Next Article in Journal
Co3O4-rGO—Synthesis, Characterization, and Evaluation of Photocatalytic Activities
Next Article in Special Issue
Molybdenum-Catalyzed Enantioselective Ring-Closing Metathesis/Kinetic Resolution of Racemic Planar-Chiral 1,1′-Diallylferrocenes
Previous Article in Journal
Mechanistic and Compositional Aspects of Industrial Catalysts for Selective CO2 Hydrogenation Processes
Previous Article in Special Issue
Exploring Long Range para-Phenyl Effects in Unsymmetrically Fused bis(imino)pyridine-Cobalt Ethylene Polymerization Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Acyclic Diene Metathesis (ADMET) Polymerization for the Synthesis of Chemically Recyclable Bio-Based Aliphatic Polyesters

by
Kotohiro Nomura
* and
Xiuxiu Wang
Department of Chemistry, Tokyo Metropolitan University, Hachioji, Tokyo 1920397, Japan
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(2), 97; https://doi.org/10.3390/catal14020097
Submission received: 5 January 2024 / Revised: 21 January 2024 / Accepted: 22 January 2024 / Published: 24 January 2024

Abstract

:
The recent developments of the synthesis of bio-based long-chain aliphatic polyesters by the acyclic diene metathesis (ADMET) polymerization of α,ω-dienes, derived from plant oils and bio-based chemicals, like bis(10-undecenoate) with isosorbide, using ruthenium-carbene catalysts are reviewed. The development of subsequent (one-pot) tandem hydrogenation produced saturated polyesters under mild conditions. The polymerizations under bulk (without solvent, 80–90 °C) or in ionic liquids (50 °C) under vacuum conditions enabled the synthesis of high molar mass polymers (Mn > 30,000 g/mol). The polymerization performed by the molybdenum-alkylidene catalyst afforded the highest-molecular-weight polyesters (44,000–49,400 g/mol, in toluene at 25 °C) exhibiting promising tensile properties (strength and elongation at break) compared to polyethylene and polypropylene. Depolymerizations of these polyesters, including closed-loop chemical recycling, were also demonstrated. Catalyst developments (more active, under mild conditions) play a key role in the efficient synthesis of these materials.

Graphical Abstract

1. Introduction

The development of sustainable polymers from renewable feedstocks attracts considerable attention from the viewpoints of the circular economy as well as green sustainable chemistry. Hydrocarbon-rich molecular biomasses, such as vegetable oils (castor, coconut, linseed, olive, palm, soybean, sunflower, etc.) presented as triglycerides with fatty acids, or fatty acid esters (FAEs) are naturally abundant and are recognized as low-cost molecular biomass products [1,2,3,4,5,6,7,8,9,10,11]. A study on bio-based advanced polyesters (exhibiting tunable mechanical properties and biodegradability), in particular, long-chain aliphatic polyesters (LCAPEs), are semicrystalline materials considered as a promising alternative of polyethylene [6,8]. The melting temperatures (Tm values) of polyesters are generally influenced by the methylene length (and the direction of dipoles called the odd–even effect) [6,12,13,14]; the placement of longer methylene units should be effective for the obtainment of polyesters without softening them at elevated temperatures. It has been considered that the precise polymerization technique provides a new strategy and methodology for the design of macromolecular architectures.
Two condensation polymerization approaches—(i) condensation polymerization by transesterification (dicarboxylic acid and diol, etc.) and (ii) acyclic diene metathesis (ADMET) polymerization (nonconjugated α,ω-dienes)—and subsequent hydrogenation (Scheme 1) are considered for the synthesis from FAEs [6,8]. The ring-opening polymerization (ROP) approach from cyclic monomers can also be considered, but the method has a limited monomer scope; the method also faces the difficulty of catalysts enabling the synthesis of high molar mass polymers [15,16]. Studies on alternative approaches to polymers are also under investigation [17,18,19,20]. Moreover, the recent progress in the development of olefin metathesis catalysts for the conversion of plant oils (FAEs) is well known [21,22,23,24].
The conventional condensation polymerization approach through transesterification (ester bond exchange) requires high temperatures with the efficient removal of by-products (such as alcohols and water) to obtain high molar mass polymers with a high degree of polymerization (DPn). For example, the synthesis of poly(ethylene terephthalate) from terephthalic acid (which must be purified) with excess ethylene glycol requires high temperatures up to 290 °C under a reduced pressure [25]. This method, however, seems to be difficult to apply for the synthesis of LCAPEs due to the difficulty of removing diols with high boiling points (e.g., 1,12-dodecanediol, 189 °C/12 Torr; 1,16-hexadecane diol 197–199 °C/3 mmHg). Moreover, precise stoichiometric control (hydroxy and carboxylic groups) is needed for this purpose [6,26,27,28]; polymerization with the precise stoichiometric ratios of diols (algae oil) and diesters (C17 and C19) is required to create high molar mass polymers (Mn = 4.0 × 104) possessing a Tm value of 99 °C [28].
The synthesis of bio-based aliphatic polyesters by adopting the ADMET polymerization [29,30,31] approach, especially using commercially available (Grubbs-type) ruthenium carbene catalysts, has been explored by many researchers, since the reactions do not require severe conditions for condensation polymerization through transesterification (described above). In this mini review, reports concerning the synthesis of bio-based aliphatic polyesters by the ADMET approach are summarized.

