Wittig reaction

Last updated

Contents

Wittig reaction
Named after Georg Wittig
Reaction type Coupling reaction
Reaction
aldehyde or ketone
+
triphenyl phosphonium ylide
alkene
+
(Ph)3P=O
Conditions
Typical solventstypically THF or diethyl ether
Identifiers
March's Advanced Organic Chemistry16–44 (6th ed.)
Organic Chemistry Portal wittig-reaction
RSC ontology ID RXNO:0000015  Yes check.svgY
 X mark.svgN (what is this?)    (verify)

The Wittig reaction or Wittig olefination is a chemical reaction of an aldehyde or ketone with a triphenyl phosphonium ylide called a Wittig reagent. Wittig reactions are most commonly used to convert aldehydes and ketones to alkenes. [1] [2] [3] Most often, the Wittig reaction is used to introduce a methylene group using methylenetriphenylphosphorane (Ph3P=CH2). Using this reagent, even a sterically hindered ketone such as camphor can be converted to its methylene derivative.

Wittig Reaktion.svg

Reaction mechanism

Mechanistic studies have focused on unstabilized ylides, because the intermediates can be followed by NMR spectroscopy. The existence and interconversion of the betaine (3a and 3b) is subject of ongoing research. [4] For lithium-free Wittig reactions, studies support a concerted formation of the oxaphosphetane without intervention of a betaine. In particular, phosphonium ylides 1 react with carbonyl compounds 2 via a [2+2] cycloaddition that is sometimes described as having [π2s+π2a] topology to directly form the oxaphosphetanes 4a and 4b. Under lithium-free conditions, the stereochemistry of the product 5 is due to the kinetically controlled addition of the ylide 1 to the carbonyl 2. When lithium is present, there may be equilibration of the intermediates, possibly via betaine species 3a and 3b. [5] [6] [7] Bruce E. Maryanoff and A. B. Reitz identified the issue about equilibration of Wittig intermediates and termed the process "stereochemical drift". For many years, the stereochemistry of the Wittig reaction, in terms of carbon-carbon bond formation, had been assumed to correspond directly with the Z/E stereochemistry of the alkene products. However, certain reactants do not follow this simple pattern. Lithium salts can also exert a profound effect on the stereochemical outcome. [8]

The mechanism of the Wittig reaction Wittig Reaction Mechanism Update.png
The mechanism of the Wittig reaction

Mechanisms differ for aliphatic and aromatic aldehydes and for aromatic and aliphatic phosphonium ylides. Evidence suggests that the Wittig reaction of unbranched aldehydes under lithium-salt-free conditions do not equilibrate and are therefore under kinetic reaction control. [9] [10] E. Vedejs has put forth a theory to explain the stereoselectivity of stabilized and unstabilized Wittig reactions. [11]

Strong evidence indicated that under Li-free conditions, Wittig reactions involving unstabilized (R1= alkyl, H), semistabilized (R1 = aryl), and stabilized (R1 = EWG) Wittig reagents all proceed via a [2+2]/retro-[2+2] mechanism under kinetic control, with oxaphosphetane as the one and only intermediate. [12]

Scope and limitations

Functional group tolerance

The Schlosser variant of the Wittig reaction Schlosser Wittig.gif
The Schlosser variant of the Wittig reaction

The Wittig reagents generally tolerate carbonyl compounds containing several kinds of functional groups such as OH, OR, aromatic nitro, epoxide, and ester groups. Even C=O and nitrile groups can be present if conjugated with the ylide- these are the stabilised ylides mentioned above. Bis-ylides (containing two P=C bonds) have also been made and used successfully. [13] There can be a problem with sterically hindered ketones, where the reaction may be slow and give poor yields, particularly with stabilized ylides, and in such cases the Horner–Wadsworth–Emmons (HWE) reaction (using phosphonate esters) is preferred. Another reported limitation is the often labile nature of aldehydes, which can oxidize, polymerize or decompose. In a so-called tandem oxidation-Wittig process the aldehyde is formed in situ by oxidation of the corresponding alcohol. [14]

