Next Article in Journal
Photodegradation of Wastewater Containing Organic Dyes Using Modified G-C3N4-Doped ZrO2 Nanostructures: Towards Safe Water for Human Beings
Previous Article in Journal
Pioneering the Future: A Trailblazing Review of the Fusion of Computational Fluid Dynamics and Machine Learning Revolutionizing Plasma Catalysis and Non-Thermal Plasma Reactor Design
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reaction Kinetics and Mechanism for the Synthesis of Glycerol Carbonate from Glycerol and Urea Using ZnSO4 as a Catalyst

1
Hubei Key Laboratory of Material Chemistry & Service Failure, School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
2
Key Laboratory for Material Chemistry for Energy Conversion and Storage, Ministry of Education, Huazhong University of Science and Technology, Wuhan 430074, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(1), 41; https://doi.org/10.3390/catal14010041
Submission received: 25 November 2023 / Revised: 26 December 2023 / Accepted: 4 January 2024 / Published: 6 January 2024
(This article belongs to the Section Catalytic Reaction Engineering)

Abstract

:
A series of Zn salts were used as catalysts for the reaction of glycerol and urea to produce glycerol carbonate and it was found that ZnSO4 showed the highest catalytic activity. Furthermore, the effects of reaction parameters on the glycerol conversion and glycerol carbonate yield were studied in detail. The results indicated that the glycerol conversion and glycerol carbonate yield were increased with the reaction temperature, reaction time, and catalyst amount while the optimal reaction conditions were 140 °C, 240 min, catalyst amount of 5 wt% (based on the glycerol weight), and urea-to-glycerol molar ratio of 1.1:1. During the reaction, the ZnSO4 catalyst is transformed into Zn(NH3)2SO4 at the initial stage of the reaction and then further transformed into Zn(C3H6O3). Zn(C3H6O3) and (NH4)2SO4 may be the true active species for the activation of urea and glycerol, respectively. The reaction mechanism is proposed in this article. Based on the experimental results, a reaction kinetics model considering the change in volume of the reaction system was also established, and the model parameters were obtained by fitting the experimental data. The statistical results showed that the established kinetics model is accurate.

1. Introduction

Recently, biodiesel, as one of the renewable and clean energies with plentiful raw material sources, high heat of combustion, low emission amount of pollutants, and a high flash point, has begun to attract more and more attention from researchers [1,2,3,4,5]. According to statistics, the global production of biodiesel in 2020 was about 50 million tons [6]. In industry, the main synthesis method for biodiesel is the transesterification of animal and vegetable fats and oils or kitchen waste oils with low-carbon alcohol (mainly methanol) [7,8,9,10]. During the reaction, glycerol (GL) is generated as a byproduct with a ratio of 0.1 kg to 1.0 kg of biodiesel. So, a huge amount of GL will be produced with the production of biodiesel [11]. Converting GL into derivatives with high added value cannot only effectively maintain the balance of the GL market but can also promote the sustainable development of the biodiesel industry [12,13]. GL can be transformed into propylene glycol, acrolein, dichloropropanol, dihydroxyacetone, glyceric acid, and glycerol carbonate (GC) through the reactions of dehydration, chlorination, oxidation, and transesterification, respectively [14,15,16,17,18,19,20]. Among these derivatives of GL, GC is one of the most promising chemicals. GC is a non-toxic, colorless, and easily biodegradable viscous liquid with a high flash point, high boiling point, and high water solubility [21,22,23]. It can be used as a polar solvent, lithium ion battery liquid, component of a separation membrane, intermediate for polyester, fuel additive, and so on [24,25,26,27,28].
GC can be synthesized from GL by using methods such as (i) the reaction of phosgene with GL; (ii) oxidative carbonylation of CO, O2, and GL; (iii) transesterification of dialkyl carbonate and GL; (iv) carbonylation of CO2 and GL; and (v) reaction of urea and GL. However, the reaction of phosgene and GL and oxidative carbonylation of CO, O2, and GL are limited due to the toxicity of reactants and high reaction temperature and pressure [29,30]. The carbonylation of CO2 and GL also suffers with low GL conversion due to a severe thermodynamic limitation [31,32]. In addition, the transesterification of dialkyl carbonate and GL is limited by expensive reactants and a catalyst with low activity [33]. Compared with other methods, the reaction of urea and GL possesses the following advantages: (i) cheap reactant (urea); (ii) mild reaction conditions; and (iii) high atomic economy, because the ammonia produced in this process can be reacted with greenhouse gas CO2 under certain conditions to produce more urea for use as the reactant (Scheme 1) [34].
Currently, the development of the catalyst for the reaction of GL and urea is a focus of research within academia and industry. Two types of catalysts for the reaction of GL and urea have been developed: homogeneous and heterogeneous. The homogeneous catalysts, including some zinc salts (such as zinc chloride, zinc bromide, zinc iodide, zinc sulfate) [35,36], show higher activity for reaction of GL with urea. The heterogeneous catalysts mainly include metal oxide (such as magnesium oxide, calcium oxide, lanthanum oxide, and zinc oxide) [37,38,39,40], mixed metal oxide catalysts [41], heteropolyacid salts [42], supported ionic liquid catalysts [43], supported metal nanoparticle catalysts (such as Au/MgO nanoparticle catalyst) [44], hydrotalcite catalysts [45,46], and so on. Zinc chloride and zinc sulfate are frequently used as the homogeneous catalysts for the reaction of GL and urea. Park et al. studied the reaction of GL and urea using ZnCl2 as the catalyst and found that during the reaction, ZnCl2 firstly was transformed into Zn(NH3)2Cl2, which further was transformed into Zn(C3H6O3) and NH4Cl [47]. Therefore, they proposed that Zn(C3H6O3) and NH4Cl were the true catalytic active species. However, Fujita and his co-workers considered that the active species was Zn(NCO)2(NH3)2 [35]. Hence, further work is required to determine the catalytic active species for the reaction of GL and urea over a homogeneous zinc salt catalyst. In addition, the reaction kinetics are important for the development of the catalyst and design of an industry reactor [38,48]. However, up to now, few works have been carried out on the kinetics of the reaction of GL and urea. In addition, in reaction process of GL and urea, low pressure or inert gas flow is needed to remove the ammonia gas generated during the reaction, to shift the reaction equilibrium to right side. However, escape of NH3 will reduce the volume of the reaction system, which was ignored in the previous reaction kinetics model.
In this study, the kinetics of the reaction of GL and urea to produce GC using ZnSO4 as the catalyst were studied. The changes in catalyst in the reaction process were deeply investigated and the catalytic active species for the catalyst and the reaction intermediate species were determined according to XRD, FT-IR, and ESI-MS results. The reaction mechanism was also proposed. Based on the above results, a reaction kinetics model considering the change in volume was established, and the model parameters were obtained.

2. Results and Discussion

2.1. Activity of Different Zn Salt Catalysts

The catalytic activities of certain catalysts for the reaction of GL and urea are listed in Table 1. A GL conversion of 42.24% and GC selectivity of 32.29% can be obtained without any catalyst. Meanwhile, the strongly basic KOH and neutral KNO3 show negative activity and there is barely any GL conversion or GC yield for the two catalysts. On the contrary, the amphoteric Zn-based catalysts (zinc oxide and zinc salts) exhibit higher catalytic activities for the reaction of GL and urea. The GC yields for these Zn-based catalysts increase in the following order: Zn(NO3)·6H2O < Zn3(PO4)2 < ZnO < ZnI2 < ZnBr2 < ZnCl2 < ZnSO4. Among these catalysts, ZnSO4 exhibits the highest activity. So, ZnSO4 was selected as the catalyst for the following investigation.

2.2. Effects of Reaction Parameters

2.2.1. Effect of Reaction Temperature

Figure 1 shows the effect of reaction temperature on the reaction of GL and urea over a ZnSO4 catalyst. With an increase in the reaction temperature from 120 °C to 150 °C, both the GL conversion and GC yield increase from 30.24% and 27.57% to 89.51% and 79.47%, respectively. In general, a higher reaction temperature can increase the collision probability of reactant molecules, to accelerate the reaction rate. Furthermore, the reaction of GL and urea is an endothermic reaction and so a higher reaction temperature will also shift the reaction equilibrium to the product side. Therefore, the GL conversion and GC yield increase with the reaction temperature. Meanwhile, the selectivity of GC remains near-constant, at about 90%, with increasing reaction temperature. Urea will decompose when the reaction temperature is over 150 °C and so a suitable reaction temperature is around 140 °C for the reaction of GL and urea [49,50].

