banner
News center
Express delivery

Magnetic BiFeO3 nanoparticles: a robust and efficient nanocatalyst for the green one-pot three-component synthesis of highly substituted 3,4-dihydropyrimidine-2(1H)-one/thione derivatives | Scientific Reports

Oct 14, 2024

Scientific Reports volume 14, Article number: 22201 (2024) Cite this article

383 Accesses

8 Altmetric

Metrics details

In this research, magnetic bismuth ferrite nanoparticles (BFO MNPs) were prepared through a convenient method and characterized. The structure and morphological characteristics of the prepared nanomaterial were confirmed through analyses using Fourier-transform infrared (FTIR) spectroscopy, field emission scanning electron microscopy (FESEM), energy-dispersive X-ray spectroscopy (EDS), elemental mapping, powder X-ray diffraction (XRD), N2 adsorption–desorption isotherms and vibrating sample magnetometry (VSM) techniques. The obtained magnetic BFO nanomaterial was investigated, as a heterogeneous Lewis acid, in three component synthesis of 3,4-dihydropyrimidin-2 (1H)-ones/thiones (DHPMs/DHPMTs). It was found that the BFO MNPs exhibit remarkable efficacy in the synthesis of various DHPMs as well as their thione analogues. It is noteworthy that this research features low catalyst loading, good to excellent yields, environmentally friendly conditions, short reaction time, simple and straightforward work-up, and the reusability of the catalyst, distinguishing it from other recently reported protocols. Additionally, the structure of the DHPMs/DHPMTs was confirmed through 1H NMR, FTIR, and melting point analyses. This environmentally-benign methodology demonstrates the potential of the catalyst for more sustainable and efficient practices in green chemistry.

The synthesis of nitrogen-containing 3,4-dihydropyrimidine-2(1H)-one/thione compounds (DHPMs/DHPMTs) commenced several years ago, driven by the recognition of the 3,4-dihydropyrimidinone skeleton in compounds with diverse biological properties, corrosion inhibition, etc1,2,3. Beyond their structural characteristics, nitrogen-containing heterocycles serve as essential building blocks in the synthesis of various biologically active compounds. These compounds exemplify the significant intersection of structure and function within the broader scope of medicinal chemistry and drug design, making them essential components4,5,6,7,8,9,10. While some initial progresses have been made, there is still a need for further exploration and expansion of their synthetic process, thus justifying continued investigation. Indeed, heterocyclic compounds occupy a crucial and indispensable domain within the field of chemistry2,3,4,5,6,7. Their fundamental role as key compounds has a significant impact on various scientific and applied fields8. Notably, a significant number of medicinal and naturally occurring compounds possessing biological activity contain inherent heterocyclic structures9. These compounds act as essential scaffolds in drug discovery and development, playing a crucial role in the pharmaceutical industry. The diverse and complex structures of heterocyclic compounds enhance their versatility, making them essential components in the synthesis of novel therapeutic agents10. Recognizing the pivotal role of heterocyclic compounds sheds light on their profound influence in advancing both theoretical understanding and practical applications in the field of chemistry11.

The versatility of DHPMs/DHPMTs stems from the ability to modify their chemical structure and subsequent enhancing of their pharmacological properties and therapeutic potential (Fig. 1). For example, (S)-L-771688 (Fig. 1A), a compound derived from the DHPMs/DHPMTs structure, has been studied as an α1a-adrenoceptor selective antagonist for the treatment of benign prostatic hyperplasia (BPH)5,6,7. In addition to BPH treatment, DHPMs/DHPMTs compounds have shown promise in other medical applications. In this context, nitractin (Fig. 1B) effectively combats the trachoma virus and displays antibacterial activity8. Idoxuridine (Fig. 1c) was also initially designed for cancer but transformed into a topical antiviral for herpes simplex keratitis. In addition, the compound (R)-SQ 32,926 (Fig. 1D), structurally similar to widely used calcium channel blockers, exhibits intriguing hypertensive effects9. On the other hand, emivirine (Fig. 1E) was developed as an HIV treatment, functioning as a non-nucleoside reverse transcriptase inhibitor10,11. Also, the discovery of the anticancer properties of 5-fluorouracil (Fig. 1F) led to the incorporation of fluorinated groups in medicinal chemistry. This compound disrupts DNA synthesis by irreversibly binding to the thymidylate synthase enzyme due to its structural similarity to uracil. Finally, monastrol (Fig. 1G) and its derivatives including enastron (Fig. 1H) and piperastrol (Fig. 1I) are notable DHPMs/DHPMTs compounds12. They exhibit promise for cancer chemotherapy by inhibiting kinesin-5, a key protein involved in mitosis regulation. Overall, DHPMs/DHPMTs compounds have demonstrated their versatility and therapeutic potential in various medical applications, from BPH treatment to antivirals, hypertension management, and cancer chemotherapy. Therefore, their unique chemical structure and pharmacological properties continue to be studied and utilized in the development of new and effective medications.

Typical examples of biologically active 3,4-dihydropyrimidine-2(1H)-one/thione (DHPM/DHPMT) derivatives.

The inherent biological potential of DHPM/DHPMT derivatives has led to the development of numerous synthetic methods for their preparation. Among the diverse strategies employed, the most notably efficient, straightforward and adaptable method for preparing of diverse DHPM/DHPMT derivatives is the one-pot three-component Biginelli condensation involving aldehydes, 1,3-dicarbonyl compounds, and urea/thiourea. In the meantime, the pivotal role of various catalysts especially heterogeneous catalysts with nano dimensions in facilitating the synthesis of nitrogen containing heterocyclic compounds through the Biginelli reaction is unequivocal13,14,15,16,17. These catalysts play a crucial role in enhancing the efficiency and selectivity of the synthesis process, underscoring their undeniable importance in the field18. Thus far, a diverse array of homogeneous or heterogeneous catalysts, including 1-dodecyl-3-(3-sulfopropyl)-imidazolium hydrogen sulfate [C3SO3HDoim]HSO420, l-proline nitrate21, 1-sulfopyridinium chloride [pyridine-SO3H]Cl22, zirconium (IV)-salophen perfluorooctanesulfonate23, triethylammonium acetate ionic liquid [Et3NH][CH3COO]24, MNPs-IL-HSO425, [bmim(SO3H) (OTf)]26, (BMI)BF4, (BMI)PF6 and (BMI)NTf227, coconut husk ash twisted graphene28, 1-butyl-3-carboxymethyl-benzotriazolium trifluoroacetate [C2O2BBTA][TFA]29, (3-(2-carboxy benzoyl)-1-methyl-1H-imidazol-3-ium chloride [Cbmim]Cl30, COF-IM-SO3H31, Fe3O4@Nb2O532, 1-butyl-1,3-thiazolidine-2-thione p-toluene sulfate [Btto][p-TS]33, p-sulfonic acid calix[4]arene (CX4)34, preyssler heteropoly acid H14NaP5W29MoO110 over SiO2 (PASi)35, boric acid [B(OH)3] supported on Fe3O4@MCM-4136, Fe3O4@SiO2@GP/Picolylamine-Cu(II)37, Bi(NO3)3, ZnO@SBA-1538, ZnO@SBA-15 (Si/Al = 7)39, isocyanurate-based periodic mesoporous organosilica40, iron oxide41, l-asparagine–EDTA–amide silica-coated MNPs42, silicasulfuric acid6, sulfamic acid pyromellitic diamide-functionalized MCM-4143, NFS-PRS9, zwitterionic MSI and TRIP44, [P4-VP]-Fe3O4-HSO4 ionic liquid45, nano-y-Fe2O3-SO3H46, MgFe2O4/cellulose/SO3H11, Magnetic Boron Nitride47, Zn coordination polymer48, perovskite photocatalyst under visible-light49, photoexcited Na2 Eosin Y50, 1,3,5-tris(2-hydroxyethyl)isocyanurate functionalized graphene oxide51, have been employed in the synthesis of DHPM/DHPMT derivatives.

The synthesis of DHPM/DHPMT derivatives faces challenges such as the use of toxic solvents, expensive catalysts, and tedious work-up processes. The unique characteristics of DHPM/DHPMT derivatives and the ongoing challenges in their synthesis have encouraged us for an innovative and efficient approach. In continuation of our interest in the synthesis of organic heterocycles according to the principles of green chemistry in the presence of MNPs40,42, 43, 51,52,53,54,55,56,57,58,59,60,61,62,63,64, we have investigated a novel method that demonstrates the catalytic capabilities of magnetic bismuth ferrite nanoparticles (BFO MNPs) to address present challenges and improves the entire synthetic process (Fig. 2).

One-pot three-component synthesis of highly substituted 3,4-dihydropyrimidine-2(1H)-one/thione derivatives catalyzed by the magnetic BiFeO3 nanoparticles (BFO MNPs) under green conditions.

All used chemicals and reagents including ethyl acetoacetate, urea or thiourea, and aldehydes were purchased from Merck or Sigma Aldrich. Additionally, pure distilled H2O and EtOH 96% were used as solvents. A UV lamp emitting light at a wavelength of 254 nm was utilized in conducting the thin-layer chromatography (TLC) experiments. The identification of the BFO MNPs as well as products was achieved on a Shimadzu FTIR 8400S spectrometer by using KBr disks. Furthermore, 1H NMR spectra were recorded using a Bruker Avance 500 in DMSO-d6 solvent at ambient temperature. The melting points were measured using a 9100 Electrothermal apparatus and are uncorrected. The reported yields are calculated based on the isolated products obtained after the purification process.

Preparation of the BFO MNPs was carried out through the solid-state thermal decomposition method65,66. Initially, a 100 ml round-bottom flask was charged with glycine (8.0 mmol), Fe(NO3)3 (4.0 mmol), Bi(NO3)3 (4.0 mmol), and deionized water (40.0 ml) and the obtained mixture was heated at 120 °C for one h. Subsequently, HNO3 (63%, d = 1.4 g ml−1, 2.80 ml) was added dropwise to the reaction mixture. The reaction mixture was subjected to heating at 120 °C, resulting in the formation of a brown solid. Finally, the brown solid was put in a furnace for one h at 350 °C and further one h at 550 °C to prepare bismuth ferrite nanoparticles (BFO MNPs).

In a 5 ml flask, aldehyde (1a–s, 1.0 mmol), ethyl acetoacetate (2, 1.0 mmol), urea/thiourea (3a–b, 1.10 mmol) and BFO MNPs (312.82 g mol−1, 4.0 mg, 1.28 mol%) were added to a mixture of H2O:EtOH (3.0 ml, v:v, 3:1). The reaction mixture was heated under reflux conditions for the appropriate time as indicated in Table 2. The reaction progress was monitored using thin-layer chromatography (TLC). After completion of the reaction, EtOAc (2 × 2 ml) was added and stirred for 10 min. The BFO MNPs were separated from the obtained mixture by applying an external magnet and collected for next runs by keeping them in an oven at 70 °C for 2 h. The filtrate was separated by using a separatory funnel. The EtOAc phase was kept in a 10 ml beaker at ambient temperature to afford recrystallized DHPMs/DHPMTs products with high purities.

Ethyl 6-methyl-4-(3-nitrophenyl)-3,4-dihydropyrimidine-2(1H)-one-5-carboxylate; White solid; M.P: 220–221 °C; 1H NMR (500 MHz, DMSO-d6) δ (ppm): 1.10 (t, J = 7.3 Hz, 3H, CH3), 2.28 (s, 3H, CH3), 4.00 (q, J = 7.3 Hz, 2H, CH2), 5.31 (d, J = 3.0 Hz, 1H, CH), 7.66 (t, J = 8.0 Hz, 1H, Ar–H), 7.70 (d, J = 7.5 Hz, 1H, Ar–H), 7.90 (brs, 1H, NH), 8.09 (s, 1H, Ar–H), 8.14 (d, J = 8.0 Hz, 1H, Ar–H), 9.37 (s, 1H, NH).

Ethyl 4-(3,4-dichlorophenyl)-6-methyl-3,4-dihydropyrimidine-2(1H)-one-5-carboxylate; White solid; M.P: 228–230 °C; 1H NMR (500 MHz, DMSO-d6) δ (ppm): 1.01 (t, J = 7.0 Hz, 3H, CH3), 2.30 (s, 3H, CH3), 3.90 (q, J = 7.0 Hz, 2H, CH2), 5.60 (s, 1H, CH), 7.32 (d, J = 8.3 Hz, 2H, Ar–H), 7.42 (d, J = 8.3 Hz, 2H, Ar–H), 7.57 (s, 1H, Ar–H), 7.57 (s, 1H, NH), 9.34 (s, 1H, NH).

Ethyl 4-(3,4-dimethoxyphenyl)-6-methyl-3,4-dihydropyrimidine-2(1H)-one-5-carboxylate; White solid; M.P: 173–175 °C; 1H NMR (500 MHz, DMSO-d6) δ (ppm): 1.12 (t, J = 6.8 Hz, 3H, CH3), 2.25 (s, 3H, CH3), 3.77 (s, 6H, 2 OCH3), 4.00 (q, J = 6.8 Hz, 2H, CH2), 5.11 (s, 1H, CH), 6.73 (d, J = 8.0 Hz, 1H, Ar–H), 6.85 (s, 1H, Ar–H), 6.90 (d, J = 8.0 Hz, 1H, Ar–H), 7.67 (s, 1H, NH), 9.15 (s, 1H, NH).

Ethyl 6-methyl-4-(thiophen-2-yl)-3,4-dihydropyrimidine-2(1H)-one-5-carboxylate; White solid; M.P: 214–216 °C; 1H NMR (500 MHz, DMSO-d6) δ (ppm): 1.17 (t, J = 7.0 Hz, 3H, CH3), 2.23 (s, 3H, CH3), 4.07 (q, J = 7.0 Hz, 2H, CH2), 5.42 (d, J = 3.0 Hz, 1H, CH), 6.89 (m, 1H, Het Ar–H), 6.94 (t, J = 4.0 Hz, 1H, Het Ar–H), 7.33 (d, J = 5.0 Hz, 1H, Ar–H), 7.90 (s, 1H, NH), 9.31 (s, 1H, NH).

Ethyl 6-methyl-4-phenyl-3,4-dihydropyrimidine-2(1H)-2-thione-5-carboxylate; White solid; M.P: 211–213 °C; 1H NMR (500 MHz, DMSO-d6) δ (ppm): 1.10 (t, J = 7.0 Hz, 3H, CH3), 2.30 (s, 3H, CH3), 4.00 (q, J = 7.0 Hz, 2H, CH2), 5.19 (s, H, CH), 7.23 (d, J = 7.0 Hz, 2H, Ar–H), 7.28 (t, J = 7.5 Hz, 1H, Ar–H), 7.35 (t, J = 8.0 Hz, 2H, Ar–H), 9.65 (s, 1H, NH), 10.33 (s, 1H, NH).

Characterization of nanoparticles typically involves various techniques to understand their size, shape, composition, surface properties, and other relevant characteristics. To confirm the structure and surface characteristics of the prepared magnetic BFO MNPs nanoparticles, several analyses were employed. These analyses include Fourier transform infrared spectroscopy (FTIR) for identifying functional groups through vibrational modes, field-emission scanning electron microscopy (FESEM) for detailed imaging to reveal information about size and morphology, energy-dispersive X-ray spectroscopy (EDS) for determining the elemental composition of the nanoparticles, X-ray powder diffraction (XRD) for determining the crystallinity degree of structure and providing insights into composition and arrangement, Brunauer–Emmett–Teller (BET) for assessing specific surface area and porosity, and vibrating sample magnetometry (VSM) technique for measuring magnetic properties including magnetic moment, magnetic susceptibility, and coercivity.

The FTIR spectrum of BFO MNPs nanoparticles prepared utilizing glycine, as a chelating agent, demonstrates apparent characteristics spanning the spectral range of 400–4000 cm⁻1 (Fig. 3). The obvious broad peak positioned at 3300–3500 cm⁻1 indicates the stretching vibrations of the hydroxyl (OH) groups located on the surface of the obtained nanomaterial. Additionally, the presence of peaks at 550 and 450 cm⁻1 is ascribed to the stretching vibrations associated with the iron-oxygen (Fe‒O) and bismuth-oxygen (Bi‒O) bonds stretching vibrations, respectively. These findings in the FTIR spectrum provide valuable insights into the structural and chemical composition of the prepared BiFeO₃ nanoparticles67.

FTIR spectrum of the BFO MNPs.

FESEM images of BiFeO₃ nanoparticles (at the scales of 50, 5 and 3 μm) indicate a spongy network with attached frog-like units, which having high porosity and microporous dimensions. These pores originate from the swift release of gases generated during nitrate combustion. Furthermore, upon analyzing of the images at the scales less than 3 μm, including 1 μm as well as 700 and 500 nm, it becomes apparent that the morphology of BFO MNPs consists of nanolayers arranged in close nearby, with micro-sized cavities formed between the layers (Fig. 4).

FESEM images of BiFeO3 nanoparticles at (A) 50 μm, (B) 5 nm, (C) 3 μm and (D) 1 μm, (E) 700 nm, and (F) 500 nm scales.

As shown in the Fig. 5, the elemental mapping analysis of the BFO MNPs catalyst clearly confirms the existence of Fe, O, and Bi elements in the composition of the nanomaterial. In addition, the mapping elemental analysis demonstrates a uniform distribution of desired elements. Furthermore, the identification of Bi, Fe, and O elements in the EDS analysis serves as additional evidence supporting the successful preparation of BFO MNPs (Fig. 6).

EDS elemental mapping analysis of the BFO MNPs.

Elemental analysis by energy-dispersive X-ray spectroscopy (EDS) of the BFO MNPs.

The X-ray diffraction (XRD) analysis of the bismuth ferrite nanostructure is illustrated in Fig. 7. The XRD pattern of the BFO MNPs exhibits characteristic peaks at 2θ of 22.2, 27.8, 31.8, 39.2, 51.1, 56.1, and 86.7°, offering detailed information about the crystalline structure of BFO MNPs. By comparing the findings of the XRD analysis with recent reports, it becomes evident that the crystal structure of BFO MNPs has been successfully established (JCPDS card number 01-073-0548). Also, employing Scherer's equation for particle size calculation at 2θ = 22.2 and a full width at half maximum (FWHM) of 0.448 shows a determined particle size of 18.9 nm68.

XRD pattern of the BFO MNPs.

The magnetization curve for the magnetic BFO MNPs is presented in Fig. 8, offering a comprehensive visualization of the magnetic properties of the prepared nanoparticles. A noteworthy ferromagnetic response is detected in the BFO MNPs, exhibiting a saturation magnetization of 4.60 emu.g−1. This observed behavior aligns cohesively with the findings reported in earlier studies on BiFeO3 nanoparticles, particularly when the average diameter size is maintained below 62.0 nm. This consistency in the magnetic properties underscores the reliability and reproducibility of the observed ferromagnetic characteristics, contributing to the growing body of knowledge surrounding magnetic nanoparticles, specifically those composed of bismuth ferrite69,70.

Hysteresis diagram of the magnetic bismuth ferrite nanoparticles (BFO MNPs).

The nitrogen adsorption–desorption isotherm of the BFO MNPs was studied to determine their surface area (Fig. S14). The specific BET surface area for the prepared BFO MNPs sample was found to be 24.81 m2.g−1. The proper surface to volume ration of the BFO MNPs, contributing to a substantial surface area, serves as a significant indicator of the appropriate characteristics of this particular surface. Furthermore, Barrett-Joyner-Halenda (BJH) pore size was mainly found to be at the range of 8.0 and 12.0 nm (Fig. S15).

To study the synthesis of DHPM/DHPMT derivatives in the presence of BFO MNPs, a model reaction comprising of benzaldehyde (1a), ethyl acetoacetate (2), and urea (3a) was chosen. Various influencing parameters, including catalyst and solvent type, catalyst loading, and temperature were systematically studied. The results of these experiments have been summarized in Table 1. Initially, the progress of the model reaction was examined under conditions without a catalyst, both at room temperature and under reflux conditions, using water as the solvent. The results showed that only trace amounts of the desired product, ethyl 6-methyl-4-(phenyl)-3,4-dihydropyrimidine-2(1H)-one-5-carboxylate (4a), were obtained even after 6 h (Table 1, entries 1,2). In the next experiments, the catalytic performance of different nanoparticles, including Fe3O4, MnFe2O4, CuFe2O4, ZnFe2O4, NiFe2O4, and BiFeO3 (BFO MNPs), was investigated in the model reaction at room temperature (Table 1, entries 3–8). The obtained findings indicated the superior performance of the BFO MNPs compared to other magnetic nanoparticles in the model reaction (Table 1, entry 8). After selecting of the BFO MNPs as the optimal catalyst, the efficiency of various solvents such as MeOH, EtOH, EtOAc and CH3CN as well as mixtures of H2O:MeOH (3:1) and H2O: EtOH (3:1) on the yield of the model reaction was evaluated (Table 1, entries 9–14). The obtained results revealed that the mixture of H2O: EtOH (3:1), as a solvent, exhibited the most significant impact, leading to a product yield of 67% (Table 1, entry 14). This observation can be explained by a combination of factors: the improved dispersibility of the BFO MNPs in the H2O: EtOH medium, the solubility of the substrates in this solvent, and the fact that the resulting DHPMs with higher molecular weights do not dissolve in this mixture. Increasing the catalyst loading to improve the model reaction, particularly in terms of yield and reaction time, was investigated in the next step. It was found that increasing the catalyst loading from 2.0 to 4.0 mg enhances the model reaction yield from 67 to 84%, accompanied by a notable reduction in reaction time from 45 to 20 min (Table 1, entry 15). Finally, the model reaction was examined under the reflux conditions. By increasing the reaction temperature from 25 °C to reflux conditions, higher yield of the desired product (93%) was obtained (Table 1, entry 16). Therefore, the optimized conditions were considered to be 4.0 mg loading of BFO MNPs in a mixture of H2O: EtOH (3.0 ml, 3:1) under reflux conditions.

To demonstrate the efficiency and general scope of the optimized conditions, the conditions was applied to the synthesis of various derivatives of DHPMs/DHPMTs 4a–y. This was achieved through a one-pot three-component reaction involving different aryl/heteroaryl aldehydes 1a–s, ethyl acetoacetate (2), and urea/ thiourea (3a–b) components. The obtained results have been summarized in Table 2. Interestingly, all studied aldehydes 1, even sensitive aldehydes to polymerization under acidic conditions such as thiophen-2-carbaldehyde (1q) and furfural (1r), survived very well under the optimized conditions to afford corresponding DHPMs/DHPMTs 4a–y in high to excellent yields.

According to the data presented in Table 2, aromatic aldehydes with electron-withdrawing groups (‒NO2, ‒Cl, ‒Br and ‒CHO) show faster formation of the desired products 4b–i and 4s–u compared to the benzaldehyde as well as aromatic aldehydes containing electron-donating groups (–OMe, –OH, –Me) to afford 4a, 4n–p and 4v–w, respectively. This can be attributed to the higher carbonyl activity of aldehydes with electron-withdrawing groups, which facilitate the reaction and leads to accelerated product formation. In this context, electron-rich heterocyclic aromatic aldehydes including thiophen-2-carbaldehyde (1q) and furfural (1r) exhibit prolonged reaction times compared to benzaldehyde (1a). The observed behavior can be assigned to their higher electron density and lower carbonyl activity in these compounds compared to aromatic carbocyclic aldehydes. Indeed, the increased electron density in these five-membered heterocyclic aldehydes hinders the attack of both enol form of ethyl acetoacetate (2) and urea/ thiourea (3a-b), as electrophile, on their carbonyl groups. Consequently, slower product formation and lower yields would be observed. These trends in reaction rates and the obtained yields is in consistence with the observed reaction mechanism, which suggests that the attack of an electrophile on the carbonyl group is the rate determining step.

The proposed mechanism for the one-pot multi-component synthesis of DHPM/DHPMT derivatives in the presence of BiFeO3 magnetic nanoparticles has been depicted in Fig. 9. In the first step of the mechanism, the carbonyl group of the aromatic aldehyde 1 is activated by BiFeO3 nanoparticles. The bismuth as well as Fe cations on the surface of the nanoparticles act as active Lewis acid centers and coordinate with the carbonyl group of aldehydes 1, enhancing their reactivity. Then, the nitrogen from urea or thiourea (3a–b) attacks, as a nucleophile, onto the activated carbonyl group by the BFO MNPs leading to the formation of hemiaminal intermediate (Int. 1). Then, this intermediate produces the imine intermediate (Int. 2) by omitting a water molecule. The next step involves the nucleophilic addition of the enol form of ethyl acetoacetate (2’), with proper concentration produced by the BFO MNPs, to the imine intermediate. This Michael addition leads to the formation of intermediate (Int. 3). Finally, the desired DHPMs/DHPMTs products 4a–y are formed through intramolecular condensation catalyzed by the BFO MNPs. This condensation involves the reaction between the remaining amine group of urea or thiourea and the keto group of ethyl acetoacetate moiety, resulting in the removal of H2O and the formation of the final products 4a–y.

Proposed mechanism for the synthesis of 3,4-dihydropyrimidine-2(1H)-one/thione derivatives in the presence of magnetic BFO MNPs nanocatalyst.

The recyclability and reusability of catalysts are important aspects in the field of green chemistry. Heterogeneous catalysts, despite their lower efficiency compared to homogeneous catalysts, are preferred due to their ability to be easily separated and reused. To verify the heterogeneity of the BFO MNPs nanocatalyst in this study, hot filtration test was conducted. After 5 min from the start of the reaction, the magnetic catalyst was separated from the reaction mixture by using an external magnet. The reaction mixture was then stirred for an additional 30 min at room temperature without the catalyst. The results showed that only 51% of the desired product was obtained after this duration, suggesting that the Bi and Fe ions present in the structure of BFO MNPs nanoparticles were absent in the reaction medium and confirms its heterogeneity. This finding is significant as it demonstrates the potential for the BFO MNPs nanocatalyst to be separated and reused, enhancing its practical applications and aligning with the principles of green chemistry.

The recyclability of the BFO MNPs was assessed to determine their potential for multiple uses. The BFO MNPs were isolated from the reaction mixture using an external magnet and then washed with EtOAc and EtOH to remove any impurities. After drying at 70 °C for 2 h, the BFO MNPs were used again in the same model reaction. As depicted in Fig. 10, repeated runs of the model reaction by using the recycled BFO MNPs catalyst showed high conversion rates, indicating their appropriate catalytic activity. After five replications of the model reaction using the recycled BFO MNPs catalyst, an average yield of 86% was obtained. These results highlight the stability and recyclability of the BFO MNPs, as a proper heterogeneous nanocatalyst, for the Biginelli three-component reaction. The fact that the recycled catalyst consistently maintained high conversion rates suggests that it can be reused multiple times without significant loss of its activity. Also, XRD analysis of the recycled catalyst, after its washing and drying, was performed. The result has been shown in Fig. S18. Interestingly, the result demonstrated that the catalyst structure remained almost intact through multiple reuse cycles. Overall, these findings demonstrate the promising potential of BFO MNPs nanoparticles as a stable and efficient heterogeneous nanocatalyst under optimal reaction conditions.

The results of recyclability and reusability of the BFO MNPs nanoparticles in the model reaction.

Ultimately, the effectiveness of the BFO MNPs catalyst was assessed in comparison to previously reported protocols. As outlined in Table 3, the BFO MNPs nanocatalyst developed in this study exhibits notable efficiency in the rapid synthesis of DHPMs/DHPMTs. Additionally, an individual characteristic of this catalyst is its capability to facilitate the formation of DHPMs/DHPMTs products under environmentally friendly conditions and green chemistry principles.

In brief, the efficacy of magnetic bismuth ferrite magnetic nanoparticles (BFO MNPs), as a heterogeneous Lewis acidic catalyst, for the synthesis of diverse derivatives of pharmacologically-active 3,4-dihydropyrimidin-2 (1H)-ones/thiones (DHPMs/DHPMTs) was explored. The catalytic performance of the BFO MNPs demonstrated notable dependence to the solvent and temperature. The corresponding DHPMs/DHPMTs were obtained in a mixture of H2O:EtOH, as a green solvent, through the one-pot Biginelli three-component reaction strategy. The presented method offers several noteworthy advantages including high to excellent yields of the desired products, low catalyst loading and short reaction times. Moreover, this approach eliminates the need for toxic organic solvents and ensures a simplified purification of the desired products, facilitates a straightforward workup process, and allows for the recyclability and reusability of the BFO MNPs catalyst for at least five runs without significant loss of its catalytic activity. This environmentally-benign methodology demonstrates the potential of the catalyst for more sustainable and efficient practices in green chemistry.

All data generated or analyzed during this study are included in this published article [and its supplementary information files].

Matos, L. H. S., Masson, F. T., Simeoni, L. A. & Homem-de-Mello, M. Biological activity of dihydropyrimidinone (DHPM) derivatives: A systematic review. Eur. J. Med. Chem. 143, 1779–1789 (2018).

Article CAS PubMed Google Scholar

Srivastava, J. K., Pillai, G. G., Bhat, H. R., Verma, A. & Singh, U. P. Design and discovery of novel monastrol-1,3,5-triazines as potent anti-breast cancer agent via attenuating Epidermal Growth Factor Receptor tyrosine kinase. Sci. Rep. 7, 5851. https://doi.org/10.1038/s41598-017-05934-5 (2017).

Article ADS CAS PubMed Central PubMed Google Scholar

Gupta, S. K. et al. Electrochemical, surface morphological and computational evaluation on carbohydrazide Schiff bases as corrosion inhibitor for mild steel in acidic medium. Sci. Rep. 13, 15108. https://doi.org/10.1038/s41598-023-41975-9 (2023).

Article ADS CAS PubMed Central PubMed Google Scholar

Kazemi, M. & Mohammadi, M. Magnetically recoverable catalysts: Catalysis in synthesis of polyhydroquinolines. Appl. Organomet. Chem. 34, e5400 (2020).

Article CAS Google Scholar

Li, N. et al. Synthesis of (S)-L-771688. Synfacts 2010, 0268–0268 (2010).

Article Google Scholar

Moghanian, H., Fard, M. A. B., Mobinikhaledi, A. & Ahadi, N. Bis (p-sulfoanilino) triazine-functionalized silica-coated magnetite nanoparticles as an efficient and magnetically reusable nano-catalyst for Biginelli-type reaction. Res. Chem. Intermed. 44, 4083–4101 (2018).

Article CAS Google Scholar

Chang, R. S. et al. In vitro studies on L-771,688 (SNAP 6383), a new potent and selective α1A-adrenoceptor antagonist. Eur. J. Pharmacol. 409, 301–312 (2000).

Article CAS PubMed Google Scholar

Javidi, J., Esmaeilpour, M. & Dodeji, F. N. Immobilization of phosphomolybdic acid nanoparticles on imidazole functionalized Fe 3 O 4@ SiO 2: A novel and reusable nanocatalyst for one-pot synthesis of Biginelli-type 3, 4-dihydro-pyrimidine-2-(1 H)-ones/thiones under solvent-free conditions. RSC Adv. 5, 308–315 (2015).

Article ADS CAS Google Scholar

Eshghi, H. et al. Preyssler heteropolyacid supported on silica coated NiFe2O4 nanoparticles for the catalytic synthesis of bis (dihydropyrimidinone) benzene and 3, 4-dihydropyrimidin-2 (1H)-ones. Chin. J. Catal. 36, 299–307 (2015).

Article CAS Google Scholar

Hopkins, A. L. et al. Design of MKC-442 (emivirine) analogues with improved activity against drug-resistant HIV mutants. J. Med. Chem. 42, 4500–4505 (1999).

Article CAS PubMed Google Scholar

Maleki, A., Jafari, A. A. & Yousefi, S. MgFe 2 O 4/cellulose/SO 3 H nanocomposite: A new biopolymer-based nanocatalyst for one-pot multicomponent syntheses of polysubstituted tetrahydropyridines and dihydropyrimidinones. J. Iran. Chem. Soc. 14, 1801–1813 (2017).

Article CAS Google Scholar

Salehi, P., Ali Zolfigol, M., Shirini, F. & Baghbanzadeh, M. Silica sulfuric acid and silica chloride as efficient reagents for organic reactions. Curr. Org. Chem. 10, 2171–2189 (2006).

Nagarajaiah, H., Mukhopadhyay, A. & Moorthy, J. N. Biginelli reaction: An overview. Tetrahedron Lett. 57, 5135–5149 (2016).

Article CAS Google Scholar

Esmaili, S., Khazaei, A., Ghorbani-Choghamarani, A. & Mohammadi, M. Silica sulfuric acid coated on SnFe 2 O 4 MNPs: synthesis, characterization and catalytic applications in the synthesis of polyhydroquinolines. RSC Adv. 12, 14397–14410 (2022).

Article ADS CAS PubMed Central PubMed Google Scholar

Ghobakhloo, F., Azarifar, D., Mohammadi, M., Keypour, H. & Zeynali, H. Copper (II) Schiff-base complex modified UiO-66-NH2 (Zr) metal–organic framework catalysts for Knoevenagel condensation–Michael addition–cyclization reactions. Inorg. Chem. 61, 4825–4841 (2022).

Article CAS PubMed Google Scholar

Saha, A., Pal, A., Mukherjee, D., Pal, S. C. & Das, M. C. Two-dimensional Cu (II)-MOF with Lewis Acid–base BIFUNCTIONAL SITES FOR CHEMICAL FIXATION of CO2 and bioactive 1, 4-DHP synthesis via Hantzsch Condensation. Inorg. Chem. (2024).

Ramish, S. M., Ghorbani-Choghamarani, A. & Mohammadi, M. Microporous hierarchically Zn-MOF as an efficient catalyst for the Hantzsch synthesis of polyhydroquinolines. Sci. Rep. 12, 1479 (2022).

Article ADS CAS PubMed Central PubMed Google Scholar

Kundu, S. K. in AIP Conference Proceedings (AIP Publishing).

Heravi, M. M., Asadi, S. & Lashkariani, B. M. Recent progress in asymmetric Biginelli reaction. Mol. Divers. 17, 389–407 (2013).

Article CAS PubMed Google Scholar

Zhou, Z.-L., Wang, P.-C. & Lu, M. Bronsted acidic ionic liquid [C3SO3HDoim] HSO4 catalyzed one-pot three-component Biginelli-type reaction: An efficient and solvent-free synthesis of pyrimidinone derivatives and its mechanistic study. Chin. Chem. Lett. 27, 226–230 (2016).

Article CAS Google Scholar

Bahekar, S. P., Sarode, P. B., Wadekar, M. P. & Chandak, H. S. Simple and efficient synthesis of 3, 4-dihydropyrimidin-2 (1H)-thiones utilizing L-proline nitrate as a proficient, recyclable and eco-friendly catalyst. J. Saudi Chem. Soc. 21, 415–419 (2017).

Article CAS Google Scholar

Velpula, R., Banothu, J., Gali, R., Deshineni, R. & Bavantula, R. 1-Sulfopyridinium chloride: Green and expeditious ionic liquid for the one-pot synthesis of fused 3, 4-dihydropyrimidin-2 (1H)-ones and thiones under solvent-free conditions. Chin. Chem. Lett. 26, 309–312 (2015).

Article CAS Google Scholar

Li, N. et al. Air-stable zirconium (IV)-salophen perfluorooctanesulfonate as a highly efficient and reusable catalyst for the synthesis of 3,4-dihydropyrimidin-2-(1H)-ones/thiones under solvent-free conditions. Appl. Organometal. Chem. 34, e5454. https://doi.org/10.1002/aoc.5454 (2020).

Article CAS Google Scholar

Attri, P. et al. Triethylammonium acetate ionic liquid assisted one-pot synthesis of dihydropyrimidinones and evaluation of their antioxidant and antibacterial activities. Arab. J. Chem. 10, 206–214 (2017).

Article CAS Google Scholar

Safari, J. & Zarnegar, Z. Brønsted acidic ionic liquid based magnetic nanoparticles: A new promoter for the Biginelli synthesis of 3, 4-dihydropyrimidin-2 (1 H)-ones/thiones. New J. Chem. 38, 358–365 (2014).

Article CAS Google Scholar

Savanur, H. M., Kalkhambkar, R. G., Aridoss, G. & Laali, K. K. [bmim (SO3H)][OTf]/[bmim][X] and Zn (NTf2) 2/[bmim][X](X= PF6 and BF4); efficient catalytic systems for the synthesis of tetrahydropyrimidin-ones (-thiones) via the Biginelli reaction. Tetrahedron Lett. 57, 3029–3035 (2016).

Article CAS Google Scholar

Ramos, L. M. et al. The Biginelli reaction with an imidazolium–tagged recyclable iron catalyst: Kinetics, mechanism, and antitumoral activity. Chem. A Eur. J. 19, 4156–4168 (2013).

Narayanan, D. P. et al. A green approach for the synthesis of coconut husk ash–twisted graphene nanocomposites: Novel catalysts for solvent-free biginelli reaction. ChemistrySelect 4, 4785–4796 (2019).

Article CAS Google Scholar

Liu, Z., Ma, R., Cao, D. & Liu, C. New efficient synthesis of 3, 4-dihydropyrimidin-2 (1 H)-ones catalyzed by benzotriazolium-based ionic liquids under solvent-free conditions. Molecules 21, 462 (2016).

Article PubMed Central PubMed Google Scholar

Heidarizadeh, F., Nezhad, E. R. & Sajjadifar, S. Novel acidic ionic liquid as a catalyst and solvent for green synthesis of dihydropyrimidine derivatives. Sci. Iran. 20, 561–565 (2013).

CAS Google Scholar

Yao, B.-J., Wu, W.-X., Ding, L.-G. & Dong, Y.-B. Sulfonic acid and ionic liquid functionalized covalent organic framework for efficient catalysis of the biginelli reaction. J. Org. Chem. 86, 3024–3032. https://doi.org/10.1021/acs.joc.0c02423 (2021).

Article CAS PubMed Google Scholar

Lima, C. G. et al. Highly efficient and magnetically recoverable niobium nanocatalyst for the multicomponent Biginelli reaction. ChemCatChem 6, 3455–3463 (2014).

Article CAS Google Scholar

Zhang, Y., Wang, B., Zhang, X., Huang, J. & Liu, C. An efficient synthesis of 3, 4-dihydropyrimidin-2 (1 H)-ones and thiones catalyzed by a novel brønsted acidic ionic liquid under solvent-free conditions. Molecules 20, 3811–3820 (2015).

Article CAS PubMed Central PubMed Google Scholar

Zacchi, C. H. C., Vieira, S. S., Ardisson, J. D., Araujo, M. H. & de Fatima, A. Synthesis of environmentally friendly, magnetic acid-type calix [4] arene catalyst for obtaining Biginelli adducts. J. Saudi Chem. Soc. 23, 1060–1069 (2019).

Article CAS Google Scholar

Portilla-Zuñiga, O. M. et al. Synthesis of Biginelli adducts using a Preyssler heteropolyacid in silica matrix from biomass building block. Sustain. Chem. Pharm. 10, 50–55 (2018).

Article Google Scholar

Ramazani, Z., Elhamifar, D., Norouzi, M. & Mirbagheri, R. Magnetic mesoporous MCM-41 supported boric acid: A novel, efficient and ecofriendly nanocomposite. Compos. Part B Eng. 164, 10–17 (2019).

Article CAS Google Scholar

Rezayati, S. et al. Magnetic silica-coated picolylamine copper complex [Fe3O4@SiO2@GP/Picolylamine-Cu(II)]-catalyzed biginelli annulation reaction. Inorg. Chem. 61, 992–1010. https://doi.org/10.1021/acs.inorgchem.1c03042 (2022).

Article CAS PubMed Google Scholar

Bhuyan, D., Saikia, M. & Saikia, L. ZnO nanoparticles embedded in SBA-15 as an efficient heterogeneous catalyst for the synthesis of dihydropyrimidinones via Biginelli condensation reaction. Microp. Mesop. Mater. 256, 39–48 (2018).

Article CAS Google Scholar

Mahato, B. N. & Krithiga, T. Mesoporous ZnO/AlSBA-15 (7) nanocomposite as an efficient catalyst for synthesis of 3, 4-dihydropyrimidin-2 (1H)-one via Biginelli reaction and their biological activity study. Bull. Chem. React. Eng. Catal. 14, 634–645 (2019).

Article CAS Google Scholar

Yaghoubi, A. & Dekamin, M. G. Green and Facile Synthesis of 4H‐Pyran Scaffold Catalyzed by Pure Nano‐Ordered Periodic Mesoporous Organosilica with Isocyanurate Framework (PMO‐ICS). ChemistrySelect2, 9236-9243 (2017).

Article CAS Google Scholar

Mondal, J., Sen, T. & Bhaumik, A. Fe 3 O 4@ mesoporous SBA-15: a robust and magnetically recoverable catalyst for one-pot synthesis of 3, 4-dihydropyrimidin-2 (1 H)-ones via the Biginelli reaction. Dalton Trans. 41, 6173–6181 (2012).

Article CAS PubMed Google Scholar

Rostami, N., Dekamin, M. G., Valiey, E. & FaniMoghadam, H. l-Asparagine–EDTA–amide silica-coated MNPs: a highly efficient and nano-ordered multifunctional core–shell organocatalyst for green synthesis of 3,4-dihydropyrimidin-2(1H)-one compounds. RSC Adv. 12, 21742–21759. https://doi.org/10.1039/D2RA02935A (2022).

Article ADS CAS PubMed Central PubMed Google Scholar

Valiey, E., Dekamin, M. G. & Alirezvani, Z. Sulfamic acid pyromellitic diamide-functionalized MCM-41 as a multifunctional hybrid catalyst for melting-assisted solvent-free synthesis of bioactive 3,4-dihydropyrimidin-2-(1H)-ones. Sci. Rep. 11, 11199. https://doi.org/10.1038/s41598-021-89572-y (2021).

Article ADS CAS PubMed Central PubMed Google Scholar

Alvim, H. G. et al. Combined role of the asymmetric counteranion-directed catalysis (acdc) and ionic liquid effect for the enantioselective biginelli multicomponent reaction. J. Org. Chem. 83, 12143–12153 (2018).

Article CAS PubMed Google Scholar

Zarchi, M. K. & Hamidi, Z. Synthesis of [P4-VP]-Fe3O4 supported Brønsted acid ionic liquid and its application as a highly efficient heterogeneous and reusable nanocatalyst in the Biginelli reaction under solvent-free conditions. React Kinet. Mech. Catal. 125, 1023–1037 (2018).

Article Google Scholar

Kolvari, E., Koukabi, N. & Armandpour, O. A simple and efficient synthesis of 3, 4-dihydropyrimidin-2-(1H)-ones via Biginelli reaction catalyzed by nanomagnetic-supported sulfonic acid. Tetrahedron 70, 1383–1386 (2014).

Article CAS Google Scholar

Rana, P. et al. Magnetic boron nitride nanosheets decorated with cobalt nanoparticles as catalyst for the synthesis of 3,4-dihydropyrimidin-2(1H)-ones/thiones. ACS Appl. Nano Mater. 5, 4875–4886. https://doi.org/10.1021/acsanm.1c04438 (2022).

Article CAS Google Scholar

Silva, G. C. et al. The Biginelli reaction under batch and continuous flow conditions: Catalysis, mechanism and antitumoral activity. RSC Adv. 5, 48506–48515 (2015).

Article ADS CAS Google Scholar

Mohamadpour, F. Recyclable photocatalyst perovskite as a single-electron redox mediator for visible-light-driven photocatalysis gram-scale synthesis of 3,4-dihydropyrimidin-2-(1H)-ones/thiones in air atmosphere. Sci. Rep. 13, 10262. https://doi.org/10.1038/s41598-023-37526-x (2023).

Article ADS CAS PubMed Central PubMed Google Scholar

Mohamadpour, F. Visible-light-induced radical condensation cyclization to synthesize 3,4-Dihydropyrimidin-2-(1H)-ones/thiones Using Photoexcited Na2 Eosin Y as a direct hydrogen atom transfer (HAT) catalyst. ACS Omega 7, 8429–8436. https://doi.org/10.1021/acsomega.1c05808 (2022).

Article CAS PubMed Central PubMed Google Scholar

Dekamin, M. G., Mehdipoor, F. & Yaghoubi, A. 1,3,5-Tris(2-hydroxyethyl)isocyanurate functionalized graphene oxide: A novel and efficient nanocatalyst for the one-pot synthesis of 3,4-dihydropyrimidin-2(1H)-ones. New J. Chem. 41, 6893–6901. https://doi.org/10.1039/C7NJ00632B (2017).

Article CAS Google Scholar

Dekamin, M. G. & Eslami, M. Highly efficient organocatalytic synthesis of diverse and densely functionalized 2-amino-3-cyano-4 H-pyrans under mechanochemical ball milling. Green Chem. 16, 4914–4921 (2014).

Article CAS Google Scholar

Dekamin, M. G., Eslami, M. & Maleki, A. Potassium phthalimide-N-oxyl: A novel, efficient, and simple organocatalyst for the one-pot three-component synthesis of various 2-amino-4H-chromene derivatives in water. Tetrahedron 69, 1074–1085 (2013).

Article CAS Google Scholar

Malihishoja, A., Dekamin, M. G. & Eslami, M. Magnetic polyborate nanoparticles as a green and efficient catalyst for one-pot four-component synthesis of highly substituted imidazole derivatives. RSC Adv. 13, 16584–16601 (2023).

Article ADS CAS PubMed Central PubMed Google Scholar

Matloubi Moghaddam, F., Eslami, M. & Hoda, G. Cysteic acid grafted to magnetic graphene oxide as a promising recoverable solid acid catalyst for the synthesis of diverse 4 H-chromene. Sci. Rep. 10, 20968 (2020).

Article CAS PubMed Central PubMed Google Scholar

Fattahi, B. & Dekamin, M. G. Fe3O4/SiO2 decorated trimesic acid-melamine nanocomposite: a reusable supramolecular organocatalyst for efficient multicomponent synthesis of imidazole derivatives. Sci. Rep. 13, 401. https://doi.org/10.1038/s41598-023-27408-7 (2023).

Article ADS CAS PubMed Central PubMed Google Scholar

Eslami, M., Dekamin, M. G. & Mahdavi, E. C (sp2)‒Hal activation for efficient Ullmann-type C‒O coupling in water by in situ reduction of supramolecular Cu (II) species stabilized on a modified GO nanosheets support. Surf. Interfaces 104363 (2024).

Moghaddam, F. M., Jarahiyan, A., Eslami, M. & Pourjavadi, A. A novel magnetic polyacrylonotrile-based palladium Core–Shell complex: A highly efficientcatalyst for Synthesis of Diaryl ethers. J. Organometal. Chem. 916, 121266 (2020).

Article CAS Google Scholar

Akbari, A., Dekamin, M. G., Yaghoubi, A. & Naimi-Jamal, M. R. Novel magnetic propylsulfonic acid-anchored isocyanurate-based periodic mesoporous organosilica (Iron oxide@PMO-ICS-PrSO3H) as a highly efficient and reusable nanoreactor for the sustainable synthesis of imidazopyrimidine derivatives. Sci. Rep. 10, 10646. https://doi.org/10.1038/s41598-020-67592-4 (2020).

Article ADS CAS PubMed Central PubMed Google Scholar

Ishani, M., Dekamin, M. G. & Alirezvani, Z. Superparamagnetic silica core-shell hybrid attached to graphene oxide as a promising recoverable catalyst for expeditious synthesis of TMS-protected cyanohydrins. J. Colloid Interface Sci. 521, 232–241. https://doi.org/10.1016/j.jcis.2018.02.060 (2018).

Article ADS CAS PubMed Google Scholar

Karami, S., Dekamin, M. G., Valiey, E. & Shakib, P. DABA MNPs: a new and efficient magnetic bifunctional nanocatalyst for the green synthesis of biologically active pyrano[2,3-c]pyrazole and benzylpyrazolyl coumarin derivatives. New J. Chem. 44, 13952–13961. https://doi.org/10.1039/D0NJ02666B (2020).

Article CAS Google Scholar

Shakib, P., Dekamin, M. G., Valiey, E., Karami, S. & Dohendou, M. Ultrasound-Promoted preparation and application of novel bifunctional core/shell Fe3O4@SiO2@PTS-APG as a robust catalyst in the expeditious synthesis of Hantzsch esters. Sci. Rep. 13, 8016. https://doi.org/10.1038/s41598-023-33990-7 (2023).

Article ADS CAS PubMed Central PubMed Google Scholar

Sourkouhi, R. P., Dekamin, M. G., Valiey, E. & Dohendou, M. Magnetic decorated 5-sulfosalicylic acid grafted to chitosan: A solid acid organocatalyst for green synthesis of quinazoline derivatives. Carbohydrate Polym. Technol. Appl. 7, 100420. https://doi.org/10.1016/j.carpta.2023.100420 (2024).

Article CAS Google Scholar

Dekamin, M. G., Varmira, K., Farahmand, M., Sagheb-Asl, S. & Karimi, Z. Organocatalytic, rapid and facile cyclotrimerization of isocyanates using tetrabutylammonium phthalimide-N-oxyl and tetraethylammonium 2-(carbamoyl)benzoate under solvent-free conditions. Catal. Commun. 12, 226–230 (2010).

Article CAS Google Scholar

Sharma, P., Diwan, P. & Pandey, O. Impact of environment on the kinetics involved in the solid-state synthesis of bismuth ferrite. Mater. Chem. Phys. 233, 171–179 (2019).

Article CAS Google Scholar

Luo, J. et al. Novel bismuth ferrite nanopowder prepared by polyethylene glycol-assisted two-step solid-state reaction: Synthesis and magneto-optical properties. Ceram. Int. 47, 3514–3519 (2021).

Article CAS Google Scholar

Shi, X., Quan, S., Yang, L., Liu, C. & Shi, F. Anchoring Co3O4 on BiFeO3: Achieving high photocatalytic reduction in Cr (VI) and low cobalt leaching. J. Mater. Sci. 54, 12424–12436 (2019).

Article ADS CAS Google Scholar

Hu, Y. et al. Synthesis of bismuth ferrite nanoparticles via a wet chemical route at low temperature. J. Nanomater. 2011, 1–6 (2011).

Article Google Scholar

Wang, X., Zhang, Y. G. & Wu, Z. Magnetic and optical properties of multiferroic bismuth ferrite nanoparticles by tartaric acid-assisted sol–gel strategy. Mater. Lett. 64, 486–488 (2010).

Article ADS CAS Google Scholar

Park, T.-J., Papaefthymiou, G. C., Viescas, A. J., Moodenbaugh, A. R. & Wong, S. S. Size-dependent magnetic properties of single-crystalline multiferroic BiFeO3 nanoparticles. Nano Lett. 7, 766–772 (2007).

Article ADS CAS PubMed Google Scholar

Zhao, J. et al. Microemulsion polymerization of cationic pyrroles bearing an imidazolum-ionic liquid moiety. J. Polym. Sci. Part A Polym. Chem. 47, 746–753 (2009).

Article ADS CAS Google Scholar

Alirezvani, Z., Dekamin, M. G., Davoodi, F. & Valiey, E. Melamine‐Functionalized Chitosan: A New Bio‐Based Reusable Bifunctional Organocatalyst for the Synthesis of Cyanocinnamonitrile Intermediates and Densely Functionalized Nicotinonitrile Derivatives. ChemistrySelect 3, 10450–10463. https://doi.org/10.1002/slct.201802010 (2018).

Article CAS Google Scholar

Chopda, L. V. & Dave, P. N. Fe (III)/bentonite as a heterogeneous catalyst for the synthesis of 3, 4-dihydropyrimidin-2-(1H)-ones. ChemistrySelect 5, 14161–14167 (2020).

Article CAS Google Scholar

Rahmatpour, A. & Donyapeyma, G. Titanium tetrachloride immobilized on cross-linked poly (N-vinyl-2-pyrrolidone) as a recyclable heterogeneous catalyst for one-pot three component synthesis of 3, 4-dihydropyrimidin-2 (1 H)-ones/thiones. Synth. Commun. 52, 678–693 (2022).

Article CAS Google Scholar

Hajipour, A. R., Khazdooz, L. & Zarei, A. Brønsted acidic ionic liquid–catalyzed one-pot synthesis of 3, 4-dihydropyrimidin-2 (1 H)-ones and thiones under solvent-free conditions. Synth. Commun. 41, 2200–2208 (2011).

Article CAS Google Scholar

Esmaeili, R., Kafi-Ahmadi, L. & Khademinia, S. A highly efficient one-pot multicomponent synthesis of 3, 4-dihydropyrimidin-2-(1H)-ones/thiones catalyzed by strontium pyroarsenate nano-plates. J. Mol. Struct. 1216, 128124 (2020).

Article CAS Google Scholar

Bais, J. et al. One pot synthesis of micromolar BACE-1 inhibitors based on the dihydropyrimidinone scaffold and their thia and imino analogues. Molecules 25, 4152 (2020).

Article CAS PubMed Central PubMed Google Scholar

Roy, D. K., Tamuli, K. J. & Bordoloi, M. Exploiting silver trifluoromethanesulfonate as efficient and reusable catalyst for the synthesis of dihydropyrimidine derivatives under different reaction environments. J. Heterocyclic Chem. 56, 3313–3323 (2019).

Article CAS Google Scholar

Murata, H., Ishitani, H. & Iwamoto, M. Synthesis of Biginelli dihydropyrimidinone derivatives with various substituents on aluminium-planted mesoporous silica catalyst. Org. Biomol. Chem. 8, 1202–1211 (2010).

Article CAS PubMed Google Scholar

Rode, N., Tantray, A., Shelar, A., Patil, R. & Terdale, S. Amino acid ionic liquid-catalyzed synthesis, anti-Leishmania activity, molecular docking, molecular dynamic simulation, and ADME study of 3, 4-dihydropyrimidin-2 (1 H)-(thio) ones. Synth. Commun. 52, 190–204 (2022).

Article CAS Google Scholar

Yang, J. et al. A vaccine targeting the RBD of the S protein of SARS-CoV-2 induces protective immunity. Nature 586, 572–577 (2020).

Article ADS CAS PubMed Google Scholar

Shaibuna, M., Kuniyil, M. J. K. & Sreekumar, K. Deep eutectic solvent assisted synthesis of dihydropyrimidinones/thiones via Biginelli reaction: Theoretical investigations on their electronic and global reactivity descriptors. N. J. Chem. 45, 20765–20775 (2021).

Article CAS Google Scholar

Moradi, L., Najafi, G. R. & Saeidiroshan, H. New method for preparation of MWCNT-SO3H as an efficient and reusable catalyst for the solvent-free synthesis of 3, 4-dihydropyrimidin-2 (1H)-ones/thiones. Iran. J. Catal. 5, 357–364 (2015).

CAS Google Scholar

Tayebee, R. et al. Phosphotungstic acid grafted zeolite imidazolate framework as an effective heterogeneous nanocatalyst for the one-pot solvent-free synthesis of 3, 4-dihydropyrimidinones. Appl. Organometal. Chem. 33, e4959 (2019).

Article Google Scholar

Moussa, S., Mehri, A. & Badraoui, B. Magnesium modified calcium hydroxyapatite: An efficient and recyclable catalyst for the one-pot Biginelli condensation. J. Mol. Struct. 1200, 127111 (2020).

Article Google Scholar

Kong, R. et al. Highly efficient synthesis of substituted 3, 4-dihydropyrimidin-2-(1 H)-ones (DHPMs) catalyzed by Hf (OTf) 4: Mechanistic insights into reaction pathways under metal Lewis acid catalysis and solvent-free conditions. Molecules 24, 364 (2019).

Article PubMed Central PubMed Google Scholar

Patel, H. A., Sawant, A. M., Rao, V. J., Patel, A. L. & Bedekar, A. V. Polyaniline supported FeCl 3: An effective heterogeneous catalyst for Biginelli reaction. Catal. Lett. 147, 2306–2312 (2017).

Article CAS Google Scholar

Sadjadi, S. & Koohestani, F. Composite of cross-linked chitosan beads and a cyclodextrin nanosponge: A metal-free catalyst for promoting ultrasonic-assisted chemical transformations in aqueous media. J. Phys. Chem. Solids 156, 110157 (2021).

Article CAS Google Scholar

Rostamnia, S. & Morsali, A. Basic isoreticular nanoporous metal–organic framework for Biginelli and Hantzsch coupling: IRMOF-3 as a green and recoverable heterogeneous catalyst in solvent-free conditions. RSC Adv. 4, 10514–10518 (2014).

Article ADS CAS Google Scholar

Pavithran, K. K. et al. An efficient synthesis of 3, 4-dihydro-2 (1H)-pyrimidinones from aldehyde bisulfite adducts using recyclable wang resin supported sulfonic acid catalyst. Lett. Org. Chem. 19, 463–470 (2022).

Article CAS Google Scholar

do Nascimento, L. G. et al. Niobium oxides as heterogeneous catalysts for Biginelli multicomponent reaction. J. Org. Chem. 85, 11170–11180 (2020).

Article CAS PubMed Google Scholar

Download references

The authors would like to express their gratitude to the Research Council of Iran University of Science and Technology, Tehran (IUST) for their financial support (grant no 160/23372). We would also like to acknowledge the support of the Iran Nanotechnology Initiative Council (INIC) and Iran National Science Foundation (INSF).

Pharmaceutical and Heterocyclic Compounds Research Laboratory, Department of Chemistry, Iran University of Science and Technology, Tehran, 16846-13114, Iran

Safa Hanifi & Mohammad G. Dekamin

Department of Chemistry, Behbahan Khatam Alanbia University of Technology, Behbahan, 63616-63973, Iran

Mohammad Eslami

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

Safa Hanifi: Investigation, Formal analysis, Writing-original draft; Mohammad G. Dekamin: Conceptualization, Formal analysis, Financial, Editing-Final draft; Mohammad Eslami: Conceptualization, Formal analysis, Investigation, Writing-original draft.

Correspondence to Mohammad G. Dekamin.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

Hanifi, S., Dekamin, M.G. & Eslami, M. Magnetic BiFeO3 nanoparticles: a robust and efficient nanocatalyst for the green one-pot three-component synthesis of highly substituted 3,4-dihydropyrimidine-2(1H)-one/thione derivatives. Sci Rep 14, 22201 (2024). https://doi.org/10.1038/s41598-024-72407-x

Download citation

Received: 06 March 2024

Accepted: 06 September 2024

Published: 27 September 2024

DOI: https://doi.org/10.1038/s41598-024-72407-x

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative