Next Article in Journal
Coupling Interface Construction of Ni(OH)2/MoS2 Composite Electrode for Efficient Alkaline Oxygen Evolution Reaction
Previous Article in Journal
An Evaluation of the Kinetic Properties Controlling the Combined Chemical and Biological Treatment of Toxic Recalcitrant Organic Compounds from Aqueous Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chitosan Capped Copper Oxide Nanocomposite: Efficient, Recyclable, Heterogeneous Base Catalyst for Synthesis of Nitroolefins

by
Mohamed A. El-Atawy
1,2,
Khaled D. Khalil
2,3,* and
Ali H. Bashal
2
1
Department of Chemistry, Faculty of Science, Alexandria University, Alexandria 21321, Egypt
2
Department of Chemistry, Faculty of Science, Taibah University, Al-Madinah Almunawarah, Yanbu 46423, Saudi Arabia
3
Department of Chemistry, Faculty of Science, Cairo University, Giza 12613, Egypt
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(9), 964; https://doi.org/10.3390/catal12090964
Submission received: 3 August 2022 / Revised: 22 August 2022 / Accepted: 26 August 2022 / Published: 29 August 2022
(This article belongs to the Section Catalysis in Organic and Polymer Chemistry)

Abstract

:
In this article, chitosan copper oxide nanocomposite was synthesized by the solution casting method under microwave irradiation. The nanocomposite solution was microwave irradiated at 300 watt for 3 min under optimal irradiation conditions. By suppressing particle agglomeration, the chitosan matrix was successfully used as a metal oxide stabilizer. The goal of this research was to create, characterize, and test the catalytic potency of these hybrid nanocomposites in a number of well-known organic processes. The prepared CS-CuO nanocomposites were analyzed by different techniques, including Fourier-transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), and field emission scanning electron microscopy (FESEM). Moreover, energy-dispersive X-ray spectroscopy (EDS) was used to measure the copper content in the prepared nanocomposite film. The finger-print peaks in the FTIR spectrum at around 632–502 cm−1 confirmed the existence of the CuO phase. The CS-CuO nanocomposite has been shown to be an efficient base promoter for nitroolefin synthesis via the nitroaldol reaction (Henry reaction) in high yields. The reaction variables were studied to improve the catalytic approach. Higher reaction yields, shorter reaction times, and milder reaction conditions are all advantages of the technique, as is the catalyst’s reusability for several uses.

