Next Article in Journal
Reductive Catalytic Fractionation of Flax Shive over Ru/C Catalysts
Previous Article in Journal
Integrating Syngas Fermentation into a Single-Cell Microbial Electrosynthesis (MES) Reactor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enantioselective Henry Reaction Catalyzed by Copper(II) Complex of Bis(trans-cyclohexane-1,2-diamine)-Based Ligand

Department of Organic Chemistry, Faculty of Chemical Technology, University of Chemistry and Technology, Technická 5, 166 28 Prague 6, Czech Republic
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(1), 41; https://doi.org/10.3390/catal11010041
Submission received: 2 December 2020 / Revised: 25 December 2020 / Accepted: 28 December 2020 / Published: 31 December 2020
(This article belongs to the Section Catalysis in Organic and Polymer Chemistry)

Abstract

:
Copper(II) complex of the ligand possessing two enantiomerically pure trans-cyclohexane-1,2-diamine units proved to be an efficient catalyst for the enantioselective Henry reaction of aromatic aldehydes with nitromethane. The effect of various reaction conditions on yield and enantioselectivity of the Henry reaction was studied. The results suggest that only one cyclohexane-1,2-diamine unit is involved in catalysis of the Henry reaction.

Graphical Abstract

1. Introduction

Base-catalyzed nucleophilic addition of nitroalkane to carbonyl compound is an important carbon–carbon bond-forming reaction [1,2,3]. The reaction is named Henry reaction according to its discoverer [4] and because of its similarity with the aldol reaction it is also referred as the nitroaldol reaction. A significant feature of the Henry reaction is a possibility to conduct the reaction enantio- or diastereoselectively. Chiral products of the Henry reaction, 2-nitroalcohols, are important intermediates in syntheses of many biologically active compounds, such as antiasthmatic drug (R)-salmeterol [5], antibiotics L-acosamine [6] and bestatin [7], or fungicide (S)-spirobrassinin [8]. In the past, development of the field of stereoselective Henry reaction had been rather slow [9]. One major obstacle is that the reaction components, nitroalkane and a carbonyl compound, lack of suitable points for attachment of chiral auxiliaries. In addition, the reversible nature of the reaction and relatively easy racemization at the stereogenic center (located at the α-position to the nitro group) have hampered the progress in stereocontrol of the Henry reaction. In general, enantioselectivity of the reaction can be controlled by chirality of a substrate, chiral auxiliary, or catalyst. Among these approaches, employment of the chiral catalyst is highly advantageous since the usually expensive and heavily accessible enantiomerically pure compound is used in less than stoichiometric amount.
The first catalyst-controlled enantioselective Henry reaction was reported by Shibasaki [10] in 1992. A large number of catalysts capable of catalyzing the nitroaldol reaction enantioselectively has been developed since then, including both metal catalysts [11] and organocatalysts [12]. For example, copper(II) complexes of bis(oxazolidine) ligands [13], lanthanide complexes of 1,1′-binaphthalene-2,2′-diol derivatives [14], dinuclear zinc(II) complexes of C2-symmetric alcohols [15], and copper(II) complexes of chiral ethane-1,2-diamine derivatives [16,17] were successfully used in the metal-based catalysis. Chiral guanidines [18], bis(thioureas) [19,20] or cinchona alkaloids [21] can serve as examples of efficient organocatalysts.
Recently, mononuclear copper complexes of secondary amines derived from enantiomerically pure trans-cyclohexane-1,2-diamine have been reported to achieve modest to excellent enantioselectivities in the Henry reaction [22,23,24,25,26,27,28,29,30]. Zhang et al. [31] have developed efficient binuclear copper(II) complex of the C2-symmetric ligand, in which two chiral β-amino alcohols are connected at the 1,4-positions to the central benzene ring. Their binuclear complex exhibited significantly higher enantioselectivities over the mononuclear one. Based on these published results, we decided to prepare ligand 1 possessing two trans-cyclohexane-1,2-diamine units attached to the central benzene unit (Figure 1) and study the effect of the second metal center in the binuclear complex on catalytic activity and enantioselectivity of the Henry reaction.

2. Results

2.1. Synthesis of Ligand 1

Preparation of ligand 1 was previously published by Bellemin-Laponnaz et al. [32] and we used the same synthetic strategy with some experimental modifications. The synthesis started from (1S,2S)-trans-cyclohexane-1,2-diamine, which was first monoprotected with the tert-butoxycarbonyl group [33]. Carbamate 2 was condensed with terephthaldehyde, and the resulting Schiff base 3 was reduced with NaBH4. The obtained amine 4 was deprotected and tetraamine 5 was condensed with two molar equivalents of benzaldehyde to give Schiff base, whose two imine double bonds were finally reduced with NaBH4 giving ligand 1 in 36% overall yield (Scheme 1).
Contrary to original authors [32], we had to isolate and carefully purify all intermediates in the reaction sequence. If crude intermediates were used in subsequent steps, we were not able to obtain ligand 1 in sufficient purity.

2.2. Catalytic Studies of the Henry Reaction

The reaction of benzaldehyde with 10 molar equivalents of nitromethane was selected for initial experiments to elucidate the catalytic activity of the copper(II) complex of ligand 1 in the Henry reaction. The copper(II) complexes were generated in situ by mixing the appropriate amount of ligand and copper(II) salt prior to the Henry reaction. After the complex formation, the solution was cooled to 0 °C and benzaldehyde and nitromethane were added. At the end of the reaction, the product was isolated by column chromatography and the enantiomer ratio was determined by HPLC on a column with chiral stationary phase. Isolated yield is reported throughout the article.
In the first experiments, the effect of ligand: copper(II) ratio on enantioselectivity of the reaction was studied using Cu(OAc)2·2H2O as a source of copper(II) ions and ethanol as a solvent (Table 1).
The temperature of 0 °C was chosen based on the results of a preliminary reaction screening. Low enantioselectivity was observed at 20 °C, and the reaction was too slow at temperatures below 0 °C. The reaction mixtures consisted merely of 2-nitroalcohol and unreacted benzaldehyde in all experiments. As can be seen from Table 1, ligand 1 or copper(II) acetate alone (Table 1, entries 1 and 2) does not catalyze the reaction. Because ligand 1 contains two binding sites for the coordination of copper(II) ions, the catalytic properties of complexes prepared at different ratios of ligand 1 to Cu(OAc)2·2H2O were studied (Table 1, entries 3–6). At first, concentration of the Cu(II) ions was kept constant (20 molar %) and the desired ratio was achieved by decreasing the concentration of ligand 1 (Table 1, entries 3–5). The highest yield of the product and the best enantioselectivity were achieved for a 1:1 ratio (Table 1, entry 3). When the ligand 1: Cu(II) ratios of 1:2 or 1:10 were used (Table 1, entries 4 or 5), yield of the Henry reaction significantly dropped down but only a very small decrease in enantioselectivity was observed. If we consider that only one copper(II) ion in the dinuclear complex might be catalytically active, then actual concentration of the "catalytically active site" could decrease as we proceed from the entry 3 to entry 5 in Table 1. A lower concentration of the “catalytically active site” might lead to a lower reaction yield. Therefore, another experiment was performed, in which concentration of ligand 1 was kept constant (compared with the entry 3 in Table 1) and the desired ligand: Cu(II) ratio was achieved by increasing the concentration of Cu(II) ions (Table 1, entry 6). In that case, yield of the reaction was almost the same as in the experiment using a ligand 1: Cu(II) ratio 1:1 and 20 molar % of both components (Table 1, entry 3). In addition, the effect of complex concentration was studied for the 1:1 ligand: copper(II) ratio (Table 1, entries 7 and 8). The results indicate that the concentration of the copper(II) complex in the reaction mixture has little effect on enantioselectivity but significantly affects yield of the Henry reaction.
To find optimal reaction conditions, three different solvents and three different copper(II) salts were tested (Table 2).
Copper(II) perchlorate formed an insoluble complex with ligand 1 in both ethanol and ethyl acetate. In such case, no product formation was observed. Copper(II) acetate proved to be the best source of the copper(II) ions, and the Henry reaction proceeded faster in ethanol then in tetrahydrofuran or ethyl acetate (Table 2, entry 1). Therefore, those two reagents were used in all subsequent experiments.
The above mentioned results confirmed that the copper(II) complex of ligand 1 catalyzes the Henry reaction of benzaldehyde with nitromethane giving (R)-2-nitro-1-phenylethanol in good yield and high enantioselectivity. However, the reaction is rather slow. In order to speed-up the reaction, influence of a base additive was explored. Many successful catalytic systems catalyzing the enantioselective Henry reaction contain achiral base [11,23,26,34]. Therefore, the effect of a base addition to the Cu(II) complex of ligand 1 has been studied in the reaction of benzaldehyde with nitromethane. Three bases were tested as an additive: N,N-diethylethanamine (TEA), N-ethyl-N-(propan-2-yl)propan-2-amine (DIPEA), and pyrrolidine (Table 3).
The results summarized in Table 3 suggest that addition of an achiral base led to a decrease in enantioselectivity. The best result was obtained in the experiment with 10 molar % of DIPEA added (Table 3, entry 4). In this case, the product was obtained in higher yield but enantiomeric excess decreased by 3% compared to the reaction performed in the absence of base (Table 3, entry 1). The addition of TEA or pyrrolidine did not help accelerate the reaction.
Since the best catalytic efficiency in the Henry reaction of benzaldehyde with nitromethane was observed for the catalytic system ligand 1: Cu(OAc)2·2H2O (1:1) in ethanol at 0 °C, the effect of the aldehyde structure was studied under these conditions (Table 4).
The results of this study indicate that both aromatic and aliphatic aldehydes gave relatively good enantioselectivities. However, yield of the Henry reaction was significantly lower when the starting aldehyde was the aliphatic one. Among all studied aldehydes, 4-bromobenzaldehyde exhibited the highest yield but the lowest enantioselectivity. The presence of an electron-donating group in the para-position of the aromatic aldehyde appears to have a positive effect on the reaction yield.

