Next Article in Journal
Rapid Quantification and Quantitation of Alkaloids in Xinjiang Fritillaria by Ultra Performance Liquid Chromatography-Quadrupole Time-of-Flight Mass Spectrometry
Next Article in Special Issue
Ortho-Nitro Effect on the Diastereoselective Control in Sulfa-Staudinger and Staudinger Cycloadditions
Previous Article in Journal
Hibiscus sabdariffa L. and Its Bioactive Constituents Exhibit Antiviral Activity against HSV-2 and Anti-enzymatic Properties against Urease by an ESI-MS Based Assay
Previous Article in Special Issue
Reactions of 5-Indolizyl Lithium Compounds with Some Bielectrophiles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rhodium(I)-Complexes Catalyzed 1,4-Conjugate Addition of Arylzinc Chlorides to N-Boc-4-pyridone

1
Department of Chemistry, Winston-Salem State University, Winston Salem, NC 27110, USA
2
School of Chemical and Environmental Engineering, Shanghai Institute of Technology, Shanghai 201418, China
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(5), 723; https://doi.org/10.3390/molecules22050723
Submission received: 15 March 2017 / Revised: 27 April 2017 / Accepted: 27 April 2017 / Published: 1 May 2017
(This article belongs to the Collection Heterocyclic Compounds)

Abstract

:
Rhodium(I)-complexes catalyzed the 1,4-conjugate addition of arylzinc chlorides to N-Boc-4-pyridone in the presence of chlorotrimethylsilane (TMSCl). A combination of [RhCl(C2H4)2]2 and BINAP was determined to be the most effective catalyst to promote the 1,4-conjugate addition reactions of arylzinc chlorides to N-Boc-4-pyridone. A broad scope of arylzinc reagents with both electron-withdrawing and electron-donating substituents on the aromatic ring successfully underwent 1,4-conjugate addition to N-Boc-4-pyridone to afford versatile 1,4-adducts 2-substituted-2,3-dihydropyridones in good to excellent yields (up to 91%) and excellent ee (up to 96%) when (S)-BINAP was used as chiral ligand.

Graphical Abstract

1. Introduction

1,4-conjugate addition of organometallic reagents to enones is one of the most reliable and widely used carbon-carbon bond formation processes to afford β-substituted carbonyl compounds, which are highly useful synthons for further organic transformations [1,2,3,4,5]. The combination of metal catalysts with an organometallic reagent has been particularly effective in promoting 1,4-addition to enones. Among the metal catalysts, copper is the most commonly used species [3]. Other metal catalysts including nickel and palladium have also been reported to be effective in 1,4-conjugate addition of organozinc, organoaluminum, organoziconium, and organomercury compounds to α,β-unsaturated enones. Organometallic reagents such as organolithiums, Grignard reagents and diorganozinc reagents have also been widely used in this regard and high yields of 1,4-adducts can be achieved in most cases. However, the competing 1,2-additions as well as 1,6-additions accompanied with these organometallic reagents have limited their applications [6,7,8,9,10,11]. Over the last 30 years, significant progresses have been made in asymmetric 1,4-conjugate additions, especially in the addition of organozinc or Grignard reagents using copper(I) catalysts in combination with chiral phosphorous ligands [12,13,14]. Although high yields and high entioselectivity can be achieved in these copper-catalyzed reactions, the substituents introduced to the β-position are limited to alkyl groups [12,13,14]. In recent years, there has been a growing interest in rhodium-catalyzed C-C bond forming reactions of organometallic reagents, due to the mild reaction conditions and toleration of various functional groups [15,16]. Since the first report by Miyaura in 1997 [17], rhodium-catalyzed 1,4-conjugate addition of organometallic reagents to unsaturated substrates has emerged as a versatile and efficient methodology for the formation of C-C bonds due to their tolerance with water and a wide range of substrates [18,19]. Rhodium-catalyzed 1,4-conjugate addition reactions utilize mild organometallic reagents such as organoboron and arylzinc reagents under mild reaction conditions, which are extremely useful in introducing versatile aromatic groups to the β-position. For the abovementioned reasons, considerable efforts have been devoted to developing the rhodium-catalyzed 1,4-conjugate addition of alkenyl(aryl)boronic acids as well as arylzinc reagents. A broad scope of substrates such as cyclic or acyclic enones and enoates have been reported [18,19]. Rhodium(I) complexes have been demonstrated to be excellent catalysts for 1,4-conjugate addition of alkenyl- and arylboronic acids to α,β-unsaturated ketones, esters, and even less reactive amides [18,19,20,21,22,23,24,25,26,27].
In the course of our investigation on access to highly versatile 2-substituted-2,3-dihydropyridones 2, we are intrigued by the possibility of rhodium(I)-complex catalyzed 1,4-addition of arylzinc cholorides or arylboronic acids to N-protected-4-pyridones 1 (Scheme 1). We are interested in N-heterocycles such as 2-substituted-2,3-dihyrdopyrdones 2 because they are an important class of medicinal compounds, and many medicinally important compounds in clinical or pre-clinical studies contain piperidine subunits [28,29,30,31]. 2-Substituted-2,3-dihyrdopyrdines are precursors for medicinally important N-heterocycles such as pyridines, piperidones, piperidines, indolizidenes, and quinolizidenes.
Due to their importance in medicinal chemistry, considerable efforts have been devoted to develop efficient synthetic strategies for the synthesis of piperidine derivatives. Most of these strategies involve using chiral auxiliaries/chiral starting materials and have been developed into effective methods for the synthesis of a wide variety of N-heterocycles [32,33,34,35,36,37,38,39,40]. The synthesis of 2,3-dihydro-4-pyridones via conjugate addition reactions have also been reported [38,39,40,41,42,43,44,45,46,47,48,49,50,51]. Great progress has been made using dialkylzinc reagents in conjunction with copper catalysts, but challenges still remain in increasing reactivity and general applicability [40,41,42,43,44,45,46]. Recent developments in copper-catalyzed asymmetric conjugate additions of Grignard reagents [47,50,51] and rhodium-catalyzed 1,4-addition of arylboroxines/boronic acids promise greater reactivity and versatility [40]. Recent advances in the NHC-Cu-catalyzed conjugate arylation of β-substututed cyclic enones to afford quaternary stereogenic centers has also been reported [52]. Despite the fact that rhodium(I)-complexes-catalyzed 1,4-conjugate addition of boronic acids is one of the most potent methods for C-C bond formation, there are no reported examples on 1,4-conjugate addition of boronic acids to N-Boc-4-pyridones or substrates that contain piperidine subunits, presumably due to the unreactive nature of these nitrogen-containing substrates with boronic acids [20,21,22,23,24,25,26,27].

2. Results and Discussions

As a starting point, we employed the standard rhodium(I)-BINAP reaction conditions (Scheme 2) [20,53]. Initially, the reaction of N-Boc-4-pyridone 3A with 2.0 equiv of p-tolylboronic acid 4a was carried out in dioxane/water (10 to 1 ratio) in the presence of 1.5 mol % of [RhCl(C2H4)2]2 and 3 mol % of ligands such as phosphoramidite ligands (A, B) [12,13,14], 1,5-diphenyl-1,5-cyclooctadiene (C) [54], and BINAP (D) [18,19]. Under these standard reactions conditions, no 1,4-adduct 5Aa was formed (Scheme 2, entries 1–2, 5, 7). No desired 1,4-adduct 5Aa was formed even after 24 h at 100 °C in dioxane/water (10 to 1 ratio) with the increased catalyst loading of 15 mol % of [RhCl(C2H4)2]2 and 30 mol % of ligands (Scheme 2, entries 3–4, 6). Only a trace amount of 1,4-adduct 5Aa was observed after extended heating with 15 mol % [RhCl(C2H4)2]2 and 30 mol % of BINAP at 100 °C for 48 h (Scheme 2, entry 8). The substituents on nitrogen were also investigated (Scheme 2, entries 9–10). With ethyl carbonate as a protecting group, a trace amount of 1,4-adduct 5Ba was observed. No 1,4-adduct 5Ca was attained with N-methylated substrate 3C (Scheme 2, entry 10).
We then turned our attention to arylzinc chlorides, which have been reported to be more reactive nucleophiles towards 1-benzyloxycarbonyl-4-quinolone and structurally similar 2,3-dihydro-4-pyridones under rhodium(I)-complexes-catalyzed reaction conditions [18,19]. There is also one isolated example on rhodium(I)-BINAP-catalyzed 1,4-conjugate addition to N-tert-butoxycarbonyl-4-pyridone [55]. Other approaches involving 2,3-dihydro-4-pyridones via direct conjugate addition of organocuprates and Grignard reagents have also been reported [56,57]. In our study, ligands such as phosphoramidite ligands (A, B) [12,13,14], 1,5-diphenyl-1,5-cyclooctadiene (C) [54], and BINAP (D) [18,19] were initially studied in rhodium(I)-catalyzed conjugate addition of p-tolylZnCl 6a to N-Boc 4-pyridone 3A at −5 °C. As shown in Scheme 3, p-tolylZnCl 6a underwent 1,4-conjugate addition to N-Boc-4-pyridone 3A catalyzed by rhodium(I)-phosphoramidite in the presence of chlorotrimethyl silane (TMSCl) with low chemical yield (Scheme 3, entry 1). Slightly higher chemical yield was observed when phosphoramidite B was used as a ligand (Scheme 3, entry 2). When 1,5-diphenyl-1,5-cyclootadiene was used with [RhCl(C2H4)2]2, a higher chemical yield 32% can be attained under similar reaction conditions (entry 4). With both ligands A and B, no significant increases in chemical yields were obtained, even with 15 mol % [RhCl(C2H4)2]2 and 30 mol % of ligands (Scheme 3, entries 3 and 5). This ligand screening showed that when 3 mol % of BINAP was used as a ligand, 1,4-adduct 5Aa can be attained in good chemical yield (89%, Scheme 3, entry 6). We also investigated the effect of substituents on nitrogen (Scheme 3, entries 8–9). With ethyl carbonate as a protecting group, the yield of 1,4-adduct 5Ba was much lower (Scheme 3, entry 8). No 1,4-adduct 5Ca was attained with N-methylated substrate 3C (Scheme 3, entry 9). Notably no desired 1,4-adduct was observed without the addition of TMSCl (Scheme 3, entry 7). It has been reported that the addition of chlorotrimethylsilane as a Lewis acid may facilitate the activation of the substrate toward 1,4-addition and also stabilize the product by forming a silyl enol ether, which is then converted to the carbonyl group under acidic work-up conditions [58].
With the optimized reaction conditions in hand, we next examined the scope of rhodium-BINAP-catalyzed arylzinc reagents conjugate addition (Scheme 4). In general, arylzinc reagents underwent smooth conjugate addition to N-Boc-4-pyridone with good to excellent chemical yields. Simple arylzinc reagents such as phenylzinc and naphthylzinc chlorides added to N-Boc-4-pyridone with excellent chemical yields (Scheme 4, entries 3–5). The arylzinc reagents with electron donating substituents usually showed higher reactivity (Scheme 4, entries 1, 2, 6–9). The arylzinc reagents with strong electron withdrawing groups such as fluoro, trifluoromethyl, and bistrifluromethyl groups also underwent conjugate addition but with lower chemical yields (entries 10–12). Compared to para- and meta-substituted arylzinc reagents, the ortho-substituted arylzinc reagents gave lower yields due to steric hindrance (entries 2, 8). We also conducted asymmetric conjugate addition of arylzinc reagents to N-Boc-4-pyridone. When (S)-BINAP was used as the chiral ligand, excellent ee can be achieved (entries 1 and 3, up to 96% ee).

3. Materials and Methods

3.1. General Procedures, Materials and Intrumentation

The 1H- and 13C-NMR spectra were recorded on a BRUKER 300 NMR spectrometer (BRUKER, Winston Salem, NC, USA), operating at 300 MHz for 1H, 75 MHz for 13C and 282 MHz for 19F. Samples for NMR spectra were dissolved in deuterated chloroform (with TMS). Infrared (IR) spectra were recorded on a Nicolet iS10 FT-IR spectrometer as neat samples (thin films). Analytical thin layer chromatography (TLC) was performed on silica gel plates, 60 µ mesh with F254 indicator. Visualization was accomplished by UV light (254 nm), and/or a 10% ethanol solution of phosphomolybdic acid and/or KMnO4 stain prepared by dissolving 1.5 g KMnO4, 10 g potassium carbonate, and 1.25 mL 10% sodium hydroxide in 200 mL water. Flash chromatography was performed with 200–400 µ silica gel.

3.1.1. Materials

Unless stated otherwise, solvents and chemicals were obtained from commercial sources and used without further purification. Anhydrous tetrahydrofuran (THF) was purchased from Sigma Aldrich. TMSCl was distilled from CaH2 under a positive nitrogen atmosphere. Arylzinc reagents were prepared from the corresponding aryllithium reagent and ZnCl2. Aryllithium reagents were prepared from the corresponding arylhalides and t-BuLi (1.70 M in pentane). t-BuLi (1.70 M in pentane) was commercially available and titrated using sec-BuOH and 1,10-phenanthroline monohydrate in THF. All glassware was flamed-dried under high vacuum and purged with argon and then cooled under a dry nitrogen atmosphere. Low temperature baths were prepared using ice NaCl water bath, or dry ice-isopropanol slush bath mixtures. All ArZnCl 1,4-conjugate addition reactions were conducted under a positive, dry argon atmosphere in anhydrous solvents in flasks fitted with rubber septa.

3.1.2. General Procedure A

Rh(I)-BINAP-catalyzed 1,4-conjugate addition reactions. This method was modified from the procedure reported by Hayashi [18]. Starting N-Boc-4-pyridone (0.5 mmol) was added to a solution of [RhCl(C2H4)2]2 (0.015 mol %, 0.0075 mmol) and BINAP (0.033 mol %, 0.0165 mmol) in dry THF (1.0 mL) at −5 °C under argon with continuous stirring. After stirring for 15 min, ArZnCl (1.0 M in THF, 3.0 quiv, 1.5 mmol, prepared from corresponding arylbromides in a separate flask [59]) and TMSCl (1.0 M in THF, 3.0 equiv, 1.5 mmol) were added simultaneously dropwise over 10 min to this solution. The resulting mixture was then warmed up to room temperature and stirred for 20 h at room temperature. Then, the reaction mixture was diluted with dichloromethane (4.0 mL), quenched with saturated aqueous NH4Cl (4.0 mL) and extracted with dichloromethane (3 × 8.0 mL). The combined organic phase was washed with water (8.0 mL), brine (8.0 mL), then dried over anhydrous Na2SO4, filtered, concentrated in vacuo, and purified by flash column chromatography (silica, 10%–20% ethyl acetate in hexanes, v/v) to give pure compounds.
HRMS data for compounds 5Aa, 5Ad, 5AfAg, 5Ai, 5AkAl were analyzed by TOF MS, see supplementary. Compounds 5AbAc, 5Ae, 5Ah, and 5Aj have been fully characterized and reported [55,57,60].

3.2. Synthesis of Adducts 5Aa, 5Ad, 5Af–Ag, 5Ai, 5Ag, 5Ak–Al

N-Boc-2-(4-methylphenyl)-2,3-dihydro-4-pyridone (5Aa). Employing General Procedure A and using N-Boc-4-pyridone (97 mg, 0.5 mmol), [RhCl(C2H4)2]2 (2.92 mg, 0.0075 mmol), (S)-BINAP (10.3 mg, 0.0165 mmol), ArZnCl (1.0 M in THF, 3.0 equiv, 1.5 mmol), TMSCl (163 mg, 1.5 mmol) and the resultant reaction mixture was slowly warmed to 0 °C and stirred for an additional 24 h at 0 °C, after purification by flash column chromatography (silica, 10%–20% ethyl acetate:hexanes, v/v) gave white solid 5Aa (128 mg, 89%). The ee was determined on a Daicel Chiralcel OD-H column with a solvent system of hexanes/2-propanol (9:1 ratio), flow rate = 1.0 mL/min. tr(major) = 7.1 min., tr(minor) = 7.8 min. 92% ee. m.p. 82.6–84.3 °C; IR (neat) 3081 (w), 2982 (w), 2926 (w), 1729 (s), 1661 (s), 1597 (s), 1412 (w), 1369 (m), 1339 (s), 1299 (s), 1253 (m), 1210 (m), 1145 (s), 1096 (m), 1010 (m), 923 (m), 845 (m), 815 (m), 774 (m) cm−1; 1H-NMR δ 1.33 (s, 9 H), 2.17 (s, 3H), 2.63 (td, J = 1.5, 16.5 Hz, 1 H), 2.98 (dd, J = 7.50, 16.5 Hz, 1 H), 5.20 (dd, J = 1.17, 8.4 Hz, 1 H), 5.49 (d, J = 7.50 Hz, 1 H), 6.96 (s, 4 H), 7.79 (d, J = 8.4 Hz, 1 H); 13C-NMR δ 21.0, 28.0, 41.9, 55.4, 83.6, 107.0, 125.8, 129.4, 135.8, 137.6, 142.9, 151.5, 192.3. HRMS (EI-ion trap) m/z: [M]+ calcd for C17H21NO3, 287.1520; found 287.1521.
N-Boc-2-(2-napthanyl)-2,3-dihydro-4-pyridone (5Ad). Employing General Procedure A and using N-Boc-4-pyridone (97 mg, 0.5 mmol), [RhCl(C2H4)2]2 (2.92 mg, 0.0075 mmol), BINAP (10.3 mg, 0.0165 mmol), ArZnCl (1.0 M in THF, 3.0 equiv, 1.5 mmol) and TMSCl (163 mg, 1.5 mmol) after purification by flash column chromatography (silica, 10%–20% ethyl acetate:hexanes, v/v) gave white solid 5Ad (147 mg, 91%): m.p. 113.8–115.1 °C; IR (neat) 3065 (w), 2981 (w), 1717 (s), 1667 (s), 1606 (s), 1451 (m), 1369 (m), 1309 (s), 1257 (m), 1222 (m), 1142 (s), 950 (m), 851 (m), 757 (s) cm−1; 1H-NMR (300 MHz, CDCl3) δ 1.40 (br s, 9H), 2.84 (d, J = 16.5 Hz, 1H), 3.16 (dd, J = 7.8, 16.5 Hz, 1H), 5.34 (d, J = 8.4 Hz, 1H), 5.76 (d, J = 7.2 Hz, 1H), 7.30 (dd, J = 1.5, 8.4 Hz, 1H), 7.35-7.43 (m, 2H), 7.68-7.76 (m, 3H), 7.97 (d, J = 8.4 Hz, 1H); 13C-NMR (75 MHz, CDCl3) δ 28.0, 41.8, 55.9, 83.8, 107.1, 123.9, 124.6, 126.2, 126.4, 127.6, 128.0, 128.9, 132.9, 133.2, 136.2, 142.9, 151.5, 192.0. HRMS (EI-ion trap) m/z: [M]+ calcd. for C20H21NO3, 323.1521; found 323.1529.
N-Boc-2-(3,5-dimethylphenyl)-2,3-dihydro-4-pyridone (5Af). Employing General Procedure A and using N-Boc-4-pyridone (97 mg, 0.5 mmol), [RhCl(C2H4)2]2 (2.92 mg, 0.0075 mmol), BINAP (10.3 mg, 0.0165 mmol), ArZnCl (1.0 M in THF, 3.0 equiv, 1.5 mmol) and TMSCl (163 mg, 1.5 mmol) after purification by flash column chromatography (silica, 10%–20% ethyl acetate:hexanes, v/v) gave white solid 5Af (129 mg, 86%): m.p. 100.0–101.7 °C; IR (neat) 3012 (w), 2978 (w), 2918 (w), 1713 (s), 1663 (s), 1596 (s), 1460 (w), 1418 (m), 1369 (m), 1337 (m), 1315 (s), 1257 (m), 1220 (m), 1142 (s), 1017 (m), 843 (s), 762 (s), 703 (w) cm−1; 1H-NMR (300 MHz, CDCl3) δ 1.32 (br s, 9H), 1.13 (s, 6H), 2.64 (td, J = 1.5, 16.5 Hz, 1H), 2.98 (dd, J = 7.8, 16.5 Hz, 1H), 5.21 (dd, J = 1.2, 8.4 Hz, 1H), 5.44 (d, J = 7.8 Hz, 1H), 6.65 (s, 2H), 6.75 (s, 1H), 7.82 (d, J = 8.4 Hz, 1H); 13C-NMR (75 MHz, CDCl3) δ 21.4, 28.0, 41.9, 55.6, 83.6, 107.0, 123.5, 128.2, 129.0, 129.5, 138.3, 138.8, 143.0, 151.5, 192.3. HRMS (EI-ion trap) m/z: [M]+ calcd for C18H23NO3, 301.1678; found 301.1683.
N-Boc-2-(4-methoxylphenyl)-2,3-dihydro-4-pyridone (5Ag). Employing General Procedure A and using N-Boc-4-pyridone (97 mg, 0.5 mmol), [RhCl(C2H4)2]2 (2.92 mg, 0.0075 mmol), BINAP (10.3 mg, 0.0165 mmol), ArZnCl (1.0 M in THF, 3.0 equiv, 1.5 mmol) and TMSCl (163 mg, 1.5 mmol) after purification by flash column chromatography (silica, 10%–20% ethyl acetate:hexanes, v/v) gave white amorphous solid 5Ag (138 mg, 91%): m.p. 70.0−71.9 °C; IR (neat) 2984 (w), 2932 (w), 2836 (w), 1719 (s), 1663 (s), 1594 (s), 1508 (s), 1508 (w), 1366 (w), 1342 (s), 1285 (s), 1246 (s), 1171 (s), 1104 (s), 1027 (s), 981 (m), 852 (w), 787 (m), 769 (s), 645 (m) cm−1; 1H-NMR (300 MHz, CDCl3) δ 1.47 (br s, 9H), 2.74 (td, J = 1.5, 16.5 Hz, 1H), 3.11 (dd, J = 7.5, 16.5 Hz, 1H), 3.77 (s, 3H), 5.34 (dd, J = 1.5, 8.4 Hz, 1H), 5.61 (d, J = 7.5 Hz, 1H), 6.81 (d, J = 8.7 Hz, 2H), 7.14 (dd, J = 8.7 Hz, 2H),7.90 (d, J = 8.4 Hz, 1H); 13C-NMR δ 28.0, 41.9, 55.1, 55.3, 83.7, 106.9, 114.1, 127.2, 131.1, 142.8, 151.5, 159.2, 192.4. HRMS (EI-ion trap) m/z: [M]+ calcd. for C17H21NO4, 303.1471; found 303.1471.
N-Boc-2-(3,5-dimethoxylphenyl)-2,3-dihydro-4-pyridone (5Ai). Employing General Procedure A and using N-Boc-4-pyridone (97 mg, 0.5 mmol), [RhCl(C2H4)2]2 (2.92 mg, 0.0075 mmol), BINAP (10.3 mg, 0.0165 mmol), ArZnCl (1.0 M in THF, 3.0 equiv, 1.5 mmol) and TMSCl (163 mg, 1.5 mmol) after purification by flash column chromatography (silica, 10%–20% ethyl acetate:hexanes, v/v) gave clear sticky oil 5Ai (141 mg, 85%): IR (neat) 2975 (w), 2838 (w), 1719 (s), 1664 (s), 1592 (s), 1457 (m), 1420 (m), 1368 (m), 1288 (s), 1203 (m), 1144 (s), 1066 (m), 999 (m), 940 (w); 833 (m), 759 (m), 697 (m) cm−1; 1H-NMR (300 MHz, CDCl3) δ 1.46 (br s, 9H), 2.76 (d, J = 16.5 Hz, 1H), 3.11 (dd, J = 7.8, 16.5 Hz, 1H), 3.74 (s, 6H), 5.33 (d, J = 8.4 Hz, 1H), 5.58 (d, J = 7.8 Hz, 1H), 6.33 (s, 3H), 7.94 (d, J = 8.4 Hz, 1H); 13C-NMR (75 MHz, CDCl3) δ 28.0, 41.9, 55.3, 55.7, 83.8, 99.3, 104.1, 107.0, 125.3, 128.2, 129.0, 141.3, 142.9, 151.4, 161.1, 192.0. HRMS (EI-ion trap) m/z: [M]+ calcd for C18H23NO5, 333.1576; found 333.1578.
N-Boc-2-(4-trifluoromethylphenyl)-2,3-dihydro-4-pyridone (5Ak). Employing General Procedure A and using N-Boc-4-pyridone (97 mg, 0.5 mmol), [RhCl(C2H4)2]2 (2.92 mg, 0.0075 mmol), BINAP (10.3 mg, 0.0165 mmol), ArZnCl (1.0 M in THF, 1.5 equiv, 0.75 mmol) and TMSCl (163 mg, 1.5 mmol) after purification by flash column chromatography (silica, 10%–20% ethyl acetate:hexanes, v/v) gave white amorphous solid 5Ak (80 mg, 47%): m.p. 78.1−79.5 °C; IR (neat) 3076 (w), 2980 (w), 2930 (w), 1728 (s), 1661 (s), 1604 (s)1474 (w), 1454 (m), 1394 (m), 1309 (s), 1258 (m), 1217 (m), 1147 (s), 1110 (s), 1069 (s), 1017 (m), 979 (w), 839 (m), 759 (m) cm−1; 1H-NMR (300 MHz, CDCl3) δ 1.42 (br s, 9H), 2.71 (d, J = 16.8 Hz, 1H), 3.14 (dd, J = 7.2, 16.8 Hz, 1H), 5.32 (d, J = 8.4 Hz, 1H), 5.66 (d, J = 7.2 Hz, 1H), 7.28 (d, J = 8.1 Hz, 2H), 7.53 (d, J = 8.1 Hz, 2H),7.92 (d, J = 8.4 Hz, 1H); 13C-NMR (75 MHz, CDCl3) δ 28.0, 41.6, 55.4, 84.2, 107.2, 125.87, 125.93, 126.2, 130.1, 130.5, 142.8, 151.2, 191.3. 19F-NMR (282 MHz, CDCl3) δ −62.7; HRMS (EI-ion trap) m/z: [M]+ calcd for C17H18NO3F3, 341.1239; found 341.1242.
N-Boc-2-(3,5-ditrifluoromethylphenyl)-2,3-dihydro-4-pyridone (5Al). Employing General Procedure A and using N-Boc-4-pyridone (97 mg, 0.5 mmol), [RhCl(C2H4)2]2 (2.92 mg, 0.0075 mmol), BINAP (10.3 mg, 0.0165 mmol), ArZnCl (1.0 M in THF, 1.5 equiv, 0.75 mmol) and TMSCl (163 mg, 1.5 mmol) after purification by flash column chromatography (silica, 10%–20% ethyl acetate:hexanes, v/v) gave yellow sticky oil 5Al (84 mg, 41%): IR (neat) 2981 (w), 1725 (m), 1671 (m), 1601 (m), 1459 (w), 1417 (w), 1372 (w), 1303 (m), 1275 (s), 1213 (m), 1125 (s), 1014 (m), 897 (m), 846 (m), 766 (m) cm−1; 1H-NMR (300 MHz, CDCl3) δ 1.41 (br s, 9H), 2.73 (d, J = 16.5 Hz, 1H), 3.15 (dd, J = 7.8, 16.5 Hz, 1H), 5.34 (d, J = 8.4 Hz, 1H), 5.69 (d, J = 7.8 Hz, 1H), 7.57 (s, 2H), 7.90 (s, 1H), 7.92 (d, J = 8.4 Hz, 1H); 13C-NMR (75 MHz, CDCl3) δ 27.9, 41.2, 55.0, 84.8, 107.4, 117.6, 121.2, 122.2, 124.8, 126.1, 132.2, 132.6, 141.7, 142.6, 150.9, 190.6. 19F-NMR (282 MHz, CDCl3) δ −63.0; HRMS (EI-ion trap) m/z: [M]+ calcd for C18H17NO3F6, 409.1113; found 409.1119.

