Fact-checked by Grok 2 weeks ago

CH3

The , denoted as CH₃•, is the simplest free , consisting of a carbon atom bonded to three hydrogen atoms and possessing an on the carbon. It exhibits a planar trigonal with D3*h , featuring C-H lengths of approximately 1.079 and H-C-H angles of 120°. This structure arises from sp2 hybridization of the carbon atom, resulting in a highly reactive due to the . It is a metastable gas with a molecular weight of 15.03 g/. The plays a pivotal role as a transient in numerous chemical processes. In chemistry, it is essential in oxidation pathways, influencing ignition and flame propagation in methane-based fuels. In , it is involved in radical chain reactions that degrade pollutants and contribute to formation. In , enable efficient reactions for constructing carbon-carbon bonds, particularly in late-stage functionalization of pharmaceuticals and materials, often generated via photo-, electro-, or transition-metal-catalyzed methods. Its reactivity includes hydrogen abstraction, to unsaturated bonds, and recombination to form , underscoring its versatility across , astro-, and surface chemistries.

Structure and bonding

Geometry and hybridization

The free methyl radical (CH₃•) adopts sp² hybridization for the carbon atom, resulting in a trigonal planar geometry with H-C-H bond angles of 120° and C-H bond lengths of approximately 1.079 Å. The three sp² hybrid orbitals lie in the molecular plane and form σ bonds with the hydrogen 1s orbitals, while the unhybridized 2p_z orbital perpendicular to the plane holds the unpaired electron. This planar D3h structure is the ground state, slightly more stable than a hypothetical pyramidal sp³ arrangement due to the lower energy of sp² hybrids and delocalization of the unpaired electron. The geometry has been confirmed spectroscopically, including through microwave spectroscopy revealing rotational constants consistent with the trigonal planar arrangement. In contrast, the (-CH₃) in organic molecules such as or features sp³ hybridization of the carbon atom, resulting in a geometry with H-C-H bond angles of approximately 109.5°. This sp³ hybridization involves the of the carbon atom's 2s orbital and three 2p orbitals (2p_x, 2p_y, and 2p_z), producing four equivalent sp³ hybrid atomic orbitals. These orbitals point toward the vertices of a regular , maximizing separation between bonding s to minimize repulsion per valence shell repulsion () theory. In the , three sp³ hybrid orbitals each overlap end-to-end with a 1s orbital to form C-H σ bonds, while the fourth forms a σ bond with an adjacent atom.

Bonding characteristics

The three C-H bonds in the methyl radical (CH₃•) are equivalent covalent (σ) bonds, formed by end-to-end overlap of sp² hybrid orbitals on the central carbon atom with 1s orbitals on the atoms. These bonds have a length of approximately 1.079 and lie in the trigonal plane, with the shared concentrated along the internuclear axis. The resides in the perpendicular 2p_z orbital, contributing to the radical's high reactivity. In comparison, the three C-H bonds in the saturated (-CH₃) are equivalent covalent (σ) bonds, each formed by the end-to-end overlap of an sp³ hybrid orbital on the central carbon atom with a 1s orbital on a . This overlap creates strong, directional bonds characteristic of tetrahedral , with the shared concentrated along the internuclear axis. These C-H bonds in the methyl group have a length of 1.09 Å, typical for sp³-hybridized carbon-hydrogen sigma bonds in simple alkyl groups. The sp³ hybridization of the carbon atom in the methyl group, which enables this bonding arrangement, is discussed in the geometry and hybridization section. The strength of C-H bonds related to the methyl radical is reflected in their dissociation energies. The energy required to break one C-H bond in methane (CH₄), producing CH₃• and H•, is 439 kJ/mol at 298 K.

Physical properties

Spectroscopic properties

() spectroscopy provides insights into the vibrational modes of the , which arise from its planar D_{3h} and sp² hybridization. The CH stretching consist of a symmetric mode (a₁') at approximately 3004 cm⁻¹ and a degenerate asymmetric mode (e') at 3161 cm⁻¹. The characteristic out-of-plane umbrella bending mode (a₂'') occurs at 606 cm⁻¹, while the degenerate in-plane deformation (e') is around 1396 cm⁻¹. These modes are used to identify the radical in gas-phase and matrix isolation studies. Electron paramagnetic resonance (EPR) spectroscopy is crucial for detecting the in the methyl radical. The spectrum exhibits a symmetric due to hyperfine coupling with the three equivalent protons (nuclear spin I=1/2), with a proton hyperfine coupling constant of approximately 23 G and a g-factor of 2.0025. The equal spacing and 1:3:3:1 intensity ratios confirm the planarity and equivalence of the hydrogens. For the methyl radical (CH₃•), ultraviolet-visible (UV-Vis) spectroscopy detects weak absorption bands arising from electronic transitions involving the . The spectrum features a broad, relatively weak band centered near 216 nm, attributed to a promotion from the (X ²A₂'') to an with π* character, analogous to n→π* transitions in other radicals due to the non-bonding electron's involvement. This absorption has a low molar absorptivity (ε ≈ 300 M⁻¹ cm⁻¹), making it suitable for sensitive detection in gas-phase studies, such as processes. The planar D₃h geometry of the radical influences the transition intensities, with forbidden or weakly allowed bands dominating the spectrum below 300 nm.

Thermodynamic properties

The f298) for the gas-phase methyl (CH₃•) is +146.5 kJ/mol, reflecting its endothermic nature relative to carbon and . This value, determined through high-level quantum chemical calculations and experimental benchmarks, underscores the energetic cost of generating the from stable precursors like . The (S°) of the gas-phase methyl radical at 298 K and 1 pressure is 194.2 J/mol·K, consistent with its planar structure and three equivalent atoms contributing to rotational and vibrational . The corresponding standard of formation (Δf298) is +148.7 kJ/mol, calculated from the enthalpy and entropy data using the relation ΔfG° = ΔfH° - T ΔfS°. As an unstable intermediate species, the isolated methyl radical does not exhibit defined phase transitions under standard conditions and exists solely in the gas phase at .

Methyl derivatives

Methyl cation

The methyl cation, \ce{CH3+}, possesses a planar D_{3h} geometry, with the central carbon atom sp²-hybridized and bonded to three equivalent hydrogen atoms, featuring an empty p-orbital perpendicular to the plane that imparts and results in C-H bond angles of 120°. This configuration contrasts with the neutral methyl radical, which shares a similar planar arrangement but includes a singly occupied p-orbital. The empty p-orbital renders \ce{CH3+} highly reactive as a strong electrophile, prone to rapid reactions with nucleophiles in both laboratory and astrophysical environments. In laboratory settings, \ce{CH3+} forms in the gas phase or media via of (\ce{CH4}), initially yielding the protonated \ce{CH5+}, which then undergoes elimination of molecular to generate the cation. This process underscores its inherent instability in condensed phases, where it persists only under isolated conditions like matrix isolation or . Astronomically, \ce{CH3+} plays a pivotal role in interstellar chemistry as a precursor to complex hydrocarbons, initiating gas-phase organic synthesis through reactions with species like \ce{H2} and electrons. Its first detection occurred in 2023 using the , which captured rovibrational emission lines in the within the star-forming region, approximately 1,350 light-years from ; these observations confirm its abundance in UV-irradiated disk atmospheres, where photochemical processes drive its formation and sustain its concentration at levels enabling detectable signals.

Methyl anion

The methyl anion, CH₃⁻, is a highly reactive characterized by a pyramidal with C_{3v} , analogous to the structure of . The central carbon atom adopts sp³ hybridization, featuring three σ bonds to atoms and a in the fourth sp³ orbital, resulting in bond angles of approximately 110° and an out-of-plane inversion barrier of about 1.7 kcal/. This pyramidal configuration distinguishes it from the planar methyl cation, highlighting the anion's nucleophilic nature driven by the localized negative charge on carbon. Due to its extreme basicity, with the pK_a of its conjugate acid estimated at approximately 50, the methyl anion acts as a and is inherently unstable in conventional solution environments, readily abstracting protons from even weak acids. It is primarily generated and studied in the gas phase via techniques such as ion cyclotron resonance or pulsed plasma discharge, or isolated in low-temperature matrices to prevent decomposition. Quantum chemical calculations, including methods and , reveal a low for the methyl radical (approximately 0.08 or 1.8 kcal/), indicating marginal stability of the anion relative to into CH₃ and an electron. These computations also underscore the anion's strong proton abstraction tendency, with potential energy surfaces showing low barriers for reactions, reinforcing its role as an elusive in gas-phase chemistry.

Methyl radical

The methyl radical (CH₃•) adopts a planar with D_{3h} , arising from sp² hybridization at the central carbon atom, which positions the three hydrogen atoms in a trigonal arrangement. The resides in a p-orbital to this plane, resulting in C-H angles of approximately 120° that minimize steric repulsion between the hydrogens while maximizing orbital overlap. This radical species is primarily generated in the gas phase through homolytic cleavage of the C-H in , as in the process CH₄ → CH₃• + H•, often initiated by , photochemical, or plasma-driven . In dilute gas-phase conditions, the methyl radical exhibits a short lifetime, typically around 1.4 milliseconds, limited by its high reactivity toward self-combination. The dominant decay pathway for the methyl radical in the gas phase is dimerization to form ethane via the recombination reaction: $2 \text{CH}_3^\bullet \rightarrow \text{C}_2\text{H}_6 This process is barrierless, with an activation energy of 0 ± 0.7 kcal/mol, reflecting the favorable coupling of the two unpaired electrons without an energetic barrier. The rate constant in the high-pressure limit follows a negative temperature dependence, expressed as k_\infty = (5.66 \pm 0.43) \times 10^{-11} (T/298)^{-0.37} cm³ molecule⁻¹ s⁻¹ over 292–714 K, underscoring the reaction's efficiency at lower temperatures and higher pressures where third-body stabilization prevents redissociation.

Reactivity

Electrophilic behavior

The exhibits electrophilic behavior primarily through equivalents of the methyl cation (CH₃⁺), which serves as a in various substitution reactions. One prominent example is the of using (CH₂N₂), where the reagent acts as a source of electrophilic methyl under acidic conditions. The mechanism begins with proton transfer from the (RCOOH) to diazomethane, generating an electrophilic diazonium ion (CH₃N₂⁺) and the corresponding (RCOO⁻). This is followed by nucleophilic attack of the carboxylate on the diazonium ion, displacing N₂ to form the (RCOOCH₃). This process is highly selective and tolerant of various functional groups, proceeding under mild conditions with N₂ as the sole byproduct. In , methyl halides such as methyl chloride (CH₃Cl) can introduce the onto aromatic rings via Friedel-Crafts , facilitated by Lewis acids like AlCl₃. The mechanism involves coordination of the Lewis acid to the halide, promoting to form a complexed methyl cation equivalent, such as [CH₃-Cl-AlCl₃]⁺. This then attacks the aromatic π-system, leading to a σ-complex , followed by to yield the methylated arene, such as from . Unlike higher alkyl halides, methylations avoid carbocation rearrangements due to the stability and simplicity of CH₃⁺, though polyalkylation can occur if not controlled. This reaction is particularly useful for preparing methyl-substituted aromatics and has been studied extensively in the context of AlCl₃-catalyzed systems. SN1 pathways involving methyl intermediates are characterized by unimolecular rate-determining dissociation of the methyl halide (CH₃X) to generate CH₃⁺, followed by rapid nucleophilic capture. The for such processes is , expressed as rate = k [CH₃X], independent of concentration, reflecting the slow formation of the highly reactive but unstable methyl cation. While methyl halides predominantly favor SN2 mechanisms in due to the poor of CH₃⁺, SN1-like has been observed in solvolysis studies under conditions promoting , such as in highly polar solvents or with strong acids, where relative rates underscore the energetic barrier to formation (e.g., methyl << primary < secondary < tertiary). The methyl cation's planar, empty p-orbital structure facilitates backside attack but leads to racemization in chiral contexts when applicable.

Nucleophilic behavior

The methyl group displays nucleophilic behavior predominantly through its deprotonated form, the methyl anion (\ce{CH3^-}), which acts as a highly reactive nucleophile in synthetic applications. The C-H bonds of the methyl group, as in methane (\ce{CH4}), possess a pK_a value of approximately 50, underscoring their extremely weak acidity and necessitating exceptionally strong bases for deprotonation to generate the carbanion. Direct deprotonation of methane using bases like alkyllithiums such as n-BuLi is thermodynamically demanding and rarely practical in standard solution conditions due to the high pK_a, but the methyl anion is routinely accessed via organolithium reagents like methyllithium (\ce{CH3Li}), which serve as effective carbanion equivalents. Methyllithium, prepared from methyl halides and lithium metal, delivers the nucleophilic methyl unit in a variety of transformations, exhibiting both strong basicity and nucleophilicity owing to the polarized C-Li bond. These reagents enable the formation of new carbon-carbon bonds by transferring the methyl group to electrophiles. The methyl anion engages in nucleophilic substitution reactions, favoring the SN2 pathway due to its compact size and high electron density at carbon, which promotes backside attack on electrophilic centers. This reactivity is particularly pronounced with unhindered substrates like methyl halides (\ce{CH3X}), where the absence of steric hindrance at the substrate carbon accelerates the bimolecular displacement, yielding products such as ethane from \ce{CH3^- + CH3I -> CH3CH3 + I^-}. In broader synthetic contexts, performs SN2 alkylations on primary alkyl halides or other electrophiles, such as carbonyls, with rates enhanced by solvents that solvate the , minimizing pairing. The inherent nucleophilicity of the methyl anion, characterized by its low methyl anion affinity toward common electrophiles and pyramidal geometry that concentrates negative charge, positions it among the strongest carbanionic nucleophiles, though its instability limits isolation to gas-phase or matrix conditions.

Radical reactions

The (CH₃•) plays a key role in initiating free of methyl acrylates, such as (CH₂=CHCOOCH₃), by adding to the electron-deficient double bond of the monomer. This addition step forms a carbon-centered on the α-carbon, typically represented as CH₃CH₂ĊHCOOCH₃, which serves as the propagating in chain growth. Theoretical studies using methods have calculated the barrier for this initiation reaction at approximately 10-15 kJ/mol in the gas phase, highlighting its feasibility under typical polymerization conditions. In the phase, the radical adds to additional units, extending the chain, while by CH₃• sets the end-group structure. Termination occurs primarily through combination of two growing radicals or , yielding with controlled molecular weight distribution. Experimental and computational assessments confirm that the rate constant for CH₃• addition to is on the order of 10⁶-10⁷ L mol⁻¹ s⁻¹ at 300 K, influencing the overall kinetics and properties like . In processes, the participates in hydrogen abstraction reactions as a step, where CH₃• + RH → CH₄ + R• sustains radical chain branching in fuels. This reaction is critical in low-temperature regimes, with evaluated rate constants varying by the C-H bond strength of RH; for example, abstraction from (CH₄) has k ≈ 1.1 × 10⁻¹⁴ exp(-4080/T) cm³ molecule⁻¹ s⁻¹, while for alkanes like , it is faster due to weaker secondary C-H bonds. Such abstractions contribute to the formation of reactive alkyl (R•), accelerating oxidation and influencing ignition delays in engines. Coupling reactions involving the are exemplified in the Kharasch addition (atom transfer radical addition, ATRA), where CH₃• initiates the process by adding to terminal , facilitating subsequent transfer of electrophilic groups like or chalcogens from additives such as alkyl iodides or xanthates. In this mechanism, of di-tert-butylhyponitrite generates CH₃•, which adds to the to form a secondary radical; this then abstracts a (e.g., I) from the additive, propagating the chain and yielding α-functionalized products with high . Recent advancements demonstrate broad substrate scope, including unactivated , under mild reflux conditions in , with yields often exceeding 80% and minimal byproducts.

Applications and occurrence

In organic synthesis

The methyl radical CH₃• serves as a key intermediate in radical-mediated reactions, enabling the selective installation of methyl groups, often in late-stage functionalization of complex molecules. These methods typically involve the generation of CH₃• from precursors like , (DMSO), or peroxides via photo-, electro-, or transition-metal catalysis, avoiding the limitations of traditional ionic methylating agents. One prominent application is the Minisci-type C-H of heteroarenes, such as pyridines and quinolines. For example, silver-mediated of acetic acid or of peroxides generates CH₃•, which adds to the electron-deficient ring, followed by rearomatization to yield methylated products. This approach is valuable for diversifying pharmaceutical scaffolds. has advanced CH₃• generation for milder conditions. In a method developed by MacMillan et al., dual photoredox and hydrogen atom transfer (HAT) activates as a methyl source, allowing C-H of (hetero)arenes with photocatalysts and HAT agents, achieving high site selectivity in drug-like molecules. Similarly, Glorius et al. employed photocatalytic reduction of DMSO-derived species for . These techniques support the "magic methyl" effect, where adding a enhances in . Transition-metal-catalyzed processes further expand applications. or catalysts facilitate CH₃• addition in cross-couplings or conjugate additions, such as Ni/photoredox systems using dimethylnickel precursors for sp³ C-methylation. These methods are applied in synthesizing methylated fragments and materials.

Biological significance

The methyl radical CH₃• has no established direct role in biological systems due to its high reactivity and short lifetime, which prevent stable incorporation . However, it appears as a transient in certain processes studied in biochemical contexts, such as mechanistic investigations of DMSO-derived radicals in cellular environments. Unlike the methyl group (-CH₃) in or , CH₃• does not participate in enzymatic methyl transfers. Its biological relevance is primarily indirect, through or radical scavenging pathways, but specific functions remain underexplored as of 2025.

Special topics

Chiral methyl groups

Isotopically substituted methyl groups, such as - and -, acquire when the three hydrogen positions are differentiated by () and/or () substitution, particularly when attached to a prochiral center like a carbonyl carbon in derivatives. In the case of -CHTD, the methyl carbon becomes a stereogenic center with four distinct substituents—the attachment point, H, D, and T—rendering the group inherently chiral and non-superimposable on its . For -CHD₂, emerges in the context of stereospecific at a prochiral center, where the non-equivalent H and D atoms allow discrimination between and faces during enzymatic addition or abstraction, enabling detailed stereochemical analysis without the group itself being tetrahedral chiral. This isotopic has been pivotal in elucidating the of biochemical transformations involving methyl groups. Synthesis of these chiral methyl groups often relies on enzymatic resolution to achieve high enantiopurity. Seminal work by Cornforth involved the kinetic resolution of racemic [1-²H₁, ²-³H₁] using , selectively oxidizing one to the corresponding chiral while leaving the other intact; subsequent chemical transformations yield substrates like chiral S-adenosylmethionine analogs with defined () or () configuration at the methyl carbon. Asymmetric deuteration methods complement this by employing chiral catalysts for stereoselective incorporation of , such as rhodium-phosphine complexes in the of prochiral precursors to generate enantioenriched -CHD₂ or -CHTD groups, achieving enantioselectivities exceeding 99% in some cases. These approaches ensure the isotopic labels are introduced with precise stereocontrol, avoiding during synthesis. In (NMR) spectroscopy, these chiral methyl groups serve as probes for determining , particularly in reactions involving or transfer at methyl sites. For instance, in studies of radical SAM methyltransferases like Fom3 in fosfomycin biosynthesis, feeding substrates with (methyl-R)- or (methyl-S)-[methyl-²H₁, ³H₁] reveals retention of configuration through analysis of tritium retention ratios (F values) and deuterium isotope effects, confirmed via ³H NMR of degradation products. Similarly, stereospecifically labeled -CHD₂ groups in protein side chains (e.g., or ) enable assignment of pro-R versus pro-S methyl positions, facilitating NMR-based mapping of enzymatic active sites and their chiral preferences in high-molecular-weight systems. This methodology has established that many , such as , operate with strict pro-R or pro-S specificity at methyl groups, providing insights into catalytic mechanisms.

Internal rotation

The internal rotation of the around the C–X bond in methyl-substituted compounds, such as (CH₃–CH₃), is characterized by a low barrier of approximately 12 kJ/mol (or 2.9 kcal/mol), primarily due to torsional arising from the repulsion between adjacent atoms in eclipsed conformations. This barrier represents the difference between the staggered (minimum ) and eclipsed () conformations, enabling rapid rotation at ambient temperatures with a timescale on the order of picoseconds. The sp³ hybridization of the carbon atoms facilitates this nearly free rotation by providing equivalent bonding orbitals. Microwave spectroscopy has been instrumental in quantifying this rotational dynamics through measurements of the molecule's rotational constants (A, B, and C), which describe the moments of and reveal splittings in the caused by the internal motion of the methyl groups. For instance, in partially deuterated (CH₃CHD₂), high-resolution spectra of torsional transitions allowed precise determination of the barrier height as V₃ = 1010.39 ± 0.10 cm⁻¹ (equivalent to ~12.05 kJ/mol), confirming the torsional potential's threefold . These rotational constants, typically on the order of several GHz for , provide direct experimental access to the structural and dynamic parameters without relying solely on computational models. Quantum mechanically, the internal rotation is described by the hindered rotor model, which treats the as a symmetric rotor subject to a periodic potential barrier, leading to a that includes kinetic and potential terms for the torsional angle θ. The for this one-dimensional hindered rotor is solved to yield quantized torsional energy levels, with the barrier height modulating the splitting between symmetric and antisymmetric states; for , this model accurately reproduces the observed splittings and thermodynamic contributions from hindered rotation. This framework, originally developed in the 1930s for , accounts for the quantum tunneling through the barrier, which further lowers the effective rotational freedom compared to a classical free rotor.

History and nomenclature

Discovery and etymology

The , denoted as CH₃, was first identified as a distinct chemical entity in the early through investigations into the composition of organic compounds, particularly during studies of wood-derived substances. In 1834, French chemists Jean-Baptiste-André Dumas and Eugène-Melchior Péligot isolated pure (then called "wood spirit" or esprit de bois) by careful of wood, determining its elemental composition as consisting of carbon, , and oxygen in the ratio corresponding to CH₄O. They further prepared several derivatives, including methyl chloride and methyl iodide, by reacting with halogens under specific conditions, which revealed a common univalent radical equivalent to CH₃ present across these compounds. This work, detailed in their seminal memoir presented to the Académie des Sciences on October 27 and November 3, 1834, marked the initial recognition of the as a fundamental building block in . In the context of the emerging radical theory of organic structure proposed by , their findings demonstrated that the behaved consistently in substitution reactions across alcohols, ethers, and halides, supporting the idea of persistent radicals (meaning atomic groups) in organic compounds and paving the way for systematic classification of . The free methyl radical (CH₃•), however, was not isolated until the 20th century. In 1929, German chemists Friedrich A. Paneth and Wilhelm Hofeditz provided the first experimental evidence for its existence using the "mirror technique." They decomposed tetramethyllead vapor (Pb(CH₃)₄) at low pressure in a flow system, observing that metallic mirrors (e.g., ) placed downstream were removed and could be recoated farther along the tube, indicating the transient presence of highly reactive CH₃• radicals that abstracted atoms from the mirror surface. This demonstration confirmed the methyl radical as a true free radical , distinct from the conceptual group identified earlier, and advanced the understanding of free radical chemistry. The etymology of "methyl" traces directly to this historical context, combining the Greek methy (μέθυ, meaning "wine") with hylē (ὕλη, meaning "wood" or "matter"), to denote the alcohol derived from wood in contrast to ethyl alcohol from fermented wine. Dumas and Péligot initially termed the alcohol "méthylique" and the related divalent group CH₂ as "méthylène" in their 1835 publication, establishing nomenclature that emphasized the substance's origin. By 1840, the term "methyl" had been adopted for the CH₃ radical itself (in the group sense), solidifying its place in chemical terminology.

Nomenclature

In , the , represented as −CH₃, is the simplest derived from (CH₄) by the removal of one . According to IUPAC substitutive , it is named "methyl" by replacing the final "-e" of the parent hydride name with the suffix "-yl," and it functions as a in naming more complex compounds. For example, in alkanes, a chain with a attached to the second carbon of is named 2-methylpropane. When multiple methyl groups are present, multiplicative prefixes such as "di-," "tri-," or "tetra-" are used, and they are arranged in alphabetical order with other substituents, ignoring the multiplicative prefixes for alphabetization. Retained names like "isopropyl" (for (CH₃)₂CH−) are permitted alongside systematic names, but "methyl" itself is a retained applicable without restriction. The , denoted as CH₃•, follows similar IUPAC conventions for , where it is named "methyl" as a retained name for the monovalent obtained by removing one from , with the at the carbon center. The systematic name is "methanidyl," but the "methyl" is preferred and widely used in chemical literature for this species. In radicofunctional , it may appear in names like "" for CH₃OH, treating the as a separate component.

References

  1. [1]
    Recent Progress in Methyl-Radical-Mediated Methylation or Demethylation Reactions
    ### Summary of Methyl Radical from Abstract and Introduction
  2. [2]
    Methyl | CH3 | CID 3034819 - PubChem - NIH
    Methyl | CH3 | CID 3034819 - structure, chemical names, physical and chemical properties, classification, patents, literature, biological activities, ...
  3. [3]
    Variationally Computed IR Line List for the Methyl Radical CH3
    May 3, 2019 · The methyl radical CH3 is a free radical of major importance in many areas of science, such as hydrocarbon combustion processes, (1) ...
  4. [4]
  5. [5]
    Physisorption-Induced C-H Bond Elongation in Methane
    Apr 13, 2006 · Experimental x-ray absorption spectra of methane adsorbed on a Pt surface. ... length from 1.09 Å to 1.18 ± 0.05 Å with the molecule at a ...
  6. [6]
    Active Thermochemical Tables: Sequential Bond Dissociation ...
    Mar 11, 2015 · The ATcT bond dissociation enthalpy of methane is BDE298(H–CH3) = 438.892 ± 0.065 kJ/mol (432.373 kJ/mol at 0 K), see Table 1. The quoted ...Introduction · Results and Discussion · Conclusions · References
  7. [7]
    [PDF] Chapter 5. IR Spectroscopy and Raman Scattering
    The bend vibration for methyl groups is a doublet (methylene is a singlet), 1450 (a) and 1375 (s). The symmetric bend for a methyl group is called an umbrella ...
  8. [8]
    [PDF] IR_lectureNotes.pdf
    6.a methyl group may be identified with C-H absorption at 1380 cm-1. This band is split into a doublet for isopropyl(gem-dimethyl) ...
  9. [9]
    Proton NMR Table - MSU chemistry
    The b-carbon of the double bond is shifted to lower field by 20 to 30 ppm, and the carbonyl carbon is shifted to higher field by 5 to 15 ppm. Examples of the ...
  10. [10]
    Determination of the absorption cross section for methyl at 216.36 ...
    Determination of the absorption cross section for methyl at 216.36 nm and the absolute rate constant for methyl radical recombination over the temperature range ...
  11. [11]
    Methyl radical - the NIST WebBook
    , Absorption spectrum of free CH3 and CD3 radicals, Can. J. Phys ... Spectrum of Gaseous Methyl Radical by Rapid Scan Spectroscopy, J. Chem. Phys ...
  12. [12]
    Methyl Enthalpy of Formation
    ### Summary of Thermodynamic Data for CH3 (g)
  13. [13]
    Methyl radical - the NIST WebBook
    Entropy of gas at standard conditions (1 bar). ΔfH°gas, Enthalpy of formation of gas at standard conditions. Data from NIST Standard Reference Database 69 ...
  14. [14]
  15. [15]
    Electron affinity of the methyl radical: Structures of CH3 and ... - PNAS
    The methyl anion is unstable with respect to the limit of CH3 plus an electron by 2-8 kcal/mole.
  16. [16]
    An experimental determination of the geometry and electron affinity ...
    ... Methyl Radical to Unsaturated Compounds: The Methyl Affinities Revisited. Journal of the American Chemical Society 1997, 119 (52) , 12869-12878. https://doi ...
  17. [17]
    [PDF] pka-compilation-williams.pdf - Organic Chemistry Data
    Apr 7, 2022 · 3-methyl-5,5-pentamethylene-2-thio-. 6-hydroxy. 7.52, 3.65 39. 11.23. 42 ... CH3=CH(CH3)2 (CH2)5. 12.95. 13.60. Provided by the ACS, Organic ...
  18. [18]
    Gas-phase reactions of the methyl anion - ACS Publications
    Gas-phase reactions of the methyl anion | Journal of the American Chemical Society.
  19. [19]
    Gas-phase coupling of methyl radicals during the catalytic partial ...
    Dec 8, 1987 · It is generally accepted that the initial step in the catalytic oxidative coupling of methane involves the homolytic cleavage of a C-H bond.Missing: lifetime | Show results with:lifetime
  20. [20]
    Diode laser study of the ν2 band of the methyl radical - AIP Publishing
    Dec 1, 1981 · The methyl radical was generated by a 60 Hz discharge in di‐tert‐butyl peroxide and was found to have a lifetime of about 1.4 msec; its ...
  21. [21]
    The Rate Constant of Ethane Formation from Methyl Radicals
    The resulting value of the recombination rate constant is 4.5×1013 (moles/cc)−1 sec−1, with an activation energy of E=0±700 cal regardless of radical source.
  22. [22]
    Reaction CH3 + CH3 → C2H6 Studied over the 292–714 K ...
    The rate constant of reaction 1 in the high-pressure limit has a negative temperature dependence: k1,inf = (5.66 ± 0.43) × 10–11(T/298 K)−0.37 cm3 molecule ...
  23. [23]
  24. [24]
    [PDF] Chem 14D – Spring 2013 pKa Table
    Mar 19, 2013 · pKa Table. (see Vollhardt: p. 60 in 5th Edition or 6th Edition). Name of Acid. Chemical Formula. pKa. Hydrogen iodide. HI ... Methane. CH4. ~50.
  25. [25]
    Bordwell pKa Table - Organic Chemistry Data
    Below are tables that include determined pKa values for various acids as determined in water, DMSO and in the gas Phase. These tables are compiled in PDF files ...Missing: CH4 | Show results with:CH4
  26. [26]
    The Dual Electronic Behavior of the Alkyl Group - ACS Publications
    Oct 28, 2025 · Although a trigonal geometry is predicted for the methyl radical (24), (79) pyramidal geometries are calculated for the ethyl (25) ...
  27. [27]
    A monomeric methyllithium complex: synthesis and structure
    May 20, 2021 · Methyllithium (MeLi) is the parent archetypal member of the RLi family, which has been widely used as a methyl synthon, a nucleophilic reagent ...
  28. [28]
    Gas-Phase Identity SN2 Reactions of Halide Anions with Methyl ...
    Gas-Phase Identity SN2 Reactions of Halide Anions with Methyl Halides: A High-Level Computational Study. Click to copy article linkArticle link copied! Mikhail ...Missing: involving | Show results with:involving
  29. [29]
    A theoretical study of the gas-phase ion pair SN2 reactions of lithium ...
    Dec 15, 2003 · A theoretical study of the gas-phase ion pair SN2 reactions of lithium halide and methyl halide with inversion and retention mechanisms. Author ...
  30. [30]
    [PDF] Evaluated Kinetic Data for Combustion Modelling
    This compilation contains critically evaluated kinetic data for 196 gas phase reactions for use in computer modeling of combustion processes.
  31. [31]
  32. [32]
    Magic methylation with methyl-containing peroxides - ScienceDirect
    Nov 18, 2024 · Frequently utilised methylating reagents include methyl iodide, dimethyl sulfate, dimethyl carbonate, methyl triflate, and various methyl ...
  33. [33]
    Trimethylaluminum - Suzuki - Major Reference Works
    Sep 6, 2017 · Lewis acid with methylation ability; 2-11, 31-40, 46-53 can effect useful CC bond forming reactions including conjugate addition, 41-45 cross-coupling ...
  34. [34]
    Exhaustive methylation by trimethylaluminium - Semantic Scholar
    Trimethylaluminium exhaustively methylates tertiary alcohols to alkanes, ketones to gem-dimethyl compounds, and carboxylic acids to t-butyl compounds.14 Citations · Biologisch Aktive 3... · 3d‐printing Inside The...Missing: review | Show results with:review
  35. [35]
    Asymmetric alkylation reactions of chiral imide enolates. A practical ...
    Evans Enolates: Solution Structures of Lithiated Oxazolidinone-Derived Enolates. Journal of the American Chemical Society 2015, 137 (40) , 13087-13095 ...
  36. [36]
    Applications of oxazolidinones as chiral auxiliaries ... - RSC Publishing
    Various chiral oxazolidinones (Evans' oxazolidinones) have been employed as effective chiral auxiliaries in the asymmetric alkylation of different enolates.Missing: stereoselective methylation
  37. [37]
    Recent Advances in Transition Metal‐Catalyzed Methylation ...
    Apr 23, 2015 · In the past decades, the transition metal-catalyzed cross-coupling reaction has emerged as a powerful tool for the formation of carbon-carbon ...
  38. [38]
    DNA Methylation and Its Basic Function - PMC - PubMed Central
    Jul 11, 2012 · DNA methylation regulates gene expression by recruiting proteins involved in gene repression or by inhibiting the binding of transcription ...
  39. [39]
    DNA methylation and regulation of gene expression - PubMed Central
    Aug 16, 2021 · Based on the sites of DNA methylation in a locus, it can serve as a repressive or activation mark for gene expression.
  40. [40]
    Methylation across the central dogma in health and diseases - Nature
    Aug 25, 2023 · There are ~200 genes in the human genome encoding known or putative SAM-dependent methyltransferases, which have been grouped according to ...
  41. [41]
    DNA methylation-environment interactions in the human genome
    May 19, 2023 · DNA methylation is sensitive to the environment, can remain stable across cell divisions, and in some contexts, can alter gene regulation.
  42. [42]
    One-Carbon Metabolism in Health and Disease - PMC - NIH
    After donating a methyl group to homocysteine, betaine becomes dimethylglycine, which can yield additional 1C units via folate-dependent mitochondrial reactions ...
  43. [43]
    One Carbon Metabolism and Epigenetics: Understanding the ... - NIH
    Composed of the folate and methionine cycle, it generates S-adenosylmethionine (SAM), which is the universal methyl donor for methylation reactions, including ...
  44. [44]
    Partitioning of One-Carbon Units in Folate and Methionine ...
    Nov 14, 2017 · Folate one-carbon metabolism (FOCM) comprises an interlinked network of reactions that transfer one-carbon (1C) units for numerous cellular ...
  45. [45]
    Modulation of DNA methylation by one-carbon metabolism
    Folate-mediated one-carbon metabolism produces thymidylate and purines for the synthesis and repair of DNA [9]. In this pathway, serine is the methyl donor, and ...
  46. [46]
    N-Methylation of Dopamine to Epinine in Brain Tissue using ... - Nature
    Aug 16, 1972 · DURING biosynthetis of catecholamines, the final step in adrenaline formation involves transfer of a methyl group from S-adenosylmethionine ...
  47. [47]
    N-Methylation of dopamine to epinine in adrenal medulla - PubMed
    N-Methylation of dopamine to epinine in adrenal medulla: a new model for the biosynthesis of adrenaline.
  48. [48]
    Methylation reactions at dopaminergic nerve endings, serving ... - NIH
    Methylation of dopamine within the synaptic cleft may serve as a phasic down regulating switch for dopamine neurotransmission. The methylation of free DA (f ...Missing: epinine | Show results with:epinine
  49. [49]
  50. [50]
    Dependent Radical SAM Methyl Transfer in Fosfomycin Biosynthesis
    May 4, 2021 · ... chiral methyl analysis. These methods facilitate the broader use of in vitro chiral methyl analysis techniques to investigate the mechanisms ...
  51. [51]
  52. [52]
    Sir John Warcup Cornforth (1917–2013) - Wiley Online Library
    Mar 5, 2014 · The basis of this work was the synthesis of “chiral” acetic acid by a route that is wonderfully simple. ... methyl group, was catalyzed by the ...
  53. [53]
    Asymmetric catalytic synthesis of chiral deuterated compounds
    This article primarily reviews the asymmetric deuterium synthesis methods reported in recent years, focusing on the following strategies: one-step reductive ...Missing: methyl | Show results with:methyl
  54. [54]
  55. [55]
  56. [56]
    [PDF] Stereochemistry of the incorporation of valine methyl groups into ...
    Analysis of the stereochemistry ofsubstitution into a chiral methyl group by 3H n.m.r. be expected to exhibit a deuterium kinetic isotope effect, so that ...<|control11|><|separator|>
  57. [57]
    The Rotational Barrier in Ethane: A Molecular Orbital Study - PMC
    The existence of a rotational barrier of 2.875 kcal·mol−1 about the C-C bond in ethane has been known for many years [1,2,3,4,5,6,7,8,9]. There are two main ...Missing: strain | Show results with:strain
  58. [58]
    Theoretical Analysis of the Rotational Barrier of Ethane
    We developed a computational approach to probe the rotation barriers of ethane and its congeners in terms of steric repulsion, hyperconjugative interaction,
  59. [59]
    [PDF] Origin of methyl torsional potential barrier &#x2014; An overview
    This paper presents the evolution of views on methyl internal rotation potential barrier. Various mechanisms proposed for the origin of torsional barrier in.
  60. [60]
    Barrier to internal rotation in ethane from the microwave spectrum of ...
    Aug 1, 1979 · Rotational spectra of CH3CHD2 were observed by using a source‐modulation microwave spectrometer. Of 26 observed transitions, ...
  61. [61]
    Experimental data for C 2 H 6 (Ethane) - CCCBDB
    Calculated rotational constants for C2H6 (Ethane). Product of moments of inertia moments of inertia. 4070.847, amu3Å6, 1.86402951984082E-116, gm3 cm6. Geometric ...
  62. [62]
    The hindered rotor theory: A review - Dzib - 2022
    Oct 27, 2021 · The hindered rotor theory allows the thermodynamic properties of molecules with hindered rotations to be calculated.THE HINDERED ROTOR · HOW TO IDENTIFY AND... · UNCERTAINTY OF THE...
  63. [63]
    The barrier to internal rotation in ethane - ACS Publications
    Origin of Methyl Internal Rotation Barriers. Accounts of Chemical Research ... Methyl group dynamics in the crystalline alanine dipeptide: A combined ...
  64. [64]
    [PDF] Ideal Gas Thermodynamic Properties of Ethane and Propane
    Oct 29, 2009 · The hindered internal rotation of the two methyl groups in ethane has been a subject of intensive research since the 1930's.
  65. [65]
    Alkane Nomenclature
    By 1840 the French chemist Regnault was already calling the CH3 radical METHYL, where the meth part comes from Dumas's methylene and the yl does double duty ...
  66. [66]
    [PDF] Brief Guide to the Nomenclature of Organic Chemistry - IUPAC
    For naming purposes, a chemical compound is treated as a combination of a parent compound. (Section 5) and characteristic (functional) groups, one of which is.
  67. [67]
    REVISED NOMENCLATURE FOR RADICALS, IONS ... - iupac
    These recommendations are concerned only with formal operations for naming radicals, ions, radical ions, and neutral compounds with compensa- ting ionic centers ...<|control11|><|separator|>
  68. [68]
    Rule C-21 (Radicofunctional Nomenclature) - ACD/Labs
    Radical, CH3: Methyl. Complete name: Methyl alcohol. Note: In English and French the functional name and the radical are named as separate words. The radical ...