Dual enantioselective control by heterocycles of (S)-indoline derivatives*

Similar documents
Stereoselective reactions of enolates

Asymmetric Catalysis by Lewis Acids and Amines

Corey-Bakshi. Bakshi-Shibata Reduction. Name Reaction Nilanjana Majumdar

Story Behind the Well-Developed Chiral Lewis Acid in Asymmetric Diels-Alder reaction

Advanced Organic Chemistry

Stereoselective reactions of the carbonyl group

N-Heterocyclic Carbene Catalysis via Azolium Dienolates: An Efficient Strategy for Enantioselective Remote Functionalizations

The aldol reaction with metal enolates proceeds by a chair-like, pericyclic process: favored. disfavored. favored. disfavored

Glyoxylic acid derivatives in asymmetric synthesis*

Chiral Supramolecular Catalyst for Asymmetric Reaction

Diels-Alder Reaction

Racemic catalysis through asymmetric activation*

Metal-catalyzed asymmetric hetero-diels-alder reactions of unactivated dienes with glyoxylates

Copper-mediated asymmetric transformations*

Chapter 4 Functional Group Transformations: Oxidation and Reduction. 4.8 Terminology for Reduction of Carbonyl Compounds

Use of Cp 2 TiCl in Synthesis

Chiral Catalyst II. Palladium Catalysed Allylic Displacement ( -allyl complexes) 1. L n Pd(0) 2. Nuc

Chiral Auxiliaries. attach auxiliary Substrate Substrate Auxiliary

Enantioselective Borylations. David Kornfilt Denmark Group Meeting Sept. 14 th 2010

A Tandem Semipinacol Rearrangement/Alkylation of a-epoxy Alcohols: An Efficient and Stereoselective Approach to Multifunctional 1,3-Diols

Stereoselective reactions of enolates: auxiliaries

Chiral Amplification. Literature Talk Fabian Schneider Konstanz, Universität Konstanz

Chiral Catalysis. Chiral Catalyst. Substrate. Chiral Catalyst

Page 1 of 9. Sessional Examination (November 2017) Max Marks: 20 Date: Time: One Hour. Model Answers

Diels-Alder Reactions

Highlights of Schmidt Reaction in the Last Ten Years

Radical cascade reactions triggered by single electron transfer

Tips for taking exams in 852

Chromium Arene Complexes

Enantioselective Protonations

Homogeneous Catalysis - B. List

B X A X. In this case the star denotes a chiral center.

P. Wipf - Chem /29/2006. To achieve high diastereo- and enantioselectivity, it is necessary to:

Organic Chemistry I (Chem340), Spring Final Exam

Lecture Notes Chem 51B S. King I. Conjugation

1. Theoretical Investigation of Mechanisms and Stereoselectivities of Synthetic Organic Reactions

Chiral Brønsted Acid Catalysis

Tautomerism and Keto Enol Equilibrium

Organic Chemistry Lecture 2 - Hydrocarbons, Alcohols, Substitutions

A Highly Efficient Organocatalyst for Direct Aldol Reactions of Ketones and Aldehydes

Suggested solutions for Chapter 41

Synthetic Efforts Towards Anthracyclines Using the Asymmetric Diels-Alder Reaction

Department of Chemistry, University of Saskatchewan Saskatoon SK S7N 4C9, Canada. Wipf Group. Tyler E. Benedum Current Literature February 26, 2005

Stereodivergent Catalysis. Aragorn Laverny SED Group Meeting July

INTRODUCTION. Development of simple and convenient methodologies for an efficient asymmetric

Nuggets of Knowledge for Chapter 17 Dienes and Aromaticity Chem 2320

Stereoselective Organic Synthesis

Organic Chemistry II / CHEM 252 Chapter 13 Conjugated Unsaturated Systems

PHOTOCATALYSIS: FORMATIONS OF RINGS

Suggested solutions for Chapter 34

sp 3 C-H insertion by α-oxo Gold Carbene B4 Kei Ito

Recent advances in transition metal-catalyzed or -mediated cyclization of 2,3-allenoic acids: New methodologies for the synthesis of butenolides*

O + k 2. H(D) Ar. MeO H(D) rate-determining. step?

-catalyzed reactions utilizing isocyanides as a C 1

Citation for published version (APA): Otto, S. (1998). Catalysis of Diels-Alder reactions in water Groningen: s.n.

Catalytic asymmetric borane reduction of prochiral ketones using chiral camphor-derived mercapto-alcohols

Chiral Brønsted Acid Catalysis

Organometallic Compounds of Magnesium *

Pericyclic Reactions and Organic Photochemistry S. Sankararaman Department of Chemistry Indian Institute of Technology, Madras

Chapter 14. Principles of Catalysis

COURSE OBJECTIVES / OUTCOMES / COMPETENCIES.

Catalytic Enantioselective Diels-Alder Reactions Chwee T.S 1 and Wong M.W 2

C h a p t e r 1. Enantioselective LUMO-Lowering Organocatalysis. The presentation of the Nobel Prize in 2001 to William S. Knowles, Ryoji Noyori,

Lecture 3: Aldehydes and ketones

Catalytic Asymmetric [4+1] Annulation of Sulfur Ylides with Copper Allenylidene Intermediates. Reporter: Jie Wang Checker: Shubo Hu Date: 2016/08/02

molecules ISSN

Electrophilic Carbenes

Measuring enzyme (enantio)selectivity

Chapter 13 Conjugated Unsaturated Systems

Lewis Base Catalysis: the Aldol Reaction (Scott Denmark) Tom Blaisdell Friday, January 17 th 2014 Topic Talk

Titanacyclopropanes as versatile intermediates for carbon carbon bond formation in reactions with unsaturated compounds*

Diels Alder Cycloaddition Strategy for Kinetic Resolution of Chiral Pyrazolidinones

Chiral phenyl-bis(oxazoline) as an efficient auxiliary for asymmetric catalysis*

Organolithium Compounds *

Mechanistic Studies of Allylsilane Rearrangement

The Vinylogous Aldol Reaction

Planar-Chiral Phosphine-Olefin Ligands Exploiting a (Cyclopentadienyl)manganese(I) Scaffold to Achieve High Robustness and High Enantioselectivity

Catalytic Asymmetric Addition to Ketones Using Organozinc Reagents

CHEM 261 HOME WORK Lecture Topics: MODULE 1: The Basics: Bonding and Molecular Structure Text Sections (N0 1.9, 9-11) Homework: Chapter 1:

Chapter 15 Dienes, Resonance, and Aromaticity

Functionalized main-group organometallics for organic synthesis*

Shi Asymmetric Epoxidation

The problem is that your product still has a-protons, and can keep on forming enolates to get more methyl groups added:

Molybdenum-Catalyzed Asymmetric Allylic Alkylation

Ethers. Chapter 14: Ethers, Epoxides, & Sulfides. General Formula: Types: a) Symmetrical: Examples: b) Unsymmetrical: Examples: Physical Properties:

Application of Ionic Liquids in Michael Addition Reactions

Lecture Notes Chem 51C S. King. Chapter 20 Introduction to Carbonyl Chemistry; Organometallic Reagents; Oxidation & Reduction

II. Special Topics IIA. Enolate Chemistry & the Aldol Reaction

1. Radical Substitution on Alkanes. 2. Radical Substitution with Alkenes. 3. Electrophilic Addition

Nucleophilic Heterocyclic Carbene Catalysis. Nathan Werner Denmark Group Meeting September 22 th, 2009

Exam 1 (Monday, July 6, 2015)

GENERAL METHODS OF ORGANIC CHEMISTRY; APPARATUS THEREFOR (preparation of carboxylic acid esters by telomerisation C07C 67/47; telomerisation C08F)

