Fact-checked by Grok 2 weeks ago

Reaction mechanism

In , a reaction mechanism is the detailed, step-by-step sequence of elementary reactions that describes how reactants are transformed into products during a . This pathway reveals the formed and consumed along the way, as well as the movement of electrons and atoms involved in bond breaking and formation. Unlike a balanced , which only summarizes the overall change, the mechanism provides insight into the and governing the reaction's progression. Chemical reactions typically proceed through one or more elementary steps, each representing a single molecular event where bonds are formed or broken in a concerted manner. These steps are classified by their molecularity: unimolecular steps involve one decomposing or rearranging; bimolecular steps entail two colliding and reacting; and termolecular steps, which are rare due to the low probability of three meeting simultaneously, involve three. The overall reaction is the net result of these summed elementary steps, often requiring the identification of transient intermediates—short-lived species produced in one step and consumed in the next. A critical feature of many mechanisms is the rate-determining step, the slowest elementary step that controls the overall , analogous to a in a . For instance, in the reaction between and below 225 °C, the rate law aligns with the slow bimolecular step NO₂ + NO₂ → NO₃ + NO, followed by faster steps leading to products. Above this temperature, a different mechanism emerges, consisting of a single bimolecular step between NO₂ and CO (NO₂ + CO → NO + CO₂) as the rate-determining step, with rate law rate = k[NO₂][CO]. Such variations highlight how and conditions can alter mechanistic pathways. Mechanisms are elucidated experimentally by comparing observed rate laws—derived from kinetic studies—with those predicted by proposed sequences of elementary steps. The rate law for an elementary step directly corresponds to its : first-order for unimolecular (rate = k[reactant]) and second-order for bimolecular (rate = k[reactant₁][reactant₂]). Additional evidence comes from , trapping of intermediates, stereochemical analysis, and computational modeling, ensuring the proposed mechanism accounts for all experimental data. In , reaction mechanisms are particularly emphasized for understanding transformations at functional groups, often depicted using curved arrow notation (arrow pushing) to track electron flow between nucleophiles, electrophiles, and intermediates like carbocations or carbanions. Common classes include addition reactions, where π-bonds are converted to σ-bonds; elimination reactions, forming π-bonds by removing atoms or groups; substitution reactions, exchanging one group for another; and rearrangement reactions, yielding isomers without net bond count change. Factors such as solvent polarity, catalysts, and electronic effects influence these pathways, enabling prediction of product and .

Fundamentals

Definition and overview

A reaction mechanism describes the detailed, step-by-step sequence of elementary reactions by which starting materials, known as reactants, are transformed into final products, specifying the bonds that break and form along the pathway. This contrasts with the overall reaction, which simply balances the net without revealing the intermediate processes involved. For instance, in reactions, the proceeds via a single concerted step with backside attack by the , leading to inversion of , whereas the involves two steps: first, departure of the to form a , followed by attack, often resulting in . The concept of reaction mechanisms emerged in the late 19th century through foundational work in . In 1884, published "Studies of Chemical Dynamics," introducing differential rate equations and the temperature dependence of equilibrium constants, laying the groundwork for understanding reaction paths. Five years later, in 1889, extended this by proposing the , which incorporated an barrier to explain rate variations with temperature, marking a key advance in conceptualizing energy requirements for reactions. The modern framework evolved significantly after the with the advent of , enabling atomic-level descriptions of bond changes. This culminated in 1935 with the development of by Henry Eyring, , and Meredith Gwynne Evans, which modeled reactions as passing through high-energy transition states on potential energy surfaces derived from quantum principles. Understanding mechanisms is crucial in chemistry, as it elucidates patterns of reactivity and selectivity, allowing prediction of product distributions and design of synthetic routes. Mechanisms often involve fleeting intermediates, and serves as a primary tool for elucidating them.

Elementary reactions

An is defined as a chemical process that occurs in a single step on a molecular scale, with no detectable or postulated reaction , such that the of the reaction corresponds directly to its stoichiometric coefficients. This means that the reactants transform directly into products through a single kinetic event, without any persistent intermediate . Elementary reactions are inherently concerted, involving the simultaneous breaking and formation of bonds in a single transition state, in contrast to stepwise mechanisms that proceed through sequential steps with intermediates. They are classified by molecularity—the number of reactant molecules involved in the collision leading to products. Unimolecular elementary reactions involve a single molecule undergoing decomposition or rearrangement, such as the thermal ring opening of cyclobutane: \ce{C4H8(g) -> 2 C2H4(g)}, where the rate law is \text{rate} = k [\ce{C4H8}]. Bimolecular reactions, the most common type, require collision between two molecules, as in the gas-phase reaction between hydrogen and iodine: \ce{H2(g) + I2(g) -> 2 HI(g)}, with the rate law \text{rate} = k [\ce{H2}][\ce{I2}]. Termolecular reactions, involving three molecules colliding simultaneously, are rare due to the low probability of such events. In practice, purely elementary reactions are uncommon in complex chemical systems; most observed reactions are composites of multiple elementary steps, where the overall process is described by a built from these fundamental units. Within an elementary step, the serves as the fleeting high-energy configuration, and no stable intermediates are formed, distinguishing it from multi-step pathways.

