Next Article in Journal
Soy Isoflavones Induce Cell Death by Copper-Mediated Mechanism: Understanding Its Anticancer Properties
Next Article in Special Issue
Recent Advances of N-2,2,2-Trifluoroethylisatin Ketimines in Organic Synthesis
Previous Article in Journal
Discrepancy Study of the Chemical Constituents of Panax Ginseng from Different Growth Environments with UPLC-MS-Based Metabolomics Strategy
Previous Article in Special Issue
Chiral Phosphine Catalyzed Allylic Alkylation of Benzylidene Succinimides with Morita–Baylis–Hillman Carbonates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Palladium-Catalyzed Stereoselective Construction of 1,3-Stereocenters Displaying Axial and Central Chirality via Asymmetric Alkylations

State Key Laboratory of Fine Chemicals, Department of Pharmaceutical Sciences, School of Chemical Engineering, Dalian University of Technology, 2 Linggong Road, Dalian 116024, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(7), 2927; https://doi.org/10.3390/molecules28072927
Submission received: 28 February 2023 / Revised: 22 March 2023 / Accepted: 22 March 2023 / Published: 24 March 2023
(This article belongs to the Special Issue Recent Advances of Catalytic Asymmetric Synthesis)

Abstract

:
The concurrent construction of 1,3-stereocenters remains a challenge. Herein, we report the development of stereoselective union of a point chiral center with allenyl axial chirality in 1,3-position by Pd-catalyzed asymmetric allenylic alkylation between racemic allenyl carbonates and indanone-derived β-ketoesters. Various target products bearing a broad range of functional groups were afforded in high yield (up to 99%) with excellent enantioselectivities (up to 98% ee) and good diastereoselectivities (up to 13:1 dr).

1. Introduction

Over the past decades, extensive demands for chiral non-racemic compounds from various fields have significantly motivated the development of asymmetric catalysis [1,2,3,4,5]. Hence numerous catalytic asymmetric methodologies have been developed for enantioselective construction of chiral structures [6,7,8,9]. Whereas many methods are available for the synthesis of chiral molecules containing one single stereocenter or even two adjacent stereocenters (Scheme 1a), much less efforts have been paid to the concurrent creation of nonadjacent stereocenters in an enantio- and diastereoselective manner, due in part to the difficulties in high levels of simultaneous stereocontrol posed by increased distance between the two chiral centers [10,11,12,13,14]. Of note, the limited reports on enantio- and diastereoselective construction of 1,3-stereocenters focused mostly on molecules bearing two point chiral centers, while those containing different types of chirality, for example, central and axial chiral motifs, have been rarely explored.
As a prominent example of axial chiral molecules, chiral allenes constitute an important structural motif widely present in a variety of organic molecules including natural products, pharmaceutical agents, chiral catalysts, and ligands for coordination chemistry etc., as represented by the selected compounds in Scheme 1b [15,16,17,18,19]. Interestingly, Enprostil [20], a prostaglandin analogue used for the treatment of acute duodenal ulcer disease, shows a distinct 1,3-stereocenters consisting of a point chiral center and allenyl axial chirality.
Functionalized alkynes are most commonly employed for the synthesis of chiral allenes by classic synthetic methods such as addition, elimination, substitution, and rearrangement (Scheme 1c) [5,21,22,23,24,25]. Moreover, transition metal-catalyzed 1,4-addition of enynes serves as a powerful strategy for the concise and efficient synthesis of multi-substituted chiral allenes [26,27,28,29,30]. It is worth noting that although great progress has been achieved for the enantioselective construction of allenyl axial chirality [31,32,33,34,35,36,37,38], only two reports documented the asymmetric concurrent creation of 1,3-stereocenters bearing allenyl axial chirality and central chirality, contributed by the Trost group [11] and the Ma/Zhang group [39], respectively. Both studies exploited the strategy of Pd-catalyzed asymmetric allenylic alkylation employing racemic allenyl acetate electrophile through dynamic kinetic transformation. In connection with our interest in the asymmetric functionalization of indanone derivatives [40,41], herein we report Pd-catalyzed asymmetric allenylic alkylation between racemic allenylic carbonates and indanone-derived β-ketoesters to construct 1,3-stereocenters bearing allenyl axial and central chirality (Scheme 1d), providing alkylation products in good yields (up to 99%) with excellent enantioselectivities (up to 98% ee) and good diastereoselectivities (up to 13:1 dr).

2. Results and Discussion

2.1. Optimization of the Reaction Conditions

In our preliminary investigation, allenylic carbonate 2a was selected as the model substrate for the asymmetric alkylation with β-ketoester 1a in the presence of Cs2CO3 using a palladium catalyst (Table 1, entries 1–26). Initially, several classic biphosphine ligands were screened to explore the effect of the chiral ligands on the Pd-catalyzed allenylic alkylation reaction. With L1 or L3 as the ligand, product 3a could be afforded in high yields (84% and 89%) with good enantioselectivities (−71% ee and −67% ee) (entries 1 and 3). In the reaction with Trost ligand L2, no product was observed (entry 2), while L4 afforded product 3a in 55% yield with −73% ee and 15:1 dr (entry 4). Further screening led to the observation that ligands L5L7 were sluggish for this allenylation, whereas ligand L6 afforded the product 3a with high diastereoselectivity (11:1) but low yield and enantioselectivity (20%, −7% ee) (entry 6). Then, the ligand (R)- and (S)-SegPhos were applied in this reaction. Interestingly, using ligand (S)-SegPhos (L9) the product 3a was obtained in 90% yield with 4:1 dr and 91% ee (entry 9). Subsequently, increasing the amount of 2a further increased the yield of 3a to 97%, and the enantioselectivity of 3a to 93%, respectively (entry 10). With L9 as the ligand, subsequent investigation on the solvent effect showed that THF was the best one, affording 97% yield 3a with 6:1 dr and 92% ee (entry 15). Next, we investigated the effect of base, and observed that the product 3a was formed in 98% yield, with 7:1 dr and 93% ee (entry 20), when the base of the reaction was NaHCO3. Further optimization was performed and found that the reaction at 0.05 M could give product 3a in 98% yield with 8:1 dr and 96% ee at −10 °C.

2.2. Substrate Scope for the Asymmetric Alkylations of β-Ketoester 1

Having established the optimized reaction conditions (Table 1, entry 26), the scope of β-ketoester 1 was then examined with allenylic carbonate 2a (Scheme 2). We evaluated the electronic nature and position of the substituents on the 1H-indanone backbone of the substrates. The products 3b, 3c, and 3g bearing electron-donating substituents were obtained in high yields (93–96% yield) with excellent enantioselectivities (95–96% ee) and good diastereoselectivities (6:1–9:1 dr). Furthermore, the disubstituted substrate 1d was also successfully transformed into 3d in 99% yield, with 97% ee and 9:1 dr. Meanwhile, substrates (1e, 1f, and 1h1l) bearing halogen groups also engaged in the asymmetric alkylation reaction well and afforded products (3e, 3f, and 3h3l) in 96–99% yields, with high enantioselectivities (95–97% ee) and diastereoselectivities (7:1–12:1 dr). Notably, the six-membered cyclic β-ketoester (1m) underwent the alkylation reaction smoothly to afford product 3m in high yield and excellent enantioselectivity (99%, 97% ee), though with a moderate dr (2:1).

2.3. Substrate Scope for the Asymmetric Alkylations of Allenylic Carbonates 2

Next, the substrate scope with respect to allenylic carbonates 2 were investigated (Scheme 3). The electronic nature and position of the substituents on the arylallenyl carbonates backbone of the substrates were also examined. When a chloro group was present at the C2 position in the aryl, high yield and enantioselectivity were observed for the formation of the alkylation product 3n (80% yield, 97% ee). The products (3o and 3p) bearing electron-donating substituents were obtained in 99% yield with 93–97% ee. Meanwhile, halogenation in the aryl C4 position of the substrates (2f and 2g) were well-tolerated, giving the corresponding products (3r and 3s) with 91–97% ee and a slight decrease in yield (84–87%). Moreover, 2-naphthyl-substituted allenylic carbonate also worked smoothly (73% yield, 97% ee). Compared to model substrate 2a, various substituted arylallenyl carbonates (2b2h) could afford the target products (3n3t) with relatively higher diastereoselectivities as expected. In addition, substrates 2i with an n-propyl group and 2j with isopropyl afforded products 3u and 3v with slightly decreased enantioselectivities and diastereoselectivities. The asymmetric alkylation of allenylic carbonate with symmetrical cyclohexyl substituent delivered the product 3w in moderate enantioselectivity and diastereoselectivity, albeit with a lower yield.

2.4. Gram-Scale Reaction and Product Derivatization

To demonstrate the practicality of the transformation, a gram-scale reaction was conducted and found that the reaction of β-ketoester 1j (2.5 mmol) with allenylic carbonate 2a (3.0 mmol) gave product 3j (1.01 g) in 98% yield, with 97% ee and 12:1 dr (Scheme 4). Target product 3x was found to undergo Suzuki coupling smoothly with phenyl boronic acid, affording compound 4a (Scheme 4) with a biphenyl motif (70% yield, 92% ee, 9:1 dr) [42].

