Next Article in Journal
Synthesis of Chalcones with Anticancer Activities
Next Article in Special Issue
Catalytic Asymmetric Hydrogenation of 3-Substituted Benzisoxazoles
Previous Article in Journal
Chemical Constituents of the Ethyl Acetate Extract of Belamcanda chinensis (L.) DC Roots and Their Antitumor Activities
Previous Article in Special Issue
Chiral Aminophosphines as Catalysts for Enantioselective Double-Michael Indoline Syntheses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Chiral Bis(Imidazolidine)Pyridine-Cu Complex-Catalyzed Enantioselective [3+2]-Cycloaddition of Azomethine Imines with Propiolates

Department of Chemistry, Graduate School of Science, Chiba University, 1-33, Yayoi, Inage, Chiba 263-8522, Japan
*
Author to whom correspondence should be addressed.
Molecules 2012, 17(5), 6170-6178; https://doi.org/10.3390/molecules17056170
Submission received: 27 April 2012 / Revised: 18 May 2012 / Accepted: 21 May 2012 / Published: 24 May 2012
(This article belongs to the Special Issue Asymmetric Catalysis)

Abstract

:
[3+2] Cycloaddition of azomethine imines with electron-deficient terminal alkynes was smoothly catalyzed by a chiral bis(imidazolidine)pyridine-CuOAc complex to give bicyclic pyrazolo[1,2-a]pyrazolone derivatives with up to 74% ee.

Graphical Abstract

1. Introduction

Highly functionalized complex molecules are key tools for promoting biochemical research and developing pharmaceuticals, because the network of chiralities, the positions of heteroatoms, and the direction of lone-pairs in the molecules correlate strictly with their biological activities. The unique pyrazolone heterocycles, five-membered-ring lactams containing N-N conjunction, are examples of highly functionalized complex molecules, which have been used for the development of fine chemicals (e.g., dyes, textile, photography, and cosmetics) and the pharmaceuticals (Figure 1) [1,2,3]. Anti-inflammatory activity has been studied on phenylbutazones [4,5,6]. BW357U shows anorectic activity [7]. Inhibitory activities of lipoxygenase and cyclooxygenase are also reported on phenidone and BW755C [8,9]. More complex bicyclic pyrazolo[1,2-a]pyrazolone derivatives exhibit diverse bioactivities, and utilized as antibacterial agents, herbicides, pesticides, antitumor agents, calcitonin agonists, and potent drugs to relieve Alzheimer’s disease. LY173013 and LY186826 are the representative examples of antibacterial agents, which have been developed as the analogues of penicillin and cephalosporin antibiotics [10,11,12,13,14].
Figure 1. Representative examples of biologically active pyrazolone heterocycles.
Figure 1. Representative examples of biologically active pyrazolone heterocycles.
Molecules 17 06170 g001
For accessing the fascinating bicyclic pyrazolo[1,2-a]pyrazolone skeleton, the [3+2] cycloaddition of azomethine imines has apparent advantage for the straightforward construction of the core structure (Scheme 1) [15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]. Due to the importance of these molecules on the biological study, the synthesis of these molecules in chiral form is highly demanded. In 2003, Fu et al. succeeded in the first catalytic asymmetric cyclization of azomethine imines and terminal alkynes using phosphaferrocene-oxazoline-CuI catalyst [30]. After the pioneering work of Fu’s group [30,31], several remarkable catalytic asymmetric [3+2] cyclizations using azomethine imines were reported by Sibi [32], Maruoka [33], Suga [34], and Chen [35,36].
Scheme 1. General scheme of [3+2] cycloaddition of azomethine imines.
Scheme 1. General scheme of [3+2] cycloaddition of azomethine imines.
Molecules 17 06170 g002
In 2010 we succeeded in the development of bis(imidazolidine)pyridine, abbreviated as PyBidine, for the catalytic asymmetric [3+2] cycloaddition of iminoesters with nitroalkene (Scheme 2) [37]. We envisioned that the new chiral PyBidine ligand would have wide potential to promote various asymmetric reactions, and the [3+2] cycloaddition using azomethine imines was set to study for constructing the fascinating bicyclic pyrazolo[1,2-a]pyrazolones in this report.
Scheme 2. Asymmetric [3+2] cycloaddition catalyzed by PyBidine-Cu(OTf)2 complex.
Scheme 2. Asymmetric [3+2] cycloaddition catalyzed by PyBidine-Cu(OTf)2 complex.
Molecules 17 06170 g003

