Next Article in Journal
Characterizing the Ee Jahn–Teller Potential Energy Surfaces by Differential Geometry Tools
Previous Article in Journal
Statistical Damage Model of Rock Structural Plane Considering Void Compaction and Failure Modes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microwave-Assisted Synthesis of 2-Methyl-1H-indole-3-carboxylate Derivatives via Pd-Catalyzed Heterocyclization

1
Department of Pharmacy, University of Naples “Federico II”, Via D. Montesano 49, 80131 Naples, Italy
2
KelAda Pharmachem Ltd. A1.01, Science Centre South, Belfield, D04 Dublin, Ireland
3
Centre for Synthesis and Chemical Biology, Department of Chemistry, The Royal College of Surgeons in Ireland, 123 St. Stephen′s Green, D02 YN77 Dublin, Ireland
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Symmetry 2022, 14(3), 435; https://doi.org/10.3390/sym14030435
Submission received: 31 January 2022 / Revised: 13 February 2022 / Accepted: 19 February 2022 / Published: 22 February 2022
(This article belongs to the Section Chemistry: Symmetry/Asymmetry)

Abstract

:
Indole moiety is well-known as a superlative framework in many natural products and synthetic pharmaceuticals. Herein, we report an efficient procedure to synthesize a series of functionalized 2-methyl-1H-indole-3-carboxylate derivatives from commercially available anilines properly functionalized by different electron-withdrawing and -donating groups through a palladium-catalyzed intramolecular oxidative coupling. The conversion of a variety of enamines into the relevant indole was optimized by exposing the neat mixture of reactants to microwave irradiation, obtaining the desired products in excellent yields and high regioselectivity. The synthesized compounds were confirmed by 1H and 13C spectroscopic means as well as by high-resolution mass spectrometry.

1. Introduction

The indole nucleus is a very widespread motif used in drug discovery and found in many pharmacologically active compounds [1,2]. Bioactive indole-based molecules have been isolated from plants, bacteria, fungi, and marine products, such as tryptamine and serotonin derivatives [3,4], bufotenine [3], ergot and vinca alkaloids [5,6]. In addition, a large pool of drugs containing the indole ring was approved by the Food and Drug Administration (FDA) as antiviral, anticancer, antimalarial, and antitubercular agents [7,8]. Some of these molecules are mainly featured by axially chiral indole-based units [9,10], which are typical building blocks present in natural alkaloids, chiral phosphine ligands, and catalysts [11,12].
The construction of the traditional indole core may occur through conventional strategies, such as those described by Fischer [13], Julia and Bartoli [14,15], or through organometallic catalyzed cross-coupling C–N/C–C bond formations [16,17], which are highly versatile and suitable for indole derivatization. The catalytic asymmetric approaches can be also used to access chiral indole-based compounds [18]. These substituted indole derivatives can be considered key intermediates for the synthesis of molecules of medicinal chemistry interest (Figure 1). For instance, 3-nitroindoles are essential structural moiety for the development and synthesis of novel antidiabetic agents [19], whereas halogenated indoles represent a key structural moiety of human 15-lipoxygenase-1 inhibitors [20], and indole-3-carboxylic acids, along with their related esters, are key moieties for mast cell tryptase inhibitors [21]. As part of chiral-indole-based skeletons, there are chiral tryptamines that constitute bioactive molecules acting in the central nervous system [22], while many axially chiral indoles have proved potent anticancer activity [23].
An efficient route to obtain indole-3-carboxylate derivatives from several substituted anilines has been developed by Wurtz et al. by performing palladium-catalyzed intramolecular oxidation [24]. This strategy involves the use of four players: (i) a copper source, (ii) a ligand, (iii) a base, and (iv) an N-aryl enamine, to efficiently get a series of 2-methyl-1H-indole-3-carboxylate derivatives [24]. Starting from this study and considering the numerous advantages the microwave (μW)-assisted synthesis has over the traditional organic synthesis, which includes reduction of reaction time, improved conversions, and cleaner product formation [25,26,27,28], we have explored the palladium-catalyzed oxidative cyclization reaction for the preparation of several indole-3-carboxylate derivatives via μW-heating technology. Previously, μW-heating has been exploited for the preparation of indole analogs in the classical reaction of Fischer [28], and other transition metal-mediated cyclizations, leading to final products yields >80% and excellent purity (>90%) [29]. In this light, we have synthesized N-aryl enamine carboxylates starting from commercially available anilines bearing different electron-withdrawing (-NO2, -Cl, -Br) and donating groups (-CH3, -OPh) (Scheme 1), then converted into their corresponding indoles under both μW-assisted and conventional heating conditions by palladium-catalyzed oxidation. By varying both the amount of oxidant agent (copper source) and the type of solvent, 2-methyl-1H-indole-3-carboxylate derivatives have been obtained with improved yields (>90%) and strikingly reduced reaction time with respect to conventional conditions.

2. Materials and Methods

2.1. Materials and General Procedures

All solvents and commercial reagents were purchased from Sigma-Aldrich (Saint Louis, MO, USA). TLC sheets (silica gel 60 F254 with plates 5 × 20, 0.25 mm) were purchased from Merck (Kenilworth, NJ, USA). High-resolution MS analysis (positive mode) was performed on a Thermo LTQ Orbitrap XL mass spectrometer (Thermo-Fisher, San Josè, CA, USA) through the infusion of compounds 1128 into the ESI source using MeOH as solvent. 1H (700, 600, and 400 MHz) and 13C (175 and 125 MHz). NMR spectra were recorded on a Bruker Avance Neo spectrometer equipped with an RT-DR-BF/1H-5 mm-OZ SmartProbe (Bruker BioSpin Corporation, Billerica, MA, USA); chemical shifts (δH and δC) were referenced to the residual CHCl3 signal (δH = 7.26 and δC = 77.0). High-performance liquid chromatography (HPLC) analyses were performed on a Knauer K-501 instrument endowed with a Knauer K-2301 RI detector purchased from LabService Analytica s.r.l., (Anzola dell′Emilia, Italy). Microwave-assisted reactions were performed on an Initiator+ microwave apparatus purchased from Biotage (Bergamo, Italy). The instrument was set on the high-absorption level operating at a frequency of 2.45 GHz with a continuous irradiation power (0–300 W) and a standard absorbance level of 300 W.

2.2. Synthesis of Enamine Derived Compounds 1119

Each substituted aniline (19, 2.00 mL, 22.0 mmol) was stirred together with methyl acetoacetate (CH3COCH2COOCH3, 10, 2.37 mL, 22.0 mmol, 1 eq) and 10% acetic acid (CH3CO2H, 250 μL) into an oven-dried round bottom flask, at room temperature (rt). After 12 h, the crude material was washed with water and extracted with ethyl acetate (EtOAc). The organic phase was washed with water (×3), brine (×3), dried over Na2SO4, filtered, and evaporated under reduced pressure. All enamine intermediates (1119) were obtained with high yield, ranging from 9396%, and stored at 46 °C. Compounds 1119 were confirmed by 1H and 13C spectroscopic as well as by HRMS (ESI) analyses.
Methyl-(Z)-3-(phenylimino)butanoate) (11): white crystals (4.03 g, 96%); 1H NMR (CDCl3): 10.4 (-NH, brs); 7.32 (2H, t, J = 15 Hz); 7.15 (1H, t, J = 15Hz); 7.10 (2H, d, J = 8 Hz); 4.73 (1H, s); 3.69 (3H, s); 1.99 (3H, s). 13C NMR (CDCl3): 170.6, 159.0, 139.3, 129.1, 125.0, 124.5, 85.6, 50.3, 20.4. HRMS (ESI): m/z 192.1018 [M+H]+ (calcd. for C11H14NO2: 192.1019) (Figures S1–S3). δH and δC agree with literature data [30].
Methyl (Z)-3-(4-bromophenyl)amino)but-2-enoate) (12): dark yellow crystals (5.58 g, 94%); 1H NMR (CDCl3): 10.3 (-NH, brs); 7.41 (2H, d, J = 8.3 Hz); 6.94 (2H, d, J = 8.3 Hz); 4.72 (1H, s); 3.67 (3H, s); 1.97 (3H, s). 13C NMR (CDCl3): 170.8, 158.4, 138.4, 132.2, 125.9, 118.2, 86.7, 50.4, 20.2. HRMS (ESI): m/z 270.0131 [M+H]+ (calcd. for C11H13BrNO2: 270.0124) (Figures S4–S6). δH and δC agree with literature data [31].
Methyl (Z)-3-((3-bromophenyl)amino)but-2-enoate) (13): dark brown oil (5.56 g, 94%); 1H NMR (CDCl3): 10.3 (-NH, brs); 7.21 (1H); 7.14 (1H); 6.95 (1H, dd, J = 17 Hz, 2.6 Hz), 6.79 (1H), 4.73 (1H, s); 3.66 (3H, s); 1.98 (3H, s). 13C NMR (CDCl3): 170.6, 158.2, 140.8, 130.5, 130.3, 127.7, 122.6, 113.6, 87.2, 50.5, 20.4. HRMS (ESI): m/z 270.0130 [M+H]+ (calcd. for C11H13BrNO2: 270.0124) (Figures S7–S9). δH and δC agree with literature data [31].
Methyl (Z)-3-(naphthalen-1-ylamino)but-2-enoate) (14): dark violet powder (4.93 g, 93%); 1H NMR (CDCl3): 10.5 (-NH, brs); 7.83–7.79 (1H, m); 7.55–7.51 (2H, m); 7.32–7.27 (3H, m); 6.80 (1H, dd, J = 7.2 Hz, 1Hz); 4.82 (1H, s); 3.75 (3H, s); 1.86 (3H, s). 13C NMR (CDCl3): 171.1, 160.7, 135.3, 134.3, 130.5, 128.3, 126.8, 126.7, 126.5, 125.3, 123.7, 122.7, 85.1, 50.4, 20.1 (Figures S10 and S11). δH and δC agree with literature data [30].
Methyl (Z)-3-(4-chlorophenyl)amino)but-2-enoate) (15): orange powder (4.72 g, 95%); 1H NMR (CDCl3): 10.2 (-NH, brs); 7.16 (2H, d, J = 8.0 Hz); 6.91 (2H, d, J = 8.0 Hz); 4.58 (1H, s); 3.58 (3H, s); 2.12 (3H, s). 13C NMR (CDCl3): 170.5, 158.8, 138.1, 129.1, 125.4, 116.5, 89.2, 50.4, 20.0 (Figures S12 and S13). δH and δC agree with literature data [31].
Methyl (Z)-3-((2,4-dichlorophenyl)amino)but-2-enoate) (16): grey crystals (5.35 g, 94%); 1H NMR (CDCl3): 10.3 (-NH, brs); 7.24 (1H, s); 7.02 (1H, d, J = 8Hz); 6.65 (1H, d, J = 8 Hz); 4.81 (1H, s); 3.71 (3H, s); 1.95 (3H, s). 13C NMR (CDCl3): 170.6, 157.9, 135.5, 130.7, 129.8, 128.8, 127.4, 116.6, 88.2, 50.7, 20.3. HRMS (ESI): m/z 260.0241 [M+H]+ (calcd. for C11H11Cl2NO2: 260.0215) (Figures S14–S16).
Methyl (Z)-3-((4-nitrophenyl)amino)but-2-enoate) (17): yellow crystals (4.83 g, 93%); 1H NMR (CDCl3): 10.9 (-NH, brs); 8.19 (2H, d, J = 8.3 Hz); 7.13 (2H, d, J = 8.3 Hz); 4.89 (1H, s); 3.72 (3H, s); 2.21 (3H, s). 13C NMR (CDCl3): 170.6, 156.1, 145.9, 125.5, 120.8, 113.4, 91.1, 50.8, 21.0. HRMS (ESI): m/z 237.0870 [M+H]+ (calcd. for C11H13N2O4: 237.0869) (Figures S17–S19). δH and δC agree with literature data [30].
Methyl (Z)-3-((2,4-dimethylphenyl)amino)but-2-enoate) (18): red-brick oil (4.62 g, 96%); 1H NMR (CDCl3): 10.0 (-NH, brs); 7.02–6.91 (3H, overlapped); 4.70 (1H, s); 3.66 (3H, s); 2.30 (3H, s); 2.23 (3H, s); 1.81 (3H, s). 13C NMR (CDCl3): 170.8, 160.2, 135.8, 135.3, 133.8, 131.4, 127.0, 126.6, 84.4, 50.0, 20.8, 19.9, 17.8. HRMS (ESI): m/z 242.1142 [M+Na]+ (calcd. For C13H17NO2Na: 242.1152) (Figures S20–S22). δH and δC agree with literature data [31].
Methyl (Z)-3-((4-phenoxyphenyl)amino)but-2-enoate) (19): orange oil (5.91 g, 95%); 1H NMR (CDCl3): 10.1 (-NH, brs); 7.20 (2H, m); 6.97–6.91 (3H, overlapped); 6.87–6.82 (4H, overlapped); 4.54 (1H, s); 3.53 (3H, s); 1.82 (3H, s). 13C NMR (CDCl3): 170.6, 159.4, 156.9, 154.7, 134.4, 129.7, 126.3, 123.3, 119.1, 118.7, 89.5, 50.9, 20.9. HRMS (ESI): m/z 284.1275 [M+H]+ (calcd. for C17H18NO3: 284.1281) (Figures S23–S25).

