Next Article in Journal
Substitutions of Fluorine Atoms and Phenoxy Groups in the Synthesis of Quinoxaline 1,4-di-N-oxide Derivatives
Previous Article in Journal
Synthesis and Antiplasmodial Activity of 3-Furyl and 3-Thienylquinoxaline-2-carbonitrile 1,4-Di-N-oxide Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unexpected Reduction of Ethyl 3-Phenylquinoxaline-2- carboxylate 1,4-Di-N-oxide Derivatives by Amines

by
Lidia M. Lima
1,
Esther Vicente
1,
Beatriz Solano
1,
Silvia Pérez-Silanes
1,
Ignacio Aldana
1,* and
Antonio Monge
1
1
Unidad en Investigación y Desarrollo de Medicamentos, Centro de Investigación en Farmacobiología Aplicada (CIFA), University of Navarra, 31080 Pamplona, Spain
2
Laboratório de Avaliação e Síntese de Substâncias Bioativas (LASSBio), Faculdade de Farmácia, Universidade Federal do Rio de Janeiro (UFRJ), Rio de Janeiro, PO Box 68024, RJ 21944-970, Brazil
*
Author to whom correspondence should be addressed.
Molecules 2008, 13(1), 78-85; https://doi.org/10.3390/molecules13010078
Submission received: 21 December 2007 / Revised: 16 January 2008 / Accepted: 16 January 2008 / Published: 17 January 2008
(This article belongs to the Section Organic Chemistry)

Abstract

:
The unexpected tendency of amines and functionalized hydrazines to reduce ethyl 3-phenylquinoxaline-2-carboxylate 1,4-di-N-oxide (1) to afford a quinoxaline 1c and mono-oxide quinoxalines 1a and 1b is described. The experimental conditions were standardized to the use of two equivalents of amine in ethanol under reflux for two hours, with the aim of studying the distinct reductive profiles of the amines and the chemoselectivity of the process. With the exception of hydrazine hydrate, which reduced compound 1 to a 3-phenyl-2-quinoxalinecarbohydrazide derivative, the amines only acted as reducing agents.

Introduction

Quinoxaline and quinoxaline 1,4-di-N-oxide are heterocycles that are often used in the synthesis of biologically active compounds [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31]. The former is described as a bioisoster of the quinoline, naphthyl, benzothienyl and other aromatic rings [32], and it can be found in the structure of anti-bacterial [1], anti-tuberculosis [5,6,7,9,10,11,13,14,16,22], anti-cancer [8,18,20,28], anti-malarial [12,24,26,29,30,31], anti-Chagas [15,25] and anti-inflammatory [19,27] drug candidates. The widespread activity of quinoxaline 1,4-di-N-oxides can be associated with the generation of free radicals [33]. In our continuing efforts to find quinoxaline-1,4-di-N-oxide derivatives with anti-malarial activity [12,23,24,26,29,30,31], the conversion of compound 1 into hydrazides 2 and amide derivatives 3 was carried out by means of hydrazinolysis and aminolysis reactions, in which unexpected results were obtained. In this work we describe the reducing profiles of amine derivatives when they act upon the ethyl 3-phenylquinoxaline-2-carboxylate 1,4-di-N-oxide derivative 1.
Scheme 1. Design of new quinoxaline 1,4-di-N-oxide derivatives.
Scheme 1. Design of new quinoxaline 1,4-di-N-oxide derivatives.
Molecules 13 00078 g001