Stereochemistry

For the reaction with aldehydes, the double bond geometry is readily predicted based on the nature of the ylide. With unstabilised ylides (R3 = alkyl) this results in (Z)-alkene product with moderate to high selectivity. If the reaction is performed in dimethylformamide in the presence of lithium iodide or sodium iodide, the product is almost exclusively the Z-isomer. [15] With stabilized ylides (R3 = ester or ketone), the (E)-alkene is formed with high selectivity. The (E)/(Z) selectivity is often poor with semistabilized ylides (R3 = aryl). [16]

To obtain the (E)-alkene for unstabilized ylides, the Schlosser modification of the Wittig reaction can be used. Alternatively, the Julia olefination and its variants also provide the (E)-alkene selectively. Ordinarily, the Horner–Wadsworth–Emmons reaction provides the (E)-enoate (α,β-unsaturated ester), just as the Wittig reaction does. To obtain the (Z)-enolate, the Still-Gennari modification of the Horner-Wadsworth-Emmons reaction can be used.

Schlosser modification

The main limitation of the traditional Wittig reaction is that the reaction proceeds mainly via the erythro betaine intermediate, which leads to the Z-alkene. The erythro betaine can be converted to the threo betaine using phenyllithium at low temperature. [17] This modification affords the E-alkene.

Allylic alcohols can be prepared by reaction of the betaine ylide with a second aldehyde. [18] For example:

An example of the Schlosser variant of the Wittig reaction Corey Schlosser Wittig.png
An example of the Schlosser variant of the Wittig reaction

Example

An example of its use is in the synthesis of leukotriene A methyl ester. [19] [20] The first step uses a stabilised ylide, where the carbonyl group is conjugated with the ylide preventing self condensation, although unexpectedly this gives mainly the cis product. The second Wittig reaction uses a non-stabilised Wittig reagent, and as expected this gives mainly the cis product.

An example of the use of the Wittig reaction in synthesis, making leukotriene A methyl ester LeukotrieneA synthesis.png
An example of the use of the Wittig reaction in synthesis, making leukotriene A methyl ester

History

The Wittig reaction was reported in 1954 by Georg Wittig and his coworker Ulrich Schöllkopf. In part for this contribution, Wittig was awarded the Nobel Prize in Chemistry in 1979. [21] [22]

See also

Related Research Articles

<span class="mw-page-title-main">Elias James Corey</span> American chemist (born 1928)

Elias James Corey is an American organic chemist. In 1990, he won the Nobel Prize in Chemistry "for his development of the theory and methodology of organic synthesis", specifically retrosynthetic analysis. Regarded by many as one of the greatest living chemists, he has developed numerous synthetic reagents, methodologies and total syntheses and has advanced the science of organic synthesis considerably.

Hydroboration–oxidation reaction is a two-step hydration reaction that converts an alkene into an alcohol. The process results in the syn addition of a hydrogen and a hydroxyl group where the double bond had been. Hydroboration–oxidation is an anti-Markovnikov reaction, with the hydroxyl group attaching to the less-substituted carbon. The reaction thus provides a more stereospecific and complementary regiochemical alternative to other hydration reactions such as acid-catalyzed addition and the oxymercuration–reduction process. The reaction was first reported by Herbert C. Brown in the late 1950s and it was recognized in his receiving the Nobel Prize in Chemistry in 1979.

An ylide or ylid is a neutral dipolar molecule containing a formally negatively charged atom (usually a carbanion) directly attached to a heteroatom with a formal positive charge (usually nitrogen, phosphorus or sulfur), and in which both atoms have full octets of electrons. The result can be viewed as a structure in which two adjacent atoms are connected by both a covalent and an ionic bond; normally written X+–Y. Ylides are thus 1,2-dipolar compounds, and a subclass of zwitterions. They appear in organic chemistry as reagents or reactive intermediates.

In organic chemistry, a nucleophilic addition reaction is an addition reaction where a chemical compound with an electrophilic double or triple bond reacts with a nucleophile, such that the double or triple bond is broken. Nucleophilic additions differ from electrophilic additions in that the former reactions involve the group to which atoms are added accepting electron pairs, whereas the latter reactions involve the group donating electron pairs.

The 1,3-dipolar cycloaddition is a chemical reaction between a 1,3-dipole and a dipolarophile to form a five-membered ring. The earliest 1,3-dipolar cycloadditions were described in the late 19th century to the early 20th century, following the discovery of 1,3-dipoles. Mechanistic investigation and synthetic application were established in the 1960s, primarily through the work of Rolf Huisgen. Hence, the reaction is sometimes referred to as the Huisgen cycloaddition. 1,3-dipolar cycloaddition is an important route to the regio- and stereoselective synthesis of five-membered heterocycles and their ring-opened acyclic derivatives. The dipolarophile is typically an alkene or alkyne, but can be other pi systems. When the dipolarophile is an alkyne, aromatic rings are generally produced.

<span class="mw-page-title-main">Organoboron chemistry</span> Study of compounds containing a boron-carbon bond

Organoboron chemistry or organoborane chemistry is the chemistry of organoboron compounds or organoboranes, which are chemical compounds of boron and carbon that are organic derivatives of borane (BH3), for example trialkyl boranes..

<span class="mw-page-title-main">Baeyer–Villiger oxidation</span> Organic reaction

The Baeyer–Villiger oxidation is an organic reaction that forms an ester from a ketone or a lactone from a cyclic ketone, using peroxyacids or peroxides as the oxidant. The reaction is named after Adolf von Baeyer and Victor Villiger who first reported the reaction in 1899.

<span class="mw-page-title-main">Nicolaou Taxol total synthesis</span>

The Nicolaou Taxol total synthesis, published by K. C. Nicolaou and his group in 1994 concerns the total synthesis of taxol. Taxol is an important drug in the treatment of cancer but also expensive because the compound is harvested from a scarce resource, namely the pacific yew.

<span class="mw-page-title-main">Peterson olefination</span> Chemical reaction

The Peterson olefination is the chemical reaction of α-silyl carbanions with ketones to form a β-hydroxysilane (2) which eliminates to form alkenes (3).

<span class="mw-page-title-main">Johnson–Corey–Chaykovsky reaction</span> Chemical reaction in organic chemistry

The Johnson–Corey–Chaykovsky reaction is a chemical reaction used in organic chemistry for the synthesis of epoxides, aziridines, and cyclopropanes. It was discovered in 1961 by A. William Johnson and developed significantly by E. J. Corey and Michael Chaykovsky. The reaction involves addition of a sulfur ylide to a ketone, aldehyde, imine, or enone to produce the corresponding 3-membered ring. The reaction is diastereoselective favoring trans substitution in the product regardless of the initial stereochemistry. The synthesis of epoxides via this method serves as an important retrosynthetic alternative to the traditional epoxidation reactions of olefins.

<span class="mw-page-title-main">Danishefsky Taxol total synthesis</span>

The Danishefsky Taxol total synthesis in organic chemistry is an important third Taxol synthesis published by the group of Samuel Danishefsky in 1996 two years after the first two efforts described in the Holton Taxol total synthesis and the Nicolaou Taxol total synthesis. Combined they provide a good insight in the application of organic chemistry in total synthesis.

The Barton–Kellogg reaction is a coupling reaction between a diazo compound and a thioketone, giving an alkene by way of an episulfide intermediate. The Barton–Kellogg reaction is also known as Barton–Kellogg olefination and Barton olefin synthesis.

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

<span class="mw-page-title-main">Tebbe's reagent</span> Chemical compound

Tebbe's reagent is the organometallic compound with the formula (C5H5)2TiCH2ClAl(CH3)2. It is used in the methylidenation of carbonyl compounds, that is it converts organic compounds containing the R2C=O group into the related R2C=CH2 derivative. It is a red solid that is pyrophoric in the air, and thus is typically handled with air-free techniques. It was originally synthesized by Fred Tebbe at DuPont Central Research.

<span class="mw-page-title-main">Petasis reagent</span> Chemical compound

The Petasis reagent, named after Nicos A. Petasis, is an organotitanium compound with the formula Cp2Ti(CH3)2. It is an orange-colored solid.

The Julia olefination (also known as the Julia–Lythgoe olefination) is the chemical reaction used in organic chemistry of phenyl sulfones (1) with aldehydes (or ketones) to give alkenes (olefins)(3) after alcohol functionalization and reductive elimination using sodium amalgam or SmI2. The reaction is named after the French chemist Marc Julia.

<span class="mw-page-title-main">Takai olefination</span>

Takai olefination in organic chemistry describes the organic reaction of an aldehyde with a diorganochromium compound to form an alkene. It is a name reaction, referencing Kazuhiko Takai, who first reported it in 1986. In the original reaction, the organochromium species is generated from iodoform or bromoform and an excess of chromium(II) chloride and the product is a vinyl halide. One main advantage of this reaction is the E-configuration of the double bond that is formed. According to the original report, existing alternatives such as the Wittig reaction only gave mixtures.

Desulfonylation reactions are chemical reactions leading to the removal of a sulfonyl group from organic compounds. As the sulfonyl functional group is electron-withdrawing, methods for cleaving the sulfur–carbon bonds of sulfones are typically reductive in nature. Olefination or replacement with hydrogen may be accomplished using reductive desulfonylation methods.

<span class="mw-page-title-main">Methylenetriphenylphosphorane</span> Chemical compound

Methylenetriphenylphosphorane is an organophosphorus compound with the formula Ph3PCH2. It is the parent member of the phosphorus ylides, popularly known as Wittig reagents. It is a highly polar, highly basic species.

Carbonyl olefin metathesis is a type of metathesis reaction that entails, formally, the redistribution of fragments of an alkene and a carbonyl by the scission and regeneration of carbon-carbon and carbon-oxygen double bonds respectively. It is a powerful method in organic synthesis using simple carbonyls and olefins and converting them into less accessible products with higher structural complexity.

References

  1. Maercker, A. Org. React.1965, 14, 270–490.
  2. W. Carruthers, Some Modern Methods of Organic Synthesis, Cambridge University Press, Cambridge, UK, 1971, 81–90. ( ISBN   0-521-31117-9)
  3. R. W. Hoffmann (2001). "Wittig and His Accomplishments: Still Relevant Beyond His 100th Birthday". Angewandte Chemie International Edition . 40 (8): 1411–1416. doi:10.1002/1521-3773(20010417)40:8<1411::AID-ANIE1411>3.0.CO;2-U. PMID   11317288.
  4. E. Vedejs & C. F. Marth (1990). "Mechanism of Wittig reaction: evidence against betaine intermediates". J. Am. Chem. Soc. 112 (10): 3905–3909. doi:10.1021/ja00166a026.
  5. Bruce E. Maryanoff, A. B. Reitz, M. S. Mutter, R. R. Inners, and H. R. Almond, Jr., "Detailed Rate Studies on the Wittig Reaction of Non-Stabilized Phosphorus Ylides via 31P, 1H, and 13C NMR Spectroscopy. Insight into Kinetic vs. Thermodynamic Control of Stereochemistry", J. Am. Chem. Soc., 107, 1068–1070 (1985)
  6. Bruce E. Maryanoff, A. B. Reitz, D. W. Graden, and H. R. Almond, Jr., "NMR Rate Study on the Wittig Reaction of 2,2-Dimethylpropanal and Tributylbutylidene-phosphorane", Tetrahedron Lett., 30, 1361–1364 (1989)
  7. Bruce E. Maryanoff, A. B. Reitz, M. S. Mutter, R. R. Inners, H. R. Almond, Jr., R. R. Whittle, and R. A. Olofson, "Stereochemistry and Mechanism of the Wittig Reaction. Diastereomeric Reaction Intermediates and Analysis of the Reaction Course", J. Am. Chem. Soc., 108, 7664–7678 (1986)
  8. A. B. Reitz, S. O. Nortey, A. D. Jordan, Jr., M. S. Mutter, and Bruce E. Maryanoff, "Dramatic Concentration Dependence of Stereochemistry in the Wittig Reaction. Examination of the Lithium-Salt Effect", J. Org. Chem., 51, 3302–3308 (1986)
  9. E. Vedejs, C. F. Marth and R. Ruggeri (1988). "Substituent effects and the Wittig mechanism: the case of stereospecific oxaphosphetane decomposition". J. Am. Chem. Soc. 110 (12): 3940–48. doi:10.1021/ja00220a036.
  10. E. Vedejs & C. F. Marth (1988). "Mechanism of the Wittig reaction: the role of substituents at phosphorus". J. Am. Chem. Soc. 110 (12): 3948–3958. doi:10.1021/ja00220a037.
  11. Vedejs, E.; Peterson, M. J. Top. Stereochem.1994, 21, 1.
  12. Byrne, Peter A.; Gilheany, Declan G. (2013). "The modern interpretation of the Wittig reaction mechanism". Chemical Society Reviews. 42 (16): 6670–96. doi:10.1039/c3cs60105f. hdl: 10197/4939 . ISSN   0306-0012. PMID   23673458.
  13. B. E. Maryanoff & A. B. Reitz (1989). "The Wittig olefination reaction and modifications involving phosphoryl-stabilized carbanions. Stereochemistry, mechanism, and selected synthetic aspects". Chem. Rev. 89 (4): 863–927. doi:10.1021/cr00094a007.
  14. Richard J. K. Taylor, Leonie Campbell, and Graeme D. McAllister (2008). "(±) trans-3,3'-(1,2-Cyclopropanediyl)bis-2-(E)-propenoic Acid, Diethyl Ester: Tandem Oxidation Procedure (TOP) using MnO2 Oxidation-Stabilized Phosphorane Trapping" (PDF). Organic Syntheses . 85: 15–26.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  15. L. D. Bergelson & M. M. Shemyakin (1964). "Synthesis of Naturally Occurring Unsaturated Fatty Acids by Sterically Controlled Carbonyl Olefination". Angew. Chem. 3 (4): 250–260. doi:10.1002/anie.196402501.
  16. Robiette, Raphaël; Richardson, Jeffery; Aggarwal, Varinder K.; Harvey, Jeremy N. (1 February 2006). "Reactivity and Selectivity in the Wittig Reaction: A Computational Study". Journal of the American Chemical Society. 128 (7): 2394–2409. doi:10.1021/ja056650q. ISSN   0002-7863. PMID   16478195.
  17. M. Schlosser & K. F. Christmann (1966). "Trans-Selective Olefin Syntheses". Angewandte Chemie International Edition in English . 5 (1): 126. doi:10.1002/anie.196601261.
  18. E. J. Corey and H. Yamamoto (1970). "Modification of the Wittig reaction to permit the stereospecific synthesis of certain trisubstituted olefins. Stereospecific synthesis of α-santalol". J. Am. Chem. Soc. 92 (1): 226–228. doi:10.1021/ja00704a052.
  19. I. Ernest, A. J. Main and R. Menasse (1982). "Synthesis of the 7-cis isomer of the natural leukotriene d4". Tetrahedron Letters . 23 (2): 167–170. doi:10.1016/S0040-4039(00)86776-3.
  20. E. J. Corey, D. A. Clark, G. Goto, A. Marfat, C. Mioskowski, B. Samuelsson and S. Hammarstroem (1980). "Stereospecific total synthesis of a "slow reacting substance" of anaphylaxis, leukotriene C-1". J. Am. Chem. Soc. 102 (4): 1436–1439. doi:10.1021/ja00524a045.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  21. Georg Wittig, Ulrich Schöllkopf (1954). "Über Triphenyl-phosphin-methylene als olefinbildende Reagenzien I". Chemische Berichte . 87 (9): 1318. doi:10.1002/cber.19540870919.
  22. Georg Wittig; Werner Haag (1955). "Über Triphenyl-phosphin-methylene als olefinbildende Reagenzien II". Chemische Berichte . 88 (11): 1654–1666. doi:10.1002/cber.19550881110.