2. Synthesis of Bio-Based Aliphatic Polyesters by ADMET Polymerization

2.1. Synthesis of Aliphatic Polyesters by ADMET Polymerization and Hydrogenation

Reports on the synthesis of bio-based polyesters by ADMET polymerization, especially using commercially available (Grubbs-type) ruthenium carbene catalysts RuCl2(PCy3)2(CHPh) (G1; Cy = cyclohexyl), RuCl2(PCy3)(IMesH2)(CHPh) (G2; IMesH2 = 1,3-bis(2,4,6-trimethylphenyl)imidazolin-2-ylidene), and RuCl2(IMesH2)(CH-2-OiPr-C6H4) (HG2), shown in Scheme 2, are well known. These ruthenium catalysts have been employed [8] because these complexes can be readily available and do not require treatment with the strict Schlenk technique due to their insensitivities toward water and oxygen (better functional group tolerance) [32,33,34,35]. More recently, the example using a molybdenum-alkylidene catalyst (Mo cat.) [36,37,38], shown below, also demonstrates the synthesis of high molar mass polymers that exhibit good tensile properties [39].
The synthesis of a bio-based polyester, expressed as PE1, by the ADMET polymerization of undec-10-en-1-yl undec-10-enoate (M1), prepared by the reaction of 10-undecenoic acid with 10-undecenol (derived from castor oil) was reported by the group of Meier in 2008 [40]. The resultant PE1 synthesized by G2 (0.5 or 1.0 mol%, 80 °C, 24 h, Scheme 3) possessed a rather high molecular weight (Mn = 22,000, 26,500), and the Mn values were controlled by the addition of terminal olefins, such as methyl 10-undecenoate and stearyl acrylate [40]. In contrast, the group reported that the polymerization of bis(undec-10-enoate) with isosorbide (M2, Scheme 3) conducted at 70–100 °C under bulk conditions yielded rather low-molecular-weight polymers (PE2, Table 1) [41], whereas the Mn values seemed to improve when the polymerizations were conducted at high temperatures and/or under nitrogen-purge conditions (for the removal of by-produced ethylene). This was probably due to the catalyst decomposition caused by conducting the reaction at 70–100 °C [42,43,44,45,46,47], because these ruthenium catalysts are known to decompose under these conditions to produce ruthenium-hydride species [44] and/or nanoparticles [46], which induce olefin isomerization and/or certain side reactions by the formed radicals [42,43,44,45,46,47,48]. G2 showed a more significant degree of olefin isomerization compared to G1 and a higher percentage of isomerization (estimated by GC-MS, after treating the mixture with MeOH-H2SO4 under reflux conditions) [41]. Later, the degree of isomerization was extensively suppressed when the polymerizations were conducted in the presence of benzoquinone [48].
The ADMET polymerization of M1 by G1 under high-vacuum conditions for two days produced PE1 (Mn = 28,000, Mw/Mn = 1.9) and a subsequent hydrogenation step (Pd/C, 50 bar H2, 60 ºC) produced a saturated polyester (HPE1, PE-20.20, Scheme 4) [49]. The Tm value (103 °C) achieved was somewhat low compared to the HPE1 prepared by the condensation polymerization of 1,20-eicosanedioic acid with eicosane-1,20-diol (Tm = 108 °C) to form ‘regio-regular’ ester groups, C(O)-O, aligned with the polymer chain (Scheme 4). It was thus suggested that the microstructural control directly affected the thermal property, as described above [6,14]. ADMET polymerizations of α,ω-dienes with different methylene chain lengths, di(icos-19-en-1-yl)tricosanedioate (M3) and di(tricos-22-en-1-yl)tricosanedioate (M4), using G1 and the subsequent olefin hydrogenation conducted by Ru(CHOEt)Cl2(PCy)2 (40 bar H2, 70 °C, 2 d), prepared from G1, yielded the corresponding values of PE-38.23 (HPE3) and PE-44.23 (HPE4), respectively (Scheme 4) [50]. The polycondensation of 1,26-hexacosanedioate, prepared by the cross-metathesis of erucic acid, with the corresponding diol (produced by a reduction with LiAlH4) with Ti(OBu)4 also produced the corresponding polyester (HPE5, PE-26.26, Tm = 114 °C) [51]. The thermal properties (Tm values) of the resultant LCAPEs with different methylene lengths, prepared by ADMET [50] and polycondensation [51,52] approaches, revealed that the Tm values achieved a constant value (Figure 1a) [50]. A linear relationship between the Tm values and the number of ester groups in 1000 carbon atoms was observed (Figure 1b) [50]. Polyesters PE-26.26, PE-12.26 and PE-4.26 [51], and PE-18,18 [53] were also prepared by polycondensation.
The one-pot synthetic method used for the bio-based aliphatic polyesters by ADMET polymerization and subsequent hydrogenation was demonstrated (Scheme 5) [54]. The polymerization of bis(undec-10-enoate)s with isosorbide (M2), isomannide (M6), 1,3-propanediol (M7), and 1,4-cyclohexanedimethanol (M8), derived from castor oil and glucose in chloroform by G2 or HG2 under a reduced pressure at 50 °C produced unsaturated polymers (expressed as PE2 and PE6PE8, respectively) [54]. The Mn values in the produced polymers (Mn = 11,900–15,900) were somewhat higher than those reported previously (Mn = 4400–8400), conducted at 70–100 °C [41], and the Mn values did not change, even under rather scaled-up conditions [54]. One reason for the obtainment of high-molecular-weight product could be that the degree of the catalyst decomposition was significantly suppressed by conducting the polymerization at 50 °C (and the polymerization was conducted under a continuously reduced pressure) [54].
As described above (Scheme 4) and below [55], conventional olefin hydrogenation requires a high hydrogen pressure and high temperature after the isolation of unsaturated polyesters after ADMET polymerization [49,50,55]. In contrast, one-pot hydrogenation under rather mild conditions (1.0 MPa, 50 °C, 3 h) was demonstrated following the addition of a small amount of Al2O3 (ca. 1 wt%) to the reaction mixture (Scheme 5). The completion of the olefin hydrogenation was confirmed by DSC thermograms (uniform compositions) due to the difficulty (accuracy of the integration of olefinic protons) of obtaining the 1H NMR spectra. No significant differences in the Mn and Mw/Mn values were observed before/after hydrogenation [54].
As shown in Figure 1b, the melting temperatures (Tm values) of the polyesters are influenced by the methylene unit number (n). As shown in Scheme 6, the copolymerization of M1 with undeca-1,10-diene (UDD) followed by olefin hydrogenation (H2 40 bar, 110 °C, 2 d) produced various LCAPEs with different chain lengths (ranging from 0.9 to 52.6 ester groups per 1000 carbon atoms), expressed as H2-poly(M1-co-UDD) [55]. A linear correlation of the melting temperatures (Tm values) with the average number of ester groups per methylene unit was thus demonstrated, whereas the ester group was incorporated in a random manner. A similar trend was observed in the copolymerization of M2 with 1,9-decadiene (DD) and the subsequent one-pot hydrogenation [56]. The saturated polymers possessed Tm values in the range of 71.7–107.6 °C, depending on the molar ratios of M2 and DD.
The polymerization of bis(undec-10-enoate)s with D-xylose (1,2-O-isopropylidene-α-D-xylofuranose, M9c), and D-mannose (M10) by G2 was studied under a dynamic-vacuum (0.1 mbar) condition without solvent (bulk) conditions (60–90 °C, 20 h, Scheme 7) [57]. The molecular weights of the resultant polymers (PE9c, PE10) were affected by the polymerization temperature employed and the monomer/Ru molar ratios. Conducting the polymerization at 90 ºC under a low Ru concentration (0.1 mol%) seemed to be the optimized condition (PE9c: Ru, Mn = 7.14–7.16 × 104, Mw/Mn = 2.2–2.3, PE10: Mn = 3.24 × 104, Mw/Mn = 2.4) [49]. Due to the fact that the polymerization was conducted without a solvent, the PDI (Mw/Mn) values were rather high due to the difficulty pf controlling the stirring [57]. Later, the polymerizations of D-xylose diester analogs with different methylene lengths (M9, x = 0, 2, 8, Scheme 7) and the corresponding diether analogs (M11) were explored [58]. The Mn values of the resultant polymers decreased upon decreasing the methylene length, and the monomers did not possess a methylene spacer [58]. Some polymerization runs failed due to precipitation or the difficulty of performing isolations [58]. The resultant unsaturated polymers were amorphous, except PE11a, and both glass transition temperatures (Tg) increased after reducing the olefinic double bonds by treating them with p-toluenesulfonyl hydrazide as a reducing agent; most of the resultant saturated polymers (HPE9 and HPE11) were amorphous, except HPE9a and HPE11a derived from the castor oil (10-undecenoate), suggesting that the placement of the methylene spacer was important (as shown in Figure 1a and Figure 2) [58]. The resultant hydrogenated polymer films, especially the HPE11a-oriented film, exhibited a good tensile strength (43 MPa) with an elongation at a break of 155%; but, the hot-press film showed a much weaker tensile strength (7.8 MPa) with and improved elongation at the break (667%) [58].
The syntheses of polyesters containing vanillin (PE12) [59] afforded high-molecular-weight PE12 (Mn = 10,000, Mw/Mn = 1.6) possessing a Tg value of 4 °C (Scheme 7), whereas the polymerization of 4-allyl-2-methoxyphenyl 10-undecenoate (M13) by G2 produced amorphous high molar mass polymers with low PDIs (Mw/Mn) with Tg at −9.6 °C [60]. The ADMET polymerization of M13 in the presence of 5-formylbenzene-1,2,3-triyl tris(undec-10-enoate) produced rather high molar mass network polymers [60].
The polymerization of trehalose bis(10-undecenoate) (M14) by HG2 (4.0 mol%) in THF at 45 ºC for 24 h (Scheme 8) produced semicrystalline polymers (PE14) possessing high molecular weights with unimodal molecular-weight distributions (Mn = 13,200, Mw/Mn = 2.1) with higher Tm values (156 °C) [61]. Both the molecular weights and melting temperatures (Tm values) of the resulting copolyesters with undec-10-en-1-yl undec-10-enoate (M1) decreased with the increase in the percentage of M1 [61].
The polymerization of bis(10-undecenoate)s with isosorbide (M2) and glucarodilactone (M15) and copolymerizations with different molar ratios were conducted in the presence of methyl-10-undecenoate (MU, 1.0 mol%) by using G2 (1.0 mol%) at 80 °C for 16 h under a reduced pressure (Scheme 8) [62]. MU was employed as the monofunctional chain stopper (chain transfer reagent by the placement of the MU unit as the end group) [62]. The resultant polymers possessed high molecular weights with unimodal molecular-weight distributions. Copolymerizations with bis(hydroxymethylfuran) undecenoate (M16) were conducted [63]. The resultant PE2 possessed a low Tg value (−10 °C) compared to PE15 (Tg = 32 °C), and the homopolymers, PE2 and PE15, were brittle materials, whereas the copolyesters were rubbery materials possessing better tensile properties, an elastic behavior, as well as shape memory properties.
The copolymerizations of α,ω-dienes (linear M17 and n-hexyl branched M18), derived from castor oil and vernonia oil, by G1 at 85 °C, produced LCAPEs containing branches at a certain percentage (after subsequent hydrogenation by Pd/C, Scheme 8) [64]. These polymers were considered as LLDPE (linear low-density polyethylene) and VLDPEs (very low-density polyethylene) mimics. However, their DSC thermograms presented multiple melting temperatures, suggesting the compositions of the resultant copolymers were not uniform [64].

