Fact-checked by Grok 2 weeks ago

Arrhenius equation

The Arrhenius equation is an in that quantifies the dependence of the rate constant for a , building on observations by Dutch chemist Jacobus van 't Hoff and originally proposed by Swedish scientist in 1889 during his investigation of the acid-catalyzed inversion of cane sugar (). The equation takes the form k = A \exp\left(-\frac{E_a}{RT}\right), where k is the rate constant, A is the pre-exponential (or frequency) factor representing the frequency of collisions with proper orientation, E_a is the (the minimum energy barrier for the reaction), R is the universal gas constant, and T is the absolute in . This relationship demonstrates that reaction rates increase exponentially with temperature, as higher temperatures provide more molecules with energy exceeding the activation barrier, consistent with the of molecular energies. Arrhenius derived the equation from experimental data on hydrolysis, fitting logarithmic plots of rate constants versus temperature to yield straight lines, from which energies could be calculated. Although initially empirical, the equation later found theoretical justification through , which attributes the pre-exponential factor to and steric factors, and , which refines the concept. The Arrhenius equation remains a cornerstone of , enabling the prediction of reaction behavior across temperature ranges and the design of processes in industrial catalysis, pharmaceutical stability, and environmental modeling. It applies primarily to elementary reactions or overall rate constants approximated as such, though modifications account for complex mechanisms or non-ideal conditions like pressure effects in gases. Arrhenius's work earned him the 1903 , partly for advancing electrolyte theory.

Introduction and Formulation

Historical Development

In the late , emerged as a distinct field amid growing interest in the of chemical s and the behavior of electrolytes in , driven by pioneers seeking to quantify how influences velocities. This period saw intense focus on dissociation theory and equilibrium constants, with researchers like exploring the temperature dependence of chemical equilibria through empirical observations. Svante Arrhenius, a Swedish chemist, published his seminal work in 1889, deriving an empirical relationship for reaction rates from van 't Hoff's earlier studies on equilibrium constants and their temperature sensitivity. In the paper "Über die Reaktionsgeschwindigkeit bei der Inversion von Rohrzucker durch Säuren," Arrhenius analyzed data from acid-catalyzed inversion of cane sugar, proposing a form that linked rate constants exponentially to temperature, motivated by van 't Hoff's 1884 notation of similar dependencies. Early support came from van 't Hoff and Wilhelm Ostwald, who validated Arrhenius's ideas through their own experiments on ionic solutions during Arrhenius's visits to their laboratories between 1886 and 1888. Ostwald provided kinetic evidence reinforcing the role of dissociated ions in reaction speeds, while van 't Hoff confirmed the temperature effects in equilibrium systems, helping establish the equation's reliability for solution chemistry. Initially applied to ionic dissociation in aqueous solutions, the equation explained temperature-driven variations in reaction velocities, such as in electrolytic processes and acid-base , laying groundwork for broader kinetic studies. Arrhenius's electrolytic theory earned him the in , recognizing his advancements in understanding solution conductivity and dynamics, while his work on rates represented a significant separate contribution to .

Basic Mathematical Formulation

The Arrhenius equation expresses the temperature dependence of the rate constant for a in the standard form k = A \exp\left( -\frac{E_a}{RT} \right), where this empirical relation was originally proposed by in 1889 based on studies of inversion. In this equation, k denotes the rate constant, a proportionality factor in the rate law that determines the speed of the reaction for a given concentration of reactants; it applies directly to elementary reactions within integrated rate laws. The units of k vary with reaction order—for instance, s^{-1} for unimolecular () processes and M^{-1} s^{-1} for bimolecular (second-order) ones—to ensure dimensional consistency in rate expressions. The A, often termed the factor, quantifies the rate of molecular collisions with appropriate orientation and proximity for initiation, reflecting the inherent efficiency of reactant encounters independent of . It carries the same units as k. The E_a represents the energy barrier that reactants must surmount to form products, corresponding to the difference between the average energy of reactants and the . Typically reported in J ^{-1} or ^{-1}, E_a governs the exponential sensitivity of the to . The parameter R is the universal gas constant with value 8.314 J mol^{-1} K^{-1}, serving to normalize the activation energy on a per-mole and per-kelvin basis, while T is the absolute temperature in kelvin, ensuring the equation captures thermal effects accurately across conditions.

