Next Article in Journal
Development of a Binder-Free Tetra-Metallic Oxide Electrocatalyst for Efficient Oxygen Evolution Reaction
Previous Article in Journal
Formulation of the Polymeric Double Networks (DNs) for Biomedical Applications with Physicochemical Properties to Resemble a Biological Tissue
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in RO(CO)P of Bio-Based Monomers

Dipartimento di Biotecnologie e Scienze della Vita, Università degli Studi dell’Insubria, Via J. H. Dunant 3, 21100 Varese, Italy
*
Author to whom correspondence should be addressed.
Sustain. Chem. 2022, 3(2), 259-285; https://doi.org/10.3390/suschem3020017
Submission received: 8 April 2022 / Revised: 24 May 2022 / Accepted: 30 May 2022 / Published: 31 May 2022

Abstract

:
Synthetic polymers have a key role in modern society as they have allowed for great technological advancement since their discovery. However, the use of fossil-fuel-based raw materials and the pollution derived from plastics accumulation in the environment raised enormous concern, driving research efforts toward the identification of more sustainable alternatives. Bio-based functional molecules susceptible to ring-opening (co)polymerisation [RO(C)OP], such as lactones, cyclic carbonates, and oxiranes, represent an attractive source of monomers for the synthesis of more sustainable polymers. In this review, we describe the main advancement in this research field reported during the last seven years. In particular, we describe the preparation of monomers from (renewable) bio-sources such as sugars, terpenes, fatty acids, and carbon dioxide with a focus on structurally novel substrates. Both metal-mediated and organo-catalytic RO(CO)P methods are described, and the properties of derived functional polymers are discussed when relevant.

1. Introduction

Technological and material advancements have been used to classify certain periods of human history, from ancient stone, bronze, and iron ages to the closer in time atomic, space and information ages [1]. There can be no doubt that synthetic plastics represent one of the greatest technological advancements in the modern era. Unfortunately, their use is accompanied by environmental pollution at such a level that the presence of plastic in nonbioturbated sediment could be used as a stratigraphic indicator for the Anthropocene [2].
Environmental accumulation of plastics due to the high stability of most commonly used polymers represents a priority problem and, at the same time, the vast majority of currently used polymers are based on non-renewable, fossil resources [3]. In the attempt to find a solution, besides recycling, the upcycling of waste polymers is emerging as highly promising strategy for mitigating plastic pollution [4]. In parallel, great attention has been directed to the identification of new, renewable alternative feedstocks for the production of (more) sustainable polymers [5]. In particular, the creation of renewable monomers has been proposed from numerous bio-sources such as terpenes [6], sugars [7], vegetable oils [8], lignin [9], cellulose [10], and starch [11].
Different classes of bio-based polymers have been obtained by means of several polymerisation techniques such as: polyhydroxyalkanoates (PHBs) by fermentation [12], polyolefins by anionic [13], and coordination-insertion polymerisation [14], polyacrilates by (controlled) radical polymerisation [15], and poly(hydroxyurethane)s by polyaddition reactions [16].
Among other polymerisation techniques, ring-opening (co)polymerisation [RO(CO)P] has been largely investigated in the last decades, and numerous catalytic systems have been developed to promote this reaction [17,18,19,20,21,22,23,24]. In particular, systems based on environmentally benign metals such as aluminium, zinc, magnesium, and iron are attracting high interest. Depending on the chemical identity of the starting monomers, RO(CO)P gives access to the synthesis of polyethers, polyesters, and polycarbonates.
In this review, we summarize the major advancements in RO(CO)P application for the conversion of bio-based monomers. In particular, the description is divided on the basis of the polymer product identity. We mainly covered the period from 2015 to the beginning of 2022, to avoid overlapping with other exhaustive review papers.

2. Polyesters

One of the most appealing classes of sustainable polymers is represented by aliphatic polyesters (APEs), because of their general biocompatibility combined with a relatively easy hydrolytic degradation [25,26]. The synthesis of APEs can be conveniently performed via ROP of cyclic esters or ROCOP of cyclic anhydrides and epoxides (Scheme 1). Notwithstanding these methods have been deeply investigated, there is still a lack of suitable monomers from renewable resources. Hence, in this section, we divide the description of sustainable APEs into those obtained from bio-based lactones and those obtained from bio-based epoxides and cyclic anhydrides.

2.1. Polyesters from ROP of Lactones

Monomers deriving from renewable biomass feedstocks, such as terpenes, sugars, and fatty acids, have been recognized as ideal candidates for the development of sustainable materials (Figure 1) [6,27,28].
Although lactide (LA) and ε-caprolactone (CL) are undoubtedly the most studied of the series, their congeners such as δ-hexalactone (HA) and β-lactones (BLs) have been less investigated [29,30]. The corresponding polyesters are used in many fields, in particular for medicinal and pharmaceutical applications [31,32]. For this reason, important efforts have been dedicated to the development of catalytic systems based on non-toxic metals. Currently, the industrial production of polylactide (PLA) employs tin(II) octanoate (Sn(Oct)2) as the catalyst, also approved by the U.S. Food and Drug Administration (FDA) [33]. Nevertheless, alternative systems capable of outperforming such catalysts have been developed during the past two decades. In this scenario, complexes based on Al, Mg, and Zn exhibit outstanding catalytic performances and have been extensively reviewed [34,35,36].
Along with these metals, Fe-based catalysts represent ideal candidates; nonetheless, applications of such systems as ROP catalysts are scant. Indeed, after the early reports by Tolman [37,38], and Gibson [39,40], the field has remained largely unexplored for almost a decade. In 2013, Byers et al. reported on the ROP of lactide promoted by bis(imino)pyridine Fe(II) bis(alkoxide) complexes [41]. The catalyst activity proved highly dependent on the nature of the exogenous alcohol employed as the co-activator. The obtained PLAs had relatively narrow molecular weight distribution (dispersity, Đ, between 1.06 and 1.45), and the process exhibited a rather living character. Nevertheless, the systems proved poorly active, allowing for 60% conversion of 100 equivalents of monomer in 24 h. The analogous ferric compound was completely inactive, enabling the possibility of an on/off redox switchable process. Later, the same group disclosed analogue ferrous species bearing a bis(amidinato)-N-hetero-cyclic carbene ligand [42]. Compared to the previously reported bis(imino)pyridine-based congener, PLAs with higher Mn were achieved in shorter reaction times. This was accounted to faster activation rates granted by the more electron-donating NHC scaffold.
Tetradentate bis(imino)pyridine iron(II) complexes active in the ROP of lactide were lately reported by Herres-Pawlis et al. (Fe1–3, Figure 2) [43]. Such species allowed for high conversion (87%) of technical grade monomer within 30 h under solvent-free conditions, affording PLAs with a number average molecular weight (Mn) lower than the theoretical values and rather large dispersity (1.5 ≤ Đ ≤ 1.9). Two years later, the same group disclosed novel iron-guanidine complexes exhibiting superior catalytic activity in the ROP of lactide (Fe4–6, Figure 2) [44]. Remarkably, such species outperformed Sn(Oct)2 under industrially relevant conditions (i.e., technical grade monomer and bulk), affording high molecular weight PLA (Mn > 90 kDa) having thermal properties matching those reported in the literature. Analysis of the polymer chain ends indicated that, in the absence of an external co-activators (alcohol), the reaction could be initiated either by adventitious water or by the guanidine moiety of the complex.
In 2017, a Fe(η6-C6H6)-bis(arylimino)-acenaphthene (BIAN) system was reported by Long et al. (Fe7, Figure 3) [45].
Complex Fe7 was tested in the ROP of L-LA, affording high molecular weight PLLA (Mn up to 300 kDa) in the absence of exogeneous alcohol with moderate dispersity values (Đ = 1.7). Upon introducing 4-methoxyphenol as the co-activator, a drop of polymer Mn consistent with the concentration of alcohol was observed. In turn, 1H NMR spectroscopy analysis revealed the presence of 4-methoxyphenoxy-terminal groups, indicating the alcohol as the actual initiator of the process. Nevertheless, no clear evidence of in situ formation of iron alkoxide species was obtained. The actual polymerisation mechanism, as well as the role of the iron complex, could not be clarified due to the formation of paramagnetic species hampering in-depth NMR spectroscopy analyses.
Concomitantly, Pang et al. disclosed a series of air-stable Fe(III)-salen catalysts for the ROP of r-LA and CL (Fe8–15, Figure 4) [46,47].
Although no polymerisation was achieved in toluene, all complexes proved active upon using propylene oxide (PO) as the solvent. MALDI-ToF analysis of the polymers revealed chain terminations compatible with chloropropanol moieties, suggesting that the catalyst activation proceeds by ring-opening of the epoxide affording a Fe-alkoxide species. Remarkably, the presence of poly(propylene oxide) blocks was not observed, indicating that, aside from the activation step, the epoxide remains completely unreacted. This was in striking contrast with the reactivity observed with analogues Al-complexes. In addition, the Fe-compounds proved better performing than Al-systems with similar structures. This was accounted for by a stronger donor property of the Fe(III) centre. The activity of Fe9 was higher than that of Fe8, possibly because of the greater flexibility of the C3 bridging unit. On the other hand, all polymers exhibited comparable Mn and Ð. The unsubstituted complex Fe11 exhibited the lowest activity, while the presence of an electron-withdrawing substituent in Fe14–15 determined an improvement of the catalytic performance due to the reduced electronic density around the metal centre. In turn, tert-Bu substituted complexes Fe12–13 allowed for narrow polydispersity and moderate Mn (1.13 and 10.5 kDa, respectively). The obtained PLAs were somewhat isotactic (0.68 < Pm < 0.78) or heterotactic depending on the ligand structure. In particular, hindered aromatic moieties granted higher isotacticity. Extensive mechanistic studies indicated that the stereoselectivity of r-LA polymerisation proceeds according to a chain-end control mechanism.
Along with linear polyesters, polymers with cyclic structures are attractive materials due to their interesting properties and absence of terminations [48]. Although relevant examples are present in the literature [49,50,51] the synthesis of cyclic polyesters remains challenging. For example, Sn-based species allowed for the synthesis of cyclic poly(L-lactide) (cPLA) displaying Mn as high as 31 kDa, albeit with very broad dispersity (Ð > 6) [52,53]. Remarkably, the industrial catalyst Sn(Oct)2 also proved to produce cPLA with good activity under specific conditions [54].
In this scenario, Capacchione and co-workers showed that the bis-thioether-bis-phenolate [OSSO]-type Fe(III) complex Fe16 (Figure 5) is capable of producing cPLA at 80 °C with LA:Fe ratios as high as 10,000 in either PO or cyclohexene oxide (CHO) as the solvent [55]. Moreover, in this case, the epoxide acted as the co-activator of the polymerisation. In fact, the first step of the mechanism was proposed to be the activation of the epoxide by the Fe-catalyst, as previously observed with the same system [56,57]. The Mn’s of the polymers were as high as 6.4 kDa with rather narrow dispersity (1.16 < Ð <1.45). Interestingly, kinetic investigations indicated a zero-order dependence on the monomer concentration. The complex was also capable of affording cyclic poly(ε-caprolactone) (cPCL)and cyclic poly(β-butyrolactone) (cPBL).
Lately, Williams, Arnold et al. reported on the synthesis of cPLA promoted by N-heterocyclic carbene (NHC) Ce(III) complexes (Ce1–2, Figure 5) [58].
Complex Ce2, bearing a saturated NHC backbone, proved more active than its unsaturated congener Ce1. Nevertheless, both complexes exhibited high polymerisation rates at very low catalyst loading (200 ppm), allowing for the isolation of cPLA displaying Mn up to 250 kDa and rather narrow polydispersity (Ð = 1.59). In addition, CL and BL were also successfully converted into cyclic polyesters. Interestingly, liner polymers could be obtained upon introducing exogenous alcohol (iPrOH), albeit with a lack of control and lower Mn (73 kDa)
Homo- and hetero-multimetallic complexes have often exhibited higher activities in the ROP of LA and CL when compared to their monometallic counterparts [59,60]. Such difference is thought to arise from cooperative effects between the different metal centres. A comprehensive overview of such systems has been recently reported by Garden and Gruszka [61].
Polyesters derived from the homo/copolymerisation of larger lactones (i.e., ω-pentadecalactone, PDL), have been proposed as viable biodegradable alternatives to polyolefins. Indeed, the mechanical and thermal properties of poly(pentadecalactone) (PPDL) were found to be comparable to that of low-density polyethylene (LDPE) [62]. As for LA and CL, most of the catalytic systems employed in the ROP of PDL are based on Al- [63], Zn- [64], and Mg-complexes [65,66].
Very recently, a family of sodium complexes bearing a dibenzhydryl-substituted phenoxide and 15-crown-5-ether ligands exhibited interesting activity in the ROP of PDL (Na1–4, Figure 6) [67].
Such systems allowed for a complete conversion of 100 equiv. of monomer within 5 h in the presence benzhydrol as co-activator. In terms of catalytic activity, all complexes proved equally performing, affording polymers with Mn as high as 50 kDa and polydispersity values spanning from 1.69 to 2.57. The size of the crown-ether cavity was thought to allow only for the polymer chain growth, inhibiting undesired transesterification processes.
With respect to organocatalyzed routes, phosphazane superbases proved viable ROP promoters [68,69]. Indeed, PPDL having Mn matching the theoretical values has been obtained with moderate to good conversions even at room temperature in the presence of exogeneous alcohol (BnOH) as the co-activator. Interestingly, the melting temperatures of the polymers (spanning from 90 to 93 °C) were found close to that of liner PE.
A dual catalyst approach involving an organic base in combination with a Lewis acid has been developed by Dove et al. [70]. In this case, N-heterocyclic carbenes (NHC1–3), 1,8-diazabicycloundec-7-ene (DBU) and 4-dimethylaminopyridine (DMAP) were selected as the bases (Figure 7), while MgX2 (X = Cl, Br, I), FeCl3, ZnCl2, B(Ph)3, Bi(OTf)3, YCl3 and AlCl3 were employed as the Lewis acids. For the latter, the activity order was found to be Mg > Y > Al, while the other species proved completely unreactive. Such trend was tentatively explained considering the moderate solubility of such reactants in the reaction medium (toluene); nevertheless, the occurrence of processes leading to the saturation of the electronic vacancy of the Lewis acid in competition with monomer coordination was not excluded. In turn, MgI2 proved more active than its bromide- and chloride analogues, allegedly because of its cationic character resulting from the dissociation of one iodine ligand.
Concerning the organic base different NHCs, in combination with MgX2, allowed for similar reaction rates and monomer conversions. Nevertheless, the use of the more donating carbene NHC3 led to a broadening of the molecular weight distribution, indicating a lack of control. Finally, the combination of MgI2 and DBU exhibited the highest catalytic activity, allowing for a complete conversion of 200 equiv. of PDL within 30 min at 110 °C in the presence of BnOH as the co-activator, affording a polymer having Mn and Ð of 71 kDa and 1.80, respectively.

