Next Article in Journal
Visible Light-Induced Photocatalyst with Au/TiO2 Nanocomposites Fabricated through Pulsed Laser-Induced Photolysis
Previous Article in Journal
Preparation and Characterization of Supported Molybdenum Doped TiO2 on α-Al2O3 Ceramic Substrate for the Photocatalytic Degradation of Ibuprofen (IBU) under UV Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Achievements in the Synthesis of Cyclic Carbonates from Olefins and CO2: The Rational Design of the Homogeneous and Heterogeneous Catalytic System

1
Key Laboratory of Hubei Province for Coal Conversion and New Carbon Materials, School of Chemistry and Chemical Engineering, Wuhan University of Science and Technology, Wuhan 430081, China
2
Hubei Key Laboratory of Biomass Fibers and Eco-Dyeing & Finishing, Wuhan Textile University, Wuhan 430073, China
3
State Key Laboratory of High-Efficiency Coal Utilization and Green Chemical Engineering, College of Chemistry and Chemical Engineering, Ningxia University, Yinchuan 750021, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2022, 12(5), 563; https://doi.org/10.3390/catal12050563
Submission received: 30 April 2022 / Revised: 16 May 2022 / Accepted: 18 May 2022 / Published: 20 May 2022
(This article belongs to the Topic Catalysis: Homogeneous and Heterogeneous)

Abstract

:
With the consumption of fossil fuels, the level of CO2 in the atmosphere is growing rapidly, which leads to global warming. Hence, the chemical conversion of CO2 into high value-added products is one of the most important approaches to reducing CO2 emissions. Due to being simple, inexpensive and environmentally friendly, the direct synthesis of cyclic carbonates from olefins and CO2 is a promising project for industrial application. In this review, we discuss the design of the homogeneous and heterogeneous catalytic system for the synthesis of cyclic carbonates from the reaction of olefins and CO2. Usually, the catalyst contains the epoxidation active site and the cycloaddition active site, which could achieve the oxidation of oleifins and the CO2-insert, respectively. This review will provide a comprehensive overview of the direct synthesis of cyclic carbonates from olefins and CO2 catalyzed by homogeneous and heterogeneous catalysts. The focus mainly lies on the rational fabrication of multifunctional catalysts, and provides a new perspective for the design of catalysts.

1. Introduction

The use of CO2 as a building block for the synthesis of high value-added chemicals has become an important current area of focus in CO2 conversion [1]. As a cheap, non-toxic and renewable C1 resource, CO2 has been successfully converted into many high value-added chemicals such as alcohols [2], carboxylic acids [3], formamides [4] and organic carbonates [5]. Considering environmental awareness and the atom-economic strategy, the direct synthesis of cyclic carbonates from CO2 and epoxides may be an alternative way of utilizing CO2. However, the high price of epoxides prevents the application of cycloaddition reaction of CO2. As a result, direct one-pot preparation of cyclic carbonates from olefins and CO2 has become an important issue for the conversion of CO2 [6].
Recently, a lot of homogeneous and heterogeneous catalytic systems have been applied in the reaction of olefins and CO2 [7]. Usually, these catalytic systems have two functionalized active sites: (i) the epoxidation active site for the oxidation of olefins; and (ii) the cycloaddition active site for the coupling reaction between epoxides and CO2 (Scheme 1) [8,9]. However, how to design catalysts reasonably, and how to promote the synergistic catalysis effect, is a big challenge for the direct oxidative carboxylation between olefins and CO2. Herein, we will try to give some suggestions on how to construct the multifunctional catalyst system in the synthesis of cyclic carbonates from olefins and CO2. This review will be divided into two major parts. The first part will give a summary of the homogeneous catalytic system in recent years, while the second part will focus on the heterogeneous catalytic system and the synergistic effect between the different catalytic moieties.

2. Homogeneous Catalytic System

2.1. Ionic Liquids Catalytic System

The straightforward synthesis of cyclic carbonates by direct oxidative carboxylation from olefins and CO2 is a promising method of fixation and utilization of CO2. Ionic liquids, with the advantages of negligible vapor pressure, unique solvation properties and the design-ability of anion and cation, have been widely used in the epoxidation and CO2 cycloaddition [10,11,12]. However, constructing the multifunctional ionic liquids catalytic system with both catalytic ability for olefins epoxidation and CO2 cycloaddition is still a big challenge. The direct formation of styrene carbonate (SC) from styrene and CO2 by employing 1-n-butyl-3-methylimidazolium bromide (BMImBr) as the catalyst and tert-butyl hydroperoxide (TBHP) as an oxidant was achieved (Scheme 2) [13]. The one-pot reaction could give a moderate conversion and selectivity (Conv. 90% and Sel. 40%), while a slightly higher conversion and selectivity (Conv. 99% and Sel. 63%) could be obtained by the two-step reaction.
Due to the lack of the epoxidation active sites, the selectivity of styrene oxide is still low, which may significantly decrease the yield of SC. Considering that bicarbonate may react with hyperoxide to form the active oxidant of peroxymonocarbonate ion (HCO4) [14], Zhang et al. reported the synthesis of SC from styrene and CO2 catalyzed by imidazolium hydrogen carbonate ionic liquids ([CnCmIm][HCO3] ILs) [8]. Owing to the epoxidation active sites and cycloaddition active sites in [CnCmIm][HCO3] (Scheme 3), the excellent yield of cyclic carbonates could be obtained by the one-pot reaction. Recently, Perosa et al. developed the tungstate assisted-tandem reaction system by combining ammonium tungstate ionic liquid catalyst ([N8,8,8,1]2[WO4]) and tetrabutyl-ammonium iodide ([N4,4,4,4]I), which could give a 94% yield of cyclic carbonate using H2O2 as oxidant under 10 bar CO2 (Scheme 4) [15]. In contrast, aromatic and cyclic olefins such as styrene and cyclohexene suffered from easy overoxidation and consequent low selectivity towards the carbonates in the tungstate ionic liquid catalyst. Although the ionic liquids catalyst systems have good design ability, the development of highly efficient ionic liquids catalysts should be carefully designed.

2.2. Metal Complex Catalytic System

It is well-known that some catalysts are inspired by nature and a lot of biological biomimetic catalyst has been applied in the olefin epoxidation reaction [16,17]. Hence, the direct oxidative carboxylation of olefins and CO2 by a metal complex catalytic system seems to be an efficient method. Jing et al. reported the one-pot synthesis of synthesis carbonate from styrene, O2 and CO2 with dioxo(tetraphenylporphyrinato) ruthenium [Ru(TPP)(O)2]/quaternary onium salt [18]. The [Ru(TPP)(O)2] can transform the oxygen atom from the catalyst to form the epoxide which could act as a substrate for the coupling reaction with CO2 (Scheme 5). It is well known that the selectivity of the epoxidation reaction has a great impact on the yield of cyclic carbonate. For example, the yield of styrene carbonate was only 3.6% with benzoic acid (55.3%) and benzaldehyde (37.5%) as major products when using a Fe(TPP)Cl as a comparable catalyst [18]. However, the Ru(Salen)(PPh3)2 catalyst showed a 40.7% yield of styrene carbonate under the same conditions, which meant that the new active site of Ru(Salen)(O) could generate from the Ru(Salen)(PPh3)2 in situ. As a result, the suitable oxidative property and coupling property in the catalyst may play an important role in the catalyst design.
The functionalization of the ligand of metal-complex could have a great effect on the conversion and selectivity of the reaction, which could be an alternative method to optimize catalyst design. Ghosh et al. developed a manganese (III) complex of an amido-amine ligand/tetrabutylammonium bromide (TBAB) catalytic system, which could efficiently convert the alkene to the epoxide and further react with CO2 to form cyclic carbonate in the presence of CO2 and TBHP [19]. The high selectivity of the epoxide may originate from the transformation of the manganese (III) complex of an amido-amine ligand to a non-heme high-valent manganese-oxo intermediate in situ (Scheme 6). Besides, the Lewis acidity in the manganese (III) complex could also accelerate the CO2 cycloaddition reaction. Based on recent research, the one-pot synthesis of cyclic carbonate catalyzed by the biomimetic transition metals (such as Fe, Mn, Ru) catalytic systems usually includes two parts. Firstly, the metal center is oxidized to form the oxo species, which could act as an intermediate to oxidize the olefin. Secondly, the coupling reaction between the epoxides and CO2 is conducted. Although the homogeneous catalytic system showed excellent activity in the olefins epoxidation and CO2 cycloaddition respectively, the integration of the epoxidation active sites and cycloaddition active sites is not easy work.
In addition to the biomimetic transition metals, selective epoxidation of olefins by alkyl hydroperoxide catalyzed by d0 transition metal complexes (Mo(VI), V(V) and Ti(IV)) has been used in the epoxidation reaction [20]. Hu et al. found that a MoO2(acac)2-quaternary ammonium salt catalytic system could achieve the direct synthesis of cyclic carbonates from olefins and CO2 with TBHP as an oxidant [21]. The one-pot synthesis of cyclic carbonates contained two separate reactions catalyzed by the different active sites. During the epoxidation process, the formation of the alkyl-peroxo M-OOR intermediate could coordinate with the olefin to form a three-membered transition state (Scheme 7). After the transformation of oxygen from M-OOR to olefin, epoxide could be obtained. In the following cycloaddition process, the oxygen atom of the epoxide was activated by the [n-Bu4N]+ cation of the TBAB and then the nucleophilic Br- attacked the less hindered carbon atom of the epoxide to form an oxygen anion intermediate. The oxygen atom of the oxygen anion intermediate nucleophilic attacked the carbon atom of CO2 affords the alkyl carbonate intermediate, which was stabilized by the [n-Bu4N]+ cation. Finally, the cyclic carbonate was obtained after the ring closure reaction and the catalyst and TBAB were regenerated.
Table 1 summarizes the homogeneous catalytic system for the conversion of olefins and CO2 to cyclic carbonates. Compared with the different homogeneous catalyst systems, it is found that using the polyoxometalates or metal complexes may be a benefit for the epoxidation reaction and the quaternary ammonium salts may promote the cycloaddition reaction (Table 1, entries 3 and 4). Although ionic liquids have good design ability and can integrate the epoxidation active site and cycloaddition active site into one catalyst, the yield of cyclic carbonates must continue to improve (Table 1, entries 1 and 2). Interestingly, we found that the reaction temperature had a significant influence on the yield and selectivity of cyclic carbonates during the one-pot reaction [8]. With the increasing temperature, the selectivity of cyclic carbonates is decreasing, which indicates that the high temperature is not beneficial for the epoxidation reaction, and the main byproduct is benzaldehyde. If benzaldehyde is formed, the subsequent CO2 cycloaddition cannot take place. Besides, the pressure of O2 is also an important reaction parameter. The high pressure of O2 may lead to over-oxidation and the byproducts such as benzaldehyde, benzoic acid and acetophenone [19]. As we know, the CO2 cycloaddition is hard to achieve, which means that the CO2-insert reaction is unfavorable under low pressure of CO2. However, further increasing the pressure of CO2 may decrease the yield of cyclic carbonates. Hence, utilization of O2 as an oxidant and without the addition of a co-catalyst is still a huge change for the direct oxidative carboxylation between olefins and CO2.

3. Heterogeneous Catalytic System

3.1. Zeolite Catalytic System

Due to the facile and efficient recovery of the catalyst from the reaction mixture, the heterogeneous method has shown enormous potential in industrial application. Zeolites, the microporous crystalline maters, attract a lot of attention in catalysts due to their tunable catalytic properties and shape selectivities [22]. Recent studies have found that a series of titanosilicate sieves (TS-1 [23], Ti-Beta [24], Ti-MWW, etc. [25]) reveal excellent selective oxidation during the epoxidation reaction. The real active sites of the oxidation reaction could be ascribed to the transformation of the tetrahedrally coordinated titanium atoms to Ti-OOH species [26,27], which may further react with olefins to selectively produce the epoxides (Scheme 8). After adding co-catalyst or functionalized modification of the titanosilicate sieves, the direct oxidative carboxylation of olefins and CO2 can proceed. Srivastava et al. first reported the synthesis of cyclic carbonates from olefin, oxidant (H2O2 and TBHP, respectively) and CO2 catalyzed by the microporous titanium-silicalite-1 (TS-1) and mesoporous Ti-MCM-41 [28]. With Ti-MCM-41/TBHP and 4-dimethylaminopyridine (DMAP) as a co-catalyst, a 33% yield of SC can obtain under 6.9 bar CO2 and 120 °C through two steps reaction. He et al. prepared a quaternary ammonium modified delaminated titanosilicate (Del-Ti-MWW-N+Bu3Br) which gave a moderate yield (48% for cyclic carbonate) in a one-pot reaction using H2O2 as an oxidant [29]. However, the pore blocking by heavy byproducts may lead to a decrease in catalytic activity and selectivity. Therefore, the development of highly efficient, easily recyclable and hierarchical porous catalysts for the direct synthesis of cyclic carbonates from olefins and CO2 may overcome these problems.
Because mesoporous materials have a larger pore size than microporous materials, some researchers have designed many mesoporous titanium-silicates applications in oxidative carboxylation of olefins with oxidants and CO2 [20]. For example, Kholdeeva et al. prepared a mesoporous titanium-silicates Ti-MMM-E by evaporation-induced self-assembly (EISA) method [30]. The mesoporous titanium-silicates Ti-MMM-E exhibited superior catalytic performance in the one-step oxidation carboxylation of styrene and 4-methylstyrene with TBHP, TBAB and CO2 under mild conditions (Scheme 9). Moreover, it was found that bromide ions in TBAB may occupy the Lewis acid sites and prevent styrene oxide from further oxidation to benzaldehyde in the presence of TBHP and Ti-MMM-E. To avoid pore blocking, Kholdeeva et al. explored a new method to prepare titanium-grafted mesoporous silica catalysts by metallic titanium clusters deposition on silica supports [31]. The non-single-site Ti(IV)nOx-like silica-supported catalyst (TixOn-MCM-41 and TixOn-MMM-2) from Ti13 metal clusters deposited onto ordered mesoporous silicas showed a high yield of cyclic carbonate. Compared with the traditional titanium-silicates, these grafted TinOx-silica materials had better accessibility to titanium active sites exposed on the surface of the catalyst (Scheme 10).
As we know, large pore size molecular sieves are beneficial for the substrate diffusion and the dispersion of the active sites. Since the well-ordered hexagonal mesoporous silica structures (SBA-15) with uniform pore sizes were discovered by Zhao et al. [32], SBA-15 has attracted increasing attention and is considered an important class of nanostructured support materials in heterogeneous catalysis [33]. Xu et al. reported amino-functionalized ILs@SBA-15 catalysts (N(CH3)2-[Im][Br]@SBA-15), which showed impressive catalytic activity for the one-pot synthesis of cyclic carbonates from olefins and CO2 [34]. The dual function of the N(CH3)2-[Im][Br]@SBA-15 could not only catalyze the oxidation of olefins but also promote the CO2 cycloaddition reaction (Scheme 11). In the oxidation cycle, Br was first oxidized by TBHP to hypobromite (BrO), which could act as an active [O] site to react with styrene to form bromohydrin. After the deprotonation of bromohydrin, styrene oxide was obtained, which was activated by the C2-H on the imidazole ring via the hydrogen bond interaction. Then the C-O bond in the styrene oxide caused polarization. Meanwhile, Br nucleophilic attached the less hindered carbon atom of the styrene oxide, forming an oxyanion intermediate. Subsequently, CO2 may interact with the oxyanion intermediate to form a new alkyl carbonate compound. The cyclic carbonate can be obtained via the intramolecular ring-closure process. As a result, the synergistic effects between Br and N(CH3)2-[Im][Br] effectively promoted the direct oxidative carboxylation reaction of olefins and CO2.
Dendritic fibrous nanosilica (DFNS), with the properties of morphology-controlled and intrinsic mesoporous characteristics, have been widely used in catalysis, gas capture, solar-energy harvesting, energy storage, and so on [35]. Sadeghzadeh et al. developed a melamine-Pd(II) Schiff base complex functionalized FeNi3/DFNS core-shell NPs (FeNi3/DFNS/salen/Pd(II) MNPs) (Scheme 12), which showed 94% yield of styrene carbonate even after 10 runs [36]. Recently, Motavalizadehkakhky et al. applied the DFNS-supported platinum (II) complexes (DFNS/Pt(II) NPS) in the synthesis of cyclic carbonate from olefins and CO2 [37]. It was found that the Pt(II) complex in the DFNS played an important role in the direct oxidative carboxylation of CO2 and styrene to styrene carbonate.