Results and Discussion

In a continuing effort to synthesize new anti-malarial drug candidates, we wished to substitute the carboethoxy moiety present in lead-compound 1, with a functionalized hydrazide subunit 2 (Scheme 1). For this purpose compound 1 was first treated with phenylhydrazine in the presence of ethanol under reflux for two hours. After workup, the crude oil was analyzed by 1H-NMR, which revealed the presence of four different carboethoxy group signals. This crude oil was purified by silica gel column chromatography and the separated products analyzed by 1H-NMR, IR, mass spectra and elemental analyses. From these analyses, it was observed that the reaction with phenylhydrazine had failed to produce the functionalized hydrazide derivative 2, but rather this reaction surprisingly gave a mixture of quinoxaline 1c, quinoxalines N-4 monoxide 1a, N-1 monoxide 1b and 1,4-di-N-oxide 1 (Table 1). While the reducing activity of hydrazine hydrate [34,35] is well known, the aforementioned information was not as clear for phenylhydrazine. In an attempt to determine if the reduction of 1 by phenylhydrazine could be influenced by the electronic profile of quinoxaline-ester 1, the derivatives 4-7, attached to electron-withdrawing and electron-donating groups, were treated with phenylhydrazine in ethanol under reflux for two hours; the results are given in Table 1 (entries 2, 3, 4 and 5). These results showed no selective reduction for substrates 5-7, with the exception of compound 4 (4’-nitro, σp = 0.81) that could be selectively reduced to quinoxaline 4c (entry 2). In this particular case, different reducing profiles were observed for phenylhydrazine and hydrazine hydrate because, with the latter no chemoselectivity was observed, and consequently, the nitro group was also reduced and the hydrazinolysis product was formed (data not shown). Intrigued by these results, the possibility that other amines could act as reducing agents towards the quinoxaline 1,4-di-N-oxide system was then studied. Derivative 1, used as template, was treated with different amines (two equivalents) in the presence of ethanol under reflux for two hours (Table 1). The results clearly demonstrated the ability of hydroxylamine (entry 9), methylamine (entry 10), ethanolamine (entry 11), methyl hydrazine (entry 6), and 2,4-dinitrophenylhydrazine (entry 7) to reduce compound 1 to the mono-oxide derivatives 1a and 1b and to quinoxaline 1c, while amines such as triethylamine (entry 12) and aniline (entry 13), were unable to reduce compound 1, as no significant difference was noted when compared to the experiment carried out in the absence of amine (entry 14).
Table 1. Reduction of ethyl 3-phenylquinoxaline-2-carboxylate 1,4-di-N-oxide (1). Molecules 13 00078 i001
Table 1. Reduction of ethyl 3-phenylquinoxaline-2-carboxylate 1,4-di-N-oxide (1). Molecules 13 00078 i001
EntryReducing AgentWn=n1=1
(1, 4-7)
n=0, n1=1
(1b, 4-7a)
n=1, n1=0
(1c, 4-7b)
n=n1=0
(1a, 4-7c)
1PhNHNH2H18.1%30.1%28.4%23.4%
2PhNHNH2NO216.7%0%0%83.3%
3PhNHNH2CF322.6%30,0%23.7%23.7%
4PhNHNH2OCH325.7%28.6%17.6%28.1%
5PhNHNH2CH315.1%31.8%21.9%31.2%
6H3CNHNH2H43.8%21.8%22.5%11.9%
72,4-diNO2PhNHNH2H57.4%8.9%24.1%9.6%
8N2H4.H2OH0%0%0%100%*
9NH2OH.H2OH22.6%8.1%33.8%35.5%
10NH2CH3.H2OH33.8%8.0%49.7%8.5%
11NH2(CH2)2OHH0%34.2%36.0%29.8%
12Et3NH80.8%0%19.2%0%
13PhNH2H81.2%0%18.8%0%
14EtOHH85.5%0%14.5%0%
15P(OCH3)3H0%100%0%0%
16Na2S2O4H0%0%0%100%
*formation of 3-phenyl-2-quinoxalinecarbohydrazide; absence of chemioselectivity.
The yields were determined by 1H-NMR analysis of total crude product mixtures obtained after the work-up of each reaction. The methyl (OCH2CH3) resonances for the four compounds 1, 1a, 1b and 1c, were observed at different chemical shifts, and thus, integration of the methyl region (OCH2CH3) allowed relative molar percentages to be readily ascertained. The unequivocal structural assignments of the N-4 oxide 1a, N-1 oxide 1b and quinoxaline 1c derivatives were accomplished by 1H-NMR, IR, mass spectra and elemental analyses. In an attempt to specifically identify mono-oxide quinoxaline derivatives (1a versus 1b), a selective mono-deoxygenation of compound 1 was performed (entry 14) using trimethylphosphite (entry 15), as previously described by Kluge and coworkers [36]. By this methodology, it was only possible to obtain the N-4 oxide derivative 1a, which was characterized by 1H-NMR; the data was used for direct comparison with the N-oxides obtained from the reaction with the amines. In a similar way, the quinoxaline di-N-oxide 1 was totally reduced to quinoxaline 1c, using Na2S2O4 (entry 16) in a mixture of methanol and water [37,24]; the chemical shift of compound 1c was measured by 1H-NMR and compared with the products obtained from amine reductions. Compounds 1a, 1b and 1c were evaluated for their ability to inhibit P. falciparum (chloroquine sensitive), and were shown to be inactive as anti-malarial agents (data not shown).

