Quang Nhat Trinhab,
Linh Dieu Nguyenab and
Hai Truong Nguyen
*ab
aDepartment of Organic Chemistry, Faculty of Chemistry, University of Science, Ho Chi Minh City, 700000, Vietnam. E-mail: ngthai@hcmus.edu.vn; Tel: +84-908-108-824
bVietnam National University, Ho Chi Minh City, 700000, Vietnam
First published on 7th July 2025
An IL@SGO catalyst was synthesized from leaf residues of Syzygium nervosum via the modification of sulfur-doped graphene oxide (SGO) with dual-acid ionic liquid 3-(3-sulfopropyl)-1-(3-(triethoxysilyl)propyl)-1H-imidazol-3-ium hydrosulfate (IL) and used for the synthesis of a 4-thiazolidinone framework with diverse biological activities. The structure and morphology of IL@SGO were determined using modern physical techniques such as FT-IR spectroscopy, Raman spectroscopy, XRD analysis, ICP-MS analysis, scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDX), and thermogravimetric analysis (TGA). It is anticipated that in the future, the IL@SGO catalyst will emerge as a promising catalyst with its unique and environmentally friendly characteristics. The reaction was carried out at 80 °C for 9 hours in the presence of toluene (5 mL) and IL@SGO (7 mg), and a major product was obtained in a yield ranging from 10% to 49%. Derived from Syzygium nervosum leaves, the IL@SGO catalyst exhibited enhanced acidity and proton conductivity, thereby improving the yield and enabling catalyst recovery and reuse that contribute to environmental preservation.
Graphene is a noteworthy carbon material with a 2D structure composed of sp2 hybridized carbon atoms arranged in a hexagonal pattern, creating a large surface area on both sides of the flat plane. Graphene-derived materials, such as few-layer graphene, GO, and rGO, represent structurally and chemically modified forms with tailored functionalities.3 GO are 2D materials like bulk graphene, with stable covalent bonds within the layers, which is synthesized through the oxidation of graphene. After oxidation, the material's surface forms oxygen-containing anionic groups such as epoxy (–C–O–C), hydroxyl (–OH), carboxyl (–COOH), and carbonyl (–CO) groups. Additionally, GO has unique outstanding properties such as mechanical stability, high surface area, and good dispersion in various environments. Therefore, it frequently serves as a catalyst.4–6 To synthesize GO from biomass, a carbonization process is commonly used, primarily utilizing agricultural and forestry waste rich in cellulose, hemicellulose, and lignin. These materials, due to their high carbon content, can be converted into porous carbon through carbonization at high pyrolysis temperatures and under oxygen-limited conditions. Studies have successfully used sources such as wood, tea leaves, sugarcane bagasse, fruit peels, and chicken bones to modify graphite into GO using Hummer's synthesis method.7 The advantages of GO materials are high surface area, electrical conductivity, mechanical strength, and the ability to be easily functionalized, which are key factors that make GO an attractive platform for the development of advanced materials.8 Traditional liquid acid catalysts show some disadvantages such as corrosion, environmental impact, and difficulty in separation from the reaction mixture.9 The materials containing Brønsted acid sites derived via GO modification signify a notable advancement in materials science, merging the outstanding characteristics of GO with the catalytic benefits of Brønsted acids.10 The modification of GO with Brønsted acid sites can be made through various methods such as direct functionalization of GO with acidic compounds, grafting of acid-functionalized polymers, or impregnation with acidic salts.11 The high surface area of GO facilitates effective charge storage in supercapacitors and batteries, and the acidic sites help in the elimination of contaminants from wastewater via adsorption or catalytic degradation. Moreover, the combination of Brønsted acid sites with GO improves the mechanical characteristics of the material, facilitating the creation of stronger and more resilient composites and coatings.12,13
The synthesis of 4-thiazolidinone plays a vital role in the field of pharmaceutical due to its potential biological activity and its presence in various bioactive nature compounds.14 4-Thiazolidinone is a five-membered heterocyclic compound containing both sulfur (S) and nitrogen (N) atoms in its structure, making it highly significant due to its pharmacological properties.14 The 4-thiazolidinone derivatives exhibit many important biological activities including antimicrobial,15 antitumor,16,17 anticonvulsant,18 anti-inflammatory,19 and antioxidant20 effects. Based on the versatility and bioactivity of this compound, the development of efficient and novel synthetic pathways becomes a key focus in synthetic chemistry. Several synthetic approaches have been proposed to construct the 4-thiazolidinone frame, each method using different starting materials and reaction conditions. Some of the most common methods are cyclization reactions with thiol derivatives or thioureas, and isothiocyanates or α,β-unsaturated carbonyl compounds.21 One of the most widely used methods to synthesize 4-thiazolidinone is the cyclization reaction of thiourea derivatives and α,β-unsaturated carbonyl compounds.21 The reaction commonly proceeds via a nucleophilic attack, leading to the formation of a five-membered thiazolidine ring with good yields and selectivity for 4-thiazolidinone, and is applied to a wide range of substrates, allowing for the functionalization of the desired products. Another synthetic pathway involves the reaction of isothiocyanates with amines or thiols, leading to the formation of a thiazolidine ring with the sulfur atom incorporated into the structure.22 Researchers had also explored multi-component reactions for the synthesis of 4-thiazolidinones, which provided the advantage of a one-pot synthesis.23 A typical multi-component reaction (MCR) may include the condensation of thiols, amines, and α,β-unsaturated carbonyl compounds to directly produce 4-thiazolidinones without isolating intermediate compounds. This approach enhances the yield, reduces the number of steps, and decreases waste, making it a highly appealing option for large-scale synthesis.24 4-Thiazolidinone can be modified to produce several derivatives with improved or customized biological characteristics. The ability to modify the structure at different positions of the thiazolidinone ring, particularly at the 2- or 5-positions, provides opportunities for developing compounds with specific interactions with biological targets.25 Integrating green chemistry concepts such as solvent-free reactions, catalytic techniques, and sustainable reagents can enhance the efficiency and ecological sustainability of 4-thiazolidinone synthesis. In conclusion, synthesizing 4-thiazolidinone is vital for drug development, offering various synthetic routes to create diverse derivatives with significant biological activities. Ongoing advancements in synthetic techniques and continued research on these derivatives in preclinical and clinical settings suggest they will play an important role in developing new therapies for various disorders.
