Fact-checked by Grok 2 weeks ago

Transition state theory

Transition state theory (), also known as activated complex theory, is a foundational framework in that explains the rates of elementary chemical reactions through the concept of a transient, high-energy formed by the reactants as they approach the point of bond breaking and formation. This theory posits that the reaction proceeds via the at the of the , where the system is in quasi-equilibrium with the reactants, and the rate is governed by the frequency at which this complex decomposes into products along the . Developed independently in 1935 by Henry Eyring at and by Meredith Gwynne Evans and at Manchester University, TST built upon earlier ideas from , potential energy surfaces, and the to provide a statistical mechanical basis for absolute reaction rates. The core assumptions include the existence of an equilibrium between reactants and the (valid under the quasi-equilibrium approximation), treatment of the vibration as a translation across the barrier, and an initial of unity, meaning no recrossing of the barrier. The theory's principal achievement is the derivation of the Eyring equation for the rate constant of a bimolecular reaction:
k = \frac{k_B T}{h} K^\ddagger = \frac{k_B T}{h} e^{-\Delta G^\ddagger / RT},
where k_B is Boltzmann's constant, T is temperature, h is Planck's constant, K^\ddagger is the equilibrium constant between reactants and the transition state, and \Delta G^\ddagger is the standard Gibbs free energy of activation. This equation highlights how both enthalpic (energy barrier) and entropic (configurational) factors contribute to the activation process, enabling the interpretation of experimental rate data in terms of thermodynamic quantities.
TST has profoundly influenced by providing a predictive tool for reaction rates in diverse systems, including gas-phase collisions, solution-phase processes, and enzymatic , though refinements like variational TST and quantum mechanical corrections address limitations such as tunneling and barrier recrossing in modern applications. Its integration with computational methods, such as locating transition states on surfaces, has become essential for modeling complex reaction mechanisms in fields like and .

Fundamentals

Core Concepts

Transition state theory (TST) provides a foundational framework for understanding chemical reaction rates by focusing on the energetic pathway that reactants follow to form products. Central to this theory is the concept of the , which represents the path of minimum connecting the reactant and product states in a multidimensional . Along this coordinate, reactions encounter an energy barrier, depicted in potential energy diagrams as a peak separating the lower-energy minima of reactants and products; this barrier arises from the need to rearrange molecular bonds and geometries during the transformation. The height of this energy barrier is quantified by the , E_a, defined as the difference in between the reactants and the highest point along the reaction coordinate. This directly influences the rate constant of a reaction, as higher barriers result in fewer molecules possessing sufficient to surmount them, leading to slower rates; empirically, this relationship is captured in the precursor , k = A \exp(-E_a / RT), where A is a , R is the , and T is temperature. TST addresses limitations in earlier , which assumed reactions occur solely through direct, energetic molecular collisions but struggled to explain observed rate discrepancies without adjustments like a . Instead, TST posits that reactions proceed through a transient, high-energy —an at the saddle point of the —where bonds are partially formed and broken, rather than via simple collisions. This exists momentarily before decomposing into products, with the rate determined by the population of this state and its subsequent decomposition frequency. Conceptually, the theory yields a rate constant of the form k = \frac{k_B T}{[h](/page/H+)} \exp\left(-\frac{\Delta G^\ddagger}{RT}\right), where k_B is Boltzmann's constant, [h](/page/H+) is Planck's constant, and \Delta G^\ddagger is the of activation, emphasizing the thermodynamic control over kinetics.

Transition State Definition

In transition state theory, the , often referred to as the , is defined as the high-energy configuration of atoms located at the on the along the minimum energy path. This point represents the critical where the system achieves the maximum relative to the reactants, marking the apex of the barrier that must be surmounted for the to proceed to products. The serves as the multidimensional framework for identifying this state, separating reactant and product regions through the configuration. Key characteristics of the transition state include its extreme instability and fleeting existence, with a lifetime on the order of a few femtoseconds, corresponding to the timescale of vibrational motions and redistribution during reconfiguration. At this state, the molecular structure features partial breaking and forming, where reactant bonds are elongated and weakened while incipient product bonds begin to develop, resulting in a transient arrangement that cannot be isolated or observed directly. This instability arises from its position at an energy maximum along the , making it prone to immediate evolution toward either reactants or products without lingering. Computationally, transition states are identified as first-order saddle points through analysis of the , the second derivative of the with respect to coordinates, which yields one negative eigenvalue indicative of negative along the reaction mode. This manifests as a single imaginary vibrational frequency in analysis, confirming the structure's role as the barrier top, whereas all other frequencies are real and positive, akin to those of stable species. Such criteria ensure the optimized geometry is not a minimum but the precise dividing point on the energy landscape. A crucial distinction exists between the transition state and reaction intermediates: while intermediates correspond to true local minima on the with all positive eigenvalues and thus some degree of , the is a maximum along the (though a minimum in perpendicular directions), rendering it inherently unstable and non-isolable. This differentiates the as a purely theoretical construct at the crest, in contrast to intermediates that may accumulate and be detectable under certain conditions.

Historical Development

Thermodynamic Formulations

The thermodynamic formulations of transition state theory emerged in the late , building on the foundational work of J. Willard Gibbs, who established the principles of chemical potentials and constants in heterogeneous systems, providing a basis for treating reaction intermediates thermodynamically. These ideas were extended in the early by researchers such as P. Kohnstamm and F. E. C. Scheffer, who in 1911 proposed that reaction rates could be linked to an exponential dependence on the change, rather than the potential itself, introducing a thermodynamic perspective on activation barriers. Jens Anton Christiansen further advanced this approach in the 1920s and 1930s, particularly through his 1924 analysis of reaction rates in ionic solutions, where he considered intermediates in with reactants, allowing the application of thermodynamics to rate processes. A key aspect of these thermodynamic treatments is the conceptualization of the —also termed the —as a distinct thermodynamic in quasi- with the reactants, despite its fleeting existence. This equilibrium assumption enables the definition of an K^\ddagger for the formation of the transition state from reactants, such that the free energy of activation is given by \Delta G^\ddagger = -RT \ln K^\ddagger, where R is the and T is the absolute temperature. This formulation treats the transition state analogously to a , facilitating the use of standard thermodynamic relations. The free energy of activation \Delta G^\ddagger relates to the of activation \Delta H^\ddagger and \Delta S^\ddagger via \Delta G^\ddagger = \Delta H^\ddagger - T \Delta S^\ddagger, providing insights into the energetic and entropic contributions to the reaction barrier. Early experimental validations of these thermodynamic ideas relied on analyzing temperature-dependent reaction rates, often through Arrhenius plots, to extract activation parameters that aligned with the predicted \Delta H^\ddagger and \Delta S^\ddagger. For instance, in the 1930s, M. G. Evans and M. Polanyi applied the thermodynamic framework to gas-phase reactions, demonstrating consistency between observed rate constants and free energy changes derived from potential energy surfaces. These validations confirmed the utility of the quasi-equilibrium approximation for many systems, though later statistical mechanical refinements would address limitations in dynamic aspects.

Kinetic and Statistical Mechanical Treatments

The kinetic treatment of transition state theory (TST) evolved from thermodynamic formulations by incorporating to describe reaction rates at the molecular level, assuming the transition state is in quasi-equilibrium with reactants. This approach bridged macroscopic observables to microscopic dynamics, enabling predictions of absolute reaction rates without empirical fitting. In the early 1930s, and collaborators advanced in TST by applying quantum principles to landscapes, emphasizing the role of vibrational and electronic states in rate-determining steps. Polanyi's work with Eyring in 1931 laid groundwork for quantum corrections to classical rate expressions, highlighting how quantum effects influence the stability and passage through the . Simultaneously, contributed pivotal quantum mechanical treatments, particularly in 1932 with Pelzer, where they used the time-dependent to derive rate constants for association reactions, incorporating quantum tunneling and barrier penetration probabilities. Wigner's 1937 analysis further refined these ideas, providing a quantum statistical framework for elementary reactions that accounted for distributions near the . The seminal 1935 contributions by Henry Eyring, and independently by Meredith Gwynne Evans and , formalized the statistical mechanical basis of TST through the theory of absolute rates. Eyring's formulation treated the as a loosely bound complex, with the rate constant derived from the flux of activated complexes across the dividing surface. Evans and Polanyi's parallel work emphasized applications to bimolecular s, integrating statistical weights for molecular orientations. This era marked a shift toward partition function-based expressions, where the rate constant k for a is given by k = \frac{k_B T}{h} \frac{Q^\ddagger}{\prod Q_i} \exp\left(-\frac{\Delta E_0}{k_B T}\right), with Q^\ddagger as the partition function of the transition state and Q_i those of the reactants; these partition functions factor into translational (Q_t = (2\pi m k_B T / h^2)^{3/2} V), rotational (Q_r = (8\pi^2 I k_B T / \sigma h^2) for linear molecules), and vibrational (Q_v = \prod 1 / (1 - \exp(-h\nu_j / k_B T))) components, adjusted for the loss of one vibrational mode at the transition state. Early statistical models introduced transmission coefficients to correct for trajectories that recross the dividing surface, rather than proceeding to products. Eyring initially set this coefficient \kappa to unity for classical cases but recognized its role in accounting for inefficiencies, such as in tunneling or steric hindrances; Wigner's quantum treatments estimated \kappa via semiclassical approximations, often yielding values near 1 for barrierless paths but lower for narrow barriers. This coefficient became essential for refining TST predictions in diverse media. The progression from classical to quantum statistics in TST involved replacing Boltzmann distributions with quantum partition functions, particularly for low-temperature regimes where vibrational quantization dominates. Classical treatments, prevalent pre-1930, assumed continuous levels, but Wigner and Polanyi's integrations of ensured accurate handling of molecular , such as anharmonic vibrations and zero-point energies, enhancing the theory's applicability to gas-phase reactions.

Potential Energy Surfaces

In transition state theory, the (PES) provides a multidimensional representing the E(\mathbf{R}) of a molecular system as a function of the coordinates \mathbf{R}, derived from solving the time-independent within the Born-Oppenheimer approximation. The full \hat{H} = \hat{T}_N + \hat{H}_{el}, where \hat{T}_N is the and \hat{H}_{el} is the , yields the PES as the V(\mathbf{R}) = E(\mathbf{R}) for , assuming fixed nuclei during calculations. This separation enables the of landscapes governing chemical reactions, with valleys corresponding to stable minima (reactants or products) and barriers indicating activation energies. Key features of the PES include the minimum energy path (MEP), which traces the lowest-energy trajectory connecting reactant and product minima via steepest descent in mass-weighted coordinates from a saddle point; this MEP often serves as the reaction path in simplified models. The , a where the energy gradient vanishes (\nabla V = 0) and the has exactly one negative eigenvalue, marks the : it acts as a maximum along the but a minimum in all orthogonal directions, forming a col or bottleneck for reactive flux. In collinear configurations, such as the H + H₂ exchange reaction, the PES reduces to a two-dimensional surface (using two independent internuclear distances), revealing a clear barrier height of approximately 0.42 on early semi-empirical surfaces like the London-Eyring-Polanyi (LEP) form. For multidimensional cases, the full three-atom H + H₂ PES (six-dimensional, reduced by symmetry and center-of-mass constraints) exhibits vibrational adiabatic effects and corner-cutting trajectories that lower effective barriers compared to collinear paths. Prior to the 1980s, PES construction relied on ab initio quantum chemical methods, primarily self-consistent field (SCF) Hartree-Fock theory supplemented by configuration interaction (CI) for electron correlation, applied to small systems due to computational limitations. Seminal calculations include Liu's 1973 collinear H₃ surface using SCF with basis sets up to double-zeta quality, and Siegbahn and Liu's 1978 three-dimensional H₃ PES via multi-reference CI, achieving chemical accuracy (errors < 1 kcal/mol) for the first time on a reactive surface. Density functional theory (DFT), though formulated in the 1930s, saw limited practical use for PES before the 1980s, with early applications mostly to atomic systems rather than molecular hypersurfaces. The PES validates transition state theory by geometrically defining a dividing surface—typically a hyperplane perpendicular to the reaction path at the saddle point—that separates reactant and product basins, with the reactive flux calculated as the one-way transmission through this surface assuming no classical recrossings. This structure confirms the theory's core assumption of a rate-determining bottleneck, as trajectories must surmount the saddle to proceed reactively. In statistical mechanical treatments, the PES underpins the computation of vibrational and rotational partition functions for species along the reaction path.

Central Equations

Eyring Equation Derivation

The derivation of the Eyring equation begins with the assumption of a quasi-equilibrium between the reactants and the transition state species, often denoted as the activated complex [M‡]. For a bimolecular reaction A + B → products, the equilibrium constant K^\ddagger relates the concentration of the transition state to the reactants: [M^\ddagger] = K^\ddagger [A][B], where K^\ddagger = \exp(-\Delta G^\ddagger / RT). This expression arises from thermodynamic considerations, with \Delta G^\ddagger being the standard Gibbs free energy of activation, R the gas constant, and T the temperature. In statistical mechanical terms, K^\ddagger is expressed using partition functions: K^\ddagger = \frac{Q^\ddagger}{Q_A Q_B} \exp(-\Delta E_0 / RT), where Q^\ddagger, Q_A, and Q_B are the molecular partition functions for the transition state and reactants, respectively, and \Delta E_0 is the difference in zero-point energies between the transition state and reactants. The partition function for the transition state excludes the degree of freedom along the , treating it as a separable translational mode perpendicular to the dividing surface at the saddle point. This formulation assumes a classical limit for vibrational modes except for the imaginary frequency along the reaction coordinate, which is handled by integrating over velocities crossing the barrier. The reaction rate is then given by the unidirectional flux of activated complexes through the transition state to form products: rate = \nu [M^\ddagger], where \nu is the frequency at which complexes cross the barrier. Analogous to a one-dimensional translational motion, \nu is derived from the mean thermal velocity component normal to the dividing surface. Using the , the average kinetic energy along this coordinate is \frac{1}{2} k_B T, leading to \nu = \frac{k_B T}{h}, with k_B and h . This frequency originates from the and the vibrational analogy, where the imaginary frequency mode \nu^\ddagger = i |\nu| / 2\pi yields the same factor upon integration over the barrier position (fixed at the saddle point) and positive velocities. Combining these, the rate constant k for the forward reaction is k = \frac{k_B T}{h} K^\ddagger = \frac{k_B T}{h} \exp(-\Delta G^\ddagger / RT). Incorporating the partition functions gives the equivalent form k = \frac{k_B T}{h} \frac{Q^\ddagger}{Q_A Q_B} \exp(-\Delta E_0 / RT). This assumes no recrossing of the dividing surface (transmission coefficient \kappa = 1) and neglects quantum tunneling effects. For unimolecular reactions, the form simplifies similarly, with appropriate stoichiometric adjustments. Regarding units, the equation yields k in concentration units consistent with the reaction order (e.g., M^{-1} s^{-1} for bimolecular), as the partition functions are dimensionless when referenced to a standard state (typically 1 M). The temperature dependence emerges from the prefactor k_B T / h, which increases linearly with T, and the exponential term, modulated by the enthalpic and entropic contributions to \Delta G^\ddagger = \Delta H^\ddagger - T \Delta S^\ddagger. This results in an overall weak temperature dependence in the prefactor compared to the activation barrier term.

