Mannich-Type Reaction of Aldimines with 2-Silyloxydienes Catalyzed by Ammonium Chloride

Abstract

Reaction of imines with 2-silyloxydiene catalyzed by ammonium chloride has been perfectly proceeded under environmentally friendly conditions to give Mannich-type product selectively. The reaction would proceed via Mannich-type mechanism, not cyclization/ring-opening process. Cyclopropanation of the corresponding Mannich-type product gave the precursor of prasugrel skeleton in high yield.

Share and Cite:

Fukumoto, S. , Shigenobu, M. and Ishimaru, K. (2019) Mannich-Type Reaction of Aldimines with 2-Silyloxydienes Catalyzed by Ammonium Chloride. International Journal of Organic Chemistry, 9, 163-173. doi: 10.4236/ijoc.2019.94014.

1. Introduction

Mannich-type reactions are widely recognized as a powerful method for constructing a variety of b-aminoketones [1] - [6]. However, Mannich-type reaction of imines with 2-silyloxydienes, which provides easy access to β-aminoketones having a terminal olefin, is still challenging because [4 + 2] type cycloadducts [7] - [15] or mixtures of Mannich-type products and cycloadducts [16] [17] [18] are obtained in most cases, as shown in Figure 1. Previously we first reported a highly effective Mannich-type reaction of imine with 2-silyloxydiene in the presence of zinc triflate and water [19] [20] [21] [22], which gave the corresponding β'-amino-α,β-enones as attractive skeletons for pharmaceutically useful compounds [23] [24] [25] [26]. Although many vinylogous Mannich-type reactions have been developed [27] - [36], only a few examples that describe the selective preparation of β'-amino-α,β-enones by the reaction of imine with 2-silyloxybutadiene have been reported so far. Thus, Kawęcki isolated the open-chain products from the aza-Diels-Alder reaction of sulfinimines with the

Figure 1. Reaction of imine with 2-silyloxydiene.

Rawal diene [37]. Pan et al. reported the addition of an α,β-unsaturated ketone-derived enolate to chiral N-phosphonyl imines [38], and Prasad et al. developed the reaction of chiral sulfinimines with silyloxydiene using TMSOTf [39]. In spite of these recent achievements, a more economical and environmentally benign synthetic methodology using green and sustainable catalysts has not been reported yet that offer alternatives to metal catalysts. Here we report the ammonium chloride-catalyzed Mannich-type reaction of imines with 2-silyloxybutadienes under mild conditions.

2. Results and Discussion

Initially, we examined the reaction of imine 1a, derived from benzaldehyde and o-anisidine, with 2-silyloxybutadiene 2a (Table 1). In contrast to the similar aza-Diels-Alder reaction of electron-rich Danishefsky’s diene, reported by Ding et al., which afforded the cyclic product in MeOH in the absence of any acids [40], the reaction of imine 1a with 2-silyloxybutadiene 2a in EtOH or MeOH without additives gave no product (Table 1, entry 1). This result suggested that

Table 1. Mannich-type reacton of imine (1a) with 2-silyloxydiene (2a)a.

aConditions: imine 1 (1 mmol), 2-silyloxydiene 2 (1.2 mmol), catalyst (0.1 mmol) in dry solvent (1 mL), r.t., 1 day. bIsolated yields.

using additive or catalyst was necessary to promote the reaction. We found that reaction with ammonium chloride (10 mol%) as a catalyst in EtOH gave the corresponding Mannich-type product 3a selectively in 95% isolated yield (entry 2). Interestingly, no trace of cycloadduct was detected by 500 MHz 1H NMR spectroscopy in the crude product. The previously reported reaction of imines having the N-benzyl group [19] did not give any products using ammonium chloride in EtOH, indicating that the reactivity of the imine is largely dependent on the N-protecting group. Finally, the attempt to perform the reaction using other additives and solvents was unsuccessful (Table 1, entries 3 - 7).