The aim of the research is to outline the novel synthesis route of 3-(3-sulfopropyl)-1-(3-(triethoxysilyl)propyl)-1H-imidazol-3-ium hydrosulfate (IL) with sulfur-doped graphene oxide (SGO) derived from Syzygium nervosum leaves to generate 4-thiazolidinone derivatives. The reaction involving benzaldehydes (1 mmol), anilines (1 mmol), and mercaptoacetic acid (1 mmol) was selected as a model to synthesize 4-thiazolidinone derivatives, owing to its significance in evaluating the catalytic efficiency of IL@SGO. IL@SGO, a composite material containing Brønsted acid sites derived via graphene oxide modification, had a highly reactive surface for the reaction, facilitating an efficient and environmentally benign method for thiazolidinone production. This reaction acts as a realistic standard to illustrate the material's catalytic efficacy under mild conditions. This catalyst is reusable and promotes eco-friendly practices via a one-pot, multi-component reaction process.
The FT-IR spectra presented the comparison between several materials, which were prepared from each stage, including graphite, graphene oxide (GO), sulfur-doped graphene oxide (SGO), a double acidic IL, and IL-functionalized sulfur-doped graphene oxide (IL@SGO) (Fig. 1). The broad absorption band around 3500–3300 cm−1 was most prominent in the spectra of graphite and GO, which indicated the presence of hydroxyl groups of the phenol and alcohol. In GO materials, these hydroxyl groups resulted from the oxidative treatment of graphite, which introduced oxygen-containing functionalities to the basal planes and edges of the GO sheets. The –OH band weakened in the SGO material because some hydroxyl groups were partially removed or chemically transformed during the sulfur doping process. In the IL@SGO material, the –OH band diminished further, suggesting that the IL functionalization either consumed these groups or masked their vibrational signal due to the introduction of new bulky organic groups. The peak at 1720 cm−1 was assigned to the stretching vibrations of carbonyl groups (CO) in carboxylic acids or esters, formed during the oxidative process of the GO material alongside hydroxyl and epoxy groups. In the SGO material, the C
O peak slightly shifted and reduced in intensity, indicating that sulfur doping partially decreased carboxyl groups or changed them into sulfonic acid or thiol groups. The peaks observed in the range of 1200–1250 cm−1 corresponded to C–O–C vibrations from epoxy and ether groups, which were prominent in GO due to the oxidative introduction of these oxygen-containing functionalities. In IL@SGO material, these peaks were almost absent, reflecting further chemical modifications during IL functionalization. The reduction or modification of these groups enhanced hydrophobicity and reduced oxygenated surface chemistry. Absorption bands between 3000 and 2800 cm−1 were characteristic of stretching vibrations from aliphatic C–H bonds. These bands were most intense in IL and IL@SGO, indicating the introduction of organic groups from the IL. The peaks in the range of 1000–1100 cm−1 were attributed to sulfur-oxygen bonds such as sulfoxides (S
O) or sulfates. These peaks were a clear indication of sulfur doping in SGO, where sulfur atoms are incorporated into the structure through chemical modifications. In the IL@SGO material, the intensity of these peaks increased, which showed that the IL functionalization stabilized the sulfur-containing groups or enhanced their vibrational signals. Peaks around 1600–1650 cm−1 were observed in the spectra of IL and IL@SGO, which were associated with the stretching vibrations of C
C and C
N bonds within the imidazole ring of the IL. A comparative analysis between GO and SGO revealed that sulfur doping reduced oxygen-containing groups such as hydroxyls, carbonyls, and epoxies. The decrease was evidenced by the decreased intensity of the –OH, C
O and C–O–C signals in the SGO material. The incorporation of sulfur likely replaced these groups with sulfur-containing functionalities, which improved the chemical reactivity.30 The FT-IR spectra demonstrated a systematic evolution of chemical functionalities from GO to SGO and finally to IL@SGO.31 The successful incorporation of sulfur and ILs suggests that IL@SGO is a multifunctional material with potential uses in catalysis, energy storage, and environmental remediation.32–34
Fig. 2 presents the X-ray Diffraction (XRD) patterns of five different materials, including graphite, GO (graphene oxide), SGO (sulfonated graphene oxide), IL (ionic liquid), and IL@SGO (IL encapsulated on sulfonated GO). The XRD pattern of graphite displayed many sharp and intense peaks, especially peaks at 2θ around 26.5°, 43.5° and 54.5° that correlate with the (002), (101), and (004) planes, respectively. The (002) peak position was the result of interlayer stacking carbon sheets of graphene over a hexagonal lattice along the c-axis, which was indicative of a high degree of crystalline and regular stacking of carbon layers. The resolution and pronounced characteristics of peaks confirmed that graphite was highly ordered, with high crystallinity and a low defect structure. This demonstrated that the GO presented substantial structural alterations from pristine graphite, with a sharp peak corresponding to the (001) plane at approximately 2θ ≈ 10–12°. The lower angles compared to the (002) peak of graphite