2.2. Synthesis of High Molecular-Weight Polymers Exhibiting Tensile Properties beyond Polyethylene and Polypropylene

There are many reports on the synthesis of bio-based aliphatic polyesters by the ADMET polymerization of α,ω-diene monomers containing carbohydrate units (such as M2, M6, M9, M10, M14, and M15) using ruthenium catalysts [41,54,57,58,61,62,63]; however, the reports on the synthesis of high-molecular-weight polymers (ca. Mn ≥ 30,000 considered for their better mechanical properties, such as their films, shown below) are limited to date (Scheme 9) [58,62,63]. Catalyst decomposition was highly considered when metathesis polymerizations (reactions) were conducted at high temperatures (70–100 °C) and the subsequent isomerization and/or undesired side reaction caused by formed radicals were known [42,43,44,45,46,47]. The catalyst decomposition also caused the difficulty of separating the metal (present as ruthenium metal particles) from the resultant polymers, and this is often observed in metathesis polymerization chemistry, especially when using ruthenium catalysts. Moreover, the reported synthetic methods were conducted under direct-vacuum and bulk conditions without a solvent [58,62,63]; the methods thus presented the difficulty of stirring high-viscosity products [58] and was applicable to for the synthesis of amorphous or semicrystalline materials with Tm values below 90 °C. Therefore, the development of the methods for the solution polymerization in the presence of appropriate solvent seemed to be better in terms of the process control (by lowering the viscosity of the reaction mixture under rather mild conditions to avoid catalyst decomposition) and the wide monomer scope.
ADMET polymerization is a condensation polymerization method that by-produces small molecules (ethylenes), and the removal is quite effective for the obtainment of high molar mass polymers under certain equilibrated conditions. Conducting the polymerization under continuous dynamic-vacuum and bulk conditions [58,62,63] is thus effective for this purpose. A consideration of these points shows that ionic liquids (ILs) can be considered as ideal solvents, not only due to their absent (or extremely low) vapor pressure and ability to provide homogeneous conditions due to their good miscibility with polymers, organic compounds, and metal catalysts, but also due to their high stability ranging from −30 to >300 °C [65,66,67,68,69,70,71,72]. Although olefin metathesis reactions in ILs are known, the reported examples for ADMET polymerizations are still limited [73,74,75,76,77,78].
More recently, the synthesis of high-molecular-weight polymers (PE2, Mn = 32,200–39,200) was demonstrated in the polymerization of α,ω-diene monomer (M2, dianhydro-D-glucityl bis(undec-10-enoate)) using the HG2 catalyst in ionic liquids (ILs) under continuous-vacuum conditions at 50 °C (Scheme 10) [79]. The Mn values were apparently higher than those reported previously (Mn = 5600–14,700) [41,54]. 1-n-Butyl-3-methyl imidazolium hexafluorophosphate, [Bmim]PF6, and 1-n-hexyl-3-methyl imidazolium bis(trifluoromethanesulfonyl)imide, [Hmim]TFSI, were found to be effective as solvents among a series of imidazolium salts and pyridinium salts. As summarized in Table 2, the method was also effective for the syntheses of high molar mass polymers containing isomannide (PE6), 1,4-cyclohexanedimethanol (PE8), and 1,4-butanediol (PE7) units as the diol segments used instead of isosorbide (PE2); the Mn values did not decrease, even under the scale-up conditions (300 mg → 1.0 g scale) [79]. The tandem hydrogenation of the resultant unsaturated polymers (PE2) in the [Bmim]PF6–toluene biphasic system upon the addition of Al2O3 (H2 1.0 MPa at 50 °C) produced the corresponding saturated polymers (HPE2).
As described above, the polymerization of M2 conducted in ILs with the continuous removal of by-produced ethylene afforded high molar mass polymers (Scheme 10) [79], whereas the polymerizations conducted in toluene or CHCl3 (even under optimized conditions with the careful removal of ethylene) afforded polymers with Mn values up to 15,000 [54]. The development of the method without using (expensive) ILs is favorable from a practical point of view.
We more recently demonstrated that the synthesis of higher molar-mass polymers (Mn = 44,000–49,400 g/mol) could be achieved by polymerization in toluene using the molybdenum-alkylidene catalyst, Mo(CHCMe2Ph)(2,6-Me2C6H3)[OC(CH3)(CF3)2] (Mo cat., Scheme 11) [39]. As summarized by the results in Table 3, the Mn values are affected by the M2/Mo molar ratios and amount of toluene used. As observed in the conventional ADMET polymerization, polymerization with low catalyst loading under high initial monomer conditions was suited to the condensation polymerization; it seemed that the Mn value in PE2 increased when the reaction scale was increased (90.5 (43.5 mg) → 261 μmol (543 mg)) with the increase in the initial monomer concentration (by varying the amount of toluene) [39]. This method is applicable to the other monomers (M6, M19). Olefinic double bonds in the resultant polymers were hydrogenated by using a rhodium catalyst under mild conditions (1.0 MPa, 50 °C), and no significant changes in the Mn or PDI values of the polymers after hydrogenation were observed.
It should be noted that both the tensile strength (stress) and elongation at break (strain) in the prepared polymer films of HPE2 increased remarkably upon increasing the Mn value (Figure 3) [39]; a fairly good linear correlation was observed between the stress and strain; and the HPE2 sample with the highest Mn value (Mn = 48,200) exhibited a tensile strength of 39.7 MPa along with an elongation at break of 436%. The value was not only higher than PE-18,18, prepared with C18 dimethyl dicarboxylate and the corresponding diol by condensation polymerization [9], but also poly(lactic acid) (PLA), poly(ethylene terephthalate) (PET), high-density polyethylene (HDPE), low-density polyethylene (LDPE), and polypropylene (PP) [39,80]. The PE2 sample before hydrogenation showed a higher strain (elongation at break) with less stress (tensile strength) compared to HPE2, and the isomannide-based HPE6 showed a similar tensile property to the isosorbide-based HPE2 [39]. The importance of the development of a synthetic method for the synthesis of high molar mass polymers by ADMET polymerization was thus demonstrated [39].

