Fact-checked by Grok 2 weeks ago

Chemical kinetics

Chemical kinetics is the branch of concerned with the rates of chemical reactions, quantifying how quickly reactants are transformed into products through the measurement of changes in concentration over time. The rate of a reaction is defined as the change in the concentration of a reactant or product per unit time, often expressed as an instantaneous rate derived from the slope of a concentration-versus-time plot. Unlike , which predicts the spontaneity and equilibrium position of reactions, chemical kinetics elucidates the dynamic pathways, , and temporal aspects of reactions, providing essential insights into how reactions proceed at the molecular level. Several key factors influence reaction rates, including the concentrations of reactants, , the presence of catalysts, the physical state and surface area of reactants, and the inherent chemical nature of the substances involved. Increasing reactant concentrations generally accelerates the rate by enhancing molecular collision frequency, as described by . Temperature effects are captured by the Arrhenius equation, k = A e^{-E_a / RT}, where k is the rate constant, A is the , E_a is the , R is the , and T is the absolute ; higher temperatures exponentially increase k by providing more molecules with sufficient energy to overcome the activation barrier. Catalysts lower E_a without being consumed, thereby speeding up reactions, while factors like surface area in heterogeneous reactions promote more effective collisions. Central to chemical kinetics are rate laws, mathematical expressions relating the to reactant concentrations, such as rate = k [A]^m [B]^n, where the orders m and n are determined experimentally rather than from . These laws, along with reaction orders (zero, first, second, etc.), reveal information about the —the sequence of elementary steps—and the rate-determining step, which is the slowest and governs the overall rate. Understanding these elements is crucial for applications in fields like , where kinetics informs pollutant degradation, and , such as optimizing reaction conditions for efficiency.

Fundamental Concepts

Reaction Rates

In chemical kinetics, the quantifies the speed of a by measuring the change in concentration of a reactant or product over time. This fundamental concept allows scientists to describe how quickly reactants are consumed or products are formed during a . Reaction rates can be expressed as either or instantaneous values. The is determined by dividing the change in concentration of a (Δ[species]) by the corresponding change in time (Δt), providing an overall measure for a specific interval. In contrast, the instantaneous captures the rate at a precise moment and is given by the of concentration with respect to time: for a reactant, it is -d[reactant]/dt (the negative sign accounts for the decrease in concentration), and for a product, it is d[product]/dt. The units of reaction rate are typically moles per liter per second ( L⁻¹ s⁻¹, or M s⁻¹), reflecting concentration change per unit time. For reactions involving multiple species, must be considered to ensure consistent rate expressions; in a balanced equation like aA + bB → cC, the rate is defined such that - (1/a) d[A]/dt = - (1/b) d[B]/dt = (1/c) d[C]/dt, making the rate independent of the species chosen. Consider the simple decomposition reaction . If the concentration of A decreases from 0.10 M to 0.05 M over 20 seconds, the average rate of disappearance of A is (0.10 M - 0.05 M)/20 s = 0.0025 M s⁻¹, and since the is 1:1, this equals the average rate of appearance of B. The instantaneous rate at any time would be the of the concentration-time curve at that point.

Rate Laws and Orders

In chemical kinetics, a rate law is a mathematical expression that relates the of a to the concentrations of the reactants. For a general aA + bB \to products, the rate law takes the form \text{rate} = k [A]^m [B]^n, where k is the rate constant, [A] and [B] are the concentrations of the reactants, and m and n are the partial reaction orders with respect to A and B, respectively; the overall reaction order is the sum m + n. This empirical relationship must be determined experimentally, as it cannot be predicted solely from the of the balanced equation. Reaction orders can be zero, first, second, or even fractional, depending on the experimental data. A zero-order reaction has a rate independent of reactant concentrations, expressed as \text{[rate](/page/Rate)} = k; an example is the decomposition of on a surface, where the remains constant until the reactant is depleted. For a first-order reaction, the is proportional to the concentration of one reactant, \text{[rate](/page/Rate)} = k [A], as seen in the decomposition of N₂O₅ to NO₂ and O₂. Second-order reactions involve either two molecules of the same reactant (\text{[rate](/page/Rate)} = k [A]^2, e.g., the decomposition of to H₂ and I₂) or one molecule each of two different reactants (\text{[rate](/page/Rate)} = k [A][B], e.g., the reaction of CH₃Br with OH⁻). The (t_{1/2}), the time required for the reactant concentration to halve, is particularly characteristic for reactions: t_{1/2} = \frac{\ln 2}{k} \approx \frac{0.693}{k}, which is of initial concentration, unlike higher-order reactions. To determine the reaction order, experimental methods such as the initial rates approach or integrated rate laws are employed. In the initial rates method, the concentrations of reactants are varied systematically while measuring the initial rate (the instantaneous rate at t = 0) under conditions where changes are minimal; if doubling [A] doubles the rate while [B] is held constant, then m = 1. Integrated rate laws provide an alternative by analyzing concentration-time data through linear plots: for zero-order, a plot of [A] versus time t is linear with slope -k; for first-order, \ln[A] versus t is linear with slope -k; and for second-order, $1/[A] versus t is linear with slope k. The plot that yields a straight line identifies the order, and the slope gives the rate constant. It is important to distinguish reaction order, which is an experimental quantity derived from rate data, from molecularity, a theoretical concept describing the number of reactant molecules involved in an elementary step (unimolecular, bimolecular, etc.). While the order may coincide with molecularity for simple elementary reactions, complex mechanisms often result in orders that do not match stoichiometric coefficients or molecularity directly.

