1,2-rearrangement

Revision as of 13:46, 4 September 2012 by WikiBot (talk | contribs) (Robot: Automated text replacement (-{{WikiDoc Cardiology Network Infobox}} +, -<references /> +{{reflist|2}}, -{{reflist}} +{{reflist|2}}))
(diff) ← Older revision | Latest revision (diff) | Newer revision → (diff)
Jump to navigation Jump to search
The printable version is no longer supported and may have rendering errors. Please update your browser bookmarks and please use the default browser print function instead.

A 1,2-rearrangement or 1,2-shift or Whitmore 1,2-shift [1] is an organic reaction where a substituent moves from one atom to another atom in a chemical compound. In a 1,2 shift the movement involves two adjacent atoms but moves over larger distances are possible. In the example below the substituent R moves from carbon atom C1 to C2.

1,2-shift
1,2-shift

The rearrangement is intramolecular and the starting compound and reaction product are structural isomers. The 1,2-rearrangement belongs to a broad class of chemical reactions called rearrangement reactions.

Reaction mechanism

A 1,2-rearrangement is often initialised by the formation of a reactive intermediate such as:

The driving force for the actual migration of a substituent in step two of the rearrangement is the formation of a more stable intermediate. For instance a tertiary carbocation is more stable than a secondary carbocation and therefore the SN1 reaction of neopentyl bromide with ethanol yields tert-pentyl ethyl ether.

Carbocation rearrangements are more common than the carbanion or radical counterparts. This observation can be explained on the basis of Hückel's rule. A cyclic carbocationic transition state is aromatic and stabilized because it holds 2 electrons. In an anionic transition state on the other hand 4 electrons are present thus antiaromatic and destabilized. A radical transition state is neither stabilized or destabilized.

The most important carbocation 1,2-shift is the Wagner-Meerwein rearrangement. A carbanionic 1,2-shift is involved in the benzilic acid rearrangement.

Radical 1,2-rearrangements

The first radical 1,2-rearrangement reported by Heinrich Otto Wieland in 1911 [2] was the conversion of bis(triphenylmethyl)peroxide 1 to the tetraphenylethane 2.

Radical 1,2-rearrangement
Radical 1,2-rearrangement

The reaction proceeds through the triphenylmethoxyl radical A, a rearrangement to diphenylphenoxymethyl C and its dimerization. It is unclear to this day whether in this rearrangement the cyclohexadienyl radical intermediate B is a transition state or a reactive intermediate as it (or any other such species) has thus far eluded detection by ESR spectroscopy [3].

An example of a less common radical 1,2-shift can be found in the gas phase pyrolysis of certain polycyclic aromatic compounds [4]. The energy required in an aryl radical for the 1,2-shift can be high (up to 60 kcal/mol or 250 kJ/mol) but much less than that required for a proton abstraction to an aryne (82 Kcal/mole). In alkene radicals proton abstraction to an alkyne is preferred.

Aryl radical 1,2-shift in a helicene
Aryl radical 1,2-shift in a helicene

1,2 Rearrangements

The following mechanisms involve a 1,2-rearrangement:

1,3-Rearrangements

1,3-rearrangements take place over 3 carbon atoms. Examples:

References

  1. Whitmore, Frank C. (1932). "The common basis of molecular rearrangements". J. Am. Chem. Soc. 54 (8): 3274–3283. doi:10.1021/ja01347a037.
  2. Wieland, H. Chem. Ber. 1911, 44, 2550-2556.
  3. Isomerization of Triphenylmethoxyl and 1,1-Diphenylethoxyl Radicals. Revised Assignment of the Electron-Spin Resonance Spectra of Purported Intermediates Formed during the Ceric Ammonium Nitrate Mediated Photooxidation of Aryl Carbinols K. U. Ingold, Manuel Smeu, and Gino A. DiLabio J. Org. Chem.; 2006; 71(26) pp 9906 - 9908; (Note) doi:10.1021/jo061898z
  4. Brooks, Michele A. (1999). "1,2-Shifts of Hydrogen Atoms in Aryl Radicals". J. Am. Chem. Soc. 121 (23): 5444–5449. doi:10.1021/ja984472d. Unknown parameter |coauthors= ignored (help)

Template:WikiDoc Sources