Next Article in Journal
Reduced Graphene Oxide Supported Cobalt-Calcium Phosphate Composite for Electrochemical Water Oxidation
Previous Article in Journal
Propylene Polymerization and Deactivation Processes with Isoselective {Cp/Flu} Zirconocene Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Terpolymerization of CO2 with Epoxides and Cyclic Organic Anhydrides or Cyclic Esters

by
David Hermann Lamparelli
and
Carmine Capacchione
*
Dipartimento di Chimica e Biologia “A. Zambelli”, Università degli Studi di Salerno, Via Giovanni Paolo II n°132, 84184 Fisciano, Italy
*
Author to whom correspondence should be addressed.
Submission received: 20 July 2021 / Revised: 31 July 2021 / Accepted: 9 August 2021 / Published: 11 August 2021

Abstract

:
The synthesis of polymeric materials starting from CO2 as a feedstock is an active task of research. In particular, the copolymerization of CO2 with epoxides via ring-opening copolymerization (ROCOP) offers a simple, efficient route to synthesize aliphatic polycarbonates (APC). In many cases, APC display poor physical and chemical properties, limiting their range of application. The terpolymerization of CO2 with epoxides and organic anhydrides or cyclic esters offers the possibility, combining the ROCOP with ring-opening polymerization (ROP), to access a wide range of materials containing polycarbonate and polyester segments along the polymer chain, showing enhanced properties with respect to the simple APC. This review will cover the last advancements in the field, evidencing the crucial role of the catalytic system in determining the microstructural features of the final polymer.

1. Introduction

The pervasiveness of polymers in our daily life is a consolidated reality in recent decades. Indeed, the reason for the fortune of polymeric materials is due to both their unique physical and chemical properties and their cheapness compared to other structural materials. In the last few years, however, there is a growing body of evidence that the large success of these materials has determined a negative effect on terrestrial and marine ecosystems with the presence of microplastics that have become ubiquitous on our planet [1]. This situation has engendered the growing attention of the polymer industries and the scientific community to find biodegradable, more sustainable polymeric materials. Parallel to this trend, the use of CO2 as a carbon feedstock has also gained momentum due to the rising interest in using such an inexpensive, non-toxic molecule as a starting material for the synthesis of polymers [2,3].
In particular, the alternating ring-opening copolymerization (ROCOP) of CO2 with epoxides has offered a conceptually simple route for the synthesis of aliphatic polycarbonates (APC), which show a clear advantage in terms of biodegradability with respect to polyolefins [4,5,6]. APC, however, often display poor chemical and mechanical properties compared to aromatic polycarbonates, and the incorporation of epoxides with various structural features does not always result in an improvement in the final properties of APC [7,8,9]. Aliphatic polyesters are another important class of biopolymers that can be conveniently obtained by the ring-opening polymerization (ROP) of cyclic esters [10,11,12,13] or by the ROCOP of epoxides with cyclic anhydrides [14,15,16,17]. Transition metal complexes generally catalyze all these polymerization processes through a coordination–insertion mechanism. Actually, taking a closer look at the polymerization mechanism of the ROCOP of epoxides with CO2 or cyclic anhydrides depicted in Scheme 1, it is evident that the formation of the metal-alkoxo bond (a’ in Scheme 1) is a key intermediate in the propagation process. In analogy, the ROP of cyclic esters proceeds via the formation of a metal-alkoxo bond (a in Scheme 1) that allows the ring-opening of the following monomer unit.
Given this mechanistic scenario, it is easy to imagine that it is possible to design a metal complex able to promote both types of polymerization, allowing us to obtain various block-copolymers. In principle, it is possible to obtain copolymers with polycarbonate and polyester segments and modulate the nature of the polycarbonate and polyester blocks, permitting the synthesis of new materials with tailored properties.
Notwithstanding the potential of such an approach, the efforts to develop efficient catalytic systems that cannot only incorporate CO2 but also give rise to unprecedented new materials have raised good results only recently.
This review will cover the last advancements (since 2003) in the metal-catalyzed and metal-free terpolymerization of CO2 with epoxides and cyclic esters or cyclic organic anhydrides for the obtaining of polycarbonate–polyester copolymers.

2. Terpolymerization of CO2 with Epoxides and Cyclic Anhydrides

Polymeric materials containing ester and carbonate linkages have shown potential as biodegradable implants and, in addition, it is possible to adjust the degradation rate by regulating the length of the polyester and polycarbonate blocks. The first approach to synthesize poly(ester-block-carbonate)s in a one-pot reaction was the copolymerization, via ROP, of cyclic esters and cyclic carbonates promoted by stannous octanoate [18]. This potentally straightforwrd way has some limitations because six or seven-membered cyclic carbonates suitable for ROP must be synthesized through time-consuming multistep protocols [19,20].

2.1. Zinc Complexes

Only in 2006, Liu and coworkers reported on the terpolymerization of propylene oxide (PO) with CO2 and maleic anhydride (MA) catalyzed by polymer-supported bimetallic catalyst (PBM) 1 of general formula P-Zn[Fe(CN)6]aCl2-3a(H2O)b, with P being the polyether type chelating agent and a ≈ 0.5 and b ≈ 0.76 [21]. Notably, the catalyst was inactive in the PO/MA copolymerization whereas it gave the poly(propylenecarbonate) (PPC) from PO/CO2. In the terpolymerization experiments (pCO2 = 4.0 MPa, T = 60 °C, t = 24 h), the polymer yield increased up to a 5:3 ratio between PO and MA, and a further increase in the MA content was detrimental for the polymerization activity. 1H and 13C NMR and IR spectroscopy revealed a random microstructure characterizing the resulting copolymers (Figure 1). The DSC thermograms showed a single transition with Tg values (29.1–56.1 °C) increasing by increasing the MA content in agreement with the microstructure revealed by NMR spectroscopy.
Later on, in 2008, the development of an efficient homogeneous catalytic system by Coates based on β-diiminate (bdi) zinc complex 2 (Figure 2) allowed the synthesis of a poly(ester-block-carbonate) by the terpolymerization (pCO2 = 0.3–5.4 MPa, T = 50 °C, t = 0.2–3 h) of cyclohexeneoxide (CHO) with diglycolic anhydride (DGA) and CO2 [22].
Interestingly, in spite of the polymerization feed simultaneously containing all three monomers, the final polymer was a diblock copolymer with a polyester block followed by a poly(cyclohexenecarbonte) (PCHC) block. Indeed, by following the polymerization reaction by in situ IR spectroscopy, the exclusive formation of the polyester block up to the total consumption of DGA was evident. From the mechanistic point of view, it was clear that the first step was the formation of zinc alkoxide by the ring-opening of CHO followed by the preferential and irreversible insertion of DGA until this monomer was completely consumed and only after the zinc alkoxide can allow the insertion of CO2 with the growth of the polycarbonate block. The polymerization of succinic anhydride (SA), albeit with lower reactivity, and vinyl-CHO was also accomplished.
In 2010, Zhang and coworkers showed that the heterogeneous double metal cyanide complex (DMCC) 3 obtained by the reaction of K3Co(CN)6 with ZnCl2 promotes the terpolymerization of CO2 with CHO and MA [23]. The proposed active site for this catalyst is a Zn atom in a tetrahedral structure with the Co atom playing a spectator role. The catalyst was highly active and selective, giving a complete conversion of CHO (pCO2 = 4.0 MPa, T = 90 °C, t = 5 h) and a complete selectivity toward the polymeric product. The resulting polymer was poly(ester-block-carbonate), but it also contained a variable amount of polyether linkages (2.9–11.7%) depending on the reaction conditions. In particular, the use of THF as a solvent inhibits the formation of polyether linkages due to the coordination of the THF molecule to the Zn2+ center. The mechanism proposed for CO2/CHO/MA terpolymerization catalyzed by a Zn–Co(III) DMCC catalyst is depicted in Scheme 2.
Unusually, notwithstanding the heterogeneous nature of the catalyst, the dispersity was narrow (Ð = 1.4–1.7) and the Mn was up to 14.1 kg mol−1.
Zinc glutarate (ZnGA) 4 was also found to be a versatile catalyst for the terpolymerization of CO2 with PO and various cyclic anhydrides. Indeed, in 2014, Meng reported on the synthesis of PO/phthalic anhydride (PA)/CO2 copolymers (pCO2 = 5.0 MPa, T = 75 °C, t = 15 h) using toluene as a solvent [24]. Intruguingly, in this case the formation of polycarbonate is favored over the polyester formation; consequently, the resulting terpolymers (Mn up to 221 kg mol−1, Ð = 2.1–3.9) consist of a polycarbonate chain randomly interrupted by PA units (4.8–10.5%) and a lower amount of polyether linkages (2.7–7.1%). On one hand, the lower reactivity of PA vs. CO2 was explained with the slower insertion of the PA into the Zr-alkoxo bond; on the other hand, the increase in the polymer yield and Mn observed when PA is present in the feed was explained with a faster insertion of PO into the zinc benzoate growing chain with respect to the zinc carbonate chain (Scheme 3).
These opposite effects determine an ideal value for the PA/PO ratio giving the maximum activity and highest molecular weight, and this ratio in the feed was experimentally found to be 1:8. However, the observed decrease in the Mn could also be explained considering the presence of diacid impurities in the anhydride that acts as a chain-transfer agent.
DSC thermograms show that the introduction of an aromatic ter-monomer in the polymer sensibly enhances the Tg with respect to the corresponding polycarbonate with an increase of 6 °C incorporating 5.6% of PA.
The same catalytic system 4 was also used for the synthesis of pseudo-interpenetrating poly(propylenecarbonate) by the terpolymerization (pCO2 = 5.4 MPa, T = 70 °C, t = 36 h) of CO2 with PO and pyromellitic dianhydride (PMDA) up to 4% (in this case, Mn increased up to 862 kg mol−1), resulting in a noticeable improvement in the mechanical and thermal properties with respect to the corresponding polycarbonate [25].
More recently, Williams et al. described the synthesis of a new dinuclear zinc complex 5 (Figure 3) that promotes the terpolymerization of CHO/PA/CO2 (pCO2 = 3.0 MPa, T = 100 °C, t = 18 h) [26].
By monitoring the reaction after 2 h, it was evident that there was the exclusive formation of polyether linkages and no formation of PCHC, and after 18 h the formation of a poly(ester-block-carbonate) (Mn up to 7 kg mol−1, Ð = 1.20) was evident and confirmed by size-exclusion chromatography (SEC) analysis and diffusion-ordered spectroscopy (DOSY) experiments.
Similar results were also obtained by Castro-Osma and coworkers by using dinuclear zinc complexes 68 supported by heteroscorpionate ligands (Figure 4) [27].
The complexes 68 were active, in the presence of DMAP, in the terpolymerization of CHO/PA/CO2 (pCO2 = 4.0 MPa, T = 80 °C, t = 16 h), giving a poly(ester-block-carbonate) material (Mn up to 3.9 kg mol−1, Ð = 1.41–1.64).