The aza-baylis-hillman Reaction: Mechanism, Asymmetric Catalysis, & Abnormal Adducts. Larry Wolf SED Group Meeting

Reactivity in Organic Chemistry. Mid term test October 31 st 2011, 9:30-12:30. Problem 1 (25p)

A Modular Approach to Polyketide Building Blocks: Cycloadditions of Nitrile Oxides and Homoallylic Alcohols

Development of Small Organic Molecules as Catalysts for Asymmetric

Predicative Design of Stereoinduction in Catalytic, Asymmetric Reactions

Amphoteric Molecules < Chemistry of Andrei K. Yudin > Hyung Min Chi 17 JUNE 2014

Copper(II)-Catalyzed Exo and Enantioselective Cycloadditions of Azomethine Imines

Functionalization of C(sp 3 ) H Bonds Using a Transient Directing Group

Transcription:

Pure Appl. Chem., Vol. 77, No. 12, pp. 2053 2059, 2005. DOI: 10.1351/pac200577122053 2005 IUPAC Dual enantioselective control by heterocycles of (S)-indoline derivatives* Yong Hae Kim, Doo Young Jung, So Won Youn, Sam Min Kim, and Doo Han Park Center for Molecular Design and Synthesis, Department of Chemistry, Korea Advanced Institute of Science and Technology, Taejon 305-701, Korea Abstract: Diastereo-and enantioselective pinacol coupling reactions of chiral α-ketoamides mediated by samarium diiodide afforded extremely high diastereoselectivities. Enantiopure (S,S)- or (R,R)-2,3-dialkyltartaric acid and derivatives can be synthesized. Diels Alder cycloadditions of S-indoline chiral acrylamides with cyclopentadiene in the presence of Lewis acids proceed with high diastereofacial selectivity, giving either endo-r or endo-s products depending on Lewis acid and the structures of chiral dienophiles. Keywords: Enantioselective; pinacol coupling; α-ketoamides; samarium diiodide; S-indoline. In chemical transformations, the synthesis of individual enantiomers is generally achieved by using natural chiral sources. Because sometimes natural sources of one of the enantiomers may be limited, it is desirable and important to obtain both enantiomers by stereocontrolled reactions of prochiral compounds that utilize the same chiral source. Only a few examples have been reported where the same reaction system gives both enantiomers depending on reaction conditions. (S)-Indoline-2-carboxylic acid has been found to be one of the best starting materials for the synthesis of the new chiral materials, especially for obtaining both R- and S-enantiomers from asymmetric reactions starting from the same starting material [1]. There are several examples of production of both diastereomers or enantiomers with high stereoselectivity using chiral catalysts or ligands derived from the same starting material of (S)-indoline-2-carboxylic acid [2,3]. The intermolecular coupling of various aldehydes or ketones to the corresponding pinacols has been extensively studied. However, enantiopure pinacols have not really been obtained using this type of coupling [4]. To obtain chiral 2,3-dialkyltartaric acids, which can be utilized in design of chiral ligands or chiral auxiliaries, we studied the pinacol coupling reaction of chiral α-ketoamides 1 with SmI 2 to obtain pinacol product 2 in extremely high diastereoselectivities. α-ketoamides 1 in the presence of SmI 2, HMPA, and tbuoh in tetrahydrofuran (THF) dimerized to give pinacol 2 with extremely high diastereoselectivity (>98 % de in some cases; Scheme 1). Also, we have found that (2S,3aS,7aS)-N-pyruvoyl-2-(tert-butyldiphenylsilyloxy)octahydroindoline 1a' (99.8 % ee), affords the opposite configuration of 3a' (R,R diol: 97 % de). The ratio of R,R:S,S:meso is 98.5:~0.5:1.5 [2]. *Paper based on a presentation at the 7 th IUPAC International Conference on Heteroatom Chemistry (ICHAC-7), Shanghai, China, 21 25 August 2004. Other presentations are published in this issue, pp. 1985 2132. Corresponding author 2053

2054 Y. H. KIM et al. Scheme 1 The reaction appears to be initiated by formation of a ketyl radical A with a one-electron transfer to SmI 2 followed by coupling of the two ketyl radicals (Fig. 1). The coupling takes place at the less hindered si face and gives high diastereoselectivities. The different configurations of the adducts produced can be rationalized by the different intermediates formed from 1 and 1a' with SmI 2. Instead of A, the intermediate A' is favorable in the case of 1a' more so than A" due to the steric repulsion between the chair conformation of the cyclohexane ring and the amide carbonyl moiety (Fig. 1). When the reaction was performed with l-proline-derived chiral α-ketoamides, only poor selectivity was observed, indicating that the phenyl ring of indoline-2-carboxylic acid helped A to adopt the depicted conformation. Fig. 1 In connection with the pinacol coupling reaction, α,β-unsaturated amides have been found to react with SmI 2 to form dimerized products containing two chiral carbons which are first obtained as the adjacent chiral hydrocarbons (Scheme 2). Scheme 2 In the Lewis acid-catalyzed Diels Alder reactions of chiral dienophiles, generally the S-form of the chiral dienophile (auxiliary) exclusively affords the endo-r adduct over the endo-s one, and the R-form exclusively gives the S-adduct over the endo-r one [5]. In the hope of obtaining the opposite configuration of the endo adduct and understanding the mechanism, three different dienophiles 6, 7, and

Dual enantioselective control 2055 Scheme 3 8 were prepared and reacted with dienes 9 and 10 in the presence of various Lewis acids, as shown in Scheme 3 [3]. Extremely high levels of asymmetric induction can be achieved in Diels Alder cycloadditions of 6 or 8 with 9; in contrast to other general dienophiles, 6 containing a carboxylate moiety reacts with 9 to give differently configured adducts depending on the Lewis acids employed; in the presence of TiCl 4, Ti(OPr i ) 4, or SnCl 4, 11a was obtained as the major diastereomer (11a:11b = endo-r:endo-s = >99:1), but with AlEt 2 Cl, ZnCl 2, or BF 3 Et 2 O, the opposite configuration of 10b was obtained (11a:11b = 1:>99). In particular, 8 containing a diphenyl-substituted tertiary alcohol moiety affords exceptionally high diastereofacial selectivities (13a:13b = >99:1, yield = >90 %) regardless of the nature of the Lewis acid. The differently configured adducts produced can be rationalized by different intermediates formed between 6 8 and the metals of the Lewis acids. Compounds 6 8 react with 9 to favor formation of endo-r species 11a or 13a with TiCl 4, Ti(OPr i ) 4, SnCl 4, or ZrCl 4 probably via formation of sevenmembered ring chelates with the acryloyl moiety of 15 or 16 having a cisoid conformation (Fig 2). On the other hand, 6 or 7 prefer endo-s formation 11b or 12b with ZnCl 2, AlEtCl 2, or BF 3 Et 2 O, with high diastereofacial selectivity probably resulting from intermediate 14, as shown in Fig. 2. Species 6 and 8 also reacted with less reactive acyclic diene 10 at 25 C to result in the same trend: for 6 with TiCl 4 the ratio of endo-r:endo-s was 97:3, while with EtAlCl 2 the ratio was reversed to 3:97, for 8 with both TiCl 4 and Et 2 AlCl, endo-r:endo-s = 97:3 and 94:6, respectively. Fig. 2

2056 Y. H. KIM et al. There are also examples of employing (S)-indoline-2-carboxylic acid-derived amino alcohols as chiral catalysts in asymmetric reactions [6,7]. Asymmetric reduction of ketones by amino alcohol borane systems and asymmetric alkylation of aldehydes using dialkyl zinc reagents and amino alcohol catalysts are those examples. In both of the reactions, by just slight modification of indoline moiety, dual enantioselective control was achieved. Numerous β-amino alcohols were tested as chiral ligands for the enantioselective reduction of ketones. Although several efficient catalysts have been developed, most of the oxazaborolidine catalysts derived from l-amino acids give access to only one enantiomer of secondary alcohols; namely, (S)-amino alcohol borane systems have been reported to give arylalkyl alcohols with the (R)-configuration and (R)-amino alcohol borane systems give (S)-arylalkyl alcohol, respectively [8]. However, in contrast to l-amino acids, which are readily available from natural sources, enantiopure d-amino acids are often difficult to obtain. When the new β-amino alcohols 17, 18 are employed as catalysts in the asymmetric reduction of ketones, a remarkable reversal of enantiofacial selectivity is demonstrated [6] (Scheme 4). When acetophenone was reduced in the presence of catalytic amount (10 mol %) of 17b, (R)-phenylethanol of high optical purity was obtained. On the other hand, when the catalyst was 18a, phenylethanol of opposite configuration [(S)-phenylethanol] was obtained in comparable optical purity. The chiral catalysts could be recycled without any significant loss of activity or optical purity. The same trend remained through other substrates tested, including even aliphatic ketone. Scheme 4 The following model can rationalize the obtained results. In the presence of catalyst 17b, the transition state B is favored the competing transition state B' due to the steric effect of diphenyl group (Fig. 3). In this case, the phenyl ring of indoline moiety does not exert significant steric interactions in the oxazaborolidine ketone borane complex. On the contrary, in the case of 18a, the steric effect of the cyclohexyl group appears to enforce the approach of chiral oxazaborolidine complex to the opposite face of carbonyl compounds as in the transition state C. The transition state C is favored over C', because the cyclohexyl group from the oxazaborolidine repels the methyl or other R S of substrates. These

Dual enantioselective control 2057 Fig. 3 transition states are well supported by the well-established fact that Lewis acidic boron atom favors the coordination of ketone with the large substituent R L anti to itself [9]. Despite numerous extensive studies regarding the asymmetric addition of organometallic reagents to aldehydes using chiral ligands [10], most of the chiral catalysts derived from one enantiomer of the natural or synthetic chiral sources, only allow the preparation of only one enantiomer of products. For example, β-amino alcohol catalysts of (S)-configuration produce only (S)-configured secondary alcohols in dialkyl zinc addition reactions. Because it is very important in good asymmetric catalysis that preparation of both of the enantiomers is efficiently achieved from the chiral catalysts derived from the same natural sources, the new chiral amino alcohols 17, 18 were again employed in the asymmetric alkylation of various aldehydes. Again in this case, the remarkable reversal of enantiofacial selectivity was observed [7]. In the amino alcohol-catalyzed diethyl zinc addition to benzaldehyde, 17b catalyzed the addition to give (S)-phenylethanol in 96 % e.e., whereas 18a catalyzed to give (R)-phenylethanol in 90 % e.e. Also, they showed excellent ability to work as chiral catalysts against other aldehydes, including cinnamaldehyde and aliphatic aldehydes. Scheme 5

2058 Y. H. KIM et al. The possible mechanisms for the diethyl zinc addition to aldehydes by 17b or 18a are shown in Fig. 4. As in the models developed for chiral reduction of ketones, with 17b, the diphenyl group exerts powerful steric effect on the transition state, thus favoring D over D'. Due to the steric factor of diphenyl group, the diethyl zinc reagents can only approach si face of the benzaldehyde. With 18a, the cyclohexyl group directs the course of the reaction favoring transition state E with Re face attack of the zinc reagent. Fig. 4 In conclusion, starting from the same starting material, (S)-indoline-2-carboxylic acid, dual enantioselective control, the process that results in both R-and S-enantiomers in the same reaction system, has been achieved for the first time. Enantiopure (S,S)- or (R,R)-2,3-dialkyltartaric acid and derivatives can be synthesized. Also, asymmetric Diels Alder cyclizations of chiral amides with Lewis acids afforded extremely high diastereoselectivity of both opposite configuration of diastereomers depending on the Lewis used. Also, amino alcohols derived from (S)-indoline-2-carboxylic acid were used as efficient catalysts in asymmetric reactions. They also exhibited dual enantioselective control in the asymmetric reduction of ketones and asymmetric alkylation of aldehydes. As shown in the examples above, careful design and selection of natural chiral sources, (S)-indoline-2-carboxylic acid, enabled the first dual enantioselective control in catalytic and stoichiometric asymmetric processes. Also, when compared with related l-proline derivatives, indoline derivatives not only showed dual enantioselective control, but also provided better reactivities. From these facts, it is apparent that efforts to develop new chiral ligands and auxiliaries should be directed toward dual enantioselective control and improved reactivity with the tool of rational design. REFERENCES 1. (a) Y. H. Kim. Acc. Chem. Res. 34, 955 (2001); (b) Y. H. Kim, D. H. Park, I. S. Byun. J. Org. Chem. 58, 4511 (1993); (c) Y. H. Kim and J. Y. Choi. Tetrahedron Lett. 37, 5543 (1996). 2. S. M. Kim, I. S. Byun, Y. H. Kim. Angew. Chem., Int. Ed. Engl. 39, 728 (2000). 3. D. H. Park, S. H. Kim, J. D. Kim, Y. H. Kim. J. Chem. Soc., Chem. Commun. 963 (1999). 4. (a) T. Wirth. Angew. Chem., Int. Ed. Engl. 35, 61 (1996); (b) A. W. Konradi, S. J. Kemp, S. F. Pedersen. J. Am. Chem. Soc. 116, 1316 (1994); (c) B. Kammermeier, G. Beck, D. Jacobi, H. Jendralla. Angew. Chem., Int. Ed. Engl. 33, 685 (1994). 5. For reviews, see: (a) L. A. Paquett. Asymmetric Synthesis, Vol. 3, J. D. Morrison (Ed.), Academic Press, New York (1994); (b) W. Oppolzer. Angew. Chem., Int. Ed. Engl. 23, 876 (1984); (c) H. Waldmann. Synthesis 535 (1994). 6. Y. H. Kim, D. H. Park, I. S. Byun, I. K. Yoon, C. S. Park. J. Org. Chem. 56, 4511 (1993). 7. Y. H. Kim, D. H. Park, I. S. Byun. Heteroatom Chem. 3, 51 (1992).

Dual enantioselective control 2059 8. (a) S. Itsuno, M. Nakano, K. Ito, A. Hirao, M. Owa, N. Kanda, S. Nakahama. J. Chem. Soc., Perkin Trans. 1 2615 (1985); (b) E. J. Corey, R. K. Bakshi, S. Shibata, C. P. Chen, V. K. Singh. J. Am. Chem. Soc. 109, 7925 (1987); (c) I. K. Youn, S. W. Lee, C. S. Park. Tetrahedron Lett. 29, 4453 (1988); (d) J. Martens, Ch. Dauelsberg, W. Behnen, S. Wallbaum. Tetrahedron: Asymmetry 3, 347 (1992) 9. (a) E. J. Corey, R. K. Bakshi, S. Shibata. J. Am. Chem. Soc. 109, 5551 (1987); (b) V. Nevalinen. Tetrahedron: Asymmetry 2, 63 (1991). 10. (a) M. Kitamura, S. Suga, K. Kawai, R. Noyori. J. Am. Chem. Soc. 108, 6071 (1986); (b) K. Soai, A. Ookawa, T. Kaba, K. Ogawa. J. Am. Chem. Soc. 109, 7111 (1987); (c) E. J. Corey, P.-W. Yuen, F. J. Hannon, D. A. Wierda. J. Org. Chem. 55, 784 (1990).