Sophie M.
Guillaume
* and
Jean-François
Carpentier
Organometallics: Materials and Catalysis, Institut des Sciences Chimiques de Rennes, CNRS – Université de Rennes 1 (UMR 6226), Campus de Beaulieu, 35042 Rennes Cedex, France. E-mail: sophie.guillaume@univ-rennes1.fr; Fax: +33 2 2323 6939; Tel: +33 2 2323 5880
First published on 29th February 2012
The ring-opening polymerization of cyclic carbonate monomers derived from biomass feedstocks constitutes an attractive way to prepare polycarbonates that are potentially useful as high-tech materials or commodity “bioplastics”. This process can be catalyzed by inherently different systems ranging from simple basic organocatalysts, simple Lewis acidic metallic salts such as triflates, or more sophisticated discrete metallo-organic complexes derived from oxophilic metals (zinc, yttrium). In this perspective, the most recent achievements in this chemistry are reviewed. The quite different performances within reach of those catalytic systems, which are assessed in terms of intrinsic reactivity, robustness and regioselectivity towards dissymmetric monomers, are highlighted.
Polyesters/polycarbonates are typically prepared from the condensation of a diol and a diacid/phosgene or dialkyl carbonate, respectively. Although industrially used and suitable with a large choice of monomer feedstocks, this step-growth polymerization suffers from major drawbacks: high temperatures are required and the leaving group/side-product (generally water) constantly has to be removed; also, long reaction times are required for the reaction to be completed. Such operating conditions unfortunately favor undesirable side-reactions, which in turn result in the lack of chain length control, no tuning of chain-end functions and limited access to copolymers.2 In comparison, the ring-opening polymerization (ROP) of cyclic esters/carbonates, most usually driven by the relief of the ring strain,3 is a much better controlled chain-growth polymerization process allowing the synthesis of well-defined and finely tuned (co)polymers. ROP investigations are being extensively developed academically and are now gaining significant industrial considerations, in particular with poly(L-lactide) (PLLA) which occupies a prominent place with a worldwide production of ca. 150000 metric tons p.a.4 In contrast to polycondensations, ROP reactions require milder operating parameters (lower temperature and/or shorter reaction time) to reach high molar mass polymers displaying controlled molecular characteristics (well-defined and tuned microstructure, regio/stereo-selectivity, end-group fidelity, predictable
n and narrow
w/
n values) with limited side-reactions. Tailored (co)polymers are thus accessible with various topologies: linear, star, branched, block, random, etc.2 Cationic, anionic and pseudo-anionic “coordination–insertion” ROP processes have been developed, with the latter catalytic route providing better controlled and “living” polymerizations. The catalytic systems required for ROP are based on either organometallic, inorganic or organic derivatives. The compliance requirements of the catalytic system ideally include cheap, easily accessible, robust and non-toxic component(s) (metal, ligands and co-catalyst), high activity, productivity and selectivity, and minimized promotion of undesirable side-reactions like transesterification or epimerization (in the case of chiral monomers/polymers). Of upmost interest for a variety of reasons (high productivity, low residues), the optimized catalytic system should be fully effective when used in minimized loadings. In this regard, we have recently revisited the “immortal” ring-opening polymerization (iROP) pioneered by Inoue,5 to extend and generalize it to the truly catalytic polymerization of cyclic esters such as lactides, β-lactones and carbonates.6
Such an “immortal” ROP process is clearly differentiated from a “classical” ROP (Scheme 1). In a “classical” ROP mediated by an organometallic catalyst, the number of growing polymer chains is given by the number of initiating functions, that is at most the number of nucleophilic groups (typically alkoxide moieties) attached to the metal center; i.e., it cannot exceed the largest oxidation state of the metal, namely four for the elements usually involved in such a coordination–insertion approach. Thus, in a “classical” ROP, the molar mass of the polymer is, in principle (i.e., for a “living” polymerization), anticipated from the monomer-to-initiator ratio. This therefore significantly limits the productivity of this route. Another issue associated with this approach is that impurities eventually present in the reaction medium (originating from the monomer, solvent, etc.) may irreversibly deactivate part (or all) of the initiating alkoxide moieties, which will therefore affect the final molar mass of the polymer. The other major weakness is the possible traces of metallic/ligand residues recovered in the final polymeric material, which could be a significant issue for a variety of applications (especially microelectronics, biomedical, food packaging).
![]() | ||
Scheme 1 Schematic illustration of “classical” ROP vs. “immortal” ROP. |
In comparison, “immortal” ROP makes use of a true catalyst (the metal or organic species) and an initiator acting also as a chain transfer agent (CTA). The latter initiator/CTA is a protic source, typically an alcohol or more seldom an amine, introduced in a large excess relative to the catalyst species. In such iROP processes,5 provided initiation is faster than propagation, termination reactions are avoided, and quite importantly, chain transfer reactions between dormant and active macromolecular chains are reversible and fast relative to propagation, the “immortal” polymerization is also “living”. Thus, the polymer molar mass is dictated only by the initial monomer-to-initiator/CTA ratio, independently of the catalytic loading. So, the larger the amount of initiator/CTA introduced, the lower the molar mass of the polymer. Nevertheless, greater loadings in monomer units still allow the preparation of high molar mass (in practical terms, desirable) polymers, thereby improving significantly the productivity relatively to “classical” ROP processes. Therefore, in iROP procedures, the catalyst is introduced in minor loadings already at the early stage of the iROP, i.e. in catalytic amounts vs. both the monomer and the initiator/CTA, and it is present during the propagation also in catalytic amounts vs. the polymer chains as well. As a consequence, catalytic residues are even less likely to be found in the final precipitated polymers, thus allowing to alleviate possible issues revolving around toxicity or contaminations. The major challenge in iROP yet remains the identification of a catalyst that stays stable relative to the large excess of initiator/CTA, while still active.6
Whereas polyesters such as polylactones and especially polylactides have been broadly studied,4 polycarbonates have been less commonly investigated.2 However, within the last decade, they have drawn a significant regain of attention, especially in light of their possible bio-sourced nature (Scheme 2). In particular, trimethylene carbonate (1,3-dioxane-2-one, hereafter abbreviated TMC), the most common cyclic carbonate, can be derived from glycerol, a by-product formed during the production of biodiesel originating from vegetable oils and animals fats and cheaply available in large volumes, and more generally from bio-sourced 1,3-propanediol (Scheme 2).7 Other six- and seven-membered ring cyclic carbonates have also been synthesized from the biomass. Among these, α-methyl-substituted TMC (rac-4-methyl-1,3-dioxan-2-one, α-MeTMC) can be similarly prepared via 1,3-butanediol derived from ethanol,8 whereas 2,2-dimethoxytrimethylene carbonate (2,2-dimethoxypropane-1,3-diol carbonate, TMC(OMe)2, a masked hydroxy-TMC equivalent potentially leading to hydrophilic polycarbonates) originates from the naturally occurring 1,3-dihydroxyacetone (DHA; Scheme 2).9 Likewise, the more recently developed seven-membered ring carbonates, 4-methyl- and 5-methyl-1,3-dioxepan-2-one (α-Me7CC and β-Me7CC, respectively), can be similarly formed upon cyclization of the corresponding α,ω-diols, namely 1,4-pentanediol and 2-methyl-1,4-butanediol, themselves arising from the bio-sourced levulinic and itaconic acids, respectively.10 Actually, glycerol, ethanol, levulinic or itaconic acids all belong to the US Department Of Energy's initial “top 10” or latter “top 10-revisited” value chemical opportunities from carbohydrate-based biorefineries.11 While DHA is not included in these classifications, it can be produced from glycerol.7c Noteworthy, all these chemicals presently under consideration are issued from non-edible biomass, allowing a non-controversial economical valorization.
![]() | ||
Scheme 2 Some bio-sourced six- and seven-membered cyclic carbonates studied in (i)ROP. |
Therefore, when such polymers are being eco-conceived thanks to sustainable polymerization approaches such as bulk (i.e. solvent-free) iROP, they are sometimes referred to as “green” polymers or “bioplastics” that could also provide a safe alternative to the production of commodity polymers presently originating from depleting fossil resources. In this context, our ongoing efforts aim at developing original poly(carbonate)s potentially derived from the biomass.
This contribution is an account of our most recent results in the field of iROP of cyclic carbonates derived from biomass resources, with a variety of catalysts including metallo-organic, metallic and organic examples, highlighting their respective catalytic performance/efficiencies and overall rewards. Possible outlooks of this work are also given.
![]() | ||
Scheme 3 Metal- and organo-catalyzed iROP of cyclic carbonates. |
A first class of catalysts that were investigated in depth are the discrete metallo-organic complexes such as the zinc-{β-diiminate} amido complex [(BDIiPr)Zn(N(SiMe3)2)] ((BDIiPr) = 2-((2,6-diisopropylphenyl)amido)-4-((2,6-diisopropylphenyl)-imino)-2-pentene) or the yttrium-{amino-alkoxy-bis(phenolate)} amido complex [(ONOOtBu)Y(N(SiHMe2)2)(THF)] (Scheme 4).8,10,12,15 The selection of these catalyst systems was based on their prior performances in the ROP of lactides as well as of the more reluctant to polymerize β-butyrolactone, which include high activity and regio- and stereo-selectivity.16,17 In particular, while the yttrium compound proved to be highly stereoselective for the formation of the syndiotactic corresponding poly(hydroxyalkanoate)s, the zinc-based initiator afforded atactic polymers.16
![]() | ||
Scheme 4 Some metal-based and organic catalysts used in iROP of cyclic carbonates. |
On the other hand, robust catalytic systems were highly desirable to better withstand drastic industrial applications with technical-grade monomer batches. Therefore, other classes of catalysts were also examined, including simple metallic salts, notably some metallic triflates, (M(OTf)3, M = Sc, Al, Bi),6,13 and some organocatalysts, essentially of the amine, guanidine or phosphazene type such as 4-N,N-dimethylaminopyridine (DMAP), 1.5.7-triazabicyclo-[4.4.0]dec-5-ene (TBD) or 2-tert-butylimino-2-diethylamino-1,3-dimethylperhydro-1,3,2-diazaphosphorine (BEMP), respectively (Scheme 4).6,12c,14,15 These were picked for their known resistance (definitely better than that of the metallo-organic complexes) towards impurities as well as for their recognized catalytic ability in several organic transformations, including polymerization.18,19
Although the former metallo-organic systems operate through a coordination-insertion mechanism (CIM), the latter metal triflates usually act through an activated-monomer mechanism (AMM);4a,6 for organo-catalyzed ROP, a variety of mechanistic pathways are encountered: nucleophilic/electrophilic activation of the monomer, basic activation of the initiating/propagating alcohol, or bifunctional activation of the monomer and alcohol.19 The significant benefit imparted by the metallo-organic complexes like the above mentioned zinc and yttrium derivatives is their high activity and especially their high selectivity, as evidenced in the iROP of asymmetric six- and seven-membered carbonates (vide infra, Scheme 5). This latter remarkable feature relies on the stereo-electronically favorable surrounding of the bulky β-diiminate or alkoxy-amino-bis(phenolate) ligands that allows a better control of the regioselectivity in the ring-opening of the monomer units, as further demonstrated in the stereoselectivity control with other cyclic monomers, especially with chiral racemic β-lactones and lactides.16
![]() | ||
Scheme 5 Possible regioselectivities for the ROP of a dissymmetric cyclic carbonate illustrated in the case of α-MeTMC. |
For comparison purposes, the bulk iROP of each carbonate monomer with these different catalyst systems has been investigated at a given ratio [carbonate]0/[catalyst]0/[BnOH]0 = 500:
1
:
5. Selected results obtained at a temperature ranging from 20 °C to 150 °C are reported in Table 1. It is noteworthy that these experiments were carried out using rigorously purified monomers, unless otherwise stated.
Carbonate | [Catalyst] | Temp./°C | Reaction Timea/min | Conv.b (%) |
![]() |
![]() |
![]() ![]() |
TOFf/h−1 |
---|---|---|---|---|---|---|---|---|
a Reaction times were not necessarily optimized.
b Monomer conversion determined by 1H NMR.
c Calculated from [carbonate]0/[BnOH]0 × monomer conversion × Mcarbonate + MBnOH, with MTMC = 102 g mol−1, MTMC(OMe)2 = 162 g mol−1, MMeTMC = 116 g mol−1, MMe7CC = 130 g mol−1, MBnOH = 108 g mol−1.
d Number average molar mass determined by SEC vs. polystyrene standards and corrected by a factor of 0.73 for TMC26 or uncorrected for polycarbonates derived from TMC(OMe)2, α-MeTMC and α,β-Me7CC.
e Molar mass distribution calculated from SEC traces.