Molecularity and Reaction Mechanisms

In chemical kinetics, an elementary reaction is defined as a single-step process that occurs exactly as written, involving the direct collision or transformation of reactant molecules without intermediates./Kinetics/03:_Rate_Laws/3.02:_Reaction_Mechanisms/3.2.01:_Elementary_Reactions) The molecularity of an elementary reaction specifies the number of reactant molecules (or atoms) participating in that step: a unimolecular reaction involves one molecule decomposing or isomerizing, a bimolecular reaction entails two species colliding to form products, and a termolecular reaction requires three species to meet simultaneously, which is exceedingly rare due to the low probability of such collisions. For elementary steps, the rate law directly reflects the molecularity, with unimolecular reactions following first-order kinetics and bimolecular reactions exhibiting second-order kinetics. A describes the complete sequence of elementary steps that lead from reactants to products, providing a molecular-level interpretation of the observed overall reaction. In multi-step mechanisms, the (RDS) is the slowest elementary step, which governs the overall , as subsequent faster steps cannot proceed more quickly than the bottleneck./Kinetics/03:_Rate_Laws/3.02:_Reaction_Mechanisms/3.2.03:_Rate_Determining_Step) If the RDS occurs early in the sequence, its rate law often approximates the overall rate law; otherwise, the influence of prior steps must be accounted for through approximations. To obtain the overall from a proposed , especially when are involved, the steady-state (SSA) is commonly applied. The SSA assumes that the concentration of reactive remains nearly over time because their rates of formation and are balanced, leading to \frac{d[\ce{intermediate}]}{dt} \approx 0./Kinetics/03:_Rate_Laws/3.02:_Reaction_Mechanisms/3.2.06:_Steady_State_Approximation) This simplifies the equations for concentrations, allowing expression of the in terms of observable species. For instance, in a with an \ce{I}: \frac{d[\ce{I}]}{dt} = k_1 [\ce{A}] - k_2 [\ce{I}][\ce{B}] - k_3 [\ce{I}] = 0 Solving for [\ce{I}] yields [\ce{I}] = \frac{k_1 [\ce{A}]}{k_2 [\ce{B}] + k_3}, and the overall rate (e.g., -\frac{d[\ce{A}]}{dt} = k_1 [\ce{A}]) can then be substituted to derive a rate law like rate = \frac{k_1 k_3 [\ce{A}]}{k_2 [\ce{B}] + k_3}./Kinetics/03:_Rate_Laws/3.02:_Reaction_Mechanisms/3.2.06:_Steady_State_Approximation) A classic example is the for unimolecular gas-phase reactions, which reconciles the apparent kinetics of such processes with their dependence on collision partners. Proposed by Frederick Lindemann in 1926, the mechanism posits that the reactant \ce{A} first undergoes bimolecular activation by a collision with another \ce{M} (often \ce{A} itself) to form an energized \ce{A^*}, followed by deactivation or : \ce{A + M ⇌[k_1][k_{-1}] A^* + M} \ce{A^* ->[k_2] products} \ce{A^* + M ->[k_{-1}] A + M} Applying the SSA to [\ce{A^*}]: \frac{d[\ce{A^*}]}{dt} = k_1 [\ce{A}][\ce{M}] - k_{-1} [\ce{A^*}][\ce{M}] - k_2 [\ce{A^*}] = 0 Thus, [\ce{A^*}] = \frac{k_1 [\ce{A}][\ce{M}]}{k_{-1} [\ce{M}] + k_2} The overall rate is -\frac{d[\ce{A}]}{dt} = k_2 [\ce{A^*}] = \frac{k_1 k_2 [\ce{A}][\ce{M}]}{k_{-1} [\ce{M}] + k_2} = k [\ce{A}], where k = \frac{k_1 k_2 [\ce{M}]}{k_{-1} [\ce{M}] + k_2}./29:_Chemical_Kinetics_II-_Reaction_Mechanisms/29.06:_The_Lindemann_Mechanism) At (high [\ce{M}]), the rate becomes first-order in [\ce{A}] as deactivation dominates, while at low pressure, it shifts to second-order, reflecting the step as rate-limiting. This laid the foundation for later theories of unimolecular reactions.

Factors Influencing Reaction Rates

Concentration and Pressure Effects

In chemical kinetics, the effect of reactant concentrations on reaction rates is primarily described by rate laws, which follow the formulated by Cato Maximilian Guldberg and Peter Waage. This principle states that the rate of a reaction is proportional to the product of the concentrations of the reactants raised to powers corresponding to their stoichiometric coefficients in the elementary step, leading to a power-law dependence. For a general reaction aA + bB \rightarrow products, the rate law takes the form rate = k [A]^m [B]^n, where m and n are the reaction orders with respect to A and B, respectively, and k is the rate constant. The specific impact of concentration changes depends on the reaction order. In zero-order reactions, the rate is independent of reactant concentration (rate = k), often observed in where the catalyst surface becomes saturated, limiting the reaction to a constant pace regardless of excess reactant. For example, the of on wire at high temperatures exhibits zero-order , with the remaining constant even if concentration doubles. In contrast, reactions show a linear dependence ([rate](/page/Rate) = k [A]), so doubling the concentration of the reactant doubles the ; a classic case is the of to propene. Higher-order reactions amplify this effect: for second-order ([rate](/page/Rate) = k [A]^2), doubling the concentration quadruples the , as seen in the of ($2HI \rightarrow H_2 + I_2), where experimental data confirm the law [rate](/page/Rate) = k [HI]^2. The physical state of reactants also influences reaction rates. Reactions in gaseous or solution phases are typically faster due to greater molecular mobility and compared to those involving solids. For solid reactants, increasing the surface area—such as by grinding into a fine —enhances the rate by exposing more reactant molecules to potential collisions with other species. For reversible reactions, increasing the concentration of a reactant primarily accelerates the forward rate according to its specific rate law, while the reverse rate increases based on product concentrations, though the net rate toward equilibrium adjusts per . In gaseous reactions, pressure influences rates indirectly through its effect on concentrations, as partial pressures determine effective reactant densities via the ([A] = P_A / RT). Rate laws for gas-phase reactions are thus often expressed in terms of partial pressures (rate = k P_A^m P_B^n), where increasing total raises partial pressures proportionally for a fixed composition. This results in an increased reaction rate for orders greater than zero, as higher enhances molecular collisions; for instance, in a second-order gas-phase reaction like the decomposition of , elevating from 1 atm to 2 atm would quadruple the rate by doubling partial pressures. Zero-order gaseous reactions, however, remain unaffected by changes.

Temperature Dependence

The rate of a chemical reaction typically increases exponentially with temperature, as higher temperatures provide reactant molecules with greater kinetic energy, leading to more frequent and energetic collisions. A common rule of thumb for many reactions near room temperature is that the reaction rate approximately doubles for every 10°C rise in temperature, though this factor can vary depending on the specific reaction and its activation energy. This temperature dependence of the reaction rate constant k is quantitatively described by the , empirically derived by in 1889: k = A e^{-E_a / RT} where A is the representing the frequency of collisions with proper orientation, E_a is the (in J/mol), R is the (8.314 J/mol·K), and T is the absolute temperature (in K). The E_a physically represents the minimum energy barrier that reactants must overcome to form products, corresponding to the energy required to reach the in the reaction pathway. To determine E_a experimentally, rate constants are measured at several temperatures, and a plot of \ln k versus $1/T (known as an ) is constructed; the slope of the resulting straight line equals -E_a / R, allowing E_a to be calculated as E_a = -R \times \text{slope}. For example, consider the reaction between oxalate ion and permanganate ion, where rate constants were measured as follows:
Temperature (°C)k (M⁻¹·s⁻¹)
24$1.3 \times 10^{-3}
28$2.0 \times 10^{-3}
32$3.0 \times 10^{-3}
36$4.4 \times 10^{-3}
40$6.4 \times 10^{-3}
Plotting \ln k against $1/T (with T in K) yields a slope of -9253 K, so E_a = -(8.314 \, \text{J/mol·K}) \times (-9253 \, \text{K}) \approx 77 \, \text{kJ/mol}. This method provides a reliable way to quantify the barrier for diverse , with typical E_a values ranging from 20 to 100 kJ/mol for elementary steps.