3.2. MOFs Catalytic System

Metal-organic frameworks (MOFs), as a new kind of porous materials with high crystalline, high metal content, large surface area and tunable pore structure, have shown great potential applications in gas storage [38], energy storage [39] and catalysis [40]. Some pioneering research has been reported in the olefins epoxidation reaction and CO2 cycloaddition reaction respectively [41,42,43,44,45]. The direct synthesis of cyclic carbonates from olefins and CO2 catalyzed by MOFs usually contained two parts, including the epoxidation reaction and cycloaddition reaction [46]. As a result, combining these two reactions into one functionalized MOFs catalyst may not only avoid the co-catalyst but also promote the conversion and selectivity. Duan et al. used MOF as support and incorporated a pyrrolidine moiety as a chiral organocatalyst and a polyoxometalate as an oxidation catalyst (ZnW-PYIs, ZnW: ZnW12O406) for the direct synthesis of cyclic carbonates from olefins and CO2 (Scheme 13) [47]. The oxygen-enriched surface of ZnW12O406− could accelerate the transformation of the catalytic precursors to the active intermediate of the epoxidation [48]. Meanwhile, pyrrolidine-2-yl-imidazole (PYI) could act as a cooperative catalytic site, which may enhance the activities of the oxidants and drive the catalysis asymmetrically [49]. Based on the synergistic effect, 72–92% yields of cyclic carbonates and 55–80% enantiomeric excesses were obtained with styrene and its derivatives in the presence of TBHP and TBAB. Compared with the polyoxometalate, Mn-porphyrin is also a suitable candidate for epoxidation reaction [50]. Nagaraja et al. constructed a 3D Mn(II)-porphyrin metal-organic framework ([{Mn2(TCPP)·2H2O}·DMF]n, TCCP = 5,10,15,20-tetrakis(4-benzoate)porphyrin), which showed excellent selective capture of CO2 and further transformed CO2 and olefins to cyclic carbonates [51]. The non-metallated porphyrin linker in the [{Mn2(TCPP)·2H2O}·DMF]n can adsorb two CO2 molecules by the interaction of CO2 and the pyrrole -N-H groups of the porphyrin linker through the -N-H---O=C=O bond. Moreover, the implantation of Fe(III) ions in the [{Mn2(TCPP)·2H2O}·DMF]n can give 98% conversion of styrene with about 65% selectivity for cyclic carbonate under mild conditions. Similarly, Li et al. developed a molybdenum (VI) oxide incorporated cooperative catalyst (CuMo-BPY: Cu33-OH)2(4,4′-BPY)(MoO4)2), which exhibited catalytic activity for the synthesis of cyclic carbonates through the two-step catalytic process [52]. The spatial matching between the Mo(VI) and alkaline μ3-OH tricopper (II) cores could provide the compatibility with the Mo=Ot-activated epoxidation intermediate during the epoxidation reaction. In the CO2 cycloaddition reaction, the μ3-OH tricopper (II) core could increase the adsorption of CO2 with the formation of a bidentate bridge or tridentate linear adsorption state [53], which could promote the CO2-insertion reaction. Hence, improving the adsorption or desorption property of a catalyst can significantly influence the catalytic activity [54].
The rational design of multifunctional MOFs can achieve selective capture and conversion of CO2 under mild conditions [55], but combining CO2 capture with its conversion is still a problem. Recently, Nguyen et al. prepared lanthanide-based metal-organic frameworks (MOF-591 and MOF-591), which showed good adsorption of CO2 [56]. Additionally, MOF-590 revealed excellent catalytic activity for the one-pot oxidative carboxylation of olefins with CO2. Usually, the oxidative carboxylation of styrene and CO2 catalyzed by the MOF-590 consisted of two steps, including the epoxidation catalyzed by MOF-590/nBu4NBr-TBHP and the cycloaddition catalyzed by MOF-590/nBu4NBr (Scheme 14). The Nd3+ clusters of MOF-590 were first oxidized to Nd4+-peroxy species and Nd4+-peroxy species released a t-butoxy radical to oxidize styrene to obtain the styrene oxide through the butoxperoxy intermediate [42]. Secondly, the styrene oxide was first coordinated by the Nd unit in the MOF-90 and the C-O bond in the styrene oxide was polarization and elongated. Then Br attacked the less steric hindrance C atom in the styrene oxide forming the metal-coordinated bromoalkoxide intermediate, which was stabilized by the Nd unit. The metal-carbonate intermediates were formed by the insertion of CO2 into the metal-coordinated bromoalkoxide intermediate. After the ring-closing reaction, styrene carbonate was formed and the MOF-590 and nBu4NBr were regenerated.
To avoid the addition of a co-catalyst, Ahn et al. constructed a zirconium-based porphyrinic MOF-545 (Zr6O8(H2O)8(H2-TCPP)2) using a tetrakis(4-carboxyopheny) porphyrin (H6-TCPP) ligand and Zr6 clusters, followed by Mn(III) metalation to the porphyrin units for the epoxidation reaction (Scheme 15) [57]. Then, 1-methyl-3-(2-carboxyethyl)imidazolium bromide (ImBr) ionic liquid was introduced to the Zr6 clusters forming ImBr-functionalized MOF-545(Mn) (ImBr-MOF-545(Mn)) for the CO2 cycloaddition reaction. The MOF-545 could act as a support platform that integrated the Lewis acidic Mn(III)/Zr(IV) ions and Br in the adjacent position and exhibited a high yield of cyclic carbonates in the absence of any co-catalyst under mild reaction conditions.

3.3. Metal Nanoparticles Catalytic System

Metal nanoparticles (NPs) or single-atom catalytic sites (SACs) have been wildly used in a variety of catalytic reactions due to their excellent activity [58,59]. For example, Sun and co-workers prepared a series of Au-supported catalysts which could synthesize the cyclic carbonates from olefins and CO2 [60,61]. They applied the Au/SiO2-ZnBr2/Bu4NBr catalyst system for the preparation of cyclic styrene carbonate from styrene and CO2 [60], which could give a 42% selectivity towards cyclic styrene carbonate under mild reaction conditions (80 °C, 1 MPa CO2). Moreover, it was found that cumene hydroperoxide was a better oxidant than Bu4NBr, indicating that choosing the suitable oxidation may have a significant effect on the yield of styrene carbonate. Using a strong basic resin R201 as support to replace the SiO2, a 51% yield of styrene carbonate was obtained in the Au/R201 catalyst with anhydrous TBHP as an oxidant. The different catalytic activity between the Au/SiO2-ZnBr2/Bu4NBr and Au/R201 catalyst may originate from the different particle sizes of Au nanoparticles, which may play important role in the epoxidation reaction. The work function of Au species with different atomicity is strongly dependent on the particle size [58]. The appropriate size of Au nanoparticles could enhance the selectivity of epoxides during the epoxidation reaction [62]. Recently, Zhang et al. fabricated a double-shell microencapsulated nanoreactor as a bifunctional catalyst for the one-pot synthesis of cyclic carbonates from olefins [63]. Au nanoparticels (NPs) were confined between an inter shell of nitrogen-doped porous carbon rich in Zn single-atom catalytic sites (SACs) and an outer shell of mesoporous SiO2 (noted as Zn-N-C/Au@mSiO2, Scheme 16). The synergistic catalysis of the ultrafine Au NPs and Zn SACs showed high catalytic efficiency and tandem catalytic function for the direct oxidative carboxylation from olefins and CO2.

3.4. Others

Frustrated Lewis pairs (FLPs) are combinations of Lewis acids and Lewis bases that are prevented from forming a stable adduct by steric or electronic hindrance [64]. Due to the sterically hindered Lewis acid and Lewis base, the FLPs have shown the ability in the activation and transformation of inert small molecules such as H2 and CO2 [65,66,67]. Qu et al. reported a defect-enriched CeO2 with constructed interfacial FLPs (two adjacent Ce3+···O2), which could activate CO2 via the interactions between C/Lewis basic lattice O2 and the two O atoms in CO2/two adjacent Lewis acidic Ce3+ ions [68]. When increasing the surface of CeO2, an 88% yield of styrene carbonate can be obtained catalyzed by porous nanorods of CeO2 (PN-CeO2) under 80 °C and 24 h. The excellent performance of the selective tandem conversion of olefins and CO2 to cyclic carbonates may derive from the efficient CO2 activation for cycloaddition and weakening of the adsorption of in situ-produced epoxides to suppress their hydrolysis (Scheme 17).
Hou et al. prepared the 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD) functionalized phosphonate salts with different cations and applied them in the synthesis of styrene carbonate from styrene and CO2 [69]. Experiment and DFT calculation demonstrated that TBA-TBD-P (TBA: tetrabutylammonium salts) could capture CO2 by organic phosphate anion sites and the adsorption energy of TBA-TBD-P at phosphonate was ~2 Kcal/mol lower than that of the TBA-TBD-P with CO2 being absorbed in nitrogen atom of TBD ring (Scheme 18). The formation of carboxyphosphate intermediate may be a crucial active intermediate species, which plays a key role in CO2 activation and conversion. The TBA-TBD-P/ZnBr2 catalyst system showed a 69.6% yield of cyclic carbonates under mild conditions (0.1 MPa CO2, 80 °C) with TBHP as an oxidant.
Considering the utilization of “carbon-neutral” resources, using a biomass-based catalyst for the synthesis of styrene carbonate from styrene and CO2 may be a green and sustainable method. Due to the abundant hydroxyl groups in the biomass-based catalyst (such as cellulose, chitosan and lignin), these biomass-based catalysts have been applied in the CO2 cycloaddition reaction [5,70,71,72]. Although the biomass-based catalysts exhibit excellent catalytic activity for the CO2 cycloaddition reaction, how to introduce the epoxidation active sites in the biomass-based catalysts and achieve the direct synthesis of cyclic carbonates remains a big challenge. Jain et al. Fabricated magnetically separable chitosan (MCS) modified with Co(II) acetylacetonate and triphenylphosphonium bromide catalyst (Co(acac)2-QPB@MCS) for the direct synthesis of cyclic carbonates from olefins and CO2 with O2 as an oxidant (Scheme 19) [73]. They used chitosan as a support platform to combine the epoxidation active site (cobalt(II) acetylacetonate complex, Co(acac)2) and the cycloaddition active site (quaternary triphenylphosphonium bromide, P+Ph3Br), which could achieve 73.1% yield of styrene carbonate. However, the yield of styrene carbonate was lower when using QPB@MCS alone or a physical mixture of Co(acac)2@MCS and QPB@MCS (1:1) as a catalyst, which indicated that the covalent grafting of Co(acac)2 and P+Ph3Br moieties to MCS showed obvious synergistic catalytic effect during the one-pot reaction.
Table 2 summarizes the heterogeneous catalytic system for the conversion of olefins and CO2 to cyclic carbonates. It is found that most of the heterogeneous catalyst with good yields and selectivity towards the cyclic carbonates usually includes two important parts: (i) a high selectivity epoxidation active site for the oxidation of olefins, (b) the Lewis acid and base sites for the CO2 cycloaddition reaction (Table 2, entries 10–17). In addition, the suitable catalyst support with high surface area, controllable pore size, and the easy diffusion of substrate molecules can accelerate the direct oxidative carboxylation of olefins with CO2 (Table 2, entries 4–9). Interestingly, some of the heterogeneous catalysts integrating the different catalytic moieties show a higher catalytic activity than the physical mixture of the different catalytic moieties, which indicates that the stereoconfinment effect and the distribution of the different active sites can efficiently stabilize and transform the intermediates and promote the reaction (Table 2, entries 10–17).
The reaction parameters such as oxidant, temperature and CO2 pressure can significantly affect the activity and selectivity for both the epoxidation and cycloaddition reaction during the oxidative carboxylation of styrene. Usually, the organic oxidants (such as TBHP, CHP and PhIO) exhibit higher activity than H2O2 or O2. For example, the oxidation of TBHP can enhance the interaction with the substrate and react with the catalyst forming the main driver of epoxidation [52]. Besides, the high temperature is a disadvantage in improving the selectivity of the epoxidation process. Hence, the low temperature is usually used for the epoxidation reaction (Table 2, entries 1–3, 13 and 19). However, the high temperature benefits the activation of CO2 during the CO2 cycloaddition reaction [29]. The high temperature of the direct oxidative carboxylation of olefins may lead to the over-oxidation of styrene oxide and polymerization of styrene [57]. It is found that the higher cyclic carbonates can be obtained with a higher initial CO2 pressure (Table 2, entries 8 and 9). The higher CO2 pressure can enhance the contact between CO2 molecules and substrate, so it is necessary to develop the catalyst with a higher CO2 capture capacity.

4. Conclusions and Perspectives

The direct oxidative carboxylation of olefins with CO2 is an economic and sustainable strategy for the synthesis of cyclic carbonates. However, the one-pot reaction usually contains two different reactions: (i) the epoxidation reaction; (ii) the CO2 cycloaddition reaction and the optimal reaction conditions of these two reactions are obvious different. For example, the optimal reaction conditions of the epoxidation reaction are 50 °C with TBHP as an oxidant (Table 2, entry 19). However, the optimal reaction conditions of the CO2 cycloaddition reaction can be obtained at 150 °C with 40 bar CO2. Hence, how to combine these two different reactions into a one-pot reaction may be a huge challenge. Although utilization of two different catalysts can achieve the one-pot reaction, the low conversion and selectivity may prevent the large-scale reaction (Table 2, entries 1, 2 and 4–5). As a result, the integration of two different catalytic sites into one catalyst may be an alternative method to avoid the problems mentioned above (Table 2, entries 17 and 23).
In this review, we carefully investigate the method in the design of the homogeneous and heterogeneous catalytic system for the direct oxidative carboxylation of olefins with CO2. Considering the oxidation reaction is the first step, the effectively transferring active [O] atom from oxidation to olefins via the epoxidation active site may play a significant role in the selectivity of the epoxides (Scheme 3 and Scheme 6). Compared with the TBHP, the use of H2O2 or O2 as oxidation is safe and environmentally sustainable. However, using H2O2 or O2 as oxidation needs well-designed epoxidation active sites such as POMs or Mn-porphyrin (Table 2, entries 10 and 17). The effective activation and transformation of active [O] species are important for the epoxidation reactions. For example, the high selectivity of epoxidation reaction with H2O2 as oxidation must activate H2O2 without radical production (Table 2, entry 3 and Scheme 8). The heterolytically cleaved the O-O band was a benefit for the epoxidation reaction. For the CO2 cycloaddition reaction, the catalyst should both have the Lewis acid and Lewis base sites, which could activate the epoxides and CO2 respectively (Table 2, entry 21). However, CO2 is hard to convert due to its thermodynamically stable and kinetically inert. As a result, a high temperature or CO2 pressure is used in the CO2 cycloaddition reaction. However, a high temperature may affect the oxidation reaction in the first step because some by-products (acid or aldehyde) can be found at a higher temperature (Table 2, entries 1–3, 13 and 19). Therefore, the integration of the suitable epoxidation active site and cycloaddition active site into a porous material could be a benefit for the direct oxidative carboxylation of olefins with CO2.

Author Contributions

Introduction and homogeneous catalytic system, T.Y.; heterogeneous catalytic system, Z.W.; Writing—review & editing, L.G.; supervision, Funding acquisition, R.Z. and H.Z.; supervision, J.C.; supervision, Funding acquisition and Writing—review & editing, X.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 21902053 and 22002114), the Research Project supported by Foundation of State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (Grant No. 2021-K73). This work was supported by scientific research project of the Education Department of Hubei Province (Q20201701).

Acknowledgments

All authors are thankful to their representative universities/institutes for literature services.

Conflicts of Interest

The authors report no conflict of interest in any capacity, i.e., competing or financial.

References

  1. Song, Q.W.; Zhou, Z.H.; He, L.N. Efficient, selective and sustainable catalysis of carbon dioxide. Green Chem. 2017, 19, 3707–3728. [Google Scholar] [CrossRef]
  2. Chen, H.; Cui, H.; Lv, Y.; Liu, P.; Hao, F.; Xiong, W.; Luo, H. CO2 hydrogenation to methanol over Cu/ZnO/ZrO2 catalysts: Effects of ZnO morphology and oxygen vacancy. Fuel 2022, 314, 123035–123047. [Google Scholar] [CrossRef]
  3. Dong, X.; Liu, X.; Chen, Y.; Zhang, M. Screening of bimetallic M-Cu-BTC MOFs for CO2 activation and mechanistic study of CO2 hydrogenation to formic acid: A DFT study. J. CO2 Util. 2018, 24, 64–72. [Google Scholar] [CrossRef]
  4. Shen, X.; Wang, W.; Wang, Q.; Liu, J.; Huang, F.; Sun, C.; Yang, C.; Chen, D. Mechanism of iron complexes catalyzed in the N-formylation of amines with CO2 and H2: The superior performance of N-H ligand methylated complexes. Phys. Chem. Chem. Phys. 2021, 23, 16675–16689. [Google Scholar] [CrossRef] [PubMed]
  5. Guo, L.; Xiong, Y.; Zhang, R.; Zhan, H.; Chang, D.; Yi, L.; Chen, J.; Wu, X. Catalytic coupling of CO2 and epoxides by lignin-based catalysts: A combined experimental and theoretical study. J. CO2 Util. 2022, 56, 101863–101870. [Google Scholar] [CrossRef]
  6. 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]
  7. Calmanti, R.; Selva, M.; Perosa, A. Tandem catalysis: One-pot synthesis of cyclic organic carbonates from olefins and carbon dioxide. Green Chem. 2021, 23, 1921–1941. [Google Scholar] [CrossRef]
  8. Liu, J.; Yang, G.; Liu, Y.; Wu, D.; Hu, X.; Zhang, Z. Metal-free imidazolium hydrogen carbonate ionic liquids as bifunctional catalysts for the one-pot synthesis of cyclic carbonates from olefins and CO2. Green Chem. 2019, 21, 3834–3838. [Google Scholar] [CrossRef]
  9. Ke, S.C.; Luo, T.T.; Chang, G.G.; Huang, K.X.; Li, J.X.; Ma, X.C.; Wu, J.; Chen, J.; Yang, X.Y. Spatially ordered arrangement of multifunctional sites at molecule level in a single catalyst for tandem synthesis of cyclic carbonates. Inorg. Chem. 2020, 59, 1736–1745. [Google Scholar] [CrossRef]
  10. Qiao, Y.; Hou, Z.; Li, H.; Hu, Y.; Feng, B.; Wang, X.; Hua, L.; Huang, Q. Polyoxometalate-based protic alkylimidazolium salts as reaction-induced phase-separation catalysts for olefin epoxidation. Green Chem. 2009, 11, 1955–1960. [Google Scholar] [CrossRef]
  11. Santos, F.M.; Magina, S.P.; Nogueira, H.I.S.; Cavaleiro, A.M.V. Synthesis and characterization of metal-substituted tetraalkylphosphonium polyoxometalate ionic liquids. New J. Chem. 2016, 40, 945–953. [Google Scholar] [CrossRef]
  12. Dai, W.; Yang, W.; Zhang, Y.; Wang, D.; Luo, X.; Tu, X. Novel isothiouronium ionic liquid as efficient catalysts for the synthesis of cyclic carbonates from CO2 and epoxides. J. CO2 Util. 2017, 17, 256–262. [Google Scholar] [CrossRef]
  13. Girard, A.L.; Simon, N.; Zanatta, M.; Marmitt, S.; Gonçalves, P.; Dupont, J. Insights on recyclable catalytic system composed of task-specific ionic liquids for the chemical fixation of carbon dioxide. Green Chem. 2014, 16, 2815–2825. [Google Scholar] [CrossRef]
  14. Yao, H.; Richardson, D.E. Epoxidation of alkenes with bicarbonate-activated hydrogen peroxide. JACS 2000, 122, 3220–3221. [Google Scholar] [CrossRef]
  15. Calmanti, R.; Selva, M.; Perosa, A. Direct oxidative carboxylation of terminal olefins to cyclic carbonates by tungstate assisted-tandem catalysis. Green Chem. 2021, 23, 7609–7619. [Google Scholar] [CrossRef]
  16. Zima, A.M.; Lyakin, O.Y.; Ottenbacher, R.V.; Bryliakov, K.P.; Talsi, E.P. Dramatic effect of carboxylic acid on the electronic structure of the active species in Fe(PDP)-catalyzed asymmetric epoxidation. ACS Catal. 2016, 6, 5399–5404. [Google Scholar] [CrossRef]
  17. Papastergiou, M.; Stamatis, A.; Simaioforidou, A.; Louloudi, M. Bio-inspired Mn-catalysts immobilized on silica surface: The influence of the ligand synthesis on catalytic behavior. Catal. Commun. 2018, 108, 33–40. [Google Scholar] [CrossRef]
  18. Bai, D.; Jing, H. Aerobic oxidative carboxylation of olefins with metalloporphyrin catalysts. Green Chem. 2010, 12, 39–41. [Google Scholar] [CrossRef]
  19. Ramidi, P.; Felton, C.M.; Subedi, B.P.; Zhou, H.; Tian, Z.R.; Gartia, Y.; Pierce, B.S.; Ghosh, A. Synthesis and characterization of manganese(III) and high-valent manganese-oxo complexes and their roles in conversion of alkenes to cyclic carbonates. J. CO2 Util. 2015, 9, 48–57. [Google Scholar] [CrossRef]
  20. Han, F.; Li, H.; Zhuang, H.; Hou, Q.; Yang, Q.; Zhang, B.; Miao, C. Direct synthesis of cyclic carbonates from olefins and CO2: Single- or multi-component catalytic systems via epoxide or halohydrin intermediate. J. CO2 Util. 2021, 53, 101742–101758. [Google Scholar] [CrossRef]
  21. Chen, F.; Dong, T.; Xu, T.; Li, X.; Hu, C. Direct synthesis of cyclic carbonates from olefins and CO2 catalyzed by a MoO2(acac)2-quaternary ammonium salt system. Green Chem. 2011, 13, 2518–2524. [Google Scholar] [CrossRef]
  22. Xing, J.; Yuan, D.; Liu, H.; Tong, Y.; Xu, Y.; Liu, Z. Synthesis of TS-1 zeolites from a polymer containing titanium and silicon. J. Mater. Chem. A 2021, 9, 6205–6213. [Google Scholar] [CrossRef]
  23. Chen, Z.; Zhang, L.; Yu, Y.; Liu, D.; Fang, N.; Lin, Y.; Xu, D.; Li, F.; Liu, Y.; He, M. TS-1 zeolite with homogeneous distribution of Ti atoms in the framework: Synthesis, crystallization mechanism and its catalytic performance. J. Catal. 2021, 404, 990–998. [Google Scholar] [CrossRef]
  24. You, Q.; Wang, X.; Wu, Y.; Bi, C.; Yang, X.; Sun, M.; Zhang, J.; Hao, Q.; Chen, H.; Ma, X. Hierarchical Ti-beta with a three-dimensional ordered mesoporosity for catalytic epoxidation of bulky cyclic olefins. New J. Chem. 2021, 45, 10303–10314. [Google Scholar] [CrossRef]
  25. Zhang, J.; Jin, S.; Deng, D.; Liu, W.; Tao, G.; Luo, Q.; Sun, H.; Yang, W. Insight into the formation of framework titanium species during acid treatment of MWW-type titanosilicate and the effect of framework titanium state on olefin epoxidation. Microporous Mesoporous Mater. 2021, 314, 110862. [Google Scholar] [CrossRef]
  26. Bregante, D.T.; Tan, J.Z.; Schultz, R.L.; Ayla, E.Z.; Potts, D.S.; Torres, C.; Flaherty, D.W. Catalytic consequences of oxidant, alkene, and pore structures on alkene epoxidations within titanium silicates. ACS Catal. 2020, 10, 10169–10184. [Google Scholar] [CrossRef]
  27. Solé-Daura, A.; Zhang, T.; Fouilloux, H.; Robert, C.; Thomas, C.M.; Chamoreau, L.-M.; Carbó, J.J.; Proust, A.; Guillemot, G.; Poblet, J.M. Catalyst design for alkene epoxidation by molecular analogues of heterogeneous titanium-silicalite catalysts. ACS Catal. 2020, 10, 4737–4750. [Google Scholar] [CrossRef]
  28. Srivastava, R.; Srinivas, D.; Ratnasamy, P. Synthesis of polycarbonate precursors over titanosilicate molecular sieves. Catal. Lett. 2003, 91, 133–139. [Google Scholar] [CrossRef]
  29. Zhang, J.; Liu, Y.; Li, N.; Wu, H.; Li, X.; Xie, W.; Zhao, Z.; Wu, P.; He, M. Synthesis of propylene carbonate on a bifunctional titanosilicate modified with quaternary ammonium halides. Chin. J. Catal. 2008, 29, 589–591. [Google Scholar] [CrossRef]
  30. Maksimchuk, N.V.; Ivanchikova, I.D.; Ayupov, A.B.; Kholdeeva, O.A. One-step solvent-free synthesis of cyclic carbonates by oxidative carboxylation of styrenes over a recyclable Ti-containing catalyst. Appl. Catal. B Environ. 2016, 181, 363–370. [Google Scholar] [CrossRef]
  31. Evangelisti, C.; Guidotti, M.; Tiozzo, C.; Psaro, R.; Maksimchuk, N.; Ivanchikova, I.; Shmakov, A.N.; Kholdeeva, O. Titanium-silica catalyst derived from defined metallic titanium cluster precursor: Synthesis and catalytic properties in selective oxidations. Inorg. Chim. Acta 2018, 470, 393–401. [Google Scholar] [CrossRef]
  32. Zhao, D.; Feng, J.; Huo, Q.; Melosh, N.; Fredrickson, G.H.; Chmelka, B.F.; Stucky, G.D. Triblock copolymer syntheses of mesoporous silica with periodic 50 to 300 angstrom pores. Science 1998, 279, 548–552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Verma, P.; Kuwahara, Y.; Mori, K.; Raja, R.; Yamashita, H. Functionalized mesoporous SBA-15 silica: Recent trends and catalytic applications. Nanoscale 2020, 12, 11333–11363. [Google Scholar] [CrossRef] [PubMed]
  34. Long, G.; Su, K.; Dong, H.; Zhao, T.; Yang, C.; Liu, F.; Hu, X. Straightforward construction of amino-functionalized ILs@SBA-15 catalysts via mechanochemical grafting for one-pot synthesis of cyclic carbonates from aromatic olefins and CO2. J. CO2 Util. 2022, 59, 101962–101969. [Google Scholar] [CrossRef]
  35. Maity, A.; Polshettiwar, V. Dendritic fibrous nanosilica for catalysis, energy harvesting, carbon dioxide mitigation, drug delivery, and sensing. ChemSusChem 2017, 10, 3866–3913. [Google Scholar] [CrossRef]
  36. Qiu, J.; Yu, L.; Ni, J.; Fei, Z.; Li, W.; Sadeghzadeh, S.M. Palladium–salen-bridged ionic networks immobilized on magnetic dendritic silica fibers for the synthesis of cyclic carbonates by oxidative carboxylation. New J. Chem. 2020, 44, 1269–1277. [Google Scholar] [CrossRef]
  37. Abassian, M.; Zhiani, R.; Motavalizadehkakhky, A.; Eshghi, H.; Mehrzad, J. A new class of organoplatinum-based DFNS for the production of cyclic carbonates from olefins and CO2. RSC Adv. 2020, 10, 15044–15051. [Google Scholar] [CrossRef] [Green Version]
  38. Li, B.; Wen, H.M.; Zhou, W.; Chen, B. Porous metal-organic frameworks for gas storage and separation: What, How, and Why? J. Phys. Chem. Lett. 2014, 5, 3468–3479. [Google Scholar] [CrossRef] [PubMed]
  39. Wu, Y.; Qiu, X.; Liang, F.; Zhang, Q.; Koo, A.; Dai, Y.; Lei, Y.; Sun, X. A metal-organic framework-derived bifunctional catalyst for hybrid sodium-air batteries. Appl. Catal. B: Environ. 2019, 241, 407–414. [Google Scholar] [CrossRef]
  40. Huang, Y.B.; Liang, J.; Wang, X.S.; Cao, R. Multifunctional metal-organic framework catalysts: Synergistic catalysis and tandem reactions. Chem. Soc. Rev. 2017, 46, 126–157. [Google Scholar] [CrossRef]
  41. Song, F.; Wang, C.; Falkowski, J.M.; Ma, L.; Lin, W. Isoreticular chiral metal-organic frameworks for asymmetric alkene epoxidation: Tuning catalytic activity by controlling framework catenation and varying open channel sizes. JACS 2010, 132, 15390–15398. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, J.; Biradar, A.V.; Pramanik, S.; Emge, T.J.; Asefa, T.; Li, J. A new layered metal-organic framework as a promising heterogeneous catalyst for olefin epoxidation reactions. Chem. Commun. 2012, 48, 6541–6543. [Google Scholar] [CrossRef] [PubMed]
  43. Stubbs, A.W.; Braglia, L.; Borfecchia, E.; Meyer, R.J.; Román- Leshkov, Y.; Lamberti, C.; Dincă, M. Selective catalytic olefin epoxidation with MnII-exchanged MOF-5. ACS Catal. 2018, 8, 596–601. [Google Scholar] [CrossRef] [Green Version]
  44. Gao, W.Y.; Chen, Y.; Niu, Y.; Williams, K.; Cash, L.; Perez, P.J.; Wojtas, L.; Cai, J.; Chen, Y.S.; Ma, S. Crystal engineering of an nbo topology metal-organic framework for chemical fixation of CO2 under ambient conditions. Angew. Chem. Int. Edit. 2014, 53, 2615–2619. [Google Scholar] [CrossRef] [PubMed]
  45. Song, J.; Zhang, Z.; Hu, S.; Wu, T.; Jiang, T.; Han, B. MOF-5/n-Bu4NBr: An efficient catalyst system for the synthesis of cyclic carbonates from epoxides and CO2 under mild conditions. Green Chem. 2009, 11, 1031–1036. [Google Scholar] [CrossRef]
  46. Jin, H.G.; Chen, F.; Zhang, H.; Xu, W.; Wang, Y.; Fan, J.; Chao, Z.S. Postmetalation of a new porphyrin ligand-based metal–organic framework for catalytic oxidative carboxylation of olefins. J. Mater. Sci. 2020, 55, 16184–16196. [Google Scholar] [CrossRef]
  47. Han, Q.; Qi, B.; Ren, W.; He, C.; Niu, J.; Duan, C. Polyoxometalate-based homochiral metal-organic frameworks for tandem asymmetric transformation of cyclic carbonates from olefins. Nature Commun. 2015, 6, 10007–100014. [Google Scholar] [CrossRef]
  48. Mizuno, N.; Yamaguchi, K.; Kamata, K. Epoxidation of olefins with hydrogen peroxide catalyzed by polyoxometalates. Coord. Chem. Rev. 2005, 249, 1944–1956. [Google Scholar] [CrossRef]
  49. Melchiorre, P.; Marigo, M.; Carlone, A.; Bartoli, G. Asymmetric aminocatalysis-gold rush in organic chemistry. Angew. Chem. Int. Edit. 2008, 47, 6138–6171. [Google Scholar] [CrossRef]
  50. Berijani, K.; Hosseini-Monfared, H. Aerobic enantioselective epoxidation of olefins mediated by an easy-to-prepare recyclable manganese-porphyrin. Mol. Catal. 2017, 433, 136–144. [Google Scholar] [CrossRef]
  51. Sharma, N.; Dhankhar, S.S.; Kumar, S.; Kumar, T.J.D.; Nagaraja, C.M. Rational design of a 3D MnII-metal-organic framework based on a nonmetallated porphyrin linker for selective capture of CO2 and one-pot synthesis of styrene carbonates. Eur. J. Inorg. Chem. 2018, 24, 16662–16669. [Google Scholar] [CrossRef]
  52. Shi, Z.; Niu, G.; Han, Q.; Shi, X.; Li, M. A molybdate-incorporated cooperative catalyst: High efficiency in the assisted tandem catalytic synthesis of cyclic carbonates from CO2 and olefins. Mol. Catal. 2018, 461, 10–18. [Google Scholar] [CrossRef]
  53. Auroux, A.; Gervasini, A. Microcalorimetric study of the acidity and basicity of metal oxide surfaces. J. Phys. Chem. 1990, 94, 6371–6379. [Google Scholar] [CrossRef]
  54. Li, S.J.; Zhou, Y.T.; Kang, X.; Liu, D.X.; Gu, L.; Zhang, Q.H.; Yan, J.M.; Jiang, Q. A simple and effective principle for a rational design of heterogeneous catalysts for dehydrogenation of formic acid. Adv. Mater. 2019, 31, 1806781–1806787. [Google Scholar] [CrossRef] [PubMed]
  55. Fracaroli, A.M.; Furukawa, H.; Suzuki, M.; Dodd, M.; Okajima, S.; Gándara, F.; Reimer, J.A.; Yaghi, O.M. Metal-organic frameworks with precisely designed interior for carbon dioxide capture in the presence of water. JACS 2014, 136, 8863–8866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Nguyen, H.T.D.; Tran, Y.B.N.; Nguyen, H.N.; Nguyen, T.C.; Gándara, F.; Nguyen, P.T.K. A series of metal-organic frameworks for selective CO2 capture and catalytic oxidative carboxylation of olefins. Inorg. Chem. 2018, 57, 13772–13782. [Google Scholar] [CrossRef] [Green Version]
  57. Yu, K.; Puthiaraj, P.; Ahn, W.S. One-pot catalytic transformation of olefins into cyclic carbonates over an imidazolium bromide-functionalized Mn(III)-porphyrin metal-organic framework. Appl. Catal. B Environ. 2020, 273, 119059–119072. [Google Scholar] [CrossRef]
  58. Liu, L.; Corma, A. Metal catalysts for heterogeneous catalysis: From single atoms to nanoclusters and nanoparticles. Chem. Rev. 2018, 118, 4981–5079. [Google Scholar] [CrossRef] [Green Version]
  59. Yang, X.F.; Wang, A.; Qiao, B.; Li, J.; Liu, J.; Zhang, T. Single-atom catalysts: A new frontier in heterogeneous catalysis. Acc. Chem. Res. 2013, 46, 1740–1748. [Google Scholar] [CrossRef]
  60. Sun, J.; Fujita, S.; Zhao, F.; Hasegawa, M.; Arai, M. A direct synthesis of styrene carbonate from styrene with the Au/SiO2-ZnBr2/Bu4NBr catalyst system. J. Catal. 2005, 230, 398–405. [Google Scholar] [CrossRef]
  61. Xiang, D.; Liu, X.; Sun, J.; Xiao, F.S.; Sun, J. A novel route for synthesis of styrene carbonate using styrene and CO2 as substrates over basic resin R201 supported Au catalyst. Catal. Today 2009, 148, 383–388. [Google Scholar] [CrossRef]
  62. Guo, L.; Zhang, R.; Chen, C.; Chen, J.; Zhao, X.; Chen, A.; Liu, X.; Xiu, Y.; Hou, Z. Gold nanoparticles embedded in silica hollow nanospheres induced by compressed CO2 as an efficient catalyst for selective oxidation. Phys. Chem. Chem. Phys. 2015, 17, 6406–6414. [Google Scholar] [CrossRef] [PubMed]
  63. Hou, Y.; Wang, X.; Guo, Y.; Zhang, X. Double-shell microcapsules with spatially arranged Au nanoparticles and single Zn atoms for tandem synthesis of cyclic carbonates. Nanoscale 2021, 13, 18695–18701. [Google Scholar] [CrossRef] [PubMed]
  64. Stephan, D.W.; Erker, G. Frustrated lewis pair chemistry: Development and perspectives. Angew. Chem. Int. Edit. 2015, 54, 6400–6441. [Google Scholar] [CrossRef]
  65. Stephan Douglas, W. The broadening reach of frustrated Lewis pair chemistry. Science 2016, 354, 1248–1257. [Google Scholar] [CrossRef]
  66. Wang, X.; Kehr, G.; Daniliuc, C.G.; Erker, G. Internal adduct formation of active intramolecular C4-bridged frustrated phosphane/borane lewis pairs. J. Am. Chem. Soc. 2014, 136, 3293–3303. [Google Scholar] [CrossRef]
  67. Chen, L.; Liu, R.; Yan, Q. Polymer meets frustrated lewis pair: Second-generation CO2-responsive nanosystem for sustainable CO2 conversion. Angew. Chem. Int. Edit. 2018, 57, 9336–9340. [Google Scholar] [CrossRef]
  68. Zhang, S.; Xia, Z.; Zou, Y.; Cao, F.; Liu, Y.; Ma, Y.; Qu, Y. Interfacial frustrated lewis pairs of CeO2 activate CO2 for selective tandem transformation of olefins and CO2 into cyclic carbonates. JACS 2019, 141, 11353–11357. [Google Scholar] [CrossRef]
  69. Fang, J.; Li, K.; Wang, Z.; Li, D.; Ma, Y.; Gong, X.; Hou, Z. Direct oxidative carboxylation of olefins into cyclic carbonates at ambient pressure. J. CO2 Util. 2020, 40, 101204–101214. [Google Scholar] [CrossRef]
  70. Guo, L.; Dou, R.; Wu, Y.; Zhang, R.; Wang, L.; Wang, Y.; Gong, Z.; Chen, J.; Wu, X. From lignin waste to efficient catalyst: Illuminating the impact of lignin structure on catalytic activity of cycloaddition reaction. ACS Sustain. Chem. Eng. 2019, 7, 16585–16594. [Google Scholar] [CrossRef]
  71. Chen, S.; Pudukudy, M.; Yue, Z.; Zhang, H.; Zhi, Y.; Ni, Y.; Shan, S.; Jia, Q. Nonmetal schiff-base complex-anchored cellulose as a novel and reusable catalyst for the solvent-free ring-opening addition of CO2 with epoxides. Ind. Eng. Chem. Res. 2019, 58, 17255–17265. [Google Scholar] [CrossRef]
  72. Wu, X.; Wang, M.; Xie, Y.; Chen, C.; Li, K.; Yuan, M.; Zhao, X.; Hou, Z. Carboxymethyl cellulose supported ionic liquid as a heterogeneous catalyst for the cycloaddition of CO2 to cyclic carbonate. Appl. Catal. A Gen. 2016, 519, 146–154. [Google Scholar] [CrossRef]
  73. Kumar, S.; Singhal, N.; Singh, R.K.; Gupta, P.; Singh, R.; Jain, S.L. Dual catalysis with magnetic chitosan: Direct synthesis of cyclic carbonates from olefins with carbon dioxide using isobutyraldehyde as the sacrificial reductant. Dalton Trans. 2015, 44, 11860–11866. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthesis of cyclic carbonate from olefins and CO2.
Scheme 1. Synthesis of cyclic carbonate from olefins and CO2.
Catalysts 12 00563 sch001
Scheme 2. Direct synthesis of cyclic carbonate from olefin and CO2.
Scheme 2. Direct synthesis of cyclic carbonate from olefin and CO2.
Catalysts 12 00563 sch002
Scheme 3. The bifunctional catalyst of [CnCmIm][HCO3].
Scheme 3. The bifunctional catalyst of [CnCmIm][HCO3].
Catalysts 12 00563 sch003
Scheme 4. Schematic representation of the main mechanistic pathways for the formation of cyclic carbonate. A+ represents a Lewis acid or Brønsted acid.
Scheme 4. Schematic representation of the main mechanistic pathways for the formation of cyclic carbonate. A+ represents a Lewis acid or Brønsted acid.
Catalysts 12 00563 sch004
Scheme 5. Proposed mechanism of aerobic oxidative carboxylation of olefins.
Scheme 5. Proposed mechanism of aerobic oxidative carboxylation of olefins.
Catalysts 12 00563 sch005
Scheme 6. Structure of manganese (III) complex of an amido-amine ligand and high-valent manganese-oxo intermediate.
Scheme 6. Structure of manganese (III) complex of an amido-amine ligand and high-valent manganese-oxo intermediate.
Catalysts 12 00563 sch006
Scheme 7. The proposed mechanism of the oxidative carboxylation of olefins catalyzed by the MoO2(acac)2/TBHP-TBAB system.
Scheme 7. The proposed mechanism of the oxidative carboxylation of olefins catalyzed by the MoO2(acac)2/TBHP-TBAB system.
Catalysts 12 00563 sch007
Scheme 8. Ti active sites and proposed mechanism for the epoxidation of olefins.
Scheme 8. Ti active sites and proposed mechanism for the epoxidation of olefins.
Catalysts 12 00563 sch008
Scheme 9. Oxidative carboxylation of styrene over Ti-MMM-E.
Scheme 9. Oxidative carboxylation of styrene over Ti-MMM-E.
Catalysts 12 00563 sch009
Scheme 10. Preparation sequence for TinOx-MCM-41 and the oxidative carboxylation of styrene over titanium-grafted mesoporous silica catalyst.
Scheme 10. Preparation sequence for TinOx-MCM-41 and the oxidative carboxylation of styrene over titanium-grafted mesoporous silica catalyst.
Catalysts 12 00563 sch010
Scheme 11. Possible mechanism for the oxidative carboxylation of styrene and CO2 over N(CH3)2-[Im][Br]@SBA-15.
Scheme 11. Possible mechanism for the oxidative carboxylation of styrene and CO2 over N(CH3)2-[Im][Br]@SBA-15.
Catalysts 12 00563 sch011
Scheme 12. Schematic illustration of the preparation of the FeNi3/DFNS/salen/Pd(II) MNPs.
Scheme 12. Schematic illustration of the preparation of the FeNi3/DFNS/salen/Pd(II) MNPs.
Catalysts 12 00563 sch012
Scheme 13. The schematic representation of tandem catalysis for the asymmetric cyclic carbonate transformation fromolefins and carbon dioxide.
Scheme 13. The schematic representation of tandem catalysis for the asymmetric cyclic carbonate transformation fromolefins and carbon dioxide.
Catalysts 12 00563 sch013
Scheme 14. Proposed mechanism for the oxidative carboxylation of styrene and CO2 catalyzed by MOF-590.
Scheme 14. Proposed mechanism for the oxidative carboxylation of styrene and CO2 catalyzed by MOF-590.
Catalysts 12 00563 sch014
Scheme 15. The synthetic procedure of ImBr-MOF-545(Mn) by post-synthetic Mn metalation followed by ImBr ionic liquid functionalization.
Scheme 15. The synthetic procedure of ImBr-MOF-545(Mn) by post-synthetic Mn metalation followed by ImBr ionic liquid functionalization.
Catalysts 12 00563 sch015
Scheme 16. Synthesis procedure of Zn-N-C/Au@mSiO2.
Scheme 16. Synthesis procedure of Zn-N-C/Au@mSiO2.
Catalysts 12 00563 sch016
Scheme 17. Adsorption energy of propylene oxide on defective CeO2(110).
Scheme 17. Adsorption energy of propylene oxide on defective CeO2(110).
Catalysts 12 00563 sch017
Scheme 18. The reaction procedure of TBA-TBD-P with CO2 and calculated two different ways of activating CO2 on TBA-TBD-P.
Scheme 18. The reaction procedure of TBA-TBD-P with CO2 and calculated two different ways of activating CO2 on TBA-TBD-P.
Catalysts 12 00563 sch018
Scheme 19. Synthesis procedure of Co(acac)2-QPB@MCS.
Scheme 19. Synthesis procedure of Co(acac)2-QPB@MCS.
Catalysts 12 00563 sch019
Table 1. The direct oxidative carboxylation of olefins catalyzed by the homogeneous catalytic system.
Table 1. The direct oxidative carboxylation of olefins catalyzed by the homogeneous catalytic system.
EntryCatalystCo-CatalystOxidantCO2
(bar)
T
(°C)
t
(h)
Conv. a
(%)
Sel. b
(%)
Yield c
(%)
Ref.
1BMImBr--TBHP51506904036[13]
2[C1C4Im][HCO3]--TBHP20653091.282.375.1[8]
3 d[N8,8,8,1]2[WO4][N4,4,4,4]IH2O21853/5----94[15]
4[Ru(TPP)(O)2]TBAIO2113048----89[18]
5Fe(TPP)ClTBAIO2113048----3.6[18]
6Ru(Salen)(PPh3)2TBAIO2113048----40.7[18]
7Mn(III)-complexTBABTBHP171006----43[19]
8 eMoO2(acac)2TBABTBHP30100/1401/1----68[21]
a Conversion of styrene. b Selectivity towards styrene carbonate. c Yield of styrene carbonate. d substrate: 1-decene, the reaction performed at 85 °C for 3 h followed by the raw addition of [N4,4,4,4]I and CO2 without any intermediate work-up. The reaction was further performed for 5 h. e 100 °C and 1 h for epoxidation reaction and 140 °C and 1 h for cycloaddition reaction.
Table 2. The direct oxidative carboxylation of olefins catalyzed by the heterogeneous catalytic system.
Table 2. The direct oxidative carboxylation of olefins catalyzed by the heterogeneous catalytic system.
EntryCatalystCo-CatalystOxidantCO2
(bar)
T
(°C)
t
(h)
Conv. a
(%)
Sel. b
(%)
Yield c
(%)
Ref.
1 dTS-1DMAPH2O26.960/12024/454.655.6--[28]
2 dTiMCM-41DMAPTBHP6.960/12024/444.083.4--[28]
3 eDel-Ti-MWW-
N+Bu3Br
--H2O22040/1401/6----48[29]
4Ti-MMM-ETBABTBHP870489270--[30]
5TinOx-MCM-41TBABTBHP870488940--[31]
6TinOx-MMM-2TBABTBHP870488151--[31]
7N(CH3)2-[Im][Br]@
SBA-15
--TBHP108028>99--79[34]
8FeNi3/DFNS/sale
/Pd(II) MNPs
anilineTBHP208010----98[36]
9DFNS/Pt(II) NPSanilineTBHP2010010----97[37]
10ZnW-PYIs--TBHP550120----92[47]
11[{Mn2(TCPP)·2H2O}·DMF]nTBABPhIO8502447.339.3--[51]
12Fe@[{Mn2(TCPP)·2H2O}·DMF]nTBABPhIO8502498.665.3--[51]
13 fCuMo-BPYTBABTBHP550/5048/48----90[52]
14MOF-590TBABTBHP18010939487[56]
15MOF-591TBABTBHP18010958581[56]
16MOF-592TBABTBHP18010988280[56]
17ImBr-MOF-545(Mn)--O25601099.194.8--[57]
18Au/SiO2TBAB
/ZnBr2
CHP108047642--[60]
19 gAu/R201--TBHP4080/1503/498.0--50.6[61]
20Zn-N-C/Au@mSiO2TBABTBHP1801293.2--92.9[63]
21PN-CeO2TBABTBHP2080208193--[68]
22TBA-TBD-PZnBr2TBHP1801099.270.269.6[69]
23Co(acac)2-QPB@
MCS
--O230100495.273.169.6[73]
a Conversion of styrene. b Selectivity towards styrene carbonate. c Yield of styrene carbonate. d 60 °C and 24 h for epoxidation reaction and 120 °C and 4 h for cycloaddition reaction. e 40 °C and 1 h for epoxidation reaction and 140 °C and 6 h for cycloaddition reaction. f 50 °C and 48 h for epoxidation reaction and 50 °C and 48 h for cycloaddition reaction. g 80 °C and 3 h for epoxidation reaction and 150 °C and 4 h for cycloaddition reaction.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yan, T.; Wang, Z.; Guo, L.; Zhang, R.; Zhan, H.; Chen, J.; Wu, X. Recent Achievements in the Synthesis of Cyclic Carbonates from Olefins and CO2: The Rational Design of the Homogeneous and Heterogeneous Catalytic System. Catalysts 2022, 12, 563. https://doi.org/10.3390/catal12050563

AMA Style

Yan T, Wang Z, Guo L, Zhang R, Zhan H, Chen J, Wu X. Recent Achievements in the Synthesis of Cyclic Carbonates from Olefins and CO2: The Rational Design of the Homogeneous and Heterogeneous Catalytic System. Catalysts. 2022; 12(5):563. https://doi.org/10.3390/catal12050563

Chicago/Turabian Style

Yan, Ting, Zixian Wang, Li Guo, Ran Zhang, Haijuan Zhan, Jialing Chen, and Xiaoqin Wu. 2022. "Recent Achievements in the Synthesis of Cyclic Carbonates from Olefins and CO2: The Rational Design of the Homogeneous and Heterogeneous Catalytic System" Catalysts 12, no. 5: 563. https://doi.org/10.3390/catal12050563

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