2.2. Chromium and Cobalt Complexes

In 2011, Duchateu and coworkers reported on the terpolymerization (pCO2 = 5.0 MPa, T = 80 °C, t = 18 h) of CHO with CO2 and various anhydrides (SA, cyclopropane-1,2-dicarboxylic acid anhydride (CPrA), cyclopentane-1,2-dicarboxylic acid anhydride (CPA) or PA) promoted by two chromium complexes (Figure 5): tetraphenylporphyrinato chromium chloride 9 and salophen chromium chloride 10 (where salophen = N,N’-bis(3,5-di-tert-butylsalicylidenyl)-1,2-phenylenediamine) activated by DMAP (4-(N,N-dimethylamino) pyridine) [28].
In analogy to the polymerization process observed by Coates in the case of the (bdi)Zn complexes, the formation of the polyester is favored over the formation of the polycarbonate, resulting in the formation of a poly(ester-block-carbonate). The authors also showed that DSC of poly(ester-block-carbonate) is inconclusive in giving information about the blocky microstructure of the copolymer because the polyester and polycarbonate phases are completely miscible, giving a single value for the Tg. Furthermore, in the case of complex 10, the authors noticed that the presence of CO2 in the polymerization feed completely suppresses the formation of polyether linkages. In particular, by copolymerizing the equimolar amount of CHO and CPrA in the presence of CO2, the pure polyester was obtained, while without CO2 an amount of 15–30% of polyether linkages was observed. For all terpolymerizations, Mn (up to 19.2 kg mol−1) showed a linear correlation with conversion and the Ð was ≤1.6, indicating controlled behavior.
Soon after, Darensbourg, using a related (salan) CrCl complex 11 activated by PPNN3 in the terpolymerization of CHO/PA/CO2, observed similar results (Mn up to 18 kg mol−1, Ð = 1.07–1.13) [29]. In this case, the poly(ester-block-carbonate) showed two distinct Tg values (48 °C and 115 °C). Intriguingly, the major reactivity of the anhydride vis à vis CO2 was explained in terms of a slower ring-opening step of the metal-carbonate intermediate with the epoxide monomer instead of a faster insertion of the anhydride in the metal-alkoxo bond (Scheme 4).
Chromium(III) complex 12 (TPPCrCl, Figure 6) with a porphyrin ligand in combination with PPNCl was successfully used by Chisolm and coworkers for the terpolymerization of CO2/PO/SA (pCO2 = 4.0–5.0 MPa, T = 25 °C, t = 3–18 h) [30]. It is worth noting that the PPNCl/Cr ratio is crucial to avoid the formation of polyether linkages; indeed, when 0.5 equiv. of PPNCl have been used, the formation of polyether linkages is favored (up to 42%) over the polyester and polycarbonate linkages, whereas with 1.0 equiv. of PPNCl the amount of polyether linkages is drastically reduced (<2%). In analogy to other chromium systems, for this system the polyester formation is also faster than the polycarbonate one, leading to copolymers with a tapered/diblock microstructure. The authors attributed the higher reactivity of SA over CO2 to the higher solubility of the anhydride in the reaction medium.
A major breakthrough in this field was the use of the single component Co(III) complex 13 (Figure 7) tethering four quaternary ammonium salts [31].
This complex displayed one of the highest activities in the CO2/PO copolymerization, reaching TOF up to 16,000 h−1. In the presence of CO2/PO/PA, this complex also shows high reactivity with a total conversion of PO only after 3.0 h (pCO2 = 3.5 MPa, T = 80 °C) and a calculated TOF = 12,000 h−1. The resulting copolymers have a gradient poly(1,2-propylene carbonate-co-phthalate)s microstructure since, due to the highest reactivity of PA compared to CO2, the polymeric chains formed in the initial stages are richer in PA, but the consumption of this comonomer favors the formation of polycarbonate chains in the last stages. The resulting copolymers have a very narrow dispersity (Đ = 1.03–1.22) and high molecular weight (Mn up to 354 kg mol−1). As previously observed, the incorporation of PA in the polymeric chain enhances the thermal properties of the final polymer with respect to the corresponding PPC.
A dinuclear Cr(III) salen complex 14 (Figure 8) was reported by Lu and coworkers to promote, in the presence of 2 equiv. PPNCl, the terpolymerization of CO2/CHO/PA (pCO2 = 1 MPa, T = 80 °C, t = 0.5–6 h) [32]. In the first 2 h, the system only produced the polyester segment with no incorporation of CO2, and only after the total consumption of PA the polycarbonate block was formed, also giving, in this case, a diblock polymer. The produced copolymers have a very narrow dispersity (Ð = 1.19–1.22) and the molecular weight increases with the polymerization time (Mn up to 21.2 kg mol−1).

