B. Solubility of Organic Compounds in Water

22
Stereoselective Organic Reactions in Water Ulf M. Lindstro ¨m* Bioorganic Chemistry, Center for Chemistry and Chemical Engineering, Lund University, P.O. Box 124, SE-221 00 Lund, Sweden Received December 18, 2001 Contents I. Introduction 2751 II. Water as Solvent 2752 A. Why Water? 2752 B. Solubility of Organic Compounds in Water 2752 1. Organic Cosolvents 2753 2. Ionic Derivatization (pH control) 2753 3. Surfactants 2753 4. Hydrophilic Auxiliaries 2753 III. Catalysis in Water 2754 A. Homogeneous Catalysis 2754 1. Brønsted Catalysis 2754 2. Lewis-Acid Catalysis 2754 B. Heterogeneous Catalysis 2756 C. Micellar Catalysis 2757 IV. Stereoselectivity in Water 2757 A. Stereocontrol in Water 2758 1. Chelation Control 2758 2. Hydrophobic Control 2759 B. Water-Soluble Chiral Auxiliaries 2759 V. Additions to C-C Multiple Bonds 2760 A. Cycloadditions 2760 1. Diels-Alder Reaction 2760 2. Hetero Diels-Alder Reaction 2761 B. Michael-Type Addition 2762 C. Dihydroxylation 2762 D. Aminohydroxylation 2763 E. Epoxidation 2763 F. Cyclopropanation 2763 VI. Additions to C-X Multiple Bonds 2764 A. Aldol-Type Reactions 2764 B. Allylation Reaction 2765 C. Mannich Reaction 2766 D. Hydride Addition 2766 VII. Substitution Reactions 2766 VIII. Rearrangements 2767 A. Claisen Rearrangement 2767 IX. Hydrogenations 2767 X. Radical Reactions 2768 XI. Conclusions 2769 XII. Acknowledgments 2769 XIII. References 2769 I. Introduction Many organic chemists, myself included, will re- member being told in early undergraduate years “in organic synthesis water is a contaminant”. From the very beginning it was made clear to you that in order to become proficient in organic synthesis, the first step was knowing how to dry your glassware, re- agents, and solvents. Extreme measures should always be taken to avoid even trace quantities of water in your reaction flask, unless of course water was used as a reactant. In undergraduate labs, water was always the usual suspect whenever a reaction gave an unexpected outcome; “My reaction did not work”! “Well, was your glassware properly dried?” With the paradigm being that water has an adverse effect on most organic reactions, it may require a significant mind leap to think of water as a versatile solvent for organic synthesis. Nevertheless, in the most recent decades, chemists have begun investigating the possibility of using water as solvent for organic reactions, with some- times surprising and unforeseen results. Without overlooking some earlier contributions, the discover- * To whom correspondence should be addressed. Phone: (+46) 462228211. Fax: (+46) 462228209. E-mail: ulf.lindstrom@ bioorganic.lth.se. Ulf Lindstro ¨ m was born in 1971 in Malmo ¨ , Sweden. He studied Chemistry and Biology at Lund University to receive his B.Sc. degree in 1996. After earning a licentiate degree in Engineering at Lund Institute of Technology in 1998, he moved to Stockholm University to complete his doctoral thesis in 2000 under the supervision of Professor Peter Somfai. His Ph.D. work mainly concerned the development of new methods of preparing optically active nitrogen heterocycles. He then moved to California and Stanford University for postdoctoral research with Professor Eric T. Kool. There he was exposed to different aspects of nucleic acid chemistry, for example, the synthesis and molecular recognition of circular DNA. He joined the Department of Bioorganic Chemistry at Lund University as Assistant Professor in the spring of 2002. His research interest, in general, is molecular recognition and reactivity in water and, in particular, the development of chiral ligands for aqueous asymmetric synthesis. 2751 Chem. Rev. 2002, 102, 2751-2772 10.1021/cr010122p CCC: $39.75 © 2002 American Chemical Society Published on Web 06/19/2002

Transcript of B. Solubility of Organic Compounds in Water

Page 1: B. Solubility of Organic Compounds in Water

Stereoselective Organic Reactions in Water

Ulf M. Lindstrom*

Bioorganic Chemistry, Center for Chemistry and Chemical Engineering, Lund University, P.O. Box 124, SE-221 00 Lund, Sweden

Received December 18, 2001

ContentsI. Introduction 2751II. Water as Solvent 2752

A. Why Water? 2752B. Solubility of Organic Compounds in Water 2752

1. Organic Cosolvents 27532. Ionic Derivatization (pH control) 27533. Surfactants 27534. Hydrophilic Auxiliaries 2753

III. Catalysis in Water 2754A. Homogeneous Catalysis 2754

1. Brønsted Catalysis 27542. Lewis-Acid Catalysis 2754

B. Heterogeneous Catalysis 2756C. Micellar Catalysis 2757

IV. Stereoselectivity in Water 2757A. Stereocontrol in Water 2758

1. Chelation Control 27582. Hydrophobic Control 2759

B. Water-Soluble Chiral Auxiliaries 2759V. Additions to C−C Multiple Bonds 2760

A. Cycloadditions 27601. Diels−Alder Reaction 27602. Hetero Diels−Alder Reaction 2761

B. Michael-Type Addition 2762C. Dihydroxylation 2762D. Aminohydroxylation 2763E. Epoxidation 2763F. Cyclopropanation 2763

VI. Additions to C−X Multiple Bonds 2764A. Aldol-Type Reactions 2764B. Allylation Reaction 2765C. Mannich Reaction 2766D. Hydride Addition 2766

VII. Substitution Reactions 2766VIII. Rearrangements 2767

A. Claisen Rearrangement 2767IX. Hydrogenations 2767X. Radical Reactions 2768XI. Conclusions 2769XII. Acknowledgments 2769XIII. References 2769

I. IntroductionMany organic chemists, myself included, will re-

member being told in early undergraduate years “in

organic synthesis water is a contaminant”. From thevery beginning it was made clear to you that in orderto become proficient in organic synthesis, the firststep was knowing how to dry your glassware, re-agents, and solvents. Extreme measures shouldalways be taken to avoid even trace quantities ofwater in your reaction flask, unless of course waterwas used as a reactant. In undergraduate labs, waterwas always the usual suspect whenever a reactiongave an unexpected outcome; “My reaction did notwork”! “Well, was your glassware properly dried?”With the paradigm being that water has an adverseeffect on most organic reactions, it may require asignificant mind leap to think of water as a versatilesolvent for organic synthesis.

Nevertheless, in the most recent decades, chemistshave begun investigating the possibility of usingwater as solvent for organic reactions, with some-times surprising and unforeseen results. Withoutoverlooking some earlier contributions, the discover-

* To whom correspondence should be addressed. Phone: (+46)462228211. Fax: (+46) 462228209. E-mail: [email protected].

Ulf Lindstrom was born in 1971 in Malmo, Sweden. He studied Chemistryand Biology at Lund University to receive his B.Sc. degree in 1996. Afterearning a licentiate degree in Engineering at Lund Institute of Technologyin 1998, he moved to Stockholm University to complete his doctoral thesisin 2000 under the supervision of Professor Peter Somfai. His Ph.D. workmainly concerned the development of new methods of preparing opticallyactive nitrogen heterocycles. He then moved to California and StanfordUniversity for postdoctoral research with Professor Eric T. Kool. Therehe was exposed to different aspects of nucleic acid chemistry, for example,the synthesis and molecular recognition of circular DNA. He joined theDepartment of Bioorganic Chemistry at Lund University as AssistantProfessor in the spring of 2002. His research interest, in general, ismolecular recognition and reactivity in water and, in particular, thedevelopment of chiral ligands for aqueous asymmetric synthesis.

2751Chem. Rev. 2002, 102, 2751−2772

10.1021/cr010122p CCC: $39.75 © 2002 American Chemical SocietyPublished on Web 06/19/2002

Page 2: B. Solubility of Organic Compounds in Water

ies made in the laboratories of Breslow1-3 andGrieco4,5 in the early 1980s on the positive effect ofwater on rates and selectivities of Diels-Alder reac-tions are often recognized as the “Big Bang” inaqueous synthesis that triggered a more widespreadinterest in the field. Since then, significant progresshas been made in the field of organic chemistry inwater/aqueous media, and new additions are con-tinuously being made to the list of organic transfor-mations that can be performed efficiently in aqueoussolvent. Besides the Diels-Alder reaction, otherexamples include Claisen-rearrangements,6,7 aldolreactions,8,9 allylation reactions,10-12 and oxidations13-15

and hydrogenations16,17 of alkenes, to mention a few.These types of reactions have been among the mostuseful to the synthetic chemist for many decades and,indeed, still are. Their usefulness is reflected by theeffort that has been devoted to the discovery ofmethods with which they can be executed catalyti-cally and stereoselectively. With a few notable excep-tions, however, the development of chiral ligands andcatalysts for asymmetric synthesis has been carriedout almost exclusively in organic media.

Is water, then, a viable alternative solvent? Overthe last 5-10 years, the concept of efficient andstereoselective synthesis in water has been solidifiedas the rates, yields, and selectivities observed formany reactions in water have begun to match, oreven surpass, those observed in organic solvents. Inview of recent achievements, many of which will behighlighted in this article, the answer to the abovequestion is “yes”. From a less synthetic standpoint,increased appreciation of organic reactions in watermay also contribute to our understanding of the basicmechanisms of life.

This review will focus on recent advances inperforming diastereo- and enantioselective reactionsin aqueous media. Examples of poorly selectivetransformations are only included if they add extraweight to a discussion or to illustrate an importantconcept. The reader will also be briefly introduced tofields of pertinence to any aspect of organic chemistryin water, such as means of solubilizing organiccompounds and catalysis in water. Biocatalysis inaqueous media is beyond the scope of this article.Simple hydrolysis and phase-transfer reactions havealso been arbitrarily excluded. Several review articleson organic synthesis in water have appeared duringthe past decade, and they are recommended toreaders seeking a broader introduction to the topic.18-21

II. Water as Solvent

A. Why Water?Until recently, the use of water as solvent for

organic reactions was mainly restricted to simplehydrolysis reactions. Accordingly, most reagents andcatalysts in organic synthesis have been imperiouslydeveloped for use in anhydrous, organic reactionmedia. Why should we now spend time “rediscover-ing” reactions for use in water that already work wellin familiar organic solvents such as THF, toluene,or methylene chloride? Because there are manypotential advantages of replacing these and other

unnatural solvents with water. The most obvious arethe following. (1) Cost. It does not get any cheaperthan water! (2) Safety. Most of the organic solventsused in the lab today are associated with risks:Flammables, explosives, carcinogenics, etc. (3) En-vironmental concerns. The chemical industry is amajor contributor to environmental pollution. Withincreasing regulatory pressure focusing on organicsolvents, the development of nonhazardous alterna-tives is of great importance.

It is, however, important that the above listedbenefits are not gained at the expense of syntheticefficiency. Even a small decrease in yield, catalystturnover, or selectivity of a reaction can lead to asubstantial increase in cost and amount of wastegenerated. Fortunately, many theoretical and practi-cal advantages of the use of water as solvent fororganic synthesis do exist. These will be elaboratedupon below but are briefly introduced here. First,experimental procedures may be simplified sinceisolation of organic products and recycling of water-soluble catalysts and other reagents could be madeby simple phase separation. Second, laborious pro-tecting-group strategies for functionalities containingacidic hydrogens may be reduced. Third, water-soluble compounds could be used in their “native”form without the need for hydrophobic derivatization,again eliminating tedious protection-deprotectionsteps from the synthetic route. Fourth, as will beamply exemplified in this review, the unique solvat-ing properties of water have been shown to havebeneficial effects on many types of organic reactionsin terms of both rate and selectivity.

B. Solubility of Organic Compounds in WaterMost chemical reactions are performed in solvents.

The solvent provides a reaction medium in whichreactants can be mixed over a very wide concentra-tion range. In general, a good solvent should readilydissolve all or most of the participating reactants,should not interact adversely with the reaction, andshould be easily separated during workup for facileisolation of products. On the basis of his/her knowl-edge of the chemical properties of the reactants, thechemist chooses a solvent which is suitable to meetthese criteria. From this perspective, it is not sur-prising that water has found limited use as solventfor organic reactions. In truth, the poor solubility ofreactants and the deleterious effect on many organictransformations are the main obstacles to the use ofwater as reaction solvent. Nonetheless, the fact thatmany of the most desirable target molecules, e.g.,carbohydrates, peptides, nucleotides, and their syn-thetic analogues, as well as many alkaloids andimportant drugs are readily soluble in water isinconsistent with the disproportionate bias towardthe use of organic solvents for their preparation. Itcan be argued that our shortcomings as syntheticchemists prompt the use of exhaustive protecting-group strategies, thus limiting the possibility of usingwater as solvent because of low solubility of thereactants. Moreover, with the limited arsenal oforganic transformations in water that is presentlyavailable to the synthetic chemist, intermediates

2752 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 3: B. Solubility of Organic Compounds in Water

soluble in organic solvents are preferred to thosesoluble in water, even if it means adding extrasynthetic steps for derivatization. This may haveparticular relevance in carbohydrate chemistry.