Conclusions

In summary, this work clearly demonstrates a tendency of amines and functionalized hydrazines to reduce ethyl 3-phenylquinoxaline-2-carboxylate 1,4-di-N-oxide (1) to a quinoxaline 1c and mono-oxides quinoxalines 1a and 1b. Although the starting material 1 was recovered in most of the reactions (entries 1-7, 9-10, 12-14), suggesting that a longer time reaction could be necessary, the experimental conditions were standardized in two hours, using two equivalents of amine, with the aim of studying the distinct reductive profile of the amine used and the chemoselectivity of the process. With the exception of hydrazine hydrate, a well-known reducing agent, which reduced compound 1 to a 3-phenyl-2-quinoxalinecarbohydrazide derivative, the amines were unable to act as nucleophiles and acted exclusively as reducing agents. Compounds 1a, 1b and 1c were inactive as antimalarial agents.

Experimental

General

All of the synthesized compounds were chemically characterized by thin layer chromatography (TLC), infrared (IR), nuclear magnetic resonance (1H-NMR), mass spectra (MS) and elemental microanalysis (CHN). Alugram SIL G/UV254 (0.2 mm layer, Macherey-Nagel GmbH & Co. KG. Düren, Germany) was used for TLC and Silica gel 60 (0.040-0.063 mm, Merck) for Flash Column Chromatography. 1H-NMR spectra were recorded on a Bruker 400 Ultrashield instrument (400 MHz), using TMS as the internal standard with CDCl3 as the solvent; the chemical shifts are reported in ppm (δ) and coupling constants (J) values are given in Hertz (Hz). Signal multiplicities are represented by: s (singlet), d (doublet), t (triplet), q (quadruplet), dd (double doublet) and m (multiplet). IR spectra were recorded on a Nicolet Nexus FTIR (Thermo, Madison, USA) in KBr pellets. Mass spectra were measured on a MSD/DS 5973N G2577A mass spectrometer (Agilent Technologies, Waldbronn, Germany) with a direct insertion probe (DIP). The ionization method was electron impact (EI, 70 eV). Elemental microanalyses were obtained on a Leco CHN-900 Elemental Analyzer (Leco, Tres Cantos, Spain) from vacuum-dried samples. The analytical results for C, H, and N, were within ± 0.4 of the theoretical values. Chemicals were purchased from Panreac Química S.A. (Barcelona, Spain), Sigma-Aldrich Química, S.A. (Alcobendas, Spain), Acros Organics (Janssen Pharmaceuticalaan, Geel, Belgium) and Lancaster (Bischheim-Strasbourg, France).

General procedure for the reduction of ethyl 3-phenylquinoxaline-2-carboxylate 1,4-di-N-oxide (1)

Ethyl 3-phenylquinoxaline-2-carboxylate 1,4-di-N-oxide (1, 1 mmol) was added to ethanol (10 mL) and the amine derivative (1 mmol). The mixture was refluxed for two hours, then the reactions were worked-up by adding CH2Cl2 (50 mL), followed by extraction with 10% aqueous HCl (4 x 15 mL). The organic layer was dried (Na2SO4), filtered, and evaporated to dryness. The yields were determined by 1H-NMR analysis of the total crude product mixture. The residue was later purified by silica gel column chromatography (n-hexane-ethyl acetate).
Ethyl 3-phenylquinoxaline-2-carboxylate 1,4-di-N-oxide (1). 1H-NMR: δ 1.08 (t, J= 7.2 Hz, 3H, OCH2CH3); 4.25 (t, J= 7.2 Hz, 2H, OCH2CH3); 7.53 (m, 3H, H3’-H5’); 7.61 (m, 2H, H2’ and H6’); 7.91 (m, 2H, H6 and H7); 8.65 (m, 2H, H5 and H8) ppm; 13C-NMR: δ 13.98 (OCH2CH3), 63.65 (OCH2CH3), 120.89 (C5), 121.08 (C8), 127.84 (C1’), 129.16 (C3’ and C5’), 130.15 (C2’ and C6’), 131.26 (C4’), 132.51 (C6), 132.53 (C7), 136.53 (C2), 137.73 (C10), 138.81 (C3), 140.08 (C9), 159.66 (CO2Et) ppm; IR: 2978 (ArC-H), 1746 (C=O), 1352 (N-oxide), 701 and 666 (monosubstituted phenyl) cm-1; MS: 310 (m/z, 100%), 294 (M+, 6%), 249 (M+, 51%), 221 (M+, 46%), 77 (M+, 46%); Anal. calcd. for C17H14N2O4: C, 65.80; H, 4.52; N, 9.03. Found: C, 65.65; H, 4.57; N, 8.98.
Ethyl 3-phenylquinoxaline-2-carboxylate 4-N-oxide (1a). 1H-NMR: δ 1.04 (t, J=7.2 Hz; 3H, OCH2CH3); 4.20 (q, J=7.2 Hz; 2H, OCH2CH3); 7.56 (m, 3H, H3’-H5’); 7.61 (m, 2H, H2’and H6’); 7.88 (m, 2H, H6 and H7); 8.26 (dd, J= 1.2, 8.2 Hz; 1H, H5); 8.65 (dd, J= 1.2, 7.6 Hz; 1H, H8); IR: 2981 (ArC-H), 1742 (C=O), 1359 (N-oxide), 701 and 666 (monosubstituted phenyl) cm-1; MS: 294 (m/z, 65%), 265 (M+, 13%), 249 (M+, 26%), 221 (M+, 100%), 77 (M+, 18%); Anal. calcd. for C17H14N2O3: C, 69.38; H, 4.79; N, 9.52. Found: C, 69.37; H, 4.77; N, 9.51.
Ethyl 3-phenylquinoxaline-2-carboxylate 1-N-oxide (1b). 1H-NMR: δ 1.25 (t, J=7.2 Hz; 3H, OCH2CH3); 4.42 (q, J=7.2 Hz; 2H, OCH2CH3); 7.54 (m, 3H, H3’-H5’); 7.80 (m, 3H, H2’and H6’ and H7); 7.89 (dt, J= 8.4, 7.6 Hz; 1H, H6); 8.20 (d, J= 8.4 Hz; 1H, H5); 8.61 (d, J= 8.0 Hz; 1H, H8); IR (KBr): 2979 (ArC-H), 1735 (C=O), 1363 (N-oxide), 701 and 671 (monosubstituted phenyl) cm-1; MS: 294 (m/z, 60%), 265 (M+, 9%), 249 (M+, 31%), 221 (M+, 100%), 77 (M+, 26%); Anal. calcd. for C17H14N2O3: C, 69.38; H, 4.79; N, 9.52. Found: C, 69.39; H, 4.80; N, 9.51.
Ethyl 3-phenylquinoxaline-2-carboxylate (1c). 1H-NMR: δ 1.19 (t, J=7.2 Hz; 3H, OCH2CH3); 4.34 (q, J=7.2 Hz; 2H, OCH2CH3); 7.53 (m, 3H, H3’-H5’); 7.76 (m, 2H, H2’and H6’); 7.85 (m, 2H, H6 and H7); 8.20 (d, J= 8.0 Hz; 1H, H5); 8.24 (d, J= 7.6 Hz; 1H, H8); IR: 2990 (ArC-H), 1730 (C=O), 711 and 669 (monosubstituted phenyl) cm-1; MS: 278 (m/z, 33%), 249 (M+, 26%), 234 (M+, 15%), 206 (M+, 100%), 77 (M+, 33%); Anal. calcd. for C17H14N2O2: C, 73.37; H, 5.07; N, 10.07. Found: C, 73.35; H, 5.08; N, 10.05.

Acknowledgements

This work has been carried out with the financial support of the FIS project (1051005, October 2005), the Instituto de Salud Carlos III: Red de centros de cancer RTICCC (C03/10) and the PiUNA project (University of Navarra). We also thank the “Coordenação de Aperfeiçoamento de Pessoal de Nível Superior” (CAPES; BR) for the fellowship (to LML; BEX0520/04-7) received.

References

  1. Dirlam, J. P.; Presslitz, J. E. Synthesis and antibacterial activity of isomeric 6 and 7-acetyl-3-methyl-2-quinoxalinecarboxamide 1,4-dioxides. J. Med. Chem. 1978, 21, 483–485. [Google Scholar] [CrossRef]
  2. Dirlam, J. P.; Czuba, L. J.; Dominy, B. W.; James, R. B.; Pezzullo, R. M.; Presslitz, J. E.; Windisch, W. W. Synthesis and antibacterial activity of 1-hydroxy-1-methyl-1,3-dihydrofuro[3,4-b]quinoxaline 4,9-dioxide and related compounds. J. Med. Chem. 1979, 22, 1118–1121. [Google Scholar] [CrossRef]
  3. Monge, A.; Martinez-Crespo, F. J.; De Cerain, A. L.; Palop, J. A.; Narro, S.; Senador, V.; Marin, A.; Sainz, Y.; Gonzalez, M.; Hamilton, E.; Barker, A. J. Hypoxia-selective agents derived from 2-quinoxalinecarbonitrile 1,4-di-N-oxides. 2. J. Med. Chem. 1995, 38, 4488–4494. [Google Scholar] [CrossRef]
  4. Monge, A.; Palop, J. A.; De Cerain, A. L.; Senador, V.; Martinez-Crespo, F. J.; Sainz, Y.; Narro, S.; Garcia, E.; De Miguel, C.; Gonzalez, M.; Hamilton, E.; Barker, A. J.; Clarke, E. D.; Greenhow, D. T. Hypoxia-selective agents derived from quinoxaline 1,4-di-N-oxides. J. Med. Chem. 1995, 38, 1786–1792. [Google Scholar] [CrossRef]
  5. Montoya, M. E.; Sainz, Y.; Ortega, M. A.; De Cerain, A. L.; Monge, A. Synthesis and antituberculosis activity of some new 2-quinoxalinecarbonitriles. Farmaco 1998, 53, 570–573. [Google Scholar] [CrossRef]
  6. Ortega, M. A.; Sainz, Y.; Montoya, M. E.; De Cerain, A. L.; Monge, A. Synthesis and antituberculosis activity of new 2-quinoxalinecarbonitrile 1,4-di-N-oxides. Pharmazie 1999, 54, 24–25. [Google Scholar]
  7. Sainz, Y.; Montoya, M. E.; Martinez-Crespo, F. J.; Ortega, M. A.; de Cerain, A. L.; Monge, A. New quinoxaline 1,4-di-N-oxides for treatment of tuberculosis. Arzneim.-Forsch. 1999, 49, 55–59. [Google Scholar]
  8. Ortega, M. A.; Morancho, M. J.; Martinez-Crespo, F. J.; Sainz, Y.; Montoya, M. E.; de Cerain, A. L.; Monge, A. New quinoxalinecarbonitrile 1,4-di-N-oxide derivatives as hypoxic-cytotoxic agents. Eur. J. Med. Chem. 2000, 35, 21–30. [Google Scholar]
  9. Ortega, M. A.; Montoya, M. E.; Jaso, A.; Zarranz, B.; Tirapu, I.; Aldana, I.; Monge, A. Antimycobacterial activity of new quinoxaline-2-carbonitrile and quinoxaline-2-carbonitrile 1,4-di-N-oxide derivatives. Pharmazie 2001, 56, 205–207. [Google Scholar]
  10. Carta, A.; Paglietti, G.; Rahbar Nikookar, M. E.; Sanna, P.; Sechi, L.; Zanetti, S. Novel substituted quinoxaline 1,4-dioxides with in vitro antimycobacterial and anticandida activity. Eur. J. Med. Chem. 2002, 37, 355–366. [Google Scholar]
  11. Ortega, M. A.; Sainz, Y.; Montoya, M. E.; Jaso, A.; Zarranz, B.; Aldana, I.; Monge, A. Anti-Mycobacterium tuberculosis agents derived from quinoxaline-2-carbonitrile and quinoxaline-2-carbonitrile 1,4-di-N-oxide. Arzneim.-Forsch. 2002, 52, 113–119. [Google Scholar]
  12. Aldana, I.; Ortega, M. A.; Jaso, A.; Zarranz, B.; Oporto, P.; Gimenez, A.; Monge, A.; Deharo, E. Anti-malarial activity of some 7-chloro-2-quinoxalinecarbonitrile-1,4-di-N-oxide derivatives. Pharmazie 2003, 58, 68–69. [Google Scholar]
  13. Jaso, A.; Zarranz, B.; Aldana, I.; Monge, A. Synthesis of new 2-acetyl and 2-benzoyl quinoxaline 1,4-di-N-oxide derivatives as anti-Mycobacterium tuberculosis agents. Eur. J. Med. Chem. 2003, 38, 791–800. [Google Scholar] [CrossRef]
  14. Zarranz, B.; Jaso, A.; Aldana, I.; Monge, A. Synthesis and antimycobacterial activity of new quinoxaline-2-carboxamide 1,4-di-N-oxide derivatives. Bioorg. Med. Chem. 2003, 11, 2149–2156. [Google Scholar] [CrossRef]
  15. Aguirre, G.; Cerecetto, H.; Di Maio, R.; Gonzalez, M.; Alfaro, M. E. M.; Jaso, A.; Zarranz, B.; Ortega, M. A.; Aldana, I.; Monge-Vega, A. Quinoxaline N,N'-dioxide derivatives and related compounds as growth inhibitors of Trypanosoma cruzi. Structure-activity relationships. Bioorg. Med. Chem. Lett. 2004, 14, 3835–3839. [Google Scholar] [CrossRef]
  16. Carta, A.; Loriga, M.; Paglietti, G.; Mattana, A.; Fiori, P. L.; Mollicotti, P.; Sechi, L.; Zanetti, S. Synthesis, anti-mycobacterial, anti-trichomonas and anti-candida in vitro activities of 2-substituted-6,7-difluoro-3-methylquinoxaline 1,4-dioxides. Eur. J. Med. Chem. 2004, 39, 195–203. [Google Scholar] [CrossRef]
  17. Kim, Y. B.; Kim, Y. H.; Park, J. Y.; Kim, S. K. Synthesis and biological activity of new quinoxaline antibiotics of echinomycin analogues. Bioorg. Med. Chem. Lett. 2004, 14, 541–544. [Google Scholar] [CrossRef]
  18. Perez-Melero, C.; Maya, A. B.; del Rey, B.; Pelaez, R.; Caballero, E.; Medarde, M. A new family of quinoline and quinoxaline analogues of combretastatins. Bioorg. Med. Chem. Lett. 2004, 14, 3771–3774. [Google Scholar]
  19. Singh, S. K.; Saibaba, V.; Ravikumar, V.; Rudrawar, S. V.; Daga, P.; Rao, C. S.; Akhila, V.; Hegde, P.; Rao, Y. K. Synthesis and biological evaluation of 2,3-diarylpyrazines and quinoxalines as selective COX-2 inhibitors. Bioorg. Med. Chem. 2004, 12, 1881–1893. [Google Scholar] [CrossRef]
  20. Zarranz, B.; Jaso, A.; Aldana, I.; Monge, A. Synthesis and anticancer activity evaluation of new 2-alkylcarbonyl and 2-benzoyl-3-trifluoromethyl-quinoxaline 1,4-di-N-oxide derivatives. Bioorg. Med. Chem. 2004, 12, 3711–3721. [Google Scholar] [CrossRef]
  21. Carta, A.; Corona, P.; Loriga, M. Quinoxaline 1,4-dioxide: A versatile scaffold endowed with manifold activities. Curr. Med. Chem. 2005, 12, 2259–2272. [Google Scholar] [CrossRef]
  22. Jaso, A.; Zarranz, B.; Aldana, I.; Monge, A. Synthesis of new quinoxaline-2-carboxylate 1,4-dioxide derivatives as anti-Mycobacterium tuberculosis agents. J. Med. Chem. 2005, 48, 2019–2025. [Google Scholar] [CrossRef]
  23. Lima, L. M.; Zarranz, B.; Marin, A.; Solano, B.; Vicente, E.; Silanes, S. P.; Aldana, I.; Monge, A. Comparative use of solvent-free KF-Al2O3 and K2CO3 in acetone in the synthesis of quinoxaline 1,4-dioxide derivatives designed as antimalarial drug candidates. J. Heterocycl. Chem. 2005, 42, 1381–1385. [Google Scholar] [CrossRef]
  24. Zarranz, B.; Jaso, A.; Aldana, I.; Monge, A.; Maurel, S.; Deharo, E.; Jullian, V.; Sauvain, M. Synthesis and antimalarial activity of new 3-arylquinoxaline-2-carbonitrile derivatives. Arzneim.-Forsch. 2005, 55, 754–761. [Google Scholar]
  25. Urquiola, C.; Vieites, M.; Aguirre, G.; Marin, A.; Solano, B.; Arrambide, G.; Noblia, P.; Lavaggi, M. L.; Torre, M. H.; Gonzalez, M.; Monge, A.; Gambino, D.; Cerecetto, H. Improving anti-trypanosomal activity of 3-aminoquinoxaline-2-carbonitrile N1,N4-dioxide derivatives by complexation with vanadium. Bioorg. Med. Chem. 2006, 14, 5503–5509. [Google Scholar] [CrossRef]
  26. Zarranz, B.; Jaso, A.; Lima, L. M.; Aldana, I.; Monge, A.; Maurel, S.; Sauvain, M. Antiplasmodial activity of 3-trifluoromethyl-2-carbonylquinoxaline di-N-oxide derivatives. Braz. J. Pharm. Sci. 2006, 42, 357–361. [Google Scholar]
  27. Burguete, A.; Pontiki, E.; Hadjipavlou-Litina, D.; Villar, R.; Vicente, E.; Solano, B.; Ancizu, S.; Perez-Silanes, S.; Aldana, I.; Monge, A. Synthesis and anti-inflammatory/antioxidant activities of some new ring substituted 3-phenyl-1-(1,4-di-N-oxide quinoxalin-2-yl)-2-propen-1-one derivatives and of their 4,5-dihydro-(1H)-pyrazole analogues. Bioorg. Med. Chem. Lett. 2007, 17, 6439–6443. [Google Scholar] [CrossRef]
  28. Solano, B.; Junnotula, V.; Marín, A.; Villar, R.; Burguete, A.; Vicente, E.; Pérez-Silanes, S.; Aldana, I.; Monge, A.; Dutta, S.; Sarkar, U.; Gates, K. S. Synthesis and biological evaluation of new 2-arylcarbonyl-3-trifluoromethylquinoxaline 1,4-di-N-oxide derivatives and their reduced analogs. J. Med. Chem. 2007, 50, 5485–5492. [Google Scholar] [CrossRef]
  29. Marín, A.; Lima, L. M.; Solano, B.; Vicente, E.; Pérez-Silanes, S.; Maurel, S.; Sauvain, M.; Aldana, I.; Monge, A.; Deharo, E. Antiplasmodial structure-activity relationship of 3-trifluoro-methyl-2-arylcarbonylquinoxaline 1,4-di-N-oxide derivatives. Exp. Parasitol. 2008, 118, 25–31. [Google Scholar] [CrossRef]
  30. Vicente, E.; Charnaud, S.; Bongard, E.; Villar, R.; Burguete, A.; Solano, B.; Ancizu, S.; Pérez-Silanes, S.; Aldana, I.; Vivas, L.; Monge, A. Synthesis and antiplasmodial activity of 3-furyl and 3-thienylquinoxaline-2-carbonitrile 1,4-di-N-oxide derivatives. Molecules 2008, 13, 69–77. [Google Scholar] [CrossRef] [Green Version]
  31. Vicente, E.; Lima, L. M.; Bongard, E.; Charnaud, S.; Villar, R.; Solano, B.; Burguete, A.; Pérez-Silanes, S.; Aldana, I.; Vivas, L.; Monge, A. Synthesis and structure-activity relationship of 3-phenylquinoxaline 1,4-di-N-oxide derivatives as antimalarial agents. Eur. J. Med. Chem. 2008. [Google Scholar] [CrossRef]
  32. Lima, L. M.; Barreiro, E. J. Bioisosterism: A useful strategy for molecular modification and drug design. Curr. Med. Chem. 2005, 12, 23–49. [Google Scholar] [CrossRef]
  33. Inbaraj, J. J.; Motten, A. G.; Chignell, C. F. Photochemical and photobiological studies of tirapazamine (SR 4233) and related quinoxaline 1,4-di-N-oxide analogues. Chem. Res. Toxicol. 2003, 16, 164–170. [Google Scholar] [CrossRef]
  34. Kuhn, L. P. Catalytic reduction with hydrazine. J. Am. Chem. Soc. 1951, 73, 1510–1512. [Google Scholar] [CrossRef]
  35. Abul-Hajj, Y. J. Stereospecific reduction of steroidal 4-ene-3-ols with hydrazine. J. Org. Chem. 1971, 36, 2730. [Google Scholar] [CrossRef]
  36. Kluge, A. F.; Maddox, M. L.; Lewis, G. S. Formation of quinoxaline monoxides from reaction of benzofurazan oxide with enones and C-13 NMR correlations of quinoxaline N-oxides. J. Org. Chem. 1980, 45, 1909–1914. [Google Scholar] [CrossRef]
  37. Haddadin, M. J.; Zahr, G. E.; Rawdah, T. N.; Chelhot, N. C.; Issidorides, C. H. Deoxygenation of quinoxaline N-oxides and related compounds. Tetrahedron 1974, 30, 659–666. [Google Scholar] [CrossRef]
  • Sample avalibility: Contact the authors

Share and Cite

MDPI and ACS Style

Lima, L.M.; Vicente, E.; Solano, B.; Pérez-Silanes, S.; Aldana, I.; Monge, A. Unexpected Reduction of Ethyl 3-Phenylquinoxaline-2- carboxylate 1,4-Di-N-oxide Derivatives by Amines. Molecules 2008, 13, 78-85. https://doi.org/10.3390/molecules13010078

AMA Style

Lima LM, Vicente E, Solano B, Pérez-Silanes S, Aldana I, Monge A. Unexpected Reduction of Ethyl 3-Phenylquinoxaline-2- carboxylate 1,4-Di-N-oxide Derivatives by Amines. Molecules. 2008; 13(1):78-85. https://doi.org/10.3390/molecules13010078

Chicago/Turabian Style

Lima, Lidia M., Esther Vicente, Beatriz Solano, Silvia Pérez-Silanes, Ignacio Aldana, and Antonio Monge. 2008. "Unexpected Reduction of Ethyl 3-Phenylquinoxaline-2- carboxylate 1,4-Di-N-oxide Derivatives by Amines" Molecules 13, no. 1: 78-85. https://doi.org/10.3390/molecules13010078

Article Metrics

Back to TopTop