Conia-ene reaction

Last updated
Conia-ene reaction
Named afterJean-Marie Conia
Reaction type Cycloaddition
Identifiers
Organic Chemistry Portal conia-ene-reaction

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. [1] 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.

Contents

Conia-ene overview.png

History and mechanism

In the late 1960s, the laboratory of chemist Jean-Marie Conia investigated small carbocyclic molecules, specifically as products of ene-type reactions with carbonyls. [2] These efforts culminated in a 1975 review paper titled “The Thermal Cyclisation of Unsaturated Carbonyl Compounds.” [1]

In its original manifestation, the Conia-ene reaction comprised the intramolecular cyclization of ε,ζ-unsaturated ketones or aldehydes to functionalized cyclopentanes upon intense heating. [1] [3] The proposed mechanism invoked a six-membered, ene reaction-like transition state in which the enol tautomer reacts concertedly with the pendant alkene.

Original Conia-ene mechanism.png

The same conditions were found to give six- and nine-membered rings with the appropriate substrates, although with lower yields and diastereoselectivity. In the case of γ,δ- and δ,ε-unsaturated ketones, equilibrium favored the linear product over the cyclopropane or cyclobutane. Alkynyl ketones were also found to cyclize under thermal conditions, giving a mixture of the conjugated and skipped cyclic enones.

Original Conia-ene reaction scope.png

Two key drawbacks prevented wider implementation of the initial Conia-ene reaction. First, molecules with additional functional groups were often incompatible with the high temperatures required for conversion. Second, regio- and diastereoselectivity depended entirely on the substrate, offering little to no control over the orientation of the product.

Advancements

In the decades after the discovery of the Conia-ene reaction, several improvements allowed for milder reaction conditions and greater control of product stereo- and regiochemistry. For example, the carbonyl component, formerly a ketone or aldehyde, became a substituted β-ketoester or malonate ester. Such carbonyls enolize much more readily, yielding better access to the desired enol tautomer. [4] Additionally, the alkene component was replaced with an alkyne, which not only gave better cyclization in accordance with Baldwin’s rules, but also furnished a product containing an alkene that served as a useful handle for further transformations. [5] Finally, recent efforts have featured metal-mediated and metal-catalyzed Conia-ene reactions that can be rendered asymmetric using chiral ligands. [6]

Activation modes

These advancements have produced five main types of Conia-ene reactions characterized by the operative activation mode: namely, enolate, alkyne, or ene-yne activation, and one- or two-metal dual activation. Note that though the mechanisms of Conia-ene variants differ from the initial ene-like cyclization, they are still considered Conia-ene or Conia-ene-type reactions. [6] In addition, due to the complexity of some Conia-ene reaction systems, the true mechanism may lay somewhere between several different activation modes.

Enolate activation

Enolate activation is the simplest Conia-ene activation mode. In this mode, the carbonyl starting material is treated with a strong base, such as nBuLi, NaH, or tBuOK, to form a metal-stabilized enolate, which then attacks the tethered alkyne and transfers the metal cation. An early example of enolate activation was reported by Taguchi and coworkers in 1999. [7] The authors found that in the presence of catalytic base, alkynyl-substituted malonate esters undergo facile cyclization to the corresponding cyclopentanes. High yields were also obtained with substituted cyanoacetate, sulfonylacetate, and phosphonoacetate analogs.

Enolate activation.png

Alkyne activation

In Conia-ene reactions proceeding via alkyne activation, a suitable late transition metal (Au, Ag, Pt, Pd) coordinates to the alkyne and increases its electrophilicity; thus, the enol tautomer of the carbonyl can attack more readily. Toste et al. pioneered two of the first examples of alkyne activation in 2004. [8] Using a cationic Au(I) complex, the authors formed a wide variety of cyclized products from linear β-ketoester starting materials. Notably, the reactions are run under mild conditions and give high diastereoselectivity. Moreover, by shortening the alkyne tether from three carbons to two, substituted cyclopentenes can also be accessed. [9]

Alkyne activation.png

Ene-yne activation

In ene-yne activation, the least common of the five modes, a single metal species coordinates with the enol alkene and the tethered alkyne, simultaneously activating both moieties for reaction. Nickel, cobalt, and rhenium complexes have all been employed in this manner. [6] A representative example was reported by Malacria et al. in 1994, in which an alkynyl substituted β-ketoester was treated with cyclopentadienyl cobalt complex and irradiation to give disubstituted methylene cyclopentane. [10]

