Methyl radical

Summary

Methyl radical is an organic compound with the chemical formula CH
3
(also written as [CH
3
]
). It is a metastable colourless gas, which is mainly produced in situ as a precursor to other hydrocarbons in the petroleum cracking industry. It can act as either a strong oxidant or a strong reductant, and is quite corrosive to metals.

Methyl radical
Names
IUPAC name
Methyl[1]
Identifiers
  • 2229-07-4 checkY
3D model (JSmol)
  • Interactive image
1696831
ChEBI
  • CHEBI:29309
ChemSpider
  • 2299212 checkY
57
MeSH Methyl+radical
  • 3034819
UNII
  • S19006ZD7R checkY
  • DTXSID60176836 Edit this at Wikidata
  • InChI=1S/CH3/h1H3 checkY
    Key: WCYWZMWISLQXQU-UHFFFAOYSA-N checkY
  • [CH3]
Properties
CH3
Molar mass 15.035 g·mol−1
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
Infobox references

Chemical properties edit

Its first ionization potential (yielding the methenium ion, CH+
3
) is 9.837±0.005 eV.[2]

Redox behaviour edit

The carbon centre in methyl can bond with electron-donating molecules by reacting:

CH
3
+ RRCH
3

Because of the capture of the nucleophile (R), methyl has oxidising character. Methyl is a strong oxidant with organic chemicals. However, it is equally a strong reductant with chemicals such as water. It does not form aqueous solutions, as it reduces water to produce methanol and elemental hydrogen:

CH
3
+ 2 H
2
O
→ 2 CH
3
OH
+ H
2

Structure edit

The molecular geometry of the methyl radical is trigonal planar (bond angles are 120°), although the energy cost of distortion to a pyramidal geometry is small. All other electron-neutral, non-conjugated alkyl radicals are pyramidalized to some extent, though with very small inversion barriers. For instance, the t-butyl radical has a bond angle of 118° with a 0.7 kcal/mol (2.9 kJ/mol) barrier to pyramidal inversion. On the other hand, substitution of hydrogen atoms by more electronegative substituents leads to radicals with a strongly pyramidal geometry (112°), such as the trifluoromethyl radical, CF
3
, with a much more substantial inversion barrier of around 25 kcal/mol (100 kJ/mol).[3]

Chemical reactions edit

Methyl undergoes the typical chemical reactions of a radical. Below approximately 1,100 °C (1,400 K), it rapidly dimerises to form ethane. Upon treatment with an alcohol, it converts to methane and either an alkoxy or hydroxyalkyl. Reduction of methyl gives methane. When heated above, at most, 1,400 °C (1,700 K), methyl decomposes to produce methylidyne and elemental hydrogen, or to produce methylene and atomic hydrogen:

CH
3
→ CH + H
2
CH
3
CH
2
+ H

Methyl is very corrosive to metals, forming methylated metal compounds:

M + n CH
3
→ M(CH3)n

Production edit

Biosynthesis edit

Some radical SAM enzymes generate methyl radicals by reduction of S-adenosylmethionine.[4]

Acetone photolysis edit

It can be produced by the ultraviolet photodissociation of acetone vapour at 193 nm:[5]

C
3
H
6
O
→ CO + 2 CH
3

Halomethane photolysis edit

It is also produced by the ultraviolet dissociation of halomethanes:

CH
3
X
→ X + CH
3

Methane oxidation edit

It can also be produced by the reaction of methane with the hydroxyl radical:

OH + CH4CH
3
+ H2O

This process begins the major removal mechanism of methane from the atmosphere. The reaction occurs in the troposphere or stratosphere. In addition to being the largest known sink for atmospheric methane, this reaction is one of the most important sources of water vapor in the upper atmosphere.

This reaction in the troposphere gives a methane lifetime of 9.6 years. Two more minor sinks are soil sinks (160 year lifetime) and stratospheric loss by reaction with OH, Cl and O1D in the stratosphere (120 year lifetime), giving a net lifetime of 8.4 years.[6]

Azomethane pyrolysis edit

Methyl radicals can also be obtained by pyrolysis of azomethane, CH3N=NCH3, in a low-pressure system.

In the interstellar medium edit

Methyl was discovered in interstellar medium in 2000 by a team led by Helmut Feuchtgruber who detected it using the Infrared Space Observatory. It was first detected in molecular clouds toward the centre of the Milky Way.[7]

References edit

  1. ^ International Union of Pure and Applied Chemistry (2014). Nomenclature of Organic Chemistry: IUPAC Recommendations and Preferred Names 2013. The Royal Society of Chemistry. p. 1051. doi:10.1039/9781849733069. ISBN 978-0-85404-182-4.
  2. ^ Golob, L.; Jonathan, N.; Morris, A.; Okuda, M.; Ross, K.J. (1972). "The first ionization potential of the methyl radical as determined by photoelectron spectroscopy". Journal of Electron Spectroscopy and Related Phenomena. 1 (5): 506–508. doi:10.1016/0368-2048(72)80022-7.
  3. ^ Anslyn E.V. and Dougherty D.A., Modern Physical Organic Chemistry (University Science Books, 2006), p.57
  4. ^ Ribbe, M. W.; Hu, Y.; Hodgson, K. O.; Hedman, B. (2014). "Biosynthesis of Nitrogenase Metalloclusters". Chemical Reviews. 114 (8): 4063–4080. doi:10.1021/cr400463x. PMC 3999185. PMID 24328215.
  5. ^ Hall, G. E.; Vanden Bout, D.; Sears, Trevor J. (1991). "Photodissociation of acetone at 193 nm: Rotational- and vibrational-state distributions of methyl fragments by diode laser absorption/gain spectroscopy". The Journal of Chemical Physics. AIP Publishing. 94 (6): 4182. Bibcode:1991JChPh..94.4182H. doi:10.1063/1.460741.
  6. ^ "Trace Gases: Current Observations, Trends, and Budgets". Climate Change 2001, IPCC Third Assessment Report. IPCC/United Nations Environment Programme.
  7. ^ "ISO detects a new molecule in interstellar space". Science & Technology. European Space Agency. Retrieved 17 June 2013.