Next Article in Journal
Tungsten Catalysts for Visible Light Driven Ofloxacin Photocatalytic Degradation and Hydrogen Production
Next Article in Special Issue
Copper Catalyzed Inverse Electron Demand [4+2] Cycloaddition for the Synthesis of Oxazines
Previous Article in Journal
Natural Sunlight Driven Photocatalytic Removal of Toxic Textile Dyes in Water Using B-Doped ZnO/TiO2 Nanocomposites
Previous Article in Special Issue
Selective Synthesis of 2-(1,2,3-Triazoyl) Quinazolinones through Copper-Catalyzed Multicomponent Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthesis of Propylene Carbonate by Urea Alcoholysis—Recent Advances

1
Łukasiewicz Research Network—Institute of Heavy Organic Synthesis “Blachownia”, Energetyków 9, 47-225 Kędzierzyn-Koźle, Poland
2
Department of Chemical Organic Technology and Petrochemistry, PhD School, Silesian University of Technology, Akademicka 2A, 44-100 Gliwice, Poland
3
Department of Chemical Organic Technology and Petrochemistry, Faculty of Chemistry, Silesian University of Technology, Krzywoustego 4, 44-100 Gliwice, Poland
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(3), 309; https://doi.org/10.3390/catal12030309
Submission received: 3 February 2022 / Revised: 4 March 2022 / Accepted: 7 March 2022 / Published: 9 March 2022
(This article belongs to the Special Issue Catalysis in Green Chemistry and Organic Synthesis)

Abstract

:
Organic carbonates are considered the chemicals of the future. In particular, propylene carbonate is widely used as a non-reactive solvent, plasticizer, fuel additive, and reagent, especially in the production of environmentally friendly polymers that are not harmful to human health. This paper reviews recent literature findings regarding the development of propylene carbonate synthetic methods starting from propane-1,2-diol and urea. The ammonia formed during the synthesis is recycled to obtain urea from carbon dioxide.

Graphical Abstract

1. Introduction

In recent decades, increasing attention has been paid to environmental protection practices. The progression of industrialization, overexploitation of earth’s resources, and continuous interference with the natural environment, which gained momentum during the nineteenth century, led to many detrimental phenomena, including global warming-induced climate change, a gaping hole in the ozone layer, and vast pollution of ecosystems. One of the greatest challenges facing mankind in the 21st century is reducing the concentration of carbon dioxide (CO2) in the atmosphere. The emission of this greenhouse gas due to human activity in 2019 amounted to 33.4 Gt, and in 2020 it was “only” 31.5 Gt; the main reason for the decline in emissions was the outbreak of the COVID-19 pandemic [1]. It is urgently necessary to reduce CO2 emissions by replacing conventional power plants with nuclear reactors, solar panels, and/or wind turbines, and at the same time, it is imperative to protect the forests that serve as the lungs of our planet.
One important tool in the fight against global warming is the reuse of CO2 emitted by humanity as a raw material for the synthesis of chemical compounds. Carbon dioxide as a renewable source of elemental carbon and a safe, cheap, and readily-available compound that can be an alternative to other C1 building blocks, e.g., phosgene [2,3,4,5,6,7,8,9], carbon monoxide (CO) [10,11,12,13], methane [14,15,16,17], or methanol [18,19,20,21]. Unfortunately, the disadvantages of CO2 lie in its low reactivity and enthalpy of formation (ΔH0 = −394 kJ/mol), which strongly limit the use of carbon dioxide as a substrate [22,23]. Urea, which can be obtained on an industrial scale from carbon dioxide and ammonia [24], is a more reactive compound, and is also safe and environmentally friendly.
Currently, the scientific community is striving to introduce ecological products and technologies to replace currently used strategies and goods that have a negative impact on the natural environment and humanity. Highly toxic compounds are still used in the chemical industry [25], and one particularly difficult task is the disposal of spent volatile organic compounds (VOCs) [26]; these compounds must be replaced in order to protect the environment.
Organic carbonates represent possible alternative to VOCs [27] because they are ecological compounds with a wide range of potential applications. These compounds can be used to synthesize important products, including polymers [28,29] and surfactants [30,31], whereas organic carbonates themselves can be used as plasticizers [32,33] and fuel additives [34,35]. One of the most widely used organic carbonates is propylene carbonate (PC), which is a biodegradable and non-toxic organic solvent. It has a high boiling and ignition point, and other selected properties of PC are presented in Table 1.
Propylene carbonate has many applications. For example, it is a component of cosmetics and body care products [39], it can be a solvent for electrolytes in lithium-ion batteries [40,41], and it is a component of fusible polyurethane compositions used as binding agents in the footwear industry [42]. Moreover, PC improves the properties of polyurethanes obtained by the reaction injection molding (RIM) method [43]. It is often used as a hardener in the production of sand silicate forms [44], and it can also be used as a component that extracts metals from acidic solutions [45]. In addition, PC is present in antifreeze agents [46], and it can also be used in brake fluids owing to its low hygroscopicity and low-temperature viscosity [47]. Importantly, PC is also a substrate for the synthesis of other organic carbonates, such as dimethyl carbonate (DMC) [48].
On an industrial scale, propylene carbonate is obtained mainly through the carboxylation of propylene oxide (PO) (Scheme 1) [49]. This method is characterized by 100% atom economy. Moreover, the carboxylation method uses carbon dioxide which is both a renewable source of C1 and an abundant greenhouse gas. On the other hand, the second substrate, propylene oxide, is a highly reactive and unstable chemical compound, which means that the carboxylation method requires the integration of a propylene oxide synthesis unit within the propylene carbonate plant.
The phosgenation of propylene glycol (PG) was widely used for the synthesis of organic carbonates [4]. However, because of the toxicity of phosgene and the hydrogen chloride formed during the process, which corrodes the equipment, this method has lost its importance in favor of the carboxylation of propylene oxide.
A variety of new non-phosgene-based methods have been developed for the synthesis of organic carbonates, including propylene carbonate. One of them is the method based on the reaction of diols and carbon dioxide [50,51]. Although the raw materials are cheap and readily available, the thermodynamics of this process is unfavorable, and the water formed as a by-product deactivates the catalyst. An interesting method for the preparation of propylene carbonate is the oxidative carboxylation of olefins in which propylene, carbon dioxide and an oxidant, e.g., oxygen, are used as substrates [52,53]. During the process, propylene oxide is formed as an intermediate that does not need to be isolated. However, the problem is the difficulty in achieving high selectivity and yield of propylene carbonate.
Other routes to obtain propylene carbonate include the reaction between a halohydrin and carbon dioxide [54,55] or propan-1,2-diol and dimethyl carbonate [56,57].
Researchers have devoted significant efforts to the preparation of PC via urea alcoholysis owing to the inexpensive, stable, and readily-available nature of the required raw materials (e.g., urea and PG) and the favorable thermodynamics of the reaction. Moreover, ammonia, which is a co-product in the synthesis of propylene carbonate, can be returned to the urea plant by reacting it with CO2. Cyclic alkylene carbonates were obtained through the urea alcoholysis method for the first time in 1991 by Su and Speranza [58] who used a tin salt (i.e., dibutyltin dilaurate (Bu2Sn(C11H23COO)2) as a catalyst.
To our knowledge, there is no review covering the recent advancements in the synthesis of propylene carbonate by urea alcoholysis in the scientific literature; a review on this topic can be found in the Chinese scientific literature, but it is more than ten years old [59]. Shukla and Srivastava [60] carried out an extensive literature review highlighting the synthesis of selected organic carbonates (e.g., dimethyl, diethyl, glycerin, ethylene, and propylene) by urea alcoholysis.
This article focuses on research findings related to the synthesis of propylene carbonate from urea and propylene glycol. Herein, we analyze the last three decades of progress regarding catalytic systems and the prospects for applying urea alcoholysis for the large-scale synthesis of propylene carbonate.

2. Types of Catalysts Used for Propylene Carbonate Synthesis by Urea Alcoholysis

2.1. Metal Oxides

Most scientific reports focusing on the production of propylene carbonate from urea and PG refer synthetic methods involving metal oxide catalysts, which are often used as catalysts owing to their simple synthesis, availability, and various properties [61].
Zinc-based oxides are the most common catalysts used for PC synthesis from propylene glycol and urea [62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89]. ZnO has favorable acid-base properties, which make it an active and selective catalyst in the synthesis of PC. The proposed mechanism of the ZnO-catalyzed reaction is shown in Scheme 2 [85].
The urea molecule is stable because of its resonance structures [90]. The zinc atom activates the urea by coordinating with the oxygen atom of its carbonyl group. Electron transfer leads to the formation of zinc isocyanate complexes, which are soluble in the reaction mixture and react further with the propylene glycol molecule. This results in the generation of an intermediate product (i.e., 2-hydroxypropyl carbamate (HPC)), which cyclizes to form propylene carbonate. The cyclization of HPC is also catalyzed by ZnO through the formation of a coordination bond with the carbonyl oxygen of HPC.
Li et al. [85] synthesized propylene carbonate from urea and PG using various metal oxides and zinc sulfide (Table 2). The best results were obtained with ZnO (yield = 98.9%). Furthermore, researchers determined the acidity and basicity of the catalysts based on temperature-programmed desorption (TPD) tests (Table 3). These measurements indicated that the excellent results of the zinc oxide-catalyzed synthesis were due to low acidity and basicity. According to the authors, the acidic nature of certain catalysts (e.g., Al2O3, ZrO2, ZnS) led to low catalytic activity, which resulted in large amounts of the intermediate product, HPC, in the post-reaction mixture. In contrast, catalysts with low acidity (e.g., ZnO, MgO, CaO, La2O3) were active, and as a result, the HPC and urea contents in the product were low. This work also revealed that catalysts dominated by basic character (i.e., A/B < 1) were less selective. After using CaO, MgO, or La2O3 as the catalyst, 4-methyl-2-oxazolidone (MOD) was present in the post-reaction mixture. The authors explained that the strongly basic catalyst active sites tended to interact with the hydrogen of the HPC hydroxyl group, which increased the electrophilicity of the carbon atom bound to the hydroxyl group. Subsequently, HPC cyclization, followed by the elimination of water when the nitrogen atom’s electron pair attacked the carbon bound to the hydroxyl group, resulted in the formation of the MOD molecule (Scheme 3).
Table 4 presents the results of the ZnO-catalyzed synthesis of propylene carbonate, where the reactions were carried out with an excess of PG relative to urea. This was due to the minimization of the formation of nitrogen-containing byproducts, e.g., the biuret reagent and its homologues, and MOD. These reactions were carried out for 2–5 h. Zhao et al. [88] found that extending the reaction time reduced the yield of PC synthesis resulting from its polymerization. The syntheses of propylene carbonate were carried out at temperatures between 138–170 °C; higher reaction temperatures favored both PC polymerization [89] and MOD formation [78].
The problem with dissolving the zinc (ZnO) catalyst in the reaction mixture was noticed by Hao et al. [77], who reported that during the urea alcoholysis process, a Zn(NCO)2(NH3)2 complex soluble in the reaction mixture was formed. Losses resulting from catalyst dissolution prompted scientists to conduct research on the effective immobilization of active ingredients on appropriate carriers.
Yu et al. [84] synthesized PC using ZnO immobilized on NaY zeolite (ZnO/NaY). The process was carried out in a tubular reactor with a fixed catalyst bed. The catalyst was obtained by impregnating the support in an aqueous solution of Zn(NO3)2. The precatalyst obtained after the impregnation process was then dried and calcined in air for 3 h at 500 °C. The obtained catalyst was highly active and selective at 150 °C, achieving a PC yield of 82.3%, corresponding to a urea conversion of 86.2%. A four-fold molar excess of PG relative to urea was used for this reaction, and the catalyst was stable under the applied conditions; there was no significant decrease in the yield of PC after 35 h of running the synthetic process (the yield remained >80%). The authors determined that the properties and activity of the catalyst depended strongly on the amount of immobilized ZnO. The best results were obtained when ZnO accounted for 5 wt.% of the catalyst, whereas higher or lower ZnO contents resulted in lower catalytic activity and lower selectivity. This study also showed that ZnO/NaY containing 5 wt.% ZnO exhibited the most balanced strength of the acidic and basic sites, which was crucial for high PC production performance.
Researchers have investigated numerous catalytic systems comprising ZnO and other oxides (Table 5). Zhang et al. [81] used a catalyst composed of zinc and magnesium oxides (ZnO-MgO), which was prepared by co-precipitation from an aqueous solution of zinc and magnesium nitrates using urea as a precipitating agent [91]. The obtained precipitate was then washed and calcined at 600 °C for 6 h. Experimental results showed that the increased basicity and specific surface area positively impacted the final reaction yield. The authors evaluated the catalytic properties of systems containing various ZnO:MgO molar ratios. For catalysts with a molar ratio in the range from 4:1 to 1:4, an increase in specific surface areas (from 3.49 m2/g to 15.06 m2/g, respectively) was observed along with an increase in MgO content. Higher content of magnesium caused a reduction in ZnO coalescence. The highest yield of PC (94.8%) was obtained when the ratio was 1:4. However, a further increase in the MgO content (Zn/Mg 0.1) led to a lower area (3.83 m2/g) and lower PC yield (80.4%) as an aggregation of the crystals was observed. According to the authors, the higher catalytic activity exhibited by the two-component oxide system could be attributed to the modification of the catalysts’ electronic properties and porosities. Experimental results showed that the increased basicity and specific surface area positively impacted the final reaction yield.
Liu et al. [79] investigated a three-component catalytic system comprising zinc, calcium, and aluminum oxides (i.e., ZnO-CaO-Al2O3). The CaO-Al2O3 system was obtained by mixing powdered CaO and γ-Al2O3 in a 10% HNO3 solution. The resulting paste was then extruded and dried, and the extrudates were calcined at 540 °C for 4 h. The obtained CaO-Al2O3 was then impregnated in an ethanolic Zn(NO3)2 solution. After the impregnation process, the precatalyst was dried and calcined at 540 °C for 4 h. The authors evaluated the influence of the catalyst composition on the efficiency of the process. Their results confirmed that the catalyst should embody balanced acid-base properties. Based on catalyst composition screening tests, they determined that the mass percentage of CaO should be 30% (in relation to γ-Al2O3), and the mass percentage of ZnO should be 20% (in relation to the CaO-Al2O3 system). The highest reaction yield (90.8%) was obtained in the synthesis of PC over 3 h at 184 °C using a three-fold molar excess of PG relative to urea. The authors reported that it was not possible to obtain a higher yield for propylene carbonate synthesis owing to the presence of CaO, the basic nature of which caused HPC dehydration to form MOD, as well as PC decomposition.
Zinc-free oxide catalysts have also been described in the literature (Table 6) [66,73,74,75,76,85,88]. Many studies in this group describe catalytic systems containing magnesium oxide (MgO), which has specific properties that make it an active catalyst for PC synthesis via urea alcoholysis; however, this compound can promote the formation of MOD [85]. Lead-based catalysts also have favorable catalytic properties. For example, high selectivity for PC synthesis (93.9%) was observed when using lead dioxide (PbO2) as a catalyst [76]. Unfortunately, the significant disadvantage of such catalysts lies in their toxicity. Other oxides, such as Al2O3, ZrO2 [85], or dubityl tin oxide (Bu2SnO) [88] were not active catalysts for this synthesis.
Synthetic compounds with perovskite structures represent a special type of oxides that have been tested for their activity as catalysts for the synthesis of propylene carbonate by urea alcoholysis (Table 7) [93,94]. These oxides have the general formula, ABO3, where A is a large-diameter metal cation, and B is a small-diameter metal cation [95]. Du et al. [94] investigated how the MgTiO3 preparation method influenced its catalytic properties in terms of PC synthesis. The catalyst was prepared using a sol-gel method; specifically, a solution of magnesium nitrate (Mg(NO3)2) was added dropwise to a solution of tetrabutoxy titanium (IV) (Ti(OBu)4) in acetic acid. Anhydrous ethanol was added to this mixture, which was then concentrated in a vacuum evaporator, and the resulting gel was calcined at various temperatures (600, 700, 800, or 900 °C). The MgTiO3 catalyst with the highest activity was obtained when the Mg:Ti molar ratio was 1:1 and the gel calcination was carried out at 700 °C for 3 h. Under these conditions, MgTiO3 was obtained with the highest quantity of basic centers, which resulted in an increase in catalytic activity and selectivity. However, it was also shown that MgO aggregated on the catalyst surface when Ti(OBu)4 and Mg(NO3)2 were used in a 1:1 molar ratio during catalyst preparation. As a result, after the fourth catalytic cycle, the yield of propylene carbonate decreased from 93.5% to approx. 60% because of the leaching of MgO from the catalyst surface. XRD tests showed that the intensity of the MgO diffraction peak (2Φ = 42.9°) decreased with a decrease in the molar ratio of Mg:Ti, whereas for the MTO-0.8 catalyst (with a molar ratio of Mg:Ti = 0.8:1) no peaks corresponding to MgO were detected. Researchers also showed that in the case of the MTO-1.5 catalyst (Mg:Ti = 1.5:1), the peaks corresponding to MgTiO3 were shifted to the right, which may be due to distortion of the structure.
To date, there are few publications in the scientific literature describing the kinetics of PC synthesis from urea and PG. To implement this method into industrial practice, it is necessary to determine the kinetic equation corresponding to the reaction where propylene carbonate is obtained through urea alcoholysis.
Shi et al. [96] designed a two-step process for the preparation of DMC. The first stage of the process involved the synthesis of propylene carbonate via urea alcoholysis, and the second stage involved PC transesterification with methanol. The authors considered Wang’s previous findings [97], and on that basis, they proposed the kinetic Equation (1) for the synthesis of PC from urea and PG catalyzed by 2 wt.% MgO:
r PC = 1.5888 × exp ( 562.602 T ) × C urea × C PG
rPC—reaction rate of propylene carbonate synthesis, mol/(L·min)
T—reaction temperature, K
Curea—urea concentration, mol/L
CPG—concentration of propylene glycol in the reaction mixture, mol/L.
The authors also determined the total annual operating cost of the designed installation for several synthetic variants. On the basis of calculations, they showed that obtaining propylene carbonate by reactive distillation was more profitable than using a cascade of three continuous stirred-tank reactors (CSTRs). Analogous simulations were carried out by Patraşcu et al. [98] using the same kinetic equation for the PC synthesis reaction. Overall, studies have shown that the synthesis of DMC from propylene carbonate and methanol can proceed with excess PC, which facilitates the preparation of high-purity DMC.
In the patent literature, there are several examples of PC synthesis catalyzed by metal oxides in the CSTR cascade [64,70,99]. One innovative solution involves the use of a horizontal reactor with partitions increasing the turbulence of the flow and helping the reaction system imitate a cascade of reactors [67]. Another method of PC synthesis in a reactive column with a metal oxide catalyst was also patented [100]. The reports discussed herein suggest that the preparation of PC through urea alcoholysis catalyzed by metal oxides is relatively advanced.