Justification of Assumptions

The quasi-equilibrium assumption in (TST) posits that the transition state is in rapid equilibrium with the reactants, despite the overall reaction being far from equilibrium, because the lifetime of the transition state is extremely short—on the order of vibrational periods (approximately $10^{-13} seconds)—allowing statistical mechanical treatment via . This assumption is justified conceptually by the fact that, at equilibrium, no species can be depleted, ensuring the preservation of as long as the critical activation energy exceeds about 5kT, where k is the and T is temperature, making the reactive population sparse and the equilibrium perturbation negligible. Early validations, such as those by in 1932, supported this by showing that the transition state acts as a dividing surface where forward flux dominates without significant backward motion, aligning with the no-recrossing condition inherent to the quasi-equilibrium. A key plausibility argument for the quasi-equilibrium draws from analogies in unimolecular dissociation rates, as described by Rice-Ramsperger-Kassel-Marcus (RRKM) theory, which assumes rapid intramolecular vibrational energy redistribution within energized molecules, establishing a statistical (quasi-equilibrium) distribution prior to crossing the transition state barrier—mirroring TST's treatment of the activated complex as being in equilibrium with reactants before decomposition. This parallel reinforces the validity of TST's statistical mechanics foundation for barrier-crossing events, particularly in isolated systems where energy randomization is fast compared to reaction timescales. The transmission coefficient \kappa, introduced to account for any deviation from perfect no-recrossing (i.e., trajectories that cross the transition state but return to reactants), is approximated as \kappa \approx 1 for most elementary reactions, implying efficient forward passage. Early estimates by in 1935 suggested \kappa values near unity for simple gas-phase reactions, based on classical trajectory considerations along the reaction coordinate, with minor adjustments (e.g., \kappa \approx 0.5 for cases with loose transition states) derived from potential energy surface analyses. Experimental support for these assumptions comes from kinetic isotope effects (KIEs), where the observed primary KIEs for hydrogen/deuterium substitutions in reactions like H abstraction match TST predictions of differences in zero-point energies between ground and transition states, providing a sensitive probe that validates the equilibrium treatment without invoking recrossing. Temperature-dependence studies further corroborate this, as the linear Arrhenius plots over wide ranges (e.g., for bimolecular gas-phase reactions) align with the quasi-equilibrium-derived , indicating consistent barrier heights and minimal dynamic corrections. The 1935 formulation by Eyring and by Evans and Polanyi provided conceptual validation through semi-empirical potential energy surface constructions for reactions like H + HBr, demonstrating that the quasi-equilibrium and no-recrossing assumptions yield rate constants in quantitative agreement with experimental data, establishing TST as a predictive framework.

Relation to Other Theories

Connection to Arrhenius Equation

The Arrhenius equation empirically describes the temperature dependence of reaction rate constants as
k = A \exp\left( -\frac{E_a}{RT} \right),
where A is the pre-exponential factor, E_a is the activation energy, R is the gas constant, and T is the absolute temperature. This form, introduced in 1889, provided a practical tool for analyzing reaction kinetics but lacked a molecular interpretation until the development of transition state theory (TST).
Transition state theory, through the Eyring equation, establishes a direct theoretical connection to the by interpreting E_a and A in terms of thermodynamic properties of the transition state. Specifically, the activation energy relates to the enthalpy of activation as E_a = \Delta H^\ddagger + RT, reflecting the energy barrier height adjusted for the work of forming the activated complex. The pre-exponential factor is given by A \approx \frac{e k_B T}{h} \exp\left( \frac{\Delta S^\ddagger}{R} \right), where k_B is , h is , and \Delta S^\ddagger is the entropy of activation; this links A to the entropy change associated with reaching the transition state, providing a statistical mechanical basis for its magnitude, typically around $10^{13} s^{-1} for bimolecular reactions. These relations emerge from assuming quasi-equilibrium between reactants and the transition state, with the rate determined by the flux over the barrier. In the high-temperature limit, where quantum effects are negligible and the system behaves classically, TST quantitatively reduces to the , validating the empirical form as an approximation of the more fundamental TST rate expression. This equivalence highlights TST's role in mechanistically grounding the observed exponential temperature dependence. However, at low temperatures, quantum mechanical effects such as vibrational zero-point energy and tunneling introduce deviations, predicting upward curvature in Arrhenius plots (ln k vs. 1/T) due to enhanced rates beyond classical expectations. The post-1930s formulation of TST marked a pivotal historical shift from the purely empirical to a mechanistic understanding rooted in potential energy surfaces and statistical mechanics, enabling predictions of rate constants from molecular properties rather than fitting experimental data alone. This advancement, spearheaded by Eyring's work in 1935, transformed chemical kinetics into a predictive science.

Inferences and Predictions

Transition state theory (TST) predicts that variations in activation parameters often exhibit enthalpy-entropy compensation, where changes in the activation enthalpy (ΔH‡) are linearly correlated with changes in the activation entropy (ΔS‡), such that the free energy of activation (ΔG‡ = ΔH‡ - TΔS‡) remains relatively constant across related reaction series. This compensation arises from the quasi-equilibrium assumption between reactants and the transition state, where environmental factors like solvent reorganization or structural variations in reactants influence both enthalpic and entropic contributions in a coupled manner. For instance, in series of homologous reactions, such as alkyl halide hydrolyses, increasing chain length leads to parallel increases in ΔH‡ and TΔS‡, maintaining similar rate constants as observed in experimental studies. TST further infers primary kinetic isotope effects (KIEs) from differences in zero-point energies (ZPE) between isotopic variants in the transition state relative to reactants. In reactions involving bond breaking or formation at the isotopically substituted site, differences in zero-point energies (ZPE) lead to primary kinetic isotope effects (KIEs), where the lighter isotope has a higher ZPE in reactants and a greater decrease upon forming the looser transition state, lowering its activation barrier relative to the heavier isotope and yielding k_H/k_D values typically between 2 and 7 for deuterium substitutions. This prediction stems from the vibrational analysis within TST, where the force constants along the reaction coordinate diminish at the transition state, amplifying ZPE differences and leading to observable rate ratios that validate the theory's assumptions about potential energy surfaces. Pressure effects on reaction rates are encapsulated in TST through the activation volume (ΔV‡), defined as the difference in partial molar volumes between the transition state and reactants, which influences the rate constant via the relation ∂ln k / ∂P = -ΔV‡ / RT. Negative ΔV‡ values, common in associative reactions, indicate volume contraction or electrostriction at the transition state and predict rate increases with increasing pressure, while positive values for dissociative processes indicate volume expansion and suggest rate decreases. Similarly, solvent effects modulate the activation free energy (ΔG‡) by differentially solvating the transition state compared to reactants, often stabilizing charge-separated transition states in polar media and enhancing rates for . These predictions align with observations in protic solvents, where hydrogen bonding lowers ΔG‡ more for the transition state than for neutral reactants. TST also underpins linear free energy relationships, exemplified by the Hammond postulate, which posits that the transition state structure resembles the nearer stable species in energy along the reaction coordinate. For exothermic reactions, the early transition state mirrors reactants, while endothermic ones feature late, product-like transition states, correlating substituent effects on rates with equilibrium constants. This qualitative inference facilitates quantitative structure-reactivity trends, such as Bronsted correlations, by linking variations in ΔG‡ to electronic perturbations at the transition state. Experimental data from the 1940s to 1960s provided key validations of these TST inferences, including temperature-dependent pre-exponential factors matching predicted entropic contributions in gas-phase reactions and solvent-dependent rates for and processes aligning with ΔG‡ shifts. Isotope effect measurements on hydrogen transfers confirmed ZPE-based KIE predictions, while high-pressure studies on ester hydrolyses demonstrated ΔV‡ values consistent with transition state volume changes. These empirical tests, often reducing to forms under simplifying assumptions, affirmed TST's predictive power for diverse kinetic phenomena.

