Next Article in Journal
Three-Component Condensation of β-Ketonitriles, 4-Fluorobenzaldehyde, and Secondary Cyclic Amines
Next Article in Special Issue
Reduction and Cycloaddition of Heteroalkenes at Ga(I) Bisamide Center
Previous Article in Journal
Selective O-alkylation of Phenol Using Dimethyl Ether
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of Multiple Functionalized Cyclohexanone Using Diels–Alder Reaction of α-Nitrocinnamate

1
School of Environmental Science and Engineering, Kochi University of Technology, Tosayamada, Kami, Kochi 782-8502, Japan
2
Research Center for Molecular Design, Kochi University of Technology, Tosayamada, Kami, Kochi 782-8502, Japan
*
Author to whom correspondence should be addressed.
Reactions 2022, 3(4), 615-624; https://doi.org/10.3390/reactions3040041
Submission received: 7 November 2022 / Revised: 17 November 2022 / Accepted: 29 November 2022 / Published: 2 December 2022
(This article belongs to the Special Issue Cycloaddition Reactions at the Beginning of the Second Millennium)

Abstract

:
A systematic study of the Diels–Alder reaction of α-nitrocinnamate was performed. The reaction of p-substituted α-nitrocinnamate with 2,3-dimethyl-1,3-butadienes smoothly proceeded regardless of the p-substituent, which was either an electron-donating or -withdrawing group. A control reaction revealed that α-nitrocinnamate isomerized during the reaction. Danishefsly’s diene (1-methoxy-3-trimethylsiloxy-1,3-butadiene) facilitated cycloaddition under mild conditions to afford a cycloadduct without the alternation of the diastereomeric ratio. Moreover, the desilylation of the cycloadduct furnished multiple functionalized cyclohexanones.

Graphical Abstract

1. Introduction

Nitroalkenes are one of the important building blocks for multiple functionalized compounds. The electron-withdrawing effect of the nitro group enhances the electron acceptability of the alkene moiety, which facilitates conjugate addition [1,2], the Diels–Alder reaction [1,3,4] and 1,3-dipolar cycloaddition [5]. Further chemical conversions of the conjugate adducts and cycloadducts can furnish amines [6], alkenes [1,7], and carbonyl compounds (Nef reaction) [8] via reduction, denitration, and hydrolysis, respectively. The introduction of an electron-withdrawing group such as a carbonyl functionality at the α position of nitroalkenes enhances electron acceptability [9]. For instance, α-nitrocinnamate 1 facilitates conjugate addition by hetero-aromatic compounds, which are less nucleophilic species [10,11]. α-carbonylated nitroalkenes should also be suitable for the Diels–Alder reaction as dienophiles because the carbonyl group decreases the LUMO level of the nitroalkenes. However, there have been few examples of the Diels–Alder reaction using α-carbonylated nitroalkenes [12,13,14,15,16,17,18,19] despite abundant studies on that using α-unsubstituted nitroalkenes [1,3,4]. However, related studies of heterocyclic compounds such as coumarin [20,21], quinolone [22,23,24], pyridone [24] and pyridazine [25] possessing an α-carbonylated nitroalkene moiety as a partial structure have been reported. Accordingly, systematic studies using α-carbonylated nitroalkene can help expand the synthetic utility of the Diels–Alder reaction. In the present work, a series of α-nitrocinnamates 1 were reacted with dienes, and their reactivity and regioselectivity were investigated.

2. Results and Discussion

When a toluene solution of ethyl α-nitrocinnamate 1a was heated with 1,3-butadiene 2 at 150 °C under microwave irradiation, product 3a was isolated in a 13% yield (Table 1, Entry 1). In the 1H NMR of 3a, two singlet signals were inequivalently observed at around 1.7 ppm, indicating that symmetrical butadiene was converted into an unsymmetrical structure. Other spectral data also supported the idea that difunctionalized cyclohexene 3a was successfully obtained. To increase the yield of 3a, several solvents were tested, and acetonitrile was found to be the most suitable (Entries 1–4). Though six hours were necessary to consume 1a at 150 °C (Entry 5), the reaction was completed within two hours when the reaction temperature was elevated to 180 °C (Entry 6).
The optimal conditions were applied to several ethyl α-nitrocinnamates 1bf possessing different substituent at the para position (Scheme 1). Each reaction used two diastereomeric products with different positions of the nitro and ethoxycarbonyl groups at the 3 position. Reactions efficiently proceeded, with substrates substituting either the electron-donating or -withdrawing group to afford the corresponding products 3be; however, the reaction of nitro-substituted substrate 1f was complicated due to overreactions such as the elimination of a nitrous acid and subsequent oxidation. This disadvantage was overcome by reducing the reaction temperature to 150 °C, and cycloadduct 3f was quantitatively obtained. Among substrates 1af, methoxy- and bromo-substituted cinnamates 1b and 1d were the least reactive, which was presumably due to the electron-donating resonance effect that increased the single bond character of the double-bond moiety. Systematic studies showed that the diastereomeric ratio of products 3af was about 60/40 even when only a Z isomer was used as a starting material, which means that isomerization occurred during the reaction. To reveal this phenomenon, only the Z isomer of 1a was heated at 180 °C, which afforded a Z/E mixture of 1a with a 60/40 ratio (Scheme 2). This result indicated that the isomerization of 1 occurred under reaction conditions and not during the Diels–Alder reaction.
Next, α-cyano and α-acetylcinnamates 4 and 5 were employed as dienophiles instead of α-nitro derivative 1 (Scheme 3). Even when the nitro group was replaced with less electron-withdrawing acetyl groups, the cycloaddition proceeded to furnish the corresponding cycloadduct 6 with a 87/13 diastereomeric ratio. In the case of substrate 5 with a less hindered cyano group, cycloadduct 7 was quantitatively produced as a single isomer because the E isomer was not converted into an unstable Z isomer, even under conditions of progressive isomerization. When cyclopenta-1,3-diene 8 was subjected to the reaction with 1a, cycloadduct 9 was obtained in a 81% yield as a mixture of four diastereomers.
So far, nitrocinnamate 1 was found to serve as an excellent dienophile that could efficiently undergo the Diels–Alder reaction, which led to multiple substituted cyclohexene 3 products. Our next challenge was focused on the synthesis of more densely functionalized cyclohexenes. For this purpose, 1-methoxy-3-trimethylsiloxy-1,3-butadiene (Danishefsky’s diene, CAS No. 54125-02-9) 10 was employed because its silyl enol ether moiety can be converted into a carbonyl group upon desilylation and two electron-donating groups can improve its reactivity. The Diels–Alder reaction of 1a with diene 10 efficiently proceeded without heating to yield cycloadduct 11 regardless of the Z/E ratio of 1a (Scheme 4). The diastereomeric ratio of 11 was consistent with the Z/E ratio of 1a, and close to a single isomer (95/5) was obtained when only the Z form of 1a was used. Cycloadduct 11 easily underwent desilylation to afford cyclohexanone 12 in air or during treatment with chromatography on silica gel. The formation of 12 was confirmed by NMR and IR measurements. In the 1H NMR spectra, a doublet at 5.2 ppm assigned to an alkenyl proton of 11 disappeared, and the absorption of a carbonyl group newly appeared in the IR spectrum.
Since cyclohexanone 12 is an oil, it was converted into solid hydrazone 13 to determine its stereochemistry with single X-ray crystallography. A single diastereomer of cyclohexanone 12 obtained from the Z isomer of 1a was reacted with 2,4-dinitrophylhydrazine in ethanol in the presence of hydrochloric acid, which afforded hydrazone 13 without any stereochemical isomerization. During the reaction, methanol was eliminated. The recrystallization of 13 from toluene–hexane successfully yielded orange needles as a single crystal, and X-ray crystallography revealed that the nitro group and 4-methylphenyl group were substituted in the cis form, which was the same relationship as the Z form in 1a (Figure 1, Supplementary Materials). These results confirmed that the Diels–Alder reaction concertedly proceeded.

3. Conclusions

A systematic study of the Diels–Alder reactions of α-nitrocinnamate 1 was performed. Cinnamate 1 efficiently reacted with 2,3-dimethyl-1,3-butadiene 2 to afford cyclohexene 3 in high yields. This reaction was not influenced by the p-substituent of the phenyl group. Cycloadduct 3 was obtained as a mixture of stereoisomers, which was found to be due to E/Z isomerization at high temperatures. In the case of the more electron-rich Danishefsly’s diene 10, cycloaddition efficiently proceeded under mild conditions, which furnished highly functionalized cyclohexene 11. Multiple functionalized cyclohexanones 12 were obtained via the subsequent desilylation of 11. The insights obtained in this study are surely valuable in the field of organic synthesis.

4. Experimental Section

All reagents were purchased from commercial sources and used without further purification. 1H and 13C NMR spectra were recorded with Bruker DPX-400 and JEOL JMN-ECZ400S spectrometers (400 MHz and 100 MHz, respectively) using TMS as an internal standard. The assignments of the 13C NMR spectra were performed with DEPT experiments. IR spectra were recorded with a JASCO FT/IR-4200 spectrometer equipped with an ATM detector. High-resolution mass spectra were obtained with an AB SCEIX Triplet TOF 4600 mass spectrometer. Microwave heating was performed with an Anton Paar Microwave 300 (850 W, 2455 MHz) and an Anton Paar Microwave 400 (850 W, 2450 MHz) using a 10 mL glass vessel. Diffraction data were collected at 93 K under a cold N2 gas stream with a Rigaku XtaLAB Synergy-S/Mo system (λ = 0.71073 Å (Mo-Kα)). The integrated data were analyzed by using a Yadokari-XG software package. The structures were solved with the ShelXT structure solution program using intrinsic phasing and refined with the ShelXL refinement package using least-squares minimization. Anisotropic refinement was performed for all non-hydrogen atoms, and all the hydrogen atoms were put in calculated positions.
Synthesis of α-nitrocinnamate 1a. α-Nitrocinnamate 1 was synthesized using a somewhat modified method previously described in the literature. Aniline (0.91 mL, 10.0 mmol) and magnesium sulfate (400 mg) were added to a solution of 4-methylbenzaldehyde (10.0 mmol) in THF (5 mL), and the mixture was stirred at room temperature for 5 h. After filtrations of magnesium sulfate, the filtrate was concentrated under reduced pressure to afford imine (2.0 g, 9.7 mmol, 97%) as a brown solid, which was used for the next step without further purification. A solution of imine (2.0 g, 9.7 mmol) and nitroacetate (1.08 mL, 9.7 mmol, CAS No. 626-35-7) in acetic anhydride (5 mL) was heated at 60 °C for 18 h. The reaction mixture was poured into water (100 mL) and extracted with dichloromethane (50 mL × 3). Organic layers were dried over magnesium sulfate, filtered, and concentrated under reduced pressure. The residue was treated with flash column chromatography on silica gel (hexane/ethyl acetate = 9/1) to afford α-nitrocinnamate 1a (1520 mg, 6.5 mmol, 65%) as a yellow oil. When the aldehyde could not be completely separated, distillation was performed to remove it. The recrystallization of the product using hexane/chloroform afforded (Z) isomer. Other cinnamates 1bf, 4, and 5 were synthesized in the same way.
Ethyl 3-(4-methylphenyl)-2-nitropropenoate (1a) [26]. Yellow plates. 1H NMR (400 MHz, CDCl3) δ 7.50 (s, 1H), 7.32 (d, J = 8.2 Hz, 2H), 7.22 (d, J = 8.2 Hz, 2H), 4.37 (q, J = 7.3 Hz, 2H), 2.39 (s, 3H) 1.36 (t, J = 7.3 Hz, 3H).
Ethyl 3-(4-methylphenyl)-2-nitropropenoate (1b) [27]. Brown plates. 1H NMR (400 MHz, CDCl3) δ 7.46 (s, 1H), 7.39 (d, J = 8.8 Hz, 2H), 6.92 (d, J = 8.2 Hz, 2H), 4.36 (q, J = 7.3 Hz, 2H), 3.85 (s, 3H), 1.36 (t, J = 7.3 Hz, 3H).
Ethyl 2-nitro-3-phenylpropenoate (1c) [26]. Yellow plates. 1H NMR (400 MHz, CDCl3) δ 7.59 (s, 1H), 7.40–7.51 (m, 5H), 4.39 (q, J = 6.9 Hz, 2H), 1.37 (t, J = 6.9 Hz, 3H).
Ethyl 3-(4-bromophenyl)-2-nitropropenoate (1d) [26]. Yellow plates. 1H NMR (400 MHz, CDCl3) δ 7.57 (d, J = 8.5 Hz, 2H), 7.47 (s, 1H), 7.28 (d, J = 8.5 Hz, 2H), 4.29 (q, J = 7.3 Hz, 2H), 1.37 (t, J = 7.3 Hz, 3H).
Ethyl 3-[(4-trifluoromethyl)phenyl]-2-nitropropenoate (1e) [26]. Yellow plates. 1H NMR (400 MHz, CDCl3) δ 7.69 (d, J = 8.3 Hz, 2H), 7.50 (s, 1H), 7.53 (d, J = 8.3 Hz, 2H), 4.05 (q, J = 7.0 Hz, 2H), 1.38 (t, J = 7.0 Hz, 3H).
Ethyl 2-nitro-3-(4-nitrophenyl)propenoate (1f) [27]. White solid. 1H NMR (400 MHz, CDCl3) δ 8.28 (d, J = 8.7 Hz, 2H), 7.60 (s, 1H), 7.59 (d, J = 8.7 Hz, 2H), 4.42 (q, J = 6.9 Hz, 2H), 1.39 (t, J = 6.9 Hz, 3H).
Ethyl 2-ethanoyl-3-(4-methylphenyl)propenoate (4) [28]. Colorless plates. 1H NMR (400 MHz, CDCl3) δ 7.63 (s, 1H), 7.29 (d, J = 8.3 Hz, 2H), 7.17 (d, J = 8.3 Hz, 2H), 4.29 (q, J = 7.0 Hz, 2H), 2.36 (s, 3H), 2.35 (s, 3H), 1.32 (t, J = 7.0 Hz, 3H).
Ethyl 2-cyano-3-(4-methylphenyl)propenoate (5) [29]. Colorless needles. 1H NMR (400 MHz, CDCl3) δ 8.23 (s, 1H), 7.91 (d, J = 8.0 Hz, 2H), 7.31 (d, J = 8.0 Hz, 2H), 3.93 (s, 3H), 2.44 (s, 3H).
Diels–Alder reaction of α-nitrocinnamate. 2,3-dimethyl-1,3-butadiene 2 (0.28 mL, 2.5 mmol) was added to a solution of α-nitrocinnamate 1a (117 mg, 0.5 mmol) in MeCN (1 mL), and the resultant solution was heated at 180 °C for 2 h under microwave irradiation. After the removal of the solvent under reduced pressure, the residue was subjected to flash column chromatography on silica gel (hexane/ethyl acetate = 95/5) to afford cycloadduct 3a (151 mg, 0.475 mmol, 95%) as a pale-yellow oil. When other substrates were used or conditions were changed, the reaction was conducted in the same way.
4-Ethoxycarbonyl-1,2-dimethyl-5-(4-methylphenyl)-4-nitrocyclohexene (3a). Pale-yellow oil (dr = 56/44). Major isomer: 1H NMR (400 MHz, CDCl3) δ 6.9–7.9 (m, 4H), 4.07 (q, J = 7.2 Hz, 2H), 4.05 (d, J = 8.0 Hz, 1H), 2.7–3.0 (m, 3H), 2.30 (s, 3H), 2.28 (br d, J = 18.8 Hz, 1H), 1.72 (s, 3H), 1.68 (s, 3H), 1.16 (t, J = 7.2 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 165.9 (C), 137.6 (C), 137.1 (C), 129.5 (CH), 128.4 (CH), 125.7 (C), 121.2 (C), 95.7 (C), 62.7 (CH2), 42.8 (CH), 36.2 (CH2), 34.0 (CH2), 21.2 (CH3), 19.2 (CH3), 18.6 (CH3), 13.8 (CH3). Minor isomer: 1H NMR (400 MHz, CDCl3) δ 6.9–7.9 (m, 4H), 4.24 (dq, J = 7.2, 14.4 Hz, 1H), 4.21 (dq, J = 7.2, 14.4 Hz, 1H), 3.95 (d, J = 7.6 Hz, 1H), 2.7–3.0 (m, 3H), 2.39 (br d, J = 17.2 Hz, 1H), 2.30 (s, 3H), 1.74 (s, 3H), 1.69 (s, 3H), 1.22 (dd, J = 7.2, 7.2 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 166.2 (C), 137.6 (C), 136.5 (C), 129.4 (CH), 128.4 (CH), 126.4 (C), 121.2 (C), 95.3 (C), 62.8 (CH2), 43.8 (CH), 37.7 (CH2), 33.9 (CH2), 21.2 (CH3), 19.4 (CH3), 18.4 (CH3), 13.8 (CH3). IR (KBr/cm−1) 1753, 1553; HRMS (ESI/TOF) calculated for (M + H+) C18H24NO4: 318.1700, found: 318.1696.
4-Ethoxycarbonyl-5-(4-methoxyphenyl)-1,2-dimethyl-4-nitrocyclohexene (3b). Pale-yellow oil (dr = 70/30). Major isomer: 1H NMR (400 MHz, CDCl3) δ 7.07 (d, J = 8.8 Hz, 2H), 6.79 (d, J = 8.8 Hz, 2H), 4.11 (q, J = 7.2 Hz, 2H), 4.05 (d, J = 7.6 Hz, 1H), 3.77 (s, 3H), 2.7–3.1 (m, 3H), 2.27 (br d, J = 18.4 Hz, 1H), 1.72 (s, 3H), 1.67 (s, 3H), 1.56 (t, J = 7.2 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 165.9 (C), 159.2 (C), 132.1 (C), 129.6 (CH), 125.7 (C), 121.2 (C), 114.2 (CH), 95.6 (C), 62.7 (CH2), 55.3 (CH3), 42.4 (CH), 36.3 (CH2), 34.0 (CH2), 19.2 (CH3), 18.6 (CH3), 13.9 (CH3). Minor isomer: 1H NMR (400 MHz, CDCl3) δ 7.02 (d, J = 8.8 Hz, 2H), 6.79 (d, J = 8.8 Hz, 2H), 4.23 (dq, J = 7.2, 14.4 Hz, 1H),4.22 (dq, J = 7.2, 14.4 Hz, 1H), 3.93 (d, J = 7.2 Hz, 1H), 3.77 (s, 3H), 2.7–3.1 (m, 3H), 2.27 (br d, J = 17.2 Hz, 1H), 1.72 (s, 3H), 1.67 (s, 3H), 1.22 (dd, J = 7.2, 7.2 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 166.2 (C), 159.2 (C), 131.5 (C), 129.6 (CH), 125.7 (C), 121.2 (C), 114.0 (CH), 95.3 (C), 62.7 (CH2), 55.3 (CH3), 43.5 (CH), 37.8 (CH2), 34.0 (CH2), 19.4 (CH3), 18.4 (CH3), 13.9 (CH3); HRMS (ESI/TOF) calculated for (M + H+) C18H24NO5: 334.1649, found: 334.1648.
4-Ethoxycarbonyl-1,2-dimethyl-4-nitro-5-phenylcyclohexene (3c). Pale-yellow oil (dr = 62/38). Major isomer: 1H NMR (400 MHz, CDCl3) δ 7.28–7.10 (m, 5H), 4.08 (br d, J = 7.3 Hz, 1H), 4.05 (q, J = 7.3 Hz, 2H), 2.72–3.00 (m, 3H), 2.30 (br d, J = 18.3 Hz, 1H), 1.68 (s, 3H), 1.72 (s, 3H), 1.12 (t, J = 7.3 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 165.9 (C), 140.2 (C), 128.8 (CH), 128.5 (CH), 127.9 (CH), 125.7 (C), 121.3 (C), 95.6 (C), 62.7 (CH2), 43.1 (CH), 36.1 (CH2), 33.9 (CH2), 19.3 (CH3), 18.6 (CH3), 13.8 (CH3). Minor isomer: 1H NMR (400 MHz, CDCl3) δ 7.28–7.10 (m, 5H), 4.22 (dq, J = 7.3, 10.6 Hz, 1H), 4.21 (dq, J = 7.3, 10.6 Hz, 1H), 3.97 (br dd, J = 7.8, 2.3 Hz, 1H), 2.72–3.00 (m, 3H), 2.41 (br d, J = 17.4 Hz, 1H), 1.73 (s, 3H), 1.69 (s, 3H), 1.12 (dd, J = 7.3, 7.3 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 166.2 (C), 139.5 (C), 128.7 (CH), 128.5 (CH), 127.9 (CH), 126.4 (C), 121.2 (C), 95.2 (C), 62.9 (CH2), 44.2 (CH), 37.6 (CH2), 34.0 (CH2), 19.5 (CH3), 18.4 (CH3), 13.8 (CH3); HRMS (ESI/TOF) calculated for (M + H+) C17H22NO4: 304.1543, found: 304.1540.
5-(4-Bromophenyl)-4-ethoxycarbonyl-1,2-dimethyl-4-nitrocyclohexene (3d). Pale-yellow solid (dr = 52/48). Major isomer: 1H NMR (400 MHz, CDCl3) δ 7.38 (d, J = 8.4 Hz, 2H), 7.00 (d, J = 8.4 Hz, 2H), 4.23 (dq, J = 7.2, 10.8 Hz, 1H), 4.20 (dq, J = 7.2, 10.8 Hz, 1H), 3.94 (d, J = 7.2 Hz,1H), 2.7–3.1 (m, 3H), 2.36 (br d, J = 18.0 Hz, 1H), 1.72 (s, 3H), 1.68 (s, 3H), 1.21 (dd, J = 7.2, 7.2 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 165.9 (C), 138.5 (C), 131.8 (CH), 130.3 (CH), 126.1 (C), 122.0 (C), 121.4 (C), 94.9 (C), 63.0 (CH2), 43.6 (CH), 37.3 (CH2), 34.1 (CH2), 19.4 (CH3), 18.4 (CH3), 13.8 (CH3). Minor isomer: 1H NMR (400 MHz, CDCl3) δ 7.39 (d, J = 8.4 Hz, 2H), 7.04 (d, J = 8.4 Hz, 2H), 4.08 (q, J = 7.2 Hz, 2H), 4.03 (br d, J = 7.6 Hz, 1H), 2.7–3.1 (m, 3H), 2.26 (br d, J = 19.2 Hz, 1H), 1.72 (s, 3H), 1.67 (s, 3H), 1.56 (t, J = 7.2 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 165.6 (C) 139.1 (C), 131.9 (CH), 130.3 (CH), 125.5 (C), 122.0 (C), 121.4 (C), 95.2 (C), 62.8 (CH2), 42.8 (CH), 35.9 (CH2), 34.1 (CH2), 19.2 (CH3), 18.5 (CH3), 13.8 (CH3). IR (KBr/cm−1) 1752, 1554; HRMS (ESI/TOF) calculated for (M + H+) C17H20BrNO4: 382.0649, found: 382.0644.
4-Ethoxycarbonyl-5-[4-(trifluoromethyl)phenyl]-1,2-dimethyl-4-nitrocyclohexene (3e). Pale-yellow oil (dr = 62/38). Major isomer: 1H NMR (400 MHz, CDCl3) δ 7.54 (d, J = 8.4 Hz, 2H), 7.29 (d, J = 8.4 Hz, 2H), 4.15 (d, J = 7.2 Hz, 1H), 4.08 (q, J = 7.2 Hz, 2H), 2.6–3.1 (m, 3H), 2.31 (d, J = 18.4 Hz, 1H), 1.74 (s, 3H), 1.69 (s, 3H), 1.13 (t, J = 7.2 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 165.6 (C), 144.2 (C), 130.2 (C, q, J = 33 Hz), 129.1 (CH), 125.7 (CH, q, J = 4 Hz), 125.5 (C), 124.1 (C, q, J = 270 Hz), 121.6 (C), 95.1 (C), 62.9 (CH2), 43.2 (CH), 35.9 (CH2), 34.1 (CH2), 19.2 (CH3), 18.5 (CH3), 13.7 (CH3). Minor isomer: 1H NMR (400 MHz, CDCl3) δ 7.32 (d, J = 7.6 Hz, 2H), 7.27 (d, J = 7.6 Hz, 2H), 4.25 (dq, J = 7.2, 10.8 Hz, 1H), 4.21 (dq, J = 7.2, 10.8 Hz, 1H), 4.04 (dd, J = 2.8, 7.2 Hz, 1H), 2.6–3.1 (m, 3H), 2.41 (d, J = 18.8 Hz, 1H), 1.74 (s, 3H), 1.69 (s, 3H), 1.21 (dd, J = 7.2, 7.2 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 165.8 (C), 144.2 (C), 130.2 (C, q, J = 32 Hz), 129.1 (CH), 126.1 (C), 125.6 (CH, q, J = 4 Hz), 124.1 (C, q, J = 271 Hz), 121.5 (C), 94.8 (C), 63.1 (CH2), 44.0 (CH), 37.2 (CH2), 33.9 (CH2), 19.4 (CH3), 18.4 (CH3), 13.8 (CH3); IR (KBr/cm−1) 1753, 1556, 1167, 1326; HRMS (ESI/TOF) calculated for (M + Na+) C18H19NO4F3Na: 394.1237, found: 339.1237.
4-Ethoxycarbonyl-1,2-dimethyl-4-nitro-5-[4-nitrophenyl]cyclohexene (3f). Pale-yellow oil (dr = 65/35). Major isomer: 1H NMR (400 MHz, CDCl3) δ 8.13 (d, J = 8.7 Hz, 2H), 7.37 (d, J = 8.7 Hz, 2H), 4.18 (d, J = 7.3 Hz, 1H), 4.10 (q, J = 7.3 Hz, 2H), 2.80–3.09 (m, 3H), 2.33 (d, J = 17.8 Hz, 1H), 1.75 (s, 3H), 1.70 (s, 3H), 1.22 (t, J = 7.3 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 165.3 (C), 147.6 (C), 147.4 (C), 129.7 (CH), 125.4 (C), 124.0 (CH), 121.7 (C), 94.8 (C), 63.1 (CH2), 43.3 (CH), 35.8 (CH2), 34.3 (CH2), 19.2 (CH3), 18.6 (CH3), 13.9 (CH3). Minor isomer: 1H NMR (400 MHz, CDCl3) δ 8.14 (d, J = 8.7 Hz, 2H), 7.32 (d, J = 8.7 Hz, 2H), 4.25 (dq, J = 7.3, 10.5 Hz, 1H), 4.22 (dq, J = 7.3, 10.5 Hz, 1H), 2.80–3.09 (m, 3H), 2.40 (d, J = 18.3 Hz, 1H), 1.76 (s, 3H), 1.71 (s, 3H), 1.22 (dd, J = 7.3, 7.3 Hz, 3H), One signal could not be observed presumably due to overlapping; 13C NMR (101 MHz, CDCl3) δ 165.6 (C), 147.6 (C), 146.8 (C), 129.7 (CH), 125.9 (C), 123.8 (CH), 121.6 (C), 94.6 (C), 63.3 (CH2), 43.9 (CH), 37.0 (CH2), 34.2 (CH2), 19.4 (CH3), 18.4 (CH3), 13.9 (CH3); HRMS (ESI/TOF) calculated for (M + Na+) C17H20N2O6Na: 371.1214, found: 371.1210.
4-Ethanoyl-4-ethoxycarbonyl-5-(4-methylphenyl)-1,2-dimethylcyclohexene (6). Pale-yellow oil (dr = 58/42). Major isomer: 1H NMR (400 MHz, CDCl3) δ 7.04–6.99 (m, 4H), 4.01 (dq, J = 7.3, 10.6 Hz, 1H), 4.06 (dq, J = 7.3, 10.6 Hz, 1H), 3.77 (br d, J = 7.3 Hz, 1H), 2.87–2.43 (m, 3H), 2.29 (s, 3H), 2.17–2.10 (m, 1H), 2.08 (s, 3H), 1.74 (s, 3H), 1.64 (s, 3H), 1.18 (dd, 7.3, 7.3 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 204.0 (C), 170.5 (C), 140.0 (C), 136.4 (C), 129.0 (CH), 128.5 (CH), 126.7 (C), 122.5 (C), 64.1 (C), 61.3 (CH2), 41.7 (CH), 36.4 (CH2), 32.4 (CH2), 26.5 (CH3), 21.1 (CH3), 19.6 (CH3), 18.7 (CH3), 14.1 (CH3). Minor isomer: 1H NMR (400 MHz, CDCl3) δ 7.04–6.99 (m, 4H), 4.18 (dq, J = 7.3, 10.6 Hz, 1H), 4.14 (dq, J = 7.3, 10.6 Hz, 1H), 3.77 (br d, J = 7.3 Hz, 1H), 2.87–2.43 (m, 3H), 2.28 (s, 3H), 2.17–2.10 (m, 1H), 1.95 (s, 3H), 1.74 (s, 3H), 1.65 (s, 3H), 1.19 (dd, 7.3, 7.3 Hz, 3H); 13C NMR (101 MHz, CDCl3) δ 205.4 (C), 171.7 (C), 139.3 (C), 136.5 (C), 129.1 (CH), 128.4 (CH), 125.6 (C), 123.3 (C), 64.2 (C), 61.4 (CH2), 42.7 (CH), 37.0 (CH2), 32.4 (CH2), 27.4 (CH3), 21.1 (CH3), 19.5 (CH3), 18.7 (CH3), 14.1 (CH3); HRMS (ESI/TOF) calculated for (M + Na+) C20H26O3Na: 337.1774, found: 337.1774.
4-Cyano-4-methoxycarbonyl-1,2-dimethyl-5-(4-methylphenyl)cyclohexene (7). Pale-yellow oil (dr = 100/0). 1H NMR (400 MHz, CDCl3) δ 7.25 (d, J = 7.6 Hz, 2H), 7.12 (d, J = 7.6 Hz, 2H), 3.49 (s, 3H), 3.22 (dd, J = 7.2, 12.0 Hz, 1H), 2.92 (br d, J = 16.8 Hz, 1H), 2.73 (br dd, J = 17.6, 12.0 Hz, 1H), 2.46 (d, J = 16.8 Hz, 1H), 2.32 (s, 3H), 2.23 (dd, J = 7.2, 17.6 Hz, 1H), 1.71 (br s, 6H); 13C NMR (101 MHz, CDCl3) δ 169.4 (C), 137.7 (C), 136.1 (C), 129.4 (CH), 127.9 (CH), 126.5 (C), 121.0 (C), 118.7 (C), 53.2 (CH3), 50.6 (C), 45.9 (CH), 41.0 (CH2), 36.3 (CH2), 21.2 (CH3), 18.9 (CH3), 18.6 (CH3); IR (KBr/cm−1) 1740; HRMS (ESI/TOF) calculated for (M + H+) C18H22NO2: 284.1645, found: 284.1645.
5-Ethoxycarbonyl-6-(4-methylphenyl)-5-nitrobicyclo [2.2.1]hept-2-ene (9). Yellow oil (dr = 59/25/15/1). 1H NMR (400 MHz, CDCl3) Major isomer: δ 7.20 (d, J = 8.2 Hz, 2H), 7.09 (d, J = 8.2 Hz, 2H), 6.66 (dd, J = 3.2, 5.5 Hz, 1H), 6.07 (dd, J = 2.7, 5.5 Hz, 1H), 4.04 (d, J = 2.7 Hz, 1H), 3.76 (d, J = 2.7 Hz, 1H), 3.73 (dq, J = 7.3, 10.5 Hz, 1H), 3.61 (dq, J = 7.3, 10.5 Hz, 1H), 3.16 (br s, 1H), 2.72 (d, J = 9.6 Hz, 1H), 2.30 (s, 3H), 1.95 (d, J = 9.6 Hz, 1H), 0.82 (dd, J = 7.3, 7.3 Hz, 3H); Minor isomer 1: δ 7.02 (d, J = 8.2 Hz, 2H), 6.98 (d, J = 8.2 Hz, 2H), 6.66–6.69 (m, 1H), 6.49 (dd, J = 3.2, 5.5 Hz, 1H), 4.58 (d, J = 3.2 Hz, 1H), 4.30 (q, J = 6.9 Hz, 2H), 3.62–3.64 (m, 1H), 3.10–3.15 (m, 1H), 2.27 (s, 3H), 1.62–1.70 (m, 2H), 1.28 (t, J = 6.9 Hz, 3H); Minor isomer 2: δ 7.02–7.07 (m, 4H), 6.62–6.65 (m, 1H), 6.47 (dd, J = 2.7, 5.5 Hz, 1H), 4.74 (d, J = 3.2 Hz, 1H), 3.80–3.83 (m, 1H), 3.58–3.79 (m, 2H), 3.11–3.14 (m, 1H), 2.28 (s, 3H), 1.62–1.70 (m, 2H), 0.69 (t, J = 7.3 Hz, 3H); HRMS (ESI/TOF) calculated for (M + H+) C17H19NO4: 324.1206, found: 324.1196.
4-Ethoxycarbonyl-3-methoxy-5-(4-methylphenyl)-1-(trimethylsiloxy)-4-nitrocyclohexene (11). Pale-yellow oil (dr = 86/14). 1H NMR (400 MHz, CDCl3) δ 7.14 (d, J = 8.0 Hz, 2H), 7.08 (d, J = 8.0 Hz, 2H), 5.17 (d, J = 4.0 Hz, 1H), 4.54 (d, J = 4.0 Hz, 1H), 4.24 (dq, J = 7.2, 10.8 Hz, 1H), 4.16 (dq, J = 7.2, 10.8 Hz, 1H), 3.94 (dd, J = 6.4, 8.0 Hz, 1H), 3.41 (s, 3H), 2.38 (dd, J = 6.4, 18.0 Hz, 1H), 2.39 (dd, J = 8.0, 18.0 Hz, 1H), 2.32 (s, 3H), 1.23 (dd, J = 7.2, 7.2 Hz, 3H), 0.26 (s, 9H); 13C NMR (101 MHz, CDCl3) δ 164.5 (C), 153.5 (C), 137.7 (C), 135.3 (C), 129.4 (CH), 128.9 (CH), 100.3 (CH), 96.2 (C), 76.4 (CH), 62.3 (CH2), 57.4 (CH3), 42.7 (CH), 35.3 (CH2), 21.2 (CH3), 14.0 (CH3), 0.4 (CH3); IR (KBr/cm−1) 1766, 1733, 1549, 1244, 1220, 1084, 849; HRMS (ESI/TOF) calculated for (M + Na+) C20H29NO6SiNa: 430.1656, found: 430.1651.
4-Ethoxycarbonyl-3-methoxy-5-(4-methylphenyl)-4-nitrocyclohexanone (12). Pale-yellow oil (dr = 95/5). 1H NMR (400 MHz, CD3Cl) δ 7.06 (s, 4H), 4.49 (dd, J = 4.1, 4.1 Hz, 1H), 4.27 (dq, J = 7.3, 11.0 Hz, 1H), 4.21 (dd, J = 5.5, 11.4 Hz, 1H), 4.20 (dq, J = 7.3, 11.0 Hz, 1H), 3.39 (dd, J = 4.1, 15.6 Hz, 1H), 3.35 (3H, s), 2.90 (dd, J = 11.4, 15.6 Hz, 1H), 2.81 (ddd, J = 1.8, 4.1, 15.6 Hz, 1H), 2.72 (ddd, J = 1.8, 5.5, 15.6 Hz, 1H), 2.28 (s, 3H), 1.23 (dd, J = 7.3, 7.3 Hz, 3H); 13C NMR (101 MHz) δ 205.5 (C), 163.9 (C), 137.9 (C), 134.0 (C), 128.9 (CH), 128.6 (CH), 95.5 (C), 79.4 (CH), 62.6 (CH2), 57.6 (CH), 43.9 (CH3), 43.5 (CH2), 41.1 (CH2), 20.9 (CH3), 13.7 (CH3); IR (KBr/cm−1) 1766, 1729, 1548, 1243, 1083; HRMS (ESI/TOF) calculated for (M + H+) C17H22NO6: 336.1442, found: 336.1443.
3-Ethoxycarbonyl-4-(4-methylphenyl)-3-nitro-6-[(2,4dinitrophenyl)hydrazino)]cyclohexene (13). Orange needles, m.p. 110–111 °C. 1H NMR (400 MHz, CD3Cl) δ 11.24 (br s, 1H), 9.11 (d, J = 2.4 Hz, 1H), 8.37 (dd, J = 2.4, 9.6 Hz, 1H), 8.06 (d, J = 9.6 Hz, 1H), 7.05 (s, 4H), 6.83 (d, J = 9.2 Hz, 1H), 6.65 (dd, J = 0.9, 9.2 Hz, 1H), 4.42 (dq, J = 7.3, 10.1 Hz, 1H), 4.42 (ddd, J = 0.9, 2.7, 6.9 Hz, 1H), 4.30 (dq, J = 7.3, 10.1 Hz, 1H), 3.42 (dd, J = 6.9, 17.4 Hz, 1H), 3.04 (dd, J = 2.7, 17.4 Hz, 1H), 2.27 (s, 3H), 1.29 (dd, J = 7.3, 7.3 Hz, 3H); 13C NMR (101 MHz, CD3Cl) δ 164.8 (C), 148.1 (C), 144.4 (C), 139.2(C), 138.7 (C), 134.4 (C), 132.1 (CH), 130.5 (C), 130.3 (CH), 129.8 (CH), 128.3 (CH), 127.1 (CH), 123.4 (CH), 116.8 (CH), 93.8 (C), 63.9 (CH2), 43.3 (CH), 28.9 (CH2), 21.2 (CH3), 13.9 (CH3); IR (KBr/cm−1) 1750, 1616, 1337, 771; HRMS (ESI/TOF) calculated for (M + H+) C22H22NO2: 484.1463, found: 484.1453.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/reactions3040041/s1. Copies of the NMR and HRMS spectra of compounds 1, 3, 6, 7, 9, and 1113, as well as the crystallographic data of 13.

Author Contributions

T.H. wrote the original draft and conducted experiments. K.I. performed X-ray analysis. K.I. and N.N. reviewed and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barrett, G.A.; Graboski, G.G. Conjugated nitroalkenes: Versatile intermediates in organic synthesis. Chem. Rev. 1986, 86, 751–762. [Google Scholar] [CrossRef]
  2. Berner, M.O.; Tedeschi, L.; Enders, D. Asymmetric Michael additions to nitroalkenes. Eur. J. Org. Chem. 2002, 2002, 1877–1894. [Google Scholar] [CrossRef]
  3. Denmark, E.S.; Thorarensen, A. Tandem [4+2]/[3+2] cycloadditions of nitroalkenes. Chem. Rev. 1996, 96, 137–166. [Google Scholar] [CrossRef] [PubMed]
  4. Halimehjani, Z.A.; Namboothiri, N.I.; Hooshmand, E.S. Nitroalkenes in the synthesis of carbocyclic compounds. RSC Adv. 2014, 4, 31261–31299. [Google Scholar] [CrossRef]
  5. Parambil, P.S.; Veettil, P.S.; Dehaen, W. The synthesis of five-membered N-heterocycles by cycloaddition of nitroalkenes with (in)organic azides and other 1,3-dipoles. Synthesis 2022, 54, 910–924. [Google Scholar] [CrossRef]
  6. Ballini, R.; Marcantoni, E.; Petrini, M. Nitroalkenes as Amination Tools. In Amino Group Chemistry: From Synthesis to the Life Sciences; Ricci, A., Ed.; Wiley-VCH: Weinheim, Germany, 2008; pp. 93–148. [Google Scholar] [CrossRef]
  7. Marčeková, M.; Ferko, B.; Detková, K.R.; Jakubec, P. Denitrative cross-couplings of nitrostyrenes. Molecules 2020, 25, 3390. [Google Scholar] [CrossRef]
  8. Ballini, R.; Petrini, M. The nitro to carbonyl conversion (Nef reaction): New perspectives for a classical transformation. Adv. Synth. Catal. 2015, 357, 2371–2402. [Google Scholar] [CrossRef]
  9. Nair, K.D.; Kumar, T.; Namboothiri, N.N.I. α-Functionalization of nitroalkenes and its applications in organic synthesis. Synlett 2016, 27, 2425–2442. [Google Scholar] [CrossRef]
  10. Bartoli, G.; Bosco, M.; Giuli, S.; Giuliani, A.; Lucarelli, L.; Marcantoni, E.; Sambri, L.; Torregiani, E. Efficient preparation of 2-indolyl-1-nitroalkane derivatives employing nitroalkenes as versatile Michael acceptors:  new practical linear approach to alkyl 9H-β-Carboline-4-carboxylate. J. Org. Chem. 2005, 70, 1941–1944. [Google Scholar] [CrossRef]
  11. Itoh, K.; Kishimoto, S. The reaction of β-nitrostyrenes with 2-methoxyfuran: A novel formation of isoxazoline N-oxide together with Michael adducts N. J. Chem. 2000, 24, 347–349. [Google Scholar] [CrossRef]
  12. Umezawa, S.; Kinoshita, M.; Yanagisawa, H. The synthesis of cyclic α-amino acids. I. Bull. Chem. Soc. Jpn. 1967, 40, 209–214. [Google Scholar] [CrossRef] [Green Version]
  13. Fairhurst, J.; Horwell, C.D.; Timms, H.G. Novel bicyclo[2.2.2]octanyl-1,4-benzodiazepinones. Their syntheses and rearrangement to bicyclooct-2-enylbenzimidazoles. J. Heterocycl. Chem. 1977, 14, 1199–1201. [Google Scholar] [CrossRef]
  14. Boger, D.L.; Lerner, R.A.; Cravatt, B.F. Synthesis of a functionalized rigid bicyclo[2.2.1]heptane: A useful hapten for eliciting catalytic antibodies. J. Org. Chem. 1994, 59, 5078–5079. [Google Scholar] [CrossRef]
  15. Baichurin, R.I.; Aboskalova, N.I.; Berkova, G.A.; Berestovitskaya, V.M. Synthesis of aryl- and heteryl- containing gem-acylnitrocyclohexenes. Russ. J. Org. Chem. 2009, 45, 1185–1188. [Google Scholar] [CrossRef]
  16. Ruffoni, A.; Casoni, A.; Pellegrino, S.; Gelmi, L.M.; Soave, M.; Clerici, F. Sulfanyl-methylene-5(4H)-oxazolones and β-sulfanyl-α-nitroacrylates as appealing dienophiles for the synthesis of conformationally constrained cysteine analogues. Tetrahedron 2012, 68, 1951–1962. [Google Scholar] [CrossRef]
  17. Jasiński, R.; Kwiatkowska, M.; Sharnin, V.; Barański, A. Experimental and theoretical studies of Diels–Alder reaction between methyl (Z)-2-nitro-3-(4-nitrophenyl)-2-propenoate and cyclopentadiene. Monatsh. Chem. 2013, 144, 327–335. [Google Scholar] [CrossRef] [Green Version]
  18. Jaykanthan, K.; Vankar, Y.D. Synthesis of conformationally constrained C-glycosyl α-and β-amino acids and sugar-carbamino sugar hybrids via Diels-Alder reaction. Org. Lett. 2005, 7, 5441–5444. [Google Scholar] [CrossRef]
  19. Lago-Santomé, H.; Martínez-Bescos, P.; Fernández-González, M.; Ozores-Viturro, L.; Cagide-Fagín, F.; Alonso, R. β-Aryl-α-nitro-α,β-enals as heterodienes and dienophiles. J. Org. Chem. 2014, 79, 8645–8651. [Google Scholar] [CrossRef]
  20. Lanari, D.; Ballini, R.; Palmieri, A.; Pizzo, F.; Vaccaro, L. Diastereoselective three-step route to o-(6-nitrocyclohex-3-en-1-yl)phenol and tetrahydro-6H-benzo[c]chromen-6-ol derivatives from salicylaldehydes. Eur. J. Org. Chem. 2011, 15, 2874–2884. [Google Scholar] [CrossRef]
