Aqua mediated SnO2 nanoparticles: A recyclable and benign catalyst for the synthesis of Quinoxalines

Document Type : Reasearch Paper

Authors

1 Department of Chemistry, Ayatollah Amoli Branch, Islamic Azad University, Amol, Iran.

2 School of Science and Technology, The University of Georgia, Tbilisi, Georgia.

Abstract

An efficient and mild synthesis of quinoxalines in cludingcyclo-condensation of 1, 2-phenylenediamine and 1, 2-diketonesin the presence of1mol% catalytic amount of SnO2 nanoparticles (1 mol%) in water at room temperature is established. On the whole, this study introduced at this point is substantial in terms of using water as solvent, low reaction time (5 to 10 minutes), high yields of products (85-88%), reusability of catalyst (three cycles), eco-friendliness, effortlessness of performance and it displays along the line of green chemistry.

Keywords


INTRODUCTION

Lately, being focused on green chemistry by using environmentally mild reagents and conditions is one of the most attractive improvements in synthesis of broadly applied organic compounds. Therefore, the use of water as a favorable solvent for organic reactions has received significant attention [1-3]. Water is a valuable solvent in many methods and carrying out organic reactions in this medium is of great interest [4]. It is certainly the most low-cost among numerous solvents applied in organic synthesis. The absence of explosive, inflammable, mutagenic and carcinogenic properties is a satisfactory aspect of water in laboratories. Additionally, water is considered as one of the appropriate solvents from an eco-friendly point of opinion [5].

Due to developing green concerns, the progress of clean synthetic processes has become essential and serious investigation. For this reason, heterogeneous organic reactions have many advantages, for example recycling, comfort of handling separation and eco-friendly safe removal [6, 7].

Quinoxalines represent a substantial category of nitrogen heterocyclic compounds as they constitute valuable intermediates in organic synthesis and are appropriate dyes [8]. Some of them show biological activities containing antiprotozoal, anti-bacterial, anti-HIV, anti-inflammatory, anti-viral, antidepressant, anti-cancer and as kinase inhibitors [9-27]. In addition, they were displayed to be NMDA receptor antagonist, PDGF-RTK inhibitor, IL-8 receptor antagonist and5-HT3receptor antagonist in the similar mode [28-33]. Several synthetic approaches and catalysts have been studied for the preparation of quinoxalines such as 6-amino-2,3-dichloroquinoxaline loaded on AMEBA resin [34], lead oxide [35], (NH4)6Mo7O24.4H2O [36], [PBBS] and [TBBDA] [37], Zr(OTf)4 [38], silica bonded S-sulfonic acid [39], solid-phase synthesis on SynphaseTM Lanterns [40], nano-flake ZnO [41],[MIMPS]3PW12O40, [TEAPS]3PW12O40 [42], Basolites [43] and Porous carbons [44].

The investigation of metal oxides has attracted the consideration of materials experts because of their mechanical, electrical, magnetic, optical, catalytic suitable relations [45] and the removal of toxic metals and dyes [46], which make them scientifically valuable. The efficacy of metal nanoparticles as reagents or catalysts in chemistry has remarkable potential in organic synthesis and materials science [47-50]. Tin oxide is a significant material owing to its properties for example strong physical and chemical interaction with adsorbed species, low operating temperature, high degree of transparency in the visible spectrum and strong thermal stability in air (up to 500 oC) [51]. Furthermore, tin oxide is broadly applied in electrochemical properties [52] and catalysis field [53-58].

As part of our efforts to investigate the usefulness of nano metal oxide catalysts for the synthesis of organic and heterocyclic compounds [53-63], we report an efficient process for the synthesis of α quinoxalines from the cyclo-condensation reaction between 1, 2-phenylenediamine and 1, 2-diketones by using 1 mol% aqua mediated SnO2 nanoparticles at room temperature (Fig. 1).

MATERIALS AND METHODS

General methods

SnO2 nanoparticles were purchased from commercial centers and characterized by using various techniques [53-55, 63]. All reagents were purchased from Merck and Aldrich and applied without more purification. All solvents were reagent grade. All yields refer to the isolated products after purification. The selected products were identified by comparison with authentic samples and by using spectroscopic data include FT-IR spectra, 1H NMR and 13C NMR analyses and melting point. FT-IR spectra were recorded on FT-IR Bruker (WQF-510) spectrometer. 1H NMR and 13C NMR spectra were recorded on Bruker DRX‐400 MHz by using TMS as the internal standard. All melting points were taken on a Thermo Scientific apparatus and were uncorrected. TLC was attained on aluminum sheets silica gel F254.

General procedure for the synthesis of quinoxalines by using aqua mediated SnO2 nanoparticles

To a mixture of 1, 2-phenylenediamine (1.0 mmol) and 1, 2-diketones (1.0 mmol) was added SnO2 nanoparticles (1 mol%) in H2O (2 mL). The mixture was stirred at room temperature for the known time (Table 2). The progress of the reaction was checked by TLC (n-hexane/ethyl acetate; 5 : 2). After completion of the reaction, water was removed and the product was heated in ethanol. SnO2 nanoparticles was filtered (the product was soluble in hot ethanol and the catalyst was insoluble). Finally, the crude product was purified by recrystallized from ethanol to afford the pure product.

Spectral data for selected products

2, 3-Diphenylquinoxaline (Table 2, 3a): Yellow crystalline solid, M.p.: 129-131 oC; Yield: 86%; FT-IR (KBr) (υmax, cm-1): 2956, 1695, 1460, 1377; 1H NMR (400 MHz, DMSO-d6): δ = 7.50-7.67 (dd, 4H, J1 = 16 Hz and J2 = 8.0 Hz, Ar–H), 7.92-7.94 (d, 2H, = 8.0 Hz, Ar–H), 7.95-7.99 (dd, 4H, J1= 16 Hz and J2 = 8.0 Hz, Ar–H), 8.02-8.04 (d, 2H, J = 8.0 Hz, Ar–H), 8.69-8.71 (d, 2H, = 8.0 Hz, Ar–H); 13C NMR (100 MHz, DMSO-d6): 148.5, 146.3, 131.2, 131.1, 130.1, 129.5, 129.1, 127.7.

Dibenzo[a,c]phenazine (Table 2, 3e): White crystalline solid, M.p.: 225-228 oC; Yield: 88%; FT-IR (KBr) (υmax, cm-1): 3051, 1601, 1475, 1354, 1221; 1H NMR (400 MHz, DMSO-d6): δ = 7.73-7.75 (d, 2H, = 8.0 Hz, Ar–H), 7.76-7.82 (dd, 2H, J1 = 16 Hz and J2 = 8.0 Hz, Ar–H), 7.84-7.88 (dd, 2H, J1= 16 Hz and J2 = 8.0 Hz, Ar–H), 8.31-8.36 (dd, 2H, J1= 16 Hz and J2 = 8.0 Hz, Ar–H), 8.51-8.56 (d, 2H, J = 8.0 Hz, Ar–H), 9.38-9.41 (d, 2H, = 8.0 Hz, Ar–H); 13C NMR (100 MHz, DMSO-d6): 122.9, 126.2, 127.9, 129.4, 129.7, 130.2, 132.0, 142.2, 142.4, 146.3.

RESULTS AND DISCUSSION

Firstly, we studied the activity of SnO2 nanoparticles in the cyclo-condensation of 1, 2-phenylenediamine and phenanthrene-9, 10-dione as a model reaction (Table 1). To our enchantment, the expected product 3e was achieved in 88% isolated yield after5 minutes in the presence of a 1 mol% catalytic amount of Aqua mediated SnO2 nanoparticles at room temperature (Table 1, entry 1).No desired product was synthesized in the absence of SnO2 nanoparticles.

It is identified that the reaction medium plays a significant role in the catalytic reaction. The investigation of the effect of the nature of solvent for this reaction by using SnO2 nanoparticles was performed at room temperature in several solvents (Table 1, entries 1-4). The highest reaction activity was attained in the system by using water as a solvent in comparison to other solvents under same reaction conditions. The above model reaction was achieved under solvent-free condition with the SnO2 nanoparticles to provide low yield.

Catalyst concentration is an important factor that fully affects the reaction rate and product yield. To optimize the catalyst loading, 0.5 mol%, 2 mol% and 3 mol% of SnO2nanoparticleswas studied but the yields were not appropriate (Table 1, entries 5-7). A 1 mol% loading of SnO2 nanoparticles was satisfactory to push the reaction forward and higher amounts of catalyst did not increase the yields meaningfully.

We next studied the effect of temperature on the rate of model reaction. For this aim, the reaction was performed in higher temperatures and under reflux condition. But, by increasing the temperature unsuccessful to improve the reaction rate significantly. As it happens, higher temperatures were lowered the product yield rather, attended by approximately impurities. Moreover, the appropriate result was attained with a 1 : 1 molar ratio for 1, 2-phenylenediamine and 9, 10-phenantrolinedione.

The SnO2 nanoparticles catalyst could be recycled and reused without decrease of catalytic activity. At the end of the reaction, hot ethanol was added to the reaction mixture and SnO2 nanoparticles was filtered (the product was soluble in hot ethanol and the catalyst was insoluble), washed three times with hot ethanol, dried at 80 oC for 120 minutes, and reused. The recycled catalyst was used to the synthesis of 3e and the yield was reserved at 86-88% via three cycles of catalyst recovering.

To consider the scope and overview of the method, a variety of 1, 2-diketones bearing electroneutral and electron-releasing groups were reacted with 1, 2-phenylenediamine as nucleophilic substrates under optimized reaction conditions and the results are showed in Table 2. No apparent electronic effect of the substituents of 1, 2-diketones was obvious. The corresponding quinoxalines were produced efficiently with facility and suitable yields (85-88%).

The proposed mechanism for this reaction is displayed in Fig. 2 [37-39]. Initially, SnO2nanoparticles as a catalyst activates the carbonyl functional group of 1, 2-diketones 2 to give intermediate 2’. At that point, nucleophilic attacks of 1, 2-phenylenediamine 1 on the intermediate 2’ to provide intermediate 4. The condensation of intermediate leads to the form of intermediate 5 via elimination of one molecule of water. In the next step, intra-cyclization of intermediate via nucleophilic attack of amine group on the activated carbonyl functional group to afford intermediate 6 which rearranges through elimination of second molecule of water to quinoxalines 3.

To display the noteworthy properties of our investigation, we have compared our result with the known data from the other works for the synthesis of 2, 3-diphenylquinoxaline (Table 2, 3a). For comparison, this target molecule was selected, as shown in Table 3. No use of hazardous solvent, short reaction time and high yields of target product make this process as a valuable method for the synthesis of library of quinoxalines.

CONCLUSION

In summary, we have introduced a facile and eco-friendly mild process for the synthesis of quinoxalines. This cyclo-condensation reaction between 1, 2-phenylenediamine and 1,2-diketones is efficiently catalyzed by aqua mediated SnO2 nanoparticles at room temperature. Easy operation, improved rates, high isolated yields of the pure products and benign reaction conditions are important advantages of the procedure presented here. More uses of SnO2 nanoparticles on the extension of this approach are continuing in our group.

SUPPORTING INFORMATION

The supporting information includes spectral images of FT-IR, 1H NMR and 13C NMR of selected product.

ACKNOWLEDGEMENTS

The authors are thankful for the facilities provided to carry out research in chemistry research laboratory at Ayatollah Amoli Branch, Islamic Azad University.

CONFLICT OF INTEREST

The authors declare no conflict of interest.

 

[1] Chen H., Shi D., (2010), Efficient one-pot synthesis of novel spirooxindole derivatives via three-component reaction in aqueous medium. J. Comb. Chem. 12: 571-576.
[2] Safari J., Banitaba S. H., Khalili S. D., (2012), Ultrasound-promoted an efficient method for one-pot synthesis of 2-amino-4, 6-diphenylnicotinonitriles in water: A rapid procedure without catalyst. Ultrason. Sonochem. 19: 1061-1069.
[3] Zou Y., Wu H., Hu Y., Liu H., Zhao X., Ji H., Shi D., (2011), A novel and environment-friendly method for preparing dihydropyrano [2, 3-c] pyrazoles in water under ultrasound irradiation. Ultrason. Sonochem. 18: 708-712.
[4] Simon M. O., Li C. J., (2012), Green chemistry oriented organic synthesis in water. Chem. Soc. Rev. 41: 1415-1427.
[5] Lindstrom, U. M., (2007), Organic Reactions in Water, Principles, Strategies and Applications. p. 60. Oxford, UK: Blackwell Publishing.
[6] Lu J., Toy P. H., (2009), Organic polymer supports for synthesis and for reagent and catalyst immobilization. Chem. Rev. 109: 815-838.
[7] Zhang Z. H., Lü H. Y., Yang S. H., Gao J. W., (2010), Synthesis of 2, 3-dihydroquinazolin-4(1H)-ones by three-component coupling of isatoic anhydride, amines, and aldehydes catalyzed by magnetic Fe3O4 nanoparticles in water. J. Comb. Chem. 12: 643-646.
[8]Brock E. D., Lewis D. M., Yousaf T. I., Harper H. H., (1999), The Procter and Gamble Company USA, WO9951688.
[9] Sakata G., Makino K., (1988), Regent progress in the quinoxaline chemistry. Synthesis and biological activity. Heterocycles. 27: 2481-2515.
[10] Loriga M., Piras S., Sanna P., Paglietti G., (1999), Quinoxaline chemistry. Part 7. 2-[aminobenzoates]- and 2-[aminobenzoylglutamate]-quinoxalines as classical antifolate agents. Synthesis and evaluation of in vitro anticancer, anti-HIV and antifungal activity. Farmaco. 52: 157-166.
[11] Seitz L. E., Suling W. J., Reynolds R. C., (2002), Synthesis and antimycobacterial activity of pyrazine and quinoxaline derivatives. J. Med. Chem. 45: 5604-5606.
[12] Kim Y. B., Kim Y. H., Park J. Y., Kim S. K., (2004), Synthesis and biological activity of new quinoxaline antibiotics of echinomycin analogues. Bioorg. Med. Chem. Lett. 14: 541-544.
[13] Hui X., Desrivot J., Bories C., Loiseau P. M., Franck X., Hocquemiller R., Figadere B., (2006), Synthesis and antiprotozoal activity of some new synthetic substituted quinoxalines. Bioorg. Med. Chem. Lett. 16: 815-820.
[14] Lindsley C. W., Zhao Z., Leister W. H., Robinson R. G., Barnett S. F., Defeo-Jones D., Jones R. E., Hartman G. D., Huff J. R., Huber H. E., Duggan M. E., (2005), Allosteric Akt (PKB) inhibitors: Discovery and SAR of isozyme selective inhibitors. Bioorg. Med. Chem. Lett. 15: 761-764.
[15] Labarbera D. V., Skibo E. B., (2005), Synthesis of imidazo [1, 5, 4-de] quinoxalin-9-ones, benzimidazole analogues of pyrroloimino quinone marine natural products. Bioorg. Med. Chem. 13: 387-395.
[16] Sarges R., Howard H. R., Browne R. G., Lebel L. A., Seymour P. A., Koe B. K., (1990), 4-Amino [1, 2, 4] triazole [4, 3-a] quinoxalines. A novel class of potent adenosine receptor antagonists and potential rapid-onset antidepressants. J. Med. Chem. 33: 2240-2254.
[17] Srinivas C., Kumar C. N. S. S. P., Rao V. J., Palaniappan S., (2007), Efficient, convenient and reusable polyaniline-sulfate salt catalyst for the synthesis of quinoxaline derivatives. J. Mol. Catal. A: Chem. 265: 227-230.
[18]  Ghomsi N. T., Ahabchane N. E. H., Es-Safi N. E., Garrigues B., Essassi E. M., (2007), Synthesis and spectroscopic structural elucidation of new quinoxaline derivatives. Spectroscopy Lett. 40: 741-751.
[19] Montana M., Mathias F., Terme T., Vanelle P., (2019), Antitumoral activity of quinoxaline derivatives: A systematic review. Eur. J. Med. Chem. 163: 136-147.
[20] Srinivasarao S., Nandikolla A., Suresh A., Ewa A. K., Głogowska A., Ghosh B., Kumar B. K., Murugesan S., Pulya S., Aggarwal H., Chandra-Sekhar K. V. G., (2020), Discovery of 1, 2, 3-triazole based quinoxaline-1, 4-di-N-oxide derivatives as potential anti-tubercular agents. Bioorg. Chem. 100: 103955-103961.
[21] Benhiba F., Benzekri Z., Guenbour A., Tabyaoui M., Bellaouchou A., Boukhris S., Oudda H., Warad I., Zarrouk A., (2020), Combined electronic/atomic level computational, surface (SEM/EDS), chemical and electrochemical studies of the mild steel surface by quinoxalines derivatives anti-corrosion properties in 1 mol⋅L-1HCl solution. Chinese J. Chem. Eng. 03: 002-009.
[22] Laabaissi T., Benhiba F., Missioui M., Rouifi Z., Rbaa M., Oudda H., Ramli Y., Guenbour A., WaradI., Zarrouk A., (2020), Coupling of chemical, electrochemical and theoretical approach to study the corrosion inhibition of mild steel by new quinoxaline compounds in 1 M HCl. Heliyon. 6: e03939.
[23] Jone-Kirubavathy S., Chitra S., (2020), Synthesis, characterization, DFT, In-vitro anti-microbial, cytotoxicity evaluation, and DNA binding interactions of transition metal complexes of quinoxaline schiff base ligand. Mater. Today-Proc. 04: 699-711.
[24] Chauhan D. S., Singh P., Quraishi M. A., (2020), Quinoxaline derivatives as efficient corrosion inhibitors: Current status, challenges and future perspectives. J. Mol. Liq. 320: 114387-114392.
[25] Seung Chan K.,Pulla Reddy B., Ha Na Y., So Young K., Jun Min J., Yeon Su K.,Gi Min P., Sang Ho M., In Su K., Young Hoon J., (2020), Synthesis and biological evaluation of quinoxaline derivatives as specific c-Met kinase inhibitors. Bioorg. Med. Chem. Lett. 127189.
[26] Ono Y., Ninomiya M., Kaneko D., Sonawane A. D., Udagawa T., Tanaka K., Nishina A., Koketsu M., (2020), Design and synthesis of quinoxaline-1, 3, 4-oxadiazole hybrid derivatives as potent inhibitors of the anti-apoptotic Bcl-2 protein. Bioorg. Chem. 104: 104245-104251.
[27] Sagar S. R., Singh D. P., Das R. D., Panchal N. B., Sudarsanam V., Nivsarkar M., Vasu K. K., (2019), Pharmacological investigation of quinoxaline-bisthiazoles as multitarget directed ligands for the treatment of Alzheimer’s disease. Bioorg. Chem. 89: 102992-102998.
[28] Monge A., Palop J. A., Del Castillo J. C., Caldero J. M., Roca J., Romero G., Del Rio J., Lasheras B., (1993), Novel antagonists of 5-HT3 receptors. Synthesis and biological evaluation of piperazinyl quinoxaline derivatives. J. Med. Chem. 36: 2745-2750.
[29] Baudy R. B., Greenblatt L. P., Jirkovsky I. L., Conklin M., Russo R. J., Bramlett D. R., Emrey T. A., Simmonds J. T., Kowal D. M., Stein R. P., Tasse R. P., (1993), Potent quinoxaline-spaced phosphono alpha-amino acids of the AP-6 type as competitive NMDA antagonists: Synthesis and biological evaluation. J. Med. Chem. 36: 331-342.
[30] Gazit A., App H., McMahon G., Chen J., Levitzki A., Bohmer F. D., (1996), Tyrphostins. 5. Potent inhibitors of platelet-derived growth factor receptor tyrosine kinase:  Structure-activity relationships in quinoxalines, quinolines, and indoletyrphostins. J. Med. Chem. 39: 2170-2177.
[31] Myers M. R., He W., Hanney B., Setzer N., Maguire M. P., Zulli A., Bilder G., Galzcinski H., Amin D., Needle S., Spada A. P., (2003), Potent quinoxaline-based inhibitors of PDGF receptor tyrosine kinase activity. Part 1: SAR exploration and effective bioisosteric replacement of a phenyl substituent. Bioorg. Med. Chem. Lett. 13: 3091-3095.
[32] He W., Myers M. R., Hanney B., Spada A. P., Bilder G., Galzcinski H., Amin, D., Needle S., Page K., Jayyosi Z., Perrone M. H., (2003), Potent quinoxaline-based inhibitors of PDGF receptor tyrosine kinase activity. Part 2: The synthesis and biological activities of RPR127963 an orally bioavailable inhibitor. Bioorg. Med. Chem. Lett. 13: 3097-3100.
[33] Li J. J., Carson K. G., Trivedi B. K., Yue W. S., Ye Q., Glynn R. A., Miller S. R., Connor D. T., Roth B. D., Luly J. R., Low J. E., Heilig D. J., Yang W., Qin S., Hunt S., (2003), Synthesis and structure-activity relationship of 2-amino-3-heteroaryl-quinoxalines as non-peptide, small-Molecule antagonists for interleukin-8 receptor. Bioorg. Med. Chem. 11: 3777-3790.
[34] Jeon M. K., Kim D. S., La H. J., Gong Y. D., (2006), Solid-phase synthesis of quinoxaline derivatives using 6-amino-2, 3-dichloroquinoxaline loaded on AMEBA resin. Tetrahedron Lett. 46: 4979-4983.
[35] Kotharkar S. A., Shinde D. B., (2006), Lead oxide (PbO) mediated synthesis of quinoxaline. J. Iran. Chem. Soc. 3: 267-271.
[36] Hasaninejad A., Zare A., Mohammadizadeh M. R., Karami Z., (2009), Synthesis of quinoxaline derivatives via condensation of aryl-1, 2-diamines with 1, 2-Diketones using (NH4)6Mo7O24.4H2O as an efficient, mild and reusable catalyst. J. Iran. Chem. Soc. 6: 153-158.
[37] Ghorbani-Vaghei R., Hajinazari S., (2013), Poly (N, N-dibromo-N-ethyl-benzene-1, 3-disulphonamide) and N, N, N, N-tetrabromobenzene-1, 3-disulphonamide as novel catalysts for synthesis of quinoxaline derivatives. J. Chem. Sci. 125: 353-358.
[38] Kolvari E., Zolfigol M. A., Koukabi N., Gilandust M., Kordi A. V., (2013), Zirconium triflate: An efficient catalyst for the synthesis of quinolines and quinoxalines. J. Iran. Chem. Soc. 10: 1183-1191.
[39] Niknam K., Saberi D., Mohagheghnejad M., (2009), Silica bonded S-sulfonic acid: a recyclable catalyst for the synthesis of quinoxalines at room temperature. Molecules. 14: 1915-1926.
[40] Wu Z., Ede N. J., (2001), Solid-phase synthesis of quinoxalines on SynPhase™ Lanterns. Tetrahedron Lett. 42: 8115-8118.
[41] Hosseini-Sarvari M., (2011), Synthesis of quinolines using Nano-flake ZnO as a new catalyst under solvent-free conditions. J. Iran. Chem. Soc. 8: S119-S128.
[42] Vahdat S. M., Baghery S., (2013), A green and efficient protocol for the synthesis of quinoxaline, benzoxazole and benzimidazole derivatives using heteropolyanion-based ionic liquids: As a recyclable solid catalyst. Comb. Chem. High Throughput Screen. 16: 618-627.
[43] Godino-Ojer M., Shamzhy M., Cejka J., Perez-Mayoral E., (2020), Basolites: A type of metal organic frameworks highly efficient in the one-pot synthesis of quinoxalines from α-hydroxy ketones under aerobic conditions. Catal. Today. 345: 258-266.
[44] Godino-Ojer M., Blazques-Garcia R., Matos I., Bernardo M., Fonseca I. M. Perez-Mayoral E., (2020), Porous carbons-derived from vegetal biomass in the synthesis of quinoxalines. Mechanistic insights. Catal. Today. 354: 90-99.
[45] McCarrol W. H., Ramanujachary K. V. (2005),Oxides: Solid-state chemistry, In King RB (Ed.), Encyclopedia of Inorganic Chemistry 2nd ed. (pp. 4006-4053). Wiley.
[46] Oyewo O. A., Elemike E. E., Onwudiwe D. C., Onyango M. C., (2020), Metal oxide-cellulose nanocomposites for the removal of toxic metals and dyes from wastewater. Int. J. Biol.Macromol. 164: 2477-2496.
[47] Chen M. S., Goodman D. W., (2004), The structure of catalytically active gold on titania. Science. 306: 252-255.
[48] Astruc D., Lu F., Aranzaes J. R., (2005), Nanoparticles as recyclable catalysts: the frontier between homogeneous and heterogeneous catalysis. Angew. Chem. Int. Ed. 44: 7852-7872.
[49] Walters G., Parkin I. P., (2009), The incorporation of noble metal nanoparticles into host matrix thin films: Synthesis, characterization and applications. J. Mater. Chem. 19: 574-590.
[50] Astruc D., (2007), Palladium nanoparticles as efficient green homogeneous and heterogeneous carbon−carbon coupling precatalysts: A unifying view. Inorg. Chem. 46: 1884-1894.
[51] Zhu J., Tay B. Y., Ma J., (2006), Synthesis of mesoporous tin oxide on neutral surfactant templates. J. Mater. Lett. 60: 1003-1010.
[52] Kobayashi H., Uebou Y., Ishida T., Tamura S., Mochizuki S., Mihara T., Tabuchi M., Kageyama H., Yamamoto Y., (2001), Electrochemical property of tin oxide thin film by photo-CVD process. J. Power Sources. 97: 229-231.
[53] Vahdat S. M., Ghafouri Raz Sh., Baghery S., (2014), Application of nano SnO2 as a green and recyclable catalyst for the synthesis of 2-aryl or alkylbenzoxazole derivatives under ambient temperature. J. Chem. Sci. 126: 579-585.
[54] Vahdat S. M., Chekin F., Hatami M., Khavarpour M., Baghery S., Roshan-Kouhi Z., (2013), Synthesis of polyhydroquinoline derivatives via a four-compone. Chin. J. Catal. 34: 758-763.
[55] Zolfigol M. A., Baghery S., Moosavi-Zare A. R., Vahdat S. M., Alinezhad H., Norouzi M., (2015), Design of 1-methylimidazolium tricyanomethanide as the first nanostructured molten salt and its catalytic application in the condensation reaction of various aromatic aldehydes, amides and β-naphthol compared with tin dioxide nanoparticles. RSC Adv. 5: 45027-45037.
[56] Manjunathan P., Marakatti V. S., Chandra P., Kulal A. B., Umbarkar S. B., Ravishankar R., Shanbhag G. V., (2018), Mesoporous tin oxide: An efficient catalyst with versatile applications in acid and oxidation catalysis. Catal. Today. 309: 61-76.
[57] Manjunathan P., Marakatti V. S., Chandra P., Kulal A. B., Umbarkar S. B., Ravishankar R., Shanbhag G. V., (2016), Sulfated tin oxide (STO)-structural properties & application in catalysis: A review. Arab. J. Chem. 9: 550-573.
[58] Zolfigol M. A., Baghery S., Moosavi-Zare A. R., Vahdat S. M., (2015), Synthesis and characterization of new 1-(α-aminoalkyl)- 2-naphthols using pyrazine-1, 4-diium trinitromethanide {[1, 4-DHPyrazine] [C(NO2)3]2} as a novel nano-structured molten salt and catalyst in compared with Ag-TiO2 nano composite. J. Mol. Catal. A: Chem. 409: 216-226.
[59] Maleki B., Baghayeri M., Vahdat S. M., Mohammadzadeh A., Akhoondi S., (2015), Ag@TiO2nanocomposite; synthesis, characterization and its application as a novel and recyclable catalyst for the one-pot synthesis of benzoxazole derivatives in aqueous media. RSC Adv. 5: 46545-46551.
[60] Chekin F., Vahdat S. M., Asadi M. J., (2016), Green synthesis and characterization of cobalt oxide nanoparticles and its electrocatalytic behavior. Russian J. Appl. Chem. 89: 816-822.
[61]     Yazdani S., Hatami M., Vahdat S. M., (2014), The chemistry concerned with the sonochemical-assisted synthesis of CeO2/poly (amic acid) nanocomposites. Turk. J. Chem. 38: 388-401.
[62] Vahdat S. M., Khavarpour M., Mohanazadeh F., (2015), A facile and highly efficient three component synthesis of pyran and chromene derivatives in the presence of nano SnO2 as a catalyst. J. Appl. Chem. 9: 41-46.
[63] Suresh G., Sathishkumar R., Iruson B., Sathyaseelan B., Senthilnathan K., Manikandan E., (2019), Study on structural, luminescence properties and hall effect of SnO2 nanoparticles obtained by a Co-precipitation technique. Int. J. Nano Dimens.10: 242-251.
[64] Das J. P., Sarkar S., (2011), An efficient synthesis of quinoxalines in water mediated by tetraethylammonium bromate. Int. J. Chem. Res. 3: 56-60.
[65] Lin P. Y., Hou R. S., Wang H. M., Kang L. J., Chen L. C., (2009), Hypervalent iodine(III) sulfonate mediated synthesis of quinoxalines in liquid PEG-400. J. Chin. Chem. Soc. 56: 683-687.
[66] Ziarani G. M., Badiei A., Haddadpour M., (2011), Application of sulfonic acid functionalized nanoporous silica (SBA-Pr-SO3H) for one-pot synthesis of quinoxaline derivatives. Int. J. Chem. 3: 87-94.