Notwithstanding the above, many organic targetsand their intermediates have very low solubility inwater, which may lead to thwarting of reactions dueto phase separation and inefficient mixing of reac-tants, although heterogeneous mixtures may retainat least partly the positive influence of water, some-times with the aid of sonication or microwave heat-ing. A variety of strategies have been investigatedin order to expand the scope of water-based organicsynthesis to embrace also highly hydrophobic reac-tants, and these will be briefly discussed below. Fora more comprehensive treatise, a thorough review ofsolubility and solubilization in aqueous media wasrecently published by Yalkowsky.22

1. Organic Cosolvents

One of the more efficient and versatile methods ofincreasing solubility and one that does not requiremodification of the solute is to use an organic cosol-vent. The cosolvent reduces the hydrogen-bond den-sity of aqueous systems, so that it is less effective insqueezing out nonpolar solutes from solution. Cosol-vents can be structurally diverse, but they all carryhydrogen-bond donor and/or acceptor groups foraqueous solubility and a small hydrocarbon regionthat serves to disrupt the strong hydrogen-bondnetwork of pure water, thereby increasing the solu-bility of nonpolar reactants.22 Some of the mostcommonly used cosolvents are the lower alcohols,DMF, acetone, and acetonitrile. The increase insolubility, however, comes at the cost of many of theproperties that make water a unique solvent forsynthesis such as high polarity, high cohesive energydensity, and the hydrophobic effect. The significanceof this erosion of the bulk properties of water dependson the nature of the chemical reaction. Reactions thatinvolve charged or highly polar species will suffermore by a decrease in solvent polarity than reactionsinvolving only uncharged species. Likewise, reactionswith a negative activation volume (Diels-Alder,Claisen rearrangement) are expected to be adverselyaffected by the addition of cosolvent because of theconcurrent decrease in cohesive energy density.Nevertheless, because of the efficiency and flexibilityof cosolvents in solubilizing organic solutes, the majorpart of the development of reactions in aqueousmedia has been made with water-cosolvent mix-tures.

2. Ionic Derivatization (pH control)

Adding a positive or negative charge to an ionizablesolute generally brings about a substantial increasein its solubility in water. Adjustment of solution pHis therefore an efficient method of solubilizing weakelectrolytes in aqueous media.23 This approach, ofcourse, changes the chemical nature of the reactantand may limit its use as a method of solubilizationfor synthetic purposes. For some types of reactions,however, the presence of a charged, highly polarmoiety can have a very positive effect. For example,

the reactions of diene-carboxylates,24 -sulfonates,25

or -ammonium26 salts with dienophiles in aqueousDiels-Alder reactions display significantly higherreaction rates over the corresponding neutral dienes.It is generally accepted that the rate enhancementof the Diels-Alder reaction in water is at least partlydue to the influence of the hydrophobic effect in thismedia (see section V.A.1). In making the dieneamphiphilic, increased solubility comes with theadded bonus of an enhanced hydrophobic effect andfaster reaction. This is principally the same effect onewould achieve by adding “salting-out” agents knownto increase the hydrophobic effect,27 only in this case,the salt is the reactant itself. When a bufferedreaction is feasible it is one of the more efficient waysto keep organic molecules solubilized. A potentialpractical advantage of using pH control in organicreactions is that the product may be recovered fromsolution by precipitation upon suitable adjustmentof pH or by extraction after addition of appropriatephase-transfer counterions.

3. Surfactants

An intriguing means of achieving aqueous solubil-ity is by using surfactants. These are amphiphilicmolecules, that is, they contain one distinctly polarand one distinctly nonpolar region. In water, surfac-tants tend to orient themselves so that they minimizecontacts between the nonpolar region and the polarwater molecules and when the concentration ofsurfactant monomer exceeds a certain critical value(critical micelle concentration, CMC), micellizationoccurs. Micelles are spherical arrangements of sur-factant monomers with a highly hydrophobic interiorand a polar, water-exposed surface. Organic solutesinteract with micelles according to their polarity;nonpolar solutes are buried in the interior of themicelle, moderately polar molecules locate them-selves closer to the polar surface, while distinctlypolar solutes will be found at the surface of themicelle. This compartmentalization of solutes isbelieved to be responsible for the observed catalyticor inhibitory influence on organic reactions in micel-lar media.28 Micellar catalysis of organic reactionswill be further discussed in section III.C.

4. Hydrophilic Auxiliaries

Another method to increase the solubility of aque-ous organic reactions is by grafting hydrophilicgroups onto insoluble reactants. This strategy hasbeen only cautiously explored for synthetic purposesbut has a pivotal role in medicinal chemistry andmodern drug design because of the low water solubil-ity of many drugs, which causes limited bioavailabil-ity and thus reduced therapeutic efficacy. One wayof improving solubility of drugs is by converting theminto water-soluble prodrugs through covalent attach-ment of a hydrophilic auxiliary. Ideally, the attach-ment should be of a transient and reversible nature,allowing for release of the parent drug from theauxiliary upon distribution, either by enzymatic orchemical means. The size and nature of the solubi-lizing function ranges from small to medium-sized,acidic and basic ionizable moieties (e.g., carboxylic

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2753

Page 4: B. Solubility of Organic Compounds in Water

acids, amines) to large, nonionizable side chains (e.g.,poly(ethylene glycol) chains).29

Lubineau and co-workers demonstrated the useful-ness of carbohydrates as solubilizing auxiliaries inthe Diels-Alder reaction30-33 and the Claisen rear-rangement.34,35 A more recent example of a removablehydrophilic auxiliary that may be of synthetic poten-tial is the 2-pyridyldimethylsilyl group.36 In its pro-tonated form this group becomes miscible with water,and when attached to a hydrophobic side chain itinduces molecular aggregation, much like a surfac-tant, creating a local hydrophobic microenvironmentin which reactions can take place. The Diels-Alderreaction of 1 with p-benzoquinone in water containing0.2 equiv of HCl was complete within 1 h to give thecycloadduct 2 as a single stereoisomer in 81% yield(Scheme 1). The auxiliary may be removed afterreaction by oxidation with H2O2 to afford the corre-sponding alcohol.

III. Catalysis in Water

The development of novel catalysts is an importantaspect of modern stereoselective synthesis. Consider-able research has been focused on the developmentof chiral catalysts capable of selectively convertingprochiral starting materials into homochiral productswith great efficiency. A major driving force in thisarea has been the chemical industry and its continu-ous search for more cost- and performance-efficientprocesses. The chemical industry is, however, a majorcontributor to environmental pollution, largely dueto the ubiquitous use of hazardous solvents. Out ofthe top 10 chemicals released or disposed of by thechemical industry in the mid-1990s, five were sol-vents, namely, methanol, toluene, xylene, methylethyl ketone, and methylene chloride.37 With sharp-ened regulatory pressure focusing on organic sol-vents, the search for alternatives is of increasingimportance. In this respect, the development ofwater-tolerant catalysts and water-soluble ligandshas rapidly become an area of intense research.38-42

The field of chemical catalysis can be divided intotwo main categories: homogeneous catalysis, inwhich the catalyst is completely miscible with thesolvent, and heterogeneous catalysis, in which thecatalyst is a solid and catalysis occurs on a surface.There are clear advantages and disadvantages toboth approaches. Homogeneous catalysis is charac-

terized by superior activity and selectivity but usuallyinvolves a cumbersome process of separation ofcatalyst from reaction products and lowered activityof the recovered catalyst. On the other hand, hetero-geneous catalysis allows for easy separation of thecatalyst, but the reactions are often hampered by lowactivity and/or selectivity. Attempts aimed at het-erogenizing homogeneous catalysts through attach-ment to organic or inorganic supports have so far notbeen very fruitful for a variety of reasons, such asmetal leaching, poor catalyst efficacy, low reproduc-ibility of activity and selectivity, and degradation ofthe polymer support.40 A different method of com-bining high catalytic efficiency with easy recoveryand reuse of the catalytic species is to use water asa second phase in which the organic products of thereaction are poorly soluble.43 Employing water-soluble catalysts thus allows for isolation of organicproducts and quantitative recovery of the catalystthrough simple phase separation. Recently, water hasbeen shown to be a promising medium for heterog-enized homogeneous catalysis, averting some of theproblems mentioned above. This will be covered insection III.B.

A. Homogeneous Catalysis

1. Brønsted Catalysis

When feasible, the use of a Brønsted acid is one ofthe more convenient and environmentally benignmethods of catalyzing organic reactions in water. Apotential synthetic advantage of water over organicsolvents in Brønsted-catalyzed reactions is that thenucleophilicity of the corresponding base, which canlead to undesired side reactions, may be of lessconcern due to extensive solvation and diffusion ofcharge by hydrogen-bonding water molecules. Inorganic solvents such problems are usually avertedby using a sterically hindered acid/base. Many reac-tions developed for use in water have been shown tobenefit from the addition of catalytic amounts ofBrønsted acids.44-46

2. Lewis-Acid Catalysis

Lewis-acid catalysis is one of the most usefulmethods in modern stereoselective synthesis. Thecoordination and activation properties of Lewis acidsmake them suitable not only for intraligand asym-metric induction, in which the absolute configurationof the newly formed stereocenter is determined byan element of chirality within the same ligand (thesubstrate) on the metal, but also for interligandasymmetric induction, in which chiral induction isgoverned by a ligand on the metal other than thesubstrate.47 The former usually involves a chiralauxiliary, while the latter is accomplished throughthe attachment of chiral organic ligands to the centralmetal atom. Interligand asymmetric induction throughthe use of Lewis acids carrying chiral ligands is atthe core of asymmetric catalysis. Consequently, tofully extend the concepts of interligand asymmetricinduction and asymmetric catalysis to water-basedorganic reactions, it is imperative that metal specieswhich retain Lewis-acid activity, even in pure water,

Scheme 1a

a Reprinted with permission from ref 36. Copyright 2001 Wiley-VCH.

2754 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 5: B. Solubility of Organic Compounds in Water

are identified and that water-soluble ligands for thesemetals are developed.

(a) Metal Salts as Water-Tolerant Lewis Acids.Many of the most commonly used Lewis acids are ofa highly water-labile nature, and their use in organicsynthesis is restricted to reactions performed understrictly anhydrous conditions. Lewis acids operatethrough coordination with one or more Lewis-basicsites of a reactant, usually a nitrogen atom or oxygenatom, thus inducing polarization of the reactant byshifting electron density from the reactant toward thecatalyst. Herein lies the greatest obstacle to efficientLewis-acid catalysis in aqueous media; water, too,has a Lewis-basic site and easily hydrates most Lewisacids to preclude binding to the organic reactant,especially when present in large excess as a solvent.Lewis-acid-promoted reaction in water is, neverthe-less, far from inconceivable. For example, substrateswith more than one Lewis-basic coordination sitehave been demonstrated to more effectively competewith water for Lewis-acid complexation.48 Extensivestudies by Kobayashi and co-workers revealed thata wide range of lanthanide triflates,49 rare-earthmetal triflates,9 and some other metal salts can beuseful as Lewis-acid promoters in aqueous media.50

These water-tolerant Lewis-acid catalysts have beensuccessfully applied to various types of water-basedreactions (vide infra). On the basis of these ground-breaking observations, certain guidelines of whatconstitutes a good water-tolerant Lewis-acid catalysthave been formulated.50,51 First, the metal salt shouldhave an intermediate hydrolysis constant (Kh). Whenthe pKh value is too low, cations are easily hydrolyzedand oxonium ions are generated (Scheme 2). How-ever, when the pKh value is too high, the Lewisacidity is generally too weak for efficient catalysis.Second, the water exchange rate constant (WERC),which is a measure of the exchange rate for substitu-tion of inner-sphere water molecules, should be largeenough to allow for rapid exchange between watermolecules and basic sites on the reactant. In theMukaiyama aldol reaction between benzaldehyde andsilyl enol ether 3, the Lewis acids that were definedas active (>50% yield of adduct 4) had pKh values inthe range 4.3-10.08 and WERC values greater than3.2 × 106 M-1 s-1. The correlation found betweencatalytic activity and pKh and WERC allows for some

general qualitative assessment to be made of whichmetal compounds may serve as effective Lewis acidsin water. The lanthanide triflates, Ln(OTf)3, gaveconsistently good yields and seemed to be particularlysuitable to function as water-stable Lewis acids. Thiscan be explained by their low hydrolysis constants,pKh, and soft Lewis-acid character, which suggestsa strong affinity for carbonyl oxygens.50 In addition,lanthanide triflates are in fact more soluble in waterthan in organic solvents, which allows for simple andquantitative recovery of the catalyst. Other metalsthat were capable of catalyzing the aqueous Mu-kaiyama-aldol reaction and thus of potential aswater-tolerant Lewis acids include Fe(II), Cu(II),Pb(II), Zn(II), Cd(II), as well as the rare-earth metalsSc(III) and Y(III).

(b) Water-Soluble Chiral Ligands. The vastmajority of chiral catalysts applied in modern syn-thesis possess low water solubility. Several strategieshave been pursued to enhance solubility of metal-ligand complexes in water, three of which stand out:(i) perform the reaction in the presence of micelle-forming surfactants (see also sections II.B.2 andIII.C), (ii) add solubilizing functionalities, commonlyionic groups, to poorly soluble ligands, and (iii)employ easily accessible, highly water-soluble sourcesof chirality from nature, such as carbohydrates oramino acids as ligands. The de novo design of chiralligands specifically targeted for metal-assisted ca-talysis in water is an attractive but largely unex-plored avenue of research.52

The need to develop water-soluble catalysts hasbeen a concern addressed mainly by the chemicalindustry, likely because it is in large-scale processingthat the most obvious advantages of two-phasecatalysis exists. Hence, the range of reactions forwhich new catalysts for water-based synthesis havebeen prepared and applied has been narrow andlargely confined to some of the most importantcatalytic reactions in industry, such as hydrogenationof CdC and CdO bonds and hydroformylation, whererhodium and ruthenium are the most commonly usedmetals.42 Bisphosphine ligands are of versatile usein coordination chemistry and have been used withgreat success in many catalytic processes. Hence,building on the history of these ligands in metalcatalysis, by far the most investigated type of ligandsfor water-based catalysis are bisphosphines carryingone or more polar substituent, such as sulfonate,carboxylate, ammonium, phosphonate, or hydroxylgroups. These were reviewed by Papadogianakis andSheldon in 1997.40 Driven by early success and lackof better performing alternatives, notwithstandingconsiderable research leading to a large number ofnew ligands, the most commonly employed water-soluble ligands are the sulfonated bisphosphines.Synthesis, separation, and characterization of thesecompounds, however, are often difficult.

Achiral and chiral water-soluble ligands that havebeen prepared and used in the literature up to 1996have been summarized by Li and Chan.19 Water-soluble metal complexes for enantioselective hydro-genation of dehydroamino acids were reviewed byNagel and Albrecht in 1998.17 For a selection of more

Scheme 2

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2755

Page 6: B. Solubility of Organic Compounds in Water

recent literature on enantioselective hydrogenationsin water, see section IX of this review. Discussionherein will be exclusive to chiral ligands that have adocumented effect of ligand-induced stereoselectivecatalysis in water; water solubility alone does notjustify inclusion. Most recently, water-soluble bis-phosphine-metal complexes have been used withsome success in catalytic asymmetric reactions otherthan hydrogenations. Henry and co-workers devel-oped a palladium(II)-catalyzed, modified Wackeroxidation of olefins leading to the excess formationof chlorohydrins rather than aldehydes normallyobtained through this process.53,54 When the Pd(II)catalyst was coordinated to a chiral bisphosphineligand, (R)-Tol-BINAP, 5, good asymmetric inductionof up to 82% ee was achieved in a 1:2 mixture ofwater/THF (Scheme 3). The reactions could be per-formed in solvent of higher water content (water/THF4:1) with similar catalytic efficiency and selectivityby using the corresponding sulfonated ligand 6.Gratifyingly, in the preparation of the catalyst,sulfonation of (R)-Tol-BINAP did not lead to mixtureswith mono-, di-, and trisulfonated products otherwisecommonly encountered in aromatic sulfonation reac-tions, which further increases the utility of this newchiral ligand in aqueous asymmetric catalysis.

Aside from synthetic ligands, naturally occurringwater-soluble compounds have also been investigatedas catalytic ligands, albeit to a much lesser extent.For example, mono- and disaccharides have beenused as precursors for the construction of chiralwater-soluble ligands of bidentate coordination, usu-ally bisphosphines,55,56 but P,N-57 and N,N-deriva-tives58 have also been prepared. In general, ligandsbased on monosaccharides have only moderate solu-bility in water and often require incorporation ofextra hydrophilic groups on the ligand or addition ofa cosolvent or a surfactant for optimum efficiency.Disaccharide ligands have the advantage of higheraqueous solubility, and very high enantioselectivitieshave indeed been observed in reactions of these inpure water. For recent examples of the use of chiralligands derived from sugars, see sections III.C, V.E,VII, and IX.

The use of unprotected amino acids as chiralelements in metal-catalyzed reactions is rare. Thismay be due to solubility problems in organic solventstraditionally used for such transformations. In water,however, this is not a concern and “naked” aminoacids should be of great potential as asymmetricligands in aqueous media. This concept was pursuedby Micskei and co-workers, who used chromium(II)-

amino acid complexes in the enantioselective reduc-tion of acetophenone in water/DMF 1:1 to givemethylphenylcarbinols (R)-7 and (S)-7 in up to 94%yield and 75% enantiomeric excess (Scheme 4).59 Theauthors found that the enantioselectivities observedwere highly dependent on the structure of the ligandsand the stoichiometry of the metal-ligand complex8 (Figure 1), where controlled conditions leading toCrL2 complexes gave the best ee’s. Engberts and co-workers employed amino acids as ligands in the firstcatalytic, enantioselective Diels-Alder reaction inwater (section V.A.1).60

B. Heterogeneous CatalysisIn terms of catalytic efficiency, heterogeneous

catalysis is generally inferior to homogeneous cataly-sis.40 The advantage of solid or immobilized catalystsis mainly due to practical and economic reasons asthey are more easily recovered from reaction mix-tures and can be reused many times without loss ofcatalytic activity. It has long been recognized thatan ideal catalytic system would be to use highlyactive and selective homogeneous catalysts that havebeen ‘heterogenized’ for simple recovery and reuse.Even so, as mentioned above, progress toward acommercially viable process has been slow. In thepast few years, however, catalysts heterogenizedthrough covalent attachment to organic or inorganicsupports such as polymers,61 silica,62,63 and layeredclays64 have been used with some success in aqueousmedia.65 So far, the most promising approach seemsto be through the use of organic polymers. Uozumiand co-workers achieved high enantioselectivities inPd-catalyzed allylic alkylations (detailed in sectionVII) by attaching a chiral phosphine ligand to anamphiphilic polystyrene-poly(ethylene glycol) co-polymer resin (Figure 2, 9).66 The described protocolapproaches the realization of what may be considered

Scheme 3 Scheme 4a

a Reprinted with permission from ref 59. Copyright 1999Elsevier Science.

Figure 1. Reprinted with permission from ref 59. Copy-right 1999 Elsevier Science.

Figure 2.

2756 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 7: B. Solubility of Organic Compounds in Water

an ideal catalytic asymmetric reaction: (i) chemicaland optical yields ranged from high to excellent; (ii)reactions were performed in water without the useof any cosolvent; (iii) the catalyst was quantitativelyrecovered through simple filtration; and (iv) thecatalyst could be reused many times without anyobservable loss of activity.

C. Micellar CatalysisThe concept of micellar catalysis goes back to the

late 1960s.67 However, the recent discovery in 199268

that rhodium-catalyzed asymmetric hydrogenationsin water proceeded with higher enantioselectivitywhen micelle-forming surfactants were present in thereaction mixture has led to a newfound interest inthe use of aqueous micelles to promote organicreactions. Since then, this strategy has been appliedto enhance the performance of several other typesof water-based reactions, such as Diels-Alderreactions,69-72 aldol reactions,73,74 and allylation reac-tions.45 The addition of surfactant often leads to asignificant increase in rate and/or selectivity and,sometimes, exceptional results are obtained. In onerecent example, the enantiomeric excess obtained inan asymmetric hydrogenation reaction increasedfrom 1.5% to 77%!75 Recent mechanistic investiga-tions have provided some insight into the underlyingprocesses of this phenomenon.76 Although some gen-eral explanations can be invoked, such as solubili-zation and concentration of reactants, the detailedmode of action in any given system depends on thenature of the micelles and the substrates and thetype of reaction. For example, it was observed thatmicelles were unable to catalyze Diels-Alder reac-tions,77,78 actually often retarding them.79 In thepresence of Lewis acid, however, surfactant-aidedrate acceleration becomes significant.70,71 The originof the increase in selectivity of many reactions,however, remains elusive despite extensive researchefforts. In studies of rhodium(II)-catalyzed asym-metric hydrogenations of dehydroamino acid deriva-tives, enantioselectivities were shown to vary withthe size of the cyclic chelate of the metal-bisphos-phine catalyst (Scheme 5).80 It was found that theobserved ratios of the enantiomeric products (er) were

almost unaffected by the addition of surfactant (10%sodium dodecyl sulfate, SDS) in reactions catalyzedby 10 and 11, which forms rigid five- and six-membered cyclic chelates, respectively, while cata-lysts 12 and 13 that involve more flexible seven-membered chelates catalyzed the reactions to give aproduct distribution of much higher enantiomericratio in the presence of SDS than in pure water. Itwas argued that the nature of the surroundingsolvent influences the conformational equilibria of thecatalytic species, which may extend to effect also thecatalyst-substrate complexes and, hence, the enan-tiomeric product distribution. Such an effect wouldlogically be less operative on a small ring, in whichconformational freedom is more limited.

Kobayashi and co-workers studied Lewis acids thatwork in aqueous media (vide supra). To avoid the useof cosolvents, which are often required for bestefficiency, the effect of anionic surfactants such asSDS on Lewis-acid-catalyzed aldol reactions81 andallylation reactions82 in water was investigated.These reactions, too, benefited greatly from theaddition of surfactants. Taking this concept one stepfurther, Kobayashi and co-workers developed a newtype of Lewis acids in which the active metal cationcarries long anionic hydrocarbon sulfate or sulfonateligands (Figure 3) that makes them form micellaraggregates in water. These so-called Lewis-acid-surfactant combined catalysts (LASCs) have beensuccessfully employed in aqueous Diels-Alder reac-tions, aldol reactions, Mannich-type reactions, andallylation reactions.83 Micellar catalysis of organicreactions in water has been reviewed extensively.28,84,85

IV. Stereoselectivity in WaterWith a few notable exceptions, the majority of

asymmetric reactions are performed in apolar andaprotic media, which precludes the use of water-soluble compounds. It is imperative that water, too,is fully explored as a reaction medium for asymmetricsynthesis. Provided that substrates and reagents canbe used that do not react with water, how will theselectivity of reactions in inert solvents be affectedby the progression to a participating solvent such aswater? Learning how to take advantage of theuniquely complex solvating properties of water maylead to new concepts and possibilities in asymmetricsynthesis. Initial work in this area have indeed ledto some interesting and sometimes surprising results,which will be discussed here.

Scheme 5a

a Reprinted with permission from ref 80. Copyright 2001 Wiley-VCH.

Figure 3.

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2757

Page 8: B. Solubility of Organic Compounds in Water

A. Stereocontrol in WaterIn stereoselective synthesis, efficiency may be

defined as to have knowledge and control of thefactors that influence the spatial arrangement of thereactants as a reaction proceeds. These factors areseemingly easier to recognize and manipulate whensolvent-reactant interactions are so insignificantthat they can be excluded from the mechanisticpicture. Such is often the case with aprotic solventsof low polarity. Water, however, is a participatingsolvent, exerted through hydrophobic and/or electro-static interactions, and as such must be consideredin any model attempting to explain or predict theoutcome of a stereodifferentiating reaction. A mostfundamental question may be to what extent we canuse established rules, such as the Cram or Felkin-Anh models, which are based on observations instrictly anhydrous media, to account for selectivitiesobserved in water?

1. Chelation Control

One of the main concerns of synthesis in water maybe whether stereoselection through chelation controlwould still be operative. Recent experimental datasuggests, perhaps contrary to what was expected,that chelation effects may indeed be relevant also inwater. The most extensive studies in later years onchelation effects in organometallic addition reactionsin water have been pursued by Paquette and co-workers.86-90 Specifically, they investigated chelationcontrol in indium-assisted allylations of R- and â-het-eroatom-substituted aldehydes and ketones in or-ganic, aqueous, and mixed solvent systems. Somerepresentative examples are collected in Table 1. Themain lesson of these studies was that water does notinhibit chelation control in indium-mediated reac-tions, and selectivities in aqueous media were sig-nificantly greater than those attained in anhydrousmedia using the corresponding reagents of othermetals such as magnesium, cerium, and chromium.91

For the acyclic R-heterosubstituted substrates, itwas presumed that the predominantly observed syn-selectivity was the result of chelation controlledaddition according to Cram’s chelate model (Figure4, A). When the anti-products were observed, anonchelate, sterically controlled Felkin-Anh transi-tion state was more likely to have been involved(Figure 4, B). Additional evidence for the involvementof chelation control was provided by comparison ofreaction rates with a similar substrate known not tobe involved in chelation. Previous studies have shownthat when chelated intermediates are involved in areaction, there will be an increase in the reaction ratedue to preformation of the transition-state com-plex.92,93 Hence, for the Felkin-Anh (nonchelation)products, no rate acceleration was observed relativeto a nonchelating standard. In the allylation reactionof dimethylamino aldehyde 19, excellent syn-selectiv-ity of >99% is observed (Table 1, entry 12). However,substituting methyl for benzyl, as in 18, leads to areversal of diastereofacial preference to favor theanti-isomer in a ratio of 3.3:1 (entry 11) in accordancewith the Felkin-Anh model. Allylation of R-hydroxyaldehyde 16 leads to the 1,2-diol of significant syn-

predominance (entry 7). Again, heteroatom substitu-tion leads to erosion of syn-selectivity (14 and 15,entries 1 and 4), which suggests that protecting-group strategies may not only be unnecessary inmany aqueous reactions, but even have a pronouncednegative effect. It was also found that â-hydroxy-aldehyde 23 underwent addition by allylindiumreagents, in this case to give the 1,3-diol with a syn/anti ratio of 8.5:1 (entry 22). According to theproposed transition-state model (Figure 4, C), biden-tate coordination of the indium metal to the â-hy-droxyl group and the carbonyl oxygen directs additionto occur syn to the â-substituent, leading to the anti-product. In comparing R-oxy, R-amino, and R-thioaldehydes in the indium-mediated allylation reactionin water, it appears that amino derivatives are thebest at forming chelates with the metal. The allyla-tions of the R-thia substrates 20 and 21 gave productsof reversed selectivity (entries 13 and 16), i.e., favor-ing the anti-products, suggesting that the reactionis under nonchelation control, proceeding through theFelkin-Anh transition state. An important observa-tion in these reactions is that when chelation control

Table 1.

2758 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 9: B. Solubility of Organic Compounds in Water

appears to be involved, water as solvent is preferredin terms of selectivity over THF or a THF/watermixture.

2. Hydrophobic ControlBeyond chelation control, experimental data sug-

gests that the unique solvating properties of watercan sometimes be used to enhance selectivity. Forexample, hydrophobic interactions have been used toexplain improved selectivities of chiral recognitionevents on transition from organic to aqueous sol-vent.94-98 In the cycloaddition of 24 with acrolein,hydrophobic effects were invoked to account for theobserved enhancement of diastereoselectivity in wa-ter compared to toluene.96 One of the diastereotopicfaces of the reactive conformation appears to be morehydrophilic and thus more heavily hydrated (Figure5, A). Reaction therefore occurs preferentially fromthe opposite, more hydrophobic side. Also, as ex-pected from Diels-Alder reactions, the reaction inwater was 50 times faster than in toluene. In a

similar experiment, methacrolein reacted with thechiral diene carboxylate 25 in pure water to giveadducts of a 4.7:1 diastereomeric ratio.97 Assumingthat the transition state adopts the conformation ofthe most stable allylic rotamer as shown (Figure 5,B), the dienophile approaches the diene mainly fromthe side opposite the hydrophilic carboxylate group.Highly stereoselective indium-promoted allylationreactions in aqueous media under apparent hydro-phobic, nonchelation control have also been reported.In synthetic studies toward dysiherbaine, Loh andco-workers found that the allylindation of ketoester26 in THF/water 1:1 with various allylic bromidesproceeded to give 1,3-amino alcohols 27 in good yields(60-80%) and with excellent 1,3-induction (72-99%de, Scheme 6).94 It was argued that the high selec-tivities observed were based on a remote substituenteffect due to a conformational preference as shownin Scheme 6. That the selectivity was not due to anyfavorable π-π interactions between the aromatic ringand the carbonyl group was confirmed by substitut-ing the phenyl group for a cyclohexyl ring, a changethat did not lead to any significant difference indiastereoselectivities.