Analysis and Visualization

Empirical Derivation

The empirical derivation of the Arrhenius equation originated from observations of temperature effects on chemical equilibria and reaction rates in the late 19th century. Jacobus Henricus van 't Hoff established the foundational relation for equilibrium constants in his 1884 work, proposing that the temperature dependence follows the differential form \frac{d \ln K}{dT} = \frac{\Delta H}{RT^2}, where K is the equilibrium constant, \Delta H is the standard enthalpy change of the reaction, R is the gas constant, and T is the absolute temperature. This equation, derived from thermodynamic principles assuming constant \Delta H, described how equilibria shift with temperature for various systems, such as dissociation reactions. Svante Arrhenius extended van 't Hoff's relation to in 1889 by recognizing that rate constants exhibit analogous temperature sensitivity. For an in , the K equals the ratio of the forward rate constant k_f to the reverse rate constant k_r, so \ln K = \ln k_f - \ln k_r. Differentiating yields \frac{d \ln K}{dT} = \frac{d \ln k_f}{dT} - \frac{d \ln k_r}{dT}. Assuming the E_{a,f} and E_{a,r} for the forward and reverse processes satisfy E_{a,f} - E_{a,r} = \Delta H, Arrhenius posited that each rate constant obeys a similar differential form: \frac{d \ln k}{dT} = \frac{E_a}{RT^2}, where E_a represents the . This assumption bridged to empirically, treating the rate constant's variation as mirroring that of K but scaled by the energy barrier E_a. Integrating the differential equation \frac{d \ln k}{dT} = \frac{E_a}{RT^2} under the approximation of constant E_a provides the explicit temperature dependence. Rearranging gives d \ln k = \frac{E_a}{R} \frac{dT}{T^2}, and integration from a reference temperature where k \to A (the pre-exponential factor) to T yields: \ln k = \ln A - \frac{E_a}{RT} or equivalently, k = A e^{-E_a / RT}. This integrated form captures the exponential increase in rate with temperature, with A representing the frequency factor derived from the limiting behavior at infinite temperature. The derivation relied on the validity of the constant-energy assumption over moderate temperature ranges, which aligned with early experimental observations. To validate this relation empirically, Arrhenius applied it to measured rate data from the acid-catalyzed inversion of ( to glucose and ), conducting experiments at temperatures between 0°C and 70°C. He calculated rate constants k from time-dependent concentration changes and constructed plots of \ln k versus $1/T. These yielded straight lines, confirming the linear relationship \ln k = \ln A - (E_a / R)(1/T), with the providing -E_a / R (typically around 107 /mol for this reaction) and the giving \ln A. This graphical fitting demonstrated the equation's predictive power across diverse reactions, establishing it as a cornerstone of without invoking molecular theories at the time. Subsequent studies affirmed the linearity for many systems, though deviations occur at extreme temperatures.

Arrhenius Plot and Graphical Interpretation

The Arrhenius plot is a graphical representation used to visualize and analyze the temperature dependence of reaction rate constants as described by the Arrhenius equation. It is constructed as a semi-logarithmic plot with the natural logarithm of the rate constant, \ln k, on the y-axis and the reciprocal of the absolute temperature, $1/T (where T is in Kelvin), on the x-axis. Under ideal conditions assuming the basic Arrhenius model applies without complications, the data points form a straight line, reflecting the linear relationship derived from the equation's integrated form./Kinetics/06:_Modeling_Reaction_Kinetics/6.02:_Temperature_Dependence_of_Reaction_Rates/6.2.03:_The_Arrhenius_Law/6.2.3.04:The_Arrhenius_Law-_Arrhenius_Plots) From this linear plot, key parameters are extracted via linear regression. The slope m of the line is m = -E_a / R, where E_a is the activation energy and R is the universal gas constant (typically 8.314 J/mol·K); thus, E_a = -m \cdot R. The y-intercept b equals \ln A, allowing the pre-exponential factor A to be calculated as A = e^b. This method provides a straightforward way to quantify E_a and A from experimental data spanning a range of temperatures./Kinetics/06:_Modeling_Reaction_Kinetics/6.02:_Temperature_Dependence_of_Reaction_Rates/6.2.03:_The_Arrhenius_Law/6.2.3.04:The_Arrhenius_Law-_Arrhenius_Plots) In practical applications, experimental rate constants measured at different temperatures often show scatter due to inherent measurement uncertainties, such as imprecise or variability in reactant concentrations. To address this, the data points are fitted using least-squares regression to determine the best-fit line, which minimizes the sum of squared residuals and yields the and intercept with associated standard errors. For example, in kinetic studies of reactions or processes, scattered data from replicate experiments at temperatures like 300 , 310 , and 320 can be plotted and regressed to estimate E_a reliably, provided the temperature is sufficient to capture meaningful variation in k. Error analysis is essential for validating the reliability of extracted parameters, particularly the slope, as it directly influences E_a. Sources of error include inaccuracies in determining k from initial rate methods or nonlinear effects from unaccounted variables, which can lead to biased slopes if unweighted regression is used on heteroscedastic data. Weighted least-squares methods, where points with higher precision (e.g., from lower-variance measurements) are given more influence, help mitigate this; for instance, errors in E_a can be reduced by 10-20% in typical kinetic datasets by incorporating uncertainties in both \ln k and $1/T. The propagation of these errors follows standard formulas, such as \Delta E_a = R \cdot \Delta m \cdot \sqrt{1 + (m \cdot \Delta x / \Delta m)^2}, emphasizing the need for multiple data points across a broad temperature range. Non-linearity in the Arrhenius plot, such as or breaks in the line, signals potential failure of the classical model to describe the data adequately, often prompting evaluation of alternative mechanisms or extended equations.