2.3. Metal-Free Catalysts

Since the discovery by Feng and coworkers that triethyl borane (TEB) in combination with onium halides or alkoxides promotes the formation of polycarbonates by coupling CO2 with PO or CHO, the efforts to extend the use of this metal-free system to the terpolymerization of CO2 with epoxides and anhydrides resulted in the synthesis of terpolymers having various microstructural features [33].
In 2020, Meng reported the quadripolymerization of CO2 with PA, PO and CHO in the presence of TEB and PPNCl, resulting in the formation of the copolymer (pCO2 = 1 MPa, T = 70 °C, t = 24–96 h) with good selectivity (94%) with respect to the cyclic product [34,35]. The microstructure of the resulting quadripolymer was clarified by 1H and 13C NMR showing the presence of four main blocks, i.e., poly(PA-alt-CHO), poly(PA-alt-PO), poly(propylene carbonate) (PPC), and poly(cyclohexene carbonate) (PCHC), and a very low amount of polyether linkages (<1%). In addition, in this case the formation of the polycarbonate segments only starts after the complete PA conversion and thus after the formation of the polyester segments. The resulting polymers display narrow dispersity (Ð = 1.14–1.21) and a high molecular weight (Mn up to 77 kg mol−1). It is worth reporting that the Tg can be easily tuned by regulating the feed ratio with a wide temperature range (Tg = 82–116 °C).
Afterward, Li and coworkers reported on the terpolymerization of CO2 with PA and CHO, in the presence of TEB and PPNCl (pCO2 = 0.1 MPa, T = 80 °C, t = 0.25–2 h) [36]. Additionally, in this case the polycarbonate block starts forming only after the complete consumption of PA in the feed, resulting in a poly(ester-b-carbonate) copolymer with little tapering, as shown by NMR spectra. The same catalytic system also allows the synthesis of poly(ester-b-carbonate) without tapering by sequential monomer addition. The resulting copolymers possess narrow dispersity (Ð = 1.09–1.15) and Mn up to 23.5 kg mol−1.
Lately, Feng obtained similar results by using TEB in combination with Bu4NN3 (pCO2 = 0.1 MPa, T = 60 °C, t = 0.75–18 h) for the terpolymerization of CO2 with PO and SA/PA. The PO/SA/CO2 terpolymerization clearly shows higher reactivity toward the oxoanion of SA over CO2, leading to the preferential formation of the polyester resulting in tapered poly(ester-b-carbonate) [37]. Only at a low concentration of SA (SA:PO = 1:20), a random poly(ester-co-carbonate) copolymer was obtained with 51% of polyester and 49% of carbonate. The PO/SA/CO2 terpolymerization leads to random copolymers also at PA:PO = 20:200, and only with the presence of 40% of PA in the feed the resulting copolymer displays a blocky nature. The terpolymerization of CHO/PA/CO2 activating TEB with PPNCl (pCO2 = 0.1 MPa, T = 80 °C, t = 17–18 h) shows the analogous behavior of PO preferentially producing copolymers with a random microstructure and blocky copolymers only with a high content of PA in the feed. Genuine poly(ester-b-carbonate)s can be obtained by sequential monomer addition both in the case of PA/PO and CHO/PA followed by feeding CO2. The Tg of the resulting copolymers can be tuned by regulating the PA content in the final copolymer with values ranging from 32.5 °C to 46.2 °C in the case of the PO/PA/CO2 copolymers and from 121 °C to 135.1 °C in the case of the CHO/PA/CO2 copolymers.
In Table 1, the results obtained in the CO2/epoxide/cyclic anhydrides’ terpolymerization discussed in this first part are summarized.

3. Terpolymerization of CO2 with Epoxides and Cyclic Esters

The synthesis of polyester-co-polycarbonate was also attempted by the terpolymerization of CO2 with epoxides and cyclic esters combining the ROCOP and ROP mechanisms [38]. This approach gives access to microstructures not accessible via ROCOP with organic anhydrides and has the advantage of using largely available monomers ε-caprolactone (CL), DL-lactide (LA) and β-butyrolactone (BBL) [14,16,36,38,39,40].

3.1. Zinc Complexes

ZnGA 4, obtained by the reaction of zinc oxide and glutaric acid, was active in the terpolymerization of CO2/PO/CL (pCO2 = 2.8 MPa, T = 60 °C, t = 40 h), resulting in high molecular weight polymers (Mn up to 27.5 kg mol−1) with narrow dispersity (Ð = 1.50–2.97) [41]. The catalytic activity decreases by increasing the content of CL beyond the 50% in mol in the feed. Notably, the system was inactive in the polymerization of CL alone and the production of cyclic carbonate contaminant was not observed. The 13C NMR analysis reveals a diblock microstructure with CL units directly linked to PC units and CL units in homosequences. Accordingly, the DSC thermograms display two transitions: one relative to the Tg of the PPC block (Tg = 5.4–17.7 °C) and the Tm of the PCL block (Tm = 51.0–57.2 °C). These polymers show excellent enzymatic biodegradability catalyzed by various lipases. The same catalytic system using glycidol terminated -PCL as a macromonomer produced the corresponding grafted copolymers in the presence of CO2/PO (pCO2 = 1.0 MPa, T = 60 °C, t = 6 h) [42].
In 2006, Doring reported the first example of the terpolymerization of CO2/CHO/LA by using zinc acetate complexes 1522 with aminoimidoacrylate (AIA) ligands (Figure 9) [43].
In order to obtain a terpolymer with an appreciable amount of polycarbonate linkages, an excess of CHO in the feed was necessary (CHO:LA = 3:1, pCO2 = 4.0 MPa, T = 90 °C, t = 16 h), giving high molecular weight polymers (Mn = 11.3–41.6 kg mol−1) with narrow dispersity (Ð = 1.09–1.96). The copolymers obtained by using L-LA instead of rac-LA show crystallinity with a melting point around 167 °C. The authors also reported the terpolymerization by using the (bdi) Zn catalysts developed by Coates (see Figure 2), showing, in this case, even a major tendency to incorporate a higher amount of polycarbonate linkages (up to 80%).
The polymer-supported bimetallic catalyst (PBM) 1 of general formula P-Zn[Fe(CN)6]aCl2-3a(H2O)b (a ≈ 0.5 and b ≈ 0.76) developed by Liu was also effective in the terpolymerization (pCO2 = 4.0 MPa, T = 50–90 °C, t = 16 h) of CO2/PO/CL, giving materials also containing polyether linkages [44].
A ternary system composed of Y(CCl3COO)3/ZnEt2/glycerin 23 was used by Xianhong and coworkers to synthesize (pCO2 = 4.0 MPa, T = 70 °C, t = 10 h) terpolymers CO2/PO/L-LA with a high molecular weight (Mn = 7.2–15.4 kg mol−1) and broad dispersity (Ð = 4.2–9.9), with the molecular weight increasing by decreasing the L-LA content in the feed [45]. It is worth noting that the presence of L-LA in the polymeric backbone even at a low content (2.4% mol) results in a considerable increase in the mechanical and thermal properties.
A major advance came in 2014 when Williams and coworkers reported that the dizinc complex 24 bearing a reduced Robson-type macrocyclic ligand promotes the ROCOP of CO2/CHO and the ROP of CL (pCO2 = 0.1 MPa, T = 80 °C, t = 2–21 h) only when activated by CHO, and intriguingly, a polymerization feed composed by a mixture of CO2/CHO/CHO only leads to the exclusive formation of PCHC (Scheme 5) [46]. Indeed, the synthesis of PCL-b-PCHC was only possible by sequential monomer addition by introducing CO2 after the consumption of CL in the presence of CHO or by reverse order completely removing CO2 after the formation of the PCHC block. The molecular weight of the resulting polymers was rather low (Mn up to 4.8 kg mol−1), with narrow dispersity (Ð = 1.38–1.49). This ability to selectively polymerize only one kind of monomer from a mixture and the ability to oscillate between the ROCOP and ROP mechanisms led to the definition of “switch catalysis” [47].
By performing the ROCOP of CO2/CHO, it was also possible to obtain a polycarbonate polyol (pCO2 = 0.1 MPa, T = 80 °C, t = 16–25 h) that, after removing CO2, can be used for the synthesis of ABA triblock copoly(caprolactone-b-cyclohexene carbonate-b-caprolactone) by adding CL. Remarkably, from the thermal behavior, it was also evident that the presence of the PCHC block disturbs or, at a higher percentage, suppresses the crystallinity of the PCL blocks, allowing the preparation of amorphous polymer films with good transparency [48].
The same catalytic system 24 was also used to obtain pentablock copolymers by alternating ROCOP (anhydrides/epoxide), ROP (lactone) and ROCOP (CO2/epoxide) by using various epoxides (CHO and VCHO), anhydrides (PA, NA), and DL (ε-decalactone). The resulting pentablock copolymers show a single Tg (from −35 to 20 °C), low molecular weight (10–16 kg mol−1) [49] and Ð = 1.06–1.16.
A more sophisticated technique was necessary to synthesize ABA block copolymers having poly(limonene-carbonate) (PLC) blocks because of the incapability of the dizinc complex to catalyze the polymerization of limonene oxide (LO) with CO2 [50]. In order to circumvent this problem, a dual catalytic system was used: (1) the dizinc complex 25 promotes, in the presence of 1,2-cyclohexane diol (CHD), the formation of a hydroxyl-telechelic PDL by the ROP of DL. This macroinitiator was then used, after the modification of the end groups for the synthesis of the PLC blocks, by using a second catalytic system based on the Al aminotriphenolate complex 26 developed by Kleij [51], as shown in Scheme 6.
The resulting biopolymers PLC-b-PDL-b-PLC have molar masses Mn spanning from 50.700 to 114.6 kg mol−1 and narrow dispersity (Ð = 1.38–1.49). The thermal and mechanical properties are superior compared to PLC, and these terpolymers show good chemical recyclability through depolymerization with the same dizinc catalyst affording the starting monomers.
Lately, a heterodinuclear Zn/Mg catalyst 27 (Figure 10) with the same ligand framework promoted the formation of ABA triblock copolymers by using DL with high activity [52].
In particular, by performing the ROP of DL a dihydroxyl telechelic PDL was obtained that, in the presence of CO2, undergoes the transformation into the ABA triblock copolymer PCHC-b-PDL-b-PCHC. The raw copolymers can incorporate a high amount of CO2 (up to 23%) and possess a high molecular weight (38.0–71.9 kg mol−1) with narrow dispersity (Ð = 1.07–1.16). These materials display a single Tg (from −44 to −50 °C), evidencing the amorphous nature of the blocks and their complete miscibility, and only the polymers with a higher content of PCHC (>50%) show a second transition at higher temperatures (81; 110 °C). These materials show promising thermal and mechanical properties compared to PCHC and the possibility to modulate them by regulating the length of the blocks in the final polymer, potentially giving a wide range of applications.
Rieger and coworkers were able, by using a (bdi)-zinc complex 28 (Scheme 7), to obtain copolymers by the terpolymerization of CO2/BBL/CHO [53]. In particular, also in this case the CO2 acts as a switching agent: (A) at pCO2 = 4.0 MPa, the polymerization proceeds with the exclusive production of PCHC and the formation of the poly(hydroxybutyrate) (PHB) only starts after releasing CO2 pressure, leading finally to a diblock copolymer PCHC-b-PHB. (B) In the absence of CO2, obviously, the system evolves to the formation of PHB and before to the total consumption of BBL feeding CO2 (pCO2 = 4.0 MPa) with the formation of a PCHC block, finally releasing the CO2 the “residual” BBL polymerizes, giving, at the end, an ABA triblock copolymer PCHC-b-PHB-b-PCHC. (C) By lowering the CO2 pressure to pCO2 = 0.3 MPa, the rates of the ROCOP and ROP processes are comparable and therefore a statistical copolymer was formed.
The copolymers’ molecular weights, in the case of the block copolymers, are higher (Mn = 77.0–166 kg mol−1) than those obtained in the case of statistical copolymers (Mn = 34.0–69.0 kg mol−1), in both cases showing narrow dispersity (Ð = 1.2–1.8). PCHC-b-PHB and PCHC-b-PHB-b-PCHC display two Tg values relative to the PHB and PCHC blocks, respectively (Tg1 = 1–2 °C and Tg2 = 116–118 °C), as a consequence of phase separation between the polycarbonate and polyester blocks, also confirmed by atom force microscopy (AFM). Conversely, the random copolymers display a single transition (Tg = 36–91 °C) that increases by increasing the amount of carbonate linkages in the polymer chain. Similar results were obtained with cyclopenteneoxide (CPO), but in this case the polymerization at a higher pressure (pCO2 = 4.0–5.0 MPa) results in the formation of a gradient copolymer rather than a diblock copolymer. The kinetic study evidenced a change in the reaction order with respect to CO2 with a zero order dependence at high pressure (between pCO2 = 0.5–1 MPa) and first-order at lower pressure (pCO2 < 0.5 MPa), indicating that under the latter conditions the insertion of CO2 became the rate-limiting step [54]. As expected, the incorporation of polyester segments in both the statistical and block copolymers leads to an improvement in the mechanical properties compared to the brittle PCHC with a decrease in the Young modulus and tensile strength and an increase in the elongation at break for polymers with a high molecular weight (>100 kg mol−1). Efforts to terpolymerize CO2/BBL/LO (limonene oxide) evidenced that due to the low ceiling temperature (60 °C) of the polylimonenecarbonate (PLC), the only way to obtain block copolymers is to first obtain the PHB block via the ROP of BBL and then feed CO2 for the formation of the PLC block. Actually, the PHB-b-PLC copolymers possess a high molecular weight (Mn up to 233 kg mol−1) and narrow dispersity (Ð = 1.23–1.39), showing two Tg = 1–3/26–133 °C. Statistical copolymers were also obtained by adjusting the CO2 pressure (pCO2 = 0.9 MPa), resulting in low conversion (up to 22% LO and 26% BBL in 22 h) and low molecular weight polymers (Mn = 9.0 kg mol−1).

3.2. Cobalt Complexes

Salen cobalt complexes are highly active catalysts in the ROCOP of CO2 with epoxides and therefore are, in principle, viable candidates for the terpolymerization of CO2 with epoxides and lactones. Unfortunately, these complexes are inactive in the ROP of cyclic esters, and consequently, the implementation of an active catalytic system for obtaining polycarbonate-b-polyester copolymers requires the use of multi-component systems able to synthesize the desired polymeric product.
An elegant strategy was developed by Darensbourg and Lu that used a combination of the bifunctional Co(III) salen complex 29 and DBU (1,8-diazabicyclo[5.4.0]undec-7-ene) (Figure 11) for the terpolymerization (pCO2 = 1.5 MPa, T = 25 °C, t = 2–6 h) of CO2/SO/LA [55].
Indeed, the cobalt complex 29 produced with high activity poly(styrene carbonate) (PSC) from CO2 and SO, after the complete consumption of SO, a given amount of H2O, which was added to the reaction mixture, resulting in the production of hydroxy-terminated PSC. Consequently, the polycarbonate formed, having the hydroxy chain-end, can act as a macroinitiator for the ROP of LA catalyzed by DBU. As a matter of fact, after adding two equivalent of H2O with respect to the cobalt catalyst and the removal of CO2, the addition of LA and DBU results in the production of the AB copolymer PSC-b-PLA with molecular weight Mn up to 17.2 kg mol−1 and narrow dispersity (Ð = 1.04–1.12). The copolymers obtained from rac-LA display a single Tg at lower values with respect to PSC (60–72 °C), and by polymerizing D-LA, the resulting diblock copolymers also display a Tm = 133–137 °C depending on the length of the PLA block.
The same authors also used this strategy to synthesize ABA block copolymers from CO2/PO/LA. They used the system composed of the salen Co(III) complex 29 activated by PPNY (Y = CF3COO) and DBU [56]. In this case, the obtaining of PLA-b-PPC-b-PLA was possible because of the addition of an excess of H2O (5–20 equiv. with respect to CO) to stop the CO2/PO copolymerization of a PPC with two α,ω hydroxy groups. Indeed, the presence of a hydroxy group on both chain-ends allows the growth of two PLA blocks, giving the desired PLA-b-PPC-b-PLA triblock copolymers. The molecular weight was rather low, Mn up to 20.1 kg mol−1, with narrow dispersity (Ð = 1.02–1.04). Additionally, in this case, on one hand the copolymers obtained from rac-LA displayed a single Tg at higher values with respect to PPC (43–44 °C), and on the other hand, by polymerizing D-LA, the resulting ABA copolymers also displayed a Tm = 110–128 °C depending on the length of the PLA blocks.
Later on, Pang and coworkers developed a ternary system composed by the dinuclear Co(II) and the Co(III) complexes with salen ligands (Figure 12) and PPNCl [57].
Indeed, the Co(II) complexes (30a32a) are active in the ROP of LA, and the Co(III) complexes (30b32b), in combination with PPNCl, are active in the ROCOP of CO2 with various epoxides (PO, CHO, SO). The terpolymerization was possible for the chain transfer between the two metal centers (Scheme 8).
By using 30a and 30b and PPNCl in an equimolar amount, the terpolymerization of CO2/LA/PO gives terpolymers, as revealed by 1H and 13C NMR analysis, possessing a multiblock microstructure with Mn up to 13.6 kg mol−1 and narrow dispersity (Ð = 1.19–1.47). Notably, the dispersity broadens in the absence of PPNCl (Ð = 3.15) and with two equivalents of PPNCl (Ð = 2.28), confirming the crucial role of the onium salt in the chain-transfer between the metal centers.
More recently, the same authors further developed this ternary system by changing the Co(II) and Co(III) complexes (30a and 30b), obtaining a more active system or combining a salen Co(III) complex with ZnGA and PPNCl [58].
Finally, in the presence of an enantiopure chiral salenCo(III) complex 33 (Figure 13) in combination with PPN-DNP (PPN = bis(triphenylphosphine)iminium, DNP = 2,4-dinitrophenoxide), Lu and coworkers also succeeded in producing CO2/CHO/BBL terpolymers with isotactic -PCHC blocks [59].
More in detail, when an equimolar amount of CHO and BBL is present in the feed, the terpolymerization proceeds smoothly with the good conversion of both monomers (pCO2 = 2 MPa, T = 40 °C, t = 2–4 h). The resulting copolymers display, in the 1H NMR spectra, the signals relative to the carbonate-ester linkages, indicating a multiblock structure. Molecular weights are rather low, Mn = 3.3–14.6 kg mol−1, with narrow dispersity (Ð = 1.19–1.44), and display thermal behavior with a Tm = 204–220 °C, evidencing the presence of stereoregular crystalline blocks along the polymer chain.
In Table 2, the main data relating to the terpolymerizations of CO2 with epoxides and cyclic esters discussed in this second part are summarized.

4. Conclusions

The possibility to terpolymerize CO2 with epoxides and other cyclic monomers (cyclic esters, organic anhydrides) offers not only a simple way to obtain a wide range of materials with unprecedented properties, but also the possibility to have such material in a completely sustainable way, combining CO2 with monomers originating from biomasses. The last decade has witnessed tremendous efforts in the development of efficient catalytic systems able to combine the ROP of cyclic esters and the ROCOP of CO2 or cyclic organic anhydrides with epoxides, allowing us to obtain polymers with various microstructural features spanning from statistical, to AB, ABA, and even more complex architectures. Notwithstanding these endeavors, however, fine control of the microstructure and the molecular weight is still a major challenge in the field. Furthermore, the number of metal centers active in the terpolymerization of CO2 with epoxides and cyclic esters of anhydrides is still limited, offering active catalysts only in the case of Zn, Cr and Co, and, only in the case of the terpolymerization with cyclic anhydrides, in the presence of metal-free borane-based catalysts.
Therefore, this review is not only an overview on the progress in the field, but also shows that there is a large space for further developments. More precisely, higher control over the polymer microstructure, an extension to a wider range of monomers and the development of new catalytic systems based on other metal centers to improve the activity and the control of the polymerization process will be highly desirable targets in future developments.

Author Contributions

Conceptualization, C.C.; data curation, C.C. and D.H.L.; writing—original draft preparation, C.C.; writing—review and editing, C.C. and D.H.L. Both authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jambeck, J.R.; Geyer, R.; Wilcox, C.; Siegler, T.R.; Perryman, M.; Andrady, A.; Narayan, R.; Law, K.L. Plastic Waste Inputs from Land into the Ocean. Science 2015, 347, 768–771. [Google Scholar] [CrossRef]
  2. Paradiso, V.; Capaccio, V.; Lamparelli, D.H.; Capacchione, C. Metal Complexes Bearing Sulfur-Containing Ligands as Catalysts in the Reaction of CO2 with Epoxides. Catalysts 2020, 10, 825. [Google Scholar] [CrossRef]
  3. Grignard, B.; Gennen, S.; Jérôme, C.; Kleij, A.W.; Detrembleur, C. Advances in the Use of CO2 as a Renewable Feedstock for the Synthesis of Polymers. Chem. Soc. Rev. 2019, 48, 4466–4514. [Google Scholar] [CrossRef]
  4. Kernbichl, S.; Rieger, B. Aliphatic Polycarbonates Derived from Epoxides and CO2. In Engineering Solutions for CO2 Conversion; Reina, T.R., Arellano-Garcia, H., Odriozola, J.A., Eds.; Wiley: Weinheim, Germany, 2021; pp. 385–406. ISBN 978-3-527-34652-3. [Google Scholar]
  5. Yu, W.; Maynard, E.; Chiaradia, V.; Arno, M.C.; Dove, A.P. Aliphatic Polycarbonates from Cyclic Carbonate Monomers and Their Application as Biomaterials. Chem. Rev. 2021. [Google Scholar] [CrossRef]
  6. Janeta, M.; Lis, T.; Szafert, S. Zinc Imine Polyhedral Oligomeric Silsesquioxane as a Quattro-Site Catalyst for the Synthesis of Cyclic Carbonates from Epoxides and Low-Pressure CO2. Chem. Eur. J. 2020, 26, 13686–13697. [Google Scholar] [CrossRef]
  7. Della Monica, F.; Paradiso, V.; Grassi, A.; Milione, S.; Cavallo, L.; Capacchione, C. A Novel [OSSO]-Type Chromium(III) Complex as a Versatile Catalyst for Copolymerization of Carbon Dioxide with Epoxides. Chem. Eur. J. 2020, 26, 5347–5353. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, Y.; Darensbourg, D.J. Carbon Dioxide-Based Functional Polycarbonates: Metal Catalyzed Copolymerization of CO2 and Epoxides. Coord. Chem. Rev. 2018, 372, 85–100. [Google Scholar] [CrossRef]
  9. Della Monica, F.; Kleij, A.W. From Terpenes to Sustainable and Functional Polymers. Polym. Chem. 2020, 11, 5109–5127. [Google Scholar] [CrossRef]
  10. Lecomte, P.; Jérôme, C. Recent Developments in Ring-Opening Polymerization of Lactones. In Synthetic Biodegradable Polymers; Rieger, B., Künkel, A., Coates, G.W., Reichardt, R., Dinjus, E., Zevaco, T.A., Eds.; Advances in Polymer Science; Springer: Berlin/Heidelberg, Germany, 2011; Volume 245, pp. 173–217. ISBN 978-3-642-27153-3. [Google Scholar]
  11. Santoro, O.; Zhang, X.; Redshaw, C. Synthesis of Biodegradable Polymers: A Review on the Use of Schiff-Base Metal Complexes as Catalysts for the Ring Opening Polymerization (ROP) of Cyclic Esters. Catalysts 2020, 10, 800. [Google Scholar] [CrossRef]
  12. Gao, J.; Zhu, D.; Zhang, W.; Solan, G.A.; Ma, Y.; Sun, W.-H. Recent Progress in the Application of Group 1, 2 & 13 Metal Complexes as Catalysts for the Ring Opening Polymerization of Cyclic Esters. Inorg. Chem. Front. 2019, 6, 2619–2652. [Google Scholar] [CrossRef]
  13. Di Iulio, C.; Jones, M.D.; Mahon, M.F.; Apperley, D.C. Zinc(II) Silsesquioxane Complexes and Their Application for the Ring-Opening Polymerization of Rac-Lactide. Inorg. Chem. 2010, 49, 10232–10234. [Google Scholar] [CrossRef]
  14. Paul, S.; Zhu, Y.; Romain, C.; Brooks, R.; Saini, P.K.; Williams, C.K. Ring-Opening Copolymerization (ROCOP): Synthesis and Properties of Polyesters and Polycarbonates. Chem. Commun. 2015, 51, 6459–6479. [Google Scholar] [CrossRef] [Green Version]
  15. Longo, J.M.; Sanford, M.J.; Coates, G.W. Ring-Opening Copolymerization of Epoxides and Cyclic Anhydrides with Discrete Metal Complexes: Structure–Property Relationships. Chem. Rev. 2016, 116, 15167–15197. [Google Scholar] [CrossRef]
  16. Liang, J.; Ye, S.; Wang, S.; Xiao, M.; Meng, Y. Design and Structure of Catalysts: Syntheses of Carbon Dioxide-Based Copolymers with Cyclic Anhydrides and/or Cyclic Esters. Polym. J. 2021, 53, 3–27. [Google Scholar] [CrossRef]
  17. Ryzhakov, D.; Printz, G.; Jacques, B.; Messaoudi, S.; Dumas, F.; Dagorne, S.; Le Bideau, F. Organo-Catalyzed/Initiated Ring Opening Co-Polymerization of Cyclic Anhydrides and Epoxides: An Emerging Story. Polym. Chem. 2021, 12, 2932–2946. [Google Scholar] [CrossRef]
  18. Li, Y.; Hong, J.; Wei, R.; Zhang, Y.; Tong, Z.; Zhang, X.; Du, B.; Xu, J.; Fan, Z. Highly Efficient One-Pot/One-Step Synthesis of Multiblock Copolymers from Three-Component Polymerization of Carbon Dioxide, Epoxide and Lactone. Chem. Sci. 2015, 6, 1530–1536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Gregory, G.L.; Ulmann, M.; Buchard, A. Synthesis of 6-Membered Cyclic Carbonates from 1,3-Diols and Low CO2 Pressure: A Novel Mild Strategy to Replace Phosgene Reagents. RSC Adv. 2015, 5, 39404–39408. [Google Scholar] [CrossRef] [Green Version]
  20. McGuire, T.M.; López-Vidal, E.M.; Gregory, G.L.; Buchard, A. Synthesis of 5- to 8-Membered Cyclic Carbonates from Diols and CO2: A One-Step, Atmospheric Pressure and Ambient Temperature Procedure. J. CO2 Util. 2018, 27, 283–288. [Google Scholar] [CrossRef]
  21. Liu, Y.; Huang, K.; Peng, D.; Wu, H. Synthesis, Characterization and Hydrolysis of an Aliphatic Polycarbonate by Terpolymerization of Carbon Dioxide, Propylene Oxide and Maleic Anhydride. Polymer 2006, 47, 8453–8461. [Google Scholar] [CrossRef]
  22. Jeske, R.C.; Rowley, J.M.; Coates, G.W. Pre-Rate-Determining Selectivity in the Terpolymerization of Epoxides, Cyclic Anhydrides, and CO2: A One-Step Route to Diblock Copolymers. Angew. Chem. Int. Ed. 2008, 47, 6041–6044. [Google Scholar] [CrossRef]
  23. Sun, X.-K.; Zhang, X.-H.; Chen, S.; Du, B.-Y.; Wang, Q.; Fan, Z.-Q.; Qi, G.-R. One-Pot Terpolymerization of CO2, Cyclohexene Oxide and Maleic Anhydride Using a Highly Active Heterogeneous Double Metal Cyanide Complex Catalyst. Polymer 2010, 51, 5719–5725. [Google Scholar] [CrossRef]
  24. Liu, Y.; Xiao, M.; Wang, S.; Xia, L.; Hang, D.; Cui, G.; Meng, Y. Mechanism Studies of Terpolymerization of Phthalic Anhydride, Propylene Epoxide, and Carbon Dioxide Catalyzed by ZnGA. RSC Adv. 2014, 4, 9503–9508. [Google Scholar] [CrossRef]
  25. Gao, L.; Feng, J. A One-Step Strategy for Thermally and Mechanically Reinforced Pseudo-Interpenetrating Poly(Propylene Carbonate) Networks by Terpolymerization of CO2, Propylene Oxide and Pyromellitic Dianhydride. J. Mater. Chem. A 2013, 1, 3556. [Google Scholar] [CrossRef]
  26. Thevenon, A.; Garden, J.A.; White, A.J.P.; Williams, C.K. Dinuclear Zinc Salen Catalysts for the Ring Opening Copolymerization of Epoxides and Carbon Dioxide or Anhydrides. Inorg. Chem. 2015, 54, 11906–11915. [Google Scholar] [CrossRef] [Green Version]
  27. de la Cruz-Martínez, F.; Martínez de Sarasa Buchaca, M.; Martínez, J.; Tejeda, J.; Fernández-Baeza, J.; Alonso-Moreno, C.; Rodríguez, A.M.; Castro-Osma, J.A.; Lara-Sánchez, A. Bimetallic Zinc Catalysts for Ring-Opening Copolymerization Processes. Inorg. Chem. 2020, 59, 8412–8423. [Google Scholar] [CrossRef]
  28. Huijser, S.; HosseiniNejad, E.; Sablong, R.; de Jong, C.; Koning, C.E.; Duchateau, R. Ring-Opening Co- and Terpolymerization of an Alicyclic Oxirane with Carboxylic Acid Anhydrides and CO2 in the Presence of Chromium Porphyrinato and Salen Catalysts. Macromolecules 2011, 44, 1132–1139. [Google Scholar] [CrossRef]
  29. Darensbourg, D.J.; Poland, R.R.; Escobedo, C. Kinetic Studies of the Alternating Copolymerization of Cyclic Acid Anhydrides and Epoxides, and the Terpolymerization of Cyclic Acid Anhydrides, Epoxides, and CO2 Catalyzed by (Salen)CrIIICl. Macromolecules 2012, 45, 2242–2248. [Google Scholar] [CrossRef]
  30. Bernard, A.; Chatterjee, C.; Chisholm, M.H. The Influence of the Metal (Al, Cr and Co) and the Substituents of the Porphyrin in Controlling the Reactions Involved in the Copolymerization of Propylene Oxide and Cyclic Anhydrides by Porphyrin Metal(III) Complexes. Polymer 2013, 54, 2639–2646. [Google Scholar] [CrossRef]
  31. Jeon, J.Y.; Eo, S.C.; Varghese, J.K.; Lee, B.Y. Copolymerization and Terpolymerization of Carbon Dioxide/Propylene Oxide/Phthalic Anhydride Using a (Salen)Co(III) Complex Tethering Four Quaternary Ammonium Salts. Beilstein J. Org. Chem. 2014, 10, 1787–1795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Liu, Y.; Guo, J.-Z.; Lu, H.-W.; Wang, H.-B.; Lu, X.-B. Making Various Degradable Polymers from Epoxides Using a Versatile Dinuclear Chromium Catalyst. Macromolecules 2018, 51, 771–778. [Google Scholar] [CrossRef]
  33. Zhang, D.; Boopathi, S.K.; Hadjichristidis, N.; Gnanou, Y.; Feng, X. Metal-Free Alternating Copolymerization of CO2 with Epoxides: Fulfilling “Green” Synthesis and Activity. J. Am. Chem. Soc. 2016, 138, 11117–11120. [Google Scholar] [CrossRef] [PubMed]
  34. Ye, S.; Wang, W.; Liang, J.; Wang, S.; Xiao, M.; Meng, Y. Metal-Free Approach for a One-Pot Construction of Biodegradable Block Copolymers from Epoxides, Phthalic Anhydride, and CO2. ACS Sustain. Chem. Eng. 2020, 8, 17860–17867. [Google Scholar] [CrossRef]
  35. Liang, J.; Ye, S.; Wang, W.; Fan, C.; Wang, S.; Han, D.; Liu, W.; Cui, Y.; Hao, L.; Xiao, M.; et al. Performance Tailorable Terpolymers Synthesized from Carbon Dioxide, Phthalic Anhydride and Propylene Oxide Using Lewis Acid-Base Dual Catalysts. J. CO2 Util. 2021, 49, 101558. [Google Scholar] [CrossRef]
  36. Zhang, J.; Wang, L.; Liu, S.; Kang, X.; Li, Z. A Lewis Pair as Organocatalyst for One-Pot Synthesis of Block Copolymers from a Mixture of Epoxide, Anhydride, and CO2. Macromolecules 2021, 54, 763–772. [Google Scholar] [CrossRef]
  37. Chidara, V.K.; Boopathi, S.K.; Hadjichristidis, N.; Gnanou, Y.; Feng, X. Triethylborane-Assisted Synthesis of Random and Block Poly(Ester-Carbonate)s through One-Pot Terpolymerization of Epoxides, CO2, and Cyclic Anhydrides. Macromolecules 2021, 54, 2711–2719. [Google Scholar] [CrossRef]
  38. Li, Y.; Zhang, Y.-Y.; Hu, L.-F.; Zhang, X.-H.; Du, B.-Y.; Xu, J.-T. Carbon Dioxide-Based Copolymers with Various Architectures. Prog. Polym. Sci. 2018, 82, 120–157. [Google Scholar] [CrossRef]
  39. Koning, C.E.; Sablong, R.J.; Nejad, E.H.; Duchateau, R.; Buijsen, P. Novel Coating Resins Based on Polycarbonates and Poly(Ester-Co-Carbonate)s Made by Catalytic Chain Growth Polymerization of Epoxides with CO2 and with Anhydride/CO2. Prog. Org. Coat. 2013, 76, 1704–1711. [Google Scholar] [CrossRef]
  40. Liu, Y.; Wu, J.; Hu, X.; Zhu, N.; Guo, K. Advances, Challenges, and Opportunities of Poly (γ-Butyrolactone)-Based Recyclable Polymers. ACS Macro Lett. 2021, 10, 284–296. [Google Scholar] [CrossRef]
  41. Hwang, Y.; Jung, J.; Ree, M.; Kim, H. Terpolymerization of CO2 with Propylene Oxide and ε-Caprolactone Using Zinc Glutarate Catalyst. Macromolecules 2003, 36, 8210–8212. [Google Scholar] [CrossRef]
  42. Zhou, J.; Wang, W.; Villarroya, S.; Thurecht, K.J.; Howdle, S.M. Epoxy Functionalised Poly (ε-Caprolactone): Synthesis and Application. Chem. Commun. 2008, 44, 5806–5808. [Google Scholar] [CrossRef] [PubMed]
  43. Kröger, M.; Folli, C.; Walter, O.; Döring, M. Alternating Copolymerization of Carbon Dioxide and Cyclohexene Oxide and Their Terpolymerization with Lactide Catalyzed by Zinc Complexes of N,N Ligands. Adv. Synth. Catal. 2006, 348, 1908–1918. [Google Scholar] [CrossRef]
  44. Liu, S.; Xiao, H.; Huang, K.; Lu, L.; Huang, Q. Terpolymerization of Carbon Dioxide with Propylene Oxide and ε-Caprolactone: Synthesis, Characterization and Biodegradability. Polym. Bull. 2006, 56, 53–62. [Google Scholar] [CrossRef]
  45. Gu, L.; Qin, Y.; Gao, Y.; Wang, X.; Wang, F. One-Pot Terpolymerization of CO2, Propylene Oxide and Lactide Using Rare-Earth Ternary Catalyst. Chin. J. Chem. 2012, 30, 2121–2125. [Google Scholar] [CrossRef]
  46. Romain, C.; Williams, C.K. Chemoselective Polymerization Control: From Mixed-Monomer Feedstock to Copolymers. Angew. Chem. Int. Ed. 2014, 53, 1607–1610. [Google Scholar] [CrossRef] [Green Version]
  47. Stößer, T.; Chen, T.T.D.; Zhu, Y.; Williams, C.K. ‘Switch’ Catalysis: From Monomer Mixtures to Sequence-Controlled Block Copolymers. Philos. Trans. R. Soc. 2018, 376, 20170066. [Google Scholar] [CrossRef] [Green Version]
  48. Paul, S.; Romain, C.; Shaw, J.; Williams, C.K. Sequence Selective Polymerization Catalysis: A New Route to ABA Block Copoly (Ester-b-Carbonate-b-Ester). Macromolecules 2015, 48, 6047–6056. [Google Scholar] [CrossRef] [Green Version]
  49. Chen, T.T.D.; Zhu, Y.; Williams, C.K. Pentablock Copolymer from Tetracomponent Monomer Mixture Using a Switchable Dizinc Catalyst. Macromolecules 2018, 51, 5346–5351. [Google Scholar] [CrossRef] [Green Version]
  50. Carrodeguas, L.P.; Chen, T.T.D.; Gregory, G.L.; Sulley, G.S.; Williams, C.K. High Elasticity, Chemically Recyclable, Thermoplastics from Bio-Based Monomers: Carbon Dioxide, Limonene Oxide and ε-Decalactone. Green Chem. 2020, 22, 8298–8307. [Google Scholar] [CrossRef]
  51. Peña Carrodeguas, L.; González-Fabra, J.; Castro-Gómez, F.; Bo, C.; Kleij, A.W. AlIII-Catalysed Formation of Poly (Limonene) Carbonate: DFT Analysis of the Origin of Stereoregularity. Chem. Eur. J. 2015, 21, 6115–6122. [Google Scholar] [CrossRef]
  52. Sulley, G.S.; Gregory, G.L.; Chen, T.T.D.; Peña Carrodeguas, L.; Trott, G.; Santmarti, A.; Lee, K.-Y.; Terrill, N.J.; Williams, C.K. Switchable Catalysis Improves the Properties of CO2-Derived Polymers: Poly(Cyclohexene Carbonate-b-ε-Decalactone-b-Cyclohexene Carbonate) Adhesives, Elastomers, and Toughened Plastics. J. Am. Chem. Soc. 2020, 142, 4367–4378. [Google Scholar] [CrossRef] [Green Version]
  53. Kernbichl, S.; Reiter, M.; Adams, F.; Vagin, S.; Rieger, B. CO2-Controlled One-Pot Synthesis of AB, ABA Block, and Statistical Terpolymers from β-Butyrolactone, Epoxides, and CO2. J. Am. Chem. Soc. 2017, 139, 6787–6790. [Google Scholar] [CrossRef] [PubMed]
  54. Kernbichl, S.; Reiter, M.; Mock, J.; Rieger, B. Terpolymerization of β-Butyrolactone, Epoxides, and CO2: Chemoselective CO2-Switch and Its Impact on Kinetics and Material Properties. Macromolecules 2019, 52, 8476–8483. [Google Scholar] [CrossRef]
  55. Wu, G.-P.; Darensbourg, D.J.; Lu, X.-B. Tandem Metal-Coordination Copolymerization and Organocatalytic Ring-Opening Polymerization via Water to Synthesize Diblock Copolymers of Styrene Oxide/CO2 and Lactide. J. Am. Chem. Soc. 2012, 134, 17739–17745. [Google Scholar] [CrossRef] [PubMed]
  56. Darensbourg, D.J.; Wu, G.-P. A One-Pot Synthesis of a Triblock Copolymer from Propylene Oxide/Carbon Dioxide and Lactide: Intermediacy of Polyol Initiators. Angew. Chem. Int. Ed. 2013, 52, 10602–10606. [Google Scholar] [CrossRef] [PubMed]
  57. Li, X.; Hu, C.; Pang, X.; Duan, R.; Chen, X. One-Pot Copolymerization of Epoxides/Carbon Dioxide and Lactide Using a Ternary Catalyst System. Catal. Sci. Technol. 2018, 8, 6452–6457. [Google Scholar] [CrossRef]
  58. Li, X.; Duan, R.; Hu, C.; Pang, X.; Deng, M. Copolymerization of Lactide, Epoxides and Carbon Dioxide: A Highly Efficient Heterogeneous Ternary Catalyst System. Polym. Chem. 2021, 12, 1700–1706. [Google Scholar] [CrossRef]
  59. Li, W.-B.; Liu, Y.; Lu, X.-B. Bimetallic Cobalt Complex-Mediated Enantioselective Terpolymerizations of Carbon Dioxide, Cyclohexene Oxide, and β-Butyrolactone. Organometallics 2020, 39, 1628–1633. [Google Scholar] [CrossRef]
Scheme 1. Representative mechanisms of the copolymerization of epoxides, CO2, and cyclic esters (I) and the ROCOP of CO2 or cyclic anhydrides with epoxides (II).
Scheme 1. Representative mechanisms of the copolymerization of epoxides, CO2, and cyclic esters (I) and the ROCOP of CO2 or cyclic anhydrides with epoxides (II).
Catalysts 11 00961 sch001
Figure 1. The structure of CO2/PO/MA terpolymer obtained by Liu using complex 1 [21].
Figure 1. The structure of CO2/PO/MA terpolymer obtained by Liu using complex 1 [21].
Catalysts 11 00961 g001
Figure 2. Terpolymerization of CHO, DGA and CO2 with β-diiminate zinc complex 2 by Coates. Reproduced with modification and permission from ref. [22]. Copyright (2008) John Wiley and Sons.
Figure 2. Terpolymerization of CHO, DGA and CO2 with β-diiminate zinc complex 2 by Coates. Reproduced with modification and permission from ref. [22]. Copyright (2008) John Wiley and Sons.
Catalysts 11 00961 g002
Scheme 2. Proposed mechanism for CHO, MA and CO2 terpolymerization using Zn-Co(III) DMCC catalyst 3. Reproduced with modification and permission from ref. [23]. Copyright (2010) Elsevier.
Scheme 2. Proposed mechanism for CHO, MA and CO2 terpolymerization using Zn-Co(III) DMCC catalyst 3. Reproduced with modification and permission from ref. [23]. Copyright (2010) Elsevier.
Catalysts 11 00961 sch002
Scheme 3. Mechanism for terpolymerization of PA, PO and CO2 with ZnGA 4. Reproduced with modification and permission from ref. [24]. Copyright (2014) Royal Society of Chemistry.
Scheme 3. Mechanism for terpolymerization of PA, PO and CO2 with ZnGA 4. Reproduced with modification and permission from ref. [24]. Copyright (2014) Royal Society of Chemistry.
Catalysts 11 00961 sch003
Figure 3. Synthesis of dinuclear zinc complex 5 by Williams et al. Reproduced with modification and permission from ref [26]. Copyright (2015) American Chemical Society.
Figure 3. Synthesis of dinuclear zinc complex 5 by Williams et al. Reproduced with modification and permission from ref [26]. Copyright (2015) American Chemical Society.
Catalysts 11 00961 g003
Figure 4. Synthesis of bimetallic zinc acetate catalysts 68 supported by heteroscorpionate ligands [27].
Figure 4. Synthesis of bimetallic zinc acetate catalysts 68 supported by heteroscorpionate ligands [27].
Catalysts 11 00961 g004
Figure 5. Catalysts 9 and 10 used by Duchateu in the terpolymerization of CO2, CHO and dicarboxylic acid anhydrides (SA, CPrA, CPA, PA) [28].
Figure 5. Catalysts 9 and 10 used by Duchateu in the terpolymerization of CO2, CHO and dicarboxylic acid anhydrides (SA, CPrA, CPA, PA) [28].
Catalysts 11 00961 g005
Scheme 4. CO2 insertion vs. cyclic anhydride insertion into the metal alkoxide intermediate. Reproduced with modification and permission from ref [29]. Copyright (2012) American Chemical Society.
Scheme 4. CO2 insertion vs. cyclic anhydride insertion into the metal alkoxide intermediate. Reproduced with modification and permission from ref [29]. Copyright (2012) American Chemical Society.
Catalysts 11 00961 sch004
Figure 6. The (porphyrin)Cr(III)Cl complex 12 employed by Chisolm and coworkers [30].
Figure 6. The (porphyrin)Cr(III)Cl complex 12 employed by Chisolm and coworkers [30].
Catalysts 11 00961 g006
Figure 7. The (salen)Co(III) complex 13 with quaternary ammonium salts in a molecule [31].
Figure 7. The (salen)Co(III) complex 13 with quaternary ammonium salts in a molecule [31].
Catalysts 11 00961 g007
Figure 8. Dinuclear chromium catalyst 11 developed by Lu et al. [32].
Figure 8. Dinuclear chromium catalyst 11 developed by Lu et al. [32].
Catalysts 11 00961 g008
Figure 9. (AIA)Zn(OAc) complexes 1522 [43].
Figure 9. (AIA)Zn(OAc) complexes 1522 [43].
Catalysts 11 00961 g009
Scheme 5. The switch catalysis mechanism, ROCOP and ROP promoted by 24. Reproduced with modification and permission from ref [46]. Copyright (2014) John Wiley and Sons.
Scheme 5. The switch catalysis mechanism, ROCOP and ROP promoted by 24. Reproduced with modification and permission from ref [46]. Copyright (2014) John Wiley and Sons.
Catalysts 11 00961 sch005
Scheme 6. Block polymer synthesis using a dual catalytic system. Reproduced with modification and permission from ref. [50]. Copyright (2020) Royal Society of Chemistry.
Scheme 6. Block polymer synthesis using a dual catalytic system. Reproduced with modification and permission from ref. [50]. Copyright (2020) Royal Society of Chemistry.
Catalysts 11 00961 sch006
Figure 10. Heterodinuclear Zn/Mg catalyst 27 [52].
Figure 10. Heterodinuclear Zn/Mg catalyst 27 [52].
Catalysts 11 00961 g010
Scheme 7. Different reaction pathway of 28 toward ROP of BBL, copolymerization of CPO with CO2, and terpolymerization of CHO, CO2 and BBL. Reproduced with modification and permission from ref [53]. Copyright (2017) American Chemical Society.
Scheme 7. Different reaction pathway of 28 toward ROP of BBL, copolymerization of CPO with CO2, and terpolymerization of CHO, CO2 and BBL. Reproduced with modification and permission from ref [53]. Copyright (2017) American Chemical Society.
Catalysts 11 00961 sch007
Figure 11. Complex 29 and organocatalyst DBU [55].
Figure 11. Complex 29 and organocatalyst DBU [55].
Catalysts 11 00961 g011
Figure 12. Structures of salenCoII complexes (30a32a) and salenCoIII complexes (30b32b) [57].
Figure 12. Structures of salenCoII complexes (30a32a) and salenCoIII complexes (30b32b) [57].
Catalysts 11 00961 g012
Scheme 8. Copolymerization of LLA (L-Lactide), PO and CO2: the proposed cycle. Reproduced with modification and permission from ref [57]. Copyright (2018) Royal Society of Chemistry.
Scheme 8. Copolymerization of LLA (L-Lactide), PO and CO2: the proposed cycle. Reproduced with modification and permission from ref [57]. Copyright (2018) Royal Society of Chemistry.
Catalysts 11 00961 sch008
Figure 13. Enantiopure chiral dinuclear (S,S,S,S)-33 complex [59].
Figure 13. Enantiopure chiral dinuclear (S,S,S,S)-33 complex [59].
Catalysts 11 00961 g013
Table 1. Summary of terpolymerization of CO2 with epoxides and cyclic organic anhydrides.
Table 1. Summary of terpolymerization of CO2 with epoxides and cyclic organic anhydrides.
Catalyst MonomersPolymerization Conditions
pCO2, T, t
Mn * (Ð)
kg mol−1
Polymer MicrostructureRef.
1CO2, PO, MA4.0 MPa, 60 °C, 24 h-random[21]
2CO2, CHO, DGA0.3–5.4 MPa, 50 °C, 0.2–3 h37 (1.2–1.4)block[22]
3CO2, CHO, MA4.0 MPa, 90 °C, 5 h14.1 (1.4–1.7)block[23]
4CO2, PO, PA5.0 MPa, 75 °C, 15 h221 (2.1–3.9)random[24]
4CO2, PO, PMDA5.4 MPa, 70 °C, 36 h862 (2.0–3.8)random[25]
5CO2, CHO, PA3.0 MPa, 100 °C, 18 h7 (1.20)block[26]
68CO2, CHO, PA4.0 MPa, 80 °C, 16 h3.9 (1.41–1.64)block[27]
910CO2, CHO, SA/CPrA/CPA/PA5.0 MPa, 80 °C, 18 h19.2 (1.1–1.6)block[28]
11CO2, CHO, PA3.5 MPa, 80 °C, 12 h18 (1.07–1.13)block[29]
12CO2, PO, SA4–5.0 MPa, 25 °C, 3–18 h-tapered/block[30]
13CO2, PO, PA3.5 MPa, 80 °C, 3 h354 (1.03–1.22)gradient[31]
14CO2, CHO, PA1 MPa, 80 °C, 0.5–6 h21.2 (1.19–1.22)block[32]
Metal-free catalyst
TEB + PPNClCO2, PO, PA, CHO1 MPa, 70 °C, 24–96 h77.7 (1.14–1.21)alternated[34,35]
CO2, CHO, PA0.1 MPa, 80 °C, 0.25–2 h23.5 (1.09–1.15)block[36]
TEB + Bu4NN3CO2, PO, SA/PA0.1 MPa, 60 °C, 0.75–18 h17.3 (1.04–1.2)tapered/random[37]
TEB + PPNClCO2, CHO, SA/PA0.1 MPa, 80 °C, 16–17 h22.7 (1.06–1.09)tapered/block[37]
* The highest reported value.
Table 2. Summary of terpolymerization of CO2 with epoxides and cyclic esters.
Table 2. Summary of terpolymerization of CO2 with epoxides and cyclic esters.
CatalystMonomersPolymerization Conditions
pCO2, T, t
Mn * (Ð)
kg mol−1
Polymer MicrostructureRef.
1CO2, PO, CL4.0 MPa, 50–90 °C, 16 h-random[44]
4CO2, PO, CL2.8 MPa, 60 °C, 40 h27.5 (1.50–2.97)block[41]
4CO2, PO, CL1.0 MPa, 60 °C, 6 h10.8 (1.3–1.6)grafted[42]
1522CO2, CHO, LA4.0 MPa, 90 °C, 16 h41.6 (1.09–1.96)alternated/random[43]
23CO2, PO, L-LA4.0 MPa, 70 °C, 10 h15.4 (4.2–9.9)tapered/random[45]
24CO2, CHO, CL0.1 MPa, 80 °C, 2–21 h4.8 (1.38–1.49)block[46]
24CO2, CHO, CL0.1 MPa, 80 °C, 16–25 h13.8 (1.29–1.49)block[48]
24CO2, CHO, VCHO, PA/NA, CL0.1 MPa, 100 °C16 (1.06–1.16)block[49]
2526CO2, CHD, DL2 MPa, 40–100 °C, 2–24 h114 (1.38–1.49)block[50]
27CO2, CHO, DL2 MPa, 80 °C, 21 h71.9 (1.07–1.16)block[52]
28CO2, BBL, CHO/CPO0.3–4 MPa, 60 °C, 0.1–7 h166 (1.2–1.8)random/block[53]
28CO2, BBL, LO0.9–4 MPa, 40–60 °C, 8–22 h233 (1.23–1.39)random/block[54]
29CO2, SO, LA1.5 MPa, 25 °C, 2–6 h17.2 (1.04–1.12)block[55]
29CO2, PO, LA1.5 MPa, 25 °C, 1–4 h20.1 (1.02–1.04)block[56]
3032CO2, PO/CHO/SO, LA2 MPa, 60 °C, 4–48 h13.6 (1.19–3.15)block[57]
33CO2, CHO, BBL2.0 MPa, 40 °C, 2–4 h14.6 (1.19–1.44)block[58]
* The highest reported value.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lamparelli, D.H.; Capacchione, C. Terpolymerization of CO2 with Epoxides and Cyclic Organic Anhydrides or Cyclic Esters. Catalysts 2021, 11, 961. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11080961

AMA Style

Lamparelli DH, Capacchione C. Terpolymerization of CO2 with Epoxides and Cyclic Organic Anhydrides or Cyclic Esters. Catalysts. 2021; 11(8):961. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11080961

Chicago/Turabian Style

Lamparelli, David Hermann, and Carmine Capacchione. 2021. "Terpolymerization of CO2 with Epoxides and Cyclic Organic Anhydrides or Cyclic Esters" Catalysts 11, no. 8: 961. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11080961

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