f Apparent turnover frequency expressed in molcarbonate molcatalyst−1 h−1.21
g Reaction run with unpurified TMC at [TMC]0/[catalyst]0/[BnOH]0=10![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
||||||||
TMC | Al(OTf)3 | 110 | 60 | 96 | 9900 | 13![]() |
1.62 | 480 |
TMC | Al(OTf)3 | 150 | 5 | 98 | 10![]() |
10![]() |
1.55 | 5880 |
TMCg | Al(OTf)3 | 150 | 20 | 92 | 94![]() |
61![]() |
1.42 | 27![]() |
TMCh | Al(OTf)3 | 150 | 15 × 60 | 87 | 88![]() |
41![]() |
1.66 | 5800 |
TMC | (BDIiPr)Zn(N(SiMe3)2) | 60 | 7 | 99 | 10![]() |
12![]() |
1.55 | 4240 |
TMC | (BDIiPr)Zn(N(SiMe3)2) | 110 | 3 | 100 | 10![]() |
11![]() |
1.77 | 10![]() |
TMCi | (BDIiPr)Zn(N(SiMe3)2) | 60 | 180 | 89 | 45![]() |
43![]() |
1.90 | 2967 |
TMCj | (BDIiPr)Zn(N(SiMe3)2) | |||||||
TMC | BEMP | 60 | 60 | 85 | 8780 | 8250 | 1.25 | 425 |
TMC | BEMP | 110 | 5 | 80 | 8270 | 6950 | 1.43 | 4800 |
TMCi | BEMP | 110 | 60 | 60 | 30![]() |
29![]() |
1.40 | 6000 |
TMCi | BEMP | 150 | 30 | 65 | 33![]() |
30![]() |
1.61 | 13![]() |
TMCh | BEMP | 150 | 15 × 60 | 95 | 97![]() |
24![]() |
1.51 | 6330 |
TMC | TBD | 60 | 30 | 99 | 10![]() |
10![]() |
1.85 | 990 |
TMC | TBD | 110 | 5 | 99 | 10![]() |
12![]() |
1.52 | 5940 |
TMCh | TBD | 150 | 15 × 60 | 91 | 92![]() |
25![]() |
1.54 | 6067 |
TMCi | TBD | 110 | 120 | 98 | 50![]() |
44![]() |
1.52 | 4900 |
TMCi | TBD | 150 | 10 | 82 | 41![]() |
19![]() |
1.65 | 49![]() |
TMC | DMAP | 60 | 30 | 50 | 5210 | 5050 | 1.18 | 500 |
TMC | DMAP | 110 | 15 | 97 | 10![]() |
13![]() |
1.53 | 1940 |
TMCi | DMAP | 110 | 120 | 97 | 49![]() |
42![]() |
1.56 | 4850 |
TMCi | DMAP | 150 | 10 | 93 | 47![]() |
30![]() |
1.47 | 55![]() |
TMC(OMe)2 | Al(OTf)3 | 60 | 330 | 28 | 4 650 | — | — | 25 |
TMC(OMe)2 | Al(OTf)3 | 90 | 330 | 96 | 15![]() |
10![]() |
1.18 | 87 |
TMC(OMe)2 | Al(OTf)3 | 110 | 330 | 0 | — | — | — | 0 |
TMC(OMe)2 | (BDIiPr)Zn(N(SiMe3)2) | 60 | 90 | 95 | 15![]() |
14![]() |
1.28 | 320 |
TMC(OMe)2 | (BDIiPr)Zn(N(SiMe3)2) | 90 | 60 | 93 | 15![]() |
17![]() |
1.25 | 465 |
TMC(OMe)2 | BEMP | 90 | 180 | 100 | 16![]() |
14![]() |
1.53 | 167 |
TMC(OMe)2 | TBD | 90 | 180 | 99 | 16![]() |
14![]() |
1.71 | 165 |
TMC(OMe)2 | DMAP | 90 | 330 | 45 | 7400 | 6800 | 1.23 | 41 |
α-MeTMC | Al(OTf)3 | 110 | 150 | 81 | 9500 | 9300 | 1.18 | 162 |
α-MeTMC | Al(OTf)3 | 150 | 15 | 77 | 9100 | 5700 | 1.18 | 1540 |
α-MeTMC | (BDIiPr)Zn(N(SiMe3)2) | 60 | 7 | 94 | 11![]() |
12![]() |
1.28 | 4028 |
α-MeTMC | BEMP | 60 | 180 | 72 | 8450 | 6000 | 1.19 | 120 |
α-MeTMC | TBD | 60 | 10 | 93 | 10![]() |
10![]() |
1.19 | 2790 |
α-MeTMC | TBD | 110 | 5 | 99 | 11![]() |
8500 | 1.55 | 5940 |
α-MeTMC | DMAP | 60 | 180 | 30 | 3600 | 2500 | 1.25 | 50 |
α-Me7CC | Al(OTf)3 | 40 | 85 | 84 | 11![]() |
5950 | 1.43 | 296 |
α-Me7CC | Al(OTf)3 | 110 | 15 | 100 | 13![]() |
4900 | 1.15 | 2000 |
α-Me7CC | (BDIiPr)Zn(N(SiMe3)2) | 60 | 30 | 94 | 12![]() |
8050 | 1.27 | 940 |
α-Me7CCi | (BDIiPr)Zn(N(SiMe3)2) | 20 | 10 | 100 | 26![]() |
8600 | 1.68 | 1200 |
β-Me7CC | Al(OTf)3 | 110 | 15 | 100 | 13![]() |
22![]() |
1.18 | 2000 |
β-Me7CC | (BDIiPr)Zn(N(SiMe3)2) | 20 | 15 | 17 | 2300 | — | — | 340 |
β-Me7CCk | TBD | 110 | 60 | 100 | 13![]() |
10![]() |
1.40 | 100 |
β-Me7CCk | BEM | 110 | 60 | 10 | 13![]() |
7800 | 1.20 | 100 |
β-Me7CCk | DMAP | 110 | 60 | 87 | 11![]() |
8500 | 1.21 | 87 |
The overall control of the molecular features of the polycarbonates produced by iROP, that is (i) the match of the experimental molar mass determined by SEC (nSEC) and/or by NMR (
nNMR) with the one calculated based on the initial [monomer]0/[initiator/CTA]0 ratio and on the monomer conversion (
ntheo), and (ii) the relatively narrow molar mass distribution values (
w/
n),20 was quite good. Some undesirable side-reactions occurred, as expected to be more likely for a bulk process as opposed to a solution one, however to a rather limited extent. In the iROP of TMC, the monomer we most extensively investigated,6,12–14 the range of
w/
n values generally remained within 1.10 to 1.90 for molar mass values ranging up to
nSEC = 185
200 g mol−1.12a,b
Studies on all other carbonates were focused on the identification of viable catalytic systems effective and possibly regioselective (in the case of α-MeTMC and α/β-Me7CC) in their controlled iROP, rather than on the optimization of the catalyst performances in terms of activity and productivity. For TMC(OMe)2, polycarbonates of molar mass as high as nSEC = 70
200 g mol−1 and of
w/
n values in the range 1.12–1.71 were thus prepared.15,20 Regarding the methyl-substituted six- and seven-membered cyclic carbonates α-MeTMC, α-Me7CC and β-Me7CC, our efforts were aimed at establishing some relationship between the substitution site and the regioselectivity of the catalytic system in ring-opening the cyclic monomer at the oxygen–acyl bond close or remote from the Me group position. The corresponding polymers feature molar mass values in a range allowing NMR investigations which were required to assess the regioselectivity, namely
nSEC ≤ 28
300 g mol−1 and 1.12 <
w/
n < 1.55 for α-MeTMC,8,20 and
nSEC ≤ 17
000 g mol−1 and 1.11 <
w/
n < 1.73 for α/β-Me7CC.10,20
In each carbonate monomer, the expected α-hydroxy, ω-alkoxyester polymer chain-ends were always observed, as evidenced by detailed NMR and MALDI-ToF-MS structural analyses of the polycarbonates. Whichever the mechanism involved (coordination–insertion mechanism or activated-monomer mechanism, vide supra),4a,6 formation of such telechelic polymers was expected (Scheme 2). The alkoxyester end-capping group always originates from the alkoxide moiety of the alcohol used as initiator/CTA. On the other hand, as a result of a different mechanistic approach of the ROP which depends on the catalytic system selected, the origin of the hydroxyl end-group differs based on whether a coordination–insertion mechanism, an activated-monomer mechanism or a mechanism involved for an organocatalyst takes place. In the case of a coordination–insertion mechanism, the hydroxyl chain-end is generated while quenching (i.e., during the final deactivation workup) the polymerization upon protonation of the M–O(alkoxide) bond of the active metal–polymeryl propagating species. On the contrary, the terminal hydroxyl end-function is formed earlier in an activated-monomer mechanism or an organo-catalyzed procedure: as soon as the first (activated) carbonate monomer is ring-opened by the exogenous nucleophilic protic initiator, the carbonate segment to be grown next already bears the terminal hydroxy unit. This reactive α-hydroxy group is then maintained throughout propagation and eventually recovered in the final polymer.6
Several other metallic complexes including Mg[N(SiMe3)2]2, Ca[N(SiMe3)2]2(THF)2, Y[N(SiMe3)2]3, [(BDIiPr)Fe(N(SiMe3)2)], Fe[N(SiMe3)2]2, Fe[N(SiMe3)2]3, Zn[N(SiMe3)2]2, and ZnEt2 were also successfully combined with an alcohol such as isopropanol or benzyl alcohol into bi-component catalytic systems for the bulk iROP of TMC at 60–110 °C.12b The magnesium, calcium and yttrium amido derivatives investigated, which are more active than [(BDIiPr)Zn[N(SiMe3)2], remain yet more sensitive and undergo rapid deactivation when the TMC loading is increased, while the iron amido systems are less active. Thus, among all these metallo-organic precursors, the {β-diiminate}–zinc amido complex provides an ideal compromise affording the overall best performances combining both a high activity and a relatively good stability. Under optimized conditions, as much as 50000 equiv. of purified TMC could be fully converted from as little as 20 ppm of this metallic precursor, with as many as 300 growing polycarbonate chains, allowing the preparation of a PTMC of a molar mass as high as 185
200 g mol−1 with a relatively narrow molar mass distribution (
w/
n = 1.68).12a,20 Also, a double monomer feed experiment carried out with this same {β-diiminate}–zinc/BnOH catalytic system further proved the “living” character of the polymerization.12b
The overall efficiency of these robust metal triflates and organocatalyst systems towards technical grade TMC is particularly relevant and significant for industrial applications which do not provide high purity reagents/operating conditions. Remarkably, under our experimental conditions, the growth of as many as 800 polymer chains from the catalyst, introduced at loadings as low as 10 ppm (i.e., 2 mg of BEMP or TBD), was possible by converting quantitatively 74.4 g of technical-grade TMC, which resulted in a PTMC with a molar mass as high as 61200 g mol−1.
With unpurified TMC, some of the impurities (supposedly protic reagents, such as water or residual 1,3-propanediol) present in the monomer in larger amounts with larger loadings were demonstrated to behave as additional initiators/CTAs in a similar manner to the purposely added exogenous alcohol. In that case, the total number of growing polymer chains is then determined by the sum of added alcohol CTA molecules and “transfer active” impurities. This increase in active species in turn, at a given conversion, proportionally decreases the molar mass of PTMCs resulting from such polymerizations. We established that an accurate theoretical molar mass of the polymer () can be determined upon taking into account the amount of these “transfer active” impurities
(
= 0.056% with the TMC batch used).13,14 This approach further allowed us to predict and thereby to tune “at will” the polymer molar mass from a given loading of TMC, based on the [TMC]0/[Al(OTf)3]0/[BnOH]0 ratio and the
content of “active transferring” impurities contained within all reagents (monomer and initiator/CTA).
The decreased polymerization activities observed with TMC(OMe)2 (as compared to TMC) for all these catalysts most likely result from unfavorable interactions of the additional methoxy functionalities with either the Lewis acidic metal (Zn, Al) centers or the organocatalysts, which compete with the effective activation of the carbonate moiety.
The ROP of β-Me7CC using the organocatalysts DMAP, TBD or BEMP in the presence of BnOH ([carbonate]0/[catalyst]0/[BnOH]0 = 100:
1
:
1) was found to be effective only at higher temperatures: whereas no polymer was formed at 20 or 60 °C, the polymerization proceeded smoothly at 110 °C. All three organocatalytic systems then exhibited rather similar activities (TOF ≤ 87–100 h−1).8,10
13C NMR investigations of the microstructure of the polycarbonates revealed particularly informative. The intensity of the three resonances observed in the carbonyl region proved diagnostic of the monomer diad sequences, allowing us to differentiate and to quantify the enchainment of two monomer units, i.e. the regioregularity of cleavage of either of the two O–C(O)O bonds in α-MeTMC, α-Me7CC or β-Me7CC. The intensity of the two smaller upfield and downfield resonances (B and C, arising from regioirregular enchainment of the monomer units) relative to the central one (A and D, arising from regioregular enchainment of the monomer units) enabled to calculate the regioregularity Xreg in the ROP.24 This extent of regioregularity provided by the catalytic system was independently confirmed with the resonances in the methine (OCHCH3) and methylene (OCH2) regions of the 13C{1H} NMR spectra.8,10
Considering poly(α-MeTMC)s, the Xreg values illustrated that the [(BDIiPr)Zn(N(SiMe3)2)]/BnOH system affords the best regioselectivity (Xreg > 0.98), as opposed to the Al(OTf)3/BnOH and TBD/BnOH systems which are less or hardly regioselective (Xreg = ca. 0.62 and 0.35–0.39, respectively).8 The same trend in regioselectivity values was observed in the iROP of α-Me7CC but to somewhat different extents: the [(BDIiPr)Zn(N(SiMe3)2)]/BnOH system achieved the most regioselective (Xreg = 0.71) iROP, while [(ONOOtBu)Y(N(SiHMe2)2)(THF)]/BnOH and especially Al(OTf)3/BnOH were less regioselective (Xreg = ca. 0.58 and 0.27, respectively).10 Apparently, the smaller ring size of α-MeTMC (as compared to α-Me7CC) and the bulkier structure of the zinc (vs. aluminium) catalyst favored a regioselective process. The lack of regioselectivity of TBD in the iROP of α-MeTMC, although inherently different in nature, is in line with the poor stereoselectivity that this and most other organocatalysts have been thus far able to achieve in ROP of chiral cyclic esters.25 Finally, we observed that, in contrast to poly(α-Me7CC), the ROP of β-Me7CC always proceeds in a perfectly statistical fashion (i.e., Xreg = 0), irrespective of the operating conditions and catalyst systems used.10
Overall, these microstructural investigations revealed: (i) the higher regioselectivity in the ROP of α-MeTMC and of α-Me7CC—with preferential ring-opening at the most hindered oxygen-acyl O–C(O)O bond, i.e. closest to the α-Me substituent—of the zinc-based system, most likely as a result of the favorable steric hindrance of this bulky catalyst; (ii) the absence of regioselectivity of the Al(OTf)3 or TBD catalyst systems in the ROP of α-MeTMC and α-Me7CC; and (iii) the absence of regioselectivity in the ROP of β-Me7CC, whichever the catalyst system, most likely as a result of the OC(O)O-further remote substitution site.
The overall order of catalytic efficiency is [(BDIiPr)Zn(N(SiMe3)2)] > Al(OTf)3 ≈ TBD > BEMP ≈ DMAP, whichever the carbonate monomer, but provided a rigorously purified sample is used. While the [(BDIiPr)Zn(N(SiMe3)2)]/BnOH system features high activity and productivity at typically 60–110 °C when using purified TMC loadings, Al(OTf)3 and all organocatalysts remain by far the most effective catalysts with technical grade TMC batches. The robustness and achievements of these metallic triflates and organocatalysts are remarkable and of high relevance for industrial applications. They are most effective at higher temperatures in the range 110–150 °C, more appropriate from industrial considerations. However, they lack selectivity and, as especially pointed out in recent studies, they do not impart regioselectivity in the ring-opening of dissymmetric cyclic carbonates such as α-MeTMC, α-Me7CC and β-Me7CC, as opposed to the more selective but much more sensitive {β-diiminate}–zinc and {alkoxy-amino-bis(phenolate)}-yttrium precursors.
According to the desired achievements and applications, one may thus select one catalyst system over another in order to reach the best compromise between sensitivity, activity and selectivity. Obviously, the design of an ideal catalytic system combining selectivities and high effectiveness still remains a challenge. One first approach might be considered in light of the recently developed, so-called “dual organic/organometallic ring-opening polymerization” which provides a three-component catalytic system: a metallic Lewis acidic center (a discrete cationic {diaminophenolato}–zinc complex) that activates the monomer (lactide), and an organic base (a tertiary amine) that activates the initiating/propagating alcohol.27 It is envisioned that the role of the Lewis acidic catalyst in inducing selectivity in such a dual polymerization [which can even proceed under immortal and living ROP conditions] could be promoted by the surrounding ancillary ligand(s).
This journal is © The Royal Society of Chemistry 2012 |