molecule level; the probability of forward versus backward steps reflects the free-energy input from the fuel. Similarly, in Brownian ratchets, external driving modulates potential landscapes to favor one direction, but reversibility dictates that the steady-state currents and dissipation align with cycle affinities, preventing machines. These systems illustrate how reversibility enables efficiency bounds, such as the Carnot limit in isothermal conditions, for nanomachines operating in viscous media. Schnakenberg network theory provides a graph-theoretic framework to analyze steady-state currents in complex reaction networks, incorporating microscopic reversibility via cycle affinities. In this approach, the network of states and transitions is decomposed into fundamental cycles, with each cycle's affinity defined as the log-ratio of forward to reverse path probabilities, quantifying the thermodynamic driving force; at , the net is the sum of current-affinity products over these cycles. This decomposition reveals how reversibility enforces zero-affinity cycles in while allowing non-zero affinities in non-equilibrium steady states, applicable to biochemical networks where multiple drives sustain fluxes. Modern interpretations formalize these ideas using integrals over ensembles of microscopic trajectories, emphasizing reversibility in probability measures. The action functional for probabilities incorporates the time-reversal operator, such that the measure for a equals that of its reverse under conditions, but in non-equilibrium, it weights paths by their ; this path-integral formulation unifies fluctuation theorems and LDB by integrating over all reversible microscopic dynamics to yield macroscopic irreversibility. Such approaches have been pivotal in deriving universal bounds on in engines.

Limitations and Exceptions

Influence of External Fields

External magnetic fields break the time-reversal symmetry () underlying microscopic reversibility because, under time reversal, particle velocities reverse while the magnetic field direction remains unchanged, requiring a reversal of the field \mathbf{B} \to -\mathbf{B} to restore invariance. This violation modifies the , transforming the standard symmetry L_{ij} = L_{ji} into L_{ij}(\mathbf{B}) = L_{ji}(-\mathbf{B}), where L_{ij} are the phenomenological coefficients relating thermodynamic fluxes and forces. A prominent example is the in electrical conductors, where the \mathbf{F} = q \mathbf{v} \times \mathbf{B} deflects charged carriers transversely to both current and field directions, producing a measurable Hall voltage. The resulting cycloidal trajectories of particles are non-reversible under time reversal without field inversion, as the force direction flips only when both velocity and \mathbf{B} are reversed, leading to opposite deflection. Coriolis forces in rotating reference frames act analogously to magnetic fields, introducing a velocity-dependent deflection that breaks and past-future symmetry in microscopic dynamics. This pseudo-magnetic effect modifies reciprocal relations similarly, with relations holding under reversal of the rate, and is relevant in geophysical and astrophysical contexts where rotation induces directional biases in particle motion. Ratchet mechanisms exploit such broken symmetries from external fields to enable directed motion of particles, such as in , without net equilibrium forces. In , an applied perpendicular breaks inversion in periodic structures, converting unbiased fluctuations or drives into net transport currents, as demonstrated in devices. Experimental verification of this asymmetry appears in motion studies, where charged particles orbit in the direction dictated by the , and the rotation sense reverses upon field inversion, confirming non-reversibility without \mathbf{B} \to -\mathbf{B}. Such observations, routine in particle accelerators and experiments, align with the predicted violation and underpin applications like Hall sensors.