2.2.2. Effect of Urea/GL Molar Ratio

Figure 2 shows the effect of the urea-to-GL molar ratio on the reaction of GL and urea over a ZnSO4 catalyst. When the urea-to-GL molar ratio increases from 1:1 to 1.5:1, the GL conversion increases from 78.18% to 85.93% while the GC yield increases from 63.41% to 75.81% firstly and then decreases to 71.13%. Meanwhile, the GC selectivity also firstly increase from 90.61% to 94.37% and then decreases to 82.8% with an increase in the urea-to-GL molar ratio from 1:1 to 1.5:1. For the reaction of urea and GL, the increase in the urea-to-GL molar ratio can shift the reaction equilibrium to the right side, so the conversion of GL increases. However, as the urea-to-GL molar ratio is over 1.1:1, which is larger than the stoichiometric ratio of urea to GL of 1:1, the product GC may further react with unreacted urea to produce the byproduct (2-oxo-1,3-dioxolane 4-yl)methyl carbamate, which is an overreaction product of GC with urea [51]. So, the GC yield and GC selectivity will decrease with the urea-to-GL molar ratio. Hence, the suitable urea-to-GL molar ratio is 1.1:1.

2.2.3. Effect of Catalyst Amount

Figure 3 shows the effect of the catalyst amount on the reaction of GL and urea over ZnSO4 catalyst. As the catalyst amount increases from 0.2 wt% to 5 wt%, the GL conversion and GC yield increase from 54.58% and 23.42% to 80.33% and 75.8%, respectively. With the further increase in catalyst amount to 7 wt%, the GL conversion and GC yield remain near-constant. Meanwhile, the GC selectivity increases with the catalyst amount. ZnSO4 can completely dissolve in the reaction mixture and the catalyst is a homogeneous one, so the increase in catalyst amount can increase the concentration of catalytic active sites in the reaction mixture. Therefore, the GL conversion and GC yield increase with catalyst amount. Hence, the suitable catalyst amount is about 5 wt%.

2.2.4. Effect of Reaction Time

The reaction time also shows an important effect on the reaction of GL and urea over a ZnSO4 catalyst (Figure 4). When the reaction time increases to 120 min, the GL conversion and GC yield increase rapidly to 68.94% and 67.97%, respectively. With a further increase in reaction time to 300 min, the GL conversion and GC yield slowly increase to 79.61% and 75.66%, respectively. The GC selectivity increases with an increase in the reaction time firstly and then slowly decreases with time. The above results indicate that the optimal reaction conditions for the reaction of GL and urea are a reaction temperature of 140 °C, urea-to-GL molar ratio of 1.1:1, catalyst amount of 5 wt% (based on GL weight), and reaction time of 240 min. Under the optimal conditions, the GL conversion, GC yield, and GC selectivity reach to 80.33%, 75.81%, and 94.37%, respectively. In order to check the repeatability of the results, the experiment was repeated three times and the relative error for the GC yield was less than 2.0%. This suggests that the results are reliable. Figure S1 shows the gas chromatograms of the reaction samples (see the Supplementary Materials). A good peak separation is achieved for all the components. Under analytical conditions, urea does not peak. Furthermore, except for GL and GC, the peaks of the other materials, such as byproducts or intermediates, can barely be observed, meaning the concentrations of these byproducts or intermediates are very low. Under reaction conditions, the possible intermediate component may be glyceryl carbamate (GCM) while the byproducts may be glycidol, which is produced from the decomposition of GC, and (2-oxo-1,3-dioxolan-4-yl)methyl carbamate [52], which is generated from the reaction of GC with surplus urea.

2.3. Reaction Mechanism

During the reaction process, firstly, the ZnSO4 catalyst was dissolved quickly in the reaction mixture, and the reaction proceeded in a homogeneous state. However, when the reaction time passed 5 h, some white solids were precipitated from the reaction mixture. This indicates that the ZnSO4 catalyst may be changed into another insoluble species during the reaction process. Meanwhile, when methanol was added to the homogeneous reaction sample at different times, some white solids again precipitated from the reaction sample. In order to determine the components of the white solids, XRD was applied to characterize these precipitates, and the results are illustrated in Figure 5. As shown in Figure 5a, the main phase for the fresh ZnSO4 catalyst is ZnSO4 (at 2θ = 18.6, 26.1, 26.6, 26.9, 29.1, 35.6, see PDF# 00-033-1476). After reaction for 30 min, the precipitated solids are remarkably different from fresh ZnSO4, having been transformed into a mixture of ZnSO4 phase (at 2θ = 18.1, 27.18, 35.9) and Zn(NH3)2SO4 phase (at 2θ = 15.8, 19.2, 20.5, 21.3, 23.6, 26.2, 29.3, 32.2, 36.9, 40.5, see PDF# 00-035-0767). After reaction for 240 min, the precipitated solids are transformed into Zn(C3H6O3) phase (at 2θ = 11.0, 17.3, 20.8, 23.8, 24.8, 27.7) and little (NH4)2SO4 phase (at 2θ = 33.7, 38.8, 41.6, see PDF# 01-084-0130). These results suggest that the ZnSO4 catalyst is transformed into Zn(NH3)2SO4 at the initial stage of the reaction and then further transformed into Zn(C3H6O3).
The ESI-MS spectra of the reaction solution of GL and urea using ZnSO4 as the catalyst with different reaction times are illustrated in Figure 6. The peak at 118 of the m/Z+ ratio may be attributed to GC while the peak at 158 of the m/Z+ ratio is assigned to the reaction intermediate, GCM. With the increase in the reaction time, the peak of GCM becomes small and almost disappears at 300 min. On the contrary, when the reaction time increases from 60 min to 180 min, the peak of GC becomes strong. The results indicate that the reaction of GL and urea firstly produce intermediate GCM, which then is transformed into GC. In addition, the peak at 201 of the m/Z+ ratio may be ascribed to ZnSO4, which quickly disappears after 180 min. The peak at 216 or 234 of the m/Z+ ratio may be assigned to the Zn(NH3)2SO4 species. The peak at 173 of the m/Z+ ratio belongs to the Zn(C3H6O3) species, which becomes strong from 180 min to 300 min. These results suggest that the ZnSO4 catalyst is transformed into Zn(NH3)2SO4, which then is transformed into Zn(C3H6O3). These results are in accordance with those of the XRD.
Figure 7 shows the FT-IR spectra of the reaction solution of GL and urea over the ZnSO4 catalyst with different times. The peak at 1715 cm−1 is attributed to -CH2- of GCM [53], which becomes gradually weaker with the reaction time. The peak at 1790 cm−1 is assigned to C=O of GC while the peaks at 1400 cm−1 and 1185 cm−1 are ascribed to –OH and C-C of GC [35], whose intensities increase with the reaction time. These results suggest that the GL reacts with urea to generate GCM firstly and then GCM is transformed into GC, which is in accordance with the results of ESI-MS. In addition, the peak at 2200 cm−1 is attributed to the N=C=O group of HNCO [54], which is produced from the decomposition of urea.
The above results strongly indicate that during the reaction, the catalyst ZnSO4 firstly reacts with NH3 produced from decomposition of urea to transform into Zn(NH3)2SO4, which then reacts with GL to generate Zn(C3H6O3) and (NH4)2SO4. Zn(C3H6O3) and (NH4)2SO4 are the true catalysts for the reaction of GL and urea. Table 2 shows the catalytic activities of some catalysts for the reaction of GL and urea. Although Zn(C3H6O3) and (NH4)2SO4 show lower activity than ZnSO4, the mixture of Zn(C3H6O3) and (NH4)2SO4 exhibits similar activity to ZnSO4.
In order to clarify the roles of Zn(C3H6O3) and (NH4)2SO4 for the activation of GL and urea, the FT-IR spectra are used to characterize the interaction of Zn(C3H6O3) or (NH4)2SO4 with GL or urea, and the results are illustrated in Figure 8 and Figure 9. In Figure 8a, the peak at 3432 cm−1 is attributed to hydroxyl of GL, which shifts to a lower frequency of 3402 cm−1 when GL is mixed with (NH4)2SO4 at 30 min. As the mixing time increases to 120 min, the peak further shifts to a lower frequency of 3394 cm−1, meaning that the interaction between (NH4)2SO4 and GL becomes strong (Figure 8c–e). The results suggest that the hydrogen-bond interaction may be yielded between (NH4)2SO4 and hydroxyl of GL, which results in the activation of molecular GL. Figure 9 shows the FTIR spectra of a mixture of urea and Zn(C3H6O3). The two materials are solid, so a certain amount of urea was mixed with a certain amount of Zn(C3H6O3) in an agate mortar and ground to give samples with different grinding times. In Figure 9a, the peaks at 1672 cm−1 are attributed to the C=O stretching vibration of urea, which also shift to a lower frequency of 1668 cm−1 as urea is mixed with Zn(C3H6O3) for 120 min [54]. Meanwhile, the peaks at 3440 cm−1 of -NH2 of urea also shift to the high-frequency side. These results suggest that there may be a hydrogen-bond interaction between urea and Zn(C3H6O3), which leads to the activation of molecular urea.
Furthermore, NMR analysis was used to study the interactions between GL and (NH4)2SO4 and between urea and Zn(C3H6O3). 1H NMR analysis of the mixture of GL and (NH4)2SO4 with the contact time of 2 h was performed and the results are shown in Figure 10. Meanwhile, 13C NMR analysis of the mixture of urea and Zn(C3H6O3) with the contact time of 2 h was also conducted and the results are shown in Figure 11. In Figure 10, the peak at 4.41 ppm can be attributed to the hydrogen atom of the primary hydroxyl of pure GL, which shifts to 4.40 ppm when GL is mixed with (NH4)2SO4. This means that there may be a hydrogen bond interaction between (NH4)2SO4 and the primary hydroxyl of GL. Meanwhile, as shown in Figure 11, the peak at 160.0911 ppm is assigned to the carbon atom of the carbonyl of pure urea, which also shifts to 160.0385 ppm when urea is mixed with Zn(C3H6O3) for 2 h. These results are in accordance with those of the FTIR and suggest that there may be an interaction between urea and Zn(C3H6O3).
Based on the above results, the possible reaction mechanism for GL and urea over a ZnSO4 catalyst is proposed (Scheme 2). During the reaction process, NH3 firstly is produced through the decomposition of urea. Meanwhile, NH3 also may be generated by the reaction of GL with urea without a catalyst. Then, the soluble ZnSO4 catalyst reacts with NH3 to give Zn(NH3)2SO4, which further reacts with GL to produce Zn(C3H6O3) and (NH4)2SO4. Afterward, molecular GL is activated by the hydrogen-bond interaction of primary hydroxyl of GL with (NH4)2SO4 while molecular urea is also activated by the interaction between C=O of urea and Zn(C3H6O3). The oxygen atom of primary hydroxyl for activated GL launches a nucleophilic attack on the carbon atom of C=O of urea to give the intermediate GCM and NH3. Finally, GC is obtained via an intramolecular nucleophilic attack and cyclization of GCM, and another molecule of NH3 is also generated.

