Next Article in Journal
Ectopic Overexpression of a Novel R2R3-MYB, NtMYB2 from Chinese Narcissus Represses Anthocyanin Biosynthesis in Tobacco
Next Article in Special Issue
Indium-Catalyzed Annulation of o-Acylanilines with Alkoxyheteroarenes: Synthesis of Heteroaryl[b]quinolines and Subsequent Transformation to Cryptolepine Derivatives
Previous Article in Journal
Computational Modeling of In Vitro Swelling of Mitochondria: A Biophysical Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Catalytic Annulation of Epoxides with Heterocumulenes by the Indium-Tin System

Research Center for Environmental Preservation, Osaka University, Osaka 565-0871, Japan
*
Author to whom correspondence should be addressed.
Submission received: 15 March 2018 / Revised: 27 March 2018 / Accepted: 28 March 2018 / Published: 28 March 2018
(This article belongs to the Special Issue Indium in Organic Synthesis)

Abstract

:
In the synthesis of five-membered heterocycles by the annulation of epoxides with heterocumulenes such as carbon dioxide and isocyanates, we developed the indium-tin catalytic system and synthesized various cyclic adducts including novel types products under mild reaction conditions.

Graphical Abstract

1. Introduction

Synthesis of five-membered heterocycles by the annulation of epoxides with heterocumulenes such as carbon dioxide and isocyanates has been intensively studied [1,2,3]. For the use of carbon dioxide, its catalytic transformation to useful organic compounds has attracted much attention [4,5,6,7,8]. In particular, fixation of carbon dioxide to cyclic carbonates is one of the most important processes of high atom efficiency [9]. Cyclic carbonates are known to be efficient aprotic polar solvents [10,11,12], electrolytes in lithium ion batteries [13,14] and materials for producing polycarbonates [15]. Instead of the products from carbon dioxide, the cyclic adducts of epoxides with isocyanates such as 2-oxazolidinones [16,17] are important biologically-active compounds [18,19,20,21,22] and synthetic intermediates such as precursors of amino alcohols and chiral auxiliaries [23,24,25,26,27,28,29,30,31]. We present here a simple, easily available catalyst, the indium-tin system, and provide the environmentally-benign process to annulated adducts under mild conditions.

2. Results

2.1. Synthesis of Cyclic Carbonates

Until now, various catalysts such as transition metal compounds, ionic liquid, onium salts and alkali metal salts, etc., have been developed for the reaction of epoxides with carbon dioxide [5,32,33]. However, these methods suffer from either the need for co-solvent, the requirement for high temperature, high CO2 pressure or expensive catalyst. Indium reagents and catalysts have been applied in modern organic synthesis by their mildness and easy handling character [34]. By using the indium halide-phosphine complex, we have already developed the reaction of terminal epoxides to give cyclic carbonates under atmospheric CO2 pressure at room temperature [35], which indicated that indium halide-based catalysts have efficient catalytic activities. As shown in Table 1, we screened indium halide catalytic systems in the reaction of epoxide 1a with CO2 (3.9 MPa) at room temperature. The sole use of InCl3 did not have catalytic activity (Entry 1). Interestingly, the combination of Bu2SnI2 with InCl3 increased the yield of carbonate 2a (Entry 2). The sole use of Bu2SnI2 was not effective at all (Entry 3). Thus, the InCl3-Bu2SnI2 system showed a high catalytic activity. Of particular interest is that the reaction proceeded well even at room temperature. The choice of acetonitrile as a solvent is essential because no reaction proceeded when other solvents such as hexane, benzene, CHCl3 and THF were used. In acetonitrile, the reaction proceeded very well, and various cyclic carbonates 2 were obtained from epoxides 1b1f catalyzed by InCl3-Bu2SnI2. Epoxides having aliphatic and aromatic substituents were reactive to afford the corresponding cyclic carbonates 2b2c (Entries 4 and 5). High chemoselectivities were observed because of the mild conditions. Functionalized cyclic carbonates 2d2f were synthesized from epoxides having halogen and oxygen substitutes (Entries 6–8).

2.2. Synthesis of 2-Oxazolidinones

Instead of using carbon dioxide, the cyclic adducts of epoxides with isocyanates are highly in demand [18,19,20,21,22,23,24,25,26,27,28,29,30,31]. To effect the reaction, various catalysts have been used such as lithium halides [36,37,38,39,40,41,42], quaternary ammonium salts [43,44], phosphonium salt [45], AlCl3 [46], magnesium halides [47], tetraphenylantimony iodide [48,49,50] and the chromium(Salphen) complex [51]. These catalysts promoting reactions require relatively severe conditions (over 100 °C). The Pd-catalyzed reaction enabled mild conditions and accomplished an asymmetric reaction; however, in the reaction, only vinyl-substituted epoxides could be applicable where π-allyl palladium intermediates should be generated [52]. We have already reported that the Bu3SnI-Ph3PO or Ph4SbI catalyzes the annulation of epoxides with aromatic isocyanates to give 3,5-disubstututed-2-oxazolidinones [53,54,55,56], which indicated that tin halide-based catalysts would afford efficient catalytic activity. In view of these backgrounds, we tested the catalytic activity of the Bu2SnI2-InCl3 system as shown in Table 2. In the reaction of epoxybutane 1b with tert-BuN=C=O (3a), a quantitative yield of 2-oxazolidinone 4a was obtained (Entry 1). Interestingly, steric hindrance of isocyanates was not a problem to give 2-oxazolidinones 4. The use of either Bu2SnI2 or InCl3 was not effective (Entries 2 and 3). Thus, it was clear that the InCl3-Bu2SnI2 catalytic system showed a high activity even for the synthesis of 2-oxazolidinones 4. Other epoxides such as epichlorohydrin 1d and glycidylic ethers 1e, 1f also reacted well (Entries 4–6). Of course, primary aliphatic isocyanate 3b and phenyl isocyanate 3c also gave the desired products 4e and 4f, although the yields were moderate owing to the trimerization of an isocyanate as a side reaction (Entries 7 and 8) [51,57]. Thus, higher yields of 4ad in the reaction of tert-BuN=C=O (3a) were achieved because the trimerization would be depressed by steric hindrance of 3a. In all cases, regioselective ring opening of epoxides took place at the less substituted site to give 3,5-disubstituted-2-oxazolidinones 4.
In the reaction with diphenyl carbodiimide, an analogue of isocyanates, the oxazolidin-2-imine 5, was obtained in good yield (Scheme 1).

3. Discussion

As shown in Figure 1, the structure of the Bu2SnI2-InCl3 system could be supposed by the measurement of 119Sn-NMR spectra in acetonitrile. The addition of equimolar InCl3 to Bu2SnI2 in acetonitrile clearly changed the 119Sn peak from a strong one at −38 ppm to a broad one at 3 ppm. This downfield shift indicates that tin species had a positive character by the combination with InCl3 [58,59].
The catalytic cycle is explained as shown in Scheme 2. By the interaction of tin and indium species, Lewis acidic tin species like Bu2SnIδ+[InCl3I]δ would be generated, which activate the epoxide ring [60,61,62,63]. This active bimetallic species is plausible because transmetallation between tin and indium reagents easily takes place [64,65,66,67,68]. The ring opening of an epoxide to A proceeds regioselectively at the less substituted carbon. Next, the tin-oxygen bond in A is added to heterocumulene to give an adduct B. In the case of isocyanates, the addition occurs at the C=N group selectively to give a stannylcarbamate B [69,70,71]. At the last stage, the Sn-X (X=O, NR) bond in the intermediate B attacks the terminal alkyl iodide [72,73] to afford cyclic carbonates 2 and 2-oxazolidinones 4 and regenerate the catalyst. The 1H-NMR of products 4 showed single regio isomers because of the regioselective ring opening of epoxides. For example, 5-H and 4-H peaks for 4f were found at 4.59 (1H), 4.08–3.66 (2H) ppm, respectively. On the other hand, it has been reported by us that 1H-NMR for another regio isomer showed its 5-H and 4-H peaks at 4.65–4.30 (2H), 4.22–4.04 (1H) ppm, respectively [49].

4. Materials and Methods

4.1. Analysis

FTIR spectra were recorded as a thin film on a Nicolet IS5 spectrometer (Thermo Electron Scientific Instruments LLC, Madison, WI, USA). All 1H and 13C-NMR spectra were recorded with a JEOL JMTC-400/54/SS (400 and 100 MHz, respectively) in deuteriochloroform (CDCl3) containing 0.03% (w/v) of tetramethylsilane as an internal standard. Temperatures shown in schemes or tables were controlled by a constant-temperature oil bath. Yields were determined by 1H-NMR using 1,1,1,2-tetrachloroethane or 1,1,2,2-tetrachloroethane as an internal standard. Mass spectra were recorded on a JEOL JMS-DS-303 spectrometer (JEOL Ltd., Tokyo, Japan). Flash column chromatography was performed by Yamazen YFLC-AI-580 using Hi-Flash Silica gel 2L Hi-Flash Column 20 mL/min eluted by hexane/EtOAc with the gradation mode changing from 9/1–3/7 depending on Rf values of each compound. Bulb-to-bulb distillation (Kugelrohr) was accomplished at the oven temperature and pressure indicated.
Dehydrated acetonitrile (MeCN) was purchased from commercial sources and used as obtained. Deuterated acetonitrile was also purchased and stored drying over 4 Å molecular sieves. All epoxides, isocyanates, carbodiimide and InCl3 were also purchased and used as obtained. Bu2SnI2 was prepared according to the previous report [74].

4.2. General Procedure for Synthesis of Cyclic Carbonates 2af from Epoxides 1 with CO2

To a 50-mL autoclave, InCl3 (0.5 mmol), Bu2SnI2 (1.0 mmol) and epoxide 1 (10 mmol) were added in MeCN (3 mL). The autoclave was flushed with CO2 (3.9 MPa) and stirred at room temperature for 5–10 h. After release of the CO2 gas, the reaction mixture was quenched with H2O (20 mL) and extracted with Et2O (3 × 20 mL). The collected organic layer was dried over MgSO4. After filtration, the mixture was concentrated in vacuo. The residue was purified by column chromatography. Further purification was performed by Kugelrohr distillation to give a pure product 2.
4-Methyl-1.3-dioxolan-2-one (2a). Colorless liquid. The NMR data of 1H and 13C agreed with the previous report. [75]. 1H-NMR: (270 MHz, CDCl3) δ 4.70 (ddd, J = 6.4, 7.3, 7.8 Hz, 1H, 4-H), 4.38 (dd, J = 7.8, 8.3 Hz, 1H, 5HH), 3.84 (dd, J = 7.3, 8.3 Hz, 1H, 5HH), 1.29 (d, J = 6.34 Hz, 3H, 4-CH3). 13C-NMR: (67.5 MHz, CDCl3) δ 155.00 (C-2), 73.55 (C-4), 70.58 (C-5), 19.18 (4-CH3).
4-Ethyl-1.3-dioxolan-2-one (2b). Colorless liquid. The NMR data of 1H and 13C agreed with the previous report [76]. 1H-NMR: (270 MHz, CDCl3) δ 4.70–4.65 (m, 1H, 4-H), 4.56 (dd, J = 8.3, 8.3 Hz, 1H, 5HH), 4.12 (dd, J = 7.8, 7.8 Hz, 1H, 5HH) 1.88–1.68 (m, 2H, 4-CH2CH3), 1.03 (t, J = 7.3 Hz, 3H, 4-CH2CH3). 13C NMR: (67.5 MHz, CDCl3) δ 154.91 (C-2), 77.88 (C-4), 68.83 (C-5), 26.59 (4-CH2CH3), 8.20 (4-CH2CH3).
4-Phenyl-1.3-dioxolan-2-one (2c). Colorless liquid. The NMR data of 1H and 13C agreed with the previous report [75]. 1H-NMR: (270 MHz, CDCl3) δ 7.44–7.18 (m, 5H, Ph), 5.66 (t, J = 8.1 Hz, 1H, 4-H) 4.78 (dd, J = 8.3, 8.3 Hz, 1H, 5HH), 4.33 (dd, J = 7.8, 7.8 Hz, 1H, 5HH). 13C-NMR: (67.5 MHz, CDCl3) δ 154.71 (C-2), 135.67 (i), 129.62 (m), 129.12 (p), 125.77 (o), 77.92 (C-4), 71.10 (C-5).
4-(Chloromethyl)-1,3-dioxolan-2-one (2d). Colorless liquid. The NMR data of 1H and 13C agreed with the previous report [75]. 1H-NMR: (270 MHz, CDCl3) δ 5.10–5.02 (m, 1H, 4-H), 4.63 (dd, J = 8.8, 8.8 Hz, 1H, 5HH), 4.41 (dd, J = 5.8, 8.8 Hz, 1H, 5HH), 3.88 (dd, J = 4.4, 12.2 Hz, 1H, CHHCl), 3.76 (dd, J = 3.4, 12.2 Hz, 1H, CHHCl). 13C-NMR: (67.5 MHz, CDCl3) δ 154.33 (C-2), 74.33 (C-4), 66.70 (C-5), 44.01 (CH2Cl).
4-(Phenoxymethyl)-1,3-dioxolan-2-one (2e). Colorless liquid. The NMR data of 1H and 13C agreed with the previous report [76]. 1H-NMR: (270 MHz, CDCl3) δ 7.35–6.87 (m, 5H, Ph), 5.08–4.99 (m, 1H, 4-H), 4.62 (dd, J = 8.3, 8.8 Hz, 1H, CHHOPh), 4.54 (dd, J = 6.4, 8.8 Hz, 1H, CHHOPh), 4.25 (dd, J = 3.9, 10.3 Hz, 1H, 5HH), 4.15 (dd, J = 3.9, 10.3 Hz, 1H, 5HH). 13C-NMR: (67.5 MHz, CDCl3) δ 157.72 (o), 154.67 (C-2), 129.55 (p), 121.98 (m), 114.59 (m), 74.10 (CH2OPh), 66.87 (C-4), 66.23 (C-5).
4-(Methoxymethyl)-1,3-dioxolan-2-one (2f). Colorless liquid. The NMR data of 1H and 13C agreed with the previous report [35]. 1H-NMR: (270 MHz, CDCl3) δ 4.88–4.80 (m, 1H, 4-H), 4.51 (dd, J = 8.3, 8.8 Hz, 1H, 5HH), 4.37 (dd, J = 6.3, 8.3 Hz, 1H, 5HH), 3.67 (dd, J = 3.4, 11.2 Hz, 1H, CHHOCH3), 3.54 (dd, J = 3.9, 11.2 Hz, 1H, CHHOCH3), 3.43 (s, 3H, OCH3). 13C-NMR: (67.5 MHz, CDCl3) δ 154.91 (C-2), 75.01 (C-4), 71.27 (CH2OCH3), 65.97 (C-5), 59.34 (OCH3).

4.3. General Procedure for Synthesis of 2-Oxazolidinones 4 from Epoxides 1bf with Isocyanates 3 and Oxazolidin-2-imine 5

To a two-neck 10-mL reaction vessel, InCl3 (0.25 mmol), Bu2SnI2 (0.50 mmol), epoxide 1 (5 mmol) and isocyanate 3 (5.5 mmol) were added in MeCN (1.5 mL) under N2 atmosphere. The reaction mixture was stirred at room temperature or 60 °C for 3–20 h. After completion of the reaction, the mixture was quenched with H2O (20 mL) and extracted with Et2O (3 × 20 mL). The collected organic layer was dried over MgSO4. After filtration, the mixture was concentrated in vacuo. The residue was purified by column chromatography. For some cases, further purification was performed by distillation to give a pure product 4.
3-(tert-Butyl)-5-ethyloxazolidin-2-one (4a). Colorless liquid. bp: 78 °C/2 mmHg. IR (neat): 1743.33 cm−1. HRMS: (EI+, 70 eV) Calculated (C9H17NO2) 171.1259 (M+) Found: 171.1257. 1H-NMR: (270 MHz, CDCl3) δ 4.30 (ddd, J = 6.4, 7.8, 8.8 Hz, 1H, 5-H), 3.64 (dd, J = 8.3, 8.8 Hz, 1H, 4HH), 3.20 (dd, J = 7.8, 8.3 Hz, 1H, 4HH), 1.70 (m, 2H, CH2CH3), 1.38 (s, 9H, NC(CH3)3), 0.99 (t, J = 7.3 Hz, 3H, CH2CH3), 13C-NMR: (67.5 MHz, CDCl3) δ 159.76 (C-2), 73.21 (C-5), 52.92 (NC(CH3)3), 48.06 (4-C), 27.45 (CH2CH3), 27.11 (NC(CH3)3), 8.55 (CH2CH3).
3-(tert-Butyl)-5-(chloromethyl)oxazolidin-2-one (4b). Colorless liquid. bp: 80 °C/2 mmHg. IR (neat): 1743.33 cm−1. HRMS: (EI+, 70 eV) Calculated (C8H14ClNO2) 191.0713 (M+) Found: 191.0714. 1H-NMR: (270 MHz, CDCl3) δ 4.64–4.55 (m, 1H, 5-H), 3.74 (dd, J = 8.8, 8.8 Hz, 1H, 4HH), 3.65 (dd, J = 5.9, 6.4 Hz, 2H, CH2Cl), 3.52 (dd, J = 5.9, 8.8 Hz, 1H, 4HH), 1.40 (s, 9H, NC(CH3)3). 13C-NMR: (67.5 MHz, CDCl3) δ 155.62 (C-2), 70.09 (C-5), 53.52 (NC(CH3)3), 46.00 (CH2Cl), 44.70 (C-4), 27.16 (NC(CH3)3).
3-(tert-Butyl)-5-(phenoxymethyl)oxazolidin-2-one (4c). Colorless liquid. bp: 140 °C/2 mmHg. IR (neat): 1727.91 cm−1. HRMS: (EI+, 70 eV) Calculated (C9H17NO2) 249.1365 (M+) Found: 249.1371. 1H-NMR: (270 MHz, CDCl3) δ 7.10 (m, 5H, Ph), 4.72 (m, 1H, 5-H), 4.10 (d, J = 2.9 Hz, 2H, OCH2), 3.76 (dd, J = 8.8, 8.8 Hz, 1H, 4HH), 3.59 (dd, J = 5.9, 8.8 Hz, 1H, 4HH), 1.40 (s, 9H, NC(CH3)3). 13C-NMR: (67.5 MHz, CDCl3) δ 158.09 (C-2), 156.24, 129.53, 121.45, 114.14, 69.57 (OCH2), 67.97 (C-5), 53.45 (NC(CH3)3), 45.55 (4-C), 27.36 (NC(CH3)3).
3-(tert-Butyl)-5-(methoxymethyl)oxazolidin-2-one (4d). Colorless liquid. bp: 75 °C/0.07 mmHg. IR (neat): 1743.33 cm−1. HRMS: (EI+, 70 eV) Calculated (C9H17NO3) 187.1208 (M+) Found: 187.1212. 1H-NMR: (270 MHz, CDCl3) δ 4.51 (m, 1H, 5-H), 3.64 (dd, J = 8.3, 8.8 Hz, 1H, 4HH), 3.52 (dd, J = 4.9 Hz, 2H, CH2O), 3.43 (dd, J = 6.4, 8.3 Hz, 1H, 4HH), 3.40 (s, 3H, OCH3), 1.38 (s, 9H, NC(CH3)3). 13C-NMR: (67.5 MHz, CDCl3) δ 156.30 (C-2), 72.66 (OCH2), 70.38 (C-5), 59.92 (OCH3), 53.13 (NC(CH3)3), 45.14 (4-C), 27.14 (NC(CH3)3).
3-Butyl-5-ethyloxazolidin-2-one (4e). Colorless liquid. bp: 85 °C/2 mmHg. IR (neat): 1751.05 cm−1. HRMS: (EI+, 70 eV) Calculated (C9H17NO2) 171.1259 (M+) Found: 171.1257. 1H-NMR: (270 MHz, CDCl3) δ 4.72 (m, 1H, 5-H), 3.59 (dd, J = 8.3, 8.3 Hz, 1H, 4HH), 3.25 (m, 2H, NCH2), 3.14 (dd, J = 6.8, 8.3 Hz, 1H, 4HH), 1.83–1.64 (m, 2H, NCH2CH2), 1.53 (dt, J = 7.3, 7.3 Hz, 2H, CH2CH3), 1.34 (tq, J = 7.3, 7.3 Hz, 2H, NCH2CH2CH2), 1.00 (t, J = 7.3 Hz, 3H, CH2CH3), 0.94 (t, J = 7.3 Hz, 3H, NCH2CH2CH2CH3). 13C-NMR: (67.5 MHz, CDCl3) δ 157.81 (C-2), 74.18 (C-5), 49.06 (C-4), 43.39 (NCH2), 29.02 (NCH2CH2), 27.70 (CH2CH3), 19.46 (NCH2CH2CH2), 13.34 (NCH2CH2CH2CH3), 8.39 (CH2CH3).
5-Ethyl-3-phenyloxazolidin-2-one (4f). Colorless liquid. The NMR data of 1H and 13C agreed with the previous report [77]. 1H-NMR: (270 MHz, CDCl3) δ 7.56–7.10 (m, 5H, Ph), 4.59 (ddt, J = 6.4, 7.3, 8.3 Hz, 1H, 5-H), 4.08 (dd, J = 8.3, 8.8 Hz, 1H, 4HH), 3.66 (dd, J = 7.3, 8.8 Hz, 1H, 4HH), 1.96–1.71 (dq, J = 6.4, 7.3 Hz, 2H, CH2CH3), 1.07 (t, J = 7.3 Hz, 3H, CH2CH3). 13C-NMR: (67.5 MHz, CDCl3) δ 154.88 (C-2), 138.29, 128.91, 123.79, 118.06, 74.05 (C-5), 49.94 (C-4), 27.88 (CH2CH3), 8.62 (CH2CH3).
5-Ethyl-N,3-diphenyloxazolidin-2-imine (5). Colorless liquid. The NMR data of 1H and 13C agreed with the previous report [56]. 1H-NMR: (270 MHz, CDCl3) δ 8.23–6.60 (m, 10H, Ph × 2), 4.70–4.15 (m, 1H, 5-H), 3.92 (t, J = 8.1 Hz, 4-HH), 3.52 (dd, J = 6.8, 8.1 Hz, 1H, 4HH), 1.95–1.35 (m, 2H, CH2CH3), 0.97 (t, J = 6.8 Hz, 3H, CH2CH3). 13C-NMR: (67.5 MHz, CDCl3) δ 148.79, 148.05, 140.44, 139.34, 129.03, 124.03, 123.30, 122.82, 119.34, 76.16 (C-5), 50.74 (C-4), 27.44 (CH2CH3), 8.78 (CH2CH3).

4.4. Observation of Tin-Indium System by 119Sn NMR

To a two-neck 10-mL reaction vessel, InCl3 (1.0 mmol) and Bu2SnI2 (1.0 mmol) were added in MeCN-d3 (1 mL) and stirred at room temperature for several minutes. After transferring of the reaction mixture into an NMR test tube and addition of tetramethyl stannane as an internal standard, 119Sn NMR was recorded.

5. Conclusions

A novel type of indium-tin species Bu2SnIδ+[InCl3I]δ was revealed to be effective for the catalytic annulation of epoxides. In the reaction with carbon dioxide, the fixation of CO2 proceeded well even at room temperature. In the reaction with isocyanates, novel types of 2-oxazolidiones were obtained in good yields.

Acknowledgments

We are grateful for financial support from The Naito Foundation. We also thank the Instrumental Analysis Center, Faculty of Engineering, Osaka University, for assistance with collecting the spectral data.

