Next Article in Journal
Model Organisms in the Fight against Muscular Dystrophy: Lessons from Drosophila and Zebrafish
Previous Article in Journal
Covalent Modification of Human Serum Albumin by the Natural Sesquiterpene Lactone Parthenolide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cu (II)-Catalyzed Asymmetric Henry Reaction with a Novel C1-Symmetric Aminopinane-Derived Ligand

by
Liudmila Filippova
1,
Yngve Stenstrøm
1,* and
Trond Vidar Hansen
1,2
1
Department of Chemistry, Biology and Food Science, Norwegian University of Life Sciences, P.O. Box 5003, N-1432 Ås, Norway
2
Department of Pharmaceutical Chemistry, School of Pharmacy, University of Oslo, P.O. Box 1068, Blindern, N-0316 Oslo, Norway
*
Author to whom correspondence should be addressed.
Molecules 2015, 20(4), 6224-6236; https://doi.org/10.3390/molecules20046224
Submission received: 19 February 2015 / Revised: 30 March 2015 / Accepted: 3 April 2015 / Published: 9 April 2015
(This article belongs to the Section Organic Chemistry)

Abstract

:
A novel C1-symmetric dinitrogen ligand was synthesized in high yield from commercially available (1R,2R,3R,5S)-(−)-isopinocampheylamine and 1-methyl-2-imidazolecarboxaldehyde. In combination with Cu(OAc)2∙H2O, this new ligand promote the reaction between nitromethane and aliphatic aldehydes with high yields (up to 97%) and moderate enantioselectivities (up to 67% ee). The reactions with benzaldehyde required prolonged reaction time that resulted in diminished yields, but accompanied with ee-values in the 55%–76% range.

Graphical Abstract

1. Introduction

The asymmetric Henry (nitroaldol) reaction provides a straightforward entry to enantiomerically enriched β-nitro alcohols, which are valuable intermediates in the synthesis of natural compounds and biologically interesting molecules [1,2]. In particular, the nitroaldol products can be reduced into vicinal amino-alcohols which is a common structural motif found in many pharmaceuticals, such as (−)-pindolol [3], (−)-arbutamine [4], ritonavir [5], (R)-salmeterol [6], and epinephrine [7]. In addition, the β-amino alcohol functionality is present in long-chain lipids such as sphingosines [8,9]. Thus, the development of an efficient asymmetric protocol for this type of reaction is of current interest [10,11].
The first asymmetric procedure for nitroaldol reaction employing a bimetallic Li-La BINOL complex was reported in 1992 by the Shibasaki group [3]. Since then, extensive collection of organocatalysts [12,13,14,15] and transition metal based catalytic systems, i.e., Zn [4,16], Co [17,18] Cr [19,20], Cu [21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50], have been developed with variable success. The majority of these metal-based catalysts reported involved the use of copper complexes with either a bi- or polydentate aza-containing chiral ligand (such as BOX-type [21,22,23], diamines [7,24,25,26,27,28,29,30,31,32], Schiff bases [17,33,34,35] amino-alcohols [9,36,37,38,39], amino/iminopyridines [40,41,42,43,44], sulfonamides [45,46]). These systems have gained some applications, but some limitations still exist. For example, multistep synthetic procedures are often necessary for the preparation of some of the ligands, and obtaining the catalyst in both enantiomeric forms is sometimes a challenge. Moreover, many of the catalytic systems display poor enantioselectivity together with low yields of the products when employing aliphatic aldehydes as substrates. Therefore, the development of easily obtainable, novel ligands, in both enantiomeric forms, is still desired for this class of aldehydes. In this view, the chiral camphor and pinane based terpenes are examples of convenient building blocks for the development of effective chiral auxiliaries. For instance, the copper complexes with camphor-based dinitrogen ligands 1a [40], 1b [7], 1c [47] and 1d [48] have been reported to give high level (up to 98%) of asymmetric induction when reacting nitromethane with a broad range of aldehydes, including aliphatic ones (Figure 1).
Figure 1. Some camphor and pinane-derived N,N-ligands.
Figure 1. Some camphor and pinane-derived N,N-ligands.
Molecules 20 06224 g001
Compounds bearing a pinane core have been utilized in some asymmetric transformations such as the enantioselective addition of diethylzinc to aldehydes [51,52,53], aldol condensation [54] and reduction of ketones [55]. Recently, the copper complex of C2-symmetric ligand 1e was reported to promote the asymmetric Henry reaction between nitromethane and aliphatic aldehydes affording nitroaldol products with enantiomeric excess from 57% to 93% ee [56]. Based on the synthesis of 1c [47], we designed the novel pinane-derived C1-symmetric ligand 4 (Scheme 1). Herein, we describe its application in a copper-catalyzed asymmetric nitroaldol reaction.

