Horner–Wadsworth–Emmons reaction

Source: Wikipedia, the free encyclopedia.

Horner–Wadsworth–Emmons reaction
Named after Leopold Horner
William S. Wadsworth
William D. Emmons
Reaction type Coupling reaction
Identifiers
Organic Chemistry Portal wittig-horner-reaction
RSC ontology ID RXNO:0000056

The Horner–Wadsworth–Emmons (HWE) reaction is a chemical reaction used in organic chemistry of stabilized phosphonate carbanions with aldehydes (or ketones) to produce predominantly E-alkenes.[1]

The Horner–Wadsworth–Emmons reaction
The Horner–Wadsworth–Emmons reaction

In 1958, Leopold Horner published a modified Wittig reaction using phosphonate-stabilized carbanions.[2][3] William S. Wadsworth and William D. Emmons further defined the reaction.[4][5]

In contrast to

extraction
.

Several reviews have been published.[6][7][8][9][10][11]

Reaction mechanism

Deprotonation by base (B-) to generate the phosphonate carbanion

The Horner–Wadsworth–Emmons reaction begins with the

rate-limiting step.[12] If R2 = H, then intermediates 3a and 4a and intermediates 3b and 4b can interconvert with each other.[13] The final elimination of oxaphosphetanes 4a and 4b yield (E)-alkene 5 and (Z)-alkene 6, with the by-product being a dialkyl-phosphate
.

The mechanism of the Horner-Wadsworth-Emmons reaction
The mechanism of the Horner-Wadsworth-Emmons reaction

The ratio of alkene

stereochemical outcome of the initial carbanion addition and upon the ability of the intermediates to equilibrate
.

The

Stereoselectivity

The Horner–Wadsworth–Emmons reaction favours the formation of (E)-alkenes. In general, the more equilibration amongst intermediates, the higher the selectivity for (E)-alkene formation.

Disubstituted alkenes

Thompson and Heathcock have performed a systematic study of the reaction of methyl 2-(dimethylphosphono)acetate with various aldehydes.[16] While each effect was small, they had a cumulative effect making it possible to modify the stereochemical outcome without modifying the structure of the phosphonate. They found greater (E)-stereoselectivity with the following conditions:

  • Increasing steric bulk of the aldehyde
  • Higher reaction temperatures (23 °C over −78 °C)
  • Li > Na > K salts


In a separate study, it was found that bulky phosphonate and bulky electron-withdrawing groups enhance E-alkene selectivity.

Trisubstituted alkenes

The steric bulk of the phosphonate and electron-withdrawing groups plays a critical role in the reaction of α-branched phosphonates with aliphatic aldehydes.[17]

Example of the Horner–Wadsworth–Emmons reaction with branched phosphonates
Example of the Horner–Wadsworth–Emmons reaction with branched phosphonates
R1 R2 Ratio of alkenes
( E : Z )
Methyl
Methyl 5 : 95
Methyl Ethyl 10 : 90
Ethyl Ethyl 40 : 60
Isopropyl
Ethyl 90 : 10
Isopropyl Isopropyl 95 : 5

Aromatic
aldehydes produce almost exclusively (E)-alkenes. In case (Z)-alkenes from aromatic aldehydes are needed, the Still–Gennari modification (see below) can be used.

Olefination of ketones

The stereoselectivity of the Horner–Wadsworth–Emmons reaction of ketones is poor to modest.

Variations

Base sensitive substrates

Since many substrates are not stable to

DBU.[18] Rathke extended this to lithium or magnesium halides with triethylamine.[19] Several other bases have been found effective.[20][21][22]

Still modification

) nearly exclusive Z-alkene production can be achieved.

The Still modification of the Horner–Wadsworth–Emmons reaction
The Still modification of the Horner–Wadsworth–Emmons reaction

Ando has suggested that the use of electron-deficient phosphonates accelerates the elimination of the oxaphosphetane intermediates.[26]

See also

References

  1. ^ Leopold Horner; Hoffmann, H. M. R.; Wippel, H. G. Ber. 1958, 91, 61–63.
  2. ^ Horner, L.; Hoffmann, H. M. R.; Wippel, H. G.; Klahre, G. Ber. 1959, 92, 2499–2505.
  3. )
  4. ^ Wadsworth, W. S., Jr.; Emmons, W. D. Organic Syntheses, Coll. Vol. 5, p. 547 (1973); Vol. 45, p. 44 (1965). (Article)
  5. ^ Wadsworth, W. S., Jr. Org. React. 1977, 25, 73–253. (Review)
  6. )
  7. ^ Kelly, S. E. Compr. Org. Synth. 1991, 1, 729–817. (Review)
  8. )
  9. Curr. Org. Chem.
    2012, 16, 2206–2230 (Review)
  10. ^ Bisceglia, J. A., Orelli, L. R. Curr. Org. Chem. 2015, 19, 744–775 (Review)
  11. ^ Larsen, R. O.; Aksnes, G. Phosphorus Sulfur 1983, 15, 218–219.
  12. J. Chem Soc., Chem. Commun.
    1970, 1308–09.
  13. )
  14. )
  15. )
  16. ^ Nagaoka, H.; Kishi, Y. Tetrahedron 1981, 37, 3873–3888.
  17. ^ Blanchette, M. A.; Choy, W.; Davis, J. T.; Essenfeld, A. P.; Masamune, S.; Roush, W. R.; Sakai, T. Tetrahedron Letters 1984, 25, 2183–2186.
  18. )
  19. ^ Paterson, I.; Yeung, K.-S.; Smaill, J. B. Synlett 1993, 774.
  20. ^ Simoni, D.; Rossi, M.; Rondanin, R.; Mazzali, A.; Baruchello, R.; Malagutti, C.; Roberti, M.; Invidiata, F. P. Org. Letters 2000, 2, 3765–3768.
  21. ^ Blasdel, L. K.; Myers, A. G. Org. Letters 2005, 7, 4281–4283.
  22. ^ Still, W. C.; Gennari, C. Tetrahedron Letters 1983, 24, 4405–4408.
  23. S2CID 216228029
    .
  24. ^ Patois, C.; Savignac, P.; About-Jaudet, E.; Collignon, N. Organic Syntheses, Coll. Vol. 9, p. 88 (1998); Vol. 73, p. 152 (1996). (Article)
  25. )