suggest an expansion of interlayer. A weak shoulder around 2θ ≈ 43° might indicate the residual graphitic domains or partial graphitization. Compared to GO, the (001) peak observed in SGO was broader and less intense, revealing loss of structural order resulting from sulfonation. The additional inclusion of sulfonic acid (–SO3H) groups introduced defects and more disorder by breaking the regular stacking up even further. The wider peaks indicated decreasing crystallinity to give a more graphene-like arrangement in the material. The sulfonation procedure made SGO more hydrophilic and chemically reactive, therefore making it more applicable for catalytic and ion-exchange applications. In comparison to the other materials, the XRD pattern of IL was unique, with no sharp peaks and many broad background signals. This means that IL was not a crystalline material, therefore having no long-range periodic order. The general features seen here were typical for disordered molecular materials with ions located randomly throughout the structure. This fluidic phase was a characteristic feature of ILs due to the use of bulky, asymmetric cations and anions that impede the formation of a well-ordered regular lattice. The absence of crystalline features indicated that the material is mainly governed by weak van der Waals forces and ionic interactions that affect its thermal and physical properties. The IL@SGO composite material inherited the structural features of both SGO and IL, the pattern showing the broad features corresponding to the amorphous character of IL coupled with the low-intensity reflections corresponding to the partially crystalline nature of SGO. The presence of weak or fully suppressed peaks associated with the (001) and (002) planes confirmed strong interactions between the IL molecules and the SGO matrix. The disruption of stacking of GO layers could be further achieved with the encapsulation of IL within SGO, resulting in a more disordered structure of the hybrid material. A few weak diffraction peaks corresponded to the standard reference (JCPDS 00-041-1487), confirming that there are small crystalline phases, which may be due to impurities or partial crystallization in the synthesis process.
This image represented a Thermogravimetric Analysis (TGA) plot, which demonstrated the thermal stability and weight loss behavior of different materials as a function of temperature (Fig. 3). The current data referred to the thermal decomposition and stability of five materials, namely graphite, GO, SGO, IL, and IL@SGO. As can be seen, the graphite line had the most stable sample and shows almost no noticeable weight reduction during the heating period. Rather, it holds status for temperatures not exceeding 800 °C and even beyond. This implies that graphite can be used in extreme thermal applications and begins degenerating at above 800 °C, hence holding a minor weight loss, which can be assigned to the elimination of surface impurities or volatiles. The first weight loss in the TGA curves of GO observed at around 100 °C had been demonstrated as the absorption of water.35 GO began to show noticeable weight loss starting around 200–300 °C. This behavior was typically associated with the thermal decomposition of oxygen-containing functional groups, such as hydroxyl (–OH), carboxyl (–COOH), and epoxide (–C–O–C–) groups, which were present on the surface and edges of GO sheets.36 The loss of weight continued until about 400 °C, when approximately 60% of the weight remains at 800 °C, which implied the existence of thermally labile groups in GO, these remaining stable at elevated temperatures, except for the carbon backbone. The shape of the SGO curve followed that of GO very closely, but with slight improvement in thermal stability, which is attributed to sulfuric acid. The structural sulfonic acid groups introduced by sulfonation enhance the chemical bonds and improved the stability from the thermal degradation. The thermal decomposition of SGO also started at around 200 °C but was much slower than that of GO, while it has the highest residual weight compared to the others at 800 °C. IL showed the most violent and rapid weight loss compared to other samples. The decomposition began at a very low temperature (about 200 °C) and total weight loss is observed by 500 °C, unlike most of the ILs, which decomposed at moderate temperatures because of the rupturing of the ionic framework or any other organic component. The sharp weight drop indicated that the IL is primarily composed of volatile components that were prone to evaporation or degradation under heat. The hybrid material IL@SGO demonstrates a three-step thermal decomposition process: The first stage (∼100–200 °C) described the evaporation of water molecules, and the second stage (∼200–400 °C) corresponded to the decomposition of IL components, like the behavior observed in the IL sample. Stage three (∼400–700 °C) was ascribed to the heat degradation of the SGO backbone. The thermal stability of IL was also improved after incorporation into the SGO matrix in IL@SGO, as the weight loss was less notable than the IL alone. This phenomenon can be interpreted due to the interaction of the IL with the functional groups on SGO that limit the mobility and volatility of the IL molecules. Besides, based on the TGA results, the amount of ionic liquid attached can be predicted to be approximately 20%.
