Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Metal-catalyzed stereoselective ring-opening polymerization of functional β-lactones: methylene-alkoxy-fluorinated polyhydroxyalkanoates unveil the role of non-covalent interactions

Ali Dhaini a, Rama M. Shakaroun a, Ali Alaaeddine b, Jean-François Carpentier *a and Sophie M. Guillaume *a
aUniv. Rennes, CNRS, Institut des Sciences Chimiques de Rennes, UMR 6226, F-35042 Rennes, France. E-mail: jean-francois.carpentier@univ-rennes.fr; sophie.guillaume@univ-rennes.fr
bUniv. Libanaise, Campus Universitaire Rafic Hariri Hadath, Faculté des Sciences, Laboratoire de Chimie Médicinale et des Produits Naturels, Beirut, Lebanon

Received 30th December 2023 , Accepted 27th January 2024

First published on 31st January 2024


Abstract

Non-covalent interactions (NCIs) can play a major role in the stereoselective ring-opening polymerization (ROP) of racemic β-lactones mediated by achiral metal catalysts, towards the formation of the corresponding polyhydroxyalkanoates (PHAs) with a specific microstructure. Our longstanding endeavor to better understand the factors governing the stereoselective ROP of 4-substituted β-propiolactone monomers rac-BPLFGs (resulting in PBPLFGs; FG = (non)functional group), from catalyst systems based on diamino- or amino-alkoxy-bis(ortho,para-substituted)phenolate rare earth complexes associated with an exogenous alcohol as an initiator, indicated that the formation of iso-enriched polyesters requires the use of ortho-halogen-substituted phenolate ligands (halogen = F, Cl, and Br) along with a methylene alkoxide pending substituent (FG = –CH2OCH2R*) on the lactone. Any other ligand/β-lactone combination resulted in either syndiotactic or atactic PBPLFGs. We report herein the controlled ROP of 4-methylene-alkoxy-fluorinated substituted β-propiolactone, rac-BPLCH2OCF2CHF2, catalyzed by diamino-bis(ortho,para-R,R-substituted)phenolate yttrium catalyst systems Y{ONNOR2}/iPrOH (with R = Me, Cl, tBu, cumyl; namely 1a–d/iPrOH, respectively). This monomer does not feature any outer methylene hydrogen in the alkoxide moiety (–CH2O[C with combining low line][F with combining low line]2CHF2vs. –CH2O[C with combining low line][H with combining low line]2R*), thus enabling the evaluation of the role of outer CH2O(R*)[C with combining low line][H with combining low line]2R NCIs in the stereocontrol. The tBu catalyst 1c showed the highest ROP activity (TOF up to 4650 h−1) out of the four catalyst systems investigated. The fluoroalkyl PHAs were recovered with molar mass values up to Mn,NMR = 106[thin space (1/6-em)]000 g mol−1 and narrow dispersity (ĐM = 1.02–1.24). Detailed NMR and mass spectrometry characterization studies supported the formation of α-isopropoxy,ω-hydroxy telechelic PBPLCH2OCF2CHF2 chains. Catalysts 1c and 1d flanked with bulky phenolate substituents (R = tBu, cumyl) and the chloro-substituted one 1b all returned syndio-enriched PBPLCH2OCF2CHF2s with Pr up to 0.87, as assessed by 13C NMR spectroscopy. Only the methyl substituted catalyst 1a gave atactic PHAs. Eventually, these findings support the hypothesis that both inner and outer methylene hydrogens within an alkoxymethylene exocyclic group on the rac-β-lactone monomer ([C with combining low line][H with combining low line]2O[C with combining low line][H with combining low line]2R*) are required, along with an ortho-chloro-substituted yttrium bisphenolate catalyst (Y{ONNOCl2}), to induce NCIs (Cl[H with combining low line]2[C with combining low line]-O-[C with combining low line](R*)[H with combining low line]2Cl), ultimately resulting in unique isotactic synthetic PHAs.


Introduction

The ring-opening polymerization (ROP) of chiral cyclic esters into the corresponding polyesters that exhibit stereoregularity along their backbone is highly desirable. Such polymers exhibiting adjacent stereocenters having the same (isotactic structure), or opposite/alternating (syndiotactic/heterotactic structure) configuration, are indeed valuable for their tunable thermal, mechanical and degradation properties, in comparison with those exhibiting randomly distributed stereocenters (atactic structure).1–4 Such polyesters typically include polylactide (PLA) and polyhydroxyalkanoates (PHAs), which are commonly prepared by ROP of lactide (LA) and β-lactones, respectively. One advantage (among others), over PLA, of the most common synthetic PHAs, which are derived from 4-substituted β-propiolactones (namely BPLFGs, FG = non-functional or functional group), arises from the FG exocyclic side chain. Indeed, the desired functionality can be easily introduced into monomers by chemical modification, ultimately resulting in a large range of functional PBPLFGs with correspondingly diverse and tunable properties. Ideally, such stereoregular (isotactic, syndiotactic or heterotactic) PLAs and PHAs are more attractive if they can be obtained from a racemic mixture of enantiopure monomers, namely from racemic lactide and β-propiolactones (rac-LA and rac-BPLFGs), respectively, provided a stereoselective catalyst system is used.5–8

Through our ongoing interest in promoting the formation of well-defined PLA and PHAs with controlled stereoregularity, we have established that achiral tetradentate diamino- or amino-alkoxy-bis(ortho,para-R,R-substituted)phenolate rare earth metal (M) complexes associated with an exogenous alcohol as a co-initiator (typically isopropanol), namely M{ONXOR2}/iPrOH (with X = OMe or NMe2), effectively promote the stereoselective ROP of rac-LA and rac-BPLFGs.9

Moderately-to-highly heterotactic PLA (Pr = 0.56–0.96; Pr/m = probability of racemo/meso enchainment between adjacent monomer units as determined by 13C NMR, with Pr + Pm = 1; Pr is identical to Psyndio and Pm is identical to Piso; Pm = 1 or Pr = 1 for a perfectly isotactic or syndiotactic polymer, respectively, and Pr = Pm = 0.5 for an atactic one) was prepared from the ROP of rac-LA promoted at room temperature using a rare earth amido complex (M = La, Nd, Y) featuring amino-alkoxy-bisphenolate ligands with various ortho-R and para-R substituents {ONOOR2}2−. The bulkier the substituents on the metal ancillary ligand (R = Me ≪ tBu < adamantyl < CMe2Ph < CMe2tBu < CPh3), the more hetero-enriched the PLAs. This initial study also revealed that the smaller yttrium (ionic radius: Y < Nd < La) afforded significantly higher stereoselectivity over the larger neodymium or lanthanum (Pr = 0.91, 0.65, and 0.60, respectively). Also, the stereoelectronic contribution of the remote phenolate para-R substituents was demonstrated to not significantly impact the control of LA unit enchainment.9 Accordingly, we later focused our efforts on the more easily synthetically accessible bisphenolate complexes flanked with identical ortho,para-R,R substituents, using preferentially yttrium complexes, namely Y{ONXOR2}/iPrOH catalyst systems.

The extended series of 4-substituted β-propiolactones that we have been investigating using these latter catalyst systems aimed at covering a wide range of exocyclic substituents, in order to probe the contribution of stereoelectronically diverse lactone side-chains to the stereocontrol of the ROP. Thus, alkyl (FG = Me), (di)methylene (thio)alkoxide (FG = CH2OMe, CH2OiPr, CH2OtBu, CH2OAll, CH2OBn, CH2OSitBuMe2, CH2OPh, CH2SPh, and CH2CH2OBn), and ester (FG = CO2Me, CO2All, and CO2Bn) substituted β-propiolactones enabled the estimation of the influence of short/long, polar/non-polar, alkyl/functional lactone pending groups on their ROP mediated by the Y{ONXOR2}/iPrOH catalyst platform (Scheme 1).7–9,10–17 Extended insights into the stereoselectivity of the resulting PHAs by 13C NMR analyses first suggested that, by itself, neither the length, steric bulkiness, electronic environment nor chemical functionality (alkyl vs. ether vs. ester) provided by the FG exocyclic substituent, may enable the preparation of PHAs other than atactic or syndiotactic PHAs. While the chemical synthesis of syndiotactic PHAs − that arise from a regular “chain-end stereocontrol mechanism” thanks to the sterically bulky substituents installed on the ligand of the catalyst,5,7–17 and that otherwise cannot be found as natural polymers since the PHAs produced naturally or biosynthetically from microorganisms or enzymes are invariably isotactic polymers − was thereby rewardingly established, isotactic polymers were only exceptionally obtained from these Y{ONXOR2}/iPrOH catalyst systems.18 Indeed, only the 4-alkoxymethylene-substituted β-propiolactones, simultaneously featuring inner and outer methylene groups apart from the side-chain oxygen atom, namely rac-BPL[C with combining low line][H with combining low line][2 with combining low line]O[C with combining low line][H with combining low line][2 with combining low line]R* (R* = H, CH[double bond, length as m-dash]CH2, Ph), returned the isotactic corresponding PBPL[C with combining low line][H with combining low line][2 with combining low line]O[C with combining low line][H with combining low line][2 with combining low line]R*s, provided the yttrium bisphenolate ligand was bearing ortho-halogen substituents (R = F, Cl, Br).11,15 DFT insights corroborated these experimental observations upon evidencing the formation of “second-sphere” interactions, referred to as non-covalent interactions (NCIs), between the phenolate ortho-halogen and the acidic hydrogens from both the inner and outer methylenes on the last inserted monomer unit (Cl[H with combining low line]2[C with combining low line]-O-[C with combining low line](R*)[H with combining low line]2Cl), i.e., on the repeating unit closest to the metal center and its surrounding ligand. Of note, these computational results emphasized the negligible role of the phenolate para-substituents that are unlikely to be involved in NCIs with the monomer exocyclic side chain, as they are sterically too distant from this active site. Furthermore, this in silico information suggested that, besides, yet in complement of, the stereoelectronic environment in close proximity of the active center, such as (ligand)Cl[H with combining low line]2[C with combining low line]-O NCIs that are attractive in nature, is essential to drive the catalyst isoselectivity.18d,19–24


image file: d3py01430d-s1.tif
Scheme 1 Stereoselective ROP of various chiral racemic 4-substituted-β-propiolactones, rac-BPLFGs (FG = Me, CH2CH2OCH2Ph, CH2OCH2R* with R* = Me, nBu, allyl, Bn, CH2OZ with Z = iPr, tBu, Ph, SitBuMe2, CO2R* with R* = Me, Allyl, Bn) performed with Y{ONXOCl2}/iPrOH catalyst systems, illustrating the importance of the –CH2OCH2R* exocyclic side chain to impart stereoregularity on these dichloro-substituted bisphenolate yttrium catalysts.9,11–17

