Next Article in Journal
Condensation of Diacetyl with Alkyl Amines: Synthesis and Reactivity of p-Iminobenzoquinones and p-Diiminobenzoquinones
Next Article in Special Issue
An Integrated Experimental/Theoretical Study of Structurally Related Poly-Thiophenes Used in Photovoltaic Systems
Previous Article in Journal
Cinidium officinale and its Bioactive Compound, Butylidenephthalide, Inhibit Laser-Induced Choroidal Neovascularization in a Rat Model
Previous Article in Special Issue
Resin-Immobilized Palladium Nanoparticle Catalysts for Organic Reactions in Aqueous Media: Morphological Aspects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

An Efficient and Recyclable Nanoparticle-Supported Cobalt Catalyst for Quinoxaline Synthesis

1
Department of Science, Payame Noor University, P. O. Box: 19395-4697, Tehran 19569, Iran
2
Laboratório de Síntese Orgânica Limpa—LASOL, Universidade Federal de Pelotas UFPEL, Pelotas, CEP 96010-900, Brazil
3
Departamento de Quimica Organica, Universidad de Cordoba, Campus de Rabanales, Edificio Marie Curie (C3), Ctra Nnal IV-A, Km 396, Cordoba E14014, Spain
*
Author to whom correspondence should be addressed.
Molecules 2015, 20(11), 20709-20718; https://doi.org/10.3390/molecules201119731
Submission received: 23 September 2015 / Revised: 3 November 2015 / Accepted: 12 November 2015 / Published: 19 November 2015
(This article belongs to the Special Issue Metal Nanocatalysts in Green Synthesis and Energy Applications)

Abstract

:
The syntheses of quinoxalines derived from 1,2-diamine and 1,2-dicarbonyl compounds under mild reaction conditions was carried out using a nanoparticle-supported cobalt catalyst. The supported nanocatalyst exhibited excellent activity and stability and it could be reused for at least ten times without any loss of activity. No cobalt contamination could be detected in the products by AAS measurements, pointing to the excellent activity and stability of the Co nanomaterial.

1. Introduction

Quinoxaline derivatives are attractive N-containing heterocycles and these scaffolds have attracted much attention, not only in synthetic chemistry [1,2,3] but also in the medicinal field [4,5,6,7,8,9,10,11]. These compounds exhibit diverse biological activities, such as antiviral [4,5], antibacterial [6], anti-inflammatory [7], antitumoral [8,9] and anti-HIV properties [10,11]. Examples of quinoxaline- containing pharmacological entities are shown in Figure 1. In addition, quinoxalines have been applied as building blocks for the development of macrocyclic molecular receptors [12,13], semiconducting materials [14,15,16,17,18,19,20], dyes [21], cavitands [22] and luminescent materials [23].
Figure 1. Biologically important quinoxalines.
Figure 1. Biologically important quinoxalines.
Molecules 20 19731 g001
Generally, quinoxalines can be prepared via a double condensation of 1,2-phenylenediamines with 1,2-diketones [24,25,26,27,28]. A number of reagents have been shown to catalyze these reactions such as acidic alumina [29], citric acid [30], magnetic Fe3O4 nanoparticles in H2O [31], silica-bonded sulfonic acid [32], among others [33,34]. Other protocols to synthesize quinoxalines mainly involve the oxidative trapping of vicinal diols or α-hydroxy ketones with 1,2-diamines [35,36,37,38,39,40,41,42], 1,4-addition of 1,2-diamines to diazenylbutenes [43], coupling of epoxides with ene-1,2-diamines [44,45], 2-nitroanilines with phenethylamines [46], alkynes or ketones with 1,2-diamines via a key oxidation process [47,48,49,50,51]. Therefore, the development of efficient methods for accessing quinoxalines derivatives continues to be an active area of research.
Nanoparticle-supported catalysts can offer important advantages as compared to homogeneous transition metal systems and colloidal nanoparticles. These include a good reusability coupled with high activities and specificities in different chemistries based on their excelling properties (high surface areas, degenerated density of energy states and plasmon) [52,53,54]. In this regard, Co/supported catalysts were previously reported to be highly active and versatile for acid and redox catalyzed processes [54,55].
To the best of our knowledge, there is no protocol describing the preparation of quinoxaline derivatives using a nanoparticle-supported cobalt catalyst. In view of the explained above, we decided to examine the synthesis of substituted quinoxalines by reaction of 1,2-diketones with 1,2-phenylenediamines using a nanoparticle-supported cobalt catalyst (Scheme 1).
Scheme 1. General scheme of the reactions.
Scheme 1. General scheme of the reactions.
Molecules 20 19731 g002

