Next Article in Journal
Controlled Growth of Unusual Nanocarbon Allotropes by Molten Electrolysis of CO2
Next Article in Special Issue
Research Progress and Reaction Mechanism of CO2 Methanation over Ni-Based Catalysts at Low Temperature: A Review
Previous Article in Journal
Recent Advances in Seawater Electrolysis
Previous Article in Special Issue
Tandem Reactions Based on the Cyclization of Carbon Dioxide and Propargylic Alcohols: Derivative Applications of α-Alkylidene Carbonates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

En Route to CO2-Based (a)Cyclic Carbonates and Polycarbonates from Alcohols Substrates by Direct and Indirect Approaches

by
Antoine Brege
1,2,
Bruno Grignard
2,
Raphaël Méreau
1,
Christophe Detrembleur
2,*,
Christine Jerome
2,* and
Thierry Tassaing
1,*
1
Institut des Sciences Moléculaires, UMR 5255 CNRS-Université de Bordeaux, 351, Cours de la Libération, CEDEX, F-33405 Talence, France
2
Center for Education and Research on Macromolecules (CERM), CESAM Research Unit, University of Liege, Allée de la Chimie, B6a, B-4000 Liege, Belgium
*
Authors to whom correspondence should be addressed.
Submission received: 7 December 2021 / Revised: 10 January 2022 / Accepted: 17 January 2022 / Published: 20 January 2022

Abstract

:
This review is dedicated to the state-of-the art routes used for the synthesis of CO2-based (a)cyclic carbonates and polycarbonates from alcohol substrates, with an emphasis on their respective main advantages and limitations. The first section reviews the synthesis of organic carbonates such as dialkyl carbonates or cyclic carbonates from the carbonation of alcohols. Many different synthetic strategies have been reported (dehydrative condensation, the alkylation route, the “leaving group” strategy, the carbodiimide route, the protected alcohols route, etc.) with various substrates (mono-alcohols, diols, allyl alcohols, halohydrins, propargylic alcohols, etc.). The second section reviews the formation of polycarbonates via the direct copolymerization of CO2 with diols, as well as the ring-opening polymerization route. Finally, polycondensation processes involving CO2-based dimethyl and diphenyl carbonates with aliphatic and aromatic diols are described.

Graphical Abstract

1. Introduction

Today, our modern society is facing climate change and ocean acidification caused by the continuous increase in CO2 levels within the atmosphere. Our annual anthropogenic emissions are now approaching 40 Gt/year, and it is recognized that alongside other gases, including water vapor, methane and fluoro-compounds, CO2 has become a major contributing factor to the greenhouse effect [1,2,3]. Legally binding obligations, governmental incentives and demands from industrial stakeholders are now strong drivers to fight global warming and to shift our societal paradigm based on fossil resources towards a more sustainable and circular vision, including CO2 utilization. Traditionally seen as a waste, CO2 may paradoxically become the key to “greening” our economy. Unlike biomass, the growth of which is inter-seasonal and requires multi-step post-treatment (extraction, fractionation, fermentation, etc.) prior to valorization, CO2 is a cheap and renewable carbon source that is regenerated instantaneously and locally. This resource may undergo numerous chemical processes and syntheses to design existing organic molecules or polymers or new (sophisticated) ones [4,5,6].
However, the main thermodynamic barrier to the large scale utilization of CO2 as a C1 building block lies in its stability, having a standard enthalpy of formation of ΔfH° = −394 kJ/mol [7]. To surpass this chemical inertness, the use of high-energy (co)reactants and the development of new catalytic methods and external energetic inputs are required to turn CO2 into valuable products.
The production of salicylic acid via the Kolbe–Schmitt process, urea from ammonia and CO2 or polyols by ring-opening copolymerization of oxiranes with CO2 are just few examples of competitive products that have already been trademarked in the chemical industry.
CO2 is also being used to fabricate the next generation of fuels and to store renewable energy via the so-called power-to-fuel technology. Its electrochemical reduction provides carbon monoxide, a key molecule in the Fischer–Tropsch process to produce alkanes. Numerous studies have revisited Sabatier’s reaction, which has been known for more than a century, to fabricate synthetic methane from renewable H2, while the partial hydrogenation of CO2 provides methanol. A myriad of sophisticated molecules are also accessible from the chemical conversion of CO2 via C-O or C-N bond formation. This route, as reviewed by others, is very promising, as (highly) reactive organic precursors counterbalance the low reactivity of CO2. The (photo)catalytic carboxylation of alkanes, alkenes or alkynes provides a diverse portfolio of carboxylic acids. The methylation of primary amines or the formylation of secondary ones is conducted with CO2 and a reductant (H2, PhSiH3, etc.).
The (a)cyclic organic carbonates and carbamates with ample structural diversity are fabricated with a high level of selectivity and a 100% atom economy pathway via non-reductive catalytic transformations of CO2 with oxiranes [8,9,10,11], aziridines [12] and amines [13]. The insertion of CO2 within epoxides is by far the most extensively explored transformation process for CO2, as it provides 5-membered cyclic carbonates used as solvents, electrolytes for Li-ion batteries or intermediates in fine organic chemistry and polymer sciences. Despite being very attractive and easily produced via facile oxidation of olefins, including natural ones derived from vegetable oils or terpenes, epoxides display acute toxicity, as they act as alkylating agents capable of binding with DNA [14,15]. In this context, the replacement of epoxides with safe and easy-to-handle (renewable) substrates to synthesize organic carbonates and polycarbonates is now gaining huge interest in the scientific community.
Alcohols are ubiquitous scaffolds with low toxicity made from multiple synthetic and natural sources. This makes them ideal candidates within the field of CO2 transformation to replace epoxides. This review provides an overview of the progress made in the synthesis of (a)cyclic and polymeric carbonates from alcohols substrates. The first section is dedicated to the main pathways that are used to upgrade CO2 with mono-, di- or polyols; haloalcohols; or allylic or propargylic alcohols into organic scaffolds, and is complemented by mechanistic guidelines. The second section highlights the synthesis of polycarbonates either by direct polycondensation of CO2 with alcohols or via the polymerization of CO2-sourced monomers through ring-opening or step-growth methods. An overview of the most relevant routes discussed in this review for the direct coupling of CO2 with alcohols to produce (a)cyclic carbonates and polycarbonates is presented in Scheme 1.

2. CO2-Based Synthons: Organic Carbonate Synthesis from the Carbonation of Alcohols

Industrially, the oxidative carbonylation of alcohols by phosgenation is the most facile and economically viable route for fabricating (a)cyclic organic carbonates. However, the toxicity of phosgene; the protective measures required for its safe transportation, storage and handling; and the large volume of chlorinated waste generated during the reaction are strong limitations that support the development of a more sustainable carbonation process. In this context, the formation of carbonates via the coupling of alcohols and CO2 looks very attractive but remains an elusive endeavour, as this approach faces two severe limitations linked to: (i) the difficulties in identifying ways to properly activate CO2, as it will not spontaneously react with alcohols; (ii) the formation of water as a by-product, which affects the reaction equilibrium (Scheme 2).
To surpass these hurdles, three main strategies have been envisioned:
(i)
The “dehydrative condensation” approach, utilizing the synergy between a catalyst, which facilitates the CO2 fixation onto the alcohol moiety, with physical or reactive dehydrating agents to trap or consume water;
(ii)
The “alkylation” strategy, involving the activation of CO2 under the form of a hemi-carbonate ion in the presence of an excess of base prior to further reaction with an alkyl halide via nucleophilic substitution. As a variant, the “leaving group” approach proposes the in situ formation of a highly reactive carbonate intermediate from the ter-reaction between CO2, an alcohol and a dihalide, which undergoes facile transcarbonation into an acyclic carbonate;
(iii)
The “protected alcohols” route, which generates in situ alcohols from ketals, trimethyl phosphates, or orthoesters via water consumption prior to reaction with CO2

2.1. Synthesis of Acyclic Organic Carbonates

The synthesis of dimethyl carbonate (DMC) is by far the most widely studied pathway to illustrate the conversion of alcohols and CO2 into useful organic scaffolds. However, the high standard Gibbs-free energy of DMC formation from methanol and CO2RG ≈ +37 kJ/mol) indicates that the reaction is thermodynamically unfavorable, limiting the yield of the overall reaction to below 1% [16,17,18]. Various approaches have been proposed and optimized to produce DMC with good yields and selectivity.

2.1.1. DMC Synthesis by Dehydrative Condensation

Since the pioneering studies by Kizlink [19,20], who demonstrated the low yield but selective formation of DMC from CO2 and MeOH using organotin catalysts, a tremendous number of metal alkoxide catalysts have been developed. Titanium-(Ti(OEt)4, Ti(OBu)4) [21], nickel-(Ni(OAc)2·4H2O) [22], niobium-([Nb(OMe)5]2) [23] and magnesium-based catalysts [24] are some examples of homogenous systems that have proven their utility in driving the formation of the acyclic organic carbonates. Heterogeneous amphoteric metal oxide systems combining acidic and basic sites such as ZrO2 [25,26,27]; ZrO2–phosphoric acid dual catalysts [28,29], ceria oxides [30,31] and other mixed oxides [32,33,34]; or Cu-Fe, Cu-Ni and bimetallic oxides immobilized onto graphite oxide and graphene nanosheets [35,36,37] also displayed some catalytic activity. However, for all systems, the formation of water as a by-product severely limited the reaction yields (up to 5% for the H3PW12O40/Ce0.1Ti0.9O2 system [38]). Some catalysts also faced selectivity issues with the high-temperature concomitant formation of formaldehyde and carbon monoxide from CO2 and methanol. To overcome these limitations, the utilization of metal-based catalysts that operate in synergy with “physical” or reactive desiccants or dehydration system emerged as a valuable option to push the reaction manifold toward high yield and the selective formation of DMC.
  • Physical drying systems;
Membrane catalytic reactors combing a Cu catalyst immobilized onto a mixed oxide carrier (MgO-SiO2) with hydrophilic polyimide–silica or polyimide–titania hybrid separating membranes [39] or inorganic absorbents such as molecular sieves or zeolites combined with organotin catalysts [40,41] were engineered to get rid of water and displace the reaction equilibrium towards the DMC formation. Among them, the Bu2Sn(OMe)2–3A molecular sieves provided the best DMC yields (45%) under operative conditions of 180 °C and 30 MPa of CO2 in 80 h. This approach highlights the benefits of the desiccant for the carbonylation of methanol, with a 10-fold increase in DMC yield [40,42]. The catalytic cycle of Bu2Sn(OMe)2 is shown in Scheme 3.
  • Reactive dehydrative systems;
To tackle the water issue and push the reaction equilibrium towards DMC formation, two scenarios involving stoichiometric or excessive reactive dehydrative systems were developed. In the first one, DMC was synthesized from CO2 and methanol (PCO2 = 15 MPa, T = 180 °C, t = 2–10 h) using (Bu)2SnO in combination with CaC2, which spontaneously consumed water to form acetylene as a by-product. DMC was formed with yields up to 11% with 13 mol% of CaC2 [43]. In the second scenario, a single catalyst was used to simultaneously drive the alcohol carbonation and hydration of the desiccant, for example the catalytic hydration of nitriles into amides. This elegant concept was pioneered by Tomishige, who used CeO2 as metal oxide catalyst and nitriles as reactive desiccants in synergy [44,45]. Zirconia-based catalysts [26,29] and amphoteric CeO2 simultaneously display active acidic and basic sites that activate CO2 and methanol. These acidic and basic sites were characterized by Lavalley et al., who differentiated low-coordination Ce cations resulting from oxygen defects as acid sites, and oxygen atoms adjacent to these low-coordination Ce cations as basic sites [46,47] that can properly fix alcohols and CO2 [48]. Interestingly, these acid-base properties can efficiently catalyze nitrile hydration into amides. Among nitriles, acetonitrile was first considered. In situ FT-IR studies revealed the presence of CH3CN adsorbed over CeO2. When water is added to the system, a simultaneous decrease in CH3CN species with increases in the acetamide anion CH3CONH and amide CH3CONH2 can be observed, meaning that CeO2 successfully catalyzes this reaction [44]. Mechanistic insights revealed that water is dissociated over CeO2 and the OH species generated on the acid sites, which react with the adsorbed CH3CN to form the acetamide anion. Reprotonation occurs thanks to the H+ species generated on basic sites from H2O dissociation over CeO2.
When applied to the synthesis of acyclic alkyl carbonates from alcohol and CO2, the CeO2–acetonitrile system effectively increased the yield of carbonate [49,50]. DMC was obtained with a 8.9% yield and 64% selectivity and diethyl carbonate was obtained (DEC) with a 7.5% yield and 94% selectivity under relatively mild conditions (150 °C and 0.2–0.5 MPa). The low selectivity for DMC formation arose from side reactions driven by the by-produced amide. The latter can react with alcohols such as methanol to form esters with the release of ammonia. NH3 further reacts with DMC to form methyl carbamate. To overcome these problems, the generated amide has to be more stable than acetamide. Benzonitrile was then tested, since the phenolate ring would increase the stability of the benzamide and possibly prevent side reactions. At 150 °C and 1 MPa CO2 pressure, the CeO2–benzonitrile dual system afforded DMC in 47% yield with 78% selectivity [51]. To increase the yield, efforts were made to better understand the hydration reaction of nitriles over CeO2. Tamura et al. successfully demonstrated that a heteroatom in the β-position of the nitrile function would considerably increase its adsorption over CeO2 by coordinating this heteroatom, as was demonstrated for pyridine [52]. Combining all of these aspects, 2-furonitrile and 2-cyanopyridine were selected as the best candidates and showed excellent hydration yields at mild temperatures (90% at 80 °C after 7 h and 100% at 30 °C after 1 h, respectively [44]). A mechanism proposed by the authors is presented in Scheme 4. The best result was obtained with 2-cyanopyridine, whereby DMC was formed at 94% yield with 96% selectivity after 12 h at 120 °C and 5 MPa CO2 pressure. A methanol/2-cyanopyridine ratio of 2:1 was needed, highlighting the need for a stoichiometric amount of dehydration agents. It should be noted that enhancement of the catalytic activity by increasing the amount of low-coordination Ce sites was achieved by preheating at 600 °C under air atmosphere for 3 h [53]. The protocol was extended to other alcohols and successfully demonstrated its versatility by affording a huge library of linear carbonates [54] (see Scheme 5).

2.1.2. Synthesis of Acyclic Carbonates via the Alkylation Route

Another way to promote the formation of linear carbonates from the coupling of alcohols and CO2 is to use co-reagents in stoichiometric amounts that will bypass the formation of water. Conceptually, alcohols can be activated or even deprotonated by bases to generate a methoxy species that can bind spontaneously with CO2 to form the corresponding ROC(O)O hemi-carbonate. This nucleophilic anion further reacts with a halide reagent R′-X, which undergoes a nucleophilic substitution to produce the envisioned ROC(O)OR′ species.
Inorganic bases: Inorganic carbonates have been studied for the three-component coupling of CO2, alcohols and alkyl or benzyl halides. Jung et al. proposed a comparative study of different inorganic carbonates with tetrabutylammonium iodide (TBAI) as an additive and n-butyl bromide as a co-reagent based on the model carbonation of 4-phenyl-1-butanol with CO2 [55]. Interestingly, Na2CO3, Rb2CO3, K2CO3 and Cs2CO3 successfully demonstrated their activity in mild conditions (23 °C, atmospheric pressure of CO2 bubbling). No reaction was observed with Li2CO3. The yield remained low for sodium- and potassium-based carbonates (2%), as well as for Rb2CO3 (10%). Excellent yield was displayed for Cs2CO3 (85%) in only 6 h. Good enhancement of the base activity was attributed to the known “cesium effect”. Due to its larger ionic radius, the cesium cation weakly bonds to the methoxy species generated through alcohol deprotonation. TBAI plays its role in stabilizing the carbonate anion generated after CO2 insertion and prevents non-desired alkylation. The scope was then extended to other alcohols and alkyl or benzyl halides were more sterically hindered, affording a wide variety of linear carbonates. By using excess base (5 eq), TBAI (5 eq) and R-X (5 eq), cheaper base K2CO3 could afford linear carbonates from various substrates (up to 82% with benzyl alcohol and benzyl bromide) at 60 °C [56]. The utilization of other additives such as crown ethers, KI or tetrabutylammonium bromide was found to be detrimental to the reaction yields.
Organic bases: As with inorganic bases, organic superbases drove the synthesis of acyclic carbonate from alcohols, CO2 and alkyl halides. With KI or KI/crown ethers as additives added in a catalytic amount (10 mol% regarding the alcohol), dibenzyl carbonate was successfully synthetized in moderate yields (25% and 30% respectively), demonstrating the efficiency of DBU for CO2 fixation onto alcohols [56]. It should be noted that in this case, benzyl chloride was used instead of benzyl bromide to prevent possible side quaternization of DBU [57]. Bicyclic guanidines were, thus, tested as bases for the coupling of alcohols and CO2. Merging computational and experimental studies allowed the reaction mechanism to be elucidated. Computational insights (conducted at the B3LYP level using the 6–31 g(d) basis set) aided in comprehending the mechanism of the carbonate formation using DBU as a base. Wang et al. suggested the four possible mechanisms shown in Scheme 6 [58]. Considering thermodynamic and kinetic aspects (ΔG = −16.20 kJ/mol and ΔG = 30.47 kJ/mol, rate constant of 2.9 × 107 s−1), the plausible mechanism involved a one-step trimolecular reaction between CO2 the alcohol and DBU (mechanism 4 in Scheme 6), even though activation of CO2 by DBU (mechanism 3 in Scheme 6) is theoretically feasible prior to formation of the acyclic carbonate via its addition onto the alkyl halide. However, it should not be generalized to other amidines and guanidines. Indeed, when using TBD, the zwitterionic adduct with CO2 was still observed under strictly anhydrous conditions [59]. Thanks to the presence of both sp2 amine and secondary sp3 amine functions, the adduct was stabilized and could be isolated and well characterized by X-ray diffraction to determine the crystal structure (see Scheme 7). In this case, the adduct was more stable than in the case of DBU thanks to the hydrogen bonding between secondary amine of the N′ labelled atom with the O′ atom of CO2 bound to TBD. However, when applied to the alcohol/CO2 system, which was propanediol in this case, the synthesis of the corresponding carbonate is more likely to happen with that activation of the alcohol rather than CO2 [60].
Ionic liquids derived from Hünig’s base-type cations were used as recyclable bases in replacement of DBU or TDB for the fixation of CO2 onto alcohols with benzyl bromide as the co-reagent [61]. Due to their non-nucleophilic characters, no side reaction between the base and the benzyl halide was detected. The high basicity of the ionic liquids allowed the spontaneous deprotonation of the alcohol moiety. The generated methoxy species is able to attack CO2, and SN2 on benzyl bromide will generate the corresponding non-symmetrical linear carbonate. At 25 °C and under 1 MPa CO2 pressure after 72 h, good yields were obtained for ethanol (82%), allyl alcohol (80%), benzyl alcohol (83%), phenol (78%), propargylic alcohol (76%) and 2-butanol (67%), with an alcohol/halide/IL molar ratio of 1:0.5:0.6. The ionic liquid was successfully recycled by distillation or extraction, with no loss of activity when reused.

2.1.3. Synthesis of Acyclic Carbonates Using the “Leaving Group” Strategy

The formation of acyclic carbonate using the leaving group strategy shares conceptual similarities with the alkylation route. The hemi-carbonate ion resulting from the trimolecular reaction between the alcohol, base and CO2 attacks a dihalide species such as CH2X2 through a nucleophilic substitution to produce the corresponding acyclic carbonate. This highly reactive ROC(O)OCH2X further undergoes a transcarbonation reaction with a second alcohol, providing the desired ROC(O)OR carbonate. This concept was applied to fabricate dibenzyl carbonate from CO2 and benzyl alcohol using DBU with the ionic liquid bmimPF6 as an additive in CH2Br2 at 70 °C under 1 MPa CO2 pressure [62]. Dibenzyl carbonate was obtained in 69% yield with 2 equivalents of DBU and 0.1 mL of bmimPF6 for 0.5 mmol of benzyl alcohol. Equimolar amounts of [DBUH]Br salt and BrCH2OH were concomitantly generated as by-products. The authors did not mention any particular role of the ionic liquid in the mechanism itself, but rather as a promotor for CO2 solubilization in the reaction media. When bmimPF6 was not added to the system, the yield decreased from 69% to 25%. Similarly, when CH2Cl2 was used instead of CH2Br2, the yield decreased to 25%, supporting the SN2 mechanism, which is more favorable with bromine versus chlorine in terms of its reactivity. No carbonate was produced with more reactive halides such as CH3I, supposedly due to quaternization of DBU. The optimized protocol was applied to other benzyl alcohols derivatives and could afford the corresponding carbonates in moderate to good yields (from 47% to 73%). DMC was also synthetized with 48% yield.

2.1.4. Synthesis of Acyclic Carbonates via the Carbodiimide Route

Aresta et al. designed an elegant method for synthesizing acyclic carbonates using carbodiimides acting both as the co-reagent and base [63]. Carbodiimides react with the alcohol to produce an isolable O-alkyl isourea intermediate [64]. This reaction can be catalyzed by transition-metal-based complexes (mostly copper-based) such as CuCl, Cu2I2, CuO, Cu(OAc)2 and ZnO2; organic superbases such as TBD; or alkali–metal TBD complexes [65]. Then, the O-alkyl isourea participates in a trimolecular interaction between the alcohol and CO2 via its remaining basic iminoether moiety, leading to an isourea–hemicarbonate adduct that allows for the carbonation of a second alcohol molecule. The latter will react with the alkyl part fixed on the O-alkyl isourea to generate the corresponding linear alkyl carbonate (see Scheme 8). Following this approach, DMC was synthetized in yields of up to 62% in 6 h at 65 °C with 5 MPa of CO2. Allyl alcohol successfully provided the corresponding carbonates as well, with the best yield of 93% using an alcohol/DCC ratio of 15:1 and DCC/Cu2I2 ratio of 500 at 65 °C and 5 MPa CO2 pressure after only 4 h.

2.1.5. Synthesis of Acyclic Carbonates from the “Protected Alcohols” Route

Ketals undergo reversible hydrolysis to form alcohols and ketones. Thus, they can be seen as the protected form of dehydrated alcohols. Upon water release, the ketal decomposes in situ into the former ketone and an alcohol that is subsequently involved in a cascade coupling reaction with CO2 to afford the acyclic carbonate (Scheme 9). Notably, systems merging the utilization of 2,2-dimethoxypropane or 2,2-diethoxypropane with CeO2 or CeO2-ZrO (mixite) oxide catalysts and CO2 [66,67] gave access to DMC with yields of 13.8–59.3 mmol·gcat−1 in 2 h at 110–140 °C and 5–6 MPa of CO2. However, the main bottleneck of these systems arose from the slow hydration rate of the 2,2-dialkoxypropane. To improve the performances, acidic co-catalysts were used as additives to boost the ketal hydration. Heterogeneous acids such as zeolites (H-FAU zeolite with SiO2/Al2O3 = 5.5) used in combination with 2,2-diethoxypropane and CeO2 catalyst remarkably increased the DEC yields up to 72% based on 2,2-diethoxypropane (T = 120 °C and 9.5 MPa of CO2) [68].
As with ketals, the structural features of orthoesters make them sensitive to hydrolysis and decompose into esters and alcohols. Orthoesters are good candidates for consuming water while releasing the alcohols during the fabrication of the acyclic carbonates in situ. As a representative example, DMC was synthesized by combining trimethyl orthoacetate with supercritical CO2 using binary catalysts made of organotin complexes and ammonium, phosphonium salt additives [69] or ceria–zirconia mixte oxides catalysts [70]. The highest DMC yield of 70% was obtained after 72 h at 180 °C, under 30 MPa of CO2 pressure using the Bu2Sn(OMe)2/Bu4PI catalytic system, while lower DMC yields of 1.8–10.4% were obtained at 12 MPa after 34 h when 1,1,1-trimethyl orthoformiate was combined with ceria–zirconia oxides. Recently, Fukaya revisited this conceptual route by using water-sensitive an alkoxysilane, i.e., tetraethyl orthosilicate, as the alcohol precursor and a Zr(OEt)4 catalyst, which favors the ethoxylation of CO2 [71]. The DEC yield reached 58% after 24 h at 180 °C under 5 MPa of CO2. As with the last method, trimethylphosphate was tested as a water scavenger to synthesize DMC from in-situ-generated methanol. The transformation of CO2 into the acyclic carbonate was performed at 180 °C for 20 h under 2.8 MPa of CO2 using catalytic amounts of Bu2Sn(OBu)2 and provided DMC with a moderate yield of 47% [19].

2.2. Synthesis of Cyclic Organic Carbonates

Cyclic carbonates are useful chemicals deserving huge interest as green solvents, electrolyte for batteries or monomers to design polycarbonate or isocyanate-free polyurethanes. Most of them are easily produced from vicinal or 1,3-diols and highly reactive phosgene derivatives, diphenyl carbonate (DPC) or DMC [72,73,74], which can be CO2-based, as demonstrated previously in this review. Finding efficient routes to convert alcohols directly into their corresponding cyclic carbonate has become a challenge in CO2 chemistry. This section is dedicated to the various strategies that have been engineered to synthesize these heterocycles from various alcohols substrates, including vicinal and 1,3-diols, allylic and propargylic alcohols and halohydrins. One has to keep in mind that 5-membered and 6-membered cyclic carbonates will be exclusively discussed in this section, since their formation is thermodynamically and kinetically more favorable than smaller or larger rings.

2.2.1. Synthesis of Cyclic Carbonates from Diols

Diols are ubiquitous scaffolds that are readily available from biomass, making them attractive substrates for the synthesis of cyclic carbonates [75,76]. However, the direct carbonation of vicinal or 1,3-diols suffers from the same thermodynamic limitation as that observed for the synthesis of acyclic carbonates from monoalcohols, i.e., the formation of water by-products [77]. Therefore, most of the strategies proposed for the fabrication of acyclic carbonates may be transposed to diol substrates. However, the large diversity of substrates investigated in the scientific literature to produce cyclic carbonates from diols made it difficult to summarize and rationalize the results. This may confer to the reader of the following sections the perception of a lack of coherence in the manuscript.

2.2.1.1. Dehydrative Condensation

Physical drying systems: Glycerol, a waste bio-based polyol from the production of biodiesel, was converted by Aresta et al. into glycerol carbonate with CO2 via the dehydrative condensation approach. By using n-Bu2Sn(OCH3)2 in combination with molecular sieves at 180 °C under 5 MPa CO2 pressure, the cyclic carbonate was produced with a low yield of ~7% in 15 h (see Scheme 10) [78].
Switching to n-Bu2SnO was detrimental, with glycerol carbonate yields decreasing to ~3%. The catalytic activity of n-Bu2SnO could be enhanced via the addition of methanol [41]. Glycerol carbonate was obtained selectively with a yield of 35% in 4 h with 13X zeolite as the dehydration agent at 120 °C under 13.5 MPa CO2. It was postulated that methanol played a dual role. It acted as a solvent, providing a homogeneous reaction mixture. Methanol also transformed Bu2SnO in situ into the more active n-Bu2Sn(OCH3)2 catalyst. The catalytic cycle involved is presented in Scheme 11. This protocol was extended to the selective synthesis of ethylene carbonate and propylene carbonate from ethylene glycol and propylene glycol, respectively, in good yields (61% and 42%, respectively), using zeolite as the drying system [79].
Reactive dehydrative systems: The activity of dual catalysts mixing a metal oxide (ZrO2, Al2O3, TiO2, polyhedral CeO2 or ZnO) with 2-cyanopyridine was evaluated for the synthesis of glycerol carbonate from glycerol and CO2 (T = 170 °C, 10 MPa of CO2 for 12 h). The benchmarking of the glycerol carbonate yields evolving from ~10.4 to 14.2% enabled the classification of the catalysts regarding their activity in the order of ZrO2~Al2O3 < TiO2 < ZnO < CeO2. In the quest for the optimization of the glycerol carbonate production, the influence of the reactive desiccant on the reaction was evaluated with polyhedral CeO2 and revealed that replacing 2-cyanopyridine by 3- or 4-cyanopyridine and acetonitrile was detrimental, with yields below ~1–5%. The crucial role of the CeO2 crystallinity [30] in the catalytic performance was also highlighted. Mesoporous CeO2 and nanocubes displayed the lowest efficiency, with similar glycerol carbonate yields of ~10.3%. CeO2 nanorods and nanowires increased the cyclic carbonate yields to ~13%. CeO2 with polyhedral crystallinity remained the most effective catalyst [80]. Tomishige et al. further extended the substrates to other mono- or disubstituted vicinal or 1,3-diols using the CeO2/2-cyanopyridine catalytic system (T = 130–150 °C, PCO2 = 5 MPa, t = 1–8 h). Remarkably, the corresponding substituted 5-membered and 6-membered cyclic carbonates were synthesized in good to excellent yields (from 62% to 99% for 14 examples) [81], demonstrating the versatility of such systems (see Scheme 12).
As 2-cyanopyridine is rather expensive and used in large excess, acetonitrile was also considered as a more affordable solvent and dehydration agent. In combination with an organic base such as TBD (2.5 mol%), propylene carbonate could be synthetized with yields up to 22.5% but with a low selectivity rate of 60.3% [82], due to the side reaction of propylene glycol with the hydrolysis product of acetonitrile, namely acetic acid (ammonia is also generated). Ammonium carbonate (NH4)2CO3 was found to considerably increase the selectivity toward propylene carbonate by preventing the decomposition of acetamide into ammonia and acetic acid. However, the yield was lowered to 15.6%, due to decomposition of (NH4)2CO3 into NH4+, which inhibits the hydrolysis of acetonitrile.

2.2.1.2. Cyclic Carbonates Synthesis via the Leaving Group Strategy or Alkylation Route

Alike acyclic carbonates (Section 2.1.2 and Section 2.1.3), produced via strategies utilizing bases and alkyl or aryl halides R-X as co-reagents to bypass the formation of water, were used for CO2 fixation onto vicinal or 1,3-diols. Mechanistically, the hemi-carbonate ion formed by fixation of CO2 onto one OH moiety undergoes an alkylation by SN2 substitution with an alkyl halide. Then, the second OH group is involved in an intramolecular transcarbonation reaction promoting the ring closure of the molecule (see Scheme 13).
Organic systems combining DBU with CH2Br2 or BuBr as alkylation agents and ionic liquid bmimPF6 or CH2Cl2 as solvents promoted the formation of a large diversity of (substituted) 5- or 6-membered cyclic carbonates (glycerol, styrene, propylene, cis-cyclohexene, trimethylene, carbonates) under mild operative conditions (T = 25–70 °C, PCO2 = 0.1–1 MPa, t = 18–24 h) [62,83], with yields ranging from 60 to 90%. Unfortunately, when diols are chosen as substrates to fix CO2, the carbonation of both alcohol moieties followed by double alkylation drives the concomitant formation of an acyclic bis-carbonate by-product. The competition between the mono- and double carbonations seems dependent on both the substrate and the nature of the base. A second challenge lies in the formation of a second type of acyclic carbonate. Upon cyclization, an alcoholate group is released in the reaction medium. This species is then capable of fixing one molecule of CO2 and reacting via SN2 with the alkyl halide co-reagent to form a (reactive) linear dialkyl carbonate. Inspired by these studies, Buchard et al. utilized highly reactive tosyl chloride (TsCl) as a co-reagent in combination with organic bases to promote the fixation of CO2 onto 1,3-diols [84]. TsCl reacts easily with O-nucleophiles to form a tosylate known to be a very good leaving group [85]. However, TsCl is not compatible with DBU and spontaneously formed a [DBUTs]Cl, meaning that a sequential procedure was using to fabricate the 6-membered cyclic carbonates. Firstly, the carbonation of one of the two alcohol moieties by CO2 fixation was promoted by using 1 equivalent of DBU. Once all the DBU was reacted, TsCl and TEA were introduced to perform the intramolecular cyclization. A mechanistic proposal was proposed by DFT calculations and supported by carbonation experiments of optically active diols. Similarly to what was proposed with alkyl or aryl halides, TsCl reacted more preferentially with the generated anionic carbonate than with the remaining alcohol moiety, exclusively showing stereochemistry retention (see Scheme 14). Such approach enabled the synthesis of up to ten 6-membered cyclic carbonates from substituted 1,3-diols with moderate to good isolated yields at room temperature with atmospheric pressure of CO2 in chloroform.
Later, the same group decoupled the reaction and investigated in detail the influence of various tertiary amines and solvents on the carbonation step of neopentyl glycol [86]. With less-toxic acetonitrile instead of chloroform, 1 eq of DBU afforded 85% of CO2 insertion with a mono/double carbonation ratio of 59:41. 2,2,6,6-Tetramethylpiperidine (TMP) and TEA showed very low CO2 insertion rates of 2% and 4%, respectively, but with 100% selectivity toward monoinsertion. No carbonation was observed with pyridine.
In accordance with these results, they adapted their synthetic procedure in a one-pot and one-step fashion by utilizing TsCl (1 eq) with TEA or TMP as the base (2 eq) for the carbonation of 1,x-diols (with x = 2–4), including challenging tri- or tetrasubstituted substrates. The non-nucleophilic character of these organic bases allowed the simultaneous addition of TsCl and the base without undesirable side quaternization of the amine. Here, 5-, 6- and even 7-membered cyclic carbonates were fabricated with excellent isolated yields using this simple approach, with similar or even better yields than those reported using the phosgene route. Nevertheless, with 1,4-diols, non-negligible amounts of oligomers were noted, possibly due to the poor stability of 7-membered cyclic carbonates in basic conditions or a competition between the cyclization and the polymerization. In the same manner, a critical study balancing two metal-free dual activating systems, namely DBU/EtBr and TEA/TsCl, was recently proposed for the coupling of CO2 with 1,x-diols to afford (a)cyclic carbonates [87]. In situ ATR-IR monitoring correlated with DFT calculations was performed to investigate the mechanism and reaction kinetics under various experimental conditions, leading to an optimized synthetic protocol. By choosing the suitable organic dual-activation system, it was possible to control the product selectivity to substituted ethylene- or trimethylene carbonate or acyclic compounds, providing a powerful tool to synthesize CO2-based precursors that are highly relevant for organic and polymer chemistry from ubiquitous building blocks.
N-heterocyclic carbenes (NHC) are known to activate CO2 under the form of imidazolium carboxylates [88,89] and can be used in combination with a base and an alkylating agent to form cyclic carbonates from diols. Dyson et al. illustrated how thiazolium carbene (20 mol%) generated in situ by Cs2CO3 (3 eq) promoted the formation of styrene carbonate (yield of 61%) when used in combination with 2 eq of BuBr (T = 90°C, 0.1 MPa of CO2 in DMF) [90]. The authors suggested two competitive mechanisms leading to both cyclic carbonates. In the first one, CO2 was activated via the formation of a thiazolium carboxylate diadduct that attacked the alkyl halide R-X through nucleophilic substitution, forming a carbonate ion. Deprotonation of an alcohol moiety by the base allows transesterification on this carboxylate with RO as the leaving group. The cyclization is promoted with a second equivalent of base (Scheme 15). In the second plausible mechanism, dibutyl carbonate was formed from the reaction of CO32− with BuBr. Then, an oxidative carboxylation occurred via the reaction between the diol and dibutyl carbonate to produce the cyclic carbonate. In this case, CO2 did not participate in the carbonation of the diol (Scheme 16). The validity of this hypothesis was supported by the formation of the cyclic carbonate, even in the absence of CO2, as the cyclic carbonate yield considerably decreased but was still significant (25%).

2.2.2. Synthesis of Cyclic Carbonates from Allyl Alcohols

The pioneering study by Cardillo et al. in 1981 demonstrated the efficiency of allyl alcohols as valuable substrates for the formation of cyclic carbonates with CO2 [91]. In their study, various 5- and 6-membered cyclic iodocarbonates were synthetized from allylic and homoallylic alcohols and CO2, using n-butyllithium as a base and I2 as a promotor for cyclization. Functionalization of the double bond by I2 is highly regioselective, which allows the formation of the product in very good yields under mild conditions (CO2 bubbling in THF; T = 25 °C, t = 12 h). With an alcohol/iodine/base molar ratio of 10:20:6.6, various 5-membered cyclic iodocarbonates were obtained from primary, secondary and even tertiary allylic alcohols in moderate to good yields (60–90%). Furthermore, 6-membered cyclic iodocarbonates were also obtained from primary and secondary homoallylic alcohols in yields ranging from 72% to 80%. The generated iodocarbonate could be converted into iodine-free carbonate via treatment with Bu3SnH. Conversion of homoallylic alcohols into 6-membered cyclic iodocarbonates was also investigated using N-iodosuccinimide as the iodine agent (1.1 eq with regard to the alcohol) [92]. The catalytic system used relies on a Brønsted acid-base bifunctional salt with best results using pyrrolidine-substituted bis(amidine) incorporating trans-stilbene diamine (StilbPBAM, see Scheme 17) as the base and triflimidic acid HNTf2 as the acid. Good to excellent yields were obtained for a large substrate range of primary homoallylic alcohols in toluene at −20 °C with CO2 bubbling (Scheme 18). High enantiomeric excess was observed in most cases. Molecular sieves were added to prevent deactivation of the catalyst in the presence of water and CO2. The use of an acid is assumed to stabilize the CO2–alcohol adduct. To overcome the use of toxic or expensive iodinating agents, Minakata et al. proposed the use of tBuOI, which can be generated in situ from tBuOCl and NaI, for the conversion of allylic and homoallylic alcohols into 5- and 6-membered iodocarbonates [93]. In this case, tBuOI abstracts the proton of a weak acid that is replaced by an iodine atom. The thermodynamically disfavored synthesis of alkylcarbonic acid from the CO2–alcohol coupling can then be turned into a highly reactive iodinated intermediate, with which cyclization is promoted by the presence of the double bond (Scheme 19). Thus, if the alkylcarbonic acid can be trapped in some way, the equilibrium should be displaced toward the formation of cyclic iodocarbonates. This proposed mechanism should be consolidated with more theoretical and experimental studies. The substrate scope proved the efficiency of this non-catalytic method, affording cyclic carbonate scaffolds with good yields (from 57% to 93%) when 2 eq of tBuOI was used at −20 °C with CO2 bubbling.

2.2.3. Synthesis of Cyclic Carbonates from Halohydrin

A more direct route for the synthesis of 5- and 6-membered cyclic carbonates is the use of β- and γ-halohydrins in combination with a base (1.1 eq) and CO2. Ethylene carbonate and trimethylene carbonate are examples of scaffolds obtained from 2-chloroethanol and 3-chloropropanol using this approach, providing excellent yields (>95% yield by simple bubbling of CO2 at 40 °C) [85]. The proposed mechanism is quite simple. Cs2CO3 deprotonates the alcohol to form an alcoholate that fixes CO2. Then, the ring-closure occurs via intramolecular nucleophilic substitution with Cl as the leaving group. Less reactive bases such as K2CO3 afforded lower carbonate yields (54%), as well as the more reactive KOtBu (20%), once again highlighting the excellent properties of cesium for these reactions. Slightly lower yields were obtained when replacing 2-chloroethanol with 2-bromo or 2-iodoethanol. Interestingly, when a tosylate OT substituent was used instead of a halide, the corresponding cyclic carbonate was also obtained in very good yields for a range of substrates (from 65% to 95%). When (±)-3-chloro-1,2-propanediol was used as the substrate, exclusive formation of glycerol carbonate was observed, demonstrating the excellent selectivity of this reaction. As K2CO3 is a cheaper and more affordable base than Cs2CO3, Wang et al. investigated the effects of additives on the formation of cyclic carbonate from halohydrins and CO2 with K2CO3 [94]. With Poly(ethylene glycol) of Mn = 400 g/mol (PEG-400), styrene carbonate, propylene carbonate, ethylene carbonate and chloromethyl-ethylene carbonate were formed quantitatively after 3 h at 50 °C under 2 MPa CO2 pressure using 0.5 eq of K2CO3. The more challenging cyclohexene carbonate was obtained with 72% yield in the same conditions. The role of the additive was assumed to coordinate potassium cations favoring the salt dissociation in the medium, thereby increasing the basicity of K2CO3. Non-nucleophilic basic ionic liquids tested for the synthesis of linear carbonates by Goodrich et al. were also tested on 3-bromopropanol and 2-chloroethanol, successfully producing ethylene carbonate and trimethylene carbonate in good yields (>90%) with excellent selectivity toward cyclic products [61].
As halohydrins are not widely available and are generally synthetized from oxidation of olefins, a one-step synthesis from terminal alkenes was proposed by Li et al. [95] (Scheme 20). This method relies on the in situ generation of bromohydrin using water as both the solvent and oxidation agent, N-bromosuccinimide as the bromination agent (1 eq) and DBU as the base (2 eq) for CO2 fixation onto the generated alcohol. The 5-membered cyclic carbonates were obtained in moderate to excellent yields after several hours of reaction at 60 °C under 2 MPa CO2 pressure. From these results, a catalytic process was then envisioned, since bromide ion can be oxidized back to bromination agents under oxidative conditions, allowing a catalytic cycle for the in situ generation of bromohydrin from olefins [96]. Hydrogen peroxide was used to regenerate the bromination agent. Several bromination agents were tested, and the best results were obtained with tetrabutylamonnium bromide or NaBr. Styrene carbonate could be obtained at 70% yield using an olefin/base/bromination agent molar ratio of 1.5:0.2:0.15 after 15 h at 50 °C under 2 MPa of CO2 pressure.

2.2.4. Synthesis of Cyclic Carbonates from Propargylic Alcohols

The carboxylation coupling of propargylic alcohols with CO2 is highly attractive. This 100% atom economy synthetic pathway provides organic cyclic carbonates that slightly differ from conventional cyclocarbonate scaffolds due to the presence of an additional exovinylene moiety (Scheme 21). A large diversity of catalysts has been reported to promote this reaction. These catalysts have to fulfil a dual role: the activation of the alcohol moiety or CO2 activation to facilitate the carbonation step, and the activation of the triple bond to drive the intramolecular cyclization and ring formation. The chemical structure of the substrate also strongly influences the formation of the cyclic scaffold. The carboxylative cyclization of tertiary propargylic alcohols is driven by the presence of 2 alkyl groups that facilitate the ring closure via the Thorpe–Ingold effect. Currently, the carbonation of secondary or primary propargylic alcohols remains very challenging.
Organocatalysts like organophosphorous compounds such as trialkyl- or triphenyl-phosphine can catalyze the carboxylation of propargylic alcohols in supercritical CO2 (PCO2 = 10 MPa and T = 100 °C). With a low (nBu)3P loading rate of 5 mol%, which appears to be the most efficient phosphine, internal and terminal tertiary propargylic alcohols were converted into the corresponding exovinylene cyclic carbonate (also called α-alkylidene cyclic carbonate) in very good yields (up to 11 examples in the yield range of 64–99%) [98]. For internal propargylic alcohols, products with olefins of (Z)-configuration were exclusively observed. The plausible mechanism suggested by the authors proposed a possible phosphine–CO2 adduct formation in supercritical conditions that would help CO2 insertion. Attempts to obtain optically active cyclic carbonates with chiral phosphine proved its efficiency [99]. Phosphorous ylides were also investigated, as they are known to form an adduct with CO2 thanks to the work by Matthews et al. [100]. Internal tertiary propargylic alcohols could be converted into exovinylene cyclic carbonates in good yields at 80 °C under atmospheric pressure of CO2.
N-heterocyclic carbenes (NHC) also successfully catalyze the carboxylation of propargylic alcohols in a similar manner, with typical imidazolium-2-carboxylates acting as intermediates in the CO2 fixation. Exovinylene cyclic carbonates were obtained under supercritical conditions and milder conditions (PCO2 = 4.5 MPa, T = 60 °C) in good yields [101]. To enhance the CO2 activation, N-heterocyclic olefins (NHO) such as ketene aminals are good candidates, since they display a highly polarized C=C bond that can act as a nucleophile, while stabilizing the positive charge due to electron delocalization via the heterocycle. After kinetic considerations, terminal and internal tertiary propargylic alcohols were successfully converted into the corresponding exovinylene cyclic carbonates in good yields (from 50% to 98% for 7 examples) using 5 mol% of NHO at 60 °C under 2 MPa of CO2 pressure in neat conditions. Although NHO/CO2 adduct generation occurs, regarding the increasing CO2 solubility in the reaction media via its fast sequestration, computational studies suggest a mechanism based on OH activation (hemi-carbonate ion formation) rather than nucleophilic attack of the alcohol moiety on the NHO/CO2 adduct [102,103]. In this case, the in-situ-generated [NHOH]+[RCO3] species is more reactive compared to the analogue when NHCs are used. Moreover, NHOs are less basic than NHCs, which limits possible side reactions such as ring-opening of the exovinylene cyclic carbonate by the unreacted propargylic alcohols (see Scheme 22), which is promoted by the formation of a [NHO]+[RO] ionic pair [104].
Amidines and guanidines are efficient promotors for the carboxylation of CO2 onto alcohols. These organic superbases were tested with propargylic alcohols by our group [105]. Mechanistic and kinetic studies were conducted to better understand side reactions (see Scheme 22) and disparities in selectivities. Optimal conditions of 100 °C and 3 MPa CO2 pressure in acetonitrile led to quantitative conversion of terminal tertiary propargylic alcohol with 7-methyl-1,5,7-triazabicyclo[4.4.0]dec-5-ene (MTBD) as the catalyst (5 mol%). By tuning the reaction time, the best selectivity of 80% toward the exovinylene cyclic carbonate was observed. Replacing MTBD with TBD led to the formation of an oxo-alkyl carbonate (resulting from the ring-opening of the exovinylene carbonate by unreacted propargylic alcohol) with a selectivity above 85%. This arose from the capability of TBD to act both as a base and a proton donor, as proven by DFT calculations.
Basic organic salts and ionic liquids with easy tunable structures and basicity by changing both the anion and the cation structure were then investigated as cheap catalysts to selectively design exovinylene cyclic carbonates. The use of ionic liquid also offers the benefit of increasing the CO2 solubility within the reaction medium, which is favorable for the fast and selective coupling of CO2 with propargylic alcohols substrates. Excellent results were obtained, with alkylammonium salts displaying basic counter anions such as acetate, azide, phenolate and isocyanate, with almost quantitative formation of exovinylene cyclic carbonate after 8 h at 80 °C and 3 MPa CO2 pressure in neat conditions [106]. Deeper investigations of the influence of the cation–anion pair of organic salts were later described by our group. Within the range of tested catalysts, the optimum activity was found for a controlled ionic separation pair made of sterically hindered cations (ideally ammonium or phosphonium cations tetrabutyl chains) and basic anions with pKa values of ~4.5–9 to promote the reaction without detrimental effects on the selectivity [107]. Several terminal tertiary propargylic alcohols were successfully converted, as well as bifunctional alkynols, in order to quantitatively produce bis-cyclic carbonates, which are useful monomers for the fabrication of world-class polymers, i.e., poly(thio)carbonates and polyurethanes [108]. Azole–anion-based aprotic ionic liquids such as tetrabutylphosphonium/2-methyl Imidazole, tetrabutylammonium phenolate and phosphonium levulinate are just a few examples of ionic liquid that proved their efficiency to convert various terminal tertiary propargylic alcohols into their corresponding cyclic carbonate, generally under non-demanding conditions (PCO2 = 0.1–5 MPa, T = 25–80 °C, t = 1 to 24 h) [109,110].
Copper-based catalysts. The synergy between a metal salt with a base or basic ionic liquid or salt represents an elegant method to boost the catalytic performance while allowing the selective formation of exovinylene cyclic carbonates. The benefits of using these binary catalysts is due to the dual activation of the alcohol substrate, i.e., the OH activation by the base, ionic liquid or organic salt, which facilitates the formation of a carbonate ion and the triple bond activation by interaction with the metal cation to favor the ring closure of the molecule. As selected examples, an ionic liquid with bio-derived levulinate anions combined with CuI provided series of exovinylene cyclic carbonates from terminal tertiary propargylic alcohols [111] with excellent yields of 80% to 99% at 25 °C and 1 MPa of CO2 within 4 h. CuI also operate in synergy with N,N-diisopropylethylamine at temperatures of 80–120 °C under 4 MPa CO2 to provide series of cyclic carbonates with internal exocyclic olefin in Z configuration, with yields from 43% to 96% [112]. Remarkably, this dual-catalyst system also enabled the conversion of a challenging terminal secondary alcohol (R1 = H; R2 = isopropyl) with 95% yield.
Silver-based catalysts. As with copper salts, Ag salts, which are by far the most widely used metal salts in the literature, have proven their utility in enhancing the efficiency of the catalyst. The pioneering study by Kikuchi et al. illustrated that terminal and internal tertiary propargylic alcohols were converted into cyclic carbonates in very good yields (up to 14 examples) using AgOAc as the catalyst (10 mol%) and DBU (1 eq) as the base at room temperature under 1 MPa of CO2 [113]. A more recent study showed a different selectivity for internal secondary propargylic alcohols by tuning the amount of base [114]. In this particular case, great selectivity toward vinylene carbonate was obtained when lowering the amount of DBU (0.4 eq instead of 1 eq), while the presence of exovinylene carbonates was not observed (see Scheme 23). Other examples using silver-based catalysts such as AgI/OAc [115], silver sulfadiazine/Et4NBr [116], silver oxides with KI and carbenes [117] and silver nanoparticles supported by polymers [118,119,120] proved their efficiency for terminal and internal tertiary propargylic alcohols. To date, only Schaub et al. have developed a remarkable dual catalyst combining a silver salt with DavePhos ligand to synthesize exovinylene cyclic carbonate from challenging primary propargylic alcohols at room temperature and 2 MPa of CO2 in acetonitrile [121]. In this study, the use of a bulky donor ligand (phosphine) provokes an angle compression of the alcohol substrate, which mimics the Thorpe–Ingold effect [122]. The versatility of this elegant catalytic system afforded a large portfolio of exovinylene cyclic carbonates with external or internal olefin (with exclusively Z configuration) from primary, secondary and tertiary substrates in very good yields from 65% to 95%.
Zinc catalyst. Ma et al. reported on the only zinc-based catalysts that provided good efficiency when associated with TEA for the carboxylative cyclization of terminal tertiary propargylic alcohols [123,124]. The ZnI2/NEt3 binary catalyst operated at low temperature (30 °C) and 1 MPa of CO2 and provided the desired cyclic carbonate scaffolds. The major weakness of this system arose from the high content of the Zn salt (20 mol%) and NEt3 (200 mol%). The mechanistic insights provided by DFT calculations suggested that the OH of the substrate was activated by the formation of a propargylic alcohol–ZnI2–Net3 complex, which drove the formation of a carbonate ion. The Zn salt also took part in the ring closure step by interacting with the alkyne, facilitating the formation of the envisioned cyclic carbonate [125] (see Scheme 24).

3. Strategies to Afford CO2-Based Polycarbonates

3.1. Direct Polymerization of CO2 with Diols

The design of polycarbonates (PC) involving the direct condensation of diols with CO2 faces the same bottlenecks as the synthesis of (a)cyclic carbonates, i.e., the formation of water as a by-product and the proper activation of the OH moieties or CO2. A second challenge lies in the intrinsic structural features of the diols, which are involved in the polymerization manifold. Indeed, the length and nature of the spacer between the alcohol moieties are decisive: 1,2- and 1,3-diols are prone to the thermodynamically favored formation of 5- and 6-membered cyclic carbonates, whereas longer spacers or vicinal diols inducing the formation of constrained cyclic carbonate substrates would drive the formation of oligomers or even polymers.

3.1.1. Direct Copolymerization by Dehydrative Polycondensation

The only relevant catalytic system used to fabricate polycarbonates from the direct carboxylative coupling of diols with CO2 was proposed by Tomishige et al. [126]. The synergistic utilization of CeO2 with excess of 2-cyanopyridine, which was already performed well when synthesizing cyclic carbonates from 1,2- and 1,3-diols, was naturally extended to the dehydrative polycondensation of various 1,x-diol substrates (with x > 3). Aliphatic C4–C10 α,ω-diols provided the corresponding polycarbonates with microstructures free of ether defects in 8 h at 130 °C and 5 MPa of CO2. However, the polymer chains displayed low molar masses (Mn~1650 g/mol). With diols with rigid structures such as 1,4-cyclohexanedimethanol and 1,4-benzenedimethanol, oligomers with Mn of only 500–600 g/mol were obtained, while with the dimethyl-substituted 2,5-dimethyl-2,5-hexanediol, no polymer was formed. To surpass these bottlenecks, the same group evaluated the influence of other nitrile desiccants, while at the same time deepening the structural characterization of the polymers. Switching from 2-cyanopyridine to 2-furonitrile enabled the fabrication of polycarbonates with Mn values of up to 3900 g/mol from 1,6-hexane diol with a ten-fold excess of nitrile (compared to the diol) at 130 °C under 5 MPa CO2 pressure. The fine microstructural characterization of the polymers highlighted the presence of chain end defects, causing termination reactions. The amide produced from the hydration of the cyano-derivatives was involved in a “transesterification” reaction with the hydroxyl chain ends to form a polycarbonate (PC) terminated by an unreactive ester moiety. This side reaction was accompanied by the formation of ammonia, which provoked a scission of the PC. This released short dead chains end-capped by an unreactive amide moiety and a second hydroxyl-terminated PC fragment that was capable of further growth [127]. The design of the CeO2 catalyst plays an important role in its activity. In this case, CeO2 nanorods enhanced the activity toward the formation of polybutylene carbonate as the capacity to uptake CO2 was increased, along with increased oxygen vacancy sites [128].

3.1.2. Direct Copolymerization via Alkylation

The polycondensation of CO2 with xylylene glycol using a base/CX4/R3P activating system was the subject of one of the pioneering studies on the synthesis of PCs from diols [129]. Polymers with Mw values up to 3800 g/mol were fabricated with CBr4 and n-Bu3P as polymerization promotors and N-cyclohexyl-N′,N′,N″,N″-tetraethylguanidine (CyTMG) as the base at room temperature in DMF. Remarkably, PCs free of ether linkages defects were synthesized. The scope of diols was further extended to di(ethylene glycol) using Ph3P and CBrCl3 to deliver the corresponding polycarbonate with Mn values of up to 6000 g/mol [130]. As with the synthesis of (a)cyclic organic carbonates, the polymerization manifold relied first on the CO2 fixation onto the diol thanks to the OH activation by the base. The [R3P-CX3]+[X] diadduct formed in situ from the R3P/CX4 dual system underwent an ionic exchange with the generated RCO3 species, leading to the [RCO3-PR3] intermediate. Then, the chain growth proceeded via successive transcarbonation reactions with another diol molecule thanks to the good leaving group ability of the R3P+-O- fragment (see Scheme 25).
In a similar concept, Du and coworkers proposed a straightforward methodology to copolymerize various diols with CO2 using Cs2CO3 as the base (4 eq) and dichloromethane as the co-reagent (6.2 eq) in NMP [131]. Interestingly, the nature of the diol dictates the microstructural features of the polymer backbone toward either a polycarbonate or a poly(ether carbonate), which alternates carbonates and ether linkages. For benzylic diols such as 1,4-benzenedimethanol, a high conversion rate (>99%) was observed after 24 h at 100 °C and a polycarbonate with very few ether linkages was obtained (63.7% yield; Mn = 1800 g/mol). With the aliphatic 1,4-cyclohexanediol, the formation of a polymer (Mn = 5000 g/mol) with a pattern of alternating ether and carbonate linkages representing 93% of the structure was synthesized. The origin of the ether defects arises from side reactions on the chloromethylated intermediate. The polymerization mechanism is analogous to the one described for the cyclic carbonate formation from vicinal diols. A hemi-carbonate anion formed by CO2 fixation onto one alcohol moiety subsequently undergoes an alkylation reaction with dichloromethane. Then, depending on the diol, the chain growth occurs either via the carbonyl attack to provide a carbonate linkage or via the α-methylene attack forming ether linkages (see Scheme 26). By using the DBU/CH2Br2/bmimPF6 ternary system, Jang and coworkers designed biorenewable PCs via copolymerization of a monosaccharide, i.e., α-methyl D-glucopyranoside (MDG), with CO2 [132]. As with most sugars, this polyol possesses both primary and secondary alcohol functions. As both types of alcohols display different reactivity levels, linear polyglycocarbonates with Mn values up to 7000 g/mol were fabricated, without any protection of the hydroxyl moieties. The NMR structural characterization highlighted the absence of branching, as the carbonate linkages mostly emanated from the C-6 primary alcohol and the C-3 or C-2 secondary hydroxyls.

3.1.3. Terpolymerization Methods

To overcome the limitations brought about by the formation of water for the direct diol–CO2 coupling, the terpolymerization approach utilizing dihalides as comonomers was pioneered by Ikeda et al. More precisely, diolate potassium salts generated from diols mixed with α,ω-dibromo compounds and crown ethers led to the formation of polycarbonates under CO2 pressure in dioxane [133,134]. The 18-crown-6-ether is believed to promote the ion separation of diolate salts by forming stable complexes with metal cations. The proposed mechanism relies on CO2 fixation onto the diolate, followed by the nucleophilic substitution of the dihalide by the carboxylate species (see Scheme 27). After experimental investigations, it appears that crown ethers do not play a key role for CO2 fixation onto the diolate, although they considerably increase the kinetic nucleophilic substitution on dihalides. Interestingly, this pathway provided selective polycarbonate chains free of any ether linkages.
Later, this terpolymerization concept was revisited by Gnanou et al. using Cs2CO3 as a base to generate the diolate [135]. As mentioned previously, the so-called “cesium effect” increases the nucleophilicity of the reactive diolate anions, which facilitates the fixation of CO2 and the nucleophilic attack of the carbonate ion onto the halide. The protocol was valid for dihalides with spacers of at least 4 carbons to prevent the risk of back-biting reactions causing the formation of cyclic carbonate by-products. Other bases such as K2CO3, DBU and Li2CO3 were found to be less effective. Under the optimum parameters (T = 100 °C, PCO2 = 1 MPa, t = 48 h and a diol/dihalide/Cs2CO3 ratio of 1:1.05:4), 97% of diols were converted into polycarbonates with Mn values up to 22,000 g/mol and a PDI of 3.6. PCs with excellent Mn values of up to 43,000 g/mol were designed, involving the proper selection of both aliphatic and aromatic diols and dihalides, as was the case for the 1,6-hexanediol–1,4-bis(chloromethyl)benzene couple. A more sterically hindered biorenewable isosorbide bearing two secondary alcohol was also successfully converted into the poly(isosorbide carbonate) (Mn = 18,000 g/mol). Similarly, the terpolymerization of α-methyl-2,3-di-O-methyl d-glucopyranoside (MDMG), a naturally occurring monosaccharide, with dibromobutane or 1,4-bis(chloromethyl)benzene and Cs2CO3 furnished PCs with Mn of 8000 g/mol and 14,000 g/mol, respectively. To obtain deeper insights into the role of the dihalide structure, Bian et al. proposed a comparative study of the use of dihalomethanes and dihaloethanes [136]. As a general result, higher monomer conversions were noted for dihaloethanes than for dihalomethanes. In the case of dihaloethane, alternating polycarbonates incorporating both the diol and dihalide structure are more likely to occur.

3.2. Ring-Opening Polymerization of Cyclic Carbonates

As the synthesis of PCs via direct copolymerization of alcohols with CO2 remains an elusive endeavour, one option to surpass these hurdles lies in the fixation of CO2 into cyclic carbonate structures that undergo facile ring-opening polymerization (ROP). This technique relies on a chain-polymerization mechanism [137], which allegedly allows accurate control of the polymer molar mass, dispersity and properties. Even though most of the studies on the synthesis of polycarbonates via ROP were performed on non-CO2 based cyclic carbonates, the general principles also apply to CO2-based cyclic carbonates (see Section 2.2), as they present identical structural features. Among them, the polymerization enthalpy is one of the key parameters to consider, since it heavily depends on the ring size and ring constraint. The following discussion will focus on the ROP of 5- and 6-membered cyclic carbonates and exovinylene cyclic carbonates, as these substrates are well described in the literature [138].

