Next Article in Journal
A Cox Model-Based Workflow for Increased Accuracy in Activity-Travel Patterns Generation
Previous Article in Journal
Surgical Treatment of Distal Radius Fractures Using Minimally Invasive Plate Osteosynthesis or Open Reduction and Internal Fixation: A Five-Year Comparative Follow-Up Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Experimental Study of Glycerol Carbonate Synthesis over g-C3N4 Catalysts

by
Mirna Lea Charif
1,
Dragoș Mihael Ciuparu
1,
Ioana Lavinia Lixandru Matei
2,3,
Gabriel Vasilievici
4,
Ionuț Banu
5,
Marian Băjan
1,
Dorin Bomboș
6,*,
Cristina Dușescu-Vasile
1,*,
Iuliana Veronica Ghețiu
7,
Cașen Panaitescu
8 and
Rami Doukeh
7
1
Department of Petroleum Refining Engineering and Environmental Protection, Petroleum-Gas University of Ploiesti, 39 Bucharest Blvd., 100680 Ploiesti, Romania
2
Botanical SRL, 7 Trandafirilor Street, 107059 Ploiesti, Romania
3
Faculty of Chemical Engineering and Biotechnologies, National University of Science and Technology POLITEHNICA Bucharest, 313 Splaiul Independentei, 060042 Bucharest, Romania
4
National Institute for Research Development for Chemistry and Petrochemistry-ICECHIM-București, 202 Spl. Independenței, 060021 Bucharest, Romania
5
Department of Chemical and Biochemical Engineering, National University of Science and Technology POLITEHNICA Bucharest, 060042 Bucharest, Romania
6
Chemistry Department, Petroleum-Gas University of Ploiesti, 39 Bucharest Blvd., 100680 Ploiesti, Romania
7
Department of Well Drilling, Extraction and Transport of Hydrocarbons, Petroleum-Gas University of Ploiesti, 39 Bucharest Blvd., 100680 Ploiesti, Romania
8
Department of Petroleum Geology and Reservoir Engineering, Petroleum-Gas University of Ploiesti, 39 Bucharest Blvd., 100680 Ploiesti, Romania
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2025, 15(11), 6236; https://doi.org/10.3390/app15116236
Submission received: 25 April 2025 / Revised: 29 May 2025 / Accepted: 30 May 2025 / Published: 1 June 2025

Abstract

This study examines a catalyst based on graphitic carbon nitride (g-C3N4) for synthesizing glycerol carbonate through the coupling reaction of glycerol and CO2. In this research, we focus on simultaneously improving CO2 emission reduction and glycerol valorization by co-doping g-C3N4 with phosphorus (P), sulfur (S), magnesium (Mg), and lithium (Li) for a better catalytic performance. The catalysts were prepared through a one-step thermal condensation process and characterized using XRD, SEM, TGA, FTIR, and nitrogen adsorption–desorption techniques. The co-doping further enhanced the surface chemical properties, Lewis acidity, basicity, and thermal stability, evidenced by the lower crystallinity, wider pore, and better catalytic performance as assessed through glycerol carbonylation reaction, optimized using a Box–Behnken design. The MgPSCN catalyst exhibited the highest glycerol conversion (68.72%) and glycerol carbonate yield (44.90%) at 250 °C, using 50 mg catalyst and 10 bar pressure. The model accuracy was validated by ANOVA (R2 > 0.99; p values < 0.0001). The results indicated that doping significantly enhanced the catalytic performance, most likely due to improved electron charge transfer and structural distortions within the g-C3N4 framework. Such a process highlights the potential of co-doped g-C3N4 catalysts for the sustainable glycerol utilization and valorization of CO2 through a scalable pathway toward green chemical synthesis—an approach that comes in line with worldwide decarbonization goals.

1. Introduction

In 2023, global carbon dioxide emissions related to energy rose by 1.1%, setting a new peak at 37.4 gigatons—moving further away from the decline needed to align with the climate targets outlined in the Paris Agreement [1]. To mitigate these emissions, two primary technological approaches are employed: capturing CO2 and either storing it or repurposing it. Forecasts indicate that by 2050, approximately 5.6 gigatons of CO2 will need to be captured and securely stored each year [2]. Nevertheless, transforming CO2 into useful chemical products presents a more cost-effective alternative compared to underground sequestration [3]. A key strategy in emission reduction involves the adoption of renewable sources, such as biodiesel derived from biomass. This option is gaining traction due to its favorable properties—including being non-toxic, renewable, and free of sulfur and aromatic compounds [4].
As a byproduct of biodiesel production, glycerol has become increasingly surplus due to the expansion of biodiesel applications. Given the need for sustainable development, converting glycerol into high-value chemicals has become a challenge for researchers [5,6]. Glycerol is highly valued for its extensive use across various industrial sectors, including pharmaceuticals, personal care products, the food industry, and cosmetics [7]. Additionally, it shows significant potential as a primary chemical reactant in synthesizing various organic molecules. Glycerol can serve as a renewable feedstock in numerous chemical processes, such as carbonation, carbonylation, transesterification, and oxidation. Among these transformations, the synthesis of glycerol carbonate is particularly compelling from an environmental perspective as it utilizes CO2 as both a carbon source and a C1 feedstock. This approach is pivotal in advancing decarbonization efforts and accelerating progress toward achieving net-zero emission targets [2,8].
Glycerol carbonate is emerging as a promising platform chemical and an eco-friendly alternative to traditional petrochemical-based substances. Its applications are diverse, serving as a solvent in detergents and paints, and playing a crucial role in producing valuable intermediates like glycidol which are used in textiles, plastics, pharmaceuticals, and cosmetics [9]. Additionally, glycerol carbonate is utilized in lithium-ion battery electrolytes, surfactants, and as a general-purpose solvent [10]. Among the various synthesis routes, the direct reaction between glycerol and CO2 is the most environmentally favorable as it exclusively produces the target compound without unwanted byproducts. However, this method has not yet been implemented on an industrial scale due to challenges such as low conversion efficiency, limited catalyst stability, and thermodynamic constraints of the reaction. These issues may be mitigated by optimizing reaction pathways and using water-removing agents to shift the equilibrium towards product formation [3,11,12].
To enhance the synthesis of glycerol carbonate, a variety of heterogeneous catalytic systems have employed dehydrating agents such as acetonitrile, benzonitrile, 2-cyanopyridine, adiponitrile, dimethylformamide (DMF), calcium oxide (CaO), and magnesium carbonate (MgCO3) [13,14,15,16,17,18]. These systems include titanosilicate-based ETS-10 zeolite catalysts, which are modified by impregnating active metals such as Co, Zn, Cu, Ni, Zr, Ce, and Fe [19]. Cerium oxide (CeO2) catalysts have also been synthesized through precipitation, hydrothermal techniques, and citrate sol–gel methods [3,15]. In another approach, finely distributed MgInCe mixed metal oxides (MgInCe-MMOs) were synthesized via a topological transformation from their layered double hydroxide precursors (MgInCe-LDHs) [20]. Hydrothermal synthesis was also used to fabricate a range of Cu1−xZrxO2 bimetallic oxides with varying Cu-to-Zr ratios [21]. ZnAlCe catalysts were obtained by co-precipitation [22], while gold nanoparticles (AuNPs) were incorporated into a metal–organic framework (MOF) constructed from mixed carboxylate linkers, including terephthalic acid and 1,3,5-benzenetricarboxylic acid [23]. La2O2CO3/ZnO catalysts were created through both co-precipitation and ethylene glycol combustion methods [24], and metal oxides such as MgO, CaO, and Al2O3 were synthesized via precipitation [25]. Finally, phosphorus and sulfur co-doped graphitic carbon nitride (g-C3N4) was produced through a one-step thermal condensation process [26].
Graphitic carbon nitride (g-C3N4) has drawn considerable interest for its potential to tackle energy and environmental issues. Its electron-rich framework, surface basicity, and hydrogen-bonding capabilities—stemming from its nitrogen and hydrogen constituents—render it a versatile companion to carbon in various advanced material applications [27]. As a result, g-C3N4 plays a prominent role in numerous catalytic processes, including carbon dioxide reduction [28], oxygen reduction reactions (ORRs) [29], hydrogen production from aqueous glycerol solutions [30], and hydrogen evolution reactions [31]. However, the performance of unmodified g-C3N4 is hampered by several inherent limitations such as (i) poor electrical conductivity, (ii) weak interfacial compatibility with other materials, (iii) a limited number of catalytically active sites combined with low Lewis acidity and basicity, and (iv) an inefficient electronic structure that hinders charge transport and reduces its catalytic efficiency [26].
To overcome the inherent drawbacks of pristine g-C3N4, scientists investigated various modification strategies, including elemental-doping and substitution. Non-metal dopants such as boron, oxygen, phosphorus, and sulfur have been included in the support network to boost its physicochemical properties [28,32,33]. Additionally, transition metals like iron, manganese, titanium, and nickel have been used to enhance its catalytic performance [34,35,36,37,38,39]. Alkaline and alkaline earth metal oxides—including lithium, barium, calcium, and magnesium—have also been introduced into g-C3N4 structures to further boost activity [40,41,42]. These modifications collectively aim to refine the material’s surface morphology, chemical reactivity, and electronic behavior while also shedding light on the nature and distribution of active catalytic sites.
While the application of modified g-C3N4 materials in photocatalytic processes has been extensively studied, their use in catalyzing the transformation of glycerol into glycerol carbonate remains relatively underexplored. Fao et al. [26] analyzed a phosphorus and sulfur co-doped g-C3N4 catalyst (PSCN) using a one-step thermal condensation method and assessed its catalytic activity for converting glycerol and CO2 into glycerol carbonate. The reaction, undertaken at 210 °C using acetonitrile as a dehydrating agent, yielded an 85% glycerol conversion with PSCN, significantly higher than the 60% achieved using unmodified g-C3N4 (CNN). In another study, Reisi and Najafi Chermahini [43] investigated MgO-supported g-C3N4 catalysts for the transesterification reaction between glycerol and dimethyl carbonate (DMC). The optimal catalyst, with a MgO to g-C3N4 weight ratio of 6:1, delivered a maximum glycerol carbonate yield of 95.47%. This performance was attributed to its strong basicity and favorable conditions, including a reaction temperature of 80 °C, a 4 h duration, a 3:1 DMC-to-glycerol molar ratio, a 30 mg catalyst dose, and excellent reusability.
The main goal of our work was to compare the catalytic performance of a series of catalysts obtained by doping g-C3N4 with phosphorus (P), sulfur (S), magnesium (Mg), and lithium (Li) in the glycerol carbonate synthesis. To the best of our knowledge, there is no study in the literature comparing all these promoters in the reaction of glycerol carbonylation reaction. Also, the multi-element doping strategy represents a distinct and innovative approach to enhancing both the structural and electronic properties of g-C3N4, aiming to improve its catalytic activity in the direct coupling reaction of glycerol with CO2. The synthesized catalysts were characterized using a variety of techniques to thoroughly evaluate their properties. Additionally, an experimental design approach was employed for identifying the best operating conditions to maximize the yield in glycerol carbonate. This comprehensive characterization and systematic evaluation aimed to enhance the efficiency and effectiveness of the glycerol carbonate production process.

2. Experimental

2.1. Materials and Methods

Melamine (99%, Alfa-Aesar, Ward Hill, MA, USA), phosphonitrilic chloride trimer (98%, Alfa-Aesar), sulfur powder (99.5%, Alfa-Aesar), Mg(NO3)2·6H2O (Merck-Sigma-Aldrich, BioUltra, Burlington, MA, USA, ≥99.0%), LiNO3 (Merck, 99.99%), carbon dioxide CO2 (SIAD Romanian, Bucharest, Romania, purity: ≥99.5%), glycerol (Gly) (Merck, ≥99.5%), acetonitrile (Amex-lab, New York, NY, USA, ≥99.9%), and glycerol carbonate (GC) (Merck-Sigma-Aldrich, ≥99%) were acquired.

2.2. Catalysts Preparation

In our study, the catalysts were synthesized through a one-step thermal condensation process [44] to improve their Lewis acidic and basic properties [26]. Bare g-C3N4 (denoted as CN) was prepared by direct thermal treatment of melamine at 550 °C at a heating rate of 2 °C/min for 5 h. The g-C3N4 catalyst co-doped with phosphorus (P) and sulfur (S) was synthesized by mixing and grinding 250 g of melamine, 25 g of phosphonitrilic chloride trimer, and 25 g of sulfur powder. This mixture was then subjected to thermal treatment in a furnace at 550 °C for 5 h, with a heating rate of 2 °C/min. After cooling, the sample was washed with a 1:2 volumetric mixture of deionized water and ethanol, and subsequently dried at 80 °C overnight. The product was labeled as PSCN.
For comparative purposes, additional catalysts—g-C4N3 co-doped with S, P, and Mg, and another variant co-doped with S, P, and Li—were prepared using the same method. Specifically, 26.5 g of Mg (NO3)2 and 25 g of LiNO3 were added to the melamine, S, and P mixture during the mixing and grinding process to introduce 2.5 g of Mg and Li metals into the catalysts, respectively. We designated the resulting catalysts as PSCN, MgPSCN, and LiPSCN, respectively.

2.3. Catalyst Characterization Equipment

The catalyst was analyzed using various techniques, such as FTIR spectroscopy, X-ray diffraction (XRD), textural analysis (N2 adsorption/desorption), scanning electron microscopy (SEM), and thermogravimetric analysis (TGA-DTG). XRD analysis was carried out at ambient temperature using a Bruker X-ray diffractometer (Karlsruhe, Germany; θ-θ type) equipped with a Cu-Kα radiation source (λ = 1.5418 Å) operating at 40 kV and 5 mA. The scans covered a 2θ range of 5–80° at a rate of 10°/min. Microstructural morphology was examined using a SEM instrument from FEI Company (Hillsboro, OR, USA). Thermogravimetric and derivative thermogravimetric analyses (TGA-DTG) were performed on a METTLER TOLEDO TGA/IST Thermal Analysis System (Greifensee, Switzerland) within the temperature range of 25–600 °C, under a nitrogen atmosphere, and at a heating rate of 10 °C/min.
The textural properties of the samples were evaluated using a Quantachrome NOVA 2200 e Gas Sorption Analyzer (Boynton Beach, FL, USA). The nitrogen adsorption/desorption isotherms were recorded at 77.35 K across a relative pressure (p/p0) range of 0.005 to 1.0. Data processing was carried out using NovaWin version 11.03 software. The specific surface area was calculated using the BET (Brunauer–Emmett–Teller) method, while the total pore volume was estimated from the desorbed volume at a relative pressure near unity, utilizing the BJH (Barrett—Joyner—Halenda).
FTIR spectroscopy was performed using a Shimadzu IRAffinity-1S spectrophotometer (Kyoto, Japan) equipped with a GladiATR-10 accessory, allowing spectra acquisition through the ATR (attenuated total reflectance) technique. FTIR spectra were captured across a wavenumber range of 400–4000 cm−1, with a spectral resolution of 4 cm−1.

