Next Article in Journal
Ballistic Properties and Izod Impact Resistance of Novel Epoxy Composites Reinforced with Caranan Fiber (Mauritiella armata)
Previous Article in Journal
Advanced Characterization of Ceramic State Polymer Electrolyte at Radio Frequencies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient, Recyclable, and Heterogeneous Base Nanocatalyst for Thiazoles with a Chitosan-Capped Calcium Oxide Nanocomposite

1
Department of Chemistry, Faculty of Science, Cairo University, Giza 12613, Egypt
2
Department of Chemistry, Faculty of Science, Taibah University, Al-Madinah Almunawarah, Yanbu 46423, Saudi Arabia
3
Institute of Organic Chemistry (IOC), Karlsruhe Institute of Technology (KIT), Fritz-Haber-Weg 6, 76133 Karlsruhe, Germany
4
Institute of Biological and Chemical Systems-Functional Molecular Systems (IBCS-FMS), Director Hermann-von-Helmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, Germany
5
Department of Chemistry, College of Science, Jouf University, Sakaka 72341, Saudi Arabia
6
Department of Chemistry, Faculty of Science, Islamic University of Madinah, Madinah 42351, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Polymers 2022, 14(16), 3347; https://doi.org/10.3390/polym14163347
Submission received: 21 June 2022 / Revised: 31 July 2022 / Accepted: 9 August 2022 / Published: 17 August 2022
(This article belongs to the Section Biomacromolecules, Biobased and Biodegradable Polymers)

Abstract

:
Calcium oxide (CaO) nanoparticles have recently gained much interest in recent research due to their remarkable catalytic activity in various chemical transformations. In this article, a chitosan calcium oxide nanocomposite was created by the solution casting method under microwave irradiation. The microwave power and heating time were adjusted to 400 watts for 3 min. As it suppresses particle aggregation, the chitosan (CS) biopolymer acted as a metal oxide stabilizer. In this study, we aimed to synthesize, characterize, and investigate the catalytic potency of chitosan–calcium oxide hybrid nanocomposites in several organic transformations. The produced CS–CaO nanocomposite was analyzed by applying different analytical techniques, including Fourier-transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), and field-emission scanning electron microscopy (FESEM). In addition, the calcium content of the nanocomposite film was measured using energy-dispersive X-ray spectroscopy (EDS). Fortunately, the CS–CaO nanocomposite (15 wt%) was demonstrated to be a good heterogeneous base promoter for high-yield thiazole production. Various reaction factors were studied to maximize the conditions of the catalytic technique. High reaction yields, fast reaction times, and mild reaction conditions are all advantages of the used protocol, as is the reusability of the catalyst; it was reused multiple times without a significant loss of potency.

Graphical Abstract

1. Introduction

Organic synthesis is one of the most important areas where catalysts are used, especially on a commercial scale, because they allow for a more cost-effective preparation of various organic compounds than would otherwise be possible. In their natural state, organo-catalysts are homogenous, but when modified as nanocatalysts, they exhibit heterogeneous catalytic behavior [1]. Although popular traditional catalysts such as pyridine and piperidine as well as triethyl amine have been utilized in organic synthesis transformations for many years, nanoparticles are currently being used for greener and novel organic reactions [2,3,4,5,6].
Nanocatalysts have recently piqued the interest of many academics as they have emerged as a viable, long-term, and low-cost alternative to the old commercial catalysts widely utilized in organic synthesis [4]. They display increased catalytic activity and regioselectivity in several chemical transformations due to their particle size (1–100 nm) and, therefore, their highly exposed surface area with a larger number of reactive sites, which causes the reaction to run to completion with a higher efficiency. Another advantage is the insolubility of nanocatalysts in reaction mixtures, which makes the separation, recovery, and reuse of the nanocatalysts easier. For the above reasons, many researchers have attempted to create new hybrid materials containing metal oxide nanoparticles due to their promising catalytic capabilities in various organic reactions [7,8].
Although calcium oxide (CaO) is a plentiful, environmentally friendly, and non-corrosive substance, its utility remains limited in organic reactions. Due to its low solubility, high alkalinity, and—most importantly—its ability to be recycled following the reaction, it is simple to handle. As a result, it has been identified as one of the heterogeneous catalysts most frequently utilized in the manufacture of biodiesel [9]. In addition, calcium oxide nanoparticles have been efficiently used as base catalysts in many organic transformations such as the green synthesis of polyfunctionalized pyridine derivatives [10], dihydropyrazole derivatives [11], chalcones [12], and 4H-pyran derivatives [13] as well as in the hydrogenation of ethylene [14].
Chitosan and its derivatives can be used as efficient, recyclable, and environmentally friendly catalyst supports in various applications [15,16,17,18,19,20]. Khaled et al. [18] investigated the catalytic activity of chitosan-grafted poly(2-cyano-1-(pyridine-3yl)allyl acrylate)-Cs-g-PCPA with 64% grafting as a novel, efficient, eco-friendly, and recyclable biocatalyst for Michael addition reactions and examined the use of the iron complex of the grafted chitosan as a promoter for alkyl oxidation reactions.
Due to its strong adsorption capacity through the NH2 and OH binding sites, chitosan, a natural polysaccharide, has been widely exploited as an efficient stabilizer for immobilizing metal oxide nanoparticles in recent decades [21,22,23,24]. A chitosan–CuO nanocomposite was prepared via the simple solution cast method and successfully used as a heterogenous basic catalyst in the synthesis of 1,3,4-pyrazoles [25] and 1,2,3-triazoles [26]. Moreover, chitosan–MgO nanocomposites were efficiently used as a heterogeneous basic catalyst in Michael additions [27] and also for the synthesis of thiazoles and 1,3,4-thiadiazoles in a comparative study with a traditional catalyst [28].
Abdel-Naby and coworkers synthesized a chitosan–Al2O3 nanocomposite under microwave conditions; the nanocomposite was then used as a green heterogeneous catalyst to synthesize novel imidazopyrazolyl thione derivatives [29]. Recently, Khalil et al. [30] investigated the catalytic activity of a chitosan–SrO nanocomposite as a powerful base promoter for the synthesis of 2-hydrazono-1,3,4-thiadiazole derivatives.
Shorter reaction times, a simple experimental approach, excellent yields, greater selectivity, and clean processes are all advantages of ultrasound-assisted reactions [31,32,33]. One of the many advantages of ultrasonic irradiation is its ability to play a critical role in chemistry, especially in cases where conventional methods demand extreme conditions or extended reaction times [34,35,36].
The 1,3-thiazole core structure, on the other hand, has been widely investigated in identifying new lead compounds for drug discovery. Thiazole-containing therapies such as tiazofurin (an inhibitor of IMP dehydrogenase) [37], dasatinib (a Bcr-Abl tyrosine kinase inhibitor) [38], dabrafenib (an inhibitor of enzyme B-RAF) [39], ixabepilone (microtubule stabilization) [40], and epothilone (microtubule function inhibition) (Figure 1) are among the best-selling drugs or have served as lead structures [41]. Several drug discovery teams have investigated the possible applicability of a thiazole scaffold in the design of anti-cancer drugs [42,43,44,45].
Considering the results above, and in continuation with our studies on nanocatalysis [7,8,25,26,27,28,29,30], herein, we introduce an eco-friendly protocol for synthesizing 1,3-thiazole derivatives in the presence of chitosan–CaO nanocomposites as an efficient and recyclable base catalyst.

