Vinylcyclopropane rearrangement

Last updated

The vinylcyclopropane rearrangement or vinylcyclopropane-cyclopentene rearrangement is a ring expansion reaction, converting a vinyl-substituted cyclopropane ring into a cyclopentene ring. [1] [2] [3]

Contents

Cyclopropyl groups adjacent to vinyl groups can undergo ring expansion reactions. This reactivity can be exploited to generate unusual cyclic compounds, such as cyclobutenes, [4] or bicyclic species such as the cycloheptene shown below. [5]

CyclopropaneCycloaddition.png


Vinylcyclopropane Rearrangement VinylcyclopropaneRearrangement Header 1.png
Vinylcyclopropane Rearrangement

Experimental and computational investigations show that mechanistically, the vinylcyclopropane rearrangement can be thought of as either a diradical-mediated two-step and/or orbital-symmetry-controlled pericyclic process. The amount by which each of the two mechanisms is operative is highly dependent on the substrate.

Due to its ability to form cyclopentene rings the vinylcyclopropane rearrangement has served several times as a key reaction in complex natural product synthesis.

Origins and history

In 1959, a young research chemist with Humble Oil and Refining (Esso, now Exxon) named Norman P. Neureiter was instructed to find new uses for the excess butadiene produced from one of the refinery processes. Discussions about carbene chemistry with one of the company's most respectable consultants at the time, William von Eggers Doering, then a professor at Yale, led the young Ph.D. graduate from Northwestern University to follow a recent procedure combining both, carbenes and butadiene. [6] In particular the procedure described the reaction of 1,3-butadiene with carbenes generated from the action of base on chloroform or bromoform, which had been studied previously by Doering. [7] Neureiter then took the resulting 1,1-dichloro-2,2-dimethylcyclopropane and under pyrolysis conditions (above 400 °C) discovered a rearrangement to 4,4-dichlorocyclopentene which today is considered to be the first thermal vinylcyclopropane-cyclopentene rearrangement in history. [8]

Vinylcyclopropane Rearrangement of 1,1-dichloro-2,2-dimethylcyclopropane to 4,4-dichlorocyclopentene 1 History Neureiter.png
Vinylcyclopropane Rearrangement of 1,1-dichloro-2,2-dimethylcyclopropane to 4,4-dichlorocyclopentene

The corresponding all-carbon version of the reaction was independently reported by Emanuel Vogel [9] and Overberger & Borchert just one year after the Neureiter publication appeared. [10] [11] Doering, although interacting with Humble Oil and Refining - and therefore also with Neureiter - as a consultant, in a 1963 publication stated the following : "Credit for discovery that vinylcyclopropane rearranges to cyclopentene is due to Overberger and Borchert, and Vogel et al., who appear to have developed several examples of the rearrangement independently." [12] The development of further vinylcyclopropane rearrangement variants didn't take long as demonstrated by Atkinson & Rees in 1967, [13] Lwowski in 1968. [14] and Paladini & Chuche in 1971. [15]

Historical Overview Vinylcyclopropane Rearrangement 1 2 History Vogel.png
Historical Overview Vinylcyclopropane Rearrangement 1

The classical vinylcyclopropane rearrangement was discovered well after two of its heteroatom variants had already been reported. Although it is believed that the vinylcyclopropane rearrangement must have occurred during Nikolay Demyanov's preparation of vinylcyclopropane by Hofmann elimination at elevated temperatures in 1922, [16] the cyclopropylimine-pyrroline rearrangement by Cloke in 1929 [17] and Wilson's cyclopropylcarbaldehyde-2,3-dihydrofuran rearrangement in 1947 [18] are really the only examples of vinylcyclopropane-like rearrangements.

Historical Overview Vinylcyclopropane Rearrangement 2 3 History Cloke.png
Historical Overview Vinylcyclopropane Rearrangement 2

This last reaction type is also known as the Cloke–Wilson Rearrangement [19]

Mechanism

The mechanistic discussion on whether the vinylcyclopropane rearrangement proceeds through a diradical-mediated two-step or a fully concerted orbital-symmetry-controlled mechanism has been going on for more than half a century. Kinetic data together with the secondary kinetic isotope effects observed at the vinyl terminus of the vinylcyclopropane suggest a concerted mechanism whereas product distribution indicates a stepwise-diradical mechanism. [20] In the 1960s, shortly after the rearrangement was discovered, it was established that the activation energy for the vinylcyclopropane rearrangement is around 50 kcal/mol. [21] The kinetic data obtained for this rearrangement were consistent with a concerted mechanism where cleavage of the cyclopropyl carbon-carbon bond was rate-limiting. Albeit a concerted mechanism seemed likely it was shortly recognized that the activation energy to break the carbon-carbon bond in unsubstituted cyclopropane was with 63 kcal/mol [22] exactly 13 kcal/mol higher in energy than the parent activation energy, a difference remarkably similar to the resonance energy of the allyl radical. [23] Immediately people started to appreciate the possibility for a diradical intermediate arising from homolytic cleavage of the weak C1-C2-cyclopropane bond under thermal conditions.