2.3. μW-Assisted Palladium-Catalyzed Oxidative Cyclization Yielding Indole Derivatives 2028

Enamine intermediates (11–19) (30.5 mg, 0.16 mmol, 1 eq), catalytic amount of palladium acetate [Pd(OAc)2 (3.6 mg, 0.016 mmol, 0.1 eq], copper acetate [Cu(OAc)2, 29.1 mg, 0.16 mmol, 1 eq], and potassium carbonate (K2CO3, 55.3 mg, 0.40 mmol, 2.5 eq) were dissolved in DMF (2 mL), and placed in a μW reaction vessel, under argon atmosphere (Table 1 and Table 2). The resulting mixture was stirred at 60 °C under μW irradiation. As reported in Table 2, when the synthesis of the indole moiety did not succeed in DMF, this was replaced by ACN (2 mL) or DMSO (2 mL). Then, the reaction was cooled down to rt, diluted with DCM, and filtered through a short pad of celite, which was then washed with DCM. The solvent was removed in vacuo and indole derivatives (20–28) were isolated by flash chromatography or semi-preparative HPLC using an isocratic mobile phase of H2O (0.1% TFA) and solvent B (0.1% TFA) (95:5, v/v), Luna column (10 μm, C18), and a flow rate of 3 mL/min. The conversion of substituted anilines (2028, Table 1 and Table 2) was also performed by conventional oil bath heating (T = 80 °C), as the only change with the respect to the procedure described above.
Methyl 2-methyl-1H-indole-3-carboxylate (20). The crude material was purified by flash chromatography (n-hexane/EtOAc, 7:3) to give the product as orangewish powder; 1H NMR (CDCl3): 8.48 (-NH, brs); 8.09 (1H, dd, J = 7.5 Hz, 2 Hz); 7.30 (1H, dd, J = 7.5Hz, 2 Hz); 7.21 (1H, qd, J = 14.5, 7.5, 1.8 Hz); 7.19 (1H, qd, J = 14.5, 7.5, 1.8 Hz); 3.94 (3H, s); 2.74 (3H, s). 13C NMR (CDCl3): 166.7, 144.3, 134.6, 127.2, 122.3, 121.6, 121.2, 110.6, 104.3, 50.8, 14.2. HRMS (ESI): m/z 190.0874 [M+H]+ (calcd. for C11H12NO2: 190.0863) (Figures S26–S28). δH and δC agree with literature data [24].
Methyl 5-bromo-2-methyl-1H-indole-3-carboxylate (21). The crude material was purified by semi-preparative HPLC (solvent B, MeOH; tR = 5.2 min). The product was obtained as brown oil; 1H NMR (CDCl3): 8.41 (-NH, brs); 8.21 (1H, s); 7.28 (1H, d, J = 8 Hz); 7.17 (1H, d, J = 8 Hz); 3.93 (3H, s); 2.74 (3H, s). 13C NMR (CDCl3): 166.1, 145.0, 133.1, 128.8, 125.6, 124.1, 115.4, 112.1, 104.7, 51.2, 14.2. HRMS (ESI): m/z 289.9785 [M+Na]+ (calcd. for C11H10BrNO2Na: 289.9787) (Figures S29–S31). δH and δC agree with literature data [32].
Methyl 6-bromo-2-methyl-1H-indole-3-carboxylate (22). The crude material was purified by semi-preparative HPLC (solvent B, MeOH; tR = 6.1 min). The product was obtained as dark brown oil; 1H NMR (CDCl3): 8.34 (-NH, brs); 7.94 (1H, d, J = 9 Hz); 7.45 (1H, d, J = 1.8 Hz); 7.31 (1H, dd, J = 9 Hz, 1.8 Hz); 3.92 (3H, s); 2.73 (3H, s). 13C NMR (CDCl3): 166.1, 144.2, 135.2, 126.0, 125.0, 122.7, 115.7, 113.4, 104.9, 50.9, 14.2. HRMS (ESI): m/z 289.9782 [M+Na]+ (calcd. for C11H10BrNO2Na: 289.9787) (Figures S32–S34). δH and δC agree with literature data [32].
Methyl 2-methyl-1H-benzo[g]indole-3-carboxylate (23). The crude material was purified by semi-preparative HPLC (H2O, solvent B, MeOH, 90:10, v,v; tR = 6.0 min). The product was obtained as dark violet powder; 1H NMR (CDCl3): 8.99 (-NH, brs); 8.21 (1H, d, J = 8.5 Hz); 7.97–7.93 (2H, dd, J = 16.5 Hz, 8 Hz); 7.63 (1H, d, J = 8.5 Hz); 7.54 (1H, td, J = 14.7 Hz, 7.6 Hz, 1 Hz); 7.44 (1H, td, J = 14.7 Hz, 7.6 Hz, 1 Hz); 3.97 (3H, s); 2.85 (3H, s). 13C NMR (CDCl3): 166.5, 141.3, 130.4, 129.9, 128.9, 125.8, 124.1, 123.5, 122.3, 120.9, 120.9, 119.1, 106.3, 50.9, 14.4. HRMS (ESI): m/z 240.1018 [M+H]+ (calcd. for C15H14NO2: 240.1019) (Figures S35–S37). δH and δC agree with literature data [24].
Methyl 5-chloro-2-methyl-1H-indole-3-carboxylate (24). The crude material was purified by semi-preparative HPLC (solvent B, ACN; tR = 5.5 min). The product was obtained as dark yellow powder; 1H NMR (CDCl3): 8.43 (-NH, brs); 8.05 (1H, d, J = 2 Hz); 7.21 (1H, d, J = 8.5 Hz); 7.14 (1H, dd, J = 8.5 Hz, 2 Hz); 3.94 (3H, s); 2.74 (3H, s). 13C NMR (CDCl3): 166.1, 145.2, 132.8, 128.3, 127.7, 122.7, 121.1, 111.5, 104.6, 50.9, 14.2. HRMS (ESI): m/z 246.0291 [M+Na]+ (calcd. for C11H10ClNO2Na: 246.0292) (Figures S38–S40). δH and δC agree with literature data [24].
Methyl 5,7-dichloro-2-methyl-1H-indole-3-carboxylate (25). The crude material was purified by semi-preparative HPLC (solvent B, MeOH; tR = 7.1 min). The product was obtained as dark brown oil; 1H NMR (CDCl3): 8.53 (-NH, brs); 7.96 (1H, s); 7.19 (1H, s); 3.93 (3H, s); 2.77 (3H, s). 13C NMR (CDCl3): 165.5, 145.6, 130.4, 128.9, 127.6, 121.9, 119.8, 116.3, 105.7, 51.2, 14.2. HRMS (ESI): m/z 258.0084 [M+H]+ (calcd. for C11H10Cl2NO2: 258.0083) (Figures S41–S43).
Methyl 2-methyl-5-nitro-1H-indole-3-carboxylate (26). The crude material was purified by flash chromatography (n-hexane/EtOAc, 7:3) and the product was obtained as yellow crystals; 1H NMR (CDCl3): 8.99 (1H, d, J = 2 Hz); 8.62 (-NH, brs); 8.12 (1H, dd, J = 9 Hz, 2 Hz); 7.36 (1H, d, J = 9 Hz); 4.00 (3H, s); 2.81 (3H, s). 13C NMR (CDCl3): 165.4, 147.0, 143.5, 137.5, 126.5, 118.5, 118.3, 113.4, 110.7, 106.7, 51.3, 14.3. HRMS (ESI): m/z 257.0534 [M+Na]+ (calcd. for C11H10N2O4Na: 257.0533) (Figures S44–S46). δH and δC agree with literature data [32].
Methyl 2,5,7-trimethyl-1H-indole-3-carboxylate (27). The crude material was purified by semi-preparative HPLC (solvent B, MeOH, tR = 5.9 min). The product was obtained as red powder; 1H NMR (CDCl3): 8.21 (-NH, brs); 7.72 (1H, s); 6.84 (1H, s); 3.93 (3H, s); 2.74 (3H, s); 2.44 (6H, s). 13C NMR (CDCl3): 166.7, 143.5, 132.2, 131.3, 127.1, 124.7, 119.2, 118.7, 104.6, 50.8, 21.6, 16.5, 14.3. HRMS (ESI): m/z 218.1178 [M+H]+ (calcd. for C13H16NO2: 218.1176) (Figures S47–S49). δH and δC agree with literature data [24].
Methyl 2-methyl-5-phenoxy-1H-indole-3-carboxylate (28). The crude material was purified by semi-preparative HPLC (H2O, solvent B, MeOH, 90:10, v,v; tR = 6.8 min). The product was obtained as dark brown oil; orange oil; 1H NMR (CDCl3): 8.37 (-NH, brs); 7.78 (1H, d, J = 1.5 Hz); 7.28–7.27 (3H, overlapped); 7.02 (1H, t, J = 15 Hz); 6.96 (2H, d, J = 8.5 Hz); 6.93 (1H, d, J = 8 Hz); 3.87 (3H, s); 2.76 (3H, s). 13C NMR (CDCl3): 166.2, 159.2, 151.5, 145.1, 131.3, 129.5, 128.2, 122.2, 117.4, 115.9, 112.6, 111.4, 104.8, 50.9, 14.3. HRMS (ESI): m/z 282.1126 [M+H]+ (calcd. for C17H16NO3: 282.1125) (Figures S50–S52). δH and δC agree with literature data [33].