B. Water-Soluble Chiral AuxiliariesAs discussed above, attachment of a water-soluble

auxiliary is one way of solubilizing hydrophobiccompounds. If the auxiliary is of a chiral nature,asymmetric induction may be achieved. Lubineauand co-workers demonstrated this concept by usingsugars to simultaneously introduce water solubilityand an element of chirality.30-33 Diels-Alder reac-tions of dienes that were attached at the anomericposition of a carbohydrate proceeded much faster andwith higher endo/exo selectivity in water than thecorresponding reactions of dienes attached to a per-acetylated carbohydrate in toluene. A representativeexample is shown in Scheme 7. Although the asym-metric induction was low (20% de), separation of thecrystalline diastereomers and hydrolytic cleavage ofthe sugar moiety gave enantiomerically pure adducts.It is noteworthy that in these reactions diastereofa-cial preference was determined by the anomericconfiguration of the sugar. The use of carbohydratesas chiral, solubilizing auxiliaries has also been dem-onstrated to give modest asymmetric induction (20%de) in Claisen rearrangements in water.34,35

Figure 4.

Figure 5.

Scheme 6a

a Reprinted with permission from ref 94. Copyright 2000Elsevier Science.

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2759

Page 10: B. Solubility of Organic Compounds in Water

Amino acids and their derivatives are of interestas potential chiral auxiliaries in aqueous synthesis.Waldman and co-workers have shown that aminoacid esters can be used to induce chirality in aza-Diels-Alder reactions.99-101 For example, in water/THF 1:1, the condensation reaction between (S)-isoleucine methyl ester hydrochloride, 28, and form-aldehyde, followed by addition of cyclopentadiene,results in the azanorbornene products 29 and 30 ina diastereomeric ratio of 93:7 (Scheme 8). Theobserved stereoselectivity was explained by invokinga transition state as illustrated in Scheme 8, wherethe addition of diene is controlled by the stericinfluence of the amino acid side chain and secondaryorbital interactions with the diene.

V. Additions to C−C Multiple Bonds

A. Cycloadditions

Cycloaddition reactions are among the most im-portant and useful processes in organic chemistryinasmuch as they enable the construction of complexpolycyclic compounds, often with a high degree ofregio- and stereocontrol. A seminal discovery inwater-based organic synthesis was that the Diels-Alder reaction and other cycloaddition reactions oftenproceed faster and with higher endo/exo selectivityin water than in organic media. Consequently, cy-cloaddition reactions in water have been subjectedto more scrutiny than any other type of reaction inwater.70 In view of the fact that several reviews onDiels-Alder reactions in water have appeared inlater years,20,48,102-104 the discussion on this topicbelow will be more stringently limited to the mostsignificant and recent achievements.

1. Diels−Alder Reaction

Prior to 1980, this powerful transformation wasmainly performed in organic solvents while onlyincidental reports were made of the use of water assolvent. In pioneering studies by Breslow and Griecoon water-based Diels-Alder reactions, surprising andremarkable discoveries were made about the uniquesolvent effect of water on this reaction.1-5 Comparedto reactions carried out in organic solvent, the cor-responding reactions in water displayed significantrate acceleration. Substantial experimental77,78,105-108

and theoretical109-113 evidence indicates that theorigin of the water effect on the rate of Diels-Alderreactions is due to differential solvation of reactantsand transition states. This explanation is in contrastwith the generally accepted concept that concertedreactions, such as the Diels-Alder reaction, displayvery modest solvent effects, reflecting a small differ-ence in polarity between the initial state and thetransition state.114,115 Hence, the surprising discover-ies made with water as solvent have subsequentlybeen attributed not so much to the increase inpolarity116 but to hydrophobic interactions and hy-drogen bonding, although a complete qualitative andquantitative understanding of how these forces con-tribute individually and collectively is still not athand. Terms such as enhanced hydrogen bonding andenforced hydrophobic interactions have been intro-duced to emphasize that these interactions occurbecause they are an integral part of the activationprocess. Hydrogen bonding between water and theactivating group of the dienophile is likely to beresponsible for part of the rate enhancement, prob-ably similar to the way a Lewis acid would oper-ate.77,78,107,112 The suggested positive influence of thehydrophobic effect is presumed to originate in thenegative activation volume of the Diels-Alder reac-tion, which leads to a compacting of the reactants inthe transition state compared to the initial state, thusremoving water molecules from the hydrophobichydration shells of the reactants as the reactionproceeds.105,108,117,118

Alternatively, the cohesive energy density (CED)is a measure of the energy required to create a cavityin a solvent and as such relates to all the inter-molecular forces acting within a solvent. The rateconstants of several Diels-Alder reactions haveindeed been correlated with the CED of the sol-vent.114,119 It was also found that the establishedpreference for the formation of endo-cycloadducts wasenhanced in water.2,3 Again, the hydrophobic effectis presumed to be intrinsically involved by favoringthe compact endo transition state more than theextended exo transition state. Hydrogen bonding isalso believed to play an important part, once moredrawing on the analogy with Lewis-acid catalysis,which is known to have a dramatic effect on endo/exo selectivity.120,121 In addition, positive effects onregio- and diastereoselectivity when switchingfrom organic to aqueous solvent have been re-ported.96,116,122,123

In an attempt to combine the beneficial effects ofwater and Lewis-acid catalysis on the Diels-Alderreaction, Engberts and co-workers investigated Di-

Scheme 7

Scheme 8

2760 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 11: B. Solubility of Organic Compounds in Water

els-Alder reactions catalyzed by Lewis acids inwater.60,124,125 Four divalent metal cations (Co2+, Ni2+,Cu2+, Zn2+) were screened for their effect on the rateand endo/exo selectivity of the reaction between thebidentate dienophile 31 and cyclopentadiene (Scheme9). All four metals studied successfully catalyzed thecycloaddition, and the catalytic efficiency of the Lewisacids was found to be of the order Co2+ < Ni2+ < Cu2+

. Zn2+. However, the rate-enhancing effect of wateron the catalyzed reaction was much less pronouncedthan the corresponding effect on the uncatalyzedreaction. Also, the increase of the endo/exo selectivityobserved for the uncatalyzed reaction in water iscompletely thwarted for the catalyzed reaction.

These observations imply that the positive influ-ence of hydrogen-bonding and hydrophobic inter-actions on Diels-Alder reactions in water is attenu-ated in the presence of Lewis acids. Hence, theconclusion was that the beneficial effects of water andLewis-acid catalysis in these reactions are not addi-tive. Nevertheless, the Cu2+-catalyzed reaction be-tween 31 and cyclopentadiene was 79 300 timesfaster than the same uncatalyzed reaction in aceto-nitrile. Having established that Lewis-acid-catalyzedDiels-Alder reactions in water proceed efficiently,focus was turned to examining possible ligand effectsin these reactions. For this purpose, a series of chiralR-amino acid ligands was screened for rate andenantioselectivity in the Diels-Alder reaction be-tween 31 and cyclopentadiene (Scheme 9).60,124 It wasfound that R-amino acids carrying an aromatic sub-stituent caused a significant increase in the Ka of theligand-Cu2+-dienophile complex (Table 2). The au-thors speculate that the increase in Ka is due toattractive arene-arene interactions between dieno-phile and ligand. Additionally, in the projected

transition-state complex (Figure 6), one face of thedienophile is effectively shielded from attack and,upon reaction with cyclopentadiene, addition product32 was indeed obtained in enantiomeric excess (Table2). This study also demonstrated that enantioselec-tivity is enhanced in water as compared to organicsolvents. It may be argued that any favorable inter-action operating between two arenes is strengthenedin water due to the hydrophobic effect. Support forthe crucial participation of a stabilizing arene-areneinteraction came from the fact that only insignificantenantioselectivities were obtained with nonaromaticamino acids. A strongly nonlinear dependence of eeon the ligand/catalyst ratio suggested that ligand-accelerated catalysis was taking place. Even when50% of the Cu(II) catalyst was present in its achiralhydrated form, most of the reaction was still medi-ated by the chiral Cu(II)-ligand complex to give highee. These results were the first examples of enantio-selective Lewis-acid-catalyzed Diels-Alder reactionsin water. Chiral induction in uncatalyzed Diels-Alder reactions in water has, however, been achievedthrough the use of chiral auxiliaries attached eitherto the dienophile126 or to the diene.31 The latterapproach was exemplified in section IV.2.B.

2. Hetero Diels−Alder Reaction

The hetero Diels-Alder cycloaddition is a poten-tially very useful reaction for the construction ofheterocycles of defined relative and absolute config-uration. However, the use of hetero Diels-Alderreactions in organic synthesis has been of limitedscope because the normally unreactive dienophilegenerally requires some external activation. This canbe accomplished through the use of electron-with-drawing substituents on the dienophile, throughLewis-acid catalysis, or by using very reactive dienes.Alternatively, in aqueous media, these reactions canproceed under milder conditions, with simple proto-nation of the dienophile being the most convenientapproach.127 The potential of using reactions betweenprotonated imines and dienes in water as an efficiententry to various heterocycles and alkaloids wasestablished in the mid-1980s through work by Griecoand Larsen.128 For example, the condensation ofiminium ion 33 with (E,E)-2,4-hexadiene in waterafforded the tetrahydropyridine derivative 34 as asingle diastereomer (Scheme 10, A). In an intra-molecular variant, the dienyl aldehyde 35 reactedwith benzylamine hydrochloride in ethanol/water 1:1at 70 °C to give the reduced quinoline derivatives 36and 37 in 63% combined yield and a 2.5:1 diastere-omeric ratio (B). Also, in a more recent study, Griecoand Kaufman reacted diene 38 in hot water to givethe tricyclic amine 39 as the exclusive diastereomer

Scheme 9

Table 2.

ligand Ka (M-1) k2 (M-1 s-1) ee (%)

water 1.2 × 10-3 2.56 0glycine 7.4 × 10-4 1.89 0L-valine 5.7 × 10-4 1.90 3L-leucine 5.1 × 10-4 2.01 3L-phenylalanine 8.7 × 10-4 2.01 14L-tyrosine 1.4 × 10-3 1.68 26L-tryptophan 3.0 × 10-3 1.44 25L-abrine 5.0 × 10-3 1.47 74

Figure 6.

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2761

Page 12: B. Solubility of Organic Compounds in Water

in 80% yield (C).129 According to Grieco, water ap-pears to be the polar solvent of choice for intra-molecular imino-Diels-Alder reactions as described.Lanthanide Lewis acids can also be used as catalystsin the aqueous imino-Diels-Alder reaction, as dem-onstrated by Wang and co-workers in their synthesesof azasugars and their analogues, compounds whichare of interest as potential glycosidase inhibitors.130

High asymmetric induction has recently beenachieved in hetero Diels-Alder reactions of inverseelectron demand. Fringuelli and co-workers per-formed [4+2] cycloadditions of nitroalkenes withvinyl ethers to provide synthetically useful cyclic six-membered nitronates (Scheme 11).131 The reaction ofnitroalkene 40 with enantiopure vinyl ether 41 inpure water proceeded with total regio- and diaste-reoselectivity to afford cis-substituted nitronate 42in good yield. The strong preference for cis-productswas rationalized by secondary orbital interactionsbetween the electron-rich oxygen on the vinyl etherand the positively charged nitrogen on the nitro-alkene. In related work, conjugated 3-diazenylbut-2-enes were found to be good substrates in theinverse electron demand hetero Diels-Alder reactionin water, leading to tetrahydropyridazines with highendo/exo selectivity.132

B. Michael-Type AdditionMany variations of this important reaction for the

formation of new carbon-carbon and carbon-heterobonds have been reported to work well in water/aqueous media.133-140 Both basic catalysis and Lewis-

acid catalysis have been applied, and nucleophilesrange from neutral amines, alcohols, and thiols toanionic species including carbanions. It is thereforesomewhat surprising that little effort has been madeto perform stereoselective conjugate additions inaqueous media. One study, however, found thatmoderate diastereoselectivities were achieved in theMichael addition of thiophenol to nitro olefins in basicmedia (Scheme 12).133 The reaction of (E)-1-methyl-1-nitrostyrene 43a with thiophenol in 0.5 M aqueousNaHCO3 gave the syn- and anti-addition products44a and 45a with 73:27 selectivity in a combinedyield of 95%. The analogous reaction with (E)-2-nitro-2-pentene, 43b, gave even higher anti-selectivity(44b/45b 80:20), but when the slightly bulkier iso-propyl substituent was present, as in 43c, theselectivity was lower and of reversed nature (44c/45c 38:62). The best result was obtained with thecyclic substrate 46, which gave exclusively the cis-adduct 47 in 70% yield upon reaction with thiophe-nol. The selectivities observed in these additions wereexplained by a conformational preference of theintermediate nitronate anion to orient the largersubstituent perpendicular to the double bond. Pro-tonation will thus preferentially occur from theopposite, less hindered side.

C. Dihydroxylation

The single most common and powerful method ofstereoselectively oxidizing unactivated olefins is theasymmetric dihydroxylation (AD) reaction developedby Sharpless and co-workers.14,141 In contrast withthe Sharpless asymmetric epoxidation of allylic al-cohols, water-free conditions are not required. In fact,the best medium for the catalytic AD is usually awater/tert-butyl alcohol mixture as it suppresses theundesired second catalytic cycle, thus preventing theformation of racemic products by keeping osmiumtetroxide as the only oxidant in the organic phase.The most interesting modification of this reaction inrecent years may be the immobilization of osmiumtetroxide to allow for recovery of the highly toxic andexpensive catalyst, which cannot be achieved soeasily in solution. Using a special microencapsulationtechnique, Kobayashi and co-workers attached os-mium tetroxide to phenoxyethoxymethyl-polysty-rene (PEM) polymers, which could then be used to

Scheme 10

Scheme 11

Scheme 12a

a Reprinted with permission from ref 132. Copyright 2001Brazilian Chemical Society.

2762 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 13: B. Solubility of Organic Compounds in Water

execute AD of olefins in water/acetone (1:1) withcomplete recovery of catalyst and retention of cata-lytic activity.142 In most cases, the diols were obtainedin good yields with high enantiomeric excesses. Thestereoselective dihydroxylation of olefins leading toracemic anti-1,2-diols can be achieved in aqueoussolution using hydrogen peroxide in the presence oftungstic oxide.143

D. AminohydroxylationThe direct conversion of achiral alkenes into the

synthetically useful 1,2-amino alcohols of high enan-tiomeric purity can effectively be achieved via asym-metric aminohydroxylation (AA).144-146 Similar to theAD reaction, AA can be performed with catalyticamounts of osmium tetroxide in aqueous media (e.g.,water/tert-butyl alcohol or water/acetonitrile). Re-cently, Sharpless and Fokin found that R,â-unsatur-ated acids were excellent substrates for the catalyticaminohydroxylation, leading to the exclusive forma-tion of syn-1,2-amino alcohols in near quantitativeyields (Scheme 13).13 Using the sodium salts of theacids, the reactions could be performed in highconcentrations in water without any organic cosol-vent. When unsymmetrically substituted alkeneswere employed, the reactions afforded exclusively theregioisomer with nitrogen on the less substitutedcarbon. However, the lack of ligand acceleration, evenin the presence of a large excess of ligand, led toracemic products only.

E. EpoxidationEnantiomerically pure epoxides are very useful

synthetic intermediates. The most powerful meansby which these are obtained are the Sharplessepoxidation and the Jacobsen epoxidation. Both ofthese reactions, however, demand strictly anhydrousconditions for optimum performance. Recently, acatalytic asymmetric version of the highly stereospe-cific, dioxirane-mediated epoxidation of alkenes inaqueous media was developed. Yang and co-workersobtained good to high enantioselectivity (71-95% ee)in the epoxidation of disubstituted and trisubstitutedolefins using BINAP-derived ketones (R)-48a-c ascatalysts and Oxone as the stoichiometric oxidant(Scheme 14).147 Almost simultaneously, Shi and co-workers found that the fructose-derived chiral ketone49 was an efficient catalyst in the epoxidation ofconjugated dienes leading to synthetically usefulvinyl epoxides of high enantiomeric excess (89-97%).148 A few representative examples are shown inScheme 15. The reactions were generally carried out

in water/dimethoxymethane/acetonitrile 2:2:1, and inthis system even highly hydrophobic dienes weregood substrates. As the authors pointed out, theobserved regioselective epoxidation of the alkenedistal to the electron-withdrawing group nicely comple-ments the Sharpless epoxidation of conjugated dienes,which preferably oxidizes the olefin proximal to thehydroxyl functionality. It also complements the chiral(salen)Mn-catalyzed epoxidation of conjugated dienes,in which cis-olefins are preferentially epoxidized.Additionally, the catalyst 49 was found to be efficientin the asymmetric epoxidation of a wider range ofsubstrates, including enynes, allylic alcohols, ho-moallylic alcohols, and bishomoallylic alcohols.15,149

F. Cyclopropanation

The ruthenium-catalyzed asymmetric cyclopropa-nation of styrene with menthyl diazoacetates inaqueous media was recently disclosed by Nishiyamaand co-workers.150 The water-soluble, chiral ligandbis(hydroxymethyl-dihydroxyoxazolyl)pyridine 50(Scheme 16) was used in these transformations.While the reaction proceeded with low enantioselec-tivity (8% ee) in pure THF or toluene, the additionof water dramatically increased the selectivity, pro-ducing cyclopropane 51 in 78% ee in THF/water 2:1and 94% ee in toluene/water 1:1.

Scheme 13a

a Reprinted with permission from ref 13. Copyright 2001 Wiley-VCH.

Scheme 14

Scheme 15

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2763

Page 14: B. Solubility of Organic Compounds in Water

VI. Additions to C−X Multiple BondsThe use of organometallic reagents to effect addi-

tion reactions to carbon-heteroatom multiple bondsis among the most important C-C bond-formingreaction available to the organic chemist. Organo-metallic additions have been particularly instrumen-tal in the development of asymmetric synthesis. Thepresence of a coordinating metal often leads to atransition state of higher structural order due tochelation of the metal to Lewis-basic atoms in thesubstrate. Attack of the nucleophile from the steri-cally less challenging side of the preorganized com-plex is then usually preferred. Hence, when chelatedintermediates are involved, mechanistic analysis isoften simplified. As discussed above and exemplifiedextensively below, when water-tolerant reagents areused, metal coordination chemistry and chelationeffects can be expected to operate in aqueous media.

A. Aldol-Type ReactionsStereoselective aldol and aldol-type reactions have

been carried out successfully in aqueous media.8,9

While the classical aldol addition commonly employsa basic catalyst in protic solvents, undesired byprod-ucts are frequently formed that reduce the syntheticvalue of the reaction. In this respect, Lewis-acid-catalyzed aldol-type reactions of silyl enol ethers withaldehydes and ketones have been found useful dueto the higher regio- and stereoselectivities that cangenerally be obtained.151 The rapid deactivation ofmany promoters and the decomposition of silyl enolethers in protic solvents, however, seemed to excludethe use of water as solvent. Pioneering experimentsby Lubineau and co-workers demonstrated that theuncatalyzed aldol reaction of silyl enol ethers in watercould proceed, though yields were poor.152 The de-velopment of Lewis acids tolerant of aqueous solventin the past decade (see also section III.A.2) hasallowed for successful catalytic, stereoselective aldol-type reactions in water.52,73,153-156

Early work on aldol reactions in aqueous mediafocused primarily on lanthanides as catalysts. In arecent example, various lanthanide triflates werecomplexed with the chiral crown ether 52 (Figure 7)and screened for their ability to induce diastereo- andenantioselectivity in the reaction between benzalde-hyde and silyl enol ether 3 in EtOH/water 9:1.52 Itwas shown that for larger lanthanide cations, suchas Ce(II), La(II), Pr(II), and Nd(II), which may have

a better size fit for the cyclic ligand than smaller ions,both diastereomeric (78-86% de) and enantiomeric(75-82% ee) selectivity were good (Table 3). A rangeof other metals have also been shown to executeaqueous aldol-type reactions with good stereoselec-tivity. Kobayashi and co-workers found that, inEtOH/water 9:1, lead(II)-triflate efficiently catalyzedthe aldol reaction of benzaldehyde and 3 upon com-plexation to the chiral 18-crown-6 ligand 53 (Figure7).153 Again, good syn/anti-selectivities (80-88% de)and enantioselectivities (75-87% ee) were obtained(Table 3).

Boron has been shown to be an efficient mediatorof stereoselective aldol reactions.73 In water/SDSmixtures, 10 mol % of diphenylborinic acid, Ph2BOH,catalyzed the reaction between aldehydes and silylenol ethers to give syn-substituted â-hydroxy ketonesin high diastereomeric excesses (80-94% de, Scheme17). While the reaction proceeded sluggishly withonly Ph2BOH, the addition of 0.01 equiv of benzoicacid dramatically improved the yields. In organicsolvents (Et2O, CH2Cl2) the reaction was almostcompletely thwarted. A mechanism was proposed inwhich a boron enolate generated by silicon-metalexchange was the reactive intermediate. The positiveeffect of benzoic acid on the reaction rate was

Scheme 16a

a Reprinted with permission from ref 149. Copyright 2001 RoyalChemical Society.

Figure 7.

Table 3.

catalyst R de (%) ee (%)

Ce(OTf)2-52 Ph 86 82Pr(OTf)2-52 Ph 80 79La(OTf)2-52 Ph 78 79Nd(OTf)2-52 Ph 86 75Pb(OTf)2-53 Ph 80 55Pb(OTf)2-53 n-hexyl 84 80Pb(OTf)2-53 n-nonyl 80 82Pb(OTf)2-53 isovaleryl 88 87

Scheme 17a

a Reprinted with permission from ref 13. Copyright 2001 Wiley-VCH.

2764 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 15: B. Solubility of Organic Compounds in Water

presumably due to acceleration of the Si-B exchange,which was thought to be the rate-determining step.The high syn-selectivities observed when (Z)-enolateswere used were rationalized by invoking a chairlikesix-membered cyclic transition state for the reactionbetween the aldehyde and the boron enolate.

Copper(II)-catalyzed aldol reactions have also beenreported recently. Indeed, the first catalytic asym-metric aldol reactions in aqueous media were per-formed with Cu(OTf)2 as catalyst and bisoxazolinesas chiral ligands.154,157 Other metals used with somesuccess as catalysts in aqueous aldol reactions arebismuth158 and indium.159 Interestingly, the predomi-nance of syn-aldol products in the water-based aldolreactions discussed above is in contrast with theanalogous reactions run under anhydrous conditionswhere the anti-isomer is usually the major product.

B. Allylation ReactionThe fairly recent, unexpected discovery that metal-

mediated allylations can be performed in water ledto a surge of interest in these reactions, and recentliterature is abundant with potentially useful ex-amples. Particular attention has been focused on theallylation of carbonyls and imines to give the corre-sponding homoallylic alcohols or amines. While anumber of metals, such as Zn,160-163 Sn,164-169 Sb,170

Co,171 Mn,172,173 Mg,174,175 or Hg,176 have been reportedto be useful for this purpose, indium (In) has showedthe most promise due to its unique properties thatmake it particularly suitable for use in water: (i) itis stable even in boiling water; (ii) it is resistant tooxidation by air; and (iii) it has an unusually low firstionization potential. The literature on indium-medi-ated reactions in aqueous media up to 1998 has beenrecently reviewed12 and will not be discussed here.More recent work on indium-mediated allylationsshow significant progress toward achieving highlevels of stereocontrol for the preparation of naturalproducts and synthetic intermediates. For example,Loh and Zhou described the first enantioselectiveindium-mediated allylation of aldehydes.11 Using(S,S)-2,6-bis(4-isopropyl-2-oxazolin-2-yl)pyridine, 54,as a chiral ligand and Ce(OTf)4 hydrate as Lewis-acid promoter, the reaction of allylindium with ben-zaldehyde in water/ethanol (1:1) afforded the homo-allylic alcohol (R)-55 in 90% yield and 92% ee(Scheme 18). The same authors later reported acatalytic enantioselective version of the same trans-formation using allyltributyltin in place of allyl-indium and a modified version of Yamamoto-Yanag-isawa’s catalyst (S)-BINAP‚AgNO3, 56.164 Initial at-

tempts using 5 mol % of catalyst 56 in various solventsystems resulted only in moderate yields and enan-tioselectivities, but the use of 5 mol % (S)-Tol-BINAP‚AgNO3, 57, in 1:9 water/ethanol transformed aro-matic aldehydes into (S)-allylic alcohols in nearquantitative yields and good selectivities (71-81% ee,Scheme 19). Much lower selectivities were observedwith aliphatic, olefinic, and acetylenic aldehydes.

Excellent asymmetric induction has been achievedby grafting a chiral auxiliary onto the substrate. Forexample, Cho and co-workers studied the indium-mediated allylation reactions of R-ketoimide 58 de-rived from Oppolzer’s sultam with various allylbromides and found that in water/THF 3:1, dia-stereoselectivities of >99% were obtained (Scheme20).177 The high diastereoselectivity was explained bytransition state 59 in which the carbonyl oxygen isinvolved in a six-membered ring chelation withindium, which may be further coordinated with oneof the oxygens of the sulfoxide. Allylation then occursfrom the least hindered side. The chelating propertiesof indium in water leading to enhanced stereoselec-tivity considerably increases the synthetic value ofthe indium-mediated allylation reaction. Canac andco-workers investigated this new methodology in aneffort to find an efficient entry to the syntheticallyuseful C-branched sugars.178 The reaction of 4-bromo-2-enopyranoside 60 with benzaldehyde in pure watercontaining powdered indium metal afforded the ad-dition product 61 as a unique stereoisomer in 95%yield (Scheme 21). Again, a six-membered cyclictransition state, 62, involving indium chelation wasinvoked to explain the observed selectivity. Withbutyraldehyde, however, lower yield (71%) and noselectivity was observed, suggesting that less reactive

Scheme 18a

a Reprinted with permission from ref 11. Copyright 1999Elsevier Science.

Scheme 19a

a Reprinted with permission from ref 163. Copyright 2000Elsevier Science.

Scheme 20a

a Reprinted with permission from ref 176. Copyright 2001Elsevier Science.

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2765

Page 16: B. Solubility of Organic Compounds in Water

and/or sterically less demanding aldehydes are poorersubstrates in this reaction. In the reaction between60 and the furanose aldehyde 63 to give the C-disaccharide 64, a complete reversal of the diaste-reoselectivity was observed compared to the reactionwith benzaldehyde (Scheme 22). Using the cyclictransition-state model, this outcome was explainedby additional chelation of the indium atom eitherwith the exocyclic C-3′ oxygen (Figure 8, 65) or withthe endocyclic oxygen (66), both favoring the sameproduct. Most noteworthy, in these conformations thechelating factor overrides the steric effect since thesugar substituent is in an unfavorable axial position,thus again indicating that significant chelation con-trol can be expected in indium-mediated allylationsin water. The effect of neighboring groups and solventin allylindation reactions has recently been studiedin detail (see section IV.A.1).10,87-89,91,94,179-181

C. Mannich ReactionThe development of Mannich-type reactions in

aqueous media has been made with the incentive offinding a milder and more convenient approachtoward the construction of â-amino ketones or esters.Classical protocols for Mannich reactions are some-times of limited synthetic potential as they ofteninvolve harsh reaction conditions and are plaguedwith severe side reactions as well as poor regio- andstereocontrol.182 Toward this end, several researchgroups have recently reported on the one-pot Man-nich-type reaction in water to give â-amino carbonylcompounds using either Lewis-acid83,183-185 or Brøn-sted catalysis,186-190 with or without the addition ofsurfactants. Both types of reactions generally proceedsmoothly in good yields, albeit with diastereoselec-tivities that are usually moderate at best. Good syn-

selectivities of up to 95:5 were obtained, however, inHBF4-catalyzed Mannich reactions between aldi-mines such as 67 and ketene silyl acetal 68 in awater/SDS mix (Scheme 23).186

D. Hydride AdditionThe stereoselective reduction of prochiral ketones

with NaBH4 as the hydride source to give secondaryalcohols has been accomplished in aqueous mediacontaining carbohydrate-based amphiphiles. The pres-ence of the glycosidic amphiphiles serves two pur-poses: they are solubilizing agents, probably throughformation of aggregates similar to micelles, and theyalso provide a stereodiscriminating environment forselective recognition of enantio- or diastereotopicfaces of the ketone. A range of ketones was selectivelyreduced through this approach by Plusquellec andco-workers.95 In their work, high regio- and stereo-selectivities were achieved in the reduction of cyclicR,â-unsaturated ketones using various types of sugar-based additives. In a more recent example, Rico-Lattes and co-workers took this concept one stepfurther by incorporating D-gluconolactone into poly-amidoamine (PAMAM) dendrimers.191 Supported bya solid core, these amphiphilic dendrimers acted asunimolecular micelles with a chiral surface. In animpressive demonstration of this strategy, acetophe-none was reduced by NaBH4 in water in 92% yieldand 98% ee.192

VII. Substitution Reactions

Little attention has been devoted to the develop-ment of substitution reactions in water. A likelyexplanation for this is that in many types of nucleo-philic reactions in water, hydrolysis of the electro-phile may compete with the desired reaction path-way. General hydrolysis reactions, in which thereaction between water and an electrophile is thepreferred outcome, will not be discussed here. Nev-ertheless, a few reports of palladium-catalyzed allylicsubstitutions in water have surfaced in recentyears.63,193-195 Uemura and co-workers used a carbo-hydrate-based phosphinite-oxazoline ligand 69 inthe palladium-catalyzed substitution of 1,3-diphenyl-3-acetoxyprop-1-ene with both carbon and nitrogennucleophiles.57 A representative example is shown inScheme 24. The substitution products were obtainedin moderate to high yields (66-95%) and in goodenantiomeric excess when performed in water alone(85% ee) or in water/acetonitrile mixes (77-84% ee),but the best result was obtained in acetonitrile alone(92% ee). A year later, Uozumi and Shibatomireported of an immobilized palladium complex of aP,N-chelate chiral ligand (9, Figure 2), which cata-

Scheme 21

Scheme 22

Figure 8.

Scheme 23a

a Reprinted with permission from ref 185. Copyright 2001Elsevier Science.

2766 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 17: B. Solubility of Organic Compounds in Water

lyzed the asymmetric alkylation of allylic esters in0.9 M aqueous Li2CO3 with up to 99% enantioselec-tivity (Scheme 25).66 The catalyst was found to beeffective for both cyclic and acyclic substrates. Con-veniently, the catalyst could be recovered by simplefiltration and reused without any loss of activity orstereoselectivity.

VIII. Rearrangements

A. Claisen Rearrangement

Water has a fundamental influence on the Claisenrearrangement.6,7 Experimental data have indicatedthat the Claisen rearrangement of allyl vinyl etheris 1000 times faster in water than in the gasphase.196,197 Because of the relative simplicity of thereaction, extensive theoretical studies have also beenperformed on the influence of various solvents, waterin particular.198 Both experimental and theoreticaldata have suggested that upon hydration, the Gibbsenergy of activation is reduced by between 3.5 and4.7 kcal/mol. Similar to the Diels-Alder reaction,differential solvation of the initial state and transi-tion state seems to be intrinsically involved in effect-ing the observed rate enhancement. However, incontrast to the Diels-Alder reaction, the explorationof aqueous Claisen rearrangements for syntheticpurposes has been minimal. The reason behind thisdiscrepancy is not clear, especially considering theimportance of the Claisen rearrangement in syntheticorganic chemistry. In the late 1980s, Grieco and co-workers investigated the accelerating effect of wateron Claisen rearrangements and also demonstratedits potential in organic synthesis on a number ofsubstrates.199 For example, rearrangement of theallyl vinyl ether 70 in basic water/methanol 2.5:1afforded the aldehyde 71 in 85% isolated yield,without the need for protection of the hydroxyl groups(Scheme 26). Rearrangement of the protected sub-strate in organic media was more difficult and led toelimination of acetaldehyde.

IX. HydrogenationsCatalytic asymmetric hydrogenation have been

among the most important reactions in the manu-facturing of chiral compounds from achiral startingmaterials ever since its inception in the early 1970s.Over the years, many hundreds of chiral catalystshave been prepared and screened against alkenes ofdiverse structural composition. Water-soluble chiralcatalysts have also been developed. The aqueous-phase asymmetric hydrogenations of olefins, ketones,imines, and R,â-unsaturated acids have been welldocumented in the literature.16 Only recently, how-ever, have reports appeared describing enantioselec-tivities that can match those achieved in the corre-sponding transformations in organic solvents. Aseminal observation was that higher ee’s are gener-ally obtained in aqueous micellar media than inwater alone. For example, Grassert and co-workersinvestigated the asymmetric hydrogenation of dialkyl1-benzamido-2-phenyl-ethenephosphonates with chiralrhodium complexes as catalysts and found that waterwas a poor solvent leading to low activities andenantioselectivities.200 With the addition of surfac-tants, however, ee’s of up to 99% were obtained.Studies into the mechanism behind the effect ofamphiphiles on the enantioselectivity of rhodium-catalyzed hydrogenations have been made. These arediscussed in further detail in section III.C.

The enantioselective reduction of dehydroaminoacid derivatives to yield optically pure amino acidshas been by far the most investigated asymmetrichydrogenation reaction.17 Several groups have re-cently reported on the use of water as solvent for thisreaction.55,56,201,202 For example, Uemura and co-workers described water-soluble chiral Rh(I) com-plexes derived from R,R- and â,â-trehalose (Figure9), which proved to be highly efficient catalysts of theasymmetric hydrogenation of enamides.56,201 Again,higher enantioselectivities were observed with water/surfactant mixtures than with water alone. In water/SDS mixtures, enantioselectivities of more than 99%could be achieved. Interestingly, these catalystsdisplayed amphiphilic nature in solubility, whichmade them equally good hydrogenation catalysts in1,2-dichloroethane as in water/SDS.

Scheme 24

Scheme 25

Scheme 26

Figure 9.

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2767

Page 18: B. Solubility of Organic Compounds in Water

Phosphine ligands have been used successfully inruthenium- and rhodium-catalyzed aqueous hydro-genations. Hiemstra and co-workers developed thediphosphines BIFAP, 72, and the sulfonated, water-soluble BIFAPS, 73 (Figure 10), and investigatedthem as ligands for the ruthenium-catalyzed hydro-genations of (Z)-acetamidocinnamic acid, 74, andmethyl acetoacetate.44 For the reaction of (Z)-aceta-midocinnamic acid, the use of (R)-BIFAPS in metha-nol or a 1:1 water/ethyl acetate mixture resulted inhydrogenated product (R)-75 of 72% ee (Scheme 27).The reaction rate in water/ethyl acetate was, how-ever, much lower than that in methanol. The hydro-genation of methyl acetoacetate with (R)-BIFAPS inwater proceeded to give 86% ee after the addition of1% sulfuric acid. For both substrates, however, thenonsulfonated BIFAP, 72, in methanol was superiorto BIFAPS, 73, in water. Without sulfuric acid in thelatter reaction the ee dropped dramatically, but thenature of the acid effect was not clear to the authors.The tetrahydroxylated, water-soluble analogue of thewidely used DuPHOS ligand named BASPHOS, 76,was remarkably efficient in the Rh(I)-catalyzed asym-metric hydrogenation of 2-acetamido acrylic acid, 77(99.6% ee, Scheme 28), and its methyl ester, 78(93.6% ee).203

A slightly different strategy of solubilizing hydro-genation catalysts has been pursued by Malmstromand Andersson.204,205 They prepared acrylic acid

polymers coupled to chiral rhodium-bisphosphinecomplexes and applied them to the asymmetrichydrogenation of dehydro amino acids and theiresters. While good yields and moderate enantio-selectivities could be obtained in water (pH 8) in thepresence of 1% of Rh catalyst, quantitative yields andee’s of up to 89% were observed when the samereaction was run in a 1:1 mixture of water/ethylacetate.

X. Radical ReactionsThe use of water as solvent for radical reactions

in organic synthesis is rare. This may be becauseearly attempts did not reveal any special effect ofwater compared to organic solvents. In general,radical reactions are not believed to be influenced bythe reaction medium to any greater extent. This isin contrast with many other reaction types developedfor use in aqueous media where the solvent effect isusually significant. Nevertheless, water should be agood medium for radical reactions considering thatit has no functionalities that are reactive toward freeradicals and a strong OH bond that makes undesiredhydrogen abstraction unlikely. The recent surge ofgeneral interest in water-based synthesis may leadto a reexamination of water as solvent for radicalreactions.206-211 Toward this end, significant solventeffects have in fact recently been reported in radicalcyclizations, especially in water. For example, thetriethylborane-mediated intramolecular radical cy-clization reaction of allyl iodoacetates 79a-c to giveγ-butyrolactones 80a-c was shown to be much moreefficient in water than in traditional solvents suchas benzene or hexane (Scheme 29).212 Indeed, noformation of lactone was observed in hexane orbenzene, but a yield of 78% of 80a was obtained atlow concentration (0.01 M) in water. Significantdiastereoselectivity was observed in the analogouscyclization reaction with the substituted allyl iodo-acetates 79b and 79c, yielding lactones 80b (trans/cis 86:14) and 80c (trans/cis 88:12), respectively.Unfortunately, the yields dropped dramatically withincreasing chain length of the 5-substituent (80b67%, 80c 32%), and the reaction was completelythwarted when R ) n-butyl. In general, increasingthe hydrophobicity of the substrates tended to sharplydecrease reactivity. This concern was addressed in arecent paper by Kita and co-workers where they usedwater-soluble radical initiators in water/surfactantsystems to effect the radical cyclization of a varietyof hydrophobic substrates.213 However, while yieldswere dramatically improved by addition of surfactant,the diastereoselectivities observed were only moder-ate at best. The Et3B-induced radical cyclization has

Figure 10. Reprinted with permission from ref 44. Copy-right 1999 Wiley-VCH.

Scheme 27a

a Reprinted with permission from ref 44. Copyright 1999 Wiley-VCH.

Scheme 28a

a Reprinted with permission from ref 203. Copyright 1999Elsevier Science.

Scheme 29

2768 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 19: B. Solubility of Organic Compounds in Water

found further application in the cyclization of N-allyl-2-iodo amides in water to provide synthetically at-tractive γ-lactams. Again, good trans-selectivitieswere obtained, especially with an electron-withdraw-ing sulfonate substituent on the nitrogen (90% ds).214

XI. ConclusionsThe past decade has seen extraordinary strides in

water-based organic synthesis, and research in thisfield still seems to be in an exponential mode ofdevelopment. This article has attempted to sum-marize key discoveries in this area as well as someof the most promising new methods of achievingstereocontrol in organic reactions developed in thelast 5 years. In this period, many catalytic stereo-selective reactions in water have been realized,mainly through the rational development of water-stable catalysts and water-soluble chiral ligands.Many of the most fundamentally useful reactions inasymmetric synthesis such as aldol reactions, al-lylations, aminohydroxylations, cycloadditions, cy-clopropanations, dihydroxylations, epoxidations, hy-drogenations, and others can now be performed withsimilar, or even improved, rates, yields, and selectivi-ties in aqueous media compared to the correspondingreactions in organic solvents. The unique hydrophobicproperties of water have led to the formulation of theconcept of hydrophobic control, an intriguing newmode of achieving stereoselectivity that has yet tobe extensively explored.

Concerns of the solubility of organic compounds inwater were also addressed. Many attractive targetmolecules are actually highly soluble in water, andwith improved synthetic methods in water, these maybe accessible directly without the need for derivati-zation. For compounds insoluble in water, the use ofcosolvents, buffers, surfactants, or hydrophilic aux-iliaries may be useful. Here it is necessary to em-phasize that the ultimate goal of developing syntheticmethods that are amenable with the use of water assolvent should not be to substitute organic solventsfor water in every transformation known but ratherto expand the options available to the syntheticchemist. We will arguably never get close to realizingthe full potential and scope of synthetic organicchemistry through the exclusive use of organic sol-vents.

Aqueous chemistry predominates in biological pro-cesses, and the development of synthetic aqueouschemistry may also aid our understanding of thedetailed mechanisms of the chemistry of life, e.g.biocatalytic processes. This may have implications forbiotechnological applications, such as artificial bio-mimetic systems.

On a final note, the greatest restriction to the widerimplementation of aqueous synthesis may be one ofa mental nature. It is hoped that this article willserve to rectify some of the misconceptions that mightpersist with many chemists regarding the inadequacyof water as solvent for organic reactions.

XII. AcknowledgmentsI am sincerely grateful to Professor Eric T. Kool

for his support of this time-consuming work. I also

thank Mr. Gregory Miller for linguistic improvementsof the preliminary manuscript. The Swedish Re-search Council is thankfully acknowledged for aPostdoctoral Fellowship.

XIII. References(1) Rideout, D. C.; Breslow, R. J. Am. Chem. Soc. 1980, 102, 7816.(2) Breslow, R.; Maitra, U.; Rideout, D. C. Tetrahedron Lett. 1983,

24, 1901.(3) Breslow, R.; Maitra, U. Tetrahedron Lett. 1984, 25, 1239.(4) Grieco, P. A.; Garner, P.; He, Z. Tetrahedron Lett. 1983, 25, 1897.(5) Grieco, P. A.; Yoshida, K.; Garner, P. J. Org. Chem. 1983, 48,

3137.(6) Gajewski, J. J. In Organic Synthesis in Water; Grieco, P. A., Ed.;

Blackie Academic & Professional: London, 1998.(7) Gajewski, J. J. Acc. Chem. Res. 1997, 30, 219.(8) Fringuelli, F.; Piermatti, O.; Pizzo, F. In Organic Synthesis in

Water; Grieco, P. A., Ed.; Blackie Academic & Professional:London, 1998.

(9) Kobayashi, S. In Organic Synthesis in Water; Grieco, P. A., Ed.;Blackie Academic & Professional: London, 1998.

(10) Lu, W.; Chan, T. H. J. Org. Chem. 2001, 66, 3467.(11) Loh, T.-P.; Zhou, J.-R. Tetrahedron Lett. 1999, 40, 9115.(12) Li, C.-J.; Chan, T.-H. Tetrahedron 1999, 55, 11149.(13) Fokin, V. V.; Sharpless, K. B. Angew. Chem., Int. Ed. Engl. 2001,

40, 3455.(14) Johnson, R. A.; Sharpless, K. B. In Catalytic Asymmetric

Synthesis; Ojima, I., Ed.; VCH: Weinheim, 2000.(15) Frohn, M.; Shi, Y. Synthesis 2000, 1979.(16) Cornils, B.; Herrmann, W. A. Aqueous Phase Organometallic

Catalysis-Concept and Applications; Wiley-VCH: Weinheim,1998.

(17) Nagel, U.; Albrecht, J. Top. Catal. 1998, 5, 3.(18) Grieco, P. A.; Editor Organic Synthesis in Water; Blackie

Academic & Professional: London, 1998.(19) Li, C.-J.; Chan, T.-H. Organic Reactions in Aqueous Media; John

Wiley & Sons: New York, 1997.(20) Lubineau, A.; Auge, J.; Queneau, Y. Synthesis 1994, 741.(21) Li, C. J. Chem. Rev. 1993, 93, 2023.(22) Yalkowsky, S. H. Solubility and Solubilization in Aqueous

Media; Oxford University Press: New York, 1999.(23) Anderson, B. D.; Flora, K. P. In The Practice of Medicinal

Chemistry; Wermuth, C. G., Ed.; Academic Press: London, 1996.(24) Grieco, P. A.; Yoshida, K.; Garner, P. J. Org. Chem. 1983, 48,

3137.(25) Keana, J. F. W.; Guzikowski, A. P.; Morat, C.; Volwerk, J. J. J.

Org. Chem. 1983, 48, 2661.(26) Grieco, P. A.; Galatsis, P.; Spohn, R. F. Tetrahedron 1986, 42,

2847.(27) Breslow, R.; Rizzo, C. J. J. Am. Chem. Soc. 1991, 113, 4340. See

also ref 102.(28) Tascioglu, S. Tetrahedron 1996, 52, 11113.(29) Wermuth, C. G. In The Practice of Medicinal Chemistry; Wer-

muth, C. G., Ed.; Academic Press: London, 1996.(30) Lubineau, A.; Bienayme, H.; Queneau, Y.; Scherrmann, M. C.

New J. Chem. 1994, 18, 279.(31) Lubineau, A.; Queneau, Y. Tetrahedron 1989, 45, 6697.(32) Lubineau, A.; Queneau, Y. J. Org. Chem. 1987, 52, 1001.(33) Lubineau, A.; Queneau, Y. Tetrahedron Lett. 1985, 26, 2653.(34) Lubineau, A.; Auge, J.; Bellanger, N.; Caillebourdin, S. J. Chem.

Soc., Perkin Trans. 1 1992, 1631.(35) Lubineau, A.; Auge, J.; Bellanger, N.; Caillebourdin, S. Tetra-

hedron Lett. 1990, 31, 4147.(36) Itami, K.; Nokami, T.; Yoshida, J.-I. Angew. Chem., Int. Ed. Engl.

2001, 40, 1074.(37) Anastas, P. T.; Heine, L. G.; Williamson, T. C. Green Chemical

Syntheses and Processes; American Chemical Society: Wash-ington, D.C., 2000.

(38) Paterniti, D. P.; Francisco, L. W.; Atwood, J. D. Organometallics1999, 18, 123.

(39) Hanson, B. E. Coord. Chem. Rev. 1999, 185, 795.(40) Papadogianakis, G.; Sheldon, R. A. Catalysis 1997, 13, 114.(41) Papadogianakis, G.; Sheldon, R. A. New J. Chem. 1996, 20, 175.(42) Herrmann, W. A.; Kohlpaintner, C. W. Angew. Chem., Int. Ed.

Engl. 1993, 32, 1524.(43) Herrmann, W. A.; Kohlpaintner, C. W. Angew. Chem., Int. Ed.

Engl. 1997, 36, 1049.(44) Gelpke, A. E. S.; Kooijman, H.; Spek, A. L.; Hiemstra, H. Chem.

Eur. J. 1999, 5, 2472.(45) Manabe, K.; Mori, Y.; Nagayama, S.; Odashima, K.; Kobayashi,

S. Inorg. Chim. Acta 1999, 296, 158.(46) Manabe, K.; Kobayashi, S. Tetrahedron Lett. 1999, 40, 3773.(47) Gawley, R. E.; Aube, J. Principles of Asymmetric Synthesis;

Elsevier Science: Oxford, 1996.(48) Otto, S.; Engberts, J. B. F. N. Pure Appl. Chem. 2000, 72, 1365.

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2769

Page 20: B. Solubility of Organic Compounds in Water

(49) Kobayashi, S. Top. Organomet. Chem. 1999, 2, 63.(50) Kobayashi, S.; Manabe, K. Pure Appl. Chem. 2000, 72, 1373.(51) Kobayashi, S.; Nagayama, S.; Busujima, T. J. Am. Chem. Soc.

1998, 120, 8287.(52) Kobayashi, S.; Hamada, T.; Nagayama, S.; Manabe, K. Org. Lett.

2001, 3, 165.(53) El-Qisairi, A.; Hamed, O.; Henry, P. M. J. Org. Chem. 1998, 63,

2790.(54) Hamed, O.; Henry, P. M. Organometallics 1998, 17, 5184.(55) Yan, Y.-Y.; RajanBabu, T. V. J. Org. Chem. 2001, 66, 3277.(56) Yonehara, K.; Hashizume, T.; Mori, K.; Ohe, K.; Uemura, S. J.

Org. Chem. 1999, 64, 5593.(57) Hashizume, T.; Yonehara, K.; Ohe, K.; Uemura, S. J. Org. Chem.

2000, 65, 5197.(58) Ferrara, M. L.; Orabona, I.; Ruffo, F.; Funicello, M.; Panunzi,

A. Organometallics 1998, 17, 3832.(59) Gyarmati, J.; Hajdu, C.; Dinya, Z.; Micskei, K.; Zucchi, C.; Palyi,

G. J. Organomet. Chem. 1999, 586, 106.(60) Otto, S.; Boccaletti, G.; Engberts, J. B. F. N. J. Am. Chem. Soc.

1998, 120, 4238.(61) Gu, W.; Zhou, W.-J.; Gin, D. L. Chem. Mater. 2001, 13, 1949.(62) Jamis, J.; Anderson, J. R.; Dickson, R. S.; Campi, E. M.; Jackson,

W. R. J. Organomet. Chem. 2000, 603, 80.(63) Dos Santos, S.; Tong, Y.; Quignard, F.; Choplin, A.; Sinou, D.;

Dutasta, J. P. Organometallics 1998, 17, 78.(64) Sartori, G.; Bigi, F.; Maggi, R.; Mazzacani, A.; Oppici, G. Eur.

J. Org. Chem. 2001, 2513.(65) De Vos, D. E.; Vankelecom, I. F. J.; Jacobs, P. A. Chiral Catalyst

Immobilization and Recycling; Wiley-VCH: Weinheim, 2000.(66) Uozumi, Y.; Shibatomi, K. J. Am. Chem. Soc. 2001, 123, 2919.(67) Cordes, E. H.; Dunlap, R. B. Acc. Chem. Res. 1969, 2, 329.(68) Oehme, G.; Paetzold, E.; Selke, R. J. Mol. Catal. 1992, 71, L1.(69) Otto, S.; Engberts, J. B. F. N. In Reactions and Synthesis in

Surfactant Systems; Texter, J., Ed.; Marcel Dekker: New York,2001.

(70) Rispens, T.; Engberts, J. B. F. N. Org. Lett. 2001, 3, 941.(71) Otto, S.; Engberts, J. B. F. N.; Kwak, J. C. T. J. Am. Chem. Soc.

1998, 120, 9517.(72) Diego-Castro, M. J.; Hailes, H. C. Tetrahedron Lett. 1998, 39,

2211.(73) Mori, Y.; Manabe, K.; Kobayashi, S. Angew. Chem., Int. Ed. Engl.

2001, 40, 2815.(74) Tian, H. Y.; Chen, Y. J.; Wang, D.; Bu, Y. P.; Li, C. J. Tetrahedron

Lett. 2001, 42, 1803.(75) Selke, R.; Holz, J.; Riepe, A.; Borner, A. Chem. Eur. J. 1998, 4,

769.(76) Grassert, I.; Oehme, G. J. Organomet. Chem. 2001, 621, 158.(77) Van der Wel, G. K.; Wijnen, J. W.; Engberts, J. B. F. N. J. Org.

Chem. 1996, 61, 9001.(78) Wijnen, J. W.; Engberts, J. B. F. N. Liebigs Ann./Recl. 1997,

1085.(79) Wijnen, J. W.; Engberts, J. B. F. N. J. Org. Chem. 1997, 62,

2039.(80) Ludwig, M.; Kadyrov, R.; Fiedler, H.; Haage, K.; Selke, R. Chem.

Eur. J. 2001, 7, 3298.(81) Kobayashi, S.; Wakabayashi, T.; Nagayama, S.; Oyamada, H.

Tetrahedron Lett. 1997, 38, 4559.(82) Kobayashi, S.; Wakabayashi, K.; Oyamada, H. Chem. Lett. 1997,

831.(83) Manabe, K.; Mori, Y.; Wakabayashi, T.; Nagayama, S.; Koba-

yashi, S. J. Am. Chem. Soc. 2000, 122, 7202.(84) Oehme, G.; Grassert, I.; Paetzold, E.; Meisel, R.; Drexler, K.;

Fuhrmann, H. Coord. Chem. Rev. 1999, 185, 585.(85) Oehme, G. In Aqueous-Phase Organometallic Catalysis-Concept

and Applications; Cornils, B., Herrmann, C. W., Eds.; Wiley-VCH: Weinheim, 1998.

(86) Paquette, L. A. In Green Chemical Syntheses and Processes;Anastas, P. T., Heine, L. G., Williamson, T. C., Eds.; Oxford:New York, 2000.

(87) Paquette, L. A.; Lobben, P. C. J. Org. Chem. 1998, 63, 5604.(88) Paquette, L. A.; Mitzel, T. M.; Isaac, M. B.; Crasto, C. F.;

Schomer, W. W. J. Org. Chem. 1997, 62, 4293.(89) Paquette, L. A.; Mitzel, T. M. J. Am. Chem. Soc. 1996, 118, 1931.(90) For earlier efforts in this area, see, for example: Enoch, K.;

Gordon, D. M.; Schmid, W.; Whitesides, G. M. J. Org. Chem.1993, 58, 5500. Chan, T. H.; Li, C. J. Can. J. Chem. 1992, 70,2726. Chan, T. H.; Li, C. J. J. Chem. Soc., Chem. Commun. 1992,747.

(91) Paquette, L. A.; Lobben, P. C. J. Am. Chem. Soc. 1996, 118, 1917.(92) Chen, X.; Hortelano, E. R.; Eliel, E. L.; Frye, S. V. J. Am. Chem.

Soc. 1990, 112, 6130.(93) Frye, S. V.; Eliel, E. L.; Cloux, R. J. Am. Chem. Soc. 1987, 109,

1862.(94) Loh, T. P.; Huang, J. M.; Xu, K. C.; Goh, S. H.; Vittal, J. J.

Tetrahedron Lett. 2000, 41, 6511.(95) Denis, C.; Laignel, B.; Plusquellec, D.; Le Marouille, J.-Y.; Botrel,

A. Tetrahedron Lett. 1996, 37, 53.(96) Lubineau, A.; Auge, J.; Lubin, N. J. Chem. Soc., Perkin Trans.

1 1990, 3011.

(97) Brandes, E.; Grieco, P. A.; Garner, P. J. Chem. Soc., Chem.Commun. 1988, 500.

(98) Hiraki, Y.; Tai, A. Chem. Lett. 1982, 341.(99) Waldmann, H.; Braun, M. Gazz. Chim. Ital. 1991, 121, 277.

(100) Waldmann, H. Angew. Chem., Int. Ed. Engl. 1989, 27, 274.(101) Waldmann, H. Liebigs Ann. Chem. 1989, 231.(102) Kumar, A. Chem. Rev. 2001, 101, 1.(103) Fringuelli, F.; Piermatti, O.; Pizzo, F.; Vaccaro, L. Eur. J. Org.

Chem. 2001, 439.(104) Garner, P. P. In Organic Synthesis in Water; Grieco, P. A., Ed.;

Blackie Academic & Professional: London, 1998.(105) Meijer, A.; Otto, S.; Engberts, J. B. F. N. J. Org. Chem. 1998,

63, 8989.(106) Engberts, J. B. F. N. Pure Appl. Chem. 1995, 67, 823.(107) Otto, S.; Blokzijl, W.; Engberts, J. B. F. N. J. Org. Chem. 1994,

59, 5372.(108) Blokzijl, W.; Engberts, J. B. F. N. In Enforced Hydrophobic

Interactions and Hydrogen Bonding in the Acceleration of Diels-Alder Reactions in Water; Cramer, C. J., Truhlar, D. G., Eds.;American Chemical Society: Washington, D.C., 1994.

(109) Kong, S.; Evanseck, J. D. J. Am. Chem. Soc. 2000, 122, 10418.(110) Cativiela, C.; Garcıa, J. I.; Mayoral, J. A.; Salvatella, L. Chem.

Soc. Rev. 1996, 209.(111) Furlani, T. R.; Gao, J. J. Org. Chem. 1996, 61, 5492.(112) Blake, J. F.; Lim, D.; Jorgensen, W. L. J. Org. Chem. 1994, 59,

803.(113) Jorgensen, W. L.; Blake, J. F.; Lim, D.; Severance, D. L. J. Chem.

Soc., Faraday Trans. 1994, 90, 1727.(114) Gajewski, J. J. J. Org. Chem. 1992, 57, 5500.(115) Reichardt, C. Solvents and Solvent Effects in Organic Chemistry,

2nd ed.; VCH: Weinheim, 1988.(116) For a discussion of polarity effects, see: Cativiela, C.; Garcia, J.

I.; Gil, J.; Martinez, R. M.; Mayoral, J. A.; Salvatella, L.; Urieta,J. S.; Mainar, A. M.; Abraham, M. H. J. Chem. Soc., PerkinTrans. 2 1997, 653 and references therein.

(117) Blokzijl, W.; Engberts, J. B. F. N. Angew. Chem., Int. Ed. Engl.1993, 32, 1545.

(118) Blokzijl, W.; Blandamer, M. J.; Engberts, J. B. F. N. J. Am.Chem. Soc. 1991, 113, 4241.

(119) Desimoni, G.; Faita, G.; Righetti, P. P.; Toma, L. Tetrahedron1990, 46, 7951.

(120) Schlachter, I.; Mattay, J.; Suer, J.; Hoeweler, U.; Wuerthwein,G.; Wuerthwein, E.-U. Tetrahedron 1997, 53, 119.

(121) Cativiela, C.; Garcıa, J. I.; Mayoral, J. A.; Salvatella, L. J. Chem.Soc., Perkin Trans. 2 1994, 847.

(122) Arseniyadis, S.; Rodriguez, R.; Yashunsky, D. V.; Camara, J.;Ourisson, G. Tetrahedron Lett. 1994, 35, 4843.

(123) Cativiela, C.; Garcıa, J. I.; Mayoral, J. A.; Royo, A. J.; Salvatella,L.; Assfeld, X.; Ruiz-Lopez, M. F. J. Phys. Org. Chem. 1992, 5,230.

(124) Otto, S.; Engberts, J. B. F. N. J. Am. Chem. Soc. 1999, 121, 6798.(125) Otto, S.; Bertoncin, F.; Engberts, J. B. F. N. J. Am. Chem. Soc.

1996, 118, 7702.(126) Waldmann, H.; Drager, M. Ann. Chem. 1990, 681.(127) Fringuelli, F.; Piermatti, O.; Pizzo, F. Targets Heterocycl. Syst.

1997, 1, 57.(128) Larsen, S. D.; Grieco, P. A. J. Am. Chem. Soc. 1985, 107, 1768.(129) Grieco, P. A.; Kaufman, M. D. J. Org. Chem. 1999, 64, 6041.(130) Yu, L.; Li, J.; Ramirez, J.; Chen, D.; Wang, P. G. J. Org. Chem.

1997, 62, 903.(131) Fringuelli, F.; Matteucci, M.; Piermatti, O.; Pizzo, F.; Burla, M.

C. J. Org. Chem. 2001, 66, 4661.(132) Attanasi, O. A.; De Crescentini, L.; Filippone, P.; Fringuelli, F.;

Mantellini, F.; Matteucci, M.; Piermatti, O.; Pizzo, F. Helv. Chim.Acta 2001, 84, 513.

(133) Da Silva, F. M.; Jones, J., Jr. J. Braz. Chem. Soc. 2001, 12, 135.(134) Venkatraman, S.; Meng, Y.; Li, C. J. Tetrahedron Lett. 2001,

42, 4459.(135) Venkatraman, S.; Meng, Y.; Li, C. J. Tetrahedron Lett. 2001,

42, 4459.(136) Shimizu, S.; Shirakawa, S.; Suzuki, T.; Sasaki, Y. Tetrahedron

2001, 57, 6169.(137) Bensa, D.; Brunel, J.-M.; Buono, G.; Rodriguez, J. Synlett 2001,

715.(138) Mori, Y.; Kakumoto, K.; Manabe, K.; Kobayashi, S. Tetrahedron

Lett. 2000, 41, 3107.(139) Moghaddam, F. M.; Mohammadi, M.; Hosseinnia, A. Synth.

Commun. 2000, 30, 643.(140) Da Silva, F. M.; Gomes, A. K.; Jones, J., Jr. Can. J. Chem. 1999,

77, 624.(141) Kolb, H. C.; Van Nieuwenhze, M. S.; Sharpless, K. B. Chem. Rev.

1994, 94, 2483.(142) Kobayashi, S.; Ishida, T.; Akiyama, R. Org. Lett. 2001, 3, 2649.(143) Mugdan, M.; Young, P. D. J. Chem. Soc. 1949, 2988.(144) Schlingloff, G.; Sharpless, K. B. In Asymmetric Oxidation

Reactions: A Practical Approach; Katsuki, T., Ed.; OxfordUniversity Press: Oxford, 2001.

2770 Chemical Reviews, 2002, Vol. 102, No. 8 Lindstrom

Page 21: B. Solubility of Organic Compounds in Water

(145) Kolb, H. C.; Sharpless, K. B. In Transition Metals for FineChemicals and Organic Synthesis; Beller, M., Bolm, C., Eds.;Wiley-VCH: Weinheim, 1998; Vol. 2.

(146) Li, G.; Chang, H.-T.; Sharpless, K. B. Angew. Chem., Int. Ed.Engl. 1996, 35, 451.

(147) Yang, D.; Wong, M.-K.; Yip, Y.-C.; Wang, X.-C.; Tang, M.-W.;Zheng, J.-H.; Cheung, K.-K. J. Am. Chem. Soc. 1998, 120, 5943.

(148) Frohn, M.; Dalkiewicz, M.; Tu, Y.; Wang, Z.-X.; Shi, Y. J. Org.Chem. 1998, 63, 2948.

(149) Wang, Z.-X.; Shi, Y. J. Org. Chem. 1998, 63, 3099.(150) Iwasa, S.; Takezawa, F.; Tuchiya, Y.; Nishiyama, H. Chem.

Commun. 2001, 59.(151) Groger, H.; Vogl, E. M.; Shibasaki, M. Chem. Eur. J. 1998, 4,

1137.(152) Lubineau, A. J. Org. Chem. 1986, 51, 2142.(153) Nagayama, S.; Kobayashi, S. J. Am. Chem. Soc. 2000, 122,

11531.(154) Kobayashi, S.; Nagayama, S.; Busujima, T. Chem. Lett. 1999,

71.(155) Manabe, K.; Mori, Y.; Kobayashi, S. Tetrahedron 1999, 55,

11203.(156) Kobayashi, S.; Nagayama, S.; Busujima, T. Tetrahedron 1999,

55, 8739.(157) Kobayashi, S.; Mori, Y.; Nagayama, S.; Manabe, K. Green Chem.

1999, 1, 175.(158) Le Roux, C.; Ciliberti, L.; Laurent-Robert, H.; Laporterie, A.;

Dubac, J. Synlett 1998, 1249.(159) Loh, T.-P.; Chua, G.-L.; Vittal, J. J.; Wong, M.-W. Chem.

Commun. 1998, 861.(160) Alcaide, B.; Almendros, P.; Aragoncillo, C.; Rodrıguez-Acebes,

R. J. Org. Chem. 2001, 66, 5208.(161) Lu, W.; Chan, T. H. J. Org. Chem. 2000, 65, 8589.(162) Cho, Y. S.; Lee, J. E.; Pae, A. N.; Choi, K. I.; Koh, H. Y.

Tetrahedron Lett. 1999, 40, 1725.(163) Hanessian, S.; Park, H.; Yang, R. Y. Synlett 1997, 353.(164) Loh, T. P.; Zhou, J.-R. Tetrahedron Lett. 2000, 41, 5261.(165) Loh, T.-P.; Xu, J.; Hu, Q.-Y.; Vittal, J. J. Tetrahedron: Asym-

metry 2000, 11, 1565.(166) Chang, H.-M.; Cheng, C.-H. Org. Lett. 2000, 2, 3439.(167) Chan, T. H.; Yang, Y.; Li, C. J. J. Org. Chem. 1999, 64, 4452.(168) Loh, T.-P.; Li, X.-R. Eur. J. Org. Chem. 1999, 1893, 3.(169) Loh, T.-P.; Xu, J. Tetrahedron Lett. 1999, 40, 2431.(170) Li, L. H.; Chan, T. H. Tetrahedron Lett. 2000, 41, 5009.(171) Khan, R. H.; Prasada Rao, T. S. R. J. Chem. Res., Synop. 1998,

202.(172) Li, C.-J.; Meng, Y.; Yi, X.-H.; Ma, J.; Chan, T.-H. J. Org. Chem.

1998, 63, 7498.(173) Li, C.-J.; Meng, Y.; Yi, X.-H.; Ma, J.; Chan, T.-H. J. Org. Chem.

1997, 62, 8632.(174) Zhang, W.-C.; Li, C.-J. J. Org. Chem. 1999, 64, 3230.(175) Li, C.-J.; Zhang, W.-C. J. Am. Chem. Soc. 1998, 120, 9102.(176) Chan, T. H.; Yang, Y. Tetrahedron Lett. 1999, 40, 3863.(177) Shin, J. A.; Cha, J. H.; Pae, A. N.; Choi, K. I.; Koh, H. Y.; Kang,

H. Y.; Cho, Y. S. Tetrahedron Lett. 2001, 42, 5489.(178) Canac, Y.; Levoirier, E.; Lubineau, A. J. Org. Chem. 2001, 66,

3206.(179) Araki, S.; Shiraki, F.; Tanaka, T.; Nakano, H.; Subburaj, K.;

Hirashita, T.; Yamamura, H.; Kawai, M. Chem. Eur. J. 2001,7, 2784.

(180) Bernardelli, P.; Paquette, L. A. J. Org. Chem. 1997, 62, 8284.(181) Paquette, L. A.; Mitzel, T. M. Tetrahedron Lett. 1995, 36, 6863.(182) Risch, N.; Arend, M.; Westermann, B. Angew. Chem., Int. Ed.

Engl. 1998, 37, 1044.(183) Loh, T.-P.; Liung, S. B. K. W.; Tan, K.-L.; Wei, L.-L. Tetrahedron

2000, 56, 3227.(184) Kobayashi, S.; Busujima, T.; Nagayama, S. Synlett 1999, 545.(185) Loh, T.-P.; Wei, L.-L. Tetrahedron Lett. 1998, 39, 323.(186) Akiyama, T.; Takaya, J.; Kagoshima, H. Tetrahedron Lett. 2001,

42, 4025.(187) Manabe, K.; Mori, Y.; Kobayashi, S. Tetrahedron 2001, 57, 2537.(188) Manabe, K.; Kobayashi, S. Org. Lett. 1999, 1, 1965.(189) Akiyama, T.; Takaya, J.; Kagoshima, H. Synlett 1999, 1426.(190) Akiyama, T.; Takaya, J.; Kagoshima, H. Synlett 1999, 1045.(191) Schmitzer, A.; Perez, E.; Rico-Lattes, I.; Lattes, A.; Rosca, S.

Langmuir 1999, 15, 4397.(192) Schmitzer, A. R.; Franceschi, S.; Perez, E.; Rico-Lattes, I.; Lattes,

A.; Thion, L.; Erard, M.; Vidal, C. J. Am. Chem. Soc. 2001, 113,5956.

(193) Feuerstein, M.; Laurenti, D.; Doucet, H.; Santelli, M. Tetrahe-dron Lett. 2001, 42, 2313.

(194) Genet, J. P.; Savignac, M. J. Organomet. Chem. 1999, 576, 305.(195) Uozumi, Y.; Danjo, H.; Hayashi, T. Tetrahedron Lett. 1998, 39,

8303.(196) Severance, D. L.; Jorgensen, W. L. J. Am. Chem. Soc. 1992, 114,

10966.(197) Brandes, E. B.; Grieco, P. A.; Gajewski, J. J. J. Org. Chem. 1989,

54, 515.(198) Hu, H.; Kobrak, M. N.; Xu, C.; Hammes-Schiffer, S. J. Phys.

Chem. A 2000, 104, 8058.(199) Grieco, P. A.; Brandes, E.; McCann, S.; Clark, J. D. J. Org. Chem.

1989, 54, 5849.(200) Grassert, I.; Schmidt, U.; Ziegler, S.; Fischer, C.; Oehme, G.

Tetrahedron: Asymmetry 1998, 9, 4193.(201) Yonehara, K.; Ohe, K.; Uemura, S. J. Org. Chem. 1999, 64, 9381.(202) Shin, S.; RajanBabu, T. V. Org. Lett. 1999, 1, 1229.(203) Holz, J.; Heller, D.; Sturmer, R.; Borner, A. Tetrahedron Lett.

1999, 40, 7059.(204) Malmstrom, T.; Andersson, C. J. Mol. Catal. A: Chem. 2000,

157, 79.(205) Malmstrom, T.; Andersson, C. J. Mol. Catal. A: Chem. 1999,

139, 259.(206) Yorimitsu, H.; Shinokubo, H.; Oshima, K. Bull. Chem. Soc. Jpn.

2001, 74, 225.(207) Miyabe, H.; Ueda, M.; Naito, T. Chem. Commun. 2000, 2059.(208) Miyabe, H.; Ueda, M.; Naito, T. J. Org. Chem. 2000, 65, 5043.(209) Jang, D. O. Tetrahedron Lett. 1998, 39, 2957.(210) Shaw, H.; Perlmutter, H. D.; Gu, C.; Arco, S. D.; Quibuyen, T.

O. J. Org. Chem. 1997, 62, 236.(211) Yamazaki, O.; Togo, H.; Nogami, G.; Yokoyama, M. Bull. Chem.

Soc. Jpn. 1997, 70, 2519.(212) Yorimitsu, H.; Nakamura, T.; Shinokubo, H.; Oshima, K.; Omoto,

K.; Fujimoto, H. J. Am. Chem. Soc. 2000, 122, 11041.(213) Kita, Y.; Nambu, H.; Ramesh, N. G.; Anilkumar, G.; Matsugi,

M. Org. Lett. 2001, 3, 1157.(214) Wakabayashi, K.; Yorimitsu, H.; Shinokubo, H.; Oshima, K. Bull.

Chem. Soc. Jpn. 2000, 73, 2377.

CR010122P

Stereoselective Organic Reactions in Water Chemical Reviews, 2002, Vol. 102, No. 8 2771

Page 22: B. Solubility of Organic Compounds in Water