2. Materials and Methods

2.1. Apparatus and Instrumentation

Details were inserted in the Supplementary Materials.

2.2. Preparation of the Chitosan–Based CaO Nanocomposite Films

A chitosan solution (2%, w/v) was prepared by swirling chitosan (CS) in a 1% (v/v) acetic acid solution on a magnetic stirrer for 12 h at room temperature. After completely dissolving, the pH of the resulting CS solution was adjusted to the range of 6–7 by adding a calculated amount of 1 M NaOH solution under continuous stirring. Portion-by-portion, a suspension of the estimated amount of the CaO nanopowder in a tiny amount of double-distilled water was then added to the CS solution under continuous stirring. After being cast into a 100 mm Petri dish and dried overnight at 70 °C to remove any traces of acetic acid, the mixture was microwaved for 3 min at 400 watts. The CS–CaO nanocomposite film was then removed. After completely drying, the CS–CaO nanocomposite film was removed, rinsed with distilled water, and dried at 70 °C. The characterizations were performed using the formed films.

2.3. Synthesis of Thiazole Derivatives 5a,b

In 30 mL of EtOH, a mixture of 2-(4-formyl-3-methoxyphenoxy)-N-phenylacetamide derivatives 3a,b (10 mmol), and thiosemicarbazide 4 (0.91 g, 10 mmol) was treated with catalytic quantities of concentrated hydrochloric acid. An ultrasonic generator was used to irradiate the reaction mixture in a water bath at 50 °C for 20 min. After cooling, the precipitate was filtered, washed with ethanol, and recrystallized from acetic acid to produce thiosemicarbazones 5a,b.
2-(4-((2-Carbamothioylhydrazineylidene)methyl)-3-methoxyphenoxy)-N-(4-chlorophenyl)acetamide (5a).
A yellowish-white solid of 74% yield; m.p. 214–216 °C; IR (KBr): v 3403, 3350, 3312, 3277 (2NH, NH2), 3042, 2926 (CH), 1682 (C=O), 1607 (C=N) cm−1; 1H NMR (DMSO-d6): δ 3.89 (s, 3H, OCH3), 4.73 (s, 2H, –CH2O), 6.91–7.89 (m, 9H, Ar–H, NH2), 8.51 (s, 1H, CH=N), 10.35 (br s, 1H, NH), 11.49 (br s, 1H, NH) ppm; MS m/z (%): 392 (M+, 26). Elemental analysis calculated for C17H17ClN4O3S (392.86): C, 51.97; H, 4.36; N, 14.26. Found: C, 51.81; H, 4.27; N, 14.14%.
N-(4-Bromophenyl)-2-(4-((2-carbamothioylhydrazinylidene)methyl)-3-methoxyphenoxy)acetamide (5b)
A yellowish-white solid of 77% yield; m.p. 201–203 °C; IR (KBr): v 3411, 3369, 3327, 3295 (2NH, NH2), 3051, 2923 (CH), 1686 (C=O), 1613 (C=N) cm−1; 1H NMR (DMSO-d6): δ 3.87 (s, 3H, OCH3), 4.70 (s, 2H, –CH2O), 6.95–7.80 (m, 9H, Ar–H, NH2), 8.47 (s, 1H, CH=N), 10.29 (br s, 1H, NH), 11.53 (br s, 1H, NH) ppm; MS m/z (%): 437 (M+, 54). Elemental analysis calculated for C17H17BrN4O3S (437.31): C, 46.69; H, 3.92; N, 12.81. Found: C, 46.58; H, 3.82; N, 12.69%.

2.4. Synthesis of Thiazole Derivatives 8a–g

Method A: A few drops of triethylamine TEA were added to a mixture of equimolar volumes of suitable hydrazonoyl chlorides 6a–d (1 mmol) and thiosemicarbazones 5a,b (1 mmol) in 20 mL EtOH. An ultrasonic generator was used to irradiate the produced solution in a water bath at 50 °C for 20–60 min (irradiation was continued until all of the starting materials disappeared and the product was formed; this was monitored by TLC). After cooling, the red precipitate was filtered off, washed with EtOH, dried, and recrystallized from EtOH to yield the thiazoles 8a–g. Below are the physical constants for the products 8a–g.
Method B: A CS–CaO nanocomposite film (0.26 g, 15 wt%) was added to a mixture of equimolar proportions of suitable hydrazonoyl chlorides 6a–d (1 mmol) and thiosemicarbazones 5a,b (1 mmol) in 20 mL EtOH. An ultrasonic generator was used to irradiate the produced solution in a water bath at 50 °C for 20–40 min. To remove the CS–CaO, the heated solution was filtered and the surplus solvent was extracted under a reduced pressure. The residue was treated with methanol to obtain authentic samples of compounds 8a–g and the formed solid was filtered and recrystallized from EtOH (m.p., mixed m.p., IR, and TLC).
N-(4-Chlorophenyl)-2-(3-methoxy-4-((2-(4-methyl-5-((E)-phenyldiazenyl)thiazol-2-yl)hydrazono)methyl)phenoxy)acetamide (8a).
A red solid; m.p. 151–153 °C; IR (KBr): v 3406, 3327 (2NH), 3043, 2929 (CH), 1683 (C=O), 1611 (C=N) cm−1; 1H NMR (DMSO-d6): δ 2.55 (s, 3H, CH3), 3.85 (s, 3H, OCH3), 4.71 (s, 2H,–CH2O), 6.91–8.2 (m, 12H, Ar–H), 8.56 (s, 1H, CH=N), 10.37 (br s, 1H, NH), 11.34 (br s, 1H, NH) ppm; MS m/z (%): 535 (M+, 71). Elemental analysis calculated for C26H23ClN6O3S (535.02): C, 58.37; H, 4.33; N, 15.71. Found: C, 58.29; H, 4.20; N, 15.58%.
N-(4-Chlorophenyl)-2-(3-methoxy-4-((2-(5-((4-methoxyphenyl)diazenyl)-4-methylthiazol-2-yl)hydrazono)methyl)phenoxy)acetamide (8b).
A red solid; m.p. 137–139 °C; IR (KBr): v 3412, 3315 (2NH), 3051, 2924 (CH), 1677 (C=O), 1606 (C=N) cm−1; 1H NMR (DMSO-d6): δ 2.53 (s, 3H, CH3), 3.78 (s, 3H, OCH3), 3.83 (s, 3H, OCH3), 4.75 (s, 2H, –CH2O), 6.90–8.16 (m, 11H, Ar–H), 8.54 (s, 1H, CH=N), 10.58 (br s, 1H, NH), 11.33 (br s, 1H, NH) ppm; MS m/z (%): 565 (M+, 52). Elemental analysis calculated for C27H25ClN6O4S (565.04): C, 57.39; H, 4.46; N, 14.87. Found: C, 57.27; H, 4.41; N, 14.68%.
N-(4-Chlorophenyl)-2-(4-((2-(5-((4-chlorophenyl)diazenyl)-4-methylthiazol-2-yl)hydrazono)methyl)-3-methoxyphenoxy)acetamide (8c).
A red solid; m.p. 170–172 °C; IR (KBr): v 3403, 3294 (2NH), 3071, 2931 (CH), 1679 (C=O), 1609 (C=N) cm−1; 1H NMR (DMSO-d6): δ 2.54 (s, 3H, CH3), 3.83 (s, 3H, OCH3), 4.71 (s, 2H, –CH2O), 6.86–8.18 (m, 11H, Ar–H), 8.56 (s, 1H, CH=N), 10.27 (br s, 1H, NH), 11.34 (br s, 1H, NH) ppm; MS m/z (%): 569 (M+, 28). Elemental analysis calculated for C26H22Cl2N6O3S (569.46): C, 54.84; H, 3.89; N, 14.76. Found: C, 54.89; H, 3.77; N, 14.59%.
N-(4-Chlorophenyl)-2-(4-((2-(5-((2,4-dichlorophenyl)diazenyl)-4-methylthiazol-2-yl)hydrazono)methyl)-3-methoxyphenoxy)acetamide (8d).
A red solid; m.p. 177–179 °C; IR (KBr): v 3408, 3329 (2NH), 3020, 2933 (CH), 1684 (C=O), 1615 (C=N) cm−1; 1H NMR (DMSO-d6): δ 2.56 (s, 3H, CH3), 3.83 (s, 3H, OCH3), 4.76 (s, 2H, –CH2O), 6.88–8.03 (m, 10H, Ar–H), 8.21 (s, 1H, CH=N), 10.32 (br s, 1H, NH), 11.33 (br s, 1H, NH) ppm; MS m/z (%): 603 (M+, 57). Elemental analysis calculated for C26H21Cl3N6O3S (603.91): C, 51.71; H, 3.50; N, 13.92. Found: C, 51.64; H, 3.43; N, 13.81%.
N-(4-Bromophenyl)-2-(3-methoxy-4-((2-(4-methyl-5-(phenyldiazenyl)thiazol-2-yl)hydrazono)methyl)phenoxy)acetamide (8e).
A red solid; m.p. 145–147 °C; IR (KBr): v 3434, 3285 (2NH), 3045, 2937 (CH), 1675 (C=O), 1608 (C=N) cm−1; 1H NMR (DMSO-d6): δ 2.55 (s, 3H, CH3), 3.83 (s, 3H, OCH3), 4.80 (s, 2H, –CH2O), 6.90–8.02 (m, 12H, Ar–H), 8.56 (s, 1H, CH=N), 10.36 (br s, 1H, NH), 11.33 (br s, 1H, NH); MS m/z (%): 579 (M+, 52). Elemental analysis calculated for C26H23BrN6O3S (579.47): C, 53.89; H, 4.00; N, 14.50. Found: C, 53.73; H, 4.13; N, 14.37%.
N-(4-Bromophenyl)-2-(3-methoxy-4-((2-(5-((4-methoxyphenyl)diazenyl)-4-methylthiazol-2-yl)hydrazono)methyl)phenoxy)acetamide (8f).
A red solid; m.p. 155–157 °C; IR (KBr): v 3406, 3290 (2NH), 3041, 2934 (CH), 1679 (C=O), 1613 (C=N) cm−1; 1H NMR (DMSO-d6): δ 2.52 (s, 3H, CH3), 3.77 (s, 3H, OCH3), 3.82 (s, 3H, OCH3), 4.84 (s, 2H, –CH2O), 6.89–8.02 (m, 11H, Ar–H), 8.17 (s, 1H, CH=N), 10.64 (br s, 1H, NH), 11.34 (br s, 1H, NH) ppm; MS m/z (%): 609 (M+, 37). Elemental analysis calculated for C27H25BrN6O4S (609.49): C, 53.21; H, 4.13; N, 13.79. Found: C, 53.03; H, 4.07; N, 13.64%.
N-(4-Bromophenyl)-2-(4-((2-(5-((4-chlorophenyl)diazenyl)-4-methylthiazol-2-yl)hydrazono)methyl)-3-methoxyphenoxy)acetamide (8g).
A red solid; m.p. 181–183 °C; IR (KBr): v 3436, 3293 (2NH), 3049, 2928 (CH), 1682 (C=O), 1609 (C=N) cm−1; 1H NMR (DMSO-d6): δ 2.55 (s, 3H, CH3), 3.83 (s, 3H, OCH3), 4.78 (s, 2H, –CH2O), 6.88–8.03 (m, 11H, Ar–H), 8.19 (s, 1H, CH=N), 10.25 (br s, 1H, NH), 11.34 (br s, 1H, NH) ppm; MS m/z (%): 613 (M+, 55). Elemental analysis calculated for C26H22BrClN6O3S (613.91): C, 50.87; H, 3.61; N, 13.69. Found: C, 50.74; H, 3.51; 5.77; N, 13.50%.

3. Results and Discussion

3.1. Preparation of the Chitosan–CaO Nanocomposite

The CS–CaO nanocomposite was prepared using a simple solution casting process [25,26,28] under microwave irradiation, as indicated in Scheme 1.

3.1.1. FTIR Characterization

As shown in Figure 2, the FTIR spectra for chitosan, the CaO nanoparticles, and the chitosan–CaO nanocomposites (C) were measured. The spectrum of pure CS [25,26,27,28] revealed a broad band at 3426 cm−1 due to the intermolecular H-bonding of the O–H and NH2 stretching bands positioned in the same region. The amide characteristic bands could be seen at υ = 1655 and 1606 cm−1 whilst the CH bands along the CS chain appeared clearly at 2915, 2876, and 1372 cm−1 in the spectrum. As shown in Figure 2B, the CaO nanoparticles showed two strong bands at υ = 605 and 516 cm−1 corresponding with the expected Ca–O stretching vibrations in the range of 620–420 cm−1 as reported in [46], which has been attributed to the monoclinic phase of CaO nanoparticles. The FTIR spectra of the CaO sample showed additional bands that corresponded with the CaCO3 source; these peaks appeared in the range of 2000 and 1200 cm−1, which are the characteristic peaks of the C–O stretching and bending modes of CaCO3 [47]. The broad band at >3300 cm−1 could be attributed to the OH stretching vibration from water. Although the CaCO3 raw material was calcined to high temperatures greater than 800 °C for 10 h, those bands still appeared. Figure 2C depicts the noisy-like shape of the hybrid chitosan–CaO nanocomposite. The presence of the combination of chitosan and CaO characteristic bands at 3400 cm−1, two distinct bands at 2918 and 2875 cm−1, and the visible alteration in the chitosan fingerprint region (especially by the CaO bands) at 629 and 538 cm−1 was strong evidence of structural changes due to the incorporation of the calcium oxide nanoparticles.

3.1.2. X-ray Diffraction (XRD)

Figure 3 shows the structural characteristics of the native chitosan and the chitosan–CaO nanocomposite from the application of the XRD technique. Chitosan displayed two typical peaks, one strong at 2θ = 19–21° and the other weak at 2θ = 36°, which conformed with the values in the literature of the hydrated crystalline structure of chitosan [8,25,26,27,28]. The same distinctive peaks of both chitosan (2θ = 19–21°) and the CaO nanoparticles (2θ = 34° and 38.5°) [47] were disturbed with a clear shift in pattern, indicating the interaction of the CaO molecules with the chitosan chain. The Debye–Scherrer formula [48] was used to calculate the average grain size from the XRD patterns:
D nm = 0.9 λ β   cos θ
where D (nm) represents the crystalline size in nm and λ is the wavelength of Cu-kα1 = 1.54060 Å. For the CS–CaO nanocomposite pattern, β could be determined for the most intense peak. The average particle size was calculated to be 42.2 nm using this equation.

3.1.3. FESEM and Morphological Changes

A SEM instrument was used to investigate the morphological alterations of the CS–CaO nanocomposites compared with unmodified chitosan. As illustrated in Figure 4, the non-porous, smooth membranous phase of chitosan consisted of dome-shaped orifices, microfibrils, and crystallites [8,25,26,27,28]. However, due to the coordination of the chitosan binding sites with the CaO molecules, the picture of the CS–CaO nanocomposite revealed a drastic morphological change.

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

The presence of Ca within the chitosan matrix was confirmed by the EDS graph of the chitosan–CaO nanocomposites (Figure 5), which revealed characteristic Ca signals. The calcium content in the manufactured sample was found to be 15.54 wt%.

3.2. Synthesis of Thiazole Derivatives Using the CS–CaO Nanocomposite Film as Basic Heterogeneous Catalyst

In continuation of our previous work, which has designed and synthesized bioactive heterocyclic compounds under mild conditions [26,33,35,49,50,51,52,53,54], here, we wished to report a mild and efficient procedure for synthesizing novel thiazoles. Initially, the active key thiosemicarbazone derivatives 5a,b were prepared by the condensation of 2-(4-formyl-3-methoxyphenoxy)-N-phenylacetamide derivatives 3a,b [55] with thiosemicarbazide 4 in an acidic ethanol solution (Scheme 2). The structure of thiosemicarbazone derivatives 5a,b were assured, depending on the data extracted from the spectra (IR, 1H NMR, and MS).
New thiazole derivatives 8a–g were synthesized from the reaction of an equimolar quantity of thiosemicarbazone derivatives 3a,b with an equivalent of 2-oxo-N-arylpropanehydrazonoyl chloride 4a–g in EtOH in the presence of two different bases (CS or CS–CaO) (Scheme 1). The chemical structure of all newly produced thiazoles 8a–g was confirmed based on spectral and elemental investigations. Their structure was confirmed by 1H NMR of all isolated 8a–g derivatives, which displayed the predicted signals for the postulated structure. For instance, we detected the characteristic four singlet signals in the 1H NMR of compound 8a at 2.55 (CH3), 3.85 (OCH3), 2.55 (CH3), 4.71 (–CH2O), and 8.56 (CH=N) in addition to exchangeable protons at 10.37 and 11.34 due to two NHs and twelve multiplet aromatic protons at 6.91–8.20 ppm. The 13C NMR of the same derivative 8a revealed characteristic 22 non-equivalent carbon signals, as illustrated in the experimental section.
Thin-layer chromatography (TLC) was used to track the progress of all the reactions. The study for the best basic catalyst began at the outset (Table 1).
As shown in Table 1, the CS–CaO nanocomposite was a better basic catalyst than traditional TEA under ultrasonic irradiation. When triethylamine was replaced with the CS–CaO nanocomposite, the yields of the desired products of 8a–g increased; the reaction time decreased under the same reaction conditions.
We discovered the optimal experimental conditions and variables (such as the catalyst loading, temperature, solvent, and reaction duration) for the reaction of 5 + 6 in the presence of a catalytic quantity of the CS–CaO nanocomposite under USI to obtain thiazole derivative 8a.
The effect of the amount of catalyst on the synthesis of component 8a was investigated in the first step (Table 2, entries 1–3). The best results (91%) were obtained with a catalyst concentration of 20 mol% (Table 2, entry 3). Lower yields were achieved by using less catalyst (Table 2). The efficiency of the various solvents was then investigated using USI (Table 2, entries 3, 4, and 5). The production of product 8a proceeded with the best yield with the fastest reaction rate in EtOH, according to a screening of several solvents (Table 2, entry 3).
In addition, the reaction time was evaluated under USI (Table 2, entries 3, 6, and 7). The best time to form product 8a was 23 min (Table 2, entry 3).
Moreover, the influence of temperature on the reaction was investigated; the results are provided in Table 2 (entries 3, 8, 9, and 10). Table 2 shows that by increasing the reaction temperature from 25 to 35 to 50 °C whilst using USI increased the product yields from 73 to 87 to 91%, respectively. Finally, 40 °C was chosen as the ideal temperature (Table 2, entry 3).
Moreover, the recyclability of the CS–CaO nanocomposite as a basic catalyst was also examined. After each run, the catalyst film was rinsed with distilled water and dried at 60 °C for 30 min before reuse. Under ideal conditions, the catalyst was reused three times without a significant loss of catalytic performance (15 wt% and 23 min; Table 3).
In other words, the catalyst retained almost 93.4% of its performance after three runs, which indicated less than a 1% reduction after each run. In the fourth run, it lost about 50% of its performance.
The optimum reaction conditions for the synthesis of product 8a, as indicated in Table 2, were a reaction of 5a + 6a in EtOH under USI in the presence of the CS–CaO nanocomposite (15 wt%) at 40 °C for 23 min. Thus, the irradiation of 5a,b + 6b–d under the optimum conditions led to the formation of thiazole derivatives 8b–g (Scheme 1).

4. Conclusions

This study used the microwave-assisted solution casting method to efficiently prepare a chitosan–CaO nanocomposite. The produced nanocomposite film was thoroughly examined using FTIR, XRD, FESEM, and EDS measurements. All the data confirmed the presence of calcium oxide nanoparticles within the chitosan matrix. The FTIR spectra showed an obvious change, especially in the fingerprint region attributed to Ca–O bending vibrations. A combination of chitosan and CaO characteristic peaks could be seen in the XRD pattern. The SEM image of the nanocomposite also revealed a clear uniform surface alteration of the chitosan upon coordination with the CaO molecules. The CS–CaO hybrid nanocomposite film was efficiently used as an eco-friendly heterogeneous basic catalyst in synthesizing thiazole derivatives, which has a significant industrial impact. Thus, the invented catalyst is promising due to its non-toxic nature and economic impact and it may be used in the industrial production of the reported compounds. We concluded that the nanocomposite basic catalyst could be used to efficiently synthesize a variety of heterocycles that have previously been manufactured via non-green methods.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym14163347/s1, Figure S1: 1H-NMR spectrum of compound 8a, Figure S2: 1H-NMR spectrum of compound 8b, Figure S3: 1H-NMR spectrum of compound 8c, Figure S4: 1H-NMR spectrum of compound 8d, Figure S5: 1H-NMR spectrum of compound 8e, Figure S6: 1H-NMR spectrum of compound 8f, Figure S7: 1H-NMR spectrum of compound 8g.

Author Contributions

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

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors acknowledge support from the KIT-Publication Fund of the Karlsruhe Institute of Technology. Stefan Bräse is grateful for support from the DFGfunded cluster program “3D Matter Made To Order” under Germany’s Excellence Strategy -2082/1- 390761711. The authors acknowledge grants from the Science and Technology Commission of Shanghai Municipality (19440741300).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Atodiresei, I.; Vila, C.; Rueping, M. Asymmetric Organocatalysis in Continuous Flow: Opportunities for Impacting Industrial Catalysis. ACS Catal. 2015, 5, 1972–1985. [Google Scholar] [CrossRef]
  2. Bell, A.T. The Impact of Nanoscience on Heterogeneous Catalysis. Science 2003, 299, 1688–1691. [Google Scholar] [CrossRef] [PubMed]
  3. Nongkhlaw, R.L.; Rahman, N. Recent advances in organo-nano catalysis in the field of green organic synthesis. Arkivoc 2018, 6, 272–313. [Google Scholar] [CrossRef]
  4. Gawande, M.B.; Zboril, R.; Malgras, V.; Yamauchi, Y. Integrated nanocatalysts: A unique class of heterogeneous catalysts. J. Mater. Chem. A 2015, 3, 8241–8245. [Google Scholar] [CrossRef]
  5. 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]
  6. Xu, Q.; Song, Y.; Li, Y.; Liu, Z. Nanocatalysis for Organic Chemistry. Curr. Org. Chem. 2016, 20, 2013–2021. [Google Scholar] [CrossRef]
  7. 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]
  8. 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] [CrossRef]
  9. Ling, J.S.J.; Tan, Y.H.; Mubarak, N.M.; Kansedo, J.; Saptoro, A.; Nolasco-Hipolito, C. A review of heterogeneous calcium oxide based catalyst from waste for biodiesel synthesis. SN Appl. Sci. 2019, 1, 810. [Google Scholar] [CrossRef]
  10. Safaei-Ghomi, J.; Ghasemzadeh, M.A.; Mehrabi, M. Calcium oxide nanoparticles catalyzed one-step multicomponent synthesis of highly substituted pyridines in aqueous ethanol media. Sci. Iran. C 2013, 20, 549–554. [Google Scholar]
  11. Kasama, K.; Inoue, Y.; Yasumori, I. Catalysis by alkaline earth metal oxides. I. the mechanism of ethylene hydrogenation on calcium oxide. Bull. Chem. Soc. Jpn. 1980, 53, 1842–1847. [Google Scholar] [CrossRef]
  12. Kulkarni, P. Calcium Oxide Catalyzed Synthesis of Chalcone Under Microwave Condition. Curr. Microw. Chem. 2015, 2, 144–149. [Google Scholar] [CrossRef]
  13. Dhakar, A.; Goyal, R.; Rajput, A.; Kaurava, M.S.; Tomar, V.S.; Agarwal, D.D. Multicomponent synthesis of 4H-pyran derivatives using KOH loaded calcium oxide as catalyst in solvent free condition. Curr. Chem. Lett. 2019, 8, 125–136. [Google Scholar] [CrossRef]
  14. Tang, Y.; Zhang, L.; Wang, S.; Zhang, J.; Miao, Y. Synthesis of dihydropyrazole derivatives using modified calcium oxide as a solid basic catalyst. Prog. React. Kinet. Mech. 2015, 40, 409–418. [Google Scholar] [CrossRef]
  15. Lee, M.; Chen, B.-Y.; Den, W. Chitosan as a Natural Polymer for Heterogeneous Catalysts Support: A Short Review on Its Applications. Appl. Sci. 2015, 5, 1272–1283. [Google Scholar] [CrossRef]
  16. Guibal, E. Heterogeneous catalysis on chitosan-based materials: A review. Prog. Polym. Sci. 2005, 30, 71–109. [Google Scholar] [CrossRef]
  17. El Kadib, A. Chitosan as a Sustainable Organocatalyst: A Concise Overview. ChemSusChem 2014, 8, 217–244. [Google Scholar] [CrossRef] [PubMed]
  18. Khalil, K.; Al-Matar, H. Chitosan Based Heterogeneous Catalyses: Chitosan-Grafted-Poly(4-Vinylpyridne) as an Efficient Catalyst for Michael Additions and Alkylpyridazinyl Carbonitrile Oxidation. Molecules 2013, 18, 5288–5305. [Google Scholar] [CrossRef]
  19. Al-Matar, H.M.; Khalil, K.D.; Meier, H.; Elnagdi, M.H. Chitosan as heterogeneous catalyst in Michael additions: The reaction of cinnamonitriles with active methyls, active methylenes and phenols. Arkivoc 2008, 16, 288–301. [Google Scholar] [CrossRef]
  20. Khalil, K.; Al-Matar, H.; Elnagdi, M. Chitosan as an eco-friendly heterogeneous catalyst for Michael type addition reactions. A simple and efficient route to pyridones and phthalazines. Eur. J. Chem. 2010, 1, 252–258. [Google Scholar] [CrossRef]
  21. 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]
  22. 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]
  23. 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]
  24. Elazab, H.A.; Remiz, Y.G.; Radwan, M.A.; Radwan, M.A.; Sadek, M.A. Synthesis and Characterization of Chitosan based Catalyst for Catalysis Applications. Int. J. Adv. Trends Comput. Sci. Eng. 2020, 9, 521–527. [Google Scholar] [CrossRef]
  25. 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]
  26. 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]
  27. 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, 17, 3583. [Google Scholar] [CrossRef] [PubMed]
  28. Riyadh, S.M.; Khalil, K.D.; Aljuhani, A. Chitosan-MgO nanocomposite: One pot preparation and its utility as an ecofriendly biocatalyst in the synthesis of thiazoles and [1,3,4] thiadiazoles. Nanomaterials 2018, 8, 928. [Google Scholar] [CrossRef] [PubMed]
  29. Abdel-Naby, A.S.; Nabil, S.; Aldulaijan, S.; Ababutain, I.M.; Alghamdi, A.I.; Almubayedh, S.; Khalil, K.D. Synthesis, Characterization of Chitosan-Aluminum Oxide Nanocomposite for Green Synthesis of Annulated Imidazopyrazol Thione Derivatives. Polymers 2021, 13, 1160. [Google Scholar] [CrossRef]
  30. Khalil, K.D.; Riyadh, S.M.; Alkayal, N.S.; Bashal, A.H.; Alharbi, K.H.; Alharbi, W. Chitosan-Strontium Oxide Nanocomposite: Preparation, Characterization, and Catalytic Potency in Thiadiazoles Synthesis. Polymers 2022, 14, 2827. [Google Scholar] [CrossRef] [PubMed]
  31. Xu, H.; Liao, W.M.; Li, H.F. A mild and efficient ultrasound-assisted synthesis of diaryl ethers without any catalyst. Ultrason. Sonochem. 2007, 14, 779–782. [Google Scholar] [CrossRef]
  32. Jarag, K.J.; Pinjari, D.V.; Pandit, A.B.; Shankarling, G.S. Synthesis of chalcone (3-(4-fluorophenyl)-1-(4-methoxyphenyl) prop-2-en-1-one): Advantage of sonochemical method over conventional method. Ultrason. Sonochem. 2011, 18, 617–623. [Google Scholar] [CrossRef] [PubMed]
  33. Alshabanah, L.A.; Al-Mutabagani, L.A.; Gomha, S.M.; Ahmed, H.A. Three-component synthesis of some new coumarin derivatives as anti-cancer agents. Front. Chem. 2022, 9, 762248. [Google Scholar] [CrossRef] [PubMed]
  34. Cravotto, G.; Fokin, V.V.; Garella, D.; Binello, A.; Boffa, L.; Barge, A. Ultrasound-Promoted Copper-Catalyzed Azide-Alkyne Cycloaddition. J. Comb. Chem. 2010, 12, 13–15. [Google Scholar] [CrossRef]
  35. Gomha, S.M.; Khalil, K.D. A Convenient ultrasound-promoted synthesis and cytotoxic activity of some new thiazole derivatives bearing a coumarin nucleus. Molecules 2012, 17, 9335–9347. [Google Scholar] [CrossRef]
  36. Pizzuti, L.; Martins, P.L.G.; Ribeiro, B.A.; Quina, F.H.; Pinto, E.; Flores, A.F.C.; Venzke, D.; Pereira, C.M.P. Efficient sonochemical synthesis of novel 3,5-diaryl-4,5-dihydro-1H-pyrazole-1-carboximidamides. Ultrason. Sonochem. 2010, 17, 34–37. [Google Scholar] [CrossRef] [PubMed]
  37. Franchetti, P.; Cappellacci, L.; Grifantini, M.; Barzi, A.; Nocentini, G.; Yang, H.; O’Connor, A.; Jayaram, H.N.; Carrell, C.; Goldstein, B.M. Furanfurin and thiophenfurin: Two novel tiazofurin analogues. Synthesis, structure, antitumour activity, and interactions with inosine monophosphate dehydrogenase. J. Med. Chem. 1995, 38, 3829–3837. [Google Scholar]
  38. Li, X.; He, Y.; Ruiz, C.H.; Koenig, M.; Cameron, M.D. Characterization of dasatinib and its structural analogs as CYP3A4 mechanism-based inactivators and the proposed bioactivation pathways. Drug Metab. Dispos. 2009, 37, 1242–1250. [Google Scholar] [CrossRef]
  39. Hu-Lieskovan, S.; Mok, S.; Homet Moreno, B.; Tsoi, J.; Robert, L.; Goedert, L.; Pinheiro, E.M.; Koya, R.C.; Graeber, T.G.; Comin-Anduix, B.; et al. Improved antitumour activity of immunotherapy with B-RAF and MEK inhibitors in BRAF(V600E) melanoma. Sci. Transl. Med. 2015, 18, 279ra41. [Google Scholar]
  40. Yao, Y.; Chen, S.; Zhou, X.; Xie, L.; Chen, A. 5-FU and ixabepilone modify the microRNA expression profiles in MDA-MB-453 triple-negative breast cancer cells. Oncol. Lett. 2014, 7, 541–547. [Google Scholar] [CrossRef] [PubMed]
  41. Altmann, K.H. Epothilone B and its analogs—A new family of anti-cancer agents. Mini Rev. Med. Chem. 2003, 3, 149–158. [Google Scholar] [CrossRef] [PubMed]
  42. Aliabadi, A.; Shamsa, F.; Ostad, S.N.; Emami, S.; Shafiee, A.; Davoodi, J.; Foroumadi, A. Synthesis and biological evaluation of 2-phenylthiazole-4-carboxamide derivatives as anti-cancer agents. Eur. J. Med. Chem. 2010, 45, 5384–5389. [Google Scholar] [CrossRef] [PubMed]
  43. Banimustafa, M.; Kheirollahi, A.; Safavi, M.; Ardestani, S.K.; Aryapour, H.; Foroumadi, A.; Emami, S. Synthesis and biological evaluation of 3-(trimethoxyphenyl)-2(3H)-thiazole thiones as combretastatin analogs. Eur. J. Med. Chem. 2013, 70, 692–702. [Google Scholar] [CrossRef]
  44. Ayati, A.; Emami, S.; Asadipour, A.; Shafiee, A.; Foroumadi, A. Recent applications of 1,3-thiazole core structure in the identification of new lead compounds and drug discovery. Eur. J. Med. Chem. 2015, 97, 699–718. [Google Scholar] [CrossRef]
  45. Das, D.; Sikdar, P.; Bairagi, M. Recent developments of 2-aminothiazoles in medicinal chemistry. Eur. J. Med. Chem. 2016, 109, 89–98. [Google Scholar] [CrossRef]
  46. Habte, L.; Shiferaw, N.; Mulatu, D.; Thenepalli, T.; Chilakala, R.; Ahn, J.W. Synthesis of Nano-Calcium Oxide from Waste Eggshell by Sol-Gel Method. Sustainability 2019, 11, 3196. [Google Scholar] [CrossRef]
  47. Margaretha, Y.Y.; Prastyo, H.S.; Ayucitra, A.; Ismadji, S. Calcium oxide from Pomacea sp. shell as a catalyst for biodiesel production. Int. J. Energy Environ. Eng. 2012, 3, 33. [Google Scholar] [CrossRef]
  48. Klug, H.P.; Alexander, L.E. X-ray Diffraction Procedures; John Wiley and Sons Inc.: New York, NY, USA, 1954; p. 633. [Google Scholar]
  49. Abbas, E.M.H.; Gomha, S.M.; Farghaly, T.A. Multicomponent reactions for synthesis of bioactive polyheterocyclic ring systems under controlled microwave irradiation. Arab. J. Chem. 2014, 7, 623–629. [Google Scholar] [CrossRef]
  50. Gomha, S.M.; Riyadh, S.M. Synthesis of triazolo[4,3-b][1,2,4,5]tetrazines and triazolo[3,4-b][1,3,4]thiadiazines using chitosan as ecofriendly catalyst under microwave irradiation. Arkivoc 2009, 6, 58–68. [Google Scholar] [CrossRef]
  51. Abu-Melha, S.; Gomha, S.M.; Abouzied, A.S.; Edrees, M.M.; Abo Dena, A.S.; Muhammad, Z.A. Microwave-assisted one pot three-component synthesis of novel bioactive thiazolyl-pyridazinediones as potential antimicrobial agents against antibiotic-resistant bacteria. Molecules 2021, 26, 4260. [Google Scholar] [CrossRef] [PubMed]
  52. Rashdan, H.R.M.; Gomha, S.M.; El-Gendey, M.S.; El-Hashash, M.A.; Soliman, A.M.M. Eco-friendly one-pot synthesis of some new pyrazolo[1,2-b]phthalazinediones with antiproliferative efficacy on human hepatic cancer cell lines. Green Chem. Lett. Rev. 2018, 11, 264–274. [Google Scholar] [CrossRef]
  53. Sayed, A.R.; Gomha, S.M.; Taher, E.A.; Muhammad, Z.A.; El-Seedi, H.R.; Gaber, H.M.; Ahmed, M.M. One-pot synthesis of novel thiazoles as potential anti-cancer agents. Drug Des. Dev. Ther. 2020, 14, 1363–1375. [Google Scholar] [CrossRef] [PubMed]
  54. Gomha, S.M.; Edrees, M.M.; Muhammad, Z.A.; El-Reedy, A.A.M. 5-(Thiophen-2-yl)-1,3,4-thiadiazole derivatives: Synthesis, molecular docking and in-vitro cytotoxicity evaluation as potential anti-cancer agents. Drug Des. Dev. Ther. 2018, 12, 1511–1523. [Google Scholar] [CrossRef] [PubMed]
  55. Omar, R.S.; Ragheb, M.A.; Elwahy, A.H.M.; Abdelhamid, I.A. Facile one-pot, three-component synthesis of novel fused 4H-pyrans incorporating 2-phenoxy-N-phenylacetamide core as novel hybrid molecules via Michael addition reaction. Arkivoc 2021, 10, 183–198. [Google Scholar] [CrossRef]
Figure 1. Examples of thiazole-bearing anti-cancer drugs.
Figure 1. Examples of thiazole-bearing anti-cancer drugs.
Polymers 14 03347 g001
Scheme 1. A simplified view of a chitosan–CaO nanocomposite.
Scheme 1. A simplified view of a chitosan–CaO nanocomposite.
Polymers 14 03347 sch001
Figure 2. FTIR of (A) chitosan, (B) CaO nanoparticles, and (C) chitosan–CaO nanocomposites (15 wt%).
Figure 2. FTIR of (A) chitosan, (B) CaO nanoparticles, and (C) chitosan–CaO nanocomposites (15 wt%).
Polymers 14 03347 g002
Figure 3. XRD of chitosan (A) and the chitosan–CaO nanocomposite (15 wt%) (B).
Figure 3. XRD of chitosan (A) and the chitosan–CaO nanocomposite (15 wt%) (B).
Polymers 14 03347 g003
Figure 4. FESEM of chitosan (A), and the chitosan–CaO nanocomposite (15 wt%) (B).
Figure 4. FESEM of chitosan (A), and the chitosan–CaO nanocomposite (15 wt%) (B).
Polymers 14 03347 g004
Figure 5. EDS of the chitosan–CaO nanocomposites (15 wt%).
Figure 5. EDS of the chitosan–CaO nanocomposites (15 wt%).
Polymers 14 03347 g005
Scheme 2. Synthesis of arylazothiazoles 8a–g.
Scheme 2. Synthesis of arylazothiazoles 8a–g.
Polymers 14 03347 sch002
Table 1. The time of reaction and the yield % of the products compared in the synthesis of thiazoles 8a–g under ultrasonic irradiation utilizing two different basic catalysts.
Table 1. The time of reaction and the yield % of the products compared in the synthesis of thiazoles 8a–g under ultrasonic irradiation utilizing two different basic catalysts.
Compound No.XArTEACS–CaO Nanocomposite
Time (Min)(%) YieldTime (Min)(%) Yield
8aClC6H538722191
8bCl4-MeOC6H443702488
8cCl4-ClC6H429722190
8dCl2,4-diClC6H325752092
8eBrC6H542712488
8fBr4-MeOC6H444692787
8gBr4-ClC6H433732489
Table 2. Optimization of the reaction conditions (catalyst loading, solvent, reaction time, and temperature) for the prepared component 8a.
Table 2. Optimization of the reaction conditions (catalyst loading, solvent, reaction time, and temperature) for the prepared component 8a.
EntryCatalyst (Mol%)SolventTime (Min)Temperature (°C)Yield (%)
15 (0.065 g)EtOH234061
210 (0.13 g)EtOH234083
3 a20 (0.26 g)EtOH234091
420 (0.26 g)Dioxane234085
520 (0.26 g)DMSO234082
620 (0.26 g)EtOH204087
720 (0.26 g)EtOH304090
820 (0.26 g)EtOH232573
920 (0.26 g)EtOH233587
1020 (0.26 g)EtOH235091
a The best reaction condition for the synthesis of component 8a.
Table 3. Recyclability of the CS–CaO nanocomposite as a basic catalyst.
Table 3. Recyclability of the CS–CaO nanocomposite as a basic catalyst.
State of CatalystFresh CatalystRecycled
1
Recycled 2Recycled
3
Recycled 4
Product 8a (% Yield)9190888553
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khalil, K.D.; Ahmed, H.A.; Bashal, A.H.; Bräse, S.; Nayl, A.A.; Gomha, S.M. Efficient, Recyclable, and Heterogeneous Base Nanocatalyst for Thiazoles with a Chitosan-Capped Calcium Oxide Nanocomposite. Polymers 2022, 14, 3347. https://doi.org/10.3390/polym14163347

AMA Style

Khalil KD, Ahmed HA, Bashal AH, Bräse S, Nayl AA, Gomha SM. Efficient, Recyclable, and Heterogeneous Base Nanocatalyst for Thiazoles with a Chitosan-Capped Calcium Oxide Nanocomposite. Polymers. 2022; 14(16):3347. https://doi.org/10.3390/polym14163347

Chicago/Turabian Style

Khalil, Khaled D., Hoda A. Ahmed, Ali H. Bashal, Stefan Bräse, AbdElAziz A. Nayl, and Sobhi M. Gomha. 2022. "Efficient, Recyclable, and Heterogeneous Base Nanocatalyst for Thiazoles with a Chitosan-Capped Calcium Oxide Nanocomposite" Polymers 14, no. 16: 3347. https://doi.org/10.3390/polym14163347

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