Ene-yne activation.png

One-metal dual activation

To effect dual activation by a single metal, the same metal species that activates the enolate also interacts with the alkyne. Though the precise mechanisms are poorly understood and likely vary from case to case, metals such as In, Zn, Fe, and Cu are proposed to operate via this mode. [6] One reaction system thought to proceed via one-metal dual activation is that developed by Shaw et al. in 2014. Using a catalytic Fe(III)-(Salen) complex, Shaw and coworkers were able to access chiral cyclopentanes from an array of alkynyl-tethered β-ketoesters and analogs thereof. [11] The reaction tolerated a wide range of ketones (phenyl, homoallyl, cyclopropyl, 2-furyl, etc.), esters (ethyl, tert-butyl, etc.), and ester analogs (nitro, phosphono, cyano, sulfonyl, etc.).

One-metal dual activation.png

Two-metal dual activation

Two-metal dual activation represents the combination of the enolate activation mode and the alkyne activation mode into a single reaction system. Generally, a hard, oxophilic metal (K, Na, Ag) activates the enolate oxygen, while a soft, carbophilic metal (Pd, Cu, Mo) coordinates with the alkyne. In some instances, however, the precise role of each metal is unclear. [6] For example, in a 2005 study Toste et al. found that treatment of an alkynyl-tethered β-ketoester with a Pd(II) phosphine complex and Yb(OTf)3 effected asymmetric cyclization to the corresponding cyclopentane. [12] It is proposed that a Pd-enolate adds into a Yb-activated alkyne, though there is also precedent for Pd activation of alkynes. [13]

Two-metal dual activation mode of the Conia-ene reaction.png

Applications in total synthesis

Following their development of Au-catalyzed Conia-ene reactions, Toste and coworkers employed such a transformation toward the alkaloid natural product lycopladine A. [14] Starting from chiral cyclohexenone 1, a series of enone functionalizations gave silyl enol ether 2 as the Conia-ene precursor. To effect cyclization, 2 was treated with catalytic AuCl(PPh3) and AgBF4 to furnish vinyl iodide 3 in high yield as a single diastereomer. The remainder of the molecule was completed in three steps to give (+)-lycopladine A in eight steps and 17% overall yield.

Conia-ene lycopladine A.png

In 2012, Carreira et al. synthesized racemic gomerone C, a halogenated terpene isolated from the red algae Laurencia majuscula, and employed Au-catalyzed Conia-ene cyclization as the penultimate step. [15] Having obtained silyl enol ether 7 in 11 steps from bicycle 6, itself the product of a Diels–Alder cycloaddition between siloxydiene4 and enone 5, the authors subject 7 to 50 mol% Echavarren’s catalyst to deliver tricycle 8 in 65% yield. This compound is then elaborated to (±)-gomerone C by chlorination of the exo-methylene.

Conia-ene gomerone C.png

In 2020, Yang and coworkers employed a diastereoselective Conia-ene reaction during their asymmetric synthesis of (+)-waihoensene, a structurally dense terpenoid from Podocarpus totara var. waihoensis, first synthesized by the Lee group in 2017. [16] Vinylogous ester 9 was first functionalized in six steps to chiral Conia-ene precursor 10. Subsequent treatment of 10 with tBuOK in DMSO gave bicycle 11 in 83% yield as a single diastereomer. This compound then required eight additional transformations to reach (+)-waihoensene in 15 steps and 4% overall yield.

Conia-ene waihoensene.png

Related Research Articles

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

The aldol reaction is a reaction in organic chemistry that combines two carbonyl compounds to form a new β-hydroxy carbonyl compound. Its simplest form might involve the nucleophilic addition of an enolized ketone to another:

<span class="mw-page-title-main">Enol</span> Organic compound with a C=C–OH group

In organic chemistry, alkenols are a type of reactive structure or intermediate in organic chemistry that is represented as an alkene (olefin) with a hydroxyl group attached to one end of the alkene double bond. The terms enol and alkenol are portmanteaus deriving from "-ene"/"alkene" and the "-ol" suffix indicating the hydroxyl group of alcohols, dropping the terminal "-e" of the first term. Generation of enols often involves deprotonation at the α position to the carbonyl group—i.e., removal of the hydrogen atom there as a proton H+. When this proton is not returned at the end of the stepwise process, the result is an anion termed an enolate. The enolate structures shown are schematic; a more modern representation considers the molecular orbitals that are formed and occupied by electrons in the enolate. Similarly, generation of the enol often is accompanied by "trapping" or masking of the hydroxy group as an ether, such as a silyl enol ether.

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

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

In organic chemistry, the Michael reaction or Michael 1,4 addition is a reaction between a Michael donor and a Michael acceptor to produce a Michael adduct by creating a carbon-carbon bond at the acceptor's β-carbon. It belongs to the larger class of conjugate additions and is widely used for the mild formation of carbon-carbon bonds.

The Robinson annulation is a chemical reaction used in organic chemistry for ring formation. It was discovered by Robert Robinson in 1935 as a method to create a six membered ring by forming three new carbon–carbon bonds. The method uses a ketone and a methyl vinyl ketone to form an α,β-unsaturated ketone in a cyclohexane ring by a Michael addition followed by an aldol condensation. This procedure is one of the key methods to form fused ring systems.

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.

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 Reformatsky reaction is an organic reaction which condenses aldehydes or ketones with α-halo esters using metallic zinc to form β-hydroxy-esters:

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 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.

The Meyer–Schuster rearrangement is the chemical reaction described as an acid-catalyzed rearrangement of secondary and tertiary propargyl alcohols to α,β-unsaturated ketones if the alkyne group is internal and α,β-unsaturated aldehydes if the alkyne group is terminal. Reviews have been published by Swaminathan and Narayan, Vartanyan and Banbanyan, and Engel and Dudley, the last of which describes ways to promote the Meyer–Schuster rearrangement over other reactions available to propargyl alcohols.

In organic chemistry, aldol reactions are acid- or base-catalyzed reactions of aldehydes or ketones.

Selenoxide elimination is a method for the chemical synthesis of alkenes from selenoxides. It is most commonly used to synthesize α,β-unsaturated carbonyl compounds from the corresponding saturated analogues. It is mechanistically related to the Cope reaction.

In organic chemistry, α-halo ketones can be reduced with loss of the halogen atom to form enolates. The α-halo ketones are readily prepared from ketones by various ketone halogenation reactions, and the products are reactive intermediates that can be used for a variety of other chemical reactions.

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.

The Danheiser benzannulation is a chemical reaction used in organic chemistry to generate highly substituted phenols in a single step. It is named after Rick L. Danheiser who developed the reaction.

Cycloisomerization is any isomerization in which the cyclic isomer of the substrate is produced in the reaction coordinate. The greatest advantage of cycloisomerization reactions is its atom economical nature, by design nothing is wasted, as every atom in the starting material is present in the product. In most cases these reactions are mediated by a transition metal catalyst, in few cases organocatalysts and rarely do they occur under thermal conditions. These cyclizations are able to be performed with excellent levels of selectivity in numerous cases and have transformed cycloisomerization into a powerful tool for unique and complex molecular construction. Cycloisomerization is a very broad topic in organic synthesis and many reactions that would be categorized as such exist. Two basic classes of these reactions are intramolecular Michael addition and Intramolecular Diels–Alder reactions. Under the umbrella of cycloisomerization, enyne and related olefin cycloisomerizations are the most widely used and studied reactions.

<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.

The ketimine Mannich reaction is an asymmetric synthetic technique using differences in starting material to push a Mannich reaction to create an enantiomeric product with steric and electronic effects, through the creation of a ketimine group. Typically, this is done with a reaction with proline or another nitrogen-containing heterocycle, which control chirality with that of the catalyst. This has been theorized to be caused by the restriction of undesired (E)-isomer by preventing the ketone from accessing non-reactive tautomers. Generally, a Mannich reaction is the combination of an amine, a ketone with a β-acidic proton and aldehyde to create a condensed product in a β-addition to the ketone. This occurs through an attack on the ketone with a suitable catalytic-amine unto its electron-starved carbon, from which an imine is created. This then undergoes electrophilic addition with a compound containing an acidic proton. It is theoretically possible for either of the carbonyl-containing molecules to create diastereomers, but with the addition of catalysts which restrict addition as of the enamine creation, it is possible to extract a single product with limited purification steps and in some cases as reported by List et al.; practical one-pot syntheses are possible. The process of selecting a carbonyl-group gives the reaction a direct versus indirect distinction, wherein the latter case represents pre-formed products restricting the reaction's pathway and the other does not. Ketimines selects a reaction group, and circumvent a requirement for indirect pathways.

References

  1. 1 2 3 Conia, J. M.; Le Perchec, P. (1975). "The thermal cyclisation of unsaturated carbonyl compounds". Synthesis. 1 (1): 1–19. doi:10.1055/s-1975-23652. S2CID   94889581.
  2. Conia, J. M. (1968). "Syntheses of cyclopropylcarbonyl compounds". Angew. Chem. Int. Ed. 7 (8): 570–575. doi:10.1002/anie.196805701.
  3. "Conia-Ene Reaction". Organic Chemistry Portal.
  4. Matthews, W. S.; Bares, J. E.; Bartmess, J. E.; Bordwell, F. G.; Cornforth, F. J.; Drucker, G. E.; Margolin, Z.; McCallum, R. J.; McCollum, G. J.; Vanier, N. R. (1975). "Equilibrium acidities of carbon acids. VI. Establishment of an absolute scale of acidities in sulfoxide solution". J. Am. Chem. Soc. 97 (24): 7006–7014. doi:10.1021/ja00857a010.
  5. Baldwin, J. E. (1976). "Rules for ring closure". J. Chem. Soc., Chem. Commun. 18 (18): 734–736. doi:10.1039/c39760000734.
  6. 1 2 3 4 5 Hack, D.; Blümel, M.; Chauhan, P.; Philipps, A. R.; Enders, D (2015). "Catalytic Conia-ene and related reactions". Chem. Soc. Rev. 44 (17): 6059–6093. doi:10.1039/c5cs00097a. PMID   26031492.
  7. Kitagawa, O.; Suzuki, T.; Fujiwara, H.; Fujita, M.; Taguchi, T. (1999). "Alkaline metallic reagent-catalyzed hydrocarbocyclization reaction of various active methine compounds having an unactivated 4-alkynyl or allenyl group". Tetrahedron Lett. 40 (24): 4585–4527. doi:10.1016/S0040-4039(99)00797-2.
  8. Kennedy-Smith, J. J.; Staben, S. T.; Toste, F. D. (2004). "Gold{I}-catalyzed Conia-ene reaction of β-ketoesters with alkynes". J. Am. Chem. Soc. 126 (13): 4526–4527. doi:10.1021/ja049487s. PMID   15070364.
  9. Staben, S. T.; Kennedy-Smith, J. J.; Toste, F. D. (2004). "Gold(I)-catalyzed 5-endo-dig carbocyclization of acetylenic dicarbonyl compounds". Angew. Chem. Int. Ed. 43 (40): 5350–5352. doi:10.1002/anie.200460844. PMID   15468061.
  10. Stammler, R.; Malacria, M. (1994). "New cobalt catalyzed cycloisomerization of β-ketoester ε-acetylenic". Synlett. 1: 92. doi:10.1055/s-1994-22751. S2CID   94650335.
  11. Shaw, S.; White, J. D. (2014). "A new iron(III)-Salen catalyst for enantioselective Conia-ene carbocyclization". J. Am. Chem. Soc. 136 (39): 13578–13581. doi:10.1021/ja507853f. PMID   25213211.
  12. Corkey, B. K.; Toste, F. D. (2005). "Catalytic enantioselective Conia-ene reaction". J. Am. Chem. Soc. 127 (49): 17168–17169. doi:10.1021/ja055059q. PMID   16332048.
  13. Streuff, J.; White, D. E.; Virgil, S. C.; Stoltz, B. M (2010). "A palladium-catalyzed enolate alkylation cascade for the formation of adjacent quaternary and tertiary stereocenters". Nat. Chem. 2 (3): 192–196. Bibcode:2010NatCh...2..192S. doi:10.1038/nchem.518. PMC   2917108 . PMID   20697457.
  14. Staben, S. T.; Kennedy-Smith, J. J.; Huang, D.; Corkey, B. K.; LaLonde, R. L.; Toste, F. D. (2006). "Gold(I)-catalyzed cyclizations of silyl enol ethers: applications to the synthesis of (+)-lycopladine A". Angew. Chem. Int. Ed. 45 (36): 5991–5994. doi:10.1002/anie.200602035. PMID   16888820.
  15. Huwyler, N.; Carreira, E. M. (2012). "Total synthesis and stereochemical revision of the chlorinated sesquiterpene (±)-gomerone C". Angew. Chem. Int. Ed. 51 (52): 13066–13069. doi:10.1002/anie.201207203. PMID   23161813.
  16. Qu, Y.; Wang, Z.; Zhang, Z.; Zhang, W.; Huang, J.; Yang, Z. (2020). "Asymmetric total synthesis of (+)-waihoensene". J. Am. Chem. Soc. 142 (14): 6511–6515. doi:10.1021/jacs.0c02143. PMID   32203659. S2CID   263581627.