Graphical Abstract

1. Introduction

Due to their unique features and vast range of applications as potent catalytic agents in various organic transformations, nano catalysis has attracted the interest of many researchers [1,2,3]. Nano catalysts have emerged as a low-cost, long-term alternative to hazardous commercial catalysts since they contain nanosized particles with a large, exposed surface area, which increases catalytic activity and regioselectivity in a variety of chemical processes [4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19].
For the above-mentioned reasons, and because of their vast range of applications in numerous disciplines, many efforts have been made to synthesize new hybrid materials, including metal oxide nanoparticles. In particular, these hybrid nanocomposites confer a considerable synergistic effect, as compared to their constituents, that increases their potential use in many applications, especially in the field of heterogenous catalysis [14,15,16,17,18,19,20,21]. Solid catalysts have unique characteristics such as a large surface area, a high atom economy, and the ability to be easily removed from the product, recovered, and reused multiple times without losing catalytic activity. Chitosan, a natural bio-stabilizer, has recently been employed as a good template for fabricating metal oxide nanoparticles [20,21,22,23,24,25]. The existence of many binding sites, notably NH2 and OH groups along the chains, is responsible for chitosan’s high adsorption capability. That explains why it is so good at forming stable complexes with metal ions and metal oxides [15,16,17,18,20,21]. On the other hand, among all metal oxides, copper oxide nanoparticles (CuONPs) have gained more attention due to their unique properties, as an inexpensive and efficient heterogeneous catalyst, and various applications [26].
Copper oxide nanoparticles’ catalytic activity allowed them to be used as a potent base catalyst in aromatic nucleophilic substitution processes [6,7] and in various heterocyclic synthesis methods [8,9,10,11,12,13,14,15,16]. In nucleophilic substitutions, Jammi et al. described a simple, general, and efficient procedure for the cross-coupling of nitrogen, oxygen, and sulfur nucleophiles with aryl iodides using CuO nanoparticles under ligand-free conditions [6]. On the other hand, Kidwai et al. have developed an economical method for the synthesis of 3-arylpentane-2,4-diones and diethyl 2-aryl-malonates has been developed using CuO-nanoparticles as the catalyst [7]. Satish et al. have developed a nanocrystalline CuO catalyzed coupling of 2-iodoaniline, carbon disulfide, and nitrogen nucleophile under ligand-free conditions in good yields [9].
Shaterian et al. have synthesized an efficient one-pot quantitative procedure for the synthesis of benzo[a]pyrano[2,3-c]phenazine derivatives in the presence of nano CuO as the catalyst [11].
Nitroolefins are useful organic compounds that are used as intermediates both in synthetic chemistry and in industrial chemistry. They are commonly employed in various carbon–carbon bond formations, including Diels–Alder cycloaddition, Michael reaction, and Morita–Baylis–Hillman reaction. Furthermore, the nitro group may readily be transformed into other useful functionalities. They are widely used in agrochemicals, materials, medicines, and other industries. As a result, researchers are interested in developing efficient procedures for the synthesis of nitroolefin. Recently, several useful approaches have been established, mainly including nitration of vinylic C–H bonds, nitro-decarboxylation of α, β-unsaturated aromatic carboxylic acids, and the traditional approach, particularly the Henry reaction [27,28].
The Henry reaction is a classical name reaction known for more than a century. It depends on base-mediated aldehyde or ketone condensation using nitroalkanes. The extensive studies performed by organic chemists and the commercial availability of the relatively low-cost starting materials make it a versatile and widely used reaction. Indeed, the Henry reaction is one of the most simple and economical sets of reaction conditions reported in the literature.
As a result of the aforementioned findings, we introduce an environmentally friendly nitroolefin synthesis method using chitosan-CuO nanocomposites as an efficient heterogeneous base catalyst.

2. Materials and Methods

2.1. Apparatus and Instrumentations

Melting points were recorded on Gallen Kamp apparatus and are reported uncorrected. The Fourier Transform Infrared (FTIR) spectra were recorded in KBr pellets on a JASCO FT-IR-6300 system at a resolution of 4 cm−1, ranging from 400 to 4000 cm−1. Field Emission Scanning Electron Microscopy (FESEM) was carried out using a model Leo (Zeiss) Remotely Operational Variable Pressure Field Emission SEM. The X-ray Diffraction (XRD) measurements were performed at room temperature on a Siemens diffractometer model D500 (Germany) operating in the reflection mode. Cu-Kα radiation (35 kV, 30 mA) was used for the analysis. Microwave experiments were carried out using a CEM Discover Labmate microwave apparatus (Discover, SP, NC, USA, 300 W). Sodium hydroxide was purchased from Sigma-Aldrich (St. Louis, MO, USA). Triple distilled water was used in all solution preparations. Chitosan of medium molecular weight and with 85% deacetylation was purchased from Sigma-Aldrich. Copper oxide (nanopowder, <50 nm particle size (TEM), 544868) was purchased from Sigma-Aldrich.

2.2. Preparation of Chitosan-Based CuO Nanocomposite Films

Chitosan solution 2 w/v% was made by stirring chitosan (CS) in a 1 v/v% acetic acid solution for 12 h at room temperature on a magnetic stirrer. The pH of the resultant CS solution was adjusted to the range of 6–7 by adding a calculated amount of 1 M NaOH solution under continuous stirring after full dissolution. Then, under continuous stirring, a suspension of 0.5 g CuO nanopowder in a small amount of double-distilled water was added portion by portion to the CS solution. The mixture was then subjected to microwave irradiation at a power of 300, 400, and 500 watts for 3, 4, and 5 min, then the solution was cast into a 100 mm Petri dish and dried overnight at 70 °C in order to remove any traces of acetic acid. The CS-CuO nanocomposite film was detached after complete drying, rinsed with distilled water, and then dried for 1 h at 70 °C. Finally, characterizations were performed on the formed films.

