Fact-checked by Grok 2 weeks ago

Reaction intermediate

In , a reaction intermediate is a transient, short-lived formed during the multi-step of a , produced in an early elementary step and rapidly consumed in a subsequent step to ultimately yield the final products, without appearing in the overall balanced equation. These intermediates play a crucial role in elucidating reaction pathways, as their identification helps chemists predict reactivity, , and in processes ranging from to biochemical transformations. Common types of reaction intermediates in include carbocations, carbanions, free radicals, and carbenes, each characterized by unique electronic structures and reactivity profiles. Carbocations are positively charged carbon species with an empty p-orbital, making them electrophilic and planar, with stability increasing from primary to tertiary due to and inductive effects from alkyl groups. In contrast, carbanions feature a negatively charged carbon with a , rendering them nucleophilic and typically pyramidal, though their stability decreases with more substituted carbons owing to electron repulsion. Free radicals, neutral species with an unpaired electron, exhibit high reactivity and are often generated in chain reactions like , with stability similarly enhanced by adjacent alkyl groups. Carbenes, divalent carbon atoms with six valence electrons, can act as either electrophiles or nucleophiles depending on substituents, and are key in reactions such as . Beyond these, other notable intermediates include nitrenes (monovalent nitrogen species analogous to carbenes, involved in reactions like aziridination) and (aryne intermediates with diradical character, key in ). Due to their fleeting existence—lifetimes varying from picoseconds to milliseconds depending on the species—reaction intermediates are rarely isolated but can be detected using advanced techniques such as (e.g., for radicals) or computational modeling, providing insights into energy barriers and transition states. Understanding these species is essential for designing efficient catalysts and predicting side reactions in industrial and biological contexts, such as enzyme mechanisms in where phosphorylated sugars serve as intermediates.

Definition and Fundamentals

IUPAC Definition

According to the International Union of Pure and Applied Chemistry (IUPAC), a is defined as a with a lifetime appreciably longer than a (corresponding to a local minimum of depth greater than RT) that is formed (directly or indirectly) from the reactants and reacts further to give (either directly or indirectly) the products of a ; this includes the corresponding . This definition was established in the IUPAC Recommendations for published in 1994, providing a standardized for transient in reaction pathways. A key distinction exists between reaction intermediates and transition states: intermediates correspond to potential energy minima and possess a measurable finite lifetime, whereas transition states represent energy maxima on the with lifetimes shorter than a single , lacking discrete existence as stable entities. In , reaction intermediates function as transient species within composite reaction mechanisms, often analyzed using approximations like the steady-state hypothesis to derive overall rate laws, where their concentrations are assumed constant over much of the reaction progress.

Key Characteristics and Stability

Reaction intermediates possess a transient , existing only briefly during chemical transformations, with typical lifetimes spanning from picoseconds (10^{-12} seconds) to milliseconds (10^{-3} seconds), influenced by the specific type and environmental conditions. This short duration arises because intermediates are rapidly consumed in subsequent reaction steps, preventing their accumulation. The lifetime \tau of an intermediate can be estimated using the equation \tau = \frac{1}{k} where k represents the rate constant for its decay, derived from first-order kinetics principles applicable to unimolecular decomposition or reaction processes. In potential energy surface (PES) representations of reactions, intermediates correspond to local minima, distinct from the global minima of reactants and products, and are separated from these by activation energy barriers that dictate the reaction pathway. These minima reflect temporary energy wells where the intermediate achieves a fleeting stability before overcoming the subsequent barrier to form the next species or product. Such depictions, often visualized in energy diagrams, underscore how intermediates occupy higher-energy states than the overall reaction endpoints, facilitating multi-step mechanisms. The stability of reaction intermediates is governed by several electronic and environmental factors, including , which delocalizes electrons through sigma-bond overlap to lower energy; , enabling charge or electron distribution across conjugated systems; and inductive effects from adjacent substituents that either donate or withdraw . Additionally, play a crucial role, as polar protic solvents can stabilize charged intermediates via and hydrogen bonding, thereby extending their persistence compared to non-polar media. Unlike reactants and products, which are stable, isolable species present at the start and end of a reaction and appearing in the balanced equation, intermediates are not isolable under standard conditions due to their high reactivity and short lifetimes, though they can be trapped or indirectly observed through specialized techniques. This distinction highlights their role as ephemeral bridges in the mechanistic pathway, essential for understanding reaction kinetics without being detectable as primary components.

Classification of Reaction Intermediates

Ionic Intermediates

Ionic intermediates are charged species that play a crucial role in many chemical reactions, particularly in organic and inorganic mechanisms. These intermediates include carbocations, which are positively charged carbon atoms with three substituents and an empty p-orbital, denoted as R₃C⁺; carbanions, negatively charged carbon atoms with a and three substituents, R₃C⁻; and oxonium ions, oxygen-containing cations such as alkyloxonium ions (R₃O⁺) where oxygen bears a positive charge and three bonds. These species form primarily through heterolytic bond cleavage, where a breaks unevenly, with both electrons transferring to one atom, generating oppositely charged fragments, or via processes such as or . For instance, the general heterolysis of an alkyl halide can be represented as: \text{R--X} \rightarrow \text{R}^+ + \text{X}^- This process is endothermic and typically requires , often facilitated by polar solvents or catalysts that stabilize the charges. Electron transfer mechanisms involve the gain or loss of an electron to or from a species, leading to radical ions that may further dissociate into ionic intermediates. Due to their charges, ionic intermediates exhibit distinct reactivity patterns: carbocations and other cations act as electrophiles, seeking electron-rich sites for nucleophilic attack, while carbanions and anions function as nucleophiles, donating electrons to electrophilic centers. This polar reactivity contrasts with the non-directional behavior of neutral intermediates. Stability of these species varies significantly; for carbocations, the order follows > secondary > primary, attributed to and inductive effects from alkyl substituents that donate to the electron-deficient carbon. In contrast, for carbanions, stability follows the order primary > secondary > , as alkyl groups exert a destabilizing inductive (+I) effect on the negative charge. Oxonium ions similarly gain through or interactions. In , the (NO₂⁺), a linear species with nitrogen bearing the positive charge, exemplifies an ionic intermediate in reactions like , where it forms from the interaction of nitric and sulfuric acids and attacks electron-rich aromatic rings. This ion's reactivity underscores the role of ionic intermediates in facilitating regioselective transformations across diverse chemical contexts.

Radical Intermediates

Radical intermediates, also known as free , are neutral species characterized by the presence of one or more electrons, resulting in high reactivity due to their open-shell . Examples include the methyl (•CH₃) and the atom (•Cl), which possess a single unpaired electron in a p-orbital, often represented with a to denote the radical site. These intermediates are transient and typically short-lived in solution or gas phase, though some, like (NO•), exhibit greater persistence. Free radicals form primarily through homolytic cleavage of covalent bonds, where each fragment retains one electron from the shared pair, often induced by , , or ; for instance, the bond dissociation of Cl₂ requires 243 kJ/mol to yield two •Cl atoms. They can also arise from processes, such as one-electron transfer reactions that generate radical ions or neutral s from stable precursors. Due to the , radicals exhibit high reactivity, favoring addition to unsaturated bonds or hydrogen abstraction from saturated molecules to achieve an octet configuration. The stability of carbon-centered radicals follows the order tertiary > secondary > primary, attributed to hyperconjugation where adjacent C-H or C-C σ bonds donate electron density to the half-filled p-orbital, with tertiary radicals benefiting from three such alkyl substituents. Allylic and benzylic radicals are further stabilized by resonance delocalization of the unpaired electron into adjacent π-systems, making them more stable than simple tertiary radicals; for example, a benzylic radical can distribute its electron density across five resonance structures involving the aromatic ring. In many reactions, radicals participate in chain processes consisting of , , and termination steps. Initiation generates the first s, often via homolysis of a precursor like Cl₂ under UV light to form •Cl. Propagation maintains the radical count through cyclic steps, such as in : \ce{R-H + X^\bullet -> R^\bullet + H-X} \ce{R^\bullet + X2 -> R-X + X^\bullet} where X is a atom, allowing the chain to continue efficiently. Termination occurs when two radicals combine, such as 2•Cl → Cl₂, reducing the radical population and halting the chain. Inorganic radicals, such as the (•OH), play crucial roles in ; in the , •OH forms via photolysis of followed by reaction with and acts as the primary oxidant, cleansing the atmosphere by reacting with trace gases like and .

Neutral Reactive Intermediates

Neutral reactive intermediates encompass uncharged, highly reactive molecular species featuring atoms with incomplete octets or elevated energy states, distinguishing them from charged or counterparts. Prominent examples include carbenes, which contain a divalent carbon with only six electrons, such as methylene (:CH₂), and nitrenes, characterized by a monovalent nitrogen also with six electrons, exemplified by (:NH). These species are inherently unstable due to their electron deficiency, leading to fleeting lifetimes under standard conditions unless stabilized by specific substituents. The electronic configuration of neutral reactive intermediates allows them to adopt either singlet or triplet spin states, profoundly affecting their reactivity profiles. In the singlet state, the non-bonding electrons occupy the same orbital in a paired, closed-shell arrangement, imparting electrophilic character that facilitates concerted reactions like insertions and additions. Conversely, the triplet state features two unpaired electrons in separate orbitals, resembling a diradical with milder reactivity, often involving stepwise mechanisms or hydrogen abstraction. The singlet-triplet energy gap (ΔE_ST) typically ranges from 10 to 40 kcal/mol depending on substituents, with electron-withdrawing groups favoring the singlet ground state in carbenes, while donor groups stabilize the triplet; this gap dictates the dominant state and thus the synthetic pathway. Generation of these intermediates commonly proceeds via photolysis, thermolysis, or elimination processes that expel a to form the electron-deficient center. For carbenes, α-elimination from polyhalomethanes using a strong base is a standard method, as illustrated by the formation of dibromocarbene from : \text{CHBr}_3 + \text{base} \rightarrow : \text{CBr}_2 + \text{HBr} This reaction proceeds through to a haloform anion followed by expulsion, yielding a carbene under typical conditions. Nitrenes are analogously produced by thermal or photochemical decomposition of azides (R-N₃ → R-N: + N₂), often generating triplet states that may interconvert to . Reactivity of neutral reactive intermediates centers on their ability to undergo insertion into σ-bonds (e.g., C-H or Si-H), cycloaddition to π-systems (e.g., forming cyclopropanes from alkenes), or intramolecular rearrangements, with states enabling stereospecific, concerted processes and triplets promoting biradical-like paths. Substituent effects significantly modulate stability and selectivity; , such as :CBr₂ or :CCl₂, exhibit enhanced persistence relative to :CH₂ due to halogen stabilization through inductive withdrawal and π-donation, enabling room-temperature applications in . Nitrenes display comparable versatility, inserting into C-H bonds for or adding to alkenes for aziridination, with their reactivity tuned by metal coordination in catalytic cycles. Another important example is benzyne (dehydrobenzene), a highly strained intermediate featuring a formal carbon-carbon in a six-membered ring, generated typically by ortho-elimination from aryl halides under strong basic conditions. Benzynes act as electrophiles in nucleophilic additions, leading to products, and their reactivity is influenced by the orthogonal π-bonds that limit stabilization. In inorganic contexts, atomic oxygen (O) exemplifies a intermediate in , where its ground state abstracts hydrogen from hydrocarbons to propagate oxidation chains, influencing flame speeds and formation. These properties underpin their utility in for efficient bond construction, as seen in carbene-mediated for pharmaceutical intermediates.

Role in Reaction Mechanisms

Intermediates in Electrophilic Additions

In reactions, an initially adds to the π-bond of an , forming a such as a or a bridged , which is then attacked by a to yield the final product. This two-step ensures and , with the intermediate's structure dictating the reaction pathway. A classic example is the addition of (HBr) to an , where the proton (H⁺) acts as the and adds to the less substituted carbon of the , generating a on the more substituted carbon. This follows , which states that the hydrogen attaches to the carbon with more hydrogens, resulting in the more stable —typically tertiary > secondary > primary—due to and inductive effects from alkyl groups. For propene (CH₃CH=CH₂ + HBr), the is the secondary CH₃CH⁺CH₃, leading to as the major product. In contrast, halogen addition to s, such as (Br₂), proceeds via a bridged intermediate rather than a classical , which shields one face of the molecule and promotes anti addition while preventing rearrangements. The electrophilic Br⁺ approaches the π-bond, forming a three-membered ring with the carbons, followed by nucleophilic attack from the bromide ion on the opposite side. For , the reaction is: \ce{CH2=CH2 + Br2 -> [CH2-CH2Br]+ Br- -> CH2Br-CH2Br} This bromonium ion intermediate ensures stereospecific trans addition, as confirmed by the formation of meso-2,3-dibromobutane from trans-2-butene. In unsymmetrical alkenes undergoing carbocation-mediated additions like acid-catalyzed hydration, the intermediate can undergo rearrangements such as 1,2-hydride or alkyl shifts to form a more stable carbocation, altering the product distribution. For 3-methyl-1-butene with HBr, the initial secondary carbocation rearranges via a hydride shift to a tertiary one, yielding 2-bromo-2-methylbutane predominantly. These shifts highlight the dynamic nature of carbocation intermediates in achieving thermodynamic stability.

Intermediates in Nucleophilic Substitutions and Eliminations

In reactions, the typically features a attached to a carbon atom, which is attacked by a , leading to replacement. These reactions proceed via two primary mechanisms: SN1 and SN2. The SN1 mechanism involves a rate-determining step where the departs, forming a intermediate, followed by rapid nucleophilic attack. For example, the of with ion proceeds through this pathway: the chloride leaves to generate the tert-butyl , which then combines with OH⁻ to form tert-butanol. The planar nature of the intermediate allows nucleophilic attack from either face, resulting in if the is chiral. The overall SN1 can be represented as: \mathrm{R_3C-Cl \xrightarrow{k_1} R_3C^+ + Cl^-} \mathrm{R_3C^+ + Nu^- \xrightarrow{k_2} R_3C-Nu} where the first step is rate-determining, making the reaction unimolecular. In contrast, the SN2 is concerted, occurring in a single step without a discrete ; the attacks the carbon from the backside as the departs simultaneously. This process features a resembling a pentacoordinate , with inversion of configuration at the carbon center. Nucleophilic elimination reactions, which form alkenes by removing a and a β-hydrogen, also follow unimolecular (E1) or bimolecular (E2) pathways. The E1 mechanism shares the intermediate with SN1, where from an adjacent carbon occurs after carbocation formation. A classic example is the acid-catalyzed of alcohols, such as tert-butanol, where of the OH group facilitates water departure to form the tert-butyl , followed by loss of a proton to yield isobutene. The E2 mechanism, however, is concerted and requires anti-periplanar alignment of the and β-hydrogen for optimal orbital overlap in the , proceeding without intermediates. SN1 and E1 pathways often compete when carbocations form, as the intermediate can partition between nucleophilic capture () and (elimination). play a key role in this competition: polar protic solvents stabilize the and leaving group ions, favoring SN1/E1 over SN2/E2, while also enhancing elimination by solvating the base weakly. In such media, higher temperatures further promote E1 over SN1 by increasing the likelihood of .

Detection and Characterization

Spectroscopic Techniques

Spectroscopic techniques play a crucial role in detecting and characterizing reaction intermediates, particularly transient with short lifetimes, by providing structural and dynamic information through their unique spectral signatures. These methods exploit interactions between intermediates and , allowing identification of electronic, vibrational, or magnetic properties that distinguish them from stable molecules. Time-resolved variants are essential for capturing fleeting , often with resolutions down to femtoseconds, enabling measurement of lifetimes and reaction pathways. Ultraviolet-visible (UV-Vis) is widely used to detect ionic intermediates like carbocations via their characteristic bands arising from π-π* transitions. For instance, the trityl cation (Ph₃C⁺) exhibits a strong maximum at approximately 435 nm in acidic media, attributed to its delocalized positive charge across the phenyl rings. This shift from the UV region of neutral (around 260 nm) confirms the formation of the cationic intermediate during solvolysis or reactions. Electron paramagnetic resonance (EPR, also known as ESR) spectroscopy detects radical intermediates by measuring the magnetic interactions of unpaired electrons with nuclear spins, producing hyperfine splitting patterns. The methyl radical (•CH₃) displays a quartet spectrum due to three equivalent hydrogen nuclei (I = 1/2 each), with an isotropic hyperfine coupling constant a_H of 23 G, a hallmark signature observed in gas-phase or matrix-trapped samples. This technique is particularly sensitive for concentrations as low as 10⁻⁹ M and has been instrumental in confirming radical mechanisms in chain reactions. Infrared (IR) and Raman spectroscopy provide vibrational signatures for neutral reactive intermediates such as carbenes and nitrenes, revealing bond strengths and geometries through characteristic stretching or bending modes. For the methylene carbene (:CH₂), matrix-isolated IR spectra show asymmetric CH₂ stretching at around 2870 cm⁻¹ for the triplet state, distinct from alkane C-H stretches near 2900 cm⁻¹, indicating the divalent carbon's sp² hybridization. Similarly, Raman spectroscopy has been applied to triplet arylnitrenes, where time-resolved measurements capture N=C stretching modes near 1300 cm⁻¹ during photolysis, aiding identification of their diradical character. Time-resolved spectroscopic methods, such as laser flash photolysis, are vital for studying intermediates with lifetimes shorter than 1 μs by generating them via pulsed excitation and monitoring decay kinetics. This technique, pioneered in the , records transient absorption spectra with resolution, quantifying rate constants for reactions like recombination (often 10⁶–10⁹ M⁻¹ s⁻¹). For example, it has measured the lifetime of nitrenium ions at around 10–100 μs in aqueous solutions, linking spectral changes to effects. Matrix isolation complements this by trapping intermediates in inert matrices at 4–20 K, stabilizing them for extended IR or UV-Vis analysis; photolyzed in matrices yields persistent :CH₂ spectra for detailed characterization. Nuclear magnetic resonance (NMR) techniques, including low-temperature methods and chemically induced dynamic nuclear polarization (CIDNP), enable detection of through enhanced signals from spin polarization during radical pair formation. CIDNP produces anomalous emission or absorption lines in ¹H NMR spectra due to unequal population of nuclear spin states in escaping , with examples like the photoinduced radical pairs in flavin systems showing polarized methyl signals at δ 2.5 ppm. This hyperpolarization boosts by factors of 10³–10⁴, allowing observation of transient at micromolar concentrations without direct detection. A specific application of involves the observation of the phenyl radical (C₆H₅•), evidenced by its characteristic from , , and hydrogens (a_H ≈ 6–18 G). This detection confirms radical intermediates in aromatic reaction pathways.

Computational and Theoretical Methods

Computational and theoretical methods play a pivotal role in elucidating the structures, stabilities, and reactivities of reaction intermediates, which are often too short-lived for direct experimental observation. These approaches, rooted in , enable the prediction of energy profiles along reaction coordinates, revealing the positions of intermediates relative to reactants, products, and transition states. methods, which solve the without empirical parameters, and (DFT), which approximates electron correlation via exchange-correlation functionals, are cornerstone techniques for such analyses. For instance, ab initio calculations at the MP2 level have been used alongside DFT to explore the structures of alkyl carbocations, confirming bridged versus open forms based on correlation effects. A key application involves mapping potential energy surfaces (PES), multidimensional landscapes that depict the energy as a function of nuclear coordinates. On a PES, reaction intermediates correspond to local minima, distinct from the saddle points representing transition states, allowing theorists to delineate stepwise mechanisms. Geometry optimizations and transition state searches, often performed using gradient-based algorithms, facilitate the location of these features; for example, ab initio methods have been employed to construct PES for ion-molecule reactions, identifying intermediates like HPSH⁺ as global minima from which products emerge. Validation of these computations frequently involves comparing predicted properties, such as vibrational frequencies or barrier heights, with sparse experimental data, ensuring reliability in mechanistic predictions. For dynamic aspects, particularly in radical intermediates, (MD) simulations provide insights into time-dependent behaviors like steps in solution. Reactive classical MD, synergized with potentials, has been used to model free- , capturing chain growth and termination events that involve transient radical species. These simulations reveal solvent effects on and reactivity, complementing static quantum calculations by incorporating and entropy contributions. Computed lifetimes of intermediates offer a direct benchmark against experimental measurements, enhancing method validation. For the methylene (:CH₂), a prototypical , high-level calculations predict rates that align with its observed singlet lifetime on the order of nanoseconds, underscoring the accuracy of quantum mechanical treatments for spin-forbidden processes. Such comparisons guide the selection of basis sets and correlation levels for reliable predictions. Software packages like Gaussian facilitate these electronic structure computations, supporting a range of and DFT methods for optimizing intermediate geometries and scanning PES. For larger systems, such as enzyme-bound intermediates, the ONIOM (Our own N-layered Integrated and ) method partitions the system into layers treated at varying theory levels, enabling hybrid treatments of reactivity in complex environments. Historically, the 1970s marked a with early Hartree-Fock applications to simple carbocations, like the ethyl cation, which provided initial theoretical support for nonclassical structures proposed experimentally.

Applications and Significance

Biological Reaction Intermediates

In biological systems, reaction intermediates play crucial roles in enzymatic and metabolic pathways, enabling selective transformations under physiological conditions. Enzymes often stabilize these transient species through interactions, facilitating reactions that would otherwise be too slow or nonspecific. In metabolic processes like and biosynthesis, intermediates such as carbocations, radicals, and enediols are precisely controlled to ensure efficiency and avoid side reactions. Enzyme-bound carbocation-like intermediates are prominent in terpene biosynthesis, where cyclases orchestrate the folding and cyclization of linear precursors like farnesyl pyrophosphate (FPP). In the cyclization of FPP to form cyclic terpenoids, the enzyme initiates ionization to generate a delocalized carbocation, which propagates through a series of rearrangements stabilized by aromatic residues and metal ions in the active site, promoting regioselective ring closure. For instance, terpene synthases use π-electron interactions from tyrosine or tryptophan residues to delocalize the positive charge, enhancing the lifetime of these high-energy intermediates and directing stereospecific product formation. Similarly, in B12-dependent methionine synthase, the methyl transfer from methylcobalamin to homocysteine involves a protein radical cage formed by nearby residues, which suppresses homolysis to methyl (•CH3) and cobalamin radicals, ensuring controlled organometallic transfer while mitigating unwanted radical reactivity. Reactive oxygen species (ROS) such as (•O2⁻) and (•OH) serve as key intermediates in responses and signaling pathways. These radicals arise from partial reduction of molecular oxygen during enzymatic reactions, like those catalyzed by NADPH oxidases, and act as diffusible intermediates that propagate signals or cause cellular damage when dysregulated. In enzymatic contexts, proteins stabilize intermediates to enable carbon-carbon bond formation; for example, in class I fructose-1,6-bisphosphate aldolase, residues including (forming a ), , and provide electrostatic stabilization and hydrogen bonding to the derived from , extending its lifetime for with glyceraldehyde-3-phosphate. A example occurs in , where phosphoglucose catalyzes the interconversion of glucose-6-phosphate (G6P) and fructose-6-phosphate (F6P) via a cis-enediol : \ce{(CH2OPO3^{2-})(CHOH)_4(CHO) ->[PGI] [enediol intermediate] -> (CH2OPO3^{2-})(CO)(CHOH)_3(CH2OH)} This proton abstraction-addition mechanism, facilitated by a glutamate residue as the base, ensures reversible isomerization without accumulation of the high-energy enediol. Uncontrolled accumulation of reactive intermediates, particularly ROS like •O2⁻ and •OH, can lead to oxidative damage, including DNA strand breaks and base modifications that contribute to mutagenesis and diseases such as cancer and neurodegeneration. Enzymatic control thus underscores the significance of these intermediates in maintaining metabolic fidelity while highlighting their potential toxicity when stabilization fails.

Industrial and Synthetic Applications

In industrial production, intermediates play a central role in free-radical vinyl polymerization processes, such as the synthesis of from styrene . The begins with the thermal or photochemical of an initiator to generate , which then add to the vinyl double bond of styrene, forming a carbon-centered that propagates the chain by successive additions, ultimately yielding high-molecular-weight used in and . This is particularly efficient in , where water-soluble initiators produce that enter monomer-swollen micelles to initiate growth. A key step involves the to the , as represented by: \text{M} + \text{I}^\bullet \rightarrow \text{IM}^\bullet followed by chain to form the growing . This method enables high yields and controlled particle sizes, producing latexes for paints and adhesives on a multimillion-ton scale annually. In catalytic olefin polymerization, such as the Ziegler-Natta process for and , alkyltitanium intermediates facilitate stereoselective monomer insertion. The , typically a titanium chloride compound activated by an organoaluminum cocatalyst, generates alkyltitanium species that coordinate olefins, enabling migratory insertions to build linear, isotactic chains with minimal branching. This process accounts for approximately 240 million tons of polyolefins produced yearly as of 2025, underpinning plastics for automotive and consumer goods. Pharmaceutical leverages intermediates for selective C-H insertions, enabling the construction of complex drug scaffolds from simple precursors. Transition-metal-catalyzed decomposition generates carbenoids that insert into unactivated C-H bonds with high stereocontrol, as seen in the of agents and inhibitors, where yields exceed 80% for key bond-forming steps. This approach minimizes synthetic steps and waste, accelerating routes to molecules like those in protease inhibitors. Process optimization in large-scale operations often involves managing intermediates to suppress side reactions and enhance selectivity. In the cumene process for phenol and acetone production, air oxidation of cumene generates isopropylbenzene s that form as the key , which is then cleaved under acidic conditions; careful control of avoids to unwanted byproducts like , achieving over 95% selectivity. Trapping or stabilizing such intermediates through additives or temperature regulation ensures high-purity outputs in this cornerstone of the resins industry. The strategic exploitation of intermediates drives in , enabling high-yield routes that convert feedstocks like into commodities with margins improved by 20-30% through optimized mechanisms. For instance, and pathways in and oxidation processes support a global market valued at over $500 billion, reducing energy costs and enabling scalable production of derivatives like plastics and solvents.

References

  1. [1]
    3.2.5: Reaction Intermediates - Chemistry LibreTexts
    Feb 12, 2023 · A reaction intermediate is transient species within a multi-step reaction mechanism that is produced in the preceding step and consumed in a subsequent step.
  2. [2]
    5.6. Reactive intermediates | Organic Chemistry 1: An open textbook
    In chemistry, a reactive intermediate or an intermediate is a short-lived, high-energy, highly reactive molecule. When generated in a chemical reaction, ...<|control11|><|separator|>
  3. [3]
    Reaction Intermediate - an overview | ScienceDirect Topics
    Reaction intermediates are defined as short-lived species that can be detected experimentally, which form during the steps of a reaction.
  4. [4]
  5. [5]
  6. [6]
    lifetime (L03515) - IUPAC Gold Book
    Mathematical definition: τ = 1 k = 1 ∑ i k i with k i the first-order rate constants for all decay processes of the decaying state. · Lifetime is used sometimes ...
  7. [7]
    Potential Energy Surface - an overview | ScienceDirect Topics
    Potential energy surfaces may be roughly classified into those with deep enough minima to support one or more strongly bound vibrational states and those ...
  8. [8]
    Hyperconjugation: A More Coherent Approach - ACS Publications
    May 15, 2012 · The stability of such anions in the gas phase (free from solvent effects, counterion effects, etc.) increases as methyl substitution increases.
  9. [9]
    3 Factors That Stabilize Carbocations - Master Organic Chemistry
    Jan 29, 2025 · This is usually done in a highly polar solvent such as H2O or acetic acid which can help to stabilize the charged carbocation intermediate.
  10. [10]
  11. [11]
    5.9: Reactive Intermediates- Carbanions and Carbon Acids
    May 30, 2020 · A carbanion is an anion in which carbon has an unshared pair of electrons and bears a negative charge usually with three substituents for a total of eight ...
  12. [12]
    7.9 Carbocation Structure and Stability - Organic Chemistry | OpenStax
    Sep 20, 2023 · As shown in Figure 7.11, tertiary alkyl halides dissociate to give carbocations more easily than secondary or primary ones. Thus, trisubstituted ...
  13. [13]
    9.1: Homolytic and Heterolytic Cleavage - Chemistry LibreTexts
    Dec 15, 2021 · Heterolytic cleavage takes both electrons of a bond, while homolytic cleavage takes one electron from each part of the bond.
  14. [14]
    Explaining the nitration of benzene - electrophilic substitution
    If you are going to substitute an -NO2 group into the ring, then the electrophile must be NO2+. This is called the "nitronium ion" or the "nitryl cation", and ...<|separator|>
  15. [15]
    Nitration Reactions | Greener Organic Transformations - Books
    May 20, 2022 · It is now known that nitration of benzene is irreversible and the role of sulfuric acid is to generate the active nitronium ion (NO2+). The ...
  16. [16]
    Free Radicals - Chemistry LibreTexts
    Jan 22, 2023 · A free radical is an atom, molecule, or ion with unpaired valence electrons, making them highly reactive. They are also defined as transient, ...Depiction in chemical reactions · Formation · Reactivity · Combustion
  17. [17]
    6.3 Free Radicals - Chemistry LibreTexts
    Jun 5, 2019 · The formation of radicals may involve breaking of covalent bonds homolytically, a process that requires significant amounts of energy. For ...
  18. [18]
    9.3: Stability of Alkyl Radicals - Chemistry LibreTexts
    Dec 15, 2021 · Both benzylic and allylic radicals are more stable than the tertiary alkyl radicals because of resonance effects. "" This page titled 9.3: ...
  19. [19]
    6.3: Radical Reactions - Chemistry LibreTexts
    Apr 4, 2024 · Radical chain reactions have three distinct phases: initiation, propagation, and termination. generic mechanism for radical chain reaction ...
  20. [20]
    7.4.5 The Hydroxyl Radical - AR4 WGI Chapter 7
    The hydroxyl radical (OH) is the primary cleansing agent of the lower atmosphere, acting as a sink for greenhouse gases and pollutants.
  21. [21]
    Carbenes and nitrenes. An overview - ScienceDirect.com
    Feb 15, 2013 · A brief survey of the structure, methods of generation and reactivity of the singlet and triplet carbenes and nitrenes is given.
  22. [22]
    Carbon Dichloride as an Intermediate in the Basic Hydrolysis of ...
    Carbon dichloride as an intermediate in the basic hydrolysis of chloroform. A mechanism for substitution reactions at a saturated carbon atom.
  23. [23]
    [PDF] Energetic and chemical reactivity of atomic and molecular oxygen.
    Jun 28, 2010 · Atomic oxygen has two reactive sites and can abstract hydrogen. Molecular oxygen is a bi-radical, and its ground state is not highly reactive ...
  24. [24]
    Alkene Reactivity - MSU chemistry
    The carbocation intermediate formed in the first step of the addition ... mechanism in which a discrete carbocation intermediate is generated in the first step.
  25. [25]
    [PDF] Reactions of Alkenes
    Electrophilic addition is probably the most common reaction of alkenes. ... The regiochemistry is explained by the intermediate carbocation: The secondary ...
  26. [26]
    [PDF] 10. Alkenes and Alkynes. Addition Reactions - Organic Chemistry
    The first step of the mechanism involves the transfer of a proton from H-X to the alkene to give a carbocation intermediate (Figure 10.07)[next page]. Page 8. ( ...
  27. [27]
    10.8: Markovnikov's Rule - Chemistry LibreTexts
    Nov 2, 2021 · When an asymmetrical alkene undergoes electrophilic addition, the product that predominates is the one that results from the more stable of the ...
  28. [28]
    Hydrohalogenation of Alkenes and Markovnikov's Rule
    Feb 8, 2013 · A better reformulation of Markovnikov's rule is therefore that addition of HX to alkenes will proceed through the most stable carbocation, which ...
  29. [29]
    8.2: Halogenation of Alkenes - Addition of X₂ - Chemistry LibreTexts
    Apr 3, 2024 · Once formed, the bromonium ion is susceptible to attack by two nucleophiles—chloride ion and bromide ion—and, in fact, a mixture of two ...
  30. [30]
    Bromination of Alkenes - The Mechanism - Master Organic Chemistry
    Mar 15, 2013 · Halogenation of alkenes with Cl2 and Br2 goes through a halonium ion intermediate to give anti addition products. Halohydrins form in H2O.
  31. [31]
    7.12: Evidence for the Mechanism of Electrophilic Additions
    Apr 3, 2024 · This rearrangement can be achieved by either a hydride shift, where a hydrogen atom migrates from one carbon atom to the next, taking a pair of ...
  32. [32]
    Rearrangements in Alkene Addition Reactions - Chemistry Steps
    Rearrangements such as hydride, methyl shift and ring expansion occur in addition reactions of alkene that go through carbocation intermediates.
  33. [33]
    Electrophilic Addition to Alkenes EA2. Cations in ... - csbsju
    Alkenes can be protonated by acids to form carbon-centred cations or carbocations. · Sometimes, the anion left by the acid combines with the carbocation.
  34. [34]
    Rearrangement Reactions (1) - Hydride Shifts
    Jun 23, 2025 · One rearrangement pathway where an unstable carbocation can be transformed into a more stable carbocation is called a hydride shift.
  35. [35]
    55. Mechanism of substitution at a saturated carbon atom. Part IV. A ...
    A discussion of constitutional and solvent effects on the mechanism, kinetics, velocity, and orientation of substitution. E. D. Hughes and C. K. Ingold, J. Chem ...Missing: nucleophilic | Show results with:nucleophilic
  36. [36]
    Mechanism for Nucleophilic Substitution and Elimination Reactions ...
    Mechanism for Nucleophilic Substitution and Elimination Reactions at Tertiary Carbon in Largely Aqueous Solutions: Lifetime of a Simple Tertiary Carbocation.
  37. [37]
    Optical nature of non‐substituted triphenylmethyl cation: Crystalline ...
    Sep 24, 2021 · The UV-vis spectrum of the Y-crystal is nearly identical to that of a solution of the trityl cation whereas the spectrum of the O-crystal ...Missing: seminal | Show results with:seminal
  38. [38]
    [PDF] Spectroscopic study of carbenes in low-temperature matrices
    the IR absorption of the carbene slowly weakened and new bands appeared in the spectrum, which the authors45 assigned to the carbonyl oxide (LXV). Compound ...
  39. [39]
    Laser photolysis and spectroscopy: a new technique for the study of ...
    The technique is a hundred times faster than present flash spectrographic instrumentation and provides time-resolved absorption spectra over a wide spectral ...
  40. [40]
    Flash Photolysis Observation and Lifetimes of 2-Fluorenyl
    Flash Photolysis Observation and Lifetimes of 2-Fluorenyl- and 4-Biphenylylacetylnitrenium Ions in Aqueous Solution. Click to copy article linkArticle link ...
  41. [41]
    Matrix isolation infrared spectroscopy | NIST
    Sep 4, 2009 · The application of matrix isolation sampling for the stabilization and spectroscopic study of reaction intermediates has recently been reviewed ...
  42. [42]
    Continuous‐Wave (CW) Photo‐CIDNP NMR Spectroscopy: A Tutorial
    Sep 4, 2025 · The generation of CIDNP results in a nuclear spin system exhibiting non-Boltzmann spin state distributions. This so-called nuclear polarisation ...
  43. [43]
    Unimolecular thermal decomposition of phenol and d 5
    Jan 26, 2012 · demonstrated phenoxy radical and cyclopentadienyl radical formation from electron paramagnetic resonance (EPR) studies of phenol pyrolysis ...
  44. [44]
    The 4-Methyl-2-pentyl Cation | The Journal of Physical Chemistry A
    In the investigations where the DFT-B3LYP method was used, we examined at least some of the structures by standard ab initio calculations at the MP2 level, to ...
  45. [45]
    [PDF] Exploring Potential Energy Surfaces for Chemical Reactions
    Molecular structures correspond to the positions of the minima in the valleys on a potential energy surface. The energetics of reactions are easily calculated ...
  46. [46]
    A Theoretical Study of the P+ + SH2 Reaction: Potential Energy ...
    The results even point to a HPSH+(1A') species, which is the global minimum of the potential surface, as the reaction intermediate from which both products are ...
  47. [47]
    Capturing Free-Radical Polymerization by Synergetic Ab Initio ...
    Feb 15, 2022 · The free radical polymerization of photocurable compounds is studied using reactive classical molecular dynamics combined with a dynamical approach.Introduction · Computational Details · Results · Conclusions
  48. [48]
    Methylene: A Paradigm for Computational Quantum Chemistry - OUCI
    The effect of Renner–Teller coupling between the ã 1A1 and b̃ 1B1 states of CH2 on the rotational structure of the ã 1A1 bending vibrational levels has been ...
  49. [49]
    About Gaussian 16
    Jul 5, 2017 · Gaussian 16 is the latest version of the Gaussian series of electronic structure programs, used by chemists, chemical engineers, biochemists, physicists and ...
  50. [50]
    The ONIOM Method and Its Applications | Chemical Reviews
    Apr 8, 2015 · The ONIOM extrapolative method can be considered as a hierarchical method combining both the size of the molecule (or system) and the accuracy ...
  51. [51]
    Reaction mechanism of the farnesyl pyrophosphate C ... - Nature
    Feb 4, 2021 · During the cyclization cascade of the terpene cyclase, the carbocation intermediates can be stabilized by the π electrons of the aromatic ...
  52. [52]
    General base-general acid catalysis by terpenoid cyclases - Nature
    Apr 13, 2016 · These enzymes catalyze complex cyclization cascades, typically proceeding through multiple carbocation intermediates, to generate myriad ...
  53. [53]
    A Protein Radical Cage Slows Photolysis of Methylcobalamin in ...
    In the catalytic cycle, the methyl group of methylcobalamin is transferred to homocysteine, generating methionine and cob(I)-alamin, and cob(I)alamin is then ...Missing: mechanism | Show results with:mechanism
  54. [54]
    Structure of a Class I Tagatose-1,6-bisphosphate Aldolase
    The carbanion is extensively stabilized by numerous electrostatic interactions and hydrogen bonds with active site residues that are nearly identical to those ...
  55. [55]
    Evidence supporting a cis-enediol-based mechanism for ... - PubMed
    In contrast, the structurally distinct PGIs of eukaryotic or bacterial origin are thought to catalyse isomerisation via a cis-enediol intermediate. We have ...
  56. [56]
    ROS homeostasis and metabolism: a dangerous liason in cancer cells
    Jun 9, 2016 · This review outlines the metabolic-dependent mechanisms that tumors engage in when faced with oxidative stress conditions that are critical for cancer ...
  57. [57]
    Introduction to polymers: 4.3.1 Initiation | OpenLearn - Open University
    Styrene can also be polymerized by this compound: This mechanism is called anionic polymerization. The next two sections refer to free radical polymerization.
  58. [58]
    [PDF] Free Radical Polymerization - The Knowles Group
    Jan 11, 2020 · Interaction of radicals and monomers heavily depends on the structure of both radical and monomer. Head. Head addition. For reactions with ...
  59. [59]
    Fundamentals of Emulsion Polymerization | Biomacromolecules
    Jun 16, 2020 · Polymerization is initiated by adding a water-soluble free-radical initiator, and as the polymerization proceeds, polymer particles stabilized ...
  60. [60]
    Emulsion Polymerization Mechanism - IntechOpen
    Jan 17, 2018 · Emulsion polymerization is a free radical polymerization protocol occurs in three distinct steps; initiation, propagation, and termination. 3.1.
  61. [61]
    14.4.1: Ziegler-Natta Polymerizations - Chemistry LibreTexts
    May 3, 2023 · The Ziegler-Natta (ZN) catalyst, named after two chemists: Karl Ziegler and Giulio Natta, is a powerful tool to polymerize α-olefins with high linearity and ...
  62. [62]
    31.2 Stereochemistry of Polymerization: Ziegler–Natta Catalysts
    Sep 20, 2023 · The active form of a Ziegler–Natta catalyst is an alkyltitanium intermediate with a vacant coordination site on the metal. Coordination of ...
  63. [63]
    Catalytic Carbene Insertion into C−H Bonds | Chemical Reviews
    The insertion of a carbene into a carbon−hydrogen bond has attracted considerable interest because of its potential in forming carbon−carbon bonds.Introduction · Catalysts · Intramolecular Reactions · Intermolecular Reactions
  64. [64]
    Catalytic C–H Functionalization by Metal Carbenoid and Nitrenoid ...
    A very promising C–H functionalization method is the insertion of metal carbenes and nitrenes into C–H bonds. This chemistry has experienced considerable growth ...
  65. [65]
    3.1.10: Phenols and Their Uses - Chemistry LibreTexts
    Oct 4, 2022 · Treatment of cumene with oxygen in air generates 2-hydroperoxypropan-2-ylbenzene (cumene hydroperoxide) through a radical pathway.
  66. [66]
    Cumene Hydroperoxide - an overview | ScienceDirect Topics
    The first step in phenol production is oxidation of cumene to cumene-hydroperoxide (CHP) with 25-30% conversion.
  67. [67]
    Impact of New Manufacturing Technologies on the Petrochemical ...
    May 3, 2016 · Improvements in process technology or lower feedstock costs may enable new technologies to become cost-competitive with existing production ...
  68. [68]
    Petrochemical Industry for the Future - ScienceDirect.com
    Production processes are designed and improved by optimizing feedstocks, momentum transfer, heat transfer, mass transfer, and chemical reaction engineering.