2.3. Chemical Recycling of Polyesters

PE18,18, prepared by the condensation polymerization of 1,18-octadeca dicarboxylic acid with 1,18-octadecanediol, was treated with MeOH (150 °C, 12 h) to produce a solid mixture consisting of dicarboxylic acid and diol after MeOH removal. The resultant solid was used for the subsequent condensation polymerization with Ti(OnBu)4 to yield recycled PE18,18 with a high molecular weight (Mn = 79,000, Mw/Mn = 1.9, Scheme 12) [9]. Moreover, the treatment of polycarbonate (PC18, Mn = 90,000, Mw/Mn = 2.7), prepared by the condensation polymerization of 1,18-octadecane diol with diethyl carbonate (DEC) in the presence of LiH, with a 10 wt% KOH ethanol solution (at 120 °C, 24 h) exclusively produced 1,18-octadecanediol (yield: 98%, and purity: 99% after recrystallization from MeOH). The subsequent polycondensation with DEC produced recycled PC18 without a reduction in the Mn value (Mn = 70,000, Mw/Mn = 3.4), which exhibited similar properties as the fresh sample [9]. These results indicate the possibility of closed-loop chemical recycling.
Although the conventional method for the depolymerization of polyester requires excess acid or base materials (or MeOH under high temperatures), more recently, exclusive acid- and base-free chemical conversions of polyesters (poly(ethylene adipate) (PEA), poly(butylene adipate) (PBA), poly(ethylene terephthalate) (PET), and poly(butylene terephthalate) (PBT)) into the corresponding monomers (diethyl adipate, diethyl terephthalate, ethylene glycol, and 1,4-butane diol) by transesterification with ethanol using the Cp’TiCl3 (Cp’ = Cp, Cp*) catalyst were demonstrated [81,82]. The depolymerizations proceeded the completed conversions (>99%) of PET and PBT to afford diethyl terephthalate and ethylene glycol or 1,4-butanediol exclusively (selectivity > 99%, 150–170 °C, Ti 1.0 or 2.0 mol%) [82]. The resultant reaction mixture after the depolymerization of PBA with ethanol by the CpTiCl3 catalyst (1.0 mol%, 150 °C, 3 h), consisting of diethyl adipate and 1,4-butanediol, was heated at 150 °C in vacuo for 24 h to afford high-molecular-weight recycled PBA with a unimodal molecular-weight distribution (Mn = 11,800, Mw/Mn = 1.6, Scheme 12), demonstrating the possibility of one-pot (acid and base free) closed-loop chemical recycling [82]. The method can also be applicable to bio-based aliphatic polyesters; the reaction of HPE2 with ethanol by CpTiCl3 afforded the corresponding dicarboxylic acid and isosorbide products exclusively [79]. Since the depolymerization (transesterification) method by titanium catalysts can be applied to various polyesters, including bio-based ones [79,81,82], the importance of the basic concept of the one-pot method can be emphasized.
Scheme 12. Closed-loop chemical recycling of polyesters [9,82].
Scheme 12. Closed-loop chemical recycling of polyesters [9,82].
Catalysts 14 00097 sch012

3. Concluding Remarks

This review summarizes the recent developments for the synthesis of bio-based LCAPEs by the acyclic diene metathesis (ADMET) polymerization of α,ω-dienes, derived from plant oils and bio-based chemicals (carbohydrates and their derivatives) in the presence of ruthenium-carbene catalysts (G1, G2, HG2; Scheme 2). The development of subsequent (one-pot) tandem hydrogenation afforded saturated polyesters under mild conditions. Reported examples for the synthesis of high-molecular-weight polymers are still limited; polymerizations under bulk conditions (without solvent, 80–90 °C) or in ionic liquids (50 °C) under vacuum conditions enabled the synthesis of high molar mass polymers (Mn > 30,000) that exhibited better mechanical properties than films. However, high-temperature polymerizations (at 70–100 °C) created the possibility of catalyst decompositions. The polymerizations using the molybdenum-alkylidene catalyst afforded the highest molecular-weight polyesters (44,000–49,400 g/mol) even in toluene at 25 °C. Hydrogenated polyester films, prepared by the polymerization of bis(10-undecenoate) with isosorbide and the subsequent hydrogenation phase, exhibited promising tensile properties (strength and elongation at break) beyond polyethylene and polypropylene. The significant effects of molecular weight on the tensile properties were demonstrated, clearly indicating the importance of the synthesis of high molar mass polymers to produce better materials properties. The reported procedures for the closed-loop chemical recycling of polyesters by depolymerization and re-polymerization methods were also introduced. The depolymerization of poly(butylene adipate) (PBA) with ethanol using the CpTiCl3 catalyst afforded diethyl adipate and 1,4-butandiol exclusively, and the subsequent polycondensation produced PBA without a loss of the Mn value. Catalyst developments (more active, under mild conditions) play a key role in efficient synthesis practices.

Author Contributions

Conceptualization, project administration, funding acquisition: K.N.; writing—original draft preparation: X.W. and K.N.; writing—review and editing: K.N. All authors have read and agreed to the published version of the manuscript.

Funding

This project was partly supported by JST-CREST (Grant Number: JPMJCR21L5), JST SICORP (Grant Number: JPMJSC19E2), Japan, and the Tokyo Metropolitan Government Advanced Research (Grant Number: R2-1). XW thanks the Tokyo Metropolitan government (Tokyo Human Resources Fund for City Diplomacy) for the pre-doctoral fellowship.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