2.3. Synthesis of Nitroolefin

Method A: In a conical flask, 10 mmol of aldehyde and 5 mmol of ammonium acetate were dissolved in 5 mL of nitromethane. The mixture was refluxed with stirring for 4–5 h and the consumption of the aldehyde was checked by TLC. The reaction mixture was evaporated under vacuum and the residue was dissolved in methylene chloride and washed with water. The organic layer was dried and finally purification over a short silica column with EtOAc: hexane as eluent afforded the nitroolefin [29,30].
Method B: The same procedure as Method A, using a chitosan-CuO nanocomposite (15 wt.%) (0.1 g) instead of ammonium acetate and ethanol-water (70% solution), was used as solvent in this reaction. After 3–4 h of refluxing the reaction mixture, the hot solution was filtered to remove chitosan-CuO nanocomposite and the filtrate was evaporated. The residue was treated with DCM and washed with water. The organic layer was dried and finally purified over a short silica column with EtOAc: hexane as eluent to give authentic samples of nitroolefin (m.p, mixed m.p, IR, and TLC). The catalyst film was removed from the reaction mixture and thoroughly washed with ethanol before being dried and put to use in additional reactions.

3. Results and Discussion

3.1. Preparation of Chitosan-CuO Nanocomposite

In fact, in our previous work, we used the simple solution casting method to prepare the chitosan-copper oxide nanocomposite, but in this paper, we attempted to modify this methodology by using microwave irradiation to achieve a more homogeneous distribution of the copper oxide nanoparticles within the bulk of the polymer matrix. Thus, in this article, the (CS-CuO) nanocomposite was prepared using a modified microwave assisted simple solution casting process [15,16], as indicated in Scheme 1.

3.1.1. FTIR Characterization

Infrared spectroscopy provides more structural information by the presence or absence of the main characteristic absorption bands that help to detect the main functional groups in the compounds. Herein, the FTIR spectra were measured for chitosan (A), CuO nanoparticles (B), and chitosan-CuO nanocomposites (C), and they are shown in Figure 1. The spectrum of pure CS [15,16,17,18,31] (A) revealed a broad band at 3426 cm−1 due to the intermolecular H-bonding of the O-H stretching and NH2 stretching bands, that are overlapped together in the same region. Furthermore, the amide characteristic bands were clearly visible at υ = 1655 and 1604 cm−1 in the spectrum, whereas the CH bands throughout the CS chain were clearly visible at υ = 2915, 2876, and 1370 cm−1. In Figure 1B, for CuO nanoparticles, two peaks, one at 3418 cm−1 of the O-H stretching and the other at 1622 cm−1 of the O-H bending, appeared due to the presence of a trace amount of water in the copper oxide nanoparticles. There are two peaks at 1381 cm-1 and 1060 cm−1 linked to the asymmetric C-O in the copper oxide structure. Additionally, the CuO spectra showed the presence of two characteristic bands, at υ = 606 and 456 cm−1 that correspond to the expected Cu-O stretching vibrations, in the range of 620–420 cm−1 as previously reported [32,33], which are attributed to the monoclinic phase of CuO nanoparticles. Figure 1C, for the hybrid chitosan-CuO nanocomposite, showed the noisy shape with the presence of the broad band at 3438 cm−1 and the two characteristic bands at 2115 and 2362 cm−1 that are strong evidence for the coordination of CuO molecules within the binding sites (NH2 and OH groups) along the chitosan backbone. Finally, the visible shift in the chitosan fingerprint region, especially at 632 and 502 cm−1, is strong evidence for structural changes caused by the incorporation of copper oxide nanoparticles.

3.1.2. X-ray Diffraction (XRD)

The structural features of native chitosan (A), pure CuO nanoparticles (B) chitosan-CuO nanocomposites (15 wt.%), and (C) were analyzed using the XRD technique as shown in Figure 2. The chitosan (A) displayed the fundamental typical peak at 2θ = 20° (17–23°), which was typical of the hydrated crystalline structure of chitosan reported earlier [15,16,17,18]. In Figure 2B, the same characteristic peaks of both chitosan and CuO molecules, 2θ of chitosan at 20° and for CuO nanoparticles at 35.6° and 38.8° [34], were interfered with a clear shift, which confirms the coordination between CuO molecules and the chitosan chain. Furthermore, the Debye–Scherrer formula was used to calculate the average grain size from the XRD patterns [35]:
D n m = 0.9 × λ   β · cos θ
where, D (nm): represents the crystalline size in nm, λ is the wavelength of Cu-Kα1 = 1.54060 A°, β can be calculated for the most intense peak for the CS-CuO nanocomposite pattern. Applying the equation, the average particle size was found to be 35.2 nm.

3.1.3. SEM and Morphological Changes

SEM images were used to investigate the morphological alterations of CS-CuO nanocomposites in comparison to unmodified chitosan. As shown in Figure 3, the nonporous, smooth membranous phase of chitosan (A) contained dome-shaped orifices, microfibrils, and crystallites, as shown in the SEM image [15,16,17,18]. On the other hand, the image of the CS-CuO nanocomposite (B) exhibited distinct modifications due to coordination with CuO molecules that are uniformly dispersed over the chitosan surface.

3.1.4. Energy-Dispersive X-ray Spectroscopy (EDS) and Estimation of Copper

Figure 4 shows an EDS graph of chitosan-CuO nanocomposites that was used to quantify the copper content within the chitosan. The hybrid material’s EDS revealed the appearance of the standard Cu signals, which indicated the amount of copper in the chitosan matrix. As shown in Figure 4, the Cu content in the prepared sample was about 13.62 (wt.%).

3.2. CS-CuO Nanocomposite Film as Basic Heterogeneous Catalyst in Synthesis of Nitroolefin

3.2.1. Optimization of the Reaction Conditions

Compound 1a was selected as a model compound for the optimization of the experimental conditions. CS-CuO was selected as the catalyst for the optimization reactions. The basic character of this catalyst and the catalytic activity of copper oxide nanoparticles promoted us to use it as a powerful catalyst in comparison to ammonium acetate (experimental Section 2.3) in the nitroaldol condensation reaction (Scheme 2)
The nature of the solvent and catalyst loading have both been optimized and the results obtained are reported in Table 1 and Figure 5.

Nature of Solvent

In the screening of the solvents (Table 1), different solvents were studied, including nonpolar, polar protic, and aprotic solvents. THF and toluene exhibited the lowest catalytic results. This could be explained by the fact that the charged intermediate formed during the reaction can’t be stabilized either in an aprotic polar or in a nonpolar solvent. Ethanol showed very poor yield to the desired nitroolefin. However, ethanol-water mixture as solvent showed the best result which suggests that water molecules are essential for enhancing the basicity of CuO nanoparticles. The other three tested polar aprotic solvents showed poor results. Thus, we selected an ethanol-water mixture as the best choice for this reaction.

Amount of Catalyst (Catalyst Loading)

We initially set the catalyst/benzaldehyde weight percent (wt.%) to be 5. The nitroolefin 1a was successfully isolated from the reaction mixture and its structure has been fully identified by NMR, mass spectroscopy, and elemental analysis.
As shown in Figure 5, the catalyst loading of 15 wt.%. was found to be the optimal quantity, as this percentage of catalyst provided the highest yield of product 1a (90%) in 3 h. After the reaction was complete, the used catalyst film was detached, purified with hot ethanol, and recovered to be usefully utilized four more times with minor loss in catalytic activity (Figure 6).
By summing up the data from the optimization, the best result was obtained by employing CS-CuO (15 wt.%), at 120 °C for 3 h in (1:5) EtOH-H2O as the solvent.

3.2.2. Scope of the Reaction

The scope of the reaction has been explored (Table 2). Different substituted benzaldehydes and nitroalkanes were subjected to our optimum conditions. It is interesting to note that the reaction was found to be tolerant to both electron donating (NMe2, OCH3, and CH3) and electron attracting groups (F, Cl, Br, and NO2) that are present in the phenyl ring of benzaldehyde.

4. Conclusions

The present study used a simple solution casting approach to efficiently prepare chitosan-copper oxide (CS-CuO) nanocomposite. The produced nanocomposite film was thoroughly examined using FTIR, XRD, FESEM, and EDS measurements. The existence of copper oxide nanoparticles within the chitosan matrix was confirmed by all the tools’ data. The existence of CuO molecules was demonstrated by FTIR spectra that exhibited a noticeable alteration in the fingerprint region. In addition, both chitosan and CuO characteristic peaks were visible in the XRD pattern. Furthermore, the SEM picture of the CS-CuO nanocomposite revealed a clear uniform surface modification due to CuO molecule coordination. This CS-CuO hybrid nanocomposite was effectively employed as an environmentally friendly heterogeneous basic catalyst in the manufacture of a high-impact nitroolefin. The created catalyst is a promising catalyst not only because of its non-toxic nature but also because of its economic impact, which means it might be employed in the industrial manufacturing of the compounds mentioned. Finally, we conclude that the nanocomposite base catalyst can be used to efficiently synthesize a variety of heterocycles that have previously been generated via non-green methods.

Author Contributions

Data curation, M.A.E.-A., K.D.K. and A.H.B.; formal analysis, M.A.E.-A., K.D.K. and A.H.B.; investigation, M.A.E.-A., K.D.K. and A.H.B.; methodology, M.A.E.-A., K.D.K., and A.H.B.; resources, M.A.E.-A., K.D.K. and A.H.B.; software, M.A.E.-A. and K.D.K.; supervision, M.A.E.-A., K.D.K. and A.H.B.; validation, K.D.K.; visualization, M.A.E.-A.; writing—original draft, M.A.E.-A. and K.D.K.; writing—review and editing, M.A.E.-A., K.D.K. and A.H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

There are no conflict to declare.

References

  1. Bell, A.T. The Impact of Nanoscience on Heterogeneous Catalysis. Science 2003, 299, 1688–1691. [Google Scholar] [PubMed]
  2. Narayanan, R. Synthesis of green nanocatalysts and industrially important green reactions. Green Chem. Lett. Rev. 2012, 5, 707–725. [Google Scholar] [CrossRef]
  3. Somwanshi, S.B.; Somvanshi, S.B.; Kharat, P.B. Nanocatalyst: A Brief Review on Synthesis to Applications. J. Phys. Conf. Ser. 2020, 1644, 012046. [Google Scholar] [CrossRef]
  4. Bhanage, B.M.; Arai, M. Catalyst product separation techniques in heck reaction. Catal. Rev. 2001, 43, 315–344. [Google Scholar] [CrossRef]
  5. Hemalatha, K.; Madhumitha, G.; Kajbafvala, A.; Anupama, N.; Sompalle, R.; Mohana, R.S. Function of Nanocatalyst in Chemistry of Organic Compounds Revolution: An Overview. J. Nanomater. 2013, 2013, 4. [Google Scholar]
  6. Jammi, S.; Sakthivel, S.; Rout, L.; Mukherjee, T.; Mandal, S.; Mitra, R.; Saha, P.; Punniyamurthy, T. CuO Nanoparticles Catalyzed C-N, C-O, and C-S Cross-Coupling Reactions: Scope and Mechanism. J. Org. Chem. 2009, 74, 1974–1976. [Google Scholar]
  7. Kidwai, M.; Bhardwaj, S.; Poddar, R. C-Arylation reactions catalyzed by CuO-nanoparticles under ligand free conditions. Beil. J. Org. Chem. 2010, 6, 35–41. [Google Scholar]
  8. Ahmady, A.Z.; Keshavarz, M.; Kardani, M.; Mohtasham, N. CuO Nanoparticles as an Efficient Catalyst for the Synthesis of Flavanones. Orient. J. Chem. 2015, 31, 1841–1846. [Google Scholar]
  9. Satish, G.; Reddy, K.H.V.; Ramesh, K.; Karnakar, K.; Nageswar, Y.V.D. Synthesis of 2-N-substituted benzothiazoles via domino condensation-hetero cyclization process, mediated by copper oxide nanoparticles under ligand-free conditions. Tetrahedron Lett. 2012, 53, 2518–2521. [Google Scholar]
  10. Gaddam, S.; Kasireddy, H.R.; Konkala, K.; Katla, R.; Durga, N.Y.V. Synthesis of N-substituted-2-aminobenzothiazoles using nano copper oxide as a recyclable catalyst under ligand-free conditions, in reusable PEG-400 medium. Chin. Chem. Lett. 2014, 25, 732–736. [Google Scholar]
  11. Shaterian, H.R.; Moradi, F.; Mohammadnia, M. Nano copper(II) oxide catalyzed four-component synthesis of functionalized benzo[a]pyrano [2,3-c]phenazine derivatives. Comptes Rendus Chim. 2012, 15, 1055–1059. [Google Scholar] [CrossRef]
  12. Rawat, M.; Rawat, D.S. Copper oxide nanoparticle catalysed synthesis of imidazo[1,2-a] pyrimidine derivatives, their optical properties and selective fluorescent sensor towards zinc ion. Tetrahedron Lett. 2018, 59, 2341–2346. [Google Scholar] [CrossRef]
  13. Zhou, K.; Wang, R.; Xu, B.; Li, Y. Synthesis, characterization and catalytic properties of CuO nanocrystals with various shapes. Nanotechnology 2006, 17, 3939–3943. [Google Scholar] [CrossRef]
  14. Varzi, Z.; Esmaeili, M.S.; Taheri-Ledari, R.; Maleki, A. Facile synthesis of imidazoles by an efficient and eco-friendly heterogeneous catalytic system constructed of FeO and CuO nanoparticles, and guarana as a natural basis. Inorg. Chem. Commun. 2021, 125, 108465–108470. [Google Scholar] [CrossRef]
  15. Khalil, K.D.; Riyadh, S.M.; Gomha, S.M.; Ali, I. Synthesis, characterization and application of copper oxide chitosan nanocomposite for green regioselective synthesis of [1,2,3]triazoles. Int. J. Biol. Macromol. 2019, 130, 928–937. [Google Scholar] [CrossRef]
  16. Aljuhani, A.; Riyadh, S.M.; Khalil, K.D. Chitosan/CuO nanocomposite films mediated regioselective synthesis of 1,3,4-trisubstituted pyrazoles under microwave irradiation. J. Saudi Chem. Soc. 2021, 25, 101276. [Google Scholar] [CrossRef]
  17. Khalil, K.D.; Ibrahim, E.I.; Al-Sagheer, F.A. A novel, efficient, and recyclable biocatalyst for Michael addition reactions and its iron(III) complex as promoter for alkyl oxidation reactions. Catal. Sci. Technol. 2016, 6, 1410–1416. [Google Scholar]
  18. Madkour, M.; Khalil, K.D.; Al-Sagheer, F.A. Heterogeneous Hybrid Nanocomposite Based on Chitosan/Magnesia Hybrid Films: Ecofriendly and Recyclable Solid Catalysts for Organic Reactions. Polymers 2021, 13, 3583. [Google Scholar] [CrossRef]
  19. Riyadh, S.M.; Khalil, K.D.; Bashal, A.H. Structural Properties and Catalytic Activity of Binary Poly (vinyl alcohol)/Al2O3 Nanocomposite Film for Synthesis of Thiazoles. Catalysts 2020, 10, 100. [Google Scholar] [CrossRef]
  20. Mizwari, Z.M.; Oladipo, A.A.; Yilmaz, E. Chitosan/metal oxide nanocomposites: Synthesis, characterization, and antibacterial activity. Int. J. Poly. Mater. Poly. Biomater. 2020, 69, 1–9. [Google Scholar] [CrossRef]
  21. Sunil, D. Recennt Advances on Chitosan-Metal Oxide Nanoparticles and their Biological Application. Mater. Sci. Forum 2013, 754, 99–108. [Google Scholar] [CrossRef]
  22. Fu, C.-C.; Hung, T.-C.; Su, C.-H.; Suryani, D.; Wu, W.-T.; Dai, W.-C.; Yeh, Y.-T. Immobilization of calcium oxide onto chitosan beads as a heterogeneous catalyst for biodiesel production. Polym. Int. 2011, 60, 957–962. [Google Scholar] [CrossRef]
  23. Guibal, E. Heterogeneous catalysis on chitosan-based materials: A review. Prog. Polym. Sci. 2005, 30, 71–109. [Google Scholar] [CrossRef]
  24. Mostafa, M.A.; Ismail, M.M.; Morsy, J.M.; Hassanin, H.M.; Abdelrazek, M.M. Synthesis, characterization, anticancer, and antioxidant activities of chitosan Schiff bases bearing quinolinone or pyranoquinolinone and their silver nanoparticles derivatives. Polym. Bull. 2022, 1–25. [Google Scholar] [CrossRef]
  25. Julianto, T.S.; Mumpuni, R.A. Chitosan and N-Alkyl chitosan as a heterogeneous based catalyst in the transesterification reaction of used cooking oil. IOP Conf. Ser. Mater. Sci. Eng. 2016, 107, 012004. [Google Scholar] [CrossRef]
  26. Singh, J.; Kaur, G.; Rawat, M. A Brief Review on Synthesis and Characterization of Copper Oxide Nanoparticles and its Applications. J. Bioelectron. Nanotechnol. 2016, 1, 9. [Google Scholar]
  27. El-Atawy, M.A.; Ferretti, F.; Ragaini, F. A Synthetic Methodology for Pyrroles from Nitrodienes. Eur. J. Org. Chem. 2018, 34, 4818–4825. [Google Scholar] [CrossRef]
  28. El-Atawy, M.A.; Ferretti, F.; Ragaini, F. Palladium-Catalyzed Intramolecular Cyclization of Nitroalkenes: Synthesis of Thienopyrroles. Eur. J. Org. Chem. 2017, 14, 1902–1910. [Google Scholar] [CrossRef]
  29. Munoz, L.; Rodriguez, A.M.; Rosell, G.; Bosch, M.P.; Guerrero, A. Enzymatic enantiomeric resolution of phenylethylamines structurally related to amphetamine. Org. Biomol. Chem. 2011, 9, 8171–8177. [Google Scholar] [CrossRef]
  30. Goodman, M.M.; Kabalka, G.W.; Marks, R.C.; Knapp, F.F.; Lee, J.; Liang, Y. Synthesis and evaluation of radioiodinated 2-(2(RS)-aminopropyl)-5-iodothiophenes as brain imaging agents. J. Med. Chem. 1992, 35, 280–285. [Google Scholar] [CrossRef]
  31. Cardenas, G.; Miranda, S.P. FTIR and TGA studies of chitosan composite films. J. Chilean Chem. Soc. 2004, 49, 291–295. [Google Scholar] [CrossRef]
  32. Vaseem, M.; Umar, A.; Kim, S.H.; Hahn, Y.-B. Low-Temperature Synthesis of Flower-Shaped CuO Nanostructures by Solution Process:  Formation Mechanism and Structural Properties. J. Phys. Chem. 2008, 112, 5729–5735. [Google Scholar] [CrossRef]
  33. Sudha, V.; Murugadoss, G.; Thangamuthu, R. Structural and morphological tuning of Cu-based metal oxide nanoparticles by a facile chemical method and highly electrochemical sensing of sulphite. Sci. Rep. 2021, 11, 3413. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, Y.; Liao, L.; Li, J.; Pan, C. From Copper Nanocrystalline to CuO Nanoneedle Array:  Synthesis, Growth Mechanism, and Properties. J. Phys. Chem. 2007, 111, 5050–5056. [Google Scholar] [CrossRef]
  35. Klug, H.P.; Alexander, L.E. X-ray Diffraction Procedures; John Wiley and Sons Inc.: New York, NY, USA, 1954; p. 633. [Google Scholar] [CrossRef]
Scheme 1. Illustrative view of chitosan-CuO nanocomposite.
Scheme 1. Illustrative view of chitosan-CuO nanocomposite.
Catalysts 12 00964 sch001
Figure 1. FTIR of (A) chitosan, (B) CuO nanoparticles, (C) Chitosan-CuO nanocomposites (15 wt.%).
Figure 1. FTIR of (A) chitosan, (B) CuO nanoparticles, (C) Chitosan-CuO nanocomposites (15 wt.%).
Catalysts 12 00964 g001
Figure 2. XRD of chitosan (A), and chitosan-CuO nanocomposite, (15 wt.%) (B).
Figure 2. XRD of chitosan (A), and chitosan-CuO nanocomposite, (15 wt.%) (B).
Catalysts 12 00964 g002
Figure 3. SEM of chitosan (A), and chitosan-CuO nanocomposites, (15 wt.%) (B).
Figure 3. SEM of chitosan (A), and chitosan-CuO nanocomposites, (15 wt.%) (B).
Catalysts 12 00964 g003
Figure 4. EDS of chitosan-CuO nanocomposites (15 wt.%).
Figure 4. EDS of chitosan-CuO nanocomposites (15 wt.%).
Catalysts 12 00964 g004
Scheme 2. Nitroaldol condensation reaction.
Scheme 2. Nitroaldol condensation reaction.
Catalysts 12 00964 sch002
Figure 5. Henry reaction of benzaldehyde and nitromethane catalyzed by CS-CuO: Optimization of catalyst loading (wt.%). Experimental conditions: catalyst CS-CuO (wt.% = 5–20), solvent = EtOH-H2O (1:5), T = 120 °C, and reaction time = 3 h.
Figure 5. Henry reaction of benzaldehyde and nitromethane catalyzed by CS-CuO: Optimization of catalyst loading (wt.%). Experimental conditions: catalyst CS-CuO (wt.% = 5–20), solvent = EtOH-H2O (1:5), T = 120 °C, and reaction time = 3 h.
Catalysts 12 00964 g005
Figure 6. Recyclability of CS-CuO (15 wt.%) nanocatalyst in synthesis of product 1a.
Figure 6. Recyclability of CS-CuO (15 wt.%) nanocatalyst in synthesis of product 1a.
Catalysts 12 00964 g006
Table 1. Henry reaction of benzaldehyde and nitromethane catalyzed by CS-CuO: effect of the solvent a.
Table 1. Henry reaction of benzaldehyde and nitromethane catalyzed by CS-CuO: effect of the solvent a.
EntryCatalyst (mol%)SolventTime (h)Temperature (°C)Yield (%)
110THF31208
210Toluene312013
310DMF312020
410CH3CN312025
510EtOH312014
610EtOH-H2O (1:1)312058
710EtOH-H2O (1:3)312065
8 a10EtOH-H2O (1:5)312080
a Experimental conditions: catalyst CS-CuO (wt.% = 10), T = 120 °C, and reaction time = 3 h.
Table 2. Scope of the reaction.
Table 2. Scope of the reaction.
Catalysts 12 00964 i001
EntryR1R2ProductYield (%)
14-ClCH3 Catalysts 12 00964 i00265
24-FCH3 Catalysts 12 00964 i00372
34-NO2CH3-CH2 Catalysts 12 00964 i00468
43-ClCH3 Catalysts 12 00964 i00574
53-BrCH3 Catalysts 12 00964 i00673
64-OCH3CH3 Catalysts 12 00964 i00760
74-CH3CH3 Catalysts 12 00964 i00866
84-NMe2CH3 Catalysts 12 00964 i00949
Experimental conditions: catalyst CS-CuO (15 wt.%), solvent = EtOH-H2O, T = 120 °C, and reaction time = 3 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

El-Atawy, M.A.; Khalil, K.D.; Bashal, A.H. Chitosan Capped Copper Oxide Nanocomposite: Efficient, Recyclable, Heterogeneous Base Catalyst for Synthesis of Nitroolefins. Catalysts 2022, 12, 964. https://doi.org/10.3390/catal12090964

AMA Style

El-Atawy MA, Khalil KD, Bashal AH. Chitosan Capped Copper Oxide Nanocomposite: Efficient, Recyclable, Heterogeneous Base Catalyst for Synthesis of Nitroolefins. Catalysts. 2022; 12(9):964. https://doi.org/10.3390/catal12090964

Chicago/Turabian Style

El-Atawy, Mohamed A., Khaled D. Khalil, and Ali H. Bashal. 2022. "Chitosan Capped Copper Oxide Nanocomposite: Efficient, Recyclable, Heterogeneous Base Catalyst for Synthesis of Nitroolefins" Catalysts 12, no. 9: 964. https://doi.org/10.3390/catal12090964

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop