Next Article in Journal
Iron Sludge-Derived Photo-Fenton Reaction for Laundry Wastewater Effluent Oxidation and Process Optimization into Industrial Ecology Symbiosis
Previous Article in Journal
Plasmon-Driven Catalytic Inhibition of pATP Oxidation as a Mechanism for Indirect Fe²⁺ Detection on a SERS-Active Platform
Previous Article in Special Issue
Catalytic Conversion of Ethyl Levulinate to γ-Valerolactone Under Mild Conditions over Zr-Beta Acidic Zeolite Prepared by Hydrothermal Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Catalyst Innovation, Mechanism Exploration, and Process Optimization for Synthesis of Glycerol Carbonate

by
Honglei Sun
1,
Zhenyu Lei
1,
Jinghui Shi
2,* and
Mingjun Jia
1,*
1
Key Laboratory of Surface and Interface Chemistry of Jilin Province, College of Chemistry, Jilin University, Changchun 130012, China
2
Key Laboratory of Preparation and Applications of Environmental Friendly Materials, Jilin Normal University, Ministry of Education, Changchun 130103, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(7), 668; https://doi.org/10.3390/catal15070668
Submission received: 7 June 2025 / Revised: 1 July 2025 / Accepted: 7 July 2025 / Published: 9 July 2025

Abstract

The catalytic conversion of bio-based glycerol (Gly) into high-value glycerol carbonate (GC) has received great attention from both the academic and industrial fields. The development of highly efficient catalysts and economical industrial processes remains a challenging subject. In this mini-review, we summary the recent advances in catalyst design, characterization, mechanism, and catalytic process optimization, including the various synthetic strategies of GC, such as the coupling of CO2 and Gly or its derivatives like glycidol (GD), the transesterification of Gly with small carbonate-containing molecules, and the carbonylation of Gly with urea. The main difficulties and challenges faced by constructing high-performance catalysts and achieving scale production of GC have been put forward, and the future research directions and opportunities in catalyst innovation, reaction mechanism exploration, and continuous catalytic process improvement have also been suggested.

Graphical Abstract

1. Introduction

The development of biodiesel is highly desirable for meeting the needs of green chemistry and the energy industry. The production of biodiesel could generate glycerol (Gly) as a by-product. Although traditional petroleum-based Gly is an important raw material for various industrial uses, continuous biodiesel production has led to a serious surplus of bio-based Gly [1]. A major drawback of bio-based Gly is its relatively low purity, commonly containing metal ion residues like Na+ and K+, and various fatty acids and their saponification products [2,3]. Such complex impurities make the energy consumption for the purification of bio-based glycerol considerably higher than that of petroleum-based Gly [4]. In addition, the traditional Gly supply chain has already been well-established and the practical demand for Gly is still quite limited, which makes it impossible to solve the problem of Gly surplus. Therefore, there is an urgent need to develop novel technologies and methods for converting bio-based glycerol to highly evaluated products with wide application.
Recently, interest in developing green and efficient production processes for glycerol carbonate (GC) has been continuously growing, since it might be an ideal way to combine bio-based Gly valorization with greenhouse gas CO2 utilization [5]. As one of the most important platform compounds, GC has shown desirable features in a variety of application fields such as medicine, cosmetics, textiles, coatings, plastics, machinery, integrated circuits, new materials, and new energy [6,7,8,9]. In particular, the active groups (carbonyl, hydroxyl) in GC can react with compounds containing functional groups such as hydroxyl and thiol to synthesize important pharmaceutical and chemical intermediates. GC is also an ideal choice for replacing epoxides to modify polymer monomers and to produce functional polymer materials like polycarbonates and polyurethanes [8]. In addition, GC has been proposed as a more suitable electrolyte carrier owing to its lower flammability and boiling point, as well as improved conductivity. However, the current high cost of GC, which is a result of high production costs, has limited its widespread application. Therefore, more recent effort has been devoted to developing a new efficient and low-cost GC production process that could meet the constantly growing demand in global resource utilization and environmental protection [10,11].
From the perspective of developing the green chemical industry, GC could be synthesized from Gly by several ecofriendly catalytic processes, such as transesterification with small organic carbonates, carbonation with carbon dioxide, and carbonylation with urea. At present, the transesterification of organic carbonates (e.g., dimethyl carbonate, ethylene carbonate) with Gly is the major method for the industrial production GC, which is conducted in the presence of alkaline catalysts such as calcium oxide and sodium hydroxide [12,13,14,15]. This reaction can occur spontaneously in thermodynamics with mild reaction conditions and simple operation procedures. The main disadvantage of this production process is the usage of expensive organic carbonates, leading to a higher production costs. In addition, the alkaline catalysts used in this process mainly exist in homogeneous form, making it difficult to realize product separation and scalable continuous production. Therefore, the transesterification production process of GC still faces multiple challenges such as catalyst innovation, cost control, and process improvement [16,17,18,19].
In principle, the direct coupling of glycerol with CO2 (carbonation) would be the most straightforward and attractive method for producing GC. Numerous catalyst systems, including various metal oxides, ionic liquids, and supported solid base catalysts, have been investigated for catalyzing this process [20,21,22]. However, such a direct route is not regarded as a feasible way for industrial implementation since the carbonation reaction is highly constrained by thermodynamics [7,23]. A relatively low GC yield is commonly obtained even when harsh operating conditions are adopted. For achieving successful industrial production, better heterogeneous catalysts and process control (e.g., shifting chemical balance) are still required, which are the major challenges faced by this research field.
Alternatively, the utilization of glycidol (GD), a derivative of Gly, to react with carbon dioxide seems to be a better choice for GC production [24,25]. A variety of metal-based and organocatalysts have been investigated for the coupling of CO2 and GD, and some of them could act as effective homogeneous catalysts for converting GD to GC under mild conditions [26]. In comparison with Gly, the carbonation of GD with CO2 is much easier, although the search of more efficient and recyclable heterogeneous catalysts still remains a challenging subject for this process. In addition, the production of GD from bio-based Gly also deserves further study, and is a determinative factor for achieving large-scale production of GC through the carbonation of GD with CO2. Traditionally, GD could be produced by using petroleum-based Gly as a raw material through a multiple-step route in the presence of Brønsted acid and Lewis acid [27,28,29,30,31]. By establishing an intensified and scalable flow process, Monbaliu and co-workers have successfully achieved the effective conversion of bio-based Gly to GD with enhanced efficiency and selectivity in a two-step continuous flow setup [32].
The carbonylation of Gly with urea is another choice for the effective production of GC. Urea itself is non-toxic and has a low price. The by-product of ammonia could be recycled to produce urea through the reaction with CO2, thereby forming a glycerol–CO2 comprehensive utilization system with ammonia as the medium. Essentially, the product of GC is produced by continuously consuming glycerol and CO2. It was reported that a number of metal oxides, metal salts, hydrotalcites, and ionic liquids can catalyze the glycerol–urea carbonylation reaction [33,34,35,36,37,38]. Notably, some heterogenous Zn-based catalysts have received more attention owing to their relatively high catalytic activity and selectivity. However, most of them conducted the catalytic reaction in homogeneous rather than heterogenous conditions, since the dissolved Zn species has been identified as the main active site for the carbonylation reaction. Moreover, the reaction is also limited by thermodynamic equilibrium. It is necessary to shift the equilibrium towards the product by reducing the system pressure (i.e., removing the NH3). Therefore, much work is still required in order to improve the stability of heterogeneous Zn-based catalysts without compromising catalytic activity and selectivity. In addition, more information revealing the nature of the catalytically active sites and reaction mechanisms needs to be explored, which can be helpful for designing novel and efficient catalysts for the carbonylation of Gly with urea.
In this review, we aim to summarize the very recent research advances involving the catalytic synthesis of GC through the pathways mentioned above. More attention has been drawn to the prominent progress made in the following aspects: (1) catalyst innovation, e.g., designing novel heterogeneous catalysts that can solve the common drawbacks like lower catalytic efficiency and poor stability/recyclability; (2) mechanism exploration, proposing a profound understanding clarifying the nature of catalytically active sites and reaction mechanisms by combining various characterization results and theory calculations; (3) catalytic process improvement, developing more efficient catalytic reaction units or equipment, which can achieve the sustainable and economical scale-up of GC production. In addition, the main challenges and opportunities in future work shall be proposed, which could hopefully provide appropriate guidance on rationally designing and developing more efficient catalysts and low-cost GC production processes.

2. Synthesis of GC from the Carbonation of Gly with CO2 and the Thermodynamic Analyses of GC Production Through Different Routes

The continuous increase of CO2 concentration in the atmosphere is becoming a global crisis. Converting CO2 into high-value-added chemicals provides an environmentally friendly approach to reduce carbon emissions [39]. Significantly, the effective utilization of CO2 and bio-based Gly to catalytically synthesize GC could not only reduce the level of atmospheric CO2, but also significantly increase the appliable value of bio-based Gly. Many research works have demonstrated the great potential of this catalytic technology for achieving carbon capture and utilization [40,41].
As already mentioned above, the direct carbonation of Gly with CO2 has been thought of as the most ideal method for producing GC. Various catalysts have been investigated for the direct carbonation reaction, including metal oxides, ionic liquids, and supported solid base catalysts. Examples include protic ionic liquids, Mg/ZIF-8 and CeZrO2 solid solutions catalysts, etc. [21,42,43]. In a few recent review articles, the research progress and the main challenges in the future involving the synthesis of GC through the direct carbonation of Gly with CO2 have been well-summarized [5]. As claimed by Manyar and coauthors, although a number of heterogeneous catalysts have shown relatively catalytic activity in the direct carbonation reaction, their catalytic efficiencies are still unsatisfactory. Relatively low GC yields are usually obtained, even when higher reaction temperatures and pressures have been adopted. The main reason should be attributed to the fact that direct carbonation is constrained by thermodynamics.
Previously, Wang et al. [44] investigated the chemical equilibrium of GC synthesis from Gly. The results of thermodynamic analysis revealed that the direct carbonation of Gly with CO2 is thermodynamically unfavorable under ambient conditions, showing that ΔrGθm = +23.92 kJ/mol at T = 298.15 K and 101,325 Pa. This phenomenon mainly stems from the high stability of CO2 and the formation of water-induced reverse reactions. In general, the synthesis of GC via the direct carbonation route requires an increase in system pressure and temperature in order to overcome the energy barrier, which can shift the equilibrium toward GC production. In addition, the introduction of a dehydrating agent (e.g., acetonitrile) can significantly reduce the concentration of the product in the reaction system, thus contributing to a further shift in the equilibrium in the direction of GC generation.
For instance, Ke et al. [45] reported that by using the hydrothermally synthesized CuO as catalyst, the Gly conversion and GC selectivity could reach 89.0% and 69.4%, respectively, under the optimized reaction conditions (Gly 0.92 g, 2-cyanopyridine 3.26 g, DMF 10 mL, 120 °C, 5 h, P(CO2) = 3 Mpa), which are much higher than the previously reported catalyst systems. The enhanced performance of this CuO catalyst could be mainly related to the stabilized valence state of Cu2+ and the modulation of its lattice oxygen to oxygen vacancy ratio. In another work reported recently, it was demonstrated that Mg-incorporated ZIF-8 (Mg/ZIF-8) could achieve 75.7% Gly conversion and 40% GC yield at 175 °C and 0.3 Mpa CO2 for 6 h, upon using acetonitrile as dehydrating agent [42]. The relatively high catalytic activity of Mg/ZIF-8 might be mainly attributed to the fact that the incorporated Mg2+ could result in structural disruption, thus generating more exposed active Mg2+ sites and coordinatively unsaturated nitrogen atoms, which could considerably boost the Lewis basicity of the catalyst. Despite this progress, more efforts are still required to conduct the efficient synthesis of GC from the direct carbonation route. The main challenge of this route is to develop more efficient heterogeneous catalysts which can work well (with high activity, selectivity, and stability) under the optimized operation conditions (suitable reaction temperature and pressure).
The coupling of CO2 and glycerol-derived glycidol (GD) has been regarded as a more appropriate route for GC synthesis due to its relatively high atom economy and mild reaction pathway [24]. Thermodynamic analysis reveals that, unlike the direct carbonation of Gly with CO2, the coupling reaction between GD and CO2 to form GC is thermodynamically spontaneous (ΔrGθm < 0). The activation energy barrier predominantly depends on the synergistic mechanism between CO2 activation and epoxide ring opening [44]. Crucially, the operated reaction temperature requires precise modulation: an insufficient thermodynamic driving force prevails at lower temperatures, while elevated temperatures promote the side reactions (e.g., hydrolysis, polymerization) or product decomposition. Consequently, optimizing temperature to reconcile thermodynamic and kinetic constraints could substantially enhance the coupling efficiency between GD and CO2 for GC synthesis in the case where an effective catalyst is adopted.
Alternatively, to improve the efficiency of GC generation, an indirect pathway can be adopted, for instance, converting CO2 to urea or DMC firstly and then reacting it with Gly to synthesize GC. This cascade reaction pathway might greatly reduce the process’s energy barriers. As reported in the literature [44], the transesterification route (glycerol + DMC) exhibits near-equilibrium thermodynamics (ΔrGθm = −1.83 kJ/mol at T = 298.15 K and 101,325 Pa), suggesting that this process is thermodynamically favorable. Notably, the chemical equilibrium constant increases with elevated temperature, indicating that adopting a heating strategy can effectively improve the thermodynamic driving force. Furthermore, removing the by-product of methanol may further overcome the equilibrium constraints, driving the reaction forward and thereby enhancing GC yield.
The thermodynamic analysis of Gly–urea carbonation reveals the positive Gibbs free energy (ΔrGθm = 32.43 kJ/mol) under ambient temperature and pressure conditions (T = 298.15 K, p = 101,325 Pa), confirming the non-spontaneous nature of this reaction pathway. Reduced system pressure facilitates the ammonia evolution step, thereby promoting GC formation. Further analysis of the temperature dependence indicates that the reaction equilibrium constant increases substantially with rising temperature. Consequently, the synergistic effect of elevated temperature and reduced pressure is thermodynamically imperative to shift the equilibrium toward products. Moreover, the prompt removal of gaseous ammonia by-products remains equally critical to circumventing equilibrium limitations and achieving viable process yields.
Overall, the reaction feasibility of GC synthesis depends on balancing thermodynamic driving forces (equilibrium conversion), by-product management (e.g., water/methanol removal), and operational conditions (temperature/pressure). Catalyst choice could play a crucial role in overcoming the kinetic barriers while maintaining high GC selectivity under optimized thermodynamic constraints.

3. Synthesis of GC from the Coupling of CO2 and Glycerol-Derived Glycidol

As mentioned above, the direct carbonation of Gly with CO2 is the most attractive path for producing GC, since it is an ideal process targeting high-value-added chemicals and feeding upon bio-based Gly renewable sources and waste effluent CO2 [46,47,48,49,50]. However, this route is constrained by both thermodynamics and kinetics, and high reaction pressure and dehydrant are generally required in order to obtain a higher yield of GC. Currently, more efficient catalysts and operating processes are still desired for promoting industrial implementation, which remains a challenging research subject in the field of catalysis and the chemical industry.
Alternatively, the coupling of CO2 and glycerol-derived glycidol (GD) has been proposed as an economically viable process [24,51,52]. Compared to Gly, GD could react with CO2 to form general GC under much milder conditions. Traditionally, GD could be produced by using petroleum-based Gly as a raw material through a multiple-step route in the presence of Brønsted acid and Lewis acid [27,28,29,30,31]. In general, Gly is first converted into an active intermediate of chloropropanediol (or allyl alcohol) via methods such as chlorination (or oxidation) in the presence of catalysts like pimelic acid (or HZSM-5 zeolite), then GD can be obtained by further dichlorination (or epoxidation) of the corresponding intermediate. Significally, Monbaliu et al. developed a tandem continuous-flow process where Gly is first converted to chlorohydrins via concentrated aqueous hydrochloric acid and using pimelic acid as catalyst, then dehydrochlorinated with a concentrated aqueous solution of NaOH to yield a GD/epichlorohydrin mixture, with integrated extraction enhancing separation. This route seems quite attractive considering the fact that an intensified and scalable flow process for effectively converting Gly into GD has been established [32]. In addition, recent progress has also revealed many types of catalysts, including metal catalysts, organic catalysts, and ionic liquid catalysts, that can catalyze the carbonation of GD with CO2 to GC with relatively high yields [53,54,55,56,57,58,59,60,61]. We list some of the catalysts for the coupling reaction of CO2 and GD in Table 1. Special attention has been paid to having intelligent control over the microenvironment of the catalysts to achieve a desirable performance, and to the integration and application of the technological processes [62,63,64].
For instance, a simple and clean approach has been found to precisely control the defect density in Metal–Organic Frameworks (MOFs) to address the requirement of the carbonation of GD with CO2 [65]. Through functional group engineering on the linkers, the researchers systematically fabricated a series of Ce-UiO66-X materials with varying concentrations of ligand defects (Figure 1a–c). The four isomorphic Ce-UiO-66-X samples (where X = OMe, Me, H, F) generated in this way have flexibly adjustable ligand defect densities (that is, unsaturated Ce sites). This characteristic is beneficial for establishing the structure−activity relationship. Among them, CeUiO-66-F, which contains the superior ligand defect density, presents the highest GC yield of 96% under mild reaction conditions, whereas Ce-UiO-66-OMe, with the lowest defect density, shows the lowest yield of 54% (Figure 2a).
To comprehensively evaluate the catalytic performance, the activation energies (Ea) for the target reaction over Ce-UiO-66-X were calculated by the kinetic equations. The activation energy was calculated by analyzing the linear relationship between ln k and 1/T (Figure 2b). An inverse correlation was observed between the MOF ligand defect concentration and Ea (Figure 2c), indicating that coordinatively unsaturated Ce sites are responsible for catalyzing CO2 cycloaddition.
Both experimental and theoretical studies have shown that the electron-withdrawing/donating properties of linker functional groups significantly influence the binding affinity between Ce-oxo clusters and BDC-X ligands, thereby enabling the tuning of ligand defect densities. When electron-donating groups (such as OMe and Me) are introduced to terephthalic acid (BDC), the coordination bonding strength between the Ce-oxo node and the BDC linker in Ce-UiO-66-OMe and Ce-UiO-66-Me is weakened. In contrast, electron-withdrawing substituents (e.g., fluorine) strengthen metal–ligand coordination bonds, leading to varied ligand defect concentrations (2.0–29.5%) in CeUiO-66-F (Figure 1b). Notably, increased defect density substantially distorts MOF crystallinity while simultaneously decreasing both the coordination number and oxidation state of Ce centers (Figure 1c). Furthermore, the inverse relationship between defect concentration and activation energy for CO2 cycloaddition confirms that ligand-defect-generated unsaturated Ce sites enhance catalytic performance.
By combining the DFT calculations, a deep mechanistic understanding of the aforementioned correlation has been obtained. As shown in Figure 2d, the carbonation process may undergo three transition states to obtain the main product of GC. The high-defect model (defect-H, two missing linkers) demonstrates greater exothermicity compared to the low-defect model (defect-L, one missing linker), despite showing comparable activation energies (DE = 6.30 vs. 6.92 kcal/mol). Notably, glycidol adsorption at the unsaturated Ce site in defect-H yields a more favorable binding energy (−21.48 kcal/mol), indicating enhanced reactivity at higher defect concentrations. Furthermore, the entire carbonation pathway remains more exothermic for defect-H, corroborating experimental observations that structural defects facilitate the CO2–glycidol coupling reaction.
This work represents the first attempt at establishing a convenient and clean methodology to accurately control the defect density in MOFs. However, the strategy is primarily validated for UiO-66-type MOFs; its applicability to other MOF topologies or more complex reactions remains unexplored, which requires further validation for broader applicability and practical robustness.
In another work reported by Monbaliu et al., they developed a continuous flow process toward the large-scale production of GC, relying on readily available nitrogen-containing Barton’s base as a potent homogeneous organocatalyst, which is an extension and in-depth study of the work previously reported in bio-based Gly to produce GD [26].
The effect of temperature, CO2 flow rate, and counter-pressure on the catalytic conversion of GD (2) to GC (3) was first investigated in a microfluidic reactor with 5 mol% of TBD (1,5,7-Triazabicyclo[4.4.0]dec-5-ene) in MEK (methyl ethyl ketone). The reaction conditions were optimized, involving processing at 140 °C (10 bar) with a 2 min residence time and 10 mLN/min of CO2, which struck a balance between yield and selectivity under intensified conditions (Figure 3a). Subsequently, a range of homogeneous organic bases were screened as organocatalysts, such as TMG (1,1,3,3-tetramethylguanidine), DBU (1,8-diazabicyclo[5.4.0]undec-7-ene), BB (2-tert-butyl-1,1,3,3-tetra-methylguanidine), and so on. Among them, BB showed the best catalytic performance, with a 91% GC yield in the case of a 5 mol% loading of the catalyst (Figure 3b).
The scalability of the carbonation process was then determined in commercial mesofluidic glass reactors (Corning Advanced-Flow Reactors) in the temperature range of 120–150 °C and 10 bar with a feed solution of GD (1.8 M in MEK) and 1 equiv. of CO2 in the presence of 1 mol% of BB. Similar trends to the microfluidic experiments were obtained. After that, the carbonation process was further evaluated in a pilot mesofluidic reactor with BB, and the typical flow chart is present in Figure 4. Further intensification could lead to a 78% GC yield within a much shorter residence time (i.e., 28 s) at 140 °C. This result outperforms all conditions reported in the previous literature. The significance of this process is emphasized by its high throughput and low footprint, leading to an E-factor of 4.7 and an STY (Space Time Yield) of 2.7 kg h−1 L−1 for the upstream carbonation process.
DFT calculations were employed to explain the necessity of both a β-hydroxyl group on the substrate and an organocatalyst with significant Brønsted basicity. Computation results highlighted the distinctive characteristics of a double H-shuffle mechanism for the coupling of CO2 and GD following a two-step route (Figure 5a). For step 1, the basicity of the catalyst is of the utmost significance (Figure 5b). In fact, a high pKaH guarantees a rapid and favorable CO2 fixation, resulting in the formation of a highly stable conjugated acid along with a linear carbonate intermediate. In step 2, the presence of the conjugated acid significantly accelerates the reaction (Figure 5c). This conjugated acid serves as a general Brønsted acid catalyst to activate the epoxide, although step 2 remains the overall rate-determining step. Opposing trends are seen in the activation energies of the two carbonation steps. This finding offers an understanding of why Barton’s base showed the most favorable outcomes for the entire transformation process. When a stronger base is used, it gives rise to a weaker conjugate acid, which then acts as a less efficient Brønsted acid for epoxide activation.
The above results verified that the carbonation of GD with CO2 was validated at the pilot scale in a commercial mesofluidic reactor upon using Barton’s base catalyst. Relatively high GC selectivity and yield could be achieved within a very short residence time with a low catalyst loading (1 mol%), demonstrating the unprecedented productivity of this intensified continuous flow process. This work represents a substantial enhancement of the existing conditions toward the intensified and scalable production of GC. Further efforts are still required in order to achieve large-scale industrial production, mainly involving the catalyst separation/recycling, solvent choices, and process optimization. In addition, the mechanistic insights should also be strengthened for promoting future catalyst development and the industrialization process.
In short, the coupling of CO2 and glycidol has shown significant potential for the industrial production of GC. Considerable efforts are still required in order to design and develop more efficient, stable, and recyclable catalysts for this process. To achieve this aim, it is necessary to deeply explore the catalytic mechanism and to analyze the key structural factors driving the generation of GC, which can provide guidance on further optimizing the catalytically active sites of the catalysts. Future research should also focus on driving technological development for producing glycidol from waste Gly, and on exploring low-cost, sustainable, and reusable catalysts that can adapt to low-purity glycidol and CO2 in industrial waste gas. Meanwhile, scalable process technologies need to be introduced in the early stage of technological research and development to promote the industrialization process in this field.

4. Synthesis of Glycerol Carbonate from Glycerol and Organic Carbonates

Currently, the transesterification of organic carbonates with glycerol (Gly) is a major route for industrial production of GC [68,69]. Several carbonates like ethylene carbonate (EC), dimethyl carbonate (DMC), propylene carbonate (PC), and diethyl carbonate (DEC) could be used for the transesterification process. Among them, the transesterification of Gly and DMC represents a highly attractive and sustainable route to GC, largely owing to DMC’s status as a “green reagent” [70,71]. In general, the transesterification of DMC with Gly could be conducted in the presence of homogeneous alkaline catalysts such as calcium oxide and sodium hydroxide [12,13,14,15]. In order to overcome the drawback of the homogeneous catalysis system, more recent effort has been devoted to the development of efficient heterogeneous catalysts with enhanced recyclability [5,72,73]. To enlarge the scale of industrial implementation, it also needs to decrease the production cost of the reagents, since the organic carbonates, including DMC, are relatively expensive in comparison with other carbonylation agents (e.g., urea).
So far, a variety of metal oxides [74,75], mixed metal oxides [76,77], hydrotalcites [78,79], and zeolitic materials [80,81] have been investigated as solid catalysts for the transesterification reaction. We have listed in Table 2 a selection of catalysts mentioned in the transesterification of glycerol and organic carbonates to facilitate a literature data comparison. However, most of them showed unsatisfied catalytic efficiency and/or poor recyclability. In addition, the nature of the catalytically active sites and the catalytic mechanisms remain inadequately explored. This knowledge gap can impede the rational design of cost-effective, high-efficiency catalysts for this transesterification reaction.
Recently, Fan et al. [82] synthesized a series of La2O3-based catalysts doped with alkali and alkaline earth metals (M/La2O3), and systematically investigated their catalytic performance for the transesterification of DMC and Gly. In their work, a simple wet impregnation strategy was employed to modify La2O3 with alkali and alkaline earth metals, and synthetic mechanisms were investigated through Na/La2O3 catalysts (Scheme 1). During impregnation, Na+ anchors onto the La2O3 carrier via pore diffusion and surface adsorption, while partial ions can form surface ion pairs at the oxygen/hydroxyl sites. During the drying process, solvent evaporation induces dopant migration and promotes crystalline reconstruction, enabling the formation of uniformly distributed Nax(NO3)y(OH)z crystals on La2O3 [or La(OH)3/La2O2CO3] surfaces. During calcination, Nax(NO3)y(OH)z decomposes into a molten salt phase, facilitating the incorporation of Na+ into the carrier lattice. The final products consist of sodium oxide (Na2O) and reconstituted La2O3.
The XRD patterns suggest that the alkali metals dispersed uniformly on the La2O3 surface without forming additional crystalline oxide phases. In contrast, the alkaline-earth-metal-doped La2O3 samples exhibited much weaker and broader diffraction signals, accompanied by the appearance of a CaO, SrCO3, or BaCO3 phase, indicating that heavy alkaline earth metal doping could inhibit the formation of the La2O3 phase. The XPS profiles shown in Figure 6 reveal the changes in the binding energies of the La 3d and O 1s spectra induced by alkali metals or alkaline earth metals. The La 3d5/2 peak shifted towards a lower binding energy for the alkali-metal-doped catalysts, whereas no obvious shift occurred for the alkaline-earth-metal-doped samples. Meanwhile, the O 1s binding energies shifted towards higher values for both types of doped catalysts, and the shift induced by alkaline earth metals was more pronounced, indicating that alkaline earth metals have stronger interactions with the O species existing in La2O3. In addition, the CO2-TPD measurements revealed that doping alkali or alkaline earth metals could significantly enhance the surface basicity of La2O3. Compared to their alkaline-earth-metal-doped analogs, the La2O3 samples modified with alkali metals exhibited a higher density of basic sites.
Among all the doped La2O3 catalysts, Na/La2O3 exhibit optimal catalytic performance (85% Gly conversion, 60% GC yield) for the transesterification of DMC with Gly. For La2O3 catalysts doped with different metals (Figure 7a), the catalytic performance shows no proportionality to the surface basic site concentration. For instance, Na/ La2O3 and Ca/ La2O3 exhibit comparable basic site densities (5.60 vs. 5.34 μmol/m2), yet Na/La2O3 achieves about 40% higher Gly conversion than Ca/La2O3. Similarly, the basic site densities of Li- and K-doped catalysts are also close to those of Na-doped La2O3 catalysts, but deliver significantly lower Gly conversions and GC yields. Figure 7b shows the relationship between glycerol conversion and the ionic radius ratio of dopant and La, which could illustrate more clearly the reason why Na/La2O3 exhibits optimal catalytic performance. This enhancement may stem from the close match between the Na+ ionic radius (1.02 Å) and the La3+ ionic radius (1.03 Å), which promotes stable active site formation. In contrast, dopants with ionic radii significantly smaller or larger than that of La3+ may cause structural deformation around the doped site, thus affecting the stability of the active site. Furthermore, XPS spectra (Figure 6) demonstrate that alkali metals are more likely to donate their electrons than alkaline earth metals, while alkaline earth metals predominantly modulate electron distributions around O sites due to the extra phase of the alkaline earth metal oxides formed. This electronic behavior correlates with catalytic activity trends wherein alkali-metal-doped La2O3 catalysts exhibit superior performance. These results suggest that the ionic radius and valence of the dopant are the decisive factors, whereas the surface basicity, although crucial for the activation of Gly/DMC, does not determine the catalytic performance of the metal-doped La2O3 catalysts.
The recyclability of the spent Na/La2O3 catalyst was tested, showing that the catalyst could be reused with a slight decrease in catalytic activity after two cycles. To investigate the deactivation reason, the effects of the regeneration solvent and Na+ leaching were evaluated by additional experiments. The results indicated that the decrease in activity was due mainly to the surface wearing rather than the regeneration process or Na+ leaching. The surface wearing of the catalyst was mainly caused by the friction during the high-rate-stirring process, which was required for uniform dispersion in the immiscible Gly/DMC phase.
As shown in Figure 8, the plausible reaction mechanism for Gly and DMC transesterification over Na/La2O3 catalysts was proposed. The carbonyl group of DMC and the hydroxyl group of Gly undergo activation at the La and O sites, respectively. Subsequently, the activated Gly anion attacks the carbonyl carbon of activated DMC, forming the intermediate of 1-(o-methoxy-carbonyl) glycerol complex (intermediate 1) and methanol. Intermediate 1 undergoes cyclization to form GC with another methanol molecule. The incorporation of alkali metals creates M-La centers that facilitate DMC activation and intermediate 1 cyclization, consequently leading to a higher GC yield than the pristine La2O3.
The authors believed that this research could help to understand the promotional role of alkali metals and alkaline earth metals and that the results may guide the future design of more efficient and stable metal oxide catalysts for the transesterification of DMC and Gly to produce GC. It is worth noting that the plausible mechanism (Figure 8) is proposed based on the literature and has not yet been verified by in situ characterization of the intermediates (e.g., in situ FTIR). Verifying these intermediates would significantly advance the understanding of the reaction mechanism for alkali-/alkaline-earth-metal-doped oxides, offering new insights for rational catalyst design.
As part of the ongoing development of green GC synthesis methods, Wang et al. [83] recently utilized an easily available and biocompatible sodium citrate (Na3-citrate) catalyst for the efficient production of GC and ethylene glycol (EG) through the transesterification of Gly and ethylene carbonate (EC) (Scheme 2).
Under the test conditions, 0.01 wt% sodium citrate demonstrates exceptional catalytic activity, achieving >99% Gly conversion and >99% GC yield with a very high turnover frequency (TOF) of 16,925 gGC·g−1Cat·h−1. The reaction time screening reveals that 45 min suffices for a >99% GC yield, while prolonging it to 60 min maintains >99% yields of both GC and EG. Notably, increasing the EC/Gly molar ratio from 1:1 to 4:1 elevates the GC yield from 58% to >99% (Figure 9). Furthermore, a kinetic model for the transesterification reaction was established, revealing an activation energy of 51.3 kJ⋅mol−1 for the Na3-citrate catalyzation process. Combining the experimental results and the DFT calculations, it is demonstrated that Na3-citrate promotes transesterification by forming strong hydrogen bonds with Gly.
The FT-IR spectra show that the hydroxyl vibrational band of Gly is red-shifted in the Gly-(Na3-citrate) complex, along with there being a decrease in peak intensity. It is tentatively suggested that Gly can be activated by sodium citrate through hydrogen bonding interactions. DFT calculations revealed that the O1-H1 bond length of Gly increased from 0.962 Å to 0.986 Å in the optimized structure of the Gly-(Na3-citrate) complex, as shown in Figure 10a,b. ESP analysis showed in Figure 10c,d that the activated (Gly) exhibited a stronger negative potential near the O1 atom than that of the inactivated (Gly) (−29.94 vs. −22.99 kcal·mol−1). The interaction between Gly and sodium citrate is mainly realized through a strong hydrogen bond (blue part in Figure 10e) as a way to achieve the goal of activating the hydroxyl group of Gly. The characterization results reveal that Na3-citrate activates the hydroxyl group of Gly, enhancing the nucleophilic attack on the electrophilic carbonyl carbon of EC to facilitate the transesterification process.
Notably, DFT calculations elucidated the potential energy profiles, optimized geometries, and possible catalytic mechanism for the Na3-citrate-catalyzed transesterification of Gly and EC (Figure 11a), which provide a clearer and more intuitive understanding of the reaction process. Reaction path analysis reveals a succession of steps separated by four transition states (TS2-1 to TS2-4) and five intermediates (IM2-0 to IM2-4). IRC analysis (Figure 11b) describes in detail the process of bond breaking and formation at each step: First, the formation of a strong hydrogen bond between the H1 atom of Gly and the O1 of the -COO- group of Na3-citrate leads to the breaking of the H1-O1 bond of Gly, and at the same time, the carbonyl oxygen (O1) of Gly becomes close to the carbon (C1) of EC. Subsequently, new O-H (O2-H1) and C-O (O1-C1) bonds are formed in IM2-1. The formation of IM2-2 involves the breakage of the C1-O4 and H1-O3 bonds, accompanied by the formation of the O4-H1 bond. After IM2-2 undergoes the breakage of the H2-O5 bond, both O5-C1 and O3-H2 bonds are formed to give the intermediate IM2-3. Finally, IM2-3 breaks the C1-O6 bond and the H2-O3 bond, forming the O6-H2 bond and giving rise to the intermediate IM2-4.
In short, this work revealed that the readily available Na3-citrate could act as an efficient catalyst for the transesterification of Gly and EC to produce GC. The established kinetic model confirms that the reaction could be easily achieved under a lower activation energy of 51.3 kJ⋅mol−1. DFT calculations demonstrated that the catalyst of Na3-citrate facilitated transesterification via the activation of Gly through a strong hydrogen bond interaction. The authors believed that this study may have offered an effective strategy for the catalytic transformation of Gly to GC. However, as a homogeneous catalyst, sodium citrate still presents challenges in post-reaction separation. Future research should focus on developing sodium citrate immobilization techniques (e.g., on magnetic carriers), exploring continuous flow processes, optimizing by-product control, and conducting economic analysis.
Although significant progress has been made in the transesterification synthesis of GC, the development of novel catalysts with high activity, excellent cyclic stability, and low cost remains a core challenge in this field. To solve this challenge, it is necessary to deeply analyze the micro-mechanisms of multi-step reactions in heterogeneous catalytic systems, and to elucidate the reaction pathways, deceleration steps, and inactivation kinetics through in situ characterization combined with kinetic simulations. This would be helpful to integrate the catalyst design with the reaction mechanism, and to then guide the precise design of active sites and the synergistic optimization of reaction parameters. In addition, much effort should be devoted to further reducing feedstock costs and advancing industrial sustainability. For instance, the feedstock of DMC could be produced from the captured CO2, thus building a closed-loop industrial chain of “carbon capture–chemical conversion–high-value product synthesis”. In this case, the Gly-DMC route would become a truly green, intelligent, and low-cost industrial production process for GC.

5. Catalysts for Glycerol Carbonate Synthesis from Glycerol and Urea

The carbonylation of Gly with urea to produce GC has gained considerable attention for its mild conditions, low cost, and easy product separation [6,7,84]. We list some of the catalysts for the carbonylation of Gly with urea in Table 3. Up to now, various catalysts have been tested for this process, including zinc-based catalysts [8,38,85,86], magnesium-based catalysts [87], tungsten-based catalysts [88,89], ionic liquid catalysts [90,91,92], and so on. In particular, zinc-based catalysts dominate this field due to their tunable Lewis acidity, stability, and availability. Inorganic components such as ZnO, ZnCl2, and ZnSO4 have shown a tendency to achieve high Gly conversion and GC selectivity. This tendency stems from the easy reactivity of Zn species with urea, which causes them to dissolve in the reaction medium. In order to solve the separation problem of these homogeneous catalysts, various solid Zn-based catalysts have been prepared and investigated for the carbonylation reaction. However, most Zn-based catalysts claimed to be heterogeneous fundamentally operate homogeneously [33,36,85,93,94,95].
Consequently, the design of novel heterogeneous catalysts, characterized by an abundance of stable and efficient high-density active sites, remains a highly attractive endeavor in the field of glycerol carbonylation. Some new progress has been made in GC synthesis in terms of catalyst design, characterization, and reaction mechanism studies. For instance, Shin and his collaborators synthesized disordered ZnAl2O4 spinel catalysts using the polymeric citrate complex method with various polyol precursors, providing a conceptual representation of the partially inverted ZnAl2O4 catalysts [96].
As shown in Figure 12, long-chain polyols with multiple adjacent hydroxyl groups increased disordered ZnAl2O4 spinel formation. The disorder degree in the disordered ZnAl2O4 spinel structure was determined by the order parameters and lattice parameters shown in Figure 12b.
The differences between a normal ordered spinel lattice structure and a partially inverted one, which arise from the substitution of Zn2+ and Al3+ ions. The incorporation of ZnO6 and AlO4 sites via partial inversion enhances structural disorder and promotes surface acidity (Scheme 3). The surface AlO4 sites are associated with varying degrees of acidity, including medium, medium-strong, and strong acidic sites. In addition, the disordered s–ZnAl2O4 structure generated both acidic and basic sites. The superior performance of the s–ZnAl2O4 catalyst, as evidenced by its high GC yield and low (2) selectivity of 41% and 25%, respectively, further highlighted the relationship between the surface properties and the catalytic properties of the catalysts, as presented in Figure 13.
The heightened degree of the disordered ZnAl2O4 structure makes it easier to form more acidic sites (Figure 14). These acidic sites function as active sites for the urea molecule during the system. Furthermore, the disordered ZnAl2O4 generates more basic sites. These basic sites act as active sites for converting glycerol to accelerate the degradation of intermediate (2) and the formation of the main product GC. Consequently, the highly disordered s–ZnAl2O4 catalyst exhibits increased acidic and basic sites. These sites critically promote urea glycerolysis and subsequent GC generation, achieving the highest GC yield.
The above study demonstrates that polyols with longer chains and adjacent hydroxyl groups (e.g., sorbitol) may enhance the structural disorder of ZnAl2O4, thus leading to higher surface acidity and catalytic activity for GC production. However, the catalytic performance of the disordered ZnAl2O4 still needs to be further improved, since the selectivity of the ally by-product (i.e., 2,3-dihydroxypropyl carbamate) is relatively high (~25% selectivity) under the test conditions.
In another study reported by our research group, it was discovered that a single Zn atom catalyst supported on N-doped porous carbon (Zn1/NC) demonstrated outstanding catalytic activity and stability in the carbonylation of Gly with urea to yield GC [97]. As depicted in Scheme 4, the Zn1/NC catalyst was fabricated by immobilizing zinc salt on a type of N-doped porous carbon material (NC), followed by a thermal treatment procedure.
The single-atom structure of the Zn species could be validated through HR-TEM imaging, XANES, and EXAFS measurements (Figure 15). The Zn1/NC shows a single intensity peak at approximately 3.75 Å−1, which can be attributed to the Zn-N/O contribution in the first nearest-neighbor coordination shell. The absence of detectable Zn-Zn signals indicates that the zinc atoms are isolated. These zinc atoms are anchored to the carbon support via Zn-O and/or Zn-N bonds. The fitting result of EXAFS suggests that the Zn atom is in a hexagonal coordination environment. Based on the structural characteristics of the NC support, Zn species in the Zn1/NC catalyst likely adopt a coordination structure where single Zn2+ ions are coplanarly coordinated to four N/O atoms from pyridine-like and carbonyl-like groups. This may involve additional weak coordination with two perpendicular O atoms from water molecules, forming a Zn1-N4-xOx(H2O)2 configuration.
Interestingly, the Zn1/NC catalyst demonstrates exceptionally high catalytic efficiency, achieving 94.8% glycerol conversion and 95.0% GC selectivity within 6 h, resulting in an impressive GC yield of 90.1%, and could be reused at least five times without any noticeable decline in either catalytic activity or GC selectivity (Figure 16).
In principle, seven possible single-Zn-site models (Zn1-N4-xOx, where x = 0–4) can be established using DFT, as illustrated in Figure 17a. It is observed that only the adsorption energy (−2.42 kcal/mol) of urea on Zn1-N2O2-1 is lower than that of glycerol (7.93 kcal/mol), indicating that urea molecules are more easily adsorbed on a single Zn site with this configuration (Figure 17b). In the other cases, the single Zn sites preferentially adsorb glycerol molecules. Hence, it is reasonable to infer that the Zn1-N2O2-1 model represents the preferred configuration of the primary active sites within the Zn1/NC catalyst.
The FT-IR (3D-FTIR) profiles are also studied to obtain the clear evolution process of the carbonylation reaction in three-dimensional conditions, indicating that the Zn1/NC catalyst demonstrates a significantly greater capacity for effectively adsorbing and activating urea to produce the active intermediate and the primary product GC.
The study achieved highly efficient, selective, and stable catalytic performance in the glycerol carbonylation reaction with urea by designing a nitrogen-doped, porous, carbon-supported, and single-atom Zn catalyst. The mechanistic investigation was thorough, demonstrating significant academic value and application potential. However, the catalyst preparation involved multiple steps, including high-temperature carbonization, acid treatment, and wet impregnation, which may have increased the cost and difficulty of large-scale production.
Similarly, Butburee et al. reported that a kind of transition metal single-atom-loaded zinc oxide quantum dots (M1-ZnO QDs), which are synthesized by facile wet chemical synthesis (Figure 18), may act as an efficient catalyst for the carbonylation reaction [98]. Their structure has been downsized to a 0 D quantum dot to maximize the exposure of the active sites and promote solubility/homogeneity with the reactants. The trace amount of single metal atoms creates significant numbers of defects on ZnO QDs, which are beneficial for the carbonylation of Gly and urea molecules. As revealed by the TEM images (Figure 18c,e), M1-ZnO QDs (M = Fe, Co, Ni, and Cu) are uniform with an average size of 5.0 ± 0.6 nm, and the metal dopants are incorporated in the ZnO structure in an atomically dispersed form. In addition, M1 could substitute some Zn atoms, as confirmed by the high-resolution TEM images (Figure 18e, f). And the substitution of some Zn atoms by the metal dopants (M1) could generate a significant number of defects in the structure, as evidently indicated by the dark-contrast spots, suggesting that some atoms are replaced or removed from the plan (red arrows in Figure 18f).
The defective structure of the doped ZnO QDs could be further confirmed by XRD, XAS, and XPS results (Figure 19a–f). This feature is consistent with the previous reports, showing that a trace amount of Fe doping in ZnO can appear in a single hexagonal wurtzite phase, in which Fe replaces the Zn atom to be incorporated into the lattice oxygen atoms in ZnO in the form of Zn1−xFexO [99].
As shown in Figure 20, M1-ZnO QDs exhibited both high activity and high selectivity for the carbonylation reaction, compared to the pristine ZnO QDs. Fe1-ZnO QDs with enriched oxygen vacancies and isolated Fe atoms especially showed the highest performance among the investigated metals (Fe, Co, Ni, and Cu), with a 90.7% Gly conversion and a nearly 100% GC selectivity, respectively (Figure 20a,b). The additional cycling experiments showed that the recovered catalyst was structurally stable, although the catalytic performance of the spent catalyst decreased somewhat during the recycling test, possibly related to the difficulty in separating the tiny particles from the reaction.
The catalytic behaviors were insightfully investigated by in situ Diffused Reflectance Infrared Fourier Transform Spectroscopy combined with the DFT calculations (Figure 20c,d). The congruential results verified that the superior catalytic efficiency could be mainly attributed to the enriched Lewis acidity at Zn around the Fe1-Vo, which facilitated the formation of *GlyU, the key intermediate for the formation of GC. In addition, this active site was supported on the tiny ZnO QDs that could maximize the exposure to the reactants, leading to an impressive Gly-to-GC efficiency.
This study demonstrates that the highly efficient and selective conversion of Gly to GC could be conducted by designing Lewis acid defect sites through single-atom doping on ZnO quantum dots. However, the small size of the ZnO quantum dots (5 nm) may bring serious separation problems; this might be the main reason why approximately 20–30% of catalyst loss has been detected during the recycling tests, which is unfavorable for achieving practical applications.
In short, the above results suggest that the precise regulation of the structure and properties of the active centers may create highly active and stable heterogeneous Zn-based catalysts for the carbonylation of Gly with urea. In spite of this progress, more efforts are still required in order to further optimize the catalytic performance, especially the stability/recyclability of the heterogenous catalysts. In addition, some model catalysts with defined coordination structures and environments (e.g., Zn-containing complexes) should be synthesized to simulate the structure of active Zn species, which can be useful for deeply understanding the nature of catalytically active sites and reaction mechanisms. Moreover, integrated technologies, such as coupling the carbonylation processes with the recycling of the by-product ammonia, are highly attractive for developing more economic and effective GC production processes.

6. Conclusions

In this review, we summarize the recent research advances in the catalytic production of glycerol carbonate from bio-based glycerol involving different synthesis routes, including the coupling of CO2- and glycerol-derived glycidol, the transesterification of glycerol with organic carbonates, and the carbonylation of glycerol with urea. Significant progress has been made in catalyst design, reaction mechanism exploration, and catalytic process optimization. By choosing suitable catalyst preparation methods and combining this with catalytic process optimization, some novel homogeneous and heterogeneous catalytic systems with improved catalytic efficiency and stability/recyclability have been developed, which can achieve higher glycerol conversion and glycerol carbonate yields under optimized reaction conditions. This progress, though still quite limited, demonstrates clearly the great feasibility for the implementation of large-scale industrial applications in the near future. By combining a variety of characterization results and DFT calculations, the nature of the catalytic actives and the catalytic reaction mechanisms have also been explored; the obtained knowledge could guide the precise design and construction of more efficient catalysts for glycerol carbonate production.
To further promote the development of the related research and application fields, several challenges and opportunities involving catalyst preparation, reaction mechanism studies, and catalytic process optimization shall be addressed in future work. Firstly, maintaining the high catalytic efficiency of heterogeneous catalysts in long-term operations is challenging, especially in harsh reaction conditions (i.e., higher temperatures and pressures). This requires continuous innovation in catalyst design and preparation through systematically tuning the microstructure, composition, and surface acidity–basicity of the active centers.
Secondly, it is necessary to deeply explore the catalytic mechanisms involving a special reaction system for further guiding the design and preparation of new catalytic materials. For instance, there is still significant controversy regarding the understanding of the nature of active centers and reaction mechanisms for the heterogeneous Zn-based catalysts applied in the carbonylation of glycerol with urea. To deal with this issue, some model catalysts like Zn-containing complexes, which have a defined coordination structure and microenvironment, should be synthesized. This would be helpful for precisely simulating the structure of active Zn species and revealing the mechanism of active centers in multi-step reaction processes through the combination of various in situ characterization techniques, kinetic experiments, and DFT calculations.
Thirdly, much effort should be devoted to conducting the catalytic synthesis of glycerol carbonate using the true bio-based glycerol (crude biogenic glycerol) as a raw material. For a specific catalytic system, the utilization of the crude biogenic glycerol might be challenging, since it usually contains methanol, water, salts, fatty acids, and soaps, which may bring various negative effects, such as poisoning catalysts, promoting side reactions, and reducing product yields. In addition, the impurity profile varies depending on the biodiesel feedstock (e.g., vegetable oils vs. animal fats), leading to an unpredictable influence on the catalytic performance. Removing these contaminants (e.g., distillation, ion exchange) can certainly increase the energy consumption and process complexity, reducing economic viability. Hence, to enable large-scale deployment, it is necessary to develop cost-effective purification processes for a given bio-based glycerol, and to advance the catalytic technologies for the direct conversion of crude glycerol.
Moreover, it is highly desirable to integrate the related industrial production technologies, such as coupling the production process of glycerol carbonate (e.g., carbonylation or transesterification process) with the production process of the feedstock, for instance, to produce urea or DMC by reacting the by-product (ammonia or methanol) with the captured CO2. This would be highly attractive for developing a truly green, intelligent, and low-cost industrial production process for glycerol carbonate. Therefore, there is still a very large space and many opportunities in catalyst innovation, reaction mechanism research, and catalytic process optimization, which shall be essential to address the current challenges faced by the catalytic production of glycerol carbonate and to promote the development of the green chemical industry.

Author Contributions

H.S.: investigation and writing—original draft; Z.L.: formal analysis; J.S. and M.J.: writing—review, editing, and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Nature Science Foundation of China (22172058) and the Project of the Education Department of Jilin Province, China (No. JJKH20240562KJ).

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
Glyglycerol
GCglycerol carbonate
GDglycidol
DMCdimethyl carbonate
ECethylene carbonate
DECdiethyl carbonate
PCdiethyl carbonate
DFTdensity functional theory
MOFsMetal–Organic Frameworks
BDCterephthalic acid
MEKmethyl ethyl ketone
TOFturnover frequency

References

  1. Quispe, C.A.G.; Coronado, C.J.R.; Carvalho Jr., J.A. Glycerol: Production, Consumption, Prices, Characterization and New Trends in Combustion. Renew. Sustain. Energy Rev. 2013, 27, 475–493. [Google Scholar]
  2. El Doukkali, M.; Dumeignil, F. Advances in Catalyst Design and Reaction Strategies for Carbon-Neutral Conversion of Bioglycerol to Propylene, 1,2-Propanediol, and Hydrogen. J. Am. Oil Chem. Soc. 2025, 102, 483–508. [Google Scholar]
  3. Yazdani, S.S.; Gonzalez, R. Anaerobic Fermentation of Glycerol: A Path to Economic Viability for the Biofuels Industry. Curr. Opin. Biotechnol. 2007, 18, 213–219. [Google Scholar]
  4. Zhou, C.-H.; Beltramini, J.N.; Fan, Y.-X.; Lu, G.Q. Chemoselective Catalytic Conversion of Glycerol as a Biorenewable Source to Valuable Commodity Chemicals. Chem. Soc. Rev. 2008, 37, 527–549. [Google Scholar]
  5. 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]
  6. Li, Q.; Zhang, W.; Zhao, N.; Wei, W.; Sun, Y. Synthesis of Cyclic Carbonates from Urea and Diols over Metal Oxides. Catal. Today 2006, 115, 111–116. [Google Scholar]
  7. Ochoa-Gómez, J.R.; Gómez-Jiménez-Aberasturi, O.; Ramírez-López, C.; Belsué, M. A Brief Review on Industrial Alternatives for the Manufacturing of Glycerol Carbonate, a Green Chemical. Org. Process Res. Dev. 2012, 16, 389–399. [Google Scholar]
  8. Nguyen-Phu, H.; Shin, E.W. Investigating Time-Dependent Zn Species over Zn-Based Catalysts in Glycerol Carbonylation with Urea and Their Roles in the Reaction Mechanism. Appl. Catal. Gen. 2018, 561, 28–40. [Google Scholar]
  9. Pagliaro, M.; Ciriminna, R.; Kimura, H.; Rossi, M.; Della Pina, C. From Glycerol to Value-Added Products. Angew. Chem. Int. Ed. 2007, 46, 4434–4440. [Google Scholar]
  10. Yadav, N.; Yadav, G.; Ahmaruzzaman, M. Catalytic Conversion and Mechanism of Glycerol into Various Value-Added Products: A Critical Review. Ind. Crops Prod. 2024, 210, 117999. [Google Scholar]
  11. Wu, P.-J.; Chiu, H.-H.; Su, Z.-C.; Lin, S.-T.; Yu, B.-Y. Safety Considerations in CO2 Conversion: Production of Glycerol Carbonate via an Indirect Pathway. J. Ind. Eng. Chem. 2025, 147, 744–754. [Google Scholar]
  12. Aresta, M.; Dibenedetto, A.; Nocito, F.; Ferragina, C. Valorization of Bio-Glycerol: New Catalytic Materials for the Synthesis of Glycerol Carbonate via Glycerolysis of Urea. J. Catal. 2009, 268, 106–114. [Google Scholar]
  13. Cholapandian, K.; Gurunathan, B.; Rajendran, N. Investigation of CaO Nanocatalyst Synthesized from Acalypha Indica Leaves and Its Application in Biodiesel Production Using Waste Cooking Oil. Fuel 2022, 312, 122958. [Google Scholar]
  14. Hadiyanto, H.; Lestari, S.P.; Abdullah, A.; Widayat, W.; Sutanto, H. The Development of Fly Ash-Supported CaO Derived from Mollusk Shell of Anadara Granosa and Paphia Undulata as Heterogeneous CaO Catalyst in Biodiesel Synthesis. Int. J. Energy Environ. Eng. 2016, 7, 297–305. [Google Scholar]
  15. Okoye, P.U.; Abdullah, A.Z.; Hameed, B.H. Stabilized Ladle Furnace Steel Slag for Glycerol Carbonate Synthesis via Glycerol Transesterification Reaction with Dimethyl Carbonate. Energy Convers. Manag. 2017, 133, 477–485. [Google Scholar]
  16. Li, H.; Gao, D.; Gao, P.; Wang, F.; Zhao, N.; Xiao, F.; Wei, W.; Sun, Y. The Synthesis of Glycerol Carbonate from Glycerol and CO2 over La2O2CO3–ZnO Catalysts. Catal. Sci. Technol. 2013, 3, 2801–2809. [Google Scholar]
  17. Park, C.; 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]
  18. Manjunathan, P.; Ravishankar, R.; Shanbhag, G.V. Novel Bifunctional Zn–Sn Composite Oxide Catalyst for the Selective Synthesis of Glycerol Carbonate by Carbonylation of Glycerol with Urea. ChemCatChem 2016, 8, 631–639. [Google Scholar]
  19. Del-Mazo-Alvarado, O.; Prieto, C.; Sánchez, A.; Ramírez-Márquez, C.; Bonilla-Petriciolet, A.; Martín, M. An Integrated Process Analysis for Producing Glycerol Carbonate from CO2 and Glycerol. ChemSusChem 2024, 17, e202301546. [Google Scholar]
  20. Kan, L.; Zhang, L.; Dong, L.-Z.; Wang, X.-H.; Li, R.-H.; Guo, C.; Li, X.; Yan, Y.; Li, S.-L.; Lan, Y.-Q. Bridging the Homogeneous and Heterogeneous Catalysis by Supramolecular Metal-Organic Cages with Varied Packing Modes. Adv. Mater. 2024, 36, 2310061. [Google Scholar]
  21. Luo, C.; Wang, J.; Lu, H.; Wu, K.; Liu, Y.; Zhu, Y.; Wang, B.; Liang, B. Atmospheric-Pressure Synthesis of Glycerol Carbonate from CO2 and Glycerol Catalyzed by Protic Ionic Liquids. Green Chem. 2022, 24, 8292–8301. [Google Scholar]
  22. Lei, L.; Wang, Y.; Zhang, Z.; An, J.; Wang, F. Transformations of Biomass, Its Derivatives, and Downstream Chemicals over Ceria Catalysts. ACS Catal. 2020, 10, 8788–8814. [Google Scholar]
  23. Ishak, Z.I.; Sairi, N.A.; Alias, Y.; Aroua, M.K.T.; Yusoff, R. Production of Glycerol Carbonate from Glycerol with Aid of Ionic Liquid as Catalyst. Chem. Eng. J. 2016, 297, 128–138. [Google Scholar]
  24. Muzyka, C.; Silva-Brenes, D.V.; Grignard, B.; Detrembleur, C.; Monbaliu, J.-C.M. The Catalytic Coupling of CO2 and Glycidol toward Glycerol Carbonate. ACS Catal. 2024, 14, 12454–12493. [Google Scholar]
  25. Yang, Q.; Peng, H.; Zhang, Q.; Qian, X.; Chen, X.; Tang, X.; Dai, S.; Zhao, J.; Jiang, K.; Yang, Q.; et al. Atomically Dispersed High-Density Al–N4 Sites in Porous Carbon for Efficient Photodriven CO2 Cycloaddition. Adv. Mater. 2021, 33, 2103186. [Google Scholar]
  26. Muzyka, C.; Renson, S.; Grignard, B.; Detrembleur, C.; Monbaliu, J.-C.M. Intensified Continuous Flow Process for the Scalable Production of Bio-Based Glycerol Carbonate. Angew. Chem. Int. Ed. 2024, 63, e202319060. [Google Scholar]
  27. Cespi, D.; Cucciniello, R.; Ricciardi, M.; Capacchione, C.; Vassura, I.; Passarini, F.; Proto, A. A Simplified Early Stage Assessment of Process Intensification: Glycidol as a Value-Added Product from Epichlorohydrin Industry Wastes. Green Chem. 2016, 18, 4559–4570. [Google Scholar]
  28. Milewski, A.; Czechowicz, D.; Jakóbik-Kolon, A.; Dydo, P. Preparation of Glycidol via Dehydrohalogenation of 3-Chloro-1,2-Popanediol Using Bipolar Membrane Electrodialysis. ACS Sustain. Chem. Eng. 2019, 7, 18640–18646. [Google Scholar]
  29. Gérardy, R.; Debecker, D.P.; Estager, J.; Luis, P.; Monbaliu, J.-C.M. Continuous Flow Upgrading of Selected C2–C6 Platform Chemicals Derived from Biomass. Chem. Rev. 2020, 120, 7219–7347. [Google Scholar]
  30. Tshibalonza, N.N.; Monbaliu, J.-C.M. Revisiting the Deoxydehydration of Glycerol towards Allyl Alcohol under Continuous-Flow Conditions. Green Chem. 2017, 19, 3006–3013. [Google Scholar]
  31. Kostyniuk, A.; Bajec, D.; Djinović, P.; Likozar, B. One-Step Synthesis of Glycidol from Glycerol in a Gas-Phase Packed-Bed Continuous Flow Reactor over HZSM-5 Zeolite Catalysts Modified by CsNO3. Chem. Eng. J. 2020, 394, 124945. [Google Scholar]
  32. Morodo, R.; Gérardy, R.; Petit, G.; Monbaliu, J.-C.M. Continuous Flow Upgrading of Glycerol toward Oxiranes and Active Pharmaceutical Ingredients Thereof. Green Chem. 2019, 21, 4422–4433. [Google Scholar]
  33. Fujita, S.; Yamanishi, Y.; Arai, M. Synthesis of Glycerol Carbonate from Glycerol and Urea Using Zinc-Containing Solid Catalysts: A Homogeneous Reaction. J. Catal. 2013, 297, 137–141. [Google Scholar]
  34. Huang, R.; Rintjema, J.; González-Fabra, J.; Martín, E.; Escudero-Adán, E.C.; Bo, C.; Urakawa, A.; Kleij, A.W. Deciphering Key Intermediates in the Transformation of Carbon Dioxide into Heterocyclic Products. Nat. Catal. 2019, 2, 62–70. [Google Scholar]
  35. Park, J.-H.; Choi, J.S.; Woo, S.K.; Lee, S.D.; Cheong, M.; Kim, H.S.; Lee, H. Isolation and Characterization of Intermediate Catalytic Species in the Zn-Catalyzed Glycerolysis of Urea. Appl. Catal. Gen. 2012, 433–434, 35–40. [Google Scholar]
  36. Turney, T.W.; Patti, A.; Gates, W.; Shaheen, U.; Kulasegaram, S. Formation of Glycerol Carbonate from Glycerol and Urea Catalysed by Metal Monoglycerolates. Green Chem. 2013, 15, 1925–1931. [Google Scholar]
  37. Rubio-Marcos, F.; Calvino-Casilda, V.; Bañares, M.A.; Fernandez, J.F. Novel Hierarchical Co3O4/ZnO Mixtures by Dry Nanodispersion and Their Catalytic Application in the Carbonylation of Glycerol. J. Catal. 2010, 275, 288–293. [Google Scholar]
  38. Ryu, Y.B.; Kim, J.S.; Kim, K.H.; Kim, Y.; Lee, M.S. Synthesis of Zn/Al Mixed-Oxide Catalyst for Carbonylation of Glycerol with Urea. Res. Chem. Intermed. 2016, 42, 83–93. [Google Scholar]
  39. 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]
  40. Procopio, D.; Di Gioia, M.L. An Overview of the Latest Advances in the Catalytic Synthesis of Glycerol Carbonate. Catalysts 2022, 12, 50. [Google Scholar] [CrossRef]
  41. 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. Biorefinery 2022, 12, 4665–4682. [Google Scholar]
  42. Hu, C.; Chang, A.-L.; Catherine, H.N.; Lee, Y.-L.; Andrew Lin, K.-Y.; Chou, Y.-J.; Yu, W.-Y. Enhanced Lewis Basicity of ZIF-8 from Metal Incorporation (Mg, Cu, or Ce) for Glycerol Carboxylation Using CO2 as a Feedstock. Mater. Today Sustain. 2023, 24, 100511. [Google Scholar]
  43. Liu, J.; Li, Y.; Liu, H.; He, D. Transformation of CO2 and Glycerol to Glycerol Carbonate over CeO2ZrO2 Solid Solution—Effect of Zr Doping. Biomass Bioenergy 2018, 118, 74–83. [Google Scholar]
  44. Li, J.; Wang, T. Chemical Equilibrium of Glycerol Carbonate Synthesis from Glycerol. J. Chem. Thermodyn. 2011, 43, 731–736. [Google Scholar]
  45. Ke, Y.-H.; Xu, H.; Wang, X.; Liu, H.; Yuan, H. Production of Glycerol Carbonate by Coupling Glycerol and CO2 over Various Metal Oxide Catalyst. J. CO2 Util. 2024, 83, 102813. [Google Scholar]
  46. Dabral, S.; Schaub, T. The Use of Carbon Dioxide (CO2) as a Building Block in Organic Synthesis from an Industrial Perspective. Adv. Synth. Catal. 2019, 361, 223–246. [Google Scholar]
  47. Grignard, B.; Gennen, S.; Jérôme, C.; Kleij, A.W.; Detrembleur, C. Advances in the Use of CO2 as a Renewable Feedstock for the Synthesis of Polymers. Chem. Soc. Rev. 2019, 48, 4466–4514. [Google Scholar]
  48. Seo, H.; Nguyen, L.V.; Jamison, T.F. Using Carbon Dioxide as a Building Block in Continuous Flow Synthesis. Adv. Synth. Catal. 2019, 361, 247–264. [Google Scholar]
  49. Limburg, B.; Cristòfol, À.; Della Monica, F.; Kleij, A.W. Unlocking the Potential of Substrate-Directed CO2 Activation and Conversion: Pushing the Boundaries of Catalytic Cyclic Carbonate and Carbamate Formation. ChemSusChem 2020, 13, 6056–6065. [Google Scholar]
  50. Hasan, M.M.F.; Rossi, L.M.; Debecker, D.P.; Leonard, K.C.; Li, Z.; Makhubela, B.C.E.; Zhao, C.; Kleij, A. Can CO2 and Renewable Carbon Be Primary Resources for Sustainable Fuels and Chemicals? ACS Sustain. Chem. Eng. 2021, 9, 12427–12430. [Google Scholar]
  51. Liu, Y.; Li, S.; Yu, X.; Chen, Y.; Tang, X.; Hu, T.; Shi, L.; Pudukudy, M.; Shan, S.; Zhi, Y. Cellulose Nanofibers (CNF) Supported (Salen)Cr(III) Composite as an Efficient Heterogeneous Catalyst for CO2 Cycloaddition. Mol. Catal. 2023, 547, 113344. [Google Scholar]
  52. Tong, Y.; Cheng, R.; Dong, H.; Liu, B. Efficient Cycloaddition of CO2 and Epoxides to Cyclic Carbonates Using Salen-Based Covalent Organic Framework as a Heterogeneous Catalyst. J. Porous Mater. 2022, 29, 1253–1263. [Google Scholar]
  53. Li, Y.; Weng, S.; Wang, S.; Zhang, G.; Liu, F.; Liu, M. Engineering the Activity and Stability of ZIF-8(Zn/Co)@g-C3N4 Nanocomposites and Their Synergistic Action in Converting Atmospheric CO2 into Cyclic Carbonates. J. Colloid Interface Sci. 2024, 656, 24–34. [Google Scholar] [PubMed]
  54. Sperandio, C.; Rodriguez, J.; Quintard, A. Organocatalytic Carbon Dioxide Fixation to Epoxides by Perfluorinated 1,3,5-Triols Catalysts. Org. Biomol. Chem. 2020, 18, 2637–2640. [Google Scholar] [PubMed]
  55. Álvarez-Miguel, L.; Burgoa, J.D.; Mosquera, M.E.G.; Hamilton, A.; Whiteoak, C.J. Catalytic Formation of Cyclic Carbonates Using Gallium Aminotrisphenolate Compounds and Comparison to Their Aluminium Congeners: A Combined Experimental and Computational Study. ChemCatChem 2021, 13, 4099–4110. [Google Scholar]
  56. He, Y.; Li, X.; Cai, W.; Lu, H.; Ding, J.; Li, H.; Wan, H.; Guan, G. One-Pot Multiple-Step Integration Strategy for Efficient Fixation of CO2 into Chain Carbonates by Azolide Anions Poly(Ionic Liquid)s. ACS Sustain. Chem. Eng. 2021, 9, 7074–7085. [Google Scholar]
  57. Sinchow, M.; Semakul, N.; Konno, T.; Rujiwatra, A. Lanthanide Coordination Polymers through Design for Exceptional Catalytic Performances in CO2 Cycloaddition Reactions. ACS Sustain. Chem. Eng. 2021, 9, 8581–8591. [Google Scholar]
  58. Rasal, K.B.; Yadav, G.D.; Koskinen, R.; Keiski, R. Solventless Synthesis of Cyclic Carbonates by Direct Utilization of CO2 Using Nanocrystalline Lithium Promoted Magnesia. Mol. Catal. 2018, 451, 200–208. [Google Scholar]
  59. Zhang, S.; Han, F.; Yan, S.; He, M.; Miao, C.; He, L.-N. Efficient Catalysts In Situ Generated from Zinc, Amide and Benzyl Bromide for Epoxide/CO2 Coupling Reaction at Atmospheric Pressure. Eur. J. Org. Chem. 2019, 2019, 1311–1316. [Google Scholar]
  60. Zhang, Z.; Li, J.; Yu, G.; Zeng, C.; Wang, M.; Huang, S.; Wang, L.; Zhang, J. Efficient Synthesis of Cyclic Carbonates under Atmospheric CO2 by DMAP-Based Ionic Liquids: The Difference of Inert Hydrogen Atom and Active Hydrogen Atom in Cation. Green Chem. Eng. 2023, 4, 285–293. [Google Scholar]
  61. Wang, T.; Zheng, D.; An, B.; Liu, Y.; Ren, T.; Ågren, H.; Wang, L.; Zhang, J.; Ahlquist, M.S.G. Dual-Ionic Imidazolium Salts to Promote Synthesis of Cyclic Carbonates at Atmospheric Pressure. Green Energy Environ. 2022, 7, 1327–1339. [Google Scholar]
  62. Chen, C.; Sun, T.; Chen, Y.; Zhang, Y.; Feng, N.; Wan, H.; Guan, G.; Ma, J. Anchoring High-Density Cooperative Catalytic Sites within Triethylenediamine-Based Ionic-Liquid Polymers via Microenvironment Modulation for Efficient CO2 Fixation. Sep. Purif. Technol. 2024, 330, 125348. [Google Scholar]
  63. Rintjema, J.; Epping, R.; Fiorani, G.; Martín, E.; Escudero-Adán, E.C.; Kleij, A.W. Corrigendum: Substrate-Controlled Product Divergence: Conversion of CO2 into Heterocyclic Products. Angew. Chem. Int. Ed. 2016, 55, 12136. [Google Scholar]
  64. Ho, J.; Nord, M.T.; Stafford, J.P.; Stylianou, K.C. Ruthenium-Based Metal–Organic Framework Catalyst for CO2 Fixation onto Epoxides. Catal. Sci. Technol. 2022, 12, 6998–7002. [Google Scholar]
  65. Yang, Q.; Wang, Y.; Tang, X.; Zhang, Q.; Dai, S.; Peng, H.; Lin, Y.; Tian, Z.; Lu, Z.; Chen, L. Ligand Defect Density Regulation in Metal–Organic Frameworks by Functional Group Engineering on Linkers. Nano Lett. 2022, 22, 838–845. [Google Scholar] [PubMed]
  66. Schwesinger, R.; Willaredt, J.; Schlemper, H.; Keller, M.; Schmitt, D.; Fritz, H. Novel, Very Strong, Uncharged Auxiliary Bases; Design and Synthesis of Monomeric and Polymer-Bound Triaminoiminophosphorane Bases of Broadly Varied Steric Demand. Chem. Ber. 1994, 127, 2435–2454. [Google Scholar]
  67. Ishikawwa, T. Superbases for Organic Synthesis: Guanidines, Amidines and Phosphazenes and Related Organocatalysts; Wiley: Chichester, UK, 2009. [Google Scholar]
  68. Jaiswal, S.; Sharma, Y.C. Ni Modified Distillation Waste Derived Heterogeneous Catalyst Utilized for the Production of Glycerol Carbonate from a Biodiesel By-Product Glycerol: Optimization and Green Metric Studies. Waste Manag. 2023, 156, 148–158. [Google Scholar]
  69. Tran, N.K.; Senthil Raja, D.; Lee, Y.-T.; Khoa Le, T.; Tsai, D.-H. Metal-Organic Framework-Derived Base Catalyst for Conversion of Dimethyl Carbonate to Glycerol Carbonate. Appl. Catal. Gen. 2024, 685, 119878. [Google Scholar]
  70. Banu, I.; Bumbac, G.; Bombos, D.; Velea, S.; Gălan, A.-M.; Bozga, G. Glycerol Acetylation with Acetic Acid over Purolite CT-275. Product Yields and Process Kinetics. Renew. Energy 2020, 148, 548–557. [Google Scholar]
  71. Chang, C.-W.; Gong, Z.-J.; Huang, N.-C.; Wang, C.-Y.; Yu, W.-Y. MgO Nanoparticles Confined in ZIF-8 as Acid-Base Bifunctional Catalysts for Enhanced Glycerol Carbonate Production from Transesterification of Glycerol and Dimethyl Carbonate. Catal. Today 2020, 351, 21–29. [Google Scholar]
  72. Gade, S.M.; Saptal, V.B.; Bhanage, B.M. Perception of Glycerol Carbonate as Green Chemical: Synthesis and Applications. Catal. Commun. 2022, 172, 106542. [Google Scholar]
  73. Sahani, S.; Jaiswal, S.; Mishra, S.; Sharma, Y.C.; Han, S.S. Recent Advances in Bio-Glycerol Valorization to Glycerol Carbonate by Heterogenous Base-Catalyzed Transesterification. Mol. Catal. 2023, 550, 113508. [Google Scholar]
  74. Karthikeyan, M.; Sundararaman, S.; Deivasigamani, P.; Rajasimman, M. Synthesis and Characterization of Barium Doped CaO Heterogeneous Nanocatalyst for the Production of Biodiesel from Catharanthus Roseus Seeds: Kinetics, Optimization and Performance Evaluation. Environ. Res. 2023, 222, 115336. [Google Scholar]
  75. Reisi, B.; Najafi Chermahini, A. Modification G-C3N4 by MgO and Its Application for Glycerol Carbonate Synthesis from Glycerol and Dimethyl Carbonate. Environ. Prog. Sustain. Energy 2023, 42, e14007. [Google Scholar]
  76. Argüello, D.S.; Barroso-Martín, I.; Bálsamo, N.F.; Rodríguez-Castellón, E.; Eimer, G.A.; Crivello, M.E. Cu-Ni Synergy in Physicochemical Properties of the Mg-Al Oxides Matrix to Selective Glycerol Carbonate Production. Chem. Eng. Res. Des. 2024, 212, 555–568. [Google Scholar]
  77. Deshmukh, G.P.; Yadav, G.D. Tuneable Transesterification of Glycerol with Dimethyl Carbonate for Synthesis of Glycerol Carbonate and Glycidol on MnO2 Nanorods and Efficacy of Different Polymorphs. Mol. Catal. 2021, 515, 111934. [Google Scholar]
  78. Ramesh, S.; Devred, F.; van den Biggelaar, L.; Debecker, D.P. Hydrotalcites Promoted by NaAlO2 as Strongly Basic Catalysts with Record Activity in Glycerol Carbonate Synthesis. ChemCatChem 2018, 10, 1398–1405. [Google Scholar]
  79. She, Q.M.; Huang, W.J.; Talebian-Kiakalaieh, A.; Yang, H.; Zhou, C.H. Layered Double Hydroxide Uniformly Coated on Mesoporous Silica with Tunable Morphorlogies for Catalytic Transesterification of Glycerol with Dimethyl Carbonate. Appl. Clay Sci. 2021, 210, 106135. [Google Scholar]
  80. Zhu, J.; Chen, D.; Wang, Z.; Wu, Q.; Yin, Z.; Wei, Z. Synthesis of Glycerol Carbonate from Glycerol and Dimethyl Carbonate over CaO-SBA-15 Catalyst. Chem. Eng. Sci. 2022, 258, 117760. [Google Scholar]
  81. Jitjamnong, J.; Khongprom, P.; Ratanawilai, T.; Ratanawilai, S. Glycerol Carbonate Synthesis via Transesterification of Enriched Glycerol and Dimethyl Carbonate Using a Li-Incorporated MCM-41 Framework. RSC Adv. 2024, 14, 5941–5958. [Google Scholar]
  82. Yu, J.; Wang, K.; Shao, S.; Li, W.; Du, S.; Chen, X.; Chao, C.; Fan, X. Effect of Ionic Radius and Valence State of Alkali and Alkaline Earth Metals on Promoting the Catalytic Performance of La2O3 Catalysts for Glycerol Carbonate Production. Chem. Eng. J. 2023, 458, 141486. [Google Scholar]
  83. Tang, H.; Luo, C.; Lu, H.; Wu, K.; Liu, Y.; Zhu, Y.; Wang, B.; Liang, B. Readily Available, Biocompatible Sodium Citrate Catalyst for Efficient Glycerol Carbonate Production through Transesterification of Glycerol and Ethylene Carbonate. Chem. Eng. J. 2024, 481, 148552. [Google Scholar]
  84. Leus, K.; Liu, Y.-Y.; Meledina, M.; Turner, S.; Van Tendeloo, G.; Van Der Voort, P. A MoVI Grafted Metal Organic Framework: Synthesis, Characterization and Catalytic Investigations. J. Catal. 2014, 316, 201–209. [Google Scholar]
  85. Nguyen-Phu, H.; Park, C.; Shin, E.W. Dual Catalysis over ZnAl Mixed Oxides in the Glycerolysis of Urea: Homogeneous and Heterogeneous Reaction Routes. Appl. Catal. Gen. 2018, 552, 1–10. [Google Scholar]
  86. Nguyen-Phu, H.; Shin, E.W. Disordered Structure of ZnAl2O4 Phase and the Formation of a Zn NCO Complex in ZnAl Mixed Oxide Catalysts for Glycerol Carbonylation with Urea. J. Catal. 2019, 373, 147–160. [Google Scholar]
  87. Wang, D.; Zhang, X.; Cong, X.; Liu, S.; Zhou, D. Influence of Zr on the Performance of Mg-Al Catalysts via Hydrotalcite-like Precursors for the Synthesis of Glycerol Carbonate from Urea and Glycerol. Appl. Catal. Gen. 2018, 555, 36–46. [Google Scholar]
  88. Jagadeeswaraiah, K.; Kumar, C.R.; Prasad, P.S.S.; Loridant, S.; Lingaiah, N. Synthesis of Glycerol Carbonate from Glycerol and Urea over Tin-Tungsten Mixed Oxide Catalysts. Appl. Catal. Gen. 2014, 469, 165–172. [Google Scholar]
  89. Jagadeeswaraiah, K.; Ramesh Kumar, C.; Rajashekar, A.; Srivani, A.; Lingaiah, N. The Role of Tungsten Oxide Species Supported on Titania Catalysts for the Synthesis of Glycerol Carbonate from Glycerol and Urea. Catal. Lett. 2016, 146, 692–700. [Google Scholar]
  90. Kim, D.-W.; Kim, M.-J.; Roshith, K.; Kim, M.-I.; Kwak, J.-Y.; Park, D.-W. Comparative Catalytic Activity of Supported ZnBr2-Containing Ionic Liquid Catalysts for Preparation of Glycerol Carbonate by Glycerolysis of Urea. Korean J. Chem. Eng. 2014, 31, 972–980. [Google Scholar]
  91. Lee, S.-D.; Park, M.-S.; Kim, D.-W.; Kim, I.; Park, D.-W. Catalytic Performance of Ion-Exchanged Montmorillonite with Quaternary Ammonium Salts for the Glycerolysis of Urea. Catal. Today 2014, 232, 127–133. [Google Scholar]
  92. Kim, D.-W.; Park, K.-A.; Kim, M.-J.; Kang, D.-H.; Yang, J.-G.; Park, D.-W. Synthesis of Glycerol Carbonate from Urea and Glycerol Using Polymer-Supported Metal Containing Ionic Liquid Catalysts. Appl. Catal. Gen. 2014, 473, 31–40. [Google Scholar]
  93. Climent, M.J.; Corma, A.; De Frutos, P.; Iborra, S.; Noy, M.; Velty, A.; Concepción, P. Chemicals from Biomass: Synthesis of Glycerol Carbonate by Transesterification and Carbonylation with Urea with Hydrotalcite Catalysts. The Role of Acid–Base Pairs. J. Catal. 2010, 269, 140–149. [Google Scholar]
  94. Nguyen-Phu, H.; Park, C.; Eun, W.S. Activated Red Mud-Supported Zn/Al Oxide Catalysts for Catalytic Conversion of Glycerol to Glycerol Carbonate: FTIR Analysis. Catal. Commun. 2016, 85, 52–56. [Google Scholar]
  95. Zhao, W.; Peng, W.; Wang, D.; Zhao, N.; Li, J.; Xiao, F.; Wei, W.; Sun, Y. Zinc Oxide as the Precursor of Homogenous Catalyst for Synthesis of Dialkyl Carbonate from Urea and Alcohols. Catal. Commun. 2009, 10, 655–658. [Google Scholar]
  96. Pham-Ngoc, N.; Nguyen-Phu, H.; Lee, Y.; Shin, E.W. The Effect of Different Polyol Precursors on Disordered Spinel ZnAl2O4 Structure Prepared by the Polymeric Citrate Complex Method and the Corresponding Catalytic Behavior in the Glycerolysis of Urea. Small Struct. 2024, 5, 2300337. [Google Scholar]
  97. Shi, J.; Cui, Y.; Sun, H.; Wang, H.; Liu, C.; Xue, X.; Li, C.; Geng, L.; Liu, J.; Jia, M. N-Doped Porous Carbon-Anchored Zinc Single-Atom as an Efficient and Robust Heterogeneous Catalyst for Glycerol Carbonylation with Urea. Chem. Eng. J. 2023, 466, 143317. [Google Scholar]
  98. Butburee, T.; Prasert, A.; Rungtaweevoranit, B.; Khemthong, P.; Mano, P.; Youngjan, S.; Phanthasri, J.; Namuangruk, S.; Faungnawakij, K.; Zhang, L.; et al. Engineering Lewis-Acid Defects on ZnO Quantum Dots by Trace Transition-Metal Single Atoms for High Glycerol-to-Glycerol Carbonate Conversion. Small 2024, 20, 2403661. [Google Scholar]
  99. Wei, X.X.; Song, C.; Geng, K.W.; Zeng, F.; He, B.; Pan, F. Local Fe Structure and Ferromagnetism in Fe-Doped ZnO Films. J. Phys. Condens. Matter 2006, 18, 7471. [Google Scholar]
Figure 1. Synthetic strategy and defect indicator of M-UiO-66-X (M = Ce, Zr, or Hf, X = OMe, Me, H, or F). (a) Schematic illustration showing the fabrication process of M-UiO-66-X with different ligand defect densities. (b) Experimental measurement of ligand defect densities in M-UiO-66-X via the dissolution/NMR approach. (c) Theoretical bonding strength between BDC-X and Ce-oxo nodes. Catalytic performance of Ce-UiO-66-X (X = F, H, Me, OMe) for CO2 cycloaddition reactions. Reproduced with the permission of ref. [65]. Copyright 2022, American Chemical Society.
Figure 1. Synthetic strategy and defect indicator of M-UiO-66-X (M = Ce, Zr, or Hf, X = OMe, Me, H, or F). (a) Schematic illustration showing the fabrication process of M-UiO-66-X with different ligand defect densities. (b) Experimental measurement of ligand defect densities in M-UiO-66-X via the dissolution/NMR approach. (c) Theoretical bonding strength between BDC-X and Ce-oxo nodes. Catalytic performance of Ce-UiO-66-X (X = F, H, Me, OMe) for CO2 cycloaddition reactions. Reproduced with the permission of ref. [65]. Copyright 2022, American Chemical Society.
Catalysts 15 00668 g001
Figure 2. Catalytic performance of Ce-UiO-66-X (X = F, H, Me, OMe) for CO2 cycloaddition reactions. (a) Catalytic yields for the cycloaddition reaction of glycidol with CO2 in the presence of Ce-UiO-66-X catalysts. Reaction conditions: 1 mmol epoxides, 1 bar CO2, 2 mL of N,N dimethylacetamide (DMA), 50 mg of Ce-UiO-66-X, 32 mg of tetrabutylammonium bromide (TBAB), 60 °C. (b) Arrhenius plot for the determination of Ea for four Ce-UiO-66-X catalysts. (c) Correlation between the Ea and ligand defect densities of Ce-UiO-66-X samples. (d) Computed reaction energy profiles of the CO2 cycloaddition with glycidol on Ce-UiO-66-H with different ligand defect densities (The blue represents Cerium (Ce), the red represents Oxygen (O), the gray represents Carbon (C), the white represents Hydrogen (H), and the purple represents X in UiO-66-X (X = H)). Reproduced with the permission of ref. [65]. Copyright 2022, American Chemical Society.
Figure 2. Catalytic performance of Ce-UiO-66-X (X = F, H, Me, OMe) for CO2 cycloaddition reactions. (a) Catalytic yields for the cycloaddition reaction of glycidol with CO2 in the presence of Ce-UiO-66-X catalysts. Reaction conditions: 1 mmol epoxides, 1 bar CO2, 2 mL of N,N dimethylacetamide (DMA), 50 mg of Ce-UiO-66-X, 32 mg of tetrabutylammonium bromide (TBAB), 60 °C. (b) Arrhenius plot for the determination of Ea for four Ce-UiO-66-X catalysts. (c) Correlation between the Ea and ligand defect densities of Ce-UiO-66-X samples. (d) Computed reaction energy profiles of the CO2 cycloaddition with glycidol on Ce-UiO-66-H with different ligand defect densities (The blue represents Cerium (Ce), the red represents Oxygen (O), the gray represents Carbon (C), the white represents Hydrogen (H), and the purple represents X in UiO-66-X (X = H)). Reproduced with the permission of ref. [65]. Copyright 2022, American Chemical Society.
Catalysts 15 00668 g002
Figure 3. (a) Microfluidic prototype for the screening of the catalyst. Conditions: GD (1.8 M in MEK with 5 mol % cat.) at 0.25 mL min−1; CO2 (gas, 1 equiv.) at 10 mLN min−1, 140 °C, 10 bar, and 2 min estimated residence time. MFC = Mass Flow Controller. (b) Screening of various homogeneous nitrogen-containing organic bases for the coupling of bio-based GD with CO2 (see (a)). [a] Indicative values for experimental pKaH in MeCN reported in the literature [66,67]. [b] Yield of GC determined by 1H NMR. [c] Yield of Gly determined by 1H NMR. Reproduced with the permission of ref. [26]. Copyright 2024, Wiley-VCH GmbH.
Figure 3. (a) Microfluidic prototype for the screening of the catalyst. Conditions: GD (1.8 M in MEK with 5 mol % cat.) at 0.25 mL min−1; CO2 (gas, 1 equiv.) at 10 mLN min−1, 140 °C, 10 bar, and 2 min estimated residence time. MFC = Mass Flow Controller. (b) Screening of various homogeneous nitrogen-containing organic bases for the coupling of bio-based GD with CO2 (see (a)). [a] Indicative values for experimental pKaH in MeCN reported in the literature [66,67]. [b] Yield of GC determined by 1H NMR. [c] Yield of Gly determined by 1H NMR. Reproduced with the permission of ref. [26]. Copyright 2024, Wiley-VCH GmbH.
Catalysts 15 00668 g003
Figure 4. Intensified mesofluidic pilot-scale process (Corning® AFR™ G1, 7 glass fluidic modules in series) for the coupling of GD and CO2. Conditions: GD (1.8 M in MEK with 1 mol % BB) at 16.18 mL min−1; CO2 (gas, 1 equiv.) at 648 mLN min−1, 140 °C, 10 bar, and 28 s estimated residence time. FM = fluidic module, MFC = Mass Flow Controller. Up to 4 % Gly was detected in the crude reactor effluent. Reproduced with the permission of ref. [26]. Copyright 2024, Wiley-VCH GmbH.
Figure 4. Intensified mesofluidic pilot-scale process (Corning® AFR™ G1, 7 glass fluidic modules in series) for the coupling of GD and CO2. Conditions: GD (1.8 M in MEK with 1 mol % BB) at 16.18 mL min−1; CO2 (gas, 1 equiv.) at 648 mLN min−1, 140 °C, 10 bar, and 28 s estimated residence time. FM = fluidic module, MFC = Mass Flow Controller. Up to 4 % Gly was detected in the crude reactor effluent. Reproduced with the permission of ref. [26]. Copyright 2024, Wiley-VCH GmbH.
Catalysts 15 00668 g004
Figure 5. (a) Mechanism for the organocatalyzed coupling of CO2 and 2 toward 3, computed at the B3LYP-GD3BJ/6–31+G*//M08HX/6–311++G** level of theory (SMD = MEK, 413 K). Activation barriers (ΔG) are given in kcalmol 1 and mentioned near the acronyms of the various bases (The blue oxygen represents the oxygen in carbon dioxide, the green oxygen represents the hydroxyl oxygen in GD, and the red oxygen represents the oxygen atom in the epoxy group in GD). (b) Free Gibbs energy of reaction (ΔG°, in red) and activation barriers (ΔG, in blue) as a function of the pKaH of the catalysts for step 1. (c) Calculated Free Gibbs energy of reaction (ΔG°) for step 2. Reproduced with the permission of ref. [26]. Copyright 2024, Wiley-VCH GmbH.
Figure 5. (a) Mechanism for the organocatalyzed coupling of CO2 and 2 toward 3, computed at the B3LYP-GD3BJ/6–31+G*//M08HX/6–311++G** level of theory (SMD = MEK, 413 K). Activation barriers (ΔG) are given in kcalmol 1 and mentioned near the acronyms of the various bases (The blue oxygen represents the oxygen in carbon dioxide, the green oxygen represents the hydroxyl oxygen in GD, and the red oxygen represents the oxygen atom in the epoxy group in GD). (b) Free Gibbs energy of reaction (ΔG°, in red) and activation barriers (ΔG, in blue) as a function of the pKaH of the catalysts for step 1. (c) Calculated Free Gibbs energy of reaction (ΔG°) for step 2. Reproduced with the permission of ref. [26]. Copyright 2024, Wiley-VCH GmbH.
Catalysts 15 00668 g005
Scheme 1. Synthesis process of modifying the La2O3 catalyst by Na. Reproduced with the permission of ref. [82]. Copyright 2023, Elsevier.
Scheme 1. Synthesis process of modifying the La2O3 catalyst by Na. Reproduced with the permission of ref. [82]. Copyright 2023, Elsevier.
Catalysts 15 00668 sch001
Figure 6. XPS spectra of pristine La2O3 samples and La2O3 samples modified by alkali and alkaline earth metals: (a) La 3d, where peak I is the main peak and peak II and III are the satellite peaks; and (b) O 1s, where peak I, II, and III are represented for lattice O2− (OL), chemisorbed surface O (OS), and weakly adsorbed OH and CO32− (OA) species, respectively. Reproduced with the permission of ref. [82]. Copyright 2023, Elsevier.
Figure 6. XPS spectra of pristine La2O3 samples and La2O3 samples modified by alkali and alkaline earth metals: (a) La 3d, where peak I is the main peak and peak II and III are the satellite peaks; and (b) O 1s, where peak I, II, and III are represented for lattice O2− (OL), chemisorbed surface O (OS), and weakly adsorbed OH and CO32− (OA) species, respectively. Reproduced with the permission of ref. [82]. Copyright 2023, Elsevier.
Catalysts 15 00668 g006
Figure 7. (a) Correlation between basic site density and glycerol conversion, where the glycerol conversion was carried out at the same condition for each catalyst (reaction condition: DMC/GL = 3:1, 0.10 g catalyst, 70 °C, and 2 h). (b) Correlation between GL conversion and the ionic radius ratio of dopant and La, where the glycerol conversion was carried out at the same condition for each catalyst (reaction condition: DMC/GL = 3:1, 0.10 g catalyst, 70 °C, and 2 h). Reproduced with the permission of ref. [82]. Copyright 2023, Elsevier.
Figure 7. (a) Correlation between basic site density and glycerol conversion, where the glycerol conversion was carried out at the same condition for each catalyst (reaction condition: DMC/GL = 3:1, 0.10 g catalyst, 70 °C, and 2 h). (b) Correlation between GL conversion and the ionic radius ratio of dopant and La, where the glycerol conversion was carried out at the same condition for each catalyst (reaction condition: DMC/GL = 3:1, 0.10 g catalyst, 70 °C, and 2 h). Reproduced with the permission of ref. [82]. Copyright 2023, Elsevier.
Catalysts 15 00668 g007
Figure 8. Plausible mechanism of Gly and DMC transesterification on Na/La2O3 catalyst. Reproduced with permission of ref. [82]. Copyright 2023, Elsevier.
Figure 8. Plausible mechanism of Gly and DMC transesterification on Na/La2O3 catalyst. Reproduced with permission of ref. [82]. Copyright 2023, Elsevier.
Catalysts 15 00668 g008
Scheme 2. Catalytic transesterification of Gly and EC. Reproduced with the permission of ref. [83]. Copyright 2024, Elsevier.
Scheme 2. Catalytic transesterification of Gly and EC. Reproduced with the permission of ref. [83]. Copyright 2024, Elsevier.
Catalysts 15 00668 sch002
Figure 9. Effect of reaction parameters on the Na3-citrate-catalyzed synthesis of GC from the transesterification of Gly and EC. (a) Reaction time, Gly = 20 mmol, EC = 80 mmol, catalyst loading = 0.01 wt% (based on Gly), reaction temperature = 120 °C. (b) EC to Gly molar ratio, Gly = 20 mmol, catalyst loading = 0.01 wt% (based on Gly), reaction temperature = 120 °C, reaction time = 45 min. Reproduced with the permission of ref. [83]. Copyright 2024, Elsevier.
Figure 9. Effect of reaction parameters on the Na3-citrate-catalyzed synthesis of GC from the transesterification of Gly and EC. (a) Reaction time, Gly = 20 mmol, EC = 80 mmol, catalyst loading = 0.01 wt% (based on Gly), reaction temperature = 120 °C. (b) EC to Gly molar ratio, Gly = 20 mmol, catalyst loading = 0.01 wt% (based on Gly), reaction temperature = 120 °C, reaction time = 45 min. Reproduced with the permission of ref. [83]. Copyright 2024, Elsevier.
Catalysts 15 00668 g009
Figure 10. (a) Optimized structures of Gly and (b) the Gly-(Na3-citrate) complex. (c) Electrostatic potential surface maps of Gly and (d) the Gly-(Na3-citrate) complex. Color code for atoms and ions: gray = carbon, red = oxygen, white = hydrogen, purple = sodium. (e) NCI analysis of Gly-(Na3-citrate) complex. Color code for atoms and ions: cyan = carbon; red = oxygen; white = hydrogen; coffee-colored = sodium. (f) Reduced density gradient as a function of signλ2ρ. Reproduced with the permission of ref. [83]. Copyright 2024, Elsevier.
Figure 10. (a) Optimized structures of Gly and (b) the Gly-(Na3-citrate) complex. (c) Electrostatic potential surface maps of Gly and (d) the Gly-(Na3-citrate) complex. Color code for atoms and ions: gray = carbon, red = oxygen, white = hydrogen, purple = sodium. (e) NCI analysis of Gly-(Na3-citrate) complex. Color code for atoms and ions: cyan = carbon; red = oxygen; white = hydrogen; coffee-colored = sodium. (f) Reduced density gradient as a function of signλ2ρ. Reproduced with the permission of ref. [83]. Copyright 2024, Elsevier.
Catalysts 15 00668 g010
Figure 11. (a) Free energy profiles and optimized geometries for the Na3-citrate-catalyzed transesterification of Gly and EC. All energies are relative to Gly+EC+Na3-citrate. Color code for atoms and ions: gray = carbon, red = oxygen, white = hydrogen, purple = sodium. (b) Description of the reaction mechanism for the Na3-citrate-catalyzed transesterification of Gly and EC. Reproduced with the permission of ref. [83]. Copyright 2024, Elsevier.
Figure 11. (a) Free energy profiles and optimized geometries for the Na3-citrate-catalyzed transesterification of Gly and EC. All energies are relative to Gly+EC+Na3-citrate. Color code for atoms and ions: gray = carbon, red = oxygen, white = hydrogen, purple = sodium. (b) Description of the reaction mechanism for the Na3-citrate-catalyzed transesterification of Gly and EC. Reproduced with the permission of ref. [83]. Copyright 2024, Elsevier.
Catalysts 15 00668 g011
Figure 12. (a) Enlarged XRD patterns of the fresh catalysts: [a] i–ZnAl2O4, [b] eg–ZnAl2O4, [c] pg200–ZnAl2O4, [d] gl–ZnAl2O4, and [e] s–ZnAl2O4. (b) Variation in the order parameters and lattice parameters of the ZnAl2O4. Reproduced with the permission of ref. [96]. Copyright 2023, Wiley-VCH GmbH.
Figure 12. (a) Enlarged XRD patterns of the fresh catalysts: [a] i–ZnAl2O4, [b] eg–ZnAl2O4, [c] pg200–ZnAl2O4, [d] gl–ZnAl2O4, and [e] s–ZnAl2O4. (b) Variation in the order parameters and lattice parameters of the ZnAl2O4. Reproduced with the permission of ref. [96]. Copyright 2023, Wiley-VCH GmbH.
Catalysts 15 00668 g012
Scheme 3. (a) The ordered lattice and (b) disordered lattice of spinel ZnAl2O4. The acidic sites of ZnAl2O4 spinel: (i) the weak Lewis acid site, (ii) the medium-weak Lewis acid site, (iii) the medium-strong Lewis acid site, and (iv) the strong Lewis acid site. Reproduced with the permission of ref. [96]. Copyright 2023, Wiley-VCH GmbH.
Scheme 3. (a) The ordered lattice and (b) disordered lattice of spinel ZnAl2O4. The acidic sites of ZnAl2O4 spinel: (i) the weak Lewis acid site, (ii) the medium-weak Lewis acid site, (iii) the medium-strong Lewis acid site, and (iv) the strong Lewis acid site. Reproduced with the permission of ref. [96]. Copyright 2023, Wiley-VCH GmbH.
Catalysts 15 00668 sch003
Figure 13. The relationships between acidic sites versus glycerol carbonate yield and basic sites versus (2) selectivity. Reproduced with the permission of ref. [96]. Copyright 2023, Wiley-VCH GmbH.
Figure 13. The relationships between acidic sites versus glycerol carbonate yield and basic sites versus (2) selectivity. Reproduced with the permission of ref. [96]. Copyright 2023, Wiley-VCH GmbH.
Catalysts 15 00668 g013
Figure 14. Mechanistic pathway for the alcoholysis of urea with glycerol over X-ZnAl2O4 catalysts. Reproduced with the permission of ref. [96]. Copyright 2023, Wiley-VCH GmbH.
Figure 14. Mechanistic pathway for the alcoholysis of urea with glycerol over X-ZnAl2O4 catalysts. Reproduced with the permission of ref. [96]. Copyright 2023, Wiley-VCH GmbH.
Catalysts 15 00668 g014
Scheme 4. Schematic illustration of the synthesis of NC and corresponding Zn1/NC catalyst. Reproduced with the permission of ref. [97]. Copyright 2023, Elsevier.
Scheme 4. Schematic illustration of the synthesis of NC and corresponding Zn1/NC catalyst. Reproduced with the permission of ref. [97]. Copyright 2023, Elsevier.
Catalysts 15 00668 sch004
Figure 15. (a) Schematic illustration for the construction of Zn1/NC. SEM, TEM, and STEM-EDS elemental mapping; (b) Zn K-edge: XANES spectra, FT-EXAFS spectra, and Zn K-edge EXAFS k2 χ(k) oscillation functions and FT-EXAFS of the Zn1/NC. Reproduced with the permission of ref. [97]. Copyright 2023, Elsevier.
Figure 15. (a) Schematic illustration for the construction of Zn1/NC. SEM, TEM, and STEM-EDS elemental mapping; (b) Zn K-edge: XANES spectra, FT-EXAFS spectra, and Zn K-edge EXAFS k2 χ(k) oscillation functions and FT-EXAFS of the Zn1/NC. Reproduced with the permission of ref. [97]. Copyright 2023, Elsevier.
Catalysts 15 00668 g015
Figure 16. (a) The reaction route of glycerol carbonylation with urea. (b) Catalytic performance comparison of the different catalysts after a 6 h reaction. (c) Reactivity of the Zn1/NC catalyst during five reaction cycles. Reaction conditions: 120 °C, 7 kPa, glycerol/urea ratio = 1.0:1.5. (d) GC selectivity as a function of the glycerol conversion for Zn(NO3)2 and Zn1/NC catalysts. Reproduced with the permission of ref. [97]. Copyright 2023, Elsevier.
Figure 16. (a) The reaction route of glycerol carbonylation with urea. (b) Catalytic performance comparison of the different catalysts after a 6 h reaction. (c) Reactivity of the Zn1/NC catalyst during five reaction cycles. Reaction conditions: 120 °C, 7 kPa, glycerol/urea ratio = 1.0:1.5. (d) GC selectivity as a function of the glycerol conversion for Zn(NO3)2 and Zn1/NC catalysts. Reproduced with the permission of ref. [97]. Copyright 2023, Elsevier.
Catalysts 15 00668 g016
Figure 17. (a) Possible models of the single Zn site (Zn1-N4−xOx, x = 0–4). (b) The adsorption structures and adsorption energies of glycerol and urea on the Zn1-N4−xOx at 393 K. Reproduced with the permission of ref. [97]. Copyright 2023, Elsevier.
Figure 17. (a) Possible models of the single Zn site (Zn1-N4−xOx, x = 0–4). (b) The adsorption structures and adsorption energies of glycerol and urea on the Zn1-N4−xOx at 393 K. Reproduced with the permission of ref. [97]. Copyright 2023, Elsevier.
Catalysts 15 00668 g017
Figure 18. (a) As-prepared ZnO QDs and M1-ZnO QDs. (b) ZnO QDs and M1-ZnO QDs precipitated by adding ethyl acetate. (c) Low-magnification TEM image showing the morphology and particle size distribution of the example M1-ZnO QDs (Fe1-ZnO). (d) Schematic illustration comparing the structures of ZnO QDs and M1-ZnO QDs. (e,f) are atomic-resolution TEM images showing the corresponding structures of ZnO QDs and the example M1-ZnO QDs (Fe1-ZnO QDs). Reproduced with the permission of ref. [98]. Copyright 2024, Wiley-VCH GmbH.
Figure 18. (a) As-prepared ZnO QDs and M1-ZnO QDs. (b) ZnO QDs and M1-ZnO QDs precipitated by adding ethyl acetate. (c) Low-magnification TEM image showing the morphology and particle size distribution of the example M1-ZnO QDs (Fe1-ZnO). (d) Schematic illustration comparing the structures of ZnO QDs and M1-ZnO QDs. (e,f) are atomic-resolution TEM images showing the corresponding structures of ZnO QDs and the example M1-ZnO QDs (Fe1-ZnO QDs). Reproduced with the permission of ref. [98]. Copyright 2024, Wiley-VCH GmbH.
Catalysts 15 00668 g018
Figure 19. (a) XRD patterns of Fe1-ZnO QDs compared to ZnO QDs and the commercial ZnO. (b) XPS spectra of Fe1-ZnO QDs compared to the commercial ZnO at the Zn 2p regions. (c) XPS spectra of Fe1-ZnO QDs compared to the commercial ZnO at the O 1s regions. (d) XANES spectra of Fe1-ZnO QDs compared to the Fe-based standard materials. (e) Fe K-edge FT-EXAFS spectra of Fe1-ZnO QDs compared to the Fe-based standard materials, as well as their theoretical fitting (f). Reproduced with the permission of ref. [98]. Copyright 2024, Wiley-VCH GmbH.
Figure 19. (a) XRD patterns of Fe1-ZnO QDs compared to ZnO QDs and the commercial ZnO. (b) XPS spectra of Fe1-ZnO QDs compared to the commercial ZnO at the Zn 2p regions. (c) XPS spectra of Fe1-ZnO QDs compared to the commercial ZnO at the O 1s regions. (d) XANES spectra of Fe1-ZnO QDs compared to the Fe-based standard materials. (e) Fe K-edge FT-EXAFS spectra of Fe1-ZnO QDs compared to the Fe-based standard materials, as well as their theoretical fitting (f). Reproduced with the permission of ref. [98]. Copyright 2024, Wiley-VCH GmbH.
Catalysts 15 00668 g019
Figure 20. (a) Comparison of yield, conversion, and selectivity of the Fe1-ZnO QD catalyst with ZnO QDs and the commercial ZnO. (b) Comparison of the catalytic performance of various M1-ZnO QDs. (Testing condition: Gly/Urea = 1:1.5 mole; catalyst weight = 50 mg; reaction temperature = 150 °C; reaction time = 3 h; N2 flow.) (c) The proposed reaction mechanism based on the observed vibration mode of the in situ DRIFTS. The reaction energy pathway for Gly-to-GlyC (inset) and the corresponding intermediate structures calculated by DFT. (d) NH3 adsorption energies at the possible active sites on the surface of the commercial ZnO (without Fe, left) and Fe1-ZnO QDs (with Fe dopant, right). Reproduced with the permission of ref. [98]. Copyright 2024, Wiley-VCH GmbH.
Figure 20. (a) Comparison of yield, conversion, and selectivity of the Fe1-ZnO QD catalyst with ZnO QDs and the commercial ZnO. (b) Comparison of the catalytic performance of various M1-ZnO QDs. (Testing condition: Gly/Urea = 1:1.5 mole; catalyst weight = 50 mg; reaction temperature = 150 °C; reaction time = 3 h; N2 flow.) (c) The proposed reaction mechanism based on the observed vibration mode of the in situ DRIFTS. The reaction energy pathway for Gly-to-GlyC (inset) and the corresponding intermediate structures calculated by DFT. (d) NH3 adsorption energies at the possible active sites on the surface of the commercial ZnO (without Fe, left) and Fe1-ZnO QDs (with Fe dopant, right). Reproduced with the permission of ref. [98]. Copyright 2024, Wiley-VCH GmbH.
Catalysts 15 00668 g020
Table 1. Some of the catalysts mentioned in the coupling of CO2 and GD.
Table 1. Some of the catalysts mentioned in the coupling of CO2 and GD.
CatalystReaction ConditionsGD Conversion/GC Yield/GC
Selectivity/%
Ref
ZIF-8(Zn/Co)@g-C3N4GD: (34.5 mmol), CO2 (2 MPa), catalyst (5 wt%), TBAB * (1 mol%), 80 °C, 6 h.94/92/≥98[53]
P-DVB-AEImImGD (15 mmol), CH3OH/GD (10:1, mol/mol), CO2 (10 bar), catalysts (40 mg), 120 °C, 2 h97/65/67[56]
0.75% (w/w) Li-MgOGD (127 mmol), CO2 (3 MPa), catalyst (0.1 g/cm3), 130 °C, 4 h.98.9/96.5/97.6[58]
ZnBr2/[DBDMA]BrGD (5.0 mmol), DMF (500 mg), CO2 (1 atm), catalyst (5 wt%), 80 °C, 12 h.63/60/95[59]
CeUiO-66-FGD (1.0 mmol), DMA * (2 mL), TBAB (32 mg), CO2 (1 bar), catalyst (50 mg), 60 °C, 6 h.-/96/-[65]
2-tert-butyl-1,1,3,3-tetra-methylguanidineGD (1.8 M in MEK *), CO2 (16.18 mL min−1, 10 bar), catalyst (1 mol%), 140 °C.87/78/90[26]
* N,N dimethylacetamide (DMA), 2-butanone (methyl ethyl ketone (TBAB)), 2-butanone (methyl ethyl ketone (MEK)).
Table 2. Some of the catalysts mentioned in the transesterification of glycerol and organic carbonates.
Table 2. Some of the catalysts mentioned in the transesterification of glycerol and organic carbonates.
CatalystReaction ConditionsGly Conversion/GC Yield/GC
Selectivity/%
Ref
Ni/CaO (NDW)DMC: Gly mole ratio of 3:1, catalyst amount of 3.0% (glycerol weight), 90 °C, 1.5 h99.2/-/94.9[68]
6MgO/g-C3N4DMC: Gly mole ratio of 2:1, catalyst amount of 20 mg, 80 °C, 4 h98.3/97.1/95.5[75]
MMO-Cu15Ni15DMC: Gly mole ratio of 2:1, catalyst amount of 7.5% (glycerol weight), 85 °C, 4.5 h88.9/84.6/95.3[76]
10 wt% NaAlO2/hydrotalciteDMC: Gly mole ratio of 2:1, catalyst amount of 3.0% (glycerol weight), 90 °C, 0.5 h92/-/ almost 100[78]
0.3CaO-SBA-15DMC: Gly mole ratio of 4:1, catalyst amount of 7.0% (glycerol weight), 95 °C, 1.5 h-/99/-[80]
25 mol% Na/La2O3DMC: Gly mole ratio of 3:1, catalyst amount of 100 mg, 70 °C, 2 h85/60/-[82]
Na3-citratesEC *: Gly mole ratio of 4:1, catalyst amount of 0.01% (glycerol weight), 110 °C, 0.75 h>99/>99/-[83]
* EC = ethylene carbonate.
Table 3. Some of the catalysts mentioned in the carbonylation of Gly with urea.
Table 3. Some of the catalysts mentioned in the carbonylation of Gly with urea.
CatalystReaction ConditionsGly Conversion/GC Yield/GC
Selectivity/%
Ref
ZnAl mixed oxideGly/Urea (1:1, mol/mol), catalyst (5 wt%), 3 kPa, 140 °C, 5 h83/73/88[85]
Mg-Al-Zr mixed oxideGly/Urea (1:1, mol/mol), catalyst (4 wt%), 10 kPa, 140 °C, 3 h96.7/87.8/91[87]
PS-(Im)2ZnI2Gly/Urea (1:1, mol/mol), catalyst (5 wt%), 14.7 kPa, 140 °C, 6 h71.7/60.3/84[92]
s–ZnAl2O4Gly/Urea (1:1, mol/mol), catalyst (5 wt%), 3 kPa, 140 °C, 5 h78.2/41.2/52.7[96]
Zn1/NCGly/Urea (1:1.5, mol/mol), catalyst (5 wt%), 7 kPa, 120 °C, 5 h94.8/90.1/95[97]
Fe1-ZnO QDGly/Urea (1:1.5, mol/mol), catalyst (50 mg), N2 flow, 150 °C, 3 h90.7/90.6/100[98]
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

Sun, H.; Lei, Z.; Shi, J.; Jia, M. Recent Advances in Catalyst Innovation, Mechanism Exploration, and Process Optimization for Synthesis of Glycerol Carbonate. Catalysts 2025, 15, 668. https://doi.org/10.3390/catal15070668

AMA Style

Sun H, Lei Z, Shi J, Jia M. Recent Advances in Catalyst Innovation, Mechanism Exploration, and Process Optimization for Synthesis of Glycerol Carbonate. Catalysts. 2025; 15(7):668. https://doi.org/10.3390/catal15070668

Chicago/Turabian Style

Sun, Honglei, Zhenyu Lei, Jinghui Shi, and Mingjun Jia. 2025. "Recent Advances in Catalyst Innovation, Mechanism Exploration, and Process Optimization for Synthesis of Glycerol Carbonate" Catalysts 15, no. 7: 668. https://doi.org/10.3390/catal15070668

APA Style

Sun, H., Lei, Z., Shi, J., & Jia, M. (2025). Recent Advances in Catalyst Innovation, Mechanism Exploration, and Process Optimization for Synthesis of Glycerol Carbonate. Catalysts, 15(7), 668. https://doi.org/10.3390/catal15070668

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