Next Article in Journal
Evaluation of Marine Microalga Diacronema vlkianum Biomass Fatty Acid Assimilation in Wistar Rats
Next Article in Special Issue
Special Issue: Asymmetric Synthesis 2017
Previous Article in Journal
Kinetics, Mechanism and Theoretical Studies of Norbornene-Ethylene Alternating Copolymerization Catalyzed by Organopalladium(II) Complexes Bearing Hemilabile α-Amino–pyridine
Previous Article in Special Issue
Recent Advances in Substrate-Controlled Asymmetric Cyclization for Natural Product Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enantioselective Michael Addition of Cyclic β-Diones to α,β-Unsaturated Enones Catalyzed by Quinine-Based Organocatalysts

1
College of Chemistry and Environment Protection Engineering, Southwest Minzu University, Chengdu 610041, China
2
Faculty of Geosciences and Environmental Engineering, Southwest Jiaotong University, Chengdu 610031, China
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(7), 1096; https://doi.org/10.3390/molecules22071096
Submission received: 16 May 2017 / Revised: 25 June 2017 / Accepted: 26 June 2017 / Published: 30 June 2017
(This article belongs to the Special Issue Asymmetric Synthesis 2017)

Abstract

:
An enantioselective (52–98% ee) Michael addition between cyclic β-diones and α,β-unsaturated enones was established in the presence of quinine-based primary amine or squaramide. A variety of cinnamones were smoothly converted into the desired 3,4-dihydropyrans in moderate to high yields (63–99%). Chalcones were also suitable acceptors and gave rise to the expected adducts in satisfactory yields (31–99%). The resulting adducts readily underwent further modification to form fused 4H-pyran or 2,3-dihydrofuran.

Graphical Abstract

1. Introduction

The Michael addition of α,β-unsaturated compounds is an atom-economic carbon-carbon bond-forming reaction in organic synthesis, and the development of the enantioselective catalytic approach for this transformation, has attracted intensive attention [1,2,3,4]. Among the often-used acceptors, unactivated α,β-unsaturated enones always exhibit relatively sluggish reactivity and emerge as a class of historically challenging substrates for metal- and organocatalytic approaches [5,6,7,8,9,10,11]. In this context, the elegantly-designed chiral primary amines, especially those based on cinchona alkaloids, provide a particularly efficient LUMO-lowering (LUMO: lowest unoccupied molecular orbital) activation mode through the formation of iminium ions with these unsaturated enones [12,13,14]. Therefore, a broad range of conjugate additions of α,β-unsaturated enones with various different nucleophiles have been successfully established with constantly high enantiocontrol [15,16,17,18,19]. However, the asymmetric Michael addition of cyclic 1,3-dicarbonyl compounds [20,21,22,23,24,25,26,27], except for 4-hydroxycoumarin and its analogues [7,28,29,30,31,32,33,34,35,36], to α,β-unsaturated enones, especially chalcones, generally draws less attention in comparison with other type of donors [37,38,39], albeit the adducts of such a conjugate addition reaction are versatile precursors to construct several classes of compounds possessing enormous bioactivities [40,41]. Liu and Feng have successfully developed an efficient Michael addition between dimedone and cinnamones employing the unmodified chiral diphenylethylenediamine (DPEN) [37]. In contrast, only moderate enantioselectivity was obtained for the Michael addition of dimedone to unfunctionalized chalcone according to Singh’s protocol [38]. Consequently, the development of the enantioselective Michael addition of cyclic β-diones to α,β-unsaturated enone is still highly sought.
Based on our continuous interest in asymmetric Michael reactions involving α,β-unsaturated enones [42,43,44,45], herein we would like to further extend the scope of donor to cyclic β-diones [20,21,22,23,46,47,48,49]. Dimedone and its analogues smoothly react with a variety of α,β-unsaturated enones, furnishing the corresponding adducts in good yields and high levels of optical purities. The synthetic potential of the desired Michael adduct is demonstrated by the easy formation of enantioenriched 4H-pyran and 2,3-dihydrofuran.

2. Results and Discussion

We were pleased to find that the Michael addition of dimedone 1a to cinnamone 2a proceeded smoothly in the presence of 9-amino(9-deoxy)-epi-quinine 3a (Figure 1) in combination with a series of different acid co-catalysts. It was documented that the acid co-catalyst had a great influence on the yield and enantioinduction [16]. The aromatic carboxylic acids displayed superior catalytic effect compared with sulfonic acid and aliphatic acids (Table 1, entries 4–9 vs. entries 1–3). The desired 3,4-dihydropyran 4a was generated with good to excellent enantioselectivities (87–90% ee) in the presence of various aromatic acids. In contrast, salicylic acid (SA) afforded an optimal yield (99%) and a superior enantioselectivity (90% ee) (entry 9 vs. entries 4–8) [50]. Having identified salicylic acid as the preferential acid co-catalyst, we turned our attention to evaluate the effect of other primary amines 3b and 3c (Figure 1) derived from naturally occurring cinchona alkaloids [51]. Both 3b and 3c delivered the expected 3,4-dihydropyran 4a possessing opposite configurations to the adduct afforded by 3a (entries 10 and 11). Moreover, these two pseudo-enantiomers displayed poorer catalytic activities and enantioselectivities compared with 9-amino(9-deoxy)-epi-quinine 3a (entries 10 and 11 vs. entry 9). Subsequently, we examined the effect of the solvent with a combination of 3a and salicylic acid. Tetrahydrofuran (THF) emerged as the favorable one in terms of reactivity and enantioselectivity (entry 15 vs. entries 9, 12–14). Notably, the model process proceeded equally smoothly when the amount of cinnamone was decreased to 1.2 equivalents (entry 16 vs. entry 15). Reducing the reaction temperature (0 °C) led to a slightly higher enantioselectivity (97% ee) (entry 17).
With the optimal reaction conditions in hand, various cinnamones 2 were treated with dimedone 1a to determine the scope and generality of this Michael addition. As presented in Table 2, the electronic property exerted marginal impact on this asymmetric process. The electron-deficient cinnamones 2c2f generally provided the corresponding 3,4-dihydropyrans in slightly higher chemical yields, in contrast with the electron-rich acceptors 2g and 2h (Table 2, entries 3–6 vs. entries 7 and 8). Meanwhile, all these enones gave rise to the desired adducts with excellent enantioselectivities (96–97% ee) irrespective of electronic nature (entries 3–8). On the other hand, the steric hindrance slightly impaired the reactivity of this conjugate addition reaction. In this context, the ortho-substituted enone 2b afforded somewhat poorer conversion (87% yield) in comparison with other electron-poor cinnamones 2c2f (96–99% yield) (entry 2 vs. entries 3–6). Gratifyingly, 2i and 2j, both possessing a bulky naphthyl group at the β-site, were also compatible with this catalytic system (entries 9 and 10). The resulting adducts 4ai and 4aj were formed in excellent yields and with high levels of enantioselectivities. The heteroaromatic enones 2k and 2l were all suitable partners for this Michael reaction (entries 11 and 12). The alkyl-substituted enones 2m and 2n were found to react relatively slowly with dimedone, however, synthetically useful yields and satisfactory enantiocontrol were still obtained (entries 13 and 14). Remarkably, cyclic enone 2o was also a competent acceptor, furnishing the bridged-ring compound 4ao in 91% yield and 98% ee (entry 15) [37]. The ketone substituent (R2) could also be varied from methyl group to ethyl group. Although relatively poorer conversion was detected, excellent enantioselectivity was maintained for this sterically more hindered acceptor (entry 16). It seemed to be an effect of increased steric bulk on the ketone, retarding the acceptor to approach the catalyst, thereby slowing down the reaction rate. On the other hand, the unsubstituted cyclic β-dione, 1,3-cyclohexanedione, was also tolerated by this catalytic system. Acceptable yields (67–78%) and high degrees of enantiomeric excesses (94–96% ee) were successfully achieved (entries 17–19), despite its relatively lower reactivity in contrast with dimedone [20,52]. Notably, a one mmole-scale Michael addition of cinnamone 2a and dimedone 1a was performed under optimal reaction conditions. Excellent chemical yield (95%) and enantiopurity (94% ee) were both obtained (entry 1).
Having identified cinnamones as the suitable acceptors, we successively turned our attention to chalcone (Scheme 1), a class of challenging substrates for iminium ion activation [16]. Different than cinnamones, the bulky benzene group might retard the later annulation process, therefore only the initial Michael adduct was accessed. Considering the unstability of the Michael adduct due to aerobic oxidation [53], a subsequent acetylation was conducted after the initial conjugate addition in a one-pot manner. To our disappointment, the titled process allowed access to the final acetyl derivative 6aa in fairly low yield (<20%), even when the initial Michael addition was performed at room temperature.
Fortunately, we finally found that the Michael addition of chalcone worked properly in the presence of squaramide 7 derived from quinine (see supporting material) [54,55]. As outlined in Table 3, this Michael addition was independent of the electronic nature of the substituents on the aromatic rings. Both the electron-rich acceptors 5b and 5f and the electron-deficient acceptors 5c and 5g generated the expected adducts in satisfactory yields and excellent optical purities (Table 3, entries 2 and 6 vs. entries 3 and 7). Moreover, the steric hindrance exerted influence on this Michael addition to a certain extent. The enone 5e possessing a naphthyl group afforded relatively lower isolated yield (68%) even after a prolonged reaction time, albeit accompanied by outstanding enantioselectivity (entry 5). Heteroaromatic chalcones 5d and 5h were also favorable partners, giving rise to the final acetyl derivatives with high levels of enantiopurities (entries 4 and 8). Except for dimedone, 1,3-cyclohexanedione 1b was a competent donor as well, albeit a longer reaction time was required in order to achieve complete conversion (entry 9). In contrast with Singh’s precedent study (72% ee for 6aa) [38], our protocol efficiently improved the enantioselectivity and displayed a wide substrate generality for this Michael addition of cyclic β-dione to chalcone [56]. Moreover, a one mmole-scale Michael addition of chalcone 5a with dimedone 1a proceeded smoothly as well. The expected acetyl derivate 6aa was formed in an almost quantitative yield and with satisfactory enantioselectivity (entry 1). The alkyl-substituted enone 5i was also a suitable acceptor, albeit unsatisfactory enantioselectivity was obtained for the resulting Michael adduct (entry 10).
The five-membered cyclic dione, 1,3-cyclopentadione 1c, was also tolerated by our catalytic protocol (Scheme 2). However, it proved to be an inferior donor in terms of reactivity and enantioselectivity, in contrast with the six-membered cyclic dione. The related acetyl derivative 6ca was obtained in an unsatisfactory yield and with moderate optical purity.
To demonstrate the synthetic potential of this Michael reaction, product modification was performed on the Michael adducts. 3,4-Dihydropyran 4aa readily underwent a dehydrating procedure to afford 4H-pyran 8 without the loss of optical purity (Scheme 3a) [20]. The Michael adduct of chalcone could be utilized for the facile preparation of the biologically interesting 2,3-dihydrofuran 9 via a successive stereoselective oxidative cyclization process (Scheme 3b) [57]. The fused 2,3-dihydrofuran 9 was obtained as a single trans-diastereomer in a synthetically useful yield and with excellent enantioselectivity.
The absolute configuration of the Michael adduct 4ad (Table 2, entry 4) was determined to be S via comparison of the optical rotation value and HPLC traces with that of the previous literature reports [37]. On the other hand, the absolute configuration of 6aa (Table 3, entry 1) was established as R by the analysis of the optical rotation value with Singh’s protocol [38]. To account for the observed stereochemical outcome of these Michael reactions, the corresponding transition state models were proposed and described in Scheme 4. The primary amine motif of 9-amino(9-deoxy)-epi-quinine 3a was engaged in iminium formation with the carbonyl group of benzalacetone 1a. Meanwhile, dimedone was deprotonated by the tertiary amine moiety of aminocatalyst 3a and orientated via hydrogen-bonding, thereby leading to a favorable attack toward the si-face of cinnamon 1a. As a result, the desired S-configured product 4a was obtained. On the other hand, chalcone 5a was efficiently activated via hydrogen-bonding interactions between the NH moiety of the squaramide 7 and the carbonyl group of chalcone. Furthermore, the re-face approach of dimedone was induced by the tertiary amine of the squaramide 7 and led to the formation of the major stereoisomer with the R configuration [58].

3. Materials and Methods

3.1. General Remarks

1H- and 13C-NMR spectra were recorded on Varian 400 MHz spectrometers. Chemical shifts (δ) are reported in ppm downfield from CDCl3 (δ = 7.26 ppm) for 1H-NMR and relative to the central CDCl3 resonance (δ = 77.0 ppm) for 13C-NMR spectroscopy. Coupling constants (J) are given in Hz. ESI-HRMS spectrometry was performed with a Bruker Daltonics LCQDECA ion trap mass spectrometer. Enantiomeric excess was determined by HPLC analysis on Chiralpak AD-H, OD-H, and IC columns in comparison with the authentic racemates. Optical rotation data were recorded on a Rudolph Autopol I automatic polarimeter. Commercial grade solvents were dried and purified by standard procedures as specified in reference [59]. THF (AR grade) was used as received. All other reagents were purchased from commercial sources and were used without further purification.

3.2. General Procedure for the Asymmetric Michael Reaction of Cinnamones

9-Amino-epi-quinine 3a (6.5 mg, 0.02 mmol), α,β-unsaturated enones (0.12 mmol), dimedone (14.0 mg, 0.1 mmol), and salicylic acid (4.9 mg, 0.04 mmol) were dissolved in THF (1 mL) without stirring. Once the solution was cooled down to 0 °C, the reaction mixture was stirred for 96 h. After the solvent was removed in vacuo, the residue was purified by flash chromatography on silica gel (EtOAc/petroleum ether) to afford the desired 3,4-dihydropyran.
2-Hydroxy-2,7,7-trimethyl-4-phenyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4aa) [37]. Colorless oil; 99% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.30–7.23 (m, 2H), 7.19–7.12 (m, 3H), 4.03 (br s, 0.6H), 3.84 (t, J = 8.4 Hz, 0.6H), 3.31 (br s, 0.4H), 3.16–3.12 (m, 0.4H), 2.50–2.15 (m, 6H), 1.48 (s, 1.7H), 1.46 (s, 1.3H), 1.19 (s, 1.7H), 1.16 (s, 1.3H), 1.11 (s, 1.7H), 1.07 (s, 1.3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.3, 196.9, 169.6, 168.6, 144.9, 142.9, 128.8, 128.2, 127.8, 127.7, 126.9, 126.8, 126.5, 125.7, 113.0, 110.5, 99.8, 99.2, 50.6, 50.5, 42.9, 42.8, 42.7, 40.5, 33.9, 32.8, 31.9, 31.4, 29.5, 28.6, 28.3, 27.8, 27.4, 27.1; 97% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 4.820 min, tmajor = 7.627 min; [ α ] D 20 = −4.2° (c = 0.028, EtOH).
4-(2-Chlorophenyl)-2-hydroxy-2,7,7-trimethyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ab) [60]. Colorless oil; 87% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.36–7.29 (m, 1H), 7.10–7.05 (m, 3H), 4.36–4.24 (m, 1H), 3.83 (br s, 0.5H), 3.43 (br s, 0.5H), 2.49–2.11 (m, 6H), 1.49 (s, 1.4H), 1.48 (s, 1.6H), 1.18 (s, 1.6H), 1.16 (s, 1.4H), 1.09 (s, 1.6H), 1.07 (s, 1.4H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 196.7, 196.5, 169.7, 140.5, 133.6, 129.9, 129.5, 127.8, 127.6, 126.9, 126.7, 126.6, 112.6, 110.4, 99.7, 98.1, 50.7, 50.6, 42.9, 42.8, 37.7, 31.9, 31.5, 29.4, 28.9, 28.2, 27.9, 27.6, 27.3; 91% ee was determined by HPLC on AD-H column, hexane/i-propanol (90/10), 1.0 mL/min, UV 254 nm, tminor = 7.703 min, tmajor = 9.353 min; [ α ] D 20 = −32.8° (c = 0.021, EtOH).
4-(3-chlorophenyl)-2-hydroxy-2,7,7-trimethyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ac) [37]. Colorless oil; 99% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.19–7.09 (m, 3H), 7.02 (t, J = 7.8 Hz, 1H), 4.34 (br s, 0.5H), 3.89 (t, J = 4.8 Hz, 0.5H), 3.82–3.77 (m, 1H), 2.48–2.09 (m, 6H), 1.44 (s, 3H), 1.18 (s, 1.3H), 1.15 (s, 1.7H), 1.09 (s, 1.3H), 1.07 (s, 1.7H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.3, 197.0, 169.9, 169.1, 147.3, 145.9, 134.2, 133.9, 129.6, 129.5, 129.1, 127.9, 127.5, 127.1, 126.4, 125.9, 125.27, 125.25, 112.6, 110.3, 99.6, 98.1, 50.6, 50.5, 42.9, 42.8, 42.5, 40.5, 33.9, 33.3, 31.9, 31.5, 29.9, 29.5, 28.6, 28.3, 27.8, 27.4, 26.9; 96% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 5.267 min, tmajor = 7.583 min; [ α ] D 20 = +5.5° (c = 0.039, EtOH).
4-(4-Chlorophenyl)-2-hydroxy-2,7,7-trimethyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ad) [37]. Colorless oil; 99% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.23 (d, J = 8.8 Hz, 1H), 7.21 (d, J = 8.8 Hz, 1H), 7.11 (d, J = 8.4 Hz, 1H), 7.07 (d, J = 8.0 Hz, 1H), 3.95 (d, J = 4.4 Hz, 0.5H), 3.81 (d, J = 8.6 Hz, 0.5H), 3.16–3.09 (m, 0.5H), 2.95 (br s, 0.4H), 2.49–2.17 (m, 6H), 1.53 (s, 1.5H), 1.50 (s, 1.5H), 1.18 (s, 1.5H), 1.15 (s, 1.5H), 1.10 (s, 1.4H), 1.08 (s, 1.6H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 196.9, 196.6, 169.3, 168.3, 143.5, 141.9, 132.1, 131.3, 129.1, 128.8, 128.49, 128.47, 128.3, 127.9, 112.9, 110.5, 99.5, 97.9, 50.74, 50.72, 42.9, 42.8, 42.5, 40.3, 33.6, 32.6, 32.0, 31.5, 29.5, 28.7, 28.3, 28.1, 27.5, 27.4; 97% ee was determined by HPLC on OD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 5.493 min, tmajor = 8.417 min; [ α ] D 20 = +13.8° (c = 0.039, EtOH), [α]D2° = +10.5° (c = 0.039, DCM).
4-(4-Fluorophenyl)-2-hydroxy-2,7,7-trimethyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ae) [37]. Colorless oil; 96% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.13 (t, J = 6.4 Hz, 1H), 7.08 (t, J = 6.4 Hz, 1H), 6.94 (t, J = 9.4 Hz, 1H), 6.92 (t, J = 8.8 Hz, 1H), 3.94 (br s, 0.5H), 3.82 (dd, J = 9.8, 8.2 Hz, 0.5H), 3.51–3.27 (m, 1H), 2.48–2.11 (m, 6H), 1.47 (s, 3H), 1.78 (s, 1.4H), 1.14 (s, 1.6H), 1.09 (s, 1.4H), 1.07 (s, 1.6H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.2, 196.9, 169.5, 168.6, 161.2 (d, J1C-F = 243.3 Hz), 160.9 (d, J1C-F = 241.8 Hz), 140.6 (d, J4C-F = 3.2 Hz), 139.0 (d, J4C-F = 3.3 Hz), 129.2, 129.1, 128.5 (d, J3C-F = 7.8 Hz), 128.2 (d, J3C-F = 7.8 Hz), 115.4 (d, J2C-F = 21.1 Hz), 115.0 (d, J2C-F = 21.2 Hz), 114.6, 114.4, 113.0, 110.7, 99.6, 98.1, 60.4, 50.7, 50.6, 42.9, 42.8, 42.7, 40.6, 33.4, 32.6, 31.9, 31.4, 29.5, 28.6, 28.3, 27.9, 27.3, 27.0, 20.9, 14.1; 96% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 5.043 min, tmajor = 9.330 min; [ α ] D 20 = −5.4° (c = 0.041, EtOH).
4-(4-Bromophenyl)-2-hydroxy-2,7,7-trimethyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4af) [37]. Colorless oil; 99% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.37 (d, J = 8.4 Hz, 1H), 7.35 (d, J = 8.4 Hz, 1H), 7.04 (d, J = 8.0 Hz, 1H), 7.00 (d, J = 8.4 Hz, 1H), 3.90 (t, J = 5.4 Hz, 0.5H), 3.72 (pseudo triple, J = 5.4 Hz, 0.6H), 3.50 (br s, 0.5H), 3.19 (br s, 0.4H), 2.48–2.07 (m, 6H), 1.49 (s, 1.6H), 1.48 (s, 1.4H), 1.17 (s, 1.4H), 1.14 (s, 1.6H), 1.09 (s, 1.4H), 1.07 (s, 1.6H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.0, 196.7, 169.4, 168.4, 144.1, 142.6, 131.6, 131.4, 130.9, 129.6, 128.9, 128.7, 120.0, 119.4, 112.8, 110.4, 99.5, 97.9, 50.7, 50.6, 42.9, 42.8, 42.4, 40.3, 33.7, 32.8, 31.9, 31.5, 29.5, 28.7, 28.3, 28.1, 27.4, 27.3; 97% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 5.720 min, tmajor = 10.570 min; [ α ] D 20 = +13.9° (c = 0.010, EtOH).
2-Hydroxy-2,7,7-trimethyl-4-(p-tolyl)-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ag) [60]. Colorless oil; 95% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.16–7.01 (m, 4H), 3.99 (br s, 0.6H), 3.80 (pseudo triple, J = 8.8 Hz, 0.6H), 3.33 (pseudo double, J = 9.6 Hz, 0.8H), 2.50–2.13 (m, 9H), 1.47 (s, 3H), 1.19 (s, 1.7H), 1.15 (s, 1.3H), 1.10 (s, 1.7H), 1.07(s, 1.3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.1, 196.7, 169.2, 169.1, 141.8, 139.6, 136.2, 135.1, 129.8, 129.0, 128.7, 127.5, 126.8, 126.7, 113.3, 110.5, 99.7, 98.1, 50.7, 42.9, 42.8, 42.7, 40.3, 33.6, 31.9, 31.5, 29.5, 28.8, 28.3, 27.9, 27.4, 27.3, 21.0, 20.9; 97% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 5.313 min, tmajor = 7.733 min; [ α ] D 20 = +6.2° (c = 0.041, EtOH).
2-Hydroxy-4-(4-methoxyphenyl)-2,7,7-trimethyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ah) [37]. Colorless oil; 89% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.10 (d, J = 8.8 Hz, 1.1H), 7.05 (d, J = 8.8 Hz, 0.9H), 6.82 (d, J = 7.6 Hz, 1.1H), 6.79 (d, J = 8.4 Hz, 0.9H), 3.99 (br s, 0.6H), 3.80 (pseudo triple, J = 8.8 Hz, 0.6H), 3.75 (s, 3H), 3.59 (br s, 0.4H), 3.31 (br s, 0.6H), 2.49–2.11 (m, 6H), 1.47 (s, 1.7H), 1.46 (s, 1.3H), 1.19 (s, 1.7H), 1.15 (s, 1.3H), 1.10 (s, 1.7H), 1.07 (s, 1.3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.0, 196.7, 169.1, 167.9, 158.2, 157.6, 136.8, 134.4, 128.7, 127.9, 127.8, 114.5, 113.8, 113.4, 113.3, 110.6, 99.7, 88.1, 55.2, 55.1, 50.8, 42.9, 42.8, 42.7, 40.1, 33.2, 31.9, 31.49, 31.47, 29.5, 28.8, 28.3, 27.9, 27.5; 97% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 6.353 min, tmajor = 12.270 min; [ α ] D 20 = +6.5° (c = 0.030, EtOH).
2-Hydroxy-2,7,7-trimethyl-4-(naphthalen-1-yl)-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ai). Colorless oil; 89% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 8.18 (d, J = 8.0 Hz, 0.3H), 8.13 (d, J = 8.4 Hz, 0.7H), 7.90 (d, J = 8.0 Hz, 0.7H), 7.85 (d, J = 7.6 Hz, 0.3H), 7.74 (d, J = 8.4 Hz, 0.7H), 7.68 (d, J = 8.0 Hz, 0.3H), 7.65–7.54 (m, 2.5H), 7.34 (t, J = 7.2 Hz, 1H), 7.27 (d, J = 5.6 Hz, 0.5H), 4.82 (d, J = 6.8 Hz, 0.8H), 4.70 (br s, 0.5H), 3.39 (br s, 0.8H), 2.63–2.23 (m, 6H), 1.47 (s, 3H), 1.27 (s, 2H), 1.24 (s, 1H), 1.17 (s, 2H), 1.12 (s, 1H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 196.9, 196.7, 169.7, 138.6, 134.7, 130.8, 129.3, 128.9, 127.9, 126.2, 125.8, 125.4, 125.3, 125.2, 123.3, 122.5, 118.8, 117.1, 110.2, 99.5, 98.3, 50.8, 50.7, 43.1, 42.9, 37.9, 32.1, 31.5, 29.4, 29.2, 28.0, 27.9, 27.83, 27.80; ESI-HRMS calcd. for C22H24O3 + H+ 337.1804, found 337.1798; 96% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 4.927 min, tmajor = 6.790 min; [ α ] D 20 = −88.9° (c = 0.045, EtOH).
2-Hydroxy-2,7,7-trimethyl-4-(naphthalen-2-yl)-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4aj) [37]. Colorless oil; 98% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.81–7.69 (m, 3H), 7.59 (s, 0.5H), 7.55 (s, 0.5H), 7.44–7.35 (m, 2.5H), 7.27–7.26 (m, 0.5H), 4.19 (br s, 0.5H), 4.01 (pseudo triple, J = 8.6 Hz, 0.5H), 3.29–3.17 (m, 1H), 2.57–2.17 (m, 6H), 1.49 (s, 3H), 1.26 (s, 2H), 1.19 (s, 1H), 1.13 (s, 1.6H), 1.09 (s, 1.4H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.2, 196.9, 169.7, 168.6, 142.3, 140.5, 133.6, 133.5, 132.3, 132.1, 128.8, 127.9, 127.7, 127.53, 127.52, 127.4, 126.1, 125.7, 125.65, 125.57, 125.52, 125.4, 124.9, 124.8, 113.0, 110.4, 99.8, 98.2, 50.7, 50.6, 42.9, 42.8, 42.5, 40.0, 34.1, 32.9, 32.0, 31.5, 29.9, 29.5, 28.7, 28.4, 27.9, 27.5, 27.2; 98% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 7.073 min, tmajor = 13.053 min; [ α ] D 20 = +60.2° (c = 0.051, EtOH).
4-(Furan-2-yl)-2-hydroxy-2,7,7-trimethyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ak) [37]. Colorless oil; 78% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.36 (s, 0.6H), 7.29 (s, 0.4H), 6.28 (dd, J = 3.0, 1.8 Hz, 1H), 6.02 (d, J = 3.2 Hz, 0.7H), 5.95 (d, J = 2.0 Hz, 0.3H), 4.15 (d, J = 6.8 Hz, 1H ), 4.04 (br s, 1H), 2.55–2.22 (m, 6H), 1.55 (s, 2.2 H), 1.41 (s, 0.8H), 1.17 (s, 3H), 1.11 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 196.7, 168.9, 155.5, 141.9, 140.3, 110.6, 110.3, 108.4, 106.0, 105.3, 98.3, 50.7, 42.8, 35.9, 32.0, 28.6, 28.2, 27.9, 26.1; 95% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 5.327 min, tmajor = 6.257 min; [ α ] D 20 = −12.9° (c = 0.015, EtOH).
2-Hydroxy-2,7,7-trimethyl-4-(thiophen-2-yl)-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4al). Brown oil; 97% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.17 (d, J = 4.8 Hz, 0.6H), 7.06 (d, J = 4.8 Hz, 0.6H), 6.89–6.87 (m, 1H), 6.81 (s, 1H), 4.32 (d, J = 5.6 Hz, 0.6H), 4.23 (pseudo triple, J = 8.0 Hz, 0.4H), 3.53 (br s, 1H), 2.51-2.11 (m, 6H), 1.52 (s, 2H), 1.46 (s, 1H), 1.19 (s, 2H), 1.16 (s, 1H), 1.10 (s, 2H), 1.07 (s, 1H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 196.8, 196.7, 169.1, 168.1, 148.7, 147.2, 127.0, 126.4, 124.6, 124.2, 123.8, 123.7, 123.2, 122.4, 112.9, 110.7, 99.4, 98.4, 50.6, 42.8, 42.7, 39.9, 31.9, 31.4, 29.4, 29.3, 28.5, 28.4, 27.7, 27.66, 27.61, 27.3; ESI-HRMS calcd. for C16H20O3S + H+ 293.1211, found 293.1206; 91% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 5.850 min, tmajor = 7.423 min; [ α ] D 20 = −18.9° (c = 0.047, EtOH).
2-Hydroxy-2,4,7,7-tetramethyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4am). Colorless oil; 69% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 3.66–3.35 (m, 1H), 2.77–2.66 (m, 1H), 2.29–2.06 (m, 6H), 1.55 (s, 1.4H), 1.50 (s, 1.6H), 1.23 (d, J = 7.6 Hz, 2H), 1.21 (d, J = 7.6 Hz, 1H), 1.05 (s, 3H), 1.02 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 213.5, 198.5, 197.9, 170.7, 166.7, 166.5, 117.1, 114.9, 114.3, 99.1, 97.8, 51.3, 51.2, 51.1, 49.2, 43.2, 42.8, 42.7, 41.5, 39.2, 31.9, 31.4, 29.8, 29.2, 28.4, 28.1, 28.0, 27.2, 26.9, 24.3, 22.9, 22.1, 19.7, 19.4, 18.5; ESI-HRMS calcd. for C13H20O3 + H+ 225.1491, found 225.1485; 91% ee was determined by HPLC on IC column, hexane/i-propanol (95/5), 1.0 mL/min, UV 254 nm, tminor = 49.740 min, tmajor = 77.910 min; [ α ] D 20 = −5.5° (c = 0.017, EtOH).
4-Butyl-2-hydroxy-2,7,7-trimethyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4an). Colorless oil; 75% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 3.19–3.10 (m, 1H), 2.93–2.87 (m, 0.5H), 2.64 (s, 0.3H), 2.59 (s, 0.3H), 2.30–2.09 (m, 6H), 2.01–1.89 (m, 1H), 1.79–1.73 (m, 1H), 1.65–1.45 (m, 3H), 1.35–1.19 (m, 4H), 1.06–1.02 (m, 6H), 0.89–0.81 (m, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 213.8, 198.6, 197.8, 171.5, 167.1, 166.7, 115.7, 114.2, 113.9, 99.3, 97.9, 51.4, 51.3, 51.2, 48.3, 43.2, 42.9, 42.8, 38.2, 35.1, 32.1, 31.8, 31.7, 31.3, 30.9, 30.5, 29.8, 29.7, 29.4, 29.3, 28.5, 28.3, 28.2, 28.1, 28.0, 27.9, 27.1, 26.8, 22.8, 22.7, 22.4, 14.1, 14.0; ESI-HRMS calcd. for C16H26O3 + H+ 267.1960, found 267.1964; 94% ee was determined by HPLC on IC column, hexane/i-propanol (70/30), 1.0 mL/min, UV 254 nm, tminor = 4.440 min, tmajor = 8.093 min; [ α ] D 20 = −13.0° (c = 0.034, EtOH).
2-Hydroxy-9,9-dimethyl-2,3,4,5,6,8,9,10-octahydro-7H-2,6-methanobenzo[b]oxocin-7-one (4ao) [37]. White solid; 91% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 4.46 (s, 1H ), 3.15 (s, 1H), 2.29 (s, 2H), 2.19 (s, 2H), 2.02 (d, J = 12.8 Hz, 1H), 1.93 (d, J = 12.4 Hz, 1H), 1.74 (d, J = 15.2 Hz, 1H), 1.16 (dd, J = 13.0, 3.8 Hz, 1H), 1.59 (d, J = 11.2 Hz, 2H), 1.45–1.33 (m, 2H), 1.04 (s, 6H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 196.7, 171.2, 112.3, 101.3, 50.3, 42.0, 38.7, 36.2, 32.3, 28.5, 28.4, 28.2, 26.9, 19.2; 98% ee was determined by HPLC on IC column, hexane/i-propanol (90/10), 1.0 mL/min, UV 254 nm, tmajor = 12.987 min, tminor = 14.423 min; [ α ] D 20 = +4.7° (c = 0.023, EtOH).
2-Ethyl-2-hydroxy-7,7-dimethyl-4-phenyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ap) [37]. Colorless oil; 63% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.24–7.16 (m, 2.5H), 7.12 (d, J = 7.6 Hz, 1H), 7.09–7.06 (m, 1.5H), 3.99 (br s, 0.5H), 3.76 (pseudo triple, J = 9.0 Hz, 0.5H), 3.15–3.00 (m, 1H), 2.44–2.09 (m, 6H), 1.69–1.63 (m, 2H), 1.13 (s, 1.6H), 1.09 (s, 1.4H), 1.04 (s, 1.6H), 1.00 (s, 1.4H), 0.89 (0.86) (t, J = 7.6 Hz, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 196.9, 196.5, 169.3, 168.0, 145.1, 142.9, 129.0, 128.3, 126.9, 126.7, 125.8, 113.3, 110.4, 101.3, 99.7, 50.8, 42.9, 42.8, 40.3, 38.0, 33.9, 33.6, 33.2, 32.0, 31.9, 31.5, 29.5, 28.9, 28.3, 27.5, 7.3, 7.2; 98% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 5.247 min, tmajor = 11.667 min; [ α ] D 20 = −5.4° (c = 0.018, EtOH).
2-Hydroxy-2-methyl-4-phenyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4ba) [60]. White solid; 71% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.28 (t, J = 7.2 Hz, 1H), 7.24 (d, J = 7.2 Hz, 1H), 7.19–7.12 (m, 3H), 4.02 (br s, 0.5H), 3.84 (pseudo triple, J = 8.8 Hz, 0.5H), 3.36–3.31(m, 0.5H), 2.63–2.45 (m, 2H), 2.42–2.30 (m, 2H), 2.27–2.15 (m, 2H), 2.11–1.95 (m, 2H), 1.47 (s, 1.5H), 1.45 (s, 1.5H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.2, 196.9, 171.1, 170.1, 144.8, 142.7, 128.9, 128.3, 127.9, 127.7, 126.79, 126.77, 126.6, 125.8, 114.4, 111.6, 99.6, 97.9, 42.8, 40.4, 36.9, 33.9, 32.5, 29.3, 29.2, 27.9, 27.2, 20.8, 20.2; 94% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 5.343 min, tmajor = 6.693 min; [ α ] D 20 = −8.2° (c = 0.019, EtOH).
4-(4-chlorophenyl)-2-hydroxy-2-methyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4bd): White solid; 78% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.25 (d, J = 7.2 Hz, 0.6H), 7.21 (d, J = 8.8 Hz, 1.5H), 7.11 (d, J = 8.8 Hz, 0.6H), 7.07 (d, J = 7.6 Hz, 1.4H), 3.95 (pseudo triple, J = 4.8 Hz, 0.5H), 3.82 (pseudo triple, J = 9.0 Hz, 0.5H), 3.33 (br s, 0.5H), 3.05 (br s, 0.5H), 2.61–2.33 (m, 4H), 2.26–2.17 (m, 2H), 2.07–1.97 (m, 2H), 1.51 (s, 1.6H), 1.49 (s, 1.4H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.2, 196.8, 171.2, 170.2, 143.4, 141.8, 132.0, 131.2, 128.8, 128.4, 128.3, 128.2, 114.2, 111.6, 99.4, 97.8, 42.5, 40.3, 36.9, 33.5, 32.6, 29.3, 29.2, 28.1, 27.2, 20.7, 20.2; ESI-HRMS calcd. for C16H17ClO3 + H+ 293.0944, found 293.0937; 95% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 6.397 min, tmajor = 8.430 min; [ α ] D 20 = −11.9° (c = 0.007, EtOH).
4-(4-Bromophenyl)-2-hydroxy-2-methyl-2,3,4,6,7,8-hexahydro-5H-chromen-5-one (4bf). White solid; 67% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.38 (d, J = 8.8 Hz, 0.8H), 7.36 (d, J = 8.8 Hz, 1.2H), 7.05 (d, J = 10.0 Hz, 0.8H), 7.03 (d, J = 8.8 Hz, 1.2H), 3.93 (pseudo triple, J = 5.0 Hz, 0.5H), 3.81 (pseudo triple, J = 9.2 Hz, 0.5H), 3.23 (br s, 0.5H), 2.99 (br s, 0.5H), 2.61–2.33 (m, 4H), 2.28–2.17 (m, 2H), 2.07–1.97 (m, 2H), 1.52 (s, 1.6H), 1.49 (s, 1.4H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.2, 196.8, 171.2, 170.2, 143.9, 142.4, 131.7, 131.4, 130.9, 129.6, 128.8, 128.6, 120.1, 119.4, 114.2, 111.6, 99.4, 97.8, 42.5, 40.3, 36.9, 36.8, 33.6, 32.7, 29.3, 29.2, 28.1, 27.2, 20.7, 20.2; ESI-HRMS calcd. for C16H17BrO3 + H+ 337.0439, found 337.0438; 96% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 6.780 min, tmajor = 8.967 min; [ α ] D 20 = −8.5° (c = 0.008, EtOH).

3.3. Procedure for the Asymmetric Michael Reaction of Chalcones

Dimedone 1a (14.0 mg, 0.1 mmol), chalcone 5a (25.0 mg, 0.12 mmol), and quinine-based squaramide 7 (12.5 mg, 0.02 mmol) were dissolved in chloroform (1.0 mL). After stirring at room temperature for 120 h, triethylamine (41.7 μL, 0.3 mmol) was added in one portion. Subsequently, acetyl chloride (14.2 μL, 0.2 mmol) was added dropwise. Once the reaction completed (1 h), the crude product was purified over silica gel by column chromatography (EtOAc/petroleum ether) to afford 6aa (37.1 mg, 95 % yield) as a colorless oil.
5,5-Dimethyl-3-oxo-2-(3-oxo-1,3-diphenylpropyl)cyclohex-1-en-1-yl acetate (6aa) [38,53]. 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.96 (d, J = 8.0 Hz, 2H), 7.54 (t, J = 7.2 Hz, 1H), 7.44 (t, J = 7.6 Hz, 2H), 7.28–7.22 (m, 4H), 7.15 (t, J = 6.6 Hz, 1H), 4.79 (t, J = 7.4 Hz, 1H), 3.82 (ABX, JAB = 17.2 Hz, JAX = 8.0 Hz, 1H), 3.75 (ABX, JAB = 17.2, JBX = 6.8 Hz, 1H), 2.53 (AB, JAB = 18.0 Hz, 1H), 2.42 (AB, JAB = 17.2 Hz, 1H), 2.24 (s, 2H), 2.15 (s, 3H), 1.02 (s, 3H), 1.00 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 198.7, 198.6, 167.3, 163.9, 142.0, 136.8, 132.9, 128.9, 128.5, 128.2, 128.1, 127.4, 126.1, 116.4, 51.7, 42.7, 40.8, 35.7, 32.5, 28.1, 27.9, 20.9; 93% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 7.467 min, tmajor = 10.760 min; [ α ] D 20 = +41.5° (c = 0.032, CHCl3).
2-(1-(4-Chlorophenyl)-3-oxo-3-phenylpropyl)-5,5-dimethyl-3-oxocyclohex-1-en-1-yl acetate (6ab). Colorless oil; 99% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.95 (d, J = 7.6 Hz, 2H), 7.55 (t, J = 7.2 Hz, 1H), 7.45 (t, J = 7.4 Hz, 2H), 7.24–7.17 (m, 4H), 4.74 (t, J = 7.2 Hz, 1H), 3.79 (ABX, JAB = 17.6 Hz, JAX = 7.2 Hz, 1H), 3.72 (ABX, JAB = 17.8 Hz, JBX = 7.8 Hz, 1H), 2.53 (AB, JAB = 18.0 Hz, 1H), 2.42 (AB, JAB = 17.6 Hz, 1H), 2.24 (s, 2H), 2.20 (s, 3H), 1.02 (s, 3H), 1.00 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 198.8, 198.3, 167.4, 164.0, 140.5, 136.7, 133.1, 131.9, 128.9, 128.7, 128.6, 128.3, 128.1, 51.7, 42.7, 40.6, 35.3, 32.6, 27.9, 27.8, 20.9; ESI-HRMS calcd. for C25H26ClO4 + H+ 425.1514, found 425.1514; 94% ee was determined by HPLC on AD-H column, hexane/i-propanol (70/30), 1.0 mL/min, UV 254 nm, tminor = 6.560 min, tmajor = 11.080 min; [ α ] D 20 = +79.6° (c = 0.019, EtOH).
5,5-Dimethyl-3-oxo-2-(3-oxo-3-phenyl-1-(p-tolyl)propyl)cyclohex-1-en-1-yl acetate (6ac). Colorless oil; 96% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.96 (d, J = 7.2 Hz, 2H), 7.54 (t, J = 7.4 Hz, 1H), 7.44 (t, J = 7.6 Hz, 2H), 7.16 (d, J = 8.0 Hz, 2H), 7.05 (d, J = 8.0 Hz, 2H), 4.74 (t, J = 7.4 Hz, 1H), 3.82 (ABX, JAB = 17.2 Hz, JAX = 8.0 Hz, 1H), 3.73 (ABX, JAB = 17.2 Hz, JBX = 6.8 Hz, 1H), 2.53 (AB, JAB = 17.6 Hz, 1H), 2.43 (AB, JAB = 17.6 Hz, 1H), 2.28 (s, 3H), 2.24 (s, 2H), 2.17 (s, 3H), 1.02 (s, 3H), 1.01 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 198.9, 198.8, 167.5, 163.8, 138.9, 136.9, 135.7, 132.9, 129.1, 128.9, 128.5, 128.1, 127.4, 51.8 42.8, 40.9, 35.5, 32.6, 28.0, 27.9, 20.9; ESI-HRMS calcd. for C26H29O4 + H+ 405.2060, found 405.2061; 93% ee was determined by HPLC on AD-H column, hexane/i-propanol (70/30), 1.0 mL/min, UV 254 nm, tminor = 6.467 min, tmajor = 13.007 min; [ α ] D 20 = +70.1° (c = 0.018, EtOH).
5,5-Dimethyl-3-oxo-2-(3-oxo-3-phenyl-1-(thiophen-2-yl)propyl)cyclohex-1-en-1-yl acetate (6ad). Colorless oil; 95% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.72–7.67 (m, 1H), 7.54 (d, J = 4.0 Hz, 1H), 7.20–7.16 (m, 4H), 7.10–7.09 (m, 1H), 7.05–7.00 (m, 1H), 4.71 (t, J = 7.4 Hz, 1H), 3.72 (ABX, JAB = 16.2 Hz, JAX = 8.4 Hz, 1H), 3.57 (ABX, JAB = 16.4 Hz, JBX = 6.4 Hz, 1H), 2.45 (AB, JAB = 18.0 Hz, 1H), 2.36 (AB, JAB = 17.6 Hz, 1H), 2.16 (s, 2H), 2.12 (s, 3H), 0.93 (s, 6H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 198.8, 191.7, 167.4, 164.1, 144.3, 141.8, 133.7, 132.1, 128.7, 128.2, 128.0, 127.4, 126.2, 51.7, 42.7, 41.3, 36.0, 32.6, 27.9, 27.8, 20.9; ESI-HRMS calcd. for C23H25O4S + H+ 397.1468, found 397.1468; 91% ee was determined by HPLC on AD-H column, hexane/i-propanol (70/30), 1.0 mL/min, UV 254 nm, tminor = 6.903 min, tmajor = 10.567 min; [ α ] D 20 = +29.6° (c = 0.016, EtOH).
5,5-Dimethyl-2-(1-(naphthalen-2-yl)-3-oxo-3-phenylpropyl)-3-oxocyclohex-1-en-1-yl acetate (6ae). Colorless oil; 68% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 8.00 (d, J = 8.0 Hz, 2H), 7.78–7.73 (m, 4H), 7.56 (t, J =7.4 Hz, 1H), 7.46 (t, J = 7.4 Hz, 2H), 7.43–7.39 (m, 3H), 4.97 (t, J = 7.2 Hz, 1H), 3.94 (ABX, JAB = 17.6 Hz, JAX = 8.4 Hz, 1H), 3.88 (ABX, JAB = 17.0 Hz, JBX = 7.0 Hz, 1H), 2.56 (AB, JAB = 18.0 Hz, 1H), 2.43 (AB, JAB = 18.0 Hz, 1H), 2.28 (AB, JAB = 16.8 Hz, 1H), 2.24 (AB, JAB = 17.6 Hz, 1H), 2.15 (s, 3H), 1.04 (s, 3H), 1.01 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 198.8, 198.6, 167.4, 164.1, 139.5, 136.9, 133.3, 133.0, 132.0, 128.8, 128.5, 128.1, 127.9, 127.7, 127.5, 126.5, 125.8, 125.6, 125.3, 51.7, 42.8, 40.8, 35.9, 32.6, 27.9, 27.8, 20.9; ESI-HRMS calcd. for C29H29O4 + H+ 441.2060, found 441.2061; 91% ee was determined by HPLC on AD-H column, hexane/i-propanol (70/30), 1.0 mL/min, UV 254 nm, tminor = 7.943 min, tmajor = 11.667 min; [ α ] D 20 = +79.7° (c = 0.011, EtOH).
2-(3-(4-Chlorophenyl)-3-oxo-1-phenylpropyl)-5,5-dimethyl-3-oxocyclohex-1-en-1-yl acetate (6af). White solid; 98% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.90 (d, J = 8.4 Hz, 2H), 7.41 (d, J = 8.4 Hz, 2H), 7.26–7.23 (m, 4H), 7.20–7.23 (m, 1H), 4.77 (t, J = 7.4 Hz, 1H), 3.82 (ABX, JAB = 17.2 Hz, JAX = 8.0 Hz, 1H), 3.70 (ABX, JAB = 17.2 Hz, JBX = 6.8 Hz, 1H), 2.54 (AB, JAB = 17.6 Hz, 1H), 2.42 (AB, JAB = 18.0 Hz, 1H), 2.24 (s, 2H), 2.17 (s, 3H), 1.02 (s, 3H), 1.00 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 198.9, 197.5, 167.3, 164.0, 141.8, 139.3, 135.1, 129.5, 128.75, 128.72, 128.2, 127.4, 126.2, 51.7, 42.7, 40.7, 35.8, 32.5, 27.8, 20.9; ESI-HRMS calcd. for C25H26ClO4 + H+ 425.1514, found 425.1514; 87% ee was determined by HPLC on AD-H column, hexane/i-propanol (70/30), 1.0 mL/min, UV 254 nm, tminor = 9.913 min, tmajor = 15.547 min; [ α ] D 20 = +57.9° (c = 0.016, EtOH).
5,5-Dimethyl-3-oxo-2-(3-oxo-1-phenyl-3-(p-tolyl)propyl)cyclohex-1-en-1-yl acetate (6ag). Colorless oil; 99% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.87 (d, J = 8.0 Hz, 2H), 7.29–7.23 (m, 6H), 7.19–7.11 (m, 1H), 4.79 (t, J = 7.4 Hz, 1H), 3.79 (ABX, JAB = 16.4 Hz, JAX = 7.6 Hz, 1H), 3.73 (ABX, JAB = 16.8 Hz, JBX = 6.8 Hz, 1H), 2.53 (AB, JAB = 18.0 Hz, 1H), 2.42 (AB, JAB = 17.6 Hz, 1H), 2.40 (s, 3H), 2.25 (s, 2H), 2.17 (s, 3H), 1.03 (s, 3H), 1.01 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 198.8, 198.3, 167.4, 163.8, 143.7, 142.1, 134.4, 129.1, 129.0, 128.2, 128.1, 127.5, 126.1, 51.7, 42.7, 40.6, 35.7, 32.6, 27.9, 21.6, 20.9; ESI-HRMS calcd. for C26H29O4 + H+ 405.2060, found 405.2061; 95% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 9.573 min, tmajor = 13.540 min; [ α ] D 20 = +77.1° (c = 0.018, EtOH)
5,5-Dimethyl-3-oxo-2-(3-oxo-1-phenyl-3-(thiophen-2-yl)propyl)cyclohex-1-en-1-yl acetate (6ah). Colorless oil; 99% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.88 (d, J = 7.6 Hz, 2H), 7.47 (t, J = 7.2 Hz, 1H), 7.36 (t, J = 7.4 Hz, 2H), 7.03–6.97 (m, 1H), 6.83–6.74 (m, 2H), 4.97 (t, J = 7.0 Hz, 1H), 3.82 (ABX, JAB = 17.4 Hz, JAX = 7.8 Hz, 1H), 3.70 (ABX, JAB = 17.6 Hz, JBX = 6.4 Hz, 1H), 2.49 (AB, JAB = 18.0 Hz, 1H), 2.38 (AB, JAB = 17.6 Hz, 1H), 2.19 (s, 2H), 2.15 (s, 3H), 0.97 (s, 3H), 0.95 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 198.4, 198.0, 167.3, 164.1, 145.7, 136.6, 133.0, 128.5, 128.2, 128.1, 126.4, 124.1, 123.3, 51.6, 42.6, 42.3, 32.6, 31.5, 27.9, 27.8, 20.9; ESI-HRMS calcd. for C23H25O4S + H+ 397.1468, found 397.1469; 97% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 7.623 min, tmajor = 9.193 min; [ α ] D 20 = +102.9° (c = 0.019, EtOH)
3-Oxo-2-(3-oxo-1,3-diphenylpropyl)cyclohex-1-en-1-yl acetate (6ba). Colorless oil; 93% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.97 (d, J = 7.6 Hz, 2H), 7.55 (t, J = 7.2 Hz, 1H), 7.44 (t, J =7.4 Hz, 2H), 7.29–7.23 (m, 4H), 7.16 (t, J = 6.6 Hz, 1H), 4.80 (t, J = 7.2 Hz, 1H), 3.81 (ABX, JAB = 17.8 Hz, JAX = 6.6 Hz, 1H), 3.75 (ABX, JAB = 17.6 Hz, JBX = 7.6 Hz, 1H), 2.65 (dt, J = 18.0, 6.4 Hz, 1H), 2.53 (dt, J = 18.0, 5.9 Hz, 1H), 2.37 (t, J = 6.4 Hz, 2H), 2.17 (s, 3H), 1.98–1.89 (m, 2H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 198.8, 198.7, 167.3, 165.6, 142.0, 136.8, 132.9, 130.1, 128.5, 128.1, 128.0, 127.5, 126.1, 40.8, 37.9, 35.7, 29.0, 20.9, 20.7; ESI-HRMS calcd. for C23H23O4 + H+ 363.1591, found 363.1591; 91% ee was determined by HPLC on AD-H column, hexane/i-propanol (70/30), 1.0 mL/min, UV 254 nm, tminor = 7.667 min, tmajor = 10.660 min; [ α ] D 20 = +92.4° (c = 0.016, EtOH).
3-Oxo-2-(3-oxo-1,3-diphenylpropyl)cyclopent-1-en-1-yl acetate (6ca). Colorless oil; 31% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.94 (d, J = 7.6 Hz, 2H), 7.52 (t, J = 7.2 Hz, 1H), 7.42 (t, J = 7.0 Hz, 2H), 7.36–7.34 (m, 2H), 7.26 (t, J = 6.8 Hz, 2H), 7.18 (t, J = 7.0 Hz, 1H), 4.49 (t, J = 7.2 Hz, 1H), 4.06 (ABX, JAB = 17.6 Hz, JBX = 8.8 Hz, 1H), 3.56 (ABX, JAB = 17.6 Hz, JBX = 6.0 Hz, 1H), 2.93–2.79 (m, 2H), 2.47–2.38 (m, 2H), 2.21 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 205.3, 198.3, 176.7, 166.5, 141.9, 136.8, 133.1, 129.9, 128.6, 128.5, 128.0, 127.8, 126.7, 40.8, 35.8, 34.7, 26.9, 21.1; ESI-HRMS calcd. for C22H22O4 + H+ 349.1440, found 349.1437; 57% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 8.787 min, tmajor = 12.983 min; [ α ] D 20 = +29.0° (c = 0.014, EtOH).
5,5-Dimethyl-3-oxo-2-(1-oxo-1-phenylhexan-3-yl)cyclohex-1-en-1-yl acetate (6ai). Colorless oil; 93% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.91 (d, J = 7.2 Hz, 2H), 7.52 (t, J = 7.2 Hz, 1H), 7.42 (t, J = 7.4 Hz, 2H), 3.32 (ABX, JAB = 16.4 Hz, JBX = 6.8 Hz, 1H), 3.23 (ABX, JAB = 16.0 Hz, JBX = 6.4 Hz, 1H), 2.46 (AB, JAB = 18.0 Hz, 1H), 2.41 (AB, JAB = 18.0 Hz, 1H), 2.27 (AB, JAB = 15.6 Hz, 1H), 2.21 (AB, JAB = 15.2 Hz, 1H), 2.20 (s, 3H), 1.77–1.68 (m, 2H), 1.54–1.47 (m, 1H), 1.21–1.16 (m, 2H), 1.06 (s, 3H), 0.99 (s, 3H), 0.85 (t, J = 7.2 Hz, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 199.7, 199.4, 167.7, 163.8, 137.2, 132.8, 128.6, 128.4, 128.1, 52.1, 42.7, 42.0, 35.0, 32.4, 31.6, 28.0, 27.9, 21.1, 20.9, 13.9; ESI-HRMS calcd. for C22H29O4 + H+ 357.2066, found 357.2064; 52% ee was determined by HPLC on IC column, hexane/i-propanol (99/1), 1.0 mL/min, UV 254 nm, tmajor = 39.457 min, tminor = 42.963 min; [ α ] D 20 = +4.75° (c = 0.022, EtOH).

3.4. Preparation of 4H-Pyran via Dehydrating

Thionyl chloride (7.3 μL, 11.9 mg, 0.1 mmol) was added dropwise to a solution of 4a (28.6 mg, 0.1 mmol, 97% ee) and pyridine (14.1 μL, 15.8 mg, 0.2 mmol) in DCM (1.0 mL) at rt. After the reaction completed, the solvent was removed under reduced pressure. The residue was subjected to silica gel flash chromatography (EtOAc/petroleum ether) to provide 8 (19.3 mg, 72% yield) as a white solid.
2,7,7-Trimethyl-4-phenyl-4,6,7,8-tetrahydro-5H-chromen-5-one (8): 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.29–7.24 (m, 4H), 7.17–7.13 (m, 1H), 4.90 (d, J = 4.8 Hz, 2H), 4.28 (d, J = 4.0 Hz, 1H), 2.39 (s, 2H), 2.22 (AB, JAB = 16.0 Hz, 1H), 2.16 (AB, JAB = 16.4 Hz, 1H), 1.88 (s, 3H), 1.09 (s, 3H), 1.03 (s, 3H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 197.3, 164.7, 145.8, 145.6, 128.2, 127.8, 126.2, 112.2, 104.5, 50.8, 41.4, 35.2, 31.9, 29.1, 27.6, 18.6; ESI-HRMS calcd. for C18H20O2 + H+ 269.1542, found 269.1541; 98% ee was determined by HPLC on OD-H column, hexane/i-propanol (90/10), 1.0 mL/min, UV 254 nm, tminor = 5.880 min, tmajor = 8.550 min; [ α ] D 20 = −182.7° (c = 0.024, EtOH).

3.5. Preparation of Fused Dihydrofuran via Stereoselective Oxidative Cyclization

After the initial Michael addition between 5a (49.9 mg, 0.24 mmol) and 1a (28.0 mg, 0.2 mmol) was completed, the corresponding adduct was purified via flash column chromatography. Subsequently, the mixture of PhIO (66 mg, 0.3 mmol) and Michael adduct (69.6 mg, 0.2 mmol) in H2O (1 mL) was treated with Bu4NI (111 mg, 0.3 mmol). The reaction mixture was warmed up to 30 °C and allowed to stir for 16 h. The reaction was followed by TLC until completion. The reaction mixture was successively quenched with saturated Na2S2O3 (25 mL) and extracted by dichloromethane (25 mL × 3). The organic layer was dried over Na2SO4 and concentrated in vacuo. The residue was purified by column chromatography on silica gel (EtOAc/petroleum ether) to furnish 2,3-dihydrobenzofuran 9 in 61% yield as a colorless oil.
2-Benzoyl-6,6-dimethyl-3-phenyl-3,5,6,7-tetrahydrobenzofuran-4(2H)-one (9) [61]. White solid; 61% yield purified by flash column chromatography; 1H-NMR (400 MHz, CDCl3) δ (ppm) 7.82 (d, J = 7.2 Hz, 2H), 7.61 (t, J = 7.4 Hz, 1H), 7.46 (t, J = 7.6 Hz, 2H), 7.36 (t, J =7.2 Hz, 2H), 7.30–7.28 (m, 1H), 7.25–7.23 (m, 2H), 5.89 (d, J = 4.8 Hz, 1H), 4.40 (d, J = 4.0 Hz, 1H), 2.63 (ABX, JAB = 18.0 Hz, JAX = 1.6 Hz, 1H), 2.53 (AB, JAB = 17.6 Hz, 1H), 2.25 (AB, JAB = 16.4 Hz, 1H), 2.18 (AB, JAB = 16.4 Hz, 1H), 1.17 (s, 6H); 13C-NMR (100 MHz, CDCl3) δ (ppm) 193.5, 192.8, 176.3, 141.2, 134.1, 133.1, 129.0, 128.89, 128.86, 127.6, 127.3, 115.1, 91.8, 51.1, 48.9, 37.6, 34.3, 29.0, 28.3; 93% ee was determined by HPLC on AD-H column, hexane/i-propanol (80/20), 1.0 mL/min, UV 254 nm, tminor = 10.843 min, tmajor = 15.107 min; [ α ] D 20 = −44.6° (c = 0.021, EtOH).

4. Conclusions

In summary, we have successfully developed an enantioselective Michael addition of cyclic β-diones to α,β-unsaturated enones in the presence of quinine-based primary amine or squaramide. These asymmetric processes displayed especially broad substrate generalities, and various cinnamones and chalcones furnished the desired adducts in good to high yields. Although chalcones proved to be a class of challenging acceptors in the precedent study [38], good reactivities and excellent enantiopurities were achieved in the case of their Michael addition with cyclic β-diones via our protocol.

Supplementary Materials

The supplementary materials are available online.

Acknowledgments

This work is financially supported by the National Natural Science Foundation of China (No. 21402163) and the Fundamental Research Funds for the Central Universities of Southwest Minzu University (No. 2016NGJPY02). Wang Qingqing gratefully acknowledges the Graduate Innovation Project of Southwest Minzu University (No. CX2016SZ058).

Author Contributions

Q.W. and X.L. (Xuefeng Li) designed and planned the research. Q.W. and W.W. performed the experiments and analyzed the data. L.Y., X.Y., X.L. (Xinying Li), and Z.Z. characterized the compounds. X.L. (Xuefeng Li) prepared the manuscript for publication. All authors discussed the results and commented on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sibi, M.P.; Manyem, S. Enantioselective conjugate additions. Tetrahedron 2000, 56, 8033–8061. [Google Scholar] [CrossRef]
  2. Christoffers, J.; Koripelly, G.; Rosiak, A.; Rössle, M. Recent advances in metal-catalyzed asymmetric conjugate additions. Synthesis 2007, 9, 1279–1300. [Google Scholar] [CrossRef]
  3. Krause, N.; Hoffmann-Röder, A. Recent advances in catalytic enantioselective Michael additions. Synthesis 2001, 2, 171–196. [Google Scholar] [CrossRef]
  4. Almaşi, D.; Alonso, D.A.; Najera, C. Organocatalytic asymmetric conjugate additions. Tetrahedron Asymmetry 2007, 18, 299–365. [Google Scholar] [CrossRef]
  5. Halland, N.; Hazell, R.G.; Jørgensen, K.A. Organocatalytic asymmetric conjugate addition of nitroalkanes to α,β-unsaturated enones using novel imidazoline catalysts. J. Org. Chem. 2002, 67, 8331–8338. [Google Scholar] [CrossRef] [PubMed]
  6. Halland, N.; Aburel, P.S.; Jørgensen, K.A. Highly Enantioselective Organocatalytic Conjugate Addition of Malonates toAcyclic α,β-Unsaturated Enones. Angew. Chem. Int. Ed. 2003, 42, 661–665. [Google Scholar] [CrossRef] [PubMed]
  7. Halland, N.; Hansen, T.; Jørgensen, K.A. Organocatalytic Asymmetric Michael Reaction of Cyclic 1,3-Dicarbonyl Compounds and α,β-Unsaturated Ketones—A Highly Atom-Economic Catalytic One-Step Formation of Optically Active Warfarin Anticoagulant. Angew. Chem. Int. Ed. 2003, 42, 4955–4957. [Google Scholar] [CrossRef] [PubMed]
  8. Halland, N.; Aburel, P.S.; Jørgensen, K.A. Highly Enantio- and Diastereoselective Organocatalytic Asymmetric Domino Michael-Aldol Reaction of β-Ketoesters and α,β-Unsaturated Ketones. Angew. Chem. Int. Ed. 2004, 43, 1272–1277. [Google Scholar]
  9. Pulkkinen, J.; Aburel, P.S.; Halland, N.; Jørgensen, K.A. Highly Enantio-and Diastereoselective Organocatalytic Domino Michael-Aldol Reactions of β-Diketone and β-Ketosulfone Nucleophiles with α,β-Unsaturated Ketones. Adv. Synth. Catal. 2004, 346, 1077–1080. [Google Scholar] [CrossRef]
  10. Prieto, A.; Halland, N.; Jørgensen, K.A. Novel imidazolidine-tetrazole organocatalyst for asymmetric conjugate addition of nitroalkanes. Org. Lett. 2005, 7, 3897–3900. [Google Scholar] [CrossRef] [PubMed]
  11. Kwiatkowski, P.; Dudzinski, K.; Łyżwa, D. Effect of High Pressure on the Organocatalytic Asymmetric Michael Reaction: Highly Enantioselective Synthesis of γ-Nitroketones with Quaternary Stereogenic Centers. Org. Lett. 2011, 13, 3624–3627. [Google Scholar] [CrossRef] [PubMed]
  12. Melchiorre, P. Cinchona-based Primary Amine Catalysis in the Asymmetric Functionalization of Carbonyl Compounds. Angew. Chem. Int. Ed. 2012, 51, 9748–9770. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, L.-W.; Luo, J.; Lu, Y. Asymmetric catalysis with chiral primary amine-based organocatalysts. Chem. Commun. 2009, 1807–1821. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, Y.-C. The Development of AsymmetricPrimary Amine Catalysts Based on Cinchona Alkaloids. Synlett 2008, 2008, 1919–1930. [Google Scholar] [CrossRef]
  15. Xie, J.W.; Chen, W.; Li, R.; Zeng, M.; Du, W.; Yue, L.; Chen, Y.C.; Wu, Y.; Zhu, J.; Deng, J.G. Highly Asymmetric Michael Addition to α,β-Unsaturated Ketones Catalyzed by 9-Amino-9-deoxyepiquinine. Angew. Chem. Int. Ed. 2007, 46, 389–392. [Google Scholar] [CrossRef] [PubMed]
  16. Bartoli, G.; Bosco, M.; Carlone, A.; Pesciaioli, F.; Sambri, L.; Melchiorre, P. Organocatalytic asymmetric Friedel-Crafts alkylation of indoles with simple α,β-unsaturated ketones. Org. Lett. 2007, 9, 1403–1405. [Google Scholar] [CrossRef] [PubMed]
  17. Pesciaioli, F.; De Vincentiis, F.; Galzerano, P.; Bencivenni, G.; Bartoli, G.; Mazzanti, A.; Melchiorre, P. Organocatalytic Asymmetric Aziridination of Enones. Angew. Chem. Int. Ed. 2008, 47, 8703–8706. [Google Scholar] [CrossRef] [PubMed]
  18. Lu, X.; Liu, Y.; Sun, B.; Cindric, B.; Deng, L. Catalytic Enantioselective Peroxidation of α,β-Unsaturated Ketones. J. Am. Chem. Soc. 2008, 130, 8134–8135. [Google Scholar] [CrossRef] [PubMed]
  19. McCooey, S.H.; Connon, S.J. Readily accessible 9-epi-amino cinchona alkaloid derivatives promote efficient, highly enantioselective additions of aldehydes and ketones to nitroolefins. Org. Lett. 2007, 9, 599–602. [Google Scholar] [CrossRef] [PubMed]
  20. Calter, M.A.; Wang, J. Catalytic, asymmetric Michael reactions of cyclic diketones with β, γ-unsaturated α-ketoesters. Org. Lett. 2009, 11, 2205–2208. [Google Scholar] [CrossRef] [PubMed]
  21. Yu, C.; Zheng, F.; Ye, H.; Zhong, W. Enantioselective synthesis of functionalized 3,4-dihydropyran derivatives organocatalyzed by a novel fluorinated-diphenylprolinolether. Tetrahedron 2009, 65, 10016–10021. [Google Scholar] [CrossRef]
  22. Dong, Z.; Feng, J.; Fu, X.; Liu, X.; Lin, L.; Feng, X. Highly Enantioselective Conjugate Addition of Cyclic Diketones to β,γ-Unsaturated α-Ketoesters Catalyzed by an N,N′-Dioxide-Cu(OTf)2 Complex. Chem. Eur. J. 2011, 17, 1118–1121. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, G.; Wang, Y.; Zhang, W.; Xu, X.; Zhong, A.; Xu, D. Organocatalytic Michael Addition of Naphthoquinone with α,β-Unsaturated Ketones: Primary Amine Catalyzed Asymmetric Synthesis of Lapachol Analogues. Eur. J. Org. Chem. 2011, 2011, 2142–2147. [Google Scholar] [CrossRef]
  24. Tan, B.; Lu, Y.; Zeng, X.; Chua, P.J.; Zhong, G. Facile Domino Access to Chiral Bicyclo[3.2.1]octanes and Discovery of a New Catalytic Activation Mode. Org. Lett. 2010, 12, 2682–2685. [Google Scholar] [CrossRef] [PubMed]
  25. Tsakos, M.; Elsegood, M.R.J.; Kokotos, C.G. Organocatalytic asymmetric domino Michael-Henry reaction for the synthesis of substituted bicyclo[3.2.1]octan-2-ones. Chem. Commun. 2013, 49, 2219–2221. [Google Scholar] [CrossRef] [PubMed]
  26. Franke, P.T.; Richter, B.; Jørgensen, K.A. Organocatalytic Asymmetric Synthesis of Functionalized 3,4-Dihydropyran Derivatives. Chem. Eur. J. 2008, 14, 6317–6321. [Google Scholar] [CrossRef] [PubMed]
  27. Rueping, M.; Sugiono, E.; Merino, E. Asymmetric Organocatalysis: An Efficient Enantioselective Access to Benzopyranes and Chromenes. Chem. Eur. J. 2008, 14, 6329–6332. [Google Scholar] [CrossRef] [PubMed]
  28. Kim, H.; Yen, C.; Preston, P.; Chin, J. Substrate-directed stereoselectivity in vicinal diamine-catalyzed synthesis of warfarin. Org. Lett. 2006, 8, 5239–5242. [Google Scholar] [CrossRef] [PubMed]
  29. Xie, J.-W.; Yue, L.; Chen, W.; Du, W.; Zhu, J.; Deng, J.-G.; Chen, Y.-C. Highly enantioselective Michael addition of cyclic 1, 3-dicarbonyl compounds to α,β-unsaturated ketones. Org. Lett. 2007, 9, 413–415. [Google Scholar] [CrossRef] [PubMed]
  30. Dong, Z.; Wang, L.; Chen, X.; Liu, X.; Lin, L.; Feng, X. Organocatalytic Enantioselective Michael Addition of 4-Hydroxycoumarin to α,β-Unsaturated Ketones: A Simple Synthesis of Warfarin. Eur. J. Org. Chem. 2009, 2009, 5192–5197. [Google Scholar] [CrossRef]
  31. Kristensen, T.E.; Vestli, K.; Hansen, F.K.; Hansen, T. New Phenylglycine-Derived Primary Amine Organocatalysts for the Preparation of Optically Active Warfarin. Eur. J. Org. Chem. 2009, 2009, 5185–5191. [Google Scholar] [CrossRef]
  32. Zhu, X.; Lin, A.; Shi, Y.; Guo, J.; Zhu, C.; Cheng, Y. Enantioselective Synthesis of Polycyclic Coumarin Derivatives Catalyzed by an in Situ Formed Primary Amine-Imine Catalyst. Org. Lett. 2011, 13, 4382–4385. [Google Scholar] [CrossRef] [PubMed]
  33. Mei, R.-Q.; Xu, X.-Y.; Li, Y.-C.; Fu, J.-Y.; Huang, Q.-C.; Wang, L.-X. Highly effective and enantioselective Michael addition of 4-hydroxycoumarin to α,β-unsaturated ketones promoted by simple chiral primary amine thiourea bifunctional catalysts. Tetrahedron Lett. 2011, 52, 1566–1568. [Google Scholar] [CrossRef]
  34. Ray, S.K.; Singh, P.K.; Molleti, N.; Singh, V.K. Enantioselective Synthesis of Coumarin Derivatives by PYBOX-DIPH-Zn(II) Complex Catalyzed Michael Reaction. J. Org. Chem. 2012, 77, 8802–8808. [Google Scholar] [CrossRef] [PubMed]
  35. Dong, J.; Du, D.-M. Highly enantioselective synthesis of Warfarin and its analogs catalysed by primary amine-phosphinamide bifunctional catalysts. Org. Biomol. Chem. 2012, 10, 8125–8131. [Google Scholar] [CrossRef] [PubMed]
  36. Işık, M.; Akkoca, H.U.; Akhmedov, İ.M.; Tanyeli, C. A bis-Lewis basic 2-aminoDMAP/prolinamide organocatalyst for application to the enantioselective synthesis of Warfarin and derivatives. Tetrahedron Asymmetry 2016, 27, 384–388. [Google Scholar] [CrossRef]
  37. Liu, Y.; Liu, X.; Wang, M.; He, P.; Lin, L.; Feng, X. Enantioselective Synthesis of 3,4-Dihydropyran Derivatives via Organocatalytic Michael Reaction of α,β-Unsaturated Enones. J. Org. Chem. 2012, 77, 4136–4142. [Google Scholar] [CrossRef] [PubMed]
  38. Molleti, N.; Allu, S.; Ray, S.K.; Singh, V.K. Bifunctional chiral urea catalyzed highly enantioselective Michael addition of cyclic 1,3-dicarbonyl compounds to 2-enoylpyridines. Tetrahedron Lett. 2013, 54, 3241–3244. [Google Scholar] [CrossRef]
  39. Ray, S.K.; Rout, S.; Singh, V.K. Enantioselective synthesis of 3, 4-dihydropyran derivatives via a Michael addition reaction catalysed by chiral pybox–diph–Zn (II) complex. Org. Biomol. Chem. 2013, 11, 2412–2416. [Google Scholar] [CrossRef] [PubMed]
  40. Kumar, A.; Sharma, S.; Tripathi, V.D.; Maurya, R.A.; Srivastava, S.P.; Bhatia, G.; Tamrakar, A.K.; Srivastava, A.K. Design and synthesis of 2,4-disubstituted polyhydroquinolines as prospective antihyperglycemic and lipid modulating agents. Biorg. Med. Chem. 2010, 18, 4138–4148. [Google Scholar] [CrossRef] [PubMed]
  41. Rodrigo, G.; Standen, N. ATP-sensitive potassium channels. Curr. Pharm. Des. 2005, 11, 1915–1940. [Google Scholar] [CrossRef] [PubMed]
  42. Li, X.; Cun, L.; Lian, C.; Zhong, L.; Chen, Y.; Liao, J.; Zhu, J.; Deng, J. Highly enantioselective Michael addition of malononitrile to alpha,beta-unsaturated ketones. Org. Biomol. Chem. 2008, 6, 349–353. [Google Scholar] [CrossRef] [PubMed]
  43. Li, X.; Ma, Y.; Xing, Z.; Tang, N.; Zhu, J.; Deng, J. The asymmetric addition of malononitrile to α,β-unsaturated ketones catalyzed by RuCl2[(R,R)-DPEN](PPh3)2 as the precatalyst. Tetrahedron Lett. 2014, 55, 3868–3872. [Google Scholar] [CrossRef]
  44. Zheng, Y.; Yao, Y.; Ye, L.; Shi, Z.; Li, X.; Zhao, Z.; Li, X. Highly enantioselective Michael addition of malononitrile to α,β-unsaturated pyrazolamides catalyzed by a bifunctional thiourea. Tetrahedron 2016, 72, 973–978. [Google Scholar] [CrossRef]
  45. Liu, S.; Wang, Q.; Ye, L.; Shi, Z.; Zhao, Z.; Yang, X.; Ding, K.; Li, X. A general, highly enantioselective Michael addition of nitroalkanes to α,β-unsaturated enones catalyzed by 9-amino (9-deoxy)-epi-quinine: A remarkable additive effect. Tetrahedron 2016, 72, 5115–5120. [Google Scholar] [CrossRef]
  46. Chen, X.K.; Zheng, C.W.; Zhao, S.L.; Chai, Z.; Yang, Y.Q.; Zhao, G.; Cao, W.G. Highly Enantioselective Michael Addition of Cyclic 1, 3-Dicarbonyl Compounds to β, γ-Unsaturated α-Keto Esters. Adv. Synth. Catal. 2010, 352, 1648–1652. [Google Scholar] [CrossRef]
  47. Wang, Y.F.; Wang, K.; Zhang, W.; Zhang, B.B.; Zhang, C.X.; Xu, D.Q. Enantioselective Asymmetric Michael Addition of Cyclic Diketones to β, γ-Unsaturated α-Keto Esters. Eur. J. Org. Chem. 2012, 2012, 3691–3696. [Google Scholar] [CrossRef]
  48. Zhang, G.; Zhang, Y.; Yan, J.; Chen, R.; Wang, S.; Ma, Y.; Wang, R. One-Pot Enantioselective Synthesis of Functionalized Pyranocoumarins and 2-Amino-4H-chromenes: Discovery of a Type of Potent Antibacterial Agent. J. Org. Chem. 2012, 77, 878–888. [Google Scholar] [CrossRef] [PubMed]
  49. Gao, Y.; Du, D.-M. Enantioselective synthesis of 2-amino-5,6,7,8-tetrahydro-5-oxo-4H-chromene-3-carbonitriles using squaramide as the catalyst. Tetrahedron Asymmetry 2012, 23, 1343–1349. [Google Scholar] [CrossRef]
  50. Feng, X.; Zhou, Z.; Zhou, R.; Zhou, Q.-Q.; Dong, L.; Chen, Y.-C. Stereodivergence in Amine-Catalyzed Regioselective [4 + 2] Cycloadditions of β-Substituted Cyclic Enones and Polyconjugated Malononitriles. J. Am. Chem. Soc. 2012, 134, 19942–19947. [Google Scholar] [CrossRef] [PubMed]
  51. Vakulya, B.; Varga, S.; Csámpai, A.; Soós, T. Highly enantioselective conjugate addition of nitromethane to chalcones using bifunctional cinchona organocatalysts. Org. Lett. 2005, 7, 1967–1969. [Google Scholar] [CrossRef] [PubMed]
  52. Rueping, M.; Parra, A.; Uria, U.; Besselievre, F.; Merino, E. Catalytic asymmetric domino Michael addition– alkylation reaction: Enantioselective synthesis of dihydrofurans. Org. Lett. 2010, 12, 5680–5683. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, G.-W.; Lu, Q.-Q.; Xia, J.-J. Three Types of Products Obtained Unexpectedly from the Reaction of Dimedone with Chalcones. Eur. J. Org. Chem. 2011, 2011, 4429–4438. [Google Scholar] [CrossRef]
  54. Alemán, J.; Parra, A.; Jiang, H.; Jørgensen, K.A. Squaramides: Bridging from Molecular Recognition to Bifunctional Organocatalysis. Chem. Eur. J. 2011, 17, 6890–6899. [Google Scholar] [CrossRef] [PubMed]
  55. Tsakos, M.; Kokotos, C.G. Primary and secondary amine-(thio)ureas and squaramides and their applications in asymmetric organocatalysis. Tetrahedron 2013, 69, 10199–10222. [Google Scholar] [CrossRef]
  56. Wang, J.; Li, H.; Zu, L.; Jiang, W.; Xie, H.; Duan, W.; Wang, W. Organocatalytic enantioselective conjugate additions to enones. J. Am. Chem. Soc. 2006, 128, 12652–12653. [Google Scholar] [CrossRef] [PubMed]
  57. Ye, Y.; Wang, L.; Fan, R. Aqueous iodine (III)-mediated stereoselective oxidative cyclization for the synthesis of functionalized fused dihydrofuran derivatives. J. Org. Chem. 2010, 75, 1760–1763. [Google Scholar] [CrossRef] [PubMed]
  58. Yang, W.; Jia, Y.; Du, D.-M. Squaramide-catalyzed enantioselective Michael addition of malononitrile to chalcones. Org. Biomol. Chem. 2012, 10, 332–338. [Google Scholar] [CrossRef] [PubMed]
  59. Armarego, W.L.F.; Chai, C. Purification of Laboratory Chemicals, 7th ed.; Butterworth-Heinemann: Oxford, UK, 2012. [Google Scholar]
  60. Jiang, L.; Wang, B.; Li, R.-R.; Shen, S.; Yu, H.-W.; Ye, L.-D. Catalytic promiscuity of Escherichia coli BioH esterase: Application in the synthesis of 3,4-dihydropyran derivatives. Process Biochem. 2014, 49, 1135–1138. [Google Scholar] [CrossRef]
  61. Ye, Y.; Wang, L.; Fan, R. Aqueous Iodine(III)-Mediated Stereoselective Oxidative Cyclization for the Synthesis of Functionalized Fused Dihydrofuran Derivatives. J. Org. Chem. 2010, 75, 1760–1763. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds 4, 6, 8 and 9 are available from the authors.
Figure 1. Structures of the chiral primary amine catalysts used.
Figure 1. Structures of the chiral primary amine catalysts used.
Molecules 22 01096 g001
Scheme 1. Michael addition of dimedone to chalcone.
Scheme 1. Michael addition of dimedone to chalcone.
Molecules 22 01096 sch001
Scheme 2. Michael addition of 1,3-cyclopentadione to chalcone.
Scheme 2. Michael addition of 1,3-cyclopentadione to chalcone.
Molecules 22 01096 sch002
Scheme 3. Synthetic elaborations of the Michael adducts.
Scheme 3. Synthetic elaborations of the Michael adducts.
Molecules 22 01096 sch003
Scheme 4. Proposed transition state models.
Scheme 4. Proposed transition state models.
Molecules 22 01096 sch004
Table 1. Optimization of reaction conditions for the Michael addition of dimedone 1a to cinnamone 2a. a
Table 1. Optimization of reaction conditions for the Michael addition of dimedone 1a to cinnamone 2a. a
Molecules 22 01096 i001
EntryCat.AcidSolventTime (h)Yield (%) bee (%) c
13aTsOHtoluene486066
23aTFAtoluene486575
33aAcOHtoluene488285
43aBAtoluene248990
53aONBAtoluene369190
63aPNBAtoluene369689
73aOFBAtoluene249688
83ap-MeOC6H4CO2Htoluene369187
93aSAtoluene249990
103bSAtoluene9689−82
113cSAtoluene3689−82
123aSAPhCF3249685
133aSADCM245583
143aSAEtOH249969
153aSATHF249994
16 d3aSATHF249994
17 d,e3aSATHF969997
a Unless otherwise noted, the reaction was performed with 0.1 mmol of 1a, 0.15 mmol of 2a, 20 mol % of 3a, and 40 mol % of acid in 1 mL of solvent at room temperature (r.t.). TsOH = p-toluenesulfonic acid, TFA = trifluoroacetic acid, BA = benzoic acid, ONBA = o-nitrobenzoic acid, PNBA = p-nitrobenzoic acid, OFBA = o-fluorobenzoic acid, DCM = dichloromethane. b Isolated yield after flash chromatography on silica gel. c Determined by HPLC analysis on a chiral stationary phase (Chiralcel AD-H). d 0.12 mmol of 2a was employed. e Carried out at 0 °C.
Table 2. Substrate scope of the Michael addition of cyclic β-diones to cinnamones and its analogues. a
Table 2. Substrate scope of the Michael addition of cyclic β-diones to cinnamones and its analogues. a
Molecules 22 01096 i002
Entry1R2R324Yield (%) bee (%) c
11aPhMe2a4aa99 (95) d97 (94) d
21ao-ClC6H4Me2b4ab8791
31am-ClC6H4Me2c4ac9996
41ap-ClC6H4Me2d4ad9997 (S) e
51ap-FC6H4Me2e4ae9696
61ap-BrC6H4Me2f4af9997
71ap-MeC6H4Me2g4ag9597
81ap-MeOC6H4Me2h4ah8997
91a1-naphthyl Me2i4ai8996
101a2-naphthyl Me 2j4aj9898
111a2-furanylMe2k4ak7895
121a2-thiophenylMe2l4al9791
131aMeMe2m4am6991
141an-BuMe2n4an7594
15 1a-C3H6-2o4ao9198
16 1aPhEt2p4ap6398
17 1bPh Me 2a4ba7194
18 1bp-ClC6H4Me 2d4bd7895
19 1bp-BrC6H4Me 2f4bf6796
a Unless otherwise noted, the reaction was performed with 0.1 mmol of 1a, 0.12 mmol of 2a, 20 mol% of 3a, and 40 mol % of salicylic acid in 1 mL of THF at 0 °C for 96 h. b Isolated yield after flash chromatography on silica gel. c Determined by HPLC analysis on a chiral stationary phase. d Data within parentheses is that performed on an one-mmole scale. e Configuration of 4ad.
Table 3. Substrate scope of the Michael addition of cyclic β-diones to chalcones. a
Table 3. Substrate scope of the Michael addition of cyclic β-diones to chalcones. a
Molecules 22 01096 i003
Entry1R2 Ar1 56Yield (%) bee (%) c
11aPh Ph 5a6aa95 (99) d93 (91) d (R) e
21ap-ClC6H4Ph 5b6ab9994
31ap-MeC6H4Ph5c6ac9693
4 f1a2-thiophenylPh 5d6ad9591
5 f1a2-naphthylPh 5e6ae6891
61aPhp-ClC6H45f6af9887
71aPhp-MeC6H45g6ag9995
81aPh2-thiophenyl5h6ah9997
9 f1bPhPh5a6ba9391
10 g1an-PrPh5i6ai9352
a Unless otherwise noted, the Michael addition was performed with 0.1 mmol of 1, 0.12 mmol of 5, and 20 mol % of 7 in 1 mL of chloroform at rt for 120 h. b Isolated yield after flash chromatography on silica gel. c Determined by HPLC analysis on a chiral stationary phase. d Data within parentheses is that performed on a one-mmole scale. e Configuration of 6aa. f Performed with 168 h. g Performed with 72 h.

Share and Cite

MDPI and ACS Style

Wang, Q.; Wang, W.; Ye, L.; Yang, X.; Li, X.; Zhao, Z.; Li, X. Enantioselective Michael Addition of Cyclic β-Diones to α,β-Unsaturated Enones Catalyzed by Quinine-Based Organocatalysts. Molecules 2017, 22, 1096. https://doi.org/10.3390/molecules22071096

AMA Style

Wang Q, Wang W, Ye L, Yang X, Li X, Zhao Z, Li X. Enantioselective Michael Addition of Cyclic β-Diones to α,β-Unsaturated Enones Catalyzed by Quinine-Based Organocatalysts. Molecules. 2017; 22(7):1096. https://doi.org/10.3390/molecules22071096

Chicago/Turabian Style

Wang, Qingqing, Wei Wang, Ling Ye, Xuejun Yang, Xinying Li, Zhigang Zhao, and Xuefeng Li. 2017. "Enantioselective Michael Addition of Cyclic β-Diones to α,β-Unsaturated Enones Catalyzed by Quinine-Based Organocatalysts" Molecules 22, no. 7: 1096. https://doi.org/10.3390/molecules22071096

Article Metrics

Back to TopTop