Empirical Extensions

Modified Arrhenius Equation

The modified Arrhenius equation addresses limitations in the basic form by incorporating empirical adjustments to enhance accuracy for systems where the exhibits dependence, such as in certain gas-phase and solution-phase reactions. A widely used empirical modification takes the form k = A T^{n} \exp\left(-\frac{E_a}{RT}\right), where n is a fitted exponent that captures deviations from ideality, often ranging from -1 to 2 depending on the reaction type. This adjustment stems from experimental observations that the A varies with temperature, necessitating corrections for factors like molecular collision frequencies or phase-specific influences in non-ideal conditions. In gas-phase reactions, values of n around 0.5 are commonly observed for many bimolecular processes due to considerations, and modified forms provide a better fit to measured rate constants across ranges compared to the linear Arrhenius plot assumption. For diatomic processes, such as those involving vibrational energy distributions, modified expressions like this account for empirical in rate data, improving predictions in high-temperature environments. Another specific empirical form is the Evans-Polanyi , which correlates activation energy with as E_a = \alpha \Delta H_r + \beta, where \alpha (typically 0.5–1) and \beta are fitted parameters; this enables estimation of rate constants for analogous reactions by assuming a shared , particularly useful in solution-phase systems where solvent interactions alter energetics. In solution-phase contexts, such modifications often incorporate or effects to refine rate predictions, as demonstrated in studies of organic transfers. Parameters in these modified equations are typically determined through non-linear least-squares regression on experimental rate constant data versus , minimizing residuals to yield optimized A, n, and E_a values while ensuring physical reasonableness (e.g., positive n for endothermic pathways). This fitting approach has been applied successfully in databases to compile rate coefficients for hundreds of gas-phase organics, highlighting its practical utility in empirical modeling.

Temperature-Dependent Prefactor Adjustments

In gas-phase reactions, empirical adjustments to the A in the Arrhenius equation often incorporate a T^{1/2} dependence, arising from the Maxwell-Boltzmann distribution of molecular speeds within . This term reflects the increased average of reactant molecules with rising temperature, enhancing collision rates even before accounting for the activation barrier. For bimolecular gas-phase processes, the modified rate constant thus takes the form k = A T^{1/2} \exp(-E_a / RT), where the T^{1/2} factor provides a more accurate description of low-activation-energy reactions compared to the temperature-independent A. For reactions in , particularly those limited by reactant , modifications to A account for and solvent through the Stokes-Einstein relation, yielding A \propto T / \eta. Here, \eta is the dynamic , which typically decreases with following its own Arrhenius-like behavior, thereby amplifying the effective at higher temperatures via faster . This adjustment is crucial for diffusion-controlled reactions, such as recombination or enzyme-substrate encounters in aqueous media, where the Smoluchowski equation predicts the encounter constant as k_\text{diff} = (8k_B T)/(3\eta) for spherical particles, linking directly to the pre-exponential term. Specific applications highlight these adjustments; in , empirical models frequently extend the Arrhenius form to k = A T^m \exp(-E_a / RT) with m > 0 (often around 1–2) to capture the combined effects of collisional frequency and structural flexibility, improving fits across physiological temperature ranges without invoking full thermodynamic profiles. In , such temperature-dependent prefactors are essential for parameterizing rate constants in tropospheric models, where varying altitudes introduce T gradients that influence oxidant-reactant interactions, as seen in OH kinetics with alkenols. When fitting experimental data to these adjusted Arrhenius expressions, strong correlations emerge between E_a and \log A, known as the kinetic compensation , where an increase in is often offset by a larger to maintain consistent rate behavior across datasets. This trade-off complicates parameter estimation, as it can lead to multiple equivalent fits with varying temperature sensitivities, necessitating careful validation through Arrhenius plots or to ensure physical interpretability.

Theoretical Interpretations

Activation Energy in Arrhenius's Original Concept

In his seminal 1889 paper on the acid-catalyzed inversion of , conceptualized , denoted as E_a, as the minimum energy required by molecules to form an activated for the to occur. This idea stemmed from the proportion of molecules that have energy exceeding a certain threshold, which he derived using the , leading to the \exp(-E_a / [RT](/page/RT)) in the expression, where R is the and T is the . Arrhenius grounded this concept in thermodynamics by connecting it to the heat of reaction and principles of . For reversible reactions, he established that the difference between the activation energies of the forward and reverse processes equals the overall heat of reaction \Delta H, under the approximation that E_a \approx \Delta H^\ddagger, the of , thereby linking kinetic barriers to thermodynamic driving forces. Within the framework of Arrhenius's ionic theory, introduced two years earlier, the activation energy was contextualized as relating to the energy barrier for dissociation in solution, as the observed reaction rates for ionic catalysts like acids depended on the temperature-sensitive degree of dissociation. This led to an early interpretation where part of the temperature dependence was attributed to dissociation energetics rather than solely the reaction step itself. Initially, Arrhenius's formulation of was purely empirical, fitted to experimental rate data from sucrose inversion without any quantum mechanical underpinnings, reflecting the pre-quantum era understanding of molecular energies and dynamics.

Collision Theory Perspective

interprets the Arrhenius equation by modeling chemical reactions as resulting from collisions between reactant molecules in the gas phase, where only a of these collisions lead to successful reactions. According to this classical framework, the rate constant k of a bimolecular is given by k = p Z e^{-E_a / RT}, where p is the accounting for the orientation requirement of the collision, Z is the , and e^{-E_a / RT} represents the of collisions with sufficient energy to surmount the activation barrier. The collision frequency Z arises from kinetic molecular theory and, for unlike molecules, is expressed as Z = N_A \sigma_{AB}^2 \left( \frac{8 \pi k_B T}{\mu} \right)^{1/2}, where N_A is Avogadro's number, \sigma_{AB} is the average collision diameter, k_B is Boltzmann's constant, and \mu is the ; this yields a temperature dependence of Z \propto T^{1/2}, which approximates the pre-exponential factor A \approx p Z in the Arrhenius equation. The exponential term derives from the Maxwell-Boltzmann of molecular speeds and energies, where the proportion of molecules with translational exceeding E_a along the line of centers is approximately e^{-E_a / RT} when E_a \gg RT. In this perspective, the E_a corresponds to the minimum required in the collision for the molecules to overcome the repulsive forces and access the reactive part of the , often modeled as a hard-sphere with a central barrier. This interpretation links the thermodynamic activation to a for initiation. The theory relies on several key assumptions, including the hard-sphere model for molecular s, behavior with no intermolecular attractions, and reactions proceeding solely via collisions without considering multi-body effects. These simplifications hold reasonably for simple gas-phase at moderate temperatures but overlook quantum mechanical phenomena, such as tunneling through the energy barrier, which can enhance rates for involving or other light atoms.

Transition State Theory Framework

Transition State Theory (TST), developed by Henry Eyring and others in the 1930s, provides a statistical mechanical foundation for the Arrhenius equation by modeling chemical reactions as proceeding through a high-energy or . In TST, the reactants are assumed to be in quasi- with this , where the represents the on the separating reactants from products. The rate of reaction is then determined by the rate at which the cross the barrier along the , treated as a one-dimensional with imaginary frequency. This framework uses partition functions to describe the equilibrium distribution: the partition function for the Q^\ddagger (excluding the ) relative to the reactants' partition function Q yields the K^\ddagger = \frac{Q^\ddagger}{Q} e^{-\Delta E_0 / RT}, where \Delta E_0 is the difference and Q^\ddagger, Q are molar partition functions. The foundational result of TST is the Eyring equation for the rate constant k: k = \frac{k_B T}{h} \exp\left( -\frac{\Delta G^\ddagger}{RT} \right) Here, k_B is Boltzmann's constant, h is , R is the , T is the absolute temperature, and \Delta G^\ddagger is the standard of activation for forming the from reactants. This expression arises from the flux of activated complexes through the dividing surface, with the prefactor k_B T / h representing the universal frequency of crossing the barrier under classical assumptions. The quasi-equilibrium approximation allows \Delta G^\ddagger to incorporate both enthalpic and entropic contributions, derived from the partition functions as \Delta G^\ddagger = -RT \ln K^\ddagger. The Eyring equation connects directly to the Arrhenius form under typical conditions. Substituting \Delta G^\ddagger = \Delta H^\ddagger - T \Delta S^\ddagger, where \Delta H^\ddagger and \Delta S^\ddagger are the and , yields: k \approx \frac{k_B T}{h} \exp\left( \frac{\Delta S^\ddagger}{R} \right) \exp\left( -\frac{\Delta H^\ddagger}{RT} \right) At high temperatures, where the temperature dependence of the T prefactor is relatively weak compared to the exponential term, this approximates the Arrhenius equation k = A \exp(-E_a / RT), with the pre-exponential factor A \propto T \exp(\Delta S^\ddagger / R) and E_a \approx \Delta H^\ddagger + RT. This relation shows how TST interprets the empirical Arrhenius parameters: E_a primarily reflects the enthalpic barrier, while A captures entropic effects through \Delta S^\ddagger, providing a thermodynamic basis absent in simpler models. A key advantage of the TST framework is its ability to account for contributions to the A, which can vary significantly for reactions involving changes in or , unlike collision theory's more rigid frequency-based estimate. This makes TST particularly applicable to complex reactions in solution or enzymatic systems, where partition functions can be computed or estimated from spectroscopic data to predict rates.

Limitations and Advanced Considerations

Shortcomings of the Classical Model

The classical Arrhenius equation, which posits a linear relationship between the natural logarithm of the rate constant and the reciprocal , assumes a constant across the studied range. However, this assumption breaks down at low temperatures, where Arrhenius plots often exhibit upward curvature, leading to higher-than-expected reaction rates. This deviation arises because the model does not account for mechanisms like quantum tunneling, which allow reactions to occur more readily than predicted by classical over-barrier crossing. Such curvature has been observed in various gas-phase reactions, limiting the equation's reliability for extrapolations to cryogenic conditions. In condensed phases, particularly and highly viscous media, the Arrhenius equation fails to adequately describe reaction kinetics due to the dominance of processes. Here, reactant is severely restricted, making the rate-limiting factor rather than the barrier inherent to the Arrhenius form. For instance, in glassy or polymeric systems, diffusion-controlled rates lead to non-Arrhenius temperature dependencies, where the apparent varies significantly with changes or increases. This poor fit underscores the classical model's theoretical foundations in gas-phase assumptions (such as ), rendering it less applicable to heterogeneous or confined environments. Complex reaction mechanisms, especially multi-step processes, further expose the shortcomings of the classical model by producing overall rates that deviate from simple exponential behavior. In such cases, the rate-determining step may shift with temperature, resulting in composite Arrhenius parameters that mask underlying complexities. A notable manifestation is the kinetic compensation effect, where variations in E_a across related reactions are linearly correlated with changes in the \ln A, often artifactual due to experimental correlations or mechanistic overlaps rather than true physical compensation. This effect complicates the interpretation of Arrhenius plots for catalytic or enzymatic systems involving multiple pathways. Historical experiments in the early on highlighted these limitations, as deviations from predicted Arrhenius behavior were frequently reported in heterogeneous reactions. For example, studies on reactions over catalysts, such as those conducted around the and , showed non-linear dependencies attributed to surface coverage variations and multi-step surface processes, challenging the universality of the model. These early observations spurred refinements in kinetic , emphasizing the need for context-specific applications of the Arrhenius equation.

Quantum and Non-Arrhenius Behaviors

In , quantum tunneling allows reactants to bypass classical energy barriers, particularly at low temperatures, leading to deviations from the linear Arrhenius plot observed in classical theory. This effect is most pronounced in reactions involving light atoms like , where the wave-like nature of particles enables penetration through potential barriers rather than requiring thermal activation over them. A seminal quantum correction to the rate constant incorporates tunneling via the Wigner factor, which approximates the as \kappa_W \approx 1 + \frac{1}{[24](/page/24)} \left( \frac{[h](/page/H+) \nu^\ddagger}{k_B [T](/page/Temperature)} \right)^2, where \nu^\ddagger is the imaginary at the , [h](/page/H+) is Planck's constant, k_B is Boltzmann's constant, and [T](/page/Temperature) is . This correction arises from semiclassical considerations of imaginary contributions in the integral, enhancing the rate beyond classical predictions and resulting in upward curvature in Arrhenius plots (ln k vs. 1/T) at low temperatures, where tunneling dominates. Non-Arrhenius behaviors emerge in complex systems where cooperative effects or structural heterogeneity alter the temperature dependence of rates. In supercooled liquids and , the Vogel-Fulcher-Tammann (VFT) equation describes these deviations, with the rate constant given by k = A \exp\left( -\frac{B}{T - T_0} \right), where T_0 is a characteristic temperature below the (often ~50 K lower), B relates to fragility, and A is the . This form captures the divergence of relaxation times as T approaches T_0, reflecting cooperative dynamics in amorphous materials, unlike the simple inverse-temperature dependence in the Arrhenius equation. In biological systems like proteins, super-Arrhenius behavior—characterized by activation energies that increase with decreasing temperature—arises from coupled conformational changes and solvent interactions, leading to fragile dynamics approximated by VFT-like fits above the but transitioning to stronger (Arrhenius-like) behavior at lower or temperatures. Modern computational advances have enabled precise incorporation of quantum effects into Arrhenius parameters. Density functional theory (DFT) calculations determine activation energies E_a by optimizing transition state geometries and potential energy surfaces, often revealing quantum corrections that refine empirical fits for catalytic reactions; for instance, DFT-derived E_a values for hydrogen abstraction in hydrocarbon oxidation align closely with experimental Arrhenius parameters when tunneling is included. In the 2020s, machine learning (ML) models have accelerated fits of complex rate data to generalized Arrhenius forms, using neural networks to predict E_a and prefactors from molecular descriptors in catalysis, such as in ammonia synthesis over metal surfaces, where ML-guided optimization reduces computational costs while capturing non-Arrhenius trends in multi-element catalysts. Representative examples highlight these quantum and non-Arrhenius phenomena. In transfer reactions, such as malonaldehyde tautomerization or enzymatic C-H cleavages, tunneling dominates at cryogenic temperatures, yielding curved Arrhenius plots with effective E_a approaching zero, as evidenced by kinetic effects exceeding classical limits (KIE > 7 for at low T). Post-2000 advancements in quantum theory provide a path-integral framework for exact thermal rates, approximating the tunneling exponent via the action along the minimum-energy path, which has been applied to multidimensional systems like proton transfer in water clusters, yielding rates that deviate from Arrhenius by factors of 10-100 at T < 200 K compared to Wigner approximations. These developments address classical Arrhenius limitations at low temperatures, where barrier underestimation fails without quantum corrections.

References

  1. [1]
    [PDF] The origin and status of the Arrhenius equation
    The classic paper (3) which is the main focus of attention here, appeared in 1889 under the title, "Uber die Reaktionsgeschwindigkeit bei der Inuersion uon ...
  2. [2]
    18.5 Collision Theory and the Effect of Temperature on Reaction Rate
    The Arrhenius equation describes the relation between a reaction's rate constant and its activation energy, temperature, and dependence on collision orientation ...
  3. [3]
    Thermodynamics and Kinetics
    The Arrhenius equation does not tell you the rate of the reaction; it tells you the rate constant for an elementary step of the reaction. The variable Ea is ...
  4. [4]
    [PDF] Reaction Rates and Temperature; Arrhenius Theory
    Arhenius discovered that most reaction-rate data obeyed an equation based on three factors: (1) The number of collisions per unit time.Missing: definition | Show results with:definition
  5. [5]
    [PDF] Reaction Kinetics - Claire Vallance
    For an elementary reaction, both Ea and A have definite physical meanings; in particular, the activation energy may be interpreted as the energy difference.
  6. [6]
    Electrochemical Contributions: Svante August Arrhenius (1859–1927)
    Jul 15, 2024 · The Arrhenius equation provides relationship between a rate constant of a chemical reaction, k, and absolute temperature, T (Kelvin scale):.
  7. [7]
    On the Reaction Velocity of the Inversion of Cane Sugar by Acids
    In his notable book Studies in Chemical Dynamics van't Hoff gives a theoretically-based formulation of the influence of temperature on the rate of reaction.
  8. [8]
    The Nobel Prize in Chemistry 1903 - NobelPrize.org
    The Nobel Prize in Chemistry 1903 was awarded to Svante August Arrhenius in recognition of the extraordinary services he has rendered to the advancement of ...Missing: reason | Show results with:reason
  9. [9]
    Arrhenius Equation - an overview | ScienceDirect Topics
    The Arrhenius equation is a simple, but remarkably accurate, formula for the temperature dependence of the chemical reaction rate constant.
  10. [10]
    Arrhenius Equation - an overview | ScienceDirect Topics
    It is an exponential function of the parameter B, the activation temperature, and the parameter A is a multiplicative constant called the Arrhenius frequency ...
  11. [11]
    Modified Arrhenius Equation in Materials Science, Chemistry and ...
    According to the Arrhenius equation, the reaction rates at very low temperatures are very small but non-zero. The Arrhenius equation is also used in materials ...
  12. [12]
    molar gas constant - CODATA Value
    molar gas constant $R$ ; Numerical value, 8.314 462 618... J mol-1 K ; Standard uncertainty, (exact).
  13. [13]
    Etudes de dynamique chimique : Hoff, J. H. van't (Jacobus Henricus ...
    May 28, 2008 · Etudes de dynamique chimique. by: Hoff, J. H. van't (Jacobus Henricus van't), 1852-1911. Publication date: 1884 ... B/W PDF download · download 1 ...
  14. [14]
    The origin and status of the Arrhenius equation - ACS Publications
    Optimizing Arrhenius parameters for multi-step reactions via metaheuristic algorithms. Journal of Mathematical Chemistry 2025, 63 (4) , 1065-1092. https ...Missing: explanation | Show results with:explanation
  15. [15]
    <title>kinetics - csbsju
    The activation energy may then be extracted from a plot of ln k vs. 1/T, which should be linear. This plot is called an "Arrhenius plot". Problem MK1.1 ...Missing: construction | Show results with:construction
  16. [16]
    [PDF] A First Course on Kinetics and Reaction Engineering Example 4.6
    This example illustrates the use of an Arrhenius plot to determine the best values of a pre- exponential factor and an activation energy. Problem Statement.
  17. [17]
    [PDF] Determining the Energy of Activation Parameters from Dynamic ...
    A second common method of determining the energy of activation (Ea) is by performing an Arrhenius Plot. If one knows the exchange rate constant (kr) at several ...
  18. [18]
    Errors in the evaluation of Arrhenius and van't Hoff plots - PMC - NIH
    Errors in the numerical values of activation or normal enthalpies, entropies and free enthalpies calculated from Arrhenius or van't Hoff plots, respectively ...
  19. [19]
    [PDF] Assessment of Uncertainty in the Determination of Activation Energy ...
    This method uses linear regression analysis to determine the relationship between log heating rate and re- ciprocal temperature. The slope of the plot resulting.
  20. [20]
    Activation Energies and Beyond | The Journal of Physical Chemistry A
    Jun 10, 2019 · There are important limitations to obtaining the activation energy by an Arrhenius analysis. ... Convex Arrhenius Plots and their Interpretation.
  21. [21]
    New modified Arrhenius equation to describe the temperature ...
    Jul 15, 2025 · We propose a new modified Arrhenius equation that can correctly describe the temperature trends of liquid phase rate constants from room temperature up to the ...
  22. [22]
    [PDF] Kinetics and Mechanisms of the Gas-Phase Reactions of the NO ...
    Although empirical, a modified Arrhenius expression, k = CT²e-DIT, has been used in this evaluation, and a unit-weighted least-squares analysis of the data ...
  23. [23]
    Formulation and numerical analysis of diatomic molecular ...
    Jan 5, 2007 · Finally, we obtain a modified Arrhenius rate equation for the dissociation of A 2 from Eq. (50). In the calculation, the vibrational quantum ...
  24. [24]
    Bimodal Evans–Polanyi Relationships in Hydrogen Atom Transfer ...
    The Evans–Polanyi (EP) relationship (also known as the Bell–Evans–Polanyi relationship) correlates reaction rate constants (or other activation parameters) with ...
  25. [25]
    Bimodal Evans–Polanyi Relationships in Hydrogen Atom Transfer ...
    Jul 26, 2021 · The slope of the Evans-Polanyi plot is 0.38, which indicates that the reactions are mildly exothermic or almost thermoneutral. This finding is ...Introduction · Results · Discussion · Conclusions
  26. [26]
    Reliable Gas Phase Reaction Rates at Affordable Cost by Means of ...
    May 31, 2023 · If a reaction obeys the Arrhenius equation, then a plot of ln k versus 1T should produce a straight line, whose slope and intercept at the ...
  27. [27]
    Database for the kinetics of the gas-phase atmospheric reactions of ...
    Jun 3, 2020 · We present a digital, freely available, searchable, and evaluated compilation of rate coefficients for the gas-phase reactions of organic compounds.
  28. [28]
    [PDF] UPDATED 01/30/05
    Collision theory predicts the dependence of the rate constant on temperature of the type T 1/2 e - E / R T and allows actual prediction of the values for the ...
  29. [29]
    [PDF] Chemical kinetic analysis of the activation energy of diffusion ...
    where k is the rate constant of chemical reaction, A is the pre-exponential factor, ' ... Einstein-Stokes equation, so the diffusion coefficient of sulfate ions ...
  30. [30]
    Chapter 2 Diffusion-Controlled Reactions in Solution - ScienceDirect
    This chapter focuses on the concepts important in the understanding of diffusion-controlled reactions in solution.Missing: pre- | Show results with:pre-
  31. [31]
    On the Temperature Dependence of Enzyme-Catalyzed Rates
    Feb 16, 2016 · It follows that for ΔCp‡ < 0, the rate of an enzyme catalyzed reaction initially rises with temperature and then reaches an optimum temperature ...
  32. [32]
    Temperature-dependent aqueous OH kinetics of C2–C10 linear and ...
    Jan 17, 2024 · The calculated atmospheric lifetimes indicate that terpenoic alcohols and diols can react with the OH in aerosol, cloud, and fog water with ...
  33. [33]
    Microscopic interpretation of Arrhenius parameters - Oxford Academic
    This chapter reviews the microscopic interpretation of the pre-exponential factor and the activation energy in rate constant expressions of the Arrhenius form.Missing: offs | Show results with:offs
  34. [34]
    [PDF] Über die Reaktionsgeschwindigkeit bei der Inversion von ... - Zenodo
    Svante Arrhenius. Die Geschwindigkeit der Reaktionen, welche von Elektrolyten (Säuren und Basen) bewirkt werden, steht in einem sehr engen Zusammenhange.
  35. [35]
    Arrhenius - an overview | ScienceDirect Topics
    The Arrhenius relation is defined as an empirical formula that correlates the velocity of a chemical reaction with absolute temperature, based on the concept ...
  36. [36]
    [PDF] The development of the Arrhenius equation
    (i). An equation in which the pre-exponential factor also has a temperature dependence was first proposed in 1893 by Kooij. (29) ...Missing: validations | Show results with:validations
  37. [37]
    Das Gesetz der Reaktionsgeschwindigkeit und der Gleichgewichte ...
    References · p1_1 M. Trautz, Die Theorie der chemischen Reaktionssgeschwindigkeit und ein neues Grenzgesetz für ideale Gase: Die Additivität der inneren ...