3. Discussion

The present study of the catalytic properties of the copper(II) complexes of bis(trans-cyclohexane-1,2-diamine)-based ligand 1 answered a question of whether the second copper (II) ion in the dinuclear copper(II) complex might somehow contribute to catalysis of the enantioselective Henry reaction. A comparison of our preliminary results with the catalytic properties of the copper(II) complexes of (1R,2R)-N,N′-dibenzylcyclohexane-1,2-diamine (6, Figure 2) published in the literature [22,23] raised expectations that second copper(II) ion might participate in the catalytic process.
Diamine 6 is structurally similar to ligand 1 but possess only one enantiomerically pure cyclohexane-1,2-diamine unit. In 2007, Bandini et al. [22] reported only 56% conversion and 51% e.e. in the reaction of benzaldehyde with nitromethane catalyzed by 12% of the Cu(OAc)2-6 catalyst. The reaction was carried out in ethanol at room temperature for 1.2 h. Later, Jeong et al. [23] published the same reaction catalyzed by 10% of the CuCl2-6 complex with the addition of 6% of achiral base. They run the reaction in methanol at room temperature for one week using TEA as an additional base and isolated (S)-2-nitro-1-phenylethanol in 53% yield and 39% e.e. Significantly higher enantioselectivities (in the range of 85–90% e.e.) were observed in our preliminary experiments, in which we catalyzed the same reaction with the Cu(OAc)2-1 complex.
However, our further studies (Table 1) revealed that the best results were obtained for the ligand 1: copper(II) acetate ratio = 1:1, where the mononuclear complex is the main catalytic species. This observation contradicted the presumed idea of cooperation between the two metal centers in the course of the Henry reaction. To obtain directly comparable data, we repeated the Henry reaction of benzaldehyde with nitromethane catalyzed by the Cu(OAc)2-6 complex under the same reaction conditions as optimized for ligand 1. The catalyst was formed in situ by mixing Cu(OAc)2·2H2O and ligand 6 in the 1:1 ratio. The reaction was carried out in ethanol at 0 °C for 72 h using 20 molar % of catalyst without any additional base. After work-up, (S)-2-nitro-1-phenylethanol was isolated in 64% yield and 89% e.e., i.e., in distinctly higher e.e. than was previously reported [22,23]. Supposedly, higher catalyst loading and lower temperature are responsible for this improvement in e.e. The reaction catalyzed by the Cu(OAc)2-6 complex proceeded more slowly than that of catalyzed by the Cu(OAc)2-1 complex, and yield of the product was slightly lower as well. However, the enantioselectivities achieved were almost the same for both catalysts. This observation complemented our previous results and led us to conclude that only one copper(II) ion is engaged in catalysis of the Henry reaction. The faster reaction rate of the Henry reaction catalyzed by the Cu(OAc)2-1 (1:1) complex compared to the reaction catalyzed by the Cu(OAc)2-6 complex might be explained by the presence of basic secondary amino groups in the close proximity of the catalytic center.
In conclusion, we demonstrated that the Cu(OAc)2-1 (1:1) complex catalyzes the enantioselective Henry reaction of various aldehydes with nitromethane. To achieve good enantioselectivities, relatively high catalyst loading (20 molar %) and low temperatures are necessary. The best results were obtained in the Henry reaction of aromatic aldehydes using copper(II) acetate as a copper(II) salt and ethanol as a solvent in the absence of achiral base.

4. Materials and Methods

4.1. General Methods

Melting points were determined on an Electrothermal IA 9100 melting point apparatus. Elemental analyses were done in the Analytical Laboratory of the Institute of Organic Chemistry and Biochemistry AS CR using a Perkin Elmer 2400 II instrument (Waltham, MA, USA). Optical rotations were recorded on a Perkin-Elmer 241 polarimeter; specific optical rotations [α] and concentrations c are given in deg·10−1·m2·g−1 and g/100 mL, respectively. FTIR spectra were measured on a Nicolet 6700 instrument (Thermo Scientific, Waltham, MA, USA). NMR spectra were recorded on a Varian (Palo Alto, CA, USA) Mercury Plus spectrometer (1H NMR at 299.97 MHz; 13C NMR at 75.43 MHz) and on an Agilent (Santa Clara, CA, USA) MR DDR2 spectrometer (1H NMR at 399.94 MHz; 13C NMR at 100.58 MHz). All NMR measurements were carried out at rt. Chemical shifts δ are reported in ppm and are referenced to the signals of residual non-deuterated solvents (CDCl3 δH 7.26, δC 77.16; DMSO-d6 δH 2.50, δC 39.50). Coupling constants J are given in Hz. Mass spectra were measured on a Waters (Milford, MA, USA) Micromass ZQ (low resolution) or a Thermo Scientific (Waltham, MA, USA) LTQ Orbitrap Velos (high resolution) spectrometer with electrospray ionization. TLC analyses were carried out on aluminum foils coated with Silica gel 60 F254 (Merck, Darmstadt, Germany) and compounds were detected under UV light (λ = 254 nm) or visualized by a treatment with 3% aqueous KMnO4 and heating. Silica gel (Fluka, 63–200 μm) was used for preparative column chromatography. HPLC analyses were done on an Ecom (Prague, Czech Republic) instrument (pump LCP4100, UV detector LCD2083, software CSW32) using the Supelcosil LC-Si (250 × 4.6 mm; 5 μm), YMC Chiral ART Amylose-SA (250 × 4.6 mm; 5 μm) and YMC Chiral Cellulose-SB (250 × 4.6 mm; 5 μm) columns. Dry ethanol was obtained from a PureSOLV PS-MD-7 (VWE International) drying column. Benzaldehyde was successively washed with 10% aqueous Na2CO3, saturated aqueous Na2SO3, and water, and then was dried over MgSO4. Upon the drying agent removal, benzaldehyde was distilled under reduced pressure. Nitromethane was distilled and kept under 5 °C prior to use. Compound 2 was prepared according to literature procedure [33]. Other chemicals and solvents were used as received from commercial sources (Sigma-Aldrich, Darmstadt, Germany; Fluorochem, Hadfield, UK; Penta s.r.o., Prague, Czech Republic). Syntheses of racemic standards of 2-nitroalcohols can be found in the Supplementary Materials.

4.2. Di-tert-butyl [(1S,1′S,2S,2′S)-{[1,4-phenylenebis(methaneylylidene)]bis(azaneylylidene)} bis(cyclohexane-2,1-diyl)]dicarbamate (3)

A solution of terephthaldehyde (1.09 g, 8.1 mmol) in acetonitrile (30 mL) was added to a solution of carbamate 2 (3.48 g, 16.2 mmol) in acetonitrile (100 mL). The reaction mixture was stirred at rt for 18 h. The white precipitate was isolated by filtration and washed with cold acetonitrile (50 mL). The precipitate was re-crystallized from ethyl acetate giving product 3 (3.45 g, 6.5 mmol, 80%) as a white solid; M.p. 238–239 °C; Anal. for C30H46N4O4 calcd. (%) C 68.41, H 8.80, N 10.64; found (%) C 67.9, H 8.6, N 10.3; [α]D20 +94.2 (c 0.4, CH2Cl2); 1H NMR (300 MHz, CDCl3): δ 8.26 (2H, s, CH=N), 7.75 (4H, s, ArH), 4.26 (2H, br s, CONH), 3.76–3.50 (2H, m, CHN), 3.13–2.89 (2H, m, CHN), 2.21–1.98 (2H, m, CH2), 1.93–1.64 (8H, m, CH2), 1.53–1.13 (6H, m, CH2), 1.27 (18H, s, CH3); 13C NMR (101 MHz, CDCl3): δ 159.7, 155.2, 138.1, 128.4, 79.0, 74.8, 54.2, 33.3, 32.0, 28.3, 25.0, 24.2; MS (ESI) m/z 549 ([M + Na]+).

4.3. Di-tert-butyl [(1S,1′S,2S,2′S)-{[1,4-phenylenebis(methylene)]bis(azanediyl)}bis(cyclohexane-2,1-diyl)] dicarbamate (4)

Sodium borohydride (0.88 g, 23.0 mmol) was added to a solution of diimine 5 (3.23 g, 6.1 mmol) in a methanol/tetrahydrofuran mixture (125 mL/125 mL). The reaction mixture was stirred at rt for 18 h. Water (100 mL) was added and the product was extracted into dichloromethane (3 × 100 mL). The combined organic phases were dried over Na2SO4, filtered, and the solvent was evaporated in vacuo. The crude product was crystallized from a dichloromethane/acetonitrile (1:1) mixture at 5 °C. The white crystals were isolated and dried in vacuo to afford dicarbamate 4 (3.07 g, 5.8 mmol, 95%); M.p. 180–181 °C; Anal. for C30H50N4O4 calcd. (%) C 67.89, H 9.50, N 10.56, found (%) C 67.5, H 9.4, N 10.4; [α]D20 +28.6 (c 1.0, CH2Cl2); 1H NMR (300 MHz, CDCl3): δ 7.26 (4H, s, ArH), 4.47 (2H, br s, CONH), 3.88 (2H, d, J 13.2 Hz, ArCHaHbN), 3.66 (2H, d, J 13.2 Hz, ArCHaHbN), 3.44–3.22 (2H, m, CHN), 2.34–2.22 (2H, m, CHN), 2.17–1.96 (4H, m, CH2), 1.78–1.59 (6H, m, CH2 + CH2NHCH), 1.45 (18H, s, CH3), 1.34–0.96 (8H, m, CH2); 13C NMR (101 MHz, CDCl3): 156.0, 139.5, 128.1, 79.1, 60.4, 54.2, 50.2, 32.9, 31.6, 28.4, 24.9, 24.6; MS (ESI) m/z 532 ([M + H]+). NMR spectra are in accordance with those previously published [32].

4.4. (1S,1′S,2S,2′S)-N1,N1′-[1,4-Phenylenebis(methylene)]bis(cyclohexane-1,2-diamine) (5)

Trifluoroacetic acid (20 mL) was added to a solution of dicarbamate 6 (800 mg, 1.5 mmol) in dichloromethane (20 mL) and the mixture was stirred at rt for 3 h. The solvents were evaporated in vacuo and the obtained oil was vigorously stirred with a saturated aqueous solution of KOH (10 mL) for 10 min. The mixture was extracted with ethyl acetate (3 × 60 mL). The combined organic layers were dried over Na2SO4, filtered, and the solvent was evaporated in vacuo to give tetraamine 5 (486 mg, 1.5 mmol, 98%) as yellow oil; 1H NMR (300 MHz, DMSO-d6): δ 7.24 (4H, s, ArH), 3.77 (2H, d, J 13.3 Hz, ArCHaHbN), 3.55 (2H, d, J 13.3 Hz, ArCHaHbN), 2.28–2.15 (2H, m, CHN), 2.03–1.84 (4H, m, CHN + CH2), 1.79–1.67 (2H, m, CH2), 1.62–1.56 (4H, m, CH2), 1.22–0.83 (8H, m, CH2); 13C NMR (101 MHz, DMSO-d6): δ 140.1, 128.1, 62.8, 55.1, 50.4, 35.4, 31.0, 25.3, 25.2; MS (ESI) m/z 331 ([M + H]+).

4.5. (1S,1′S,2S,2′S)-N1,N1′-[(1,4-Phenylenebis(methylene)]bis(N2-benzylcyclohexane-1,2-diamine) (1)

Benzaldehyde (382 mg, 0.37 mL, 3.60 mmol) was added to a solution of tetramine 5 (540 mg, 1.64 mmol) in acetonitrile (20 mL) and the reaction mixture was stirred at rt for 18 h. The solvent was evaporated in vacuo and the residue (crude diimine) was dissolved in a methanol/tetrahydrofuran (20 mL/20 mL) mixture. Sodium borohydride (244 mg, 6.46 mmol) was added and the reaction mixture was stirred at rt for 18 h. Water (20 mL) was added and product was extracted into dichloromethane (3 × 60 mL). The combined organic layers were dried over MgSO4, filtered, and evaporated in vacuo. The crude product was obtained as orange oil (800 mg). It was dissolved in water (20 mL) and pH of the solution was adjusted to 6 using conc. HCl while vigorously stirring. The mixture was stirred at rt for additional 20 min, then a solution of NH4PF6 (1.46 g, 8.9 mmol) in water (5 mL) was added and the mixture was cooled to 5 °C for 18 h. The white precipitate was isolated and dissolved in a vigorously stirred biphasic mixture of dichloromethane (20 mL) and saturated aqueous solution of NaOH (20 mL). The organic layer was separated and the aqueous layer was extracted with dichloromethane (2 × 20 mL). The combined organic layers were dried over MgSO4, filtered and the solvent was evaporated in vacuo to give ligand 1 as yellowish oil (491 mg, 0.97 mmol, 59%); [α]D20 +87.7 (c 0.3, CHCl3); IR νmax (liquid film): 3297, 3025, 2926, 2854, 1453, 1275, 1260, 1116 cm−1; 1H NMR (400 MHz, DMSO-d6): δ 7.30–7.10 (14H, m, ArH), 3.75 (4H, dd, J 13.3, 6.3 Hz, ArCH2N), 3.55 (4H, dd, J 13.3, 7.4 Hz, ArCH2N), 2.22–2.09 (4H, m, CHN), 2.06–1.92 (4H, m, CH2), 1.68–1.53 (4H, m, CH2), 1.20–1.06 (4H, m, CH2), 1.05–0.80 (4H, m, CH2); 13C NMR (101 MHz, DMSO-d6): δ 141.8, 139.9, 128.5, 128.3, 128.1, 126.9, 67.5, 60.6, 50.3, 50.1, 31.1, 31.0, 25.6, 25.0; HRMS (ESI) for C34H46N4 ([M + H]+) calcd. 511.3795, found 511.3792.

4.6. Typical Procedure for the Henry Reaction

Ligand 1 (51 mg, 0.1 mmol) and Cu(OAc)2·2H2O (22 mg, 0.1 mmol) were dissolved in dry ethanol (4.0 mL) and the solution was stirred at rt for 15 min. The mixture was cooled to 0 °C and benzaldehyde (53 mg, 51 μL, 0.5 mmol) was added, followed by addition of nitromethane (122 mg, 107 μL, 2 mmol). The reaction mixture was stirred at 0 °C for 48 h and then concentrated in vacuo. The residue was subjected to a column chromatography in a hexane/ethyl acetate (4:1) mixture. 2-Nitro-1-phenylethan-1-ol was obtained as a light yellow oil (59 mg, 0.35 mmol, 71%); TLC Rf (hexane/ethyl acetate 85:15) 0.2; 1H NMR (300 MHz, CDCl3): δ 7.53–7.30 (5H, m, ArH), 5.48 (1H, dd, J 9.4, 3.2 Hz, CH), 4.68–4.45 (2H, m, CH2), 2.83 (1H, br s, OH); 13C NMR (400 MHz, CDCl3) δ 138.3, 129.1, 129.0, 126.1, 81.3, 71.1. The analytical data are in accordance with those published previously [35]. Enantiomeric excess was determined by means of HPLC analysis on the YMC Chiral Cellulose-SB column (250 × 4.6 mm; 5 μm) using a heptane/propan-2-ol (9:1) mixture at flow rate of 1 mL/min with UV detection at 210 nm. The individual peaks were assigned to particular enantiomers by comparison with HPLC analyses done on a column with chemically equivalent chiral stationary phase [35]: t(R) = 15.1 min (major), t(S) = 17.7 min (minor); e.e. 91%.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/11/1/41/s1. Syntheses of compounds 2, 6 and racemic 2-nitroalcohols, copies of 1H NMR, 13C NMR and MS spectra and HPLC chromatograms.

Author Contributions

Conceptualization, J.H.; methodology, J.H.; investigation, D.T. and M.N.; writing—original draft preparation, D.T. and J.H.; writing—review and editing, J.H.; supervision, J.H.; project administration, J.H.; funding acquisition, J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Czech Science Foundation, grant number 18-09824S, and the specific university research, grant number A2_FCHT_2020_036.

Acknowledgments

The authors thank the Analytical Laboratory of the Institute of Organic Chemistry and Biochemistry of the Academy of Sciences of the Czech Republic headed by Stanislava Matějková for conducting the elemental analyses.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Seebach, D.C.; Ernest, W.; Leher, F.; Weller, T. Nitroaliphatic Compounds—Ideal Intermediates in Organic Synthesis? Chimia 1979, 33, 1–18. [Google Scholar]
  2. Rosini, G. Comprehensive Organic Synthesis; Trost, B.M., Ed.; Pergamon: Oxford, UK, 1991; Volume 2, p. 321. [Google Scholar]
  3. Luzzio, F.A. The Henry reaction: Recent examples. Tetrahedron 2001, 57, 915–945. [Google Scholar] [CrossRef]
  4. Henry, L. Formation synthétique d’alcools nitrés. C. R. 1895, 120, 1265–1268. [Google Scholar]
  5. Guo, Z.-L.; Deng, Y.-Q.; Zhong, S.; Lu, G. Enantioselective synthesis of (R)-salmeterol employing an asymmetric Henry reaction as the key step. Tetrahedron Asymmetry 2011, 22, 1395–1399. [Google Scholar] [CrossRef]
  6. Ginesta, X.; Pastó, M.; Pericàs, M.A.; Riera, A. New Stereodivergent Approach to 3-Amino-2,3,6-trideoxysugars. Enantioselective Synthesis of Daunosamine, Ristosamine, Acosamine, and Epi-daunosamine. Org. Lett. 2003, 5, 3001–3004. [Google Scholar] [CrossRef]
  7. Gogoi, N.; Boruwa, J.; Barua, N.C. A total synthesis of (−)-bestatin using Shibasaki’s asymmetric Henry reaction. Tetrahedron Lett. 2005, 46, 7581–7582. [Google Scholar] [CrossRef]
  8. Liu, L.; Zhang, S.; Xue, F.; Lou, G.; Zhang, H.; Ma, S.; Duan, W.; Wang, W. Catalytic Enantioselective Henry Reactions of Isatins: Application in the Concise Synthesis of (S)-(−)-Spirobrassinin. Chem. Eur. J. 2011, 17, 7791–7795. [Google Scholar] [CrossRef]
  9. Palomo, C.; Oiarbide, M.; Laso, A. Recent Advances in the Catalytic Asymmetric Nitroaldol (Henry) Reaction. Eur. J. Org. Chem. 2007, 2007, 2561–2574. [Google Scholar] [CrossRef]
  10. Sasai, H.; Suzuki, T.; Arai, S.; Arai, T.; Shibasaki, M. Basic character of rare earth metal alkoxides. Utilization in catalytic carbon-carbon bond-forming reactions and catalytic asymmetric nitroaldol reactions. J. Am. Chem. Soc. 1992, 114, 4418–4420. [Google Scholar] [CrossRef]
  11. Ananthi, N.; Velmathi, S. Asymmetric Henry reaction catalysed by transition metal complexes: A short review. Ind. J. Chem. 2013, 52B, 87–108. [Google Scholar]
  12. Alvarez-Casao, Y.; Marques-Lopez, E.; Herrera, R.P. Organocatalytic Enantioselective Henry Reactions. Symmetry 2011, 3, 220–245. [Google Scholar] [CrossRef] [Green Version]
  13. Evans, D.A.; Seidel, D.; Rueping, M.; Lam, H.W.; Shaw, J.T.; Downey, C.W. A New Copper Acetate-Bis(oxazoline)-Catalyzed, Enantioselective Henry Reaction. J. Am. Chem. Soc. 2003, 125, 12692–12693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Sasai, H.; Suzuki, T.; Itoh, N.; Shibasaki, M. Catalytic asymmetric nitroaldol reactions. A new practical method for the preparation of the optically active lanthanum complex. Tetrahedron Lett. 1993, 34, 851–854. [Google Scholar] [CrossRef]
  15. Trost, B.M.; Yeh, V.S.C. A Dinuclear Zn Catalyst for the Asymmetric Nitroaldol (Henry) Reaction. Angew. Chem. Int. Ed. 2002, 41, 861–863. [Google Scholar] [CrossRef]
  16. Xiong, Y.; Wang, F.; Huang, X.; Wen, Y.; Feng, X. A New Copper(I)–Tetrahydrosalen-Catalyzed Asymmetric Henry Reaction and Its Extension to the Synthesis of (S)-Norphenylephrine. Chem. Eur. J. 2007, 13, 829–833. [Google Scholar] [CrossRef] [PubMed]
  17. Arai, T.; Watanabe, M.; Yanagisawa, A. Practical Asymmetric Henry Reaction Catalyzed by a Chiral Diamine-Cu(OAc)2 Complex. Org. Lett. 2007, 9, 3595–3597. [Google Scholar] [CrossRef]
  18. Chinchilla, R.; Nájera, C.; Sánchez-Agulló, P. Enantiomerically pure guanidine-catalysed asymmetric nitroaldol reaction. Tetrahedron Asymmetry 1994, 5, 1393–1402. [Google Scholar] [CrossRef]
  19. Liu, X.-G.; Jiang, J.-J.; Shi, M. Development of axially chiral bis(arylthiourea)-based organocatalysts and their application in the enantioselective Henry reaction. Tetrahedron Asymmetry 2007, 18, 2773–2781. [Google Scholar] [CrossRef]
  20. Otevrel, J.; Bobal, P. Biphenyl-Based Bis(thiourea) Organocatalyst for Asymmetric and syn-Selective Henry Reaction. Synthesis 2017, 49, 593–603. [Google Scholar]
  21. Li, H.; Wang, B.; Deng, L. Enantioselective Nitroaldol Reaction of α-Ketoesters Catalyzed by Cinchona Alkaloids. J. Am. Chem. Soc. 2006, 128, 732–733. [Google Scholar] [CrossRef] [Green Version]
  22. Bandini, M.; Piccinelli, F.; Tommasi, S.; Umani-Ronchi, A.; Ventrici, C. Highly enantioselective nitroaldol reaction catalyzed by new chiral copper complexes. Chem. Commun. 2007, 14, 616–618. [Google Scholar] [CrossRef] [PubMed]
  23. Nguyen, Q.T.; Jeong, J.H. Syntheses and X-ray structures of Cu(II) and Zn(II) complexes of N,N′-dibenzyl-(R,R)-1,2-diaminocyclohexane and application to nitroaldol reaction. Polyhedron 2008, 27, 3227–3230. [Google Scholar] [CrossRef]
  24. Tanaka, K.; Hachiken, S. Enantioselective Henry reaction catalyzed by trianglamine–Cu(OAc)2 complex under solvent-free conditions. Tetrahedron Lett. 2008, 49, 2533–2536. [Google Scholar] [CrossRef]
  25. Kowalczyk, R.; Sidorowicz, Ł.; Skarżewski, J. Asymmetric Henry reaction catalyzed by chiral secondary diamine-copper(II) complexes. Tetrahedron Asymmetry 2008, 19, 2310–2315. [Google Scholar] [CrossRef]
  26. Kowalczyk, R.; Skarżewski, J. Asymmetric nitroaldol reaction catalyzed by copper–diamine complexes: Selective construction of two contiguous stereogenic centers. Tetrahedron Asymmetry 2009, 20, 2467–2473. [Google Scholar] [CrossRef]
  27. Jin, W.; Li, X.; Wan, B. A Highly Diastereo- and Enantioselective Copper(I)-Catalyzed Henry Reaction Using a Bis(sulfonamide)−Diamine Ligand. J. Org. Chem. 2011, 76, 484–491. [Google Scholar] [CrossRef] [PubMed]
  28. Chunhong, Z.; Liu, F.; Gou, S. Application of chiral N,N′-dialkyl-1,2-cyclohexanediamine derivatives in asymmetric copper(II)-catalyzed Henry reactions. Tetrahedron Asymmetry 2014, 25, 278–283. [Google Scholar] [CrossRef]
  29. Cho, J.; Lee, G.H.; Nayab, S.; Jeong, J.H. Copper complexes bearing methylthiophenyl and methylfuranyl derivatives of (R,R)-1,2-diaminocyclohexane: X-ray structures and catalytic exploitation in Henry reaction. Polyhedron 2015, 99, 198–203. [Google Scholar] [CrossRef]
  30. Khlebnikova, T.B.; Konev, V.N.; Pai, Z.P. Levopimaric acid derived 1,2-diamines and their application in the copper-catalyzed asymmetric Henry reaction. Tetrahedron 2018, 74, 260–267. [Google Scholar] [CrossRef]
  31. Lu, G.; Zheng, F.; Wang, L.; Guo, Y.; Li, X.; Cao, X.; Wang, C.; Chi, H.; Dong, Y.; Zhang, Z. Asymmetric Henry reaction catalyzed by Cu(II)-based chiral amino alcohol complexes with C2-symmetry. Tetrahedron Asymmetry 2016, 27, 732–739. [Google Scholar] [CrossRef]
  32. Bissessar, D.; Achard, T.; Bellemin-Laponnaz, S. Robust and Recyclable Self-Supported Chiral Nickel Catalyst for the Enantioselective Michael Addition. Adv. Synth. Catal. 2016, 358, 1982–1988. [Google Scholar] [CrossRef]
  33. Lee, D.W.; Ha, H.J.; Lee, W.K. Selective Mono-BOC Protection of Diamines. Synth. Commun. 2007, 37, 737–742. [Google Scholar] [CrossRef]
  34. Tanaka, K.; Iwashita, T.; Yoshida, E.; Ishikawa, T.; Otuka, S.; Urbanczyk-Lipkowska, Z.; Takahashi, H. Solvent-dependent strong asymmetric amplification in the catalytic enantioselective Henry reaction using the trans-N,N′-bis-biphenyl-4-ylmethyl-cyclohexane-1,2-diamine-CuCl2 complex. Chem. Commun. 2015, 51, 7907–7910. [Google Scholar] [CrossRef] [PubMed]
  35. Ao, C.; Men, J.; Wang, Y.; Shao, T.; Huang, Y.; Huo, J.; Gao, G. The development of new amine–amide ligands for application in Cu(II)-catalyzed enantioselective Henry reactions. Tetrahedron Asymmetry 2016, 27, 589–595. [Google Scholar] [CrossRef]
Figure 1. Structure of ligand 1.
Figure 1. Structure of ligand 1.
Catalysts 11 00041 g001
Scheme 1. Synthesis of ligand 1.
Scheme 1. Synthesis of ligand 1.
Catalysts 11 00041 sch001
Figure 2. Structure of ligand 6.
Figure 2. Structure of ligand 6.
Catalysts 11 00041 g002
Table 1. Effect of molar ratios ligand 1: copper(II) acetate: benzaldehyde on the reaction of benzaldehyde with nitromethane a.
Table 1. Effect of molar ratios ligand 1: copper(II) acetate: benzaldehyde on the reaction of benzaldehyde with nitromethane a.
Catalysts 11 00041 i001
EntryLigand 1
[molar %] b
Cu(OAc)2·2H2O
[molar %] b
Molar Ratio
Ligand 1:Cu(II)
Yield c
[%]
e.e. d
[%]
120----
2-20---
320201:17191
410201:23986
52201:102387
6202001:106984
7551:14089
810101:14688
a Reaction conditions: benzaldehyde (0.2 mmol), nitromethane (2 mmol), ethanol (2.0 mL), 0 °C, 48 h. b Related to benzaldehyde. c Isolated yield. d Determined by HPLC on a column with chiral stationary phase.
Table 2. Effect of counteranion and solvent on the reaction of benzaldehyde with nitromethane catalyzed by the 1:1 ligand 1: Cu(II) complex a.
Table 2. Effect of counteranion and solvent on the reaction of benzaldehyde with nitromethane catalyzed by the 1:1 ligand 1: Cu(II) complex a.
Catalysts 11 00041 i002
EntryCu(II) SaltSolventTemperature
[°C]
Reaction Time [h]Yield c [%]e.e. d [%]
1Cu(OAc)2·2H2OEtOH0487191
2CuCl2EtOH0483893
3 bCu(ClO4)2·6H2OEtOH0---
4Cu(OAc)2·2H2OTHF0726092
5CuCl2THF0720-
6CuCl2THF253363875
7Cu(ClO4)2·6H2OTHF0720-
8Cu(ClO4)2·6H2OTHF25336420
9Cu(OAc)2·2H2OEtOAc0481490
10CuCl2EtOAc072791
11 bCu(ClO4)2·6H2OEtOAc0---
a Reaction conditions: benzaldehyde (0.5 mmol), nitromethane (5 mmol), ligand 1 (0.1 mmol), Cu(II) salt (0.1 mmol), solvent (4.3 mL). b Insoluble complex formed. c Isolated yield. d Determined by HPLC on a column with chiral stationary phase.
Table 3. Influence of a base addition on the reaction of benzaldehyde with nitromethane a catalyzed by the copper(II) complex of ligand 1.
Table 3. Influence of a base addition on the reaction of benzaldehyde with nitromethane a catalyzed by the copper(II) complex of ligand 1.
Catalysts 11 00041 i003
EntryBaseMolar Equiv. of Base bYield c
[%]
e.e. d
[%]
1--7488
2TEA0.17583
3TEA1.06678
4DIPEA0.18385
5DIPEA1.08960
6pyrrolidine0.14583
7pyrrolidine1.03674
a Reaction conditions: benzaldehyde (0.2 mmol), nitromethane (2 mmol), ligand 1 (0.04 mmol), copper(II) acetate (0.04 mmol), ethanol (2.0 mL), 0 °C, 72 h. b Related to benzaldehyde. c Isolated yield. d Determined by HPLC on a column with chiral stationary phase.
Table 4. Effect of the aldehyde structure on yield and enantioselectivity of the Henry reaction a.
Table 4. Effect of the aldehyde structure on yield and enantioselectivity of the Henry reaction a.
Catalysts 11 00041 i004
EntryAldehydeYield b
[%]
e.e. c
[%]
14-bromobenzaldehyde8275
24-methylbenzaldehyde6885
34-nitrobenzaldehyde5085
42-methylpropanal1885
5hexanal1285
6cyclohexanecarbaldehyde2183
73-phenylpropanal2588
a Reaction conditions: aldehyde (0.2 mmol), nitromethane (2 mmol), ligand 1 (0.04 mmol), copper(II) acetate (0.04 mmol), ethanol (2.0 mL), 0 °C, 48 h. b Isolated yield. c Determined by HPLC on a column with chiral stationary phase.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tetour, D.; Novotná, M.; Hodačová, J. Enantioselective Henry Reaction Catalyzed by Copper(II) Complex of Bis(trans-cyclohexane-1,2-diamine)-Based Ligand. Catalysts 2021, 11, 41. https://doi.org/10.3390/catal11010041

AMA Style

Tetour D, Novotná M, Hodačová J. Enantioselective Henry Reaction Catalyzed by Copper(II) Complex of Bis(trans-cyclohexane-1,2-diamine)-Based Ligand. Catalysts. 2021; 11(1):41. https://doi.org/10.3390/catal11010041

Chicago/Turabian Style

Tetour, David, Marika Novotná, and Jana Hodačová. 2021. "Enantioselective Henry Reaction Catalyzed by Copper(II) Complex of Bis(trans-cyclohexane-1,2-diamine)-Based Ligand" Catalysts 11, no. 1: 41. https://doi.org/10.3390/catal11010041

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