Key Components

Reaction intermediates

In mechanisms, intermediates are transient that form during an elementary step and are subsequently consumed in a later step, serving as bridges between reactants and products in multi-step processes. These do not appear in the overall balanced equation of the reaction but are essential for elucidating the pathway and of the transformation./12%3A_Kinetics/12.06%3A_Reaction_Mechanisms) Reaction intermediates encompass a variety of types, including free radicals, which possess an and participate in reactions; ionic species such as carbocations (positively charged carbon centers) and carbanions; and excited states, where molecules occupy higher-energy electronic configurations following photon absorption. Free radicals, for instance, are key in propagation steps of radical mechanisms, enabling efficient transformations in and processes. Carbocations exemplify ionic intermediates, often stabilized by adjacent alkyl groups or molecules, while excited states drive photochemical reactions by facilitating bond breaking or rearrangement.Complete_and_Semesters_I_and_II/Map%253A_Organic_Chemistry(Wade)/05%253A_An_Introduction_to_Organic_Reactions_using_Free_Radical_Halogenation_of_Alkanes/5.07%253A_Reactive_Intermediates_-_Carbocations) A classic example of a intermediate occurs in solvolysis reactions proceeding via the SN1 mechanism, where an alkyl halide in a polar undergoes heterolytic cleavage to generate a planar , which then reacts with the as to form the substitution product. In biochemical contexts, enzyme-substrate complexes represent relatively stable intermediates, wherein the enzyme's binds the to form a transient complex that lowers the for catalysis, as seen in Michaelis-Menten kinetics. These examples illustrate how intermediates dictate reaction , , and efficiency./09%3A_Reactions_of_Alkyl_Halides-_Nucleophilic_Substitutions_and_Eliminations/9.06%3A_Characteristics_of_the_SN1_Reaction) The lifetimes of reaction intermediates span a broad range, typically from femtoseconds for highly reactive excited states or radicals to seconds for more stable species like enzyme-substrate complexes, influenced by factors such as , , and . Short-lived radicals in chain reactions, for example, persist only long enough to propagate the chain, often on the order of nanoseconds, while carbocations in solvolysis can last microseconds before capture by nucleophiles. Detecting reaction intermediates poses significant challenges due to their inherent instability and low steady-state concentrations, often necessitating indirect methods for . Kinetic studies provide through rate laws and isotope effects that imply the presence and reactivity of intermediates, while trapping techniques use to capture and identify them, converting short-lived species into observable products. Spectroscopic methods, such as time-resolved UV-Vis or , enable direct observation under ultrafast conditions, though success depends on the intermediate's lifetime and environment.

Transition states

In transition state theory, the (TS) represents the highest-energy configuration of atoms along the in an , occurring at the apex of the activation barrier where bonds are simultaneously breaking and forming. This hypothetical arrangement is a fleeting, unstable species that separates reactants from products, serving as the "point of no return" in the reaction pathway. Formulated by Henry Eyring in 1935, the concept posits that the rate of a reaction is determined by the equilibrium concentration of molecules reaching this . The is characterized by partial bonds, with bond lengths and angles intermediate between those of reactants and products, rendering it non-isolable under normal conditions. Its lifetime is extremely short, on the order of 10^{-13} seconds—comparable to a single —preventing observation or isolation. Conventionally denoted with a double dagger symbol (e.g., [AB]‡), the TS embodies the kinetic bottleneck of the reaction, where the system possesses sufficient energy to surmount the barrier but has not yet committed to product formation. The energy relationship of the is central to understanding reaction rates: the (E_a) is defined as the difference between the energy of the reactants and the TS, quantifying the barrier height that must be overcome. According to the Hammond postulate, proposed by George S. Hammond in 1955, the structure of the TS resembles the nearer stable species on the energy profile—for exothermic reactions, an "early" TS akin to reactants; for endothermic ones, a "late" TS resembling products. This principle aids in predicting TS geometry based on reaction . Representative examples illustrate TS diversity. In bimolecular nucleophilic substitution (S_N2) reactions, the TS features a linear arrangement of the nucleophile, central carbon, and leaving group in a trigonal bipyramidal geometry, with partial bonds to both the incoming nucleophile and departing group. In contrast, pericyclic reactions like the Diels-Alder cycloaddition involve a cyclic TS, where multiple bonds form concertedly through a delocalized, aromatic-like transition structure. These configurations highlight how TS geometry dictates stereochemistry and selectivity in mechanistic pathways.