3. Results and Discussion

A smart approach for the construction of functionalized indoles consists in transforming substituted enamines via palladium-catalyzed cyclization by oil bath heating conditions [24]. Herein, we investigated and optimized the reaction conditions of the oxidative enamine-cyclization under μW irradiation, in order to achieve a high-yielding conversion in differently substituted indoles. Given numerous advantages offered by μW irradiation, including high yields, selectivity, and low quantities of side products, many organic reactions are well-assisted by μW radiation, such as the Heck and Suzuki reactions [34]. In our study, we initially explored the palladium-catalyzed intramolecular oxidative coupling of the no-substituted indole 20 under μW irradiation, and then the impact of substituents, such as electron-withdrawing (-NO2, -Cl, -Br) and -donating (-CH3, -OPh) groups on the transformation into indoles were investigated under both conventional heating and μW irradiation.
Firstly, aniline (1), as well as commercially available para-, meta-, and 2,4-substituted anilines and 1-naphtylamine (19), were converted into their corresponding enamine derivatives (1119) following an efficient synthetic protocol (Scheme 2). In particular, a solution of each aniline, methyl acetoacetate (10, 1 eq), and 10% CH3CO2H was prepared and stirred at rt for 12 h [24]. Then, the solvent removal in vacuo afforded crystals for each crude material, which were differently colored depending on the bearing substituents. These were washed with distilled water and extracted with EtOAc. The resulting residues were preliminarily analyzed by NMR spectroscopy (both 1H and 13C) and high-resolution mass spectrometry (HRMS (ESI)). All enamines were obtained with excellent yields (ranging from 93–96%) allowing the desired products (1119) to be directly used for the next synthetic step without further purification. Interestingly, this protocol has also proved to be efficient for the synthesis of two new enamine-derived compounds 16 and 19, starting from two commercial anilines, that is the 2,4-dichloroaniline (6) and p-phenoxyaniline (9).
Our initial studies were focused on the conversion of the methyl-(Z)-3-(phenylimino)butanoate (11) into indole product 20 under μW radiation (Table 1), which had previously been synthesized by Wurtz et al. using a mixture of catalysts including: (i) Pd(OAc)2 (10 mol%) as the catalyst, (ii) Cu(OAc)2 as the oxidant (3 eq), (iii) K2CO3 (3 eq) as the base required for the catalytic cycle to operate. DMF was selected as a solvent and the reaction was run at 80 °C under argon for 3h (yield 72%) [24].
In our experiments, we successfully demonstrated that a significantly lower amount of Cu II (3 eq→1 eq) was required to improve the quality of the methodology and proven to be environmentally superior; while Pd (II) was kept at the same loading (10% mol) because when we performed the reaction using a low amount of 5% mol, the yield was drastically reduced (19%, entry 4). When the reaction was carried out under conventional oil bath heating (80 °C) the cyclization proceeded slowly and indole 20 was obtained in good yield (76%, Table 1, entry 1) after 16 h.
However, excellent results were obtained by performing the same interconversion under μW irradiation (Table 1, entries 2, 3). The μW-assisted palladium-catalyzed cyclization of enamine 11 was carried out under sub-boiling point conditions, setting a temperature of 60 °C for DMF. In addition, the reaction was monitored at different reaction times to determine the optimum reaction time.
Results collected (Table 1) showed that the conversion of 11 required only 0.5 h providing product 20 in 93% isolated yields (Table 1, entry 2); limiting the reaction time at 0.25 h, produced a significant reduction of the yield of 20 (Table 1, entry 3) which was obtained in only 63% yield. With this set of optimized conditions in hand, we set up to determine the scope of reaction by subjecting enamines 1219, bearing a variety of different electron-donating (-CH3, -OPh) and -withdrawing groups (-NO2, -Cl, -Br), into desired indoles 2128.
The preparation of indoles 2128 was independently studied in various solvents (DMF, DMSO, and ACN) in order to investigate the effect of the media on the reaction rate. It is important to notice that irradiation after full conversion may lead to the formation of products of decomposition, hence, it was important to determine the end point of the reaction in order to obtain compounds as free as possible from products of decomposition. For this reason, the reaction time was gradually reduced from 3h to 0.5h when possible (Table 2). Of all three tested solvents, DMF, DMSO, and ACN, DMF gave the best results for most of the substituted N-aryl enamine. Additionally, to assess the improvement of the proposed protocol, each µW-assisted cyclization has been compared with the same reaction performed under conventional heating in an oil bath.
It is noteworthy that indole-3-carboxylate derivatives such as 21, 22, 24, 25 (Table 2), bearing halogens (-Br and -Cl) and compound 26 bearing a nitro group are key intermediates for the preparation of biologically active compounds [35]. In particular, the formation of brominated indoles in high yields and in short times could be greatly advantageous since these are the main scaffolds for Suzuki–Miyaura cross-coupling reactions, which are widely used for the synthesis of bioactive molecules [32]. In our work, we performed and optimized the synthesis of both 5-bromo and 6-bromo indole 3-carboxylate derivatives under μW irradiation, starting from methyl (Z)-3-((4-bromophenyl)amino)but-2-enoate (12) and methyl (Z)-3-((3-bromophenyl)amino)but-2-enoate) (13).
Entries 1 and 2 (Table 2) show that the μW irradiation provided a notable reduction of reaction time and high yields. In the case of para-Br atom substitution to the benzene moiety, the μW heating allowed to synthesize 21 in 3 h (Table 2, entry 1, 94% yield,) vs. 12 h in an oil bath (Table 2, entry 3, 89% yield). Interestingly, our microwave-assisted synthetic strategy did not promote the formation of debrominated indole as a side product, in contrast to under conventional heating as reported by Newman et al. [36]. Specifically, when the conversion was performed in an oil bath at 80 °C, a significant amount of debrominated indole was detected even after 8 min [36]. On the contrary, in our experiments, the formation of this byproduct was not observed, even after a long time of 3 h.
Attractive results regarding the regioselectivity were obtained in the case of the meta-Br atom substitution to the benzene moiety. Under conventional heating, the conversion of enamine 13 proceeded very slowly and various solvents were used in order to obtain the best protocol. When we performed the conversion of enamine 13 in DMF as a solvent, the synthesis of indole product 22 was unsatisfactory and required a long reaction time (16 h), yielding 17% of the product (Table 2, entry 8). The synthetic trend resulted similar under μW irradiation because we observed a yield of 29% for 22 after 3 h exposure which remained constant over time (Table 2, entry 5). At this stage, on the basis of the high solubility of enamine 13 in solvents such as ACN and DMSO, we decided to change the solvent to improve the cyclization outcome. Firstly, we considered that the use of ACN as solvent makes the reaction mixture more polar, increasing its ability to absorb the μW energy taking advantage of “microwave dielectric heating” phenomena such as dipolar polarization or ionic conduction mechanisms [37]. In this context, we performed the reaction under sub-boiling point conditions setting a temperature of 40 °C for ACN and 60 °C for DMSO.
Our results showed that the use of ACN was unideal since 22 was not observed in the oil bath whereas a small amount was isolated from μW synthesis (Table 2, entry 7, 8% yield). Instead, better results were obtained using DMSO (Table 2, entry 9). In particular, the yield of 22 increased to 33% stirring the reaction overnight in an oil bath; meanwhile, 22 was obtained in 74% yield after 3 h (Table 2, entry 6) under μW conditions. It is important to note that under all tested conditions, we did not observe the formation of two regioisomers, but the formation of 22 only, as demonstrated by NMR spectra and by HPLC analysis of the crude reaction mixture. Moreover, our optimized conditions allowed us to avoid, as for 21, the formation of a debrominated side product in contrast to what was previously reported [36].
This striking regioselectivity was also observed for the formation of indole 23, derived from α-naphthylamine, which contains both α and β C-H bonds available for cyclization. The use of μW-heating allowed to obtain selectively the α- regioisomer in excellent yields (95%) after just 1 h under microwave heating (Table 2, entry 11).
When para -Cl substrates were employed even faster cyclizations were observed, as a more electron-withdrawing substituent makes the palladium-catalyzed reaction faster compared to a Br in the same position. Hence, compound 24 was obtained after 1 h in DMF with an improved yield (Table 2, entry 14, 90%,) under microwave irradiation, compared to conventional heating that afforded 24 in 73% yield in 16 h (Table 2, entry 16). Similarly, the presence of two chlorine atoms in both orto and para positions also made the conversion of enamine 16 to indole 25 in a shorter time and higher yield under μW-heating (Table 2, entries 17 and 18).
Interesting results were achieved in presence of a strong electron-withdrawing substituent such as the NO2 group on the aromatic part of the enaminone. This group, indeed, represents a clear example of a rather sensitive group that can negatively influence oxidative coupling. In the first attempt, our proposed microwave-assisted synthetic strategy afforded the desired indole 26 in a very low yield (Table 2, entry 21, 18%,) in DMF, as well as stirring the reaction mixture overnight at 80 °C in an oil bath just 12% of the desired product was observed (entry 23). At this stage, our second attempt consisted in changing the solvent reaction by switching to DMSO and ACN in order to increase the product yield. When we used DMSO no traces of the indole were isolated, neither in the oil bath (entry 24) nor in the μW reactor (entry 22). In contrast, when the enaminone derivative 17 was coupled with the metal catalysts in ACN, which has a lower dielectric constant than DMSO but is comparable with DMF, better results were achieved. In particular, under conventional heating oil bath conditions, ACN did not improve the yield (Table 2, entry 27, 24%). On the contrary, it becomes interesting by μW irradiation. As reported in Table 2, the conversion of 17 into the methyl 2-methyl-5-nitro-1H-indole-3-carboxylate resulted strongly improved when the oxidative coupling was performed using μW energy (entry 25). Indeed, the yield increased up to 91% after 3 h.
We also have explored the effect of electron-donating groups such as methyl (-CH3) and phenoxy (-OPh) groups. Wurtz et al. have previously been described the synthesis of indole product 27 through the cyclization of methyl (Z)-3-((2,4-dimethylphenyl)amino)but-2-enoate (18) under conventional heating (110 °C) by adding Cu(OAc)2 (3 eq), Pd(OAc)2 (10% mol), K2CO3 (3 eq) in DMF. The indole derivative is afforded in 62% yield [24]. Our protocol, which runs with limited amounts of Cu II provided 27 in high yield (89%, entry 30) when the reaction was performed in DMF at 80 °C for 16 h. Intriguingly, when this cyclization was explored under μW heating conditions (DMF, 60 °C), indole 27 was obtained in an enhanced 94% yield (within 3 h, entry 28).
The palladium-catalyzed oxidation proceeded in high yields also in the presence of -OPh in para position as an electron-donating substituent on enamine 19. Recently, the literature reported the preparation of indole 28 recurring to iridium-based photoredox catalyst in 73% yield in DMF at 120 °C with irradiation from an 11 W lightbulb for 24 h [34]. Herein, we obtained the indole compound 28 in 95% yield within 1h under μW irradiation (entry 32), in comparison to conventional heating (90% yield within 3 h).

4. Conclusions

In conclusion, we have developed a microwave-assisted methodology that provided functionalized 2-methyl-1H-indole-3-carboxylate derivatives (i) using significantly reduced amounts of Cu II oxidant, (ii) required shorted reaction times, and (iii) provided substituted indoles in high isolated yields and, in most cases, pure enough for a further reaction; (iv) run regioselectivity. The reaction was performed by exposing the N-aryl enamines properly functionalized by different electron-withdrawing and -donating groups to μW irradiations. All the synthesized indoles were obtained in higher yields and reduced reaction time under μW irradiation in comparison to conventional heating. The versatility of the protocol and the possibility of varying the solvent in an eco-friendly manner, with good conversion results, will allow this methodology to be applied to countless generations of suitably functionalized enamines. Indeed, by using the optimized synthetic protocol with a reduced amount of copper source (just 1 eq) as oxidant, the enamines 13 and 14 with α and β C-H bonds underwent a completely regioselective heterocyclization affording only one of the feasible regioisomeric indoles (22 and 23, entries 510 and 1113, Table 2). Moreover, merely by μW heating and in ACN as a solvent, the negative influence that the p-NO2 group of 17 exerts on the oxidative coupling to afford 26 (entry 25) has been overpassed increasing the yield up to 91%. Thus, this high-speed μW approach can be considered a fast and highly efficient method to synthesize many indole derivatives with a key role in the development of chiral indole-based compounds.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/sym14030435/s1, 1H and 13C NMR spectra and HRESIMS of compounds 11–28 (Figures S1–S52).

Author Contributions

Conceptualization, M.G.-H., B.G.K., M.F.A.A. and P.G.; methodology, R.B., M.C., N.G. and F.M.; validation, R.B., M.C. and N.G.; formal analysis, R.B., M.C. and N.G.; investigation, M.M., F.M. and P.G.; data curation, R.B. and M.C.; writing—original draft preparation, R.B. and M.C.; writing—review and editing, F.M. and P.G.; visualization, M.M., F.M. and P.G; supervision, M.G.-H., B.G.K., M.F.A.A. and P.G.; project administration, M.G.-H., B.G.K., M.F.A.A. and P.G.; funding acquisition, M.G.-H., M.F.A.A. and P.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant from the HORIZON 2020-MSCA RISE 2018 program for the “GreenX4Drug” project, grant number 823939.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kaushik, N.K.; Kaushik, N.; Attri, P.; Kumar, N.; Kim, C.H.; Verma, A.K.; Choi, E.H. Biomedical Importance of Indoles. Molecules 2013, 18, 6620–6662. [Google Scholar] [CrossRef] [PubMed]
  2. Chadha, N.; Silakari, O. Indoles as therapeutics of interest in medicinal chemistry: Bird’s eye view. Eur. J. Med. Chem. 2017, 134, 159–184. [Google Scholar] [CrossRef] [PubMed]
  3. Chilton, W.S.; Bigwood, J.; Jensen, R.E. Psilocin, Bufotenine and Serotonin: Historical and Biosynthetic Observations. J. Psychedelic Drugs 1979, 11, 61–69. [Google Scholar] [CrossRef] [PubMed]
  4. Jones, R.S. Tryptamine: A neuromodulator or neurotransmitter in mammalian brain? Progress. Neurobiol. 1982, 19, 117–139. [Google Scholar] [CrossRef]
  5. Negård, M.; Uhlig, S.; Kauserud, H.; Andersen, T.; Høiland, K.; Vrålstad, T. Links between genetic groups, indole alkaloid pro-files and ecology within the grass-parasitic Claviceps purpurea species complex. Toxins 2015, 7, 1431–1456. [Google Scholar] [CrossRef] [Green Version]
  6. Moudi, M.; Go, R.; Yien, C.Y.; Nazre, M. Vinca alkaloids. Int. J. Prev. Med. 2013, 4, 1231–1235. [Google Scholar]
  7. Desai, N.; Somani, H.; Trivedi, A.; Bhatt, K.; Nawale, L.; Khedkar, V.M.; Jha, P.C.; Sarkar, D. Synthesis, biological evaluation and molecular docking study of some novel indole and pyridine based 1,3,4-oxadiazole derivatives as potential antitubercular agents. Bioorg. Med. Chem. Lett. 2016, 26, 1776–1783. [Google Scholar] [CrossRef]
  8. Li, T.; Liu, S.; Tan, W.; Shi, F. Catalytic Asymmetric Construction of Axially Chiral Indole-Based Frameworks: An Emerging Area. Chem.–A Eur. J. 2020, 26, 15779–15792. [Google Scholar] [CrossRef]
  9. Luz, J.G.; Carson, M.W.; Condon, B.; Clawson, D.; Pustilnik, A.; Kohlman, D.T.; Barr, R.J.; Bean, J.S.; Dill, M.J.; Sindelar, D.K.; et al. Indole Glucocorticoid Receptor Antagonists Active in a Model of Dyslipidemia Act via a Unique Association with an Agonist Binding Site. J. Med. Chem. 2015, 58, 6607–6618. [Google Scholar] [CrossRef]
  10. Rios, I.G.; Rosas-Hernández, A.; Martin, E. Recent Advances in the Application of Chiral Phosphine Ligands in Pd-Catalysed Asymmetric Allylic Alkylation. Molecules 2011, 16, 970–1010. [Google Scholar] [CrossRef]
  11. Bringmann, G.; Tasler, S.; Endress, H.; Kraus, J.; Messer, K.; Wohlfarth, M.; Lobin, W. Murrastifoline-F: First Total Synthesis, Atropo-Enantiomer Resolution, and Stereoanalysis of an Axially Chiral N,C-Coupled Biaryl Alkaloid. J. Am. Chem. Soc. 2001, 123, 2703–2711. [Google Scholar] [CrossRef] [PubMed]
  12. Dedashpour, S.; Emami, S. Indole in the target-based design of anticancer agents: A versatile scaffold with diverse mechanisms. Eur. J. Med. Chem. 2018, 150, 9–29. [Google Scholar] [CrossRef] [PubMed]
  13. Robinson, B. The Fischer Indole Synthesis; John Wiley and Sons: New York, NY, USA, 1982; pp. 48–59. [Google Scholar]
  14. Baudin, J.-B.; Julia, S.A. Synthesis of indoles from N-aryl-1-alkenylsulphinamides. Tetrahedron Lett. 1986, 27, 837–840. [Google Scholar] [CrossRef]
  15. Bartoli, G.; Palmieri, G.; Bosco, M.; Dalpozzo, R. The reaction of vinyl grignard reagents with 2-substituted nitroarenes: A new approach to the synthesis of 7-substituted indoles. Tetrahedron Lett. 1989, 30, 2129–2132. [Google Scholar] [CrossRef]
  16. Willis, M.C.; Brace, G.N.; Holmes, I.P. Palladium-Catalyzed Tandem Alkenyl and Aryl C?N Bond Formation: A Cascade N-Annulation Route to 1-Functionalized Indoles. Angew. Chem. Int. Ed. 2005, 44, 403–406. [Google Scholar] [CrossRef]
  17. Zhang, Y.-C.; Jiang, F.; Shi, F. Organocatalytic Asymmetric Synthesis of Indole-Based Chiral Heterocycles: Strategies, Reactions, and Outreach. Accounts Chem. Res. 2019, 53, 425–446. [Google Scholar] [CrossRef]
  18. Kumari, A.; Singh, R.K. Medicinal chemistry of indole derivatives: Current to future therapeutic prospectives. Bioorg. Chem. 2019, 89, 103021. [Google Scholar] [CrossRef]
  19. Bahekar, R.H.; Jain, M.R.; Goel, A.; Patel, D.N.; Prajapati, V.M.; Gupta, A.A.; Jadav, P.A.; Patel, P.R. Design, synthesis, and biological evaluation of substituted-N-(thieno[2,3-b]pyridin-3-yl)-guanidines, N-(1H-pyrrolo[2,3-b]pyridin-3-yl)-guanidines, and N-(1H-indol-3-yl)-guanidines. Bioorg. Med. Chem. 2007, 15, 3248–3265. [Google Scholar] [CrossRef]
  20. Eleftheriadis, N.; Neochoritis, C.G.; Leus, N.G.J.; Van Der Wouden, P.E.; Dömling, A.; Dekker, F.J. Rational Development of a Potent 15-Lipoxygenase-1 Inhibitor with in Vitro and ex Vivo Anti-inflammatory Properties. J. Med. Chem. 2015, 58, 7850–7862. [Google Scholar] [CrossRef] [Green Version]
  21. Hopkins, C.R.; Czekaj, M.; Kaye, S.S.; Gao, Z.; Pribish, J.; Pauls, H.; Liang, G.; Sides, K.; Cramer, D.; Cairns, J.; et al. Design, synthesis, and biological activity of potent and selective inhibitors of mast cell tryptase. Bioorg. Med. Chem. Lett. 2005, 15, 2734–2737. [Google Scholar] [CrossRef]
  22. Wolfard, J.; Xu, J.; Zhang, H.; Chung, C.K. Synthesis of chiral tryptamines via a regioselective indole alkylation. Org. Lett. 2018, 20, 5431–5434. [Google Scholar] [CrossRef] [PubMed]
  23. Jiang, F.; Chen, K.; Wu, P.; Zhang, Y.; Jiao, Y.; Shi, F. A Strategy for Synthesizing Axially Chiral Naphthyl-Indoles: Catalytic Asymmetric Addition Reactions of Racemic Substrates. Angew. Chem. Int. Ed. 2019, 58, 15104–15110. [Google Scholar] [CrossRef] [PubMed]
  24. Würtz, S.; Rakshit, S.; Neumann, J.J.; Dröge, T.; Glorius, F. Palladium-Catalyzed Oxidative Cyclization ofN-Aryl Enamines: From Anilines to Indoles. Angew. Chem. Int. Ed. 2008, 47, 7230–7233. [Google Scholar] [CrossRef] [PubMed]
  25. Patil, S.A.; Miller, D.D. Microwave-Assisted Synthesis of Medicinally Relevant Indoles. Curr. Med. Chem. 2011, 18, 615–637. [Google Scholar] [CrossRef] [PubMed]
  26. Wilson, N.S.; Roth, G.P. Recent trends in microwave-assisted synthesis. Curr. Opin. Drug Discov. Dev. 2002, 5, 620–629. [Google Scholar] [CrossRef]
  27. Wathey, B.; Tierney, J.; Lidström, P.; Westman, J. The impact of microwave-assisted organic chemistry on drug discovery. Drug Discov. Today 2002, 7, 373–380. [Google Scholar] [CrossRef]
  28. Berrino, E.; Supuran, C.T. Advances in microwave-assisted synthesis and the impact of novel drug discovery. Expert Opin. Drug Discov. 2018, 13, 861–873. [Google Scholar] [CrossRef]
  29. Franco, L.H.; Palermo, J.A. Synthesis of 2-(Pyrimidin-4-yl)indoles. Chem. Pharm. Bull. 2003, 5, 975–977. [Google Scholar] [CrossRef] [Green Version]
  30. Wang, W.; Zhang, S.; Shi, G.; Chen, Z. Electrochemical synthesis of 1,2,4,5-tetrasubstituted imidazoles from enamines and benzylamines. Org. Biomol. Chem. 2021, 19, 6682–6686. [Google Scholar] [CrossRef]
  31. Huang, W.; Zhu, C.; Li, M.; Yu, Y.; Wu, W.; Tu, Z.; Jiang, H. TBAI or KI-Promoted Oxidative Coupling of Enamines and N -Tosylhydrazine: An Unconventional Method toward 1,5- and 1,4,5-Substituted 1,2,3-Triazoles. Adv. Synth. Catal. 2018, 360, 3117–3123. [Google Scholar] [CrossRef]
  32. Neumann, J.J.; Rakshit, S.; Dröge, T.; Würtz, S.; Glorius, F. Exploring the Oxidative Cyclization of Substituted N -Aryl Enamines: Pd-Catalyzed Formation of Indoles from Anilines. Chem.–A Eur. J. 2011, 17, 7298–7303. [Google Scholar] [CrossRef] [PubMed]
  33. Zoller, J.; Fabry, D.C.; Ronge, M.A.; Rueping, M. Synthesis of Indoles Using Visible Light: Photoredox Catalysis for Palla-dium-Catalyzed CH Activation. Angew. Chem. Int. Ed. 2014, 53, 13264–13268. [Google Scholar] [CrossRef] [PubMed]
  34. Beletskaya, I.; Cheprakov, A.V. The Heck Reaction as a Sharpening Stone of Palladium Catalysis. Chem. Rev. 2000, 100, 3009–3066. [Google Scholar] [CrossRef] [PubMed]
  35. Beccalli, E.M.; Broggini, G.; Martinelli, M.; Sottocornola, S. C-C, C-O, C-N bond formation on sp2 Carbon by Pd(II)-Catalyzed reactions involving on oxidant agents. Chem. Rev. 2007, 107, 5318–5365. [Google Scholar] [CrossRef] [PubMed]
  36. Newman, S.G.; Lautens, M. The Role of Reversible Oxidative Addition in Selective Palladium(0)-Catalyzed Intramolecular Cross-Couplings of Polyhalogenated Substrates: Synthesis of Brominated Indoles. J. Am. Chem. Soc. 2010, 132, 11416–11417. [Google Scholar] [CrossRef]
  37. Dallinger, D.; Kappe, C.O. Microwave-Assisted Synthesis in Water as Solvent. Chem. Rev. 2007, 107, 2563–2591. [Google Scholar] [CrossRef]
Figure 1. Representation of some indole-based compounds containing biologically active compounds.
Figure 1. Representation of some indole-based compounds containing biologically active compounds.
Symmetry 14 00435 g001
Scheme 1. Microwave-assisted palladium-catalyzed cyclization of N-aryl enamine substituted with electron-withdrawing (EWG) and -donating (EDG) groups that were used in our study.
Scheme 1. Microwave-assisted palladium-catalyzed cyclization of N-aryl enamine substituted with electron-withdrawing (EWG) and -donating (EDG) groups that were used in our study.
Symmetry 14 00435 sch001
Scheme 2. Synthesis of N-aryl enamine derivatives (1119) from substituted anilines (19).
Scheme 2. Synthesis of N-aryl enamine derivatives (1119) from substituted anilines (19).
Symmetry 14 00435 sch002
Table 1. Optimized preparation of 2-methyl-1H-indole-3-carboxylate (20), passing from conventional to μW-heating conditions.
Table 1. Optimized preparation of 2-methyl-1H-indole-3-carboxylate (20), passing from conventional to μW-heating conditions.
Symmetry 14 00435 i001
EntrySolventEnergy InputTime (h)Isolated Yields (%)
1DMFa1676
2DMFb0.5 93
3DMFb0.2567
4DMFc0.519
[a] Standard reaction conditions: 11 (0.16 mmol), Pd(OAc)2 (10 mol%), Cu(OAc)2 (0.16 mmol, 1 eq), K2CO3 (0.4 mmol), DMF (2 mL), 80 °C. [b] Microwave conditions: 11 (0.16 mmol), Pd(OAc)2 (10 mol%), Cu(OAc)2 (0.16 mmol), K2CO3 (0.4 mmol), DMF (2 mL), 60 °C. [c] Microwave conditions: 11 (0.16 mmol), Pd(OAc)2 (5 mol%), Cu(OAc)2 (0.16 mmol), K2CO3 (0.4 mmol), DMF (2 mL), 60 °C.
Table 2. Screening of solvent conditions for the cyclization reaction of arylenamides 12–19 to indole-based compounds 21–28 under conventional and μW conditions.
Table 2. Screening of solvent conditions for the cyclization reaction of arylenamides 12–19 to indole-based compounds 21–28 under conventional and μW conditions.
Symmetry 14 00435 i002
CompoundMicrowaveConventional Heating
EntrySolventTime (h)Isolated Yield (%)EntrySolventTime (h)Isolated Yield (%)
Symmetry 14 00435 i003
21
1DMF3943DMF1689
2DMF1554DMF6 51
Symmetry 14 00435 i004
22
5DMF3298DMF1617
6DMSO3749DMSO1633
7ACN a3810ACN b16ND
Symmetry 14 00435 i005
23
11DMF19512DMF691
13DMF678
Symmetry 14 00435 i006
24
14DMF19016DMF1673
15DMF0.582
Symmetry 14 00435 i007
25
17DMF19519DMF391
18DMF0.58220DMF123
Symmetry 14 00435 i008
26
21DMF31823DMF1612
22DMSO3ND24DMSO16ND
25ACN a39127ACN b1624
26ACN a112
Symmetry 14 00435 i009
27
28DMF39430DMF689
29DMF13831DMF358
Symmetry 14 00435 i010
28
32DMF19534DMF390
33DMF0.56235DMF143
Standard reaction conditions: enamine intermediate 12–19 (0.16 mmol), Pd(OAc)2 (10% mol), Cu(OAc)2 (0.16 mmol), K2CO3 (0.4 mmol), solvent (1.5 mL), 80 °C. Microwave reaction conditions: enamine intermediate 12–19 (0.16 mmol), Pd(OAc)2 (10% mol), Cu(OAc)2 (0.016 mmol), K2CO3 (0.4 mmol), solvent (2 mL), 60 °C. a T = 40 °C; b T = 60 °C; ND= not detected.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bellavita, R.; Casertano, M.; Grasso, N.; Gillick-Healy, M.; Kelly, B.G.; Adamo, M.F.A.; Menna, M.; Merlino, F.; Grieco, P. Microwave-Assisted Synthesis of 2-Methyl-1H-indole-3-carboxylate Derivatives via Pd-Catalyzed Heterocyclization. Symmetry 2022, 14, 435. https://doi.org/10.3390/sym14030435

AMA Style

Bellavita R, Casertano M, Grasso N, Gillick-Healy M, Kelly BG, Adamo MFA, Menna M, Merlino F, Grieco P. Microwave-Assisted Synthesis of 2-Methyl-1H-indole-3-carboxylate Derivatives via Pd-Catalyzed Heterocyclization. Symmetry. 2022; 14(3):435. https://doi.org/10.3390/sym14030435

Chicago/Turabian Style

Bellavita, Rosa, Marcello Casertano, Nicola Grasso, Malachi Gillick-Healy, Brian G. Kelly, Mauro F. A. Adamo, Marialuisa Menna, Francesco Merlino, and Paolo Grieco. 2022. "Microwave-Assisted Synthesis of 2-Methyl-1H-indole-3-carboxylate Derivatives via Pd-Catalyzed Heterocyclization" Symmetry 14, no. 3: 435. https://doi.org/10.3390/sym14030435

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop