Sharpless asymmetric dihydroxylation

Last updated
Sharpless asymmetric dihydroxylation
Named after Karl Barry Sharpless
Reaction type Addition reaction
Reaction
Alkene
+
OsO4
+
Chiral quinine ligand
1,2-diol (main product)
Identifiers
Organic Chemistry Portal sharpless-dihydroxylation
RSC ontology ID RXNO:0000142

Sharpless asymmetric dihydroxylation (also called the Sharpless bishydroxylation) 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, with the chiral outcome controlled by the choice of dihydroquinidine (DHQD) vs dihydroquinine (DHQ) as the ligand. Asymmetric dihydroxylation reactions are also highly site selective, providing products derived from reaction of the most electron-rich double bond in the substrate. [1] [2] [3]

Contents

The Sharpless dihydroxylation.
RL = Largest substituent; RM = Medium-sized substituent; RS = Smallest substituent Sharpless Dihydroxylation Scheme.png
The Sharpless dihydroxylation.
RL = Largest substituent; RM = Medium-sized substituent; RS = Smallest substituent

It is common practice to perform this reaction using a catalytic amount of osmium tetroxide, which after reaction is regenerated with reoxidants such as potassium ferricyanide [4] [5] or N-methylmorpholine N-oxide. [6] [7] This dramatically reduces the amount of the highly toxic and very expensive osmium tetroxide needed. These four reagents are commercially available premixed ("AD-mix"). The mixture containing (DHQ)2-PHAL is called AD-mix-α, and the mixture containing (DHQD)2-PHAL is called AD-mix-β. [8]

Such chiral diols are important in organic synthesis. The introduction of chirality into nonchiral reactants through usage of chiral catalysts is an important concept in organic synthesis. This reaction was developed principally by K. Barry Sharpless building on the already known racemic Upjohn dihydroxylation, for which he was awarded a share of the 2001 Nobel Prize in Chemistry.

Background

Alkene dihydroxylation by osmium tetroxide is an old and extremely useful method for the functionalization of olefins. However, since osmium(VIII) reagents like osmium tetroxide (OsO4) are expensive and extremely toxic, it has become desirable to develop catalytic variants of this reaction. Some stoichiometric terminal oxidants that have been employed in these catalytic reactions include potassium chlorate, hydrogen peroxide (Milas hydroxylation), N-Methylmorpholine N-oxide (NMO, Upjohn dihydroxylation), tert-butyl hydroperoxide (tBHP), and potassium ferricyanide (K3Fe(CN)6). K. Barry Sharpless was the first to develop a general, reliable enantioselective alkene dihydroxylation, referred to as the Sharpless asymmetric dihydroxylation (SAD). Low levels of OsO4 are combined with a stoichiometric ferricyanide oxidant in the presence of chiral nitrogenous ligands to create an asymmetric environment around the oxidant.

Reaction mechanism

The reaction mechanism of the Sharpless dihydroxylation begins with the formation of the osmium tetroxide – ligand complex (2). A [3+2]-cycloaddition with the alkene (3) gives the cyclic intermediate 4. [9] [10] Basic hydrolysis liberates the diol (5) and the reduced osmate (6). Methanesulfonamide (CH3SO2NH2) has been identified as a catalyst to accelerate this step of the catalytic cycle and if frequently used as an additive to allow non-terminal alkene substrates to react efficiently at 0 °C. [8] Finally, the stoichiometric oxidant regenerates the osmium tetroxide – ligand complex (2).

The reaction mechanism of the Sharpless dihydroxylation Sharpless Dihydroxylation Mechanism.png
The reaction mechanism of the Sharpless dihydroxylation

The mechanism of the Sharpless asymmetric dihydroxylation has been extensively studied and a potential secondary catalytic cycle has been identified (see below). [11] [12] If the osmylate ester intermediate is oxidized before it dissociates, then an osmium(VIII)-diol complex is formed which may then dihydroxylate another alkene. [13] Dihydroxylations resulting from this secondary pathway generally suffer lower enantioselectivities than those resulting from the primary pathway. A schematic showing this secondary catalytic pathway is shown below. This secondary pathway may be suppressed by using a higher molar concentration of ligand.