Having established the optimal reaction conditions for the Mannich-type reaction, we subsequently explored the scope of the reaction with respect to the imine substrates (Table 2). Imines 1b and 1c bearing an o- or p-tolyl group reacted to provide the corresponding products 3b (90% yield) and 3c (88% yield), respectively. Meanwhile, imines having an electron-donating or an electron-withdrawing group all reacted in a satisfactory way to provide the corresponding products 3d-3g in high yield. The reaction with 2-silyloxybutadiene 2b, derived from mesityl oxide, also proceeded to give 3h-3k in 87% - 97% yields. Further investigation of the reaction with 2-silyloxydiene 2c derived from acetylcyclohexene afforded the corresponding β'-amino-α,β-enones in 95% - 98% yields (Table 3).

To investigate the reaction mechanism, the reaction of 1a and 2a was quenched after 1 h and analyzed using 500 MHz 1H NMR spectroscopy. No cycloadduct was detected but a Mannich-type product and starting materials were observed. We suspect that the Mannich-type products are not formed via cyclization/ring-opening mechanism, as in the case of our previous reaction between N-benzyl-protected imine and 2-silyloxybutadiene using zinc triflate and water (Scheme 1) [19]. Additionally, further HCl work-up of the acyclic product 3a gave no cycloadducts but the β'-amino-α,β-enone was recovered. However, N-benzyl-protected acyclic products afforded piperidones upon reaction with HCl [19], indicating that the acyclic product 3a is stable under acidic conditions.

To demonstrate the synthetic utility of the Mannich-type products, we performed the preparation of a precursor of the prasugrel skeleton [41] [42], as

Scheme 1. Plausible mechanism.

Table 2. Scope of imines (1) and 2-silyloxydienes (2a)a.

aConditions: imine 1 (1 mmol), 2-silyloxydiene 2 (1.2 mmol), ammonium chloride (0.1 mmol) in dry EtOH (1 mL), r.t., 1 day. bIsolated yields.

shown in Scheme 2. Thus, the reaction of the imine (1h) derived from o-fluorobenzaldehyde with 2-silyloxybutadiene 2b proceeded smoothly to afford

Table 3. Reaction of imines (1) with 2-silyloxydiene (2c) derived from acetylcyclohexenea.

aConditions: imine 1 (1 mmol), 2-silyloxydiene 2c (1.2 mmol), ammonium chloride (0.1 mmol) in dry EtOH (1 mL), r.t., 1 day. bIsolated yields.

the corresponding acyclic product 3o, which was then cyclopropanated to give 4o in 65% yield.

3. Conclusion

In summary, a Mannich-type reaction of imine with 2-silyloxybutadiene that provides access to versatile β'-amino-α,β-enones has been developed under green conditions. We found that the use of ammonium chloride in EtOH is critical for the efficient outcome of the reaction, and this non-metal method is useful for various dienes and aldimines compared to reports so far [37] [38] [39]. According to our results, the reaction proceeds via Mannich-type mechanism, instead of through a cyclization/ring-opening process. Additionally, we prepared a precursor of the prasugrel skeleton by cyclopropanation of 3o. Further investigation of the applicability of this reaction and mechanistic elucidation is currently in progress.

4. Experimental

Typical Procedure for Mannich-Type reaction of imine 1 with 2-silyloxydiene 2

To a stirred solution of NH4Cl (0.006 g, 0.1 mmol), imine 1 (1 mmol) in dry ethanol (1 mL) was added 2-silyloxydiene 2 (1.2 mmol) at r.t.. The reaction mixture was stirred at r.t. for 24 h, the solvent was evaporated at reduced pressure. The crude product was purified by flash column chromatography (ethyl acetate: n-hexane = 1:6) to afford 3.

Data for 3a; colorless oil, 0.34 g, 95%. 1H NMR (CDCl3, 500 MHz) δ 3.17 (dd, 1H, J = 15.4 Hz, 5.7 Hz), 3.23 (dd, 1H, J = 15.4 Hz, 7.5 Hz), 3.85 (s, 3H), 4.97 (dd, 1H, J = 6.9 Hz, 6.9 Hz), 6.44 (d, 1H, J = 7.5 Hz), 6.45 - 6.76 (m, 4H), 7.22 - 7.25 (m, 1H), 7.32 - 7.49 (m, 10H); HRMS (EI) m/z [M]+ calcd. for C24H23NO2 357.1729, found: 357.1727.

Data for 3b; colorless oil. 0.30 g, 90%. 1H NMR (CDCl3, 300 MHz) δ 2.02 (s, 3H), 2.87 (dd, 1H, J = 15.4 Hz, 5.9 Hz), 2.95 (dd, 1H, J = 15.4 Hz, 6.9 Hz), 3.54 (s, 3H), 4.69 (dd, 1H, J = 6.6 Hz, 6.6 Hz), 6.23 (d, 1H, J = 6.9 Hz), 6.35 - 6.51 (m, 4H), 6.78 - 6.87 (m, 2H), 7.04 - 7.24 (m, 8H); HRMS (EI) m/z [M]+ calcd. for C25H25NO2 371.1885, found: 371.1894.

Data for 3c; colorless oil. 0.32 g, 88%. 1H NMR (CDCl3, 300MHz) δ 2.06 (s, 3H), 2.92 (dd, 1H, J = 15.4 Hz, 5.9 Hz), 2.96 (dd, 1H, J = 15.4 Hz, 7.0 Hz), 3.58 (s, 3H), 4.67 (dd, 1H, J = 6.3 Hz, 6.3 Hz), 6.23 (d, 1H, J = 7.7 Hz), 6.35 - 6.51 (m, 3H), 6.78 - 6.85 (m, 1H), 6.87 - 7.07 (m, 4H), 7.08 - 7.18 (m, 2H), 7.20 - 7.26 (m, 3H); HRMS (EI) m/z [M]+ calcd. for C25H25NO2 371.1885, found: 371.1902.

Data for 3d; colorless oil. 0.32 g, 85%. 1H NMR (CDCl3, 500 MHz) δ 3.12 (dd, 1H, J = 15.5 Hz, 5.7 Hz), 3.20 (dd, 1H, J = 15.5 Hz, 6.9 Hz), 3.74 (s, 3H), 3.83 (s, 3H), 4.91 (dd, 1H, J = 6.3 Hz, 6.3 Hz), 6.46 (d, 1H, J = 1.7 Hz), 6.61 - 6.63 (m, 4H), 6.81 - 6.85 (m, 2H), 7.25 - 7.37 (m, 5H), 7.42 - 7.47 (m, 3H); HRMS (EI) m/z [M]+ calcd. for C25H25NO3 387.1834, found: 387.1838.

Data for 3e; colorless oil. 0.36 g, 93%. 1H NMR (CDCl3, 300MHz) δ 3.12 (dd, 1H, J = 15.7 Hz, 5.5 Hz), 3.20 (dd, 1H, J =15.7 Hz, 5.5 Hz), 3.85 (s, 3H), 4.93 - 4.99 (m, 2H), 6.36 (d, 1H, J = 6.2 Hz), 6.61 - 6.76 (m, 4H), 7.26 - 7.53 (m, 10H); HRMS (EI) m/z [M]+ calcd. for C24H22NO2Cl 391.1339, found: 391.1328.

Data for 3f; colorless oil. 0.32 g, 81%. 1H NMR (CDCl3, 500 MHz) δ 3.02 (dd, 1H, J = 14.9 Hz, 8.6 Hz), 3.28 (d, 1H, J = 3.5 Hz), 3.71 (s, 3H), 5.28 (dd, 1H, J = 9.1 Hz, 3.4 Hz), 6.21 (dd, 1H, J = 6.2 Hz, 1.8 Hz), 6.59 - 6.74 (m, 4H), 7.17 - 7.59 (m, 10H); HRMS (EI) m/z [M]+ calcd. for C24H22NO2Cl 391.1339, found: 391.1335.

Data for 3g; colorless oil. 0.33 g, 87%. 1H NMR (CDCl3, 500 MHz) δ 3.13 (dd, 1H, J = 15.5 Hz, 5.7 Hz), 3.20 (dd, 1H, J = 15.5 Hz, 5.7 Hz), 3.83 (s, 3H), 4.93 (dd, 1H, J = 6.3 Hz, 6.3 Hz), 6.35 - 6.39 (m, 1H), 6.59 - 6.80 (m, 4H), 6.95 - 6.99 (m, 2H), 7.18 - 7.25 (m, 3H), 7.33 - 7.38 (m, 3H), 7.45 - 7.53 (m, 2H); HRMS (EI) m/z [M]+ calcd. for C24H22NO2 375.1634, found: 375.1628.

Data for 3h; colorless oil. 0.29 g, 95%. 1H NMR (CDCl3, 300 MHz) δ 1.72 (s, 3H), 2.00 (s, 3H), 2.76 (dd, 1H, J = 16.9 Hz, 5.9 Hz), 2.82 (dd, 1H, J = 16.9 Hz, 5.9 Hz), 3.81 (s, 3H), 4.77 (dd, 1H, J = 6.6 Hz, 6.6 Hz), 5.91 (brs, 1H), 6.30 - 6.35 (m, 1H), 6.45 - 6.55 (m, 1H), 6.58 - 6.64 (m, 2H), 7.06 - 7.15 (m, 1H), 7.17 - 7.20 (m, 2H), 7.26 - 7.28 (m, 2H); HRMS (EI) m/z [M]+ calcd. for C20H23NO2 309.1729, found: 309.1722.

Data for 3i; colorless oil. 0.31 g, 97%. 1H NMR (CDCl3, 500 MHz) δ 1.85 (s, 3H), 2.11 (s, 3H), 2.30 (s, 3H), 2.90 (dd, 1H, J = 13.2 Hz, 5.7 Hz), 2.93 (dd, 1H, J = 13.2 Hz, 5.7 Hz), 3.85 (s, 3H), 4.85 (dd, 1H, J = 6.9 Hz, 6.9 Hz), 6.04 (brs, 1H), 6.44 - 6.50 (m, 1H), 6.58 - 6.52 (m, 1H), 6.55 - 6.73 (m, 2H), 7.10 - 7.13 (m, 2H), 7.20 - 7.28 (m, 2H); HRMS (EI) m/z [M]+ calcd. for C21H24NO2 323.1885, found: 323.1894.

Data for 3j; colorless oil. 0.31 g, 92%. 1H NMR (CDCl3, 300 MHz) δ 1.80 (s, 3H), 2.08 (s, 3H), 2.80 (dd, 1H, J = 15.4 Hz, 6.2 Hz), 2.90 (dd, 1H, J = 15.4 Hz, 6.2 Hz), 3.70 (s, 3H), 3.80 (s, 3H), 4.81 (dd, 1H, J = 6.6 Hz, 6.6 Hz), 5.95 - 6.00 (m, 1H), 6.34 - 6.44 (m, 1H), 6.52 - 6.58 (m, 1H), 6.67 - 6.72 (m, 2H), 6.79 - 6.84 (m, 2H), 7.24 - 7.28 (m, 2H); HRMS (EI) m/z [M]+ calcd. for C21H25NO3 339.1834, found: 339.1836.

Data for 3k; colorless oil. 0.29 g, 89%. 1H NMR (CDCl3, 300 MHz) δ 1.85 (s, 3H), 2.10 (s, 3H), 2.88 (dd, 1H, J = 15.4 Hz, 5.9 Hz), 2.92 (dd, 1H, J = 15.4 Hz, 5.9 Hz), 3.11 (s, 3H), 4.84 (dd, 1H, J = 6.6 Hz, 6.6 Hz), 5.99 - 6.01 (m, 1H), 6.36 (d, 1H, J = 7.7 Hz), 6.59 - 6.76 (m, 3H), 6.95 - 7.01 (m, 2H), 7.32 - 7.37 (m, 2H); HRMS (EI) m/z [M]+ calcd. for C20H22NO2F 327.1635, found: 327.1628.

Data for 3l; colorless oil. 0.33 g, 98%. 1H NMR (CDCl3, 300 MHz) δ 1.50 - 1.65 (m, 4H), 2.15 - 2.28 (m, 4H), 3.12 (d, 2H, J = 6.6 Hz), 3.85 (s, 3H), 4.82 - 4.90 (m, 1H), 4.92 - 5.05 (m, 1H), 6.37 (dd, 1H, J = 5.9 Hz, 1.8 Hz), 6.57 - 6.84 (m, 4H), 7.18 - 7.39 (m, 5H); HRMS (EI) m/z [M]+ calcd. for C22H25NO2 335.1885, found: 335.1876.

Data for 3m; colorless oil. 0.27 g, 78%. 1H NMR (CDCl3, 300 MHz) δ 1.50 - 1.65 (m, 4H), 2.15 - 2.25 (m, 4H), 2.30 (s, 3H), 3.11 (d, 2H, J = 6.6 Hz), 3.85 (s, 3H), 4.82 (dd, 1H, J = 6.6 Hz, 6.6 Hz), 6.40 (d, 1H, J = 6.2 Hz), 6.57 - 6.75 (m, 3H), 6.83 - 6.86 (m, 1H), 7.10 (d, 2H, J = 8.0 Hz), 7.27 (d, 2H, J = 8.4 Hz); HRMS (EI) m/z [M]+ calcd. for C23H27NO2 349.2042, found: 349.2044.

Data for 3n; colorless oil. 0.34 g, 95%. 1H NMR (CDCl3, 300 MHz) δ 1.56 - 1.59 (m, 4H), 2.17 - 2.21 (m, 4H), 3.11 (dd, 2H, J = 5.9 Hz, 1.8 Hz), 3.86 (s, 3H), 4.80 - 4.87 (m, 1H), 4.96 (brs, 1H), 6.33 (dd, 1H, J = 7.7 Hz, 1.5 Hz), 6.59 - 6.77 (m, 3H), 6.82 - 6.87 (m, 1H), 6.93 - 7.02 (m, 2H), 7.33 - 7.37 (m, 2H); HRMS (EI) m/z [M]+ calcd. for C22H24NO2F 353.1791, found: 353.1797.

Data for 3o; colorless oil. 0.24 g, 97%. 1H NMR (CDCl3, 500 MHz) δ 1.85 (s, 3H), 2.09 (s, 3H), 2.92 (dd, 1H, J = 15.4 Hz, 7.7 Hz), 2.97 (dd, 1H, J = 15.4 Hz, 7.7 Hz), 3.86 (s, 3H), 5.14 - 5.18 (m, 1H), 6.06 - 6.07 (m, 1H), 6.40 - 6.45 (m, 1H), 6.58 - 6.65 (m, 1H), 6.68 - 6.80 (m, 2H), 7.00 - 7.08 (m, 2H), 7.17 - 7.25 (m, 1H), 7.35 - 7.37 (m, 1H); HRMS (EI) m/z [M]+ calcd. for C20H22NO2F 327.1635, found: 327.1637.

Cyclopropanation of 3o using trimethyloxosulfonium iodide

To a stirred solution of 3o (0.243 g, 0.74 mmol) in DMSO (1 mL) was added trimethyloxosulfonium iodide (0.22 g, 1.0 mmol) and NaH (0.024 g, 1.0 mmol) at r.t. The mixture was stirred for 1 day and quenched with ice-water (20 mL). The mixture was extracted with ether, washed twice with water, and the organic layers were dried by Na2SO4. The solvent was removed at reduced pressure to give the product 4o as a white solid (0.21g, 65%).

Data for 4o; 1H NMR(CDCl3, 500MHz) δ 0.71 - 0.72 (m, 1H), 0.86 (s, 3H), 1.06 (s, 3H), 1.15 - 1.16 (m, 1H), 1.75 - 1.80 (m, 1H), 3.06 (dd, 1H, J = 16.1 Hz, 5.7 Hz), 3.09 (dd, 1H, J = 16.1 Hz, 5.7 Hz), 3.77 (s, 3H), 5.08 (brs, 1H), 5.16 (dd, 1H, J = 6.3 Hz, 6.3 Hz), 6.40 - 6.42 (m, 1H), 6.60 - 6.65 (m, 1H), 6.71 - 6.74 (m, 2H), 7.01 - 7.05 (m, 2H), 7.17 - 7.25 (m, 1H), 7.36 - 7.37 (m, 1H); HRMS (EI) m/z [M]+ calcd. for C21H24NO2F 341.1791, found: 341.1796.

Acknowledgements

We acknowledge JEOL Ltd. and Nihon Waters K. K. for measurement of HRMS.

Conflicts of Interest

The authors declare no conflicts of interest regarding the publication of this paper.

References

[1] Saranya, S., Harry, N.A., Krishman, K.K. and Anilkumar, G. (2018) Developments and Perspectives in the Asymmetric Mannich Reaction. Asian Journal of Organic Chemistry, 7, 613-633.
https://doi.org/10.1002/ajoc.201700679
[2] Bala, S., Sharma, N., Kajal, A., Kamboj, S. and Saini, V. (2014) Mannich Bases: An Important Pharmacophore in Present Scenario. International Journal of Medicinal Chemistry, 2014, Article ID: 191072.
https://doi.org/10.1155/2014/191072
[3] Karimi, B., Enders, D. and Jafari, E. (2013) Recent Advances in Metal-Catalyzed Asymmetric Mannich Reactions. Synthesis, 45, 2769-2812.
https://doi.org/10.1055/s-0033-1339479
[4] Subramaniapillai, S.G. (2013) Mannich Reaction: A Versatile and Convenient Approach to Bioactive Skeletons. Journal of Chemical Science, 125, 467-482.
https://doi.org/10.1007/s12039-013-0405-y
[5] Córdova, A. (2004) The Direct Catalytic Asymmetric Mannich Reaction. Accounts of Chemical Research, 37, 102-112.
https://doi.org/10.1021/ar030231l
[6] Roman, G. (2015) Mannich Bases in Medicinal Chemistry and Drug Design. European Journal of Medicinal Chemistry, 89, 743-816.
https://doi.org/10.1016/j.ejmech.2014.10.076
[7] Boger, D.L. and Weinreb, S.N. (2012) In Hetero-Diels-Alder Methodology in Organic Synthesis. Academic Press, Inc., London.
[8] Buonora, P., Olsen, J.-C. and Oh, T. (2001) Recent Developments in Imino Diels-Alder Reactions. Tetrahedron, 57, 6099-6138.
https://doi.org/10.1016/S0040-4020(01)00438-0
[9] Kumatabara, Y., Kaneko, S., Nakata, S., Shirakawa, S. and Maruoka, K. (2016) Hydrogen-Bonding Catalysis of Tetraalkylammonium Salts in an Aza-Diels-Alder Reaction. Chemistry: An Asian Jour-nal, 11, 2126-2129.
https://doi.org/10.1002/asia.201600781
[10] Shang, D., Xin, J., Liu, Y., Zhou, X., Liu, X. and Feng, X. (2008) Enantioselective Aza-Diels-Alder Reaction of Aldimine with “Danishefsky-Type Diene” Catalyzed by Chiral Scandium(III)-N,N’-Dioxide Complexes. Journal of Organic Chemistry, 73, 630-637.
https://doi.org/10.1021/jo7021263
[11] Itoh, J., Fuchibe, K. and Akiyama, T. (2006) Chiral Bronsted Acid Catalyzed Enantioselective Aza-Diels-Alder Reaction of Brassard’s Doeme with Imines. Angewandte Chemie International Edition, 45, 4796-4798.
https://doi.org/10.1002/anie.200601345
[12] Sundén, H., Ibrahem, I., Eriksson, L. and Córdova, A. (2005) Direct Catalytic Enantioselective Aza-Diels-Alder Reactions. Angewandte Chemie International Edition, 44, 4877-4880.
https://doi.org/10.1002/anie.200500811
[13] Loncaric, C., Manabe, K. and Kobayashi, S. (2003) AgOTf-Catalyzed Aza-Diels-Alder Reaction of Danishefsky’s Diene with Imines in Water. Advanced Synthesis and Catalysis, 345, 475-477.
https://doi.org/10.1002/adsc.200390052
[14] Mancheno, O.G., Arrayás, R.G. and Carretero, J.C. (2004) Chiral Copper Complexes of Phosphino Sulfenyl Ferrocenes as Efficient Catalysts for Enantioselective Formal Aza Diels-Alder Reaction of N-Sulfonyl Imines. Journal of American Chemical Society, 126, 456-457.
https://doi.org/10.1021/ja038494y
[15] Kobayashi, S., Kusakabe, K. and Ishitani, H. (2000) Chiral Catalyst Optimization Using Both Solid-Phase and Liquid-Phase Methods in Asymmetric Aza Diels-Alder Reactions. Organic Letters, 2, 1225-1227.
https://doi.org/10.1021/ol005656b
[16] Girling, P.R., Kiyoi, T. and Whiting, A. (2011) Mannich-Michael versus Formal Aza-Diels-Alder Approaches to Piperidone Derivatives. Organic and Biomolecular Chemistry, 9, 3105-3121.
https://doi.org/10.1039/c0ob00996b
[17] Waldmann, H. and Braun, M. (1992) Asymmetric Tandem Mannich Reactions of Amino Acid Ester Imines with Danishefsky’s Diene. Journal of Organic Chemistry, 57, 4444-4451.
https://doi.org/10.1021/jo00042a027
[18] Kunz, H. and Pfrengle, W. (1989) Carbohydrates as Chiral Templates: Stere-oselective Tandem Mannich-Michael Reactions for the Synthesis of Piperidine Alkaloids. Angewandte Chemie International Edition, 28, 1067-1068.
https://doi.org/10.1002/anie.198910671
[19] Ishimaru, K. and Kojima, T. (2000) A Novel Approach for Mannich-Type Bases Having a Terminal Olefin: Zinc Triflate and Water-Promoted Cyclization/C-N Bond Cleavage Process. Journal of Organic Chemistry, 65, 8395-8398.
https://doi.org/10.1021/jo0011888
[20] Ishimaru, K. and Kojima, T. (2003) Stereoselective Mannih-Type Reaction of Chiral Aldimines with 2-Silyloxybutadienes by Using Trifluoromethanesulfonic Acid. Tetrahedron Letters, 44, 5441-5444.
https://doi.org/10.1016/S0040-4039(03)01314-5
[21] Ishimaru, K. and Kojima, T. (2003) Addition of Water to Zinc Triflate Promotes a Novel Reaction: Stereoselective Mannich-Type Reaction of Chiral Aldimines with 2-Silyloxybutadiene. Journal of Organic Chemistry, 68, 4959-4962.
https://doi.org/10.1021/jo0300338
[22] Ishimaru, K. and Kojima, T. (2001) Stereoselective Synthesis of Mannich-Type Products Having a Terminal Olefin by Use of Benzaldiminetricarbonylchromium Derivatives. Tetrahedron Letters, 42, 5037-5039.
https://doi.org/10.1016/S0040-4039(01)00912-1
[23] Adamo, I., Benedetti, F., Berti, F. and Campaner, P. (2006) Stereoselective Hydroazidation of Amino Enones: Synthesis of the Ritonavir/Lopinavir Core. Organic Letters, 8, 51-54.
https://doi.org/10.1021/ol0524104
[24] Edwards, M.L., Ritter, H.W., Stemerick, D.M. and Stewart, K.T. (1983) Mannich Bases of 4-phenyl-3-buten-2-one. A New Class of Antiherpes Agent. Journal of Medicinal Chemistry, 26, 431-436.
https://doi.org/10.1021/jm00357a020
[25] Dimmock, J.R., Nyathi, C.B. and Smith, P.J. (1979) Synthesis and Bioactivities of 1-(hydroxyphenyl)-I-nonen-3-ones and Related Ethers and Esters. Journal of Pharmaceutical Science, 68, 1216-1221.
https://doi.org/10.1002/jps.2600681006
[26] Dimmock, J.R., Nyathi, C.B. and Smith, P.J. (1978) Synthesis and Evaluation of 1-(hydroxyphenyl)-I-nonen-3-ones and Related Compounds for Antineoplastic and Antimicrobial Activities. Journal of Pharmaceutical Science, 67, 1543-1546.
https://doi.org/10.1002/jps.2600671113
[27] Roselló, M.S., Pozo, C. and Fustero, S. (2016) A Decade of Advance in the Asymmetric Vinylogous Mannich Reaction. Synthesis, 48, 2553-2571.
https://doi.org/10.1055/s-0035-1561650
[28] Lindemann, C. and Schneider, C. (2016) Quinolizine-Based Alkaloids: A General Catalytic, Highly Enantio and Diastereoselective Synthetic Approach. Synthesis, 48, 828-844.
https://doi.org/10.1055/s-0035-1561289
[29] Ye, J.-L., Zhang, Y.-F., Liu, Y., Zhang, J.-Y., Ruan, Y.-P. and Huang, P.-Q. (2015) Studies on the Asymmetric Synthesis of Pandamarliactonines: An Unexpected syn-Selective Vinylogous Mannich Reaction of N-tert-cutanesulfimines. Organic Chemistry Frontiers, 2, 697-704.
https://doi.org/10.1039/C5QO00098J
[30] Wang, Q., Gemmeren, M. and List, B. (2014) Asymmetric Disul-fonimide-Catalyzed Synthesis of d-amino-b-ketoester Derivatives by Vinylogous Mukaiyama-Mannich Reactions. Angewandte Chemie International Edition, 53, 13592-13595.
https://doi.org/10.1002/anie.201407532
[31] Liu, L.-J. and Liu, J.-T. (2014) 2-Chlorotetrafluoroethanesulfinamide Induced Asymmetric Vinylogous Mannich Reaction. Tetrahedron, 70, 1236-1245.
https://doi.org/10.1016/j.tet.2013.12.071
[32] Guo, Y., Zhang, Y., Qi, L., Tian, F. and Wang, L. (2014) Organocatalytic Direct Asymmetric Vinylogous Mannich Reactions of g-butenolides with Isatin-Derived Ketimines. RSC Advances, 4, 27286-27289.
https://doi.org/10.1039/c4ra04824e
[33] Sickert, M., Abels, F., Lang, M., Sieler, J., Birkemeyer, C. and Schneider, C. (2010) Bronsted Acid Catalyzed, Enantioselective Vinylogous Mannich Reaction. Chemistry: An European Journal, 16, 2806-2818.
https://doi.org/10.1002/chem.200902537
[34] Gu, C.-L., Liu, L., Wang, D. and Chen, Y.-J. (2009) Tunable and Highly Regio- and Diastereoselective Vinylogous Mannich-Type Reaction of Dioxinone-Derived Silyl Dienolate. Journal of Organic Chemistry, 74, 5754-5757.
https://doi.org/10.1021/jo900977y
[35] Sickert, M. and Schneider, C. (2008) The Enantioselective, Bronsted Acid Catalyzed, Vinylogous Mannich Reaction. Angewandte Chemie International Edition, 47, 3631-3634.
https://doi.org/10.1002/anie.200800103
[36] Martin, S.F. (2002) Evolution of the Vinylogous Mannich Reaction as a Key Construction for Alkaloid Synthesis. Accounts of Chemical Research, 35, 895-904.
[37] Kawecki, R. (2006) Aza Diels-Alder Re-actions of Sulfinimines with the Rawal Diene. Tetrahedron: Asymmetry, 17, 1420-1423.
https://doi.org/10.1016/j.tetasy.2006.04.026
[38] Xiong, Y., Mei, H., Han, J., Li, G. and Pan, Y. (2014) Asymmetric Synthesis of β’-Amino-α,β-Enones via Addition of α,β-Unsaturated Ketone-Derived Enolates to Chiral N-Phosphonyl Imines. Tetrahedron Letters, 55, 2476-2479.
https://doi.org/10.1016/j.tetlet.2014.03.005
[39] Reddy, A.A., Reddy, P.O. and Prasad, K.R. (2016) Synthesis of β-Amino-Substituted Enones by Addition of Substituted Methyl Enones to Sulfinimines: Application to the Total Synthesis of Alkaloids (+)-Lasubine II and (+)-241D and the Formal Total Synthesis of (-)-Lasubine I. Journal of Organic Chemistry, 81, 11363-11371.
https://doi.org/10.1021/acs.joc.6b01541
[40] Yuan, Y., Li, X. and Ding, K. (2002) Acid-Free Aza Diels-Alder Reaction of Danishefsky’s Diene with Imines. Organic Letters, 4, 3309-3311.
https://doi.org/10.1021/ol0265822
[41] Pan, X., Huang, R., Zhang, J., Ding, L., Li, W., Zhang, Q. and Liu, F. (2012) Efficient Synthesis of Prasugrel, a Novel P2Y12 Receptor Inhibitor. Tetrahedron Letters, 53, 5364-5366.
https://doi.org/10.1016/j.tetlet.2012.07.071
[42] Jakubowski, J.A., Winters, K.J., Naganuma, H. and Wallentin, L. (2007) Prasugrel: A Novel Thienopyridine Antiplatelet Agent. A Review of Preclinical and Clinical Studies and the Mechanistic Bases for Its Distinct Antiplatelet Profile. Cardiovascular Drug Reviews, 25, 357-374.
https://doi.org/10.1111/j.1527-3466.2007.00027.x

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.