2.4. Procedure for Catalytic Test in the Glycerol Carbonylation Reaction

A mixture of 25 g of glycerin and 50 mL of acetonitrile, used as a dehydrating agent to shift the unfavorable equilibrium towards the product side and improve the selectivity for glycerol carbonate [13,45], was added in the reactor. The catalyst was added at room temperature and then the temperature gradually increased to the desired value while the glycerin–catalyst–acetonitrile mixture was mechanically stirred. Subsequently, carbon dioxide gas was introduced until the desired pressure was reached. The reaction was allowed to proceed for 4 h.
After each experiment, the autoclave was cooled to ambient temperature, depressurized, and the liquid phase was isolated from the catalyst through centrifugation. Further, the catalyst was rinsed with an ethanol–deionized water mixture, then dried at 80 °C overnight. The liquid phase was analyzed by gas chromatography flame ionization detector (GC-FID, Varian CP-3800, Varian. Inc., Walnut Creek, CA, USA) equipped with a ZB-5ms column (L = 30 m, D = 250 μm, d = 0.25 μm) and a carrier gas (N2) flow rate of 1 mL/min. To determine the concentrations of glycerol carbonate a calibration curve was constructed using acetonitrile as the solvent. The curve was generated over a concentration range of 1 to 90% by weight and is provided in one of our recently published papers [46].
The yield was evaluated by the following equation:
% Yield   of   GC = moles   GC moles   GLY   introduced

3. Results and Discussions

3.1. Catalyst Characterization

The structure and phase identification of the samples, including bare g-C3N4 (CN), g-C3N4 doped with P and S (PSCN), g-C3N4 doped with P, S, and Mg (MgPSCN), and g-C3N4 doped with P, S, and Li (LiPSCN), were determined by using X-ray diffraction analysis.
The XRD patterns for these materials are given in Figure 1. Two prominent peaks with high intensity were observed at 2θ ≈ 12.24° and 27.32°, which correspond to the diffraction planes of bare g-C3N4. These diffraction values agree with the data in the literature [47,48,49,50]. The pronounced peak at 2θ ≈ 27.24° corresponds to the (002) diffraction plane, attributed to the interlayer stacking of conjugated aromatic structures, with an interlayer distance of 0.324 nm. The lower-intensity peak at 2θ ≈ 12.24° corresponds to the (100) diffraction plane, indicating the in-plane structural arrangement of tri-s-triazine units with an interlayer distance of 0.680 nm [26,28,32].
For the samples doped with P and S, the intensity of the (002) diffraction peak decreased, indicating a reduction in crystallinity compared to bare g-C3N4 likely due to the disruption of crystal growth by the addition of P and S. The intensity of the (100) peak also diminished, suggesting that the incorporation of these dopants distorted the formation of nitride pores and altered the interlayer distance, indicating strong interactions between the dopants and the host g-C3N4 matrix [28,51].
The g-C3N4 samples doped with Li and Mg exhibited similar diffraction peaks to those of bare g-C3N4, indicating that the overall structure of g-C3N4 remained largely unchanged by Li and Mg doping [52]. Additionally, no separate diffraction peaks corresponding to Li or Mg were detected in the MgPSCN and LiPSCN samples, confirming that Li and Mg were successfully doped into the g-C3N4 framework [40].
After the addition of Li and Mg the peak intensities for both the (002) and (100) diffraction planes decreased, suggesting that the presence of these dopants further reduced the crystallinity of the material [40]. Moreover, the distortion of nitride pores and the alteration of the interlayer distance further indicate strong interactions between the dopants and the host g-C3N4 [52,53].
The average crystallite size (D) of the catalysts was determined using the Debye–Scherrer Equation (1), which is as follows [54]:
D = k λ β cos θ
where λ is the wavelength of the X-rays used (Cu Kα = 0.15406 nm), β is the full width at half maximum (FWHM) of the diffraction peak, θ is the Bragg diffraction angle, and k is the shape factor, typically assumed to be 0.98. The average crystallite sizes of CN, PSCN, MgPSCN, and LiPSCN, calculated using this equation, were found to be 10.94, 17.57, 10.71, and 6.86 nm, respectively (Table 1).
The full X-ray photoelectron spectroscopy (XPS) survey of the MgPSCN catalyst before use, as illustrated in Figure 2, is evidence of the presence of C, N, O, P, S, and Mg elements in MgPSCN. The observed peaks at 285 eV, 395 eV, 529 eV, 136 eV, 166 eV, 49.7 eV, and 1303 eV correspond to C1s, N1s, O1s, P2p, S2p, Mg2p, and Mg1s, respectively [52,55].
The peak at 285 eV is attributed to the sp2 hybridization of C=C bonds, while the peak at approximately 136.0 eV corresponds to P–N/P–O bonds. Additionally, peaks in the sulfur spectrum at 166 eV and 228 eV were observed, corresponding to S2p and S1s, indicating the incorporation of sulfur atoms (S-doped) in nitrogen sites within the structure [56].
The peak at 395 eV suggests the presence of tri-coordinated nitrogen in N-(C)3 groups. For the Mg 1s spectrum, the peak at 1303 eV is evidence of the presence of Mg(II) ions. In the Mg 2p spectrum, the peak at 49.7 eV corresponds to elemental magnesium (Mg0) [40,52].
After two cycles of use under optimal reaction conditions, no significant changes were observed in the spectral peaks except for the disappearance of the peak at 166 eV, which corresponds to S2p, suggesting a potential loss of sulfur atoms during the reaction. These findings confirm the stability of the catalyst, which was further supported by microscopic analysis.
Figure 3 presents SEM images of the porous samples, highlighting distinct morphological characteristics. The morphology of bare g-C3N4 (CN) (Figure 3A) exhibits an irregular granular and layered structure with numerous pores, as also shown in the previous literature [28,57]. This porous structure is attributed to the liberation of ammonia (NH3) and carbon dioxide (CO2) gases during the thermal decomposition of melamine [44]. In contrast, CN doped with phosphorus and sulfur (PSCN) (Figure 3B) demonstrates a thinner, plate-like structure with a higher surface porosity compared to undoped g-C3N4. A significant difference is observed in the layer distribution and pore size, with the pores in PSCN increasing in dimension as confirmed by textural analysis.
Doping with magnesium, phosphorus, and sulfur (MgPSCN) (Figure 3C) results in the formation of numerous grains with varying nano-scale dimensions, displaying smaller diameters than those in PSCN. Additionally, notable distortions in the layered morphology are evident. In contrast, lithium-doped g-C3N4 (LiPSCN) (Figure 3D) shows a less uniform and smaller layered structure compared to PSCN, although the LiPSCN sample remains more uniform, larger in size, and more compactly bonded than MgPSCN [40].
The elemental mapping results of the samples showed well-defined distributions for C, N, P, S, Mg, and Li, as indicated in Figure 4. The elemental analysis indicates the presence of C and N in CN samples, with a C/N ratio of 0.48%. Upon doping with phosphorus (P) and sulfur (S), an increase in the carbon-to-nitrogen ratio to 0.80% was observed in the PSCN sample. Similarly, for the MgPSCN and LiPSCN catalysts, the carbon-to-nitrogen ratios increased compared to the CN catalyst, with values of 0.663% and 0.551%, respectively, when doped with magnesium (Mg), phosphorus (P), and sulfur (S). This suggests that deamination has taken place due to the co-doping of Mg and Li in the pure g-C3N4 matrix [40]. The above discussion further supports the successful co-doping of pure g- C3N4 with P, S, Mg, and Li.
To analyze the specific surface area and pore structure of the samples, nitrogen adsorption–desorption measurements were conducted. As illustrated in Figure 5, all samples display type-IV isotherms, which are characteristic of mesoporous materials according to the IUPAC classification, indicating the presence of well-defined mesopores [57,58]. The hysteresis loop observed at a high relative pressure (P/P0) of 0.5–1.0, categorized as type H3, suggests the presence of slit-shaped pores associated with plate-like particles [59]. This characteristic is consistent with the layered morphology also observed in the SEM analysis. Similar results have been reported in previous studies [32,40,53]. Additionally, variations in N2 adsorption volumes indicated differences in surface area among the samples. Table 1 summarizes the surface area, pore volume, and pore width of the synthesized catalysts. A decrease in the specific surface area was observed for PSCN compared to CN, from 10.85 to 10.10 m2/g, with an even further reduction to 9.19 m2/g in the case of MgPSCN, which is consistent with previous studies [26,28,40]. In contrast, LiPSCN exhibited a significant increase in surface area, reaching 17.32 m2/g. The comparable specific surface area of these samples did not have a significant impact on their catalytic activity. Furthermore, a substantial increase in pore width was observed, approximately 6-fold for both PSCN and LiPSCN compared to CN, while MgPSCN showed an increase of approximately 3-fold.
The thermal stabilities of CN, PSCN, MgPSCN, and LiPSCN samples were assessed using thermogravimetric analysis (TGA) under a nitrogen (N2) atmosphere, as shown in Figure 6. An initial weight loss of 5%, 6.1%, 7.4%, and 6.5% was observed for CN, PSCN, MgPSCN, and LiPSCN, respectively, up to 200 °C. This weight loss is attributed to the evaporation of water molecules adsorbed physically and chemically, along with the removal of hydroxyl groups [60,61]. In the temperature range of 200–400 °C, CN exhibited a minor weight loss of approximately 1%. However, doping with sulfur and phosphorus in PSCN led to an increased weight loss of around 3%, which can be explained by the lower thermal stability of the P–O–C bond compared to the C–C bond, thus reducing the initial decomposition temperature [62]. For MgPSCN and LiPSCN, the weight loss in this temperature range increased to 3.5% and 8%, respectively. This is likely due to the enhanced thermal conductivity of metals and their oxides, facilitating more uniform heat transfer and promoting the catalytic decomposition of PSCN in this range [62]. The effect is particularly pronounced in LiPSCN between 400 and 600 °C, where a significant weight loss of 23.6% was recorded, compared to 15.2% for MgPSCN and 14.5% for PSCN in the same temperature range. Therefore, based on the previous results, PSCN and MgPSCN are considered to be more thermally stable than LiPSCN.
The FTIR spectrum of the catalysts, particularly for the CN catalyst shown in Figure 7, exhibits broad peaks observed near 3000–3600 cm−1, corresponding to N–H stretching vibrations. The peaks in the range of 1200–1650 cm−1 are associated with the aromatic heterocyclic C–N rings. The peak at 800 cm−1 is attributed to the breathing mode of triazine rings. These findings confirm the structural integrity of g-C3N4, consistent with previously reported results [40,52]. For PSCN, MgPSCN, and LiPSCN the reduced intensity of the C–N and N–H vibrational peaks indicate structural distortions caused by the incorporation of phosphorus and sulfur additives. Additionally, the slight shifts in the characteristic peaks reflect strong interactions between the active materials and the g-C3N4 framework. These results are similar to those of previous studies [50], which found that the addition of phosphorus and sulfur altered the electronic properties and enhanced the Lewis basicity. A notable observation for MgPSCN is the further decrease in peak intensity compared to PSCN, indicating stronger structural distortions attributed to Mg–N coordination. This coordination modifies the electronic environment of the material, aligning with findings reported in earlier studies [52,53]. LiPSCN exhibits a spectral behavior similar to MgPSCN but with less pronounced reductions in peak intensity and smaller shifts, suggesting weaker interactions between Li and the g-C3N4 framework [40]. Overall, the FTIR analysis demonstrates the successful incorporation of P, S, Mg, and Li into the g-C3N4 framework, highlighting the effectiveness of tailored doping strategies in enhancing the catalytic and electronic properties for advanced applications.