Catalytic cycle of the Sharpless asymmetric dihydroxylation Sharpless Asymmetric Dihydroxylation catalytic cycle.png
Catalytic cycle of the Sharpless asymmetric dihydroxylation

[2+2] vs [3+2] debate

In his original report Sharpless suggested the reaction proceeded via a [2+2] cycloaddition of OsO4 onto the alkene to give an osmaoxetane intermediate (see below). [14] This intermediate would then undergo a 1,1- migratory insertion to form an osmylate ester which after hydrolysis would give the corresponding diol. In 1989 E. J. Corey published a slightly different variant of this reaction and suggested that the reaction most likely proceeded via a [3+2] cycloaddition of OsO4 with the alkene to directly generate the osmylate ester. [15] Corey's suggestion was based on a previous computational study done by Jorgensen and Hoffmann which determined the [3+2] reaction pathway to be the lower energy pathway. In addition Corey reasoned that steric repulsions in the octahedral intermediate would disfavor the [2+2] pathway.

Osmium tetroxide dihydroxylation proposed and correct mechanism Osmium tetroxide dihydroxylation proposed and correct mechanism.png
Osmium tetroxide dihydroxylation proposed and correct mechanism

The next ten years saw numerous publications by both Corey and Sharpless, each supporting their own version of the mechanism. While these studies were not able to distinguish between the two proposed cyclization pathways, they were successful in shedding light on the mechanism in other ways. For example, Sharpless provided evidence for the reaction proceeding via a step-wise mechanism. [16] Additionally both Sharpless and Corey showed that the active catalyst possesses a U-shaped chiral binding pocket. [17] [18] [19] Corey also showed that the catalyst obeys Michaelis-Menten kinetics and acts like an enzyme pocket with a pre-equilibrium. [20] In the February 1997 issue of the Journal of the American Chemical Society Sharpless published the results of a study (a Hammett analysis) which he claimed supported a [2+2] cyclization over a [3+2]. [21] In the October issue of the same year, however, Sharpless also published the results of another study conducted in collaboration with Ken Houk and Singleton which provided conclusive evidence for the [3+2] mechanism. [10] Thus Sharpless was forced to concede the decade-long debate.

Catalyst structure

Allyl benzoate bound within the U-shaped binding pocket of the active dihydroquinidine catalyst, osmium tetroxide interacting with the Re face. SAD active catalyst conformation.png
Allyl benzoate bound within the U-shaped binding pocket of the active dihydroquinidine catalyst, osmium tetroxide interacting with the Re face.

Crystallographic evidence has shown that the active catalyst possesses a pentacoordinate osmium species held in a U-shaped binding pocket. The nitrogenous ligand holds OsO4 in a chiral environment making approach of one side of the olefin sterically hindered while the other is not. [20]

Catalytic systems

Numerous catalytic systems and modifications have been developed for the SAD. Given below is a brief overview of the various components of the catalytic system:

  1. Catalytic Oxidant: This is always OsO4, however certain additives can coordinate to the osmium(VIII) and modify its electronic properties. OsO4 is often generated in situ from K2OsO2(OH)4 (an Os(VI) species) due to safety concerns.
  2. Chiral Auxiliary: This is usually some kind of cinchona alkaloid.
  3. Stoichiometric Oxidant:
    • Peroxides were among the first stoichiometric oxidants to be used in this catalytic cycle; see the Milas hydroxylation. Drawbacks of peroxides include chemoselectivity issues. [13]
    • Trialkylammonium N-oxides, such as NMO—as in the Upjohn Reaction—and trimethylamine N-oxide. [13]
    • Potassium ferricyanide (K3Fe(CN)6) is the most commonly used stoichiometric oxidant for the reaction, and is the oxidant that comes in the commercially available AD-mix preparations.
  4. Additive:
    • Citric acid: Osmium tetroxide is an electrophilic oxidant and as such reacts slowly with electron-deficient olefins. It has been found that the rate of oxidation of electron-deficient olefins can be accelerated by maintaining the pH of the reaction slightly acidic. [13] On the other hand, a high pH can increase the rate of oxidation of internal olefins, and also increase the enantiomeric excess (e.e.) for the oxidation of terminal olefins. [13]

Regioselectivity

In general Sharpless asymmetric dihydroxylation favors oxidation of the more electron-rich alkene (scheme 1). [22]

SAD scheme 1 SAD scheme 1.png
SAD scheme 1

In this example SAD gives the diol of the alkene closest to the (electron-withdrawing) para-methoxybenzoyl group, albeit in low yield. This is likely due to the ability of the aryl ring to interact favorably with the active site of the catalyst via π-stacking. In this manner the aryl substituent can act as a directing group. [23]

SAD scheme 2 SAD scheme 2.png
SAD scheme 2

Stereoselectivity

The diastereoselectivity of SAD is set primarily by the choice of ligand (i.e. AD-mix-α versus AD-mix-β), however factors such as pre-existing chirality in the substrate or neighboring functional groups may also play a role. In the example shown below, the para-methoxybenzoyl substituent serves primarily as a source of steric bulk to allow the catalyst to differentiate the two faces of the alkene. [23]

SAD scheme 3 SAD scheme 3.png
SAD scheme 3

It is often difficult to obtain high diastereoselectivity on cis-disubstituted alkenes when both ends of the olefin have similar steric environments.

Further reading

See also

Related Research Articles

<span class="mw-page-title-main">Osmium tetroxide</span> Chemical compound

Osmium tetroxide (also osmium(VIII) oxide) is the chemical compound with the formula OsO4. The compound is noteworthy for its many uses, despite its toxicity and the rarity of osmium. It also has a number of unusual properties, one being that the solid is volatile. The compound is colourless, but most samples appear yellow. This is most likely due to the presence of the impurity OsO2, which is yellow-brown in colour. In biology, its property of binding to lipids has made it a widely-used stain in electron microscopy.

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

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.

<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.

<span class="mw-page-title-main">AD-mix</span>

In organic chemistry, AD-mix is a commercially available mixture of reagents that acts as an asymmetric catalyst for various chemical reactions, including the Sharpless asymmetric dihydroxylation of alkenes. The two letters AD, stand for asymmetric dihydroxylation. The mix is available in two variations, "AD-mix α" and "AD-mix β" following ingredient lists published by Barry Sharpless.

(<i>E</i>)-Stilbene Chemical compound

(E)-Stilbene, commonly known as trans-stilbene, is an organic compound represented by the condensed structural formula C6H5CH=CHC6H5. Classified as a diarylethene, it features a central ethylene moiety with one phenyl group substituent on each end of the carbon–carbon double bond. It has an (E) stereochemistry, meaning that the phenyl groups are located on opposite sides of the double bond, the opposite of its geometric isomer, cis-stilbene. Trans-stilbene occurs as a white crystalline solid at room temperature and is highly soluble in organic solvents. It can be converted to cis-stilbene photochemically, and further reacted to produce phenanthrene.

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.

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

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.

Ruthenium tetroxide is the inorganic compound with the formula RuO4. It is a yellow volatile solid that melts near room temperature. It has the odor of ozone. Samples are typically black due to impurities. The analogous OsO4 is more widely used and better known. It is also the anhydride of hyperruthenic acid (H2RuO5). One of the few solvents in which RuO4 forms stable solutions is CCl4.

The Sharpless oxyamination is the chemical reaction that converts an alkene to a vicinal amino alcohol. The reaction is related to the Sharpless dihydroxylation, which converts alkenes to vicinal diols. Vicinal amino-alcohols are important products in organic synthesis and recurring pharmacophores in drug discovery.