Quantum and Relativistic Considerations

In , microscopic reversibility manifests through the CPT theorem, which asserts that any Lorentz-invariant is invariant under the combined operation of charge conjugation (C), (P), and time reversal (T). This theorem ensures overall time-reversal symmetry in closed systems, preserving the reversibility of microscopic dynamics at a fundamental level. However, subtle violations arise in specific interactions; for instance, observed in the decay of neutral implies a corresponding T violation via the CPT theorem, as demonstrated in the 1964 experiment where the long-lived kaon (K_L) decayed into two pions with a branching inconsistent with CP conservation. This T violation, while minute (on the order of 10^{-3}), represents a departure from strict microscopic reversibility in weak interactions. At equilibrium, quantum detailed balance extends the classical principle to density matrix formalism, requiring that for the equilibrium density operator ρ and any observable A, the commutator with the Hamiltonian H vanishes in trace: \tr\left(\rho [H, A]\right) = 0. This condition ensures that expectation values of observables are stationary, upholding time-reversal invariance in thermal equilibrium states. It aligns with the Kubo-Martin-Schwinger (KMS) condition, linking detailed balance to the canonical ensemble and preventing net fluxes in reversible quantum dynamics. Relativistic considerations further refine microscopic reversibility under , where time reversal acts differently on particles and s due to their charge conjugation properties, yet the symmetry holds for closed systems overall. In , the anti-unitary time-reversal operator T maps a particle state to its time-reversed counterpart without interchanging it with an , preserving Lorentz invariance and unitarity in isolated evolutions. However, topological effects can introduce apparent violations; the Aharonov-Bohm effect, for example, induces a phase shift in quantum interference patterns dependent on the enclosed via the , breaking time-reversal symmetry in the presence of non-zero flux and leading to asymmetric statistics. Modern extensions to open quantum systems adapt microscopic reversibility through quantum fluctuation relations, which generalize classical fluctuation-dissipation theorems to account for and . These relations, derived from microreversibility principles, equate the probabilities of forward and reversed trajectories in the presence of , with ratios involving work or , thus maintaining a form of reversibility despite environmental coupling. For instance, in driven , they quantify rare violations as exponential tails, bridging constraints to non-equilibrium .

References

  1. [1]
    5.18: The Principle of Microscopic Reversibility - Chemistry LibreTexts
    Jul 7, 2024 · The principle of microscopic reversibility states that any molecular process and its reverse occur with equal rates at equilibrium.
  2. [2]
    The Principle of Microscopic Reversibility - PNAS
    For migration in either direction the top of the range must be crossed, in analogy with the equal energies of the activated molecules for thetwo opposing ...
  3. [3]
    [PDF] Microscopic Reversibility Goes Quantum - PHYSICS - APS.org
    Oct 20, 2022 · Microscopic Reversibility Goes. Quantum. A fundamental principle in statistical mechanics called microscopic reversibility has been extended ...
  4. [4]
    Quantum coherence and the principle of microscopic reversibility
    The principle of microscopic reversibility is a fundamental element in the formulation of fluctuation relations and Onsager reciprocal relations.<|control11|><|separator|>
  5. [5]
  6. [6]
    Philosophy of Statistical Mechanics
    Jan 10, 2023 · Based on the Stosszahlansatz, Boltzmann derived what is now known as the Boltzmann Equation: ... Boltzmann prove with the H-theorem? Under which ...
  7. [7]
    [PDF] A quick introduction to kinetic theory. - Applied Mathematics
    hard-sphere gas is reversible in time. However, in the limit N → ∞, the effective dynamics modeled by the Boltzmann equation is irreversible. Clearly ...
  8. [8]
    Chapter 6: Ludwig Boltzmann. You Can't Put the Toothpaste Back in ...
    Microscopic collisions can be reversible. A faster molecule can transfer energy to a slower one when they collide or vice versa. Film the former, then run ...
  9. [9]
    Microscopic reversibility - Wikipedia
    Tolman. In chemistry, J. H. van't Hoff (1884) came up with the idea that equilibrium has dynamical nature and is a result of the balance between the ...Missing: Jacobus osmotic pressure
  10. [10]
    [PDF] Osmotic pressure and chemical equilibrium - Nobel Prize
    JACOBUS H. V A N'T HOFF. Osmotic pressure and chemical equilibrium. Nobel ... 1901 J.H.VAN'T HOFF remembers the vital role played by osmotic pressure in ...Missing: 1884 | Show results with:1884
  11. [11]
    A New Principle of Equilibrium - PNAS
    A New Principle of Equilibrium. Gilbert N. LewisAuthors Info & Affiliations. March 15, 1925. 11 (3) 179-183. https://doi.org/10.1073/pnas.11.3.179.
  12. [12]
    A Review of the Concept of Time Reversal and the Direction of Time
    Jun 30, 2024 · That is, time reversal reparametrizes the time coordinate, changes the sign of the p i (conjugate momenta), and leaves the q i unchanged ( ...
  13. [13]
    [PDF] Time Reversal - Bryan W. Roberts May 30, 2018 - PhilSci-Archive
    May 30, 2018 · Time reversal invariance in this context means that, given some force F, if x(t) is a solution to Newton's equation, then so is its time-reverse ...<|control11|><|separator|>
  14. [14]
    [PDF] Liouville's theorem and the foundation of classical mechanics - HAL
    Apr 13, 2022 · The theorem is interpreted to define the condition that describe the conservation of information in classical mechanics. The Hamilton equations ...
  15. [15]
    [PDF] Time Reversal Invariance in Quantum Mechanics by Reza Moulavi ...
    I will present and examine my own alternative account of time reversal at the end. Also based on the invariance of the Schrödinger equation under time reversal ...
  16. [16]
    [PDF] When we do (and do not) have a classical arrow of time - LSE
    In the Newtonian “force” formulation of classical mechanics, time reversal is simply the reversal of the order of events in a trajectory x(t). That is, if x(t) ...
  17. [17]
    microscopic reversibility at equilibrium (M03916) - IUPAC
    The principle of microscopic reversibility at equilibrium states that, in a system at equilibrium, any molecular process and the reverse of that process occur, ...
  18. [18]
    The Principle of Microscopic Reversibility - PNAS
    Optimism is a psychological attribute characterized as the general expectation that good things will happen, or the belief that the future will be favorable ...
  19. [19]
    Local detailed balance: a microscopic derivation - IOPscience
    Dec 8, 2014 · The ergodic hypothesis states that a generic trajectory of the system will asymptotically cover the energy hypersurface uniformly. To be ...
  20. [20]
    [PDF] Local Detailed Balance : A Microscopic Derivation - arXiv
    Dec 28, 2014 · The ergodic hypothesis states that a generic trajectory of the system will asymptotically cover the energy hypersurface uniformly. To be ...
  21. [21]
    [PDF] Local detailed balance - KU Leuven
    We review the physical meaning and mathematical implementation of the con- dition of local detailed balance for a class of nonequilibrium mesoscopic ...
  22. [22]
    [PDF] Local detailed balance Abstract Contents - SciPost
    Jul 6, 2021 · We review the physical meaning and mathematical implementation of the condition of local detailed balance for a class of nonequilibrium ...
  23. [23]
    [PDF] 2 Further Studies on the Thermal Equilibrium of Gas Molecules
    [This statement is now known as the H- theorem.] .Emust approach a minimum value and remain constant thereafter, and the corresponding final value of/will ...
  24. [24]
    [PDF] the expansion of
    VAN KAMPEN. This is the equation for the macroscopic part of n, that is, the macroscopic rate equation. Fourth step: the next order determines the fluctuations.
  25. [25]
    [PDF] arXiv:1407.6927v2 [cond-mat.stat-mech] 8 Sep 2014
    Sep 8, 2014 · Abstract. We develop a general framework for the discussion of detailed balance and analyse its microscopic background. We.
  26. [26]
    The Principle of Microscopic Reversibility
    If the mechanism in one direction is known, then the mechanism in the opposite direction is known. The lowest-energy pathway in the forward direction will be ...Missing: physics | Show results with:physics
  27. [27]
    Kinetic consequences of the principle of microscopic reversibility
    The principle of microscopic reversibility was formulated for elementary reactions at equilibrium. It follows from the principle that for any system at ...<|control11|><|separator|>
  28. [28]
    Generalized microscopic reversibility, kinetic co-operativity of ...
    Dec 1, 1978 · Generalized microscopic reversibility implies that the apparent rate of any catalytic process in a complex mechanism is paralleled by ...
  29. [29]
    [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.
  30. [30]
    Thermodynamics and Kinetics of Molecular Motors - PMC
    By use of microscopic reversibility, it is shown that the ratio between the number of forward steps and the number of backward steps in any sufficiently long ...
  31. [31]
    None
    Nothing is retrieved...<|separator|>
  32. [32]
    Cyclotron motion without magnetic field - IOPscience
    Aug 12, 2019 · Here we show that, in addition to a dc Hall effect, anomalous Hall materials possess circulating currents and cyclotron motion without magnetic field.Missing: asymmetry experimental
  33. [33]
    Time reversal symmetry for classical, non-relativistic quantum and ...
    Onsager stated that: “the principle of dynamical reversibility does not apply when (external) magnetic fields or Coriolis forces are present, and the reciprocal ...
  34. [34]
    Electron ratchets: State of the field and future challenges
    Electron ratchets are non-equilibrium electronic devices that break inversion symmetry to produce currents from non-directional and random perturbations.
  35. [35]
    [PDF] Electron Ratchets: State of the Field and Future Challenges
    May 22, 2020 · Symmetry breaking by magnetic fields. Ratcheting is predicated on breaking the inversion symmetry along the direction of transport. One way ...
  36. [36]
    [PDF] Notes on the CPT theorem 1. Introduction In relativistic quantum field ...
    Mar 2, 2021 · In relativistic quantum field theory, the CPT theorem is a collection of results showing that the dynamics are invariant under a transformation ...
  37. [37]
    Deciphering the Algebraic CPT Theorem - PhilSci-Archive
    Jun 20, 2019 · The CPT theorem states that any causal, Lorentz-invariant, thermodynamically well-behaved quantum field theory must also be invariant under a ...
  38. [38]
    Evidence for the Decay of the Meson | Phys. Rev. Lett.
    Nov 8, 2013 · A 1964 experiment on an unusual particle showed a violation of symmetry and led to the conclusion that matter and antimatter are not quite equivalent.
  39. [39]
    [PDF] Lecture 8 CPT theorem and CP violation - SMU Physics
    CPT theorem implies that the violation of CP invariance implies violation of time reversal symmetry T. • One of the most important unresolved questions today is ...
  40. [40]
    Physics Quantum Detailed Balance and KMS Condition
    In this paper, we give a definition of detailed balance for a quantum dynamical semigroup of a PF*- algebra, which extends the analogous notion proposed in [3] ...
  41. [41]
    Detailed balance as a quantum-group symmetry of Kraus operators
    Feb 14, 2018 · The density matrix ρ0 plays the role of asymptotic equilibrium state, and thermodynamical quantities can be defined in terms of ρ0. However, ...
  42. [42]
    [PDF] Time reversal in classical electromagnetism
    According to standard quantum field theory textbooks this is not so: time reversal does not turn particles into anti-particles. Feynman's view is ...
  43. [43]
    [PDF] Statistics of energy levels without time-reversal symmetry: Aharonov ...
    Abstract. A planar domain D contains a single line of magnetic flux a. Switching on @ breaks time-reversal symmetry (T) for quantal particles with charge q ...
  44. [44]
    [1106.1982] Microscopic reversibility of quantum open systems - arXiv
    Jun 10, 2011 · Microscopic reversibility equates the time forward and reversed probabilities, and therefore appears as a thermodynamic symmetry for open ...Missing: fluctuation relations
  45. [45]
    Quantum coherence and the principle of microscopic reversibility
    Nov 15, 2023 · The principle of microscopic reversibility is a fundamental element in the formulation of fluctuation relations and Onsager reciprocal relations.