Next Article in Journal
Developing an Assisting Device to Reduce the Vibration on the Hands of Elders
Next Article in Special Issue
Nanostructured Na2CaP2O7: A New and Efficient Catalyst for One-Pot Synthesis of 2-Amino-3-Cyanopyridine Derivatives and Evaluation of Their Antibacterial Activity
Previous Article in Journal
Photovoltaic Prediction Software: Evaluation with Real Data from Northern Spain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Five-Membered Cyclic Carbonates: Versatility for Applications in Organic Synthesis, Pharmaceutical, and Materials Sciences

1
Université d’Orléans et CNRS, Institut de Chimie Organique et Analytique, UMR 7311, BP 6759, F-45067 Orléans, France
2
Laboratório de Síntese Orgânica Limpa LASOL, Centro de Ciências Químicas, Farmacêuticas e de Alimentos CCQFA, Universidade Federal de Pelotas UFPel, P.O. Box 354, Pelotas 96010-900, RS, Brazil
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2021, 11(11), 5024; https://doi.org/10.3390/app11115024
Submission received: 30 April 2021 / Revised: 24 May 2021 / Accepted: 27 May 2021 / Published: 29 May 2021

Abstract

:
This review presents the recent advances involving several applications of five-membered cyclic carbonates and derivatives. With more than 150 references, it covers the period from 2012 to 2020, with special emphasis on the use of five-membered cyclic carbonates as building blocks for organic synthesis and material elaboration. We demonstrate the application of cyclic carbonates in several important chemical transformations, such as decarboxylation, hydrogenation, and transesterification reactions, among others. The presence of cyclic carbonates in molecules with high biological potential is also displayed, together with the importance of these compounds in the preparation of materials such as urethanes, polyurethanes, and flame retardants.

1. Introduction

Organic carbonates are important synthetic intermediates in a number of reactions [1,2]. They are characterized by presenting a carbonyl group flanked by two oxygen atoms and may be linear or cyclic. Five-membered cyclic carbonates can be obtained through the combination of CO2 with diols and epoxides [3,4,5,6] or by conversion of glycerol into glycerol carbonate and derivatives (Figure 1) [7].
Among the above-mentioned substrates, glycerol derivatives are particularly important starting materials, which have been widely exploited for different applications, such as their potent antiwear action when added to fuels [8]. For example, glycerol carbonate (GC) 1a was used as a monomer in the synthesis of hyperbranched aliphatic polyethers [9] and polyurethanes [10].
The versatility of GC-derived five-membered cyclic carbonates as synthons in organic synthesis was explored by Rollin and co-workers [11,12,13,14]. However, their extensive reactivity was reviewed only recently [15,16,17,18,19,20,21]. Therefore, being interested in the chemical versatility of the carbonate moiety, we provide here an updated review of the recent advances involving different applications of five-membered cyclic carbonates and derivatives in organic synthesis.
For an optimized discussion, this review presents papers published between 2012 and 2020, organized in three sections, according to the potential application of the synthesized compounds in chemistry, biology, and material science studies.

2. Applications in Chemistry

2.1. Decarboxylation of Cyclic Carbonates

The decarboxylation reaction of glycerol carbonate (GC) 1a with concomitant generation of CO2 is a typical method used to obtain the 1,2-anhydroglycerol derivative denominated glycidol, which has a variety of industrial uses, such as in surface coatings, as a gelation agent in solid propellants, as a stabilizer for natural oils and vinyl polymers, and as a demulsifier (emulsion breaker). Glycidol is also used as an intermediate in the synthesis of esters and amines and it can be of high-value in the production of epoxy resins, polyurethanes, and in the controlled etherification of glycerol to form polyglycerols [22,23].
At the industrial level, glycidol can be produced through oxidation of allyl alcohol using hydrogen peroxide in the presence of a titanium silicate catalyst [24]. However, all conventional processes used to prepare glycidol generate undesirable byproducts; therefore, producing glycidol from glycerol derivatives stands as an interesting alternative to minimize the generation of waste.
Kim, Lee, and co-workers [25] performed the decarboxylation of GC 1a to glycidol 2a using various ionic liquids (ILs) as catalysts (0.5 mol%) at 175 °C under reduced pressure (2.67 kPa) for 30 min, releasing CO2 as the only co-product (Scheme 1). After studying the catalytic efficiency of anions (with a range of sodium salts) and cations (with a range of metallic chlorides), the authors discovered that both the yield and selectivity in producing glycidol 2a could be improved by using ILs. The best results were attained using [bmim] nitrate and iodide (59–73% yields), due to the strong interaction of the anion of the IL with 1a and the weak interaction with 2a. The effects of the N-cations were less pronounced, while [bdmim]NO3, [bmim]NO3, [Bu4N]NO3, and [MPPyr]NO3 all proved similarly satisfactory for the production of 2a. The addition of Lewis acid Zn(NO3)2 to the reaction medium caused an increase in the glycidol yield, which was attributed to the weakening of the hydrogen-bonding interactions between the nitrate anion of [bmim]NO3 and the hydroxyl groups of 1a and 2a.
In a continuous process, a solvent was used to minimize the contact between the IL and glycidol while simultaneously removing 2a as soon as it formed. The use of high-boiling solvents, such as glycol dimethyl ether, dibenzyl ether, and dibutyl phthalate, increased the yield (83–98%) and selectivity in both batch and continuous reactions. The mechanism of the decarboxylation of 1a was theoretically investigated under [EMIm]NO3 catalysis and computational calculations showed that the decarboxylation of 1a to 2a to proceeded in two steps: (1) nitrate-assisted ring-opening of 1a, followed by formation of a 3-membered ring; (2) proton-transfer from HNO3 to an oxygen atom of the carbonate group, along with the simultaneous loss of CO2. It is worth mentioning that Li and co-workers [26] recently developed a similar protocol for synthesizing epoxybutane from 1,2-butanediol by using the ionic liquid 1-butyl-3-methylimidazolium bromide in batch or continuous processes.
With a view to getting away from conventional heating, the conversion of 1a to 2a was studied through alternative methodologies: solvent-free thermal activation at 150 °C over 23 h (about 35% conversion yield), ultrasound activation at ca. 60 °C (44–48% conversion yield), and microwave activation in liquid phase at 150 °C over 30 min (48–71% conversion yield). In all cases, the authors used a ZSM-5 zeolite or a zinc oxide-supported nanoscaled cobalt oxide (RT-10CoZn, 5 wt%) as catalysts (Scheme 2) [27]. In thermal activation, the ZSM-5 zeolite showed 100% selectivity for 2a, while it did not exceed 60% with RT-10CoZn. In ultrasonic activation, RT-10CoZn showed 100% selectivity for 2a, whereas in microwave activation, RT-10CoZn reaches 71% conversion yield to 2a with total selectivity, indicating that this catalyst delivers the best performance in the shortest reaction time (30 min vs. 6–7 h under ultrasonic activation and 23 h under thermal activation). The hypothetical mechanism is based on the polymerization of 1a to poly-(glyceryl-1,2-carbonate), which then forms 2a through CO2 release. The reuse of both heterogeneous catalysts was not mentioned.
Lee and co-workers [28] studied the use of a binary zinc–lanthanum catalytic system (ZnO/La2O3; Zn:La ratio = 1:9) in the decarboxylative process. When 10 wt% of the mixed oxides was employed in the reaction, 2a was obtained with 76% yield after 2 h reaction under reduced pressure (0.05 MPa) in a nonspecific solvent below 180 °C (Scheme 2). The evaluation of the catalytic activity showed that an increase in the La molar ratio correlates with a yield increase of 2a, due to equilibrium between acidic and basic sites on the surface of the catalyst. However, when the reaction was performed in the presence of La2O3 alone, 2a was obtained with 47% yield only.
Mahajani and co-workers [29] produced 2a by using the same ZnO/La2O3 system, albeit with a molar ratio for Zn/La of 4:1. In this work, the performance of the catalyst was studied up to four re-uses (about 64–97% conversion) and a decrease of activity was noted after each cycle, because of the adsorption of carbonaceous side products on the catalytic basic sites. This problem was minimized by calcination of the catalyst to restore its activity. As the major objective of the authors was to synthesize 1a by dimethyl carbonate transesterification with glycerol, no additional studies were carried out to increase the catalyst’s activity in preparing 2a.
Hyunjoo Lee and co-workers [30] described a one-pot consecutive glycidol synthesis method from glycerol, whereby carbonylation of glycerol 3a with urea first takes place under high pressure using a homogeneous zinc catalyst to produce GC 1a. Subsequently, decarboxylation is used to obtain the desired 2a product (Scheme 3). This second reaction proceeds in acetonitrile in the presence of zinc acetate to give glycidol 2a with 67% yield after 1.5 h. According to the authors, the high catalytic activity of Zn(OAc)2 can be attributed to the formation of the intermediate species [Zn(NH3)x•(OAc)2].
An original decarboxylation process involving GC 1 was described by Falivene and co-workers [31] through an organocatalytic coupling reaction between the bromolactide 4 and carbonates 1, in the presence of the heterocyclic carbene 1,3-bis(2,4,6-trimethylphenyl)imidazol-2-ylidene (IMes), resulting in the formation of chiral keto-diesters 5 (Scheme 4). The reaction was found to proceed at 100 °C in times ranging from 13 h to 15 h, delivering the desired products in high yields (72–93%).
More recently, the use of palladium catalysis for the decarboxylation of GC 1a was reported [32,33]. Kleij and co-workers [32] described the synthesis of ethers via a Pd-catalyzed reaction of carbonates 6 with phenols 7 in THF in the presence of cesium carbonate (Cs2CO3) and phosphoramidite or phosphine ligands (Scheme 5). The authors observed that the choice of the ligand influences the regioselectivity of the reaction. Thus, by changing the ligands it was possible to direct the reaction to prepare either branched- 8 or linear allylic aryl ethers 8. Reacting the cyclic carbonates at 0 °C for 36 h led to aryl ethers 8, while running the same reaction at room temperature for 12 h mainly provided 9. The products were obtained in moderate to excellent yields (49–98%) and with considerable enantio-induction. In a closely related work, Guo and co-workers [33] developed an efficient protocol to construct densely functionalized quaternary carbons with the concomitant formation of a new Csp3-Csp3 bond through decarboxylation of carbonate 6. The reaction was carried out in DMF at room temperature for 12 h, using palladium catalysis with the L2 ligand and resulted in the formation of branched aldehydes 10 (32–91%) and substituted acroleins 11 (26–84%).

2.2. Hydrogenation Reaction

Catalytic hydrogenation of organic carbonates is of synthetic interest as an environmentally benign approach to fundamental synthetic building blocks, such as alcohols and diols [34,35]. In 2012, Han and co-workers [36] developed a process to obtain ethylene glycol and methanol via catalytic hydrogenation of cyclic carbonates by using (PNP)RuII pincer complexes as catalysts (Scheme 6). In this work, ethylene glycol was prepared through insertion of CO2 into epoxides (omega process), as an option for the recycling of CO2. This catalytic process allowed them to obtain a variety of ethylene glycols (EGs) 12 with 96–99% yields, with co-production of methanol (up to 99% yield) from different cyclic carbonates 1. The reactions proceeded over 4–10 h in THF at 140 °C under 50 atm of H2 and in the presence of catalytic amounts of the RuII complex I (0.02–0.10 mol%) and tBuOK (1.0 equivalent relative to the catalyst) in a 125 mL autoclave. It is noteworthy that when using GC 1a protected with a methyl or a benzyl group, the corresponding EGs 12 were obtained with 99% and 97% yields, respectively, after 4 h. Additionally, it was observed that sterically hindered cyclic carbonates 1 required longer reaction times or higher amounts of the catalyst. The protocol was successfully applied to a six-membered ring carbonate, giving the respective diol after 2 h of reaction. Moreover, the protocol was used in the depolymerization of poly(propylene carbonate) (PPC), resulting in propane-1,2-diol and methanol after 24 h, both with 99% conversion.
Similarly, the Ru-lutidine-bridged N-heterocyclic carbene complex (Ru complex II) was used as a catalyst in the hydrogenation of ethylene carbonate 1c (R = H) (Scheme 6) [37]. In this work, glycol and methanol were obtained with 92% and 42% yields, with 99% conversion of the carbonate under 5 MPa of H2 over 12 h at 130 °C. A N-butyl-substituted CNC–pincer ligand and RuHCl(CO)(PPh3)3 were applied to prepare the Ru-NHC complex, formed in situ in the presence of tBuOK (10 mol%) as a base. From the optimization studies, the authors concluded that the high catalytic activity might be ascribed to the metal–ligand cooperation through aromatization–dearomatization of the lutidine framework. Furthermore, air and moisture-stable Ru-based complexes were applied for indirect conversion of CO2.
In the same line, the pincer pyridine-bridged N-heterocyclic carbene-ruthenium complex (Ru complex III) exhibited high catalytic activity in the hydrogenation of cyclic carbonates 13 (Scheme 7) [38]. When carried out in THF in the presence of K3PO4 and H2 (50 atm) at 140 °C, with a catalyst loading of 0.5 mol%, the reaction provided nearly quantitative yields of the corresponding diol 12 and methanol. Even when using a catalytic amount of a weak base and a low catalyst loading, a range of substrates with varying steric congestion underwent the transformation.
The production of diols and methanol was reported by using copper–silica nanocomposite catalysts with a uniform Cu dispersion, prepared using a precipitation–gel method (Cu-SiO2-PG) (Scheme 8) [39]. This system promoted hydrogenation of cyclic carbonates 1 (87–97% conversion), providing methanol and diols 12 at conversion rates of 29–98% in both batch and fixed-bed continuous flow reactors, under relatively mild conditions. All the cyclic carbonates involved gave the respective diols in high yield, with the exception of GC 1a, which provided glycerol with 29% yield only.
A catalytic process using a basic zwitterionic salt as an alternative to inorganic bases was employed to synthesize glycols through hydrolysis of cyclic carbonates [40]. Six examples of glycols 12 were prepared with 64–98% yields by using 1–6 wt% zwitterionic 1,3-dimethylimidazolium-2-carboxylate (DMImC), cyclic carbonate 1/H2O = 1:1.5 (molar ratio), at 140 °C and under 0.4 MPa for 30 min to 6 h (Scheme 8). A reaction time of 24 h was required to convert 4-phenyl-1,3-dioxolan-2-one 1b into the corresponding vicinal diol 12a with 90% yield. The catalyst DMImC was recovered and re-used six times at 140 °C, 0.4 MPa for 30 min, keeping good selectivity and high yield.
Based on the experimental results, NMR spectroscopy, and theoretical calculations, a possible catalytic cycle was suggested (Scheme 9). Generated by formation of the conjugated acid DMImCH, the hydroxide ion promotes opening of the cyclic carbonate 1b, forming the intermediate I, which is converted to II by intramolecular hydrogen transfer. Then, II is converted into a possible imidazolium alkylcarbonate salt III or IV by reaction with DMImCH. Subsequent elimination of DMImC and production of II occurs. The reaction of II with OH- affords V with CO2 and H2O release to the reaction medium. Finally, a proton from DMImCH is captured by V, producing the desired glycol 12a while regenerating the catalyst for a new cycle.
The indirect hydrogenation of CO2 via the ethylene carbonate intermediate 1c to produce ethylene glycol 12b and methanol was proposed by using mesoporous silica (KIT-6, MCM-41, and SBA-15) as the support for copper-based catalysts (Scheme 10, condition a) [41]. This attractive approach showed 100% conversion of carbonate 1c, yielding 94.7% of diol 12b and 62.3% of CH3OH 7a in only 4 h reaction time over Cu/SBA-15 under the optimal conditions. This copper-based catalyst exhibited superior catalytic activity with turnover numbers of 22.0 and 11.4 (mol mol−1) towards diol and methanol, respectively. It is noteworthy that the higher catalytic performance was due to the synergic effect between the Cu0 and Cu+ species that co-exist on the surface of the catalysts and also to a larger metallic surface area on SBA-15 when compared with KIT-6 and MCM-41. Additionally, the thermodynamic calculation showed the ethylene glycol hydrogenation to be exothermic (ΔrHθm = −71.59 kJ mol−1) and thermodynamically favorable (ΔrGθm = −25.62 kJ mol−1).
In another study, the same authors reported the selective hydrogenation of ethylene carbonate 1c with 100% of conversion over Cu/SiO2 catalysts, prepared using the ammonia evaporation method, with silica sol as the silica source (Scheme 10, condition b) [42]. The 10%Cu/SiO2-AE catalyst with moderate copper loading exhibited better catalytic activity to co-produce methanol 7a with 70.8% yield and EG 12c at 98.0% yield. After catalyst characterization, the authors disclosed the synergic effects of Cu0 and Cu+ species in this reaction; while Cu0 species promoted H2 dissociation, Cu+ species were adsorbed on the carbonyl group of EC. Moreover, in catalyst reusability experiments, the 10%Cu/SiO2-AE catalyst showed agglomeration of Cu and Cu2O particles, which were responsible for the gradually decreasing activity after consecutive recycling runs.
Similarly, nanoarray Cu/SiO2 catalysts embedded in monolithic channels (CuSi-NAM) were applied in hydrogenation reactions (Scheme 10, condition c) [43]. Monolithic Cu-based catalysts exhibited high activity in the conversion of ethylene carbonate 1c (99%) into EG 12c and methanol 7a selectivity, with yields of 97% and 50%, respectively.
The most frequently used methods to produce diols are described above, which involve catalytic hydrogenation or hydrolysis of cyclic carbonates using organic solvents. Fu and co-workers recently published a new procedure for the conversion of cyclic carbonates 1 to diols 12 in overheated water, without an external catalyst or additive (Scheme 11) [44]. The reaction of ethylene carbonate proceeded smoothly to give the desired glycol in high yield (99%) along with CO2 as a byproduct. The reaction was studied at different temperatures (100 °C to 250 °C) with water filling of 25% for 2 h. In general, a higher temperature was better in promoting the reaction and the methodology was extended to methyl- and ethyl-substituted ethylene carbonates. With these substrates, 10 h was needed to provide the corresponding diols with 59% and 62% yield, respectively.
The same diols can be obtained from cyclic carbonates via a deprotection reaction. In this sense, the new heterogeneous zinc/imidazole catalyst was evaluated in various reactions such as transesterification and transcarbonation [45]. A small amount of the suitable zinc/imidazole catalyst immobilized on polystyrene resin proved effective to prepare 3-phenoxypropane-1,2-diol 12 from 4-(phenoxymethyl)-1,3-dioxolan-2-one 1d after 6 h in refluxing methanol (Scheme 11).
An extension of the deprotection reaction was recently described by Perin and co-workers [46], whereby 4-chalcogenomethyl-1,3-dioxolan-2-ones 1 were synthesized from GC 1a (Section 2.4), then converted into diols 12 by using potassium carbonate in PEG-400 for 12 h at 25 °C (Scheme 12). Six vicinal diols bearing an arylsulfanyl or arylselenanyl moiety were obtained in excellent yields.
Pongrácz and Kollár described the asymmetric hydroformylation of racemic 4-vinyl-1,3-dioxolan-2-one 1e to prepare the aldehyde 1f (Scheme 13) [47]. The authors reported the chemoselective and regiospecific formation of 1f by using the “pre-formed” PtCl2(diphosphine) and tin(II) chloride catalyst and (2S,4S)-2,4-bis(diphenylphosphinopentane) [(S,S)-BDPP)] as an enantiopure diphosphine ligand. The platinum-containing catalytic systems provided nearly complete conversion of compound 1e after 3, 19, and 72 h at 100, 60, and 40 °C, respectively. In addition, a hydrogenated product 1g was formed, even under optimal conditions. In this work, the kinetic resolution of the aldehyde indicated that (S)-(1f) was preferably formed in the temperature range of 40–100 °C.
Werner and co-workers reported on the first use of an earth-abundant metal complex as a catalyst for the transfer hydrogenation of carbonate 14, opening an indirect route for the reduction of CO2 to methanol (Scheme 14) [48]. An advantage of this method, according to the authors, is the use of isopropanol as a hydrogen source, avoiding the use of flammable hydrogen under high pressures. Using a metal-based catalyst, it was possible to perform the transfer hydrogenation of GC 1a selectively, using the PNP pincer-type iron complex C17H38BrFeNOP2 15. This reaction occurred at 140 °C over 6 to 24 h, obtaining the respective diols 12 and MeOH as products with 51–97% and 39–92% yield, respectively.

2.3. Transesterification of Cyclic Carbonates

The transesterification of cyclic carbonates, especially ethylene carbonate 1c, to give non cyclic carbonates has attracted much attention, due to their role as building blocks for the production of chemical intermediates in the polymer industry [49].
Hahm and co-workers [50] reported the synthesis of dimethyl carbonate 16a via transesterification of ethylene carbonate 1c using an excess of methanol (1:15) over 4 h with ionic liquid catalysts supported on the mesoporous cellular foam (MCF) (Scheme 15). In this work, 4-diazobicyclo[2.2.2]octane (DABCO) provided the highest yield of dimethyl carbonate 16a (84%), while the heterogeneous catalyst [DABCO]OH@MCF gave around 77% yield for the expected product. The ionic-liquid-supported catalyst was re-used four times with only 8% decrease of 16a yield. At higher temperatures (up to 150 °C), decomposition of ethylene carbonate 1c or degradation of the immobilized species was observed, causing a decrease in 16a yield. Furthermore, the chemically immobilized [DABCO]OH@MCF catalyst proved to be more stable than the physically impregnated DABCO/MCF.
In a closely related study, Saptal and Bhanage [51] developed the synthesis of a series of bifunctional, DABCO-based, hydroxyl-functionalized ILs, which were applied in the conversion of carbon dioxide into valuable chemicals, such as cyclic carbonates, the transesterification of which was also considered. Here, transesterification of ethylene carbonate 1c with methanol to prepare dimethyl carbonate 16a proceeded at a moderate temperature in the presence of different ionic liquids. The [DABCO–PDO][I], functionalized with propane diol functional groups, gave the highest conversion rate and yield for dimethyl carbonate 16a, with 93% yield after 4 h at 80 °C (Scheme 15). One major advantage of this protocol is that no transition metal, solvent, or additive is used.