2.4. Reaction Kinetics

2.4.1. Reaction Rate

The reaction kinetics were further studied based on the above results to derive the reaction rate constant and active energy. The reaction process includes a complex cascade reaction: firstly, GL reacts with urea to produce intermediate GCM and NH3 over the catalyst, and then, GC is generated from the intramolecular cyclization of GCM, and another molecule of NH3 is also released. The reaction is carried out in a vacuum and NH3 is continually drawn out from the reaction system, so the above two reactions are irreversible. The reaction formulas are listed as follows:
R 1 :   GL + urea GCM + NH 3 ( g )
R 2 :   GCM GC + NH 3 ( g )
Meanwhile, the following reactions may take place in the reaction process:
ZnSO 4 + 2 NH 3 Zn ( NH 3 ) 2 SO 4
Zn ( NH 3 ) 2 SO 4 + GL Zn ( C 3 H 6 O 3 ) + ( NH 4 ) 2 SO 4
Reactions (3) and (4) are the conversion reaction of the catalyst. Since obtaining the concentrations of the catalyst components in the reaction is difficult and the reaction degrees of Reactions (3) and (4) are small, they are ignored in the kinetics model. Side reactions, such as the reaction of GC with urea and the decomposition of GC to produce glycidol, may also take place. However, the concentrations of the byproducts are so small that they cannot be found in a gas chromatogram or ESI-MS spectra. So, these side reactions are also not considered in the kinetics model.
According to the reaction mechanism, the main catalysts may be Zn(C3H6O3) and (NH4)2SO4. The reaction process has the following steps: firstly, GL and urea are activated through interaction with the catalyst, respectively, and then the GCM and one molecule of NH3 are generated by the reaction of activated GL and urea. Finally, GC and another molecule of NH3 are obtained through an intramolecular nucleophilic attack and cyclization of GCM. The above elementary reactions in the reaction process are listed as follows:
GL + cat GL *
urea + cat urea *
GL * + urea * GCM + NH 3 ( g )
GCM + cat GCM *
GCM * GC + NH 3 ( g )
where ‘cat’ denotes the catalyst, and GL*, urea*, and GCM* denote the activated GL, activated urea, and activated GCM molecules, respectively. Among the above Reactions (5)~(9), the Reactions (5), (6), and (8) are the activation processes of GL, urea, and GCM, respectively, whose reaction rates are very quick, meaning reaction equilibriums are reached quickly. In addition, for the reaction of GL and urea, though GL is liquid and urea is solid, urea is dissolved quickly in the GL phase when urea is mixed with GL. Meanwhile, though the catalyst ZnSO4 is solid, it is also dissolved in the reaction mixture as ZnSO4 is added to the reaction system. Furthermore, though the solid of Zn(C3H6O3)2 is separated from the reaction system when the reaction time passes 5 h, its amount is small. Therefore, the reaction can be regarded as a homogeneous one. So, we have the following expressions:
C G L * = K 1 C G L C c a t
C u r e a * = K 2 C u r e a C c a t
C G C M * = K 4 C G C M C c a t
where K1, K2, and K4 are the equilibrium constants of Reactions (5), (6), and (8), respectively. Ci is the component concentration, mol/L. According to the law of mass action, the reaction rates of Reactions (7) and (9) can be calculated by using the following equations, respectively:
r 1 = k 1 C G L * C u r e a * = k 1 K 1 C G L C c a t K 2 C u r e a C c a t
r 2 = k 2 C G C M * = k 2 K 4 C G C M C c a t
Although the catalyst ZnSO4 is transformed into Zn(C3H6O3) and (NH4)2SO4 during the reaction, the concentration of the catalyst, Ccat, can be considered as constant during the reaction because there is a large number of active sites. By applying commands k 1 = k 1 K 1 K 2 C c a t C c a t and k 2 = k 2 K 4 C c a t , we obtain
r 1 = k 1 C G L C u r e a = k 1 C A C B
r 2 = k 2 C G C M = k 2 C c
where r1 and r2 are the reaction rates of Reactions (7) and (9), mol/(L·min), respectively; k1 and k2 are the apparent reaction rate constants of Reactions (7) and (9), respectively; and A, B, C, and D denote the GL, urea, GCM, and GC, respectively. Accordingly, it can be found that the above reaction rate model is well-connected with the proposed reaction mechanism.
The Arrhenius equation is used to calculate reaction rate constant:
k j = k j * exp ( E a j R T )
where kj* is the pre-exponential factor of reaction rate constant kj; Eaj is the active energy of reaction j, kJ/mol; R is the gas constant (R = 8.314 J/(mol·K); and T is the reaction temperature, K.
The reaction is carried out under vacuum conditions of 5 kPa to remove the generated NH3, so the mass of the reaction mixture should decrease. Meanwhile, the volume of the reaction system also decreases with the reaction time due to the molar volume of GC or GMC being less than the total molar volumes of GL and urea. So, the variation in the volume of the reaction mixture should also be considered in the kinetics model. Some assumptions are made that the volume of the reaction mixture may be obtained by summing the volumes of every component and that all NH3 are removed from the reaction system by the vacuum. Meanwhile, the effect of reaction temperature on the molar volumes of components is ignored. According to the material balance for component i, the reactor model is obtained, as follows (The detailed process of the establishment for the kinetics model is given in Appendix A):
d C A d t = 1 + C A ( V C ¯ V A ¯ V B ¯ ) r 1 C A ( V D ¯ V C ¯ ) r 2
d C B d t = 1 + C B ( V C ¯ V A ¯ V B ¯ ) r 1 C B ( V D ¯ V C ¯ ) r 2
d C C d t = 1 C c ( V C ¯ V A ¯ V B ¯ ) r 1 1 + C C ( V D ¯ V C ¯ ) r 2
d C D d t = C D ( V C ¯ V A ¯ V B ¯ ) r 1 + 1 C D ( V D ¯ V C ¯ ) r 2
d V d t = ( V C ¯ V A ¯ V B ¯ ) r 1 + ( V D ¯ V C ¯ ) r 2 V
The initial condition is
t = 0, CA = CA0, CB = CB0, CC = CD = 0, V = V0.
where V i ¯ is the molar volume of component i, L/mol; V is the volume of reaction mixture, L; CA0 and CB0 are the initial concentrations of GL and urea, mol/L, respectively; and V0 is the initial volume of the reaction mixture, L.