2.5. Plausible Mechanism of the Palladium-Catalyzed Alkylation of β-Ketoester 1

A plausible mechanism for the palladium-catalyzed allenylic alkylation of indanone β-ketoester is depicted in Scheme 5 [43,44]. Oxidative addition of 2a with Pd(0) would immediately undergo delocalization to yield intermediate A. In the presence of NaHCO3, the nucleophile 1a attacks the terminal carbon atom of intermediate A to afford the target product 3a and regenerate Pd(0).

3. Materials and Methods

3.1. General Information

Unless otherwise noted, materials were purchased from commercial suppliers and used without further purification. Column chromatography was performed on silica gel (200~300 mesh). Enantiomeric excesses (ee) were determined by HPLC (Agilent, Palo Alto, CA, USA) using corresponding commercial chiral columns as stated at 30 °C with UV detector at 254 nm. Optical rotations (JiaHang Instruments, Shanghai, China) were reported as follows: [ α ] D T (c g/100 mL, solvent). All 1H NMR and 19F NMR spectra were recorded on a Bruker Avance II 400 MHz (Bruker, Karlsruhe, Germany) and Bruker Avance III 600 MHz (Bruker, Karlsruhe, Germany), respectively, 13C NMR spectra were recorded on a Bruker Avance II 101 MHz or Bruker Avance III 151 MHz with chemical shifts reported as ppm (in CDCl3, TMS as an internal standard). Data for 1H NMR are recorded as follows: chemical shift (δ, ppm), multiplicity (s = singlet, d = doublet, t = triplet, m = multiplet, br = broad singlet, dd = double doublet, coupling constants in Hz, integration). HRMS (ESI) was obtained with a HRMS/MS instrument (LTQ Orbitrap XL TM, Agilent, Palo Alto, CA, USA). The characterization data is available in Supplementary Material.

3.2. Procedure for the Synthesis of Compounds 3

(S)-SegPhos (L9) (5 mol%) and Pd2dba3 (2.5 mol%) were stirred in THF (4 mL) in a Schlenk tube under a nitrogen atmosphere at room temperature for 10 min. To this Schlenk tube were added 1 (0.20 mmol, 1.0 equiv), NaHCO3 (0.20 mmol, 1.0 equiv), and 2 (0.24 mmol, 1.2 equiv), then the reaction mixture was stirred at −10 °C. When compound 1 was consumed as checked by TLC, the reaction was stopped and purified by column chromatography (petroleum ether/ethyl acetate = 30:1) on silica gel directly to give product 3.
Ethyl 1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3a)
Prepared according to the procedure within 48 h as light yellow liquid (65.8 mg, 99% yield, dr = 8:1). [ α ] D 17 = 84.324 (c 0.37, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.77 (d, J = 7.7 Hz, 1H), 7.06–7.55 (m, 1H), 7.51–7.33 (m, 3H), 7.31–7.27 (m, 1H), 7.21–7.16 (m, 3H), 6.08 (dt, J = 5.9, 2.7 Hz, 1H), 5.48 (q, J = 6.9 Hz, 1H), 4.09 (q, J = 7.1 Hz, 2H), 3.75 (d, J = 17.4 Hz, 1H), 3.24 (d, J = 17.4 Hz, 1H), 3.01 (ddd, J = 14.7, 7.0, 2.8 Hz, 1H), 2.68 (ddd, J = 14.7, 7.2, 2.7 Hz, 1H), 1.17 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.5, 201.8, 170.4, 153.2, 135.3, 135.3, 134.1, 128.5, 127.7, 127.0, 126.8, 126.4, 124.7, 95.5, 90.0, 61.8, 60.3, 36.5, 34.2, 14.0. HRMS (ESI) m/z Calcd. for C22H20NaO3 ([M + Na]+) 335.1305, found 335.1298. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 26.2 min, tminor = 24.5 min).
Ethyl 6-methyl-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3b)
Prepared according to the procedure within 60 h as light yellow liquid (64.4 mg, 93% yield, dr = 9:1). [ α ] D 17 = 77.698 (c 0.56, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.58 (s, 1H), 7.45–7.39 (m, 1H), 7.36–7.26 (m, 3H), 7.24–7.18 (m, 3H), 6.10 (dt, J = 6.1, 2.7 Hz, 1H), 5.49 (q, J = 6.8 Hz, 1H), 4.11 (q, J = 7.1 Hz, 2H), 3.71 (d, J = 17.2 Hz, 1H), 3.21 (d, J = 17.2 Hz, 1H), 3.01 (ddd, J = 14.6, 7.1, 2.8 Hz, 1H), 2.70 (ddd, J = 14.5, 7.2, 2.6 Hz, 1H), 2.42 (s, 3H), 1.19 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.5, 202.0, 170.6, 150.7, 137.7, 136.7, 135.4, 134.1, 128.5, 127.0, 126.8, 126.1, 124.6, 95.4, 90.1, 61.7, 60.6, 36.2, 34.2, 21.1, 14.0. HRMS (ESI) m/z Calcd. for C23H22NaO3 ([M + Na]+) 369.1461, found 369.1456. Enantiomeric excess was determined to be 95% (determined by HPLC using chiral AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 27.9 min, tminor = 22.4 min).
Ethyl 6-methoxy-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3c)
Prepared according to the procedure within 48 h as light yellow liquid (68.8 mg, 95% yield, dr = 8:1). [ α ] D 17 = 71.161 (c 0.27, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.40–7.32 (m, 4H), 7.27 (d, J = 1.9 Hz, 2H), 7.25–7.15 (m, 2H), 6.16 (dt, J = 5.8, 2.7 Hz, 1H), 5.55 (q, J = 6.9 Hz, 1H), 4.18 (q, J = 7.1 Hz, 2H), 3.91 (s, 3H), 3.72 (d, J = 17.0 Hz, 1H), 3.24 (d, J = 17.0 Hz, 1H), 3.06 (ddd, J = 14.6, 7.1, 2.8 Hz, 1H), 2.78 (ddd, J = 14.6, 7.2, 2.7 Hz, 1H), 1.25 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.5, 201.9, 170.5, 159.7, 146.2, 136.5, 134.1, 128.5, 127.1, 127.0, 126.8, 124.9, 105.7, 95.4, 90.0, 61.7, 61.0, 55.6, 35.8, 34.2, 14.0. HRMS (ESI) m/z Calcd. for C23H22NaO4 ([M + Na]+) 385.1410, found 385.1404. Enantiomeric excess was determined to be 95% (determined by HPLC using chiral OD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 31.3 min, tminor = 35.8 min).
Ethyl 5,6-dimethoxy-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3d)
Prepared according to the procedure within 72 h as light yellow liquid (77.6 mg, 99% yield, dr = 9:1). [ α ] D 17 = 100.27 (c 0.74, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.32–7.22 (m, 2H), 7.20–7.10 (m, 4H), 6.79 (s, 1H), 6.05 (dt, J = 6.1, 2.8 Hz, 1H), 5.48 (q, J = 6.9 Hz, 1H), 4.10 (q, J = 7.1 Hz, 2H), 3.92 (s, 3H), 3.89 (s, 3H), 3.63 (d, J = 17.1 Hz, 1H), 3.20–3.09 (m, 1H), 2.97 (ddd, J = 14.8, 7.1, 2.7 Hz, 1H), 2.71 (ddd, J = 14.7, 7.0, 2.8 Hz, 1H), 1.18 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.3, 200.3, 170.7, 156.0, 149.7, 148.8, 134.1, 128.5, 128.0, 126.9, 126.8, 107.3, 104.9, 95.4, 90.1, 61.7, 60.6, 56.2, 56.1, 36.1, 34.0, 14.0. HRMS (ESI) m/z Calcd. for C24H24NaO5 ([M + Na]+) 415.1516, found 415.1507. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral OD-H column, hexane/2-propanol = 7/3, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 25.3 min, tminor = 33.9 min).
Ethyl 6-fluoro-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3e)
Prepared according to the procedure within 72 h as light yellow liquid (69.3 mg, 99% yield, dr = 7:1). [ α ] D 18 = 76.074 (c 0.65, CH2Cl2); 1H NMR (600 MHz, chloroform-d) δ 7.37 (dd, J = 7.4, 2.5 Hz, 1H), 7.33 (dd, J = 8.5, 4.3 Hz, 1H), 7.29–7.22 (m, 3H), 7.20–7.13 (m, 3H), 6.05 (dt, J = 6.0, 2.7 Hz, 1H), 5.47 (q, J = 6.7 Hz, 1H), 4.08 (qd, J = 7.1, 1.2 Hz, 2H), 3.67 (d, J = 17.1 Hz, 1H), 3.19 (d, J = 17.1 Hz, 1H), 2.97 (ddd, J = 14.9, 7.0, 2.8 Hz, 1H), 2.73 (ddd, J = 14.9, 7.0, 2.8 Hz, 1H), 1.16 (t, J = 7.2 Hz, 3H); 13C NMR (151 MHz, Chloroform-d) δ 206.3, 201.1, 170.1, 162.4 (d, J = 166.7 Hz), 148.6, 137.0, 133.9, 128.6, 127.9 (d, J = 5.1 Hz), 127.1, 126.8, 122.9, 110.3 (d, J = 15.2 Hz), 95.8, 89.9, 61.9, 61.2, 35.9, 34.0, 14.0; 19F NMR (376 MHz, Chloroform-d) δ −104.2–−123.6 (m). HRMS (ESI) m/z Calcd. for C22H19FNaO3 ([M + Na]+) 373.1210, found 373.1202. Enantiomeric excess was determined to be 96% (determined by HPLC using chiral IC-IB-H column, hexane/2-propanol = 9/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 33.4 min, tminor = 32.0 min).
Ethyl 6-chloro-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3f)
Prepared according to the procedure within 96 h as light yellow liquid (72.5 mg, 99% yield, dr = 8:1). [ α ] D 16 = 104.13 (c 0.70, CH2Cl2); 1H NMR (600 MHz, chloroform-d) δ 7.69 (d, J = 2.1 Hz, 1H), 7.47 (dd, J = 8.1, 2.1 Hz, 1H), 7.29 (d, J = 8.2 Hz, 1H), 7.28–7.24 (m, 2H), 7.20–7.16 (m, 1H), 7.15–7.12 (m, 2H), 6.04 (dt, J = 6.0, 2.7 Hz, 1H), 5.46 (q, J = 6.7 Hz, 1H), 4.08 (q, J = 7.1 Hz, 2H), 3.67 (d, J = 17.4 Hz, 1H), 3.18 (d, J = 17.4 Hz, 1H), 2.96 (ddd, J = 15.0, 7.0, 2.7 Hz, 1H), 2.74 (ddd, J = 14.9, 6.9, 2.8 Hz, 1H), 1.15 (t, J = 7.1 Hz, 3H); 13C NMR (151 MHz, Chloroform-d) δ 206.2, 200.7, 170.1, 151.3, 136.9, 135.2, 134.1, 133.8, 128.5, 127.6, 127.1, 126.8, 124.3, 95.9, 89.8, 62.0, 60.8, 36.0, 34.0, 14.0. HRMS (ESI) m/z Calcd. for C22H19ClNaO3 ([M + Na]+) 389.0915, found 389.0911. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 37.9 min, tminor = 30.4 min).
Ethyl 5-methyl-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3g)
Prepared according to the procedure within 72 h as light yellow liquid (66.5 mg, 96% yield, dr = 6:1); [ α ] D 16 = 69.372 (c 0.38, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.65 (d, J = 7.8 Hz, 1H), 7.30–7.23 (m, 2H), 7.18 (d, J = 7.1 Hz, 5H), 6.06 (dt, J = 5.9, 2.8 Hz, 1H), 5.47 (q, J = 6.9 Hz, 1H), 4.08 (q, J = 7.1 Hz, 2H), 3.68 (d, J = 17.4 Hz, 1H), 3.17 (d, J = 17.3 Hz, 1H), 2.98 (ddd, J = 14.7, 7.0, 2.8 Hz, 1H), 2.67 (ddd, J = 14.6, 7.2, 2.7 Hz, 1H), 2.40 (s, 3H), 1.16 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.4, 201.3, 170.6, 153.8, 146.8, 143.3, 134.1, 133.0, 129.1, 128.5, 127.0, 126.8, 124.5, 95.4, 90.1, 61.7, 60.4, 36.3, 34.2, 22.1, 14.0. HRMS (ESI) m/z Calcd. for C23H22NaO3 ([M + Na]+) 369.1461, found 369.1455. Enantiomeric excess was determined to be 96% (determined by HPLC using chiral AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 33.3 min, tminor = 30.3 min).
Ethyl 5-fluoro-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3h)
Prepared according to the procedure within 96 h as light yellow liquid (67.2 mg, 96% yield, dr = 10:1). [ α ] D 17 = 84.648 (c 0.47, CH2Cl2); 1H NMR (600 MHz, chloroform-d) δ 7.75 (dd, J = 8.4, 5.2 Hz, 1H), 7.31–7.25 (m, 2H), 7.22–7.15 (m, 3H), 7.08–7.01 (m, 2H), 6.06 (dt, J = 6.0, 2.8 Hz, 1H), 5.47 (q, J = 6.8 Hz, 1H), 4.09 (q, J = 7.1 Hz, 2H), 3.71 (d, J = 17.5 Hz, 1H), 3.20 (d, J = 17.6 Hz, 1H), 2.98 (ddd, J = 14.9, 7.0, 2.8 Hz, 1H), 2.70 (ddd, J = 14.9, 6.9, 2.7 Hz, 1H), 1.17 (t, J = 7.1 Hz, 3H); 13C NMR (151 MHz, Chloroform-d) δ 206.3, 199.9, 170.1, 167.5 (d, J = 257.6 Hz), 156.2 (d, J = 9.9 Hz), 133.9, 131.7, 128.6, 127.1, 127.0 (d, J = 10.8 Hz), 126.8, 116.2 (d, J = 23.9 Hz), 113.2 (d, J = 22.7 Hz), 95.7, 89.9, 61.9, 60.6, 36.2, 34.0, 14.0; 19F NMR (377 MHz, Chloroform-d) δ -101.5 (t, J = 9.4 Hz). HRMS (ESI) m/z Calcd. for C22H19FNaO3 ([M + Na]+) 373.1210, found 373.1202. Enantiomeric excess was determined to be 96% (determined by HPLC using chiral AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 26.1 min, tminor = 24.0 min).
Ethyl 5-chloro-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3i)
Prepared according to the procedure within 48 h as light yellow liquid (72.5 mg, 99% yield, dr = 11:1). [ α ] D 18 = 69.762 (c 0.51, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.70 (d, J = 8.1 Hz, 1H), 7.51–7.41 (m, 1H), 7.35 (dd, J = 10.2, 2.0 Hz, 2H), 7.28 (d, J = 1.8 Hz, 1H), 7.22 (d, J = 7.2 Hz, 1H), 7.20–7.15 (m, 2H), 6.07 (dt, J = 6.0, 2.9 Hz, 1H), 5.49 (q, J = 6.6 Hz, 1H), 4.11 (q, J = 7.1 Hz, 2H), 3.71 (d, J = 17.5 Hz, 1H), 3.22 (d, J = 17.5 Hz, 1H), 2.99 (ddd, J = 14.9, 7.0, 2.8 Hz, 1H), 2.75 (ddd, J = 14.9, 6.8, 2.8 Hz, 1H), 1.19 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.3, 200.4, 170.1, 154.6, 141.9, 133.9, 133.8, 128.6, 128.6, 127.2, 126.7, 126.7, 125.7, 95.9, 89.8, 61.9, 60.5, 36.1, 33.9, 14.0. HRMS (ESI) m/z Calcd. for C22H19ClNaO3 ([M + Na]+) 389.0915, found 389.0908. Enantiomeric excess was determined to be 95% (determined by HPLC using chiral AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 25.3 min, tminor = 22.8 min).
Ethyl 5-bromo-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3j)
Prepared according to the procedure within 48 h as light yellow liquid (81.2 mg, 99% yield, dr = 12:1). [ α ] D 16 = 48.041 (c 0.69, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.59 (d, J = 8.2 Hz, 1H), 7.54–7.46 (m, 2H), 7.30–7.24 (m, 2H), 7.22–7.12 (m, 3H), 6.03 (dt, J = 6.1, 2.8 Hz, 1H), 5.46 (q, J = 6.7 Hz, 1H), 4.08 (q, J = 7.1 Hz, 2H), 3.68 (d, J = 17.5 Hz, 1H), 3.19 (d, J = 17.5 Hz, 1H), 2.96 (ddd, J = 14.9, 6.9, 2.8 Hz, 1H), 2.73 (ddd, J = 14.9, 6.8, 2.9 Hz, 1H), 1.16 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.2, 200.7, 170.0, 154.7, 134.3, 133.8, 131.4, 130.8, 129.8, 128.6, 127.2, 126.7, 125.7, 95.9, 89.8, 62.0, 60.4, 36.0, 33.9, 14.0. HRMS (ESI) m/z Calcd. for C22H19BrNaO3 ([M + Na]+) 433.0410, found 433.0403. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 27.2min, tminor = 24.8 min).
Ethyl 4-chloro-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3k)
Prepared according to the procedure within 60 h as light yellow liquid (72.5 mg, 99% yield, dr = 9:1). [ α ] D 15 = 227.30 (c 0.67, CH2Cl2); 1H NMR (600 MHz, chloroform-d) δ 7.65 (d, J = 7.6 Hz, 1H), 7.52 (d, J = 7.7 Hz, 1H), 7.33–7.30 (m, 1H), 7.29–7.23 (m, 2H), 7.20–7.13 (m, 3H), 6.06 (dt, J = 6.0, 2.7 Hz, 1H), 5.48 (q, J = 6.8 Hz, 1H), 4.10 (qd, J = 7.2, 2.4 Hz, 2H), 3.69 (d, J = 17.8 Hz, 1H), 3.23 (d, J = 17.8 Hz, 1H), 3.00 (ddd, J = 14.8, 6.7, 2.8 Hz, 1H), 2.70 (ddd, J = 14.9, 7.2, 2.7 Hz, 1H), 1.17 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.4, 201.1, 170.0, 150.7, 137.3, 134.8, 133.8, 132.8, 129.3, 128.5, 127.1, 126.8, 122.8, 95.9, 89.8, 62.0, 60.2, 35.6, 34.1, 14.0. HRMS (ESI) m/z Calcd. for C22H19ClNaO3 ([M + Na]+) 389.0915, found 389.0908. Enantiomeric excess was determined to be 95% (determined by HPLC using chiral AD-OD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 47.7 min, tminor = 51.0 min).
Ethyl 4-bromo-1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3l)
Prepared according to the procedure within 72 h as light yellow liquid (81.2 mg, 99% yield, dr = 8:1). [ α ] D 16 = 76.255 (c 0.74, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.69 (d, J = 7.7 Hz, 2H), 7.29–7.22 (m, 3H), 7.21–7.14 (m, 3H), 6.06 (dt, J = 6.1, 2.8 Hz, 1H), 5.48 (q, J = 6.8 Hz, 1H), 4.11 (qd, J = 7.1, 1.5 Hz, 2H), 3.64 (d, J = 17.9 Hz, 1H), 3.18 (d, J = 17.9 Hz, 1H), 3.00 (ddd, J = 14.8, 6.7, 2.9 Hz, 1H), 2.69 (ddd, J = 14.8, 7.3, 2.7 Hz, 1H), 1.18 (t, J = 7.2 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.4, 201.2, 170.0, 152.8, 138.0, 137.3, 133.8, 129.5, 128.6, 127.1, 126.8, 123.5, 122.0, 95.9, 89.8, 62.0, 60.3, 37.6, 34.2, 14.0. HRMS (ESI) m/z Calcd. for C22H19BrNaO3 ([M + Na]+) 433.0410, found 433.0404. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral AD-OD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 48.8 min, tminor = 52.0 min).
Ethyl 1-oxo-2-(4-phenylbuta-2,3-dien-1-yl)-1,2,3,4-tetrahydronaphthalene-2-carboxylate (3m)
Prepared according to the procedure within 48 h as light yellow liquid (68.5 mg, 99% yield, dr = 2:1). [ α ] D 16 = 9.667 (c 0.30, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 8.05 (dd, J = 7.9, 1.5 Hz, 1H), 7.51–7.38 (m, 1H), 7.35–7.21 (m, 5H), 7.21–7.17 (m, 2H), 6.12 (dt, J = 6.5, 2.3 Hz, 1H), 5.60 (q, J = 7.4 Hz, 1H), 4.16 (qt, J = 7.1, 1.4 Hz, 2H), 3.11–2.86 (m, 2H), 2.79 (ddd, J = 8.1, 5.5, 2.4 Hz, 2H), 2.62 (dt, J = 13.9, 5.5 Hz, 1H), 2.34 (ddd, J = 14.1, 9.5, 5.0 Hz, 1H), 1.17 (t, J = 7.1Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 207.0, 195.0, 171.5, 143.2, 134.3, 133.6, 131.9, 128.8, 128.6, 128.4, 128.1, 127.0, 126.8, 94.7, 90.2, 61.5, 57.5, 33.8, 30.4, 25.7, 14.1. HRMS (ESI) m/z Calcd. for C23H22NaO3 ([M + Na]+) 369.1461, found 369.1454. Enantiomeric excess was determined to be 97%/75% (determined by HPLC using chiral OJ-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 76.9/68.2 min, tminor = 57.7/82.0 min).
Ethyl 2-(4-(2-chlorophenyl) buta-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3n)
Prepared according to the procedure within 72 h as light yellow liquid (58.6 mg, 80% yield, dr = 10:1). [ α ] D 15 = 300.81 (c 0.25, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.78–7.76 (m, 1H), 7.61–7.57 (m, 1H), 7.46–7.33 (m, 3H), 7.29 (dd, J = 8.0, 1.4 Hz, 1H), 7.21–7.16 (m, 1H), 7.13–7.08 (m, 1H), 6.54 (dt, J = 6.5, 2.7 Hz, 1H), 5.50 (q, J = 7.0 Hz, 1H), 4.09 (qd, J = 7.1, 1.4 Hz, 2H), 3.73 (d, J = 17.4 Hz, 1H), 3.22 (d, J = 17.3 Hz, 1H), 2.99 (ddd, J = 14.7, 7.2, 2.8 Hz, 1H), 2.73 (ddd, J = 14.6, 7.1, 2.8 Hz, 1H), 1.16 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 207.3, 201.8, 170.4, 153.2, 135.4, 135.3, 132.1, 131.8, 129.7, 128.3, 128.0, 127.8, 126.7, 126.4, 124.8, 91.8, 90.2, 61.8, 60.2, 36.5, 33.9, 14.0. HRMS (ESI) m/z Calcd. for C22H19ClNaO3 ([M + Na]+) 389.0915, found 389.0912. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral OJ-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 48.5 min, tminor = 38.6 min).
Ethyl 2-(4-(3-methoxyphenyl) buta-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3o)
Prepared according to the procedure within 48 h as light yellow liquid (71.6 mg, 99% yield, dr = 9:1). [ α ] D 21 = 64.833 (c 0.51, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.76 (d, J = 7.7 Hz, 1H), 7.59–7.55 (m, 1H), 7.45–7.33 (m, 2H), 7.20–7.16 (m, 1H), 6.86–6.68 (m, 3H), 6.05 (dt, J = 6.1, 2.8 Hz, 1H), 5.48 (q, J = 6.8 Hz, 1H), 4.09 (qd, J = 7.2, 1.1 Hz, 2H), 3.81 (s, 3H), 3.74 (d, J = 17.3 Hz, 1H), 3.24 (d, J = 17.4 Hz, 1H), 3.00 (ddd, J = 14.7, 6.9, 2.9 Hz, 1H), 2.67 (ddd, J = 14.7, 7.2, 2.6 Hz, 1H), 1.16 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.5, 201.9, 170.4, 159.9, 153.2, 135.5, 135.4, 135.2, 129.5, 127.8, 126.5, 124.8, 119.5, 113.0, 111.8, 95.5, 90.2, 61.8, 60.3, 55.2, 36.5, 34.2, 14.0. HRMS (ESI) m/z Calcd. for C23H22NaO4 ([M + Na]+) 385.1410, found 385.1407. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral IB-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 30.8 min, tminor = 26.8 min).
Ethyl 1-oxo-2-(4-(p-tolyl) buta-2,3-dien-1-yl)-2,3-dihydro-1H-indene-2-carboxylate (3p)
Prepared according to the procedure within 72 h as light yellow liquid (68.5 mg, 99% yield, dr = 13:1). [ α ] D 13 = 64.270 (c 0.45, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.77 (d, J = 7.6 Hz, 1H), 7.65–7.52 (m, 1H), 7.48–7.33 (m, 3H), 7.09 (s, 3H), 6.05 (dt, J = 6.4, 2.7 Hz, 1H), 5.45 (q, J = 6.8 Hz, 1H), 4.10 (q, J = 7.1 Hz, 2H), 3.74 (d, J = 17.4 Hz, 1H), 3.24 (d, J = 17.4 Hz, 1H), 3.00 (ddd, J = 14.6, 7.0, 2.8 Hz, 1H), 2.66 (ddd, J = 14.5, 7.2, 2.6 Hz, 1H), 2.32 (s, 3H), 1.17 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.3, 201.9, 170.4, 153.3, 136.8, 135.4, 135.2, 131.0, 129.3, 127.7, 126.7, 126.5, 124.8, 95.3, 89.9, 61.8, 60.4, 36.5, 34.3, 21.2, 14.0. HRMS (ESI) m/z Calcd. for C23H22NaO3 ([M + Na]+) 369.1461, found 369.1455. Enantiomeric excess was determined to be 93% (determined by HPLC using chiral AD-OD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 59.1 min, tminor = 54.5 min).
Ethyl 2-(4-(4-cyanophenyl) buta-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3q)