3.2. Optimization of Catalyst Activity Through Experimental Design Methods

The effectiveness of glycerol carboxylation with CO2 using the MgPSCN catalyst was assessed by examining glycerol conversion and glycerol carbonate yield. This evaluation employed a Box–Behnken experimental design, which considered the three key independent variables of reaction temperature (X1), catalyst amount (X2), and reaction pressure (X3). The Box–Behnken design is a robust statistical method that has been successfully utilized in various studies to optimize complex chemical reactions. By systematically varying these factors researchers can identify the conditions for the maximization of the yield of the carboxylation process. This approach has proven to be valuable in enhancing the understanding and performance of catalytic systems in the literature [63,64,65]. The experimental design matrix corresponding to the Box–Behnken design consisted of a series of 15 experiments. The reaction time for these experiments was set to 4 h. The natural values as well as the coded factor values used in our experimental design matrix are given in Table 2.
The impact of the factors on the responses was assessed using a polynomial model of the following form:
Y i = β 0 + β 1 X 1 + β 2 X 2 + β 3 X 3 + β 12 X 1 X 2 + β 13 X 1 X 3 + β 23 X 2 X 3 + β 11 X 1 2 + β 22 X 2 2 + β 33 X 3 2
The validation of the equations was performed using statistical methods, specifically analysis of variance (ANOVA). Additionally, response surface methodology (RSM) was employed to analyze the experimental design, utilizing the experimental values in Table 3. The Fisher test value (F-value) calculated for the model was compared with the theoretical Fisher test value (Ft(p − 1, np)) at a 5% significance level. This comparison was made to assess the model’s validity, taking into account the number of experiments conducted (n) and the number of terms included in the model (p). By evaluating the F-value against the theoretical threshold, researchers can determine whether the model provides a statistically significant explanation of the observed data, ensuring the robustness and reliability of the experimental results [66].
In other words, as highlighted in the literature, the p-value for each parameter should be below 5% to ensure that each independent variable in the model significantly contributes to predicting the output. This threshold indicates that the variable’s effect on the response is statistically significant, thereby validating its inclusion in the model [67]. The coefficients of the models, along with the ANOVA tests, were thoroughly evaluated using Design-Expert Version 8.
Both glycerol conversion (Y1) and glycerol carbonate yield (Y2) were considered as dependent variables. To find the best operating conditions which ensure the highest glycerol conversion as well as glycerol carbonate yield, a polynomial model based on Equation (2) was developed. The analysis of variance (ANOVA) for both models given in Table 4 proves that the adequacy of these models is in good agreement with the experimental data. The adequacy is supported by a p-value lower than 5%, a correlation coefficient (R2) above 0.95, as well as the shapes of the parity diagrams presented in Figure 8.
The thorough analysis of the values of residuals, providing the differences between the calculated and experimental values, is another justification regarding the model accuracy if these values are enough small (Figure 8). By considering the normal probability plot, one can evaluate the data’s normal distribution. Figure 9 presents both the internally studentized residuals and the normal probability plot for glycerol conversion and GC yield.
The estimated coefficients together with their confidence intervals are given by relation (2) for glycerol conversion and by relation (3) for GC yield.
The effects of the main and binary interactions can be evaluated using the estimated coefficients. Coefficients that did not pass the statistical significance test (based on p-value) were removed, except for those necessary to maintain the model’s hierarchy. The main interactions were significant for both dependent variables. Equations (3) and (4) are as follows:
Y 1 = 40.21 1 ± 0.056 + 13.87 1 ± 0.152 X 1 + 18.69 1 ± 0.113 X 2 + 9.54 1 ± 0.221 X 3 + 10.76 1 ± 0.278 X 1 X 2 + 8.13 1 ± 0.369 X 1 X 3 14.66 1 ± 0.0.211 X 1 2
Y 2 = 27.99 1 ± 0.084 + 8.91 1 ± 0.162 X 1 + 10.7 1 ± 0.135 X 2 + 7.18 1 ± 0.203 X 3 + 6.25 1 ± 0.329 X 1 X 2 + 6.9 1 ± 0.298 X 2 X 3 6.31 1 ± 0.339 X 1 2 3.81 1 ± 0.559 X 2 2 5.36 1 ± 0.397 X 3 2
A positive influence means that the parameter is increasing with the factor and a negative influence means that the parameter decreases with the factor.
Given the coefficients in Equations (3) and (4) and plots in Figure 10, both glycerol conversion and GC yield are positively influenced by reaction temperature (X1), catalyst amount (X2), and reaction pressure (X3). The impact of the key factors on glycerol conversion and GC yield is illustrated in Figure 10. The ANOVA for the GC yield model is given in Table 5. Based on these values, an F-value 114.3 for the model suggests that this is significant, and p-values below 0.05 indicate that the model terms are significant.
The experimental results given in Table 6 are determined using the optimal condition calculated based on the mathematical model, and the highest values for glycerol conversion as well as GC yield were for the MgPSCN catalyst. The optimal values of the factors obtained by maximizing both glycerol conversion and GC yield are X 1 = 0.83 ,   X 2 = 1 ,   and   X 3 = 0.5 that provides a glycerol conversion of 71% and a GC yield of 46%, values close to the experimental values presented in Table 6 for the same catalyst.
The reusability of the MgPSCN catalyst in the synthesis of glycerol carbonate was investigated over four consecutive reaction cycles under optimal conditions. Before each reuse, the catalyst was rinsed with ethanol and subsequently dried at 85 °C for 60 min to ensure the removal of any residual byproducts or impurities that could influence its catalytic performance.
The results demonstrated a gradual decline in glycerol carbonate yield with each successive reuse of the catalyst (Figure 11). This reduction can be attributed to the accumulation of secondary or undesired compounds on the catalyst surface, leading to active site blockage and a decrease in catalytic efficiency despite ethanol washing. The plotted data illustrate this downward trend, highlighting the impact of surface and chemical modifications on the catalyst’s effectiveness over multiple uses.
To examine potential morphological changes in the MgPSCN catalyst after two consecutive cycles of the glycerol carbonate synthesis reaction (coupling glycerol with carbon dioxide), the catalyst was studied by SEM before the reaction. Prior to characterization, the catalyst was flushed with ethanol and dried at 85 °C for 60 min. SEM images revealed nanometer-sized granules on the surface of irregularly structured layers with multiple pores (Figure 11 and Figure 12), resembling the morphological structure of the catalyst before use. Furthermore, elemental mapping results of the catalyst after two cycles of use showed lower distributions of C, N, P, S, and Mg compared to those observed in the fresh catalyst (see Figure 13). This loss can be assigned to the detachment of nanoparticles from the surface of the nitride layers, which correlated with a decrease in glycerol carbonate yield after use.

4. Conclusions

This work introduces a new catalytic system with potential for scalable, environmentally friendly synthesis of glycerol carbonate and provides both experimental data and statistical modeling to support its performance. In this respect, four graphitic–carbon–nitride (g-C3N4) catalysts co-doped with phosphorus (P), sulfur (S), magnesium (Mg), and lithium (Li) were synthesized using a one-step thermal condensation method. The catalysts were characterized using XRD, SEM, TGA, FTIR, and nitrogen adsorption–desorption techniques and were employed for the synthesis of glycerol carbonate through the coupling reaction of glycerol and CO2. Co-doping did not significantly modify the average crystallite size, except for lithium doping which favored the shrinkage of crystallites. The BET surface area of the pores was also maintained at similar values, except for Li doping which almost doubled it. Thermogravimetric analysis revealed that doping with sulfur and phosphorus led to increased weight loss, which can be elucidated by the lesser value of the homolytic dissociation energy of the P–O–C bond compared to that of the C–C bond. The incorporation of phosphorus and sulfur additives to the PSCN, MgPSCN, and LiPSCN catalysts led to a reduction in the intensity of the C–N and N–H vibrational peaks, as revealed by FTIR analysis. SEM analysis revealed that doping with magnesium, phosphorus, and sulfur favored the formation of numerous clusters with variable sizes at the nanometric scale, with diameters smaller than those in PSCN, while doping with lithium favored the uniformization of the structure compared to PSCN. The parameters of the glycerol carbonate synthesis process were optimized using the response surface methodology based on a Box–Behnken design, with reaction temperature, catalyst amount, and operating pressure as independent variables. All catalysts considered in this study were tested under these conditions, with the experimental data obtained highlighting the increase in conversion and yield in glycerol carbonate for the doped catalysts. Thus, highest values for glycerol conversion of 70.05% and GC yield of 45.85% were obtained on the MgPSCN catalyst using the optimal factors values.

Author Contributions

Conceptualization, D.M.C.; Data curation, I.B., C.P. and R.D.; Formal analysis, G.V., I.L.L.M., M.B. and I.V.G.; Investigation, M.L.C., I.L.L.M., G.V., M.B., C.D.-V. and R.D.; Methodology, M.L.C., D.B. and C.P.; Project administration, D.M.C.; Resources, D.B and I.L.L.M.; Software, I.B. and R.D.; Supervision, D.M.C.; Validation, G.V. and D.B.; Visualization, D.M.C.; Writing—original draft, M.L.C., I.B., D.B., C.D.-V. and R.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was carried out through the PN 23.06 Core Program—ChemNewDeal within the National Plan for Research, Development, and Innovation 2022–2027, developed with the support of the Ministry of Research, Innovation, and Digitization, project no. PN 23.06.02.01 (InteGral).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Author Ioana Lavinia Lixandru Matei was employed by the company Botanical SRL. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. International Energy Agency. CO2 Emissions in 2023; International Energy Agency: Paris, France, 2024. [Google Scholar]
  2. Usman, M.; Rehman, A.; Saleem, F.; Abbas, A.; Eze, V.C.; Harvey, A. Synthesis of cyclic carbonates from CO2 cycloaddition to bio-based epoxides and glycerol: An overview of recent development. RSC Adv. 2023, 13, 22717–22743. [Google Scholar] [CrossRef] [PubMed]
  3. Gao, J.; Yang, J.-W.; Ma, T.; Wang, J.; Xia, D.; Du, B.; Cui, Y.; Yang, C. Mechanism study on direct synthesis of glycerol carbonate from CO2 and glycerol over shaped CeO2 model catalysts. Chin. Chem. Lett. 2023, 34, 108395. [Google Scholar] [CrossRef]
  4. Li, Y.; Liu, J.; He, D. Catalytic synthesis of glycerol carbonate from biomass-based glycerol and dimethyl carbonate over Li-La2O3 catalysts. Appl. Catal. A Gen. 2018, 564, 234–242. [Google Scholar] [CrossRef]
  5. Fan, X.; Burton, R.; Zhou, Y. Glycerol (Byproduct of Biodiesel Production) as a Source for Fuels and Chemicals–Mini Review. Open Fuels Energy Sci. J. 2010, 3, 17–22. [Google Scholar] [CrossRef]
  6. Kaur, J.; Sarma, A.K.; Jha, M.K.; Gera, P. Valorisation of crude glycerol to value-added products: Perspectives of process technology, economics and environmental issues. Biotechnol. Rep. 2020, 27, e00487. [Google Scholar] [CrossRef] [PubMed]
  7. Tan, H.W.; Abdul Aziz, A.R.; Aroua, M.K. Glycerol production and its applications as a raw material: A review. Renew. Sustain. Energy Rev. 2013, 27, 118–127. [Google Scholar] [CrossRef]
  8. Inrirai, P.; Keogh, J.; Centeno-Pedrazo, A.; Artioli, N.; Manyar, H. Recent advances in processes and catalysts for glycerol carbonate production via direct and indirect use of CO2. J. CO2 Util. 2024, 80, 102693. [Google Scholar] [CrossRef]
  9. Nda-Umar, U.I.; Ramli, I.; Taufiq-Yap, Y.H.; Muhamad, E.N. An Overview of Recent Research in the Conversion of Glycerol into Biofuels, Fuel Additives and other Bio-Based Chemicals. Catalysts 2019, 9, 15. [Google Scholar] [CrossRef]
  10. Sonnati, M.O.; Amigoni, S.; Taffin de Givenchy, E.P.; Darmanin, T.; Choulet, O.; Guittard, F. Glycerol carbonate as a versatile building block for tomorrow: Synthesis, reactivity, properties and applications. Green Chem. 2013, 15, 283–306. [Google Scholar] [CrossRef]
  11. Christy, S.; Noschese, A.; Lomelí-Rodriguez, M.; Greeves, N.; Lopez-Sanchez, J.A. Recent progress in the synthesis and applications of glycerol carbonate. Curr. Opin. Green Sustain. Chem. 2018, 14, 99–107. [Google Scholar] [CrossRef]
  12. Li, J.; Wang, T. Chemical equilibrium of glycerol carbonate synthesis from glycerol. J. Chem. Thermodyn. 2011, 43, 731–736. [Google Scholar] [CrossRef]
  13. Razali, N.; McGregor, J. Improving Product Yield in the Direct Carboxylation of Glycerol with CO2 through the Tailored Selection of Dehydrating Agents. Catalysts 2021, 11, 138. [Google Scholar] [CrossRef]
  14. Honda, M.; Tamura, M.; Nakao, K.; Suzuki, K.; Nakagawa, Y.; Tomishige, K. Direct Cyclic Carbonate Synthesis from CO2 and Diol over Carboxylation/Hydration Cascade Catalyst of CeO2 with 2-Cyanopyridine. ACS Catal. 2014, 4, 1893–1896. [Google Scholar] [CrossRef]
  15. Liu, J.; Li, Y.; Zhang, J.; He, D. Glycerol carbonylation with CO2 to glycerol carbonate over CeO2 catalyst and the influence of CeO2 preparation methods and reaction parameters. Appl. Catal. A Gen. 2016, 513, 9–18. [Google Scholar] [CrossRef]
  16. Hu, C.; Chang, C.-W.; Yoshida, M.; Wang, K.-H. Lanthanum nanocluster/ZIF-8 for boosting catalytic CO2/glycerol conversion using MgCO3 as a dehydrating agent. J. Mater. Chem. A 2021, 9, 7048–7058. [Google Scholar] [CrossRef]
  17. Takeuchi, K.; Matsumoto, K.; Fukaya, N.; Sato, K.; Choi, J.C. Synthesis of Glycerol Carbonate from Glycerol and CO2 Using CaO as a Dehydrating Agent. Asian J. Org. Chem. 2022, 11, e202200212. [Google Scholar] [CrossRef]
  18. Rozulan, N.; Halim, S.A.; Razali, N.; Lam, S.S. A review on direct carboxylation of glycerol waste to glycerol carbonate and its applications. Biomass Convers. Biorefin. 2022, 12, 4665–4682. [Google Scholar] [CrossRef]
  19. Gao, Z.; Xiang, M.; He, M.; Zhou, W.; Chen, J.; Lu, J.; Wu, Z.; Su, Y. Transformation of CO2 with Glycerol to Glycerol Carbonate over ETS-10 Zeolite-Based Catalyst. Molecules 2023, 28, 2272. [Google Scholar] [CrossRef] [PubMed]
  20. Chen, X.; Shu, X.; Zhu, Y.; Zhang, J.; Chai, Z.; Song, H.; An, Z.; He, J. Highly dispersed MgInCe-mixed metal oxides catalyzed direct carbonylation of glycerol and CO2 into glycerol carbonate. Chin. J. Chem. Eng. 2024, 72, 153–163. [Google Scholar] [CrossRef]
  21. Xu, H.; Ke, Y. Synthesis of glycerol carbonate from glycerol and CO2 over Cu-Zr complex oxide. J. Fuel Chem. Technol. 2024, 52, 171–182. [Google Scholar] [CrossRef]
  22. Cao, Z.; Wu, G.; Wang, X.; Deng, L.; Wan, J.; Liu, Y.; Kan, J.; Shang, C.; Guo, Z. Resourcification of CO2 to high-value-added glycerol carbonate by ZnAlCe composite oxides with frustrated Lewis pairs. Mol. Catal. 2024, 564, 114348. [Google Scholar] [CrossRef]
  23. Lukato, S.; Wójcik, M.; Krogul-Sobczak, A.; Litwinienko, G. Enhancing the Green Synthesis of Glycerol Carbonate: Carboxylation of Glycerol with CO2 Catalyzed by Metal Nanoparticles Encapsulated in Cerium Metal–Organic Frameworks. Nanomaterials 2024, 14, 650. [Google Scholar] [CrossRef] [PubMed]
  24. Park, C.-y.; Nguyen-Phu, H.; Shin, E.W. Glycerol carbonation with CO2 and La2O2CO3/ZnO catalysts prepared by two different methods: Preferred reaction route depending on crystalline structure. Mol. Catal. 2017, 435, 99–109. [Google Scholar] [CrossRef]
  25. Koranian, P.; Kumar Dalai, A.; Sammynaiken, R. Production of glycerol carbonate from glycerol and carbon dioxide using metal oxide catalysts. Chem. Eng. Sci. 2024, 286, 119687. [Google Scholar] [CrossRef]
  26. Fao, G.D.; Catherine, H.N.; Huang, C.-H.; Lee, Y.-L.; Jiang, J.-C.; Hu, C. Unraveling the effects of P and S doping over g-C3N4 in strengthening Lewis basicity for CO2/glycerol conversion: A theoretical and experimental study. Carbon 2023, 201, 129–140. [Google Scholar] [CrossRef]
  27. Zhu, J.; Xiao, P.; Li, H.; Carabineiro, S.A.C. Graphitic Carbon Nitride: Synthesis, Properties, and Applications in Catalysis. ACS Appl. Mater. Interfaces 2014, 6, 16449–16465. [Google Scholar] [CrossRef]
  28. Arumugam, M.; Tahir, M.; Praserthdam, P. Effect of nonmetals (B, O, P, and S) doped with porous g-C3N4 for improved electron transfer towards photocatalytic CO2 reduction with water into CH4. Chemosphere 2022, 286, 131765. [Google Scholar] [CrossRef]
  29. Niu, H.; Wang, X.; Shao, C.; Liu, Y.; Zhang, Z.; Guo, Y. Revealing the oxygen reduction reaction activity origin of single atoms supported on g-C3N4 monolayers: A first-principles study. J. Mater. Chem. A 2020, 8, 6555–6563. [Google Scholar] [CrossRef]
  30. Mohamed, R.M.; Kadi, M.W. Generation of Hydrogen Gas Using CuCr2O4-g-C3N4 Nanocomposites under Illumination by Visible Light. ACS Omega 2021, 6, 4485–4494. [Google Scholar] [CrossRef]
  31. Rajput, Y.; Kumar, P.; Zhang, T.C.; Kumar, D.; Nemiwal, M. Recent advances in g-C3N4-based photocatalysts for hydrogen evolution reactions. Int. J. Hydrogen Energy 2022, 47, 38533–38555. [Google Scholar] [CrossRef]
  32. Aboubakr, A.E.A.; El Rouby, W.M.A.; Khan, M.D.; Revaprasadu, N.; Millet, P. Effect of morphology and non-metal doping (P and S) on the activity of graphitic carbon nitride toward photoelectrochemical water oxidation. Sol. Energy Mater. Sol. Cells 2021, 232, 111326. [Google Scholar] [CrossRef]
  33. Chu, Y.-C.; Lin, T.-J.; Lin, Y.-R.; Chiu, W.-L.; Nguyen, B.-S.; Hu, C. Influence of P,S,O-Doping on g-C3N4 for hydrogel formation and photocatalysis: An experimental and theoretical study. Carbon 2020, 169, 338–348. [Google Scholar] [CrossRef]
  34. Gao, J.; Wang, Y.; Zhou, S.; Lin, W.; Kong, Y. A Facile One-Step Synthesis of Fe-Doped g-C3N4 Nanosheets and Their Improved Visible-Light Photocatalytic Performance. ChemCatChem 2017, 9, 1708–1715. [Google Scholar] [CrossRef]
  35. Wang, J.-C.; Cui, C.-X.; Li, Y.; Liu, L.; Zhang, Y.-P.; Shi, W. Porous Mn doped g-C3N4 photocatalysts for enhanced synergetic degradation under visible-light illumination. J. Hazard. Mater. 2017, 339, 43–53. [Google Scholar] [CrossRef] [PubMed]
  36. Budiman, A.W.; Sembodo, B.S.T.; Alviansyah, A.; Putri, I.B.S. Application of titanium doped g-C3N4 for the degradation of toxic dyes in textiles wastewater. AIP Conf. Proc. 2020, 2217, 030127. [Google Scholar] [CrossRef]
  37. Zhang, R.; Niu, S.; Zhang, X.; Jiang, Z.; Zheng, J.; Guo, C. Combination of experimental and theoretical investigation on Ti-doped g-C3N4 with improved photo-catalytic activity. Appl. Surf. Sci. 2019, 489, 427–434. [Google Scholar] [CrossRef]
  38. Xu, Y.; Ge, F.; Chen, Z.; Huang, S.; Wei, W.; Xie, M.; Xu, H.; Li, H. One-step synthesis of Fe-doped surface-alkalinized g-C3N4 and their improved visible-light photocatalytic performance. Appl. Surf. Sci. 2019, 469, 739–746. [Google Scholar] [CrossRef]
  39. Deng, P.; Xiong, J.; Lei, S.; Wang, W.; Ou, X.; Xu, Y.; Xiao, Y.; Cheng, B. Nickel formate induced high-level in situ Ni-doping of g-C3N4 for a tunable band structure and enhanced photocatalytic performance. J. Mater. Chem. A 2019, 7, 22385–22397. [Google Scholar] [CrossRef]
  40. Paul, D.R.; Sharma, R.; Singh, S.; Singh, P.; Panchal, P.; Sharma, A.; Devi, P.; Nehra, S.P. Mg/Li Co-doped g-C3N4: An excellent photocatalyst for wastewater remediation and hydrogen production applications towards sustainable development. Int. J. Hydrogen Energy 2023, 48, 37746–37761. [Google Scholar] [CrossRef]
  41. Zhang, W.; Zhang, Z.; Choi, S.H.; Yang, W. Facile enhancement of photocatalytic efficiency of g-C3N4 by Li-intercalation. Catal. Today 2019, 321–322, 67–73. [Google Scholar] [CrossRef]
  42. Vuggili, S.B.; Khanth, S.K.; Kadiya, K.; Gaur, U.K.; Sharma, M. Improvement in visible light stimulated photocatalysis by the inducement of magnesium dopant inside graphitic carbon nitride frameworks. J. Environ. Chem. Eng. 2019, 7, 103440. [Google Scholar] [CrossRef]
  43. Reisi, B.; Najafi Chermahini, A. Modification g-CN by MgO and its application for glycerol carbonate synthesis from glycerol and dimethyl carbonate. Environ. Prog. Sustain. Energy 2023, 42, e14007. [Google Scholar] [CrossRef]
  44. Liu, G.; Qiao, X.; Gondal, M.A.; Liu, Y.; Shen, K.; Xu, Q. Comparative Study of Pure g-C3N4 and Sulfur-Doped g-C3N4 Catalyst Performance in Photo-Degradation of Persistent Pollutant Under Visible Light. J. Nanosci. Nanotechnol. 2018, 18, 4142–4154. [Google Scholar] [CrossRef] [PubMed]
  45. Lukato, S.; Kasozi, G.N.; Naziriwo, B.; Tebandeke, E. Glycerol carbonylation with CO2 to form glycerol carbonate: A review of recent developments and challenges. Curr. Res. Green Sustain. Chem. 2021, 4, 100199. [Google Scholar] [CrossRef]
  46. Charif, M.L.; Doukeh, R.; Ciuparu, D.M. The Catalytic Performance of Metal-Oxide-Based Catalysts in the Synthesis of Glycerol Carbonate: Toward the Green Valorization of Glycerol. Catalysts 2025, 15, 534. [Google Scholar] [CrossRef]
  47. Liu, D.Y.; Dong, J.H.; Liu, F.M.; Gao, X.F.; Yu, Y.; Zhang, S.B.; Dong, L.M.; Guo, Y.K. Synthesis and photocatalytic performance of g-C3N4 composites. J. Ovonic Res. Vol. 2019, 15, 239–246. [Google Scholar]
  48. Cao, J.; Qin, C.; Wang, Y.; Zhang, H.; Sun, G.; Zhang, Z. Solid-State Method Synthesis of SnO2-Decorated g-C3N4 Nanocomposites with Enhanced Gas-Sensing Property to Ethanol. Materials 2017, 10, 604. [Google Scholar] [CrossRef]
  49. Jiang, J.; Zheng, B.; Yu, W.; Wu, X.; Mi, R.; Huang, Z.; Liu, Y.-g.; Fang, M.; Min, X. Synthesis and photocatalytic performance of composite g-C3N4 with functionalized multi-walled carbon nanotubes. J. Alloys Compd. 2023, 968, 171707. [Google Scholar] [CrossRef]
  50. Liu, Y.; Liu, P.; Sun, C.; Wang, T.; Tao, K.; Gao, D. P dopants induced ferromagnetism in g-C3N4 nanosheets: Experiments and calculations. Appl. Phys. Lett. 2017, 110, 222403. [Google Scholar] [CrossRef]
  51. Jiang, L.; Yuan, X.; Zeng, G.; Chen, X.; Wu, Z.; Liang, J.; Zhang, J.; Wang, H.; Wang, H. Phosphorus- and Sulfur-Codoped g-C3N4: Facile Preparation, Mechanism Insight, and Application as Efficient Photocatalyst for Tetracycline and Methyl Orange Degradation under Visible Light Irradiation. ACS Sustain. Chem. Eng. 2017, 5, 5831–5841. [Google Scholar] [CrossRef]
  52. Dong, X.; Zhang, S.; Wu, H.; Kang, Z.; Wang, L. Facile one-pot synthesis of Mg-doped g-C3N4 for photocatalytic reduction of CO2. RSC Adv. 2019, 9, 28894–28901. [Google Scholar] [CrossRef]
  53. Duan, Y.; Li, J.; Jia, D.; Yao, H.; Shang, X.; Li, C. One-step synthesis of Mg-doped g-C3N4 nanosheets for efficient photo-Fenton-like catalysis. Diam. Relat. Mater. 2022, 129, 109313. [Google Scholar] [CrossRef]
  54. Sedaghati-Jamalabad, G.; Bagheri-Mohagheghi, M.M. A study on the structural and optical properties of the SnFe2O4 spinel compound as anode electrode in Li ion-battery: The optical and dielectric parameters via synthesis methods. Opt. Quantum Electron. 2024, 56, 965. [Google Scholar] [CrossRef]
  55. Yang, Y.; Yan, J.; Zhang, Y.; Xing, S.; Ran, J.; Ma, Y.; Li, X. S/P co-doped g-C3N4 with secondary calcination for excellent photocatalytic performance. Int. J. Hydrogen Energy 2024, 51, 962–974. [Google Scholar] [CrossRef]
  56. Cako, E.; Dudziak, S.; Głuchowski, P.; Trykowski, G.; Pisarek, M.; Fiszka Borzyszkowska, A.; Sikora, K.; Zielińska-Jurek, A. Heterojunction of (P, S) co-doped g-C3N4 and 2D TiO2 for improved carbamazepine and acetaminophen photocatalytic degradation. Sep. Purif. Technol. 2023, 311, 123320. [Google Scholar] [CrossRef]
  57. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef]
  58. Chen, K.; Zhang, T.; Chen, X.; He, Y.; Liang, X. Model construction of micro-pores in shale: A case study of Silurian Longmaxi Formation shale in Dianqianbei area, SW China. Pet. Explor. Dev. 2018, 45, 412–421. [Google Scholar] [CrossRef]
  59. Yang, Y.; Zhang, J.; Xu, L.; Li, P.; Liu, Y.; Dang, W. Pore Structure and Fractal Characteristics of Deep Shale: A Case Study from Permian Shanxi Formation Shale, from the Ordos Basin. ACS Omega 2022, 7, 9229–9243. [Google Scholar] [CrossRef]
  60. Zhao, C.; Yan, Q.; Wang, S.; Dong, P.; Zhang, L. Regenerable g-C3N4–chitosan beads with enhanced photocatalytic activity and stability. RSC Adv. 2018, 8, 27516–27524. [Google Scholar] [CrossRef]
  61. Shirman, R.; Chakraborty, S.; Sasson, Y. Ru/GCN Nanocomposite as an Efficient Catalyst for Hydrogen Generation from Sodium Hypophosphite. Nanomaterials 2024, 14, 1187. [Google Scholar] [CrossRef]
  62. Li, Z.; Cheng, X.; Liu, Y.; Liu, H.; Jiang, Y.; Wang, N. Intumescent flame retardancy and smoke suppression of Eucommia ulmoides gum/natural rubber blends based on synergistic g-C3N4@Fe3O4 nanocomposites. RSC Adv. 2022, 12, 21704–21712. [Google Scholar] [CrossRef] [PubMed]
  63. Stoica-Guzun, A.; Stroescu, M.; Jinga, S.I.; Mihalache, N.; Botez, A.; Matei, C.; Berger, D.; Damian, C.M.; Ionita, V. Box-Behnken experimental design for chromium(VI) ions removal by bacterial cellulose-magnetite composites. Int. J. Biol. Macromol. 2016, 91, 1062–1072. [Google Scholar] [CrossRef] [PubMed]
  64. Perveen, R.; Latif, F.; Abbas, M.; Ayub, K.; Sarfaraz, S.; Saeed, M.; Rana, S.; Al-Rashida, M.; Nawaz, M.A.; Hameed, A. Optimization study and application of box-behnken model for probing eggshell supported transition metals based catalysts to synthesize hydrazone & dihydropyrimidinones. Sci. Rep. 2024, 14, 23270. [Google Scholar] [CrossRef]
  65. Matei, P.L.; Deleanu, I.; Brezoiu, A.M.; Chira, N.A.; Busuioc, C.; Isopencu, G.; Cîlțea-Udrescu, M.; Alexandrescu, E.; Stoica-Guzun, A. Ultrasound-Assisted Extraction of Blackberry Seed Oil: Optimization and Oil Characterization. Molecules 2023, 28, 2486. [Google Scholar] [CrossRef]
  66. Kamaroddin, M.F.; Rahaman, A.; Gilmour, D.J.; Zimmerman, W.B. Optimization and cost estimation of microalgal lipid extraction using ozone-rich microbubbles for biodiesel production. Biocatal. Agric. Biotechnol. 2020, 23, 101462. [Google Scholar] [CrossRef]
  67. Montgomery, D.C. Design and Analysis of Experiments, 8th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2013. [Google Scholar]
Figure 1. The X-ray diffraction (XRD) patterns of catalysts.
Figure 1. The X-ray diffraction (XRD) patterns of catalysts.
Applsci 15 06236 g001
Figure 2. Survey scan of XPS of MgPSCN before and after use.
Figure 2. Survey scan of XPS of MgPSCN before and after use.
Applsci 15 06236 g002
Figure 3. Scanning electron microscopy images of CN (A), PSCN (B), MgPSCN (C), and LiPSCN (D) catalysts.
Figure 3. Scanning electron microscopy images of CN (A), PSCN (B), MgPSCN (C), and LiPSCN (D) catalysts.
Applsci 15 06236 g003
Figure 4. EDX spectra of CN (A), PSCN (B), MgPSCN (C), and LiPSCN (D) catalysts.
Figure 4. EDX spectra of CN (A), PSCN (B), MgPSCN (C), and LiPSCN (D) catalysts.
Applsci 15 06236 g004
Figure 5. Adsorption–desorption isotherms and pore volume distribution for CN, PSCN, MgPSCN, and LiPSCN.
Figure 5. Adsorption–desorption isotherms and pore volume distribution for CN, PSCN, MgPSCN, and LiPSCN.
Applsci 15 06236 g005
Figure 6. TGA-DTG patterns of the CN, PSCN, MgPSCN, and LiPSCN.
Figure 6. TGA-DTG patterns of the CN, PSCN, MgPSCN, and LiPSCN.
Applsci 15 06236 g006
Figure 7. FTIR spectra of the CN, PSCN, MgPSCN, and LiPSCN.
Figure 7. FTIR spectra of the CN, PSCN, MgPSCN, and LiPSCN.
Applsci 15 06236 g007
Figure 8. Parity diagrams for glycerol conversion (A) and glycerol carbonate yield (B).
Figure 8. Parity diagrams for glycerol conversion (A) and glycerol carbonate yield (B).
Applsci 15 06236 g008
Figure 9. Normal probability plot ((A)—glycerol conversion; (C)—GC yield) and internally studentized residuals ((B)—glycerol conversion, (D)—GC yield).
Figure 9. Normal probability plot ((A)—glycerol conversion; (C)—GC yield) and internally studentized residuals ((B)—glycerol conversion, (D)—GC yield).
Applsci 15 06236 g009
Figure 10. Influence of the main factors on the glycerol conversion and GC yield.
Figure 10. Influence of the main factors on the glycerol conversion and GC yield.
Applsci 15 06236 g010
Figure 11. Effect of catalyst reuse on glycerol carbonate yield over successive reaction cycles.
Figure 11. Effect of catalyst reuse on glycerol carbonate yield over successive reaction cycles.
Applsci 15 06236 g011
Figure 12. SEM images of the MgPSCN catalyst after two reaction cycles at optimum parameters.
Figure 12. SEM images of the MgPSCN catalyst after two reaction cycles at optimum parameters.
Applsci 15 06236 g012
Figure 13. EDX spectra of the MgPSCN catalyst after two reaction cycles at optimum parameters.
Figure 13. EDX spectra of the MgPSCN catalyst after two reaction cycles at optimum parameters.
Applsci 15 06236 g013
Table 1. Textural properties and crystallite sizes.
Table 1. Textural properties and crystallite sizes.
CatalystsBET Surface Area (m2/g)Pore Volume (cm3/g)Average Pore Width (nm)Average Crystallite Size (nm)
CN10.850.0625.2010.94
PSCN10.100.08529.39617.57
MgPSCN9.190.06316.0910.71
LiPSCN17.320.12829.3966.86
Table 2. Numerical values assigned to the independent variables within the experimental design matrix.
Table 2. Numerical values assigned to the independent variables within the experimental design matrix.
FactorValue/Level
LowMediumHigh
Natural (Z)Coded (X)Natural (Z)Coded (X)Natural (Z)Coded (X)
Reaction temperature (°C), X1150−12000250+1
Catalyst amount (%), X210−130050+1
Operating pressure (bar), X33−17010+1
Table 3. Box–Behnken experimental design data.
Table 3. Box–Behnken experimental design data.
Nr. ExpX1X2X3Y1Y2
1−11018.5114.28
211068.7244.9
3−10114.8212.98
400041.4927.75
510157.8533.02
600042.9428.15
70−1128.39.93
81−1011.858.98
90−1−110.228.37
1001163.4543.08
1101−153.8813.94
12−10−18.722.15
1310−119.2517.15
1400041.2128.08
15−1−104.673.34
Table 4. ANOVA parameters used to evaluate the suitability of the model.
Table 4. ANOVA parameters used to evaluate the suitability of the model.
Independent VariableCorrelation Coefficient, R2Adjusted R2Calculated F-Valuep-Value
Y10.9910.985162.9<0.0001
Y20.9930.984114.3<0.0001
Table 5. The ANOVA analysis table for GC yield model.
Table 5. The ANOVA analysis table for GC yield model.
SourceSum of SquaresDfMean SquareF-Valuep-Value
Model2577.198322.15114.30<0.0001significant
X 1 635.461635.46225.47<0.0001significant
X 2 915.491915.49324.83<0.0001significant
X 3 411.851411.85146.13<0.0001significant
X 1 X 2 156.001156.0055.350.0003significant
X 2 X 3 190.161190.1667.470.0002significant
X 1 2 147.091147.0952.190.0004significant
X 2 2 53.50153.5018.980.0048significant
X 3 2 105.951105.9537.590.0009significant
Cor Total2594.1014
Table 6. Conversion and GC yield at optimum conditions.
Table 6. Conversion and GC yield at optimum conditions.
CatalystsGlycerol ConversionGC Yield
CN47.2411.91
PSCN61.8829.74
MgPSCN70.0545.85
LiPSCN58.4227.34
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Charif, M.L.; Ciuparu, D.M.; Lixandru Matei, I.L.; Vasilievici, G.; Banu, I.; Băjan, M.; Bomboș, D.; Dușescu-Vasile, C.; Ghețiu, I.V.; Panaitescu, C.; et al. An Experimental Study of Glycerol Carbonate Synthesis over g-C3N4 Catalysts. Appl. Sci. 2025, 15, 6236. https://doi.org/10.3390/app15116236

AMA Style

Charif ML, Ciuparu DM, Lixandru Matei IL, Vasilievici G, Banu I, Băjan M, Bomboș D, Dușescu-Vasile C, Ghețiu IV, Panaitescu C, et al. An Experimental Study of Glycerol Carbonate Synthesis over g-C3N4 Catalysts. Applied Sciences. 2025; 15(11):6236. https://doi.org/10.3390/app15116236

Chicago/Turabian Style

Charif, Mirna Lea, Dragoș Mihael Ciuparu, Ioana Lavinia Lixandru Matei, Gabriel Vasilievici, Ionuț Banu, Marian Băjan, Dorin Bomboș, Cristina Dușescu-Vasile, Iuliana Veronica Ghețiu, Cașen Panaitescu, and et al. 2025. "An Experimental Study of Glycerol Carbonate Synthesis over g-C3N4 Catalysts" Applied Sciences 15, no. 11: 6236. https://doi.org/10.3390/app15116236

APA Style

Charif, M. L., Ciuparu, D. M., Lixandru Matei, I. L., Vasilievici, G., Banu, I., Băjan, M., Bomboș, D., Dușescu-Vasile, C., Ghețiu, I. V., Panaitescu, C., & Doukeh, R. (2025). An Experimental Study of Glycerol Carbonate Synthesis over g-C3N4 Catalysts. Applied Sciences, 15(11), 6236. https://doi.org/10.3390/app15116236

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop