Epoxidation of allylic alcohols

Last updated

The epoxidation of allylic alcohols is a class of epoxidation reactions in organic chemistry. One implementation of this reaction is the Sharpless epoxidation. Early work showed that allylic alcohols give facial selectivity when using meta-chloroperoxybenzoic acid (m-CPBA) as an oxidant. This selectivity was reversed when the allylic alcohol was acetylated. This finding leads to the conclusion that hydrogen bonding played a key role in selectivity and the following model was proposed. [1]

Peracid H bonding model.png

For cyclic allylic alcohols, greater selectivity is seen when the alcohol is locked in the pseudo equatorial position rather than the pseudo axial position. [2] However, it was found that for metal catalyzed systems such as those based on vanadium, reaction rates were accelerated when the hydroxyl group was in the axial position by a factor of 34. Substrates which were locked in the pseudo equatorial position were shown to undergo oxidation to form the ene-one. In both cases of vanadium catalyzed epoxidations, the epoxidized product showed excellent selectivity for the syn diastereomer. [3]

Conformationally locked directing groups.png

In the absence of hydrogen bonding, steric effects direct peroxide addition to the opposite face. However, perfluoric peracids are still able to hydrogen bond with protected alcohols and give normal selectivity with the hydrogen present on the peracid. [4]

Fluro binding model.png

Although the presence of an allylic alcohol does lead to increased stereoselectivity, the rates of these reactions are slower than systems lacking alcohols. However, the reaction rates of substrates with a hydrogen bonding group are still faster than the equivalent protected substrates. This observation is attributed to a balance of two factors. The first is the stabilization of the transition state as a result of the hydrogen bonding. The second is the electron-withdrawing nature of the oxygen, which draws electron density away from the alkene, lowering its reactivity. [5]

Kinetics with allylic hetero atoms.png

Acyclic allylic alcohols exhibit good selectivity as well. In these systems both A1,2 (steric interactions with vinyl) and A1,3 strain are considered. It has been shown that a dihedral angle of 120 best directs substrates which hydrogen bond with the directing group. This geometry allows for the peroxide to be properly positioned, as well as to allow minimal donation from the C-C pi into the C-O sigma star. [6] This donation would lower the electron density of the alkene, and deactivate the reaction. However, vanadium complexes do not hydrogen bond with their substrates. Instead they coordinate with the alcohol. This means that a dihedral angle of 40 allows for ideal position of the peroxide sigma star orbital. [7]

Roation stuff.png

In systems that hydrogen bond, A1,3 strain plays a larger role because the required geometry forces any allylic substituents to have severe A1,3 interactions, but avoids A1,2. This leads to syn addition of the resulting epoxide. In the vanadium case, the required geometry leads to severe A1,2 interactions, but avoids A1,3, leading to formation of the epoxide anti to the directing group. Vanadium catalyzed epoxidations have been shown to be very sensitive to the steric bulk of the vinyl group. [8] [9] [10]

Vinyl Bulk.png

Homoallylic alcohols are effective directing groups for epoxidations in both cyclic and acyclic systems for substrates which show hydrogen bonding. However these reactions tend to have lower levels of selectivity. [11] [12]

Homoallylic peracid.png

While hydrogen bonding substrates give the same type of selectivity in allylic and homoallylic cases, the opposite is true of vanadium catalysts.

Homoallylic VO.png

A transition state proposed by Mihelich shows that for these reactions, the driving force for selectivity is minimizing A1,3 strain in a pseudo chair structure.

The proposed transition state shows that the substrate will try to assume a conformation which minimizes the allyic strain. To do this, the least sterically bulky R group will rotate to assume the R4 position. [13]

Although peracids and metal catalyzed epoxidations show different selectivity in acyclic systems, they show relatively similar selectivity in cyclic systems For cyclic ring systems that are smaller seven or smaller or 10 or lager, similar patterns of selectivity are observed. However it has been shown that for medium-sized rings (eight and nine) peracid oxidizers show reverse selectivity, while vanadium catalyzed reactions continue to show formation of the syn epoxide. [14]

Ring size on selectivity.png

Although it is the least reactive metal catalyst for epoxidations, vanadium is highly selective for alkenes with allylic alcohols. Early work done by Sharpless shows its preference for reacting with alkenes with allylic alcohols over more substituted electron dense alkenes. In this case, Vanadium showed reverse regioselectivity from both m-CPBA and the more reactive molybdenum species. Although vanadium is generally less reactive than other metal complexes, in the presence of allylic alcohols, the rate of the reaction is accelerated beyond that of molybdenum, the most reactive metal for epoxidations. [15]

VO selectivity.png

Related Research Articles

Sharpless epoxidation

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

Sharpless asymmetric dihydroxylation is the chemical reaction of an alkene with osmium tetroxide in the presence of a chiral quinine ligand to form a vicinal diol. The reaction has been applied to alkenes of virtually every substitution, often high enantioselectivities are realized. Asymmetric dihydroxylation reactions are also highly site selective, providing products derived from reaction of the most electron-rich double bond in the substrate.

Epoxide

An epoxide is a cyclic ether with a three-atom ring. This ring approximates an equilateral triangle, which makes it strained, and hence 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.

Enamine

An enamine is an unsaturated compound derived by the condensation of an aldehyde or ketone with a secondary amine. Enamines are versatile intermediates.

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.

Dihydroxylation is the process by which an alkene is converted into a vicinal diol. Although there are many routes to accomplish this oxidation, the most common and direct processes use a high-oxidation-state transition metal. The metal is often used as a catalyst, with some other stoichiometric oxidant present. In addition, other transition metals and non-transition metal methods have been developed and used to catalyze the reaction.

Jacobsen epoxidation

The Jacobsen epoxidation, sometimes also referred to as Jacobsen-Katsuki epoxidation is a chemical reaction which allows enantioselective epoxidation of unfunctionalized alkyl- and aryl- substituted alkenes. It is complementary to the Sharpless epoxidation (used to form epoxides from the double bond in allylic alcohols). The Jacobsen epoxidation gains its stereoselectivity from a C2 symmetric manganese(III) salen-like ligand, which is used in catalytic amounts. The manganese atom transfers an oxygen atom from chlorine bleach or similar oxidant. The reaction takes its name from its inventor, Eric Jacobsen, with Tsutomu Katsuki sometimes being included. Chiral-directing catalysts are useful to organic chemists trying to control the stereochemistry of biologically active compounds and develop enantiopure drugs.

Hydroperoxide Class of chemical compounds

Hydroperoxides or peroxols are compounds containing the hydroperoxide functional group (ROOH). If the R is organic, the compounds are called organic hydroperoxides. Such compounds are a subset of organic peroxides, which have the formula ROOR. Organic hydroperoxides can either intentionally or unintentionally initiate explosive polymerisation in materials with unsaturated chemical bonds.

Wharton reaction

The Wharton olefin synthesis or the Wharton reaction is a chemical reaction that involves the reduction of α,β-epoxy ketones using hydrazine to give allylic alcohols. This reaction, introduced in 1961 by P. S. Wharton, is an extension of the Wolff–Kishner reduction. The general features of this synthesis are: 1) the epoxidation of α,β-unsaturated ketones is achieved usually in basic conditions using hydrogen peroxide solution in high yield; 2) the epoxy ketone is treated with 2–3 equivalents of a hydrazine hydrate in presence of substoichiometric amounts of acetic acid. This reaction occurs rapidly at room temperature with the evolution of nitrogen and the formation of an allylic alcohol. It can be used to synthesize carenol compounds. Wharton's initial procedure has been improved.

Shi epoxidation

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. The reaction was first discovered by Yian Shi of Colorado State University in 1996.

Asymmetric induction

In stereochemistry, asymmetric induction describes the preferential formation in a chemical reaction of one enantiomer or diastereoisomer over the other as a result of the influence of a chiral feature present in the substrate, reagent, catalyst or environment. Asymmetric induction is a key element in asymmetric synthesis.

In organic chemistry, kinetic resolution is a means of differentiating two enantiomers in a racemic mixture. In kinetic resolution, two enantiomers react with different reaction rates in a chemical reaction with a chiral catalyst or reagent, resulting in an enantioenriched sample of the less reactive enantiomer. As opposed to chiral resolution, kinetic resolution does not rely on different physical properties of diastereomeric products, but rather on the different chemical properties of the racemic starting materials. This enantiomeric excess (ee) of the unreacted starting material continually rises as more product is formed, reaching 100% just before full completion of the reaction. Kinetic resolution relies upon differences in reactivity between enantiomers or enantiomeric complexes. Kinetic resolution is a concept in organic chemistry and can be used for the preparation of chiral molecules in organic synthesis. Kinetic resolution reactions utilizing purely synthetic reagents and catalysts are much less common than the use of enzymatic kinetic resolution in application towards organic synthesis, although a number of useful synthetic techniques have been developed in the past 30 years.

Prilezhaev reaction

The Prilezhaev reaction, also known as the Prileschajew reaction or Prilezhaev epoxidation, is the chemical reaction of an alkene with a peroxy acid to form epoxides. It is named after Nikolai Prilezhaev, who first reported this reaction in 1909. A widely used peroxy acid for this reaction is meta-chloroperoxybenzoic acid (m-CPBA), due to its stability and good solubility in most organic solvents. An illustrative example is the epoxidation of trans-2-butene with m-CPBA to give trans-2,3-epoxybutane:

Electrophilic substitution of unsaturated silanes involves attack of an electrophile on an allyl- or vinylsilane. An allyl or vinyl group is incorporated at the electrophilic center after loss of the silyl group.

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.

Desulfonylation reactions are chemical reactions leading to the removal of a sulfonyl group from organic compounds. As the sulfonyl functional group is electron-withdrawing, methods for cleaving the sulfur-carbon bonds of sulfones are typically reductive in nature. Olefination or replacement with hydrogen may be accomplished using reductive desulfonylation methods.

Reductions with hydrosilanes are methods used for hydrogenations and hydrogenolysis of organic compounds. The approach is a subset of Ionic hydrogenations. In this particular method, the substrate is treated with a hydrosilane and auxiliary reagent, often a strong acid, resulting in formal transfer of hydride from silicon to carbon. This style of reduction with hydrosilanes enjoys diverse if specialized applications.

Oxaziridine Chemical compound

An oxaziridine is an organic molecule that features a three-membered heterocycle containing oxygen, nitrogen, and carbon. In their largest application, oxaziridines are intermediates in the industrial production of hydrazine. Oxaziridine derivatives are also used as specialized reagents in organic chemistry for a variety of oxidations, including alpha hydroxylation of enolates, epoxidation and aziridination of olefins, and other heteroatom transfer reactions. Oxaziridines also serve as precursors to amides and participate in [3+2] cycloadditions with various heterocumulenes to form substituted five-membered heterocycles. Chiral oxaziridine derivatives effect asymmetric oxygen transfer to prochiral enolates as well as other substrates. Some oxaziridines also have the property of a high barrier to inversion of the nitrogen, allowing for the possibility of chirality at the nitrogen center.

Vanadyl acetylacetonate Chemical compound

Vanadyl acetylacetonate is the chemical compound with the formula VO(acac)2, where acac is the conjugate base of acetylacetone. It is a blue-green solid that dissolves in polar organic solvents. The coordination complex consists of the vanadyl group, VO2+, bound to two acac ligands via the two oxygen atoms on each. Like other charge-neutral acetylacetonate complexes, it is not soluble in water.

Metal-catalyzed C–H borylation reactions are transition metal catalyzed organic reactions that produce an organoboron compound through functionalization of aliphatic and aromatic C–H bonds and are therefore useful reactions for carbon–hydrogen bond activation. Metal-catalyzed C–H borylation reactions utilize transition metals to directly convert a C–H bond into a C–B bond. This route can be advantageous compared to traditional borylation reactions by making use of cheap and abundant hydrocarbon starting material, limiting prefunctionalized organic compounds, reducing toxic byproducts, and streamlining the synthesis of biologically important molecules. Boronic acids, and boronic esters are common boryl groups incorporated into organic molecules through borylation reactions. Boronic acids are trivalent boron-containing organic compounds that possess one alkyl substituent and two hydroxyl groups. Similarly, boronic esters possess one alkyl substituent and two ester groups. Boronic acids and esters are classified depending on the type of carbon group (R) directly bonded to boron, for example alkyl-, alkenyl-, alkynyl-, and aryl-boronic esters. The most common type of starting materials that incorporate boronic esters into organic compounds for transition metal catalyzed borylation reactions have the general formula (RO)2B-B(OR)2. For example, bis(pinacolato)diboron (B2Pin2), and bis(catecholato)diborane (B2Cat2) are common boron sources of this general formula.

References

  1. Henbest, H. B.; Wilson, R. A. L. (1957). "376. Aspects of stereochemistry. Part I. Stereospecificity in formation of epoxides from cyclic allylic alcohols". Journal of the Chemical Society (Resumed): 1958. doi:10.1039/JR9570001958.
  2. Chamberlain, P.; Roberts, M. L.; Whitham, G. H. (1970). "Epoxidation of allylic alcohols with peroxy-acids. Attempts to define transition state geometry". Journal of the Chemical Society B: Physical Organic: 1374. doi:10.1039/J29700001374.
  3. Weyerstahl, Peter; Marschall-Weyerstahl, Helga; Penninger, Josef; Walther, Lutz (1987). "Terpenes and terpene derivatives-22". Tetrahedron. 43 (22): 5287–5298. doi:10.1016/S0040-4020(01)87705-X.
  4. McKittrick, Brian A.; Ganem, Bruce (1985). "Syn-stereoselective epoxidation of allylic ethers using CF3CO3H". Tetrahedron Letters. 26 (40): 4895–4898. doi:10.1016/S0040-4039(00)94979-7.
  5. Hoveyda, Amir H.; Evans, David A.; Fu, Gregory C. (1993). "Substrate-directable chemical reactions". Chemical Reviews. 93 (4): 1307–1370. doi:10.1021/cr00020a002.
  6. Houk, K. N.; Paddon-Row, M. N.; Rondan, N. G.; Wu, Y. D.; Brown, F. K.; Spellmeyer, D. C.; Metz, J. T.; Li, Y .; Loncharich, R. J. Science, 1986, 231, 1108-1117.
  7. Waldemar, A.; Wirth, T. Accounts of Chemical Research, 1999, 32.8, 703-710.
  8. Mihelich, Edward D. (1979). "Vanadium-catalyzed epoxidations. I. A new selectivity pattern for acyclic allylic alcohols". Tetrahedron Letters. 20 (49): 4729–4732. doi:10.1016/S0040-4039(01)86695-8.
  9. Rossiter, B.E.; Verhoeven, T.R.; Sharpless, K.B. (1979). "Stereoselective epoxidation of acyclic allylic alcohols. A correction of our previous work". Tetrahedron Letters. 20 (49): 4733–4736. doi:10.1016/S0040-4039(01)86696-X.
  10. Narula, Acharan S. (1982). "Stereoselective introduction of chiral centres in acyclic precursors: A probe into the transition state for V5+-catalyzed t-butylhydroperoxide (TBHP) epoxidation of acyclic allylic alcohols and its synthetic implications". Tetrahedron Letters. 23 (52): 5579–5582. doi:10.1016/S0040-4039(00)85899-2.
  11. Cragg, G. M. L.; Meakins, G. D. (1965). "366. Steroids of unnatural configuration. Part IX. Oxidation of 9α-lumisterol (pyrocalciferol) and 9β-ergosterol (isopyrocalciferol) with perbenzoic acid". J. Chem. Soc.: 2054–2063. doi:10.1039/JR9650002054.
  12. Johnson, M. R.; Kishi, Y. Tetrahedron Lett., 1979, 4347-4350.
  13. Mihelich, Edward D.; Daniels, Karen; Eickhoff, David J. (1981). "Vanadium-catalyzed epoxidations. 2. Highly stereoselective epoxidations of acyclic homoallylic alcohols predicted by a detailed transition-state model". Journal of the American Chemical Society. 103 (25): 7690–7692. doi:10.1021/ja00415a067.
  14. Itoh, Takashi; Jitsukawa, Koichiro; Kaneda, Kiyotomi; Teranishi, Shiichiro (1979). "Vanadium-catalyzed epoxidation of cyclic allylic alcohols. Stereoselectivity and stereocontrol mechanism". Journal of the American Chemical Society. 101: 159–169. doi:10.1021/ja00495a027.
  15. Sharpless, K. B.; Michaelson, R. C. J. Am. Chem. Soc., 1973, 95 (18)