2.4.2. Solution of Kinetics Model

The kinetics model (18~22) is a set of ordinary differential equations. The four-order Runge–Kutta method can be used to solve these equations [55]. By minimizing the following expression (24), the reaction rate constants can be obtained:
ψ = j N ( C i , j exp C i , j c a l ) 2 + ( V j exp V j c a l ) 2
which is the squared sum of the residuals between the experimental data of the component concentration, C i , j exp (i = GL and GC), the volume of the reaction system, Vjexp, and the model predictions, C i , j c a l and Vjcal. MATLAB software was used to write the calculation program and the Levenberg–Marquardt algorithm was used to perform the correlation. The molar volume of GCM was calculated by using Aspen plus software. The rate constants at every temperature were firstly correlated and then their expressions were obtained from the Arrhenius Equation (17), and the activation energies of every reaction were also obtained. The results are listed in Table 3. The correlation coefficients for k1 and k2 are 0.9995 (in the fourth column, Table 3), meaning the fitting accuracy is high. The activation energies of R1 and R2 are 143.39 kJ/mol and 87.29 mol/L, respectively. The activation energies of the reaction of GL with urea over other catalysts are listed in Table 4. The activation energies for ZnSO4 were closest to those over the catalyst MgO.
Figure 12 gives a comparison of the experimental and calculated concentrations for every component at different temperatures. In these figures, the points denote the experimental concentrations and lines denote the calculated concentrations. The intermediate GCM cannot be measured with a gas chromatograph, so only the calculated concentrations for GCM are given. The experimental concentrations of urea were calculated from the concentrations of GL. The calculated concentrations are close to the experimental values. Furthermore, the concentration of GCM rapidly increases firstly and then decreases gradually with increasing reaction time, which matches the change characteristic of intermediate species. Figure 13 shows a comparison of experimental and calculated volumes of the reaction mixture at different temperatures. The relative deviation for volume is −10%~+20%.
A statistical test was used to further evaluate the kinetics model, and the results are listed in Table 5. The F value is the ratio of the mean regression sum of squares to the mean residual sum of squares, which can be calculated by using the following equation:
F = i = 1 N [ ( C i c a l ) 2 + ( V i c a l ) 2 ] / p i = 1 N [ ( C i exp C i c a l ) 2 + ( V i exp V i c a l ) 2 ] / ( N p )
where N is the experimental run number and p is the number of parameters in the kinetics model. FT is the tabulated value corresponding to degrees of freedom at a 5% confidence level. In general, if F > 10 FT, a model is considered acceptable [57,58]. As shown in Table 5, the F value is 216.37, which is >10 FT (=23.6). Hence, the established kinetics model is suitable for the reaction of GL with urea over a ZnSO4 catalyst.

3. Materials and Methods

3.1. Materials

Glycerol (GL) (>99%), urea (AR), potassium hydroxide (KOH) (AR), potassium nitrate (KNO3) (AR), ammonium sulfate ((NH4)2SO4) (AR), zinc acetate dihydrate ((CH3COO)2Zn∙2H2O) (AR), zinc nitrate hexahydrate (Zn(NO3)2∙6H2O) (AR), zinc oxide (ZnO) (AR), and methanol (>99.5%) were purchased from Sinopharm Chemical Regent Co., Ltd., Shanghai, China. Glycerol carbonate (GC) (90.0%) was obtained from TCI Shanghai. Zinc sulfate (ZnSO4) (AR) and zinc phosphate (Zn3(PO4)2) (AR) were purchased from Shanghai yuanye Bio-Technology Co., Ltd. (Shanghai, China). Tetraethylene glycol (TEG) (99%), n-Butyl alcohol (99.9%), zinc bromide (ZnBr2) (AR), and zinc iodide (ZnI2) (AR) were purchased from Shanghai Aladdin Biochemical Technology Co., Ltd (Shanghai, China). Zinc chloride (ZnCl2) was sourced from Chengdu Chron Chemical Co., Ltd. (Chengdu, China). N2 (99.9%), H2 (99.9%), and air (99.9%) were obtained from Wuhan Zhongxin Ruiyuan Gas Co., Ltd. (Wuhan, China). All the chemicals were used without further purification.
Zn(C3H6O3) was prepared from the reaction of GL and zinc acetate dihydrate, using a method that can be found in the literature [49].

3.2. Reaction Procedure

The synthesis of GC from the reaction of GL with urea by using ZnSO4 as a catalyst was carried out in a 25 mL round-bottomed 3-neck glass flask as a batch reactor equipped with a magnetic stirrer, thermometer, reflux condenser, and vacuum pump. Energy was applied to the glass flask with a constant temperature oil bath. The accuracy of the temperature measurements was ± 0.5 °C. In a typical run, 0.1 mol of GL (9.21 g) was added to the flask and heated to 140 °C under a reduced pressure with a stirring speed of 600 rpm. Then, 0.11 mol of urea (6.61 g) and 0.46 g of catalyst were loaded into the reactor quickly and the reaction mixture was kept at 140 °C and a reduced pressure of 5 kPa for 4 h with continuous stirring. The vacuum system was used to remove NH3 produced during the reaction. After the reaction, the reaction mixture was cooled to room temperature and its mass and volume were measured. The liquid mixture was sampled for analysis.
It was found that the volume of the reaction system decreased continuously due to the removal of ammonia. In order to obtain accurate concentrations of the reaction components at different reaction times for the kinetics investigation, an experiment at every reaction time was carried out individually.
The Fuli GC-9790 II Gas Chromatograph (Fuli 9790-II, Fuli, Wenling, China) with a flame ionization detector (FID) and KB-WAX capillary column (30 m long and 0.25 mm I.D.) was used to analyze the reaction samples. The column temperature was programmed at 70 °C for 2 min, and then increased to 250 °C with 15 °C/min speed and held there for 5 min. Tetraethylene glycol was used as the internal standard for GL and GC, and n-butyl alcohol was used for the other byproducts. The approximately 0.7 g of methanol used as a solvent was added to the sample.
The GL conversion, XGL, GC yield, YGC, GC selectivity, and SGC were calculated using the following expressions:
X GL = n GL in n GL out n GL in × 100 %
Y GC = n GC out n GL in × 100 %
S GC = n GC out n GL in n GL out × 100 %
where n G L i n and n G L o u t are the initial molar number of GL and the molar number of GL after the reaction, respectively, and n G C o u t is the molar number of GC after the reaction.

3.3. Characterization

The FT-IR spectra of the samples were measured using a Bruker VERTEX 70 instrument (Bruker, Karlsruhe, Germany). The scanning range is 4000~400 cm−1 and the resolution ratio is 2 cm−1.
The X-ray diffraction (XRD) patterns of the catalysts were measured on a PANalytical X’Pert Pro X-ray diffractometer (X’Pert PRO, PANalytical B.V., Almelo, Netherlands) with Cu Kα radiation at 30 kV and 15 mA, over the range of 5 to 90°.
ESI-MS spectra were measured by using an UltiMate-3000-micro TOF instrument (Thermofisher-Bruker Daltonics Inc. Blerica, Massachusetts, MA, USA; Bremen, Germany) with the positive ion mode at 50~1200 m/Z, a dry heater temperature of 200 °C, a nebulizer pressure of 1 bar, and a dry gas flow velocity of 6 L/min.

