Next Article in Journal
Expanding the Scope of the Cleavable N-(Methoxy)oxazolidine Linker for the Synthesis of Oligonucleotide Conjugates
Next Article in Special Issue
Trending Topics on Coumarin and Its Derivatives in 2020
Previous Article in Journal
Chemical Diversity and Bioactivities of Monoterpene Indole Alkaloids (MIAs) from Six Apocynaceae Genera
Previous Article in Special Issue
Synthetic Routes to Coumarin(Benzopyrone)-Fused Five-Membered Aromatic Heterocycles Built on the α-Pyrone Moiety. Part 1: Five-Membered Aromatic Rings with One Heteroatom
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Chiral Tertiary Amine Catalyzed Asymmetric [4 + 2] Cyclization of 3-Aroylcoumarines with 2,3-Butadienoate

1
Key Laboratory of Small Functional Organic Molecule, Ministry of Education, College of Chemistry and Chemical Engineering, Jiangxi Normal University, Nanchang 330022, China
2
Jiangxi Provincial Research of Targeting Pharmaceutical Engineering Technology, Shangrao Normal University, Shangrao 334001, China
*
Author to whom correspondence should be addressed.
Submission received: 28 November 2020 / Revised: 7 January 2021 / Accepted: 14 January 2021 / Published: 18 January 2021
(This article belongs to the Special Issue Coumarin and Its Derivatives)

Abstract

:
Coumarins and 2H-pyran derivatives are among the most commonly found structural units in natural products. Therefore, the introduction of 2H-pyran moiety into the coumarin structural unit, i.e., dihydrocoumarin-fused dihydropyranones, is a potentially successful route for the identification of novel bioactive structures, and the synthesis of these structures has attracted continuing research interest. Herein, a chiral tertiary amine catalyzed [4 + 2] cyclization of 3-aroylcoumarines with benzyl 2,3-butadienoate was reported. In the presence of Kumar’s 6’-(4-biphenyl)-β-iso-cinchonine, the desired dihydrocoumarin-fused dihydropyranone products could be obtained in up to 97% yield and 90% ee values.

1. Introduction

Coumarin derivatives are among the most commonly found structural units in natural products, pharmaceuticals, and functional materials [1,2,3,4,5]. Therefore, numerous endeavors have been devoted to develop effective methods for the synthesis of coumarin based compounds [6,7,8,9,10,11]. On the other hand, 2H-pyran moieties also play a vital role in natural and unnatural bioactive compounds. Therefore, the introduction of 2H-pyran moiety into coumarin structural unit is a highly potential route for the identification of novel bioactive structures and the synthesis of these structures, i.e., dihydrocoumarin-fused dihydropyranones, have attracted continuing research interest. Among the developed methods, the [4 + 2] reaction of 3-aroylcoumarins are the most commonly used [12,13,14].
Early in 2012, Shi and co-worker described the first [4 + 2] cyclization of 3-aroylcoumarines (1) with ethyl 2,3-butadienoate (2a) to construct racemic dihydrocoumarin-fused dihydropyranones 3 in 79–95% yield using DABCO as the Lewis base catalyst (Figure 1a) [15]. This [4 + 2] process was initiated by the necleophilic attack of tertiary amine to 2,3-butadienoate to generate zwitterionic I. The γ-carbanion of I then attacks the β-carbon of enones 1 to give II with Z configuration to avoid the interaction of the ester group with the 3-position substituent. In the following, an intramolecular nucleophilic substitution of II could give cycloadduct 3 and regenerate the tertiary amine catalyst. Soon after that, the Ye group reported that chiral dihydrocoumarin-fused dihydropyranones could be accessed in a highly enantioselective manner via chiral NHC catalyzed [4 + 2] cycloaddition of ketenes and 3-aroylcoumarins [16]. In 2016, Lu et al. achieved a phosphine-catalyzed [4 + 2] annulation of allenones with 3-aroylcoumarins to afford chiral dihydrocoumarin-fused dihydropyrans [17]. Chen and co-workers reported the synthesis of chiral dihydrocoumarin-fused dihydropyrans through dienamine catalysis, but only two examples were explored [18]. Despite the above achievements, the identification of new protocols using easily available starting materials and chiral catalysts for the enantioselective construction of dihydrocoumarin-fused dihydropyrans are still highly desirable.
Inspired by Shi’s pioneering work and based on our interest in the synthesis of chiral coumarin derivatives, we envision that the replacement of DABCO with a suitable chiral tertiary amine catalyst to mediate the [4 + 2] cyclization of 3-aroylcoumarines with 2,3-butadienoate might offer a new method for the synthesis of chiral dihydrocoumarin-fused dihydropyrans (Figure 1b).