2.2. Metal Salts

Metal salts have also been tested for their potential applications as catalysts for PC synthesis from urea and PG. As in the case of metal oxides, most of the salts used as catalysts are advantageous because they are inexpensive and readily available.
Among the salts used for this reaction, compounds of zinc [76,88,101,102], magnesium [76,102], lead [101], and tin [103] exhibited the best catalytic properties. Doya et al. [76] used various zinc salts (e.g., ZnCO3, ZnCl2, Zn(NO3)2, and Zn(OAc)2) and magnesium acetate (Mg(OAc)2) and compared their catalytic properties with those of Bu2Sn(C11H23COO)2, which was employed by Su and Speranza in the first published studies on cyclic alkylene carbonate synthesis [58]. The results showed that the zinc and magnesium salts had similar catalytic properties (PC selectivities ranged from 88.1% to 89.5%) and were more selective than Bu2Sn(C11H23COO)2, which only achieves up to 64.6% selectivity for PC synthesis. In the patent, Sun et al. [103] presented examples of PC synthesis using various salts, among which the highest yield (93.47%) was obtained for tin(II) chloride; however, it should be noted that the reactions with other catalysts were carried out under conditions that may not be optimal for the desired process.
The highest yield of PC (96.5%) was achieved by Gao et al. [102] using magnesium chloride (MgCl2) as a catalyst. Zhou et al. [101] obtained a 93.5% yield using a mixture of basic zinc carbonate and lead carbonate prepared using a co-precipitation method. According to the authors, the most active catalytic system contained a 1:2 molar ratio of PbCO3:Zn5(CO3)2(OH)6.
Zhao et al. [88] conducted research on the selection of an active catalyst for the synthesis of propylene carbonate. Among the investigated compounds, anhydrous zinc acetate (Zn(OAc)2) exhibited the best activity; therefore, further studies were carried out to evaluate the influence of certain parameters on the efficiency of the process. Under the optimized conditions (mole ratio of urea:PG = 1:4; reaction time = 3 h; temperature = 170 °C), a 94% yield of PC was obtained. The authors also studied the effect of Zn(OAc)2 immobilization on various supports, e.g., activated carbon (AC), zeolite (zeo), and γ-alumina (γ-Al2O3). The Zn(OAc)2 was immobilized on these supports by an incipient impregnation process. The catalytic activity of the catalyst was particularly influenced by the physical properties of the support. The best results were achieved for catalysts immobilized on AC, which showed the highest specific surface among the tested carriers, and on γ-Al2O3 with the highest average pore size. Using AC and γ-Al2O3 as the carrier, the yields of PC synthesis were 77.5 and 77.9%, respectively. Unfortunately, the applied immobilization method was not effective because Zn(OAc)2 easily leached from the support. According to the authors, the fresh Zn(OAc)2/AC catalyst contained 3 wt.% zinc, and after PC synthesis, the Zn content dropped to 2.3 wt.%. After the zinc acetate was washed off of the catalyst, the yield of PC was only 66% once the catalyst was recycled.
Additional studies on the synthesis of propylene carbonate have used potassium silicate [104] and zinc sulfide as catalysts [85], but these compounds were characterized by low catalytic activity. The relevant patent [105] describes a method for obtaining organic carbonates by urea alcoholysis using silicates as catalysts. Table 8 shows the results of studies probing the catalytic activity of these salts in the synthesis of propylene carbonate.
Polyoxometalates represent a special type of metal salts. Their catalytic activity for the synthesis of propylene carbonate was patented by Liu et al. [106]. The catalysts were obtained by mixing a heteropolyacid with a suitable metal carbonate in water, and then evaporating the water and calcining the precipitated precatalyst at 200–400 °C. The authors emphasized that the separation of the precatalyst from the aqueous mixture was also possible via partial evaporation of the water, cooling to 0–5 °C, and centrifuging to isolate the precipitate. Phosphomolybdic acid (H3PMo12O40), phosphotungstic acid (H3PW12O40), or silicon tungsten acid (H4SiW12O40) could be applied with an appropriate carbonate or hydroxycarbonate (e.g., ZnCO3, MgCO3, 4MgCO3 × Mg(OH)2, CaCO3, K2CO3) to synthesize these catalysts. The catalytic activities of heteropolyoxometalates are presented in Table 9. In general, these heteropolyoxometalates do not have high catalytic activity; the highest PC yield (78.71%) was obtained using Zn3(PMo12O40)2.

2.3. Modified Hydroxyapatites

Modified hydroxyapatites are another group of chemical compounds that can be used as catalysts for the preparation of propylene carbonate via urea alcoholysis. Owing to their basic nature, these compounds can also be used as catalysts in the transesterification process [107] or in the Michael addition reaction [108].
Du et al. [109] developed a method for obtaining propylene carbonate and ethylene carbonate through urea alcoholysis using metal-modified hydroxyapatites as catalysts, and the catalytic activities of these systems were compared with those of the corresponding metal oxides [110]. Hydroxyapatite (Ca10(PO4)6(OH)2; HAP) was obtained by co-precipitation from a solution containing ammonium hydrogen phosphate ((NH4)2HPO4) and calcium nitrate (Ca(NO3)2). The resulting precipitate was then filtered, dried, and calcined at 900 °C. The obtained HAP was impregnated in an aqueous solution of the metal nitrate, then dried and calcined at 400 °C. The authors explained that during HAP impregnation, ion exchange occurred, which involved the replacement of Ca2+ cations in the crystal structure of the hydroxyapatite with ions of a given metal. Lanthanum-modified HAP (La-HAP) demonstrated the best catalytic properties among the evaluated catalysts. The yield of PC catalyzed by La-HAP was 91.5%; for comparison, the yield was lower (79.6%) when using the corresponding metal oxide, La2O3. The BET surface analysis showed that the HAP surface area after the impregnation process in the La3(NO3)3 solution decreased from 9.0 m2/g to 7.3 m2/g. The researchers explained that this was due to the amorphous lanthanum blocking the pores The authors reported that the high activity of the La-HAP catalyst was due to the presence of strongly basic adsorption centers. XRD analysis showed that the diffraction peaks of HAP and La/HAP were similar, indicating that the major phase of La/HAP was HAP. As the authors explain, some La3+ ions were introduced into the HAP crystal lattice by ion exchange with Ca2+ ions, but the HAP structure was preserved. In the case of La3+, which were not exchanged in the HAP crystal lattice, they were strongly dispersed on the HAP surface.
Table 10 summarizes the catalytic activities of selected modified hydroxyapatites. These catalysts are characterized by high selectivity (over 90%), which is undoubtedly their great advantage.

2.4. Ionic Liquids

Ionic liquids (ILs) are chemical compounds composed of ions with a melting point lower than 100 °C [111]. Gabriel and Weiner (1888) were the first to document the synthesis of an ionic liquid, specifically, ethanol-ammonium nitrate (mp = 52–55 °C) [112], although Walden is often considered the father of ILs, following the reported synthesis of nitrate ethylammonium liquid at room temperature (mp = 12 °C) in 1914 [113]. However, scientific interest in ionic liquids increased significantly around the beginning of the 21st century and has continued to this day [114].
Because ionic liquids can be designed by selecting the appropriate cations and anions, it is possible to obtain a chemical compound with the desired properties. This advantage means that ionic liquids can be used in many fields of science; however, only some of them are currently produced on an industrial scale [115]. In general, the low availability of starting materials limits their production possibilities. The low cost of key reagents and abundance of their manufacturers are essential for a given ionic liquid to be produced on a commercial scale.
The scientific literature contains thousands of publications discussing the use of ionic liquids as catalysts in chemical synthesis. However, to date, few scientific reports have detailed the synthesis of propylene carbonate by ionic liquid-catalyzed urea alcoholysis. Kuznetsov et al. [116] synthesized cyclic alkylene carbonates using a two-phase reaction system comprising a deep eutectic solvent (DES) [117,118] and a chlorinated organic solvent. The DES was an equimolar mixture of zinc chloride, urea, and glycol, which formed a transparent colorless solution when heated to above 50 °C. The synthesis of propylene carbonate was carried out at low temperature (84 °C) and required a long reaction time (24 h) to ultimately reach a maximum yield of 70%.
A patent from 2016 [119] describes a method for obtaining PC through a reactive distillation method, in which the catalytic system consisted of zinc oxide, a zinc salt, and a quaternary ammonium salt. Another patent [120] details the synthesis of cyclic alkylene carbonates using a catalytic system containing a metal salt and an ionic liquid with an imidazolium cation. The preparation of PC by urea alcoholysis at 160 °C for 3 h under reduced pressure (150 kPa) using a catalytic system containing 1-hexadecyl-3-methylimidazolium chloride and zinc chloride ([C16mim]Cl-ZnCl2) led to a 94.1% yield of PC. The catalytic system did not lose significant activity in this case; after its 5th cycle, the PC yield was 90.1%. The proposed mechanism of the reaction catalyzed by this system (Scheme 4) [121] assumes an increase in urea reactivity due to (i) the interaction between the oxygen of the urea carbonyl group and Zn2+ and (ii) the formation of a hydrogen bond between this oxygen and the hydrogen of the imidazole ring at the C(2) position.
Owing to the relative novelty of PC synthesis via urea alcoholysis using ionic liquid catalysts, it can be expected that there will be more publications on this subject in the coming years.