2. Results and Discussion

The C1-symmetric ligand 4 was easily prepared from commercially available (1R,2R,3R,5S)-(−)-isopinocampheylamine (2) and 1-methyl-2-imidazolecarboxalehyde in a one-pot reaction as outlined in Scheme 1.
Scheme 1. Synthesis of the ligand 4
Scheme 1. Synthesis of the ligand 4
Molecules 20 06224 g002
Reagents and conditions: (i) 1-methyl-2-imidazolecarboxaldehyde (1 equiv.), MeOH, rt; (ii) NaBH4 (3 equiv), MeOH, 0 °C to rt.
Initially, in connection with our interest in the synthesis of aliphatic nitroalcohols, we examined the ability of 4 to promote the nitroaldol reaction between valeraldehyde (5a) and nitromethane in the presence of CuCl2∙2H2O (5 mol %) at 0 °С with DIPEA (1 equiv.) as base additive. The results are summarized in Table 1. When THF was used as solvent, the reaction proceeded smoothly resulting in high yield (92%), but moderate enantioselectivity was observed (Table 1, entry 1). Then Cu(OAc)2∙H2O was tested to see if there were any influence of the counterions on the reaction. Both copper salts gave similar results in terms of enantioselectivity, but the yields varied significantly depending on the solvent (Table 1, entries 1–4). Most likely, this is due to differences in the solubility of the formed complexes. Hence, the use of other solvents was investigated. Of the solvents tested, i-PrOH was the best choice (Table 1, entry 6, 97% yield, 55% ee). Lowering the temperature of the reaction to −20 °С resulted in a significant decrease of the reaction rate with no improvements in enantioselectivity (Table 1, entries 8 and 9). The absolute configuration in the product 6a was determined based on comparison of the specific optical rotation value with literature [24,25,26,27,28,29,30,31,32].
Table 1. Selected experimental conditions for asymmetric Henry reaction between pentanal and nitromethane.
Molecules 20 06224 i001
Table 1. Selected experimental conditions for asymmetric Henry reaction between pentanal and nitromethane.
Molecules 20 06224 i001
Entry aCopper SaltSolventT (°C)Time (h)DIPEA (mol %)Yield (%) be.e. (%) c
1CuCl2∙2H2OTHF4201009053
2Cu(OAc)2∙H2OTHF4201005252
3CuCl2∙2H2OEtOH4201008833
4Cu(OAc)2∙H2OEtOH4201009237
5Cu(OAc)2∙H2OCH2Cl24201006847
6Cu(OAc)2∙H2Oi-PrOH4201009755
7Cu(OAc)2∙H2OEt2O4201006346
8Cu(OAc)2∙H2OTHF−2520100<5n.d. d
9Cu(OAc)2∙H2Oi-PrOH−25721008857
10Cu(OAc)2∙H2Oi-PrOH421006857
11Cu(OAc)2∙H2Oi-PrOH441008856
12Cu(OAc)2∙H2Oi-PrOH461008555
13Cu(OAc)2∙H2Oi-PrOH481008957
15Cu(OAc)2∙H2Oi-PrOH42003656
16Cu(OAc)2∙H2Oi-PrOH42058857
17Cu(OAc)2∙H2Oi-PrOH420109856
a The reactions were carried out on 0.5 mmol scale of valeraldehyde with 10 equiv. of nitromethane in 2 mL of solvent. b Isolated product. c Enantiomeric excesses by HPLC analyis using a column with a chiral stationary phase. d Not determined.
The Henry reaction is reversible [1]. The enantiomeric excess of the nitroaldol adducts can decrease as a result of the retro-nitroaldol reaction. Hence, a time-course study was performed to investigate the possibility of reversibility. The reaction was almost complete within four hours, while the ee-values of the product 6a remained constant in the conditions explored (Table 1, entry 10–13), indicating the absence of reversibility.
An external base is usually added to promote the reaction. Considering the still low ee of the product observed, and the fact that the excess of additional base can enhance the non-enantioselective reaction pathway [26], the effect of base loading was examined. The addition of even small amounts of external base noticeably improved the reaction rate, while the ee-values of the products were unchanged with increasing amounts of base added. These experiments verified that inhibition of an unselective pathway occurred (Table 1, entries 15–17). The loading of 10 mol % of external base was sufficient to promote the reaction until completion.
Based on the conditions investigated, different aliphatic aldehydes were tested. As shown in Table 2, all of the aldehydes smoothly converted into the nitroaldol products with good to excellent yields. Moderate ee-values ranging from 49% to 67% were observed; the highest ee-value was observed with 2,2-dimethylpropanal (5c) (Table 2, entry 3, 67% ee). Again, the absolute configuration in the products 6a–6h was assigned based on specific optical rotation values.
Then we tested the catalyst system with benzaldehyde (5h) (Table 2, entries 8–10). The reaction between 5h and nitromethane in the presence of ligand 4 and CuCl2∙2H2O did not reach completion even after 20 hours at 4 °С (Table 2, entry 8). Lowering the temperature of the reaction improved the level of enantioinduction to 76% ee, but drastically diminished the yields to impractical values (Table 2, entries 9–10). Hence, other aromatic aldehydes were not investigated.
Finally, in view to explore the nitroalkane substrate scope, nitroethane was reacted with aliphatic aldehydes employing optimized catalytic procedure (Table 3). The corresponding nitroaldol adducts bearing two stereogenic centers were obtained in high yields (up to 94%) albeit moderate diastereoselectivities favoring syn-product and enantioselectivities were observed. Diastereoselectivity was improved when α-branched aldehyde was used (Table 3, entry 2).
Table 2. Enantioselective Henry reaction of various aldehydes with nitromethane.
Molecules 20 06224 i002
Table 2. Enantioselective Henry reaction of various aldehydes with nitromethane.
Molecules 20 06224 i002
Entry aRAldehydeProductYield (%) be.e. (%) c
1n-Bu5a6a9157
2i-Bu5b6b8949
3t-Bu5c6c8867
4i-Pr5d6d8360
5n-nonyl5e6e7555
6n-undecyl5f6f8253
7trans-2-decenyl5g6g6547
8 dphenyl5h6h8038
9 ephenyl5h6h6345
10 fphenyl5h6h1076
a The reactions were performed on 0.5 mmol scale. b Isolated yield. c Determined by HPLC using a column with a chiral stationary phase. d CuCl2∙2H2O was used as a copper source. e The reaction was carried out at −20 °С for 48 h in the presence of 1 equiv. of DIPEA. f The reaction was carried out at −40 °С for 48 h in the presence of 1 equiv. of DIPEA.
Table 3. Diastereoselective Henry reaction using nitroethane.
Molecules 20 06224 i003
Table 3. Diastereoselective Henry reaction using nitroethane.
Molecules 20 06224 i003
Entry aRProductYield (%) bsyn/anti (%) cee (%) d
1i-Bu7a9358/4254/38
2i-Pr7b9481:1966/34
a The reactions were performed on 0.5 mmol scale. b Isolated yield. c Determined by 1H-NMR spectroscopy analysis of isolated compound. d Determined by HPLC analysis using a column with a chiral stationary phase.