The SEM images and particle size distribution graphs offer a comparative interpretation of the structure of GO, SGO, and IL@SGO, which clearly represented the structural and morphological evolutions occurring during sulfur doping and ionic liquid modification (Fig. 4). As shown in Fig. 4, GO exhibited excellent dispersion and a large surface area due to its smooth, porous, and layered structures with low aggregation, in contrast to the rougher, wrinkled morphology of SGO. The addition of functional groups increased chemical reactivity and catalytic surface area. IL@SGO showed significant aggregation with many spherical particles forming on its surface, leading to a denser structure that may alter its conductivity and mechanical properties. Particle size analysis revealed that GO has the smallest size distribution (20–60 μm), indicating better dispersion, while SGO had a wider distribution with larger particles due to sulfur doping. IL@SGO had the broadest size range (20–100 μm) primarily due to ionic liquid-induced aggregation. Thus, while GO offered excellent dispersity for applications in adsorption, membrane technology, and sensors, the rougher texture of SGO made it suitable for catalytic reactions and energy storage. In contrast, the bulkier morphology of IL@SGO may enhance the supercapacitor performance, lubricants, or composite materials that required improved mechanical properties. This transition from GO to SGO to IL@SGO signified a trade-off between dispersion and aggregation, affecting their functional applications.
![]() | ||
Fig. 4 SEM images and particle size distributions of three components (graphite, GO, and SGO) and IL@SGO. |
The image provided in Fig. 5 presents the EDX (Energy-Dispersive X-ray) spectra and corresponding elemental compositions of four samples: graphite, GO, SGO, and IL@SGO. The spectra and compositional data highlighted the structural and chemical changes during the transformation of graphite into GO, SGO, and IL@SGO. The EDX spectrum of graphite showed a dominant carbon peak (85.91% by mass, 89.39% by atom), indicating purity, with minimal oxygen (13.05% by mass) and trace elements of Si, S, and Cl probably due to environmental contamination. Upon converting graphite to GO, the carbon content decreased significantly to 65.47% by mass (72.53% by atom), while the oxygen content increased to 31.04% by mass (26.12% by atom), indicating the formation of functional groups from oxidation. Small peaks for sulfur (2.75%) and chlorine (0.2%) were attributed to Hummers' method or oxidation processes. The increase in oxygen disrupted the sp2 hybridized structure, creating hydrophilic GO sheets. In IL@SGO, the carbon content further decreased to 26.04% by mass (32.38% by atom), with the oxygen content increasing to 68.87% by mass (64.29% by atom), suggesting enhanced functionalization. Sulfur peaks (1.51% by mass and 0.7% by atom) confirmed successful doping, while nitrogen (1.69% by mass and 1.8% by atom) indicated the presence of nitrogen-containing groups. IL@SGO showed carbon at 62.66% by mass (71.44% by atom) and oxygen at 25.1% by mass (21.49% by atom), with a higher sulfur fraction (4.33% by mass, 1.85% by atom) from the ionic liquid, along with increased chlorine (2.33%) and nitrogen (3.26%) contents due to functional groups from the IL. Overall, the transition from graphite to GO marked a shift from a carbon-rich structure to a highly oxidized material, initiating the functionalization process. Doping exemplified by the introduction of sulfur into SGO affords excellent doping, altering specific electronic and chemical properties of GO. The final modification with IL (IL@SGO) integrated additional functional groups, increasing sulfur and nitrogen contents while reducing excessive oxygen groups, resulting in a more balanced and functional material.37–39
![]() | ||
Fig. 5 Energy-dispersive X-ray (EDX) spectroscopy of three components (graphite, GO, and SGO) and IL@SGO. |
The EDX mapping results revealed the elemental distribution and compositions of the samples (Fig. 6). Graphite and GO were found to have a high homogeneous intensity in the mapping of C (C-K), highlighting the fact that these materials contain more carbon than any other element. However, SGO and IL@SGO caused a tiny decrease in the carbon intensity due to the entrapment of new functional groups. The existence of Oxygen (O-K) was evident in GO, indicating the oxide formation of graphite; however, its intensity was diminished in SGO and IL@SGO, which might indicate the partial replacement or reaction of oxygen with functional groups. The content of K (potassium) was low in graphite and GO, and the amount of K increased markedly in SGO and IL@SGO (ion-liquid-modified reduced graphene oxide), confirming successful N-containing functionalization. Silicon (Si-K) popped only in SGO and IL@SGO, indicating silane-based modification(s) being implemented. IL@SGO showed a higher amount of sulfur (S-K) and chlorine (Cl-K), which may be attributed to the presence of sulfur-derived ILs or chlorinated species during the functionalization process. Specifically, graphite was mainly carbon, whereas GO added oxygen, SGO added nitrogen and silicon, and IL@SGO is the most functionally rich material, containing nitrogen, silicon, sulfur, and chlorine. Such results indicated the increasing chemical modification in the studied samples, and the most abundant elemental composition belongs to IL@SGO.
The Raman spectrum offered a comprehensive insight into the sample's chemical structure and composition (Fig. 7). The presence of peaks within this wavenumber range suggested the existence of silicon–oxygen–carbon (Si–O–C) bonds. These bonds were indicative of organosilicon compounds or silicon-containing framework materials. For example, a significant peak at about 511 cm−1 was related to vibrations of silicon observed in some silicon carbide (SiC) coatings. The D band (∼1350 cm−1), corresponding to the breathing modes of sp2 carbon atoms in rings, was activated by defects or disorder in the carbon lattice. To quantify the degree of disorder, the intensity of the D band (ID) was compared to that of the G band (IG), resulting in the ID/IG ratio. This ratio served as a metric for assessing the level of structural imperfections. This means that a higher ID/IG ratio was suggestive of a higher amount of disorder or defects in the carbon framework. The band G (1600 cm−1) was associated with the in-plane stretching modes of sp2-hybridized carbon atoms. An obvious G band indicated graphitic or ordered carbon domains. The D and G band intensity ratio (ID/IG) typically correlated with the crystallinity and defect density of the carbon material. Based on previous research, the ratio intensity between D and G of graphene oxide was approximately 0.85.40 In this study, as shown in the Raman spectra, the ID/IG value was calculated to be about 0.85, which illustrated the more amorphous carbon. This can be explained by the existence of sulfur in the structure of catalyst through the synthesis process. Besides, to confirm the appearance of sulfur in the structure of catalyst IL@SGO, the amount of sulfur in the product of each step was determined by ICP-OES. Consequently, the amount of sulfur in graphene oxide preparation step was calculated at about 35376 mg g−1, while the values of sulfur at the sulfonated stage and ionic liquid at the grafted stage were about 307
449 mg g−1 and 424
832 mg g−1, respectively. The Raman spectrum and ICP-OES results of sulfur are strong evidence to prove the successful synthesis of the catalyst IL@SGO. The widespread peak in this area was characteristic of hydroxyl (–OH) groups (3000–3500 cm−1), which corresponded to hydroxyl functionalities sorbed on the surface of material or water molecules adsorbed onto it. These groups may affect the hydrophilicity and reactivity of the material. The lower ratio implied a more oriented graphitic structure, while a higher ratio indicates more defects or more amorphous carbon.
BET surface characterization is an important aspect to analyze the surface area and porosity of the synthesized material. The characteristics of a matter are very much affected by the surface area-to-volume ratio. The GO, SGO, and IL@SGO catalytic pores and specific surface area were assessed (Fig. 8). The plots of GO showed a type II adsorption isotherm, which was identified for non-porous or macroporous adsorbents exhibiting unrestricted monolayer-multilayer adsorption. The adsorption volume initially increased rapidly at low relative pressures when the adsorbate molecules engaged with the higher energy regions, subsequently interacting with the lower energy areas. After treating GO with acid, the plots of SGO demonstrated type I adsorption isotherm, which were applicable to adsorbents with extremely small pores or micropores. In this scenario, adsorption occurs through the filling of these micropores. IL@SGO illustrated that the quantity absorbed value remains at 0 for different relative pressures. It can be explained by the fact that grafting an ionic liquid reduces the porosity of the material. The catalyst surface area was measured to be 44.1 m2 g−1 for the GO catalyst and 135.1 m2 g−1 for the SGO catalyst, respectively. This suggested that the specific surface area of SGO and GO was much larger than that of IL@SGO, confirming the results of SEM characterization. The specific surface area of GO, SGO and IL@SGO was lower than the range for the theoretically monolayer graphene oxide reported in the literature that is from 2 to 1000 m2 g−1, probably due to the agglomeration of the graphene sheets, giving partial overlap and coalescence, especially between the smaller sheets, thus reducing the surface area of the material.41
An experimental study of the isolated yield of a reaction under different temperatures, reaction times, solvent types, solvent volumes, catalyst types, and catalyst loadings is summarized in Table 1. Temperature is one of the key parameters that governs reaction kinetics and thermodynamics. At room temperature (r.t.), the reaction yield was only 16% (Table 1, Entry 1). As the temperature increased to 80 °C, yield increased up to 32% (Table 1, Entry 3). Higher temperatures may accelerate reaction rates by providing enough energy to overcome activation barriers. Increasing the temperature to 100 °C reduced the yield to 30% and to 29% at 120 °C (Entries 4–5, Table 1). The decline may be due to thermal degradation, evaporation of volatiles, or side reactions. The data indicated that the highest yield occurred at 80 °C. At 80 °C, shorter reaction times of 5 and 6 hours yielded 27% and 28%, respectively (Entries 6 and 7, Table 1), suggesting that the reaction had not yet reached completion. By extending the time to 9 hours, the yield reached a maximum of 37% (Entry 9, Table 1). It is worth mentioning, however, that increasing the reaction time to 10 hours did not improve the conversion (36%, Entry 10, Table 1). These results indicated equilibration after 9 hours, and that extending the reaction time was unlikely to give improved yields. Minimizing the resource consumption while maximizing the yield is critical to scale-up, achieved via the optimization reaction, i.e. reaction time for efficient reaction. The solvent selection influences the surrounding reaction environment such as solubility, polarity, and stabilization of intermediates, and therefore, it is the key to the reaction outcome. The best solvent, toluene, yielded 37% (Entry 9, Table 1). This may be attributed to its moderate polarity and compatibility with the reaction mechanism. In contrast, polar solvents such as ethanol (3%, Entry 11, Table 1), THF (13%, Entry 12, Table 1), and water (1%, Entry 13, Table 1) yielded lower results due to probable repulsive interactions with the catalyst or reactants. Similarly, nonpolar solvents such as cyclohexane (2%, Entry 17, Table 1) and n-hexane (5%, Entry 18, Table 1) performed poorly compared to extreme polar solvents. Chemical transformations often need to be optimized to balance solubility, reaction kinetics, and selectivity; therefore, solvent optimization is an important part of such optimization. Catalyst loading is a crucial parameter as it directly affects the availability of active sites, thus governing the efficiency of the reaction. Using IL@SGO as the catalyst, the yield increased from 30% with 1 mg (Entry 29, Table 1) to 42% with 7 mg (Entry 32, Table 1). Beyond this point, further increasing the catalyst loading to 10 mg resulted in a slight decline to 41% (Entry 33, Table 1). The observed trend suggested that too much catalyst may result in aggregation and lead to a lower catalysis surface area or diffusion limitations. The optimum catalyst loading was also highly dependent on the concentration of the catalytic component, as the overuse of resources would not only affect the cost-effectiveness of the process but also make it unfeasible to achieve a high throughput, posing a high barrier to commercialization. Catalyst performance was evaluated at 80 °C for 9 h. The hybrid catalyst IL@SGO gave the highest yield of 37% (Entry 9, Table 1) compared to GO (25%, Entry 20, Table 1), SGO (32%, Entry 21, Table 1), and IL (31%, Entry 22, Table 1). While H2SO4 (Entry 23, Table 1), TsOH (Entry 24, Table 1), and HCl (Entry 25, Table 1) are common acid catalysts known to characterize similar sugar polymerization products (29–32%), they did not offer the synergistic advantages of IL@SGO. This superiority is due to the integrated functions of SGO (enhanced stability and surface area) and IL (enhanced reaction selectivity). The design of hybrid catalysts such as IL@SGO is a good strategy for improving the reaction efficiency. The volume of the solvent is one of the most significant factors that affects the yield of the main product. In particular, the yield increased significantly going from 2 to 4 mL of toluene (Entries 35 and 37, Table 1), ultimately affording a 41% yield. However, by increasing the volume to 6 mL, a slight decrease was observed (38%, Entry 38, Table 1). This means that overly dilute conditions suppress the reactant-catalyst collisions. Solvent volumes should be optimized to maintain most of the kinetic benefits of mixing and progress but avoid excessive dilution. Control experiments in the absence of a solvent (Entry 28, Table 1) or a catalyst (Entry 27, Table 1) yielded negligible results (5% and 14%, respectively). This highlights the importance of having both elements for the reaction to occur properly. These controls confirmed the pivotal role of IL@SGO and toluene in mediating the reaction mechanism. Additionally, the reactions between each pair of substrates were also carried out under the same optimal condition and checked by TLC for comparison of the main product 2,3-diphenyl-4-thiazolidinone to explain the incomplete conversion of the reaction. The yield of the reactions can be affected by some side products, which are formed through the reaction of each pair.42,43 The side reactions are mentioned in the ESI Scheme S2 and Section S2.†
Entry | Temp. (°C) | Time (h) | Solvents | Volume of solvent (mL) | Catalysts | Loading of catalyst (mg) | Isolated yieldb (%) |
---|---|---|---|---|---|---|---|
a Reaction conditions: benzaldehyde (1 mmol), aniline (1 mmol), and mercaptoacetic acid (1 mmol).b Isolated yield by crystallization in ethanol (5 mL).c The reactions were conducted by microwave. | |||||||
1 | RT | 8 | Toluene | 5 | IL@SGO | 5 | 16 |
2 | 60 | 21 | |||||
3 | 80 | 32 | |||||
4 | 100 | 30 | |||||
5 | 120 | 29 | |||||
6 | 80 | 5 | Toluene | 5 | IL@SGO | 5 | 27 |
7 | 6 | 28 | |||||
8 | 7 | 28 | |||||
9 | 9 | 37 | |||||
10 | 10 | 36 | |||||
11 | 80 | 9 | Ethanol | 5 | IL@SGO | 5 | 3(5)c |
12 | THF | 13 | |||||
13 | H2O | 1(1)c | |||||
14 | Ethyl acetate | 24 | |||||
15 | Acetonitrile | 1(1)c | |||||
16 | DMF | 2(7)c | |||||
17 | Cyclohexane | 2(1)c | |||||
18 | n-Hexane | 5(9)c | |||||
19 | None | 0 | 7(5)c | ||||
20 | 80 | 9 | Toluene | 5 | GO | 5 | 25 |
21 | SGO | 32 | |||||
22 | IL | 31 | |||||
23 | H2SO4 | 29 | |||||
24 | TsOH | 30 | |||||
25 | HCl | 32 | |||||
26 | H3PO4 | 22 | |||||
27 | None | 0 | 14 | ||||
28 | 80 | 9 | None | 0 | None | 0 | 5(5)c |
29 | Toluene | 5 | IL@SGO | 1 | 30 | ||
30 | 3 | 34 | |||||
31 | 5 | 37 | |||||
32 | 7 | 42 | |||||
33 | 10 | 41 | |||||
34 | 80 | 9 | None | 0 | IL@SGO | 7 | 15 |
35 | Toluene | 2 | 22 | ||||
36 | 3 | 40 | |||||
37 | 4 | 41 | |||||
38 | 6 | 38 |
The reaction conditions investigated were used to investigate the efficiency and utility of the synthesis procedure based on several substrates. The aim of this study was to evaluate the efficiency and versatility of the method in the synthesis of various 4-thiazolidinone derivatives using IL@SGO as the catalyst. The synthesis of 4-thiazolidinone derivatives was carried out between aldehydes (1.0 mmol), anilines (1.0 mmol), and mercaptoacetic acid (1 mmol) in the presence of toluene (5 mL) and IL@SGO (7 mg) at 80 °C for 9 h, and the results were recorded as acceptable to good (Scheme 3). These substituents, in a broad sense, can be classified as electron-donating groups (EDGs) and electron-withdrawing groups (EWGs), both of which significantly alter the electronic and steric environments of the reactions. EWGs include groups such as halogens (–F, –Cl, and –Br) and nitro (–NO2) that remove electron density from the aromatic ring, mainly via inductive effects. Under certain conditions, these groups can stabilize transition states or intermediates through delocalization of negative charges, thereby enhancing the reactivity. However, if one introduces a strong deactivating group, then one can do oxidization of the ring, and the reaction will cease as with the example above. Electron-withdrawing groups (EWGs) such as –NO2 withdraw the electron density from the aromatic system through inductive or resonance effects. Groups that donate electrons to the aromatic system increase the nucleophilicity of the ring and promote reactivity. However, steric hindrance caused by these groups, especially at the ortho position, can hinder accessibility to reactive sites, thereby cutting yields. This duality often results in varied and position-dependent impacts on yields. Halogens are widely known for their moderate electron-withdrawing properties and relatively small steric footprints. As can be seen from Scheme 3, halogen-substituted compounds (4ca, 4da, and 4ea) displayed yields ranging from 38% to 41%. These results illustrated the trade-off between electronic stabilization and steric hindrance (38% yield of compound 4ca). This small size of fluorine reduces steric hindrance, but some strong electronegativity of intermediates, making the reactivity lower than that of other larger sizes such as chlorine and bromine. The yield obtained with compound 4da was 41%, slightly better than what was obtained with fluorine. Chlorine's larger size and weaker electronegativity than fluorine give a better balance between steric and electronic effects. For example, the proposed compound, 4ea, gave a yield of 40% like chlorine. Bromine has an even larger atomic radius and more steric bulk and also greater polarizability, which may stabilize transition states. Nitro (–NO2) groups are among the most potent electron-withdrawing substituents due to their capacity to delocalize negative charge via resonance and inductive effects. The moderate reactivity of 4fa (31% yield) was due to the para-positioned nitro group, which reduced steric hindrance. However, the yield was lower than that of halogenated derivatives because of the strong electron-withdrawing effect. The ortho-positioned nitro group in 4ac (10% yield) caused a lot of steric hindrance, and the fact that it pulled electrons away from the molecule makes the yield much lower. The methyl (–Me) group was a weak electron-donating group, contributing electron density via its inductive effect. It turned out that the para-positioned methyl group had the highest yield (49%), which means that its electron-donating properties made it more reactive without creating any steric problems. Positional differences (e.g., ortho) may diminish yields due to steric obstruction. The methoxy (–OMe) group was a potent electron-donating group, providing electron density via resonance while exerting inductive withdrawal. The reaction was almost non-existent for methoxy-substituted compounds indicate that steric hindrance at the ortho position or excessive electron donation may interfere with the reaction process. Positional isomerism (ortho, meta or para) plays an important role in influencing the yield of products through the stereochemical and electron-donating effects of the reaction sites. When substituents are in the ortho position of compounds, they often face a lot of stereochemical hindrance, which makes them less reactive. This is why the product yields were low at ortho-NO2 and ortho-OCH3 positions. When aldehydes were used to synthesize compounds with substituents at the para position on the arene ring, there was less steric hindrance, which means that more of the product was formed. However, the electronic effect was stronger, which achieved a high yield of about 49% with 4ha. When anilines were added to products, the substituents were put in the ortho, meta, and para positions behind the amino group. This is because of the electronic effects and a strong steric hindrance effect, which made the amino group much less reactive.
Fig. S1† presents the FT-IR spectra comparing IL@SGO and its recovery process. In the IL@SGO recovery spectrum, the decline in O–H band at 3200–3600 cm−1 suggested a partial loss of hydroxyl groups. Furthermore, the decline in the intensity of aliphatic and aromatic C–H stretching vibrations indicated that the IL may have experienced partial removal or decomposition. A more intense peak in the IL@SGO spectrum suggested stronger IL-GO interactions, while shifts and reduced intensity in the recovery spectrum imply structural changes like esterification or chemical degradation. The peaks associated with CC and C
N stretching (1500–1600 cm−1) were assigned to the aromatic rings and imine groups, with the decline in these peaks hinting at disrupted IL-GO π–π stacking interactions. Additionally, a notable drop in S
O band's strength in the recovery spectrum suggested sulfur group separation, implying instability of the IL under these conditions. The fingerprint area (500–1500 cm−1) showed notable differences between the two spectra, with the IL@SGO spectrum displaying more complex vibrations than the less intense IL@SGO recovery spectrum, indicating a loss of functional variety and structural breakdown of the IL-GO composite during recovery. The shifts and intensity reductions in characteristic peaks (–OH, C
O, C–H, S
O) suggested that the recovery process adversely affects IL-GO interactions, leading to structural changes and loss of functional groups, particularly oxygen-containing ones.
Fig. S2† shows the TGA (Thermogravimetric Analysis) curves for IL@SGO (red line) and IL@SGO recovery (blue line). Both materials experience a minor weight loss below ∼150 °C due to evaporated moisture, indicating similar surface moisture levels. However, IL@SGO recovery showed faster weight loss, particularly beyond 300 °C, suggesting reduced thermal stability from structural changes during recovery. Above ∼500 °C, the weight loss indicated breakdown of stable elements such as sulfur and carbon. IL@SGO maintains a higher residual weight, indicating greater stability and possibly higher carbon content than the recovered material. This illustrated how recovery affected thermal stability: IL@SGO demonstrated greater resistance and slower decomposition, while recovery may compromise structural integrity and thermal behavior. Both materials remained stable up to ∼300 °C, making them suitable for low-temperature applications.
Fig. S3† presents the SEM (Scanning Electron Microscopy) images and corresponding particle size distribution histograms for IL@SGO (top row) and IL@SGO recovery (bottom row). In contrast, the SEM image of recovered IL@SGO showed a less porous, more aggregated structure with smoother surfaces and less distinct particles, indicating partial collapse of the porous framework during recovery. The original IL@SGO catalyst had a peak particle size distribution in the 40–60 μm range, indicating homogeneity, while the recovered IL@SGO peaks at 60–80 μm reflect a broader size distribution due to particle agglomeration, likely from thermal or mechanical factors. This process decreased porosity and increased aggregation, potentially compromising the material's catalytic performance, as larger particles from the IL@SGO recovery may reduce the dispersion efficiency.
EDX spectroscopy revealed changes in the chemical composition of IL@SGO before and after recovery (Fig. S4†). After recovery, the C content decreased from 62.66 to 42.35%, suggesting loss or degradation of organic materials, while the O content increased, indicating possible oxidation or rearrangement. Notably, the S content dramatically increased from 4.33% to 25.42%, suggesting the adsorption of sulfur-rich species during recovery. The Si concentration dropped slightly to 1.02%, potentially indicating the partial dissociation of the GO structure, while the Cl levels remained unchanged. Overall, the recovery process led to a significant decrease in carbon content and a marked increase in sulfur content.
The microscopic images of IL@SGO before and after recovery revealed notable structural changes; IL@SGO seems to be quite porous with unequal particle distribution, which qualified for adsorption or catalysis uses (Fig. S5†). After recovery, the intensity and homogeneity of C and O decreased, aligning with organic material loss, while S became strong and localized, indicating its build-up. Cl remained less prominent but showed some intensity change, and Si followed a similar trend, suggesting decreased GO separation. Post-recovery, sulfur prominently formed localized areas, while carbon distribution diminished, confirming partial rearrangement of the IL and GO structures along with sulfur adsorption.
![]() | ||
Fig. 10 Procedure for leaching tests: (I) reaction stopped; (II) reaction without IL@SGO; and (III) continuous reaction with IL@SGO. |
In section (I), IL@SGO was separated from the reaction mixture, and the product solution was subjected to crystallization in 5 mL of ethanol. Subsequently, the main product was obtained with a yield of approximately 10%, and the sulfur content retained in the catalyst was determined to be 2.208 mmol g−1.
In section (II), IL@SGO was separated from the reaction mixture, and the reaction proceeded with continuous stirring for the remaining 4.5 hours. Upon completion, the reaction mixture was crystallized in 5 mL of ethanol, yielding approximately 12% of the main product. At this stage, ICP-OES analysis confirmed the absence of sulfur leaching from the catalyst into the solution.
In section (III), the reaction mixture was allowed to continue for the remaining 4.5 hours. After the reaction was completed, IL@SGO was removed, and crystallization of the reaction mixture was carried out in 5 mL of ethanol. For the remaining reaction under optimal conditions, the yield of the main product was about 40% and the content of sulfur which remained in the catalyst was determined to be about 3.089 mmol g−1.
The study also assessed the environmental sustainability of the process using green chemistry metrics. IL@SGO-based catalysis exhibited high atom economy (89.28%) and a low environmental factor (E-factor = 1.28), supporting its potential as an eco-friendly alternative to organic synthesis. However, moderate carbon efficiency (49%) and reaction mass efficiency (43.74%) indicated room for further optimization. IL@SGO proves to be an effective, reusable, and environmentally sustainable catalyst for the synthesis of 4-thiazolidinones, contributing to the advancement of green chemistry. Future research may be focused on improving its stability and optimizing the reaction conditions to enhance the efficiency and sustainability.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra02881g |
This journal is © The Royal Society of Chemistry 2025 |