K.N. would like to express his heartfelt thanks to the laboratory members who contributed as co-authors in the cited references for their wonderful contributions.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Gandini, A. Polymers from renewable resources: A Challenge for the future of macromolecular materials. Macromolecules 2008, 41, 9491–9504. [Google Scholar] [CrossRef]
  2. Meier, M.A.R.; Metzger, J.O.; Schubert, U.S. Plant oil renewable resources as green alternatives in polymer science. Chem. Soc. Rev. 2007, 36, 1788–1802. [Google Scholar] [CrossRef]
  3. Xia, Y.; Larock, R.C. Vegetable oil-based polymeric materials: Synthesis, properties, and applications. Green Chem. 2010, 12, 1893–1909. [Google Scholar] [CrossRef]
  4. Biermann, U.; Bornscheuer, U.; Meier, M.A.R.; Metzger, J.O.; Schäfer, H.J. Oils and fats as renewable raw materials in chemistry. Angew. Chem. Int. Ed. 2011, 50, 3854–3871. [Google Scholar] [CrossRef]
  5. Hillmyer, M.A.; Tolman, W.B. Aliphatic polyester block polymers: Renewable, degradable, and sustainable. Acc. Chem. Res. 2014, 47, 2390–2396. [Google Scholar] [CrossRef]
  6. Stempfle, F.; Ortmann, P.; Mecking, S. Long-chain aliphatic polymers to bridge the gap between semicrystalline polyolefins and traditional polycondensates. Chem. Rev. 2016, 116, 4597–4641. [Google Scholar] [CrossRef]
  7. Monomers and polymers from chemically modified plant oils and their fatty acids. In Polymers from Plant Oils, 2nd ed.; Gandini, A.; Lacerda, T.M. (Eds.) John Wiley & Sons, Inc.: Hoboken, NJ, USA; Scrivener Publishing LLC: Beverly, MA, USA, 2019; pp. 33–82. [Google Scholar]
  8. Nomura, K.; Awang, N.W.B. Synthesis of bio-based aliphatic polyesters from plant oils by efficient molecular catalysis: A selected survey from recent reports. ACS Sustain. Chem. Eng. 2021, 9, 5486–5505. [Google Scholar] [CrossRef]
  9. Häußler, M.; Eck, M.; Rothauer, D.; Mecking, S. Closed-loop recycling of polyethylene-like materials. Nature 2021, 590, 423–427. [Google Scholar] [CrossRef]
  10. Biermann, U.; Bornscheuer, U.T.; Feussner, I.; Meier, M.A.R.; Metzger, J.O. Fatty acids and their derivatives as renewable platform molecules for the chemical industry. Angew. Chem. Int. Ed. 2021, 60, 20144–20165. [Google Scholar] [CrossRef]
  11. Worch, J.C.; Dove, A.P. 100th Anniversary of macromolecular science viewpoint: Toward catalytic chemical recycling of waste (and future) plastics. ACS Macro Lett. 2020, 9, 1494–1506. [Google Scholar] [CrossRef]
  12. Ishioka, R.; Kitakuni, E.; Ichikawa, Y. Aliphatic polyesters: “Bionolle”. In Biopolymers; Steinbüchel, A., Doi, Y., Eds.; Wiley-VCH: Weinheim, Germany, 2002; Volume 4, pp. 275–297. [Google Scholar]
  13. Korshak, W.V.; Vinogradova, S.V. Polyesters of 1,20-eicosanediol. Bull. Acad. Sci. USSR Div. Chem. Sci. 1953, 2, 995–998. [Google Scholar] [CrossRef]
  14. Bunn, C.W. The melting points of chain polymers. J. Polym. Sci. 1955, 16, 323–343. [Google Scholar] [CrossRef]
  15. Gallagher, J.J.; Hillmyer, M.A.; Reineke, T.M. Isosorbide-based polymethacrylates. ACS Sustain. Chem. Eng. 2015, 3, 662–667. [Google Scholar] [CrossRef]
  16. Wang, J.; Mahmud, S.; Zhang, X.; Zhu, J.; Shen, Z.; Liu, X. Bio-based amorphous polyesters with high Tg: Trade-off between rigid and flexible cyclic diols. ACS Sustain. Chem. Eng. 2019, 7, 6401–6411. [Google Scholar] [CrossRef]
  17. Walther, G.; Deutsch, J.; Martin, A.; Baumann, F.; Fridag, D.; Franke, R.; Kçckritz, A. α,ω-Functionalized C19 Monomers. ChemSusChem 2011, 4, 1052–1054. [Google Scholar] [CrossRef]
  18. Witt, T.; Häußler, M.; Kulpa, S.; Mecking, S. Chain multiplication of fatty acids to precise telechelic polyethylene. Angew. Chem. Int. Ed. 2017, 56, 7589–7594. [Google Scholar] [CrossRef]
  19. Herrmann, N.; Köhnke, K.; Seidensticker, T. Selective product crystallization for concurrent product separation and catalyst recycling in the isomerizing methoxycarbonylation of methyl oleate. ACS Sustain. Chem. Eng. 2020, 8, 10633–10638. [Google Scholar] [CrossRef]
  20. Pandey, S.; Rajput, B.S.; Chikkali, S.H. Refining plant oils and sugars to platform chemicals, monomers, and polymers. Green Chem. 2021, 23, 4255–4295. [Google Scholar] [CrossRef]
  21. Wyrębek, P.; Małecki, P.; Sytniczuk, A.; Kośnik, W.; Gawin, A.; Kostrzewa, J.; Kajetanowicz, A.; Grela, K. Looking for the noncyclic(amino)(alkyl)carbene ruthenium catalyst for ethenolysis of cthyl oleate: Selectivity is on target. ACS Omega 2018, 3, 18481–18488. [Google Scholar] [CrossRef]
  22. Sytniczuk, A.; Kajetanowicz, A.; Grela, K. “Inverted” cyclic(alkyl)(amino)carbene ligands allow olefin metathesis with ethylene at parts-per-billion catalyst loading. Chem. Catal. 2023, 3, 100713. [Google Scholar] [CrossRef]
  23. Gawin, R.; Tracz, A.; Krajczy, P.; Kozakiewicz-Piekarz, A.; Martínez, J.P.; Trzaskowski, B. Inhibition of the decomposition pathways of ruthenium olefin metathesis catalysts: Development of highly efficient catalysts for ethenolysis. J. Am. Chem. Soc. 2023, 145, 25010–25021. [Google Scholar] [CrossRef]
  24. Del Vecchio, A.; Talcik, J.; Colombel-Rouen, S.; Lorkowski, J.; Serrato, M.R.; Roisnel, T.; Vanthuyne, N.; Bertrand, G.; Jazzar, R.; Mauduit, M. Highly robust and efficient Blechert-type cyclic(alkyl)(amino)carbene ruthenium complexes for olefin metathesis. ACS Catal. 2023, 13, 6195–6206. [Google Scholar] [CrossRef]
  25. Rogers, M.E.; Long, T.E. Synthetic Methods in Step-Growth Polymers; Wiley-Interscience: Hoboken, NJ, USA, 2003. [Google Scholar]
  26. Le Fevere de Ten Hove, C.; Penelle, J.; Ivanov, D.A.; Jonas, A.M. Encoding crystal microstructure and chain folding in the chemical structure of synthetic polymers. Nat. Mater. 2004, 3, 33–37. [Google Scholar] [CrossRef]
  27. Menges, M.G.; Penelle, J.; Le Fevere de Ten Hove, C.; Jonas, A.M.; Schmidt-Rohr, K. Characterization of long-chain aliphatic polyesters:  Crystalline and supramolecular structure of PE22,4 elucidated by X-ray scattering and nuclear magnetic resonance. Macromolecules 2007, 40, 8714–8725. [Google Scholar] [CrossRef]
  28. Roesle, P.; Stempfle, F.; Hess, S.K.; Zimmerer, J.; Río Bartulos, C.; Lepetit, B.; Eckert, A.; Kroth, P.G.; Mecking, S. Synthetic polyester from Algae oil. Angew. Chem. Int. Ed. 2014, 53, 6800–6804. [Google Scholar] [CrossRef]
  29. Atallah, P.; Wagener, K.B.; Schulz, M.D. ADMET: The future revealed. Macromolecules 2013, 46, 4735–4741. [Google Scholar] [CrossRef]
  30. Pribyl, J.; Wagener, K.B.; Rojas, G. ADMET polymers: Synthesis, structure elucidation, and function. Mater. Chem. Front. 2021, 5, 14–43. [Google Scholar] [CrossRef]
  31. Chen, Y.; Abdellatif, M.M.; Nomura, K. Olefin metathesis polymerization: Some recent developments in the precise polymerizations for synthesis of advanced materials (by ROMP, ADMET). Tetrahedron 2018, 74, 619–692. [Google Scholar] [CrossRef]
  32. Trnka, T.M.; Grubbs, R.H. The development of L2X2Ru=CHR olefin metathesis catalysts:  An organometallic success story. Acc. Chem. Res. 2001, 34, 18–29. [Google Scholar] [CrossRef]
  33. Samojzowicz, C.; Bieniek, M.; Grela, K. Ruthenium-based olefin metathesis catalysts bearing N-heterocyclic carbene ligands. Chem. Rev. 2009, 109, 3708–3742. [Google Scholar] [CrossRef]
  34. Grubbs, R.H.; Wenzel, A.G.; O’Leary, D.J.; Khosravi, E. (Eds.) Handbook of Metathesis, 2nd ed.; Wiley-VCH: Weinheim, Germany, 2015. [Google Scholar]
  35. Vougioukalakis, G.; Grubbs, R.H. Ruthenium-based heterocyclic carbene-coordinated olefin metathesis catalysts. Chem. Rev. 2010, 110, 1746–1787. [Google Scholar] [CrossRef]
  36. Oskam, H.J.; Fox, H.H.; Yap, B.K.; McConville, H.D.; O’Dell, R.; Lichtenstein, J.B.; Schrock, R.R. Ligand variation in alkylidene complexes of the type Mo(CHR)(NR′)(OR″)2. J. Organomet. Chem. 1993, 459, 185–198. [Google Scholar] [CrossRef]
  37. Schrock, R.R.; Hoveyda, A.H. Molybdenum and tungsten imido alkylidene complexes as efficient olefin-metathesis catalysts. Angew. Chem. Int. Ed. 2003, 42, 4592–4633. [Google Scholar] [CrossRef] [PubMed]
  38. Miyashita, T.; Kunisawa, M.; Sueki, S.; Nomura, K. Synthesis of poly(arylene vinylene)s containing different end groups by combined acyclic diene metathesis polymerization with Wittig-type coupling. Angew. Chem. Int. Ed. 2017, 56, 5288–5293. [Google Scholar] [CrossRef] [PubMed]
  39. Kojima, K.; Wang, X.; Go, L.; Makino, R.; Matsumoto, Y.; Shimoyama, D.; Abdellatif, M.M.; Kadota, J.; Hirano, H.; Nomura, K. Synthesis of high molecular weight bio-based aliphatic polyesters exhibiting tensile properties beyond polyethylene. ACS Macro Lett. 2023, 12, 1403–1408. [Google Scholar] [CrossRef] [PubMed]
  40. Rybak, A.; Meier, M.A.R. Acyclic diene metathesis with a monomer from renewable resources: Control of molecular weight and one-step preparation of block copolymers. ChemSusChem 2008, 1, 542–547. [Google Scholar] [CrossRef] [PubMed]
  41. Fokou, P.A.; Meier, M.A.R. Use of a renewable and degradable monomer to study the temperature-dependent olefin isomerization during ADMET polymerizations. J. Am. Chem. Soc. 2009, 131, 1664–1665. [Google Scholar] [CrossRef] [PubMed]
  42. Ulman, M.; Grubbs, R.H. Ruthenium carbene-based olefin metathesis initiators:  Catalyst decomposition and longevity. J. Org. Chem. 1999, 64, 7202–7207. [Google Scholar] [CrossRef]
  43. Lehman, S.E.; Wagener, K.B. Comparison of the kinetics of acyclic diene metathesis promoted by Grubbs ruthenium olefin metathesis catalysts. Macromolecules 2001, 35, 48–53. [Google Scholar] [CrossRef]
  44. Schmidt, B. Catalysis at the interface of ruthenium carbene and ruthenium hydride chemistry: Organometallic aspects and applications to organic synthesis. Eur. J. Org. Chem. 2004, 2004, 1865–1880. [Google Scholar] [CrossRef]
  45. Hong, S.H.; Wenzel, A.G.; Salguero, T.T.; Day, M.W.; Grubbs, R.H. Decomposition of ruthenium olefin metathesis catalysts. J. Am. Chem. Soc. 2007, 129, 7961–7968. [Google Scholar] [CrossRef] [PubMed]
  46. Higman, C.S.; Lanterna, A.E.; Marin, M.L.; Scaiano, J.C.; Fogg, D.E. Catalyst decomposition during olefin metathesis yields isomerization-active ruthenium nanoparticles. ChemCatChem 2016, 8, 2446–2449. [Google Scholar] [CrossRef]
  47. Jawiczuk, M.; Marczyk, A.; Trzaskowski, B. Decomposition of ruthenium olefin metathesis catalyst. Catalysts 2020, 10, 887. [Google Scholar] [CrossRef]
  48. Fokou, P.A.; Meier, M.A.R. Studying and suppressing olefin isomerization side reactions during ADMET polymerizations. Macromol. Rapid Commun. 2010, 31, 368–373. [Google Scholar] [CrossRef] [PubMed]
  49. Trzaskowski, J.; Quinzler, D.; Bährle, C.; Mecking, S. Aliphatic long-chain C20 polyesters from olefin metathesis. Macromol. Rapid Commun. 2011, 32, 1352–1356. [Google Scholar] [CrossRef] [PubMed]
  50. Stempfle, F.; Ortmann, P.; Mecking, S. Which polyesters can mimic polyethylene? Macromol. Rapid Commun. 2013, 34, 47–50. [Google Scholar] [CrossRef]
  51. Vilela, C.; Silvestre, A.J.D.; Meier, M.A.R. Plant oil-based long-chain C26 monomers and their polymers. Macromol. Chem. Phys. 2012, 213, 2220–2227. [Google Scholar] [CrossRef]
  52. Stempfle, F.; Quinzler, D.; Heckler, I.; Mecking, S. Long-chain linear C19 and C23 monomers and polycondensates from unsaturated fatty acid esters. Macromolecules 2011, 44, 4159–4166. [Google Scholar] [CrossRef]
  53. Roumanet, P.-J.; Jarroux, N.; Goujard, L.; Le Petit, J.; Raoul, Y.; Bennevault, V.; Guégan, P. Synthesis of linear polyesters from monomers based on 1,18-(Z)-octadec-9-enedioic acid and their biodegradability. ACS Sustain. Chem. Eng. 2020, 8, 16853–16860. [Google Scholar] [CrossRef]
  54. Nomura, K.; Chaijaroen, P.; Abdellatif, M.M. Synthesis of bio-based long-chain polyesters by acyclic diene metathesis polymerization and tandem hydrogenation and depolymerization with ethylene. ACS Omega 2020, 5, 18301–18312. [Google Scholar] [CrossRef]
  55. Ortmann, P.; Mecking, S. Long-spaced aliphatic polyesters. Macromolecules 2013, 46, 7213–7218. [Google Scholar] [CrossRef]
  56. Kojima, M.; Abdellatif, M.M.; Nomura, K. Synthesis of semi-crystalline long chain aliphatic polyesters by ADMET copolymerization of dianhydro-D-glucityl bis(undec-10-enoate) with 1,9-decadiene and tandem hydrogenation. Catalysts 2021, 11, 1098–1106. [Google Scholar] [CrossRef]
  57. Piccini, M.; Leak, D.J.; Chuck, C.J.; Buchard, A. Polymers from sugars and unsaturated fatty acids: ADMET polymerisation of monomers derived from d-xylose, d-mannose and castor oil. Polym. Chem. 2020, 11, 2681–2691. [Google Scholar] [CrossRef]
  58. Piccini, M.; Lightfoot, J.; Castro, D.B.; Buchard, A. Xylose-Based Polyethers and Polyesters Via ADMET Polymerizationtoward Polyethylene-Like Materials. ACS Appl. Polym. Mater. 2021, 3, 5870–5881. [Google Scholar] [CrossRef]
  59. Llevot, A.; Grau, E.; Carlotti, S.; Greliera, S.; Cramail, H. ADMET polymerization of bio-based biphenyl compounds. Polym. Chem. 2015, 6, 7693–7700. [Google Scholar] [CrossRef]
  60. Le, D.; Samart, C.; Kongparakul, S.; Nomura, K. Synthesis of new polyesters by acyclic diene metathesis polymerization of bio-based α,ω-dienes prepared from eugenol and castor oil (undecenoate). RSC Adv. 2019, 9, 10245–10252. [Google Scholar] [CrossRef] [PubMed]
  61. Hibert, G.; Grau, E.; Pintori, D.; Lecommandoux, S.; Cramail, H. ADMET polymerization of α,ω-unsaturated glycolipids: Synthesis and physico-chemical properties of the resulting polymers. Polym. Chem. 2017, 8, 3731–3739. [Google Scholar] [CrossRef]
  62. Shearouse, W.C.; Lillie, L.M.; Reineke, T.M.; Tolman, W.B. Sustainable Polyesters Derived from Glucose and Castor Oil: BuildingBlock Structure Impacts Properties. ACS Macro Lett. 2015, 4, 284–288. [Google Scholar] [CrossRef]
  63. Lillie, L.M.W.; Tolman, B.T.; Reineke, M. Structure/property relationships in copolymers comprising renewable isosorbide, glucarodilactone, and 2,5-bis(hydroxymethyl)-furan subunits. Polym. Chem. 2017, 8, 3746–3754. [Google Scholar] [CrossRef]
  64. Lebarbé, T.; Neqal, M.; Grau, E.; Alfos, C.; Cramail, H. Branched polyethylene mimicry by metathesis copolymerization of fatty acid-based α,ω-dienes. Green Chem. 2014, 16, 1755–1758. [Google Scholar] [CrossRef]
  65. Zhang, S.; Wang, J.; Lu, X.; Zhou, Q. Structures and Interactions of Ionic Liquids; Springer: Berlin/Heidelberg, Germany, 2014. [Google Scholar] [CrossRef]
  66. Zhang, S.; Lu, X.; Zhou, Q.; Li, X.; Zhang, X.; Li, S. Ionic Liquids. Physicochemical Properties; Elsevier B.V.: Amsterdam, The Netherlands, 2009. [Google Scholar] [CrossRef]
  67. Dong, K.; Liu, X.; Dong, H.; Zhang, X.; Zhang, S. Multiscale studies on ionic liquids. Chem. Rev. 2017, 117, 6636–6695. [Google Scholar] [CrossRef] [PubMed]
  68. Wang, Y.; He, H.; Wang, C.; Lu, Y.; Dong, K.; Huo, F.; Zhang, S. Insights into ionic liquids: From Z-bonds to quasi-liquids. JACS Au 2022, 2, 543–561. [Google Scholar] [CrossRef] [PubMed]
  69. Yang, J.; Lu, X.; Zhou, Q.; Xu, J.; Xin, J.; Zhang, S. Efficient biomass pretreatment process based on the simple reuse of a low-viscosity ionic-liquid solvent system. ACS Sustain. Chem. Eng. 2022, 10, 12738–12750. [Google Scholar] [CrossRef]
  70. Wang, X.; Zeng, J.; Lu, X.; Xin, J.; Zhang, S. High aluminum content beta zeolite as an active Lewis acid catalyst for γ-valerolactone decarboxylation. Ind. Eng. Chem. Res. 2019, 58, 11841–11848. [Google Scholar] [CrossRef]
  71. Amesho, K.T.T.; Lin, Y.-C.; Mohan, S.V.; Halder, S.; Ponnusamy, V.K.; Jhang, S.-R. Deep eutectic solvents in the transformation of biomass into biofuels and fine chemicals: A review. Environ. Chem. Lett. 2022, 21, 183–230. [Google Scholar] [CrossRef]
  72. Yang, Y.; Yang, S.; Yao, X.; Kang, Y.; Xin, J.; Ibrahim, T.; Xu, J.; Lu, X. A Renewable co-solvent promoting the selective removal of lignin by increasing the total number of hydrogen bonds. Green Chem. 2020, 22, 6393–6403. [Google Scholar] [CrossRef]
  73. Simocko, C.; Yang, Y.; Swager, T.M.; Wagener, K.B. Metathesis step-growth polymerizations in ionic liquid. ACS Macro Lett. 2013, 2, 1061–1064. [Google Scholar] [CrossRef]
  74. Zhao, J.; Wang, D.; Autenrieth, B.; Buchmeiser, M.R. First acyclic diene metathesis polymerization under biphasic conditions using a dicationic ruthenium alkylidene: Access to high-molecular-weight polymers with very low ruthenium contamination. Macromol. Rapid Commun. 2015, 36, 190–194. [Google Scholar] [CrossRef]
  75. Ponkratov, D.O.; Shaplov, A.S.; Vygodskii, Y.S. Metathesis polymerization in ionic media. Polym. Sci. Ser. C 2019, 61, 2–16. [Google Scholar] [CrossRef]
  76. Weychardt, H.; Plenio, H. Acyclic diene metathesis polymerization of divinylarenes and divinylferrocenes with Grubbs-type olefin metathesis catalysts. Organometallics 2008, 27, 1479–1485. [Google Scholar] [CrossRef]
  77. Sydlik, S.A.; Delgado, P.A.; Inomata, S.; VanVeller, B.; Yang, Y.; Swager, T.M.; Wagener, K.B. Triptycene-containing polyetherolefins via acyclic diene metathesis polymerization. J. Polym. Sci. Part A Polym. Chem. 2013, 51, 1695–1706. [Google Scholar] [CrossRef]
  78. Lucero, J.M.; Romero, Z.; Moreno, A.; Huber, D.L.; Simocko, C. ADMET polymerization in affordable, commercially available, high boiling solvents. SN Appl. Sci. 2020, 2, 647. [Google Scholar] [CrossRef]
  79. Wang, X.; Zhao, W.; Nomura, K. Synthesis of high molecular weight bio-based aliphatic polyesters by acyclic diene metathesis polymerization in ionic liquids. ACS Omega 2023, 8, 7222–7233. [Google Scholar] [CrossRef] [PubMed]
  80. Wang, X.; Chin, A.L.; Zhou, J.; Wang, H.; Tong, R. Resilient poly(α-hydroxy acids) with improved strength and ductility via scalable stereosequence-controlled polymerization. J. Am. Chem. Soc. 2021, 143, 16813–16823. [Google Scholar] [CrossRef]
  81. Nomura, K.; Aoki, T.; Ohki, Y.; Kikkawa, S.; Yamazoe, S. Transesterification of methyl-10-undecenoate and poly(ethylene adipate) catalyzed by (cyclopentadienyl)titanium trichlorides as model chemical conversions of plant oils and acid-, base-free chemical recycling of aliphatic polyesters. ACS Sustain. Chem. Eng. 2022, 10, 12504–12509. [Google Scholar] [CrossRef]
  82. Ohki, Y.; Ogiwara, Y.; Nomura, K. Depolymerization of polyesters by transesterification with ethanol using (cyclopentadienyl)titanium trichlorides. Catalysts 2023, 13, 421. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of bio-based polyesters from plant oil (triglycerides).
Scheme 1. Synthesis of bio-based polyesters from plant oil (triglycerides).
Catalysts 14 00097 sch001
Scheme 2. Ruthenium-carbene and molybdenum-alkylidene catalysts for the synthesis of aliphatic polyesters by ADMET polymerization.
Scheme 2. Ruthenium-carbene and molybdenum-alkylidene catalysts for the synthesis of aliphatic polyesters by ADMET polymerization.
Catalysts 14 00097 sch002
Scheme 3. ADMET polymerization of castor oil-derived monomers (M1, M2) [40,41].
Scheme 3. ADMET polymerization of castor oil-derived monomers (M1, M2) [40,41].
Catalysts 14 00097 sch003
Scheme 4. Synthesis of linear polyesters (LCAPEs) [49,50,51].
Scheme 4. Synthesis of linear polyesters (LCAPEs) [49,50,51].
Catalysts 14 00097 sch004
Figure 1. Plots of melting temperature (Tm) vs. number of (a) methylene units (x) in diol and (b) ester groups per 1000 C (methylene) in PE-23.x [50].
Figure 1. Plots of melting temperature (Tm) vs. number of (a) methylene units (x) in diol and (b) ester groups per 1000 C (methylene) in PE-23.x [50].
Catalysts 14 00097 g001
Scheme 5. One-pot synthesis of bio-based polyesters by Ru-catalyzed ADMET polymerization and hydrogenation [46].
Scheme 5. One-pot synthesis of bio-based polyesters by Ru-catalyzed ADMET polymerization and hydrogenation [46].
Catalysts 14 00097 sch005
Scheme 6. ADMET copolymerization of undec-10-en-1-yl undec-10-enoate (M1) or bis(undec-10-enoate) with isosorbide (M2) with nonconjugated dienes, and subsequent hydrogenation [55,56].
Scheme 6. ADMET copolymerization of undec-10-en-1-yl undec-10-enoate (M1) or bis(undec-10-enoate) with isosorbide (M2) with nonconjugated dienes, and subsequent hydrogenation [55,56].
Catalysts 14 00097 sch006
Figure 2. Plots of melting temperature (Tm) vs. number of ester groups per 1000 C (methylene units) in the hydrogenated copolymers, H2-poly(M1-co-UDD)s [55].
Figure 2. Plots of melting temperature (Tm) vs. number of ester groups per 1000 C (methylene units) in the hydrogenated copolymers, H2-poly(M1-co-UDD)s [55].
Catalysts 14 00097 g002
Scheme 7. ADMET polymerization of α,ω-dienes containing D-xylose, D-mannose, vanillin, and eugenol as the monomer units [57,58,59,60].
Scheme 7. ADMET polymerization of α,ω-dienes containing D-xylose, D-mannose, vanillin, and eugenol as the monomer units [57,58,59,60].
Catalysts 14 00097 sch007
Scheme 8. Synthesis of bio-based copolyesters with different molar ratios [61,62,63,64].
Scheme 8. Synthesis of bio-based copolyesters with different molar ratios [61,62,63,64].
Catalysts 14 00097 sch008
Scheme 9. Selected reports for the synthesis of high-molecular-weight aliphatic polyesters by the acyclic diene metathesis (ADMET) polymerization of α,ω-diene monomers containing carbohydrate units [58,62,63].
Scheme 9. Selected reports for the synthesis of high-molecular-weight aliphatic polyesters by the acyclic diene metathesis (ADMET) polymerization of α,ω-diene monomers containing carbohydrate units [58,62,63].
Catalysts 14 00097 sch009
Scheme 10. Synthesis of high-molecular-weight bio-based polyesters by ADMET polymerization in ionic liquids (ILs) and tandem hydrogenation, and depolymerization by olefin metathesis and transesterification [79].
Scheme 10. Synthesis of high-molecular-weight bio-based polyesters by ADMET polymerization in ionic liquids (ILs) and tandem hydrogenation, and depolymerization by olefin metathesis and transesterification [79].
Catalysts 14 00097 sch010
Scheme 11. Synthesis of high molecular mass polyesters by ADMET polymerization using the molybdenum catalyst [39].
Scheme 11. Synthesis of high molecular mass polyesters by ADMET polymerization using the molybdenum catalyst [39].
Catalysts 14 00097 sch011
Figure 3. Plots of tensile (fracture) strengths and strains (elongations at breaks) of HPE2 with different Mn values. The plots of PE18.18 (polyester—18.18) [9], commercially available polyethylene terephthalate (PET), poly(lactic acid) (PLA), high-density polyethylene (HDPE), low-density polyethylene (LDPE), and polypropylene (PP) [39].
Figure 3. Plots of tensile (fracture) strengths and strains (elongations at breaks) of HPE2 with different Mn values. The plots of PE18.18 (polyester—18.18) [9], commercially available polyethylene terephthalate (PET), poly(lactic acid) (PLA), high-density polyethylene (HDPE), low-density polyethylene (LDPE), and polypropylene (PP) [39].
Catalysts 14 00097 g003
Table 1. Synthesis of PE2 by ADMET polymerization using ruthenium catalysts [41] 1.
Table 1. Synthesis of PE2 by ADMET polymerization using ruthenium catalysts [41] 1.
Ru cat.Temp.
/°C
Nitrogen Purge 2Mn 3Mw/Mn 3Isomerization 4
/%
G260no56001.6548
G170no44001.573
G270no60001.7149
G180no47501.564
G280no61001.6169
G180yes66001.773
G280yes84001.7576
G190no54501.693
G290no62001.6566
G1100no50001.6142 5
1 Conditions: Ru cat 1.0 mol%, 5 h. 2 N2 purge during polymerization. 3 GPC in THF vs. polystyrene stds. 4 Isomerized diesters (%) estimated with GC-MS after transesterification. 5 Unidentified side products.
Table 2. ADMET polymerizations of M2, M6M8 by HG2 in [Hmim]TFSI [79] 1.
Table 2. ADMET polymerizations of M2, M6M8 by HG2 in [Hmim]TFSI [79] 1.
MonomerYield 2/%Mn 3Mw/Mn 3
M29339,2001.95
M248637,5001.91
M69226,0001.95
M78933,4002.30
M7 48734,9001.82
M89438,8003.38
1 Conditions: monomer (300 mg) in IL 0.14 mL (initial conc. 4.48 M (M2), 4.48 M (M6), 5.07 M (M7), 4.69 M (M8)), HG2 1.0 mol%, 50 °C in vacuo. 2 Isolated yield. 3 GPC data in THF versus polystyrene standards. 4 Reaction scale: monomer (1.0 g) in [Hmim]TFSI 0.30 mL (initial concentrations: 6.97 M (M2) and 7.80 M (M7)).
Table 3. ADMET polymerizations of M2, M6, and M19 with the molybdenum catalyst (25 °C, 6 h) [39] 1.
Table 3. ADMET polymerizations of M2, M6, and M19 with the molybdenum catalyst (25 °C, 6 h) [39] 1.
Monomer (μmol)cat./mol%Yield 2/%Mn 3/g·mol−1Mw/Mn 3
M2 (90.5)5.09916,0001.79
M2 (90.5)2.59025,1001.43
M2 (90.5)1.08834,4001.49
M2 (272)1.08846,1002.08
M2 (272)1.09146,1001.84
M6 (272)1.08734,8001.87
M19 (272)1.09967,2002.27
M2 (272)0.59048,7002.04
M2 (543) 40.59149,4002.47
1 Conditions: Mo(CHCMe2Ph)(N-2,6-Me2C6H3)[OC(CH3)(CF3)2]2 (Mo), toluene (0.72 mL), quenched by C6H5CHO or 4-Me3SiOC6H3CHO (for termination through Wittig-type cleavage). 2 Isolated yield (as MeOH insoluble fraction). 3 GPC data in THF (at 40 °C) vs. polystyrene standards. 4 Toluene: 1.0 mL.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nomura, K.; Wang, X. Acyclic Diene Metathesis (ADMET) Polymerization for the Synthesis of Chemically Recyclable Bio-Based Aliphatic Polyesters. Catalysts 2024, 14, 97. https://doi.org/10.3390/catal14020097

AMA Style

Nomura K, Wang X. Acyclic Diene Metathesis (ADMET) Polymerization for the Synthesis of Chemically Recyclable Bio-Based Aliphatic Polyesters. Catalysts. 2024; 14(2):97. https://doi.org/10.3390/catal14020097

Chicago/Turabian Style

Nomura, Kotohiro, and Xiuxiu Wang. 2024. "Acyclic Diene Metathesis (ADMET) Polymerization for the Synthesis of Chemically Recyclable Bio-Based Aliphatic Polyesters" Catalysts 14, no. 2: 97. https://doi.org/10.3390/catal14020097

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop