Skip to main content
Scientific Reports logoLink to Scientific Reports
. 2025 Nov 17;15:40184. doi: 10.1038/s41598-025-23849-4

Functionalized magnetic nanoparticles with deep eutectic solvents as efficient catalyst for the green synthesis of dihydroquinazolinones

Zohreh Valipour 1, Rahman Hosseinzadeh 1,, Yaghoub Sarrafi 1, Behrooz Maleki 1
PMCID: PMC12623933  PMID: 41249193

Abstract

In this study, deep eutectic solvents were attached to SiO2-coated magnetic nanoparticles through a linker. The resulting nanomagnet-immobilized arginine/choline chloride designated as MNP@Arg/ChCl was characterized using FT-IR, XRD, TGA, VSM, and SEM techniques. The catalyst showed an irregular spherical morphology, and the mean particle diameter was determined to be about 39.9 nm. TGA analysis also confirmed that the catalyst remained thermally stable up to approximately 225 °C. MNP@Arg/ChCl demonstrated excellent catalytic activity in the solvent-free synthesis of 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-one derivatives via a three-component reaction involving isatoic anhydride, amines, and aldehydes. The scope of this method was assessed using various aromatic aldehydes and aromatic amines. The 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-one derivatives were obtained in yields ranging from 85% to 93% within a short reaction time. Due to the magnetization properties of MNP@Arg/ChCl, the catalyst can be easily separated from the reaction mixture and reused several times without significant loss of catalytic activity. This method is particularly attractive in the context of green chemistry due to its high catalytic efficiency, thermal stability, reusability, ease of recovery, and environmentally friendly features.

Supplementary Information

The online version contains supplementary material available at 10.1038/s41598-025-23849-4.

Keywords: Natural deep eutectic solvent, Magnetic nanocatalyst, Three-component reactions, Quinazolinones

Subject terms: Catalysis, Green chemistry, Organic chemistry

Introduction

Over the past two decades, deep eutectic solvents (DESs) have attracted a lot of attention as catalysts and reaction media due to their unique properties, including tunable physicochemical properties, non-toxicity, negligible vapor pressure, simple preparation, nonflammability, high thermal stability, and biodegradability1,2. In general, DESs are straightforwardly prepared by heating and mixing a hydrogen bond acceptor (HBA) with a hydrogen-bond donor (HBD), facilitated by hydrogen bonding, halogen bonding, π-π interactions, or electrostatic interactions3,4. Choi, Paiva and co-workers described natural deep eutectic solvents (NADES) derived from natural origins such as alcohols, sugars, carboxylic acids, and amino acids5,6. NADESs are composed of components that are in our daily food, making them safe, sustainable, and cost-effective7. Arginine (Arg) is a naturally occurring, readily available, non-toxic, and biologically active amino acid. It consists of three parts: an alpha-amino group, a carboxyl group, and a guanidine moiety, the latter of which imparts its basic nature and facilitates its participation in various biochemical reactions. The N-H bonds of the guanidine moiety can easily act as hydrogen bond donors8.

While DESs have shown considerable promise as solvent and catalyst in organic transformations over the past decade914, their industrial applications remain limited due to the need for substantial quantities of DES, challenges with reusability, and associated high costs15. To overcome these challenges, DESs are being supported on solid substrates such as metal organic frameworks (MOFs)16, SBA-1517, graphene oxide18, and pectin19. Heterogeneous catalysis offers significant advantages over homogeneous catalysis in the synthesis of organic compounds20. One key benefit is the straightforward recovery and reuse of the catalyst. Moreover, heterogeneous catalysts are generally more robust and compatible with a broader range of reaction conditions compared to their homogeneous counterparts2123. However, the disadvantages of heterogeneous catalysts include their lower selectivity and catalytic activity, which result from poor dispersion and accessibility of heterogeneous catalysts. Recently, DES catalysts supported on magnetic nanoparticles have gained attention due to their high surface area, good stability, straightforward synthesis, ease of operation, and convenient separation using an external permanent magnet2429.

The development and synthesis of complex organic molecules through green and sustainable methods is a fundamental goal in organic chemistry30. Multicomponent reactions (MCR) have presented as an attractive technique that allows for the facile synthesis of complex molecules in a one-pot synthesis, without the need to isolate or purify intermediates31. Compared to traditional methodologies, MCRs offer several advantages, including shorter reaction times, energy efficiency, cost-effectiveness, and the capability to produce a diverse array of products32. As a result, MCR has profoundly impacted pharmaceutical and chemical synthesis.

2,3-Dihydroquinazolin-4(1 H)-ones are an important class of heterocyclic compounds known for their diverse range of biological activities and pharmacological properties, including, antitumor33, anti-inflammatory34, antibacterial35, antihistamine36, antidiabetic and antioxidant37, antimalarial38, and antifungal39. Due to the importance of these compounds, considerable efforts have been made to improve the synthesis of 2,3-dihydroquinazolin-4(1 H)-ones through a three-component reaction involving isatoic anhydride, aldehydes, amines, or ammonium salts, utilizing various catalysts, including glutamic acid40, carboxymethyl cellulose@MnFe2O441, Fe3O4@SiO2@Pr-PABA42, zinc (II) perfluorooctanoate43, CoFe2O4‑Spirulina‑SO3H44, CS-TDI-SSA-Fe3O445, MCM-41@serine@Cu(II)46, Fe3O4/TiCl2/cellulose47, and Co-aminobenzamid@Al-SBA-1548. Although the methods described offer considerable benefits, some of them suffer from one or more limitations, including laborious work-up procedures, difficulties in catalyst preparation, high temperatures, the use of toxic solvents and reagents, prolonged reaction times, and low yields.

Building on previous studies focused on the development of new heterogeneous DES catalysts1619,2429 and continuing our research into green and sustainable methods for synthesizing heterocyclic compounds4953, this study presents an environmentally friendly approach for immobilizing NADES on magnetic nanoparticles to synthesize 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-one derivatives, as outlined in Scheme 1. The MNP@Arg/ChCl was designed and synthesized by functionalizing the Fe3O4@SiO2 surface with 3-chloropropyltrimethoxysilane (CPTMS) and arginine/choline chloride. The applicability of the synthesized MNP@Arg/ChCl was assessed as a reusable heterogeneous catalyst in the three-component synthesis of 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-ones under mild and solvent-free conditions.

Scheme 1.

Scheme 1

Synthesis of 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-one derivatives by MNP@Arg/ChCl.

Results and discussion

Synthesis and characterization of MNPs@Arg/ChCl

MNPs@Arg/ChCl were successfully synthesized through hydrogen bonding interactions between arginine attached to silica-coated magnetite nanoparticles and choline chloride. As illustrated in Scheme 2, the process began with the synthesis of silica-coated nanomagnetite using Fe3O4 and tetraethoxysilane (TEOS). Subsequently, the core-shell nanoparticle system was functionalized with 3-chloropropyltrimethoxysilane (CPTMS). The subsequent reaction of arginine with the chloro-functionalized SiO2 led to the immobilization of arginine on the nanoparticle surface. Finally, choline chloride is added as a hydrogen bond acceptor. The resulting novel MNP@Arg/ChCl was characterized by FT-IR, EDX, element mapping images, FE-SEM, XRD, and TGA.

Scheme 2.

Scheme 2

Synthesis route of MNP@Arg/ChCl.

The Fourier-transform infrared spectroscopy (FT-IR) analysis was employed to confirm the structure by identifying the functional groups. The IR spectra of Fe₃O₄@SiO₂, Fe₃O₄@SiO₂@CPTMS, Fe₃O₄@SiO₂@CPTMS@Arg, and MNP@Arg/ChCl are collectively displayed in Fig. 1. The absorption band at 559 cm⁻¹ corresponds to the stretching vibration of the Fe–O bond in Fe₃O₄. Additionally, the Si–O–Si stretching vibrations at 1075 cm⁻¹ indicate the presence of a silica shell surrounding Fe₃O₄ (Fig. 1A). In the spectra presented in Fig. 1B, the Fe–O stretching vibrations are observed at 554 cm⁻¹, and Si–O–Si stretching vibrations at 1066 cm⁻¹. Also, the stretching vibrations at 2985 and 2921 cm⁻¹ pertain to C–H bonds, which result from the presence of the propyl group, thereby demonstrating the successful bonding of CPTMS to Fe₃O₄@SiO₂. The FT-IR spectrum of Fe3O4@SiO2@CPTMS@Arg (Fig. 1C) displays absorption bands in the regions of 3500 –3000 cm⁻¹, as well as at 2927 cm⁻¹ and 2856 cm⁻¹, which are attributed to the stretching vibrations of N-H bonds and O-H bonds and the stretching vibrations of C-H bonds in arginine. The asymmetric COO⁻ stretching vibrations were observed at 1570 cm⁻¹, and the symmetric COO⁻ stretching vibrations at 1414 cm⁻¹. Additionally, the bending vibrations of NH and the stretching vibrations of C = N are noted at 1680 cm⁻¹ and 1610 cm⁻¹, respectively, which are associated with the guanidine functional group54. The presence of characteristic peaks of arginine indicates the successful binding of Fe3O4@SiO2@CPTMS with L-arginine. Finally, the alkyl group vibration bands at 1472 cm⁻¹ and the C-N⁺(CH₃)₃ vibrations at 954 cm⁻¹, along with the broadening of the absorption bands in the range of 3600 –2600 cm⁻¹, confirm the successful interaction of choline chloride with Fe₃O₄@SiO₂@CPTMS@Arg (Fig. 1D).

Fig. 1.

Fig. 1

FT-IR spectra of (A) Fe₃O₄@SiO₂, (B) Fe₃O₄@SiO₂@CPTMS, (C) Fe₃O₄@SiO₂@CPTMS@Arg and (D) MNP@Arg/ChCl.

Energy dispersive X-ray spectroscopy (EDX) was utilized to confirm the formation of the MNP@Arg/ChCl nanocomposite. The peaks corresponding to Fe, Si, O, C, N, and Cl observed in the EDX spectrum (Fig. 2), along with the quantitative data regarding the elemental composition presented in Table 1, validate the successful synthesis of the MNP@Arg/ChCl nanocomposites. Furthermore, elemental mapping analysis demonstrated a uniform distribution of atoms throughout the structure of the nanocomposites (Fig. 3).

Fig. 2.

Fig. 2

EDX spectrum of MNP@Arg/ChCl.

Table 1.

Quantitative results of EDS analysis.

Elemental Fe Si O C N Cl
W% 13.17 4.72 35.01 29.68 16.27 1.16

Fig. 3.

Fig. 3

Elemental mapping analysis of MNP@Arg/ChCl.

To investigate the morphology and surface characteristics of the catalyst, scanning electron microscopy (SEM) images of MNP@Arg/ChCl were recorded (Fig. 4). The obtained images revealed that the MNP@Arg/ChCl catalyst exhibits irregular spherical shapes. Furthermore, the SEM micrographs indicate slight aggregation, which may be attributed to interactions between the layers of DES surrounding the nanoparticles. Additionally, the average particle size derived from the histogram indicates a core diameter of approximately 39.9 nm for MNP@Arg/ChCl.

Fig. 4.

Fig. 4

SEM images and particle size analysis of MNP@Arg/ChCl.

The crystallinity and phase analysis of magnetic Fe₃O₄ and MNP@Arg/ChCl were examined using X-ray diffraction (XRD). The peaks observed at 2θ = 11.04°, 18.32°, 19.16°, 19.48°, 23.08°, and 27.44° were attributed to L-arginine (Fig. 5), which matched the standard pattern (JCPDS-PDF, No. 00-030-1527)55. The XRD patterns of Fe₃O₄ magnetic nanoparticles reveal six prominent diffraction peaks at 2θ values of 30.3°, 35.6°, 43.3°, 53.7°, 57.3°, and 62.8°, corresponding to the (220), (311), (400), (422), (511), and (440) planes, respectively (Fig. 5). This pattern is consistent with the standard XRD data for the crystal structure of Fe₃O₄ (JCPDS card No. 19–0629)56. Furthermore, the presence of Fe₃O₄ diffraction peaks indicates that no phase changes occurred in the Fe₃O₄ nanoparticles after the preparation of MNP@Arg/ChCl. Also, using the Scherrer equation, the average crystallite size of the MNP@Arg/ChCl nanoparticles was determined to be 41.1 nm.

Fig. 5.

Fig. 5

XRD patterns of (a) Fe3O4, and (b) MNP@Arg/ChCl.

The thermal stability of the MNP@Arg/ChCl catalyst was evaluated using thermogravimetric analysis (TGA) over a temperature range of 20 to 800 °C under an argon atmosphere. The TGA curve shows a weight loss of about 8 wt% in the range of 20–180 °C, which is mainly due to moisture loss. This is followed by a weight loss of about 35 wt% in the temperature range of 280–700 °C, attributed to the thermal decomposition of organic components. This suggests the presence of substantial organic layers on the surface of the catalyst (Fig. 6).

Fig. 6.

Fig. 6

TGA curve of MNP@Arg/ChCl.

The vibrating sample magnetometer (VSM) analysis was utilized to investigate the magnetic properties of MNP@Arg/ChCl. Based on the magnetic curves shown in Fig. 7, the saturation magnetic values for Fe3O4 and MNP@Arg/ChCl were 67.13 and 20.21 emug− 1, respectively. The observed decrease in magnetic properties with increasing amounts of organic materials loaded on the Fe3O4 surface confirms the successful functionalization of the nanoparticles. Furthermore, the catalyst’s magnetization is sufficient to enable its efficient recovery from the reaction using an external magnet.

Fig. 7.

Fig. 7

Magnetization curves of Fe3O4 and MNP@Arg/ChCl.

Catalytic activity MNP@Arg/ChCl

We initiated our study on the catalytic efficiency of the synthesized MNP@Arg/ChCl in a three-component reaction for the synthesis of 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-ones. Accordingly, the reaction between isatoic anhydride 1 (0.50 mmol), aniline 2a (0.60 mmol), and 4-chlorobenzaldehyde 3a (0.50 mmol) was investigated as a model reaction.

The effects of various factors such as solvent, reaction temperature, catalyst amount, and type of catalyst, were investigated on the yield and rate of the reaction (Table 2). Initially, the reaction was carried out in the absence of catalyst, under solvent-free conditions, and at reflux in ethanol. After 2 h no product was observed (Table 2, entries 1 and 2). This confirmed that the presence of the catalyst is essential for this reaction. To investigate the effect of solvents, the model reaction was conducted under both solvent-free conditions and in the presence of various solvents, including ethanol, water, acetonitrile, and n-hexane, under reflux conditions (Table 2, entries 3–6). According to the results presented in Table 2, the reaction conducted under solvent-free conditions achieved a high yield in the presence of 30 mg of catalyst at 60 min (entry 7). In the next step, temperature was examined as a parameter to achieve better efficiency. Increasing the temperature to 100 °C enhanced the product yield to 92% and reduced the reaction time (Table 2, entry 8). Also, it should be noted that increasing the temperature to 120 °C did not result in a significant change in either reaction yield or time (Table 2, entry 9). Next, the model reaction was conducted using varying amounts of catalyst. As shown in Table 2, entry 10, reducing the catalyst amount from 30 mg to 15 mg resulted in a decrease in yield, whereas increasing it to 50 mg did not affect the reaction yield (Table 2, entry 11). Model reactions employing ChCl, MNP@Arg, and Fe3O4 individually were performed to evaluate the distinct contributions of each catalyst components (entries 12–14). The results indicated that the overall efficiency of the catalyst in the three-component reaction for synthesizing 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-ones is superior to that of its individual components (Table 2, entry 8). The synergistic effect arising from the interaction between ChCl and MNP@Arg, combined with the ionic nature of the deep eutectic solvent, enhanced the catalytic performance of the system. The use of SiO₂-coated magnetic nanoparticles (MNPs) in our catalyst design is justified not only by the ease and speed of catalyst recovery through magnetic separation, but also by the fact that the multifunctional catalyst MNP@Arg/ChCl delivers significantly higher yields than its individual components.

Table 2.

Optimization of the model reaction.

graphic file with name 41598_2025_23849_Figa_HTML.jpg
Entry Solvent Catalyst catalyst amount Temp. (oC) Time (min.) Yielda (%)
1 80 120
2 EtOH reflux 120
3 EtOH MNP@Arg/ChCl 30 reflux 120 72
4 H2O MNP@Arg/ChCl 30 reflux 120 50
5 CH3CN MNP@Arg/ChCl 30 reflux 120 70
6 Hexane MNP@Arg/ChCl 30 reflux 120 25
7 MNP@Arg/ChCl 30 80 60 78
8 MNP@Arg/ChCl 30 100 30 92
9 MNP@Arg/ChCl 30 120 30 93
10 MNP@Arg/ChCl 15 100 30 55
11 MNP@Arg/ChCl 50 100 30 93
12 Fe3O4 30 100 30 35
13 MNP@Arg 30 100 30 50
14 ChCl 30 100 30 27

Reaction conditions: Isatoic anhydride (0.50 mmol), aniline (0.60 mmol), and 4-chlorobenzaldehyde (0.50 mmol).

aIsolated yields.

Under the optimal reaction conditions (Table 2, entry 8), the scope and versatility of the three-component reaction involving isatoic anhydride, various substituted amines and aldehydes were investigated. As shown in Table 3, a wide variety of benzaldehydes containing different substituents, including electron-withdrawing groups (− Cl, −Br, −CN, and − NO2) and electron-donating groups (− OCH3, −CH3, and − N(CH3)2) successfully reacted. The results indicated that both electron-donating and electron-withdrawing groups at the para position of the benzaldehyde ring had no significant effect on the reaction time and gave products in good to excellent yields (Table 3, entries 1–7). In contrast, ortho-substituted aromatic aldehydes exhibited longer reaction times due to steric hindrance (Table 3, entries 8–10). Notably, the reaction with heteroaryls, such as thiophene and furan carbaldehydes, gave the desired products in a short reaction time and excellent yields (Table 3, entries 11 and 12). Next, we examined a variety of aromatic amine substrates. As shown in Table 3, anilines with electron-withdrawing (-Cl, -Br and -COOH) and electron-donating (-OCH3 and -CH3) afforded the desired products in high yields. However, the presence of electron-withdrawing groups led to a slight increase in the reaction times (Table 3, entries 14–18).

Table 3.

Preparation of 2,3-dihydroquinazolin-4(1 H)-one derivatives via a three-component reaction.

Entry R 1 R 2 Product Time (min) Yielda (%) M.P (˚C)
Found Reported
1 4-ClC6H4 C6H5 graphic file with name 41598_2025_23849_Figb_HTML.gif 30 92 223–225 220–22240
2 4-BrC6H4 C6H5 graphic file with name 41598_2025_23849_Figc_HTML.gif 30 91 220–222 221–22340
3 4-NO2C6H4 C6H5 graphic file with name 41598_2025_23849_Figd_HTML.gif 35 90 194–196 196–19840
4 4-CN C6H4 C6H5 graphic file with name 41598_2025_23849_Fige_HTML.gif 35 88 200–202 203–20457
5 4-MeOC6H4 C6H5 graphic file with name 41598_2025_23849_Figf_HTML.gif 35 93 201–204 203–20540
6 4-MeC6H4 C6H5 graphic file with name 41598_2025_23849_Figg_HTML.gif 30 91 212–214 215–21645
7 4-(Me)2NC6H4 C6H5 graphic file with name 41598_2025_23849_Figh_HTML.gif 30 89 187–190 184–18543
8 3-MeOC6H4 C6H5 graphic file with name 41598_2025_23849_Figi_HTML.gif 35 90 186–188 187–18957
9 2-ClC6H4 C6H5 graphic file with name 41598_2025_23849_Figj_HTML.gif 45 85 170–173 174–17640
10 2-MeOC6H4 C6H5 graphic file with name 41598_2025_23849_Figk_HTML.gif 45 80 147–149 144–14658
11 C4H3S C6H5 graphic file with name 41598_2025_23849_Figl_HTML.gif 35 93 203–206 198–20144
12 C4H3O C6H5 graphic file with name 41598_2025_23849_Figm_HTML.gif 40 90 139–141 135–13759
13 C6H5 C6H5 graphic file with name 41598_2025_23849_Fign_HTML.gif 30 93 199–203 202–20640
14 C6H5 4-ClC6H4 graphic file with name 41598_2025_23849_Figo_HTML.gif 45 89 207–209 210–21243
15 C6H5 4-BrC6H4 graphic file with name 41598_2025_23849_Figp_HTML.gif 40 91 230–232 228–22957
16 C6H5 4-HOOCC6H4 graphic file with name 41598_2025_23849_Figq_HTML.gif 60 90 237–240 238–24060
17 C6H5 4-MeC6H4 graphic file with name 41598_2025_23849_Figr_HTML.gif 30 91 194–197 196–19943
18 C6H5 4-MeOC6H4 graphic file with name 41598_2025_23849_Figs_HTML.gif 30 93 213–216 209–21143
19 4-ClC6H4 4-MeC6H4 graphic file with name 41598_2025_23849_Figt_HTML.gif 30 90 249–250 251–25345

Reaction conditions: Isatoic anhydride (0.5 mmol), amine (0.6 mmol), and aldehyde (0.5 mmol) in MNP@Arg/ChCl (30 mg) at 100 °C.

aIsolated yields.

Based on published studies19,61,62, a plausible mechanism for the synthesis of 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-one derivatives through one-pot reactions involving aldehydes, isatoic anhydride, and anilines in the presence of MNP@Arg/ChCl is proposed and illustrated in Scheme 3. MNP@Arg/ChCl, as a multifunctional catalyst, activates carbonyl, amine, and imine functional groups through extensive hydrogen bonding. Additionally, Fe3O4 nanoparticles act as Lewis acids, enhancing the electrophilicity of aldehydes and imines through coordination with the Fe centers. Moreover, the guanidine groups in arginine serve as basic sites, promoting the deprotonation steps and facilitating the reaction. Initially, the catalyst’s hydrogen-bond donating and accepting abilities, combined with the Lewis acidity of the Fe3O4 nanoparticles, facilitate the nucleophilic attack of the aromatic amine on the activated isatoic anhydride. This is followed by decarboxylation, leading to the formation of 2-amino-N-arylbenzamide (2). In the next step, the catalyst activates the reaction between the aldehyde and 2-amino-N-arylbenzamide (2) via hydrogen bonding and coordination with Fe3O4 nanoparticles, resulting in the formation of intermediate a. Then, the elimination of a water molecule produces the imine intermediate b, which is activated by the catalyst to undergo an intramolecular nucleophilic reaction forming intermediate c. Proton transfer facilitated by the multifunctional catalyst ultimately leads to the formation of the desired product (4).

Scheme 3.

Scheme 3

A plausible mechanism for the synthesis of 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-ones.

The recyclability and reusability of the magnetic catalyst in the model reaction were evaluated under optimized conditions. After the reaction was completed, the catalyst was easily recovered using an external magnet and washed several times with ethyl acetate. It was then dried in an oven at 80 °C. Based on our experimental findings, the recycled magnetic nanocatalyst can be utilized for five consecutive cycles in the model reaction with no substantial decline in catalytic efficiency (Fig. 8). The catalyst’s stability and recyclability are primarily attributed to the SiO2 shell covering the nano-Fe3O4, which protects it from oxidation during the cycling process. The comparison of the XRD patterns of the recycled catalyst and the fresh catalyst (Fig. 9A) indicates that no phase transformation occurred in the Fe3O4 nanoparticles during the recycling process. Additionally, the comparison of the FT-IR spectra of the recycled and fresh catalysts demonstrates that the recycled catalyst retains the characteristic bands, confirming that the nanocatalyst structure remains unchanged after five consecutive reaction cycles (Fig. 9B). SEM images also did not show a significant change in the morphology in the surface of recovered catalyst after the 5th recycles (Fig. 10). EDX analysis was also conducted to confirm the presence of choline chloride in the recycled catalyst. The EDX spectrum clearly displays the chlorine peak, verifying the presence of choline chloride onto recycled catalyst (Figs. 10 and 11).

Fig. 8.

Fig. 8

Recycling study of MNP@Arg/ChCl for the model reaction of 4-chlorobenzaldehyde, isatoic anhydrides, and aniline.

Fig. 9.

Fig. 9

(A) XRD patterns and (B) FT-IR spectra of (a) MNP@Arg/ChCl, (b) reused MNP@Arg/ChCl.

Fig. 10.

Fig. 10

SEM images of (a) fresh MNP@Arg/ChCl and (b) reused MNP@Arg/ChCl after five cycles.

Fig. 11.

Fig. 11

EDX analysis of reused MNP@Arg/ChCl.

To assess the heterogeneity of the catalyst and evaluate the stability of MNPs@Arg/ChCl in the reaction medium, a hot filtration test was conducted. For this purpose, the model reaction was performed in ethanol (Table 2, entry 3). Before the reaction was completed (60 min), the catalyst was separated using an external magnet. The reaction mixture was then maintained under the same conditions for one hour. During this time, the reaction did not proceed further (monitored by TLC), indicating that the catalyst remained insoluble in the solvent. (Fig. 12).

Fig. 12.

Fig. 12

Time conversion curve for the hot-filtration test (a) after separation of catalyst, and (b) the presence of catalyst.

To evaluate the effectiveness of the MNPs@Arg/ChCl in synthesizing 2,3-diphenyl-2,3-dihydroquinazolin-4(1 H)-one (4q) compared to other catalysts reported in the literature, the results are summarized in Table 4. The results indicate that this method provides the product (4q) in a higher yield and shorter reaction time compared to other methods. Notably, the absence of solvents, metals, toxic reagents, and microwave radiation, combined with the ease of operation and separation, as well as the use of two completely natural and environmentally friendly materials, highlights the significant advantages of the introduced nanocatalyst.

Table 4.

Comparison of the catalytic activity of MNP@Arg/ChCl with selected reported catalysts for the synthesis of 4q.

Entry Catalyst Conditions Time (min) Yield Refs.
graphic file with name 41598_2025_23849_Figu_HTML.gif 1 Glutamic acid EtOH/MW 8 85% 40
2 CCMC@MnFe2O4 EtOH/Reflux 40 90% 41
3 Fe3O4@SiO2@Pr-PABA H2O/Reflux 25 85% 42
4 CoFe2O4-Sp-SO3H H2O/60 °C 60 93% 44
5 CS-TDI- SSA-Fe3O4 EtOH/Reflux 23 87% 45
6 Fe3O4@Sap/Cu(II) H2O/RT 30 91% 57
7 N-Bromosaccharin H2O/Reflux 40 87% 58
6 [PyPS]3PW12O40 [2] MW, solvent free/80 °C 15 92% 63
9 ProMSAa Solvent free/120 °C 60 91% 64
10 MNP@Arg/ChCl Solvent free/100 °C 30 93% This work

aProlinium methane sulphonate.

Conclusion

In conclusion, we have developed magnetic deep eutectic solvent as a novel biocompatible nanocatalyst. The nanomagnet-stabilized arginine/choline chloride (MNP@Arg/ChCl) was simply prepared and characterized by FT-IR, XRD, TGA, VSM, and SEM techniques. The synthesized catalyst demonstrated remarkable performance in the synthesis of 2,3-dihydroquinazolin-4(1 H)-one derivatives through three-component reactions involving isatoic anhydride, aniline derivatives, and aldehyde derivatives under solvent-free conditions at 100 °C. The desired products were obtained in yields ranging from 85% to 93% within a short reaction time. Furthermore, the catalyst could be easily separated from the reaction mixture and reused five times without a significant decline in catalytic activity. We believe that this approach can make a valuable contribution to the field of green chemistry for the synthesis of heterocyclic compounds due to its high catalytic efficiency, ease of recovery, reusability, and environmentally friendly properties.

Experimental

The details of the reagents and instruments utilized in this research are presented in Appendix, along with additional supplementary material.

Preparation of Fe3O4 nanoparticles

The synthesis of Fe3O4 nanoparticles was conducted as described in the literature65. A total of 3.78 g of FeCl3·6H2O and 2.22 g of FeSO4·7H2O were dissolved in 150 mL of deionized water. The resulting solution was stirred under N2 at 80 °C for 1 h. Subsequently, 20 mL of NH4OH solution was added, leading to the formation of a black solid. The final mixture was stirred at 80 °C under N2 for 1 h. The precipitate was separated using an external magnet and washed four times with distilled water and ethanol. Finally, the magnetic nanoparticles (MNPs) were dried in a vacuum oven at 60 °C.

Preparation of Fe3O4@SiO2-CPTMS

An amount of 1.0 g of Fe3O4 was dispersed by sonication for 15 min in the mixture of water (6 mL) and ethanol (35 mL). Subsequently, 1.5 mL of tetraethyl orthosilicate (TEOS) was added to the mixture, and sonication was continued for an additional 10 min. Following sonication, 1.4 mL of ammonium hydroxide (NH4OH, 28%) was gradually added. After stirring at 80 °C for 12 h, the silica-coated magnetic nanoparticles (Fe3O4@SiO2) were separated via a permanent magnet, rinsed several times with deionized water and methanol, and then dried at 25 °C under vacuum56. In the second step, 1.0 g of Fe3O4@SiO2 was suspended in dry toluene through sonication for 35 min. Subsequently, 3 mL of 3-chloropropyltrimethoxysilane (CPTMS) was slowly added dropwise to the solution. The mixture was kept at reflux and stirred under an argon atmosphere for 12 h. Afterward, the resulting solid was easily separated using an external magnet, washed with ethanol and acetone, and then dried in a vacuum at 80 °C66.

Preparation of MNPs@Arg/ChCl

The synthesized Fe3O4@SiO2-CPTMS (1.0 g) was dispersed in dry toluene over a period of 20 min. Subsequently, 1.5 g of L-arginine and 1.2 mmol of triethylamine were added to the mixture, which was stirred for 48 h at 100 °C. Afterward, the nanoparticles (Fe3O4@SiO2-CPTMS@Arg) were filtered, washed with deionized (DI) water and ethanol, and dried at vacuum oven 100 °C for 12 h67. The synthesis of the magnetic DES was carried out according to the procedures described in the literature2729. Specifically, 0.419 g of choline chloride (3.0 mmol) was added to 1.0 g of Fe3O4@SiO2-CPTMS@Arg in 30 mL of toluene and the mixture was refluxed for 18 h. Upon completion of the reaction, the resulting catalyst (named as MNPs@Arg/ChCl) was washed with 50 mL of n-hexane and 20 mL of ethanol, and then allowed to air dry.

General procedure for the synthesis of 2,3-disubstituted-2,3-dihydroquinazolin-4(1 H)-ones

A mixture of isatoic anhydride (0.5 mmol), aldehyde (0.5 mmol), aniline (0.6 mmol), and 30 mg of MNPs@Arg/ChCl was stirred at 100°. The progress of the reaction was monitored using TLC with a mobile phase of n-hexane and ethyl acetate (EtOAc) in a 3:1 ratio. Upon completion, ethanol was added and the nanocatalyst was easily separated using an external magnet, and the residue was concentrated to yield the crude product. Finally, recrystallization of the crude product from 5 mL of ethanol resulted in a pure product.

The structures of the products were confirmed by comparing their physical data and 1H and 13C NMR spectra (see Figures S1-S38 Supplementary Information).

Selected data for the products: 2-(4-Chlorophenyl)-3-phenyl-2,3-dihydroquinazolin-4(1H)-one (4a): White solid; mp 223–225 °C; 1H NMR (300 MHz, DMSO-d6) (ppm): δ 7.74–7.69 (m, 2 H, HAr and NH), 7.39–7.21 (m, 10 H), 6.78–6.71 (m, 2 H), 6.34 (d, J = 2.7 Hz, 1H). 13C NMR (100 MHz, DMSO-d6) δC: 162.6, 146.8, 141.0, 140.1, 134.3, 133.3, 129.1, 129.0, 128.8, 128.4, 126.7, 126.6, 118.1, 115.7, 115.3, 72.3.

2-(4-Nitrophenyl)-3-phenyl-2,3-dihydroquinazolin-4(1H)-one (4c): White solid; mp 194–196 °C; 1H NMR (300 MHz, DMSO-d6) (ppm): δ 8.20 (d, J = 8.7 Hz, 2 H), 7.85 (d, J = 3.3 Hz, 1H, NH), 7.74 (dd, J = 7.8, 1.8 Hz, 1H), 7.66 (d, J = 8.7 Hz, 2 H), 7.40–7.23 (m, 5 H), 7.26–7.17 (m, 1H), 6.79–6.72 (m, 2 H), 6.51 (d, J = 3.0 Hz, 1H). 13C NMR (100 MHz, DMSO-d6) δC: 162.4, 148.4, 147.8, 140.9, 134.5, 129.2, 128.5, 128.3, 126.7, 126.5, 124.6, 124.1, 118.4, 115.7, 115.4, 72.1.

3-Phenyl-2-(p-tolyl)-2,3-dihydroquinazolin-4(1H)-one (4f): White solid; mp 212–214 °C; 1H NMR (300 MHz, DMSO-d6) (ppm): δ 7.73 (dd, J = 7.8, 1.5 Hz, 1H), 7.63 (d, J = 2.7 Hz, 1H, NH), 7.36–7.25 (m, 7 H), 7.19 (td, J = 6.9, 1.5 Hz, 1H), 7.11 (d, J = 7.8 Hz, 2 H), 6.77–6.69 (m, 2 H), 6.25 (d, J = 2.7 Hz, 1H), 2.30 (s, 3 H). 13C NMR (100 MHz, DMSO-d6) δC: 162.7, 147.0, 141.3, 138.1, 134.1, 129.4, 129.1, 129.0, 128.4, 126.9, 126.5, 126.4, 117.9, 115.8, 115.2, 72.8, 21.0.

2-(3-Methoxyphenyl)-3-phenyl-2,3-dihydroquinazolin-4(1H)-one (4k): White solid; mp 186–188 °C; 1H NMR (300 MHz, DMSO-d6) (ppm): δ 7.76 (dd, J = 7.8, 1.8 Hz, 1H), 7.36–7.17 (m, 9 H, HAr and NH), 7.00 (d, J = 8.4 Hz, 1H), 6.87 (t, J = 7.5 Hz, 1H), 6.79 (d, J = 8.1 Hz, 1H), 6.72 (t, J = 8.4 Hz, 1H), 6.43 (d, J = 2.4 Hz, 1H), 3.77 (s, 3 H). 13C NMR (100 MHz, DMSO-d6) δC: 162.9, 156.4, 147.0, 141.2, 134.1, 130.3, 129.1, 128.3, 128.0, 127.0, 126.8, 126.6, 120.5, 117.8, 115.3, 111.8, 68.6, 56.0.

2-(Thiophene-2-yl)-3-phenyl-2,3-dihydroquinazolin-4(1H)-one (4n): White solid; mp 203–206 °C; 1H NMR (300 MHz, DMSO-d6) (ppm): δ 7.75 (dd, J = 7.8, 1.5 Hz, 1H), 7.63 (d, J = 3.0 Hz, 1H, NH),7.41–7.31 (m, 6 H), 7.25 (td, J = 7.2, 1.5 Hz, 1H), 6.97 (dd, J = 5.1, 1.5 Hz, 1H), 6.89 (t, J = 4.2 Hz, 1H), 6.84 (d, J = 7.8 Hz, 1H), 6.79 (td, J = 7.8, 1.2 Hz, 1H), 6.55 (d, J = 2.7 Hz, 1H). 13C NMR (100 MHz, DMSO-d6) δC: 162.0, 146.8, 145.1, 140.9, 134.3, 129.1, 128.4, 126.9, 126.8, 126.70, 126.3, 118.5, 116.0, 115.7, 69.9.

2-(4-Chlorophenyl)-3-(p-tolyl)-2,3-dihydroquinazolin-4(1H)-one (4w): White solid; mp 249–250 °C; 1H NMR (300 MHz, DMSO-d6) (ppm): δ 7.72 (dd, J = 7.8, 1.8 Hz, 1H), 7.63 (d, J = 3.0 Hz, 1H), 7.38 (s, 4 H), 7.29 (td, J = 7.8, 1.8 Hz, 1H), 7.14 (s, 4 H), 6.77–6.70 (m, 2 H), 6.28 (d, J = 2.7 Hz, 1H), 2.27 (s, 3 H). 13C NMR (100 MHz, DMSO-d6) δC: 162.6, 146.8, 140.2, 138.4, 135.9, 134.2, 129.6, 128.9, 128.8, 126.6, 118.1, 115.2, 110.2, 72.5, 21.0.

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary Material 1 (6.4MB, docx)

Acknowledgements

The authors thank the University of Mazandaran for partial support of this project.

Author contributions

**Z.V: ** Writing – original draft, Methodology, Investigation, Formal analysis, Data curation. **R.H: ** Writing – review & editing, Supervision, Project administration, Conceptualization. **Y.S: ** Conceptualization, Supervision. **B.M: ** Conceptualization, Validation.

Funding

None.

Data availability

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

Declarations

Competing interests

The authors declare no competing interests.

Ethical approval

This work does not contain any studies with human participants or animals performed by any of the authors.

Footnotes

Publisher’s note

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

References

  • 1.Smith, E. L., Abbott, A. P. & Ryder, K. S. Deep eutectic solvents (DESs) and their applications. Chem. Rev.114, 11060–11082 (2014). [DOI] [PubMed] [Google Scholar]
  • 2.Hansen, B. B. et al. Deep eutectic solvents: a review of fundamentals and applications. Chem. Rev.121, 1232–1285 (2020). [DOI] [PubMed] [Google Scholar]
  • 3.Yu, D., Mou, H., Zhao, X., Wang, Y. & Mu, T. Eutectic molecular liquids based on hydrogen bonding and π–π interaction for exfoliating two-dimensional materials and recycling polymers. Chem. Asian J.14, 3350–3356 (2019). [DOI] [PubMed] [Google Scholar]
  • 4.Yu, D. et al. Inverted deep eutectic solvents based on host-guest interactions. Chem. Asian J.14, 4183–4188 (2019). [DOI] [PubMed] [Google Scholar]
  • 5.Choi, Y. H. et al. Are natural deep eutectic solvents the missing link in Understanding cellular metabolism and physiology. Plant. physiol.156, 1701–1705 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Paiva, A. et al. Natural deep eutectic solvents– solvents for the 21st century, ACS sustain. Chem. Eng.2, 1063–1071 (2014). [Google Scholar]
  • 7.Dai, Y., Witkamp, G. J., Verpoorte, R. & Choi, Y. H. Natural deep eutectic solvents as a new extraction media for phenolic metabolites in Carthamus tinctorius L. Anal. chem.85, 6272–6278 (2013). [DOI] [PubMed] [Google Scholar]
  • 8.Shimoni, L. & Glusker, J. P. Hydrogen bonding motifs of protein side chains: descriptions of binding of arginine and amide groups. Protein Sci.4, 65–74 (1995). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Alonso, D. A. et al. Deep eutectic solvents: the organic reaction medium of the century, Eur. J. Org. Chem. 612–632 (2016). (2016).
  • 10.Liu, P., Hao, J. W., Mo, L. P. & Zhang, Z. H. Recent advances in the application of deep eutectic solvents as sustainable media as well as catalysts in organic reactions. RSC Adv.5, 48675–48704 (2015). [Google Scholar]
  • 11.Falcini, C. & Gonzalo, G. D. Deep eutectic solvents as catalysts in the synthesis of active pharmaceutical ingredients and precursors. Catalysts14, 120 (2024). [Google Scholar]
  • 12.Inaloo, I. D. & Majnooni Carbon dioxide utilization in the efficient synthesis of carbamates by deep eutectic solvents (DES) as green and attractive solvent/catalyst systems. New. J. Chem.43, 11275–11281 (2019). [Google Scholar]
  • 13.Inaloo, I. D., Majnooni, S. & Esmaeilpour, M. Superparamagnetic Fe3O4 nanoparticles in a deep eutectic solvent: An efficient and recyclable catalytic system for the synthesis of primary carbamates and monosubstituted ureas. Eur. J. Org. Chem. 3481–3488 (2018). (2018).
  • 14.Bagherzadeh, N., Sardarian, A. R. & Inaloo, I. D. Green and efficient synthesis of thioureas, Ureas, primary O-thiocarbamates, and carbamates in deep eutectic solvent/catalyst systems using thiourea and Urea. New. J. Chem.45, 11852–11858 (2021). [Google Scholar]
  • 15.Makoś-Chełstowska, P., Kaykhaii, M., Płotka-Wasylka, J. & de la Guardia, M. Magnetic deep eutectic solvents–fundamentals and applications. J. Mol. Liq. 365, 120158 (2022). [Google Scholar]
  • 16.Akbarian, M., Sanchooli, E., Oveisi, A. R. & Daliran, S. Choline chloride-coated UiO-66-Urea MOF: a novel multifunctional heterogeneous catalyst for efficient one-pot three-component synthesis of 2-amino-4H-chromenes. J. Mol. Liq. 325, 115228 (2021). [Google Scholar]
  • 17.Azizi, N. & Edrisi, M. Deep eutectic solvent immobilized on SBA-15 as a novel separable catalyst for one-pot three-component Mannich reaction. Microporous Mesoporous Mater.240, 130–136 (2017). [Google Scholar]
  • 18.Lamei, N., Ezoddin, M., Ardestani, M. S. & Abdi, K. Dispersion of magnetic graphene oxide nanoparticles coated with a deep eutectic solvent using ultrasound assistance for preconcentration of methadone in biological and water samples followed by GC–FID and GC–MS. Anal. Bioanal Chem.409, 6113–6121 (2017). [DOI] [PubMed] [Google Scholar]
  • 19.Bakhtiarian, M. & Khodaei, M. M. Synthesis of 2,3-dihydro-4(1 H) Quinazolinones using a magnetic pectin-supported deep eutectic solvent. Colloids Surf. A: Physicochem Eng. Asp. 641, 128569 (2022). [Google Scholar]
  • 20.Rajendran, A., Anand, S., Sunajadevi, K. R. P., Pinheiro, D. & Chundattu, S. J. Heterogeneous catalysis in the synthesis of nitrogen-containing heterocyclics. ChemistrySelect9(41), e202403656 (2024).
  • 21.Rabiei, K., Pouramiri, B., Neshati, M. & Imanvand, A. Eco-friendly synthesis of various tri-substituted imidazole derivatives by using a novel functionalized NCP@ SiO3PrNHPrSiO3TiO2 under solvent-free conditions. J. Organomet. Chem.1001, 122888 (2023). [Google Scholar]
  • 22.Mohammadkhani, Z., Rabiei, K., Keypour, H., Kouhdareh, J. & Karakaya, I. Synthesizing and post-synthetically modifying metal-organic frameworks (Co (BDC)-NH2) for carbonylative Sonogashira coupling reaction. J. Organomet. Chem.999, 122822 (2023). [Google Scholar]
  • 23.Karimi-Nami, R. et al. Pd immobilization phenanthroline-2, 9-dicarbaldehyde modified magnetic CuBDC MOF as a reusable heterogeneous catalyst for suzuki-miyaura cross-coupling reactions. Polycycl. Aromat. Compd.44 (6), 4065–4080 (2024). [Google Scholar]
  • 24.Zamani, P. et al. Magnetic nanoparticle anchored deep eutectic solvents as a catalyst for the etherification and amination of naphthols. Adv. Synth. Catal.360, 4372–4380 (2018). [Google Scholar]
  • 25.Nguyen, T. T., Le, N. & Tran, P. T. An efficient multicomponent synthesis of 2,4,5-trisubstituted and 1,2,4,5-tetrasubstituted imidazoles catalyzed by a magnetic nanoparticle supported Lewis acidic deep eutectic solvent. RSC Adv.9, 38148–38153 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Tavakol, H. & Keshavarzipour, F. Preparation of choline chloride–urea deep eutectic solvent-modified magnetic nanoparticles for synthesis of various 2-amino-4H-pyran derivatives in water solution. Appl. Organomet. Chem.31, e3811 (2017). [Google Scholar]
  • 27.Anizadeh, M. R., Zolfigol, M. A., Torabi, M., Yarie, M. & Notash, B. Urea-dithiocarbamic acid functionalized magnetic nanoparticles modified with chcl: catalytic application for the synthesis of novel hybrid pyridones via cooperative geminal-vinylogous anomeric-based oxidation. J. Mol. Liq. 364, 120016 (2022). [Google Scholar]
  • 28.Benrashid, A., Habibi, D. & Beiranvand, M. Deep eutectic solvent-immobilized magnetic nanoparticles as a capable catalyst for the green synthesis of Triazolopyrimidines. J. Mol. Liq. 409, 12550 (2024). [Google Scholar]
  • 29.Zarei, N., Zolfigol, M. A., Torabi, M. & Yarie, M. Synthesis of new hybrid pyridines catalyzed by Fe3O4@SiO2@urea-riched ligand/ChCl. Sci. Rep.13, 9486 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Sharma, S., Gallou, F. & Handa, S. Towards a sustainable tomorrow: advancing green practices in organic chemistry. Green. Chem.26, 6289–6317 (2024). [Google Scholar]
  • 31.Nandi, S. et al. One-pot multicomponent reaction: a highly versatile strategy for the construction of valuable nitrogen-containing heterocycles. ChemistrySelect7, e202201901 (2022). [Google Scholar]
  • 32.Younus, H. A. et al. Multicomponent reactions (MCR) in medicinal chemistry: a patent review (2010–2020). Expert Opin. Ther. Pat.31, 267 (2021). [DOI] [PubMed] [Google Scholar]
  • 33.Birch, H. L. et al. Novel 7-methoxy-6-oxazol-5-yl-2,3-dihydro-1H-quinazolin-4-ones as IMPDH inhibitors. Bioorg. Med. Chem. Lett.15, 5335–5339 (2005). [DOI] [PubMed] [Google Scholar]
  • 34.Abdel-Aziz, A. et al. Synthesis, anti-inflammatory, analgesic, COX-1/2 inhibitory activities and molecular Docking studies of substituted 2-mercapto-4(3H)-quinazolinones. J. Med. Chem.121, 410–421 (2016). [DOI] [PubMed] [Google Scholar]
  • 35.Poudapally, S. et al. Cu-Free Sonogashira type cross-coupling of 6-halo-2-cyclopropyl-3-(pyridyl-3-ylmethyl) quinazolin-4(3H)-ones as potential antimicrobial agents. J. Heterocycl. Chem.54, 2272–2286 (2017). [Google Scholar]
  • 36.Alagarsamy, V., Solomon, V. R. & Murugan, M. Synthesis and Pharmacological investigation of novel 4-benzyl-1-substituted-4H-[1,2,4]triazolo[4,3-a]quinazolin-5-ones as new class of H1-antihistaminic agents. Bioorg. Med. Chem.15, 4009–4015 (2007). [DOI] [PubMed] [Google Scholar]
  • 37.Barmak, A., Niknam, K. & Mohebbi, G. Synthesis, structural studies, and α-glucosidase inhibitory, antidiabetic, and antioxidant activities of 2,3-dihydroquinazolin-4(1H)-ones derived from Pyrazol-4-carbaldehyde and anilines. ACS Omega. 4, 18087–18099 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Mohammadi, A. A., Rohi, H. & Soorki, A. A. Synthesis and in vitro antibacterial activities of novel 2-aryl-3-(phenylamino)-2,3-dihydroquinazolin-4(1H)-one derivatives. J. Heterocycl. Chem.50, 1129–1133 (2013). [Google Scholar]
  • 39.Xu, Z. et al. Antifungal Quinazolinones from marine-derived Bacillus cereus and their Preparation. Med. Chem. Lette. 21, 4005–4007 (2011). [DOI] [PubMed] [Google Scholar]
  • 40.Mane, R. et al. Glutamic acid-catalyzed synthesis of dihydroquinazolinone: anticancer activity, electrochemical behavior, molecular docking, dynamics, simulations and drug-likeness studies. Res. Chem. Intermed. 50, 3271–3303 (2024). [Google Scholar]
  • 41.Rahimi, S. et al. Synthesis of antioxidant and antibacterial active Quinazolinones by carboxymethyl cellulose@MnFe2O4 biocatalyst. Inorg. Chem. Commun.158, 111556 (2023). [Google Scholar]
  • 42.FaniMoghadam, H., Dekamin, M. G. & Rostami, N. Para-aminobenzoic acid grafted on silica-coated magnetic nanoparticles: a highly efficient and synergistic organocatalyst for on-water synthesis of 2,3-dihydroquinazolin-4(1H)-ones. Res. Chem. Intermed. 48, 3061–3089 (2022). [Google Scholar]
  • 43.Wang, L. M., Hu, L., Shao, J. H., Yu, J. & Zhang, L. A novel catalyst zinc(II) perfluorooctanoate [Zn(PFO)2]-catalyzed three-component one-pot reaction: synthesis of Quinazolinone derivatives in aqueous micellar media. J. Fluor. Chem.129, 1139–1145 (2008). [Google Scholar]
  • 44.Mashhadi, E. Safaei-Ghomi. J. Sulfonated magnetic spirulina nanobiomaterial as a novel and environmentally friendly catalyst for the synthesis of dihydroquinazolin-4(1H)-ones in aqueous medium. Sci. Rep.14, 2296 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.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. Carbohydr. Polym. Technol. Appl.7, 100420 (2024). [Google Scholar]
  • 46.Tamoradi, T., Ghadermazi, M. & Ghorbani-Choghamarani, A. Synthesis of polyhydroquinoline, 2,3-dihydroquinazolin-4(1H)-one, sulfide and sulfoxide derivatives catalyzed by new copper complex supported on MCM-41. Catal. Lett.148, 857–872 (2018). [Google Scholar]
  • 47.Mirjalili, B. B. F., Bamoniri, A. & Azad, S. Synthesis of 2,3-dihydroquinazolin-4(1H)-ones catalyzed by nano-Fe3O4/TiCl2/cellulose as a bio-based magnetic catalyst. J. Iran. Chem. Soc.14, 47–55 (2017). [Google Scholar]
  • 48.Safaei-Ghomi, J., Teymuri, R., Bakhtiari, A. Co-aminobenzamid@Al-SBA-15: a favorable catalyst in synthesis of 2,3-dihydroquinazolin-4(1H)-ones. BMC Chem.13, 1–12 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Valipour, Z., Hosseinzadeh, R., Sarrafi, Y. & Maleki, B. Natural deep eutectic solvent as a green catalyst for the one-pot synthesis of Chromene and 4H-pyran derivatives. Org. Prep Proced. Int.56, 105–117 (2024). [Google Scholar]
  • 50.Hosseinzadeh, R., Zarei, S., Valipour, Z. & Maleki, B. A vitamin C-based natural deep eutectic solvent for the synthesis of 2,3-dihydroquinazolin-4(1H)-one derivatives. Heliyon10, e37170 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Jannatabadi, S. A., Hosseinzadeh, R. & Maleki, B. One-pot synthesis of Dihydropyrimidinone and polyhydroquinoline derivatives with natural deep eutectic solvents (NADESs): alternative to toxic organic solvents and environmental eco-friendly. Results Chem.12, 101848 (2024). [Google Scholar]
  • 52.Hosseinzadeh, R., Karimi, S., Valipour, Z. & Aghili, N. Synthesis of 2,3-dihydroquinazolinones via condensation of 2-aminobenzamides with aldehydes under ball milling conditions. Org. Prep Proced. Int.56, 307–316 (2024). [Google Scholar]
  • 53.Roustaie, E., Tajbakhsh, M. & Hosseinzadeh, R. Green synthesis of the ZnO/ZSM-5 composite by using convolvulus persicus extract: as an efficient and reusable catalyst in the Biginelli reaction. J. Organomet. Chem.1001, 122885 (2023). [Google Scholar]
  • 54.Kumar, S. & Rai, S. B. Spectroscopic studies of L-arginine molecule. Indian J. Pure Appl. Phys.48, 251–255 (2010). [Google Scholar]
  • 55.Amirnejat, S., Nosrati, A., Javanshir, S. & Naimi-Jamal, M. R. Superparamagnetic alginate-based nanocomposite modified by L-arginine: an eco-friendly bifunctional catalysts and an efficient antibacterial agent. Int. J. Biol. Macromol.152, 834–845 (2020). [DOI] [PubMed] [Google Scholar]
  • 56.Hosseinzadeh, R., Mavvaji, M., Tajbakhsh, M. & Lasemi, Z. Synthesis and characterization of N-hydroxyphthalimide immobilized on SiO2-coated Fe3O4 nanoparticles as magnetic catalyst for oxidation of benzyl alcohols and hydrocarbons. J. Iran. Chem. Soc.15, 893–904 (2018). [Google Scholar]
  • 57.Razavi, N. & Akhlaghinia, B. Hydroxyapatite nanoparticles (HAP NPs): a green and efficient heterogeneous catalyst for three-component one-pot synthesis of 2,3-dihydroquinazolin-4(1H)-one derivatives in aqueous media. New. J. Chem.40, 447–457 (2016). [Google Scholar]
  • 58.Hutait, S., Maikhuri, V. K., Kumar Pandey, A., Shukla, P. & Srivastava, S. Versatile Quinazolinone synthesis: N-bromosaccharin-catalyzed divergence in 2,3-dihydroquinazolin-4(1H)-one and quinazolin-4(3H)-one formation. ChemistrySelect9, e202304087 (2024). [Google Scholar]
  • 59.Zhang, J., Ren, D., Ma, Y., Wang, W. & Wu, H. CuO nanoparticles catalyzed simple and efficient synthesis of 2,3-dihydroquinazolin-4(1H)-ones and quinazolin-4(3H)-ones under ultrasound irradiation in aqueous ethanol under ultrasound irradiation in aqueous ethanol. Tetrahedron70, 5274 (2014). [Google Scholar]
  • 60.Bollini, M. et al. Synthesis, 2D-QSAR studies and biological evaluation of Quinazoline derivatives as potent anti-trypanosoma Cruzi agents. Med. Chem.15, 265–276 (2019). [DOI] [PubMed] [Google Scholar]
  • 61.Badolato, M., Aiello, F. & Neamati, N. 2,3-Dihydroquinazolin-4(1H)-one as a privileged scaffold in drug design. RSC Adv.8, 20894–20921 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Zhang, Z. H., Lu, H. Y., Yang, S. H. & Gao, J. W. Synthesis of 2, 3-dihydroquinazolin-4 (1 H)-ones by three-component coupling of isatoic anhydride, amines, and aldehydes catalyzed by magnetic Fe3O4 nanoparticles in water. J. Comb. Chem.12, 643–646 (2010). [DOI] [PubMed] [Google Scholar]
  • 63.Yang, Y., Fu, R., Liu, Y., Cai, J. & Zeng, X. Microwave-promoted one-pot three-component synthesis of 2,3-dihydroquinazolin-4(1H)-ones catalyzed by heteropolyanion-based ionic liquids under solvent-free conditions. Tetrahedron76, 131312 (2020). [Google Scholar]
  • 64.Rode, N. R. et al. Identification of potential antileishmanial derivatives of 2,3-dihydroquinazolin-4(1H)-one synthesized using AAIL catalyst and molecular Docking studies. Res. Chem. Intermed. 50, 1125–1145 (2024). [Google Scholar]
  • 65.Atacan, K. & Özacar, M. Characterization and immobilization of trypsin on Tannic acid modified Fe3O4 nanoparticles. Colloids Surf. B Biointerfaces. 128, 227–236 (2015). [DOI] [PubMed] [Google Scholar]
  • 66.Tajbakhsh, M., Farhang, M., Hosseinzadeh, R. & Sarrafi, Y. Nano Fe3O4 supported biimidazole Cu(i) complex as a retrievable catalyst for the synthesis of imidazo[1,2-a]pyridines in aqueous medium. RSC Adv.4, 23116–23124 (2014). [Google Scholar]
  • 67.Kargar Karkhah, M., Kefayati, H. & Shariati, S. Enantioselective synthesis of 3-amino-1-aryl-1H-benzo[f]chromene-2-carbonitrile derivatives by Fe3O4@PS-arginine as an efficient chiral magnetic nanocatalyst. Appl. Organomet. Chem.33, e5139 (2019). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Material 1 (6.4MB, docx)

Data Availability Statement

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


Articles from Scientific Reports are provided here courtesy of Nature Publishing Group

RESOURCES