fbpx
Wikipedia

Sharpless asymmetric dihydroxylation

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]

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
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 edit

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 edit

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

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

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

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

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 edit

 
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 edit

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 edit

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

 
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

Stereoselectivity edit

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

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

Further reading edit

  • Jacobsen, E. N.; Marko, I.; Mungall, W. S.; Schroeder, G.; Sharpless, K. B. (1988). "Asymmetric dihydroxylation via ligand-accelerated catalysis". J. Am. Chem. Soc. 110 (6): 1968–1970. doi:10.1021/ja00214a053.

See also edit

References edit

  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. ^ a b 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. ^ a b 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. ^ a b c d e 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. ^ a b 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. ^ a b 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.

sharpless, asymmetric, dihydroxylation, also, called, sharpless, bishydroxylation, chemical, reaction, alkene, with, osmium, tetroxide, presence, chiral, quinine, ligand, form, vicinal, diol, reaction, been, applied, alkenes, virtually, every, substitution, of. 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 Sharpless asymmetric dihydroxylationNamed after Karl Barry SharplessReaction type Addition reactionReactionAlkene OsO4 Chiral quinine ligand 1 2 diol main product IdentifiersOrganic Chemistry Portal sharpless dihydroxylationRSC ontology ID RXNO 0000142 The Sharpless dihydroxylation RL Largest substituent RM Medium sized substituent RS Smallest substituentIt 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 a and the mixture containing DHQD 2 PHAL is called AD mix b 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 Contents 1 Background 2 Reaction mechanism 2 1 2 2 vs 3 2 debate 2 2 Catalyst structure 3 Catalytic systems 4 Regioselectivity 5 Stereoselectivity 6 Further reading 7 See also 8 ReferencesBackground editAlkene 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 editThe 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 nbsp The reaction mechanism of the Sharpless dihydroxylationThe 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 nbsp Catalytic cycle of the Sharpless asymmetric dihydroxylation 2 2 vs 3 2 debate edit 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 nbsp Osmium tetroxide dihydroxylation proposed and correct mechanismThe 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 edit nbsp 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 editNumerous catalytic systems and modifications have been developed for the SAD Given below is a brief overview of the various components of the catalytic system 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 Chiral Auxiliary This is usually some kind of cinchona alkaloid 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 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 editIn general Sharpless asymmetric dihydroxylation favors oxidation of the more electron rich alkene scheme 1 22 nbsp SAD scheme 1In 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 p stacking In this manner the aryl substituent can act as a directing group 23 nbsp SAD scheme 2Stereoselectivity editThe diastereoselectivity of SAD is set primarily by the choice of ligand i e AD mix a versus AD mix b 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 nbsp SAD scheme 3It is often difficult to obtain high diastereoselectivity on cis disubstituted alkenes when both ends of the olefin have similar steric environments Further reading editJacobsen E N Marko I Mungall W S Schroeder G Sharpless K B 1988 Asymmetric dihydroxylation via ligand accelerated catalysis J Am Chem Soc 110 6 1968 1970 doi 10 1021 ja00214a053 See also editAsymmetric catalytic oxidation Milas hydroxylation Upjohn dihydroxylation Sharpless aminohydroxylationReferences edit 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 Kolb H C Van Nieuwenhze M S Sharpless K B 1994 Catalytic Asymmetric Dihydroxylation Chem Rev 94 8 2483 2547 doi 10 1021 cr00032a009 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 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 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 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 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 a b 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 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 a b 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 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 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 a b c d e Sundermeier U Dobler C Beller M Recent developments in the osmium catalyzed dihydroxylation of olefins Modern Oxidation Methods 2004 WILEY VCH Verlag GmbH amp Co KGaA Weinheim ISBN 3 527 30642 0 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 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 Thomas G Sharpless K B ACIEE 1993 32 1329 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 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 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 a b 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 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 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 a b 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 Retrieved from https en wikipedia org w index php title Sharpless asymmetric dihydroxylation amp oldid 1187636019, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.