2.4. Substitution Reaction

Cyclic carbonates can be used to prepare new organochalcogen derivatives. The interest in compounds containing sulfur, selenium, and tellurium atoms has been continuously increasing, due to their potential applications as synthons for the synthesis of heterocycles [52], in cross-coupling reactions [53,54], and in asymmetric synthesis [55,56]. Those compounds have applications in materials science [57,58,59], and as key intermediates in the synthesis of natural products [60,61,62].
Perin and co-workers [63] recently reported the synthesis of 4-arylsulfanylmethyl-1,3-dioxolan-2-ones under microwave irradiation (MW) using an environmentally benign KF/Al2O3/PEG-400 system [64]. By applying KF/Al2O3 as an efficient heterogeneous base, 3-O-tosyl glycerol 1,2-carbonate 1h (TGC) reacted with alkyl, aryl, or heteroaryl thiols 17 to provide good yields of the corresponding sulfanylated carbonates 1 after short times of MW irradiation at 60 °C. This green methodology was extended to thiophenols bearing either electron-donating (CH3) or electron-withdrawing groups (Cl and F) in the para- and ortho-positions, to provide with good yields the respective 4-arylthio-1,2-glycerol carbonates. However, when the strong electron-donating amino group was present in the ortho-position (2-aminobenzenethiol), only 65% yield of the desired product was obtained, due to the disfavoring proton abstraction from the thiol (Scheme 16). The conversion was extended, albeit in lower yields, to 2-mercaptobenzothiazolyl and dodecyl derivatives.
The possibility of developing a selective method for the synthesis of 1,3-bis-arylthiopropan-2-ols 18 by incorporation of two thiol units was explored (Scheme 17). By reacting a second equivalent of thiol 17 under 15 h conventional heating at 90 °C, the corresponding disubstituted products 18 were obtained in good to excellent yields. When conducted at 90 °C for 3 h in a microwave reactor (irradiation power of 200 W), the same compounds 18 were isolated in slightly lower yields [63].
The same year, Perin’s group reported the selective synthesis under stoichiometric control of 4-chalcogenomethyl-1,3-dioxolan-2-ones and 1,3-bis(organylchalcogenyl)propan-2-ols [46]. To obtain the GC-derived chalcogenoethers 1, diaryl diselenides 19 or disulfides 20 (0.6 equivalent) were first reduced using NaBH4/PEG-400 under inert atmosphere, before the addition of TGC 1h (Scheme 18). A library of eleven compounds 1 was prepared with 55–92% yield after 2 h reaction at 50 °C. The protocol was extended to aliphatic diselenides, and when dibutyl diselenide was used, the corresponding selenide was isolated with 74% yield. In general, diaryl disulfides lead to better yields when compared to the seleno-analogues.
During the optimization process, it was observed that increasing the amount of diaryl diselenides 19 or disulfides 20 to 1.2 equivalent led to ring-opening of the carbonate unit of 1h over 1 h, resulting in symmetrical and unsymmetrical 1,3-bis(organochalcogenyl)propan-2-ols 18 in moderate to excellent yields (Scheme 19). The reaction works well when either electron-donating or electron-withdrawing groups are present in the aromatic unit R. It is worth noting that when using dimesityl diselenide in the synthesis of symmetrical compounds 18, 3 h reaction was needed to reach 94% yield of the expected product. Moreover when using 1,2-bis(2-chlorophenyl)disulfide in the synthesis of unsymmetrical compounds 18′, the respective product was isolated with 40% yield after 1 h reaction.
The stereoselective construction of cyclic carbonate systems is another attractive line of studies that has been explored. In this sense, racemic vinyl-substituted cyclic carbonates 6 are important substrates for copper-mediated SN2 reactions to access useful hydroxyl functionalities 9 and 21 in a single step (Table 1, entries 1–3) [65], or for combining a palladium catalytic system and a boron reagent to obtain chiral alcohols 12 and ethers 8 (Table 1, entries 4–6) [66].
Miralles and co-workers [65] explored the allyl–boryl coupling of a series of substituted vinyl cyclic carbonates 6 with B2pin2 22 in the presence of CuCl, Cs2CO3, and the N-heterocyclic carbene ligand L1 (Table 1, Figure 2). Several E-configured borylated products 9 were obtained with 33–65% yields within 16 h at room temperature (Table 1, entry 1). When performed in the absence of ligand and with a higher amount of base, the allylalkyl coupling reaction with diborylmethane 23 generated compound 9 with 48–82% yields (Table 1, entry 2). Under similar conditions but using 1,2-bis(di-tert-butylphosphinomethyl)benzene L2 (Figure 2) as the ligand, the selective allyl–boryl coupling of vinyl cyclic carbonates 6 with B2pin2 22 led to the corresponding (Z)-stereoisomers, which were converted with 37–55% yields into the boracycles 20 (Table 1, entry 3) [65]. Finally, the versatility of (E)-allylic and homoallylic borylated products 9 was demonstrated through the in situ copper-catalyzed SN2′ allyl–boryl and allyl–alkyl couplings, followed by oxidative work-up. Thus, (E)-configured but-2-ene-1,4-diols and pent-2-ene-1,5-diols were isolated with 27–59% and 42–58% yields, respectively.
In the presence of a synergistic catalytic system formed by an in situ generated chiral palladium complex and a boron reagent, the asymmetric allylic substitution of racemic vinyl cyclic carbonates 6 with water and hydroxyl reagents provided a series of tertiary alcohols and ethers 8 in high yields (Table 1) [66]. The synthesis of tertiary alcohols 12 was performed by using a system containing phenylboronic acid 24, Pd2(dba)3.CHCl3, Zhou’s ligand [(R)-L3] (Figure 2), and water (10 equivalent). The corresponding products were obtained with 82–98% yields (80–98% ee) after 16 h at 40 °C in THF (Table 1, entry 4). By changing to Feringa’s ligand (S,S,S)-L4 (Figure 2) and to triethylborane 25 as the boron reagent, the alcohols 12 were isolated with 61–87% yields and high enantioselectivity (80–98% ee) (Table 1, entry 5). An enantioselective allylic etherification process was developed, involving reaction of hydroxyl reagents with racemic vinyl cyclic carbonates 6 in the presence of (R)-L3 and triethylborane 25 to provide the desired ethers 8 (56–97% yields; 37–98% ee) (Table 1, entry 6).

2.5. Metal-Catalyzed Miscellaneous Reactions

In recent years, the introduction of transition metal catalysis into diverse reactions has provided profitable results in organic chemistry. The most typical examples involve the use of palladium [67,68,69,70,71], platinum [72,73,74], rhodium [75,76,77], and magnesium [78] complexes. In addition, many studies have involved GC or vinylated carbonates as starting materials in coupling reactions to generate polymers and other important compounds.
Targeting the synthesis of new glyceryl carbonate ethers through coupling reactions, Sauthier, Visseaux, and co-workers reported the telomerization between GC 1a and 1,3-butadiene 26 (2 equivalent) in the presence of a catalytic system containing palladium bis(acetylacetonate), triphenylphosphine, and triethylamine as a base [79]. The reactions were performed over 17 h at temperatures ranging from −15 to 80 °C, giving the expected octadienylglyceryl carbonate 1i with 85% yield (Scheme 20, step 1). Compound 1i was subsequently hydrogenated under Pd/C catalysis to the corresponding octyl ether 1j (Scheme 20, step 2). The unsaturated glycerol carbonate ether 1i was co-polymerized with ε-caprolactone 27 in the presence of a neodymium initiator [Nd(2,6-tertBu2OC6H3)3] to form the polymeric derivative 28 with 15–82% yields (Scheme 20, step 3). This polymer-bearing carbonate unit presents good thermomechanical properties.
Several studies showing the efficiency of the palladium catalysis in the allylic cycloaddition of vinylated carbonates 6 have been published [80,81,82,83,84]. For example, Zhang and co-workers [80] demonstrated the interesting Pd-catalyzed decarboxylative heterocyclization of [60]fullerene for the preparation of novel vinyl-substituted [60]fullerene-fused tetrahydrofurans, pyrans, or quinolines. In 2018, through a cooperative N-heterocyclic carbene (NHC) organocatalysis and palladium catalysis, Singha and co-workers [81] performed enantioselective [5+2] annulations of enals 11 with vinylated carbonates 6 (VECs) for the synthesis of lactone 29. The reaction was performed over 10–12 h at room temperature in toluene solution in the presence of a cooperative N-heterocyclic carbene (NHC)/Pd catalytic system, along with (R)-Tol-BINAP L1 as a ligand and N-methylpiperidine. A library of twenty-six examples was obtained with 30–99% yields and 92–99% ee (Scheme 21).
Lan, Zhao, and co-workers [82] also used a Pd2(dba)3-containing system for the synthesis of nine-membered cyclic compounds 30, starting from azadienes 31 and VECs 6. In the presence of ligand L2 (5 mol%), this formal [5+4] cycloaddition led to the enantioselective preparation of twenty-four new heterocyclic compounds 30 with 70–95% yields (80–92% ee) (Scheme 21).
In order to obtain CF3-substituted tetrahydrobenzoxazonines 32, Shibata and co-workers [85] reacted VECs 6 and trifluoromethyl benzoxazinone 33 in refluxing dichloroethane in the presence of tetrakis(triphenylphosphine)-palladium(0). Using this catalytic system, various nine-membered heterocyclic compounds 32 were isolated with 33–91% yields.
In 2016, the catalytic and stereoselective formation of (Z)-1,4-but-2-enediols 9 was reported through decarboxylative hydration of VECs 6 in the presence of water as a nucleophile [86]. The reaction was performed in DMF/H2O at room temperature, using a catalytic Pd precursor, Pd/bis(sulfoxide) and a bidentate phosphine ligand. The expected unsaturated 1,4-diols 9 were obtained with 50–98% yields with >99:1 Z-stereoselectivity (Scheme 22). While different groups such as aryl, naphthyl, and heteroaryl were efficiently incorporated into diols 9, it is important to highlight that some specific substitutions, e.g., o-haloaryl or 3-pyridyl, required forced conditions with higher temperatures and larger amounts of Pd-catalyst (5 mol%) and phosphine ligand (10 mol%) in order to reach preparative yields. In order to illustrate the good reactivity of the (Z)-1,4-but-2-enediols 8, some chemical modifications were made, such as m-CPBA oxidation to the epoxide 2b, high-yield conversion into the cyclic orthoester 34, and copper-catalyzed chemoselective oxidocyclization to the isomeric lactones 35a and 35b (Scheme 23).
Racemic VECs 6 were used for the stereoselective O- or C-allylation of phenols and N-allylation of 2-hydroxypyridines by use of 1–5 mol% of [Rh(cod)Cl]2 and 2–10 mol% of a chiral diphosphine as the ligand [(R,R)-QuinoxP*; (S,S)-DIOP or (R,R)-TBDM-SLIOP] [87]. Performed in dichloroethane or toluene over 24–48 h at temperatures ranging from r.t. to 80 °C, the reaction allowed exploration of an ample range of substrates, with yields exceeding 70% in most cases (see Figure 3 for selected examples). In this extensive study, carbonate 1e was converted into compound 39a (94% yield; 97% ee) by asymmetric O-allylation of 4-bromophenol 7b (Scheme 24).
Rueping and co-workers [88] described the alkaline earth metal-catalyzed reduction of cyclic and linear organic carbonates using a readily available magnesium catalyst (Scheme 25). These mild reaction conditions provided an indirect route for the conversion of CO2 into valuable alcohols. By using a low catalyst loading and avoiding ligands, a broad scope of compounds 40 was prepared via the selective hydroboration with HBPin 41, all with excellent yields. Specifically, cyclic carbonates 1 bearing substituents such as phenyl, methyl, hexyl, tBu, as well as glycerol derivatives, were efficiently reduced with excellent yields, ranging from 88% to 95% within 3 h.
With a view to promoting C-H olefination of arenes and heteroarenes 42 by rhodium catalysis, Xu, Zhang, and co-workers [89] explored the use of dichloro(pentamethylcyclopentadienyl)rhodium(III) dimer {[Cp*RhCl2]2} as a catalyst with copper(II) acetate as the oxidant and additional silver hexafluoroantimonate. By using VEC 1e, a series of o-olefinated benzamides such as 43 (Scheme 26) were obtained (49–94% yields).
The authors also developed olefination of arylpyridines and indoles with 30–91% yield using unactivated aliphatic olefins under the same conditions, albeit using pivalic acid as a promoter. A selection of the prepared compounds is shown in Figure 4.
Under manganese catalysis, VEC 1e was employed to functionalize aryl compounds bearing electron donor groups to obtain the hydroxybutenylated derivatives 47ac [90]. The manganese complex [MnBr(CO)5] was used in the presence of sodium acetate over 5 h reaction time, allowing the synthesis of twelve allylic alcohols with moderate to excellent yields (54–97%) (Scheme 27).
With a view to synthesizing spiroketals, orthoesters, and orthocarbonates, Lacour and co-workers [91] resorted to ruthenium catalysis, thanks to the combined use of the salt [CpRu(CH3CN)3][BArF] (2.5 mol%) and 1,10-phenanthroline (2.5 mol%). For the synthesis of orthocarbonates 48 (27–72% yields), decomposition of α-diazo-β-ketoesters 49 (3.0 equivalent) in the presence of 1 was performed in dichloromethane for 3 days at 60 °C (Scheme 28). The lower reactivity of cyclic carbonates 1 when compared to cyclic ketones and lactones (3–24 h) led the authors to suggest the occurrence of a stronger complexation of the catalyst [CpRu(phen)][BArF] with 1, which would slow down the catalytic cycle due to the slower release of the complex [CpRu(phen)][BArF]. Moreover, compared to ketones, cyclic carbonates display reduced nucleophilicity in the capture of the transient electrophilic carbene.
Another example of ruthenium catalysis was described by Guillaume, Carpentier, and co-workers [92] to obtain α- and α,ω-telechelic cyclocarbonate polycyclooctenes 50 and 51, macromolecules having two reactive end groups that are able to react selectively with other molecules to create new bonds [93]. The high-yielding reaction between cyclooctene 52 and VEC 1e or acryloyl derivative 1k was promoted by a Grubbs 2nd generation ruthenium catalyst in THF at 40 °C (Scheme 29). However, the authors did not define the polymerization degrees. The polymers characterized by 1H NMR are important as a starting backbone for the preparation of new materials for applications in different fronts, such as surfactants or NIPUs.
Silylating agents are broadly used to insert different functional groups on the surfaces of organic or inorganic materials. In this line, the synthesis of functionalized symmetrical and unsymmetrical disilazanes was reported with excellent yields (81–95%) by using the Karstedt catalyst Pt2[(CH2=CHSi(CH3)2)2O]3 [94]. Kucinski and co-workers reported the reaction between olefins and the 1,1,3,3-tetramethyldisilazane 53 in toluene at 60–110 °C over 3–24 h (Scheme 30). Among the olefins involved, when VEC 1e was reacted with 53, the bifunctional unsymmetrical disilazane 54 was obtained with 85% yield after only 5 h, thus showing the relevant tolerance of the cyclic carbonate unit.
To perform the C-H allylation of indoles by insertion of VEC 1e, Yu and co-workers [95] described the use of mechanochemistry, cobalt(III) catalyst [Cp*Co(MeCN)3][SbF6]2, silver acetate, and silica gel (Scheme 31). This reaction allowed the functionalization of fifteen indole derivatives with 1e under solvent-free conditions, delivering the desired products 55 in modest to high yields (28–98%). It is worth highlighting that the reactions were run for only 2 h using a ball mill (800 rpm), which characterizes this method as an environmentally friendly protocol.
In 2012, Bensemhoun and Condon [96] reported the first creation of carbon–carbon bonds with glycerol 1,2-carbonate (Scheme 32). They proposed conjugate addition reactions of activated glycerol 1,2-carbonate derivatives 1 onto electron-deficient olefins 57 based on an electrochemical process using nickel complexes as catalysts in conjunction with a consumable anode. Halogen (I, Br) or pseudohalide (N3, OTs) derivatives 1 and certain electron-deficient olefins 57 such as methyl vinyl ketone, butyl acrylate, and acrylonitrile were used as substrates and the reactions were conducted in an undivided cell fitted with a nickel grid as the cathode and an iron rod as the anode. The first step of the reaction was a short pre-electrolysis process at room temperature; after 15 min, the starting materials and NiBr2·3H2O were added to the system for the electrolysis at 76 °C under a constant 0.2 A current, until full consumption of the GC. The regioselective conjugate addition was effective and provided the new carbonates in good yields. This sacrificial anode-based electroreductive process is compatible with the presence of the carbonate function. A comparison of product yields using DMF or propylene carbonate (PC) as a green solvent gave good results in both cases.
In 2020, Glorius and co-workers [97] described the synthesis of α,β-disubstituted γ-butyrolactones 58 through cooperative iridium catalysis and N-heterocyclic carbene organocatalysis (Scheme 33). Considering the importance and prevalence of γ-butyrolactones in biologically active natural products, in this study the authors prepared several compounds 58 with moderate to excellent yields, with control of the relative and absolute configuration of the two formed stereocenters, using an enantio- and diastereodivergent [3 + 2] annulation reaction. These lactones were prepared from VEC 1e and trans-cinnamaldehydes 59 using a solution of [Ir(COD)Cl]2 (2 mol%), Et3N, L4 (8 mol%), and 4C (10 mol%) in toluene (0.1 M) under Ar for 48–96 h. This combination of catalysts was responsible for the effectiveness and selectivity, with the Ir–π-allyl intermediates seeming well suited for nucleophilic addition of the NHC-enolate, giving the desired compounds. An extended study involving two chiral catalysts allowed selective access to all four possible stereoisomers of the γ-butyrolactone products. The usefulness of this strategy was illustrated in the synthesis of the naturally occurring lignan ()-hinokinin.

2.6. Other Reactions

Due to their favorable nutritional properties, 1,3-diglycerides, i.e., 1,3-sn-diacylglycerols 3a, are popular in the food industry. Interestingly, these diglycerides may be synthetized by an efficient and benign two-step, one-pot process, without any solvent (Scheme 34) [98]. A practical method used fatty acid anhydrides 60 and GC 1a under DABCO catalysis to produce the industrially relevant diglycerides 3a with good yields and high purity. Long-chain anhydrides 60 of lauric, myristic, palmitic, and stearic acids 61 led in all cases to the esterified carbonates, which were isolated in similar yields after 7 h reaction. Two plausible mechanisms were suggested for their transformation into 3a. First, the liberated fatty acid 61 reacts with DABCO to form an ammonium salt. The carboxylate anion can subsequently attack either the alkylene carbon (route 1) or the carbonyl group of the esterified carbonates (route 2), providing the desired 3a. The formation of the product was monitored by observing the liberation of CO2 and the disappearance of the carbonate IR band of the esterified carbonates.
Poly(ionic liquid)s (PILs) represent an emerging interdisciplinary topic due to their potential applications in chemistry, physics, materials science, and electrochemistry [99,100]. In this sense, the application of nonporous PIL beads with swelling properties as catalysts for the reaction of ethylene carbonate with aniline was proposed by Zhang and co-workers (Scheme 35) [101]. After preparation of cross-linked poly(ionic liquid)s by direct radical copolymerization of N-vinyl imidazolium ionic liquids with sodium acrylate, the authors used them in the conversion of various carbonates. Due to its swelling ability, P[VBIM][AA] presents catalytic activities similar to the homogeneous ionic liquid monomers. However, cyclic carbonates with a low dielectric constant, such as propylene- and 1,2-butylene carbonate, showed different swelling abilities, thus decreasing the catalytic activity and providing the corresponding oxazolidones 62 in lower yields (50% and 2%) than with ethylene carbonate 1 (99%).
With a view to avoiding purification steps, Truscello and co-workers [102] reported the one-pot synthesis of glycols 12 via in situ formation of glycerol carbonate 1a in the absence of any solvent. By reacting different phenols 7 with glycerol 3 (3 equivalent) and diethyl carbonate 16b (1.4 equivalent) under K2CO3 catalysis, glycols 12 were obtained with 72–82% yields (NMR evaluation) under conventional conditions (Scheme 36). 1-Naphthol was also successfully converted into the corresponding diol with 74% yield after 18 h.
The proposed method allowed seven aryloxypropanediols to be obtained, three of which were pharmaceutically important products, namely guaiphenesin, mephenesin, and chlorphenesin (Figure 5). To check the recyclability of excess glycerol and of the catalyst after the first run, mephenesin was extracted with toluene, then 1 equivalent of o-cresol, 1.4 equivalent of glycerol, and 1.4 equivalent of diethyl carbonate were added to the recovered residue. The resulting mixture was then stirred at 105–110 °C over 28 h. Although mephenesin could be obtained with good yields (76–80%) after three consecutive runs, the study of the re-use of the glycerol and base was not further developed.
A tetrabutylammonium-fluoride-mediated hydroxyalkylation process for phenols 7 with carbonate 1c was developed to access different aryl β-hydroxyethyl ethers 64 [103]. In this process, the corresponding products were obtained with 88–98% yields by using TBAF.3H2O (1 mol%) in DMF as the solvent at 170 °C, under N2 atmosphere over 13 min to 21 h of reaction (Scheme 37). Different substituted phenols were employed and good yields were obtained. The authors also extended this protocol to gram-scale and flow synthesis of multiflorol 64e, with 90% isolated yield.
Olsén and co-workers [104] reported the synthesis of densely functionalized carbamates 65 and 65′ through ring-opening of cyclic carbonates 14. The highlight of this study was the fact that unprotected α-amino acids, such as glycine 66a, were used in aqueous medium to provide a variety of new alkenes, carboxylic acids, thiols, and alcohols linked to a central carbamate motif (Scheme 38). The reactions were performed over 2 h at room temperature in the presence of triethylamine (2 equivalent). Optimization studies showed the necessity of 4 equivalent of the amino acid 66 to achieve opening of the cyclic carbonate ring. On reaction of the cyclic carbonates 14 with glycine 66a, the corresponding carbamates 65 and 65′ were obtained with 75–91% yields, while different amino acids 66b, containing one stereocenter, led to carbamates 65′’a-d with 94–97% conversion and 20–70% yields (NMR evaluation).
Recently, Luo and co-workers [105] reported the asymmetric allylic alkylation of α-branched β-ketocarbonyls, using an arene-coordinating chiral primary amine as a dual aminocatalyst and ligand. Different ketoesters 67 were reacted with VEC 1e to give the enantiomerically enriched allylic adducts 68 with good yields and excellent enantioselectivities (Scheme 39). To access the (S)-isomers, the reactions were performed in the presence of a bulky tertiary aminocatalyst (catalyst A)/TfOH (20 mol%), Pd(PPh3)4 (2.5 mol%), and (S)-2,2′-bis(diphenylphosphino)-1,1′-binaphthyl [(S)-BINAP] (6.25 mol%) in acetonitrile as the solvent, at 40 °C over 36 to 96 h. (R)-Isomers were produced from (catalyst B)/TfOH (20 mol%), Pd2dba3 (2.5 mol%), and tris(4-methoxyphenyl)phosphine [(PMP)3P] (10 mol%) in acetonitrile at room temperature for 16–72 h. Some of the vinylated carbonates presented better efficiency in providing the (R)-enantiomers (Scheme 40).

2.7. Alkene Polymerization

The synthesis of different monomers with at least two reactive groups for further polymerization is an attractive approach, especially from the point of view of the polymer industry. While one of the reactive groups can be used for copolymerization, the other one can be set aside for further transformations, namely for crosslinking [106]. In this sense, the presence of an unsaturated site on the monomer facilitates subsequent polymerization. On the other hand, the presence of a cyclic carbonate moiety as a second reactive site results in polymers with low reactivity, which allows better control of crosslinking compared with isocyanate or epoxide functions, for example. It is worth noting that the reactions on the double bond of the monomer, which can keep the cyclic carbonate unit intact, are critical for later functionalization of these monomers [19].
In this sense, Lapinte and co-workers [107] described the efficient synthesis of glycerol-based co-oligomers functionalized with alcohol, cyclocarbonate, or acetal pendant groups. The process involves free radical chain transfer polymerization using 2-mercaptoethanol (ME) as the chain transfer agent and glycerol-derived acrylate monomers. Initially, GC 1a was reacted with acryloyl chloride 69 in the presence of Et3N to provide glycerol carbonate acrylate (GCA) 1k with 80% yield (Scheme 41). The free radical telomerization of GCA to oligoGCAs was performed in the presence of AIBN and ME at 80 °C. During the telomerization process, the decomposition of AIBN causes the formation of a thiyl radical from ME, which further reacts with a monomer unit to form radical species, resulting in the product carrying a hydroxyl head group. A similar procedure was conducted using solketal instead of GC. Those materials did not exhibit branching or gelation and were totally soluble in the tested solvents. Additionally, the carbonate ring or acetal protective groups on oligomers can be selectively removed under basic or acidic conditions.
Seeking to obtain amphiphilic polymers, Robin and co-workers [108] have prepared thiol monomers 71 by the esterification of oleic and lauric acids with 2-mercaptoethanol. These intermediates were employed in AIBN-induced free-radical polymerization with glycerin carbonate acrylate (GCA) 1k in acetonitrile solution at 80 °C. In this telomerization process, two oleate polymers 72 were obtained with 65% (for n = 27) and 73% (for n = 65) yield, together with one laurate polymer with 62% yield (for n = 24), as shown in Scheme 42. The fatty acid fragment is responsible for the hydrophobic character of the polymer; in order to prepare amphiphilic derivatives 73 from telomers 72, the base-catalyzed hydrolysis of the cyclic carbonate unit was used to unmask hydroxyl groups. In aqueous media, the hydrolyzed polymers 73 were able to form aggregates with different numbers of micellar units, with critical micellization concentrations in the range of 10 to 60 mg·L−1. These values increase with the length of the hydrophilic moiety and decrease with the growth of the hydrophobic moiety length and the presence of double bonds, which leads to control of the diameter of the micelles.
In order to obtain fluorinated terpolymers 74 and copolymers 75, Carpentier and co-workers [109] reacted glycerol carbonate vinyl ether 1l with either chlorotrifluoroethylene 76 and vinyl ether 77 or with 76 alone, respectively, in the presence of tert-butylperoxypivalate (TBPPi) in 1,1,1,3,3-pentafluorobutane at 74 °C in an autoclave, in the presence of 3 mol% of potassium carbonate (Scheme 43). The combined presence of fluorine atoms and the GC unit in a polymer led to several possibilities for posterior functionalization. This versatility was explored by the authors via the synthesis of fluorinated polymers bearing hydroxyurethane moieties PFHU 78ab and 79ab through the ring-opening of the GC unit with isopropylamine in DCM at 50 °C. Complete conversion of the terpolymers 74 to PFHU 78ab derivatives was achieved using a large excess (20–30 equivalent) of base, while full conversion of 75 into 79ab only required 1.2 equivalent of base.
Patti and co-workers [110] promoted the reaction of the double bond of polystyrene polymers bearing pendant cyclic carbonate groups BMD 80a and BBD 80b via radical photopolymerization to develop new alternative materials for application in nanocomposites of sodium smectite, such as the one used in saline leachates, keeping the cyclic carbonate unit intact in the polymerization step. The synthesis of BMD 80a was performed via the reaction between GC 1a and 4-vinylbenzyl bromide 81a in the presence of sodium hydride in THF (Scheme 44). The analog BBD 80b was similarly synthesized by reacting 4-(4-hydroxybutyl)-1,3-dioxolan-2-one with 81a. The monomers 80a and 80b were photopolymerized over 24 h under nitrogen atmosphere using AIBN and a 12-cm-high UV lamp. The corresponding polymers were obtained with 93% (PBMD 82a) and 74% yields (PBBD 82b). On treatment of 82a with a saline solution (3 M NaCl), a nanocomposite with disordered intercalation was formed. Alternatively, 82a and 82b were intercalated in the presence of the same saline solution, providing partially disordered intercalates nanocomposites.
Plasseraud and co-workers [111] prepared monoglycerides 83 by reacting CG 1a or glycidol 2a with carboxylic acids 61. These intermediates 83 were converted into the corresponding monoglyceride bis-acrylates 84, which underwent polymerization under UV irradiation in the presence of 2-hydroxy-2-methyl-1-phenyl-propan-1-one (Darocur 1173®) at room temperature (Scheme 45). When GC 1a was employed (step 1), the desired product was isolated with 14% yield only (even after 24 h of reaction) due to side-reactions. In contrast, when glycidol 2a was used in excess (carboxylic acid 61:2a = 1:2), the intermediate 83 was obtained in yields above 80% after only 50 min.

2.8. Ring-Opening Polymerization

The class of five- and six-membered cyclic alkylene carbonates is large and varied, offering a wide range of monomeric systems for ring-opening polymerization (ROP), leading to new polymers [112,113,114]. Gnaneshwar and co-workers [115] reported the synthesis of a α-(cyclic carbonate), ω-Hydroxyl/Itaconic acid asymmetric telechelic poly(ε-caprolactone), by ROP followed by modification of the cyclic carbonates with 2-phenylethylamine (Scheme 46). Firstly, compound 86 was prepared with 96% yield by reacting GC 1a as a functional bio-based initiator with ε-caprolactone 27 in the presence of tin 2-ethylhexanoate (110 °C, argon flow). The authors reported that the ROP technique allows control of the molecular weights, as well as a linear correlation of the molecular weight vs. monomer conversion. The reactivity of compound 86 was demonstrated through the modification with 2-phenylethylamine to give hydroxylurethane 87 using a non-isocyanate route.
In 2016, Möller and co-workers developed an anionic ring-opening copolymerization technique for cyclic carbonate 1c and tert-butyl glycidyl ether 2c, followed by the removal of the protective group (Scheme 47) [116]. The hydroxy-functionalized PEG derivatives were submitted to sequential transcarbonylation, transalkylation, and fragmentation, aiming to replace carbonate groups with ether linkages. Poly(ethylene oxide-co-tert-butyl glycidyl ether)hydroxyl telechelic copolymer 88a was synthetized via the reaction between 1c and 2c under mono-mode microwave irradiation, using CsOH as an initiator. After 15 h at 180 °C, 1c was totally consumed, and the product 88a, with molecular weight values of 750–1700 g.mol−1, was obtained. Further, poly(ethylene oxide-co-glycidyl ether) 88b was obtained after the removal of the tert-butyl protective group, generating hydrophilic reactive hydroxyl sites in the copolymer.
The synthesis of a new polyether by cationic and anionic ROP of a cyclic carbonate-containing epoxide 1m under different conditions was achieved by Gnaneshwar and co-workers [115] to improve the hydraulic barrier properties of Na–smectite liners to saline leachates (Scheme 48). By using a strong non-nucleophilic base, it was possible to perform concomitant ring-opening of both the epoxide and the cyclic carbonate, providing the respective poly[oxy(ethane-1,2-diyl)] 89 with 87% yield. In cationic conditions, poly[-oxy(ethane-1-{[(methyloxy)methyl]oxirane}-1,2-diyl)] 90 was obtained with 52% yield by selective ring-opening of the cyclic carbonate functionality. In contrast, the treatment under nitrogen of the monomer with boron trifluoride resulted in ring-opening of the epoxide to give the poly(oxy{-ethane-1-[(4-methoxymethyl)-1,3-dioxolan-2-one]-1,2-diyl}) 91 with 63% yield. Thereafter, nanocomposites of sodium smectite were formed from those polymers via clay polymerization in situ and solution intercalation methods. The resulting nanocomposite 91 was shown to be more resistant to leaching in 3M NaCl than its respective cyclic carbonate 1m.
In 2017, Marcos-Fernández and co-workers [117] described the degradation of poly(ethylene terephthalate) by ethylene carbonate (EC) in the presence of potassium hydroxide, a low-cost catalyst. Degradation of PET at 130 °C produced short-chain oligomers with approximately 33% of carbonate units inserted in their structure, resulting from the ring-opening of EC with partial decarboxylation. Furthermore, the reaction of bis(2-hydroxyethyl) terephthalate (BHET) with EC in the absence of KOH produced difunctional oligomers with hydroxyl terminal groups, bearing aromatic rings esterified with polyether short chains. Generally, such oligomers could be applied as macroglycols for the preparation of thermoplastic polyurethanes.

3. Potential Applications in Biological Studies

3.1. Raw Material Derivatives

A raw material is a basic material that is used for the generation of a new material with greater application potential. Currently, glycerol represents one of the most versatile and available raw materials due to the large production of biodiesel, from which large tonnages of glycerol co-product are obtained [118]. Another highly abundant raw material is xylan, a major biopolymer associated with plant hemicelluloses [119]. Taking into account their abundance, it is interesting to explore ecologically sound applications for these raw materials.
In 2016, two methodologies describing the functionalization of xylan through reaction with cyclic carbonate 1e were reported. Saake and co-workers [120] first described the functionalization of xylan 92 to deliver modified polysaccharides 93 (Scheme 49A) through a DBU-catalyzed reaction in DMSO at 140 °C. Two different xylans were evaluated, with a special focus on the effects of the reaction time, temperature, and amount of xylan in the reaction mixture. Samples were analyzed mainly using NMR spectroscopy regarding their degree of substitution (DS) and molar substitution (MS). The products were further characterized by FTIR spectroscopy, mass spectrometry, and SEC, and also regarding their water solubility.
A few months later, the same group [121] reported that differences in temperature and time were determining factors in the synthesis of xylan derivatives. Considering the decarboxylation process that occurs in the reaction medium, the authors observed a significant difference between reactions performed at 140 °C and at 160 °C. When the reaction proceeds at 140 °C, derivatives with low substitution levels are preferably obtained, with longer reaction times being required to obtain highly substituted xylanes 92. Moreover, as DMSO undergoes degradation at high temperatures, a reaction proceeding at 140 °C is preferred. The remaining hydroxyl group on the hydroxyvinylethyl appendage of xylan 93 then reacts with another vinyl carbonate molecule, leading to product 94 (Scheme 49B).
In 2016, Oliveira and co-workers [122] used two renewable materials for the synthesis of the glyceryl glycoside 95. The Ferrier allylic rearrangement involving GC 1a and tri-O-acetyl-D-glucal 96 provided a mixture of diastereomers 95 with 31% yield, together with degradation products (Scheme 50). Considering their possible effects on the reaction, both the reaction time and the quantity of Lewis acid were reduced, resulting in a modest increase in the yield to 40%. Finally, an optimized set of mild conditions was established—by stirring a mixture of glucal, GC 1a (1.2 equivalent), BF3.Et2O (0.1 equivalent). and 4Å molecular sieves in CH2Cl2 at −40 to −20 °C for 10 min provided glycoside 95 with 84% yield (diastereoisomeric mixture) after flash column chromatography. This protocol was used to prepare the starting material for a CuAAC reaction to induce the formation of glucoglycerol-1,2,3-triazole derivatives.

3.2. Heterocyclic Compounds

Heterocyclic systems represent the most general structural unit of many natural compounds; they have valuable biologically properties and a wide range of applications in the pharmaceutical industry [123,124]. Oxa- and aza-heterocycles atoms are the most representative, which have been used to access new compounds in chemical industries [125]. In the following sections, we mention studies from recent years that have shown efficient synthetic methodologies for the functionalization of heterocycles via the insertion of different substituents in the chemical structure.
The oxazolidin-2-ones 62 were prepared through the reaction between cyclic carbonates 1 and aryl amines 63 in the presence of 1-butyl-3-methylimidazolium acetate ([bmim]OAc) as an eco-efficient catalyst (Scheme 51) [126]. NMR spectroscopy and DFT calculations demonstrated that both the anion and the cation of the ionic liquid catalyst cooperatively activate the starting materials via formation of hydrogen bonds. Thus, several compounds 62 were obtained with good to excellent yields after 9 h at 130 °C. Ethylene-, propylene-, and 4-phenoxymethylethylene carbonates 1 were reacted with a range of arylamines 63. The carbonylation reaction using different cyclic carbonates showed that steric hindrance negatively influenced the reactivity in the following order: ethylene carbonate > propylene carbonate > 4-phenoxymethylethylene carbonate.
In the same work, a [bmim]OAc-catalyzed synthesis approach for aryl bis-oxazolidin-2-ones 97 from cyclic carbonates 1 and aryl- or heteroaryldiamines 98 was devised. Ethylene- and propylene carbonate smoothly reacted with aromatic diamines to form the corresponding bis-oxazolidin-2-ones 97 in good yields (Scheme 52). In addition, the imidazolium-based ionic liquids also exhibited high catalytic activity to prepare 3-aryl-[1,3]-oxazinan-2-ones from trimethylene carbonate and [1,3]-dithiolan-2-ylidene-arylamines from ethylene trithiocarbonate [126].
Due to the versatility of glycerol carbonates, several investigations have been conducted, aiming to develop protocols for the introduction of the GC unit in heterocyclic systems. In 2012, Sackus and co-workers [127] reported the functionalization of 1-phenyl-1H-pyrazol-3-ol 99 with TGC 1h. The synthesis of phenylpyrazole derivatives has attracted the attention of synthetic organic chemists because of their broad spectrum of biological activities. In this protocol, 1h reacts with compound 99 in DMF at room temperature in the presence of K2CO3 to provide chemoselectively for 4-{[(1-phenyl-1H-pyrazol-3-yl)oxy]methyl}-1,3-dioxolan-2-one 100 with 71% yield (Scheme 53).
The same group [14] reported the use of 1h as a bis-electrophile for the N-glyceryl functionalization of diverse aza-aromatic systems 101106 (Scheme 54). The ambident benzimidazole 101 was converted into the carbonate 107 with 76% yield after 5 h, whereas under the same conditions, benzotriazole 102 provided a mixture of 1- and 2-substituted regioisomers (85% global yield). For less acidic aza-heterocyclic systems, a decrease in the yield was observed, e.g., 1H-indole 108 led to the N-alkylated compound 109 with only 34% yield. Comparatively, methyl 1H-indole-3-carboxylate 104, with the N-H bond under the influence of the electron-withdrawing group, was more reactive than unsubstituted indoles, providing the expected product with 48% yield. Carbazole 105 reacted in a solvent-dependent way to provides mixtures with 60% yield for N-alkylated product 110 and the isomeric glycidyl carbamate. When reacting 1,2,3,4-tetrahydrocyclopenta[b]indole as the substrate in DMF, the corresponding glycidyl carbamate was obtained exclusively with 51% yield.
Various chemistshave evaluated the performance of benzimidazole-derived carbonates as building blocks for useful transformations. The synthetic potential of the prepared compounds was evaluated in reductive or hydrolytic conditions, involving selective carbonate ring-cleavage using diverse nucleophiles (Figure 6) [14].
A new series of N-phenyl-4,5-dibromopyrrolamides and N-phenylindolamides was synthetized and a selected set of compounds was evaluated against the enzyme DNA gyrase from Escherichia coli and Staphylococcus aureus [128]. Zidar and co-workers used ethylene carbonate to prepare N-aryloxazolidin-2-ones according to Scheme 55. Initially, 3-(4-nitrophenyl)oxazolidin-2-one 62a was obtained with 81% yield by reacting 4-nitroaniline 63c with ethylene carbonate 1c at 100 °C in the presence of DBU. Subsequent reduction of the nitro group of 62a provided the corresponding arylamine, which was further coupled with 4,5-dibromo-pyrrole-2-carboxylic acid or indole-2-carboxylic acid to give the target arylcarboxamides 115 and 116, respectively. However, these compounds showed weaker activities against both E. coli and S. aureus DNA gyrase compared to the reference molecule 2-{[4-(4,5-dibromo-1H-pyrrole-2-carboxamido)phenyl]amino}-2-oxoacetic acid.
In 2016, Kim and co-workers [129] described a rhodium(III)-catalyzed, heteroatom-directed C-H allylation process with allylic phosphonates and allylic carbonates. Specifically, the mild and site-selective C-H allylation of 2-arylbenzo[d]thiazole 117 with vinyl carbonate 1e provided the allylic alcohol 118 (90% yield) with an E/Z ratio of 7.5:1 (Scheme 56). This protocol was suitable for a wide range of substrates and may be applicable for accessing new bioactive heterocyclic compounds.
A one-pot synthesis method for new N-3-substituted 2-thiohydantoins 119 was reported starting from GC (Scheme 57) [130]. In this study, a tandem Staudinger reaction was devised to convert primary azides into isothiocyanates, which were further reacted with α-amino esters. Namely, glycerol carbonate azide 1n led to isothiocyanate 1o, which was reacted without isolation with α-amino ester 120 to provide the N-3-glycerylated compound 119 with 68% overall yield.
In 2017, Zhao and co-workers [131] described the construction of nine-membered heterocycles via a [5+4] cycloaddition reaction between N-tosyl azadienes 31 and substituted vinylethylene carbonates 6 under Pd catalysis, involving highly diastereoselective functionalization (Scheme 58). The benzofuran-fused heterocycles 30 were isolated with 85–92% yields after 15 h of reaction. This strategy was applied with high efficiency to various carbonates and azadienes bearing para-, meta-, or ortho- substituents at the aryl ring with both electron-withdrawing and electron-donating characters. When the unsubstituted vinyl carbonate 1e was reacted with various N-tosyl azadienes 31, the kinetically preferred five-membered ring structure was formed through [3+2] cyclization, and the spirocyclic products 121 were isolated in good yields and with high diastereoselectivity.
The same group [132] reported the synthesis of [5,5]- and [6,5]-spiro-heterocycles 123 and 124 through the reaction of aurones 122 with vinylethylene carbonates 6 under Lewis-acid-assisted Pd catalysis (Scheme 59). Using a simple approach and using commercially available catalysts, the scope of the [3+2] cycloaddition was extended to several enones 122 and carbonates 6 bearing electron-donating or electron-withdrawing substituents at either para- or ortho-positions on the aryl group, as well as to naphthyl- and furyl derivatives. Under the optimized conditions, substituted [5,5] products 123 were obtained in high yields, except when alkyl vinylethylene carbonates 6 were used. Additionally, the protocol was applied in [4+2] cycloadditions, targeting the formation of [6,5] spirocyclic structures 124. A range of substrates bearing different functionalities was examined, as well as other thia- or aza-heterocyclic-based enones. Generally, the expected products were obtained in very good yields and uniformly excellent diastereoselectivity (>20:1) was observed for all prepared compounds.
Guo and co-workers developed an efficient Pd-catalyzed formal [5+3] cycloaddition process to prepare N,O-containing eight-membered heterocycles 126 (Scheme 60) [133]. In this work, zwitterionic allylpalladium intermediates were generated in situ from vinylethylene carbonates 6, which acted as 1,3-dipolarophiles in the cycloaddition. Under this Pd-catalyzed protocol, the reaction between N-iminoquinazolinium ylides 125 and carbonates 6 provided tricyclic compounds 126 in high yields with excellent regioselectivities. The optimal conditions were applied to alkyl- or aryl-substituted vinylethylene carbonates and different azomethine imines, including N-quinazolinium and N-isoquinolinium ylides, providing the expected [5+3] cycloadducts. The reaction mechanism involves the formation of the zwitterionic allylpalladium intermediate A, which adds to the azomethine imine. The allylpalladium intermediate B formed subsequently undergoes N-alkylation, leading to the [5+3] cycloaddition product 126 (Scheme 60).
Ackermann and co-workers developed an air- and water-tolerant manganese(I) catalyst to perform the C-H allylation of heteroarenes 127 and arenes 128 with vinyl carbonates 14 through “in-water” decarboxylative C-H/C-O cleavages (Scheme 61) [134]. In a typical procedure, aromatic compounds were reacted at 100 °C for 16 h with an excess of carbonate 14 in the presence of MnBr(CO)5 (10 mol%) and NaOAc (20 mol%), using trifluoroethanol (TFE) or water as the solvent. C-H Allylations of indoles 127, for instance, proceeded with good yields and high levels of chemo-, site-, and regioselectivity.
In 2017, Wang and co-workers [135] explored the domino C−H/N−H allylation of arylimidates under cobalt(III) catalysis (Scheme 62). In this work, the combined C−H and N−H allylation was reported for the one-step synthesis of isoquinolines. The reaction between different vinyl carbonates 14 and substituted arylimidates 131 delivered the desired decorated vinyl isoquinolines 132 in good yields. The protocol was extended to arylimidates 131 bearing electron-donating or electron-withdrawing groups, such as fluorine, chlorine, ester, ketone, and amide groups. The method was also suitable for carbonates 14 bearing aryl groups.
In 2019, He and co-workers reported the synthesis of oxazolidinones 133 through the carboxylative cyclization of anilines 63 and cyclic carbonates 1 (Scheme 63) [136]. The protic ionic liquid 1,8-diazobicyclo[5.4.0]-7-undecenium imidazoline [HDBU][Im] was used as a bifunctional catalyst under mild reaction conditions and without the addition of conventional solvents. Using this new protocol, a total of seventeen oxazolidinones 133 with different functional groups were obtained, such as Cl, Br, CH3, OCH3, and NO2, with yields in the range of 61–92% after 6 h of reaction. Some of the advantages of this protocol are the low catalyst load, the wide functionality tolerance, and the convenient recycling of the catalyst.
In the same year, Ji and co-workers described the synthesis of a new heterogeneous catalyst, the ionic liquid 1-butyl-3-methylimidazolium acetate ([bmim][OAc]) immobilized on MIL-101-NH2, denoted as IL(OAc)-MIL-101-NH2. This catalyst was used in the reaction between phenylamine 63 and propylene carbonate 1b to generate the oxazolidinones of interest 133 (Scheme 64) [137]. The method was extended to several anilines, allowing the synthesis of ten new oxazolidinones 133 with yields that varied from 68 to 94% after 9 h of reaction.

3.3. Bioactive Urethanes and Derivatives

Carbamates or urethanes have attracted much attention due to their biological relevance as mimic fragments of pharmaceutical compounds, such as efavirenz [138] and retigabine (Figure 7) [139]. In other respects, urethanes are monomers that are largely used in the synthesis of polyurethanes, one of the most used raw materials in global industry [10]. Polyurethanes are also versatile thermoplastic elastomers employed in the preparation of biocompatible components, including artificial hearts, feeding tubes, catheters, surgical drains, and others [140].
In 2012, Endo and co-workers [141] reported the synthesis of a branched cationic polyurethane 134 via the polycondensation of 4-chloromethyl-1,3-dioxolan-2-one 1o with diethylenetriamine (DETA) 63d in ionic liquid as the solvent (Scheme 65). The reaction of 1o and DETA 63d was achieved using a 2:1 molar ratio for the monomers. When carried out in molten imidazolium salts, the polymerization yields were significantly higher than those obtained using conventional solvents. The authors reported that due to its ammonium group, the polymer can serve as a catalyst and a capsule for DNA, whereby it can be used as a gene delivery carrier. The complexation with DNA was confirmed by dynamic light scattering (DLS) experiments.
By reacting primary amines 63e, alkyl halides, and diverse carbonates 1 bearing cationic, hydrophobic, or amphiphilic groups, Möller and co-workers [142] prepared poly(ethyleneimine)s (PEI) and amphiphilic monodisperse compounds (Scheme 66). These compounds were assayed for their ability to inhibit the proliferation of microorganisms (E. coli, S. aureus, and B. subtilis) and for their hemolytic activity, which was influenced by the microstructure and length of the PEI alkyl chain. The results led to the conclusion that the microstructure of the amphiphilic monodisperse compounds has a significant influence on their antibacterial properties and the hemolytic activity. Compounds with alkyl chains directly linked to cationic groups have high antibacterial efficiency for long (C-14 to C-18) alkyl chains.

3.4. Other Compounds with Biological Activity

Several reactions have been explored for the synthesis of glycerol carbonate derivatives with interesting biological activities. In this sense, Sebastiano and co-workers [143] investigated the condensation of ferrocenemethanol 139 with glycerol derivatives and related amines in the absence of solvent and catalyst. The reactions were performed under neat homogenous conditions and eleven novel compounds were described. Among the tested alcohols, GC 1a was reacted with 139 in the presence of catalytic CO2 or without any catalyst, leading to the corresponding ferrocenylmethyl ether 140 with 92% and 97% yields, respectively (Scheme 67). Ferrocenylmethyl ether 140 three other ferrocenyl analogues were tested in vitro as fungicides and were shown to inhibit fungi growth against Botrytis cinerea and Penicillium spp.
Recently, Zhang and co-workers [144] described the copper-catalyzed enantioselective synthesis of highly substituted β-amino alcohols 141 starting from ethynylated carbonates 6 and arylamines 63. The reaction was performed at −20 °C in toluene and in the presence of triethylamine and (S,S)-L as a copper ligand (Scheme 68). Using this procedure, fourteen novel β-amino-β-ethynyl alcohols 141 were obtained with 65–99% yields. When reacting ethynylated carbonates bearing naphthyl or heterocyclic groups under the optimal conditions, the efficacy of the method was maintained, giving the desired products with 93% and 98% yields, respectively. In contrast, the only aliphatic amine tested, tbutylamine, yielded the respective β-amino alcohol with only 75% yield. The reaction enantioselectivity (determined by HPLC) presented er ratios varying from 77:23 to 94:6.
In order to illustrate the potential usefulness of the β-amino-β-ethynyl alcohols 141, the authors tested several reactions, such as hydrogenation or Sonogashira coupling. They also performed a click reaction between 141a and antiretroviral Zidovudine 142 to efficiently prepare a triazolyl derivative of 142 (85% yield; 92:8 er) (Scheme 69). Zidovudine, a FDA-approved drug used for preventing and treating AIDS, is included in the World Health Organization’s List of Essential Medicines [145].

4. Precursors to Materials

4.1. Urethanes and Polyurethanes

As already mentioned in Section 2.3, polyurethanes (PUs) are a valuable group of materials that have applications in a range of industrial segments, in competition with other plastics, ceramics, metals, and rubbers, both in domestic and industrial fields [146]. Due to their multiple applications and high global consumption, the production of polyurethanes (24 Mtons by 2020) represents around 6% of all the plastics produced every year. Many different kinds of polyurethanes, such as flexible and rigid foams, elastomers, adhesives, and coatings, are broadly manufactured and applied, namely as rollers, shoe soles, seals, and engineering components for electrical encapsulation and in the mining industry [147,148]. They also have important applicability as components of medical implants in surgery [146].
The traditional route for the synthesis of PUs involves the reaction of isocyanates with alcohols (Scheme 70) [147]. Despite the versatility of isocyanates as building blocks, these compounds are carcinogenic, mutagenic, and reprotoxic and are classified as “very harmful” chemicals by the European Union. Additionally, the synthesis of isocyanates themselves involves a reaction between the lethal phosgene gas and an amine salt [148]. For these reasons, the development of new routes to access PUs involving less toxic reagents became synonymous with the synthesis of non-isocyanate polyurethanes. The main alternative methods developed so far are (i) the cyclic carbonate pathway (polyaddition), (ii) transurethanization (polycondensation), and (iii) ring-opening polymerization and rearrangements [146,147,149,150]. In particular, the synthesis of PUs via the polyaddition of amines on cyclic carbonates has been explored in recent years as an environmentally friendly route (Scheme 70).
The studies published between 2012 and 2020 on the isocyanate-free construction of the urethane nucleus (the carbamate function) can be grouped according to the structure of the cyclic carbonate reacting with the amine: glycerol carbonate 1a, polycyclic carbonate (poly-1), ethylene carbonate 1c, and ethylene carbonate derivatives (deriv-1) (Scheme 71).
The reported studies include the preparation of polymeric and non-polymeric materials, for which the synthesis has been described [151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166], as well as studies on the properties [167,168,169,170] and structural characterization [171,172,173,174,175]. With regard to the materials prepared from cyclic carbonate building blocks 1, the properties observed so far include thermal stability [176], stiffness [177], elongation and tensile strength [178], rheological qualities [179,180,181], resistance to corrosion [182], water absorption [183], and optical resistance [184]. These diverse properties are compatible with applications in interesting fields, such as adhesives [185,186,187,188].

4.2. Flame Retardants

Reducing the occurrence and severity of fire is a major issue, which has fostered the development of new products able to deter or stop flame propagation. Flame retardants are chemicals added into a combustible material in order to increase the ignition temperature, reduce the rate of burning, limit the spread of flames, and minimize, if not eliminate, the generation of smoke. The most important active species in flame-retardant families are halogens (chlorine and bromine), phosphorus, and water [189,190].
In 2014, Morgan and co-workers [191] evaluated the heat release of polyurethanes modified with boron- and phosphorus-containing flame retardants using pyrolysis combustion flow calorimetry (PCFC). The observed results demonstrated that both the flame retardant’s structure and its environment in the polymer greatly affect the flammability behavior. Among the tested potential flame retardants, the phosphorylated glycerol carbonate 145 (Scheme 72) did not inhibit melt flow in the polyurethane during decomposition, thus reflecting the low flame retardant potential of 145 when incorporated in a PU.
In a subsequent study, the same group [192] reported the synthesis and flammability testing of a new class of phosphorus-containing flame-retardant molecules incorporated into a PU. A new mixed phosphate ester 148 was prepared in two steps, by first converting GC 1a into its dichlorophosphate 147, which was further reacted with glycidol 2a (2 equivalent) to give 148, containing two epoxide units and a glycerol carbonate moiety (Scheme 73). The fire-retardant performance of 148 as an additive was evaluated using PCFC, showing middling levels of flame-retardant potency.
In 2016, Sonnenschein and co-workers reported the alkoxylation of bisphenols S and A by propylene carbonate to prepare new hydroxylated polymerizable building blocks 149–159 (Scheme 74) [193]. The thermo-oxidative evaluation of these useful low molecular weight molecules showed significantly better thermo-oxidative stability for sulfone polyols compared to other bisphenol-A derivatives, as measured by gas evolution and pyrolysis GC–mass spectrometry techniques. Moreover, bisphenol-S polyols possess higher viscosities than their non-sulfone counterparts of similar molecular weight—a property which is presumably linked to the increased interchain hydrogen bonding potential provided by the sulfone moiety.

4.3. Energy and Electronics

For further development of lithium-ion batteries, it is essential to prepare less flammable and easily processable materials. In this sense, Rieger and co-workers [194] studied the structure and ionic conductivity of liquid crystals containing propylene carbonate units (Scheme 75 and Scheme 76). The interest in this class of compounds is due to their thermal stability, high dielectric constant, and good lithium ion conductivity at room temperature [195]. The authors prepared type 160 liquid crystalline molecules bearing a cyclic carbonate endpoint (Scheme 75). The ionic conductivity of the tetrafluorinated compound 160a was compared with that of the non-fluorinated analogue 161a (Scheme 76), while 160b was compared to the tetra(ethylene oxide) (TEO) analogue 162a (Scheme 75), which bears the same fluorinated mesogenic core. The fluorinated precursor 163 was obtained in 2 steps from 2,3,5,6-tetrafluorophenol, while the non-fluorinated precursor 164a was prepared from 4-hydroxybenzoic acid. Liquid crystals 160a and 161a were obtained with satisfactory yields via the esterification with GC 1a under mild reaction conditions. The TEO analogue 162 was synthesized with 65% yield via a sequential BH3-THF reduction of 163c, followed by the Cs2CO3-catalyzed etherification of the benzylic alcohol 166 using the TEO tosylate 167a.
After mixing the synthesized compounds with lithium bis(trifluoro-methanesulfonyl)imide (LiTFSI), the authors observed that Li+ self-assembly was driven by the interaction with the carbonate units via coordination with the free electron pairs of oxygen. Furthermore, the cyclic carbonate moiety present in 160 and 161 enhanced the lithium salt dissociation as compared to the TEO derivative 162. Measurements by small- and wide-angle X-ray scattering (SAXS/WAXS) demonstrated that the compounds align spontaneously in the smectic phase when cooled from the isotropic melt. The formation of a smectic layer improves the ionic conductivity by enhancing the ion mobility parallel to the layer plane.
Another interesting study involving GC 1a was reported by Kim and co-workers in 2016 [196]. In this work, 1a was reacted with furfuryl amine 63g to provide the urethanes 168ab (Scheme 77). After a few more reaction steps, a surface of polydimethylsiloxane-based polyurethane cross-linked by Diels-Alder adducts 169 was prepared. In the sequence, silver nanowire (AgNW) was deposited on this surface via vapor-assisted solution process, resulting in a highly flexible, transparent, and healable electrode.
In 2020, Cui and co-workers [197] synthesized several fluorescent macrocycles 171 through the iridium-catalyzed intermolecular decarboxylative coupling of vinylated carbonate 1e with readily available isatoic anhydrides 170 (Scheme 78). The best reaction conditions were obtained with [Ir(cod)(OMe)]2 (2.0 mol%) as the catalyst, 2,2′-bipyridine as the chelating N,N-donor ligand, and a mixture DCM/acetone (3:1) as the solvent. The sequential allyl-amination–macrolactonization upon extrusion of CO2 provided the expected benzo-fused, 14-membered aza-macrolides 171 with moderate yields after 12 h of reaction. The scope was evaluated using isatoic anhydrides 170 bearing electron-donating (Me and CH3) and electron-withdrawing groups (halides and CF3). Azamacrolides 171 showed favorable fluorescence emission in the 397–430 nm range.
The proposed reaction mechanism starts with the Ir(I)-catalyzed decarboxylation of 1e to generate the zwitterionic p-allyl iridium intermediate A. After tautomerization of A, an electrophilic six-membered iridacyclic species B is formed. Then, the intermediate B is trapped by the nucleophilic amino group of 170a, giving the intermediate C and the active Ir(I) catalyst for the next catalytic cycle. Lastly, intermediate C undergoes intermolecular ester exchange with a second molecule of C to deliver product 171a after CO2 elimination (Scheme 79).
An online Raman analysis technique for monitoring of the production of biofuels was developed to study the transesterification reaction of canola oil with dimethyl carbonate [198]. The reaction was monitored by measuring the height of the C=C stretching mode of the unsaturated fatty acids present in the oil. The transesterification of the triglycerides was catalyzed using a triazabicyclodecene-modified, Mg-Al layered double hydroxide (TBD-LDH). The methyl fatty esters 172 were obtained together with esterified glycerol carbonates, which further reacted with dimethyl carbonate to form additional 172 and the bis-carbonate 1p (Scheme 80).

4.4. Non-Ionic Surfactants

Surface-active agents (surfactants) are a class of compounds that present outstanding physical properties as solubilizers, with a wide range of application in the most varied industry sectors, for example being used in the soap industry, in cosmetics, and in the food area [199]. Among the range of surfactants, non-ionic structures are used in diverse areas, such as in cleaning products, textile processing, pulp and paper processing, emulsion polymerization, and agrochemicals. An important characteristic of these materials is the absence of a hydrophilic group with filler bound to their hydrophobic chain [200,201].
In this regard, Corma, Iborra, and co-workers [202] prepared glycerol-carbonate-derived fatty esters 1q using a hybrid Nafion–silica composite (Nafion SAC-13) as an acid catalyst in the absence of solvent [203]. In this work, the desired compounds were prepared through direct esterification of GC 1a with carboxylic acids of different chain lengths; the protocol led to the formation of several side-products, such as monoglycerides 12 and 173, as well as the triglyceride 174. Two different possible reaction routes were proposed by the authors:
GC 1a and the acid 61 firstly react under acid catalysis to give the GC carboxylate 1q; this intermediate is subsequently decarbonylated to form two isomeric monoglycerides, 12 and 173, which can further undergo stepwise esterification to provide diglycerides 173′ and 174″ and the triglyceride 174 (Scheme 81, route I);
Firstly, 1a undergoes hydrolysis to glycerol, which similarly reacts with 58 to yield monoglycerides 174 and 12, diglycerides 173′ and 173″, and triglyceride 174 (Scheme 81, route II).
Several parameters were evaluated in this reaction, such as the solvent, heterogeneous catalysts, zeolites, and hybrid organic–inorganic acids. The best results for the esterification of GC 1a were obtained by using 3.5 wt% of Nafion–silica (with respect to the total weight of the reactants) at 100 °C. The highest activity was exhibited by the Nafion SAC-13 nanocomposite, which proved to be an excellent esterification catalyst, while other ion exchange resins showed lower activity and selectivity. It has to be pointed out that when increasing the chain length in 61 from hexanoic to lauric and palmitic acids, the conversion rate decreased considerably from 98% to 71% and 31%, respectively.
Being interested in the synthesis of an important non-ionic surfactant, glycerol monostearate 12f (GMS), Han and Wang [204] devised a solvent-free esterification approach for GC 1a with stearic acid 61a catalyzed by copper 4-toluenesulfonate (CPTS) to prepare the intermediate (2-oxo-1,3-dioxolan-4-yl)methyl stearate (ODOMS) 1r. This compound presents good thermal and oxidative stability, in addition to surfactant properties. Using Et3N-catalyzed hydrolysis at 140 °C, ODOMS 1r was converted into GMS 12f with 64% yield (Scheme 82).

4.5. Miscellaneous

In addition to the plethora of chemical transformations presented in the previous sections in which glycerol carbonate is involved, the versatility of this glycerol derivative has allowed it to be employed in singular polymerization reactions, leading to diversified products with significant potential for application in various materials. In this line, the synthesis of new fluorinated terpolymers bearing cyclocarbonate pendant groups to ensure the crosslinking of the polymer chains was reported [205]. The new cyclocarbonate- and triazole-functionalized fluorinated terpolymer 176 was obtained in three steps from chlorotrifluoroethylene 76, as shown in Scheme 83. The glycerol carbonate vinyl ether monomer 1l was obtained with 45% yield via a transetherification reaction between GC 1a and ethoxyethylene 177 under palladium catalysis. The key co-polymerization step involving 1l, 76, and 2-chloroethyl vinyl ether 178 was catalyzed by nbutylperoxypivalate and proceeded in a HFP solution at 74 °C, leading to the terpolymer 179. After two more steps, the triazolyl-modified terpolymer 176 was obtained as a membrane, which was removable from the glassy substrate via immersion in 1M aqueous hydrochloric acid. After the final conditioning process, it was verified that the developed semi-interpenetrating polymer network (semi-IPN) blended membranes presented lower water uptake and better mechanical properties compared to non-crosslinked membranes.
In the same year, Sacripante and co-workers [206] described the preparation of polyester resins derived from rosin acids for application as a xerographic toner. Rosin acids obtained from pulp byproduct (tall oil), gum, or wood rosins are more sustainable monomers than those from food biomass. Reacting renewable dehydroabietic acid 180 with GC 1a (3.3 equivalent) in the presence of catalytic tetraethylammonium iodide at 150 °C provided a mixture of dehydroabietates 11e, 181, and 182. After adding more 180, the reaction was continued for 8 h at the same temperature to deliver a mixture of the monoesters 12e and 181 (61.9% yield) and 182 (12.5%), together with the isomeric diesters 183 (12.4%) and 184 (5.3%). This mixture of 182, 183, 184, and 12e rosin adducts was reacted with various organic diacids to provide novel polyester resins with a glass transition of 35–53.9 °C, which is the suitable temperature range for toner applications (Scheme 84).
A promising approach in green and sustainable chemistry is the direct use of CO2 by combining the coupling reaction and free-radical polymerization to efficiently produce polyacrylates bearing cyclic carbonate groups. In 2016, Zhang, Du. and co-workers [207] described the use of a new nano-lamellar double-metal cyanide complex, Zn-Co(III) DMCC/CTAB, a catalyst that allows the preparation of highly transparent polyacrylates with specific UV absorption. The reaction of CO2 with glycidyl methacrylate 185 followed by free radical polymerization (FRP) of the resulting alkenyl carbonate ester DOMA 1s provided new polyacrylates, known as PDOMAs (Scheme 85). AIBN (1.5 mol% related to 1s) was used as the radical initiator in the FRP, which was conducted in DMF at 80 °C. The resulting polymer, poly(2-oxo-1,3-dioxolane-4-yl)methyl methacrylate (PDOMA) 186, which is soluble in strong polar solvents, has tunable number-average molecular weights (12.0–132.0 kg·mol−1), high glass transition temperatures (121.0–140.4 °C), and a high thermal decomposition temperature (5 wt%, 257 °C). PDOMAs 186 presented excellent visible light transmittance (93%) and nearly 100% UV absorbance in the wavelength range of 200–313 nm.
In 2016, Carbonnier and co-workers [208] designed a synthetic route involving GC methacrylate 1s and glycol bis-methacrylate 189 for an FRP preparation of monolithic columns. Further surface modification with allylamine and thiols provided functionalized polymeric monoliths through a cyclic carbonate opening and thiol-ene click addition (Scheme 86). Subsequent possible immobilization of metallic nanoparticles allowed these materials to be successfully used as supports for chromatographic applications or as microreactors for flow catalysis.

5. Conclusions

The versatility of five-membered cyclic carbonates in organic synthesis and materials science has been recognized for years, as demonstrated in a number of papers. In this review, the studies reported from 2012 to the present were covered as an update of previously published contributions [5,15,16,17,18,19,20,21].
The cyclic carbonate nucleus has attracted considerable attention among researchers in different areas since it can be implemented as a substrate to access several classes of compounds with broad functional diversity. Our review covers miscellaneous uses of this structural unit as a starting material for the preparation of biologically active compounds, attractive molecules for the materials industry, and versatile building blocks in organic synthesis. In particular, the preparation of urethanes and polyurethanes should be highlighted as one of the most important applications of five-membered cyclic carbonates, since this represents an isocyanate-free and environmentally friendly route to these valuable polymers.
The possibility of using glycerol and CO2 (as demonstrated more recently) as readily available, cheap, renewable starting materials to prepare cyclic carbonates reinforces the role of such versatile motifs in the chemical industry. It is expected that investigating the reactivity and the physicochemical properties of cyclic carbonates will continue to attract the attention of synthetic and organic materials chemists for years to come. There are still some challenges to be overcome in the comprehensive development about cyclic carbonates, namely increases in the range and selectivity of the reactions involved in polymer synthesis and the greenness of most of the protocols engaged in the functionalization of this building block.
We hope the present review can awaken in the reader an interest in the development of new reactions and applications for the well-established range of cyclic carbonates, especially those derived from glycerol. This might well become a viable way to add value to glycerol issued from regrowing resources, with nobler applications for chemicals, pharmaceuticals, agrochemicals, and materials.

Author Contributions

Conceptualization, P.R. and G.P.; methodology, P.R., E.J.L., and G.P.; formal analysis, L.K.S., A.M.B., and D.R.A.; resources, E.J.L., R.G.J., and G.P.; writing—original draft preparation, L.K.S., A.M.B., and D.R.A.; writing—review and editing, P.R., E.J.L., R.G.J., and G.P.; visualization, P.R., E.J.L., R.G.J., and G.P.; supervision, P.R. and G.P.; project administration, G.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We are grateful to Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), and Fundação de Amparo à Pesquisa do Estado do Rio Grande do Sul (FAPERGS) for financial support. CNPq is also acknowledged for fellowships for G.P., R.G.J., and E.J.L. This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior Brasil (CAPES) Finance Code 001.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ricapito, N.G.; Ghobril, C.; Zhang, H.; Grinstaff, M.W.; Putnam, D. Synthetic Biomaterials from Metabolically Derived Synthons. Chem. Rev. 2016, 116, 2664–2704. [Google Scholar] [CrossRef] [PubMed]
  2. Guo, W.; Gómez, J.E.; Cristòfol, À.; Xie, J.; Kleij, A.W. Catalytic Transformations of Functionalized Cyclic Organic Carbonates. Angew. Chem. Int. Ed. 2018, 57, 13735–13747. [Google Scholar] [CrossRef]
  3. Schäffner, B.; Schäffner, F.; Verevkin, S.P.; Borner, A. Organic Carbonates as Solvents in Synthesis and Catalysis. Chem. Rev. 2010, 110, 4554–4581. [Google Scholar] [CrossRef] [PubMed]
  4. Gérardy, R.; Estager, J.; Luis, P.; Debecker, D.P.; Monbaliu, J.-C.M. Versatile and Scalable Synthesis of Cyclic Organic Carbonates under Organocatalytic Continuous Flow Conditions. Catal. Sci. Technol. 2019, 9, 6841–6851. [Google Scholar] [CrossRef]
  5. Pescarmona, P.P. Cyclic Carbonates Synthesised from CO2: Applications, Challenges and Recent Research Trends. Curr. Opin. Green Sustain. Chem. 2021, 29, 100457. [Google Scholar] [CrossRef]
  6. Marciniak, A.A.; Lamb, K.J.; Ozorio, L.P.; Mota, C.J.A.; North, M. Heterogeneous Catalysts for Cyclic Carbonate Synthesis from Carbon Dioxide and Epoxides. Curr. Opin. Green Sustain. Chem. 2020, 26, 100365. [Google Scholar] [CrossRef]
  7. Okoye, P.U.; Hammed, B.H. Review on Recent Progress in Catalytic Carboxylation and Acetylation of Glycerol as a Byproduct of Biodiesel Production. Renew. Sust. Energ. Rev. 2016, 53, 558–574. [Google Scholar] [CrossRef]
  8. Samoilov, V.O.; Ramazanov, D.N.; Nekhaev, A.I.; Maximov, A.L.; Bagdasarov, L.N. Heterogeneous Catalytic Conversion of Glycerol to Oxygenated Fuel Additives. Fuel 2016, 172, 310–319. [Google Scholar] [CrossRef]
  9. Rokicki, G.; Rakoczy, P.; Parzuchowski, P.; Sobiecki, M. Hyperbranched Aliphatic Polyethers Obtained from Environmentally Benign Monomer: Glycerol Carbonate. Green Chem. 2005, 7, 529–539. [Google Scholar] [CrossRef]
  10. Maisonneuve, L.; Lamarzelle, O.; Rix, E.; Grau, E.; Cramail, H. Isocyanate-Free Routes to Polyurethanes and Poly(hydroxy Urethane)s. Chem. Rev. 2015, 115, 12407–12439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Simão, A.-C.; Lynikaite-Pukleviciene, B.; Rousseau, C.; Tatibouët, A.; Cassel, S.; Šačkus, A.; Rauter, A.P.; Rollin, P. 1,2-Glycerol Carbonate: A Versatile Renewable Synthon. Lett. Org. Chem. 2006, 3, 744–748. [Google Scholar] [CrossRef]
  12. Rousseau, J.; Rousseau, C.; Lynikaitė, B.; Šačkus, A.; de Leon, C.; Rollin, P.; Tatibouët, A. Tosylated Glycerol Carbonate, a Versatile Bis-electrophile to Access New Functionalized Glycidol Derivatives. Tetrahedron 2009, 65, 8571–8581. [Google Scholar] [CrossRef]
  13. Balieu, S.; Zein, A.E.; Sousa, R.; Jèrôme, F.; Tatibouët, A.; Gatard, S.; Pouilloux, Y.; Barrault, J.; Rollin, P.; Bouquillon, S. One-Step Surface Decoration of Poly(propyleneimines) (PPIs) with the Glyceryl Moiety: New Way for Recycling Homogeneous Dendrimer-Based Catalysts. Adv. Synth. Catal. 2010, 352, 1826–1833. [Google Scholar] [CrossRef]
  14. Vilkauskaitė, G.; Krikštolaitytė, S.; Paliulis, O.; Rollin, P.; Tatibouët, A.; Šačkus, A. Use of Tosylated Glycerol Carbonate to Access N-Glycerylated Aza-aromatic Species. Tetrahedron 2013, 69, 3721–3727. [Google Scholar] [CrossRef]
  15. Pham, P.D.; Monge, S.; Lapinte, V.; Raoul, Y.; Robin, J.J. Various Radical Polymerizations of Glycerol-based Monomers. Eur. J. Lipid Sci. Technol. 2013, 115, 28–40. [Google Scholar] [CrossRef] [Green Version]
  16. Zhang, H.; Grinstaff, M.W. Recent Advances in Glycerol Polymers: Chemistry and Biomedical Applications. Macromol. Rapid Commun. 2014, 35, 1906–1924. [Google Scholar] [CrossRef]
  17. Ochoa-Gómez, J.R.; Gómez-Jiménez-Aberasturi, O.; Ramírez-López, C.; Belsué, M. A Brief Review on Industrial Alternatives for the Manufacturing of Glycerol Carbonate, a Green Chemical. Org. Process. Res. Dev. 2012, 16, 389–399. [Google Scholar] [CrossRef]
  18. Sonnati, M.O.; Amigoni, S.; Givenchy, E.P.T.; Darmanin, T.; Choulet, O.; Guittard, F. Glycerol Carbonate as a Versatile Building Block for Tomorrow: Synthesis, Reactivity, Properties and Applications. Green Chem. 2013, 15, 283–306. [Google Scholar] [CrossRef]
  19. Besse, V.; Camara, F.; Voirin, C.; Auvergne, R.; Caillol, S.; Boutevin, B. Synthesis and applications of unsaturated cyclocarbonates. Polym. Chem. 2013, 4, 4545–4561. [Google Scholar] [CrossRef]
  20. Duval, A.; Avérous, L. Cyclic Carbonates as Safe and Versatile Etherifying Reagents for the Functionalization of Lignins and Tannins. ACS Sustain. Chem. Eng. 2017, 5, 7334–7343. [Google Scholar] [CrossRef]
  21. Tabanelli, T.; Giliberti, C.; Mazzoni, R.; Cucciniello, R.; Cavani, F. An Innovative Synthesis Pathway to Benzodioxanes: The Peculiar Reactivity of Glycerol Carbonate and Catechol. Green Chem. 2019, 21, 329–338. [Google Scholar] [CrossRef]
  22. Pagliaro, M.; Rossi, M. The Future of Glycerol, New Usages for a Versatile Raw Material; RSC Green Chemistry Book Series; RSC Publishing: Cambridge, UK, 2010; p. 108. [Google Scholar]
  23. Jérôme, F.; Pouilloux, Y.; Barrault, J. Rational Design of Solid Catalysts for the Selective Use of Glycerol as a Natural Organic Building Block. ChemSusChem 2008, 1, 586–613. [Google Scholar] [CrossRef]
  24. Hutchings, G.J.; Lee, D.F.; Minihan, A.R. Epoxidation of Allyl Alcohol to Glycidol Using Titanium Silicalite TS-1: Effect of the Reaction Conditions and Catalyst Acidity. Catal. Lett. 1996, 39, 83–90. [Google Scholar] [CrossRef]
  25. Choi, J.S.; Simanjuntaka, F.S.H.; Oh, J.H.; Lee, K.I.; Lee, S.D.; Cheong, M.; Kim, H.S.; Lee, H. Ionic-liquid-catalyzed decarboxylation of glycerol carbonate to glycidol. J. Catal. 2013, 297, 248–255. [Google Scholar] [CrossRef]
  26. Niu, X.; Wang, L.; Cao, J.; Cao, Y.; He, P.; Zhou, J.; Li, H. Efficient Synthesis of Epoxybutane from Butanediol via a Two-step Process. RSC Adv. 2019, 9, 10072–10080. [Google Scholar] [CrossRef] [Green Version]
  27. Bolívar-Diaz, C.L.; Calvino-Casilda, V.; Rubio-Marcos, F.; Fernández, J.F.; Beñares, M.A. New Concepts for Process Intensification in the Conversion of Glycerol Carbonate to Glycidol. Appl. Catal. B Environ. 2013, 129, 575–579. [Google Scholar] [CrossRef]
  28. Ha, J.H.; Kim, J.S.; Kim, M.H.; Lee, K.Y.; Lee, M.S. Synthesis of Glycidol by Decarboxylation of Glycerol Carbonate Over Zn–La Catalysts with Different Molar Ratio. J. Nanosci. Nanotechnol. 2016, 16, 10898–10902. [Google Scholar] [CrossRef]
  29. Singh, D.; Reddy, B.; Ganesh, A.; Mahajani, S. Zinc/Lanthanum Mixed-Oxide Catalyst for the Synthesis of Glycerol Carbonate by Transesterification of Glycerol. Ind. Eng. Chem. Res. 2014, 53, 18786–18795. [Google Scholar] [CrossRef]
  30. Endah, Y.K.; Kim, M.S.; Choi, J.; Jae, J.; Lee, S.D.; Lee, H. Consecutive Carbonylation and Decarboxylation of Glycerol with Urea for the Synthesis of Glycidol via Glycerol Carbonate. Catal. Today 2017, 293–294, 136–141. [Google Scholar] [CrossRef]
  31. Zhu, J.-B.; Tang, X.; Falivene, L.; Caporaso, L.; Cavallo, L.; Chen, E.Y.-X. Organocatalytic Coupling of Bromo-Lactide with Cyclic Ethers and Carbonates to Chiral Bromo-Diesters: NHC or Anion Catalysis? ACS Catal. 2017, 7, 3929–3933. [Google Scholar] [CrossRef]
  32. Xie, J.; Guo, W.; Cai, A.; Escudero-Adán, E.C.; Kleij, A.W. Pd-Catalyzed Enantio- and Regioselective Formation of Allylic Aryl Ethers. Org. Lett. 2017, 19, 6388–6391. [Google Scholar] [CrossRef] [PubMed]
  33. Guo, W.; Kuniyil, R.; Gómez, J.E.; Maseras, F.; Kleij, A.W. A Domino Process toward Functionally Dense Quaternary Carbons through Pd-Catalyzed Decarboxylative C(sp3)–C(sp3) Bond Formation. J. Am. Chem. Soc. 2018, 140, 3981–3987. [Google Scholar] [CrossRef] [PubMed]
  34. Balaraman, E.; Fogler, E.; Milstein, D. Efficient Hydrogenation of Biomass-derived Cyclic Di-esters to 1,2-Diols. Chem. Commun. 2012, 48, 1111–1113. [Google Scholar] [CrossRef] [PubMed]
  35. Balaraman, E.; Gunanathan, C.; Zhang, J.; Shimon, L.J.W.; Milstein, D. Efficient Hydrogenation of Organic Carbonates, Carbamates and Formates Indicates Alternative Routes to Methanol Based on CO2 and CO. Nat. Chem. 2011, 3, 609–614. [Google Scholar] [CrossRef]
  36. Han, Z.; Rong, L.; Wu, J.; Zhang, L.; Wang, Z.; Ding, K. Catalytic Hydrogenation of Cyclic Carbonates: A Practical Approach from CO2 and Epoxides to Methanol and Diols. Angew. Chem. Int. Ed. 2012, 51, 13041–13045. [Google Scholar] [CrossRef]
  37. Wu, X.; Ji, L.; Ji, Y.; Elageed, E.H.M.; Gao, G. Hydrogenation of Ethylene Carbonate Catalyzed by Lutidine-bridged N-Heterocyclic Carbene Ligands and Ruthenium Precursors. Catal. Commun. 2016, 85, 57–60. [Google Scholar] [CrossRef]
  38. Chen, J.; Zhu, H.; Chen, J.; Le, Z.-G.; Tu, T. Synthesis, Characterization and Catalytic Application of Pyridine-Bridged N-Heterocyclic Carbene–Ruthenium Complexes in the Hydrogenation of Carbonates. Chem. Asian J. 2017, 12, 2809–2812. [Google Scholar] [CrossRef]
  39. Liu, H.; Huang, Z.; Han, Z.; Ding, K.; Liu, H.; Xia, C.; Chen, J. Efficient Production of Methanol and Diols via the Hydrogenation of Cyclic Carbonates using Copper–Silica Nanocomposite Catalysts. Green Chem. 2015, 17, 4281–4290. [Google Scholar] [CrossRef]
  40. Sun, J.; Yao, X.; Cheng, W.; Zhang, S. 1,3-Dimethylimidazolium-2-carboxylate: A Zwitterionic Salt for the Efficient Synthesis of Vicinal Diols from Cyclic Carbonates. Green Chem. 2014, 16, 3297–3304. [Google Scholar] [CrossRef]
  41. Li, F.; Wang, L.; Han, X.; He, P.; Cao, Y.; Li, H. Influence of Support on the Performance of Copper Catalysts for the Effective Hydrogenation of Ethylene Carbonate to Synthesize Ethylene Glycol and Methanol. RSC Adv. 2016, 6, 45894–45906. [Google Scholar] [CrossRef]
  42. Li, F.; Wang, L.; Han, X.; Cao, Y.; He, P.; Li, H. Selective Hydrogenation of Ethylene Carbonate to Methanol and Ethylene Glycol over Cu/SiO2 Catalysts Prepared by Ammonia Evaporation Method. Int. J. Hydrog. Energy 2017, 42, 2144–2156. [Google Scholar] [CrossRef]
  43. Zhou, M.; Shi, Y.; Ma, K.; Tang, S.; Liu, C.; Yue, H.; Liang, B. Nanoarray Cu/SiO2 Catalysts Embedded in Monolithic Channels for the Stable and Efficient Hydrogenation of CO2-Derived Ethylene Carbonate. Ind. Eng. Chem. Res. 2018, 57, 1924–1934. [Google Scholar] [CrossRef]
  44. Fu, J.; Jiang, N.; Ren, D.; Song, Z.; Li, L.; Huo, Z. Hydrothermal Conversion of Ethylene Carbonate to Ethylene Glycol. Int. J. Hydrog. Energy 2016, 41, 9118–9122. [Google Scholar] [CrossRef]
  45. Nakatake, D.; Yazaki, R.; Matsushima, Y.; Ohshima, T. Transesterification Reactions Catalyzed by a Recyclable Heterogeneous Zinc/Imidazole Catalyst. Adv. Synth. Catal. 2016, 358, 2569–2574. [Google Scholar] [CrossRef]
  46. Perin, G.; Silva, C.M.; Borges, E.L.; Goulart, H.A.; Silva, R.B.; Jacob, R.G.; Silva, M.S.; Alves, D.; Schumacher, R.F. Selective Synthesis of 4-Chalcogenylmethyl-1,3-dioxolan-2-ones and 1,3-Bis(organylchalcogenyl)propan-2-ols from 3-O-Tosyl Glycerol 1,2-Carbonate. ChemistrySelect 2016, 1, 6238–6242. [Google Scholar] [CrossRef]
  47. Pongrácz, P.; Kollár, L. Asymmetric Hydroformylation of 4-Vinyl-1,3-dioxolan-2-one. J. Heterocycl. Chem. 2017, 54, 1430–1436. [Google Scholar] [CrossRef]
  48. Liu, X.; Vries, G.J.; Werner, T. Transfer Hydrogenation of Cyclic Carbonates and Polycarbonate to Methanol and Diols by Iron Pincer Catalysts. Green Chem. 2019, 21, 5248–5255. [Google Scholar] [CrossRef]
  49. Huang, S.; Yan, B.; Wang, S.; Ma, X. Recent Advances in Dialkyl Carbonates Synthesis and Applications. Chem. Soc. Rev. 2015, 44, 3079–3116. [Google Scholar] [CrossRef] [PubMed]
  50. Lee, K.-H.; Lee, S.; Shin, D.; Hahm, H.-S. Dimethyl Carbonate Synthesis via Transesterification of Ethylene Carbonate and Methanol using Ionic Liquid Catalysts Immobilized on Mesoporous Cellular Foams. Res. Chem. Intermed. 2016, 42, 109–121. [Google Scholar] [CrossRef]
  51. Saptal, V.B.; Bhanage, B.M. Bifunctional Ionic Liquids for the Multitask Fixation of Carbon Dioxide into Valuable Chemicals. ChemCatChem 2016, 8, 244–250. [Google Scholar] [CrossRef]
  52. Jagodzinski, T.S. Thioamides as Useful Synthons in the Synthesis of Heterocycles. Chem. Rev. 2003, 103, 197–227. [Google Scholar] [CrossRef]
  53. Beletskaya, I.P.; Ananikov, V.P. Transition-Metal-Catalyzed C−S, C−Se, and C−Te Bond Formation via Cross-Coupling and Atom-Economic Addition Reactions. Chem. Rev. 2011, 111, 1596–1636. [Google Scholar] [CrossRef]
  54. Zeni, G.; Lüdtke, D.S.; Panatieri, R.B.; Braga, A.L. Vinylic Tellurides:  From Preparation to Their Applicability in Organic Synthesis. Chem. Rev. 2006, 106, 1032–1076. [Google Scholar] [CrossRef]
  55. Wirth, T. Organoselenium Chemistry in Stereoselective Reactions. Angew. Chem. Int. Ed. 2000, 39, 3740–3749. [Google Scholar] [CrossRef]
  56. Santi, C.; Santoro, S.; Battistelli, B. Organoselenium Compounds as Catalysts in Nature and Laboratory. Curr. Org. Chem. 2010, 14, 2442–2462. [Google Scholar] [CrossRef]
  57. Goswami, S.; Hazra, A.; Chakrabarty, R.; Fun, H.-K. Recognition of Carboxylate Anions and Carboxylic Acids by Selenium-Based New Chromogenic Fluorescent Sensor: A Remarkable Fluorescence Enhancement of Hindered Carboxylates. Org. Lett. 2009, 11, 4350–4353. [Google Scholar] [CrossRef]
  58. Rasale, D.B.; Maity, I.; Das, A.K. In Situ Generation of Redox Active Peptides Driven by Selenoester Mediated Native Chemical Ligation. Chem. Commun. 2014, 50, 11397–11400. [Google Scholar] [CrossRef]
  59. Ahuja, B.L.; Raykar, V.; Joshi, R.; Tiwari, S.; Talreja, S.; Choudhary, G. Electronic Properties and Momentum Densities of Tin Chalcogenides: Validation of PBEsol Exchange-Correlation Potential. Phys. B 2015, 465, 21–28. [Google Scholar] [CrossRef]
  60. Trost, B.M.; Waser, J.; Meyer, A. Total Synthesis of (−)-Pseudolaric Acid B. J. Am. Chem. Soc. 2007, 129, 14556–14557. [Google Scholar] [CrossRef] [Green Version]
  61. Wever, W.J.; Bogart, J.W.; Baccile, J.A.; Chan, A.N.; Schroeder, F.C.; Bowers, A.A. Chemoenzymatic Synthesis of Thiazolyl Peptide Natural Products Featuring an Enzyme-Catalyzed Formal [4+2] Cycloaddition. J. Am. Chem. Soc. 2015, 137, 3494–3497. [Google Scholar] [CrossRef] [Green Version]
  62. Perin, G.; Barcellos, A.M.; Luz, E.Q.; Borges, E.L.; Jacob, R.G.; Lenardão, E.J.; Sancineto, L.; Santi, C. Green Hydroselenation of Aryl Alkynes: Divinyl Selenides as a Precursor of Resveratrol. Molecules 2017, 22, 327. [Google Scholar] [CrossRef] [Green Version]
  63. Perin, G.; Silva, C.M.; Borges, E.L.; Duarte, J.E.G.; Goulart, H.A.; Silva, R.B.; Schumacher, R.F. Selective Synthesis of 4-Thiomethyl-1,3-dioxolan-2-ones under Microwave Irradiation using an Environmentally Benign KF/Al2O3/PEG-400 System. Res. Chem. Intermed. 2016, 42, 5873–5885. [Google Scholar] [CrossRef]
  64. Lenardão, E.J.; Soares, L.K.; Barcellos, A.M.; Perin, G. KF/Al2O3 as a Green System for the Synthesis of Organochalcogen Compounds. Curr. Green Chem. 2016, 3, 4–17. [Google Scholar] [CrossRef]
  65. Miralles, N.; Gómez, J.E.; Kleij, A.W.; Fernández, E. Copper-Mediated SN2′ Allyl–Alkyl and Allyl–Boryl Couplings of Vinyl Cyclic Carbonates. Org. Lett. 2017, 19, 6096–6099. [Google Scholar] [CrossRef]
  66. Khan, A.; Khan, S.; Zhao, C.; Mao, Y.; Chen, Y.; Zhang, Y.J. Enantioselective Construction of Tertiary C–O Bond via Allylic Substitution of Vinylethylene Carbonates with Water and Alcohols. J. Am. Chem. Soc. 2017, 139, 10733–10741. [Google Scholar] [CrossRef]
  67. Wu, X.F.; Neumann, H.; Beller, M. Synthesis of Heterocycles via Palladium-Catalyzed Carbonylations. Chem. Rev. 2013, 113, 1–35. [Google Scholar] [CrossRef]
  68. Chinchilla, R.; Najera, C. Chemicals from Alkynes with Palladium Catalysts. Chem. Rev. 2014, 114, 1783–1826. [Google Scholar] [CrossRef] [Green Version]
  69. Musaev, D.G.; Figg, T.M.; Kaledin, A.L. Versatile Reactivity of Pd-Catalysts: Mechanistic Features of the Mono-N-protected Amino Acid Ligand and Cesium-Halide Base in Pd-Catalyzed C–H Bond Functionalization. Chem. Soc. Rev. 2014, 43, 5009–5031. [Google Scholar] [CrossRef]
  70. Sperger, T.; Sanhueza, I.A.; Kalvet, I.; Schoenebeck, F. Computational Studies of Synthetically Relevant Homogeneous Organometallic Catalysis Involving Ni, Pd, Ir, and Rh: An Overview of Commonly Employed DFT Methods and Mechanistic Insights. Chem. Rev. 2015, 115, 9532–9586. [Google Scholar] [CrossRef]
  71. Zuo, L.; Liu, T.; Chang, X.; Guo, W. An Update of Transition Metal-Catalyzed Decarboxylative Transformations of Cyclic Carbonates and Carbamates. Molecules 2019, 24, 3930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Fürstner, A. From Understanding to Prediction: Gold- and Platinum-Based π-Acid Catalysis for Target Oriented Synthesis. Acc. Chem. Res. 2014, 47, 925–938. [Google Scholar] [CrossRef]
  73. Kakati, N.; Maiti, J.; Lee, S.H.; Jee, S.H.; Viswanathan, B.; Yoon, Y.S. Anode Catalysts for Direct Methanol Fuel Cells in Acidic Media: Do We Have Any Alternative for Pt or Pt–Ru? Chem. Rev. 2014, 114, 12397–12429. [Google Scholar] [CrossRef] [PubMed]
  74. Nie, Y.; Li, L.; Wei, Z. Recent Advancements in Pt and Pt-Free Catalysts for Oxygen Reduction Reaction. Chem. Soc. Rev. 2015, 44, 2168–2201. [Google Scholar] [CrossRef] [PubMed]
  75. Song, G.; Wang, F.; Li, X. C–C, C–O and C–N Bond Formation via Rhodium(III)-Catalyzed Oxidative C–H Activation. Chem. Soc. Rev. 2012, 41, 3651–3678. [Google Scholar] [CrossRef]
  76. Etayo, P.; Vidal-Ferran, A. Rhodium-Catalysed Asymmetric Hydrogenation as a Valuable Synthetic Tool for the Preparation of Chiral Drugs. Chem. Soc. Rev. 2013, 42, 728–754. [Google Scholar] [CrossRef] [PubMed]
  77. Ye, B.; Cramer, N. Chiral Cyclopentadienyls: Enabling Ligands for Asymmetric Rh(III)-Catalyzed C–H Functionalizations. Acc. Chem. Res. 2015, 48, 1308–1318. [Google Scholar] [CrossRef] [PubMed]
  78. Cao, W.; Wang, W.; Lu, K.; Yao, W.; Xue, F.; Ma, M. Magnesium-Catalyzed Hydroboration of Organic Carbonates, carbon Dioxide and Esters. Dalton Trans. 2020, 49, 2776–2780. [Google Scholar] [CrossRef] [PubMed]
  79. Ibn El Alami, M.S.; Suisse, I.; Fadlallah, S.; Sauthier, M.; Visseaux, M.C.R. Telomerization of 1,3-Butadiene with Glycerol Carbonate and Subsequent Ring-Opening Lactone Co-polymerization. Chimie 2016, 19, 299–305. [Google Scholar] [CrossRef] [Green Version]
  80. Liu, Q.; Liu, T.-X.; Ru, Y.; Zhu, X.; Zhang, G. Palladium-Catalyzed Decarboxylative Heterocyclizations of [60]Fullerene: Preparation of Novel Vinyl-substituted [60]Fullerene-fused Tetrahydrofurans/Pyrans/Quinolines. Chem. Commun. 2019, 55, 14498–14501. [Google Scholar] [CrossRef]
  81. Singha, S.; Patra, T.; Daniliuc, C.G.; Glorius, F. Highly Enantioselective [5+2] Annulations through Cooperative N-Heterocyclic Carbene (NHC) Organocatalysis and Palladium Catalysis. J. Am. Chem. Soc. 2018, 140, 3551–3554. [Google Scholar] [CrossRef]
  82. Rong, Z.-Q.; Yang, L.-C.; Liu, S.; Yu, Z.; Wang, Y.-N.; Tan, Z.Y.; Huang, R.-Z.; Lan, Y.; Zhao, Y. Nine-Membered Benzofuran-Fused Heterocycles: Enantioselective Synthesis by Pd-Catalysis and Rearrangement via Transannular Bond Formation. J. Am. Chem. Soc. 2017, 139, 15304–15307. [Google Scholar] [CrossRef] [PubMed]
  83. Zhao, C.; Shah, B.H.; Khan, I.; Kan, Y.; Zhang, Y. Enantioselective Synthesis of Isoxazoline N-Oxides via Pd-Catalyzed Asymmetric Allylic Cycloaddition of Nitro-Containing Allylic Carbonates. J. Org. Lett. 2019, 21, 9045–9049. [Google Scholar] [CrossRef] [PubMed]
  84. Zhao, C.; Khan, I.; Zhang, Y.J. Enantioselective Total Synthesis of Furofuran Lignans via Pd-Catalyzed Asymmetric Allylic Cycloadditon of Vinylethylene Carbonates with 2-Nitroacrylates. Chem. Commun. 2020, 56, 12431–12434. [Google Scholar] [CrossRef]
  85. Das, P.; Gondo, S.; Nagender, P.; Uno, H.; Tokunaga, E.; Shibata, N. Access to Benzo-fused Nine-membered Heterocyclic Alkenes with a Trifluoromethyl Carbinol Moiety via a Double Decarboxylative Formal Ring-Expansion Process under Palladium Catalysis. Chem. Sci. 2018, 9, 3276–3281. [Google Scholar] [CrossRef] [Green Version]
  86. Guo, W.; Martínez-Rodríguez, L.; Martin, E.; Escudero-Adán, E.C.; Kleij, A.W. Highly Efficient Catalytic Formation of (Z)-1,4-But-2-ene Diols Using Water as a Nucleophile. Angew. Chem. Int. Ed. 2016, 55, 11037–11040. [Google Scholar] [CrossRef]
  87. Li, C.; Breit, B. Rhodium-Catalyzed Dynamic Kinetic Asymmetric Allylation of Phenols and 2-Hydroxypyridines. Chem. Eur. J. 2016, 22, 14655–14663. [Google Scholar] [CrossRef]
  88. Szewczyk, M.; Magre, M.; Zubar, V.; Rueping, M. Reduction of Cyclic and Linear Organic Carbonates Using a Readily Available Magnesium Catalyst. ACS Catal. 2019, 9, 11634–11639. [Google Scholar] [CrossRef]
  89. Xue, X.; Xu, J.; Zhang, L.; Xu, C.; Pan, Y.; Xu, L.; Li, H.; Zhang, W. Rhodium(III)-Catalyzed Direct C-H Olefination of Arenes with Aliphatic Olefins. Adv. Synth. Catal. 2016, 358, 573–583. [Google Scholar] [CrossRef]
  90. Lu, Q.; Klauck, F.J.R.; Glorius, F. Manganese-Catalyzed Allylation via Sequential C–H and C–C/C–Het Bond Activation. Chem. Sci. 2017, 8, 3379–3383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Tortoreto, C.; Achard, T.; Egger, L.; Guénée, L.; Lacour, J. Synthesis of Spiro Ketals, Orthoesters, and Orthocarbonates by CpRu-Catalyzed Decomposition of α-Diazo-β-ketoesters. Org. Lett. 2016, 18, 240–243. [Google Scholar] [CrossRef]
  92. Annunziata, L.; Fouquay, S.; Michaud, G.; Simon, F.; Guillaume, S.M.; Carpentier, J.-F. Mono- and Di-cyclocarbonate Telechelic Polyolefins Synthesized from ROMP Using Glycerol Carbonate Derivatives as Chain-transfer Agents. Polym. Chem. 2013, 4, 1313–1316. [Google Scholar] [CrossRef]
  93. Yagci, Y.; Nuyken, O.; Graubner, V.-M. Encyclopedia of Polymer Science and Technology. In Telechelic Polymers; Wiley Online Library: Hoboken, NJ, USA, 2004. [Google Scholar] [CrossRef]
  94. Kuciński, K.; Szudkowska-Frątczak, J.; Hreczycho, G. Pt-Catalyzed Synthesis of Functionalized Symmetrical and Unsymmetrical Disilazanes. Chem. Eur. J. 2016, 22, 13046–13049. [Google Scholar] [CrossRef] [PubMed]
  95. Jiang, X.; Chen, J.; Zhu, W.; Zhu, W.; Cheng, K.; Liu, Y.; Su, W.-K.; Yu, C.J. Cobalt(III)-Catalyzed Fast and Solvent-Free C–H Allylation of Indoles Using Mechanochemistry. Org. Chem. 2017, 82, 10665–10672. [Google Scholar] [CrossRef]
  96. Bensemhoun, J.; Condon, S. Valorization of Glycerol 1,2-Carbonate as a Precursor for the Development of New Synthons in Organic Chemistry. Green Chem. 2012, 14, 2595–2599. [Google Scholar] [CrossRef]
  97. Singha, S.; Serrano, E.; Mondal, S.; Daniliuc, C.G.; Glorius, F. Diastereodivergent Synthesis of Enantioenriched α,β-Disubstituted γ-Butyrolactones via Cooperative N-Heterocyclic Carbene and Ir Catalysis. Nat. Catal. 2020, 3, 48–54. [Google Scholar] [CrossRef]
  98. Kargar, M.; Hekmatshoar, R.; Ghandi, M.; Mostashari, A. Use of Glycerol Carbonate in an Efficient, One-Pot and Solvent Free Synthesis of 1,3-sn-Diglycerides. J. Am. Oil Chem. Soc. 2012, 90, 259–264. [Google Scholar] [CrossRef]
  99. Yuan, J.; Antonietti, M. Poly(ionic liquid)s: Polymers Expanding Classical Property Profiles. Polymer 2011, 52, 1469–1482. [Google Scholar] [CrossRef] [Green Version]
  100. Yuan, J.; Mecerreyes, D.; Antonietti, M. Poly(ionic liquid)s: An Update. Prog. Polym. Sci. 2013, 38, 1009–1036. [Google Scholar] [CrossRef]
  101. Zhang, Y.; Wang, B.; Elageed, E.H.M.; Qin, L.; Ni, B.; Liu, X.; Gao, G. Swelling Poly(ionic liquid)s: Synthesis and Application as Quasi-Homogeneous Catalysts in the Reaction of Ethylene Carbonate with Aniline. ACS Macro Lett. 2016, 5, 435–438. [Google Scholar] [CrossRef]
  102. Truscello, A.M.; Gambarotti, C.; Lauria, M.; Auricchio, S.; Leonardi, G.; Shisodia, S.U.; Citterio, A. One-pot synthesis of aryloxypropanediols from glycerol: Towards valuable chemicals from renewable sources. Green Chem. 2013, 15, 625–628. [Google Scholar] [CrossRef]
  103. Kao, S.-C.; Lin, Y.-C.; Ryu, I.; Wu, Y.-K. Revisiting Hydroxyalkylation of Phenols with Cyclic Carbonates. Adv. Synth. Catal. 2019, 361, 3639–3644. [Google Scholar] [CrossRef]
  104. Olsén, P.; Oschmann, M.; Johnston, E.V.; Åkermark, B. Synthesis of Highly Functional Carbamates Through Ring-Opening of Cyclic Carbonates with Unprotected α-Amino Acids in Water. Green Chem. 2018, 20, 469–475. [Google Scholar] [CrossRef]
  105. Wang, Y.; Chai, J.; You, C.; Zhang, J.; Mi, X.; Zhang, L.; Luo, S. π-Coordinating Chiral Primary Amine/Palladium Synergistic Catalysis for Asymmetric Allylic Alkylation. J. Am. Chem. Soc. 2020, 142, 3184–3195. [Google Scholar] [CrossRef]
  106. Tillet, G.; Boutevin, B.; Ameduri, B. Chemical Reactions of Polymer Crosslinking and Post-crosslinking at Room and Medium Temperature. Prog. Polym. Sci. 2011, 36, 191–217. [Google Scholar] [CrossRef]
  107. Pham, P.D.; Monge, S.; Lapinte, V.; Raoul, Y.; Robin, J.J. Glycerol-based Co-oligomers by Free-radical Chain Transfer Polymerization: Towards Reactive Polymers bearing Acetal and/or Carbonate Groups with Enhanced Properties. Eur. Polym. J. 2017, 95, 491–502. [Google Scholar] [CrossRef] [Green Version]
  108. Rotta, J.; Pham, P.D.; Lapinte, V.; Borsali, R.; Minatti, E.; Robin, J.-J. Synthesis of Amphiphilic Polymers Based on Fatty Acids and Glycerol-Derived Monomers—A Study of Their Self-Assembly in Water. Macromol. Chem. Phys. 2014, 215, 131–139. [Google Scholar] [CrossRef]
  109. Hamiye, R.; Alaaeddine, A.; Awada, M.; Campagne, B.; Caillol, S.; Guillaume, S.M.; Ameduri, B.; Carpentier, J.-F. From Glycidyl Carbonate to Hydroxyurethane Side-groups in Alternating Fluorinated Copolymers. Polym. Chem. 2014, 5, 5089–5099. [Google Scholar] [CrossRef]
  110. Shaheen, U.; Turney, T.W.; Saito, K.; Gates, W.P.; Patti, A.F. Pendant Cyclic Carbonate-polymer/Na-Smectite Nanocomposites via In Situ Intercalative Polymerization and Solution Intercalation. J. Polym. Sci. Part A Polym. Chem. 2016, 54, 2421–2429. [Google Scholar] [CrossRef]
  111. Mhanna, A.; Sadaka, F.; Boni, G.; Brachais, C.-H.; Brachais, L.; Couvercelle, J.-P.; Plasseraud, L.; Lecamp, L. Photopolymerizable Synthons from Glycerol Derivatives. J. Am. Oil Chem. Soc. 2014, 91, 337–348. [Google Scholar] [CrossRef]
  112. Dubois, P.; Coulembier, O.; Raquez, J.-M. Handbook of Ring-Opening Polymerization; WILEY-VCH: Weinheim, Germany, 2009. [Google Scholar]
  113. Clements, J.H. Reactive Applications of Cyclic Alkylene Carbonates. Ind. Eng. Chem. Res. 2003, 42, 663–674. [Google Scholar] [CrossRef]
  114. Yadav, N.; Seidi, F.; Crespy, D.; D’Elia, V. Polymers Based on Cyclic Carbonates as Trait d’Union Between Polymer Chemistry and Sustainable CO2 Utilization. ChemSusChem 2019, 12, 724–754. [Google Scholar] [CrossRef] [Green Version]
  115. Patil, R.M.; Hong, H.; Chai, C.L.L.; Ghanwat, A.A.; Ganugapati, S.; Gnaneshwar, R. Synthesis and Characterization of α-(Cyclic Carbonate), ω-Hydroxyl/Itaconic Acid Asymmetric Telechelic Poly(ε-caprolactone). Polym. Bull. 2015, 72, 2489–2501. [Google Scholar] [CrossRef]
  116. Kapiti, G.; Keul, H.; Möller, M. Functional PEG Building Blocks via Copolymerization of Ethylene Carbonate and tert-Butyl Glycidyl Ether. Polym. Chem. 2016, 7, 5050–5059. [Google Scholar] [CrossRef]
  117. García, K.E.; Navarro, R.; Ramírez-Hernández, F.; Marcos-Fernández, Á. New Routes to Difunctional Macroglycols Using Ethylene Carbonate: Reaction with Bis-(2-hydroxyethyl) Terephthalate and Degradation of Poly(Ethylene Terephthalate). Polym. Degrad. Stab. 2017, 144, 195–206. [Google Scholar] [CrossRef]
  118. Kong, P.S.; Aroua, M.K.; Daud, W.M.A.W. Conversion of Crude and Pure Glycerol into Derivatives: A Feasibility Evaluation. Renew. Sustain. Energy Rev. 2016, 63, 533–555. [Google Scholar] [CrossRef]
  119. Cunha, A.G.; Gandini, A. Turning Polysaccharides into Hydrophobic Materials: A Critical Review. Part 2. Hemicelluloses, Chitin/Chitosan, Starch, Pectin and Alginates. Cellulose 2010, 17, 1045–1065. [Google Scholar] [CrossRef]
  120. Akil, Y.; Lorenz, D.; Lehnen, R.; Saake, B. Safe and Non-toxic Hydroxyalkylation of Xylan Using Propylene Carbonate. Eur. Polym. J. 2016, 77, 88–97. [Google Scholar] [CrossRef]
  121. Akil, Y.; Lehnen, R.; Saake, B. Novel Synthesis of Hydroxyvinylethyl Xylan Using 4-Vinyl-1,3-dioxolan-2-one. Tetrahedron Lett. 2016, 57, 4200–4202. [Google Scholar] [CrossRef]
  122. Costa, P.L.F.; Melo, V.N.; Guimarães, B.M.; Schuler, M.; Pimenta, V.; Rollin, P.; Tatibouët, A.; Oliveira, R.N. Glycerol Carbonate in Ferrier Reaction: Access to New Enantiopure Building Blocks to Develop Glycoglycerolipid Analogues. Carbohydr. Res. 2016, 436, 1–10. [Google Scholar] [CrossRef]
  123. Vitaku, E.; Smith, D.T.; Njardarson, J.T. Analysis of the Structural Diversity, Substitution Patterns, and Frequency of Nitrogen Heterocycles among U.S. FDA Approved Pharmaceuticals. J. Med. Chem. 2014, 57, 10257–10274. [Google Scholar] [CrossRef] [PubMed]
  124. Neha; Dwivedi, A.R.; Kumar, R.; Kumar, V. Recent Synthetic Strategies for Monocyclic Azole Nucleus and Its Role in Drug Discovery and Development. Curr. Org. Synth. 2018, 15, 321–340. [Google Scholar] [CrossRef]
  125. Eicher, T.; Hauptmann, S. The Chemistry of Heterocycles; Wiley/VCH: Weinheim, Germany, 2003. [Google Scholar]
  126. Wang, B.; Yang, S.; Min, L.; Gu, Y.; Zhang, Y.; Wu, X.; Zhang, L.; Elageed, E.H.M.; Wu, S.; Gao, G. Eco-Efficient Synthesis of Cyclic Carbamates/Dithiocarbonimidates from Cyclic Carbonates/Trithiocarbonate and Aromatic Amines Catalyzed by Ionic Liquid BmimOAc. Adv. Synth. Catal. 2014, 356, 3125–3134. [Google Scholar] [CrossRef]
  127. Vilkauskaitė, G.; Holzer, W.; Šačkus, A. 4-{[(1-Phenyl-1H-pyrazol-3-yl)oxy]methyl}-1,3-dioxolan-2-one. Molbank 2012, 2012, M786. [Google Scholar] [CrossRef] [Green Version]
  128. Zidar, N.; Tomašič, T.; Macut, H.; Sirc, A.; Brvar, M.; Montalvão, S.; Tammela, P.; Ilaš, J.; Kikelj, D. New N-Phenyl-4,5-dibromopyrrolamides and N-Phenylindolamides as ATPase Inhibitors of DNA Gyrase. Eur. J. Med. Chem. 2016, 117, 197–211. [Google Scholar] [CrossRef]
  129. Jo, H.; Han, S.; Park, J.; Choi, M.; Han, S.H.; Jeong, T.; Lee, S.-Y.; Kwak, J.H.; Jung, Y.H.; Kim, I.S. Rhodium(III)-Catalyzed Heteroatom-directed C–H Allylation with Allylic Phosphonates and Allylic Carbonates at Room Temperature. Tetrahedron 2016, 72, 571–578. [Google Scholar] [CrossRef]
  130. Gosling, S.; El Amri, C.; Tatibouët, A. Staudinger Condensation for the Preparation of Thiohydantoins. Synthesis 2014, 46, 1079–1084. [Google Scholar] [CrossRef]
  131. Yang, L.-C.; Rong, Z.-Q.; Wang, Y.-N.; Tan, Z.Y.; Wang, M.; Zhao, Y. Construction of Nine-Membered Heterocycles through Palladium-Catalyzed Formal [5+4] Cycloaddition. Angew. Chem. Int. Ed. 2017, 56, 2927–2931. [Google Scholar] [CrossRef] [PubMed]
  132. Yang, L.-C.; Tan, Z.Y.; Rong, Z.-Q.; Liu, R.; Wang, Y.-N.; Zhao, Y. Palladium-Titanium Relay Catalysis Enables Switch from Alkoxide-pAllyl to Dienolate Reactivity for Spiro-Heterocycle Synthesis. Angew. Chem. Int. Ed. 2018, 57, 7860–7864. [Google Scholar] [CrossRef]
  133. Yuan, C.; Wu, Y.; Wang, D.; Zhang, Z.; Wang, C.; Zhou, L.; Zhang, C.; Song, B.; Guo, H. Formal [5+3] Cycloaddition of Zwitterionic Allylpalladium Intermediates with Azomethine Imines for Construction of N,O-Containing Eight-Membered Heterocycles. Adv. Synth. Catal. 2018, 360, 652–658. [Google Scholar] [CrossRef]
  134. Wang, H.; Lorion, M.M.; Ackermann, L. Air-Stable Manganese(I)-Catalyzed C-H Activation for Decarboxylative C-H/C-O Cleavages in Water. Angew. Chem. Int. Ed. 2017, 56, 6339–6342. [Google Scholar] [CrossRef]
  135. Wang, H.; Lorion, M.M.; Ackermann, L. Domino C–H/N–H Allylations of Imidates by Cobalt Catalysis. ACS Catal. 2017, 7, 3430–3433. [Google Scholar] [CrossRef]
  136. Lang, X.-D.; Li, Z.-M.; He, L.-N. Protic Ionic Liquid-Catalyzed Synthesis of Oxazolidinones Using Cyclic Carbonates as both CO2 Surrogate and Sustainable Solvent. Catal. Today 2019, 324, 167–173. [Google Scholar] [CrossRef]
  137. Ji, M.; Lv, H.Y.; Cheng, L.C.; Wang, T.T.; Chong, S.Y. Metal-Organic Framework MIL-101-NH2-Supported Acetate-Based Butylimidazolium Ionic Liquid as a Highly Efficient Heterogeneous Catalyst for the Synthesis of 3-Aryl-2-oxazolidinones. Langmuir 2019, 35, 495–503. [Google Scholar] [CrossRef]
  138. Rakhmaninam, N.Y.; van den Anker, J.N. Efavirenz in the Therapy of HIV Infection. Expert Opin. Drug Metab. Toxicol. 2010, 6, 95–103. [Google Scholar] [CrossRef] [Green Version]
  139. Splinter, M.Y. Ezogabine (Retigabine) and its Role in the Treatment of Partial-onset Seizures: A Review. Clin. Ther. 2012, 34, 1845–1856. [Google Scholar] [CrossRef]
  140. Cooper, S.L.; Guan, J. Advances in Polyurethane Biomaterials; Woodhead Publishing: Cambridge, UK, 2016; ISBN 9780081006221. [Google Scholar]
  141. Ochiai, B.; Koda, K.; Endo, T. Branched Cationic Polyurethane Prepared by Polyaddition of Chloromethylated Five-Membered Cyclic Carbonate and Diethylenetriamine in Molten Salts. J. Polym. Sci. A Polym. Chem. 2012, 50, 47–51. [Google Scholar] [CrossRef]
  142. He, Y.; Heine, E.; Keusgen, N.; Keul, H.; Möller, M. Synthesis and Characterization of Amphiphilic Monodisperse Compounds and Poly(Ethylene Imine)s: Influence of their Microstructures on the Antimicrobial Properties. Biomacromolecules 2012, 13, 612–623. [Google Scholar] [CrossRef] [PubMed]
  143. Shisodia, S.U.; Auricchio, S.; Citterio, A.; Grassi, M.; Sebastiano, R. New Examples of Template Catalysis based Processes: Glycerol-like Units as Efficient Promoters for Dehydrative Nucleophilic Substitutions of Ferrocenylmethanol. Tetrahedron Lett. 2014, 55, 869–872. [Google Scholar] [CrossRef]
  144. Tian, L.; Gong, L.; Zhang, X. Copper-Catalyzed Enantioselective Synthesis of β-Amino Alcohols Featuring Tetrasubstituted Tertiary Carbons. Adv. Synth. Catal. 2018, 360, 2055–2059. [Google Scholar] [CrossRef]
  145. Kinch, M.S.; Patridge, E. An Analysis of FDA-Approved Drugs for Infectious Disease: HIV/AIDS Drugs. Drug Discov. Today 2014, 19, 1510–1513. [Google Scholar] [CrossRef]
  146. Clemitson, I. Castable Polyurethane Elastomers; CRC Press: New York, NY, USA, 2008. [Google Scholar]
  147. Suryawanshi, Y.; Sanap, P.; Wani, V. Advances in the Synthesis of Non-isocyanate Polyurethanes. Polym. Bull. 2018, 76, 3233–3246. [Google Scholar] [CrossRef]
  148. Sonnenschein, M.F. Polyurethanes—Science, Technology, Markets and Trends; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015. [Google Scholar]
  149. Ghasemlou, M.; Daver, F.; Ivanova, E.P.; Adhikari, B. Bio-based Routes to Synthesize Cyclic Carbonates and Polyamines Precursors of Non-isocyanate Polyurethanes: A review. Eur. Polym. J. 2019, 118, 668–684. [Google Scholar] [CrossRef]
  150. Monie, F.; Grignard, B.; Thomassin, J.-M.; Mereau, R.; Tassaing, T.; Jerome, C.; Detrembleur, C. Chemo- and Regioselective Additions of Nucleophiles to Cyclic Carbonates for the Preparation of Self-Blowing Non-Isocyanate Polyurethane Foams. Angew. Chem. Int. Ed. 2020, 59, 17033–17041. [Google Scholar] [CrossRef] [PubMed]
  151. Nohra, B.; Candy, L.; Blanco, J.F.; Raoul, Y.; Mouloungui, Z. Synthesis of Five and Six-membered Cyclic Glycerilic Carbonates bearing Exocyclic Urethane Functions. Eur. J. Lipid Sci. Technol. 2013, 115, 111–122. [Google Scholar] [CrossRef]
  152. Bassam, N.; Laure, C.; Jean-François, B.; Yann, R.; Zephirin, M. Aza-Michael versus Aminolysis Reactions of Glycerol Carbonate Acrylate. Green Chem. 2013, 15, 1900–1909. [Google Scholar] [CrossRef]
  153. Nohra, B.; Candy, L.; Blanco, J.-F.; Raoul, Y.; Mouloungui, Z. Aminolysis Reaction of Glycerol Carbonate in Organic and Hydroorganic Medium. J. Am. Oil Chem. Soc. 2012, 89, 1125–1133. [Google Scholar] [CrossRef] [Green Version]
  154. Lombardo, V.M.; Dhulst, E.A.; Leitsch, E.K.; Wilmot, N.; Heath, W.H.; Gies, A.P.; Miller, M.D.; Torkelson, J.M.; Scheidt, K.A. Cooperative Catalysis of Cyclic Carbonate Ring Opening: Application Towards Non-Isocyanate Polyurethane Materials. Eur. J. Org. Chem. 2015, 2015, 2791–2795. [Google Scholar] [CrossRef]
  155. Van Velthoven, J.L.J.; Gootjes, L.; Van Es, D.S.; Noordover, B.A.J.; Meuldijk, J. Poly(Hydroxy Urethane)s Based on Renewable Diglycerol Dicarbonate. Eur. Polym. J. 2015, 70, 125–135. [Google Scholar] [CrossRef]
  156. Blain, M.; Yau, H.; Jean-Gérard, L.; Auvergne, R.; Benazet, D.; Schreiner, P.R.; Caillol, S.; Andrioletti, B. Urea- and Thiourea- Catalyzed Aminolysis of Carbonates. ChemSusChem 2016, 9, 2269–2272. [Google Scholar] [CrossRef]
  157. Lee, J.Y.; Saito, R. Polyurethane-Silica Nanocomposites Provided from Perhydropolysilazane: Polymerization Mechanism. Polymer (Guildf) 2018, 135, 251–260. [Google Scholar] [CrossRef]
  158. Annunziata, L.; Diallo, A.K.; Fouquay, S.; Michaud, G.; Simon, F.; Brusson, J.M.; Carpentier, J.F.; Guillaume, S.M. α,ω-Di(glycerol arbonate) Telechelic Polyesters and Polyolefins as Precursors to Polyhydroxyurethanes: An Isocyanate-free Approach. Green Chem. 2014, 16, 1947–1956. [Google Scholar] [CrossRef] [Green Version]
  159. Aricò, F.; Bravo, S.; Crisma, M.; Tundo, P. 1,3-Oxazinan-2-ones via Carbonate Chemistry: A Facile, High Yielding Synthetic Approach. Pure Appl. Chem. 2016, 88, 227–237. [Google Scholar] [CrossRef] [Green Version]
  160. Mei, C.; Zhao, Y.; Zou, K.; Cao, C.; Pang, G.; Shi, Y. Synthesis of N-Aryl-2-oxazolidinones from Cyclic Carbonates and Aromatic Amines Catalyzed by Bio-catalyst. Res. Chem. Intermed. 2018, 44, 2179–2194. [Google Scholar] [CrossRef]
  161. Blain, M.; Jean-Gérard, L.; Auvergne, R.; Benazet, D.; Caillol, S.; Andrioletti, B. Rational Investigations in the Ring Opening of Cyclic Carbonates by Amines. Green Chem. 2014, 16, 4286–4291. [Google Scholar] [CrossRef]
  162. Mommer, S.; Lamberts, K.; Keul, H.; Möller, M. A Novel Multifunctional Coupler: The Concept of Coupling and Proof of Principle. Chem. Commun. 2013, 49, 3288–3290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Legros, V.; Taing, G.; Buisson, P.; Schuler, M.; Bostyn, S.; Rousseau, J.; Sinturel, C.; Tatibouët, A. Activated Glycerol Carbonates, Versatile Reagents with Aliphatic Amines: Formation and Reactivity of Glycidyl Carbamates and Trialkylamines. Eur. J. Org. Chem. 2017, 2017, 5032–5043. [Google Scholar] [CrossRef]
  164. Magliozzi, F.; Chollet, G.; Grau, E.; Cramail, H. Benefit of the Reactive Extrusion in the Course of Polyhydroxyurethanes Synthesis by Aminolysis of Cyclic Carbonates. ACS Sustain. Chem. Eng. 2019, 7, 17282–17292. [Google Scholar] [CrossRef] [Green Version]
  165. Quérette, T.; Fleury, E.; Sintes-Zydowicz, N. Non-isocyanate Polyurethane Nanoparticles Prepared by Nanoprecipitation. Eur. Polym. J. 2019, 114, 434–445. [Google Scholar] [CrossRef]
  166. Quienne, B.; Kasmi, N.; Dieden, R.; Caillol, S.; Habibi, Y. Isocyanate-Free Fully Biobased Star Polyester-Urethanes: Synthesis and Thermal Properties. Biomacromolecules 2020, 21, 1943–1951. [Google Scholar] [CrossRef] [PubMed]
  167. Pyo, K.-h.; Lee, D.H.; Kim, Y.; Kim, J.-W. Extremely Rapid and Simple Healing of a Transparent Conductor Based on Ag Nanowires and Polyurethane with a Diels-Alder Network. J. Mater. Chem. C 2016, 4, 972–977. [Google Scholar] [CrossRef]
  168. Jaratrotkamjorn, R.; Nourry, A.; Pasetto, P.; Choppé, E.; Panwiriyarat, W.; Tanrattanakul, V.; Pilard, J.-F. Synthesis and Characterization of Elastomeric, Biobased, Nonisocyanate Polyurethane from Natural Rubber. J. Appl. Polym. Sci. 2017, 134, 45427–45439. [Google Scholar] [CrossRef]
  169. Rix, E.; Grau, E.; Chollet, G.; Cramail, H. Synthesis of Fatty Acid-based Non-isocyanate Polyurethanes, NIPUs, in Bulk and Mini-Emulsion. Eur. Polym. J. 2016, 84, 863–872. [Google Scholar] [CrossRef] [Green Version]
  170. Bao, Y.M.; Shen, G.R.; He, J.; Li, Y.S. Water-soluble Hyperbranched Poly(Ester Urethane)s Based on D,L-Alanine: Isocyanate-free Synthesis, Post-functionalization and Application. Green Chem. 2012, 14, 2243–2250. [Google Scholar] [CrossRef]
  171. Besse, V.; Camara, F.; Méchin, F.; Fleury, E.; Caillol, S.; Pascault, J.P.; Boutevin, B. How to Explain Low Molar Masses in PolyHydroxyUrethanes (PHUs). Eur. Polym. J. 2015, 71, 1–11. [Google Scholar] [CrossRef]
  172. Zhang, K.; Nelson, A.M.; Talley, S.J.; Chen, M.; Margaretta, E.; Hudson, A.G.; Moore, R.B.; Long, T.E. Non-isocyanate Poly(Amide-hydroxyurethane)s from Sustainable Resources. Green Chem. 2016, 18, 4667–4681. [Google Scholar] [CrossRef] [Green Version]
  173. Lambeth, R.H.; Rizvi, A. Mechanical and Adhesive Properties of Hybrid Epoxy-Polyhydroxyurethane Network Polymers. Polymer 2019, 183, 121881–121888. [Google Scholar] [CrossRef]
  174. Dai, J.; Wu, Z.; Tang, L.; Qu, J. Preparation of Five-membered Bis(Cyclic Carbonate)s at Atmospheric Pressure for Polyhydroxyurethane Coatings. J. Appl. Polym. Sci. 2019, 136, 47957–47965. [Google Scholar] [CrossRef]
  175. Wu, Z.; Tang, L.; Dai, J.; Qu, J. Synthesis and Properties of Aqueous Cyclic Carbonate Dispersion and Nonisocyanate Polyurethanes under Atmospheric Pressure. Prog. Org. Coat. 2019, 136, 105209–105217. [Google Scholar] [CrossRef]
  176. Duval, C.; Kébir, N.; Jauseau, R.; Burel, F. Organocatalytic Synthesis of Novel Renewable Non-isocyanate Polyhydroxyurethanes. J. Polym. Sci. Part A Polym. Chem. 2016, 54, 758–764. [Google Scholar] [CrossRef]
  177. Fleischer, M.; Blattmann, H.; Mülhaupt, R. Glycerol-, Pentaerythritol- and Trimethylolpropane-based Polyurethanes and their Cellulose Carbonate Composites Prepared via the Non-isocyanate Route with Catalytic Carbon Dioxide Fixation. Green Chem. 2013, 15, 934–942. [Google Scholar] [CrossRef]
  178. Carré, C.; Zoccheddu, H.; Delalande, S.; Pichon, P.; Avérous, L. Synthesis and Characterization of Advanced Biobased Thermoplastic Nonisocyanate Polyurethanes, with Controlled Aromatic-Aliphatic Architectures. Eur. Polym. J. 2016, 84, 759–769. [Google Scholar] [CrossRef]
  179. Tryznowski, M.; Świderska, A.; Zołek-Tryznowska, Z.; Gołofit, T.; Parzuchowski, P.G. Facile Route to Multigram Synthesis of Environmentally Friendly Non-isocyanate Polyurethanes. Polymer (Guildf) 2015, 80, 228–236. [Google Scholar] [CrossRef]
  180. Benyahya, S.; Habas, J.P.; Auvergne, R.; Lapinte, V.; Caillol, S. Structure–Property Relationships in Polyhydroxyurethanes Produced from Terephthaloyl Dicyclocarbonate with Various Polyamines. Polym. Int. 2012, 61, 1666–1674. [Google Scholar] [CrossRef]
  181. Lee, S.H.; Lee, D.S. Self-Healing and Rheological Properties of Polyhydroxyurethane Elastomers Based on Glycerol Carbonate Capped Prepolymers. Macromol. Res. 2019, 27, 460–469. [Google Scholar] [CrossRef]
  182. Wazarkar, K.; Kathalewar, M.; Sabnis, A. Development of Epoxy-Urethane Hybrid Coatings via Non-isocyanate Route. Eur. Polym. J. 2016, 84, 812–827. [Google Scholar] [CrossRef]
  183. Wang, Z.; Zhang, X.; Zhang, L.; Tan, T.; Fong, H. Nonisocyanate Biobased Poly(ester urethanes) with Tunable properties synthesized via an environment-friendly route. ACS Sustain. Chem. Eng. 2016, 4–5, 2762–2770. [Google Scholar] [CrossRef]
  184. Tale, N.V.; Jagtap, R.N.; Tathe, D.S. An Efficient Approach for the Synthesis of Thermoset Polyurethane Acrylate Polymer and its Film Properties. Des. Monomers Polym. 2014, 17, 147–155. [Google Scholar] [CrossRef] [Green Version]
  185. Mamiński, M.Ł.; Szymański, R.; Parzuchowski, P.; Antczak, A.; Szymona, K. Hyperbranched Polyglycerols with Bisphenol A Core as Glycerol-derived Components of Polyurethane Wood Adhesives. BioResources 2012, 7, 1440–1451. [Google Scholar] [CrossRef] [Green Version]
  186. Dolci, E.; Michaud, G.; Simon, F.; Boutevin, B.; Fouquay, S.; Caillol, S. Remendable Thermosetting Polymers for Isocyanate-free Adhesives: A Preliminary Study. Polym. Chem. 2015, 6, 7851–7861. [Google Scholar] [CrossRef]
  187. Leitsch, E.K.; Heath, W.H.; Torkelson, J.M. Polyurethane/Polyhydroxyurethane Hybrid Polymers and their Applications as Adhesive Bonding Agents. Int. J. Adhes. Adhes. 2016, 64, 1–8. [Google Scholar] [CrossRef] [Green Version]
  188. Ju, Y.H.; Lee, H.-J.; Han, C.J.; Lee, C.-R.; Kim, Y.; Kim, J.-W. Water-responsive Pressure-sensitive Adhesive with Reversibly Changeable Adhesion for Fabrication of Stretchable Devices. Mater. Des. 2020, 195, 108995. [Google Scholar] [CrossRef]
  189. Weil, E.D.; Levchik, S.V. Flame Retardants for Plastics and Textiles; Hanser Publishers: Munich, Germany, 2009. [Google Scholar]
  190. Dufton, P.W. Flame Retardants for Plastics Market. Report; Rapra Technology Limited: West Midlands, UK, 2003. [Google Scholar]
  191. Benin, V.; Gardelle, B.; Morgan, A.B. Heat Release of Polyurethanes Containing Potential Flame Retardants Based on Boron and Phosphorus Chemistries. Polym. Degrad. Stab. 2014, 106, 108–121. [Google Scholar] [CrossRef]
  192. Benin, V.; Cui, X.; Morgan, A.B.; Seiwert, K.J. Synthesis and Flammability Testing of Epoxy Functionalized Phosphorous-based Flame Retardants. Appl. Polym. Sci. 2015, 132, 42296–42305. [Google Scholar] [CrossRef]
  193. Sonnenschein, M.F.; Virgili, J.M.; Babb, D.A.; Bell, B.M.; Nickless, B.C. Synthesis and Flame Retardant Potential of Polyols Based on Bisphenol-S. J. Polym. Sci. Part A Polym. Chem. 2016, 54, 2102–2108. [Google Scholar] [CrossRef]
  194. Eisele, A.; Kyriakos, K.; Bhandary, R.; Schönhoff, M.; Papadakis, C.M.; Rieger, B. Structure and Ionic Conductivity of Liquid Crystals having Propylene Carbonate Units. J. Mater. Chem. A 2015, 3, 2942–2953. [Google Scholar] [CrossRef]
  195. Zhu, Z.; Einset, A.G.; Yang, C.-Y.; Chen, W.-X.; Wnek, G.E. Synthesis of Polysiloxanes Bearing Cyclic Carbonate Side Chains. Dielectric Properties and Ionic Conductivities of Lithium Triflate Complexes. Macromolecules 1994, 27, 4076–4079. [Google Scholar] [CrossRef]
  196. Lee, D.H.; Heo, G.; Pyo, K.; Kim, Y.; Kim, J.-W. Mechanically Robust and Healable Transparent Electrode Fabricated via Vapor-Assisted Solution Process. ACS Appl. Mater. Interfaces 2016, 8, 8129–8136. [Google Scholar] [CrossRef]
  197. Fu, W.; Wang, L.; Yang, Z.; Shen, J.-S.; Tang, F.; Zhang, J.; Cui, X. Facile Access to Versatile Aza-macrolides Through Iridium-Catalysed Cascade Allyl-amination/Macrolactonization. Chem. Commun. 2020, 56, 960–963. [Google Scholar] [CrossRef] [PubMed]
  198. Mollaeian, K.; Wei, S.; Islam, M.R.; Dickerson, B.; Holmes, W.E.; Benson, T.J. Development of an Online Raman Analysis Technique for Monitoring the Production of Biofuels. Energy Fuels 2016, 30, 4112–4117. [Google Scholar] [CrossRef]
  199. Steber, J. Handbook for Cleaning/Decontamination of Surfaces: The Ecotoxicity of Cleaning Product Ingredients; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  200. Castro, M.J.L.; Ojeda, C.; Cirelli, A.F. Advances in Surfactants for Agrochemicals. Environ. Chem. Lett. 2014, 12, 85–95. [Google Scholar] [CrossRef]
  201. Sorella, G.; Strukul, G.; Scarso, A. Recent Advances in Catalysis in Micellar Media. Green Chem. 2015, 17, 644–683. [Google Scholar] [CrossRef]
  202. Climent, M.J.; Corma, A.; Iborra, S.; Martínez-Silvestre, S.; Velty, A. Preparation of Glycerol Carbonate Esters by using Hybrid Nafion-Silica Catalyst. ChemSusChem 2013, 6, 1224–1234. [Google Scholar] [CrossRef] [Green Version]
  203. Mauritz, K.A.; Moore, R.B. State of Understanding of Nafion. Chem. Rev. 2004, 104, 4535–4586. [Google Scholar] [CrossRef] [PubMed]
  204. Han, L.; Wang, T. Preparation of Glycerol Monostearate from Glycerol Carbonate and Stearic Acid. RSC Adv. 2016, 6, 34137–34145. [Google Scholar] [CrossRef]
  205. Campagne, B.; David, G.; Améduri, B.; Jones, D.J.; Rozière, J.; Roche, I. New Semi-IPN PEMFC Membranes Composed of Crosslinked Fluorinated Copolymer bearing Triazole Groups and sPEEK for Operation at Low Relative Humidity. Int. J. Hydrog. Energy 2015, 40, 16797–16813. [Google Scholar] [CrossRef]
  206. Sacripante, G.G.; Zhou, K.; Farooque, M. Sustainable Polyester Resins Derived from Rosins. Macromolecules 2015, 48, 6876–6881. [Google Scholar] [CrossRef]
  207. Liu, B.; Zhang, Y.-Y.; Zhang, X.-H.; Du, B.-Y.; Fan, Z.-Q. Fixation of carbon dioxide concurrently or in tandem with free radical polymerization for highly transparent polyacrylates with specific UV absorption. Polym. Chem. 2016, 7, 3731–3739. [Google Scholar] [CrossRef]
  208. Poupart, R.; Houda, D.N.E.; Chellapermal, D.; Guerrouache, M.; Carbonnier, B.; Droumaguet, B.L. Novel In-capillary Polymeric Monoliths Arising from Glycerol Carbonate Methacrylate for Flow-through Catalytic and Chromatographic Applications. RSC Adv. 2016, 6, 13614–13617. [Google Scholar] [CrossRef]
Figure 1. Versatility of five-membered cyclic carbonates.
Figure 1. Versatility of five-membered cyclic carbonates.
Applsci 11 05024 g001
Scheme 1. The decarboxylation of GC 1a to glycidol 2a.
Scheme 1. The decarboxylation of GC 1a to glycidol 2a.
Applsci 11 05024 sch001
Scheme 2. Decarboxylation reactions using US, MW and ZnO/La2O3.
Scheme 2. Decarboxylation reactions using US, MW and ZnO/La2O3.
Applsci 11 05024 sch002
Scheme 3. One-pot consecutive glycidol synthesis method from glycerol.
Scheme 3. One-pot consecutive glycidol synthesis method from glycerol.
Applsci 11 05024 sch003
Scheme 4. Organocatalytic coupling reaction between the bromolactide 4 and carbonates 1.
Scheme 4. Organocatalytic coupling reaction between the bromolactide 4 and carbonates 1.
Applsci 11 05024 sch004
Scheme 5. Synthesis of ethers via a Pd-catalyzed reaction of carbonates 6 with phenols 7.
Scheme 5. Synthesis of ethers via a Pd-catalyzed reaction of carbonates 6 with phenols 7.
Applsci 11 05024 sch005
Scheme 6. Catalytic hydrogenation of cyclic carbonates by using (PNP)RuII pincer complexes as catalysts.
Scheme 6. Catalytic hydrogenation of cyclic carbonates by using (PNP)RuII pincer complexes as catalysts.
Applsci 11 05024 sch006
Scheme 7. Hydrogenation of ethylene carbonate 1c.
Scheme 7. Hydrogenation of ethylene carbonate 1c.
Applsci 11 05024 sch007
Scheme 8. Hydrogenation of cyclic carbonates 1, providing methanol and diols 12.
Scheme 8. Hydrogenation of cyclic carbonates 1, providing methanol and diols 12.
Applsci 11 05024 sch008
Scheme 9. Glycols synthesis through hydrolysis of cyclic carbonates.
Scheme 9. Glycols synthesis through hydrolysis of cyclic carbonates.
Applsci 11 05024 sch009
Scheme 10. The indirect hydrogenation of CO2 via the ethylene carbonate intermediate 1c to produce ethylene glycol 12b and methanol 7a.
Scheme 10. The indirect hydrogenation of CO2 via the ethylene carbonate intermediate 1c to produce ethylene glycol 12b and methanol 7a.
Applsci 11 05024 sch010
Scheme 11. Catalytic hydrogenation or hydrolysis of cyclic carbonates using organic solvents.
Scheme 11. Catalytic hydrogenation or hydrolysis of cyclic carbonates using organic solvents.
Applsci 11 05024 sch011
Scheme 12. Synthesis of diols 12 from 4-chalcogenomethyl-1,3-dioxalan-2-ones 1.
Scheme 12. Synthesis of diols 12 from 4-chalcogenomethyl-1,3-dioxalan-2-ones 1.
Applsci 11 05024 sch012
Scheme 13. Hydroformylation of racemic 4-vinyl-1,3-dioxolan-2-one 1e to prepare the aldehyde 1f.
Scheme 13. Hydroformylation of racemic 4-vinyl-1,3-dioxolan-2-one 1e to prepare the aldehyde 1f.
Applsci 11 05024 sch013
Scheme 14. Metal complex as a catalyst for the transfer hydrogenation of carbonate 14.
Scheme 14. Metal complex as a catalyst for the transfer hydrogenation of carbonate 14.
Applsci 11 05024 sch014
Scheme 15. Synthesis of dimethyl carbonate 16a via transesterification of ethylene carbonate 1c.
Scheme 15. Synthesis of dimethyl carbonate 16a via transesterification of ethylene carbonate 1c.
Applsci 11 05024 sch015
Scheme 16. Synthesis of 4-arylsulfanylmethyl-1,3-dioxolan-2-ones 1 under MW.
Scheme 16. Synthesis of 4-arylsulfanylmethyl-1,3-dioxolan-2-ones 1 under MW.
Applsci 11 05024 sch016
Scheme 17. Synthesis of 1,3-bis-arylthiopropan-2-ols 18 by incorporation of two thiol.
Scheme 17. Synthesis of 1,3-bis-arylthiopropan-2-ols 18 by incorporation of two thiol.
Applsci 11 05024 sch017
Scheme 18. Selective synthesis under stoichiometric control of chalcogenoethers 1.
Scheme 18. Selective synthesis under stoichiometric control of chalcogenoethers 1.
Applsci 11 05024 sch018
Scheme 19. Synthesis of symmetrical and unsymmetrical 3-bis(organochalcogenyl)propan-2-ols.
Scheme 19. Synthesis of symmetrical and unsymmetrical 3-bis(organochalcogenyl)propan-2-ols.
Applsci 11 05024 sch019
Figure 2. Scope of the ligands employed in the synthesis of compounds 8, 9, 12, and 21.
Figure 2. Scope of the ligands employed in the synthesis of compounds 8, 9, 12, and 21.
Applsci 11 05024 g002
Scheme 20. Telomerization between GC 1a and 1,3-butadiene 26.
Scheme 20. Telomerization between GC 1a and 1,3-butadiene 26.
Applsci 11 05024 sch020
Scheme 21. Synthesis of nine-membered cyclic compounds 30, starting from azadienes 31 and VECs 6.
Scheme 21. Synthesis of nine-membered cyclic compounds 30, starting from azadienes 31 and VECs 6.
Applsci 11 05024 sch021
Scheme 22. Decarboxylative hydration of VECs 6.
Scheme 22. Decarboxylative hydration of VECs 6.
Applsci 11 05024 sch022
Scheme 23. Oxidation to the epoxide 2b, high-yield conversion into the cyclic orthoester 34, and chemoselective oxidocyclization to the isomeric lactones 35a and 35b.
Scheme 23. Oxidation to the epoxide 2b, high-yield conversion into the cyclic orthoester 34, and chemoselective oxidocyclization to the isomeric lactones 35a and 35b.
Applsci 11 05024 sch023
Figure 3. Products of O-, C- and N-allylation 36a, 37a and 38a, respectively.
Figure 3. Products of O-, C- and N-allylation 36a, 37a and 38a, respectively.
Applsci 11 05024 g003
Scheme 24. Conversion of carbonate 1e by asymmetric O-allylation of 4-bromophenol 7b.
Scheme 24. Conversion of carbonate 1e by asymmetric O-allylation of 4-bromophenol 7b.
Applsci 11 05024 sch024
Scheme 25. The alkaline earth metal-catalyzed reduction of cyclic and linear organic carbonates.
Scheme 25. The alkaline earth metal-catalyzed reduction of cyclic and linear organic carbonates.
Applsci 11 05024 sch025
Scheme 26. C-H olefination of arenes and heteroarenes 42 by rhodium catalysis.
Scheme 26. C-H olefination of arenes and heteroarenes 42 by rhodium catalysis.
Applsci 11 05024 sch026
Figure 4. Selected compounds prepared by Xu, Zhang, and co-workers [89].
Figure 4. Selected compounds prepared by Xu, Zhang, and co-workers [89].
Applsci 11 05024 g004
Scheme 27. Synthesis of twelve allylic alcohols.
Scheme 27. Synthesis of twelve allylic alcohols.
Applsci 11 05024 sch027
Scheme 28. Ruthenium catalysis for the synthesis of orthocarbonates 48.
Scheme 28. Ruthenium catalysis for the synthesis of orthocarbonates 48.
Applsci 11 05024 sch028
Scheme 29. Synthesis of α- and α,ω-telechelic cyclocarbonate polycyclooctenes 50 and 51.
Scheme 29. Synthesis of α- and α,ω-telechelic cyclocarbonate polycyclooctenes 50 and 51.
Applsci 11 05024 sch029
Scheme 30. Synthesis of bifunctional unsymmetrical disilazane 54.
Scheme 30. Synthesis of bifunctional unsymmetrical disilazane 54.
Applsci 11 05024 sch030
Scheme 31. C-H Allylation of indoles using VEC.
Scheme 31. C-H Allylation of indoles using VEC.
Applsci 11 05024 sch031
Scheme 32. Conjugate addition reactions of activated glycerol 1,2-carbonate derivatives 1 onto electron-deficient olefins 57.
Scheme 32. Conjugate addition reactions of activated glycerol 1,2-carbonate derivatives 1 onto electron-deficient olefins 57.
Applsci 11 05024 sch032
Scheme 33. Synthesis of α,β-disubstituted γ-butyrolactones 58.
Scheme 33. Synthesis of α,β-disubstituted γ-butyrolactones 58.
Applsci 11 05024 sch033
Scheme 34. Synthesis of diglycerides 3a from fatty acid anhydrides 60 and GC 1a under DABCO catalysis.
Scheme 34. Synthesis of diglycerides 3a from fatty acid anhydrides 60 and GC 1a under DABCO catalysis.
Applsci 11 05024 sch034
Scheme 35. Synthesis of oxazolidones 62 using P[VBIM][AA].
Scheme 35. Synthesis of oxazolidones 62 using P[VBIM][AA].
Applsci 11 05024 sch035
Scheme 36. One-pot synthesis of glycols 12.
Scheme 36. One-pot synthesis of glycols 12.
Applsci 11 05024 sch036
Figure 5. Structure of pharmaceutically important products.
Figure 5. Structure of pharmaceutically important products.
Applsci 11 05024 g005
Scheme 37. Obtaining different aryl β-hydroxyethyl ethers 64.
Scheme 37. Obtaining different aryl β-hydroxyethyl ethers 64.
Applsci 11 05024 sch037
Scheme 38. Synthesis of densely functionalized carbamates.
Scheme 38. Synthesis of densely functionalized carbamates.
Applsci 11 05024 sch038
Scheme 39. Synthesis of enantiomerically enriched allylic adducts 68.
Scheme 39. Synthesis of enantiomerically enriched allylic adducts 68.
Applsci 11 05024 sch039
Scheme 40. Selected examples.
Scheme 40. Selected examples.
Applsci 11 05024 sch040
Scheme 41. Synthesis of glycerol carbonate 1k and 70.
Scheme 41. Synthesis of glycerol carbonate 1k and 70.
Applsci 11 05024 sch041
Scheme 42. Preparing amphiphilic derivatives 73.
Scheme 42. Preparing amphiphilic derivatives 73.
Applsci 11 05024 sch042
Scheme 43. Synthesis of fluorinated terpolymers 7475 and 7879.
Scheme 43. Synthesis of fluorinated terpolymers 7475 and 7879.
Applsci 11 05024 sch043
Scheme 44. Obtaining the polymers 80 and 82.
Scheme 44. Obtaining the polymers 80 and 82.
Applsci 11 05024 sch044
Scheme 45. Synthesis of monoglycerides 83 and subsequent conversion to monoglyceride bis-acrylates 84.
Scheme 45. Synthesis of monoglycerides 83 and subsequent conversion to monoglyceride bis-acrylates 84.
Applsci 11 05024 sch045
Scheme 46. Synthesis of compound 86 and subsequent modification to give hydroxylurethane 87 using a non-isocyanate route.
Scheme 46. Synthesis of compound 86 and subsequent modification to give hydroxylurethane 87 using a non-isocyanate route.
Applsci 11 05024 sch046
Scheme 47. Anionic ring-opening copolymerization.
Scheme 47. Anionic ring-opening copolymerization.
Applsci 11 05024 sch047
Scheme 48. Synthesis of a new polyether by cationic and anionic ROP of a cyclic carbonate-containing epoxide 1m.
Scheme 48. Synthesis of a new polyether by cationic and anionic ROP of a cyclic carbonate-containing epoxide 1m.
Applsci 11 05024 sch048
Scheme 49. Synthesis of xylan derivatives by route (A) or (B).
Scheme 49. Synthesis of xylan derivatives by route (A) or (B).
Applsci 11 05024 sch049
Scheme 50. Synthesis of the glyceryl glycoside 95.
Scheme 50. Synthesis of the glyceryl glycoside 95.
Applsci 11 05024 sch050
Scheme 51. Obtaining oxazolidin-2-ones 62 from cyclic carbonates 1 and aryl amines 63.
Scheme 51. Obtaining oxazolidin-2-ones 62 from cyclic carbonates 1 and aryl amines 63.
Applsci 11 05024 sch051
Scheme 52. [bmim]OAc-Catalyzed synthesis approach for aryl bis-oxazolidin-2-ones 97.
Scheme 52. [bmim]OAc-Catalyzed synthesis approach for aryl bis-oxazolidin-2-ones 97.
Applsci 11 05024 sch052
Scheme 53. Obtaining pyrazole derivatives 100.
Scheme 53. Obtaining pyrazole derivatives 100.
Applsci 11 05024 sch053
Scheme 54. Use of 1h as a bis-electrophile for the N-glyceryl functionalization of diverse aza-aromatic systems 101106.
Scheme 54. Use of 1h as a bis-electrophile for the N-glyceryl functionalization of diverse aza-aromatic systems 101106.
Applsci 11 05024 sch054
Figure 6. Products obtained from benzimidazole-derived carbonate 107.
Figure 6. Products obtained from benzimidazole-derived carbonate 107.
Applsci 11 05024 g006
Scheme 55. Synthesis of a new series of N-phenyl-4,5-dibromopyrrolamides and N-phenylindolamides.
Scheme 55. Synthesis of a new series of N-phenyl-4,5-dibromopyrrolamides and N-phenylindolamides.
Applsci 11 05024 sch055
Scheme 56. C-H Allylation of 2-arylbenzo[d]thiazole 117 with vinyl carbonate 1e to afford the allylic alcohol 118.
Scheme 56. C-H Allylation of 2-arylbenzo[d]thiazole 117 with vinyl carbonate 1e to afford the allylic alcohol 118.
Applsci 11 05024 sch056
Scheme 57. One-pot synthesis method for new N-3-substituted 2-thiohydantoins 119.
Scheme 57. One-pot synthesis method for new N-3-substituted 2-thiohydantoins 119.
Applsci 11 05024 sch057
Scheme 58. Synthesis of benzofuran-fused heterocycles 30 and spirocyclic products 121.
Scheme 58. Synthesis of benzofuran-fused heterocycles 30 and spirocyclic products 121.
Applsci 11 05024 sch058
Scheme 59. Synthesis of [5,5]- and [6,5]-spiro-heterocycles 123 and 124.
Scheme 59. Synthesis of [5,5]- and [6,5]-spiro-heterocycles 123 and 124.
Applsci 11 05024 sch059
Scheme 60. Pd-catalyzed formal [5+3] cycloaddition process to prepare N,O-containing eight-membered heterocycles 126.
Scheme 60. Pd-catalyzed formal [5+3] cycloaddition process to prepare N,O-containing eight-membered heterocycles 126.
Applsci 11 05024 sch060
Scheme 61. C-H allylation of heteroarenes 127 and arenes 128 with vinyl carbonates 14.
Scheme 61. C-H allylation of heteroarenes 127 and arenes 128 with vinyl carbonates 14.
Applsci 11 05024 sch061
Scheme 62. Domino C−H/N−H allylation of arylimidates.
Scheme 62. Domino C−H/N−H allylation of arylimidates.
Applsci 11 05024 sch062
Scheme 63. Carboxylative cyclization of anilines 63 and cyclic carbonates 1.
Scheme 63. Carboxylative cyclization of anilines 63 and cyclic carbonates 1.
Applsci 11 05024 sch063
Scheme 64. Reaction between phenylamine 63 and propylene carbonate 1b.
Scheme 64. Reaction between phenylamine 63 and propylene carbonate 1b.
Applsci 11 05024 sch064
Figure 7. Structures of efavirenz (component of first-line antiretroviral therapy) [138] and retigabine (anticonvulsant medication) [139].
Figure 7. Structures of efavirenz (component of first-line antiretroviral therapy) [138] and retigabine (anticonvulsant medication) [139].
Applsci 11 05024 g007
Scheme 65. The synthesis of a branched cationic polyurethane 134.
Scheme 65. The synthesis of a branched cationic polyurethane 134.
Applsci 11 05024 sch065
Scheme 66. Preparation of poly(ethyleneimine)s (PEI) and amphiphilic monodisperse compounds.
Scheme 66. Preparation of poly(ethyleneimine)s (PEI) and amphiphilic monodisperse compounds.
Applsci 11 05024 sch066
Scheme 67. The condensation of ferrocenemethanol 139 with glycerol derivatives.
Scheme 67. The condensation of ferrocenemethanol 139 with glycerol derivatives.
Applsci 11 05024 sch067
Scheme 68. The copper-catalyzed enantioselective synthesis of highly substituted β-amino alcohols 141.
Scheme 68. The copper-catalyzed enantioselective synthesis of highly substituted β-amino alcohols 141.
Applsci 11 05024 sch068
Scheme 69. Click reaction between 141a and antiretroviral Zidovudine 142.
Scheme 69. Click reaction between 141a and antiretroviral Zidovudine 142.
Applsci 11 05024 sch069
Scheme 70. The routes for the synthesis of PUs.
Scheme 70. The routes for the synthesis of PUs.
Applsci 11 05024 sch070
Scheme 71. The isocyanate-free construction of the urethane nucleus. a Please, check references [151,152,153,167,168,169,170,171,176]; b Please, check references [154,155,156,157,158,173,178,179]; c Please, check references [159,160,183]; d Please, check reference [161,162,163,164,165,166,172,182,186].
Scheme 71. The isocyanate-free construction of the urethane nucleus. a Please, check references [151,152,153,167,168,169,170,171,176]; b Please, check references [154,155,156,157,158,173,178,179]; c Please, check references [159,160,183]; d Please, check reference [161,162,163,164,165,166,172,182,186].
Applsci 11 05024 sch071
Scheme 72. Synthesis of the phosphorylated glycerol carbonate 145.
Scheme 72. Synthesis of the phosphorylated glycerol carbonate 145.
Applsci 11 05024 sch072
Scheme 73. Conversion of GC 1a into its dichlorophosphate 147, which was further reacted with glycidol 2a to give 148.
Scheme 73. Conversion of GC 1a into its dichlorophosphate 147, which was further reacted with glycidol 2a to give 148.
Applsci 11 05024 sch073
Scheme 74. Alkoxylation of bisphenols S and A by propylene carbonate.
Scheme 74. Alkoxylation of bisphenols S and A by propylene carbonate.
Applsci 11 05024 sch074
Scheme 75. Synthesis of liquid crystalline molecules bearing a cyclic carbonate endpoint 160.
Scheme 75. Synthesis of liquid crystalline molecules bearing a cyclic carbonate endpoint 160.
Applsci 11 05024 sch075
Scheme 76. Synthesis of the non-fluorinated analogue 161.
Scheme 76. Synthesis of the non-fluorinated analogue 161.
Applsci 11 05024 sch076
Scheme 77. Synthesis of urethanes 168ab.
Scheme 77. Synthesis of urethanes 168ab.
Applsci 11 05024 sch077
Scheme 78. Iridium-catalyzed intermolecular decarboxylative coupling of vinylated carbonate 1e with readily available isatoic anhydrides 170.
Scheme 78. Iridium-catalyzed intermolecular decarboxylative coupling of vinylated carbonate 1e with readily available isatoic anhydrides 170.
Applsci 11 05024 sch078
Scheme 79. The proposed reaction mechanism.
Scheme 79. The proposed reaction mechanism.
Applsci 11 05024 sch079
Scheme 80. The transesterification reaction of canola oil with dimethyl carbonate.
Scheme 80. The transesterification reaction of canola oil with dimethyl carbonate.
Applsci 11 05024 sch080
Scheme 81. Prepared glycerol-carbonate-derived fatty esters.
Scheme 81. Prepared glycerol-carbonate-derived fatty esters.
Applsci 11 05024 sch081
Scheme 82. Esterification approach for GC 1a with stearic acid 61a and ODOMS 1r conversion into GMS 12f.
Scheme 82. Esterification approach for GC 1a with stearic acid 61a and ODOMS 1r conversion into GMS 12f.
Applsci 11 05024 sch082
Scheme 83. Synthesis of cyclocarbonate- and triazole-functionalized fluorinated terpolymer 176.
Scheme 83. Synthesis of cyclocarbonate- and triazole-functionalized fluorinated terpolymer 176.
Applsci 11 05024 sch083
Scheme 84. Preparation of polyester resins. Conditions for i: 180 (0.330 mol), Et4NI (0.005 mol) and 1a (1.1 mol) were reacted in a 1 L Parr 4020 reactor at 150 °C under N2 atmosphere for 2 h; then, more acid 180 (0.670 mol) was added and the reaction was maintained at 150 °C for 8 h (>99% conversion).
Scheme 84. Preparation of polyester resins. Conditions for i: 180 (0.330 mol), Et4NI (0.005 mol) and 1a (1.1 mol) were reacted in a 1 L Parr 4020 reactor at 150 °C under N2 atmosphere for 2 h; then, more acid 180 (0.670 mol) was added and the reaction was maintained at 150 °C for 8 h (>99% conversion).
Applsci 11 05024 sch084
Scheme 85. Preparation of highly transparent polyacrylates.
Scheme 85. Preparation of highly transparent polyacrylates.
Applsci 11 05024 sch085
Scheme 86. Synthetic route involving GC methacrylate 1s and glycol bis-methacrylate 189 for an FRP preparation of monolithic columns.
Scheme 86. Synthetic route involving GC methacrylate 1s and glycol bis-methacrylate 189 for an FRP preparation of monolithic columns.
Applsci 11 05024 sch086
Table 1. Synthesis of alcohols and ethers compounds from cyclic carbonate 6.
Table 1. Synthesis of alcohols and ethers compounds from cyclic carbonate 6.
Applsci 11 05024 i001
EntryConditionRProductYield (%)
1B2pin2 22 (1.2 equivalent)
CuCl (9 mol%)/
L1 (13 mol%), Cs2CO3 (15 mol%), CH3OH, r.t., 16 h
aryl Applsci 11 05024 i00233–65
2Bpin(CH2)Bpin 23, (1.2 equivalent)
CuCl (9 mol%), Cs2CO3 (50 mol%), CH3OH, r.t., 16 h
aryl, heteroaryl, vinyl Applsci 11 05024 i00348–82
3B2pin2 22 (1.2 equivalent)
CuCl (9 mol%)/
L2 (13 mol%), Cs2CO3 (15 mol%), CH3OH, r.t., 16 h
alkyl, aryl, heteroaryl Applsci 11 05024 i00437–55
4Pd2(dba)3.CHCl3 (2.5 mol%)
C6H5B(OH)2 24 (20 mol%), (R)-L3 (10 mol%), THF
H2O (10 equivalent), 16 h, 40 °C
aryl, heteroaryl Applsci 11 05024 i00582–98%
(80–98% ee)
5Pd2(dba)3.CHCl3 (2.5 mol%)
B(C2H5)3 25 (10 mol%), (S,S,S)-L4 (10 mol%), THF
H2O (10 equivalent), 24 h, 20 °C
alkyl, benzyl, heteroalkyl Applsci 11 05024 i00661–87%
(80–98% ee)
6R1OH 7, Pd2(dba)3.CHCl3 (2.5 mol%), B(C2H5)3 25 (5 mol%), (R)-L3 (10 mol%), 40 °C
toluene, 4Å MS, 16 h
R = alkyl, aryl, heteroaryl, heteroalkyl
R1 = alkyl, benzyl, heteroalkyl, allyl
Applsci 11 05024 i00756–97%
(37–99% ee)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rollin, P.; Soares, L.K.; Barcellos, A.M.; Araujo, D.R.; Lenardão, E.J.; Jacob, R.G.; Perin, G. Five-Membered Cyclic Carbonates: Versatility for Applications in Organic Synthesis, Pharmaceutical, and Materials Sciences. Appl. Sci. 2021, 11, 5024. https://doi.org/10.3390/app11115024

AMA Style

Rollin P, Soares LK, Barcellos AM, Araujo DR, Lenardão EJ, Jacob RG, Perin G. Five-Membered Cyclic Carbonates: Versatility for Applications in Organic Synthesis, Pharmaceutical, and Materials Sciences. Applied Sciences. 2021; 11(11):5024. https://doi.org/10.3390/app11115024

Chicago/Turabian Style

Rollin, Patrick, Liane K. Soares, Angelita M. Barcellos, Daniela R. Araujo, Eder J. Lenardão, Raquel G. Jacob, and Gelson Perin. 2021. "Five-Membered Cyclic Carbonates: Versatility for Applications in Organic Synthesis, Pharmaceutical, and Materials Sciences" Applied Sciences 11, no. 11: 5024. https://doi.org/10.3390/app11115024

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