Johnson–Corey–Chaykovsky reaction

Last updated

Contents

Johnson-Corey–Chaykovsky reaction
Named afterA. William Johnson
Elias James Corey
Michael Chaykovsky
Reaction type Ring forming reaction
Identifiers
Organic Chemistry Portal corey-chaykovsky-reaction

The Johnson–Corey–Chaykovsky reaction (sometimes referred to as the Corey–Chaykovsky reaction or CCR) 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.

Johnson-Corey-Chaykovsky Reaction JCCRintro.png
Johnson–Corey–Chaykovsky Reaction

The reaction is most often employed for epoxidation via methylene transfer, and to this end has been used in several notable total syntheses (See Synthesis of epoxides below). Additionally detailed below are the history, mechanism, scope, and enantioselective variants of the reaction. Several reviews have been published. [1] [2] [3] [4] [5] [6]

History

The original publication by Johnson concerned the reaction of 9-dimethylsulfonium fluorenylide with substituted benzaldehyde derivatives. The attempted Wittig-like reaction failed and a benzalfluorene oxide was obtained instead, noting that "reaction between the sulfur ylid and benzaldehydes did not afford benzalfluorenes as had the phosphorus and arsenic ylids." [7]

The first example of the Johnson-Corey-Chaykovsky reaction Johnson-Corey-Chaykovsky Reaction Example.png
The first example of the Johnson–Corey–Chaykovsky reaction

The subsequent development of (dimethyloxosulfaniumyl)methanide, (CH3)2SOCH2 and (dimethylsulfaniumyl)methanide, (CH3)2SCH2 (known as Corey–Chaykovsky reagents) by Corey and Chaykovsky as efficient methylene-transfer reagents established the reaction as a part of the organic canon. [8]

Corey-Chaykovsky Reagent CCReagent.png
Corey–Chaykovsky Reagent

Mechanism

The reaction mechanism for the Johnson–Corey–Chaykovsky reaction consists of nucleophilic addition of the ylide to the carbonyl or imine group. A negative charge is transferred to the heteroatom and because the sulfonium cation is a good leaving group it gets expelled forming the ring. In the related Wittig reaction, the formation of the much stronger phosphorus-oxygen double bond prevents oxirane formation and instead, olefination takes place through a 4-membered cyclic intermediate. [4]

Mechanism of the Johnson-Corey-Chaykovsky reaction CCR Mechanism V.1.svg
Mechanism of the Johnson–Corey–Chaykovsky reaction

The trans diastereoselectivity observed results from the reversibility of the initial addition, allowing equilibration to the favored anti betaine over the syn betaine. Initial addition of the ylide results in a betaine with adjacent charges; density functional theory calculations have shown that the rate-limiting step is rotation of the central bond into the conformer necessary for backside attack on the sulfonium. [1]

Selectivity in the Johnson-Corey-Chaykovsky reaction CCR Mechanism (Newman) V2.png
Selectivity in the Johnson–Corey–Chaykovsky reaction

The degree of reversibility in the initial step (and therefore the diastereoselectivity) depends on four factors, with greater reversibility corresponding to higher selectivity: [1]

  1. Stability of the substrate with higher stability leading to greater reversibility by favoring the starting material over the betaine.
  2. Stability of the ylide with higher stability similarly leading to greater reversibility.
  3. Steric hindrance in the betaine with greater hindrance leading to greater reversibility by disfavoring formation of the intermediate and slowing the rate-limiting rotation of the central bond.
  4. Solvation of charges in the betaine by counterions such as lithium with greater solvation allowing more facile rotation in the betaine intermediate, lowering the amount of reversibility.

Scope

The application of the Johnson–Corey–Chaykovsky reaction in organic synthesis is diverse. The reaction has come to encompass reactions of many types of sulfur ylides with electrophiles well beyond the original publications. It has seen use in a number of high-profile total syntheses, as detailed below, and is generally recognized as a powerful transformative tool in the organic repertoire.

