Shapiro reaction

Last updated
Shapiro reaction
Named after Robert H. Shapiro
Reaction type Coupling reaction
Identifiers
Organic Chemistry Portal shapiro-reaction
RSC ontology ID RXNO:0000125

The Shapiro reaction or tosylhydrazone decomposition is an organic reaction in which a ketone or aldehyde is converted to an alkene through an intermediate hydrazone in the presence of 2 equivalents of organolithium reagent. [1] [2] [3] The reaction was discovered by Robert H. Shapiro in 1967. [4] The Shapiro reaction was used in the Nicolaou Taxol total synthesis. [5] This reaction is very similar to the Bamford–Stevens reaction, which also involves the basic decomposition of tosyl hydrazones.

Contents

The Shapiro reaction Shapiro reaction.svg
The Shapiro reaction

Reaction mechanism

In a prelude to the actual Shapiro reaction, a ketone or an aldehyde (1) is reacted with p-toluenesulfonylhydrazide [6] (2) to form a p-toluenesulfonylhydrazone (or tosylhydrazone) which is a hydrazone (3). Two equivalents of strong base such as n-butyllithium abstract the proton from the hydrazone (4) followed by the less acidic proton α to the hydrazone carbon (5), forming a carbanion. The carbanion then undergoes an elimination reaction producing a carbon–carbon double bond and ejecting the tosyl anion, forming a diazonium anion (6). This diazonium anion is then lost as molecular nitrogen resulting in a vinyllithium species (7), which can then be reacted with various electrophiles, including simple neutralization with water or an acid (8).

Shapiro Reaction.png

Scope

The position of the alkene in the product is controlled by the site of deprotonation by the organolithium base. In general, the kinetically favored, less substituted site of differentially substituted tosylhydrazones is deprotonated selectively, leading to the less substituted vinyllithium intermediate. Although many secondary reactions exist for the vinyllithium functional group, in the Shapiro reaction in particular water is added, resulting in protonation to the alkene. [7] Other reactions of vinyllithium compounds include alkylation reactions with for instance alkyl halides. [8]

Shapiro reactions starting from camphor (1) through the intermediate hydrazone (2) to the vinyllithium (3). Addition of water (c) results in 2-bornene (4) and addition of an alkyl bromide (d) gives 5 ShapiroReaction borneneSynthesis.svg
Shapiro reactions starting from camphor (1) through the intermediate hydrazone (2) to the vinyllithium (3). Addition of water (c) results in 2-bornene (4) and addition of an alkyl bromide (d) gives 5

Importantly, the Shapiro reaction cannot be used to synthesize 1-lithioalkenes (and the resulting functionalized derivatives), as sulfonylhydrazones derived from aldehydes undergo exclusive addition of the organolithium base to the carbon of the C–N double bond. [9]

Catalytic Shapiro reaction

Traditional Shapiro reactions require stoichiometric (sometimes excess) amounts of base to generate the alkenyllithium reagents. To combat this problem, Yamamoto and coworkers developed an efficient stereoselective and regioselective route to alkenes using a combination of ketone phenylaziridinylhydrazones as arenesulfonylhydrazone equivalents with a catalytic amount of lithium amides. The required phenylaziridinylhydrazone was prepared from the condensation of undecan-6-one with 1-amino-2-phenylaziridine. Treatment of the phenylaziridinylhydrazone with 0.3 equivalents of LDA in ether resulted in the alkene shown below with a cis:trans ratio of 99.4:0.6. The ratio was determined by capillary GLC analysis after conversion to the corresponding epoxides with mCPBA. The catalyst loading can be reduced to 0.05 equivalents in the case of a 30mmol scale reaction.

The high stereoselectivity is obtained by the preferential abstraction of the α-methylene hydrogen syn to the phenylaziridine, and is also accounted for by the internal chelation of the lithiated intermediated. [10]

The Shapiro reaction with N-aziridinylhydrazones produces the alkene product, as well as stryrene and gaseous nitrogen as byproducts. The cycle of the catalytic Shapiro reaction is also shown. Catalytic Shapiro reaction.jpg
The Shapiro reaction with N-aziridinylhydrazones produces the alkene product, as well as stryrene and gaseous nitrogen as byproducts. The cycle of the catalytic Shapiro reaction is also shown.

A one pot in situ combined Shapiro-Suzuki reaction

The Shapiro reaction can also be combined with the Suzuki reaction to produce a variety of olefin products. Keay and coworkers have developed methodology that combines these reactions in a one pot process that does not require the isolation of the boronic acid, a setback of the traditional Suzuki coupling. This reaction has a wide scope, tolerating a slew of trisylhydrazones and aryl halides, as well as several solvents and Pd sources. [11]

The Shapiro and Suzuki reactions are combined to yield a variety of alkene products. Shapiro-Suzuki reaction.jpg
The Shapiro and Suzuki reactions are combined to yield a variety of alkene products.

An application of the Shapiro reaction in total synthesis

The Shapiro reaction has been used to generate olefins towards to complex natural products. K. Mori and coworkers wanted to determine the absolute configuration of the phytocassane group of a class of natural products called phytoalexins. This was accomplished by preparing the naturally occurring (–)-phytocassane D from (R)-Wieland-Miescher ketone. On the way to (–)-phytocassane D, a tricyclic ketone was subjected to Shapiro reaction conditions to yield the cyclic alkene product. [12]

Use of the Shapiro reaction in the synthesis of (-)-phytocassane D Application in total synthesis.jpg
Use of the Shapiro reaction in the synthesis of (–)-phytocassane D

See also

Related Research Articles

<span class="mw-page-title-main">Alkene</span> Hydrocarbon compound containing one or more C=C bonds

In organic chemistry, an alkene is a hydrocarbon containing a carbon–carbon double bond. The double bond may be internal or in the terminal position. Terminal alkenes are also known as α-olefins.

<span class="mw-page-title-main">Carboxylic acid</span> Organic compound containing a –C(=O)OH group

In organic chemistry, a carboxylic acid is an organic acid that contains a carboxyl group attached to an R-group. The general formula of a carboxylic acid is R−COOH or R−CO2H, with R referring to the alkyl, alkenyl, aryl, or other group. Carboxylic acids occur widely. Important examples include the amino acids and fatty acids. Deprotonation of a carboxylic acid gives a carboxylate anion.

<span class="mw-page-title-main">Hydrazone</span> Organic compounds - Hydrazones

Hydrazones are a class of organic compounds with the structure R1R2C=N−NH2. They are related to ketones and aldehydes by the replacement of the oxygen =O with the =N−NH2 functional group. They are formed usually by the action of hydrazine on ketones or aldehydes.

<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.

The Wolff–Kishner reduction is a reaction used in organic chemistry to convert carbonyl functionalities into methylene groups. In the context of complex molecule synthesis, it is most frequently employed to remove a carbonyl group after it has served its synthetic purpose of activating an intermediate in a preceding step. As such, there is no obvious retron for this reaction. The reaction was reported by Nikolai Kischner in 1911 and Ludwig Wolff in 1912.

<span class="mw-page-title-main">Enolate</span> Organic anion formed by deprotonating a carbonyl (>C=O) compound

In organic chemistry, enolates are organic anions derived from the deprotonation of carbonyl compounds. Rarely isolated, they are widely used as reagents in the synthesis of organic compounds.

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. 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.

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

Organoboron chemistry or organoborane chemistry studies organoboron compounds, also called organoboranes. These chemical compounds combine boron and carbon; typically, they are organic derivatives of borane (BH3), as in the trialkyl boranes.

<span class="mw-page-title-main">Bamford–Stevens reaction</span> Synthesis of alkenes by base-catalysed decomposition of tosylhydrazones

The Bamford–Stevens reaction is a chemical reaction whereby treatment of tosylhydrazones with strong base gives alkenes. It is named for the British chemist William Randall Bamford and the Scottish chemist Thomas Stevens Stevens (1900–2000). The usage of aprotic solvents gives predominantly Z-alkenes, while protic solvent gives a mixture of E- and Z-alkenes. As an alkene-generating transformation, the Bamford–Stevens reaction has broad utility in synthetic methodology and complex molecule 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.

<span class="mw-page-title-main">Seyferth–Gilbert homologation</span> Chemical compound

The Seyferth–Gilbert homologation is a chemical reaction of an aryl ketone 1 with dimethyl (diazomethyl)phosphonate 2 and potassium tert-butoxide to give substituted alkynes 3. Dimethyl (diazomethyl)phosphonate 2 is often called the Seyferth–Gilbert reagent.

<span class="mw-page-title-main">Grignard reagent</span> Organometallic compounds used in organic synthesis

A Grignard reagent or Grignard compound is a chemical compound with the general formula R−Mg−X, where X is a halogen and R is an organic group, normally an alkyl or aryl. Two typical examples are methylmagnesium chloride Cl−Mg−CH3 and phenylmagnesium bromide (C6H5)−Mg−Br. They are a subclass of the organomagnesium compounds.

<span class="mw-page-title-main">Meerwein arylation</span> Organic reaction

The Meerwein arylation is an organic reaction involving the addition of an aryl diazonium salt (ArN2X) to an electron-poor alkene usually supported by a metal salt. The reaction product is an alkylated arene compound. The reaction is named after Hans Meerwein, one of its inventors who first published it in 1939.

<span class="mw-page-title-main">Nozaki–Hiyama–Kishi reaction</span> Coupling reaction used in organic synthesis

The Nozaki–Hiyama–Kishi reaction is a nickel/chromium coupling reaction forming an alcohol from the reaction of an aldehyde with an allyl or vinyl halide. In their original 1977 publication, Tamejiro Hiyama and Hitoshi Nozaki reported on a chromium(II) salt solution prepared by reduction of chromic chloride by lithium aluminium hydride to which was added benzaldehyde and allyl chloride:

Heteroatom-promoted lateral lithiation is the site-selective replacement of a benzylic hydrogen atom for lithium for the purpose of further functionalization. Heteroatom-containing substituents may direct metalation to the benzylic site closest to the heteroatom or increase the acidity of the ring carbons via an inductive effect.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

A tosylhydrazone in organic chemistry is a functional group with the general structure RR'C=N-NH-Ts where Ts is a tosyl group. Organic compounds having this functional group can be accessed by reaction of an aldehyde or ketone with tosylhydrazine.

<span class="mw-page-title-main">Enders SAMP/RAMP hydrazone-alkylation reaction</span>

The Enders SAMP/RAMP hydrazone alkylation reaction is an asymmetric carbon-carbon bond formation reaction facilitated by pyrrolidine chiral auxiliaries. It was pioneered by E. J. Corey and D. Enders in 1976, and was further developed by D. Enders and his group. This method is usually a three-step sequence. The first step is to form the hydrazone between (S)-1-amino-2-methoxymethylpyrrolidine (SAMP) or (R)-1-amino-2-methoxymethylpyrrolidine (RAMP) and a ketone or aldehyde. Afterwards, the hydrazone is deprotonated by lithium diisopropylamide (LDA) to form an azaenolate, which reacts with alkyl halides or other suitable electrophiles to give alkylated hydrazone species with the simultaneous generation of a new chiral center. Finally, the alkylated ketone or aldehyde can be regenerated by ozonolysis or hydrolysis.

<span class="mw-page-title-main">Vinyl iodide functional group</span>

In organic chemistry, a vinyl iodide functional group is an alkene with one or more iodide substituents. Vinyl iodides are versatile molecules that serve as important building blocks and precursors in organic synthesis. They are commonly used in carbon-carbon forming reactions in transition-metal catalyzed cross-coupling reactions, such as Stille reaction, Heck reaction, Sonogashira coupling, and Suzuki coupling. Synthesis of well-defined geometry or complexity vinyl iodide is important in stereoselective synthesis of natural products and drugs.

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. Shapiro, R. H.; Lipton, M. F.; Kolonko, K. J.; Buswell, R. L.; Capuano, L. A. (1975). "Tosylhydrazones and alkyllithium reagents: More on the regiospecificity of the reaction and the trapping of three intermediates". Tetrahedron Lett. 16 (22–23): 1811–1814. doi:10.1016/S0040-4039(00)75263-4.
  2. Shapiro, Robert H. (1976). "Alkenes from Tosylhydrazones". Org. React. 23 (3): 405–507. doi:10.1002/0471264180.or023.03. ISBN   978-0471264187.
  3. Adlington, Robert M.; Barret, Anthony G. M. (1983). "Recent applications of the Shapiro reaction". Acc. Chem. Res. 16 (2): 55–59. doi:10.1021/ar00086a004.
  4. Shapiro, Robert H.; Heath, Marsha J. (1967). "Tosylhydrazones. V. Reaction of Tosylhydrazones with Alkyllithium Reagents. A New Olefin Synthesis". J. Am. Chem. Soc. 89 (22): 5734–5735. doi:10.1021/ja00998a601.
  5. Nicolaou, Kyriacos C.; Sorensen, Erik J. (1996). Classics in Total Synthesis: Targets, Strategies, Methods. Wiley. ISBN   9783527292318.
  6. Friedman, Lester; Litle, Robert L.; Reichle, Walter R. (1960). "p-Toluenesulfonylhydrazide". Organic Syntheses . 40: 93.; Collective Volume, vol. 5, p. 1055
  7. Shapiro, R. H.; Duncan, J. H. (1971). "2-Bornene (1,7,7-Trimethylbicyclo[2.2.1]hept-2-ene)". Organic Syntheses . 51: 66. doi:10.15227/orgsyn.051.0066.; Collective Volume, vol. 6
  8. Chamberlin, A. Richard; Liotta, Ellen L.; Bond, F. Thomas (1983). "Generation and Reactions of Alkenyllithium Reagents: 2-Butylbornene". Organic Syntheses . 61: 141. doi:10.15227/orgsyn.061.0141.; Collective Volume, vol. 7, p. 77
  9. Chamberlin, A. Richard; Bloom, Steven H. (1990). "Lithioalkenes from Arenesulfonylhydrazones". Org. React. 39 (1): 1–83. doi:10.1002/0471264180.or039.01.
  10. Maruoka, Keiji; Oishi, Masataka; Yamamoto, Hisashi (1991). "The Catalytic Shapiro Reaction". J. Am. Chem. Soc. 118 (9): 2289–2290. doi:10.1021/ja951422p.
  11. Passafaro, Marco S.; Keay, Brian A. (1996). "A one pot in situ combined Shapiro-Suzuki reaction". Tetrahedron Lett. 37 (4): 429–432. doi:10.1016/0040-4039(95)02210-4.
  12. Yajima, Arata; Mori, Kenji (2000). "Synthesis and absolute configuration of (–)-phytocassane D, a diterpene phytoalexin isolated from the rice plant, Oryza sativa". Eur. J. Org. Chem. 2000 (24): 4079–4091. doi:10.1002/1099-0690(200012)2000:24<4079::AID-EJOC4079>3.0.CO;2-R.