Interestingly, stereoregular PHAs have also been prepared by ROP of the eight-membered R-substituted cyclic diolide (DLR).25 Rewardingly, ROP of rac-DLMe mediated by yttrium-salen complexes generated perfectly isotactic well-defined high molar mass PDLMe (aka PHB, poly(3-hydroxybutyrate), with Pm > 0.99, Mn = 154[thin space (1/6-em)]000 g mol−1 and ĐM = 1.01).26 Isotactic PHA copolymers were similarly produced from the ring-opening copolymerization of rac-DLMe with rac-DLR′ (R′ = Et, nBu) (Pm > 0.99) or through the design of unsymmetrically disubstituted eight-membered diolides rac-DLR,R′, with R ≠ R′, returning alternating isotactic PHAs.27–29 Syndiotactic stereodiblock PHAs were also successfully obtained from the ROP of meso-DLR diastereomers (Scheme 2).30


image file: d3py01430d-s2.tif
Scheme 2 Yttrium-salen-mediated stereoselective ROP of rac- and meso-DLR/R′s (R/R′ = Me, Et, nBu, Bn) producing stereoregular isotactic or syndiotactic PHAs.25–30

Of note, amino-alkoxy-bisphenolate yttrium catalyst systems have also been successfully reported to give stereoregular polymers from the ROP of other racemic cyclic monomers. These include the five-membered γ-butyrolactone (GBL), the six-membered ring δ-valerolactone, and the related fused six-five bicyclic lactones, namely 3,4-trans-cyclohexyl and 4,5-trans-cyclohexyl fused γ-butyrolactones, respectively. Similarly, ROP of glycolide (GA), lactide/glycolide, and O-carboxyanhydride (OCA) monomers with such yttrium-based catalyst systems has been reported to provide stereoregular poly(α-hydroxyalkanoate)s5–7,31 and references therein.

In order to further verify our above-mentioned hypothesis of an intimate stereoelectronic relationship between the β-lactone exocyclic substituent and the suitable yttrium R ancillary substituents, which can impart isoselectivity in the ROP of 4-substituted rac-BPLCH2OCH2R*, we thus report herein the ROP of the 4-methylene-alkoxy-fluorinated substituted β-propiolactone, rac-BPLCH2OCF2CHF2. The ROP reactions were conducted using diamino-bisphenolate yttrium catalyst systems Y{ONNOR2}/iPrOH bearing various ortho,para-R,R substituents (R = Me, Cl, tBu, CMe2Ph, namely 1a–d/iPrOH) (Scheme 3). In rac-BPLCH2OCF2CHF2, the outer methylene hydrogens are replaced by fluorine in order to inhibit the outer O[C with combining low line]R*[H with combining low line]2Cl(ligand) NCIs, and ultimately to evaluate the impact of such an interaction on the resulting PHA stereoregularity. Hence, the ROP of rac-BPLCH2OCF2CHF2 was studied: (i) to determine and rationalize the activity of catalysts, especially in relation to the stereoelectronic effects of the phenolate R substituents, (ii) to fully characterize the resulting PBPLCH2OCF2CHF2s at the molecular and microstructural levels by NMR spectroscopy, SEC, mass spectrometry, DSC and TGA thermal analyses, and (iii) finally to assess the catalyst ability in affording stereoregular PHAs. Ultimately, the suggested contribution of the combined stereoelectronic parameters and involvement of NCIs to the control of stereoregularity of the resulting PHAs was successfully verified.


image file: d3py01430d-s3.tif
Scheme 3 Stereoselective ROP of racemic 4-((2,2,3,3-tetrafluoropropyl)methylene) β-propiolactone, rac-BPLCH2OCF2CHF2, catalyzed by diamino-bisphenolate yttrium complexes Y{ONNOR2} 1a–d in the presence of iPrOH as the co-initiator.

Fluorinated polymers are attractive for their thermal, chemical and electronic stabilities. They are typically very hydrophobic, not flammable, inert to acids, bases, solvents, and oils, and highly resistant to aging and oxidation, and exhibit low dielectric constants, low refractive indexes and low surface tension. Ranging from semi-crystalline to amorphous forms and from thermoplastic to elastomeric materials, they have found various applications from building and construction, automotive, petrochemical, aeronautics and aerospace, photonics, electronics, and biomedical systems. Side-chain fluorinated polymers are valuable for their surface properties and are often employed as polymer dispersions in water used as coatings applied to textiles, carpets, nonwovens and paper to provide water, soil, oil and stain resistance.32–35 In addition, to our knowledge, besides microbial PHAs containing fluorinated side-chain substituents,36–38 there are only two examples of chemically synthesized PHAs featuring a fluorinated pendant substituent on the repeating units or a fluorinated end-capping group, which are still being obtained by post-polymerization functionalization.39,40 While the novel PBPLCH2OCF2CHF2s reported herein add to the previously established short series of fluoroalkyl PHAs, all of which were atactic polymers,41 they represent the first extended series of stereoregular fluorinated PHAs chemically prepared by ROP of a β-lactone monomer.

Experimental section

Methods and materials

All manipulations involving organometallic catalysts were performed under an inert atmosphere (argon, <3 ppm O2) using standard Schlenk, vacuum line, and glovebox techniques. Solvents were freshly distilled from Na/benzophenone under argon and degassed thoroughly by freeze–thaw–vacuum cycles prior to use. Isopropyl alcohol (Acros) was distilled over Mg turnings under an argon atmosphere and placed over activated 3–4 Å molecular sieves. Bisphenol proligands {ONXOR2}H2, yttrium amide precursors Y[N(SiHMe2)2]3(THF)2 of 1a–d,10,42 [Salph(Cr(THF)2)][Co(CO)4]43–45 and (BDI)ZnN(SiMe3)2[thin space (1/6-em)]46,47 (BDI = β-diketiminate) were synthesized according to the literature procedure. Racemic 2-((1,1,2,2-tetrafluoroethoxy)methyl)oxirane (rac-GCH2OCF2CHF2) (ABCR) was dried over CaH2 and then stored under argon at −27 °C (1H, J-MOD and 19F NMR data – refer to Fig. S2–S4). Enantioenriched (S)-GCH2OCF2CHF2 (>90% ee) was prepared as a colorless oil (1.6 g, 40%) by hydrolytic kinetic resolution (HKR) of rac-GCH2OCF2CHF2, following a reported procedure (refer to the ESI, Scheme S1 and Fig. S1).48 All other reagents were purchased from Aldrich, Sigma or Acros and used as received.

Instrumentation and measurements

1H (500 and 400 MHz), 13C{1H} (125 and 100 MHz) and 2D (COSY, HMBC, HSQC, and DOSY) NMR spectra were recorded on Bruker Avance AM 500 or Ascend 400 spectrometers at 25 °C. 1H and 13C{1H} NMR spectra were referenced internally relative to SiMe4 (δ 0 ppm) using the residual solvent resonances. Coupling constants are reported in Hz.

Monomer conversions were calculated from the 1H NMR spectra of the crude polymer samples in CDCl3 or (CD3)2CO by using the integration (Int.) ratios Int.PBPL(CH2OCF2CHF2)/[Int.PBPL(CH2OCF2CHF2) + Int.BPL(CH2OCF2CHF2)] of the methine hydrogen signal of BPLCH2OCF2CHF2 and PBPLCH2OCF2CHF2 (the corresponding methine hydrogen signal of the polymer (see above) and of the monomer δ (ppm) 4.71).

Chiral gas chromatography analysis of (S)-GCH2OCF2CHF2 was performed on a GC/FID VARIAN CP-3380 chromatograph equipped with a Chiralsil-Dex CB Varian CP7502 Chrompack.

Number-average molar mass (Mn,SEC), weight-average molar mass (Mw,SEC) and dispersity (ĐM = Mw/Mn) values of the PBPLFGs were determined by size-exclusion chromatography (SEC) in THF at 30 °C (flow rate = 1.0 mL min−1) on Polymer Laboratories PL50 apparatus equipped with a refractive index detector and a set of two ResiPore PLgel 3 μm MIXED-D 300 × 7.5 mm columns. The polymer samples were dissolved in THF (5 mg mL−1). All elution curves were calibrated with polystyrene standards (Mp = 290[thin space (1/6-em)]300, 126[thin space (1/6-em)]000, 70[thin space (1/6-em)]500, 30[thin space (1/6-em)]230, 19[thin space (1/6-em)]920, 9960, 4900, 3320, 1180, and 580 g mol−1); Mn,SEC values of the PBPLFGs were uncorrected for the possible difference in the hydrodynamic radius vs. those of polystyrene.

The molar mass of PBPLFG samples was also determined by 1H NMR analysis in CDCl3 or (CD3)2CO from the relative intensities of the signals of the PBPLCH2OCF2CHF2 repeating unit methine hydrogen (δ (ppm): 5.48 –OCH(CH2OCF2CHF2)CH2), and the isopropyl chain-end (δ (ppm): 4.98 (CH3)2CHO−, 1.21 (CH3)2CHO−).

High resolution matrix-assisted laser desorption ionization-time of flight, MALDI-ToF, mass spectra of the polymers were recorded using an ULTRAFLEX III TOF/TOF spectrometer (Bruker Daltonik Gmbh, Bremen, Germany) in positive ionization mode. Spectra were recorded using reflectron mode and an accelerating voltage of 25 kV. A mixture of a freshly prepared solution of the polymer in THF or CH2Cl2 (HPLC grade, 10 mg mL−1) and DCTB (trans-2-(3-(4-tert-butylphenyl)-2-methyl-2-propenylidene)malononitrile), and a MeOH solution of the cationizing agent (NaI, 10 mg mL−1) was prepared. These solutions were combined in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v/v ratio of matrix-to-sample-to-cationizing agent. The resulting solution (ca. 0.25–0.5 μL) was deposited onto the sample target (Prespotted AnchorChip PAC II 384/96 HCCA) and air or vacuum dried.

Differential scanning calorimetry (DSC) analyses were performed with a DSC2500 TA Instrument calibrated with indium, at a rate of 10 °C min−1, under a continuous flow of helium (25 mL min−1), using aluminum capsules (40 μL). The thermograms were recorded according to the following cycles: −80 to +200 °C at 10 °C min−1; +200 to −80 °C at 10 °C min−1; −80 °C for 5 min; −80 to +200 °C at 10 °C min−1; +200 to −80 °C at 10 °C min−1. Melting-transition (Tm) and glass-transition (Tg) temperatures were measured from the second heating run.

Thermogravimetric analyses (TGA) were performed on a Mettler Toledo TGA/DSC1 by heating the polymer samples at a rate of 10 °C min−1 from +25 to +500 °C under a dynamic nitrogen atmosphere (flow rate = 10 mL min−1). The onset decomposition temperature (Td) was defined as the temperature for 5% weight loss.

Carbonylation of racemic 2-((1,1,2,2-tetrafluoroethoxy)methyl)oxirane (rac-GCH2OCF2CHF2) into racemic 4-((1,1,2,2-tetrafluoroethoxy)methyl)oxetan-2-one (rac-BPLCH2OCF2CHF2)41,49. The procedure used was adapted from a previously reported synthesis, especially for the purification of epoxide (previously performed by distillation at 70 °C) and scale-up synthesis (0.96 g, 79%).41 In a typical experiment, in a glovebox, a Schlenk flask was charged with [Salph(Cr(THF)2)][Co(CO)4] (248 mg, 0.274 mmol). On a vacuum line, dry DME (15 mL) was syringed in and the resulting solution was cannulated into a degassed high-pressure stainless steel reactor, which was pressurized with carbon monoxide (30 bars), and stirred with a magnetic bar for 15 min before depressurization. A solution of rac-GCH2OCF2CHF2 (3.01 g, 17.28 mmol, 63 equiv. vs. Cr) in dry DME (15 mL) was transferred into the reactor, which was then pressurized with CO to 40 bars. The reaction mixture was stirred for 2 days at 20 °C. The reactor was then vented to atmospheric pressure, volatiles were removed under vacuum and the crude product was purified through a silica column (CH2Cl2, 2 × 300 mL). Evaporation of volatiles afforded 4-((1,1,2,2-tetrafluoroethoxy)methyl)oxetan-2-one, rac-BPLCH2OCF2CHF2, as a yellowish oil (2.60 g; 75%). 1H NMR (400 MHz, (CD3)2CO, 25 °C) δ (ppm): 6.27 (tt, 3JH–F = 3.0, 2JH–F = 52.0, 1H, OCF2CHF2), 4.87 (m, 2H, CH2OCHF2), 4.47 (dd, 3JH–H = 4.4, 2JH–H = 12.0, 1H, CHCHHO), 4.36 (dd, 3JH–H = 5.5, 2JH–H = 12.0, 1H, CHCHHO), 3.65 (dd, 3JH–H = 6.0, 2JH–H = 17.0, 1H, C(O)CHHCH), 3.43 (dd, 3JH–H = 4.0, 2JH–H = 16.5, 1H, C(O)CHHCH) (Fig. S5). 13C J-MOD NMR (125 MHz, (CD3)2CO, 25 °C) δ (ppm): 167.7 (C[double bond, length as m-dash]O), 118.4 (OCF2CF2H), 109.0 (tt, 1JC–F = 247, 2JC–F = 40, OCF2CF2H), 68.5 (CHOC(O)), 65.3 (4JH–F = 5, CH2OCF2), 40.4 (OC(O)CH2CH) (Fig. S6). 19F NMR (376 MHz, (CD3)2CO, 25 °C) δ (ppm): −138.4 (dt, 2JF–H = 52.7, 3JF–F = 5.6, 2F, OCF2CF2H), −92.3 (dt, 3JF–H = 3.0, 3JF–F = 5.6, 2F, OCF2CF2H) (Fig. S7).

The carbonylation of (S)-2-((1,1,2,2-tetrafluoroethoxy)methyl)oxirane (>90% ee) was performed similarly and (S)-BPLCH2OCF2CHF2 (>90% ee) was obtained as a colorless oil (0.59 g, 74%), which NMR spectra were identical to those of rac-BPLCH2OCF2CHF2 (Fig. S2–S4). Both rac-BPLCH2OCF2CHF2 and (S)-BPLCH2OCF2CHF2 were stored under argon in a fridge at −4 °C.

Typical rac-BPLCH2OCF2CHF2 polymerization procedure10–17

In a typical experiment (Table 1, entry 4), in a glovebox, a Schlenk flask was charged with [Y(N(SiHMe2)2)3](THF)2 (8.8 mg, 14.0 μmol) and {ONNOtBu2}H2 (1c, 6.2 mg, 9.8 μmol), and toluene (0.4 mL) was next added. To this solution, iPrOH (75.7 μL of a 1% (v/v) solution in toluene, 1 equiv. vs. Y) was added under stirring at room temperature (ca. 20 °C). After 5 min of stirring, a solution of rac-BPLCH2OCF2CHF2 (200 mg, 0.98 mmol, 100 equiv. vs. Y) in toluene (0.5 mL) was added rapidly and the mixture was stirred at 20 °C overnight (Fig. 1 and 2). The reaction was quenched by addition of acetic acid (ca. 0.5 mL of a 1.6 mol L−1 solution in toluene). The resulting mixture was concentrated to dryness under vacuum and the conversion was determined by 1H NMR analysis of the residue in (CD3)2CO. The crude polymer was then dissolved in acetone (ca. 1 mL) and precipitated in cold pentane (ca. 5 mL), filtered and dried. PBPLCH2OCF2CHF2 was recovered as a white solid (120 mg, 60%). 1H NMR (400 MHz, (CD3)2CO, 25 °C): δ (ppm) 6.23 (tt, 3JH–F = 3, 1JH–F = 53, 1H, OCF2CHF2), 5.47 (m, 1H, CH2CHCH2O), 4.99 (septet, 2JH–H = 7, 1H terminal group, (CH3)2CHC(O)), 4.27 (m, 2H, CHCH2OCHF2), 2.82 (m, 2H, C(O)CH2CH), 1.21 (d, 6H terminal group, (CH3)2CHOC(O)); 13C J-MOD NMR (100 MHz, (CD3)2CO, 25 °C) δ (ppm): 169.6 (C[double bond, length as m-dash]O), 118.3 (tt, 1JC–F = 250, 2JC–F = 42, OCF2CF2H), 109.3 (tt, 1JC–F = 269, 2JC–F = 32, OCF2CF2H), 69.1 (C(O)CH2CH), 66.7 ((CH3)2CHO), 65.5 (CHCH2O), 35.5 (C(O)CH2CH), 22.0 ((CH3)2CHO); 19F{1H} NMR (376 MHz, (CD3)2CO, 25 °C) δ (ppm): −138.2 (OCF2CF2H), −92.1 (OCF2CF2H) (Fig. 3). All recovered polymers were then thoroughly analyzed by NMR spectroscopy (Fig. 2, 3, 5 and S8–S13), SEC (Fig. 2 and S14–S16), mass spectrometry (Fig. 4), DSC (Fig. S17–S19) and TGA (Fig. S20–S22).
image file: d3py01430d-f1.tif
Fig. 1 Semi-logarithmic first-order plot for the ROP of rac-BPLCH2OCF2CHF2 mediated by 1b/iPrOH (20 °C, toluene; [BPLCH2OCF2CHF2]0/{[1c]0/[iPrOH]0} = 250[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1; Table 1 entries 11–14).

image file: d3py01430d-f2.tif
Fig. 2 Variation of Mn,NMRimage file: d3py01430d-u1.tif, Mn,SECimage file: d3py01430d-u2.tif, and Mn,theo ● (solid line) values of PBPLCH2OCF2CHF2 synthesized from the ROP of rac-BPLCH2OCF2CHF2 mediated by the 1c/iPrOH (1[thin space (1/6-em)]:[thin space (1/6-em)]1) catalyst system as a function of the BPLCH2OCF2CHF2 monomer loading (Table 1, entries 8, 10, 15, and 16).

image file: d3py01430d-f3.tif
Fig. 3 1H NMR (400 MHz, (CD3)2CO, 25 °C), 13C{1H} J-MOD NMR (100 MHz, (CD3)2CO, 25 °C) and 19F{1H} NMR (376 MHz, (CD3)2CO, 25 °C) spectra of a PBPLCH2OCF2CHF2 polymer prepared from the ROP of rac-BPLCH2OCF2CHF2 mediated by the 1c/iPrOH (1[thin space (1/6-em)]:[thin space (1/6-em)]1) catalyst system (Table 1, entry 10). * and ** stand for resonances of the residual solvent and/or catalyst.

image file: d3py01430d-f4.tif
Fig. 4 High-resolution MALDI-ToF mass spectrum (top; positive mode, DCTB matrix, Na+ cationizing salt) of a PBPLCH2OCF2CHF2 sample prepared from the ROP of rac-BPLCH2OCF2CHF2 (M = 202 g mol−1) using 1d/iPrOH (Table 1, entry 17), showing the zoomed regions of different populations I–IV (middle), and the corresponding correlation with the simulated spectra (bottom). * refers to the same series minus one H.

image file: d3py01430d-f5.tif
Fig. 5 Zoomed carbonyl region of the 13C J-MOD NMR spectra (125 MHz, (CD3)2CO, 23 °C) of PBPLCH2OCF2CHF2s prepared by ROP of rac-BPLCH2OCF2CHF2, except for the top spectra of enantiopure (S)-BPLCH2OCF2CHF2, mediated by (BDI)ZnN(SiMe3)2/iPrOH and 1a–d/iPrOH catalyst systems in toluene at 20 °C (Table 1, entries 1, 2, 3, 5, 8 and 18).
Table 1 Characteristics of the PBPLCH2OCF2CHF2s synthesized by ROP of rac-BPLCH2OCF2CHF2 mediated by 1a–d/iPrOH catalyst systemsa
Entry Catalyst Solvent [M]0/[1]0[thin space (1/6-em)]a Timeb (h) Conv.c (%) TOFd (h−1) M n,theo[thin space (1/6-em)]e (g mol−1) M n,NMR[thin space (1/6-em)]f (g mol−1) M n,SEC[thin space (1/6-em)]g (g mol−1) Đ M P r [thin space (1/6-em)]h T g [thin space (1/6-em)]i (°C)
a Reactions performed with [BPLCH2OCF2CHF2]0 = [M]0 = 1.0 M in toluene, with [1]0/[iPrOH]0 = 1, at room temperature. b Reaction times were not necessarily optimized. c Conversion of BPLCH2OCF2CHF2 as determined by 1H NMR analysis of the crude reaction mixture. d Unoptimized turnover frequency, in mol(BPLCH2OCF2CHF2) mol(Y)−1 h−1, as determined from TOF = conversion × [M]0/[1]0/reaction time. e Molar mass calculated according to Mn,theo = ([BPLCH2OCF2CHF2]0/[1]0 × conv.BPLCH2OCF2CHF2 × MBPLCH2OCF2CHF2) + MiPrOH with MBPLCH2OCF2CHF2 = 202 g mol−1 and MiPrOH = 60 g mol−1. f Molar mass determined by 1H NMR analysis of the isolated polymer, from the resonances of the terminal OiPr group. g Number-average molar mass (uncorrected values) and dispersity (ĐM = Mw/Mn) determined by SEC analysis in THF at 30 °C vs. polystyrene standards. h [thin space (1/6-em)]Pr is the probability of racemic linkages between BPLCH2OCF2CHF2 units as determined by 13C{1H} NMR analysis of the isolated PBPLCH2OCF2CHF2 (refer to the Experimental section). i Glass transition temperature as determined by DSC analysis. j Zn = (BDI)Zn(N(SiMe3)2)/iPrOH. k ROP of enantioenriched (S)-BPLCH2OCF2CHF2 (>90% ee) that gives an isotactic-rich polymer50 with Tm = 101.9 °C, Tc = 67.2 °C and ΔHm = 0.42 J g−1. l n.o.: no signal observed. m n.d.: not determined.
1 Zn[thin space (1/6-em)]j Toluene 100 15.75 82 5.2 16[thin space (1/6-em)]500 14[thin space (1/6-em)]800 9100 1.03 0.45 −25.0
2 1d[thin space (1/6-em)]k Toluene 100 27 100 >3.7 20[thin space (1/6-em)]200 14[thin space (1/6-em)]200 13[thin space (1/6-em)]000 1.07 0.09 −7.8k
3 1a Toluene 50 48 100 >1.0 10[thin space (1/6-em)]100 7500 9300 1.12 0.57 −20.7
4 1a Toluene 100 46 47 1.0 9500 6200 5300 1.43 0.53 n.o.l
5 1b Toluene 50 16 100 >6.2 10[thin space (1/6-em)]100 8100 8400 1.06 0.77 −19.2
6 1b Toluene 100 29 97 >3.3 20[thin space (1/6-em)]000 15[thin space (1/6-em)]500 16[thin space (1/6-em)]200 1.15 0.69 −17.4
7 1b Toluene 100 18 100 >5.5 20[thin space (1/6-em)]200 21[thin space (1/6-em)]800 17[thin space (1/6-em)]400 1.14 0.74 −11.9
8 1c Toluene 60 0.5 97 116 11[thin space (1/6-em)]800 11[thin space (1/6-em)]200 11[thin space (1/6-em)]200 1.04 0.86 −22.8
9 1c Toluene 60 3 min 94 1128 11[thin space (1/6-em)]500 10[thin space (1/6-em)]900 11[thin space (1/6-em)]600 1.06 0.85 −19.4
10 1c Toluene 100 18.5 100 >5.4 20[thin space (1/6-em)]200 16[thin space (1/6-em)]000 11[thin space (1/6-em)]800 1.03 0.79 −22.9
11 1c Toluene 250 30 s 13 3900 6700 7100 9500 1.05 0.73 −16.8
12 1c Toluene 250 1 min 31 4650 15[thin space (1/6-em)]700 13[thin space (1/6-em)]800 14[thin space (1/6-em)]200 1.02 0.85 −15.6
13 1c Toluene 250 2 min 49 3675 24[thin space (1/6-em)]700 24[thin space (1/6-em)]500 25[thin space (1/6-em)]300 1.04 0.83 −13.6
14 1c Toluene 250 5 min 54 1620 27[thin space (1/6-em)]400 20[thin space (1/6-em)]900 25[thin space (1/6-em)]300 1.05 0.84 −14.4
15 1c Toluene 250 6 100 >41.6 50[thin space (1/6-em)]600 53[thin space (1/6-em)]600 61[thin space (1/6-em)]600 1.09 0.79 −7.5
16 1c Toluene 500 4 100 >125 101[thin space (1/6-em)]100 106[thin space (1/6-em)]000 102[thin space (1/6-em)]600 1.24 0.79 −5.8
17 1d Toluene 50 2 100 >25 10[thin space (1/6-em)]100 10[thin space (1/6-em)]900 11[thin space (1/6-em)]400 1.05 0.85 −17.0
18 1d Toluene 100 27.3 100 >3.7 20[thin space (1/6-em)]200 16[thin space (1/6-em)]500 12[thin space (1/6-em)]900 1.07 0.86 −17.9
19 1b C6H5Cl 50 72 100 >0.7 10[thin space (1/6-em)]100 6500 13[thin space (1/6-em)]000 1.06 0.67 n.d.m
20 1d C6H5Cl 100 1 100 >100 20[thin space (1/6-em)]200 21[thin space (1/6-em)]500 14[thin space (1/6-em)]500 1.13 0.87 n.d.m
21 1b THF 20 48 24 0.1 1300 1100 800 1.02 0.59 n.d.m
22 1b THF 100 29 17 0.6 3560 2200 2500 1.21 0.66 n.d.m
23 1d THF 100 29 48 1.6 9700 5300 5000 1.09 0.83 n.d.m


Results and discussion

The herein investigated fluorinated functional 4-substituted-β-propiolactone rac-BPLCH2OCF2CHF2 was first synthesized in good yields on a multi-gram scale (up to 2.6 g, ca. 75%) from the carbonylation reaction of the parent epoxide, rac-BPLCH2OCF2CHF2, according to the adaptation of a typical procedure (Fig. S5–S7).41 While the synthesis of rac-BPLCH2OCF2CHF2 was previously reported by Coates and coworkers, its preliminary homopolymerization therein mediated by a zinc β-diketiminate catalyst (BDI)Zn(OiPr) (6 h at 50 °C) returned a single PBPLCH2OCF2CHF2 sample (95% monomer conversion, Mn,SEC(THF) = 20[thin space (1/6-em)]000 g mol−1 and ĐM = 1.14, Tg = −12 °C, no Tm observed; atactic polymer).41 No further ROP was therein reported. For the sake of comparison with our data, we have reproduced this latter experiment under the conditions implemented in our study, i.e., in toluene at room temperature and using the in situ catalyst combination (BDI)Zn(N(SiMe3)2)/iPrOH (Table 1, entry 1).

The ROP of rac-BPLCH2OCF2CHF2 promoted by the yttrium catalyst systems Y{ONNOR2}/iPrOH (R = Me, Cl, tBu, CMe2Ph, 1a–d/iPrOH) was then investigated using the previously established optimized conditions, namely in toluene solution at room temperature (ca. 20 °C) (Scheme 3). The catalysts were conveniently generated in situ upon the addition of 1 equiv. of iPrOH, added as a co-initiator to a mixture of the proligand {ONNOR2}H2 and the yttrium amido precursor [Y(N(SiHMe2)2)3](THF)2, respectively, as typically performed.10–17 Of note, the exploration of the impact of different para-R substituents on the phenolate proligand of complexes 1 was not undertaken in the present study, since they were previously demonstrated to have, in contrast to ortho-substituents, a minimal/negligible effect on the catalyst/initiator activity and stereoselectivity in the ROP of alike β-propiolactones.9–17 Also, reaction conditions affording controlled, maximized or optimized molar mass and dispersity values were not sought after in the present work. The main objective of the current study was to gain further insights into the contribution to the stereocontrol of the ROP of potential NCIs between hydrogens from the methylene within the –[C with combining low line][H with combining low line]2O[C with combining low line][H with combining low line]2R* moiety of the functional side group on the repeating unit of the growing polymer chain, and the ortho-aryl R substituents of the {ONNOR2} ligand. The ROP performance of the catalyst systems was first assessed, including their ability to promote stereoselective polymerization, as assessed by NMR and mass spectrometry analyses of the macromolecular characteristics of the resulting polyesters. The most significant data for the synthesized PBPLCH2OCF2CHF2s are gathered in Table 1 and discussed thereafter.

Among the different catalyst systems 1a–d/iPrOH, the dimethylated Y{ONNOMe2} 1a was found to be the least active in the ROP of rac-BPLCH2OCF2CHF2, only consuming 50 monomer units in 2 days with a turnover frequency (TOF expressed in molBPLCH2OCF2CHF2 molcatalyst−1 h−1) TOF1a of 1.04 h−1 (Table 1, entries 3 and 4). The similarly uncrowded dichloro Y{ONNOCl2} 1b system showed a slightly better, yet still low activity, enabling the complete consumption of 100 monomer equiv. in 18 h with TOF1b = 5.6 h−1 (Table 1, entries 5–7). Similar modest activities of these two catalyst systems were previously observed in the ROP of rac-LA and rac-4-substituted β-lactones and accounted for the likely formation of dinuclear/aggregated yttrium species, due to insufficient bulkiness of the ortho-substituents of the ligand.9,11 The alike ROP from 1b performed in chlorobenzene or THF turned out to be even slower – and poorly controlled in terms of molar mass and dispersity – regardless of the solvent polarity (TOF1b/C6H5Cl = 0.7 h−1; TOF1b/THF < 0.6 h−1; Table 1, entries 19, 21 and 22). Then, the bulkier cumyl-substituted yttrium catalyst 1d (R = CMe2Ph), surprisingly, did not improve the ROP kinetics much, regardless of the solvent used (toluene, chlorobenzene or THF; TOF1d < 4 h−1; Table 1, entries 2, 17, 18, 20 and 23). Nevertheless, ROP performed with 1d in slightly polar non-coordinating chlorobenzene was significantly faster than that in THF, in which the solvent competed with the monomer to coordinate with the yttrium center, a phenomenon already well established.9–17 This rather poor activity of the bulky catalyst 1d contrasts with the one observed for the similarly bulky, yet more active tBu-substituted catalyst system Y{ONNOtBu2} 1c. The latter one, expectedly,9,11 was found to be significantly more active, enabling the ready, complete conversion of 60 to 500 equiv. of rac-BPLCH2OCF2CHF2 in toluene at room temperature (Table 1, entries 8–16). While 500 monomer equiv. were fully consumed in less than 4 h (TOF1c ≫ 125 h−1; Table 1, entry 16; note that reaction times herein reported were not optimized), the highest activity was observed for the conversion of 77 monomer equiv. within 30 s (TOF1c/77 equiv. = 4650 h−1; Table 1, entry 12). Only the highest activity previously achieved in the ROP of the parent β-thiobutyrolactone (TOF1c/50 equiv. up to 3000 h−1)51 and the correlated malolactonates rac-BPLCO2Me,CO2All,CO2Bn (TOF100 equiv. = 3000 h−1),17 yet mediated by the related tBu-substituted yttrium catalyst featuring a methoxy cap {ON([O with combining low line][M with combining low line][e with combining low line])OtBu2} in place of the amido cap {ON([N with combining low line][M with combining low line][e with combining low line]2)OtBu2} as in the present catalyst 1c, is close to this present record value herein established.8,10–17 Monitoring the polymerization of rac-BPLCH2OCF2CHF2 performed with 1c/iPrOH returned a linear semi-logarithmic plot, thus revealing that the reaction was first order in the monomer with an apparent rate constant kapp = 2.84 ± 0.034 in h−1 (Fig. 1). Hence, overall, the ROP of rac-BPLCH2OCF2CHF2 proceeded more rapidly with the more sterically encumbered tBu-substituted catalyst 1c and falls within the range of rates recorded for other related functionalized BPLFGs (FG = CH2OMe, CH2OAll, CH2OPh, CH2CH2OBn, CH2CH2OiPr, CH2CH2OtBu, and CH2CH2OSitBuMe2), using the same diamino- and amino-alkoxy-bisphenolate yttrium catalysts.9–17 The unexpected low activity of the cumyl-substituted system 1d possibly suggests detrimental (in terms of activity) electronic interactions involving the phenyl rings in the cumyl-substituents; attractive CH⋯π electronic interactions involving the phenyl rings of the cumyl-substituted catalyst have already been suggested in the ROP of rac-BPLMe with similar Y{ONXOcumyl2} systems,9 as well in the ROP of rac-LA with {BDIaryl}Mg(OR) catalyst systems, as supported by DFT calculations.19 Yet, these were not probed in the present study.

The catalyst systems 1a–d/iPrOH showed good control over the macromolecular characteristics of the polyesters, especially in terms of molar mass and dispersity. The experimental molar mass values, as determined by 1H NMR and SEC analyses of the PBPLCH2OCF2CHF2 samples (Mn,NMR and Mn,SEC, respectively), were generally matching the calculated data (Mn,theo) established from the monomer conversion as estimated by NMR analysis of the crude samples (i.e., before precipitation of the polymer; refer to the Experimental section) (Table 1). PBPLCH2OCF2CHF2 samples with Mn,NMR values ranging from 2300 to 106[thin space (1/6-em)]000 g mol−1 were thus isolated with a fairly narrow dispersity value ranging from 1.02 to 1.24 (Table 1). As depicted in Fig. 2 for the ROP of rac-BPLCH2OCF2CHF2 promoted by 1c/iPrOH, the experimental Mn,NMR values varied linearly with the monomer loading/conversion, in agreement with the theoretical data as well. Altogether, these observations thereby confirmed the ability of these Y{ONXOR2} systems to limit the occurrence of side reactions, namely inter- and intra-molecular transesterifications, i.e., reshuffling and back-biting reactions, respectively, as typically encountered in ROP of cyclic esters or carbonates.

The PBPLCH2OCF2CHF2 polymers were characterized by 1D, 2D NMR and mass spectrometry analyses. The 1H, 13C J-MOD, 19F NMR and 2D spectra showed the expected characteristic signals corresponding to the BPLCH2OCF2CHF2 repeating units (Fig. 3 and S7–S9). In particular, the methine (δ 5.47 ppm) and methylene (δ 2.82 ppm) typical 1H signals of the polyester backbone, along with the distinctive signals of the –CH2OCF2CHF2 pendent moiety observed in 1H, 13C and 19F NMR spectra (δCH2O 4.27 ppm, δCHF2 6.23 ppm, δCH2O 65.5 ppm, δOCF2 118.4 ppm, δCHF2 109.4 ppm, δOCF2 −92.1 ppm, and δCHF2 −138.2 ppm), all supported the formation of the awaited PBPLCH2OCF2CHF2. The characteristic isopropoxycarbonyl chain-end group signature was also unambiguously observed (δOCH 4.99 ppm, δOCHMe2 1.21 ppm; δOCHMe2 66.7 ppm, δOCHMe2 22.0 ppm), thus supporting the formation of linear macromolecules and that the in situ prepared {ONNOR2}Y(OiPr) alkoxide complex is indeed the active species that initiates the polymerization (Scheme 3). Furthermore, the DOSY NMR spectrum of a PBPLCH2OCF2CHF2 sample evidences the presence of a single population of macromolecules that are end-capped by isopropoxy end-groups (Table 1, entry 17; Fig. S10). This selective formation of isopropoxy terminated macromolecules is also corroborated by the good match between theoretical Mn values and experimental Mn values determined by NMR considering the presence of an isopropoxy end-group for each macromolecule (vide supra, Table 1).

MALDI-ToF mass spectrometry investigations of a low molar mass sample prepared from the 1d/iPrOH catalyst enabled us to gain deeper insights into the macromolecular structure and topology of PBPLCH2OCF2CHF2. A typical mass spectrum, recorded using a DCTB matrix, for a sample prepared from the ROP of rac-BPLCH2OCF2CHF2 using 1d/iPrOH (Table 1, entry 17), is illustrated in Fig. 4. It shows a set of four distinct populations of macromolecules ionized by Na+, all featuring the anticipated repeating unit of m/z 202 corresponding to the monomer (MBPL(CH2OCF2CHF2)) (Fig. 4). The first population of linear macromolecules featuring α-isopropoxyl and ω-hydroxyl end groups was observed (green population I), as anticipated for the ROP of rac-BPLCH2OCF2CHF2 promoted by 1d/iPrOH. The second population was assigned to linear α-isopropoxy,ω-crotonate telechelic PBPLCH2OCF2CHF2 chains, arising from ω-dehydration of the former population I (orange population II). Such a crotonate chain-end formation upon H2O elimination is not surprising, as it is a known side reaction reported to be occurring during the synthesis of polyesters by ROP of such β-lactones.13,16,52 Another population derived from population I upon abstraction of the isopropyl end-group, showed linear macromolecules with α-carboxylic acid and ω-hydroxyl end-capping groups (blue population III). A careful analysis of NMR spectra of various polymer samples did not reveal any distinctive alkene signals (anticipated δ1Hca. 6.25, 5.77 ppm, δ13Cca. 144.9, 120.5 ppm),52 nor any carboxylic acid signals (anticipated δ1Hca. 10.00 ppm, δ13Cca. 196.4 ppm),52 corresponding to such α-OiPr, ω-COCH2CH[double bond, length as m-dash]CHOCF2CHF2 or α-COOH,ω-OH telechelic PBPLCH2OCF2CHF2 chains, respectively (Fig. 4). This observation thus suggested that these latter two populations II and III were most likely formed during the mass spectrometry analysis (i.e., promoted by the acidic DCTB matrix), or that they actually represented minor populations undetectable by NMR, i.e., they corresponded to negligible species present within the isolated polymers which are overexpressed in the MS analysis. Finally, the last population observed was attributed to cyclic PBPLCH2OCF2CHF2 chains (red population IV; undetectable by NMR analysis since they feature the same repeating unit as linear macromolecules). The observation of a significant population of cyclic macromolecules by MALDI-ToF MS most likely results from the sample preparation (in an acidic DCTB matrix, which shall promote easy transesterification/cyclization) and/or incidental over-expression (co-crystallization, ionization) of cyclic macromolecules (over linear ones). These four populations of macromolecular chains were unequivocally confirmed by the close match of the high-resolution experimental spectrum with the corresponding isotopic simulations, as illustrated in Fig. 4; for example, for population I – [(CH3)2CHO(COCH2CH(CH2OCF2CHF2)O)nH]·Na+ with m/z calculated 1497.2213 vs. found 1497.213 for n = 7; population II – [(CH3)2CHO(COCH2CH(CH2OCF2CHF2)O)n (COCH2CH[double bond, length as m-dash]CHOCF2CHF2)]·Na+ with m/z calculated 1479.2110 vs. found 1479.215 for n = 6; population III – [HO(COCH2CH(CH2OCF2CHF2)O)nH]·Na+ with m/z calculated 1455.1752 vs. found 1455.172 for n = 7; and population IV – [(COCH2CH(CH2OCF2CHF2)O)n]·Na+ with m/z calculated m/z 1437.1632 vs. found 1437.168 for n = 5.

In order to assess the ability of the catalyst systems 1a–d/iPrOH to stereoselectively promote the ROP of rac-BPLCH2OCF2CHF2, and in turn to evaluate the possible role of NCIs in promoting stereocontrol, 13C J-MOD NMR spectra of the samples were closely examined. The spectrum of an isotactic-rich PBPLCH2OCF2CHF2, synthesized from the ROP of enantio-enriched (S)-BPLCH2OCF2CHF2 using 1d/iPrOH, was first acquired as a reference of an isotactic polymer (Table 1, entry 2).50 From our previous studies on ROP of chiral β-lactones mediated by such achiral yttrium diamino- or amino-alkoxy-bisphenolate complexes, we have established that the stereocontrol of these ROPs proceed through the so-called chain-end mechanism (CEM). The CEM involves tuning and selection of the chirality of the next monomer to be inserted, as dictated by the chirality of the last (first-order) or the last and penultimate (second-order) monomer unit inserted into the propagating yttrium-polymeryl species. Thus, a regular CEM of the ROP of such a chiral racemic BPLCH2OCF2CHF2 β-propiolactone, would generate syndiotactic PBPLCH2OCF2CHF2 upon inserting the monomer of configuration opposite to that of the last inserted one, thereby minimizing steric hindrance in the transition state.6–9 The 13C NMR spectra of PBPLCH2OCF2CHF2 recovered from the ROP of (S)- and rac-BPLCH2OCF2CHF2 promoted by 1a–d/iPrOH are depicted in Fig. 5 and S9, comparatively showing the carbonyl, and methylene and methine signals of the repeating units (C3OC4H2C5H(CH2OCF2CHF2)).

The highly isotactic PBPLCH2OCF2CHF2 sample provided the reference spectrum from which the catalysts stereoselectivity was thus assessed (Fig. 5 – top spectrum). It exhibits two distinct signals in the carbonyl region (δ ca. 169.60 ppm), assigned to the meso (m) and racemo (r) diads, with the meso signal being the most intense (Pm = 0.91;50δC[double bond, length as m-dash]Or ca. 169.60, δC[double bond, length as m-dash]Om ca.169.55). In comparison, the methine and methylene signals (δ 69.13, 65.53, and 35.54 ppm, respectively) did not show any deconvolution at the diad level, possibly due to the carbon coupling to fluorine (Fig. S11). Thus, the tacticity of the polymers was evaluated from the intensity of meso and racemo diad signals of the carbonyl region, from which Pm and Pr values were determined. As expected, the ROP mediated by (BDI)ZnN(SiMe3)2/iPrOH, known as a non-stereoselective catalyst towards rac-β-lactones,6,41 returned an atactic polymer (Pr = 0.45; the two diad signals are similarly intense and broadened vs. the corresponding isotactic signals). Moving to the {ONNOR2}Y catalyst systems, the uncrowded dimethyl-substituted complex 1a did not show any stereoselectivity, giving polymers with Pr = ca. 0.55, thus following the common trend we reported earlier.9,11 Note that within the series of our reported BPLFGs, high syndioselectivity using 1a was only observed − in place of the common non-stereoselective ROP affording atactic PHAs − in the ROP of the functionalized monomers with FG = CH2OiPr, CH2OtBu, CH2OPh, CH2OSitBuMe2, CH2SPh, CO2Me, CO2All, and CO2Bn (Table 2).12,13 The sterically near-equivalent dichloro catalyst 1b gave more syndio-enriched polymers with Pr = ca. 0.73. The overall stereoselective ability of catalyst 1b observed in the ROP of alike BPLFGs returns either atactic (FG = Me, CH2CH2OBn) or more commonly syndiotactic (FG = CH2OiPr, CH2OtBu, CH2OSitBuMe2, CH2OPh, CH2SPh, CO2Me, CO2All, CO2Bn) PHAs, while more unusual isotactic ones have only been reported with the 4-alkoxymethylene-β-propiolactones (FG = CH2OMe/All/Bn) (Table 2). On the other hand, both yttrium complexes bearing ligands flanked with bulky ortho,para-R,R substituents (R = tBu and cumyl; namely 1c and 1d, respectively) consistently induced the formation of highly syndio-regular polymers with Pr = 0.79–0.87. This latter propensity of the sterically bulkier yttrium catalysts to promote the formation of syndiotactic PHAs is invariably observed in the similarly yttrium-catalyzed ROP of all the BPLFGs we have investigated so far, regardless of the chemical nature of the functional pending substituent, be it methylene-alkyl, -fluoroalkyl, -alkoxy, -silyloxy, or alkoxycarbonyl exocyclic side-groups. Note that, changing the solvent from toluene to THF or chlorobenzene did not affect significantly the syndioselectivity of the resulting PBPLCH2OCF2CHF2, as assessed from 1b and 1d. However, no iso-enriched PBPLCH2OCF2CHF2 could thus be prepared under these operating conditions.

Table 2 Stereoselectivity (Pr) and thermal characteristics (Tg, Tm, and Td) of PBPLFGs with FG = Me, CO2R (R = Me, All, Bn), CH2OR (R = Me, All, Bn, Ph, iPr, tBu, SiMe2tBu), CH2SPh, CH2CH2OBn, and CH2OCF2CHF2, prepared by ROP of the corresponding rac-BPLFGs promoted by yttrium catalyst systems Y{ONXOR2} 1a–d/iPrOH (reference numbers for each β-lactone are provided below the formulae)
Cat.1 (R′ = R′′)

image file: d3py01430d-u3.tif

image file: d3py01430d-u4.tif

image file: d3py01430d-u5.tif

image file: d3py01430d-u6.tif

image file: d3py01430d-u7.tif

image file: d3py01430d-u8.tif

image file: d3py01430d-u9.tif

image file: d3py01430d-u10.tif

image file: d3py01430d-u11.tif

image file: d3py01430d-u12.tif

image file: d3py01430d-u13.tif

image file: d3py01430d-u14.tif

rac-BPLFGs
rac-BPLCH3 rac-MLAR R = Me, Bn, All rac-BPLCH2OMe rac-BPLCH2OAll rac-BPLCH2OBn rac-BPLCH2OPh rac-BPLCH2SPh rac-BPLCH2CH2OBn rac-BPLCH2OiPr rac-BPLCH2OtBu rac-BPLCH2OTBDMS rac-BPLCH2OCF2CHF2
7 and 53 7, 16 and 17 15 15 15 13 13 13 12 12 12 (This work)
a [thin space (1/6-em)]Table 1, entries 8–18. b [thin space (1/6-em)]Table 1, entries 3 and 4. c [thin space (1/6-em)]Table 1, entries 5–7. d n.d. = not determined. e Y{ONOOR2} [amino-alkoxy-bisphenolate]yttrium-amido complexes.
Crowded (cumyl, tBu) (1c and 1d)
Syndiotactic Syndiotactic Syndiotactic Syndiotactic Syndiotactic Syndiotactic Syndiotactic Syndiotactic Syndiotactic Syndiotactic Syndiotactic Syndiotactic
P r = 0.80–0.91 P r = 0.76–0.82 P r = 0.78–0.81 P r = 0.81–0.84 P r = 0.85–0.90 P r = 0.81–0.87 P r = 0.83–0.87 P r = 0.77–0.85 P r = 0.82–0.86 P r = 0.78–0.84 P r = 0.81–0.87 P r = 0.70–0.86
T m = 133–178 °C T m = 49–173 °C T m = 116 °C T m = 85 °C No Tm obsv. No Tm obsv No Tm obsv No Tm obsv No Tm obsv No Tm obsv T m = 119 °C No Tm obsv
T g = n.dd T g = 30–40 °C T g = −12 °C T g = −38 °C T g = 0 °C T g = 37–40 °C T g = 12–14 °C T g = −11–(−12) °C T g = −18 °C T g = −6 °C T g = 9 °C T g = −6–(−23) °C
M n (g mol −1 ) ≈35–60[thin space (1/6-em)]000 ≈5–15[thin space (1/6-em)]000 ≈5000 ≈7–10[thin space (1/6-em)]000 ≈10[thin space (1/6-em)]000 ≈10–80[thin space (1/6-em)]000 ≈10–80[thin space (1/6-em)]000 ≈5–20[thin space (1/6-em)]000 ≈60[thin space (1/6-em)]000 ≈5000 ≈35–90[thin space (1/6-em)]000 ≈10–100[thin space (1/6-em)]000
 
Aliphatic non-crowded
rxn n.d.d Atactic Atactic Atactic Syndiotactic Syndiotactic Atactic Syndiotactic Syndiotactic Syndiotactic Atactic
P r = 0.56 rxn n.d.d P r = 0.49 P r = 0.49 P r = 0.50 P r = 0.76 P r = 0.74–0.76 P r = 0.49 P r = 0.71–0.72 P r = 0.74–0.75 P r = 0.77 P r = 0.53–0.57
T m = n.d.d rxn n.d.d No Tm obsv. No Tm obsv. No Tm obsv. No Tm obsv. No Tm obsv. No Tm obsv. T m = n.d.d T m = n.d.d T m = n.d.d No Tm obsv.
T g = n.d.d rxn n.d.d T g = −18 °C T g = −40 °C T g = −6 °C T g = 30 °C T g = 8–9 °C T g = −12 °C T g = n.d.d T g = n.d.d T g = n.d.d T g = −21 °C
M n (g mol −1 ) ≈5000 ≈7–10[thin space (1/6-em)]000 ≈10[thin space (1/6-em)]000 ≈10[thin space (1/6-em)]000 ≈5–10[thin space (1/6-em)]000 ≈7000 ≈10[thin space (1/6-em)]000
 
Halogenated non-crowded (Cl, 1b)
Atactic Syndiotactic Isotactic Isotactic Isotactic Syndiotactic Syndiotactic Atactic Syndiotactic Syndiotactic Syndiotactic Syndiotactic
P r = 0.42–0.45 P r = 0.89–0.95+ P r = 0.10 P r = 0.09 P r = 0.10 P r = 0.75–0.77 P r = 0.73–0.74 P r = 0.49 P r = 0.70 P r = 0.70–0.71 P r = 0.76 P r = 0.69–0.77
T m = n.d. d T m = 111–207 °C No Tm obsv. No Tm obsv. No Tm obsv. No Tm obsv. No Tm obsv. No Tm obsv. T m = n.d.d T m = n.d.d T m = n.d.d No Tm obsv.
T g = n.d.d T g = 30–40 °C T g = −18 °C T g = −39 °C T g = 0 °C T g = 21–22 °C T g = 9 °C T g = −15 °C T g = n.d.d T g = n.d.d T g = n.d.d T g = −12–(−17) °C
M n (g mol −1 ) ≈10–15[thin space (1/6-em)]000 ≈5000 ≈7–10[thin space (1/6-em)]000 ≈10[thin space (1/6-em)]000 ≈5000 ≈50[thin space (1/6-em)]000 ≈5000 ≈20[thin space (1/6-em)]000
T d = 256 °C T d = n.d.d T d = n.d.d T d = n.d.d T d = n.d.d T d = 272 °C T d = 271 °C T d = 226 °C T d = 211 °C T d = 197 °C T d = 195 °C T d = 195–203 °C


Noteworthily, the dichloro-substituted catalyst 1b favored a similar syndio-enrichment in PBPLCH2OCF2CHF2 (Pr = 0.69–0.77) to that in PBPLCH2OPh (Pr = 0.73–0.77), both PHAs arising from 4-alkoxymethylene-substituted β-propiolactones, and each monomer being depleted of any outer methylene hydrogens within the terminal alkoxy moiety, namely CH2O[C with combining low line][F with combining low line]2CHF2 and CH2[O with combining low line][C with combining low line]6[H with combining low line]5, respectively (Table 2).13 Such replacement within the exocyclic alkoxy moiety of BPLCH2OC[H with combining low line][2 with combining low line]R* of hydrogen by fluorine or by a quaternary sp2 carbon, clearly impeded the formation of NCIs with the yttrium phenolate chloro substituent that was expected/understood to induce stereoselectivity in the returned PHAs. Hence, this observation further supports our previously suggested hypothesis – based on former experimental and in silico results –, that both the inner and outer two hydrogens, as in BPL[C with combining low line][H with combining low line]2O[C with combining low line][H with combining low line]2R*, are required to induce the necessary NCIs with the phenolate ortho-chloro substituent on the yttrium catalyst. Unfortunately, the perfluorinated β-lactone with any inner and outer methylene hydrogens depleted remained synthetically inaccessible.

Note that, while both inner and outer methylene hydrogens in BPL[C with combining low line][H with combining low line]2O[C with combining low line][H with combining low line]2R* experimentally appear to be simultaneously required to achieve isoselectivity (Table 2), DFT computations suggested that the NCIs of the ortho-chloro substituent are stronger with the inner methylene hydrogens (Cl[H with combining low line]2[C with combining low line]OCH2CHCH2 distance = 2.556 Å) than with the outer methylene hydrogens (Cl[H with combining low line]2[C with combining low line](CHCH2)OCH2 distance = 3.083 Å).15 Increasing the bulkiness in an outer methylene-free alkoxide BPLFG as with FG = CH2O[C with combining low line][R with combining low line]# = CH2O[i with combining low line][P with combining low line][r with combining low line], CH2O[t with combining low line][B with combining low line][u with combining low line], CH2[O with combining low line][P with combining low line][h with combining low line], CH2[O with combining low line][S with combining low line][i with combining low line][t with combining low line][B with combining low line][u with combining low line][M with combining low line][e with combining low line]2 also revealed that a sterically encumbered FG substituent could not promote isoselectivity by itself.12 Replacing oxygen (alkoxy) with sulfur (thioether) in the outer moiety (i.e., FG = CH2[S with combining low line][P with combining low line][h with combining low line] vs. CH2[O with combining low line][P with combining low line][h with combining low line]) similarly did not provide an appropriate stereoelectronic environment favorable for the formation of Cl[H with combining low line]2[C with combining low line]O NCIs.13 Lengthening the inner methylene of the alkoxymethylene functional group using an additional methylene (i.e., FG = [C with combining low line][H with combining low line]2[C with combining low line][H with combining low line]2OBn vs. [C with combining low line][H with combining low line]2OBn) was also found to be unsuccessful for iso-enrichment of the resulting PHA, in spite of the presence of both inner and outer methylenes apart the central oxygen.13 Finally, the introduction of a chemically distinct functional group (vs. alkoxymethylene FG as in BPL[C with combining low line][H with combining low line]2[O with combining low line][C with combining low line][H with combining low line]2[R with combining low line]*) using an ester substituent into the β-propiolactone monomer, as within the related alkyl β-malolactonate series (BPLFG with FG = CO2Me, CO2All, CO2Bn), also switched stereoselectivity from isotactic to syndiotactic.16,17 Last but not least, moving from classical oxolactone to its parent thiolactone upon exchanging the endocyclic oxygen with sulfur (i.e, going from β-butyrolactone to β-thiobutyrolactone) did not suitably impact the electronic environment on the exocyclic methyl substituent to provide isotacticity.51 Consequently, iso-enriched PBPLFGs have so far only been obtained from {ONXOF2,Cl2,Br2}Y-catalyzed ROP of β-propiolactone monomers featuring both inner and outer methylene hydrogens within the 4-alkoxymethylene exocyclic substituent, namely BPL[C with combining low line][H with combining low line]2[O with combining low line][C with combining low line][H with combining low line]2[R with combining low line]* with R* = CH3 (FG = CH2OMe), CH2-CH[double bond, length as m-dash]CH2 (FG = CH2OAll), or Ph (FG = CH2OBn) (Table 2).

In addition, as the crystallinity of a polymer is closely dependent on its tacticity, PBPLCH2OCF2CHF2 samples with different tacticities were analyzed by differential scanning calorimetry (DSC; Table 1 and Fig. S17–S19). All thermograms showed a sub-0 °C glass transition temperature (Tg) ranging from −25.0 to −5.8 °C, while no melting temperature (Tm) was observed below 200 °C in any of the samples (Table 1). Typically, the Tg values of PBPLCH2OCF2CHF averaged to ca. −20 °C for Mn ≤ 25[thin space (1/6-em)]000 g mol−1, while a higher molar mass resulted, as anticipated, in higher Tg values (typically −5.8 °C for Mn,NMR = 106[thin space (1/6-em)]000 g mol−1) (Table 1, entry 16). The absence of a melting transition for PBPLCH2OCF2CHF2 suggested that all the fluorinated PHAs herein prepared were amorphous, regardless of their microstructure. This is in agreement with the thermal signature of the other parent perfluoroalkylated PBPLCH2Ys with Y = CF(CF3)2, (CF2)3CF3, (CF2)2CF(CF3)2, C6F5, previously synthesized using the non-stereoselective (BDI)Zn(OiPr) catalyst − hence returning atactic polymers −, which similarly did not exhibit Tm.41 Note that among all the fluorinated PHAs prepared in this latter work, only the iso-enriched polymer (Mn,SEC = 16[thin space (1/6-em)]000 g mol−1, ĐM = 1.1) prepared from the optically pure (S)-BPLCH2OCF2CHF2, was reported to show a minimal semi-crystalline signature with Tm = 101.9 °C and ΔHm = 0.30 J g−1. However, in this former literature report, no Tg was detected, whereas our analogous sample showed a Tg value (−7.8 °C; Table 1, entry 2) but no Tm. The lack of crystallinity in PBPLFGs was similarly observed in other syndiotactic PBPLFGs, with FG = CH2OiPr, CH2OtBu, CH2OPh, CH2OBn (Table 2). Actually, among the PBPLFGs obtained by ROP of the corresponding rac-BPLFGs from 1c and 1d/iPrOH catalyst systems, the only syndiotactic PHAs that featured a Tm value are the PBPLFGs with FG = CH3, CH2OMe, CH2OAll, CH2OSitBuMe2, CO2Me, CO2All, and CO2Bn (Table 2).6,11,12,16 This lack of crystallinity in PBPLCH2OCF2CHF, no matter the tacticity of the PHA main chain, may be inherent to the relatively long fluorinated pendent moiety. Finally, the thermal degradation profile of PBPLCH2OCF2CHF2, evaluated by TGA, typically exhibited a degradation temperature (at 5% mass loss of the polymer) of Td = ca. 200 °C, within the lowest values of the 195–272 °C range determined for the related PBPLFGs (Fig. S20–S22 and Table 2).

Conclusion

ROP of fluorinated 4-methylene-alkoxy-substituted β-propiolactone rac-BPLCH2OCF2CHF2 mediated by Y{ONNOR2}/iPrOH catalyst systems effectively proceeded towards the formation of the corresponding fluoroalkylated PHAs. Complete monomer consumption was observed at room temperature, yet at a different rate depending on the ortho,para-R,R phenolate substituents, with catalysts bearing a bulkier tert-butyl substituent being more active than those bearing smaller chloride and methyl ones, and also, unexpectedly, the cumyl substituent. Well-defined high molar mass and narrowly dispersed fluorinated PHAs (Mn,NMR = 2300 to 106[thin space (1/6-em)]000 g mol−1; ĐM = 1.02–1.24) were isolated, showcasing a rather good control of the polymerization, especially emphasizing tunable molar mass values and limited undesirable transfer and/or transesterification side reactions. Molecular characterization of the polymers by NMR and mass spectrometry analyses confirmed the formation of α-isopropoxy,ω-hydroxy telechelic PBPLCH2OCF2CHF2 chains; in particular, the iPrO– and HO– chain end groups were unambiguously identified. The thermal signature of PBPLCH2OCF2CHF2 samples revealed amorphous thermoplastic fluorinated polyesters, similar to previously reported perfluroalkyl-substituted PHAs (PBPLCH2Z, vide supra).41

Detailed analyses of the 13C NMR spectra of the various polymers formed from the different catalyst systems showed that the dimethyl-substituted {ONNOMe2}Y complex 1a promoted the formation of atactic polymers, while the dichloro {ONNOCl2}Y catalyst 1b and the bulkier tert-butyl- and cumyl-substituted catalysts 1c and 1d, respectively, all returned syndio-enriched PHAs with Pr up to 0.87. The thus recovered PBPLCH2OCF2CHF2s represent the first examples of stereoregular fluorinated PHAs that were chemically synthesized by ROP of the corresponding β-lactone. This stereoselectivity behavior, for a given rac-BPLCH2OCF2CHF2 monomer, mainly dictated by the nature of the ortho-R substituent on the yttrium bisphenolate surrounding ligand, is in line with the general tendency which prevails in the ROP of such parent BPLFGs: a “simple” modification of this ligand ortho-substituent enables the tuning of the stereoselectivity from atactic to syndiotactic PHAs. The {ONNOR2}Y catalyst systems based on 1a–d failed to provide isotactic PBPLCH2OCF2CHF2; the only examples of such iso-enriched related PHAs, chemically synthesized by ROP of a functional or non-functional β-lactone, remain, to date, the unique series of 4-alkoxymethylene functionalized BPLCH2OCH2R*s with R* = H, CH[double bond, length as m-dash]CH2, or Ph. Inhibiting NCIs between the ortho-chloro-substituted phenolate and the outer methylene hydrogens within the exocyclic 4-alkoxymethylene pending substituent upon fluorinating the alkyl (i.e., CH2O[C with combining low line][F with combining low line]2CHF2vs. CH2O[C with combining low line][H with combining low line]2R*), while maintaining the inner methylene hydrogens available for such Cl[H with combining low line]2[C with combining low line]O attractive interactions, was thus found to be inefficient to impart isoselectivity during the ROP process. Combining our earlier findings gained from the BPLFG series with FG = Me, CH2OiPr, CH2OtBu, CH2OPh, CH2OSi(tBuMe2), CH2SPh, CO2Me, CO2All, and CO2Bn (Scheme 1), the experimental results reported in the present work further highlight our initial assumption: methylene hydrogens have to be present at both the adjacent positions of oxygen (i.e., both moieties bonded to oxygen) within the functional side-substituent on the monomer (FG = [C with combining low line][H with combining low line]2O[C with combining low line][H with combining low line]2R*), and the chloro(halogeno)-substituted phenolate yttrium complex 1b has to be used to enable access to isotactic PBPLFGs (Scheme 4). Would these two conditions not be simultaneously satisfied, either atactic (most typically using 1a) or syndio-enriched (commonly prepared from 1c or 1d) PBPLFGs would be recovered. The screening of this whole range of FGs in the evaluation of the efficiency of the stereoselectivity of the typical catalyst systems based on 1a–d hence confirms that the combined stereoelectronic contribution of the lactone-substituent FG and of the catalyst substituents R, along with possible NCIs between FG and R, are all determinant and interdependent. This fine FGR cooperation is ultimately confirmed as being the key parameter to understand and tune the fine microstructure of functional and non-functional PHAs prepared by chemical ROP of the β-lactone monomers.


image file: d3py01430d-s4.tif
Scheme 4 Stereoselective ROP of chiral racemic 4-substituted-β-propiolactones, rac-BPLFGs with (left) FG = [C with combining low line][H with combining low line]2O[C with combining low line][H with combining low line]2R* and R* = H, CH[double bond, length as m-dash]CH2, Ph, or (right) FG = [C with combining low line][H with combining low line]2O[C with combining low line][F with combining low line]2CHF2, performed with the Y{ON(X)OCl2}/iPrOH catalyst system, returning iso-enriched PBPLCH2OCH2R* or syndio-enriched PBPLCH2OCF2CHF2, respectively, and illustrating the importance of the methylene-alkoxy-methylene moiety within the exocyclic side chain to impart isoselectivity through attractive NCIs.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

We are grateful for the financial support provided by the University of Rennes and Lebanese University (Ph.D. grants to R. M. S. and A. D.). We thank Jérôme Ollivier and Philippe Jéhan, and Christine Deponge and Elsa Caytan for MS and NMR analyses, respectively, partly developed at ScanMAT UAR 2025 CNRS-Université de Rennes.

References

  1. L. Al-Shok, D. M. Haddleton and F. Adams, Advances in Polymer Science, Springer, Berlin, Heidelberg, 2022.  DOI:10.1007/12_2021_111.
  2. E. Glöckler, S. Ghosh and S. Schulz, Polym. Rev., 2023, 63, 478–514 CrossRef.
  3. Yu. Wang, X. Wang, W. Zhang and W. H. Sun, Organometallics, 2023, 42, 1680–1692 CrossRef CAS.
  4. D. J. Walsh, M. G. Hyatt, S. A. Miller and D. Guironnet, ACS Catal., 2019, 9, 11153–11188 CrossRef CAS.
  5. M. J.-L. Tschan, R. M. Gauvin and C. M. Thomas, Chem. Soc. Rev., 2021, 50, 13587–13608 RSC.
  6. W. Josh, P. Hannah, J. Setuhn, B. Panagiotis, B. Matthew and D. Andrew, Nat. Rev. Chem., 2019, 3, 514–535 CrossRef.
  7. H. Li, R. M. Shakaroun, S. M. Guillaume and J.-F. Carpentier, Chem. – Eur. J., 2020, 26, 128–138 CrossRef CAS PubMed.
  8. S. M. Guillaume, E. Kirilov, Y. Sarazin and J.-F. Carpentier, Chem. – Eur. J., 2015, 21, 7988–8003 CrossRef CAS PubMed.
  9. J.-F. Carpentier, Organometallics, 2015, 34, 4175–4189 CrossRef CAS.
  10. A. Amgoune, C. M. Thomas, S. Ilinca, T. Roisnel and J.-F. Carpentier, Angew. Chem., Int. Ed., 2006, 45, 2782–2784 CrossRef CAS PubMed.
  11. R. Ligny, M. M. Hänninen, S. M. Guillaume and J.-F. Carpentier, Chem. Commun., 2018, 54, 8024–8031 RSC.
  12. R. M. Shakaroun, A. Dhaini, R. Ligny, A. Alaaeddine, S. M. Guillaume and J.-F. Carpentier, Polym. Chem., 2023, 14, 720–727 RSC.
  13. R. M. Shakaroun, H. Li, P. Jéhan, M. Blot, A. Alaaeddine, J.-F. Carpentier and S. M. Guillaume, Polym. Chem., 2021, 12, 4022–4034 RSC.
  14. R. Ligny, S. M. Guillaume and J.-F. Carpentier, Chem. – Eur. J., 2019, 25, 6412–6424 CrossRef CAS PubMed.
  15. R. Ligny, M. M. Hänninen, S. M. Guillaume and J.-F. Carpentier, Angew. Chem., Int. Ed., 2017, 56, 10388–10393 CrossRef CAS PubMed.
  16. C. G. Jaffredo, Y. Chapurina, E. Kirillov, J.-F. Carpentier and S. M. Guillaume, Chem. – Eur. J., 2016, 22, 7629–7641 CrossRef CAS PubMed.
  17. C. G. Jaffredo, Y. Chapurina, S. M. Guillaume and J.-F. Carpentier, Angew. Chem., Int. Ed., 2014, 53, 2687–2691 CrossRef CAS PubMed.
  18. Note that the isoselective ROP of BPLMe (i.e., β-butyrolactone) was previously reported from other catalytic systems. Refer to: (a) J. Bruckmoser, S. Pongratz, L. Stieglitz and B. Rieger, J. Am. Chem. Soc., 2023, 145, 11494–11498 CrossRef CAS PubMed; (b) X. Dong, A. M. Brown, A. J. Woodside and J. R. N. Robinson, Chem. Commun., 2022, 58, 2854–2857 RSC; (c) X. Dong and J. R. Robinson, Chem. Sci., 2020, 11, 8184–8195 RSC; (d) Z. Zhuo, C. Zhang, Y. Luo, Y. Wang, Y. Yao, D. Yuan and D. Cui, Chem. Commun., 2018, 54, 11998–12001 RSC; (e) S. Vagin, M. Winnacker, A. Kronast, P. T. Altenbuchner, P. Deglmann, C. Sinkel, R. Loos and B. Rieger, ChemCatChem, 2015, 7, 3963–3971 CrossRef CAS; (f) N. Ajellal, G. Durieux, L. Delevoye, G. Tricot, C. Dujardin, C. M. Thomas and R. M. Gauvin, Chem. Commun., 2010, 46, 1032–1034 RSC; (g) R. Reichardt, S. Vagin, R. Reithmeier, A. K. Ott and B. Rieger, Macromolecules, 2010, 43, 9311–9317 CrossRef CAS; (h) M. Zintl, F. Molnar, T. Urban, V. Bernhart, P. Preishuber-Pflugl and B. Rieger, Angew. Chem., Int. Ed., 2008, 47, 3458–3460 CrossRef CAS PubMed.
  19. L. Marshall, V. C. Gibson and H. S. Rzepa, J. Am. Chem. Soc., 2005, 127, 6048–6051 CrossRef PubMed.
  20. S. Gesslbauer, R. Savela, Y. Chen, A. J. P. White and C. Romain, ACS Catal., 2019, 9, 7912–7920 CrossRef CAS.
  21. H.-Y. Huang, W. Xiong, Y.-T. Huang, K. Li, Z. Cai and J.-B. Zhu, Nat. Catal., 2023, 6, 720–728 CrossRef CAS.
  22. J.-F. Brière, S. Oudeyer, V. Dallab and V. Levacher, Chem. Soc. Rev., 2012, 41, 1696–1707 RSC.
  23. O. I. Kazakov, P. P. Datta, M. Isajani, E. T. Kiesewetter and M. K. Kiesewetter, Macromolecules, 2014, 47, 7463–7468 CrossRef CAS PubMed.
  24. M. E. G. Mosquera, M. P. Cebrián and M. F. Millan, Influence of Noncovalent Interactions in Catalytic Ring-opening Polymerization Processes, in Noncovalent Interactions in Catalysis, RSC Catal. Ser., 2019, ch. 19 Search PubMed.
  25. X. Tang, A. H. Westlie, E. M. Watson and E. Y.-X. Chen, Science, 2019, 366, 754–758 CrossRef CAS PubMed.
  26. X. Tang and E. Y.-X. Chen, Nat. Commun., 2018, 9, 2345 CrossRef PubMed.
  27. A. H. Westlie, S. A. Hesse, X. Tang, E. C. Quinn, C. R. Parker, C. J. Takacs, C. J. Tassone and E. Y.-X. Chen, ACS Macro Lett., 2023, 12, 619–625 CrossRef CAS PubMed.
  28. Z. Zhang, C. Shi, M. Scoti, X. Tang and E. Y.-X. Chen, J. Am. Chem. Soc., 2022, 144, 20016–20024 CrossRef CAS PubMed.
  29. X. Tang, A. H. Westlie, L. Caporaso, L. Cavallo, L. Falivene and E. Y.-X. Chen, Angew. Chem., Int. Ed., 2020, 20, 7881–7890 CrossRef PubMed.
  30. A. H. Westlie and E. Y.-X. Chen, Macromolecules, 2020, 53, 9906–9915 CrossRef CAS.
  31. H. Westlie, E. C. Quinn, C. R. Parker and E. Y.-X. Chen, Prog. Polym. Sci., 2022, 134, 101608 CrossRef.
  32. Fluorinated Polymers: Synthesis and Applications, Organofluorine Compounds, ed. R. Schwarz and C. E. Diesendruck, John Wiley & Sons, Ltd., 2022,  DOI:10.1002/9780470682531.pat1008.
  33. B. Ameduri and H. Sawada, Fluorinated Polymers, The Royal Society of Chemistry, Cambridge, 2017, vol. 2, p. 372 Search PubMed.
  34. G. Hougham, P. E. Cassidy, K. Johns and T. Davidson, Fluoropolymers 2: Properties, Kluwer/Plenum, New York, 1999 Search PubMed.
  35. J. Scheirs, in Modern Fluoropolymers, ed. J. Scheirs, John Wiley & Sons, New York, 1997, ch. 24, pp. 435–486 Search PubMed.
  36. B. W. Thuronyi, T. M. Privalsky and M. C. Y. Chang, Angew. Chem., Int. Ed., 2017, 56, 13637–13640 CrossRef CAS PubMed.
  37. Y. Takagi, R. Yasuda, A. Maehara and T. Yamane, Eur. Polym. J., 2004, 40, 1551–1557 CrossRef CAS.
  38. O. Kim, R. A. Gross, W. J. Hammar and R. A. Newmark, Macromolecules, 1996, 29, 4572–4581 CrossRef CAS.
  39. M. Le Gal, A. Rios De Anda, L. Michely, C. S. Colin, E. Renard and V. Langlois, Biomacromolecules, 2021, 22, 4510–4520 CrossRef CAS PubMed.
  40. G. Foli, M. D. Esposti, M. Toselli, D. Morselli and P. Fabbri, Analyst, 2019, 144, 2087–2096 RSC.
  41. J. W. Kramer and G. W. Coates, Tetrahedron, 2008, 64, 6973–6978 CrossRef CAS PubMed.
  42. T. Xu, G. Yang, C. Liu and X. Lu, Macromolecules, 2017, 50, 515–522 CrossRef CAS.
  43. J. W. Kramer, E. B. Lobkovsky and G. W. Coates, Org. Lett., 2006, 8, 3709–3712 CrossRef CAS PubMed.
  44. J. Wöltinger, J. E. Bäckvall and Á. Zsigmond, Chem. – Eur. J., 1999, 5, 1460–1467 CrossRef.
  45. W. F. Edgell and J. Lyford IV, Inorg. Chem., 1970, 9, 1932–1933 CrossRef CAS.
  46. M. Bochmann, G. Bwembya and K. J. Webb, Inorg. Synth., 1997, 31, 19–24 CrossRef CAS.
  47. B. M. Chamberlain, M. Cheng, D. R. Moore, T. M. Ovitt, E. B. Lobkovsky and G. W. Coates, J. Am. Chem. Soc., 2001, 123, 3229–3238 CrossRef CAS PubMed.
  48. S. E. Schaus, B. D. Brandes, J. F. Larrow, M. Tokunaga, K. B. Hansen, A. E. Gould, M. E. Furrow and E. N. Jacobsen, J. Am. Chem. Soc., 2002, 124, 1307–1315 CrossRef CAS PubMed.
  49. J. A. Schmidt, E. B. Lobkovsky and G. W. Coates, J. Am. Chem. Soc., 2005, 127, 11426–11435 CrossRef CAS PubMed.
  50. The value of Pr = 0.09 for the isotactic-enriched polymer evidences stereoerrors that arise from the incomplete enantiopurity of the initial monomer (S)-BPLCH2OCF2CHF2 which was prepared from the corresponding enantio-enriched epoxide GCH2OCF2CHF2, obtained itself by hydrolytic kinetic resolution of the racemic compound (refer to the ESI, Scheme 1); the latter epoxide was clearly much enantio-enriched but, despite our efforts, chiral GLC did not enable us to achieve complete separation of the peaks of two enantiomers and thus to provide an accurate value of % ee; the enantiopurity of epoxide GCH2OCF2CHF2, and therefore of monomer (S)-BPLCH2OCF2CHF2, is estimated to >90% ee.
  51. H. Li, J. Ollivier, S. M. Guillaume and J.-F. Carpentier, Angew. Chem., Int. Ed., 2022, 61, e202202386 CrossRef CAS PubMed.
  52. R. M. Shakaroun, H. Li, P. Jéhan, A. Alaaeddine, J.-F. Carpentier and S. M. Guillaume, Polym. Chem., 2020, 11, 2640–2652 RSC.
  53. N. Ajellal, M. Bouyahyi, A. Amgoune, C. M. Thomas, A. Bondon, I. Pillin, Y. Grohens and J.-F. Carpentier, Macromolecules, 2009, 42, 987–993 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3py01430d

This journal is © The Royal Society of Chemistry 2024