3. Experimental Section

All starting reagents and solvents were obtained from Sigma Aldrich and used as purchased without further purification. Analytical TLC was performed using silica gel 60 F254 plates (Merck). Flash column chromatography was performed on silica gel 60 (40–63 nm). NMR spectra were recorded on Varian Gemini and Bruker AscendTM 400 spectrometers 300 MHz and 400 MHz for 1H-NMR and 75 MHz and 100 MHz for 13C-NMR, respectively. Chemical shifts are reported in ppm downfield from tetramethylsilane relative to CDCl3 as internal standard (7.24 ppm for 1H and 77.00 for 13C). IR spectra (4000–600 cm−1) were recorded on a Perkin-Elmer Spectrum BX series FT-IR spectrophotometer using a reflectance cell (HATR). Optical rotations were measured using a 1 mL cell with 1.0 dm path length on Perkin Elmer 341 polarimeter in dedicated solvent. HPLC analyzes were performed on Agilent 1200 Series instrument using chiral OD-H or AD-H columns.

3.1. Preparation of Ligands

3.1.1. Synthesis of (E)-1-(1-Methyl-1H-imidazol-2-yl)-N-((1S,2S,3S,5R)-2,6,6-trimethylbicyclo[3.1.1]heptan-3-yl)methanimine (3)

To (1R,2R,3R,5S)-(−)-isopinocampheylamine (2) (612 mg, 4.0 mmol) was added of 1-methyl-2-imidazolecarboxaldehyde (440 mg, 4.0 mmol) in 20 ml of dry methanol. The reaction mixture was left stirring at room temperature until no more starting material was detected (controlled by TLC, visualized by ninhydrine stain). Then mixture was directly concentrated under reduced pressure and purified by column chromatography eluting with hexane and ethyl acetate to obtain 930 mg of compound 3 as colourless oil (95% yield). [ α ] D 20 = −45.1 (c = 2.0, EtOH). 1H-NMR (CDCl3, 400MHz, TMS): δ = 8.16 (s, 1H), 7.07 (s, 1H), 6.88 (s, 1H), 3.99 (s, 3H), 3.36–3.33 (m, 1H), 2.38–2.25 (m, 2H), 2.10–2.01 (m, 1H), 1.98–1.91 (m, 1H), 1.89–1.81 (m, 2H), 1.22 (s, 3H), 1.47 (d, J = 9.3 Hz, 1H), 1.03 (s, 3H), 0.98 (d, J = 8.7 Hz, 3H) ppm. 13C-NMR (75 MHz, CDCl3, TMS): δ = 149.7, 143.2, 128.9, 124.4, 70.5, 47.5, 44.1, 41.6, 38.7, 36.2, 35.4, 33.6, 27.9, 23.5, 19.8 ppm.

3.1.2. Synthesis of (1S,2S,3S,5R)-2,6,6-Trimethyl-N-((1-methyl-1H-imidazol-2-yl)methyl)-bicyclo[3.1.1]heptan-3-amine (4)

Sodium borohydride (456 mg, 12.1 mmol, 3.2 equiv.) was added in portions to a chilled 0 °C solution of 2 (930 mg, 3.8 mmol) in methanol. The mixture then was allowed to warm to room temperature and left stirred until no more starting material was detected (controlled by TLC, UV detection). The reaction was quenched with aqueous HCl, and MeOH solvent was removed under reduced pressure. The aqueous part was basified until pH = 9–10 and extracted by ethyl acetate, dried over anhydrous Na2SO4 and evaporated. The crude product was purified by flash column chromatography (hexane: EtOAc 7:3) to obtain 884 mg (94% yield) of 4 as a solid compound. [ α ] D 20 = −66.9 (c = 1.85, EtOH). 1H-NMR (CDCl3, 400 MHz, TMS): δ = 6.88 (s, 1H), 6.78 (s, 1H), 3.80 (dd, J1 = 13.2 Hz, J2 = 24.5 Hz, 2H), 3.69 (s, 3H), 2.88–2.81 (m, 1H), 2.41–2.24 (m, 2H), 1.93–1.90 (m, H), 1.78–1.74 (m, H), 1.64–1.57 (m, H), 1.18 (s, 3H), 1.06 (d, J = 7.5 Hz, 3H), 0.93 (s, 3H) ppm. 13C-NMR (75 MHz, CDCl3, TMS): δ = 147.5, 127.7, 121.7, 57.6, 48.4, 45.7, 44.9, 42.3, 39.1, 37.1, 34.3, 33.3, 28.4, 24.0, 22.1 ppm. HRMS (TOF ES+): exact mass calculated for C15H26N3 [M + H]+ 248.2126; found 248.2124.

3.2. General Procedure for the Enantioselective Henry Reaction

Ligand 4 (6.2 mg, 5 mol %) and Cu(OAc)2∙H2O (5 mg, 5 mol %) were added to the test tube containing 2 mL of i-PrOH and stirred for an hour to obtain a blue solution. Then the test tubes were transferred to a bath at the given reaction temperature and 0.5 mmol of aldehyde, 10 equivalents of nitroalkane and 0.1 equivalent of DIPEA (9 μL) were added. The reaction mixture was left stirring for the 24 h, after that the volatile compounds were removed under reduced pressure and the residue was directly purified on silica gel column eluting with hexane:EtOAc to afford the corresponding product.

3.2.1. R-(−)-1-Nitro-2-hexanol (6a)

Purified by column chromatography on silica (hexane:EtOAc 9:1), colorless oil, 90% yield, 57% ee. Enantiomeric excess was determined by HPLC (Chiralcel OD-H column, hexane/isopropanol, 90/10 v/v, 0.5 mL/min, 210 nm): tR (major) = 8.34 min, tR (minor) = 9.95 min; [ α ] D 20 = −6.5° (c = 1.2, CH2Cl2), (lit. [21] [ α ] D 20 = −9.3° (c = 2.73, CH2Cl2, 93% ee (R))).1H-NMR (CDCl3, 300 MHz, TMS): δ = 4.44–4.25 (m, 3H), 2.51 (d, J = 4.7 Hz, 1H), 1.49–1.31 (m, 6H), 0.90 (t, J = 7.0 Hz, 3H). 13C-NMR (100 MHz, CDCl3, TMS): δ = 80.6, 68.7, 33.4, 27.3, 22.4, 13.9 ppm.

3.2.2. R-(+)-4-Methyl-1-nitropentan-2-ol (6b)

Purified by column chromatography (hexane:EtOAc 10:1), colourless oil, 89% yield, 49% ee. Enantiomeric excess was determined by HPLC (Chiralcel AD-H column, hexane/isopropanol, 90/10 v/v, 0.5 mL/min, 23 °C, 210 nm): tR (major) = 13.53 min, tR (minor) = 17.72 min; [ α ] D 20 = +1.0° (c = 1.0, CH2Cl2), (lit. [4] [ α ] D 20 = +2.3° (c = 2.5, CH2Cl2, 82% ee (R))). 1H-NMR (CDCl3, 400 MHz, TMS): δ = 4.40–4.32 (m, 3H), 2.53 (d, J = 4.0 Hz, 1H), 1.84–1.78 (m, 1H), 1.51–1.44 (m, 1H), 1.23–1.17 (m, 1H), 0.94 (app t, J = 6.2 Hz, 6H). 13C-NMR 100 MHz, CDCl3, TMS): δ = 81.0, 67.0, 42.4, 24.3, 23.2, 21.7.

3.2.3. R-(−)-3,3-Dimethyl-1-nitrobutan-2-ol (6c)

Purified by column chromatography (hexane:EtOAc 10:1), colourless oil, 88% yield, 67% ee. Enantiomeric excess was determined by HPLC (Chiralcel OD-H column, hexane/isopropanol, 98/2 v/v, 1.0 mL/min, 23 °C, 210 nm): tR (major) = 16.39 min, tR (minor) = 19.31 min; [ α ] D 20 = −29.3° (c = 0.15, CH2Cl2), (lit. [57] [ α ] D 20 = −28.1° (c = 1.78, CH2Cl2, 69% ee (R))). 1H-NMR (CDCl3, 400 MHz, TMS): δ = 4.5 (dd, J1 = 13.0 Hz, J2 = 2.0 Hz, 1H), 4.35 (dd, J1 = 13 Hz, J2 = 10.1 Hz, 1H), 4.01 (ddd, J1 = 10.1 Hz, J2 = 4.7 Hz, J3 = 2 Hz), 2.39 (d, J = 4.7 1H), 0.96 (s, 9H). 13C-NMR (75 MHz, CDCl3, TMS): δ = 78.2, 76.2, 34.3, 25.6.

3.2.4. R-(−)-3-Methyl-1-nitrobutan-2-ol (6d)

Purified by column chromatography (hexane:EtOAc 10:1), colourless oil, 83% yield, 60% ee. Enantiomeric excess was determined by HPLC (Chiralcel OD-H column, hexane/isopropanol, 98/2 v/v, 1.0 mL/min, 23 °C, 210 nm): tR (major) = 20.46 min, tR (minor) = 22.90 min; [ α ] D 20 = −17.9° (c = 0.2, CH2Cl2), (lit. [45] [ α ] D 20 = −30.6° (c = 0.69, CH2Cl2, 99% ee (R))). 1H-NMR (CDCl3, 400 MHz, TMS): δ = 4.53–4.36 (m, 2H), 4.10–4.07 (m, 1H), 2.51 (br s, 1H), 1.82–1.74 (m, 1H), 0.97 (app t, J = 6.3 Hz). 13C-NMR (100 MHz, CDCl3, TMS): δ = 79.3, 73.4, 31.7, 18.4, 17.5.

3.2.5. R-(−)-1-Nitroundecan-2-ol (6e)

Purified by column chromatography (hexane:EtOAc 10:1), yellowish oil, 82% yield; 55% ee. Enantiomeric excess was determined by HPLC (Chiralcel AD-H column, hexane/isopropanol, 90/10 v/v, 0.5 mL/min, 23 °C, 210 nm): tR (major) = 13.19 min, tR (minor) = 17.99 min; [ α ] D 20 = −2.8° (c = 1.9, CH2Cl2), (lit. [45] [ α ] D 20 = −4.9° (c = 1.04, CH2Cl2, 97% ee (R))). 1H-NMR (CDCl3, 400 MHz, TMS): δ = 4.43–4.28 (m, 3H), 2.47 (d, J = 4.5 Hz, 1H), 1.57–1.44 (m, 3H), 1.37–1.24 (m, 13H), 0.86 (t, J = 6.7 Hz, 3H). 13C-NMR (100 MHz, CDCl3, TMS): δ = 80.6, 68.7, 33.7, 31.9, 29.5, 29.4, 29.3, 29.3, 25.2, 22.7, 14.1.

3.2.6. R-(−)-1-Nitrotridecan-2-ol (6f)

Purified by column chromarography (hexane:EtOAc 10:1), pale yellow oil, 75% yield, 57% ee. Enantiomeric excess was determined by HPLC (Chiralcel AD-H column, hexane/isopropanol, 90/10 v/v, 0.5 mL/min, 23 °C, 210 nm): tR (major) = 11.51 min, tR (minor) = 14.79 min; [ α ] D 20 = −2.5° (c = 0.4, CH2Cl2). 1H-NMR (CDCl3, 400 MHz, TMS): δ = 4.35–4.32 (m, 2H), 4.31–4.27 (m, 1H), 2.47 (d, J = 4.6 Hz, 1H), 1.56–1.44 (m, 2H), 1.34–1.24 (m, 18H), 0.86 (t, J = 6.7 Hz). 13C-NMR (100 MHz, CDCl3, TMS): δ = 80.6, 68.7, 33.7, 31.9, 29.6, 29.5, 29.4, 29.3 (2H), 25.2, 22.7, 14.1.