Prepared according to the procedure within 72 h as light yellow liquid (57.9 mg, 81% yield, dr = 12:1). [ α ] D 15 = 88.378 (c 0.37, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.69 (d, J = 7.7 Hz, 1H), 7.51 (dd, J = 7.5, 1.2 Hz, 1H), 7.46 (d, J = 8.1 Hz, 2H), 7.39–7.28 (m, 2H), 7.23–7.14 (m, 2H), 6.01 (dt, J = 6.0, 2.8 Hz, 1H), 5.50 (q, J = 6.9 Hz, 1H), 4.01 (q, J = 7.1 Hz, 2H), 3.65 (d, J = 17.3 Hz, 1H), 3.11 (d, J = 17.3 Hz, 1H), 2.90 (ddd, J = 14.8, 7.0, 2.9 Hz, 1H), 2.69 (ddd, J = 14.8, 7.3, 2.7 Hz, 1H), 1.08 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 207.6, 201.6, 170.3, 153.0, 139.3, 135.5, 135.3, 132.3, 127.9, 127.2, 126.4, 124.8, 119.0, 110.2, 94.8, 91.0, 61.9, 60.0, 36.5, 33.6, 14.0. HRMS (ESI) m/z Calcd. for C23H19NNaO3 ([M + Na]+) 380.1257, found 380.1248. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral OD-H column, hexane/2-propanol = 95/5, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 41.0 min, tminor = 36.8 min).
Ethyl 2-(4-(4-fluorophenyl) buta-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3r)
Prepared according to the procedure within 48 h as light yellow liquid (60.9 mg, 87% yield, dr = 10:1). [ α ] D 17 = 79.750 (c 0.40, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.76 (d, J = 7.7 Hz, 1H), 7.60–7.56 (m, 1H), 7.44–7.35 (m, 2H), 7.17–7.10 (m, 2H), 7.01–6.93 (m, 2H), 6.04 (dt, J = 6.3, 2.8 Hz, 1H), 5.47 (q, J = 6.8 Hz, 1H), 4.09 (q, J = 7.1 Hz, 2H), 3.73 (d, J = 17.4 Hz, 1H), 3.21 (d, J = 17.3 Hz, 1H), 2.98 (ddd, J = 14.8, 6.9, 2.8 Hz, 1H), 2.70 (ddd, J = 14.7, 7.1, 2.7 Hz, 1H), 1.16 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.2, 201.8, 170.4, 162.0 (d, J = 247.5 Hz), 153.2, 135.3 (d, J = 6.1 Hz), 130.0, 128.3, 128.2, 127.8, 126.4, 124.7, 115.5 (d, J = 21.2 Hz), 94.6, 90.3, 61.8, 60.2, 36.5, 34.1, 14.0. 19F NMR (376 MHz, Chloroform-d) δ -102.3 (q, J = 6.5 Hz); HRMS (ESI) m/z Calcd. for C22H19FNaO3 ([M + Na]+) 373.1210, found 373.1203. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral AD-OD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 64.5 min, tminor = 61.0 min).
Ethyl 2-(4-(4-chlorophenyl) buta-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3s)
Prepared according to the procedure within 48 h as light yellow liquid (61.5 mg, 84% yield, dr = 11:1). [ α ] D 17 = 72.922 (c 0.42, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.76 (d, J = 7.7 Hz, 1H), 7.60–7.56 (m, 1H), 7.46–7.34 (m, 2H), 7.30–7.19 (m, 2H), 7.15–7.06 (m, 2H), 6.03 (dt, J = 6.2, 2.8 Hz, 1H), 5.49 (q, J = 6.9 Hz, 1H), 4.09 (q, J = 7.1 Hz, 2H), 3.73 (d, J = 17.4 Hz, 1H), 3.20 (d, J = 17.3 Hz, 1H), 2.98 (ddd, J = 14.7, 6.9, 2.9 Hz, 1H), 2.70 (ddd, J = 14.7, 7.2, 2.7 Hz, 1H), 1.16 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, Chloroform-d) δ 206.5, 201.8, 170.4, 153.1, 135.4, 135.3, 132.6, 132.6, 128.7, 128.0, 127.8, 126.4, 124.7, 94.7, 90.5, 61.8, 60.2, 36.5, 34.0, 14.0. HRMS (ESI) m/z Calcd. for C22H19ClNaO3 ([M + Na]+) 389.0915, found 389.0907. Enantiomeric excess was determined to be 91% (determined by HPLC using chiral IC-IB-H column, hexane/2-propanol = 9/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 41.9 min, tminor = 37.0 min).
Ethyl 2-(4-(naphthalen-2-yl) buta-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3t)
Prepared according to the procedure within 72 h as light yellow liquid (55.8 mg, 73% yield, dr = 9:1). [ α ] D 16 = 94.309 (c 0.25, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.84–7.70 (m, 4H), 7.54 (d, J = 7.9 Hz, 2H), 7.44 (d, J = 6.0 Hz, 2H), 7.40–7.35 (m, 3H), 6.24 (dd, J = 6.7, 3.1 Hz, 1H), 5.55 (d, J = 6.9 Hz, 1H), 4.10 (q, J = 7.4 Hz, 2H), 3.76 (d, J = 17.4 Hz, 1H), 3.27 (d, J = 17.3 Hz, 1H), 3.04 (dd, J = 15.3, 6.7 Hz, 1H), 2.73 (dd, J = 15.3, 7.6 Hz, 1H), 1.16 (t, J = 7.1 Hz, 3H); 13C NMR (101 MHz, chloroform-d) δ 207.0, 201.8, 170.4, 153.2, 135.4, 135.3, 133.7, 132.7, 131.6, 128.5, 128.2, 127.7, 126.4, 126.2, 125.7, 125.7, 125.5, 124.7, 124.7, 95.9, 90.3, 61.8, 60.4, 36.5, 34.2, 14.0. HRMS (ESI) m/z Calcd. for C26H22NaO3 ([M + Na]+) 405.1461, found 405.1455. Enantiomeric excess was determined to be 97% (determined by HPLC using chiral AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.8 mL/min, tmajor = 49.1 min, tminor = 45.2 min).
Ethyl 2-(hepta-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3u)
Prepared according to the procedure within 84 h as light yellow liquid (50.74 mg, 85% yield, dr = 2:1). [ α ] D 16 = 49.275 (c 0.14, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.76 (d, J = 7.7 Hz, 1H), 7.65–7.57 (m, 1H), 7.48 (d, J = 7.7 Hz, 1H), 7.40–7.37 (m, 1H), 5.14–4.83 (m, 2H), 4.24–4.07 (m, 2H), 3.68 (dd, J = 17.3, 2.6 Hz, 1H), 3.23 (d, J = 17.3 Hz, 1H), 2.83 (dtd, J = 15.1, 7.6, 2.5 Hz, 1H), 2.54 (dddd, J = 14.5, 7.4, 5.2, 2.6 Hz, 1H), 1.87 (qd, J = 7.1, 3.6 Hz, 2H), 1.36 (qd, J = 7.4, 2.1 Hz, 2H), 1.21 (t, J = 7.1 Hz, 3H), 0.88 (t, J = 7.4, 3H); 13C NMR (101 MHz, Chloroform-d) δ 205.7, 202.1, 170.6, 153.4, 135.4, 135.3, 127.6, 126.4, 124.7, 91.4, 85.6, 61.6, 60.6, 36.2, 34.6, 30.9, 22.3, 14.1, 13.6. HRMS (ESI) m/z Calcd. for C19H22NaO3 ([M + Na]+) 321.1461, found 321.1453. Enantiomeric excess was determined to be 79%/71% (determined by HPLC using chiral AS-AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 30.5/26.7 min, tminor = 25.6/28.2 min).
Ethyl 2-(5-methylhexa-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3v)
Prepared according to the procedure within 84 h as light yellow liquid (49.5 mg, 83% yield, dr = 2:1). [ α ] D 16 = 67.677 (c 0.20, CH2Cl2); 1H NMR (600 MHz, chloroform-d) δ 7.76 (dd, J = 7.7, 3.9 Hz, 1H), 7.65–7.59 (m, 1H), 7.48 (dd, J = 7.8, 2.9 Hz, 1H), 7.44–7.35 (m, 1H), 5.15–4.89 (m, 2H), 4.32–4.08 (m, 2H), 3.68 (d, J = 17.2 Hz, 1H), 3.24 (dd, J = 17.3, 3.7 Hz, 1H), 2.94–2.77 (m, 1H), 2.55 (ddt, J = 14.2, 7.6, 2.0 Hz, 1H), 2.20 (dp, J = 9.7, 3.1 Hz, 1H), 1.20 (t, J = 7.1 Hz, 3H), 0.94 (ddd, J = 6.5, 4.6, 1.6 Hz, 6H); 13C NMR (151 MHz, Chloroform-d) δ 204.2, 202.2, 170.7, 153.4, 135.3, 129.0, 127.7, 126.4, 124.8, 98.9, 86.8, 61.7, 60.6, 36.2, 34.8, 27.9, 22.4, 14.1. HRMS (ESI) m/z Calcd. for C19H22NaO3 ([M + Na]+) 321.1461, found 321.1453. Enantiomeric excess was determined to be 77%/55% (determined by HPLC using chiral AS-AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 27.5/24.7 min, tminor = 23.8/26.2 min).
Ethyl 2-(4-cyclohexylbuta-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3w)
Prepared according to the procedure within 84 h as light yellow liquid (52.8 mg, 78% yield, dr = 2:1). [ α ] D 16 = 43.017 (c 0.18, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.80–7.72 (m, 1H), 7.64–7.6 (m, 1H), 7.48 (d, J = 7.7 Hz, 1H), 7.44–7.34 (m, 1H), 4.99 (dtt, J = 23.8, 6.8, 3.1 Hz, 2H), 4.15 (qd, J = 7.1, 1.5 Hz, 2H), 3.68 (d, J = 17.4 Hz, 1H), 3.24 (d, J = 17.3 Hz, 1H), 2.83 (dddd, J = 14.1, 6.9, 4.1, 2.6 Hz, 1H), 2.56 (dddd, J = 14.9, 7.9, 5.5, 2.7 Hz, 1H), 1.86 (dddd, J = 11.2, 8.7, 5.9, 3.1 Hz, 1H), 1.66 (tt, J = 18.8, 7.4 Hz, 5H), 1.39–1.10 (m, 6H), 1.00 (tdd, J = 13.6, 7.7, 3.7 Hz, 2H); 13C NMR (101 MHz, Chloroform-d) δ 204.5, 202.2, 170.7, 153.4, 135.4, 129.0, 128.4, 127.7, 126.4, 124.8, 97.5, 86.4, 61.7, 60.6, 37.2, 36.2, 34.8, 32.9, 26.1, 14.1. HRMS (ESI) m/z Calcd. for C22H26NaO3 ([M + Na]+) 361.1774, found 361.1765. Enantiomeric excess was determined to be 73%/55% (determined by HPLC using chiral AS-AD-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 29.9/26.9 min, tminor = 26.3/27.9 min).
Ethyl4-bromo-2-(4-(4-chlorophenyl) buta-2,3-dien-1-yl)-1-oxo-2,3-dihydro-1H-indene-2-carboxylate (3x)
Prepared according to the procedure within 48 h as light yellow liquid (87.9 mg, 99% yield, dr = 10:1). [ α ] D 16 = 101.589 (c 1.20, CH2Cl2); 1H NMR (600 MHz, chloroform-d) δ 7.72–7.69 (m, 2H), 7.26–7.21 (m, 3H), 7.07 (d, J = 8.3 Hz, 2H), 6.01 (dt, J = 6.0, 2.8 Hz, 1H), 5.50 (q, J = 6.8 Hz, 1H), 4.11 (qd, J = 7.1, 3.2 Hz, 2H), 3.62 (d, J = 17.8 Hz, 1H), 3.14 (d, J = 17.8 Hz, 1H), 2.98 (ddd, J = 14.9, 6.7, 2.9 Hz, 1H), 2.72 (ddd, J = 14.9, 7.1, 2.8 Hz, 1H), 1.18 (t, J = 7.1 Hz, 3H). 13C NMR (101 MHz, Chloroform-d) δ 206.4, 201.2, 170.0, 152.7, 138.0, 137.3, 132.7, 132.4, 129.6, 128.7, 127.9, 123.4, 122.0, 95.1, 90.2, 62.0, 60.2, 37.6, 33.9, 14.0. HRMS (ESI) m/z Calcd. for C22H18BrClNaO3 ([M + Na]+) 467.0020, found 467.0022. Enantiomeric excess was determined to be 96% (determined by HPLC using chiral AD-OJ-H column, hexane/2-propanol = 50/1, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 92.3 min, tminor = 83.9 min).

3.3. Procedure for the Synthesis of Compounds 4a

A Schlenk flask under a nitrogen atmosphere was charged with compound 3x (222 mg, 0.50 mmol, 1.0 eq), phenylboronic acid (73.2 mg, 0.60 mmol, 1.2 eq), Cs2CO3 (244 mg, 0.75 mmol, 1.5 eq), and Pd(PPh3)4 (28.9 mg, 25.0 µmol, 5 mol%). THF (5 mL) was added and the mixture was heated to 80 °C for 18 h, when compound 3x was consumed as checked by TLC, the mixture was cooled to rt and diluted with Et2O (15 mL). The mixture was washed with water (15 mL). The aq. layer was extracted with Et2O (2 × 25 mL) and the combined org. layers were dried, filtered, and concentrated. The crude product was purified by column chromatography (petroleum ether/ethyl acetate = 20:1) yielding the title compound 4a as a slightly yellow oil.
Ethyl2-(4-(4-chlorophenyl)buta-2,3-dien-1-yl)-1-oxo-4-phenyl-2,3-dihydro-1H-indene-2-carboxylate (4a)
Prepared according to the procedure within 18 h as slightly yellow oil (154 mg, 70% yield, dr = 9:1). [ α ] D 16 = 23.014 (c 0.37, CH2Cl2); 1H NMR (400 MHz, chloroform-d) δ 7.79 (dd, J = 18.7, 7.5 Hz, 1H), 7.63 (dd, J = 16.9, 7.2 Hz, 1H), 7.55–7.42 (m, 5H), 7.38 (dd, J = 6.6, 3.1 Hz, 1H), 7.32–7.27 (m, 1H), 7.16 (dd, J = 8.6, 2.1 Hz, 2H), 7.07–7.03 (m, 1H), 6.10–5.91 (m, 1H), 5.61–5.48 (m, 1H), 4.30–4.12 (m, 2H), 3.81 (dd, J = 17.5, 11.5 Hz, 1H), 3.22 (dd, J = 17.5, 9.8 Hz, 1H), 3.01 (tdd, J = 11.7, 7.0, 3.0 Hz, 1H), 2.70 (dddd, J = 30.5, 14.6, 7.2, 2.7 Hz, 1H), 1.23 (t, J = 7.1 Hz, 3H). 13C NMR (101 MHz, Chloroform-d) δ 206.5, 201.8, 170.4, 150.6, 140.2, 138.7, 135.6, 132.6, 128.8, 128.7, 128.4, 128.4, 128.4, 127.9, 127.8, 123.8, 115.3, 94.5, 90.5, 61.9, 60.5, 36.5, 34.2, 14.1. HRMS (ESI) m/z Calcd. for C28H23ClNaO3 ([M + Na]+) 465.1228, found 465.1225. Enantiomeric excess was determined to be 92% (determined by HPLC using chiral IF-OD-H column, hexane/2-propanol = 95/5, λ = 254 nm, 30 °C, 0.6 mL/min, tmajor = 41.3 min, tminor = 44.3 min).

4. Conclusions

In conclusion, we have developed Pd(0)-catalyzed asymmetric allenylic alkylation of indanone-derived β-ketoesters by allenyl carbonates. This reaction provides a contribution toward the construction of 1,3-stereocenters bearing allenyl axial and central chirality with high levels of stereocontrol. This work features a broad substrate scope, mild reaction conditions, high efficiency, excellent enantioselectivities, and good diastereoselectivities.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28072927/s1, materials and methods, experimental procedures [33,43,45], characterization data, 1H, 13C, and 19F NMR spectra, HRMS spectrometry data, and HPLC chromatogram.

Author Contributions

A.X. performed the experiments, acquired and analyzed the original data, and wrote the preliminary manuscript. B.W. and J.Q. conceived and designed the experiments, revised all figures and schemes, analyzed the data, and reviewed and edited the manuscript. X.W. and Y.H. proofread and analyzed the data. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Fundamental Research Funds for the Central Universities (No. DUT21LAB134).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Liu, H.; Leow, D.; Huang, K.-W.; Tan, C.-H. Enantioselective Synthesis of Chiral Allenoates by Guanidine-Catalyzed Isomerization of 3-Alkynoates. J. Am. Chem. Soc. 2009, 131, 7212–7213. [Google Scholar] [CrossRef]
  2. Zhang, W.; Zheng, S.; Liu, N.; Werness, J.B.; Guzei, I.A.; Tang, W. Enantioselective Bromolactonization of Conjugated (Z)-Enynes. J. Am. Chem. Soc. 2010, 132, 3664–3665. [Google Scholar] [CrossRef]
  3. Yang, J.; Wang, Z.; He, Z.; Li, G.; Hong, L.; Sun, W.; Wang, R. Organocatalytic Enantioselective Synthesis of Tetrasubstituted α-Amino Allenoates by Dearomative γ-Addition of 2,3-Disubstituted Indoles to β,γ-Alkynyl-α-imino Esters. Angew. Chem. Int. Ed. 2020, 59, 642–647. [Google Scholar] [CrossRef]
  4. Tap, A.; Blond, A.; Wakchaure, V.N.; List, B. Chiral Allenes via Alkynylogous Mukaiyama Aldol Reaction. Angew. Chem. Int. Ed. 2016, 55, 8962–8965. [Google Scholar] [CrossRef]
  5. Zheng, Y.; Miao, B.; Qin, A.; Xiao, J.; Liu, Q.; Li, G.; Zhang, L.; Zhang, F.; Guo, Y.; Ma, S. Negishi Coupling for Highly Selective Syntheses of Allenes via Ligand Effect and Mechanistic Study via SAESI-MS/MS. Chin. J. Chem. 2019, 37, 1003–1008. [Google Scholar] [CrossRef]
  6. Neff, R.K.; Frantz, D.E. Recent Advances in the Catalytic Syntheses of Allenes: A Critical Assessment. ACS Catal. 2014, 4, 519–528. [Google Scholar] [CrossRef]
  7. Ye, J.; Ma, S. Conquering three-carbon axial chirality of allenes. Org. Chem. Front. 2014, 1, 1210–1224. [Google Scholar] [CrossRef]
  8. Chu, W.-D.; Zhang, Y.; Wang, J. Recent advances in catalytic asymmetric synthesis of allenes. Catal. Sci. Technol. 2017, 7, 4570–4579. [Google Scholar] [CrossRef]
  9. Huang, X.; Ma, S. Allenation of Terminal Alkynes with Aldehydes and Ketones. Acc. Chem. Res. 2019, 52, 1301–1312. [Google Scholar] [CrossRef]
  10. Davis, C.R.; Luvaga, I.K.; Ready, J.M. Enantioselective Allylation of Alkenyl Boronates Promotes a 1,2-Metalate Rearrangement with 1,3-Diastereocontrol. J. Am. Chem. Soc. 2021, 143, 4921–4927. [Google Scholar] [CrossRef]
  11. Trost, B.M.; Schultz, J.E.; Chang, T.; Maduabum, M.R. Chemo-, Regio-, Diastereo-, and Enantioselective Palladium Allylic Alkylation of 1,3-Dioxaboroles as Synthetic Equivalents of α-Hydroxyketones. J. Am. Chem. Soc. 2019, 141, 9521–9526. [Google Scholar] [CrossRef]
  12. Trost, B.M.; Zell, D.; Hohn, C.; Mata, G.; Maruniak, A. Enantio-and Diastereoselective Synthesis of Chiral Allenes by Palladium-Catalyzed Asymmetric [3 + 2] Cycloaddition Reactions. Angew. Chem. Int. Ed. 2018, 130, 13098–13102. [Google Scholar] [CrossRef]
  13. Li, Z.; Hu, B.; Wu, Y.; Fei, C.; Deng, L. Control of chemoselectivity in asymmetric tandem reactions: Direct synthesis of chiral amines bearing nonadjacent stereocenters. Proc. Natl. Acad. Sci. USA 2018, 115, 1730–1735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Zhu, B.; Lee, R.; Li, J.; Ye, X.; Hong, S.N.; Qiu, S.; Coote, M.L.; Jiang, Z. Chemoselective Switch in the Asymmetric Organocatalysis of 5H-Oxazol-4-ones and N-Itaconimides: Addition–Protonation or [4 + 2] Cycloaddition. Angew. Chem. Int. Ed. 2016, 128, 1321–1325. [Google Scholar] [CrossRef]
  15. Suzuki, M.; Kurosawa, E. Okamurallene, a novel halogenated C15 metabolite from the red alga laurencia okamuhai yamada. Tetrahedron Lett. 1981, 22, 3853–3856. [Google Scholar] [CrossRef]
  16. Zhu, Y.L.; Pai, S.B.; Liu, S.H.; Grove, K.L.; Jones, B.C.; Simons, C.; Zemlicka, J.; Cheng, Y.C. Inhibition of replication of hepatitis B virus by cytallene in vitro. Antimicrob. Agents Chemother. 1997, 41, 1755–1760. [Google Scholar] [CrossRef] [Green Version]
  17. Cai, F.; Pu, X.; Qi, X.; Lynch, V.; Radha, A.; Ready, J.M. Chiral Allene-Containing Phosphines in Asymmetric Catalysis. J. Am. Chem. Soc. 2011, 133, 18066–18069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Pu, X.; Qi, X.; Ready, J.M. Allenes in Asymmetric Catalysis: Asymmetric Ring Opening of meso-Epoxides Catalyzed by Allene-Containing Phosphine Oxides. J. Am. Chem. Soc. 2009, 131, 10364–10365. [Google Scholar] [CrossRef] [Green Version]
  19. Hasegawa, M.; Sone, Y.; Iwata, S.; Matsuzawa, H.; Mazaki, Y. Tetrathiafulvalenylallene: A New Class of Donor Molecules Having Strong Chiroptical Properties in Neutral and Doped States. Org. Lett. 2011, 13, 4688–4691. [Google Scholar] [CrossRef] [PubMed]
  20. Ching, C.K.; Lam, S.K. A comparison of two prostaglandin analogues (enprostil vs misoprostol) in the treatment of acute duodenal ulcer disease. J. Gastroenterol. 1995, 30, 607–614. [Google Scholar] [CrossRef]
  21. Watanabe, Y.; Yamazaki, T. Facile Preparation of CF3-Containing 1-Bromoallenes. Synlett 2009, 20, 3352–3354. [Google Scholar]
  22. Lü, B.; Jiang, X.; Fu, C.; Ma, S. Highly Regio- and Stereoselective Cyclic Iodoetherification of 4,5-Alkadienols. An Efficient Preparation of 2-(1′(Z)-Iodoalkenyl)tetrahydrofurans. J. Org. Chem. 2009, 74, 438–441. [Google Scholar] [CrossRef] [PubMed]
  23. Yokota, M.; Fuchibe, K.; Ueda, M.; Mayumi, Y.; Ichikawa, J. Facile Synthesis of 1,1-Difluoroallenes via the Difluorovinylidenation of Aldehydes and Ketones. Org. Lett. 2009, 11, 3994–3997. [Google Scholar] [CrossRef] [PubMed]
  24. Li, C.-Y.; Sun, X.-L.; Jing, Q.; Tang, Y. Enantioselective synthesis of allenic esters via an ylide route. Chem. Commun. 2006, 28, 2980–2982. [Google Scholar] [CrossRef]
  25. Skattebøfl, L. Chemistry of gem-dihalocyclopropanes—VI: A novel synthesis of cyclopentadienes and fulvenes. Tetrahedron 1967, 23, 1107–1117. [Google Scholar] [CrossRef]
  26. Woerly, E.M.; Cherney, A.H.; Davis, E.K.; Burke, M.D. Stereoretentive Suzuki−Miyaura Coupling of Haloallenes Enables Fully Stereocontrolled Access to (−)-Peridinin. J. Am. Chem. Soc. 2010, 132, 6941–6943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. He, G.; Xue, C.; Fu, C.; Ma, S. An Efficient Synthesis of Allenyl Perfluoroalkyl Ketones from Mono-1, 2-Addition-Elimination Reaction of Allenoates with RfMgX. Synlett 2010, 2, 281–285. [Google Scholar] [CrossRef]
  28. Deska, J.; del Pozo Ochoa, C.; Bäckvall, J.-E. Chemoenzymatic Dynamic Kinetic Resolution of Axially Chiral Allenes. Chem. Eur. J. 2010, 16, 4447–4451. [Google Scholar] [CrossRef]
  29. Qian, H.; Yu, X.; Zhang, J.; Sun, J. Organocatalytic Enantioselective Synthesis of 2,3-Allenoates by Intermolecular Addition of Nitroalkanes to Activated Enynes. J. Am. Chem. Soc. 2013, 135, 18020–18023. [Google Scholar] [CrossRef]
  30. Poulsen, P.H.; Li, Y.; Lauridsen, V.H.; Jørgensen, D.K.B.; Palazzo, T.A.; Meazza, M.; Jørgensen, K.A. Organocatalytic Formation of Chiral Trisubstituted Allenes and Chiral Furan Derivatives. Angew. Chem. Int. Ed. 2018, 57, 10661–10665. [Google Scholar] [CrossRef] [PubMed]
  31. Wang, H.; Luo, H.; Zhang, Z.-M.; Zheng, W.-F.; Yin, Y.; Qian, H.; Zhang, J.; Ma, S. Pd-Catalyzed Enantioselective Syntheses of Trisubstituted Allenes via Coupling of Propargylic Benzoates with Organoboronic Acids. J. Am. Chem. Soc. 2020, 142, 9763–9771. [Google Scholar] [CrossRef] [PubMed]
  32. Fu, L.; Greßies, S.; Chen, P.; Liu, G. Recent Advances and Perspectives in Transition Metal-Catalyzed 1,4-Functionalizations of Unactivated 1,3-Enynes for the Synthesis of Allenes. Chin. J. Chem. 2020, 38, 91–100. [Google Scholar] [CrossRef]
  33. Song, S.; Zhou, J.; Fu, C.; Ma, S. Catalytic enantioselective construction of axial chirality in 1,3-disubstituted allenes. Nat. Commun. 2019, 10, 507. [Google Scholar] [CrossRef] [Green Version]
  34. Liu, H.-C.; Hu, Y.-Z.; Wang, Z.-F.; Tao, H.-Y.; Wang, C.-J. Synergistic Cu/Pd-Catalyzed Asymmetric Allenylic Alkylation of Azomethine Ylides for the Construction of α-Allene-Substituted Nonproteinogenic α-Amino Acids. Chem. Eur. J. 2019, 25, 8681–8685. [Google Scholar] [CrossRef]
  35. Xiao, J.; Xu, H.; Huo, X.; Zhang, W.; Ma, S. One Stone Two Birds—Enantioselective Bimetallic Catalysis for α-Amino Acid Derivatives with an Allene Unit. Chin. J. Chem. 2021, 39, 1958–1964. [Google Scholar] [CrossRef]
  36. Wan, B.; Ma, S. Enantioselective decarboxylative amination: Synthesis of axially chiral allenyl amines. Angew. Chem. Int. Ed. 2013, 125, 459–463. [Google Scholar] [CrossRef]
  37. Zhu, C.; Chu, H.; Li, G.; Ma, S.; Zhang, J. Pd-Catalyzed Enantioselective Heck Reaction of Aryl Triflates and Alkynes. J. Am. Chem. Soc. 2019, 141, 19246–19251. [Google Scholar] [CrossRef]
  38. Chu, W.-D.; Zhang, L.; Zhang, Z.; Zhou, Q.; Mo, F.; Zhang, Y.; Wang, J. Enantioselective Synthesis of Trisubstituted Allenes via Cu(I)-Catalyzed Coupling of Diazoalkanes with Terminal Alkynes. J. Am. Chem. Soc. 2016, 138, 14558–14561. [Google Scholar] [CrossRef]
  39. Zhang, J.; Huo, X.; Xiao, J.; Zhao, L.; Ma, S.; Zhang, W. Enantio- and Diastereodivergent Construction of 1,3-Nonadjacent Stereocenters Bearing Axial and Central Chirality through Synergistic Pd/Cu Catalysis. J. Am. Chem. Soc. 2021, 143, 12622–12632. [Google Scholar] [CrossRef] [PubMed]
  40. Zou, L.; Wang, B.; Mu, H.; Zhang, H.; Song, Y.; Qu, J. Development of Tartaric Acid Derived Chiral Guanidines and Their Application to Catalytic Enantioselective α-Hydroxylation of β-Dicarbonyl Compounds. Org. Lett. 2013, 15, 3106–3109. [Google Scholar] [CrossRef] [PubMed]
  41. Zou, L.; Bao, X.; Zhang, H.; Song, Y.; Qu, J.; Wang, B. Novel Tartrate-Based Guanidines for Enantioselective Fluorination of 1,3-Dicarbonyl and α-Cyano Carbonyl Compounds. Aust. J. Chem. 2014, 67, 1115–1118. [Google Scholar] [CrossRef]
  42. Witzig, R.M.; Fäseke, V.C.; Häussinger, D.; Sparr, C. Atroposelective synthesis of tetra-ortho-substituted biaryls by catalyst-controlled non-canonical polyketide cyclizations. Nat. Catal. 2019, 2, 925–930. [Google Scholar] [CrossRef]
  43. Song, S.; Ma, S. Highly Selective Nucleophilic 4-Aryl-2,3-allenylation of Malonates†. Chin. J. Chem. 2020, 38, 1233–1238. [Google Scholar] [CrossRef]
  44. Zhang, Y.; Zhang, X.; Ma, S. Stretchable chiral pockets for palladium-catalyzed highly chemo- and enantioselective allenylation. Nat. Commun. 2021, 12, 2416. [Google Scholar] [CrossRef]
  45. Fernández, M. Studies on the Chemistry of 2-[3-(2-Nitrophenyl)-2-oxopropyl] benzaldehydes: Novel Syntheses of 5H-Benzo [b] carbazole-6, 11-diones and Indolo [1, 2-b] isoquinoline-6, 11-diones. Synthesis 2009, 2009, 3051–3060. [Google Scholar] [CrossRef]
Scheme 1. Representative allenes with different functions, different methods for the synthesis of allenes and construction of 1,3-stereocenters bearing allenyl axial and central chirality.
Scheme 1. Representative allenes with different functions, different methods for the synthesis of allenes and construction of 1,3-stereocenters bearing allenyl axial and central chirality.
Molecules 28 02927 sch001
Scheme 2. Substrate scope for the reactions of β-ketoesters 1 with arylallenyl carbonate 2a. The reaction was carried out on a 0.2 mmol scale with Pd2dba3 (2.5 mol%) and L9 (5 mol%) in 4.0 mL THF with NaHCO3 (0.2 mmol); the ratio of 1/2a is 1.0/1.2; isolated yields are given; the dr was determined by 1H NMR of crude product; the ee was determined by chiral HPLC.
Scheme 2. Substrate scope for the reactions of β-ketoesters 1 with arylallenyl carbonate 2a. The reaction was carried out on a 0.2 mmol scale with Pd2dba3 (2.5 mol%) and L9 (5 mol%) in 4.0 mL THF with NaHCO3 (0.2 mmol); the ratio of 1/2a is 1.0/1.2; isolated yields are given; the dr was determined by 1H NMR of crude product; the ee was determined by chiral HPLC.
Molecules 28 02927 sch002
Scheme 3. Substrate scope for the reactions of β-ketoester 1a with allenylic carbonates 2. The reaction was carried out on a 0.2 mmol scale with Pd2dba3 (2.5 mol%) and L9 (5 mol%) in 4.0 mL THF with NaHCO3 (0.2 mmol); the ratio of 1a/2 is 1.0/1.2; isolated yields are given; the dr was determined by 1H NMR of crude product; the ee was determined by chiral HPLC.
Scheme 3. Substrate scope for the reactions of β-ketoester 1a with allenylic carbonates 2. The reaction was carried out on a 0.2 mmol scale with Pd2dba3 (2.5 mol%) and L9 (5 mol%) in 4.0 mL THF with NaHCO3 (0.2 mmol); the ratio of 1a/2 is 1.0/1.2; isolated yields are given; the dr was determined by 1H NMR of crude product; the ee was determined by chiral HPLC.
Molecules 28 02927 sch003
Scheme 4. Gram-scale reaction and product derivatization.
Scheme 4. Gram-scale reaction and product derivatization.
Molecules 28 02927 sch004
Scheme 5. Plausible mechanism of the palladium-catalyzed alkylation of β-ketoester 1.
Scheme 5. Plausible mechanism of the palladium-catalyzed alkylation of β-ketoester 1.
Molecules 28 02927 sch005
Table 1. Optimization of reaction conditions.
Table 1. Optimization of reaction conditions.
Molecules 28 02927 i001
Molecules 28 02927 i002
Entry aLigandBaseSol.T [°C]t [h]Yield [%] bdr cee [%] d
1L1Cs2CO3DCM250.58410:1−71/−39
2L2Cs2CO3DCM2512trace--
3L3Cs2CO3DCM253893:1−67/−69
4L4Cs2CO3DCM25125515:1−73/−17
5L5Cs2CO3DCM2548214:1−23/−33
6L6Cs2CO3DCM25242011:1−7/−5
7L7Cs2CO3DCM257.5595:1−69/−69
8L8Cs2CO3DCM2524295:1−89/−51
9L9Cs2CO3DCM2524904:191/81
10 eL9Cs2CO3DCM256974:193/77
11 eL9Cs2CO3CHCl3259925:192/61
12 eL9Cs2CO3DCE2512924:191/75
13 eL9Cs2CO3MeCN2512965:192/81
14 eL9Cs2CO3Tol.259985:192/73
15 eL9Cs2CO3THF256976:192/65
16 eL9Cs2CO3Dioxane256984:193/61
17 eL9Et3NTHF2510706:194/71
18 eL9C4H9OKTHF254907:192/71
19 eL9C2H5ONaTHF256983:190/73
20 eL9NaHCO3THF256987:193/67
21 eL9Na2CO3THF2510615:192/51
22 eL9K2CO3THF256982:193/79
23 eL9NaHCO3THF024986:195/79
24 eL9NaHCO3THF−1042987:195/70
25 e,fL9NaHCO3THF−1012987:196/71
26 e,gL9NaHCO3THF−1048988:196/77
a The reaction was conducted with 1a (0.1 mmol), 2a (0.11 mmol), base (0.1 mmol), Pd2dba3 (2.5 mol%) and ligand (5 mol%) in solvent (1.0 mL). b Isolated yield. c Detected by 1H NMR of the crude product. d Detected by chiral HPLC analysis. e 2a (0.12 mmol). f THF (0.5 mL) was used. g THF (2.0 mL) was used.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xue, A.; Wei, X.; Huang, Y.; Qu, J.; Wang, B. Palladium-Catalyzed Stereoselective Construction of 1,3-Stereocenters Displaying Axial and Central Chirality via Asymmetric Alkylations. Molecules 2023, 28, 2927. https://doi.org/10.3390/molecules28072927

AMA Style

Xue A, Wei X, Huang Y, Qu J, Wang B. Palladium-Catalyzed Stereoselective Construction of 1,3-Stereocenters Displaying Axial and Central Chirality via Asymmetric Alkylations. Molecules. 2023; 28(7):2927. https://doi.org/10.3390/molecules28072927

Chicago/Turabian Style

Xue, Aiqi, Xingfu Wei, Yue Huang, Jingping Qu, and Baomin Wang. 2023. "Palladium-Catalyzed Stereoselective Construction of 1,3-Stereocenters Displaying Axial and Central Chirality via Asymmetric Alkylations" Molecules 28, no. 7: 2927. https://doi.org/10.3390/molecules28072927

Article Metrics

Back to TopTop