Catalysts and Inhibitors

A is a substance that increases the rate of a without undergoing any net change itself, providing an alternative reaction pathway that allows the reaction to proceed more rapidly. achieve this by lowering the required for the reaction through stabilization of the or intermediate species along the new pathway. In contrast, an is a substance that decreases the by interfering with the , often by binding to reactants, intermediates, or , thereby impeding the formation of products. Catalysis is classified as homogeneous or heterogeneous based on the phase of the catalyst relative to the reactants. In homogeneous catalysis, the catalyst exists in the same phase as the reactants, typically in solution, allowing for molecular-level interactions that facilitate the reaction. A common example is acid-base catalysis, where acids or bases donate or accept protons to stabilize charged intermediates; for instance, in the hydrolysis of esters, hydronium ions act as catalysts by protonating the carbonyl oxygen, enhancing nucleophilic attack by water. Heterogeneous catalysis involves a catalyst in a different phase, usually a solid surface interacting with gaseous or liquid reactants, where adsorption of reactants onto the surface promotes bond breaking and forming. Platinum metal, for example, catalyzes the oxidation of hydrogen and oxygen to water in fuel cells by adsorbing both gases, lowering the energy barrier for their combination. Enzyme catalysis represents a specialized form of homogeneous catalysis in biological systems, where protein-based enzymes accelerate reactions essential for life processes. The kinetics of enzyme-catalyzed reactions are often described by the Michaelis-Menten model, which relates the reaction rate to substrate concentration through parameters such as the maximum rate V_{\max} = k_{\text{cat}} [E]_{\text{total}} and the Michaelis constant K_m, reflecting the enzyme-substrate affinity. Here, k_{\text{cat}} is the turnover frequency, quantifying the number of substrate molecules converted to product per enzyme molecule per second, typically ranging from 1 to 10^6 s^{-1} depending on the enzyme. The efficiency of catalysts is often evaluated by their turnover frequency, which measures catalytic activity under saturation conditions, providing insight into how effectively the alternative pathway enhances the rate compared to the uncatalyzed reaction. Inhibitors in chemical kinetics can be competitive or non-competitive, depending on their binding mode. Competitive inhibitors bind to the same as the or reactant, increasing the apparent by reducing the effective concentration available for reaction; for example, in , succinate acts as a competitive of by mimicking the substrate's structure. Non-competitive inhibitors bind to a different site, altering the catalyst's conformation or blocking a subsequent step without affecting substrate binding, thus decreasing the maximum ; in , poisons like compounds irreversibly adsorb onto metal surfaces, such as in automotive exhaust converters, deactivating sites and drastically reducing turnover frequency.

Theoretical Models

Collision Theory

Collision theory posits that chemical reactions occur when reactant molecules collide with each other, but only those collisions possessing sufficient to overcome the barrier and the correct orientation relative to one another result in product formation. This model, developed independently by Max Trautz in 1916 and William Lewis in 1918, applies primarily to bimolecular reactions in the gas phase and provides a framework linking reaction rates to molecular collision dynamics based on kinetic molecular theory. The derivation of the rate constant begins with the Z, which represents the number of collisions per unit volume per unit time between reactant molecules A and B. According to kinetic , for hard-sphere molecules, Z = \pi d^2 \sqrt{\frac{8 k_B T}{\pi \mu}} N_A [A][B], where d is the molecular , \mu is the , k_B is Boltzmann's constant, T is , N_A is Avogadro's number, and [A], [B] are concentrations. The of collisions with exceeding the E_a follows from the Maxwell-Boltzmann distribution, yielding e^{-E_a / RT}, where R is the . To account for the requirement of proper molecular orientation, a p (typically between 0 and 1) is introduced. Thus, the rate law for a bimolecular reaction is given by: \text{Rate} = p Z e^{-E_a / RT} [A][B] The rate constant k is therefore k = p Z' e^{-E_a / RT}, where Z' is the concentration-independent part of Z. This exponential dependence empirically aligns with the Arrhenius equation for temperature effects on rates. Key assumptions underlying collision theory include treating molecules as rigid hard spheres that interact only upon contact and assuming their kinetic energies follow the Maxwell-Boltzmann distribution, with no consideration of internal degrees of freedom or attractive/repulsive forces beyond contact. Despite its foundational insights, has limitations, particularly in predicting accurate values for low steric factors where p is much less than 1 due to complex molecular geometries, and it inadequately describes reactions involving intricate multi-step mechanisms rather than simple direct collisions.

Transition state theory (TST), also known as activated complex theory, posits that chemical reactions proceed through a high-energy intermediate configuration called the or , which exists momentarily at the of the along the . This represents a configuration where the system has partial bond breaking and forming, marking the highest energy barrier separating reactants from products; it is a maximum in energy with respect to the but a minimum in the orthogonal directions. The theory, developed in 1935 by Henry Eyring and independently by Meredith Gwynne Evans and , relies on to describe the distribution of molecular configurations and treats the as a dividing surface in . The cornerstone of TST is the Eyring equation, which expresses the rate constant k for an as k = \frac{k_B T}{h} e^{-\Delta G^\ddagger / RT}, where k_B is Boltzmann's constant, T is the absolute temperature, h is Planck's constant, R is the , and \Delta G^\ddagger is the standard of activation required to reach the from the reactants. This equation derives from the assumption that the activated complex vibrates along the with a frequency of k_B T / h, and the concentration of the complex is governed by a quasi-equilibrium constant related to \Delta G^\ddagger. The free energy term \Delta G^\ddagger = \Delta H^\ddagger - T \Delta S^\ddagger incorporates both the \Delta H^\ddagger and \Delta S^\ddagger of activation, providing a thermodynamic basis for the reaction barrier. TST connects directly to the empirical Arrhenius equation k = A e^{-E_a / RT}, where the pre-exponential factor A approximates \frac{k_B T}{h} e^{\Delta S^\ddagger / R}, linking the frequency factor to the entropy change upon forming the transition state; a positive \Delta S^\ddagger increases A by reflecting greater disorder in the activated complex, while negative values decrease it. This relationship reveals the molecular origins of the Arrhenius parameters, with E_a \approx \Delta H^\ddagger + RT for unimolecular reactions, allowing experimental activation energies to be interpreted in terms of transition state properties. In applications to solution-phase reactions and , TST employs the quasi-equilibrium assumption, positing that the remains in rapid with reactants despite the overall nonequilibrium nature of the , enabling rate predictions from profiles along the path. For reactions, multidimensional surfaces account for solvent effects via potentials of mean force, optimizing the dividing surface to minimize recrossings and incorporating coefficients for quantum tunneling. In enzymes, TST quantifies by comparing stabilization relative to the uncatalyzed , often using variational formulations to locate the optimal and predict rate enhancements from altered \Delta [G](/page/G)^\ddagger.

Experimental Methods

Measuring Reaction Rates

The rate of a chemical reaction is experimentally determined by monitoring the change in concentration of a reactant or product over time, typically expressed as the negative change in reactant concentration divided by the time interval for reactants or the positive change for products. This approach allows for the calculation of average rates, instantaneous rates via graphical slopes, or initial rates to minimize complications from changing concentrations. Selection of the monitoring depends on the reaction; often, the species with the most distinct change is chosen to ensure accurate and sensitive measurements. Several standard laboratory techniques are employed to track these concentration changes on timescales from seconds to hours. , particularly UV-visible absorption, is widely used for reactions involving colored species or those absorbing in the UV range; it relies on Beer's law, where is directly proportional to concentration, enabling continuous real-time monitoring with high precision. Conductometry measures variations in the electrical conductivity of the , which correlates with the number and of ions; this is particularly effective for ionic reactions in solution, such as acid-base neutralizations, where conductivity decreases as ions combine. measurements are applied to gas-producing or gas-consuming reactions in closed systems, where the change in pressure at constant volume and temperature is proportional to the change in gas moles via the , suitable for reactions like decompositions yielding gases. For reactions without easily monitorable physical changes, sampling followed by provides quantitative data; aliquots are withdrawn at timed intervals and analyzed by acid-base, , or to determine concentrations, often using back-titration to quench the reaction and account for ongoing progress during sampling. To isolate the effects of individual reactants and determine rate laws, specific experimental designs are essential. The method of initial rates involves conducting multiple experiments with varying initial concentrations of one reactant while keeping others constant, then measuring the initial from the earliest data points; the of rate changes reveals the with respect to that reactant, allowing construction of the overall rate law. Pseudo-order conditions simplify multi-reactant kinetics by using a large excess of one reactant, rendering its concentration effectively constant throughout the reaction; this transforms the rate law into a pseudo- form with respect to the limiting reactant, facilitating easier analysis via first-order integrated equations. Once data are collected, analysis focuses on determining the and through graphical methods. For integrated rate laws, plots of concentration versus time yield a straight line for zero- reactions, natural log of concentration versus time for , and reciprocal concentration versus time for second-; the linear plot identifies the , while the provides the ./Kinetics/05%3A_Experimental_Methods/5.07%3A_Using_Graphs_to_Determine_Integrated_Rate_Laws) Alternatively, the plots the logarithm of the against the logarithm of concentration for each reactant, where the equals the . considerations are critical, including instrumental (e.g., spectrophotometer ), systematic issues like temperature variations affecting , and random from sampling inconsistencies; these are minimized through replicates, statistical fitting (e.g., R² values), and control of experimental conditions to ensure reliable and determination. A representative example is the between and ions, where the rate is monitored by the time elapsed until a sudden color change from colorless to blue-black occurs due to iodine formation complexing with ; shorter times indicate faster rates, allowing initial rate calculations under varied concentrations to determine the second-order rate law. For reactions on faster timescales, these standard methods may require adaptations, but specialized techniques extend the approach further.

Techniques for Fast Reactions

Flow methods enable the study of reactions occurring on timescales by rapidly mixing reactants and observing the through spectroscopic detection. In continuous techniques, reactants are propelled through a mixing chamber into an observation tube at a constant , allowing the reaction progress to be monitored at discrete points along the tube corresponding to specific reaction times, typically from about 1 ms to seconds. This approach, pioneered by Hartridge and Roughton in 1923, consumes relatively large volumes of but provides a straightforward means to resolve fast without halting the . Stopped- methods improve upon this by using high-pressure syringes to mix reactants in milliseconds and abruptly stopping the against a barrier, creating a stable observation volume for time-resolved measurements down to approximately 1 ms. Developed by Britton Chance in the , stopped-flow has become a cornerstone for investigating rapid processes like enzyme-substrate binding and solvolysis s, often coupled with UV-visible or detection to track transient species. Relaxation methods perturb a system at or near and monitor its return to a new , suitable for reactions as fast as . The temperature-jump (T-jump) technique rapidly increases the solution temperature—typically by 1–10 °C using electrical discharge, heating, or irradiation—shifting the and revealing relaxation times around $10^{-6} s or faster for reversible processes. Introduced by in the , T-jump has elucidated proton transfer and conformational changes in biomolecules. Similarly, pressure-jump (P-jump) methods apply sudden pressure changes (up to 100 MPa) via mechanical or explosive drivers to alter equilibria in compressible systems, probing relaxation kinetics in the regime for reactions sensitive to volume changes, such as or binding. These techniques assume small perturbations and linear response, yielding rate constants from fits to the observed signals. Ultrafast techniques extend observations to nanoseconds and below, capturing transient intermediates in photochemical and photoinitiated reactions. involves a high-intensity (from a or ) to excite the sample, generating short-lived whose is monitored by a probe beam, resolving from microseconds to nanoseconds. Developed by Ronald Norrish and in the , this method revolutionized the study of radicals and excited states in gas and solution phases. advances this further using ultrashort (10–100 fs duration) in pump-probe configurations to initiate and interrogate bond breaking, , and electron transfer on attosecond-to-femtosecond scales ($10^{-15} s). Pioneered by in the 1980s, via these has visualized real-time , including vibrational coherence and transition-state evolution. These methods have been applied to benchmark fast reactions, such as the SN1 solvolysis of carbocations, where stopped-flow monitors the rapid capture of intermediates by nucleophiles in polar solvents, yielding second-order rate constants up to $10^9 M^{-1} s^{-1} for highly stabilized ions. In biological systems, femtosecond spectroscopy has resolved the initial in photosynthetic centers, revealing sub-100 migration from chlorophylls to the pair dimer in , followed by charge separation in 3–10 ps.

Connections to Thermodynamics

Kinetics Near Equilibrium

In reversible chemical reactions, the net reaction rate is given by the difference between the forward rate and the reverse rate, expressed as v_{\text{net}} = k_f [ \text{reactants} ] - k_r [ \text{products} ], where k_f and k_r are the forward and reverse rate constants, respectively. At , the forward and reverse rates become equal, resulting in a net rate of zero and constant concentrations of . This dynamic balance arises from the application of rate laws to both directions of the . The K for an elementary is directly related to these rate constants by K = k_f / k_r, which determines the position of based on the relative magnitudes of k_f and k_r. This relationship holds because at equilibrium, the ratio of product to reactant concentrations equals K, mirroring the ratio of the rate constants. Near equilibrium, small perturbations—such as a sudden change in or —displace the system slightly from balance, and the return to follows an characterized by the relaxation time \tau. For a simple reversible reaction like \ce{A ⇌ B}, the relaxation time is \tau = \frac{1}{k_f + k_r}, representing the for concentrations to approach their values after the . This \tau provides a measure of how quickly the system relaxes, with shorter times indicating faster . Relaxation methods exploit this to determine individual rate constants for reactions too fast for direct observation. A representative example is the isomerization between cis- and trans-1,2-difluoroethene (\ce{(cis-)CHF=CHF ⇌ (trans-)CHF=CHF}) at 623 K, where the system approaches an equilibrium mixture of approximately 66.7% cis and 33.3% trans isomers, with the ratio [trans]/[cis] = 0.5. The forward and reverse rates balance to yield K = 0.5, and perturbations reveal the relaxation dynamics governed by k_f and k_r. Catalysts like iodine vapor accelerate the approach to this equilibrium without altering K.

Free Energy Relationships

The rate of a is fundamentally governed by the activation , ΔG‡, which quantifies the barrier between reactants and the . Within , the expresses this relationship explicitly, stating that the rate constant k for a reaction is k = \frac{k_B T}{h} \exp\left(-\frac{\Delta G^\ddagger}{RT}\right), where k_B is the , h is Planck's constant, T is the absolute , and R is the . This formulation reveals that smaller values of ΔG‡ correspond to faster reactions, as the exponential term dominates the temperature dependence. The activation can be decomposed into enthalpic (ΔH‡) and entropic (ΔS‡) contributions via ΔG‡ = ΔH‡ - TΔS‡, providing a thermodynamic lens to interpret kinetic data. Hammond's postulate further connects reaction to characteristics by positing that the structure of the more closely resembles the adjacent stable species along the . Specifically, in reactions with a large negative overall change (ΔG_reaction < 0, exothermic), the occurs early and thus structurally and energetically resembles the reactants, leading to a lower sensitivity of the rate to the nature of the products. Conversely, for endothermic reactions (ΔG_reaction > 0), the late mirrors the products, making the barrier more dependent on product . This , grounded in the idea that states near extrema in the profile have minimal structural deviation from those extrema, aids in rationalizing how thermodynamic driving forces influence kinetic barriers without requiring full computational profiling. Linear relationships (LFER) extend these concepts by quantifying how perturbations, such as s, linearly affect changes and thus reaction rates. The exemplifies this approach for aromatic systems, relating the logarithm of the rate constant ratio to a parameter: \log \left( \frac{k}{k_0} \right) = \rho \sigma, where k_0 is the rate constant for the unsubstituted reference compound, \sigma measures the electronic influence of the (positive for electron-withdrawing, negative for donating), and \rho reflects the reaction's susceptibility to these effects (positive for rates accelerated by electron withdrawal). Developed initially for constants and extended to , this assumes additive impacts on the , proportional to their effects on a benchmark process. The slope \rho serves as a mechanistic probe: values greater than 1 indicate substantial charge development in the , while near-zero \rho suggests minimal electronic involvement. A representative application of LFER appears in solvolysis reactions, where solvent acts as the in ionizing substrates like substituted benzyl chlorides. For the aqueous ethanolysis of meta- and para-substituted benzyl tosylates at 50°C, a Hammett plot yields a linear with \rho \approx -4.0, demonstrating that electron-donating substituents (e.g., methoxy, \sigma = -0.27) lower ΔG‡ by stabilizing the developing positive charge in the rate-determining ionization step. Similarly, in the solvolysis of 2-arylethyl tosylates, k versus \sigma gives \rho = -3.6 in 80% at 75°C, highlighting how aryl substituents modulate the carbocation-like and enabling rate predictions for unstudied analogs. These not only validate SN1 mechanisms but also underscore the predictive power of LFER in linking thermodynamic substituent effects to kinetic outcomes.

Historical Development

Early Observations

The foundations of chemical kinetics emerged in the mid-19th century through empirical studies of reaction rates, beginning with Ludwig Wilhelmy's investigation of the -catalyzed inversion of in 1850. Wilhelmy quantitatively measured the rate of this reaction using , demonstrating that the rate was proportional to the concentrations of both and the catalyzing , marking the first derivation of a mathematical rate law for a chemical process. This work, published in Annalen der Physik und Chemie, established that reaction progress could be described by differential equations, challenging the prevailing view that reactions proceeded instantaneously to completion. In the 1860s, further empirical observations expanded on these ideas, with and Léon Péan de Saint-Gilles examining the of reversible reactions, particularly the formation and of esters such as from and . Their 1862 studies revealed that rates depended on the concentrations of reactants and that reactions reached rather than complete conversion, with the forward and reverse rates balancing at this point; they quantified how environmental factors like temperature influenced these rates in homogeneous systems. Concurrently, in 1864, Cato Maximilian Guldberg and Peter Waage formulated the based on analyses of compositions in various reactions, proposing that the rate of a chemical transformation is proportional to the product of the concentrations of the reacting species raised to powers reflecting their stoichiometric coefficients. This principle, initially presented in Norwegian in Forhandlinger: Videnskabs-Selskabet i Christiania, provided a general framework linking speeds to molecular concentrations and anticipated the connection between and . By 1884, advanced these observations in his book Études de Dynamique Chimique, introducing the principle of mobile equilibrium, which explicitly linked the rates of forward and reverse reactions to the . Van 't Hoff showed that the ratio of these rates equals the equilibrium constant, derived from experimental data on dissociation reactions, thereby integrating kinetic observations with thermodynamic principles and demonstrating how perturbations shift equilibria dynamically. Into the early , Max Bodenstein's experiments on the gas-phase reaction between hydrogen and iodine to form (2 ⇌ H₂ + I₂) provided key empirical validation of bimolecular rate laws. In his 1899 study, Bodenstein precisely measured reaction rates at various temperatures using a static reactor and optical analysis of iodine concentrations, confirming second-order for both formation and and establishing the reaction as a classic example of elementary bimolecular behavior without chain mechanisms. These observations solidified the empirical basis for understanding rate dependencies in gaseous systems.

Key Theoretical Advances

In 1889, introduced the concept of , proposing that the rate constant of a depends exponentially on through the relation k = A e^{-E_a / RT}, where E_a represents the minimum energy barrier that reactants must overcome to form products. This empirical relationship provided a foundational framework for understanding temperature effects on reaction rates, bridging experimental observations with a thermodynamic interpretation of reactivity. Building on this, Frederick Lindemann proposed in 1922 a for unimolecular reactions in the gas phase, suggesting that such reactions proceed via activation of a single through collisions with other molecules, followed by if sufficient energy is acquired. This two-step process—activation (A + M ⇌ A* + M) and (A* → products)—resolved the of how unimolecular reactions exhibit second-order at low pressures and at high pressures, laying the groundwork for statistical theories of reaction rates. Transition state theory (TST), developed independently in 1935 by Henry Eyring and by Meredith Gwynne Evans and , advanced the field by conceptualizing reactions as proceeding through a transient at the of the . Eyring's formulation derived the rate constant as k = \frac{k_B T}{h} e^{-\Delta G^\ddagger / RT}, where the free energy of \Delta G^\ddagger quantifies the energy difference between reactants and the , emphasizing equilibrium between reactants and the activated complex. Evans and Polanyi complemented this by applying semi-empirical surfaces to specific reactions, such as exchanges, enabling predictions of energies from bond dissociation data. TST, which succeeded earlier by incorporating quantum mechanical insights, remains a cornerstone for interpreting reaction mechanisms across diverse systems. Post-World War II developments refined unimolecular reaction theories, with the Rice-Ramsperger-Kassel-Marcus (RRKM) theory emerging as a statistical extension of Lindemann's ideas. Initially formulated by Oscar Rice and Herman Ramsperger in 1927 and Louis Kassel in 1928, it was significantly advanced in 1951 by Rudolph Marcus and , who incorporated detailed energy distribution among vibrational modes to predict microcanonical rate constants for energized molecules. RRKM assumes rapid intramolecular vibrational energy redistribution, allowing the to depend on the above the critical , and has been validated for a wide range of processes. Concurrently, Bigeleisen's work in the 1940s established the theoretical basis for kinetic isotope effects, deriving expressions for rate ratios between isotopologues based on differences and vibrational frequency shifts in the . His 1949 analysis showed that heavier s typically slow reactions due to reduced tunneling and altered force constants, providing a powerful tool for probing structures in kinetic studies.

Applications

Industrial Processes

Chemical kinetics plays a pivotal role in the design of industrial , where rate laws derived from experimental data guide the selection and sizing of reactor types to optimize production efficiency and yield. Batch reactors, which operate as closed systems where reactants are loaded at the start and the reaction proceeds over a defined time without material inflow or outflow, are commonly used for small-scale or variable product runs, allowing precise control over reaction progress based on integrated rate equations. In contrast, continuous stirred-tank (CSTRs) maintain steady-state conditions with continuous addition of reactants and removal of products, enabling high-throughput production; their design relies on rate laws to calculate residence times and volumes that achieve desired levels, often requiring larger sizes than batch systems for equivalent output due to incomplete mixing and back-mixing effects. In large-scale optimization, such as the Haber-Bosch process for synthesis, kinetic studies inform selection and operating conditions to maximize s under high-pressure, high-temperature environments. The process employs iron-based s promoted with and alumina to enhance the rate of and combination into , with kinetic models revealing that the depends on partial pressures and follows a involving surface adsorption steps; this has enabled industrial plants to achieve conversions of around 10-15% per pass while recycling unreacted gases for overall efficiencies exceeding 95%. In pharmaceutical development, chemical kinetics is essential for assessing drug stability and controlling reaction rates to ensure product and efficacy. Stability studies apply zero-order or rate laws to predict under accelerated conditions like and , allowing formulators to select excipients and packaging that minimize or oxidation rates; for instance, kinetic profiling helps establish expiration dates by extrapolating from Arrhenius plots of constants. Polymerization reactions in plastics exemplify kinetic control for tailoring material properties, where chain-growth mechanisms dictate molecular weight and yield. In the of or , initiators trigger , and rate laws incorporating and termination steps guide adjustments in concentration and temperature to achieve desired lengths and production rates, often in continuous processes to scale output efficiently.

Modeling and Numerical Simulation

Modeling chemical kinetics often requires solving systems of ordinary differential equations (ODEs) derived from rate laws, which describe the of concentrations in networks. These equations arise from postulated reaction mechanisms, serving as inputs to the models. For a simple second-order reaction A \to products, the rate law yields the ODE \frac{d[A]}{dt} = -k [A]^2, where k is the rate constant and [A] is the concentration of A; more complex mechanisms produce coupled systems of such equations. Numerical integration methods are essential for solving these ODEs, particularly when analytical solutions are unavailable. Explicit methods like the fourth-order Runge-Kutta (RK4) algorithm provide accurate approximations for non-stiff systems by evaluating the rate function multiple times per step, balancing computational efficiency and precision in simulating reaction progress. However, chemical kinetic systems frequently exhibit stiffness due to disparate timescales from fast and slow reactions, leading to instability in explicit integrators unless tiny steps are used. For stiff problems, implicit methods such as Gear's algorithm (also known as the backward differentiation formula, BDF) are preferred, as they solve nonlinear equations at each step to maintain stability over larger timesteps, making them suitable for large-scale kinetic simulations. In systems with low molecule numbers, such as cellular reactions, deterministic ODEs overlook fluctuations, necessitating stochastic approaches like simulations. The , an exact method, models the by randomly selecting reaction events based on their propensity functions and waiting times, accurately capturing noise in small-volume or low-concentration regimes. Software tools facilitate these simulations for practical applications. COPASI, a widely used platform for biochemical networks, integrates deterministic (e.g., Runge-Kutta) and (e.g., ) solvers to analyze complex pathways, supporting tasks like parameter estimation and analysis. In modeling, numerical methods combining detailed kinetics with transport equations enable predictions of ignition delays and flame speeds; for instance, adaptive integrators handle the stiff ODEs from mechanisms involving hundreds of species.

References

  1. [1]
    Chemical Kinetics - Purdue University
    Chemical kinetics studies the rate of reaction, which is the change in concentration of reactants per unit of time, as reactants transform into products.Chemical Kinetics · Instantaneous Rates of... · Different Ways of Expressing...
  2. [2]
    Kinetics - gchem
    Chemical kinetics is the study of reaction rates. Thermodynamics is a powerful tool in understanding chemical processes, but it is not the complete picture.
  3. [3]
    18.2 Factors Affecting Reaction Rates – Chemistry Fundamentals
    We can identify five factors that affect the rates of chemical reactions: the chemical nature of the reacting substances, the state of subdivision (one large ...
  4. [4]
    The Collision Model of Chemical Kinetics
    Equation 14.39 is known as the Arrhenius equation and summarizes the collision model of chemical kinetics, where T is the absolute temperature (in K) and R ...
  5. [5]
    [PDF] Chemical Kinetics The Study Of Reaction Rates In Solution
    Chemical kinetics plays a significant role in environmental chemistry, particularly in understanding pollutant degradation and reaction rates in natural waters.
  6. [6]
    Chemical Kinetics: Reaction Rates
    The term reaction rate refers to how fast the chemical reaction occurs. The rate of reaction, r, describes how fast the concentrations of reactants and products ...
  7. [7]
    [PDF] Chapter 14 Chemical Kinetics - MSU chemistry
    Chemical kinetics studies how fast a reaction happens, which is determined by the energy barrier (Ea) and is distinct from thermodynamics. Rate is a change in ...
  8. [8]
    Determining Reaction Rates
    We do not need the minus sign when calculating average rates from products. An instantaneous rate is the rate at some instant in time. An instantaneous rate is ...
  9. [9]
    [PDF] Chemical Kinetics
    (instantaneous rate) is the slope of the tangent to the curve. • Instantaneous rate is different from average rate, so when we refer to the rate of a reaction, ...
  10. [10]
    18.1 Chemical Reaction Rates – Chemistry Fundamentals
    The rate of a reaction can be expressed either in terms of the decrease in the amount of a reactant or the increase in the amount of a product per unit time.
  11. [11]
    CHEM 245 - Chemical kinetics: an introduction
    Chemical kinetics covers reaction rates, rate law equations, mechanisms, and reaction coordinate diagrams, including average and instantaneous rates.
  12. [12]
    Half Lives
    For a first order reaction A Arrow products , rate = k[A]:. The half life of a first order reaction is independent of concentration. For a second order reaction ...Equations for half lives · Determining a half life · Converting a half life to a rate...
  13. [13]
    Integrated Rate laws
    The graph that is linear indicates the order of the reaction with respect to A. Then, you can choose the correct rate equation.
  14. [14]
  15. [15]
    On the theory of unimolecular reactions - Journals
    Hollingsworth C, Seybold P, Kier L and Cheng C (2004) First‐order stochastic cellular automata simulations of the lindemann mechanism, International Journal of ...
  16. [16]
    Cato Guldberg and Peter Waage, the history of the Law of Mass ...
    According to the Law of Mass Action, the velocity of a chemical reaction depends on the concentrations of the reactants. At equilibrium the concentrations of ...
  17. [17]
    [PDF] EXPERIMENT 2 - Reaction Order
    In this experiment we will find the reaction order graphically by plotting the decrease in reactant concentration over time and use it to write a new rate law, ...
  18. [18]
    [PDF] CHAPTER 10 – CHEMICAL KINETICS - WebAssign
    A second order reaction that is second order in HI. c. 2 N2O5 → 4 NO2 + ... The decomposition of HI is second order in HI, so a plot of 1/[HI] versus ...
  19. [19]
    Equilibrium and LeChatelier's Principle - Background - Harper College
    A system is at equilibrium when the rates of the forward and reverse reactions are equal. If additional reactant is added the rate of the forward reaction ...
  20. [20]
    [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.
  21. [21]
    [PDF] Kinetics: The Rates of Reactions - Web Pages
    The Effect of Temperature. – For most reactions, the reaction rate increases almost exponentially with T (rate ~ doubles for every 10°C of T↑). – T affects ...
  22. [22]
    [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.
  23. [23]
    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 ...
  24. [24]
    Arrhenius Theory and Reaction Coordinates
    The Arrhenius theory and reaction coordinates relate to a particular perspective on chemical kinetics that imagines the path taken by the reactants as they are ...
  25. [25]
    Determining Activation Energy
    The activation energy can be determined by finding the rate constant of a reaction at several different temperatures.
  26. [26]
    Activation Energy
    "Two-Point Form" of the Arrhenius Equation​​ The activation energy can also be found algebraically by substituting two rate constants (k1, k2) and the two ...
  27. [27]
    [PDF] Enzyme Kinetics
    This is the Michaelis-Menten Rate Law where Km is the Michaelis constant. As was indicated before, Vmax is the maximum rate attainable.
  28. [28]
    18.7 Catalysis – Chemistry Fundamentals
    Catalysts can be homogenous (in the same phase as the reactants) or heterogeneous (a different phase than the reactants).
  29. [29]
    [PDF] Intro to Catalysis - Ready Lab - UT Southwestern, Dallas, Texas
    Inhibitor: A substance that decreases the TOF. Can be competitive. (aka reversible) or non-competitive (not reversible, usually covalent modification of ...
  30. [30]
    [PDF] Lecture 34: Kinetics: Catalysts - MIT OpenCourseWare
    Dec 8, 2014 · Homogeneous catalysts: reactants and catalysts are in the same phase. Example: chlorofluocarbons catalyze the depletion of O3 (all gas phase).
  31. [31]
    Chapter 7: Catalytic Mechanisms of Enzymes - Chemistry
    Acid-Base Catalysis is involved in any reaction mechanism that requires the transfer of a proton from one molecule to another. It is very common to see this ...
  32. [32]
    Effect of catalysts - Chemistry 302
    Catalysts in the same phase are called homogeneous catalysts, while those in different phases are called heterogeneous catalysts. For example, if we have Pt ...Missing: definition | Show results with:definition
  33. [33]
    [PDF] Introduction to Enzyme kinetics - Rose-Hulman
    The Michaelis-Menten equation has two parameters, Vmax and Km. Vmax = kcat[E]total, and therefore is a function of both the total enzyme concentration and of ...
  34. [34]
    [PDF] Computational analysis of a Mn-based electrocatalyst with primary ...
    May 6, 2022 · This lower energy pathway facilitates CO2 reduction hence the turnover frequency (TOF) of 901.4s-1 for 1-Mn determined experimentally, which ...
  35. [35]
  36. [36]
    11.10: Collision Theory
    ### Summary of Collision Theory (11.10: Collision Theory - Chemistry LibreTexts)
  37. [37]
    [PDF] Collision Theory of Reaction Rates and Its Limitations | Dalal Institute
    In 1916, a German chemist Max Trautz proposed a theory based on the collisions of reacting molecules to explain reaction kinetics.Missing: original | Show results with:original
  38. [38]
  39. [39]
  40. [40]
    Development of transition-state theory | The Journal of Physical ...
    This article is cited by 854 publications. Chonghe Zhang, Agustin Valles, Junyi Wang, Noah D. McMillion, Christopher C. Cummins, Robert J. Gilliard, ...
  41. [41]
  42. [42]
    12.1 Chemical Reaction Rates - Chemistry 2e | OpenStax
    Feb 14, 2019 · For reactions involving aqueous electrolytes, rates may be measured via changes in a solution's conductivity.
  43. [43]
  44. [44]
    Video: Measuring Reaction Rates - JoVE
    Sep 24, 2020 · Concentration changes can be measured by experimental techniques like polarimetry, spectroscopy, or pressure measurements. Polarimetry uses ...
  45. [45]
    Chemical Kinetics: Method of Initial Rates
    The initial rate, r0, is the rate of reaction at time t = 0. In the Method of Initial Rates, the differential rate law is determined by varying the initial ...
  46. [46]
    [PDF] Pseudo-first Order Kinetics – Determination of a rate law
    This strategy of flooding the reaction mixture with one reactant allows us to isolate the effect of the other reactant on the rate of change or reaction rate.
  47. [47]
    Experimental errors in kinetic tests and its influence on the precision ...
    This work analyzes how the experimental errors affect kinetic studies based on catalytic tests when oscillations of the input variables are the main sources of ...
  48. [48]
    Iodine clock reaction demonstration method - RSC Education
    Use this iodine clock reaction demonstration to introduce your students to rates of reaction and kinetics. Includes kit list and safety instructions.
  49. [49]
    A method of measuring the velocity of very rapid chemical reactions
    In devising methods for determining the velocity of any chemical reaction there are two experimental problems which invariably arise.
  50. [50]
    The Stopped-flow Method and Chemical Intermediates in Enzyme ...
    This Personal perspective recites the steps taken to minimize the volumes expended in rapid flow studies, together with the improvements in electronic ...Missing: seminal | Show results with:seminal
  51. [51]
    [PDF] Manfred Eigen - Nobel Lecture
    Periodic relaxation methods have added greatly to our knowledge of the kinetics of chemical reactions, particularly in the time range from micro- seconds to ...
  52. [52]
    Pressure Jump Relaxation Method for the Kinetic Study of Fast ...
    Jul 13, 2023 · In the pressure jump technique, the pressure on a reaction system in solution is changed rapidly and the position of equilibrium thereby shifts adiabatically ...
  53. [53]
    [PDF] R. G. Norrish - Nobel Lecture
    The opportunity for the application of the methods of flash photolysis to chemical kinetics, not only in the gas phase, but also to the study of photo- chemical ...
  54. [54]
    [PDF] Ahmed H. Zewail - Nobel Lecture
    Timeline of some events in the history of measurements of time, from yearly calendars to the femtosecond regime (see text). ear's ability to recognize a tone.
  55. [55]
    Kinetics of the Reactions of Halide Anions with Carbocations
    Stopped-flow and laser flash techniques have been employed to determine absolute rate constants for the reactions of benzhydrylium ions with solvents (k2 in ...
  56. [56]
    Femtosecond spectroscopy of excitation energy transfer and initial ...
    The transfer of excitation energy between H, B, and P occurs in slightly less than 100 fsec and leads to the ultrafast formation of an excited state of P. This ...
  57. [57]
    2.1.4: Relaxation Methods - Chemistry LibreTexts
    Feb 12, 2023 · A temperature jump of as much as 10 oC for example can be produced in an aqueous solution by an electric discharge that lasts no more than 1 μs.
  58. [58]
    15.2: The Equilibrium Constant (K) - Chemistry LibreTexts
    Aug 14, 2020 · The equilibrium constant is equal to the rate constant for the forward reaction divided by the rate constant for the reverse reaction.Developing an Equilibrium... · Example 15 . 2 . 1 : Equilibrium... · Exercise 15 . 2 . 1
  59. [59]
    13.2: The Equilibrium State - Chemistry LibreTexts
    Jul 17, 2023 · A simple and instructive example of a chemical equilibrium is provided by the interconversion of the cis and trans isomers of difluoroethene.
  60. [60]
    A Brief Introduction to the History of Chemical Kinetics - IntechOpen
    The following text provides brief historical background to chemical kinetics, lays the foundation of transition state theory (TST), and reaction thermodynamics.
  61. [61]
    Ludwig Ferdinand Wilhelmy (1812-1864) - Le Moyne
    A necessary conclusion from the equation, moreover, is that when T = °, can become zero, hence the inversion of the sugar, strictly speaking, is never ended, ...Missing: first rate
  62. [62]
    [PDF] Marcelin Berthelot and Léon Péan de Saint-Gilles ... - Le Moyne
    These conditions and their effects are perhaps most impressive in the reactions of the ethers, especially their formations and their decompositions. Ethers ...
  63. [63]
    [PDF] Studies concerning affinity
    This article presents an English translation of "Studies Con- cerning Affinity" (I), Guldherg and Waage's first presentation, in 1864, of the law of mass action ...
  64. [64]
    Jacobus H. van 't Hoff – Biographical - NobelPrize.org
    This principle of mobile equilibrium was subsequently (1885) put in a general form by Le Chatelier, who extended the principle to include compensation, by ...
  65. [65]
    Bodenstein kinetics - Le Moyne
    Max Bodenstein made detailed measurements of the thermodynamics and kinetics of the gas-phase decomposition of hydrogen iodide 2 HI < --- > H 2 + I 2.Missing: key experiment
  66. [66]
    Discussion on “the radiation theory of chemical action”
    Discussion on “the radiation theory of chemical action”. F. A. Lindemann, S. Arrhenius, I. Langmuir, N. R. Dhar, J. Perrin and W. C. McC. Lewis, Trans.
  67. [67]
    Reactor Design & Selection for Pharma Continuous Manufacturing
    May 18, 2021 · There are three main categories of ideal reactors: batch reactors, plug flow reactors (PFR, also known as tubular reactors), and continuous- ...
  68. [68]
    Types of Reactors - Chemical Engineering World
    Jul 8, 2020 · The biggest advantage of using a C.S.T.R in industries is that it can produce a large amount of products and being a continuous steady state ...
  69. [69]
    Reaction Mechanisms, Kinetics, and Improved Catalysts for ...
    Apr 6, 2022 · The Haber–Bosch (HB) process is the primary chemical synthesis technique for industrial production of ammonia (NH3) for manufacturing ...
  70. [70]
    Reaction kinetics for ammonia synthesis using ruthenium and iron ...
    In this work, the reaction kinetics of ruthenium- and iron-based catalysts are determined experimentally at pressures between 10 to 80 bar and at temperatures ...
  71. [71]
    Systematic strategies for degradation kinetic study of pharmaceuticals
    Mar 16, 2022 · The term stability primarily refers to a drug's ability to maintain its chemical nature, strength, and efficacy within pharmacopoeial limits.
  72. [72]
    Chemical Kinetics - Pharmlabs
    where, The Arrhenius equation can be rewritten as: Again, an equation of the form y = mx + b is generated, indicating that a semi-log plot of Kapp vs. the ...<|control11|><|separator|>
  73. [73]
    Introduction to Polymerization and Depolymerization
    Apr 1, 2025 · Reaction Kinetics: Polymerization reactions can be complex, with varying rates of polymer growth and competing side reactions. Understanding and ...
  74. [74]
    Frontal Polymerization of Epoxy Resins: Kinetic Modeling, Rate ...
    Frontal polymerization (FP) technology has attracted significant attention as an efficient, low-energy curing method for thermosetting resins.2. Experiment And Methods · 3. Results And Discussion · 3.2. Curing Kinetics Of...
  75. [75]
    [PDF] 1.2 Boxes and arrows to differential equations
    Problem 14: Imagine that you do two experiments in chemical kinetics. In one ... Then the kinetic equations become d[S] dt. = −k+[S][E] + k−. [ES].
  76. [76]
    An Introduction to the Use of Numerical Methods in Chemical Kinetics
    The principles of numerical integration are briefly presented. The Runge–Kutta algorithm is used to explore two simple mechanisms as well as an auto-catalyzed ...
  77. [77]
    [PDF] Chemical Rates - Duke Mathematics Department
    By solving the differential equations, we can predict the concentrations at future times. Part 1: A simple first-order reaction. Consider a chemical reaction in ...<|separator|>
  78. [78]
    [PDF] Gear's method - Introduction to Chemical Engineering Computing
    In Gear's method, the nonlinear equations are solved. If they cannot be solved, then the step size is reduced and you try again. Keep reducing the step size ...
  79. [79]
    Exact stochastic simulation of coupled chemical reactions
    Gillespie-Driven kinetic Monte Carlo Algorithms to Model Events for Bulk or Solution (Bio)Chemical Systems Containing Elemental and Distributed Species.
  80. [80]
    COPASI—a COmplex PAthway SImulator - Oxford Academic
    Here, we present COPASI, a platform-independent and user-friendly biochemical simulator that offers several unique features.
  81. [81]
    Numerical Modeling of Chemical Kinetics, Spray Dynamics, and ...
    In this study, we use numerical simulations to study several aspects of jet fuel combustion, including laminar flames, ignition, and turbulent spray flames.