2. Results and Discussion

The study started with an exploration of the most appropriate metal salts for the [3+2] cycloaddition of azomethine imines with ethyl propiolate (Table 1).
Table 1. Catalyst activity of PyBidine-metal salt complex for [3+2] cycloaddition of azomethine imine with ethyl propiolate. Molecules 17 06170 i001
Table 1. Catalyst activity of PyBidine-metal salt complex for [3+2] cycloaddition of azomethine imine with ethyl propiolate. Molecules 17 06170 i001
EntryMetal saltTime (h)Yield (%)ee (%)
1CuI39944
2CuBr58060
3CuOAc29960
4Cu(OAc)239946
5Cu(OTf)2246921
6Co(OAc)231trace-
7Ni(OAc)232trace-
8Zn(OAc)250trace-
As reported in the former papers [30,31], Cu(I) salt showed good catalytic activity, and the PyBidine-CuI gave the product in 99% yield with 44% ee (rt, 3h). The PyBidine-CuOAc complex improved catalyst activity to give the product in 99% yield with 60% ee (entry 3). It was noteworthy that the Cu(II) salts also had catalyst activity, although the stereoselectivities of the product were reduced [27]. Because the other PyBidine-metal complex derived from the divalent acetate salts of the first row transition metals (Ni(OAc)2, Co(OAc)2, and Zn(OAc)2) gives traces of product, the copper element has unique and essential role for the [3+2] cycloaddition using azomethine imines. In all cases we examined, (S, S, S, S)-PyBidine-Cu catalyst gave the (R)-enriched product.
Next, the reaction conditions were optimized for the PyBidine-CuOAc catalyzed reaction (Table 2). The solvent effects were examined in the reaction using methyl propiolate. Various solvents (CHCl3, CH2Cl2, PhMe, THF, dioxane, EtOH) can be utilized to give the compound with good chemical yield, though the use of MeCN resulted in low yield. Among the solvents we examined, the reaction in CH2Cl2 gave the product with the best stereoselectivity (entry 2). 3,3-Dimethylpyrazolidinone also transformed to the bicyclic pyrazolo[1,2-a]pyrazolone in CH2Cl2 with 53% ee (entry 8). The azomethine imines prepared from substituted 4-iodo- or 3-chlorobenzaldehyde were examined in entries 11–14. A lower reaction temperature is helpful for improving the enantiomeric excesses in some cases (entries 9–10, and 13–14), though the reaction time was prolonged. To accelerate the reaction at low temperature, the addition of bases was also examined. The addition of DIPEA significantly accelerated the reaction to give the product with slightly improved enantioselectivity (entries 9 and 15), although the addition of pyridine or K2CO3 resulted in low stereoselectivity. For the ethyl propiolate, when the reaction was performed at −20 °C with DIPEA, the PyBidine-CuOAc gave the product with 74% ee (entry 19).
Table 2. PyBidine-CuOAc catalyzed [3+2] cycloaddition of azomethine imine with methyl propiolate a. Molecules 17 06170 i002
Table 2. PyBidine-CuOAc catalyzed [3+2] cycloaddition of azomethine imine with methyl propiolate a. Molecules 17 06170 i002
EntryArR1R2Base bSolventTemp (°C)Time (h)Yield (%)ee (%)
1PhHMe-CHCl3rt246655
2PhHMe-CH2Cl2rt49567
3PhHMe-PhMert2.59956
4PhHMe-MeCNrt273952
5PhHMe-THFrt249661
6PhHMe-dioxanert249535
7PhHMe-EtOHrt288844
8PhMeMe-CH2Cl2rt246753
9PhHEt-CH2Cl2rt29960
10PhHEt-CH2Cl2−10459765
114-IC6H4-HEt-CH2Cl2rt29959
124-IC6H4-HEt-CH2Cl2−10235854
133-ClC6H4-HEt-CH2Cl2rt38847
143-ClC6H4-HEt-CH2Cl2−10236262
15PhHEtDIPEACH2Cl2rt19970
16PhHEtPyridineCH2Cl2rt19940
17PhHEtK2CO3CH2Cl2rt169032
18PhHEtDIPEACH2Cl20257872
19PhHEtDIPEACH2Cl2−20586874
a The absolute configurations of the products are provided by analogy with the products reported in ref. [30]; b 0.5 eq. of base were used to azomethine imine.
Although the enantiomeric excesses of the products are moderate, the reaction mechanism is assumed to be as summarized in Scheme 3. As proposed in former reports, the formation of copper acetylide would be a key for promoting the catalyst, though the Cu(II) salts also showed significant catalyst activity in our study. When the copper acetylide is generated on the (S,S,S,S)-PyBidine-Cu catalyst, the approach of (Z)-azomethine imine from an upper-side is considered on the reaction sphere. The approach of (Z)-azomethine imine to the Cu-acetylide would be assisted using the Lewis acidic apical cite of the Cu center. In a model A, a phenyl-ring of the PyBidine (emphasized by purple color) would have a steric repulsion with the aromatic ring of the (Z)-azomethine imine. To avoid the repulsion, the reaction would proceed via a model B to give the (R)-bicyclic pyrazolo[1,2-a]pyrazolone.
Scheme 3. Plausible mechanism of asymmetric [3+2] cycloaddition catalyzed by PyBidine-CuOAc complex.
Scheme 3. Plausible mechanism of asymmetric [3+2] cycloaddition catalyzed by PyBidine-CuOAc complex.
Molecules 17 06170 g004