2. Results and Discussion

We started our investigations by carrying out the reaction between 3-benzoylcoumarine (1a) and benzyl 2,3-butadienoate (2b) in ClCH2CH2Cl at 25 °C (Table 1). Initially, a series of cinchona alkaloids, including quinine (4a), cinchonine (4b), and C2-symmetric (bis)cinchona alkaloid (4cf) were screened (entries 1–6), and the corresponding chiral dihydrocoumarin-fused dihydropyran product (3a) was obtained in up to only 45% ee values when 4c was used, but the yield of 3a was pretty low even after 48 h (entry 3). To our delight, when the bifunctional β-isocupreidine (4g) was tested in our reaction, the reaction was greatly accelerated to complete with 8 h and delivered 3a in almost quantitative yield with promising 52% ee value (entry 7) [19,20]. Based on this result, we turn our attention to modify β-isocupreidine, so as to improve the enantiocontrol of the reaction. According the methods reported by Kumar and co-workers, a series of 6’-aryl-β-iso-cinchonine (4hl) were prepared and examined in the current reaction [21]. It was observed that 6’-phenyl-β-iso-cinchonine 4h could facilitate the model reaction to give 82% yield for 3a with improved 60% ee (entry 8). Further variation of the phenyl into more steric aryl groups turned out to be ineffective, as is demonstrated by the 29–35% ee values obtained from catalyst 4ik. A slightly improved 63% ee was achieved when β-iso-cinchonine (4l) bearing a longer 4-biphenyl group at the 6’ position was tried (entry 12), but no more improvement was obtained when further increase the length the substituent (entry 13).
In the following, the solvent effects were examined using 4l as the catalyst. The reaction was found to be more effective in solvent with moderate polarity (entries 14–21) and EtOAc was found to be the best, which afforded the desired product 3a in 87% yield and 80% ee (entry 18). In the following, we tried to lower the reaction temperature to 0 °C to improve the ee value of 3a (entry 22). To our surprise, the reaction finished within 1 h and gave an improved 93% yield, but no improvement of the ee value was observed. The observed higher reactivity at 0 °C than at 25 °C might be attributed to the competitive nucleophilic addition of quinoline nitrogen atoms of the catalyst to 2,3-butadienoate at higher temperature, which deactivate the catalyst and retard the reaction catalytic cycle.
Based on the above optimization, the scope of this tertiary amine catalyzed enantioselective [4 + 2] cyclization of 3-aroylcoumarines with benzyl 2,3-butadienoate (2b); we then evaluated this using 10 mol% of 4l as the catalyst in EtOAc at 0 °C (Figure 2). It was observed that the reaction outcome was significantly affected by the electronic properties of the substituents on the coumarin benzene ring. In general, substrates bearing electron-donating groups (EDG, such as Me, OMe) were relatively less reactive and afforded slightly higher ee values. As is shown by the 79–98% yields and 80–88% ee values for products 3be. The more steric 1f substrate could give the corresponding product 3f in highest 90% ee under standard conditions. In contrast, the reactions of electron-withdrawing group (EWG) substituted substrates are found to be more reactive but delivered relatively lower enantioselectivities. For example, products 3gi were obtained in excellent yield but with only around 70% ee. Additionally, coumarins bearing both EDG and EWG on the benzyl ring of the aroyl group were also well tolerated under standard conditions, but the enantiomeric excess was significantly affected by the steric effect. The para-substituted products (3k,n) could be obtained in much higher ee values than the meta-substituted products (3l,o). Moreover, the current reaction is also suitable for the reaction of (1,1’-biphenyl)-4-carbonyl and thiophene-2-carbonyl substituted coumarins, which afforded the desired products 3o and 3p in 79% and 81% ee values, respectively.
The Z/E configuration of the products was determined by the converting product 3m into the known compound 5 and comparing their NMR spectrum (Figure 3). Under the above optimized conditions, the reaction of 3-benzoyl coumarin 1m with ethyl 2,3-butadienoate 2a afforded product 5 with 45% ee value. The same product 5 could also be obtained via a 3-step sequence from 3m and 2b in 79% ee value. The NMR spectrum of these newly synthesized products 5 were identical with the previous report by Shi and co-workers. Thus, the configuration of the products 3 were assigned to be E. This process also highlighted the synthetic potential of product 1 to be elaborated into other dihydrocoumarin-fused dihydropyran derivatives. We also tried to recrystallize products 3 and determined their absolute configuration by X-ray crystallography analysis, but turned out to be unsuccessful.
In order to demonstrate the practicability of the current method, we conducted a gram-scale reaction of 3-benzoyl coumarin 1f with benzyl 2,3-butadienoate 2b (Figure 4). In the presence of only 2.5 mol% of 4l as the catalyst, the reaction of 2.5 mmol of 1f with 1.5 equivalents of 2b could give rise to the desired dihydrocoumarin-fused dihydropyran 3f in 91% yield (1.215 g) with slightly improved 93% ee.
In summary, a series of chiral dihydrocoumarin-fused dihydropyranones were sysnthesized via the enantioselective [4 + 2] cyclization of 3-aroylcoumarines with benzyl 2,3-butadienoate. In the presence of 10 mol% of Kumar’s 6’-(4-biphenyl)-β-iso-cinchonine as the chiral tertiary amine catalyst, the desired products could be obtained in up to 97% yield and 90% ee values under mild conditions. The current method used an easily available chiral catalyst and starting materials and could be conducted on gram-scale without loss of enantiomeric excess. The thus obtained products are potential in the construction of other dihydrocoumarin-fused dihydropyran derivatives. Considering the wide existence of coumarins and 2H-pyran moieties in natural products and pharmaceuticals, the thus obtained optically active dihydrocoumarin-fused dihydropyranones should be of interest to medicinal chemists.

3. Materials and Methods

3.1. General Information

Reactions were monitored by thin layer chromatography using UV light or KMnO4 to visualize the course of reaction. Purification of reaction products was carried out by flash chromatography on silica gel. Chemical yields refer to pure isolated substances. The [α]D was recorded using PolAAr 3005 High Accuracy Polarimeter (Optical Activity Ltd., Huntingdon, England). Infrared (IR) spectra were obtained using a Bruker tensor 27 infrared spectrometer (Bruker, Borken, Germany). 1H, 13C and 19F NMR spectra were obtained using Bruker DPX-400 spectrometer (Bruker UK Limited, Coventry, UK). Chiral HPLC analyses were obtained using Agilent Technologies 1260 Infinity series (Agilent Technologies, Inc., Waldbronn, Germany) and DAICEL CHIRALPAK columns (CPI Company, Tokyo, Japan). Chemical shifts were reported in ppm from tetramethylsilane with the solvent resonance as the internal standard. The following abbreviations were used to designate chemical shift multiplicities: s = singlet, d = doublet, t = triplet, q = quartet, h = heptet, m = multiplet, br = broad.

3.2. Tertiary Amine Catalyzed Asymmetric [4 + 2] Cyclization

General procedure: To a 4 mL vial was sequentially added 3-aroylcoumarines 1 (0.2 mmol), catalyst 4l (0.02 mmol, 10 mol%), and EtOAc (1.0 mL); the mixture was stirred at 0 °C for 15 min before benzyl buta-2,3-dienoate 2b (0.3 mmol, 1.5 equiv.) was charged. The reaction was monitored by TLC analysis. After completion of the reaction, the solvent was removed by rotary evaporation and the residue was directly subjected to column chromatography using PE/EtOAc (20:1–15:1) as the eluent to afford product 3.
Benzyl (E)-2-(5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene) acetate (3a).
White solid (m.p. 121.4–122.1 °C); 1H NMR (400 MHz, CDCl3) δ 7.51–7.31 (m, 12H), 7.20 (t, J = 7.2 Hz, 1H), 7.10 (d, J = 8.0 Hz, 1H), 5.90 (d, J = 1.6 Hz, 1H), 5.23 (s, 2H), 4.87 (dd, J = 14.8, 6.0 Hz, 1H), 4.01 (dd, J = 12.2, 5.6 Hz, 1H), 2.56–2.49 (m, 1H); 13C NMR (101 MHz, CDCl3) δ 166.6, 164.2, 162.3, 161.5, 150.8, 136.0, 133.1, 130.9, 129.2, 129.1, 128.8, 128.5, 128.3, 128.3, 125.8, 124.9, 122.7, 117.3, 102.2, 101.5, 66.4, 30.3, 26.1; [α]D26.0 = + 44.3 (c = 0.26, CHCl3); The enantiomeric purity of 3a was determined by HPLC analysis (DAICEL CHIRALPAK AD–H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 11.2 min (major) and 14.5 min (minor)); HRMS (ESI): Exact mass calcd for C27H21O5 [M+H]+: 425.1389, found: 425.1392.
Benzyl (E)-2-(9-methoxy-5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene) acetate (3b).
White solid (m.p. 131.1–132.8 °C); 1H NMR (400 MHz, CDCl3) δ 7.51–7.36 (m, 10H), 7.03 (d, J = 8.8 Hz, 1H), 6.91 (d, J = 2.0 Hz, 1H), 6.85 (dd, J = 8.8, 2.8 Hz, 1H), 5.91 (d, J = 1.6 Hz, 1H), 5.23 (s, 2H), 4.80 (dd, J = 14.6, 5.6 Hz, 1H), 3.98 (dd, J = 12.0, 5.6 Hz, 1H), 3.83 (s, 3H), 2.58–2.51(m,1H); 13C NMR (101 MHz, CDCl3) δ 166.6, 164.1, 162.1, 161.9, 156.8, 144.7, 136.0, 133.0, 130.8, 129.1, 128.8, 128.5, 128.3, 128.3, 123.7, 118.0, 114.2, 111.1, 102.2, 101.4, 66.4 56.0, 30.5, 25.9; [α]D26.0 = +242.2 (c = 0.49, CHCl3); The enantiomeric purity of 3b was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 15.2 min (major) and 21.0 min (minor)); HRMS (ESI): Exact mass calcd for C28H23O6 [M+H]+: 455.1495, found: 455.1496.
Benzyl (E)-2-(7-methoxy-5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3c).
White solid (m.p. 108.7–110.2 °C); 1H NMR (400 MHz, CDCl3) δ 7.53–7.34 (m, 10H), 7.14 (t, J = 8.4 Hz, 1H), 6.95 (dd, J = 13.0, 8.4 Hz, 2H), 5.91 (d, J = 1.2 Hz, 1H), 5.24 (s, 2H), 4.83 (dd, J = 14.2, 6.0 Hz, 1H), 4.00 (dd, J = 12.4, 6.0 Hz, 1H), 3.90 (s, 3H), 2.57–2.50 (m, 1H); 13C NMR (101 MHz, CDCl3) δ 166.5, 164.2, 162.1, 161.0, 147.8, 140.2, 136.0, 132.9, 130.8, 129.1, 128.7, 128.4, 128.3, 128.2, 124.7, 123.8, 117.0, 111.7, 102.1, 101.2, 66.3, 56.2, 30.5, 26.0; [α]D26.0 = +86.5 (c = 0.5, CHCl3); The enantiomeric purity of 3c was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 13.0 min (major) and 17.0 min (minor)). HRMS (ESI): Exact mass calcd for C28H23O6 [M+H]+: 455.1495, found: 455.1493.
Benzyl (E)-2-(8-methoxy-5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3d).
White solid (m.p. 101.9–102.5 °C); 1H NMR (400 MHz, CDCl3) δ 7.50–7.34 (m, 10H), 7.27–7.24 (m, 1H), 6.74 (dd, J = 8.6, 2.4 Hz, 1H), 6.63 (d, J = 2.4 Hz, 1H), 5.87 (d, J = 1.6 Hz, 1H), 5.21 (s, 2H), 4.81 (dd, J = 14.8, 5.6 Hz, 1H), 3.92 (dd, J = 12.4, 5.6 Hz, 1H), 3.80 (s, 3H), 2.47–2.44 (m, 1H); 13C NMR (101 MHz, CDCl3) δ 166.6, 164.3, 162.2, 161.5, 160.3, 151.5, 136.0, 133.1, 130.8, 129.0, 128.8, 128.4, 128.3, 126.5, 114.5, 111.1, 102.5, 102.0, 101.7, 66.3, 55.7, 29.7, 26.4; [α]D26.0 = −44.3 (c = 0.26, CHCl3); The enantiomeric purity of 3d was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 12.2 min (major) and 17.7 min (minor)). HRMS (ESI): Exact mass calcd for C28H23O6 [M+H]+: 455.1495, found: 455.1498.
Benzyl (E)-2-(9-methyl-5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3e).
White solid (m.p. 127.6–128.9 °C); 1H NMR (400 MHz, CDCl3) δ 7.51–7.35 (m, 10H), 7.18 (s, 1H), 7.12 (d, J = 8.4 Hz, 1H), 6.99 (d, J = 8.0 Hz, 1H), 5.91 (d, J = 1.6 Hz, 1H), 5.24 (s, 2H), 4.87 (dd, J = 14.8, 5.6 Hz, 1H), 3.97 (dd, J = 12.4, 5.6 Hz, 1H), 2.53–2.46 (m, 1H), 2.37 (s,3H); 13CNMR (101 MHz, CDCl3) δ 166.7, 164.4, 162.1, 161.7, 148.7, 136.0, 134.6, 133.1, 130.8, 129.7, 129.0, 128.8, 128.5, 128.3, 128.2, 126.2, 122.2, 117.0, 102.0, 101.6, 66.4, 30.2, 26.1, 21.0; [α]D26.0 = +155.3 (c = 0.50, CHCl3); The enantiomeric purity of 3e was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 12.7 min (major) and 18.5 min (minor)). HRMS (ESI): Exact mass calcd for C28H23O5 [M+H]+: 439.1545, found: 439.1548.
Benzyl (E)-2-(7,9-di-tert-butyl-5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3f).
Yellow solid (m.p. 83.7–84.9 °C); 1H NMR (400 MHz, CDCl3) δ 7.54–7.37 (m, 11H), 7.28 (s, 1H), 5.95 (s, 1H), 5.28 (s, 2H), 4.76 (dd, J = 14.8, 6.0 Hz, 1H), 4.01 (dd, J = 11.6, 6.0 Hz, 1H), 2.77 (t, J = 13.2 Hz, 1H), 1.49 (s, 9H), 1.39 (s, 9H); 13C NMR (101 MHz, CDCl3) δ 166.6, 164.1, 162.1, 161.9, 156.8, 144.7, 136.0, 133.0, 130.9, 129.1, 128.8, 128.5, 128.3, 128.3 123.7, 118.0, 114.2, 111.1, 102.2, 101.4, 66.4, 56.0, 30.5, 25.9; [α]D26.0 = +58.6 (c = 0.52, CHCl3); The enantiomeric purity of 3f was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 6.0 min (major) and 5.1 min (minor)). HRMS (ESI): Exact mass calcd for C35H37O5 [M+H]+: 537.2641, found: 537.2643.
Benzyl (E)-2-(9-fluoro-5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3g).
Yellow solid (m.p. 118.4–119.7 °C); 1H NMR (400 MHz, CDCl3) δ 7.53–7.39 (m, 10H), 7.11–7.00 (m, 3H), 5.93 (d, J = 1.6 Hz, 1H), 5.25 (s, 2H), 4.80 (dd, J = 14.6, 5.6 Hz, 1H), 3.97 (dd, J = 12.4, 5.6 Hz, 1H), 2.56–2.49 (m, 1H); 13C NMR (101 MHz, CDCl3) δ 166.3, 163.5, 162.6, 161.1, 159.4 (d, 1J = 243 Hz), 146.7, 135.9, 132.8, 130.9, 129.0, 128.7, 128.4, 128.2, 128.2, 124.3, 124.2, 118.5, 118.4, 115.8 (d, 2J = 23 Hz), 112.6 (d, 3J = 25 Hz), 102.4, 100.4, 66.4, 30.3, 25.7; [α]D26.0 = −22.8 (c = 0.47, CHCl3); The enantiomeric purity of 3g was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 11.4 min (major) and 14.1 min (minor)). HRMS (ESI): Exact mass calcd for C27H20FO5 [M+H]+: 443.1295, found: 443.1297.
Benzyl (E)-2-(9-chloro-5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3h).
White solid (m.p. 127.4–128.9 °C); 1H NMR (400 MHz, CDCl3) δ 7.50–7.47 (m, 3H), 7.43–7.40 (m, 8H), 7.39–7.28 (m, 1H), 7.03 (d, J = 8.4 Hz, 1H), 5.92 (d, J = 2.0 Hz, 1H), 5.24 (s, 2H), 4.83 (dd, J = 14.6, 5.6 Hz, 1H), 3.98 (dd, J = 12.4, 5.6 Hz, 1H), 2.55–2.47 (m, 1H); 13C NMR (101 MHz, CDCl3) δ 166.4, 163.5, 162.9, 161.0, 149.3, 135.9, 132.8, 131.0, 130.1, 129.2, 129.0, 128.8, 128.5, 128.3, 128.3, 125.9, 124.3, 118.6, 102.6, 100.4, 66.5, 30.3, 25.8; [α]D26.0 = −56.8 (c = 0.19, CHCl3); The enantiomeric purity of 3h was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 13.7 min (major) and 16.4 min (minor)). HRMS (ESI): Exact mass calcd for C27H20ClO5 [M+H]+: 459.0999, found: 459.0997.
Benzyl (E)-2-(9-bromo-5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3i).
White solid (m.p. 131.4–132.7 °C); 1H NMR (400 MHz, CDCl3) δ 7.51–7.35 (m, 12H), 6.98 (d, J = 8.8 Hz, 1H), 5.92 (d, J = 1.2 Hz, 1H), 5.24 (s, 2H), 4.83 (dd, J = 14.6, 6.0 Hz, 1H), 3.99 (dd, J = 12.2, 6.0 Hz, 1H), 2.55–2.48 (m, 1H); 13C NMR (101 MHz, CDCl3) δ 166.4, 163.5, 162.9, 160.9, 149.9, 135.9, 132.8, 132.2, 131.0, 129.0, 128.8, 128.8, 128.5, 128.3, 128.3, 124.8, 119.0, 117.6, 102.6, 100.4, 66.5, 30.3, 25.8; [α]D26.0 = −17.8 (c = 0.22, CHCl3); The enantiomeric purity of 3i was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 17.1 min (major) and 14.2 min (minor)). HRMS (ESI): Exact mass calcd for C27H20BrO5 [M+H]+: 503.0494, found: 503.0490.
Benzyl (E)-2-(8-bromo-5-oxo-4-phenyl-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3j).
White solid (m.p. 119.2–120.7 °C); 1H NMR (400 MHz, CDCl3) δ 7.49–7.46 (m, 3H), 7.45–7.29 (m, 8H),7.24–7.20 (m, 2H), 5.90 (d, J = 2.0 Hz, 1H), 5.21 (s, 2H), 4.80 (dd, J = 14.8, 5.6 Hz, 1H), 3.90 (dd, J = 12.2, 5.6 Hz, 1H), 2.51–2.44 (m, 1H); 13C NMR (101 MHz, CDCl3) δ 166.4, 163.6, 162.8, 160.7, 151.2, 135.9, 132.7, 130.9, 129.0, 128.7, 128.4, 128.3, 127.8, 127.1, 122.0, 121.8, 120.4, 102.4, 100.5, 66.4, 30.0, 25.8; [α]D26.0 = +67.4 (c = 0.52, CHCl3); The enantiomeric purity of 3j was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 14.7 min (major) and 18.0 min (minor)). HRMS (ESI): Exact mass calcd for C27H20BrO5 [M+H]+: 503.0494, found: 503.0497.
Benzyl (E)-2-(5-oxo-4-(p-tolyl)-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3k).
Yellow solid (m.p. 133.2–133.9 °C); 1H NMR (400 MHz, CDCl3) δ 7.44–7.32 (m, 9H), 7.24–7.19 (m, 3H), 7.11 (d, J = 8.0 Hz, 1H), 5.91 (d, J = 1.2 Hz, 1H), 5.24 (s, 2H), 4.87 (dd, J = 14.8, 5.6 Hz, 1H), 3.99 (dd, J = 12.2, 6.0 Hz, 1H), 2.55–2.48 (m, 1H), 2.41 (s, 3H); 13C NMR (101 MHz, CDCl3) δ 166.5, 164.3, 162.4, 161.7, 150.8, 141.3, 136.0, 130.0, 129.1, 129.0, 128.7, 128.4, 128.3, 125.7, 124.8, 122.8, 117.2, 102.0, 100.9, 66.3, 30.3, 26.0, 21.7; [α]D26.0 = +32.2 (c = 0.54, CHCl3); The enantiomeric purity of 3k was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 12.1 min (major) and 18.0 min (minor)). HRMS (ESI): Exact mass calcd for C28H23O5 [M+H]+: 439.1545, found: 439.1543.
Benzyl (E)-2-(5-oxo-4-(m-tolyl)-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3l).
Yellow solid (m.p. 135.6–136.8 °C); 1H NMR (400 MHz, CDCl3) δ 7.43–7.31 (m, 9H), 7.24–7.19 (m, 3H), 7.11 (d, J = 8.4 Hz, 1H), 5.91 (d, J = 1.2 Hz, 1H), 5.24 (s, 2H), 4.87 (dd, J = 14.8, 6.0 Hz, 1H), 3.99 (dd, J = 12.4, 5.6 Hz, 1H). 2.55–2.48 (m, 1H), 2.41 (s, 3H); 13C NMR (101 MHz, CDCl3) δ 166.5, 164.3, 162.4, 161.7, 150.8, 141.3, 136.0, 130.0, 129.1, 129.0, 128.7, 128.4, 128.3, 125.7, 124.8, 122.8, 117.2, 102.0, 100.9, 66.3, 30.3, 26.0, 21.7; [α]D26.0 = +42.2 (c = 0.52, CHCl3); The enantiomeric purity of 3l was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 9.9 min (major) and 13.2 min (minor)). HRMS (ESI): Exact mass calcd for C28H23O5 [M+H]+: 439.1545, found: 439.1544.
Benzyl (E)-2-(4-(4-chlorophenyl)-5-oxo-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3m).
Yellow solid (m.p. 122.5–123.7 °C); 1H NMR (400 MHz, CDCl3) δ 7.46–7.32 (m, 11H), 7.23–7.19 (m, 1H), 7.10 (d, J = 8.0 Hz, 1H), 5.90 (d, J = 1.6 Hz, 1H), 5.23 (s, 2H), 4.86 (dd, J = 14.8, 6.0 Hz, 1H), 4.00 (dd, J = 12.4, 5.6 Hz, 1H), 2.55–2.48 (m, 1H); 13C NMR (101 MHz, CDCl3) δ 166.4, 163.9, 161.4, 161.1, 150.7, 137.0, 136.0, 130.5, 129.3, 128.8, 128.6, 128.5, 128.3, 125.8, 125.0, 122.5, 117.3, 102.3, 101.9, 66.4, 30.3, 26.0; [α]D26.0 = −148.3 (c = 0.49, CHCl3); The enantiomeric purity of 3m was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 13.9 min (major) and 21.5 min (minor)). HRMS (ESI): Exact mass calcd for C27H20ClO5 [M+H]+: 459.0999, found: 459.0999.
Benzyl (E)-2-(4-(3-chlorophenyl)-5-oxo-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3n).
Yellow solid (m.p. 124.5–125.3 °C); 1H NMR (400 MHz, CDCl3) δ 7.49–7.34 (m, 11H), 7.23–7.21 (m, 1H), 7.11–7.09 (m, 1H), 5.92 (d, J = 2.0 Hz, 1H), 5.24 (d, J = 2.0 Hz, 2H), 4.90–4.85 (m, 1H), 4.01–3.98 (m, 1H), 2.52 (t, J = 12.8, 1H); 13C NMR (101 MHz, CDCl3) δ 166.3, 163.8, 161.1, 160.6, 150.6, 135.9, 134.7, 134.2, 131.4, 130.8, 129.5, 129.2, 129.0, 128.7, 128.4, 128.3, 127.4, 125.8, 125.0, 122.3, 117.2, 102.4, 102.2, 66.4, 30.2, 25.9; [α]D26.0 = +100.7 (c = 0.55, CHCl3); The enantiomeric purity of 3n was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 12.4 min (major) and 13.9 min (minor)). HRMS (ESI): Exact mass calcd for C27H20ClO5 [M+H]+: 459.0999, found: 459.1003.
Benzyl (E)-2-(4-([1,1’-biphenyl]-4-yl)-5-oxo-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3o).
White solid (m.p. 130.7–131.4 °C); 1H NMR (400 MHz, CDCl3) δ 7.66–7.60(m, 6H), 7.49–7.33 (m, 10H), 7.22 (t, J = 7.6 Hz, 1H), 7.13 (d, J = 8.0 Hz, 1H), 5.95 (s, 1H), 5.26 (s, 2H), 4.89 (dd, J = 14.8, 5.6 Hz, 1H), 4.03 (dd, J = 12.0, 5.6 Hz, 1H), 2.55 (t, J = 14.0 Hz, 1H); 13C NMR (101 MHz, CDCl3) δ 166.5, 164.2, 162.0, 161.6, 150.8, 143.7, 140.3, 136.0, 129.6, 129.1, 128.9, 128.7, 128.4, 128.3, 127.3, 126.9, 125.8, 117.2, 102.1, 101.4, 66.3, 30.3, 26.0; [α]D26.0 = +91.1 (c = 0,34, CHCl3); The enantiomeric purity of 3o was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 17.9 min (major) and 31.4 min (minor)). HRMS (ESI): Exact mass calcd for C33H25O5 [M+H]+: 501.1702, found: 501.1708.
Benzyl (E)-2-(5-oxo-4-(thiophen-3-yl)-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (3p).
White solid (m.p. 135.7–136.2 °C); 1H NMR (400 MHz, CDCl3) δ 7.83 (d, J = 3.6 Hz, 1H), 7.53 (d, J = 5.2 Hz, 1H), 7.41–7.30 (m, 7H), 7.18 (t, J = 7.6 Hz, 1H), 7.09 (dd, J = 8.4, 4.8 Hz, 2H), 5.92 (d, J = 1.2 Hz, 1H), 5.27–5.20 (m,2H), 4.70 (dd, J = 15.0, 6.0 Hz, 1H), 4.02 (dd, J = 12.6, 5.6 Hz, 1H), 2.69–2.62 (m,1H); 13C NMR (101 MHz, CDCl3) δ 166.5, 164.0, 163.0, 161.6, 155.2, 150.6, 136.0, 133.6, 132.8, 130.7, 129.1, 128.8, 128.5, 128.3, 127.3, 125.7, 124.9, 122.6, 117.1, 101.8, 100.3, 66.4, 30.9, 29.8, 26.4; [α]D26.0 = +123.7 (c = 0.22, CHCl3); The enantiomeric purity of 3p was determined by HPLC analysis (DAICEL CHIRALPAK AD-H (0.46 cmφ × 25 cm), hexane:2-propanol = 60:40, flow rate = 0.75 mL/min, retention time: 12.4 min (major) and 14.6 min (minor)). HRMS (ESI): Exact mass calcd for C25H19O5S [M+H]+: 459.0999, found: 459.1003.

3.3. Synthesis of 5 from 1m and 2b

To the reaction mixture obtained under standard condition using 1m (1 mmol) and 2b (1.25 mmol) was added Pd/C (wt. 10%), then the mixture was stirred under H2 atmosphere (H2 balloon) at room temperature for 48 h. The resulting reaction mixture was filtered through a pad of Celite and eiluted with EtOAc. The filtration was concentrated under reduced pressure and the residue was purified by silica gel column chromatography (PE:EtOAc = 5:1–2:1) to afford the free acid intermediate, which was then dissolved in CH2Cl2 (5 mL). This solution was cooled to 0 °C before DCC (2.0 equiv.), DMAP (2.0 equiv.) and EtOH (2 mL) were added. After that, the reaction mixture was moved to rt and stirred overnight. After completion of the reaction by TLC analysis, the solvent was removed by rotary evaporation and the residue was directly subjected to column chromatography using PE/EtOAc (15:1-9:1) as the eluent to afford product 5.
Ethyl (E)-2-(4-(4-chlorophenyl)-5-oxo-1,10b-dihydro-2H,5H-pyrano[3,4-c]chromen-2-ylidene)acetate (5) [4].
1H NMR (400 MHz, CDCl3) δ 7.46–7.31 (m, 6H), 7.21 (t, J = 7.6 Hz, 1H), 7.09 (d, J = 8.0 Hz, 1H), 5.83 (s, 1H), 4.87 (dd, J = 14.8, 5.6 Hz, 1H), 4.24 (q, J = 7.2 Hz, 2H), 4.00 (dd, J = 12.0, 5.6 Hz, 1H), 2.50 (t, J = 14.4 Hz, 1H), 1.33 (t, J = 7.2 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 166.6, 163.4, 161.5, 161.1, 150.6, 136.9, 131.4, 130.5, 129.2, 128.6, 125.8, 125.0, 122.5, 117.2, 102.6, 101.7, 60.6, 30.3, 25.8, 14.4; [α]D26.0 = +78.6 (c = 0.38, CHCl3); The enantiomeric purity of 5 was determined by HPLC analysis (DAICEL CHIRALPAK OD-3 (4.6 mmφ × 150 mml), hexane:2-propanol = 80:20, flow rate = 0.75 mL/min, retention time: 6.1 min (major) and 7.6 min (minor)).

Supplementary Materials

The following are available online: 1H and 13C NMR spectra and HPLC data of compounds 3a-p and 5.

Author Contributions

X.-P.Z. conceived and designed the project and wrote the paper after discussing with Y.-Y.P. and C.-B.J.; J.-L.L., X.-H.W. and J.-C.S. performed the experiments and analyzed the data. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (21702080), Natural Science Foundation of Jiangxi Province of China (20181BAB213003), the Open Project Program of Key Laboratory of Functional Small Organic Molecule, Ministry of Education, Jiangxi Normal University (KLFS-KF-201603) and the foundation of Jiangxi Educational Committee (170223).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Samples of the compounds are not available from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Hu, X.-L.; Gao, C.; Xu, Z.; Liu, M.-L.; Feng, L.-S.; Zhang, G.-D. Recent Development of Coumarin Derivatives as Potential Antiplasmodial and Antimalarial Agents. Med. Chem. 2018, 18, 114–123. [Google Scholar] [CrossRef]
  2. Srikrishna, D.; Godugu, C.; Dubey, P.K. A Review on Pharmacological Properties of Coumarins. Mini-Rev. Med. Chem. 2018, 18, 113–141. [Google Scholar]
  3. Cao, D.; Liu, Z.; Verwilst, P.; Koo, S.; Jangjili, P.; Kim, J.S.; Lin, W. Coumarinbased Small-molecule Fluorescent Chemosensors. Chem. Rev. 2019, 119, 10403–10519. [Google Scholar] [CrossRef] [PubMed]
  4. Kasperkiewicz, K.; Ponczek, M.B.; Owczarek, J.; Guga, P.; Budzisz, E. Antagonists of Vitamin K-Popular Coumarin Drugs and New Synthetic and Natural Coumarin Derivatives. Molecules 2020, 25, 1465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Stefanachi, A.; Leonetti, F.; Pisani, L.; Catto, M.; Carotti, A. Coumarin: A Natural, Privileged and Versatile Scaffold for Bioactive Compounds. Molecules 2018, 23, 250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Wang, Z.-H.; Zhu, M.-W.; Chen, Q.-A.; Zhou, Y.-G. Catalytic Biomimetic Asymmetric Reduction of Alkenes and Imines Enabled by Chiral and Regenerable NAD(P)H Models. Angew. Chem. Int. Ed. 2019, 58, 1813–1817. [Google Scholar] [CrossRef]
  7. Zhang, H.; Luo, Y.; Li, D.-W.; Yao, Q.; Dong, S.-X.; Liu, X.H.; Feng, X.M. Enantioselective Synthesis of 4-Hydroxy-dihydrocoumarins via Catalytic Ring Opening/Cycloaddition of Cyclobutenones. Org. Lett. 2019, 21, 2388–2392. [Google Scholar] [CrossRef]
  8. Chen, Y.-R.; Ganapuram, M.R.; Hsieh, K.-H.; Chen, K.-H.; Karanam, P.; Vagh, S.S.; Liou, Y.-C.; Lin, W.-W. 3-Homoacyl Coumarin: An All Carbon 1,3-Dipole for Enantioselective Concerted (3 + 2) Cycloaddition. Chem. Commun. 2018, 54, 12702–12705. [Google Scholar] [CrossRef]
  9. Zhang, X.-Z.; Gan, K.-J.; Liu, X.-X.; Deng, Y.-H.; Wang, F.-X.; Yu, K.-Y.; Zhang, J.; Fan, C.-A. Enantioselective Synthesis of Functionalized 4-Aryl Hydrocoumarins and 4-Aryl Hydroquinolin-2-ones via Intramolecular Vinylogous Rauhut-currier Reaction of para-Quinone Methides. Org. Lett. 2017, 19, 3207–3210. [Google Scholar] [CrossRef]
  10. Jin, J.-H.; Li, X.-Y.; Luo, X.-Y.; Fossey, J.S.; Deng, W.-P. Asymmetric Synthesis of cis-3,4-Dihydrocoumarins via [4+ 2] Cycloadditions Catalyzed by Amidine Derivatives. J. Org. Chem. 2017, 82, 5424–5432. [Google Scholar] [CrossRef]
  11. Zhao, Y.-L.; Lou, Q.-X.; Wang, L.-S.; Hu, W.-H.; Zhao, J.-L. Organocatalytic Friedel-crafts Alkylation/Lactonization Reaction of Naphthols with 3-Trifluoroe-thylidene Oxindoles: The Asymmetric Synthesis of Dihydrocoumarins. Angew. Chem. Int. Ed. 2017, 56, 338–342. [Google Scholar] [CrossRef] [PubMed]
  12. Matthias, H.; Geoffrey, A.; Cordell, N.-R.; Payom, T. Candenatone, a Novel Purple Pigment from Dalbergia candenatensis. J. Org. Chem. 1988, 53, 4162–4165. [Google Scholar]
  13. Wang, W.-J.; Xue, J.-J.; Tian, T.; Zhang, J.; Wei, L.-P.; Shao, J.-D.; Xie, Z.-X.; Li, Y. Total Synthesis of (±)-δ-Rubromycin. Org. Lett. 2013, 15, 2402–2405. [Google Scholar] [CrossRef] [PubMed]
  14. Li, J.; Zhang, W.-H.; Zhang, F.; Chen, Y.; Li, A. Total Synthesis of Longeracinphyllin A. J. Am. Chem. Soc. 2017, 139, 14893–14896. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, Y.; Yu, H.-Z.; Zheng, H.-F.; Shi, D.-Q. DABCO and Bu3P catalyzed [4 + 2] and [3 + 2] cycloadditions of 3-acyl-2Hchromen-ones and ethyl 2,3-butadienoate. Org. Biomol. Chem. 2012, 10, 7739–7746. [Google Scholar] [CrossRef]
  16. Jian, T.-Y.; Chen, X.-Y.; Sun, L.-H.; Ye, S. N-heterocyclic carbene-catalyzed [4 + 2] cycloaddition of ketenes and 3-aroylcoumarins: Highly enantioselective synthesis of dihydrocoumarin-fused dihydropyranones. Org. Biomol. Chem. 2013, 11, 158–163. [Google Scholar] [CrossRef]
  17. Han, X.-Y.; Ni, H.-Z.; Chan, W.-L.; Gai, X.-K.; Wang, Y.-J.; Lu, Y.-X. Highly enantioselective synthesis of dihydrocoumarin-fused dihydropyrans via the phosphine-catalyzed [4 + 2] annulation of allenones with 3-aroylcoumarins. Org. Biomol. Chem. 2016, 14, 5059–5064. [Google Scholar] [CrossRef]
  18. Shi, M.-L.; Zhan, G.; Zhou, S.-L.; Du, W.; Chen, Y.-C. Asymmetric Inverse-Electron-Demand Oxa-Diels−Alder Reaction of Allylic Ketones through Dienamine Catalysis. Org. Lett. 2016, 18, 6480–6483. [Google Scholar] [CrossRef]
  19. Wang, F.; Li, Z.; Wang, J.; Li, X.; Cheng, J.-P. Enantioselective Synthesis of Dihydropyran-Fused Indoles through [4 + 2] Cycloaddition between Allenoates and 3-Olefinic Oxindoles. J. Org. Chem. 2015, 80, 5279–5286. [Google Scholar] [CrossRef]
  20. Wang, F.; Yang, C.; Xue, X.-S.; Li, X.; Cheng, J.-P. A Highly Efficient Chirality Switchable Synthesis of Dihydropyran-Fused Benzofurans by Fine-Tuning the Phenolic Proton of β-Isocupreidine (β-ICD) Catalyst with Methyl. Chem. Eur. J. 2015, 21, 10443–10449. [Google Scholar] [CrossRef]
  21. Heiko, D.; Vivek, K.; Herbert, W.; Kamal, K. Lewis Base Catalyzed [4 + 2] Annulation of Electron-Deficient Chromone-Derived Heterodienes and Acetylenes. Chem. Eur. J. 2011, 17, 5130–5137. [Google Scholar]
Figure 1. (a) DABCO catalyzed [4 + 2] cyclization of 3-aroylcoumarines with 2,3-butadienoate; (b) Chiral tertiary amine catalyzed [4 + 2] cyclization of 3-aroylcoumarines with 2,3-butadienoate
Figure 1. (a) DABCO catalyzed [4 + 2] cyclization of 3-aroylcoumarines with 2,3-butadienoate; (b) Chiral tertiary amine catalyzed [4 + 2] cyclization of 3-aroylcoumarines with 2,3-butadienoate
Molecules 26 00489 g001
Figure 2. Substrate scope of tertiary amine catalyzed asymmetric [4 + 2] cyclization.
Figure 2. Substrate scope of tertiary amine catalyzed asymmetric [4 + 2] cyclization.
Molecules 26 00489 g002
Figure 3. Z/E Configuration determination.
Figure 3. Z/E Configuration determination.
Molecules 26 00489 g003
Figure 4. Gram-scale synthesis and product elaboration.
Figure 4. Gram-scale synthesis and product elaboration.
Molecules 26 00489 g004
Table 1. Condition optimization for the catalytic asymmetric [4 + 2] cyclization.
Table 1. Condition optimization for the catalytic asymmetric [4 + 2] cyclization.
Molecules 26 00489 i001
EntryCat.SolventTemp. (°C)t (h)Yield (%)Ee (%)
1Quinine (4a)DCE2548639
2Cinchonine (4b)DCE25483316
34cDCE2548845
44dDCE25488327
54eDCE2548trace
64fDCE25482322
74gDCE2589652
84hDCE25118260
94iDCE25237835
104jDCE25236034
114kDCE25236729
124lDCE25118763
134mDCE25167847
144ln-C6H122548trace
154ltoluene25486879
164lTHF25487766
174lAcetone252.57958
184lEtOAc257.58779
194lCH3CN252.59556
204lDMF252.58853
214lMeOH2548trace--
224lEtOAc019379
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, J.-L.; Wang, X.-H.; Sun, J.-C.; Peng, Y.-Y.; Ji, C.-B.; Zeng, X.-P. Chiral Tertiary Amine Catalyzed Asymmetric [4 + 2] Cyclization of 3-Aroylcoumarines with 2,3-Butadienoate. Molecules 2021, 26, 489. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26020489

AMA Style

Li J-L, Wang X-H, Sun J-C, Peng Y-Y, Ji C-B, Zeng X-P. Chiral Tertiary Amine Catalyzed Asymmetric [4 + 2] Cyclization of 3-Aroylcoumarines with 2,3-Butadienoate. Molecules. 2021; 26(2):489. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26020489

Chicago/Turabian Style

Li, Jun-Lin, Xiao-Hui Wang, Jun-Chao Sun, Yi-Yuan Peng, Cong-Bin Ji, and Xing-Ping Zeng. 2021. "Chiral Tertiary Amine Catalyzed Asymmetric [4 + 2] Cyclization of 3-Aroylcoumarines with 2,3-Butadienoate" Molecules 26, no. 2: 489. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26020489

Article Metrics

Back to TopTop