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.

Asymmetric catalytic oxidation is a technique of oxidizing various substrates to give an enantio-enriched product using a catalyst. Typically, but not necessarily, asymmetry is induced by the chirality of the catalyst. Typically, but again not necessarily, the methodology applies to organic substrates. Functional groups that can be prochiral and readily susceptible to oxidation include certain alkenes and thioethers. Challenging but pervasive prochiral substrates are C-H bonds of alkanes. Instead of introducing oxygen, some catalysts, biological and otherwise, enantioselectively introduce halogens, another form of oxidation.

The Upjohn dihydroxylation is an organic reaction which converts an alkene to a cis vicinal diol. It was developed by V. VanRheenen, R. C. Kelly and D. Y. Cha of the Upjohn Company in 1976. It is a catalytic system using N-methylmorpholine N-oxide (NMO) as stoichiometric re-oxidant for the osmium tetroxide. It is superior to previous catalytic methods.

The Milas hydroxylation is an organic reaction converting an alkene to a vicinal diol, and was developed by Nicholas A. Milas in the 1930s. The cis-diol is formed by reaction of alkenes with hydrogen peroxide and either ultraviolet light or a catalytic osmium tetroxide, vanadium pentoxide, or chromium trioxide.

The Lemieux–Johnson or Malaprade–Lemieux–Johnson oxidation is a chemical reaction in which an olefin undergoes oxidative cleavage to form two aldehyde or ketone units. The reaction is named after its inventors, Raymond Urgel Lemieux and William Summer Johnson, who published it in 1956. The reaction proceeds in a two step manner, beginning with dihydroxylation of the alkene by osmium tetroxide, followed by a Malaprade reaction to cleave the diol using periodate. Excess periodate is used to regenerate the osmium tetroxide, allowing it to be used in catalytic amounts. The Lemieux–Johnson reaction ceases at the aldehyde stage of oxidation and therefore produces the same results as ozonolysis.

<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.

<span class="mw-page-title-main">DuPhos</span> Class of chemical compounds

DuPhos is a class of organophosphorus compound that are used ligands for asymmetric synthesis. The name DuPhos is derived from (1) the chemical company that sponsored the research leading to this ligand's invention, DuPont and (2) the compound is a diphosphine ligand type. Specifically it is classified as a C2-symmetric ligand, consisting of two phospholanes rings affixed to a benzene ring.

Alcohol oxidation is a collection of oxidation reactions in organic chemistry that convert alcohols to aldehydes, ketones, carboxylic acids, and esters where the carbon carries a higher oxidation state. The reaction mainly applies to primary and secondary alcohols. Secondary alcohols form ketones, while primary alcohols form aldehydes or carboxylic acids.

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.

<span class="mw-page-title-main">Potassium osmate</span> Chemical compound

Potassium osmate is the inorganic compound with the formula K2[OsO2(OH)4]. This diamagnetic purple salt contains osmium in the VI (6+) oxidation state. When dissolved in water a pink solution is formed but when dissolved in methanol, the salt gives a blue solution. The salt gained attention as a catalyst for the asymmetric dihydroxylation of olefins.

In homogeneous catalysis, C2-symmetric ligands refer to ligands that lack mirror symmetry but have C2 symmetry. Such ligands are usually bidentate and are valuable in catalysis. The C2 symmetry of ligands limits the number of possible reaction pathways and thereby increases enantioselectivity, relative to asymmetrical analogues. C2-symmetric ligands are a subset of chiral ligands. Chiral ligands, including C2-symmetric ligands, combine with metals or other groups to form chiral catalysts. These catalysts engage in enantioselective chemical synthesis, in which chirality in the catalyst yields chirality in the reaction product.

References

  1. Noe, Mark C.; Letavic, Michael A.; Snow, Sheri L. (15 December 2005). "Asymmetric Dihydroxylation of Alkenes". Org. React. 66 (109): 109–625. doi:10.1002/0471264180.or066.02. ISBN   0471264180.
  2. Kolb, H. C.; Van Nieuwenhze, M. S.; Sharpless, K. B. (1994). "Catalytic Asymmetric Dihydroxylation". Chem. Rev. 94 (8): 2483–2547. doi:10.1021/cr00032a009.
  3. Gonzalez, Javier; Aurigemma, Christine; Truesdale, Larry (2004). "Synthesis of (+)-(1S,2R)- and (−)-(1R,2S)-trans-2-Phenylcyclohexanol Via Sharpless Asymmetric Dihydroxylation (AD)". Organic Syntheses . 79: 93. doi:10.15227/orgsyn.079.0093.
  4. Minato, M.; Yamamoto, K.; Tsuji, J. (1990). "Osmium tetraoxide catalyzed vicinal hydroxylation of higher olefins by using hexacyanoferrate(III) ion as a cooxidant". J. Org. Chem. 55 (2): 766–768. doi:10.1021/jo00289a066.
  5. Oi, R.; Sharpless, K. B. (1996). "3-[(1S)-1,2-Dihydroxyethyl]-1,5-Dihydro-3H-2,4-Benzodioxepine". Organic Syntheses . 73: 1. doi:10.15227/orgsyn.073.0001.; Collective Volume, vol. 9, p. 251
  6. VanRheenen, V.; Kelly, R. C.; Cha, D. Y. (1976). "An improved catalytic OsO4 oxidation of olefins to cis-1,2-glycols using tertiary amine oxides as the oxidant". Tetrahedron Lett. 17 (23): 1973–1976. doi:10.1016/s0040-4039(00)78093-2.
  7. McKee, B. H.; Gilheany, D. G.; Sharpless, K. B. (1992). "(R,R)-1,2-Diphenyl-1,2-ethanediol (Stilbene diol)". Organic Syntheses . 70: 47. doi:10.15227/orgsyn.070.0047.; Collective Volume, vol. 9, p. 383
  8. 1 2 Sharpless, K. B.; Amberg, Willi; Bennani, Youssef L.; et al. (1992). "The osmium-catalyzed asymmetric dihydroxylation: A new ligand class and a process improvement". J. Org. Chem. 57 (10): 2768–2771. doi:10.1021/jo00036a003.
  9. Corey, E.J.; Noe, M. C.; Grogan, M. J. (1996). "Experimental test of the [3+2]- and [2+2]-cycloaddition pathways for the bis-cinchona alkaloid-OsO4 catalyzed dihydroxylation of olefins by means of kinetic isotope effects". Tetrahedron Lett. 37 (28): 4899–4902. doi:10.1016/0040-4039(96)01005-2.
  10. 1 2 DelMonte, A. J.; Haller, J.; Houk, K. N.; Sharpless, K. B.; Singleton, D. A.; Strassner, T.; Thomas, A. A. (1997). "Experimental and Theoretical Kinetic Isotope Effects for Asymmetric Dihydroxylation. Evidence Supporting a Rate-Limiting "(3 + 2)" Cycloaddition". J. Am. Chem. Soc. 119 (41): 9907–9908. doi:10.1021/ja971650e.
  11. Ogino, Y.; Chen, H.; Kwong, H.-L.; Sharpless, K. B. (1991). "On the timing of hydrolysis / reoxidation in the osmium-catalyzed asymmetric dihydroxylation of olefins using potassium ferricyanide as the reoxidant". Tetrahedron Lett. 3 (2): 3965–3968. doi:10.1016/0040-4039(91)80601-2.
  12. Wai, J. S. M.; Marko, I.; Svendsen, J. N.; Finn, M. G.; Jacobsen, E. N.; Sharpless, K. Barry (1989). "A mechanistic insight leads to a greatly improved osmium-catalyzed asymmetric dihydroxylation process". J. Am. Chem. Soc. 111 (3): 1123. doi:10.1021/ja00185a050.
  13. 1 2 3 4 5 Sundermeier, U., Dobler, C., Beller, M. Recent developments in the osmium-catalyzed dihydroxylation of olefins. Modern Oxidation Methods. 2004 WILEY-VCH Verlag GmbH & Co. KGaA,Weinheim. ISBN   3-527-30642-0
  14. Hentges, Steven G.; Sharpless, K. Barry (June 1980). "Asymmetric induction in the reaction of osmium tetroxide with olefins". J. Am. Chem. Soc. 102 (12): 4263. doi:10.1021/ja00532a050.
  15. Corey, E. J.; DaSilva Jardine, Paul; Virgil, Scott; Yuen, Po Wai; Connell, Richard D. (December 1989). "Enantioselective vicinal hydroxylation of terminal and E-1,2-disubstituted olefins by a chiral complex of osmium tetroxide. An effective controller system and a rational mechanistic model". J. Am. Chem. Soc. 111 (26): 9243. doi:10.1021/ja00208a025.
  16. Thomas, G.; Sharpless, K. B. ACIEE 1993, 32, 1329
  17. Corey, E. J.; Noe, Mark C. (December 1993). "Rigid and highly enantioselective catalyst for the dihydroxylation of olefins using osmium tetraoxide clarifies the origin of enantiospecificity". J. Am. Chem. Soc. 26 (115): 12579. doi:10.1021/ja00079a045.
  18. Kolb, H. C.; Anderson, P. G.; Sharpless, K. B. (February 1994). "Toward an Understanding of the High Enantioselectivity in the Osmium-Catalyzed Asymmetric Dihydroxylation (AD). 1. Kinetics". J. Am. Chem. Soc. 116 (1278): 1278. doi:10.1021/ja00083a014.
  19. Corey, E. J.; Noe, Mark C.; Sarshar, Sepehr (1994). "X-ray crystallographic studies provide additional evidence that an enzyme-like binding pocket is crucial to the enantioselective dihydroxylation of olefins by OsO4-bis-cinchona alkaloid complexes". Tetrahedron Letters . 35 (18): 2861. doi:10.1016/s0040-4039(00)76644-5.
  20. 1 2 Corey, E. J.; Noe, M. C. (17 January 1996). "Kinetic Investigations Provide Additional Evidence That an Enzyme-like Binding Pocket Is Crucial for High Enantioselectivity in the Bis-Cinchona Alkaloid Catalyzed Asymmetric Dihydroxylation of Olefins". J. Am. Chem. Soc. 118 (2): 319. doi:10.1021/ja952567z.
  21. Sharpless, K. B.; Gypser, Andreas; Ho, Pui Tong; Kolb, Hartmuth C.; Kondo, Teruyuki; Kwong, Hoi-Lun; McGrath, Dominic V.; Rubin, A. Erik; Norrby, Per-Ola; Gable, Kevin P.; Sharpless, K. Barry (1997). "Toward an Understanding of the High Enantioselectivity in the Osmium-Catalyzed Asymmetric Dihydroxylation. 4. Electronic Effects in Amine-Accelerated Osmylations". J. Am. Chem. Soc. 119 (8): 1840. doi:10.1021/ja961464t.
  22. Xu, D.; Crispino, G. A.; Sharpless, K. B. (September 1992). "Selective asymmetric dihydroxylation (AD) of dienes". J. Am. Chem. Soc. 114 (19): 7570–7571. doi:10.1021/ja00045a043.
  23. 1 2 Corey, E. J.; Guzman-Perez, Angel; Noe, Mark C. (November 1995). "The application of a mechanistic model leads to the extension of the Sharpless asymmetric dihydroxylation to allylic 4-methoxybenzoates and conformationally related amine and homoallylic alcohol derivatives". J. Am. Chem. Soc. 117 (44): 10805–10816. doi:10.1021/ja00149a003.