Types of ylides

General form of ylide reagent used CCReagentGeneral.png
General form of ylide reagent used

Many types of ylides can be prepared with various functional groups both on the anionic carbon center and on the sulfur. The substitution pattern can influence the ease of preparation for the reagents (typically from the sulfonium halide, e.g. trimethylsulfonium iodide) and overall reaction rate in various ways. The general format for the reagent is shown on the right. [1]

Use of a sulfoxonium allows more facile preparation of the reagent using weaker bases as compared to sulfonium ylides. (The difference being that a sulfoxonium contains a doubly bonded oxygen whereas the sulfonium does not.) The former react slower due to their increased stability. In addition, the dialkylsulfoxide by-products of sulfoxonium reagents are greatly preferred to the significantly more toxic, volatile, and odorous dialkylsulfide by-products from sulfonium reagents. [1]

The vast majority of reagents are monosubstituted at the ylide carbon (either R1 or R2 as hydrogen). Disubstituted reagents are much rarer but have been described: [1]

  1. If the ylide carbon is substituted with an electron-withdrawing group (EWG), the reagent is referred to as a stabilized ylide. These, similarly to sulfoxonium reagents, react much slower and are typically easier to prepare. These are limited in their usefulness as the reaction can become prohibitively sluggish: examples involving amides are widespread, with many fewer involving esters and virtually no examples involving other EWG's. For these, the related Darzens reaction is typically more appropriate.
  2. If the ylide carbon is substituted with an aryl or allyl group, the reagent is referred to as a semi-stabilized ylide. These have been developed extensively, second only to the classical methylene reagents (R1=R2=H). The substitution pattern on aryl reagents can heavily influence the selectivity of the reaction as per the criteria above.
  3. If the ylide carbon is substituted with an alkyl group the reagent is referred to as an unstabilized ylide. The size of the alkyl groups are the major factors in selectivity with these reagents.

The R-groups on the sulfur, though typically methyls, have been used to synthesize reagents that can perform enantioselective variants of the reaction (See Variations below). The size of the groups can also influence diastereoselectivity in alicyclic substrates. [1]

Synthesis of epoxides

Reactions of sulfur ylides with ketones and aldehydes to form epoxides are by far the most common application of the Johnson–Corey–Chaykovsky reaction. Examples involving complex substrates and 'exotic' ylides have been reported, as shown below. [9] [10]

Example 1 of epoxidation with CCR CCREpox1.png
Example 1 of epoxidation with CCR
Example 1 of epoxidation with CCR CCREpox2.png
Example 1 of epoxidation with CCR

The reaction has been used in a number of notable total syntheses including the Danishefsky Taxol total synthesis, which produces the chemotherapeutic drug taxol, and the Kuehne Strychnine total synthesis which produces the pesticide strychnine. [11] [12]

Taxol synthesis CCR step TaxolCCR.png
Taxol synthesis CCR step
Strychnine synthesis CCR step StrychnineCCR.png
Strychnine synthesis CCR step

Synthesis of aziridines

The synthesis of aziridines from imines is another important application of the Johnson–Corey–Chaykovsky reaction and provides an alternative to amine transfer from oxaziridines. Though less widely applied, the reaction has a similar substrate scope and functional group tolerance to the carbonyl equivalent. The examples shown below are representative; in the latter, an aziridine forms in situ and is opened via nucleophilic attack to form the corresponding amine. [3] [9]

Aziridination with the Johnson-Corey-Chaykovsky reaction AziridinationCCR.png
Aziridination with the Johnson–Corey–Chaykovsky reaction

Synthesis of cyclopropanes

For addition of sulfur ylides to enones, higher 1,4-selectivity is typically obtained with sulfoxonium reagents than with sulfonium reagents. Many electron-withdrawing groups have been shown compatible with the reaction including ketones, esters, and amides (the example below involves a Weinreb amide). With further conjugated systems 1,6-addition tends to predominate over 1,4-addition. [3] [9]

Cyclopropanation with the Johnson-Corey-Chaykovsky reaction CCRcycloprop.png
Cyclopropanation with the Johnson–Corey–Chaykovsky reaction

Other reactions

In addition to the reactions originally reported by Johnson, Corey, and Chaykovsky, sulfur ylides have been used for a number of related homologation reactions that tend to be grouped under the same name.

Oxetane and Azitidine synthesis with the Johnson-Corey-Chaykovsky reaction CCRoxetane.png
Oxetane and Azitidine synthesis with the Johnson–Corey–Chaykovsky reaction
[4+1] cycloaddition with Corey-Chaykovsky reagent CCR41.png
[4+1] cycloaddition with Corey–Chaykovsky reagent
Living polymerization with Johnson-Corey-Chaykovsky Reaction CCRPolymer.png
Living polymerization with Johnson–Corey–Chaykovsky Reaction

Enantioselective variations

The development of an enantioselective (i.e. yielding an enantiomeric excess, which is labelled as "ee") variant of the Johnson–Corey–Chaykovsky reaction remains an active area of academic research. The use of chiral sulfides in a stoichiometric fashion has proved more successful than the corresponding catalytic variants, but the substrate scope is still limited in all cases. The catalytic variants have been developed almost exclusively for enantioselective purposes; typical organosulfide reagents are not prohibitively expensive and the racemic reactions can be carried out with equimolar amounts of ylide without raising costs significantly. Chiral sulfides, on the other hand, are more costly to prepare, spurring the advancement of catalytic enantioselective methods. [2]

Stoichiometric reagents

The most successful reagents employed in a stoichiometric fashion are shown below. The first is a bicyclic oxathiane that has been employed in the synthesis of the β-adrenergic compound dichloroisoproterenol (DCI) but is limited by the availability of only one enantiomer of the reagent. The synthesis of the axial diastereomer is rationalized via the 1,3-anomeric effect which reduces the nucleophilicity of the equatorial lone pair. The conformation of the ylide is limited by transannular strain and approach of the aldehyde is limited to one face of the ylide by steric interactions with the methyl substituents. [5] [2]

chiral oxathiane reagent for the Johnson-Corey-Chaykovsky reaction CCRoxathiane.png
chiral oxathiane reagent for the Johnson–Corey–Chaykovsky reaction

The other major reagent is a camphor-derived reagent developed by Varinder Aggarwal of the University of Bristol. Both enantiomers are easily synthesized, although the yields are lower than for the oxathiane reagent. The ylide conformation is determined by interaction with the bridgehead hydrogens and approach of the aldehyde is blocked by the camphor moiety. The reaction employs a phosphazene base to promote formation of the ylide. [5] [2]

chiral camphor-derived reagent for the Johnson-Corey-Chaykovsky reaction CCRcamphor.png
chiral camphor-derived reagent for the Johnson–Corey–Chaykovsky reaction

Catalytic reagents

Catalytic reagents have been less successful, with most variations suffering from poor yield, poor enantioselectivity, or both. There are also issues with substrate scope, most having limitations with methylene transfer and aliphatic aldehydes. The trouble stems from the need for a nucleophilic sulfide that efficiently generates the ylide which can also act as a good leaving group to form the epoxide. Since the factors underlying these desiderata are at odds, tuning of the catalyst properties has proven difficult. Shown below are several of the most successful catalysts along with the yields and enantiomeric excess for their use in synthesis of (E)-stilbene oxide. [5] [2]

chiral catalysts for the Johnson-Corey-Chaykovsky reaction CCRchiralcat.png
chiral catalysts for the Johnson–Corey–Chaykovsky reaction

Aggarwal has developed an alternative method employing the same sulfide as above and a novel alkylation involving a rhodium carbenoid formed in situ . The method too has limited substrate scope, failing for any electrophiles possessing basic substituents due to competitive consumption of the carbenoid. [2]

chiral catalyst with carbenoid alkylation for the Johnson-Corey-Chaykovsky reaction CCRcatcarbene.png
chiral catalyst with carbenoid alkylation for the Johnson–Corey–Chaykovsky reaction

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">Epoxide</span> Organic compounds with a carbon-carbon-oxygen ring

In organic chemistry, an epoxide is a cyclic ether, where the ether forms a three-atom ring: two atoms of carbon and one atom of oxygen. This triangular structure has substantial ring strain, making epoxides highly reactive, more so than other ethers. They are produced on a large scale for many applications. In general, low molecular weight epoxides are colourless and nonpolar, and often volatile.

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.

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.

In chemistry, stereoselectivity is the property of a chemical reaction in which a single reactant forms an unequal mixture of stereoisomers during a non-stereospecific creation of a new stereocenter or during a non-stereospecific transformation of a pre-existing one. The selectivity arises from differences in steric and electronic effects in the mechanistic pathways leading to the different products. Stereoselectivity can vary in degree but it can never be total since the activation energy difference between the two pathways is finite: both products are at least possible and merely differ in amount. However, in favorable cases, the minor stereoisomer may not be detectable by the analytic methods used.

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">Henry reaction</span>

The Henry reaction is a classic carbon–carbon bond formation reaction in organic chemistry. Discovered in 1895 by the Belgian chemist Louis Henry (1834–1913), it is the combination of a nitroalkane and an aldehyde or ketone in the presence of a base to form β-nitro alcohols. This type of reaction is also referred to as a nitroaldol reaction. It is nearly analogous to the aldol reaction that had been discovered 23 years prior that couples two carbonyl compounds to form β-hydroxy carbonyl compounds known as "aldols". The Henry reaction is a useful technique in the area of organic chemistry due to the synthetic utility of its corresponding products, as they can be easily converted to other useful synthetic intermediates. These conversions include subsequent dehydration to yield nitroalkenes, oxidation of the secondary alcohol to yield α-nitro ketones, or reduction of the nitro group to yield β-amino alcohols.

A pinacol coupling reaction is an organic reaction in which a carbon–carbon bond is formed between the carbonyl groups of an aldehyde or a ketone in presence of an electron donor in a free radical process. The reaction product is a vicinal diol. The reaction is named after pinacol, which is the product of this reaction when done with acetone as reagent. The reaction is usually a homocoupling but intramolecular cross-coupling reactions are also possible. Pinacol was discovered by Wilhelm Rudolph Fittig in 1859.

<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.

<span class="mw-page-title-main">Shi epoxidation</span>

The Shi epoxidation is a chemical reaction described as the asymmetric epoxidation of alkenes with oxone and a fructose-derived catalyst (1). This reaction is thought to proceed via a dioxirane intermediate, generated from the catalyst ketone by oxone. The addition of the sulfate group by the oxone facilitates the formation of the dioxirane by acting as a good leaving group during ring closure. It is notable for its use of a non-metal catalyst and represents an early example of organocatalysis.

<span class="mw-page-title-main">Darzens reaction</span>

The Darzens reaction is the chemical reaction of a ketone or aldehyde with an α-haloester in the presence of a base to form an α,β-epoxy ester, also called a "glycidic ester". This reaction was discovered by the organic chemist Auguste Georges Darzens in 1904.

The Stetter reaction is a reaction used in organic chemistry to form carbon-carbon bonds through a 1,4-addition reaction utilizing a nucleophilic catalyst. While the related 1,2-addition reaction, the benzoin condensation, was known since the 1830s, the Stetter reaction was not reported until 1973 by Dr. Hermann Stetter. The reaction provides synthetically useful 1,4-dicarbonyl compounds and related derivatives from aldehydes and Michael acceptors. Unlike 1,3-dicarbonyls, which are easily accessed through the Claisen condensation, or 1,5-dicarbonyls, which are commonly made using a Michael reaction, 1,4-dicarbonyls are challenging substrates to synthesize, yet are valuable starting materials for several organic transformations, including the Paal–Knorr synthesis of furans and pyrroles. Traditionally utilized catalysts for the Stetter reaction are thiazolium salts and cyanide anion, but more recent work toward the asymmetric Stetter reaction has found triazolium salts to be effective. The Stetter reaction is an example of umpolung chemistry, as the inherent polarity of the aldehyde is reversed by the addition of the catalyst to the aldehyde, rendering the carbon center nucleophilic rather than electrophilic.

<span class="mw-page-title-main">Aziridines</span> Functional group made of a carbon-carbon-nitrogen heterocycle

In organic chemistry, aziridines are organic compounds containing the aziridine functional group, a three-membered heterocycle with one amine and two methylene bridges. The parent compound is aziridine, with molecular formula C2H4NH. Several drugs feature aziridine rings, including mitomycin C, porfiromycin, and azinomycin B (carzinophilin).

Chiral Lewis acids (CLAs) are a type of Lewis acid catalyst. These acids affect the chirality of the substrate as they react with it. In such reactions, synthesis favors the formation of a specific enantiomer or diastereomer. The method is an enantioselective asymmetric synthesis reaction. Since they affect chirality, they produce optically active products from optically inactive or mixed starting materials. This type of preferential formation of one enantiomer or diastereomer over the other is formally known as asymmetric induction. In this kind of Lewis acid, the electron-accepting atom is typically a metal, such as indium, zinc, lithium, aluminium, titanium, or boron. The chiral-altering ligands employed for synthesizing these acids often have multiple Lewis basic sites that allow the formation of a ring structure involving the metal atom.

The Parikh–Doering oxidation is an oxidation reaction that transforms primary and secondary alcohols into aldehydes and ketones, respectively. The procedure uses dimethyl sulfoxide (DMSO) as the oxidant and the solvent, activated by the sulfur trioxide pyridine complex (SO3•C5H5N) in the presence of triethylamine or diisopropylethylamine as base. Dichloromethane is frequently used as a cosolvent for the reaction.

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

Jacobsen's catalyst is the common name for N,N'-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexane­diaminomanganese(III) chloride, a coordination compound of manganese and a salen-type ligand. It is used as an asymmetric catalyst in the Jacobsen epoxidation, which is renowned for its ability to enantioselectively transform prochiral alkenes into epoxides. Before its development, catalysts for the asymmetric epoxidation of alkenes required the substrate to have a directing functional group, such as an alcohol as seen in the Sharpless epoxidation. This compound has two enantiomers, which give the appropriate epoxide product from the alkene starting material.

Nucleophilic epoxidation is the formation of epoxides from electron-deficient double bonds through the action of nucleophilic oxidants. Nucleophilic epoxidation methods represent a viable alternative to electrophilic methods, many of which do not epoxidize electron-poor double bonds efficiently.

The Juliá–Colonna epoxidation is an asymmetric poly-leucine catalyzed nucleophilic epoxidation of electron deficient olefins in a triphasic system. The reaction was reported by Sebastian Juliá at the Chemical Institute of Sarriá in 1980, with further elaboration by both Juliá and Stefano Colonna.

In organic chemistry, the Baylis–Hillman, Morita–Baylis–Hillman, or MBH reaction is a carbon-carbon bond-forming reaction between an activated alkene and a carbon electrophile in the presence of a nucleophilic catalyst, such as a tertiary amine or phosphine. The product is densely functionalized, joining the alkene at the α-position to a reduced form of the electrophile.

In organic chemistry, the Jocic reaction, also called the Jocic–Reeve reaction is a name reaction that generates α-substituted carboxylic acids from trichloromethylcarbinols and corresponding nucleophiles in the presence of sodium hydroxide. The reaction involves nucleophilic displacement of the hydroxyl group in a 1,1,1-trichloro-2-hydroxyalkyl structure with concomitant conversion of the trichloromethyl portion to a carboxylic acid or other acyl group.

References

  1. 1 2 3 4 5 6 7 Aggarwal, V. K.; Richardson, J. (2003). "The complexity of catalysis: origins of enantio- and diastereocontrol in sulfur ylide mediated epoxidation reactions". Chemical Communications (21): 2644–2651. doi:10.1039/b304625g. PMID   14649793.
  2. 1 2 3 4 5 6 Aggarwal, V. K.; Winn, C. L. (2004). "Catalytic, Asymmetric Sulfur Ylide-Mediated Epoxidation of Carbonyl Compounds: Scope, Selectivity, and Applications in Synthesis". Accounts of Chemical Research . 37 (8): 611–620. doi:10.1021/ar030045f. PMID   15311960.
  3. 1 2 3 Gololobov, Y. G.; Nesmeyanov, A. N.; lysenko, V. P.; Boldeskul, I. E. (1987). "Twenty-five years of dimethylsulfoxonium ethylide (corey's reagent)". Tetrahedron . 43 (12): 2609–2651. doi:10.1016/s0040-4020(01)86869-1.
  4. 1 2 Li, A.-H.; Dai, L.-X.; Aggarwal, V. K. (1997). "Asymmetric Ylide Reactions: Epoxidation, Cyclopropanation, Aziridination, Olefination, and Rearrangement". Chemical Reviews . 97 (6): 2341–2372. doi:10.1021/cr960411r. PMID   11848902.
  5. 1 2 3 4 Aggarwal, Varinder K.; Ford, J. Gair; Fonguerna, Sílvia; Adams, Harry; Jones, Ray V. H.; Fieldhouse, Robin (1998-08-08). "Catalytic Asymmetric Epoxidation of Aldehydes. Optimization, Mechanism, and Discovery of Stereoelectronic Control Involving a Combination of Anomeric and Cieplak Effects in Sulfur Ylide Epoxidations with Chiral 1,3-Oxathianes". Journal of the American Chemical Society . 120 (33): 8328–8339. doi:10.1021/ja9812150.
  6. McGarrigle, E. M.; Myers, E. L.; Illa, O.; Shaw, M. A.; Riches, S. L.; Aggarwal, V. K. (2007). "Chalcogenides as Organocatalysts". Chemical Reviews . 107 (12): 5841–5883. doi:10.1021/cr068402y. PMID   18072810.
  7. Johnson, A.W.; LaCount, R.B. (1961). "The Chemistry of Ylids. VI. Dimethylsulfonium Fluorenylide—A Synthesis of Epoxides". J. Am. Chem. Soc. 83 (2): 417–423. doi:10.1021/ja01463a040.
  8. Corey, E. J.; Chaykovsky, M. (1965). "Dimethyloxosulfonium Methylide ((CH3)2SOCH2) and Dimethylsulfonium Methylide ((CH3)2SCH2). Formation and Application to Organic Synthesis". J. Am. Chem. Soc. 87 (6): 1353–1364. doi:10.1021/ja01084a034.
  9. 1 2 3 4 5 Li, Jack Jie (2005). Named Reactions in Heterocyclic Chemistry. Hoboken, New Jersey: John Wiley & Sons, Inc. pp. 2–14. ISBN   9780471704140.
  10. Mundy, Bradford, P.; Ellerd, Michael D.; Favaloro, Frank G. Jr. (2005). Name Reactions and Reagents in Organic Chemistry (2 ed.). Hoboken, New Jersey: John Wiley & Sons, Inc. pp. 174–175, 743. ISBN   9780471739869.{{cite book}}: CS1 maint: multiple names: authors list (link)
  11. Danishefsky, S. J.; Masters, J. J.; Young, W. B.; Link, J. T.; Snyder, L. B.; Magee, T. V.; Jung, D. K.; Isaacs, R. C. A.; Bornmann, W. G.; Alaimo, C. A.; Coburn, C. A.; Di Grandi, M. J. (1996). "Total Synthesis of Baccatin III and Taxol". Journal of the American Chemical Society . 118 (12): 2843–2859. doi:10.1021/ja952692a.
  12. Kuehne, M. E.; Xu, F. (1993). "Total synthesis of strychnan and aspidospermatan alkaloids. 3. The total synthesis of (.+-.)-strychnine". The Journal of Organic Chemistry . 58 (26): 7490–7497. doi:10.1021/jo00078a030.
  13. Luo, J.; Shea, K. J. (2010). "Polyhomologation. A Living C1 Polymerization". Accounts of Chemical Research . 43 (11): 1420–1433. doi:10.1021/ar100062a. PMID   20825177.