2.2. Polyesters from ROCOP of Epoxides and Cyclic Anhydrides

The ROCOP of epoxides and cyclic anhydrides is another effective approach to access polyesters [20,71]. Aside from benchmark monomers such as propylene oxide (PO), cyclohexene oxide (CHO) and phthalic anhydride (PA), terpene-anhydrides and epoxides have been gaining increasing attention (Figure 8). An example of fully terpene-based polyester was reported by Thomas et al. in 2011 [72]. Indeed, the ROCOP of camphoric anhydride (CA) and limonene oxide (LO) was achieved in the presence of Al-salen complex Al1 (Figure 9). Attempts to post-functionalize the polymer by olefin cross-metathesis were unsuccessful [73].
Tricyclic anhydrides AH1 and AH2, derived from α-pinene and α-phellandrene, respectively, have been employed in the ROCOP with PO to afford polyesters exhibiting high Tg [74]. Catalysts investigated in such a study were Cr-, Co- and Al-complexes bearing a salphen scaffold (Figure 9). Improved thermal properties (higher Tg values) have been achieved upon replacing PO with a more rigid epoxide such as CHO [75]. Reports concerning the use of terpene-derived epoxides (i.e., APO, LO, Figure 8b) remain scant since their reactivity is generally hampered by steric congestion.
Early examples of LO copolymerisation with phthalic anhydride catalysed by Cr1 required high temperatures [76]. In turn, the use of the Fe-aminotriphenolate complex Fe17 allowed for the copolymerisation of PA with various terpene-derived epoxides (Figure 8b) under mild reaction conditions, both in solution and in bulk [77].
In 2020, Della Monica and Kleij reported the ROCOP of new terpene-based epoxides obtained from β-elemene in the synthesis of highly functional semi-aromatic polyesters (Figure 10) [78]. This terpene, naturally occurring in citrus fruit, is usually extracted from ginger root and demonstrated to have anticancer activity [79]. Thanks to the different reactivity of the double bonds in β-elemene, it was possible to obtain the corresponding mono-, di- and tri-epoxides (BEM, BED, and BET respectively). In particular, BED was obtained as a regio-isomers mixture in a 7:3 composition. Copolymerisation of BEM with PA was investigated using complexes Fe17 and Al3, activated by (triphenylphosphine)-iminium chloride (PPNCl), obtaining poly(BEM-alt-PA) with 3.5 ≤ Mn ≤ 8 kDa and 1.13 ≤ Đ ≤ 1.28.
The functional polyester poly(BEM-alt-PA), bearing two pendant double bonds with different reactivity, was subject to post-functionalization via epoxidation, obtaining the corresponding poly(BED-alt-PA) and poly(BET-alt-PA) that cannot be prepared by direct reaction of BED or BET (Figure 9). Indeed, ROCOP of BED with PA was also performed under conditions similar to BEM, but an insoluble, crosslinked polyester was obtained, with a Tg of 125 °C. Notably, the post-synthetic epoxidation affects the thermal properties, with the Tg raising from 68 to 121 °C.
Fatty acids represent a highly attractive renewable feedstock for chemical industries [80] and became interesting also for the development of sustainable polymers [81,82]. For example, an efficient approach is the valorisation of epoxidized-products derived from fatty acids [83,84,85].
In this regard, very recently the ROCOP of epoxidized fatty acids esters for the synthesis of functional polyesters was also described [86]. First, copolymerisation of methyl oleate oxide (MOO) with PA was investigated, using combinations of Fe17 and Al3 with PPNCl and DMAP as the catalytic system, obtaining polyesters with 4 ≤ Mn ≤ 12 kDa and 1.15 ≤ Đ ≤ 1.19. Notably, the reaction was also possible using MOO based on oleic acid obtained from food industry waste. Structurally different polyesters were obtained starting from various epoxides/cyclic anhydrides combinations (Figure 11), and the flexible nature of the pendant side-chains resulted in materials with −45 ≤ Tg ≤ −5 °C.
In addition, two different methods were used for the creation of thermally stable, and thermo-reversible cross-linking networks. In the first case, the polymer based on allyl oleate oxide (AOO) was reacted with a bis-azide via Cu(I)-catalyzed azide−alkyne cycloaddition, obtaining an insoluble material with Tg of 37 °C, higher than that of the pristine polyester (Tg = −22 °C). In the second case, the polymer based on furfuryl oleate oxide (FOO) was reacted with bis-maleimide in Diels-Alder (DA) reaction. In this case, the crosslinked polyester also showed an increase in Tg (from −19 to 95 °C), however thermal curing at 160 °C for 16 h induced a retro-DA reaction giving PE with the same properties as the original sample.

3. Polycarbonates

Polycarbonates (PCs) are classified into aromatic PCs and aliphatic PCs. Commercial aromatic PCs are known as BPA-PCs, based on bis-phenol A (BPA) structural motif. In recent years, concern has arisen regarding BPA-PC because it degrades under environmental temperature and normal pH, releasing BPA which is known to act as an endocrine disruptor [87]. In addition, industrial PC production relies on the use of highly toxic phosgene. On the contrary, aliphatic PCs seem to represent a more sustainable alternative, due to their general biodegradability and biocompatibility [88,89]. However, most common aliphatic PCs, such as poly(propylene carbonate), suffer from scarce material properties limiting their possible applications [90]. In this context, the successful synthesis of aliphatic polycarbonates achieved either via direct ROP of cyclic carbonates (CCs) or ROCOP of epoxides and carbon dioxide is highly attractive (Scheme 2) [17,18,19,20,21,22,91,92].
Notwithstanding these methods are highly attractive, they both present some critical aspects to be considered when designing new monomers. In the case of CCs, it is important to notice that not all CCs can be polymerised, and the success of the reaction often depends on ring-size, substitution pattern and stereochemistry of the monomer. For example, it is well known that the trans-form of cyclohexene carbonate can be polymerised while the cis-isomer is inert toward polymerisation [93,94]. In the case of ROCOP, the epoxide nature also plays an important role in directing the reaction toward the formation of PC or CC [95]. In addition, the search for catalytic systems based on sustainable metals is still highly active since the most active systems producing PCs are based on metals such as Co and Cr. Hence, in this section, we divide the description of sustainable PCs into those obtained from bio-based CCs and those obtained from bio-based epoxides.