3.2.7. R-(−)-(E)-1-Nitroundec-3-en-2-ol (6g)

Purified by column chromatography (hexane:EtOAc 95:5), yellow oil, 65% yield, 47% ee. Enantiomeric excess was determined by HPLC (Chiralcel AD-H column, hexane/isopropanol, 98/2 v/v, 0.8 mL/min, 23 °C, 210 nm): tR (major) = 28.80 min, tR (minor) = 30.60 min; [ α ] D 20 = −0.4° (c = 0.7, CH2Cl2). 1H-NMR (CDCl3, 400 MHz, TMS): δ = 5.86 (dtd, J1 = 15.4 Hz, J2 = 6.8 Hz, J3 = 1.1 Hz, 1H), 5.42 (ddt, J1 = 15.4 Hz, J2 = 6.6 Hz, J3 = 1.5 Hz, 1H), 4.79 (pentet, J = 5.6 Hz, 1H), 4.40 (d, J = 5.7 Hz, 2H), 2.40 (d, J = 4.4 Hz, 1H), 2.03 (dd, J1 = 14.2 Hz, J2 = 7.2 Hz, 2H), 1.37–1.25 (m, 10H), 0.86 (t, J = 6.7 Hz, 3H). 13C-NMR (100 MHz, CDCl3, TMS): δ = 136.3, 126.0, 8.1, 70.0, 32.2, 31.8, 29.1, 29.1, 28.8, 22.6, 14.1.

3.2.8. (R)-(−)-1-Phenyl-2-nitroethanol (6h)

Prepared according general procedure and purified by column chromatography (hexane:EtOAc, 8:1), colorless oil, 63% yield, 45% ee. Enantiomeric excess was determined by HPLC (Chiralcel OD-H column, hexane/isopropanol, 90/10 v/v, 0.5 mL/min, 254 nm): tR (major) = 19.88 min, tR (minor) = 24.00 min. [ α ] D 20 = −23.1° (c = 1.2 CH2Cl2), (lit. [45] [ α ] D 20 = −53.1° (c = 1.06 CH2Cl2, 96% ee (R))). 1H-NMR (CDCl3, 300MHz, TMS): δ = 7.40–7.34 (m, 5H), 5.46 (dt, J1 = 9.4 Hz, J2 = 3.5 Hz, 1H), 4.60 (dd, J1 = 13.3 Hz, J2 = 9.3 Hz, 1H), 4.50 (dd, J1 = 13.4 Hz, J2 = 3.2 Hz, 1H), 2.78 (br s, 1H). 13C-NMR (75 MHz, CDCl3, TMS): δ = 138.2, 129.0, 128.9, 126.0, 81.2, 71.0.

3.2.9. (2R,3R)-5-Methyl-2-nitrohexane-3-ol (7a)

Prepared according general procedure and purified by column chromatography (hexane:EtOAc 10:1), colourless oil, 93% yield. Enantiomeric excess was determined by HPLC (Chiralcel AD-H column, hexane/isopropanol, 98/2 v/v, 1.0 mL/min, 23 °C, 220 nm): tR (anti minor) = 16.40 min, tR (anti major) = 17.62 min, tR (syn major) = 21.28 min, tR (syn minor) = 22.85 min. [ α ] D 20 = +15.2° (c = 0.91 EtOH), (lit. [58] [ α ] D 20 = +12.5° (c = 0.96 EtOH, syn/anti 93/7, 99% ee)). Diastereomeric ratio (syn/anti) was determined by 1H-NMR (CDCl3, 400 MHz, TMS): δ = 4.52–4.43 (m, 1H), 4.25–4.22 (m, 0.42H, anti), 3.96–3.91 (m, 0.58H, syn), 2.18 (br s, 1H), 1.88–1.76 (m, 1H), 1.53 (dd, J1 = 6.7 Hz, J2 = 3.8 Hz, 3H), 1.42–1.35 (m, 1H), 1.24–1.16 (m, 1H), 0.95–0.90 (m, 6H). 13C-NMR (100 MHz, CDCl3, TMS): δ = 88.1 (syn), 86.7 (anti), 71.2 (syn), 70.2 (anti), 42.0 (syn), 41.8 (anti), 24.6 (anti), 24.3 (syn), 23.6 (syn), 23.3 (anti), 21.7 (anti), 21.4 (syn), 16.2 (syn), 12.4 (anti).

3.2.10. (3R, 4R)-2-methyl-4-nitropentan-3-ol (7b)

Prepared according general procedure and purified by column chromatography (hexane:EtOAc 10:1), colourless oil, 94% yield. Enantiomeric excess was determined by HPLC (Chiralcel OD-H column, hexane/isopropanol, 99/1 v/v, 1.0 mL/min, 23 °C, 220 nm): tR (anti major) = 18.56 min, tR (syn major) = 19.26 min, tR (anti minor) = 21.66 min, tR (anti minor) = 22.36 min. [ α ] D 20 = −3.4° (c = 1.12 CHCl3), (lit. [58] [ α ] D 20 = −2.0° (c = 0.85 CHCl3, syn/anti 97/3, 90% ee)). Diastereomeric ratio (syn/anti) was determined by 1H-NMR (CDCl3, 400 MHz, TMS): δ = 4.68–4.61 (m, 1H), 3.86 (dd, J1 = 8.0 Hz, J2 = 3.2 Hz, 0.19H, anti), 3.67 (dd, J1 = 7.2 Hz, J2 = 4.5 Hz, 0.81H, syn), 2.18 (br s 1H), 1.82–1.74 (m, 0.81H, syn), 1.72–1.64 (m, 0.19H, anti), 1.53 (d, J = 6.8 Hz, 3H), 1.02 (d, J = 6.8 Hz, 3H), 0.91 (d, J = 6.7 Hz, 3H). 13C-NMR (100 MHz, CDCl3, TMS): δ = 86.2 (syn), 84.6 (anti), 77.3 (syn), 77.2 (anti), 30.8 (anti), 29.9 (syn), 19.8 (anti), 18.8 (syn), 18.6 (anti), 16.4 (syn), 15.5 (syn), 12.1 (anti).

4. Conclusions

In conclusion, a new pinane-based dinitrogen C1-symmetric ligand has been synthesized via an experimental simple one pot procedure. This ligand was found to be reasonably selective for the copper catalyzed asymmetric Henry reaction between nitromethane and aliphatic aldehydes. Of merit, high yields were observed for this class of aldehydes. The ease of availability of ligand 4 in both enantiomeric forms, together with the simple and scalable procedure for its preparation, is of advantage for the present catalytic system. Currently, this catalyst system is investigated towards the synthesis of vicinal amino-alcohol derived lipids.

Acknowledgments

Financial support from the University of Life Sciences to L.F. (Quota Scheme) and a grant from the Norwegian Research Council, NFR 209335 are acknowledged.

Author Contributions

LF performed all the experiments. TVH designed the concept of the study. LF, YS, and TVH wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ono, N. The Nitrogroup in Organic Synthesis; Wiley-VCH: New York, USA, 2001. [Google Scholar]
  2. Luzzio, F.A. The Henry reaction: Recent examples. Tetrahedron 2001, 57, 915–945. [Google Scholar] [CrossRef]
  3. Sasai, H.; Suzuki, T.; Arai, S.; Shibasaki, M. Basic character of rare earth metal alkoxides. Utilization in catalytic C-C-bond-forming reactions and catalytic asymmetric nitroaldol reactions. J. Am. Chem. Soc. 1992, 114, 4418–4420. [Google Scholar] [CrossRef]
  4. Trost, B.M.; Yeh, V.S.C.; Ito, H.; Bremeyer, N. Effect of ligand structure on zinc-catalyzed Henry reaction. Asymmertic synthesis of (−)-denopamine and (−)-arbutamine. Org. Lett. 2002, 4, 2621–2623. [Google Scholar] [CrossRef] [PubMed]
  5. Roy, A.; Reddy, L.A.; Dwivedi, N.; Naram, J.; Swapna, R.; Malakondaiah, G.C.; Ravikumar, M.; Bhalerao, D.; Pratap, T.P.; Reddy, P.P.; et al. Diastereoselective synthesis of a core fragment of ritonavir and lopinavir. Tetrahedron Lett. 2011, 52, 6968–6970. [Google Scholar] [CrossRef]
  6. 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]
  7. Zhou, Y.; Dong, J.; Zhang, F.; Gong, Y. Synthesis of C1-symmetric chiral secondary diamines and their applications in the asymmetric copper(II)-catalyzed Henry (nitroaldol) reactions. J. Org. Chem. 2011, 76, 588–600. [Google Scholar] [CrossRef] [PubMed]
  8. Uraguchi, D.; Nakamura, S.; Ooi, T. Catalytic asymmetric direct Henry reaction of ynals: Short synthesis of (2S,3R)-(+)-xestoaminol C and (−)-codonopsinines. Angew. Chem. Int. Ed. 2010, 49, 7562–7565. [Google Scholar] [CrossRef]
  9. Xu, K.; Lai, G.; Zha, Z.; Pan, S.; Chen, H.; Wang, Z. A highly anti-selective asymmetric Henry reaction catalyzed by a chiral copper complex: Applications to the syntheses of spisulisine and a pyrroloisoquinoline derivative. Chem. Eur. J. 2012, 18, 12357–12362. [Google Scholar] [CrossRef] [PubMed]
  10. Palomo, C.; Oiarbide, M.; Laso, A. Recent advances in catalytic asymmetric nitroaldol (Henry) reaction. Eur. J. Org. Chem. 2007, 2561–2574. [Google Scholar] [CrossRef]
  11. Boruwa, J.; Gogoi, N.; Saikia, P.P.; Barua, N.C. Catalytic asymmetric Henry reaction. Tetrahedron: Asymmetry 2006, 17, 3315–3326. [Google Scholar] [CrossRef]
  12. Alvarez-Casao, Y.; Marques-Lopez, E.; Herrera, R.P. Organocatalytic enantioselective Henry reaction. Symmetry 2011, 3, 220–245. [Google Scholar] [CrossRef] [Green Version]
  13. Ube, H.; Terada, M. Enantioselective Henry (nitroaldol) reaction catalyzed by axially chiral guanidines. Bioorg. Med. Chem. Lett. 2009, 19, 3895–3898. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, L.; Zhang, F.; Xue, F.; Lou, G.; Zhang, H.; Ma, S.; Duan, W.; Wang, W. Catalytic enantioselective Henry reaction of isatins: Application in concise synthesis of (S)-(−)-spirobrassinin. Chem. Eur. J. 2011, 17, 7791–7795. [Google Scholar] [CrossRef] [PubMed]
  15. Marcelli, T.; Haas, R.N. S.; Maarseveen, J.H.; Hiemstra, H. Asymmetric organocatalytic Henry reaction. Angew. Chem. Int. Ed. 2006, 45, 929–931. [Google Scholar] [CrossRef]
  16. Trost, B.; Yeh, V. A dinuclear Zn calalyst for the asymmetric nitroaldol (Henry) reaction. Angew. Chem. Int. Ed. 2002, 41, 861–863. [Google Scholar] [CrossRef]
  17. Lang, K.; Park, J.; Hong, S. Urea/transition-metal cooperative catalyst for anti-selective asymmetric nitroaldol reaction. Angew. Chem. Int. Ed. 2012, 51, 1620–1624. [Google Scholar] [CrossRef]
  18. Kogami, Y.; Nakajima, T.; Ikeno, T.; Yamada, T. Enantioselective Henry reaction catalyzed by salen-cobalt complexes. Synthesis 2004, 1947–1950. [Google Scholar]
  19. Zulauf, A.; Mallah, M.; Schulz, E. New chiral thiophene chromium complexes for asymmetric Henry reaction. J. Org. Chem. 2009, 74, 2242–2245. [Google Scholar] [CrossRef] [PubMed]
  20. Kowalczyk, R.; Sidorowicz, L.; Skarzewski, J. Asymmetric nitroaldol reaction catalyzes by a chromium(III)-salen system. Tetrahedron: Asymmetry 2007, 18, 2581–2586. [Google Scholar] [CrossRef]
  21. 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]
  22. Ginotra, S.K.; Singh, V.K. Enantioselective Henry reaction catalyzed by a C2-symmetric bis(oxazoline)- Cu(OAc)2∙H2O complex. Org. Biomol. Chem. 2007, 5, 3932–3937. [Google Scholar] [CrossRef] [PubMed]
  23. Lee, J.M.; Kim, J.; Shin, Y.; Yeom, C.E.; Lee, J.E.; Hueon, T.; Kim, B.M. Heterogeneous asymmetric Henry reaction using a chiral bis(oxazoline)-copper complex immobilized in magnetically separable mesocellular mesoporous silica support. Tetrahedron: Asymmetry 2010, 21, 285–291. [Google Scholar] [CrossRef]
  24. Kowalczyk, R.; Sidirowicz, L.; Skarzewski, J. Asymmetric Henry reaction catalyzed by chiral secondary diamine-copper(II) complexes. Tetrahedron: Asymmetry 2008, 19, 2310–2315. [Google Scholar] [CrossRef]
  25. Luo, M.; Yan, B. Enantioselective Henry reaction catalyzed cy chiral N-metal complexes containing R(+)/S(−)-α-ethylphenyl amines. Tetrahedron Lett. 2010, 51, 5577–5580. [Google Scholar] [CrossRef]
  26. Kodama, K.; Sugawara, K.; Hirose, T. Benzylidene acetal type bridged nucleic acids: Changes in properties upon cleavage of the bridged triggered by external stimuli. Chem. Eur. J. 2011, 17, 7918–7926. [Google Scholar] [CrossRef] [PubMed]
  27. Selvakumar, S.; Sevasankaran, D.; Singh, V.K. Enantioselective Henry reaction catalyzed by C2-symmetric chiral diamine-copper(II) complex. Org. Biomol. Chem. 2009, 7, 3156–3162. [Google Scholar] [CrossRef]
  28. Arai, T.; Watanabe, M.; Yanadisawa, A. Practical Asymmetric Henry reaction catalyzed by a chiral diamine- Cu(OAc)2 complex. Org. Lett. 2007, 9, 3595–3597. [Google Scholar] [CrossRef] [PubMed]
  29. Noole, A.; Lippur, K.; Metsala, A.; Lopp, M.; Kanger, T. Enantioselective Henry reaction catalyzed by Cu (II) salt and bipiperidine. J. Org. Chem. 2010, 75, 1313–1316. [Google Scholar] [CrossRef] [PubMed]
  30. Chougnet, A.; Zhang, G.; Liu, K.; Haussinger, D.; Kagi, A.; Allmendinger, T.; Woggon, W.D. Diastereoselective and highly enantioselective Henry reactions using C1-symmertical copper(II) complexes. Adv. Synth. Catal. 2011, 353, 1797–1806. [Google Scholar] [CrossRef]
  31. 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]
  32. Liu, F.; Gou, S.; Li, L. Asymmetric Henry reaction of aldehydes vith various nitroalkanes catalyzed by copper(II) complexes on novel chiral N-monoalkyl cyclohexane-1,2-diamines. Appl. Organometal. Chem. 2014, 28, 186–193. [Google Scholar] [CrossRef]
  33. Wei, Y.; Yao, L.; Zhang, B.; He, W.; Zhang, S. Novel Schiff base ligand derived from cinchona alkaloids for Cu(II)-catalyzed asymmetric Henry reaction. Tetrahedron 2011, 67, 8552–8558. [Google Scholar] [CrossRef]
  34. Boobalan, R.; Lee, G.H.; Chen, C. Copper complex of aminoisoborneol Schiff base Cu2(SBAIB-d)2: An efficient catalyst for direct catalytic asymmetric nitroaldol (Henry) reaction. Adv. Synth. Catal. 2012, 354, 2511–2520. [Google Scholar]
  35. Yang, W.; Liu, H.; Du, D.M. Efficient in situ three-component formation of chiral oxazoline-Schiff base copper(II) complexes: Towards combinatorial library of chiral catalysts for asymmetric Henry reaction. Org. Biomol. Chem. 2010, 8, 2956–2960. [Google Scholar] [CrossRef] [PubMed]
  36. Lai, G.; Guo, F.; Zheng, Y.; Fang, Y.; Song, H.; Xu, K.; Wang, S.; Zha, Z.; Wang, Z. Highly enantioselective Henry reactions in water catalyzed by a copper tertiary amine complex and applied in the synthesis of (S)-N-trans-feruloylctopamine. Chem. Eur. J. 2011, 17, 1114–1117. [Google Scholar] [CrossRef] [PubMed]
  37. Guo, Z.L.; Zhong, S.; Li, Y.B.; Lu, G. Chiral 1,1'-binaphthylazepine derived amino-alcohol catalyzed asymmetric Henry reaction. Tetrahedron: Asymmetry 2011, 22, 238–245. [Google Scholar] [CrossRef]
  38. Qin, D.D.; Yu, W.; Zhou, J.D.; Zhan, Y.C.; Ruan, Y.P.; Zhou, Z.H.; Chen, H.B. Syn- and enantioselective Henry reactions of aliphatic aldehydes and application to the synthesis of safingol. Chem. Eur. J. 2013, 19, 16541–16544. [Google Scholar] [CrossRef] [PubMed]
  39. Reddy, B.S.; Reddy, S.M.; Manisha, S.; Madan, C. Asymmetric Henry reaction catalyzed by a chiral Cu(II) complex: A facile enantioselective synthesis of (S)-2-nitro-1-arylethanols. Tetrahedron: Asymmetry 2011, 22, 530–535. [Google Scholar] [CrossRef]
  40. Blay, G.; Climent, E.; Fernandez, I.; Hernandez-Olmos, V.; Pedro, J.R. Enantioselective Henry reaction catalyzed with copper(II)-iminopyridine complexes. Tetrahedron: Asymmetry 2007, 18, 1603–1612. [Google Scholar] [CrossRef]
  41. Blay, G.; Domingo, L.R.; Hernandez-Olmos, V.; Pedro, J.R. New highly asymmetric Henry reaction catalyzed by CuII and a C1-symmetric aminopyridine ligand, and its application to the synthesis of miconazole. Chem. Eur. J. 2008, 14, 4725–4730. [Google Scholar] [CrossRef] [PubMed]
  42. Ji, Y.Q.; Qi, G.; Judeh, Z.M.A. Efficient asymmetric copper(I)-catalyzed Henry reaction using chiral N-alkyl-C1-tetrahydro-1,1'-bisisoquinolines. Eur. J. Org. Chem. 2011, 4892–4898. [Google Scholar]
  43. Ji, Y.Q.; Qi, G.; Judeh, Z.M.A. Catalytic anti-selective asymmetric Henry (nitroaldol) reaction catalyzed by Cu(I)-amine-imine complexes. Tetrahedron: Asymmetry 2011, 22, 2065–2070. [Google Scholar] [CrossRef]
  44. Blay, G.; Hernandez-Olmos, V.; Pedro, J.R. Synthesis of (S)-(+)-sotalol and R-(−)-isoproterenol via a catalytic enantioselective Henry reaction. Tetrahedron: Asymmetry 2010, 21, 578–581. [Google Scholar] [CrossRef]
  45. Jin, W.; Li, X.; Huang, Y.; Wu, F.; Wan, B. A highly effective bis(sulfonamide)-diamine ligand: A unique chiral skeleton for the enantioselective Cu-catalyzed Henry reaction. Chem. Eur. J. 2010, 16, 8259–8261. [Google Scholar] [CrossRef] [PubMed]
  46. Arai, T.; Takashita, R.; Endo, Y.; Watanabe, M.; Yanagisawa, A. Asymmetric syn-selective Henry reaction catalyzed by the sulfonyldiamine-CuCl-pyridine system. J. Org. Chem. 2008, 73, 4903–4906. [Google Scholar] [CrossRef] [PubMed]
  47. Zhou, Y.; Gong, Y. Asymmetric copper(II)-catalysed nitroaldol (Henry) reactions utilizing a chiral C1-symmetric dinitrogen ligand. Eur. J. Org. Chem. 2011, 6092–6099. [Google Scholar] [CrossRef]
  48. Sanjeevakumar, N.; Periasamy, M. Highlly enantioselective Henry reaction catalyzed by a new chiral C2-symmetric N,N'-bis(isobornyl)ethylendiamine-copper complex. Tetrahedron: Asymmetry 2009, 20, 1842–1847. [Google Scholar] [CrossRef]
  49. Ni, B.; He, J. Highly asymmetric Henry reaction catalyzed by chiral copper(II) complexes. Tetrahedron Lett. 2013, 54, 462–465. [Google Scholar] [CrossRef]
  50. Sema, H.A.; Bez, G.; Karmakar, S. Asymmetric Henry reaction catalyzed by L-proline derivatives in the presence of Cu(OAc)2: Isolation and characterization of an in situ formed Cu(II) complex. Appl. Organometal. Chem. 2014, 28, 290–297. [Google Scholar] [CrossRef]
  51. Csillag, K.; Nemeth, L.; Martinek, T.A.; Szakonyi, Z.; Fulop, F. Stereoselective synthesis of pinane-type tridentate aminodiols and their application in the enantioselective addition of diethylzinc to benzaldehyde. Tetrahedron: Asymmetry 2012, 23, 144–150. [Google Scholar] [CrossRef]
  52. Binder, C.M.; Bautista, A.; Zaidlewicz, M.; Krzeminski, M.P.; Oliver, A.; Singaram, B. Dual stereoselectivity in the dialkylzinc reaction using (−)-β-pinene derived amino alcohol chiral auxiliaries. J. Org. Chem. 1999, 74, 2337–2343. [Google Scholar] [CrossRef]
  53. Cherng, Y.J.; Fang, J.M.; Lu, T.J. Pinane-type reagents for enantioselective reactions: Reduction of ketones and addition of diethylzinc to aldehydes. J. Org. Chem. 1999, 64, 3207–3212. [Google Scholar] [CrossRef] [PubMed]
  54. Siyutkin, D.E.; Kucherenko, A.S.; Frolova, L.L.; Kuchin, A.V.; Zlotin, S.G. 2-Hydroxy-3-[(S)-prolinamido]pinanes as novel bifunctional organocatalysts for asymmetric aldol reactions in aqueous media. Tetrahedron: Asymmetry 2011, 22, 1320–1324. [Google Scholar] [CrossRef]
  55. Krzeminski, M.P.; Cwiklinska, M. Chiral terpene auxiliaries II. Spiroborate esters derived from α-pinene—new catalysts for asymmetric borane reduction of prochiral ketones. Tetrahedron Lett. 2011, 52, 3919–3921. [Google Scholar] [CrossRef]
  56. Malhotra, S.V.; Brown, H.C. C2-symmetric N,N'-bis(terpenyl)ethylenediamines synthesis and application in the enantioselective nitroaldol reaction. RSC Adv. 2014, 4, 14264–14269. [Google Scholar] [CrossRef]
  57. Dhahagani, K.; Rajesh, J.; Kanan, R.; Rajagopal, G. Asymmetric Henry reaction of aldehydes catalyzed by recyclable MCM-41 supported copper(II) salen complex. Tetrahedron: Asymmetry 2011, 22, 857–865. [Google Scholar] [CrossRef]
  58. Sohtome, Y.; Hasimoto, Y.; Nagasawa, K. Diastereoselective and enantioselective Henry (nitroaldol) reaction utilizing a guanidine-thiourea bifunctional organocatalyst. Eur. J. Org. Chem. 2006, 13, 2894–2897. [Google Scholar] [CrossRef]
  • Sample Availability: Sample of the ligand 4 is available from the authors.

Share and Cite

MDPI and ACS Style

Filippova, L.; Stenstrøm, Y.; Hansen, T.V. Cu (II)-Catalyzed Asymmetric Henry Reaction with a Novel C1-Symmetric Aminopinane-Derived Ligand. Molecules 2015, 20, 6224-6236. https://doi.org/10.3390/molecules20046224

AMA Style

Filippova L, Stenstrøm Y, Hansen TV. Cu (II)-Catalyzed Asymmetric Henry Reaction with a Novel C1-Symmetric Aminopinane-Derived Ligand. Molecules. 2015; 20(4):6224-6236. https://doi.org/10.3390/molecules20046224

Chicago/Turabian Style

Filippova, Liudmila, Yngve Stenstrøm, and Trond Vidar Hansen. 2015. "Cu (II)-Catalyzed Asymmetric Henry Reaction with a Novel C1-Symmetric Aminopinane-Derived Ligand" Molecules 20, no. 4: 6224-6236. https://doi.org/10.3390/molecules20046224

Article Metrics

Back to TopTop