  21. Nakao, R.; Fujii, Y.; Hayakawa, I.; Mizoguchi, H. Kinetic resolution of α-nitrolactones by catalytic asymmetric hydrolysis or ester–amide exchange reaction. Synlett 2020, 31, 2018–2022. [Google Scholar] [CrossRef]
  22. Fujita, R.; Watanabe, K.; Yoshisuji, T.; Kabuto, C.; Matsuzaki, H.; Hongo, H. Diels-Alder cycloadditions of 2(1H)-quinolones having an electron-withdrawing group at the 3-position acting as dienophiles with dienes. Chem. Pharm. Bull. 2001, 49, 893–899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Asahara, M.; Nagamatsu, M.; Tohda, Y.; Nishiwaki, N.; Ariga, M. Diels-Alder reaction of 1-methyl-3,6,8-trinitro-2-quinolone. J. Heterocycl. Chem. 2004, 41, 803–805. [Google Scholar] [CrossRef]
  24. Fujita, R.; Yoshisuji, T.; Wakayanagi, S.; Wakamatsu, H. Synthesis of 5(6H)-phenanthridones using Diels–Alder reaction of 3-nitro-2(1H)-quinolones acting as dienophiles. Chem. Pharm. Bull. 2006, 54, 204–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Piaz, V.D.; Giovannoni, M.P.; Ciciani, G.; Giomi, D.; Nesi, R. Diels-Alder reactions on 5-acetyl-2-methyl-4-nitro-6-phenylpyridazin-3(2H)-one: A new facet of the pyridazine system. Tetrahedron Lett. 1993, 34, 161–162. [Google Scholar] [CrossRef]
  26. Liu, J.; Gong, L.; Meggers, E. Asymmetric Friedel–Crafts alkylation of indoles with 2-nitro-3-arylacrylates catalyzed by a metal-templated hydrogen bonding catalyst. Tetrahedron Lett. 2015, 56, 4653–4656. [Google Scholar] [CrossRef]
  27. Baichurin, R.I.; Baichurina, L.V.; Aboskalva, N.I.; Berestovitskaya, V.M. Synthesis and structure of β-aryl-α-nitroacrylates. Russ. J. Gen. Chem. 2013, 83, 1764–1770. [Google Scholar] [CrossRef]
  28. Li, Z.-N.; Chen, X.-L.; Fu, Y.-J.; Wang, W.; Luo, M. A facile synthesis of trisubstituted alkenes from β-diketones and aldehydes with AlCl3 as catalyst. Res. Chem. Intermed. 2012, 38, 25–35. [Google Scholar] [CrossRef]
  29. Tukhtaev, H.B.; Ivanov, K.L.; Bezzubov, S.I.; Cheshkov, D.A.; Melnikov, M.Y.; Budynina, E.M. aza-Wittig reaction with nitriles: How carbonyl function switches from reacting to activating. Org. Lett. 2019, 21, 1087–1092. [Google Scholar] [CrossRef]
Scheme 1. Diels–Alder reactions of ethyl α-nitrocinnamates possessing a para-substituent.
Scheme 1. Diels–Alder reactions of ethyl α-nitrocinnamates possessing a para-substituent.
Reactions 03 00041 sch001
Scheme 2. Isomerization of 1a at 180 °C.
Scheme 2. Isomerization of 1a at 180 °C.
Reactions 03 00041 sch002
Scheme 3. Diels–Alder reactions using other dienophiles and diene.
Scheme 3. Diels–Alder reactions using other dienophiles and diene.
Reactions 03 00041 sch003
Scheme 4. Diels–Alder reaction of ethyl α-nitrocinnamate 1a with Danishefsky’s diene 10.
Scheme 4. Diels–Alder reaction of ethyl α-nitrocinnamate 1a with Danishefsky’s diene 10.
Reactions 03 00041 sch004
Figure 1. The crystal structure of 13 with ORTEP drawing (50% probability). Atoms are colored as follows: carbon, gray; nitrogen, blue; oxygen, red. Triclinic, P-1, Z = 2, GOF = 1.062, R1[I > 2σ(I)] = 0.0369, wR2 (all data) = 0.1055.
Figure 1. The crystal structure of 13 with ORTEP drawing (50% probability). Atoms are colored as follows: carbon, gray; nitrogen, blue; oxygen, red. Triclinic, P-1, Z = 2, GOF = 1.062, R1[I > 2σ(I)] = 0.0369, wR2 (all data) = 0.1055.
Reactions 03 00041 g001
Table 1. Optimal conditions of the Diels–Alder reaction with ethyl α-nitrocinnamate 1a and 2,3-dimethyl-1,3-butadiene (2).
Table 1. Optimal conditions of the Diels–Alder reaction with ethyl α-nitrocinnamate 1a and 2,3-dimethyl-1,3-butadiene (2).
Reactions 03 00041 i001
EntrySolv.Temp. (°C)Time (h)Yield of 3a (%) 1,2Recovery of 1a (%) 1
1toluene1501.513(13) 386
2THF1501.52961
3CHCl31501.55740
4MeCN1501.56732
5MeCN1506.0960
6MeCN1802.0quant.0
1 Determined by 1H NMR using internal standard (1,1,2,2-tetrachloroethane); 2 total yield of diastereoisomers; 3 isolated yield.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hamada, T.; Iwai, K.; Nishiwaki, N. Synthesis and Characterization of Multiple Functionalized Cyclohexanone Using Diels–Alder Reaction of α-Nitrocinnamate. Reactions 2022, 3, 615-624. https://doi.org/10.3390/reactions3040041

AMA Style

Hamada T, Iwai K, Nishiwaki N. Synthesis and Characterization of Multiple Functionalized Cyclohexanone Using Diels–Alder Reaction of α-Nitrocinnamate. Reactions. 2022; 3(4):615-624. https://doi.org/10.3390/reactions3040041

Chicago/Turabian Style

Hamada, Takumi, Kento Iwai, and Nagatoshi Nishiwaki. 2022. "Synthesis and Characterization of Multiple Functionalized Cyclohexanone Using Diels–Alder Reaction of α-Nitrocinnamate" Reactions 3, no. 4: 615-624. https://doi.org/10.3390/reactions3040041

Article Metrics

Back to TopTop