Open Access Article
Majidreza Gerami,
Mahnaz Farahi
* and
Bahador Karami
Department of Chemistry, Yasouj University, Yasouj, 75918-74831, Iran. E-mail: farahimb@yu.ac.ir; Fax: +98 7412242167
First published on 25th November 2025
A novel Brønsted acidic ionic liquid immobilized on modified magnetic lignin (Fe3O4/MPL–[IL]) was successfully synthesized and comprehensively characterized by FT-IR, XRD, EDX, VSM, FE-SEM, TGA, and TEM analyses. The prepared nanocatalyst was utilized for preparing 2-amino-4-aryl-4H-benzo[f]chromene-3-carbonitriles via the reaction between aryl aldehydes, β-naphthol, and active methylene compounds (malononitrile or ethyl cyanoacetate), affording high yields of 87–97% within short reaction times (20–80 min) under mild conditions. Necessary experiments were conducted for the recyclability test of the prepared catalyst, and the results showed that the Fe3O4/MPL–[IL] nanocatalyst can be reused five times while maintaining about 95% of its initial catalytic activity, with excellent stability.
Ionic liquids (ILs) are salts composed of large asymmetric organic cations and inorganic or organic anions that remain in the liquid state at or near room temperature.27 Ionic liquids have different properties from molecular liquids, making them favorable materials for use in a diversity of fields. During the past decade, they have been central to many areas of chemistry because of their broad applicability and unique properties such as a wide liquid range, low vapor pressure, good conductivity, high thermal stability, and non-flammability. Their physicochemical properties are largely governed by their special structure and the interaction of ions.28–30 Ionic liquids decrease environmental impact by minimizing solvent releases and improving reaction efficiency. Owing to negligible vapor pressures and their non-flammable nature, they have been excellent alternatives to traditional organic solvents in chemical reactions. Their structures can be improved by changing the ions, allowing chemists to design them for specific functions. Furthermore, they are widely used in a variety of applications, such as electrochemistry,31 organic synthesis catalysis,32 extraction,33 and separation processes.34 They also show a role in processing biomass by way of dissolving tough materials. Researchers are also exploring their use in drug delivery systems and nanotechnology, where their ability to dissolve a wide range of substances and interact selectively with biomolecules offers exciting potential.35–39 One vital application of ionic liquids is that they show wide usage as a catalyst in many organic transformations. However, when employed alone as homogeneous catalysts, ILs may face limitations such as difficult recovery and potential leaching. Therefore, immobilizing ionic liquids on solid substrates, such as magnetic lignin, has emerged as a promising strategy to improve their recyclability and stability in numerous catalytic and separation processes.40,41 This synergy often leads to enhanced catalytic activity, better selectivity, and greater process efficiency.42,43 A combination of fascinating features of ionic liquid with those of the supporting material will develop novel performances when the synergistic effects appear. ILs have been supported on different solid supports such as MCM-41,44 chitosan,45 SBA-15,46 Merrifield resin,47 silica gel,48 alumina,49 molecular sieves,50 clays,51 carbon nanotubes,52 etc. The obtained materials are known as supported ionic liquids (SILs). Different methods have been used for IL immobilization, such as the sol–gel method,53 impregnation,54 encapsulation,55 and grafting method.56
In recent years, there has been a growing interest in the syntheses of benzo[f]chromene and its derivatives because most of the compounds with biological activity are derived from these compounds. Benzo[f]chromenes play a fundamental role in both organic chemistry and pharmaceutical development due to their diverse chemical properties and biological activities.57,58 Their rigid polycyclic framework also renders them valuable scaffolds in medicinal chemistry and material science. Compounds containing the benzo[f]chromene skeleton are known to have antitumor properties.59 Moreover, several prospective non-sedative anxiolytic agents with the benzo[f]chromene structure have been discovered.60 These heterocyclic compounds widely occur in nature in the form of alkaloids, vitamins, pigments, and as constituents of plant and animal cells.61 Therefore, several methods have been reported for the preparation of benzo[f]chromene derivatives.62,63
Following our continuous interest in introducing new and safe heterogeneous catalysts,64–70 and by considering all the reasons and importance of clean synthetic procedures, this work presents the design and synthesis of Fe3O4/MPL-supported Brønsted acidic ionic liquid system as a heterogeneous catalyst, which combines the benefits of easy recovery, eco-friendliness, and cost-effectiveness compared to conventional homogeneous catalysts.71–73 The structural and chemical features of the catalyst were thoroughly characterized, and its performance was evaluated in the one-pot synthesis of 2-amino-4-aryl-4H-benzo[f]chromene-3-carbonitrile derivatives. In addition, the catalyst recyclability and operational stability were examined to assess its practical applicability in green synthetic processes.
Finally, the Fe3O4/MPL–[IL] was produced via radical copolymerization between IL and Fe3O4/MPL (Scheme 2). The resulting nanocomposite was analyzed by various characterization techniques, including FT-IR, XRD, EDX, VSM, FE-SEM, TGA, and TEM analysis.
The synthesis of Fe3O4/MPL–[IL] nanocatalyst was monitored using FT-IR spectroscopy (Fig. 1). As illustrated in Fig. 1a, broad absorption near 3410 cm−1 signifies the O–H stretching of phenolic groups inherent to the lignin matrix. The signal at 2943 cm−1 corresponds to vibrations of aliphatic C–H bonds. A band at 1627 cm−1 evidences the presence of carbonyl groups. Aromatic C
C stretching vibrations appear at 1408 cm−1, confirming the benzene ring structure (Fig. 1a).74 The FT-IR spectrum of magnetite (Fe3O4) is characterized by absorption bands in the low-wavenumber region, which arise from lattice vibrations of metal–oxygen bonds. A strong and well-defined band typically appears at approximately 570–590 cm−1 corresponding to the stretching vibrations of Fe–O bonds (Fig. 1b).75 In Fig. 1c, characteristic bands for symmetric and asymmetric Si–O stretching are found at 1044 and 1122 cm−1, respectively, and a peak indicative of S–H bonds is observed at 2550 cm−1.76 The presence of the Fe–O peak in all samples implies that Fe3O4 nanoparticles were retained throughout modification steps (Fig. 1d and f). In Fig. 1e, the absorption bands at 1613 cm−1 correspond to the stretching vibration of a C
N band, and the peaks around 3154 cm−1 is related to sp2 C–H of imidazolium ring. Vibrational bands located at 1042 cm−1 and 1125 cm−1 are attributed to C–S and S
O bonds, which confirms the existence of the –SO3H group (Fig. 1e).77 Notably, in the FT-IR spectrum of the final Fe3O4/MPL–[IL] nanocatalyst (Fig. 1f), several diagnostic absorption bands confirm the successful integration of the composite constituents. The signals located at approximately 2945 cm−1 and 3418 cm−1 are attributed to the stretching vibrations of aliphatic –CH2 groups and hydroxyl functionalities, respectively, which indicate the presence of lignin-derived structural units as well as surface –OH groups originating from the modification steps. A distinct absorption at 1613 cm−1 corresponds to the C
N stretching vibration, providing clear evidence for the incorporation of the imidazolium-based ionic liquid into the hybrid structure. The simultaneous presence of Fe–O vibrational modes alongside these organic functional groups confirms that Fe3O4 nanoparticles remained intact throughout the modification process and were effectively stabilized within the lignin/ionic liquid matrix. Collectively, these spectral features demonstrate the preservation of lignin's inherent functionalities, the successful anchoring of ionic liquid moieties, and the structural integrity of Fe3O4, thereby validating the synthesis of the targeted Fe3O4/MPL–[IL] nanocatalyst.
![]() | ||
| Fig. 1 The FT-IR spectrum of (a) lignin, (b) Fe3O4, (c) MPL, (d) Fe3O4/MPL, (e) [IL], and (f) Fe3O4/MPL–[IL]. | ||
XRD analysis was used to examine the structural properties of Fe3O4, Fe3O4–lignin, and the Fe3O4/MPL–[IL] (Fig. 2). For Fe3O4 (Fig. 2a), six prominent peaks appeared at 2θ values of 30.27°, 35.6°, 42.9°, 53.61°, 57.12°, and 63.31°, corresponding to the (220), (311), (400), (422), (511), and (440) planes, respectively, confirming the formation of a pure spinel crystal structure.78 The XRD pattern of the Fe3O4–lignin composite structure, as shown in Fig. 2b, closely matches previous research findings on Fe3O4-coated lignin structures. This pattern displays six distinct peaks at 30.28°, 35.6°, 43.3°, 53.57°, 57.47°, and 63.31°, which correspond respectively to the (220), (311), (400), (422), (511), and (440) crystallographic planes.79 The XRD pattern for the Fe3O4/MPL–[IL] catalyst (Fig. 2c) demonstrates significant changes upon functionalization. The structural fingerprint of lignin was altered after modification with layered materials and incorporation of Fe3O4. Comparison with the earlier spectra (Fig. 2a and b) verifies that Fe3O4 nanoparticles were successfully integrated and stabilized with ionic liquid (IL) on the lignin matrix, confirming the successful synthesis of the desired nanocatalyst.
The surface morphology and particle size of the Fe3O4/MPL–[IL] nanocatalyst were evaluated using FE-SEM images (Fig. 3). According to the FE-SEM images, the prepared Fe3O4/MPL–[IL] nanocatalyst has a spherical morphology and uniform particle distribution, with an orderly size of less than 70 nm. The SEM image of pure Fe3O4 nanoparticles exhibits a nearly spherical morphology with uniform size distribution. This confirms that the Fe3O4 core structure remains stable during the subsequent MPL and IL modification processes used in the Fe3O4–MPL/IL catalyst synthesis.
To gain insight into the chemical composition of the Fe3O4/MPL–[IL] nanoparticles, an EDX analysis was performed, and the results are shown in Fig. 4. The spectrum reveals a prominent signal for iron (Fe), identifying it as the primary metallic constituent. A notable oxygen (O) peak also appears, reflecting a high presence of oxygen-containing groups. Additionally, the signals corresponding to carbon (C), sulfur (S), silicon (Si), and nitrogen (N) are detected. These non-metallic elements likely originate from the lignin component, supporting the effective attachment of both lignin and ionic liquid onto the ferrite nanoparticles.
Fig. 5 presents the elemental mapping results for the Fe3O4/MPL–[IL] nanocatalyst. The distribution maps clearly demonstrate that all key elements, carbon (C), silicon (Si), iron (Fe), oxygen (O), sulfur (S), and nitrogen (N), are evenly dispersed across the catalyst's surface. This uniform elemental spread strongly supports the effective incorporation and stable anchoring of the intended components onto the lignin matrix during the synthesis process.
Fig. 6 shows TEM images of the Fe3O4/MPL–[IL] nanocatalyst. These images reveal magnetite NPs with black cores surrounded by a gray shell of modified lignin. Additionally, the images show that the nanoparticles mainly consist of small, nearly spherical particles.
The magnetic behavior of the prepared nanocatalyst was examined by vibrating sample magnetometry (VSM), as depicted in Fig. 7. According to the VSM curves, the magnetic saturation values for Fe3O4 and Fe3O4/MPL–[IL] composite are 52 and 8 emu g−1, respectively. This reduction in magnetization is attributed to the presence of a lignin coating and the incorporation of an ionic liquid, which contributes to the overall decrease in magnetic intensity compared to the uncoated magnetic nanoparticles. However, the magnetic sensitivity of the nanocatalyst is sufficient for its easy magnetic recovery from various reaction mixtures. Therefore, the prepared catalyst can be readily recovered using magnets.
The thermal behavior of the Fe3O4/MPL–[IL] nanocatalyst was assessed through thermogravimetric analysis (TGA), as shown in Fig. 8. The measurements were conducted over a temperature range of 20 to 1000 °C. The resulting TGA profile reflects the structural stability of the material and confirms the presence of functional groups anchored to the nanostructure. An initial mass loss of approximately 4% occurs below 150 °C, which is attributed to the release of physically adsorbed moisture, surface-bound hydroxyl species, and interstitial water molecules. A more pronounced weight reduction, approximately 25%, is observed between 150 °C and 550 °C, likely due to the thermal decomposition of organic moieties, including amine functionalities and sulfate groups attached to the catalyst surface. Finally, a smaller weight decrease of about 6% in the 550–1000 °C range is linked to the breakdown of covalently bonded organic groups on the lignin framework, further indicating the thermal robustness of the synthesized nanocatalyst.
The efficiency of the Fe3O4/MPL–[IL] nanocatalyst, after its properties were identified, was assessed in producing 2-amino-4-aryl-4H-benzo[f]chromen-3-carbonitriles 4 via a reaction between aromatic aldehydes 1, malononitrile 2, and 2-naphthol 3 (Scheme 3).
![]() | ||
| Scheme 3 Synthesis of 2-amino-4-aryl-4H-benzo[f]chromen-3-carbonitriles 4 using Fe3O4/MPL–[IL] nanocatalyst. | ||
Initially, to optimize the conditions, the reaction between benzaldehyde (1 mmol), malononitrile (1 mmol), and 2-naphthol (1 mmol) was selected as a model system (Table 1). Without the catalyst, only low yields were achieved, even at higher temperatures, emphasizing the essential role of catalysis in this process. Catalyst loading testing revealed that 0.004 g was optimal, yielding the highest yield at 80 °C. Raising the temperature to 95 °C with this catalyst amount further increased the yield to 97%. Then, the effect of various solvents was investigated to test the model reaction, and the highest yield was observed under solvent-free conditions. Overall, the optimal conditions are 0.004 g of Fe3O4/MPL–[IL] at 95 °C in a solvent-free environment, yielding excellent catalytic performance. Building on these results, a range of aryl-substituted aldehydes was employed to synthesize a series of 2-amino-4-aryl-4H-benzo[f]chromen-3-carbonitrile derivatives. According to the data presented in Table 2, these compounds were formed in good to excellent yields.
| Entry | Catalyst loading (g) | Solvent | Temp. (°C) | Yieldb (%) |
|---|---|---|---|---|
| a Reaction conditions: benzaldehyde (1 mmol), malononitrile (1 mmol), 2-naphthol (1 mmol). Time: 40 min.b Isolated yields. | ||||
| 1 | Cat. 1 (—) | — | 25 | — |
| 2 | Cat. 1 (—) | — | 80 | 10 |
| 3 | Cat. 1 (0.002) | — | 80 | 60 |
| 4 | Cat. 1 (0.003) | — | 80 | 75 |
| 5 | Cat. 1 (0.004) | — | 80 | 80 |
| 6 | Cat. 1 (0.005) | — | 80 | 70 |
| 7 | Cat. 1 (0.004) | — | 70 | 80 |
| 8 | Cat. 1 (0.004) | — | 95 | 97 |
| 9 | Cat. 1 (0.004) | — | 100 | 90 |
| 10 | Cat. 1 (0.004) | EtOH | Reflux | 80 |
| 11 | Cat. 1 (0.004) | H2O | 95 | 70 |
| 12 | Cat. 1 (0.004) | Methanol | Reflux | 80 |
| 13 | Cat. 1 (0.004) | Toluene | 95 | 60 |
| 14 | Cat. 1 (0.004) | EtOH/H2O | 95 | 75 |
| 15 | Fe3O4 (0.004) | Solvent-free | 95 | — |
| 16 | Fe3O4/MPL (0.004) | Solvent-free | 95 | 55 |
| 17 | IL (0.004) | Solvent-free | 95 | 40 |
| Entry | Product 4 | M.p. (°C) | Yieldb (%) | TONc | TOFd |
|---|---|---|---|---|---|
| a Reaction conditions: aldehyde (1 mmol), malononitrile or ethyl cyanoacetate (1 mmol), 2-naphthol (1 mmol), Fe3O4/MPL–[IL] (0.004 g), 95 °C, 20–80 min.b Isolated yields.c Turnover number [all TONs were calculated by this equation: yield (%)/Cat. (mol%)].d Turnover frequency [all TOFs were calculated by this equation: TON/time (min)]. | |||||
| 4a | ![]() |
282–284 (ref. 80) | 97 | 122 | 3.05 |
| 4b | ![]() |
241–243 (ref. 80) | 89 | 112 | 2.03 |
| 4c | ![]() |
261–263 (ref. 80) | 96 | 121 | 3.02 |
| 4d | ![]() |
183–185 (ref. 80) | 97 | 122 | 3.05 |
| 4e | ![]() |
240–242 (ref. 81) | 95 | 119 | 2.97 |
| 4f | ![]() |
255–257 (ref. 82) | 96 | 121 | 3.02 |
| 4g | ![]() |
209–211 (ref. 82) | 87 | 109 | 1.81 |
| 4h | ![]() |
259–261 (ref. 82) | 88 | 110 | 1.83 |
| 4i | ![]() |
238–240 (ref. 82) | 92 | 116 | 2.57 |
| 4j | ![]() |
250–252 (ref. 82) | 95 | 119 | 2.97 |
| 4k | ![]() |
173–175 (ref. 83) | 96 | 121 | 4.84 |
| 4l | ![]() |
141–143 (ref. 83) | 95 | 119 | 4.76 |
| 4m | ![]() |
150–152 (ref. 84) | 95 | 119 | 4.76 |
Based on the findings, Scheme 4 illustrates a proposed pathway for the reaction. Initially, Fe3O4/MPL–[IL] magnetic nanoparticles serve as an acidic nanocatalyst to activate the aldehyde's carbonyl group, facilitating the formation of α,β-unsaturated intermediate I (2-benzylidene malononitrile) through a Knoevenagel condensation.85 Subsequently, a Michael addition between 2-naphthol and intermediate I yields intermediate II. Finally, the intramolecular cyclization of the adduct II to intermediate III and its tautomerization (1,3-H shift) gives the corresponding products.
The potential for reusing Fe3O4/MPL–[IL], an aspect critical from both environmental and cost perspectives, was investigated in the reaction of benzaldehyde, malononitrile, and 2-naphthol under optimum reaction conditions. Upon completion of the reaction, hot ethanol was added, and the catalyst was isolated from the mixture by employing a magnetic field. It was then rinsed with water and ethanol, dried, and applied again in subsequent reactions. The effectiveness of Fe3O4/MPL–[IL] over multiple reaction cycles was evaluated (Fig. 9). It maintained substantial catalytic function over five consecutive uses, demonstrating its structural integrity. To examine whether Fe3O4/MPL–[IL] operates in a homogeneous or heterogeneous manner, a filtration test has been done in the model reaction under optimized reaction conditions. After nearly 50% of the reaction progress, we separated the Fe3O4/MPL–[IL] catalyst. Then, the mixture residue continued under optimal conditions, but no substantial increase in product conversion was observed. In this regard, no considerable reaction progress was observed, indicating that the active catalytic centers were not washed from the support during the reaction, and the catalyst most likely worked in a heterogeneous manner.
The crystallographic characteristics of the reused catalyst were analyzed through XRD pattern evaluation (Fig. 10). The consistent peak positions and their relative intensities validate the preservation of the catalyst's structure. Furthermore, Fig. 11 displays the FT-IR spectrum of the recycled Fe3O4/MPL–[IL] nanocatalyst. This spectral data confirms that the structural integrity of the catalyst remains intact even after five cycles of use.
Additionally, the catalytic efficiency of Fe3O4/MPL–[IL] was compared to previously reported catalysts for the same transformation. As shown in Table 3, this nanocatalyst shows comparable efficiency to other catalysts in terms of reaction times, conditions, and product yield.
| Entry | Catalyst | Conditions | Time (min)/yielda (%) |
|---|---|---|---|
| a Isolated yields.b This work. | |||
| 1 | N-Octyl phenylphosphinate | Fluorene 0.6 mmol, O2 balloon, 130 °C | 900/72 (ref. 86) |
| 2 | ZSM | H2O2, H2O, 100 °C | 240/52 (ref. 87) |
| 3 | Au–Pd | H2O, MeOH, H2O2, 120 °C | 480/94 (ref. 88) |
| 4 | Mg/Al HT | 0.1 g, EtOH, r.t. | 60/87 (ref. 89) |
| 5 | Sodium dodecyl sulfate | 0.05 mmol, H2O, reflux | 60/87 (ref. 90) |
| 6 | Fe3O4/MPL–[IL] | Solvent-free, 95 °C | 40/97b |
Supplementary information (SI) is available. See DOI: https://doi.org/10.1039/d5ra07067h.
| This journal is © The Royal Society of Chemistry 2025 |