2,3-Wittig rearrangement

Last updated

The [2,3]-Wittig rearrangement is the transformation of an allylic ether into a homoallylic alcohol via a concerted, pericyclic process. Because the reaction is concerted, it exhibits a high degree of stereocontrol, and can be employed early in a synthetic route to establish stereochemistry. The Wittig rearrangement requires strongly basic conditions, however, as a carbanion intermediate is essential. [1,2]-Wittig rearrangement is a competitive process. [1]

Contents

Introduction

[2,3]-Sigmatropic rearrangements occur for a variety of groups X and Y (see below). When X is a carbanion and Y an alkoxide, the rearrangement is called the [2,3]-Wittig rearrangement and the products are pent-1-en-5-ols. The [1,2]-Wittig rearrangement, which produces isomeric pent-5-en-1-ols, is a competitive process that takes place at high temperatures. [2] Because of the high atom economy and stereoselectivity of the [2,3]-rearrangement, it has gained considerable synthetic utility. The carbanion is generated by direct lithiation of moderately acidic substrates, tin transmetallation, or reductive lithiation of O,S-acetals. Stereoselective methods employing chiral starting materials have been used to effect either asymmetric induction or simple diastereoselection [3]

(1)

23Gen.png

Mechanism and stereochemistry

Prevailing mechanism

After carbanion formation, the [2,3]-Wittig rearrangement is rapid and selective at low temperatures. However, if the reaction mixture is allowed to reach temperatures above −60 °C, [1,2]-rearrangement becomes competitive. [4]

(2)

23Mech1.png

The postulated transition state possesses a five-membered, envelope-like structure. [5] The group attached to the carbanion (G) can occupy either a pseudoequatorial or pseudoaxial position, although the former is usually preferred. Large substituents on the other side of the ether oxygen prefer to occupy the exo position (RE) to avoid A1,3 strain. These restrictions lead to a preference for the syn product from (Z) isomers and anti products from (E) isomers; however, some exceptions to this rule are known. [6]

(3)

23Mech2.png

Stereoselective variants

Stereoselective variants of the [2,3]-Wittig rearrangement have employed three strategies: diastereoselection based on an existing, established stereocenter, placement of a chiral auxiliary on the starting material whose configuration is unaffected by the reaction, and the use of a chiral base. The relative diastereoselection strategy works well only for a limited number of G groups, but usually results in high yields because no chiral auxiliary group needs to be removed or modified. The stereocenter opposite the carbanion usually must be tertiary (rather than quaternary) in order to enforce the placement of the largest substituent in the RE position. [7]

(4)

23Stereo1.png

The asymmetric induction approach relies on stereocenters already set in the starting material that are unaffected by the reaction (chiral auxiliaries). The most success has been achieved by placing these stereocenters either in the G group [8] or in a substituent attached to the end of the double bond. [9] Diastereomeric ratios in excess of 90:10 are common for these reactions; however, removal of the chiral auxiliary is sometimes difficult. [10]

(5)

23Stereo2.png

The use of chiral bases has afforded enantioenriched rearrangement products in a few cases, [11] although this method does not appear to be general. Enantioselectivity in these reactions is often low, suggesting that the association between the conjugate acid of the base and the rearranging carbanion is likely weak.

(6)

23Stereo3.png

Scope and limitations

A variety of allylic ethers undergo the Wittig rearrangement—the fundamental requirement is the ability to generate the appropriate carbanion in the substrate. This demands either acidic hydrogens, a reducible functional group, or a carbon-metal bond. Historically, alkenyl, alkynyl, and phenyl groups have been used to acidify the α position. Free terminal alkynes are tolerated, although yields are higher when silyl-protected alkynes are used. [12]

(7)

23Scope3.png

When an alkene is used as the anion-stabilizing group G, issues of selectivity arise concerning the site of the carbanion. Anion-stabilizing groups such as (trimethyl)silyl or methylthio provide essentially complete site selectivity. [13]

(8)

23Scope1.png

Carbonyl groups may also be used as the anion-stabilizing group; carbonyl groups are particularly useful for asymmetric rearrangements that employ chiral auxiliaries. [14]

(9)

23Scope4.png

A highly enantioselective method employing chromium carbonyl complexes involves the use of the acidified phenyl ring as an anion-stabilizing group. [15]

(10)

23Scope5.png

That the substrate must contain acidic hydrogens adjacent to the ether oxygen was a significant limitation of the original reaction. Thus, the development of transmetallation methods that allowed the selective generation of carbanions from carbon-tin bonds represented a profound methodological advance. The scope of the groups that could be attached to the anionic center expanded dramatically as a result. [6]

(11)

23Scope2.png

Synthetic applications

The products of the [2,3]-Wittig rearrangement of bis(allylic) ethers are 1,5-dien-3-ols. These substrates may undergo the oxy-Cope rearrangement upon deprotonation, affording δ,ε-unsaturated carbonyls. This tandem sigmatropic strategy has been employed in the synthesis of some natural products, including brevicomine and oxocrinol. [16]

(12)

23Synth.png

Variations

One variation of the 2,3-Wittig rearrangement is the Wittig-Still rearrangement [17]

Experimental conditions and procedure

Typical conditions

Rearrangements must be carried out at temperatures below −60 °C to avoid competitive [1,2]-rearrangement. Typically, simple treatment of the substrate with n-butyllithium is sufficient to cause rearrangement. Reactions involving butyllithium should be carried out under nitrogen or argon with strict exclusion of water.

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.

<span class="mw-page-title-main">Sharpless epoxidation</span> Chemical reaction

The Sharpless epoxidation reaction is an enantioselective chemical reaction to prepare 2,3-epoxyalcohols from primary and secondary allylic alcohols. The oxidizing agent is tert-butyl hydroperoxide. The method relies on a catalyst formed from titanium tetra(isopropoxide) and diethyl tartrate.

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.

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

<span class="mw-page-title-main">Ene reaction</span> Reaction in organic chemistry

In organic chemistry, the ene reaction is a chemical reaction between an alkene with an allylic hydrogen and a compound containing a multiple bond, in order to form a new σ-bond with migration of the ene double bond and 1,5 hydrogen shift. The product is a substituted alkene with the double bond shifted to the allylic position.

The Simmons–Smith reaction is an organic cheletropic reaction involving an organozinc carbenoid that reacts with an alkene to form a cyclopropane. It is named after Howard Ensign Simmons, Jr. and Ronald D. Smith. It uses a methylene free radical intermediate that is delivered to both carbons of the alkene simultaneously, therefore the configuration of the double bond is preserved in the product and the reaction is stereospecific.

<span class="mw-page-title-main">Claisen rearrangement</span> Chemical reaction

The Claisen rearrangement is a powerful carbon–carbon bond-forming chemical reaction discovered by Rainer Ludwig Claisen. The heating of an allyl vinyl ether will initiate a [3,3]-sigmatropic rearrangement to give a γ,δ-unsaturated carbonyl, driven by exergonically favored carbonyl CO bond formation.

The Carroll rearrangement is a rearrangement reaction in organic chemistry and involves the transformation of a β-keto allyl ester into a α-allyl-β-ketocarboxylic acid. This organic reaction is accompanied by decarboxylation and the final product is a γ,δ-allylketone. The Carroll rearrangement is an adaptation of the Claisen rearrangement and effectively a decarboxylative allylation.

The Overman rearrangement is a chemical reaction that can be described as a Claisen rearrangement of allylic alcohols to give allylic trichloroacetamides through an imidate intermediate. The Overman rearrangement was discovered in 1974 by Larry Overman.

<span class="mw-page-title-main">Chiral auxiliary</span> Stereogenic group placed on a molecule to encourage stereoselectivity in reactions

In stereochemistry, a chiral auxiliary is a stereogenic group or unit that is temporarily incorporated into an organic compound in order to control the stereochemical outcome of the synthesis. The chirality present in the auxiliary can bias the stereoselectivity of one or more subsequent reactions. The auxiliary can then be typically recovered for future use.

<span class="mw-page-title-main">2,3-sigmatropic rearrangement</span> Class of chemical reaction

2,3-Sigmatropic rearrangements are a type of sigmatropic rearrangements and can be classified into two types. Rearrangements of allylic sulfoxides, amine oxides, selenoxides are neutral. Rearrangements of carbanions of allyl ethers are anionic. The general scheme for this kind of rearrangement is:

The Nazarov cyclization reaction is a chemical reaction used in organic chemistry for the synthesis of cyclopentenones. The reaction is typically divided into classical and modern variants, depending on the reagents and substrates employed. It was originally discovered by Ivan Nikolaevich Nazarov (1906–1957) in 1941 while studying the rearrangements of allyl vinyl ketones.

The Kornblum–DeLaMare rearrangement is a rearrangement reaction in organic chemistry in which a primary or secondary organic peroxide is converted to the corresponding ketone and alcohol under acid or base catalysis. The reaction is relevant as a tool in organic synthesis and is a key step in the biosynthesis of prostaglandins.

The Rubottom oxidation is a useful, high-yielding chemical reaction between silyl enol ethers and peroxyacids to give the corresponding α-hydroxy carbonyl product. The mechanism of the reaction was proposed in its original disclosure by A.G. Brook with further evidence later supplied by George M. Rubottom. After a Prilezhaev-type oxidation of the silyl enol ether with the peroxyacid to form the siloxy oxirane intermediate, acid-catalyzed ring-opening yields an oxocarbenium ion. This intermediate then participates in a 1,4-silyl migration to give an α-siloxy carbonyl derivative that can be readily converted to the α-hydroxy carbonyl compound in the presence of acid, base, or a fluoride source.

<span class="mw-page-title-main">Allylic strain</span> Type of strain energy in organic chemistry

Allylic strain in organic chemistry is a type of strain energy resulting from the interaction between a substituent on one end of an olefin with an allylic substituent on the other end. If the substituents are large enough in size, they can sterically interfere with each other such that one conformer is greatly favored over the other. Allylic strain was first recognized in the literature in 1965 by Johnson and Malhotra. The authors were investigating cyclohexane conformations including endocyclic and exocylic double bonds when they noticed certain conformations were disfavored due to the geometry constraints caused by the double bond. Organic chemists capitalize on the rigidity resulting from allylic strain for use in asymmetric reactions.

Organostannane addition reactions comprise the nucleophilic addition of an allyl-, allenyl-, or propargylstannane to an aldehyde, imine, or, in rare cases, a ketone. The reaction is widely used for carbonyl allylation.

Electrophilic amination is a chemical process involving the formation of a carbon–nitrogen bond through the reaction of a nucleophilic carbanion with an electrophilic source of nitrogen.

The Tsuji–Trost reaction is a palladium-catalysed substitution reaction involving a substrate that contains a leaving group in an allylic position. The palladium catalyst first coordinates with the allyl group and then undergoes oxidative addition, forming the π-allyl complex. This allyl complex can then be attacked by a nucleophile, resulting in the substituted product.

William Clark Still is an American organic chemist. As a distinguished professor at Columbia University, Clark Still made significant contributions to the field of organic chemistry, particularly in the areas of natural product synthesis, reaction development, conformational analysis, macrocyclic stereocontrol, and computational chemistry. Still and coworkers also developed the purification technique known as flash column chromatography which is widely used for the purification of organic compounds.

In organic chemistry, the oxy-Cope rearrangement is a chemical reaction. It involves reorganization of the skeleton of certain unsaturated alcohols. It is a variation of the Cope rearrangement in which 1,5-dien-3-ols are converted to unsaturated carbonyl compounds by a mechanism typical for such a [3,3]-sigmatropic rearrangement.

References

  1. Nakai, T.; Mikami, K. (2004). "The [2,3]-Wittig Rearrangement". Organic Reactions. doi:10.1002/0471264180.or046.02. ISBN   0471264180.
  2. Baldwin, J. E.; Patrick, J. E. (1971). "Stereochemistry of [2,3]-sigmatropic reactions. Wittig rearrangement". J. Am. Chem. Soc. 93 (14): 3556. doi:10.1021/ja00743a060.
  3. Nakai, T.; Mikami, K.; Taya, S.; Fujita, Y. (1981). "[2,3]-Wittig rearrangement of unsymmetrical bis-allylic ethers. Facile method for regio- and stereoselective synthesis of 1,5-dien-3-ols". J. Am. Chem. Soc. 103 (21): 6492. doi:10.1021/ja00411a038.
  4. Schollkopf, U.; Fellenberger, K.; Rizk, M. (1970). "1.2-Wanderungen zum Atom mit freiem Elektronenpaar, VIII.ortho-Isomerisation bei anionisierten Äthern und Wanderungsmechanismus eines Propargyl-Restes bei der Wittig-Umlagerung". Justus Liebigs Ann. Chem. 734: 106–115. doi:10.1002/jlac.19707340111.
  5. Mikami, K.; Kimura, Y.; Kishi, N.; Nakai, T. (1983). "Acyclic diastereoselection of the [2,3]-Wittig sigmatropic rearrangement of a series of isomeric crotyl ethers. A conceptual model for the transition-state geometry". J. Org. Chem. 48 (2): 279. doi:10.1021/jo00150a033.
  6. 1 2 Still, W. C.; Mitra, A. (1978). "A highly stereoselective synthesis of Z-trisubstituted olefins via [2,3]-sigmatropic rearrangement. Preference for a pseudoaxially substituted transition state". J. Am. Chem. Soc. 100 (6): 1927. doi:10.1021/ja00474a049.
  7. Sayo, N.; Azuma, K.; Mikami, K.; Nakai, T. (1984). "Acyclic stereocontrol via asymmetric [2,3]-Wittig rearrangement with high enantio- and erythro-selectivity and its use in the chiral synthesis of insect pheromones". Tetrahedron Lett. 25 (5): 565. doi:10.1016/S0040-4039(00)99939-8.
  8. Mikami, K.; Fujimoto, K.; Kasuga, T.; Nakai, T. (1984). "Asymmetric [2,3]Wittig sigmatropic rearrangement involving a chiral azaenolate as the migrating terminus. A simple synthesis of (+)-verrucarinolactone". Tetrahedron Lett. 25 (52): 6011. doi:10.1016/S0040-4039(01)81746-9.
  9. Priepke, H.; Bruckner, R.; Harms, K. (1990). "Asymmetric Induction in the Wittig-Still Rearrangement of Ethers Containing an Allylic Stereocenter – Diastereocontrol by Allylic Nitrogen". Chem. Ber. 123 (3): 555. doi:10.1002/cber.19901230323.
  10. Paquette, L. A.; Wright, J.; Drtina, G. J.; Roberts, R. A. (1987). "Enantiospecific total synthesis of natural (−)-retigeranic acid a and two (−)-retigeranic acid B candidates". J. Org. Chem. 52 (13): 2960. doi:10.1021/jo00389a070.
  11. Marshall, J. A.; Lebreton, J. (1988). "Enantioselective synthesis of macrocyclic propargylic alcohols by [2,3] Wittig ring contraction. Synthesis of (+)-aristolactone and cembranoid precursors". J. Am. Chem. Soc. 110 (9): 2925. doi:10.1021/ja00217a039.
  12. Castedo, L.; Granja, J. R.; Mourino, A. (1985). "(2,3)-Wittig sigmatropic rearrangements in steroid synthesis. New stereocontrolled approach to steroidal side chains at C-20". Tetrahedron Lett. 26 (40): 4959. doi:10.1016/S0040-4039(00)94997-9.
  13. Mikami, K.; Kishi, N.; Nakai, T. (1989). "Silicon-directed regiocontrol in Witting rearrangements of bis-allyl ethers and allyl propargyl ethers". Chem. Lett. 18 (9): 1683–1686. doi:10.1246/cl.1989.1683.
  14. Takahashi, O.; Mikami, K.; Nakai, T. (1987). "Asymmetric [2,3]-Wittig rearrangement involving a chiral ester enolate terminus. A chiral synthesis of erythro-.ALPHA.-hydroxy-.BETA.-alkyl carboxylic acid derivatives". Chem. Lett. 16 (1): 69–72. doi: 10.1246/cl.1987.69 .
  15. Uemura, M.; Nishimura, H.; Minami, T.; Hayashi, Y. (1991). "(.eta.6-Arene)chromium complexes in organic synthesis: Synthesis of (.+-.)-dihydroxyserrulatic acid". J. Am. Chem. Soc. 113 (14): 5402. doi:10.1021/ja00014a036.
  16. Mikami, K.; Nakai, T. (1982). "Applications of the tandem (2,3)-Wittig-oxy-Cope rearrangement to syntheses of exo-brevicomin and oxocrinol. The scope and limitation of the sigmatropic sequences as a synthetic method for δ,ε-unsaturated ketones". Chem. Lett. 11 (9): 1349–1352. doi:10.1246/cl.1982.1349.
  17. Rycek Lukas, Hudlicky Tomas (2017). "Applications of the Wittig-Still Rearrangement in Organic Synthesis". Angewandte Chemie International Edition. 56 (22): 6022–6066. doi:10.1002/anie.201611329. PMID   28211171.