Fact-checked by Grok 2 weeks ago

Kinetic isotope effect

The kinetic isotope effect (KIE) is a phenomenon in where the reaction rate of a process changes upon isotopic substitution of one or more s in the reactants, primarily due to mass-dependent differences in molecular vibrations and zero-point energies without altering the electronic structure. This effect, first theoretically described in the early following the discovery of , arises from quantum mechanical principles, where heavier isotopes exhibit lower vibrational frequencies, leading to lower zero-point energies and typically slower reaction rates compared to lighter isotopes. KIEs are quantified as the ratio of rate constants (k_light/k_heavy), with primary KIEs occurring when the isotopic substitution is at a being broken or formed in the rate-determining step, often yielding large values (e.g., k_H/k_D up to 7–8 for / substitutions), while secondary KIEs involve atoms not directly involved in bond breaking and result in smaller effects (typically 1.01–1.25). Heavy atom KIEs, such as those involving /13 or /18, are subtler (1.02–1.10) but still diagnostically useful. Theoretically grounded in and , KIEs reflect differences in the of activation between isotopologues, influenced by factors like tunneling in proton transfers, where effects can exceed classical predictions (e.g., up to 700 in enzymatic reactions at low temperatures). Historically, the effect was formalized by Jacob Bigeleisen and in 1947, building on earlier work by Urey, Eyring, and Polanyi, and it has since become a for elucidating reaction mechanisms, identifying rate-limiting steps, and probing geometries in organic, inorganic, and biochemical systems. Applications span synthetic chemistry, where labeling optimizes (e.g., in norzoanthamine ), to enzymology, where KIEs reveal proton transfer pathways, and computational modeling, employing methods like semiclassical theory for accurate predictions. Notable examples include the bromination of acetone, where H/D substitution at the alpha position yields a primary KIE of approximately 7, and E2 eliminations, where beta- effects distinguish concerted from stepwise mechanisms. Overall, KIEs provide precise, isotope-specific insights into reaction dynamics, making them indispensable for advancing mechanistic understanding across chemical disciplines.

Introduction

Definition and Principles

The kinetic isotope effect (KIE) is defined as the change in the rate of a that results from the substitution of an atom in the reactants with one of its isotopes, quantified as the ratio of the rate constants for the unlabeled and isotopically labeled , typically expressed as k_\text{light}/k_\text{heavy} . This effect arises primarily from differences in , which influence the vibrational properties of molecular bonds and, consequently, the energy barriers to reaction . Isotopes are variants of a that differ in number but share the same , leading to distinctions between (non-radioactive) and radioactive (unstable) forms . KIE studies predominantly employ isotopes due to their persistence in experimental systems, with common examples including hydrogen-1 (¹H, natural abundance ~99.985%) and (²H or D, ~0.015%), (¹²C, ~98.93%) and (¹³C, ~1.07%), and nitrogen-14 (¹⁴N, ~99.632%) and nitrogen-15 (¹⁵N, ~0.368%) . These substitutions minimally alter electronic structure but significantly impact nuclear mass-dependent properties. At the core of KIE principles lies the quantum mechanical behavior of molecular vibrations, modeled as harmonic oscillators where vibrational frequency \nu scales inversely with the of the \mu, such that heavier isotopes exhibit lower frequencies . This results in differences in (ZPE), the residual vibrational energy at , which is higher for lighter isotopes; consequently, isotopic substitution can lower the ZPE in the relative to the reactants, altering the \Delta E_a . The magnitude of the KIE is approximately given by k_1/k_2 \approx \exp(\Delta E_a / RT), where \Delta E_a is the activation energy difference induced by the substitution, R is the , and T is the . The significance of KIEs lies in their utility for elucidating reaction mechanisms, particularly by revealing the extent of weakening or changes in vibrational modes at the without directly observing it . By comparing rate ratios, researchers can infer the symmetry and nature of the , providing insights into whether a is being broken or formed in the rate-determining step .

Historical Development

The discovery of the kinetic isotope effect (KIE) emerged in the early , closely tied to the identification of by and his collaborators at , who reported spectroscopic evidence for a heavy isotope of in 1932. This breakthrough, for which Urey received the in 1934, enabled the preparation of deuterated compounds and prompted investigations into isotopic differences in chemical behavior. Initial theoretical predictions of rate variations arose in 1933, when Henry Eyring and proposed that differences between and could lead to distinct reaction rates in proton-transfer processes. Early experimental observations of KIEs focused on - exchange reactions, with the first reports of rate differences appearing between 1932 and 1934. Karl-Friedrich Bonhoeffer and colleagues at the University of Berlin documented slower exchange rates for compared to in aqueous systems in 1935, providing direct evidence of kinetic distinctions in simple isotopic substitutions. Concurrently, Urey's group explored isotopic in exchange equilibria and , laying groundwork for recognizing KIEs as tools for probing reaction mechanisms. In the late 1940s, at the , extended these studies to carbon isotope effects in , such as the decomposition of , demonstrating measurable rate variations that highlighted KIEs' utility in tracing reaction pathways. Theoretical formalisms advanced significantly in the 1950s and 1960s through the work of Jacob Bigeleisen and Max Wolfsberg, who developed semiclassical models for predicting KIE magnitudes based on vibrational frequency shifts in transition states. Bigeleisen's 1949 extension of isotope theory to provided a foundational framework, while their 1958 collaboration refined approximations for heavy-atom and hydrogen KIEs, emphasizing temperature-independent factors. Lars Melander's 1960 monograph, Isotope Effects on Reaction Rates, synthesized these developments and introduced primary hydrogen-deuterium KIEs as diagnostic probes for bond-breaking steps in , marking a shift toward mechanistic applications. The 1970s saw expanded exploration of secondary KIEs, where isotopic substitution at non-reacting positions revealed subtle vibrational and steric influences on rates, as detailed in reviews by Kenneth B. Wiberg and Frank H. Westheimer. Recognition of inverse KIEs—where the heavier isotope reacts faster—gained traction in the mid-20th century, particularly as discussed in Melander's 1960 , often linked to pre-equilibrium steps or compressed transition states, as observed in certain nucleophilic substitutions and enzymatic processes. By the 1990s, integration with computational methods, including quantum mechanical calculations of vibrational frequencies, transformed KIEs from empirical observables into predictive tools for transition-state modeling, exemplified by David A. Singleton's NMR-based competitive methods. This evolution elevated KIEs from isolated observations to a of mechanistic .