Extensions via Kramers Theory

Brownian Motion Model

In 1940, Hendrik A. Kramers developed a model that extended (TST) to describe reaction rates in solution by treating chemical reactions as the diffusive escape of a from a potential well over an energy barrier. This approach addressed the limitations of gas-phase TST in viscous media, where solvent interactions significantly influence dynamics. Kramers' framework posits that the reacting system can be analogous to a particle subject to random thermal fluctuations from the surrounding solvent, enabling it to surmount the barrier despite the stabilizing potential. Central to the model is the analogy of an overdamped Brownian particle attempting to escape a potential well, where the particle's motion is governed by a balance of deterministic forces from the potential and stochastic forces from the solvent. The key assumption is that frictional damping from the solvent dominates over inertial effects, rendering the particle's velocity quickly relaxed to a local equilibrium distribution, such as the . This overdamped regime simplifies the dynamics to a diffusion process along the reaction coordinate, contrasting with the ballistic motion in low-friction environments. In classical , which serves as the low-friction limit of Kramers' theory, the rate is determined by equilibrium populations at the transition state without diffusive corrections. Kramers' model marks a transition from the gas-phase TST, where reactions proceed via direct collisions with minimal solvent interference, to the diffusive regime prevalent in liquids, where high viscosity leads to prolonged residence times near the barrier. This shift emphasizes the role of solvent friction in modulating barrier crossing, particularly for reactions in condensed phases. The probability of barrier crossing in this framework relates directly to the , which describes the diffusion of the particle's position in the presence of the potential field, yielding a steady-state flux over the barrier under quasi-equilibrium conditions.

Rate Expressions for Viscous Media

In the high-friction limit of Kramers theory, applicable to reactions in viscous media where the friction coefficient \gamma greatly exceeds the characteristic frequencies of the potential energy surface, the reaction rate is significantly reduced compared to the transition state theory prediction due to the diffusive character of barrier crossing. The rate constant takes the form k = \frac{\omega_R |\omega^\ddagger|}{2\pi \gamma} \exp\left( -\frac{\Delta E}{RT} \right), where \omega_R is the angular frequency associated with the curvature at the bottom of the reactant well, \omega^\ddagger is the imaginary angular frequency at the (reflecting the inverted parabolic barrier), \Delta E is the activation energy barrier height, R is the gas constant, and T is the temperature.90098-2) This expression modifies the transition state theory pre-factor k_B T / h by replacing it with a term inversely proportional to \gamma, emphasizing the role of solvent drag in limiting the successful transmission over the barrier.90098-2) The friction coefficient \gamma connects directly to the macroscopic solvent viscosity \eta through the Stokes-Einstein relation, \gamma = 6\pi \eta r for translational motion of a spherical solute of radius r, thereby making the rate inversely dependent on \eta in viscous solutions. For rotational barriers, such as in molecular isomerizations, the relevant rotational friction is \gamma_\text{rot} = 8\pi \eta r^3, leading to a cubic dependence on solute size and further highlighting viscosity's influence on reorientation rates in solution.90098-2) In the low-friction regime, characteristic of less viscous environments where energy exchange with the bath dominates over momentum dissipation, the system enters an energy diffusion limit, and the rate expression shifts to k \approx \frac{E_a}{2\pi \rho k_B T} \frac{|\omega^\ddagger|}{\omega_R} \exp\left( -\frac{\Delta E}{RT} \right), with E_a denoting the activation energy (approximately \Delta E), \rho the moment of inertia (for rotational degrees of freedom) or an effective reduced mass parameter (for translational), and k_B Boltzmann's constant.90098-2) Here, the rate increases linearly with \gamma (and thus weakly with \eta) because minimal friction is needed to facilitate energy redistribution without excessive recrossing.90098-2) The transition between these limits exhibits crossover behavior, with the full Kramers solution involving solutions to the Klein-Kramers equation that interpolate via modified Bessel functions, yielding numerical factors of order unity and a maximum rate (Kramers turnover) at intermediate \gamma \approx 2 |\omega^\ddagger|.90098-2) This turnover underscores the non-monotonic viscosity dependence, optimizing rates in moderately viscous media. Applications include modeling translational escape in solvated ions and rotational barriers in viscous organic solvents, where the high-friction form accurately predicts rates for barrier curvatures \omega_R, |\omega^\ddagger| \sim 10^{12}--$10^{13} s^{-1}.90098-2)

Limitations

Quasi-Equilibrium Breakdown

In transition state theory (TST), the quasi-equilibrium assumption posits that the reactants and the transition state complex maintain an equilibrium distribution, enabling the reaction rate to be formulated in terms of the equilibrium constant for transition state formation despite the overall system being far from equilibrium. This core assumption breaks down in certain scenarios, particularly in condensed phases, where dynamical processes prevent the rapid establishment or maintenance of this equilibrium. For instance, in solution reactions involving fast solvent relaxation, such as excited-state proton transfer in substituted 3-hydroxyflavones in polar solvents like methanol, the reorientation of solvent dipoles occurs on a timescale slower than the reaction dynamics near the transition state, leading to a non-equilibrium solvation environment that disrupts the assumed Boltzmann distribution of the transition state population. Similarly, reactions with high energy barriers can fail to maintain equilibrium if the low density of states near the transition state hinders sufficient collisional or bath-mediated redistribution before trajectories depart from the dividing surface. Early critiques of the quasi-equilibrium validity emerged in the 1950s, with N. B. Slater challenging the statistical foundations of TST in his dynamical theory of unimolecular reactions, arguing that the assumption of ergodic behavior and equipartition in the transition state complex is untenable due to correlated molecular motions that prevent full intramolecular energy randomization. Slater's work highlighted that classical TST over-relies on equilibrium statistical mechanics, which may not capture the non-statistical flux through the transition state in isolated or low-collision environments, prompting a shift toward more dynamical treatments. These concerns extended to condensed phases, where intermolecular interactions further complicate the equilibration process. The breakdown of quasi-equilibrium leads to deviations from TST predictions, often resulting in underestimation of reaction rates in condensed phases, as non-equilibrium distributions can facilitate higher effective fluxes through the transition state than equilibrium statistics would suggest. Experimental evidence for this includes anomalous temperature dependence observed in solution-phase protein conformational dynamics, where prefactors exceeding the canonical TST value of approximately $10^{13} s^{-1} (derived from vibrational frequencies) indicate enhanced rates due to non-equilibrated landscapes, as seen in studies of impaired protein folding under cellular conditions. Such anomalies manifest as curved Arrhenius plots or unusually low apparent activation energies, underscoring the failure to maintain reactant-transition state equilibrium in viscous or structured solvents. This limitation ties directly to non-equilibrium thermodynamics, where steady-state flux formulations or generalized master equations are needed to account for persistent gradients and irreversible entropy production beyond detailed balance.

Recrossing and Tunneling Effects

In transition state theory (TST), the core assumption is that all trajectories crossing the dividing surface at the transition state proceed to products without returning to reactants, corresponding to a transmission coefficient \kappa = 1. However, classical dynamical effects can cause recrossing, where a fraction of trajectories cross the surface but reverse direction due to solvent friction, anharmonicities, or multidimensional barrier curvature, reducing the effective rate and requiring \kappa < 1 as a correction factor in the rate expression k = \kappa k^\ddagger. Recrossing is particularly evident in trajectory simulations, where it is defined as a barrier crossing followed by return within a short time window, such as 0.2 ps. Classical molecular dynamics (MD) studies quantify recrossing by tracking the ratio of reactive to total crossings. For the gas-phase isomerization NCCN \rightleftharpoons NCNC, simulations over microsecond timescales reveal recrossings occurring roughly twice as frequently as successful reactive events, yielding \kappa \approx 0.5 due to prolonged residence times near the barrier compared to first-passage times. In the more complex CH_3CN \rightleftharpoons CH_3NC isomerization, recrossings outnumber reactive events by about 10:1, highlighting greater dynamical bottlenecks in systems with broader barriers. In contrast, for enzymatic hydride transfer in human , MD simulations at the equicommittor dividing surface show milder recrossing, with \kappa \approx 0.94 in the enzyme versus 0.88 in water, implying 6–12% of flux arises from recrossing pairs that contribute to non-reactive returns. Such computational trajectory analyses across systems indicate recrossing fractions of 10–20% in many condensed-phase reactions, underscoring the need for dynamical corrections to TST. Quantum tunneling introduces another deviation from classical TST by enabling particles to penetrate energy barriers, especially in hydrogen-atom or proton-transfer reactions where the light mass and narrow, high barriers favor non-classical over thermal activation. This effect enhances reaction rates, particularly at low temperatures, and is incorporated via a tunneling correction factor \Gamma > 1 multiplied into the TST rate. The seminal Bell correction provides an analytical estimate for parabolic barriers, approximating the transmission probability through the inverted barrier potential and accounting for the particle's wave-like . For general barrier shapes, semiclassical estimates employ the to compute the tunneling probability: P(E) \approx \exp\left( -\frac{2}{\hbar} \int_{x_1}^{x_2} \sqrt{2m \left( V(x) - E \right)} \, dx \right), where the integral spans the turning points x_1 and x_2 of the forbidden region under the barrier V(x) for energy E, with m the reduced mass; this integral quantifies the exponential suppression of tunneling, which diminishes rapidly with increasing barrier width or mass. Tunneling notably amplifies kinetic isotope effects (KIEs) in H/D substitutions, as the lighter isotope tunnels more efficiently than , yielding KIE values (k_H / k_D) far exceeding classical predictions based solely on differences. In enzymatic proton transfers, such as those in , tunneling contributions can elevate primary KIEs to over 20 at and above 100 at cryogenic conditions, providing direct evidence of quantum enhancement beyond 's classical framework. This isotope-dependent tunneling is a hallmark of quantum-corrected applications in biological and involving light-particle transfers.

Generalized Theories

Variational TST Approaches

Variational transition state theory (VTST) extends classical theory by optimizing the location of the dividing surface along the reaction path to minimize the rate constant and reduce recrossing of reactive trajectories. The principle involves selecting a generalized transition state at the position of maximum along the minimum energy path (MEP), ensuring that the forward flux through this surface approximates the net reactive flux without significant recrossings. This optimization addresses limitations in conventional TST, where the dividing surface at the may allow trajectories to recross, overestimating the rate. In microcanonical VTST (μVTST), the dividing surface is varied at a fixed total energy, employing RRKM-like densities of states to compute energy-specific rate constants. The microcanonical rate constant is given by k_{\mu \text{VT}}(E) = \min_s \frac{N_{\text{GT}}(E,s)}{h \rho_R(E)}, where N_{\text{GT}}(E,s) is the phase-space volume of the generalized transition state at reaction coordinate s and energy E, \rho_R(E) is the reactant density of states, and h is Planck's constant; the thermal rate constant is then obtained by Boltzmann averaging over the energy distribution. This approach is particularly useful for systems where energy-specific dynamics are important, such as in unimolecular reactions. Canonical VTST (CVT), in contrast, incorporates thermal averaging by locating the dividing surface at the position that minimizes the canonical rate constant along the reaction path. The CVT rate constant is expressed as k_{\text{CVT}}(T) = \min_s \frac{k_B T}{h} \frac{Q_{\text{GT}}(s^\dagger, T)}{Q_R(T)} \exp\left(-\frac{V_{\text{MEP}}(s^\dagger)}{k_B T}\right), where s^\dagger is the optimizing , Q_{\text{GT}} and Q_R are the partition functions of the generalized and reactants, respectively, V_{\text{MEP}}(s) is the along the MEP, k_B is Boltzmann's constant, and T is temperature; for bimolecular reactions, Q_R(T) involves the product of reactant partition functions. This method provides a thermally averaged optimization, making it suitable for temperature-dependent rate predictions. Further improvements in VTST include rate-constant optimization (RCO), which refines the by dynamically adjusting the dividing surface to account for residual recrossing effects, thereby enhancing the accuracy of the predicted s. RCO builds on the by incorporating trajectory-based corrections to the , often reducing errors in systems with complex surfaces. These developments, pioneered in the foundational work of Truhlar and Garrett, have established VTST as a for reliable classical rate calculations. Recent advances as of 2025 integrate , such as object-aware equivariant diffusion models and PSI-Net neural networks, to accelerate prediction on surfaces. Additionally, optimal transport methods have been applied to generate states more efficiently for complex reactions.

Semiclassical and Nonadiabatic Variants

With foundational work in the 1970s and further developments in the 1990s, semiclassical transition state theory (SCTST) extends classical TST by incorporating quantum effects through semiclassical approximations, particularly for tunneling through barriers along reaction paths. SCTST evaluates reaction rates using the properties of classical trajectories on an inverted , akin to paths that represent the most probable tunneling trajectory. This approach addresses limitations of classical TST in low-temperature regimes where quantum tunneling dominates, providing rate constants that include and tunneling corrections without full quantum dynamical calculations. A key component of SCTST is theory, which models tunneling as periodic orbits on the upside-down barrier potential. Seminal applications of instanton methods to chemical reactions trace back to adaptations of field-theoretic s for metastable decay, where the instanton determines the tunneling factor in the rate expression. In practice, the instanton rate constant k is approximated as k = \frac{k_B T}{h} e^{-S/\hbar}, with S the instanton , enabling accurate predictions for barrier tunneling in systems like H + H_2 reactions at cryogenic temperatures. Recent refinements, such as uniform semiclassical theory, resolve divergences near crossover temperatures by integrating over energy-dependent probabilities, improving accuracy across a wide range. Nonadiabatic variants of TST account for transitions between states, crucial for processes like where between and motion breaks the Born-Oppenheimer approximation. Marcus theory provides the foundational framework for nonadiabatic rates, expressing the rate as k = \frac{2\pi}{\hbar} |V|^2 \frac{1}{\sqrt{4\pi \lambda k_B T}} \exp\left( -\frac{(\Delta G^0 + \lambda)^2}{4\lambda k_B T} \right), where V is the , \lambda the reorganization energy, and \Delta G^0 the change. This semiclassical model assumes weak and treats motion classically, predicting inverted region behavior observed experimentally in outer-sphere transfers. To incorporate nonadiabatic transitions explicitly, is often combined with the Landau-Zener formula, which estimates the probability P of staying on the same adiabatic surface as P = \exp\left( -\frac{2\pi V^2}{\hbar |dE/dR| v} \right), with v the nuclear at the crossing point. This hybrid approach, known as Marcus-Landau-Zener theory, quantifies the transition probability at conical intersections or avoided crossings, enabling rate calculations for photoinduced electron transfers in photosynthetic systems. Surface hopping methods further enhance nonadiabatic by simulating trajectories that hop between surfaces based on LZ probabilities, providing dynamical corrections to statistical rates in multidimensional nonadiabatic regimes. Improvements to variational TST (CVT) incorporate microcanonical optimized (μO) multidimensional tunneling for barriers in polyatomic systems. In CVT/μOMT, the dividing surface is variationally optimized at the level, while tunneling is treated microcanonically along the reaction path, using the minimum path for the imaginary-frequency mode and Eckart barriers for others. This captures corner-cutting effects in multidimensional space, yielding rate constants over 200–500 K with mean deviations of 17% and up to 23% compared to benchmark calculations for reactions like H + CH_4. For instance, in bimolecular abstractions, μOMT enhances coefficients by factors up to 10 at low temperatures, emphasizing tunneling's role without overestimating recrossing. Hybrid quantum-classical methods, such as ring-polymer (RP-), approximate quantum thermal rates by mapping the system to a classical ring in an extended . RP- defines the dividing surface through the , yielding rates that converge to classical at high temperatures and include quantum delocalization and tunneling at low ones. Pioneered in the mid-2000s, this approach computes thermal flux correlation functions via on the free-particle ring-polymer , achieving errors below 5% for H + H_2 compared to exact quantum scattering. Post-2000 developments include centroid , which uses the ring-polymer centroid to enforce quantum Boltzmann statistics in the short-time flux-side limit. This QTST variant derives rates from centroid dynamics, ensuring positive-definite statistics and equivalence to ring-polymer MD-TST in the t \to 0^+ limit. For activated processes, centroid-density methods treat recrossing via quasi-classical trajectory corrections, providing quantum-corrected rates for barrier crossings in condensed phases with deviations under 15% from path-integral benchmarks.

Applications

Enzymatic Reactions

In the context of enzymatic reactions, transition state theory (TST) posits that enzymes accelerate reaction rates by stabilizing the more effectively than the substrates, thereby lowering the activation barrier ΔG‡. This concept was first articulated by in 1948, who proposed that enzymes achieve their catalytic proficiency through structural complementarity that allows tighter binding to the high-energy complex compared to the substrates, effectively reducing the difference between the enzyme-substrate complex and the . Pauling's idea shifted the understanding of enzyme specificity from mere substrate recognition to active involvement in stabilization, providing a foundational framework for interpreting catalytic rate enhancements observed in biological systems. The integration of TST with Michaelis-Menten kinetics further elucidates enzymatic turnover, where the catalytic rate constant k_cat is expressed according to the Eyring formulation adapted for enzyme-bound reactions: k_\text{cat} = \frac{k_B T}{h} \exp\left(-\frac{\Delta G^\ddagger_\text{enz}}{RT}\right) Here, ΔG‡_enz represents the activation free energy for the enzyme-catalyzed process, which is diminished relative to the uncatalyzed reaction due to transition state binding. This relationship highlights how enzymes can achieve rate accelerations of up to 10^17-fold by compressing the transition state energy landscape, with k_cat reflecting the frequency of successful barrier crossings once the enzyme-substrate complex forms. analogs, stable molecular mimics of the transition state geometry, exploit this principle as potent inhibitors by binding tightly to the enzyme's , often with dissociation constants in the picomolar to nanomolar range, thereby blocking . A representative example of transition state stabilization occurs in serine proteases such as , where the oxyanion hole—formed by backbone amide s from glycine 193 and serine 195—provides bonding to stabilize the negatively charged in the tetrahedral during . This interaction lowers the activation barrier by approximately 5-10 kcal/mol, facilitating the nucleophilic attack by the catalytic serine. However, TST assumptions can deviate in certain enzymatic processes, particularly those involving light atom transfers. For instance, in hydride transfer reactions catalyzed by , quantum mechanical tunneling contributes significantly to the reaction rate, allowing the hydride to bypass the classical barrier and enhancing the observed kinetic isotope effects beyond what classical TST predicts. Such quantum effects underscore the need for semiclassical extensions to TST in biological transfer mechanisms.

Surface Reactions and Catalysis

Transition state theory (TST) provides a for understanding the of surface reactions in , where reactants adsorb onto the catalyst surface before undergoing transformation at a . In , the theory accounts for the adsorption step as a precursor to the , with the rate determined by the barrier between adsorbed reactants and the . Adsorption isotherms, particularly the Langmuir model, describe the coverage of surface sites by reactants, which directly influences reaction rates in catalytic processes. The Langmuir isotherm assumes monolayer adsorption on homogeneous sites without lateral interactions, leading to a fractional coverage θ = K P / (1 + K P), where K is the and P is the . In TST applications, this coverage dependence modifies rate expressions for surface reactions, as the probability of forming the scales with the availability of adjacent occupied sites, resulting in rate laws like r = k θ_A θ_B for bimolecular Langmuir-Hinshelwood mechanisms. Coverage-dependent rates arise because high coverages can block sites or alter activation barriers through adsorbate interactions, deviating from ideal low-coverage assumptions. For dissociative adsorption, a key step in many catalytic cycles, TST identifies the transition state as the configuration where the adsorbing molecule reaches the on the (PES) en route to separated surface-bound fragments. A classic example is the dissociative adsorption of H₂ on surfaces like Cu or Pd, where the transition state involves stretched H-H bonds and partial to the metal, with the activation barrier often low or absent for late s. Reversible work TST variants refine this by incorporating configurational and vibrational modes at the transition state to predict sticking probabilities, showing good agreement with experimental dissociation rates for H₂ on Cu(111). The Brønsted-Evans-Polanyi (BEP) relation links activation barriers to overall reaction energetics in surface , stating that the E_a scales linearly with the reaction energy ΔE: E_a = α ΔE + β, where α (0 < α < 1) reflects the position of the along the , and β is a constant. This empirical correlation, derived from PES features, enables prediction of barriers from adsorption energies without full searches, aiding catalyst screening. In , BEP relations often yield volcano plots of activity versus , optimizing catalysts where the balances weak and strong adsorption. TST and BEP principles illuminate the ammonia synthesis reaction on iron catalysts, where N₂ dissociation is rate-limiting, with the transition state involving side-on N₂ coordination and partial N-N bond breaking on Fe(111) sites. The overall mechanism follows a Langmuir-Hinshelwood pathway, with hydrogen addition steps having lower barriers, and BEP scaling confirms Fe's position near the volcano peak due to moderate N binding. Experimental turnover frequencies align with TST-derived rates, emphasizing promoter effects like K or Al₂O₃ in lowering the N₂ dissociation barrier. In CO oxidation on catalysts, femtosecond spectroscopy probes the transient , revealing a bent CO-O intermediate where C-O bond formation occurs atop the surface. TST models this as the rate-determining step in the Mars-van Krevelen or Langmuir-Hinshelwood cycles, with BEP relations correlating O₂ activation barriers to formation energies, guiding design of active nanoparticles. Michael Polanyi's work in pioneered semiempirical construction of PES for surface reactions, collaborating with Eyring to apply and experimental data to map energy landscapes for gas-metal interactions, laying groundwork for TST in adsorption dynamics.

Computational Implementations

Computational implementations of transition state theory (TST) have become integral to , enabling the prediction of reaction rates by locating (TS) and computing associated properties on surfaces (PES). These methods typically involve optimizing geometries to find saddle points, performing frequency analyses to confirm a single imaginary frequency, and applying variational principles to refine rate constants. Seminal algorithms and software packages facilitate this process, often integrating or (DFT) levels for accuracy. Transition state search algorithms are essential for navigating the PES to identify saddle points. The quadratic synchronous transit (QST) method, introduced by Peng and Schlegel, constructs an initial path by interpolating between reactant and product geometries, then alternates between maximizing energy along the reaction coordinate and minimizing in orthogonal directions using quasi-Newton updates. This approach is particularly effective for systems where approximate TS structures are available, converging to the TS with empirical Hessian estimates. Complementing QST, the nudged elastic band (NEB) method, developed by Henkelman, Uberuaga, and Jónsson, discretizes the reaction path into intermediate "images" connected by springs, optimizing them to minimize the total energy while projecting out components parallel to the path to avoid sliding. The climbing image variant of NEB enhances convergence to the highest-energy saddle point, making it robust for complex pathways. Popular quantum chemistry software implements these algorithms for TS optimization and validation. In Gaussian, the Opt=QST2 or QST3 keywords automate TS searches starting from reactant/product pairs or an initial guess, followed by Freq calculations to compute vibrational frequencies and confirm the TS nature via one imaginary mode. Similarly, supports TS optimization through its internal algorithms, including NEB-TS for direct saddle-point location, with analytical or numerical frequency computations via the NumFreq module to verify the . For variational TST (VTST), which minimizes the along the reaction path to improve conventional TST predictions, the POLYRATE program suite provides comprehensive tools, computing canonical and microcanonical VTST rates with multidimensional tunneling corrections using reaction-path curvatures from input PES data. Gaussian also incorporates VTST via its VTST module for variational rate evaluations. PES construction underpins these computations, traditionally relying on ab initio methods like coupled-cluster or DFT for scanning reaction coordinates. DFT, such as B3LYP or ωB97X-D functionals, efficiently maps PES for medium-sized molecules, providing energies and gradients for TS optimization and rate calculations, though it requires benchmark validation against higher-level theories for barrier heights. For larger systems, post-2010 advances in (ML) potentials have accelerated TST applications by approximating quantum-mechanical PES with neural networks or Gaussian processes trained on ab initio data, enabling simulations of thousands of atoms while retaining near-DFT accuracy for rate predictions. These ML models, exemplified by the DeePMD or frameworks, reduce computational cost by orders of magnitude, facilitating direct dynamics TST for condensed-phase s. Recent developments as of 2025, including equivariant graph neural networks like , further improve transferability and accuracy for biomolecular TST applications. Case studies illustrate these implementations. In the Diels-Alder cycloaddition of and , DFT-based QST searches in Gaussian locate the concerted with a barrier of approximately 25 kcal/mol at B3LYP/6-31G(d), enabling TST rate predictions that match experimental kinetics when variational effects are included via POLYRATE. For SN2 displacements, such as Cl⁻ + CH₃Cl in aqueous solution, NEB in combined with explicit solvent models (e.g., ) identifies the TS, revealing solvent stabilization that lowers the barrier by 10-15 kcal/mol compared to gas phase, with VTST rates incorporating polarization effects via polarizable continuum models aligning closely with measured rate constants.

References

  1. [1]
    The Activated Complex and the Absolute Rate of Chemical Reactions.
    Applying Transition-State Theory to Explore Transport and Selectivity in Salt-Rejecting Membranes: A Critical Review. Environmental Science & Technology ...Missing: original | Show results with:original
  2. [2]
    [PDF] The Development of Transition-State Theory
    Transition-state theory, as developed in 1935 by Eyring and by Evans and Polanyi, is the culmination of a series of investigations that are conveniently ...
  3. [3]
    The Activated Complex in Chemical Reactions - AIP Publishing
    The probability of the activated state is calculated using ordinary statistical mechanics. This probability multiplied by the rate of decomposition gives the ...
  4. [4]
    Activation Energy - Chemical Kinetics
    The activation energy is the change in the internal energy of the system, which is not quite true. E a measures the change in the potential energy of a pair of ...
  5. [5]
    [PDF] The Transition State and Cognate Concepts - University of Bath
    Contents. 1. Introduction. 2. Conceptual power of the transition state. 3. Glossary of Terms. 3.1 Potential-energy surface (PES). 3.2 Saddle point.
  6. [6]
    [PDF] Potential Energy Surfaces and Dynamics Calculations
    Conventionally the transition state is defined as a surface in configuration space separating the reactant and product regions and located at the saddle.
  7. [7]
    Enzymatic Transition States, Transition-State Analogs, Dynamics ...
    Chemical transition states have short lifetimes of a few femtoseconds, the time required for electron redistribution and the time to convert the restoring mode ...
  8. [8]
    Transition States in Chemical Reactions
    Transition states correspond to saddle points on the potential energy surface. Like minima, the saddle points are stationary points with all forces zero.
  9. [9]
  10. [10]
  11. [11]
    [PDF] The Potential Energy Surface in molecular quantum mechanics - arXiv
    Apr 9, 2013 · Abstract—The idea of a Potential Energy Surface (PES) forms the basis of almost all accounts of the mechanisms of chemical.
  12. [12]
    [PDF] Potential Energy Surfaces
    In such a region there is a nonzero quantum mechanical probability that the system “hops” to another surface. Based on this probability one portion of the quan-.
  13. [13]
    Crossing the dividing surface of transition state theory. III. Once and ...
    Sep 14, 2015 · The theory is most easily implemented if the dividing surface is reduced to a plane in configuration space, passing through a saddle point of ...Missing: validate | Show results with:validate
  14. [14]
    Potential Energy Surface for H a - AIP Publishing
    The diatomic Coulomb and exchange integrals are estimated from accurate values for the H2 molecule by a modified London-Eyring-Sato procedure, while the three- ...
  15. [15]
  16. [16]
    An accurate three‐dimensional potential energy surface for H3
    Mar 1, 1978 · An accurate three‐dimensional potential energy surface for H 3 has been obtained by the configuration interaction (CI) method.
  17. [17]
    [PDF] generalized transition state theory
    The most complete calculation we perform is to find the variationally best dividing surface for each total energy;20.21 this is called microcanonical ...Missing: validate | Show results with:validate
  18. [18]
  19. [19]
    Theory of enthalpy-entropy compensation - ACS Publications
    Glass transition memorized by the enthalpy-entropy compensation in the shear thinning of supercooled metallic liquids. Journal of Physics: Condensed Matter ...
  20. [20]
    The Nature of the Enthalpy–Entropy Compensation and “Exotic ...
    It is shown that EEC results from the linear correlations of enthalpy and entropy of activation with dimer length, which are derived directly from the ...
  21. [21]
    Validation of Enthalpy–Entropy Compensation Mechanism in Partial ...
    Apr 13, 2020 · Eyring analysis was then used to convert the chemical exchange rates to determine the transition state enthalpy and entropy of the molecules.
  22. [22]
    [PDF] Kinetic Isotope Effects in Organic Chemistry - Macmillan Group
    Sep 14, 2005 · The difference in energy between the AZPE's in the ground state and the transition state determine the magnitude of the kinetic isotope effect.
  23. [23]
    Initial-state and transition-state solvent effects on reaction rates and ...
    Initial-state and transition-state solvent effects on reaction rates and the use of thermodynamic transfer functions. Click to copy article linkArticle link ...
  24. [24]
    A Correlation of Reaction Rates | Journal of the American Chemical ...
    The role of elemental factors in the reactivity of combination reactions between heavy vinylidene-like molecules and carbene analogues.Missing: postulate | Show results with:postulate
  25. [25]
    [PDF] Rudolph A. Marcus - Nobel Lecture
    In chemists' terminology this line represents the “transition state” of the reaction. In transition state theory a quasi-equilibrium between the transition ...
  26. [26]
    Brownian motion in a field of force and the diffusion model of ...
    A particle which is caught in a potential hole and which, through the shuttling action of Brownian motion, can escape over a potential barrier yields a ...
  27. [27]
    Reaction-rate theory: fifty years after Kramers | Rev. Mod. Phys.
    Apr 1, 1990 · Fifty years after Kramers' seminal paper on thermally activated barrier crossing, the authors report, extend, and interpret much of our current understanding.Missing: critiques | Show results with:critiques
  28. [28]
    Impaired protein conformational landscapes as revealed in ... - PNAS
    Jun 13, 2011 · Values of Aobs far in excess of 1013 s-1 indicate an unusual temperature dependence for chemical reactions modeled by transition state theory ( ...
  29. [29]
    Thermodynamics of the transition state and the application to ...
    According to transition-state theory, when the equilibrium hypothesis is valid, activated complexes formed from reactants may be treated thermodynamically ...Missing: formulations Christiansen Evans
  30. [30]
    Quantifying the limits of transition state theory in enzymatic catalysis
    Nov 3, 2017 · We present a computational strategy to quantify the error associated to transition state theory from the number of recrossings observed at the equicommittor.<|control11|><|separator|>
  31. [31]
    THE TUNNEL EFFECT CORRECTION FOR PARABOLIC ...
    The transition-state formulation of chemical kinetics involves a quantum correc- tion which can be expressed in terms of the curvatures of the energy ...
  32. [32]
    [PDF] Semiclassical Multidimensional Tunnelling Calculations
    The considerations in the previous paragraph imply that the tunnelling correction to transition state theory should involve barrier heights deter- mined by ...
  33. [33]
    Large kinetic isotope effects in enzymatic proton transfer and ... - NIH
    When tunneling is turned on, the additional transfer mechanism increases the rate and lnk starts to curve upward. The Arrhenius prefactor is the projection at β ...
  34. [34]
  35. [35]
    VARIATIONAL TRANSITION STATE THEORY - Annual Reviews
    Transition state theory (TST)l is the most widely used theory for calculating rates of bimolecular reactions occurring in the gas phase and in condensed.
  36. [36]
    Variational transition state theory: theoretical framework and recent ...
    Nov 22, 2017 · This article reviews the fundamentals of variational transition state theory (VTST), its recent theoretical development, and some modern applications.Missing: seminal papers
  37. [37]
    Semiclassical transition state theory. A new perspective
    The semiclassical transition state theory (SCTST) introduced by Miller, Hernandez, Handy, Jayatilaka and Willetts requires the inversion of an (effectively ...
  38. [38]
    Semiclassical limit of quantum mechanical transition state theory for ...
    Mar 1, 1975 · The semiclassical limit of quantum mechanical transition state theory is derived by invoking the classical path approximation for the Boltzmann density ...Missing: 1983 | Show results with:1983
  39. [39]
    Semiclassical instanton theory for reaction rates at any temperature
    Sep 4, 2024 · This paper presents a semiclassical theory for reaction rates valid at any temperature, solving the crossover temperature problem by smoothly ...Missing: seminal | Show results with:seminal
  40. [40]
    Uniform Semiclassical Instanton Rate Theory - ACS Publications
    Oct 31, 2023 · The uniform semiclassical instanton rate theory derives thermal transmission probability, using a uniform energy-dependent transmission ...Missing: seminal | Show results with:seminal
  41. [41]
    The Landau−Zener Formula | The Journal of Physical Chemistry B
    The Landau−Zener formula for the probability that a nonadiabatic transition has taken place is derived without solving directly the usual second-order ...
  42. [42]
    Crossover between the adiabatic and nonadiabatic electron transfer ...
    Jan 19, 2021 · The semiclassical models of nonadiabatic transition were proposed first by Landau and Zener in 1932, and have been widely used in the study ...
  43. [43]
    A nonadiabatic description of electron transfer reactions involving ...
    The nonadiabatic model presented here predicts an almost linear dependence of lnk upon the free energy change ΔF0 in the abnormal free energy region in contrast ...
  44. [44]
    [PDF] Test of Variational Transition State Theory with Multidimensional ...
    Table I shows an average absolute deviation between the CVT/μOMT and accurate rate constants of only 17% over the five temperatures, with a maximum deviation of ...
  45. [45]
    Microcanonical Tunneling Rates from Density-of-States Instanton ...
    Dec 22, 2020 · Canonical variational theory with microcanonically optimized multidimensional tunnelling was used for bimol. rates, modeling the gas-phase ...
  46. [46]
    Chemical reaction rates from ring polymer molecular dynamics
    Craig, David E. Manolopoulos; Chemical reaction rates from ring polymer molecular dynamics. J. Chem. Phys. 22 February 2005; 122 (8): 084106. https://doi ...
  47. [47]
    Chemical Reaction Rate Coefficients from Ring Polymer Molecular ...
    Sep 14, 2016 · This Feature Article presents an overview of the current status of ring polymer molecular dynamics (RPMD) rate theory.
  48. [48]
    Derivation of a true (t → 0+) quantum transition-state theory. II ...
    This t → 0+ limit appears to be unique in giving positive-definite Boltzmann statistics, and is identical to ring-polymer molecular dynamics (RPMD) TST. Here, ...Missing: post- | Show results with:post-
  49. [49]
    Centroid‐density quantum rate theory: Dynamical treatment of ...
    A new method is presented for the calculation of quantum mechanical rate constants for activated processes. This method is a hybrid approach involving ...Missing: TST | Show results with:TST
  50. [50]
    Molecular Architecture and Biological Reactions - ACS Publications
    Molecular Architecture and Biological Reactions. Click to copy article link ... Specificity in Transition State Binding: The Pauling Model Revisited.
  51. [51]
    Looking Back: A Short History of the Discovery of Enzymes and How ...
    Sep 9, 2020 · In 1948 Linus Pauling proposed that enzymes had to stabilize the transition state rather than the substrate as proposed by Fischer. The ...Missing: tightly | Show results with:tightly
  52. [52]
    Transition States and Transition State Analogue Interactions with ...
    Apr 7, 2015 · The transition state structure, defined by the position of all atoms participating in the reaction coordinate when the transition state is ...
  53. [53]
    Transition-state stabilization at the oxyanion binding sites of serine ...
    Studies of thioamide effects on serine protease activity enable two-site stabilization of cancer imaging peptides.
  54. [54]
    Hydrogen Tunneling in Enzyme Reactions - Science
    Semiclassical mass considerations that do not include tunneling effects would predict that kH/kT = (kD/kT)3.26, where kH, kD, and kT are the rate constants for ...
  55. [55]
    [PDF] FEATURE ARTICLE
    The transition states to elementary reaction steps on surfaces have received little attention because of the lack of experimental probes of their structure ...<|separator|>
  56. [56]
    [PDF] A context to reexamine the Langmuir-Hinshelwood kinetic description
    The Langmuir-Hinshelwood method describes catalytic surface reactions of adsorbed species obeying the conditions of the Langmuir theory of adsorption: (I) ...
  57. [57]
    (PDF) Reversible Work Transition-State Theory—Application to ...
    Aug 31, 2019 · ... dissociative sticking probability of H2 on a metal surface within the transition state theory. The reversible work involved in shifting the ...
  58. [58]
    Brønsted−Evans−Polanyi Relation of Multistep Reactions and ...
    A no. of elementary reactions at metal surfaces show a linear Bronsted-Evans-Polanyi relation between the activation energy and the reaction energy, and ...
  59. [59]
    Surface chemistry. Probing the transition state region in catalytic CO ...
    Feb 27, 2015 · Femtosecond x-ray laser pulses are used to probe the carbon monoxide (CO) oxidation reaction on ruthenium (Ru) initiated by an optical laser pulse.
  60. [60]
    Working tools for theoretical chemistry: Polanyi, eyring, and debates ...
    Dec 1, 2006 · This paper was written in Karlsruhe, and it uses the notion of radiation from the aether (p. 230).19 The work on transition-state theory began ...Abstract · Introduction · Polanyi and Eyring's... · Reception and Defense of the...
  61. [61]
    Machine Learning Force Fields | Chemical Reviews
    Mar 11, 2021 · ML-FFs combine the accuracy of ab initio methods and the efficiency of classical FFs. They provide easy access to a system's potential energy ...
  62. [62]
    [PDF] Combining Synchronous Transit and Quasi-Newton Methods
    In this paper we build on these ideas and our earlier efforts to assemble a transition state searching method that combines a linear or quadratic synchronous.Missing: seminal | Show results with:seminal
  63. [63]
    [PDF] CHAPTER 16 Nudged elastic band method for finding minimum en
    The Nudged Elastic Band (NEB) method finds the lowest energy path for atom rearrangements, used to define reaction coordinates. It is efficient for parallel ...
  64. [64]
    Transition State Optimizations with Opt=QST2 | Gaussian.com
    Jun 21, 2017 · This method will converge efficiently to the actual transition structure using an empirical estimate of the Hessian and suitable starting structures.Missing: seminal | Show results with:seminal
  65. [65]
    Freq - Gaussian.com
    Dec 16, 2020 · The Freq keyword computes force constants and vibrational frequencies, and intensities. It is best to compute frequencies after a geometry  ...
  66. [66]
    4.3. Transition State Searches - ORCA 6.1 Manual
    The transition state search is separated into two different optimization problems. The energy is maximized along one Hessian eigenmode and minimized along the ...
  67. [67]
    7.27. Frequency calculations - numerical and analytical - ORCA 6.0 ...
    The calculation of frequencies through the numerical or analytical Hessian is done via the orca_numfreq module and the combination of the orca_propint, orca_ ...
  68. [68]
    Polyrate 2023: A computer program for the calculation of chemical ...
    Polyrate can perform VTST calculations on gas-phase reactions with both tight and loose transition states. For tight transition states it uses the reaction-path ...
  69. [69]
    Quantum mechanical study of solvent effects in a prototype SN2 ...
    Both gas phase and solution results with the SRP model agree very well with experiment and provide insight into the specific role of solvent on the reaction ...Missing: displacement TST