3. Experimental

3.1. General

Infrared (IR) spectra were recorded on a JASCO FT/IR-4100 Fourier transform infrared spectrophotometer. NMR spectra were recorded on a JEOL JNM-ECA500 or JEOL JNM-ECA400 spectrometer, operating at 500 MHz or 400 MHz for 1H-NMR and 125 MHz or 100 MHz for 13C-NMR. Chemical shifts in CDCl3 were reported downfield from TMS (=0 ppm) for 1H-NMR. For 13C-NMR, chemical shifts were reported downfield from TMS (=0 ppm) or in the scale relative to CHCl3 (77.00 ppm for 13C-NMR) as an internal reference. Optical rotations were measured on a JASCO P-1020 Polarimeter. The enantiomeric excess (ee) was determined by HPLC analysis. Column, DAICEL CHIRALPAK AD-H; mobile phase, hexane-i-PrOH; flow rate, 1.0 mL/min. General experimental details for synthesis of PyBidine ligand see reference in the manuscript [37]. Azomethine imine were synthesized according to known procedure [30].

3.2. General Procedure

PyBidine (0.0075 mmol) and CuOAc (0.00825 mmol) were added to a round flask containing a stir bar under Ar. Next CH2Cl2 (1 mL) was added to the flask and the mixture was stirred over 12 h. To the resulting solution, azomethine imine (0.15 mmol) were added, followed by ethyl propiolate (0.18 mmol) that was added dropwise at rt. After being stirred for an appropriate time, the reaction mixture was purified by silica gel column chromatography (Et2O/hexane = 2/1 to Et2O) to afford the corresponding cycloadduct. The enantiomeric excess of the products was determined by chiral stationary-phase HPLC analysis (Daicel CHIRALPAK AD-H, i-PrOH/hexane 20/80, flow rate 1.0 mL/min).
(R)-Ethyl 5-oxo-1-phenyl-1,5,6,7-tetrahydropyrazolo[1,2-a]pyrazole-2-carboxylate. Yellow oil; FTIR (neat) 3,403, 3,087, 3,063, 3,031, 2,980, 2,928, 2,853, 1,699, 1,602, 1,551, 1,499, 1,455, 1,433, 1,414, 1,388, 1,364, 1,321, 1,300, 1,251, 1,206, 1,173, 1,104, 1,077, 1,039, 1,027, 986, 914, 874, 842, 819, 765, 752, 725, 700, 663 cm−1; 1H-NMR (CDCl3) δ 7.54 (d, J = 0.5 Hz, 1H), 7.44–7.25 (m, 5H), 5.15 (d, J = 1.7 Hz, 1H), 4.15–3.95 (m, 2H) 3.41–3.36 (m, 1H) 3.08–3.00 (m, 1H) 2.95–2.83 (m, 1H), 2.79–2.70 (m, 1H), 1.13 (t, J = 7.2 Hz, 3H); 13C-NMR (CDCl3) δ 164.7, 163.4, 138.2, 128.5, 128.4, 128.3, 128.2, 118.5, 73.3, 60.4, 51.6, 35.7, 14.0; HRMS [ESI] Calcd. for C15H16N2O3Na+ [M+Na]+: 295.1053, found 295.1048; HPLC analysis: tR 10.7 min (major, 87%) and 13.1 min (minor, 13%), detection at 330 nm]; [α]D24 −207 (c 1.01, CHCl3).