Classification

Primary Kinetic Isotope Effects

Primary kinetic isotope effects (primary KIEs) arise from isotopic at atoms directly involved in the making or breaking during the rate-determining step of a . This alters the primarily due to differences in (ZPE) between the isotopologues, with the heavier typically leading to a slower . Unlike secondary KIEs, which involve remote substitutions and yield smaller effects, primary KIEs are larger because the isotopic change directly impacts the vibrational modes of the reacting . The magnitude of primary KIEs depends on factors such as , temperature, and the extent of bond cleavage or formation in the . For /deuterium (H/D) substitutions, typical primary KIE values range from 2 to 8 at 25°C, reflecting the significant difference (doubling from H to D) and its effect on vibrational frequencies. For heavier atoms like , primary KIEs are much smaller, typically 1.01 to 1.10, due to the minor relative change (about 8%). These effects can be quantified using the Bigeleisen-Mayer equation, which relates the natural logarithm of the KIE to differences in vibrational frequencies between reactant and transition states. A common semiclassical approximation is \ln(KIE) \approx \frac{\Delta \mathrm{ZPE}}{RT}, where \Delta \mathrm{ZPE} is the difference in zero-point energy between the light and heavy isotopologues for the reacting bond, R is the gas constant, and T is temperature; more precise calculations involve u_i = h c \omega_i / kT with \omega_i as the frequency (in cm⁻¹). Symmetric transition states maximize the KIE, while temperature dependence arises from the thermal population of vibrational levels, reducing the effect at higher temperatures. The degree of bond change further modulates the magnitude: maximal KIE occurs when the bond is half-broken in a linear, symmetric transition state. Interpretation of primary KIE magnitudes provides insight into structure via the Hammond postulate, which posits that the resembles the nearest extremum in energy (reactant or product). A large primary KIE (e.g., approaching the maximum for H/D) indicates a late with significant bond cleavage, as in exothermic reactions where the is product-like. Conversely, a small KIE suggests an early with minimal bond change, typical of endothermic processes. For nucleophilic substitutions, an SN1 mechanism (rate-determining departure of ) might exhibit a modest primary KIE if the isotope is on the departing group, reflecting partial bond breaking, whereas an SN2 mechanism (concerted) could show a larger KIE for isotopes directly at the reaction center, indicating a more symmetric, central . Primary KIEs can be limited in multi-step reactions or those involving pre-equilibria, where the observed effect may be masked or attenuated if the isotopic step is not rate-determining. For instance, a pre-equilibrium isotope exchange before the slow step can lead to an apparent or reduced KIE, complicating direct attribution to bond breaking. In complex mechanisms with multiple isotopically sensitive steps, the composite KIE requires to isolate the primary contribution, often necessitating additional experimental probes.

Secondary Kinetic Isotope Effects

Secondary kinetic isotope effects (SKIEs) occur when isotopic takes place at atoms not directly involved in the bond-making or bond-breaking process of the rate-determining step, yet still influence the through indirect vibrational or structural changes. These effects are distinct from primary kinetic isotope effects, which arise at the reaction center and serve as a for comparison due to their larger magnitudes. SKIEs are generally small and provide valuable insights into the and electronic structure of the without altering the primary reaction pathway. SKIEs are categorized based on the position of isotopic relative to the reaction center: alpha-secondary effects involve substitution at the adjacent to the of bond change, while beta- and gamma-secondary effects occur at further removed positions. The underlying mechanisms vary by type; alpha-SKIEs often stem from inductive effects or changes in vibrational modes due to hybridization shifts, beta-SKIEs frequently involve where sigma adjacent to the reaction center donate to the developing orbital, and gamma-SKIEs may arise from long-range inductive or steric influences. Additionally, steric release mechanisms can contribute, particularly in crowded transition states where isotopic alleviates spatial constraints. The magnitude of SKIEs is typically modest, with hydrogen/deuterium (H/D) effects ranging from 1.1 to 1.4 for (k_H/k_D > 1) cases and occasionally (k_H/k_D < 1) values around 0.8–0.9, while carbon-13 effects are even smaller, between 1.00 and 1.03. Unlike some primary KIEs that may exhibit complex temperature profiles due to tunneling, SKIE magnitudes generally decrease with increasing temperature, reflecting a reduced influence of zero-point energy differences at higher thermal energies. Interpretation of SKIEs focuses on probing subtle transition state features, such as rehybridization from sp³ to sp², which alters out-of-plane bending frequencies and leads to inverse effects as deuterium's lower vibrational amplitude better accommodates the flatter geometry. Steric interactions can also be assessed, as isotopic substitution affects molecular conformation through vibrational differences. A common approximation for the SKIE accounts for these vibrational adjustments: \text{KIE} \approx \exp\left[ -\frac{\Delta \text{ZPE} + \Delta E_\text{bend}}{RT} \right] Here, \Delta \text{ZPE} represents the differential zero-point energy change, and \Delta E_\text{bend} captures shifts in bending mode energies between the reactant and transition state, emphasizing non-stretching vibrational contributions. Special cases include steric isotope effects, where zero-point energy differences between C-H and C-D bonds influence conformational preferences, potentially accelerating or decelerating the reaction; for instance, the lower amplitude of C-D vibrations results in effectively shorter bonds, reducing steric repulsion in tight transition states like those in S_N2 reactions. These effects, first systematically explored in mid-20th-century reviews, underscore SKIEs' utility in distinguishing mechanistic nuances such as associative versus dissociative pathways.

Inverse Kinetic Isotope Effects

Inverse kinetic isotope effects (IKIEs) occur when the heavier isotope in a reactant reacts faster than the lighter one, resulting in a kinetic isotope effect ratio less than unity (k_light/k_heavy > 1, or equivalently k_heavy/k_light < 1). This phenomenon is relatively rare compared to normal kinetic isotope effects, as it requires specific conditions where the zero-point energy (ZPE) difference between isotopologues is smaller in the transition state than in the ground state (ΔZPE_TS < ΔZPE_GS). In such cases, the heavier isotope experiences less of a ZPE penalty in progressing to the transition state, leading to a lower activation barrier relative to the lighter isotope. Mechanistically, IKIEs are associated with tight transition states, where vibrational force constants are higher than in the reactants, compressing the ZPE difference. In contrast, loose transition states typically produce normal effects due to greater ZPE separation. A common origin for inverse alpha-secondary effects involves sp²-hybridized carbons in the ground state (e.g., in alkenes or carbonyls), where out-of-plane bending modes have larger ZPE differences; progression to an sp³-like transition state reduces this disparity, favoring the heavier isotope. This simplified model highlights that IKIEs arise when \Delta \text{ZPE}_\text{TS} < \Delta \text{ZPE}_\text{reactants}, as the activation energy difference favors the heavier isotopologue. Such effects are particularly noted for heavy isotopes like ¹³C or ¹⁸O in associative mechanisms, including certain , where the transition state tightens bonds involving these atoms. The magnitude of IKIEs is typically small and close to unity, often ranging from 0.9 to 1.0 per atom for heavy isotopes, reflecting subtle ZPE shifts. For deuterium in secondary positions, values around 0.8–0.9 are common. These effects are often temperature-independent, serving as a diagnostic tool for tight transition states, unlike normal KIEs which may vary with temperature due to entropic contributions. This distinguishes IKIEs from equilibrium isotope effects (EIEs), which are generally normal (k_heavy/k_light < 1) as heavier isotopes preferentially occupy lower ZPE ground states in thermodynamic equilibria.

