Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2022 Jul 21;55(15):1997–2010. doi: 10.1021/acs.accounts.2c00197

Synergic Heterodinuclear Catalysts for the Ring-Opening Copolymerization (ROCOP) of Epoxides, Carbon Dioxide, and Anhydrides

Wilfred T Diment 1, Wouter Lindeboom 1, Francesca Fiorentini 1, Arron C Deacy 1, Charlotte K Williams 1,*
PMCID: PMC9350912  PMID: 35863044

Conspectus

graphic file with name ar2c00197_0013.jpg

The development of sustainable plastic materials is an essential target of chemistry in the 21st century. Key objectives toward this goal include utilizing sustainable monomers and the development of polymers that can be chemically recycled/degraded. Polycarbonates synthesized from the ring-opening copolymerization (ROCOP) of epoxides and CO2, and polyesters synthesized from the ROCOP of epoxides and anhydrides, meet these criteria. Despite this, designing efficient catalysts for these processes remains challenging. Typical issues include the requirement for high catalyst loading; low catalytic activities in comparison with other commercialized polymerizations; and the requirement of costly, toxic cocatalysts. The development of efficient catalysts for both types of ROCOP is highly desirable. This Account details our work on the development of catalysts for these two related polymerizations and, in particular, focuses on dinuclear complexes, which are typically applied without any cocatalyst. We have developed mechanistic hypotheses in tandem with our catalysts, and throughout the Account, we describe the kinetic, computational, and structure–activity studies that underpin the performance of these catalysts. Our initial research on homodinuclear M(II)M(II) complexes for cyclohexene oxide (CHO)/CO2 ROCOP provided data to support a chain shuttling catalytic mechanism, which implied different roles for the two metals in the catalysis. This mechanistic hypothesis inspired the development of mixed-metal, heterodinuclear catalysts. The first of this class of catalysts was a heterodinuclear Zn(II)Mg(II) complex, which showed higher rates than either of the homodinuclear [Zn(II)Zn(II) and Mg(II)Mg(II)] analogues for CHO/CO2 ROCOP. Expanding on this finding, we subsequently developed a Co(II)Mg(II) complex that showed field leading rates for CHO/CO2 ROCOP and allowed for unique insight into the role of the two metals in this complex, where it was established that the Mg(II) center reduced transition state entropy and the Co(II) center reduced transition state enthalpy. Following these discoveries, we subsequently developed a range of heterodinuclear M(III)M(I) catalysts that were capable of catalyzing a broad range of copolymerizations, including the ring-opening copolymerization of CHO/CO2, propylene oxide (PO)/CO2, and CHO/phthalic anhydride (PA). Catalysts featuring Co(III)K(I) and Al(III)K(I) were found to be exceptionally effective for PO/CO2 and CHO/PA ROCOP, respectively. Such M(III)M(I) complexes operate through a dinuclear metalate mechanism, where the M(III) binds and activates monomers while the M(I) species binds the polymer change in close proximity to allow for insertion into the activated monomer. Our research illustrates how careful catalyst design can yield highly efficient systems and how the development of mechanistic understanding aids this process. Avenues of future research are also discussed, including the applicability of these heterodinuclear catalysts in the synthesis of sustainable materials.

Key References

  • Kember M. R.; Knight P. D.; Reung P. T. R.; Williams C. K.. Highly Active Dizinc Catalyst for the Copolymerization of Carbon Dioxide and Cyclohexene Oxide at One Atmosphere Pressure. Angew. Chem. Int. Ed., 2009, 48, 931–933.1

    A highly active Zn(II)Zn(II) complex, coordinated by a macrocyclic ligand, for cyclohexene oxide/CO2ring-opening copolymerization (ROCOP); the catalyst turns over at low CO2pressure (1 bar).

  • Deacy A. C.; Kilpatrick A. F. R.; Regoutz A.; Williams C. K.. Understanding metal synergy in heterodinuclear catalysts for the copolymerization of CO2 and epoxides. Nat. Chem., 2020, 12, 372–380.2

    A highly active heterodinuclear Mg(II)Co(II) catalyst for cyclohexene oxide/CO2ROCOP. Kinetic investigations reveal that synergy depends upon different roles for each metal.

  • Deacy A. C.; Moreby E.; Phanopoulos A.; Williams C. K.. Co(III)/Alkali-Metal(I) Heterodinuclear Catalysts for the Ring-Opening Copolymerization of CO2 and Propylene Oxide. J. Am. Chem. Soc., 2020, 142, 19150–19160.3

    A series of heterodinuclear catalysts, Co(III)M(I), coordinated by a different macrocyclic ligand, for propylene oxide/CO2ROCOP. The Co(III)K(I) catalyst shows both high activities and selectivities and it operates with chain transfer agents and a range of epoxide monomers.

  • Diment W. T.; Gregory G. L.; Kerr R. W. F.; Phanopoulos A.; Buchard A.; Williams C. K.. Catalytic Synergy Using Al(III) and Group 1 Metals to Accelerate Epoxide and Anhydride Ring-Opening Copolymerizations. ACS Catal. 2021, 11, 12532–12542.4

    A series of heterodinuclear Al(III)M(I) catalysts, coordinated by a Schiff base ligand, for epoxide/anhydride ROCOP. The Al(III)K(I) catalyst shows exceptional activities, loading tolerance, and broad monomer scope. Kinetic and computational calculations reveal a mechanism operating via a metalate intermediate.

Introduction

The efficient conversion of carbon dioxide to useful products is a linchpin of sustainable chemistry. Carbon dioxide is an abundant and low-cost C1 feedstock and is a waste product of many industrial processes, including combustion and fermentation.5 One promising avenue for its valorization and recycling is to produce polycarbonates via the ring-opening copolymerization (ROCOP) of carbon dioxide with epoxides (Scheme 1).6 The related ROCOP of epoxides and anhydrides is a broadly applicable and controlled process delivering polyesters. In both reactions, the polymerization catalyst is essential for cost-effective manufacturing and to control the polymers’ properties.

Scheme 1. Summary of Ring-Opening Copolymerization (ROCOP) for Epoxide/CO2.

Scheme 1

An analogous mechanism is followed for epoxide/anhydride ROCOP.

This Account describes some of our discoveries, developments, and growing understanding of dinuclear ring-opening copolymerization catalysts. Prior to presenting such a personal perspective, we would like to acknowledge the pioneering initial discoveries of R.F. Fischer of Shell Development Company (epoxide/anhydride ROCOP) and Shohei Inoue of the University of Tokyo (carbon dioxide/epoxide ROCOP).7,8 This field of polymerization catalysis has since benefited from the innovation and creativity of many other groups worldwide, including those of Kuran, Darensbourg, Nozaki, Coates, Lu, Lee, Wang, Koning, Gürtler, Hadjichristidis, Gnanou, Chisholm, Kleij, Frey, Grinstaff, Reiger, Leitner, Wu, Kozak, Luinstra, Ree, and many more. Our Account cannot describe all their innovations and influences; rather, the reader is directed to some excellent comprehensive reviews.6,913

Before designing a new catalyst, it is essential to consider the polymerizations’ elementary steps. Epoxide/carbon dioxide ROCOP involves this series of reactions: initiation, propagation, and chain transfer, with backbiting and polyether formation both possible as side-reactions. Epoxide/anhydride ROCOP follows an analogous mechanism with CO2 replaced by anhydride (Scheme 1).6 During initiation, an activated epoxide molecule is ring-opened by a nucleophile or initiating group, X (typically an acetate, alkoxide, halide, or carboxylate), which generates a metal alkoxide intermediate. Subsequent, rapid insertion of CO2 into the metal alkoxide generates a metal–carbonate intermediate and starts the propagation catalytic cycle. During propagation, the carbonate intermediate ring-opens another activated epoxide to (re)form a metal alkoxide, which then inserts CO2 to (re)form the carbonate; multiple repetitions build up the polymer.

We target controlled polymerization catalysts, meaning that the polymer molar mass values are predictable, evolve linearly, and show narrow molar mass distributions (Đ < 1.2). These are not strictly living polymerizations, but are sometimes described as “immortal,” because chain transfer reactions are endemic (Scheme 1).14 They involve a rapid exchange of metal alkoxide intermediates with protic species that are either present in or added to the reaction. Common chain transfer agents (CTAs) are alcohols, and the exchange (transfer) reactions typically occur faster than propagation and therefore do not influence polymerization kinetics.15 Rather, chain transfer controls molar mass and chain end-group chemistry without increasing the quantity of catalyst. Although this Account principally focuses on the activity and selectivity of our catalysts, it should be noted that in all cases our catalysts display good reaction control (Đ < 1.2) and immortal polymerization behavior, which allows for controllable molar mass polymers from polyols (<5 kg mol–1) up to higher molar mass samples (50 kg mol–1) depending on the amount of CTA applied to the polymerizations.

In carbon dioxide/epoxide ROCOP, the major side product is trans-cyclic carbonate, which is usually formed by intramolecular reaction (backbiting) of the metal alkoxide with the polymer chain (Scheme 1). Another side-reaction is the metal alkoxide reaction with epoxide to form (poly)ether linkages. This process decreases the overall CO2 uptake but can be useful to moderate polymers’ thermal-mechanical properties.16 In epoxide/anhydride ROCOP, in contrast to epoxide/CO2 ROCOP, backbiting reactions usually have higher energy barriers, so cyclic side products (lactones) are not usually observed; however, polyether formation can still be observed as a side reaction.

Effective polymerization catalysts show high productivity [turnover number (TON)] and activity [turnover frequency (TOF)] at minimal catalyst loadings and, for reactions involving CO2, at pressures of <10 bar, which allows for the retrofitting of current industrial plants. They deliver polymers with predictable molar mass, monomer sequence (% carbonate/ester selectivity), end-group chemistry, and minimal side products (% polymer selectivity). They should also be inexpensive, lightweight, straightforward to synthesize, nontoxic, redox inactive, and removable/recoverable after use. Our team has investigated dinuclear catalysts comprising light main-group (particularly s-block) and first row transition metal elements. These complexes apply ancillary ligands featuring phenolate and amines/imines since these “hard” donors form stable, activated complexes. The internuclear separations of these complexes are in the range of 3–4 Å, which enables intermetallic electronic communication and avoids the need for an addition of cocatalyst, the use of which can limit activity at low catalyst loading.

Our initial hypothesis was that dinuclear catalysts should show lower barriers to epoxide ring-opening and heteroallene insertion reactions compared with mononuclear active sites. It was inspired by nature’s bimetallic metalloenzymes, e.g., di-Zn(II) hydrolases, phosphatases, or peptidases, which follow mechanisms involving dinuclear substrate activation and nucleophilic attack steps (often with water).17,18 It was also underpinned by outstanding work from other researchers in the field investigating di- or multinuclear catalysts. Since the late 1970s, there were reports of multinuclear Zn(II) catalysts,19 and particular mention is warranted of Zn(II)phenoxides from Darensbourg et al., Zn(II) β-diimine (BDI) dimeric catalysts from Coates et al., anilido aniline di-Zn(II) catalysts from Lee et al., and di-Zn(II) complexes from Nozaki et al.2024 The selection of the ancillary ligand is critical to control both catalytic stability and performance. Our approach has been to explore dinucleating Schiff base ligands; this Account focuses on three of them: L1 for cyclohexene oxide (CHO)/CO2; L2 for propylene oxide (PO)/CO2; and L3 for CHO/phthalic anhydride (PA) (Figure 1). Henceforth, research using these ligands is described, and for efficiency, complexes are referred to as M1(oxidation state)M2(oxidation state).

Figure 1.

Figure 1

Ligands and some of the key metal combinations used in dinuclear catalysts for epoxide/CO2 and epoxide/anhydride ROCOP. X = anionic coligand (typically acetate, halide, or benzoate).

Epoxide/CO2 ROCOP with L1

In 2009, we reported a Zn(II)Zn(II) complex, using L1 as an ancillary ligand, that was active for the copolymerization of carbon dioxide with CHO under 1 bar of pressure.1 It showed a TOF of 25 h–1 when using 0.1 mol % catalyst at 100 °C (1:1000, catalyst/CHO), even when using unoptimized conditions, i.e., a glass Schlenk line and magnetic stirring. It also showed quantitative carbon dioxide uptake and high polymer selectivity; it was a rare example of a ROCOP catalyst active at low CO2 pressure. The polymer product, poly(cyclohexene carbonate) (PCHC), shows a high glass transition temperature (Tg) (105– 115 °C), tensile strength (42 MPa), and Young’s modulus (3.6 GPa).25

The macrocyclic ancillary ligand was also very effective in coordinating a range of s-block and first row transition metals with a metal–metal distance of ∼3.0 Å, as confirmed by single crystal X-ray diffraction.2630 Our subsequent reports described a range of homodinuclear catalysts, in some cases as the first examples of catalysts using those metals/oxidation states: Mg(II)Mg(II),31 Fe(III)Fe(III),27 Co(II)Co(II), and Co(II)Co(III).26,28 We were particularly interested in the Mg(II)Mg(II) catalyst, which was twice as active as the Zn(II)Zn(II) catalyst, yet Mg(II) is 1/3 of Zn(II)’s mass and also shows 2000 times higher elemental abundance.

These dinuclear catalysts showed unparalleled activity using low-pressure carbon dioxide sources, thereby obviating the use of specialized high-pressure steel equipment, and are tolerant of large amounts of CTA, including water. These features are attractive for the coupling of carbon dioxide use with captured emissions. At the time of the study, the UK operated an amine-based carbon capture and storage demonstrator plant at Ferrybridge, a coal-fired power station, and we were able to source captured carbon dioxide emissions from the plant in a Tedlar bag. Both the Zn(II)Zn(II) and Mg(II)Mg(II) catalysts showed impressive performances using this impure carbon dioxide gas, even under unoptimized conditions, i.e., using heavy textbooks as weights to force the gas from the Tedlar bag into the Schlenk tube. We also investigated the effects of deliberately added impurities to the carbon dioxide: the catalysts remained active using carbon dioxide contaminated by high concentrations of water, carbon monoxide, nitrogen, oxygen, amines, and thiols. The study demonstrates the potential to couple carbon dioxide utilization with large-scale capture and storage technologies.32

A detailed kinetic investigation including elucidation of the polymerization rate law using the Zn(II)Zn(II) catalyst was undertaken to understand the performances of these dinuclear catalysts (Figure 2, top).33 The copolymerization shows first-order rate dependencies in both the catalyst and epoxide concentrations, but is zero-order in the carbon dioxide pressure (1–40 bar). The rate-limiting step was proposed as CHO ring-opening by the Zn-carbonate intermediate.

Figure 2.

Figure 2

Data for CHO/CO2 ROCOP kinetics and its proposed mechanism using Zn(II)Zn(II). Kinetic studies, top panel: (a) plot of initial rate versus catalyst concentration, which indicates a first-order dependence; (b) plot of initial rate versus epoxide concentration, which indicates a first-order dependence; (c) plot of initial rate versus carbon dioxide pressure, which indicates zero-order dependence. All images adapted with permission from ref (33). Copyright 2011 American Chemical Society. Computational mechanism, bottom panel: the potential energy surface, determined by DFT calculations [theory level: ωB97XD/6-31G(d)], for the alternating copolymerization.

In collaboration with Henry Rzepa at Imperial College London, a DFT investigation of the polymerization mechanism examined all possible permutations of four monomer insertions; the calculations accounted for the polar reaction medium (neat epoxide) and temperature (Figure 1, bottom).34 The study suggests that initiation occurs from Zn(II)Zn(II) featuring two bridging acetate ligands; in situ IR spectroscopy supported this speciation (1). CHO coordination at Zn(II) (M2) (2) is slightly exothermic (−4.0 kcal mol–1) and is followed by nucleophilic attack of a Zn-acetate (or M1-carbonate) (3TS) to form the Zn-alkoxide (M2) (4). In situ IR spectroscopy suggests that only one of the two acetate groups initiates, while the other remains coordinated and counterbalances the change in coordination chemistry of the growing polymer chain. Carbon dioxide insertion proceeds via initial ligand activation and generation of a zwitterionic intermediate, which rearranges (5TS) to a Zn-carbonate (M1) species (6). The calculations show that CHO ring-opening is the rate-determining step, with a transition state barrier of 22.9 kcal mol–1, compared with a lower barrier to carbon dioxide insertion (16.6 kcal mol–1). The DFT barrier for the rate-limiting step is close to the value determined experimentally (25.7 kcal mol–1), thereby giving confidence in the mechanism.33 The mechanism shows that the growing polymer chain exchanges twice between the two metal centers per catalytic cycle; it was, therefore, termed the “chain shuttling mechanism.”34 The movement of the growing polymer chain between metal sites is facilitated by the bridging acetate ligand; in situ IR studies reveal both Zn-coordinated and polymer chain end-functionalized acetate groups during propagation. The mechanism shows different roles for the two Zn(II) centers (hence the use of M1 and M2): (1) epoxide coordination and activation and/or (2) provision of a labile, nucleophilic carbonate group. The chain shuttling mechanism inspired the investigation of heterodinuclear metal complexes, i.e., where M1 differs from M2.

The initial investigation targeted Mg(II)Zn(II) catalysts because of the precedent for the homodinuclear catalysts, the ability to monitor complexation using 1H NMR spectroscopy, and the rate enhancement shown by Mg(II)Mg(II).35 The initial complex synthesis involved adding each metal precursor sequentially to the macrocycle at room temperature. It formed a 1:2:1 mixture of Zn(II)Zn(II), Mg(II)Zn(II), and Mg(II)Mg(II), as determined by both 1H NMR spectroscopy and mass spectrometry. The statistical mixture of catalysts showed higher rates, under identical conditions, than either of the homodinuclear catalysts or a 1:1 mixture of the two catalysts. These results suggested some intermetallic synergy and motivated the isolation of a pure heterodinuclear catalyst.

Adapting the synthetic procedure allowed isolation of a pure heterodinuclear Mg(II)Zn(II) catalyst.36 The Mg(II)Zn(II) catalyst showed a TOF of 34 h–1 (1 bar of CO2, 80 °C) and was twice the activity of Mg(II)Mg(II) and 5 times faster than a 1:1 mixture of Mg(II)Mg(II) and Zn(II)Zn(II) (Table 1, entries 1–3).

Table 1. Homo- and Heterodinuclear CHO/CO2 ROCOP Catalysts Using L1a.

# catalyst time (h) conv.b(%) CO2c (%) polym.d (%) TONe TOFf (h–1) Mn [Đ]g (g mol–1)
1 Mg(II)Mg(II) 10 15 >99 >99 151 15 800 [1.13]
2 Mg(II)Zn(II) 10 34 >99 >99 344 34 3100 [1.14]
3 Mg(II)Mg(II)/Zn(II)Zn(II)(1:1) 10 7 >99 >99 72 7 <500
4h Mg(II)Zn(II) 3 37 >99 >99 372 124 10 200 [1.02] 4750 [1.08]
5i Mg(II)Zn(II) 0.5 44 >99 >99 4415 8830 44 400 [1.04] 21 200 [1.05]
a

Reaction conditions: [Cat]/[CHO] = 1:1000, neat epoxide, 1 bar CO2, 80 °C.36 Note that under these conditions, the Zn(II)Zn(II) did not initiate polymerization, as the study utilized bromide coligands.

b

Conversion of epoxide, determined by 1H NMR.

c

Selectivity for carbonate vs ether, determined by 1H NMR.

d

Selectivity for polymer vs cyclic carbonate, determined by 1H NMR.

e

Turnover number (TON) = total number of moles of epoxide consumed/mol of catalyst.

f

Turnover frequency (TOF) = TON/time (hours).

g

Number average molecular weight [dispersity], determined by GPC.

h

Conditions: [Cat]/[CHO] = 1:1000, 1 bar CO2, 80 °C.37

i

Conditions: [Cat]/[CHO] = 1:10000, 20 bar CO2, 120 °C.37

Subsequent optimization of the catalyst coligands and of the reaction conditions yielded increases in catalytic activities.37 At 1 bar of CO2 pressure and 80 °C, a Mg(II)Zn(II) catalyst bearing nitro-benzoate coligands gave a TOF of 124 h–1 (Table 1, entry 4). The same catalyst yielded a TOF of 8830 h–1 when CO2 pressure was increased to 20 bar, the temperature was increased to 120 °C, and the reaction was performed with impeller stirring (Table 1, entry 5).

Following this initial study, in order to better understand which metal combinations were synergic, Zn(II) was combined with other alkali [Li(I), Na(I), K(I)], alkaline earth [Ca(II)], and Group 13 [Al(III), Ga(III), In(III)] metals to form heterodinuclear complexes.29,30 This work sheds light upon synthesis procedures: reactions were conducted either under kinetic (−78 °C) or thermodynamic (100 °C) control in the second metalation step. All the reactions resulted in quantitative heterodinuclear complex formation (>95% by NMR spectroscopy), and almost all the complexes, except Zn(II)Na(I), are thermodynamically more stable than their homodinuclear analogues. The high selectivity and stability for heterodinuclear complexes was at first quite surprising because of the symmetrical macrocycle with identical metal coordination sites, but is attributed to the additional stability arising from polarized M1–O–M2 interactions enabled by the phenoxide bridges. Nonetheless, in nearly all the cases, the heterodinuclear complexes were less active than the Zn(II)Zn(II) catalyst, with the exception being the synergic Zn(II)Mg(II) catalyst.

Rationalizing the lack of synergy is, of course, challenging but may arise from factors such as saturated metal coordination spheres (Group 13), large ionic radii metals preventing planar coordination [Ca(II)], or small ionic radii metals forming heterodinuclear aggregates [Li(I)]. A molecular structure of a Mg(II)Zn(II) precatalyst, determined by single crystal X-ray diffraction, showed THF coordination at Mg(II), which may well be relevant to epoxide binding during catalysis (Figure 3a).

Figure 3.

Figure 3

Data for heterodinuclear M(II)M(II) complexes applied in CHO/CO2 ROCOP. (a) Molecular structure of a Zn(II)Mg(II) complex showing THF bound to the Mg(II) center as a mimic for epoxide coordination. Image adopted with permission from ref (29). Copyright 2018 American Chemical Society. (b) Cyclic voltammetry plots showing oxidations and establishing the Mg(II)Co(II) complex. Image adopted with permission from ref (2). Copyright 2020 Springer Nature. (c) Plot comparing transition state enthalpy (ΔH) and entropy (ΔS) barriers for Mg(II)Mg(II), Co(II)Co(II), and Mg(II)Co(II), which indicates the basis for heterodinuclear synergy. Image adopted with permission from ref (2). Copyright 2020 Springer Nature. (d) Plot showing the variation in propagation rate constants (kp) for various heterodinuclear complexes, Mg(II)M(II), in CHO/CO2 ROCOP. Image adopted with permission from ref (38). Copyright 2022 Wiley.

Following the investigation of Zn(II)-main group element combinations, we targeted Mg(II)M′(II) complexes, where M′ is a first row transition metal. Some years earlier, we reported a Co(II)Co(II) catalyst that showed good activity, and understanding whether synergic effects could boost its performance was a priority. Thus, the Mg(II)Co(II) catalyst was synthesized using the sequential metalation approach; quantifying the heterodinuclear complex formation was challenging because of the paramagnetism of Co(II).2 The complex was characterized using SQUID, cyclic voltammetry, and XPS; all techniques indicated a pure heterodinuclear complex formation. For example, Co(II)Co(II) shows two cobalt-centered oxidations, while Mg(II)Mg(II) shows only an irreversible, high-energy ligand oxidation, and Mg(II)Co(II) displays only one metal oxidation (Figure 3b).

Under 1 bar of CO2 pressure and at 120 °C, Mg(II)Co(II) is significantly faster than Mg(II)Mg(II), Co(II)Co(II), or physical mixtures of both (Table 2, entries 1–3). It showed a maximum activity of 1205 h–1, maintained quantitative CO2 uptake and polymer selectivity, and is at the forefront of low-pressure catalysts in the field.

Table 2. Homo- and Heterodinuclear CHO/CO2 ROCOP Catalysts Using L1a.

# catalyst p(CO2) (bar) conv.b (%) CO2c (%) polym.d (%) TONe TOF (h–1)f
1 Mg(II)Mg(II) 1 18 >99 >99 368 368
2 Co(II)Co(II) 1 36 >99 96 712 712
3 Mg(II)Co(II) 1 25 >99 99 502 1205
4 Mg(II)Mg(II) 20 91 >99 >99 1820 1060
5 Co(II)Co(II) 20 94 >99 >99 1880 4200
6 Mg(II)Co(II) 20 95 >99 >99 1900 7200
7g Mg(II)Co(II) 20 96 >99 >99 1920 12460
a

Reaction conditions: [Cat]/[trans-1,2-cyclohexane diol (CHD)]/[CHO] = 1:20:2000, neat epoxide, 120 °C.

b

Conversion of epoxide, determined by 1H NMR.

c

Selectivity for carbonate vs ether, determined by 1H NMR.

d

Selectivity for polymer vs cyclic carbonate, determined by 1H NMR.

e

Turnover number (TON) = total number of moles of epoxide consumed/mol of catalyst.

f

Turnover frequency (TOF) = TON/time (hours).

g

[Cat]/[CHD]/[CHO] = 1:20:2000, 3 M CHO in diethylcarbonate, 140 °C.2

Increasing the CO2 pressure (20 bar) while maintaining a temperature of 120 °C significantly increased the rate, primarily because of the enhanced mechanical stirring in high pressure equipment, which overcame the diffusion limitations at low pressure (Table 2, entries 4–6). Such high-pressure polymerizations can also be conducted at higher temperature (140 °C) without forming any cyclic carbonate side products. Under such conditions, a very high TOF of 12 460 h–1 was observed: this remains the most active catalyst yet reported for perfectly alternating PCHC (Table 2, entry 7). Further, the impressive activity was maintained under high dilution, and conditions were optimized to allow for complete epoxide conversion (TON = 1920).

Undertaking polymerization kinetic analyses at different temperatures enabled determination of the transition state enthalpy (ΔH) and entropy (ΔS) barriers (Figure 3c). Comparing the barriers for the homo- vs heterodinuclear catalysts revealed that Co(II)Co(II) has a lower transition state enthalpy and that Mg(II)Mg(II) has a lower entropy barrier. Accordingly, Mg(II)Co(II) shows the best of both metals, with the Co(II) center reducing the reaction enthalpy, and the Mg(II) center reducing the reaction entropy. The data imply that each metal has a distinctive role in catalysis and that the bridging phenolate groups help moderate the active site electronics. The rate-determining step involves CHO coordination at Mg(II), which is attacked by the Co(II)-carbonate, and the reaction follows the chain shuttling mechanism.

The synergic combination of s-block Mg(II) and Co(II) motivated the investigation of other Mg(II) transition-metal(II) catalysts.38 All transition metals were applied in the +2 oxidation state to (1) allow for a direct comparison with Mg(II)Co(II) and Mg(II)Zn(II) catalysts and (2) ensure all catalysts applied commercial M(OAc)2 precursors; hence, the metals featured common initiating groups. The catalysts were synthesized using the sequential metalation route, whereby Mg(II) was first coordinated by the macrocycle, followed by the addition of the second transition metal and heating at 100 °C for 16 h. The complexes were characterized by mass spectrometry, IR spectroscopy, and cyclic voltammetry. All the heterodinuclear complexes were catalysts for the ring-opening copolymerization of carbon dioxide with CHO and most were more active than Mg(II)Mg(II). The Mg(II)Mn(II) catalyst was highly active, which was unexpected since prior research showed that Mn(III) porphyrin or salen complexes are poorly active.39 The Mg(II)Fe(II) catalyst displayed exceptional activity, equivalent to that obtained for the Mg(II)Co(II) catalyst, and is significant because iron is ∼3000 times more abundant, cheaper, and less toxic than cobalt. The series of complexes show a “volcano” type activity profile: the best performances arise using midrow transition metals (Figure 3d).

Epoxide/CO2 ROCOP using L2

Several research teams have investigated s-block heterodinuclear metal complexes of L2, particularly focusing on the influence of local electric fields on electron, proton, and hydrogen atom transfers.4043 We selected L2 for heterodinuclear polymerization catalysis using transition metals(II/III) and s-block metals(I/II) (Figure 4, left). Our prior work using L1 showed that these metal combinations were ineffective, but the “crown ether” binding pocket in L2 is much better suited to s-block metal coordination chemistry. First, L2 complexes using four different metal combinations were investigated for CHO/CO2 ROCOP: Zn(II)Na(I), Mg(II)Na(I), Co(III)Na(I), and Zn(II)Mg(II).44 The Zn(II)Mg(II) complex was completely inactive, in contrast to the synergy observed using these metals with L1 (vide supra).35 Mg(II)Na(I) also showed a low activity, but Zn(II)Na(I) showed promising performances (TOF = 29 h–1, PCHC selectivity = 94%, 1 bar CO2, 0.025 mol % catalyst). The best catalyst was Co(III)Na(I), which showed a very high activity (TOF = 1590 h–1, PCHC selectivity >99%, 100 °C, 1 bar CO2, 0.025 mol % catalyst); it is the most active low-pressure CHO/CO2 ROCOP catalyst reported to date.

Figure 4.

Figure 4

Heterodinuclear Co(III)M(I) catalysts, coordinated by L2, for epoxide/CO2 ROCOP (left). Solid-state structure of Co(III)K(I) with acetate bond lengths and representative structure showing anionic binding of acetates to Co(III) center (right).

Our group also investigated propene oxide (PO)/CO2 ROCOP, since the monomer is a commodity chemical, and poly(propene carbonate) is an elastomer/plastomer (Tg = 30–40 °C).45 PO/CO2 ROCOP is challenging because PO has a lower ring strain than CHO, which reduces the driving force for polycarbonate formation, and because it also shows a lower barrier to cyclic carbonate formation. As a result, catalytic activities and selectivities for PO/CO2 ROCOP are generally lower than for equivalent reactions using CHO, and many L1 catalysts are inactive for PO/CO2 ROCOP.

In 2020, we reported the first heterodinuclear catalysts for PO/CO2 ROCOP.46 The complexes feature Co(III) coordinated within the Schiff-base pocket and a Group 1 metal bound by the ether groups.

Co(III)K(I) and Co(III)Na(I) both show “cobaltate” solid-state structures, i.e., where both the acetate ligands coordinate as anions to the Co(III) (Figure 4, right).44,46 The X-ray diffraction data show significantly longer acetate C–O bond lengths for the oxygen atoms bound to Co(III) (i.e., single bonds and O anionic donors) and shorter acetate C–O bond lengths for the oxygen atoms bound to K(I) or Na(I) centers (i.e., double bonds and O dative covalent donors). The cobaltate structures suggest that the Co(III) centers are strong Lewis acids and are important to understanding the polymerization mechanisms, for which DFT investigations in collaboration with Buchard and Phanopoulos are underway. Further, both complexes have intermetallic separations between 3–4 Å [M(I) = Na, 3.388 Å; K, 3.698 Å], similar to other highly active dinuclear catalysts we have reported (vide supra).

All the new heterodinuclear complexes are active catalysts for PO/CO2 ROCOP (Table 3). The Co(III)K(I) catalyst is the most active and selective and achieves a TOF of 800 h–1 at low catalyst loading (0.025 mol % catalyst, 30 bar CO2, 70 °C; Table 3, entry 2). Heterodinuclear catalysts applying the larger ionic radii of Rb(I) or Cs(I) were much less active than analogous Na(I) or K(I) complexes, perhaps due to “poor fit” of the s-block metal with the crown ether portion of L2 (Table 3, entries 1–5). The exceptional performance of Co(III)K(I) likely arises from a beneficial combination of metal sizes, intermetallic communication, and coordination chemistry; it is noted that K(I) has the highest binding affinity for 18-crown-6 and thus may be the most compatible s-block metal for L2. The importance of using L2 is demonstrated by the finding that a 1:1 mixture of [(salen)Co(III)X] complex and 18-crown-6/KI, reported earlier by Lu and co-workers, mostly formed cyclic carbonate (41% PPC selectivity) with much lower activity than Co(III)K(I) (Table 3, entry 6). This result highlights the importance of the ligand coordination chemistry and adjacency effects between the metals to maximize both polymerization selectivity and activity.

Table 3. Heterodinuclear PO/CO2 ROCOP catalysts using L2a.

# catalyst time (h) conv.b (%) CO2c (%) polym.d (%) TONe TOFf (h–1) Mn [Đ]g (g mol–1)
1 Co(III)Na(I) 5.0 15 >99 79 600 120 2300 [1.08]
2 Co(III)K(I) 4.0 34 >99 98 1360 340 5900 [1.10]
3h Co(III)K(I) 1.4 28 >99 93 1120 800 5800 [1.07]
4 Co(III)Rb(I) 23 31 >99 91 1240 54 6500 [1.07]
5 Co(III)Cs(I) 23 27 >99 84 1080 57 5600 [1.08]
6i [(salen)Co(2,4-DNP)] + 18C6/KI 3.0 27 >99 41 540 182 4700 [1.43]
a

Reaction conditions: [Cat]/[CHD]/[PO] = 1:20:4000, neat epoxide, 20 bar CO2, 50 °C.3

b

Conversion of epoxide, determined by 1H NMR.

c

Selectivity for carbonate vs ether, determined by 1H NMR.

d

Selectivity for polymer vs cyclic carbonate, determined by 1H NMR.

e

Turnover number (TON) = total number of moles of epoxide consumed/mol of catalyst.

f

Turnover frequency (TOF) = TON/time (hours).

g

Number-average molecular weight [dispersity], determined by GPC.

h

Reaction conditions: [Cat]/[CHD]/[PO] = 1:20:4000, neat epoxide, 30 bar CO2, 70 °C.

i

Reaction conditions: [Cat]/[KI]/[PO] = 1:1:2000, 15 bar CO2, 25 °C.47

The Co(III)K(I) catalyst also showed excellent tolerance to copolymerizations conducted using CTA: high activity was maintained even when 250 equiv of diol (CHD) were added (vs catalyst). The use of diols controls the poly(propene carbonate) molar mass and selects for hydroxyl telechelic end-groups; at high loadings polyols are produced which may be useful for subsequent polyurethane manufacture. Co(III)K(I) can also be used with other epoxides, including cyclic (CHO, vinyl-CHO, and cyclopentene oxide) and acyclic (vinyl-PO, allyl glycidyl ether, and tert-butyl glycidyl ether) epoxides to form a range of different polycarbonates.

Kinetic analysis of the Co(III)K(I) catalyst shows a first-order dependence on catalyst and PO concentrations and a zero-order in carbon dioxide pressure (5–30 bar), indicative that epoxide ring-opening from a nucleophilic carbonate is the rate-determining step. The use of Eyring analysis gives a ΔH of +13.4 kcal mol–1 and ΔS of −26.9 cal mol–1 K–1, which results in a ΔG323 K of +22.1 kcal mol–1. We are currently undertaking detailed computation and kinetic experiments in order to better understand the polymerization mechanism—such insights are critical to accelerating discovery of other heterodinuclear catalysts using these ancillary ligands. The catalysts will also be tested in various other polymerization catalyses including heterocycle ring-opening polymerization; heterocycle/heteroallene ROCOP; and switchable catalyses, which bridge both mechanisms.

Epoxide/Anhydride ROCOP with L3

The success of heterodinuclear catalysis in yielding effective systems for epoxide/CO2 ROCOP naturally led to investigation into the related epoxide/anhydride ROCOP. Polyesters are important sustainable polymers since many monomers are, or could in future be, biobased; materials are recyclable and, after use, can be (bio)degraded to small molecules/monomers.48 Further, with the right catalysts, epoxide/anhydride ROCOP can be well controlled to produce semiaromatic, rigid, functionalized, and/or aliphatic polyesters; catalyst development is a priority. The two ROCOP reactions are similar but not identical, thereby meaning that catalysts must be considered and optimized for polyester production. For example, Al(III) catalysts typically underperform in CO2/epoxide ROCOP but many are excellent for anhydride/epoxide ROCOP.4951 The difference is proposed to arise from a low barrier to Al-alkoxide bond dissociation that normally results in cyclic carbonate formation when carbon dioxide is applied, but since this side reaction is not prevalent when anhydrides are used, facilitates epoxide coordination and activation.6,49

In 2015, we demonstrated that L1Zn(II)Mg(II) was an effective catalyst for CHO/PA ROCOP, with the complex showing a TOF of 188 h–1 (1 mol % catalyst vs PA, 100 °C).36 The heterodinuclear complex outperforms the homodinuclear Zn(II)Zn(II) and Mg(II)Mg(II) congeners—the same heterodinuclear synergy as was observed for epoxide/CO2 ROCOP (vide supra).52 In the same year, we also reported a homodinuclear Zn(II)Zn(II) complex coordinated by Schiff base ligand L3 that showed a higher activity for CHO/PA ROCOP (TOF = 198 h–1, 1 mol % catalyst vs PA, T = 100 °C) than L1Zn(II)Mg(II), despite being a homodinuclear complex (Figure 5).53 Hence, subsequent investigations using L3 targeted heterodinuclear complexes.54 The heterodinuclear Zn(II)Mg(II) was not isolable because metal scrambling formed the Zn(II)Zn(II) and Mg(II)Mg(II) homodinuclear species, i.e., the heterodinuclear complex was not stable using this ligand.

Figure 5.

Figure 5

Dinuclear catalysts using L3 for PA/CHO ROCOP.

Other heterodinuclear complexes using either Mg(II) or Zn(II) were successfully synthesized and were stable with respect to homodinuclear complexation. None of these heterodinuclear complexes, nor the homodinuclear Mg(II)Mg(II) complex, showed any higher activity than Zn(II)Zn(II). The most effective was Zn(II)Cd(II), which showed a moderate activity (TOF = 67 h–1, 1 mol % catalyst vs anhydride, 100 °C). Besides this lower catalytic activity [in comparison with Zn(II)Zn(II)], complexes containing heavy, expensive, and toxic metals such as Cd(II) are not ideal catalytic targets, and so further catalyst exploration was warranted.

In search for effective metal combinations, M(III)M′(I) heterodinuclear complexes were proposed. We had already investigated mononuclear Al(III) complexes coordinated by L3, which showed good epoxide/anhydride ROCOP activity but required a cocatalyst.55,56 In that work, only the Schiff-base (N,N,O,O)-groups were used to coordinate Al(III); hence, a logical next step was to investigate Al(III)M(I) complexes [M(I) = Na(I), K(I), Rb(I), Cs(I)] utilizing the ether-coordination pocket to coordinate the s-block metal (Figure 6, left).4

Figure 6.

Figure 6

Heterodinuclear Al(III)M(I) catalysts coordinated by L3 for epoxide/anhydride ROCOP (left). The proposed ROCOP mechanism is based on kinetic studies, solid-state structural data, and DFT calculations [theory level: ωb97xD/6-31+G(d,p)-H,C/6-311+G(d)-Al,K,O,N] (right).

The resultant complexes were exceptionally active ROCOP catalysts, with the most effective being Al(III)K(I) and Al(III)Rb(I) with TOFs of 1072 h–1 and 1136 h–1, respectively (Table 4, entries 2 and 3). Al(III)Na(I) and Al(III)Cs(I), although slower than the best complexes, also displayed very high activity for this field of catalysis (Table 4, entries 1 and 4). All the catalysts were quantitative in their selectivity for polyester, forming no polyether linkages. Subsequent study was carried out utilizing the Al(III)K(I) complex because K is significantly lighter and more earth-abundant than Rb. The Al(III)K(I) catalyst showed high activity using other epoxides, including alkylene oxides, like PO, and anhydrides. This broad substrate scope allows it to furnish polyesters with a range of properties. It was also tolerant to very low catalyst loading and maintained equivalent activity at 0.005 mol % loading (vs CHO, i.e., [Cat]/[PA]/[CHO] = 1:4000:20000).

Table 4. Heterodinuclear PA/CHO ROCOP Catalysts Using L3a.

# catalyst time (min) conv.b (%)b polyesterc (%) TONd TOFe (h–1) Mn [Đ]f (g mol–1)
1 Al(III)Na(I) 15 45 >99 180 720 9300 [1.06]
2 Al(III)K(I) 15 67 >99 268 1072 14300 [1.06]
3 Al(III)Rb(I) 15 71 >99 284 1136 14400 [1.06]
4 Al(III)Cs(I) 15 54 >99 216 875 10400 [1.05]
a

Reaction conditions: [Cat]/[PA]/[CHO] = 1:400:2000, neat epoxide, 100 °C.

b

Conversion of anhydride, determined by 1H NMR.

c

Selectivity for polyester vs polyether, determined by 1H NMR.

d

Turnover number (TON) = number of moles of anhydride consumed/number of moles of catalyst.

e

Turnover frequency (TOF) = TON/time (hours).

f

Number-average molecular weight [dispersity], determined by GPC.

Further investigations looked to develop an understanding of the role of the two metals during catalysis (Figure 6, right). Kinetic studies revealed a first-order rate dependence on both catalyst and epoxide concentration and a zero-order dependence on anhydride concentration. Hence, it was concluded that epoxide ring-opening is the rate-determining step. Since, in order to form ester linkages, epoxide ring-opening can only occur from a carboxylate chain end, the solid-state molecular structure of Al(III)K(I), which features two acetate coligands, is a useful model for the catalyst resting state. Analysis of the solid-state molecular structure of Al(III)K(I) revealed a 3.665 Å separation of the two metals, similar to both the highly active L2Co(III)K(I) and L1Mg(II)Co(II) complexes, and further demonstrates that intermetallic separation in the range of 3–4 Å is likely key in enabling intermetallic synergy. Like the L2Co(III)K(I) catalyst for CHO/CO2 ROCOP, it shows a “metalate” structure with both acetate groups coordinated as anions to the Al(III). Subsequent DFT calculations, conducted in collaboration with Buchard at the University of Bath, applied an Al(III)K(I) complex with benzoate coligands, which better modeled the propagating phthalate chain-end present in polymerizations. These calculations suggest that the Al(III) center binds and activates the epoxide, while the K(I) center coordinates the growing polymer chain through dative covalent bonds. This coordination ensures the growing chain is close to the active site and may facilitate monomer insertions (Figure 6, right). We have termed this pathway the “dinuclear metallate mechanism.”

There are some parallels between the proposed dinuclear metalate mechanism and the hypotheses for metal salen/cocatalyst systems.56 In both cases, the catalysts show metalate resting states and the polymer chain must undergo an initially unfavorable dissociation from M(III) to allow for subsequent epoxide coordination and activation. However, while salen-type catalysts require external cocatalysts [typically bis(triphenylphosphine)iminium chloride (PPNCl)] to aid the initial polymer chain dissociation, these new heterodinuclear complexes facilitate this process through the s-block M(I) center. The requirement for an external cocatalyst of metal-salen catalyst systems renders them inactive at moderate/low catalyst loadings, while heterodinuclear catalysts show excellent loading tolerance. Further, PPNX cocatalysts are significantly heavier than group 1 metals (e.g., [PPN]+ = 538.59 g mol–1, K+ = 39.1 g mol–1) and are typically expensive and may be corrosive to steel. Therefore, it is desirable to avoid their use.

Conclusions and Outlook

This Account has described some of our approaches to the discovery, development, and understanding of dinuclear catalysts for the ROCOP of epoxides with CO2 or anhydrides, with a particular focus on heterodinuclear synergic metal combinations. The best heterodinuclear catalysts show field-leading rates, high selectivity, excellent loading tolerance, and are effective using many different monomers (Figure 7, top panel). Nonetheless, the selection of the metal combination is critical, and not all heterodinuclear complexes are synergic. Catalyst design must correctly match the right metals with the most appropriate ancillary ligands; intermetallic separation, metallic electronic communication, donor atom chemistry, and the nature of the coligands are all important parameters underpinning effective heterometallic synergy.

Figure 7.

Figure 7

Leading heterodinuclear catalysts for CHO/CO2, PO/CO2, and CHO/PA ROCOP (top). Catalytic mechanisms for the different catalysts (bottom).

The most effective polycarbonate catalysts utilize Co(II), Co(III), or Fe(III), with partner s-block metals such as Mg(II), Na(I), or K(I). The best polyester catalysts benefit from combining Al(III) with Group 1 metals. The use of earth-abundant, inexpensive, and nontoxic metals in catalysis, particularly Al(III), Fe(III), and Group 1 and 2 metals, is a significant advantage compared with many traditional catalysts, which typically apply transition metals and cocatalysts. In most cases, these earth-abundant metals show little-to-no activity by themselves; the heterodinuclear pairing is essential in yielding the high catalytic activities.

Our research implicates two different pathways by which these dinuclear catalysts may operate: (1) the chain shuttling and 2) the dinuclear metallate mechanisms (Figure 7, bottom panel). In both mechanisms, the metals’ roles are separated into epoxide coordination and activation and the provision of a reactive nucleophile. The Lewis acidic metal center activates the epoxide monomers, which is a crucial part of the rate-determining step. The second metal provides a reactive “nucleophile” and/or may also be important in (dative covalent) coordination of the growing polymer chain close to the metal-coordinated epoxide. It is important to fully investigate the catalyst’s structure, kinetics, and intermediates to understand which mechanism best describes performance. As a rule-of-thumb, where metals of different oxidation state/charge are applied, i.e., M(III)M(I) complexes, a dinuclear metalate reaction appears most likely. For a homodinuclear system, such as Zn(II)Zn(II), a chain shuttling mechanism is proposed. Heterodinuclear M(II)M(II) complexes could follow either mechanism.

Our future investigations of heterodinuclear polymerization catalysts will prioritise the use of inexpensive, abundant (easily extracted), and light elements. It is essential to identify, structurally characterize, and study the reactivity of the catalytic intermediates. We will continue to research these processes using in operando and in situ spectroscopies and a portfolio of analytical techniques. Considering the polymers produced using these catalyses, there has been excellent progress in delivery of low molar mass, hydroxyl end-capped polycarbonate/ester polyols. These oligomeric polyols are useful, and exploration of structure–activity relationships in polyurethane, surfactant, and resin/cross-linking applications are ongoing in our group. It is also possible to use the same catalysts and processes to make high molar mass polycarbonates and polyesters—this field of research is under-investigated and warrants greater future attention. One attraction of these heterodinuclear catalysts is their loading and monomer scope tolerance; thus, we will use them to study polymer viscoelastic, thermal, and mechanical properties.5760 Catalyst development will focus upon monomers that are biorenewable and/or extracted as waste coproducts. Therefore, the integration of these polymerizations with carbon capture technologies is a priority.61,62

Acknowledgments

The EPSRC (EP/S018603/1; EP/R027129/1; EP/V003321/1; EP/V038117/1), Oxford Martin School (Future of Plastics), Research England (iCAST), and the EPSRC OXICFM CDT are acknowledged for research funding. For the purpose of Open Access, the authors have applied a CC BY public copyright license to any Author Accepted Manuscript (AAM) version arising from this submission.

Biographies

Wilfred T. Diment graduated with an M.Chem from the University of Oxford in 2017. Subsequently, he completed his Ph.D. in 2021 under the supervision of Prof. Charlotte Williams at the same university, with his research focusing on the development of catalysts for polyester production. He continued with Prof. Williams as a postdoctoral researcher, with his work focusing on the control of the material properties of ROCOP-derived polyesters.

Wouter Lindeboom completed his M.Chem in 2019 under the supervision of Prof. Charlotte Williams at the University of Oxford. He is currently carrying out a Ph.D. under her supervision and developing heterodinuclear catalysts for PO/CO2 ROCOP.

Francesca Fiorentini graduated with an M.Chem in 2020 from Durham University. She is now studying for a Ph.D. at the University of Oxford, supervised by Prof. Charlotte Williams, and developing heterodinuclear catalysts for PO/CO2 ROCOP.

Arron C. Deacy completed his M.Sci at Imperial College London in 2015 and completed his master’s project under the supervision of Dr. Andrew Ashley focusing on the use of tin cations in the hydrogenation of carbon dioxide. Subsequently, he completed a Ph.D. at the University of Oxford in 2019 under the supervision of Prof. Charlotte Williams, where he investigated heterodinuclear catalysts for the epoxide/CO2 ROCOP. Arron has continued in the Williams group as a postdoctoral researcher looking at the chemical recycling of CO2-based polycarbonates.

Charlotte K. Williams is a Professor of Chemistry at the University of Oxford. Her research interests include the polymerization catalyses described in this review, as well as other carbon dioxide utilizations, explorations of sustainable polymer chemistry, metal/metal oxide nanoparticles, and the applications of polymers in energy storage. She is a Fellow of the Royal Society (2021) and was recognized with an Order of the British Empire (OBE) in 2020 for Services to Chemistry. Prior to moving to Oxford, she was an academic at Imperial College London from 2003–2016; she conducted postdoctoral research at the University of Cambridge (Andrew Holmes, Richard Friend) and the University of Minnesota (Bill Tolman, Marc Hillmyer). She was awarded her B.Sc. (1998) and Ph.D. (2001) in Chemistry from Imperial College London.

The authors declare the following competing financial interest(s): C.K.W. is a director of Econic Technologies.

Special Issue

Published as part of the Accounts of Chemical Research special issue “Sustainable Polymers”.

References

  1. Kember M. R.; Knight P. D.; Reung P. T.; Williams C. K. Highly active dizinc catalyst for the copolymerization of carbon dioxide and cyclohexene oxide at one atmosphere pressure. Angew. Chem. Int. Ed 2009, 48 (5), 931–933. 10.1002/anie.200803896. [DOI] [PubMed] [Google Scholar]
  2. Deacy A. C.; Kilpatrick A. F. R.; Regoutz A.; Williams C. K. Understanding metal synergy in heterodinuclear catalysts for the copolymerization of CO2 and epoxides. Nat. Chem. 2020, 12 (4), 372–380. 10.1038/s41557-020-0450-3. [DOI] [PubMed] [Google Scholar]
  3. Deacy A. C.; Moreby E.; Phanopoulos A.; Williams C. K. Co(III)/Alkali-Metal(I) Heterodinuclear Catalysts for the Ring-Opening Copolymerization of CO2 and Propylene Oxide. J. Am. Chem. Soc. 2020, 142 (45), 19150–19160. 10.1021/jacs.0c07980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Diment W. T.; Gregory G. L.; Kerr R. W. F.; Phanopoulos A.; Buchard A.; Williams C. K. Catalytic Synergy Using Al(III) and Group 1 Metals to Accelerate Epoxide and Anhydride Ring-Opening Copolymerizations. ACS Catal. 2021, 11, 12532–12542. 10.1021/acscatal.1c04020. [DOI] [Google Scholar]
  5. Hepburn C.; Adlen E.; Beddington J.; Carter E. A.; Fuss S.; Mac Dowell N.; Minx J. C.; Smith P.; Williams C. K. The technological and economic prospects for CO2 utilization and removal. Nature 2019, 575 (7781), 87–97. 10.1038/s41586-019-1681-6. [DOI] [PubMed] [Google Scholar]
  6. Paul S.; Zhu Y. Q.; Romain C.; Brooks R.; Saini P. K.; Williams C. K. Ring-opening copolymerization (ROCOP): synthesis and properties of polyesters and polycarbonates. Chem. Comm 2015, 51 (30), 6459–6479. 10.1039/C4CC10113H. [DOI] [PubMed] [Google Scholar]
  7. Fischer R. F. Polyesters from Epoxides and Anhydrides. J. Polym. Sci. 1960, 44, 155–172. 10.1002/pol.1960.1204414314. [DOI] [Google Scholar]
  8. Inoue S.; Koinuma H.; Tsuruta T. Copolymerization of Carbon Dioxide and Epoxide. J. Polym. Sci. B Polym. Lett. 1969, 7 (4), 287–292. 10.1002/pol.1969.110070408. [DOI] [Google Scholar]
  9. Liang X.; Tan F.; Zhu Y. Recent Developments in Ring-Opening Copolymerization of Epoxides With CO2 and Cyclic Anhydrides for Biomedical Applications. Front. Chem. 2021, 9, 647245. 10.3389/fchem.2021.647245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. 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 (24), 15167–15197. 10.1021/acs.chemrev.6b00553. [DOI] [PubMed] [Google Scholar]
  11. Wang Y.; Darensbourg D. J. Carbon dioxide-based functional polycarbonates: Metal catalyzed copolymerization of CO2 and epoxides. Coord. Chem. Rev. 2018, 372, 85–100. 10.1016/j.ccr.2018.06.004. [DOI] [Google Scholar]
  12. Kozak C. M.; Ambrose K.; Anderson T. S. Copolymerization of carbon dioxide and epoxides by metal coordination complexes. Coord. Chem. Rev. 2018, 376, 565–587. 10.1016/j.ccr.2018.08.019. [DOI] [Google Scholar]
  13. Andrea K. A.; Plommer H.; Kerton F. M. Ring-opening polymerizations and copolymerizations of epoxides using aluminum- and boron-centered catalysts. Eur. Polym. J. 2019, 120, 109202. 10.1016/j.eurpolymj.2019.08.029. [DOI] [Google Scholar]
  14. Inoue S. Immortal polymerization: The outset, development, and application. J. Polym. Sci. Part A: Polym. Chem. 2000, 38 (16), 2861–2871. . [DOI] [Google Scholar]
  15. Darensbourg D. J. Chain transfer agents utilized in epoxide and CO2 copolymerization processes. Green Chem. 2019, 21 (9), 2214–2223. 10.1039/C9GC00620F. [DOI] [Google Scholar]
  16. Scharfenberg M.; Hilf J.; Frey H. Functional Polycarbonates from Carbon Dioxide and Tailored Epoxide Monomers: Degradable Materials and Their Application Potential. Adv. Funct. Mater. 2018, 28 (10), 1704302. 10.1002/adfm.201704302. [DOI] [Google Scholar]
  17. Tahirov T. H.; Oki H.; Tsukihara T.; Ogasahara K.; Yutani K.; Ogata K.; Izu Y.; Tsunasawa S.; Kato I. Crystal structure of methionine aminopeptidase from hyperthermophile, Pyrococcus furiosus. J. Mol. Biol. 1998, 284 (1), 101–124. 10.1006/jmbi.1998.2146. [DOI] [PubMed] [Google Scholar]
  18. Balasubramanian A.; Durairajpandian V.; Elumalai S.; Mathivanan N.; Munirajan A. K.; Ponnuraj K. Structural and functional studies on urease from pigeon pea (Cajanus cajan). Int. J. Bio. Macromol. 2013, 58, 301–309. 10.1016/j.ijbiomac.2013.04.055. [DOI] [PubMed] [Google Scholar]
  19. Rokicki A.; Kuran W. Copolymerization of Carbon Dioxide with Propylene Oxide in the Presence of Catalysts based on Alkylmetal Compounds and Pyrogallol. Makromol. Chem. 1979, 180 (9), 2153–2161. 10.1002/macp.1979.021800910. [DOI] [Google Scholar]
  20. Darensbourg D. J.; Wildeson J. R.; Yarbrough J. C.; Reibenspies J. H. Bis 2,6-difluorophenoxide Dimeric Complexes of Zinc and Cadmium and Their Phosphine Adducts: Lessons Learned Relative to Carbon Dioxide/Cyclohexene Oxide Alternating Copolymerization Processes Catalyzed by Zinc Phenoxides. J. Am. Chem. Soc. 2000, 122, 12487–12496. 10.1021/ja002855h. [DOI] [Google Scholar]
  21. Cheng M.; Moore D. R.; Reczek J. J.; Chamberlain B. M.; Lobkovsky E. B.; Coates G. W. Single-Site beta-Diiminate Zinc Catalysts for the Alternating Copolymerization of CO2 and Epoxides: Catalyst Synthesis and Unprecedented Polymerization Activity. J. Am. Chem. Soc. 2001, 123, 8738–8749. 10.1021/ja003850n. [DOI] [PubMed] [Google Scholar]
  22. Moore D. R.; Cheng M.; Lobkovsky E. B.; Coates G. W. Mechanism of the Alternating Copolymerization of Epoxides and CO2 Using β-Diiminate Zinc Catalysts: Evidence for a Bimetallic Epoxide Enchainment. J. Am. Chem. Soc. 2003, 125, 11911–11924. 10.1021/ja030085e. [DOI] [PubMed] [Google Scholar]
  23. Nozaki K.; Nakano K.; Hiyama T. Optically Active Polycarbonates: Asymmetric Alternating Copolymerization of Cyclohexene Oxide and Carbon Dioxide. J. Am. Chem. Soc. 1999, 121, 11008–11009. 10.1021/ja992433b. [DOI] [Google Scholar]
  24. Lee B. Y.; Kwon H. Y.; Lee S. Y.; Na S. J.; Han S.; Yun H.; Lee H.; Park Y.-W. Bimetallic Anilido-Aldimine Zinc Complexes for Epoxide/CO2 Copolymerization. J. Am. Chem. Soc. 2005, 127 (9), 3031–3037. 10.1021/ja0435135. [DOI] [PubMed] [Google Scholar]
  25. Koning C; Wildeson J; Parton R; Plum B; Steeman P; Darensbourg D.J Synthesis and physical characterization of poly(cyclohexane carbonate), synthesized from CO2 and cyclohexene oxide. Polymer 2001, 42 (9), 3995–4004. 10.1016/S0032-3861(00)00709-6. [DOI] [Google Scholar]
  26. Kember M. R.; White A. J. P.; Williams C. K. Highly Active Di- and Trimetallic Cobalt Catalysts for the Copolymerization of CHO and CO2 at Atmospheric Pressure. Macromolecules 2010, 43 (5), 2291–2298. 10.1021/ma902582m. [DOI] [Google Scholar]
  27. Buchard A.; Kember M. R.; Sandeman K. G.; Williams C. K. A bimetallic iron(III) catalyst for CO2/epoxide coupling. Chem. Commun. 2011, 47 (1), 212–214. 10.1039/C0CC02205E. [DOI] [PubMed] [Google Scholar]
  28. Kember M. R.; Jutz F.; Buchard A.; White A. J. P.; Williams C. K. Di-cobalt(ii) catalysts for the copolymerisation of CO2 and cyclohexene oxide: support for a dinuclear mechanism?. Chem. Sci. 2012, 3 (4), 1245–1255. 10.1039/c2sc00802e. [DOI] [Google Scholar]
  29. Deacy A. C.; Durr C. B.; Garden J. A.; White A. J. P.; Williams C. K. Groups 1, 2 and Zn(II) Heterodinuclear Catalysts for Epoxide/CO2 Ring-Opening Copolymerization. Inorg. Chem. 2018, 57 (24), 15575–15583. 10.1021/acs.inorgchem.8b02923. [DOI] [PubMed] [Google Scholar]
  30. Deacy A. C.; Durr C. B.; Williams C. K. Heterodinuclear complexes featuring Zn(ii) and M = Al(iii), Ga(iii) or In(iii) for cyclohexene oxide and CO2 copolymerisation. Dalton Trans 2020, 49 (1), 223–231. 10.1039/C9DT02918D. [DOI] [PubMed] [Google Scholar]
  31. Kember M. R.; Williams C. K. Efficient magnesium catalysts for the copolymerization of epoxides and CO2; using water to synthesize polycarbonate polyols. J. Am. Chem. Soc. 2012, 134 (38), 15676–15679. 10.1021/ja307096m. [DOI] [PubMed] [Google Scholar]
  32. Chapman A. M.; Keyworth C.; Kember M. R.; Lennox A. J. J.; Williams C. K. Adding Value to Power Station Captured CO2: Tolerant Zn and Mg Homogeneous Catalysts for Polycarbonate Polyol Production. ACS Catal. 2015, 5 (3), 1581–1588. 10.1021/cs501798s. [DOI] [Google Scholar]
  33. Jutz F.; Buchard A.; Kember M. R.; Fredriksen S. B.; Williams C. K. Mechanistic investigation and reaction kinetics of the low-pressure copolymerization of cyclohexene oxide and carbon dioxide catalyzed by a dizinc complex. J. Am. Chem. Soc. 2011, 133 (43), 17395–17405. 10.1021/ja206352x. [DOI] [PubMed] [Google Scholar]
  34. Buchard A.; Jutz F.; Kember M. R.; White A. J. P.; Rzepa H. S.; Williams C. K. Experimental and Computational Investigation of the Mechanism of Carbon Dioxide/Cyclohexene Oxide Copolymerization Using a Dizinc Catalyst. Macromolecules 2012, 45 (17), 6781–6795. 10.1021/ma300803b. [DOI] [Google Scholar]
  35. Saini P. K.; Romain C.; Williams C. K. Dinuclear metal catalysts: Improved performance of heterodinuclear mixed catalysts for CO2-epoxide copolymerization. Chem. Commun. 2014, 50, 4164–4167. 10.1039/C3CC49158G. [DOI] [PubMed] [Google Scholar]
  36. Garden J. A.; Saini P. K.; Williams C. K. Greater than the Sum of Its Parts: A Heterodinuclear Polymerization Catalyst. J. Am. Chem. Soc. 2015, 137 (48), 15078–15081. 10.1021/jacs.5b09913. [DOI] [PubMed] [Google Scholar]
  37. Trott G.; Garden J. A.; Williams C. K. Heterodinuclear zinc and magnesium catalysts for epoxide/CO2 ring opening copolymerizations. Chem. Sci. 2019, 10 (17), 4618–4627. 10.1039/C9SC00385A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Reis N. V.; Deacy A. C.; Rosetto G.; Durr C. B.; Williams C. K. Heterodinuclear Mg(II)M(II) (M = Cr, Mn, Fe, Co, Ni, Cu and Zn) Complexes for the Ring Opening Copolymerization of Carbon Dioxide/Epoxide and Anhydride/Epoxide. Chem. Eur. J. 2022, 28 (14), e202104198 10.1002/chem.202104198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Darensbourg D. J.; Frantz E. B. X-Ray crystal structures of five-coordinate (salen)MnN3 derivatives and their binding abilities towards epoxides: chemistry relevant to the epoxide–CO2copolymerization process. Dalton Trans 2008, 37, 5031–5036. 10.1039/b806364h. [DOI] [PubMed] [Google Scholar]
  40. Martinez A.; Hemmert C.; Gornitzka H.; Meunier B. Synthesis and activity of macrocyclized chiral Mn(III)-Schiff-base epoxidation catalysts. J. Organomet. Chem. 2005, 690, 2163–2171. 10.1016/j.jorganchem.2005.01.055. [DOI] [Google Scholar]
  41. Leonard N. G.; Chantarojsiri T.; Ziller J. W.; Yang J. Y. Cationic Effects on the Net Hydrogen Atom Bond Dissociation Free Energy of High-Valent Manganese Imido Complexes. J. Am. Chem. Soc. 2022, 144 (4), 1503–1508. 10.1021/jacs.1c09583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Chantarojsiri T.; Ziller J. W.; Yang J. Y. Incorporation of redox-inactive cations promotes iron catalyzed aerobic C–H oxidation at mild potentials. Chem. Sci. 2018, 9 (9), 2567–2574. 10.1039/C7SC04486K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Du K.; Thorarinsdottir A. E.; Harris T. D. Selective Binding and Quantitation of Calcium with a Cobalt-Based Magnetic Resonance Probe. J. Am. Chem. Soc. 2019, 141 (17), 7163–7172. 10.1021/jacs.9b02661. [DOI] [PubMed] [Google Scholar]
  44. Lindeboom W.; Fraser D. A. X.; Durr C. B.; Williams C. K. Heterodinuclear Zn(II), Mg(II) or Co(III) with Na(I) Catalysts for Carbon Dioxide and Cyclohexene Oxide Ring Opening Copolymerizations. Chemistry 2021, 27, 12224. 10.1002/chem.202101140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Bernhard M.; Anton J.; Schmidt F.; Sandkaulen F.; Pascaly M. Über den Technologiewandel in der Propylenoxid-Herstellung. Chem. unserer Zeit 2017, 51 (3), 198–209. 10.1002/ciuz.201700764. [DOI] [Google Scholar]
  46. Deacy A. C.; Moreby E.; Phanopoulos A.; Williams C. K. Co(III)/Alkali-Metal(I) Heterodinuclear Catalysts for the Ring-Opening Copolymerization of CO2and Propylene Oxide. J. Am. Chem. Soc. 2020, 142, 19150. 10.1021/jacs.0c07980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Lu X.-B.; Shi L.; Wang Y.-M.; Zhang R.; Zhang Y.-J.; Peng X.-J.; Zhang Z.-C.; Li B. Design of Highly Active Binary Catalyst Systems for CO2/Epoxide Copolymerization: Polymer Selectivity, Enantioselectivity, and Stereochemistry Control. J. Am. Chem. Soc. 2006, 128, 1664–1674. 10.1021/ja056383o. [DOI] [PubMed] [Google Scholar]
  48. Zhu Y. Q.; Romain C.; Williams C. K. Sustainable polymers from renewable resources. Nature 2016, 540 (7633), 354–362. 10.1038/nature21001. [DOI] [PubMed] [Google Scholar]
  49. Luinstra G. A.; Haas G. R.; Molnar F.; Bernhart V.; Eberhardt R.; Rieger B. On the formation of aliphatic polycarbonates from epoxides with chromium(III) and aluminum(III) metal-salen complexes. Chem. Eur. J. 2005, 11 (21), 6298–6314. 10.1002/chem.200500356. [DOI] [PubMed] [Google Scholar]
  50. Abel B. A.; Lidston C. A. L.; Coates G. W. Mechanism-Inspired Design of Bifunctional Catalysts for the Alternating Ring-Opening Copolymerization of Epoxides and Cyclic Anhydrides. J. Am. Chem. Soc. 2019, 141 (32), 12760–12769. 10.1021/jacs.9b05570. [DOI] [PubMed] [Google Scholar]
  51. Li J.; Liu Y.; Ren W.-M.; Lu X.-B. Asymmetric Alternating Copolymerization of Meso-epoxides and Cyclic Anhydrides: Efficient Access to Enantiopure Polyesters. J. Am. Chem. Soc. 2016, 138 (36), 11493–11496. 10.1021/jacs.6b07520. [DOI] [PubMed] [Google Scholar]
  52. Saini P. K.; Romain C.; Zhu Y. Q.; Williams C. K. Di-magnesium and zinc catalysts for the copolymerization of phthalic anhydride and cyclohexene oxide. Polym. Chem. 2014, 5 (20), 6068–6075. 10.1039/C4PY00748D. [DOI] [Google Scholar]
  53. 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 (24), 11906–11915. 10.1021/acs.inorgchem.5b02233. [DOI] [PubMed] [Google Scholar]
  54. Deacy A. C.; Durr C. B.; Kerr R. W. F.; Williams C. K. Heterodinuclear catalysts Zn(ii)/M and Mg(ii)/M, where M = Na(i), Ca(ii) or Cd(ii), for phthalic anhydride/cyclohexene oxide ring opening copolymerisation. Catal. Sci. Technol. 2021, 11 (9), 3109–3118. 10.1039/D1CY00238D. [DOI] [Google Scholar]
  55. Diment W. T.; Stosser T.; Kerr R. W. F.; Phanopoulos A.; Durr C. B.; Williams C. K. Ortho-vanillin derived Al(iii) and Co(iii) catalyst systems for switchable catalysis using epsilon-decalactone, phthalic anhydride and cyclohexene oxide. Catal. Sci. Technol. 2021, 11 (5), 1737–1745. 10.1039/D0CY02164D. [DOI] [Google Scholar]
  56. Fieser M. E.; Sanford M. J.; Mitchell L. A.; Dunbar C. R.; Mandal M.; Van Zee N. J.; Urness D. M.; Cramer C. J.; Coates G. W.; Tolman W. B. Mechanistic Insights into the Alternating Copolymerization of Epoxides and Cyclic Anhydrides Using a (Salph)AICI and Iminium Salt Catalytic System. J. Am. Chem. Soc. 2017, 139 (42), 15222–15231. 10.1021/jacs.7b09079. [DOI] [PubMed] [Google Scholar]
  57. Rosetto G.; Deacy A. C.; Williams C. K. Mg(ii) heterodinuclear catalysts delivering carbon dioxide derived multi-block polymers. Chem. Sci. 2021, 12, 12315–12325. 10.1039/D1SC03856G. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Sulley G. S.; Gregory G. L.; Chen T. T. D.; Carrodeguas L. P.; 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-epsilon-decalactone-b-cyclohexene carbonate) Adhesives, Elastomers, and Toughened Plastics. J. Am. Chem. Soc. 2020, 142 (9), 4367–4378. 10.1021/jacs.9b13106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Gregory G. L.; Sulley G. S.; Carrodeguas L. P.; Chen T. T. D.; Santmarti A.; Terrill N. J.; Lee K. Y.; Williams C. K. Triblock polyester thermoplastic elastomers with semi-aromatic polymer end blocks by ring-opening copolymerization. Chem. Sci. 2020, 11 (25), 6567–6581. 10.1039/D0SC00463D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Yi N.; Chen T. T. D.; Unruangsri J.; Zhu Y.; Williams C. K. Orthogonal functionalization of alternating polyesters: selective patterning of (AB) n sequences. Chem. Sci. 2019, 10 (43), 9974–9980. 10.1039/C9SC03756J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Chen T. T. D.; Carrodeguas L. P.; Sulley G. S.; Gregory G. L.; Williams C. K. Bio-based and Degradable Block Polyester Pressure-Sensitive Adhesives. Angew. Chem. Int. Ed 2020, 59 (52), 23450–23455. 10.1002/anie.202006807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Peña Carrodeguas L.; Martín C.; Kleij A. W. Semiaromatic Polyesters Derived from Renewable Terpene Oxides with High Glass Transitions. Macromolecules 2017, 50 (14), 5337–5345. 10.1021/acs.macromol.7b00862. [DOI] [Google Scholar]

Articles from Accounts of Chemical Research are provided here courtesy of American Chemical Society

RESOURCES