(R)-Ethyl 1-(3-chlorophenyl)-5-oxo-1,5,6,7-tetrahydropyrazolo[1,2-a]pyrazole-2-carboxylate. Yellow oil; FTIR (neat) 3,403, 3,086, 2,981, 2,927, 2,851, 1,697, 1,599, 1,474, 1,433, 1,387, 1,363, 1,319, 1,248, 1,198, 1,171, 1,105, 1,055, 1,024, 986, 934, 883, 842, 820, 785, 763, 735, 718, 693, 663 cm−1; 1H-NMR (CDCl3) δ 7.53 (s, 1H), 7.42 (s, 1H), 7.30 (d, J = 1.4 Hz, 3H), 5.12 (d, J = 1.8 Hz, 1H), 4.15–3.95 (m, 2H), 3.50–3.45 (m, 1H), 3.09–3.02 (m, 1H), 2.98–2.89 (m, 1H), 2.78–2.72 (m, 1H), 1.16 (t, J =7.2 Hz, 3H); 13C-NMR (CDCl3) δ 164.7, 163.2, 140.9, 134.4, 129.6, 128.7, 128.5, 128.3, 126.5, 73.0, 60.5, 52.1, 35.5, 14.0; HRMS [ESI] calcd for C15H16N2O3Cl+ [M+H]+: 307.0844, found 307.0839; HPLC analysis: tR 9.1 min (major, 81%) and 11.1 min (minor, 19%), detection at 330 nm]; [α]D24 −362 (c 1.00, CHCl3).
(R)-Ethyl 1-(4-iodophenyl)-5-oxo-1,5,6,7-tetrahydropyrazolo[1,2-a]pyrazole-2-carboxylate. Yellow oil; FTIR (neat) 3,403, 3,084, 2,981, 2,925, 2,851, 2,309, 1,696, 1,604, 1,509, 1,434, 1,388, 1,364, 1,319, 1,250, 1,171, 1,105, 1,026, 986, 921, 849, 796, 758, 732, 664 cm−1; 1H-NMR (CDCl3) δ 7.53–7.52 (m, 1H), 7.40–7.36 (m, 2H), 7.09–7.03 (m, 2H), 5.14 (d, J = 1.6 Hz, 1H), 4.13–4.01 (m, 2H) 3.46–3.41 (m, 1H), 3.08–3.01 (m, 1H), 2.97–2.90 (m, 1H), 2.79–2.72 (m, 1H), 1.15 (t, J = 7.2, 3H); 13C-NMR (CDCl3) δ 164.6, 163.3, 134.4, 129.86, 129.79, 128.5, 118.1, 115.5, 115.3, 72.8, 60.4, 51.8, 35.6, 14.0; HPLC analysis: tR 9.4 min (major, 77%) and 13.2 min (minor, 23%), detection at 330 nm]; [α]D24 −290 (c 1.00, CHCl3).
(R)-Methyl 5-oxo-1-phenyl-1,5,6,7-tetrahydropyrazolo[1,2-a]pyrazole-2-carboxylate. Yellow solid; FTIR (neat) 3,373, 3,083, 2,921, 2,862, 1,995, 1,714, 1,604, 1,496, 1,456, 1,407, 1,305, 1,246, 1,210, 1,167, 1,105, 958, 910, 836, 808, 771, 753, 735, 700 cm−1; 1H-NMR (CDCl3) δ 7.53 (brs, 1H), 7.38–7.31 (m, 5H), 5.17 (d, J = 1.6 Hz, 1H), 3.62 (s, 3H), 3.41–3.36 (m, 1H), 3.08–3.01 (m, 1H), 2.94–2.85 (m, 1H), 2.79–2.72 (m, 1H); 13C-NMR (CDCl3) δ 164.7, 163.8, 138.1, 128.6, 128.5, 128.2, 117.9, 73.1, 51.5, 35.7; HRMS [ESI] Calcd for C14H15N2O3+ [M+H]+: 259.1077, Found 259.1073; HPLC analysis: tR 11.3 min (major, 84%) and 12.7 min (minor, 16%), detection at 330 nm]; [α]D24 −296 (c 1.00, CHCl3).
(R)-Methyl 7,7-dimethyl-5-oxo-1-phenyl-1,5,6,7-tetrahydropyrazolo[1,2-a]pyrazole-2-carboxylate. Yellow solid; FTIR (neat) 3,414, 3,076, 3,030, 2,950, 1,695, 1,600, 1,495, 1,455, 1,385, 1,322, 1,256, 1,227, 1,120, 1,102, 1,042, 1,028, 1,006, 962, 913, 888, 847, 803, 755, 715, 698 cm−1; 1H-NMR (CDCl3) δ 7.52 (d, J = 1.6 Hz, 1H), 7.46–7.26 (m, 5H), 5.46 (brs, 1H), 3.62 (s, 3H), 2.87 (d, J = 15.6 Hz, 1H), 2.40 (d, J = 15.9 Hz, 1H), 1.24 (s, 3H), 1.15 (s, 1H); 13C-NMR (CDCl3) δ 166.6, 164.1, 142.0, 129.4, 128.4, 127.9, 127.8, 116.8, 64.51, 64.46, 51.4, 49.3, 24.9, 18.9; HPLC analysis (i-PrOH/hexane 10/90, flow rate 1.1 mL/min: tR 9.0 min (minor, 24%) and 9.8 min (major, 76%), detection at 330 nm.

4. Conclusions

In conclusion, we have examined the application of PyBidine-Cu(OAc) complex for the [3+2] cycloaddition of azomethine ylide and propiolate for the construction the bicyclic pyrazolo[1,2-a]pyrazolone skeletons with 74% ee. This shows the broad utility of the PyBidine ligand for various metal-catalyzed reactions. In addition to a detailed study on the reaction mechanism, the extensive application to other catalytic reactions based on the unique functions of PyBidine is in progress.

Acknowledgments

This work was supported by a Grant-in-Aid for Scientific Research from the Ministry of Education, Culture, Sports, Science and Technology, Japan, and the Takeda Science Foundation.

References and Notes

  1. Eicher, T.; Hauptmann, S. The Chemistry of Heterocycles, 2nd ed; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar]
  2. Varvounis, G.; Fiamegos, Y.; Pilidis, G. Pyrazol-3-ones part 1: Synthesis and applications. Adv. Heterocycl. Chem. 2001, 80, 73–156. [Google Scholar] [CrossRef]
  3. Elguero, J. Pyrazoles: Comprehensive Heterocyclic Chemistry II; Katritzky, A.R., Rees, C.W., Scriven, E.F.V., Eds.; Elsevier: Oxford, UK, 1996; Volume 3, pp. 1–75. [Google Scholar]
  4. Marchand-Brynaert, J.; Ghosez, L. Non β-Lactam Analogs of Penicillins and Cephalosporins. In Recent Progress in the Chemical Synthesis of Antibiotics; Lukacs, G., Ohno, M., Eds.; Springer: Berlin, Germany, 1990; pp. 727–794. [Google Scholar]
  5. Hanessian, S.; McNaughton-Smith, G.; Lombart, H.-G.; Lubell, W.D. Design and synthesis of conformationally constrained amino acids as versatile scaffolds and peptide mimetics. Tetrahedron 1997, 53, 12789–12854. [Google Scholar] [CrossRef]
  6. Konaklieva, M.I.; Plotkin, B.J. Bioisosters of β-Lactams as Anti-Infectives. Curr. Med. Chem. Anti-Infect. Agents 2003, 2, 287–302. [Google Scholar] [CrossRef]
  7. White, H.L.; Howard, J.L.; Cooper, B.R.; Soroko, F.E.; McDermed, J.D.; Ingold, K.J.; Maxwell, R.A. A Novel inhibitor of gamma-aminobutyrate aminotransferase with anorectic activity. J. Neurochem. 1982, 39, 271–273. [Google Scholar] [CrossRef]
  8. Cucurou, C.; Battioni, J.P.; Thang, D.C.; Nam, N.H.; Mansuy, D. Mechanisms of inactivation of lipoxygenases by phenidone and BW755C. Biochemistry 1991, 30, 8964–8970. [Google Scholar] [CrossRef]
  9. Kawai, K.; Hasegawa, J.; Shiojiri, H.; Nozawa, Y.; Tsurumi, K.; Fujimura, H. Interaction of BW755C, a potent inhibitor of lipoxygenase and cyclo-oxygenase, with mitochondrial cytochrome oxidase. Experientia 1985, 41, 490–492. [Google Scholar] [CrossRef]
  10. Jungheim, L.N.; Sigmund, S.K.; Fisher, J.W. Bicyclic pyrazolidinones, a new class of antibacterial agent based on the β-lactam model. Tetrahedron Lett. 1987, 28, 285–288. [Google Scholar] [CrossRef]
  11. Jungheim, L.N.; Sigmund, S.K. 1,3-Dipolar cycloaddition reactions of pyrazolidinium ylides with acetylenes. Synthesis of a new class of antibacterial agents. J. Org. Chem. 1987, 52, 4007–4013. [Google Scholar] [CrossRef]
  12. Ternansky, R.J.; Draheim, S.E.; Pike, A.J.; Counter, F.T.; Eudaly, J.A.; Kasher, J.S. Structure-activity relationship within a series of pyrazolidinone antibacterial agents. 2. Effect of side-chain modification on in vitro activity and pharmacokinetic parameters. J. Med. Chem. 1993, 36, 3224–3229. [Google Scholar] [CrossRef]
  13. Claramunt, R.M.; Elguero, J. The chemistry of pyrazolidinones. A review. Org. Prep. Proc. Int. 1991, 23, 273–320. [Google Scholar] [CrossRef]
  14. Kosower, E.M.; Radkowsky, A.E.; Fairlamb, A.H.; Croft, S.L.; Nea, R.A. Bimane cyclic esters, possible stereologues of trypanothione as antitrypanosomal agents. Bimanes 29. Eur. J. Med. Chem. 1995, 30, 659–671. [Google Scholar] [CrossRef]
  15. Karlsson, S.; Hogberg, H.-E. Asymmetric 1,3-Dipolar Cycloadditions for the Construction of Enantiomerically Pure Heterocycles. A Review. Org. Prep. Proc. Int. 2001, 33, 103–172. [Google Scholar] [CrossRef]
  16. Gothelf, K.V.; Jørgensen, K.A. Asymmetric 1,3-Dipolar cycloaddition reactions. Chem. Rev. 1998, 98, 863–909. [Google Scholar]
  17. Kanemasa, S. Cornerstone works for catalytic 1,3-Dipolar cycloaddition reactions. Heterocycles 2010, 82, 87–200. [Google Scholar] [CrossRef]
  18. Taylor, E.C.; Haley, N.F.; Clemens, R.J. Synthesis and Properties of 3-Oxo-1,2-diazetidinium Ylides. J. Am. Chem. Soc 1981, 103, 7743–7752. [Google Scholar] [CrossRef]
  19. Dorn, H. 1,3-Dipolar cycloaddition of pyrazolid-3-one azomethinimines and E-β-nitrostyrene—elucidation of the generation of seemingly “non cisoid” adducts. Tetrahedron Lett. 1985, 26, 5123–5126. [Google Scholar] [CrossRef]
  20. Huisgen, R.; Weinberger, R. Are any non-stereospecific 1,3-Dipolar cycloaddition known? A revision. Tetrahedron Lett. 1985, 26, 5119–5122. [Google Scholar] [CrossRef]
  21. Keller, M.; Sido, A.S.S.; Pale, P.; Sommer, J. Copper(I) Zeolites as Heterogeneous and Ligand-Free Catalysts: [3+2] Cycloaddition of Azomethine Imines. Chem. Eur. J. 2009, 15, 2810–2817. [Google Scholar] [CrossRef]
  22. Pezdirc, L.; Stanovnik, B.; Svete, J. Copper(I) Iodide-Catalyzed Cycloadditions of (1Z, 4R*,5R*)-4-Benzamido-5-phenylpyrazolidin-3-on-1-azomethine Imines to Ethyl Propiolate. Aust. J. Chem. 2009, 62, 1661–1666. [Google Scholar] [CrossRef]
  23. Yamashita, Y.; Kobayashi, S. NO cations as highly efficient catalysts for carbon-carbon bond forming reactions. Chem. Lett. 2009, 38, 678–679. [Google Scholar] [CrossRef]
  24. Shi, F.; Mancuso, R.; Larock, R.C. 1,3-Dipolar cycloaddition of arynes with azomethine imines: synthesis of 1,2-dihydropyrazolo[1,2-a]indazol-3(9H)-ones. Tetrahedron. Lett. 2009, 50, 4067–4070. [Google Scholar] [CrossRef]
  25. Oishi, T.; Yoshimura, K.; Yamaguchi, K.; Mizuno, N. An efficient copper-mediated 1,3-Dipolar Cycloaddition of pyrazolidinone-based dipoles to terminal alkynes to produce N,N-Bicyclic pyrazolidinone derivatives. Chem. Lett. 2010, 39, 1086–1087. [Google Scholar] [CrossRef]
  26. Na, R.; Jing, C.; Xu, Q.; Jiang, H.; Wu, X.; Shi, J.; Zhong, J.; Wang, M.; Benitez, D.; Tkatchouk, E.; et al. Phosphine-Catalyzed Annulations of Azomethine Imines: Allene-Dependent [3+2], [3+3], [4+3], and [3+2+3] Pathways. J. Am. Chem. Soc. 2011, 133, 13337–13348. [Google Scholar]
  27. Yoshimura, K.; Oishi, T.; Yamaguchi, K.; Mizuno, N. An efficient, ligand-free, heterogeneous supported copper hydroxide catalyst for the synthesis of N,N-Bicylic pyrazolidinone derivatives. Chem. Eur. J. 2011, 17, 3827–3831. [Google Scholar] [CrossRef]
  28. Xin, Y.; Zhao, J.; Gu, J.; Zhu, S. Catalyst free 1,3-dipolar cycloaddition of 3-oxo-1,2-pyrazolidinium ylides to β-trifluoroacetyl vinyl ethyl ether: Synthesis of 6-trifluoroacetyl substituted bicyclic pyrazolidinones. J. Fluorine Chem. 2011, 132, 402–408. [Google Scholar] [CrossRef]
  29. Koptelov, Y.B.; Sednev, M.V. Stable azomethine imines derived from pyrazolidin-3-one and their cycloaddition to N-Arylmaleimides. Russ. J. Org. Chem. 2011, 47, 547–555. [Google Scholar] [CrossRef]
  30. Shintani, R.; Fu, G.C. A new copper-catalyzed [3+2] cycloaddition: Enantioselective coupling of terminal alkynes with azomethine imines to generate five-membered nitrogen heterocycles. J. Am. Chem. Soc. 2003, 125, 10778–10779. [Google Scholar] [CrossRef]
  31. Suarez, A.; Downey, C.W.; Fu, G.C. Kinetic resolutions of azomethine imines via copper-catalyzed [3+2] cycloadditions. J. Am. Chem. Soc. 2005, 127, 11244–11245. [Google Scholar] [CrossRef]
  32. Sibi, M.P.; Rane, D.; Stanley, L.M.; Soeta, T. Copper(II)-Catalyzed Exo and enantioselective cycloadditions of azomethine imines. Org. Lett. 2008, 10, 2971–2974. [Google Scholar] [CrossRef]
  33. Hashimoto, T.; Maeda, Y.; Omote, M.; Nakatsu, H.; Maruoka, K. Catalytic enantioselective 1,3-Dipolar cycloaddition of C,N-Cyclic azomethine imines with α,β-Unsaturated aldehydes. J. Am. Chem. Soc. 2010, 132, 4076–4077. [Google Scholar] [CrossRef]
  34. Suga, H.; Funyu, A.; Kakehi, A. Highly enantioselective and diastereoselective 1,3-Dipolar cycloaddition reactions between azomethine imines and 3-Acryloyl-2-oxazolidinone catalyzed by Binaphthyldiimine-Ni(II) complexes. Org. Lett. 2007, 9, 97–100. [Google Scholar] [CrossRef]
  35. Chen, W.; Du, W.; Duan, Y.-Z.; Wu, Y.; Yang, S.-Y.; Chen, Y.-C. Enantioselective 1,3-Dipolar cycloaddition of cyclic enones catalyzed by multifunctional primary amines: Beneficial effects of hydrogen bonding. Angew. Chem. Int. Ed. 2007, 46, 7667–7670. [Google Scholar] [CrossRef]
  36. Chen, W.; Yuan, X.-H.; Li, R.; Du, W.; Wu, Y.; Ding, L.-S.; Chen, Y.-C. Organocatalytic and stereoselective [3+2] cycloadditions of azomethine imines with α,β-Unsaturated aldehydes. Adv. Synth. Catal. 2006, 348, 1818–1822. [Google Scholar] [CrossRef]
  37. Arai, T.; Mishiro, A.; Yokoyama, N.; Suzuki, K.; Sato, H. Chiral Bis(imidazolidine)pyridine-Cu(OTf)2: Catalytic asymmetric endo-selective [3+2] cycloaddition of imino esters with nitroalkenes. J. Am. Chem. Soc. 2010, 132, 5338–5339. [Google Scholar]

Share and Cite

MDPI and ACS Style

Arai, T.; Ogino, Y. Chiral Bis(Imidazolidine)Pyridine-Cu Complex-Catalyzed Enantioselective [3+2]-Cycloaddition of Azomethine Imines with Propiolates. Molecules 2012, 17, 6170-6178. https://doi.org/10.3390/molecules17056170

AMA Style

Arai T, Ogino Y. Chiral Bis(Imidazolidine)Pyridine-Cu Complex-Catalyzed Enantioselective [3+2]-Cycloaddition of Azomethine Imines with Propiolates. Molecules. 2012; 17(5):6170-6178. https://doi.org/10.3390/molecules17056170

Chicago/Turabian Style

Arai, Takayoshi, and Yuta Ogino. 2012. "Chiral Bis(Imidazolidine)Pyridine-Cu Complex-Catalyzed Enantioselective [3+2]-Cycloaddition of Azomethine Imines with Propiolates" Molecules 17, no. 5: 6170-6178. https://doi.org/10.3390/molecules17056170

Article Metrics

Back to TopTop