3.2.1. 5-Membered Cyclic Carbonates

5-membered cyclic carbonates undergo ROP because of their high positive polymerization enthalpy [139,140]. However, by combining a high temperature of 170 °C with metal alkoxides catalysts such as dibutyltin dimethoxide, titanium tetrapropanolate or zirconium tetrabutanolate, polymers were obtained from ethylene carbonate with yields from 55% to 100% and Mn values ranging from 3000 to 22,000 g/mol. The structural elucidation of the polymers showed the typical features of poly(ethylene oxide-co-ethylene carbonate). Interestingly, the content of carbonate linkages could be controlled via the choice of the catalysts, and the incorporation of 50 mol% of ethylene carbonate units within the polymer backbone was achieved using dibutyltin dimethoxide (catalytic loading of 2 mol%). A careful investigation of the polymerization enthalpy and entropy to afford homopolycarbonates indicated that the formation of the cyclic monomer was more favored than the polymerization itself. However, producing poly(ethylene oxide-co-ethylene carbonate) is slightly more favored. In fact, the polymerization mechanism involves a CO2 loss to afford the ether repeating unit, which compensates the entropy loss and facilitates the formation of an alternating polymer. On the one hand, at 170 °C, an enthalpy of 112.5 kJ/mol was calculated for the formation of pure poly(ethylene carbonate). On the other hand, the same polymerization enthalpy dropped to −11.7 kJ/mol for a random poly(ethylene oxide-co-ethylene carbonate). Deeper investigations of the mechanism for the ROP of ethylene carbonate were conducted by Storey and Hoffman [141], who highlighted the role of the catalysts, namely organotin catalysts and sodium stannate trihydrate catalysts. As with the ROP of lactones [142,143], organotin complexes act both as initiators and catalysts for the ROP of ethylene carbonate. For sodium stannate trihydrate, the use of a diol as an initiator was required. In both cases, the mechanism relies on nucleophilic attack of the initiator on the carbonyl of the cyclic carbonates, generating a reactive alcoholate moiety which will further attack another carbonyl group (Scheme 28, blue pathway). Due to the high polymerization enthalpy of this reaction, a competitive pathway (Scheme 28, red pathway) is also likely to occur, leading to formation of ether linkages driven by the loss of CO2. More insights were given by Clements [144] on the effect of the initiator’s structure. As an example, aromatic alcohols involve a systematic CO2 loss in the initiation step, whereas aliphatic alcohols preferentially favor the formation of a carbonate linkage. Lee et al. also described the several hypothetical pathways for the formation of ether linkages in the ROP of ethylene carbonate, as well as possible depolymerisation processes, using KOH as an initiator [145]. Other catalysts such as phosphazenes [146] and ZnEt2 [147] were tested to design polymers via the ROP of various 5-membered cyclic carbonates. Both catalysts required high polymerization temperature (T = 180 °C) to display sufficient activity and only provided poly(ether-co-carbonates) with low carbonate contents of 21–54% and Mn values of 4300–14,000 g/mol. With ionic liquids, the ROP of ethylene carbonate proceeded at relatively low temperature (from 80 to 120 °C) but only afforded oligomers with low carbonate contents of up to 34%. Haba and Guillaume introduced an elegant concept based on the increase in ring strain of 5-membered cyclic carbonates to construct ether-free PCs by direct ROP. A known example is the ROP of ethylene carbonate fused in a trans configuration onto a cyclohexane ring, i.e., trans-cyclohexane carbonate [148,149], whereas its cis-analogue does not polymerize [150]. The Zn(diphenolate)-promoted ROP of this bicyclic monomer provided a pure isotactic poly(cyclohexene carbonate) with Mn values of up to 21,000 g/mol at 60 °C. Similarly, 4,6-O-benzylidene-2,3-O-carbonyl-r-d-glucopyranoside, i.e., ethylene carbonate fused onto a biorenewable glucopyranoside, furnished polycarbonates with Mn values of up to 16,000 g/mol at 25 °C using a superbase organocatalyst [151].

3.2.2. 6-Membered Cyclic Carbonates

The ROP of 6-membered cyclic carbonates is thermodynamically much more favored than their 5-membered analogues due to increased ring strain, yielding in most cases a negative polymerization enthalpy that compensates the positive polymerization entropy. Indeed, the carbonyl group displays a flat geometry, which increases the strain of the 6-membered cyclic compounds [152]. The ROP of 6-membered cyclic carbonates was extensively reviewed by others and will not be discussed in detail in this section. Cationic polymerization utilizes Brønsted or Lewis acids to activate the monomer and generate species capable of propagation. However, in many cases, propagation is accompanied by decarboxylation, which induces ether defects in the polymer skeleton, as well as backbiting termination reactions, as was first demonstrated by Kricheldorf et al. in their attempt to polymerize trimethylene carbonate using methyl triflate as the initiator [138,153]. Efforts to prevent such decarboxylation reactions and to obtain pure polycarbonates were successfully conducted with alkyl halides initiators (namely methyl iodide, benzyl bromide and allyl iodide) [154], methanesulfonic acids and phosphoric acids [155,156], but only afforded polymers with Mn values in the range of 3000–16,000 g/mol.
Anionic ROP provided better polymerization control and enabled the synthesis of high molar mass polycarbonates. Alkali metal alcoholates and sec-butyllithium [157] were efficient initiators, yielding ether-free polymer chains. However, backbiting termination reactions could not be supressed and oligomers with cyclic structures were concomitantly formed. Taking advantage of the facile ROP with anionic mechanisms, substituted 6-membered cyclic carbonates with pendant groups such as amino-acid (Mn values of up to 24,000 g/mol) [158], 2-ethyl-2-hydroxymethyl [159], norbornene [160], aromatic [161] and more [162,163] were polymerized efficiently.
To prevent non-desired side reactions arising from the high reactivity of the ionic active centers in anionic and cationic ROP of 6-membered cyclic carbonates, coordination–insertion mechanisms have been proposed and studied in detail by Kricheldorf et al. and Darensbourg et al. Organometallic complexes containing Al [164], Ca [165,166,167], Sn [168,169,170,171] or lanthanides [172] were engineered and showed great activity for the synthesis of defect-free polycarbonates. The general polymerization mechanism is based on (i) the ring-opening of the cyclic carbonate via the addition of a nucleophile initiator to the metal-activated carbonyl of the ester group, followed by the (ii) monomer insertion into the oxygen–metal bond generated after initiation, which is then (iii) prone to propagation. The authors mention that in the propagation step, the acyl–oxygen bond is broken rather than the alkyl–oxygen bond for monomer insertion [165]. These coordination–insertion ROP provide a “quasi-living” mechanism, where a linear increase in polymer Mn with the monomer conversion is observed.
Organo-ROP represents the most attractive route to design metal-free PCs with controlled molecular parameters and architectures. Systems comprising benzyl alcohol as initiators and catalysts derived from amidines, guanidines, n-heterocyclic carbenes or olefins (eventually coupled to hydrogen bond donors such as thioureas) successfully polymerize trimethylene carbonate in a very controlled manner at ambient temperatures [173,174,175]. In an attempt to produce CO2-sourced polycarbonates, Buchard et al. designed 6-membered cyclic carbonates via the carbonation of polyol sugars (see Section 2.2.1.2) and further studied their organocatalytic ring-opening polymerization using TBD as the catalyst. Promising results were obtained from D-mannose (Mn values of up to 33,000 g/mol) [176], 2-deoxy-D-ribose (Mn values of up to 42,000 g/mol) [177], thymidine (Mn values of up to 17,000 g/mol) [178] and others [179]. Similarly, Gnanou et al. also reported on the ROP of 6-membered sugar-based cyclic carbonates obtained from the coupling of glycosides and CO2 (based on the DBU/CH2Br2 carbonation protocol at 70 °C). Various ROP organocatalysts were selected, i.e., TBD/4-methyl benzyl alcohol and DBU/1-(3,5-bis(trifluoromethyl)phenyl)-3-cyclohexylthiourea/4-methyl benzyl alcohol, yielding pure polycarbonates with Mn values of up to 4300 g/mol [180].

3.3. Polycondensation of Alcohols with (a)Cyclic Carbonates

While ROP is an appealing method to produce polycarbonates in a controlled manner, it is limited to a narrow library of aliphatic polymer microstructures, due to the thermodynamic constraints, which mostly restrict the polymerization to 6-membered cyclic carbonates. Even though substrates with a large variety of pendant groups have successfully been polymerized by ROP, the urge to develop more diversified architectural structures with cheap and easily scalable processes has brought about step-growth polymerization as a practical solution for various industries. Currently, the mass production of polycarbonates involves the polycondensation of diols with phosgene. DMC, DEC or DPC, which are easily accessible from CO2 (see Section 2.1), are now emerging as valuable alternatives to phosgene, and their polycondensation with naturally occurring ubiquitous diols offers a route for the design of polycarbonates with reduced carbon footprints. Attempts to valorize 5-membered cyclic carbonates as useful carbonylation synthons for the production of polycarbonates were also reported but remained elusive. As highlighted for ROP, they generally face the same limitations, such as competitive methylene and carbonyl attacks, generating poly(ether-co-carbonates) [181,182,183].

3.3.1. DMC and DPC as Carbonylating Agents

The formation of polycarbonates from diols and DMC or DPC by polycondensation generally involves a two-step approach, usually performed in a one-pot fashion and in solvent-free conditions (see Scheme 29). First, oligocarbonates of low molar masses are fabricated via transcarbonation between the diol and the acyclic carbonate (DMC or DPC) at moderate temperatures (high enough to maintain a molten reaction media) and ambient pressure. The resulting telechelic oligomers end-capped with both alcohol and carbonate moieties are then chain-extended by applying a high vacuum shifting the reaction equilibrium (step-growth polymerization mechanism) via elimination of methanol (for DMC) or phenol (for DPC). Temperatures are progressively increased during the process to overcome mass transfer limitations, reflecting the viscosity and molar mass increases. The nature of the initial diol(s) monomer dictates the final polymer microstructure (i.e., aliphatic or non-aliphatic polycarbonates).

Aliphatic PC

Aliphatic PCs have recently gained huge interest in research thanks to their biodegradability and biocompatibility, making these materials suitable for the biomedical field [184]. Depending on the structure of the aliphatic diols, the occurrence of side reactions was highlighted by Xie and coworkers when dialkylcarbonate was used as the carbonylating agent [185], limiting the access to high molar mass aliphatic PCs. However, side alkylation reactions of the alcohol chain end either via the methylene attacks of DEC or via backbiting side reactions furnished polycarbonates with unreactive ether chain ends or cyclic by-products, causing the death of the chain growth. These side reactions were especially highlighted for poly(ethylene carbonate) and poly(propylene carbonate) made from ethylene glycol and 1,3-propanediol, but were considerably limited or even totally supressed when diols with spacers larger than 3 carbon units were used as monomers. Following this approach, aliphatic PCs with moderate to high molar masses were successfully synthetized from a diverse library of linear alkyl diols with DMC when suitable catalysts were used. For example, the utilization of a binary DMAP/thioureas organocatalyst provided poly(butylene carbonate) with Mn values of up to 52,000 g/mol [186,187], while imidazolium or TBD furnished poly(hexamethylene-co-trimethylene carbonate) with Mn values of up to 7400 g/mol [188] and poly(hexamethylene carbonate) with Mn up 33,000 g/mol [189].
Metal-based catalysts such as sodium alkoxide Li(Acac), Na(Acac) and K(Acac) were also reported as efficient systems for constructing polycarbonates, with diversified microstructures and Mw values of up to 200,000 g/mol [190,191], while TiO2/SiO2-based catalysts provided poly(hexamethylene carbonate) with Mw of 160,000 g/mol and terpolymers with Mw values of up to 87,000 g/mol [192,193]. Interestingly, this concept was extended to the fabrication of hyperbranched PCs when blends of diols and triols were condensed with DMC using DMAP or Li(Acac) as catalysts [194].
Unlike DMC and DEC, the utilization of DPC enabled suppression of the side alkylations of the hydroxyl chain ends of growing PCs. Moreover, phenol may be viewed as a better leaving group than methanol for DMC, which should facilitate the construction of high molar mass polymers. However, phenol displays a high boiling point so that the polycondensation was generally performed at temperatures over 200 °C. As selected examples, poly(butylene carbonate)s with Mw values of up to 156,000 g/mol or 202,000 g/mol were synthesized using Zn(OAc)2 or basic solid MgO as catalysts, respectively, at temperatures of 200–210 °C [195,196].
A different approach was proposed by Miller and Vanderhenst, who first synthetized decane-1,10-diyl dimethyl dicarbonate from 1,10-decanediol with an excess of DMC, catalyzed by Na2CO3 at 85 °C. The bifunctional carbonate was then subjected to self-polycondensation and aliphatic polycarbonates were obtained via carbonate metathesis polymerization after isolation of these bis-carbonates. Mw values of up to 88,000 g/mol were observed with Zn(OAc)2 as the catalyst and a reaction temperature of 150 °C for the oligomerization step (12 h) and 175 °C for the chain extension (1.5 h). The mechanism proposed by the authors is presented in Scheme 30 with K2CO3 as the catalyst. A one-pot process with the diol and an excess of DMC as the starting reagent was proposed to afford aliphatic PC with Mw values of up to 56,000 g/mol [197].

Aromatic and Cycloaliphatic PCs

Aromatic PC displays remarkable thermo-mechanical and optic properties useful for plethora of applications within the construction, automotive, electronics and other sectors. Historically, high molar mass polycarbonates were firstly industrialized via the phosgenation of bisphenol A (BPA). The toxicity and safety issues linked to the utilization of phosgene pushed scientists to evaluate the potential of DMC and DPC for constructing aromatic PCs from BPA. In combination with molecular sieves, a dual DMAP/(Bu2SnCl)2O catalytic system afforded PC with a high-molecular weight (Mw values of up to 75,000 g/mol) using BPA and DMC, but a very high temperature of 320 °C was needed to maintain the molten reaction medium, leading to poor control of the dispersity (D = 3.8) [198]. Lithium- and lanthanum-based catalysts were also reported and proved their efficiency in accessing to aromatic polycarbonates from BPA [199,200]. The utilization of DPC as a carbonylation agent is now emerging at the industrial level (also known as the Asahi–Kasei process) to construct PCs derived from BPA [201]. Briefly, a carbonate source is brought about by the coupling reaction between CO2 and ethylene oxide, affording ethylene carbonate, which further undergoes methanolysis to produce DMC. The latter is converted into DPC using phenol. The melt polycondensation responds to a two-step procedure with a first oligomerization phase between DPC and BPA at 230 °C, followed by a chain extension at 265 °C to design aromatic PCs with molar mass values of up to 117,000 g/mol.
In the quest for sustainable and less toxic alternatives to BPA-based PC, emerging research studies are now focusing on the utilization of sugar-based bicyclic diols, i.e., 1,4:3,6-dianhydro-D-sorbitol, also known as isosorbide, to construct the new generation of PCs with thermomechanical properties competing with those of aromatic PC.
Due to the bicyclic structure (two fused tetrahydrofuran rings) of isosorbide, the two alcohol functions are not equivalent. The so-called “endo” OH displays intramolecular H-bonding with the O atom of a tetrahydrofuran ring, inducing disparity in the two OH reactivities (see Scheme 31). In this regard, a large range of catalysts was tested to selectively activate the endo or exo alcohol moieties and to prevent reactivity disparities that would considerably limit the molar mass of the PC. Among them, metal salts [202], metal alkoxides [203,204], inorganic carbonates [205], ionic liquids [206,207,208,209,210,211,212,213], organic bases [214] and others [215,216,217,218] enabled the facile construction of defect-free PCs derived from isosorbide with Mn values in the range of 14,000 to 55,000 g/mol.

3.3.2. Polycondensation between Bis(α-Alkylidene Cyclic Carbonates) and Diols

Unlike conventional 5-membered cyclic carbonates that poorly react with alcohol in “soft” experimental conditions, CO2-based exovinylene cyclic carbonates (see Section 2.2.4) undergo facile and regioselective alcoholysis, even at room temperature. The thermodynamic driving force of the ring-opening lies in the formation of a keto-enol equilibrium (see Scheme 32).
Taking advantage of this concept, new CO2-based bis(α-alkylidene cyclic carbonates) were designed by our group. Bis(α-alkylidene cyclic carbonates) were either prepared by direct coupling of bis-propagylic alcohols with CO2 (aliphatic spacer R2, see Scheme 33) or by heck coupling of two CO2-based exovinylene cyclic carbonates with 1,4-diiodobenzene (aromatic spacer R2, see Scheme 33). These new monomers were further copolymerized via step growth with diols to obtain regioregular PCs with microstructures free of any defects [108,219,220,221]. Indeed, unlike conventional 5-membered cyclic carbonates, the formation of ether linkages (via decarboxylation) was not observed. New poly(β-oxo carbonates) with Mn values ranging from 5500 to 25,000 g/mol were fabricated using DBU as the catalyst in DMF or chloroform (see Scheme 33). This concept was extended to the formation of poly(oxo-urethanes) and poly(hydroxyl-oxazolidone) [108]. Aliphatic poly(oxo-carbonates) synthetized with OH-terminated poly(ethylene glycol) and bis(α-alkylidene cyclic carbonates) with Mn values of up to 60,000 g/mol showed great properties for potential applications as solid electrolytes for solid Li-ion batteries [219]. An attempt to synthesize poly(β-oxo carbonates) in a one-step or one-pot strategy directly from the terpolymerization of diols, bis-propargylic alcohols and CO2 was recently reported. The key to the success lies in the utilization of a single dual system, i.e., tetrabutylammonium phenolate/AgI, capable of catalyzing two different reactions, i.e., the carboxylative cyclization of CO2 with the bispropargylic alcohol to deliver the bis(α-alkylidene cyclic carbonates) in situ, followed by its polymerization in cascade with a primary diol. Even though the results were promising, only oligomers with Mn values in the range of 1000 to 2700 g/mol were obtained due to the occurrence of side reactions [220].

4. Conclusions

This paper reviews the main strategies to afford (a)cyclic carbonates and polycarbonates from the coupling of alcohols and CO2 or CO2-based synthons. Compared to the well-known coupling of CO2 with epoxides, these systems suffer from limitations in terms of the yields and demanding experimental conditions, due to the formation of water as a by-product, which considerably affects the formation of carbonates. However, the challenging coupling of alcohols and CO2 is still very interesting, as it provides a large manifold of accessible and easy-to-handle substrates, opening the possibility for the formation of carbonates that are hardly affordable with epoxide–CO2 coupling. Many routes have been proposed to overcome the formation of water, combining dehydration agents with catalysts or co-reagents with excess base to trap the water formation. In all cases, the synthesis of (a)cyclic carbonates or polycarbonates could not be achieved without the use of stoichiometric or excess amounts of additives, as the quantitative removal of water is needed to achieve satisfying yields.

Author Contributions

Conceptualization, C.J., C.D. and T.T.; writing—original draft preparation, A.B. and B.G.; writing—review and editing, A.B., B.G. and T.T.; visualization, A.B., R.M. and C.D.; supervision, C.J. and T.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was equally funded by the Région Nouvelle Aquitaine, grant number 2015-0412, and the University of Liège. The authors from Liege also thank the "Fonds National pour la Recherche Scientifique” (F.R.S.-FNRS) and the Fonds Wetenschappelijk Onderzoek—Vlaanderen (FWO) for financial support in the frame of the EOS project n°O019618F (ID EOS: 30902231).

Data Availability Statement

Not applicable.

Acknowledgments

The authors wish to thank the Conseil Regional d‘Aquitaine and the University of Liege for the funding of the PhD fellowship of A. Brege through the international doctoral school program EJD-Funmat.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stocker, T.F. Intergovernmental Panel on Climate Change 2013: Physical Science Basis, Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK, 2013. [Google Scholar]
  2. Friedlingstein, P.; O’sullivan, M.; Jones, M.W.; Andrew, R.M.; Hauck, J.; Olsen, A.; Peters, G.P.; Peters, W.; Pongratz, J.; Sitch, S.; et al. Global Carbon Budget 2020. Earth Syst. Sci. Data 2020, 12, 3269–3340. [Google Scholar] [CrossRef]
  3. Jackson, R.B.; Le Quéré, C.; Andrew, R.M.; Canadell, J.G.; Peters, G.P.; Roy, J.; Wu, L. Warning signs for stabilizing global CO2 emissions. Environ. Res. Lett. 2017, 12, 110202. [Google Scholar] [CrossRef] [Green Version]
  4. Artz, J.; Müller, T.E.; Thenert, K.; Kleinekorte, J.; Meys, R.; Sternberg, A.; Bardow, A.; Leitner, W. Sustainable Conversion of Carbon Dioxide: An Integrated Review of Catalysis and Life Cycle Assessment. Chem. Rev. 2018, 118, 434–504. [Google Scholar] [CrossRef]
  5. Maeda, C.; Miyazaki, Y.; Ema, T. Recent progress in catalytic conversions of carbon dioxide. Catal. Sci. Technol. 2014, 4, 1482–1497. [Google Scholar] [CrossRef] [Green Version]
  6. Tappe, N.A.; Reich, R.M.; D’elia, V.; Kühn, F.E. Current advances in the catalytic conversion of carbon dioxide by molecular catalysts: An update. Dalton Trans. 2018, 47, 13281–13313. [Google Scholar] [CrossRef]
  7. Ruscic, B.; Feller, D.; Peterson, K.A. Active Thermochemical Tables: Dissociation energies of several homonuclear first-row diatomics and related thermochemical values. Theor. Chem. Acc. 2013, 133, 1415. [Google Scholar] [CrossRef] [Green Version]
  8. Martín, C.; Fiorani, G.; Kleij, A.W. Recent Advances in the Catalytic Preparation of Cyclic Organic Carbonates. ACS Catal. 2015, 5, 1353–1370. [Google Scholar] [CrossRef]
  9. North, M.; Pasquale, R.; Young, C. Synthesis of cyclic carbonates from epoxides and CO2. Green Chem. 2010, 12, 1514–1539. [Google Scholar] [CrossRef]
  10. Alves, M.; Grignard, B.; Boyaval, A.; Méreau, R.; De Winter, J.; Gerbaux, P.; Detrembleur, C.; Tassaing, T.; Jérôme, C. Organocatalytic Coupling of CO2 with Oxetane. ChemSusChem 2017, 10, 1128–1138. [Google Scholar] [CrossRef]
  11. Rintjema, J.; Guo, W.; Martin, E.; Escudero-Adan, E.C.; Kleij, A.W. Highly Chemoselective Catalytic Coupling of Substituted Oxetanes and Carbon Dioxide. Chem. Eur. J. 2015, 21, 10754–10762. [Google Scholar] [CrossRef]
  12. Katie, J.L.; Ian, D.V.I.; Michael, N.; Mani, S. Valorization of Carbon Dioxide into Oxazolidinones by Reaction with Aziridines. Curr. Green Chem. 2019, 6, 32–43. [Google Scholar]
  13. Yang, Z.-Z.; He, L.-N.; Gao, J.; Liu, A.-H.; Yu, B. Carbon dioxide utilization with C–N bond formation: Carbon dioxide capture and subsequent conversion. Energy Environ. Sci. 2012, 5, 6602. [Google Scholar] [CrossRef]
  14. Harder, A.; Escher, B.I.; Landini, P.; Tobler, N.B.; Schwarzenbach, R.P. Evaluation of Bioanalytical Assays for Toxicity Assessment and Mode of Toxic Action Classification of Reactive Chemicals. Environ. Sci. Technol. 2003, 37, 4962–4970. [Google Scholar] [CrossRef]
  15. Kostal, J.; Voutchkova-Kostal, A.; Weeks, B.; Zimmerman, J.B.; Anastas, P.T. A Free Energy Approach to the Prediction of Olefin and Epoxide Mutagenicity and Carcinogenicity. Chem. Res. Toxicol. 2012, 25, 2780–2787. [Google Scholar] [CrossRef] [PubMed]
  16. Santos, B.A.V.; Silva, V.M.T.M.; Loureiro, J.M.; Barbosa, D.; Rodrigues, A.E. Modeling of physical and chemical equilibrium for the direct synthesis of dimethyl carbonate at high pressure conditions. Fluid Phase Equilib. 2012, 336, 41–51. [Google Scholar] [CrossRef]
  17. Santos, B.a.V.; Pereira, C.S.M.; Silva, V.M.T.M.; Loureiro, J.M.; Rodrigues, A.E. Kinetic study for the direct synthesis of dimethyl carbonate from methanol and CO2 over CeO2 at high pressure conditions. Appl. Catal. Gen. 2013, 455, 219–226. [Google Scholar] [CrossRef]
  18. Müller, K.; Mokrushina, L.; Arlt, W. Thermodynamic Constraints for the Utilization of CO2. Chem. Ing. Technol. 2014, 86, 497–503. [Google Scholar] [CrossRef]
  19. Kizlink, J.; Pastucha, I. Preparation of Dimethyl Carbonate from Methanol and Carbon Dioxide in the Presence of Organotin Compounds. Collect. Czech. Chem. Commun. 1994, 59, 2116–2118. [Google Scholar] [CrossRef]
  20. Kizlink, J. Synthesis of Dimethyl Carbonate from Carbon Dioxide and Methanol in the Presence of Organotin Compounds. Collect. Czech. Chem. Commun. 1993, 58, 1399–1402. [Google Scholar] [CrossRef]
  21. Kizlink, J.; Pastucha, I. Preparation of Dimethyl Carbonate from Methanol and Carbon Dioxide in the Presence of Sn(IV) and Ti(IV) Alkoxides and Metal Acetates. Collect. Czech. Chem. Commun. 1995, 60, 687–692. [Google Scholar] [CrossRef]
  22. Zhao, T.; Han, Y.; Sun, Y. Novel reaction route for dimethyl carbonate synthesis from CO2 and methanol. Fuel Process. Technol. 2000, 62, 187–194. [Google Scholar] [CrossRef]
  23. Aresta, M.; Dibenedetto, A.; Pastore, C.; Pápai, I.; Schubert, G. Reaction mechanism of the direct carboxylation of methanol to dimethylcarbonate: Experimental and theoretical studies. Top. Catal. 2006, 40, 71–81. [Google Scholar] [CrossRef]
  24. Du, Y.; He, L.-N.; Kong, D.-L. Magnesium-catalyzed synthesis of organic carbonate from 1,2-diol/alcohol and carbon dioxide. Catal. Commun. 2008, 9, 1754–1758. [Google Scholar] [CrossRef]
  25. Tomishige, K.; Sakaihori, T.; Ikeda, Y.; Fujimoto, K. A novel method of direct synthesis of dimethyl carbonate from methanol and carbon dioxide catalyzed by zirconia. Catal. Lett. 1999, 58, 225–229. [Google Scholar] [CrossRef]
  26. Tomishige, K.; Ikeda, Y.; Sakaihori, T.; Fujimoto, K. Catalytic properties and structure of zirconia catalysts for direct synthesis of dimethyl carbonate from methanol and carbon dioxide. J. Catal. 2000, 192, 355–362. [Google Scholar] [CrossRef]
  27. Ballivet-Tkatchenko, D.; Dos Santos, J.H.Z.; Philippot, K.; Vasireddy, S. Carbon dioxide conversion to dimethyl carbonate: The effect of silica as support for SnO2 and ZrO2 catalysts. C. R. Chim. 2011, 14, 780–785. [Google Scholar] [CrossRef]
  28. Ikeda, Y.; Asadullah, M.; Fujimoto, K.; Tomishige, K. Structure of the Active Sites on H3PO4/ZrO2 Catalysts for Dimethyl Carbonate Synthesis from Methanol and Carbon Dioxide. J. Phys. Chem. 2001, 105, 10653–10658. [Google Scholar] [CrossRef]
  29. Ikeda, Y.; Sakaihori, T.; Tomishige, K.; Fujimoto, K. Promoting effect of phosphoric acid on zirconia catalysts in selective synthesis of dimethyl carbonate from methanol and carbon dioxide. Catal. Lett. 2000, 66, 59–62. [Google Scholar] [CrossRef]
  30. Yoshida, Y.; Arai, Y.; Kado, S.; Kunimori, K.; Tomishige, K. Direct synthesis of organic carbonates from the reaction of CO2 with methanol and ethanol over CeO2 catalysts. Catal. Today 2006, 115, 95–101. [Google Scholar] [CrossRef]
  31. Wang, S.; Zhao, L.; Wang, W.; Zhao, Y.; Zhang, G.; Ma, X.; Gong, J. Morphology control of ceria nanocrystals for catalytic conversion of CO2 with methanol. Nanoscale 2013, 5, 5582–5588. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, Z.-F.; Liu, Z.-T.; Liu, Z.-W.; Lu, J. DMC Formation over Ce0.5Zr0.5O2 Prepared by Complex-decomposition Method. Catal. Lett. 2009, 129, 428–436. [Google Scholar] [CrossRef]
  33. Lee, H.J.; Park, S.; Song, I.K.; Jung, J.C. Direct Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide over Ga2O3/Ce0.6Zr0.4O2 Catalysts: Effect of Acidity and Basicity of the Catalysts. Catal. Lett. 2011, 141, 531–537. [Google Scholar] [CrossRef]
  34. Lee, H.J.; Park, S.; Jung, J.C.; Song, I.K. Direct synthesis of dimethyl carbonate from methanol and carbon dioxide over H3PW12O40/CeXZr1−XO2 catalysts: Effect of acidity of the catalysts. Korean J. Chem. Eng. 2011, 28, 1518–1522. [Google Scholar] [CrossRef]
  35. Zhou, Y.; Wang, S.; Xiao, M.; Han, D.; Lu, Y.; Meng, Y. Novel Cu–Fe bimetal catalyst for the formation of dimethyl carbonate from carbon dioxide and methanol. RSC Adv. 2012, 2, 6831–6837. [Google Scholar] [CrossRef]
  36. Bian, J.; Xiao, M.; Wang, S.J.; Lu, Y.X.; Meng, Y.Z. Graphite oxide as a novel host material of catalytically active Cu–Ni bimetallic nanoparticles. Catal. Commun. 2009, 10, 1529–1533. [Google Scholar] [CrossRef]
  37. Bian, J.; Wei, X.W.; Wang, L.; Guan, Z.P. Graphene nanosheet as support of catalytically active metal particles in DMC synthesis. Chin. Chem. Lett. 2011, 22, 57–60. [Google Scholar] [CrossRef]
  38. La, K.W.; Youn, M.H.; Chung, J.S.; Baeck, S.H.; Song, I.K. Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide by Heteropolyacid/Metal Oxide Catalysts. Solid State Phenom. 2007, 119, 287–290. [Google Scholar] [CrossRef]
  39. Li, C.-F.; Zhong, S.-H. Study on application of membrane reactor in direct synthesis DMC from CO2 and CH3OH over Cu–KF/MgSiO catalyst. Catal. Today 2003, 82, 83–90. [Google Scholar] [CrossRef]
  40. Choi, J.-C.; He, L.-N.; Yasuda, H.; Sakakura, T. Selective and high yield synthesis of dimethyl carbonate directly from carbon dioxide and methanol. Green Chem. 2002, 4, 230–234. [Google Scholar] [CrossRef]
  41. George, J.; Patel, Y.; Pillai, S.M.; Munshi, P. Methanol assisted selective formation of 1,2-glycerol carbonate from glycerol and carbon dioxide using nBu2SnO as a catalyst. J. Mol. Catal. Chem. 2009, 304, 1–7. [Google Scholar] [CrossRef]
  42. Iwakabe, K.; Nakaiwa, M.; Sakakura, T.; Choi, J.-C.; Yasuda, H.; Takahashi, T.; Ooshima, Y. Reaction Rate of the Production of Dimethyl Carbonate Directly from the Supercritical CO2 and Methanol. J. Chem. Eng. Jpn. 2005, 38, 1020–1024. [Google Scholar] [CrossRef]
  43. Zhang, Z.; Liu, S.; Zhang, L.; Yin, S.; Yang, G.; Han, B. Driving dimethyl carbonate synthesis from CO2 and methanol and production of acetylene simultaneously using CaC2. Chem. Commun. 2018, 54, 4410–4412. [Google Scholar] [CrossRef]
  44. Tamura, M.; Satsuma, A.; Shimizu, K.-I. CeO2-catalyzed nitrile hydration to amide: Reaction mechanism and active sites. Catal. Sci. Technol. 2013, 3, 1386–1393. [Google Scholar] [CrossRef]
  45. Tomishige, K.; Furusawa, Y.; Ikeda, Y.; Asadullah, M.; Fujimoto, K. CeO2–ZrO2 Solid Solution Catalyst for Selective Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide. Catal. Lett. 2001, 76, 71–74. [Google Scholar] [CrossRef]
  46. Lamotte, J.; Morávek, V.; Bensitel, M.; Lavalley, J.C. FT-IR study of the structure and reactivity of methoxy species on ThO2 and CeO2. React. Kinet. Catal. Lett. 1988, 36, 113–118. [Google Scholar] [CrossRef]
  47. Badri, A.; Binet, C.; Lavalley, J.-C. Use of methanol as an IR molecular probe to study the surface of polycrystalline ceria. J. Chem. Soc. Faraday Trans. 1997, 93, 1159–1168. [Google Scholar] [CrossRef]
  48. Binet, C.; Daturi, M.; Lavalley, J.-C. IR study of polycrystalline ceria properties in oxidised and reduced states. Catal. Today 1999, 50, 207–225. [Google Scholar] [CrossRef]
  49. Honda, M.; Kuno, S.; Begum, N.; Fujimoto, K.-I.; Suzuki, K.; Nakagawa, Y.; Tomishige, K. Catalytic synthesis of dialkyl carbonate from low pressure CO2 and alcohols combined with acetonitrile hydration catalyzed by CeO2. Appl. Catal. Gen. 2010, 384, 165–170. [Google Scholar] [CrossRef]
  50. Honda, M.; Suzuki, A.; Noorjahan, B.; Fujimoto, K.-I.; Suzuki, K.; Tomishige, K. Low pressure CO2 to dimethyl carbonate by the reaction with methanol promoted by acetonitrile hydration. Chem. Commun. 2009, 30, 4596–4598. [Google Scholar] [CrossRef] [PubMed]
  51. Honda, M.; Kuno, S.; Sonehara, S.; Fujimoto, K.-I.; Suzuki, K.; Nakagawa, Y.; Tomishige, K. Tandem Carboxylation-Hydration Reaction System from Methanol, CO2 and Benzonitrile to Dimethyl Carbonate and Benzamide Catalyzed by CeO2. ChemCatChem 2011, 3, 365–370. [Google Scholar] [CrossRef]
  52. Tamura, M.; Shimizu, K.-I.; Satsuma, A. Comprehensive IR study on acid/base properties of metal oxides. Appl. Catal. Gen. 2012, 4339, 135–145. [Google Scholar] [CrossRef]
  53. Honda, M.; Tamura, M.; Nakagawa, Y.; Sonehara, S.; Suzuki, K.; Fujimoto, K.-I.; Tomishige, K. Ceria-Catalyzed Conversion of Carbon Dioxide into Dimethyl Carbonate with 2-Cyanopyridine. ChemSusChem 2013, 6, 1341–1344. [Google Scholar] [CrossRef] [PubMed]
  54. Honda, M.; Tamura, M.; Nakagawa, Y.; Nakao, K.; Suzuki, K.; Tomishige, K. Organic carbonate synthesis from CO2 and alcohol over CeO2 with 2-cyanopyridine: Scope and mechanistic studies. J. Catal. 2014, 318, 95–107. [Google Scholar] [CrossRef]
  55. Salvatore, R.N.; Chu, F.; Nagle, A.S.; Kapxhiu, E.A.; Cross, R.M.; Jung, K.W. Efficient Cs2CO3-promoted solution and solid phase synthesis of carbonates and carbamates in the presence of TBAI. Tetrahedron 2002, 58, 3329–3347. [Google Scholar] [CrossRef]
  56. Shi, M.; Shen, Y.-M. Synthesis of Mixed Carbonates via a Three-Component Coupling of Alcohols, CO2, and Alkyl Halides in the Presence of K2CO3 and Tetrabutylammonium Iodide. Molecules 2002, 7, 386–393. [Google Scholar] [CrossRef]
  57. Lethesh, K.C.; Shah, S.N.; Mutalib, M.I.A. Synthesis, Characterization, and Thermophysical Properties of 1,8-Diazobicyclo[5.4.0]undec-7-ene Based Thiocyanate Ionic Liquids. J. Chem. Eng. Data 2014, 59, 1788–1795. [Google Scholar] [CrossRef]
  58. Wang, Y.; Han, Q.; Wen, H. Theoretical discussion on the mechanism of binding CO2by DBU and alcohol. Mol. Simul. 2013, 39, 822–827. [Google Scholar] [CrossRef]
  59. Villiers, C.; Dognon, J.-P.; Pollet, R.; Thuéry, P.; Ephritikhine, M. An Isolated CO2 Adduct of a Nitrogen Base: Crystal and Electronic Structures. Angew. Chem. Int. Ed. 2010, 49, 3465–3468. [Google Scholar] [CrossRef]
  60. Ma, J.; Zhang, X.; Zhao, N.; Al-Arifi, A.S.N.; Aouak, T.; Al-Othman, Z.A.; Xiao, F.; Wei, W.; Sun, Y. Theoretical study of TBD-catalyzed carboxylation of propylene glycol with CO2. J. Mol. Catal. Chem. 2010, 315, 76–81. [Google Scholar] [CrossRef]
  61. Goodrich, P.; Gunaratne, H.Q.N.; Jin, L.; Lei, Y.; Seddon, K.R. Carbon Dioxide Utilisation for the Synthesis of Unsymmetrical Dialkyl and Cyclic Carbonates Promoted by Basic Ionic Liquids. Aust. J. Chem. 2018, 71, 181–185. [Google Scholar] [CrossRef]
  62. Lim, Y.N.; Lee, C.; Jang, H.-Y. Metal-Free Synthesis of Cyclic and Acyclic Carbonates from CO2and Alcohols. Eur. J. Org. Chem. 2014, 2014, 1823–1826. [Google Scholar] [CrossRef]
  63. Aresta, M.; Dibenedetto, A.; Fracchiolla, E.; Giannoccaro, P.; Pastore, C.; Pápai, I.; Schubert, G. Mechanism of Formation of Organic Carbonates from Aliphatic Alcohols and Carbon Dioxide under Mild Conditions Promoted by Carbodiimides. DFT Calculation and Experimental Study. J. Org. Chem. 2005, 70, 6177–6186. [Google Scholar] [PubMed]
  64. Däbritz, E. Syntheses and Reactions of O,N,N′-Trisubstituted Isoureas. Angew. Chem. Int. Ed. Engl. 1966, 5, 470–477. [Google Scholar] [CrossRef]
  65. Imberdis, A.; Lefèvre, G.; Thuéry, P.; Cantat, T. Metal-Free and Alkali-Metal-Catalyzed Synthesis of Isoureas from Alcohols and Carbodiimides. Angew. Chem. 2018, 130, 3138–3142. [Google Scholar] [CrossRef] [Green Version]
  66. Wang, S.; Zhou, J.; Zhao, S.; Zhao, Y.; Ma, X. Enhancement of Dimethyl Carbonate Synthesis with In Situ Hydrolysis of 2,2-Dimethoxy Propane. Chem. Eng. Technol. 2016, 39, 723–729. [Google Scholar] [CrossRef]
  67. Tomishige, K.; Kunimori, K. Catalytic and direct synthesis of dimethyl carbonate starting from carbon dioxide using CeO2-ZrO2 solid solution heterogeneous catalyst: Effect of H2O removal from the reaction system. Appl. Catal. Gen. 2002, 237, 103–109. [Google Scholar] [CrossRef]
  68. Chang, T.; Tamura, M.; Nakagawa, Y.; Fukaya, N.; Choi, J.-C.; Mishima, T.; Matsumoto, S.; Hamura, S.; Tomishige, K. An effective combination catalyst of CeO2 and zeolite for the direct synthesis of diethyl carbonate from CO2 and ethanol with 2,2-diethoxypropane as a dehydrating agent. Green Chem. 2020, 22, 7321–7327. [Google Scholar] [CrossRef]
  69. Sakakura, T.; Saito, Y.; Okano, M.; Choi, J.-C.; Sako, T. Selective Conversion of Carbon Dioxide to Dimethyl Carbonate by Molecular Catalysis. J. Org. Chem. 1998, 63, 7095–7096. [Google Scholar] [CrossRef]
  70. Zhang, Z.-F.; Liu, Z.-W.; Lu, J.; Liu, Z.-T. Synthesis of Dimethyl Carbonate from Carbon Dioxide and Methanol over CexZr1-xO2 and [EMIM]Br/Ce0.5Zr0.5O2. Ind. Eng. Chem. Res. 2011, 50, 1981–1988. [Google Scholar] [CrossRef]
  71. Putro, W.S.; Ikeda, A.; Shigeyasu, S.; Hamura, S.; Matsumoto, S.; Lee, V.Y.; Choi, J.-C.; Fukaya, N. Sustainable Catalytic Synthesis of Diethyl Carbonate. ChemSusChem 2020, 14, 842–846. [Google Scholar] [CrossRef]
  72. Guidi, S.; Calmanti, R.; Noè, M.; Perosa, A.; Selva, M. Thermal (Catalyst-Free) Transesterification of Diols and Glycerol with Dimethyl Carbonate: A Flexible Reaction for Batch and Continuous-Flow Applications. ACS Sustain. Chem. Eng. 2016, 4, 6144–6151. [Google Scholar] [CrossRef]
  73. Baral, E.R.; Lee, J.H.; Kim, J.G. Diphenyl Carbonate: A Highly Reactive and Green Carbonyl Source for the Synthesis of Cyclic Carbonates. J. Org. Chem. 2018, 83, 11768–11776. [Google Scholar] [CrossRef]
  74. Ochoa-Gómez, J.R.; Gómez-Jiménez-Aberasturi, O.; Ramírez-López, C.; Maestro-Madurga, B. Synthesis of glycerol 1,2-carbonate by transesterification of glycerol with dimethyl carbonate using triethylamine as a facile separable homogeneous catalyst. Green Chem. 2012, 14, 3368. [Google Scholar] [CrossRef]
  75. Schutyser, W.; Van Den Bosch, S.; Renders, T.; De Boe, T.; Koelewijn, S.F.; Dewaele, A.; Ennaert, T.; Verkinderen, O.; Goderis, B.; Courtin, C.M.; et al. Influence of bio-based solvents on the catalytic reductive fractionation of birch wood. Green Chem. 2015, 17, 5035–5045. [Google Scholar] [CrossRef]
  76. Koelewijn, S.F.; Van Den Bosch, S.; Renders, T.; Schutyser, W.; Lagrain, B.; Smet, M.; Thomas, J.; Dehaen, W.; Van Puyvelde, P.; Witters, H.; et al. Sustainable bisphenols from renewable softwood lignin feedstock for polycarbonates and cyanate ester resins. Green Chem. 2017, 19, 2561–2570. [Google Scholar] [CrossRef]
  77. Ma, J.; Liu, J.; Zhang, Z.; Han, B. Mechanisms of ethylene glycol carbonylation with carbon dioxide. Comput. Theor. Chem. 2012, 992, 103–109. [Google Scholar] [CrossRef]
  78. Aresta, M.; Dibenedetto, A.; Nocito, F.; Pastore, C. A study on the carboxylation of glycerol to glycerol carbonate with carbon dioxide: The role of the catalyst, solvent and reaction conditions. J. Mol. Catal. Chem. 2006, 257, 149–153. [Google Scholar] [CrossRef]
  79. Tamura, M.; Honda, M.; Nakagawa, Y.; Tomishige, K. Direct conversion of CO2 with diols, aminoalcohols and diamines to cyclic carbonates, cyclic carbamates and cyclic ureas using heterogeneous catalysts. J. Chem. Technol. Biotechnol. 2014, 89, 19–33. [Google Scholar] [CrossRef]
  80. Su, X.; Lin, W.; Cheng, H.; Zhang, C.; Wang, Y.; Yu, X.; Wu, Z.; Zhao, F. Metal-free catalytic conversion of CO2and glycerol to glycerol carbonate. Green Chem. 2017, 19, 1775–1781. [Google Scholar] [CrossRef]
  81. Honda, M.; Tamura, M.; Nakao, K.; Suzuki, K.; Nakagawa, Y.; Tomishige, K. Direct Cyclic Carbonate Synthesis from CO2 and Diol over Carboxylation/Hydration Cascade Catalyst of CeO2 with 2-Cyanopyridine. ACS Catal. 2014, 4, 1893–1896. [Google Scholar] [CrossRef]
  82. Huang, S.; Ma, J.; Li, J.; Zhao, N.; Wei, W.; Sun, Y. Efficient propylene carbonate synthesis from propylene glycol and carbon dioxide via organic bases. Catal. Commun. 2008, 9, 276–280. [Google Scholar] [CrossRef]
  83. Kitamura, T.; Inoue, Y.; Maeda, T.; Oyamada, J. Convenient synthesis of ethylene carbonates from carbon dioxide and 1,2-diols at atmospheric pressure of carbon dioxide. Synth. Commun. 2015, 46, 39–45. [Google Scholar] [CrossRef]
  84. Gregory, G.L.; Ulmann, M.; Buchard, A. Synthesis of 6-membered cyclic carbonates from 1,3-diols and low CO2 pressure: A novel mild strategy to replace phosgene reagents. RSC Adv. 2015, 5, 39404–39408. [Google Scholar] [CrossRef] [Green Version]
  85. Reithofer, M.R.; Sum, Y.N.; Zhang, Y. Synthesis of cyclic carbonates with carbon dioxide and cesium carbonate. Green Chem. 2013, 15, 2086–2090. [Google Scholar] [CrossRef]
  86. Mcguire, T.M.; López-Vidal, E.M.; Gregory, G.L.; Buchard, A. Synthesis of 5- to 8-membered cyclic carbonates from diols and CO2: A one-step, atmospheric pressure and ambient temperature procedure. J. CO2 Util. 2018, 27, 283–288. [Google Scholar] [CrossRef]
  87. Brege, A.; Méreau, R.; Mcgehee, K.; Grignard, B.; Detrembleur, C.; Jerome, C.; Tassaing, T. The coupling of CO2 with diols promoted by organic dual systems: Towards products divergence via benchmarking of the performance metrics. J. CO2 Util. 2020, 38, 88–98. [Google Scholar] [CrossRef]
  88. Voutchkova, A.M.; Feliz, M.; Clot, E.; Eisenstein, O.; Crabtree, R.H. Imidazolium Carboxylates as Versatile and Selective N-Heterocyclic Carbene Transfer Agents: Synthesis, Mechanism, and Applications. J. Am. Chem. Soc. 2007, 129, 12834–12846. [Google Scholar] [CrossRef]
  89. Riduan, S.N.; Zhang, Y.; Ying, J.Y. Conversion of Carbon Dioxide into Methanol with Silanes over N-Heterocyclic Carbene Catalysts. Angew. Chem. Int. Ed. 2009, 48, 3322–3325. [Google Scholar] [CrossRef] [PubMed]
  90. Bobbink, F.D.; Gruszka, W.; Hulla, M.; Das, S.; Dyson, P.J. Synthesis of cyclic carbonates from diols and CO2 catalyzed by carbenes. Chem. Commun. 2016, 52, 10787–10790. [Google Scholar] [CrossRef] [Green Version]
  91. Cardillo, G.; Orena, M.; Porzi, G.; Sandri, S. A new regio- and stereo-selective functionalization of allylic and homoallylic alcohols. J. Chem. Soc. Chem. Commun. 1981, 10, 465–466. [Google Scholar] [CrossRef]
  92. Vara, B.A.; Struble, T.J.; Wang, W.; Dobish, M.C.; Johnston, J.N. Enantioselective Small Molecule Synthesis by Carbon Dioxide Fixation using a Dual Brønsted Acid/Base Organocatalyst. J. Am. Chem. Soc. 2015, 137, 7302–7305. [Google Scholar] [CrossRef] [Green Version]
  93. Minakata, S.; Sasaki, I.; Ide, T. Atmospheric CO2 Fixation by Unsaturated Alcohols Using tBuOI under Neutral Conditions. Angew. Chem. Int. Ed. 2010, 49, 1309–1311. [Google Scholar] [CrossRef]
  94. Wang, J.-L.; He, L.-N.; Dou, X.-Y.; Wu, F. Poly(ethylene glycol): An Alternative Solvent for the Synthesis of Cyclic Carbonate from Vicinal Halohydrin and Carbon Dioxide. Aust. J. Chem. 2009, 62, 917–920. [Google Scholar] [CrossRef]
  95. Eghbali, N.; Li, C.-J. Conversion of carbon dioxide and olefins into cyclic carbonates in water. Green Chem. 2007, 9, 213–215. [Google Scholar] [CrossRef]
  96. Smith, M.B.; March, J. (Eds.) March’s Advanced Organic Chemistry: Reactions, Mechanisms and Structure, 5th; Wiley Interscience: New York, NY, USA, 2001; ISBN 0-471-58589-0. [Google Scholar]
  97. Lang, X.D.; He, L.N. Green Catalytic Process for Cyclic Carbonate Synthesis from Carbon Dioxide under Mild Conditions. Chem. Rec. 2016, 16, 1337–1352. [Google Scholar] [CrossRef]
  98. Kayaki, Y.; Yamamoto, M.; Ikariya, T. Stereoselective Formation of α-Alkylidene Cyclic Carbonates via Carboxylative Cyclization of Propargyl Alcohols in Supercritical Carbon Dioxide. J. Org. Chem. 2007, 72, 647–649. [Google Scholar] [CrossRef]
  99. Sun, Y.-L.; Wei, Y.; Shi, M. Phosphine-catalyzed fixation of CO2 with γ-hydroxyl alkynone under ambient temperature and pressure: Kinetic resolution and further conversion. Org. Chem. Front. 2019, 6, 2420–2429. [Google Scholar] [CrossRef]
  100. Matthews, C.N.; Driscoll, J.S.; Birum, G.H. Mesomeric phosphonium inner salts. Chem. Commun. 1966, 20, 736–737. [Google Scholar] [CrossRef]
  101. Kayaki, Y.; Yamamoto, M.; Ikariya, T. N-Heterocyclic Carbenes as Efficient Organocatalysts for CO2 Fixation Reactions. Angew. Chem. Int. Ed. 2009, 48, 4194–4197. [Google Scholar] [CrossRef] [PubMed]
  102. Li, W.; Yang, N.; Lyu, Y. Theoretical Insights into the Catalytic Mechanism of N-Heterocyclic Olefins in Carboxylative Cyclization of Propargyl Alcohol with CO2. J. Org. Chem. 2016, 81, 5303–5313. [Google Scholar] [CrossRef]
  103. Yan, Z.-E.; Huo, R.-P.; Guo, L.-H.; Zhang, X. The mechanisms for N-heterocyclic olefin-catalyzed formation of cyclic carbonate from CO2 and propargylic alcohols. J. Mol. Model. 2016, 22, 94. [Google Scholar] [CrossRef]
  104. Li, W.; Huang, D.; Lyu, Y. A comparative computational study of N-heterocyclic olefin and N-heterocyclic carbene mediated carboxylative cyclization of propargyl alcohols with CO2. Org. Biomol. Chem. 2016, 14, 10875–10885. [Google Scholar] [CrossRef] [PubMed]
  105. Boyaval, A.; Méreau, R.; Grignard, B.; Detrembleur, C.; Jerome, C.; Tassaing, T. Organocatalytic Coupling of CO2 with a Propargylic Alcohol: A Comprehensive Mechanistic Study. ChemSusChem 2017, 10, 1241–1248. [Google Scholar] [CrossRef]
  106. Méreau, R.; Grignard, B.; Boyaval, A.; Detrembleur, C.; Jerome, C.; Tassaing, T. Tetrabutylammonium Salts: Cheap Catalysts for the Facile and Selective Synthesis of α-Alkylidene Cyclic Carbonates from Carbon Dioxide and Alkynols. ChemCatChem 2018, 10, 956–960. [Google Scholar] [CrossRef]
  107. Grignard, B.; Ngassamtounzoua, C.; Gennen, S.; Gilbert, B.; Méreau, R.; Jerome, C.; Tassaing, T.; Detrembleur, C. Boosting the Catalytic Performance of Organic Salts for the Fast and Selective Synthesis of α-Alkylidene Cyclic Carbonates from Carbon Dioxide and Propargylic Alcohols. ChemCatChem 2018, 10, 2584–2592. [Google Scholar] [CrossRef]
  108. Gennen, S.; Grignard, B.; Tassaing, T.; Jerome, C.; Detrembleur, C. CO2-Sourced alpha-Alkylidene Cyclic Carbonates: A Step Forward in the Quest for Functional Regioregular Poly(urethane)s and Poly(carbonate)s. Angew. Chem. Int. Ed. Engl. 2017, 56, 10394–10398. [Google Scholar] [CrossRef] [PubMed]
  109. Qiu, J.; Zhao, Y.; Li, Z.; Wang, H.; Fan, M.; Wang, J. Efficient Ionic-Liquid-Promoted Chemical Fixation of CO2 into α-Alkylidene Cyclic Carbonates. ChemSusChem 2017, 10, 1120–1127. [Google Scholar] [CrossRef]
  110. Chen, K.; Shi, G.; Dao, R.; Mei, K.; Zhou, X.; Li, H.; Wang, C. Tuning the basicity of ionic liquids for efficient synthesis of alkylidene carbonates from CO2 at atmospheric pressure. Chem. Commun. 2016, 52, 7830–7833. [Google Scholar] [CrossRef] [Green Version]
  111. Hu, Y.; Song, J.; Xie, C.; Wu, H.; Jiang, T.; Yang, G.; Han, B. Transformation of CO2 into α-Alkylidene Cyclic Carbonates at Room Temperature Cocatalyzed by CuI and Ionic Liquid with Biomass-Derived Levulinate Anion. ACS Sustain. Chem. Eng. 2019, 7, 5614–5619. [Google Scholar] [CrossRef]
  112. Ouyang, L.; Tang, X.; He, H.; Qi, C.; Xiong, W.; Ren, Y.; Jiang, H. Copper-Promoted Coupling of Carbon Dioxide and Propargylic Alcohols: Expansion of Substrate Scope and Trapping of Vinyl Copper Intermediate. Adv. Synth. Catal. 2015, 357, 2556–2565. [Google Scholar] [CrossRef]
  113. Satoshi, K.; Shunsuke, Y.; Yuudai, S.; Wataru, Y.; Hau-Man, C.; Kosuke, F.; Kohei, S.; Izumi, I.; Taketo, I.; Tohru, Y. Silver-Catalyzed Carbon Dioxide Incorporation and Rearrangement on Propargylic Derivatives. Bull. Chem. Soc. Jpn. 2011, 84, 698–717. [Google Scholar]
  114. Yamada, T.; Ugajin, R.; Kikuchi, S. Silver-Catalyzed Efficient Synthesis of Vinylene Carbonate Derivatives from Carbon Dioxide. Synlett 2014, 25, 1178–1180. [Google Scholar] [CrossRef]
  115. Yuan, Y.; Xie, Y.; Zeng, C.; Song, D.; Chaemchuen, S.; Chen, C.; Verpoort, F. A recyclable AgI/OAc catalytic system for the efficient synthesis of α-alkylidene cyclic carbonates: Carbon dioxide conversion at atmospheric pressure. Green Chem. 2017, 19, 2936–2940. [Google Scholar] [CrossRef]
  116. Li, J.-Y.; Han, L.-H.; Xu, Q.-C.; Song, Q.-W.; Liu, P.; Zhang, K. Cascade Strategy for Atmospheric Pressure CO2 Fixation to Cyclic Carbonates via Silver Sulfadiazine and Et4NBr Synergistic Catalysis. ACS Sustain. Chem. Eng. 2019, 7, 3378–3388. [Google Scholar] [CrossRef]
  117. Yuan, Y.; Xie, Y.; Song, D.; Zeng, C.; Chaemchuen, S.; Chen, C.; Verpoort, F. One-pot carboxylative cyclization of propargylic alcohols and CO2 catalysed by N-heterocyclic carbene/Ag systems. Appl. Organomet. Chem. 2017, 31, e3867. [Google Scholar] [CrossRef]
  118. Cui, M.; Qian, Q.; He, Z.; Ma, J.; Kang, X.; Hu, J.; Liu, Z.; Han, B. Synthesizing Ag Nanoparticles of Small Size on a Hierarchical Porosity Support for the Carboxylative Cyclization of Propargyl Alcohols with CO2 under Ambient Conditions. Chem.–A Eur. J. 2015, 21, 15924–15928. [Google Scholar] [CrossRef]
  119. Yang, Z.; Yu, B.; Zhang, H.; Zhao, Y.; Chen, Y.; Ma, Z.; Ji, G.; Gao, X.; Han, B.; Liu, Z. Metalated Mesoporous Poly(triphenylphosphine) with Azo Functionality: Efficient Catalysts for CO2 Conversion. ACS Catal. 2016, 6, 1268–1273. [Google Scholar] [CrossRef]
  120. Zhou, Z.; He, C.; Yang, L.; Wang, Y.; Liu, T.; Duan, C. Alkyne Activation by a Porous Silver Coordination Polymer for Heterogeneous Catalysis of Carbon Dioxide Cycloaddition. ACS Catal. 2017, 7, 2248–2256. [Google Scholar] [CrossRef]
  121. Dabral, S.; Bayarmagnai, B.; Hermsen, M.; Schießl, J.; Mormul, V.; Hashmi, A.S.K.; Schaub, T. Silver-Catalyzed Carboxylative Cyclization of Primary Propargyl Alcohols with CO2. Org. Lett. 2019, 21, 1422–1425. [Google Scholar] [CrossRef]
  122. Jung, M.E.; Piizzi, G. gem-Disubstituent Effect: Theoretical Basis and Synthetic Applications. Chem. Rev. 2005, 105, 1735–1766. [Google Scholar] [CrossRef]
  123. Islam, S.S.; Salam, N.; Molla, R.A.; Riyajuddin, S.; Yasmin, N.; Das, D.; Ghosh, K.; Islam, S.M. Zinc (II) incorporated porous organic polymeric material (POPs): A mild and efficient catalyst for synthesis of dicoumarols and carboxylative cyclization of propargyl alcohols and CO2 in ambient conditions. Mol. Catal. 2019, 477, 110541. [Google Scholar] [CrossRef]
  124. Hu, J.; Ma, J.; Zhu, Q.; Qian, Q.; Han, H.; Mei, Q.; Han, B. Zinc(ii)-catalyzed reactions of carbon dioxide and propargylic alcohols to carbonates at room temperature. Green Chem. 2016, 18, 382–385. [Google Scholar] [CrossRef]
  125. Ma, J.; Lu, L.; Mei, Q.; Zhu, Q.; Hu, J.; Han, B. ZnI2/NEt3-Catalyzed Cycloaddition of CO2 with Propargylic Alcohols: Computational Study on Mechanism. ChemCatChem 2017, 9, 4090–4097. [Google Scholar] [CrossRef]
  126. Tamura, M.; Ito, K.; Honda, M.; Nakagawa, Y.; Sugimoto, H.; Tomishige, K. Direct Copolymerization of CO2 and Diols. Sci. Rep. 2016, 6, 24038. [Google Scholar] [CrossRef] [Green Version]
  127. Gu, Y.; Matsuda, K.; Nakayama, A.; Tamura, M.; Nakagawa, Y.; Tomishige, K. Direct Synthesis of Alternating Polycarbonates from CO2 and Diols by Using a Catalyst System of CeO2 and 2-Furonitrile. ACS Sustain. Chem. Eng. 2019, 7, 6304–6315. [Google Scholar] [CrossRef]
  128. Gong, Z.-J.; Li, Y.-R.; Wu, H.-L.; Lin, S.D.; Yu, W.-Y. Direct copolymerization of carbon dioxide and 1,4-butanediol enhanced by ceria nanorod catalyst. Appl. Catal. B Environ. 2020, 265, 118524. [Google Scholar] [CrossRef]
  129. Kadokawa, J.-I.; Habu, H.; Fukamachi, S.; Karasu, M.; Tagaya, H.; Chiba, K. Direct polycondensation of carbon dioxide with xylylene glycols: A new method for the synthesis of polycarbonates. Macromol. Rapid Commun. 1998, 19, 657–660. [Google Scholar] [CrossRef]
  130. Kadokawa, J.-I.; Fukamachi, S.; Tagaya, H.; Chiba, K. Direct Polycondensation of Carbon Dioxide with Various Diols Using the Triphenylphosphine/Bromotrichloromethane/N-Cyclohexyl-N′,N′,N″,N″-tetramethylguanidine System as Condensing Agent. Polym. J. 2000, 32, 703. [Google Scholar] [CrossRef] [Green Version]
  131. Bian, S.; Pagan, C.; Andrianova Artemyeva, A.A.; Du, G. Synthesis of Polycarbonates and Poly(ether carbonate)s Directly from Carbon Dioxide and Diols Promoted by a Cs2CO3/CH2Cl2 System. ACS Omega 2016, 1, 1049–1057. [Google Scholar] [CrossRef] [PubMed]
  132. Pati, D.; Chen, Z.; Feng, X.; Hadjichristidis, N.; Gnanou, Y. Synthesis of polyglycocarbonates through polycondensation of glucopyranosides with CO2. Polym. Chem. 2017, 8, 2640–2646. [Google Scholar] [CrossRef]
  133. Soga, K.; Toshida, Y.; Hosoda, S.; Ikeda, S. A convenient synthesis of a polycarbonate. Makromol. Chem. 1977, 178, 2747–2751. [Google Scholar] [CrossRef]
  134. Soga, K.; Toshida, Y.; Hosoda, S.; Ikeda, S. Synthesis of a polycarbonate directly from carbon dioxide, alkali metal diolates and α,ω-dihalo compounds. Makromol. Chem. 1978, 179, 2379–2386. [Google Scholar] [CrossRef]
  135. Chen, Z.; Hadjichristidis, N.; Feng, X.; Gnanou, Y. Cs2CO3-promoted polycondensation of CO2 with diols and dihalides for the synthesis of miscellaneous polycarbonates. Polym. Chem. 2016, 7, 4944–4952. [Google Scholar] [CrossRef]
  136. Bian, S.; Andrianova, A.A.; Kubatova, A.; Du, G. Effect of dihalides on the polymer linkages in the Cs2CO3-promoted polycondensation of 1 atm carbon dioxide and diols. Mater. Today Commun. 2019, 18, 100–109. [Google Scholar] [CrossRef]
  137. Höcker, H.; Keul, H.; Kühling, S.; Hovestadt, W.; Müller, A.; Wurm, B. Ring-opening polymerization and copolymerization of cyclic carbonates with a variety of initiating systems. Makromol. Chem. Macromol. Symp. 1993, 73, 1–5. [Google Scholar] [CrossRef]
  138. Rokicki, G. Aliphatic cyclic carbonates and spiroorthocarbonates as monomers. Prog. Polym. Sci. 2000, 25, 259–342. [Google Scholar] [CrossRef]
  139. Vogdanis, L.; Heitz, W. Carbon dioxide as a monomer, 3. The polymerization of ethylene carbonate. Die Makromol. Chem. Rapid Commun. 1986, 7, 543–547. [Google Scholar] [CrossRef]
  140. Vogdanis, L.; Martens, B.; Uchtmann, H.; Hensel, F.; Heitz, W. Synthetic and thermodynamic investigations in the polymerization of ethylene carbonate. Die Makromol. Chem. 1990, 191, 465–472. [Google Scholar] [CrossRef]
  141. Storey, R.F.; Hoffman, D.C. Formation of poly(ethylene ether carbonate) diols: Proposed mechanism and kinetic analysis. Macromolecules 1992, 25, 5369–5382. [Google Scholar] [CrossRef]
  142. Kricheldorf, H.R.; Jonté, J.M.; Berl, M. Polylactones 3. Copolymerization of glycolide with L, L-lactide and other lactones. Die Makromol. Chem. 1985, 12, 25–38. [Google Scholar] [CrossRef]
  143. Kricheldorf, H.R.; Berl, M.; Scharnagl, N. Poly(lactones). 9. Polymerization mechanism of metal alkoxide initiated polymerizations of lactide and various lactones. Macromolecules 1988, 21, 286–293. [Google Scholar] [CrossRef]
  144. Clements, J.H. Reactive Applications of Cyclic Alkylene Carbonates. Ind. Eng. Chem. Res. 2003, 42, 663–674. [Google Scholar] [CrossRef]
  145. Lee, J.-C.; Litt, M.H. Ring-Opening Polymerization of Ethylene Carbonate and Depolymerization of Poly(ethylene oxide-co-ethylene carbonate). Macromolecules 2000, 33, 1618–1627. [Google Scholar] [CrossRef]
  146. Yang, H.; Yan, M.; Pispas, S.; Zhang, G. Synthesis of Poly[(ethylene carbonate)-co-(ethylene oxide)] Copolymer by Phosphazene-Catalyzed ROP. Macromol. Chem. Phys. 2011, 212, 2589–2593. [Google Scholar] [CrossRef]
  147. Soga, K.; Tazuke, Y.; Hosoda, S.; Ikeda, S. Polymerization of propylene carbonate. J. Polym. Sci. Polym. Chem. Ed. 1977, 15, 219–229. [Google Scholar] [CrossRef]
  148. Tezuka, K.; Komatsu, K.; Haba, O. The anionic ring-opening polymerization of five-membered cyclic carbonates fused to the cyclohexane ring. Polym. J. 2013, 45, 1183–1187. [Google Scholar] [CrossRef] [Green Version]
  149. Haba, O.; Tomizuka, H.; Endo, T. Anionic Ring-Opening Polymerization of Methyl 4,6-O-Benzylidene-2,3-O- carbonyl-α-d-glucopyranoside: A First Example of Anionic Ring-Opening Polymerization of Five-Membered Cyclic Carbonate without Elimination of CO2. Macromolecules 2005, 38, 3562–3563. [Google Scholar] [CrossRef]
  150. Guerin, W.; Diallo, A.K.; Kirilov, E.; Helou, M.; Slawinski, M.; Brusson, J.-M.; Carpentier, J.-F.; Guillaume, S.M. Enantiopure Isotactic PCHC Synthesized by Ring-Opening Polymerization of Cyclohexene Carbonate. Macromolecules 2014, 47, 4230–4235. [Google Scholar] [CrossRef]
  151. Azechi, M.; Matsumoto, K.; Endo, T. Anionic ring-opening polymerization of a five-membered cyclic carbonate having a glucopyranoside structure. J. Polym. Sci. Part A Polym. Chem. 2013, 51, 1651–1655. [Google Scholar] [CrossRef]
  152. Rokicki, G.; Parzuchowski, P.G. Polymer Science: A Comprehensive Reference; Matyjaszewski, K., Möller, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2012; pp. 247–308. [Google Scholar] [CrossRef]
  153. Kricheldorf, H.R.; Jenssen, J. Polylactones. 16. Cationic Polymerization of Trimethylene Carbonate and Other Cyclic Carbonates. J. Macromol. Sci. Part A—Chem. 1989, 26, 631–644. [Google Scholar] [CrossRef]
  154. Ariga, T.; Takata, T.; Endo, T. Cationic Ring-Opening Polymerization of Cyclic Carbonates with Alkyl Halides To Yield Polycarbonate without the Ether Unit by Suppression of Elimination of Carbon Dioxide. Macromolecules 1997, 30, 737–744. [Google Scholar] [CrossRef]
  155. Delcroix, D.; Martín-Vaca, B.; Bourissou, D.; Navarro, C. Ring-Opening Polymerization of Trimethylene Carbonate Catalyzed by Methanesulfonic Acid: Activated Monomer versus Active Chain End Mechanisms. Macromolecules 2010, 43, 8828–8835. [Google Scholar] [CrossRef]
  156. Liu, J.; Cui, S.; Li, Z.; Xu, S.; Xu, J.; Pan, X.; Liu, Y.; Dong, H.; Sun, H.; Guo, K. Polymerization of trimethylene carbonates using organic phosphoric acids. Polym. Chem. 2016, 7, 5526–5535. [Google Scholar] [CrossRef]
  157. Keul, H.; Bächer, R.; Höcker, H. Anionic ring-opening polymerization of 2,2-dimethyltrimethylene carbonate. Die Makromol. Chem. 1986, 187, 2579–2589. [Google Scholar] [CrossRef]
  158. Sanda, F.; Kamatani, J.; Endo, T. Synthesis and Anionic Ring-Opening Polymerization Behavior of Amino Acid-Derived Cyclic Carbonates. Macromolecules 2001, 34, 1564–1569. [Google Scholar] [CrossRef]
  159. Kühling, S.; Keul, H.; Höcker, H.; Buysch, H.-J.; Schön, N. Synthesis of poly(2-ethyl-2-hydroxymethyltrimethylene carbonate). Die Makromol. Chem. 1991, 192, 1193–1205. [Google Scholar] [CrossRef]
  160. Murayama, M.; Sanda, F.; Endo, T. Anionic Ring-Opening Polymerization of a Cyclic Carbonate Having a Norbornene Structure with Amine Initiators. Macromolecules 1998, 31, 919–923. [Google Scholar] [CrossRef]
  161. Matsuo, J.; Sanda, F.; Endo, T. A novel observation in anionic ring-opening polymerization behavior of cyclic carbonates having aromatic substituents. Macromol. Chem. Phys. 1998, 199, 2489–2494. [Google Scholar] [CrossRef]
  162. Matsuo, J.; Aoki, K.; Sanda, F.; Endo, T. Substituent Effect on the Anionic Equilibrium Polymerization of Six-Membered Cyclic Carbonates. Macromolecules 1998, 31, 4432–4438. [Google Scholar] [CrossRef]
  163. Shen, Y.; Chen, X.; Gross, R.A. Polycarbonates from Sugars: Ring-Opening Polymerization of 1,2-O-Isopropylidene-d-Xylofuranose-3,5-Cyclic Carbonate (IPXTC). Macromolecules 1999, 32, 2799–2802. [Google Scholar] [CrossRef]
  164. Darensbourg, D.J.; Ganguly, P.; Billodeaux, D. Ring-Opening Polymerization of Trimethylene Carbonate Using Aluminum(III) and Tin(IV) Salen Chloride Catalysts. Macromolecules 2005, 38, 5406–5410. [Google Scholar] [CrossRef]
  165. Darensbourg, D.J.; Choi, W.; Ganguly, P.; Richers, C.P. Biometal Derivatives as Catalysts for the Ring-Opening Polymerization of Trimethylene Carbonate. Optimization of the Ca(II) Salen Catalyst System. Macromolecules 2006, 39, 4374–4379. [Google Scholar] [CrossRef]
  166. Darensbourg, D.J.; Choi, W.; Richers, C.P. Ring-Opening Polymerization of Cyclic Monomers by Biocompatible Metal Complexes. Production of Poly(lactide), Polycarbonates, and Their Copolymers. Macromolecules 2007, 40, 3521–3523. [Google Scholar] [CrossRef]
  167. Darensbourg, D.J.; Choi, W.; Karroonnirun, O.; Bhuvanesh, N. Ring-Opening Polymerization of Cyclic Monomers by Complexes Derived from Biocompatible Metals. Production of Poly(lactide), Poly(trimethylene carbonate), and Their Copolymers. Macromolecules 2008, 41, 3493–3502. [Google Scholar] [CrossRef]
  168. Kricheldorf, H.R.; Mahler, A. Polymers of carbonic acid 18: Polymerizations of cyclobis(hexamethylene carbonate) by means of BuSnCl3 or Sn(II)2-ethylhexanoate. Polymer 1996, 37, 4383–4388. [Google Scholar] [CrossRef]
  169. Kricheldorf, H.R.; Stricker, A. Polymers of carbonic acid, 28. SnOct2-initiated polymerizations of trimethylene carbonate (TMC, 1,3-dioxanone-2). Macromol. Chem. Phys. 2000, 201, 2557–2565. [Google Scholar] [CrossRef]
  170. Kricheldorf, H.R.; Weegen-Schulz, B. Polymers of carbonic acid. XIV. High molecular weight poly(trimethylene carbonate) by ring-opening polymerization with butyltin chlorides as initiators. J. Polym. Sci. Part A Polym. Chem. 1995, 33, 2193–2201. [Google Scholar] [CrossRef]
  171. Pêgo, A.P.; Grijpma, D.W.; Feijen, J. Enhanced mechanical properties of 1,3-trimethylene carbonate polymers and networks. Polymer 2003, 44, 6495–6504. [Google Scholar] [CrossRef]
  172. Ling, J.; Shen, Z.; Huang, Q. Novel Single Rare Earth Aryloxide Initiators for Ring-Opening Polymerization of 2,2-Dimethyltrimethylene Carbonate. Macromolecules 2001, 34, 7613–7616. [Google Scholar] [CrossRef]
  173. Nederberg, F.; Lohmeijer, B.G.G.; Leibfarth, F.; Pratt, R.C.; Choi, J.; Dove, A.P.; Waymouth, R.M.; Hedrick, J.L. Organocatalytic Ring Opening Polymerization of Trimethylene Carbonate. Biomacromolecules 2007, 8, 153–160. [Google Scholar] [CrossRef] [PubMed]
  174. Naumann, S.; Thomas, A.W.; Dove, A.P. Highly Polarized Alkenes as Organocatalysts for the Polymerization of Lactones and Trimethylene Carbonate. ACS Macro Lett. 2016, 5, 134–138. [Google Scholar] [CrossRef]
  175. Kamber, N.E.; Jeong, W.; Waymouth, R.M.; Pratt, R.C.; Lohmeijer, B.G.G.; Hedrick, J.L. Organocatalytic Ring-Opening Polymerization. Chem. Rev. 2007, 107, 5813–5840. [Google Scholar] [CrossRef] [PubMed]
  176. Gregory, G.L.; Jenisch, L.M.; Charles, B.; Kociok-Köhn, G.; Buchard, A. Polymers from Sugars and CO2: Synthesis and Polymerization of a d-Mannose-Based Cyclic Carbonate. Macromolecules 2016, 49, 7165–7169. [Google Scholar] [CrossRef] [Green Version]
  177. Gregory, G.L.; Kociok-Köhn, G.; Buchard, A. Polymers from sugars and CO2: Ring-opening polymerisation and copolymerisation of cyclic carbonates derived from 2-deoxy-d-ribose. Polym. Chem. 2017, 8, 2093–2104. [Google Scholar] [CrossRef] [Green Version]
  178. Gregory, G.L.; Hierons, E.M.; Kociok-Köhn, G.; Sharma, R.I.; Buchard, A. CO2-Driven stereochemical inversion of sugars to create thymidine-based polycarbonates by ring-opening polymerisation. Polym. Chem. 2017, 8, 1714–1721. [Google Scholar] [CrossRef] [Green Version]
  179. Gregory, G.L.; Lopez-Vidal, E.M.; Buchard, A. Polymers from sugars: Cyclic monomer synthesis, ring-opening polymerisation, material properties and applications. Chem. Commun. 2017, 53, 2198–2217. [Google Scholar] [CrossRef] [Green Version]
  180. Pati, D.; Feng, X.; Hadjichristidis, N.; Gnanou, Y. CO2 as versatile carbonation agent of glycosides: Synthesis of 5- and 6-membered cyclic glycocarbonates and investigation of their ring-opening. J. CO2 Util. 2018, 24, 564–571. [Google Scholar] [CrossRef]
  181. Harris, R.F. Molecular weight advancement of poly(ethylene ether carbonate) polyols. J. Appl. Polym. Sci. 1989, 38, 463–476. [Google Scholar] [CrossRef]
  182. Pawłowski, P.; Rokicki, G. Synthesis of oligocarbonate diols from ethylene carbonate and aliphatic diols catalyzed by alkali metal salts. Polymer 2004, 45, 3125–3137. [Google Scholar] [CrossRef]
  183. Rokicki, G.; Kowalczyk, T. Synthesis of oligocarbonate diols and their characterization by MALDI-TOF spectrometry. Polymer 2000, 41, 9013–9031. [Google Scholar] [CrossRef]
  184. Xu, J.; Feng, E.; Song, J. Renaissance of aliphatic polycarbonates: New techniques and biomedical applications. J. Appl. Polym. Sci. 2014, 131, 39822. [Google Scholar] [CrossRef] [Green Version]
  185. He, Q.; Zhang, Q.; Liao, S.; Zhao, C.; Xie, X. Understanding cyclic by-products and ether linkage formation pathways in the transesterification synthesis of aliphatic polycarbonates. Eur. Polym. J. 2017, 97, 253–262. [Google Scholar] [CrossRef]
  186. Sun, J.; Kuckling, D. Synthesis of high-molecular-weight aliphatic polycarbonates by organo-catalysis. Polym. Chem. 2016, 7, 1642–1649. [Google Scholar] [CrossRef] [Green Version]
  187. Meabe, L.; Lago, N.; Rubatat, L.; Li, C.; Müller, A.J.; Sardon, H.; Armand, M.; Mecerreyes, D. Polycondensation as a Versatile Synthetic Route to Aliphatic Polycarbonates for Solid Polymer Electrolytes. Electrochim. Acta 2017, 237, 259–266. [Google Scholar] [CrossRef]
  188. Naik, P.U.; Refes, K.; Sadaka, F.; Brachais, C.-H.; Boni, G.; Couvercelle, J.-P.; Picquet, M.; Plasseraud, L. Organo-catalyzed synthesis of aliphatic polycarbonates in solvent-free conditions. Polym. Chem. 2012, 3, 1475–1480. [Google Scholar] [CrossRef]
  189. Mutlu, H.; Ruiz, J.; Solleder, S.C.; Meier, M.a.R. TBD catalysis with dimethyl carbonate: A fruitful and sustainable alliance. Green Chem. 2012, 14, 1728–1735. [Google Scholar] [CrossRef]
  190. Park, J.H.; Jeon, J.Y.; Lee, J.J.; Jang, Y.; Varghese, J.K.; Lee, B.Y. Preparation of High-Molecular-Weight Aliphatic Polycarbonates by Condensation Polymerization of Diols and Dimethyl Carbonate. Macromolecules 2013, 46, 3301–3308. [Google Scholar] [CrossRef]
  191. Zhang, J.; Zhu, W.; Li, C.; Zhang, D.; Xiao, Y.; Guan, G.; Zheng, L. Effect of the biobased linear long-chain monomer on crystallization and biodegradation behaviors of poly(butylene carbonate)-based copolycarbonates. RSC Adv. 2015, 5, 2213–2222. [Google Scholar] [CrossRef]
  192. Zhu, W.; Huang, X.; Li, C.; Xiao, Y.; Zhang, D.; Guan, G. High-molecular-weight aliphatic polycarbonates by melt polycondensation of dimethyl carbonate and aliphatic diols: Synthesis and characterization. Polym. Int. 2011, 60, 1060–1067. [Google Scholar] [CrossRef]
  193. Zhu, W.; Zhou, W.; Li, C.; Xiao, Y.; Zhang, D.; Guan, G.; Wang, D. Synthesis, Characterization and Degradation of Novel Biodegradable Poly(butylene-co-hexamethylene carbonate) Copolycarbonates. J. Macromol. Sci. Part A 2011, 48, 583–594. [Google Scholar] [CrossRef]
  194. Sun, J.; Aly, K.I.; Kuckling, D. A novel one-pot process for the preparation of linear and hyperbranched polycarbonates of various diols and triols using dimethyl carbonate. RSC Adv. 2017, 7, 12550–12560. [Google Scholar] [CrossRef] [Green Version]
  195. Wang, Z.; Yang, X.; Liu, S.; Hu, J.; Zhang, H.; Wang, G. One-pot synthesis of high-molecular-weight aliphatic polycarbonates via melt transesterification of diphenyl carbonate and diols using Zn(OAc)2 as a catalyst. RSC Adv. 2015, 5, 87311–87319. [Google Scholar] [CrossRef]
  196. Wang, Z.; Yang, X.; Li, J.; Liu, S.; Wang, G. Synthesis of high-molecular-weight aliphatic polycarbonates from diphenyl carbonate and aliphatic diols by solid base. J. Mol. Catal. A Chem. 2016, 424, 77–84. [Google Scholar] [CrossRef]
  197. Vanderhenst, R.; Miller, S.A. Polycarbonates from biorenewable diols via carbonate metathesis polymerization. Green Mater. 2013, 1, 64–78. [Google Scholar] [CrossRef]
  198. Haban, O.; Itakura, I.; Ueda, M.; Kuze, S. Synthesis of Polycarbonate from Dimethyl Carbonate and bisphenol-A. J. Polym. Sci. 1998, 37, 2087–2093. [Google Scholar]
  199. Ignatov, V.N.; Tartari, V.; Carraro, C.; Pippa, R.; Nadali, G.; Berti, C.; Fiorini, M. New Catalysts for Bisphenol A Polycarbonate Melt Polymerisation, 2. Polymer Synthesis and Characterisation. Macromol. Chem. Phys. 2001, 202, 1946–1949. [Google Scholar] [CrossRef]
  200. Kim, Y.; Choi, K.Y.; Chamberlin, T.A. Kinetics of melt transesterification of diphenyl carbonate and bisphenol A to polycarbonate with lithium hydroxide monohydrate catalyst. Ind. Eng. Chem. Res. 1992, 31, 2118–2127. [Google Scholar] [CrossRef]
  201. Fukuoka, S.; Kawamura, M.; Komiya, K.; Tojo, M.; Hachiya, H.; Hasegawa, K.; Aminaka, M.; Okamoto, H.; Fukawa, I.; Konno, S. A novel non-phosgene polycarbonate production process using by-product CO2 as starting material. Green Chem. 2003, 5, 497–507. [Google Scholar] [CrossRef]
  202. Zhang, M.; Lai, W.; Su, L.; Lin, Y.; Wu, G. A synthetic strategy toward isosorbide polycarbonate with a high molecular weight: The effect of intermolecular hydrogen bonding between isosorbide and metal chlorides. Polym. Chem. 2019, 10, 3380–3389. [Google Scholar] [CrossRef]
  203. Yang, Z.; Liu, L.; An, H.; Li, C.; Zhang, Z.; Fang, W.; Xu, F.; Zhang, S. Cost-Effective Synthesis of High Molecular Weight Biobased Polycarbonate via Melt Polymerization of Isosorbide and Dimethyl Carbonate. ACS Sustain. Chem. Eng. 2020, 8, 9968–9979. [Google Scholar] [CrossRef]
  204. Li, Q.; Zhu, W.; Li, C.; Guan, G.; Zhang, D.; Xiao, Y.; Zheng, L. A non-phosgene process to homopolycarbonate and copolycarbonates of isosorbide using dimethyl carbonate: Synthesis, characterization, and properties. J. Polym. Sci. Part A Polym. Chem. 2013, 51, 1387–1397. [Google Scholar] [CrossRef]
  205. Eo, Y.S.; Rhee, H.-W.; Shin, S. Catalyst screening for the melt polymerization of isosorbide-based polycarbonate. J. Ind. Eng. Chem. 2016, 37, 42–46. [Google Scholar] [CrossRef]
  206. Yang, Z.; Li, X.; Xu, F.; Wang, W.; Shi, Y.; Zhang, Z.; Fang, W.; Liu, L.; Zhang, S. Synthesis of bio-based polycarbonate via one-step melt polycondensation of isosorbide and dimethyl carbonate by dual site-functionalized ionic liquid catalysts. Green Chem. 2021, 23, 447–456. [Google Scholar] [CrossRef]
  207. Qian, W.; Ma, X.; Liu, L.; Deng, L.; Su, Q.; Bai, R.; Zhang, Z.; Gou, H.; Dong, L.; Cheng, W.; et al. Efficient synthesis of bio-derived polycarbonates from dimethyl carbonate and isosorbide: Regulating exo-OH and endo-OH reactivity by ionic liquids. Green Chem. 2020, 22, 5357–5368. [Google Scholar] [CrossRef]
  208. Li, C.; Zhang, Z.; Yang, Z.; Fang, W.; An, H.; Li, T.; Xu, F. Synthesis of bio-based poly(oligoethylene glycols-co-isosorbide carbonate)s with high molecular weight and enhanced mechanical properties via ionic liquid catalyst. React. Funct. Polym. 2020, 155, 104689. [Google Scholar] [CrossRef]
  209. Qian, W.; Tan, X.; Su, Q.; Cheng, W.; Xu, F.; Dong, L.; Zhang, S. Transesterification of Isosorbide with Dimethyl Carbonate Catalyzed by Task-Specific Ionic Liquids. ChemSusChem 2019, 12, 1169–1178. [Google Scholar] [CrossRef]
  210. Qian, W.; Liu, L.; Zhang, Z.; Su, Q.; Zhao, W.; Cheng, W.; Li, D.; Yang, Z.; Bai, R.; Xu, F.; et al. Synthesis of Bioderived Polycarbonates with Molecular Weight Adjustability Catalyzed by Phenolic-derived Ionic Liquids. Green Chem. 2020, 22, 2488–2497. [Google Scholar] [CrossRef]
  211. Sun, W.; Xu, F.; Cheng, W.; Sun, J.; Ning, G.; Zhang, S. Synthesis of isosorbide-based polycarbonates via melt polycondensation catalyzed by quaternary ammonium ionic liquids. Chin. J. Catal. 2017, 38, 908–917. [Google Scholar] [CrossRef]
  212. Ma, C.; Xu, F.; Cheng, W.; Tan, X.; Su, Q.; Zhang, S. Tailoring Molecular Weight of Bioderived Polycarbonates via Bifunctional Ionic Liquids Catalysts under Metal-Free Conditions. ACS Sustain. Chem. Eng. 2018, 6, 2684–2693. [Google Scholar] [CrossRef]
  213. Zhang, Z.; Xu, F.; He, H.; Ding, W.; Fang, W.; Sun, W.; Li, Z.; Zhang, S. Synthesis of high-molecular weight isosorbide-based polycarbonates through efficient activation of endo-hydroxyl groups by an ionic liquid. Green Chem. 2019, 21, 3891–3901. [Google Scholar] [CrossRef]
  214. Fang, W.; Zhang, Z.; Yang, Z.; Zhang, Y.; Xu, F.; Li, C.; An, H.; Song, T.; Luo, Y.; Zhang, S. One-pot synthesis of bio-based polycarbonates from dimethyl carbonate and isosorbide under metal-free condition. Green Chem. 2020, 22, 4550–4560. [Google Scholar] [CrossRef]
  215. Shen, X.; Liu, S.; Wang, Q.; Zhang, H.; Wang, G. Synthesis of Poly(isosorbide carbonate) via Melt Polycondensation Catalyzed by a KF/MgO Catalyst. Chem. Res. Chin. Univ. 2019, 35, 721–728. [Google Scholar] [CrossRef]
  216. Fenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.; Pascault, J.P. Polymers from renewable 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide): A review. Prog. Polym. Sci. 2010, 35, 578–622. [Google Scholar] [CrossRef]
  217. Shen, X.-L.; Wang, Z.-Q.; Wang, Q.-Y.; Liu, S.-Y.; Wang, G.-Y. Synthesis of Poly(isosorbide carbonate) via Melt Polycondensation Catalyzed by Ca/SBA-15 Solid Base. Chin. J. Polym. Sci. 2018, 36, 1027–1035. [Google Scholar] [CrossRef]
  218. Zhang, M.; Lai, W.; Su, L.; Wu, G. Effect of Catalyst on the Molecular Structure and Thermal Properties of Isosorbide Polycarbonates. Ind. Eng. Chem. Res. 2018, 57, 4824–4831. [Google Scholar] [CrossRef]
  219. Ouhib, F.; Meabe, L.; Mahmoud, A.; Eshraghi, N.; Grignard, B.; Thomassin, J.-M.; Aqil, A.; Boschini, F.; Jérôme, C.; Mecerreyes, D.; et al. CO2-sourced polycarbonates as solid electrolytes for room temperature operating lithium batteries. J. Mater. Chem. A 2019, 7, 9844–9853. [Google Scholar] [CrossRef]
  220. Ngassam Tounzoua, C.; Grignard, B.; Brege, A.; Jerome, C.; Tassaing, T.; Mereau, R.; Detrembleur, C. A Catalytic Domino Approach toward Oxo-Alkyl Carbonates and Polycarbonates from CO2, Propargylic Alcohols, and (Mono- and Di-)Alcohols. ACS Sustain. Chem. Eng. 2020, 8, 9698–9710. [Google Scholar] [CrossRef]
  221. Siragusa, F.; Van Den Broeck, E.; Ocando, C.; Müller, A.J.; De Smet, G.; Maes, B.U.W.; De Winter, J.; Van Speybroeck, V.; Grignard, B.; Detrembleur, C. Access to Biorenewable and CO2-Based Polycarbonates from Exovinylene Cyclic Carbonates. ACS Sustain. Chem. Eng. 2021, 9, 1714–1728. [Google Scholar] [CrossRef]
Scheme 1. Main strategies used to afford (a)cyclic carbonates and polycarbonates from the coupling of alcohols and CO2.
Scheme 1. Main strategies used to afford (a)cyclic carbonates and polycarbonates from the coupling of alcohols and CO2.
Catalysts 12 00124 sch001
Scheme 2. The CO2–alcohol coupling to synthetize organic carbonates.
Scheme 2. The CO2–alcohol coupling to synthetize organic carbonates.
Catalysts 12 00124 sch002
Scheme 3. Catalytic cycle for the formation of DMC using BU2Sn(OMe)2. Adapted from [40].
Scheme 3. Catalytic cycle for the formation of DMC using BU2Sn(OMe)2. Adapted from [40].
Catalysts 12 00124 sch003
Scheme 4. Catalytic cycle for the reaction of water with 2-cyanopyridine over CeO2. Adapted from [44].
Scheme 4. Catalytic cycle for the reaction of water with 2-cyanopyridine over CeO2. Adapted from [44].
Catalysts 12 00124 sch004
Scheme 5. Library of linear carbonates afforded via the CeO2/2-cyanopyridine route (CeO2 0.34 g, alcohol/2-cyanopyridine = 20 mmol:100 mmol, T = 120–150 °C, PCO2 = 5 MPa, t = 16–48 h). Reprinted from [54].
Scheme 5. Library of linear carbonates afforded via the CeO2/2-cyanopyridine route (CeO2 0.34 g, alcohol/2-cyanopyridine = 20 mmol:100 mmol, T = 120–150 °C, PCO2 = 5 MPa, t = 16–48 h). Reprinted from [54].
Catalysts 12 00124 sch005
Scheme 6. Mechanisms considered for the computational studies of CO2 fixation onto alcohols via the organic superbase DBU. Mechanism 4 displays the most favored pathway in terms of the kinetics and thermodynamics. Adapted from [58].
Scheme 6. Mechanisms considered for the computational studies of CO2 fixation onto alcohols via the organic superbase DBU. Mechanism 4 displays the most favored pathway in terms of the kinetics and thermodynamics. Adapted from [58].
Catalysts 12 00124 sch006
Scheme 7. TBD-CO2 stabilized adduct isolated and characterized by X-ray diffraction. Adapted from [59].
Scheme 7. TBD-CO2 stabilized adduct isolated and characterized by X-ray diffraction. Adapted from [59].
Catalysts 12 00124 sch007
Scheme 8. Reaction mechanism proposed for the synthesis of acyclic carbonates from alcohols, CO2 and carbodiimides. Adapted from [63].
Scheme 8. Reaction mechanism proposed for the synthesis of acyclic carbonates from alcohols, CO2 and carbodiimides. Adapted from [63].
Catalysts 12 00124 sch008
Scheme 9. Cascade reaction for the formation of DEC from 2,2-diethoxypropane, initiated by ethanol. Adapted from [68].
Scheme 9. Cascade reaction for the formation of DEC from 2,2-diethoxypropane, initiated by ethanol. Adapted from [68].
Catalysts 12 00124 sch009
Scheme 10. Synthetic pathway for the synthesis of glycerol carbonate from glycerol catalyzed by n-Bu2Sn(OCH3)2. Adapted from [78].
Scheme 10. Synthetic pathway for the synthesis of glycerol carbonate from glycerol catalyzed by n-Bu2Sn(OCH3)2. Adapted from [78].
Catalysts 12 00124 sch010
Scheme 11. Catalytic cycle for the synthesis of glycerol carbonate from glycerol and CO2 using n-Bu2Sn(OCH3)2 as the catalyst. Adapted from [41].
Scheme 11. Catalytic cycle for the synthesis of glycerol carbonate from glycerol and CO2 using n-Bu2Sn(OCH3)2 as the catalyst. Adapted from [41].
Catalysts 12 00124 sch011
Scheme 12. Library of affordable 5- and 6-membered cyclic carbonates obtained with the a dual CeO2/2-cyanopyridine system (CeO2 = 2 mmol, diol/2-cyanopyridine = 10 mmol/100 mmol, T = 130–150 °C, PCO2 = 5 MPa, t = 1–8 h). Adapted from [81].
Scheme 12. Library of affordable 5- and 6-membered cyclic carbonates obtained with the a dual CeO2/2-cyanopyridine system (CeO2 = 2 mmol, diol/2-cyanopyridine = 10 mmol/100 mmol, T = 130–150 °C, PCO2 = 5 MPa, t = 1–8 h). Adapted from [81].
Catalysts 12 00124 sch012
Scheme 13. General “leaving group” or “alkylation” strategy used to obtain cyclic carbonates from diols.
Scheme 13. General “leaving group” or “alkylation” strategy used to obtain cyclic carbonates from diols.
Catalysts 12 00124 sch013
Scheme 14. Computational study of the synthesis of 6-membered cyclic carbonate from a diol using the DBU/TsCl system. G. L. Gregory, M. Ulmann and A. Buchard, RSC Adv., 2015, 5, 39404; Published by The Royal Society of Chemistry [84].
Scheme 14. Computational study of the synthesis of 6-membered cyclic carbonate from a diol using the DBU/TsCl system. G. L. Gregory, M. Ulmann and A. Buchard, RSC Adv., 2015, 5, 39404; Published by The Royal Society of Chemistry [84].
Catalysts 12 00124 sch014
Scheme 15. Mechanism proposed for the formation of cyclic carbonates from diols catalyzed by ionic liquids and Cs2CO3 in the presence of alkyl halides. Adapted from [90].
Scheme 15. Mechanism proposed for the formation of cyclic carbonates from diols catalyzed by ionic liquids and Cs2CO3 in the presence of alkyl halides. Adapted from [90].
Catalysts 12 00124 sch015
Scheme 16. Competitive pathway for the formation of carbonate linkages leading to non-CO2-based cyclic carbonates. Adapted from [90].
Scheme 16. Competitive pathway for the formation of carbonate linkages leading to non-CO2-based cyclic carbonates. Adapted from [90].
Catalysts 12 00124 sch016
Scheme 17. StilbPAM structure as the base for the Brønsted acid-base bifunctional salt used in the synthesis of iodocarbonates from allyl alcohols.
Scheme 17. StilbPAM structure as the base for the Brønsted acid-base bifunctional salt used in the synthesis of iodocarbonates from allyl alcohols.
Catalysts 12 00124 sch017
Scheme 18. Example of iodocarbonate synthesis from allyl alcohol using Brønsted acid-base bifunctional salt catalysts and N-iodosuccinimide as the iodinating agent. Adapted from [92].
Scheme 18. Example of iodocarbonate synthesis from allyl alcohol using Brønsted acid-base bifunctional salt catalysts and N-iodosuccinimide as the iodinating agent. Adapted from [92].
Catalysts 12 00124 sch018
Scheme 19. Proposed mechanism for the formation of iodocarbonate from allyl alcohol using t-BuOI as the iodinating agent. Adapted from [93].
Scheme 19. Proposed mechanism for the formation of iodocarbonate from allyl alcohol using t-BuOI as the iodinating agent. Adapted from [93].
Catalysts 12 00124 sch019
Scheme 20. Synthetic strategy used to obtain cyclic carbonates from terminal alkenes via in situ generation of halohydrin. Adapted from [97].
Scheme 20. Synthetic strategy used to obtain cyclic carbonates from terminal alkenes via in situ generation of halohydrin. Adapted from [97].
Catalysts 12 00124 sch020
Scheme 21. Synthesis of exovinylene cyclic carbonates from the coupling of propargylic alcohols and CO2.
Scheme 21. Synthesis of exovinylene cyclic carbonates from the coupling of propargylic alcohols and CO2.
Catalysts 12 00124 sch021
Scheme 22. Undesired side reactions of ring-opening of exovinylene cyclic carbonates by unreacted propargylic alcohols.
Scheme 22. Undesired side reactions of ring-opening of exovinylene cyclic carbonates by unreacted propargylic alcohols.
Catalysts 12 00124 sch022
Scheme 23. Selective synthesis of vinylene carbonates rather than exovinylene cyclic carbonates using DBU/AgOAc catalytic systems on secondary propargylic alcohols.
Scheme 23. Selective synthesis of vinylene carbonates rather than exovinylene cyclic carbonates using DBU/AgOAc catalytic systems on secondary propargylic alcohols.
Catalysts 12 00124 sch023
Scheme 24. Catalytic cycle of the ZnI2/NEt3 dual system for the carboxylation of tertiary propargylic alcohols to afford exovinylene cyclic carbonates. Adapted from [125].
Scheme 24. Catalytic cycle of the ZnI2/NEt3 dual system for the carboxylation of tertiary propargylic alcohols to afford exovinylene cyclic carbonates. Adapted from [125].
Catalysts 12 00124 sch024
Scheme 25. Mechanism for the formation of polycarbonates from the polycondensation of diols and CO2 promoted by the CyTMG/R3P/CX4 system.
Scheme 25. Mechanism for the formation of polycarbonates from the polycondensation of diols and CO2 promoted by the CyTMG/R3P/CX4 system.
Catalysts 12 00124 sch025
Scheme 26. Competitive pathways between carbonyl attack (blue) and methylene attack (pink) for the formation of carbonate or ether linkages. Adapted from [131].
Scheme 26. Competitive pathways between carbonyl attack (blue) and methylene attack (pink) for the formation of carbonate or ether linkages. Adapted from [131].
Catalysts 12 00124 sch026
Scheme 27. Terpolymerization synthetic pathway for the formation of polycarbonates using diolates and dihalides.
Scheme 27. Terpolymerization synthetic pathway for the formation of polycarbonates using diolates and dihalides.
Catalysts 12 00124 sch027
Scheme 28. Competitive pathways between methylene attack (red) and carbonyl attack (blue) to produce poly(ether-co-carbonates) or homopolycarbonates via ROP of 5-membered cyclic carbonates.
Scheme 28. Competitive pathways between methylene attack (red) and carbonyl attack (blue) to produce poly(ether-co-carbonates) or homopolycarbonates via ROP of 5-membered cyclic carbonates.
Catalysts 12 00124 sch028
Scheme 29. Schematic representation of polycarbonate synthesis via melt polycondensation of diols with DPC or DMC (R2 = CH3 or C6H5).
Scheme 29. Schematic representation of polycarbonate synthesis via melt polycondensation of diols with DPC or DMC (R2 = CH3 or C6H5).
Catalysts 12 00124 sch029
Scheme 30. Carbonate metathesis polymerization principle catalyzed by K2CO3. DMC is removed under vacuum to displace the equilibrium toward PC formation. Adapted from [197].
Scheme 30. Carbonate metathesis polymerization principle catalyzed by K2CO3. DMC is removed under vacuum to displace the equilibrium toward PC formation. Adapted from [197].
Catalysts 12 00124 sch030
Scheme 31. The 3D structure of isosorbide with explicit intramolecular H-bonds for endo-OH.
Scheme 31. The 3D structure of isosorbide with explicit intramolecular H-bonds for endo-OH.
Catalysts 12 00124 sch031
Scheme 32. Regioselective alcoholysis of exovinylene cyclic carbonates.
Scheme 32. Regioselective alcoholysis of exovinylene cyclic carbonates.
Catalysts 12 00124 sch032
Scheme 33. Synthetic procedure used to access regioregular poly(β-oxo carbonates). Adapted from [108].
Scheme 33. Synthetic procedure used to access regioregular poly(β-oxo carbonates). Adapted from [108].
Catalysts 12 00124 sch033
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Brege, A.; Grignard, B.; Méreau, R.; Detrembleur, C.; Jerome, C.; Tassaing, T. En Route to CO2-Based (a)Cyclic Carbonates and Polycarbonates from Alcohols Substrates by Direct and Indirect Approaches. Catalysts 2022, 12, 124. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12020124

AMA Style

Brege A, Grignard B, Méreau R, Detrembleur C, Jerome C, Tassaing T. En Route to CO2-Based (a)Cyclic Carbonates and Polycarbonates from Alcohols Substrates by Direct and Indirect Approaches. Catalysts. 2022; 12(2):124. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12020124

Chicago/Turabian Style

Brege, Antoine, Bruno Grignard, Raphaël Méreau, Christophe Detrembleur, Christine Jerome, and Thierry Tassaing. 2022. "En Route to CO2-Based (a)Cyclic Carbonates and Polycarbonates from Alcohols Substrates by Direct and Indirect Approaches" Catalysts 12, no. 2: 124. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12020124

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