4. Conclusions

We have successfully developed rhodium(I)-complexes-catalyzed 1,4-conjugate additions of arylzinc chlorides to N-Boc-4-pyridone in the presence of chlorotrimethylsilane (TMSCl). A combination of [RhCl(C2H4)2]2 and BINAP was determined to be the most effective catalyst for 1,4-conjugate addition of arylzinc chlorides to N-Boc-4-pyridones. We also demonstrated that this reaction is compatible with a broad scope of substrates with both electron-withdrawing and electron-donating substituents on the aromatic ring to afford 1,4-adducts 2-substituted-2,3-dihydropyridones in high yields. Excellent ee can be achieved when (S)-BINAP is used as the chiral ligand. These 1,4-adducts are versatile intermediates that can be utilized for the synthesis of medicinally important N-heterocycles such as pyridines, piperidones, piperidines, indolizidenes, and quinolizidenes.

Supplementary Materials

Supplementary materials are available online.

Acknowledgments

This work was generously supported in part by the National Science Foundation HBCU-UP RIA (NSF award no. 1600987). We also thank Winston Salem State University for start-up fund and Research Initiation Program for providing seed funding for this project. We also like to thank Marcus Wright from Chemistry Department, Wake Forest University, Winston Salem for assistance in attaining NMR spectra.

Author Contributions

Fenghai Guo conceived and designed the experiments; Matthew McGilvary, Malcolm Jeffries, Briana Graves, Shekinah Graham performed the experiments; Yuelin Wu analyzed the data and helped in results and discussion; Fenghai Guo wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Boctert-Butyloxycarbonyl
C-CCarbon carbon
BINAP2,2’-bis(diphenylphosphino)-1,1’-binaphthyl

References and Notes

  1. Perlmutter, P. Conjugate Addition Reactions in Organic Synthesis; Pergamon: Oxford, MS, USA, 1992. [Google Scholar]
  2. Lipshutz, B.H. Organometallics in Organic Synthesis, 2nd ed.; Schlosser, M., Ed.; John Wiley & Sons: Chichester, UK, 2002; pp. 665–815. [Google Scholar]
  3. Alexakis, A.; Benhaim, C. Enantioselective Copper-Catalyzed Conjugate Addition. Eur. J. Org. Chem. 2002, 3221–3236. [Google Scholar] [CrossRef]
  4. Christoffers, J.; Koripelly, G.; Rosiak, A.; Rossle, M. Recent Advances in Metal-Catalyzed Asymmetric Conjugate Additions. Synthesis 2007, 1279–1300. [Google Scholar] [CrossRef]
  5. Alexakis, A.; Backvall, J.-E.; Krause, N.; Pamies, O.; Dieguez, M. Enantioselective copper-catalyzed conjugate addition and allylic substitution reactions. Chem. Rev. 2008, 108, 2796–2823. [Google Scholar] [CrossRef] [PubMed]
  6. Perez, M.; Ayad, T.; Maillos, P.; Poughon, V.; Fahy, J.; Ratovelomanana-Vidal, V. Highly Diastereoselective 1,6-Conjugate Addition of Arylboronic Acids to Securinine. Synlett 2016, 27, 1077–1082. [Google Scholar]
  7. Wei, Y.; Liu, Z.; Wu, X.; Fei, J.; Gu, X.; Yuan, X.; Ye, J. Remote Construction of Chiral Vicinal Tertiary and Quaternary Centers by Catalytic Asymmetric 1,6-Conjugate Addition of Prochiral Carbon Nucleophiles to Cyclic Dienones. Chem. Eur. J. 2015, 21, 18921–18924. [Google Scholar] [CrossRef] [PubMed]
  8. Magrez-Chiquet, M.; Morin, M.S.T.; Wencel-Delord, J.; Amraoui, S.D.; Basle, O.; Alexakis, A.; Crevisy, C.; Mauduit, M. Enantioselective 1,6-Conjugate Addition of Dialkylzinc Reagents to Acyclic Dienones Catalyzed by Cu-DiPPAM Complex—Extension to Asymmetric Sequential 1,6/1,4-Conjugate Addition. Chem. Eur. J. 2013, 19, 13663–13667. [Google Scholar] [CrossRef] [PubMed]
  9. Hayashi, T.; Yamamoto, S.; Tokunaga, N. Rhodium-Catalyzed Asymmetric 1,6-Addition of Aryl Zinc Reagents to Dienones. Angew. Chem. 2005, 117, 4296–4299. [Google Scholar] [CrossRef]
  10. Hayashi, T.; Yamamoto, S.; Tokunaga, N. Rhodium-Catalyzed Asymmetric 1,6-Addition of Aryl Zinc Reagents to Dienones. Angew. Chem. Int. Ed. 2005, 44, 4224–4227. [Google Scholar] [CrossRef] [PubMed]
  11. Hayashi, T.; Tokunaga, N.; Inoue, K. Rhodium-Catalyzed Asymmetric 1,6-Addition of Aryltitanates to Enynones Giving Axially Chiral Allenes. Org. Lett. 2004, 6, 305–307. [Google Scholar]
  12. Phosphoramidite ligands and their application in copper catalyzed conjugate addition reactions: Feringa, B.L. Phosphoramidites: Marvellous ligands in catalytic asymmetric conjugate addition. Acc. Chem. Res. 2000, 33, 346–353. [Google Scholar]
  13. Phosphoramidite ligands and their application in copper catalyzed conjugate addition reactions: Yang, B.; Xie, F.; Yu, H.; Shen, K.; Ma, Z.; Zhang, W. Enantioselective Synthesis of Chiral γ-aryl α-keto Ester by Copper-catalyzed 1,4-conjugate Addition Using D2-symmetric Biphenyl Phosphoramidite Ligand. Tetrahedron 2011, 67, 6197–6201. [Google Scholar]
  14. Chiral phosphite ligands: Xing, A.P.; Bai, C.B.; Wang, L.L. Chiral phosphite ligands derived from L-(+)-tartaric acid: synthesis and application in the Cu-catalyzed 1, 4-conjugate addition of organozincs to cyclic enones. Tetrahedron 2013, 69, 455–459. [Google Scholar] [CrossRef]
  15. Fagnou, K.; Lautens, M. Rhodium-Catalyzed Carbon-Carbon Bond Forming Reactions of Organometallic Compounds. Chem. Rev. 2003, 103, 169–196. [Google Scholar] [CrossRef] [PubMed]
  16. Hayashi, T.; Yamasaki, K. Rhodium-Catalyzed Asymmetric 1,4-Addition and Its Related Asymmetric Reactions. Chem. Rev. 2003, 103, 2829–2844. [Google Scholar] [CrossRef] [PubMed]
  17. Sakai, M.; Hayashi, H.; Miyaura, N. Rhodium-Catalyzed Conjugate Addition of Aryl- or 1-Alkenylboronic Acids to Enones. Organometallics 1997, 16, 4229–4231. [Google Scholar] [CrossRef]
  18. Shintani, R.; Tokunaga, N.; Doi, H.; Hayashi, T. A New Entry of Nucleophiles in Rhodium-Catalyzed Asymmetric 1,4-Addition Reactions: Addition of Organozinc Reagents for the Synthesis of 2-Aryl-4-piperidones. J. Am. Chem. Soc. 2004, 126, 6240–6241. [Google Scholar] [CrossRef] [PubMed]
  19. Shintani, R.; Yamagami, T.; Kimura, T.; Hayashi, T. Asymmetric Synthesis of 2-Aryl-2,3-dihydro-4-quinolones by Rhodium-Catalyzed 1,4-Addition of Arylzinc Reagents in the Presence of Chlorotrimethylsilane. Org. Lett. 2005, 7, 5317–5319. [Google Scholar] [CrossRef] [PubMed]
  20. Sakuma, S.; Miyaura, N. Rhodium(I)-Catalyzed Asymmetric 1,4-Addition of Arylboronic Acids to α,β-Unsaturated Amides. J. Org. Chem. 2001, 66, 8944–8946. [Google Scholar] [CrossRef] [PubMed]
  21. Sakuma, S.; Sakai, M.; Itooka, R.; Miyaura, N. Asymmetric Conjugate 1,4-Addition of Arylboronic Acids to α,β-Unsaturated Esters Catalyzed by Rhodium(I)/(S)-BINAP. J. Org. Chem. 2000, 65, 5951–5955. [Google Scholar] [CrossRef] [PubMed]
  22. Miyaura, N.; Suzuki, A. Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds. Chem. Rev. 1995, 95, 2457–2483. [Google Scholar] [CrossRef]
  23. Hayashi, T.; Ueyama, K.; Tokunaga, N.; Yoshida, K. A Chiral Chelating Diene as A New Type of Chiral Ligand for Transition Metal Catalysts: Its Preparation and Use for the Rhodium-Catalyzed Asymmetric 1,4-Addition. J. Am. Chem. Soc. 2003, 125, 11508–11509. [Google Scholar] [CrossRef] [PubMed]
  24. Defieber, C.; Paquin, J.-F.; Serna, S.; Carreira, E.M. Chiral [2.2.2] Dienes as Ligands for Rh(I) in Conjugate Additions of Boronic Acids to a Wide Range of Acceptors. Org. Lett. 2004, 6, 3873–3876. [Google Scholar] [CrossRef] [PubMed]
  25. Kuuloja, N.; Vaismaa, M.; Franzén, R. Rh-IndOlefOx Catalyzed Conjugate Addition/Heck-type Coupling of Organoboronics to a Lactam or a Lactone. Tetrahedron 2012, 68, 2313–2318. [Google Scholar] [CrossRef]
  26. Glorius, F. Chiral Olefin Ligands-New “Spectators” in Asymmetric Catalysis. Angew. Chem. Int. Ed. 2004, 43, 3364–3366. [Google Scholar] [CrossRef] [PubMed]
  27. Croix, C.; Prié, G.; Chaulet, C.; Viaud-Massuard, M.-C. Rhodium-Catalyzed 1,4-Addition of Arylboronic Acids to 3-Benzylidene-1H-pyrrolo[2,3-b]pyridin-2(3H)-one Derivatives. J. Org. Chem. 2015, 80, 3264–3269. [Google Scholar] [CrossRef] [PubMed]
  28. Asano, N.; Nash, R.J.; Molyneux, R.J.; Fleet, G.W.J. Sugar-mimic Glycosidase Inhibitors: Natural Occurrence, Biological Activity and Prospects for Therapeutic Application. Tetrahedron Asymmetry 2000, 11, 1645–1680. [Google Scholar] [CrossRef]
  29. Watson, P.S.; Jiang, B.; Scott, B. A Diastereoselective Synthesis of 2,4-Disubstituted Piperidines: Scaffolds for Drug Discovery. Org. Lett. 2000, 2, 3679–3681. [Google Scholar] [CrossRef] [PubMed]
  30. Stinson, S.C. Driven by the Needs of the Drug Industry and Fueled by the Ingenuity of Chemists, Sales of Single-enantiomer Chiral Compounds Keep Accelerating. C&EN 2001, 79, 45–57. [Google Scholar]
  31. Thayer, A.M. Centering on Chirality. C&EN 2007, 85, 11–19. [Google Scholar]
  32. Pearson, M.S.M.; Mathe-Allainmat, M.; Fargeas, V.; Lebreton, J. Recent Advances in the Total Synthesis of Piperidine Azasugars. Eur. J. Org. Chem. 2005, 2159–2191. [Google Scholar] [CrossRef]
  33. Cossy, J. Selective Methodologies for the Synthesis of Biologically Active Piperidinic Compounds. Chem. Rec. 2005, 5, 70–80. [Google Scholar] [CrossRef] [PubMed]
  34. Zhou, P.; Chen, B.C.; Davis, F.A. Recent Advances in Asymmetric Reactions Using Sulfinimines (N-sulfinyl imines). Tetrahedron 2004, 60, 8003–8030. [Google Scholar] [CrossRef]
  35. Buffat, M.G.P. Synthesis of Piperidines. Tetrahedron 2004, 60, 1701–1729. [Google Scholar] [CrossRef]
  36. Weintraub, P.M.; Sabol, J.S.; Kane, J.M.; Borcherding, D.R. Recent Advances in the Synthesis of Piperidones and Piperidines. Tetrahedron 2003, 59, 2953–2989. [Google Scholar] [CrossRef]
  37. Laschat, S.; Dickner, T. Stereoselective Synthesis of Piperidines. Synthesis 2000, 1781–1813. [Google Scholar] [CrossRef]
  38. Bailey, P.D.; Millwood, P.A.; Smith, P.D. Asymmetric Routes to Substituted Piperidines. Chem. Commun. 1998. [Google Scholar] [CrossRef]
  39. Šebesta, R.; Pizzuti, M.; Boersma, A.J.; Minnaard, A.J.; Feringa, B.L. Catalytic Enantioselective Conjugate Addition of Dialkylzinc Reagents to N-Substituted-2,3-dehydro-4-piperidones. Chem. Commun. 2005, 1711–1713. [Google Scholar]
  40. Jagt, R.B.C.; de Vries, J.G.; Feringa, B.L.; Minnaard, A.J. Enantioselective Synthesis of 2-Aryl-4-piperidones via Rhodium/Phosphoramidite-Catalyzed Conjugate Addition of Arylboroxines. Org. Lett. 2005, 7, 2433–2435. [Google Scholar] [CrossRef] [PubMed]
  41. Alexakis, A.; Backvall, J.E.; Krause, N.; Pamies, O.; Dieguez, M. Enantioselective Copper-Catalyzed Conjugate Addition and Allylic Substitution Reactions. Chem. Rev. 2008, 108, 2796–2823. [Google Scholar] [CrossRef] [PubMed]and references cited therein.
  42. Feringa, B.L.; Naasz, R.; Imbos, R.; Arnold, L.A. Modern Organocopper Chemistry; Krause, N., Ed.; John Wiley & Sons: New York, NY, USA, 2002; Volume 7, pp. 224–258. [Google Scholar]
  43. Krause, N.; Hoffmann-Röder, A. Recent Advances in Catalytic Enantioselective Michael Additions. Synthesis 2001, 171–196. [Google Scholar] [CrossRef]
  44. Sibi, M.P.; Manyem, S. Enantioselective Conjugate Additions. Tetrahedron 2000, 56, 8033–8061. [Google Scholar] [CrossRef]
  45. Duncan, A.P.; Leighton, J.L. Enantioselective Cu-Catalyzed Conjugate Addition of Diethylzinc to Acyclic Aliphatic Enones. Org. Lett. 2004, 6, 4117–4119. [Google Scholar] [CrossRef] [PubMed]and references cited therein.
  46. Brown, M.K.; Degrado, S.J.; Hoveyda, A.H. Highly Enantioselective Cu-Catalyzed Conjugate Addition of Dialkylzinc Reagents to Unsaturated Furanones and Pyranones: Preparation of Air-Stable and Catalytically Active Cu-Peptide Complexes. Angew. Chem. Int. Ed. 2005, 44, 5306–5310. [Google Scholar] [CrossRef] [PubMed]
  47. Lopóz, F.; Harutyunyan, S.R.; Meetsma, A.; Minnaard, A.J.; Feringa, B.L. Copper-Catalyzed Enantioselective Conjugate Addition of Grignard Reagents to α,β-Unsaturated Esters. Angew. Chem. Int. Ed. 2005, 44, 2752–2756. [Google Scholar] [CrossRef] [PubMed]
  48. Brown, M.K.; Hoveyda, A.H. Enantioselective Total Synthesis of Clavirolide C. Applications of Cu-Catalyzed Asymmetric Conjugate Additions and Ru-Catalyzed Ring-Closing Metathesis. J. Am. Chem. Soc. 2008, 130, 12904–12906. [Google Scholar] [CrossRef] [PubMed]
  49. Alexakis, A.; Albrow, V.; Biswas, K.; d’Augustin, M.; Prieto, O.; Woodward, S. Highly Enantioselective Copper(I)-phosphoramidite-catalysed Additions of Organoaluminium Reagents to Enones. Chem. Comm. 2005, 2843–2845. [Google Scholar] [CrossRef] [PubMed]
  50. For reviews on Grignard reagents and copper catalysis, see; Harutyunyan, S.R.; den Hartog, T.; Geurts, K.; Minnaard, A.J.; Feringa, B.L. Catalytic Asymmetric Conjugate Addition and Allylic Alkylation with Grignard Reagents. Chem. Rev. 2008, 108, 2824–2852. [Google Scholar] [CrossRef] [PubMed] and references cited therein.
  51. For reviews on Grignard reagents and copper catalysis, see; López, F.; Minnaard, A.J.; Feringa, B.L. Catalytic Enantioselective Conjugate Addition with Grignard Reagents. Acc. Chem. Res. 2007, 40, 179–188. [Google Scholar]
  52. Lee, K.-S.; Brown, M.K.; Hird, A.W.; Hoveyda, A.H. A Practical Method for Enantioselective Synthesis of All-Carbon Quaternary Stereogenic Centers through NHC-Cu-Catalyzed Conjugate Additions of Alkyl- and Arylzinc Reagents to β-Substituted Cyclic Enones. J. Am. Chem. Soc. 2006, 128, 7182–7184. [Google Scholar] [CrossRef] [PubMed]
  53. Hayashi, T.; Takahashi, M.; Takaya, Y.; Ogasawara, M. Catalytic Cycle of Rhodium-Catalyzed Asymmetric 1,4-Addition of Organoboronic Acids. Arylrhodium, Oxa-Π-allylrhodium, and Hydroxorhodium Intermediates. J. Am. Chem. Soc. 2002, 124, 5052–5058. [Google Scholar] [CrossRef] [PubMed]
  54. Kina, A.; Ueyame, K.; Hayashi, T. Enantiomerically Pure Rhodium Complexes Bearing 1,5-Diphenyl-1,5-cyclooctadiene as a Chiral Diene Ligand. Their Use as Catalysts for Asymmetric 1,4-Addition of Phenylzinc Chloride. Org. Lett. 2005, 7, 5889–5892. [Google Scholar] [CrossRef] [PubMed]
  55. Guo, X.; Singh, R.P.; Corey, E.J. Enantioselective Synthesis of a Chiral C3-Symmetric Bridgehead Amine. Org. Lett. 2010, 12, 1812–1814. [Google Scholar] [CrossRef] [PubMed]
  56. Dieter, R.K.; Guo, F. Conjugate Addition Reactions of N-Carbamoyl-4-Pyridones with Organometallic Reagents. J. Org. Chem. 2009, 74, 3843–3848. [Google Scholar] [CrossRef] [PubMed]
  57. Guo, F.; Dhakal, R.C.; Dieter, R.K. Conjugate Addition Reactions of N-Carbamoyl-4-pyridones and 2,3-Dihydropyridones with Grignard Reagents in the Absence of Cu(I) Salts. J. Org. Chem. 2013, 78, 8451–8464. [Google Scholar] [CrossRef] [PubMed]
  58. Hayashi, T.; Yamamoto, S.; Tokunaga, N. Rhodium-Catalyzed Asymmetric 1,6-Addition of Aryl Zinc Reagents to Dienones. Angew. Chem. Int. Ed. 2005, 44, 4224–4227. [Google Scholar] [CrossRef] [PubMed]
  59. Liang, Y.; Fu, G. Catalytic Asymmetric Synthesis of Tertiary Alkyl Fluorides: Negishi Cross-Couplings of Racemic α,α-Dihaloketones. J. Am. Chem. Soc. 2014, 136, 5520–5524. [Google Scholar] [CrossRef] [PubMed]
  60. Li, H.; Wu, J. Rhodium(II)-Catalyzed Formal [3 + 2] Cycloaddition of N-Sulfonyl-1,2,3-triazoles with Isoxazoles: Entry to Polysubsituted 3-Aminopyrroles. Org. Lett. 2015, 17, 5424–5427. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds 5Ad, 5Ae, 5Af, 5Ag, 5Ah, and 5Ai are available from the authors.
Scheme 1. Rh(I)-catalyzed C-C bond formation.
Scheme 1. Rh(I)-catalyzed C-C bond formation.
Molecules 22 00723 sch001
Scheme 2. Rh-catalyzed conjugate addition of arylboronic acid.
Scheme 2. Rh-catalyzed conjugate addition of arylboronic acid.
Molecules 22 00723 sch002
Scheme 3. Optimization of Rh-catalyzed conjugate addition.
Scheme 3. Optimization of Rh-catalyzed conjugate addition.
Molecules 22 00723 sch003
Scheme 4. Rh(I)-BINAP-catalyzed conjugate addition reaction.
Scheme 4. Rh(I)-BINAP-catalyzed conjugate addition reaction.
Molecules 22 00723 sch004

Share and Cite

MDPI and ACS Style

Guo, F.; McGilvary, M.A.; Jeffries, M.C.; Graves, B.N.; Graham, S.A.; Wu, Y. Rhodium(I)-Complexes Catalyzed 1,4-Conjugate Addition of Arylzinc Chlorides to N-Boc-4-pyridone. Molecules 2017, 22, 723. https://doi.org/10.3390/molecules22050723

AMA Style

Guo F, McGilvary MA, Jeffries MC, Graves BN, Graham SA, Wu Y. Rhodium(I)-Complexes Catalyzed 1,4-Conjugate Addition of Arylzinc Chlorides to N-Boc-4-pyridone. Molecules. 2017; 22(5):723. https://doi.org/10.3390/molecules22050723

Chicago/Turabian Style

Guo, Fenghai, Matthew A. McGilvary, Malcolm C. Jeffries, Briana N. Graves, Shekinah A. Graham, and Yuelin Wu. 2017. "Rhodium(I)-Complexes Catalyzed 1,4-Conjugate Addition of Arylzinc Chlorides to N-Boc-4-pyridone" Molecules 22, no. 5: 723. https://doi.org/10.3390/molecules22050723

Article Metrics

Back to TopTop