Kinetic Principles

Chemical kinetics

Chemical kinetics plays a crucial role in elucidating reaction mechanisms by analyzing how reaction rates depend on reactant concentrations and temperature, thereby identifying the rate-determining step and inferring the composition of the transition state. The rate-determining step is the slowest elementary step in a multi-step mechanism, which governs the overall reaction rate, as subsequent faster steps cannot accelerate the process. Through kinetic studies, the activation energy derived from temperature dependence reveals the energy barrier associated with the transition state of this step, providing insights into the structural changes during the reaction. The rate law for a reaction expresses the reaction rate as a function of reactant concentrations, typically in the differential form: \text{rate} = k [\text{reactants}]^{\text{order}} where k is the rate constant and the order is the sum of the exponents for each reactant. For overall reactions, integrated rate laws describe concentration changes over time: for zero-order reactions, [\text{A}] = [\text{A}]_0 - kt; for first-order, \ln[\text{A}] = \ln[\text{A}]_0 - kt; and for second-order, \frac{1}{[\text{A}]} = \frac{1}{[\text{A}]_0} + kt. These forms allow determination of the order by plotting data and checking linearity, aiding in mechanism proposal. Reaction orders are determined experimentally using the initial rates method, where are measured at the start of reactions with varying initial concentrations while keeping others constant; the order with respect to a reactant is the exponent that matches the observed rate change factor. Integer orders often correspond to the of elementary steps, where is the number of reactant molecules involved, but observed orders can be fractional, signaling complex mechanisms with intermediates or pre-equilibria. The rate constant k follows the : k = A e^{-E_a / RT} where A is the pre-exponential factor representing collision frequency and orientation, and E_a is the activation energy; plotting \ln k versus $1/T yields a straight line with slope -E_a / R. A classic example is the for unimolecular gas-phase reactions, where the apparent order transitions from second-order at low (rate limited by bimolecular activation A + M → A* + M) to at high (rate limited by unimolecular decomposition A* → products), demonstrating dependence due to the competition between and deactivation.

Molecularity

Molecularity refers to the number of reactant molecules or atoms that participate in a single elementary step of a reaction mechanism, determined by the of that step. It classifies elementary reactions as unimolecular (involving one ), bimolecular (involving two ), or termolecular (involving three ). Unimolecular reactions typically involve the decomposition or of a single , while bimolecular reactions entail collisions between two . Termolecular reactions, requiring simultaneous collision of three , are exceedingly rare due to the low probability of such events and the associated unfavorable change from reduced translational freedom. In a reaction mechanism, each elementary step possesses a well-defined that directly corresponds to the number of reactant entities in its balanced equation, providing insight into the collision requirements for that step. However, the overall reaction, which comprises multiple elementary steps, does not have a singular , as it cannot be reduced to a single collision event. This concept applies exclusively to elementary steps, aiding in the construction and validation of proposed mechanisms. Molecularity differs fundamentally from the of a : is a theoretical property fixed by the and always an (1, 2, or rarely 3), whereas is an experimental parameter derived from the , which may be fractional or zero and reflects the overall rather than individual steps. For elementary reactions, the often predicts the , but complex mechanisms can yield orders that deviate from any single step's . Examples illustrate these classifications effectively. A classic unimolecular reaction is the thermal decomposition of cyclobutane to , where a single C₄H₈ rearranges without additional collision partners: C₄H₈ → 2 C₂H₄. Bimolecular processes are more common, such as the radical recombination in the formation of from atoms: H + HI → H₂ + I, involving two in a single collision. Termolecular examples, though scarce, include the gas-phase reaction 2 NO + O₂ → 2 NO₂, where three must collide simultaneously, highlighting the entropic disadvantage that makes such steps infrequent in mechanisms.

Experimental Methods

Kinetic studies

Kinetic studies provide essential experimental data to elucidate reaction mechanisms by measuring rates under controlled conditions and analyzing dependencies on concentrations, , and isotopic substitutions. These approaches rely on the foundational principles of to infer the sequence of elementary steps, often identifying the rate-determining step or the involvement of intermediates. Common methods include the initial rates approach, where the is measured at the outset when product concentrations are negligible and intermediates have not accumulated significantly, allowing of the rate law's dependence on reactant concentrations. This technique assumes pseudo-first-order conditions for all but one reactant, facilitating determination of reaction orders. The method complements this by using a large excess of all but one reactant to maintain constant concentrations, simplifying the expression to depend solely on the varied ; for instance, in a reaction A + B → products, excess B makes the rate proportional to [A] alone. Relaxation techniques, such as temperature-jump methods, perturb systems near to study fast . In a temperature jump, a rapid increase (typically 1–10 K via or ) shifts the , and the system's relaxation back to a new is monitored, often via spectroscopic changes; the relaxation time τ relates to rate constants by 1/τ = k_f + k_r, where k_f and k_r are forward and reverse rates. These methods access timescales from microseconds to seconds, revealing elementary steps in complex mechanisms. For mechanisms involving short-lived intermediates, the steady-state approximation simplifies rate expressions by assuming the intermediate concentration [I] changes negligibly after an initial transient, such that d[I]/dt ≈ 0. This leads to algebraic expressions for [I] in terms of reactants and products, enabling derivation of observable rate laws; originally applied to chain reactions, it has become central to analyzing multi-step processes. Pre-steady-state probes the initial phases before steady-state conditions are reached, capturing fast elementary steps. Burst phases occur when product formation is rapid in the first turnover but slows due to a rate-limiting step, observable as an initial "burst" in progress curves; single-turnover experiments, using excess over , isolate individual catalytic cycles to measure rate constants for or . These techniques, often combined with rapid mixing or stopped-flow, distinguish pre-steady-state events from overall rates. Kinetic isotope effects (KIEs) further probe structures. Primary KIEs arise when isotopic substitution (e.g., H to D) affects the bond-breaking step in the rate-determining , typically yielding k_H/k_D > 1 (up to ~7 for H/D at ) due to differences. Secondary KIEs, from substitution at non-reacting positions, reflect steric or hyperconjugative changes in the , with smaller magnitudes (k_H/k_D ≈ 1.1–1.3); both indicate involvement without direct bond cleavage. A representative example is Michaelis-Menten kinetics in mechanisms, where steady-state analysis of enzyme-substrate yields the rate law v = (k_cat [E]_t [S]) / (K_m + [S]), with K_m reflecting affinity and k_cat the ; pre-steady-state studies reveal burst kinetics if product release limits the cycle. This framework, derived from early inversion studies, underpins mechanistic inferences in biocatalysis.

Spectroscopic and trapping techniques

Spectroscopic techniques provide direct evidence for the existence and structure of intermediates by probing their , , or properties on various timescales. Ultrafast laser , utilizing pulse durations, enables the observation of transient such as transition states and short-lived intermediates in photochemical and thermal reactions, achieving resolutions down to 10^{-15} seconds to capture ultrafast dynamics like bond breaking and formation. resonance (ESR) detects unpaired electrons in radical intermediates, offering insights into their geometry and reactivity, particularly in oxidation and radical chain processes where radicals persist for microseconds or longer. () identifies longer-lived intermediates through their characteristic chemical shifts and coupling patterns, allowing real-time monitoring of progress and intermediate accumulation in solution-phase mechanisms. Trapping experiments employ to capture and stabilize reactive , preventing their further reaction and enabling their isolation for subsequent analysis. For instance, the stable nitroxide TEMPO (2,2,6,6-tetramethylpiperidin-1-yl)oxyl acts as an efficient for carbon-centered radicals, forming persistent adducts that can be characterized by or further , thus confirming radical pathways in oxidative or photolytic reactions. These methods complement direct observation by providing qualitative evidence for intermediate involvement when lifetimes are too short for spectroscopic detection alone. Product studies analyze the stereochemical and structural outcomes of reactions to infer mechanistic pathways. In bimolecular (SN2) reactions, inversion of at the chiral center, as observed in the reaction of (R)- with yielding (S)-2-butanol, supports a concerted backside attack without free intermediates. Crossover experiments, where mixed pairs ligands to produce hybrid products, indicate the presence of intimate pair intermediates in solvolysis reactions. Isotopic labeling with stable isotopes like ^{13}C or tracks atom positions and migration during reactions, distinguishing between competing mechanisms. For example, labeling in hydrogen migrations reveals kinetic isotope effects that confirm rate-determining steps involving C-H breaking, while ^{13}C enrichment allows NMR detection of specific carbon atoms to map pathway branching. Representative examples illustrate these techniques' applications. generates and detects intermediates, such as glycosylidene carbenes from precursors, by monitoring their UV absorption transients and reaction with nucleophiles on nanosecond timescales. Chemically induced dynamic nuclear polarization () in NMR observes enhanced emission or absorption signals from pair recombination products, providing for cage and in mechanisms, as seen in photochemical dissociations where spin-correlated pairs produce polarized spectra.

Theoretical Modeling

Computational approaches

Computational modeling of reaction mechanisms enables the prediction and elucidation of chemical pathways , bypassing the need for extensive physical experimentation. These approaches have evolved significantly since the mid-20th century, beginning with foundational semi-empirical techniques such as Hückel , which provided early insights into π-electron systems and pericyclic reactions during the 1930s and 1950s. Over time, advancements in computational power and parameterization have extended these methods to broader classes of and inorganic reactions, facilitating the mapping of potential energy surfaces (PES) that describe reactant, intermediate, and product states. Classical methods, particularly (MM), are widely employed for simulating reaction mechanisms in large molecular systems where full quantum treatment is impractical. MM approximates the PES using empirical force fields that model bond stretching, angle bending, torsional rotations, and non-bonded interactions via parameterized potential functions, allowing efficient exploration of conformational changes and reaction coordinates in biomolecules or polymers. For instance, force fields like MMFF94 have been instrumental in studying enzyme-catalyzed reactions by optimizing geometries and estimating energy barriers along proposed mechanisms. Semi-empirical quantum mechanical methods bridge the gap between classical approximations and more rigorous quantum calculations, offering quantum-like accuracy at reduced computational cost for medium-sized molecules. Techniques such as Austin Model 1 (AM1) and Parameterized Model 3 (PM3) achieve this by neglecting certain integrals and parameterizing others based on experimental data, enabling the study of electronic effects in reaction pathways. These methods are particularly valuable for initial screening of mechanistic hypotheses, as they can handle systems with hundreds of atoms while incorporating approximate electron correlation. A key aspect of these computational approaches is the mapping of the on the PES, which involves optimization to identify energy minima corresponding to stable intermediates and first-order saddle points representing s. Optimization algorithms, such as the quasi-Newton or Berny methods, iteratively adjust molecular coordinates to minimize energy or satisfy the conditions for a , often guided by initial guesses from experimental data. This process allows for the construction of reaction profiles that reveal activation energies and rate-determining steps. Representative applications include the modeling of the Diels-Alder cycloaddition, a concerted [4+2] , where semi-empirical methods like AM1 have predicted / and activation barriers in good agreement with experimental outcomes for reactions involving and . Solvent effects are incorporated via implicit models, such as the conductor-like screening model (), which treats the solvent as a continuum to modulate the PES without explicit solvent molecules, thus capturing polarity influences on reaction rates. Such computations are routinely validated against kinetic and spectroscopic experimental data to refine mechanistic understanding. Recent advances as of 2025 integrate techniques, such as potentials, with traditional methods to accelerate PES exploration for complex systems, enabling simulations of larger timescales and molecular sizes previously inaccessible.