  38. [38]
    Temperature Dependence of Rate Processes Beyond Arrhenius and ...
    The deviation from Arrhenius behavior in the temperature dependence of diffusion can now be evaluated from Equation (39) through the Stokes-Einstein equation( ...Transitivity Defined · The Transitivity Plot · Vogel--Fulcher--Tammann...
  39. [39]
    Uniform description of non-Arrhenius temperature dependence of ...
    Dec 18, 2013 · There will be deviations from simple Arrhenius equation in the sense that the apparent activation energy will decrease with decreasing ...Missing: shortcomings | Show results with:shortcomings
  40. [40]
    Relationship between diffusion, self-diffusion and viscosity
    May 1, 2009 · We investigate the experimental limits of validity of the Stokes–Einstein equation. There is an important difference between diffusion and ...
  41. [41]
    Assessing relationships between self-diffusion coefficient and ...
    Sep 10, 2019 · Our result shows that there is a crossover in the self-diffusion coefficient and viscosity from high-temperature Arrhenius behavior to low-temperature non- ...
  42. [42]
    Kinetic Compensation Effect - an overview | ScienceDirect Topics
    The kinetic compensation effect (KCE) refers to the observed correlation between the activation energy of a kinetic process and the pre-exponential factor ...
  43. [43]
    The mathematical origins of the kinetic compensation effect
    The kinetic compensation effect states that there is a linear relationship between Arrhenius parameters ln A and E for a family of related processes.Missing: equation | Show results with:equation
  44. [44]
    Density Functional Theory Rate Calculation of Hydrogen Abstraction ...
    The impact of quantum tunneling was investigated for a representative reaction, and it was shown that the Wigner correction provides a good approxn. to more ...
  45. [45]
    [PDF] Quantum mechanical transition state theory and tunneling corrections
    Mar 1, 1999 · TRANSITION STATE THEORY APPROXIMATION. The transition state theory of Hansen and Andersen uses the values of Cf f(0) and Cf f(0) ~where each ...
  46. [46]
    Vogel–Fulcher–Tammann plot for viscosity scaled with temperature ...
    The formula of VFT equation has a term of (T−T0)−1 instead of T−1 in the exponential function where T0 is the ideal glass transition temperature.
  47. [47]
    Aspects of Protein Reaction Dynamics: Deviations from Simple ...
    Such behavior is often referred to as a “super-Arrhenius” temperature dependence. ... Arrhenius type behavior in proteins will require additional studies.
  48. [48]
    Observation of fragile-to-strong dynamic crossover in protein ... - NIH
    At high temperatures, 〈τT〉 follows a super-Arrhenius behavior [called a “fragile” behavior (27)] describable approximately by a Vogel–Fulcher–Tammann (VFT) ...
  49. [49]
    Machine Learning in Computational Surface Science and Catalysis
    This work applies the recently published on-the-fly machine-learned force field techniques using a variant of the Gaussian approximation potentials.
  50. [50]
    Machine learning-assisted Ru-N bond regulation for ammonia ...
    Aug 21, 2025 · A series of Arrhenius plots were generated by plotting lnr against 1/T. The slope of the Arrhenius plot allowed for the calculation of the ...Results · Ammonia Synthesis... · Methods
  51. [51]
    Arrhenius curves of hydrogen transfers: tunnel effects, isotope ... - NIH
    In this paper, the Arrhenius curves of selected hydrogen-transfer reactions for which kinetic data are available in a large temperature range are reviewed.
  52. [52]
    Hydrogen Tunneling in Stoichiometric and Catalytic Reactions ...
    Nov 6, 2023 · Figure 3 shows an example representing the Arrhenius equation ... Wigner correction may be improved by the consideration of the shape ...
  53. [53]
    Perspective: Ring-polymer instanton theory - AIP Publishing
    Instanton theory provides a simple description of these processes in terms of dominant tunneling pathways. Using a ring-polymer discretization, an efficient ...Missing: post- | Show results with:post-
  54. [54]
    Quantum Tunneling Rates of Gas-Phase Reactions from On-the-Fly ...
    Oct 19, 2016 · The instanton method obtains approximate tunneling rates from the minimum-action path (known as the instanton) linking reactants to the products at a given ...Missing: post- | Show results with:post-