2. Results and Discussion

Initially, we chose 1,2-diphenylethanedione (1a)and 1,2-diamino-4-nitrobenzene (2a) as model substrates to establish the best conditions for this reaction and some experiments were performed to synthesize the corresponding quinoxaline 3a (Table 1). We started our studies reacting 1,2-diketone 1a (1.0 mmol) with 1,2-phenylenediamine 2a (1.0 mmol) at 100 °C for 2 h, without catalyst and solvent. Under these conditions, product 3a was not obtained (Table 1, entry 1). Good results were obtained however when the reactions of substrates 1a and 2a were carried out using H2O as solvent in the presence of Co NPs (2 mol %) as catalyst. Reactions performed at 100 °C and 50 °C gave the desired product in 87% and 57% yield, respectively (Table 1, entries 2 and 3). A similar result was obtained when the reaction was conducted at 100 °C, however using 1 mol% of Co NPs (86% yield) (Table 1, entry 4). Good results were also found when the reactions were performed using EtOH as solvent (Table 1, entry 5–9). Excellent yields of product 3a were achieved in reactions carried out in EtOH at 78 °C using 1 mol% of catalyst (Table 1, entries 7–8). When the amount of catalyst was reduced to 0.5 mol %, a decrease in the yield of product 3a was observed (Table 1, entry 9). Finally, the reaction performed using 1 mol % of Co NPs at 100 °C and in absence of EtOH yielded the quinoxaline 3a in 72% yield (Table 1, entry 10).
Analyzing the results shown in Table 1, we established the best reaction conditions reacting 1,2-diphenylethanedione (1a, 1.0 mmol, 0.033 g) with 1,2-diamino-4-nitrobenzene (2a, 1.0 mmol) using supported CoNPs (1 mol %) as catalyst and EtOH (5 mL) as solvent. After that, the mixture was stirred at reflux for 90 min in open atmosphere, affording 6-nitro-2,3-diphenylquinoxaline(3a) in 92% yield after crystallization.
Table 1. Optimization of reaction condition a. Molecules 20 19731 i001
Table 1. Optimization of reaction condition a. Molecules 20 19731 i001
EntryCatalyst (mol %)SolventTemperature (°C)Time (h)Yield 3a (%) b
1--1002-
22H2O100287
32H2O50257
41H2O100286
52EtOH78293
62EtOH50260
71EtOH78293
81EtOH781.592
90.5EtOH781.580
101-100272
a Reactions are performed using, 1,2-diphenylethanedione 1a (1.0 mmol) and 1,2-diamino-4-nitrobenzene 2a (1.0 mmol) in open atmosphere. b Yields are given for isolated product 3a after crystallization.
In order to extend the scope of the reaction, the best conditions were employed in reactions of 1,2-diamino-4-nitrobenzene (2a) with other 1,2-diketones 1be with different patterns of substitution and the results are summarized in Table 2. As it can be seen on Table 2 (Entries 1–5), our methodology is suitable to a range of substituted 1,2-diketones containing electron-withdrawing groups, affording excellent yields to desired products in all examples. In addition, the possibility of performing the reaction of 1,2‑diketones 1ae with o-phenylenediamine (2b) was also investigated (Table 2, entries 6–10). Using these substrates, a range of substituted quinoxalines was obtained in excellent yields using the nanoparticle-supported cobalt catalyst under optimized reaction conditions.
Table 2. Generality of the reaction of 1,2-diketones with 1,2-diamines a. Molecules 20 19731 i002
Table 2. Generality of the reaction of 1,2-diketones with 1,2-diamines a. Molecules 20 19731 i002
Entry1,2-Diketone 11,2-Diamines 2Product 3Yield (%) bM.P. (°C)
1 Molecules 20 19731 i003
1a
Molecules 20 19731 i004
2a
Molecules 20 19731 i005
3a
92188–190
2 Molecules 20 19731 i006
1b
2a Molecules 20 19731 i007
3b
92195–197
3 Molecules 20 19731 i008
1c
2a Molecules 20 19731 i009
3c
92173–175
4 Molecules 20 19731 i010
1d
2a Molecules 20 19731 i011
3d
90175–177
5 Molecules 20 19731 i012
1e
2a Molecules 20 19731 i013
3e
88143–145
6 Molecules 20 19731 i014
1a
Molecules 20 19731 i015
2b
Molecules 20 19731 i016
3f
92127–129
7 Molecules 20 19731 i017
1b
2b Molecules 20 19731 i018
3g
96127–129
8 Molecules 20 19731 i019
1c
2b Molecules 20 19731 i020
3h
94133–135
9 Molecules 20 19731 i021
1d
2b Molecules 20 19731 i022
3i
94190–192
10 Molecules 20 19731 i023
1e
2b Molecules 20 19731 i024
3j
94134–135
11 Molecules 20 19731 i025 Molecules 20 19731 i026
2c
Molecules 20 19731 i027
3k
98116–117
12 Molecules 20 19731 i0282c Molecules 20 19731 i029
3l
95163–165
a Reactions were performed using 1,2-diketones 1ae (1.0 mmol), 1,2-diamines 2ab (1.0 mmol), supported CoNPs (1 mol %, 0.033 g) and EtOH (5 mL) at reflux in open flask for 90 min. b Yields are given for isolated products after crystallization.
Reused runs were carried out under similarly optimized conditions using 5 mmol 1,2-diphenylethanedione (1a), 5 mmol of 1,2-diamino-4-nitrobenzene (2a) and supported cobalt catalyst (0.05 mmol, 0.165 g) at 78 °C in 10 mL of ethanol. The catalyst showed excellent recoverability and reusability over ten successive runs under the same conditions as the first run. It is quite remarkable that all materials discussed in this study exhibited outstanding structural stability by TGA (results not shown). The cobalt catalyst was found to be highly stable and reusable under the investigated conditions (up to 12 runs) without any significant loss of its catalytic activity (Table 3). Indeed, ICP analysis of both reaction filtrate and catalyst showed no detectable Co leaching (<0.5 ppm) in the reaction filtrate upon reaction completion, with an almost identical Co content for both fresh and reused catalyst (0.30 vs. 0.29 mmol of Co per gram of catalyst for fresh and 10-time reused material, respectively).
Table 3. Reuses of the supported CoNP catalyst in the reaction of 1,2-diphenylethanedione (1a) with 1,2-diamino-4-nitrobenzene (2a).
Table 3. Reuses of the supported CoNP catalyst in the reaction of 1,2-diphenylethanedione (1a) with 1,2-diamino-4-nitrobenzene (2a).
Run No. a12345678910
Yield (%) b94949292929091909087
a Reaction conditions: 1,2-diphenylethanedione (5.0 mmol) and 1,2-diamino-4-nitrobenzene (5.0 mmol), supported CoNPs ( 0.05 mmol, 0.165 g) in EtOH (10 mL) at reflux conditions for 90 min. b Isolated yields.
The study of the scale-up reaction (from 1 to 20 mmol of substrate) was also investigated under the optimized reaction conditions. When the amount of 1a and 2a was increased to 20 mmol, the same conversion was obtained after 90 min under optimized conditions.
The catalytic performance of our system was eventually compared to reported literature data. As can be seen in Table 4, our recoverable catalytic system possesses remarkably improved activities as compared to those of related previously reported heterogeneous systems.
Table 4. Comparison of the result in the reaction of 1,2-diphenylethanedione (1a) with 1,2-diaminobenzene (2a) with our method and the previous literature.
Table 4. Comparison of the result in the reaction of 1,2-diphenylethanedione (1a) with 1,2-diaminobenzene (2a) with our method and the previous literature.
EntryConditionTime (min)Yield (%)Reference
1Polyaniline-sulfate salt (5 wt %), DCE, r.t.2095[25]
2CAN (5 mol %), H2O, r.t.1098[26]
3I2 (10 mol %), DMSO, r.t.3595[27]
4MeOH:AcOH (9:1), MW, 160 °C599[28]
5Acidic alumina, 80 °C296[29]
6Citric acid (10 mol %), EtOH, r.t.194[30]
7Fe3O4NPs (10 mol %), H2O, r.t.15095[31]
8Silicabonded S-sulfonicacid (3.4 mol %), EtOH/H2O (70/30), r.t.596[32]
9Ga(OTf)3 (1 mol %), EtOH, r.t.599[33]
10Bi(OTf)3 (10 mol %), H2O, r.t.597[34]
11CoNP (1mol %), EtOH, reflux9092Our work

3. Experimental Section

3.1. General Information

Unless otherwise stated, all reagents and chemicals in this study were used as received and were not further purified (Sigma-Aldrich Chemie GmbH, Taufkirchen, Germany). Melting point recorded on a RY-1 microscopic melting apparatus (Hangzhou Chincan Trading Co., Shanghai, China) and uncorrected. 1H-NMR and 13C-NMR spectra were respectively recorded on 500 MHz and 125 MHz by using a Bruker Avance 500 spectrometer (Bruker BioSpin GmbH, Rheinstetten, Germany). Metal content in the materials was determined using inductively coupled plasma (ICP) in a Philips PU 70000 sequential spectrometer (Philips, Almelo, The Netherlands) equipped with an Echelle monochromator (0.0075 nm resolution). Samples were digested in HNO3 and subsequently analyzed by ICP Nitrogen adsorption measurements (Philips) were carried out at 77 K using an ASAP 2000 volumetric adsorption analyzer from Micromeritics (Micromeritics, Norcross, GA, USA). The samples were outgassed for 24 h at 100 °C under vacuum (p b 10–2 Pa) and subsequently analyzed.

3.2. Preparation of the Supported Cobalt Catalyst

CoNPs was synthesized as previously reported [55]. Briefly, salicylaldehyde (2 mmol, 0.244 g) was added to an excess of absolute MeOH, to which 3‑aminopropyl(trimethoxy)silane (2 mmol, 0.352 g) was subsequently added. The color of the solution instantly changed to yellow indicating imine formation. After 3 h, cobalt (II) acetate, Co(OAc)2·2H2O (1 mmol, 0.248 g) was added to the solution, and the mixture stirred for three additional hours to allow the new ligands to complex the cobalt. A color change from pink to olive green is observed. SBA-15 (3 g) was activated by refluxing in concentrated hydrochloric acid (6 M) and then washed thoroughly with deionized water and dried before undergoing chemical surface modification. This activation treatment readily hydrolyses the siloxane Si-O-Si bonds to Si-OH species which will be key to anchor the cobalt complex. Both complex and activated silica were then mixed and the mixture was stirred overnight. The solvent was removed using a rotary evaporator, and the resulting olive green solid dried at 80 °C overnight. The final product was washed with MeOH and water (to remove all physisorbed metal species) until the washings were colourless. Further drying of the solid product was carried out in an oven at 80 °C for 8 h. The loading of cobalt was calculated about 0.3 mmol·g−1 and surface analysis showed cobalt oxide species well dispersed on the surface of SBA-15 with 450 m2·g−1 surface area and pore size of 3.6 nm with 0.77 cm3·g−1 mesoporous pore volume.

3.3. General Reaction Procedure

To a mixture of 1,2-dicarbonyl compound 1ae (1.0 mmol) and 1,2-diamine 2ab (1.0 mmol) in ethanol (5 mL), supported CoNP (0.033g, 1 mol%) was added and the mixture was refluxed in an open flask for 90 min. Reactions were monitored by thin layer chromatography (TLC) until total disappearance of the starting material. After completion of the reaction, the reaction mixture was cooled to room temperature, and resulting solid was collected by filtration and dissolved in ethyl acetate (10 mL). The supported catalyst was recovered by filtration. After evaporation of solvent, the resulting solid product was purified by crystallization in ethanol.

3.4. Selected Spectroscopic Data

6-Nitro-2,3-diphenylquinoxaline (Table 2, Entry 1, 3a). Yellow solid; m.p. 188–190 °C (lit. [56] 193–194 °C). 1H-NMR (CDCl3): δ 7.38 (m, 6H, Ar-H), 7.56 (m, 4H, Ar-H), 8.28 (m, 1H, Ar-H), 8.45 (m, 1H, Ar-H), 9.02 (m, 1H, Ar-H); 13C-NMR (CDCl3): δ 123.27, 125.51, 128.45, 129.67, 129.85, 129.95, 130.66, 137.95, 139.87, 143.39, 147.80, 155.62, 156.18.
2,3-Diphenylquinoxaline (Table 2, Entry 6, 3f). White solid; m.p. 127–129 °C (lit. [26] 126–127 °C). 1H-NMR (CDCl3): δ 7.35 (m, 6H, Ar-H), 7.56 (m, 4H, Ar-H), 7.76 (m, 2H, Ar-H), 8.20 (m, 2H, Ar-H); 13C-NMR (CDCl3): δ 128.29, 128.89, 129.13, 129.915, 130.10, 138.92, 141.15, 154.38.
2,3-Bis(4-Fuorophenyl)quinoxaline (Table 2, Entry 8, 3h). White solid; m.p. 133–135 °C (lit. [56] 135–137 °C). 1H-NMR (CDCl3): δ 7.06 (m, 4H, Ar-H), 7.52 (m, 4H, Ar-H), 7.80 (q, J = 9.5 Hz, 1H, Ar-H), 8.16 (q, J = 9.1 Hz, 1H, Ar-H); 13C-NMR (CDCl3): δ 115.45, 115.61, 129.14, 130.22, 131.70, 131.82, 134.90, 135.02, 141.21, 152.16, 161.54, 164.80.
2,3-Bis(4-Chlorophenyl)quinoxaline (Table 2, Entry 9, 3i).White solid; m.p. 190–192 °C (lit. [32] 195–196 °C). 1H-NMR (CDCl3): δ 7.32 (m, 4H, Ar-H), 7.49 (m, 4H, Ar-H), 7.72 (m, 2H, Ar-H), 8.11 (m, 2H, Ar-H); 13C-NMR (CDCl3): δ 128.50, 128.62, 129.05, 129.11, 129.17, 130.05, 130.13, 130.26, 131.30, 134.12, 137.36, 138.62, 140.09, 141.11, 153.02, 153.18.
6-Methyl-2,3-diphenylquinoxaline (Table 2, Entry 11, 3k). Brown solid; m.p. 116–118 °C (lit. [33] 117–118 °C). 1H-NMR (CDCl3): δ 2.61 (s, 3H, Ar-CH3 ), 7.35 (s, 6H, Ar-H), 7.55 (d, J = 6.5, 4H, Ar-H), 7.60 (s, 1H, Ar-H) 7.98 (s, 1H, Ar-H), 8.09 (d, J = 8.4, 1H, Ar-H); 13C-NMR (CDCl3): δ 21.95, 128.04, 128.24, 128.67, 128.73, 129.90, 129.92, 132.32, 139.24, 139.73, 140.49, 141.29, 152.55, 153.29.

4. Conclusions

In summary, we have developed an environmentally friendly and highly active cobalt nanoparticle on mesoporous SBA-15 material for the synthesis of quinoxalinesin excellent yields from 1,2-diamine and 1,2-dicarbonyl compounds. Reactions could efficiently afford the target products after short reaction times and were run under air and mild reaction conditions and require low loadings of the supported catalyst. The catalyst was found to be highly reusable for at least ten reaction runs under the investigated conditions.

Acknowledgments

F.R. is grateful to Payame Noor University for support of this work.

Author Contributions

F.R. conducted all experimental work, D.A. and R.L. supervised, discussed and completely wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brown, D.J. Quinoxalines: Supplement II. In The Chemistry of Heterocyclic Compounds; John Wiley & Sons: Princeton, NJ, USA, 2004. [Google Scholar]
  2. Katritzky, A.R.; Pozharskii, A.F. Handbook of Heterocyclic Chemistry; Pergamon: Oxford, UK, 2000. [Google Scholar]
  3. Eicher, T.; Hauptmann, S.; Speicher, A. The Chemistry of Heterocycles; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar]
  4. Moorthy, N.; Hari, N.S.; Manivannan, E.; Karthikeyan, C.; Trivedi, P. 6H-Indolo[2,3-b]quinoxalines: DNA and protein interacting scaffold for pharmacological activities. Mini Rev. Med. Chem. 2013, 13, 1415–1420. [Google Scholar] [CrossRef] [PubMed]
  5. Rong, F.; Chow, S.; Yan, S.; Larson, G.; Hong, Z.; Wu, J. Structure–activity relationship (SAR) studies of quinoxalines as novel HCV NS5B RNA-dependent RNA polymerase inhibitors. Bioorg. Med. Chem. Lett. 2007, 17, 1663–1666. [Google Scholar] [CrossRef] [PubMed]
  6. Parhi, A.K.; Zhang, Y.; Saionz, K.W.; Pradhan, P.; Kaul, M.; Trivedi, K.; Pilch, D.S.; LaVoie, E.J. Antibacterial activity of quinoxalines, quinazolines, and 1,5-naphthyridines. Bioorg. Med. Chem. Lett. 2013, 23, 4968–4974. [Google Scholar] [CrossRef] [PubMed]
  7. Smits, R.A.; Lim, H.D.; Hanzer, A.; Zuiderveld, O.P.; Guaita, E.; Adami, M.; Coruzzi, G.; Leurs, R.; Esch, I.J.P. Fragment based design of new H4 receptor-ligands with anti-inflammatory properties in vivo. J. Med. Chem. 2008, 51, 2457–2467. [Google Scholar] [CrossRef] [PubMed]
  8. Hazeldine, S.T.; Polin, L.; Kushner, J.; Paluch, J.; White, K.; Edelstein, M.; Palomino, E.; Corbett, T.H.; Horwitz, J.P. Design, synthesis, and biological evaluation of analogues of the antitumor agent, 2-{4-[(7-chloro-2-quinoxalinyl)oxy]phenoxy}propionic acid (XK469). J. Med. Chem. 2001, 44, 1758–1776. [Google Scholar] [CrossRef] [PubMed]
  9. Rajule, R.; Bryant, V.C.; Lopez, H.; Luo, X.; Natarajan, A. Perturbing pro-survival proteins using quinoxaline derivatives: A structure–activity relationship study. Bioorg. Med. Chem. 2012, 20, 2227–2234. [Google Scholar] [CrossRef] [PubMed]
  10. Hui, X.; Desrivot, J.; Bories, C.; Loiseau, P.M.; Franck, X.; Hocquemiller, R.; Figadere, B. Synthesis and antiprotozoal activity of some new synthetic substituted quinoxalines. Bioorg. Med. Chem. Lett. 2006, 16, 815–820. [Google Scholar] [CrossRef] [PubMed]
  11. 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] [PubMed]
  12. Sessler, J.L.; Maeda, H.; Mizuno, T.; Lynch, V.M.; Furuta, H. Quinoxaline-Bridged Porphyrinoids. J. Am. Chem. Soc. 2002, 124, 13474–13479. [Google Scholar] [CrossRef] [PubMed]
  13. Azov, V.A.; Beeby, A.; Cacciarini, M.; Cheetham, A.G.; Diederich, F.; Frei, M.; Gimzewski, J.K.; Gramlich, V.; Hecht, B.; Jaun, B.; et al. Resorcin[4]arenecavitand-based molecular switches. Adv. Funct. Mater. 2006, 16, 147–156. [Google Scholar] [CrossRef]
  14. Champion, R.D.; Cheng, K.-F.; Pai, C.-L.; Chen, W.-C.; Jenekhe, S.A. Electronic properties and field-effect transistors of thiophene-based donor–acceptor conjugated copolymers. Macromol. Rapid Commun. 2005, 26, 1835–1840. [Google Scholar] [CrossRef]
  15. Wang, E.; Hou, L.; Wang, Z.; Hellström, S.; Zhang, F.; Inganäs, O.; Andersson, M.R. An easily synthesized blue polymer for high-performance polymer solar cells. Adv. Mater. 2010, 22, 5240–5244. [Google Scholar] [CrossRef] [PubMed]
  16. Chang, D.W.; Lee, H.J.; Kim, J.H.; Park, S.Y.; Park, S.-M.; Dai, L.; Baek, J.-B. Novel quinoxaline-based organic sensitizers for dye-sensitized solar cells. Org. Lett. 2011, 13, 3880–3883. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, X.; Shim, J.W.; Tiwari, S.P.; Zhang, Q.; Norton, J.E.; Wu, P.-T.; Barlow, S.; Jenekhe, S.A.; Kippelen, B.; Brédas, J.-L.; et al. Dithienopyrrole–quinoxaline/pyridopyrazinedonor-acceptorpolymers: Synthesis and electrochemical, optical, charge-transport, and photovoltaic properties. J. Mater. Chem. 2011, 21, 4971–4982. [Google Scholar] [CrossRef]
  18. Zhuang, W.; Zhen, H.; Kroon, R.; Tang, Z.; Hellström, S.; Hou, L.; Wang, E.; Gedefaw, D.; Inganäs, O.; Zhang, F.; et al. Molecular orbital energy level modulation through incorporation of selenium and fluorine into conjugated polymers for organic photovoltaic cells. J. Mater. Chem. A 2013, 1, 13422–13425. [Google Scholar] [CrossRef]
  19. Wang, K.; Zhang, Z.-G.; Fu, Q.; Li, Y. Synthesis and photovoltaic properties of a D–A copolymer based on the 2,3-di(5-hexylthiophen-2-yl)quinoxaline acceptor unit. Macromol. Chem. Phys. 2014, 215, 597–603. [Google Scholar] [CrossRef]
  20. Chen, S.; Zhang, Q.; Zhang, J.; Gu, J.; Zhang, L. Synthesis of two conjugated polymers as TNT chemosensor materials. Sens. Actuators B 2010, 149, 155–160. [Google Scholar] [CrossRef]
  21. Brock, E.D.; Lewis, D.M.; Yousaf, T.I.; Harper, H.H. Reactive Dye Compounds. U.S. Patent WO1999051688 A1, 14 October 1999. [Google Scholar]
  22. Lindner, B.D.; Zhang, Y.; Hoefle, S.; Berger, N.; Teusch, C.; Jesper, M.; Hardcastle, K.I.; Qian, X.; Lemmer, U.; Colsmann, A.; et al. N-fused quinoxalines and benzoquinoxalines as attractive emitters for organic light emitting diodes. J. Mater. Chem. C 2013, 1, 5718–5724. [Google Scholar] [CrossRef]
  23. Thomas, K.R.J.; Velusamy, M.; Lin, J.T.; Chuen, C.-H.; Tao, Y.-T. Chromophore-labeled quinoxaline derivatives as efficient electroluminescent materials. Chem. Mater. 2005, 17, 1860–1866. [Google Scholar] [CrossRef]
  24. Ayaz, M.; Xu, Z.; Hulme, C. Novel succinct routes to quinoxalines and 2-benzimidazolylquinoxalines via the Ugi reaction. Tetrahedron Lett. 2014, 55, 3406–3409. [Google Scholar] [CrossRef] [PubMed]
  25. Srinivas, C.; Kumar, C.N.S.S.P.; Rao, V.J.; Palaniappan, S. Efficient, convenient and reusable polyaniline-sulfate salt catalyst for the synthesis of quinoxaline derivatives. J. Mol. Catal. A Chem. 2007, 265, 227–230. [Google Scholar] [CrossRef]
  26. More, S.V.; Sastry, M.N.V.; Yao, C.-F. Cerium(IV) ammonium nitrate(CAN) as a catalyst in tap water: A simple, proficient and green approach for the synthesis of quinoxalines. Green Chem. 2006, 8, 91–95. [Google Scholar] [CrossRef]
  27. Bhosale, R.S.; Sarda, S.R.; Ardhapure, S.S.; Jadhav, W.N.; Bhusare, S.R.; Pawar, R.P. An efficient protocol for the synthesis of quinoxaline derivatives at room temperature using molecular iodine as the catalyst. Tetrahedron Lett. 2005, 46, 7183–7186. [Google Scholar] [CrossRef]
  28. Zhao, Z.; Wisnoski, D.D.; Wolkenberg, S.E.; Leister, W.H.; Wang, Y.; Lindsley, C.W. General microwave-assisted protocols for the expedient synthesis of quinoxalines and heterocyclic pyrazines. Tetrahedron Lett. 2004, 45, 4873–4876. [Google Scholar] [CrossRef]
  29. Jafarpour, M.; Rezaeifard, A.; Danehchin, M. Easy access to quinoxaline derivatives using alumina as an effective and reusable catalyst under solvent-free conditions. Appl. Catal. A Gen. 2011, 394, 48–51. [Google Scholar] [CrossRef]
  30. Mahesh, R.; Dhar, A.K.; Sasank, T.V.N.V.T.; Thirunavukkarsu, S.; Devadoss, T. Citric acid: An efficient and green catalyst for rapid one pot synthesis of quinoxaline derivatives at room temperature. Chin. Chem. Lett. 2011, 22, 389–392. [Google Scholar] [CrossRef]
  31. Lü, H.-Y.; Yang, S.-H.; Deng, J.; Zhang, Z.-H. Magnetic Fe3O4 nanoparticles as new, efficient, and reusable catalysts for the synthesis of quinoxalines in water. Aust. J. Chem. 2010, 63, 1290–1296. [Google Scholar] [CrossRef]
  32. Niknam, K.; Saberi, D.; Mohagheghnejad, M. Silica bonded S-sulfonic acid: A recyclable catalyst for the synthesis of quinoxalines at room temperature. Molecules 2009, 14, 1915–1926. [Google Scholar] [CrossRef] [PubMed]
  33. Cai, J.-J.; Zou, J.-P.; Pan, X.-Q.; Zhang, W. Gallium(III) triflate-catalyzed synthesis of quinoxaline derivatives. Tetrahedron Lett. 2008, 49, 7386–7390. [Google Scholar] [CrossRef]
  34. Yadav, J.S.; Subba Reddy, B.V.; Premlatha, K.; Shankar, K.S. Bismuth(III)-catalyzed rapid synthesis of 2,3-disubstituted quinoxalines in water. Synthesis 2008, 3787–3792. [Google Scholar] [CrossRef]
  35. Hille, T.; Irrgang, T.; Kempe, R. The Synthesis of benzimidazoles and quinoxalines from aromatic diamines and alcohols by iridium-catalyzed acceptorless dehydrogenative alkylation. Chem. Eur. J. 2014, 20, 5569–5572. [Google Scholar] [CrossRef] [PubMed]
  36. Cho, C.S.; Oh, S.G. A new ruthenium-catalyzed approach for quinoxalines from o-phenylenediamines and vicinal-diols. Tetrahedron Lett. 2006, 47, 5633–5636. [Google Scholar] [CrossRef]
  37. Jeena, V.; Robinson, R.S. An environmentally friendly, cost effective synthesis of quinoxalines: The influence of microwave reaction conditions. Tetrahedron Lett. 2014, 55, 642–645. [Google Scholar] [CrossRef]
  38. Sithambaram, S.; Ding, Y.; Li, W.; Shen, X.; Gaenzler, F.; Suib, S.L. Manganese octahedral molecular sieves catalyzed tandem process for synthesis of quinoxalines. Green Chem. 2008, 10, 1029–1032. [Google Scholar] [CrossRef]
  39. Robinson, R.S.; Taylor, R.J.K. Quinoxaline synthesis from α-hydroxy ketones via a tandem oxidation process using catalysed aerobic oxidation. Synlett 2005, 1003–1005. [Google Scholar]
  40. Kim, S.Y.; Park, K.H.; Chung, Y.K. Manganese(IV) dioxide-catalyzed synthesis of quinoxalines under microwave irradiation. Chem. Commun. 2005, 1321–1323. [Google Scholar] [CrossRef] [PubMed]
  41. Raw, S.A.; Wilfred, C.D.; Taylor, R.J.K. Tandem oxidation processes for the preparation of nitrogen-containing heteroaromatic and heterocyclic compounds. Org. Biomol. Chem. 2004, 2, 788–796. [Google Scholar] [CrossRef] [PubMed]
  42. Aparicio, D.; Attanasi, O.A.; Filippone, P.; Ignacio, R.; Lillini, S.; Mantellini, F.; Palacios, F.; Santos, J.M. Straightforward access to pyrazines, piperazinones, and quinoxalinesby reactions of 1,2-diaza-1,3-butadienes with 1,2-diamines under solution, solvent-free, or solid-phase conditions. J. Org. Chem. 2006, 71, 5897–5905. [Google Scholar] [CrossRef] [PubMed]
  43. Ibrahim, M.M.; Grau, D.; Hampel, F.; Tsogoeva, S.B. α-Nitro epoxides in organic synthesis: development of a one-pot organocatalytic strategy for the synthesis of quinoxalines. Eur. J. Org. Chem. 2014, 1401–1405. [Google Scholar] [CrossRef]
  44. Antoniotti, S.; Dunach, E. Direct and catalytic synthesis of quinoxaline derivatives from epoxides and ene-1,2-diamines. Tetrahedron Lett. 2002, 43, 3971–3973. [Google Scholar] [CrossRef]
  45. Nguyen, T.B.; Retailleau, P.; Al-Mourabit, A. A simple and straightforward approach to quinoxalines by iron/sulfur-catalyzed redox condensation of o-nitroanilines and phenethylamines. Org. Lett. 2013, 15, 5238–5241. [Google Scholar] [CrossRef] [PubMed]
  46. Xu, Y.; Wan, X. Ruthenium-catalyzed oxidation of alkynes to 1,2-diketones under room temperature and one-pot synthesis of quinoxalines. Tetrahedron Lett. 2013, 54, 642–645. [Google Scholar] [CrossRef]
  47. Shi, S.; Wang, T.; Yang, W.; Rudolph, M.; Hashmi, A.S.K. Gold-catalyzed synthesis of glyoxals by oxidation of terminal alkynes: one-pot synthesis of quinoxalines. Chem. Eur. J. 2013, 19, 6576–6580. [Google Scholar] [CrossRef] [PubMed]
  48. Okumura, S.; Takeda, Y.; Kiyokawa, K.; Minakata, S. Hypervalent iodine(III)-induced oxidative [4+2] annulation of o-phenylenediamines and electron-deficient alkynes: Direct synthesis of quinoxalines from alkyne substrates under metal-free conditions. Chem. Commun. 2013, 49, 9266–9268. [Google Scholar] [CrossRef] [PubMed]
  49. Chen, C.-Y.; Hu, W.-P.; Liu, M.-C.; Yan, P.-C.; Wang, J.-J.; Chung, M.-I. Efficient synthesis of quinoxalines with hypervalent iodine as a catalyst. Tetrahedron 2013, 69, 9735–9741. [Google Scholar] [CrossRef]
  50. Zhang, C.; Xu, Z.; Zhang, L.; Jiao, N. Et3N-catalyzed oxidative dehydrogenative coupling of α-unsubstituted aldehydes and ketones with aryl diamines leading to quinoxalines using molecular oxygen as oxidant. Tetrahedron 2012, 68, 5258–5262. [Google Scholar] [CrossRef]
  51. Campelo, J.M.; Luna, D.; Luque, R.; Marinas, J.M.; Romero, A.A. Sustainable preparation of supported metal nanoparticles and their applications in catalysis. ChemSusChem 2009, 2, 18–45. [Google Scholar] [CrossRef] [PubMed]
  52. White, R.J.; Luque, R.; Budarin, V.; Clark, J.H.; Macquarrie, D.J. Supported metal nanoparticles on porous materials. Methods and applications. Chem. Soc. Rev. 2009, 38, 481–494. [Google Scholar] [CrossRef] [PubMed]
  53. Astruc, D.; Lu, F.; Aranzaes, J.R. Nanoparticles as recyclable catalysts: The frontier between homogeneous and heterogeneous catalysis. Angew. Chem. Int. Ed. 2005, 44, 7852–7872. [Google Scholar] [CrossRef] [PubMed]
  54. Rajabi, F.; Feiz, A.; Luque, R. An efficient synthesis of coumarin derivatives using a SBA-15 supported Cobalt(II) nanocatalyst. Catal. Lett. 2015, 145, 1621–1625. [Google Scholar] [CrossRef]
  55. Rajabi, F.; Raessi, M.; Arancon, R.A.D.; Saidi, M.R.; Luque, R. Supported cobalt oxide nanoparticles as efficient catalyst in esterification and amidation reactions. Catal. Commun. 2015, 59, 122–126. [Google Scholar] [CrossRef]
  56. Heravi, M. M.; Taheri, S.; Bakhtiari, K.; Oskooie, H. A. Zn[(L)proline]: A powerful catalyst for the very fast synthesis of quinoxaline derivatives at room temperature. Catal. Commun. 2007, 8, 1341–1344. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds 3al are available from the authors.

Share and Cite

MDPI and ACS Style

Rajabi, F.; Alves, D.; Luque, R. An Efficient and Recyclable Nanoparticle-Supported Cobalt Catalyst for Quinoxaline Synthesis. Molecules 2015, 20, 20709-20718. https://doi.org/10.3390/molecules201119731

AMA Style

Rajabi F, Alves D, Luque R. An Efficient and Recyclable Nanoparticle-Supported Cobalt Catalyst for Quinoxaline Synthesis. Molecules. 2015; 20(11):20709-20718. https://doi.org/10.3390/molecules201119731

Chicago/Turabian Style

Rajabi, Fatemeh, Diego Alves, and Rafael Luque. 2015. "An Efficient and Recyclable Nanoparticle-Supported Cobalt Catalyst for Quinoxaline Synthesis" Molecules 20, no. 11: 20709-20718. https://doi.org/10.3390/molecules201119731

Article Metrics

Back to TopTop