Sobhan Rezayatia,
Fatemeh Kalantaria and
Ali Ramazani*ab
aDepartment of Chemistry, Faculty of Science, University of Zanjan, Zanjan 45371-38791, Iran. E-mail: aliramazani@gmail.com; aliramazani@znu.ac.ir
bDepartment of Biotechnology, Research Institute of Modern Biological Techniques (RIMBT), University of Zanjan, Zanjan 45371-38791, Iran
First published on 25th April 2023
In the current study, an environmentally friendly and facile method was proposed for designing and constructing a catalyst with Ni(II) attached to a picolylamine complex on 1,3,5-triazine-immobilized Fe3O4 core–shell magnetic nanoparticles (NiII-picolylamine/TCT/APTES@SiO2@Fe3O4) via a stepwise procedure. The as-synthesized nanocatalyst was identified and characterized via Fourier-transform infrared (FT-IR), X-ray photoelectron spectroscopy (XPS), thermogravimetric analysis (TGA), vibrating-sample magnetometry (VSM), transmission electron microscopy (TEM), X-ray diffraction (XRD), Brunauer–Emmett–Teller (BET), field-emission scanning electron microscopy (FE-SEM), inductively coupled plasma (ICP), and energy-dispersive X-ray spectrometry (EDX). The obtained results from the BET analysis indicated that the synthesized nanocatalyst had high specific area (53.61 m2 g−1) and mesoporous structure. TEM observations confirmed the particle size distribution was in the range 23–33 nm. Moreover, the binding energy peaks observed at 855.8 and 864.9 eV in the XPS analysis confirmed the successful and stable attachment of Ni(II) on the surface of the picolylamine/TCT/APTES@SiO2@Fe3O4. The as-fabricated catalyst was used to produce pyridine derivatives by the one-pot pseudo-four component reaction of malononitrile, thiophenol, and a variety of aldehyde derivatives under solvent-free conditions or EG at 80 °C. The highest yield achieved was 97% for compound 4d in EG at 80 °C with a TOF of 823 h−1 and TON of 107. It was found that the used catalyst was recyclable for eight consecutive cycles. On the basis of ICP analysis, the results indicated that the Ni leaching was approximately 1%.
One of the most important modern science areas is nanoscience. Nanotechnology allows physicians, engineers, scientists, and chemists to do work at the cellular and molecular levels to make important advances in the healthcare and science fields, among others. In recent years, due to their electrical characteristics, distinctive magnetic properties, unique size, and high surface area, the utilization of magnetic nanoparticles (MNPs) has attracted considerable attention. They also have particular properties in industrial applications and biomedical applications, such as drug delivery systems, separation of metal ions for environmental remediation, bio-isolation, saving information, biomolecular sensing, magneto-heat therapy, targeted gene therapy, and sensors.14–16 A high biocompatibility and sensitivity to postsynthetic surface functionalization have made magmite (γ-Fe2O3) and magnetite (Fe3O4) among the range of iron oxides, great candidates for many significant applications.17–22 One of the important advantages of MNPs as catalysts is their easy recovery using a magnetic field. The two intrinsic properties of paramagnetism and the insolubility of nanomagnets facilitate the separation of such magnetic catalysts from reaction mixtures. In particular, silica-modified MNPs have received widespread attention due to their various important advantages, such as ease and simplicity of synthesis, functionalization, and easy separation from the reaction medium employing a simple magnet, thermal stability, and low toxicity.
One of the best ways to synthesize new heterocyclic compounds from available and simple raw materials is through multicomponent reactions (MCRs), which have been widely employed in the preparation of biologically active molecules, natural products, and organic matter.23,24 The development of one-pot MCRs is one of the best methods for the synthesis of desired products due to certain salient features of this method, such as high selectivity, facile automation, simplicity, and high atomic economy.25–27 Other benefits of one-pot MCRs over multistep syntheses include their short reaction time and high chemical yields, which also leads to reduced waste generation, manpower, and energy consumption. Hence, MCRs have become a very powerful tool in combinational chemistry and drug discovery for the synthesis of biologically important heterocyclic compounds.
N-Heterocyclic scaffolds are very common in naturally occurring compounds. Among the heterocyclic compounds, the 2-amino-3,5-dicarbonitrile-6-thio-pyridines ring system has attracted much attention due to its widespread occurrence in numerous bioactive natural products. Such compounds have enjoyed wide applications in diverse pharmacological activities, such as anti-bacterial,28 anti-prion,29 anticancer agents,30 and anti-hepatitis B virus.31 (Scheme 1).
Furthermore, they can act as potassium channel openers for the treatment of urinary incontinence. In addition, some of these compounds were discovered to be candidates for the development of new drugs for the treatment of cancer, Parkinson's disease, epilepsy, hypoxia/ischemia, kidney disease, and asthma.32 For example, there are some naturally occurring compounds I, II, and III that bear a pyridine group as a structural subunit;33–35 while compounds IV and V are commercially available drug molecules that have the pyridine skeleton (Scheme 2).36
Scheme 2 (a) Some naturally occurring compounds (I, II, and III) bearing pyridine groups and (b) some drug molecules with a pyridine skeleton (IV and V). |
Diverse approaches have been developed for preparing pyridine derivatives with multistep methods.37–39 Recently, the reaction between aldehyde, malononitrile, and thiol via a one-pot pseudo-four component reaction was evaluated as an efficient and simple method for the preparation of 2-amino-3,5-dicarbonitrile-6-thio-pyridines in the presence of a catalyst. In recent years, more effective catalysts for the preparation of pyridine derivatives have been reported, such as ZnCl2,40 NAP-MgO,41 iodoxybenzoic acid (IBX),42 DBU,43 KF/alumina,44 scandium(III) triflate,45 [bmIm]OH,46 DABCO,47 WEB,48 Fe2O3@Fe3O4@Co3O4,49 NaNH2,50 Cu(I) nanoparticles,51 Zn(II) and Cd(II) MOFs,52 microporous molecular sieves 4A,53 and phosphotungstic acid/CTAB.54 Nevertheless, some previously reported procedures remain valuable, albeit many of these procedures have significant drawbacks, such as low yields, tedious workup procedures, high temperatures, long reaction times, and the use of phosphine ligands and copper salts. For example, Evodkimov et al. developed a new method for the synthesis of pyridines derivatives using DABCO or Et3N as a base catalyst. One of the limitations of that method was that the corresponding products were obtained in yields of only 20–48%.47 Sridhar et al. reported a new ZnCl2 as a Lewis acid catalyst for the synthesis of pyridines derivatives under exotic reaction conditions (conventional heating or microwave heating). The ZnCl2 catalyst gave the products in 45–73% and 46–67% yields, respectively.40 It was further observed that some methods used expensive or toxic catalysts and reagents.43 Therefore, the development of an environmentally friendly and green route for the synthesis of pyridine derivatives is still a desirable goal.
Continuing our endeavors for preparing and designing new magnetic nanoparticles as nanocatalysts,55–58 herein, for the first time, we report the synthesis of NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 as a heterogeneous nanocatalyst. This catalyst could be successfully used in the one-pot pseudo-four component synthesis of substituted pyridine derivatives (Scheme 3).
Scheme 3 Activity of the NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 catalyst for the synthesis of substituted pyridine derivatives. |
Fig. 1 FT-IR spectra of (a) Fe3O4, (b) SiO2@Fe3O4, (c) APTES@SiO2@Fe3O4, (d) TCT/APTES@SiO2@Fe3O4, (e) picolylamine/TCT/APTES@SiO2@Fe3O4, and (f) NiII-picolylamine/TCT/APTES@SiO2@Fe3O4. |
EDX spectroscopy is one of the most significant elemental analyses to chemically characterize samples with high accuracy. The EDX spectrum clearly showed the presence of O, Fe, Si, C, N, and Ni elements in the catalyst with mass percentages of 36.0, 39.0, 19.71, 2.83, 1.45, and 1.01, respectively, indicating the catalyst had been successfully prepared. The Ni loading onto the NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 catalyst was confirmed by the peak presence of the Ni element in the EDX spectrum (Fig. 2). Moreover, ICP was used to confirm the exact amount of Ni loading of the catalyst. The ICP results clearly showed the presence of Ni (1.77%) in the catalyst.
Further, the presence of O, Fe, Si, C, N, and Ni elements in the NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 catalyst was confirmed by another elemental analysis, namely elemental mapping analysis (Fig. 3). Fig. 3 clearly indicates the presence of O, Fe, Si, C, N, and Ni elements in the catalyst with a suitable dispersity. Moreover, the results from the elemental mapping analysis comfirmed the EDX analysis.
Fig. 3 Elemental mapping analysis of NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 catalyst: (a) C, (b) N, (c) O, (d) Si, (e) Fe, and (f) Ni. |
XPS was performed from 0 to 1200 eV for the NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 catalyst to obtain the oxidation of the Ni state together with elemental analysis (Fig. 4). Fig. 4a–f illustrate the spectra for (a) Fe 2p, (b) C 1s, (c) N 1s, (d) Ni 2p, (e) Si 2p, and (f) O 1s elements in NiII-picolylamine/TCT/APTES@SiO2@Fe3O4. Fig. 4a demonstrates the high-resolution spectra of Fe 2p with two peaks at 712.6 and 726.4 eV, attributed to 2p3/2 and 2p1/2, respectively.58 The peaks at binding energies of 283.8 and 286.4 eV were attributed to C–C and C–O, respectively (Fig. 4b).58 Fig. 4c depicts the high-resolution spectrum of N 1s with two peaks at 401.2 and 404.6 eV related to C–N and CN, respectively.58 The XPS spectrum for Ni showed doublet peaks at 855.8 and 864.9 eV related to Ni 2p3/2 and Ni 2p1/2, respectively (Fig. 4d).59,60 These peaks well confirmed the oxidation state of Ni(II) in the catalyst. The Si 2p appear at ≈102.7 eV and 104.8 eV, respectively related to the Si–C and Si–O, respectively (Fig. 4e).58 Finally, Fig. 4f reveals the O atoms contained in as-prepared nanocatalyst. O 1s at 530.2 eV, 534.8 eV and 535.2 eV corresponding to Fe–O, Si–O and O–H, respectively.58
Fig. 4 XPS spectra of NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 catalyst: (a) Fe 2p, (b) C 1s, (c) N 1s, (d) Ni 2p, (e) Si 2p, and (f) O 1s. |
The crystalline structure of NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 was investigated by powder X-ray diffraction (PXRD) (Fig. 5). Fig. 5 displays strong and sharp peaks at 2θ = 30.45°, 35.44°, 35.68°, 43.24°, 53.69°, 57.14°, and 62.86° belonging to NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 which was well consistent with the Fe3O4 nanoparticle crystal structure. Besides, the obtained result demonstrated that the designed catalyst had been successfully synthesized and there was no damage to the crystalline structure of Fe3O4, and the crystallographic faces were in good agreement with the XRD pattern of cubic Fe3O4 NP (JCPDS 88-0866).57,58 The crystallite average size and interplaner distance of the catalyst were determined to be 28.7 nm and 0.208 nm from the XRD spectra according to FWHM calculations using the Scherrer formula and Bragg equation, respectively (Table S1, ESI†).
The magnetic properties of Fe3O4 and NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 were studied at ambient temperature by VSM analysis (Fig. 6). Fig. 6a clearly indicates that these nanoparticles of (a) Fe3O4 and (b) NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 had a superparamagnetic nature. The magnetic saturation value for pure Fe3O4 was observed at 58.74 emu g−1. After coating various organic compounds on the surface of pure Fe3O4 to synthesize the NiII-picolylamine/TCT/APTES@SiO2@Fe3O4, the magnetic saturation value was observed at 37.02 emu g−1. Nevertheless, this magnetic saturation value was sufficient for the catalyst's easy separation from the reaction mixture by employing a magnetic bar (Fig. 6b).
Fig. 6 (a) VSM analysis, (b) image of the magnetic separation and redispersion of NiII-picolylamine/TCT/APTES@SiO2@Fe3O4. |
The thermal stability of Fe3O4, SiO2@Fe3O4, and NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 were investigated through TGA analysis in the temperature range of 25–600 °C (Fig. 7). Both the TGA diagram of Fe3O4 and silica-coated SiO2 indicated a small weight loss below 100 °C, probably due to the removal of surface hydroxyl groups and physically adsorbed and trapped water molecules. For NiII-picolylamine/TCT/APTES@SiO2@Fe3O4, there were two mass reductions: (i) below 100 °C, probably due to the evaporation of surface hydroxyl groups and physically adsorbed and trapped water molecules in the structure of the catalyst and (ii) between 180–600 °C, which was the main weight loss and probably due to the thermal decomposition of the functional groups and organic compounds supported on the catalyst surface. This thermal decomposition indicated the coating of various organic compounds on the iron oxide support. Moreover, the complete decomposition of the as-prepared nanocatalyst happened at about 500 °C.
The structural features, morphologies, and particle sizes of Fe3O4 and NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 were investigated by FE-SEM, TEM, and size-distributions images (Fig. 8). The FE-SEM images indicated that Fe3O4 and NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 had been synthesized with spherical shapes with the particle sizes in the 20–33 nm and of 31–42 nm ranges, respectively (Fig. 8a and b). In another study, the uniformity and morphology of Fe3O4 and NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 were further investigated by TEM (Fig. 8c and d). Fig. 8d clearly indicates that the particles had been synthesized with a uniform size and core/shell. Further, the size distribution of the NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 was measured by TEM to be over 28 nm, which suitably matched with the results obtained from the XRD analysis (Fig. 8e).
The isotherm type, total pore volume, mean pore diameter, and surface area of NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 were determined by BET analysis (Fig. 9). As can be seen from Fig. 9a, the structure of the as-prepared nanocatalyst could be ascribed to the type IV isotherm with a distinct hysteresis loop indicating mesoporous structure.57,58 Fig. 9b shows that the as-prepared nanocatalyst sat in the P/P0 range of 0.006–0.007 with a strong inflection, which may have been because the capillary condensation of N2 was reflected in the mesopores. Further, the corresponding structural parameters of the as-prepared nanocatalyst included a total pore volume, mean pore diameter, and surface area of 0.189 cm3 g−1, 22.31 nm, and 53.61 m2 g−1, respectively.
Fig. 9 (a) BET plot of the nitrogen adsorption–desorption and (b) pore-size distribution curve of the NiII-picolylamine/TCT/APTES@SiO2@Fe3O4. |
Entry | Catalyst (g) | Solvent (3 mL) | Temperature (°C) | Time (min) | Yieldb (%) |
---|---|---|---|---|---|
a Reaction conditions: malononitrile (2 mmol), thiophenol (1 mmol), benzaldehyde (1 mmol), catalyst, and solvent (3 mL).b Isolated yield. | |||||
1 | None | Solvent-free | Room temperature | 24 (h) | Trace |
2 | None | Solvent-free | 80 | 24 (h) | Trace |
3 | 0.005 | Solvent-free | 80 | 130 | 43 |
4 | 0.01 | Solvent-free | 80 | 90 | 62 |
5 | 0.03 | Solvent-free | 80 | 60 | 75 |
6 | 0.05 | Solvent-free | 80 | 30 | 95 |
7 | 0.07 | Solvent-free | 80 | 35 | 95 |
8 | 0.09 | Solvent-free | 80 | 40 | 93 |
9 | 0.1 | Solvent-free | 80 | 40 | 91 |
10 | 0.05 | MeOH | Reflux | 120 | 45 |
11 | 0.05 | H2O | Reflux | 90 | 78 |
12 | 0.05 | CH2Cl2 | Reflux | 180 | 51 |
13 | 0.05 | CHCl3 | Reflux | 180 | 44 |
14 | 0.05 | H2O:EtOH (1:1) | 80 | 60 | 83 |
15 | 0.05 | EtOH | Reflux | 60 | 92 |
16 | 0.05 | EG | 80 | 20 | 96 |
In the next study, we investigated the one-pot pseudo-four component reaction of malononitrile (2 mmol), thiophenol (1 mmol), and benzaldehyde (1 mmol), using various catalysts, such as Fe3O4, SiO2@Fe3O4, APTES@SiO2@Fe3O4, TCT/APTES@SiO2@Fe3O4, picolylamine/TCT/APTES@SiO2@Fe3O4, and NiCl2·6H2O for the synthesis of 4a under solvent-free conditions at 80 °C, and the results are listed in Table 2. As can be seen, it was found that NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 catalyze the reaction with high yield and short reaction time than the other catalysts.
Entry | Catalyst | Time (h) | Yieldb (%) |
---|---|---|---|
a Reaction conditions: malononitrile (2 mmol), thiophenol (1 mmol), benzaldehyde (1 mmol), and catalyst (0.05 g) under solvent-free conditions at 80 °C.b Isolated yield. | |||
1 | Fe3O4 | 4 | 41 |
2 | SiO2@Fe3O4 | 3 | 55 |
3 | APTES@SiO2@Fe3O4 | 1 | 71 |
4 | TCT/APTES@SiO2@Fe3O4 | 24 | 10 |
5 | Picolylamine/TCT/APTES@SiO2@Fe3O4 | 6 | 25 |
6 | NiCl2·6H2O | 6 | 45 |
7 | NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 | 30 min | 95 |
Knowing the optimized conditions (using 0.05 g of the as-prepared nanocatalyst under solvent-free conditions or EG at 80 °C), a range of aromatic aldehydes containing electron-withdrawing and electron-donating groups were reacted with malononitrile and thiophenol, and afforded the corresponding products (4a–l), and good results were obtained (Table 3). In method A, NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 as a high-performance heterogeneous nanocatalyst was used for the reaction of malononitrile, thiophenol, and various aldehydes under solvent-free conditions at 80 °C and gave satisfactory yields (91–97%) after 10–55 min. Moreover, NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 as a high-performance heterogeneous nanocatalyst showed excellent activity for the reaction of malononitrile, thiophenol, and various aldehydes in EG at 80 °C. The corresponding products were furnished in short reaction times (8–45 min) and in excellent yields (89–97%). Generally, NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 displayed remarkable efficiency for the one-pot synthesis of substituted pyridine derivatives in both methods A and B.
Entry | Product | Method Aa | Method Bb | Melting point (°C) |
---|---|---|---|---|
% Yieldc (reaction time) | % Yieldc (reaction time) | |||
a Reaction conditions: Malononitrile (2 mmol), thiophenol (1 mmol), various aldehydes (1 mmol), and NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 (0.05 g) under solvent-free conditions at 80 °C.b Reaction conditions: Malononitrile (2 mmol), thiophenol (1 mmol), various aldehydes (1 mmol), and NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 (0.05 g) in EG (3 mL) at 80 °C.c Isolated yield. | ||||
1 | 95 (30 min) | 96 (20 min) | 217–219 (ref. 51) | |
2 | 97 (15 min) | 96 (10 min) | 285–287 (ref. 51) | |
3 | 94 (35 min) | 93 (30 min) | 220–222 (ref. 51) | |
4 | 96 (10 min) | 97 (8 min) | 271–273 (ref. 51) | |
5 | 93 (50 min) | 95 (40 min) | 209–211 (ref. 51) | |
6 | 91 (30 min) | 92 (25 min) | 224–226 (ref. 53) | |
8 | 92 (50 min) | 91 (40 min) | 238–240 (ref. 51) | |
9 | 95 (35 min) | 94 (30 min) | 256–258 (ref. 51) | |
10 | 94 (55 min) | 89 (45 min) | 314–316 (ref. 51) | |
11 | 96 (30 min) | 92 (25 min) | 183–185 (ref. 51) | |
12 | 96 (20 min) | 96 (15 min) | 217–219 (ref. 51) | |
13 | 93 (20 min) | 93 (10 min) | — |
Due to environmental concerns, the catalyst reusability was tested (Fig. 10). For this, we carried out the reaction between 4-nitrobenzaldehyde (1 mmol), malononitrile (2 mmol), and thiophenol (1 mmol) in the presence of NiII-picolylamine/TCT/APTES@SiO2@Fe3O4 in EG at 80 °C. After the reaction completion, the catalyst could be separated easily from the reaction mixture by using a simple magnet. Then, the recovered catalyst was washed with EtOH (3 × 10 mL) and dried at room temperature for reused in the next run. This catalyst was reused up to eight times with only a slight decrease in activity, whereby the yields from the reaction smoothly decreased from 97% to 87% during 25–33 min (Fig. 10a). This slight decrease in the catalytic activity of the recovered nanocatalyst was due to the small leaching of Ni (1%) in the reaction after eight runs. The nanocatalyst was investigated by various analyses, including TEM and size distribution after each run for eight times. The TEM image was obtained for the recovered nanocatalyst after the eight runs (Fig. 10b), and indicated that the recovered nanocatalyst had a nearly uniform size and spherical shape, similar to the fresh catalyst. Likewise, the size distribution of the recovered nanocatalyst was measured by TEM, and was nearly same as the fresh catalyst (Fig. 10c).
Fig. 10 (a) Reusability results of NiII-picolylamine/TCT/APTES@SiO2@Fe3O4, (b) TEM image, and (c) size distribution of the reused catalyst. |
In another study, the effect of leaching of the active species in the reaction mixture on the stability of the as-prepared nanocatalyst was investigated. To address this possibility, we used ICP-AES analysis of the fresh state of the catalyst and also the recovered catalyst after eight runs. The reaction on the basis of ICP-AES analysis showed a leaching of Ni of 1 wt% after the eight runs. This result obtained could explain the negligible decrease in the yield of the product 4b with the increased recycling.
Large-scale synthesis of the reaction between 4-nitrobenzaldehyde (7 mmol, 1.05 g), malononitrile (14 mmol, 0.92 g), and thiophenol (7 mmol, 0.77 g) was performed in the presence of NiII-picolylamine/TCT/APTES@ SiO2@Fe3O4 in EG (21 mL) at 80 °C to give the corresponding 4b to prove the scalability and the prospective use of the current method (Scheme 5). The obtained results indicated that when performing the large-scale synthesis, the desired product was obtained with an 89% yield.
A plausible mechanism for obtaining the substituted pyridine derivatives in the reaction catalyzed by NiII-picolylamine/TCT/APTES@ SiO2@Fe3O4 is indicated in Scheme 6. The reaction was accelerated via the surface properties, including by Ni2+ acting as a Lewis acid. Initially, the Lewis acid site of the as-prepared nanocatalyst activated the aldehyde A and malonontrile B, and produced the 2-arylidene malononitrile C via a Knoevenagel condensation. In the next step, Ni2+ catalyzed the thiophenol D and malononitrile B and via a Michael addition to the intermediate C led to producing the intermediate F. Then, dihydropyridine G was formed via the tautomerization of the Michael product F, followed by yielding the desired product H.
Scheme 6 Plausible mechanism for the construction of substituted pyridine derivatives catalyzed by NiII-picolylamine/TCT/APTES@ SiO2@Fe3O4. |
Footnote |
† Electronic supplementary information (ESI) available: Copies of FT-IR, 1H NMR (250 MHz, CDCl3 or DMSO-d6) and 13C NMR (62.5 MHz, CDCl3 or DMSO-d6) spectra of the synthesized compounds, and XRD data for the NiII-picolylamine/TCT/APTES@SiO2@Fe3O4. See DOI: https://doi.org/10.1039/d3ra01826a |
This journal is © The Royal Society of Chemistry 2023 |