Next Article in Journal
Antihyperglycemic Effect of Orthosiphon Stamineus Benth Leaves Extract and Its Bioassay-Guided Fractions
Previous Article in Journal
Enhancement of Leaf Gas Exchange and Primary Metabolites under Carbon Dioxide Enrichment Up-Regulates the Production of Secondary Metabolites in Labisia pumila Seedlings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Asymmetric Aldol Reactions of α,β-Unsaturated Ketoester Substrates Catalyzed by Chiral Diamines

Laboratory of Chemical Synthesis and Pollution Control, College of Chemistry and Chemical Engineering, China West Normal University, Nanchong 637009, Sichuan, China
*
Author to whom correspondence should be addressed.
Molecules 2011, 16(5), 3778-3786; https://doi.org/10.3390/molecules16053778
Submission received: 14 March 2011 / Revised: 16 April 2011 / Accepted: 20 April 2011 / Published: 4 May 2011

Abstract

:
Highly efficient asymmetric aldol reactions between α,β-unsaturated keto esters and acyclic ketones catalyzed by chiral diamines are reported. The corresponding products were obtained in excellent yields with excellent enantioselectivities. The absolute configuration for the product was determined by X-ray analysis. A variety of substrates were tolerable in the present catalytic system.

1. Introduction

The asymmetric aldol reaction affording the corresponding β-hydroxy carbonyl compounds represents one of the most important methods for C-C bond formation [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16]. This transformation has been successfully performed using Lewis acids as catalysts when activated ketones were used as substrates [17]. Recently direct asymmetric aldol reactions were also realized using organocatalytic methods. The organocatalytic aldol reaction has defined the current state of art in asymmetric synthesis and received great attention from the chemical community tahks to the facile preparation of the catalysts, environmentally benign features and mild reaction conditions. Proline and its derivatives, namely the prolinamides, have been found to be versatile catalysts, while chiral amines such as cinchona derived primary amines, 1,2-diaminohexane derivatives and other amines with primary-tertiary amine structure have also been evaluated in direct aldol reactions and significant progress has been achieved [18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35]. Despite the great success achieved in the area, some major challenges still remain. One of the most desirable goals in organic chemistry is the catalytic asymmetric assembly of simple and readily available precursor molecules into structurally complex products. In this context, α,β-unsaturated ketoesters have been used for the construction of functionally diverse organic compounds [36,37,38,39]. α,β-Unsaturated ketoesters should be a suitable aldol acceptor candidates. In fact, Tang has reported an enantioselective construction of bicyclic molecules via a Michael-aldol reaction using α,β-unsaturated ketoesters as starting material [40]. However, while acyclic ketones such as acetone provided acyclic products in excellent yields (99%), enantioselectivities were extremely low (14% ee). In 2008, Zhao reported an enantioselective aldol reaction between α,β-unsaturated ketoesters and ketones [41]. Although cyclic ketones afforded excellent yields and stereoslectivities, in case of acyclic ketones like acetone only a 45% ee was obtained with the reported procedure. However, the methodology provides a rapid access to a functionalities diversed product that are amenable to diverse transformations. To our knowledge, the aldol reaction between α,β-unsaturated ketoesters and cyclic ketones remained elusive. Recently we reported a highly efficient asymmetric aldol reaction between cyclic ketones and aromatic aldehydes catalyzed by cinchona derived amines [42]. We developed a chiral amine 1c featuring a tertiary-primary amine motif and evaluated in the direct aldol reaction [43]. We surmised that structure 1 might be efficient catalyst in the direct aldol reactions between α,β-unsaturated ketoesters and ketones (Figure 1). Herein we discribed our work toward this effort.