2.5. Others

Apart from metal oxides, metal salts, modified hydroxyapatites, and ionic liquids, the catalytic properties of other substances have also been tested in the context of PC synthesis by urea alcoholysis (Table 11). In particular, several reports have described the synthesis of PC catalyzed by pure metals, such as zinc and lead. These studies verified that metallic zinc had inferior catalytic properties relative to ZnO [77,89], whereas metallic lead and PbO achieved comparable PC yields [88].
Kulasegaram et al. [122] used zinc monoglycerolate as a catalyst in the reactions of urea with various glycols. When using vicinal diols (e.g., propylene glycol, ethylene glycol), the corresponding cyclic alkylene carbonate was the main product of the reaction. In contrast, when using terminal diols as the substrate (e.g., 1,3-propanediol, 1,4-butanediol), the corresponding dicarbamates were obtained. These syntheses were carried out with an excess of urea at 140 °C. Propylene carbonate was obtained from urea and PG with 80% selectivity and almost complete conversion (99%) of PG.
Indran et al. [104] used boiler ash as a catalyst for the synthesis of cyclic alkylene carbonates. The ash was obtained from the combustion of palm fruits, kernels, shells, and fibers. Raw ash was prepared via calcination at 900 °C to obtain the “BA 900” catalyst, which contained mainly K2SiO3 (43.53 wt.%) and SiO2 (33.14 wt.%), as well as small amounts of potassium, calcium, magnesium, aluminum, iron, phosphorus, and zirconium oxides. Upon using BA 900 as a catalyst for the synthesis of PC, the yield of the reaction was 73.2%.
Xiao et al. [123] used phosphate tailings as the raw materials to prepare a catalyst for the synthesis of propylene carbonate from urea and propylene glycol. The phosphate tailings were ground and washed with deionized water, and then, the precipitate was dried and calcined at 800 °C for 4 h to obtain the PT-800-4 catalyst. This compound achieved a yield and selectivity for PC synthesis via urea alcoholysis of 81.8% and 95.6%, respectively.

3. Conclusions

This review presents recent advances regarding the synthesis of propylene carbonate by urea alcoholysis, highlighting key findings. Numerous catalysts applied in this reaction are discussed in terms of the important aspects of their preparation and potential reaction mechanisms. This review confirms that a good catalyst for the preparation of PC from urea and PG should have balanced acid-base properties. Catalysts with numerous acidic sites are less catalytically active, whereas overly basic catalysts are less selective. The catalyst of greatest interest is zinc oxide, including catalytic systems composed of ZnO and other metal oxides.
Owing to its physicochemical properties, non-toxicity, and biodegradability, propylene carbonate is an extremely desirable chemical compound, whose demand on the global market is growing. The search for new, more ecological PC production methods is therefore extremely important for increasing its global production. The main method currently used to obtain propylene carbonate, i.e., carboxylation of propylene oxide, requires permanent access to epoxide; hence, its implementation is limited to petrochemical plants that have this raw material on site.
The preparation of PC by urea alcoholysis is a sustainable, economical and convenient alternative to the currently used method on the industrial scale. Compared with the synthesis of propylene carbonate by propylene oxide carboxylation, urea alcoholysis offers the use of inexpensive, stable and readily available raw materials (urea and propylene glycol) and favorable thermodynamics of the reaction. In addition, ammonia, which is a co-product in the synthesis of propylene carbonate, can be recycled to the urea synthesis plant.
Obtaining propylene carbonate by urea alcoholysis represents an alternative approach for companies that do not produce propylene oxide. This method requires stable raw materials that can be safely imported. However, because ammonia is released during the synthesis, it is reasonable to develop chemical installations for urea alcoholysis-based PC synthesis in companies already containing a urea plant; this would allow for the recycling of NH3, thereby closing its circulation. The main beneficiaries of this solution may be chemical plants producing nitrogen fertilizers, which offer urea. However, for the urea alcoholysis method to be commonly used in the chemical industry, it is necessary to further understand the kinetics of this process, and unfortunately, there is currently not enough data on this subject.
Therefore, it seems that the urea alcoholysis method discussed herein will not replace the method of obtaining propylene carbonate by carboxylation of PO in the near future. As long as the petrochemical industry continues to develop and fossil resources, such as crude oil, natural gas, and coal are still used, the current method of producing PC will continue to dominate. Nevertheless, the method for obtaining propylene carbonate via urea alcoholysis may be an alternative for enterprises with a different raw material base.

Author Contributions

Conceptualization, Ł.K.; writing—original draft preparation, Ł.K.; writing—review and editing, Ł.K., A.S.; supervision, A.C., A.S. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by Łukasiewicz Research Network—Institute of Heavy Organic Synthesis “Blachownia”. This research was co-financed by the Ministry of Education and Science of Poland under grant No. DWD/4/55/2020-004.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Global Energy Review: CO2 Emissions in 2020. Available online: https://www.iea.org/articles/global-energy-review-co2-emissions-in-2020 (accessed on 23 March 2021).
  2. Babad, H.; Zeiler, A.G. Chemistry of phosgene. Chem. Rev. 1973, 73, 75–91. [Google Scholar] [CrossRef]
  3. Davies, A.G.; Hua-De, P.; Hawari, J.A.-A. The reaction of 1,3,2-dioxastannolans with diacyl chlorides: Decarbonylation in the reaction with oxalyl chloride. J. Organomet. Chem. 1983, 256, 251–260. [Google Scholar] [CrossRef]
  4. Van Vegten, C.H.; Ten Kate, A.J.B.; De Groot, M.T. Process for the Preparation of Cyclic Organic Carbonates. WO 2013189861, 27 December 2013. [Google Scholar]
  5. Saunders, J.H.; Slocombe, R.J. The Chemistry of the Organic Isocyanates. Chem. Rev. 1948, 43, 203–218. [Google Scholar] [CrossRef]
  6. Arnold, R.G.; Nelson, J.A.; Verbanc, J.J. Recent Advances In Isocyanate Chemistry. Chem. Rev. 1957, 57, 47–76. [Google Scholar] [CrossRef]
  7. Tilley, J.N.; Sayigh, A.A.R. Pyrolysis of N-tert-N-alkylcarbamoyl chlorides. New synthesis of isocyanates. J. Org. Chem. 1963, 28, 2076–2079. [Google Scholar] [CrossRef]
  8. Matzner, M.; Kurkjy, R.P.; Cotter, R.J. The chemistry of chloroformates. Chem. Rev. 1964, 64, 645–687. [Google Scholar] [CrossRef]
  9. Jones, J.I. The reaction of carbonyl chloride with 1,2-epoxides. J. Chem. Soc. 1957, 0, 2735–2743. [Google Scholar] [CrossRef]
  10. Drent, E. Chemical Synthesis from C1 Compounds. In Microbial Growth on C1 Compounds; Verseveld, H.W., Duine, J.A., Eds.; Springer: Dordrecht, The Netherlands, 1987; pp. 272–281. [Google Scholar]
  11. Polo-Garzon, F.; Fung, V.; Nguyen, L.; Tang, Y.; Tao, F.; Cheng, Y.; Daemen, L.L.; Ramirez-Cuesta, A.J.; Foo, G.S.; Zhu, M.; et al. Elucidation of the Reaction Mechanism for High-Temperature Water Gas Shift over an Industrial-Type Copper–Chromium–Iron Oxide Catalyst. J. Am. Chem. Soc. 2019, 141, 7990–7999. [Google Scholar] [CrossRef]
  12. Pal, D.B.; Chand, R.; Upadhyay, S.N.; Mishra, P.K. Performance of water gas shift reaction catalysts: A review. Renew. Sustain. Energy Rev. 2018, 93, 549–565. [Google Scholar] [CrossRef]
  13. Li, P.; Zhu, M.; Tian, Z.; Han, Y.; Zhang, Y.; Zhou, T.; Kang, L.; Dan, J.; Guo, X.; Yu, F.; et al. Two-dimensional layered double hydroxide derived from vermiculite waste water supported highly dispersed Ni nanoparticles for CO methanation. Catalysts 2017, 7, 79. [Google Scholar] [CrossRef] [Green Version]
  14. Wang, L.; Tao, L.; Xie, M.; Xu, G.; Huang, J.; Xu, Y. Dehydrogenation and aromatization of methane under non-oxidizing conditions. Catal. Lett. 1993, 21, 35–41. [Google Scholar] [CrossRef]
  15. Li, G.; Vollmer, I.; Liu, C.; Gascon, J.; Pidko, E.A. Structure and Reactivity of the Mo/ZSM-5 Dehydroaromatization Catalyst: An Operando Computational Study. ACS Catal. 2019, 9, 8731–8737. [Google Scholar] [CrossRef] [Green Version]
  16. Keller, G.E.; Bhasin, M.M. Synthesis of ethylene via oxidative coupling of methane I. Determination of active catalysts. J. Catal. 1892, 73, 9–19. [Google Scholar]
  17. Zavyalova, U.; Holena, M.; Schlçgl, R.; Baerns, M. Statistical Analysis of Past Catalytic Data on Oxidative Methane Coupling for New Insights into the Composition of High-Performance Catalysts. ChemCatChem 2011, 3, 1935–1947. [Google Scholar] [CrossRef]
  18. Missengue, R.N.M.; Losch, P.; Musyoka, N.M.; Louis, B.; Pale, P.; Petrik, L.F. Conversion of South African coal fly ash into high-purity ZSM-5 zeolite without additional source of silica or alumina and its application as a methanol-to-olefins catalyst. Catalysts 2018, 8, 124. [Google Scholar] [CrossRef] [Green Version]
  19. Mitran, G.; Mieritz, D.G.; Seo, D.K. Highly selective solid acid catalyst H1-xTi2(PO4)3-x(SO4)x for non-oxidative dehydrogenation of methanol and ethanol. Catalysts 2017, 7, 95. [Google Scholar] [CrossRef] [Green Version]
  20. Chang, C.D.; Silvestri, A.J. The conversion of methanol and other O-compounds to hydrocarbons over zeolite catalysts. J. Catal. 1977, 47, 249–259. [Google Scholar] [CrossRef]
  21. Yarulina, I.; Chowdhury, A.D.; Meirer, F.; Weckhuysen, B.M.; Gascon, J. Recent trends and fundamental insights in the methanol-to-hydrocarbons process. Nat. Catal. 2018, 1, 398–411. [Google Scholar] [CrossRef]
  22. Burczyk, B. Carbon dioxide in organic synthesis. Wiad. Chem. 2013, 67, 1–53. [Google Scholar]
  23. North, M. What is CO2? Thermodynamics, Basic Reactions and Physical Chemistry. In Carbon Dioxide Utilisation; Styring, P., Quadrelli, E.A., Armstrong, K., Eds.; Elsevier: Amsterdam, The Netherlands, 2015; pp. 3–17. [Google Scholar]
  24. Krase, N.W.; Gaddy, V.L. Synthesis of urea from ammonia and carbon dioxide. Ind. Eng. Chem. 1922, 14, 611–615. [Google Scholar] [CrossRef]
  25. Chemical Casualties. Toxic industrial chemicals. J. R. Army Med. Corps. 2002, 148, 371–381. Available online: https://militaryhealth.bmj.com/content/148/4/371 (accessed on 24 March 2021). [CrossRef] [Green Version]
  26. Joshi, D.R.; Adhikari, N. An overview on common organic solvents and their toxicity. J. Pharm. Res. Int. 2019, 28, 1–18. [Google Scholar] [CrossRef]
  27. Shaikh, A.-A.G.; Sivaram, S. Organic Carbonates. Chem. Rev. 1996, 96, 951–976. [Google Scholar] [CrossRef]
  28. Rokicki, G.; Rakoczy, P.; Parzuchowski, P.; Sobiecki, M. Hyperbranched aliphatic polyethers obtained from environmentally benign monomer: Glycerol carbonate. Green Chem. 2005, 7, 529–539. [Google Scholar] [CrossRef]
  29. Komiya, K.; Fukuoka, S.; Kawamura, S. Process for Preparing Aromatic Polycarbonates. U.S. Patent 5840826, 24 November 1998. [Google Scholar]
  30. Banno, T.; Toyota, T.; Matsumura, S. Creation of Novel Green Surfactants Containing Carbonate Linkages, Intech Open. 2013. Available online: https://www.intechopen.com/books/biodegradation-life-of-science/creation-of-novel-green-surfactants-containing-carbonate-linkages (accessed on 24 March 2021).
  31. Stjerndahl, M.; Holmberg, K.J. Hydrolyzable nonionic surfactants: Stability and physicochemical properties of surfactants containing carbonate, ester, and amide bonds. Colloid Interface Sci. 2005, 291, 570–576. [Google Scholar] [CrossRef] [PubMed]
  32. Gohel, K.; Kanchan, D.K.; Machhi, H.K.; Soni, S.S.; Maheshwaran, C. Gel polymer electrolyte based on PVDF-HFP:PMMA incorporated with propylene carbonate (PC) and diethyl carbonate (DEC) plasticizers: Electrical, morphology, structural and electrochemical properties. Mater. Res. Express 2020, 7, 025301. [Google Scholar] [CrossRef]
  33. Mohamed, S.A.; Al-Ghamdi, A.A.; Sharma, G.D.; El Mansy, M.K. Effect of ethylene carbonate as a plasticizer on CuI/PVA nanocomposite: Structure, optical and electrical properties. J. Adv. Res. 2014, 5, 79–86. [Google Scholar] [CrossRef] [Green Version]
  34. Abdalla, A.O.G.; Liu, D. Dimethyl carbonate as a promising oxygenated fuel for combustion: A review. Energies 2018, 11, 1552. [Google Scholar] [CrossRef] [Green Version]
  35. Pacheco, M.A.; Marshall, C.L. Review of Dimethyl Carbonate (DMC) Manufacture and Its Characteristics as a Fuel Additive. Energy Fuels 1997, 11, 2–29. [Google Scholar] [CrossRef]
  36. Yu, Z.; Xu, T.; Xing, T.; Fan, L.-Z.; Lian, F.; Qiu, W. A Raman spectroscopy investigation of the interactions of LiBOB with γ-BL as electrolyte for advanced lithium batteries. J. Power Sources 2010, 195, 4285–4289. [Google Scholar] [CrossRef]
  37. Borodin, O.; Smith, G.D. Development of Many-Body Polarizable Force Fields for Li-Battery Components: 1. Ether, Alkane, and Carbonate-Based Solvents. J. Phys. Chem. B 2006, 110, 6279–6292. [Google Scholar] [CrossRef] [PubMed]
  38. Verevkin, S.P.; Emel’yanenko, V.N.; Toktonov, A.V.; Chernyak, Y.; Schäffner, B.; Börner, A. Cyclic alkylene carbonates. Experiment and first principle calculations for prediction of thermochemical properties. J. Chem. Thermodyn. 2008, 40, 1428–1432. [Google Scholar] [CrossRef]
  39. Sun, Y.; Liu, J.-C.; Kimbleton, E.S.; Wang, J.C.T. Composition Base for Topical Therapeutic and Cosmetic Preparations. U.S. Patent 5993787, 30 November 1999. [Google Scholar]
  40. Velázquez-Martínez, O.; Valio, J.; Santasalo-Aarnio, A.; Reuter, M.; Serna-Guerrero, R. A critical review of lithium-ion battery recycling processes from a circular economy perspective. Batteries 2019, 5, 68. [Google Scholar] [CrossRef] [Green Version]
  41. Ogumi, Z.; Jeong, S.-K. SPM analysis of surface film formation on graphite negative electrodes in lithium-ion batteries. Electrochemistry 2003, 71, 1011–1017. [Google Scholar] [CrossRef] [Green Version]
  42. Albert, K.-H.; von Voithenberg, H. Hot-Melt Polyurethane Compositions and Their Use in Bonding Shoes. EP 0381897, 21 December 1989. [Google Scholar]
  43. Nodelman, N.H. Manufacture of Flexible Polyoxyalkylene-Polyurea-Polyurethane Reaction-Injection Moldings from Compositions Having Improved Flow Properties. U.S. Patent 5149458, 22 September 1992. [Google Scholar]
  44. Stevenson, J.; Machin, J.; Dyke, D.L. Sand Reclamation. U.S. Patent 4416694, 22 November 1983. [Google Scholar]
  45. Stephens, B.G. Solvent Extraction of Metals with a Cyclic Alkylene Carbonate. U.S. Patent 3912801, 14 October 1975. [Google Scholar]
  46. Keil, H. Nonfoaming Antifreeze Composition. DE 2407123, 15 February 1974. [Google Scholar]
  47. Jousset, J. Automobile Brake Fluid. FR 1508261, 22 November 1966. [Google Scholar]
  48. Deng, W.; Shi, L.; Yao, J.; Zhang, Z. A review on transesterification of propylene carbonate and methanol for dimethyl carbonate synthesis. Carbon Resour. Convers. 2019, 2, 198–212. [Google Scholar] [CrossRef]
  49. Wagner, P.; Mendoza-Frohn, C.; Buysch, H.-J. Continuous, Adiabatic, Catalytic Process for the Manufacture of Propylene Glycol Carbonate from Propylene Oxide and Carbon Dioxide. U.S. Patent 5449791, 12 September 1995. [Google Scholar]
  50. Tomishige, K.; Yasuda, H.; Yoshida, Y.; Nurunnabi, M.; Li, B.; Kunimori, K. Novel Route to Propylene Carbonate: Selective Synthesis from Propylene Glycol and Carbon Dioxide. Catal. Lett. 2004, 95, 45–49. [Google Scholar] [CrossRef]
  51. Tomishige, K.; Yasuda, H.; Yoshida, Y.; Nurunnabi, M.; Li, B.; Kunimori, K. Catalytic performance and properties of ceria based catalysts for cyclic carbonate synthesis from glycol and carbon dioxide. Green Chem. 2004, 6, 206–214. [Google Scholar] [CrossRef]
  52. Wang, L.; Que, S.; Ding, Z.; Vessally, E. Oxidative carboxylation of olefins with CO2: Environmentally benign access to five-membered cyclic carbonates. RSC Adv. 2020, 10, 9103–9115. [Google Scholar] [CrossRef] [Green Version]
  53. Siewniak, A.; Jasiak-Jaroń, K.; Kotyrba, Ł.; Baj, S. Efficient Catalytic System Involving Molybdenyl Acetylacetonate and Immobilized Tributylammonium Chloride for the Direct Synthesis of Cyclic Carbonates from Carbon Dioxide and Olefins. Catal. Lett. 2017, 6, 1567–1573. [Google Scholar] [CrossRef] [Green Version]
  54. Khokarale, S.G.; Mikkola, J.-P. Metal free synthesis of ethylene and propylene carbonate from alkylene halohydrin and CO2 at room temperature. RSC Adv. 2019, 9, 34023–34031. [Google Scholar] [CrossRef]
  55. Hirose, T.M.; Shimizu, S.; Qu, S.; Shitara, H.; Kodama, K.; Wang, L. Economical synthesis of cyclic carbonates from carbon dioxide and halohydrins using K2CO3. RSC Adv. 2016, 6, 69040–69044. [Google Scholar] [CrossRef]
  56. Lopes, E.J.C.; Ribeiro, A.P.C.; Martins, L.M.D.R.S. New Trends in the Conversion of CO2 to Cyclic Carbonates. Catalysts 2020, 10, 479. [Google Scholar] [CrossRef]
  57. Selva, M.; Caretto, A.; Noe, M.; Perosa, A. Carbonate phosphonium salts as catalysts for the transesterification of dialkyl carbonates with diols. The competition between cyclic carbonates and linear dicarbonate products. Org. Biomol. Chem. 2014, 12, 4143–4156. [Google Scholar] [CrossRef] [PubMed]
  58. Su, W.-Y.; Speranza, G.P. Process for Preparing Alkylene Carbonates. U.S. Patent 5003084, 26 March 1991. [Google Scholar]
  59. Liu, L.; Du, Z.; Yuan, H.; Jin, F.; Wu, Y. Research advances in synthesis of propylene carbonate from urea alcoholysis. Huaxue Yu Shengwu Gongcheng 2009, 26, 1–4. [Google Scholar]
  60. Shukla, K.; Srivastava, V.C. Synthesis of organic carbonates from alcoholysis of urea: A review. Catal. Rev. 2017, 59, 1–43. [Google Scholar] [CrossRef]
  61. Védrine, J.C. Heterogeneous catalysis on metal oxides. Catalysts 2017, 7, 341. [Google Scholar] [CrossRef] [Green Version]
  62. Xiao, F.; Liu, L.; Li, J.; Li, S. Method and Device for Extracting Biuret from Reaction Liquid of Urea and Polyol. CN 111153833, 15 May 2020. [Google Scholar]
  63. Pu, Y.; Li, L.; Qiao, C.; Wang, F.; Yang, H.; Xiao, Y.; Zhou, S.; Tian, Y. Method for Extracting Biuret in Process of Preparing Cyclic Carbonate by Reacting Urea and Polyalcohol. CN 110256397, 20 September 2019. [Google Scholar]
  64. Liu, X.; Chen, Y.; Jia, W.; Pei, K.; Xu, M.; Zhang, Y. Method for Preparing Dimethyl Carbonate by Urea Two-Step Method and the System Thereof. CN 107353207, 17 November 2017. [Google Scholar]
  65. Sun, Y. Preparation of Dimethyl Carbonate by Indirect Alcoholysis of Urea. CN 106083585, 9 November 2016. [Google Scholar]
  66. Li, Y.; Luo, G.; Jiang, D.; Wang, G. Catalyst and Method for Preparing Dimethyl Carbonate by Utilizing this Catalyst. CN 105251496, 20 January 2016. [Google Scholar]
  67. Xiao, F.; Yang, J.; Zhao, N.; Li, S.; Du, Y.G.; Yang, J.; Gao, W.; Zhang, J. Continuous Reaction Process, Horizontal Continuous Mixer and Horizontal Reactor for Synthesizing Cyclocarbonate from Urea. CN 105017206, 4 November 2015. [Google Scholar]
  68. He, X. Method for Preparing Propylene Carbonate from Urea and 1,2-Propanediol with Magnesium Oxide-Zinc Oxide Composite Rare Earth Oxide as Catalyst. CN 104961720, 7 October 2015. [Google Scholar]
  69. Xiao, F.; Zhang, T.; Zhao, N.; Wang, F.; Yang, J.; Shan, H.; Wan, J.; Wu, C. Propylene Carbonate Synthesis Catalyst and Its Preparation and Application. CN 103933961, 23 July 2014. [Google Scholar]
  70. Xiao, F.; Zhao, N.; Yang, J.; Wang, X.; Liu, Z. Continuous Process for Synthesis of Cyclic Carbonates from Urea. CN 104059047, 24 September 2014. [Google Scholar]
  71. Xiao, F.; Zhang, T.; Zhao, N.; Wang, F.; Yang, J.; Shan, H.; Wan, J.; Wu, C. Preparation of Propylene Carbonate in Presence of Metal Oxide Catalyst. CN 103721697, 16 April 2014. [Google Scholar]
  72. Dobrichovsky, M.; Majer, I. Process and Catalysts for Producing Alkylene and/or Dialkyl Carbonates. WO 2009143785, 3 December 2009. [Google Scholar]
  73. Zhao, X.; Zhang, Y.; Chen, Y.; Wang, Y. Process for the Preparation of Dimethyl Carbonate from Urea by Alcoholysis Using Aliphatic Dihydroxy Alcohol as Recycling Agent. CN 1569810, 26 January 2005. [Google Scholar]
  74. Sun, Y.; Wei, W.; Li, Q.; Zhao, N.; Wang, M.; Wu, Y.; Chu, Y. Process for Preparation of Propylene or Ethylene Carbonate. CN 1421431, 4 June 2003. [Google Scholar]
  75. Doya, M.; Kanbara, Y.; Kimizuka, K.-I.; Okawa, T. Process and Catalysts for the Production of Alkylene Carbonates from Alkylene Glycols and Urea. EP 0625519, 23 November 1994. [Google Scholar]
  76. Doya, M.; Ohkawa, T.; Kanbara, Y.; Okamoto, A.; Kimizuka, K. Process and Catalysts for the Industrial-Scale Preparation of Alkylene Carbonates from Glycols and Urea. EP 0581131, 30 March 1994. [Google Scholar]
  77. Hao, L.; Wang, J.; Shen, W.; Zhu, Z.; Fang, Y. Carbonylation of 1,2-propylene glycol with urea to propylene carbonate under the catalysis of Zn-Al oxide from the view of homogeneous catalysis. Mol. Catal. 2018, 452, 54–59. [Google Scholar] [CrossRef]
  78. An, H.; Zhang, G.; Zhao, X.; Wang, Y. Preparation of highly stable Ca-Zn-Al oxide catalyst and its catalytic performance for one-pot synthesis of dimethyl carbonate. Catal. Today 2018, 316, 185–192. [Google Scholar] [CrossRef]
  79. Liu, S.; Sun, S.; Tian, X.; Sun, P.; Zhang, S.; Yao, Z. Zn-Ca-Al mixed oxide as efficient catalyst for synthesis of propylene carbonate from urea and 1,2-propylene glycol. Chin. J. Chem. Eng. 2017, 25, 609–616. [Google Scholar] [CrossRef]
  80. An, H.; Ma, Y.; Zhao, X.; Wang, Y. Preparation of Zn-Al oxide catalyst and its catalytic performance in propylene carbonate synthesis from urea and propylene glycol on a fixed-bed reactor. Catal. Today 2016, 264, 136–143. [Google Scholar] [CrossRef]
  81. Zhang, T.; Zhang, B.; Li, L.; Zhao, N.; Xiao, F. Zn-Mg mixed oxide as high-efficiency catalyst for the synthesis of propylene carbonate by urea alcoholysis. Catal. Commun. 2015, 66, 38–41. [Google Scholar] [CrossRef]
  82. Zhang, G.; An, H.; Zhao, X.; Wang, Y. Preparation of Ca-Zn-Al Oxides and Their Catalytic Performance in the One-Pot Synthesis of Dimethyl Carbonate from Urea, 1,2-Propylene Glycol, and Methanol. Ind. Eng. Chem. Res. 2015, 54, 3515–3523. [Google Scholar] [CrossRef]
  83. Wu, D.; Guo, Y.; Geng, S.; Xia, Y. Synthesis of Propylene Carbonate from Urea and 1,2-Propylene Glycol in a Monolithic Stirrer Reactor. Ind. Eng. Chem. Res. 2013, 52, 1216–1223. [Google Scholar] [CrossRef]
  84. Yu, G.-L.; Chen, X.-R.; Chen, C.-L. Synthesis of propylene carbonate from urea and 1,2-propylene glycol over ZnO/NaY catalyst. React. Kinet. Catal. Lett. 2009, 97, 69–75. [Google Scholar] [CrossRef]
  85. Li, Q.; Zhao, N.; Wei, W.; Sun, Y. Catalytic performance of metal oxides for the synthesis of propylene carbonate from urea and 1,2-propanediol. J. Mol. Catal. A Chem. 2007, 270, 44–49. [Google Scholar] [CrossRef]
  86. 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] [CrossRef]
  87. Zhao, X.; Jia, Z.; Wang, Y. Clean synthesis of propylene carbonate from urea and 1,2-propylene glycol over zinc-iron double oxide catalyst. J. Chem. Technol. Biotechnol. 2006, 81, 794–798. [Google Scholar] [CrossRef]
  88. Zhao, X.; Zhang, Y.; Wang, Y. Synthesis of Propylene Carbonate from Urea and 1,2-Propylene Glycol over a Zinc Acetate Catalyst. Ind. Eng. Chem. Res. 2004, 43, 4038–4042. [Google Scholar] [CrossRef]
  89. Bhanage, B.M.; Fujita, S.-I.; Ikushima, Y.; Arai, M. Transesterification of urea and ethylene glycol to ethylene carbonate as an important step for urea based dimethyl carbonate synthesis. Green Chem. 2003, 5, 429–432. [Google Scholar] [CrossRef]
  90. Estiu, G.; Merz, K.M. The hydrolysis of urea and the proficiency of urease. J. Am. Chem. Soc. 2004, 126, 6932–6944. [Google Scholar]
  91. Qin, H.; Tan, X.; Huang, W.; Jiang, J.; Jiang, H. Application of urea precipitation method in preparation of advanced ceramic powders. Ceram. Int. 2015, 41, 11598–11604. [Google Scholar] [CrossRef]
  92. An, H.; Zhao, X.; Jia, Z.; Wu, C.; Wang, Y. Synthesis of magnetic Pb/Fe3O4/SiO2 and its catalytic activity for propylene carbonate synthesis via urea and 1,2-propylene glycol. Front. Chem. Eng. China 2009, 3, 215–218. [Google Scholar]
  93. Du, Z.; Wu, Y.; Yuan, H.; Wang, C.; Xiong, J.; Zhou, B.; Huang, L.; Huang, C. Preparation of Cyclic Carbonate by Alcoholysis of Urea. CN 102229597, 2 November 2011. [Google Scholar]
  94. Du, Z.; Chen, F.; Lin, Z.; Li, X.; Yuan, H.; Wu, Y. Effect of MgO on the catalytic performance of MgTiO3 in urea alcoholysis to propylene carbonate. Chem. Eng. J. 2015, 278, 79–84. [Google Scholar] [CrossRef]
  95. Katz, L.; Ward, R. Structure relations in mixed metal oxides. Inorg. Chem. 1964, 3, 205–211. [Google Scholar] [CrossRef]
  96. Shi, L.; Wang, S.-J.; Wong, D.S.-H.; Huang, K. Novel Process Design of Synthesizing Propylene Carbonate for Dimethyl Carbonate Production by Indirect Alcoholysis of Urea. Ind. Eng. Chem. Res. 2017, 56, 11531–11544. [Google Scholar] [CrossRef]
  97. Wang, Y. Green Process Development of Dimethyl Carbonate. Master’s Thesis, Qingdao University of Science and Technology, Qingdao, China, 2012. [Google Scholar]
  98. Patraşcu, I.; Bîldea, C.S.; Kiss, A.A. Novel eco-efficient process for dimethyl carbonate production by indirect alcoholysis of urea. Chem. Eng. Res. Des. 2020, 160, 486–498. [Google Scholar] [CrossRef]
  99. Fang, Y.; Shen, W. Preparation Method of Propylene Carbonate (or Ethylene Carbonate) from Urea and Propylene Glycol (or Ethylene Glycol) by Continuous Reactive Spraying. CN 111943927, 17 November 2020. [Google Scholar]
  100. Fang, Y.; Cao, G.; Du, W.; Wang, Y.; Xu, R.; Song, Z.; Meng, Q.; Liu, Q.; Wu, F.; Shao, C.A. Continuously Production of Propylene Carbonate or Ethylene Carbonate. CN 103420972, 4 December 2013. [Google Scholar]
  101. Zhou, J.; Wu, D.; Zhang, B.; Guo, Y. Synthesis of propylene carbonate from urea and 1,2-propylene glycol over metal carbonates. Chem. Ind. Chem. Eng. Q. 2011, 17, 323–331. [Google Scholar] [CrossRef]
  102. Gao, Z.W.; Wang, S.F.; Xia, C.G. Synthesis of propylene carbonate from urea and 1,2-propanediol. Chin. Chem. Lett. 2009, 20, 131–135. [Google Scholar] [CrossRef]
  103. Sun, W.; Wang, S.; Xia, C. Process for Preparation of Organic Carbonates. CN 101440035, 27 May 2009. [Google Scholar]
  104. Indran, V.P.; Saud, A.S.H.; Maniam, G.P.; Yusoff, M.M.; Taufiq-Yap, Y.H.; Rahim, M.H.A. Versatile boiler ash containing potassium silicate for the synthesis of organic carbonates. RSC Adv. 2016, 6, 34877–34884. [Google Scholar] [CrossRef] [Green Version]
  105. Büttner, H.; Müller, J.; Jonker, R.; Doedt, S.; Bosse, T.; Makran, S. Process for Synchronous Absorption of Carbon Dioxide from Flue Gas for Simultaneous Synthesis of Dialkyl/Alkylene Carbonates. WO 2014059961, 24 April 2014. [Google Scholar]
  106. Liu, S.; Wu, S.; Wang, G.; Lei, Y.; Li, J.; Yang, X. Preparation of Propylene Carbonate from Urea and 1,2-Propylene Glycol in Presence of Heteropolyacids. CN 101450317, 10 June 2009. [Google Scholar]
  107. Chen, W.; Huang, Z.; Liu, Y.; He, Q. Preparation and characterization of a novel solid base catalyst hydroxyapatite loaded with strontium. Catal. Commun. 2008, 9, 516–521. [Google Scholar] [CrossRef]
  108. Zahouily, M.; Abrouki, Y.; Bahlaouan, B.; Rayadh, A.; Sebti, S. Hydroxyapatite: New efficient catalyst for the Michael addition. Catal. Commun. 2003, 4, 521–524. [Google Scholar] [CrossRef]
  109. Du, Z.; Wu, Y.; Chen, Q.; Liu, L. Synthesis of Alkylene Carbonates by Cyclocondensation Reaction. CN 101544627, 30 September 2009. [Google Scholar]
  110. Du, Z.; Liu, L.; Yuan, H.; Xiong, J.; Zhou, B.; Wu, Y. Synthesis of Propylene Carbonate from Alcoholysis of Urea Catalyzed by Modified Hydroxyapatites. Chin. J. Catal. 2010, 31, 371–373. [Google Scholar] [CrossRef]
  111. Lei, Z.; Chen, B.; Koo, Y.-M.; MacFarlane, D.R. Introduction: Ionic Liquids. Chem. Rev. 2017, 117, 6633–6635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Gabriel, S.; Weiner, J. Ueber einige Abkömmlinge des Propylamins. Ber. Dtsch. Chem. Ges. 1888, 21, 2669–2679. [Google Scholar] [CrossRef] [Green Version]
  113. Walden, P. Molecular weights and electrical conductivity of several fused salts. Bull. Russ. Acad. Sci. 1914, 8, 405–422. [Google Scholar]
  114. Singh, S.K.; Savoy, A.W. Ionic liquids synthesis and applications: An overview. J. Mol. Liq. 2020, 297, 112038. [Google Scholar] [CrossRef]
  115. Greer, A.J.; Jacquemin, J.; Hardacre, C. Industrial applications of ionic liquids. Molecules 2020, 25, 5207. [Google Scholar] [CrossRef]
  116. Kuznetsov, V.A.; Pervova, M.G.; Pestov, A.V. Synthesis of alkylene carbonates in ionic liquid. Russ. J. Org. Chem. 2013, 49, 1859–1860. [Google Scholar] [CrossRef]
  117. Smith, E.L.; Abbott, A.P.; Ryder, K.S. Deep Eutectic Solvents (DESs) and Their Applications. Chem. Rev. 2014, 114, 11060–11082. [Google Scholar] [CrossRef] [Green Version]
  118. Abbott, A.P.; Barron, J.C.; Ryder, K.S.; Wilson, D. Eutectic-based ionic liquids with metal-containing anions and cations. Chem. Eur. J. 2007, 13, 6495–6501. [Google Scholar] [CrossRef]
  119. Qu, Q.; Zhao, J.; Yang, G. Method for Synthesizing Propylene Carbonate by Reacting Propanediol and Urea with Ionic Liquid Catalyst. CN 105601609, 25 May 2016. [Google Scholar]
  120. Cheng, W.; Deng, L.; Li, Z.; Dong, L.; Shi, Z.; Su, Q.; Zhang, S. Method for Synthesizing Cyclic Carbonate from Urea and Dihydric Glycol by Imidazole Ionic Liquid Catalysis. CN 110156742, 23 August 2019. [Google Scholar]
  121. Deng, L.; Sun, W.; Shi, Z.; Qian, W.; Su, Q.; Dong, L.; He, H.; Li, Z.; Cheng, W. Highly synergistic effect of ionic liquids and Zn-based catalysts for synthesis of cyclic carbonates from urea and diols. J. Mol. Liq. 2020, 316, 113883. [Google Scholar] [CrossRef]
  122. Kulasegaram, S.; Shaheen, U.; Turney, T.W.; Gates, W.P.; Patti, A.F. Zinc monoglycerolate as a catalyst for the conversion of 1,3- and higher diols to diurethanes. RSC Adv. 2015, 5, 47809–47812. [Google Scholar] [CrossRef]
  123. Xiao, Y.; Xiang, C.; Lei, H.; Jin, S.; Yin, X.; Ding, Y.; Du, Z. Effect of change of Ca, P and Mg on the surface of catalyst prepared from phosphate tailing on urea alcoholysis. Catal. Commun. 2019, 128, 105712. [Google Scholar] [CrossRef]
Scheme 1. Selected methods for synthesis of propylene carbonate.
Scheme 1. Selected methods for synthesis of propylene carbonate.
Catalysts 12 00309 sch001
Scheme 2. Mechanism of propylene carbonate synthesis via urea alcoholysis catalyzed by zinc oxide. Reprinted from [85] with permission from Elsevier.
Scheme 2. Mechanism of propylene carbonate synthesis via urea alcoholysis catalyzed by zinc oxide. Reprinted from [85] with permission from Elsevier.
Catalysts 12 00309 sch002
Scheme 3. Scheme of a side reaction of MOD formation.
Scheme 3. Scheme of a side reaction of MOD formation.
Catalysts 12 00309 sch003
Scheme 4. Mechanism of propylene carbonate synthesis via urea alcoholysis catalysed by the system [C16mim]Cl, ZnCl2. Reprinted from [121] with permission from Elsevier.
Scheme 4. Mechanism of propylene carbonate synthesis via urea alcoholysis catalysed by the system [C16mim]Cl, ZnCl2. Reprinted from [121] with permission from Elsevier.
Catalysts 12 00309 sch004
Table 1. Selected properties of propylene carbonate.
Table 1. Selected properties of propylene carbonate.
PropertiesValueReference
Melting point, °C−49[36]
Boiling point, °C242[36]
Flash point, °C128[36]
Relative permittivity66[36]
Density (30 °C), g/cm31.195[37]
Enthalpy of vaporization (150 °C), kcal/mol13.19[37]
Specific heat capacity (25 °C), J/(g·K)1.20[38]
Standard molar enthalpy of formation, kJ/mol−614.1[38]
Table 2. Catalytic performance of metal oxides for the synthesis of propylene carbonate 1. Reprinted from [85] with permission from Elsevier.
Table 2. Catalytic performance of metal oxides for the synthesis of propylene carbonate 1. Reprinted from [85] with permission from Elsevier.
CatalystsConversion, %, UreaYield, %
2-Hydroxypropyl Carbamate Propylene Carbonate4-Methyl-2-Oxazolidone
Blank89.348.640.70
Al2O390.451.538.90
ZrO292.852.740.10
ZnS89.850.339.50
ZnO100.01.198.90
MgO97.72.892.91.4
La2O394.35.686.82.2
CaO85.67.274.63.8
1 Reaction conditions: temperature: 170 °C, pressure: 280 mmHg, time: 2 h, PG: 0.75 mol, urea: 0.5 mol, amount of catalyst: 0.6 g.
Table 3. BET surface area and acid-base properties of metal oxides. Reprinted from [85] with permission from Elsevier.
Table 3. BET surface area and acid-base properties of metal oxides. Reprinted from [85] with permission from Elsevier.
EntryCatalystSBET, m2/gAcidity, μmol/gBasicity, μmol/gA/B a
1CaO9.52.3063.660.04
2La2O321.00.541.690.32
3MgO7.43.9317.820.22
4ZnO6.71.020.981.04
5ZnS35.231.310.5457.64
6ZrO233.429.103.229.04
7Al2O3136.520.906.933.02
a Acidity/basicity ratio.
Table 4. Catalytic activity of ZnO in the synthesis of propylene carbonate.
Table 4. Catalytic activity of ZnO in the synthesis of propylene carbonate.
Urea:PG Molar RatioTime, hTemperature, °CUrea Conversion, %Selectivity, %Yield, %Ref.
1:1.04313810099.1-[72]
1:2Continuous process in a flow tank reactor (residence time—3 h)17010099.4-[75]
1:2.5214510056.2-[76]
2:3217010098.9-[85]
1:1.5417099.899.8-[86]
1:65170--67[88]
1:1.3531507791-[89]
Table 5. Catalytic action of ZnO-containing metal oxide catalytic systems.
Table 5. Catalytic action of ZnO-containing metal oxide catalytic systems.
CatalystUrea:PG Molar RatioTime, hTemperature, °CUrea Conversion, %Selectivity, %Yield, %Leaching of the CatalystRef.
NiO/ZnO1:4Continuous process in a flow tank reactor with a stirrer16010095-n/a 1[64]
Fe2O3/ZnO1:43.5170–180--93.5n/a[66]
MgO/ZnO1:43.5170–180--93n/a[66]
ZnO-Al2O31:1.9Continuous process in 2 horizontal reactors1. reactor: 130
2. reactor: 180
--97n/a[67]
La2O3-CeO2/MgO-ZnO1:3.5–1:4.51.5–2.5150–17098-85n/a[68]
ZnO-CaO1:100.5170--99.1n/a[69]
ZnO-MgO1:12140--99.6n/a[69]
ZnO-Al2O3-Fe2O31:52170--96.8n/a [69]
ZnO-Al2O31:320250--95.7n/a[69]
ZnO-MgO-Fe2O31:26180--99.4n/a[69]
ZnO-ZrO2-Fe2O31:20.5250--98.0n/a[69]
ZnO-Al2O31:1.9Continuous process in a cascade of 3 flow-through tank reactors1. reactor: 130
2. reactor: 150
3. reactor: 180
100-97n/a[70]
ZnO-MgO1:12140--90n/a[71]
ZnO-Al2O3-Fe2O31:52170--98n/a[71]
ZnO-CaO1:46200--85n/a[71]
ZnO-Al2O31:320250--86n/a[71]
ZnO-MgO-Fe2O31:26180--87n/a[71]
ZnO-ZrO2-Fe2O31:20.5250--83n/a[71]
MgO-ZnO1:42170--92n/a[73]
CoO-ZnO1:43170--90n/a[73]
ZnO-Al2O32:33.3170100-96.6Zn—2.3%
Al—6.1%
[77]
ZnO-CaO-Al2O31:33184--90.8ZnO content dropped from 20.2% to 12.6% after 5th reuse[79]
ZnO-Al2O31:6Continuous process in a fixed bed reactor (LHSV = 1.0 h−1)140-97.785.1n/a[80]
ZnO-MgO1:1.50.5170--94.8n/a[81]
ZnO-Cr2O31:1.56180--97.8n/a[83]
ZnO-ZnFe2O41:42170--78.4n/a[87]
1 not available information.
Table 6. Catalytic performance of zinc-free oxide catalysts.
Table 6. Catalytic performance of zinc-free oxide catalysts.
CatalystUrea:PG Molar RatioTime, hTemperature, °CUrea Conversion, %Selectivity, %Yield, %Ref.
MgO/Al2O31:43.5170–180--92.5[66]
CaO/MgO1:43.5170–180--93.5[66]
Fe2O3-MgO1:43180--93[73]
MgO1:2119010098.9-[75]
MgO1:4116510096.5-[76]
PbO21:2216510093.9 [76]
CaO1:2216510095.4-[76]
Al2O32:3217090.4-38.9[85]
ZrO22:3217092.8-40.1[85]
MgO2:3217097.7-92.9[85]
La2O32:3217094.3-86.4[85]
CaO2:3217085.6-74.6[85]
MgO1:65170--75[88]
MgO-ZrO21:65170--45[88]
PbO1:65170--57[88]
Bu2SnO1:65170--61[88]
Pb/Fe3O4/SiO21:42180--87.7[92]
Table 7. Catalytic activities of catalysts comprising oxides with perovskite structures.
Table 7. Catalytic activities of catalysts comprising oxides with perovskite structures.
CatalystUrea:PG Molar RatioTime, hTemperature, °CSelectivity, %Yield, %Ref.
LaMnO31:4217088.375.2[93]
La0.5Pb0.5MnO31:4217090.374.7[93]
La0.8Pb0.2MnO31:4217091.570.2[93]
La0.9Ag0.1MnO31:4217085.172.9[93]
LaMn0.5Fe0.5O31:4217084.364.8[93]
La0.5Pb0.5Mn0.5Co0.5O31:4217088.260.2[93]
Ce0.5Pb0.5MnO31:4217066.755.8[93]
CaTiO31:40217090.584.7[93]
BaTiO31:20217087.271.1[93]
Ba0.5Na0.5TiO31:4217090.180.5[93]
PbZrO31:4217091.382.4[93]
CaSnO31:4217090.885.4[93]
KNbO31:4217092.388.4[93]
Na0.5K0.5NbO31:4217092.388.4[93]
MgTiO31:4217099.093.5[94]
Table 8. The catalytic effect of metal salts.
Table 8. The catalytic effect of metal salts.
CatalystUrea:PG Molar RatioTime, hTemperature, °CUrea Conversion, %Selectivity, %Yield, %Ref.
Bu2Sn(C11H23COO)21:1.5317510064.6-[76]
ZnCO31:1.5317510089.5-[76]
ZnCl21:1.5317510087.8-[76]
Zn(OAc)21:1.5317510088.3-[76]
Zn(NO3)21:1.5317510088.8-[76]
Mg(OAc)21:1.5317510088.1-[76]
ZnS2:3217089.8-39.5[85]
Zn(OAc)2 × 2H2O1:65170--54[88]
MgCO31:65170--56[88]
Mg(OAc)21:65170--38[88]
Pb(OAc)2 × 3H2O1:65170--38[88]
Pb(OAc)21:65170--58[88]
Zn(OAc)21:43170--94[88]
Zn(OAc)2/AC1:43170--77.9[88]
Zn(OAc)2/zeo1:43170--66.6[88]
Zn(OAc)2/γ-Al2O31:43170--77.5[88]
PbCO3-Zn5(CO3)2(OH)62:35180--93.5[101]
MgCl21:43160--96.5[102]
ZnCl21:43160--92.4[102]
SnCl21:6.35180--93.47[103]
Zn(NO3)21:1.63200--57.36[103]
CaCl21:3.920150--43.16[103]
FeCl31:3.25170--83.29[104]
PbCl21:7.92200--62.89[103]
K2SiO31:410150--72.8[104]
Table 9. Catalytic activity of heteropolyoxometalates [106].
Table 9. Catalytic activity of heteropolyoxometalates [106].
CatalystCarbonate Used in the Synthesis of the CatalystTime, hSelectivity, %Yield, %
Mg3(PMo12O40)24MgCO3 × Mg(OH)2314071.82
Mg3(PW12O40)24MgCO3 × Mg(OH)2514043.83
K3PMo12O40K2CO3315054.48
Mg2SiW12O404MgCO3 × Mg(OH)2315062.78
Zn3(PMo12O40)2ZnCO34.515078.71
Table 10. Catalytic activity of modified hydroxyapatites.
Table 10. Catalytic activity of modified hydroxyapatites.
CatalystUrea:PG Molar RatioTime, hTemperature, °CSelectivity, %Yield, %Ref.
Na-HAP1:4217099.088.6[109]
Li-HAP1:4217095.572.5[109]
K-HAP1:4217099.170.4[109]
Sr-HAP1:4217099.072.1[109]
Ba-HAP1:4217099.074.3[109]
Mn-HAP1:4217099.577.9[109]
Co-HAP1:4217093.832.7[109]
Pb-HAP1:4217098.269.1[109]
Zn-HAP1:4217099.673.8[109]
Ce-HAP1:4217096.350.5[109]
Zn-Pb-HAP1:4217098.485.5[101]
Pb-Mn-HAP1:4217097.680.8[101]
HAP1:42170-50.6[110]
La-HAP1:42170-91.5[110]
Table 11. Catalytic activity of other catalysts.
Table 11. Catalytic activity of other catalysts.
CatalystUrea:PG Molar RatioTime, hTemperature, °CUrea Conversion, %Selectivity, %Yield, %Ref.
Zn1:1.5317510073-[76]
Zn1:65170--57[88]
Pb1:65170--59[88]
Zinc monoglycerolate1.5:124140PG conversion 99-80[122]
BA 9001:410150--73.2[104]
PT-800-41:42170-95.681.8[123]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kotyrba, Ł.; Chrobok, A.; Siewniak, A. Synthesis of Propylene Carbonate by Urea Alcoholysis—Recent Advances. Catalysts 2022, 12, 309. https://doi.org/10.3390/catal12030309

AMA Style

Kotyrba Ł, Chrobok A, Siewniak A. Synthesis of Propylene Carbonate by Urea Alcoholysis—Recent Advances. Catalysts. 2022; 12(3):309. https://doi.org/10.3390/catal12030309

Chicago/Turabian Style

Kotyrba, Łukasz, Anna Chrobok, and Agnieszka Siewniak. 2022. "Synthesis of Propylene Carbonate by Urea Alcoholysis—Recent Advances" Catalysts 12, no. 3: 309. https://doi.org/10.3390/catal12030309

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