Seyede Elaheh Ghafouria,
Hamid Goudarziafshar*a and
Ahmad Reza Moosavi-Zare*b
aDepartment of Chemical Engineering, Hamedan University of Technology, Hamedan, 65155, Iran. E-mail: , hamid_gafshar@yahoo.com
bChemistry Department, College of Sciences, Shiraz University, Shiraz 71946-84795, Iran. E-mail: moosavizare@yahoo.com
First published on 7th July 2025
In this research, Ag@SiO2 nanoparticles were prepared by a sol–gel method (a wet chemical method) at room temperature. The nanoparticles were functionalized by chlorosulfonic acid to introduce acidic sites on the surface to obtain sulfonic acid grafted on silica-coated silver nanoparticles (Ag@SiO2–SO3H). Ag@SiO2–SO3H was identified and characterized by FT-IR, DLS, ZP, FE-SEM, EDX-mapping, TEM, XRD, BET, BJH and TGA analyses. Ag@SiO2–SO3H performed well as a reusable catalyst for the multi-component synthesis of pyrimido[4,5-b]quinoline derivatives with high efficiency and satisfactory results.
Some important pyrimido[4,5-b]quinolone derivatives were prepared by the three-component reaction of an aldehyde with dimedone and 6-amino-1,3-dimethyluracil.7 Multi-component synthesis is an important protocol in green chemistry that has recently been attracted much attention because it allows the preparation of products in one step without generating side products during the reaction; moreover, this approach prevents the waste of energy, materials, and solvents, shortens reaction times, and increases reaction efficiency.9–15
Various catalysts have been used for the multi-component preparation of pyrimido[4,5-b]quinolones, including SBA-15/PrN(CH2PO3H2)2,16 trityl chloride (TrCl),17 nano-[Fe3O4@SiO2/N-propyl-1-(thiophene-2-yl)ethanimine][ZnCl2],18 [TSSECM],19 N,N-diethyl-N-sulfoethanaminium chloride,20 [H2-DABCO][ClO4]2,21 Fe3O4@cellulose sulfuric acid,22 glycolic acid-supported cobalt ferrite,23 nano-[Fe3O4@-SiO2@R-NHMe2][H2PO4],24 [bmim]Br,25 nanocrystalline MgO,26 agar-entrapped sulfonated DABCO,27 [C4(DABCO)2]·2OH,28 DABCO,29 cellulose sulfuric acid,30 nano-[Fe3O4@SiO2@BDIL],31 Cs2.3H0.7PW10Mo2O40,32 Fe3O4@SiO2–SO3H,33 DABCO-based ionic liquid,34 nano-[Cu–4C3NSP](Cl)2,35 [Fe3O4@SiO2@RNMe2–SO3H][CF3CO2],36 GO/Fe3O4@PTRMS@BDSA@SO3H,37 TiO2/porous carbon nanocomposite,38 RuCl3·xH2O,39 and triethylbenzylammonium chloride (TEBAC)40 are some catalyst which recently used and reported for the preparation of pyrimido[4,5-b]quinolines. Given the important biological activities exhibited by pyrimido[4,5-b]quinoline derivatives, finding new methods for the synthesis of these compounds remains necessary.
Recently, various supported acidic catalysts have been introduced as catalysts in organic transformations. Some catalysts with sulfonic acid moieties were designed and used in reactions owing to their ease of use compared to common inorganic acids in organic reactions. Silica,41 mesoporous compounds,42 silica-coated magnetite nanoparticles,43 inorganic materials,44 organic materials45–47 and ionic liquids48–50 are examples of supports that are used for the preparation of supported or grafted sulfonic acid catalysts.
Silver nanoparticles are used as antimicrobial agents, biosensors, composite fibers, cosmetics, detergents, and electronic components due to their unique properties. In addition, they have attracted the attention of researchers in the fields of medical imaging, nanocomposites, filters, drug delivery, and biolabeling.51 Ag@SiO2 has also been applied as a catalyst for various transformations such as low-temperature CO oxidation,52,53 benzene combustion,54 redox reactions on surfaces,55 oxidative dehydrogenation of methanol,56 reduction of mixtures of 4-nitrophenol (4-NP) and rhodamine-B (Rh-B),57 decomposition of ozone,58 selective hydrogenation of acetylene,59 hydrogenation of 4-nitrophenol,60 diethyl oxalate hydrogenation to ethyl glycolate,61 hydrogenation of nitrobenzene,62 reduction of methylene blue,63 semi-hydrogenation of dimethyl glycolate,64 hydrogenation of dimethyl oxalate to methyl glycolate65 and preparation of methyl glycolate through hydrogenation of dimethyl oxalate.66 Moreover, catalysts have been designed using silver nanoparticles such as Ag/SiO2–Al2O3,67 Ag/SiO2–CoFe2O4,68 Pd–Ag/SiO2,69 Ni–Ag/SiO2 (ref. 70) and Cu@Ag/SiO2 (ref. 71) to catalyze various chemical reactions.
Considering the above context, in the present work, silica-coated silver nanoparticles were used to prepare sulfonic acid supported on silica-coated silver nanoparticles (Ag@SiO2–SO3H) as a heterogeneous catalyst for the efficient multi-component synthesis of pyrimido[4,5-b]quinolines (Scheme 1).
The preparation of the Ag@SiO2–SO3H was confirmed by various analyses such as Fourier transform infrared spectroscopy (FT-IR), dynamic light scattering (DLS), zeta potential (ZP), atomic force microscopy (AFM), field emission-scanning electron microscopy (FE-SEM), energy dispersive X-ray (EDX), SEM coupled EDX (SEM mapping), transmission electron microscopy (TEM), X-ray diffraction (XRD) analysis, X-ray photoelectron spectroscopy (XPS), Brunauer–Emmett–Teller analysis (BET), and thermogravimetric analysis (TGA).
FT-IR spectroscopy was performed to identify specific functional groups in the structure of the catalyst. In the FT-IR spectrum of Ag@SiO2, the characteristic absorption peak at 460 cm−1 corresponding to the Si–O bending vibration was observed.72 The identified absorption peaks at 1039, 1261 and 1384 cm−1 correspond to Si–O–Si bending vibrations, and symmetric and asymmetric Si–O–Si stretching vibrations, respectively. Absorption peaks at 1634 and 3419 cm−1 correspond to hydroxyl groups in Si–OH, OH bending and stretching vibrations, respectively.72 The presence of a Si–O functional group in the FT-IR spectrum indicates the successful formation of Ag@SiO2 nanoparticles (Fig. 2). In the FT-IR spectrum of Ag@SiO2–SO3H, characteristic absorption peaks at 887 and 1007 cm−1 correspond to S–O bending and stretching vibrations, 1070 and 1175 cm−1 correspond to SO symmetric and asymmetric bending, respectively, and 1296 cm−1 corresponds to S
O stretching vibrations. The presence of S–O bonds is evidence for the successful formation of the catalyst (Fig. 2).
The energy dispersive X-ray (EDX) spectrum of Ag@SiO2–SO3H nanoparticles is presented in Fig. 3. This pattern clearly indicates the presence of expected elements, namely, silver, silicon, sulfur and oxygen, in the structure of the core–shell catalyst nanoparticles. The extra peaks observed in this spectrum are caused by gold and platinum impurities introduced during the coating for analysis. Thus, the EDX analysis confirms the successful formation of Ag@SiO2–SO3H (Fig. 3).
To determine the distribution of elements in Ag@SiO2–SO3H, SEM-coupled EDX (SEM mapping) analysis was performed. Based on this analysis, it was found that elements including silver, oxygen, silicon and sulfur were well dispersed over the structure of Ag@SiO2–SO3H with an appropriate distribution (Fig. 4).
To determine the size and distribution of the catalyst, dynamic light scattering (DLS) analysis was performed. The average particle sizes of Ag@SiO2 and Ag@SiO2–SO3H were observed to be 22.6 and 77.5 nm, respectively. In DLS analysis, a uniform sample of nanoparticles is shown as a narrow peak at a specific size; on the other hand, a broad peak or multiple peaks can be related to a less uniform sample of nanoparticles with a range of particle sizes. The observed sharp and narrow peak for Ag@SiO2–SO3H nanoparticles in the DLS analysis confirms the uniformity of the sample. The functionalization of the Ag@SiO2 nanoparticles with SO3H increased the size and acidity of the prepared nanoparticles (Fig. 5).
To determine the stability of Ag@SiO2 and Ag@SiO2–SO3H nanoparticles, their zeta potentials were measured. The zeta potentials were 42.5 and −23.3 mV, respectively (Fig. 6). According to the standards of the DLS method, the particle size and zeta potential values indicate the stability of the catalyst (Table 1). The negative zeta potential indicates a high repulsion between nanoparticles; this repulsive force prevents aggregation and increases the long-term stability of metal colloids.73
Sample | DLS (nm) | ZP (mv) | pH |
---|---|---|---|
Ag@SiO2 | 22.6 | 42.5 | 3.7 |
Ag@SiO2–SO3H | 77.5 | −23.3 | 2.3 |
The boundary between unstable and stable suspensions is often found at 30 or −30 mV; i.e., particles with zeta potentials that are more positive than 30 mV or more negative than −30 mV. If a zeta potential is more positive than 30 mV and more negative than −30 mV, the resulting suspension is stable. A zeta potential with values more positive than −15 mV indicates that the suspension is on the verge of agglomeration.74 The coagulation or flocculation process occurs more rapidly when the zeta potential is in the range of 30 mV. The observed zeta potential thus reflects the stability of the particles.
To study the Ag@SiO2–SO3H nanoparticles in more detail, atomic force microscopy (AFM) images were recorded. The 2D and 3D topographies of the synthesized Ag@SiO2–SO3H are shown in Fig. 7. The initial image with a color bar is shown in Fig. 7A. Fig. 7B and D show 2D and enhanced images (improved with MATLAB software) that reveal large numbers of small spherical particles. The distribution of the particles is uniform and the size of the particles in this image is 42.5 nm. In the 3D image (Fig. 7C), the distribution of particles is uniform, and the image shape is improved with MATLAB software. The color change in the color bar from light to dark indicates an increase in size. The topography of the AFM images reveals that Ag@SiO2–SO3H was prepared with a high number of particles of small sizes. The surfaces show very low roughness with a homogeneous structure (Fig. 7).
Field emission-scanning electron microscopy (FE-SEM) was used to determine the size and morphology of the catalyst particles (Fig. 8). The micrographs show that the particles were prepared on the nanometer scale, typically less than 100 nm. The spherical nanoparticles that are observed in the images are related to the Ag@SiO2–SO3H nanoparticles. The distribution of particles is depicted as a histogram in Fig. 8.
Transmission electron microscopy (TEM) of Ag@SiO2–SO3H was conducted to observe the size and core–shell structure of Ag@SiO2–SO3H nanoparticles. The TEM images confirmed that the Ag@SiO2–SO3H nanoparticles have a core–shell structure. These nanoparticles have a dark core of Ag and light shell of sulfonic–silica, indicating that Ag nanoparticles are coated with silica and functionalized with sulfonic acid (Fig. 9).
The X-ray diffraction (XRD) pattern of Ag@SiO2–SO3H contained peaks at 2θ = 27.59, 31.97, 38.11, 44.30, 45.86, 54.37, 57.00, 64.44, 76.08, 77.40 and 81.54°; for silver, this corresponds to miller indices (hkl) at (111), (002), (022), (113) and (222) crystal planes, respectively, as previously reported for Ag@SiO2 nanoparticles.75 The crystal sizes of the prepared catalyst particles were calculated based on Debbie Scherer's equation and found to be 33.67, 30.75, 20.58, 11.79, 24.57, 23.67, 20.66, 5.68, 9.31, 3.97 and 2.31 nm (Fig. 10).
X-ray photoelectron spectroscopy (XPS) was conducted for the AgNPs and the binding energies of electrons observed in the XPS spectrum were compared to the reported values. The reference peak for carbon is 284.5 eV according to the references. On the other hand, the reference peak in the sample spectrum was observed at 289.2 eV; to correct for this difference, the binding energy of the observed peaks should be reduced by 4.7 eV (289.2 − 284.5 = 4.7 eV).76 The survey spectrum shows what elements are present in the sample (the carbon peak is considered as a reference peak), and the displacement of other elements relative to carbon is measured. In the survey spectrum, carbon, silver and oxygen elements are observed. The presence of silver is evidence of the synthesis of the nanoparticles whereas the presence of carbon and oxygen can be due to surface contamination introduced during sample preparation or from oxygen in the air (or due to the materials used for synthesis, such as CTAB or ascorbic acid).77 Two peaks, 3d5/2 and 3d3/2, with binding energies of 372.7 eV and 378.9 eV, (372.7 − 4.7 = 368 eV and 378.9 − 4.7 = 373.9 eV) are attributed to the presence of Ag0.76 The difference in energy between the two silver peaks is ΔBE = 5.98 eV.76 For the oxygen element, the binding energy is 537 eV (537 − 4.7 = 532.3 eV), the binding energy of 578 eV corresponds to silver 3p3 and 608 eV corresponds to the silver 3p1 layer (608 − 4.7 = 603.3 and 578 − 4.7 = 573.3 eV).76 Two broad peaks with binding energy 1007.15–981.7 eV (1007.15 − 4.7 = 1002.45 eV and 981.7 − 4.7 = 977 eV) correspond to oxygen electrons and the broad peak 1140.2 eV (1140.2 − 4.7 = 1135.5 eV) is related to silver Auger electrons (Fig. 11, 12 and Table 2).76
![]() | ||
Fig. 12 XPS survey scan (spectra) of the AgNPs and valence band spectra in the vicinity of the Fermi level. |
Peaks | Binding energy (eV) | |
---|---|---|
Reported | (Lit)71 | |
Ag Auger | 1136.0 | 1134.9 |
Ag 3p1 | 603.19 | 604.0 |
Ag 3p3 | 573.3 | 573.6 |
Ag 3d | 368.47–374.45 | 368.4 |
C 1s | 285.0 | 284.5 |
The thermal stability of the catalyst was investigated by thermal gravimetric analysis (TGA), derivative thermogravimetric (DTG) analysis and differential thermal analysis (DTA). According to the results obtained from this analysis, Ag@SiO2–SO3H can be used from room temperature up to 190 °C, with only 5.1% weight loss in this temperature range (Fig. 13). In this study, two-step weight losses were observed in the TG diagram of the catalyst. The first weight loss step is attributed to the loss of physically adsorbed solvents on the surface of the catalyst and the second weight loss occurred upon the loss of SO3H groups from the catalyst above 190 °C.78 Above this temperature, the Ag@SiO2 core–shell structure was maintained with high thermal stability.79,80 According to the TGA, DTG and DTA analyses, it is concluded that this catalyst shows acceptable thermal stability in chemical reactions.78–80
According to IUPAC classification, the hysteresis loop in the adsorption and desorption curve is of type H4, which is related to mesoporous materials, and the adsorption and desorption profiles do not match, which is caused by factors such as the shape of the cavities, the structure of the material, the method of producing the material, and the physical purity of the material. According to IUPAC, the morphology of particles with irregular internal cavities is in the form of narrow slits, which shows a mixture of non-uniform mesopores and micropores. In materials that exhibit this type of hysteresis loop, the adsorbate–adsorbate interactions are relatively strong (Fig. 14). Based on the profile of the BJH diagram (Fig. 14), the maximum size of the holes was observed in the range of 2 to 22 nm (with a peak at 9 nm), which corresponds to the diameter of the holes obtained from BET and shows the catalytic activity of the catalyst for the synthesis of acyclic reactions. The specific surface area was 1.91 (m2 g−1), the BET constant value was 357.24 nm, the total pore volume was 0.024 (cm3 g−1), and the average pore diameter was 50 nm. For the BET diagram, deviation from the straight line occurred at approximately 0.33 and the interpretable range in BET was 0.05–0.3 (Fig. 14 and Table 3).
Ag@SiO2–SO3H | |||
---|---|---|---|
aSBET (m2 g−1) | BET (C) nm | Total pore (cm3 g−1) | Mean pore diameter (nm) |
1.91 | 357.24 | 0.024 | 50.24 |
After the preparation and identification of Ag@SiO2–SO3H, its catalytic ability was tested with the condensation reaction of 2-chlorobenzaldehyde with dimedone and 6-amino-1,3-dimethyluracil to establish the best reaction conditions. Temperature, catalyst amounts and various solvents, including ethanol, chloroform, dichloromethane, ethyl acetate and dimethylformamide, were investigated; solvent-free conditions were also examined. It was found that acceptable results were obtained when the model reaction was carried out in the presence of 3 mg of catalyst under solvent-free conditions (Table 4).
Entry | Solvent | Catalyst amount (mg) | Temp. (°C) | Time (min) | Yielda (%) |
---|---|---|---|---|---|
a Isolated yield. | |||||
1 | — | — | 80 | 120 | 10 |
2 | — | 1.5 | 80 | 50 | 47 |
3 | — | 3 | 80 | 10 | 95 |
4 | — | 4.5 | 80 | 10 | 52 |
5 | Ethanol | 3 | Reflux | 10 | 66 |
6 | CH3Cl | 3 | Reflux | 10 | 88 |
7 | Ethyl acetate | 3 | Reflux | 10 | 74 |
8 | DCM | 3 | Reflux | 10 | 72 |
9 | DMF | 3 | Reflux | 10 | 74 |
10 | n-Hexane | 3 | Reflux | 10 | 52 |
11 | — | 3 | 50 | 30 | 49 |
12 | — | 3 | 110 | 10 | 70 |
With the optimized reaction conditions in hand, the generality of the method was studied with a series of aromatic aldehydes containing electron-donating or electron-withdrawing groups and halogen on their ring to prepare the corresponding pyrimido[4,5-b]quinoline derivatives (Table 5). All the products were prepared in high yields and short reaction times.
Entry | Product | Time (min) | Yielda (%) | M.p. °C (Lit.) |
---|---|---|---|---|
a Isolated yield. | ||||
1 | ![]() |
15 | 89 | 244–245 (253–255)17 |
2 | ![]() |
10 | 81 | 346–350 (—) |
3 | ![]() |
12 | 89 | 247–249 (238–240)37 |
4 | ![]() |
10 | 95 | 347–351 (350–355)17 |
5 | ![]() |
19 | 85 | 293–296 (>300)24 |
6 | ![]() |
24 | 79 | 346–349 (—) |
7 | ![]() |
25 | 78 | 289–293 (285–287)24 |
8 | ![]() |
30 | 85 | 296–298 (>300) |
9 | ![]() |
10 | 85 | 288–289 (292–294)24 |
10 | ![]() |
40 | 83 | 298–301 (300–303)17 |
11 | ![]() |
10 | 97 | 225–229 (226–229)35 |
12 | ![]() |
35 | 95 | 315–320 (>300)24 |
13 | ![]() |
20 | 93 | 278–282 (301–303)35 |
14 | ![]() |
25 | 99 | 310–314 (>300)32 |
15 | ![]() |
10 | 98 | 280–284 (208–212)19 |
A suggested mechanism, supported by previous reports,16,18,35,81 illustrates the role of the catalyst in the preparation of pyrimido[4,5-b]quinolines (Scheme 2). The enol form of dimedone first reacts with the aldehyde, which is activated with Ag@SiO2–SO3H to prepare (I). By the elimination of one molecule of H2O, intermediate (II) is generated as a Michael acceptor. In the next step, 6-amino-1,3-dimethyluracil reacts with the presented Michael acceptor (II) to prepare (III) after tautomerization. Finally, intramolecular nucleophilic attack of the amine on the carbonyl group in intermediate (III) and ring closure occur, followed by the removal of one molecule of H2O, to generate the final product (Scheme 2).
In this study, silver nanoparticles were used as a suitable support for preparing an acidic catalyst with a nanoscale core–shell structure. Coating the silver nanoparticles with silica prevents the particles from coalescing and agglomerating, and also prevents these particles from undergoing reactions such as oxidation and reduction, and protects them from interacting with the acidic part of the catalyst. Coating the silver nanoparticles with silica increases the stability of the catalyst and its lifetime. The design of this solid acidic catalyst at the nanoscale using silver nanoparticles as a support increases its contact surface, accelerates the reaction rate, and simplifies the use of the catalyst.
To investigate the reusability of Ag@SiO2–SO3H, the reaction of 4-chlorobenzaldehyde with dimedone and 6-amino-1,3-dimethyluracil was carried out in the presence of this catalyst; upon completion of the reaction, warm absolute ethanol was added and the reaction mixture was stirred to extract the remaining starting materials and product. The catalyst was centrifuged and separated from the reaction mixture. The recovered catalyst was successfully reused in four further reactions and the related product was prepared with acceptable yield and suitable reaction time (Fig. 15). The reaction time for each reaction was 15 minutes. To confirm that the structure of the catalyst was unchanged after recovery, the FT-IR spectra of fresh and reused catalysts were compared (Fig. 16). According to this investigation, the important bond vibrations related to specific bonds were retained.
The preparation of the related pyrimido[4,5-b]quinoline by the reaction of 4-nitrobenzaldehyde with dimedone and 6-amino-1,3-dimethyluracil catalyzed by Ag@SiO2@SO3H was compared in terms of yield and reaction time under a range of conditions and catalysts; the results are summarized in Table 6. The results indicate that the present catalyst and conditions delivered superior outcomes to those obtained using previously reported methods.
Entry | Catalyst (amount) | Condition | Time (min) | Yielda (%) | Ref. |
---|---|---|---|---|---|
a Isolated yield. | |||||
1 | [H2-DABCO][ClO4]2 (0.19 mmol, 0.060 g) | H2O, 75 °C | 30 | 85 | 21 |
2 | GO/Fe3O4@TRMS@BDSA@SO3H (0.05 g) | SF, 70 °C | 14 | 92 | 37 |
3 | Nano-[Cu–4C3NSP](Cl)2 (6 mol%) | EtOH, 60 °C | 45 | 80 | 35 |
4 | [Dsim]HSO4, (0.25 mmol) | EtOH, 70 °C | 15 | 92 | 81 |
5 | TiO2/PCN (0.006 g) | EtOH, reflux | 40 | 80 | 38 |
6 | Nano-[Fe3O4@SiO2/N-propyl-1-(thiophene-2-yl)ethanimine][ZnCl2] (0.02 g) | EtOH, 60 °C | 20 | 94 | 18 |
7 | RuCl3·xH2O (3 mol%) | H2O, 85 °C | 25 | 91 | 39 |
8 | TEBAC (0.15 g) | H2O, 90 °C | 13 | 86 | 40 |
9 | Ag@SiO2–SO3H (10 mol%) | SF, 80 °C | 10 | 97 | Our work |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5na00393h |
This journal is © The Royal Society of Chemistry 2025 |