2. Result and Discussion

Initially, the reaction between an α,β-unsaturated ketoester and acetone was chosen as the model reaction, and various catalysts including proline, cinchona alkaloid derived amine 1b and the amines 1c -1f developed by us were surveyed in the reaction.
An α,β-unsaturated ketoester has two reaction sites, so the enamine formed from acetone and catalysts can attack either the β-position or the carbonyl position of the ketoester. In the first case, the Michael product may undergo a further aldol reaction forming a cyclic product. However, all the catalysts surveyed in the work ubiquitously formed the aldol product 4. The results are summarized in Table 1.
In the initial investigation, 2a reacted smoothly with acetone at room temperature in the presence of proline (10 mol%) affording the corresponding product in excellent yield albeit with low enantioselectivity (62% yield, 35% ee, Table 1, entry 1). Cinchona alkaloid-derived amine 1b gave excellent yield (95% yield) with good enantioselectivity (75% ee), but a long reaction time was required (36 h). The catalyst 1c which shows highly efficiency in the reaction of ketoesters and acetone afforded an almost quantitative yield (97% yield) and almost the same enantioselectivity (74% ee) in shorter reaction times, as the reaction proceeded to completion in 12 h. The catalyst 1c was thus much more reactive than 1b. Other catalysts 1d-1e provided less satisfactory results compared to 1c. Considering the reactivity of the catalystd, 1c was chosen for further investigation. Acid screening found that TFA was the best candidate (Table 1, entries 3). The best ratio of catalyst to acid was found to be 1:1.5 (Table 1, entries 3, 11–12). The amounts of catalyst used had a significant influence on the reaction outcome. For example, increasing the catalyst loading resulted in higher yield and enantioselectivity. Ten mol % of catalyst was optimal considering both the yield and enantioselectivity.
Solvent screening found acetone was the best choice; other solvents provided enantioselectivities ranging between 27–74% ee. General speaking, aprotic solvents such as DMSO and DMF provided higher selection than other solvents used (Table 2, entries 1–9). Temperature had a significant effect on the reaction, for example, lowering the temperature to −20 °C resulted in a dramatic increase in enantioselectivity (91% ee at −20 °C vs. 83% ee at 25 °C, entry 12 vs. entry 10). Too low a temperature retarded the reaction, and at −40 °C the reaction preceded very slowly affording 83% yield after 5 days although no drop in enantioselectivity was observed.
The optimized protocol was then expanded to a wide variety of α,β-unsaturated ketoesters and the results are summarized in Table 3. Various ketoesters reacted with acetone in excellent yields and enantioselecitvity.
A brief survey of the scope of the reaction with respect to ketoester structure was carried out. Various esters including substrates substituted by electron withdrawing and donating groups at benzene ring attached to the esters uniformly provided the aldol products in near quantitative yields and with excellent enantioselectivities. For example, the introduction of methyl-, fluoro-, chloro-, or methoxy- groups at the para position of the benzene ring appended to the α,β-unsaturated ketoesters did not significantly affect the reactivity and enantioselectivity of the reaction (Table 2, entries 2–5). The steric hindrance in substrates had almost no effect on the stereoselectivity. ortho-Substituted esters such as both 2g and 2h afforded excellent yields and enantioselectivity. The results thus demonstrated good tolerance of substrates.
The absolute configuration of 4f was determined to be R by X-ray diffraction of a single crystal formed in a mixed solvent of ethyl acetate in hexane (Figure 2). While this manuscript was in preparation Chan et al. reported an enantioselective aldol reaction using the same substrates catalyzed by cinchonine derived amine. The absolute configuration of 4f was further verified to be R by the comparison of the retention time reported [44].
Cyclic ketones can also be used in this procedure. It should be noted that only aldol adducts were obtained with excellent enantioslectivity, although low diastereoselectivities were observed under our conditions and no bicyclic products were detected (Scheme 2). This was contrary to the procedure using prolinamides as catalysts [40].

3. Experimental

3.1. General

Unless otherwise noted, material were purchased from commercial suppliers and used without further purification. Acetone, cyclohexanone and cyclopentanone were freshly distilled. Flash column chromatography was performed using 200–300 mesh silica gel. 1HNMR spectra and 13CNMR spectra were recorded on a Bruker AVANCE-400 spectrophotometer. Chiral HPLC was performed on Shimadzu LC-20A with chiral columns (Chirapak AD-H and AS-H columns, Daicel Chemical Ind., Ltd.).

3.2. General Synthetic Pordecure for Aldol Reactions between α,β-unsaturated Ketoesters and Ketones

A solution of 1c (0.01 mmol), TFA (0.015 mmol) and α,β-unsaturated ketoester 2a (0.1 mmol) in acetone (1 mL) was stirred for about 24 h till TLC shows the disappearance of the ester. The reaction was quenched with saturated ammonium chloride solution (10 mL) and extracted with dichloromethane (3 × 15 mL). The combined organic layer was dried (anhydrous Na2SO4) and concentrated in vacuo. The crude product was submitted to column chromatography purification [petroleum ether-ethyl acetate = 8:1 (vv)], and the corresponding product 4a was obtained in 95% yield. 1H-NMR (400 MHz, CDCl3): δ (ppm) 7.36–7.39 (m, 2H), 7.30–7.34 (m, 2H), 7.28–7.26 (m, 1H), 6.86 (d, J =15.8 Hz, 1H), 6.16 (d, J = 15.8 Hz, 1H), 4.01 (s, 1H), 3.81 (s, 3H), 3.27 (d, J =17.4 Hz, 1H), 2.93 (d, J =17.4 Hz, 1H), 2.20 (s, 3H). 13C-NMR (100 MHz, CDCl3): δ 206.7, 174.1, 135.8, 130.7, 128.5, 128.1, 127.9, 126.6, 75.1, 53.1, 51.6, 30.5. [α]D25 = +37.8, (C = 0.24). MS Exact mass calcd. for C14H16O4Na+: 271.0946, Found: 271.0950.

4. Conclusions

In summary, we have demonstrated that enantioselective aldol reactions between α,β-unsaturated ketoesters and acetone can be successfully realized in the presence of chiral diamines and Bronsted acids. Only aldol adducts were formed and no Michael-aldol products were detected.

References and Notes

  1. Mahrwald, R. Modern Aldol Reaction; Wiley-VCH: Weinheim, Germany, 2004; Volumes 1-2. [Google Scholar]
  2. Gröger, H.; Vogl, E.M.; Shibasaki, M. New catalytic concepts for the asymmetric aldol reaction. Chem. Eur. J. 1998, 4, 1137–1141. [Google Scholar]
  3. Machajewski, T.D.; Wong, C.-H.; Lerner, R.A. The catalytic asymmetric aldol reaction. Angew. Chem. Int. Ed. 2000, 39, 1352–1374. [Google Scholar]
  4. Sakthivel, K.; Notz, W.; Bui, T.; Barbas, C.F., III. Amino Acid Catalyzed Direct Asymmetric Aldol Reactions: A Bioorganic Approach to Catalytic Asymmetric Carbon-Carbon Bond-Forming Reactions. J. Am. Chem. Soc. 2001, 123, 5260–5261. [Google Scholar]
  5. Denmark, S.E.; Stavenger, R.A. Asymmetric Catalysis of Aldol Reactions with Chiral Lewis Bases. Acc. Chem. Res. 2000, 33, 432–440. [Google Scholar]
  6. Palomo, C.; Oiarbide, M.; Garcia, J.M. Current Progress in The Asymmetric Aldol Addition Reaction. Chem. Soc. Rev. 2004, 33, 65–75. [Google Scholar]
  7. Kano, T.; Tokuda, J.; Maruoka, K. Design of an Axially Chiral Amino Acid with a Binaphthyl Backbone as an Organocatalyst for a Direct Asymmetric Aldol Reaction. Angew. Chem. Int. Ed. 2005, 44, 3055–3057. [Google Scholar]
  8. Sauer, S.J.; Garnsey, M.R.; Coltart, D.M. Direct Carbon-Carbon Bond Formation via Reductive Soft Enolization: A Kinetically Controlled syn-Aldol Addition of r-Halo Thioesters and Enolizable Aldehydes. J. Am. Chem. Soc. 2010, 132, 13997–13999. [Google Scholar]
  9. Xu, X.-Y.; Tang, Z.; Wang, Y.-Z.; Luo, S.-W.; Cun, L.-F.; Gong, L.-Z. Asymmetric Organocatalytic Direct Aldol Reactions of Ketones with α-Keto Acids and Their Application to the Synthesis of 2-Hydroxy-γ-butyrolactones. J. Org. Chem. 2007, 72, 9905–9913. [Google Scholar]
  10. Yamaoka, Y.; Yamamoto, H. Super Silyl Stereo-Directing Groups for Complete 1,5-syn and -anti Stereoselectivities in The Aldol Reactions of β-Siloxy Methyl Ketones with Aldehydes. J. Am. Chem. Soc. 2010, 132, 5354–5356. [Google Scholar]
  11. List, B.; Lerner, R.A.; Barbas, C.F., III. Proline-Catalyzed Direct Asymmetric Aldol Reactions. J. Am. Chem. Soc. 2000, 122, 2395–2396. [Google Scholar]
  12. Zhou, Y.; Shan., Z. (R)- or (S)-Bi-2-Naphthol Assisted, L-proline Catalyzed Direct Aldol Reaction. Tetrahedron: Asymmetry 2006, 17, 1671–1677. [Google Scholar]
  13. Torii, H.; Nakadai, M.; Ishihara, K.; Saito, S.; Yamamoto, H. Asymmetric Direct Aldol Reaction Assisted by Water and a Proline-derived Tetrazole Catalyst. Angew. Chem. Int. Ed. 2004, 43, 1983–1986. [Google Scholar]
  14. Martin, H.J.; List, B. N-terminal Prolyl-peptides Efficiently Catalyze Enantioselective Aldol and Michael reaction. Synlett 2003, 1901. [Google Scholar]
  15. Tang, Z.; Yang, Z.; Cun, L.; Gong, L.-Z.; Mi, A.-Q.; Jiang, Y.-Z. Small Peptides Catalyze Highly Enantioselective Direct Aldol Reactions of Aldehydes with Hydroxyacetone: Unprecedented Regiocontrol in Aqueous Media. Org. Lett. 2004, 6, 2285–2287. [Google Scholar]
  16. Cordova, A.; Zou, W.; Dziedzic, P.; Ibrahem, I.; Reyes, E.; Xu, Y. Direct Asymmetric Intermolecular Aldol Reactions Catalyzed by Amino Acids and Small Peptides. Chem. Eur. J. 2006, 12, 5383–5397. [Google Scholar]
  17. Xu, Z.-H.; Daka, P.; Wang, H. Primary Amine-Metal Lewis Acid Bifunctional Catalysts: The Application to Asymmetric Direct Aldol Reactions. Chem. Commun. 2009, 44, 6825–6827. [Google Scholar]
  18. Jiang, Z.-Q.; Lu, Y.-X. Direct Asymmetric Aldol Reaction of Acetone with α-ketoesters Catalyzed by Primary-tertiary Diamine Organocatalysts. Tetrahedron Lett. 2010, 51, 1884–1886. [Google Scholar]
  19. Wang, F.; Xiong, Y.; Liu, X.-H.; Feng, X.-M. Asymmetric direct aldol reaction of α-keto esters and acetone catalyzed by bifunctional organocatalysts. Adv. Synth. Catal. 2007, 349, 2665–2668. [Google Scholar]
  20. Liu, J.; Yang, Z.-G.; Wang, Z.; Wang, F.; Chen, X.-H.; Liu, X.-H.; Feng, X.-M.; Su, Z.-S.; Hu, C.-W. Asymmetric Direct Aldol Reaction of Functionalized Ketones Catalyzed by Amine Organocatalysts Based on Bispidine. J. Am. Chem. Soc. 2008, 130, 5654–5655. [Google Scholar]
  21. Luo, S.-Z.; Xu, H.; Li, J.-Y.; Zhang, L.; Chen, J.-P. A Simple Primary-Tertiary Diamine-Bronsted Acid Catalyst for Asymmetric Direct Aldol Reactions of Linear Aliphatic Ketones. J. Am. Chem. Soc. 2007, 129, 3074–3075. [Google Scholar]
  22. McCooey, S.H.; Connon, S.J. Readily Accessible 9-epi-Amino Cinchona Alkaloid Derivatives Promoto Efficient, Highly Enantioselective Additions of Aldehydes and Ketones to Nitroolefins. Org. Lett. 2007, 9, 599–602. [Google Scholar]
  23. Córdova, A.; Zou, W.; Ibrahem, I.; Reyes, E.; Engqvist, M.; Liao, W.-W. Acyclic Amino Acid-Catalyzed Direct Asymmetric Aldol Reactions: Alanine, the Simplest Stereoselective Organocatalyst. Chem. Commun. 2005, 28, 3586–3588. [Google Scholar]
  24. Zhou, J.; Wakchaure, V.; Kraft, P.; List, B. Primary-Amine-Catalyzed Enantioselective Intramolecular Aldolizations. Angew. Chem. Int. Ed. 2008, 47, 7656–7658. [Google Scholar]
  25. Mase, N.; Nakai, Y.; Ohara, N.; Yoda, H.; Takabe, K.; Tanaka, F.; Barbas, C.F., III. Organocatalytic Direct Asymmetric Aldol Reactions in Water. J. Am. Chem. Soc. 2006, 128, 734–735. [Google Scholar]
  26. Saito, S.; Nakadai, M.; Yamamoto, H. Diamine-Protonic Acid Catalysts for Catalytic Asymmetric Aldol Reaction. Synlett 2001, 1245–1248. [Google Scholar]
  27. Eder, U.; Sauer, G.; Wiechert, R. New Type of Asymmetric Cyclization to Optically Active Steroid CD partial Structures. Angew. Chem. Int. Ed. 1971, 10, 496–497. [Google Scholar]
  28. Notz, W.; List, B. Catalytic Asymmetric Synthesis of anti-1,2-Diols. J. Am. Chem. Soc. 2000, 122, 7386–7387. [Google Scholar]
  29. Northrup, A.B.; MacMillan, D.W.C. The First Direct and Enantioselective Cross-Aldol Reaction of Aldehydes. J. Am. Chem. Soc. 2002, 124, 6798–6799. [Google Scholar]
  30. Thayumanavan, R.; Tanaka, F.; Barbas, C.F., III. Direct Organocatalytic Asymmetric Aldol Reactions of α-Amino Aldehydes: Expedient Syntheses of Highly Enantiomerically Enriched anti-β-Hydroxy-α-amino Acids. Org. Lett. 2004, 6, 3541–3544. [Google Scholar]
  31. Ward, D.E.; Jheengut, V.; Akinnusi, O.T. Enantioselective Direct Intermolecular Aldol Reactions with Enantiotopic Group Selectivity and Dynamic Kinetic Resolution. Org. Lett. 2005, 7, 1181–1184. [Google Scholar]
  32. Alcaide, B.; Almendros, P.; Luna, A.; Torres, M.R. Proline-Catalyzed Diastereoselective Direct Aldol Reaction between 4-Oxoazetidine-2-carbaldehydes and Ketones. J. Org. Chem. 2006, 71, 4818–4822. [Google Scholar]
  33. Shen, Z.; Ma, J.; Liu, Y.; Jiao, C.; Li, M.; Zhang, Y. Beta-Cyclodextrin-Immobilized (4S)-Phenoxy-(S)-Proline as a Catalyst for Direct Asymmetric Aldol Reactions. Chirality 2005, 17, 556–558. [Google Scholar]
  34. Samanta, S.; Liu, J.; Dodda, R.; Zhao, C. C2-Symmetric Bisprolinamide as a Highly Efficient Catalyst for Direct Aldol Reaction. Org. Lett. 2005, 7, 5321–5323. [Google Scholar]
  35. Zhong, G.; Fana, J.; Barbas, C.F., III. Amino Alcohol-Catalyzed Direct Asymmetric Aldol Reactions: Enantioselective Synthesis of anti-α-Fluoro-β-Hydroxy Ketones. Tetrahedron Lett. 2004, 45, 5681–5684. [Google Scholar]
  36. Zhou, L.; Lin, L.-L.; Wang, W.-T.; Ji, J.; Liu, X.-H.; Feng, X.-M. Highly Enantioselective Michael Addition of Malonates to β, γ-Unsaturated α-Ketoesters Catalyzed by Chiral N,N′-Dioxide-yttrium(III) Complexes with Convenient Procedure. Chem. Comm. 2010, 46, 3601–3603. [Google Scholar]
  37. Rueping, M.; Nachtsheim, B.J.; Moreth, S.A.; Bolte, M. Asymmetric Brosted Acid Catalysis: Enantioselective Nucleophilic Substitutions and 1,4-Additions. Angew. Chem. Int. Ed. 2008, 47, 593–596. [Google Scholar]
  38. Yao, W.-J.; Wu, L.-J.; Ma, C. Asymmetric Synthesis of Spiro-3,4-dihydropyrans via a Domino Organocatalytic Sequence. Org. Lett. 2010, 12, 2422–2425. [Google Scholar]
  39. Faita, G.; Mella, M.; Toscanini, M.; Desimoni, G. Asymmetric Friedel-Crafts Alkylation of Activated Benzenes with Methyl (E)-2-oxo-4-aryl-3-Butenoates Catalyzed by [Pybox/Sc(OTf)3]. Tetrahedron 2010, 66, 3024–3029. [Google Scholar]
  40. Cao, C.-L.; Chun, X.-L.; Kang, Y.-B.; Tang, Y. Enantioselective Formal [3+3] Annulation for the Direct Construction of Bicyclic Skeletons with Four Stereogenic Centers. Org. Lett. 2007, 9, 4151–4154. [Google Scholar]
  41. Zheng, C.; Wu, Y.; Wang, X.; Zhao, G. Highly enantioselective organocatalyzed construction of quaternary carbon centers via cross-aldol reaction of ketones in water. Adv. Synth. Catal. 2008, 350, 2690–2694. [Google Scholar]
  42. Zheng, B.-L.; Liu, Q.-Z.; Guo, C.-S.; Wang, X.-L.; He, L. Highly Enantioselective Direct Aldol Reaction Catalyzed by Cinchona Derived Primary Amines. Org. Biomol. Chem. 2007, 9, 2913–2915. [Google Scholar]
  43. Liu, Q.-Z.; Wang, X.-L.; Luo, S.-W.; Zheng, B.-L.; Qin, D.-B. Facile Preparation of Optically Pure Diamines and Their Applications in Asymmetric Aldol Reactions. Tetrahedron Lett. 2008, 49, 7434–7437. [Google Scholar]
  44. Li, P.-F.; Zhao, J.-L.; Li, F.-B.; Chan, A.S.C.; Kwong, F.Y. Highly Enantioselective and Efficient Organocatalytic Aldol Reaction of Acetone and β, γ-Unsaturated α-Keto Ester. Org. Lett. 2010, 12, 5616–5619. [Google Scholar]
  45. CCDC 802087 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033; e-mail: [email protected]).
Sample Availability: Samples of the compounds 4a-4j are available from the authors.
Figure 1. The catalysts surveyed in the work.
Figure 1. The catalysts surveyed in the work.
Molecules 16 03778 g001
Scheme 1. Reaction between α,β-unsaturated keto ester 2a and acetone.
Scheme 1. Reaction between α,β-unsaturated keto ester 2a and acetone.
Molecules 16 03778 sch001
Figure 2. X-ray crystal structure of 4f [45].
Figure 2. X-ray crystal structure of 4f [45].
Molecules 16 03778 g002
Scheme 2. Reactions between α,β-unsaturated ketoester and cyclic ketones.
Scheme 2. Reactions between α,β-unsaturated ketoester and cyclic ketones.
Molecules 16 03778 sch002
n = 1, 98% yield, dr = 1:1, 92% and 74% ee; n = 2, 97% yield, dr = 3:1, 94% and 7% ee.
Table 1. The reaction between α,β-unsaturated keto ester and acetone.a
Table 1. The reaction between α,β-unsaturated keto ester and acetone.a
Molecules 16 03778 i001
EntryCat.Additive TimeYield bEe c
11a-486235
21bTFA369875
31cTFA129774
41dTFA129454
51eTFA208830
61fTFA259656
71cHAc129725
81cPhCO2H129719
91cTsOH425079
101cTfOH215782
11 d1cTFA129778
12 e1cTFA129583
13 fS,S-1cTFA188671
14 g1cTFA108778
15 h1cTFA109579
a) Unless otherwise stated, the reaction was carried out with 0.1 mmol of α,β-unsaturated keto- ester in the presence of 10 mol% of catalyst at room temperature with the combination of 20 mol% TFA using acetone as solvent; b) isolated yields; c) optical purity was determined using chiral HPLC; d) 10 mol% of TFA used; e) 15 mol% of TFA used; f) S,S-1c was used as catalyst under identical conditions of entry 12; g) 5 mol% of catalyst 1c in combination with 7.5 mol% TFA was used; h) 20 mol% of catalyst used.
Table 2. Solvent screening. a
Table 2. Solvent screening. a
Molecules 16 03778 i002
EntrySolventTimeYield Ee
1CH2Cl2198665
2Toluene198567
3THF239359
4CH3CN199059
5DMSO199574
6Et2O189652
7DMF189668
8CHCl3189327
9CH3CH2OH188953
10acetone129583
11bacetone219785
12cacetone189591
13dacetone1208391
a Unless otherwise stated, the reaction was carried out with 0.1 mmol of α,β-unsaturated ketoester in the presence of 10 mol% of catalyst at room temperature with the combination of 15 mol% TFA. b Reaction carried out at 0 °C; c Reaction carried at −20 °C; d Reaction carried at −40 °C.
Table 3. The reactions between α,β-unsaturated ketoesters and ketone.a
Table 3. The reactions between α,β-unsaturated ketoesters and ketone.a
Molecules 16 03778 i003
EntryProductR1Yield bEe c
14aH9591
24bp-CH39690
34cp-NO29196
44dp-OMe9593
54ep-F9392
64fp-Br9694
74gp-Cl9192
84hp-CN9695
94im-CH38692
104jm-OMe9192
a Unless otherwise stated, the reaction was carried out with 0.1 mmol of α,β-unsaturated ketoester in the presence of 10 mol% of catalyst using acetone as solvent; b isolated yields; c optical purity was determined using chiral HPLC.

Share and Cite

MDPI and ACS Style

Kan, S.-S.; Li, J.-Z.; Ni, C.-Y.; Liu, Q.-Z.; Kang, T.-R. Asymmetric Aldol Reactions of α,β-Unsaturated Ketoester Substrates Catalyzed by Chiral Diamines. Molecules 2011, 16, 3778-3786. https://doi.org/10.3390/molecules16053778

AMA Style

Kan S-S, Li J-Z, Ni C-Y, Liu Q-Z, Kang T-R. Asymmetric Aldol Reactions of α,β-Unsaturated Ketoester Substrates Catalyzed by Chiral Diamines. Molecules. 2011; 16(5):3778-3786. https://doi.org/10.3390/molecules16053778

Chicago/Turabian Style

Kan, Sha-Sha, Jian-Zhen Li, Cheng-Yan Ni, Quan-Zhong Liu, and Tai-Ran Kang. 2011. "Asymmetric Aldol Reactions of α,β-Unsaturated Ketoester Substrates Catalyzed by Chiral Diamines" Molecules 16, no. 5: 3778-3786. https://doi.org/10.3390/molecules16053778

Article Metrics

Back to TopTop