Author Contributions

Ikuya Shibata and Akio Baba conceived and designed the experiments; Akira Imakuni and Itaru Suzuki performed the experiments; Itaru Suzuki analyzed the data; Akira Imakuni contributed reagents/materials/analysis tools; Ikuya Shibata wrote the paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Behrens, C.H.; Sharpless, K.B. New transformation of 2,3-epoxy alcohols and related derivatives. Easy routes to homochiral substances. Aldrichim. Acta 1983, 16, 67–79. [Google Scholar] [CrossRef]
  2. Rao, A.S.; Paknikar, S.K.; Kirtane, J.G. Recent advances in the preparation and synthetic applications of oxiranes. Tetrahedron 1983, 39, 2323–3367. [Google Scholar] [CrossRef]
  3. Smith, J.G. Synthetically useful reactions of epoxides. Synthesis 1984, 1984, 629–656. [Google Scholar] [CrossRef]
  4. Bercaw, J.E.; Creutz, C.; Dinjus, E.; Dixon, D.A.; Domen, K.; DuBois, D.L.; Eckert, J.; Fujita, E.; Gibson, D.H.; Goddard, W.A.; et al. Catalysis research of relevance to carbon management: Progress, challenges, and opportunitie. Chem. Rev. 2001, 101, 953–996. [Google Scholar]
  5. Sakakura, T.; Choi, J.-C.; Yasuda, H. Transformation of carbon dioxide. Chem. Rev. 2007, 107, 2365–2387. [Google Scholar] [CrossRef] [PubMed]
  6. Aresta, M.; Dibenedetto, A. The contribution of the utilization option to reducing the CO2 atmospheric loading: research needed to overcome existing barriers for a full exploitation of the potential of the CO2 use. Catal. Today 2004, 98, 455–462. [Google Scholar] [CrossRef]
  7. Aresta, M.; Dibenedetto, A. Utilisation of CO2 as a chemical feedstock: Opportunities and challenges. Dalton Trans. 2007, 2975–2992. [Google Scholar] [CrossRef] [PubMed]
  8. Omae, I. Aspects of carbon dioxide utilization. Catal. Today 2006, 115, 33–52. [Google Scholar] [CrossRef]
  9. Shaikh, A.-A.G.; Sivaram, S. Organic carbonates. Chem. Rev. 1996, 96, 951–976. [Google Scholar] [CrossRef] [PubMed]
  10. Clements, J.H. Reactive applications of cyclic alkylene carbonates. Ind. Eng. Chem. Res. 2003, 42, 663–674. [Google Scholar] [CrossRef]
  11. Schaffner, B.; Holz, J.; Verevkin, S.P.; Borner, A. Organic carbonates as alternative solvents for palladium-catalyzed substitution reactions. ChemSusChem 2008, 1, 249–253. [Google Scholar] [CrossRef] [PubMed]
  12. Bayardon, J.; Holz, J.; Schaffner, B.; Andrushko, V.; Verevkin, S.; Preetz, A.; Borner, A. Propylene carbonate as a solvent for asymmetric hydrogenations. Angew. Chem. Int. Ed. 2007, 46, 5971–5974. [Google Scholar] [CrossRef] [PubMed]
  13. Lagowski, J.J. The Chemistry of Nonaqueous Solvents; Academic Press: New York, NY, USA, 1976. [Google Scholar]
  14. Inaba, M.; Siroma, Z.; Funabiki, A.; Ogumi, A.; Abe, T.; Mizutani, Y.; Asano, M. Electrochemical scanning tunneling microscopy observation of highly oriented pyrolytic graphite surface reactions in an ethylene carbonate-based electrolyte solution. Langmuir 1996, 12, 1535–1540. [Google Scholar] [CrossRef]
  15. Fukuoka, S.; Kawamura, M.; Komiya, K.; Tojo, M.; Hachiya, H.; Hasegawa, K.; Aminaka, M.; Okamoto, H.; Fukawa, I.; Konno, S. A novel non-phosgene polycarbonate production process using by-product CO2 as starting material. Green Chem. 2003, 5, 497–507. [Google Scholar] [CrossRef]
  16. Dyen, M.E.; Swern, D. 2-Oxazolidones. Chem. Rev. 1967, 67, 197–246. [Google Scholar] [CrossRef] [PubMed]
  17. Ozaki, S. Recent advances in isocyanate chemistry. Chem. Rev. 1972, 72, 457. [Google Scholar] [CrossRef]
  18. Shiozaki, M. Pharmacology of oxazolidinones in rat decerebrate rigidity, with reference to their glutamate blocking action. Gen. Pharm. 1988, 19, 163–169. [Google Scholar] [CrossRef]
  19. Barbachyn, M.R.; Ford, C.W. Oxazolidinone structure-activity relationships leading to linezolid. Angew. Chem. Int. Ed. 2003, 42, 2010–2023. [Google Scholar] [CrossRef] [PubMed]
  20. Guo, B.; Fan, H.; Xin, Q.; Chu, W.; Wang, H.; Huang, Y.; Chen, X.; Yang, Y. Solubility-driven optimization of (pyridin-3-yl) benzoxazinyl-oxazolidinones leading to a promising antibacterial agent. J. Med. Chem. 2013, 56, 2642–2650. [Google Scholar] [CrossRef] [PubMed]
  21. Trstenjak, U.; Ilas, J.; Kikelj, D. Low molecular weight dual inhibitors of factor Xa and fibrinogen binding to GPIIb/IIIa with highly overlapped pharmacophores. Eur. J. Med. Chem. 2013, 64, 302–313. [Google Scholar] [CrossRef] [PubMed]
  22. Xue, T.; Ding, S.; Guo, B.; Zhou, Y.; Sun, P.; Wang, H.; Chu, W.; Gong, G.; Wang, Y.; Chen, X.; et al. Design, synthesis, and structure-activity and structure-pharmacokinetic relationship studies of novel [6,6,5] tricyclic fused oxazolidinones leading to the discovery of a potent, selective, and orally bioavailable FXa inhibitor. J. Med. Chem. 2014, 57, 7770–7791. [Google Scholar] [CrossRef] [PubMed]
  23. Cardillo, G.; Orena, M.; Sandri, S.; Tomashini, C. An efficient synthesis of (R)-(+)- and (S)-(−)-propranolol from resolved 5-idomethyloxazo-lidin-2-ones. Tetrahedron 1987, 43, 2505–2512. [Google Scholar] [CrossRef]
  24. Rao, A.V.R.; Dhar, T.G.M.; Chakraborty, T.K.; Gurjar, M.K. A stereospecific synthesis of (4R)-4-[(E)-2-butenyl]-4, N-dimethyl-L-threonine (MeBmt). Tetrahedron Lett. 1988, 29, 2069–2072. [Google Scholar] [CrossRef]
  25. Ager, D.J.; Prakash, I.; Schaad, D.R. 1,2-Amino alcohols and their heterocyclic derivatives as chiral auxiliaries in asymmetric synthesis. Chem. Rev. 1996, 96, 835–876. [Google Scholar] [CrossRef] [PubMed]
  26. Roush, W.R.; James, R.A. Towards the synthesis of aureolic acid analogue conjugates: Synthesis and glycosidation reactions of 3-O-acetyl-4-azido-2,4,6-trideoxy-l-glucopyranose derivatives. Aust. J. Chem. 2002, 55, 141–146. [Google Scholar] [CrossRef]
  27. Aurelio, L.; Brownlee, R.T.C.; Hughes, A.B. Synthetic preparation of N-methyl-α-amino acids. Chem. Rev. 2004, 104, 5823–5846. [Google Scholar] [CrossRef] [PubMed]
  28. Mukhtar, T.A.; Wright, G.D. Streptogramins, oxazolidinones, and other inhibitors of bacterial protein synthesis. Chem. Rev. 2005, 105, 529–542. [Google Scholar] [CrossRef] [PubMed]
  29. Birrell, J.A.; Jacobsen, E.N. A practical method for the synthesis of highly enantioenriched trans-1,2-amino alcohols. Org. Lett. 2013, 15, 2895–2897. [Google Scholar] [CrossRef] [PubMed]
  30. Laserna, V.; Guo, W.; Kleij, A.W. Aluminium-catalysed oxazolidinone synthesis and their conversion into functional non-symmetrical ureas. Adv. Synth. Catal. 2015, 357, 2849–2854. [Google Scholar] [CrossRef]
  31. Heravi, M.M.; Zadsirjan, V.; Farajpour, B. Applications of oxazolidinones as chiral auxiliaries in the asymmetric alkylation reaction applied to total synthesis. RSC Adv. 2016, 6, 30498–30551. [Google Scholar] [CrossRef]
  32. Darensbourg, D.J.; Holtcamp, M.W. Catalysts for the reactions of epoxides and carbon dioxide. Coord. Chem. Rev. 1996, 153, 155–174. [Google Scholar] [CrossRef]
  33. Sun, J.; Fujita, S.; Arai, M. Development in the green synthesis of cyclic carbonate from carbon dioxide using ionic liquids. J. Organomet. Chem. 2005, 690, 3490–3497. [Google Scholar] [CrossRef]
  34. Cintas, P. Synthetic organoindium chemistry: What makes indium so appealing? Synlett 1995, 1995, 1087–1096. [Google Scholar] [CrossRef]
  35. Shibata, I.; Mitani, I.; Imakuni, A.; Baba, A. Highly efficient synthesis of cyclic carbonates from epoxides catalyzed by indium tribromide system. Tetrahedron Lett. 2011, 52, 721–723. [Google Scholar] [CrossRef]
  36. Gulbins, K.; Hamann, M. Darstellung von oxazolidonen. Angew. Chem. 1958, 70, 705. [Google Scholar] [CrossRef]
  37. Gulbins, K.; Benzing, G.; Maysenholder, R.; Hamann, K. Synthese von substituierten oxazolidone-(2). Chem. Ber. 1960, 93, 1975–1982. [Google Scholar] [CrossRef]
  38. Irwin, W.J.; Wheeler, D.L. 3,5-Diphenyloxazolidin-2-one. J. Chem. Soc. C 1971, 66, 3166–3167. [Google Scholar] [CrossRef]
  39. Herweh, J.E.; Foglia, T.A.; Swern, D.J. Synthesis and nuclear magnetic resonance spectra of 2-oxazoli-dones. J. Org. Chem. 1968, 33, 4029–4033. [Google Scholar] [CrossRef]
  40. Herweh, J.E. Bis-2-oxazolidones-preparation and characterization. J. Heterocycl. Chem. 1968, 5, 687–690. [Google Scholar] [CrossRef]
  41. Herweh, J.E.; Kaufmann, W.J. 2-Oxazolidones via the lithium bromide catalyzed reaction of isocyanates with epoxides in hydrocarbon solvents. Tetrahedron Lett. 1971, 12, 809–812. [Google Scholar] [CrossRef]
  42. Aroua, L.; Baklouti, A. α,ω-Bis(oxazolidinone)polyoxyethylene via a lithium bromide–catalyzed reaction of oligoethylene glycol diglycidyl ethers with isocyanates. Synth. Commun. 2007, 37, 1935–1942. [Google Scholar] [CrossRef]
  43. Speranza, G.P.; Peppel, W.J. Preparation of substituted 2-oxazolidones from 1,2-epoxides and isocyanates. J. Org. Chem. 1958, 23, 1922–1924. [Google Scholar] [CrossRef]
  44. Weiner, M.L. Reaction of phenyl isocyanate with phenyl glycidyl ether. J. Org. Chem. 1961, 26, 951–952. [Google Scholar] [CrossRef]
  45. Toda, Y.; Gomyou, S.; Tanaka, S.; Komiyama, Y.; Kikuchi, A.; Suga, H. Tetraarylphosphonium salt-catalyzed synthesis of oxazolidinones from Isocyanates and Epoxides. Org. Lett. 2017, 19, 5786–5789. [Google Scholar] [CrossRef] [PubMed]
  46. Braun, D.; Weinert, J. Umsetzung von epoxiden mit isocyanaten, II. Darstellung und charakterisierung von 2-oxazolidinonen. Eur. J. Org. Chem. 1979, 1979, 200–209. [Google Scholar] [CrossRef]
  47. Zhang, X.; Chen, W.; Zhao, C.; Li, C.; Wu, X.; Chen, W.Z. A Facile and efficient synthesis of 3-aryloxazolidin-2-ones from isocyanates and epoxides promoted by MgI2 etherate. Synth. Commun. 2010, 40, 3654–3659. [Google Scholar] [CrossRef]
  48. Baba, A.; Fujiwara, M.; Matsuda, H. Unusual cycloaddition of oxiranes with isocyanates catalyzed by tetraphenylstibonium iodide; selective formation of 3,4-disubstituted oxazolidinones. Tetrahedron Lett. 1986, 27, 77–80. [Google Scholar] [CrossRef]
  49. Fujiwara, M.; Baba, A.; Matsuda, H. Selective α-cleavage cycloaddition of oxiranes with heterocumulenes catalyzed by tetraphenylstibonium iodide. J. Heterocycl. Chem. 1988, 25, 1351–1357. [Google Scholar] [CrossRef]
  50. Fujiwara, M.; Baba, A.; Matsuda, H. Mechanistic studies of tetraphenylstibonium iodide-catalyzed cycloaddition of oxiranes with heterocumulenes. Bull. Chem. Soc. Jpn. 1990, 63, 1069–1073. [Google Scholar] [CrossRef]
  51. Wu, X.; Mason, J.; North, M. Isocyanurate formation during oxazolidinone synthesis from epoxides and isocyanates catalysed by a chromium(salphen) complex. Chem. Eur. J. 2017, 23, 12937–12943. [Google Scholar] [CrossRef] [PubMed]
  52. Trost, B.M.; Sudhakar, A.R. Cis Hydroxyamination equivalent. Application to the synthesis of (-)-acosamine. J. Am. Chem. Soc. 1987, 109, 3792–3794. [Google Scholar] [CrossRef]
  53. Shibata, I.; Baba, A.; Iwasaki, H.; Matsuda, H. Cycloaddition reaction of heterocumulenes with oxiranes catalyzed by organotin iodide-Lewis base complex. J. Org. Chem. 1986, 51, 2177–2184. [Google Scholar] [CrossRef]
  54. Fujiwara, M.; Baba, A.; Tomohisa, Y.; Matsuda, H. Cycloaddition reaction of 2,3-disubstituted oxiranes with isocyanates by highly activated catalyst; Ph4SbI–Bu3SnI. Chem. Lett. 1986, 15, 1963–1966. [Google Scholar] [CrossRef]
  55. Baba, A.; Seki, K.; Matsuda, H. Stereospecific cycloaddition of heterocumulenes to oxiranes catalyzed by organotin halide complexes. J. Heterocycl. Chem. 1990, 27, 1925–1930. [Google Scholar] [CrossRef]
  56. Yano, K.; Amishiro, N.; Baba, A.; Matsuda, H. Selective formation of α-cleavage cycloadduct of oxirane with heterocumulene promoted by high-coordinated trialkyltin complexes. Bull. Chem. Soc. Jpn. 1991, 64, 2661–2667. [Google Scholar] [CrossRef]
  57. Tsuzuki, R.; Ishikawa, K.; Kase, M. New reactions of organic isocyanates. I. Reaction with alkylene carbonates. J. Org. Chem. 1960, 25, 1009–1012. [Google Scholar] [CrossRef]
  58. Holecek, J.; Nádvorník, M.; Handlír, K.; Lycka, A. 13C and 119Sn NMR Study of some four- and five-coordinate triphenyltin(IV) compounds. J. Organomet. Chem. 1983, 241, 177–184. [Google Scholar] [CrossRef]
  59. Nádvorník, M.; Holecek, J.; Handlír, K.; Lycka, A. The 13C and 119Sn NMR spectra of some four- and five-coordinate tri-n-butyltin(IV) compounds. J. Organomet. Chem. 1984, 275, 43–51. [Google Scholar] [CrossRef]
  60. Bonini, C.; Righi, G. Regio- and chemoselective synthesis of halohydrins by cleavage of oxiranes with metal halides. Synthesis 1994, 1994, 225–238. [Google Scholar] [CrossRef]
  61. Backwell, J.E.; Young, M.W.; Sharpless, K.B. Vicinal acetoxychlorination of olefins by chromyl chloride in acetyl chloride. Tetrahedron Lett. 1977, 40, 3523–3526. [Google Scholar]
  62. Shibata, I.; Baba, A.; Matsuda, H. Regioselective ring cleavage of oxiranes catalyzed by organotin halide—Triphenylphosphine complex. Tetrahedron Lett. 1986, 27, 3021–3024. [Google Scholar] [CrossRef]
  63. Shibata, I.; Yoshimura, N.; Baba, A.; Matsuda, H. Remarkable dependency of the regioselectivity in the ring opening of α,β-epoxyketones upon tin halide-Lewis base complexes as catalysts. Tetrahedron Lett. 1992, 33, 7149–7152. [Google Scholar] [CrossRef]
  64. Baba, A.; Shibata, I. Dihaloindium hydride as a novel reducing agent. Chem. Rec. 2005, 5, 323–335. [Google Scholar] [CrossRef] [PubMed]
  65. Inoue, K.; Sawada, A.; Shibata, I.; Baba, A. Indium hydride: A novel radical initiator in the reduction of organic halides with tributyltin hydride. Tetrahedron Lett. 2001, 42, 4661–4663. [Google Scholar] [CrossRef]
  66. Miyai, T.; Inoue, K.; Baba, A. Indoum triiodide (InI3)-catalyzed allylation of carbonyl compounds by allylic tins. Synlett 1997, 1997, 699–700. [Google Scholar] [CrossRef]
  67. Marshall, J.A.; Hinkle, K.W. Synthesis of anti-homoallylic alcohols and monoprotected 1,2-diols through InCl3-promoted addition of allylic stannanes to aldehydes. J. Org. Chem. 1995, 60, 1920–1921. [Google Scholar] [CrossRef]
  68. Li, X.-R.; Loh, T.-P. Indium trichloride-promoted tin-mediated carbonyl allylation in water: High simple diastereo- and diastereofacial selectivity. Tetrahedron Asymmetry 1996, 7, 1535–1538. [Google Scholar] [CrossRef]
  69. Baba, A.; Kishiki, H.; Shibata, I.; Matsuda, H. Reaction of tributyltin ω-haloalkoxides with isocyanates or carbodiimides. A possibility of the addition of a tin-oxygen bond across the carbon-oxygen group of isocyanate Organometallics 1984, 4, 1329–1333. [Google Scholar]
  70. Shibata, I.; Baba, A.; Matsuda, H. Formation of N-tributylstannyl-2-oxazolidone from (Bu3Sn)2O and 2-chloroethyl isocyanate. J. Chem. Soc. Chem. Commun. 1986, 1703–1704. [Google Scholar] [CrossRef]
  71. Shibata, I.; Nakamura, K.; Baba, A.; Matsuda, H. Formation of N-tributylstannyl heterocycle from bis(tributyltin) oxide and ω-haloalkyl isocyanate. One-pot convenient synthesis of 2-oxazolidinones and tetrahydro-2H-1,3-oxazin-2-one. Bull. Chem. Soc. Jpn. 1989, 62, 853–859. [Google Scholar] [CrossRef]
  72. Delmond, B.; Pommier, J.C.; Valade, J. Halogénoalcoxyétains: I. (halogéno-2 alcoy)tributylétains, synthéses et propriétés. application á la préparation d’époxydes. J. Organomet. Chem. 1972, 35, 91–104. [Google Scholar] [CrossRef]
  73. Delmond, B.; Pommier, J.C.; Valade, J. Halogénoalcoxyétains II. halogéno-3 alcoxytributylétains, synthéses et propriétés-application à la preéparation d’oxétannes et d’alcools stanniques. J. Organomet. Chem. 1973, 47, 337–350. [Google Scholar] [CrossRef]
  74. Suzuki, I.; Uji, Y.; Ieki, R.; Kanaya, S.; Tsunoi, S.; Shibata, I. Transition-metal-free reductive coupling of 1,3-butadienes with aldehydes catalyzed by dibutyliodotin hydride. Org. Lett. 2017, 19, 5392–5394. [Google Scholar] [CrossRef] [PubMed]
  75. Whiteoak, C.J.; Martin, E.; Belmonte, M.M.; Benet-Buchholz, J.; Kleij, A.W. An efficient iron catalyst for the synthesis of five- and six-membered organic carbonates under mild conditions. Adv. Synth. Catal. 2012, 354, 469–476. [Google Scholar] [CrossRef]
  76. Liu, X.; Cao, C.; Li, Y.; Guan, P.; Yang, L.; Shi, Y. Cycloaddition of CO2 to epoxides catalyzed by N-heterocyclic carbene (NHC)–ZnBr2 System under mild conditions. Synlett 2012, 43, 1343–1348. [Google Scholar] [CrossRef]
  77. Baba, A.; Shibata, I.; Matsuda, K.; Matsuda, H. The cycloaddition of isocyanates and carbodiimides to oxiranes catalyzed by organotin iodide-Lewis base complexes. Synthesis 1985, 1985, 1144–1146. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are not available from the authors.
Scheme 1. Catalytic synthesis of oxazolidin-2-imine 5.
Scheme 1. Catalytic synthesis of oxazolidin-2-imine 5.
Molecules 23 00782 sch001
Figure 1. 119Sn NMR of the tin-indium system.
Figure 1. 119Sn NMR of the tin-indium system.
Molecules 23 00782 g001
Scheme 2. A plausible catalytic cycle.
Scheme 2. A plausible catalytic cycle.
Molecules 23 00782 sch002
Table 1. Synthesis of cyclic carbonates 2 from epoxides 1 with CO2 at rt a.
Table 1. Synthesis of cyclic carbonates 2 from epoxides 1 with CO2 at rt a.
Molecules 23 00782 i001
EntryRCat.Time (h)ProductYield 2 (%) b
1Me (1a)InCl35Molecules 23 00782 i0022atrace
2 InCl3-Bu2SnI2578
3 Bu2SnI25trace
4Et (1b)InCl3-Bu2SnI28Molecules 23 00782 i0032b85
5Ph (1c)InCl3-Bu2SnI210Molecules 23 00782 i0042c69
6CH2Cl (1d)InCl3-Bu2SnI25Molecules 23 00782 i0052d82
7CH2OPh (1e)InCl3-Bu2SnI25Molecules 23 00782 i0062e90
8CH2OMe (1f)InCl3-Bu2SnI210Molecules 23 00782 i0072f68
a InCl3 0.5 mmol, Bu2SnI2 1 mmol, epoxide 1 10 mmol, CO2 3.9 Pa, MeCN 3 mL; b Determined by 1H-NMR.
Table 2. Synthesis of 2-oxazolidinones 4 from epoxides 1bf with isocyanates 3 a.
Table 2. Synthesis of 2-oxazolidinones 4 from epoxides 1bf with isocyanates 3 a.
Molecules 23 00782 i008
EntryR1R2ConditionsProductYield 3 (%) b
1Et (1b)t-Bu (3a)rt, 10 hMolecules 23 00782 i0094a79
2 7 c
3 trace d
4CH2Cl (1d) 60 °C, 3 hMolecules 23 00782 i0104b79
5CH2OPh (1e) 60 °C, 3 hMolecules 23 00782 i0114c90
6CH2OMe (1f) 60 °C, 3 hMolecules 23 00782 i0124d99
7Et (1b)n-Bu (3b)60 °C, 7 hMolecules 23 00782 i0134e49
8Et (1b)Ph (3c)rt, 20 hMolecules 23 00782 i0144f64
a InCl3 0.25 mmol, Bu2SnI2 0.5 mmol, epoxide 1 5 mmol, isocyanate 3 5.5 mmol, MeCN 1.5 mL, under nitrogen; b Determined by 1H-NMR; c Only Bu2SnI2 was used; d Only InCl3 was used.

Share and Cite

MDPI and ACS Style

Suzuki, I.; Imakuni, A.; Baba, A.; Shibata, I. Catalytic Annulation of Epoxides with Heterocumulenes by the Indium-Tin System. Molecules 2018, 23, 782. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules23040782

AMA Style

Suzuki I, Imakuni A, Baba A, Shibata I. Catalytic Annulation of Epoxides with Heterocumulenes by the Indium-Tin System. Molecules. 2018; 23(4):782. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules23040782

Chicago/Turabian Style

Suzuki, Itaru, Akira Imakuni, Akio Baba, and Ikuya Shibata. 2018. "Catalytic Annulation of Epoxides with Heterocumulenes by the Indium-Tin System" Molecules 23, no. 4: 782. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules23040782

Article Metrics

Back to TopTop