3.1. ROP of Cyclic Carbonates

The synthesis of bio-based cyclic carbonates for application in ROP is largely based on sugars, which are converted into cyclic carbonates thanks to the presence of multiple hydroxyl groups. Thanks to their large availability and biocompatibility, carbohydrates represent one of the more attractive feedstocks in sustainable polymer chemistry [96]. Indeed, they combine a wide structural diversity with the presence of oxygenated functional groups, ideally offering the possibility to design materials with enhanced material properties and biodegradability. In the last years, the research groups of Wooley and Buchard reported the synthesis of several sugar-based CC monomers for PC synthesis. Bicyclic and tricyclic structures were obtained, containing cyclic carbonates moieties with both cis-and trans-configuration. This results in good reactivity of such sugar-based monomers, which can be polymerised in the presence of simple organocatalysts.
In 2017, the multistep synthesis of d-glucose-based monomer CC1 was described (Figure 12) [97].
The procedure involves the Ferrier rearrangement of starting tri-O-acetyl-d-glucal to yield intermediate a, followed by deprotection of the alcohol groups. Finally, the 1,3-diol b was converted into the desired monomer CC1 using triphosgene in the presence of pyridine. The corresponding polycarbonate PCC1 was successfully obtained via ROP of CC1 by using 1,8-diazabicyclo[5.4.0]-undec-7-ene (DBU), 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD) and 1-(3,5-bis-(trifluoromethyl)-phenyl)-3-cyclohexyl-2-thiourea (TU) in the presence of 4-methylbenzyl alcohol (MBA) as the initiator. Almost complete conversion was achieved in short times (1–10 min) with all the catalysts, obtaining the highest Mn of about 10 kDa with TBD. However, bimodal molecular weights distribution was observed in this case, probably due to trans-carbonation side-reaction. The PCC1 shows a relatively high Tg of about 69 °C, likely due to the presence of a rigid six-membered ring. Interestingly, this sugar-based polymer presents a double-bond that could be further used in post-synthetic transformations.
Soon after, direct synthesis of the five-membered cyclic carbonate monomer CC2 was described by reaction of a commercially available d-glucopyranoside derivative with the same procedure reported for CC1, but a very high yield (95%) was obtained in this case (Figure 13) [98].
It is important to note that, in general, five-membered CCs do not react under ROP catalytic conditions. On the contrary, in this case, the combination of trans stereochemistry with the presence of fused rings provided enough ring-strain for a successful ROP. Indeed, polycarbonate PCC2 was readily obtained under conditions similar to PCC1. However, a longer reaction time of 5 h was needed and Mn in the range of 3–6 kDa were obtained. Interestingly, PCC2 was modified by post-synthetic reactions. Indeed, removing the benzylidene protecting groups under acidic conditions led to the formation of hydroxy-functionalized PCC2-OH, which was further converted into the acyl protected PCC2-OAc using acetic anhydride. Measurement of water contact angle revealed, as expected, that the hydroxy-functionalized polycarbonate PCC2-OH is significantly more hydrophilic (38° ≤ contact angle ≤ 58°) than the protected polycarbonates (94° ≤ contact angle ≤ 128°).
More recently, the synthesis of a series of differently substituted six-membered CCs (CC3-R, Figure 14) was described, starting from the same glucopyranose derivative used for the preparation of CC2 [99]. In this case, the initial hydroxyl groups were converted into linear carbonates c by reaction with various chloroformates. Removal of benzylidene acetal protecting group, forming 1,3-diols d, followed by cyclic carbonate formation with triphosgene yield the desired CC3-R monomers. The ROP of CC3-R led to the formation of the corresponding PCC-R polycarbonates with narrow distributions (1.04 ≤ Đ ≤ 1.06) and similar molecular weights (15 ≤ Mn ≤ 16 kDa). Remarkably, the presence of structurally different substituents resulted in the fine-tuning of Tg in a wide thermal window, from 38 to 125 °C. Very interestingly, the authors were able to analyse the relationship between Tg and molecular weight for two selected polycarbonates, PCC-neoP and PCC-EtHex. In detail, the analysis was conducted at the level of discrete oligomers (from monomers to tetramers) and polymers with a polymerisation degree from 10 to 40, finding that the variation of Tg with Mn is very well described by the Flory-Fox theory [100].
Notably, the same research group analysed the regio-chemistry of CC3-R ROP, combining NMR analyses and DFT investigations, finding that the presence of carbonate pendant groups is crucial for the formation of regio-regular polymers [101]. Indeed, analogous monomers bearing ether pendant groups were also polymerized but in a regio-irregular fashion, and the pendant carbonate moieties seem to be involved in the stabilization of the transition state via intermolecular hydrogen-bonding with TBD.
All the examples of sugar-based CCs described so far demonstrated the high potential of this class of monomers but, unfortunately, their preparation is based on the use of phosgene derivatives. Consequently, the search for more sustainable reaction procedures for the synthesis of structurally complex cyclic carbonates is a topic of high interest [83]. In this regard, Buchard’s research group reported a new method for the synthesis of six-membered CCs from diols, using carbon dioxide, an organic base, and tosyl chloride (TsCl) [102,103]. This procedure was successfully applied for the synthesis of sugar-based cyclic carbonate monomers.
In 2016, the synthesis of d-mannose-based monomer CC4 was described, and its structure was also confirmed by X-ray analysis (Figure 15) [104].
As previously described, also in this case the trans-stereochemistry and the presence of a fused tricyclic structure resulted in a good reactivity of CC4 in the organocatalytic ROP promoted by TBD. The resulting PCC4 was obtained with Mn in the range 6–33 kDa and with narrow distributions (1.10 ≤ Đ ≤ 1.19). However, using CC4/TBD ratio higher than 150 and reaching Mn > 15 kDa, back-biting of the polymer chain forming smaller cyclic species was observed, supported by MALDI analysis. Notably, the ROP of CC4 proceeds in a regioselective manner probably due to the steric hindrance of the 2,3-O-isopropylidene protecting group. In addition, PCC4 exhibits a high Tg of 152 °C likely due to the rigidity of its structure. Deprotection of the 1,2-hydroxyl groups was also explored, and it was possible to remove 70% of protecting groups by treating with CF3COOH/H2O mixture.
Later, the same group investigated the use of thymidine in polycarbonate synthesis [105]. Attempts to obtain a cyclic carbonate monomer using both phosgene-reagents and the CO2 method described before were unsuccessful. This result was attributed to the high strain of the thymidine-base carbonate in the trans-configuration. Interestingly, a different multistep approach was adopted involving the protection of the secondary alcohol moiety with a tosyl group, followed by an SN2-type intramolecular reaction that, by involving the stereochemistry inversion in a 3′ position, yielded the desired CC5 monomer (Figure 16).
The ROP of CC5 to obtain the corresponding PCC5 was achieved in the presence of TBD, and the polymerisation proceeds until the establishment of the monomer–polymer equilibrium (e.g., 80% conversion at 25 °C), so the temperature dependence of the equilibrium was studied, and the polymerisation thermodynamic parameters determined equal to ΔHp = −12.3 ± 0.4 kJ ∙ mol−1 and ΔSp = −29 ± 1.1 J mol−1 ∙ K−1. As for the case of CC4, back-biting of the polymer chain was observed, however Mn up to 17 kDa with Đ = 1.3 were obtained. The hydrolytic stability of PCC5 was also investigated and, while stable in 1 M HCl for one week, the polymer hydrolyses in 1 M NaOH in 4 h to the cis-diol of 3-N-methyl thymidine. A cell attachment study was conducted with a human osteoblast cancer cell line (MG-63) to explore the possibility to use PCC5 as a tissue engineering scaffold indicating promising results, but further structural modifications are needed to obtain better results.
Soon after, the synthesis of 2-deoxy-d-ribose-based monomer CC6 was reported with a procedure similar to that used for monomer CC5 [106]. Using this method, the desired carbonate was obtained in 80% yield in the two α and β anomeric forms (Figure 17).
Polymerisation studies conducted with TBD as the catalyst revealed that only the α-form of CC6 is susceptible to ROP, yielding the insoluble polycarbonate PCC6 with Mn as high as 26 kDa, while the β-form cannot polymerise. This difference was rationalized by DFT investigations, finding that the overall thermodynamics of β-CC6 ROP is unfavourable (ΔΔG = +1.4/+2.0 kcal ∙ mol−1) compared with that of α-CC6 (ΔΔG = −0.9 to −1.9 kcal ∙ mol−1). MALDI analyses of PCC6 revealed the presence of cyclic species in high concentrations (≈ 50%). Copolymerisation of CC6 with trimethylene carbonate (TMC) was investigated, obtaining random copolymers with CC6:TMC composition from 93:7 to 14:86, 43 ≤ Mn ≤ 57 kDa and 1.37 ≤ Đ ≤ 1.35. The poly(CC6-co-TMC) copolymers showed Tg values comprised between those of PCC6 and PTMC homopolymers (58 and −25 °C, respectively), depending on composition in agreement with the Fox equation.
In 2019, Buchard et al. demonstrated that monomer CC7, prepared from glucose-derived erythritol, can be used in two different stereoselective polymerisation processes (Figure 18) [107]. In the case of TBD-mediated ROP, the original stereochemistry of the double bond is retained, yielding P-cis-CC7. In contrast, the ring-opening metathesis polymerisation (ROMP) performed with the Grubbs second-generation catalyst (GII, Figure 17) inverts the original stereochemistry, producing P-trans-CC7 with stereoselectivity up to 95%.
The two polymers have different material properties, indeed P-trans-CC7 is a soft amorphous material (Tg = −22 °C) while P-cis-CC7 is a hard, semicrystalline material (Tg = −24 °C; Tm = 115 °C).
Another interesting sugar-based precursor is represented by (R)-(-)-1,3-butanediol, obtained via glucose fermentation with very high stereoselectivity [108]. Recently Hillmyer et al. reported the preparation of carbonate monomer (R)-CC8, based on (R)-(-)-1,3-butanediol, and its ROP toward polycarbonate PCC8 (Figure 19) [109].
The ROP of CC8 was already reported before, however, only amorphous polycarbonates were obtained because of the scarce regiochemistry control [110,111], or the use of racemic monomer mixtures leading to atactic PCC8 [112,113]. On the contrary, the polymerisation of (R)-CC8 promoted by bulky Zn-β-diiminate, and Al-salen complexes (Zn1, and Al4–5, Figure 18) proceeds with virtually perfect regioselectivity (99%) under operationally simple reaction conditions (bulk, 70 °C), with Mn up to 150 kDa and Đ = 1.2 in the case of Al4. The PCC8 obtained in this way is semicrystalline, with a Tm = 73 °C. Interestingly, an attempt to form a stereocomplex by mixing PCC8 prepared from (R)- and (S)-CC8 only resulted in a little enhancement with Tm = 80 °C.
Replacing CO2 with CS2 in the DBU-based procedure described above discloses the access to structurally related thiocarbonate monomers TC1–3, derived from 2-deoxy-d-ribose or d-xylofuranose (Figure 20) [114]. Poly-thiocarbonates PTC1–3, were obtained by organocatalytic ROP of TC1–3, with 3 ≤ Mn ≤ 31 kDa and 1.3 ≤ Đ ≤ 2.2. The interest in these polycarbonate analogues is based on the observation that the substitution of some oxygen atoms with sulphur ones may result in the enhancement of certain material properties [115,116,117]. However, in this case, the Tg were similar to or lower than those of the corresponding sugar-based polycarbonates.

3.2. ROCOP of Carbon Dioxide and Oxiranes

The sustainable synthesis of polycarbonates via the direct use of carbon dioxide involves the identification of suitable bio-based co-monomers. As mentioned in the introductory paragraph, not all oxiranes can be copolymerized with CO2, because of the preferential formation of the corresponding cyclic carbonates. Consequently, the use of bio-sourced epoxides is quite limited and an exemplary case is given by limonene oxide (LO) for which, to date, only two catalytic systems are known to promote the formation of the corresponding poly(limonene carbonate) (PLC) (Figure 21) [118,119]. This topic has been recently reviewed by some of us [6].
However, in the last few years, some interesting cases of bio-based PC obtained from the ROCOP of CO2 with properly designed monomers have been reported.
In 2020, Greiner et al. reported the synthesis of menth-2-ene oxide (MO) and its copolymerisation with CO2 promoted by a zinc β-diiminato catalyst (Zn2), obtaining the new poly(menthane carbonate) PMC (Figure 22) [120].
The monomer was obtained starting from L-menthol, which was first converted to menth-2-ene j, and then epoxidized to the desired MO obtaining a mixture of the two possible stereoisomers in a 3:7 ratio. Using catalyst Zn2, copolymerisation of MO with CO2 at 25 bar and 25 °C in toluene gave good results, yielding PMC with 7 ≤ Mn ≤ 23 kDa and 1.11 ≤ Đ ≤ 1.23. Bulk polymerisation was also possible, but with a lower degree of control over the copolymerisation process obtaining bimodal molecular weights distributions. Importantly, both MO isomers were incorporated in PMC, as confirmed by NMR analyses of polycarbonate hydrolysis products. This is in contrast with the case of Zn2-promoted LO/CO2 ROCOP where the trans-epoxide is preferentially converted. The PMC has a high Tg of 144 °C and did not show any crystallinity. However, this polycarbonate exhibits a particularly high decomposition temperature (Td = 308 °C) if compared with structurally related PLC (Td = 229 °C) prepared with the same catalyst.
Very recently, Wooley et al. described the use of a d-xylose derived oxetane in polycarbonate synthesis using an interesting dual-pathway approach [121]. First, the desired xylose oxetane (XO) was prepared via intramolecular cyclization of intermediate l, obtained from d-xylose as in Figure 23.
At this point, the first approach was based on the copolymerisation of XO and carbon dioxide promoted by the combination of Cr-salen complex Cr2 with a suitable cocatalyst, which is known to catalyse this reaction for structurally related substrates (Figure 24) [122,123]. Under the best conditions (i.e., 140 °C, and 3 MPa of CO2), it was possible to obtain PXC with 80% selectivity, Mn = 7.4 kDa, and Đ = 1.12. In parallel, the xylene carbonate XC is also formed during the reaction. In order to maximise the polycarbonate selectivity, a second approach was adopted based on the initial formation of XC followed by ROP (Figure 24). The cycloaddition reaction of CO2 to XO was performed, using ZnI2 and tetra-butylammonium iodide as the catalytic system, obtaining XC with 99% selectivity. Next, the ROP of XC catalysed by TBD in the presence of MBA gave PXC with Mn = 10 kDa and Đ < 1.2. The polycarbonate obtained with the two different approaches has the same structure and shows similar thermal behaviour (122 ≤ Tg ≤ 125 °C). However, MALDI analyses showed that different chain-end groups are present depending on the approach employed. In the XO/CO2 ROCOP case, PXC was initiated with a halide group, while a 4-methylbenzyl group was observed in the case of XC ROP.

4. Polyethers

Polyethers and polyols are commonly synthesized via ROP of cyclic esters. With respect to sustainable monomers, terpene-derived epoxides represent excellent candidates for green polyethers. In this scenario, limonene oxide is widely explored, since it can be easily produced from limonene, an abundant monoterpene readily extracted from the peel of citrus fruits. However, studies regarding the homo-polymerisation of LO are scant as only a few catalysts can overcome the kinetic barrier for the activation of its internal trisubstituted epoxide group.
The cationic photoinitiated ROP of LO using diaryliodonium or triarylsulphonium salts as photoinitiators was reported by Crivello et al. [124]. Nonetheless, the process led to a mixture of low molecular weight products derived from several side reactions. In turn, α-pinene oxide (α-PO) displayed higher reactivity compared to LO, allegedly because of the simultaneous double ring-opening reaction involving both its epoxide and cyclobutene rings. However, side reactions occur also, in this case, limiting the use of α-PO in homopolymers production.
More recently, the aluminium-based catalyst Al6 was successfully employed in the synthesis of poly(limonene oxide) (PLO) by coordination/insertion ROP (Figure 25) [125]. The produced polyether shows low molecular weights, up to 1300 Da, and moderate dispersity (1.37 ≤ Ð ≤ 1.42).
However, despite the scarce literature on LO homo-polymerisation, several studies are reported on the ROCOP of anhydrides and terpene-derived epoxides (vide supra).
Another attractive bio-based monomer for polyethers synthesis is isosorbide, a molecule derived from sorbitol. Reineke et al. reported on a cationic and quasi-zwitterionic ROP of an annulated isosorbide derivative (1,4:2,5:3,6-trianhydro-d-mannitol) in the presence of Sc(OTf)3 and propylene oxide. Depending on the reaction conditions, it was possible to selectively direct the polymerisation towards either linear or cyclic macromolecular architectures (Figure 26) [126].
Vares and co-workers reported on the synthesis of isosorbide mono-epoxides with the remaining hydroxyl group capped as a methyl ether, in order to avoid difficulties in subsequent polymerisations to linear polymers. The novel isosorbide epoxides, polymerised by anionic ROP, afforded polyethers with Tg around 10–15°C and molecular weights in the range of 13–17 kDa (Figure 27) [127].
Most polyether polyols are synthesized via ROP of oxiranes such as ethylene oxide (EO), propylene oxide (PO) or butylene oxide (BO). When oxiranes are obtained from renewable sources, they are considered bio-based monomers. To this end, EO can be synthesized from biomass-derived ethanol (e.g., from sugar cane), PO and BO can be bio-synthesized using a multistep process from glucose or via fermentation of biomasses, respectively [128,129]. Oxiranes can be polymerised via anionic, coordination, and cationic ROP as well as by metal-free ROP. PEO and PPO are mainly synthesized by oxyanionic polymerisation. However, due to the drawbacks of such a process, studies on the coordination polymerisation reaction aiming to achieve high molecular weight polymers have been carried out. In this regard, the effect of many initiators on kinetics and molecular weight control has been investigated. Excellent overviews are reported on the polymerisation of EO, PO and other alkylene oxides, classically considered non-bio-based monomers, hence out of the scope of this review [130].
Similarly, THF can also be considered a bio-based monomer polymerisable by ROP to produce a polyether polyol. The preparation of poly(tetrahydrofuran) (PTHF) is known since the 1930s but nowadays, PolyTHF® commercialized by BASF is produced using THF obtained from the bio-based 1,4 butanediol cyclisation [131].

5. Conclusions

We discussed the synthesis of different classes of bio-based polymeric materials by ring-opening (co)polymerisation, with a high focus on polyesters, polycarbonates and polyethers. In particular, the great tolerance toward the presence of various functional groups on the starting monomers allows for the synthesis of structurally complex and functional materials. This, in turn, offers a wide range of possibilities in terms of possible applications, from commodity plastics to fine applications such as tissue engineering. Nevertheless, several significant aspects still need to be addressed.
With respect to the polylactones, relevant monomers such as ε-caprolactone and δ-valerolactone are still synthesized mainly from petrochemicals while polylactide, despite being biodegradable and biomass-derived, finds limited applications due to its poor mechanical properties. Polyesters derived from fully renewable higher lactones (i.e., ω-pentadecalactone) are considered possible solutions to these issues, albeit their large-scale production is still limited. Moreover, the biodegradation behaviour of this class of polymers is still widely unexplored; indeed, further insights are required to evaluate the long-term impact of such materials on the natural environment.
In the case of the polyethers, the identification of suitable monomers exhibiting good reactivity also in homopolymerisation reactions is highly desirable. This problem is related to that of polycarbonates. Indeed, in the latter case, bio-based epoxides could serve as starting material for the production of bio-based PCs. In addition, the amount of bio-based cyclic carbonates suitable for ring-opening polymerisation is still limited.
In general, there is still a tremendous lack of concrete sustainable monomers for the synthesis of oxygenated polymers.
Another aspect is related to the possibility to obtain controlled molecular weights suitable for real applications. Indeed, molecular weights are often reported in the range from 2 to 20 kg/mol and, notwithstanding these values are sufficient for certain uses like polyols in polyurethane production [132], they are too low for other applications. For example, industrial production of PLA by ROP with Sn(Oct)2 gives Mn spanning from 20 to 220 kg/mol with Đ = 1.6–2.0 [133,134].
In conclusion, with this review, we demonstrated that the use of ring-opening (co)polymerisation processes is a highly promising approach for the assessment of new sustainable polymers obtained from bio-based, renewable feedstocks, and we hope that this will trigger more significant advancements in this direction.

Author Contributions

Writing—original draft preparation, O.S., L.I., and F.D.M.; writing—review and editing, O.S., L.I., and F.D.M.; visualization, O.S., L.I., and F.D.M. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank the Università degli Studi dell’Insubria for financial support. L.I. thanks for funding under the Scheme “Fondo di Ateneo per la Ricerca” (FAR 2021).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Buchanan, R.A. History of Technology. Available online: https://www.britannica.com/technology/history-of-technology (accessed on 31 March 2022).
  2. Zalasiewicz, J.; Waters, C.N.; Ivar do Sul, J.A.; Corcoran, P.L.; Barnosky, A.D.; Cearreta, A.; Edgeworth, M.; Gałuszka, A.; Jeandel, C.; Leinfelder, R.; et al. The geological cycle of plastics and their use as a stratigraphic indicator of the Anthropocene. Anthropocene 2016, 13, 4–17. [Google Scholar] [CrossRef] [Green Version]
  3. Riechers, M.; Fanini, L.; Apicella, A.; Galván, B.C.; Blondel, E.; Espiña, B.; Kefer, S.; Keroullé, T.; Klun, K.; Pereira, T.R.; et al. Plastics in our ocean as transdisciplinary challenge. Mar. Pollut. Bull. 2021, 164, 112051. [Google Scholar] [CrossRef] [PubMed]
  4. Jehanno, C.; Alty, J.W.; Roosen, M.; De Meester, S.; Dove, A.P.; Chen, E.Y.-X.; Leibfarth, F.A.; Sardon, H. Critical advances and future opportunities in upcycling commodity polymers. Nature 2022, 603, 803–814. [Google Scholar] [CrossRef] [PubMed]
  5. Zhu, Y.; Romain, C.; Williams, C.K. Sustainable polymers from renewable resources. Nature 2016, 540, 354–362. [Google Scholar] [CrossRef] [PubMed]
  6. Della Monica, F.; Kleij, A.W. From terpenes to sustainable and functional polymers. Polym. Chem. 2020, 11, 5109–5127. [Google Scholar] [CrossRef]
  7. Galbis, J.A.; de García-Martín, M.G.; de Paz, M.V.; Galbis, E. Synthetic Polymers from Sugar-Based Monomers. Chem. Rev. 2016, 116, 1600–1636. [Google Scholar] [CrossRef]
  8. de Espinosa, M.L.; Meier, M.A.R. Plant oils: The perfect renewable resource for polymer science?! Eur. Polym. J. 2011, 47, 837–852. [Google Scholar] [CrossRef] [Green Version]
  9. Sun, Z.; Fridrich, B.; de Santi, A.; Elangovan, S.; Barta, K. Bright Side of Lignin Depolymerisation: Toward New Platform Chemicals. Chem. Rev. 2018, 118, 614–678. [Google Scholar] [CrossRef] [Green Version]
  10. Isikgora, F.H.; Becer, C.R. Lignocellulosic biomass: A sustainable platform for the production of bio-based chemicals and polymers. Polym. Chem. 2015, 6, 4497–4559. [Google Scholar] [CrossRef] [Green Version]
  11. Jiang, T.; Duan, Q.; Zhu, J.; Liu, H.; Yu, L. Starch-based biodegradable materials: Challenges and opportunities. Adv. Ind. Eng. Polym. Res. 2020, 3, 8–18. [Google Scholar] [CrossRef]
  12. Winnacker, M. Polyhydroxyalkanoates: Recent Advances in Their Synthesis and Applications. Eur. J. Lipid Sci. Technol. 2019, 121, 1900101. [Google Scholar] [CrossRef] [Green Version]
  13. Wahlen, C.; Frey, H. Anionic Polymerisation of Terpene Monomers: New Options for Bio- Based Thermoplastic Elastomers. Macromolecules 2021, 54, 7323–7336. [Google Scholar] [CrossRef]
  14. Lamparelli, D.H.; Winnacker, M.; Capacchione, C. Stereoregular Polymerisation of Acyclic Terpenes. ChemPlusChem 2022, 87, e202100366. [Google Scholar] [CrossRef] [PubMed]
  15. Veith, C.; Diot-Néant, F.; Miller, S.A. Synthesis and polymerisation of bio-based acrylates: A review. Polym. Chem. 2020, 11, 7452–7470. [Google Scholar] [CrossRef]
  16. Gomez-Lopez, A.; Panchireddy, S.; Grignard, B.; Calvo, I.; Jerome, C.; Detrembleur, C.; Sardon, H. Poly(hydroxyurethane) Adhesives and Coatings: State-of-the-Art and Future Directions. ACS Sustain. Chem. Eng. 2021, 9, 9541–9562. [Google Scholar] [CrossRef]
  17. Coates, G.W.; Moore, D.R. Discrete metal-based catalysts for the copolymerisation of CO2 and epoxides: Discovery, reactivity, optimization, and mechanism. Angew. Chem. Int. Ed. 2004, 43, 6618–6639. [Google Scholar] [CrossRef] [PubMed]
  18. Klaus, S.; Lehenmeier, M.W.; Anderson, C.E.; Rieger, B. Recent advances in CO2/epoxide copolymerisation—New strategies and cooperative mechanisms. Coord. Chem. Rev. 2011, 255, 1460–1479. [Google Scholar] [CrossRef]
  19. Lu, X.-B.; Darensbourg, D.J. Cobalt catalysts for the coupling of CO2 and epoxides to provide polycarbonates and cyclic carbonates. Chem. Soc. Rev. 2012, 41, 1462–1484. [Google Scholar] [CrossRef]
  20. Paul, S.; Zhu, Y.; Romain, C.; Brooks, R.; Saini, P.K.; Williams, C.K. Ring-opening copolymerisation (ROCOP): Synthesis and properties of polyesters and polycarbonates. Chem. Commun. 2015, 51, 6459–6479. [Google Scholar] [CrossRef] [Green Version]
  21. Kozak, C.M.; Ambrose, K.; Anderson, T.S. Copolymerisation of carbon dioxide and epoxides by metal coordination complexes. Coord. Chem. Rev. 2018, 376, 565–587. [Google Scholar] [CrossRef]
  22. Xiao, L.; Zhao, Y.; Qiao, S.; Sun, Z.; Santoro, O.; Redshaw, C. Synthesis and structures of mono- and di-nuclear aluminium and zinc complexes bearing α-diimine and related ligands, and their use in the ring opening polymerization of cyclic ester. Dalton Trans. 2020, 49, 1456–1472. [Google Scholar] [CrossRef] [PubMed]
  23. Dubois, P.; Coulembier, O.; Raquez, J.-M. Handbook of Ring-Opening Polymerisation; Wiley-VCH Verlag GmbH & Co.: Weinheim, Germany, 2009. [Google Scholar]
  24. Santoro, O.; Redshaw, C. Use of titanium complexes bearing diphenolate or calix[n]arene ligands in α-olefin polymerisation and the ROP of cyclic esters. Catalysts 2020, 10, 210. [Google Scholar] [CrossRef] [Green Version]
  25. Vert, M.; Li, S.M.; Spenlehauer, G.; Guerin, P. Bioresorbability and biocompatibility of aliphatic polyesters. J. Mater. Sci. Mater. Med. 1992, 3, 432–446. [Google Scholar] [CrossRef]
  26. Gross, R.A.; Kalra, B. Biodegradable Polymers for the Environment. Science 2002, 297, 803–807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Buntara, T.; Noel, S.; Phua, P.H.; Melián-Cabrera, I.; de Vries, J.G.; Heeres, H.J. Although ε-CL is currently obtained mainly from petroleum, promising sustainable processes have been developed. Caprolactam from Renewable Resources: Catalytic Conversion of 5-Hydroxymethylfurfural into Caprolactone. Angew. Chem. Int. Ed. 2011, 50, 7083–7087. [Google Scholar] [CrossRef]
  28. Hillmyer, M.A.; Tolman, W.B. Aliphatic Polyester Block Polymers: Renewable, Degradable, and Sustainable. Acc. Chem. Res. 2014, 47, 2390–2396. [Google Scholar] [CrossRef]
  29. Duparc, H.V.; Shakaroun, R.M.; Slawinski, M.; Carpentier, J.F.; Guillaume, S.M. Ring-opening (co)polymerisation of six-membered substituted δ-valerolactones with alkali metal alkoxides. Eur. Polym. J. 2020, 134, 10985–109864. [Google Scholar] [CrossRef]
  30. Shakaroun, R.M.; Li, H.; Jéhan, P.; Blot, M.; Alaaeddine, A.; Carpentier, J.-F.; Guillaume, S.M. Stereoselective ring-opening polymerisation of functional β-lactones: Influence of the exocyclic side-group. Polym. Chem. 2021, 12, 4022–4034. [Google Scholar] [CrossRef]
  31. Albertson, A.C.; Varma, I.K. Degradable Polymer Microspheres for Controlled Drug Delivery. Adv. Polym. Sci. 2002, 157, 67–112. [Google Scholar]
  32. Bogaert, J.-C.; Coszach, P. Poly(lactic acids): A potential solution to plastic waste dilemma. Macromol. Symp. 2000, 153, 287–303. [Google Scholar] [CrossRef]
  33. Masutani, K.; Kimura, Y. Chapter 1: PLA Synthesis. From the Monomer to the Polymer. In Poly(Lactic Acid) Science and Technology: Processing, Properties, Additives and Applications; The Royal Society of Chemistry: London, UK, 2015; pp. 1–36. [Google Scholar] [CrossRef]
  34. Fuoco, T.; Pappalardo, D. Aluminum Alkyl Complexes Bearing Salicylaldiminato Ligands: Versatile Initiators in the Ring-Opening Polymerisation of Cyclic Esters. Catalysts 2017, 7, 64. [Google Scholar] [CrossRef]
  35. Ksiazkiewicz, A.N.; Pich, A.; Kögerler, P.; Monakhov, K.Y.; Herres-Pawlis, S. Mononuclear zinc(II) Schiff base complexes as catalysts for the ring-opening polymerisation of lactide. Eur. Polym. J. 2020, 122, 109302–109308. [Google Scholar] [CrossRef]
  36. 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 Polymerisation (ROP) of Cyclic Esters. Catalysts 2020, 10, 800. [Google Scholar] [CrossRef]
  37. O’Keefe, B.J.; Monnier, S.M.; Hillmyer, M.A.; Tolman, W.B. Rapid and Controlled Polymerisation of Lactide by Structurally Characterized Ferric Alkoxides. J. Am. Chem. Soc. 2001, 123, 339–340. [Google Scholar] [CrossRef]
  38. O’Keefe, B.J.; Breyfogle, L.E.; Hillmyer, M.A.; Tolman, W.B. Mechanistic Comparison of Cyclic Ester Polymerisations by Novel Iron(III)−Alkoxide Complexes:  Single vs Multiple Site Catalysis. J. Am. Chem. Soc. 2002, 124, 4384–4393. [Google Scholar] [CrossRef]
  39. Gibson, V.C.; Marshall, E.L.; Navarro-Llobet, D.; White, A.J.P.; Williams, D.J. A well-defined iron(ii) alkoxide initiator for the controlled polymerisation of lactide. J. Chem. Soc. Dalton Trans. 2002, 4321–4322. [Google Scholar] [CrossRef]
  40. McGuinness, D.S.; Marshall, E.L.; Gibson, V.C.; Steed, J.W. Anionic iron(II) alkoxides as initiators for the controlled ring-opening polymerisation of lactide. J. Polym. Sci. Part A 2003, 41, 3798–3803. [Google Scholar] [CrossRef]
  41. Biernesser, A.B.; Li, B.; Byers, J.A. Redox-Controlled Polymerisation of Lactide Catalyzed by Bis(imino)pyridine Iron Bis(alkoxide) Complexes. J. Am. Chem. Soc. 2013, 135, 16553–16560. [Google Scholar] [CrossRef]
  42. Manna, C.M.; Kaplan, H.Z.; Li, B.; Byer, J.A. High molecular weight poly(lactic acid) produced by an efficient iron catalyst bearing a bis(amidinato)-N-heterocyclic carbene ligand. Polyhedron 2014, 84, 160–167. [Google Scholar] [CrossRef]
  43. Herber, U.; Hegner, K.; Wolters, D.; Siris, R.; Wrobel, K.; Hoffmann, A.; Lochenie, C.; Weber, B.; Kuckling, D.; Herres-Pawlis, S. Iron(II) and Zinc(II) Complexes with Tetradentate Bis(pyrazolyl)methane Ligands as Catalysts for the Ring-Opening Polymerisation of rac-Lactide. Eur. J. Inorg. Chem. 2017, 2017, 1341–1354. [Google Scholar] [CrossRef]
  44. Rittinghaus, R.D.; Schfer, P.M.; Albrecht, P.; Conrads, C.; Hoffmann, A.; Ksiazkiewicz, A.N.; Bienemann, O.; Pich, A.; Herres-Pawlis, S. New Kids in Lactide Polymerisation: Highly Active and Robust Iron Guanidine Complexes as Superior Catalysts. ChemSusChem 2019, 12, 2161–2165. [Google Scholar] [CrossRef] [PubMed]
  45. Brown, L.A.; Wekesa, F.S.; Unruh, D.K.; Findlater, M.; Long, B.K. BIAN-Fe(η6-C6H6): Synthesis, Characterization, and L-Lactide Polymerisation. J. Polym. Sci. Part A 2017, 55, 2824–2830. [Google Scholar] [CrossRef]
  46. Liang, Y.; Duan, R.-L.; Hu, C.-Y.; Lia, L.-L.; Pang, X.; Zhang, W.-X.; Chen, X.S. Salen-iron Complexes: Synthesis, Characterization and Their Reactivity with Lactide. Chin. J. Polym. Sci. 2018, 36, 185–189. [Google Scholar] [CrossRef]
  47. Duan, R.; Hu, C.; Li, X.; Pang, X.; Sun, Z.; Chen, X.; Wang, X. Air-Stable Salen−Iron Complexes: Stereoselective Catalysts for Lactide and ε-Caprolactone Polymerisation through in Situ Initiation. Macromolecules 2017, 50, 9188–9195. [Google Scholar] [CrossRef]
  48. Semlyen, J.A. Cyclic Polymers, 2nd ed.; Kluwer: Dordrecht, The Netherlands, 2000. [Google Scholar]
  49. Chang, Y.A.; Waymouth, R.M. Recent progress on the synthesis of cyclic polymers via ring-expansion strategies. J. Polym. Sci. Part A 2017, 55, 2892–2902. [Google Scholar] [CrossRef]
  50. Weil, J.; Mathers, R.T.; Getzler, Y.D.Y.L. Lactide Cyclopolymerisation by an Alumatrane-Inspired Catalyst. Macromolecules 2012, 45, 1118–1121. [Google Scholar] [CrossRef]
  51. Castro-Osma, J.A.; Alonso-Moreno, C.; García-Martinez, J.C.; Fernández-Baeza, J.; Sánchez-Barba, L.F.; Lara-Sánchez, A.; Otero, A. Ring-Opening (ROP) versus Ring-Expansion (REP) Polymerisation of ε-Caprolactone to Give Linear or Cyclic Polycaprolactones. Macromolecules 2013, 46, 6388–6394. [Google Scholar] [CrossRef]
  52. Kricheldorf, H.R.; Weidner, S.M.; Meyer, A. High Tm poly(Llactide)s via REP or ROPPOC of L-lactide. Polym. Chem. 2020, 11, 2182–2193. [Google Scholar] [CrossRef] [Green Version]
  53. Kricheldorf, H.R.; Weidner, S.M. High molar mass cyclic poly(L-lactide) via ring-expansion polymerisation with cyclic dibutyltin bisphenoxides. Eur. Polym. J. 2018, 105, 158–166. [Google Scholar] [CrossRef]
  54. Kricheldorf, H.R.; Weidner, S.M. High molar mass cyclic poly(L-lactide) obtained by means of neat tin(II) 2-ethylhexanoate. Polym. Chem. 2020, 11, 5249–5260. [Google Scholar] [CrossRef]
  55. Impemba, S.; Della Monica, F.; Grassi, A.; Capacchione, C.; Milione, S. Cyclic Polyester Formation with an [OSSO]-Type Iron(III) Catalyst. ChemSusChem 2020, 13, 141–145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Della Monica, F.; Maity, B.; Pehl, T.; Buonerba, A.; De Nisi, A.; Monari, M.; Grassi, A.; Rieger, B.; Cavallo, L.; Capacchione, C. [OSSO]-Type Iron(III) Complexes for the Low-Pressure Reaction of Carbon Dioxide with Epoxides: Catalytic Activity, Reaction Kinetics, and Computational Study. ACS Catal. 2018, 8, 6882–6893. [Google Scholar] [CrossRef]
  57. Della Monica, F.; Ricciardi, M.; Proto, A.; Cucciniello, R.; Capacchione, C. Regioselective Ring-Opening of Glycidol to Monoalkyl Glyceryl Ethers Promoted by an [OSSO]-FeIII Triflate Complex. ChemSusChem 2019, 12, 3448–3452. [Google Scholar] [CrossRef] [PubMed]
  58. Kerr, R.W.F.; Ewing, P.M.D.A.; Raman, S.K.; Smith, A.D.; Williams, C.K.; Arnold, P.L. Ultrarapid Cerium(III)–NHC Catalysts for High Molar Mass Cyclic Polylactide. ACS Catal. 2021, 11, 1563–1569. [Google Scholar] [CrossRef]
  59. Santoro, O.; Elsegood, M.R.J.; Teat, S.J.; Yamato, T.; Redshaw, C. Lithium calix[4]arenes: Structural studies and use in the ring opening polymerisation of cyclic esters. RSC Adv. 2021, 11, 11304–11317. [Google Scholar] [CrossRef]
  60. Santoro, O.; Elsegood, M.R.J.; Bedwell, E.V.; Pryce, J.A.; Redshaw, C. INSIGHTS into the structures adopted by titanocalix[6 and 8]arenes and their use in the ring opening polymerisation of cyclic esters. Dalton Trans. 2020, 49, 11978–11996. [Google Scholar] [CrossRef]
  61. Gruszka, W.; Garden, J.A. Advances in heterometallic ring-opening (co)polymerisation catalysis. Nat. Commun. 2021, 12, 3252. [Google Scholar] [CrossRef]
  62. Focarete, M.L.; Scandola, M.; Kumar, A.; Gross, R.A. Flat polymer ribbons and other shapes by electrospinning. J. Polym. Sci. Part B Polym. Phys. 2001, 39, 1721–1729. [Google Scholar] [CrossRef]
  63. Huang, H.-C.; Wang, B.; Chen, X.-L.; Pan, L.; Li, Y.-S. Ring-opening polymerisation of (Macro)lactones by highly active mononuclear salen–aluminum complexes bearing cyclic β-ketoiminato ligand. J. Polym. Sci. Part A Polym. Chem. 2019, 57, 973–981. [Google Scholar] [CrossRef]
  64. Naddeo, M.; D’Auria, I.; Viscusi, G.; Gorrasi, G.; Pellecchia, C.; Pappalardo, D. Tuning the thermal properties of poly(ethylene)-like poly(esters) by copolymerisation of ε-caprolactone with macrolactones, in the presence of a pyridylamidozinc(II) complex. J. Polym. Sci. 2020, 58, 528–539. [Google Scholar] [CrossRef]
  65. Nifant’ev, I.E.; Shlyakhtin, A.V.; Tavtorkin, A.N.; Ivchenko, P.V.; Borisov, R.S.; Churakov, A.V. Monomeric and dimeric magnesium mono-BHT complexes as effective ROP catalysts. Catal. Commun. 2016, 87, 106–111. [Google Scholar] [CrossRef]
  66. Wilson, J.A.; Hopkins, S.A.; Wright, P.M.; Dove, A.P. Synthesis of ω-Pentadecalactone Copolymers with Independently Tunable Thermal and Degradation Behavior. Macromolecules 2015, 48, 950–958. [Google Scholar] [CrossRef]
  67. Wang, X.; Wang, X.; Zhen, N.; Gu, J.; Zhang, H.; Dong, B.; Wang, F.; Liu, H. Sodium complexes bearing cavity-like conformations: A highly active and well-controlled catalytic system for macrolactone homo- and copolymerisation. Polym. Chem. 2021, 12, 1957. [Google Scholar] [CrossRef]
  68. Zhao, N.; Ren, C.; Shen, Y.; Liu, S.; Li, Z. Facile Synthesis of Aliphatic ω-Pentadecalactone Containing Diblock Copolyesters via Sequential ROP with l-Lactide, ε-Caprolactone, and δ-Valerolactone Catalyzed by Cyclic Trimeric Phosphazene Base with Inherent Tribasic Characteristics. Macromolecules 2019, 52, 1083–1091. [Google Scholar] [CrossRef]
  69. Ladelta, V.; Bilalis, P.; Gnanou, Y.; Hadjichristidis, N. Ring-opening polymerisation of ω-pentadecalactone catalyzed by phosphazene superbases. Polym. Chem. 2017, 8, 511–515. [Google Scholar] [CrossRef]
  70. Naumann, S.; Scholten, P.B.V.; Wilson, J.A.; Dove, A.P. Dual Catalysis for Selective Ring-Opening Polymerisation of Lactones: Evolution toward Simplicity. J. Am. Chem. Soc. 2015, 137, 14439–14445. [Google Scholar] [CrossRef]
  71. Longo, J.M.; Sanford, M.J.; Coates, G.W. Ring-Opening Copolymerisation of Epoxides and Cyclic Anhydrides with Discrete Metal Complexes: Structure−Property Relationships. Chem. Rev. 2016, 116, 15167–15197. [Google Scholar] [CrossRef]
  72. Robert, C.; de Montigny, F.; Thomas, C.M. Tandem synthesis of alternating polyesters from renewable resources. Nat. Commun. 2011, 2, 586. [Google Scholar] [CrossRef]
  73. Fournier, L.; Robert, C.; Pourchet, S.; Gonzalez, A.; Williams, L.; Prunet, J.; Thomas, C.M. Facile and efficient chemical functionalization of aliphatic polyesters by cross metathesis. Polym. Chem. 2016, 7, 3700–3704. [Google Scholar] [CrossRef] [Green Version]
  74. Van Zee, N.J.; Coates, G.W. Alternating Copolymerisation of Propylene Oxide with Biorenewable Terpene-Based Cyclic Anhydrides: A Sustainable Route to Aliphatic Polyesters with High Glass Transition Temperatures. Angew. Chem. Int. Ed. 2015, 54, 2665–2668. [Google Scholar] [CrossRef]
  75. Sanford, M.J.; Peña Carrodeguas, L.; Van Zee, N.J.; Kleij, A.W.; Coates, G.W. Alternating Copolymerisation of Propylene Oxide and Cyclohexene Oxide with Tricyclic Anhydrides: Access to Partially Renewable Aliphatic Polyesters with High Glass Transition Temperatures. Macromolecules 2016, 49, 6394–6400. [Google Scholar] [CrossRef]
  76. Nejad, E.H.; Paoniasari, A.; van Melis, C.G.W.; Koning, C.E.; Duchateau, R. Catalytic Ring-Opening Copolymerisation of Limonene Oxide and Phthalic Anhydride: Toward Partially Renewable Polyesters. Macromolecules 2013, 46, 631–637. [Google Scholar] [CrossRef]
  77. Carrodeguas, P.L.; Martín, C.; Kleij, A.W. Semiaromatic Polyesters Derived from Renewable Terpene Oxides with High Glass Transitions. Macromolecules 2017, 50, 5337–5345. [Google Scholar] [CrossRef]
  78. Della Monica, F.; Kleij, A.W. Synthesis and Characterization of Bio-based Polyesters with Tunable Tg by ROCOP of Beta-Elemene Oxides and Phthalic Anhydride. ACS Sustain. Chem. Eng. 2021, 9, 2619–2625. [Google Scholar] [CrossRef]
  79. Adio, A.M. (−)-trans-β-Elemene and related compounds: Occurrence, synthesis, and anticancer activity. Tetrahedron 2009, 65, 5145–5159. [Google Scholar] [CrossRef]
  80. Biermann, U.; Bornscheuer, U.T.; Feussner, I.; Meier, M.A.R.; Metzger, J.O. Fatty Acids and their Derivatives as Renewable Platform Molecules for the Chemical Industry. Angew. Chem. Int. Ed. 2021, 60, 20144–20165. [Google Scholar] [CrossRef]
  81. Nomura, K.; Awang, B.N.W. Synthesis of Bio-Based Aliphatic Polyesters from Plant Oils by Efficient Molecular Catalysis: A Selected Survey from Recent Reports. ACS Sustain. Chem. Eng. 2021, 9, 5486–5505. [Google Scholar] [CrossRef]
  82. 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]
  83. Vatcharaporn, A.; Cristòfol, À.; Della Monica, F.; Limburg, B.; D’Elia, V.; Kleij, A.W. Recent progress in the catalytic transformation of carbon dioxide into biosourced organic carbonates. Green Chem. 2021, 23, 1077–1113. [Google Scholar]
  84. Hosney, H.; Nadiem, B.; Ashour, I.; Mustafa, I.; El-Shibiny, A. Epoxidized vegetable oil and bio-based materials as PVC plasticizer. J. Appl. Polym. Sci. 2018, 135, 46270. [Google Scholar] [CrossRef] [Green Version]
  85. Clark, A.J.; Hoong, S.S. Copolymers of tetrahydrofuran and epoxidized vegetable oils: Application to elastomeric polyurethanes. Polym. Chem. 2014, 5, 3238–3244. [Google Scholar] [CrossRef]
  86. Brandolese, A.; Della Monica, F.; Pericàs, M.À; Kleij, A.W. Catalytic Ring-Opening Copolymerisation of Fatty Acid Epoxides: Access to Functional Biopolyesters. Macromolecules 2022, 55, 2566–2573. [Google Scholar] [CrossRef] [PubMed]
  87. Morin, N.; Arp, H.P.H.; Hale, S.E. Bisphenol A in Solid Waste Materials, Leachate Water, and Air Particles from Norwegian Waste-Handling Facilities: Presence and Partitioning Behavior. Environ. Sci. Technol. 2015, 49, 7675–7683. [Google Scholar] [CrossRef] [PubMed]
  88. Xu, J.; Feng, E.; Song, J. Renaissance of aliphatic polycarbonates: New techniques and biomedical applications. J. Appl. Polym. Sci. 2014, 131, 39822. [Google Scholar] [CrossRef] [Green Version]
  89. Artham, T.; Doble, M. Biodegradation of aliphatic and aromatic polycarbonates. Macromol. Biosci. 2008, 8, 14–24. [Google Scholar] [CrossRef]
  90. Luinstra, G.A.; Borchardt, E. Biodegradation of aliphatic and aromatic polycarbonates. Adv. Polym. Sci. 2012, 245, 29–48. [Google Scholar]
  91. Dove, P. Organic Catalysis for Ring-Opening Polymerisation Andrew. ACS Macro Lett. 2012, 12, 1409–1412. [Google Scholar] [CrossRef]
  92. Fukushima, K.; Nozaki, K. Organocatalysis: A Paradigm Shift in the Synthesis of Aliphatic Polyesters and Polycarbonates. Macromolecules 2020, 13, 5018–5022. [Google Scholar] [CrossRef]
  93. Tezuka, K.; Komatsu, K.; Haba, O. The anionic ring-opening polymerisation of five-membered cyclic carbonates fused to the cyclohexane ring. Polym. J. 2013, 45, 1183–1187. [Google Scholar] [CrossRef] [Green Version]
  94. Guerin, W.; Diallo, A.K.; Kirilov, E.; Helou, M.; Slawinski, M.; Brusson, J.-M.; Carpentier, J.-F.; Guillaume, S.M. Enantiopure Isotactic PCHC Synthesized by Ring-Opening Polymerisation of Cyclohexene Carbonate. Macromolecules 2014, 47, 4230–4235. [Google Scholar] [CrossRef]
  95. Della Monica, F.; Kleij, A.W. Mechanistic guidelines in nonreductive conversion of CO2: The case of cyclic carbonates. Catal. Sci. Technol. 2020, 10, 3483–3501. [Google Scholar] [CrossRef]
  96. Gregory, G.L.; López-Vidal, E.M.; Buchard, A. Polymers from sugars: Cyclic monomer synthesis, ring-opening polymerisation, material properties and applications. Chem. Commun. 2017, 53, 2198–2217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Lonnecker, A.T.; Lim, Y.H.; Wooley, K.L. Functional Polycarbonate of a D-Glucal-Derived Bicyclic Carbonate via Organocatalytic Ring-Opening Polymerisation. ACS Macro Lett. 2017, 6, 748–753. [Google Scholar] [CrossRef] [Green Version]
  98. Felder, S.E.; Redding, M.J.; Noel, A.; Grayson, S.M.; Wooley, K.L. Organocatalyzed ROP of a Glucopyranoside Derived Five-Membered Cyclic Carbonate. Macromolecules 2018, 51, 1787–1797. [Google Scholar] [CrossRef]
  99. Song, Y.; Ji, X.; Dong, M.; Li, R.; Lin, Y.-N.; Wang, H.; Wooley, K.L. Advancing the Development of Highly-Functionalizable Glucose-Based Polycarbonates by Tuning of the Glass Transition Temperature. J. Am. Chem. Soc. 2018, 140, 16053–16057. [Google Scholar] [CrossRef] [PubMed]
  100. Fox, T.G., Jr.; Flory, P.J. Second-order transition temperatures and related properties of polystyrene. I. Influence of molecular weight. J. Appl. Phys. 1950, 21, 581–591. [Google Scholar] [CrossRef]
  101. Song, Y.; Yang, X.; Shen, Y.; Dong, M.; Lin, Y.-N.; Hall, M.B.; Wooley, K.L. Invoking Side-Chain Functionality for the Mediation of Regioselectivity during Ring-Opening Polymerisation of Glucose Carbonates. J. Am. Chem. Soc. 2020, 142, 16974–16981. [Google Scholar] [CrossRef]
  102. 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]
  103. 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]
  104. Gregory, G.L.; Jenisch, L.M.; Charles, B.; Kociok-Köhn, G.; Buchard, A. Polymers from Sugars and CO2: Synthesis and Polymerisation of a D-Mannose-Based Cyclic Carbonate. Macromolecules 2016, 49, 7165–7169. [Google Scholar] [CrossRef] [Green Version]
  105. Gregory, G.L.; Hierons, E.M.; Kociok-Köhn, G.; Sharma, R.I.; Buchard, A. CO2-Driven stereochemical inversion of sugars to create thymidine-based polycarbonates by ring-opening polymerisation. Polym. Chem. 2017, 8, 1714–1721. [Google Scholar] [CrossRef] [Green Version]
  106. Gregory, G.L.; Kociok-Köhnc, G.; Buchard, A. Polymers from sugars and CO2: Ring-opening polymerisation and copolymerisation of cyclic carbonates derived from 2-deoxy-D-ribose. Polym. Chem. 2017, 8, 2093–2104. [Google Scholar] [CrossRef] [Green Version]
  107. McGuire, T.M.; Pérale, C.; Castaing, R.; Kociok-Köhn, G.; Buchard, A. Divergent Catalytic Strategies for the Cis/Trans Stereoselective Ring-Opening Polymerisation of a Dual Cyclic Carbonate/Olefin Monomer. J. Am. Chem. Soc. 2019, 141, 13301–13305. [Google Scholar] [CrossRef] [Green Version]
  108. Burgard, A.; Burk, M.J.; Osterhout, R.; Van Dien, S.; Yim, H. Development of a Commercial Scale Process for Production of 1,4- Butanediol from Sugar. Curr. Opin. Biotechnol. 2016, 42, 118–125. [Google Scholar] [CrossRef] [PubMed]
  109. DeRosa, C.A.; Luke, A.M.; Anderson, K.; Reineke, T.M.; Tolman, W.B.; Bates, F.S.; Hillmyer, M.A. Regioregular Polymers from Bio-based (R)-1,3-Butylene Carbonate. Macromolecules 2021, 54, 5974–5984. [Google Scholar] [CrossRef]
  110. Yasuda, H.; Aludin, M.-S.; Kitamura, N.; Tanabe, M.; Sirahama, H. Syntheses and Physical Properties of Novel Optically Active Poly(Ester-Carbonate)s by Copolymerisation of Substituted Trimethylene Carbonate with ε-Caprolactone and Their Biodegradation Behavior. Macromolecules 1999, 32, 6047–6057. [Google Scholar] [CrossRef]
  111. Tsutsumi, C.; Yasuda, H. Biodegradation of Copolymers Composed of Optically Active L-Lactide and (R)- or (S)-1-Methyltrimethylene Carbonate. J. Polym. Sci. Part A Polym. Chem. 2001, 39, 3916–3927. [Google Scholar] [CrossRef]
  112. Brignou, P.; Carpentier, J.-F.; Guillaume, S.M. Metal—And Organo-Catalyzed Ring-Opening Polymerisation of α-Methyl-Trimethylene Carbonate: Insights into the Microstructure of the Polycarbonate. Macromolecules 2011, 44, 5127–5135. [Google Scholar] [CrossRef]
  113. Hua, G.; Franzén, J.; Odelius, K. Phosphazene-Catalyzed Regioselective Ring-Opening Polymerisation of Rac-1-Methyl Trimethylene Carbonate: Colder and Less Is Better. Macromolecules 2019, 52, 2681–2690. [Google Scholar] [CrossRef]
  114. López-Vidal, E.M.; Gregory, G.L.; Kociok-Köhnc, G.; Buchard, A. Polymers from sugars and CS2: Synthesis and ring-opening polymerisation of sulfur-containing monomers derived from 2-deoxy-D-ribose and D-xylose. Polym. Chem. 2018, 9, 1577–1582. [Google Scholar] [CrossRef] [Green Version]
  115. Overberger, C.G.; Weise, J.K.J. Anionic ring-opening polymerization of thiolactones. Am. Chem. Soc. 1968, 90, 3533–3537. [Google Scholar] [CrossRef]
  116. Wu, H.-L.; Yang, J.-L.; Luo, M.; Wang, R.-Y.; Xu, J.-T.; Du, B.-Y.; Zhang, X.-H.; Darensbourg, D.J. Poly(trimethylene monothiocarbonate) from the Alternating Copolymerization of COS and Oxetane: A Semicrystalline Copolymer. Macromolecules 2016, 49, 8863–8868. [Google Scholar] [CrossRef]
  117. Luo, M.; Zhang, X.-H.; Darensbourg, D.J. Poly (monothiocarbonate)s from the alternating and regioselective copolymerization of carbonyl sulfide with epoxides. Acc. Chem. Res. 2016, 49, 2209–2219. [Google Scholar] [CrossRef] [PubMed]
  118. Byrne, C.M.; Allen, S.D.; Lobkovsky, E.B.; Coates, G.W.J. Alternating Copolymerization of Limonene Oxide and Carbon Dioxide. Am. Chem. Soc. 2004, 126, 11404–11405. [Google Scholar] [CrossRef] [Green Version]
  119. Carrodeguas, P.L.; González-Fabra, J.; Castro-Gómez, F.; Bo, C.; Kleij, A.W. Al(III) -catalysed formation of poly(limonene)carbonate: DFT analysis of the origin of stereoregularity. Chem. Eur. J. 2015, 21, 6115–6122. [Google Scholar] [CrossRef]
  120. Wambach, A.; Agarwal, S.; Greiner, A. Synthesis of Bio-based Polycarbonate by Copolymerisation of Menth-2-ene Oxide and CO2 with Exceptional Thermal Stability. ACS Sustain. Chem. Eng. 2020, 8, 14690–14693. [Google Scholar] [CrossRef]
  121. Tran, D.K.; Rashad, A.Z.; Darensbourg, D.J.; Wooley, K.L. Sustainable synthesis of CO2-derived polycarbonates from D-xylose. Polym. Chem. 2021, 12, 5271–5278. [Google Scholar] [CrossRef]
  122. Darensbourg, D.J.; Moncada, A.I. Tuning the Selectivity of the Oxetane and CO2 Coupling Process Catalyzed by (Salen)CrCl/n-Bu4NX: Cyclic Carbonate Formation vs Aliphatic Polycarbonate Production. Macromolecules 2010, 43, 5996–6003. [Google Scholar] [CrossRef]
  123. Darensbourg, D.J.; Moncada, A.I.; Wei, S.-H. Aliphatic Polycarbonates Produced from the Coupling of Carbon Dioxide and Oxetanes and Their Depolymerisation via Cyclic Carbonate Formation. Macromolecules 2011, 44, 2568–2576. [Google Scholar] [CrossRef]
  124. Park, H.J.; Ryu, C.Y.; Crivello, J.V. Photoinitiated cationic polymerisation of limonene 1,2-oxide and a-pinene oxide J. Polym. Sci. Part A Polym. Chem. 2013, 51, 109–117. [Google Scholar] [CrossRef]
  125. Sessini, V.; Palenzuela, M.; Damian, J.; Mosquera, M.E.G. Bio-based polyether from limonene oxide catalytic ROP as green polymeric plasticizer for PLA. Polymer 2020, 210, 123003. [Google Scholar] [CrossRef]
  126. Saxon, D.J.; Nasiri, M.; Mandal, M.; Maduskar, S.; Dauenhauer, P.J.; Cramer, C.J.; LaPointe, A.M.; Reineke, T.M. Architectural Control of Isosorbide-Based Polyethers via Ring-Opening Polymerisation. J. Am. Chem. Soc. 2019, 141, 5107–5111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Matt, L.; Liblikas, I.; Bonjour, O.; Jannasch, P.; Vares, L. Synthesis and anionic polymerisation of isosorbide mono-epoxides for linear bio-based polyethers. Polym. Chem. 2021, 12, 5937–5941. [Google Scholar] [CrossRef]
  128. Gandini, A.; Lacerda, T.M. From monomers to polymers from renewable resources: Recent advances. Prog. Polym. Sci. 2015, 48, 1–39. [Google Scholar] [CrossRef]
  129. Harvey, B.J.; Meylemans, H.A. The role of butanol in the development of sustainable fuel technologies. J. Chem. Technol. Biotechnol. 2011, 86, 2–9. [Google Scholar] [CrossRef]
  130. Herzberger, J.; Niederer, K.; Pohlit, H.; Seiwert, J.; Worm, M.; Wurm, F.R.; Frey, H. Polymerisation of Ethylene Oxide, Propylene Oxide, and Other Alkylene Oxides: Synthesis, Novel Polymer Architectures, and Bioconjugation. Chem. Rev. 2016, 116, 2170–2243. [Google Scholar] [CrossRef] [Green Version]
  131. PolyTHF® - Polytetrahydrofuran. Available online: https://chemicals.basf.com/global/en/Intermediates/Product_groups/PolyTHF.html (accessed on 31 March 2022).
  132. Sardon, H.; Mecerreyes, D.; Basterretxea, A.; Avérous, L.; Jehanno, C. From Lab to Market: Current Strategies for the Production of Bio-based Polyols. ACS Sustain. Chem. Eng. 2021, 9, 10664–10677. [Google Scholar] [CrossRef]
  133. Jacobsen, S.; Degee, P.H.; Fritz, H.G.; Jerome, D.R. Polylactide (PLA)-A New Way of Production. Polym. Eng. Sci. 1999, 39, 1311–1319. [Google Scholar] [CrossRef]
  134. Pradhan, R.; Reddy, M.; Diebel, W.; Erickson, L.; Misra, M.; Mohanty, A. Comparative compostability and biodegradation studies of various components of green composites and their blends in simulated aerobic composting bioreactor. Int. J. Plast. Technol. 2010, 14, 45–50. [Google Scholar] [CrossRef]
Scheme 1. General scheme for the ROP of cyclic esters (a) and ROCOP of epoxides with cyclic anhydrides (b).
Scheme 1. General scheme for the ROP of cyclic esters (a) and ROCOP of epoxides with cyclic anhydrides (b).
Suschem 03 00017 sch001
Figure 1. Main sustainable lactones employed as monomers in ring-opening polymerisation processes.
Figure 1. Main sustainable lactones employed as monomers in ring-opening polymerisation processes.
Suschem 03 00017 g001
Figure 2. Bis(imino)pyridine-Fe1–3, and guanidine-based Fe(II) Fe4–6 complexes active in the ROP of lactide.
Figure 2. Bis(imino)pyridine-Fe1–3, and guanidine-based Fe(II) Fe4–6 complexes active in the ROP of lactide.
Suschem 03 00017 g002
Figure 3. Synthesis of Fe(η6-C6H6)-bis(arylimino)-acenaphthene complex Fe7.
Figure 3. Synthesis of Fe(η6-C6H6)-bis(arylimino)-acenaphthene complex Fe7.
Suschem 03 00017 g003
Figure 4. General representation and structural variation of Fe(III)-salen complexes Fe8–15.
Figure 4. General representation and structural variation of Fe(III)-salen complexes Fe8–15.
Suschem 03 00017 g004
Figure 5. Fe-complex Fe16, and Ce-complexes Ce1–2 affording cyclic polyesters.
Figure 5. Fe-complex Fe16, and Ce-complexes Ce1–2 affording cyclic polyesters.
Suschem 03 00017 g005
Figure 6. Sodium complexes Na1–4 active in the ROP of ω-pentadecalactone.
Figure 6. Sodium complexes Na1–4 active in the ROP of ω-pentadecalactone.
Suschem 03 00017 g006
Figure 7. Organic bases used in the dual catalyst approach proposed for the ROP of lactones.
Figure 7. Organic bases used in the dual catalyst approach proposed for the ROP of lactones.
Suschem 03 00017 g007
Figure 8. Terpene-based anhydrides (a) and epoxides (b).
Figure 8. Terpene-based anhydrides (a) and epoxides (b).
Suschem 03 00017 g008
Figure 9. Complexes employed in the ROCOP of epoxides and anhydrides.
Figure 9. Complexes employed in the ROCOP of epoxides and anhydrides.
Suschem 03 00017 g009
Figure 10. Synthesis of β-elemene-based epoxides, corresponding functional semi-aromatic polyester poly(BEM-alt-PA), and its post-modification.
Figure 10. Synthesis of β-elemene-based epoxides, corresponding functional semi-aromatic polyester poly(BEM-alt-PA), and its post-modification.
Suschem 03 00017 g010
Figure 11. Fatty acid-based epoxides, and synthesis of corresponding functional polyesters.
Figure 11. Fatty acid-based epoxides, and synthesis of corresponding functional polyesters.
Suschem 03 00017 g011
Scheme 2. Polycarbonates synthesis via ring-opening of cyclic carbonates (a) and ring-opening copolymerisation of carbon dioxide and epoxides (b).
Scheme 2. Polycarbonates synthesis via ring-opening of cyclic carbonates (a) and ring-opening copolymerisation of carbon dioxide and epoxides (b).
Suschem 03 00017 sch002
Figure 12. Synthesis of d-glucose-based monomer CC1 and corresponding polycarbonate PCC1. Organo-catalysts structures are also shown for clarity.
Figure 12. Synthesis of d-glucose-based monomer CC1 and corresponding polycarbonate PCC1. Organo-catalysts structures are also shown for clarity.
Suschem 03 00017 g012
Figure 13. Synthesis of d-glucose-based monomer CC2 and corresponding polycarbonate PCC2, followed by post-functionalization toward hydroxylated- and acetate-polycarbonates PCC2-OH and PCC2-OAc, respectively.
Figure 13. Synthesis of d-glucose-based monomer CC2 and corresponding polycarbonate PCC2, followed by post-functionalization toward hydroxylated- and acetate-polycarbonates PCC2-OH and PCC2-OAc, respectively.
Suschem 03 00017 g013
Figure 14. Synthesis of d-glucose-based monomer CC3-R and corresponding polycarbonate PCC3-R.
Figure 14. Synthesis of d-glucose-based monomer CC3-R and corresponding polycarbonate PCC3-R.
Suschem 03 00017 g014
Figure 15. Synthesis of d-mannose-based monomer CC4 and corresponding polycarbonate PCC4.
Figure 15. Synthesis of d-mannose-based monomer CC4 and corresponding polycarbonate PCC4.
Suschem 03 00017 g015
Figure 16. Synthesis of thymidine-based monomer CC5 and corresponding polycarbonate PCC5.
Figure 16. Synthesis of thymidine-based monomer CC5 and corresponding polycarbonate PCC5.
Suschem 03 00017 g016
Figure 17. Synthesis of 2-deoxy-d-ribose-based monomer CC6 and corresponding polycarbonate PCC6.
Figure 17. Synthesis of 2-deoxy-d-ribose-based monomer CC6 and corresponding polycarbonate PCC6.
Suschem 03 00017 g017
Figure 18. Synthesis of erythritol based monomer CC7 and corresponding polycarbonates P-cis-CC7 and P-trans-CC7.
Figure 18. Synthesis of erythritol based monomer CC7 and corresponding polycarbonates P-cis-CC7 and P-trans-CC7.
Suschem 03 00017 g018
Figure 19. Synthesis of bio-1,3-butylene glycol-based monomer CC8 and corresponding polycarbonate PCC8. Structures of catalysts Zn1, Al3 and Al4 are also shown on the bottom.
Figure 19. Synthesis of bio-1,3-butylene glycol-based monomer CC8 and corresponding polycarbonate PCC8. Structures of catalysts Zn1, Al3 and Al4 are also shown on the bottom.
Suschem 03 00017 g019
Figure 20. Synthesis of 2-deoxy-d-ribose-based monomer TC1 (a), d-xylofuranose-based monomers TC2 and TC3 (b), and corresponding poly-thiocarbonates PTC1, PTC2 and PTC3.
Figure 20. Synthesis of 2-deoxy-d-ribose-based monomer TC1 (a), d-xylofuranose-based monomers TC2 and TC3 (b), and corresponding poly-thiocarbonates PTC1, PTC2 and PTC3.
Suschem 03 00017 g020
Figure 21. Conversion of the cis- and trans-forms of limonene oxide to poly(limonene carbonate).
Figure 21. Conversion of the cis- and trans-forms of limonene oxide to poly(limonene carbonate).
Suschem 03 00017 g021
Figure 22. Synthesis L-menthol-based epoxide MO, and corresponding polycarbonate PMC. Structure of catalyst Zn2 is also shown.
Figure 22. Synthesis L-menthol-based epoxide MO, and corresponding polycarbonate PMC. Structure of catalyst Zn2 is also shown.
Suschem 03 00017 g022
Figure 23. Synthesis d-xylose-based oxetane XO.
Figure 23. Synthesis d-xylose-based oxetane XO.
Suschem 03 00017 g023
Figure 24. Synthesis d-xylose-based carbonate XC, and corresponding polycarbonate PCC via XO/CO2 ROCOP (top), and XC ROP (bottom). Structure of catalyst Cr2, and cocatalysts are also shown.
Figure 24. Synthesis d-xylose-based carbonate XC, and corresponding polycarbonate PCC via XO/CO2 ROCOP (top), and XC ROP (bottom). Structure of catalyst Cr2, and cocatalysts are also shown.
Suschem 03 00017 g024
Figure 25. Polymerisation of LO to PLC catalysed by Al6.
Figure 25. Polymerisation of LO to PLC catalysed by Al6.
Suschem 03 00017 g025
Figure 26. Cationic ROP of 1,4:2,5:3,6-trianhydro-d-mannitol.
Figure 26. Cationic ROP of 1,4:2,5:3,6-trianhydro-d-mannitol.
Suschem 03 00017 g026
Figure 27. ROP of isosorbide epoxides.
Figure 27. ROP of isosorbide epoxides.
Suschem 03 00017 g027
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Santoro, O.; Izzo, L.; Della Monica, F. Recent Advances in RO(CO)P of Bio-Based Monomers. Sustain. Chem. 2022, 3, 259-285. https://doi.org/10.3390/suschem3020017

AMA Style

Santoro O, Izzo L, Della Monica F. Recent Advances in RO(CO)P of Bio-Based Monomers. Sustainable Chemistry. 2022; 3(2):259-285. https://doi.org/10.3390/suschem3020017

Chicago/Turabian Style

Santoro, Orlando, Lorella Izzo, and Francesco Della Monica. 2022. "Recent Advances in RO(CO)P of Bio-Based Monomers" Sustainable Chemistry 3, no. 2: 259-285. https://doi.org/10.3390/suschem3020017

Article Metrics

Back to TopTop