Theoretical Foundations

Semiclassical Theory

The semiclassical theory of kinetic isotope effects builds upon transition state theory (TST), adapting the Eyring equation to account for isotopic differences in molecular partition functions and zero-point energies without invoking full quantum mechanical treatments of dynamics. In TST, the rate constant for a reaction is expressed as k = \frac{k_B T}{h} e^{-\Delta G^\ddagger / RT}, where \Delta G^\ddagger is the free energy difference between reactants and the transition state. For isotopic variants, the kinetic isotope effect (KIE) is defined as \mathrm{KIE} = k_\mathrm{light}/k_\mathrm{heavy} = e^{-(\Delta G^\ddagger_\mathrm{light} - \Delta G^\ddagger_\mathrm{heavy})/RT}. Under the Born-Oppenheimer approximation, which separates electronic and nuclear motions, the free energy differences stem primarily from vibrational contributions. The full expression for the KIE in terms of molecular partition functions Q (including translational, rotational, and vibrational components) is \mathrm{KIE} = \frac{Q^\ddagger_\mathrm{light}}{Q^\ddagger_\mathrm{heavy}} \cdot \frac{Q_\mathrm{heavy}}{Q_\mathrm{light}} \cdot e^{-\Delta E_0 / RT}, where the superscript \ddagger denotes the transition state, and \Delta E_0 captures the zero-point energy (ZPE) difference between the transition state and reactants. This formulation highlights how isotopic substitution alters the effective activation energy through changes in vibrational properties at the transition state relative to the reactants. Central to the semiclassical description is the role of zero-point energy, which arises from the quantum nature of vibrations even in this framework. The ZPE for a vibrational mode is given by \frac{1}{2} h \nu, where \nu is the frequency; heavier isotopes increase the reduced mass \mu, lowering \nu via the scaling \nu \propto 1/\sqrt{\mu} and thus reducing the ZPE. This effect is most pronounced for modes involving the isotopically substituted atom that change significantly between reactants and transition state, leading to \Delta E_0 = \sum \frac{1}{2} h c \Delta \tilde{\nu} , where \tilde{\nu} denotes wavenumber and c is the speed of light. For primary KIEs, where the isotope is directly involved in bond breaking or forming, the lower ZPE at the transition state for the heavy isotope raises the effective barrier, yielding KIE values greater than unity. In contrast, secondary effects involve remote modes with smaller \Delta \tilde{\nu}, resulting in more modest KIEs. These ZPE differences dominate the temperature-independent component of the KIE in semiclassical models. To facilitate practical calculations, semiclassical approximations expand the partition function ratios logarithmically, as in the Bigeleisen equation, which is particularly effective for high-frequency vibrational modes and small isotopic mass differences. The approximate form is \ln(\mathrm{KIE}) \approx \sum_i \frac{u_i}{24} \coth\left(\frac{u_i}{2}\right) \Delta u_i, where the sum is over normal modes, u_i = h c \tilde{\nu}_i / k_B T, and \Delta u_i is the frequency shift due to isotopic substitution (with the reaction coordinate mode typically excluded or treated separately). This second-order Taylor expansion of the vibrational partition function assumes temperatures high enough that u_i \ll 1 for most modes but captures leading corrections for stiffer vibrations. The equation simplifies computation by focusing on frequency changes rather than full partition functions, emphasizing contributions from modes sensitive to the isotope. It applies well to heavy-atom KIEs but requires caution for light isotopes. These semiclassical models rest on foundational assumptions, including the Born-Oppenheimer approximation for treating nuclear motion on a fixed electronic potential energy surface and the harmonic oscillator model, which approximates vibrational potentials as quadratic to derive frequencies and partition functions. Translational and rotational partition functions also contribute minor symmetry and mass-dependent terms. However, limitations emerge for systems involving hydrogen or deuterium, where the low reduced mass amplifies frequency shifts (up to \sqrt{2} for H/D), rendering harmonic approximations inadequate due to anharmonicity and potentially violating the high-temperature limit of the expansions. Such cases often necessitate refinements beyond pure semiclassics, though the framework remains essential for baseline predictions.

Quantum Mechanical Contributions

In quantum mechanical treatments of kinetic isotope effects (KIEs), valence bond (VB) and molecular orbital (MO) theories provide a framework for understanding the isotopic dependence of transmission coefficients at the transition state. These theories describe the reaction as an overlap of electronic wavefunctions between reactant and product configurations, where isotopic substitution alters the nuclear masses and thus the vibrational components of the total wavefunction. The transmission coefficient, which accounts for reactive flux through the barrier, varies with isotope due to differences in wavefunction overlap, leading to enhanced or diminished KIEs beyond classical expectations. For instance, in proton transfer reactions, VB models like the empirical valence bond (EVB) approach capture how mass-dependent delocalization affects the probability of bond breaking and forming. Anharmonicity introduces corrections to the harmonic approximation for zero-point energy (ZPE), particularly relevant for light atom transfers like H/D substitutions. In the transition state, where force constants weaken, anharmonic effects distort the potential energy surface, shifting vibrational energy levels and partition function ratios. This is especially pronounced in bending modes during H/D transfers, where the looser transition state geometry amplifies anharmonic contributions, often resulting in inverse secondary KIEs from higher deuterium frequencies in these modes compared to stretching modes. Computational corrections for ground-state anharmonicity can adjust primary KIEs by 5-20% in model systems like triatomic ring openings. Quantum paths in reaction coordinates deviate from classical collinear trajectories through corner-cutting, where the wavepacket explores multidimensional space to find shorter barrier-crossing routes. This non-classical behavior, prominent in tunneling-dominated reactions, increases the effective transmission coefficient for lighter isotopes, magnifying KIEs by reducing the barrier width perceived by the wavefunction. In contrast, collinear paths align with the minimum energy path but underestimate quantum flux, leading to smaller predicted KIEs; corner-cutting effects are evident in H/D transfers with barriers around 5-10 kcal/mol, enhancing KIEs by factors of 2-5 relative to collinear approximations. The quantifies these paths by integrating over instanton trajectories, revealing their role in accurate KIE predictions. Ab initio computational methods calculate KIEs by evaluating isotopic partition functions with quantum corrections to vibrational frequencies, often incorporating anharmonic and multidimensional effects. Software like Gaussian performs frequency analyses at the transition state and reactants using density functional theory or post-Hartree-Fock levels (e.g., MP2/6-31+G(d,p)), yielding ZPE differences and thermal corrections via the rigid rotor-harmonic oscillator approximation, refined for isotopes. These approaches, combined with transition state theory, reproduce experimental KIEs within 10-20% for organic reactions, as seen in path-integral simulations of enzymatic proton transfers.

Transient and Tunneling Effects

Quantum tunneling refers to the quantum mechanical probability that a particle penetrates a potential energy barrier despite having insufficient classical energy to surmount it, arising from the wave-like nature of particles and their non-zero penetration into classically forbidden regions. In kinetic isotope effects, tunneling disproportionately affects lighter isotopes, such as hydrogen over deuterium, due to the inverse mass dependence of the tunneling probability, which exaggerates the rate difference and contributes to observed KIEs beyond semiclassical expectations, particularly in hydrogen transfer reactions. A key theoretical treatment of tunneling's impact on reaction rates is the Bell correction, which modifies the transition state theory rate constant to account for sub-barrier penetration assuming a parabolic barrier potential. The tunneling transmission coefficient is given by \kappa_\text{tunnel} = \frac{\sinh\left(2\pi a E_a / h v\right)}{2\pi a E_a / h v}, where a is the imaginary frequency parameter of the transition state, E_a is the classical activation energy, h is , and v is the vibrational frequency along the reaction coordinate; this factor exceeds unity and increases the computed KIE when applied differentially to isotopic variants. Tunneling contributions become more prominent at low temperatures, where thermal energy is insufficient for classical barrier crossing, leading to enhanced KIEs as the tunneling pathway dominates. This manifests in curved Arrhenius plots, with the logarithm of the rate constant deviating upward from linearity at lower temperatures, reflecting reduced temperature sensitivity of the rate. For primary H/D KIEs in hydrogen transfer reactions, values up to 100 have been reported at around -100°C, substantially larger than the room-temperature semiclassical limit of approximately 7, highlighting tunneling's role in amplifying isotope sensitivity under cryogenic conditions. Transient kinetic isotope effects occur in systems with short-lived intermediates, where isotopic scrambling or equilibration proceeds faster than the forward reaction, altering the observed KIE from the intrinsic transition state value. These effects are particularly relevant in pre-steady-state regimes, where rapid buildup of intermediates allows isotopic exchange before commitment to product formation. Theoretical models for pre-steady-state KIEs typically employ numerical integration of rate equations for isotopologues, incorporating branching ratios for equilibration versus reaction to fit experimental transients and dissect step-specific isotope sensitivities. Experimental detection of tunneling often relies on the temperature dependence of KIEs, where plots of \ln(\text{KIE}) versus $1/T exhibit greater width or curvature than semiclassical predictions, indicating increasing tunneling dominance as temperature decreases. For heavy atoms, however, tunneling effects are severely limited by their larger mass, which reduces the de Broglie wavelength and barrier penetration probability, confining observed KIEs to near-semiclassical magnitudes typically below 1.05 even at low temperatures.

Experimental Methods

Competitive Measurements

Competitive measurements of kinetic isotope effects (KIEs) rely on intermolecular competitions between isotopologues in the same reaction mixture to determine the ratio of rate constants, R = k_{\text{light}}/k_{\text{heavy}}, by monitoring the differential consumption of labeled and unlabeled substrates. This approach assumes no significant intramolecular isotope effects and utilizes the relationship derived from the fractional extent of reaction, where the natural logarithm of the ratio of remaining light to heavy substrate fractions equals R times the extent of reaction: \ln\left(\frac{f_{\text{light}}}{f_{\text{heavy}}}\right) = R \cdot x Here, f represents the fraction of unreacted substrate, and x is the fractional conversion of the reaction. This method is particularly suited for primary KIEs, where the isotopic substitution occurs at the reacting site. In practice, the setup involves preparing a mixture of isotopically labeled substrates, such as methanol isotopologues CH₃OH and CH₃OD for hydrogen/deuterium studies, and allowing the reaction to proceed to partial conversion under controlled conditions. The remaining substrate and product mixtures are then isolated and analyzed, typically via , , or , to quantify the isotopic compositions and compute R from changes in ratios relative to the initial mixture. This competitive format ensures that both isotopologues experience identical reaction conditions, minimizing external variables. The primary advantages of competitive measurements include high precision, especially for large KIEs such as those exceeding 2 for , where relative errors can be as low as ±0.3% using techniques like scintillation counting or . Additionally, remote labeling—placing the isotope at a non-reacting position—facilitates easier synthesis of labeled compounds and reduces handling challenges, particularly with radioactive isotopes like ¹⁴C or ³H, while stable isotopes like ¹⁸O enable safer applications. This method excels in providing direct rate ratios without needing separate reactions for each isotopologue. Limitations arise when products are not easily separable, as isolation is required for accurate ratio analysis, and contaminants can introduce errors in isotopic quantification. Secondary isotope effects may mask or complicate the primary KIE signal, necessitating corrections. In multi-step reactions, the observed KIE (R_{\text{obs}}) represents a weighted average across isotopically sensitive steps, expressed as R_{\text{obs}} = \sum (f_i \cdot R_i), where f_i are the fractional contributions of each step and R_i the intrinsic KIEs, potentially underestimating the effect at the rate-determining step if commitments to catalysis are high.

Natural Abundance Approaches

Natural abundance approaches measure kinetic isotope effects (KIEs) by analyzing the inherent isotopic compositions of elements like carbon-13 (¹³C, natural abundance ~1.1%) and deuterium (²H, ~0.015%) in substrates before and after partial reactions, eliminating the need for synthetic isotopic enrichment. These methods exploit high-precision isotope ratio mass spectrometry (IRMS) to quantify subtle shifts in isotopic ratios, typically expressed in δ notation (δ¹³C or δ²H in per mil, ‰), resulting from differential reaction rates of isotopologues. The underlying principle follows the Rayleigh distillation model for unidirectional reactions in closed systems, where continuous removal of product leads to isotopic fractionation in the remaining reactant or cumulative product. The enrichment factor ε, representing the deviation from unity in the fractionation (KIE = 1 + ε), is derived from the Rayleigh equation R / R_0 = f^{(\varepsilon)} for the remaining substrate (f = fraction remaining), or for low conversions and instantaneous product, approximated as ε ≈ (R_start / R_product) - 1, where R is the heavy-to-light isotope ratio (e.g., ¹³C/¹²C). This model assumes constant fractionation and low conversion extents for linear approximations, enabling KIE determination from measured δ shifts via regression of ln(R / R_0) vs. ln f. IRMS techniques form the core of these measurements, providing bulk (intermolecular) isotope ratios with precision better than 0.1‰, suitable for averaging KIEs across multiple sites in a molecule. For intramolecular (site-specific) analysis, compound-specific IRMS (CS-IRMS) interfaces with gas chromatography or elemental analyzers to isolate and measure ratios at individual positions, distinguishing local from global effects. Reactions are typically run to partial completion (e.g., 10-50% conversion), with substrates and products isolated via extraction or chromatography for sequential IRMS analysis; baseline isotopic compositions are established from unreacted controls to account for instrumental drift. These approaches have been validated against competitive labeling methods in simple systems, confirming accuracy for carbon KIEs. The primary advantages lie in applicability to complex, polydisperse molecules—such as natural products or biomolecules—where targeted labeling is impractical or cost-prohibitive, and in sensitivity to small KIEs inherent to non-hydrogen elements (e.g., ¹³C KIEs of ~1.02). For instance, natural abundance IRMS has quantified a ¹³C KIE of 1.021 ± 0.006 at the anomeric carbon in the phosphorolysis of inosine by , revealing oxocarbenium ion-like transition states without enrichment. Similarly, δD measurements via IRMS have detected hydrogen KIEs in enzyme-catalyzed reactions, leveraging the larger natural fractionation potential (up to 1.2-1.5). This non-invasive strategy also supports studies of heavy elements, like chlorine or bromine in solvents, where natural abundances suffice for detectable effects. Challenges include inherent isotopic heterogeneity from biosynthetic or environmental sources, which can mask reaction-induced shifts and require rigorous statistical corrections or multi-point sampling along the reaction progress for Rayleigh plot linearity. In molecules with multiple reactive sites, observed bulk fractionation represents a composite, necessitating computational models or selective derivatization to deconvolute site-specific ε values; baseline variations in δ¹³C (~±1‰ across commercial sources) further demand high-throughput IRMS with automated correction algorithms. Despite these, advancements in continuous-flow IRMS have enhanced throughput, making natural abundance methods routine for KIEs as small as 1.005-1.025.

Spectroscopic and Mass Spectrometry Techniques

Nuclear magnetic resonance (NMR) spectroscopy plays a central role in determining kinetic isotope effects (KIEs) through site-specific analysis of isotopic ratios at natural abundance or enriched levels. In particular, 13C NMR with 1H decoupling enables precise measurement of primary and secondary 13C KIEs by producing narrow resonances that allow quantification of subtle isotopic fractionations in reactants and products. This technique involves acquiring spectra under composite pulse decoupling conditions to minimize broadening from proton-carbon couplings, facilitating the detection of KIEs as small as 1-2% in enzymatic reactions such as glucoside hydrolysis. For distinguishing between CH and CHD groups in secondary H/D KIE studies, specialized pulse sequences like DEPT variants enhance signal discrimination based on carbon-proton multiplicities, where CHD sites exhibit altered polarization transfer due to the deuterium's lower gyromagnetic ratio. This approach is particularly useful in natural abundance deuterium NMR to probe vibrational and hyperconjugative effects in transition states. Quantitative single-pulse 13C NMR further supports natural abundance measurements by directly comparing peak intensity ratios of isotopomers before and after partial reaction conversion, yielding intermolecular KIEs without isotopic enrichment. For instance, the method integrates over multiple acquisitions to achieve reliable ratios (R/R₀) for carbon positions involved in bond breaking or forming. Mass spectrometry techniques complement NMR by providing high-throughput analysis of isotopic compositions, especially for volatile or gaseous samples. Isotope ratio mass spectrometry (IRMS), often coupled with gas chromatography (GC-IRMS), quantifies KIEs through δ-values representing deviations in isotopic ratios (e.g., δ¹³C) from international standards, capturing kinetic fractionations in processes like N₂O production by denitrifying enzymes. This method excels in measuring apparent KIEs (ε) from linear regressions of δ versus reaction progress (ln f / (1-f)), with site-specific resolution for α and β positions in molecules. For intramolecular KIEs in volatile compounds, GC-MS variants, such as GC-Orbitrap MS, resolve carbon isotope distributions within fragments, enabling precise δ¹³C determinations for positional effects in organics like ethanol derivatives. Infrared (IR) and reflection-absorption IR (RAIRS) spectroscopies contribute to KIE assessment by detecting vibrational frequency shifts due to isotopic substitution, which reflect zero-point energy differences influencing reaction rates. These techniques quantify KIEs indirectly from post-reaction peak intensity ratios in isotopically labeled samples, where reduced intensities for heavier isotopomers indicate slower reaction paths; for example, C-H/D labeling shifts CH stretching frequencies by ~50-100 cm⁻¹, allowing calibration of primary KIEs in elimination reactions. RAIRS is advantageous for surface-bound species, providing monolayer sensitivity to vibrational modes altered by isotopes. NMR methods typically achieve 1-5% precision for H/D KIEs, limited by signal-to-noise ratios and relaxation times, while mass spectrometry offers ppm-level accuracy for ¹³C KIEs (e.g., 0.1‰ in δ¹³C, equivalent to ~0.01% in ratios). Recent post-2010 advances in site-specific natural isotope fractionation NMR (SNIF-NMR) have enhanced resolution for ¹³C and ²H, integrating quantitative ¹³C NMR with multivariate analysis to map positional KIEs in complex molecules like pharmaceuticals, improving discrimination of synthetic versus natural origins. These developments build on natural abundance principles for non-destructive, high-fidelity isotopic profiling.

Applications and Examples

Organic and Organometallic Reactions

In nucleophilic substitution reactions, kinetic isotope effects (KIEs) distinguish between SN1 and SN2 mechanisms. For SN2 pathways, which involve backside attack and inversion at the carbon center, primary ^{13}C KIEs at the reacting carbon typically range from 1.05 to 1.10, reflecting significant bonding changes in the transition state. These values arise from the elongation of the C-X bond and formation of the C-Nu bond, as observed in the methanolysis of alkyl halides. In contrast, SN1 reactions proceed via carbocation intermediates, exhibiting secondary α-deuterium KIEs of approximately 1.15 due to the loss of hyperconjugation and change in hybridization from sp^3 to sp^2 at the α-carbon. This normal secondary effect is exemplified in the solvolysis of secondary alkyl sulfonates, where the KIE indicates a loose transition state resembling the carbocation. Elimination reactions further illustrate KIE applications in mechanism elucidation. In E2 processes, which are concerted and involve base abstraction of a β-hydrogen, primary β-H/D KIEs range from 3 to 7, signaling C-H bond cleavage in the rate-determining step. A representative example is the elimination of 1-bromopropane with , yielding a K_H/K_D of 6.7. For E1 mechanisms, β-H/D KIEs are smaller (around 1.4), as the hydrogen abstraction follows the rate-limiting carbocation formation. Additionally, inverse secondary ^{13}C KIEs (approximately 0.98-1.00) at the α- or β-carbon in E2 reactions arise from the sp^3 to sp^2 hybridization change, tightening vibrational modes in the transition state. In organometallic chemistry, KIEs probe steps like and in . Deuterium KIEs for in Pd systems, which often involve oxidative addition to form Pd-C bonds, typically fall between 2 and 5, indicating C-H bond breaking as rate-determining. For instance, in the direct arylation of arenes, a K_H/K_D of 5.5 confirms this step's involvement. Migratory aptitude in , such as the or , influences selectivity, with aryl groups migrating preferentially over alkyl due to lower barriers, though direct KIE measurements are less common; secondary effects help identify insertion as key in product distribution. Recent applications in C-C bond formation highlight KIE utility in confirming transition states. In the of aryl bromides with boronic acids, primary ^{13}C KIEs at the ipso carbon (1.020 under catalytic conditions) pinpoint oxidative addition as the initial irreversible step, aligning with DFT predictions and distinguishing ligand effects on the Pd(0) precatalyst. Similar measurements in transmetalation yield ^{13}C KIEs of 1.035 at the boron-bound carbon, validating a boronate intermediate and aiding optimization of these widely used synthetic methods.

Biological and Enzymatic Systems

In biological systems, kinetic isotope effects (KIEs) are particularly valuable for elucidating the mechanisms of proton and hydride transfers in enzyme-catalyzed reactions, where hydrogen atoms play central roles in biocatalysis. Primary hydrogen/deuterium (H/D) KIEs in hydride transfer steps often exhibit large values, typically ranging from 3 to 7, reflecting substantial differences in zero-point energies and potential quantum tunneling contributions that accelerate the reaction for lighter isotopes. For instance, in horse liver alcohol dehydrogenase (HLADH), which catalyzes the reversible oxidation of alcohols using NAD⁺ as a cofactor, the primary H/D KIE for the hydride transfer from NADH to acetaldehyde is approximately 3.0–3.4, indicating a late transition state with partial C–H bond cleavage. Similar large primary KIEs, up to 5–6 in some variants, are observed in other hydride-transferring enzymes like morphinone reductase and dihydrofolate reductase (DHFR), where the isotope sensitivity highlights the rate-determining nature of the transfer step. Evidence for quantum tunneling in these enzymatic proton transfers emerges from the temperature dependence of primary KIEs, which deviates from classical expectations. In many cases, the KIE shows little variation with temperature (e.g., ΔEₐ ≈ 0 for H/D), suggesting that tunneling through the energy barrier dominates, as the lighter hydrogen isotope more readily penetrates the barrier than deuterium. For example, in thermostable phosphite dehydrogenase (PTDH), temperature-independent primary ²H KIEs of ~2.5–3.0 over 25–60°C confirm a shortened donor-acceptor distance in the tunneling-ready state, enhancing the reaction rate. This phenomenon is widespread in hydride transfers, such as in HLADH and DHFR, where the observed temperature profile aligns with semiclassical models incorporating environmentally coupled vibrations that modulate the barrier width. Secondary β-deuterium KIEs provide insights into conformational dynamics during biosynthetic processes, particularly in enzymes assembling complex carbon frameworks. In polyketide synthases (PKSs), which iteratively condense acyl units to form polyketide natural products, β-secondary KIEs (typically 1.1–1.2) arise from hyperconjugative interactions and steric adjustments at the β-position relative to the transferring bond, probing the enzyme's active-site geometry and substrate orientation. These effects have been leveraged in studies of PKS modules, such as those in spinosyn biosynthesis, to distinguish between syn- and anti-elimination pathways and confirm rigid conformational constraints during chain extension. Solvent KIEs further illuminate proton transfer events in hydrolytic enzymes, where replacing H₂O with D₂O alters rates due to differences in solvent-mediated hydrogen bonding and fractionation factors. In serine proteases like α-chymotrypsin and subtilisin, solvent KIEs (k_{H₂O}/k_{D₂O}) range from 2.0 to 3.0 for the acylation step, reflecting involvement of one or more protons in the transition state, such as in the charge-relay system (Ser-His-Asp triad). Proton inventory plots, which track rate dependence on the deuterium atom fraction (n) in H₂O/D₂O mixtures, often yield linear or convex curves consistent with a single proton transfer, as seen in the hydrolysis of amide substrates where the fractionation factor φ^{TS} ≈ 0.3–0.5 indicates strengthened hydrogen bonds in the transition state. Modern computational and experimental approaches have advanced the study of KIEs in DNA repair enzymes, confirming primary isotope sensitivities in base excision repair pathways. Using hybrid quantum mechanics/molecular mechanics (QM/MM) simulations, recent work on human 8-oxoguanine DNA glycosylase (hOgg1) and formamidopyrimidine-DNA glycosylase (FPG) predicts primary H/D KIEs of 4–6 for the N-glycosidic bond cleavage, driven by proton-coupled electron transfer and tunneling at the lesion site. Cryo-enzymology techniques, combined with variable-temperature KIE measurements down to 100 K, reveal dynamic protein motions that couple to the reaction coordinate in enzymes like uracil DNA glycosylase (UDG), where primary KIEs remain elevated (~3–5) even at cryogenic temperatures, underscoring quantum effects in repair fidelity. These 2020s studies integrate QM/MM free-energy profiles with experimental KIEs to validate mechanisms, highlighting how active-site electrostatics modulate isotope discrimination in genomic maintenance.

Heavy Element and Solvent Effects

Kinetic isotope effects (KIEs) for heavy elements heavier than carbon, such as nitrogen and sulfur, are typically small, ranging from 1.01 to 1.03, due to modest differences in zero-point energies (ZPEs) compared to lighter isotopes like hydrogen or deuterium. For the 15N/14N isotope pair in nucleophilic substitution reactions, such as processes involving leaving groups like p-nitrophenyl, primary KIEs of approximately 1.020 have been measured, reflecting partial bond breaking at the transition state with limited vibrational perturbation. Similarly, in sulfuryl transfer reactions, 34S/32S KIEs around 1.02 arise from enzymatic or chemical cleavages of S-O bonds, where the heavier isotope's lower ZPE leads to slight rate enhancements, as observed in dissimilatory sulfate reduction pathways with fractionations up to 20‰ (equivalent to ε ≈ -20‰). Solvent effects on KIEs are prominent in protic media, where hydrogen/deuterium exchange in water influences reaction rates through solvation and proton transfer steps. In acid-catalyzed hydrolyses, the solvent KIE k_H₂O/k_D₂O typically ranges from 2 to 3, stemming from slower protonation or hydration by D₂O due to its higher ZPE and stronger hydrogen bonding, as seen in general acid catalysis of ester hydrolyses. Isotope fractionation also occurs within solvation shells, where preferential solvation of lighter isotopes around transition states amplifies these effects, particularly in polar protic solvents that stabilize charged intermediates. Compensating variations in KIEs across multiple isotopic sites in complex molecules can be quantified using algorithms that deconvolute site-specific effects from natural abundance measurements, such as the Singleton remote label method for 13C KIEs in multi-site substrates. Inverse solvent KIEs (k_D₂O/k_H₂O > 1) emerge in reactions involving tight pairs, where deuterium's tighter reduces separation of the pair, leading to earlier states and faster rates, as demonstrated in SN2 reactions with ion-pairing substituents in low-polarity solvents. Recent studies in the have applied and KIEs to trace environmental degradation pathways of halogenated pollutants, revealing small primary effects of 1.002 to 1.005 for reductive , which indicate C- or C- bond cleavage as rate-limiting steps in of compounds like . These values, measured via compound-specific , help distinguish abiotic versus transformation mechanisms in contaminated aquifers.

References

  1. [1]
    Kinetic Isotope Effects - Chemistry LibreTexts
    May 4, 2024 · The kinetic isotope effect (KIE) is a phenomenon associated with isotopically substituted molecules exhibiting different reaction rates.Potential Energy Surfaces · Kinetic Isotope Effects · Heavy Atom Isotope Effects
  2. [2]
    [PDF] Kinetic Isotope Effects in Organic Chemistry - Macmillan Group
    Sep 14, 2005 · What Are Kinetic Isotope Effects? □ A kinetic isotope effect is a mechanistic phenomenon wherein isotopically substituted molecules react at.
  3. [3]
    Kinetic isotope effects and how to describe them - PMC - NIH
    Dec 13, 2017 · Kinetic isotope effects (KIEs) are for many chemical processes a useful experimental observable which can be used alongside theoretical analysis ...
  4. [4]
    Harold C. Urey – Facts - NobelPrize.org
    By distilling liquid hydrogen, a hydrogen isotope was extracted in 1932 that was twice as heavy as regular hydrogen. It was called deuterium.Missing: kinetic effect exchange
  5. [5]
    Deuterium Isotope Rate Effects with Citrate-condensing Enzyme
    Bonhoeffer K.F., Salzer F. Naturwissenschaften, 23 (1935), p. 867. View in Scopus Google Scholar. 4. Steacie E.W.R.. Z. Phyzik. Chem., B27 (1934), ... Kinetics ...
  6. [6]
    [PDF] Harold C. Urey - Nobel Lecture
    Due to these discoveries the abundant isotopes of the light elements were known when the existence of the hydrogen isotope of atomic weight two, now called ...
  7. [7]
    [PDF] An Isotope Effect in a Simple Chemical Reaction - eScholarship
    AN ISOTOPE EFFECT IN A SIMPLE CHEFJIICBL REACTIOB. BY. Peter E. Yankwich and Melvin Calvin. July 13, 1948. Berkeley, California. Special Rwiw of 0uAemifi.d ...
  8. [8]
    [PDF] Jacob Bigeleisen - National Academy of Sciences
    He began by formulating in 1949 the extension of the Bigeleisen-Mayer theory to kinetic isotope effects (KIEs), and his analysis showed that KIEs provide a ...
  9. [9]
    Isotope Effects on Reaction Rates - Lars C. S. Melander - Google ...
    Title, Isotope Effects on Reaction Rates Modern concepts in chemistry ; Author, Lars C. S. Melander ; Edition, illustrated ; Publisher, Ronald Press Company, 1960.
  10. [10]
    Kinetic isotope effect - Wikipedia
    A kinetic isotope effect (KIE) is the change in the reaction rate of a chemical reaction when one of the atoms in the reactants is replaced by one of its ...Classification · Theory · Experiments · Case studies
  11. [11]
  12. [12]
    [PDF] Use of Isotopes for Studying Reaction Mechanisms
    In this article we shall discuss the factors determining the magnitude of long range isotope effects on equilibria and rates. Secondary Kinetic Isotope Effects.
  13. [13]
  14. [14]
    The Relative Reaction Velocities of Isotopic Molecules
    Reprints. Search Site. Citation. Jacob Bigeleisen; The Relative Reaction Velocities of Isotopic Molecules. J. Chem. Phys. 1 August 1949; 17 (8): 675–678. https ...
  15. [15]
    Computational physical organic chemistry using the empirical ...
    Using the empirical valence bond approach to explore kinetic isotope effects in proton and hydride transfer reactions. There has been a great deal of ...
  16. [16]
  17. [17]
    The Anharmonicity Correction for Kinetic Isotope Effect Calculations
    The kinetic isotope effect, in the ring opening of a triatomic system, is corrected for anharmonicity in the vibrational ground state. For this system it is ...
  18. [18]
    Critical evaluation of anharmonicity and configurational averaging in ...
    Jul 9, 2020 · Anharmonic effects upon vibrational frequencies and isotopic partition function ratios are modelled computationally by means of quantum mechanics/molecular ...
  19. [19]
    Quantum-instanton evaluation of the kinetic isotope effects
    A general quantum-mechanical method for computing kinetic isotope effects is presented. The method is based on the quantum-instanton approximation for the ...
  20. [20]
    Ab Initio Path-Integral Calculations of Kinetic and Equilibrium ...
    Jun 30, 2015 · The software package Gaussian 09 was used for all the electronic-structure calculations. An essential feature of Gaussian 09 that ...
  21. [21]
    Computation of kinetic isotope effects for enzymatic reactions - NIH
    We describe a computational approach, incorporating quantum mechanics into enzyme kinetics modeling with a special emphasis on computation of kinetic ...
  22. [22]
    The Tunnel Effect in Chemistry
    **Extracted Information:**
  23. [23]
    Hydrogen Tunneling Links Protein Dynamics to Enzyme Catalysis
    ... tunneling. Table 1. Deviations from semiclassical predictions for KIEs. Parameter, Prediction, Observations, References. Primary H/D KIE, <7, >100, 149, 150.
  24. [24]
    Isotope Effects and Heavy-Atom Tunneling in the Roush ... - NIH
    Large kinetic isotope effects (KIEs) in hydrogen transfer reactions have often been interpreted as a diagnostic for tunneling through the reaction barrier.
  25. [25]
    Kinetic isotope effects in Ras-catalyzed GTP hydrolysis - PNAS
    It has been pointed out that a secondary isotope effect near unity, which is observed in the hydrolysis of phosphate monoesters and is reported for Ras- ...Missing: seminal | Show results with:seminal
  26. [26]
    Advances in Kinetic Isotope Effect Measurement Techniques for ...
    We briefly review different experimental approaches for the measurement of KIEs on enzymatic reaction with an emphasis on newer techniques.
  27. [27]
    Highly Precise Measurement of Kinetic Isotope Effects Using 1H ...
    Dec 5, 2012 · We present a technique that uses 13C NMR spectroscopy to measure kinetic isotope effects on the second-order rate const. (kcat/Km) for enzyme- ...
  28. [28]
    Apparent Kinetic Isotope Effects for Multi-Step Steady-State Reactions
    Mar 27, 2025 · The apparent kinetic isotope effect (KIE) for a multistep steady-state reaction can be expressed simply as a sum of terms, one for each transition state (TS) ...
  29. [29]
    Measurement of Kinetic Isotope Effects in an Enzyme-Catalyzed ...
    The methods for measuring KIEs are either competitive or noncompetitive, with all but deuterium isotopes generally being measured by competitive methods due to ...Missing: seminal papers
  30. [30]
    [PDF] An Introduction to Isotopic Calculations
    For a kinetic isotope effect, the fractionation factor is equal to the ratio of the isotope-specific rate constants. It is sometimes denoted by β rather than α.
  31. [31]
    [PDF] Kinetic Isotope Effect: Principles and its use in mechanism ... - EPFL
    Enzyme catalyzing oxidation of amino acids to imino acids using FAD as coenzyme. • Difficulty to analyze the direct reaction with amino acid carbanion. → ...Missing: Bonhoeffer 1935
  32. [32]
    Sensitive and Accurate 13C Kinetic Isotope Effect Measurements ...
    Scheme 1. In general, SN1 mechanisms give rise to small normal or inverse 12C/13C KIEs, while SN2 mechanisms produce relatively large normal KIEs (Scheme 1).
  33. [33]
    Isotopic Fractionation and Kinetic Isotope Effects of a Purified ...
    Oct 10, 2025 · 4 Calculation of the KIE for 15Nα using the Expanded Rayleigh model yields a normal KIE value of 1.0072 ± 0.0010 and an ε value of −7.2 ± 1.0‰ ...Missing: formula | Show results with:formula
  34. [34]
    Molecular Average and Intramolecular δ 13 C Measurements of ...
    Aug 26, 2025 · We developed a novel method using gas chromatography – Orbitrap mass spectrometry (GC Orbitrap MS) to measure the intramolecular carbon isotope ...Missing: competitive | Show results with:competitive
  35. [35]
    Determination of Stretching Frequencies by Isotopic Substitution ...
    May 18, 2023 · This lab allows students the opportunity to collect Fourier transform infrared (FTIR) data and analyze the effect of isotopically labeling C–H ...Missing: RAIR | Show results with:RAIR
  36. [36]
    Application of NMR Spectroscopy in Isotope Effects Studies
    KIE is defined as kL/kH ratio, where kL and kH are rate constants of reaction with light (L) and heavy (H) isotopomer, respectively. KIE can be calculated from ...
  37. [37]
    Stable Isotope Ratio Mass Spectrometry and Site-Specific Natural ...
    In this review, we will discuss the effectiveness of stable isotope ratio mass spectrometry and/or Site-Specific Natural Isotope Fractionation-Nuclear Magnetic ...
  38. [38]
    [PDF] 678738.pdf
    Interpretation of 13C- and 37Cl-AKIEs for oxidation and SN2-type reactions ... the typical 13C-KIE range for a SN2 reaction (1.03 – 1.09)21 (Scheme 1). Abe ...
  39. [39]
    Secondary deuterium isotope effects in solvolysis of cyclopentyl p ...
    Secondary deuterium isotope effects in solvolysis of cyclopentyl p-bromobenzenesulfonate. Stereochemistry of E1 and SN1 reactions. Click to copy article link ...Missing: seminal | Show results with:seminal
  40. [40]
    An unusually large α-secondary deuterium kinetic isotope effect.
    Secondary α-deuterium KIEs arise when hydrogen(s) at the α- or reacting carbon are replaced by deuterium. Solvolysis reactions proceeding via a carbenium ...Missing: seminal | Show results with:seminal
  41. [41]
  42. [42]
  43. [43]
    Elusive transition state of alcohol dehydrogenase unveiled - PMC
    The hydride transfer catalyzed by alcohol dehydrogenase has been difficult to understand. Here we add to the large corpus of anomalous and paradoxical data.
  44. [44]
    Enzymatic Kinetic Isotope Effects from First-Principles Path ... - NIH
    Experimentally, the primary H/D kinetic isotope effect of benzyl alcohol conversion via YADH has been determined to be 3.4. Although the kinetic isotope effect ...
  45. [45]
    Primary Deuterium Kinetic Isotope Effects: A Probe for the Origin of ...
    Jun 21, 2018 · Large primary deuterium kinetic isotope effects (1? DKIEs) on enzyme-catalyzed hydride transfer may be observed when the transferred hydride ...Scheme 1 · Experimental Section · Figure 7
  46. [46]
    Temperature-Independent Kinetic Isotope Effects as Evidence for a ...
    Sep 19, 2019 · We have evaluated the role of hydride tunneling in a thermostable variant of this enzyme (17X-PTDH) by measuring the temperature dependence of the primary 2 H ...
  47. [47]
    Large kinetic isotope effects in enzymatic proton transfer and ... - PNAS
    The temperature dependence of the kinetic isotope effect is caused by a thermally excited substrate mode that modulates the barrier, as has been suggested ...
  48. [48]
    Variable temperature kinetic isotope effects demonstrate extensive ...
    Jun 12, 2023 · Proton tunneling is expected to increase the magnitude of kH and decrease its temperature dependence. This would result in larger KIEs and an ...
  49. [49]
    Evidence for an Enzyme Catalyzed Rauhut-Currier Reaction ... - NIH
    The SpnL enzyme catalyzes a biological Rauhut-Currier reaction during spinosyn A biosynthesis, making it unique in nucleophile-assisted C-C bond formation.
  50. [50]
    Solution Structure and Conformational Flexibility of a Polyketide ...
    Oct 18, 2021 · The results of this hybrid approach shed light on the solution structure of a cis-AT type PKS module as well as its inherent conformational dynamics.
  51. [51]
    The role of the medium in solvent isotope effects on serine protease ...
    Solvent isotope effects kcat(H2O)/ kcat(D2O) range from 2.1 to 3.0 with proton inventories (dependences of kcat(n) on n, the atom fraction of deuterium in the ...
  52. [52]
    Solvent Isotope Effects in H2O−D2O Mixtures (Proton Inventories ...
    The solvent isotope effects (proton inventories) determined by these data showed no differences between the wild-type and mutant enzymes, despite the fact that ...
  53. [53]
    Proton Bridging in Catalysis by and Inhibition of Serine Proteases of ...
    Solvent Isotope Effect and Proton Inventory Measurements and Techniques. Enzyme activity was monitored before and after each experiment using active-site ...Missing: KIE | Show results with:KIE
  54. [54]
    Unlocking the chemistry facilitated by enzymes that process nucleic ...
    Kinetic Isotope Effect Studies of the Reaction Catalyzed by Uracil DNA ... QM/MM MD Simulations on the Enzymatic Pathway of the Human Flap Endonuclease ...
  55. [55]
    QM/MM MD Study on the Reaction Mechanism of Thymidine ...
    Jun 17, 2024 · Here, we report mechanistic studies on type II thymidine kinase, Thermotoga maritima TmTK, aiming to predict barriers for the enzyme-catalyzed reaction.
  56. [56]
    Isotope effects in nucleophilic substitution reactions. III. The effect of ...
    A Hammett p value of +2.04 when the Z substituent in the leaving group was varied and a nitrogen kinetic isotope effect of 1.0200 showed that the Ca-N bond was ...
  57. [57]
    32S/34S isotope fractionation in plant sulphur metabolism - Tcherkez
    May 21, 2013 · Here, we review documented 32 S/ 34 S isotope fractionations that may apply to sulphur metabolism, and explain how they should yield disparities.
  58. [58]
    [PDF] Kinetic Isotope Effects in the Study of Organometallic Reaction ...
    May 5, 2011 · CONTENTS. 1. Introduction. 4857. 2. Basis of Kies. 4858. 2.1. Origin of Isotope Effects. 4858. 2.2. Magnitude of the Observed KIEs.Missing: Bonhoeffer 1935
  59. [59]
    (PDF) The proton inventory technique in a dual mechanistic system
    Aug 9, 2025 · The kinetic solvent isotope effects (k H/k D) on the water and H 2PO 4- catalyzed hydrolysis are 2.86 and 2.00, respectively. The observed ...
  60. [60]
    [PDF] Heavy-Atom Kinetic Isotope Effects: Primary Interest or Zero Point?
    ABSTRACT: Chemists have many options for elucidating reaction mechanisms. Global kinetic analysis and classic transition state probes (e.g., LFERs, ...
  61. [61]
    Isotope effects in nucleophilic substitution reactions. VI. The effect of ...
    In all three solvents in which the form of the ionic reactant changes, a smaller isotope effect and tighter transition state are found for the reaction with ...
  62. [62]
    Bromine isotope effects: Predictions and measurements
    Measured and predicted bromine kinetic isotope effects for reported transformation pathways of studied brominated contaminants. Compound, Reaction, Mechanism ...Review · 4.1. Solvolysis Of N-Butyl... · 5.1. 1. Ethylene Dibromide...
  63. [63]
    Dual C–Br Isotope Fractionation Indicates Distinct Reductive ...
    Multi-element compound-specific isotope analysis bears the potential to elucidate the biodegradation pathways of 1,2-DBA in the environment, which is crucial ...Scheme 1 · Figure 2 · Scheme 2