Quantum chemical methods

Quantum chemical methods employ to compute electronic structures and profiles of molecular systems, enabling precise elucidation of reaction mechanisms. These approaches solve the approximately to determine molecular geometries, transition states, and energy barriers, providing insights into reaction pathways that complement experimental data. By modeling electron correlations and exchange effects, they predict properties such as activation energies and stereochemical outcomes with for small to medium-sized molecules. Ab initio methods form the cornerstone of these calculations, starting with the Hartree-Fock (HF) approximation, which assumes a single wavefunction to compute the antisymmetrized product of molecular orbitals and accounts for electron exchange but neglects correlation energy. In HF theory, the molecular orbitals \phi_i satisfy the Roothaan equations: \mathbf{F} \mathbf{C} = \mathbf{S} \mathbf{C} \boldsymbol{\epsilon} where \mathbf{F} is the , \mathbf{S} the overlap matrix, \mathbf{C} the , and \boldsymbol{\epsilon} the orbital energies; this yields a mean-field solution to the many-electron problem. To incorporate electron correlation, post-HF techniques such as second-order Møller-Plesset (MP2) add corrections via Rayleigh-Schrödinger , treating correlation as a perturbation to the HF and improving energy accuracy for weakly correlated systems. For higher precision, coupled-cluster methods with single, double, and perturbative triple excitations, denoted CCSD(T), provide near-quantitative results by exponentially parameterizing the wavefunction as e^{\hat{T}} \Phi_0, where \hat{T} is the cluster operator, making it a benchmark for reaction energetics. Density functional theory (DFT) offers a computationally efficient alternative by mapping the many-electron problem onto a non-interacting electron system via the Hohenberg-Kohn theorems, which establish that the ground-state energy is a unique functional of the \rho(\mathbf{r}). The Kohn-Sham introduces fictitious orbitals \psi_i satisfying: \left[ -\frac{1}{2} \nabla^2 + v_{\text{ext}}(\mathbf{r}) + v_{\text{H}}[\rho](\mathbf{r}) + v_{\text{xc}}[\rho](\mathbf{r}) \right] \psi_i(\mathbf{r}) = \epsilon_i \psi_i(\mathbf{r}) with \rho(\mathbf{r}) = \sum_i |\psi_i(\mathbf{r})|^2, where v_{\text{H}} is the Hartree potential and v_{\text{xc}} the exchange-correlation potential; the total energy is then E = T_s[\rho] + \int v_{\text{ext}} \rho \, d\mathbf{r} + E_{\text{H}}[\rho] + E_{\text{xc}}[\rho]. Popular hybrid functionals like B3LYP combine exact HF exchange (20%) with Becke's gradient-corrected exchange and Lee-Yang-Parr correlation, balancing speed and accuracy for organic reaction profiles. These methods map potential energy surfaces (PES), which represent the as a of coordinates, to identify paths by scanning coordinates such as lengths or angles to locate minima (reactants/products) and saddle points (s). The intrinsic (IRC) traces the minimum-energy path from a to adjacent minima by following the steepest descent in mass-weighted coordinates, defined as \mathbf{q}(s) = \mathbf{q}_{\text{TS}} - s \mathbf{h} + \cdots, where s is the and \mathbf{h} the transition vector, revealing the connectivity of intermediates on the PES. Validation of quantum chemical predictions involves comparing computed activation energies E_a to experimental values, often achieving agreement within 1-2 kcal/ for CCSD(T) or high-level DFT on benchmark reactions, though discrepancies arise from anharmonic effects or influences. A key error source is basis set superposition error (BSSE), which artificially lowers interaction energies due to incomplete basis sets allowing orbitals to borrow from partners; counterpoise corrections mitigate this by computing energies with ghost atoms. In reactions like SN2, quantum calculations using CCSD(T) with augmented basis sets predict inversion barriers, such as ~15 kcal/mol for Cl⁻ + CH₃Cl, aligning with gas-phase experiments and confirming the collinear geometry. For pericyclic reactions, DFT with B3LYP elucidates , as in Diels-Alder cycloadditions, where the method predicts endo selectivity via secondary orbital interactions, matching observed diastereoselectivity in thermal [4+2] additions.

References

  1. [1]
    Reaction Mechanisms – Chemistry - UH Pressbooks
    The reaction mechanism (or reaction path) is the process, or pathway, by which a reaction occurs. A chemical reaction usually occurs in steps.
  2. [2]
    Chemical Reactivity - MSU chemistry
    Chemical Reaction: A transformation resulting in a change of composition, constitution and/or configuration of a compound ( referred to as the reactant or ...
  3. [3]
    CHEM 245 - Chemical kinetics: an introduction
    Reaction mechanisms. The rate law can be explained, or predicted, by a reaction mechanism which is a series of one or more elementary steps in which reactants ...
  4. [4]
    [PDF] 1 Elementary Reactions and Mechanisms
    The mechanism lists the elementary reactions making up a chemical pro- cess. It can be used to predict the detailed concentration vs. time behavior and.
  5. [5]
    [PDF] Useful Definitions: Mechanism
    Mechanism – A scheme that illustrates all reaction intermediates, as well as the flow of electrons and movement of atoms during bond breaking and bond making ...
  6. [6]
    [PDF] A First Course on Kinetics and Reaction Engineering Unit 6 ...
    The set of elementary reaction events that actually takes place at the molecular level is known as the reaction mechanism. Since each step in a reaction ...
  7. [7]
    7.12: Comparison of SN1 and SN2 Reactions - Chemistry LibreTexts
    May 30, 2020 · S N1 reactions are called solvolysis reactions when the solvent is the nucleophile. These patterns of reactivity are summarized in the table ...
  8. [8]
    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.
  9. [9]
    A BRIEF HISTORY OF CHEMICAL KINETICS
    Jan 31, 2001 · 1889: Arrhenius (Sweden) further analyzed the temperature dependence of reaction rate, k = A exp(-B/T), and gave it an "energy barrier" ...
  10. [10]
    Importance of Understanding Fundamental Chemical Mechanisms
    Sep 6, 2018 · Mechanisms at this level require that students understand that interactions between particles affect their kinetic and potential energies, and ...
  11. [11]
    3.2.1: Elementary Reactions - Chemistry LibreTexts
    Feb 12, 2023 · An elementary reaction is a single step reaction with a single transition state and no intermediates. Elementary reactions add up to complex ...
  12. [12]
    Difference Between Concerted and Stepwise Reactions
    Jun 24, 2020 · The key difference between concerted and stepwise reactions is that concerted reactions are single-step reactions, whereas stepwise reactions are multistep ...
  13. [13]
    Elementary Reaction Definition and Examples (Chemistry)
    Oct 2, 2022 · An elementary reaction is a chemical reaction that proceeds in a single step with only one transition state (reactants → products).
  14. [14]
    Elementary Reaction: Definition, Molecularity, and Rate Law
    Sep 2, 2021 · Example: The reaction between nitrogen dioxide (NO2) and carbon monoxide (CO) gives nitric oxide (NO) and carbon dioxide (CO2).
  15. [15]
    Radicals: Reactive Intermediates with Translational Potential
    Sep 15, 2016 · Radicals frequently enable the most direct means of reactivity umpolung. Due to their early transition states and lack of stifling aggregation ...
  16. [16]
    Excited State Anions in Organic Transformations - PMC - NIH
    Organic anions can be easily formed in the presence of base, and their rather long excited‐state lifetimes distinguish them from radical anions. Excited anionic ...
  17. [17]
    [PDF] Enzyme Kinetics - MIT OpenCourseWare
    This enzyme-substrate complex is a reaction intermediate, meaning that it is formed and consumed in the reactions, but does not appear in the overall chemical ...<|control11|><|separator|>
  18. [18]
    Enzyme Kinetics: The Enzyme Substrate Complex
    He proposed that the substrate and enzyme formed some intermediate substance which is known as the enzyme substrate complex. The reaction can be represented as: ...
  19. [19]
    Lifetimes of Dialkylcarbocations Derived from Alkanediazonium Ions ...
    Kinetics of the reactions of TMBH+ with nucleophiles such as halide ions and alcohols have been studied previously by excited state protonation of TMB (eq 2).
  20. [20]
    [PDF] Reaction of Carbocations, Carbenes and Radicals
    Successful free-radical reactions are usually chain processes. The lifetimes of the intermediate radicals are very short. To meet the synthetic requirements ...
  21. [21]
    Determination of Mechanism Chemi - csbsju
    Trapping approaches can be employed to intercept other reactive species. A classic example comes from the quest for cyclobutadiene.
  22. [22]
    New Approach to the Detection of Short-Lived Radical Intermediates
    Aug 24, 2022 · We report a new general method for trapping short-lived radicals, based on a homolytic substitution reaction SH2′.
  23. [23]
    Solving the structure of reaction intermediates by time-resolved ...
    Dec 15, 2008 · One of the methods allowing detection of intermediates that many spectroscopic techniques offer is based on the examination of isosbestic points ...
  24. [24]
    The Activated Complex in Chemical Reactions - AIP Publishing
    The Activated Complex in Chemical Reactions Available. Special Collection: JCP 90 for 90 Anniversary Collection. Henry Eyring.
  25. [25]
    Enzymatic transition states and transition state analogues
    Transition state lifetimes are a fraction of a picosecond, preventing chemical equilibrium in extended covalent systems. Thus, dynamic descriptions of ...
  26. [26]
    10.5: Transition States - Chemistry LibreTexts
    May 13, 2020 · It is customary to enclose this structure in brackets, with a double dagger at the top right. The transition state represents a point of such ...Missing: notation | Show results with:notation
  27. [27]
    6.2.3.3: The Arrhenius Law - Activation Energies
    Feb 12, 2023 · The activation energy ( E a ), labeled Δ ⁢ G ‡ in Figure 2, is the energy difference between the reactants and the activated complex, also known ...
  28. [28]
    7.2: SN2 Reaction Mechanism, Energy Diagram and Stereochemistry
    Dec 15, 2021 · In the transition state of SN2 reaction, there are five groups around the carbon and the carbon can be called “pentacoordinated”. As the OH – ...SN2 Reaction Mechanism · Energy Diagram of SN2... · The Effect of Alkyl Halide...
  29. [29]
    21.11: Pericyclic Reactions - Chemistry LibreTexts
    Jul 31, 2021 · There are numerous reactions in organic chemistry that proceed through cyclic transition states. They may be classified generally as pericyclic reactions.
  30. [30]
    3.2.3: Rate Determining Step - Chemistry LibreTexts
    Feb 12, 2023 · The rate determining step is the slowest step of a chemical reaction that determines the speed (rate) at which the overall reaction proceeds.
  31. [31]
    6.2.3.1: Arrhenius Equation - Chemistry LibreTexts
    Feb 13, 2024 · The two plots below show the effects of the activation energy (denoted here by E‡) on the rate constant. Even a modest activation energy of 50 ...
  32. [32]
    3.3.3: Reaction Order - Chemistry LibreTexts
    Feb 12, 2023 · The reaction order is the relationship between the concentrations of species and the rate of a reaction.
  33. [33]
    1.6: Integrated Rate Laws (Zeroth, First, and second order reactions)
    Jan 8, 2021 · The reaction rate of a zeroth-order reaction is independent of the concentration of the reactants. The reaction rate of a first-order ...Zeroth-Order Reactions · First-Order Reactions · Second-Order Reactions
  34. [34]
    5.2: Methods of Determining Reaction Order - Chemistry LibreTexts
    Feb 12, 2023 · Either the differential rate law or the integrated rate law can be used to determine the reaction order from experimental data.
  35. [35]
    5.14: The Lindemann-Hinshelwood Mechanism for First-order Decay
    Jul 7, 2024 · First-order kinetics for a unimolecular reaction corresponds to a constant probability that a given molecule reacts in unit time.
  36. [36]
    Pedagogical Content Knowledge of Chemical Kinetics: Experiment ...
    Jul 30, 2018 · For the assumption that all reactions are elementary, (15) those reactions have a distinctive feature: The molecularity (the number of molecules ...
  37. [37]
    17. Reaction Mechanisms - Elementary Reactions
    Bimolecular reaction – two molecules collide, interact, and undergo some kind of change. It can be two of the same or different molecules. Example: Hydrogen ...17 Reaction Mechanisms... · Molecularity · Consecutive Elementary...<|control11|><|separator|>
  38. [38]
    Molecularity - CK12-Foundation
    An example of a bimolecular reaction is the reaction between hydrogen and iodine to form hydrogen iodide: H2 + I2 → 2HI. Why is an SN2 reaction bimolecular?
  39. [39]
    The Theory of Trimolecular Reactions 1 - ACS Publications
    Termolecular reaction of nitrogen monoxide and oxygen: A still unsolved problem. ... The Energy and Entropy of Activation of the Reaction Between Nitric ...
  40. [40]
    [PDF] Methods of Determining Mechanisms | Dalal Institute
    These intermediates are usually detected by spectrophotometric methods directly; however, some unstable intermediates are first isolated by trapping them with ...
  41. [41]
    [PDF] Reaction Kinetics - Claire Vallance
    1. Introduction. Chemical reaction kinetics deals with the rates of chemical processes. Any chemical process may. be broken down into a sequence of one or more ...
  42. [42]
    Kinetic Analysis by Chemical Relaxation Methods | Rev. Mod. Phys.
    Chemical relaxation experiments, like temperature jumps, are used for kinetic studies. Mean relaxation times, especially the mean reciprocal time, are key for ...
  43. [43]
    Steady-State Approximation, Reaction Mechanism and Rate Law of ...
    This chapter has two aims: To describe the different categories of elementary reactions from which reaction mechanisms will be elaborated, To describe the ...
  44. [44]
    Steady-state, Pre-steady-state, and Single-turnover Kinetic ... - NIH
    Aug 19, 2013 · Here, we describe steady-state, pre-steady-state, and single-turnover approaches to isolate and measure specific steps during OGG1 catalytic cycling.Missing: review | Show results with:review
  45. [45]
    6 Kinetic Isotope Effects in Enzymic Reactions - ScienceDirect.com
    A primary kinetic isotope effect results when the bond to the isotopically substituted atom is broken in the transition state. A secondary isotope effect ...Missing: paper | Show results with:paper
  46. [46]
    Translation of the 1913 Michaelis–Menten Paper - ACS Publications
    Sep 2, 2011 · Here we introduce the translation, describe the historical context of the work, and show a new analysis of the original data.
  47. [47]
    Predictions of Steady-State Photo-CIDNP Enhancement by Machine ...
    Jul 28, 2025 · Photochemically induced dynamic nuclear polarization (photo-CIDNP) is a hyperpolarization method used to boost signal sensitivity in NMR ...
  48. [48]
    Applications of Electron Spin Resonance Spectrometry for Reactive ...
    ESR is used for quantitative and qualitative analysis of reactive oxygen species (ROS) and reactive nitrogen species (RNS), including short-lived free radicals.
  49. [49]
    Mechanistic analysis by NMR spectroscopy: A users guide
    NMR spectroscopy is used for in situ analysis of reaction mechanisms, providing non-destructive, quantitative analysis of organic and organometallic reactions.Mechanistic Analysis By Nmr... · 1. Introduction · 2. Practicalities And The...
  50. [50]
    Picosecond laser studies of ultrafast processes in chemistry
    Investigation of phenyl azide photochemistry by conventional and time-resolved spectroscopy. Elucidation of intermediates and reaction mechanisms. Journal ...
  51. [51]
    Recent Developments for the Deuterium and Tritium Labeling of ...
    Feb 18, 2022 · Noteworthy, secondary kinetic isotope effects offer valuable information to understand the reaction mechanisms. 5.2. Drug Development. The ...
  52. [52]
    A laser flash photolysis derived study of a glycosylidene carbene
    A laser flash photolysis derived study of a glycosylidene carbene. Click ... Research on Chemical Intermediates 1994, 20 (2) , 175-193. https://doi.org ...
  53. [53]
    The Nobel history of computational chemistry. A personal perspective
    May 3, 2024 · The evolution of computational chemistry resulted from a combination of advances in theoretical methods, the development of powerful algorithms ...
  54. [54]
    Development and use of quantum mechanical molecular models. 76 ...
    Development and use of quantum mechanical molecular models. 76. AM1: a new general purpose quantum mechanical molecular model.Missing: original | Show results with:original
  55. [55]
    Ab initio quantum chemistry: Methodology and applications - PMC
    May 10, 2005 · Hartree–Fock theory, a mean field approach, produces reasonable results for many properties but is incapable of providing a robust description ...Missing: seminal | Show results with:seminal
  56. [56]
    [PDF] Ab initio and electronic structure methods ChE210D
    Jul 31, 2025 · The Hartree-Fock approximation to the wavefunction is equivalent to a mean-field solution of. Schrodinger's equation in which each electron ...
  57. [57]
    [PDF] Ab Initio Theoretical Ch i l Ki ti emical Kinetics
    Hartree-Fock (HF). 2. Second-Order Moller Plesset Perturbation. Theory (MP2). 3. Configuration Interaction (CISD). 4. Coupled Cluster Theory [CCSD(T);. QCISD(T)].
  58. [58]
    Coupled-cluster theory in quantum chemistry | Rev. Mod. Phys.
    Feb 22, 2007 · The correlation energy is defined with respect to the Hartree-Fock energy for the given basis set, and the full correlation energies are ...Missing: seminal | Show results with:seminal
  59. [59]
    Density functional theory: Its origins, rise to prominence, and future
    Aug 25, 2015 · This paper reviews the development of density-related methods back to the early years of quantum mechanics and follows the breakthrough in their application ...
  60. [60]
    Intrinsic reaction coordinate: Calculation, bifurcation, and automated ...
    Sep 3, 2014 · The intrinsic reaction coordinate (IRC) approach has been used extensively in quantum chemical analysis and prediction of the mechanism of chemical reactions.
  61. [61]
    How Accurate are Approximate Quantum Chemical Methods at ... - NIH
    Calculations of relative interaction energies allow for cancellation of systematic errors, including basis set truncation and superposition errors, and the ...Missing: validation activation
  62. [62]
    [PDF] Basis Set Superposition Errors: Theory and Practice
    Apr 16, 2019 · Basis set superposition error (BSSE) is the unbalancing of basis sets that causes an artificial lowering of energy when modeling interactions, ...
  63. [63]
    An interacting quantum atom study of model SN2 reactions (X ...
    Nov 10, 2017 · The calculated barrier for X = Cl is in good agreement with the experimental30 one (55.2 ± 8.4 kJ/mol) and with previous high-level ab ...
  64. [64]
    A 21st Century View of Allowed and Forbidden Electrocyclic Reactions
    ... quantum chemical methods to examine allowed and forbidden pericyclic reactions. Indeed, our major focus will be following the orbitals during what would be ...