4. Conclusions

The GC yield for these Zn-based catalysts increase in the following order: Zn(NO3)·6H2O < Zn3(PO4)2 < ZnO < ZnI2 < ZnBr2 < ZnCl2 < ZnSO4. Among these catalysts, ZnSO4 possesses the highest activity. Under a reaction temperature of 140 °C, catalyst amount of 5 wt% (based on the GL weight), urea-to-GL molar ratio of 1.1:1, and reaction time of 240 min, the GL conversion and GC yield can reach 80.33% and 75.81%, respectively, over a ZnSO4 catalyst. During the reaction, the ZnSO4 catalyst is transformed into Zn(NH3)2SO4 at the initial stage of the reaction and then further transformed into Zn(C3H6O3). Zn(C3H6O3) and (NH4)2SO4 may be the true active species for the activation of urea and GL, respectively. The reaction process includes a cascade reaction: GL reacts with urea to produce intermediate GCM and NH3 firstly, and then GCM is transformed into GC through intramolecular cyclization, and another molecular of NH3 is also released. The reaction kinetics model considering the change in volume of the reaction system has been established, and the activation energies were obtained by fitting the experimental data.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14010041/s1, Figure S1: Gas chromatograms of the reaction samples: (1) methanol; (2) n-butyl alcohol; (3) GL; (4) tetraethylene glycol; (5) GC.

Author Contributions

Conceptualization, H.W. and J.M.; methodology, J.M.; software, H.W.; validation, J.M. and H.W.; formal analysis, J.M.; investigation, J.M.; resources, H.W.; data curation, J.M.; writing—original draft preparation, H.W.; writing—review and editing, H.W.; visualization, H.W.; supervision, H.W.; project administration, H.W.; funding acquisition, H.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundamental Research Funds for the Central Universities of China (2011QN117).

Data Availability Statement

The authors state that the data pertaining to the manuscript will be made available upon request.

Acknowledgments

XRD and ESI-MS analysis were performed in the Analytical and Testing Center, and FT-IR analysis was performed in the Experimental Teaching Center of Chemistry and Chemical Engineering, School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

Cimolar concentration of component i (mol/L)
Eajactivation energy of reaction j (kJ/mol)
FF-test function
FTtabulated value of F distribution
k1reaction rate constant of reaction R1 (L·min−1·mol−1)
k2reaction rate constant of reaction R2 (min−1)
kj*pre-exponential factor (L·min−1·mol−1 or min−1)
Mimolar mass of component i (kg/mol)
mimass of component i (kg)
nimolar number of component i (mol)
Nnumber of experimental runs
pnumber of parameters
Rgas constant, 8.314 J/(mol·K)
R2correlation coefficient
rjreaction rate of reaction j (mol/(L·min))
SGCselectivity of glycerol carbonate (-)
Ttemperature (K)
treaction time (min)
Vvolume of reaction mixture (L)
V i ¯ molar volume of component i (L/mol)
XGLconversion of glycerol (-)
YGCyield of glycerol carbonate (-)
GCglycerol carbonate
GCMglyceryl carbamate
GLglycerol
Greek letters
ψobjective function
ρidensity of component i (kg/L)

Appendix A

It was assumed that the volume of the reaction mixture can be obtained by summing the volume of every component, and that all NH3 are removed from the reaction system by the vacuum. Meanwhile, the effect of reaction temperature on the molar volumes of components was ignored. So, we obtained
d V d t = d ( V A + V B + V C + V D ) d t
where, Vi is the volume of component i, L, and V is the total volume of reaction mixture, L. A, B, C, and D denote the GL, urea, GCM, and GC, respectively.
We also obtained
V i = m i ρ i
where mi is the mass of component i, kg, and ρi is the density of component i, kg/L. By introducing (A2) into (A1), we obtained
d V d t = d ( m A ρ A + m B ρ B + m C ρ C + m D ρ D ) d t
We also obtained
m i = M i n i
where ni is the molar number of component i, mol, and Mi is the molar mass of component i, kg/mol. So, by introducing (A4) into (A3), we obtained
d V d t = d ( M A n A ρ A + M B n B ρ B + M C n C ρ C + M D n D ρ D ) d t = V A ¯ d n A d t + V B ¯ d n B d t + V C ¯ d n C d t + V D ¯ d n D d t
where V i ¯ is the molar volume of component i, L/mol. According to material balance for component i, we obtained
d n A d t = r A V = r 1 V
d n B d t = r B V = r 1 V
d n C d t = r C V = ( r 1 r 2 ) V
d n D d t = r D V = r 2 V
where ri (i = A, B, C, D) is the reaction rate of component i, mol/(L·min), and r1 and r2 are the reaction rates of reactions R1 and R2, mol/(L·min).
So, by introducing (A6)~(A9) into (A5), we obtained
d V d t = V A ¯ ( r 1 V ) + V B ¯ ( r 1 V ) + V C ¯ ( r 1 r 2 ) V + V D ¯ r 2 V
By combining like terms, we obtained
d V d t = ( V C ¯ V A ¯ V B ¯ ) r 1 + ( V D ¯ V C ¯ ) r 2 V
Equation (A11) justifies Equation (22).
According to the material balance for component A, we also obtained
d n A V d t = d ( V C A ) V d t = V d C A V d t + C A d V V d t
where Ci is the molar concentration of component i, mol/L. According to (A6), we obtained
d C A d t + C A d V V d t = r 1
By introducing (A11) into (A13), we obtained
d C A d t = 1 + C A ( V C ¯ V A ¯ V B ¯ ) r 1 C A ( V D ¯ V C ¯ ) r 2
Equation (A14) justifies Equation (18). In the same way, Equations (19)~(21) could also be obtained.

References

  1. Alibas, I.; Yilmaz, A.; Alibas, K.; Arslan, M.; Koc, C. The effect of different drying methods and microalgae species on the quality parameters of biodiesel obtained by transesterification technique. Biomass Bioenerg. 2023, 168, 106688. [Google Scholar] [CrossRef]
  2. Javed, F.; Zimmerman, W.B.; Fazal, T.; Hafeez, A.; Mustafa, M.; Rashid, N.; Rehman, F. Green synthesis of biodiesel from microalgae cultivated in industrial wastewater via microbubble induced esterification using bio-MOF based heterogeneous catalyst. Chem. Eng. Res. Des. 2023, 189, 707–720. [Google Scholar] [CrossRef]
  3. Chen, C.; Mira, D.; Xing, Z.; Jiang, X. Thermophysical property prediction of biodiesel mixtures at extreme conditions using molecular dynamics simulation. J. Mol. Liq. 2022, 367, 120423. [Google Scholar] [CrossRef]
  4. Masera, K.; Hossain, A.K. Biofuels and thermal barrier: A review on compression ignition engine performance, combustion and exhaust gas emission. J. Energy Inst. 2019, 92, 783–801. [Google Scholar] [CrossRef]
  5. Julio, A.A.V.; Milessi, T.S.; Batlle, E.A.O.; Lora, E.E.S.; Maya, D.M.Y.; Palacio, J.C.E. Techno-economic and environmental potential of renewable diesel as complementation for diesel and biodiesel in Brazil: A comprehensive review and perspectives. J. Clean. Prod. 2022, 371, 133431. [Google Scholar] [CrossRef]
  6. Meng, L.; Su, H.; Du, S.; Wang, Y.; Li, M. Brief discussion on the production technology and development trend of biodiesel in our country. Chem. Eng. Equip. 2019, 7, 274–276. (In Chinese) [Google Scholar]
  7. Song, S.S.; Tian, B.C.; Chen, H.; Chi, Z.; Liu, G.L.; Chi, Z.M. Transformation of corncob-derived xylose into intracellular lipid by engineered strain of Aureobasidium melanogenum P10 for biodiesel production. Renew. Energy 2022, 200, 1211–1222. [Google Scholar] [CrossRef]
  8. Akram, F.; Haq, I.; Raja, S.I.; Mir, A.S.; Qureshi, S.S.; Aqeel, A.; Shah, F.I. Current trends in biodiesel production technologies and future progressions: A possible displacement of the petro-diesel. J. Clean. Prod. 2022, 370, 133479. [Google Scholar] [CrossRef]
  9. Ashine, F.; Kiflie, Z.; Prabhu, S.V.; Tizazu, B.Z.; Varadharajan, V.; Rajasimman, M.; Joo, S.W.; Vasseghian, Y.; Jayakumar, M. Biodiesel production from Argemone mexicana oil using chicken eggshell derived CaO catalyst. Fuel 2013, 332, 126166. [Google Scholar] [CrossRef]
  10. Caro, P.; Bandres, M.; Urrutigoïty, M.; Cecutti, C.; Thiebaud-Roux, S. Recent progress in synthesis of glycerol carbonate and evaluation of its plasticizing properties. Front. Chem. 2019, 7, 308. [Google Scholar] [CrossRef]
  11. Pagliaro, M.; Ciriminna, R.; Kimura, H.; Rossi, M.; Pina, C.D. From glycerol to value-added products. Angew. Chem. Int. Ed. 2007, 46, 4434–4440. [Google Scholar] [CrossRef] [PubMed]
  12. Atadashi, I.M.; Aroua, M.K.; Abdul Aziz, A. High quality biodiesel and its diesel engine application: A review. Renew. Sust. Energy Rev. 2010, 14, 1999–2008. [Google Scholar] [CrossRef]
  13. Demirbas, A. Biodiesel production from vegetable oils via catalytic and non-catalytic supercritical methanol transesterification methods. Prog. Energy Combust. Sci. 2005, 31, 466–487. [Google Scholar] [CrossRef]
  14. Sonnati, M.O.; Amigoni, S.; Taffin de Givenchy, E.P.; Darmanin, T.; Choulet, O.; Guittard, F. Glycerol carbonate as a versatile building block for tomorrow: Synthesis, reactivity, properties and applications. Green Chem. 2013, 15, 283–306. [Google Scholar] [CrossRef]
  15. Zheng, Y.; Chen, X.; Shen, Y. Commodity chemicals derived from glycerol, an important biorefinery feedstock. Chem. Rev. 2008, 108, 5253–5277. [Google Scholar] [CrossRef] [PubMed]
  16. An, S.; Sun, Y.; Song, D.; Zhang, Q.; Guo, Y.; Shang, Q. Arenesulfonic acid-functionalized alkyl-bridged organosilica hollow nanospheres for selective esterification of glycerol with lauric acid to glycerol mono- and dilaurate. J. Catal. 2016, 342, 40–54. [Google Scholar] [CrossRef]
  17. Binhayeeding, N.; Klomklao, S.; Sangkharak, K. Utilization of waste glycerol from biodiesel process as a substrate for mono-, di-, and triacylglycerol production. Energy Procedia 2017, 138, 895–900. [Google Scholar] [CrossRef]
  18. Islam, Z.; Klein, M.; Aßkamp, M.R.; Ødum, A.S.R.; Nevoigt, E. A modular metabolic engineering approach for the production of 1,2- propanediol from glycerol by Saccharomyces cerevisiae. Metab. Eng. 2017, 44, 223–235. [Google Scholar] [CrossRef]
  19. Zhou, S.; Lama, S.; Sankaranarayanan, M.; Park, S. Metabolic engineering of Pseudomonas denitrificans for the 1,3-propanediol production from glycerol. Bioresour. Technol. 2019, 292, 121933. [Google Scholar] [CrossRef]
  20. Maina, S.; Kachrimanidou, V.; Ladakis, D.; Papanikolaou, S.; Castro, A.M.; Koutinas, A. Evaluation of 1,3-propanediol production by two Citrobacter freundii strains using crude glycerol and soybean cake hydrolysate. Environ. Sci. Pollut. Res. 2019, 26, 35523–35532. [Google Scholar] [CrossRef]
  21. Guedes, P.H.P.S.; Luz, R.F.; Cavalcante, R.M.; Young, A.F. Process simulation for technical and economic evaluation of acrolein and glycerol carbonate production from glycerol. Biomass Bioenerg. 2023, 168, 106659. [Google Scholar] [CrossRef]
  22. Rittiron, P.; Niamnuy, C.; Donphai, W.; Chareonpanich, M.; Seubsai, A. Production of glycerol carbonate from glycerol over templated-sodium-aluminate catalysts prepared using a spray-drying method. ACS Omega 2019, 4, 9001–9009. [Google Scholar] [CrossRef] [PubMed]
  23. 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] [CrossRef]
  24. Pradhan, G.; Jaiswal, S.; Sharma, Y.C. Exploring the promotional effect of transition metals (Cr and V) on the catalytic activity of MgO for glycerol carbonate synthesis. Mol. Catal. 2022, 526, 112332. [Google Scholar] [CrossRef]
  25. Nomanbhay, S.; Ong, M.Y.; Chew, K.W.; Show, P.L.; Lam, M.K.; Chen, W.H. Organic carbonate production utilizing crude glycerol derived as by-product of biodiesel production: A review. Energies 2020, 13, 1483–1506. [Google Scholar] [CrossRef]
  26. 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] [CrossRef]
  27. Arora, S.; Gosu, V.; Arun Kumar, U.K.; Subbaramaiah, V. Valorization of glycerol into glycerol carbonate using the stable heterogeneous catalyst of Li/MCM-41. J. Clean. Prod. 2021, 295, 126437. [Google Scholar] [CrossRef]
  28. Jaiswal, S.; Sahani, S.; Shar, Y.C. Enviro-benign synthesis of glycerol carbonate utilizing bio-waste glycerol over Na-Ti based heterogeneous catalyst: Kinetics and E- metrics studies. J. Environ. Chem. Eng. 2022, 10, 107485. [Google Scholar] [CrossRef]
  29. Rousseau, J.; Rousseau, C.; Lynikaite, B.; Šačkus, A.; Leon, C.; Rollin, P.; Tatibouet, A. Tosylated glycerol carbonate, a versatile bis-electrophile to access new functionalized glycidol derivatives. Tetrahedron 2009, 65, 8571–8581. [Google Scholar] [CrossRef]
  30. Hu, J.; Li, J.; Gu, Y.; Guan, Z.; Mo, W.; Ni, Y.; Li, T.; Li, G. Oxidative carbonylation of glycerol to glycerol carbonate catalyzed by PdCl2(phen)/KI. Appl. Catal. A Gen. 2010, 386, 188–193. [Google Scholar] [CrossRef]
  31. Li, J.; Wang, T. Chemical equilibrium of glycerol carbonate synthesis from glycerol. J. Chem. Thermodyn. 2011, 43, 731–736. [Google Scholar] [CrossRef]
  32. Aresta, M.; Dibenedetto, A.; Nocito, F.; Pastore, C. A study on the carboxylation of glycerol to glycerol carbonate with carbon dioxide: The role of the catalyst, solvent and reaction conditions. J. Mol. Catal. A Chem. 2006, 257, 149–153. [Google Scholar] [CrossRef]
  33. Patel, Y.; George, J.; Pillai, S.M.; Munshi, P. Effect of liophilicity of catalyst in cyclic carbonate formation by transesterification of polyhydric alcohols. Green Chem. 2009, 11, 1056–1060. [Google Scholar] [CrossRef]
  34. Kondawar, S.E.; Mane, R.B.; Vasishta, A.; More, S.B.; Dhengale, S.D.; Rode, C.V. Carbonylation of glycerol with urea to glycerol carbonate over supported Zn catalysts. Appl Petrochem. Res. 2017, 7, 41–53. [Google Scholar] [CrossRef]
  35. 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] [CrossRef]
  36. Endah, Y.K.; Kim, M.S.; Choi, J.; Jae, J.; Lee, S.D.; Lee, H. Consecutive carbonylation and decarboxylation of glycerol with urea for the synthesis of glycidol via glycerol carbonate. Catal. Today 2017, 293–294, 136–141. [Google Scholar] [CrossRef]
  37. Zhang, P.; Liu, L.; Fan, M.; Dong, Y.; Jiang, P. The value-added utilization of glycerol for the synthesis of glycerol carbonate catalyzed with a novel porous ZnO catalyst. RSC Adv. 2016, 6, 76223–76230. [Google Scholar] [CrossRef]
  38. Fernandes, G.P.; Yadav, G.D. Selective glycerolysis of urea to glycerol carbonate using combustion synthesized magnesium oxide as catalyst. Catal. Today 2018, 309, 153–160. [Google Scholar] [CrossRef]
  39. 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] [CrossRef]
  40. Wang, L.; Ma, Y.; Wang, Y.; Liu, S.; Deng, Y. Efficient synthesis of glycerol carbonate from glycerol and urea with lanthanum oxide as a solid base catalyst. Catal. Commun. 2011, 12, 1458–1462. [Google Scholar] [CrossRef]
  41. Baek, J.; Ryu, Y.B.; Kim, M.H.; Moon, M.J.; Lee, M.S. Synthesis of glycerol carbonate from biomass glycerol: Focus on the calcination temperature. Mater. Sci. Forum. 2012, 724, 374–377. [Google Scholar] [CrossRef]
  42. Kumar, C.R.; Jagadeeswaraiah, K.; Sai Prasad, P.S.; Lingaiah, N. Samarium-exchanged heteropoly tungstate: An efficient solid acid catalyst for synthesis of glycerol carbonate from glycerol and benzylation of anisole. ChemCatChem 2012, 4, 1360–1367. [Google Scholar] [CrossRef]
  43. 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. A Gen. 2014, 473, 31–40. [Google Scholar] [CrossRef]
  44. Rahim, M.H.A.; He, Q.; Lopez-Sanchez, J.A.; Hammond, C.; Dimitratos, N.; Sankar, M.; Carley, A.F.; Kiely, C.J.; Knight, D.W.; Hutchings, G.J. Gold, palladium and gold–palladium supported nanoparticles for the synthesis of glycerol carbonate from glycerol and urea. Catal. Sci. Technol. 2012, 2, 1914–1924. [Google Scholar] [CrossRef]
  45. Sun, Y.; Tong, X.; Wu, Z.; Liu, J.; Yan, Y.; Xue, S. A sustainable preparation of glycerol carbonate from glycerol and urea catalyzed by hydrotalcite-like solid catalysts. Energy Technol. 2014, 2, 263–268. [Google Scholar] [CrossRef]
  46. 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. A Gen. 2018, 555, 36–46. [Google Scholar] [CrossRef]
  47. 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. A Gen. 2012, 433–434, 35–40. [Google Scholar] [CrossRef]
  48. 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] [CrossRef]
  49. 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] [CrossRef]
  50. Schaber, P.M.; Colson, J.; Higgins, S.; Thielen, D.; Anspach, B.; Brauer, J. Thermal decomposition (pyrolysis) of urea in an open reaction vessel. Thermochim. Acta 2004, 424, 131–142. [Google Scholar] [CrossRef]
  51. 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] [CrossRef]
  52. Luo, W.; Sun, L.; Yang, Y.; Chen, Y.; Zhou, Z.; Liu, J.; Wang, F. Cu–Mn composite oxides: Highly efficient and reusable acid–base catalysts for the carbonylation reaction of glycerol with urea. Catal. Sci. Technol. 2018, 8, 6468–6477. [Google Scholar] [CrossRef]
  53. Wang, D.; Zhang, X.; Liu, C.; Cheng, T. Synthesis of glycerol carbonate from glycerol and urea over lanthanum compounds. React. Kinet. Mech. Catal. 2015, 115, 597–609. [Google Scholar] [CrossRef]
  54. Turney, T.W.; Patti, A.; Gates, W.; Shaheen, U.; Kulasegaram, S. Formation of glycerol carbonate from glycerol and urea catalysed by metal monoglycerolatest. Green Chem. 2013, 15, 1925–1931. [Google Scholar] [CrossRef]
  55. Wang, H.; Li, G. Kinetic study on the synthesis of ethyl nitrite by the reaction of C2H5OH, O2, and NO in a trickle bed reactor. Chem. Eng. J. 2010, 163, 422–428. [Google Scholar] [CrossRef]
  56. Lertlukkanasuk, N.; Phiyanalinmat, S.; Kiatkittipong, W.; Arpornwichanop, A.; Aiouache, F.; Assabumrungrat, S. Reactive distillation for synthesis of glycerol carbonate via glycerolysis of urea. Chem. Eng. Process. 2013, 70, 103–109. [Google Scholar] [CrossRef]
  57. Wang, H.; Liu, T.; Jiang, C.; Wang, Y.; Ma, J. Synthesis of glycidol and glycerol carbonate from glycerol and dimethyl carbonate using deep-eutectic solvent as a catalyst. Chem. Eng. J. 2022, 442, 136196. [Google Scholar] [CrossRef]
  58. Ayastuy, J.L.; Gutierrez-Ortiz, M.A.; Gonzalez-Marcos, J.A.; Aranzabal, A.; Gonzalez-Velasco, J.R. Kinetics of the low-temperature WGS reaction over a CuO/ZnO/Al2O3 catalyst. Ind. Eng. Chem. Res. 2005, 44, 41–50. [Google Scholar] [CrossRef]
Scheme 1. The reaction of GL and urea to produce GC.
Scheme 1. The reaction of GL and urea to produce GC.
Catalysts 14 00041 sch001
Figure 1. Effect of reaction temperature on the reaction of GL and urea using ZnSO4 as the catalyst (reaction conditions: reaction time: 4 h; catalyst amount: 5 wt% (based on GL weight); urea-to-GL molar ratio: 1.1:1; reaction pressure: 5 kPa).
Figure 1. Effect of reaction temperature on the reaction of GL and urea using ZnSO4 as the catalyst (reaction conditions: reaction time: 4 h; catalyst amount: 5 wt% (based on GL weight); urea-to-GL molar ratio: 1.1:1; reaction pressure: 5 kPa).
Catalysts 14 00041 g001
Figure 2. Effect of GL-to-urea molar ratio on the reaction of GL and urea using ZnSO4 as the catalyst (reaction conditions: reaction temperature: 140 °C; reaction time: 4 h; catalyst amount: 5 wt% (based on GL weight); reaction pressure: 5 kPa).
Figure 2. Effect of GL-to-urea molar ratio on the reaction of GL and urea using ZnSO4 as the catalyst (reaction conditions: reaction temperature: 140 °C; reaction time: 4 h; catalyst amount: 5 wt% (based on GL weight); reaction pressure: 5 kPa).
Catalysts 14 00041 g002
Figure 3. Effect of catalyst amount on the reaction of GL and urea using ZnSO4 as the catalyst (reaction conditions: reaction temperature: 140 °C; reaction time: 4 h; urea-to-GL molar ratio: 1.1:1; reaction pressure: 5 kPa).
Figure 3. Effect of catalyst amount on the reaction of GL and urea using ZnSO4 as the catalyst (reaction conditions: reaction temperature: 140 °C; reaction time: 4 h; urea-to-GL molar ratio: 1.1:1; reaction pressure: 5 kPa).
Catalysts 14 00041 g003
Figure 4. Effect of reaction time on the reaction of GL and urea using ZnSO4 as the catalyst (reaction condition: reaction temperature: 140 °C; catalyst amount: 5 wt% (based on GL weight); urea-to-GL molar ratio: 1.1:1; reaction pressure: 5 kPa).
Figure 4. Effect of reaction time on the reaction of GL and urea using ZnSO4 as the catalyst (reaction condition: reaction temperature: 140 °C; catalyst amount: 5 wt% (based on GL weight); urea-to-GL molar ratio: 1.1:1; reaction pressure: 5 kPa).
Catalysts 14 00041 g004
Figure 5. (a) XRD spectra of fresh ZnSO4; (b) standard spectra of ZnSO4; (c) precipitated solid at 30 min; (d) standard spectra of Zn(NH3)2SO4; (e) standard spectra of (NH4)2SO4; (f) precipitated solid at 240 min; (g) Zn(C3H6O3) sample (obtained via the method in [49]).
Figure 5. (a) XRD spectra of fresh ZnSO4; (b) standard spectra of ZnSO4; (c) precipitated solid at 30 min; (d) standard spectra of Zn(NH3)2SO4; (e) standard spectra of (NH4)2SO4; (f) precipitated solid at 240 min; (g) Zn(C3H6O3) sample (obtained via the method in [49]).
Catalysts 14 00041 g005
Figure 6. ESI−MS spectra of reaction solution of GL and urea over ZnSO4 catalyst with different reaction times: (a) 60 min; (b) 180 min; (c) 300 min.
Figure 6. ESI−MS spectra of reaction solution of GL and urea over ZnSO4 catalyst with different reaction times: (a) 60 min; (b) 180 min; (c) 300 min.
Catalysts 14 00041 g006
Figure 7. FT−IR spectra of reaction solution of GL and urea over ZnSO4 catalyst with different reaction times: (a) 0 min; (b) 30 min; (c) 60 min; (d) 120 min; (e) 180 min; (f) 240 min; (g) 300 min.
Figure 7. FT−IR spectra of reaction solution of GL and urea over ZnSO4 catalyst with different reaction times: (a) 0 min; (b) 30 min; (c) 60 min; (d) 120 min; (e) 180 min; (f) 240 min; (g) 300 min.
Catalysts 14 00041 g007
Figure 8. FT-IR spectra of the mixture of GL and (NH4)2SO4 with different contact times: (a) pure GL; (b) fresh (NH4)2SO4; (c) GL + (NH4)2SO4, 30 min; (d) GL + (NH4)2SO4, 60 min; (e) GL + (NH4)2SO4, 120 min.
Figure 8. FT-IR spectra of the mixture of GL and (NH4)2SO4 with different contact times: (a) pure GL; (b) fresh (NH4)2SO4; (c) GL + (NH4)2SO4, 30 min; (d) GL + (NH4)2SO4, 60 min; (e) GL + (NH4)2SO4, 120 min.
Catalysts 14 00041 g008
Figure 9. FT−IR spectra of the mixture of urea and Zn(C3H6O3) with different contact times: (a) pure urea; (b) fresh Zn(C3H6O3); (c) urea + Zn(C3H6O3), 30 min; (d) urea + Zn(C3H6O3), 60 min; (e) urea + Zn(C3H6O3), 120 min.
Figure 9. FT−IR spectra of the mixture of urea and Zn(C3H6O3) with different contact times: (a) pure urea; (b) fresh Zn(C3H6O3); (c) urea + Zn(C3H6O3), 30 min; (d) urea + Zn(C3H6O3), 60 min; (e) urea + Zn(C3H6O3), 120 min.
Catalysts 14 00041 g009
Figure 10. 1H NMR analysis of (a) pure GL and (b) mixture of GL and (NH4)2SO4 with the contact time of 2 h. The deuterated dimethylsulfoxide (DMSO)-d6 was used as a solvent in NMR analysis.
Figure 10. 1H NMR analysis of (a) pure GL and (b) mixture of GL and (NH4)2SO4 with the contact time of 2 h. The deuterated dimethylsulfoxide (DMSO)-d6 was used as a solvent in NMR analysis.
Catalysts 14 00041 g010
Figure 11. 13C NMR analysis of (a) pure urea and (b) a mixture of urea and Zn(C3H6O3) with the contact time of 2 h. The deuterated dimethylsulfoxide (DMSO)-d6 was used as a solvent in NMR analysis.
Figure 11. 13C NMR analysis of (a) pure urea and (b) a mixture of urea and Zn(C3H6O3) with the contact time of 2 h. The deuterated dimethylsulfoxide (DMSO)-d6 was used as a solvent in NMR analysis.
Catalysts 14 00041 g011
Scheme 2. The proposed reaction mechanism of GL and urea using ZnSO4 as the catalyst (the dash line denotes the hydrogen-bond interaction).
Scheme 2. The proposed reaction mechanism of GL and urea using ZnSO4 as the catalyst (the dash line denotes the hydrogen-bond interaction).
Catalysts 14 00041 sch002
Figure 12. The comparison of experimental and calculated component concentrations at different temperatures for the reaction of GL with urea over a ZnSO4 catalyst: (a) 120 °C; (b) 130 °C; (c) 140 °C; (d) 150 °C. (The green line denotes the concentration of GCM).
Figure 12. The comparison of experimental and calculated component concentrations at different temperatures for the reaction of GL with urea over a ZnSO4 catalyst: (a) 120 °C; (b) 130 °C; (c) 140 °C; (d) 150 °C. (The green line denotes the concentration of GCM).
Catalysts 14 00041 g012aCatalysts 14 00041 g012b
Figure 13. The comparison of experimental and calculated volumes of reaction mixture at different temperatures for the reaction of GL with urea over a ZnSO4 catalyst.
Figure 13. The comparison of experimental and calculated volumes of reaction mixture at different temperatures for the reaction of GL with urea over a ZnSO4 catalyst.
Catalysts 14 00041 g013
Table 1. The activities of different catalysts for the reaction of GL and urea a.
Table 1. The activities of different catalysts for the reaction of GL and urea a.
No.CatalystXGL, % bYGC, % cSGC, % d
1-42.2432.2976.44
2KOHtracetracetrace
3KNO3tracetracetrace
4ZnO71.2064.7390.91
5ZnCl281.7573.8390.31
6ZnBr278.3369.9189.25
7ZnI274.4167.9591.32
8Zn(NO3)·6H2O68.1161.1989.84
9Zn3(PO4)267.1462.3092.79
10ZnSO480.3375.8194.37
a Reaction conditions: reaction temperature: 140 °C; reaction time: 4 h; catalyst amount: 5 wt% (based on GL weight); urea-to-GL molar ratio: 1.1:1; reaction pressure: 5 kPa; stirring rate: 600 rpm. b XGL: the conversion of GL. c YGC: the yield of GC. d SGC: the selectivity of GC.
Table 2. The catalytic activity of some catalysts for the reaction of GL and urea a.
Table 2. The catalytic activity of some catalysts for the reaction of GL and urea a.
No.CatalystXGL, % bYGC, % cSGC, % d
11Zn(C3H6O3)65.7860.8492.49
12(NH4)2SO447.0138.6982.30
13Zn(C3H6O3) + (NH4)2SO476.3871.6293.77
a Reaction conditions: reaction temperature: 140 °C; reaction time: 4 h; catalyst amount: 5 wt% (based on GL weight); urea-to-GL molar ratio: 1.1:1; reaction pressure: 5 kPa; stirring rate: 600 rpm. b XGL: the conversion of GL. c YGC: the yield of GC. d SGC: the selectivity of GC.
Table 3. The reaction rate constant and activation energy for the reaction of GL with urea using ZnSO4 as the catalyst.
Table 3. The reaction rate constant and activation energy for the reaction of GL with urea using ZnSO4 as the catalyst.
ki aInchEa, kJ/mol bR2 c
k 1 = exp ( 35.297 17247 / T ) L·mol−1·min−1143.390.9995
k 2 = exp ( 29.327 10499 / T ) min−187.290.9995
a the reaction rate constant. b the activation energy. c the correlation coefficients.
Table 4. The activation energy for the reaction of GL with urea over other catalysts.
Table 4. The activation energy for the reaction of GL with urea over other catalysts.
No.Cat.Temperature, °CPressure, kPaEa, kJ/mol aRef.
1MgO135~150101.3 b117.85[38]
2Co3O4/ZnO100~160101.331.89[56]
a the activation energy. b The reaction was carried out under nitrogen flow at a flow rate of 100 mL/min.
Table 5. The statistical test results for the kinetics model.
Table 5. The statistical test results for the kinetics model.
CatalystF10 × FT
ZnSO4216.3723.6
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

Wang, H.; Ma, J. Reaction Kinetics and Mechanism for the Synthesis of Glycerol Carbonate from Glycerol and Urea Using ZnSO4 as a Catalyst. Catalysts 2024, 14, 41. https://doi.org/10.3390/catal14010041

AMA Style

Wang H, Ma J. Reaction Kinetics and Mechanism for the Synthesis of Glycerol Carbonate from Glycerol and Urea Using ZnSO4 as a Catalyst. Catalysts. 2024; 14(1):41. https://doi.org/10.3390/catal14010041

Chicago/Turabian Style

Wang, Huajun, and Jingjing Ma. 2024. "Reaction Kinetics and Mechanism for the Synthesis of Glycerol Carbonate from Glycerol and Urea Using ZnSO4 as a Catalyst" Catalysts 14, no. 1: 41. https://doi.org/10.3390/catal14010041

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