diradical mechanism Mechanism 01.png
diradical mechanism

The discussion on whether the vinylcyclopropane rearrangement proceeds via a fully concerted or a two-step, non-concerted mechanism was given further careful consideration when Woodward and Hoffmann used the vinylcyclopropane rearrangement to exemplify [1,3]-sigmatropic concerted alkyl shifts in 1969. [24] They hypothesized that if a concerted mechanism was operative the consequences of orbital-symmetry controlled factors would only allow the formation of certain products. According to their analysis of a vinylcyclopropane substituted with three R groups the antarafacial [1,3]-shift of bond 1,2 to C-5, with retention at C-2, leading to the ar cyclopentene and the suprafacial [1,3]-shift of bond 1,2 to C-5, with inversion at C-2, leading to cyclopentene si are symmetry allowed whereas the suprafacial [1,3]-shift of bond 1,2 to C-5, with retention at C-2, leading to cyclopentene sr and the antarafacial [1,3]-shift of bond 1,2 to C-5, with inversion at C-2, leading to the ai cyclopentene are symmetry-forbidden. It is important to note that Woodward and Hoffmann based their analysis solely on the principles of the conservation of orbital symmetry theory without however making any mechanistic or stereochemical prediction.

Woodward-Hoffmann analysis Mechanism 02.png
Woodward-Hoffmann analysis

The attention directed towards the vinylcyclopropane rearrangement by Woodward and Hoffmann as a representative example for [1,3]-carbon shifts clearly enhanced the interest in this reaction. Furthermore, their analysis revealed potential experiments that would allow to distinguish between a concerted or stepwise mechanism. The stereochemical consequences of a concerted reaction pathway on the reaction outcome suggested an experiment where one would correlate the obtained reaction stereochemistry with the predicted reaction stereochemistry for a model substrate. Observing the formation of ai- and sr-cyclopentene products would support the notion that a stepwise, non-concerted mechanism is operative whereas their absence would point towards a fully concerted mechanism. As it turned out finding an appropriate substituted model substrate to study the stereochemical outcome of the vinylcyclopropane rearrangement was much more challenging than initially thought since side reaction such as the homodienyl [1,5]-hydrogen shifts and more so thermal stereomutations tend to scramble stereochemical distinctions much faster than rearrangements lead to the cyclopentene products.

Stereomutations Mechanism 03.png
Stereomutations

Even though deconvolution of the complex kinetic scenarios underlying these rearrangements was difficult there have been several studies reported where exact and explicit deconvolutions of kinetic and stereochemical raw data to account for the stereochemical contributions arising from competitive stereomutations was possible. [20] [25] [26] [27]

Thereby rate constants for all four stereochemically distinct pathways of the vinylcyclopropane rearrangement could be determined.

Stereochemical distribution of products Mechanism 04 05.png
Stereochemical distribution of products

The data clearly indicated that the mechanistic preferences of the rearrangements are system dependent. Whereas trans-vinylcyclopropanes tend to form more of the symmetry-allowed ar- and si-cyclopentenes supportive of a concerted mechanism, the cis-vinylcyclopropanes preferentially yield the symmetry-forbidden ai- and sr- products suggesting a more stepwise, diradical mechanism. The influence of substituent effects on the reaction stereochemistry also becomes apparent from the data. Substituents with increased radical stabilizing ability not only lower the rearrangements activation energy but also reclosure of the initially formed diradical species becomes slower relative to the rate of cyclopentene formation resulting in an overall more concerted mechanism with less stereomutation (e.g. entry 6 & 7). In all cases though all the four products were formed indicating that both orbital-symmetry controlled pericyclic, as well as diradical-mediated two-step mechanisms are operative either way. The data is consistent with the formation of biradical species on a relatively flat potential energy surface allowing for restricted conformational flexibility before the products are formed. The amount of conformational flexibility and therefore conformational evolution accessible to the diradical species before forming product depends on the constitution of the potential energy surface. This notion is also supported by computational work. [28] One transition state with a high diradicaloid character was found. Following the potential energy surface of the lowest energy path of the reaction it was found that a very shallow regime allows the diradical species to undergo conformational changes and stereoisomerization reactions with minor energetic consequences. Furthermore, it was shown that substituents can favor stereoselective pathways by destabilizing species that allow stereochemical scrambling.

Methodology development

Arguably the biggest drawback of the vinylcyclopropane rearrangement as a synthetic method is its intrinsically high activation barrier resulting in very high reaction temperatures (500-600 °C). Not only do these high temperatures allow side reactions with similar activation energies, such as homodienyl-[1,5]-hydrogen shifts, to occur but also do they significantly limit the functional groups tolerated in the substrates. It was well recognized by the chemical community that in order for this reaction to become a useful synthetic method, hopefully applicable in complex natural product settings at some point, some reaction development had to be done. It was found that the reaction temperature could be lowered drastically when the cyclopropane ring contained a dithiane group. Even though the dithiane-substituted vinylcyclopropane substrates required two synthetic steps starting from the corresponding 1,3-dienes the method proved itself successful for the synthesis of a variety of substituted cyclopentenes. The immediate rearrangement products could be easily converted to the corresponding cyclopentenones. [29]

Corey 02 MechanisticDevelopment Corey.png
Corey

Methoxy-substituted vinylcyclopropanes show significantly faster reaction rates allowing the rearrangement to take place at 220 °C. [30]

03 MechanisticDevelopment Simpson.png

It was found that siloxyvinylcyclopropanes [31] as well as the analogous sulfinylvinylcyclopropanes [32] could be used as substrates to build annulated cyclopentene structures. Albeit these reactions still required reaction temperatures above 300 °C they were able to make useful products arising from the annulation of cyclopentene to a present ring system.

01 MechanisticDevelopment Trost.png

Vinylcyclopropane rearrangements can also be mediated photochemically. [33] [34] In a particularly intriguing example he was able to show that vinylcyclopropanes embedded within a cyclooctane core can be converted to the corresponding [5-5]-fused ring systems.

04 MechanisticDevelopment Paquette.png

Vinylcyclopropane rearrangements are amenable to transition metal catalysts. [Dirhodium acetate]] catalyzes the rearrangements from room temperature to 80 °C.y [35] [36]

05 MechanisticDevelopment Brown&Hudlicky.png

Analogous to the rate acceleration observed in the anionic-oxy-Cope rearrangement Danheiser reported a very similar effect for vinylcyclopropane substrates bearing [alkoxy] substituents. [37]

06 MechanisticDevelopment Danheiser.png

Another intriguing result was reported by Larsen in 1988. [38] He was able to promote vinylcyclopropane rearrangements with substrates such as the one shown in the reaction below at temperatures as low as -78 °C. The substrates were generated in situ upon ringcontracting thiocarbonyl Diels-Alder adducts under basic conditions. This methodology allowed the formation of numerous highly functionalized cyclopentenes in a stereoselective manner.

07 MechanisticDevelopment Larsen.png

The methodology is allows the formation of various [5-5]- as well as [5-6]-carbon scaffolds. [39]

08 MechanisticDevelopment Hudlicky.png

Use in total synthesis

Five-membered carbon rings are ubiquitous structural motifs in natural products. In contrast to the larger, fully "consonant" cyclohexane scaffold cyclopentanes and their derivatives are "dissonant" according to the Lapworth-Evans model of alternating polarities. The dissonance in polarity clearly limits the ways by which cyclopentanes can be disconnected which becomes evident in the decreased number of general methods available for making five-membered rings versus the corresponding six-membered rings. Especially the fact that there is no Diels-Alder-equivalent for the synthesis of five-membered rings has been bothering synthetic chemists for many decades. Consequentially, after the vinylcyclopropane rearrangement was discovered around 1960 it didn't take long for the synthetic community to realize the potential inherent to form cyclopentenes by means of the vinylcyclopropane rearrangement. As the vinylcyclopropane rearrangement progressed as a methodology and the reaction conditions improved during the 1970s, first total syntheses making use of the vinylcyclopropane rearrangement started to appear around 1980. Key figures to apply this reaction in total synthesis were Barry M. Trost, Elias J. Corey, Thomas Hudlicky, Leo A. Paquette,

Trost's synthesis of aphidicolin (1979)

A synthesis of Aphidicolin use methodology around the vinylcyclopropane rearrangement developed in their own laboratory . [40] A key step converts a late stage siloxyvinylcyclopropane into a cyclopentene that contained the [6-6-5]-fused carbon skeleton found within the natural product. The rearranged product into the natural product by further manipulations.

Aphidicolin Trost 1 Aphidicolin Trost.png
Aphidicolin Trost

Piers' synthesis of zizaene (1979)

The synthesis of zizaene is another early example for the application of a vinylcyclopropane rearrangement as a key disconnection. [41]

Piers Zizaene 2 Piers Zizaene.png
Piers Zizaene

Hudlicky's synthesis of hirstuene (1980) and isocomene (1984)

The methodology has also been applied to the synthesis hirsutene [42] and isocomene [43]

Triquinanes Hudlicky 3 Triquinanes Hudlicky.png
Triquinanes Hudlicky

Synthesis of alpha-Vetispirene

Cinylcyclopropane rearrangement has been used to build the spirocyclic natural product alpha-Vetispirene in 1982. [44]

Paquette vetispirene 4 Paquette vetispirene.png
Paquette vetispirene

Synthesis of Antheridiogen-An

Antheridiogen-An was prepared using a Lewis-acid mediated late-stage vinylcyclopropane rearrangement. [45]

Corey Antheridiogen 5 Corey Antheridiogen.png
Corey Antheridiogen

Synthesis of biotin

Copper-catalyzed heteroatom-vinylcyclopropane rearrangement was used to form the tetrahydrothiophene core of biotin and the thiophene unit of Plavix respectively. [46]

Njardarson Biotin&Plavix 6 Njardarson Biotin&Plavix.png
Njardarson Biotin&Plavix

Synthesis of salviasperanol

An acid-mediated vinylcyclopropane rearrangement was used to synthesize the natural product salviasperanol. [47]

Majetich Salviasperanol 7 Majetich Salviasperanol.png
Majetich Salviasperanol

See also

Related Research Articles

In organic chemistry, an electrocyclic reaction is a type of pericyclic rearrangement where the net result is one pi bond being converted into one sigma bond or vice versa. These reactions are usually categorized by the following criteria:

A sigmatropic reaction in organic chemistry is a pericyclic reaction wherein the net result is one σ-bond is changed to another σ-bond in an uncatalyzed intramolecular reaction. The name sigmatropic is the result of a compounding of the long-established sigma designation from single carbon–carbon bonds and the Greek word tropos, meaning turn. In this type of rearrangement reaction, a substituent moves from one part of a π-bonded system to another part in an intramolecular reaction with simultaneous rearrangement of the π system. True sigmatropic reactions are usually uncatalyzed, although Lewis acid catalysis is possible. Sigmatropic reactions often have transition-metal catalysts that form intermediates in analogous reactions. The most well-known of the sigmatropic rearrangements are the [3,3] Cope rearrangement, Claisen rearrangement, Carroll rearrangement, and the Fischer indole synthesis.

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.

The Cope rearrangement is an extensively studied organic reaction involving the [3,3]-sigmatropic rearrangement of 1,5-dienes. It was developed by Arthur C. Cope and Elizabeth Hardy. For example, 3-methyl-hexa-1,5-diene heated to 300 °C yields hepta-1,5-diene.

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.

<span class="mw-page-title-main">Bergman cyclization</span>

The Masamune-Bergman cyclization or Masamune-Bergman reaction or Masamune-Bergman cycloaromatization is an organic reaction and more specifically a rearrangement reaction taking place when an enediyne is heated in presence of a suitable hydrogen donor. It is the most famous and well-studied member of the general class of cycloaromatization reactions. It is named for Japanese-American chemist Satoru Masamune and American chemist Robert G. Bergman. The reaction product is a derivative of benzene.

Bullvalene is a hydrocarbon with the chemical formula C10H10. The molecule has a cage-like structure formed by the fusion of one cyclopropane and three cyclohepta-1,4-diene rings. Bullvalene is unusual as an organic molecule due to the C−C and C=C bonds forming and breaking rapidly on the NMR timescale; this property makes it a fluxional molecule.

<span class="mw-page-title-main">Wolff rearrangement</span>

The Wolff rearrangement is a reaction in organic chemistry in which an α-diazocarbonyl compound is converted into a ketene by loss of dinitrogen with accompanying 1,2-rearrangement. The Wolff rearrangement yields a ketene as an intermediate product, which can undergo nucleophilic attack with weakly acidic nucleophiles such as water, alcohols, and amines, to generate carboxylic acid derivatives or undergo [2+2] cycloaddition reactions to form four-membered rings. The mechanism of the Wolff rearrangement has been the subject of debate since its first use. No single mechanism sufficiently describes the reaction, and there are often competing concerted and carbene-mediated pathways; for simplicity, only the textbook, concerted mechanism is shown below. The reaction was discovered by Ludwig Wolff in 1902. The Wolff rearrangement has great synthetic utility due to the accessibility of α-diazocarbonyl compounds, variety of reactions from the ketene intermediate, and stereochemical retention of the migrating group. However, the Wolff rearrangement has limitations due to the highly reactive nature of α-diazocarbonyl compounds, which can undergo a variety of competing reactions.

<span class="mw-page-title-main">Cyclopropanation</span> Chemical process which generates cyclopropane rings

In organic chemistry, cyclopropanation refers to any chemical process which generates cyclopropane rings. It is an important process in modern chemistry as many useful compounds bear this motif; for example pyrethroid insecticides and a number of quinolone antibiotics. However, the high ring strain present in cyclopropanes makes them challenging to produce and generally requires the use of highly reactive species, such as carbenes, ylids and carbanions. Many of the reactions proceed in a cheletropic manner.

The divinylcyclopropane-cycloheptadiene rearrangement is an organic chemical transformation that involves the isomerization of a 1,2-divinylcyclopropane into a cycloheptadiene or -triene. It is conceptually related to the Cope rearrangement, but has the advantage of a strong thermodynamic driving force due to the release of ring strain. This thermodynamic power is recently being considered as an alternative energy source.

A (4+3) cycloaddition is a cycloaddition between a four-atom π-system and a three-atom π-system to form a seven-membered ring. Allyl or oxyallyl cations (propenylium-2-olate) are commonly used three-atom π-systems, while a diene plays the role of the four-atom π-system. It represents one of the relatively few synthetic methods available to form seven-membered rings stereoselectively in high yield.

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.:

Rearrangements, especially those that can participate in cascade reactions, such as the aza-Cope rearrangements, are of high practical as well as conceptual importance in organic chemistry, due to their ability to quickly build structural complexity out of simple starting materials. The aza-Cope rearrangements are examples of heteroatom versions of the Cope rearrangement, which is a [3,3]-sigmatropic rearrangement that shifts single and double bonds between two allylic components. In accordance with the Woodward-Hoffman rules, thermal aza-Cope rearrangements proceed suprafacially. Aza-Cope rearrangements are generally classified by the position of the nitrogen in the molecule :

<span class="mw-page-title-main">Thermal rearrangement of aromatic hydrocarbons</span>

Thermal rearrangements of aromatic hydrocarbons are considered to be unimolecular reactions that directly involve the atoms of an aromatic ring structure and require no other reagent than heat. These reactions can be categorized in two major types: one that involves a complete and permanent skeletal reorganization (isomerization), and one in which the atoms are scrambled but no net change in the aromatic ring occurs (automerization). The general reaction schemes of the two types are illustrated in Figure 1.

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.

<span class="mw-page-title-main">Activation of cyclopropanes by transition metals</span>

In organometallic chemistry, the activation of cyclopropanes by transition metals is a research theme with implications for organic synthesis and homogeneous catalysis. Being highly strained, cyclopropanes are prone to oxidative addition to transition metal complexes. The resulting metallacycles are susceptible to a variety of reactions. These reactions are rare examples of C-C bond activation. The rarity of C-C activation processes has been attributed to Steric effects that protect C-C bonds. Furthermore, the directionality of C-C bonds as compared to C-H bonds makes orbital interaction with transition metals less favorable. Thermodynamically, C-C bond activation is more favored than C-H bond activation as the strength of a typical C-C bond is around 90 kcal per mole while the strength of a typical unactivated C-H bond is around 104 kcal per mole.

<span class="mw-page-title-main">Rick L. Danheiser</span> American organic chemist

Rick L. Danheiser is an American organic chemist and is the Arthur C. Cope Professor of Chemistry at the Massachusetts Institute of Technology and chair of the MIT faculty. His research involves the invention of new methods for the synthesis of complex organic compounds. Danheiser is known for the Danheiser annulation and Danheiser benzannulation reactions.

Vinylcyclopropane [5+2] cycloaddition is a type of cycloaddition between a vinylcyclopropane (VCP) and an olefin or alkyne to form a seven-membered ring.

In organic chemistry, the Conia-ene reaction is an intramolecular cyclization reaction between an enolizable carbonyl such as an ester or ketone and an alkyne or alkene, giving a cyclic product with a new carbon-carbon bond. As initially reported by J. M. Conia and P. Le Perchec, the Conia-ene reaction is a heteroatom analog of the ene reaction that uses an enol as the ene component. Like other pericyclic reactions, the original Conia-ene reaction required high temperatures to proceed, limiting its wider application. However, subsequent improvements, particularly in metal catalysis, have led to significant expansion of reaction scope. Consequently, various forms of the Conia-ene reaction have been employed in the synthesis of complex molecules and natural products.

References

  1. Mil'vitskaya, E M; Tarakanova, A V; Plate, Alfred F (1976). "Thermal Rearrangements of Vinylcyclopropanes". Russ Chem. Rev. 45: 469–478. doi:10.1070/RC1976v045n05ABEH002675.
  2. Goldschmidt, Z.; Crammer, B. (1988). "Vinylcyclopropane rearrangements". Chem. Soc. Rev. 17: 229–267. doi:10.1039/CS9881700229.
  3. Hudlicky, Tomas; Reed, Josephine W. (2010). "From Discovery to Application: 50 Years of the Vinylcyclopropane-Cyclopentene Rearrangement and Its Impact on the Synthesis of Natural Products". Angewandte Chemie International Edition. 49 (29): 4864–76. doi:10.1002/anie.200906001. PMID   20586104.
  4. Fürstner, Alois; Aïssa, Christophe (2006). "PtCl-Catalyzed Rearrangement of Methylenecyclopropanes". Journal of the American Chemical Society. 128 (19): 6306–6307. doi:10.1021/ja061392y. hdl: 11858/00-001M-0000-0025-AE20-3 . PMID   16683781.
  5. Wender, Paul A.; Haustedt, Lars O.; Lim, Jaehong; Love, Jennifer A.; Williams, Travis J.; Yoon, Joo-Yong (May 2006). "Asymmetric Catalysis of the [5 + 2] Cycloaddition Reaction of Vinylcyclopropanes and π-Systems". Journal of the American Chemical Society. 128 (19): 6302–6303. doi:10.1021/ja058590u. PMID   16683779. S2CID   197039161.
  6. Woodworth, Robert C.; Skell, Philip S. (1957). "Reactions of bivalent carbon species. Addition of dihalocarbenes to 1,3-butadiene". J. Am. Chem. Soc. 79 (10): 2542. doi:10.1021/ja01567a048.
  7. Doering, W. von E.; Hoffman, A. Kentaro (1954). "The Addition of Dichlorocarbene to Olefins". J. Am. Chem. Soc. 76 (23): 6162. doi:10.1021/ja01652a087.
  8. Neureiter, Norman (1959). "Pyrolysis of 1,l-Dichloro-2-vinylcyclopropane. Synthesis of 2-Chlorocyclopentadiene". J. Org. Chem. 24 (12): 2044. doi:10.1021/jo01094a621.
  9. Vogel, Emanuel (1960). "Kleine Kohlenstoff-Ringe". Angewandte Chemie. 72: 4–26. doi:10.1002/ange.19600720103.
  10. Overberger, C. G.; Borchert, A. E. (1960). "Novel thermal rearrangements accompanying acetate pyrolysis in small ring systems". J. Am. Chem. Soc. 82 (4): 1007. doi:10.1021/ja01489a069.
  11. Overberger, C. G.; Borchert, A. E. (1960). "Ionic Polymerization. XVI. Reactions of 1-Cyclopropylethanol-Vinylcyclopropane". J. Am. Chem. Soc. 82 (18): 4896. doi:10.1021/ja01503a036.
  12. Doering, W. von E.; Lambert, J. B. (1963). "Thermal Reorganization of a- and b-Thujene: A degenerate rearrangement of the vinylcyclopropane type". Tetrahedron. 19 (12): 1989. doi:10.1016/0040-4020(63)85013-9.
  13. Atkinson, R. S.; Rees, C. W. (1967). "A vinylaziridine to pyrroline rearrangement". Chemical Communications (23): 1232a. doi:10.1039/C1967001232a.
  14. Lwowski, Walter; Rice, Susan N.; Lwowski, Walter (1968). "Singlet and Triplet Nitrenes. 111. The Addition of Carbethoxynitrene to 1,3-Dienes". J. Org. Chem. 33 (22): 481. doi:10.1021/jo01266a001.
  15. Paladini, J; Chuche, X. X. (1971). "Rearrangement thermique d'epoxydes vinyliques". Tetrahedron Letters. 12 (46): 4383. doi:10.1016/S0040-4039(01)97447-7.
  16. Demjanow, N. J.; Dojarenko, Marie (1922). "Über Vinylcyclopropan, einige Derivate des Methyl-cyclopropyl-carbinols und die Isomerisation des Cyclopropan-Ringes". Ber. Dtsch. Chem. Ges. B. 55 (8): 2718. doi:10.1002/cber.19220550846.
  17. Cloke, J. B.; Borchert, A. E. (1929). "The formation of pyrrolines from gamma-chloropropyl and cyclopropyl ketimines". J. Am. Chem. Soc. 51 (18): 1174. doi:10.1021/ja01379a028.
  18. Wilson, C. L.; Borchert, A. E. (1947). "Reactions of Furan Compounds. VII. Thermal Interconversionof 2,3-Dihydrofuran and Cyclopropane Aldehyde". J. Am. Chem. Soc. 69 (18): 3002. doi:10.1021/ja01204a020.
  19. Organic Syntheses Based on Name Reactions: A Practical Guide to 750 Transformations Alfred Hassner, Irishi Namboothiri Elsevier, 2012
  20. 1 2 Baldwin, John E. (2003). "Thermal Rearrangements of Vinylcyclopropanes to Cyclopentenes". Chemical Reviews. 103 (4): 1197–212. doi:10.1021/cr010020z. PMID   12683781.
  21. Flowers, M. C.; Rabinovitch, B. S. (1960). "The Thermal Unimolecular Isomerisation of Vinylcyclo- propane to Cyclopentene". J. Chem. Soc. 82 (23): 3547. doi:10.1021/ja01508a008.
  22. Schlag, E. W.; Rabinovitch, B. S. (1960). "Kinetics of the Thermal Unimolecular Isomerization Reactions of Cyclopropane-d2". J. Am. Chem. Soc. 82 (23): 5996. doi:10.1021/ja01508a008.
  23. Egger, K. W.; Golden, David M.; Benson, Sidney W. (1964). "Iodine-Catalyzed Isomerization of Olefins. 11. The Resonance Energy of the Allyl Radical and the Kinetics of the Positional Isomerization of 1-Butene". J. Am. Chem. Soc. 86 (24): 5420. doi:10.1021/ja01078a011.
  24. Woodward, R. B.; Hoffmann, R. (1969). "The Conservation of Orbital Symmetry". Angew. Chem. Int. Ed. 8 (11): 781. doi:10.1002/anie.196907811.
  25. Gajewski, Joseph J.; Squicciarini, Michael P. (1989). "Evidence for concert in the vinylcyclopropane rearrangement. A reinvestigation of the pyrolysis of trans-1-methyl-2-(1-tert-butylethenyl)cyclopropane". Journal of the American Chemical Society. 111 (17): 6717. doi:10.1021/ja00199a035.
  26. Gajewski, Joseph J.; Olson, Leif P. (1991). "Evidence for a dominant suprafacial-inversion pathway in the thermal unimolecular vinylcyclopropane to cyclopentene 1,3-sigmatropic shift". Journal of the American Chemical Society. 113 (19): 7432. doi:10.1021/ja00019a056.
  27. Gajewski, Joseph J.; Olson, Leif P.; Willcott, M. Robert (1996). "Evidence for Concert in the Thermal Unimolecular Vinylcyclopropane to Cyclopentene Sigmatropic 1,3-Shift". Journal of the American Chemical Society. 118 (2): 299. doi:10.1021/ja951578p.
  28. Houk, K. N.; Nendel, Maja; Wiest, Olaf; Storer, Joey W. (1997). "The Vinylcyclopropane−Cyclopentene Rearrangement: A Prototype Thermal Rearrangement Involving Competing Diradical Concerted and Stepwise Mechanisms". Journal of the American Chemical Society. 119 (43): 10545. doi:10.1021/ja971315q.
  29. Corey, E. J.; Walinsky, S. W. (1972). "Reaction of 1,3-dithienium fluoroborate with 1,3-dienes. Synthesis of .DELTA.3-cyclopenten-1-ones". Journal of the American Chemical Society. 94 (25): 8932. doi:10.1021/ja00780a063.
  30. Simpson, John M.; Richey, Herman G. (1973). "The effects of methoxyl and phenyl substituents on the thermal rearrangements of vinylcyclopropane". Tetrahedron Letters. 14 (27): 2545. doi:10.1016/S0040-4039(01)96201-X.
  31. Trost, Barry M.; Bogdanowicz, Mitchell J. (1973). "New synthetic reactions. IX. Facile synthesis of oxaspiropentanes, versatile synthetic intermediates". Journal of the American Chemical Society. 95 (16): 5311. doi:10.1021/ja00797a036.
  32. Trost, Barry M.; Keeley, Donald E. (1976). "New synthetic methods. A stereocontrolled approach to cyclopentane annelation". Journal of the American Chemical Society. 98: 248–250. doi:10.1021/ja00417a048.
  33. Paquette, Leo A.; Meehan, George V.; Henzel, Richard P.; Eizember, Richard F. (1973). "Photochemistry of conjugated cis-bicyclo[5.1.0]octenones, cis- and trans-bicyclo[5.2.0]non-2-en-4-ones, and their methylene analogs". The Journal of Organic Chemistry. 38 (19): 3250. doi:10.1021/jo00959a004.
  34. Paquette, Leo A.; Henzel, Richard P.; Eizember, Richard F. (1973). "Thermochemical behavior of conjugated cis-bicyclo[5.1.0]octenones, cis- and trans-bicyclo[5.2.0]non-2-en-4-ones, and their methylene analogs". The Journal of Organic Chemistry. 38 (19): 3257. doi:10.1021/jo00959a005.
  35. Hudlicky, Tomas; Koszyk, Francis F.; Kutchan, Toni M.; Sheth, Jagdish P. (1980). "Cyclopentene annulation via intramolecular addition of diazoketones to 1,3-dienes. Applications to the synthesis of cyclopentanoid terpenes". The Journal of Organic Chemistry. 45 (25): 5020. doi:10.1021/jo01313a003.
  36. Brown, Vanessa; Brown, John M.; Conneely, John A.; Golding, Bernard T.; Williamson, David H. (1975). "Synthesis and thermolysis of rhodium and iridium complexes of endo-6-vinylbicyclo[3.1.0]hex-2-ene. A metal-promoted vinylcyclopropane to cyclopentene rearrangement". Journal of the Chemical Society, Perkin Transactions 2 (1): 4. doi:10.1039/P29750000004.
  37. Danheiser, Rick L.; Martinez-Davila, Carlos; Morin, John M. (1980). "Synthesis of 3-cyclopentenols by alkoxy-accelerated vinylcyclopropane rearrangement". The Journal of Organic Chemistry. 45 (7): 1340. doi:10.1021/jo01295a045.
  38. Larsen, Scott D. (1988). "A stereoselective synthesis of functionalized cyclopentenes via base-induced ring contraction of thiocarbonyl Diels-Alder adducts". Journal of the American Chemical Society. 110 (17): 5932–5934. doi:10.1021/ja00225a072.
  39. Hudlicky, Tomas; Heard, Nina E.; Fleming, Alison (1990). "4-Siloxy-.alpha.-bromocrotonate: A new reagent for [2+3] annulation leading to oxygenated cyclopentenes at low temperatures". The Journal of Organic Chemistry. 55 (9): 2570. doi:10.1021/jo00296a004.
  40. Trost, B. M.; Nishimura, Yoshio; Yamamoto, Kagetoshi (1979). "A Total Synthesis of Aphidicolin". J. Am. Chem. Soc. 101 (5): 1328. doi:10.1021/ja00499a071.
  41. Piers, E. (1979). "Five-membered Ring Annelation via Thermal Rearrangement of a-Cyclopropyl-ab-unsaturated Ketones: a New Total Synthesis of (&)-Zizaene". J. Chem. Soc. Chem. Commun. (24): 1138. doi:10.1039/C39790001138.
  42. Hudlicky, T.; Kutchan, Toni M.; Wilson, Stephen R.; Mao, David T. (1980). "Total Synthesis of (rac)-Hirsutene". J. Am. Chem. Soc. 102 (20): 6351. doi:10.1021/ja00540a036.
  43. Hudlicky, T.; Kavka, Misha; Higgs, Leslie A.; Hudlickyl, Tomas (1984). "Stereocontrolled Total Synthesis of Isocomene Sesquiterpenes". Tetrahedron Lett. 25 (23): 2447. doi:10.1016/S0040-4039(01)81201-6.
  44. Paquette, L. A. (1982). "A Short Synthesis of (rac)-alpha-Vetispirene". Tetrahedron Lett. 23: 3227. doi:10.1016/s0040-4039(00)87576-0.
  45. Corey, E. J.; Myers, Andrew G. (1985). "Total Synthesis of (rac)-Antheridium-Inducing Factor (AAn,2) of the Fern Anemia pbylfitidis. Clarification of Stereochemistry". J. Am. Chem. Soc. 107 (19): 5574. doi:10.1021/ja00305a067.
  46. Njardarson, J. T.; Araki, H; Batory, LA; McInnis, CE; Njardarson, JT (2007). "Highly Selective Copper-Catalyzed Ring Expansion of Vinyl Thiiranes: Application to Synthesis of Biotin and the Heterocyclic Core of Plavix". J. Am. Chem. Soc. 129 (10): 2768–9. doi:10.1021/ja069059h. PMID   17302422.
  47. Majetich, G.; Zou, G; Grove, J (2008). "Total Synthesis of (−)-Salviasperanol". Org. Lett. 10 (1): 85–7. doi:10.1021/ol701743c. PMID   18052176.